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
|
---|---|---|---|---|---|---|---|---|---|
1104.3324 | 1 | 1104 | 2011-04-17T16:13:27 | S-Bases in S-Linear Algebra | [
"math.FA",
"quant-ph"
] | In this paper we define the S-bases for the spaces of tempered distributions. These new bases are the analogous of Hilbert bases of separable Hilbert spaces for the continuous case (they are indexed by m-dimensional Euclidean spaces) and enjoy properties similar to those shown by algebraic bases in the finite dimensional case. The S-bases are one possible rigorous and extremely manageable mathematical model for the "physical" bases used in Quantum Mechanics. | math.FA | math |
SBases in SLinear Algebra
David Carf`ı
Abstract
In this paper we define the S bases for the spaces of tempered distri-
butions. These new bases are the analogous of Hilbert bases of separable
Hilbert spaces for the continuous case (they are indexed by m-dimensional
Euclidean spaces) and enjoy properties similar to those shown by algebraic
bases in the finite dimensional case. The Sbases are one possible rigorous
and extremely manageable mathematical model for the "physical" bases
used in Quantum Mechanics.
1 S Linear independence
A finite family v of vectors of a vector space is said linearly independent if any
zero linear combination of v is given by a zero system of coefficients. This is
exactly the definition we generalize in the following.
Definition (of Slinear independence). Let v ∈ S(Rm, S′
n) be an Sfamily
of tempered distributions in the space S′
n indexed by the Euclidean space Rm.
The family v is said to be S linearly independent, if any coefficient distribution
a ∈ S′
m such that
ZRm
av = 0S ′
n
m. In other terms the family v is said S linearly
must be the zero distribution 0S ′
independent if and only if any zero Slinear combination of the family v has
necessarily a zero coefficient system.
We shall denote the superposition of a family of distribution v with respect
to a distributional system of coefficients a, also by the multiplication symbol
a.v.
1
Example (The Dirac family). The Dirac family in S′
n is Slinearly inde-
pendent. In fact, we have
ZRn
uδ = u,
for all u ∈ S′
n, and then the relation u.δ = 0S ′
n is equivalent to u = 0S ′
n .
Example (the Fourier families). Let a, b be two real non-zero num-
bers, recall that the (a, b)-Fourier family in the space of tempered distributions
S′(Rn, C) is the family
of smooth and bounded regular tempered distributions (recall that [f ] denote
the distribution canonically associated to a locally summable function f ) . In
the particular case a = 1 and b = −1/ (with the reduced Planck constant)
we obtain what we call the De Broglie family, i.e. the family
(cid:16)a−nhe−ib(p·)i(cid:17)p∈Rn
,
uϕ = 0S ′
n(C).
.
(cid:16)he(i/)(p·)i(cid:17)p∈Rn
ZRn
0 = (cid:18)ZRn
uϕ(cid:19) (φ) =
= u (bϕ (φ)) =
= u(cid:0)S(a,b) (φ)(cid:1) =
= F(a,b) (u) (φ) ,
All the Fourier families are Slinearly independent. In fact, let ϕ be the (a, b)-
Fourier family, and let
For every test function φ in Sn(C), we have
where S(a,b) is the Fourier-Schwartz transformation on the space Sn(C), so that
the Fourier transform of the distribution u must be zero, i.e.
F(a,b) (u) = 0S ′
n(C),
and hence u = 0S ′
erator.
n(C), being the Fourier transformation F(a,b) an injective op-
2 Linear and Slinear independence
A first elementary connection between the classic linear independence and the
new Slinear independence is given by the following theorem, which states that
2
the Slinear independence is a stronger requirement for an Sfamily than the
requirement of the simple linear independence.
Recall that an infinite family v of vectors of a vector space is said to be
linearly independent if any finite subfamily of v is linearly independent.
Theorem. Let v ∈ S(Rm, S′
n) be an Slinearly independent family. Then,
the family v is linearly independent. Consequently, the Slinear hull Sspan (v)
is an infinite dimensional subspace of S′
n.
Proof. By contradiction, assume the family v linearly dependent. Then
there exists a linearly dependent finite subfamily of v. More precisely, there are
a positive integer k ∈ N and a k-sequence α ∈ (Rm)k of points belonging to the
m-dimensional Euclidean space Rm such that the k-sequence of distributions
vα = (vαi )k
i=1, extracted from the family v by means of the index selection α is
a linearly dependent system of S′
n. Then there exists a non-zero scalar k-tuple
a ∈ Kk such that the finite linear combinationP avα is zero, that is such that
aivαi = 0S ′
n .
kXi=1
=
=
kXi=1
kXi=1
= 0S ′
n.
aivαi =
i=1 aiδαi as a coefficient distribution,
we have
Consider the tempered distribution d =Pk
dv = ZRm kXi=1
ai(cid:18)ZRm
ZRm
aiδαi! v =
δαi v(cid:19) =
Now, since the distribution d is different from the zero distribution 0S ′
m, the
preceding equality contradicts the Slinear independence of the family v, against
our assumptions. (cid:4)
3 Topology and Slinear independence
The last theorem of the above section shows that, for what concerns the Sfamilies,
the Slinear independence implies the usual linear independence. Actually, the
3
Slinear independence is more restrictive than the linear independence, as we
shall see later by a simple notable example of family which is linearly indepen-
dent but not Slinearly independent. On the contrary the Slinear independence
is less restrictive than the β(S′
n)-topological independence, as it is shown below.
Topological independence. We recall that a system of vectors v = (vi)i∈I
n)-topologically
n)-topologically free) if and only if there exists a family
n)-continuous) linear forms
n, indexed by a non-void index set I, is said β(S′
n)-continuous (respectively, σ(S′
in the space S′
free (respectively, σ(S′
L = (Li)i∈I of β(S′
on S′
n such that
Li(vk) = δik,
for any pair (i, k) ∈ I 2, where the family δ = (δik)(i,k)∈I 2 is the Kronecker family
on the square I 2. Note that the above relation can be written as L ⊗ v = δ,
where δ is the Kronecker family.
If the family v is not topologically free it is said topologically bound. If the
family v is topologically free, any family L of continuous linear forms, satisfying
the above relations is said a dual family of the family v. So, to say that a family
v is topologically free is equivalent to say that v has a dual family of linear
continuous forms.
Recalling that (by reflexivity) any continuous linear functional on the space
S′
n is canonically and univocally representable by a test function in Sn, to
say that the family v is topologically free is equivalent to say that the bi-
orthonormality condition
hgi, vki = δik,
is true, for any pair (i, k) ∈ I 2, where g is a suitable family of test functions.
Theorem. Every Sfamily in the space S′
n)-topologically bound and,
n)-topologically bound. Consequently, no Sfamily has a dual family of
n is β(S′
thus, σ(S′
test functions.
Proof. Let v be any Sfamily in the space S′
n indexed by Rm. And let L
be an arbitrary family in the dual S′′
n indexed by the same index set. Being
the Schwartz space (Sn) reflexive, for every i, there is a test function gi in Sn
canonically generating the functional Li, that is such that
Li = h., gii .
In other terms, the test function gi is such that Li(u) = u(gi), for every tempered
distribution u in S′
n. Assume the existence of an index i such that Li(vi) = 1,
then we deduce
1 = Li(vi) = vi(gi) = v(gi) (i) ,
being v an Sfamily, the function v(gi) is continuous, then there is a neighborhood
U of the point i in which the function v(gi) is strictly positive. Then, for every
point k in the neighborhood U , we have
Li(vk) = vk(gi) = v(gi) (k) > 0,
4
and then L cannot verify the condition
Li(vk) = δik,
for any pair (i, k) ∈ I 2. So we cannot find a dual family of functionals for v
and, consequently, v cannot be topologically independent. (cid:4)
Note. By the same proof, it is possible to prove that every C0-family
of tempered distributions is strongly topologically bound. Consequently every
smooth family of tempered distributions is strongly topologically bound.
4 Multiplicity of representations
It's simple to prove the following property that characterizes the Slinear depen-
dence of a family of distributions by explicit multeplicity of representations of
some member of the family itself.
Property. An Sfamily v in S′
n indexed by Rm is Slinearly dependent if and
only if there is a point index p in Rm and a tempered distribution a in S′
m
different from the Dirac delta distribution δp such that the p-th member of the
family is representable also by
vp =ZRm
av.
Proof. Necessity.
Indeed, if vp fulfills that property we have that the
Slinear combination (a − δp).v is zero with a non-zero coefficient distribution,
so that the family v is Slinearly dependent. Sufficiency. Vice versa, let, for
every index-point p, the term vp of the family be representable in a unique way
as the superposition vp = δp.v. Assume v Slinearly dependent, then there is a
coefficient system a different from zero such that a.v = 0, hence
vp = ZRm
= ZRm
= ZRm
δpv − 0 =
δpv −ZRm
(δp − a)v,
av =
since a is a non-zero distribution, the distribution δp − a is different form the
Dirac delta δp, and so the member vp is representable in another (different) way,
against the assumption. (cid:4)
5
5 Characterizations of Slinear independence
By the Dieudonn´e-Schwartz theorem we immediately deduce two characteriza-
tions.
We say an Sfamily v to be topologically exhaustive (with respect to the weak*
n)-closed (or
topology or the strong topology) if its Slinear hull Sspan (v) is σ(S′
strongly closed, which is the same).
Theorem. Let v ∈ S(Rm, S′
an Sfamily whose Slinear hull Sspan (v) is σ(S′
assertions are equivalent
n) be a topologically exhaustive family, that is
n)-closed. Then the following
• 1) the family v is Slinearly independent;
m) and σ(S′
n);
morphism for the weak* topologies σ(S′
morphism for the strong topologies β(S′
• 2) the superposition operator RRm (·, v) is an injective topological homo-
• 3) the superposition operator RRm (·, v) is an injective topological homo-
• 4) the operator bv is a surjective topological homomorphism for the weak
• 5) the operator bv is a surjective topological homomorphism of the topolog-
ical vector space (Sn) onto the space (Sm).
topologies σ(Sn) and σ(Sm);
m) and β(S′
n);
Theorem. Let v ∈ S(Rm, S′
n) be an Sfamily. Then the following assertions
are equivalent
• 1) the family v is Slinearly independent and the hull S span (v) is σ(S′
n)-
closed;
morphism for the weak* topologies σ(S′
• 2) the superposition operator RRm (·, v) is an injective topological homo-
• 3) the operator bv is a surjective topological homomorphism for the weak
• 4) the operator bv is a surjective topological homomorphism from the space
topologies σ(Sn) and σ(Sm);
(Sn) onto the space (Sm).
m) and σ(S′
n);
Remark (on the coordinate operator). In the conditions of the above
theorem, if the family v is S linearly independent, we can consider the algebraic
isomorphism from the space S′
m onto the Slinear hull Sspan (v) sending every
6
ZRm
(·, v) : S′
m → S′
n : (a, v) 7→ a.v
tempered distribution a ∈ S′
of the injection
m into the superposition a.v, that is the restriction
to the pair of sets (S′
phism by the symbol [·v].
theorem that
m,S span (v)). We shall denote the inverse of this isomor-
It is an important consequence of the preceding
• Theorem. The operator
[·v] :S span (v) → S′
m
is a topological isomorphism, with respect to the topology induced by the
n) on the Slinear hull Sspan (v) and to the weak* topol-
weak* topology σ(S′
ogy σ(S′
n)-closed, that
is if the family v is topologically exhaustive.
m), if and only if the Slinear hull Sspan (v) is σ(S′
6 S Bases
Definition (of Sbasis). Let v ∈ S(Rm, S′
a subspace of the space S′
if it is Slinearly independent and it Sgenerates V , that is
n and let V be
n. The family v is said an S basis of the subspace V
n) be an Sfamily in S′
Sspan(v) = V.
In other terms, the family v is said an S basis of the subspace V if and only
if the superposition operator of the family v is injective and its range coincides
with the subspace V .
The Dirac family δ of S′
n is an Sbasis of the whole S′
n (indeed its superpo-
sition operator is the identity on S′
n). We call the Dirac family δ the canonical
Sbasis of S′
n (because of reasons which will appear clear soon, and that can be
summarized into the relation u.δ = u, valid for any tempered distribution u) or
the Dirac basis of S′
n.
Moreover, the following complete version of the Fourier expansion-theorem,
allow us to call the Fourier families of S′(Rn, C) by the name of Fourier bases
of S′(Rn, C).
Theorem (geometric form of the Fourier expansion theorem). In the
n(C) the Fourier families are Sbases
space of complex tempered distributions S′
(of the entire space S′
n(C)).
Proof. Indeed, the superposition operators of the Fourier families in S′
n are
the Fourier transforms upon S′
n which are bijective. (cid:4)
7
7 Algebraic characterizations of S bases
The following is an elementary but meaningful generalization of the Fourier
expansion theorem.
Theorem (characterization of an Sbasis). Let v ∈ S(Rm, S′
n) be an
Sfamily. Then,
• 1) the family v Sgenerates the space S′
n if and only if the superposition
• 2) the family v is S linearly independent if and only if the superposition
operator t (bv) is surjective;
operator t (bv) is injective;
operator t (bv) is bijective.
• 3) the family v is an Sbasis of the space S′
n if and only if the superposition
Proof. The proof follows immediately from the definitions, however we see it.
Moreover, it is obvious, by the very definitions, that the family v Sgenerates
the space S′
First of all the superposition operator t (bv) is well defined because v is an Sfamily.
n if and only if the superposition operator t (bv) is surjective, and
that v is Slinearly independent if and only if the superposition operator t (bv) is
injective. (cid:4)
7.1 Example
The following point gives us an example of linearly independent family which is
not Slinearly independent and also an example of a linearly independent system
of S generators which is not an Sbasis.
Example (a system of linearly independent Sgenerators that is not
1 of the first derivatives of the
an Sbasis). Let v = (δ′
Dirac distributions. The family v is of class S, in fact
x)x∈R be the family in S′
v(φ)(x) = vx(φ) = δ′
x(φ) = −φ′(x),
and the the derivative φ′ is an S function. Consequently, the operator associated
with the family v is the derivation in the test function space S1 up to the sign
1. This
last superposition operator is a surjective operator (every tempered distribution
has a primitive) but it is not injective (every tempered distribution has many
primitives), then the family v is a system of Sgenerators for the space S′
1, but
it is not Slinearly independent. Moreover, note that, however, the family v is
and, then, the superposition operator t (bv) is the derivation in the space S′
8
linearly independent. In fact, let P be a finite subset of the real line R, and let,
for every point p0 in P , fp0 be a test function in S1 such that
f ′
p0 (p) = δp0,p,
for every index p in P , here δ(.,.) is the Kronecker delta upon the square P 2.
Now, if a = (ap)p∈P is a finite family of scalars such that
then
apδp0p = ap0,
apvp = 0S ′
1,
Xp∈P
apvp (fp0 ) =Xp∈P
0 =Xp∈P
for every index p0 in P . And so, any finite linear combination of members of v
is zero only with respect to a zero-system of coefficients.
8 Totality of Sbases
We give another way to characterize an S basis.
Theorem (characterization of an Sbasis). Let v ∈ S(Rm, S′
n) be an
Sfamily. Then:
1) the family v is a system of S generators of the entire space S′
n if and only
if the family v is total in the space Sn (in the sense that if vp(g) = 0,
for every index-point p, then g = 0) and the linear hull of the family v is
weakly* closed;
2) the family v is Slinearly independent if and only it is total in the distri-
bution space S′
m, in the sense that if a.v = 0 then a = 0;
3) a family v is an Sbasis of the space S′
n if and only if the family v is total
both in the function space Sn and distribution space S′
m.
Proof. 1) Indeed, the condition means that the operator generated by the
family v is injective and this (by the Dieudonn`e-Schwartz theorem) implies
the surjectivity of the superposition operator of v, since the image of the su-
perposition operator of v is closed. 2) This is exactly the definition of linear
independence. 3) Follows from the two before. (cid:4)
9
9 Topological characterizations of S bases
By the Dieudonn´e-Schwartz theorem we immediately take a characterization.
Theorem. Let v ∈ S(Rm, S′
n) be a family of tempered distributions. Then
the following assertions are equivalent
• 1) the family v is an S basis of the space S′
n;
m) and σ(S′
n);
m) and β(S′
n);
strong topologies β(S′
weak* topologies σ(S′
• 2) the superposition operator RRm(·, v) is a topological isomorphism for the
• 3) the superposition operator RRm(·, v) is a topological isomorphism for the
• 4) the operator bv is a topological isomorphism for the weak topologies
• 5) the operator bv is a topological isomorphism of the topological vector
space (Sn) onto (Sm).
σ(Sn) and σ(Sm);
References
[1] J. Barros-Neto, An Introduction to the theory of distributions, Marcel
Dekker, Inc. NewYork, 1981
[2] N. Boccara, Functional analysis, an introduction for physicists, Academic
press, Inc. 1990
[3] J. Horvath, Topological Vector Spaces and Distributions (Vol.I), Addison-
Wesley Publishing Company, 1966
[4] L. Schwartz, Th´eorie des distributions, Hermann, Paris 1966
[5] J. Dieudonn´e, "La dualit´e dans les espaces vectoriels topologiques",
Annales scietifiques de l'E.N.S. 3 es´erie, tome 59 p. 107 - 139, 1942.
http://www.numdam.org
[6] J. Dieudonn´e L. Schwartz, "La dualit´e dans les espaces (F ) and (LF )", An-
nales de l'institut Fourier, tome 1 p. 61 - 101, 1949. http://www.numdam.org
[7] P. A. M. Dirac, The principles of Quantum Mechanics, Oxford Claredon
press, 1930.
[8] D. Carf`ı, "S-linear operators
in quantum mechanics and in eco-
nomics", (APPS), volume 6 pp.7-20, 2004. (no.1 electronic edition, )
http://vectron.mathem.pub.ro/apps/v6/a6.htm
10
[9] D. Carf`ı, "Dirac-orthogonality in the space of tempered distributions", Jour-
nal of computational and applied mathematics, vol. 153, pp.99-107, numbers
1-2, 1 april 2003.
[10] D. Carf`ı, "S-diagonalizable operators in quantum mechanics", Glasnik
Mathematicki, vol. 40 n.2, pp. 267-307, 2005.
[11] D. Carf`ı, "On the Schodinger's equation associated with an operator...",
Rendiconti del Seminario Matematico di Messina, n. 8 serie II, 2001.
[12] D. Carf`ı, "Quantum statistical systems with a continuous range of states",
series of Advances in Mathematics for Applied Sciences, vol. 69, pp. 189 -
200, edited by World Scientific. 2005.
[13] D. Carf`ı, "Dyson formulas for Financial and Physical evolutions in S′-n",
Proceedings of the "VIII Congresso SIMAI", Ragusa, Baia Samuele, 22 - 26
Maj 2006.
[14] D. Carf`ı, "Prigogine approach to irreversibility for Financial and Phys-
ical applications", Supplemento Atti dell'Accademia Peloritana dei Peri-
colanti di Messina, Proceedings Thermocon'05, 2006. ISSN: 0365-0359.
http://antonello.unime.it/atti/
[15] D. Carf`ı, "S-convexity in the space of Schwartz distributions and applica-
tions", Supplemento ai Rendiconti del Circolo Matematico di Palermo, Serie
II - Numero 77 - pp. 107-122, Anno 2006,. ISSN: 0009-725X.
[16] D. Carf`ı, "S-Linear Algebra in Economics and Physics", Applied Sciences
(APPS), vol. 9, 2007, ISSN 1454-5101, forthcoming volume.
David Carf`ı
Faculty of Economics
University of Messina
[email protected]
11
|
1612.03775 | 2 | 1612 | 2017-07-10T17:02:10 | Angular equivalence of normed spaces | [
"math.FA"
] | Angular equivalence is introduced and shown to be an equivalence relation among the norms on a fixed real vector space. It is a finer notion than the usual (topological) notion of norm equivalence. Angularly equivalent norms share certain geometric properties: A norm that is angularly equivalent to a uniformly convex norm is itself uniformly convex. The same is true for strict convexity. Extreme points of the unit balls of angularly equivalent norms occur on the same rays, and if one unit ball is a polyhedron so is the other.
Among norms arising from inner products, two norms are angularly equivalent if and only if they are topological equivalent. But, unlike topological equivalence, angular equivalence is able to distinguish between different norms on a finite-dimensional space. In particular, no two $\ell^p$ norms on $\mathbb R^n$ are angularly equivalent. | math.FA | math |
ANGULAR EQUIVALENCE OF NORMED SPACES
EDER KIKIANTY AND GORD SINNAMON
Abstract. Angular equivalence is introduced and shown to be an equivalence
relation among the norms on a fixed real vector space.
It is a finer notion
than the usual (topological) notion of norm equivalence. Angularly equivalent
norms share certain geometric properties: A norm that is angularly equivalent
to a uniformly convex norm is itself uniformly convex. The same is true for
strict convexity. Extreme points of the unit balls of angularly equivalent norms
occur on the same rays, and if one unit ball is a polyhedron so is the other.
Among norms arising from inner products, two norms are angularly equiv-
alent if and only if they are topological equivalent. But, unlike topological
equivalence, angular equivalence is able to distinguish between different norms
on a finite-dimensional space. In particular, no two ℓp norms on Rn are angu-
larly equivalent.
1. Introduction and Definition
Two norms on a real vector space are equivalent if both give rise to the same
notion of convergence. A wide variety of functional analysis results concern only the
topology a norm generates, not the specific values taken by a given norm. In such
a setting, choosing the most convenient one among several, or many, topologically
equivalent norms can clarify arguments and simplify proofs.
In other situations, specific properties of individual norms are central to the
theory. For example, uniform convexity is an important property of a norm that
may not be shared by a topologically equivalent norm.
It is our object here to introduce a finer equivalence of norms, one that preserves
certain properties of the norm that simple topological equivalence does not. The
idea is straightforward; two norms are angularly equivalent if, over all pairs of non-
zero vectors, the angle between the pair, determined by one norm, is comparable
to the angle between the same pair, determined by the other norm. This will be
made precise shortly.
Although the theory of angles in normed spaces cannot match the elegance of its
counterpart for inner product spaces, it serves here to give us a means of defining
an equivalence of norms that compares vectors two by two rather than one at at
time, as topological equivalence does. Thus, angular equivalence emerges as a kind
of "second order" equivalence compared to the "first order" topological equivalence.
Definition 1. Two norms, k·k1 and k·k2, on a real vector space X are topologically
equivalent provided there exist positive constants m, M such that for all x, y ∈ X,
(1.1)
mkxk1 ≤ kxk2 ≤ Mkxk1.
2010 Mathematics Subject Classification. 46B (primary), 46B20 (secondary).
Supported by the Natural Sciences and Engineering Research Council of Canada.
1
2
EDER KIKIANTY AND GORD SINNAMON
Very little is needed to give the definition of angular equivalence besides an
accessible concept of angle in normed spaces. We define an angle based on the
g-functional, introduced in [5] and studied in [2, Chapter 4] and references there.
It it closely connected to smoothness and convexity properties of the unit ball and
lends itself to straightforward calculations.
1
t→0+
t (kx+tyk−kxk) and g−(x, y) = kxk lim
Fix vectors x and y in a real vector space X, with norm k·k. A few applications
of the triangle inequality are enough to show that 1
t (kx + tyk − kxk) is a non-
decreasing function of t taking (−∞, 0) ∪ (0,∞) into [−kyk,kyk]. It follows that
both
g+(x, y) = kxk lim
exist, and satisfy
(1.2)
−kxkkyk ≤ kxk(kxk−kx−yk) ≤ g−(x, y) ≤ g+(x, y) ≤ kxk(kx+yk−kxk) ≤ kxkkyk.
Definition 2. Suppose k · k is a norm on a real vector space X. The g-functional
relative to k · k is the map g : X × X → [0,∞) given by, g = 1
2 (g− + g+). If x
and y are non-zero vectors in X, the norm angle1 from x to y is θ(x, y), defined by
0 ≤ θ ≤ π and
t (kx+tyk−kxk)
t→0−
1
cos θ(x, y) =
g(x, y)
kxkkyk
.
We will refrain from referring to the norm angle "between" x and y, since the
norm angle from x to y may not coincide with the norm angle from y to x. If the
norm in X arises from an inner product, it is easy to see that norm angles agree
with angles defined by the inner product. To see that θ(x, y) does not depend on
the lengths of x and y, make the substitution s = bt/a in the defining limits to get
g(ax, by) = abg(x, y) whenever a, b > 0. A little extra care shows that this equation
holds for any a, b ∈ R so, in particular, g(x,−y) = −g(x, y).
Definition 3. Two norms, k · k1 and k · k2, on a real vector space X are angularly
equivalent provided there exists a constant C such that for all non-zero x, y ∈ X,
(1.3)
Here θ1(x, y) and θ2(x, y) are the norm angles from x to y relative to k · k1 and
k · k2, respectively. Also, tan(π/2) is taken to be +∞.
tan(θ2(x, y)/2) ≤ C tan(θ1(x, y)/2).
It is clear from the definition that angular equivalence is a reflexive and transitive
relation. To see that it is also symmetric, replace y by −y in (1.3). Then, for
j = 1, 2, gj(x,−y) = −gj(x, y), cos θj(x,−y) = − cos θj(x, y) and
1 + cos θj(x, y)(cid:19)1/2
tan(θj(x, y)/2) =(cid:18) 1 − cos θj(x, y)
1 − cos θj(x,−y)(cid:19)1/2
=(cid:18) 1 + cos θj(x,−y)
So if (1.3) holds for all non-zero x, y then so does
(1.4)
=
1
tan(θj(x,−y)/2)
.
tan(θ1(x, y)/2) =
1
tan(θ1(x,−y)/2) ≤
C
tan(θ2(x,−y)/2)
= C tan(θ2(x, y)/2).
1The term "g-angle" is in use, coined by Pavle Milici´c for an angle based on a symmetrized
g-functional.
ANGULAR EQUIVALENCE OF NORMED SPACES
3
Thus angular equivalence is an equivalence relation.
Since (1.3) implies (1.4) with the same constant C, angular equivalence has a
strong form of symmetry. It follows that the constant C is necessarily at least 1.
Remark 1. We will see later that for "most" x and y, g− and g+ coincide. When
they differ, the choice of g as the average of the two is unimportant for our pur-
poses; the notion of angular equivalence does not depend on the value of g at these
exceptional pairs. Moreover, at pairs for which g− and g+ coincide, any semi-
inner product necessarily agrees with their common value. So angular equivalence
is not dependent on the specific functional used to define it. See [2] for a thorough
discussion of semi-inner products. This remark is supported by Theorem 3.3 and
Corollaries 3.4 and 3.5 below.
The use of the trigonometric ratio tan(θ/2) in the definition of angular equiva-
lence may seem arbitrary, since there are many possible ways to express the con-
dition that two angles be comparable. Our choice is motivated by the situation
for inner product spaces, where the ratio tan(θ/2) appears in the following sharp
inequality relating the best constants for topological and angular equivalence.
Proposition 1.1 ([4]). Let X be a real vector space equipped with inner products
h·,·i1 and h·,·i2, giving rise to norms k ·k1 and k ·k2 in the usual way. Let m, M ∈
[0,∞] be the best constants in the inequality
mkxk1 ≤ kxk2 ≤ Mkxk1,
x ∈ X.
Then, for any non-zero x, y ∈ X,
tan(θ2(x, y)/2) ≤ (M/m) tan(θ1(x, y)/2).
The constant M/m is best possible in this inequality. Here θ1 and θ2 are the angles
between x and y relative to k·k1 and k·k2, respectively. (The usual angles coincide
with the norm angles in inner product spaces.)
This result is a reformulation of the generalized Wielandt inequality. The origi-
nal, based on [8] and appearing in [1], involves the action of an invertible matrix on
Cn, relating the angle between a pair of vectors and the angle between the trans-
formed vectors. Since M/m < ∞ if and only if the two norms are topologically
equivalent, the reformulated inequality shows that if two norms arise from inner
products on the same vector space, then the norms are angularly equivalent if and
only if they are topologically equivalent.
This is an indication that angular equivalence is not too much finer than topo-
logical equivalence; the two notions coincide on an important class of spaces. As
an indication that angular equivalence is nonetheless considerably finer than topo-
logical equivalence see Corollaries 2.3 and 2.4 below, where we show that, unlike
topological equivalence, norms on a finite-dimensional spaces are not all angularly
equivalent. (Of course, we have not yet established that angular equivalence is in
fact finer than topological equivalence. But we will; see Theorem 4.2.)
2. A First Look at Angular Equivalence
Many properties shared by angularly equivalent norms may be deduced from the
definition with minimal calculation, while others emerge only after careful analysis.
In this section we present a selection of the former.
4
EDER KIKIANTY AND GORD SINNAMON
Recall that if E is a subset of a real vector space X we say x ∈ E is an extreme
point of E provided x is not in any open line segment contained in E. When X is
a normed space and E is the closed unit ball of X this is of particular interest in
functional analysis. Our first result shows that the unit balls of angularly equivalent
norms share the same extreme "directions".
Theorem 2.1. Let X be a real vector space having two norms, k · k1 and k · k2.
Suppose k · k1 and k · k2 are angularly equivalent and x is a non-zero vector in X.
Then x/kxk1 is an extreme point of the k · k1-unit ball if and only if x/kxk2 is an
extreme point of the k · k2-unit ball.
Proof. We argue the contrapositive. Suppose x/kxk2 is not an extreme point of the
k · k2-unit ball. Then there are points y and z in X such that (y + z)/2 = x/kxk2
If
and the closed line segment from y to z is contained in the k · k2-unit ball.
s ∈ [0, 1] then (1 − s)y + sz and sy + (1 − s)z are on the line segment and hence in
the k · k2-unit ball. Thus,
2 = ky+zk2 = k(1−s)y+sz+sy+(1−s)zk2 ≤ k(1−s)y+szk2+ksy+(1−s)zk2 ≤ 2.
It follows that k(1 − s)y + szk2 = 1. In particular, observe that kyk2 = kzk2 = 1.
Now,
g±
2 (y, z) = lim
t→0±
= lim
s→0±
= lim
s→0±
1
1−s
t (ky + tzk2 − 1)
s (ky + s
s (k(1 − s)y + szk2 − 1 + s) = 1.
1−s zk2 − 1)
1
This shows that g2(y, z) = 1, cos(θ2(y, z)) = 1, and tan(θ2(y, z)/2) = 0. By angular
equivalence, tan(θ1(y, z)/2) = 0 as well. This implies cos(θ1(y, z)) = 1 and hence
g1(y, z) = kyk1kzk1. The last statement, which may be written as
g−
1 (y, z) + g+
1 (y, z) = 2kyk1kzk1,
combined with
g−
1 (y, z) ≤ g+
1 (y, z) ≤ kyk1(ky + zk1 − kyk1) ≤ kyk1kzk1,
from (1.2), gives
Since ky + zk1 = kyk1 + kzk1 and x/kxk2 = (y + z)/2, we have
kyk1(ky + zk1 − kyk1) = kyk1kzk1.
x
kxk1
=
y + z
ky + zk1
=
kyk1
kyk1 + kzk1
y
kyk1
+
kzk1
kyk1 + kzk1
,
z
kzk1
which is a convex combination of the points y/kyk1 and z/kzk1. Thus, x/kxk1 is
an interior point of the line segment from y/kyk1 to z/kzk1. Since the endpoints of
this segment lie in the k · k1-unit ball, convexity shows that the entire line segment
does. Thus, x/kxk1 is not an extreme point of the k · k1-unit ball.
Reversing the roles of the two norms gives the other implication.
(cid:3)
A normed space is strictly convex if every boundary point of the unit ball is an
extreme point. It is immediate that strict convexity is preserved by angular equiv-
alence. The corresponding result for uniform convexity also holds, see Corollary
2.7.
ANGULAR EQUIVALENCE OF NORMED SPACES
5
Corollary 2.2. Suppose X be a real vector space having two angularly equivalent
norms, k · k1 and k · k2. Then X is strictly convex when equipped with k · k1 if and
only if X is strictly convex when equipped with k · k2.
The unit ball of finite-dimensional normed space is polygonal if and only if it has
only finitely many extreme points, namely, the vertices of the polygon.
Corollary 2.3. Let X be a finite-dimensional real vector space having two angularly
equivalent norms, k · k1 and k · k2. Then the k · k1-unit ball is polygonal if and only
if the k · k2-unit ball is polygonal. In this case, the vertices of the two polygons lie
on the same rays.
Remark 2. If two norms are angularly equivalent on X then the restrictions of
those two norms to any subspace Y of X are also angularly equivalent. This follows
easily from the definition. However, the subspace Y may have extreme points that
the original space did not have. For every subspace Y , the rays containing extreme
points relative to Y are the same for both norms.
One of the most telling features of angular equivalence is its ability to distinguish
between norms on a finite-dimensional space. Here we see that no two ℓp norms
are angularly equivalent.
(This only applies in finite dimensions. The angular
equivalence of the usual ℓp sequence space norms is a question that does not arise
naturally, since they are not norms on the same underlying vector space.)
Corollary 2.4. Suppose p, q ∈ [1,∞]. For any integer n ≥ 2, the ℓp and ℓq norms
on Rn are angularly equivalent only if p = q.
Proof. If the two norms are angularly equivalent then their restrictions to the sub-
space R2 of Rn (embedded in the usual way) are also angularly equivalent. So we
may assume n = 2. The unit ball in the ℓ1 norm is a square with vertices at (0,±1)
and (±1, 0). The unit ball in the ℓ∞ norm is a square with vertices at (±1,±1).
When 1 < p < ∞, the unit ball in the ℓp norm is not a polygon. By Corollary 2.3,
the ℓ1 norm and the ℓ∞ norm are not angularly equivalent to each other or to any
other ℓp norm.
It remains to consider p, q ∈ (1,∞). Fix s > 0 and consider the vectors (1, 0)
and (1, s). The value of the ℓp-norm g-functional for this pair is,
1
t→0
t (k(1, 0)+t(1, s)kℓp−k(1, 0)kℓp) = lim
k(1, 0)kℓp lim
Thus θℓp ≡ θℓp ((1, 0), (1, s)) satisfies,
t→0
1
t (((1+t)p +tpsp)1/p−1) = 1.
cos θℓp =
1
(1 + sp)1/p
and
tan2 θℓp =
(1 + sp)1/p − 1
(1 + sp)1/p + 1
.
By angular equivalence of the ℓp and ℓq norms, there exists a C > 0, independent
of s such that
1
C ≤(cid:18) (1 + sp)1/p − 1
(1 + sp)1/p + 1(cid:19)(cid:18) (1 + sq)1/q + 1
(1 + sq)1/q − 1(cid:19) ≤ C
Taking the limit as s → 0+ it is easy to see that this expression remains bounded
above and below only if p = q.
(cid:3)
The next theorem may be viewed as a stability result for angular equivalence.
6
EDER KIKIANTY AND GORD SINNAMON
Theorem 2.5. Suppose k · k1 and k · k2 are angularly equivalent norms on a real
vector space X. Then the norm k·k3 = k·k1 +k·k2 is angularly equivalent to k·k1
and k · k2.
Proof. Fix non-zero x, y ∈ X. Let gj(x, y) and θj = θj(x, y) be the g-functional
and norm angle from x to y with respect to the norm k · kj, for j = 1, 2, 3. The
definition of the g-functional shows that
g3(x, y)
kxk3
=
g1(x, y)
kxk1
+
g2(x, y)
kxk2
,
kyk3 cos θ3 = kyk1 cos θ1 + kyk2 cos θ2.
or equivalently,
Therefore,
tan2(θ3/2) =
1 − cos θ3
1 + cos θ3
=
=
kyk1
kyk3
kyk1
kyk3
1 + kyk2
kyk1
1 + kyk2
kyk1
(1 − cos θ1) + kyk2
kyk3
(1 + cos θ1) + kyk2
kyk3
(1 − cos θ2)
(1 + cos θ2)
1−cos θ2
1−cos θ1
1+cos θ2
1+cos θ1
tan2(θ1/2).
But for any A, B > 0, 1+A
1+B ≤ 1 + A
B , so
tan2(θ2/2)
tan2(θ3/2) ≤(cid:18)1 +
tan2(θ1/2)(cid:19) tan2(θ1/2) ≤ (1 + C2) tan2(θ1/2),
where C is the constant of angular equivalence of k·k1 and k·k2 from (1.3). Taking
square roots shows that k · k3 is angularly equivalent to k · k1, and hence to k · k2
as well.
(cid:3)
With stability under sums of norms we can easily give an example to show that
an angular equivalency class containing inner product norms may contain other
norms as well.
Example 1. A norm that is angularly equivalent to a norm arising from an inner
product need not arise from an inner product: Consider the two norms on R2 defined
by k(ξ, η)k1 = (3ξ2 + η2)1/2 and k(ξ, η)k2 = (ξ2 + 3η2)1/2. These both arise from
inner products and are topologically equivalent so by Theorem 1.1 they are angularly
equivalent. Theorem 2.5 shows that their sum, k(ξ, η)k3 = k(ξ, η)k1 + k(ξ, η)k2 is
also angularly equivalent. However, it is easy to check that k·k3 does not arise from
an inner product. (For example, the parallelogram law fails for the vectors (1, 0)
and (0, 1).)
The stability of angular equivalence does not extend to the maximum of two
norms.
Example 2. The maximum of two angularly equivalent norms is a norm that need
not be angularly equivalent to the original two. Let k · k1 and k · k2 be the norms in
the previous example, set k(ξ, η)k4 = max(k(ξ, η)k1,k(ξ, η)k2), and let gj(x, y) and
θj = θj(x, y) be the g-functional and norm angle from x to y with respect to the
norm k · kj, for j = 1, 4. Now let s > 0 and take x = (1, 1) and y = (1 − s, 1 + s).
Calculations show that
kx + tyk1 = 2p(1 + t)2 − (1 + t)ts + t2s2
and g1(x, y) = 4 − 2s;
ANGULAR EQUIVALENCE OF NORMED SPACES
7
and also that
kx+tyk4 = 2p(1 + t)2 + (1 + t)ts + t2s2,
But kxk1 = kxk4 = 2, kyk1 = 2√1 − s + s2 and kyk4 = 2√1 + s + s2. So, using
g±
4 (x, y) = 4±2s,
and g4(x, y) = 4.
the definition of cos θ1(x, y) and cos θ4(x, y), we get,
3s2
tan2(θ1(x, y)/2) =
4(1 − s/2 + √1 − s + s2)2
and
tan2(θ4(x, y)/2) =
s + s2
(1 + √1 + s + s2)2
.
Clearly, there is no constant C for which tan(θ4(x, y)/2) ≤ C tan(θ1(x, y)/2) as
s → 0+. Thus k · k4 is not angularly equivalent to k · k1.
We close this section with the promised proof that uniform convexity is preserved
by angular equivalence. A real vector space X, with norm k·k, is uniformly convex
provided that for all ε > 0 there exists a δ > 0 such that if kxk = kyk = 1 and
kx − yk ≥ ε then k(x + y)/2k ≤ 1 − δ. First we show that unform convexity can
be characterized in terms of the norm angle. The idea appears in [7], where closely
related results are proved.
Theorem 2.6. Let X be a real vector space with norm k · k and let θ(x, y) denote
the norm angle from x to y. Then X is uniformly convex if and only if for all
ε > 0 there exists a δ > 0 such that if kxk = kyk = 1 and kx − yk ≥ ε then
tan(θ(x, y)/2) ≥ δ.
Proof. Suppose X is uniformly convex. Fix ε > 0 and choose an η > 0 so that
k(x + y)/2k ≤ 1 − η whenever kxk = kyk = 1 and kx − yk ≥ ε. Set δ = √η. If
kxk = kyk = 1 and kx−yk ≥ ε then (1.2) shows that −1 ≤ g(x, y) ≤ kx+yk−1 ≤ 1
so
tan(θ(x, y)/2) =s 1 − g(x, y)
1 + g(x, y) ≥r 1 − g(x, y)
2
≥r1 −(cid:13)(cid:13)(cid:13)
x + y
2 (cid:13)(cid:13)(cid:13) ≥ √η = δ.
For the converse, fix ε > 0 and choose an η > 0 such that if kxk = kyk = 1 and
kx− yk ≥ ε/4 then tan(θ(x, y)/2) ≥ η. Set δ = min(η2/(1 + η2), ε/4). Now suppose
kxk = kyk = 1 and kx − yk ≥ ε, and set z = x + y. If z = 0, then the desired
conclusion, kz/2k ≤ 1 − δ, is trivial. Otherwise,
k(2 − kzk)x − kzk((z/kzk) − x)k = kx − yk ≥ ε,
so either k(2 − kzk)xk ≥ 2δ or kkzk((z/kzk) − x)k ≥ ε − 2δ. The first of the two
implies that kz/2k ≤ 1 − δ and completes the proof. The choice of δ ensures that
ε − 2δ ≥ ε/2 so the second implies k(z/kzk) − xk ≥ ε/(2kzk) ≥ ε/4, and we have
η ≤ tan(θ(z/kzk, x)/2) =s 1 − g(z/kzk, x)
1 + g(z/kzk, x)
.
But this is equivalent to g(z/kzk, x) ≤ (1 − η2)/(1 + η2). Using (1.2) again, this
time with x and y replaced by z and x, respectively, gives kzk − 1 ≤ g(z/kzk, x) so
we have,
kz/2k ≤ 1
2 (1 + g(z/kzk, x)) ≤
1
2(cid:16)1 +
1 − η2
1 + η2(cid:17) = 1 −
η2
1 + η2 ≤ 1 − δ.
8
EDER KIKIANTY AND GORD SINNAMON
The next result uses the previous theorem so it is included here despite its
dependence on Theorem 4.2 below.
(cid:3)
Corollary 2.7. Suppose X be a real vector space having two angularly equivalent
norms, k · k1 and k · k2. Then X is uniformly convex when equipped with k · k1 if
and only if X is uniformly convex when equipped with k · k2.
Proof. Let C be the constant in the definition of angular equivalence so that, using
symmetry and (1.3), for all non-zero x and y in X we have tan(θ1(x, y)/2) ≤
C tan(θ2(x, y)/2).
Theorem 4.2 shows that k·k1 and k·k2 are also topologically equivalent so there
Suppose X is uniformly convex when equipped with k ·k1 and fix ε > 0. Choose
η > 0 so that, if kxk1 = kyk1 = 1 and kx− yk1 ≥ mε/(2M ), then tan(θ1(x, y)/2) ≥
η. Set δ = η/C.
Let kxk2 = kyk2 = 1 and kx − yk2 ≥ ε, and set x = x/kxk1 and y = y/kyk1.
Clearly, kxk1 = kyk1 = 1. Also, (as in the proof of the Dunkl-Williams inequality,)
exist constants m and M such that (1.1) holds.
x
ε ≤ kx − yk2 =(cid:13)(cid:13)(cid:13)(cid:13)
kxk2 −
≤ kx − yk2
kxk2
y
+
y
kxk2 −
kxk2
+ kxk2 − kyk2
kxk2
y
kyk2(cid:13)(cid:13)(cid:13)(cid:13)2
≤
2kx − yk2
.
kxk2
But kxk2 = kxk2/kxk1 ≥ m so kx − yk1 ≥ (1/M )kx − yk2 ≥ mε/(2M ). This
ensures that tan(θ1(x, y)/2) ≥ η. But θ1(x, y) = θ1(x, y) so angular equivalence
implies tan(θ2(x, y)/2) ≥ η/C = δ. This shows that X is uniformly convex when
equipped with k · k2. For the other implication, reverse the roles of k · k1 and
k · k2.
(cid:3)
3. Norms in the Plane
Since angular equivalence is defined in terms of pairs of vectors, it is natural to
study it first in two-dimensional spaces. Our analysis in the plane is more than
the investigation of a special case, however. It enables us to establish results for
real vector spaces of any dimension: It is evident from the definition that, if a pair
of norms on a real vector space X are angularly equivalent then the restrictions
of those norms to any subspace of X are also angularly equivalent. We will make
use of the equally evident converse: If the restrictions of two norms on X to any
two-dimensional subspace of X are angularly equivalent with constant C, then the
original norms are also angularly equivalent with the same constant C. Uniform
control of the constant is essential here.
In the plane we can study norms by viewing the boundary of their unit balls
as polar functions. The following setup and notation will be used throughout this
section and the next. Here and throughout, the symbol "∠" denotes the usual angle
in the plane, not the norm angle.
Suppose k·k is a norm on R2 and define r > 0 by requiring kr(t)(cos t, sin t)k = 1
for all t ∈ R. Then r is a strictly positive, π-periodic function and the unit ball,
{ρ(cos t, sin t) : 0 ≤ ρ ≤ r(t), t ∈ R}, is a closed, convex set with (0, 0) in its
interior. The convexity of the unit ball readily implies that r is continuous. Since
r is continuous and periodic it attains its minimum and maximum values; both
ANGULAR EQUIVALENCE OF NORMED SPACES
9
are strictly positive. Call them rm and rM , respectively. Let O = (0, 0) and
parametrize the boundary of the ball by setting Pt = r(t)(cos t, sin t) for all t ∈ R.
If 0 < α − β < π, set
ψ(α, β) =(∠OPαPβ,
α > β,
π − ∠OPαPβ, α < β.
Also define ϕ−(α) = limβ→α− ψ(α, β) and ϕ+(α) = limβ→α+ ψ(α, β). These exist
by property (ii) below.
Lemma 3.1. Suppose α, β ∈ R with 0 < α − β < π. Then:
(i) ψ(α, β) + α = ψ(β, α) + β;
(ii) ψ(α, t) is non-decreasing for t in (α − π, α) ∪ (α, α + π);
(iii) 0 < ψ(α, β) < π and, if 0 < α − β ≤ π/2, then
0 < tan−1(rm/rM ) ≤ ψ(α, β) ≤ π − tan−1(rm/rM ) < π;
(iv) r(α)/r(β) = cos(α − β) + sin(α − β) cot ψ(α, β) and ψ is continuous;
(v) the left derivative of r at α exists and equals r(α) cot ϕ−(α), the right
derivative of r at α exists and equals r(α) cot ϕ+(α), and r is differentiable
wherever ϕ− = ϕ+;
(vi) ϕ−(α) ≤ ϕ+(α) and if α < β then ϕ+(α) + α ≤ ϕ−(β) + β;
(vii) ϕ− is left continuous and ϕ+ is right continuous;
(viii) ϕ− and ϕ+ are continuous at all but countably many points, and ϕ− is
continuous at α if and only if ϕ+ is continuous at α if and only if ϕ−(α) =
ϕ+(α).
Proof. The angles of triangle PαOPβ add to π. If α > β they are α − β, ψ(α, β),
and π − ψ(β, α) and if α < β they are β − α, ψ(β, α) and π − ψ(α, β). This proves
(i).
To prove (ii), suppose s < t and both lie in (α − π, α) ∪ (α, α + π). If s < t < α
then, by convexity, the segment PsPα intersects the segment OPt. It follows that,
ψ(α, s) = ∠0PαPs ≤ ∠0PαPt = ψ(α, t).
If s < α < t then the segment PsPt intersects the segment OPα. Thus
ψ(α, s) + π − ψ(α, t) = ∠OPαPs + ∠OPαPt = ∠PsPαPt ≤ π.
If α < s < t then the segment PαPt intersects the segment OPs. It follows that,
π − ψ(α, s) = ∠0PαPs ≥ ∠0PαPt = π − ψ(α, t).
In each of the three cases we have ψ(α, s) ≤ ψ(α, t), as required.
Since r(α) > 0, r(β) > 0, and 0 < α − β < π, △PαOPβ is non-degenerate.
In particular, 0 < ψ(α, β) < π, the first statement of (iii). For the other, suppose
0 < α − β < π/2. Then, using (ii),
ψ(α, β) ≥ ψ(α, α − π/2) = tan−1(r(α − π/2)/r(α)) ≥ tan−1(rm/rM ) > 0
and
ψ(α, β) ≤ ψ(α, α + π/2) = π − tan−1(r(α + π/2)/r(α)) ≤ π − tan−1(rm/rM ) < π.
The sine law in △OPαPβ gives
sin ψ(α, β)
=
sin(∠OPαPβ)
=
r(β)
r(β)
sin(∠OPβ Pα)
r(α)
=
sin(π − ψ(α, β) − (α − β))
r(α)
10
EDER KIKIANTY AND GORD SINNAMON
when α > β and
sin(π − ψ(α, β))
r(β)
=
sin(∠OPαPβ)
r(β)
=
sin(∠OPβ Pα)
r(α)
=
sin(ψ(α, β) − (β − α))
r(α)
when α < β. These both reduce to the equation in (iv). Continuity of r now implies
continuity of ψ.
To prove (v) we use (iv) to get
r(α) − r(s)
α − s
= r(s)(cid:18) cos(α − s) − 1
α − s
+
sin(α − s)
α − s
cot ψ(α, s)(cid:19) .
The definitions of ϕ+ and ϕ− (and the continuity of r) show that the left and right
derivatives of r exist and are equal to,
lim
s→α−
r(α) − r(s)
α − s
= r(α) cot ϕ−(α)
and
lim
s→α+
r(α) − r(s)
α − s
= r(α) cot ϕ+(α),
respectively. It is immediate that r is differentiable wherever ϕ+ = ϕ−.
By (ii),
ϕ−(α) = lim
s→α−
ψ(α, s) ≤ lim
t→α+
ψ(α, t) = ϕ+(α).
Also, if α < β then, by (i) and (ii),
ϕ+(α) + α = lim
t→α+
ψ(α, t) + α ≤ ψ(α, β) + α
= ψ(β, α) + β ≤ lim
s→β−
ψ(β, s) + β = ϕ−(β) + β.
These prove (vi).
For (vii), observe that if γ > α, applying (vi), (ii), and then (iv) gives
ϕ+(α) ≤ lim
s→α+
= lim
s→α+
ϕ−(s) + s − α ≤ lim
lim
t→s+
ψ(s, t) ≤ lim
s→α+
ϕ+(s)
s→α+
ψ(s, γ) = ψ(α, γ).
Letting γ → α+ shows ϕ+(α) ≤ lims→α+ ϕ+(s) ≤ ϕ+(α). Thus, ϕ+ is right-
continuous at α. A similar argument shows that ϕ− is left-continuous.
By (vi), the functions s 7→ ϕ−(s) + s and t 7→ ϕ+(t) + t are both non-decreasing
and hence continuous except at countably many points. It follows that ϕ− and ϕ+
are both continuous except at countably many points, the first statement of (viii).
The second may be deduced from the following consequence of (vi) and (vii):
ϕ−(α) = lim
s→α−
ϕ−(s) ≤ lim
s→α−
ϕ+(s) ≤ ϕ−(α)
≤ ϕ+(α) ≤ lim
t→α+
ϕ−(t) ≤ lim
t→α+
ϕ+(t) = ϕ+(α).
(cid:3)
The g-functional and norm angle for a norm on R2 can be expressed in terms of
the functions r, ψ and φ.
ANGULAR EQUIVALENCE OF NORMED SPACES
11
Theorem 3.2. Suppose k · k is a norm on R2, define r, ψ and ϕ± as above, and
let Qt = (cos t, sin t) for t ∈ R, If a, b > 0 and 0 < α − β < π, then
g(aQα, bQβ) =
cos θ(aQα, bQβ) =
tan2(θ(aQα, bQβ)/2) =
ab
r(α)2 (cid:0)cos(α − β) + 1
cot(α − β) + 1
2 (cot ϕ−(α) + cot ϕ+(α)) sin(α − β)(cid:1) ,
2 (cot ϕ−(α) + cot ϕ+(α))
,
and
cot(α − β) + cot ψ(α, β)
cot ψ(α, β) − 1
2 (cot ϕ−(α) + cot ϕ+(α))
2 cot(α − β) + cot ψ(α, β) + 1
2 (cot ϕ−(α) + cot ϕ+(α))
.
If a, b > 0 and α = β then
g(aQα, bQβ) =
ab
r(α)2 ,
cos θ(aQα, bQβ) = 1,
and
tan2(θ(aQα, bQβ)/2) = 0.
ab
r(α)2 ,
If a, b > 0 and α − β = π then
cos θ(aQα, bQβ) = −1, and tan2(θ(aQα, bQβ)/2) = ∞.
g(aQα, bQβ) = −
Proof. Since a and b are non-zero, for t sufficiently close to zero we may define c
and γ ∈ (α − π/2, α + π/2) as functions of t, by requiring that cQγ = aQα + tbQβ.
Evidently, both c and γ are differentiable. Equating components, we have the
equations,
c cos γ = a cos α + tb cos β and c sin γ = a sin α + tb sin β.
Differentiating these with respect to t gives
c′ cos γ − cγ′ sin γ = b cos β and c′ sin γ + cγ′ cos γ = b sin β.
This system is readily solved to yield c′ = b cos(β − γ) and cγ′ = b sin(β − γ). The
definition of r ensures that kr(t)Qtk = 1 for all t, so
r(α)(cid:19) =
t (kcQγk − kaQαk) =
t r(α) − r(γ)−r(α)
t (cid:18) c
γ−α
r(α)r(γ)
r(γ) −
c−a
a γ−α
t
a
1
1
.
As t → 0 we have
t → aγ′(0) = b sin(β−α).
γ → α,
c → a,
Notice that γ is a strictly monotone function of t in a neighbourhood of 0. Lemma
3.1(v) shows that
t → c′(0) = b cos(β−α) and a γ−α
c−a
lim
t→0±
r(γ) − r(α)
γ − α
= r(α) cot ϕ±(α)
or
lim
t→0±
= r(α) cot ϕ∓(α),
r(γ) − r(α)
γ − α
depending on whether γ is increasing or decreasing. In either case,
g(aQα, bQβ) = 1
2kaQαk(cid:0) lim
t→0+
a
1
t (kcQγk − kaQαk) + lim
t→0−
1
t (kcQγk − kaQαk)(cid:1)
b cos(β − α)r(α) − r(α) 1
2 (cot ϕ+(α) + cot ϕ−(α))b sin(β − α)
r(α)2
=
=
r(α)
ab
r(α)2 (cid:0)cos(α − β) + 1
2 (cot ϕ−(α) + cot ϕ+(α)) sin(α − β)(cid:1) .
12
EDER KIKIANTY AND GORD SINNAMON
Using this formula and Lemma 3.1(iv), we get
cos θ(aQα, bQβ) =
=
g(aQα, bQβ)
kaQαkkbQβk
g(aQα, bQβ)
ab/(r(α)r(β))
=
But tan2(θ/2) = (1 − cos θ)/(1 + cos θ), so,
tan2(θ(aQα, bQβ)/2) =
cot ψ(α, β) − 1
cot(α − β) + 1
2 (cot ϕ−(α) + cot ϕ+(α))
.
cot(α − β) + cot ψ(α, β)
2 cot(α − β) + cot ψ(α, β) + 1
2 (cot ϕ−(α) + cot ϕ+(α))
2 (cot ϕ−(α) + cot ϕ+(α))
.
The formulas in the case α = β and α− β = π are easily established directly from
the definitions of the g-functional and norm angle. We omit the details.
(cid:3)
If we have two norms, k · k1 and k · k2, on R2 we define rj, ψj and ϕ±
j as
above based on k · kj, for j = 1, 2. The previous theorem shows that the angular
equivalence of these two norms can be expressed in the form,
(3.1)
cot ψ2(α, β) − A2
2 cot(α − β) + cot ψ2(α, β) + A2 ≤ C2
cot ψ1(α, β) − A1
2 cot(α − β) + cot ψ1(α, β) + A1
for all α, β ∈ R with 0 < α − β < π. Here Aj = 1
j (α)) for j =
1, 2. The cases β = α and β = α±π need not be included, as tan(θ2(aQα, bQβ)/2) ≤
C tan(θ1(aQα, bQβ)/2) holds trivially in those cases.
2 . Next we see
However, one can avoid discontinuities of the functions ϕ±
j (α) + cot ϕ+
1 and ϕ±
2 (cot ϕ−
2 (β),
2 (β) = ϕ−
2 (α) = ϕ−
1 (β), ϕ+
2 (α), ϕ+
1 (α), ϕ+
that it is enough to consider only values of α for which A1 and A2 simplify.
Theorem 3.3. The norms k · k1 and k · k2 on R2 are angularly equivalent with
constant C if and only if (3.1) holds for all α, β satisfying
1 (β) = ϕ−
1 (α) = ϕ−
(3.2) ϕ+
and 0 < α − β < π.
Proof. If the two norms are angularly equivalent, then by Theorem 3.2, (3.1) holds
for all α, β such that 0 < α − β < π. Conversely, suppose (3.1) holds for all α, β
satisfying (3.2) and 0 < α − β < π. We will show that (3.1) holds for all α, β
satisfying 0 < α − β < π.
2 (t)} has a
countable complement and is therefore dense in R. By hypothesis, for each α ∈ E,
inequality (3.1) holds for all β ∈ E∩((α−π, α)∪(α, α+π)). The continuity of ψ1 and
ψ2, from Lemma 3.1(iv), shows that it remains valid for all β ∈ (α−π, α)∪(α, α+π).
Fix a β ∈ R. Then (3.1) holds for all α ∈ E ∩ ((β − π, β) ∪ (β, β + π)). Taking
one-sided limits of (3.1) with respect to α, but only through points of E, gives, for
each α ∈ (β − π, β) ∪ (β, β + π), the two inequalities, (one with "+" and one with
"−")
(3.3)
By Lemma 3.1(viii) the set E = {t ∈ R : ϕ+
2 (t) = ϕ−
1 (t) = ϕ−
1 (t), ϕ+
2 (α)
cot ψ2(α, β) − cot ϕ±
2 cot(α−β) + cot ψ2(α, β) + cot ϕ±
1 (α)
Fix an α ∈ (β − π, β) ∪ (β, β + π). A convexity argument will show that for the
fixed β and α, (3.3) implies (3.1).
C2(cot ψ1(α, β) − cot ϕ±
1 (α))
2 cot(α−β) + cot ψ1(α, β) + cot ϕ±
2 (α) ≤
.
ANGULAR EQUIVALENCE OF NORMED SPACES
13
(3.4)
j (α) + z cot ϕ+
Let Aj(z) = (1 − z) cot ϕ−
cot ψ2(α, β) − A2(z)
2 cot(α − β) + cot ψ2(α, β) + A2(z) ≤ C2
j (t) for j = 1, 2. The inequality
cot ψ1(α, β) − A1(z)
2 cot(α − β) + cot ψ1(α, β) + A1(z)
holds for z = 0, 1 because of (3.3). If we show that it holds for z = 1/2 then we
have (3.1).
It is important to point out that in each of fractions in (3.4) the numerator and
denominator cannot both be zero, lest cot(α − β) + cot ψj(α, β) = 0. If this were
true, then either ψj(α, β) = β − α or ψj(α, β) = β − α + π. But by Lemma 3.1(i)
this is ψj(β, α) = 0 or ψj(β, α) = π, contrary to Lemma 3.1(iii).
Define f : [0, 1] → [0,∞) by
f (z) = C2(cot ψ1(α, β) − A1(z))(2 cot(α − β) + cot ψ2(α, β) + A2(z))
− (cot ψ2(α, β) − A2(z))(2 cot(α − β) + cot ψ1(α, β) + A1(z)).
First consider the case β < α < β + π. Lemma 3.1(ii) shows that for j = 0, 1
and 0 ≤ z ≤ 1,
and we get,
0 < ψj(α, β) ≤ ϕ−
j (α) ≤ ϕ+
j (α) ≤ ψj(α, β + π) < π
cot ψj (α, β) ≥ cot ϕ−
j (α) ≥ Aj(z) ≥ cot ϕ+
j (α) ≥ cot ψj(α, β + π).
Since r(β) = r(β + π), Lemma 3.1(iv) implies
cos(α− β) + sin(α− β) cot ψj(α, β) = cos(α− β− π) + sin(α− β− π) cot ψj(α, β + π)
so cot ψj(α, β + π) = −2 cot(α− β)− cot ψj (α, β). We conclude that the numerators
and denominators of both sides of inequality (3.4) are all non-negative. It follows
that (3.4) holds if and only if f (z) ≥ 0.
j = 0, 1 and 0 ≤ z ≤ 1,
The second case is similar. If β − π < α < β then Lemma 3.1(ii) shows that for
0 < ψj(α, β − π) ≤ ϕ−
j (α) ≤ ϕ+
j (α) ≤ ψj(α, β) < π
and we get,
cot ψj (α, β − π) ≥ cot ϕ−
j (α) ≥ Aj(z) ≥ cot ϕ+
j (α) ≥ cot ψj(α, β).
Since r(β) = r(β − π), Lemma 3.1(iv) implies
cos(α− β + π) + sin(α− β + π) cot ψj(α, β− π) = cos(α− β) + sin(α− β) cot ψj(α, β)
so cot ψj(α, β − π) = −2 cot(α − β) − cot ψj(α, β). This time the numerators and
denominators of both sides of inequality (3.4) are all non-positive but again (3.4)
holds if and only if f (z) ≥ 0.
It remains to show that f (1/2) ≥ 0.
The function f (z) is a quadratic polynomial and the coefficient of z2 is
−(C2 − 1)(cot ϕ−
1 (t) − cot ϕ+
1 (t))(cot ϕ−
2 (t) − cot ϕ+
2 (t)) ≤ 0
so it is a concave function. Since (3.4) holds for z = 0 and z = 1, f (0) ≥ 0 and
f (1) ≥ 0. It follows that f (z) ≥ 0 for all z ∈ [0, 1] and in particular f (1/2) ≥ 0. (cid:3)
14
EDER KIKIANTY AND GORD SINNAMON
This theorem may be used to simplify verification of angular equivalence of
norms on R2 but it also shows that the notion of angular equivalence does not
depend on the value of the two g-functionals at the exceptional pairs for which
either g-functional fails to satisfy g−(x, y) = g+(x, y).
Corollary 3.4. Let k·k1 and k·k2 be norms on the real vectors space X. To show
that the norms are angularly equivalent with constant C it is enough to verify (1.3)
for pairs x, y satisfying g+
2 (x, y) = g−
Proof. Suppose that (1.3) holds for pairs x, y satisfying g+
j (x, y) for
j = 1, 2. A calculation shows that this case includes all linearly dependent pairs.
So fix a pair of independent vectors x and y in X. We identify their two-dimensional
span with R2 by the map (ξ, η) 7→ ξx + ηy so that k(ξ, η)kj = kξx + ηykj is a norm
on R2 for j = 1, 2. This identification is isometric in both norms so it does not
affect g-functional calculations or norm angles.
2 (x, y)
j (x, y) = g−
1 (x, y) and g+
1 (x, y) = g−
j and ϕ−
j
j (aQα, bQβ) = g−
Using these norms on R2 we define ϕ+
for j = 1, 2. Suppose the condi-
tions (3.2) hold for some α, β satisfying 0 < α− β < π. The proof of Theorem 3.2
shows that g+
j (aQα, bQβ) for j = 1, 2 and for any a, b > 0. But
the points aQα, bQβ ∈ R2 correspond to vectors in the span of x and y. Our hy-
pothesis, shows that tan(θ2(aQα, bQβ)/2) ≤ C tan(θ1(aQα, bQβ)/2). By Theorem
3.2 inequality (3.1) also holds. Now Theorem 3.3 shows that the two norms on R2
are angularly equivalent with constant C. The same is true for the original two
norms on the span of x and y. In particular, (1.3) holds for the vectors x, y. This
completes the proof.
(cid:3)
Let X be a real vector space. A semi-inner product in the sense of Lumer-Giles,
is a map [·,·] : X × X → [0,∞) that is linear in the first variable, homogeneous
in the second, positive definite, and satisfies [x, y]2 ≤ [x, x][y, y].
It follows that
x 7→ [x, x]1/2 is a norm on X. See [2, Chapter 2] and the references there. It is
natural to define the angle θ[·,·] from x to y, associated with semi-inner product
[·,·], by requiring that 0 ≤ θ ≤ π and
cos θ[·,·] =
[y, x]
[y, y]1/2[x, x]1/2 .
Corollary 3.6 in [2] shows that any semi-inner product in the sense of Lumer-
Giles satisfies g−(x, y) ≤ [y, x] ≤ g+(x, y). In particular, [y, x] = g(x, y) whenever
g−(x, y) = g+(x, y) for all x, y ∈ X. This proves the following.
Corollary 3.5. Let X be a real vector space and let [·,·]1 and [·,·]2 be semi-inner
products in the sense of Lumer-Giles, giving rise to norms k · k1 and k · k2, respec-
tively. If (1.3) holds with θ[·,·]1 and θ[·,·]2 replacing the norm angles θ1 and θ2 then
k · k1 and k · k2 are angularly equivalent.
It is worth pointing out that if X is a normed space, setting [y, x] = g(x, y)
defines a semi-inner product in the sense of Lumer-Giles only under additional
conditions on the norm. See [2, Definition 4.2.9 and Proposition 4.2.10].
4. Angular Equivalence Implies Topological Equivalence
The techniques developed in the last section are used to prove that angular
equivalence is finer than topological equivalence. The definitions of rj , ψj, and ϕ±
j ,
ANGULAR EQUIVALENCE OF NORMED SPACES
15
relative to the norm k·kj, for j = 1, 2, that were introduced at the beginning of Sec-
tion 3 will be used throughout. We begin by showing that the angular equivalence
of two norms gives a relationship between their ψ and ϕ-functions.
Lemma 4.1. Let k · k1 and k · k2 be angularly equivalent norms on R2. If 0 <
α − β < π/2, then
(α + ϕ−
2 (β)) ≤ M C2((α + ϕ−
1 (α)) − (β + ϕ+
2 (α)) − (β + ϕ+
1 (β))),
where M = sup{csc2 ψ1(t, s) : s, t ∈ (β, α), s 6= t}.
Proof. Let E be the set of points in (β, α) at which both ϕ+
2 = ϕ−
2 .
By Lemma 3.1(viii), E contains all but countably many points of (β, α) and ϕ+
1 ,
ϕ−
1 , ϕ+
2 are continuous there. Fix s, t ∈ E and let ¯t ∈ (β, α). The cosecant
function decreases on (0, π/2) and increases on (π/2, π). So if y lies between ψ1(t, s)
and ψ1(t, ¯t) we have
2 , and ϕ−
1 and ϕ+
1 = ϕ−
csc2 y ≤ max(csc2 ψ1(t, ¯t), csc2 ψ1(t, s)) ≤ M.
The mean value theorem implies that,
cot ψ1(t, s) − cot ψ1(t, ¯t) ≤ Mψ1(t, ¯t) − ψ1(t, s).
Letting ¯t → t, we have
(4.1)
cot ψ1(t, s) − cot ϕ+
1 (t) ≤ M (ϕ+
1 (t) − ψ1(t, s).
On the other hand, the function csc2 is bounded below by 1 so
(4.2)
cot ψ2(t, s) − cot ϕ+
2 (t) ≥ ϕ+
2 (t) − ψ2(t, s).
By Lemma 3.1(iii) the functions ψ1 and ψ2 each take values in a compact subset
of (0, π). So do their limits, ϕ+
2 . But the cotangent function is unbounded
near 0. Therefore, for each ε > 0 there exists a δ > 0 such that if 0 < t − s < δ
then
1 and ϕ+
2 cot(t − s) + cot ψ2(t, s) + cot ϕ+
2 cot(t − s) + cot ψ1(t, s) + cot ϕ+
2 (t)
1 (t) ≤ 1 + ε.
Since t ∈ E, the angular equivalence of k · k1 and k · k2, in the form (3.1), becomes
cot ψ2(t, s) − cot ϕ+
2 (t)
2 (t) ≤ C2
2 cot(t − s) + cot ψ2(α, β) + cot ϕ+
Note that both sides of the inequality are non-negative. Combining it with the
previous estimate gives
2 cot(t − s) + cot ψ1(t, s) + cot ϕ+
1 (t)
cot ψ1(t, s) − cot ϕ+
1 (t)
.
(4.3)
cot ψ2(t, s) − cot ϕ+
2 (t) ≤ (1 + ε)C2 cot ψ1(t, s) − cot ϕ+
1 (t)
whenever 0 < t − s < δ.
16
EDER KIKIANTY AND GORD SINNAMON
Now suppose that s, t ∈ E and s < t. By Lemma 3.1(i), t−s = ψj(s, t)−ψj (t, s),
for j = 1, 2, so (4.3) combines with (4.2) and (4.1) to give
2 (s))
2 (t)) − (s + ϕ+
2 (t) − ψ2(t, s) + ψ2(s, t) − ϕ+
2 (t) + cot ϕ+
(t + ϕ+
≤ ϕ+
≤ cot ψ2(t, s) − cot ϕ+
≤ (1 + ε)C2( cot ψ1(t, s) − cot ϕ+
≤ (1 + ε)M C2(ϕ+
= (1 + ε)M C2(ϕ+
= (1 + ε)M C2((t + ϕ+
1 (t)) − (s + ϕ+
1 (t) + cot ϕ+
1 (s) − cot ψ1(s, t))
1 (t) − ψ1(t, s) + ψ1(s, t) − ϕ+
1 (s))
1 (t) − ψ1(t, s) + ψ1(s, t) − ϕ+
1 (s))
1 (s))).
2 (s)
2 (s) − cot ψ2(s, t)
Removal of the absolute value signs to get the second-last line is justified by Lemma
3.1(ii). Now suppose β < ¯β < ¯α < α with ¯α, ¯β ∈ E. Since E is dense (β, α) we can
choose t0 = ¯β < t1 < ··· < tn = ¯α with tk − tk−1 < δ and tk ∈ E for each k. Then
(¯α + ϕ+
n
2 (tk−1))
2 (¯α)) − ( ¯β + ϕ+
2 ( ¯β)) =
(tk + ϕ+
Xk=1
≤ (1 + ε)M C2
= (1 + ε)M C2((¯α + ϕ+
2 (tk)) − (tk−1 + ϕ+
Xk=1
(tk + ϕ+
n
1 (tk)) − (tk−1 + ϕ+
1 ( ¯β))).
1 (¯α)) − ( ¯β + ϕ+
1 (tk−1))
Letting ¯β decrease to β through E, ¯α increase to α through E, and ε → 0 we have
(α + ϕ−
2 (α)) − (β + ϕ+
2 (β)) ≤ M C2((α + ϕ−
2 (α)) − (β + ϕ+
2 (β)).
(cid:3)
Somewhat surprisingly, the next result, initially expected to be easy, turned out
to be the main result of the current article because of its involved proof. It is hoped
that greater familiarity with angular equivalence will reveal a simpler one.
Theorem 4.2. Let X be a real vector space with norms k · k1 and k · k2. If k · k1
and k · k2 are angularly equivalent then they are also topologically equivalent.
Proof. Let C ≥ 1 be the constant in the definition of angular equivalence, that is,
suppose that for all non-zero x, y ∈ X,
tan(θ2(x, y)/2) ≤ C tan(θ1(x, y)/2).
We will show that for all non-zero ¯x, ¯y ∈ X,
(4.4)
k¯xk1k¯yk2
k¯xk2k¯yk1 ≤ 40C2.
Topological equivalence follows from this by fixing any non-zero ¯y ∈ X to get
mk¯xk1 ≤ k¯xk2 with m = k¯yk2/(40C2kyk1) and, interchanging ¯x and ¯y in (4.4),
k¯xk2 ≤ Mk¯xk1 with M = 40C2k¯yk2/k¯yk1.
If ¯x and ¯y are multiples of one another then (4.4) holds trivially so we assume
henceforth that they are independent. Let Z be the two-dimensional subspace of
X spanned by ¯x and ¯y. Choose vectors x, y ∈ Z satisfying three conditions:
(4.5)
kxk2 = kyk2 = 1; kyk1 ≤ kzk1
kzk2 ≤ kxk1 for 0 6= z ∈ Z; kxk1 ≤ kx+tyk1 for t ≥ 0.
ANGULAR EQUIVALENCE OF NORMED SPACES
17
To see that this is possible, take x and y to give the maximum and minimum,
respectively, of the continuous function z 7→ kzk1 on the compact set {z ∈ Z :
kzk2 = 1}. For any non-zero z ∈ Z, z/kzk2 is in this set so the first two conditions
hold. But they also hold with y replaced by −y. The third condition must hold
either with y or with −y; otherwise, there would exist s, t > 0 such that kx+ tyk1 <
kxk1 and kx − syk1 < kxk1, which leads to the contradiction,
s+tkx + tyk1 + t
s+tkx − syk1 < kxk1.
s+t (x − sy)k1 ≤ s
kxk1 = k s
s+t (x + ty) + t
With x and y now fixed, we identify Z with R2 by the linear map (ξ, η) 7→ ξx+ηy
so that k(ξ, η)k1 = kξx + ηyk1 and k(ξ, η)k2 = kξx + ηyk2 give two norms on R2.
These make the identification isometric in both norms so g-functional calculations
are not affected. In particular, the two norms on R2 are angularly equivalent with
constant C. Define rj, ψj and ϕ±
j as above for j = 1, 2. Note that the polar graph
of rj coincides with the boundary of the k · kj-unit ball-we refer to it simply as
the rj-curve.
The first condition of (4.5) ensures that r2(0) = 1 and r2(π/2) = 1. For conve-
nience we set a = r1(0) and b = r1(π/2). The second condition implies that,
k¯xk1k¯yk2
k¯xk2k¯yk1 ≤ kxk1
kyk1
=
b
a
,
so we may complete the proof of (4.4) and the theorem by showing b/a ≤ 40C2.
Suppose instead that b/a > 40C2. First, we set up to apply Lemma 4.1 with
α = tan−1 3 and β = 0. Estimates will be needed for α + ϕ−
1 (0), α +
ϕ−
2 (α) − ϕ+
2 (0), and M = sup{csc2(ψ1(s, t)) : s, t ∈ (0, α), s 6= t}.
For the remainder of the proof we will frequently apply the definitions of ψ and
ϕ± and the properties in Lemma 3.1(i) and (ii) without explicit reference. Also, as
above, "angle" refers to the usual angle in R2, not the norm angle.
1 (α) − ϕ+
The third condition of (4.5) shows that, in the first quadrant, the r1-curve lies
to the left of the line x = a. Therefore ψ1(0, t) ≥ π/2 for t > 0 and, in the limit,
ϕ+
1 (0) ≥ π/2. Also, the r1-curve intersects the line y = 3x on the segment from
(0, 0) to (a, 3a). Therefore ψ1(π/2, α) is less than or equal to the measure of the
angle from (0, 0) to (0, b) to (a, 3a). That is,
(4.6)
ψ1(π/2, α) ≤ cot−1(b/a − 3).
Note that since C ≥ 1 our assumption that b/a > 40C2 implies b/a > 3. Now,
(4.7) α + ϕ−
1 (0) ≤ α + ψ1(α, π/2) − π/2 = ψ1(π/2, α) ≤ cot−1(b/a − 3).
1 (α) − ϕ+
To estimate ϕ+
2 (0) we use the angular equivalence hypothesis applied to the two
norm angles from (1, 0) to (cos(−t), sin(−t)), for t > 0. In the form (3.1), it implies
cot ψ2(0,−t) − A2
2 cot t + cot ψ2(0,−t) + A2 ≤ C2
cot ψ1(0,−t) − A1
2 cot t + cot ψ1(0,−t) + A1
.
18
EDER KIKIANTY AND GORD SINNAMON
where Aj = 1
ψ2 are uniformly bounded away from zero, but cot t → ∞ as t → 0+. Thus,
j (0)), j = 1, 2. By Lemma 3.1(iii), A1, A2, ψ1 and
j (0) + cot ϕ+
2 (cot ϕ−
cot ϕ−
2 (0) − cot ϕ+
2 (0)
= 2 lim
t→0+
t→0+
≤ 2 lim
= C2(cot ϕ−
cot ψ2(0,−t) − A2
2 cot t + cot ψ2(0,−t) + A2
cot ψ1(0,−t) − A1
C2
2 cot t + cot ψ1(0,−t) + A1
1 (0) − cot ϕ+
1 (0)).
(2 cot t + cot ψ2(0,−t) + A2)
(2 cot t + cot ψ2(0,−t) + A2)
Returning to the second condition of (4.5) we see that for each t,
r1(t)
r2(t)
= k(cos t)x + (sin t)yk2
k(cos t)x + (sin t)yk1 ≥
1
kxk1
= a.
Recalling that r2(0) = 1 and applying Lemma 3.1(v), we get
t→0−
0 ≤ lim
Thus, cot ϕ−
1
t→0−
r1(t)
r2(t)(cid:19) = lim
t (cid:18)a −
2 (0) ≥ cot ϕ−
cot ϕ+
2 (0) ≥ cot ϕ−
1 (0) and we have
a
r2(t) − 1
tr2(t) −
r1(t) − a
tr2(t)
= a cot ϕ−
2 (0)− a cot ϕ−
1 (0).
2 (0) − C2(cot ϕ−
1 (0) − cot ϕ+
1 (0)
1 (0))
1 (0) + C2 cot ϕ+
≥ (1 − C2) cot ϕ−
≥ (1 − C2) cot ψ1(0,−π/2) + C2 cot ψ1(0, π/2)
= (1 − C2)(a/b) + C2(−a/b)
≥ −2C2a/b.
We conclude that
ϕ+
2 (0) ≤ cot−1(−2C2a/b) = π − cot−1(2C2a/b).
The r2-curve passes through (−1, 0) and (0, 1) so, in the first quadrant, it lies
below the line y = x + 1. Thus, the r2-curve intersects the line y = 3x on the
segment from (0, 0) to (1/2, 3/2). It follows that ψ2(α, 0) is greater than or equal to
the measure of the apex angle of the isosceles triangle with vertices (0, 0), (1/2, 3/2),
and (1, 0). The latter is π − 2α so we have
(4.8)
α + ϕ−
2 (α) − ϕ+
2 (0) ≥ α + ψ2(α, 0) − ϕ+
2 (0) ≥ cot−1(2C2a/b) − α.
Finally we estimate M . Since b/a > 40C2 > 3 + 1/3, cot−1(b/a − 3) <
cot−1(1/3) = α so (4.6) implies
ψ1(α, π/2) = ψ1(π/2, α) − α + π/2 ≤ cot−1(b/a − 3) − α + π/2 ≤ π/2.
Thus, for s, t ∈ [0, α) with s 6= t,
ψ1(t, s) ≤ ψ1(t, α) = ψ1(α, t) + α − t ≤ ψ1(α, t) + α ≤ ψ1(α, π/2) + α ≤ π/2 + α.
Combining this with,
ψ1(t, s) ≥ ψ1(t, 0) = ψ1(0, t) − t ≥ ϕ+
gives π/2 − α ≤ ψ1(t, s) ≤ π/2 + α and hence
1 (0) − t ≥ π/2 − t ≥ π/2 − α
csc2 ψ1(t, s) ≤ csc2(π/2 − α) = 10.
ANGULAR EQUIVALENCE OF NORMED SPACES
19
So M ≤ 10. Now Lemma 4.1 gives,
Using (4.7) and (4.8), we have
α + ϕ−
2 (α) − ϕ+
2 (0) ≤ 10C2(α + ϕ−
1 (α) − ϕ+
1 (0)).
cot−1(2C2a/b) − α ≤ 10C2 cot−1(b/a − 3).
Since cot−1 u ≤ 1/u when u > 2/π, and C ≥ 1, our assumption b/a > 40C2 yields,
10C2 cot−1(b/a − 3) ≤ 10C2 cot−1(40C2 − 3C2) ≤ 10/37 < 0.271
and
cot−1(2C2a/b) − α ≥ cot−1(2/40) − tan−1 3 > 0.271.
This contradiction completes the proof.
(cid:3)
5. Further Work
There are many natural, fundamental questions about angular equivalence still
to be investigated. We list a few in the hope that interested readers will contribute
to the theory. Throughout, k·k1 and k·k2 are angularly equivalent norms on a real
vector space X. By Theorem 4.2 the two norms are also topologically equivalent.
Since the completions of k · k1 and k · k2 are norms on a common vector space it
makes sense to ask,
Question 1. Are the completions of angularly equivalent norms again angularly
equivalent?
If Y is a subspace of X then it is closed with respect to k · k1 if and only if it is
closed with respect to k·k2. Each of the two norms give rise to a quotient norm on
X/Y defined by, kx + Y kj = inf y∈Y kx + ykj, j = 1, 2.
Question 2. Do angularly equivalent norms induce angularly equivalent norms on
quotient spaces?
The vector space X ∗, of continuous linear functionals on X, is the same for
both norms. Thus, their respective dual norms k · k∗
2 are norms on a
common vector space. These dual norms are not, in general, angularly equivalent.
See Example 3, below. So it is natural to ask,
1 and k · k∗
Question 3. Under what conditions on angularly equivalent norms are their dual
norms also angularly equivalent?
Example 3. The dual norms of angularly equivalent norms need not be angu-
larly equivalent: Consider the two weighted ℓ1 norms, k(ξ, η)k1 = 2ξ + η and
k(ξ, η)k2 = ξ + 2η on R2. A calculation shows that, with θj denoting the k · kj-
norm angle from (ξ, η) to (µ, ν),
tan2(θ2/2) = µ − µ sgn ξ + 2(ν − ν sgn η)
µ + µ sgn ξ + 2(ν + ν sgn η)
2(µ − µ sgn ξ) + ν − ν sgn η
2(µ + µ sgn ξ) + ν + ν sgn η
≤ 4
= 4 tan2(θ1/2).
(The operator sgn takes the value 1, −1 or 0 when its argument is positive, negative
or zero, respectively.) Thus, the two norms are angularly equivalent. However, their
dual norms are given by k(ξ, η)k∗
2 = max(ξ,η/2),
respectively. These have polygonal unit balls in which the vertices are not on the
1 = max(ξ/2,η) and k(ξ, η)k∗
20
EDER KIKIANTY AND GORD SINNAMON
same rays. One unit ball has vertices at (±2,±1) and the other has vertices at
(±1,±2). By Corollary 2.3 the dual norms are not angularly equivalent.
The Cartesian product and tensor product of normed vector spaces can be given
various different, but topologically equivalent, norms. This leads to the question,
Question 4. How should the norm of a Cartesian or tensor product be defined
to ensure that product norms are angularly equivalent whenever the norms on the
factors are angularly equivalent?
An Orlicz space can be equipped with either of two topologically equivalent
norms, the Luxemburg norm or the Orlicz norm. Both are needed because each
arises naturally from the other when considering the norm on the dual space. For
the special case of Lp spaces, these two norms coincide (up to a constant multiple)
and hence are angularly equivalent. However, the examples following [3, Theorem
10] show that there are Orlicz spaces which are strictly convex when equipped with
the Luxemburg norm but are not strictly convex when equipped with the Orlicz
norm. By Corollary 2.2 the two norms are not angularly equivalent for such spaces.
So we ask,
Question 5. For which Orlicz spaces are the Luxemburg norm and the Orlicz norm
angularly equivalent?
A similar question could be posed for many other families. For example, topo-
logically equivalent norms abound on the much-studied families of Banach spaces
defined by Hardy, Sobolev, Besov, Triebel, Lizorkin and others.
References
[1] F. L. Bauer and A. S. Householder 'Some inequalities involving the euclidean condition of a
matrix', Numer. Math. 2 (1960), 308–311.
[2] S. Dragomir, Semi-Inner Products and Applications (Nova Science Publishers, Inc., Haup-
pauge, NY, 2004).
[3] A. Kami´nska 'Strict convexity of sequence Orlicz-Musielak spaces with Orlicz norm', J. Funct.
Anal. 50 (1983) 285–305.
[4] M. Lin and G. Sinnamon 'The generalized Wielandt inequality in inner product spaces',
Eurasian Math. J. 3 (2012) 72–85.
[5] P. Milici´c 'Sur le semi-produit scalaire dans quelques espaces vectorial norm´es', Mat. Vesnik
8 (1971) 181–185.
[6] P. Milici´c 'Sur le g-angle dans un espace norm´e', Mat. Vesnik 45 (1998) 43–48.
[7] P. Milici´c 'Characterizations of convexities of normed spaces by means of g-angles', Mat. Vesnik
54 (2002) 37–44.
[8] H. Wielandt 'Inclusion theorems for eigenvalues', National Bureau of Standards Appl. Math.
Series 29 (1953), 7–78 .
E. Kikianty, Department of Mathematics and Applied Mathematics, University of
Pretoria, Pretoria, South Africa
E-mail address: [email protected]
Department of Mathematics, University of Western Ontario, London, Canada
E-mail address: [email protected]
|
1706.10110 | 2 | 1706 | 2017-11-08T08:23:55 | On Using Toeplitz and Circulant Matrices for Johnson-Lindenstrauss Transforms | [
"math.FA",
"cs.CC",
"cs.DS"
] | The Johnson-Lindenstrauss lemma is one of the corner stone results in dimensionality reduction. It says that given $N$, for any set of $N$ vectors $X \subset \mathbb{R}^n$, there exists a mapping $f : X \to \mathbb{R}^m$ such that $f(X)$ preserves all pairwise distances between vectors in $X$ to within $(1 \pm \varepsilon)$ if $m = O(\varepsilon^{-2} \lg N)$. Much effort has gone into developing fast embedding algorithms, with the Fast Johnson-Lindenstrauss transform of Ailon and Chazelle being one of the most well-known techniques. The current fastest algorithm that yields the optimal $m = O(\varepsilon^{-2}\lg N)$ dimensions has an embedding time of $O(n \lg n + \varepsilon^{-2} \lg^3 N)$. An exciting approach towards improving this, due to Hinrichs and Vyb\'iral, is to use a random $m \times n$ Toeplitz matrix for the embedding. Using Fast Fourier Transform, the embedding of a vector can then be computed in $O(n \lg m)$ time. The big question is of course whether $m = O(\varepsilon^{-2} \lg N)$ dimensions suffice for this technique. If so, this would end a decades long quest to obtain faster and faster Johnson-Lindenstrauss transforms. The current best analysis of the embedding of Hinrichs and Vyb\'iral shows that $m = O(\varepsilon^{-2}\lg^2 N)$ dimensions suffices. The main result of this paper, is a proof that this analysis unfortunately cannot be tightened any further, i.e., there exists a set of $N$ vectors requiring $m = \Omega(\varepsilon^{-2} \lg^2 N)$ for the Toeplitz approach to work. | math.FA | math | On Using Toeplitz and Circulant Matrices for
Johnson-Lindenstrauss Transforms
Casper Benjamin Freksen∗
Kasper Green Larsen∗
Abstract
The Johnson-Lindenstrauss lemma is one of the corner stone results in dimensionality reduction.
It says that given N , for any set of N vectors X ⊂ Rn, there exists a mapping f : X → Rm
such that f (X) preserves all pairwise distances between vectors in X to within (1 ± ε) if m =
O(ε−2 lg N ). Much effort has gone into developing fast embedding algorithms, with the Fast Johnson-
Lindenstrauss transform of Ailon and Chazelle being one of the most well-known techniques. The
current fastest algorithm that yields the optimal m = O(ε−2 lg N ) dimensions has an embedding
time of O(n lg n + ε−2 lg3 N ). An exciting approach towards improving this, due to Hinrichs and
Vybíral, is to use a random m × n Toeplitz matrix for the embedding. Using Fast Fourier Transform,
the embedding of a vector can then be computed in O(n lg m) time. The big question is of course
whether m = O(ε−2 lg N ) dimensions suffice for this technique. If so, this would end a decades long
quest to obtain faster and faster Johnson-Lindenstrauss transforms. The current best analysis of
the embedding of Hinrichs and Vybíral shows that m = O(ε−2 lg2 N ) dimensions suffices. The main
result of this paper, is a proof that this analysis unfortunately cannot be tightened any further, i.e.,
there exists a set of N vectors requiring m = Ω(ε−2 lg2 N ) for the Toeplitz approach to work.
1
Introduction
The performance of many geometric algorithms depends heavily on the dimension of the input data. A
widely used technique to combat this "curse of dimensionality", is to preprocess the input via dimension-
ality reduction while approximately preserving important geometric properties. Running the algorithm
on the lower dimensional data then uses less resources (time, space, etc.) and an approximate result for
the high dimensional data can be derived from the low dimensional result.
Dimensionality reduction approximately preserving pairwise Euclidean distances has found uses in a
wide variety of applications, including: Nearest-neighbour search [2, 12], clustering [6, 8], linear program-
ming [22], streaming algorithms [19], compressed sensing [7, 11], numerical linear algebra [25], graph
sparsification [20], and differential privacy [5]. See more applications in [21, 14]. The most fundamental
result in this regime is the Johnson-Lindenstrauss (JL) lemma [15], which says the following:
Theorem 1 (Johnson-Lindenstrauss lemma). Let X ⊂ Rn be a set of N vectors, then for any 0 < ε <
1/2, there exists a map f : X → Rm for some m = O(ε−2 lg N ) such that
∀x, y ∈ X, (1 − ε)kx − yk2
2 ≤ kf (x) − f (y)k2
2 ≤ (1 + ε)kx − yk2
2.
7
1
0
2
v
o
N
8
]
.
A
F
h
t
a
m
[
2
v
0
1
1
0
1
.
6
0
7
1
:
v
i
X
r
a
This result dates back to 1984 and says that to preserve pairwise Euclidean distances amongst a set
of N points/vectors in Rn to within a factor (1 ± ε), it suffices to use just m = O(ε−2 lg N ) dimensions.
The bound on m was very recently proven optimal [18].
The standard technique for constructing a map with the properties of Theorem 1 is the following:
Let A be an m × n matrix with entries independently sampled as either N (0, 1) random variables (as
in [10]) or Rademacher (uniform among {−1, +1}) random variables (as in [1]). Once such entries have
been drawn, let f : Rn → Rm be defined as:
f (x) =
Ax.
1
√m
∗This research is supported by a Villum Young Investigator Grant, an AUFF Starting Grant and MADALGO, Center
for Massive Data Algorithmics, a Center of the Danish National Research Foundation, grant DNRF84. Department of
Computer Science, Aarhus University. {cfreksen,larsen}@cs.au.dk.
1
To prove that the map f satisfies the guarantees in Theorem 1, it is first shown that for any vector x,
the probability that kf (x)k2
2 is less than 1/N 2. This probability is called the
error probability and denoted δ. Using linearity of f and a union bound over all pairs x, y ∈ X, the
probability that all pairwise distances (i.e. the norm of the vector x − y) are preserved can be shown to
be at least 1/2.
2 is not within (1 ± ε)kxk2
1.1 Time Complexity
Examining the classic Johnson-Lindenstrauss reduction above, we see that to embed a vector, we need to
multiply with a dense matrix and the embedding time becomes O(nm) (or equivalently O(nε−2 lg N )).
This may be prohibitively large for many applications (recall one prime usage of dimensionality reduction
is to speed up algorithms), and much research has been devoted to obtaining faster embedding time.
Fast Johnson-Lindenstrauss Transform. Ailon and Chazelle [2] were the first to address the ques-
tion of faster Johnson-Lindenstrauss transforms. In their seminal paper, they introduced the so-called
Fast Johnson-Lindenstrauss transform for speeding up dimensionality reduction. The basic idea in their
paper is to first "precondition" the input data by multiplying with a diagonal matrix with random
signs, followed by multiplying with a Hadamard matrix. This has the effect of "spreading" out the
mass of the input vectors, allowing for the dense matrix A above to be replaced with a sparse matrix.
Since we can multiply with a Hadamard matrix using Fast Fourier Transform, this gives an embed-
ding time of O(n lg n + ε−2 lg3 N ) for embedding into the optimal m = O(ε−2 lg N ) dimensions. For
m = ε−2 lg N ≤ n1/2−γ for any constant γ > 0, the embedding complexity was improved even further
down to O(n lg m) in [3].
Another approach to achieve the O(n lg m) embedding time, but without the restriction on ε−2 lg N ≤
n1/2−γ, is to sacrifice the target dimension. This was done in [4] and later improved in [17], where the
embedding complexity was O(n lg m) at the cost of an increased target dimension m = O(ε−2 lg N lg4 n).
Sparse Vectors. Another approach to improve the performance of JL transforms, is to assume the
input data is sparse, i.e. has few non-zero coordinates. Designing an algorithm based on the work in [24],
Dasgupta et al. [9] achieved an embedding complexity of O(kxk0ε−1 lg2(mN ) lg N ), where kxk0 = {i
xi 6= 0}. This was later improved to O(kxk0ε−1 lg N ) in [16].
Toeplitz Matrices. Finally, another very exciting approach is to use Toeplitz matrices or partial
circulant matrices for the embedding. We first introduce the terminology.
An m × n Toeplitz matrix is an m × n matrix, where every entry on a diagonal has the same value:
t0
t−1
t−2
...
t1
t0
t−1
...
t2
t1
t0
...
t−(m−1)
t−(m−2)
t−(m−3)
···
···
···
. . .
···
tn−1
tn−2
tn−3
...
tn−m
A partial circulant matrix is a special kind of Toeplitz matrix, where every row, except the first, is the
previous row rotated once:
t0
tn−1
tn−2
...
t1
t0
tn−1
...
t2
t1
t0
...
tn−(m−1)
tn−(m−2)
tn−(m−3)
···
···
···
. . .
···
tn−1
tn−2
tn−3
...
tn−m
Hinrichs and Vybíral [13] proposed the following algorithm for generating a JL embedding based on a
Toeplitz matrix1: Let t−(m−1), t−(m−2), . . . , tn−1 and d1, . . . , dn be i.i.d. Rademacher random variables,
and T be a Toeplitz matrix defined from t−(m−1), t−(m−2), . . . , tn−1 such that entry (i, j) takes values
1[13] uses a partial circulant matrix but notes that a Toeplitz matrix could be used as well.
2
Table 1: Comparison of the performances of various Johnson-Lindenstrauss transform algorithms. N is
the number of input vectors, n is the dimension of the input vectors, m is the dimension of the output
vectors, ε is the distortion.
Embedding time
Type
Random projection O(nm)
Sparse
Sparse
FFT
FFT
FFT
Toeplitz
Toeplitz
Target dimension (m) Ref. Notes
O(ε−2 lg N )
O(kxk0ε−1 lg2(mN ) lg N ) O(ε−2 lg N )
O(ε−2 lg N )
O(kxk0ε−1 lg N )
O(n lg n + m lg2 N )
O(ε−2 lg N )
O(ε−2 lg N )
O(n lg m)
O(ε−2 lg N lg4 n)
O(n lg m)
O(ε−2 lg3 N )
O(n lg m)
O(ε−2 lg2 N )
O(n lg m)
[10]
[9]
[16]
[2]
[3]
[17]
[13]
[23]
m ≤ n1/2−γ
tj−i for i = 1, . . . , m and j = 1, . . . , n. Let D be an n × n diagonal matrix with the random variable di
giving the i'th diagonal entry. Define the map f as
f (x) =
1
√m
T Dx.
Multiplying with a Toeplitz matrix corresponds to computing a convolution and can be done using
Fast Fourier Transform. By appropriately blocking the input coordinates, the complexity of embedding
a vector x is just O(n lg m) for any target dimension m. The big question is of course, how low can the
target dimension m be, while preserving the distances between vectors up to a factor of 1 ± ε?
In the original paper [13], the authors proved that setting the target dimension to m = O(ε−2 lg3(1/δ)),
the norm of any vector would be preserved to within (1 ± ε) with probability at least 1 − δ. Setting
δ = 1/N 2, a union bound over all pairwise difference vectors (as in the classic construction) shows that
dimension m = O(ε−2 lg3 N ) suffices. Later, the analysis was refined in [23], which lowered the target
dimension to m = O(ε−2 lg2(1/δ)) for preserving norms to within (1 ± ε) with probability 1 − δ. Again,
setting δ = 1/N 2, this gives m = O(ε−2 lg2 N ) target dimension. Now if the analysis could be tightened
even further to give the optimal m = O(ε−2 lg N ) dimensions, this would end the decades long quest for
faster and faster embedding algorithms!
Our Contribution. Our main result unfortunately shows that the analysis of Vybíral [23] cannot be
tightened to give an even lower target dimensionality. More specifically, we prove that the upper bound
given in [23] is optimal:
Theorem 2. Let T and D be the m × n Toeplitz and n × n diagonal matrix in the embedding proposed
by [13]. For all 0 < ε < C, where C is a universal constant, and any desired error probability δ > 0, if
the following holds for every unit vector x ∈ Rn:
then it must be the case that m = Ω(ε−2 lg2(1/δ)).
Pr"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
< ε# > 1 − δ,
While Theorem 2 already shows that one cannot tighten the analysis of Vybíral for preserving the
norm of just one vector, Theorem 2 does leave open the possibility that one would not need to union
bound over all N 2 pairs of difference vectors when trying to preserve all pairwise distances amongst a
set of N vectors. It could still be the case that there somehow was a strong positive correlation between
distances being preserved (though this seems extremely unlikely, and would be something not seen in
any previous approach to JL). To complete the picture, we indeed show that this is not the case, at least
for N somewhat smaller than the dimension n:
Theorem 3. Let T and D be the m × n Toeplitz and n × n diagonal matrix in the embedding proposed
by [13]. For all 0 < ε < C, where C is a universal constant, if the following holds for every set of N
3
vectors X ⊂ Rn:
Pr"∀x, y ∈ X :(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
T Dx −
1
√m
2 − kx − yk2
≤ εkx − yk2
2# = Ω(1),
2
T Dy(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
then it must be the case that either m = Ω(ε−2 lg2 N ) or m = Ω(n/N ).
We remark that our proofs also work if we replace T be a partial circulant matrix (which was also
proposed in [13]). Furthermore, we expect that minor technical manipulations to our proof would also
show the above theorems when the entries of T and D are N (0, 1) distributed rather than Rademacher
(this was also proposed in [13]).
2 Lower Bound for One Vector
Let T be m× n Toeplitz matrix defined from random variables t−(m−1), t−(m−2), . . . , tn−1 such that entry
(i, j) takes values tj−i for i = 1, . . . , m and j = 1, . . . , n. Let D be an n × n diagonal matrix with the
random variable di giving the i'th diagonal entry. This section shows the following:
Theorem 4. Let T be m× n Toeplitz and D n× n diagonal. If t−(m−1), t−(m−2), . . . , tn−1 and d1, . . . , dn
are independently distributed Rademacher random variables for i = −(m− 1), . . . , n− 1 and j = 1, . . . , n,
then for all 0 < ε < C, where C is a universal constant, there exists a unit vector x ∈ Rn such that
and furthermore, all but the first O(√m) coordinates of x are 0.
Pr"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
> ε# ≥ 2−O(ε√m).
It follows from Theorem 4 that if we want to have probability at least 1− δ of preserving the norm of
any unit vector x to within (1± ε), it must be the case that ε√m = Ω(lg(1/δ)), i.e. m = Ω(ε−2 lg2(1/δ)).
This is precisely the statement of Theorem 2. Thus we set out to prove Theorem 4.
To prove Theorem 4, we wish to invoke the Paley-Zygmund inequality, which states, that if X is a
non-negative random variable with finite variance and 0 ≤ θ ≤ 1, then
Pr[X > θE[X]] ≥ (1 − θ)2
E2[X]
E[X 2]
.
We carefully choose a unit vector x, and define the random variable for Paley-Zygmund to be the
k'th moment of the difference between the norm of x transformed and 1.
Proof. Let k be an even positive integer less than m/4 and define s := 4k. Note that s ≤ m. Let x be
an arbitrary n-dimensional unit vector such that the first s coordinates are in {−1/√s, +1/√s}, while
the remaining n − s coordinates are 0. Define the random variable parameterized by k
Zk := (cid:13)(cid:13)(cid:13)(cid:13)
1
√m
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
2
2 − 1!k
.
Since k is even, the random variable Zk is non-negative.
We wish to lower bound E[Zk] and upper bound E[Z 2
bounds we prove are as follows:
Lemma 5. If k ≤ √m, then the random variable Zk satisfies:
and
E[Zk] ≥ m−k/2kk2−O(k)
k ] ≤ m−kk2k2O(k).
E[Z 2
4
k ] in order to invoke Paley-Zygmund. The
Before proving Lemma 5 we show how to use it together with Paley-Zygmund to complete the proof
of Theorem 4.
We start by invoking Paley-Zygmund and then rewriting the expectations according to Lemma 5,
E2[Zk]
E[Z 2
k]
E2[Zk]
E[Z 2
k]
=⇒
=⇒
Pr[Zk > E[Zk]/2] ≥ (1/4)
Pr[Z 1/k
k > (E[Zk]/2)1/k] ≥ (1/4)
C0√m# ≥ 2−O(k).
>
k
1
√m
Pr"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
Here C0 is some constant greater than 0. For any 0 < ε < 1/C0, we can now set k such that k/(C0√m) =
ε, i.e. we choose k = εC0√m. This choice of k satisfies k ≤ √m as required by Lemma 5. We have thus
shown that:
Pr"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
> ε# ≥ 2−O(ε√m).
Remark. Theorem 4 can easily be extended to partial circulant matrices. The difference between partial
circulant and Toeplitz matrices is the dependence between the values in the first m and last m columns.
However, as only the first s = 4k ≤ 4√m entries in x are nonzero, the last m columns are ignored, and
so partial circulant and Toeplitz matrices behave identically in our proof.
Proof of Lemma 5. Before we prove the two bounds in Lemma 5 individually, we rewrite E[Zk], as this
benefits both proofs.
1
m
E[Zk] = E
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
= E
= E
= E
= E
1
m
1
m
1
m
m
Xi=1
m
Xi=1
=
1
mk
2
n
n
n
m
m
j x2
Xj=1
t2
j−id2
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
Xj=1
Xi=1
Xj=1
Xj=1 Xh∈{1,...,n}\{j}
2 − 1!k
k
2
− 1
tj−idjxj
j
+
Xj=1 Xh∈{1,...,n}\{j}
j
+
Xj=1 Xh∈{1,...,n}\{j}
j − x2
k
tj−ith−idj dhxjxh
tj−ith−idjdhxjxh
E
Y(i,j,h)∈S
t2
j−id2
X
j x2
n
n
n
S∈([m]×[n]×[n])k∀(i,j,h)∈S:h6=j
k
tj−ith−idj dhxjxh
− 1
k
tj−ith−idjdhxj xh
Observe that for j > s or h > s the product becomes 0, as either xj or xh is 0. By removing all these
terms, we simplify the sum to
E[Zk] =
1
mk
S∈([m]×[s]×[s])k∀(i,j,h)∈S:h6=j
X
E
Y(i,j,h)∈S
tj−ith−idjdhxj xh
5
following two things are true:
Observe for an S ∈ ([m] × [s] × [s])k, that the value EhQ(i,j,h)∈S tj−ith−idj dhxjxhi is 0 if one of the
• A dj occurs an odd number of times in the product.
• A variable ta occurs an odd number of times in the product.
To see this, note that by the independence of the random variables, we can write the expectation of the
product, as a product of expectations where each term in the product has all the occurrences of the same
random variable. Since the dj 's and ta's are Rademachers, the expectation of any odd power of one of
these random variables is 0. Thus if just a single random variable amongst the dj 's and ta's occurs an
odd number of times, we have EhQ(i,j,h)∈S tj−ith−idjdhxjxhi = 0. Similarly, we observe that if every
random variable occurs an even number of times, then the expectation of the product is exactly 1/sk
since each xj also occurs an even number of times. If Γk denotes the number of tuples S ∈ ([m]×[s]×[s])k
such that ∀(i, j, h) ∈ S we have h 6= j and furthermore:
• For all columns a ∈ [s], {(i, j, h) ∈ S j = a ∨ h = a} mod 2 = 0.
• For all diagonals a ∈ {−(m − 1), . . . , s − 1}, {(i, j, h) ∈ S j − i = a ∨ h − i = a} mod 2 = 0.
Then we conclude
Note that Z 2
k = Z2k. Therefore,
E[Zk] =
Γk
skmk .
E[Z 2
k ] = E[Z2k] =
Γ2k
s2km2k .
(1)
(2)
To complete the proof of Lemma 5 we need lower and upper bounds for Γk and Γ2k. The bounds we
prove are
Lemma 6. If k ≤ √m, then Γk and Γ2k satisfy:
and
Γk = mk/2skkk2−O(k)
Γ2k = mks2kk2k2O(k).
The proofs of the two bounds in Lemma 6 are in Sections 2.1 and 2.2.
Substituting the bounds from Lemma 6 in (1) and (2) we get
which are the bounds we sought for Lemma 5.
E[Zk] = m−k/2kk2−O(k)
E[Z 2
k ] = m−kk2k2−O(k),
2.1 Lower Bounding Γk
We first recall that the definition of Γk is the number of tuples S ∈ ([m] × [s] × [s])k satisfying that
∀(i, j, h) ∈ S we have h 6= j and furthermore:
• For all columns a ∈ [s], {(i, j, h) ∈ S j = a ∨ h = a} mod 2 = 0.
• For all diagonals a ∈ {−(m − 1), . . . , s − 1}, {(i, j, h) ∈ S j − i = a ∨ h − i = a} mod 2 = 0.
We view a triple (i, j, h) ∈ ([m] × [s] × [s]) as two entries (i, j) and (i, h) in an m × s matrix.
Furthermore, when we say that a triple touches a column or diagonal, a matrix entry of the triple lie on
that column or diagonal, so (i, j, h) touches columns j and h and diagonals j − i and h− i. Similarly, we
say that a tuple S ∈ ([m]× [s]× [s])k touches a given column or diagonal l times, if l triples in S touches
that column or diagonal.
We intent to prove a lower bound for Γk by constructing a big family of tuples F ⊆ ([m] × [s] × [s])k,
where each tuple satisfies, that each column and diagonal touched by that tuple is touched exactly twice.
As each column and diagonal is touched an even number of times, the number of tuples in the family is
a lower bound for Γk.
6
Proof of Γk = mk/2skkk2−O(k). We describe how to construct a family of tuples F ⊆ ([m] × [s] × [s])k
satisfying that ∀S ∈ F,∀(i, j, h) ∈ S we have h 6= j and furthermore:
• For all columns a ∈ [s], {(i, j, h) ∈ S j = a ∨ h = a} ∈ {0, 2}.
• For all diagonals a ∈ {−(m − 1), . . . , s − 1}, {(i, j, h) ∈ S j − i = a ∨ h − i = a} ∈ {0, 2}.
From this and the definition of Γk it is clear that F ≤ Γk.
When constructing an S ∈ F, we view S as consisting of two halves S1 and S2, such that S1 touches
exactly the same columns and diagonals as S2 and both S1 and S2 touches each column and diagonal at
most once. To capture this, we give the following definition, where S is meant to be the family of such
halves S1 and S2.
Definition 7. Let S be the set of all tuples S ∈ ([m] × [s] × [s])k/2 such that
• ∀(i, j, h) ∈ S, j 6= h
• For all columns a ∈ [s],{(i, j, h) ∈ S j = a ∨ h = a} ≤ 1
• For all diagonals a ∈ {−(m − 1), . . . , s − 1},{(i, j, h) ∈ S j − i = a ∨ h − i = a} ≤ 1
Definition 7 mimics the definition of Γk, and the first item in Definition 7 ensures that the triples in
a tuple in S are of the same form as in Γk. The final two items ensure that each column and diagonal,
respectively, is touched at most once. This is exactly the properties we wanted of S1 and S2 individually.
We can now construct F as all pairs of (half) tuples S1, S2 ∈ S, such that S1 touches exactly the
same columns and diagonals as S2. To capture that S1 and S2 touch the same columns and diagonals,
we introduce the notion of a signature. A signature of Si is the set of columns and diagonals touched by
Si.
To have S1 and S2 touch exactly the same columns and diagonals, it is necessary and sufficient that
they have the same signature.
We introduce the following notation: B denotes the number of signatures with at least one member,
and by enumerating the signatures, we let bi denote the number of (half) tuples in S with signature i.
We recall that a (half) tuple S1 ∈ S touches each column and diagonal at most once, and if S1 and
S2 share the same signature, they touch exactly the same columns and diagonals. Therefore, using ◦ to
mean concatenation, S = S1 ◦ S2 ∈ F, as each column and diagonal touched is touched exactly twice.
Therefore F is a lower bound for Γk. Note that for a given signature i, the number of choices of S1 and
S2 with that signature is b2
i . This gives the following inequality,
Γk ≥ F =
b2
i .
B
Xi=1
We now apply the Cauchy-Schwarz inequality:
b2
i
B
Xi=1
B
Xi=1
12 ≥(cid:0)
B
Xi=1
bi(cid:1)2
=⇒
B
Xi=1
b2
i=1 bi(cid:1)2
i ≥ (cid:0)PB
PB
i=1 12
=⇒ Γk ≥ S2
B
.
(3)
To get a lower bound on S2/B (and in turn Γk), we need a lower bound on S and an upper bound
on B. These bounds are stated in the following lemmas
Lemma 8. S = Ω(mk/2sk2−k).
Lemma 9. B = O(cid:16)(cid:0)m+s
k(cid:1)(cid:17)
k/2(cid:1)sk/2(cid:0)s
Before proving any of these lemmas, we show that they together with (3) give the desired lower bound
on Γk:
Γk =
Ω(mk/2sk2−k)2
O(cid:16)(cid:0)m+s
k(cid:1)(cid:17)
k/2(cid:1)sk/2(cid:0)s
= Ω(cid:18) mks2k2−2k(k/2)k/2kk
(m + s)k/2sk/2sk (cid:19) .
(4)
7
Because s = 4k, we have (k/2)k/2
sk/2 = 2−Θ(k), and because s ≤ m:
we can simplify (4),
mk
(m+s)(k/2) = mk/22−Θ(k). With this
Γk = mk/2skkk2−O(k).
which is the lower bound we sought.
Proof of Lemma 8. Recall that S ⊆ ([m] × [s] × [s])k/2 is the set of (half) tuples that touch each column
and diagonal at most once, and, for each triple (i, j, h) in these (half) tuples, we have j 6= h.
We prove Lemma 8 by analysing how we can create a large number of distinct S ∈ S by choosing the
triples in S iteratively.
For each triple, we choose a row and two distinct entries on this row. We choose the row among any
of the m rows.
However, because S ∈ S, when choosing entries on the row, we cannot choose entries that lie on
columns or diagonals touched by previously chosen triples. Instead we choose the two entries among
any of the other entries. Therefore, whenever we choose a triple, this triple prevents at most four row
entries from being chosen for every subsequent triple, as the two diagonals and two columns touched by
the chosen triple intersect with at most four entries on the rows of the subsequent triples. This leads to
the following recurrence, describing a lower bound for the number of triples
F (r, c, t) =(r · c · (c − 1) · F (r, c − 4, t − 1)
1
if t > 0
otherwise
(5)
where r is the number of rows to choose from, c is the minimum number of choosable entries in any row,
and t is the number of triples left to choose.
Inspecting (5), we can see that F can equivalently be defined as
F (r, c, t) = rt
(c − 4i)(c − 1 − 4i).
t−1
Yi=0
(6)
If t ≤ c
8 then the terms inside the product in (6) are greater than c
2 , so we can bound F from below:
We now insert the values for r, c and t to find a lower bound for S, noting that s = 4k ensures that
t ≤ c
8 :
= rtc2t 1
4t .
c
2(cid:1)2t
F (r, c, t) ≥ rt(cid:0)
2(cid:1) ≥ mk/2sk 1
k
S ≥ F(cid:0)m, s,
4k/2 =⇒ S = Ω(mk/2sk2−k).
Proof of Lemma 9. Recall that for at triple S ∈ S we define the signature as the set of columns and
diagonals touched by S. Furthermore, viewing a triple (i, j, h) ∈ ([m] × [s] × [s]) as the two entries (i, j)
and (i, h) in an m × s matrix, we define the left endpoint as (i, min{j, h} and the right endpoint as
(i, max{j, h}).
The claim to prove is
This is proven by first showing an upper bound on the number of choices for the diagonals of left
endpoints, then diagonals of right endpoints and finally for columns.
B = O(cid:18)(cid:18)m + s
k/2 (cid:19)sk/2(cid:18)s
k(cid:19)(cid:19) .
k/2(cid:1) choices for the diagonals corresponding to left endpoints in a triple.
In an m× s matrix there are m + s different diagonals and as the chosen diagonals have to be distinct,
there are (cid:0)m+s
As the right endpoint of a triple has to be in the same row as the left endpoint, there are at most
s choices for the diagonal corresponding to the right endpoint when the left endpoint has been chosen
(which it has in our case). This gives a total of sk/2 choices for diagonals corresponding to right endpoints.
Finally, there are s columns to choose from and the chosen columns have to be distinct, and so the
total number of choices of columns is (cid:0)s
k(cid:1).
The product of these number of choices gives the upper bound sought.
8
2.2 Upper Bounding Γ2k
Proof. Recall that Γ2k is defined as the number of tuples S ∈ ([m] × [s] × [s])2k such that ∀(i, j, h) ∈ S
we have h 6= j and furthermore:
• For all columns a ∈ [s], {(i, j, h) ∈ S j = a ∨ h = a} mod 2 = 0.
• For all diagonals a ∈ {−(m − 1), . . . , s − 1}, {(i, j, h) ∈ S j − i = a ∨ h − i = a} mod 2 = 0.
Let F ⊆ ([m] × [s] × [s])2k be the family of tuples satisfying these conditions, and so F = Γ2k.
To prove an upper bound on Γ2k, we show how to encode a tuple S ∈ F using at most k lg m +
2k lg s + 2k lg k + O(k) bits, such that S can be decoded from this encoding. Since any S ∈ F can be
encoded using k lg m + 2k lg s + 2k lg k + O(k) bits and F = Γ2k, we can conclude:
Γ2k = 2k lg m+2k lg s+2k lg k+O(k) = mks2kk2k2O(k).
Let σ denote the encoding function and σ−1 denote the decoding function. If S ∈ F and t ∈ S, σ(t)
denotes the encoding of the triple t, σ(S) denotes the encoding of the entire tuple S, and σ(F ) denotes
the image of σ.
A tuple S ∈ F consists of triples t1, t2, . . . , t2k such that S = t1 ◦ t2 ◦ ··· ◦ t2k. To encode S ∈ F we
We will first describe a graph view of a tuple S which will be useful for encoding and decoding, then
encode each of the triples and store them in the same order: σ(S) = σ(t1) ◦ σ(t2) ◦ ··· ◦ σ(t2k).
we will show an encoding algorithm and finally a decoding algorithm.
Graph A tuple S ∈ F forms a (multi)graph structure, where every triple (i, j, h) ∈ S is a vertex. Since
S ∈ F, there lies an even number of triple endpoints on each diagonal. We can thus pair endpoints lying
on the same diagonal, such that every endpoint is paired with exactly one other endpoint. When two
triples have endpoints that are paired, the triples have an edge between them in the graph. As every
triple has two endpoints, every vertex has degree two, and so the graph consists entirely of simple cycles
of length at least two.
Encoding To encode an S ∈ F, we first encode each cycle by itself by defining the σ(t)'s for the triples
t in the cycle. After this, we order the defined σ(t)'s as the t's were ordered in the input.
1. For each cycle we perform the following:
(a) We pick any vertex t = (i, j, h) of the cycle and give it the type head. Define σ(t) as the
concatenation of its type head, its row i, and two columns j and h. This uses lg m+2 lg s+O(1)
bits.
(b) We iterate through the cycle starting after head and give vertices, except the last, type mid.
The last vertex just before head is given the type last.
(c) For each triple t of type mid we store its type and two columns explicitly. However, instead
of storing its row we store the index of its predecessor in the cycle order as well as how
they are connected: If we typed tr just before ts when iterating through the cycle, when
encoding ts we store r as well as whether tr and ts are connected by the left or right endpoint
in tr and left or right endpoint in ts. So define σ(t) as the concatenation of mid, the two
columns, the predecessor index and how it is connected to the predecessor. All in all we spend
lg k + 2 lg s + O(1) bits encoding each mid.
(d) Finally, to encode the triple t, which is typed last, we define σ(t) as the concatenation of its
type, its predecessors index, how it is connected to its predecessor, and the column of the
endpoint on the predecessor's diagonal. We thus spend lg k + lg s + O(1) bits to encode a last.
However, since s = 4k the number of bits per encoded last is equivalent to 2 lg k +O(1), which
turns out to simplify the analysis later.
Note that for each triple, the type is encoded in the first 2 bits2 in the encoding of the triple. This
will be important during decoding.
2To encode the type in 2 bits we could use the following scheme: 00 = head, 01 = mid, 10 = last, and 11 is unused.
9
2. After encoding all cycles, we order the encoded triples in the same order as the triples in the input,
and output the concatenation of the encoded triples: σ(t1) ◦ σ(t2) ◦ ··· ◦ σ(t2k).
To analyse the number of bits needed in total, we look at the average number of bits per triple inside
a cycle. Since all cycles have a length of at least two, each cycle has a head and a last triple. These two
use an average of lg m
2 + lg s + lg k +O(1) bits. We now claim that the number of bits per mid is bounded
by this average. Recall that we assumed √m ≥ k and s = 4k:
√m ≥ k =⇒
lg m
2 ≥ lg k =⇒
lg m
lg m
2
+ lg s + lg k + 2 ≥ lg k + 2 lg s =⇒
2
+ lg s + lg k + O(1) ≥ mid.
2 + lg s + lg k + O(1) bits, the number of bits for all
Since the average of all 2k triples is at most lg m
triples is at most k lg m + 2k lg s + 2k lg k + O(k).
Decoding We now show how to decode any σ(S) ∈ σ(F ) such that σ−1(σ(S)) = S.
we restore the order of triples. The following steps describes this algorithm in more detail.
First we need to extract the encodings of the individual triples, then we decode each cycle, and finally
1. We first extract the individual σ(ti)'s, by iterating through the bit string σ(S). By looking at the
type in the first two bits of an encoded triple we know the length of the encoded triple. From this
we can extract σ(ti) as well as know where σ(ti+1) begins.
2. We now wish to decode each cycle to get the row and two columns of every triple, so do the
following for each head ti:
(a) t ← ti
(b) Look at the first two bits in t to determine its type, and while t is not a last, do:
i. If t is a head, the row and two columns are stored explicitly.
ii. If t is a mid, the two columns are stored explicitly. From the reference to t's predecessor,
we can calculate the diagonal shared between t and its predecessor. This is can be done as
we have stored which endpoint (left or right) of the predecessor lies on the diagonal, and
as we decode in the same order as we encoded, we have already decoded the predecessor.
We know which of t's columns the shared diagonal intersects, and so we can calculate the
row.
If the predecessor is a head, take note of which diagonal is shared with it.
iii. t ← the triple that has t as its predecessor.
(c) t is now a last, so we know the predecessor's index as well as a column. However, as the graph
consists of cycles rather than just linked lists, last shares a diagonal with its predecessor as
well as sharing a diagonal with head. We have noted which diagonal has already been used
for head, so the other diagonal head touches is shared with last.
From these two diagonals and the column, it is possible to calculate the row by the intersection
between the predecessor diagonal and the column, and the other column by the intersection
of the row and the other diagonal.
3. Finally order the triples as they were, when the encoded versions were extracted in step 1.
3 Lower Bound for N Vectors
In this section, we generalize the result of Section 2 to obtain a lower bound for preserving all pairwise
distances amongst a set of N vectors. Our proof uses Theorem 4 as a building block. Recall that
Theorem 4 guarantees that there is a vector x such that
1
√m
Pr"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
10
> ε# ≥ 2−O(ε√m).
Moreover, the vector x has non-zeroes only in the first O(√m) coordinates. From such a vector x, define
x→i as the vector having its j'th coordinate equal to the (j − i)'th coordinate of x if j > i and otherwise
its j'th coordinate is 0. In words, x→i is just x with all coordinates shifted by i.
For i ≤ n − O(√m) (i small enough that all the non-zero coordinates of x stay within the n coordin-
ates), we see that T Dx and T Dx→i have the exact same distribution. Furthermore, if i ≥ m + C√m
for a big enough constant C, T Dx and T Dx→i are independent. To see this, observe that the only
random variables amongst t−(m−1), . . . , tn−1 that are multiplied with a non-zero coordinate of x in T Dx
are t−(m−1), . . . , tO(√m). Similarly, only random variables d1, . . . , dO(√m) amongst d1, . . . , dn are multi-
plied by a non-zero coordinate of x. For x→i, the same is true for variables t−(m−1)+i, . . . , tO(√m)+i and
di+1, . . . , di+O(√m). If i ≥ m + C√m, these sets of variables are disjoint and hence T Dx and T Dx→i are
independent.
With the observations above in mind, we now define a set of vectors X as follows
X := {0, x, x→m+C√m, x→2(m+C√m), . . . , x→⌊(n−C√m)/(m+C√m)⌋(m+C√m)}.
1√m T D as embedding. Furthermore, by the
The 0-vector clearly maps to the 0-vector when using
arguments above, the embeddings of all the remaining vectors are independent and have the same
distribution. It follows that
Pr"∀x→i ∈ X :(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx→i(cid:13)(cid:13)(cid:13)(cid:13)
≤ ε# ≤(cid:16)1 − 2−O(ε√m)(cid:17)X
≤ exp(cid:16)−X2−O(ε√m)(cid:17) .
Now since 0 ∈ X, it follows that to preserve all pairwise distances amongst vectors in X to within (1± ε),
we also have to preserve all norms to within ±ε. This is true since for all x of unit norm:
1
√m
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
This proves the following:
2
2 − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
T Dx(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
1
√m
T Dx −
1
√m
T D0(cid:13)(cid:13)(cid:13)(cid:13)
2
2 − kx − 0k2
.
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Theorem 10. Let T be m×n Toeplitz and D n×n diagonal. If t−(m−1), t−(m−2), . . . , tn−1 and d1, . . . , dn
are independently distributed Rademacher random variables for i = −(m− 1), . . . , n− 1 and j = 1, . . . , n,
then for all 0 < ε < C, where C is a universal constant, there exists a set of N = Ω(n/m) vectors
X ⊂ Rn such that
1
√m
T Dx −
1
√m
2 − kx − yk2
≤ εkx − yk2
2# ≤ exp(cid:16)−N 2−O(ε√m)(cid:17) .
It follows from Theorem 10 that if we want to have constant probability of successfully embedding
any set of N vectors, then either it must be the case that m = Ω(n/N ), or
Pr"∀x, y ∈ X :(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:13)(cid:13)(cid:13)(cid:13)
2
T Dy(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
where C0 is a constant. This in turn implies that
−N 2−O(ε√m) ≥ −C0,
lg N − O(ε√m) ≤ lg C0
=⇒
√m = Ω(ε−1 lg N ) =⇒
m = Ω(ε−2 lg2 N ).
This completes the proof of Theorem 3.
11
References
[1] Dimitris Achlioptas.
Database-friendly random projections:
binary coins.
doi:10.1016/S0022-0000(03)00025-4.
Journal of computer and System Sciences,
Johnson–Lindenstrauss with
66(4):671–687, June 2003.
[2] Nir Ailon and Bernard Chazelle. The fast Johnson–Lindenstrauss transform and approximate nearest
neighbors. SIAM Journal on Computing, 39(1):302–322, May 2009. doi:10.1137/060673096.
[3] Nir Ailon and Edo Liberty. Fast dimension reduction using rademacher series on dual BCH codes.
Discrete & Computational Geometry, 42(4):615–630, 2009. doi:10.1007/s00454-008-9110-x.
[4] Nir Ailon and Edo Liberty. An almost optimal unrestricted fast Johnson–Lindenstrauss transform.
ACM Trans. Algorithms, 9(3):21:1–21:12, June 2013. doi:10.1145/2483699.2483701.
[5] Jeremiah Blocki, Avrim Blum, Anupam Datta, and Or Sheffet. The Johnson–Lindenstrauss trans-
form itself preserves differential privacy. In Proceedings of the 2012 IEEE 53rd Annual Symposium
on Foundations of Computer Science, FOCS '12, pages 410–419, Washington, DC, USA, 2012. IEEE
Computer Society. doi:10.1109/FOCS.2012.67.
[6] C. Boutsidis, A. Zouzias, M. W. Mahoney, and P. Drineas. Randomized dimensionality reduction
for k-means clustering. IEEE Transactions on Information Theory, 61(2):1045–1062, February 2015.
doi:10.1109/TIT.2014.2375327.
[7] E. J. Candes, J. Romberg, and T. Tao. Robust uncertainty principles: exact signal reconstruc-
IEEE Transactions on Information Theory,
tion from highly incomplete frequency information.
52(2):489–509, February 2006. doi:10.1109/TIT.2005.862083.
[8] Michael B. Cohen, Sam Elder, Cameron Musco, Christopher Musco, and Madalina Persu. Di-
mensionality reduction for k-means clustering and low rank approximation. In Proceedings of the
Forty-seventh Annual ACM Symposium on Theory of Computing, STOC '15, pages 163–172, New
York, NY, USA, 2015. ACM. doi:10.1145/2746539.2746569.
[9] Anirban Dasgupta, Ravi Kumar, and Tamás Sarlos. A sparse Johnson–Lindenstrauss transform.
In Proceedings of the Forty-second ACM Symposium on Theory of Computing, STOC '10, pages
341–350, New York, NY, USA, 2010. ACM. doi:10.1145/1806689.1806737.
[10] Sanjoy Dasgupta and Anupam Gupta. An elementary proof of a theorem of Johnson and Linden-
strauss. Random Struct. Algorithms, 22(1):60–65, 2003. doi:10.1002/rsa.10073.
[11] D. L. Donoho. Compressed sensing. IEEE Transactions on Information Theory, 52(4):1289–1306,
April 2006. doi:10.1109/TIT.2006.871582.
[12] Sariel Har-Peled, Piotr Indyk, and Rajeev Motwani. Approximate nearest neighbor: To-
Theory of Computing, 8(14):321–350, 2012.
wards removing the curse of dimensionality.
doi:10.4086/toc.2012.v008a014.
[13] Aicke Hinrichs and Jan Vybíral. Johnson–Lindenstrauss lemma for circulant matrices. Random
Structures & Algorithms, 39(3):391–398, 2011. doi:10.1002/rsa.20360.
[14] Piotr Indyk. Algorithmic applications of low-distortion geometric embeddings. In Proceedings of the
42nd IEEE Symposium on Foundations of Computer Science, FOCS '01, pages 10–33, Washington,
DC, USA, 2001. IEEE Computer Society. doi:10.1109/SFCS.2001.959878.
[15] William B Johnson and Joram Lindenstrauss. Extensions of lipschitz mappings into a hilbert space.
Contemporary mathematics, 26:189–206, 1984. doi:10.1090/conm/026/737400.
[16] Daniel M. Kane and Jelani Nelson. Sparser Johnson–Lindenstrauss transforms. J. ACM, 61(1):4:1–
4:23, January 2014. doi:10.1145/2559902.
[17] Felix Krahmer and Rachel Ward. New and improved Johnson–Lindenstrauss embeddings via the
Restricted Isometry Property. SIAM J. Math. Anal., 43(3):1269–1281, 2011.
12
[18] Kasper Green Larsen and Jelani Nelson. Optimality of the Johnson–Lindenstrauss lemma.
In
Proceedings of the 2017 IEEE 58th Annual Symposium on Foundations of Computer Science, FOCS
'17, Washington, DC, USA, October 2017. IEEE Computer Society.
[19] S. Muthukrishnan. Data Streams: Algorithms and Applications, volume 1(2) of Foundations and
TrendsTM in Theoretical Computer Science. now Publishers Inc., Hanover, MA, USA, January 2005.
doi:10.1561/0400000002.
[20] Daniel A. Spielman and Nikhil Srivastava. Graph sparsification by effective resistances. SIAM J.
Comput., 40(6):1913–1926, December 2011. doi:10.1137/080734029.
[21] Santosh S. Vempala. The random projection method, volume 65 of DIMACS - Series in Discrete
Mathematics and Theoretical Computer Science. American Mathematical Society, Providence, RI,
USA, September 2004. doi:10.1007/978-1-4615-0013-1_16.
[22] Ky Vu, Pierre-Louis Poirion, and Leo Liberti. Using the Johnson–Lindenstrauss lemma in linear
and integer programming. ArXiv e-prints, July 2015. arXiv:1507.00990.
[23] Jan Vybíral. A variant of the Johnson–Lindenstrauss lemma for circulant matrices. Journal of
Functional Analysis, 260(4):1096–1105, 2011. doi:10.1016/j.jfa.2010.11.014.
[24] Kilian Weinberger, Anirban Dasgupta, John Langford, Alex Smola, and Josh Attenberg. Fea-
ture hashing for large scale multitask learning.
In Proceedings of the 26th Annual International
Conference on Machine Learning, ICML '09, pages 1113–1120, New York, NY, USA, 2009. ACM.
doi:10.1145/1553374.1553516.
[25] David P. Woodruff. Sketching as a Tool for Numerical Linear Algebra, volume 10(1–2) of Foundations
and TrendsTM in Theoretical Computer Science. now Publishers Inc., Hanover, MA, USA, 2014.
doi:10.1561/0400000060.
13
|
1312.7409 | 3 | 1312 | 2015-03-08T05:58:46 | Fredholm multiplication conditional type operators on L^p-spaces | [
"math.FA"
] | In this paper, first we investigate closed range multiplication conditional type operators between two Lp-spaces. Then we characterize Fredholm ones when the underlying measure space is non-atomic. Finally we give some examples. | math.FA | math |
FREDHOLM MULTIPLICATION CONDITIONAL TYPE
OPERATORS ON Lp-SPACE
Y. ESTAREMI
Abstract. In this paper, first we investigate closed range multiplication con-
ditional type operators between two Lp-spaces. Then we characterize Fred-
holm ones when the underlying measure space is non-atomic. Finally we give
some examples.
1. Introduction and Preliminaries
Let (X, Σ, µ) be a complete σ-finite measure space. For any sub-σ-finite algebra
A ⊆ Σ with 1 ≤ p ≤ ∞, the Lp-space Lp(X, A, µA) is abbreviated by Lp(A), and
its norm is denoted by k.kp. All comparisons between two functions or two sets
are to be interpreted as holding up to a µ-null set. The support of a measurable
function f is defined as S(f ) = {x ∈ X; f (x) 6= 0}. We denote the vector space of
all equivalence classes of almost everywhere finite valued measurable functions on
X by L0(Σ).
For a sub-σ-finite algebra A ⊆ Σ, the conditional expectation operator associated
with A is the mapping f → EAf , defined for all non-negative function f as well as
for all f ∈ Lp(Σ), 1 ≤ p ≤ ∞, where EAf , by the Radon-Nikodym theorem, is the
unique A-measurable function satisfying
ZA
f dµ = ZA
EAf dµ,
∀A ∈ A.
As an operator on Lp(Σ), EA is an idempotent and EA(Lp(Σ)) = Lp(A). If there
is no possibility of confusion we write E(f ) in place of EA(f ). This operator will
play a major role in our work and we list here some of its useful properties:
• If g is A-measurable, then E(f g) = E(f )g.
• E(f )p ≤ E(f p).
• If f ≥ 0, then E(f ) ≥ 0; if f > 0, then E(f ) > 0.
• E(f g) ≤ E(f p)
• For each f ≥ 0, σ(f ) ⊆ σ(E(f )).
p′ , where 1
p E(gp′
p + 1
p′ = 1 (Holder inequality).
1
1
)
A detailed discussion and verification of most of these properties may be found
in [11]. We recall that an A-atom of the measure µ is an element A ∈ A with
µ(A) > 0 such that for each F ∈ A, if F ⊆ A then either µ(F ) = 0 or µ(F ) = µ(A).
A measure space (X, Σ, µ) with no atoms is called non-atomic measure space. It
is well-known fact that every σ-finite measure space (X, A, µA ) can be partitioned
2010 Mathematics Subject Classification. 47B38, 47B37.
Key words and phrases. Conditional expectation, multiplication operators, Fredholm, closed
range.
1
2
Y. ESTAREMI
uniquely as X = (cid:0)Sn∈N An(cid:1)∪B, where {An}n∈N is a countable collection of pairwise
disjoint A-atoms and B, being disjoint from each An, is non-atomic (see [14]).
Compositions of conditional expectation operators and multiplication operators
appear often in the study of other operators such as multiplication operators and
weighted composition operators. Specifically, in [10], S.-T. C. Moy characterized
all operators on Lp of the form f → E(f g) for g in Lq with E(g) bounded. Eleven
years later, R. G. Douglas, [5], analyzed positive projections on L1 and many of his
characterizations are in terms of combinations of multiplications and conditional
expectations. More recently, P.G. Dodds, C.B. Huijsmans and B. De Pagter, [1],
extended these characterizations to the setting of function ideals and vector lat-
tices. J. Herron presented some assertions about the operator EMu on Lp spaces
in [7]. Also, some results about multiplication conditional type operators can be
found in [6, 8]. In [2, 3, 4] we investigated some classic properties of multiplication
conditional type operators MwEMu on Lp spaces. Let f ∈ L0(Σ), then f is said to
be conditionable with respect to E if f ∈ D(E) := {g ∈ L0(Σ) : E(g) ∈ L0(A)}.
Throughout this paper we take u and w in D(E). In this paper, some necessary
and sufficient conditions for closeness of range of multiplication conditional type
operators between two Lp-spaces are given. Also, Fredholm ones are characterized
when the underlying measure space is non-atomic. Some results of this paper are
a generalization of some results of [13].
Now we give a definition of multiplication conditional type operators on Lp-
spaces.
Definition 1.1. Let (X, Σ, µ) be a σ-finite measure space and let A be a σ-
subalgebra of Σ such that (X, A, µA) is also σ-finite. Let E be the corresponding
conditional expectation operator relative to A. If 1 ≤ p, q < ∞ and u, w ∈ L0(Σ)
(the spaces of Σ-measurable functions on X) such that uf is conditionable and
wE(uf ) ∈ Lq(Σ) for all f ∈ D ⊆ Lp(Σ), where D is a linear subspace, then the
corresponding multiplication conditional type operator is the linear transformation
MwEMu : D → Lq(Σ) defined by f → wE(uf ).
Here we recall some results of [1] that state our results is valid for a large class of
linear operators. Let (X, Σ, µ) be a finite measure space, then L∞(Σ) ⊆ Lp(Σ) ⊆
L1(Σ) and Lp(Σ) is an order ideal of measurable functions on (X, Σ, µ). Thus by
propositions (3.1, 3.3, 3.6) of [1] we have theorems A, B, C:
Theorem A. If T is a linear operator on Lp(Σ) for which
(i) T f ∈ L∞(Σ) whenever f ∈ L∞(Σ).
(ii) kT fnk1 → 0 for all sequences {fn}∞
n=1 ⊆ Lp(Σ) such that fn ≤ g (n =
1, 2, 3, ....) for some g ∈ Lp(Σ) and fn → 0 a.e.,
(iii) T (f.T g) = T f.T g for all f ∈ L∞(Σ) and all g ∈ Lp(Σ),
MULTIPLICATION CONDITIONAL TYPE OPERATORS
3
then there exists a σ-subalgebra A of Σ and there exists w ∈ Lp′
(Σ) such that
T f = EA(wf ) for all f ∈ Lp(Σ).(p−1 + p′−1 = 1)
Theorem B. For a linear operator T : Lp(Σ) → Lp(Σ) the following statement
are equivalent.
(i) T is positive and order continuous, T 2 = T , T 1 = 1 and the range R(T ) of T
is a sublattice.
(ii) There exist a σ-subalgebra A of Σ and a function 0 ≤ w ∈ Lp′
(Σ) with
EA(w) = 1 such that T f = EA(wf ) for all f ∈ Lp(Σ).
Theorem C. For a linear operator T : Lp(Σ) → Lp(Σ) the following statement
are equivalent.
(i) T is a positive and order continuous projection onto a sublattice such that
T 1 is strictly positive.
(ii) There exist a σ-subalgebra A of Σ, 0 ≤ w ∈ Lp′
(Σ) and a strictly positive
function k ∈ L1(Σ) with EA(wk) = 1 such that T f = kEA(wf ) for all f ∈ Lp(Σ).
Moreover, if we choose k such that EA(k) = 1, then both w and k are uniquely
determined by T .
2. CLOSED RANGE AND FREDHOLM WEIGHTED
CONDITIONAL TYPE OPERATORS
In this section first we describe closed range multiplication conditional type op-
erators MwEMu between two Lp-spaces. Let 1 ≤ p, q < ∞ and f ∈ Lp, then it is
easily seen that
kMwEMu(f )kq = kEMv(f )kq,
1
where v = u(E(wq))
q . Thus without loss of generality we can consider the opera-
tor EMv instead of MwEMu, in our discussion about closedness of range. Also, we
recall that for an operator T on a Banach space X, N (T ) = {x ∈ X : T (x) = 0}
and R(T ) = {T (x) : x ∈ X} are called null space and range of T , respectively.
Theorem 2.1. Let 1 < p < ∞ and let p′ be conjugate component to p. Then
(a) If the operator T = EMu from Lp(Σ) into itself is injective and has closed
p′ ≥ δ a.e, On S, where
range, then there exists δ > 0 such that v = (E(up′
S = {x ∈ X : v(x) 6= 0}.
1
))
4
Y. ESTAREMI
(b) If σ(E(u)) = σ(E(up′
S = σ(E(up′
closed range on Lp(Σ).
)) = {x ∈ X : E(up′
)) and there exists δ > 0 such that E(u) ≥ δ a.e. On
)(x) 6= 0}, then the operator T = EMu has
Proof. (a) Let f ∈ Lp(Σ). Then
kT f kp
p = ZX
≤ ZX
= ZX
E(uf )pdµ
(E(up′
))
p
p′ E(f p)dµ
vpf pdµ = kMvf kp
p.
Since T is injective and closed range, then there exists δ > 0 such that for
f ∈ Lp(Σ), kT f kp ≥ δkf kp. Thus
kMvf kLp(S) = kMvf kLp(X)
≥ kT f kp
≥ δkf kLp(X)
≥ δkf kLp(S)
and so kMvf kLp(S) ≥ δkf kLp(S), for all f ∈ Lp(Σ). This mean's that Mv has closed
range on Lp(X). Thus there exists β > 0 such that v ≥ β a.e. On S.
(b) Let fn, g ∈ Lp(Σ) such that kE(ufn) − gkp → 0, when n → ∞. Since
E(u) χS ∈ Lp(S)
δ a.e. On S. This implies that
E(u) ≤ 1
1
g
E(u) ≥ δ a.e. On S, then
and E(u g
g
χS)kp
kE(ufn) − E(u
E(u) χS) = g ∈ Lp(X, A, µ). Hence
p = ZX
= ZS
= ZS
E(u)
E(ufn) − E(u
g
E(u)
)χSpdµ
E(ufn) − E(u
g
E(u)
)χSpdµ
E(ufn) − gpdµ
≤ kE(ufn) − gkp
p → 0,
when n → ∞. So the operator EMu has closed range on Lp(Σ).
Theorem 2.2. Let 1 < q < p < ∞ and let p′, q′ be conjugate component to p
and q respectively. Then
(a) If the operator T = EMu from Lp(Σ) into Lq(Σ) is injective and has closed
range, then
(1) v = 0 a.e. On B and the set {n ∈ N : v(An) 6= 0} is finite.
(2) Mv from Lp(Σ) into Lq(Σ) has finite rank.
Where v = (Euq′
)
1
q′ and S = {x ∈ X : v(x) 6= 0}.
(b) Let
MULTIPLICATION CONDITIONAL TYPE OPERATORS
5
(1) The operator T = EMu has closed range.
(2) The operator T = EMu has finite rank.
(3) v = 0 a.e. On B and the set Nv = {n ∈ N : v(An) 6= 0} is finite.
(4) E(u) = 0 a.e. On B and the set NN (u) = {n ∈ N : E(u)(An) 6= 0} is finite.
Then
(3) → (2) → (1) → (4).
Proof. (a) Let f ∈ Lp(Σ). Then
kT f kq
q = ZX
≤ ZX
= ZX
E(uf )qdµ
(E(uq′
))
q
q′ E(f q)dµ
vqf qdµ
= kMvf kq
q.
Since T is injective and closed range, then there exists δ > 0 such that for f ∈
Lp(Σ), kT f kq ≥ δkf kp. Thus kMvf kq ≥ kT f kp ≥ δkf kp and so kMvf kq ≥ δkf kp,
for all f ∈ Lp(Σ). This mean's that Mv from Lp(Σ) into Lq(Σ) has closed range.
Thus by [13] we have v = 0 a.e. On B and the set {n ∈ N : v(An) 6= 0} is finite.
Also, Mv from Lp(Σ) into Lq(Σ) has finite rank.
(b) If µ(S) = 0 then EMu is the zero operator. So we assume that µ(S) > 0.
i=1Ani for some integer k > 0.
(3) → (2). If (3) holds, then S = ∪n∈Nv An = ∪k
Hence
EMu(Lp(X, Σ, µ)) ⊆ Lp(S, A, µ),
Since for any f ∈ Lp(X, Σ, µ), σ(E(uf )) ⊆ S. This implies that EMu has finite
rank.
(2) → (1) is trivial.
(1) → (4). Suppose that EMu has closed range.First we show that E(u) = 0 a.e.
On B. Assume on the contrary that µ({x ∈ B : E(u)(x) 6= 0}) > 0. Then we have
µ({x ∈ B : E(u)(x) > δ}) > 0 for some δ > 0. Set G = {x ∈ B : E(u)(x) > δ} and
E(u)(x) for x ∈ G. For every f ∈ Lp(G) and
define a function v on G by v(x) =
g ∈ Lq(G),
1
MvEMu(f ) = v(E(u)G)f = f, ME(u)GMv(g) = g.
Thus Mv is the inverse operator of EMuLp(G) = ME(u)G and EMuLp(G) =
ME(u)G is a bounded operator from Lp(G) into Lq(G) that has closed rang.
For any E ∈ AG = {A ∩ G : A ∈ A} with µ(E) < ∞. put f =
E(u)(x) χE(x),
then f ∈ Lp(G). Moreover, ME(u)f = χE and so χE ∈ ME(u)(Lp(G)). Hence,
the the range ME(u)(Lp(G)) contains all linear combinations of such χE's. Thus
1
6
Y. ESTAREMI
ME(u)(Lp(G)) = Lq(G), since ME(u) has closed range and all linear combinations
of such χE's are dense in Lq(G). This implies that Mv maps Lq(G) into Lp(G),
that is Mv is bounded from Lq(G) into Lp(G). Since G is non-atomic by theorem
1.4 of [13] we have v = 0 a.e on G. But this is a contradiction.
1
E(u)(x) for x ∈ S and consider the operator Mw. put f =
Now, we show that NE(u) is finite. Since µ(S) > 0, it follows that NE(u) 6= ∅. Put
E(u)(x) χAn(x),
w(x) =
then f ∈ Lp(G), by the same method of preceding paragraph we see that Mw maps
Lq(S) into Lp(S) that is Mw is bounded from Lq(S) into Lp(S). So by theorem
1.4 of [13] we have
1
b = sup
n∈NE(u)
1
E(u)(An)rµ(An)
= sup
n∈NE(u)
w(An)r
µ(An)
< ∞,
p + 1
r = 1
where 1
Theorem 1.3 of [13] says E(u) ∈ Lr(X, A, µ). So,
q . Since NE(u) 6= ∅, then b > 0 and E(u)(An)rµ(An) ≥ 1
b . While
Σn∈NE(u)
1
b
≤ kE(u)kr
r < ∞.
This implies that NE(u) is finite.
Theorem 2.3. Let 1 < p < q < ∞ and let p′, q′ be conjugate component to p
and q respectively.
(a) If the operator T = EMu from Lp(Σ) into Lq(Σ) is injective and has closed
range, then
(1) The set {n ∈ N : v(An) 6= 0} is finite.
(2) Mv from Lp(Σ) into Lq(Σ) has finite rank.
Where v = (Euq′
)
1
q′ and S = {x ∈ X : v(x) 6= 0}.
(b) Let
(1) The operator T = EMu has closed range.
(2) The operator T = EMu has finite rank.
(3) The set Nv = {n ∈ N : v(An) 6= 0} is finite.
(4) The set NN (u) = {n ∈ N : E(u)(An) 6= 0} is finite.
Then
(3) → (2) → (1) → (4).
Proof. (a) Let f ∈ Lp(Σ). Then kT f kq
q ≤ kMvf kq
q.
MULTIPLICATION CONDITIONAL TYPE OPERATORS
7
Since T is injective and closed range, then there exists δ > 0 such that for f ∈
Lp(Σ), kT f kq ≥ δkf kp. Thus kMvf kq ≥ kT f kp ≥ δkf kp and so kMvf kq ≥ δkf kp,
for all f ∈ Lp(Σ). This mean's that Mv from Lp(Σ) into Lq(Σ) has closed range.
Thus by [13] the set {n ∈ N : v(An) 6= 0} is finite. Also, Mv from Lp(Σ) into Lq(Σ)
has finite rank.
(b) Theorem 2.3 of [4] tells us that v = 0 a.e on B and so E(u) = 0 a.e on B. By
the same method that we used in last theorem, it is easy to see that (3) → (2) → (1).
Now, we show that (1) → (4). Suppose that NN (u) 6= ∅. If we put S = σ(E(u)),
then we can write S = ∪n∈NE(u) An. Define a function w on S by w(x) =
E(u)(x)
for x ∈ S. By the same method that is used in the proof of last Lp(S). Hence by
Theorem 1.3 of [13] we have w ∈ Ls(A), where 1
p . While Theorem 1.4 of
µ(An) < ∞. Since NN (u) 6= ∅ implies b > 0 and
[13] says that b = supn∈NE(u)
since w(An)sµ(An) = µ(An)
b for all n ∈ NN (u), it follows that
E(u)(An)s ≥ 1
1
E(u)(An)s
q + 1
s = 1
Σn∈NE(u)
1
b
≤ kwks
s < ∞
this implies that NE(u) is finite.
In the sequel we consider the function u is A-measurable. In this case E(u) = u
and EMu = Mu Lp(A). Therefore we get the following results.
Corollary 2.4. Let 1 < p < ∞ and let p′ be conjugate component to p. If
u ∈ L0(A) and T = EMu is injective on Lp(Σ). Then the operator T = EMu has
closed range if and only if there exists δ > 0 such that u ≥ δ a.e. On S. Where
S = {x ∈ X : u(x) 6= 0}.
Corollary 2.5. Let 1 < q < p < ∞ and let p′, q′ be conjugate component to p
and q respectively. If u ∈ L0(A) and T = EMu is injective from Lp(Σ) into Lq(Σ).
Then the followings are equivalent:
(1) The operator Mu has closed range.
(2) The operator Mu has finite rank.
(3) u = 0 a.e. On B and the set Nu = {n ∈ N : u(An) 6= 0} is finite.
Corollary 2.6. Let 1 < p < q < ∞ and let p′, q′ be conjugate component to p
and q respectively. If u ∈ L0(A) and T = EMu is injective from Lp(Σ) into Lq(Σ),
then the followings are equivalent:
(1) The operator EMu has closed range.
(2) The operator EMu has finite rank.
(3) The set Nu = {n ∈ N : u(An) 6= 0} is finite.
8
Y. ESTAREMI
If the multiplication conditional type operator EMu is bounded from Lp(Σ)
onto Lp(A), then µ(Z(E(up′
)(x) = 0}) = 0. Suppose that
F ⊆ Z(E(up′
)) with F ∈ A and µ(F ) < ∞. Then χF ∈ Lp(A) = R(T ) and there
exists f ∈ Lp(Σ) such that T f = χF . So by conditional-type Holder inequality we
have
)) = {x ∈ X : E(up′
µ(F ) = ZF
≤ ZF
E(u.f )pdµ
(E(up′
)
p
p′ .f pdµ = 0.
Hence µ(Z(E(up′
))) = 0.
In the next theorem we characterize Fredholm multiplication conditional type
operators on Lp-spaces when the underlying measure space is non-atomic.
Theorem 2.7. Let 1 < p < ∞ and (X, Σ, µ) be a non-atomic measure space.
If T = EMu is a bounded operator from Lp(Σ) into Lp(A), then T is Fredholm if
and only if T is invertible.
Proof. Let T be Fredholm, then T has closed range. First we show that T is
surjective. Suppose on the contrary. Let f0 ∈ Lp(A) \ R(T ). Then there exists a
bounded linear functional Lg0 on Lp(A) for some g0 ∈ Lp′
(A) (p−1 + p′−1 = 1),
which is defined as
Lg0(f ) = ZX
f ¯g0dµ,
f ∈ Lp(A),
such that
Then there exists a positive constant δ such that
Lg0 (f0) = 1,
Lg0(R(T )) = 0.
µ(Eδ = {x ∈ X : f0(x)g0(x) > δ}) > 0.
Since the underlying measure space is non-atomic, then we can find a disjoint
sequences {En}n such that En ⊆ Eδ and 0 < µ(En) < ∞. Let gn = g0.χEn.
Clearly gn ∈ Lp′
(A) and for every f ∈ Lp(A) we have
ZX
¯f T ∗(gn)dµ = ZX
= ZX
¯f ¯uE(gn)dµ
g0E(¯u ¯f .χEn)dµ
= Lg0 (T f ) = 0.
This implies that gn ∈ N (T ∗) and so N (T ∗) is infinite dimensional. Therefore the
codimension of R(T ) is not finite. This is a contradiction, therefore T is surjective.
Let 0 6= f ∈ Lp(Σ) such that T (f ) = 0. Since S(f ) ⊆ S(E(f )), µ(S(f )) > 0
and S(E(f )) is A-measurable, then there exists A ∈ A with 0 < µ(A) < ∞ and
µ(S(f ) ∩ A) > 0. Also since the underlying measure space is non-atomic, we can
find a disjoint sequence of A-measurable subsets {An}n of A with µ(S(f )∩An) > 0.
Clearly fn = f.χS(f )∩An ∈ Lp(Σ) and T (fn) = χAn .T (f ) = 0. This means N (T ) is
infinite dimensional, that is a contradiction. Thus T should be injective and so is
invertible. The converse is obvious.
MULTIPLICATION CONDITIONAL TYPE OPERATORS
9
In the sequel we present some examples of conditional expectations and corre-
sponding multiplication conditional type operators.
Examples. (a) Let (X1, Σ1, µ1) and (X2, Σ2, µ2) be two σ-finite measure spaces
and X = X1 × X2, Σ = Σ1 × Σ2 and µ = µ1 × µ2. Put A = {A × X2 : A ∈ Σ1}.
Then A is a sub-σ-algebra of Σ and for all f ∈ D(EA) we have
EA(f )(x1, x2) = ZX2
f (x1, y)dµ2(y) µ − a.e.
on X.
Mpreover, if (X, Σ, µ) is a finite measure space and k : X × X → C is a Σ ⊗ Σ-
measurable function such that the operator T : Lp(Σ) → Lq(Σ) defined by
T f (x) = ZX
k(x, y)f (y)dµ,
f ∈ Lp(Σ),
is bounded. Then T is a weighted conditional type operator. Since kf is a Σ ⊗ Σ-
measurable function, when f ∈ Lp(Σ). Then by taking u := k and f ′(x, y) = f (y),
we get that
EA(uf )(x, y) = EA(uf ′)(x, y)
= ZX
= ZX
u(x, y)f ′(x, y)dµ(y)
u(x, y)f (y)dµ(y)
= T f (x).
Hence T is a weighted conditional type operator.
operator
In particular the convolution
w ∗ u(y) = ZG
w(y − x)u(x)dm(x)
on an abelian group G with Haar measure m, can be viewed as an integral operator.
(b) This example deals with a weighted conditional expectation operator which
is in the form of an integral operator. Let X = (0, 1] × (0, 1], dµ = dxdy, Σ the
Lebesgue measurable subsets of X and let A be the σ-algebra generated by the
family of the sets of the form A × (0, 1] where A is a Lebesgue measurable subset
of (0, 1]. For f in L2((0, 1]2), we have (EAf )(x, y) = R 1
0 f (x, t)dt.
Let X = (0, ∞) × (0, ∞) and A be the σ-algebra generated by the family of the
sets of the form A × (0, ∞) where A is a Lebesgue measurable subset of (0, ∞).
Then
T f (x, y) = (EAMu(f ))(x, y) = Z ∞
0
u(x, t)f (x, t)dt.
10
Y. ESTAREMI
Hence for every function f : (0, ∞) → R we can define the function f ′ on X as
f ′(x, y) = f (y). So,
T f (x, y) = (EAMu(f ))(x, y)
= (EAMu(f ′))(x, y)
= Z ∞
0
u(x, t)f (t)dt
= T ′(f )(x).
This implies that T is an integral transform and specially by taking u(x, y) =
e−xy we obtain one of the most important classical integral transforms that is widely
used in analysis, namely the Laplace integral transform. We refer to [9] for some
applications of integral transforms, especially Laplace integral transforms.
References
[1] P.G. Dodds, C.B. Huijsmans and B. De Pagter, characterizations of conditional expectation-
type operators, Pacific J. Math. 141(1) (1990), 55-77.
[2] Y. Estaremi, Essential norm of weighted conditional type operators on Lp-spaces, positivity.
18 (2014), 41-52.
[3] Y. Estaremi and M.R. Jabbarzadeh, Compact weighted Lambert type operators on Lp-spaces,
J. Math. Anal. Appl. 420 (2014)118-123.
[4] Y. Estaremi and M.R. Jabbarzadeh, Weighted lambert type operators on Lp-spaces, Oper.
Matrices 1 (2013), 101-116.
[5] R. G. Douglas, Contractive projections on an L1 space, Pacific J. Math. 15 (1965), 443-462.
[6] J. J. Grobler and B. de Pagter, Operators representable as multiplication-conditional expec-
tation operators, J. Operator Theory 48 (2002), 15-40.
[7] J. Herron, Weighted conditional expectation operators, Oper. Matrices 1 (2011), 107-118.
[8] A. Lambert, Lp multipliers and nested sigma-algebras, Oper. Theory Adv. Appl. 104 (1998),
147-153.
[9] Y. Luchko, Integral transforms of the Mellin convolution type and their generating operators,
Integral Trans. and special Func. 11 (2008), 809-851.
[10] Shu-Teh Chen, Moy, Characterizations of conditional expectation as a transformation on
function spaces, Pacific J. Math. 4 (1954), 47-63
[11] M. M. Rao, Conditional measure and applications, Marcel Dekker, New York, 1993.
[12] J. H. Shapiro, The essential norm of a composition operator, Analls of Math. 125 (1987),
375-404.
[13] H. Takagi and K. Yokouchi, Multiplication and composition operators between two Lp-spaces,
Contemporary Math. 232(1999), 321-338.
[14] A. C. Zaanen, Integration, 2nd ed., North-Holland, Amsterdam, 1967.
E-mail address: [email protected]
Department of mathematics, Payame Noor university , P. O. Box: 19395-3697, Tehran,
Iran
|
1201.1308 | 1 | 1201 | 2012-01-05T21:27:07 | On representing and absolutely representing systems of subspaces in Banach spaces | [
"math.FA"
] | We study properties of representing and absolutely representing systems of subspaces in Banach spaces. We also present sufficient conditions for the system of subspaces to be a representing system of subspaces. | math.FA | math | УДК 517.982.22
О ПРЕДСТАВЛЯЮЩИХ И АБСОЛЮТНО ПРЕДСТАВЛЯЮЩИХ
СИСТЕМАХ ПОДПРОСТРАНСТВ В БАНАХОВЫХ
ПРОСТРАНСТВАХ
И.С.Фещенко
Аннотация. В работе изучаются свойства представляющих систем подпространств
и абсолютно представляющих систем подпространств в банаховых пространствах.
Ключевые слова: банахово пространство, представляющая система подпространств,
абсолютно представляющая система подпространств.
1. Введение
Пусть X (cid:22) линейное нормированное пространство над полем K действительных или
комплексных чисел, Xk, k > 1, (cid:22) система подпространств X (т.е. замкнутых линейных
множеств), которую мы будем обозначать S = (X; Xk, k > 1).
Определение 1.1. ([4]) Система S называется представляющей системой подпро-
странств (ПСП) в X, если произвольный x ∈ X можно представить в виде x =
k=1 xk, где xk ∈ Xk, k > 1.
Отметим, что в случае банахова пространства X в книге [12] (определение 15.21)
P∞
ПСП в X называется псевдоразложением X.
Определение 1.2. ([4]) Система S называется абсолютно представляющей систе-
мой подпространств в X, если произвольный x ∈ X можно представить в виде x =
P∞
k=1 xk, где xk ∈ Xk, k > 1 иP∞
k=1 kxkk < ∞.
ПСП и АПСП являются естественным обобщением представляющих и абсолютно
представляющих систем (все Xk одномерны) (см., например, [3]). Определения ПСП и
АПСП можно давать и в более широких классах пространств, чем линейные нормиро-
ванные (см.[4]); ПСП и АПСП в различных классах пространств изучались в работах
Ю.Ф. Коробейника, А.В. Абанина, К.А.Михайлова и др. (см., например, [4],[1],[6]). В
данной работе мы изучаем ПСП и АПСП в банаховых пространствах.
В параграфе 2 изучаются свойства ПСП в X, приведены достаточные условия для
того, чтобы счётная система подпространств была ПСП в X, доказана теорема об
устойчивости ПСП, а также показана связь между ПСП в гильбертовом пространстве
и проблемой Гальперина.
В параграфе 3 приводятся различные критерии АПСП, изучаются свойства АПСП
в X.
1
2
1
0
2
n
a
J
5
]
.
A
F
h
t
a
m
[
1
v
8
0
3
1
.
1
0
2
1
:
v
i
X
r
a
2
И.С.Фещенко
2. Представляющие системы подпространств в банаховых
пространствах
2.1. Некоторые свойства ПСП в банаховых пространствах. Пусть X (cid:22) бана-
хово пространство, S = (X; Xk, k > 1) (cid:22) система его подпространств. Введём необхо-
димые обозначения и определения. Для непустого I ⊂ N определим подпространство
S(I) как замыкание линейной оболочки подпространств Xi, i ∈ I. Обозначим X ∗ со-
пряженное пространство к X; для ϕ ∈ X ∗ обозначим ϕ(S,I) сужение ϕ на S(I). Ясно,
что ϕ(S,I) ∈ (S(I))∗. Для разбиения π = {Ik} (k пробегает конечное или счётное число
значений) множества N определим
F1(S, π, ϕ) =Xk
kϕ(S,Ik)k, ϕ ∈ X ∗.
Разбиение π множества N назовём последовательным, если оно имеет один из сле-
дующих двух видов:
возрастающей последовательности натуральных чисел n(1) < n(2) < . . ., n(0) = 0.
(1) Множества Ik = {n(k − 1) + 1, n(k − 1) + 2, . . . , n(k)}, k > 1 для некоторой
(2) Множества Ik = {n(k − 1) + 1, . . . , n(k)}, 1 6 k 6 p, Ip+1 = {n(p) + 1, n(p) + 2, . . .}
для некоторой возрастающей последовательности натуральных чисел n(1) < . . . <
n(p), n(0) = 0.
Теорема 2.1. Пусть S (cid:22) ПСП в X. Тогда существует ε > 0 такое, что для произ-
вольного последовательного разбиения π множества N выполнено
(2.1)
F1(S, π, ϕ) > εkϕk, ϕ ∈ X ∗.
Доказательство. Определим пространство
Dc = {ξ = (x1, x2, . . .) xk ∈ Xk, k > 1,
xk сходится}
∞Xk=1
с нормой kξk = supk>1 kx1 + . . . + xkk. Легко проверить, что Dc банахово. Опреде-
k=1 xk. Тогда A
ограничен (более того, kAk 6 1). Поскольку S является ПСП в X, то Im(A) = X.
Из теоремы про открытое отображение следует существование числа M > 0 такого,
что для произвольного x ∈ X существует ξ = (x1, x2, . . .) ∈ Dc, для которого x = Aξ
k=1 xk и для каждого k > 1 kx1 + . . . + xkk 6 Mkxk.
Для натуральных l 6 k определим xl,k = xl + . . . + xk, тогда kxl,kk 6 2Mkxk. Для
последовательного разбиения π первого вида и ϕ ∈ X ∗ имеем
лим линейный оператор A : Dc → X равенством A(x1, x2, . . .) = P∞
и kξk 6 Mkxk. Тогда x = P∞
∞Xk=1
ϕ(xn(k−1)+1,n(k)) 6
ϕ(x) = ϕ(
xn(k−1)+1,n(k)) 6
∞Xk=1
6
∞Xk=1
kϕ(S,Ik)kkxn(k−1)+1,n(k)k 6 2MkxkF1(S, π, ϕ),
откуда, в силу произвольности x ∈ X, F1(S, π, ϕ) > (1/(2M))kϕk. Для последователь-
ного разбиения π второго вида такая же оценка доказывается аналогично.
(cid:3)
3
Для подмножества M ⊂ X обозначим M ⊥ множество всех ϕ ∈ X ∗ таких, что
ϕ(x) = 0, x ∈ M. Напомним (см., например, параграф 15 книги [12]), что система под-
пространств Gk, k > 1, банахова пространства E называется разложением Шаудера
E, если для каждого x ∈ E существуют и единственны xk ∈ Gk, k > 1, такие, что
k=1 xk.
x =P∞
Теорема 2.2. Пусть S (cid:22) ПСП в X. Тогда система подпространств X ′
1 является разложением Шаудера в замыкании своей линейной оболочки.
k =Tj6=k X ⊥
j , k >
kϕ + ψk > εkϕk (см. теорему 15.5 в [12]).
Доказательство. Достаточно доказать, что существует ε > 0, такое, что для произ-
k выполнено
вольных натуральных n, m и произвольных ϕ ∈Pn
Из теоремы 2.1 следует, что существует ε > 0 такое, что для произвольного по-
следовательного разбиения π выполнено неравенство (2.1). Пусть ϕ ∈Pn
k, ψ ∈
Pn+m
k. Определим разбиение π так: I1 = {1, 2, . . . , n}, I2 = {n + 1, n + 2, . . .}. Тогда
kϕ + ψk > k(ϕ + ψ)(S,I1)k = kϕ(S,I1)k = kϕ(S,I1)k + kϕ(S,I2)k > εkϕk,
k=n+1 X ′
k, ψ ∈Pn+m
k=n+1 X ′
k=1 X ′
k=1 X ′
откуда следует нужное утверждение.
(cid:3)
Будем говорить, что система подпространств S является перестановочной ПСП
(ППСП) в X, если для произвольной биекции σ : N → N система подпространств
Sσ = (X; Xσ(k), k > 1) является ПСП в X. Напомним (см., например, с.534 в [12]), что
разложение Шаудера Gk, k > 1 банахова пространства E называется безусловным,
если каждый сходящийся ряд видаP∞
Теорема 2.3. Пусть S (cid:22) ППСП в X. Тогда система подпространств X ′
1 является безусловным разложением Шаудера в замыкании своей линейной оболоч-
ки.
j , k >
k =Tj6=k X ⊥
k=1 xk, xk ∈ Gk, k > 1, сходится безусловно.
Доказательство. Воспользуемся следующим утверждением (см. теорему 15.18 в [12]):
если Gk, k > 1, (cid:22) система подпространств банахова пространства E, причём замыка-
ние линейной оболочки Gk, k > 1, равно E, то Gk, k > 1, является безусловным разло-
жением Шаудера E тогда и только тогда, когда для произвольной биекции σ : N → N
система подпространств Gσ(k), k > 1, является разложением Шаудера E. Теперь из
теоремы 2.2 следует нужное утверждение.
(cid:3)
2.2. Достаточное условие для того, чтобы система подпространств была
ПСП в X. Пусть X (cid:22) линейное нормированное пространство, S = (X; Xk, k > 1) (cid:22)
система его подпространств. Для множества F ⊂ X и элемента x ∈ X обозначим
d(x, F ) расстояние от x до F . Определим множество
∆(S) = {x ∈ X lim inf
k→∞
d(x, Xk) = 0}.
4
И.С.Фещенко
Теорема 2.4. Если замыкание линейной оболочки ∆(S) равно X, то S является
ППСП в X.
Доказательство. Очевидно, для произвольной биекции σ : N → N ∆(Sσ) = ∆(S).
Поэтому достаточно доказать, что произвольная система подпространств S, удовле-
творяющая условию теоремы, является ПСП в X.
виде x =P∞
Достаточно доказать, что произвольный x ∈ X,kxk < 1 может быть представлен в
k=1 xk, где xk ∈ Xk. Итак, пусть x ∈ X,kxk < 1. Для k = 1, 2, . . . проделаем
Пусть для некоторого k > 1 у нас уже определены натуральные числа N(l, i, j) для
следующие операции.
l = 1, . . . , k − 1, i = 1, . . . , r(l), j = 1, . . . , N(l) и элементы yl,i,j ∈ XN (l,i,j), причём
kx −
k−1Xl=1
r(l)Xi=1
N (l)Xj=1
yl,i,jk < 2−(k−1)
(для k = 1 ничего не определено). Обозначим z = x −Pk−1
j=1 yl,i,j, тогда
kzk < 2−(k−1) (для k = 1 определяем z = x). Существует N(k) ∈ N и элементы
xk,j ∈ ∆(S), j = 1, . . . , N(k), такие, что
i=1PN (l)
l=1 Pr(l)
(2.2)
kz −
N (k)Xj=1
xk,jk < 2−k.
Выберем r(k) ∈ N так, чтобы
(2.3)
kxk,j/r(k)k < 2−k(N(k))−1, j = 1, . . . , N(k).
Неравенство (2.2) перепишем в виде
(2.4)
kz −(cid:16) xk,1 + . . . + xk,N (k)
r(k)
{z
Из kzk < 2−(k−1) и (2.2) следует, что kPN (k)
вольного a = 1, . . . , r(k) − 1 имеем:
(2.5)
xk,1 + . . . + xk,N (k)
+ . . . +
r(k)
r(k)
xk,1 + . . . + xk,N (k)
+ . . . +
r(k)
(cid:17)k < 2−k.
}
j=1 xk,jk < 2−(k−1) + 2−k. Поэтому для произ-
xk,1 + . . . + xk,N (k)
r(k)
k < 2−(k−1) + 2−k
a
{z
}
k
Ясно, что xk,j/r(k) ∈ ∆(S) для j = 1, . . . , N(k). Используя неравенства (2.3),(2.4),(2.5)
несложно видеть, что существуют натуральные числа N(k, i, j), i = 1, . . . , r(k); j =
1, . . . , N(k) и элементы yk,i,j ∈ XN (k,i,j), такие, что
N(k, i, j) < N(k, i′, j′) если i < i′; N(k, i, j) < N(k, i, j′) если j < j′;
(1) N(k − 1, r(k − 1), N(k − 1)) < N(k, 1, 1) (для k = 1 это условие отсутствует);
5
yl,i,jk < 2−k;
(2) kz −Pr(k)
i=1PN (k)
j=1 yk,i,jk < 2−k, т.е.
(3) для произвольного a = 1, . . . , r(k) − 1
r(l)Xi=1
N (l)Xj=1
kXl=1
N (k)Xj=1
kx −
k
aXi=1
yk,i,jk < 2−(k−1) + 2−k;
Из неравенств (2.7),(2.8),(2.9) следует δs < 6 · 2−k → 0, s → ∞. Поэтому справедливо
равенство (2.6). Дополняя его в нужных местах нулями, получим искомое разложение
(cid:3)
(4) kyk,i,jk < 2−k(N(k))−1 для всех i = 1, 2, . . . , r(k), j = 1, 2, . . . , N(k).
Действительно, сначала выберем N(k, 1, 1) и y(k, 1, 1) (достаточно близко к xk,1/r(k)),
затем N(k, 1, 2) и yk,1,2 (достаточно близко к xk,2/r(k)), . . ., затем N(k, 1, N(k)) и
yk,1,N (k) (достаточно близко к xk,N (k)/r(k)), затем переходим к выбору ¾второй груп-
пы¿: N(k, 2, 1) и y(k, 2, 1) (достаточно близко к xk,1/r(k)) и т.д.
Выполнив такие операции, получим набор элементов yl,i,j ∈ XN (l,i,j), где l = 1, 2, . . .,
i = 1, . . . , r(l), j = 1, 2 . . . , N(l). По построению N(l, i, j) < N(l′, i′, j′) если l < l′;
N(l, i, j) < N(l, i′, j′) если i < i′; N(l, i, j) < N(l, i, j′) если j < j′. Покажем, что
(2.6)
x = y1,1,1 + y1,1,2 + . . . + y1,1,N (1) + y1,2,1 + . . . + y1,2,N (1) + . . . +
+ y1,r(1),1 + . . . + y1,r(1),N (1) + y2,1,1 + . . . + y2,1,N (2) + y2,2,1 + . . .
Для этого рассмотрим сумму первых s членов ряда в правой части (2.6). Представим
s в виде s = r(1)N(1) + . . . + r(k − 1)N(k − 1) + aN(k) + b, где 0 6 a 6 r(k) − 1,
0 6 b 6 N(k) − 1. Оценим
δs = kx −
k−1Xl=1
yl,i,j −
aXi=1
N (k)Xj=1
yk,i,j −
bXj=1
yk,a+1,jk.
Из построения yl,i,j следуют оценки
N (l)Xj=1
r(l)Xi=1
r(l)Xi=1
k−1Xl=1
N (k)Xj=1
aXi=1
kx −
N (l)Xj=1
yl,i,jk < 2−(k−1),
k
yk,i,jk < 2−(k−1) + 2−k,
yk,a+1,jk < N(k)2−k(N(k))−1 = 2−k.
(2.7)
(2.8)
(2.9)
k
bXj=1
x =P∞
k=1 xk, xk ∈ Xk.
6
И.С.Фещенко
Приведём пример системы S, для которой выполнено условие теоремы 2.4. Для
двух подпространств Y, Z пространства X определим
ρ0(Y, Z) = sup{d(y, Z) y ∈ Y,kyk = 1}.
(2.10)
Пример 2.1. Пусть система подпространств Yj, j ∈ Λ (Λ (cid:22) некоторое множество ин-
дексов) такова, что замыкание линейной оболочки Yj, j ∈ Λ равно X. Пусть система
S = (X; Xk, k > 1) такова, что для каждого j ∈ Λ существует последовательность
натуральных чисел k(1) < k(2) < . . ., для которой limn→∞ ρ0(Yj, Xk(n)) = 0. Тогда для
каждого j ∈ Λ Yj ⊂ ∆(S). Поэтому S удовлетворяет условию теоремы 2.4, а значит,
является ПСП в X.
Замечание 1. Если система S удовлетворяет условию теоремы 2.4, то для каждо-
го n ∈ N система S(>n) = (X; Xk+n−1, k > 1) также удовлетворяет условию теоремы
2.4, а поэтому является ПСП в X. Поэтому для каждого n ∈ N замыкание линейной
оболочки подпространств Xk, k > n равно X. Последнее условие не является достаточ-
ным для того, чтобы S была ПСП в X. Это показывает следующий пример (который
относится к математическому фольклору).
Действительно, если f (x) = P∞
Пусть X = Lp([0, 1], dx) (p ∈ [1,∞)), подпространство Xk порождено xk, k > 0 (нам
удобнее нумеровать подпространства числами 0, 1, 2, . . ., а не 1, 2, . . .). Тогда для всех
n > 0 замыкание линейной оболочки Xk, k > n равно X, но S не есть ПСП в X.
k=0 akxk (сходимость по норме пространства X), то
kakxkk → 0, k → ∞, а поэтому для всех достаточно больших k ak 6 (kp + 1)1/p.
Поэтому f (x) ∈ C ∞([0, 1)).
2.3. Об одном достаточном условии для того, чтобы x ∈ X допускал разло-
жение по системе подпространств S в случае гильбертова пространства X.
Пусть X (cid:22) гильбертово пространство. Для подпространства Y ⊂ X обозначим PY ор-
топроектор на Y . Пусть S = (X; Xk, k > 1) (cid:22) система подпространств X. Рассмотрим
произвольный x ∈ X и попробуем разложить его по системе подпространств S, т.е.
представить в виде x =P∞
k=1 xk, где xk ∈ Xk, k > 1.
Естественно попробовать определить искомое разложение следующим образом:
x = PX1x + (I − PX1)x = PX1x + PX2(I − PX1)x + (I − PX2)(I − PX1)x = . . . =
=
PXk(I − PXk−1) . . . (I − PX1)x + (I − PXn) . . . (I − PX1)x = . . .
nXk=1
Определим операторы E0 = I, En = (I − PXn) . . . (I − PX1), n > 1. Обозначим xn =
k=1 xk +Enx, n > 1. Таким образом, получаем следующее
утверждение.
PXnEn−1x, n > 1, тогда x =Pn
Утверждение 2.1. Если Enx → 0, n → ∞, то x =P∞
Таким образом, если последовательность операторов En сходится к 0 сильно при
n → ∞, то система подпространств S является ПСП в X. Однако вопрос о сильной
сходимости En к 0 может оказаться очень сложным. Рассмотрим следующий пример.
k=1 xk.
7
Пусть N ∈ N, Hk, 1 6 k 6 N (cid:22) подпространства X, причём TN
k=1 Hk = 0. Пусть
отображение i(·) : N → {1, 2, . . . , N} таково, что i(k + 1) 6= i(k), k > 1 и для каждого
m ∈ {1, 2, . . . , N} существует бесконечно много k, для которых i(k) = m. Определим
Xk = H ⊥
i(k), k > 1. Тогда En = PHi(n) . . . PHi(1), n > 1. Известно, что En сходится к 0
слабо при n → ∞ (см.[9]). Вопрос о сильной сходимости En к 0 при n → ∞ называется
проблемой Гальперина и является чрезвычайно сложным (см., например, [10]). В то
же время из теоремы 2.4 следует, что S является ПСП в X.
2.4. Устойчивость ПСП в банаховых пространствах. Пусть X (cid:22) банахово про-
странство, S = (X; Xk, k > 1) (cid:22) система подпространств X. Мы покажем, что ес-
соответствующих результатов мы обобщим результаты параграфов 2,3 работы [7], в
которой рассматриваются системы одномерных подпространств, на произвольные си-
стемы подпространств.
ли подпространства Xk, eXk достаточно ¾близки¿, k > 1, то система подпространств
eS = (X; eXk, k > 1) также является ПСП в X. За меру ¾близости¿ подпространств
выберем величину ρ0(Xk, eXk), определённую формулой (2.10). Для доказательства
[7]). Для набора P = (x1, . . . , xn), где xk ∈ Xk, 1 6 k 6 n, определим Σ(P ) =Pn
Введём необходимые определения (обобщающие определения параграфа 2 работы
k=1 xk,
а также
ΘS(P ) = max
16k6nk
xjk.
kXj=1
Для x ∈ X, ε > 0 определим
ΘS(x, ε) = inf{ΘS(P ) kΣ(P ) − xk 6 ε}.
(Мы считаем, что inf(∅) = ∞.) Для x ∈ X определим
Θ∗
S(x) = sup{ΘS(x, ε) ε > 0} = lim
ε→0+
ΘS(x, ε).
Наконец, определим
Следующие две леммы и теорема доказываются точно так же, как леммы 1,2 и
ΘS = sup{Θ∗
S(x) kxk 6 1}.
теорема 1 в [7].
S(x) < ∞ для произвольного x ∈ X, то Θ∗
Лемма 2.1. Если Θ∗
эквивалентная k · k.
Лемма 2.2. Если для некоторых α ∈ (0, 1), B > 0 выполнено ΘS(x, αkxk) 6 Bkxk, x ∈
X, то Θ∗
S(x) (cid:22) норма на X,
S(x) 6 B
1−αkxk, x ∈ X.
Теорема 2.5. Следующие утверждения эквивалентны:
(1) S является ПСП в X,
(2) существуют α ∈ (0, 1), B > 0, такие, что для произвольного x ∈ X
(3) Θ∗
ΘS(x, αkxk) 6 Bkxk,
S(x) < ∞ для произвольного x ∈ X,
8
И.С.Фещенко
(4) ΘS < ∞.
Теперь установим теорему об устойчивости ПСП в X.
Теорема 2.6. Пусть S (cid:22) ПСП в X. Если система подпространств eS = (X; eXk, k >
1) такова, чтоP∞
k=1 ρ0(Xk, eXk) < (2ΘS)−1, то eS является ПСП в X.
Доказательство. Доказательство аналогично доказательству теоремы 2 в [7]. Пусть
x ∈ X, x 6= 0. Пусть ε > 0. Существует P = (x1, . . . , xn), xk ∈ Xk, такой, что
kx − Σ(P )k 6 ε, ΘS(P ) 6 ΘS(1 + ε)kxk.
Тогда kxkk 6 2ΘS(1 + ε)kxk, 1 6 k 6 n. Обозначим dk = ρ0(Xk, eXk), k > 1. Для
произвольного k, 1 6 k 6 n, существует exk ∈ eXk, для которого kxk −exkk 6 dk(1 +
ε)kxkk. Определим eP = (ex1, . . . ,exn). Тогда
nXk=1
kx−Σ(eP )k 6 kx−Σ(P )k+kΣ(P )−Σ(eP )k 6 ε+
dk(1+ε)kxkk 6 ε+2ΘS(1+ε)2kxk
Для произвольного m, 1 6 m 6 n, имеем
nXk=1
dk.
k
dk.
(cid:3)
mXk=1
xkk + k
mXk=1
mXk=1exkk 6 k
(exk − xk)k 6 ΘS(1 + ε)kxk + 2ΘS(1 + ε)2kxk
Зафиксируем α ∈ (2ΘSP∞
k=1 dk, 1), B > ΘS + 2ΘSP∞
ε > 0 из доказаных неравенств имеем Θ eS(x, αkxk) 6 Bkxk. Поэтому eS является ПСП
k=1 dk. При достаточно малом
mXk=1
в X.
3. Абсолютно представляющие системы подпространств в банаховых
пространствах
ностей ξ = (x1, x2, . . .), xk ∈ Xk, для которых kξk = P∞
равенством A(x1, x2, . . .) = P∞
3.1. Критерии АПСП. Пусть X (cid:22) банахово пространство, S = (X; Xk, k > 1) (cid:22)
система подпространств X. Определим l1(X1, X2, . . .) как множество последователь-
k=1 kxkk < ∞. Ясно, что
l1(X1, X2, . . .) (cid:22) банахово пространство. Определим оператор A : l1(X1, X2, . . .) → X
k=1 xk. Система S является АПСП в X тогда и только
тогда, когда Im(A) = X. Хорошо известно, что это равносильно тому, что A∗ : X ∗ →
(l1(X1, X2, . . .))∗ является изоморфным вложением, т.е. для некоторого ε > 0 kA∗ϕk >
εkϕk, ϕ ∈ X ∗. Легко видеть, что (l1(X1, X2, . . .))∗ = l∞(X ∗
2 , . . .) (cid:22) множество всех
последовательностей η = (ϕ1, ϕ2, . . .), ϕk ∈ X ∗
k , для которых kηk = supk kϕkk < ∞.
k=1 ϕk(xk). Легко видеть, что A∗ϕ = (ϕ(S,1), ϕ(S,2), . . .), где
ϕ(S,k) обозначено сужение ϕ на Xk. Таким образом, получаем следующую теорему.
При этом действие η(ξ) =P∞
1 , X ∗
Теорема 3.1. S является АПСП в X тогда и только тогда, когда существует ε > 0
такое, что
(3.1)
k kϕ(S,k)k > εkϕk, ϕ ∈ X ∗.
sup
9
Отметим, что теорему 3.1 можно сформулировать следующим образом (уменьшив
ε): S является АПСП в X тогда и только тогда, когда существует ε > 0 такое, что
для произвольного ϕ ∈ X ∗,kϕk = 1 существуют k > 1, x ∈ Xk,kxk = 1 такие, что
ϕ(x) > ε.
Понятие АПСП тесно связано с понятием абсолютно представляющего семейства
(АПСм). Напомним (см., например, [5]), что множество D ⊂ X называется АПСм в X
j=1 ajxj
j=1 kajxjk < ∞. АПСм в банаховых и гильбертовых пространствах изучались в
k=1{x ∈ Xk,kxk = 1}.
Ясно, что S является АПСП в X тогда и только тогда, когда D(S) является АПСм в
X.
если для произвольного x ∈ X существуют aj ∈ K, xj ∈ D, такие, что x =P∞
и P∞
[13],[2],[8]. Для системы подпространств S определим D(S) =Sn
Приведенный критерий для АПСП (см. абзац после теоремы 3.1) можно получить
из следующего хорошо известного критерия для АПСм (см., например, теорему 1 в
[5], теорему 2.1 в [13]), который доказывается аналогично теореме 3.1. (Множество D
называется нормированным, если kxk = 1, x ∈ D.)
Теорема 3.2. Пусть D (cid:22) нормированное множество в X. D является АПСм в X
тогда и только тогда, когда существует ε > 0 такое, что для произвольного ϕ ∈
X ∗,kϕk = 1, существует x ∈ D такой, что ϕ(x) > ε.
Далее мы докажем критерий для АПСм в X, из которого непосредственно следует
критерий для АПСП в X (вместо D надо взять D(S)). Следующая теорема обобщает
теорему 3 в [8] и показывает, что в теореме 3.2 условие произвольности ϕ ∈ X ∗ можно
ослабить. (Будем говорить, что D тотально в X, если замыкание линейной оболочки
D равно X).
Теорема 3.3. Пусть X (cid:22) банахово пространство, Y (cid:22) конечномерное подпростран-
ство X, D (cid:22) тотальное в X нормированное множество. D является АПСм в X
тогда и только тогда, когда существует ε > 0 такое, что для произвольного ϕ ∈
Y ⊥,kϕk = 1 существует x ∈ D такой, что ϕ(x) > ε.
Для доказательства нам нужна следующая лемма (которая относится к математи-
ческому фольклору).
Лемма 3.1. Пусть Y, Z (cid:22) подпространства банахова пространства X. Если Y ∩Z =
0 и Y + Z = X, то существует c > 0 такое, что для произвольных y ∈ Y, z ∈ Z
ky + zk > c(kyk + kzk).
Доказательство. Определим пространство Y ⊕ Z как множество пар ξ = (y, z), y ∈
Y, z ∈ Z, с нормой kξk = kyk + kzk. Ясно, что Y ⊕ Z банахово. Определим оператор
A : Y ⊕ Z → X равенством A(y, z) = y + z. Тогда A ограничен, ker A = 0, ImA = X.
Поэтому A обратим, откуда непосредственно следует нужное утверждение.
(cid:3)
Доказательство теоремы 3.3. Необходимость очевидна. Докажем достаточность. Пред-
положим, что D (cid:22) не АПСм в X. Поскольку Y конечномерно, то оно дополняемо в
X, т.е. существует подпространство Z такое, что Y ∩ Z = 0 и Y + Z = X. Существует
Оценим kηk. Пусть y ∈ Y , y =Pm
ϕ(y) 6
tkϕ(ek) 6
mXk=1
mXk=1
Поэтому для произвольных y ∈ Y, z ∈ Z
tk(ϕ(ek − fk) + ϕ(fk)) 6 2δ
mXk=1
tk 6 2δc−1
2 kyk.
10
И.С.Фещенко
c1 > 0 такое, что
(3.2)
ky + zk > c1(kyk + kzk), y ∈ Y, z ∈ Z.
Выберем в Y нормированный базис e1, . . . , em. Существует c2 > 0 такое, что для про-
извольных t1, . . . , tm ∈ K kPm
k=1 tkekk > c2Pm
Рассмотрим произвольное δ > 0. Поскольку D тотально в X, существуют элементы
f1, . . . , fm из линейной оболочки D, такие, что kek − fkk < δ,kfkk = 1 для k = 1, . . . , m.
Ясно, что {f1, . . . , fm}∪D не является АПСм в X. Поэтому существует ϕ ∈ X ∗, kϕk = 1
такой, что ϕ(fk) 6 δ для 1 6 k 6 m, ϕ(x) 6 δ для x ∈ D. Определим линейные
функционалы ψ, η равенствами ψ(y + z) = ϕ(z), η(y + z) = ϕ(y), y ∈ Y, z ∈ Z. Из
неравенства (3.2) следует, что ψ, η ∈ X ∗. Более того, ψ ∈ Y ⊥, η ∈ Z ⊥.
k=1 tk.
k=1 tkek. Тогда
Положим c3 = 2c−1
η(y + z) = ϕ(y) 6 2δc−1
1 c−1
2 , тогда kηk 6 c3δ.
2 kyk 6 2c−1
1 c−1
2 δky + zk.
Поскольку kϕk = 1, то kψk > (1 − c3δ). Для произвольного x ∈ D ψ(x) 6 δ(1 + c3).
Положим eψ = ψ/kψk. Тогда eψ ∈ Y ⊥, keψk = 1. Для произвольного x ∈ D eψ(x) 6
δ(1 + c3)/(1 − c3δ). При достаточно малых δ получим противоречие.
Рассмотрим АПСП в равномерно гладких пространствах (АПСм в равномерно глад-
ких пространствах изучались в [13],[2]). Напомним определение равномерно гладкого
пространства. Определим модуль гладкости пространства X равенством
ρ(τ ) = sup{(kx + yk + kx − yk)/2 − 1 kxk = 1,kyk = τ}, τ > 0.
(cid:3)
X называется равномерно гладким если ρ(τ )/τ → 0 при τ → 0. Для нас равномерно
гладкие пространства важны по следующей причине: как мы увидим при доказатель-
стве следующей теоремы, если S (cid:22) АПСП в равномерно гладком X, то для каждого
x ∈ X разложение x в абсолютно сходящийся ряд по системе подпространств S может
быть получено простым ¾естественным¿ образом.
Будем говорить, что множество A является λ-сетью для множества B, если для
произвольного x ∈ B существует y ∈ A такой, что kx − yk 6 λ. Обозначим VX =
{x ∈ X,kxk = 1}. Напомним, что для системы подпространств S = (X; Xk, k > 1)
Теорема 3.4. Пусть X (cid:22) равномерно гладкое банахово пространство. Тогда утвер-
ждения равносильны:
D(S) =S∞
k=1 VXk.
(1) S = (X; Xk, k > 1) является АПСП в X,
(2) существуют τ, λ ∈ (0, 1) такие, что τ D(S) (cid:22) λ-сеть для VX,
(3) λS = supkxk=1 inf k>1 d(x, Xk) < 1.
Доказательство. (1) ⇒ (2). Для действительного пространства X нужное утвержде-
ние следует из теоремы 3.1 в [13]. Для комплексного X рассмотрим X как простран-
ство над R и из упомянутой теоремы получим нужное.
11
(2) ⇒ (3). Очевидно.
(3) ⇒ (1). Доказательство аналогично доказательству теоремы 3 в [2]. Фиксируем
λ ∈ (λS, 1). Рассмотрим произвольный x ∈ X. Существуют i(1) ∈ N, x1 ∈ Xi(1) такие,
что kx − x1k 6 λkxk. Определим y1 = x − x1, тогда ky1k 6 λkxk и x = x1 + y1. Далее
проделаем аналогичную процедуру. Пусть мы имеем разложение x = x1 + . . . + xk + yk.
Существуют i(k + 1) ∈ N, xk+1 ∈ Xi(k+1) такие, что kyk − xk+1k 6 λkykk. Определим
yk+1 = yk − xk+1, тогда kyk+1k 6 λkykk и x = x1 + . . . + xk+1 + yk+1. Индукцией по k
легко установить, что kykk 6 λkkxk, k > 1. Поэтому
(3.3)
x =
xk.
k=1 kxkk 6 ((1 + λ)/(1− λ))kxk. Для
Ясно, что kxkk 6 λk−1(1 + λ)kxk, k > 1, поэтомуP∞
k > 1 определим zk =Pj:i(j)=k xj, тогда zk ∈ Xk. Из (3.3) имеем x =P∞
Замечание 2. В работе [2] доказано, что каждая АПС (одномерных подпространств)
в равномерно гладком X является ¾быстрой¿ ПС (см. определение 3 и теорему 3 в [2]).
Аналогичное утверждение верно для АПСП. Как следует из доказательства теоремы
3.4, (3) ⇒ (1), каждая АПСП S в равномерно гладком X является ¾быстрой¿ ПСП
(наше определение согласовано с определением 3 в [2]): существуют C > 0 и λ ∈ (0, 1)
такие, что для произвольного x ∈ X существует инъективное отображение k 7→ n(k)
k=1 zk.
(cid:3)
k=1 yk и kykk 6 Cλkkxk, k > 1.
и элементы yk ∈ Xn(k) такие, что x =P∞
∞Xk=1
Рассмотрим теперь АПСП в гильбертовых пространствах; как мы увидим далее,
критерий для АПСП в гильбертовых пространствах приобретает геометрическую на-
глядность (см. также теоремы 1,2 в [8]). Итак, пусть X гильбертово. Тогда X ∗ можно
отождествить с X: ϕ(·) = (·, ϕ). S является АПСП в X тогда и только тогда, когда
существует ε > 0 такое, что для произвольного ϕ ∈ VX существует x ∈ D(S) такой,
что (x, ϕ) > ε.
Теорема 3.5. Пусть τ > 0, ε ∈ (0, 1]. Следующие условия равносильны:
(1) для произвольного ϕ ∈ VX существует x ∈ D(S) такой, что (x, ϕ) > ε,
(2) τ D(S) является √1 + τ 2 − 2τ ε-сетью для VX .
Доказательство. Для произвольных ϕ ∈ VX, x ∈ D(S) имеем kϕ − τ xk2 = 1 + τ 2 −
2τ Re(x, ϕ). Из этого равенства очевидным образом следует нужное утверждение. (cid:3)
Следствие 3.1. Пусть τ > 0. Система подпространств S является АПСП в X то-
гда и только тогда, когда существует λ ∈ (0,√1 + τ 2) такое, что τ D(S) является
λ-сетью для VX .
3.2. Об одном свойстве АПСП. Перед тем, как сформулировать и доказать следу-
ющую теорему, напомним определение C-выпуклого пространства и некоторые свой-
ства C-выпуклых пространств.
12
И.С.Фещенко
Пусть Y (cid:22) банахово пространство над R. Обозначим c0(R) множество последова-
тельностей ξ = (z1, z2, . . .), zk ∈ R, для которых zk → 0 при k → ∞, с нормой
kξk = supk zk. Y называется C-выпуклым, если c0(R) не является финитно пред-
ставимым в Y (см., например, параграфы 5.1, 5.2 книги [11] и библиографию в конце
параграфа 5.2). Для натурального n определим
(3.4)
C(n, Y ) = inf(max(k
αkykk αk = ±1, 1 6 k 6 n) yk ∈ Y,kykk > 1, 1 6 k 6 n) .
nXk=1
В из лемм 5.2.1, 5.2.2 и теоремы 5.2.2 книги [11] следует, что Y C-выпукло тогда и
только тогда, когда C(n, Y ) → ∞ при n → ∞.
Пусть Y (cid:22) банахово пространство над C. Обозначим c0(C) множество последова-
тельностей ξ = (z1, z2, . . .), zk ∈ C, для которых zk → 0 при k → ∞, с нормой
kξk = supk zk. Y называется C-выпуклым, если c0(C) не является финитно предста-
вимым в Y . Для натурального n определим C(n, Y ) формулой (3.4); величину CC(n, Y )
формулой (3.4), только максимум берётся по αk = 1, αk ∈ C. Перенося леммы 5.2.1,
5.2.2 и теорему 5.2.2 книги [11] на случай комплексных пространств, получим, что Y
C-выпукло тогда и только тогда, когда CC(n, Y ) → ∞ при n → ∞. Поскольку для
произвольных α1, . . . , αn ∈ C,αk 6 1 и произвольных y1, . . . , yn ∈ Y
то CC(n, Y ) 6 2C(n, Y ), а поэтому Y C-выпукло тогда и только тогда, когда C(n, Y ) →
∞ при n → ∞.
Для I ⊂ N обозначим min(I) наименьший элемент множества I.
Теорема 3.6. Предположим, что X ∗ C-выпукло. Если S (cid:22) АПСП в X, то существу-
ет N0, такое, что для произвольного конечного I ⊂ N, удовлетворяющего min(I) >
N0, система подпространств Xk, k /∈ I является АПСП в X.
Теорема 3.6 очевидным образом следует из следующей леммы.
Лемма 3.2. Пусть I1, I2, . . . (cid:22) подмножества N, m ∈ N. Предположим, каждое на-
туральное n принадлежит не более чем m множествам Ij. Тогда для некоторого j
система подпространств Xk, k /∈ Ij является АПСП в X.
Доказательство. Существует ε > 0, для которого выполнено (3.1). Предположим,
утверждение леммы неверно. Тогда для произвольного j существует ϕj ∈ X ∗,kϕjk =
1, такой, что kϕ(S,k)
k 6 2−j для всех k /∈ Ij. Рассмотрим произвольное n ∈ N.
Существуют α1, . . . , αn ∈ {±1}, для которых kPn
j=1 αjϕjk > C(n, X ∗). Определим
ϕ =Pn
k 6
m + 1. Из неравенства (3.1) следует m + 1 > εC(n, X ∗), что, в силу произвольности n,
противоречит C-выпуклости X ∗.
(cid:3)
j=1 αjϕj. Тогда kϕk > C(n, X ∗) и для произвольного k kϕ(S,k)k 6Pn
j=1 kϕ(S,k)
j
j
k
nXk=1
nXk=1
nXk=1
αkykk 6 k
Re(αk)ykk + k
Im(αk)ykk 6 2 max
βk=±1k
βkykk,
nXk=1
3.3. Устойчивость АПСП.
Теорема 3.7. Если S является АПСП в X, то существует δ > 0 такое, что произ-
13
АПСП в X.
Доказательство. Пусть ε > 0 такое, что выполнено неравенство (3.1). Покажем, что
Выберем ε1 < ε, d1 > d, причём d1 < ε1. Рассмотрим произвольный ϕ ∈ X ∗,kϕk = 1.
вольная система eS = (X; eXk, k > 1), удовлетворяющая supk ρ0(Xk, eXk) < δ, является
если для системы подпространств eS d = supk ρ0(Xk, eXk) < ε, то eS является АПСП в X.
Существует k ∈ N и x ∈ Xk,kxk = 1, такие, что ϕ(x) > ε1. Существует ex ∈ eXk, для
которого kx −exk 6 d1. Имеем
откуда kϕ( eS,k)k > (ε1 − d1)/(1 + d1). Поэтому eS (cid:22) АПСП в X.
ϕ(ex) > ϕ(x) − ϕ(x −ex) > (ε1 − d1), kexk 6 (1 + d1),
Список литературы
(cid:3)
[1] Абанин А.В. Индуктивные абсолютно представляющие системы подпространств // Комплекс-
ный анализ. Теория операторов. Математическое моделирование.(cid:22) Владикавказ: Изд-во ВНЦ
РАН, 2006, С. 27-34.
[2] Вершинин Р.В. О представляющих и абсолютно представляющих системах в банаховых про-
странствах // Матем. физ., анал., геом.(cid:22) 1998.(cid:22) Т.5, №1/2.(cid:22) С.3-14.
[3] Коробейник Ю.Ф. Представляющие системы // УМН.(cid:22) 1981.(cid:22) Т.36,Вып. 1 (217).(cid:22) С.73-126.
[4] Коробейник Ю.Ф. О представляющих системах подпространств // Мат. заметки.(cid:22) 1985.(cid:22)
Т.38,№5.(cid:22) С.741-755.
[5] Коробейник Ю.Ф. Об абсолютно представляющих семействах в некоторых классах локально
выпуклых пространств // Изв. вузов. Матем.(cid:22) 2009.(cid:22) № 9.(cid:22) С.25-35.
[6] Михайлов К.А. Абсолютно представляющие системы подпространств в пространствах пробных
ультрадифференцируемых функций // Изв. Вузов. Сев. Кав. регион. Естественные науки.(cid:22)
2009.(cid:22) № 6.(cid:22) С. 8-11.
[7] Слепченко А.Н. О некоторых обобщениях базисов банаховых пространств // Матем. сб.(cid:22) 1983.(cid:22)
Т. 121 (163),№ 2 (6).(cid:22) С.272-285.
[8] Шрайфель И.С. Об абсолютно представляющих системах в гильбертовых пространствах // Изв.
вузов. Матем.(cid:22) 1995.(cid:22) №9.(cid:22) С.78-82.
[9] Amemiya I., Ando T. Convergence of random products of contractions in Hilbert space // Acta Sci.
Math. Szeged.(cid:22) 1965.(cid:22) V.26.(cid:22) P.239-244.
[10] Bauschke H.H. Projection algorithms: results and open problems // Inherently Parallel Algorithms in
Feasibility and Optimization and their Applications (Haifa 2000).(cid:22) D. Butnariu, Y. Censor, S. Reich
(editors).(cid:22) Elsevier, 2001.(cid:22) P.11-22.
[11] Kadets M.I., Kadets V.M. Series in Banach spaces. Conditional and Unconditional convergence.(cid:22)
Birkhauser Verlag, Basel, Boston,Berlin, 1997.
[12] Singer I. Bases in Banach spaces II.(cid:22) Springer Verlag, Berlin, Heidelberg, New York, 1981.(cid:22) 880 p.
[13] Vershynin R. Absolutely representing systems, uniform smoothness, and type // arXiv:
math/ 9804044v1 [math.FA] 8 Apr 1998.
|
1102.3097 | 1 | 1102 | 2011-02-15T15:21:15 | Phase Space Localization of Riesz bases for L^2(R^d) | [
"math.FA"
] | We prove a strong uncertainty principle for Riesz bases in L^2(R^d) and show that the orthonormal basis constructed by Bourgain possesses the optimal phase-space localization. | math.FA | math |
(1)
(2)
PHASE SPACE LOCALIZATION OF RIESZ BASES FOR L2(Rd)
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
Abstract. We prove a strong uncertainty principle for Riesz bases in L2(Rd) and show
that the orthonormal basis constructed by Bourgain possesses the optimal phase-space
localization.
In [B] J. Bourgain constructed an orthonormal basis for L2(R) consisting of functions
1. Introduction
fn ∈ L2(R), such that
a∈RZR
n∈N(cid:16) inf
sup
x − a2fn(x)2dx + inf
b∈RZR
ξ − b2bfn(ξ)2dξ(cid:17) < ∞ .
Bourgain remarked that the exponent 2 of x − a and ξ − b is optimal and that there
are no orthonormal bases with a better phase-space localization.
In this paper we prove the following strong uncertainty principle for Riesz bases for
L2(Rd).
Theorem 1. If {fn}∞
n=1 is a Riesz basis for L2(Rd) and s > d, then
n∈N(cid:16) inf
a∈RdZRd
sup
x − a2sfn(x)2dx + inf
b∈RdZRd
ξ − b2sbfn(ξ)2dξ(cid:17) = ∞ .
This theorem therefore asserts that the Bourgain basis possesses the best possible
phase-space localization. For the case of an orthonormal basis for L2(R) in dimension
d = 1, Bourgain outlines a proof strategy for Theorem 1. Precisely, he writes that "it has
been shown by T. Steger that L2(R) does not admit a basis of the form fj = eibj xgj(x−aj),
where gj satisfies supj kgjkAǫ < ∞, defining
kgk2
Aǫ =Z (1 + x2)1+ǫgj(x)2dx +Z (1 + ξ2)1+ǫbgj(ξ)2dξ .
Here ǫ > 0 is any strictly positive number. His argument is based on the fact that
the operations x (x-multiplication) and d/dx in the latter basis would become "almost"
diagonal operators, violating the non-commutation property [d/dx, x] = I. He also makes
use of a density computation due to Y. Meyer of the set Λ of pairs (aj, bj) in phase space.
The condition ǫ > 0 is important in Steger's argument as well as for Meyer's distribution
result to be valid", see [B].
2000 Mathematics Subject Classification. 81B99, 81S05, 42C15.
Key words and phrases. Uncertainty principle, Balian-Low theorem, Riesz basis, localized frame.
K. G. was supported in part by the project P22746-N13 of the Austrian Science Foundation (FWF).
E. M. was supported by the Research Council of Norway grant 185359/V30.
1
2
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
Some of these arguments have made their way into the literature. A density argument
related to Meyer's argument has appeared in the fundamental paper of Ramanathan and
Steger [RS] on the density of Gabor frames and has become the main technique to inves-
tigate the density of frames. See [BCHL, GR, H] for some variations of the Ramanathan-
Steger technique. The canonical commutation relations were used in Battle's elegant
proof of the Balian-Low theorem [Bat].
However, a full proof of the uncertainty principle of Theorem 1 has not yet been
given. Research has focused mainly on bases consisting of phase-space shifts fn(x) =
e2πibnxg(x−an) of a single generating function g, so-called Gabor systems. The theorem of
Balian-Low asserts that a basis satisfying (1) cannot consist of a (regular) Gabor system.
We refer the reader to proofs of the theorem in [D] and [Bat], to the survey articles
on the Balian-Low theorem and its generalizations [BHW, CP] and to the monograph
[G1] for detailed discussions of the subject. Gabor systems are somewhat easier to
handle, because one needs to control the localization of only one function in contrast to
Bourgain's case.
In this paper we offer a complete proof of Theorem 1 which extends the result men-
tioned in [B] to higher dimensions and to Riesz bases instead of orthonormal bases. For
orthogonal bases our proof follows the outline of Bourgain. The case of Riesz bases
requires additional ideas. We will apply the theory of localized frames [FG, G2] to verify
that the biorthogonal basis possesses the same localization properties as the original
basis. In a second step we use a bootstrap argument. We will show that if a Riesz basis
violates condition (2) for s > d, then we can construct a new Riesz basis with optimal
phase-space localization, for instance, with all functions in a Gelfand-Shilov space of test
functions.
It may seem a lot of effort to prove the non-existence of well-localized phase-space
bases, but several arguments are of interest in themselves. The proof combines tools
from the density theory of frames, the canonical commutation relations, the theory of
localized frames, recent phase-space methods, and a new argument of how to improve
the quality of a given basis.
One of the corollaries of Theorem 1 is that there is no Riesz basis of phase-space shifts
of the Gaussian function in L2(Rd). This fact implies that there is no subset Λ ⊂ Cd
that is both sampling and interpolating for the Bargmann-Fock space F 2(Cd). This
statement is well-known in dimension d = 1, but seems to have been open in higher
dimensions.
The paper is organized as follows: In Section 2 we show that a well-localized phase-
space basis must be indexed by a set of density one. In Section 3 we prove Theorem 1
for the special case when its biorthogonal basis is also well-localized, it includes the
case of an orthonormal basis. Section 4 contains some preparations from time-frequency
analysis.
In Section 5 we develop the necessary arguments to prove the uncertainty
principle of Theorem 1 for Riesz bases. Section 6 elaborates the non-existence of sets of
simultaneous sampling and interpolation and concludes with further remarks.
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
3
2. Density Conditions
We say that a sequence of functions {fn}∞
n=1 ⊂ L2(Rd) has phase-space localization of
magnitude s, if
In this case there exist points (an, bn) ∈ R2d, such that
sup
a∈RdZRd
n∈N(cid:16) inf
n∈N(cid:16)ZRd
sup
x − a2sfn(x)2dx + inf
x − an2sfn(x)2dx +ZRd
b∈RdZRd
ξ − b2sbfn(ξ)2dξ(cid:17) < ∞ .
ξ − bn2sbfn(ξ)2dξ(cid:17) < ∞ .
Then the set Λ = {(an, bn)}∞
are localized. Note that there is some freedom in the choice of points (an, bn) ∈ R2d.
n=1 is the set in the phase-space where the functions {fn}n
We will first estimate the density of the set Λ = {(an, bn)}∞
n=1 ⊂ R2d both for Riesz
bases and frames for L2(Rd) which have phase-space localization. The ideas we follow
are well known, see [RS, SP, LP, S].
Let Λ be a subset of R2d, we denote by D+(Λ) and D−(Λ) its upper and lower Beurling
densities,
D+(Λ) = lim sup
r→∞
sup
x∈R2d
card(λ ∩ Q(x, r))
Q(x, r)
, D−(Λ) = lim inf
r→∞
inf
x∈R2d
card(λ ∩ Q(x, r))
Q(x, r)
,
where x = (x1, x2) ∈ Rd × Rd and
Q(x, r) = {(y1, y2) ∈ Rd × Rd : x1 − y1 < r, x2 − y2 < r} .
These densities can be also defined by using dilations of cubes or balls in R2d instead of
Q(x, r), as was proved by Landau [L].
A set Λ ⊂ R2d is relatively separated, if supx∈R2d card(cid:0)Λ ∩ (x + [0, 1]2d)(cid:1) < ∞. Clearly,
if D+(Λ) < ∞, then Λ is relatively separated.
Lemma 1. Suppose that {fn}∞
ization of magnitude s, s > 0, at points {(an, bn)}∞
n=1, i.e.,
n=1 is a Riesz basis for L2(Rd) that has phase-space local-
n∈N(cid:16)ZRd
sup
x − an2sfn(x)2dx +ZRd
ξ − bn2sbfn(ξ)2dξ(cid:17) = S < ∞ .
Then Λ = {(an, bn)}∞
Proof. Fix ǫ > 0. We say that a function g ∈ L2(Rd) is ǫ-concentrated on some set
E ⊂ Rd if
n=1 ⊂ R2d is relatively separated and D+(Λ) ≤ 1.
ZE
g(x)2dx ≥ (1 − ǫ2)kgk2.
Since
Zx−an≥r
fn(x)2 dx ≤ r−2sZx−an≥r
x − an2sfn(x)2 dx ≤ r−2sS ,
there exists r = r(ǫ) such that fn is ǫ-concentrated on B(an, r) uniformly in n. Likewise
and denote QR = Q((x0, ξ0), R). Remark that if (an, bn) ∈ QR, then fn is ǫ-concentrated
bfn is ǫ-concentrated on B(bn, r) for every n. We fix (x0, ξ0) ∈ Rd×Rd, consider any R > 0
on B(x0, R + r), and bfn is ǫ-concentrated on B(ξ0, R + r), where r = r(ǫ) as above.
4
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
We now apply a standard estimate of the trace of a time-frequency restriction operator
to conclude that D+(Λ) ≤ 1, see [RS].
Let F be the Fourier transform and PE be the projection operator PEf = χE f
(multiplication of f by the characteristic function of E). The phase-space restriction
operator is defined by
L = PB(x0,R+r)(F −1PB(ξ0,R+r)F )PB(x0,R+r) = P1P2P1 .
It is well-known, see for example [FS], that
tr(L) = B(x0, R + r)B(ξ0, R + r) = QR+r.
For each fn such that (an, bn) ∈ QR, we have
kfn − Lfnk ≤ kfn − P1fnk + kP1kkfn − P2fnk + kP1P2kkfn − P1fnk ≤ 3ǫkfnk.
Now let {gn} be the biorthogonal basis for {fn}, i.e., (fn, gm) = δnm. Then
tr(L) ≥ X(an,bn)∈QR
(Lfn, gn) ≥ X(an,bn)∈Qr
((fn, gn) − (fn − Lfn, gn)) ≥ (1−3Cǫ)card(Λ∩QR),
where C = supn kfnkkgnk < ∞ (since {fn} is a Riesz basis). Thus
card(cid:16)Λ ∩ Q((x0, ξ0), R)(cid:17) ≤ (1 − 3Cǫ)−1Q((x0, ξ0), R + r).
Taking the limit R → ∞, we obtain D+(Λ) ≤ (1 − 3Cǫ)−1 for every ǫ > 0, and thus
D+(Λ) ≤ 1, and Λ is relatively separated.
(cid:3)
Remark. If {fn}∞
n=1 is a frame that has phase-space localization of magnitude s > 0 at
points {(an, bn)}∞
n=1 and satisfies kfnk2 ≤ C, then it is still true that Λ = {(an, bn)}∞
n=1 is
a relatively separated set and that D+(Λ) < ∞. This follows by compactness arguments,
see Theorem 3.5 in [JP] for a similar result in dimension d = 1.
Lemma 2. Suppose that {fn}∞
n=1 is a frame for L2(Rd) with supn∈N kfnk2 = C < ∞. If
s > d and {fn}n has phase-space localization of magnitude s at points {(an, bn)}n, then
Λ = {(an, bn)}∞
n=1 is relatively separated and D−(Λ) ≥ 1.
We remark that the Lemma does not hold for s = d. This can be seen from the
construction of an orthonormal basis in [B].
Proof. Let K(y, l) denote the cube with center y ∈ Rq and the side length 2l,
K(y, l) = {z ∈ Rq : ky − zk∞ < l},
where kzk∞ = max1≤s≤q zs, z = (z1, . . . , zq) ∈ Rq.
Fix ǫ > 0 and choose δ in the open interval (d/s, 1). This is possible by the hypothesis
s > d.
Step 1. An estimate for the coefficients (ψ, fm) of a localized function.
Assume that ψ ∈ L2(Rd), kψk2 = 1, ψ is ǫ-concentrated on K(a, R − Rδ) and its Fourier
transform is supported on K(b, R − Rδ). Set η = ψ(1 − χK(a,R−Rδ)), so that kηk2 ≤ ǫ.
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
5
If kam − ak∞ > 2kR, then the following estimate holds:
(ψ, fm)2 ≤ 2(ψ(1 − χK(a,R−Rδ )), fm)2 + 2(cid:18)ZK(a,R−Rδ)
≤ 2(η, fm)2 + 2(cid:18)((2k − 1)R + Rδ)−sZK(a,R−Rδ)
ψ(x)fm(x)dx(cid:19)2
x − amsψ(x)fm(x)dx(cid:19)2
≤ 2(η, fm)2 + 2((2k − 1)R + Rδ)−2sS .
If kbm − bk∞ > 2kR, then
(ψ, fm)2 = (bψ,cfm)2 =(cid:18)ZK(b,R−Rδ )
bψ(ξ)bfn(ξ)dξ(cid:19)2
≤ ((2k − 1)R + Rδ)−2sS .
Let M0 = {n : (an, bn) ∈ K(a, R) × K(b, R)} and M be the complement of M0,
M = {n : (an, bn) 6∈ K(a, R) × K(b, R)}. We further partition M into the sets Mk as
follows:
Mk = {n : max(kan − ak∞, kbn − bk∞) ∈ [2kR, 2k+1R)},
k ≥ 0.
Since D+(Λ) < ∞ by Lemma 1 (see also the remark after the lemma), we find that
card(Mk) ≤ C1(2kR)2d for some constant C1 large enough.Thus
Xm∈M
(ψ, fm)2 =
(ψ, fm)2
∞Xk=0 Xm∈Mk
≤ 2 Xm∈M
(η, fm)2 + 2S
∞Xk=0
C122kdR2d(cid:0)(cid:0)2k − 1(cid:1) R + Rδ(cid:1)−2s
≤ 2Bkηk2
2 + 2SC1R2d−2sδ + 2SC1R2d−2s
∞Xk=1
22kd(2k − 1)−2s ,
where B is the upper frame bound of {fn}n. Since s > d by assumption, the last sum
converges. Further, sδ > d and, by choosing R large enough, the second and third terms
can be made arbitrarily small. Given ǫ > 0 and δ ∈ (d/s, 1), we find that
(3)
Xm∈M
(ψ, fm)2 ≤ C 2
0 ǫ2
for R ≥ R0(ǫ, δ, Λ, S) ,
with the constant C0 depending only on the frame bound B of {fn}n.
Step 2. Comparison with a basis of prolate spheroidal functions. For given
ǫ > 0, δ ∈ (d/s, 1), and R ≥ R0(ǫ, δ, Λ, S), we now consider those prolate spheroidal
functions φ1, ..., φN with N = N(R) that are ǫd−1/2 concentrated on (−R + Rδ, R − Rδ)
and whose Fourier transforms are supported on (−R + Rδ, R − Rδ). We refer the reader
to [SP] and [LP] for definitions and properties of these functions. According to [LP] the
number of φj with these concentration properties satisfies limR→∞ N(R)R−2 = 1.
6
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
In higher dimensions we take tensor products of phase-space shifts of these prolate
spheroidal functions. Let σ = (n1, . . . , nd) ∈ {1, 2, . . . , N(R)}d and define
ψσ(x) =
dYj=1
e−2πibj xj φnj (xj − aj),
then we obtain an orthonormal set of N d functions {ψσ}σ that are ǫ-concentrated on
K(a, R − Rδ) and whose Fourier transforms are supported on K(b, R − Rδ).
Now let {gn}n be the dual frame of {fn}n. If A > 0 is the lower frame bound of {fn}n,
then we have
kXn
cngnk2
2 ≤ A−1kck2
2
for every c ∈ ℓ2 .
Step 3. Density estimate. We now follow the argument of Ramanathan and
Steger in [RS]. Let S be the orthogonal projection of L2(Rd) onto Ψ = span{ψσ : σ ∈
{1, . . . , N(R)}d} and T be the orthogonal projection onto G = span{gn : n ∈ M0}. We
consider U : Ψ → Ψ, U = S ◦ T . For each ψ ∈ Ψ we obtain
kψ − Uψk2 = kS(ψ − T ψ)k2 ≤ kψ − T ψk = inf
g∈G
(ψ, fn)gnk
= k Xm∈M
(ψ, fm)gmk ≤ A−1/2(cid:16) Xm∈M
kψ − gk ≤ kψ − Xn∈M0
(ψ, fm)2(cid:17)1/2
.
Since each basis function ψσ is in Ψ and satisfies the concentration assumptions from
Step 1, the estimate (3), implies that
kψσ − Uψσk2 ≤ A−1/2C0ǫ .
Consequently,
tr(U) ≥Xσ
(Uψσ, ψσ) =Xσ (cid:16)kψσk2
2 − (ψσ − Uψσ, ψσ)(cid:17) ≥ (1 − A−1/2C0ǫ)N(R)d .
On the other hand, since U is the composition of two projections, all eigenvalues of U
belong to (0, 1), and therefore tr(U) ≤ rank(U) ≤ dim(G). Thus
(1 − A−1/2C0ǫ)N(R)d ≤ tr(U) ≤ card(Λ ∩ K(a, R) × K(b, R)) .
We now use the definition of the Beurling density with cubes in R2d instead of balls,
and obtain
D−(Λ) = lim
R→∞
inf
(a,b)∈R2d
card(cid:0)Λ ∩ K(a, R) × K(b, R)(cid:1)
N(R)d
R2d = 1 − A−1/2C0ǫ .
R2d
≥ (1 − A−1/2C0ǫ) lim
R→∞
As ǫ > 0 was arbitrary, we conclude that D−(Λ) ≥ 1.
(cid:3)
Combining Lemmas 1 and 2, we obtain the density result for localized Riesz bases
(recall, however, that our aim is to prove that there are no such bases).
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
7
Corollary. If s > d and {fn}∞
calization of magnitude s at points {(an, bn)}∞
D(Λ) = D+(Λ) = D−(Λ) = 1.
n=1 is a Riesz basis for L2(Rd) that has phase-space lo-
n=1, then the density of Λ = {(an, bn)}n is
3. Uncertainty identity
We first prove Theorem 1 under the additional condition that the dual basis is also
well-localized. The proof extends Battle's elegant proof of the Balian-Low theorem [Bat]
and rediscovers Steger's argument mentioned by Bourgain in [B] (see the quote above).
The core of the argument is the following uncertainty identity (the canonical commu-
tation relations)
(xf, ∇g) + (∇f, xg) =
dXj=1
(xjf, ∂g
∂xj
) + ( ∂f
∂xj
, xjg) = −d(f, g),
which holds provided that f, g, ∂f
∂xj
, ∂g
∂xj
, xjg, xjf ∈ L2(Rd) for j = 1, . . . , d.
Lemma 3. Assume that {fn}∞
{gn}∞
n=1. If the bases satisfy the localization estimates
n=1 is a Riesz basis for L2(Rd) with the biorthogonal basis
(a) RRd x − an2sfn(x)2dx +RRd ξ − bn2sbfn(ξ)2dξ ≤ S2 < ∞ for every n;
(b) RRd x − an2sgn(x)2dx +RRd ξ − bn2sbgn(ξ)2dξ ≤ T 2 < ∞ for every n; and
n=1 ⊂ R2d is relatively separated and 0 < D−(Λ) ≤ D+(Λ) < ∞,
(c) Λ = {(an, bn)}∞
then s ≤ d.
Proof. We assume that s > d and use the uncertainty identity to derive a contradiction
−−−−→
(xf, g) ∈ Cd for the vector with components
from (a)−(c). In the following we will write
−−−−→
(∇f, g) = ( ∂f
, g)d
(xjf, g), j = 1, . . . , d. Likewise
∂xj
j=1.
Step 1. An estimate for non-diagonal coefficients. Condition (a) implies xfn ∈
L2(Rd)d, and then the sequence of vectors
cn
m =
−−−−−−→
(xfn, gm) = (xjfn, gm)d
j=1 ∈ Cd
is well defined. By the biorthogonality condition, for m 6= n
−−−−−−−−−−−→
((x − an)fn, gm).
Since {gm} is a Riesz basis, assumption (b) implies that
cn
m =
cn
m2 ≤ Bkx − an fnk2
2 ≤ BS2 .
Xm:m6=n
Xn:n6=m
Next, since cn
m =
−−−−−−→
(xgm, fn), we also have
cn
m2 ≤ Bkx − am gmk2
2 ≤ BT 2.
Here B is the upper basis constant for both Riesz bases {gm}m and {fn}n.
The coefficients
dn
m =
−−−−−−→
(ζbfn,cgm) = (2πi)−1−−−−−−→
(∇fn, gm)
8
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
enjoy similar properties.
Step 2. Commutation relations. We now apply the uncertainty identity to each
pair {fn, gn} and obtain
(4)
(xfn, gm) ·
d = −Xm (cid:16)−−−−−−→
= −2πiXm
m · dm
(cn
n − dn
m · cm
n ),
−−−−−−→
(∇gn, fm) +
−−−−−−→
(∇fn, gm) ·
−−−−−−→
(xgn, fm)(cid:17)
where λ · µ =Pd
j=1 λjµj is the standard scalar product in Cd.
For each R > 0 define N (R) = {n : an ≤ R, bn ≤ R} and N(R) = card N (R). Now
we sum up the identities (4) for all n ∈ N (R),
(−cn
m · dm
n + dn
m · cm
n )
d
2πi
(5)
N(R) = Xn∈N (R)Xm
dXj=1
= Xn,m∈N (R)
(−(cn
m)j(dm
n )j + (dn
m)j(cm
n )j) + Xn∈N (R) Xm6∈N (R)
(−cn
m · dm
n + dn
m · cm
n ).
Clearly, the first sum equals zero. We will derive a contradiction by showing that the
second sum possesses a slower growth than N(R). We divide the necessary estimates
into several steps.
Step 3. Points (an, bn) near the boundary. To estimate the second sum, we
partition N (R) into two sets,
N (R) = N (R − r) ∪(cid:0)N (R) \ N (R − r)(cid:1) ,
where r = Rδ for some δ ∈ (d/s, 1).
First, for n ∈ R(R − r, R) = N (R) \ N (R − r) we get
Xn∈R(R−r,R) Xm6∈N (R)
cn
mdm
n ≤
(6)
≤ Xn∈R(R−r,R)(cid:16) Xm:m6=n
cn
m2(cid:17)1/2(cid:16) Xm:m6=n
dm
n 2(cid:17)1/2
≤ (N(R) − N(R − r))BS1T1 .
The sum of dn
mcm
n admits the same estimate.
Step 4. Further partition of N (R) for interior (an, bn). Next we partition the
complement of N (R) into the rings Nk = N (Rk+1) \ N (Rk) where Rk = 2kR, k ≥ 0.
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
9
Then for each n ∈ N (R − r) we get
Xm6∈N (R)
cn
mdm
n =
≤
cn
mdm
n
∞Xk=0 Xm∈Nk
∞Xk=0 Xm∈Nk:am>Rk
≤(cid:16) ∞Xk=0 Xm∈Nk:am>Rk
+(cid:16) Xm:m6=n
cn
cn
mdm
n +
∞Xk=1 Xm∈Nk:bm>Rk
n 2(cid:17)1/2
n 2(cid:17)1/2
dm
dm
.
cn
m2(cid:17)1/2(cid:16) Xm:m6=n
m2(cid:17)1/2(cid:16) ∞Xk=0 Xm∈Nk:bm>Rk
cn
mdm
n
+
Step 5. Main estimate. Now we write down an estimate for cn
m when an < R − r
and am > Rj. Set h(j)
n (x) = (x − an)jfn(x)(1 − χB(R−r/2)(x)), j = 1, 2, ..., d. Then
kh(j)
n k2
2 ≤Zx>R−r/2
(x − an)j2fn(x)2dx.
Further, for x ≥ r/2 and an ≤ R − r, we have x − an ≥ r/2, and therefore
kh(j)
n k2
2 ≤ (r/2)2−2sZRd
(7)
dXj=1
Then we have
x − an2sfn(x)2dx ≤ (r/2)2−2sS2.
(8)
cn
m2 =
≤
dXj=1
(cid:0)(x − an)jfn, gm(cid:1)2
dXj=1(cid:0)2((x − an)jfnχB(R−r/2), gm)2 + 2(h(j)
n , gm)2(cid:1)
≤ 2S2kgmχB(R−r/2)k2
2 + 2
dXj=1
(h(j)
n , gm)2
(9)
≤ 2S2T 2(Rk − R + r/2)−2s + 2
dXj=1
(h(j)
n , gm)2.
And since {gm}m is a Riesz basis,
(10)
Xm
(h(j)
n , gm)2 ≤ Bkh(j)
n k2
2 .
10
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
Summing up the estimates (8) over all k and all m ∈ Nk such that am > Rk and taking
into account (7) and (10), we obtain
(11)
∞Xk=0 Xm∈Nk:am>Rk
cn
m2 ≤ 2S2T 2
∞Xk=0
N(Rk+1)(Rk − R + r/2)−2s + B(r/2)2−2sS2.
n where the summation is over all k and
m, where the sums are over all k
and m ∈ Nk such that bm > Rk and over all k and m ∈ Nk such that am > Rk, by
We can derive similar estimates ofPkPm dm
m ∈ Nk such that bm > Rk by using the localization inequality for bgn. Likewise, we
obtain the estimates for PkPm cm
using the localization conditions on gn and bfn.
Step 6. Comparison of the densities. Finally we combine the inequality obtained
in Step 4 with (11) and similar inequalities with other combinations of indices. Then we
obtain for every n ∈ N (R − r)
n and PkPm dn
Xm6∈N (R)
(cn
mdn
m + cm
n dm
n ) ≤(cid:16)C1r2−2s + C2
∞Xk=0
N(Rk+1)(Rk − R + r/2)−2s(cid:17)1/2
,
where C1 and C2 depend on S, T, B, s, and supn kfnk2 and supn kgnk2. Assumption (c)
(the estimate of the upper density D+(Λ) < ∞) implies that
N(Rk+1) ≤ D122d(k+1)R2d
and N(R − r) ≤ D1(R − r)2d
for some D1 > 0 and all R large enough. Then for R large enough we obtain
(cn
mdm
n + cm
Xn∈N (R−r) Xm6∈N (R)
N(R − r)(cid:16)C1r2−2s + C2D122dR2d(r/2)−2s + C2D1R2d
n dn
m) ≤
∞Xk=1
22d(k+1)(2kR − R + r/2)−2s(cid:17)1/2
≤ C(R − r)2dRdr−s,
where C depends on S, T, B, s, D1, d as well as supn kfnk2 and supn kgnk2.
To finish the proof we recall that r = Rδ and δ ∈ (d/s, 1). Observe that for r large
enough the estimate of the upper density implies N(R) − N(R − r) ≤ D1R2d−1r (we just
cover the set Q(0, R) \ Q(0, R − r) by cubes with side length r). Now, combining (5),
(6), and the last inequality, we obtain
N(R) ≤ C3(cid:16)N(R) − N(R − r) + R3d−δs(cid:17) ≤ C4(R2d−1+δ + R3d−s).
If we now let R go to infinity, we see that
D−(Λ) ≤ lim
R→∞
N(R)
R2d ≤ C4 lim
R→∞(cid:16)Rδ−1 + Rd−s(cid:17) = 0 .
This conclusion contradicts the assumption (c) that the lower density estimate D−(Λ)
is strictly positive.
(cid:3)
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
11
Lemma 3 concludes the proof of Theorem 1 for the case of an orthonormal basis. For
a Riesz basis we are not able to prove that the phase-space localization of magnitude
s (condition (a) of the lemma) with s > d implies the required localization for its
biorthogonal basis (stated in (b)). For the general case a more complicated argument is
presented in the next sections.
4. Some preliminaries on modulation spaces
The proof of Theorem 1 for Riesz bases requires some tools from time-frequency anal-
ysis. We give a minimalistic account of the required facts on the short-time Fourier
transform and modulation spaces. The reader can find the details and a much more
general theory of modulation spaces in [G1].
Short-time Fourier Transform. For (a, b) ∈ R2d we write
π(a, b)f (t) = e2πib·tf (t − a)
for the phase-space shift of a function f on Rd. Let g(x) = 2−d/4e−πx2
be the normalized
Gaussian function on Rd. We consider the short-time Fourier transform of a function
φ ∈ L2(Rd) with respect to g, [G1, Chapter 3]
Vgφ(x, ξ) = (φ, π(x, ξ)g) =ZRd
φ(t)g(t − x)e−2πit·ξdt,
x, ξ ∈ Rd.
The inversion formula for the short-time Fourier transform yields
φ(t) =ZZR2d
Vgφ(x, ξ)e2πiξ·tg(t − x)dx,
for every φ ∈ L2(Rd), with a weak interpretation of the vector-valued integral.
Modulation Spaces. For each s ≥ 0 let
L2
s(Rm) = {f ∈ L2(Rm) : kf k2
L2
s
f (x)2(1 + x)2sdx < ∞} .
=ZRm
The modulation space M 2
s (Rd) is defined by
M 2
s (Rd) = {φ ∈ L2(Rd) : kφkM 2
s = kVgφkL2
s(R2d) < ∞}.
The following norm equivalence identifies the modulation space M 2
Lebesgue space L2
s, see [G1, Prop. 11.3.1, 12.1.6]:
s ∩ F L2
s with the Fourier-
The adjoint operator V ∗
s ≤ kφkL2
c1kφkM 2
g of the short-time Fourier transform is defined on L2(R2d) by
s ≤ c2kφkM 2
s .
s + kbφkL2
g F (t) =ZZR2d
V ∗
F (x, ξ)e2πiξ·tg(t − x)dxdξ =ZZR2d
F (x, ξ)π(x, ξ)g(t)dxdξ .
If F ∈ L2
s(R2d), then by [G1, Prop. 11.3.2] V ∗
s (Rd) and
g F ∈ M 2
s ≤ CkF kL2
s.
kV ∗
g F kM 2
(12)
12
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
We note that the phase-space localization of magnitude s can be rephrased as
sup
n∈N
inf
(a,b)∈R2d
kπ(a, b)fnkM 2
s < ∞ .
Amalgam spaces. We define the amalgam space W (L2
the space of all continuous function on R2d for which the norm
s) ⊂ L2
s(R2d) ∩ L∞(R2d) as
kF k2
W (L2
s) = Xk,n∈Zd
sup
F (x + k, ξ + n)2 (1 + k + n)2s
x,ξ∈[0,1]d
is finite. For s = 0, kF kW (L2) ≥ kF k2 obviously. The continuity of F implies the
existence of points xkn, ξkn ∈ [0, 1]d, such that
kF k2
W (L2
s) = Xk,n∈Zd
F (k + xkn, n + ξkn)2(1 + k + n)2s .
The definition of W (L2
relatively separated, z ∈ R2d, and F ∈ W (L2
s) implies the following sampling inequality: If Λ = {λn} ⊆ R2d is
s), then
≤ sup
k∈Z2d
card(cid:0)Λ ∩ (k + [0, 1]2d)(cid:1) kF kWL2
s
.
(13)
(14)
(15)
(16)
(17)
(cid:16)Xn
F (z + λn)2 (1 + z + λn)2s(cid:17)1/2
s(Rd) with bφ ∈ L2
kVgφkW (L2
s) ≤ CkVgφkL2
s = CkφkM 2
s .
The following important inequality links modulation spaces with amalgam spaces: For
every φ ∈ L2
s(Rd) we have, e.g., by [G1, Theorem 12.2.1],
5. Basis modification
To finish the proof of Theorem 1, we will modify a given Riesz basis {fn}n for L2(Rd)
that has phase-space localization of magnitude s > d into a Riesz basis with much better
localization properties. The argument in this section may be of independent interest and
can also be used to prove positive results about frames and bases.
Proposition 1. Assume that {fn}∞
n=1 is a Riesz basis for L2(Rd) that satisfies
for some s > d. Then there exists a Riesz basis {hn}∞
sup
n∈N(cid:16)ZRd
n∈N(cid:16)ZRd
sup
x − an2sfn(x)2 +ZRd
x − an2thn(x)2 +ZRd
ξ − bn2sbfn(ξ)2(cid:17) < ∞
ξ − bn2tchn(ξ)2(cid:17) < ∞
n=1 that satisfies
for every t > 0.
Proof. The new basis is obtained by a modification of {fn}n. We use the inversion
formula for the short-time Fourier transform and truncate it. In the language of time-
frequency analysis we apply a localization operator to fn. Precisely, let R > 0 and
Q(R) = Q(0, R) = B(0, R) × B(0, R) ⊂ R2d. Then the localization operator AR is
defined by
ARf (t) =ZZQ(R)
Vgf (x, ξ)e2πiξ·tg(t − x)dxdξ,
f ∈ L2(Rd) .
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
13
Intuitively, ARf is the part of f that is concentrated on the set Q(R) in the phase space.
For more on localization operators see for instance [W, CG].
We recast the assumption as follows: {fn}n is a Riesz basis for L2(Rd), fn(x) =
e2πibnxφn(x − an), s > d and
kφnk2
M 2
s
sup
n∈N
≤ S < ∞.
We now define
(18)
ψn = ARφn =ZZQ(R)
Vgφn(x, ξ)π(x, ξ)g dxdξ
and the modified basis hn(x) = hR
n (x) = e2πibnxψn(x − an).
Claim. For R large enough {hn}∞
To prove the claim, it suffices to show that for every ǫ > 0 there exists R such that
n=1 is a Riesz basis for L2(Rd).
for hn = hR
n and every sequence {cn}n ∈ ℓ2 the inequality
≤ ǫkcnk2
cn(fn − hn)(cid:13)(cid:13)(cid:13)2
(cid:13)(cid:13)(cid:13)Xn
cnfnk2 − kXn
cnhnk2 ≥ kXn
kXn
holds. If {fn}n is a Riesz basis with the lower basis constant A > 0, then
cn(fn − hn)k2 ≥ (A − ǫ)kck2 ,
and so {hn}n is a Riesz basis.
Using once again the crucial assumption s > d, we now choose a number σ such that
d < σ < s. Using the inversion formula for the short-time Fourier transform and (18),
we write
φn − ψn =ZR2d(cid:0)1 − χQ(R)(x, ξ)(cid:1)Vgφn(x, ξ)π(x, ξ)g dxdξ ,
and estimate the M 2
σ -norm of φn − ψn with (12) as
kφn − ψnk2
M 2
σ
≤ZR2d(cid:0)1 − χQ(R)(x, ξ)(cid:1)Vgφn(x, ξ)2(1 + x + ξ)2σ dxdξ
≤ (1 + R)2(σ−s)ZR2d(cid:0)1 − χQ(R)(x, ξ)(cid:1)Vgφn(x, ξ)2(1 + x + ξ)2s dxdξ
≤ C(1 + R)2(σ−s)kφnk2
≤ C(1 + R)2(σ−s)S .
M 2
s
Choosing now R large enough, we have
kφn − ψnkM 2
σ < ǫ
for all n .
14
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
Then we have, with (13) and a suitable choice of points (xk,m, ξk,m) ∈ [0, 1]2d, that
(cid:13)(cid:13)(cid:13)Xn
2
2
2
2
2
≤ C(cid:13)(cid:13)(cid:13)Xn
cn(fn − hn)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)Xn
= X(k,m)∈Z2d(cid:12)(cid:12)(cid:12)Xn
≤ X(k,m)∈Z2d(cid:16)Xn
≤ X(k,m)∈Z2d(cid:16)Xn
cnVg(fn − hn)(cid:13)(cid:13)(cid:13)
cnVg(fn − hn)(cid:13)(cid:13)(cid:13)
cnVg(fn − hn)(xk,m + k, ξk,m + m)(cid:12)(cid:12)(cid:12)
cn Vg(φn − ψn)(xk,m + k − an, ξk,m + m − bn)(cid:17)2
(1 + k − an + m − bn)−2σ(cid:17) ×
W (L2)
2
uniformly bounded independent of k and m. Thus we obtain
×(cid:16)Xn
cn2Vg(φn − ψn)(xk,m + k − an, ξk,m + m − bn)2(1 + k − an + m − bn)2σ(cid:17) .
Since σ > d and Λ is relatively separated, the sum Pn(1 + k − an + m − bn)−2σ is
cn(fn − hn)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)Xn
cn2Xk,m
≤ CXn
Vg(φn − ψn)(xk,m + k − an, ξk,m + m − bn)2(1 + k − an + m − bn)2σ .
≤
2
2
By (14) and (15) we estimate further that
Xk,m
Vg(φn − ψn)(xk,m + k − an, ξk,m + m − bn)2(1 + k − an + m − bn)2σ
≤ C2kVg(φn − ψn)k2
σ) ≤ C3kφn − ψnk2
M 2
σ
W (L2
< C3ǫ2 .
Collecting all estimates, we arrive at
(cid:13)(cid:13)(cid:13)Xn
cn(fn − hn)(cid:13)(cid:13)(cid:13)
2
2
≤ C3ǫ2Xn
cn2 = C3kck2
2 ǫ2 .
Consequently, {hn}n is a Riesz basis for L2(Rd).
Finally, applying (12) once again, we obtain, for arbitrary t > 0,
kψnkL2
t
+ k ψnkL2
t
≤ CkψnkM 2
t
≤ C ′kVgφnχQ(R)kL2
t
≤ CtRt ,
which is (17).
(cid:3)
REMARK: The construction of ψn implies that ψn(t) ≤ Ce−αt2
for some α, β, C > 0. Thus the perturbed basis belongs to the Gelfand-Shilov space
S1/2,1/2, the smallest space of test functions that is invariant under the Fourier transform.
and cψn(ξ) ≤ Ce−βξ2
To complete the proof of Theorem 1 we will show that the biorthogonal basis {fhn}n
satisfies (16) for some s large enough and then apply Lemma 3.
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
15
The modified basis {hn}n possesses enough phase-space localization so that the theory
of localized frames [FG,G2] is applicable. We say that a frame {hλ : λ ∈ Λ} is s-localized
over the index set Λ ⊆ Rd, if its Gramian satisfies
(19)
(hµ, hλ) ≤ C(1 + λ − µ)−s
for all λ, µ ∈ Λ .
The main result about localized frames asserts that the dual frame possesses the same
type of localization. Specifically we need the following result taken from [FG, Thm. 1.1,
Cor. 3.7]
Proposition 2. Let Λ ⊆ Rd be a relatively separated set and {hλ : λ ∈ Λ} be a frame
for L2(Rd). Assume that {hλ} is s-localized for s > d. Then the (canonical) dual frame
is also s-localized, i.e.,
(20)
for all λ, µ ∈ Λ .
(fhµ,fhλ) ≤ C ′(1 + λ − µ)−s
For a Riesz basis this result can be proved directly. The Gramian matrix G of the
basis with entries (hµ, hλ) is invertible with inverse (G−1)λµ = (fhµ,fhλ). By a theorem
of Jaffard [Jaf] the polynomial off-diagonal decay is preserved under inversion, whence
follows the statement of the proposition (for a Riesz basis).
To apply Proposition 2, we need to compare the phase-space localization of magnitude
s in Bourgain's uncertainty principle with the localization defined by the off-diagonal
decay of the Gramian.
Lemma 4. (i) If {hn}n is a Riesz basis that has phase-space localization of magnitude
s > 0, then {hn}n is s-localized over the index set Λ in the sense of (19).
(ii) If {hn}n is a Riesz basis that has phase-space localization of magnitude s > 3d,
then the biorthogonal basis {fhn}n has phase-space localization of magnitude t for any
t ∈ (d, s).
Proof. (i) We choose the set Λ = {(an, bn)}n ⊂ R2d as the appropriate index set. By
Lemma 1, Λ is relatively separated. Then
(hn, hm) ≤
≤ sup
x∈Rd(cid:16)(1 + x − am)(1 + x − an)(cid:17)−s ZRd
n∈NZRd
hn(x)(1 + x − an)shm(x)(1 + x − am)s dx
≤ C(1 + am − an)−s sup
hn(x)2(1 + x − an)2s dx = (1 + am − an)−sS2 .
The argument for the Fourier transforms yields that
By combining both estimates we arrive at
(hn, hm) = (chn,chm) ≤ C(1 + bm − bn)−sS2 .
(hn, hm) ≤ C(1 + λ − µ)−s
for all λ, µ ∈ Λ ,
in other words, {hn} is s-localized over Λ.
16
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
(ii) By (i) {hn} is s-localized on Λ ⊆ R2d and s > 3d. Therefore Proposition 2 applies
and the dual frame is also s-localized as in (20). Now we have
Next let d < t < s − 2d (which is possible by s > 3d). A straightforward calculation and
(17) give
fhn =Xm
x − an2thm(x)2dx ≤ZRd
ZRd
(fhn,fhm)hm.
Ct(cid:0)x − am2t + an − am2t(cid:1)hm(x)2dx
≤ C(S + an − am2t) ≤ C ′(1 + λn − λm)2t.
With the triangle inequality for the L2
t -norm we obtain
(cid:16)ZRd
x − an2thn(x)2dx(cid:17)1/2
Xm
(hn, hm)hm(x)2x − an2t dx(cid:17)1/2
=(cid:16)ZRd
≤ CXm
≤ CXm
(hn, hm)(1 + λn − λm)t
(1 + λn − λm)t−s .
Since Λ is relatively separated and t − s < −2d, the last sum is uniformly bounded.
Similar estimates hold for the Fourier transforms. Consequently, the dual basis is
(cid:3)
localized in the sense of (16) for t ∈ (d, s).
We now can finish the proof of Theorem 1 for Riesz bases.
We start with a Riesz basis {fn}n that satisfies (16) for some s > d. Then we modify
this basis by means of Proposition 1 to a Riesz basis {hn}n that satisfies the localization
estimates (16) for all t > 0. Finally, Lemma 4 guarantees that the dual basis {fhn}n also
satisfies the localization estimates (16) for all t > 0. Thus all assumptions of Lemma 3
are satisfied whence we conclude that the localization of {hn}n is t ≤ d. This is a
contradiction to the construction of {hn}n. This means that the original basis {fn}n
cannot satisfy the strong localization estimate s > d, and the proof of Theorem 1 is
complete.
6. Sampling in Bargmann-Fock spaces and Concluding Remarks
F =RCd F (z)2 e−πz2
We finally give an application of Theorem 1 to several complex variables.
Recall that the Bargmann-Fock space F consists of all entire functions on Cd with
norm kF k2
dz. The Bargmann-Fock space possesses the reproducing
kernel Kw(z) = eπ ¯w·z for w, z ∈ Cd, so that F (w) = (F, Kw). Consequently, a set
{Kλ : λ ∈ Λ} is a Riesz basis for F , if and only if Λ ⊆ Cd is simultaneously sampling
F, and for every c ∈ ℓ2(Λ) there
and interpolating for F , i.e., Pλ F (λ)2e−πλ2 ≍ kF k2
exists a (unique) F ∈ F , such that F (λ)e−πλ2/2 = aλ.
The following result is an immediate consequence of Theorem 1.
Theorem 2. The Bargmann-Fock space does not admit a set Λ ⊆ Cd that is simultane-
ously sampling and interpolating.
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
17
Proof. We use the Bargmann transform defined as
Bf (z) = F (z) = 2d/4e−πz2/2ZRd
f (t)e−πt·te2πt·zdt
z ∈ Cd ,
and translate Theorem 1 into a statement of complex analysis. The Bargmann trans-
form is unitary from L2(Rd) onto F and maps the phase-space shifts of the Gaussian
e−2πiw2·xe−π(x−w1)2
= π(w1, −w2)g(x) to the reproducing kernel eiαKw e−w2/2 for some
phase factor c = 1. Thus Λ ⊆ Cd is a set of sampling and interpolation, if and only if
{Kλ : λ ∈ Λ} is a Riesz basis for F .
Clearly, the Gaussian satisfies the localization condition (16) for all s > 0. By Theo-
rem 1 a set of phase-space shifts of the Gaussian cannot from a Riesz basis for L2(Rd).
Consequently, no set Λ ⊆ Cd can be simultaneously sampling and interpolating.
(cid:3)
REMARK: Theorem 2 is a statement of complex analysis. Indeed, in dimension d = 1
is it well-known and can be proved with classical methods from complex variables. In
higher dimensions the result was expected, but seems to have been open so far. A proof
with different methods has been announced in [AFK].
Final remarks.
It is interesting to compare the critical value of the localization
parameter s in higher dimensions with the higher dimensional versions of the Balian-
Low theorem.
It is known that for d = 1 and every s < 2 there exists a function f
with
ZR
ZR
xsf 2 < ∞,
ξs f2 < ∞,
such that {e2πintf (x − m)}n,m∈Z is an orthogonal basis for L2(R). (A more precise result
is obtained in [BCGP], we refer the reader to [Jan] also.) Thus in one-dimensional
setting the restrictions on the localization properties of an arbitrary orthogonal basis
and of a Gabor system can be observed only at one point of our scale, p = 2. The
situation changes drastically when we consider higher dimensional spaces. There exists
an orthogonal basis {fn} for L2(Rd) that satisfies
but for every orthogonal basis of the form {e2πintf (x − m)}n,m∈Zd a multidimensional
version of Balian-Low theorem (see [CP] and references therein) implies
n∈N(cid:16) inf
a∈RdZRd
sup
x − a2dfn(x)2 + inf
b∈RdZRd
a∈RdZRd
inf
sup
n∈N
x − a2fn(x)2 = ∞ or
sup
n∈N
ξ − b2dbfn(ξ)2(cid:17) < ∞,
b∈RdZRd
inf
ξ − b2bfn(ξ)2 = ∞.
The reason for it could lie in the choice of the radial weight vp = xp. It seems that
might be a more natural weight in
the product weight ws(x) =(cid:16)(1 + x1)...(1 + xd)(cid:17)s
higher dimensions.
There are a (p, q)-version of Bourgain's theorem [BP] and a (p, q) version of the Balian-
Low theorem [Ga] for 1/p + 1/q = 1. It would be interesting to obtain a (p, q)-version of
Theorem 1.
18
KARLHEINZ GR OCHENIG AND EUGENIA MALINNIKOVA
Acknowledgment. K. G. would like to thank the Department of Mathematics of NTNU
Trondheim for its hospitality and great research environment during the research on this pa-
per. Both authors would like to thank the Centre de Recerca Matem`atica Barcelona for ideal
conditions to complete this work.
References
[AFK] G. Ascensi, H. G. Feichtinger, and N. Kaiblinger, Dilation of the Weyl symbol and Balian-Low
Theorem, in preparation.
[BCHL] R. Balan, P. G. Casazza, C. Heil, and Z. Landau, Density, overcompleteness, and localization
of frames. I. Theory, J. Fourier Anal. Appl., 12(2) (2006), 105 -- 143.
[Bat] G. Battle, Heisenberg proof of the Balian-Low theorem, Lett. Math. Phys., 15 (1988), no. 2,
175 -- 177.
[BCGP] J. Benedetto, W. Czaja, P. Gadzin'ski, and A. Powell, The Balian-Low theorem and regularity
of Gabor systems, J. Geom. Anal., 13 (2003), no. 2, 239 -- 254.
[BHW] J. Benedetto, Ch. Heil, and D. Walnut, Differentiation and the Balian-Low theorem, J. Fourier
Anal. Appl., 1 (1995), no. 4, 355 -- 402
[BP] J. Benedetto and A. Powell, A (p, q) version of Bourgain's theorem, Trans. Amer. Math. Soc., 358
(2006), no. 6, 2489 -- 2505.
[B] J. Bourgain, A Remark on the uncertainty principle for Hilbertian basis, J. Funct. Anal., 79 (1988),
136 -- 143.
[CG] E. Cordero and K. Grochenig, Time-frequency analysis of localization operators, J. Funct. Anal.,
205(1) (2003), 107 -- 131.
[CP] W. Czaja and A. Powell, Recent developments in the Balian-Low theorem. Harmonic analysis and
[D]
applications, 79 -- 100, Appl. Numer. Harmon. Anal., Birkhauser, Boston, MA, 2006.
I. Daubechies, The wavelet transform, time-frequency localization and signal analysis, IEEE Trans.
Inform. Theory, 36 (1990), no. 5, 961 -- 1005.
[FS] G. B. Folland and A. Sittaram, The Uncertainty Principle: A Mathematical survey, J. Fourier
Anal. and Appl., 3 (1997), no. 3, 208 -- 238.
[FG] M. Fornasier and K. Grochenig, Intrinsic localization of frames, Constr. Approx., 22(3) (2005),
395 -- 415.
[Ga] S. Z. Gautam, A critical-exponent Balian-Low theorem, Math. Res. Letters, 15 (2008), no. 3,
471-483.
[G1] K. Grochenig, Foundations of Time-Frequency Analysis, Birkhauser, Boston, 2001.
[G2] K. Grochenig, Localization of Frames, Banach Frames and the Invertibility of the Frame Operator,
J. Fourier Anal. and Appl., 10 (2004), no. 2, 105 -- 132.
[GR] K. Grochenig and H. Razafinjatovo, On Landau's necessary density conditions for sampling and
interpolation of band-limited functions, J. London Math. Soc. (2), 54(3) (1996), 557 -- 565.
[H] C. Heil. History and evolution of the density theorem for Gabor frames, J. Fourier Anal. Appl. ,
13(2) (2007), 113 -- 166.
[Jaf] S. Jaffard, Propri´et´es des matrices "bien localis´ees" pr`es de leur diagonale et quelques applications,
Ann. Inst. H. Poincar´e Anal. Non Lin´eaire, 7 (1990), no. 5, 461 -- 476.
[JP] P. Jaming and A. M. Powell, Time-frequency concentration of generating systems, Proc.
Amer. Math. Soc., to appear.
[Jan] A. J. E. M Janssen, A decay result for certain windows generating orthogonal Gabor bases,
J. Fourier Anal. Appl. 14 (2008), no. 1, 1 -- 15.
[LP] H. J. Landau and H. O. Pollak, Prolate spheroidal wave functions, Fourier analysis and uncertainty.
III. The dimension of the space of essentially time- and band-limited signals, Bell System Tech. J.,
41 (1962) 1295 -- 1336.
[L] H. J. Landau, Necessary density conditions for sampling and interpolation of certain entire func-
tions, Acta Math., 117 (1967), 37 -- 52.
LOCALIZATION PROPERTIES OF BASES FOR L2(Rd)
19
[RS] J. Ramanathan and T. Steger, Incompleteness of sparse coherent states, Appl. Comput. Har-
mon. Anal., 2 (1995), no. 2, 148 -- 153.
[SP] D. Slepian and H. O. Pollak, Prolate spheroidal wave functions, Fourier analysis and uncertainty,
I, Bell System Tech. J., 40 (1961), 43 -- 64.
[S] D. Slepian, Prolate spheroidal wave functions, Fourier analysis and uncertainity. IV. Extensions
to many dimensions; generalized prolate spheroidal functions. Bell System Tech. J., 43 (1964)
3009 -- 3057.
[W] M. W. Wong, Localization operators, Seoul National University Research Institute of Mathematics
Global Analysis Research Center, Seoul, 1999.
Faculty of Mathematics, University of Vienna, Nordbergstrasse 15, A-1090 Vienna,
Austria
E-mail address: [email protected]
Department of Mathematics, Norwegian University of Science and Technology, 7491,
Trondheim, Norway
E-mail address: [email protected]
|
1003.1329 | 1 | 1003 | 2010-03-05T19:40:04 | Quantization for an elliptic equation of order 2m with critical exponential non-linearity | [
"math.FA",
"math.AP"
] | On a smoothly bounded domain $\Omega\subset\R{2m}$ we consider a sequence of positive solutions $u_k\stackrel{w}{\rightharpoondown} 0$ in $H^m(\Omega)$ to the equation $(-\Delta)^m u_k=\lambda_k u_k e^{mu_k^2}$ subject to Dirichlet boundary conditions, where $0<\lambda_k\to 0$. Assuming that $$\Lambda:=\lim_{k\to\infty}\int_\Omega u_k(-\Delta)^m u_k dx<\infty,$$ we prove that $\Lambda$ is an integer multiple of $\Lambda_1:=(2m-1)!\vol(S^{2m})$, the total $Q$-curvature of the standard $2m$-dimensional sphere. | math.FA | math |
Quantization for an elliptic equation of order 2m
with critical exponential non-linearity
Luca Martinazzi∗
Centro De Giorgi, Pisa
Michael Struwe
ETH Zurich
[email protected]
[email protected]
November 18, 2018
Abstract
On a smoothly bounded domain Ω ⊂ R2m we consider a sequence
⇁ 0 in H m(Ω) to the equation (−∆)muk =
k subject to Dirichlet boundary conditions, where 0 < λk → 0.
w
of positive solutions uk
λkukemu2
Assuming that
Λ := lim
k→∞ZΩ
uk(−∆)m
ukdx < ∞,
we prove that Λ is an integer multiple of Λ1 := (2m − 1)! vol(S 2m), the
total Q-curvature of the standard 2m-dimensional sphere.
1
Introduction
Given a smoothly bounded domain Ω ⊂ R2m, suppose that for each k ∈ N we
have a smooth function uk > 0 satisfying the equation
(−∆)muk = λkukemu2
k in Ω
(1)
with
uk = ∂ν uk = . . . = ∂m−1
(2)
where 0 < λk → 0 as k → ∞. We assume that (uk) is bounded in H m(Ω).
Hence, after passing to a subsequence and integrating by parts we may assume
that as k → ∞ we have
uk = 0 on ∂Ω,
ν
ZΩ ∇muk2dx = ZΩ
uk(−∆)mukdx = λkZΩ
kemu2
u2
k dx → Λ > 0.
(3)
Note that by elliptic estimates the quantity
kuk := (cid:0)ZΩ ∇muk2dx(cid:1)1/2
= (cid:0)ZΩ Xα=m
∂αuk2dx(cid:1)1/2
defines a norm on the Beppo-Levi space H m
standard Sobolev norm.
0 (Ω) which is equivalent to the
∗The first author was supported by the ETH Research Grant no. ETH-02 08-2 and by the
Italian FIRB Ideas "Analysis and beyond".
1
Generalising previous results by Adimurthi and Struwe [3], Adimurthi and
Druet [1] and Robert and Struwe [11], the first author proved in [8] the following
theorem.
Theorem 1 Let (uk) be a sequence of positive solutions to (1), (2) with 0 <
λk → 0 as k → ∞ and satisfying (3) for some Λ > 0. Then supΩ uk → ∞ as
k → ∞ and there exist a subsequence (uk) and sequences of points x(i)
k → x(i) ∈
Ω, 1 ≤ i ≤ I, for some integer I ≤ CΛ, such that the following is true.
For every 1 ≤ i ≤ I, letting r(i)
k(x(i)
k )2mu2
λk(r(i)
k > 0 be given by
k )emu2
k(x(i)
k ) = 22m(2m − 1)!
(4)
(5)
(6)
and setting
η(i)
k (x) := uk(x(i)
k )(uk(x(i)
k )) + log 2,
k + r(i)
k x) − uk(x(i)
k → ∞ as k → ∞, and
1 + x2 in C2m−1
loc
2
(R2m) as k → ∞.
we have r(i)
k , ∂Ω)/r(i)
k → 0, dist(x(i)
η(i)
k (x) → η0(x) = log
Moreover, for i 6= j there holds
k − x(j)
x(i)
k
r(i)
k
→ ∞ as k → ∞.
such that there holds
In addition, with Rk(x) := inf 1≤i≤I x − x(i)
k(x)emu2
k (x)u2
λkR2m
k there exists a constant C > 0
k(x) ≤ C
(7)
uniformly for all x ∈ Ω, k ∈ N.
Finally uk → 0 in C2m−1
We remark that the function η0 given by (5) satisfies the Q-curvature equa-
(Ω\S), where S = {x(1), . . . , x(I)}.
loc
tion
and
(−∆)mη0 = (2m − 1)!e2mη0
(8)
(2m − 1)!ZR2m
e2mη0dx = ZS2m
QS2mdvolgS2m = (2m − 1)!S2m =: Λ1.
(9)
For a discussion of the geometric meaning of (8) we refer to [4] or to the intro-
duction of [7].
The purpose of this paper is to prove the following quantization result.
Theorem 2 Under the hypothesis of Theorem 1 we have Λ = I ∗Λ1 for some
I ∗ ∈ N\{0}.
2
The analogue of Theorem 2 was proven by O. Druet [5] in dimension 2
(m = 1) and by the second author [13] in dimension 4 (m = 2) in the case of the
Navier boundary condition uk = ∆uk = 0 on ∂Ω. Note that in the latter case
the maximum principle implies that ∆uk ≤ 0 in Ω whereas such an estimate is
not available in the case of the Dirichlet boundary condition.
Quantization results similar to Theorem 2 previously have also been obtained
for concentrating sequences of solutions uk to the Q-curvature equation
(−∆)muk = λke2muk
in Ω ⊂ R2m.
(10)
In the case of the Navier boundary condition, assuming that λk → 0 and
Λ := lim
k→∞ZΩ
λke2muk dx < ∞,
J. Wei [14] proved that when m = 2 and when Ω is convex the quantity Λ
is an integer multiple of Λ1. Moreover concentration points are simple and
isolated, in the sense that x(i) 6= x(j) for i 6= j, and I ∗ = I in the notation of
Theorems 1 and 2 above. Robert and Wei [12] proved the analogous result for a
general domain Ω and in the case of Dirichlet boundary conditions. In [9], the
first author and Petrache generalized the result of Robert and Wei to arbitrary
dimensions.
Equation (1) is more difficult to deal with analytically than equation (10);
the analogous questions whether for a blowing up sequence of solutions to (1)
the concentration points are isolated, simple and stay away from the boundary
are still open, even in dimension 2.
Our paper is organized as follows. In the next section we present the proof
of Theorem 2 in the case when Ω = BR is a ball and each function uk is radially
symmetric. In Section 3 we prove the theorem in the general case. Some useful
technical results are collected in the Appendix. The overall strategy of the proof
is very similar to the approach followed in [13], and some of the results in [13]
can be carried over almost literally to the present setting. Several key steps in
the proof, however, require conceptually new ideas in the case when m ≥ 3.
These ideas also shed new light on the previous approaches in low dimensions
and have a unifying feature.
Throughout the paper the letter C denotes a generic constant independent
of k which can change from line to line, or even within the same line.
2 Proof of Theorem 2 in the radial case
Let Ω = BR = BR(0) and assume that each uk is radially symmetric. By slight
abuse of notation we write uk(x) = uk(r) if x = r. In the notation of Theorem
1 we then have I = 1 and we can choose x(1)
k = 0 for every k > 0. In fact, as
shown in assertion (17) of Lemma 4 below, we have uk(0) = maxΩ uk.
2.1 Strategy of the proof
Set ek := λku2
k and let
kemu2
Λk(r) := ZBr
ekdx, Nk(s, t) := Λk(t) − Λk(s) = ZBt\Bs
ekdx
3
as in [13]. We shall say that the property (Hℓ) is satisfied if there exist sequences
s(0)
k
:= 0 < r(1)
k < s(1)
k < . . . < r(ℓ)
k < s(ℓ)
k ≤ R,
k ∈ N,
such that the following holds:
(Hℓ,1) limk→∞
r(j)
k
s(j)
k
= limk→∞
s(j−1)
k
r(j)
k
= 0 for 1 ≤ j ≤ ℓ,
(Hℓ,2) limk→∞
uk(s(j)
k )
uk(Lr(j)
k )
= 0 for 1 ≤ j ≤ ℓ, L > 0,
, r(j)
k /L) + Nk(Lr(j)
k , s(j)
k )(cid:1) = 0 for 1 ≤ j ≤ ℓ.
(Hℓ,3) limk→∞ Λk(s(j)
(Hℓ,4) limL→∞ limk→∞(cid:0)Nk(s(j−1)
k
k ) = jΛ1 for 1 ≤ j ≤ ℓ,
For the proof of Theorem 2 we proceed via induction from the following two
claims: (H1) holds, and if (Hℓ) holds then either (Hℓ+1) holds as well, or
lim
k→∞
Nk(s(ℓ)
k , R) = 0.
(11)
By (3) and (Hℓ,3) the induction terminates when ℓ > Λ
Λ1
largest integer such that (Hℓ0 ) holds, (Hℓ0,3) and (11) imply
. Letting ℓ0 be the
Λ = lim
k→∞
Λk(s(ℓ0)
k
) + lim
k→∞
Nk(s(ℓ0)
k
, R) = ℓ0Λ1,
and Theorem 2 in the radial case follows.
2.2 Proof of (H1)
Let rk > 0 be defined as in Theorem 1 such that
λkr2m
k u2
k(0)emu2
k(0) = 22m(2m − 1)!,
and set
We have
wk(x) := uk(0)(uk(x) − uk(0)) in BR.
(−∆)mwk = λkuk(0)ukemu2
= λkuk(0)ukemu2
k
k(0)e
2m(cid:0)1+ wk
2u2
k
(0)(cid:1)wk
=: fk in BR.
Letting also
σk(r) := ZBr
fkdx ≥ Λk(r),
then by (5) of Theorem 1 and (9) clearly we have
lim
L→∞
lim
k→∞
Λk(Lrk) = lim
L→∞
lim
k→∞
= lim
L→∞
lim
k→∞
4
σk(Lrk)
(2m − 1)!ZBL
e2mηk dx = Λ1.
(12)
For 0 < t ≤ R let gk solve the equation
∆mgk = ∆mwk in Bt
with homogeneous Dirichlet boundary data
gk = ∂ν gk = . . . = ∂m−1
ν
gk = 0 on ∂Bt.
Then Lemma 22 in the Appendix gives the identity
(−1)m∂m
ν gk(t) =
Ak(t)
ω2m−1t3m−2
(13)
similar to (20) in [13], where
Ak(t) := Z t
0
t2 ···Z tm−1
0
tmσk(tm)dtm . . . dt2,
(14)
and where ω2m−1 is the (2m − 1)-dimensional volume of S2m−1.
Lemma 3 For every b < 2 we can find L = L(b) and k0 = k0(b) such that for
k ≥ k0 we have
(−∂ν)mgk(t) ≥
Proof. Noting that
2m−1(m − 1)!b
tm
for Lrk ≤ t ≤ R.
(15)
Λ1
ω2m−12m−1(m − 1)!
= 2m(m − 1)!,
from (12) and (13) together with the identity
Z t
0
t2 ···Z tm−1
0
tmdtm ··· dt2 =
t2m−2
2m−1(m − 1)!
we obtain the claim.
(cid:3)
These estimates now yield the following result analogous to Lemma 2.1 in
[13]. Note, however, that the statement (17) below in the present case no longer
can simply be deduced from the maximum principle, as was the case in [13]. In
addition, the higher order nature of equation (1) requires substantial technical
modifications of the approach used in [13].
Lemma 4 For any b < 2 there is L = L(b) and k0 = k0(b) such that for k ≥ k0
there holds
b
t
+ tP (t)
w′
k(t) ≤ −
w′
k(t) ≤ 0
wk(t) ≤ b log(cid:16) rk
in BR\BLrk,
t (cid:17) + C in BR,
in BR,
(16)
(17)
(18)
where P is a polynomial independent of k. In particular uk is monotone de-
creasing. For any ε ∈]0, 1[ let Tk > 0 be such that uk(Tk) = εuk(0). Then we
have
(19)
= 0
lim
k→∞
rk
Tk
5
and
lim
k→∞
Λk(Tk) = lim
k→∞
σk(Tk) = Λ1.
(20)
Proof. Fix t > 0 and write wk = gk + hk, where
∆mhk = 0 in Bt, and gk = ∂ν gk = . . . = ∂m−1
ν
gk = 0 on ∂Bt.
Step 1. We claim that
∂m
ν gk(t) = tm−1 (t−1(t−1 ··· (t−1
}
m−1 times
{z
w′
k(t) )′ ··· )′)′
{z }
m−1 times
.
(21)
Indeed, subtracting ∂m
ν wk(t) from both sides of (21) we need to show
−∂m
ν hk(t) = tm−1 (t−1(t−1 ··· (t−1
}
m−1 times
{z
w′
k(t) )′ ··· )′)′
{z }
m−1 times
−∂m
ν wk(t).
Using the boundary condition ∂j
observing that on the right-hand side the terms involving ∂m
can replace wk by hk and it suffices to prove
ν hk(t) for 0 ≤ j ≤ m − 1, and
ν wk(t) cancel, we
νwk(t) = ∂j
−∂m
ν hk(t) = tm−1 (t−1(t−1 ··· (t−1
}
m−1 times
{z
k(t) )′ ··· )′)′
h′
{z }
m−1 times
−∂m
ν hk(t).
But ∆mhk = 0 and radial symmetry imply that h(r) = Pm−1
i=0 αir2i; so
(t−1(t−1 ··· (t−1
}
m−1 times
{z
h′
k(t) )′ ··· )′)′
{z }
m−1 times
= 0,
and (21) follows.
Step 2. Inserting now (15) into (21), for any given b < 2 we infer
(−1)m−1tm−1(cid:16)t−1(cid:16)t−1 ···(cid:16)t−1
}
(cid:16)w′
{z
2m−1(m − 1)!b
= (−1)m−1∂m
ν gk(t) +
m−1 times
tm
k(t) +
b
t(cid:17)(cid:17)′
···(cid:17)′(cid:17)′
{z
}
m−1 times
≤ 0 for Lrk ≤ t ≤ R,
(22)
provided that we fix L = L(b) sufficiently large and then also choose k large
enough. We now prove by induction over 1 ≤ j ≤ m that
t(cid:17)(cid:17)′
ϕj,k(t) := (−1)m−jt−1(cid:16)t−1(cid:16)t−1 ···(cid:16)t−1
}
···(cid:17)′(cid:17)′
{z
}
≤ Pj(t),
(cid:16)w′
k(t) +
m−j times
m−j times
{z
(23)
b
for Lrk ≤ t ≤ R, where Pj(t) ≥ 0 is a polynomial in t independent of k. The
case j = 1 follows at once from (22) with P1 ≡ 0. Using the Dirichlet boundary
condition (which implies ∂j
νwk(R) = 0 for 1 ≤ j ≤ m − 1) we get ϕj,k(R) ≤ Cj
6
for some constant Cj ≥ 0, 2 ≤ j ≤ m. Observing that ϕ′
2 ≤ j ≤ m, we then obtain
j,k(t) = −tϕj−1,k(t) for
ϕj,k(t) = ϕj,k(R) +Z R
t
rϕj−1,k(r)dr
≤ Cj +Z R
t
rPj−1(r)dr =: Pj(t),
that is, (23). For j = m we get
w′
k(t) ≤ −
b
t
+ tPm(t), Lrk ≤ t ≤ R.
Integrating once more, recalling that L depends on b, and that wk(Lrk) → η0(L)
as k → ∞, for sufficiently large k we find
wk(t) ≤ wk(Lrk) − b log(cid:16) t
Lrk(cid:17) + C ≤ b log(cid:16) rk
t (cid:17) + C
for Lrk ≤ t ≤ R. For 0 < t < Lrk (18) already follows from Theorem 1.
In order to prove (17), observe that (13) implies
(−∂ν)mgk(t) ≥ 0 for 0 < t ≤ R, k ∈ N,
and (21) yields
(−1)m−1tm−1 (t−1(t−1 ··· (t−1
}
m−1 times
{z
w′
k(t) )′ ··· )′)′
{z }
m−1 times
≤ 0.
(24)
In analogy with (23), for 1 ≤ j ≤ m we can show by induction that
ψj,k(t) := (−1)m−jt−1 (t−1(t−1 ··· (t−1
}
w′
k(t) )′ ··· )′)′
{z }
m−j times
m−j times
{z
≤ 0 for all 0 < t ≤ R.
Indeed ψ1,k(t) ≤ 0 by (24), while for 2 ≤ j ≤ m, we have ψj,k(R) = 0 thanks to
the boundary condition. Hence
ψj,k(t) = Z R
t
rψj−1(r)dr ≤ 0 for all 0 < t ≤ R,
and the case j = m implies (17).
Step 3. In order to prove (19), assume by contradiction that
lim inf
k→∞
Tk
rk
= L ∈ [0,∞[.
Then from Theorem 1 for a suitable subsequence on the one hand we have
uk(0)(uk(Tk) − uk(0)) + log 2 = ηk(cid:18) Tk
rk(cid:19) → log(cid:18) 2
But on the other hand, since uk(0) → ∞ we also have that
1 + L2(cid:19) as k → ∞.
uk(0)(uk(Tk) − uk(0)) = u2
k(0)(ε − 1) → −∞
7
as k → ∞, a contradiction.
It thus remains to prove (20). Using (18) and observing that
k(0) ≤ wk(r) ≤ 0 for 0 ≤ r ≤ Tk,
(ε − 1)u2
from (4) for k ≥ k0 we get
k(0)emu2
k u2
fk(r) ≤ λku2
≤ λkr2m
k(0)e
k(0)emu2
2m(cid:0)1+ wk (r)
2u2
k
(0)(cid:1)wk(r)
em(ε+1)wk(r) ≤ Cr−2m
k
k(0)r−2m
k
(cid:16) rk
r (cid:17)m(ε+1)b
.
Choosing now b < 2 such that m(ε + 1)b = 2m + ε, and integrating over BTk ,
we find
Λ1 ≤ lim
k→∞
Λk(Tk) ≤ lim
k→∞
σk(Tk)
= Λ1 + lim
L→∞
lim
k→∞ZBTk \BLrk
fkdx
≤ Λ1 + C lim
L→∞
lim
k→∞
≤ Λ1 +
C
ε
lim
L→∞
1
r2m
k ZBTk \BLrk (cid:16) rk
Lrk(cid:17)ε
= Λ1,
lim
k→∞(cid:16) rk
r (cid:17)2m+ε
dx
hence (20).
(cid:3)
According to Lemma 4 we can now choose a sequence εk → 0 as k → ∞ and
corresponding numbers sk = Tk(εk) such that uk(sk) → ∞ as k → ∞ and
lim
k→∞
rk
sk
= 0,
lim
k→∞
Λk(sk) = Λ1,
lim
L→∞
lim
k→∞
Nk(Lrk, sk) = 0.
(25)
Observing that Theorem 1 implies limk→∞
uk(Lrk)
uk(0) = 1 for every L ≥ 0, we get
lim
k→∞
uk(sk)
uk(Lrk)
= lim
k→∞
uk(sk)
uk(0)
= 0,
for all L > 0.
(26)
We also claim that
lim
L→∞
lim
k→∞
Nk(sk, Lsk) = 0.
(27)
To see this, remember that for 0 < s < t < R
Nk(s, t) = ZBt\Bs
ekdx = ω2m−1Z t
s
λkr2m−1u2
kemu2
k dr.
Now set
Pk(t) := t
∂
∂t
Nk(s, t) = tZ∂Bt
ekdσ = ω2m−1λkt2mu2
k(t)emu2
k(t).
Using the monotonicity of uk that we proved in Lemma 4 we immediately obtain
the estimate
Pk(t) = Cω2m−1λku2
k(t)emu2
k(t)Z t
t/2
r2m−1dr ≤ CNk(t/2, t) ≤ CPk(t/2)
(28)
8
analogous to (26) in [13]; hence we also conclude that
Nk(t, 2t) ≤ CNk(t/2, t)
for t ∈ [0, R/2].
(29)
Now (25) and (29) imply that for any M ∈ N
Nk(2M−1sk, 2M sk) ≤ C lim
lim
k→∞
k→∞
Nk(2M−2sk, sM−1sk)
≤ ··· ≤ CM lim
k→∞
Nk(sk/2, sk) = 0.
Therefore if 2M ≥ L we have
lim
k→∞
Nk(sk, Lsk) ≤
M
Xj=1
Nk(2j−1sk, 2jsk) = 0,
as claimed.
Setting r(1)
k
:= rk, s(1)
k
1 we see that the property (H1) is satisfied.
:= sk and taking into account (25) - (27) and Theorem
2.3 The inductive step
We now assume that (Hℓ) holds for some integer ℓ ≥ 1 and fix numbers
s(0)
k = 0 < r(1)
k < s(1)
k < . . . < r(ℓ)
k < s(ℓ)
k , k ∈ N
such that (Hℓ,1), (Hℓ,2), (Hℓ,3) and (Hℓ,4) hold true. To complete the proof
of Theorem 2 it suffices to show that either (Hℓ+1) or (11) holds. The proof
requires the following analogue of (29) in [13].
Lemma 5 There is a constant C0 = C0(Λ) such that for tk > s(ℓ)
k
there holds
Pk(tk)
m
+ C0N 2
k (s(ℓ)
k , tk) + o(1),
(30)
Nk(s(ℓ)
k , tk) ≤
with error o(1) → 0 as k → ∞
Proof. For s = s(ℓ)
k < t we integrate by parts to obtain
r2m−1λku2
kemu2
k dr
s
Nk(s, t) = ω2m−1Z t
2m (cid:0)r2mu2
kemu2
2m − ω2m−1Z t
ω2m−1
Pk(t)
= λk
≤
s
t
k(cid:1)(cid:12)(cid:12)
s −
λkr2m(cid:16) 1
m
ω2m−1
2m Z t
k(cid:17) uk
+ u2
s
λkr2m(2uk + 2mu3
k)u′
kemu2
k dr
uk(0)
kemu2
w′
k dr.
(31)
Define gk(t) as in the beginning of the proof of Lemma 4. Then (13) and (21)
imply
(t−1(t−1 ··· (t−1
}
m−1 times
{z
= (−1)m
Ak(t)
ω2m−1t4m−3 ,
k(t) )′ ··· )′)′
w′
{z }
m−1 times
9
where Ak is as in (14). Integrating this relation m − 1 times from t to R, and
using the Dirichlet boundary condition ∂j
νwk(R) = 0 for 1 ≤ j ≤ m − 1 we get
w′
k(t)
t
= −Z R
t
t1Z R
t1
t2 ···Z R
tm−2
Ak(tm−1)
ω2m−1t4m−3
m−1
dtm−1 ··· dt1;
hence
−tw′
k(t)
uk(t)
uk(0)
= t2 uk(t)
uk(0) Z R
t
t1Z R
t1
t2 ···Z R
tm−2
Ak(tm−1)
ω2m−1t4m−3
m−1
dtm−1 ··· dt1 =: I.
More explicitly,
I = t2Z R
×Z tm−1
t
0
t1
t2 ···Z R
t1Z R
ρ1 ···Z ρm−2
0
tm−2
1
ω2m−1t4m−3
m−1
ρm−1τk(ρm−1, t)dρm−1 ··· dρ1 dtm−1 ··· dt1,
where
τk(ρ, t) =
uk(t)σk(ρ)
uk(0)
= ZBρ
λkuk(t)ukemu2
k dx.
We now show that I can be bounded in terms of Nk(s, t) up to a small error.
From this the desired inequality (30) will be immediate. Split
I =: II + III,
where II corresponds to ρm−1 ≤ t. Since u′
k ≤ 0, for ρ ≤ t we have
τk(ρ, t) = ZBρ
≤ ZBs
λkuk(t)ukemu2
λkuk(s)ukemu2
λkuk(ρ)ukemu2
k dx ≤ ZBρ
k dx + Nk(s, ρ) ≤ Nk(s, t) + o(1)
k dx
(32)
with error o(1) → 0 as k → ∞. Here we used that for arbitrary L > 1 we can
bound
ZBs
λkuk(s)ukemu2
k dx ≤ Nk(Lr(ℓ)
k , s) +
uk(s)
uk(Lr(ℓ)
k )
Λk(Lr(ℓ)
k ),
and by (Hℓ,2), (Hℓ,4) the latter tends to 0, if first k → ∞ and then L → ∞.
Since
t2Z ∞
ρm−1dρm−1 ··· dt1 ≤ C
ρ1 ···Z ρm−2
t1 ···Z ∞
m−1 Z tm−1
ω2m−1t4m−3
tm−2
1
0
0
t
uniformly in t, we conclude that
II ≤ CNk(s, t) + o(1).
In order to obtain a similar bound for III, for t ≤ ρ we estimate
τk(ρ, t) =
uk(t)σk(ρ)
uk(0)
=
uk(t)
uk(ρ) + 1 ZBρ
λk(uk(ρ) + 1)ukemu2
k dx .
10
Recalling (32), we have
ZBρ
λk(uk(ρ) + 1)ukemu2
k dx ≤ τk(ρ, ρ) + o(1) ≤ Nk(s, ρ) + o(1).
Also note that by Holder's inequality we can estimate
uk(t) − uk(ρ) ≤ Z ρ
t
k(r)dr ≤ k∇ukkL2m(cid:0) log
u′
Thus, with a constant C = C(Λ) for all t ≤ ρ we obtain
ρ
2m .
t(cid:1) 2m−1
tuk(t)
ρ(uk(ρ) + 1)
=
≤
t
uk(ρ) + 1
ρ(cid:18) uk(t) − uk(ρ)
ρ(cid:18)C(cid:0) log
t(cid:1) 2m−1
ρ
t
+
uk(ρ)
uk(ρ) + 1(cid:19)
2m + 1(cid:19) ≤ C
(33)
and with C1 = C1(Λ) we can bound
t
ρ
τk(ρ, t) ≤ C1Nk(s, ρ) + o(1).
It follows that
1
ω2m−1t4m−3
m−1
1
ω2m−1t4m−3
m−1
t
t
t
t1
tm−2
III = t2Z R
×Z tm−1
= tZ R
t1Z R
×Z tm−1
≤ C1tZ R
×Z tm−1
t2 ···Z R
t1Z R
ρ1 ···Z ρm−2
t2 ···Z R
ρ1 ···Z ρm−2
t2 ···Z R
t1Z R
ρ1 ···Z ρm−2
tm−2
t1
t1
t
t
t
t
t
t
ρm−1τk(ρm−1, t)dρm−1 ··· dρ1 dtm−1 ··· dt1
ρ2
m−1
τk(ρm−1, t)dρm−1 ··· dρ1 dtm−1 ··· dt1
t
ρm−1
1
tm−2
ω2m−1t4m−3
m−1
ρ2
m−1(Nk(s, ρm−1) + o(1))dρm−1 ··· dt1.
For any L ≥ 1 we split the integral with respect to t1 and use the obvious
inequality Nk(s, ρm−1) ≤ 2Λ for large k to estimate
1
t3 ···Z R
tm−2
ω2m−1t4m−3
m−1
t
t1
t2
III ≤ C1tZ Lt
×Z tm−1
+ 2C1ΛtZ R
×Z tm−1
t1Z R
t2Z R
ρ1 ···Z ρm−2
t1Z R
t2Z R
ρ1 ···Z ρm−2
Lt
t2
t1
t
t
t
t
ρ2
m−1(Nk(s, ρm−1) + o(1))dρm−1 ··· dt1
t3 ···Z R
ρ2
m−1dρm−1 ··· dt1 + o(1)
ω2m−1t4m−3
m−1
tm−2
1
11
Observing the uniform bound
LtZ ∞
Lt
t1 ···Z ∞
tm−2
1
m−1 Z tm−1
0
ω2m−1t4m−3
ρ1 ···Z ρm−2
0
ρ2
m−1dρm−1 ··· dt1 ≤ C,
with a constant C2 = C2(Λ) we obtain
III ≤ C1tZ Lt
×Z tm−1
t1Z R
t2Z R
ρ1 ···Z ρm−2
t1
t2
t
t
t
C2Λ
L
+
+ o(1)
t3 ···Z R
tm−2
1
ω2m−1t4m−3
m−1
ρ2
m−1(Nk(s, ρm−1) + o(1))dρm−1 ··· dt1
To proceed we successively split the integral also with respect to t2, . . . , tm−1
and use the uniform bounds
LtZ Lt
0
t1 ···Z Lt
0
tj−1Z ∞
Lt
tj ···Z ∞
ρ1 ···Z ρm−2
0
tm−2
×Z tm−1
0
1
ω2m−1t4m−3
m−1
ρ2
m−1dρm−1 ··· dtj ≤ C
t
t
t
t
t1
t2
Lt
III ≤ C1tZ Lt
×Z tm−1
+ 2C1ΛtZ Lt
×Z tm−1
for 2 ≤ j < m to estimate
t1Z Lt
t2Z R
ρ1 ···Z ρm−2
t2Z R
t1Z R
ρ1 ···Z ρm−2
t1Z Lt
ρ1 ···Z ρm−2
t1Z Lt
t2Z Lt
ρ1 ···Z ρm−2
≤ ··· ≤ C1tZ Lt
×Z tm−1
+ 2C1ΛtZ Lt
×Z tm−1
t2
t2
t1
t1
t2
t
t
t
t
t
t
t
t2Z Lt
≤ CmNk(s, Lt) +
+ o(1),
t
CmΛ
L
t3 ···Z R
tm−2
1
ω2m−1t4m−3
m−1
1
L
tm−2
+ o(1)
ω2m−1t4m−3
m−1
C2Λ
ρ2
m−1(Nk(s, ρm−1) + o(1))dρm−1 ··· dt1
t3 ···Z R
ρ2
m−1dρm−1 ··· dt1 +
t3 ···Z Lt
m−1(cid:0)Nk(s, Lt) + o(1)(cid:1)dρm−1 ··· dt1
t3 ···Z R
ρ2
m−1dρm−1 ··· dt1 +
ω2m−1t4m−3
m−1
ω2m−1t4m−3
m−1
Cm−1Λ
+ o(1)
tm−2
ρ2
L
Lt
1
1
with constants Cj = Cj (Λ), 2 ≤ j ≤ m. Using (27) in case t ≤ 2s and (28) in
case t > 2s we get
Nk(s, Lt) ≤ C(L)Nk(s, t) + o(1),
12
and with the constant Cm+1 = CmΛ = Cm+1(Λ) there results
−tw′
k(t)
uk(t)
uk(0) ≤ C(L, Λ)Nk(s, t) +
Cm+1
L
+ o(1).
Inserting this into (31) we infer
Nk(s, t) ≤
Pk(t)
2m − ω2m−1Z t
λkr2m(cid:16) 1
+(cid:16)C(L, Λ)Nk(s, t) +
s
m
Pk(t)
2m
≤
+ u2
k dr
uk(0)
k(cid:17) uk
kemu2
w′
L (cid:17)Nk(s, t) + o(1).
Cm+1
(34)
Choosing L = 2Cm+1 we finally get (30) for an appropriate C0 = C0(Λ).
(cid:3)
Lemma 6 Let C0 = C0(Λ) be the constant appearing in (30).
tk ∈]s(ℓ)
k , R] there holds
If for some
0 < lim
k→∞
Nk(s(ℓ)
k , tk) =: α <
1
2C0
,
(35)
then
s(ℓ)
k
tk
lim
k→∞
= 0, lim inf
k→∞
Pk(tk) ≥
Proof. Assume that for some tk ∈]s(ℓ)
as in the proof of (27) also yields that
2
mα
, and lim
L→∞
lim
k→∞
Nk(s(ℓ)
k , tk/L) = 0.
k , R] we have (35). Since the same reasoning
lim
L→∞
lim
k→∞
Nk(s(ℓ)
k , Ls(ℓ)
k ) = 0,
necessarily s(ℓ)
Pk(tk)
lim inf
k→∞
k /tk → 0 as k → ∞. Moreover, (30) yields
k→∞(cid:16)Nk(s(ℓ)
k , tk) − C0N 2
Nk(s(ℓ)
k , tk)
≥ lim
2
m ≥ lim
α
2
k→∞
=
,
k (s(ℓ)
k , tk)(cid:17)
as claimed. Now we show that
Indeed, if we assume
lim
L→∞
lim
k→∞
Nk(s(ℓ)
k , tk/L) = 0.
lim
L→∞
lim sup
k→∞
Nk(s(ℓ)
k , tk/L) = β > 0,
we have
β
2 ≤ Nk(s(ℓ)
k , tk/L) ≤ Nk(s(ℓ)
k , tk) <
1
2C0
(36)
(37)
for any L ≥ 1 and sufficiently large k. Therefore we can apply (36) with tk/L
instead of tk for any L ≥ 1 to get
lim
k→∞
Pk(tk/L) ≥
mβ
2
.
13
Then (28) yields
C lim
k→∞
Nk(tk/(2L), tk/L) ≥ lim
k→∞
Pk(tk/L) ≥
mβ
2
.
Choosing L = 2j and summing over j for 0 ≤ j ≤ M − 1, we get
C lim
k→∞
Nk(2−M tk, tk) ≥
which contradicts (3). Therefore (37) is proven.
Λk(tk) ≥ C lim
k→∞
mM β
2 → ∞ as M → ∞,
(cid:3)
Suppose now that for some tk ≥ s(ℓ)
k
there holds
lim sup
k→∞
Nk(s(ℓ)
k , tk) > 0.
We then want to show that (Hℓ+1) holds. We can choose numbers rℓ+1
such that for a subsequence there holds
k ∈]s(ℓ)
k , tk[
0 < lim
k→∞
Nk(s(ℓ)
k , r(ℓ+1)
k
) <
1
2C0
,
(38)
where C0 is as in Lemma 6. Observe that Lemma 6 then implies
lim
k→∞
s(ℓ)
k /r(ℓ+1)
k
= lim
L→∞
lim
k→∞
Nk(s(ℓ)
k , r(ℓ+1)
k
/L) = 0,
(39)
and
lim
k→∞
Pk(r(ℓ+1)
k
) > 0.
(40)
Proposition 7 We have
η(ℓ+1)
k
(x) := uk(r(ℓ+1)
k
)(cid:0)uk(r(ℓ+1)
k
x) − uk(r(ℓ+1)
k
)(cid:1) → η(ℓ+1)
in C2m−1
loc
η(ℓ+1)
0
(R2m\{0}). Moreover, for a suitable constant c(ℓ+1) the function
:= η(ℓ+1) + c(ℓ+1) satisfies
(−∆)mη(ℓ+1)
0
= (2m − 1)!e2mη(ℓ+1)
0
,
ZR2m
(2m − 1)!e2mη(ℓ+1)
0
dx = Λ1.
The above proposition, which will be proven in the following section, implies
that
hence (39) yields
lim
L→∞
lim
k→∞
Nk(r(ℓ+1)
k
/L, Lr(ℓ+1)
k
) = Λ1;
lim
L→∞
lim
k→∞
Nk(s(ℓ)
= lim
L→∞
lim
k→∞
)
k , Lr(ℓ+1)
k
Nk(s(ℓ)
k , r(ℓ+1)
k
/L) + lim
L→∞
lim
k→∞
Nk(r(ℓ+1)
k
/L, Lr(ℓ+1)
k
)
(41)
= 0 + Λ1 = Λ1.
Then the inductive hypothesis (Hℓ,3) gives
lim
L→∞
lim
k→∞
Λk(Lr(ℓ+1)
k
) = lim
L→∞
lim
k→∞(cid:16)Λk(s(ℓ)
k ) + Nk(s(ℓ)
k , Lr(ℓ+1)
k
)(cid:17)
= (ℓ + 1)Λ1.
14
Now set w(ℓ+1)
k
(x) = uk(r(ℓ+1)
k
(−∆)mw(ℓ+1)
k
)(uk(x) − uk(r(ℓ+1)
= λkuk(r(ℓ+1)
k
)ukemu2
k
)) so that
k =: f (ℓ+1)
.
k
Similar to Lemma 4 and with the same proof (except that instead of Theorem
1 one needs to use Proposition 7) we have
Lemma 8 For any 0 < ε < 1, letting T (ℓ+1)
εuk(r(ℓ+1)
), we have
k
k
(ε) > 0 be such that uk(T (ℓ+1)
k
) =
lim
k→∞
Nk(s(ℓ)
k , T (ℓ+1)
k
) = Λ1.
(42)
Moreover r(ℓ+1)
k
/T (ℓ+1)
k
→ 0 as k → ∞.
According to Lemma 8 and (41) we can choose numbers εk → 0 and a
) → ∞ as k → ∞
(εk) we have uk(s(ℓ+1)
subsequence so that for s(ℓ+1)
and
:= T (ℓ+1)
k
k
k
lim
k→∞
r(ℓ+1)
k
s(ℓ+1)
k
= 0,
while
lim
k→∞
Λk(s(ℓ+1)
k
) = (ℓ + 1)Λ1,
lim
L→∞
lim
k→∞
Nk(Lr(ℓ+1)
k
, s(ℓ+1)
k
) = 0.
Again reasoning as in the proof of (27) we also infer
lim
k→∞
Nk(s(ℓ+1)
k
, Ls(ℓ+1)
k
) = 0 for every L ≥ 1.
Finally, observe that the definition of s(ℓ+1)
k
implies that
lim
k→∞
uk(s(ℓ+1)
)
uk(Lr(ℓ+1)
k
k
)
= 0 for every L ≥ 0.
Together with (39) this completes the proof of (Hℓ+1), and hence of Theorem 2
in the radially symmetric case.
2.4 Proof of Proposition 7
As preparation for the proof of Proposition 7 we need the following two lemmas.
Lemma 9 For r(ℓ+1)
k
as above, we have
k
vk(x) := uk(cid:0)r(ℓ+1)
Proof. We write rk = r(ℓ+1)
. Moreover, we consider only the case m > 1, the
case m = 1 being considerably easier. As in the proof of Lemma 3.2 in [13] we
have
x(cid:1) − uk(cid:0)r(ℓ+1)
(R2m\{0}).
(cid:1) → 0
in C2m−1
loc
k
k
(−∆)mvk(x) = λkr2m
k uk(rkx)emu2
k(rkx) =: gk(x) ≥ 0,
15
with gk → 0 in L∞
have
loc(R2m\{0}). By scaling and Sobolev's embedding we also
k∇2vkkLm(BR/rk
k∇mvkkL2(BR/rk
) = k∇2ukkLm(BR) ≤ C,
) = k∇mukkL2(BR) ≤ C.
(43)
loc
loc
Set wk := ∆vk. Then a subsequence wk → w weakly in H m−2
(R2m) and in
C2m−3,α
(R2m \ {0}) for some function w ∈ Lm(R2m) with ∇m−2w ∈ L2(R2m).
Clearly ∆m−1w = 0 in R2m\{0}. In fact, since the point x = 0 has vanishing
H m-capacity, as in [13] we have ∆m−1w = 0 in R2m. Recalling that w ∈
Lm(R2m) we conclude that w ≡ 0; see Lemmas 23 and 24 in the appendix.
Recalling that (∆vk) is bounded in Lm(R2m) and noting the condition
vk(1) = 0, from standard elliptic estimates we infer that (vk) is bounded in
W 2,m(B1). Hence a subsequence vk → v weakly in W 2,m(B1) and in C2m−1,α
away from x = 0. We then have ∆v = 0 and v(1) = 0, therefore v ≡ 0 on B1.
By elliptic estimates, from (43) and the condition vk(1) = 0 we also infer that
(vk) is bounded in W 2,m
loc (R2m). Therefore, we also have that vk → v weakly in
W 2,m
(R2m \ {0}), with ∆v = 0. By unique continuation
it follows that v ≡ 0. This completes the proof.
loc (R2m) and in C2m−1,α
loc
(cid:3)
Lemma 10 For any L > 0 there exists k0 = k0(L) such that for all k ≥ k0 and
any 1 ≤ j ≤ 2m − 1 there holds
uk(r(ℓ+1)
k
)ZB
∇jukdx ≤ C(Lr(ℓ+1)
k
)2m−j .
(ℓ+1)
k
\B
r
(ℓ+1)
k
/L
Lr
Proof. The proof is identical to the proof of Lemma 6 in [8], using Lemma 9
above instead of Lemma 3 in [8].
(cid:3)
Proof of Proposition 7. For simplicity of notation, we now drop the index ℓ + 1.
Step 1. We claim that ηk → η in C2m−1,α
(R2m\{0}) for some smooth function
η. For any L > 1 let ΩL := BL(0)\B1/L(0). Recall that by Lemma 9 we have
uk(x) := uk(rkx)
uk(rk) → 1 uniformly on ΩL as k → ∞. Thus by (7) with error
o(1) → 0 as k → ∞ we have
loc
0 ≤ (−∆)mηk(x) = λkr2m
≤ (L2m + o(1))λk(rkx)2mu2
k u2
k(rk)uk(x)emu2
k(rkx)emu2
k(rkx)
k(rkx) ≤ CL2m + o(1).
(44)
Split ηk = hk + lk on Ω2L, where
∆mhk = 0 on Ω2L, and lk = ∆lk = . . . = ∆m−1lk = 0 on ∂Ω2L.
Since k∆mηkkL∞(Ω2L) ≤ C = C(L), by elliptic estimates we get that lk → l in
C2m−1,α(Ω2L). Together with Lemma 10 this implies
k∇hkkL1(Ω2L) ≤ k∇lkkL1(Ω2L) + k∇ηkkL1(Ω2L) ≤ C.
Moreover, since ηk = 0 on ∂B1(0), we have
hk = lk ≤ C on ∂B1(0).
(45)
16
Then, from a Poincar´e-type inequality, we easily get khkkL1(Ω2L) ≤ C. By virtue
of Proposition 21, we infer that
khkkC j(ΩL) ≤ Cj
for every j ∈ N.
Hence a subsequence hk → h smoothly on ΩL, and
ηk → η := h + l
in C2m−1,α(ΩL),
proving our claim.
Step 2. With uk(x) := uk(rkx)
uk(rk) as above, from (44) we get
(−∆)mηk = λkr2m
k u2
k(rk)emu2
= µkukem(uk+1)ηk ,
k(rk)uk(x)em(u2
k(rk · )−u2
k(rk))
(46)
where by (40) we may assume
µk := λkr2m
k u2
k(rk)emu2
k(rk) = ω−1
2m−1Pk(rk) → µ0 > 0.
Since uk → 1 locally uniformly on R2m\{0} we may pass to the limit k → ∞ in
(46) to see that η solves the equation
(−∆)mη = µ0e2mη
on R2m\{0}
(47)
in the distribution sense. In fact, we now show that (47) holds on all of R2m.
Note that by Step 1 for any L > 1 we have
ZΩL
e2mηdx = lim
k→∞ZΩL
k→∞ZΩL
= lim
u2
kem(uk+1)ηk dx
µ−1
k uk(−∆)mηkdx
≤ µ−1
0
lim inf
k→∞ ZBLrk
uk(−∆)mukdx ≤ µ−1
0 Λ.
As L → ∞, by Fatou's lemma, we get e2mη ∈ L1(R2m). Moreover η ≥ 0 on B1,
hence η ∈ Lp(B1) for every p ∈ [1,∞[. Also note that (−∆)mηk ≥ 0 and that
from (32) we can bound
lim sup
k→∞ ZB1/L(0)
(−∆)mηk dx
= lim sup
k→∞ ZBrk /L(0)
λkuk(rk)ukemu2
k dx ≤ lim sup
k→∞
Nk(s(ℓ)
k , rk/L) → 0
(48)
as L → ∞. Since by Lemma 4 we have uk ≥ 1, ηk ≥ 0 on B1, from (46) and
(48) we also find that
lim sup
k→∞ ZB1/L(0)
ηk dx → 0 as L → ∞ .
(49)
17
By (47) and (48) for any test function ϕ ∈ C∞
ZR2m(cid:0)(−∆)mη − µ0e2mη(cid:1)ϕ dx = lim
L→∞ZR2m
(−∆)mηϕτL dx
k→∞ ZR2m(cid:0)(−∆)mη − (−∆)mηk(cid:1)ϕτL dx,
lim inf
0 (R2m) we now obtain
= lim
L→∞
(50)
where for L ∈ N we let τL(x) = τ (Lx) with a fixed cut-off function τ ∈ C∞
such that 0 ≤ τ ≤ 1 and τ ≡ 1 in B1. But by Step 1 for any L ≥ 1 we have
0 (B2)
lim inf
k→∞ ZR2m (cid:0)(−∆)mη − (−∆)mηk(cid:1)ϕτL dx
= lim inf
k→∞ ZR2m
(η − ηk)(cid:0)(−∆)mϕ(cid:1)τL dx,
and since η ∈ L1(B1) and on account of (49) the latter converges to 0 as L → ∞
for any fixed ϕ ∈ C∞
0 (R2m). From (50) we thus see that η solves (47) in the
distribution sense on R2m. By elliptic estimates, η is smooth on all of R2m; see
for instance [7], Corollary 8. The function η0 := η + 1
(2m−1)! then satisfies
2m log
µ0
(−∆)mη0 = (2m − 1)!e2mη0
in R2m,
ZR2m
e2mη0dx < ∞.
(51)
Solutions to (51) have been classified in [7], where it was shown that either
2σ
(i) η0(x) = log
(ii) m > 1 and there exist 1 ≤ j ≤ m − 1 and a 6= 0 such that
1+σx−x02 for some σ > 0, x0 ∈ R2m, or
lim
x→∞
∆jη0(x) = a,
and hence for sufficiently large L, with error o(1) → 0 as k → ∞,
(Lrk)2j−2muk(rk)ZBLrk \Brk /L ∇2jukdx = L2j−2mZBL\B1/L ∇2jηkdx
= L2j−2mZBL\B1/L ∇2jη0dx + o(1) ≥ CL2j + o(1)
(52)
for some constant C > 0 independent of L.
But (52) is incompatible with the estimate of Lemma 10 when L and k are
large. Hence case (i) occurs (with x0 = 0, by radial symmetry). In particular,
(cid:3)
we have RR2m(2m − 1)!e2mη0 dx = Λ1.
3 The general case
The following gradient bound analogous to [5], Proposition 2, and generalizing
[13], Proposition 4.1, will be crucial in the sequel. The proof will be given in
the next section.
Proposition 11 There exists a uniform constant C such that
sup
x∈Ω
inf
1≤j≤I x − x(j)
k ℓuk(x)∇ℓuk(x) ≤ C for all 1 ≤ ℓ ≤ 2m − 1, k ∈ N.
18
Fix an index i ∈ {1, . . . , I} and let xk = x(i)
k → 0 as given
by Theorem 1. After a translation we may assume that x(i) = 0. Set as before
k → x(i), rk = r(i)
ek := λku2
kemu2
k ,
fk := λkuk(0)ukemu2
k ,
and
Λk(r) := ZBr
ekdx.
In the following we will use the notation
¯f (r) := Z
∂Br
f dσ,
for any function f . Set also
ek := λk ¯u2
kem¯u2
k ≤ ¯ek.
(Here we used Jensen's inequality.) Again we let wk(x) := uk(0)(uk(x)− uk(0)),
satisfying
(−∆)m ¯wk = λkuk(0)ukemu2
k = ¯fk.
Finally set
Λk(r) := ZBr
ekdx ≤ Λk(r), σk(r) := ZBr
¯fkdx.
(53)
Again Theorem 1 implies
lim
L→∞
lim
k→∞
Λk(Lrk) = lim
L→∞
lim
k→∞
Λk(Lrk) = lim
L→∞
lim
k→∞
σk(Lrk) = Λ1.
(54)
Recalling that x(i)
k = 0 we let
ρk = ρ(i)
k := min(cid:8) inf
j6=i
x(j)
k
2
, dist(0, ∂Ωk)(cid:9);
that is, we set ρk = dist(0, ∂Ωk) if the (x(i)
Observe that by Theorem 1 we have rk = o(ρk) as k → ∞.
Note that Proposition 11 implies the uniform bound
k ) are the only concentration points.
0 ≤ sup
r/2≤x≤r
u2
k(x) − inf
r/2≤x≤r
u2
k(x) ≤ Cr sup
x=r∇u2
k(x) ≤ C
(55)
for 0 ≤ r ≤ ρk.
Lemma 12 Let 0 < ε < 1 and assume that for k ≥ k0 = k0(ε) there holds
inf
0≤r≤ρk
¯uk(r) ≤
ε¯uk(0)
2
.
Let Tk = Tk(ε) ≤ Sk = Sk(ε) ∈]0, ρk] be the smallest numbers such that
¯uk(Tk) = εuk(0), ¯uk(Sk) = εuk(0)/2, respectively. Then
lim
k→∞
rk
Tk
= lim
k→∞
Tk
Sk
= 0.
19
(56)
Moreover for any b < 2 and k ≥ k0 = k0(b) there holds
¯wk(r) ≤ b log(cid:18) rk
r (cid:19) + C for 0 ≤ r ≤ Tk,
and we have
Λk(Tk) = Λ1.
lim
k→∞
(57)
(58)
Proof. Property (56) follows from (55) and our choice of Tk and Sk.
As in the proof of Lemma 4 for a given t ≤ Tk we decompose ¯wk = gk + hk
on Bt, with
∆mhk = 0 in Bt, and gk = ∂ν gk = . . . = ∂m−1
ν
gk = 0 on ∂Bt.
By (54), we get the analogues of Lemma 3 and of (22); that is, for L ≥ L0 =
L0(b), k ≥ k0 = k0(L) there holds
(−1)m−1tm−1(cid:16)t−1(cid:16)t−1 ···(cid:16)t−1
}
m−1 times
{z
(cid:16) ¯w′
k(t) +
b
t(cid:17)(cid:17)′
···(cid:17)′(cid:17)′
}
{z
m−1 times
≤ 0
for all t ∈ [Lrk, Sk]. We now inductively integrate from t to Sk as in Lemma 4.
Using Proposition 11 to bound
∂j
r ¯wk(Sk) =
uk(0)
¯uk(Sk)
Sj
k ¯uk(Sk)∂j
Sj
k
r ¯uk(Sk)
≤
C
εSj
k
,
and recalling (56), for L ≥ L0 and k ≥ k0 we get
t ¯w′
k(t) ≤ −b +
C
ε
t2
S2
k
= −b + o(1) for all Lrk ≤ t ≤ Tk,
with error o(1) → 0 as k → ∞. Since b < 2 is arbitrary, (57) follows as before.
In order to prove (58) observe that the definition of rk gives
ek(r) ≤ Cλku2
≤ Cλkr2m
k(0)emu2
k(0)e
2m(cid:0)1+
k u2
k(0)emu2
k(0)r−2m
k
¯wk (r)
2u2
k
(0)(cid:1) ¯wk(r)
em(ε+1) ¯wk(r) ≤ Cr−2m
k (cid:18) rk
r (cid:19)m(ε+1)b
for Lrk ≤ r ≤ Tk. We then complete the proof as in the radial case.
(cid:3)
For 0 ≤ s < t ≤ ρk set
Nk(s, t) := Λk(t) − Λk(s) = ZBt\Bs
λku2
kemu2
k dx,
and let
Nk(s, t) := Λ(t) − Λ(s) = Z t
s
From (55) we infer
ω2m−1λkr2m−1 ¯u2
kem¯u2
k dr ≤ Nk(s, t).
sup
x=r
emu2
k(x) ≤ Cem¯u2
k(r)
for 0 ≤ r ≤ ρk;
20
(59)
(60)
hence we obtain
sup
x=r
k(x)emu2
u2
k(x) ≤ C(1 + ¯u2
k(r))em¯u2
k (r)
for 0 ≤ r ≤ ρk.
(61)
Then (61) implies
Nk(s, t) ≤ C Nk(s, t) + o(1) for 0 ≤ s ≤ t ≤ ρk,
(62)
with o(1) → 0 as k → ∞. Similarly, setting
ekdσ = ω2m−1λkt2m ¯u2
Pk(t) = tZ∂Bt
k(t)em¯u2
k(t) ≤ Pk(t) := tZ∂Bt
ekdσ,
we can estimate
Pk(t) ≤ C Pk(t) + o(1) for 0 ≤ t ≤ ρk,
(63)
with o(1) → 0 as k → ∞. Finally, from (61) we also obtain the analogue of
(28); that is, we have
Pk(t) ≤ CNk(t/2, t) + o(1) ≤ CPk(t/2) + o(1),
(64)
with error o(1) → 0 as k → ∞.
In particular, we obtain the following improvement of Lemma 12.
Lemma 13 For any 0 < ε < 1, if Tk = Tk(ε) ≤ ρk is as in Lemma 12, then
we have
Proof. Indeed (58) and (62) imply
lim
k→∞
Λk(Tk) = Λ1.
Nk(Lrk, Tk) ≤ C lim
which together with (54) implies the lemma.
lim
L→∞
lim
k→∞
L→∞
Nk(Lrk, Tk) = 0,
lim
k→∞
(cid:3)
If the assumptions of Lemma 12 hold for any 0 < ε < 1 we may proceed
to resolve secondary concentrations at scales o(ρk) as in the radially symmetric
case. Indeed, by Lemmas 12 and 13 we may then choose a subsequence (uk),
numbers εk → 0 as k → ∞ and corresponding numbers sk = Tk(εk) ≤ ρk with
rk/sk → 0 as k → ∞ and such that
Λk(sk) = Λ1,
Nk(Lrk, sk) = 0,
lim
L→∞
lim
k→∞
lim
k→∞
while in addition ¯uk(sk) → ∞ and
¯uk(sk)
¯uk(Lrk)
lim
k→∞
= 0 for every L > 0.
As before, by slight abuse of notation, we set rk = r(1)
k , so that
the analogue of (H1) holds, and iterate. Suppose that for some integer ℓ ≥ 1 we
already have determined numbers
k , sk = s(1)
s(0)
k
:= 0 < r(1)
k < s(1)
k < ··· < r(ℓ)
k < s(ℓ)
k = o(ρk)
satisfying the analogues of (Hℓ,1) up to (Hℓ,4). Similar to Lemma 5 we then
have the following result.
21
Lemma 14 There is a constant C0 = C0(Λ) such that for s(ℓ)
there holds
k ≤ tk = o(ρk)
+ C0 N 2
k (s(ℓ)
k , tk) + o(1),
(65)
Pk(tk)
Nk(s(ℓ)
k , tk) ≤
with error o(1) → 0 as k → ∞.
Proof. For ease of notation we write s = s(ℓ)
proof of Lemma 5, similar to (31) we find
m
k . Replacing wk with ¯wk in the
Nk(s, t) ≤
Pk(t)
2m −Z t
s
ω2m−1r2m ¯uk(r)
uk(0)
¯w′
k(r)ekdr + o(1),
with error o(1) → 0 as k → ∞, uniformly in s ≤ t. Proceeding as in Lemma 5,
from the equation
(t−1(t−1 ··· (t−1
}
m−1 times
{z
¯w′
k(t) )′ ··· )′)′
{z }
m−1 times
= (−1)m
Ak(t)
ω2m−1t4m−3 ,
where Ak is defined by (14), with σk now given by (53), we get
t ¯w′
k(t) = −t2Z ρk
t
t1Z ρk
t1
t2 ···Z ρk
tm−2
Ak(tm−1)
ω2m−1t4m−3
m−1
dtm−1 ··· dt1 + Bk(t, ρk),
where Bk(t, ρk) corresponds to the boundary terms. By arguing as in the proof
of Lemma 4 we see that Bk is a linear combination of terms of the form
t2l+2
ρ2l+2
k
ρj
k∂j
r ¯wk(ρk), 0 ≤ l ≤ m − 2, 1 ≤ j ≤ m − 1.
After multiplication with ¯uk(t)
uk(0) , the resulting terms can be written as
t2l+2
ρ2l+2
k
¯uk(t)ρj
k∂j
r ¯uk(ρk) =
t2l+1
ρ2l+1
k
t¯uk(t)
ρk(¯uk(ρk) + 1)
ρj
k(¯uk(ρk) + 1)∂j
r ¯uk(ρk).
But by Proposition 11 and the analogue of (33) we have
ρj
k(¯uk(ρk) + 1)∂j
r ¯uk(ρk) ≤ C,
t¯uk(t)
ρk(¯uk(ρk) + 1) ≤ C.
Hence for t = tk = o(ρk) we have ¯uk(t)
error o(1) → 0 as k → ∞ we obtain the identity
t2 ···Z ρk
uk(0) Z ρk
t1Z ρk
= t2 ¯uk(t)
¯uk(t)
uk(0)
−t ¯w′
k(t)
t
t1
tm−2
Ak(tm−1)
ω2m−1t4m−3
m−1
dtm−1 ··· dt1.
uk(0) Bk(t, ρk) → 0 as k → ∞, and up to an
The rest of the proof is similar to the proof of Lemma 5.
(cid:3)
On account of (62) and (63) we now obtain the analogue of Lemma 6. The
proof is the same as in the radially symmetric case.
22
Lemma 15 Let C0 = C0(Λ) be the constant appearing in (65), and let tk > s(ℓ)
k
be such that for a subsequence
lim
k→∞
tk
ρk
= 0,
0 < lim
k→∞
Nk(s(ℓ)
k , tk) =: α <
1
2C0
.
(66)
Then
s(ℓ)
k
tk
lim
k→∞
= 0, lim inf
k→∞
Pk(tk) ≥
mα
2
, and lim
L→∞
lim
k→∞
Nk(s(ℓ)
k , tk/L) = 0.
We now closely follow [6]. By the preceding result it suffices to consider the
following two cases. In Case A for any sequence tk = o(ρk) we have
sup
s(ℓ)
k <t<tk
Pk(t) → 0 as k → ∞,
and then in view of Lemma 15 also
lim
L→∞
lim
k→∞
Nk(s(ℓ)
k , ρk/L) = 0,
(67)
thus completing the concentration analysis at scales up to o(ρk).
In Case B for some s(ℓ)
k < tk ≤ ρk there holds
lim
k→∞
k , tk) > 0,
Nk(s(ℓ)
lim sup
k→∞
tk
ρk
= 0.
Then, as in the radial case, from Lemma 15 we infer that for a subsequence (uk)
and suitable numbers r(ℓ+1)
k , tk[ we have
k
∈]s(ℓ)
Nk(s(ℓ)
lim
k→∞
s(ℓ)
k
r(ℓ+1)
k
= 0,
lim
k→∞
k , r(ℓ+1)
k
) > 0, lim inf
k→∞
Pk(r(ℓ+1)
k
) > 0;
(68)
in particular, ¯uk(r(ℓ+1)
k
) → ∞ as k → ∞. Also note that
lim
L→∞
lim sup
k→∞
Nk(s(ℓ)
k , r(ℓ+1)
k
/L) = lim
k→∞
r(ℓ+1)
k
ρk
= lim
k→∞
tk
ρk
= 0.
(69)
Moreover, analoguous to Proposition 7 we have the following result, which is a
special case of Proposition 17 below.
Proposition 16 There exists a subsequence (uk) such that
η(ℓ+1)
k
(x) := ¯uk(r(ℓ+1)
k
)(uk(r(ℓ+1)
k
x) − ¯uk(r(ℓ+1)
k
)) → η(ℓ+1)(x)
in C2m−1
for a suitable constant c(ℓ+1).
(R2m \ {0}) as k → ∞, where η(ℓ+1)
loc
0
:= η(ℓ+1) + c(ℓ+1) solves (8), (9)
From Proposition 16 the desired energy quantization result at the scale r(ℓ+1)
k
follows as in the radial case.
If ρk ≥ ρ0 > 0 we can argue as in [13], p. 416, to obtain numbers s(ℓ+1)
k
satisfying
lim
L→∞
lim
k→∞
Λk(s(ℓ+1)
k
) = (ℓ + 1)Λ1,
(70)
23
and such that
lim
L→∞
lim
k→∞
(Λk(s(ℓ+1)
k
) − Λk(Lr(ℓ+1)
k
)) = lim
k→∞
r(ℓ+1)
k
s(ℓ+1)
k
= lim
k→∞
s(ℓ+1)
k
= 0,
while ¯uk(s(ℓ+1)
k
) → ∞ as k → ∞. Moreover, for any L ≥ 1 we have
lim
k→∞
¯uk(s(ℓ+1)
)
¯uk(Lr(ℓ+1)
k
k
)
= 0.
(71)
By iteration we then establish (70), (71) up to ℓ + 1 = ℓ0 for some maximal
index ℓ0 ≥ 1 where Case A occurs and thus complete the concentration analysis
near the point x(i), getting
lim
k→∞
Λk(ρ0) = ℓ0Λ1.
k
k
k
k
k
k
k
). Then as before we can define numbers s(ℓ+1)
, ρk] there holds ¯uk(t) ≥ ε0 ¯uk(r(ℓ+1)
If ρk → 0 as k → ∞, we distinguish the following two cases. In Case 1
for some ε0 ∈]0, 1[ and all t ∈ [r(ℓ+1)
). The
decay estimate that we established in Lemma 12 then remains valid through-
out this range and (70) holds true for any choice s(ℓ+1)
= o(ρk). Again the
concentration analysis at scales up to o(ρk) is complete. In Case 2, for any
ε ∈]0, 1[ there is a minimal Tk = Tk(ε) ∈ [r(ℓ+1)
, ρk] as in Lemma 12 such that
¯uk(Tk) = ε¯uk(r(ℓ+1)
< ρk with
¯uk(s(ℓ+1)
) → ∞ as k → ∞ so that (70), (71) also hold true, and we proceed by
iteration up to some maximal index ℓ0 ≥ 1 where either Case 1 or Case A holds
with final radii r(ℓ0)
For the concentration analysis at the scale ρk first assume that for some
number L ≥ 1 there is a sequence (xk) such that ρk/L ≤ Rk(xk) ≤ xk ≤ Lρk
and
(72)
By Proposition 11 we may assume that xk = ρk. Moreover, (55) implies
that dist(0, ∂Ωk)/ρk → ∞ as k → ∞. As in [13], Lemma 4.6, we then have
¯uk(ρk)/¯uk(r(ℓ0)
) → 0 as k → ∞, ruling out Case 1; that is, at scales up to o(ρk)
we end with Case A. The desired quantization result at the scale ρk then is a
consequence of the following result similar to [13], Proposition 4.7, whose proof
may be easily carried over to the present situation.
k(xk) ≥ ν0 > 0.
λkxk2mu2
k(xk)emu2
, respectively.
, s(ℓ0)
k
k
k
Proposition 17 Assuming (72), there exist a finite set S0 ⊂ R2m and a sub-
sequence (uk) such that
ηk(x) := uk(xk)(uk(ρkx) − uk(xk)) → η(x)
in C2m−1
η0 = η + c0 solves (8), (9).
loc
(R2m \ S0) as k → ∞, where for a suitable constant c0 the function
By Proposition 17 in case of (72) there holds
lim
L→∞
k→∞Z{x∈Ω;
lim
ρk
L ≤Rk(x)≤x≤Lρk}
ekdx = Λ1.
(73)
24
Letting
Xk,1 = X (i)
k,1 = {x(j)
k ;∃C > 0 : x(j)
k ≤ Cρk for all k}
and carrying out the above blow-up analysis up to scales of order o(ρk) also on
all balls of center x(j)
k ∈ Xk,1, then from (71) and (73) we have
lim
L→∞
lim
k→∞
Λk(Lρk) = Λ1(1 + I1),
where I1 is the total number of bubbles concentrating at the points x(j)
at scales o(ρk).
k ∈ X (i)
k,1
On the other hand, if (72) fails to hold clearly we have
lim
L→∞
lim sup
k→∞ Z{x∈Ω;
ρk
L ≤Rk(x)≤x≤Lρk}
ekdx = 0,
(74)
and the energy estimate at the scale ρk again is complete.
In order to deal with secondary concentrations around x(i)
k = 0 at scales
exceeding ρk, with Xk,1 defined as above we let
x(j)
k
2
ρk,1 = ρ(i)
k,1 = min(cid:8)
inf
k /∈Xk,1}
{j;x(j)
, dist(0, ∂Ωk)(cid:9);
that is, we again set ρk,1 = dist(0, ∂Ωk), if {j; x(j)
/∈ Xk,1} = ∅. From this
definition it follows that ρk,1/ρk → ∞ as k → ∞. Then, using the obvious
analogue of Lemma 15, either we have
k
lim
L→∞
lim sup
k→∞
Nk(cid:16)Lρk,
ρk,1
L (cid:17) = 0,
and we iterate to the next scale; or there exist radii tk ≤ ρk,1 such that tk/ρk →
∞, tk/ρk,1 → 0 as k → ∞ and a subsequence (uk) such that
Pk(tk) ≥ ν0 > 0 for all k.
(75)
The argument then depends on whether (72) or (74) holds. In case of (72), as in
[13], Lemma 4.6, the bound (75) and Proposition 7 imply that ¯uk(tk)/¯uk(ρk) →
0 as k → 0. Then we can argue as in Case A for r ∈ [Lρk, ρk,1] for sufficiently
large L, and we can continue as before to resolve concentrations in this range
of scales.
In case of (74) we further need to distinguish whether Case A or Case 1
holds at the final stage of our analysis at scales o(ρk). In fact, for the following
estimates we also consider all points x(j)
k . Recalling that in
Case A we have (71) (with index ℓ0 instead of ℓ + 1) and (67), on account of (74)
for a suitable sequence of numbers s(0)
k,1 → ∞ as
k → ∞ we find
k ∈ X (i)
k,1 such that s(0)
k,1/ρk → ∞, tk/s(0)
k,1 in place of x(i)
lim
L→∞
lim
k→∞(cid:16)Λ(s(0)
k,1) − Xx(j)
k ∈X (i)
k,1
k (Lr(ℓ(j)
Λ(j)
0 )
k
)(cid:17) = 0,
0 )
k (r) and r(ℓ(j)
k . In particular, with such a choice of s(0)
are computed as above with respect to the concentration
k,1 we find the intermediate
k
where Λ(j)
point x(j)
quantization result
lim
k→∞
Λk(s(0)
k,1) = Λ1I1
25
analogous to (70), where I1 is defined as above. In Case 1 we can obtain the
same conclusion by our earlier reasoning. Moreover, in Case 1 we can argue as
in [13], Lemma 4.8, to conclude that ¯uk(tk)/¯uk(Lr(ℓ(j)
0 )
) → 0 for any L ≥ 1 as
k → 0; therefore, similar to (71) in Case A, we can achieve that for any L ≥ 1
we have
k
lim
k→∞
¯uk(s(0)
k,1)
¯uk(Lr(ℓ(j)
0 )
k
)
= lim
k→∞
r(ℓ(j)
0 )
k
s(0)
k,1
= lim
k→∞
ρk
s(0)
k,1
= lim
k→∞
s(0)
k,1
tk
= 0
for all x(j)
k ∈ X (i)
k,1 where Case 1 holds, similar to (Hℓ).
We then finish the argument by iteration. For ℓ ≥ 2 we inductively define
the sets
Xk,ℓ = X (i)
k,ℓ = {x(j)
k ;∃C > 0 : x(j)
k ≤ Cρk,ℓ−1 for all k}
and we let
ρk,ℓ = ρ(i)
k,ℓ = min(cid:8)
inf
k /∈Xk,1}
{j;x(j)
x(j)
k
2
, dist(0, ∂Ωk)(cid:9);
that is, as before, we set ρk,ℓ = dist(0, ∂Ωk), if {j; x(j)
k,ℓ} = ∅. Iteratively
performing the above analysis at all scales ρk,ℓ, thereby exhausting all concen-
tration points x(j)
k , upon passing to further subsequences, we finish the proof of
Theorem 2.
/∈ X (i)
k
3.1 Proof of Proposition 11
Our proof of Proposition 11 is modelled on the proof of [5], Proposition 2.
In fact, the first steps of the proof seem almost identical to the corresponding
arguments in [5]. The special character of the present problem only enters at the
last stage, where we also need to distinguish the cases ℓ = 1 and 2 ≤ ℓ ≤ 2m− 1.
Fix any index 1 ≤ ℓ ≤ 2m − 1. The following constructions will depend on
this choice; however, for ease of notation we suppress the index ℓ in the sequel.
Set Rk(x) := inf 1≤j≤I x − x(j)
k and choose points yk such that
Rℓ
k(yk)uk(yk)∇ℓuk(yk) = sup
Ω
Rℓ
kuk∇ℓuk =: Lk.
Suppose by contradiction that Lk → ∞ as k → ∞. From Theorem 1 then it
follows that sk := Rk(yk) → 0 as k → ∞. Set
and let
Ωk := {y; yk + sky ∈ Ω}
vk(y) := uk(yk + sky),
y ∈ Ωk.
Observe that for 1 ≤ j ≤ m via Sobolev's embedding from (3) we obtain
k∇jvkk2
L
2m
j (Ωk) ≤ Ck∇mvkk2
L2(Ωk) = CZΩk
vk(−∆)mvkdx ≤ C.
(76)
26
Also let
and set
Clearly then we have
y(i)
k :=
x(i)
k − yk
sk
, 1 ≤ i ≤ I
Sk := {y(i)
k ; 1 ≤ i ≤ I}.
dist(0, Sk) = inf
1≤i≤I y(i)
k = 1
and
sup
y∈Ωk
(dist(y, Sk)ℓvk(y)∇ℓvk(y)) = vk(0)∇ℓvk(0) = Lk → ∞
(77)
as k → ∞. Moreover (7) implies
0 ≤ vk(−∆)mvk = λks2m
k v2
kemv2
k ≤
C
dist(y, Sk)2m .
(78)
Since limk→∞ sk = 0, we may assume that as k → ∞ the domains Ωk
exhaust a half-space
Ω0 = R2m−1×] − ∞, R0[,
where 0 < R0 ≤ ∞. We may also assume that either limk→∞ y(i)
limk→∞ y(i)
points of Sk, satisfying dist(0, S0) = 1. For R > 0 denote
k = ∞ or
k = y(i), 1 ≤ i ≤ I, and we let S0 be the set of these accumulation
Observing that λks2m
B1/R(y).
Kk,R := Ωk ∩ BR(0)\ [y∈S0
k → 0, from (78) we obtain that
k→∞k∆mvkkL∞(Kk,R) = 0 for every R > 0.
lim
(79)
Lemma 18 We have R0 = ∞, hence Ω0 = R2m.
Proof. Suppose by contradiction that R0 < ∞. Choosing R = 2R0 and observ-
ing that by (2) for 0 ≤ j < ℓ ≤ 2m − 1 we have ∂j
k = 0 on ∂Ωk, from Taylor's
formula and (77) we conclude
ν v2
v2
k
sup
Kk,R
vk(0)∇ℓvk(0) ≤ C = C(R).
, we then have 0 ≤ wk ≤ C on Kk,R. Using (76),
Letting wk :=
vk
√vk(0)∇vk(0)
Sobolev's embedding, (77) and (79) we infer
k∇wkkL2m(Ωk) + k∇2wkkLm(Ωk) + k∆mwkkL∞(Kk,R) → 0 as k → ∞.
Since ∂j
that wk → 0 in C2m−1,α
wk(0)∇ℓwk(0) = 1.
νwk = 0 on ∂Ωk for 0 ≤ j ≤ m − 1, it follows from elliptic regularity
(Kk,R) for 0 < α < 1, contradicting the fact that
(cid:3)
loc
27
Lemma 19 As k → ∞ we have vk(0) → ∞ and
vk
vk(0) → 1 in C2m−1,α
loc
(R2m\S0).
Proof. First observe that
ck := sup
B1/2
vk → ∞ as k → ∞.
Indeed, otherwise (76), (79) and elliptic regularity would contradict (77). Let-
ting wk := vk
ck
, from (76) and (79) for any R > 0 we have
k∇wkkL2m(Ωk) + k∇2wkkLm(Ωk) + k∆mwkkL∞(Kk,R) → 0 as k → ∞,
whence wk → w ≡ const in C2m−1,α
we obtain
loc
(R2m\S0). Recalling that dist(0, S0) = 1,
w ≡ sup
B1/2
w = lim
k→∞
sup
B1/2
wk = 1.
In particular we conclude that vk(0)
ck
vk(0) = ckwk(0) → ∞,
= wk(0) → 1 as k → ∞ and therefore
(R2m\S0), as claimed. (cid:3)
For the final argument now we need to distinguish the cases ℓ = 1 and
wk(0) → 1 in C2m−1,α
vk(0) = wk
loc
vk
2 ≤ ℓ ≤ 2m − 1. Consider first the case ℓ = 1. Set
vk(y) − vk(0)
vk(y) :=
∇vk(0)
.
From (77) and Lemma 19 we infer
∇vk(y) =
vk(0)
vk(y)
vk(y)∇vk(y)
vk(0)∇vk(0) ≤
1 + o(1)
dist(y, S0)
,
(80)
with error o(1) → 0 in C2m−1,α
(R2m\S0) as k → ∞. Since vk(0) = 0, from (80)
we conclude that vk is bounded in C1(Kk,R) for every R > 0, uniformly in k.
Moreover, (78) and Lemma 19 give
loc
∆mvk =
vk(0)
vk
vk∆mvk
vk(0)∇vk(0) ≤ C(R)
vk(0)
Lkvk → 0
(81)
loc
loc
∆mv = 0,
uniformly on Kk,R as k → ∞, for any R > 0. The sequence vk then is bounded
in C2m−1,α
(R2m\S0) for any α < 1, and by Arzel`a-Ascoli's theorem we can
assume that vk → v in C2m−1,α
v(0) = 0,
(R2m\S0), where v satisfies
∇v(0) = 1,
∇v(y) ≤
dist(y, S0)
Fix a point x0 ∈ S0. For any r ∈]0, dist(x0, S0\{x0})/2[ let ϕ ∈ C∞
0 (Br(x0))
be a function 0 ≤ ϕ ≤ 1 such that ϕ ≡ 1 in Br/2(x0), and satisfying ∇jϕ ≤
Cr−j for 0 ≤ j ≤ m. Integration by parts yields
(∇ϕvk · ∇∆m−1vk + ϕvk∆mvk)dx
ZBr (x0)
(82)
1
.
= −ZBr (x0)
ϕ∇vk · ∇∆m−1vk dx =: I.
28
(83)
Again integrating by parts m − 1 times, we obtain
I = (−1)mZBr(x0) Xα=m−1
∂α(ϕ∇vk) · ∇∂αvk dx,
so that by Holder's inequality and (76) this term may be bounded
I ≤ C X1≤j≤m
≤ C X1≤j≤m
rj−mZBr(x0) ∇jvk∇mvkdx
j k∇mvkkL2 ≤ C.
k∇jvkkL
2m
Similarly, we have
0 ≤ ZBr (x0)
ϕvk(−∆)mvk dx ≤ C,
and from (83) we conclude the bound
ZBr (x0)∇ϕvk · ∇∆m−1vkdx(cid:12)(cid:12)(cid:12) ≤ C.
(cid:12)(cid:12)(cid:12)
(84)
Observe that ∇ϕ = 0 in Br/2(x0). By Lemma 19 therefore the integral on the
left-hand side equals
ZBr (x0)∇ϕvk · ∇∆m−1vkdx
= (1 + o(1))vk(0)∇vk(0)ZBr (x0) ∇ϕ · ∇∆m−1vkdx
= −(1 + o(1))vk(0)∇vk(0)ZBr(x0)
ϕ∆mvkdx.
Since (−∆)mvk ≥ 0, it follows that
ZBr/2(x0)
(−∆)mvkdx ≤
C
vk(0)∇vk(0)
= CL−1
k → 0 as k → ∞.
loc(R2m). Therefore ∆mv ≡ 0
Recalling (81), we infer that ∆mvk → 0 in L1
in R2m. Since from (82) we have v(y) ≤ C(1 + y) for y ∈ R2m, we may
now invoke a Liouville-type theorem as in [7], Theorem 5, to see that v is a
polynomial of degree at most 2m− 2 if m > 1 and of degree at most 1 if m = 1.
But then (82) implies that v ≡ 0, contradicting the fact that ∇vk(0) = 1. This
completes the proof in the case ℓ = 1.
In the case when 2 ≤ ℓ ≤ m − 1 we set
vk(y) :=
vk(y) − vk(0)
∇ℓvk(0)
.
As shown above we have
dist(y, Sk)vk(y)∇vk(y) ≤ C sup
x∈Ω
Rk(x)uk(x)∇uk(x) ≤ C;
29
∇vk ≤
C(1 + o(1))
vk(0)∇ℓvk(0) dist(y, S0)
=
C(1 + o(1))
Lk dist(y, S0) → 0.
(85)
hence Lemma 19 implies with error o(1) → 0 in C2m−1,α
that
loc
(R2m\S0) as k → ∞
Notice that this is stronger than its analogue (80). As in the case ℓ = 1 we have
∆mvk =
vk(0)
vk
vk∆mvk
vk(0)∇ℓvk(0) ≤
C(R)
Lk → 0
(86)
uniformly on Kk,R as k → ∞, for any R > 0, hence vk → v in C2m−1,α
where v satisfies
loc
(R2m\S0),
∆mv = 0,
v(0) = 0,
∇ℓv(0) = 1.
On the other hand (85) implies ∇v ≡ 0, contradiction. This completes the
proof.
(cid:3)
Appendix
We collect here some technical results used in the above sections. The proof of
the following proposition can be found in [7], Prop. 4.
Proposition 20 Let ∆mh = 0 in B2 ⊂ Rn. For every 0 ≤ α < 1 and ℓ ≥ 0
there is a constant C(ℓ, α) independent of h such that
khkC ℓ,α(B1) ≤ C(ℓ, α)khkL1(B2).
By a simple covering argument Proposition 20 can be extended to the case
of annuli.
Proposition 21 Let ∆mh = 0 in B2L(0)\B1/2L(0) ⊂ Rn for some L ≥ 1. For
every 0 ≤ α < 1 and ℓ ≥ 0 there is a constant C = C(ℓ, α, L) such that
khkC ℓ,α(BL(0)\B1/L(0)) ≤ CkhkL1(B2L(0)\B1/2L(0)).
Lemma 22 Let g ∈ C∞(Bt), where Bt = Bt(0) ⊂ Rn for some n ∈ N, t > 0.
Assume that g is radially symmetric and satisfies
g = ∂νg = . . . = ∂m−1
ν
g = 0
on ∂Bt.
Then
Z∂Bt
tm−1∂m
ν g dσ = Z t
0
t2 ···Z tm−1
0
tm(cid:18)ZBtm
∆mgdx(cid:19)dtm . . . dt2.
Proof. For m = 1 equation (88) simply reduces to
Z∂Bt
∂νg dσ = ZBt
∆g dx.
For m = 2 consider the function ϕ(x) = x · ∇g(x) with
∆(x · ∇g)dx = ZBt
∂νϕ dσ = ZBt
∆ϕ dx = ZBt
Z∂Bt
(x · ∇∆g + 2∆g)dx
30
(87)
(88)
(89)
and note that the condition ∂νg = 0 on ∂Bt and (89) imply
Z∂Bt
∂ν ϕ dσ = Z∂Bt
∂ν(x · ∇g)dσ = Z∂Bt
t∂2
ν g dσ, and ZBt
∆g dx = 0.
Thus from Fubini's theorem we obtain the desired identity
Z∂Bt
t∂2
ν g dσ = ZBt
= Z t
0
x · ∇∆g dx
t2(cid:18)Z∂Bt2
∂ν ∆g dσ(cid:19)dt2 = Z t
0
t2(cid:18)ZBt2
∆2g dx(cid:19)dt2.
We now proceed by induction. Assume that the lemma is true for m − 1.
Choosing ϕ(x) = x · ∇g(x) with
ϕ = ∂νϕ = . . . = ∂m−2
ν
ϕ = 0 on ∂Bt
we get
tm−1∂m
Z∂Bt
= Z t
0
ν gdσ = Z∂Bt
t2 ···Z tm−2
0
tm−1ZBtm−1
tm−2∂m−1
ν
tm−2∂m−1
(t∂ν g)dσ = Z∂Bt
∆m−1(x · ∇g)dxdtm−1 . . . dt2 =: I.
ν
(x · ∇g)dσ
Observe that ∆m−1(x · ∇g) = x · ∇∆m−1g + 2(m − 1)∆m−1g, hence
I = Z t
0
t2 ···Z tm−2
+ 2(m − 1)Z t
tm−1ZBtm−1
t2 ···Z tm−2
0
0
0
(x · ∇∆m−1g)dxdtm−1 . . . dt2
tm−1ZBtm−1
∆m−1gdxdtm−1 . . . dt2
= II + III.
By inductive hypothesis and (87) the contribution from the second term is
III = 2(m − 1)Z∂Bt
tm−2∂m−1
ν
gdσ = 0,
and our claim follows from writing
ZBtm−1
x · ∇∆m−1gdx = Z tm−1
= Z tm−1
0
0
tmZ∂Btm
tmZBtm
∂ν∆m−1gdσdtm
∆mgdxdtm.
(90)
(cid:3)
Lemma 23 Let u ∈ C∞(Rn) ∩ Lp(Rn), for some p ≥ 1, satisfy ∆ju = 0 for
some integer j > 0. Then u ≡ 0.
31
Proof. We first claim that
R→∞ Z
lim
BR(ξ)
udx = 0
for every ξ ∈ Rn. Indeed by Jensen's inequality
updx(cid:19) 1
udx ≤ (cid:18) Z
Z
(cid:12)(cid:12)(cid:12)(cid:12)
udx(cid:12)(cid:12)(cid:12)(cid:12) ≤ Z
p
1
Rn/pkukLp(Rn) → 0,
≤
BR(ξ)
BR(ξ)
BR(ξ)
as R → ∞. By Pizzetti's formula (see [10]) we have constants c1, . . . , cj−1 such
that
Z
BR(ξ)
udx = u(ξ) + c1R2∆u(ξ) + ··· + cj−1R2j−2∆j−1u(ξ) =: P (R).
Taking the limit as R → ∞ we see at once that the polynomial P (R) is iden-
tically 0, and in particular u(ξ) = P (0) = 0. Since ξ was arbitrary the proof is
complete.
(cid:3)
Lemma 24 There holds
capHm ({0}) = inf{k∇mϕkL2; ϕ ∈ X} = 0,
where
X = {ϕ ∈ C∞
0 (B1(0)); 0 ≤ ϕ ≤ 1,∃r > 0 : ϕ(x) = 1 for x ≤ r}.
Proof. Let f (x) = log log log(1/x) with ∇mf ∈ L2(Be−e (0)) and fix g ∈ C∞(R)
with 0 ≤ g ≤ 1 satisfying g(s) = 0 for s ≤ 0, g(s) = 1 for s ≥ 1. Letting
ϕk(x) = g(f (x) − k), k ∈ N,
we find ϕk ∈ X for all k and k∇mϕkkL2 → 0 as k → ∞.
(cid:3)
References
[1] Adimurthi, O. Druet, Blow-up analysis in dimension 2 and a sharp form
of Trudinger-Moser inequality, Comm. Partial Differential Equations 29
(2004), 295-322.
[2] Adimurthi, F. Robert, M. Struwe, Concentration phenomena for Li-
ouville's equation in dimension 4, J. Eur. Math. Soc. 8 (2006), 171-180.
[3] Adimurthi, M. Struwe, Global compactness properties of semilinear el-
liptic equations with critical exponential growth, J. Functional Analysis 175
(2000), 125-167.
[4] S-Y. A. Chang Non-linear Elliptic Equations in Conformal Geometry,
Zurich lecture notes in advanced mathematics, EMS (2004).
32
[5] O. Druet, Multibumps analysis in dimension 2: quantification of blow-up
levels, Duke Math. J. 132 (2006), 217-269.
[6] T. Lamm, F. Robert, M. Struwe The heat flow with a critical exponential
nonlinearity, J. Functional Anal. 257 (2009) 2951-2998.
[7] L. Martinazzi Classification of solutions to the higher order Liouville's
equation in R2m, Math. Z. 263 (2009), 307-329.
[8] L. Martinazzi A threshold phenomenon for embeddings of H m
0 into Orlicz
spaces, Calc. Var. Partial Differential Equations 36 (2009), 493-506.
[9] L. Martinazzi, M. Petrache Asymptotics and quantization for a mean-
field equation of higher order, Comm. Partial Differential Equations 35
(2010), 1-22.
[10] P. Pizzetti Sulla media dei valori che una funzione dei punti dello spazio
assume alla superficie di una sfera, Rend. Lincei 18 (1909), 182-185.
[11] F. Robert, M. Struwe Asymptotic profile for a fourth order PDE with
critical exponential growth in dimension four, Adv. Nonlin. Stud. 4 (2004),
397-415.
[12] F. Robert, J.-C. Wei Asymptotic behavior of a fourth order mean field
equation with Dirichlet boundary condition, Indiana Univ. Math. J. 57
(2008), 2039-2060.
[13] M. Struwe Quantization for a fourth order equation with critical expo-
nential growth, Math. Z. 256 (2007), 397-424.
[14] J.-C. Wei Asymptotic behavior of a nonlinear fourth order eigenvalue prob-
lem, Comm. Partial Differential Equations 21 (1996), 1451-1467.
33
|
1809.08465 | 2 | 1809 | 2019-12-04T07:32:19 | The Segal-Bargmann Transform on Classical Matrix Lie Groups | [
"math.FA"
] | We study the complex-time Segal-Bargmann transform $\mathbf{B}_{s,\tau}^{K_N}$ on a compact type Lie group $K_N$, where $K_N$ is one of the following classical matrix Lie groups: the special orthogonal group $\mathrm{SO}(N,\mathbb{R})$, the special unitary group $\mathrm{SU}(N)$, or the compact symplectic group $\mathrm{Sp}(N)$. Our work complements and extends the results of Driver, Hall, and Kemp on the Segal-Bargman transform for the unitary group $\mathrm{U}(N)$. We provide an effective method of computing the action of the Segal-Bargmann transform on \emph{trace polynomials}, which comprise a subspace of smooth functions on $K_N$ extending the polynomial functional calculus. Using these results, we show that as $N\to\infty$, the finite-dimensional transform $\mathbf{B}_{s,\tau}^{K_N}$ has a meaningful limit $\mathscr{G}_{s,\tau}^{(\beta)}$ (where $\beta$ is a parameter associated with $\mathrm{SO}(N,\mathbb{R})$, $\mathrm{SU}(N)$, or $\mathrm{Sp}(N)$), which can be identified as an operator on the space of complex Laurent polynomials. | math.FA | math |
The Segal-Bargmann Transform on Classical Matrix Lie
Groups
Alice Z. Chan∗a
aDepartment of Mathematics, University of California, San Diego, La Jolla, CA 92093
Abstract
We study the complex-time Segal-Bargmann transform BKN
s,τ on a compact type Lie group
KN , where KN is one of the following classical matrix Lie groups: the special orthogonal
group SO(N, R), the special unitary group SU(N ), or the compact symplectic group
Sp(N ). Our work complements and extends the results of Driver, Hall, and Kemp on the
Segal-Bargman transform for the unitary group U(N ). We provide an effective method
of computing the action of the Segal-Bargmann transform on trace polynomials, which
comprise a subspace of smooth functions on KN extending the polynomial functional
calculus. Using these results, we show that as N → ∞, the finite-dimensional transform
BKN
s,τ has a meaningful limit Gs,τ which can be identified as an operator on the space of
complex Laurent polynomials.
Keywords: Segal-Bargmann transform; Heat kernel analysis on Lie groups
Contents
1 Introduction
1.1 The classical Segal-Bargmann transform . . . . . . . . . . . . . . . . . . .
1.2 Main definitions and theorems
. . . . . . . . . . . . . . . . . . . . . . . .
2 Background
2.1 Heat kernels on Lie groups . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Notation and definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3 The Segal-Bargmann transform on SO(N, R)
3.1 Magic formulas and derivative formulas
3.2
3.3
Intertwining formulas for ∆SO(N,R)
Intertwining formulas for ASO(N,C)
s,τ
. . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . .
2
2
3
7
7
8
10
10
16
23
∗Supported by NSF GRFP under Grant No. DGE-1650112
Email address: [email protected] (Alice Z. Chan)
Preprint submitted to Journal of Functional Analysis
December 5, 2019
4 Limit theorems for the Segal-Bargmann transform on SO(N, R)
4.1 Concentration of measures . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 The free Segal-Bargmann transform for SO(N, R) . . . . . . . . . . . . . .
5 The Segal-Bargmann transform on Sp(N )
5.1 Two realizations of the compact symplectic group . . . . . . . . . . . . . .
5.2 Magic formulas and derivative formulas
. . . . . . . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . . . . . . . .
5.3
. . . . . . . . . . . . . . . . . . . . . .
5.4
5.5 Limit theorems for the Segal-Bargmann transform on Sp(N )
. . . . . . .
Intertwining formulas for ∆Sp(N )
Intertwining formulas for ASp(N,C)
s,τ
5.6 Extending the magic formulas and intertwining formulas to fSp(N ) . . . .
6 The Segal-Bargmann transform on SU(N )
29
29
33
37
38
40
45
47
50
51
53
1. Introduction
1.1. The classical Segal-Bargmann transform
In this paper, we consider a complex-time generalization of the classical Segal-Bargmann
transform on the matrix Lie groups SO(N, R), SU(N ), and Sp(N ), and analyze its limit
as N → ∞. We begin with a brief discussion of the classical Segal-Bargmann transform
and its generalizations to Lie groups of compact type. For t > 0, let ρt denote the heat
kernel with variance t on Rd:
This has an entire analytic continuation to Cd, given by
ρt(x) = (2πt)−d/2 exp(−x2/2t).
2t (cid:17) .
(ρt)C(z) = (2πt)−d/2 exp(cid:16)−
(Btf )(z) :=ZRd
(ρt)C(z − y)f (y) dy.
z · z
If f ∈ L1
loc(Rd) is of sufficiently slow growth, we can define
(1.1)
(1.2)
The map f 7→ Btf is called the classical Segal-Bargmann transform, due to the work
of the eponymous authors in [2, 1, 17, 15, 16]. If we let µt denote the heat kernel on Cd,
now with variance t
2 :
µt(z) = (πt)−d exp(−z2/t),
then the main result is that the map
Bt : L2(Rd, ρt) → HL2(Cd, µt)
is a unitary isomorphism, where HL2(Cd, µt) is the space of square-integrable holomor-
phic functions on Cd.
2
1.2. Main definitions and theorems
In [10], Hall showed that the Segal-Bargmann transform can be extended to compact
Lie groups, and in [6], the transform was further extended by Driver to Lie groups of
compact type. Later, the behavior of the transform on the unitary group U(N ) as
N → ∞ was analyzed independently by C´ebron in [5] and Driver, Hall, and Kemp in [8].
The authors of the latter paper subsequently introduced a complex-time generalization
of the Segal-Bargmann transform for all compact type Lie groups in [9]. We use this
version of the transform (see Definition 1.1) in our analysis of the special orthogonal
group SO(N, R), the special unitary group SU(N ), and the compact symplectic group
Sp(N ). We now outline the main results of this paper, deferring a fuller discussion of
the requisite background and definitions to Section 2.
Let K denote a Lie group of compact type with Lie algebra k. Then K possesses a
left-invariant metric whose associated Laplacian ∆K is bi-invariant (cf. (2.1)). For each
s > 0, there is an associated heat kernel ρK
s ∈ C∞(K, (0,∞)) satisfying
f (k)ρK
s (xk−1) dk
for all f ∈ L2(K).
(1.3)
s
2 ∆K f(cid:1) (x) =ZK
(cid:0)e
f (xk)ρK
s (k) dk =ZK
Let KC denote the complexification of K and C+ = {τ = t + iu : t > 0, u ∈ R} denote
the right half-plane. As shown in [9, Theorem 1.2], the heat kernel (ρK
s )s>0 possesses a
unique entire analytic continuation
ρK
C : C+ × KC → C
such that ρK
C (s, x) = ρs(x) for all s > 0 and x ∈ K ⊆ KC. Thus we can proceed as in
the classical case and define the Segal-Bargmann transform for compact type Lie groups
using a group convolution formula similar to (1.1) (see [6] for details).
Definition 1.1 (Complex-time Segal-Bargmann transform). Let s > 0 and τ = t + iθ ∈
D(s, s), and let ρK
C be as above. For z ∈ KC and f ∈ L2(K, ρK
s and ρK
s ), define
(BK
s,τ f )(z) :=ZK
ρK
C (τ, zk−1)f (k) dk
for z ∈ KC.
(1.4)
The complex-time Segal-Bargmann transform BK
certain L2 and holomorphic L2 spaces on K and KC:
s,τ is an isometric isomorphism between
Theorem 1.2 (Driver, Hall, and Kemp,
compact-type Lie group, and set s > 0 and τ ∈ D(s, s). Let µKC
rescaled) heat kernel on KC defined by (2.4). Then BK
f ∈ L2(K, ρK
[9, Theorem 1.6]). Let K be a connected,
s,τ denote the (time-
s,τ f is holomorphic on KC for each
s ). Moreover,
BK
s,τ : L2(K, ρK
s ) → HL2(KC, µKC
s,τ )
(1.5)
is a unitary isomorphism.
3
If we set K = Rd and s = τ > 0, we recover the main result for the classical Segal-
Bargmann transform from Section 1.1.
Every compact Lie group K has a faithful representation. Hence, viewing K as a matrix
group, we can extend BK
s,τ to matrix-valued functions. Let MN (R) denote the algebra of
N × N real matrices with unit IN , and let MN (C) denote the algebra of N × N complex
matrices. Let GLN (C) denote the group of all invertible matrices in MN (C). Define a
scaled Hilbert-Schmidt norm on MN (C) by
kAk2
MN (C) :=
1
N
Tr(AA∗),
A ∈ MN (C),
(1.6)
where Tr denotes the usual trace.
Definition 1.3 (Boosted Segal-Bargmann transform). Let K ⊆ GLN (C) be a compact
matrix Lie group with complexification KC. Let F be an MN (C) valued function on K or
KC, and let kFkMN (C) denote the scalar-valued function A 7→ kF (A)kMN (C). Fix s > 0
and τ ∈ D(s, s), and let
L2(K, ρK
L2(KC, µKC
s ; MN (C)) = {F : K → MN (C); kFkMN (C) ∈ L2(K, ρK
s )},
s,τ ; MN (C)) = {F : KC → MN (C); kFkMN (C) ∈ L2(KC, µKC
s,τ )}.
The norms defined by
L2(K,ρK
kFk2
kFk2
L2(KC,µ
s ;MN (C)) :=ZK kF (A)k2
:=ZKC kF (A)k2
KC
s,τ ;MN (C))
MN (C) ρK
s (A) dA
MN (C) µKC
s,τ (A) dA
and
(1.7)
(1.8)
make these spaces into Hilbert spaces. Let HL2(KC, µKC
denote the subspace of matrix-valued holomorphic functions.
s,τ ; MN (C)) ⊆ L2(KC, µKC
s,τ ; MN (C))
The boosted Segal-Bargmann transform
BK
s,τ : L2(K, ρK
s ; MN (C)) → HL2(KC, µKC
s,τ ; MN (C))
is the unitary isomorphism determined by applying BN
s,τ componentwise; i.e.,
BK
s,τ (f · A) = BK
s,τ f · A for f ∈ L2(K, ρK
s ) and A ∈ MN (C).
In this paper, we consider the case when K = KN is one of SO(N, R), SU(N ), or Sp(N ).
We provide an effective method of computing BKN
s,τ on the space of polynomials, and
show that in this setting, the transform has a meaningful limit as N → ∞. However, the
boosted Segal-Bargmann transform typically does not preserve the space of polynomials
on K. For example, consider the polynomial PN (A) = A2 on SO(N, R). Let Tr denote
the usual trace and set tr = 1
N Tr. From Example 4.11, we see that
(BSO(N,R)
s,t
PN )(A) =
1
N
(1 − e−t)IN +
1
2
e−t(1 + e
4
2t
N )A2 −
N
2
e−t(−1 + e
2t
N )A tr A, (1.9)
This example highlights that in general, BK
s,τ maps the space of polynomial functions
on K into the larger space of trace polynomials, first introduced in [8, Definition 1.7],
which we now briefly describe (see Section 3.2 for the full details).
Let C[u, u−1; v] denote the algebra of polynomials in the commuting variables u, u−1, and
v = {v±1, v±2, . . .}. For each P ∈ C[u, u−1; v], there is an associated trace polynomial
N on SO(N, R). For now, it is enough to think of P (1)
P (1)
N (A) as a certain "polynomial"
function of A and tr(A). If P ∈ C[u, u−1; v] is a function of the variable u alone, then P (1)
is simply the corresponding polynomial on SO(N, R). For example, the trace polynomial
on SO(N, R) associated with P (u; v) = u5 is P (1)
N (A) = A5. On the other hand, (1.9) is
the trace polynomial on SO(N, R) associated with
N
f (u; v) =
1
N
(1 − e−t) +
1
2
e−t(1 + e
2t
N )u2 −
N
2
e−t(−1 + e
2t
N )uv1 ∈ C[u, u−1; v].
In addition, for P ∈ C[u, u−1; v], there are analogous trace polynomials P (2)
on SU(N ), and Sp(N ), resp. (see Definition 3.5 for precise definitions).
To simplify the statement of our results, we introduce the following notation: for N ∈ N,
let
N and P (4)
N
K (1)
N = SO(N, R), K (2)
N = SU(N ), K (2′)
N = U(N ), K (4)
N = Sp(N ),
(1.10)
N,C = (K (β)
and set K (β)
K (β)
s
ρN,(β)
s
and µN,(β)
= ρ
N
s,τ = µ
N )C and ∆(β)
N = ∆K (β)
N
K (β)
s,τ
N,C
for the heat kernels on K (β)
for β = 1, 2, 2′, 4.
In addition, we write
N,C, respectively.
N and K (β)
N P (β)
Our first theorem is an intertwining result showing that ∆(β)
N can be computed by
applying a certain pseudodifferential operator on C[u, u−1; v] to P and then considering
the corresponding trace polynomial.
Theorem 1.4 (Intertwining formulas for ∆SO(N,R), ∆SU(N ), and ∆Sp(N )). For β ∈
{1, 2, 4} and N ∈ N, there are pseudodifferential operators D(β)
N acting on C[u, u−1; v]
(see Definition 3.7) such that, for each P ∈ C[u, u−1; v],
N P ]N .
N P (β)
(1.11)
∆(β)
N = [D(β)
Moreover, for all τ ∈ C and P ∈ C[u, u−1; v],
N = [e
N P (β)
2 ∆(β)
e
τ
τ
2 D(β)
N P ]N
For each such β, D(β)
N takes the form
D(β)
N = L0 +
1
N L(β)
1 +
1
N 2L(β)
2 ,
where L0,L(β)
1
are first order pseudodifferential operators and L(β)
2
differential operator, all independent of N .
5
(1.12)
(1.13)
is a second order
τ
2 D(β)
N is well defined on C[u, u−1; v] (see Corollary 3.11). The proof of
The exponential e
Theorem 1.4 for SO(N, R) and Sp(N ) can be found on p. 20 and p. 45, resp. We do
not provide a full proof of the SU(N ) version of the theorem, as it is very similar to the
SO(N, R) and Sp(N ) cases; the proof is discussed on p. 55.
Our next result shows that we can use e
transform on trace polynomials.
Theorem 1.5. Let P ∈ C[u, u−1; v], N ∈ N, s > 0 and τ ∈ D(s, s). For β ∈ {1, 2, 4},
the Segal-Bargmann transform on the trace polynomial P (β)
N to explicitly compute the Segal-Bargmann
N can be computed as
τ
2 D(β)
K (β)
s,τ P (β)
N
N = [e
B
τ
2 D(β)
N P ]N .
(1.14)
The proof of Theorem 1.5 for SO(N, R) and Sp(N ) can be found on p. 23 and p. 47,
resp. The proof of this theorem for the SU(N ) case is similar, and is discussed on p. 55.
K (β)
N
s,τ
on polynomials as N → ∞. While the range
Our main result concerns the limit of B
of the boosted Segal-Bargmann transform is the space of trace polynomials, as N → ∞
the image concentrates back onto the smaller subspace of Laurent polynomials.
Theorem 1.6. Let s > 0, τ ∈ D(s, s), and f ∈ C[u, u−1]. Then there exist unique
gs,τ , hs,τ ∈ C[u, u−1] such that
K (β)
kB
K (β)
N
s,τ
N
s,τ fN − [gs,τ ]Nk2
)−1fN − [hs,τ ]Nk2
L2(K (β)
N,C,µN,(β)
s,τ
L2(K (β)
N ,ρN,(β)
s
;MN (C))
;MN (C))
k(B
N 2(cid:19) ,
= O(cid:18) 1
= O(cid:18) 1
N 2(cid:19) .
(1.15)
(1.16)
for β ∈ {1, 2, 4}.
Definition 1.7. The map Gs,τ : C[u, u−1] → C[u, u−1] given by f 7→ gs,τ is called the
free Segal-Bargmann transform, and the map Hs,τ : C[u, u−1] → C[u, u−1] given by
f 7→ hs,τ is called the free inverse Segal-Bargmann transform. Explicit formulas
for Gs,τ and Hs,τ are given in (4.22) and (4.23), respectively.
The proof of Theorem 1.6 for SO(N, R) can be found on p. 35. The proof of the theorem
for the SU(N ) and Sp(N ) cases is entirely analogous to the SO(N, R) case, so we do not
provide the full details; the proofs for SU(N ) and Sp(N ) are discussed on pp. 55 and 50.
Remark 1.8. The fact that the Segal-Bargmann transform on matrices in MN (C) has a
meaningful limit G t as N → ∞ is due to Biane [4, Proposition 13]. For the special case
when s = t > 0, the free Segal-Bargmann transform Gt,t of this paper is equivalent to
G t, which Biane called the free Hall transform. A two-parameter generalization of this
transform, called the free unitary Segal-Bargmann transform, Gs,t, where s > t/2 > 0, was
introduced in [8]. This version of the free Segal-Bargmann transform was also investigated
by Ho in [11]; in particular, he gave an integral kernel for the large-N limit of the Segal-
Bargmann transform over U(N ). In this paper, we study a slight generalization of Gs,t
by allowing the time parameter τ to be complex, incorporating results from [9] on the
complex-time Segal-Bargmann transform. When τ = t > 0, the free Segal-Bargmann
transform Gs,τ of Definition 1.7 is precisely the same as the operator Gs,t of [8].
6
Remark 1.9. Many of the results in this paper on the Segal-Bargmann transform for
SO(N, R), SU(N ), and Sp(N ) correspond to work done by Driver, Hall, and Kemp on the
Segal-Bargmann transform on U(N ). The U(N ) analogue of our intertwining formulas,
Theorem 1.4, is contained in [8, Theorem 1.18], where it is shown that ∆U(N ) has the
intertwining formula
1
1
1
2
in the SU(N ) case).
1
2
1
D(2′)
N = L0 +
on the trace polynomial space
is identically zero (as it is for
N 2L(2′)
N L(2′)
1 +
, and L(2′)
for certain pseudodifferential operators L0,L(2′)
C[u, u−1; v] (cf. Remark 3.13), where the operator L(2′)
L(2)
The same operator L0 appears in the intertwining formulas for SO(N, R), U(N ), SU(N ),
and Sp(N ). As N → ∞, it is the operator L0 that drives the large-N behavior of the
transform; this is the key to our main result on the free Segal-Bargmann transform,
Theorem 1.6, which states that the large-N limit of the Segal-Bargmann transform on
SO(N, R), SU(N ), and Sp(N ) is the same operator Gs,τ . Moreover, we emphasize that
the rate of convergence in all cases is O(1/N 2), as it is in the analogous result for the
U(N ) case, [8, Theorem 1.11]. This is somewhat surprising, since there is a nontrivial
term of order 1/N in the intertwining formulas for SO(N, R) and Sp(N ) that does not
appear in the U(N ) and SU(N ) cases. That we still obtain O(1/N 2) convergence for
SO(N, R) and Sp(N ) is due to the fact that in the scalar version of their intertwining
formulas (cf. Theorems 3.20 and 5.5), the term of order 1/N is a first order differential
operator. As a result, the 1/N terms are "washed away" in the covariance estimates
(cf. the proof of Theorem 4.10 in the SO(N, R) case for a more precise explanation; the
Sp(N ) case is similar).
2. Background
In this section, we expand on the background required to prove our main results.
In
particular, we provide a brief overview of the heat kernel results used in constructing the
Segal-Bargmann transform, and conclude by setting some notation that will be used for
the remainder of the paper.
2.1. Heat kernels on Lie groups
Definition 2.1. Let K be a Lie group with Lie algebra k. An inner product h·,·ik on k
is Ad-invariant if, for all X1, X2 ∈ k and all k ∈ K,
hAdk X1, Adk X2ik = hX1, X2ik .
A group whose Lie algebra possesses an Ad-invariant inner product is called compact
type.
The Lie groups SO(N, R), SU(N ), and Sp(N ) studied in this paper are compact type
(and are, in fact, compact).
Compact type Lie groups have the following property.
7
Proposition 2.2 ([[14], Lemma 7.5]). If K is a compact type Lie group with a fixed
Ad-invariant inner product, the K is isometrically isomorphic to a direct product group,
i.e. K ∼= K0 × Rd for some compact Lie group K0 and some nonnegative integer d.
Let K be a compact type Lie group with Lie algebra k, and fix an Ad-invariant inner
product on k. Choose a right Haar measure λ on K; we will write dx for λ(dx) and
L2(K) for L2(K, λ). If βk is an orthonormal basis for k, we define the Laplace operator
∆K on K by
where for any X ∈ k, eX is the left-invariant vector field given by
∆K = XX∈βk eX 2,
f (ketX )(cid:12)(cid:12)(cid:12)(cid:12)t=0
(eXf )(k) =
d
dt
for any smooth real or complex-valued function f on K. The operator is independent of
orthonormal basis chosen. For a detailed overview of left-invariant Laplacian operators,
see [14]; see also [6, 7, 9] for further background on heat kernels on compact type Lie
groups.
The operator ∆K is left-invariant for any Lie group K. When K is compact type, ∆K is
also bi-invariant. In addition, it is well known that ∆K on K is elliptic and essentially
self-adjoint on L2(K) with respect to any right invariant Haar measure (see [9, Section
3.2]). As a result, there exists an associated heat kernel ρK
t ∈ C∞(K, (0,∞)) satisfying
(1.3).
Definition 2.3. Let s > 0 and τ = t + iθ ∈ D(s, s). We define a second order left-
invariant differential operator AKC
s,τ on KC by
(2.1)
(2.2)
(2.3)
AKC
s,τ =
dXj=1(cid:20)(cid:18)s −
j +
t
2(cid:19) eX 2
t
j − θeXjeYj(cid:21) ,
2eY 2
j=1 is any orthonormal basis of k, and Yj = JXj where J is the operation of
where {Xj}d
multiplication of i on kC = Lie(KC).
The operator AKC
c (KC) is essentially self-adjoint on L2(KC) with respect to any right
Haar measure (see [9, Section 3.2]). There exists a corresponding heat kernel density
µKC
s,τ ∈ C∞(KC, (0,∞)) such that
s,τC∞
(cid:16)e
1
2 A
KC
s,τ f(cid:17) (z) =ZKC
µKC
s,τ (w)f (zw) dz
for all f ∈ L2(KC).
(2.4)
2.2. Notation and definitions
The classical Segal-Bargmann transform is an isometric isomorphism from L2(Rd) onto
HL2(Cd). In order to extend the Segal-Bargmann transform to Lie groups, we note that
every connected Lie group has a complexification defined by a certain representation-
theoretic universal property, mimicking the relationship between Rd and Cd (see [9,
8
Section 2]). For the present purposes, it is enough to know the concrete complexifications
of SO(N, R), SU(N ), and Sp(N ), which we describe below.
Let SO(N, R) denote the special orthogonal group, defined by SO(N, R) = {A ∈
MN (R) : AA⊺ = IN , det(A) = 1}. Its complexification, SO(N, R)C, is given by SO(N, C) =
{A ∈ MN (C) : AA⊺ = IN , det(A) = 1}. The Lie group SO(N, R) has Lie algebra
so(N, R) = {X ∈ MN (R) : X ⊺ = −X}, while SO(N, C) has Lie algebra so(N, C) =
{X ∈ MN (C) : X ⊺ = −X}.
The unitary group, U(N ), is defined by U(N ) = {A ∈ MN (C) : AA∗ = IN}. The Lie
algebra of U(N ) is u(N ) = {X ∈ MN (C) : X ∗ = −X}. The special unitary group
SU(N ) is the subgroup of U(N ) consisting of matrices with determinant 1. The complex-
ification of SU(N ) is the special linear group SL(N, C) = {A ∈ GLN (C) : det(A) = 1}.
The Lie algebra of SU(N ) is su(N ) = {X ∈ MN (C) : X ∗ = −X, Tr(X) = 0}, and the
Lie algebra of SL(N, C) is sl(N, C) = {X ∈ MN (C) : Tr(X) = 0}.
There are two standard realizations of the compact symplectic group Sp(N ), first as
a group of N × N matrices over the quaternions, and second as a group of 2N × 2N
matrices over C. It is more convenient for us to work with the latter realization, which we
introduce here; we address the former realization in Section 5.6. Our choice of realization
means that we will often have to add a normalization factor of 1
2 when working with
Sp(N ) ⊆ M2N (C).
Set
Ω0 =(cid:20)
1
0
−1 0 (cid:21) ∈ M2(C).
}
{z
N times
) ∈ M2N (C). We define the com-
For N ∈ N, we define Ω = ΩN := diag(Ω0, Ω0, . . . , Ω0
plex symplectic group by Sp(N, C) = {A ∈ M2N (C) : A⊺ΩA = Ω}. The Lie algebra
of Sp(N, C) is sp(N, C) = {X ∈ M2N (C) : ΩX ⊺Ω = X}. The compact symplectic
group Sp(N ) consists of the elements of Sp(N, C) which are also unitary, i.e.
Sp(N ) = Sp(N, C) ∩ U(2N ) = {A ∈ M2N (C) : A⊺ΩA = Ω, A∗A = I2N}.
(2.5)
The Lie algebra of Sp(N ) is sp(N ) = {X ∈ M2N (C) : ΩX ⊺Ω = X, X ∗ = −X}. The
complexification of Sp(N ) is Sp(N, C).
The groups SO(N, R), SU(N ), and Sp(N ) can be thought of as (special) unitary groups
over R, C, and H, resp. For notational convenience, throughout the paper we associate
the parameters β = 1, 2, 2′, 4 with SO(N, R), SU(N ), U(N ), and Sp(N ), resp., where β
refers to the dimension of R, C, and H as associative algebras over R.
In order for the large-N limit of the Segal-Bargmann transform to converge in a meaning-
ful way, we require the following normalizations of the trace and Hilbert-Schmidt inner
products.
Notation 2.4. For A ∈ MN (C), define
Tr(A),
Tr(A),
(2.6)
(2.7)
1
N
1
2N
9
tr(A) =
etr(A) =
where Tr denotes the usual trace. We use tr when applying the Segal-Bargmann trans-
Sp(N ).
form on SO(N, R) and SU(N ), and etr when applying the Segal-Bargmann transform on
We also define inner products on so(N, R), su(N ), and sp(N ) by
N
2
N 2
2
Tr(XY ∗) =
tr(XY ∗), X, Y ∈ so(N, R),
hX, Y iso(N,R) =
hX, Y isu(N ) = N Tr(XY ∗) = N 2 tr(XY ∗), X, Y ∈ su(N ),
hX, Y isp(N ) = N Tr(XY ∗) = 2N 2etr(XY ∗), X, Y ∈ sp(N ).
(2.8)
(2.9)
(2.10)
3. The Segal-Bargmann transform on SO(N, R)
In this section and the next, we prove Theorem 1.4, Theorem 1.5, and Theorem 1.6 for
the Segal-Bargmann transform on SO(N, R). We begin with a set of "magic formulas"
which are the key ingredient to proving the SO(N, R) version of the intertwining formula
for ∆SO(N,R) of Theorem 1.4. These formulas, which give simple expressions for certain
quadratic matrix sums, are the analogues of the "magic formulas" of [8, Proposition 3.1]
for U(N ). They allow us to compute the Segal-Bargmann transform for trace polynomials
on SO(N, R) (1.5 for SO(N, R)). We then prove a second intertwining formula for the
operator ASO(N,C)
s,τ
which is used in proving the limit theorems of Section 4.
3.1. Magic formulas and derivative formulas
Lemma 3.1 (Elementary matrix identities). Let Ei,j ∈ MN (C) denote the N ×N matrix
with a 1 in the (i, j)th entry and zeros elsewhere. For 1 ≤ i, j, k, ℓ ≤ N and A = (Ai,j ) ∈
MN (C), we have
Ei,jEk,ℓ = δj,kEi,ℓ
Ei,jAEi,j = Aj,iEi,j
Ei,j AEj,i = Aj,j Ei,i
Tr(AEi,j ) = Aj,i
(3.1)
(3.2)
(3.3)
(3.4)
Proof. For 1 ≤ a, b,≤ N ,
[Ei,jEk,ℓ]a,b =
[Ei,j]a,h[Ek,ℓ]h,b.
NXh=1
Note that [Ei,j]a,h[Ek,ℓ]h,b equals 1 precisely when a = i, b = ℓ, and j = k, and h = j = k,
and equals 0 otherwise, so [Ei,j Ek,ℓ]a,b = δj,kδa,iδb,ℓ, which proves (3.1).
We now make the following observations: for any N × N matrix A, AEi,j is the matrix
which is all zeros except for the jth column, which is equal to the ith column of A. Hence
the only (possibly) nonzero diagonal entry is the (j, j)th entry, which is Aj,i. This proves
(3.4). Similarly, Ei,jA is the matrix which is all zeros except for the i row, which is the
jth row of A. Putting these observations together yields (3.2) and (3.3).
10
Proposition 3.2 (Magic formulas for SO(N, R)). For any A, B ∈ MN (C),
X 2 = −
1
N
XAX =
XX∈βso(N,R)
XX∈βso(N,R)
XX∈βso(N,R)
tr(XA)X =
tr(XA) tr(XB) =
N − 1
N
IN = −IN +
1
N
IN
A⊺ − tr(A)IN
1
N 2 (A⊺ − A)
1
N 2 (tr(A⊺B) − tr(AB))
XX∈βso(N,R)
(3.5)
(3.6)
(3.7)
(3.8)
XAX is independent
of the choice of orthonormal basis. Hence to see (3.6), we may set βso(N,R) to be the
orthonormal basis
Proof. As shown in Theorem 3.3 of [8], the quantity PX∈βso(N,R)
(Ei,j − Ej,i)(cid:27)1≤i<j≤N
.
By Lemma 3.1,
Xi<j
βso(N,R) =(cid:26) 1
√N
(Ei,j − Ej,i)A(Ei,j − Ej,i) =Xi<j
=Xi6=j
=Xi6=j
=Xi6=j
[Ei,jAEi,j − Ei,jAEj,i − Ej,iAEi,j + Ej,iAEj,i]
(Ei,j AEi,j − Ei,jAEj,i)
Aj,iEi,j −Xi6=j
NXi=1
Aj,iEi,j +
Aj,jEi,i
Ai,iEi,i −
NXi=1
Ai,iEi,i −Xi6=j
Aj,jEi,i
= A⊺ − Tr(A)IN ,
and dividing by N proves (3.6). Equation (3.5) then follows from (3.6) by setting A = IN .
11
Similarly, we compute
Xi<j
Tr(A(Ei,j − Ej,i))(Ei,j − Ej,i) =Xi<j
=Xi6=j
=Xi6=j
=Xi6=j
[Tr(AEi,j )Ei,j − Tr(AEi,j )Ej,i
− Tr(AEj,i)Ei,j + Tr(AEj,i)Ej,i]
[Tr(AEi,j )Ei,j − Tr(AEj,i)Ei,j ]
Ai,jEi,j
Aj,iEi,j +
Aj,iEi,j −Xi6=j
NXi=1
Ai,iEi,i −Xi6=j
NXi=1
Ai,iEi,i
−
Ai,jEi,j
= A⊺ − A,
which is equivalent to (3.7). Equation (3.8) now follows from (3.7) by multiplying by B
and taking tr.
Remark 3.3. Our proof of the magic formulas for SO(N, R) is primarily computational,
relying on the elementary matrix identities of Lemma 3.1. These results can also be
(β)
obtained more abstractly. For β ∈ {1, 2, 2′, 4}, let k
N denote so(N, R), su(N ), u(N ),
(β)
N . In [13, Lemma 1.2.1],
and sp(N ), resp., and let βk
, defined as the
L´evy provides an explicit decomposition of the Casimir element Ck
tensor
be an orthonormal basis for k
(β)
N
(β)
N
Ck
(β)
N
= XX∈β
k
(β)
N
X ⊗ X.
Let
Ea,b ⊗ Eb,a
and
P =
Ea,b ⊗ Ea,b.
For K ∈ {R, C, H}, set I(K) = {1, i, j, k} ∩ K and define two elements ReK and CoK of
K ⊗R K by
γ ⊗ γ−1
and
γ ⊗ γ.
Then for β ∈ {1, 2′4} (where 2′ corresponds to the value 2), the Casimir element Ck
given by
is
(β)
N
Ck
(β)
N
=
1
βN (cid:16)−T ⊗ ReK +P ⊗ CoK(cid:17) ,
and for β = 2, Csu(N ) = Cu(N ) − 1
compute any expression of the formPX∈β
k
(β)
N
N 2 iIN ⊗ iIN . This decomposition can be used to
B(X, X), where B is an R-bilinear map.
In particular, the magic formulas for SO(N, R), SU(N ), and Sp(N ) (cf. Propositions 3.2,
6.2, and 5.10) can be obtained in this way.
12
T =
NXa,b=1
ReK = Xγ∈I(K)
NXa,b=1
CoK = Xγ∈I(K)
Proposition 3.4 (Derivative formulas for SO(N, R)). Let X ∈ βso(N,R). The following
identities hold on SO(N, R) and SO(N, C):
AjXAm−j, m < 0
mXj=1
eXAm =
AjXAm−j, m ≥ 0
0Xj=m+1
eXAm = −
eX tr(Am) = m · tr(XAm), m ∈ Z
(m − j) tr(Aj )Am−j
∆SO(N,R)Am = 21m≥2 1
m−1Xj=1
m−1Xj=1
(m − j)Am−2j −
Am, m ≥ 0
(−m + j) tr(Aj)Am−j
∆SO(N,R)Am = 21m≤−2 1
−1Xj=m+1
−1Xj=m+1
XX∈βso(N,R) eX tr(Am) · eXAp =
N
mp
N 2 (Ap−m − Ap+m), m, p ∈ Z.
(−m + j)Am−2j −
m(N − 1)
+
m(N − 1)
N
−
Am, m < 0
N
N
(3.9)
(3.10)
(3.11)
(3.12)
(3.13)
(3.14)
We emphasize that these derivative formulas hold true only in the case on SO(N, R) and
SO(N, C), in which case A−1 = A⊺.
Proof. The proof of (3.9), (3.10), and (3.11) for the U(N ) case are contained in [8]; the
proof of these equations for the SO(N, R) case is identical. Next, by (3.9), we have, for
m ≥ 0,
eX 2Am = 21m≥2 Xj,k6=0
j+k+ℓ=m
AjXAkXAℓ +
mXj=1
AjX 2Am−j.
(3.15)
13
Summing over all X ∈ βSO(N,R) and using the magic formulas (3.5) and (3.6), we have
∆SO(N,R)Am = 21m≥2 XX∈βso(N,R) Xk,j6=0
k+j+ℓ=m
AkXAjXAℓ + XX∈βso(N,R)
Ak(Aj)⊺Aℓ − tr(Aj)Ak+ℓ(cid:21) −
AjX 2Am−j
mXj=1
m(N − 1)
Am
N
(cid:20) 1
N
k+j+ℓ=m
= 21m≥2 Xk,j6=0
= 21m≥2 1
m−1Xj=1
(m − j)Am−2j −
N
m(N − 1)
Am,
N
(m − j) tr(Aj )Am−j
m−1Xj=1
−
which proves (3.12).
Similarly, if m < 0, we use (3.10) to get
eX 2Am = 21m≤−2 Xk,ℓ<0,j≤0
j+k+ℓ=m
AjXAkXAℓ +
0Xj=m+1
AjX 2Am−j.
Summing over all X ∈ βSO(N,R) and using the magic formulas (3.5) and (3.6), we have
∆SO(N,R)Am = 21m≥2 XX∈βso(N,R) Xk,ℓ<0,j≤0
j+k+ℓ=m
AkXAjXAℓ + XX∈βso(N,R)
Ak(Aj )⊺Aℓ − tr(Aj)Ak+ℓ(cid:21) +
0Xj=m+1
AjX 2Am−j
m(N − 1)
Am
N
(−m − j)Am−2j −
(−m − j) tr(Aj )Am−j
−1Xj=m+1
N
j+k+ℓ=m
(cid:20) 1
= 21m≥2 Xk,ℓ<0,j≤0
= 21m≥2 1
−1Xj=m+1
N
m(N − 1)
+
Am,
N
which proves (3.13).
For (3.14), we first assume m, p ≥ 0. Using equation (3.7) and the fact that A⊺ = A−1
14
for A ∈ SO(N, R),
XX∈βso(N,R) eX tr(Am) · eXAp = XX∈βso(N,R)
AkXAp−k!
AjXAm−j pXk=1
tr
mXj=1
tr(XAm) XAp−k
mXj=1
Ak XX∈βso(N,R)
mAk XX∈βso(N,R)
tr(XAm)X Ap−k
mp
N 2 (Ap−m − Ap+m).
pXk=1
pXk=1
=
=
=
For m < 0 and p ≥ 0, we have
tr−
Aj XAm−j pXk=1
0Xj=m+1
XX∈βso(N,R) eX tr(Am) · eXAp = XX∈βso(N,R)
tr(XAm) XAp−k
pXk=1
0Xj=m+1
Ak XX∈βso(N,R)
(−m)Ak XX∈βso(N,R)
tr(XAm)X Ap−k
pXk=1
= −
mp
N 2 (Ap−m − Ap+m).
= −
=
AkXAp−k!
For m ≥ 0 and p < 0, we have
XX∈βso(N,R) eX tr(Am) · eXAp = XX∈βso(N,R)
0Xk=p+1
0Xk=p+1
= −
mp
N 2 (Ap−m − Ap+m).
AjXAm−j−
tr
0Xk=p+1
mXj=1
tr(XAm) XAp−k
mXj=1
Ak XX∈βso(N,R)
mAk XX∈βso(N,R)
tr(XAm)X Ap−k
= −
=
AkXAp−k
15
Finally, for m, p < 0, we have
XX∈βso(N,R) eX tr(Am) · eXAp
= XX∈βso(N,R)
0Xk=p+1
0Xk=p+1
=
=
Aj XAm−j−
− tr
0Xk=p+1
0Xj=m+1
tr(XAm) XAp−k
0Xj=m+1
Ak XX∈βso(N,R)
(−m)Ak XX∈βso(N,R)
tr(XAm)X Ap−k
(−m)(−p)
N 2
(Ap−m − Ap+m).
=
This proves (3.14).
AkXAp−k
3.2. Intertwining formulas for ∆SO(N,R)
In this section, we describe the trace polynomial functional calculus that will be necessary
for the proof of Theorem 1.4, which contains the intertwining formulas for ∆SO(N,R),
∆SU(N ), and ∆Sp(N ). The space of trace polynomials was first introduced in [8, Definition
1.7], where it was used to prove analogous intertwining results for ∆U(N ). Using this
framework, we then prove Theorem 1.4 for the SO(N, R) case.
The computations in this section are related to, but not quite the same as, those used to
establish the intertwining formulas for ∆U(N ) in [8, Sections 3-4].
Definition 3.5. Let C[u, u−1] denote the algebra of Laurent polynomials in a single
variable u:
C[u, u−1] =(Xk∈Z
akuk : ak ∈ C, ak = 0 for all but finitely many k) ,
(3.16)
with the usual polynomial multiplication. Let C[u] and C[u−1] denote the subalgebras
consisting of polynomials in u and u−1, respectively.
Let C[v] denote the algebra of polynomials in countably many commuting variables v =
{v±1, v±2, . . .}, and let C[u, u−1; v] denote the algebra of polynomials in the commuting
variables {u, u−1, v±1, v±2, . . .}.
We now define the trace polynomial functional calculus: for P ∈ C[u, u−1; v], we
have
P (1)
N (A) = [P (1)]N (A) := P (u; v)u=A,vk=tr(Ak),k6=0,
P (2)
N (A) = [P (2)]N (A) := P (u; v)u=A,vk=tr(Ak),k6=0,
P (4)
N (A) = [P (4)]N (A) := P (u; v)u=A,vk= etr(Ak),k6=0,
16
A ∈ SO(N, R) or SO(N, C),
A ∈ SU(N ) or SL(N, C),
A ∈ Sp(N ) or Sp(N, C).
N , where P ∈ C[u, u−1; v], are called trace polynomials. We
N and write PN when it is clear from context
Functions of the form P (β)
will often suppress the superscripts for P (β)
which version of the trace polynomial functional calculus is being used.
As an illustrative example, the trace polynomial P (1)
P (u; v) = v1v6
N = PN on SO(N, R) associated with
−3u2 − 3v2
5 is
PN (A) = tr(A) tr(A−3)6A2 − 3 tr(A5)2IN .
We now introduce the pseudodifferential operators on C[u, u−1; v] that appear in our
intertwining formulas.
Definition 3.6. Let R± denote the positive and negative projection operators
and R− : C[u, u−1; v] → C[u−1; v]
R+ : C[u, u−1; v] → C[u; v]
defined by
R+ ∞Xk=−∞
ukqk(v)! =
ukqk(v), R− ∞Xk=−∞
∞Xk=0
ukqk(v)! =
−1Xk=0−∞
ukqk(v).
In addition, for any k ∈ Z, let Muk denote the multiplication operator defined by
Muk P (u; v) = ukP (u; v).
17
Definition 3.7. Define the following operators on C[u, u−1; v]:
1 =
1 =
2 =
Z1 = Z +
2 − Y −
1 − Y −
Y2 = Y +
N = N0 + N1 = Xk≥1
∞Xk=1
Y1 = Y +
∞Xk=2
∞Xk=2
∞Xk=1
1 = Xk≥1
2 = Xj,k≥1
2 − Z −
1 − Z −
1 − K−
Z2 = Z +
K1 = K+
K2 = K+
2 − K−
J = 2 Xk≥1
∂u∂vk
2 =
ku
∂
kvk
∂
∂vk
∂
∂u
+ u
∂uMu−kR+ −
(R+ − R−),
−1Xk=−∞
jvjvk−j ∂
−2Xk=−∞
(k − j)vk−2j ∂
∂vk −
∂vk −
vkuR+ ∂
k−1Xj=1
k−1Xj=1
u−k+1R+ ∂
ku−k+1 ∂2
∂uMu−kR+ −
∂vk∂u − Xk≥1
∂vj∂vk − Xj,k≥1
+ Xk≥1
∂vj∂vk
jkvjvk
∂2
∂2
jkvk−j
+ Xj,k≥1
,
∂vk
∂uMu−kR−,
vkuR− ∂
jvjvk−j ∂
−1Xj=k+1
(k − j)vk−2j ∂
−2Xk=−∞
−1Xj=k+1
−1Xk=−∞
u−k+1R− ∂
∂uMu−kR−,
,
∂vk
kuk+1 ∂2
∂vk∂u
,
jkvk+j
∂2
∂vj∂vk
,
(3.17)
(3.18)
(3.19)
(3.20)
(3.21)
(3.22)
(3.23)
k2vk
∂
∂vk
+ u
∂2
∂u2 + u
∂
∂u
.
(3.24)
We set
L0 = −N − 2Y1 − 2Y2,
L(1)
1 = N + 2Z1 + 2Z2,
L(2)
1 = 0,
L(4)
1 = −
L(2)
2 = −2K−
2L(1)
L(4)
1 ,
2 =
1
L(1)
2 = 2K1 + K2,
1 − K−
2 + J ,
1
4L(1)
2 .
For β ∈ {1, 2, 4}, we define
D(β)
N = L0 +
1
N L(β)
1 +
1
N 2L(β)
2 .
(3.25)
(3.26)
(3.27)
(3.28)
(3.29)
Notation 3.8. For m ∈ Z and A ∈ MN (C), let Wm(A) = Am, Vm(A) = tr(Am), and
V(A) = {Vm(A)}m≥1. The functions Wm, Vm, and V implicitly depend on N , but we
suppress the index, which should not cause confusion.
Theorem 3.9 (Partial product rule). Let α, β ∈ C. For P ∈ C[u, u−1; v] and Q ∈ C[v],
the operator αL0 + βL(1)
satisfies
1
(cid:16)αL0 + βL(1)
1 (cid:17) (P Q) =(cid:16)(cid:16)αL0 + βL(1)
1 (cid:17) P(cid:17) Q + P(cid:16)(cid:16)αL0 + βL(1)
1 (cid:17) Q(cid:17) .
18
(3.30)
Proof. From Definition 3.7, recall that
L0 = −(N0 + N1) − 2Y1 − 2Y2,
L(1)
1 = N0 + N1 + 2Z1 + 2Z2.
Observe that N0, Y2, Z1 are first order differential operators on C[v], and so satisfy the
product rule on C[u, u−1; v]. On the other hand, N1, Y1, and Z2 annihilate C[v] and
satisfy, for P ∈ C[u, u−1; v] and Q ∈ C[v],
N1(P Q) = (N1P )Q,
Y1(P Q) = (Y1P )Q,
Z2(P Q) = (Z2P )Q.
Putting this together shows (3.30).
Definition 3.10. The trace degree of a monomial in C[u, u−1; v] is defined to be
deg(cid:16)uk0vk1
1 vk−1
−1 ··· vkm
m vk−m
−m (cid:17) = k0 + X1≤j≤m
jkj.
(3.31)
We define the trace degree of any element in C[u, u−1; v] to be the highest trace degree
of any of its monomial terms. For m ≥ 0, we define the subspace Cm[u, u−1; v] ⊆
C[u, u−1; v] to be
Cm[u, u−1; v] = {P ∈ C[u, u−1; v] : deg P ≤ m},
(3.32)
Cm[u, u−1; v].
consisting of polynomials of trace degree ≤ m. Note that Cm[u, u−1; v] is finite-dimensional
and C[u, u−1; v] =Sm≥0
Corollary 3.11. Let m, N ∈ N and α, β ∈ C. The operators αL0+βL(1)
N , and
D(4)
N preserve the finite dimensional subspace Cm[u, u−1; v] ⊆ C[u, u−1; v]. It follows that
eαL0+βL(1)
N are well-defined operators on each Cm[u, u−1; v]
(via power series, cf. Remark 3.14) and hence are well-defined on all of C[u, u−1; v].
N , and eαD(4)
N , D(2)
1 , D(1)
N , eαD(2)
1 , eαD(1)
1 , D(1)
N , D(2)
N , and D(4)
Proof. Recall that αL0 + βL(1)
N are linear combinations of the
operators J , N , Y1, Y2, Z1, Z2, K1, and K2 described in Definition 3.7. Each of
these operators, in turn, are linear combinations of compositions of multiplication, dif-
ferentiation, and projection operators. The multiplication and differentiation operators
raise and lower trace degree, respectively, while projection operators R+ and R− leaves
Cm[u, u−1; v] invariant. Keeping track of multiplication and differentiation operators
in J , N , Y1, Y2, Z1, Z2, K1, and K2 shows that these operators preserve trace de-
gree. Hence αL0 + βL(1)
N preserve the finite dimensional subspace
Cm[u, u−1; v] ⊆ C[u, u−1; v], and the result for eαL0+βL(1)
N , and eαD(4)
N , and D(4)
N , D(2)
1 , D(1)
N , eαD(2)
1 , eαD(1)
N
follows.
Corollary 3.12. For any P ∈ C[u, u−1; v], Q ∈ C[v], and α, β, τ ∈ C,
1 )Q.
1 )(P Q) = e
2 (αL0+βL(1)
2 (αL0+βL(1)
2 (αL0+βL(1)
e
τ
τ
τ
1 )P · e
(3.33)
19
Proof. Suppose L is a linear operator on C[u, u−1; v] which leaves C[v] and Cm[u, u−1; v]
invariant and satisfies the partial product rule (3.30). Applying (3.30) repeatedly, we get
1
k!Lk(P Q) =
1
k!
∞Xk=0
1
=
1
1
1
j!k!
ℓ!(k − ℓ)!
(LℓP )(Lk−ℓQ)
∞Xk=0
∞Xj,k=0
= eLP · eLQ.
2(cid:16)αL0 + βL(1)
kXℓ=0(cid:18)k
ℓ(cid:19)(LℓP )(Lk−ℓQ) =
∞Xk=0
kXℓ=0
j!LjP ∞Xk=0
(Lj P )(LkQ) =
∞Xj=0
1 (cid:17) leaves C[v] and Cm[u, u−1; v] invari-
2(cid:16)αL0 + βL(1)
1 (cid:17) concludes the proof.
k!LkQ!
By Definition 3.7, it is clear that τ
ant, so setting L = τ
We can now prove our main intertwining result for ∆SO(N,R).
Proof of Theorem 1.4 for SO(N, R). It suffices to show that ∆SO(N,R)PN = [DN P ]N
holds for P (u; v) = umq(v), where m ∈ Z \ {0} and q ∈ C[v]. Then PN = Wm · q(V),
and by the product rule,
∆SO(N,R)PN = XX∈βso(N,R) eX 2(Wm · q(V))
= XX∈βso(N,R) eX(cid:16)(eXWm) · q(V) + Wm · eXq(V)(cid:17)
= XX∈βso(N,R)
(eX 2Wm) · q(V) + 2eXWm · eXq(V) + Wm · eX 2q(V)
+ 2 XX∈βso(N,R) eXWm · eXq(V) + Wm · (∆SO(N,R)q(V))
= (∆SO(N,R)Wm) · q(V)
(3.34)
For the first term in (3.34), we consider the cases m ≥ 0 and m < 0 separately. For
20
m ≥ 0, we apply (3.12) to get
(∆SO(N,R)Wm) · q(V)
(m − k)Wm−2k −
2
(m − k)VkWm−k# q(V)
m−1Xk=1
N " ∞Xk=1
∂uMu−kR+P!#N
u−k+1R+ ∂
∂uMu−kR+! P#N
+
vkuR+ ∂
m(N − 1)
N
= −
Wm · q(V)
∂
N
N − 1
+ 21m≥2" 1
m−1Xk=1
N (cid:20)u
∂uR+P(cid:21)N
− 2" ∞Xk=1
N (cid:20)u
∂uR+P(cid:21)N
N (cid:20)u
N − 1
N − 1
∂
= −
= −
2
N
+
2 P ]N − 2[Y +
[Z +
∂uR−P(cid:21)N −
2
N
2 annihilates C[u−1] while Y −
1 ,Z −
∂
1 P ]N .
(3.35)
1 P ]N .
2 P ]N + 2[Y −
[Z −
2 annihilates C[u], so for all
A similar computation, using (3.13), shows that for m < 0,
(∆SO(N,R)Wm) · q(V) =
We observe that Y +
m ∈ Z,
1 ,Z +
XX∈βso(N,R) eXWm · eXq(V)
(∆SO(N,R)Wm) · q(V) = −
N − 1
N
[N1P ]N +
2
N
[Z2P ]N − 2[Y1P ]N .
(3.36)
Next, by the chain rule and (3.14), the middle term in (3.34) is
mk
∂vk
∂vk
= XX∈βso(N,R) eXWm · Xk≥1(cid:18) ∂
q(cid:19) (V) · eXVk
XX∈βso(N,R) eXWm · eXVk(cid:18) ∂
q(cid:19) (V)
= Xk≥1
N 2 (Wm−k − Wm+k)(cid:18) ∂
q(cid:19) (V)
= Xk≥1
k(cid:18)W−k+1(cid:20) ∂
um(cid:21)N − Wk+1(cid:20) ∂
N 2 Xk≥1
N 2Xk≥1
∂u∂vk PN
k(u−k+1 − uk+1)
∂vk
∂u
∂u
∂2
=
=
1
1
=
1
N 2 [K1P ]N .
21
um(cid:21)N(cid:19)(cid:18) ∂
∂vk
q(cid:19) (V) (3.37)
(3.38)
∂u um(cid:3)N in line (3.37).
Finally, for the last term in (3.34), we have, for each X ∈ βso(N,R),
where we have used the fact that mWm+j = Wj+1(cid:2) ∂
eX 2q(V) = eXXk≥1(cid:18) ∂
= Xk≥1(cid:18) ∂
q(cid:19) (V) · eXVk
q(cid:19) (V) · eX 2Vk + Xj,k≥1(cid:18) ∂2
∂vk
∂vk
∂vj∂vk
q(cid:19) (V) · eXVj · eXVk.
(3.39)
In summing the first term in (3.39) over X ∈ βso(N,R), we break up the sum for positive
and negative k. Using (3.12) and (3.13), we have
(k − ℓ)Vk−2ℓ(cid:18) ∂
∂vk
q(cid:19) (V)
2
N
k−1Xℓ=1
∞Xk=2
q(cid:19) (V)
∂vk
XX∈βso(N,R)
∞Xk=1(cid:18) ∂
∂vk
N − 1
N
= −
and
XX∈βso(N,R)
−1Xk=−∞(cid:18) ∂
∂vk
=
N − 1
N
− 2
− 2
∂vk
q(cid:19) (V) · eX 2Vk
kVk(cid:18) ∂
q(cid:19) (V) +
∞Xk=1
(k − ℓ)VℓVk−ℓ(cid:18) ∂
k−1Xℓ=1
∞Xk=2
q(cid:19) (V) · eX 2Vk
kVk(cid:18) ∂
−1Xk=−∞
−1Xℓ=k+1
−2Xk=−∞
2
N
∂vk
(3.40)
q(cid:19) (V)
(3.41)
(3.42)
(3.43)
Summing the second term in (3.39) over X ∈ βso(N,R) and using (3.14) yields
(k − ℓ)Vk−2ℓ(cid:18) ∂
∂vk
∂vk
q(cid:19) (V) −
−2Xk=−∞
(−k + ℓ)VℓV−k−ℓ(cid:18) ∂
−1Xℓ=k+1
q(cid:19) (V).
q(cid:19) (V) · eXVj · eXVk
jk(Vk−j − Vk+j )(cid:18) ∂2
∂vj∂vk
∂vj∂vk
N − 1
N N0 +
2
N Z1 − 2Y2 +
q(cid:19) (V)
N 2K2(cid:19) P(cid:21)N
1
XX∈βso(N,R) Xj,k≥1(cid:18) ∂2
N 2 Xj,k≥1
Wm · (∆SO(N,R)q(V)) =(cid:20)(cid:18)−
=
1
Adding (3.40), (3.41), and (3.42) together and multiplying by Wm, we get
Combining (3.36), (3.38), and (3.43) proves (1.11) for SO(N, R) (β = 1). Equation (1.12)
now follows from Corollary (3.11).
22
Remark 3.13. A similar intertwining formula is proven in [8] for the U(N ) case, where
for any P ∈ C[u, u−1; v],
,
(3.44)
(3.45)
∆U(N )PN =(cid:20)(cid:18)L0 −
1
N 2 (2K−
1 + K−
2 )(cid:19) P(cid:21)N
and, subsequently, for τ ∈ C,
τ
2 ∆U(N )PN = [e
e
τ
2 (L0− 1
N 2 (2K−
1 +K−
2 ))P ]N .
Remark 3.14. In this paper, we are primarily concerned with computing the Segal-
Bargmann transform on trace polynomials. In this setting, if D is a left-invariant differ-
ential operator on K, where K is a compact type matrix Lie group, then we can compute
the action of eD on a trace polynomial f via the power series
(eDf )(x) = ∞Xn=0
Dnf! (x).
1
n!
(3.46)
In particular, if D = τ
series definition. In [9], it is shown that e
KC. Moreover, this analytic continuation is equal to BK
2 ∆K, we can define e
τ
τ
2 ∆K f as a function on K using the power
2 ∆K f has a unique analytic continuation to
s,τ f .
Proof of Theorem 1.5 for SO(N, R). By the intertwining formula (1.12), we have
τ
2 ∆SO(N,R)PN = [e
e
τ
2 D(1)
N P ]N .
τ
2 D(1)
Corollary 3.11 shows that [e
N P ]N is a trace polynomial on SO(N, R), so its ana-
lytic continuation to SO(N, C) is given by the same trace polynomial function. Hence
N P ]N , viewed as a holomorphic function on SO(N, C), is the analytic continuation
[e
2 ∆SO(N,R)PN , and so is equal to BSO(N,R)
of e
2 D(1)
PN .
s,τ
τ
τ
3.3. Intertwining formulas for ASO(N,C)
s,τ
We now prove an intertwining formula for SO(N, C) that is analogous to the intertwining
formula for SO(N, R) in Theorem 1.4. To do so, we use the word polynomial space
introduced in [8, Notation 3.21], which was used to prove a similar intertwining formula
for GLN (C) = U(N )C.
Notation 3.15. For m ∈ N, define Em to be the set of words ε : {1, . . . , m} → {±1,±∗}.
We denote the length of a word ε ∈ Em by ε = m and set E = ∪m≥0Em. We define the
word polynomial space W to be
the space of polynomials in the commuting variables {vε}ε∈Em. For j, k ∈ Z not both
zero, we define the words
W = C [{vε}ε∈E ] ,
ε(j, k) =
z
}
j times
k times
±∗, . . . ,±∗) ∈ Ej+k,
(±1, . . . ,±1,
23
{
z
}
{
(3.47)
where the first j slots contain +1 if j > 0 and −1 if j < 0, and the last k slots contain
+∗ if k > 0 and −∗ if k < 0. We set vε(0,0) = 1.
Notation 3.16. For ε = (ε1, . . . , εm) ∈ Em and A ∈ SO(N, C), we define Aε =
Aε1 ··· Aεm , where A+∗ := A∗ and A−∗ := (A∗)−1.
If P ∈ W , we define a function
PN : SO(N, C) → C by
PN (A) = P (V(A)),
where
and
V(A) = {Vε(A) : ε ∈ E }
Vε(A) = tr(Aε) = tr(Aε1 ··· Aεm ).
Notation 3.17. We define the inclusion maps ι, ι∗ : C[v] ֒→ W , with ι linear and ι∗
conjugate linear, by
(3.48)
The maps ι and ι∗ intertwine with the evaluation map in the following way: for Q ∈ C[v],
(3.49)
[ι∗(Q)]N (A) = QN (A)∗.
[ι(Q)]N (A) = QN (A),
ι∗(vk) = vε(0,k).
ι(vk) = vε(k,0),
Definition 3.18. The trace degree of a monomialQm
j=1 vkj
εj ∈ W is defined to be
deg
mYj=1
vkj
εj =
mXj=1
kjεj,
and the trace degree of an arbitrary element of W is the highest trace degree of any of
its monomial terms.
For each m ∈ N, we let Wm denote the finite dimensional subspace of W
Wm = {P ∈ W : deg(P ) ≤ m} ⊆ C[{vε}ε≤m],
Wm.
so that W =S∞
m=1
ε such that
ε
Theorem 3.19. Fix s ∈ R and τ = t + iθ ∈ C. There are collections {Qs,τ
{Rs,τ
ε,δ : ε, δ ∈ Eo in W such that:
: ε ∈ E }, andnSs,τ
1. for each ε ∈ E , Qs,τ
and Rs,τ
ε
ε
ε
are certain finite sums of monomials of trace degree
: ε ∈ E },
ASO(N,C)
s,τ
Vε = [Qs,τ
ε ]N +
1
N
[Rs,τ
ε ]N ,
(3.50)
2. for ε, δ ∈ E , Ss,τ
that
ε,δ is a certain finite sum of monomials of trace degree ε + δ such
dSO(N,R)Xℓ=1 (cid:20)(cid:18)s −
t
2(cid:19) (eXℓVε)(eXℓVδ) +
t
2
(eYℓVε)(eYℓVδ) − θ(eXℓVε)(eYℓVδ)(cid:21) =
1
N 2 [Ss,τ
ε,δ ]N .
(3.51)
24
For the proof below, we use the following conventions: let βso(N,R) = {Xℓ}
denote
an orthonormal basis for so(N, R), with β+ = βso(N,R) and β− = iβso(N,R). If X ∈ MN (C)
and A ∈ SO(N, C),
(AX)1 := AX,
(AX)∗ := X ∗A∗,
(3.52)
In addition, we will be liberal in our use of ± to denote a sign that may vary for different
terms and on different sides of an equation, since we will not require a precise formula
for our word polynomials beyond what is described in the theorem statement.
dSO(N,R)
ℓ=1
(AX)−1 := −XA−1,
(AX)−∗ := −A∗X ∗.
Proof. Fix a word ε = (ε1, . . . , εm) ∈ E . Then for each X ∈ β± and A ∈ SO(N, C),
so
tr(Aε1 ··· (AX)εj ··· Aεm ),
mXj=1
(eXVε)(A) =
mXj=1
tr(Aε1 ··· (AX 2)εj ··· Aεm )
+ 2 X1≤j<k≤m
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm )
(eX 2Vε)(A) =
(3.53)
(3.54)
(3.55)
Applying magic formula (3.5) to sum each term in (3.54) over β±, we have
XX∈β±
tr(Aε1 ··· (AX 2)εj ··· Aεm ) = ±
N − 1
N
tr(Aε1 ··· Aεj ··· Aεm ) = ±
N − 1
N
Vε(A).
(3.56)
Summing over 1 ≤ j ≤ m now gives
XX∈β±
mXj=1
tr(Aε1 ··· (AX 2)εj ··· Aεm ) =
N − 1
N
n±(ε)Vε(A),
(3.57)
where n±(ε) ∈ Z and n±(ε) ≤ ε. For (3.55), we can express each term in the sum as
(3.58)
j,k XAε1
j,k XAε2
j,k ),
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm ) = ± tr(Aε0
j,kε1
j,k, and ε2
j,k are substrings of ε such that ε0
j,k, ε1
so ε0
X ∈ β± by magic formula (3.6), we have
j,kε2
j,k = ε. Summing (3.58) over
XX∈β±
tr(Aε1 ···(AX)εj ··· (AX)εk ··· Aεm )
= ±(cid:20) 1
N
1
N
= ±
tr(Aε0
j,k (Aε1
j,k )−1Aε2
j,k ) − tr(Aε0
j,k Aε2
j,k ) tr(Aε1
j,k )(cid:21)
Vε3
j,k
(A) ± Vε0
j,kε2
j,k
(A)Vε1
j,k
(A),
25
(3.59)
(3.60)
j,k is the word of length ε such that Vε3
where ε3
summing (3.55) over X ∈ β± gives
j,k
(A) = tr(Aε0
j,k (Aε1
j,k )−1Aε2
j,k ). Thus
XX∈β± X1≤j<k≤m
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm )
N X1≤j<k≤m
(A) + X1≤j<k≤m
±Vε3
=
1
j,k
±Vε0
j,kε2
j,k
(A)Vε1
j,k
(A).
We define word polynomials corresponding to (3.61) by
Q±
Q±
ε,0 = X1≤j<k≤m
ε,1 = X1≤j<k≤m
±vε3
j,k
±vε0
j,kε2
j,k
vε1
j,k
.
(3.61)
(3.62)
Now if X ∈ β+ and Y = iX ∈ β−, then
(eY Vε)(A) =
and
mXj=1
tr(Aε1 ··· (AY )εj ··· Aεm ) = i
mXj=1
± tr(Aε1 ··· (AX)εj ··· Aεm ),
(3.63)
± tr(Aε1 ··· (AX 2)εj ··· Aεm )
(eXeY Vε)(A) =i
mXj=1
+ 2i X1≤j<k≤m
± tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm ).
(3.64)
(3.65)
A similar argument to the above shows that summing (3.64) over X ∈ β+ gives
± tr(Aε1 ··· (AX 2)εj ··· Aεm ) =
N − 1
N
η(ε)Vε(A)
(3.66)
mXj=1
where η(ε) ∈ Z and η(ε) ≤ ε. In addition, summing (3.65) over X ∈ β+, we have
XX∈β+
XX∈β+ X1≤j<k≤m
We define corresponding word polynomials
± tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm )
(A) + X1≤j<k≤m
=
1
j,k
±Vε0
j,kε2
j,k
Vε3
N X1≤j<k≤m
Qε,2 = i X1≤j<k≤m
Qε,3 = i X1≤j<k≤m
26
±vε3
j,k
±vε0
j,kε2
j,k
vε1
j,k
.
(A)Vε1
j,k
(A).
(3.67)
(3.68)
(3.69)
Putting this together, we define
Qs,τ
ε =(cid:18)s −
ε =(cid:18)s −
Rs,τ
t
t
2(cid:19)(cid:0)n+(ε)vε + 2Q+
ε,1(cid:1) +
2(cid:19)(cid:0)−n+(ε)vε + 2Q+
ε,0(cid:1) +
t
2(cid:0)n−(ε)vε + 2Q−
2(cid:0)−n−(ε)vε + 2Q−
ε,1(cid:1) − θ (iη(ε) + 2Qε,3) ,
ε,0(cid:1) − θ (−iη(ε) + 2Qε,2) .
t
(3.71)
(3.70)
By construction, Qs,τ
For part (2), fix δ ∈ E with n = δ. For each X ∈ β±,
and Rs,τ
ε
ε
are of the required form and satisfy (3.50).
(eXVδ)(A)(eXVε)(A) =
mXj=1
nXk=1
tr(Aε1 ··· (AX)εj ··· Aεm ) tr(Aδ1 ··· (AX)δk ··· Aδn ).
(3.72)
Using the invariance of the trace under cyclic permutations, we can write the terms in
the above sum as
± tr(XAε(j)) tr(XAδ(k)),
where ε(j) and δ(k) are particular cyclic permutations of ε and δ. Summing over X ∈
β± and applying magic formula (3.8), along with the fact that X ∈ so(N, C) is skew-
symmetric,
XX∈β±
(eXVδ)(A)(eXVε)(A) =
=
1
N 2
1
N 2
mXj=1
mXj=1
nXk=1
nXk=1
±(cid:16)tr((Aε(j))−1Aδ(k)) − tr(Aε(j)Aδ(k)(cid:17)
±(Vε(j)⊺δ(k)(A) − Vε(j)δ(k)(A)),
(3.73)
where ε(j)⊺ is the word satisfying Vε(j)⊺ (A) = Vε(j)(A)−1.
(For example, if ε(j) =
(−∗,−1,∗, 1, 1), then ε(j)⊺ = (−1,−1,−∗, 1,∗).) We define the associated word polyno-
mials
S±
ε,δ =
mXj=1
nXk=1
±(vε(j)⊺δ(k) − vε(j)δ(k)).
(3.74)
Using magic formula (3.8) to sum over X ∈ β+ gives
Next, if X ∈ β+ and Y = iX ∈ β−, then
dSO(N,R)Xℓ=1
(eXVε)(A)(eY Vδ)(A) = i
mXj=1
nXk=1
(eXℓVε)(A)(eYℓVδ)(A) =
mXj=1
ε,δ = i
i
N 2
S′
Hence we define the associated word polynomial
mXj=1
nXk=1
± tr(XAε(j)) tr(XAδ(k)).
(3.75)
nXk=1
±(Vε(j)⊺δ(k)(A) − Vε(j)δ(k)(A)).
(3.76)
±(vε(j)⊺δ(k) − vε(j)δ(k))
27
(3.77)
(where the signs in (3.77) do not necessarily correspond to those in (3.74)). Finally, we
define
Ss,τ
ε,δ =(cid:18)s −
t
2(cid:19) S+
ε,δ +
t
2
S−
ε,δ − θS′
ε,δ,
(3.78)
where we have constructed Ss,τ
and satisfies (3.51).
ε,δ so that it is a sum of monomials of trace degree ε + δ
s,τ
). Fix s ∈ R and τ = t + iθ ∈ C,
ε,δ : ε, δ ∈ Eo be as in Theorem 3.19.
t
2(cid:19) (eXℓVε)(eXℓVδ)
(eYℓVε)(eYℓVδ) − θ(eXℓVε)(eYℓVδ)(cid:21).
t
2
The following result will be useful for the proof of Theorem 1.6. It appears in [8] for the
GL(N, C) case, with the proof virtually unchanged for SO(N, C).
28
Rs,τ
ε
∂2
ε
ε
Qs,τ
ε
Ss,τ
ε,δ
1
1
1
2 =
1 =
0 =
∂vε∂vδ
∂
∂vε
∂
∂vε
: ε ∈ E }, {Rs,τ
Then for all N ∈ N and P ∈ W ,
Theorem 3.20 (Intertwining formula for ASO(N,C)
: ε ∈ E }, and nSs,τ
and let {Qs,τ
Define first and second order differential operators
2Xε∈E
eLs,τ
2Xε∈E
eLs,τ
2 Xε,δ∈E
eLs,τ
PN =(cid:20)(cid:18)eLs,τ
N eLs,τ
∂vε(cid:19) (V) · (eXVε)(cid:21)
∂vε(cid:19) (V) · (eX 2Vε) + Xε,δ∈E(cid:18) ∂2P
∂vε(cid:19) (V) · ASO(N,C)
∂vε∂vδ(cid:19) (V)
+ Xε,δ∈E(cid:18) ∂2P
eX 2PN =Xε∈E eX(cid:20)(cid:18) ∂P
=Xε∈E(cid:18) ∂P
PN =Xε∈E(cid:18) ∂P
dSO(N,R)Xℓ=1 (cid:20)(cid:18)s −
ASO(N,C)
s,τ
1
2ASO(N,C)
s,τ
Hence
1 +
0 +
s,τ
Vε
1
Proof. For each X ∈ MN (C), we apply the chain rule to get
The result now follows from Theorem 3.19.
+
(3.79)
(3.80)
(3.81)
.
(3.82)
2 (cid:19) P(cid:21)N
1
N 2 eLs,τ
∂vε∂vδ(cid:19) (V) · (eXVε)(eXVδ).
Lemma 3.21 ([8, Lemma 3.28]). There exists a sesquilinear form (with the second
argument conjugate linear)
B : C[u, u−1; v] × C[u, u−1; v] → W
such that, for all P, Q ∈ C[u, u−1; v], we have deg(B(P, Q)) = deg(P ) + deg(Q) and
[B(P, Q)]N (A) = tr[PN (A)QN (A)∗]
for all A ∈ SO(N, C).
4. Limit theorems for the Segal-Bargmann transform on SO(N, R)
We now use the results from the previous section to identify the limit of the Segal-
Bargmann transform on SO(N, R) as N → ∞. In Section 4.1, we prove a few prelimi-
and µSO(N,C)
nary results regarding the way in which the heat kernel measures ρSO(N,R)
concentrate their mass as N → ∞. Then, in Section 4.2, we prove the SO(N, R) version
of Theorem 1.6, the main limit theorem for BSO(N,R)
s,τ
.
s
s,τ
4.1. Concentration of measures
The following lemma is an essential component of our main limit theorems. It is a known
result (see [12, Corollary 6.2.32]); we include a direct proof here for convenience.
Lemma 4.1. Let X, Y ∈ M (N, C) and suppose k · k is a submultiplicative matrix norm.
Then
keX+Y − eXk ≤ kY kekXkekY k.
Proof. For n ≥ 0, note that
k(X + Y )n − X nk ≤ (kXk + kY k)n − kXkn,
where the inequality follows by expanding (X +Y )n, which includes an X n, then applying
the submultiplicativity of the matrix norm, and then recombining terms. Hence
X n
n! (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n!
−
(X + Y )n
keX+Y − eXk =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xn=0
∞Xn=0
∞Xn=0
1
n!k(X + Y )n − X nk
∞Xn=0
1
≤
n!
= ekXk+kY k − ekXk.
≤
[(kXk + kY k)n − kXkn]
It remains to show that ekXk+kY k− ekXk ≤ kY kekXkekY k, which is equivalent to showing
that ey − 1 ≤ yey for y ≥ 0. Rearranging, this inequality is equivalent to 1 − y ≤ e−y,
which holds, in fact, for all y ∈ R: by Bernoulli's inequality,
≥ 1 − y.
e−y = lim
y
n→∞(cid:16)1 −
29
n(cid:17)n
Corollary 4.2. Let V be a finite-dimensional normed vector space over C and suppose
that that L0, L1, and L2 are operators on V . Then there exist constants C1, C2, C3 < ∞
depending on L0, L1, L2,k · kV such that for N sufficiently large,
N L1+ 1
keL0+ 1
N 2 L2 − eL0kEnd(V ) ≤
N L1+ 1
keL0+ 1
N 2 L2 − eL0+ 1
N L1 − eL0kEnd(V ) ≤
C3
N
keL0+ 1
C1
N
N L1kEnd(V ) ≤
,
C2
N 2
(4.1)
(4.2)
(4.3)
where k · kEnd(V ) is the operator norm on End(V ) induced by k · kV . Hence if ϕ ∈ V ∗ is
a linear functional, then for N sufficiently large,
ϕ(eL0+ 1
N L0+ 1
N 2 L2v) − ϕ(eL0v) ≤
C1
N kϕkV ∗kvV ,
ϕ(eL0+ 1
N L1+ 1
N 2 L2v) − ϕ(eL0+ 1
N L1v) − ϕ(eL0 v) ≤
ϕ(eL0+ 1
N L1v) ≤
C2
N 2kϕkV ∗kvV
C3
N kϕkV ∗kvV
(4.4)
(4.5)
(4.6)
where k · kV ∗ is the dual norm on V ∗.
Proof. We set X = L0 and Y = 1
N L1+ 1
keL0+ 1
1
N 2 L2 as in Lemma 4.1, so
N L1 + 1
N kL1kEnd(V )ekL0kEnd(V ) ek 1
1
N kL1kEnd(V )ekL0kEnd(V ) e
N 2 L2 − eL0kEnd(V ) ≤
≤
kL1kEnd(V )+ 1
N 2
0
1
N0
N L1+ 1
N 2 L2kEnd(V )
1
N kL1kEnd(V )+ 1
N 2 kL2kEnd(V ) .
kL2kEnd(V ) ≤ 2. Then (4.1) holds for all
Choose N0 ∈ N such that e
N ≥ N0 by setting C1 = 2kL1kEnd(V )ekL0kEnd(V ) . Equation (4.4) then follows. The
proofs of (4.2), (4.3), and subsequently (4.5) and (4.6), are similar.
We now introduce notation and establish a few preliminary results which we will use for
proving Theorem 1.6 for SO(N, R) in the next section.
Definition 4.3. Define a family of holomorphic functions {νk}k∈Z on C by setting
ν0(τ ) ≡ 1 and for k 6= 0,
νk(τ ) = e− k
2 τ
(−τ )j
j!
kj−1(cid:18) k
j + 1(cid:19).
k−1Xj=0
For τ ∈ C, define the trace evaluation map πτ : C[u, u−1; v] → C[u, u−1] by
(πτ P )(u) = P (u; v)vk=νk(τ ),k6=0.
30
(4.7)
(4.8)
For a concrete example, if P (u; v) = v3v2
−7u6 + 9v1v5
−4, then
(πτ P )(u) = ν3(τ )ν−7(τ )2u6 + 9ν1(τ )ν−4(τ )5.
The following result is a key tool in proving our main limit theorems.
Theorem 4.4 (Biane, [3]). For each s > 0 and k ∈ Z,
N→∞ZU(N )
lim
tr(U k) ρU(N )
s
(U ) dU = νk(s).
For P ∈ C[u, u−1; v], let P (u; 1) = P (u; v)vk=1,k6=0. Using the intertwining formula
(3.45) for the U(N ) case and Theorem 4.4, we have
νk(s) = lim
N→∞ZU(N )
= lim
N→∞
(e
s
2 (L0− 1
N 2 (2K−
1 +K−
2 ))vk)(1).
tr(U k) ρU(N )
s
(U ) dU = lim
N→∞
(e
s
2 ∆U(N ) tr[(·)k])(IN )
(4.9)
(4.10)
1 , and K−
Since L0, K−
2 are linear operators which preserve the finite-dimensional vector
2 ))vk)(1) to be the evaluation of the
space Ck[u, u−1; v], we consider (e
linear functional ϕ(Q) := Q(1) on the finite-dimensional space Cm[v]. Hence Corollary
4.2 applies, so taking the limit as N → ∞, we have
N 2 (2K−
2 (L0− 1
1 +K−
s
νk(s) = (e
s
2 L0 vk)(1).
(4.11)
This yields the following version of Theorem 4.4 for the SO(N, R) case, which is due to
L´evy [13].
Lemma 4.5 (L´evy, [13]). For s > 0 and k ∈ Z,
tr(Ak) ρSO(N,R)
s
(A) dA = νk (s) .
(4.12)
N→∞ZSO(N,R)
lim
Proof. We have
N→∞ZSO(N,R)
lim
tr(Ak) ρSO(N,R)
s
(A) dA = lim
N→∞
(e
= lim
N→∞
(e
= νk (s) ,
s
2 ∆SO(N,R) tr[(·)k])(IN )
2 (L0+ 1
N L(1)
N 2 L(1)
1 + 1
s
2 )vk)(1) = (e
where we have again applied Corollary 4.2 as in (4.11).
Lemma 4.6. For Q ∈ C[v] and τ ∈ C, we have
(cid:0)e
τ
2 L0Q(cid:1) (1) = πτ Q.
31
s
2 L0 vk)(1)
(4.13)
Proof. By Theorem 3.9, e
τ
2 L0 is a homomorphism of C[v]. Hence it suffices to show that
(cid:0)eτ L0vk(cid:1) (1) = πτ (vk) = νk (τ )
(4.14)
τ
k
(cid:12)(cid:12)(cid:0)e
for each k ∈ Z. To see that it holds for all τ ∈ C, we expand the left-hand side as a power
series in τ . Fix a norm k·kWk on the finite-dimensional vector space Wk := Ck[u, u−1; v].
Let k·kEnd(Wk) denote the operator norm on End(Wk) induced by k·kWk, and let k·kW ∗
denote the corresponding dual norm on Wk. Let ϕ be the linear functional acting on Wk
defined by ϕ(P ) = P (1). Then for any τ ∈ C,
(L0)nvk(cid:19) (1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2 L0vk(cid:1) (1)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
∞Xn=0(cid:18) 1
n!(cid:16) τ
2(cid:17)n
n!(cid:18)τ
2 (cid:19)n
∞Xn=0
ϕ((L0)nvk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
and the right-hand side converges by the ratio test. Thus the function τ 7→(cid:0)e
2 L0 vk(cid:1) (1)
is analytic on the complex plane. By (4.11), this function agrees with the entire function
τ 7→ νk (τ ) for τ ∈ R, and so it also agrees for all τ ∈ C.
Lemma 4.7. Let s > 0 and τ = t + iθ ∈ D(s, s). For any Q ∈ C[v],
n!(cid:16) τ
2(cid:17)n
End(Wk)kvkkWk ,
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
k kL0kn
∞Xn=0
kϕkW ∗
≤
1
1
τ
eLs,τ
0
eLs,τ
0
(cid:16)e
(cid:16)e
ι(Q)(cid:17) (1) = πs−τ Q,
ι∗(Q)(cid:17) (1) = πs−τ Q.
(4.15)
(4.16)
(4.18)
(4.19)
(4.20)
Proof. First, observe that if f : SO(N, C) → MN (C) is holomorphic, then gJXf = ieXf
for all X ∈ so(N, R). Thus
2 eX 2 − iθeX 2(cid:21) f = (s − τ )∆SO(N,R)f.
ASO(N,C)
fSO(N,R) = XX∈βso(N,R)(cid:20)(cid:18)s −
2(cid:19) eX 2 −
s,τ
t
t
In particular, since QN is a trace polynomial,
1
2 ASO(N,C)
s,τ
e
QN = e
1
2 (s−τ )∆SO(N,R)QN .
(4.17)
Applying intertwining formulas (3.82) and (1.12) to the left side and using (3.49) gives
eLs,τ
0 + 1
N
he
eLs,τ
1 + 1
N 2
eLs,τ
2
ι(Q)iN
= [e
1
2 (s−τ )D(1)
N Q]N .
Hence for Q ∈ C[v],
eLs,τ
eLs,τ
0 + 1
N
1 + 1
N 2
eLs,τ
2
(cid:16)e
ι(Q)(cid:17) (1) =(cid:16)e
1
2 (s−τ )(L0+ 1
N L(1)
1 + 1
N 2 L(1)
2 )Q(cid:17) (1).
If deg Q = m, then we can view the left and right sides as the evaluation of a linear
functional ϕ(P ) = P (1) on the finite-dimensional spaces Wm and Cm[v], respectively.
Thus we may apply Corollary 4.2 to take the limit as N → ∞, which yields
eLs,τ
0
(cid:16)e
ι(Q)(cid:17) (1) =(cid:16)e
32
1
2 (s−τ )L0Q(cid:17) (1).
By Lemma 4.6, the expression on the right is πs−τ Q. Equation (4.16) is proved similarly,
using the fact that if f : SO(N, C) → MN (C) is antiholomorphic, thengJXf = −ieXf for
all X ∈ so(N, R).
Theorem 4.8. Let s > 0, τ ∈ D(s, s), and k ∈ Z. Then
N→∞ZSO(N,C)
lim
tr(Ak) µSO(N,C)
s,τ
(A) dA = νk (s − τ ) .
(4.21)
Proof. We use (2.4) and intertwining formula (3.82) to get
ZSO(N,C)
tr(Ak) µSO(N,C)
s,τ
1
2 ASO(N,C)
s,τ
tr[(·)k](cid:17) (IN )
eLs,τ
0 + 1
N
eLs,τ
1 + 1
N 2
eLs,τ
2
(A) dA =(cid:16)e
=(cid:16)e
ι(vk)(cid:17) (1).
Viewing the last expression as a linear functional acting on ι(vk) ∈ Wk, we apply Corollary
4.2 to take the limit as N → ∞. Combining this with Lemma 4.7, we have
N→∞ZSO(N,C)
lim
tr(Ak) µSO(N,C)
s,τ
(A) dA =(cid:16)e
eLs,τ
0
ι(vk)(cid:17) (1) = πs−τ vk = νk (s − τ ) .
4.2. The free Segal-Bargmann transform for SO(N, R)
In this section, we prove our main result on the large-N limit of the Segal-Bargmann
transform for SO(N, R). The free Segal-Bargmann transform Gs,τ and the free inverse
Segal-Bargmann transform Hs,τ appearing in Theorem 1.6 have explicit formulas, which
we first describe.
Definition 4.9. Let L0 be the pseudodifferential operator on C[u, u−1; v] from Definition
3.7. We define the free Segal-Bargmann transform to be the map Gs,τ : C[u, u−1] →
C[u, u−1] given by
(4.22)
2 L0 ,
τ
Gs,τ = πs−τ ◦ e
and we define the free inverse Segal-Bargmann transform to be the map Hs,τ :
C[u, u−1] → C[u, u−1] given by
Hs,τ = πs ◦ e− τ
2 L0.
Before proving Theorem 1.6, we require the following result.
Theorem 4.10. Let s > 0 and τ = t + iθ ∈ D(s, s). For any P ∈ C[u, u−1; v],
L2(ρSO(N,R)
s
)
kPN − [πsP ]Nk2
kPN − [πs−τ P ]Nk2
L2(µSO(N,C)
s,τ
and
= O(cid:18) 1
N 2(cid:19) ,
= O(cid:18) 1
N 2(cid:19) .
)
(4.23)
(4.24)
(4.25)
This result shows that with respect to the heat kernel measure, the space of trace poly-
nomials concentrates onto the space of polynomials as N → ∞.
33
Proof. We first show (4.25). It suffices to prove the result for polynomials of the form
P (u; v) = ukQ(v) for k ∈ Z and Q ∈ C[v]; (4.25) then holds on all of C[u, u−1; v] by the
triangle inequality. For such a polynomial,
P (u; v) − πs−τ P (u; v) = uk[Q(v) − πs−τ Q] = ukRs−τ ,
where we let Rs−τ = Q − πs−τ Q. Hence for A ∈ SO(N, C),
kPN (A) − [πs−τ P ]N (A)k2
MN (C) = tr(cid:0)Ak[Rs−τ ]N (A)[Rs−τ ]N (A)∗A∗k(cid:1)
= tr(AkA∗k)[Rs−τ ]N (A)[Rs−τ ]N (A)∗
= [vε(k,k)ι(Rs−τ )ι∗(Rs−τ )]N (A).
Along with intertwining formula (3.82), this allows us to compute
(4.26)
(4.27)
(cid:16)kPN − [πs−τ P ]Nk2
MN (C)(cid:17) (IN )
eLs,τ
eLs,τ
1 + 1
N 2
(4.28)
2 (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))(cid:17) (1)
1 (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))(cid:17) (1)(cid:12)(cid:12)(cid:12)(4.29)
(4.30)
kPN − [πs−τ P ]Nk2
L2(µSO(N,C)
s,τ
)
1
2 ASO(N,C)
s,τ
= e
eLs,τ
0 + 1
N
=(cid:16)e
By the triangle inequality, the last line is bounded by
N
N
eLs,τ
eLs,τ
eLs,τ
eLs,τ
eDs,τ
0 + 1
0 + 1
(cid:12)(cid:12)(cid:12)(cid:16)e
1 (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))(cid:17) (1)(cid:12)(cid:12)(cid:12) .
N (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))(cid:17) (1) −(cid:16)e
+(cid:12)(cid:12)(cid:12)(cid:16)e
N (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))(cid:17) (1) −(cid:16)e
N 2(cid:19) .
= O(cid:18) 1
(cid:12)(cid:12)(cid:12)(cid:16)e
To bound (4.30), recall from Theorem 3.20 that eLs,τ
operator on Wm, so e
0 + 1
0 + 1
0 + 1
eDs,τ
eLs,τ
eLs,τ
eLs,τ
eLs,τ
N
N
1
eLs,τ
0 + 1
N
e
eLs,τ
1 (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))
eLs,τ
1 vε(k,k) · e
0 + 1
= e
eLs,τ
N
If m = deg vε(k,k)ι(Rs−τ )ι∗(Rs−τ ), we can interpret (4.29) as the evaluation of the linear
functional ϕ(R) = R(1) on Wm, so by Corollary 4.2,
1 (vε(k,k)ι(Rs−τ )ι∗(Rs−τ ))(cid:17) (1)(cid:12)(cid:12)(cid:12)
N eLs,τ
1
is a first-order differential
(4.31)
is an algebra homomorphism, i.e.
eLs,τ
0 + 1
N
eLs,τ
1
ι(Rs−τ ) · e
eLs,τ
0 + 1
N
eLs,τ
1
ι∗(Rs−τ ).
(4.32)
For any T ∈ W ,
0 + 1
eLs,τ
N
(cid:12)(cid:12)(cid:12)(e
eLs,τ
1 T )(1)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(e
By Corollary 4.2, the first term on the right side of (4.33) is O(1/N ), while the second
term is a constant not depending on N . Moreover, by Lemma 4.7,
eLs,τ
0 + 1
N
eLs,τ
1 T )(1) − (e
eLs,τ
0 T )(1)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(e
eLs,τ
0 T )(1)(cid:12)(cid:12)(cid:12)
(4.33)
eLs,τ
0
(e
ι(Rs−τ ))(1) = (e
34
eLs,τ
0
ι∗(Rs−τ ))(1) = 0.
(4.34)
Applying (4.33) to each term in the product in (4.32) and using (4.34), we have that
(4.30) is O(1/N 2). Combining this with (4.31) shows (4.25).
Finally, observe that s
SO(N, R) shows (4.24).
2 ∆SO(N,R) = 1
2ASO(N,C)
s,0
. Setting τ = 0 in (4.27) and restricting to
We can now prove Theorem 1.6 for SO(N, R).
Proof of Theorem 1.6 for SO(N, R). Let f ∈ C[u, u−1]. Using Theorem 1.5 to rewrite
BSO(N,R)
s,τ
fN and applying the triangle inequality, we have
kBSO(N,R)
s,τ
)
τ
s,τ
2 D(1)
fN − [Gs,τ f ]NkL2(µSO(N,C)
N f ]N − [πs−τ ◦ e
= k[e
N f ]N − [e
≤ k[e
+ k[e
2 D(1)
τ
τ
τ
τ
2 L0 f ]NkL2(µSO(N,C)
s,τ
)
2 L0f ]NkL2(µSO(N,C)
s,τ
)
2 L0 f ]N − [πs−τ ◦ e
2 L0 f ]NkL2(µSO(N,C)
s,τ
τ
By Theorem 4.10, (4.36) is O(1/N ), so it remains to show that
k[e
τ
2 D(1)
N f ]N − [e
τ
2 L0f ]Nk2
L2(µSO(N,C)
s,τ
N 2(cid:19) .
= O(cid:18) 1
)
To this end, let m = deg f and R(N ) = e
and the sesquilinear form B of Lemma 3.21, we have
N f − e
τ
2 D(1)
τ
2 L0f ∈ Cm[u, u−1; v]. Recalling (1.8)
k[e
τ
2 D(1)
N f ]N − [e
τ
2 L0f ]Nk2
L2(µSO(N,C)
s,τ
)
).
(4.35)
(4.36)
(4.37)
1
)
s,τ
2 AN
L2(µSO(N,C)
= k[R(N )]Nk2
= e
=(cid:16)e
s,τ(cid:16)k[R(N )]Nk2
N B(R(N ), R(N ))(cid:17) (1).
MN (C)(cid:17) (IN )
eDs,τ
Fix norms k · kW2m and k · kCm[u,u−1;v] on W2m and Cm[u, u−1; v], respectively. By
Corollary 4.2 applied to the linear functional ϕ(P ) = P (1) on W2m, there exists a constant
C = C(m, s, τ ) (not dependent on N ) such that
eDs,τ
N B(R(N ), R(N ))(cid:17) (1) −(cid:16)e
(cid:12)(cid:12)(cid:12)(cid:16)e
Next, using the linear functional ψ(P ) =(cid:0)e
0 B(R(N ), R(N ))(cid:17) (1)(cid:12)(cid:12)(cid:12) ≤
eLs,τ
τ
2 L0 P )(cid:1) (1) on W2m, we have
2mkB(R(N ), R(N ))kW2m .
C
N B(R(N ), R(N ))kW2m.
Putting (4.37), (4.39), and (4.40) together yields
(cid:12)(cid:12)(cid:12)(cid:16)e
eLs,τ
0 B(R(N ), R(N ))(cid:17) (1)(cid:12)(cid:12)(cid:12) ≤ kψkW ∗
) ≤(cid:18)kψkW ∗
2 L0f ]Nk2
L2(µSO(N,C)
s,τ
τ
35
k[e
τ
2 D(1)
N f ]N − [e
2m +
C
N(cid:19)kB(R(N ), R(N ))kW2m .
(4.38)
(4.39)
(4.40)
(4.41)
Since B : Cm[u, u−1; v] × Cm[u, u−1; v] → W2m is a sesquilinear form with finite domain
and range, there exists a constant C′ = C′(m) (not dependent on N ) such that
for all P, Q ∈ Cm[u, u−1; v]. (4.42)
kB(P, Q)kW2m ≤ C′kPkCm[u,u−1;v]kQkCm[u,u−1;v]
A final application of Corollary 4.2 shows that there exists a constant C′′ depending on
L0, L(1)
1 , L(1)
2 , τ , k · kCm[u,u−1;v], and f , but again, not on N , such that
C′′
N
kR(N )kCm[u,u−1;v] = ke
2 L0 fkCm[u,u−1;v] ≤
.
(4.43)
N f − e
2 D(1)
τ
τ
Hence
kB(R(N ), R(N ))kW2m ≤ kR(N )k2
Cm[u,u−1;v] ≤
(C′′)2
N 2 .
(4.44)
Combining (4.41) and (4.44) proves (4.37).
To prove (1.16), we observe that (BSO(N,R)
triangle inequality argument using (4.24), it suffices to show
s,τ
2 ∆SO(N,R)fN . By a similar
k[e− τ
2 D(1)
N f ]N − [e− τ
2 L0 f ]Nk2
(4.45)
)−1fNSO(N,R) = e− τ
= O(cid:18) 1
N 2(cid:19) .
L2(ρSO(N,R)
)
s
Replacing τ with −τ in the definition of R(N ) and (s, τ ) with (s, 0) from (4.38) onwards,
the same argument as above proves (4.45).
For uniqueness, we first define seminorms on C[u, u−1; v] by
kPk(1)
kPk(1)
s,N = kPNkL2(ρSO(N,R)
),
s,τ,N = kPNkL2(µSO(N,C)
s,τ
s
),
In addition, for β = 1, 2, define the seminorms
kPk(2)
kPk(2)
s,N = kPNkL2(ρU(N )
s,τ,N = kPNkL2(µ
s
),
GLN (C)
s,τ
kPk(β)
kPk(β)
s,N
N→∞kPk(β)
N→∞kPk(β)
s,τ,N
s = lim
s,τ = lim
.
)
(4.46)
(4.47)
(4.48)
(4.49)
In [8, Lemma 4.7], it was shown that for β = 2, the seminorms (4.48) and (4.49) are
norms when restricted to C[u, u−1]. Since
kPk(1)
s = kPk(2)
s
(4.50)
it follows that the seminorm (4.48) is also a norm on C[u, u−1] for β = 1. An argument
completely analogous to the one used in [8, Lemma 4.7] to prove that kPk(2)
s,τ is a norm
on C[u, u−1] shows that kPk(1)
s,τ ∈ C[u, u−1]
s,τ is a norm on C[u, u−1]. Hence, if gs,τ , g′
both satisfy
kBSO(N,R)
s,τ
fN − [gs,τ ]Nk2
L2(µSO(N,C)
s,τ
= O(cid:18) 1
N 2(cid:19) = kBSO(N,R)
s,τ
36
)
fN − [g′
s,τ ]Nk2
L2(µSO(N,C)
s,τ
,
)
the triangle inequality implies that
kgs,τ − g′
s,τkL2(µSO(N,C)
s,τ
) = O(cid:18) 1
N 2(cid:19) .
s,τ = 0. Since k · k(1)
(4.51)
s,τ is a norm on C[u, u−1], it
s,τ . A similar argument shows that Hs,τ f is the unique polynomial
Taking N → ∞ shows that kgs,τ − g′
follows that gs,τ = g′
satisfying (1.16).
Example 4.11. Set P (u; v) = u2 ∈ C[u, u−1]. Then we can show that
s,τk(1)
BSO(N,R)
s,τ
τ
2 ∆SO(N,R) PN (A)
PN (A) = e
1
(1 − e−τ )IN +
N
=
1
2
e−τ (−1 + e
via the same method used in [8, Example 3.5]. By the intertwining formula e
[e
N P ]N , this corresponds to the trace polynomial
N )A2 −
e−τ (1 + e
2 D(1)
τ
2τ
2τ
N )A tr A,
τ
2 ∆SO(N,R) PN =
N
2
τ
2 D(1)
N P )(u; v) =
(e
e−τ (1 + e
2τ
N )u2 −
N
2
e−τ (−1 + e
2τ
N )uv1
1
N
(1 − e−τ ) +
1
2
= e−τ (u2 − τ uv1) + O(cid:18) 1
N 2(cid:19) .
N P at the moments
By Theorem 4.10, we compute Gs,τ P by evaluating the traces in e
νs−τ and taking the large-N limit of the resulting polynomial. Since ν1(s−τ ) = e−(s−τ )/2,
this yields
τ
2 D(1)
In [8, Example 3.5], it was shown that for the unitary case (with τ = t > 0),
Gs,τ P (u) = e−τ (u2 − τ e−(s−τ )/2u).
t
2 D(2′ )
(e
N P )(u; v) = e−t cosh(t/N )u2 − N e−t sinh(t/N )uv1
= e−τ (u2 − τ uv1) + O(cid:18) 1
N 2(cid:19) ,
which also results in the same expression for Gs,tP .
5. The Segal-Bargmann transform on Sp(N )
In this section, we prove the Sp(N ) version of Theorems 1.4, 1.5, and 1.6. which comprise
our main results. We begin with a slight digression and show that Sp(N ) can also be
realized as a subset of N × N quaternion matrices. This allows us to easily identify a par-
ticular orthonormal basis of sp(N ) ⊆ M2N (C) which we use to compute a set of magic
2 ASp(N,C)
.
formulas used in the proofs of our intertwining formulas for e
These magic formulas share key similarities with the magic formulas for SO(N, R). Con-
sequently, many of the results for the Sp(N ) case can proven using techniques similar to
those used in the SO(N, R) case, and so we do not always provide the full details. We
conclude by showing that when Sp(N ) is realized as a subset of MN (H), a version of the
magic formulas and intertwining formulas for ∆Sp(N ) also hold.
2 ∆Sp(N ) and e
s,τ
1
τ
37
5.1. Two realizations of the compact symplectic group
The quaternion algebra H is the four-dimensional associative algebra over R spanned
by 1, i, j, and k satisfying the relations
and
i2 = j2 = k2 = −1
ij = k,
jk = i,
ki = j,
ji = −k,
kj = −i,
ik = −j.
We denote the quaternion conjugate of q ∈ H by q∗. That is, if q = a1 + bi + cj + dk,
then q∗ = a1 − bi − cj − dk. If A ∈ MN (H), we define the adjoint of A to be the matrix
A∗ defined by (A∗)i,j = (Aj,i)∗. For each N ≥ 1, we define fSp(N ) as
(5.1)
Its Lie algebra is
which we endow with the real inner product
fSp(N ) = {A ∈ MN (H) : A∗A = IN}.
esp(N ) = {X ∈ MN (H) : X ∗ = −X},
(5.2)
(5.3)
,
Let
hX, Y i esp(N ) = 2N Re Tr(X ∗Y )
CN =(cid:26) 1
√4N
for all X, Y ∈ esp(N ).
Ea,a(cid:27)1≤a≤N
An orthonormal basis for esp(N ) with respect to this inner product is
(Ea,b + Eb,a)(cid:27)1≤a<b≤N ∪(cid:26) 1
√2N
(Ea,b − Eb,a)(cid:27)1≤a<b≤N ∪ iCN ∪ jCN ∪ kCN .
β esp(N ) =(cid:26) 1
√4N
where Ea,b ⊆ MN (H) is the N ×N matrix with a 1 in the (a, b) entry and zeros elsewhere.
(5.4)
From (5.1), we see that fSp(N ) is the unitary group over the quaternions. We now
explicitly construct a (real) Lie group isomorphism between fSp(N ) and Sp(N ). While
fSp(N ) is, in some sense, the more natural realization of the compact symplectic group,
for our purposes it is easier to work with the definition given in (2.5).
In particular,
the complexification of the compact symplectic group is more readily understood when
Sp(N ) is realized as a subset of M2N (C) rather than of MN (H).
The material presented below is standard (cf. [18]); we include it here for convenience.
First, observe that we can realize H as a subalgebra of M2(C). We do so by letting
ψ : H → M2(C) denote the injective homomorphism
H ∋ q = a1 + bi + cj + dk 7→ (cid:20) a + di −b − ci
a − di (cid:21) ∈ M2(C).
b − ci
38
(5.5)
The map ψ preserves conjugation: ψ(q∗) = ψ(q)∗ for all q ∈ H, where we take the
quaternion conjugate on the left and the complex conjugate transpose on the right.
Moreover, note that we can write a matrix of the form (5.5) as
(cid:20) α −β
α (cid:21)
β
(5.6)
for some α, β ∈ C, and any matrix of this form corresponds to a q ∈ H.
Now define a map Φ : MN (H) → M2N (C) by
Φ([qi,j ]1≤i,j≤N ) = [ψ(qi,j)]1≤i,j≤N .
By the properties of block multiplication, Φ is an algebra homomorphism. Since ψ is
one-to-one, it follows that Φ is also one-to-one. Moreover, since ψ preserves conjugation,
if A ∈ MN (H), then Φ(A∗) = Φ(A)∗, where again we take the quaternion
so does Φ:
conjugate transpose on the left and the complex conjugate transpose on the right.
Consequently, if A ∈fSp(N ) ⊆ MN (H), so that A∗A = IN , then
Φ(A)∗Φ(A) = Φ(A∗)Φ(A) = Φ(A∗A) = I2N .
Hence
Φ(fSp(N )) = {Z ∈ M2N (C) : Z ∗Z = I2N and ∃ A ∈ MN (H) such that Φ(A) = Z}.
We would like to show that Φ(fSp(N )) = Sp(N ), which would imply that Sp(N ) ∼=fSp(N )
as groups. Recall that since A ∈ Sp(N ) is unitary and Ω−1 = −Ω, we can rewrite the
definition of Sp(N ) as follows:
(5.7)
Sp(N ) = {Z ∈ M2N (C) : Z ∗Z = I2N ,−ΩZΩ = Z}.
Comparing (5.7) and 5.8, we see that it suffices to show that
Z = Φ(A) for some A ∈ MN (H) ⇔ −ΩZΩ = Z.
(5.8)
(5.9)
We need the following lemma.
Lemma 5.1. Let B ∈ M2(C). Then −Ω0BΩ0 = B if and only if B =(cid:20) α −β
α (cid:21) for
β
some α, β ∈ C.
Proof. Direct computation.
Consequently, for Z ∈ M2N (C),
[−ΩZΩ]i,j = X1≤k,ℓ≤N
[Ω]i,k[Z]k,ℓ[Ω]ℓ,j = −Ω0[Z]i,jΩ0,
(5.10)
where [Z]i,j denotes the (i, j)th 2 × 2 block of Z. By Lemma 5.1, we see that for each
1 ≤ i, j ≤ N , there exists qi,j ∈ H such that ψ(qi,j) = [Z]i,j precisely when −ΩZΩ = Z.
39
Theorem 5.2. The map ΦfSp(N ) :fSp(N ) → Sp(N ) is a Lie group isomorphism.
Proof. The preceding discussion shows that ΦfSp(N ) is a group isomorphism. It remains
to be shown that ΦfSp(N ) and Φ−1
are smooth, but this is clear from the definition
of Φ.
fSp(N )
The inner products (2.10) and (5.3) for sp(N ) and esp(N ), resp., are related by Φ. For
X ∈ MN (H), note that
Re Tr X =
Re Tr Φ(X) =
Tr Φ(X).
(5.11)
1
2
1
2
Hence for X, Y ∈ esp(N ),
hX, Y i esp(N ) = 2N Re Tr(XY ∗) = N Tr(Φ(XY ∗))
= N Tr(Φ(X)Φ(Y )∗) = hΦ(X), Φ(Y )isp(N ) .
(5.12)
In particular, if β esp(N ) is an orthonormal basis for esp(N ) ⊆ MN (H) with respect to (5.2),
then Φ(β esp(N )) is an orthonormal basis for sp(N ) ⊆ M2N (C) with respect to (2.10).
5.2. Magic formulas and derivative formulas
Proposition 5.3 (Magic formulas for Sp(N )). Let βsp(N ) be any orthonormal basis for
sp(N ) with respect to the inner product (2.10). For any A, B ∈ M2N (C),
X 2 = −
2N + 1
2N
I2N = −I2N −
1
2N
I2N
XAX = −
XX∈βsp(N )
XX∈βsp(N )
XX∈βsp(N )etr(XA)X =
XX∈βsp(N )etr(XA)etr(XB) =
1
2N
ΩA⊺Ω−1 −etr(A)I2N
1
4N 2 (ΩA⊺Ω−1 − A)
1
4N 2 (etr(ΩA⊺ΩB) −etr(AB))
(5.13)
(5.14)
(5.15)
(5.16)
Proof. As with the magic formulas for SO(N, R), the quantities on the left hand side of
(5.13), (5.14), (5.15), and (5.16) are independent of choice of orthonormal basis. Thus we
may compute with respect to the orthonormal basis βsp(N ) := Φ(β esp(N )), where β esp(N )
is the orthonormal basis described in (5.4).
We will use the following notation: for A ∈ M2N (C), let [A]i,j denote the (i, j)th 2 × 2
submatrix, and let Ai,j
If Ea,b ⊆ MN (H) is an
elementary matrix and γ ∈ {1, i, j, k} ⊆ H,
k,ℓ denote the (k, ℓ)th entry of [A]i,j.
F γ
a,b := Φ(γEa,b) ∈ M2N (C).
40
(5.17)
Observe that [F γ
one of
a,b]i,j contains all zeros unless (i, j) = (a, b), in which case it is equal to
0
ψ(1) =(cid:20) 1
0 −i
−i
depending on whether γ equals 1, i, j, or k, respectively.
1 (cid:21) , ψ(i) =(cid:20) 0 −1
0 (cid:21) , ψ(j) =(cid:20)
0
1
0 (cid:21) , ψ(k) =(cid:20) i
0 −i (cid:21) ,
0
Equation (5.13) follows from (5.14). For (5.14), we have
XX∈βsp(N )
XAX =
1
4N X1≤a,b≤N Xγ∈{1,i,j,k}
F 1
4N X1≤a,b≤N
−
1
F γ
a,bAF γ
a,b
a,bAF 1
b,a − Xγ∈{i,j,k}
F γ
a,bAF γ
b,a
For 1 ≤ a, b, c, d, i, j ≤ N , we have
[F γ
[F γ
a,bAF γ
c,d]i,j = X1≤k,ℓ≤N
=(02
ψ(γ)[A]b,cψ(γ)
i 6= a or j 6= d
i = a and j = d
a,b]i,k[A]k,ℓ[F γ
c,d]ℓ,j = [F γ
a,b]i,b[A]b,c[F γ
c,d]c,j
A straightforward computation shows that
Hence for the first term on the right hand side of (5.18),
1,2
Ab,c
1,1 Ab,c
2,1 Ab,c
2,2 −Ab,c
1,2 −Ab,c
ψ(1)[A]b,cψ(1) =" Ab,c
ψ(j)[A]b,cψ(j) =" −Ab,c
−Ab,c
1
4N X1≤a,b≤N Xγ∈{1,i,j,k}
2,1
2,2 # , ψ(i)[A]b,cψ(i) =" −Ab,c
1,1 # , ψ(k)[A]b,cψ(k)" −Ab,c
4N Xγ∈{1,i,j,k}
a,b
2N " −Aj,i
1,1 #
Ab,c
2,2
2,1
1,2 −Ab,c
Ab,c
2,2 #
Ab,c
1,1
1,2
2,1 −Ab,c
Ab,c
i,j
1,1 # .
Aj,i
1,2
2,2
2,1 −Aj,i
Aj,i
F γ
a,bAF γ
F γ
i,jAF γ
=
=
i,j
i,j
1
1
(5.18)
(5.19)
(5.20)
(5.21)
(5.22)
(5.23)
For the second term on the right hand side of (5.18), the (i, j)th block is 02 for i 6= j.
41
For i = j,
i,j
i,j
F γ
i,bAF γ
F γ
a,bAF γ
b,a
b,i
ψ(γ)[A]b,bψ(γ)
2,2 #
0
1,1 + Ab,b
1
=
=
1
4N
i,bAF 1
a,bAF 1
1
F 1
4N X1≤a,b≤N
b,a − Xγ∈{i,j,k}
4N
F 1
NXb=1
b,i − Xγ∈{i,j,k}
[A]b,b − Xγ∈{i,j,k}
NXb=1
NXb=1" Ab,b
" −Aj,i
" −Aj,i
XAX
2,2
0 Ab,b
1,1 + Ab,b
Tr(A)I2.
1
2N
1
2N
1
2N
=
=
=
i,j
1,1 #
Aj,i
1,2
2,2
2,1 −Aj,i
Aj,i
1,1 # − Tr(A)I2
Aj,i
2,2
1,2
2,1 −Aj,i
Aj,i
i 6= j
i = j.
Putting this together, we have
XX∈βsp(N )
On the other hand, we can compute
(cid:20)−
1
2N
ΩA⊺Ω−1 −etr(A)I2N(cid:21)i,j
1
2N (Ω0([A]j,i)⊺Ω−1
0
Ω0([A]j,i)⊺Ω−1
0 − Tr(A)I2
= −
Multiplying out (5.29) and comparing with (5.28) proves (5.14).
Next, for (5.15), we again use the basis βsp(N ) to compute
(5.24)
(5.25)
(5.26)
(5.27)
(5.28)
i 6= j,
i = j.
(5.29)
(5.30)
b,a . (5.31)
XX∈βsp(N )etr(XA)X =
1
1
−
4N X1≤a,b≤N Xγ∈{1,i,j,k}etr(F γ
etr(F 1
4N X1≤a,b≤N
etr(F i
etr(F k
1,2 + Ab,a
2,1),
1,1 + Ab,a
(Ab,a
(Ab,a
2,2),
i
2N
42
a,bA) =
1
2N
a,bA) = −
etr(F 1
etr(F j
For 1 ≤ a, b ≤ N , a simple calculation shows that
a,bA)F γ
a,b
a,bA)F 1
b,a − Xγ∈{i,j,k}etr(F γ
a,bA)F γ
a,bA) =
a,bA) =
1
2N
(Ab,a
1,2 − Ab,a
2,1)
1,1 − Ab,a
(Ab,a
2,2),
i
2N
(5.32)
(5.33)
The (i, j)th block of the right hand side of (5.30) is thus
1
4N X1≤a,b≤N Xγ∈{1,i,j,k}etr(F γ
a,bA)F γ
a,b
i,j
=
=
1
4N etr(F 1
4N 2" Aj,i
−Aj,i
1
i,j A)F 1
i,j + Xγ∈{i,j,k}etr(F γ
1,1 # .
2,2 −Aj,i
1,2
Aj,i
2,1
i,jA)F γ
Similarly, the (i, j)th block of (5.31) is
1
=
1
4N X1≤a,b≤N
etr(F 1
4N etr(F 1
4N 2" Ai,j
XX∈βsp(N )etr(XA)X
Ai,j
=
=
i,j
1
a,bA)F 1
b,a − Xγ∈{i,j,k}etr(F γ
i,j − Xγ∈{i,j,k}etr(F γ
2,2 # .
1,2
j,iA)F 1
1,1 Ai,j
2,1 Ai,j
a,bA)F γ
j,iA)F γ
i,j
b,a
i,j
i,j
1
4N 2" Aj,i
−Aj,i
2,2 − Ai,j
2,1 − Ai,j
2,2 #
1,2 − Ai,j
1,1 −Aj,i
1,1 − Ai,j
Aj,i
1,2
i,j
Combining our computations,
i,j
i,j
(5.34)
(5.35)
(5.36)
(5.37)
Again, we can expand
(cid:20) 1
4N 2 (ΩA⊺Ω−1 − A)(cid:21)i,j
B ∈ M2N (C) and taking etr proves (5.16).
=
1
4N 2 [−Ω0([A]j,i)⊺Ω−1
0 − A]i,j
to see that this is equivalent to (5.36). This proves (5.15); multiplying on the right by
While Proposition 5.3 holds for all A, B ∈ M2N (C), we now restrict to the case in which
A and B are in Sp(N ) or Sp(N, C). Note that for A ∈ Sp(N, C), ΩA⊺Ω−1 = A−1.
The magic formulas for Sp(N ) allow us to compute the following derivative formulas, as
in the SO(N, R) case.
Proposition 5.4 (Derivative formulas for Sp(N )). The following identities hold on
43
Sp(N ) and Sp(N, C):
1
1
N
(m − j)etr(Aj )Am−j
∆Sp(N )Am = 1m≥2−
m−1Xj=1
m−1Xj=1
(m − j)Am−2j − 2
+(cid:18)−1 −
2N(cid:19) mAm, m ≥ 0
(−m + j)etr(Aj)Am−j
∆Sp(N )Am = 1m≤−2−
−1Xj=m+1
−1Xj=m+1
(−m + j)Am−2j − 2
−(cid:18)−1 −
2N(cid:19) mAm, m < 0
XX∈βsp(N ) eXetr(Am) · eXAp =
mp
4N 2 (Ap−m − Ap+m), m, p ∈ Z.
1
N
1
(5.39)
(5.40)
(5.38)
Proof. For m ≥ 0, we use (3.15) and magic formulas (5.83) and (5.84) to sum over all
X ∈ βSp(N ):
∆Sp(N )Am = 21m≥2 XX∈βso(N,R) Xk,j6=0
k+j+ℓ=m
Aj X 2Am−j (5.41)
mXj=1
AkXAjXAℓ + XX∈βso(N,R)
Ak(Aj)∗Aℓ −etr(Aj )Ak+ℓ(cid:21) −
(cid:20)−
1
2N
m(2N + 1)
2N
k+j+ℓ=m
= 21m≥2 Xk,j6=0
= 1m≥2−
+(cid:18)−1 −
1
N
m−1Xj=1
(m − j)Am−2j − 2
2N(cid:19) mAm.
1
(m − j)etr(Aj )Am−j
m−1Xj=1
Am
(5.42)
(5.43)
For m < 0, we use (3.10) and magic formulas (5.83) and (5.84) to sum over all X ∈ βSp(N ):
∆Sp(N )Am = 21m≥2 XX∈βso(N,R) Xk,ℓ<0,j≤0
j+k+ℓ=m
AjX 2Am−j
0Xj=m+1
AkXAjXAℓ + XX∈βso(N,R)
Ak(Aj )∗Aℓ −etr(Aj)Ak+ℓ(cid:21) +
(−m + j)etr(Aj)Am−j
−1Xj=m+1
m(2N + 1)
Am
2N
(−m + j)Am−2j − 2
j+k+ℓ=m
1
2N
= 21m≥2 Xk,ℓ<0,j≤0
= 1m≤−2−
−(cid:18)−1 −
(cid:20)−
−1Xj=m+1
2N(cid:19) mAm.
1
N
1
44
Finally, comparing the magic formulas (3.7) and (5.15), we see that the proof of (5.40)
is identical to that of (3.14).
5.3. Intertwining formulas for ∆Sp(N )
The derivative formulas allow us to prove the intertwining formula for ∆Sp(N ).
Proof of Theorem 1.4 for Sp(N ). We proceed as in the proof of Theorem 1.4 for the
SO(N, R) case. Again, it suffices to show that ∆Sp(N )PN = [D(4)
N P ]N holds for P (u; v) =
umq(v), where m ∈ Z \ {0} and q ∈ C[v]. Then PN = Wm · q(V), and by the product
rule,
∆Sp(N )PN = (∆Sp(N )Wm) · q(V) + 2 XX∈βsp(N ) eXWm · eXq(V) + Wm · (∆Sp(N )q(V))
(5.44)
For the first term in (5.44), we consider the cases m ≥ 0 and m < 0 separately. For
m ≥ 0, we apply (5.38) to get
(∆Sp(N )Wm) · q(V) = −
m(2N + 1)
2N
2N + 1
Wm · q(V)
1
N
+ 1m≥2(cid:20) −
m−1Xk=1
2N (cid:20)u
∂uR+P(cid:21)N
2N "−2 ∞Xk=1
− 2" ∞Xk=1
vkuR+ ∂
2N (cid:20)u
∂uR+P(cid:21)N −
2N + 1
+
1
∂
∂
= −
= −
(m − k)Wm−2k − 2
(m − k)VkWm−k(cid:21)q(V)
m−1Xk=1
∂uMu−kR+! P#N
u−k+1R+ ∂
∂uMu−kR+! P#N
1
N
[Z +
2 P ]N − 2[Y +
1 P ]N .
(5.45)
Similarly, for m < 0, we apply (5.39) to get
2N + 1
(∆Sp(N )Wm) · q(V) =
2N (cid:20)u
2 annihilates C[u−1] while Y −
∂uR−P(cid:21)N
1 ,Z −
1 ,Z +
∂
Since Y +
m ∈ Z,
+
1
N
[Z −
2 P ]N + 2[Y −
1 P ]N .
2 annihilates C[u], we have that for all
(∆Sp(N )Wm) · q(V) = −
2N + 1
2N
[N1P ]N −
1
N
[Z2P ]N − 2[Y1P ]N .
(5.46)
45
For the middle term in (5.44), we note the similarity between the derivative formulas
(3.14) and (5.40) for the SO(N, R) and Sp(N ) cases, resp. Hence, as in (3.38),
1
4N 2 [K1P ]N .
XX∈βsp(N ) eXWm · eXq(V) =
q(cid:19) (V) · eX 2Vk + Xj,k≥1(cid:18) ∂2
∂vj∂vk
q(cid:19) (V) · eXVj · eXVk.,
(5.47)
(5.48)
For the last term in (5.44), we have, for each X ∈ βsp(N ),
eX 2q(V) = Xk≥1(cid:18) ∂
∂vk
as with (3.39).
In summing the first term in (5.48) over X ∈ βsp(N ), we break up the sum for positive
and negative k. Using (5.38) and (5.39), we have
XX∈βsp(N )
∞Xk=1(cid:18) ∂
∂vk
q(cid:19) (V) · eX 2Vk = −
−
1
N
2N + 1
2N
∂vk
kVk(cid:18) ∂
q(cid:19) (V)
∞Xk=1
(k − ℓ)Vk−2ℓ(cid:18) ∂
k−1Xℓ=1
∞Xk=2
(k − ℓ)VℓVk−ℓ(cid:18) ∂
k−1Xℓ=1
∞Xk=2
∂vk
∂vk
− 2
q(cid:19) (V)
q(cid:19) (V)
(5.49)
and
XX∈βsp(N )
−1Xk=−∞(cid:18) ∂
∂vk
q(cid:19) (V) · eX 2Vk =
2N + 1
2N
∂vk
kVk(cid:18) ∂
q(cid:19) (V)
−1Xk=−∞
(k − ℓ)Vk−2ℓ(cid:18) ∂
−1Xℓ=k+1
−2Xk=−∞
(−k + ℓ)VℓV−k−ℓ(cid:18) ∂
−2Xk=−∞
−1Xℓ=k+1
∂vk
1
N
−
− 2
q(cid:19) (V)
q(cid:19) (V)
∂vk
(5.50)
Summing the second term in (5.48) over X ∈ βsp(N ) and using magic formula (5.40)
yields
XX∈βsp(N ) Xj,k≥1(cid:18) ∂2
∂vj∂vk
q(cid:19) (V) · eXVj · eXVk
=
1
4N 2 Xj,k≥1
jk(Vk−j − Vk+j )(cid:18) ∂2
∂vj∂vk
q(cid:19) (V).
(5.51)
46
Adding (5.49), (5.50), and (5.51) together and multiplying by Wm, we get
Wm · (∆Sp(N )q(V)) =(cid:20)(cid:18)−
2N + 1
2N N0 −
1
N Z1 − 2Y2 +
1
4N 2K2(cid:19) P(cid:21)N
.
(5.52)
Combining (5.46), (5.47), and (5.52) proves (1.11) for Sp(N ). Equation (1.12) now
follows.
τ
2 D(4)
N P ]N . Corollary 3.11 shows that [e
Proof of Theorem 1.5 for Sp(N ). The proof for Sp(N ) is entirely analogous to the proof
2 ∆Sp(N )PN =
for SO(N, R). Using the intertwining formula for e
[e
N P ]N is a well-defined trace polynomial
on Sp(N ), so its analytic continuation to Sp(N, C) is given by the same trace polyno-
mial function. Hence [e
N P ]N , viewed as a holomorphic function on Sp(N, C), is the
2 ∆Sp(N )PN , and so is equal to BSp(N )
analytic continuation of e
2 ∆Sp(N ) (1.12), we have e
s,τ PN .
2 D(4)
2 D(4)
τ
τ
τ
τ
t
5.4. Intertwining formulas for ASp(N,C)
Theorem 5.5. Fix s ∈ R and τ = t + iθ ∈ C. There are collections {T s,τ
{U s,τ
monomials of trace degree ε such that
: ε ∈ E } such that for each ε ∈ E , T s,τ
and U s,τ
s,τ
ε
ε
ε
ε
: ε ∈ E },
are certain finite sums of
ASp(N,C)
s,τ
Vε = [T s,τ
ε
]N +
1
N
[U s,τ
ε
]N ,
(5.53)
Moreover, let {Ss,τ
3.19. Then for ε, δ ∈ E ,
ε,δ : ε, δ ∈ E } ⊆ W be the collection of word polynomials from Theorem
dsp(N )Xℓ=1 (cid:20)(cid:18)s −
t
2(cid:19) (eXℓVε)(eXℓVδ) +
t
2
(eYℓVε)(eYℓVδ) − θ(eXℓVε)(eYℓVδ)(cid:21) =
1
N 2 [Ss,τ
ε,δ ]N ,
(5.54)
dsp(N )
ℓ=1
where βsp(N ) = {Xℓ}
For the proof, we will employ the conventions, mutandis mutatis, as those used in the
proof of Theorem 3.19 (see the remarks following the theorem statement).
and Yℓ = iXℓ.
Proof. Fix a word ε = (ε1, . . . , εm) ∈ E . Then for each X ∈ β± and A ∈ Sp(N ), we
apply the product rule twice to get
(eX 2Vε)(A) =
mXj=1
tr(Aε1 ··· (AX 2)εj ··· Aεm )
+ 2 X1≤j<k≤m
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm ).
(5.55)
(5.56)
47
Applying magic formula (5.13) to each term in (5.56), we have
XX∈β±
tr(Aε1 ··· (AX 2)εj ··· Aεm ) = ±
2N + 1
2N
tr(Aε1 ··· Aεj ··· Aεm ) = ±
2N + 1
2N
Vε(A).
(5.57)
Summing over 1 ≤ j ≤ m now gives
XX∈β±
mXj=1
tr(Aε1 ··· (AX 2)εj ··· Aεm ) =
2N + 1
2N
n±(ε)Vε(A),
(5.58)
where n±(ε) ∈ Z and n±(ε) ≤ ε (where the integers n±(ε) are not necessarily the
same as in the SO(N, R) case). For (5.56), we can express each term in the sum as
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm ) = ± tr(Aε0
j,kε1
j,k, and ε2
j,k are substrings of ε such that ε0
j,k, ε1
so ε0
X ∈ β± by magic formula (5.14), we have
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm ) = ±(cid:20) −
XX∈β±
j,k XAε1
j,k XAε2
j,k ),
(5.59)
j,kε2
j,k = ε. Summing (5.59) over
1
2N
tr(Aε0
j,k (Aε1
j,k )−1Aε2
j,k )
j,k Aε2
j,k ) tr(Aε1
j,k )(cid:21)
− tr(Aε0
Vε3
j,k
1
2N
= ±
(A) ± Vε0
j,kε2
j,k
(A)Vε1
j,k
(A),
(5.60)
(A) = tr(Aε0
j,k (Aε1
j,k )−1Aε2
j,k ). Thus
j,k is the word of length ε such that Vε3
where ε3
summing (5.56) over X ∈ β± gives
j,k
XX∈β± X1≤j<k≤m
tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm )
2N X1≤j<k≤m
(A) + X1≤j<k≤m
±Vε3
=
1
j,k
±Vε1
j,k
(A)Vε0
j,kε2
j,k
(A).
(5.61)
We define word polynomials corresponding to (5.61) by
T ±
ε,0 = X1≤j<k≤m
ε,1 = X1≤j<k≤m
T ±
±vε3
j,k
±vε0
j,kε2
j,k
vε1
j,k
.
Now if X ∈ β+ and Y = iX ∈ β−, then
± tr(Aε1 ··· (AX 2)εj ··· Aεm )
(eXeY Vε)(A) =
i
2
mXj=1
+ i X1≤j<k≤m
± tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm ).
48
(5.62)
(5.63)
(5.64)
(5.65)
Summing (5.64) over X ∈ β+ and applying (5.83) gives
XX∈β+
mXj=1
± tr(Aε1 ··· (AX 2)εj ··· Aεm ) =
2N + 1
2N
η(ε)Vε(A)
(5.66)
where η(ε) ∈ Z and η(ε) ≤ ε. In addition, summing (5.65) over X ∈ β+, we have
XX∈β+ X1≤j<k≤m
± tr(Aε1 ··· (AX)εj ··· (AX)εk ··· Aεm)
(A) − X1≤j<k≤m
2N X1≤j<k≤m
= −
Vε3
1
j,k
±Vε0
j,kε2
j,k
(A)Vε1
j,k
(A).
(5.67)
We define corresponding word polynomials
Tε,2 = i X1≤j<k≤m
Tε,3 = i X1≤j<k≤m
Putting this together, we define
±vε3
j,k
±vε0
j,kε2
j,k
vε1
j,k
.
T s,τ
ε =(cid:18)s −
ε =(cid:18)s −
U s,τ
t
2(cid:19)(cid:0)n+(ε)vε + 2T +
ε,1(cid:1) +
2(cid:19)(cid:18)−
ε,0(cid:19) +
vε + T +
n+(ε)
2
t
t
2(cid:0)n−(ε)vε + 2T −
2(cid:18)−
n−(ε)
2
t
ε,1(cid:1) − θ (iη(ε) + 2Tε,3) ,
η(ε)
ε,0(cid:19) − θ(cid:18)−i
2
vε + T −
(5.68)
(5.69)
(5.70)
+ Tε,2(cid:19) .
(5.71)
By construction, T s,τ
ε
and U s,τ
ε
are of the required form and satisfy (5.53).
The proof of (5.54) is essentially identical to that of Theorem 3.19(2); this is due to the
similarity of magic formulas (3.7) and (5.85) for SO(N, R) and Sp(N ).
ε
ε
s,τ
Theorem 5.6 (Intertwining formula for ASp(N,C)
{T s,τ
be as in Theorem 3.20. Define first and second order differential operators
: ε ∈ E }, {U s,τ
). Fix s ∈ R and τ = t + iθ ∈ C. Let
: ε ∈ E }, and nSs,τ
ε,δ : ε, δ ∈ Eo be as in Theorem 5.5, and eLs,τ
2Xε∈E
2Xε∈E
Cs,τ
1 =
Cs,τ
0 =
∂
∂vε
∂
∂vε
(5.72)
(5.73)
T s,τ
ε
U s,τ
1
1
2
ε
Then for all N ∈ N and P ∈ W ,
1
2ASp(N,C)
s,τ
PN =(cid:20)(cid:18)Cs,τ
0 +
1
N Cs,τ
1 +
49
2 (cid:19) P(cid:21)N
1
N 2 eLs,τ
.
(5.74)
Proof. For each X ∈ sp(N ), we apply the chain rule to get
eX 2PN =Xε∈E eX(cid:20)(cid:18) ∂P
=Xε∈E(cid:18) ∂P
PN =Xε∈E(cid:18) ∂P
∂vε(cid:19) (V) · (eXVε)(cid:21)
∂vε(cid:19) (V) · (eX 2Vε) + Xε,δ∈E(cid:18) ∂2P
∂vε(cid:19) (V) · ASp(N,C)
∂vε∂vδ(cid:19) (V)
+ Xε,δ∈E(cid:18) ∂2P
dSO(N,R)Xℓ=1 (cid:20)(cid:18)s −
Vε
s,τ
Hence
ASp(N,C)
s,τ
The result now follows from Theorem 5.5.
+
∂vε∂vδ(cid:19) (V) · (eXVε)(eXVδ).
t
2(cid:19) (eXℓVε)(eXℓVδ)
(eYℓVε)(eYℓVδ) − θ(eXℓVε)(eYℓVδ)(cid:21).
t
2
5.5. Limit theorems for the Segal-Bargmann transform on Sp(N )
In this section, we outline the proof of Theorem 1.6 for Sp(N ). Since the techniques used
are similar to those used in Section 4 to prove the SO(N, R) case, we do not provide the
full details.
The following two lemmas are the analogues of Lemmas 4.5, 4.6, and 4.7 for Sp(N ). The
proofs are very similar, and so are not included: the key concentration result is again
Lemma 4.2, and the only change required is to replace intertwining formulas (1.11) and
(3.82) for SO(N, R) with intertwining formulas (1.11) and (5.74) for Sp(N ) wherever
applicable.
Lemma 5.7 (L´evy, [13]). For s > 0 and k ∈ Z,
tr(Ak) ρSp(N )
(A) dA = νk (s) .
(5.75)
lim
s
Lemma 5.8. Let s > 0 and τ = t + iθ ∈ D(s, s). For any Q ∈ C[v],
N→∞ZSp(N )
(cid:16)eCs,τ
0
ι(Q)(cid:17) (1) = πs−τ Q.
As in the proof of Theorem 1.6 for SO(N, R), we first require the following analogue of
Theorem 4.10.
Theorem 5.9. Let s > 0 and τ = t + iθ ∈ D(s, s). For any P ∈ C[u, u−1; v],
L2(ρSp(N )
s
)
kPN − [πsP ]Nk2
kPN − [πs−τ P ]Nk2
L2(µSp(N,C)
s,τ
50
and
N 2(cid:19) ,
= O(cid:18) 1
N 2(cid:19) .
= O(cid:18) 1
)
(5.76)
(5.77)
(5.78)
Proof. The proof is entirely analogous to the proof of Theorem 4.10. Again, we replace
with intertwining formula (5.74) for ASp(N,C)
intertwining formula (3.82) for ASO(N,C)
wherever necessary. To show the Sp(N ) version of (4.34) (where eLs,τ
is replaced with
Cs,τ
0 ), we use Lemma 5.8 in place of Lemma 4.7. The remainder of the proof is the
same.
s,τ
s,τ
0
Proof of Theorem 1.5 for Sp(N ). The proof of the limit results (1.15) and (1.16) is very
similar to the SO(N, R) case (see p. 35), where we now use the polynomial S(N ) =
2 L0 f ∈ Cm[u, u−1; v] in place of R(N ) and intertwining formula (5.74) in
e
place of (3.82) in (4.38).
N f − e
2 D(4)
τ
τ
To show uniqueness, we define seminorms on C[u, u−1; v] by
kPk(4)
kPk(4)
s,N = kPNkL2(ρSp(N )
N→∞kPk(4)
s = lim
s,N ,
s
),
kPk(4)
kPk(4)
s,τ,N = kPNkL2(µSp(N,C)
s,τ = lim
N→∞kPk(4)
s,τ,N .
s,τ
Noting that
kPk(4)
s = kPk(2)
s ,
),
(5.79)
(5.80)
(5.81)
the remainder of the proof proceeds as in the proof of uniqueness in Theorem 1.5 for
SO(N, R).
We conclude this section by showing a version of the magic formulas and intertwining
5.6. Extending the magic formulas and intertwining formulas to fSp(N )
results for Sp(N ) hold for fSp(N ) ⊆ MN (H). We have the following key relation:
(5.82)
which can be seen by the definition of ψ (see (5.5)).
etr(Φ(A)) = Re tr(A), A ∈ MN (H),
Proposition 5.10 (Magic formulas forfSp(N )). Let β esp(N ) be any orthonormal basis for
esp(N ) with respect to the inner product (5.3). For any A, B ∈ MN (H),
2N + 1
(5.83)
X 2 = −
2N
IN = −IN −
1
2N
IN
XX∈β fsp(N )
XX∈β fsp(N )
XX∈β fsp(N )
XX∈β fsp(N )
Re tr(XA)X =
Re tr(XA) Re tr(XB) =
51
XAX = −
1
2N
A∗ − Re tr(A)IN
1
4N 2 (A∗ − A)
1
4N 2 (Re tr(A∗B) − Re tr(AB))
(5.84)
(5.85)
(5.86)
Proof. We begin with the observation that for A ∈ MN (H),
ΩΦ(A)⊺Ω−1 = Φ(A∗).
(5.87)
To see why this is the case, note that Lemma 5.1 directly implies that for q ∈ H,
Ω0ψ(q)⊺Ω−1
0 = ψ(q∗).
Hence if A = [qi,j],
[ΩΦ(A)⊺Ω−1]i,j = X1≤k,ℓ≤N
[Ω]i,k[Φ(A)⊺]k,ℓ[Ω]ℓ,j = [Ω]i,i[Φ(A)⊺]i,j [Ω]j,j
= Ω0ψ(qj,i)⊺Ω−1
0 = ψ(q∗
j,i) = [Φ(A∗)]i,j ,
which shows (5.87).
Recall that if β esp(N ) is an orthonormal basis for esp(N ), Φ(β esp(N )) is an orthonormal basis
for sp(N ). Let A ∈ MN (H). Using (5.14) and (5.82), we have
Φ(X)Φ(A)Φ(X)
XX∈β fsp(N )
Φ(XAX) = XX∈β fsp(N )
1
2N
1
2N
= −
= −
= Φ(cid:18)−
ΩΦ(A)⊺Ω−1 −etr(Φ(A))I2N
Φ(A∗) − Re tr(A)Φ(IN )
A∗ − Re tr(A)IN(cid:19) .
1
2N
Since Φ is injective, (5.84) follows. The proof of (5.85) is similar, and (5.83) and (5.86)
follow from (5.84) and (5.85), resp.
defined by
AnfSp(N ) version of Theorem 1.4 also holds for an appropriately defined trace polynomial
functional calculus: for P ∈ C[u, u−1; v], we let ePN : fSp(N ) → MN (H) be the function
Proposition 5.11. Let P ∈ C[u, u−1; v] and A ∈fSp(N ) ⊆ MN (H). Then
ePN (A) := P (u; v)u=A,vk=Re tr(Ak),k6=0.
(5.89)
(5.88)
Proof. It suffices to prove the result for P (u; v) = umvn
Using (5.82), we have
k , with m, n, k ∈ Z and k 6= 0.
PN (Φ(A)) = Φ(ePN (A)).
PN (Φ(A)) = Φ(A)metr(Φ(A)k)n = Φ(Am)(Re tr(Ak))n = Φ(Am(Re tr(Ak))n) = Φ(ePN (A)).
Using this proposition, we see that the intertwining formula for Sp(N ) ⊆ MN (H) is a
direct consequence of the intertwining formula (1.11) for Sp(N ) ⊆ M2N (C).
52
Theorem 5.12 (Intertwining formulas). For any P ∈ C[u, u−1; v],
Moreover, for all τ ∈ C,
e
.
N PiN
∆fSp(N )ePN =hD(4)
2 ∆ fSp(N )ePN = [e
2 D(4)
τ
τ
N P ]N .
(5.90)
(5.91)
(5.92)
Remark 5.13. We have seen that the magic formulas (5.10) for fSp(N ) can be derived
from the magic formulas (5.3) for Sp(N ). However, the converse is not true. Suppose we
only know that the magic formula (5.84) holds. Applying Φ to both sides yields
XX∈β fsp(N )
Φ(X)Φ(A)Φ(X) = −
1
2N
Φ(A)∗ −etr(Φ(A))I2N .
We cannot replace Φ(A) with an arbitrary B ∈ M2N (C) in the formula above. For
example, consider
Then we can compute that
XY ∈βsp(1)
Y BY =
1
2(cid:20) −3
1
0
B =(cid:20) 1
1 −3 (cid:21) 6=
1 (cid:21) ∈ Sp(1, C).
2(cid:20) −3 −1
1
0
0 −3 (cid:21) = −
1
2
B∗ −etr(B)I2.
This is another reason why it is more convenient to work with Sp(N ) rather thanfSp(N ).
6. The Segal-Bargmann transform on SU(N )
In this final section, we analyze the Segal-Bargmann transform on the special unitary
group SU(N ). This case uses many of the techniques from the U(N ) case, studied in [8].
Consequently, we outline the main ideas required for the proofs of these results and do
not provide the full details.
We first prove the intertwining formulas for SU(N ). As in the SO(N, R) and Sp(N )
cases, the basis of these results is a set of magic formulas. We recall the following set of
magic formulas for U(N ), proven in [8].
Proposition 6.1 ([8, Proposition 3.1]). Let βu(N ) be any orthonormal basis for u(N )
with respect to the inner product defined by hX, Y iu(N ) = N 2 tr(XY ∗). For any A, B ∈
53
MN (C),
X 2 = −IN
XAX = − tr(A)IN
XX∈βu(N )
XX∈βu(N )
XX∈βu(N )
tr(XA)X = −
tr(XA) tr(XB) = −
1
N 2 A
1
N 2 tr(AB)
XX∈βu(N )
The magic formulas for SU(N ) follow easily from the magic formulas for U(N ).
Proposition 6.2 (Magic formulas for SU(N )). Let βsu(N ) be any orthonormal basis for
su(N ) with respect to the inner product (2.9). For any A, B ∈ MN (C),
(6.1)
(6.2)
(6.3)
(6.4)
(6.5)
(6.6)
(6.7)
(6.8)
X 2 =(cid:18)−1 +
1
N 2(cid:19) IN
XAX = − tr(A)IN +
1
N 2 A
XX∈βsu(N )
XX∈βsu(N )
XX∈βsu(N )
tr(XA)X = −
tr(XA) tr(XB) = −
1
N 2 A +
1
N 2 tr(A)IN
1
N 2 tr(AB) +
1
N 2 tr(A) tr(B)
XX∈βsu(N )
Proof. The expressions on the left side of (6.5), (6.6), (6.7), and (6.8) are independent
of orthonormal basis chosen. Fix an orthonormal basis βsu(N ) for su(N ). Observe that
the matrix i
N IN is an element of u(N ) of unit norm such that
(cid:28) i
N
IN , X(cid:29)u(N )
= 0
for all X ∈ su(N ) ⊆ u(N ).
Hence βu(N ) = βsu(N ) ∪ { i
XX∈βsu(N )
X 2 = XX∈βu(N )
N IN} is an orthonormal basis for u(N ). Thus
X 2 −(cid:18) i
N
IN(cid:19)2
=(cid:18)−1 +
1
N 2(cid:19) IN ,
which proves (6.5). The remaining magic formulas are proven similarly.
Proposition 6.3 (Derivative formulas for SU(N )). The following identities hold on
54
SU(N ) and SL(N, C):
∆SU(N )Am = 21m≤−2
∆SU(N )Am = −21m≥2
m−1Xj=1
−1Xj=m+1
XX∈βsu(N ) eX tr(Am) · eXAp =
jAj tr(Am−j) − mAm +
jAj tr(Am−j ) + mAm +
m2
N 2 Am, m ≥ 0
m2
N 2 Am, m < 0
(6.9)
(6.10)
(6.11)
mp
N 2 (−Am+p + Ap tr(Am)), m, p ∈ Z.
Proof. The proof of these results is essentially identical to the proofs of the corresponding
derivative formulas for U(N ) in [8, Theorem 3.3]; one need only to replace the magic
formulas for U(N ) by the magic formulas for SU(N ) in the proof, wherever applicable.
The derivative formulas of Proposition 6.3 comprise the key ingredient for the intertwin-
ing formulas for ∆SU(N ) and BSU(N )
(Theorems 1.4 and 1.5). The proofs of these results
are entirely analogous to the corresponding results for U(N ) (cf. [8, Theorems 1.8, 1.9]).
The only change required in the proof is to replace the magic formulas and derivative
formulas for U(N ) by the corresponding formulas for SU(N ) from Propositions 6.2 and
6.3. By keeping track of these changes, we see that for P ∈ C[u, u−1; v], the intertwining
formula for ∆SU(N ) is
s,τ
∆SU(N )PN = [D(2)
N P ]N =(cid:20)(cid:18)L0 −
1
N 2 (2K−
1 + K−
2 − J )(cid:19) P(cid:21)N
.
The only difference between this and (3.44), the intertwining formula for ∆U(N ), is that
for SU(N ), the 1/N 2 term on the right hand side contains the additional operator J .
This is a consequence of the close relationship between the magic formulas for U(N ) and
SU(N ).
Finally, our main result regarding the free Segal-Bargmann transform for SU(N ), Theo-
rem 1.6, is proven in the same way as the corresponding result for U(N ) (see [8, Theorem
1.11]); again, the only change necessary is the substitution of the magic and derivative
formulas for SU(N ) in place of those for U(N ) wherever required.
Acknowledgments
The author thanks their PhD advisor Todd Kemp for suggesting the topic of the present
paper, as well as his invaluable advice and extensive comments in completing this work.
In addition, the author thanks the referee, whose suggestions greatly improved the pre-
sentation of the results.
References
[1] Bargmann, V. On a Hilbert space of analytic functions and an associated integral transform.
Comm. Pure Appl. Math. 14 (1961), 187 -- 214.
55
[2] Bargmann, V. Remarks on a Hilbert space of analytic functions. Proc. Nat. Acad. Sci. U.S.A. 48
(1962), 199 -- 204.
[3] Biane, P. Free Brownian motion, free stochastic calculus and random matrices. In Free probability
theory (Waterloo, ON, 1995), vol. 12 of Fields Inst. Commun. Amer. Math. Soc., Providence, RI,
1997, pp. 1 -- 19.
[4] Biane, P. Segal-Bargmann transform, functional calculus on matrix spaces and the theory of
semi-circular and circular systems. J. Funct. Anal. 144, 1 (1997), 232 -- 286.
[5] C´ebron, G. Free convolution operators and free Hall transform. J. Funct. Anal. 265, 11 (2013),
2645 -- 2708.
[6] Driver, B. K. On the Kakutani-Ito-Segal-Gross and Segal-Bargmann-Hall isomorphisms. J. Funct.
Anal. 133, 1 (1995), 69 -- 128.
[7] Driver, B. K., Gross, L., and Saloff-Coste, L. Holomorphic functions and subelliptic heat
kernels over Lie groups. J. Eur. Math. Soc. (JEMS) 11, 5 (2009), 941 -- 978.
[8] Driver, B. K., Hall, B. C., and Kemp, T. The large-N limit of the Segal-Bargmann transform
on UN . J. Funct. Anal. 265, 11 (2013), 2585 -- 2644.
[9] Driver, B. K., Hall, B. C., and Kemp, T. The complex-time Segal -- Bargmann transform. J.
Funct. Anal. 278, 1 (2020).
[10] Hall, B. C. The Segal-Bargmann "coherent state" transform for compact Lie groups. Journal of
Functional Analysis 122, 1 (1994), 103 -- 151.
[11] Ho, C.-W. The two-parameter free unitary Segal-Bargmann transform and its Biane-Gross-
Malliavin identification. J. Funct. Anal. 271, 12 (2016), 3765 -- 3817.
[12] Horn, R. A., and Johnson, C. R. Topics in matrix analysis. Cambridge University Press,
Cambridge, 1994. Corrected reprint of the 1991 original.
[13] L´evy, T. The master field on the plane. Ast´erisque, 388 (2017).
[14] Milnor, J. Curvatures of left invariant metrics on Lie groups. Advances in Math. 21, 3 (1976),
293 -- 329.
[15] Segal, I. E. Mathematical characterization of the physical vacuum for a linear Bose-Einstein field.
(Foundations of the dynamics of infinite systems. III). Illinois J. Math. 6 (1962), 500 -- 523.
[16] Segal, I. E. Mathematical problems of relativistic physics, vol. 1960 of With an appendix by George
W. Mackey. Lectures in Applied Mathematics (proceedings of the Summer Seminar, Boulder, Col-
orado. American Mathematical Society, Providence, R.I., 1963.
[17] Segal, I. E. The complex-wave representation of the free boson field.
In Topics in functional
analysis (essays dedicated to M. G. Kreın on the occasion of his 70th birthday), vol. 3 of Adv. in
Math. Suppl. Stud. Academic Press, New York-London, 1978, pp. 321 -- 343.
[18] Stillwell, J. Naive Lie Theory. Undergraduate Texts in Mathematics. Springer New York, 2008.
56
|
1707.04663 | 1 | 1707 | 2017-07-14T23:30:13 | Mixing inequalities in Riesz spaces | [
"math.FA",
"math.CA",
"math.PR"
] | Various topics in stochastic processes have been considered in the abstract setting of Riesz spaces, for example martingales, martingale convergence, ergodic theory, AMARTS, Markov processes and mixingales. Here we continue the relaxation of conditional independence begun in the study of mixingales and study mixing processes. The two mixing coefficients which will be considered are the $\alpha$ (strong) and $\varphi$ (uniform) mixing coefficients. We conclude with mixing inequalities for these types of processes. In order to facilitate this development, the study of generalized $L^1$ and $L^\infty$ spaces begun by Kuo, Labuschagne and Watson will be extended. | math.FA | math |
Mixing inequalities in Riesz spaces ∗
Wen-Chi Kuo†
School of Computational and Applied Mathematics
University of the Witwatersrand
Private Bag 3, P O WITS 2050, South Africa
Michael J. Rogans
School of Statistics and Actuarial Science
University of the Witwatersrand
Private Bag 3, P O WITS 2050, South Africa
Bruce A. Watson ‡
School of Mathematics
University of the Witwatersrand
Private Bag 3, P O WITS 2050, South Africa
June 18, 2021
Abstract
Various topics in stochastic processes have been considered in the abstract setting of
Riesz spaces, for example martingales, martingale convergence, ergodic theory, AMARTS,
Markov processes and mixingales. Here we continue the relaxation of conditional inde-
pendence begun in the study of mixingales and study mixing processes. The two mixing
coefficients which will be considered are the α (strong) and ϕ (uniform) mixing coeffi-
cients. We conclude with mixing inequalities for these types of processes. In order to
facilitate this development, the study of generalized L1 and L∞ spaces begun by Kuo,
Labuschagne and Watson will be extended.
∗Keywords: Riesz spaces, vector lattices; mixing processes, dependent processes, conditional expectation
operators. Mathematics subject classification (2000): 47B60, 60G20.
†Supported in part by NRF grant number CSUR160503163733.
‡Supported in part by the Centre for Applicable Analysis and Number Theory and by NRF grant number
IFR2011032400120.
1
1
Introduction
Mixing processes are stochastic processes in which independence assumptions are re-
placed by a measure of independence called the mixing coefficient, see [1, 3, 6, 16] for
measure theoretic essentials of mixing processes. In the Riesz space (measure free) set-
ting, processes which require independence, such as Markov processes, were considered
with independence replaced by conditional independence, see [21].
In line with the
above approach, mixingales (processes with independence/conditional independence in
the limit) were considered in the Riesz space setting in [14]. In this work we will pose
α (strong) and ϕ (uniform) mixing processes in the Riesz space setting. Core to the
theory of mixing is the family of inequalities generally referred to as the mixing inequal-
ity, the conditional Riesz space analogues of which form the focus of this paper, see
Section 4. We will give mixing inequalities for both α and ϕ mixing (the one being an
easy consequence of the other). We refer the reader to [1, 6, 10, 16, 18] for the measure,
non-conditional analogues. To facilitate this study the Riesz space analogues of the Lp
spaces introduced in [15] will be revisited.
In [12], it was shown that a conditional expectation operator, T , on a Riesz space, E,
admits a unique maximal extension to a conditional expectation operator, also denoted
T , in the universal completion, Eu, of E, with domain a Dedekind complete Riesz space,
which will be denoted L1(T ). The procedure used there was based on that of de Pagter
and Grobler, [8], for the measure theoretic setting. We observe here that the range of
the maximal extension of the conditional expectation operator, i.e. R(T ) := {T f f ∈
L1(T )}, is a universally complete f -algebra and that L1(T ) is an R(T )-module. This
prompts the definition of an R(T ) (vector valued) norm k · kT,1 := T · on L1(T ). Here
the homogeneity is with respect to multiplication by elements of R(T )+. Following in a
similar manner L∞(T ) is taken to be the subspace of L1(T ) composed of R(T ) bounded
elements. An R(T ) valued norm k · kT,∞ := inf{g ∈ R(T )+ · ≤ g}, is defined on
L∞(T ). This extends on the concepts of L∞(T ) defined in [15].
In [4, Sections 39, 42 and 43] Dellacherie and Meyer gave an extension of martingale
theory to σ-finite processes. As a direct application of the material presented in Sections
2 to 4, we give, in Section 5, an extension of the theory of mixing theory to σ-finite
processes. The extension of mixing theory even to this special case, to the knowledge of
the authors, has not been considered in the literature.
Natural connections with the theory presented here are to laws of large numbers and
other convergence theorems. For the conditionally independent case in Riesz spaces we
refer the reader to Stoica [19].
2
2 Preliminaries
For general background on Riesz spaces we refer the reader to [23] and for the foundations
of stochastic processes in Riesz spaces to [11, 12, 13, 14, 15]. We will present only the
essential results from the theory of Riesz space stochastic processes required for our
consideration of mixing processes.
Definition 2.1 Let E be a Dedekind complete Riesz space with weak order unit. A
positive order continuous linear projection T on E with range R(T ), a Dedekind complete
Riesz subspace of E, is said to be a conditional expectation operator if T e is a weak order
unit of E for each weak order unit e of E.
A Riesz space E is said to be universally complete if E is Dedekind complete and every
subset of E which consists of mutually disjoint elements has a supremum in E. A Riesz
space Eu is said to be a universal completion of the Riesz space E if Eu is universally
complete and contains E as an order dense subspace. If e is a weak order unit of E then
e is also a weak order unit of Eu, see [23].
We say that a conditional expectation operator, T , on a Riesz space is strictly positive if
T f = 0 implies that f = 0. As shown in [12], a strictly positive conditional expectation
operator, T , on a Riesz space can be extended to its so called natural domain (maximal
domain to which it can be extended as a conditional expectation operator) denoted
dom(T ). We set L1(T ) := dom(T ) and we denote the extension of T to L1(T ) again by
T . This is consistent with the special case of T an expectation operator in the measure
theoretic setting, see [8]. In particular, if E is a Dedekind complete Riesz space with
weak order unit and conditional expectation operator, T , we say that E is T -universally
complete if E = L1(T ). From the definition of dom(T ), E is T -universally complete if
and only if for each upwards directed net (fα)α∈Λ in E+ such that (T fα)α∈Λ is order
bounded in Eu, we have that (fα)α∈Λ is order convergent in E.
If e is a weak order unit of E then we denote the f -algebra of e bounded elements by
Ee := {f ∈ E f ≤ ke,
for some k ∈ R+}
and set
L∞(T ) := {f ∈ L1(T ) f ≤ g, for some g ∈ R(T )+}.
We recall from [12, Theorem 5.3] that each conditional expectation operator T is an
averaging operator in the sense that if f ∈ R(T ) and g ∈ E with f g ∈ E then T (f g) =
f T (g).
Theorem 2.2 Let E be a T -universally complete Riesz space, where T is a conditional
expectation operator on E, and let e be a weak order unit for E with T e = e. Then R(T )
3
is universally complete and hence an f -algebra. In addition E = L1(T ) and L∞(T ) are
R(T )-modules.
In order to show that R(T ) is universally complete, we need to show that for
Proof:
each W ⊂ R(T )+ consisting of mutually disjoint elements, i.e.
if u, v ∈ W with u 6= v
then u ∧ v = 0, we have that w := ∨v∈W v ∈ R(T ). To this end, let W be as above and
set w := ∨v∈W v, which exists in Eu. Now u ∧ ne ∈ R(T )+ and u ∧ ne ≤ ne, for all
n ∈ N, u ∈ W . Here {u ∧ ne u ∈ W } ⊂ R(T ) is bounded above by ne ∈ R(T ) and so
the Dedekind completeness of R(T ) allows us to conclude that
Now w ∧ ne ↑ w and
w ∧ ne = ∨u∈W (u ∧ ne) ∈ R(T ).
T (w ∧ ne) = w ∧ ne ≤ w ∈ Eu, n ∈ N,
so the T -universal completeness of E gives that w ∈ E. The order continuity of T gives
T w = w and w ∈ R(T ). Thus R(T )u = R(T ) from which it follows that R(T ) is an
f -algebra.
Since f g = f +g+ + f −g− − f −g+ − f +g−, to show that L1(T ) is an R(T )-module, it
suffices to show that f g ∈ L1(T ) for each f ∈ L1(T )+ and g ∈ R(T )+. Now from the
averaging properties of T and as (f ∧ ne)g ∈ L1(T ) we have that
T ((f ∧ ne)g) = T (f ∧ ne)g ≤ T (f )g ∈ Eu
+
for all n ∈ N. However L1(T ) is T-universally complete and (f ∧ne)g ↑ f g, so f g ∈ L1(T ).
Finally, to show that L∞(T ) is an R(T )-module we, take f ∈ L∞(T )+ and g ∈ R(T )+.
Here there is F ∈ R(T ) with 0 ≤ f ≤ F . Now
as R(T ) is an algebra. Hence f g ∈ L∞(T ).
f g ≤ F g ∈ R(T )+
We note the connection with the work of Grobler and de Pagter in [8], where in the
measure theoretic case, it is shown that the range space of the maximal extension of a
classical conditional expectation operator is an algebra.
3 The Lp(T ), p = 1, ∞, spaces with T -norms
In the previous section we define the Lp(T ), p = 1, ∞, spaces, see also [12, 15]. Here we
present the corresponding vector valued generalizations of the Lp-norms. In particular,
4
these norms take their values in the positive cone of the universally complete algebra
R(T ). We also refer the reader to the Riesz semi-norm approach used by Grobler and
Labuschagne, [9], to study the space L2(T ). For some recent progress in this area we
refer the reader to [2].
Definition 3.1 Let E be a Dedekind complete Riesz space with weak order unit and T
be a strictly positive conditional expectation operator on E. If E is an R(T )-module and
φ : E → R(T )+ with
(a) φ(f ) = 0 if and only if f = 0,
(b) φ(gf ) = gφ(f ) for all f ∈ E and g ∈ R(T ),
(c) φ(f + h) ≤ φ(f ) + φ(h) for all f, h ∈ E,
then φ will be called an R(T )-valued norm on E.
Theorem 3.2 Let E be a T -universally complete Riesz space with weak order unit,
where T is a strictly positive conditional expectation operator on E. The map
defines an R(T )-valued norm on L1(T ), and the map
f 7→ T f =: kf kT,1
f 7→ kf kT,∞ = inf {g ∈ R(T )+ f ≤ g}
defines an R(T )-valued norm on L∞(T ).
Proof: For L1(T ), condition (a) of Definition 3.1 follows from the strict positivity of T
and while (c) of Definition 3.1 follows from the linearity of T and · obeying the triangle
inequality. For (b) of Definition 3.1, let f ∈ E and g ∈ R(T ), then we observe that the
terms f +g+, f −g+, f +g−, f −g− are disjoint and positive. Thus
gf = f +g+ − f −g+ − f +g− + f −g− = f +g+ + f −g+ + f +g− + f −g− = gf .
Here g ∈ R(T ), so T gf = T (gf ) = gT f , from which Definition 3.1 part (b)
follows.
We now consider the case of L∞(T ). Here (a) and (b) follow directly from the definition
of k · kT,∞. For (c), consider f, g ∈ L∞(T ). As f + g ≤ f + g, it follows that
{h ∈ R(T )+ f + g ≤ h} ⊂ {h ∈ R(T )+ f + g ≤ h}.
Therefore
kf + gkT,∞ = inf{h ∈ R(T )+ f + g ≤ h} ≤ inf{h ∈ R(T )+ f + g ≤ h}.
5
Writing h = hf + hg, for hf , hg ∈ R(T )+, it follows that
{hf + hg f ≤ hf , g ≤ hg, hf , hg ∈ R(T )+} ⊂ {h ∈ R(T )+ f + g ≤ h},
giving
inf{hf + hg f ≤ hf , g ≤ hg, hf , hg ∈ R(T )+} ≥ inf{h ∈ R(T )+ f + g ≤ h}.
Combining the above and noting that the conditions f ≤ hf and g ≤ hg are indepen-
dent, we have
kf + gkT,∞ ≤ inf{hf + hg f ≤ hf , g ≤ hg, hf , hg ∈ R(T )+}
= inf{hf f ≤ hf , hf ∈ R(T )+} + inf{hg g ≤ hg, hg ∈ R(T )+}
= kf kT,∞ + kgkT,∞.
Theorem 3.3 If f ∈ L1(T ) and g ∈ L∞(T ), then gf ∈ L1(T ) and
kgf kT,1 ≤ kgkT,∞kf kT,1.
Proof: From linearity, it suffices to show that gf ∈ L1(T ) for f ∈ L1(T )+ and g ∈
L∞(T )+. Here there exists h ∈ R(T )+ such that g ≤ h, and we note from earlier that
hf ∈ L1(T )+, but 0 ≤ gf ≤ hf , giving gf ∈ L1(T )+.
Now for each h ∈ R(T )+ with g ≤ h we have from T being an averaging operator that
kgf kT,1 = T gf = T (gf ) ≤ T (hf ) = hT f = hkf kT,1.
However, from the compatability of the multiplicative structure with the order structure,
inf{hkf kT,1 g ≤ h, h ∈ R(T )+} = inf{h ∈ R(T )+ g ≤ h}kf kT,1 = kf kT,1kgkT,∞,
from which the result follows.
Setting f = e, where e is a weak order unit with T e = e and which has been chosen as the
algebraic unit for the f -algebra structure, we have the following corollary to Theorem
3.3.
Corollary 3.4 If g ∈ L∞(T ) then kgkT,1 ≤ kgkT,∞.
To conclude this section, we give a variant of the conditional Jensen's inequality. For
additional details on conditional Jensen's inequalities in Riesz spaces, see [7].
6
Theorem 3.5 If S is a conditional expectation operator on L1(T ) compatible T (in the
sense that T S = T = ST ), then
for all f ∈ Lp(T ), p = 1, ∞.
kSf kT, p ≤ kf kT,p,
Proof: For p = 1, as S is a positive operator,
kSf kT,1 = T Sf ≤ T Sf = T f = kf kT,1.
For p = ∞, if f ≤ g ∈ R(T )+, then g = T g and from the positivity of S we have
Sf ≤ Sf so
Sf ≤ Sf ≤ Sg = ST g = T g = g.
Thus
{g ∈ R(T )+ f ≤ g} ⊂ {g ∈ R(T )+ Sf ≤ g},
from which it follows that kSf kT,∞ ≤ kf kT,∞.
4 Mixing inequalities
In this section we consider conditional versions of strong mixing (α-mixing) and uniform
mixing (ϕ-mixing) in the Riesz space setting. At the core of mixing processes is the
family of inequalities generally termed the mixing inequalities, see [1, 6, 10, 16, 18] for
the classical mixing inequalities. We begin by giving conditional definitions of strong and
uniform mixing in the measure theoretic setting. These conditional definitions of mixing
admit direct generalizations to Riesz spaces with conditional expectation operators. We
conclude with a conditional mixing inequality for conditionally strong mixing processes
in Riesz spaces. This yields, directly, conditional mixing inequalities for Riesz space con-
ditionally uniform mixing processes and conditional mixing inequalities for conditionally
strong and conditionally uniform mixing processes in the measure space setting.
In the classical measure theoretic setting, the strong mixing coefficient is defined as
follows. Let (Ω, F , µ) be a probability space and A and B be sub-σ-algebras of F. The
strong mixing coefficient between A and B is
α(A, B) = sup{µ(A ∩ B) − µ(A)µ(B) A ∈ A, B ∈ B}.
(4.1)
In place of the expectation, we could condition on a sub-σ-algebra, say C, of A ∩ B which
would result in the C-conditioned strong mixing coefficient
αC(A, B) = sup{E[IAIBC] − E[IAC]E[IBC] A ∈ A, B ∈ B].
(4.2)
7
Definition 4.1 Let E be a Dedekind complete Riesz space with weak order unit, say e,
and conditional expectation operator, T , with T e = e. If U is a conditional expectation
operators on E, with T U = T = U T , then we say that U is compatible with T . If U is a
conditional expectation on E compatible with T then we denote by B(U ) the set of band
projections P on E with P e ∈ R(U ).
In light of (4.2), we define the strong mixing coefficient in a Riesz space with conditional
expectation operator as follows.
Definition 4.2 Let E be a Dedekind complete Riesz space with weak order unit, say
e, and conditional expectation operator, T , with T e = e. We define the T -conditional
strong mixing coefficient with respect to the conditional expectation operators U and V
on E compatible with T , by
αT (U, V ) := sup{T P Qe − T P e · T Qe P ∈ B(U ), Q ∈ B(V )}.
We can now give bounds for the T -conditional strong mixing coefficient in terms of the
T -conditional norm.
Theorem 4.3 Let E be a T -universally complete Riesz space, E = L1(T ), where T is a
conditional expectation operator on E where E has a weak order unit, say e, with T e = e.
Let U and V be conditional expectation operators on E compatible with T , then
αT (U, V ) ≤ sup
kU Qe − T QekT,1 ≤ 2αT (U, V ).
Q∈B(V )
Proof: Let P ∈ B(U ) and Q ∈ B(V ) then as T is an averaging operator T P e · T Qe =
T (P e · T Qe). The f -algebra structure gives that P e · Qe = P Qe and hence
T (P e · Qe) − T P e · T Qe = T [P e · (Qe − T Qe)] = T [P (Qe − T Qe)].
From the Ando-Douglas-Radon-Nikod`ym Theorem, see [22], it follows that
T [P (Qe − T Qe)] = T [P U (Qe − T Qe)]
(4.3)
which is maximized, over P ∈ B(U ), when P = P+ is the band projection onto the band
generated by [U (Qe − T Qe)]+, in which case
T P+U (Qe − T Qe) = T [U (Qe − T Qe)]+.
(4.4)
Hence
sup{T P Qe − T P e · T Qe P ∈ B(U )} = T [U (Qe − T Qe)]+.
8
We note that (4.3) is minimized, over P ∈ B(U ), when P = P− is the band projection
onto the band generated by [U (Qe − T Qe)]−, in which case
T P−U (Qe − T Qe) = −T [U (Qe − T Qe)]−.
(4.5)
Hence
Now
sup{−T P Qe + T P e · T Qe P ∈ B(U )} = T [U (Qe − T Qe)]−.
sup{T P Qe − T P e · T Qe P ∈ B(U )}
= sup{(T P Qe − T P e · T Qe) ∨ (−T P Qe + T P e · T Qe) P ∈ B(U )}
≤ sup{(T P1Qe − T P1e · T Qe) ∨ (−T P2Qe + T P2e · T Qe) P1, P2 ∈ B(U )}
= sup{(T P Qe − T P e · T Qe) P ∈ B(U )} ∨ sup{(−T P Qe + T P e · T Qe) P ∈ B(U )}
and applying (4.4) to each of the expressions in the last line of the above gives
sup{T P Qe − T P e · T Qe P ∈ B(U )} = T ([U (Qe − T Qe)]+) ∨ T ([U (Qe − T Qe)]−).
From the linearity of the expectation and conditional expectation operators,
U [(I − Q)e − T (I − Q)e] = −U [Qe − T Qe].
(U [(I − Q)e − T (I − Q)e])+ = [U (Qe − T Qe)]−
Hence
and thus
sup{T P Qe − T P e · T Qe P ∈ B(U )}
≤ T [U (Qe − T Qe)]+ ∨ T (U [(I − Q)e − T (I − Q)e])+.
(4.6)
Since Q ∈ B(V ) implies I − Q ∈ B(V ), the first inequality of the Theorem follows from
taking the supremum of (4.6) over V ∈ B(V ).
Combining (4.3), (4.4) and (4.5), we have
kU Qe − T QekT,1 = T U (Qe − T Qe)
= T (U (Qe − T Qe))+ + T (U (Qe − T Qe))−
= T P+U (Qe − T Qe) − T P−U (Qe − T Qe).
As U is an averaging operator and P−, P+ ∈ B(U ) we have
T P±U (Qe − T Qe) = T U P±(Qe − T Qe) = T P±(Qe − T Qe).
The positivity of T and the definition of αT (U, V ) give
T P±(Qe − T Qe) ≤ T P±(Qe − T Qe) ≤ αT (U, V ).
9
Combining the above gives
kU Qe − T QekT,1 ≤ 2αT (U, V )
which proves the second inequality of the Theorem.
It should be noted here that the product T P e · T Qe exists in E, as shown in the previous
section.
We now consider the uniform mixing coefficient. In the measure theoretic setting the
uniform mixing coefficient is defined as follows. Let (Ω, F , µ) be a probability space and
A and B be sub-σ-algebras of F. The uniform mixing coefficient between A and B is
ϕ(A, B) = sup{µ(BA) − µ(B) A ∈ A, B ∈ B, µ(A) > 0}.
(4.7)
As with the strong mixing coefficient, the uniform mixing coefficient has an interesting
formulation in terms of Lp norms.
Lemma 4.4 Let (Ω, F , µ) be a probability space and A and B be sub-σ-algebras of F,
then
ϕ(A, B) = sup
B∈B
kE[IB − E[IB]A]k∞.
Proof: We begin by observing that, since E[IBA] − E[IB] is A-measurable,
kE[IBA] − E[IB]k∞ = sup(cid:26) E[IA(E[IBA] − E[IB])]
µ(A)
Now for A ∈ A with µ(A) > 0 we have
A ∈ A, µ(A) > 0(cid:27) .
(cid:12)(cid:12)(cid:12)(cid:12)
E[IA(E[IBA] − E[IB])]
µ(A)
=
E[IAIB] − E[IA]E[IB]
µ(A)
= µ(BA) − µ(B),
from which the Lemma follows.
The above Lemma leads naturally to conditional and Riesz space variants of the uniform
mixing coefficient.
Definition 4.5 Let E be a Dedekind complete Riesz space with weak order unit, say
e, T a conditional expectation operator on E and E having weak order unit say e with
T e = e. Let U and V be conditional expectation operators on E compatible with T , then
ϕT (U, V ) = sup
kU Qe − T QekT,∞.
Q∈B(V )
10
Combining Corollary 3.4 with Theorem 4.3 and Definition 4.5 we have the following
theorem.
Theorem 4.6 Let E be a T -universally complete Riesz space, E = L1(T ), where T is a
conditional expectation operator on E where E has a weak order unit, say e, with T e = e.
Let U and V be conditional expectation operators on E compatible with T , then
αT (U, V ) ≤ ϕT (U, V ).
The mixing inequalities now give bounds on the norm of the differences between the
composition of the conditional expectation operators, say U and V , compatible with the
conditional expectation operator T . It should be noted that if U and V are conditionally
independent with respect to T then U V = T = V U . The measure theoretic version was
proved in [17], wherein results from [5] were used.
Theorem 4.7 Let E be a T -universally complete Riesz space, T a conditional expec-
tation operator on E and e a weak order unit for E with e = T e. Let U and V be
conditional expectation operators on E compatible with T , then, for f ∈ R(V ) ∩ L∞(T ),
we have
kU f − T f kT,1 ≤ 4αT (U, V )kf kT,∞.
Proof: Let g := kf kT,∞ ∈ R(T )+, then f is in the order interval [−g, g]. Hence, from
[12, Theorem 4.2], there are sequences (f ±
n ) ⊂ R(V ) with 0 ≤ f ±
n ↑ f ± of the form
N ±
n
i,ng,
f ±
n =
θ±
i,nP ±
Xi=1
i,n have P ±
i,ne ∈ R(V ) and P ±
1,n < θ±
0,n < θ±
2,n < · · · < θ±
where the band projections P ±
real numbers θ±
i,n have 0 =: θ±
n ≤ g and P ±
f ±
i,ng ∈ R(V ). Set
i,nP ±
Nn,n. Also θ±
j,n = 0 for i 6= j, and the
Nn,n ≤ 1 since
then
Q±
i,n =
N ±
n
Xj=i
P ±
i,n,
f ±
n =
N ±
n
Xi=1
i,nQ±
β±
i,ng,
where 0 < β±
i,n := θ±
i,n − θ±
i−1,n. Here
N ±
n
Xi=1
β±
i,n ≤ 1 and Q±
i,ne ∈ R(V ).
11
Now
U Q±
i,ng − T Q±
i,ng = U (g · Q±
i,ne) − T (g · Q±
i,ne).
Since U and T are averaging operators with g ∈ R(T ) ⊂ R(U ) it follows that
U (g · Q±
i,ne) − T (g · Q±
i,ne) = g · (U Q±
i,ne − T Q±
i,ne) = g · U Q±
i,ne − T Q±
i,ne.
(4.8)
Hence
T U Q±
i,ng − T Q±
i,ng = T (g · U Q±
i,ne − T Q±
i,ne) = g · T U Q±
i,ne − T Q±
i,ne.
Theorem 4.3 gives
and hence
T U Q±
i,ne − T Q±
i,ne ≤ 2αT (U, V )
T U Q±
i,ng − T Q±
i,ng ≤ 2αT (U, V ) · g.
Applying the above to f ±
n gives
n − T f ±
n
T U f ±
N ±
n
β±
i,nT U Q±
i,ng − T Q±
i,ng
≤
Xi=1
≤ 2
N ±
n
Xi=1
β±
i,nαT (U, V ) · g ≤ 2αT (U, V ) · g.
Taking the order limit as n → ∞ and using the order continuity of conditional expecta-
tion operators gives
T U f ± − T f ± ≤ 2αT (U, V ) · g.
Finally
T U f − T f ≤ T U f + − T f + + T U f − − T f − ≤ 4αT (U, V ) · g
which can be rewritten as in the statement of the theorem.
Applying Theorem 4.7 to probability spaces we have the following corollary.
Corollary 4.8 Let (Ω, F , µ) be a probability space, C, G, H be sub-σ-algebras of F with
G and H containing C. For f ∈ L0(Ω, H, µ) with f bounded by g ∈ L0(Ω, C, µ) we have
E[E[f G] − E[f C] C] ≤ 4αC(G, H)g.
Setting f = V g in Theorem 4.7 and using Theorem 3.5, we obtain the following corollary.
12
Corollary 4.9 Let E be a T -universally complete Riesz space with weak order unit
e = T e, where T is a conditional expectation operator on E. Let U and V be conditional
expectation operators on E compatible with T . Then for g ∈ L∞(T )
kU V g − T gkT,1 ≤ 4αT (U, V )kgkT,∞.
The next theorem, see [18] for the measure theoretic case, arises using a similar procedure
as the above for theorem but we now proceed from (4.8) and the definition of ϕT (U, V )
as follows:
U Q±
i,ne) ≤ g · ϕT (U, V ).
i,ng − T Q±
i,ng = g · (U Q±
i,ne − T Q±
Here
U f ±
n − T f ±
n ≤
N ±
n
Xi=1
β±
i,nU Q±
i,ng − T Q±
i,ng ≤
N ±
n
Xi=1
β±
i,nϕT (U, V ) · g ≤ ϕT (U, V ) · g.
Letting n → ∞ gives
and thus
U f ± − T f ± ≤ ϕT (U, V ) · g
kU f − T f kT,∞ ≤ 2ϕT (U, V ) · kf kT,∞
from which the following theorem follows.
Theorem 4.10 Let E be a T -universally complete Riesz space with weak order unit
e = T e, where T is a conditional expectation operator on E. Let U and V be conditional
expectation operators on E compatible with T . Then for f ∈ L∞(T ) ∩ R(V )
kU f − T f kT,1 ≤ kU f − T f kT,∞ ≤ 2ϕT (U, V )kf kT,∞.
Applying Theorem 4.10 to probability spaces we have the following corollary.
Corollary 4.11 Let (Ω, F , µ) be a probability space, C, G, H be sub-σ-algebras of F with
G and H containing C. For f ∈ L0(Ω, H, µ) with f bounded by g ∈ L0(Ω, C, µ) we have
E[f G] − E[f C] ≤ 2ϕC(G, H)g.
5 Mixing for σ-finite processes
In this section we consider the simplest non-trivial application, that is, to σ-finite pro-
cesses, hence giving a theory of conditional mixing for such processes. In this concrete
13
example the spaces and operators can be clearly identified. A consideration of σ-finite
processes in the context of martingale theory can be found in the work of Dellacherie
and Meyer, [4, Sections 39, 42 and 43].
Let (Ω, A, µ) be a σ-finite measure space, which to be interesting should have µ(Ω) = ∞,
and let (Ωi)i∈N be a µ-measurable partition of Ω into sets of finite positive measure. Let
A0 be the sub-σ-algebra of A generated by (Ωi)i∈N. We take the Riesz space E =
L∞(Ω, A, µ) and the conditional expectation operator T = E[ · A0]. For f ∈ E we have
T f (ω) = RΩi
f dµ
µ(Ωi)
,
for ω ∈ Ωi.
(5.9)
Here we have that the universal completion, Eu, of E is the space of all A-measurable
functions. The T -universal completion of E is the space
L1(T ) = (cid:26)f ∈ Eu (cid:12)(cid:12)(cid:12)(cid:12)
ZΩi
f dµ < ∞ for all i ∈ N(cid:27) ,
which is characterized by f Ωi ∈ L1(Ω, A, µ), for each i ∈ N. Here T can be extended to
an L1(T ) conditional expectation operator as per (5.9). We note that E has weak order
unit e = 1, the function identically 1 on Ω, which again is a weak order unit for L1(T ),
but is not in L1(Ω, A, µ). The range of the generalized conditional expectation operator
T is
R(T ) = {f ∈ Eu f a.e. constant on Ωi, i ∈ N} ,
which is an f -algebra. The last of the spaces to be considered is
L∞(T ) = {f ∈ Eu f essentially bounded on Ωi for each i ∈ N} .
The vector norms on L1(T ) and L∞(T ) are
kf k1(ω) = T f (ω) = RΩi
f dµ
µ(Ωi)
,
for ω ∈ Ωi, f ∈ L1(T ),
kf k∞(ω) = ess supΩif ,
for ω ∈ Ωi, f ∈ L∞(T ).
(5.10)
(5.11)
We note that L1(Ω, A, µ) ( L1(T ), L∞(Ω, A, µ) ( L∞(T ), L∞(T ) ⊂ L1(T ) while
L∞(Ω, A, µ) 6⊂ L1(Ω, A, µ).
Let C and D be sub-σ-algebras of A which contain A0. The α-mixing coefficient of C
and D conditioned on A0 (which in measure theoretic terms could be denote αA0(C, D)
is αT (U, V ). Here U and V are the restrictions to L1(T ) of the extensions to L1(U ) and
L1(V ) respectively of the conditional expectation operators U and V on E conditioning
with respect to the σ-algebras C and D. In this example case these operators can be
given explicitly by
U (f ) =
V (f ) =
Ei[f IΩiC],
Ei[f IΩiD],
∞
Xi=1
Xi=1
∞
14
(5.12)
(5.13)
for f ∈ L1(T ). Here the conditional expectation Ei[f IΩiC] = Ei[f C] is the conditional
expectation on Ωi of f Ωi with respect to the probability measure µi(A) := µ(A∩Ωi)
µ(Ωi) and
the σ-algebra {C ∩ ΩiC ∈ C}, and similarly for C replaced by D. Since the structure of
the example is extremely simple, an explicit computation can be carried out to give
αT (U, V ) = αA0(C, D)
=
=
∞
Xi=1
Xi=1
∞
IΩi sup {µi(C ∩ D) − µi(C)µi(D) C ∈ C, D ∈ D}
αi(C, D)IΩi ,
where αi(C, D) is the α-mixing coefficient of σ-algebras C and D with respect to the prob-
ability measure µi. Corollary 4.9 gives that if g is µ-measurable and essential bounded
on each Ωi, i ∈ N, then
kU V g − T gkT,1 ≤ 4αT (U, V )kgkT,∞,
which in this example case can be written as
1
µ(Ωi) ZΩi (cid:12)(cid:12)(cid:12)(cid:12)
1
µ(Ωi) ZΩi
g dµ(cid:12)(cid:12)(cid:12)(cid:12)
Ei[Ei[gD]C] −
dµ ≤ 4αi(C, D)ess supΩig
for each i ∈ N. The conditional uniform mixing coefficient is given by
ϕT (U, V ) = ϕA0(C, D)
=
=
∞
Xi=1
Xi=1
∞
IΩi sup
D∈D
ess supΩi Ei[IDC] − µi(D)
ϕi(C, D)IΩi ,
where ϕi(C, D) is the ϕ-mixing coefficient of C and D relative to the probability measure
µi. For g as above, Theorem 4.10 gives
kU V g − T gkT,∞ ≤ 2ϕT (U, V )kgkT,∞,
which in the special case under consideration yields
Ei[Ei[gD]C] −
(cid:12)(cid:12)(cid:12)(cid:12)
a.e. on Ωi, for each i ∈ N.
1
µ(Ωi) ZΩi
g dµ(cid:12)(cid:12)(cid:12)(cid:12)
≤ 2ϕi(C, D) ess sup Ωig
It should be noted that the work presented here also applies to processes where the
random variables are Riesz space valued, say Lp, and the conditional expectation, T , is
generated by an arbitrary sub-σ-algebra of A. In this case we obtain a generalization of
mixing to the context of vector measure.
15
References
[1] P.P. Billingsley, Probability and Measure, John Wiley and Sons, 3rd edition,
1995.
[2] S. Cerreia-Vioglio, M. Kupper, F. Maccheroni, M. Marinacci, N. Vo-
gelpoth, Conditional Lp-spaces and the duality of modules over f-algebras, J.
Math. Anal. Appl., (2016), http://dx.doi.org/10.1016/j.jmaa.2016.06.018.
[3] P. Doukhan, Mixing: properties and examples, Lecture Notes in Statistics, 85
(1994), 15-23.
[4] C. Dellacherie, P.-A. Meyer, Probabilities and Potentials: B, Theory of Mar-
tingales, North Holland Publishing Company, 1982.
[5] A. Dvoretzky, Asymptotic normality for sums of dependent random variables,
Proc. Sixth Berkeley Symp. on Math. Statist. and Prob., 2 (1972), 513-535.
[6] G.A. Edgar, L. Sucheston, Stopping times and directed processes, Cambridge
University Press, 1992.
[7] J.J. Grobler, Jensen's and martingale inequalities in Riesz spaces, Indag. Math.,
25 (2014), 275-295.
[8] J.J. Grobler, B. de Pagter, Operators representable as multiplication-
conditional expectation operators, J. Operator Theory, 48 (2002), 15-40.
[9] J.J. Grobler, C.C.A. Labuschagne, The Ito integral for Brownian motion in
vector lattices: Part 2, J. Math. Anal. Appl., 423 (2015), 820-833.
[10] I.A. Ibragimov, Some limit theorems for stationary processes, Theory Probab.
Appl., 7 (1962), 349-382.
[11] W.-C. Kuo, C.C.A. Labuschagne, B.A. Watson, Discrete-time stochastic pro-
cesses on Riesz spaces, Indag. Math., 15 (2004), 435-451.
[12] W.-C. Kuo, C.C.A. Labuschagne, B.A. Watson, Conditional expectations on
Riesz spaces, J. Math. Anal. Appl., 303 (2005), 509-521.
[13] W.-C. Kuo, C.C.A. Labuschagne, B.A. Watson, Convergence of Riesz space
martingales, Indag. Math., 17 (2006), 271-283.
[14] W.-C. Kuo, J.J. Vardy, B.A. Watson, Mixingales on Riesz spaces, J. Math.
Anal. Appl., 402 (2013), 731-738.
[15] C.C.A. Labuschagne, B.A. Watson, Discrete stochastic integration in Riesz
spaces, Positivity, 14 (2010), 859-875.
[16] Z. Lin, Z. Bai, Probability Inequalities, Science Press Beijing and Springer-Verlag,
Berlin, Heidelberg, 2010.
[17] D.L. McLeish, A maximal inequality and dependent strong laws, Ann. Probab., 3
(1975), 829-839.
[18] R.J. Serfling, Contributions to central limit theory for dependent variables, Ann.
Math. Stat., 39 (1968), 1158-1175.
16
[19] G. Stoica, Limit laws in vector lattices, Quaetiones Math., 38 (2015), 829-833.
[20] V.G. Troitsky, F. Xanthos Spaces of regular abstract martingales, J. Math.
Anal. Appl., (2016), http://dx.doi.org/10.1016/j.jmaa.2015.09.082.
[21] J.J. Vardy, B.A. Watson, Markov processes on Riesz spaces, Positivity, 16
(2012), 373-391.
[22] B.A. Watson, An Ando-Douglas type theorem in Riesz spaces with a conditional
expectation, Positivity, 13 (2009), 543-558.
[23] A.C. Zaanen, Introduction to Operator Theory in Riesz Spaces, Springer Verlag,
1997.
17
|
1206.1613 | 1 | 1206 | 2012-06-07T20:56:27 | Average Number of Lattice Points in a Disk | [
"math.FA"
] | The difference between the number of lattice points in a disk of radius $\sqrt{t}/2\pi$ and the area of the disk $t/4\pi$ is equal to the error in the Weyl asymptotic estimate for the eigenvalue counting function of the Laplacian on the standard flat torus. We give a sharp asymptotic expression for the average value of the difference over the interval $0 \leq t \leq R$. We obtain similar results for families of ellipses. We also obtain relations to the eigenvalue counting function for the Klein bottle and projective plane. | math.FA | math |
Average Number of Lattice Points in a Disk
Sujay Jayakar
Robert S. Strichartz ∗
Abstract
The difference between the number of lattice points in a disk of
radius √t/2π and the area of the disk t/4π is equal to the error in
the Weyl asymptotic estimate for the eigenvalue counting function of
the Laplacian on the standard flat torus. We give a sharp asymptotic
expression for the average value of the difference over the interval
0 ≤ t ≤ R. We obtain similar results for families of ellipses. We
also obtain relations to the eigenvalue counting function for the Klein
bottle and projective plane.
1 The simplest case
Consider the standard flat torus [0, 1] × [0, 1] with boundaries identified.
The eigenfunctions of the Laplacian are e2πin·x for n ∈ Z2 with eigenvalues
(2π)2n2, so the eigenvalue counting function is
N (t) = #
(1.1)
t/2π of radius √t/2π about the
the number of lattice points inside the disk B√
origin. To first approximation N (t) is the area of the disk t/4π, and this is
exactly the Weyl asymptotic law. The problem of estimating the difference
,
(cid:110)
n ∈ Zd : n ≤ √t/2π
(cid:111)
D(t) = N (t) −
t
4π
(1.2)
is notoriously difficult (conjectured to be O(t1/4+) for every > 0). Here
∗Research supported in part by NSF grant #DMS-0652440
2010 Mathematics Subject Classification. Primary 35J05; Primary 42B99.
Key words and phrases: lattice points, Weyl asymptotics, Bessel function
1
Figure 1: D(t)
Figure 2: t1/4D(t)
Figure 3: A(t)
Figure 4: t1/4A(t)
we study the simpler problem of approximating the average value
(cid:90) R
A(R) =
1
R
D(t) dt.
0
(1.3)
Note that we are not taking the absolute value of D(t) in the average, so
we may exploit the cancellation from regions where N (t) is greater than and
less than t/4π. We will show that A(R) = O(R−1/4) as R → ∞, and more
precisely
A(R) = g(R
−1/2)R
−1/4 + O(R
−3/4) as R → ∞
(1.4)
where g(R) is an explicit uniformly almost periodic function of mean value
zero. Somewhat different but related ideas are given in Bleher [2, 3]. The
2
0200040006000800010000t−15−10−5051015D(t)0200040006000800010000t−2.0−1.5−1.0−0.50.00.51.01.52.0t−1/4D(t)0200040006000800010000t−0.4−0.20.00.20.4A(t)0200040006000800010000t−2.0−1.5−1.0−0.50.00.51.01.52.0t1/4A(t)following Lemma is well-known (see [4], p. 74), but we include the proof for
the convenience of the reader.
Lemma 1. We have
A(R) =
(cid:88)
n(cid:54)=0
√R)
1
πn2 J2(n
(1.5)
where n = (n1, n2) is a variable in Z2, and J2 denotes the Bessel function.
The series in (1.5) converges uniformly and absolutely.
Proof. Let χt denote the characteristic function of the ball B√
well-known that
t/2π.
It is
χt(z) =
z (cid:54)= 0
z = 0
(1.6)
Following standard methods (see [7] or [4]) we apply the Poisson summation
formula to
FR,δ =
1
R
χt ∗ ψδ dt,
0
(1.7)
where ψδ is a smooth approximate identity. The ψδ convolution makes FR,δ
smooth, but eventually we will let δ → 0. Note that
t
(cid:40) √
2πz J1(z√t)
(cid:90) R
t
4π
(cid:90) R
1
R
n∈Z2
(cid:88)
(cid:88)
(cid:90) R
1
R
n∈Z2
1
R
0
(cid:88)
n∈Z2
N (t) dt = lim
δ→0
0
χt(n) ψ(δn) dt
(cid:90) R
0
1
R
FR,δ(n)
(cid:90) R
dt +
(cid:88)
(cid:90) R
n(cid:54)=0
0
t
4π
(cid:88)
FR,δ(n).
(1.8)
(1.9)
√t
2πn
J1(n√t) dt ψ(δn)
J1(n√t) dt ψ(δn).
(1.10)
The Poisson summation formula gives
(cid:88)
n∈Z2
FR,δ(n) =
=
=
by (1.6). Combining (1.8) and (1.9) yields
√t
2πn
A(R) = lim
δ→0
1
R
n(cid:54)=0
0
3
(cid:90) R
0
(cid:90) R
0
1
R
sα+1Jα(s) ds = Rα+1Jα+1(R)
(1.11)
for α = 1, together with the change of variables s = n√t, to evaluate the
integral in (1.10)
√t
2πn
J1(n√t) dt =
1
R
R
s2
πn4 J1(s) ds
(1.12)
(cid:90) n√
0
1
=
πn2 J2(nR),
Now we use the property of Bessel functions ([6])
(cid:88)
n(cid:54)=0
and substitute this into (1.10) to obtain
A(R) = lim
δ→0
√R) ψ(δn).
1
πn2 J2(n
(1.13)
The estimate J2(n√R) = O(
n1/2R1/4 ) shows the convergence of the sum in
(1.13) without the term ψ(δn), so we can take the limit in (1.13) and obtain
(1.5).
1
Theorem 2. Consider the uniformly almost periodic function with mean
value zero
g(x) = −
−5/2 cos
n
nx −
π
4
.
(1.14)
(cid:16)
(cid:17)
(cid:88)
n(cid:54)=0
√2
π3/2
We have
A(R) = g(R1/2)R
−1/4 + O(R
−3/4) as R → ∞.
(1.15)
More generally, there exists a sequence of uniformly almost periodic functions
g1, g2, . . . with g1 = g such that for any n,
A(R) =
gj(R1/2)R
1
4− j
2 + O(R
− 1
4− n
2 ).
(1.16)
(cid:114)
Proof. We use the well-known asymptotic expression for Bessel functions
Jα(x) =
−1/2 cos
x
2
π
+ O(x
−3/2)
as x → ∞.
(1.17)
n(cid:88)
(cid:18)
j=1
x −
(cid:19)
1
2
απ −
π
4
4
Figure 5: g(√t)
(cid:114)
When α = 2 this is
Figure 6: t1/4A(t) − g(√t)
(cid:17)
x
2
π
−1/2 cos
J2(x) = −
−3/2),
and we substitute this into (1.5) with x = n√R to obtain
n2 O((n√R)
−3/2).
A(R) = g(R1/2)R
−1/4 +
x −
+ O(x
n(cid:54)=0
π
4
1
(1.18)
(1.19)
It is easy to see that the remainder term in (1.19) is O(R−3/4), so (1.19) yields
(1.15). To obtain the more refined asymptotic expression (1.16) we use the
known more refined asymptotic expansion for Bessel functions (see [6]). In
particular we note that it is possible to obtain explicit series expansions of
the functions gj; for example,
15√2
8π3/2
(cid:88)
−7/2 sin
nx −
g2(x) =
(1.20)
(cid:16)
(cid:17)
n
π
4
.
n(cid:54)=0
It is also reasonable to consider the function N ((2πr)2) that counts the
number of lattice points inside the ball Br of radius r, the difference D((2πr)2) =
N ((2πr)2) − πr2, and the average with respect to the radius variable
A(R) =
1
R
D((2πr)2) dr.
(1.21)
(cid:16)
(cid:88)
(cid:90) R
0
5
0200040006000800010000t−2.0−1.5−1.0−0.50.00.51.01.52.0g(√t)0200040006000800010000t−0.4−0.20.00.20.4t1/4A(t)−g(√t)Figure 7: √t(t1/4A(t) − g(√t))
Figure 8: A(t)
Figure 9:
√
1
2
2π A(2πt2)
Figure 10:
√
1
2
This is a different average, but a change of variable shows that
(cid:32)
(cid:90) 2πR2
2π A(2πt2) − A(t)
(cid:33)
.
(1.22)
A(R) =
1
2
R
1
2πR2
0
D(t)
dt
t1/2
Since most of the contribution to the integral occurs for values of t near
2πR2, we see that A(R) has the same asymptotics as
√
1
2
2π A(2πR2).
In Figure 1 we show the graph of D(t) and in Figure 2 the graph of
t−1/4D(t). This illustrates the rough t1/4 growth rate of D(t). In Figure 3
we show the graph of A(t), and Figure 4 the graph of t1/4A(t). Figure 5
shows the graph of g(√t), which is almost identical to Figure 4 for large t.
Figure 6 shows the difference of t1/4A(t) and g(√t), and Figure 7 shows this
6
0200040006000800010000t−3−2−10123√t(t1/4A(t)−g(√t))020406080100t−0.4−0.20.00.20.4A(t)020406080100t−0.4−0.20.00.20.4A(2πt2)/2√2π020406080100t−2.0−1.5−1.0−0.50.00.51.01.52.0A(2πt2)/2√2π−A(t)difference multiplied by t1/2. Figure 8 shows the graph of A(t). Figure 9
2π A(2πt2), which agrees with Figure 8 for large t, and
shows the graph of
Figure 10 shows the difference. For more data see the website [5].
√
1
2
(2π)2
2 The general case
Consider the general flat 2-dimensional torus, R2/L for some lattice L. The
eigenfunctions of the Laplacian (restriction of the standard R2 Laplacian)
have the form e2πix·ξ for ξ in the dual lattice L
(cid:48), with eigenvalues (2π)2ξ · ξ.
(cid:48) we can find an orthonormal
(cid:19)2(cid:33)
By diagonalizing the quadratic form ξ · ξ on L
(cid:18)n · v2
basis v1, v2 in R2 and positive constants a1, a2, such that the eigenvalues are
for n ∈ Z2.
.
(cid:19)2(cid:33)1/2
(cid:18) n · v2
(cid:19)2
≤ √t/2π
(cid:41)
(cid:19)2(cid:33)
(cid:32)(cid:18) x · v1
(cid:19)2
(cid:18) x · v2
(cid:32)(cid:18) n · v1
(cid:19)2
n ∈ Z2 :
(cid:32)(cid:18) n · v1
(cid:40)
x ∈ R2 : (2π)2
In place of disks we consider the family of ellipses
Thus the eigenvalue counting function is
N (t) = #
≤ t
.
+
a2
+
a2
Et =
a1
+
a2
(2.2)
(2.1)
a1
a1
Of course N (t) is just the number of lattice points in Et, and the volume of
Et is a1a2t
for the difference and define
the average A(R) by (1.3). The analog of (1.6) is
4π . Again we write D(t) = N (t) − a1a2t
4π
a1a2
2π√(a1z·v1)2+(a2z·v2)2 J1(
χEt(z) =
a1a2t
4π
Lemma 3. We have
A(R) =
n(cid:54)=0
π[(a1n · v1)2 + (a2n · v2)2]
a1a2
the series converging uniformly and absolutely.
7
(cid:40)
(cid:88)
(cid:112)
(a1z · v1)2 + (a2z · v2)2√t)
z (cid:54)= 0
z = 0
(2.3)
(cid:112)
(a1n · v1)2 + (a2n · v2)2√R),
J2(
(2.4)
Figure 11: g(√t), a1 = 2, a2 = 1/2
Figure 12: t1/4A(t) − g(√t)
Proof. The same proof as for Lemma 1, with (2.3) used in place of (1.6).
Theorem 4. The asymptotic expansions (1.15) and (1.16) hold, where now
g(x) = −
√2
π3/2 a1a2
(cid:0)(a1n · v1)2 + (a2n · v2)2(cid:1)−5/4
(cid:17)
(cid:88)
(cid:16)(cid:0)(a1n · v1)2 + (a2n · v2)2(cid:1)1/2
n(cid:54)=0
· cos
π
4
x −
.
(2.5)
Proof. Same as for Theorem 2, using Lemma 3 in place of Lemma 1.
See Figure 11 for g(√t) with a1 = 2 and a2 = 1/2 and Figure 12 for the
difference t1/4A(t) − g(√t) for the same family of ellipses. Plots for different
families of ellipses are available on the website [5].
3 The Klein bottle and projective plane
If we identify the vertical boundaries of the square directly, and the horizontal
boundaries with reflection, we obtain the standard flat Klein bottle KB. In
terms of functions defined on the square, we are imposing the boundary
conditions u(0, y) = u(1, y) and u(x, 0) = u(1 − x, 1) in order to have a
function on KB. We may cover KB by the rectangular torus [0, 1] × [0, 2]
with the identities
(cid:40)
(3.1)
u(x + 1, y) = u(x, y)
u(1 − x, y + 1) = u(x, y)
8
0200040006000800010000t−4−3−2−101234g(√t)0200040006000800010000t−2.0−1.5−1.0−0.50.00.51.01.52.0t1/4A(t)−g(√t)describing the lifts of functions on KB to R2. The eigenfunctions of the
Laplacian on KB lift to eigenfunctions on the rectangular torus, and so
are linear combinations of functions of the form e2πi(jx+ k
2 y) with eigenvalue
2 (y+1)) = (−1)ke2πi(−jk+ k
(2π)2(j2 + ( k
2 y).
Thus there are two families of eigenfunctions
2 )2). Now we observe that e2πi(j(1−x)+ k
e2πi k
2 y
e2πi(jx+ k
for k even (corresponding to j = 0), and
2 y) + (−1)ke2πi(−jx+ k
2 y)
for j > 0.
(3.2)
(3.3)
We can therefore see that the eigenvalue function NKB is close to one half
the counting function NT1,2 for the [0, 1] × [0, 2] torus.
Theorem 5. NKB(t) = 1
2NT1,2(t) ± 1
2.
Proof. NT1,2(t) counts all integers j, k such that j2 + ( k
(2π)2 . When
j (cid:54)= 0 the pair ±j contributes just a single eigenvalue to NKB(t). When
√
t
j = 0 we count all k such that k ≤
π in NT1,2(t), but just the even values
√
t
π } = 1
of k in NKB(t), and #{k even : k ≤
√
π } ± 1
2.
2{k : k ≤
2 )2 ≤ t
t
It is interesting to compare the Klein bottle with the projective plane
(PP) obtained from [0, 1] × [0, 1] by identifying both sets of boundary edges
with reflections. Functions on PP lift to R2 with the identities
(cid:40)
u(1 − x, y + 1) = u(x, y)
u(x + 1, 1 − y) = u(x, y)
(3.4)
and the torus [0, 2] × [0, 2] is a four-fold covering of PP. However, while it is
possible to pull back the standard Laplacian to PP, the pairs {(0, 0), (1, 1)}
and {(0, 1), (1, 0)} of identified points on PP are singularities (cone points
with total angle π) with respect to the otherwise flat metric.
Reasoning as in the KB example, we know that eigenfunctions of the
2 y)
2 )2). Imposing the conditions (3.4) leads to
Laplacian on PP must be linear combinations of the functions e2πi( j
with eigenvalue (2π)2(( j
2)2 + ( k
2 x+ k
9
four families of eigenfunctions:
−2πi k
constants (corresponding to j = 0 and k = 0)
2 y for k > 0 even (corresponding to j = 0 but k (cid:54)= 0)
2 x for j > 0 even (corresponding to k = 0 but j (cid:54)= 0)
2 y)(cid:17)
2 x− k
e2πi(− j
2 x+ k
2 y) + e2πi( j
e2ıi k
e2πi j
e2πi( j
2 y + e
2 x + e
2 x+ k
−2πi j
2 y) + e2πi(− j
2 x− k
2 y) + (−1)j+k(cid:16)
(3.5)
(3.6)
(3.7)
for j > 0 and k > 0.
(3.8)
This leads to the identity
NPP(t) =
1
4
NT2,2(t) +
1
4 ±
1
2
=
1
4
NT1,1(4t) +
1
4 ±
1
2
.
(3.9)
Similar results hold for KB and PP constructed from the tori considered
in section 2. Related questions in the context of fractal Laplacians with the
Sierpinski carpet replacing the square are discussed in [1].
References
[1] M. Begu´e, T. Kalloniatis, and R. Strichartz. Harmonic functions and the
spectrum of the Laplacian on the Sierpinski carpet. Preprint, 2012.
[2] P.M. Bleher. On the distribution of the number of lattice points inside a
family of convex ovals. Duke Math J. 67, pages 461 -- 481, 1991.
[3] P.M. Bleher. Distribution of the error term in the Weyl asymptotics
for the Laplace operator on a two-dimensional torus and related lattice
problems. Duke Math J. 70, pages 655 -- 682, 1993.
[4] H. Iwaniec and E. Kowalski. Analytic Number Theory. AMS Colloq.
Publ. vol 53, 2004.
[5] S. Jayakar and R. Strichartz. Average number of lattice points in a disk.
http://www.math.cornell.edu/~sujay/lattice, June 2012.
[6] N.N Lebedev. Special functions and their applications. Dover Publica-
tions, New York, 1965.
10
[7] E.M. Stein and R. Shakarchi. Functional Analysis. Princeton Univ. Press,
2011.
563 Malott Hall, Cornell University, Ithaca, NY 14853, USA
Email addresses: [email protected], [email protected]
11
|
1807.11865 | 1 | 1807 | 2018-07-31T15:28:08 | On extensions of symmetric operators | [
"math.FA",
"math.SP"
] | We give an explicit description of all minimal self-adjoint extensions of a densely defined, closed symmetric operator in a Hilbert space with deficiency indices $(1, 1)$. | math.FA | math | ON EXTENSIONS OF SYMMETRIC OPERATORS
NAMIG J. GULIYEV
Abstract. We give an explicit description of all minimal self-adjoint exten-
sions of a densely defined, closed symmetric operator in a Hilbert space with
deficiency indices (1, 1).
There is a widely used linearization technique in the theory of Sturm -- Liouville
problems with boundary and/or discontinuity conditions polynomially dependent
on the eigenvalue parameter. One considers a Hilbert (or Pontryagin) space of
the form L2 ⊕ Ck and constructs a self-adjoint operator in this space such that
the eigenvalue problem for this operator and the original boundary value problem
become equivalent, in the sense that their eigenvalues coincide, the eigenfunctions
of the latter problem are in one-to-one correspondence with the first components
of the eigenvectors of the former problem, and so on (see, e.g., [1], [2], [8], and the
references therein). Fulton [6, Remark 2.1] attributes this technique to Friedman
[5, pp. 205 -- 207]. The purpose of this short paper is to show that a straightforward
generalization of this technique gives an explicit description of all minimal self-
adjoint extensions of a densely defined, closed symmetric operator with deficiency
indices (1, 1) (see below for definitions).
Let A be a densely defined, closed symmetric operator in a separable Hilbert
space H with deficiency indices (1, 1). Let {C, Γ0, Γ1} be a boundary triplet for A∗.
This means that Γ0, Γ1 : D(A∗) → C are two linear mappings such that abstract
Green's identity
hA∗x, yiH − hx, A∗yiH = Γ1x · Γ0y − Γ0x · Γ1y,
x, y ∈ D(A∗)
holds and the mapping Γ : D(A∗) → C2, x 7→ (Γ0x, Γ1x) is surjective [4], [12,
Chapter 14]. The domain of A then coincides with the kernel of Γ:
D(A) = {x ∈ D(A∗) Γ0x = Γ1x = 0}.
(1)
8
1
0
2
l
u
J
1
3
]
.
A
F
h
t
a
m
[
1
v
5
6
8
1
1
.
7
0
8
1
:
v
i
X
r
a
If eA is a self-adjoint operator in a Hilbert space eH ⊃ H such that A ⊂ eA, then eA
is called a self-adjoint extension of A. A self-adjoint extension eA is called minimal
if no nontrivial subspace of eH ⊖ H is reducing for eA [10, Section 4], or equivalently
[12, Lemma 5.17]
eH = span(cid:26)(cid:16)eA − λI(cid:17)−1
x(cid:12)(cid:12)(cid:12)(cid:12) x ∈ H, λ ∈ C \ R(cid:27),
where span denotes the linear span.
2010 Mathematics Subject Classification. 47A20, 47B25.
Key words and phrases. symmetric operator, self-adjoint extension, generalized resolvent.
1
R(λ) = PH(cid:16)eA − λI(cid:17)−1(cid:12)(cid:12)(cid:12)(cid:12)H
for some self-adjoint extension eA, where PH is the orthogonal projection onto H.
For every generalized resolvent there is a unique (up to unitary equivalence) min-
imal self-adjoint extension with this property (see [10, Theorem 8]). On the other
hand, there is a one-to-one correspondence between generalized resolvents R(λ) and
functions ω(λ) holomorphic on the open upper half-plane C+ with ω(λ) ≤ 1: for
every x ∈ H the value y := R(λ)x satisfies the equation
and the condition
(see [3]). Denoting
A∗y − λy = x
(ω(λ) − 1) Γ1y − i (ω(λ) + 1) Γ0y = 0
f (λ) :=
iω(λ) + i
1 − ω(λ)
we obtain a one-to-one correspondence between generalized resolvents R(λ) and
Herglotz -- Nevanlinna functions f , i.e. functions holomorphic on C+ with Im f (λ) ≥
0 (cf. [13, Subsection 1.2]). The above condition then becomes
Γ1y + f (λ)Γ0y = 0.
(2)
The "Dirichlet" condition Γ0y = 0 corresponds to f = ∞. We will denote the gener-
alized resolvent corresponding to f by Rf (λ). It should be noted that complete pa-
rameterizations of all generalized resolvents in terms of Herglotz -- Nevanlinna func-
tions were first obtained independently by Naimark [11] and Krein [9].
We are now ready to construct our minimal self-adjoint extension corresponding
to a Herglotz -- Nevanlinna function f . This function has a unique representation of
the form [12, Appendix F]
f (λ) = h0λ + h +Z +∞
−∞ (cid:18) 1
t − λ
t
1 + t2(cid:19) dσ(t),
−
where h0 ≥ 0, h ∈ R, and
Z +∞
−∞
dσ(t)
1 + t2 < ∞
(the reader may refer to [7, Appendix A] for some examples of such representations).
If h0 > 0 then we consider the Hilbert space eH := H ⊕ L2(R; dσ) ⊕ C with inner
product given by
2
NAMIG J. GULIYEV
A holomorphic operator-valued function R(λ) on C \ R is called a generalized
resolvent of A if
for
x1(t)y1(t) dσ(t) +
x2y2
h0
−∞
hex,eyi eH := hx0, y0iH +Z +∞
ey =
ex =
,
x0
x1
x2
y0
y1
y2
∈ eH,
ON EXTENSIONS OF SYMMETRIC OPERATORS
3
and define the operator
eAex :=
Γ1x0 + hΓ0x0 +R +∞
A∗x0
tx1(t) − Γ0x0
−∞ (cid:16)x1(t) − t
1+t2 Γ0x0(cid:17) dσ(t)
with
and
D(eA) := {ex ∈ eH x0 ∈ D(A∗), tx1(t) − Γ0x0 ∈ L2(R; dσ), x2 = −h0Γ0x0}.
If h0 = 0 then we set eH := H ⊕ L2(R; dσ), and define eA by
tx1(t) − Γ0x0(cid:19)
A∗x0
eAex :=(cid:18)
D(eA) :=(cid:26)ex ∈ eH(cid:12)(cid:12)(cid:12)(cid:12) x0 ∈ D(A∗), tx1(t) − Γ0x0 ∈ L2(R; dσ),
1 + t2 Γ0x0(cid:19) dσ(t) = 0(cid:27) .
The extension eA is canonical (i.e. eH = H) if and only if f is a real constant (or ∞).
Theorem. For each Herglotz -- Nevanlinna function f the operator eA defined above
is a minimal self-adjoint extension of the operator A with the corresponding gener-
alized resolvent Rf (λ).
Γ1x0 + hΓ0x0 +Z +∞
−∞ (cid:18)x1(t) −
t
Proof. We will consider the case h0 > 0; the other case can be proved similarly. To
prove the self-adjointness, let ey, ez ∈ eH be such that
heAex,eyi eH = hex,ezi eH
for all ex ∈ D(eA). Taking into account (1) and setting x1(t) ≡ 0 = x2 we obtain
hAx0, y0iH = hx0, z0iH,
x0 ∈ D(A).
(3)
Hence y0 ∈ D(A∗) and z0 = A∗y0. By surjectivity of Γ, there exists x0 ∈ D(A∗)
with Γ0x0 = 0 and Γ1x0 = 1. Then for this x0 and x1(t) ≡ 0 the equality (3) gives
y2 = −h0Γ0y0. Setting x0 = 0 in (3) we get
Z +∞
−∞
x1(t)(ty1(t) − Γ0y0 − z1(t)) dσ(t) = 0
for all x1(t) ∈ L2(R; dσ) with tx1(t) ∈ L2(R; dσ). Since such functions are dense
in L2(R; dσ), we obtain ty1(t) − Γ0y0 = z1(t) ∈ L2(R; dσ). Finally choosing x0 ∈
D(A∗) with Γ0x0 = 1, the equality (3) yields
z2 = Γ1y0 + hΓ0y0 +Z +∞
−∞ (cid:18)y1(t) −
t
1 + t2 Γ0y0(cid:19) dσ(t).
check that if
Thus ey ∈ D(eA) and eAey =ez.
To see that the generalized resolvent corresponding to eA is Rf (λ), it suffices to
ey =
=(cid:16)eA − λI(cid:17)−1
y0
y1
y2
x
0
0
4
NAMIG J. GULIYEV
then y0 satisfies (2), and this is straightforward. Finally, to check the minimality,
we need to verify that the linear span of all ey of this form with all possible values
of x ∈ H and λ ∈ C \ R is dense in eH. To this end, let ez ∈ eH be orthogonal to all
such ey, i.e.
dσ(t) − z2Γ0y0 = 0
z1(t)
hez,eyi eH = hz0, y0iH +Z +∞
−∞
Γ0y0
t − λ
for all y0 ∈ D(A∗). Surjectivity of Γ implies z0 = 0 and
Z +∞
−∞
z1(t)
t − λ
dσ(t) = z2,
λ ∈ C \ R.
Now the Stieltjes -- Perron inversion formula [12, Theorem F.2] applied to the regular
complex Borel measure z1 dσ yields z1(t) = 0 for dσ-a.e. t and consequently z2 =
0.
(cid:3)
References
[1] C. Bartels, S. Currie, M. Nowaczyk, and B. A. Watson, Sturm -- Liouville problems with trans-
fer condition Herglotz dependent on the eigenparameter: Hilbert space formulation, Integral
Equations Operator Theory 90 (2018), no. 3, Art. 34, 20 pp. arXiv:1804.07149
[2] J. Behrndt and F. Philipp, Finite rank perturbations in Pontryagin spaces and a Sturm --
Liouville problem with λ-rational boundary conditions, Indefinite inner product spaces, Schur
analysis, and differential equations, Birkhauser/Springer, Cham, 2018, pp. 163 -- 189.
[3] V. M. Bruk, A certain class of boundary value problems with a spectral parameter in the
boundary condition (Russian), Mat. Sb. (N.S.) 100(142) (1976), no. 2, 210 -- 216.
[4] V. Derkach, Boundary triplets, Weyl functions, and the Krein formula, Operator theory,
Springer, Basel, 2015, pp. 183 -- 218.
[5] B. Friedman, Principles and techniques of applied mathematics, John Wiley & Sons, Inc.,
New York, 1956.
[6] C. T. Fulton, Two-point boundary value problems with eigenvalue parameter contained in
the boundary conditions, Proc. Roy. Soc. Edinburgh Sect. A 77 (1977), no. 3-4, 293 -- 308.
[7] F. Gesztesy and E. Tsekanovskii, On matrix-valued Herglotz functions, Math. Nachr. 218
(2000), 61 -- 138. arXiv:funct-an/9712004
[8] N. J. Guliyev, Essentially isospectral transformations and their applications, preprint.
arXiv:1708.07497
[9] M. G. Krein, On Hermitian operators with deficiency indices one (Russian), Dokl. Akad.
Nauk SSSR 43 (1944), 339 -- 342.
[10] M. A. Naimark, Spectral functions of a symmetric operator (Russian), Izv. Akad. Nauk SSSR.
Ser. Mat. 4 (1940), no. 3, 277 -- 318.
[11] M. A. Naimark, On spectral functions of a symmetric operator (Russian), Izv. Akad. Nauk
SSSR. Ser. Mat. 7 (1943), no. 6, 285 -- 296.
[12] K. Schmudgen, Unbounded self-adjoint operators on Hilbert space, Springer, Dordrecht, 2012.
[13] A. V. Shtraus, On spectral functions of differential operators (Russian), Izv. Akad. Nauk
SSSR. Ser. Mat. 19 (1955), no. 4, 201 -- 220.
Institute of Mathematics and Mechanics, Azerbaijan National Academy of Sciences,
9 B. Vahabzadeh str., AZ1141, Baku, Azerbaijan.
E-mail address: [email protected]
|
1308.6227 | 1 | 1308 | 2013-08-28T17:45:08 | On the Convergence of Regular Families of Cardinal Interpolators | [
"math.FA"
] | In this note a general way to develop a cardinal interpolant for $l^2$-data on the integer lattice $Z^n$ is shown. Further, a parameter is introduced which allows one to recover the original Paley-Weiner function from which the data came. | math.FA | math | ON THE CONVERGENCE OF REGULAR FAMILIES
OF CARDINAL INTERPOLATORS
JEFF LEDFORD
Abstract. In this note a general way to develop a cardinal in-
terpolant for l2-data on the integer lattice Zn is shown. Further,
a parameter is introduced which allows one to recover the original
Paley-Weiner function from which the data came.
Keywords. cardinal interpolation, multiquadric, multivariate in-
terpolation, Paley-Wiener functions
.
A
F
h
t
a
m
[
1
v
7
2
2
6
.
8
0
3
1
:
v
i
X
r
a
1. Introduction
In this article we prove a result similar to the ones found in [2], [3],
[4], [5], and [6]. These results may be thought of as outgrowths of
Schoenberg's work on splines (see for instance [7]). The general set
up is as follows. Suppose that we have data {f (j)} from some class,
we wish to interpolate this data with an interpolant that depends on
some parameter.
In each case what is investigated is what happens
as the parameter approaches a limiting value. The papers mentioned
above all fix a particular type of what it called in this article a cardinal
interpolator. The goal of this paper is to unify the procedures in those
and similar papers in order to provide conditions on an interpolator
which allows one to recover Paley-Wiener functions from their samples
on the integer lattice by allowing the parameter to approach a limiting
value. In doing so, more examples are uncovered.
This article is organized as follows. In the next, section several pre-
liminary definitions are laid out. We define what is meant by cardinal
interpolator in Section 3 and introduce the corresponding fundamental
function. Section 4 deals with properties of the fundamental function,
while in Section 5 regular families are introduced and contains the proof
of the main theorem. The final section consists of three examples of
regular families of cardinal interpolators.
Finally, our calculations will often involve a positive constant C,
whose value depends on its occurrence but not on various parameters
unless otherwise stated.
1
2
JEFF LEDFORD
2. Definitions and Basic Facts
In this section, we collect several definitions and general facts that
will be used in what follows. Our methods require the use of the Fourier
transform, for which we adopt the following convention.
Definition 1. If f ∈ L1(Rn) we define its Fourier transform, denoted
f (ξ) or F (f )(ξ), to be:
(1)
f (x)e−ix·ξdx,
F (f )(ξ) = f (ξ) = (2π)−n/2ZRn
where x · ξ =Pn
j=1 xjξj.
We will encounter examples which are not integrable in this case we
adopt the convention of [8], particularly Definition 8.9 on page 103,
and include it here for convenience changing the name a little to avoid
confusion.
Definition 2. For m ∈ Z+, the set of all functions γ ∈ S that satisfy
γ(ξ) = O(kξkm) for kξk → 0 will be denoted by Sm.
Here S is the Schwartz space, and k · k is the Euclidean norm on Rn.
Definition 3. Suppose that φ : Rn → C is continuous and slowly
increasing. A measurable function φ ∈ L2
loc(Rn \ {0}) is called the
specialized Fourier transform of φ if there exists an integer m ∈ Z+
such that
ZRn
φ(x)γ(x)dx =ZRn
φ(ξ)γ(ξ)dξ
is satisfied for all γ ∈ S2m. The least such m is called the order of φ.
Before we continue, a few comments are needed. In [8] this is called
the generalized Fourier transform. However, there is already a cus-
tomary definition of the generalized Fourier transform, which uses the
entire Schwartz class as test functions. To avoid this confusion, I have
If a function φ(x) ∈ L1(Rn) then its specialized
changed the name.
and classical Fourier transform coincide, and the order is 0. The same
is true for φ(x) ∈ L2(Rn). Finally, the specialized and the generalized
(defined in the usual way) Fourier transform coincide on the set S2m.
We turn to constraints on the data that we will use. It is not nec-
essary to start out assuming that the data comes from any particular
function, however it serves our purposes later. Thus throughout the
paper we will sample functions exclusively from the Paley-Wiener space
P Wπ.
REGULAR CARDINAL INTERPOLATORS
3
Definition 4. We define the Paley Wiener space, denoted P Wπ, as
(2)
P Wπ = {f ∈ L2(Rn) : supp( f ) ⊆ [−π, π]n}.
We note that the norm on this space is given by
kfkP Wπ = kfkL2(Rn).
Lemma 1. Suppose that f ∈ P Wπ, then {f (j)} ∈ l2(Zn).
Proof. This is a direct consequence of Plancherel's theorem.
(cid:3)
3. Cardinal Interpolators
Here we set out hypotheses on a function φ(x) which will be used
throughout the rest of the paper.
Definition 5. We say that a function φ(x) is a cardinal interpolator
if it satisfies the following conditions:
(H1) φ(x) is a real valued slowly increasing function on Rn,
(H2) φ(ξ) ≥ 0 and φ(ξ) ≥ δ > 0 in [−π, π]n,
(H3) φ(ξ) ∈ C n+1(Rn \ {0}),
(H4) There exists ǫ > 0 such that if α ≤ n + 1,
(H5) For any multi-index α, with α ≤ n + 1,
Dα φ(ξ) = O(kξk−(n+ǫ)) as kξk → ∞,
α
Dαj φ(ξ)
Yj=1
[ φ(ξ)]α+1 ∈ L∞([−π, π]n)
where
αj = α.
α
Xj=1
For a cardinal interpolant φ(x), we form the corresponding so-called
fundamental function of interpolation, denoted Lφ(x), by its Fourier
transform.
(3)
Lφ(ξ) = (2π)−n/2 φ(ξ)"Xj∈Zn
φ(ξ − 2πj)#−1
4. Properties of the Fundamental Function
The main goal of this section is to establish the integrability Lφ(x).
This is done through several lemmas. First we rewrite the transform
of the fundamental function as
(2π)n/2 Lφ(ξ) =
φ(ξ)
φ(x) + u(ξ)
,
4
JEFF LEDFORD
φ(ξ − 2πj). Condition (H4) implies that for all
where u(ξ) = Pj6=0
α ≤ n + 1, Dαu is well defined on [−π, π]n. Notice also that (H2)
provides the transparent bound Lφ(ξ) ≤ 1. Hypotheses (H4) and
(H5) will allow us to control the growth of the derivative of the above
expression. Before stating our first lemma we adopt the notation N0 =
N ∪ {0}, and follow the convention D0f = f .
Lemma 2. If α ∈ N0 is a multi-index, with α ≥ 1, we have
(4)
Dα Lφ(ξ) =h φ(ξ) + uξi−α−1
Daγ φ(ξ)Yδ /∈B
Ca,B Yγ∈B
Xa∈Aα,B⊆Sα
where Aα =(cid:8)(a1, a2, . . . , aα+1) : aj ∈ Nn
0 , a1 + · · · + aα+1 = α(cid:9), Sα =
{1, 2, . . . ,α+1}, and Ca,B are constants. Furthermore, the coefficients
corresponding to B = ∅ and B = Sα are 0.
Proof. We will induct on α. If α = 1, we use the quotient rule to see
that
Daδ u(ξ),
Dα Lφ(ξ) = (2π)−n/2 u(ξ)Dα φ(ξ) − φ(ξ)Dαu(ξ)
( φ(ξ) + u(ξ))2
.
This is the desired form, and since all of the terms on top are mixed
products of φ(ξ) and u(ξ) the coefficient condition holds as well. Now
if we assume that (4) holds for all α ≤ n, then applying the Leibniz
rule and collecting terms yields the result.
(cid:3)
We have the following result.
Corollary 1. If α ≤ n + 1, then Dα Lφ ∈ L∞([−π, π]n).
Proof. Combine (H5) with (4).
(cid:3)
In order to determine the growth of Lφ(ξ) away from the origin, we
examine the function
P (ξ) =
1
φ(ξ) + u(ξ)
.
Lemma 3. For ξ ∈ [−π, π]n and α a multi-index, we have
(5)
DαP (ξ) =h φ(ξ) + u(ξ)i−α−1
Xa∈Aα
where Aα =(cid:8)(a1, a2, . . . , aα) : aj ∈ Nn
0 , a1 + · · · + aα = α(cid:9).
Yj=1
α
Ca
Daj ( φ(ξ) + u(ξ)),
REGULAR CARDINAL INTERPOLATORS
5
Proof. We again induct on α. If α = 1, we use the quotient rule to
obtain
DαP (ξ) = −Dα( φ(ξ) + u(ξ))
( φ(ξ) + u(ξ))2
.
This is the desired result. Suppose now that (5) holds for all α ≤ n,
then applying the quotient rule and collecting terms yields the result.
(cid:3)
Corollary 2. For α ≤ n + 1, DαP (ξ) ∈ L∞(Rn).
Proof. Since P (ξ) is periodic, we need only combine (H5) and (5). (cid:3)
Proposition 1. If φ is a cardinal interpolator and α ≤ n + 1, then
Dα Lφ ∈ L1(Rn).
Proof. Since Lφ(ξ) = φ(ξ)P (ξ), we see that near the origin Corollary 1
controls the growth, while away from the origin Corollary 2 and (H4)
imply the result.
(cid:3)
Corollary 3. If φ is a cardinal interpolator, then Lφ(x) is continuous
and there exists a constant C > 0 such that
(6)
Proof. Continuity follows from Proposition 1 by letting α = 0. The
bound (6), follows as well since
(cid:12)(cid:12)(1 + kxkn+1)Lφ(x)(cid:12)(cid:12) ≤ C.
(cid:12)(cid:12)(1 + kxkn+1)Lφ(x)(cid:12)(cid:12) ≤ Xα≤n+1(cid:13)(cid:13)(cid:13)
Dα Lφ(ξ)(cid:13)(cid:13)(cid:13)L1(Rn)
.
(cid:3)
We now show why Lφ(x) is called a fundamental function of inter-
polation.
Lemma 4. Lφ(x) satisfies Lφ(k) = δ0,k for k ∈ Zn.
Proof. The proof is an easy calculation invoking the inversion theorem:
Lφ(ξ)eik·ξdξ
Lφ(k) =
=
=
=
1
1
1
(2π)n/2 ZRn
(2π)n/2 Xm∈ZnZ[−π,π]n
(2π)n/2 Z[−π,π]n
(2π)n Z[−π,π]n
1
Lφ(ξ − 2πm)eik·ξdξ
Lφ(ξ − 2πm)dξ
eik·ξ Xm∈Zn
eik·ξdξ = δ0,k.
6
JEFF LEDFORD
We turn now to constructing an interpolant consisting of translates
of Lφ(x).
Claim 1. If {f (j)}j∈Zn ∈ l∞(Zn) and we define Iφf (x) pointwise as
follows:
(cid:3)
Iφf (x) = Xj∈Zn
f (j)Lφ(x − j).
where φ(ξ) is a cardinal interpolator, then Iφf (x) is well defined and
continuous.
Proof. As a result of (6) we have the following:
Iφf (x) ≤ Ck{f (j)}kl∞(Zn) Xj∈Zn
1
1 + kx − jkn+1 < ∞.
Continuity follows from the Weierstrass M test and the continuity of
Lφ(x). Notice this bound holds for all x ∈ Rn because the last sum is
periodic.
(cid:3)
We are now in position to prove our first theorem.
Theorem 1. Let 1 ≤ p ≤ ∞. If {f (j)} ∈ lp(Zn) and Iφf (x) is defined
by (7), then Iφf (x) ∈ Lp(Rn).
Proof. We will use the Riesz-Thorin interpolation theorem. Consider
the linear map given by {f (j)} 7→ Iφf (x). From the above claim, we
can see that Iφ maps l∞(Zn) to C(Rn). In order to apply the theorem,
we must check the p = 1 case and the p = ∞ case. For p = 1 we have:
(7)
(8)
f (j)Lφ(x − j)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1(Rn)
kIφf (x)kL1(Rn) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj∈Zn
≤k{f (j)}kl1(Zn)kLφ(x − j)kL1(Rn)
1 + kxkn+1(cid:13)(cid:13)(cid:13)(cid:13)L1(Rn)k{f (j)}kl1(Z)
≤C(cid:13)(cid:13)(cid:13)(cid:13)
≤C1k{f (j)}kl1(Zn)
1
Here we've used Minkowski's inequality and Corollary 3. The p = ∞
case is the content of (8). Thus the Riesz-Thorin interpolation theorem
guarantees a constant Cp independent of {f (j)} such that
kIφf (x)kLp(Rn) ≤ Cpk{f (j)}klp(Zn).
(cid:3)
REGULAR CARDINAL INTERPOLATORS
7
5. Regular Families of Cardinal Interpolators
The goal of this section is to introduce regularity conditions on a
family of cardinal interpolators {φα} so that we may get recovery re-
sults similar to those found in [2], [3], [4], [5], and [6].
Definition 6. We call a family of functions {φα(ξ)}α∈A a regular fam-
ily of cardinal interpolators if for each α ∈ A ⊂ (0,∞), φα(x) is a
cardinal interpolator and in addition to this, we have:
φα(ξ + 2πj)
(R1) For j ∈ Zn \ {0}, ξ ∈ [−π, π]n, define Mj,α(ξ) =
,
φα(ξ)
Mj,α(ξ) = 0 for almost every ξ ∈
(R2) There exists Mj ∈ l1(Zn \ {0}), independent of α, such that for
then for j ∈ Zn \ {0},
[−π, π]n.
all j ∈ Zn \ {0}, Mj,α(ξ) ≤ Mj.
lim
α→∞
Depending on the interpolator φ(x), the parameter α may be contin-
uous or discrete. We will encounter examples of both in the following
section.
Proposition 2. If {φα} is a regular regular family of cardinal inter-
polators, then Lφα(ξ) → (2π)−n/2χ[−π,π]n(ξ) almost everywhere.
Proof. For ξ ∈ [−π, π]n we have:
φα(ξ)(cid:27)−1
If we can show
Lφα(ξ) = (2π)−n/2(cid:26)1 +
uα(ξ)
φα(ξ) → 0 we will be done with this part. Conditions
(R1) and (R2) allow us to apply the dominated convergence theorem
for series, which gives us the result. Now we have only to check ξ /∈
[−π, π]n. We may write ξ = 2πj + r, where j ∈ Zn \ {0} and r ∈
[−π, π]n. Then we have:
uα(ξ)
.
Lφα(ξ) = (2π)−n/2
φα(2πj + r)
φα(r) (cid:26)1 +
uα(r)
φα(r)(cid:27)−1
≤ (2π)−n/2Mj,α(ξ).
The last inequality comes from (H2), thus (R1) provides the desired
result.
(cid:3)
Proposition 3. If {φα} is regular a regular family of cardinal inter-
Lφα(ξ + 2πj) = 0 almost everywhere and in
polators, then lim
α→∞Xj6=0
L1([−π, π]n).
8
JEFF LEDFORD
Proof. We note that if j ∈ Zn \ {0}, then Lφα(ξ + 2πj) ≤ Mj,α(ξ) by
(R1), thus (R2) allows us to use the dominated convergence theorem
for series which gives the result.
(cid:3)
These propositions allow us to prove the main theorem.
Theorem 2. Suppose that { φα(ξ)}α∈A is a regular family of cardinal
interpolators and let f (x) ∈ P Wπ. Then we have the following limits:
(a) lim
(b) lim
α→∞(cid:13)(cid:13)(cid:13)(cid:13)Xj∈Zn
α→∞(cid:12)(cid:12)(cid:12)(cid:12)Xj∈Zn
f (j)Lφα(x − j) − f (x)(cid:13)(cid:13)(cid:13)(cid:13)L2(Rn)
f (j)Lφα(x − j) − f (x)(cid:12)(cid:12)(cid:12)(cid:12)
= 0,
= 0 uniformly in Rn.
Proof. (a) Since f ∈ P Wπ, we have that {f (j)}j∈Zn ∈ l2(Zn). This,
in turn, implies that Xj∈Zn
f (j)Lφα(x − j) ∈ L2(Rn). So we may use
Plancherel's theorem to estimate the norm.
2
L2(Rn)
2
f (j)Lφα(x − j) − f (x)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj∈Zn
f (j)e−ij·ξ − f (ξ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Lφα(ξ)Xj∈Zn
=Z[−π,π]n(cid:12)(cid:12)(cid:12)
f (ξ)(cid:16)(2π)n/2 Lφα(ξ) − 1(cid:17)(cid:12)(cid:12)(cid:12)
f (j)e−ij·ξ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+ZRn\[−π,π]n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Lφα(ξ)Xj∈Zn
2
dξ
L2(Rn)
2
dξ
We have used that f ∈ P Wπ and the well known L2(Rn) equality
(2π)n/2 f (ξ) = Xj∈Zn
f (j)e−ij·ξ for ξ ∈ [−π, π]n. Proposition 2 shows that
(2π)n/2 Lφα(ξ) → χ[−π,π]n(ξ) almost everywhere, thus the first integral
tends to 0 by the dominated convergence theorem, since we have the
elementary estimate (2π)n/2 Lφα(ξ) − 1 ≤ (2π)n/2 + 1. We estimate
the second integral.
REGULAR CARDINAL INTERPOLATORS
9
ZRn\[−π,π]n Lφα(ξ)Xj∈Zn
f (j)e−ij·ξ2dξ
= Xm∈Zn\{0}Z[−π,π]n Lφα(ξ + 2πm)Xj∈Zn
=(2π)n Xj∈Zn\{0}Z[−π,π]n Lφα(ξ + 2πj)2 f (ξ)2dξ
≤(2π)nZ[−π,π]n Xj∈Zn\{0}
Mj,α(ξ)2 f (ξ)2dξ
f (j)e−ij·(ξ+2πm)2dξ
This last term tends to 0 by the dominated convergence theorem as in
Proposition 3, noting that l1(Zn \ {0}) ⊂ l2(Zn \ {0}).
Schwarz inequality.
Part (b) is proved in a similar manner, with help from the Cauchy-
f (j)Lφα(x − j) − f (x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj∈Zn
f (j)e−ijξ − f (ξ)! eix·ξdξ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(2π)−(n/2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ZRn Lφα(ξ)Xj∈Zn
≤Z[−π,π]n(cid:12)(cid:12)(cid:12)
f (ξ)(cid:12)(cid:12)(cid:12)
Lφα(ξ) − (2π)−n/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+Xj6=0Z[−π,π]n(cid:12)(cid:12)(cid:12)
f (ξ)(cid:12)(cid:12)(cid:12)
Lφα(ξ + 2πj)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Lφα(ξ) − (2π)−n/2(cid:13)(cid:13)(cid:13)L2([−π,π]n)
≤kfkP Wπ(cid:13)(cid:13)(cid:13)
Mj,α(ξ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L2([−π,π]n)
+kfkP Wπ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj6=0
dξ
dξ
From Proposition 2 above we can see that the first term tends to 0.
By (R2) we have that the sum in the second term is bounded above by
kMjk2
l1(Zn\{0}), so Proposition 3 and the dominated convergence theo-
rem finish the proof.
(cid:3)
6. Examples of Regular Families of Interpolators
In this section, we exhibit three families of regular cardinal interpo-
lators. Our first example will be that of polyharmonic cardinal splines.
10
JEFF LEDFORD
The result that we attain is most similar to the one found in [3]. Our
second example will be a family of Gaussians, which has also been
studied previously, see for instance [4], [5]. Finally, we will consider a
family of multiquadrics and extend a result which is found in [6], as
well a provide what appears to be a novel example. For each example
we work out the one dimensional case, followed by the multivariate
case often suppressing the straightforward yet tedious details involved
in checking (H5).
6.1. Polyharmonic Cardinal Splines. Our first example was ex-
plored in [3], and motivated this treatment. We begin by considering
the one dimensional version of the example found there. A function
or distribution f is called k-harmonic, where k ∈ {1, 2, 3, . . .}, if it
satisfies
(9)
∆kf = 0
on Rn, where ∆ is the Laplacian operator and ∆k(f ) = ∆(∆k−1f ). A
function which satisfies (9) with k ≥ 1 is called polyharmonic. The
fundamental solution of (9) is given by a constant multiple of φk(x) =
x2k−1, whose Fourier transform is given by φk(ξ) = (2π)−1/2ξ−2k.
which satisfies
A polyharmonic cardinal spline is a function or distribution on R
(i) f ∈ C 2k−2(Rn),
(ii) ∆kf = 0
on Rn \ Zn.
The fundamental functions Lk(x) use φk(ξ) = (2π)−1/2ξ−2k. Several
properties of the fundamental function may be found in [3], including
the fact that it is a k-harmonic spline. We show that {ξ−2k}∞
k=1 is
a regular family of cardinal interpolators. Checking (H1)-(H5) is an
elementary exercise in calculus and will be omitted. We check (R1)
and (R2). For j 6= 0 and ξ < π we have
Mj,k(ξ) =
ξ2k
ξ + 2πj2k ≤ (2j − 1)−2k.
These terms tend to 0, hence (R1) is satisfied. For (R2) we note that
(2j − 1)−2k ≤ (2j − 1)−2.
Thus, if n = 1, polyharminic cardinal splines form a regular family of
cardinal interpolators. For n > 1, we have φk(ξ) = kξk−2k, where ξ ∈
Rn, in this case we must take k ≥ n+1. (H1)-(H4) are straightforward,
so we will check (H5), (R1), and (R2). We need to find the derivatives
REGULAR CARDINAL INTERPOLATORS
11
in question. To this end, we note that for j = 1, . . . , n we have
Repeating this argument for higher derivatives we arrive at
.
d
dξj (cid:0)kξk−2k(cid:1) = −2kkξk−2k−1 ξj
kξk
Dβ(cid:0)kξk−2k(cid:1) = kξk−2k−βΩ(ξ)
where β is a multi-index such that β ≤ n + 1, and Ω(λξ) = Ω(ξ).
This information allows us to check (H5) easily. We move on to (R1)
and (R2). Mimicking the one dimension case we see that
Mj,k(ξ) =(cid:18)
kξ + 2πjk(cid:19)2k
kξk
.
If ξ ∈ (−π, π)n and j 6= 0, then we have kξk < kξ + 2πjk hence these
terms go to 0. Thus we have shown that (R1) holds. For (R2) we
must be a little more careful. In this case, we see that the terms grow
like kjk−2k. Thus we may find a constant and C such that Mj,k(ξ) ≤
Ckjk−2k as long as j 6= 0. Recalling that 2k ≥ n + 1 we have that
Mj,k(ξ) ≤ Ckjk−n−1.
6.2. Gaussians. Our second example is the Gaussian. This example
was studied in [4, 5], although we change the notation somewhat. We
begin with φα(x) = e−x2/(4α), where x ∈ R, thus omitting constants we
have φα(ξ) = e−αξ2
. We take α ≥ 1 as a convenience, the arguments go
through with α ≥ α0 > 0. Again, checking conditions (H1)-(H5), is a
straightforward calculus exercise, whose details are omitted. We check
(R1) and (R2). For ξ < π and j 6= 0, we have
Mj,α(ξ) = eαξ2−α(ξ+2πj)2
.
We can see that these terms go to 0 since the exponent is negative.
Since this is true for all j 6= 0, (R1) holds. As for (R2) we note that if
α ≥ 1 we have
eαπ2−α(ξ+2πj)2
= e−απ2(4j)(j−1) ≤ e−4π2j(j−1).
is a regular family of cardinal interpolators in the
Hencene−x2/(4α)oα≥1
univariate case. In the multivariate case, we have φα(ξ) = e−αkξk2
, and
again consider α ∈ [1,∞) This function is smooth so all of the relevant
properties are easily checked. (D1) and (D2) are checked as above by
replacing the absolute values with norms. Hence, ne−x2/(4α)oα≥1
is a
regular family of cardinal interpolators in the multivariate case as well.
12
JEFF LEDFORD
6.3. Multiquadrics. Our next example is a family of 'multiquadrics',
meaning φ(x) = (c2 +kxk2)α. In this example we have a choice between
which parameter to limit over and which to keep fixed. We will work
both cases. In the univariate case, allowing c to vary is the subject of
[6]. We begin by finding the specialized Fourier transform of φ(x) =
(c2 + kxk2)α. We have
21+α
2
K n
2 +α(ckξk),
ξ 6= 0,
φ(ξ) =
Γ(−α)(cid:18)kξk
where
Kβ(r) =
e−r cosh teβtdt,
c (cid:19)−α− n
2ZR
1
for r > 0 and β ∈ R. The details of this may be found in Theorem 8.15
of [8]. We must omit the case when α ∈ {0, 1, 2, . . .} because in this
case we would get a measure in the transform space and we would like a
function. This is why we use the specialized Fourier transform and not
the more standard generalization. We first consider the univariate case
and list some properties of Kβ(r) which will be useful is our analysis.
These properties may be found on page 52 of [8]. The first is that Kβ(r)
is decreasing on (0,∞). We find the following asymptotic expansions
in [1]:
(10)
e−reβ2/(2r)
1
√r
Kβ(r) ∼
Kβ(r) ∼ r−β
(11)
Here f ∼ g means that f = (C + o(1))g, where C is a constant in-
dependent of the limiting parameter. We also have the differentiation
formula
(r → ∞),
(r → 0).
(12)
d
dz(cid:2)zβKβ(z)(cid:3) = −zβKβ−1(z).
2 and r > 0
And because we will need it later we have the following bounds for
β ≥ 1
(13)
√2πr−1/2e−re−β2/(2r)
r−1/2e−r ≤ Kβ(r) ≤
r π
2
All of this information may be found in [1] and Chapter 5 of [8], and has
been restated here for convenience and to aide in the calculations that
follow. We will consider two different families of multiquadrics. In the
first example we fix a positive number c and take {αj}∞
2,∞),
αj = ∞. We show
such that dist({αj},{1, 2, 3, . . .}) > 0 and lim
{φαj (x)}∞
j=1 is a regular family of cardinal interpolators. We begin
j=1 ⊂ [ 1
j→∞
REGULAR CARDINAL INTERPOLATORS
13
in the univariate case checking (H1)-(H5). (H1)-(H4) follow from the
definition, the differentiation formula, and the asymptotic estimates.
(H5) also follows from the asymptotic estimates, but we will show these
calculations. First, we find the derivatives in question omitting the
associated constants since they have no bearing on the overall result.
αj (ξ) = − ξ−αj− 1
φ(1)
αj (ξ) =ξ−αj− 1
φ(2)
2 Kαj + 3
2
(cξ)
(cξ)
Using the asymptotic expansion, we find the following limits.
(cξ) + ξ−αj− 3
2 Kαj + 3
2 Kαj− 5
2
2
1
φαj (ξ)
φ(1)
αj (ξ)
φ2
αj (ξ)
φ(2)
αj (ξ)
φ2
αj (ξ)
= 0
2
2
= ξαj+ 1
Kαj + 1
= ξ2αj+1
C + o(1)
(cξ)
= ξαj+ 1
(cξ)
= ξ2αj−1(cid:26) C3 + o(1)
Kαj + 3
K 2
αj + 1
2
2
2
[C1 + o(1)]2 −
= ξ2αj C2 + o(1)
[C1 + o(1)]2 = 0
[C1 + o(1)]2(cid:27)
C2 + o(1)
(ξ → 0)
(ξ → 0)
(ξ → 0)
1
2
We can see that in the last expression we must take αj ≥
in order
for the limit to be bounded. Now we turn to (R1). Let ξ < π and
i 6= 0, then
Mi,αj (ξ) ≤
παj+ 1
2
Kαj + 1
2
(cπ)
Kαj + 1
2
(cξ + 2πi)
ξ + 2πiαj+ 1
2 ≤
1
.
(2i − 1)αj+ 1
2
j=1 ⊂ [ 1
We have used the fact that Kβ(r) is decreasing. This estimate allows us
to prove (D2) as well, since the condition for every α may be replaced
with for every α large enough. In this case, large enough means αj > 1
2.
Thus we have that {(c2 + x2)αj} is a well behaved family for fixed
c > 0 and{αj}∞
2,∞), such that dist({αj},{1, 2, 3, . . .}) > 0 and
αj = ∞. Before moving on, we make note of the specific example
lim
j→∞
αk = 2k−1
for k ∈ {1, 2, 3, . . .}. This is the family of odd powers
}. This family corresponds to
'smoothed out' polyharmonic splines of order 2k. In light of the results
in [3], one may well expect that our family of multiquadrics shares a
similar convergence property. This is indeed the case as our work above
shows.
of the multiquadric {(cid:16)pc2 + x2(cid:17)2k−1
2
We now explore what happens if we fix α and allow c to vary. Again
we will need to bound the parameter away from the origin so we take
c ≥ 1. Let φc(x) = (c2 + x2)α for some fixed α ∈ (−∞,− 3
2,∞) \
2] ∪ [ 1
14
JEFF LEDFORD
{1, 2, 3, . . .}, then, omitting constants because they will divide out in
the fundamental function, we have φc(ξ) = ξ−α− 1
2 +α(cξ). The
bounds on α are a result of the restrictions imposed by the derivatives.
Since φc(0) is a real positive number if α < − 1
2, (H5) places restrictions
on the growth of φc(ξ) near the origin. To satisfy this condition, we
need to increase the rate of decay of φc(x). Most of the work to check
that this is a regular family of cardinal interpolators has already been
done. The same arguments used above can be recycled, however we
must check (R1) and (R2) separately. We check (R1) presently.
2 K 1
Mj,c(ξ) ≤
2
2
πα+ 1
2
Kα+ 1
(cξ + 2πj)
Kα+ 1
Kα+ 1
(cπ)
(cξ + 2πj)
Kα+ 1
ξ + 2πjα+ 1
≤ (2j − 1)−(α+ 1
≤ (2j − 1)−(α+1) ecπe−cξ−2πje−(α+ 1
≤ (2j − 1)−(α+1) ecπe−cξ−2πj
(cπ)
2 )
2
2
2
2 )2/(2cξ+2πj)
We can see, for j 6= 0 and ξ < π, that lim
Mj,c(ξ) = 0. This
calculation also shows that (R2) is satisfied since each term there
is bounded by (2j − 1)−(α+1)e−2π(j−1). Thus we have shown that
{(c2 + x2)α}c≥1 is a well behaved family of interpolators for a fixed
α ∈ (−∞,− 3
2 is examined
in [6].
2 ,∞)\{1, 2, 3, . . .}. The case where α = 1
2]∪[ 1
c→∞
We now visit the n-dimensional multiquadric φ(x) = (c2 + kxk2)α.
We will again consider both the case when α is the parameter and
c is fixed and the case when c is the parameter and α is fixed. We
begin by recalling the specialized n-dimensional Fourier transform of
the multiquadric,
φ(ξ) =
21+α
Γ(−α)(cid:18)kξk
c (cid:19)−α− n
2
K n
2 +α(ckξk),
ξ 6= 0.
Conditions (H1)-(H4) follow as before. We need to calculate the deriva-
tives in order to check (H5). To this end, we may use the differentiation
formula (12) together with repeated use of the product rule and the as-
ymptotic estimate (11) to see that near the origin we have the following
asymptotic estimate
(14)
Dβ φ(ξ) ∼ kξk−α− n
2 −α+ n
2 +β,
kξk → 0.
REGULAR CARDINAL INTERPOLATORS
15
Recall that f ∼ g means that f = g(C + o(1)) for an appropriate
constant C. This calculation holds no matter which parameter we are
considering fixed. Thus we can begin by fixing c > 0 and checking
(H5), (R1) and (R2) for φα(x). For (H5) we note that α > 0, hence
the asymptotic estimate (14) becomes
Dβ φ(ξ) ∼ kξk−2α−n−β,
kξk → 0.
Using this estimate, checking (H5) is now a routine matter. We note
that we must have α ≥ 1
Mj,α(ξ) =(cid:18)
2. To see (R1), note that
kξk
kξ + 2πjk(cid:19)−α− n
2 (ckξ + 2πjk)
Kα+ n
2 (ckξk)
2 Kα+ n
,
and if ξ ∈ (−π, π)n we have
lim
α→∞
α→∞(cid:18)
Mj,α(ξ) ≤ lim
kξk
kξ + 2πjk(cid:19)−α− n
2
.
These terms tend to 0 as α increases. As in the one variable case, we
must be careful about how we choose the parameter α, in particular
we take α ∈ {αj} ⊂ [ 1
2,∞) such that dist({αj},{0, 1, 2, 3, . . .}) > 0
αj = ∞. For the bounding sequence in (R2) we take α = n+1
2 .
and lim
j→∞
These terms grow like kjk−n−1, hence we may find a constant C such
that
Mj,α(ξ) ≤ Ckjk−n−1.
This bound holds for all αj ≥ n+1
We now will fix α and let c → ∞. That is, we consider φc(x) =
(c2 + kxk2)α. (H1)-(H4) hold as before. We take advantage of (14) to
find that if α ∈ (−∞,− 2n+1
2 ,∞) \ {1, 2, 3, . . .} then (H5) holds.
As for (R1) and (R2), we must be a bit more careful. We have
2 , hence (R2) holds.
] ∪ [ 1
2
and we can see from (13) that
Mj,c(ξ) =(cid:18)
Mj,c(ξ) ≤ C(cid:18)
2 Kα+ n
kξk
kξ + 2πjk(cid:19)−α− n
kξ + 2πjk(cid:19)−α− n−1
kξk
2
2 (ckξ + 2πjk)
Kα+ n
2 (ckξk)
,
ec(kξk−kξ+2πjk).
Here C is a positive constant independent of c whose particular value is
unimportant. If ξ ∈ (−π, π)n, then Mj,c(ξ) → 0 because the exponent
in the exponential is negative. To check (R2), note that these terms
decay exponentially hence we can find a bounding sequence which is in
l1(Zn \ {0}).
16
JEFF LEDFORD
References
[1] M. Abramowitz and I. Stegun, eds., Handbook of Mathematical Functions with
Formulas, Graphs, and Mathematical Tables, Dover, New York, New York,
1965 pp. 375-378.
[2] B.J.C. Baxter, The asymptotic cardinal function of the multiquadratic ϕ(r) =
[3] W. Madych, S. Nelson, Polyharmonic Cardinal Splines, J. Approximation The-
(r2 + c2)1/2 as c → ∞, Comput. Math. Appl. 24 (1992), no. 12, 1-6.
ory 60 (1990), 141-156.
[4] S. Riemenschneider and N. Sivakumar, Gaussian radial-basis functions: cardi-
nal interpolation of lp and power-growth data. Radial basis functions and their
applications Adv. Comput. Math. 11, (1999), 229-251.
[5] S. Riemenschneider and N. Sivakumar, On cardinal interpolation by Gaussian
radial-basis functions: properties of fundamental functions and estimates for
Lebesgue constants J. Anal. Math. 79, (1999), 33-61.
[6] S. Riemenschneider and N. Sivakumar, On the Cardinal-Interpolation Operator
Associated with the One-Dimensional Multiquadric East J. Approx. 7, (2001),
no. 4, 485-514.
[7] I.J. Schoenberg, Cardinal Spline Interpolation SIAM, Philadelphia, PA, 1973.
[8] H. Wendland, Scattered Data Approximation, Cambridge University Press,
Cambridge, UK, 2005.
|
1312.0986 | 2 | 1312 | 2016-12-12T11:09:33 | The short-time Fourier transform of distributions of exponential type and Tauberian theorems for shift-asymptotics | [
"math.FA"
] | We study the short-time Fourier transform on the space $\mathcal{K}_{1}'(\mathbb{R}^n)$ of distributions of exponential type. We give characterizations of $\mathcal{K}_{1}'(\mathbb{R}^n)$ and some of its subspaces in terms of modulation spaces. We also obtain various Tauberian theorems for the short-time Fourier transform. | math.FA | math |
THE SHORT-TIME FOURIER TRANSFORM OF DISTRIBUTIONS
OF EXPONENTIAL TYPE AND TAUBERIAN THEOREMS FOR
SHIFT-ASYMPTOTICS
SANJA KOSTADINOVA, STEVAN PILIPOVI ´C, KATERINA SANEVA, AND JASSON VINDAS
Abstract. We study the short-time Fourier transform on the space K′
1(Rn) of dis-
tributions of exponential type. We give characterizations of K′
1(Rn) and some of its
subspaces in terms of modulation spaces. We also obtain various Tauberian theorems
for the short-time Fourier transform.
1. Introduction
The short-time Fourier transform (STFT) is a very effective device in the study
of function spaces. The investigation of major test function spaces and their duals
through time-frequency representations has attracted much attention. For example,
the Schwartz class S(Rn) and the space of tempered distributions S ′(Rn) were studied
in [10] (cf. [9]). Characterizations of Gelfand-Shilov spaces and ultradistribution spaces
by means of the short-time Fourier transform and modulation spaces are also known
[11, 17, 26] (cf. [4, 5]).
The purpose of this paper is two folded. On the one hand we study the short-time
Fourier transform in the context of the space K′
1(Rn) of distributions of exponential
type, the dual of the space of exponentially rapidly decreasing smooth functions K1(Rn)
(see Section 2 for the definition of all spaces employed in this article). We will obtain
various characterizations of K′
1(Rn) and related spaces via the short-time Fourier trans-
form. The space K′
1(Rn) was introduced by Silva [24] and Hasumi [12] in connection
with the so-called space of Silva tempered ultradistributions U ′(Cn). Let us mention
1(Rn) and U ′(Rn) were also studied by Morimoto through the theory of ultra-
that K′
hyperfunctions [16] (cf.
[18]). We refer to [7, 14, 25, 30] for some applications of the
Silva spaces. Our second goal is to present a new kind of Tauberian theorems. In such
theorems the exponential asymptotics of functions and distributions can be obtained
from those of the short-time Fourier transform. Let us state a sample of our results. In
the next statement L stands for a locally bounded Karamata slowly varying function
2010 Mathematics Subject Classification. Primary 81S30, 40E05; Secondary 26A12, 41A60, 46F05,
46F12.
Key words and phrases. short-time Fourier transform; Tauberian theorems; modulation spaces;
1 and K1; distributions of exponential type; Silva tempered ultradistributions (tempered ultra-
K′
hyperfunctions); S-asymptotics; regularly varying functions.
Research supported by the project 174024 of the Ministry of Education and Sciences of Serbia.
S. Pilipovi´c and J. Vindas also acknowledge support from Ghent University, through the BOF-grant
number 01T00513.
1
2
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
[2, 15], namely, a positive function that is asymptotically self-similar in the sense:
lim
x→∞
L(ax)
L(x)
= 1,
∀a > 0.
Theorem 1.1. Let f be a positive non-decreasing function on [0, ∞) and let ψ be a
positive function such that ψ′′ ∈ L1
−∞(ψ(t) + ψ′(t) + ψ′′(t))eβt+εtdt < ∞,
where β ≥ 0 and ε > 0. Suppose that the limits
loc(R) andR ∞
(1.1)
lim
x→∞
f (t)ψ(t − x)e−2πiξt dt = J(ξ)
e2πiξx
eβxL(ex)Z ∞
0
exist for every ξ ∈ R, then
f (x)
=
J(0)
.
(1.2)
lim
x→∞
eβxL(ex)
−∞ ψ(t)eβtdt
R ∞
the requirements over ψ can be relaxed toR ∞
Furthermore, if L satisfies L(xy) ≤ AL(x)L(y) for all x, y > 0 and some constant A,
−∞(ψ(t) + ψ′(t) + ψ′′(t))L(et)eβtdt < ∞.
It turns out that Theorem 1.1 can be deduced from a more general type of Tauberian
theorems. In Section 6 we shall give precise descriptions of the S-asymptotic properties
[21] of a distribution in terms of the asymptotic behavior of its short-time Fourier
transform (S-asymptotics stands for shift-asymptotics). The notion of S-asymptotics
measures the asymptotic behavior of the translates of a distribution T−hf with respect
the parameter h and it is closely related to the extension of Wiener's Tauberian ideas
[15, 29] to the context of Schwartz distributions [20]. We would like to point out that
there is an extensive literature about Tauberian theorems for Schwartz distributions,
see, e.g., the monographs [21, 28] and references therein. We also mention the article
[22], where Tauberian theorems for the short-time Fourier transform were also studied,
though with a different approach.
The plan of this article is as follows. In Section 3 we shall present continuity theorems
for the STFT and its adjoint on the test function space K1(Rn) and the topological
1(Rn). We also introduce in this paper the space B′
tensor product K1(Rn)b⊗U(Cn), where U(Cn) is the space of entire rapidly decreasing
functions in any horizontal band of Cn. We then use such continuity results to develop a
framework for the STFT on K′
ω(Rn)
of ω-bounded distributions and its subspace B′
ω(Rn) with respect to an exponentially
moderate weight ω; when ω = 1, these spaces coincide with the well-known Schwartz
spaces [23, p. 200] of bounded distributions B′(Rn) and B′(Rn), which are of great
importance in the study of convolution and growth properties of distributions. Notice
that the distribution space B′(Rn) also plays an important role in Tauberian theory; see,
for instance, Beurling's theorem [6, p. 230] and the distributional Wiener Tauberian
theorem from [20]. The spaces B′
ω(Rn) will be characterized in Section
4 in terms of the short-time Fourier transform and also in terms properties of the
set of translates of their elements. Section 5 is devoted to the characterization of
K′
1(Rn) and related spaces via modulation spaces. The conclusive Section 6 deals with
Tauberian theorems, where in particular we give a proof of Theorem 1.1. Our Tauberian
ω(Rn) and B′
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
3
hypotheses are actually in terms of membership to suitable modulation spaces, this
allows us to reinterpret the S-asymptotics in the weak∗ topology of modulation spaces.
2. Preliminaries
2.1. Notation. We use the constants in the Fourier transform as
(2.1)
e−2πix·ξϕ(x)dx.
F (ϕ)(ξ) = bϕ(ξ) =ZRn
The translation and modulation operators are defined by Txf ( · ) = f ( · − x) and
Mξf ( · ) = e2πiξ · f ( · ), x, ξ ∈ Rn. The operators MξTx and TxMξ are called time-
frequency shifts and we have MξTx = e2πix·ξTxMξ. The notation hf, ϕi means dual
pairing whereas (f, ϕ)L2 stands for the L2 inner product. All dual spaces in this
article are equipped with the strong dual topology. We denote by f the function
(or distribution) f (t) = f (−t).
2.2. The STFT. The short-time Fourier transform (STFT) of a function f ∈ L2(Rn)
with respect to a window function ψ ∈ L2(Rn) is defined as
(2.2)
f (t)ψ(t − x)e−2πiξ·t dt, x, ξ ∈ Rn.
There holds kVψf k2 = kf k2 kψk2. The adjoint of Vψ is given by the mapping
Vψf (x, ξ) = hf, MξTxψi =ZRn
ψ F (t) =ZZR2n
V ∗
F (x, ξ)ψ(t − x)e2πiξ·tdxdξ,
interpreted as an L2(Rn)-valued weak integral. If ψ 6= 0 and γ ∈ L2(Rn) is a synthesis
window for ψ, namely, (γ, ψ)L2 6= 0, then for any f ∈ L2(Rn),
(2.3)
f =
Vψf (x, ξ)MξTxγdξdx.
1
(γ, ψ)L2ZZR2n
Whenever the dual pairing in (2.2) is well-defined, the definition of Vψf can be
generalized for f in larger classes than L2(Rn), for instance: f ∈ D′(Rn) and ψ ∈
D(Rn). In fact, it is enough to have ψ ∈ A(Rn) and f ∈ A′(Rn), where A(Rn) is a
time-frequency shift invariant topological vector space. Note also that the inversion
formula (2.3) holds pointwise when f is sufficiently regular, for instance, for function
in the Schwartz class S(Rn). For a complete account on the STFT, we refer to [9].
2.3. Spaces. The Hasumi-Silva [24, 12] test function space K1(Rn) consists of those
ϕ ∈ C ∞(Rn) for which all norms
νk(ϕ) := sup
t∈Rn, α≤k
ektϕ(α)(t),
k ∈ N0,
are finite. The elements of K1(Rn) are called exponentially rapidly decreasing smooth
It is easy to see that K1(Rn) is an FS-space and therefore Montel and
functions.
reflexive. The space K1(Rn) is also nuclear [12].
Note that if ϕ ∈ K1(Rn), then the Fourier transform (2.1) extends to an entire
In fact, the Fourier transform is a topological isomorphism from K1(Rn)
function.
4
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
onto U(Cn), the space of entire functions which decrease faster than any polynomial
in bands. More precisely, a entire function φ ∈ U(Cn) if and only if
(1 + z2)k/2φ(z) < ∞, ∀k ∈ N0,
νk(φ) := sup
z∈Πk
where Πk is the tube Πk = Rn + i[−k, k]n.
The dual space K′
1(Rn) consists of all distributions f of exponential type, i.e., those
of the form f =Pα≤l(es · fα)(α), where fα ∈ L∞(Rn) [12]. The Fourier transform
extends to a topological isomorphism F : K′
1(Rn) → U ′(Cn), the latter space is known
as the space of Silva tempered ultradistributions [12] (also called the space of tempered
ultra-hyperfunctions [16]). The space U ′(Cn) contains the space of analytic functionals.
See also the textbook [14] for more information about these spaces.
We introduce a generalization of the Schwartz space of bounded distributions B′(Rn)
[23, p. 200]. Let ω : Rn → (0, ∞) be an exponentially moderate weight, namely, ω is
measurable and satisfies the estimate
(2.4)
ω(x + y) ≤ Aω(y)eax,
x, y ∈ Rn,
for some constants A > 0 and a ≥ 0. For instance, any positive measurable function
ω which is submultiplicative, i.e., ω(x + y) ≤ ω(x)ω(y), and integrable near the origin
must necessarily satisfy (2.4), as follows from the standard results about subadditive
functions [1, 13]. Extending the Schwartz space DL1(Rn), we define the Fr´echet space
0 }, provided with the family of
DL1
norms
ω (Rn) = {ϕ ∈ C ∞(Rn) : ϕ(α) ∈ L1
ω(Rn), ∀α ∈ Nn
kϕk1,ω,k := sup
ϕ(α)(t)ω(t)dt,
k ∈ N0.
α≤kZRn
ω(Rn) stands for the strong dual of DL1
Then, B′
we have the dense embedding K1(Rn) ֒→ DL1
ω(Rn) the space of ω-bounded distributions. We also define B′
B′
D(Rn) in B′
ω (Rn), we have B′
ω (Rn), i.e., B′
ω(Rn).
ω(Rn) = (DL1
ω(Rn) ⊂ K′
ω(Rn))′. Since
1(Rn). We call
ω(Rn) as the closure of
norms
tained as the completion of K1(Rn) ⊗ U(Cn) in, say, the π- or the ε- topology [27].
Next, we shall consider K1(Rn)b⊗U(Cn), the topological tensor product space ob-
Explicitly, the nuclearity of K1(Rn) implies that K1(Rn)b⊗U(Cn) = K1(Rn)b⊗πU(Cn) =
K1(Rn)b⊗εU(Cn). Thus, the topology of K1(Rn)b⊗U(Cn) is given by the family of the
and we also obtain (K1(Rn)b⊗U(Cn))′ = K′
Finally, let m be a weight on R2n, that is, m : R2n → (0, ∞) is measurable and
m (R2n) consists
∂xα Φ(x, z)(cid:12)(cid:12)(cid:12)(cid:12) , k ∈ N0,
locally bounded. Then, if p, q ∈ [1, ∞], the weighted Banach space Lp,q
of all measurable functions F such that
(x,z)∈Rn×Πk, α≤k
ρk(Φ) :=
sup
∂α
ekx(1 + z2)k/2(cid:12)(cid:12)(cid:12)(cid:12)
1(Rn)b⊗U ′(Cn).
F (x, ξ)pm(x, ξ)pdx(cid:19)q/p
kF kLp,q
m := ZRn(cid:18)ZRn
dξ!1/q
< ∞.
(With the obvious modification when p = ∞ or q = ∞.)
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
5
3. Short-time Fourier transform of distributions of exponential type
In this section we study the mapping properties of the STFT on the space of distri-
butions of exponential type. Note that the STFT extends to the sesquilinear mapping
(f, ψ) 7→ Vψf and its adjoint induces the bilinear mapping (F, ψ) 7→ V ∗
We start with the test function space K1(Rn). If f, ψ ∈ K1(Rn), then we immedi-
ately get that (2.2) extends to a holomorphic function in the second variable, namely,
Vψf (x, z) is entire in z ∈ Cn. We write in the sequel z = ξ + iη with ξ, η ∈ Rn. Observe
ψ ∈ K1(Rn), then for arbitrary η ∈ Rn we may write V ∗
also that an application of the Cauchy theorem shows that if Φ ∈ K1(Rn)b⊗U(Cn) and
ψ Φ as
ψ F .
V ∗
ψ Φ(t) =ZZR2n
(3.1)
Φ(x, ξ + iη)ψ(t − x)e2πi(ξ+iη)·tdxdξ.
Our first proposition deals with the range and continuity properties of V and V ∗ on
test function spaces.
Proposition 3.1. The following mappings are continuous:
∂α
Proof. For part (i), let ϕ, ψ ∈ K1(Rn). Let k be an even integer. If (x, z) ∈ Rn × Πk
and α ≤ k, then
(i) V : K1(Rn) × K1(Rn) → K1(Rn)b⊗U(Cn).
(ii) V ∗ : (K1(Rn)b⊗U(Cn)) × K1(Rn) → K1(Rn).
∂xα Vψϕ(x, z)(cid:12)(cid:12)(cid:12)(cid:12)
(1 − ∆t)k/2(ϕ(t)ψ(α)(t − x)e2πη·t)dt(cid:12)(cid:12)(cid:12)(cid:12)
ekxZRn(cid:12)(cid:12)(cid:12)ϕ(β1)(t)ψ(α+β2)(t − x)(cid:12)(cid:12)(cid:12) e2πktdt,
ekx(1 + z2)k/2(cid:12)(cid:12)(cid:12)(cid:12)
≤ (1 + nk2)k/2ekx(cid:12)(cid:12)(cid:12)(cid:12)ZRn
≤ Ck Xβ1+β2≤k
ψ Φ(t)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (2π)αXβ≤α(cid:18)α
≤ (4π)ανk(ψ)ZZR2n
ξkekxΦ(x, ξ)dxdξ
∂α
∂tα V ∗
K1(Rn), and α ≤ k, we obtain
which shows that ρk(Vψϕ) ≤ Ckν8k(ϕ)νk(ψ). For (ii), if Φ ∈ K1(Rn)b⊗U(Cn), ψ ∈
ekt(cid:12)(cid:12)(cid:12)(cid:12)
β(cid:19)ektZZR2n
ξkΦ(x, ξ)ψ(β)(t − x)dxdξ
≤ Ak,nνk(ψ)ρk+n+1(Φ);
hence ρk(V ∗
ψ Φ) ≤ Ak,nνk(ψ)ρk+n+1(Φ).
(cid:3)
Observe that if the window ψ ∈ K1(Rn) \ {0} and γ ∈ K1(Rn) is a synthesis window,
the reconstruction formula (2.3) reads as:
(3.2)
1
(γ, ψ)L2
V ∗
γ Vψ = idK1(Rn).
6
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
We now study the STFT on K′
1(Rn). Notice that the modulation operators Mz
operate continuously on K1(Rn) even when z ∈ Cn. Thus, if f ∈ K′
1(Rn) and ψ ∈
K1(Rn) then Vψf , defined by the dual pairing in (2.2), also extends in the second
variable as an entire function Vψf (x, z) in z ∈ Cn. Furthermore, it is clear that
Vψf (x, z) is C ∞ in x ∈ Rn. We begin with a lemma.
Lemma 3.2. Let ψ ∈ K1(Rn).
(a) Let B′ ⊂ K′
1(Rn) be a bounded set. There is k = kB′ ∈ N0 such that
(3.3)
sup
f ∈B′, (x,z)∈Rn×Πλ
e−kx−2πx·ℑm z(1 + z)−kVψf (x, z) < ∞,
∀λ ≥ 0.
Proof. Part (a). By the Banach-Steinhaus theorem, B′ is equicontinuous, so that there
are C > 0 and k ∈ N0 such that hf, ϕi ≤ Cνk(ϕ), ∀f ∈ B′, ∀ϕ ∈ K1(Rn). Hence, for
all f ∈ B′ (z = ξ + iη),
Vψf (x, z) ≤ C sup
(b) For every f ∈ K′
(3.4)
1(Rn) and Φ ∈ K1(Rn)b⊗U(Cn),
hVψf, Φi =Df, V ∗
ψ ΦE .
∂tα(cid:16)e−2πiz·tψ(t − x)(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)
ekt(cid:12)(cid:12)(cid:12)(cid:12)
ekt+2πη·tXβ≤α(cid:18)α
t∈Rn,α≤k
sup
∂α
t∈Rn,α≤k
≤ (2π)kC(1 + z2)k/2
≤ (4π)kC(1 + z2)k/2ekx+2πη·xνk+1+⌊2πη⌋(ψ),
β(cid:19)(cid:12)(cid:12)ψ(β)(t − x)(cid:12)(cid:12)
where ⌊2πη⌋ stands for the integral part of 2πη.
Part (b). We first remark that the left hand side of (3.4) is well defined because
of part (a). To show (3.4), notice that the integral in (3.1), with η = 0, can be
approximated by a sequence of convergent Riemann sums in the topology of K1(Rn);
this justifies the exchange of integral and dual pairing in
(cid:28)f (t),ZZR2n
Φ(x, ξ)e−2πiξ·tψ(t − x)dxdξ(cid:29)t
=ZZR2n
which is the same as (3.4).
Φ(x, ξ)hf, MξTxψidxdξ,
(cid:3)
In particular, if B′ is a singleton, part (a) of Lemma 3.2 gives the growth order of
the function Vψf on every set Rn × Πλ.
Let us define the adjoint STFT on ∈ K′
1(Rn)b⊗U ′(Cn).
Definition 3.3. Let ψ ∈ K1(Rn). The adjoint STFT V ∗
the distribution V ∗
ψ F ∈ K′
ψ of F ∈ K′
1(Rn) whose action on test functions is given by
hV ∗
ψ F, ϕi :=(cid:10)F, Vψϕ(cid:11) , ϕ ∈ K1(Rn).
(3.5)
The next theorem summarizes our results.
Theorem 3.4. The two STFT mappings
1(Rn)b⊗U ′(Cn) is
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
7
(i) V : K′
(ii) V ∗ : (K′
1(Rn) × K1(Rn) → K′
1(Rn)b⊗U ′(Cn)
1(Rn)b⊗U ′(Cn)) × K1(Rn) → K′
1(Rn)
are hypocontinuous. Let ψ ∈ K1(Rn) \ {0} and let γ ∈ K1(Rn) be a synthesis window
for it. The following inversion and desingularization formulas hold:
(3.6)
1
(γ, ψ)L2
V ∗
γ Vψ = idK′
1(Rn),
and, for all f ∈ K′
1(Rn), ϕ ∈ K1(Rn), and η ∈ Rn,
(3.7)
hf, ϕi =
Vψf (x, ξ + iη)Vγϕ(x, −ξ − iη)dxdξ.
1
(γ, ψ)L2ZZR2n
Proof. That V and V ∗ are hypocontinuous on these spaces follows from Proposition
3.1 and the formula (3.4) from Lemma 3.2; we leave the details to the reader. By the
Cauchy theorem, it is enough to show (3.7) for η = 0. Using (3.5), (3.4), and (3.2),
we have hV ∗
ψ Vγϕi = (γ, ψ)L2hf, ϕi, namely, (3.6) and
(3.7).
(cid:3)
γ Vψf, ϕi = hVψf, Vγϕi = hf, V ∗
The next corollary gives the converse to part (a) of Lemma 3.2 under a weaker
1(Rn) in terms
inequality than (3.3), namely, a characterization of bounded sets in K′
of the STFT.
Corollary 3.5. Let B′ ⊂ K′
k ∈ N0 such that
1(Rn) and ψ ∈ K1(Rn) \ {0}. If there are η ∈ Rn and
(3.8)
sup
f ∈B′,(x,ξ)∈R2n
e−kx(1 + ξ)−kVψf (x, ξ + iη) < ∞,
then the set B′ is bounded in K′
k ∈ N0 such that (3.3) holds.
1(Rn). Conversely, if B′ is bounded in K′
1(Rn) there is
Proof. In view of the Banach-Steinhaus theorem, we only need to show that B′ is
weakly bounded. Let γ be a synthesis window for ψ and let ϕ ∈ K1(Rn). Then, by the
desingularization formula (3.7), we have
hf, ϕi ≤
sup
f ∈B′
Cη
(γ, ψ)L2ZZR2n
ekx(1 + ξ)k Vγϕ(x, −ξ − iη) dxdξ < ∞,
because Vγϕ ∈ K1(Rn)b⊗U(Cn). The converse was already shown in Lemma 3.2.
4. Characterizations of B′
ω(Rn) and B′
ω(Rn)
(cid:3)
We now turn our attention to the characterization of the space of ω-bounded dis-
ω(Rn). Recall that ω stands for an exponentially
tributions B′
moderate weight, i.e., a positive and measurable function satisfying (2.4).
ω(Rn) and its subspace B′
Theorem 4.1. Let f ∈ K′
1(Rn) and ψ ∈ K1(Rn) \ {0}.
(i) The following statements are equivalent:
(a) f ∈ B′
(b) The set {T−hf /ω(h) : h ∈ Rn} is bounded in K′
ω(Rn).
1(Rn).
8
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
(c) There is s ∈ R such that
(4.1)
sup
(x,ξ)∈R2n
(1 + ξ)−s Vψf (x, ξ)
ω(x)
< ∞.
(ii) The next three conditions are equivalent:
ω(Rn).
(a)′ f ∈ B′
(b)′ limh→∞ T−hf /ω(h) = 0 in K′
(c)′ There is s′ ∈ R such that
1(Rn).
(4.2)
lim
(x,ξ)→∞
(1 + ξ)−s′ Vψf (x, ξ)
= 0.
ω(x)
Remark 4.2. Theorem 4.1 remains valid if we replace K′
1(Rn) and K1(Rn) by D′(Rn)
and D(Rn) everywhere in the statement. Schwartz has shown in [23, p. 204] the
equivalence between (a) and (b) for ω = 1 by using a much more complicated method
involving a parametrix technique.
Proof. Part (i). (a) ⇒ (b). Let f ∈ B′
ω(Rn), since K1(Rn) is barreled, we only need to
show that the f ∗ ϕ is bounded by ω for fixed ϕ ∈ K1(Rn). Let B := {φ ∈ D(Rn) :
RRn φ(x)ω(x)dx ≤ 1}. By the assumption (2.4),
α≤kZRn
k ϕ ∗ φk1,ω,k ≤ A max
ϕ(α)(x)eaxdx,
∀k ∈ N0, ∀φ ∈ B,
namely, the set ϕ ∗ B is bounded in DL1
supφ∈B hf, ϕ ∗ φi < ∞. Since D(Rn) is dense in L1
(L1
ω(Rn). Consequently, supφ∈B hf ∗ ϕ, φi =
ω(Rn), this implies that f ∗ ϕ ∈
ω(Rn))′, i.e., suph∈Rn (f ∗ ϕ)(h)/ω(h) < ∞, as claimed.
(b) ⇒ (c). Notice that (VψT−hf )(x, z) = e2πiz·hVψf (x + h, z). Fix λ ≥ 0. By
Corollary 3.5 (cf. (3.3)), there are k ∈ N0 and Cλ > 0 such that, for all x, h ∈ Rn and
z ∈ Πλ,
e2πiz·hVψf (x + h, z) ≤ Cλω(h)(1 + z)ke(k+2πλ)x.
Taking x = 0 and ℑm z = 0, one gets (4.1).
(c) ⇒ (a). Fix a synthesis window γ ∈ K1(Rn). In view of (2.4), one has that if j is
any non-negative even integer and λ ≥ 0, then, for all ϕ ∈ DL1
ω(Rn),
We may assume that s is an even integer. By (4.1) and the previous estimate, we
obtain, for every ϕ ∈ K1(Rn),
hf, ϕi ≤
(1 + ξ)sω(x) Vγϕ(x, −ξ) dxdξ ≤ Cskϕk1,ω,s+n+1,
e−2πx·ℑm zω(x)Vγϕ(x, z)dx
ω(x)ϕ(β1)(t)γ(β2)(t − x)e2πλt−xdtdx
γ(β)(x)e(2πλ+a)xdx ≤ Cj,λkϕk1,ω,j.
sup
z∈Πλ
(1 + z2)j/2ZRn
≤ Cj Xβ1+β2≤jZZR2n
β≤jZRn
≤ A Cjkϕk1,ω,j max
C
(ψ, γ)L2ZZR2n
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
9
which yields f ∈ B′
ω(Rn).
Part (ii). Any of the conditions implies that f ∈ B′
ω(Rn). (a)′ ⇒ (b)′. Fix ϕ ∈
K1(Rn). Given fixed ε > 0, we must show that lim suph→∞ hT−hf, ϕi/ω(h) ≤ ε.
ω(Rn). Since f is in the
Notice that {Thϕ/ω(h) : h ∈ Rn} is a bounded set in DL1
closure of D(Rn) in B′
ω(Rn), there is φ ∈ D(Rn) such that hT−h(f − φ), ϕ ≤ εω(h) for
every h ∈ Rn. Consequently,
lim sup
h→∞
hT−hf, ϕi
ω(h)
≤ ε + lim
h→∞
1
ω(h)(cid:12)(cid:12)(cid:12)(cid:12)ZRn
ϕ(t − h)φ(t)dt(cid:12)(cid:12)(cid:12)(cid:12) ≤ ε.
(b)′ ⇒ (c)′. If ξ remains on a compact of K ⊂ Rn, then {Mξψ : ξ ∈ K} is compact
in K1(Rn), thus, by the Banach-Steinhaus theorem,
0 = lim
x→∞
hT−xf, Mξψi
ω(x)
= lim
x→∞
Vψf (x, ξ)
ω(x)
, uniformly in ξ ∈ K.
There is s such that (4.1) holds. Taking into account that the above limit holds for
arbitrary K, we obtain that (4.2) is satisfied for any s′ > s.
(c)′ ⇒ (a)′. We may assume that s′ is a non-negative even integer. Consider the
weight ωs′(x, ξ) = ω(x)(1 + ξ)s′. The limit relation (4.2) implies that Vψf is in the
closure of K1(Rn) ⊗ S(Rn) with respect to the norm k kL∞,∞
. Since we have the
dense embedding U(Cn) ֒→ S(Rn), there is a sequence {Φj}∞
that limj→∞ Φj = Vψf in L∞,∞
φj = V ∗
for any ϕ ∈ K1(Rn),
1/ωs′ (R2n). Let γ ∈ K1(Rn) be a synthesis window and set
γ Φj ∈ K1(Rn) (cf. Proposition 3.1). By the relations (3.7) and (3.5), we have
j=1 ⊂ K1(Rn)b⊗U(Cn) such
1/ω
s′
hf − φj, ϕi ≤
Ckϕk1,ω,s+n+1
(γ, ψ)L2
kVψf − ΦjkL∞,∞
1/ωs′
,
where C does not depend on j. Thus, φj → f in B′
f ∈ B′
ω(Rn) because D(Rn) ֒→ K1(Rn).
ω(Rn), which in turn implies that
We immediately get the ensuing result, a corollary of Theorem 4.1.
(cid:3)
ω(Rn). In particular, f ∈ D′(Rn) belongs
1(Rn) if and only if there is s ∈ R such that {e−shT−hf : h ∈ Rn} is bounded in
1(Rn) =Sω B′
ω(Rn) =Sω
Corollary 4.3. K′
to K′
D′(Rn).
B′
We present here the characterization of the spaces K1(Rn), K′
5. Characterizations through modulation spaces
1(Rn), B′
ω(Rn),
B′
ω(Rn),
U(Cn), and U ′(Cn) in terms of modulation spaces.
Let us recall the definition of the modulation spaces. There are several equivalent
ways to introduce them [9]. Here we follow the approach from [4, 5] based on Gelfand-
Shilov spaces. We are interested in modulation spaces with respect to weights that are
10
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
exponentially moderate. We denote by M the class of all weight functions m on R2n
that satisfy inequalities (for some constants A > 0 and a ≥ 0):
m(x1 + x2, ξ1 + ξ2)
m(x1, ξ1)
≤ Aea(x2+ξ2),
(x1, ξ1), (x2, ξ2) ∈ R2n.
Observe that any so-called v-moderate weight [9] belongs to M. We also consider the
Gelfand-Shilov space Σ1
1(Rn) of Beurling type (sometimes also denoted as S (1)(Rn) or
G(Rn)) and its dual (Σ1
1(Rn) consists [3] of all entire functions ϕ
1)′(Rn). The space Σ1
such that
ϕ(x)eλx < ∞ and
sup
x∈Rn
sup
ξ∈Rn
bϕ(ξ)eλξdξ < ∞,
∀λ > 0.
We refer to [19] for topological properties of Σ1
1)′(Rn) is also
known as the space of Silva ultradistributions of exponential type [14, 25] or the space
of Fourier ultra-hyperfunctions [18]. If m ∈ M, ψ ∈ Σ1
1(Rn) \ {0}, and p, q ∈ [1, ∞],
the modulation space M p,q
m (Rn) is defined as the Banach space
1(Rn). The dual space (Σ1
(5.1)
M p,q
m (Rn) = {f ∈ (Σ1
1)′(Rn) : kf kM p,q
m := kVψf kLp,q
m < ∞}.
This definition does not depend on the choice of the window ψ, as different windows
lead to equivalent norms. If p = q, then we write M p
m (Rn). The
m(Rn) (for m = 1) was original introduced by Feichtinger in [8]. We shall
space M 1
m (Rn) = {f ∈
also define
(Σ1
1)′(Rn) : lim(x,ξ)→∞ m(x, ξ)Vψf (x, ξ) = 0}.
We now connect the space of exponential distributions with the modulation spaces.
For it, we consider the weight subclass M1 ⊂ M consisting of all weights m such that
(for some s, a ≥ 0 and A > 0)
m (Rn) as the closed subspace of M ∞
m (Rn) given by M ∞
m(Rn) instead of M p,q
M ∞
(5.2)
m(x1 + x2, ξ1 + ξ2)
m(x1, ξ1)
≤ Aeax2(1 + ξ2)s,
(x1, ξ1), (x2, ξ2) ∈ R2n.
1(Rn) ֒→ K1(Rn),
Let m ∈ M1. By Proposition 3.1, K1(Rn) ⊂ M p,q
m (Rn)
we obtain that K1(Rn) is dense (weakly∗ dense if p = ∞ or q = ∞) in M p,q
1(Rn). It follows from the results of [9] that we may use
and therefore M p,q
ψ ∈ K1(Rn) \ {0} in (5.1). Also, if f ∈ M p,q
m (Rn) and ψ ∈ K1(Rn) then Vψf is an entire
function in the second variable (cf. Section 3); the next proposition describes the norm
behavior of Vψf (x, z) in the complex variable z ∈ Cn.
m (Rn). Since Σ1
m (Rn) ⊂ K′
Proposition 5.1. Let m ∈ M1, p, q ∈ [1, ∞], and ψ ∈ K1(Rn) \ {0}. If f ∈ M p,q
then (∀λ ≥ 0)
m (Rn),
(5.3)
η≤λ ZRn(cid:18)ZRn
sup
e−2πx·ηVψf (x, ξ + iη)m(x, ξ)pdx(cid:19)q/p
dξ!1/q
< Cλkf kM p,q
m .
(With obvious changes if p = ∞ or q = ∞.)
Proof. Assume that m satisfies (5.2) and set v(x, ξ) = (1 + ξ)seax. Notice first that
e−2πx·ηVψf (x, ξ + iη) = Vψη f (x, ξ), where ψη(t) = e2πη·tψ(t). As in the proof of [9,
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
11
Prop. 11.3.2, p. 234],
Vψη f Lp,q
m =
1
kψk2
L2
(VψηV ∗
ψ )Vψf Lp,q
m ≤ CVψη ψL1
vVψf Lp,q
m .
Since {ψη : η ≤ λ} is bounded in K1(Rn), we obtain that {Vψη ψ : η ≤ λ} is bounded
v < ∞.
in K1(Rn)b⊗U(Cn); hence supη≤λ VψηψL1
e−2πix·ξV bψbf (ξ, −x), we can transfer results from K′
Using the fundamental identity of time-frequency analysis, i.e. [9, p. 40] Vψf (x, ξ) =
1(Rn) into U ′(Rn) by employing the
weight class M2 = {m ∈ M : m(x, ξ) = m(ξ, x) ∈ M1}. For s, a ≥ 0, we employ the
following special classes of weights (ω satisfies the conditions imposed in Subsection
2.3):
(cid:3)
vs,a(x, ξ) := eax(1 + ξ)s
and ωs(x, ξ) := ω(x)(1 + ξ)s.
Clearly vs,a, ωs ∈ M1. Obviously, for every m ∈ M1 there are s, a ≥ 0 such that
M p,q
m (Rn) ⊆ M p,q
vs,a(Rn) ⊆ M p,q
(Rn).
1/vs,a
Proposition 5.2. Let p, q ∈ [1, ∞]. Then,
(5.4)
(5.5)
(5.6)
M p,q
m (Rn),
M p,q
m (Rn),
K′
1(Rn) = [m∈M1
K1(Rn) = \m∈M1
ω(Rn) =[s>0
B′
M ∞
1/ωs(Rn), and
U ′(Cn) = [m∈M2
U(Cn) = \m∈M2
ω(Rn) =[s>0
B′
M p,q
m (Rn),
M p,q
m (Rn),
M ∞
1/ωs(Rn).
Proof. The results for U(Cn) and U ′(Cn) follow from those for K1(Rn) and K′
1(Rn). The
equalities in (5.6) are a reformulation of the equivalences (a) ⇔ (c) and (a)′ ⇔ (c)′
from Theorem 4.1. By (5.2) and [9, Cor. 12.1.10, p. 254], given m ∈ M1, there are
(Rn) hold.
s, a > 0 such that the embeddings M ∞
Thus, part (a) from Lemma 3.2 gives the equality K′
1/vs,a(Rn) =
vs+n+1,a+ε(Rn) ⊆ M p,q
m (Rn) ⊆ M ∞
1/vs,a
m (Rn). In view of Proposition 3.1, it only remains to show that
1(Rn) = Ss,a>0 M ∞
Sm∈M1 M p,q
\m∈M1
M p,q
m (Rn) = \s,a>0
M ∞
vs,a(Rn) ⊆ K1(Rn).
(5.7)
We show the latter inclusion by proving that if f ∈ M ∞
is holomorphic in the tube Rn + i{η ∈ Rn : η < a/(2π)} and satisfies
vs,a(Rn) (with s, a > 0), then bf
In fact, choose a positive window ψ ∈ D(Rn) such that Pj∈Zn ψ(t − j) = 1 for all
t ∈ Rn. Since f =Pj∈Zn f Tjψ, we obtain bf =Pj∈Zn Vψf (j, · ), with convergence in
(1 + z2)s/2bf (z) < ∞,
a
2π
∀λ <
ℑm z≤λ
sup
.
12
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
U ′(Cn). In view of Proposition 5.1, each Vψf (j, z) is entire in z and satisfies the bounds
sup
(1 + z2)s/2Vψ(j, z) < Cλe−(a−2πλ)j.
ℑm z≤λ
The Weierstrass theorem implies that bf (z) = Pj∈Zn Vψf (j, z) is holomorphic in the
stated tube domain and we also obtain (5.7). Summing up, if f ∈ Ts,a>0 M ∞
then bf ∈ U(Cn), i.e., f ∈ K1(Rn).
The following corollary collects what was shown in the proof of Proposition 5.2.
vs,a(Rn),
(cid:3)
Corollary 5.3. Let s, a > 0. If f ∈ M ∞
Rn + i{η ∈ Rn : η < a/(2π)} and satisfies the bounds (5.7).
vs,a(Rn), then bf is holomorphic in the tube
We make a remark concerning Proposition 5.2.
Remark 5.4. Employing [26, Thrms. 3.2 and 3.4], Proposition 5.2 can be extended for
p, q ∈ (0, ∞].
6. Tauberian theorems for S-asymptotics of distributions
In this section we characterize the so-called S-asymptotic behavior of distributions
in terms of the STFT. We briefly explain this notion; we refer to [21] for a complete
treatment of the subject.
Let f ∈ K′
1(Rn). The idea of the S-asymptotics is to study the asymptotic properties
of the translates T−hf with respect to a locally bounded and measurable comparison
function c : Rn → (0, ∞). It is said that f has S-asymptotic behavior with respect to
c if there is g ∈ D′(Rn) such that
(6.1)
lim
h→∞
1
c(h)
T−hf = g
in D′(Rn).
The distribution g is not arbitrary; in fact, one can show [21] that the relation (6.1)
forces it to have the form g(t) = Ceβ·t, for some C ∈ R and β ∈ Rn. If C 6= 0, one can
also prove [21] that c must satisfy the asymptotic relation
(6.2)
lim
h→∞
c(t + h)
c(h)
= eβ·t,
uniformly for t in compact subsets of Rn.
¿From now on, we shall always assume that c satisfies (6.2). A typical example of such
a c is any function of the form c(t) = eβ·tL(et), where L is a Karamata slowly varying
function [2]. The assumption (6.2) implies [21] that (6.1) actually holds in the space
K′
1(Rn). We will use the more suggestive notation
(6.3)
f (t + h) ∼ c(h)g(t)
in K′
1(Rn) as h → ∞
for denoting (6.1), which of course means that (f ∗ ϕ)(h) ∼ c(h)RRn ϕ(t)g(t)dt as
h → ∞, for each ϕ ∈ D(Rn) (or, equivalently, ϕ ∈ K1(Rn)). In order to move further,
we give an asymptotic representation formula and Potter type estimates [2] for c:
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
13
Lemma 6.1. The locally bounded measurable function c satisfies (6.2) if and only if
there is b ∈ C ∞(Rn) such that limx→∞ b(α)(x) = 0 for every multi-index α > 0 and
(6.4)
c(x) ∼ exp (β · x + b(x))
as x → ∞.
In particular, for each ε > 0 there are constants aε, Aε > 0 such that
(6.5)
aε exp(β · t − εt) ≤
c(t + h)
c(h)
≤ Aε exp(β · t + εt),
t, h ∈ Rn.
Proof. By considering e−β·tc(t), one may assume that β = 0. Let ϕ ∈ D(Rn) be such
that RRn ϕ(t)dt = 1. Set b(x) = RRn log c(t + x)ϕ(t)dt. Clearly, b ∈ C ∞(Rn) and the
relation (6.2) implies that b(x) = log c(x)+o(1) and b(α)(x) = o(1) as x → ∞, for each
multi-index α > 0. This gives (6.4). Conversely, since c is locally bounded, we may
assume that actually c(x) = eβ·x+b(x), but b(t + h) − b(h) ≤ t maxξ∈[h,t+h] ∇b(ξ),
which gives (6.2). Using the fact that ∇b is bounded, the same argument yields
(6.5).
(cid:3)
Observe that Lemma 6.1 also tells us that the space B′
c(Rn) is well-defined for c.
We can now characterize (6.3) in terms of the STFT. The direct part of the following
theorem is an Abelian result, while the converse may be regarded as a Tauberian
theorem.
If f ∈ K′
1(Rn) has the S-
Theorem 6.2. Let f ∈ K′
asymptotic behavior (6.3) then, for every λ ≥ 0,
e2πiz·h Vψf (x + h, z)
(6.6)
1(Rn) and ψ ∈ K1(Rn)\{0}.
lim
h→∞
c(h)
= Vψg(x, z),
uniformly for z ∈ Πλ and x in compact subsets of Rn.
Conversely, suppose that the limits
(6.7)
lim
x→∞
e2πiξ·x Vψf (x, ξ)
c(x)
= J(ξ) ∈ C
exist for almost every ξ ∈ Rn. If there is s ∈ R such that
(6.8)
sup
(x,ξ)∈R2n
(1 + ξ)−s Vψf (x, ξ)
c(x)
< ∞,
then f has the S-asymptotic behavior (6.3) with g(t) = Ceβ·t, where the constant is
completely determined by the equation J(ξ) = Cbψ(−ξ + iβ/(2π)).
Remark 6.3. Assume (6.8). Consider a weight of the form mε(x, ξ) = eβ·x+εx(1 + ξ)s
It will be shown below that the asymptotics (6.3) holds in the weak∗
with ε > 0.
(Rn), i.e., (f ∗ ϕ)(h) ∼ c(h)hg, ϕi as h → ∞ for every ϕ in the
topology of M ∞
mε(Rn). Furthermore, one may use in (6.7) and (6.8) a window
modulation space M 1
ψ ∈ M 1
1/mε
mε(Rn)\{0}.
14
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
Proof. Fix λ ≥ 0 and a compact K ⊂ Rn. Note that the set
{MzTxψ : (x, z) ∈ K × Πλ}
is compact in K1(Rn). By the Banach-Steinhaus theorem,
lim
h→∞
e2πizh Vψf (x + h, z)
c(h)
= lim
h→∞D T−hf
c(h)
, MzTxψE =Dg, MzTxψE,
uniformly with respect to (x, z) ∈ K × Πλ, as asserted in (6.6).
Conversely, assume (6.7) and (6.8). Let H = {ξ ∈ Rn :
1(Rn), T−hf /c(h) converges strongly to a distribution g in K′
In view
c(Rn) or, equivalently, {T−hf /c(h) : h ∈ Rn}
of Theorem 4.1, we have that f ∈ B′
1(Rn). By the Banach-Steinhaus theorem and the Montel property
is bounded in K′
of K′
1(Rn) if and only if
limh→∞hT−hf, ϕi/c(h) exists for ϕ in a dense subspace of K1(Rn). Let D be the linear
span of {MξTxψ : (x, ξ) ∈ Rn × H}. By the desingularization formula (3.7) and the
Hahn-Banach theorem, we have that D is dense in K1(Rn). Thus, it suffices to verify
that limh→∞hT−hf, MξTxψi/c(h) exists for each (x, ξ) ∈ Rn × H. But in this case
(6.2) and (6.7) yield
(6.7) holds}.
lim
h→∞
hT−hf, MξTxψi
c(h)
= lim
h→∞
e2πiξ·h Vψf (x + h, ξ)
c(h)
e2πiξ·(x+h) Vψf (x + h, ξ)
c(h + x)
= e(β−2πiξ)·x lim
h→∞
= e(β−2πiξ)·xJ(ξ),
as required. We already know that g(t) = Ceβ·t. Comparison between (6.6) and (6.7)
leads to J(ξ) = Vψg(0, ξ) = CRRn ψ(t)eβ·t−2πiξ·tdt. To show the assertion from Remark
6.3, note first that, by using (6.5), one readily verifies that
T−hf M ∞
1/mε
c(h)
sup
h∈Rn
< ∞.
mε(Rn), we also have that D is dense
mε(Rn) and the assertion follows at once. The fact that one may use a window
mε(Rn)\{0} in (6.7) and (6.8) follows in a similar fashion because in this case
Since we have the dense embedding K1(Rn) ֒→ M 1
in M 1
ψ ∈ M 1
the desingularization formula (3.7) still holds.
(cid:3)
Let us make two addenda to Theorem 6.2. The ensuing corollary improves Remark
6.3, provided that c satisfies the extended submultiplicative condition (for some A > 0):
(6.9)
c(t + h) ≤ Ac(t)c(h).
Corollary 6.4. Assume that c satisfies (6.9) and set cs(x, ξ) = c(x)(1 + ξ)s, s ∈ R.
cs(Rn) \ {0} such that the limits (6.7) exist for
If f ∈ M ∞
1/cs
almost every ξ ∈ Rn, then, for some g, the S-asymptotic behavior (6.3) holds weakly∗
in M ∞
1/cs(Rn), that is, (f ∗ ϕ)(h) ∼ c(h)hg, ϕi as h → ∞ for every ϕ ∈ M 1
(Rn) and there is ψ ∈ M 1
cs(Rn).
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
15
Proof. We retain the notation from the proof of Theorem 6.2. The assumption f ∈
(Rn) of course tells us that (6.8) holds. Employing the hypothesis (6.9), one
M ∞
1/cs
readily sees that suph∈Rn kT−hf kM ∞
/c(h) < ∞. A similar argument to the one used
1/cs
in the proof of Theorem 6.2 yields that the set D associated to ψ is dense in M 1
cs(Rn),
which as above yields the result.
(cid:3)
In dimension n = 1, the next theorem actually obtains the ordinary asymptotic
behavior of f in case it is a regular distribution on (0, ∞) satisfying an additional
Tauberian condition. We fix mε as in Remark 6.3 and cs as in Corollary 6.4.
Theorem 6.5. Let f ∈ M ∞
(6.10)
1/cs(R). Suppose that
e2πiξ·x Vψf (x, ξ)
lim
x→∞
c(x)
= J(ξ) ∈ C,
for almost every ξ ∈ R, where ψ ∈ M 1
cs(R) \ {0} if c satisfies
(6.9)). If there is α ≥ 0 such that eαtf (t) is a positive non-decreasing function on the
interval (0, ∞), then
mε(R) \ {0} (resp. ψ ∈ M 1
(6.11)
lim
t→∞
f (t)
c(t)
= C,
where C is the constant from Theorem 6.2.
Proof. Using (6.10), the same method from Theorem 6.2 applies to show that f (t+h) ∼
1(R) as h → ∞, where g(t) = Ceβt. We may assume that α ≥ −β. Set
Cg(t) in K′
f (t) = eαtf (t), b(t) = eαtc(t), and r = α+β ≥ 0. It is enough to show that f (t) ∼ Cb(t)
as t → ∞, whence (6.11) would follow. By (6.3), we have that
f (t + h) ∼ b(h)Cert
as h → ∞ in K′
1(R),
h f (t + h), ϕ(t)i ∼ Cb(h)Z ∞
−∞
ertϕ(t)dt,
∀ϕ ∈ K1(Rn).
Let ε > 0 be arbitrary. Choose a non-negative test function ϕ ∈ D(R) such that
0 ϕ(t)dt = 1. Using the fact that f is non-decreasing on (0, ∞)
i.e.,
(6.12)
supp ϕ ⊆ (0, ε) and R ε
and (6.12), we obtain
f (h)
b(h)
lim sup
h→∞
= lim sup
h→∞
ϕ(t)dt ≤ lim
h→∞
f (h)
b(h)Z ε
0
h f (t + h), ϕ(t)i
= lim
h→∞
b(h)
= CZ ε
1
b(h)Z ε
0
f (t + h)ϕ(t)dt
ertϕ(t)dt ≤ Cerε,
taking ε → 0+, we have shown that lim suph→∞
in (6.12) a non-negative ϕ such that supp ϕ ⊆ (−ε, 0) andR 0
lim inf h→∞
f (h)/b(h) ≥ C. This shows that f (t) ∼ Cb(t) as t → ∞, as claimed.
0
f (h)/b(h) ≤ C. Similarly, choosing
−ε ϕ(t)dt = 1, one obtains
(cid:3)
We conclude this article with a proof of Theorem 1.1.
16
S. KOSTADINOVA, S. PILIPOVI ´C, K. SANEVA, AND J. VINDAS
Proof of Theorem 1.1. Set c(t) = eβtL(et) and, as before (with s = 0), c0(x, ξ) = c(x)
and mε(x, ξ) = eβx+εx. Note that (1.2) is the same as (6.11). Let us first verify that
ψ ∈ M 1
mε(R). In fact, if we take another window γ ∈ K1(R), we have
Vγψ(x, ξ)eβx+εxdxdξ =ZZR2
ψ(t − x)γ(t)eβx+εxdtdx +
ZZR2
≤ C ZZR2
(1 + ξ3)Vγψ(x, ξ)eβx+εxdx
dξ
1 + ξ3
3Xj=0ZZR2
ψ(j)(t − x)γ(3−j)(t)eβx+εxdtdx! ,
which is finite (a similar argument shows that ψ ∈ M 1
ψ′′(t))L(et)eβtdt < ∞).
M ∞
1/c0
Since f is non-decreasing, we have
(R). Let us first show the crude bound f (t) = O(c(t)). Set A1 =R ∞
−∞(ψ(t) + ψ′(t) +
In view of Theorem 6.5, it is enough to establish f ∈
0 ψ(t)dt < ∞.
c0(R) if R ∞
f (x) ≤
f (t)ψ(t − x)dt ≤ A2c(x),
1
0
0
1
f (t + x)ψ(t)dt ≤
A1Z ∞
A1Z ∞
c(t)ψ(t−x)dt ≤ c(x) AεZ ∞
−∞
because of (1.1) with ξ = 0. Thus
Vψf (x, ξ) ≤ A2Z ∞
0
eβt+εtψ(t)dt < A3c(x), ∀(x, ξ) ∈ R2
(likewise in the other case using L(xy) ≤ AL(x)L(y)), which completes the proof. (cid:3)
References
[1] A. Beurling, Sur les int´egrales de Fourier absolument convergentes et leur application `a une
transformation fonctionelle, in: IX Congr. Math. Scand., Helsingfors, (1938) 345 -- 366.
[2] N. H. Bingham, C. M. Goldie, J. L. Teugels, Regular Variation, Cambridge University Press,
Cambridge, 1987.
[3] S.-Y. Chung, D. Kim, S. Lee, Characterizations for Beurling-Bjorck space and Schwartz space,
Proc. Amer. Math. Soc. 125 (1997), 3229-3234.
[4] E. Cordero, Gelfand-Shilov window classes for weighted modulation spaces, Integral Transforms
Spec. Funct. 18 (2007), 829 -- 837.
[5] E. Cordero, S. Pilipovi´c, L. Rodino, N. Teofanov, Localization operators and exponential weights
for modulation spaces, Mediterr. J. Math. 2 (2005), 381 -- 394.
[6] W. F. Donoghue, Distributions and Fourier Transforms, Academic Press, New York-London,
1969.
[7] R. Estrada, J. Vindas, On Borel summability and analytic functionals, Rocky Mountain J. Math.
43 (2013), 895 -- 903.
[8] H. G. Feichtinger, On a new Segal algebra, Monatsh. Math. 92 (1981), 269 -- 289.
[9] K. Grochenig, Foundations of time-frequency analysis, Birkhauser Boston, Inc., Boston, MA,
2001.
[10] K. Grochenig, G. Zimmermann, Hardy's theorem and the short-time Fourier transform of
Schwartz functions, J. London Math. Soc. 63 (2001), 205 -- 214.
[11] K. Grochenig, G. Zimmermann, Spaces of test functions via the STFT, J. Funct. Spaces Appl. 2
(2004), 25 -- 53.
[12] M. Hasumi, Note on the n-dimensional tempered ultra-distributions, Tohoku Math. J. 13 (1961),
94 -- 104.
[13] E. Hille, R. S. Phillips, Functional analysis and semi-groups, American Mathematical Society,
Providence, R. I., 1957.
THE STFT OF DISTRIBUTIONS OF EXPONENTIAL TYPE
17
[14] R. F. Hoskins, J. Sousa Pinto, Theories of generalised functions. Distributions, ultradistributions
and other generalised functions, Horwood Publishing Limited, Chichester, 2005.
[15] J. Korevaar, Tauberian theory. A century of developments, Springer-Verlag, Berlin, 2004.
[16] M. Morimoto, Theory of tempered ultrahyperfunctions. I, II, Proc. Japan Acad. 51 (1975), 87 -- 91;
51 (1975), 213 -- 218.
[17] H. M. Obiedat, Z. Mustafa, F. Awawdeh, Short-time Fourier transform over the Silva space,
Inter. J. Pure Appl. Math. 44 (2008), 755 -- 764.
[18] Y. S. Park, M. Morimoto, Fourier ultra-hyperfunctions in the Euclidean n−space, J. Fac. Sci.
Univ. Tokyo Sect. IA Math. 20 (1973), 121 -- 127.
[19] S. Pilipovi´c, Tempered ultradistributions, Boll. Un. Mat. Ital. B (7) 2 (1988), 235 -- 251.
[20] S. Pilipovi´c, B. Stankovi´c, Wiener Tauberian theorems for distributions, J. London Math. Soc.
(2) 47 (1993), 507 -- 515.
[21] S. Pilipovi´c, B. Stankovi´c, J. Vindas, Asymptotic behavior of generalized functions, Series on
Analysis, Applications and Computation, 5, World Scientific Publishing Co. Pte. Ltd., Hacken-
sack, NJ, 2012.
[22] K. Saneva, R. Aceska, S. Kostadinova, Some Abelian and Tauberian results for the short-time
Fourier transform, Novi Sad J. Math. 43 (2013), 81 -- 89.
[23] L. Schwartz, Th´eorie des distributions, Hermann, Paris, 1966.
[24] J. Sebastiao e Silva, Les fonctions analytiques comme ultra-distributions dans le calcul
op´erationnel, Math. Ann. 136 (1958), 58 -- 96.
[25] J. Sebastiao e Silva, Les s´eries de multipoles des physiciens et la th´eorie des ultradistributions,
Math. Ann. 174 (1967), 109 -- 142.
[26] 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.
[27] F. Tr`eves, Topological vector spaces, distributions and kernel, Academic Press, New York, 1967.
[28] V. S. Vladimirov, Yu. N. Drozzinov, B. I. Zavialov, Tauberian theorems for generalized functions,
Kluwer Academic Publishers Group, Dordrecht, 1988.
[29] N. Wiener, Tauberian theorems, Ann. of Math. (2) 33 (1932), 1 -- 100.
[30] Z. Ziele´zny, Hypoelliptic and entire elliptic convolution equations in subspaces of the space of
distributions. II, Studia Math. 32 (1969), 47 -- 59.
Faculty of Electrical Engineering and Information Technologies, Ss. Cyril and
Methodius University, Rugjer Boshkovik bb, 1000 Skopje, Macedonia
E-mail address: [email protected]
Department of Mathematics and Informatics, University of Novi Sad, Trg Dositeja
Obradovi´ca 4, 21000 Novi Sad, Serbia
E-mail address: [email protected]
Faculty of Electrical Engineering and Information Technologies, Ss. Cyril and
Methodius University, Rugjer Boshkovik bb, 1000 Skopje, Macedonia
E-mail address: [email protected]
Department of Mathematics, Ghent University, Krijgslaan 281 Gebouw S22, 9000
Gent, Belgium
E-mail address: [email protected]
|
1606.08606 | 1 | 1606 | 2016-06-28T08:32:20 | Mityagin's Extension Problem. Progress Report | [
"math.FA",
"math.CA"
] | Given a compact set $K\subset {\Bbb R}^d,$ let ${\mathcal E}(K)$ denote the space of Whitney jets on $K$. The compact set $K$ is said to have the extension property if there exists a continuous linear extension operator $W:{\mathcal E}(K) \longrightarrow C^{\infty}({\Bbb R}^d)$. In 1961 B. S. Mityagin posed a problem to give a characterization of the extension property in geometric terms. We show that there is no such complete description in terms of densities of Hausdorff contents or related characteristics. Also the extension property cannot be characterized in terms of growth of Markov's factors for the set. | math.FA | math |
1
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
ALEXANDER GONCHAROV AND ZELIHA URAL
Abstract. Given a compact set K ⊂ Rd, let E(K) denote the space of Whitney jets
on K. The compact set K is said to have the extension property if there exists a
continuous linear extension operator W : E(K) −→ C∞(Rd). In 1961 B. S. Mityagin
posed a problem to give a characterization of the extension property in geometric
terms. We show that there is no such complete description in terms of densities of
Hausdorff contents or related characteristics. Also the extension property cannot be
characterized in terms of growth of Markov's factors for the set.
1. introduction
By the celebrated Whitney theorem [23], for each compact set K ⊂ Rd, by means of
a continuous linear operator one can extend jets of finite order from E p(K) to functions
defined on the whole space, preserving the order of differentiability. In the case p = ∞,
the possibility of such extension crucially depends on geometry of the set. Following
[20], let us say that K has the extension property (EP) if there exists a linear continuous
extension operator W : E(K) −→ C ∞(Rd). For example, any set K with an isolated
point does not have EP , since here each neighborhood of the space E(K) contains a
linear subspace, but this is not the case for C ∞(Rd).
B. S. Mityagin posed in 1961 ([13], p.124) the following problem (in our terms):
What is a geometric characterization of the extension property?
We show that there is no complete characterization of that kind in terms of densities
of Hausdorff contents of sets or analogous functions related to Hausdorff measures.
This is similar to the state in Potential Theory where R. Nevanlinna [14] and H.
Ursell [21] proved that there is no complete characterization of polarity of compact
sets in terms of Hausdorff measures. The scale of growth rate of functions h, which
define the Hausdorff measure Λh, can be decomposed into three zones. For h from the
first zone of small growth, if 0 < Λh(K) then the set K is not polar. For h from the
zone of fast growth, if Λh(K) < ∞ then the set K is polar. But between them there
is a zone of uncertainty. It is possible to take two functions with h2 ≺ h1 from this
zone and the corresponding Cantor-type sets Kj with 0 < Λhj (Kj) < ∞ for j ∈ {1, 2},
such that the large (with respect to the Hausdorff measure) set K2 is polar, whereas
the smaller K1 is not polar.
Here we present a similar example of two Cantor-type sets: the smaller set has EP
whereas the larger set does not have it.
1 Alexander Goncharov, Zeliha Ural (Bilkent University, 06800, Ankara, Turkey)
e-mail: [email protected], [email protected]
2010 Mathematics Subject Classification. 46E10, 31A15, 41A10.
Key words and phrases. Whitney functions, extension problem, Hausdorff measures, Markov's
factors.
1
2
ALEXANDER GONCHAROV AND ZELIHA URAL
Of course, such global characteristics as Hausdorff measures or Hausdorff contents
cannot be used, in general, to distinguish EP, which we observe if a compact set is not
"very small" near each its point. One can suggest for this reason to characterize EP
in terms of lower densities of Hausdorff contents of sets, because, clearly, densities of
Hausdorff measures cannot be used for this aim. We analyze a wide class of dimension
functions and show that lower densities of Hausdorff contents do not distinguish EP .
Neither EP can be characterized in terms of growth rate of Markov's factors (Mn(·))∞
for sets. Two sets are presented, K1 with EP and K2 without it, such that Mn(K1)
grows essentially faster than Mn(K2) as n → ∞.
It should be noted that, by W.
Ple´sniak [16], any Markov compact set (with a polynomial growth rate of Mn(·)) has
EP . All examples are given in terms of the sets K(γ) introduced in [9].
n=1
Our paper is organized as follows. Section 2 is a short review of main methods
of extension. Also we consider there the Tidten-Vogt linear topological characteriza-
tion of EP . In Section 3 we give some auxiliary results about the weakly equilibrium
Cantor-type set K(γ). In Section 4 we use local Newton interpolations to construct an
extension operator W . Sections 5 contains the main result, namely a characterization
of EP for E(K(γ)) in terms of a sequence related to γ. In sections 6 we compare W
with the extension operator from [11], which is given by individual extensions of ele-
ments of Schauder basis for the space E(K(γ)). In Section 7 we consider two examples
that correspond to regular and irregular behaviour of the sequence γ. In Section 8 we
calculate the Hausdorff h−measure of K(γ) for a siutable dimension function h and
present Ursell's type example for EP. In Section 9 we consider Hausdorff contents and
related characteristics. In Section 10 we compare the growth of Markov's factors and
EP for K(γ).
For the basic facts about the spaces of Whitney functions defined on closed subsets
of Rd see e.g. [3], the concepts of the theory of logarithmic potential can be found in
[17]. Throughout the paper, log denotes the natural logarithm. Given compact set K,
Cap(K) stands for the logarithmic capacity of K, Rob(K) = log(1/Cap(K)) ≤ ∞ is
the Robin constant for K. If K is not polar then µK is its equilibrium measure. For
each set A, let #(A) be the cardinality of A, A be the diameter of A. Also, [a] is
k=m(· · · ) = 1 if m > n. The symbol ∼
denotes the strong equivalence: an ∼ bn means that an = bn(1 + o(1)) for n → ∞.
the greatest integer in a, Pn
k=m(· · · ) = 0 and Qn
2. Three methods of extension
j=1 ∈ Zd
Let K ⊂ Rd be a compact set, α = (αj)d
+ be a multi-index. Let I be
a closed cube containing K and F (K, I) = {F ∈ C ∞(I) : F (α)K = 0, ∀α} be the
ideal of flat on K functions. The Whitney space E(K) of extendable jets consists of
traces on K of C ∞-functions defined on I, so it is a factor space of C ∞(I) and the
restriction operator R : C ∞(I) −→ E(K) is surjective. This means that the sequence
J−→ C ∞(I) R−→ E(K) −→ 0 is exact. If it splits then the right inverse
0 −→ F (K, I)
to R is the desired linear continuous extension operator W and K has EP . We see
that there always exists a linear extension operator (for example one can individually
extend the elements of a vector basis in E(K)) and a continuous extension operator,
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
3
by Whitney's construction. Numerous examples show that a set K has EP if K is not
"very small" near each its point, but the exact geometric meaning of "smallness" has
not been comprehended yet.
In [20] M. Tidten applied D. Vogt's theory of splitting of short exact sequences of
Fr´echet spaces (see e.g [12], Chapter 30) and presented the following important linear
topological characterization of EP : a compact set K has the extension property if and
only if the space E(K) has a dominating norm (satisfies the condition (DN)).
Recall that a Fr´echet space X with an increasing system of seminorms ( · k)∞
k=0
has a dominating norm · p if for each q ∈ N there exist r ∈ N and C ≥ 1 such that
· 2
q ≤ C · p · r.
Concerning the question "How to construct an operator W if it exists?", we can
select three main methods that can be applied for wide families of compact sets.
ements (en)∞
W (f ) = P∞
n=1 of a topological basis of E(K). Then for f = P∞
The first method goes back to B. S. Mityagin [13]: to extend individually the el-
n=1 ξn · en take
n=1 ξn · W (en). See Theorem 2.4 in [22] about possibility of suitable si-
multaneous extensions of en in the case when K has nonempty interior. The main
problem with this method is that we do not know whether each space E(K) has a
topological basis, even though E(K) is complemented in C ∞(I). This is a particular
case of the significant Mityagin-Pe lczy´nski problem: Suppose X is a nuclear Fr´echet
space with basis and E is a complemented subspace of X. Does E possess a basis? The
space X = s of rapidly decreasing sequences, which is isomorphic to C ∞(I), presents
the most important unsolved case.
The second method was suggested in [15], where W. Paw lucki and W. Ple´sniak
constructed an extension operator W in the form of a telescoping series containing La-
grange interpolation polynomials with Fekete nodes. The authors considered the family
of compact sets with polynomial cusps, but later, in [16], the result was generalized
to any Markov sets. In fact (see T.3.3 in [16]), for each C ∞ determining compact set
K, the operator W is continuous in the so-called Jackson topology τJ if and only if τJ
coincides with the natural topology τ of the space E(K) and this happens if and only if
the set K is Markov. We remark that τJ is not stronger then τ and that τJ always has
the dominating norm property, see e.g. [2]. Thus, in the case of non-Markov compact
set with EP ([5], [2]), the Paw lucki-Ple´sniak extension operator is not continuous in
τJ , but this does not exclude the possibility for it to be bounded in τ . At least for
some non-Markov compact sets, the local version of this operator is bounded in τ ([2]).
In [4] L. Frerick, E. Jord´a, and J. Wengenroth showed that, provided some condi-
tions, the classical Whitney extension operator for the space of jets of finite order can
be generalized to the case E(K). Instead of Taylors polynomials in the Whitney con-
struction, the authors used a kind of interpolation by means of certain local measures.
A linear tame extension operator was presented for E(K), provided K satisfies a local
form of Markov's inequality.
4
ALEXANDER GONCHAROV AND ZELIHA URAL
There are some other methods to construct W for closed sets, for example Seeley's
extension [18] from a half space or Stein's extension ([19], Ch 6) from sets with the
Lipschitz boundary. However these methods, in order to define W (f, x) at some point
x, essentially require existence of a line through x with a ray where f is defined, so
these methods cannot be applied for compact sets.
Here we consider rather small Cantor-type sets that are neither Markov no local
Markov. We follow [2] in our construction, so W is a local version of the Paw lucki-
Ple´sniak operator. It is interesting that, at least for small sets, W can be considered
as an operator extending basis elements of the space. Thus, for such sets, the first
method and a local version of the second method coincide.
3. Notations and auxiliary results
In what follows we will consider only perfect compact sets K ⊂ I = [0, 1], so the
Fr´echet topology τ in the space E(K) can be given by the norms
k f kq = f q,K + sup((Rq
yf )(k)(x)
x − yq−k
: x, y ∈ K, x 6= y, k = 0, 1, ...q)
for q ∈ Z+, where f q,K = sup{f (k)(x) : x ∈ K, k ≤ q} and Rq
is the Taylor remainder.
yf (x) = f (x) − T q
y f (x)
Given f ∈ E(K), let f q = inf F q,I, where the infimum is taken over all
possible extensions of f to F ∈ C ∞(I). By the Lagrange form of the Taylor remainder,
we have f q ≤ 3 F q,I for any extension F . The quotient topology τQ, given by the
norms ( · ∞
q=0), is complete and, by the open mapping theorem, is equivalent to τ.
Hence for any q there exist r ∈ N, C > 0 such that
(1)
f q ≤ C f r
for any f ∈ E(K). In general, extensions F that realize f q for a given function f ,
essentially depend on q. Of course, the extension property of K means the existence
of a simultaneous extension which is suitable for all norms.
Our main subject is the set K(γ) introduced in [9]. For the convenience of the reader
we repeat the relevant material. Given sequence γ = (γs)∞
s=1 with 0 < γs < 1/4, let
r0 = 1 and rs = γsr2
s−1 for s ∈ N. Define P2(x) = x(x − 1), P2s+1 = P2s(P2s + rs) and
Es = {x ∈ R : P2s+1(x) ≤ 0} for s ∈ N. Then Es = ∪2s
j=1Ij,s, where the s-th level basic
intervals Ij,s are disjoint and max1≤j≤2s Ij,s → 0 as s → ∞. Here, (P2s + rs/2)(Es) =
[−rs/2, rs/2], so the sets Es are polynomial inverse images of intervals. Since Es+1 ⊂ Es,
we have a Cantor-type set K(γ) := ∩∞
s=0Es.
In what follows we will consider only γ satisfying the assumptions
∞
(2)
γk ≤ 1/32
for
k ∈ N and
γk < ∞.
The lengths lj,s of the intervals Ij,s of the s−th level are not the same, but, provided
(2), we can estimate them in terms of the parameter δs = γ1γ2 · · · γs ([9], L.6):
(3)
δs < lj,s < C0 δs
for
1 ≤ j ≤ 2s,
Xk=1
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
5
where C0 = exp(16 P∞
k=1 γk). Each Ij,s contains two adjacent basic subintervals I2j−1,s+1
and I2j,s+1. Let hj,s = lj,s − l2j−1,s+1 − l2j,s+1 be the distance between them. By Lemma
4 in [9], hj,s > (1 − 4γs+1)lj,s. Therefore,
(4)
hj,s ≥ 7/8 · lj,s > 7/8 · δs
for all j.
In addition, by T.1 in [9], the level domains Ds = {z ∈ C : P2s(z) + rs/2 < rs/2}
k=1 2−k log 1
form a nested family and K(γ) = ∩∞
γk
represents the Robin constant of Ds. Therefore, the set K(γ) is non-polar if and only
s=0Ds. The value Rs = 2−s log 2 +Ps
=P∞
n=1 2−n−1 log 1
δn
< ∞.
if Rob(K(γ)) =P∞
n=1 2−n log 1
γn
We decompose all zeros of P2s into s groups. Let X0 = {x1, x2} = {0, 1}, X1 =
{x3, x4} = {l1,1, 1 − l2,1}, · · · , Xk = {l1,k, l1,k−1 − l2,k, · · · , 1 − l2k,k} for k ≤ s − 1. Thus,
Xk = {x : P2k (x) + rk = 0} contains all zeros of P2k+1 that are not zeros of P2k. Set
Ys = ∪s
#(Ys) = 2s+1 for s ∈ Z+. We refer s−th type points to the elements of Xs.
k=0Xk. Then P2s(x) = Qxk∈Ys−1(x − xk). Clearly, #(Xs) = 2s for s ∈ N and
The points from Ys can be ordered using, as in [7], the rule of increase of the
type. First we take points from X0 and X1 in the ordering given above. The set
X2 = {x5, · · · , x8} consists of the points of the second type. We take xj+4 as the point
which is the closest to xj for 1 ≤ j ≤ 4. Here, x5 = x1 + l1,2, x6 = x2 − l4,2, etc. Simi-
larly, Xk = {x2k+1, · · · , x2k+1} can be defined by the previous points. For 1 ≤ j ≤ 2k,
the point xj is an endpoint of a certain basic interval of k-th level. Let us take xj+2k
as its another endpoint. Thus, xj+2k = xj ± li,k, where the sign and i are uniquely
defined by j. In the same way, any N points can be chosen on each basic interval. For
example, suppose 2n ≤ N < 2n+1 and the points (zk)N
k=1 are chosen on Ij,s by this rule.
Then the set includes all 2n zeros of P2s+n on Ij,s (points of the type ≤ s + n − 1) and
some N − 2n points of the type s + n.
We use two technical lemmas from [11]. We suppose that γ satisfies (2).
Let 2n ≤ N < 2n+1 and Z = (zk)N +1
k=1 be chosen on a given I = Ij,s by the rule of
k=1 and C1 = 8/7 · (C0 + 1). For fixed x ∈ R and
increase of the type. Write ZN = (zk)N
finite A = (am), let dk(x, A) := x − amk ր .
Lemma A. For each x ∈ R with δ = dist(x, ZN ) ≤ δs+n and z ∈ Z we have
δs+n QN
k=2 dk(x, ZN ) ≤ C N
k=2 dk(z, Z).
1 QN +1
In the next lemma we consider the same N and Z, as above, but now we arrange
zk in increasing order. For q = 2m − 1 with m < n and 1 ≤ j ≤ N + 1 − q, let
J = {zj, · · · , zj+q} be 2m consecutive points from Z. Given j, we consider all possible
chains of strict embeddings of segments of natural numbers:
[j, j + q] = [a0, b0] ⊂
[a1, b1] ⊂ · · · ⊂ [aN −q, bN −q] = [1, N + 1], where ak = ak−1, bk = bk−1 + 1 or ak = ak−1 −
k=1 (zbk − zak ).
1, bk = bk−1 for 1 ≤ k ≤ N − q. Every chain generates the product QN −q
Lemma B. For each J ⊂ Z there exists z ∈ J such that QN +1
Potential Theory meaning: Rob(K(γ)) = P∞
We will characterize EP of K(γ) in terms of the values Bk = 2−k−1 · log 1
δk
that have
k=1 Bk. The main condition is (compare
For fixed J, let Π(J) denote the minimum of these products for all possible chains.
k=q+2 dk(z, Z) ≤ Π(J).
6
ALEXANDER GONCHAROV AND ZELIHA URAL
with (3) in [8]):
(5)
⇒ 0 as n → ∞ uniformly with respect to s.
Bn+s
k=s Bk
Pn+s
We see that this condition allows polar sets.
Example 1. Let γ1 = exp(−4B) and γk = exp(−2kB) for k ≥ 2, where B ≥ 1
4 log 32,
so (2) is valid. Here, Bk = B for all k. Hence (5) is satisfied and the set K(γ) is polar.
The condition (5) means that
(6)
∀ε ∃s0, ∃n0 : Bs+n < ε(Bs + · · · + Bs+n) for n ≥ n0, s ≥ s0.
Clearly, instead of ∃s0 one can take above ∀s0. Let us show that (6) is equivalent to
(7) ∀ε1 ∀m ∈ Z+∃N : Bs+n−m + · · · + Bs+n < ε1(Bs + · · · + Bs+n−m−1), n ≥ N, s ≥ 1.
Indeed, the value m = 0 in (7) gives (6) at once. For the converse, remark that in (7) we
can take on the right side ε1(Bs + · · · + Bs+n), so here we consider (7) in this new form.
Suppose (6) is valid. Given ε1 and m, take ε = ε1/(m + 1) and the corresponding value
n0 from (6). Take N = n0 + m. Then for n ≥ N and 0 ≤ k ≤ m we have n − k ≥ n0,
so Bs+n−k < ε(Bs + · · · + Bs+n−k) < ε(Bs + · · · + Bs+n). Summing these inequalities,
we obtain a new form of (7).
It follows that the negation of the main condition can be written as
(8)
∃ε ∃m : ∀N ∃n > N :
Also, (6) is equivalent to
s+n
Xs+n−m
Bk > ε
s+n−m−1
Xs
Bk for s = sj ↑ ∞.
(9)
∀ε ∃m, n0, s0 : Bs+n < ε(Bs+n−m + · · · + Bs+n−1) for n ≥ n0, s ≥ s0.
Indeed, comparison of right sides of inequalities shows that (9) implies (6). Conversely,
given ε, take n0 such that (6) is valid with ε/(1 + ε) instead of ε. Take m = n0.
Then for n ≥ n0, s ≥ s0 we have s = s + n − m ≥ s0 and, by (6), Bs+n = Bs+m <
ε
1+ε (Bs + · · · Bs+m), which is (9).
We will use a "geometric" version of (9) in terms of (δk)
(10)
∀M ∃m, n0, s0 : δs+n−1 δ2
s+n−2 · · · δ2m−1
s+n−m < δM
s+n for n ≥ n0, s ≥ s0.
4. Extension operator for E(K(γ))
Here, as in [2], we use the method of local Newton interpolations. Let K be shorthand
for K(γ). We fix a nondecreasing sequence of natural numbers (ns)∞
s=0 with ns ≥ 2 and
ns → ∞. Given function f on K, we interpolate f at 2n0 points that are chosen by the
rule of increase of the type on the whole set. A half of points are located on K ∩ I1,1.
We continue interpolation on this set up to the degree 2n1. Separately we do the same
on K ∩ I2,1. Continuing in this fashion, we interpolate f with higher and higher degrees
on smaller and smaller basic intervals. At each step the additional points are chosen
by the rule of increase of the type. Interpolation on Ij,s does not affect other intervals
of the same level due to the following function.
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
7
Let t > 0 and a compact set E on the line be given. Then u(·, t, E) is a C ∞−
function with the properties: u(·, t, E) ≡ 1 on E, u(x, t, E) = 0 for dist(x, E) > t and
supx∈K u(p)
xp (x, t, K) ≤ cp t−p, where the constant cp depends only on p. Let cp ր .
Given N + 1 points (zk)N +1
ωk(x) =
ΩN +1(x)
(x−zk)Ω′
k=1 on K ∩ Ij,s let LN (f, x, Ij,s) =PN +1
k=1 (x − zk).
N +1(zk) with ΩN +1(x) =QN +1
Let Ns = 2ns − 1 and Ms = 2ns−1−1 − 1 for s ≥ 1, M0 = 1. Then, for fixed s, we take
Ms + 1 ≤ N ≤ Ns, so 2n ≤ N < 2n+1 with n ∈ {ns−1 − 1, · · · , ns − 1}. For such N
and s we take tN := δs+n. Let, in addition, 1 ≤ j ≤ 2s be fixed. Then we choose N + 1
points on the interval Ij,s by the rule of increase of the type and consider for given f
k=1 f (zk) ωk(x), where
AN,j,s := [LN (f, x, Ij,s) − LN −1(f, x, Ij,s)] u(x, tN , Ij,s ∩ K).
We call Aj,s(f, x) := PNs
N =Ms+1 AN,j,s the accumulation sum. The last term here
corresponds to the interpolation on Ij,s at 2ns points. In order to continue interpolation
on subintervals of Ij,s, let us consider the transition sum
Tk,s(f, x) := [LMs+1(f, x, Ik,s+1) − LNs(f, x, Ij, s)] u(x, δs+ns−1, Ik, s+1 ∩ K),
where we suppose 1 ≤ k ≤ 2s+1, j = [ k+1
2 ] and Ij, s ⊃ Ik,s+1 ∪ Ii,s+1.
As above, we represent the difference in brackets in the telescoping form:
[LMs+1 − LNs] = −
2ns −1
XN =2ns −1
[LN (f, x, Ij, s) − LN −1(f, x, Ij, s)].
Here, the interpolating set Z for LN consists of Ms+1 + 1 points of Ys+ns−1 ∩ Ik,s+1
and N − Ms+1 points, chosen by the rule of increase of the type on Ii,s+1. The second
parameter of u is smaller than the mesh size of Z, so Tk,s(f, x) 6= 0 only near Ik,s+1.
Consider a linear operator
W (f, ·) = LM0(f, ·, I1, 0) u(·, 1, K) +
∞
2s
Xs=0(cid:2)
Xj=1
Aj,s(f, ·) +
2s+1
Xk=1
Tk,s(f, ·)(cid:3).
We remark at the outset that, for fixed x ∈ R and s, because of the choice of parameters
for the function u, at most one value Aj,s does not vanish. The same is valid for Tk,s.
Let us show that W extends functions from E(K), provided a suitable choice of
(ns)∞
s=0. Define n0 = n1 = 2 and ns = [log2 log 1
δs
] for s ≥ 2. Then ns ≤ ns+1 and
(11)
1
2
log
1
δs
< 2ns ≤ log
1
δs
for s ≥ 2.
Lemma 4.1. Let (ns)∞
we have W (f, x) = f (x).
s=0 be given as above. Then for any f ∈ E(K(γ)) and x ∈ K(γ)
Proof. Let us fix a natural number q with q > 2 + log(8C0/7), where C0 is defined in
(3). By the telescoping effect,
(12)
W (f, x) = lim
s→∞
LMs(f, x, Ij,s),
8
ALEXANDER GONCHAROV AND ZELIHA URAL
where j = j(s, x) is chosen in a such way that x ∈ Ij,s. As in [EvI],
2n
(13)
LMs(f, x, Ij,s) − f (x) ≤ f q
x − zk q ωk(x) .
Xk=1
Here n is shorthand for ns−1 − 1 and s is such that Ms = 2n − 1 > q. The interpolating
set (zk)2n
k=1 for LMs consists of all points of the type ≤ s + n − 1 on Ij,s. Given point x,
we consider the chain of basic intervals containing it: x ∈ Ijn,s+n ⊂ · · · ⊂ Ij1,s+1 ⊂ Ij,s.
We see that Ijn,s+n contains one interpolating point, Ijn−1,s+n−1 \ Ijn,s+n does one more
zi, Ijn−2,s+n−2 \ Ijn−1,s+n−1 contains two such points, etc. Thus, for fixed k, we get
2n
Yi=1,i6=k
x − zk q
x − zi ≤ l q−1
j,s
· ljn,s+n · ljn−1,s+n−1 · l2
jn−2,s+n−2 · · · l 2 n−1
j,s
.
By (3), this does not exceed C 2n+q−1
On the other hand, by a similar argument, for the denominator of ωk(x) we have
s+n−2 · · · δ2n−2
s+1 δ2n−1+q−1
δs+n δs+n−1 δ2
.
0
s
zk − z1 · · · zk − zk−1 · zk − zk+1 · · · zk − z2n ≥ lqn−1,s+n−1 · h2
qn−2,s+n−2 · · · h 2 n−1
j,s
for some indices qn−1, qn−2, · · · . The last product exceeds (7/8)2n−2δs+n−1 δ2
by (4). It follows that
s+n−2 · · · δ2n−1
s
,
LHS of (13) ≤ f q 2n C q−1
0
(8C0/7) 2 n
δs+n δq−1
s
.
The expression on the right side approaches zero as s → ∞. Indeed, 2n < log(1/δs−1),
by (11), and 2n(8C0/7) 2 n δq−1
s < 1 due to the choice of q. Thus the limit in (12) exists
and equals f (x).
(cid:3)
5. Extension property of weakly equilibrium Cantor-type sets
We need two more lemmas.
Lemma 5.1. Let γ satisfy (2), q = 2m, r = 2n with m < n and Z = (zk)r
points of the type ≤ s + n − 1 on I1,s for some s ∈ Z+. Let f (x) = Qr
for x ∈ K(γ) ∩ I1,s and f = 0 on K(γ) \ I1,s. Then f 0,K(γ) ≤ C r
n+s−2 · · · δ2n−1
δ2
, f (q)(0) ≥ q! · (7/8)r−q · δ2m
n+s−m−1 · · · δ2n−1
and f r ≤ 2 · r!.
k=1 be all
k=1(x − zk)
0 · δn+s · δn+s−1 ·
s
s
Proof. Fix x that realizes f 0,K(γ) and a chain of basic intervals containing this point:
x ∈ Ij0,n+s ⊂ Ij1,n+s−1 ⊂ · · · ⊂ Ijn,s = I1,s. Arguing as in Lemma 4.1, we see that
f 0,K(γ) ≤ lj0,n+s · lj1,n+s−1 · l2
, which, by (3), gives the desired bound.
j2,n+s−2 · · · l2n−1
In order to estimate f (q)(0), let us remark that f (q)(x) is a sum of(cid:0)r
is g(x) := Qr
the location of points from Z, we get g(0) = Qr
product has a coefficient q! and consists of r − q terms (x − zk). One of these products
k=q+1(x − zik ), where zi1 < zi2 < · · · < zir . All products are nonnegative
at x = 0, since r − q is even. From here, f (q)(0) ≥ q! · g(0). Taking into account
1,s >
(cid:3)
k=q+1 zik > h2m
, by (4). The bound of kf kr is evident.
q(cid:1) products, each
1,n+s−m−1 · · · h2n−1
n+s−m−1 · · · δ2n−1
(7/8)r−q · δ2m
1,s
s
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
9
In the next Lemma, for given 2n ≤ N < 2n+1, we consider ΩN (x) = QN
k=1(x − zk)
k=1, where the points are chosen on Ij,s by the rule of increase of the
with ZN = (zk)N
type. Let u(x) = u(x, δs+n, Ij,s ∩ K(γ)) and, as above, di(x, ZN ) := x − zki ր .
Lemma 5.2. The bound (ΩN · u)(p)(x) ≤ 2p (C0 + 1) cp δ−p+1
k=2 dk(x, ZN )
is valid for each p < N and x ∈ R.
Proof. By Leibnitz's formula, (ΩN · u)(p)(x) ≤ Pp
increases, we have Ω(i)
N (x) ≤ N !
i=0(cid:0)p
k=i+1 dk(x, ZN ). This gives
N (x) cp−iδ−p+i
s+n . Since dk
s+n N p QN
i(cid:1) Ω(i)
Yk=i+1
N
dk(x, ZN ).
(14)
(ΩN · u)(p)(x) ≤ 2p cp δ−p
s+n · max
0≤i≤p
(N δs+n)i
(N −i)!QN
The set ZN consists of 2n endpoints of subintervals of the level s+n−1 covered by Ij,s
and N −2n points of the type s+n. Here, dist(x, Ij,s ∩K) = x−x0 ≤ δs+n for some x0.
Let x0 ∈ Ii,s+n ⊂ Im,s+n−1. Then Im,s+n−1 contains from 2 to 4 points of ZN . In all cases,
d1(x, ZN ) ≤ li,s+n + δs+n ≤ (C0 + 1)δs+n, by (3). Also, δs+n/2 ≤ d2 ≤ (C0 + 1)δs+n−1.
Here the lower bound corresponds to the case #(Ii,s+n ∩ ZN ) = 2, whereas the upper
bound deals with #(Im,s+n−1 ∩ ZN ) = 2. Similarly, d3 ≥ hm,s+n−1 − δs+n. From (4)
and (2) it follows that d3 ≥ 7/8 δs+n−1 − δs+n ≥ 27 δs+n. This gives δi−1
s+ndi+1 · · · dN ≤
(C0 + 1)d2 · · · dN for 0 ≤ i ≤ p and, by (14), the lemma follows.
(cid:3)
We can now formulate our main result.
Theorem 5.3. Suppose γ satisfies (2). Then K(γ) has the extension property if and
only if (5) is valid.
Proof. Recall that the extension property of a set is equivalent to the condition (DN)
of the corresponding Whitney space. Due to L. Frerick [Fr, Prop. 3.8], E(K) satisfies
(DN) if and only if for any ε > 0 and for any q ∈ N there exist r ∈ N and C > 0 such
that · 1+ε
r. Hence, in order to prove that (5) is necessary for EP of
K(γ), we can show that (8) implies the lack of (DN) for E(K(γ)), that is there exist
ε > 0 and q such that for any r ∈ N one can find a sequence (fj) ⊂ E(K(γ)) with
q ≤ C · 0,K · ε
fj1+ε
q
fj−1
0,K(γ) fj− ε
r → ∞ as
j → ∞.
Let us fix ε and m from the condition (8) and take q = 2m. For each fixed large r
(clearly, we can take it in the form r = 2n) and sj defined by (8), we consider the
function fj given in Lemma 5.1 for s = sj. Then
C fj1+ε
r ≥ (δn+s · δn+s−1 · δ2
n+s−m−1 · · · δ2n−1
n+s−2 · · · δ2m−1
n+s−m)−1(δ2m
0,K(γ) fj− ε
fj−1
)ε,
s
q
where C does not depend on j. The right side here goes to infinity. Indeed, its logarithm
is 2n+s {2Bn+s + Bn+s−1 + · · · + Bn+s−m − ε[Bn+s−m−1 + · · · + Bs]} and the expression in
braces exceeds Bn+s by (8). Therefore the whole value exceeds 2n+sBn+s = 1
,
which goes to infinity when s = sj increases. Thus, EP of K(γ) implies (5).
2 log 1
δs+n
For the converse, we consider the extension operator W from Section 5, where (ns)
are chosen as in (11). We proceed to show that W is bounded provided (10). Let us
fix any natural number p. This p and C1 from Lemma A define M = 2p + 2 + log(2C1).
10
ALEXANDER GONCHAROV AND ZELIHA URAL
We fix m ∈ N that corresponds to M it the sense of (10). Let q = 2m − 1 and r = r(q)
be defined by (1). We will show that the bound (W (f, x))(p) ≤ C f r is valid for
some constant C = C(p) and all f ∈ E(K), x ∈ R.
Given f and x, let us consider terms of accumulation sums. For fixed s ∈ N we
choose j ≤ 2s such that x ∈ Ij,s. Fix N with 2n ≤ N < 2n+1 for ns−1 − 1 ≤ n ≤ ns − 1,
so Ms + 1 ≤ N ≤ Ns. For large enough s the value N exceeds p and q. As in Lemma A,
let Z = (zk)N +1
k=1 = ZN ∪ {zN +1}. By Newton's representation of interpolating operator
in terms of divided differences, we have
AN,j,s(f, x) = [z1, · · · , zN +1]f · ΩN (x) u(x),
where ΩN and u are taken as in Lemma 5.2. We aim to show that
(15)
Ns A(p)
N,j,s(f, x) ≤ s−2 f r
for large s. This gives convergence of the accumulation sums.
Since divided differences are symmetric in their arguments, we can use (4) from [2] :
(16)
[z1, · · · , zN +1]f ≤ 2N − q f q (Π(J0))−1,
where Π(J0) = min1≤j≤N +1−qΠ(J) for Π(J) defined in Lemma B. Fix z ∈ J0 that
corresponds to this set in the sense of Lemma B.
Applying Lemma 5.2 and Lemma A for z = z yields
N +1
(ΩN · u)(p)(x) ≤ C δ−p
s+n N p C N
1
dk(z, Z) with C = 2p(C0 + 1) cp.
Yk=2
Recall that the set Z includes all points of the type ≤ s + n − 1 on Ij,s and N − 2n
k=2 dk(z, Z) if we will
consider only distances from z to points from Ys+n−1 ∩ Ij,s. Arguing as in Lemma 4.1,
s+n−m. We observe that d1(z, Z) = 0 is not
points of the type s + n. We can only enlarge the product Qq+1
we getQq+1
included into the product on the left side. By (10), Qq+1
In order to get (15), it is enough to show that
k=2 dk(z, Z) ≤ C q
k=2 dk(z, Z) ≤ C q
0 δM
s+n.
0δs+n−1 δ2
s+n−2 · · · δ2m−1
(17)
s2 Ns N p (2C1)N δM −p
s+n → 0 as s → ∞.
. Also, (2C1)N < δ− log(2C1)
Here, by (11), Ns N p < 2ns(p+1) ≤ log(1/δs)p+1 < δ−p−1
.
Clearly, we can replace δs+n in (17) by δs. Then, because of the choice of M, the
product in (17) does not exceed s2 δs, which approaches 0 as s → ∞, since, by (2),
δs ≤ 32−s.
s
s
Similar arguments are used for terms of the transition sums.
(cid:3)
On the other hand, (16) and Lemma B for J0 give
[z1, · · · , zN +1]f ≤ 2N − q f q
Combining these we see that
N +1
Yk=q+2
d−1
k (z, Z).
q+1
A(p)
N,j,s(f, x) ≤ C f q δ−p
s+n N p (2C1)N
dk(z, Z).
Yk=2
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
11
6. Extension of basis elements
Bases in the spaces E(K(γ))
7. Two examples
First we consider regular sequences (Bk)∞
(Bk)∞
Recall that (Bk)∞
k=1 is regular if, for some k0, both sequences (Bk)∞
k=k0 and (βk)∞
k=1 has subexponential growth if βk → 0 as k → ∞.
k=1. Let βk = (log Bk)/k. We say that
k=k0 are monotone.
k = k−a, γ(2)
k = a−k, γ(3)
For example, given a > 1, let γ(1)
k = exp(−ak) for large enough
k. Then γ(j) for 1 ≤ j ≤ 3 generate regular B(j) with B(1)
k ∼
2−k−2 k2 log a, B(3)
k → − log(a/2),
so B(j) are not of subexponential growth, except B(3) for a = 2. We see that (5) is valid
in the first two cases and in the third case with a ≤ 2.
k ∼ (a/2)k+1/(a−1). Here, β(1)
k ∼ 2−k−1 a k log k, B(2)
k ր − log 2 and β(3)
k , β(2)
More generally, (5) is valid for each monotone convergent (Bk)∞
k=1. Indeed, if Bk ց
B ≥ 0, then LHS of (5) does not exceed (n + 1)−1. If Bk ր B, then we take s0 with
Bs > B/2 for s ≥ s0. Then Bs+· · · Bs+n ≥ (n+1)B/2 and LHS of (5) < 2(n+1)−1. This
covers the case of regular sequences (Bk)∞
k=1 when βk are negative. Let us show that
(5) is valid as well for divergent regular sequences (Bk)∞
k=1 of subexponential growth.
Theorem 7.1. Let (Bk)∞
and only if (Bk)∞
k=1 be regular with positive values of βk. Then (5) is valid if
k=1 has subexponential growth.
Proof. A regular sequence (Bk)∞
k=1 is not of subexponential growth, provided βk > 0,
in the following three cases: βk ր β < ∞, βk ր ∞ and βk ց ε0 > 0. We aim to show
that (5) is not valid under the circumstances.
In the first case, given s and n, let b = exp βs+n. Then b − 1 ≥ exp β1 − 1 > β1 > 0
k=s Bk < bs+n+1/(b − 1) as Bk = exp(kβk) ≤ bk for such k.
k=s Bk > (b − 1)/b > β1/ exp β, which contradicts (5).
and b ≤ exp β. Here, Ps+n
Therefore, Bs+n/Ps+n
If βk ր ∞ then, by the same argument, Bs+n/Ps+n
where s0 is fixed with exp βs0 > 2.
k=s Bk > (b−1)/b > 1/2 for s ≥ s0,
Suppose βk ց ε0. We fix indices s1 < s2 < · · · such that the intervals Ij connecting
points (sj, βsj ) and (sj+1, βsj+1) form a convex envelope of the set (k, βk) on the plane.
We start from s1 = max{s : βs = β1}. If sj is chosen, then we take sj+1 with he
property: for each k with sj ≤ k ≤ sj+1 the point (k, βk) is not over Ij. At any step
we can take the next value so large that the slopes of Ij increases to zero. In addition,
given sj, we take sj+1 such that
(18)
(4 − 2 sj/sj+1)βsj+1 ≥ (3 − sj/sj+1)βsj ,
which is possible as βk decreases to a positive limit.
For fixed j, we take s = sj and s+n = sj+1. Let βk = ak +b with a = −(βs −βs+n)/n
and b = βs + (βs − βs+n) s/n for s ≤ k ≤ s + n, so the points (k, βk) are located just on
the interval Ij. Also, let g(x) = ax2 + bx and Bk = exp g(k) = exp(k βk) on [s, s + n].
Of course, Bs = Bs and Bs+n = Bs+n.
It is easy to check that the function g increases on this interval. Hence, Ps+n
Bk < R s+n
Ps+n
[2a(n + s) + b]−1 − g(s) · [2as + b]−1 + 2aR s+n
g(x) dx + Bs+n. By integration by parts, R s+n
k=s Bk ≤
g(x) dx = g(n + s) ·
g(x)(2ax + b)−2dx. We neglect the last
k=s
s
s
s
12
ALEXANDER GONCHAROV AND ZELIHA URAL
s
term, as a < 0, and the second term, as 2as + b = g′(s) > 0. Also, 2a(n + s) + b =
g(x) dx < 2 Bs+n/ε0
(2 + s/n)βs+n − (1 + s/n)βs ≥ βs/2 ≥ ε0/2, by (18). Hence R s+n
and Bs+n/Ps+n
k=s Bk > ε0/(2 + ε0), so (5) is not valid.
growth of (Bk)∞
We proceed to show that (5) is valid for βk ց 0, that is in the case of subexponential
k=s Bk from below. Let
k=1. Here, for fixed large s and n, we estimatePs+n
b = exp βs+n. Then Bk ≥ bk for s ≤ k ≤ s + n. Therefore, Bs+n/Ps+n
If bn < 2 for the given s and n then bn+1 − 1 > (n + 1)βs+n. On the other hand,
exp βs+n − 1 < 2 βs+n for βs+n < 1. Thus the fraction above does not exceed 4/(n + 1).
Otherwise, bn ≥ 2 and bn < 2(bn+1 − 1). Here the fraction does not exceed 4 βs+n. It
k=s Bk ≤ max{4/(n + 1), 4 βn}, which is the desired conclusion.
(cid:3)
k=s Bk ≤ bn(b−1)
bn+1−1 .
follows that Bs+n/Ps+n
Our next objective is to consider irregular sequences (Bk)∞
k=1 (compare with Ex.6
in [10]). Given two sequences, (kj)∞
j=1 with
εj ց 0, let γk = (k + 5)−2 for k 6= kj and γkj = (kj + 5)−2εj. Then γ satisfies (2) with
δk = (5!/(k + 5)!)2 ε1ε2 · · · εj for kj ≤ k < kj+1. Let Aj := log
. We will consider
only sequences with the property
j=1 ⊂ N with kj+1 − kj ր ∞ and (εj)∞
ε1ε2···εj
1
(19)
j+1 · A−1
k2
j → 0 as j → ∞.
Provided this condition, Bk = 2−k log (k+5)!
In addition, an easy computation shows that for large j,
5! + 2−k−1 Aj ∼ 2−k−1 Aj for kj ≤ k < kj+1.
(20)
Bkj + Bkj+1 + · · · + Bkj+1−1 < 3 Bkj .
Now we can construct different examples of compact sets K(γ) without extension
property.
Example 2. Let Aj = 2kj , so εj = exp(−2kj + 2kj−1) for j ≥ 2 and ε1 = exp(−2k1).
In this case, (19) is valid under mild restriction 2−kj k2
j+1 → 0 as j → ∞. Let us take
s = kj, n = kj+1 − kj. Then Bs+n > 2−kj+1−1 Aj+1 = 1/2 and, by (20), Bs + · · · +
Bs+n−1 < 3 Bs < 4 · 2−kj−1 Aj = 2. This gives (8) with ε = 1/4 and m = 0.
8. Extension Property of K(γ) and Hausdorff measures
From now on, h is a dimension function, which means that h : (0, T ) → (0, ∞) is
continuous, nondecreasing and h(t) → 0 as t → 0. The h−Hausdorff content of E ⊂ R
is defined as
and the h−Hausdorff measure of E is
Mh(E) = inf{X h(Gi) : E ⊂ ∪Gi}
Λh(E) = lim inf
δ→0
{X h(Gi) : E ⊂ ∪Gi, Gi ≤ δ}.
Here we consider at most countable coverings of E by intervals (open or closed).
It is easily seen that Mh(E) = 0 if and only if Λh(E) = 0. We write h1 ≺ h2 if
h1(t) = o(h2(t)) as t → 0. Let h1 ≈ h2 if C −1h1(t) ≤ h2 ≤ Ch1(t) for some constant
C ≥ 1 and 0 < t ≤ t0 < T. We will denote by h0 the function h0(t) = (log 1
t )−1 with
0 < t < 1, which defines the logarithmic measure of sets.
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
13
A set E is called dimensional if there is at least one dimension function h that makes
E an h−set, that is 0 < Λh(E) < ∞. In our case, the set K(γ) is dimensional. In [1],
following Nevanlinna [Nev], the corresponding dimension function was presented. Let
η(δk) = k for k ∈ Z+ with δ0 := 1 and η(t) = k + log δk
for δk+1 < t < δk.
Then h(t) := 2−η(t) for 0 < t ≤ 1. Clearly, h(δk) = 2−k.
t / log δk
δk+1
Proposition 8.1. Let γ satisfy (2) and h be defined as above. Then Λh(K(γ)) = 1.
Proof. Take t = C0 δk, where C0 is given in (3). Then δk < t = C0 γk δk−1 < δk−1
for large enough k. Here, η(t) = k − log C0/ log(1/γk) and h(t) = 2−k ak with ak :=
exp log C0·log 2
log(1/γk ) . Since γk → 0, given ε > 0, there is k0 such that ak < 1 + ε for k ≥ k0.
j=1 h(lj,k) < 2k h(t) < 1 + ε provided that
j=1 h(lj,k) for each k. Since ε is arbitrary, we get
From (3) it follows that 1 = 2k h(δk) < P2k
k ≥ k0. Of course, Λh(K(γ)) ≤ P2k
Let us show that Λh(K(γ)) ≥ 1. Fix ε > 0 and choose k0 such that
Λh(K(γ)) ≤ 1.
(21)
ε log 1/γk > − log 2 · log(1 − 4γk) for k ≥ k0.
This can be done as γk → 0. Take any open covering ∪Gi of K(γ). Given ε, we can
consider only coverings with Gi < δk0 for each i. We choose a finite subcover ∪N
i=1Gi
of K(γ).
Fix i ≤ N and k with δk+1 < Gi ≤ δk. Recall that, for each j ≤ 2k, we have
hj,k > (1 − 4γk+1)lj,k. Therefore the distance between any two basic intervals from Ek+1
exceeds (1 − 4γk+1)δk. If Gi < (1 − 4γk+1)δk then Gi can intersect at most one interval
from Ek+1. In this case we can consider only Gi ≤ max1≤j≤2j+1 lj,k+1 ≤ C0δk+1, by (3).
Thus there are two possibilities: δk+1 < Gi ≤ C0δk+1 or (1 − 4γk+1)δk < Gi ≤ δk.
In the first case we have h(Gi) > 2−k−1. Here, Gi intersects at most one interval from
Ek+1 and, by construction, at most 2m−k−1 interval from Em for m > k. In turn, in the
latter case, h(Gi) > 2−k(1 − ε). Indeed, here, η(Gi) < k − log(1 − 4γk+1)/ log(1/γk+1)
and h(Gi) > 2−k a, where a = exp log(1−4γk+1)·log 2
log(1/γk+1) > (1−ε), by (21). Now Gi intersects
at most two interval from Ek+1 and at most 2m−k interval from Em.
Let us choose m so large that each basic interval from Em belongs to some Gi, perhaps
not to unique. We decompose all intervals from Em into two groups corresponding to
i 2m−k <
the cases considered above. Counting intervals gives 2m ≤ P′
2m[P′
i h(Gi)(1 − ε)−1]. From this we see that Pi h(Gi) > 1 − ε, which
is the desired conclusion, as ε and (Gi) here are arbitrary.
i h(Gi) +P′′
i 2m−k−1 +P′′
(cid:3)
The same reasoning applies to a part of K(γ) on each basic interval.
Corollary 8.2. Let γ and h be as in Proposition above, k ∈ N, 1 ≤ j ≤ 2k. Then
Λh(K(γ) ∩ Ij,k) = 2−k.
Suppose, in addition, that K(γ) is not polar. Then, by Corollary 3.2 in [1], µK(γ)(Ij,k) =
2−k, so the values of µK(γ) and the restriction of Λh on K(γ) coincide on each basic
interval. From here, by Lemma 3.3 in [1], these measures are equal on K(γ). Thus, a
non-polar set K(γ) satisfying (2) is indeed equilibrium Cantor-type set if we accept for
definition of this concept the condition µK = ΛhK(γ), which is more natural than the
14
ALEXANDER GONCHAROV AND ZELIHA URAL
definition suggested in in [WE], Section 6.
We recall that there is no complete characterization of polarity of compact sets in
terms of Hausdorff measures, see e.g Chapter V in [14]. On the one hand, a set is polar
if its logarithmic measure is finite. This defines a zone Zpol in the scale of growth rate
of dimension functions consisting of h with lim inf t→0 h(t)/h0(t) > 0. If h ∈ Zpol and
Λh(K) < ∞ then Cap(K) = 0. On the other hand, functions with R0 h(t)/t dt < ∞
form a nonpolar zone Znp : if h ∈ Znp and Λh(K) > 0 then Cap(K) > 0. But, by Ursell
[21], the remainder makes up a zone Zu of uncertainty. One can take two functions in
this zone with h2 ≺ h1 and sets K1, K2, where Kj is a hj−set, such that K2 is polar, K1
is not, though in the sense of Hausdorff measure the set K2 is larger than K1. Indeed,
Λh2(K2) > 0, but Λh2(K1) = 0 or Λh1(K2) = ∞, but Λh1(K1) < ∞.
Let us show that a similar circumstance is valid with the extension property.
Proposition 8.3. There are two dimension functions h2 ≺ h1 and two sets K1, K2,
where Kj is an hj−set for j ∈ {1, 2}, such that the smaller set K1 has the extension
property, whereas the larger set K2 does not have.
Proof. Take K1 from Example 1. Let us show that the corresponding function h1 = 2−η1
is equivalent to h0. It is enough to find C > 0 such that η0(t) − C ≤ η1(t) ≤ η0(t) + C
for small t. Here, η0(t) = (log log 1/t)/ log 2, so h0(t) = 2−η0(t). For the set K1 we have
δk = exp(−2k+1B) and η0(δk) = k + log 2B/ log 2. If δk+1 < t ≤ δk for some k, then
k ≤ η1(t) < k + 1 and k + log 2B/ log 2 ≤ η0(t) < k + 1 + log 2B/ log 2, which gives
h1 ≈ h0.
j=1 satisfies 2−kj 2j k2
In turn, let K2 be as in Example 2 with Aj = 2kj 2−j and εj = exp(−Aj + Aj+1) for
j ≥ 2. Here we suppose that (kj)∞
j+1 → 0 as j → ∞. Then (19)
and (20) are valid, which, as in Example 2, gives the lack of the extension property
for K2. Let us show that h2 ≺ h0. It is enough to check that η2(t) − η0(t) → ∞ as
t → 0. Let δk < t ≤ δk−1 with kj ≤ k < kj+1 for large enough j. Then log 1/δk =
2 log((k + 5)!/5!) + Aj < 2 Aj and η0(t) < η0(δk) < kj + 1 − j. On the other hand,
η2(t) ≥ η2(δk−1) = k − 1 ≥ kj − 1. Therefore, η2(t) − η0(t) > j − 2, which completes
the proof.
(cid:3)
One can suppose that, for the considered family of sets, the scale of growth rate of
dimension functions can be decomposed as above into three zones. If K(γ) is an h−set
for a function h with moderate growth then the set has EP .
If the corresponding
function h is large enough, then EP fails. Proposition above shows that the zone of
uncertainty here is not empty.
We see that h = h0 is not the largest function which allows EP for h−sets K(γ).
If, as in the regular case, we take Bk ր ∞ of subexponential growth, then δk =
exp(−2k+1Bk) and h0(δk) = 2−k−1B−1
k , which is essentially smaller than h(δk) = 2−k
for the corresponding function h.
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
15
Example 3. Let log(m) t denote the m−th iteration log · · · log t for large enough t.
The sequence Bk = exp(k/ log(m) k) has subexponential growth. Then the correspond-
ing sequence (γk)∞
k=1 satisfies (2), as for large k we have γk = δk/δk−1 < exp(−2kBk) <
exp(−2k) and for the previous k we can take γk = 1/32. By Theorem 7.1, the set
K(γ) has EP . Let us find a dimension function h that corresponds to this set. We
will search it in the form h(t) = hα(t)
(t). Let t = δk. Then log 1/t = 2k+1Bk, so
k ∼ (log log 1/t)/ log 2. On the other hand, h(t) = 2−k = (2k+1Bk)−α(t), which gives
α(t) ∼ 1 − (log 2 · log(m) k)−1 ∼ 1 − (log 2 · log(m+2) 1/t)−1. Clearly, h ≻ h0.
0
The next Proposition generalizes Example 3. We restrict our attention to strictly
increasing functions h of the form h = hα
0 , where α is a monotone function on [0, t0].
Since in the next section we shall be interested in considering of dimension functions
exceeding h0, let us suppose that α(t) ≤ 1. Then h ≻ tσ for each fixed σ > 0.
In addition we assume that asymptotically
(22)
h(t) ≤ 2 h(t2),
which is valid for typical dimension functions corresponding to the cases
a) α(t) = α0 ∈ (0, 1],
b) α(t) = α0 + ε(t) with α0 ∈ [0, 1),
c) α(t) = 1 − ε(t).
Here,
(23)
ε(t) ց 0 with ε(t) log log 1/t ր ∞ as t ց 0,
since for slowly increasing ε we get hα0±ε ≈ hα0
0 .
By (22), for the inverse function h−1, we have h−1(τ ) ≤ (h−1(2τ ))2 and h−1 ≺ τ M
for M given beforehand. From this, γk = h−1(2−k)/h−1(2−k+1) defines a sequence
satisfying (2). We denote the corresponding set by K α(γ). Our aim is to check EP
for this set provided regularity of the sequence Bk = 2−k−1 log(1/h−1(2−k)). We see at
once that Bk increases. In its turn, βk ց 0 if α0 = 1 in the case (a), βk ր 1/α0 − 1 in
(b) and in (a) with α0 < 1. Concerning (c), the monotonicity of βk requires additional
rather technical restrictions on ε. At least for ε(t) = εm(t) := (log(m) 1/t)−1 we have
βk ց 0. Here, m ≥ 3, as h ≈ h0 for m ∈ {1, 2}.
Proposition 8.4. Let K α(γ) be defined by a function h, as above, with a regular
sequence (Bk)∞
k=1. Then K α(γ) has the extension property if and only if
(cid:18)log
1
h−1(2−k)(cid:19)1/k
→ 2 as k → ∞.
Proof. Let us find h−1 for the case α(t) = 1−ε(t). If h(t) = τ then [1−ε(t)] log log 1/t =
log 1/τ. Let us define a function δ by the condition log log 1/t = [1 + δ(τ )] log 1/τ. Then
[1 − ε(t)][1 + δ(τ )] = 1, so δ(τ ) ց 0 as τ ց 0. Then t = h−1(τ ) = exp[−(1/τ )1+δ(τ )] and
log(1/h−1(2−k)) = 2k(1+δ(2−k )). The k−th root of this expression tends to 2. On the
other hand, (Bk)∞
k=1 here has subexponential growth as βk = (δ(2−k) − 1/k) log 2 → 0.
By Theorem 7.1, K α(γ) has EP.
16
ALEXANDER GONCHAROV AND ZELIHA URAL
Similarly, if α(t) = α0 + ε(t) with 0 < α0 < 1 then h−1(τ ) = exp[−(1/τ )1/α0−δ(τ )].
Here, (log(1/h−1(2−k)))1/k = 2(1/α0−δ(2−k)) 9 2 and βk 9 0, there is no EP. In the case
(a), the function δ vanishes.
Lastly, α0 = 0 in (b) gives h−1(τ ) = exp[−(1/τ )∆(τ )] with ∆(τ ) ր ∞ as τ ց 0.
(cid:3)
Here, (log(1/h−1(2−k)))1/k → ∞ and βk → ∞.
9. Extension Property and densities of Hausdorff contents
To decide whether a set K has EP, we have to consider a local structure of the most
rarefied parts of K. Obviously, such global characteristics as Hausdorff measures or
Hausdorff contents cannot be applied in general for this aim. Instead, one can suggest
to describe EP in terms of lower densities of Mh or related functions. Given a dimension
function h, a compact set K, x ∈ K and r > 0, let ϕh,K(x, r) := Mh(K ∩ B(x, r)) and
ϕh,K(r) := inf x∈K ϕh,K(x, r), where B(x, r) = [x−r, x+r]. One can suppose that K has
EP if and only if the corresponding function ϕh,K is not very small, in a sense, as r → 0.
Essentially, this is similar to analysis of the lower density of the Hausdorff content,
which can be defined as φh(K) := lim inf r→0 inf x∈K
. Indeed, Mh(B(x, r)) =
Mh(K∩B(x,r))
Mh(B(x,r))
h(2r) for h with h(t) ≻ t and the expression above is lim inf r→0
ϕh,K (r)
h(2r) .
In order to distinguish EP by means of φh, we have to consider large enough di-
mension functions h. Indeed, if for some h1 with h1 ≻ h there exists h1−set K1 with
EP , then h cannot be used for this aim, because Λh(K1) = 0 implies Mh(K1) = 0 and
the corresponding density vanishes contrary to our expectations. Therefore, we can
consider only functions exceeding h0.
We remark that Λh−analogs of ϕh,K or φh cannot be applied in general for distin-
guishing EP, since for fat sets (K = Int(K)) we have Λh(K ∩ B(x, r)) = ∞ provided
h(t) ≻ t.
Interestingly, it turns out that the lower density φh can be used to characterize EP
for the family of compact sets considered in [6].
with Qk ≥ 2, let K = {0} ∪S∞
Example 4. Given two sequences bk ց 0 (for brevity, we take bk = e−k) and Qk ր
k . In what follows we
will consider two cases: Qk ≤ Q with some Q and Qk ր ∞ with Qk < log k for large
k. By Theorem 4 in [6], K has the extension property in the first case and does not
have it for unbounded (Qk).
k=1 Ik, where Ik = [ak, bk], Ik = bQk
In the next lemma we consider concave dimension functions h = hα
0 for the cases
(a), (b), as above, and for more general
c ′) α(t) = α0 − ε(t) with α0 ∈ (0, 1].
We suppose now that ε is a monotone differentiable function on [0, t0] with 0 < ε(t) <
1 − α0 in (b) and 0 < ε(t) < α0/2 in (c ′). As before, we assume (23). A direct
computation shows that
(24) h′(t) < h(t) h0(t) α(t)/t for the cases (a), (b) and h′(t) < h(t) h0(t)/t for (c ′).
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
17
k=nIk) = h(bn). This means that the covering of the set ∪∞
Lemma 9.1. Suppose intervals Ik are given as in Example 4 and n is large enough.
Then Mh(∪∞
k=nIk by the
interval [0, bn] is optimal in the sense of definition of Mh.
Proof. Let us fix any open covering of K, choose its finite subcovering ∪M
i=1Gi and
enumerate sets Gi from left to right. We can suppose that G1 covers ∪∞
k=N Ik for some
N ≥ n. Indeed, if G1 covers as well some part of IN −1, then other part of IN −1 is
covered by G2. In this case, association of G1 and G1 into one interval will give better
covering, since h(b) ≤ h(x) + h(b − x) for 0 ≤ x ≤ b, by concavity of h. For the same
reason, we suppose that each Gi covers entire number of Ik. After this we reduce each
Gi to the minimal closed interval Fi containing the same intervals Ik. Thus, F1 = [0, bN ]
and F2 = [aN −1, bq] with some N − 1 ≤ q ≤ n. Our aim is to show that
(25)
h(bq) < h(bN ) + h(bq − aN −1),
so replacing F1 ∪ F2 with [0, bq] is preferable. We use the mean value theorem and the
decrease of h′. Note that h(bk) = k−α(bk).
Consider first the value q = N − 1. We will show h(bN −1) − h(bN ) < h(IN −1).
In the cases (a), (b), by (24), LHS < h′(bN )e−N (e − 1) < N −1−α(bN )α(bN )(e − 1).
On the other hand, h(IN −1) = [QN −1(N − 1)]−α(IN −1). Here, α(IN −1) < α(bN ), so
we reduce the desired inequality to (QN −1/N)α(bN )α(bN )(e − 1) < 1. It is valid, since
for α0 > 0 the first term on the left goes to zero, whereas for α0 = 0 in (b) we have
α(bN ) = ε(bN ) → 0 as N → ∞.
Similarly, in the case (c ′) the inequality [QN −1(N −1)]α0−ε(IN −1)(e−1) < N 1+α0−ε(bN )
is valid, as is easy to check.
Suppose now that q ≤ N − 2. We write (25) as h(bq) − h(bq − aN −1) < h(bN ).
Here, in all cases, by (24), LHS < h′(bq−aN −1) aN −1 < h(bq)h0(bq) aN −1
bq−aN −1
, where the
last fraction does not exceed bN −1
. On the other hand, h(bN ) ≥ N −1 as α(bN ) ≤ 1.
bq−bN −1
Hence it is enough to show that N < (eN −q−1 − 1) q1+α(bq). We neglect α(bq) and notice
that (eN −q−1 − 1) q ≥ (e − 1)(N − 2), which completes the proof of (25).
Continuing in this manner, we see that h(bn) ≤PM
Corollary 9.2. Suppose bn+1 ≤ r ≤ bn − bn+1. Then ϕh,K(r) = h(In).
i=1 h(Fi).
(cid:3)
Proof. Clearly, ϕh,K(x, r) = h(In) for each x ∈ In. If x ∈ K ∩ [0, bn+1] then B(x, r)
covers all intervals Ik with k ≥ n + 1. By Lemma, ϕh,K(x, r) = h(bn+1) > h(In). Of
course, for x ∈ Ik with k < n the value ϕh,K(x, r) also exceed h(In).
(cid:3)
Remark. The covering of two (or small number of) intervals Ik by one interval is
not optimal, since Mh(Ik ∪ Ik+1) = h(Ik) + h(Ik+1) < h(bk − ak+1).
We proceed to characterize EP for given compact sets in terms of lower densities φh
for h = hα
0 , where
α(t) = α0 ∈ (0, 1] or α(t) = α0 ± εm(t)
(26)
with 0 < α0 < 1 and εm(t) = (log(m) 1/t)−1 for m > 2, so (23) is valid.
Proposition 9.3. Let K be from the family of compact sets given in Example 4 and
h be as above. Then K has the extension property if and only if φh(K) > 0.
18
ALEXANDER GONCHAROV AND ZELIHA URAL
Proof. Suppose first that Qk ≤ Q with some Q, so K has EP . We aim to show
ϕh,K (r)
h(2r) > 0. Let e−k−1 ≤ r < e−k for some k. Then, as ϕh,K increases, ϕh,K(r) ≥
limr→0
ϕh,K(e−k−1), which is h(Ik) = (k · Qk)−α(Ik), by Corollary 9.2. On the other hand,
h(2r) < h(2e−k) = (k − log 2)−α(2e−k). Therefore,
ϕh,K(r)/h(2r) > Q−α(Ik)
k
kα(2e−k)−α(Ik) (1 − log 2/k)−α(2e−k).
The first term on the right converges to Q−α0 as k → ∞. The second and the third
terms converge to 1. Hence, φh(K) ≥ Q−α0. Besides, this value is achieved in the case
Qk = Q by the sequence rk = bk − bk+1. Thus, φh(K) > 0. In addition, for given σ > 0
we have a compact sets K with EP such that 0 < φh(K) < σ.
Similar arguments apply to the case Qk ր ∞, when K does not have EP. Here,
φh(K) ≤ limk ϕh,K(rk)/h(2rk) for rk as above. By Corollary 9.2, ϕh,K(rk) = h(Ik).
Also, h(2rk) > h(e−k). Hence, ϕh,K(rk)/h(2rk) < Q−α0/2
kα(e−k)−α(Ik), which converges
to 0 as k increases.
(cid:3)
k
Remark. For this family of sets, the extension property can be characterized as
well in terms of the Lebesgue linear measure λ. Let λ(r) := inf x∈K λ(K ∩ [x − r, x + r]).
Then K has the extension property if and only if lim inf r→0 λ(r) · r−Q > 0 for some Q.
Nevertheless, at least for dimension functions h = hα
0 with α as in (26), there is no
general characterization of EP in terms of lower densities φh. In view of Example 3
and the discussion in the beginning of the section, the value α0 = 1 can be omitted
from consideration
We treat now regular sets K(γ) with δk = exp(−bk). Here, Bk = 2−1(b/2)k. By
Theorem 7.1, K(γ) has EP if b = 2 and does not have it for b > 2.
Lemma 9.4. For each constants C ≥ 1 and h, as above, there is b > 2 such that
h(Cδk) < 2 h(δk+1) for large enough k. This inequality is valid also for b = 2.
Proof. In all cases we have h(δk) = b−k·α(δk) and the desired inequality has the form
(27)
b(k+1)α(δk+1) < 2 (bk − log C)α(Cδk).
Suppose α ≡ α0. Then (27) is valid as bα0 < 2 (1−b−k log C) for large k and b = 2+σ
with small enough σ. All the more it is valid for b = 2.
The same reasoning applies to the case α = α0 + ε(t) with ε ր as ε(δk+1) < ε(Cδk).
In the last case α(t) = α0 − εm(t) we use the following simple inequality
log(m)(Cx) − log(m)(x) < log C · [log x log(2)(x) · · · log(m−1)(x)]−1,
which is valid for all x from the domain of definition of log(m). From this we have
k · [ε(Cδk) − ε(δk+1)] → 0 as k → ∞ and (27) can be treated as in the first case.
(cid:3)
Corollary 9.5. Let k be large enough.Then the covering of each basic interval Ij,k of
K(γ) by one interval is better (in the sense of definition of Mh) than covering by two
adjacent subintervals.
Indeed, by (3), h(lj,k) < h(C0δk) < 2 h(δk+1) < h(l2j−1,k+1) + h(l2j,k+1).
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
19
Remark. It is essential that coverings of a whole basic interval are considered. For
example, for the set I1,k∪I3,k+1 we have h(l1,k)+h(l3,k+1) < h(b3,k+1), which corresponds
to the covering of the set by one interval.
Proposition 9.6. Let h = hα
Then φh(K(γ)) = b−α0.
0 with α as in (26) and K(γ) be defined by δk = exp(−bk).
Proof. For brevity, we denote here K(γ) by K. Fix x ∈ K. Let x ∈ Ij,k ⊂ Ii,k−1 and
C0 δk ≤ r ≤ 7/8 · δk−1. Then, by (3), lj,k ≤ r < hi,k−1 and K ∩ [x − r, x + r] = K ∩ Ij,k.
Arguing as in Lemma 9.1, by Lemma 9.4, we get ϕh,K(x, r) = h(lj,k). Therefore, by
monotonicity, h(δk) < ϕh,K(x, r) < h(C0δk) for each x ∈ K.
We proceed to estimate φh(K) from both sides. Suppose that C0 δk ≤ r ≤ C0 · δk−1
for some k. Then h(δk) < ϕh,K(r) < h(C0δk−1) and
h(δk)
h(2C0δk−1)
<
ϕh,K(r)
h(2r)
<
h(C0δk−1)
h(2C0δk)
.
Here, δk = δb
k−1. Analysis similar to that in the proof of Lemma 9.4 shows that the
first fraction above has the limit b−α0, whereas the last fraction tends to bα0 as k → ∞.
Moreover, the value b−α0 can be achieved as limk ϕh,K(rk)/h(2rk) for rk = 7/8·δk−1. (cid:3)
Comparison of Propositions 9.3 and 9.6 shows that, for given dimension functions,
lower densities of Hausdorff contents cannot be used in general to characterize the
extension property.
10. Extension Property and growth of Markov's factors
Let Pn denote the set of all holomorphic polynomials of degree at most n. For any
infinite compact set K ⊂ C we consider the sequence of Markov's factors
Mn(K) = inf{M : P ′0,K ≤ M P 0,K, P ∈ Pn}
for n ∈ N. We see that Mn(K) is the norm of the operator of differentiation in the space
(Pn, · 0,K). We say that a set K is Markov if the sequence (Mn(K)) is of polynomial
growth. This class of sets is of interest to us, since, by W.Ple´sniak [16], any Markov
set has EP. On the other hand, there exist non-Markov compact sets with EP ([5],
[2]). We guess that there is some extremal growth rate (mn)∞
n=1 with the property: if,
for some compact set K, Mn(K)/mn → ∞ as n → ∞ then K does not have EP. The
next proposition asserts that here, as above, there is a zone of uncertainty, in which
growth rate of Markov's factors is not related with EP. In this sense, it is an analog
of Proposition 8.3.
Proposition 10.1. There are two sets K1 with EP and K2 without it, such that
Mn(K1) grows essentially faster than Mn(K2) as n → ∞.
Proof. By Theorem 6 in [we], M2k (K(γ)) ∼ 2/δk. By monotonicity, δ−1
4 δ−1
Example 1, so δ(1)
Aj = 2kj . For simplicity, we fix kj = j2 that satisfies (19). Here, δ(2)
k < Mn(K(γ)) <
k+1 for 2k ≤ n < 2k+1 with large enough k. As in Proposition 8.3, we take K1 from
k = exp(−2k+1B) with B > 1. Also, we use K2 from Example 2 with
k > k−2k ε1ε2 · · · εj
20
ALEXANDER GONCHAROV AND ZELIHA URAL
for kj ≤ k < kj+1. We aim to show that Mn(K2)/Mn(K1) → 0 as n → ∞. Let us fix
large n with 2k ≤ n < 2k+1. For this k we fix j with kj ≤ k < kj+1. Then
(28)
Mn(K2)/Mn(K1) < 4 δ(1)
k /δ(2)
k+1.
Suppose first that k ≤ kj+1 − 2. Then RHS of (28) does not exceed 4 exp[−2k+1B +
j+1,
2(k + 1) log(k + 1) + Aj]. The expression in brackets is smaller than 2kj (1 − 2B) + k2
which is (j + 1)4 − (2B − 1) 2j 2, so it tends to −∞ as j → ∞.
If k = kj+1−1 then RHS of (28) is smaller than 4 exp[−2kj+1B+2kj+1 log kj+1+Aj+1],
(cid:3)
which goes to 0, since B > 1. This completes the proof.
Existence of a zone of uncertainty (for the extension property) in the scale of growth
rate of Markov's factors implicates the problem to find boundaries of this zone.
References
[1] G. Alpan- A. Goncharov, Two measures on Cantor sets, J. Approx. Theory 186(2014), 28-32.
[2] M. Altun-A. Goncharov, A local version of the Pawlucki-Plesniak extension operator, Journal
of Approx. Theory, 132(2004), 34-41.
[3] L. Frerick, Extension operators for spaces of infinite differentiable Whitney jets, J. Reine
Angew. Math. 602 (2007), 123-154.
[4] L. Frerick, E. Jorda and J. Wengenroth, Tame linear extension operators for smooth
Whitney functions,J. Funct. Anal., 261 (2011), no. 3, 591-603.
[5] A. Goncharov, A compact set without Markov's property but with an extension operator for
C∞-functions, Studia math., 119(1996), 27-35.
[6] A. Goncharov-M. Kocatepe, Isomorphic Classification of the Spaces of Whitney Functions,
Michigan Math. J., 44(1997), 555-577.
[7] A. Goncharov, Bases in the spaces of C∞-functions on Cantor-type sets, Constr. Approx. 23
(2006), no. 3, 351-360.
[8] A. Goncharov, On the geometric characterization of the extension property, Bull. Belg. Math.
Soc. Simon Stevin, 14(2007), 513-520.
[9] A. Goncharov,Weakly Equilibrium Cantor-Type Sets, Potential Anal., 40(2014),143-161.
[10] A. Goncharov, Best exponents in Markov's inequalities, Math. Inequal. Appl. 17(2014), no. 4,
1515-1527.
[11] A. Goncharov- Z. Ural, Bases in some spaces of whitney functions, in process...
[12] R. Meise- D. Vogt, Introduction to functional analysis, Oxford Graduate Texts in Mathematics,
2. The Clarendon Press, Oxford University Press, New York (1997), x+437 pp.
[13] B.S. Mityagin, Approximate dimension and bases in nuclear spaces, Russian Math. Surveys,
16:4(1961), 59-127.
[14] R. Nevanlinna, Analytic Functions, Die Grundlehren der mathematischen Wissenschaften, vol.
162, Springer-Verlag, New York, Berlin (1970), viii+373 pp.
[15] W. Pawlucki and W. Plesniak,Extension of C∞ functions from sets with polynomial cusps,
Studia Math., 88(1988), 279-287.
[16] W. Plesniak,Markov's Inequality and the Existence of an Extension Operator for C∞ functions,
Jour. of Approx. Theory, 61(1990), 106-117.
[17] T. Ransford, Potential theory in the complex plane, London Mathematical Society Student
Texts, 28. Cambridge University Press, Cambridge, (1995), x+232 pp.
[18] R.T. Seeley, Extension of C∞ functions defined in a half space, Proc. Amer. Math. Soc.
15(1964), 625-626.
[19] E.M.Stein, Singular integrals and differentiability properties of functions,Princeton Univ.Press
(1970).
[20] M. Tidten, Fortsetzungen von C∞- Funktionen, welche auf einer abgeschlossenen Menge in Rn
definiert sind, Manuscripta Math., 27(1979), 291-312.
MITYAGIN'S EXTENSION PROBLEM. PROGRESS REPORT
21
[21] H. D. Ursell, Note on the Transfinite Diameter, J. London Math. Soc.,s1-13 (1938),34-37.
[22] D. Vogt,Sequence space representations of spaces of test functions and distributions in Func-
tional Analysis, Holomorphy and Approximation Theory, G.I. Zapata (ed.), Lecture Notes in
Pure and Appl. Math., 83(1983), 405-443.
[23] H. Whitney,Analytic extensions of differentiable functions defined in closed sets, Trans. Amer.
Math. Soc., 36(1934),63-89.
|
1301.2606 | 1 | 1301 | 2013-01-11T21:10:17 | Affine invariant points | [
"math.FA"
] | We answer in the negative a question by Gruenbaum who asked if there exists a finite basis of affine invariant points. We give a positive answer to another question by Gruenbaum about the "size" of the set of all affine invariant points. Related, we show that the set of all convex bodies K, for which the set of affine invariant points is all of n-dimensional Euclidean space, is dense in the set of convex bodies. Crucial to establish these results, are new affine invariant points, not previously considered in the literature. | math.FA | math |
Affine invariant points. ∗
Mathieu Meyer, Carsten Schutt and Elisabeth M. Werner †
Abstract
We answer in the negative a question by Grunbaum who asked if there exists a
finite basis of affine invariant points. We give a positive answer to another question
by Grunbaum about the "size" of the set of all affine invariant points. Related, we
show that the set of all convex bodies K, for which the set of affine invariant points
is all of Rn, is dense in the set of convex bodies. Crucial to establish these results,
are new affine invariant points, not previously considered in the literature.
1
Introduction.
A number of highly influential works (see, e.g., [8, 10, 12], [14]-[18], [20]-[31], [37, 44, 51,
52, 58]) has directed much of the research in the theory of convex bodies to the study of
the affine geometry of these bodies. Even questions that had been considered Euclidean
in nature, turned out to be affine problems - among them the famous Busemann-Petty
Problem (finally laid to rest in [6, 9, 56, 57]).
The affine structure of convex bodies is closely related to the symmetry structure
of the bodies. From an affine point of view, ellipsoids are the most symmetric convex
bodies, and simplices are considered to be among the least symmetric ones. This is
reflected in many affine invariant inequalities (we give examples below) where ellipsoids
and simplices are the extremal cases. However, simplices have many affine symmetries.
Therefore, a more systematic study for symmetry of convex bodies is needed. Grunbaum,
in his seminal paper [13], initiated such a study. A crucial notion in his work, the affine
invariant point, allows to analyze the symmetry situation. In a nutshell: the more affine
invariant points, the fewer symmetries.
In this paper, we address several issues that were left open in Grunbaum's paper.
For instance, it was not even known whether there are "enough" affine invariant points.
We settle this in Theorem 3 below.
Let Kn be the set of all convex bodies in Rn (i.e., compact convex subsets of Rn with
nonempty interior). Then (see Section 2 for the precise definition) a map p : Kn → Rn
is called an affine invariant point, if p is continuous and if for every nonsingular affine
map T : Rn → Rn one has,
p(T (K)) = T (p(K)).
∗Keywords: affine invariant points, symmetry. 2010 Mathematics Subject Classification: 52A20,
53A15
†Partially supported by an NSF grant, a FRG-NSF grant and a BSF grant
1
An important example of an affine invariant point is the centroid g. More examples will
be given throughout the paper. Let Pn be the set of affine invariant points on Kn,
Pn = {p : Kn → Rn(cid:12)(cid:12) p is affine invariant}.
Observe that Pn is an affine subspace of C(Kn, Rn), the continuous functions on Kn with
values in Rn. We denote by V Pn the subspace parallel to Pn. Thus, with the centroid
g,
Grunbaum [13] posed the problem if there is a finite basis of affine invariant points, i.e.
affine invariant points pi ∈ Pn, 1 ≤ i ≤ l, such that every p ∈ Pn can be written as
V Pn = Pn − g.
p =
l
Xi=1
αipi, with αi ∈ R and
αi = 1.
l
Xi=1
We answer this question in the negative and prove:
Theorem 1. V Pn is infinite dimensional for all n ≥ 2.
In fact, we will see that, with a suitable norm, V Pn is a Banach space. Hence, by Baire's
theorem, a basis of Pn is not even countable.
For a fixed body K ∈ Kn, we let
Pn(K) = {p(K) : p ∈ Pn}.
Then Grunbaum conjectured [13] that for every K ∈ Kn,
Pn(K) = Fn(K),
(1)
where Fn(K) = {x ∈ Rn : T x = x, for all affine T with T K = K}. We give a positive
answer to this conjecture, when Pn(K) is (n − 1)-dimensional. Note also that if K has
enough symmetries, in the sense that Fn(K) is reduced to one point xK , then Pn(K) =
{xK}.
Theorem 2. Let K ∈ Kn be such that Pn(K) is (n-1)-dimensional. Then
Pn(K) = Fn(K).
Symmetry or enough symmetries, are key in many problems. The affine invariant in-
equalities connected with the affine geometry often have ellipsoids, respectively simplices
as extremal cases. Examples are the Lp affine isoperimetric inequalities of the Lp Brunn
Minkowski theory, a theory initiated by Lutwak in the groundbreaking paper [25]. For
related results we refer to e.g. [3, 32, 33], [45]-[49], [53], [54]. The corresponding Lp affine
isoperimetric inequalities, established by Lutwak [25] for p > 1 and in [53] for all other p
- the case p = 1 being the classical affine isoperimetric inequality [1] - are stronger than
the celebrated Blaschke Santal´o inequality (see e.g., [7, 43]; and e.g., [2, 36] for recent
results): the volume product of polar reciprocal convex bodies is maximized precisely by
ellipsoids.
It is an open problem which convex bodies are minimizers for the Blaschke Santal´o
inequality. Mahler conjectured that the minimum is attained for the simplex. A ma-
jor breakthrough towards Mahler's conjecture is the inequality of Bourgain-Milman [4],
2
which has been reproved with completely different methods by Kuperberg [19] and by
Nazarov [35]. See also [11, 39, 40, 42] for related results. Even more surprising is that
it is not known whether the minimizer is a polytope. The strongest indication to date
that it is indeed the case is given in [41].
Another example is the Petty projection inequality [38], a far stronger inequality
than the classical isoperimetric inequality, and its Lp analogue, the Lp Petty projection
inequality, established by Lutwak, Yang, and Zhang [27] (see also Campi and Gronchi
[5]). These inequalities were recently strengthened and extended by Haberl and Schuster
[15]. It is precisely the ellipsoids that are maximizers in all these inequalities. On the
other hand, the reverse of the Petty projection inequality, the Zhang projection inequality
[55], has the simplices as maximizers.
Grunbaum [13] also asked , whether Pn(K) = Rn, if Fn(K) = Rn. A first step toward
solving this problem, is to clarify if there is a convex body K such that Pn(K) = Rn.
Here, we answer this question in the affirmative and prove that the set of all K such that
Pn(K) = Rn, is dense in Kn and consequently the set of all K such that Pn(K) = Fn(K),
is dense in Kn.
Theorem 3. The set of all K ∈ Kn such that Pn(K) = Rn is open and dense in
(Kn, dH ).
Here, dH is the Hausdorff metric on Kn, defined as
dH (K1, K2) = min{λ ≥ 0 : K1 ⊆ K2 + λBn
2 ; K2 ⊆ K1 + λBn
2 },
2 is the Euclidean unit ball centered at 0. More generally, Bn
where Bn
2 (a, r), is the
Euclidean ball centered at a with radius r. We shall use the following well known fact.
Let Km, K ∈ Kn. Then dH (Km, K) → 0 if and only if for some εm → 0 one has
(1 − εm)(cid:0)K − g(K)(cid:1) ⊂ Km − g(Km) ⊂ (1 + εm)(cid:0)K − g(K)(cid:1) for every m.
To establish Theorems 1 - 3, we need to introduce new examples of affine invariant
(2)
(3)
points, that have not previously been considered in the literature.
2 Affine invariant points and sets: definition and
properties.
Let K ∈ Kn. Throughout the paper, int(K) will denote the interior, and ∂K the
boundary of K. The n-dimensional volume of K is voln(K), or simply K. K◦ = {y ∈
Rn : hx, yi ≤ 1 ∀x ∈ K} is the polar body of K. More generally, for x in Rn, the polar of
K with respect to x is K x = (K − x)◦ + x.
A map p : Kn → Rn is said to be continuous if it is continuous when Kn is equipped
with the Hausdorff metric and Rn with the Euclidean norm k · k.
Grunbaum [13] gives the following definition of affine invariant points. Please note
that formally we are considering maps, not points.
Definition 1. A map p : Kn → Rn is called an affine invariant point, if p is continuous
and if for every nonsingular affine map T : Rn → Rn one has
p(T (K)) = T (p(K)).
(4)
3
Let Pn the set of affine invariant points in Rn,
and for a fixed body K ∈ Kn, Pn(K) = {p(K) : p ∈ Pn}.
Pn = {p : Kn → Rn(cid:12)(cid:12) p is affine invariant},
(5)
We say an affine invariant point p ∈ Pn proper, if for all K ∈ Kn, one has
p(K) ∈ int(K).
Examples. Well known examples (see e.g.
convex body K in Rn are
(i) the centroid
[13]) of proper affine invariant points of a
g(K) = RK xdx
K
;
(6)
(ii) the Santal´o point, the unique point s(K) for which the volume product KK x
attains its minimum;
(iii) the center j(K) of the ellipsoid of maximal volume J (K) contained in K, or John
ellipsoid of K;
(iv) the center l(K) of the ellipsoid of minimal volume L(K) containing K, or Lowner
ellipsoid of K.
Note that if T (K) = K for some affine map T : Rn → Rn and some K ∈ Kn, then
for every p ∈ Pn, one has p(K) = p(T (K)) = T (p(K)). It follows that if K is centrally
symmetric or is a simplex, then p(K) = g(K) for every p ∈ Pn, hence Pn(K) = {g(K)}.
The continuity property is an essential part of Definition 1, as, without it, pathological
affine invariant points can be constructed. The next example illustrates this.
Example 1. Let Pn be the set of all convex polytopes in Kn and define for P ∈ Pn,
p(P ) =
1
m
m
Xi=1
vi(P ),
where v1(P ), . . . vm(P ) are the vertices of P . For K ∈ Kn \ Pn, let p(K) = g(K), the
centroid of K. Then p : Kn → Rn is affine invariant, but it is not continuous at any
point.
Indeed, let K ∈ Kn. We approximate K by a polytope P , and, in turn, approximate
P by a polytope Pl by replacing one vertex v of P by sufficiently many vertices v1, . . . , vl
near v. When l → ∞, p(Pl) → v ∈ ∂P . Pl is near K, but p(K) = g(K).
Next, we introduce the notion of affine invariant set mappings, or, in short, affine
invariant sets. There, continuity of a map A : Kn → Kn is meant when Kn is equipped
on both sides with the Hausdorff metric. Our Definition 2 of affine invariant sets differs
from the one given by Grunbaum [13].
Definition 2. A map A : Kn → Kn is called an affine invariant set mapping, if A is
continuous and if for every nonsingular affine map T of Rn, one has
A(T K) = T (A(K)).
4
We then call A(K), or simply the map A, an affine invariant set mappings. We denote
by Sn the set of affine invariant set mappings,
Sn = {A : Kn → Kn(cid:12)(cid:12)A is affine invariant and continuous}.
We say that A ∈ Sn is proper, if A(K) ⊂ int(K) for every K ∈ Kn.
Known examples (see e.g., [13]) of affine invariant sets are the John ellipsoid and the
(7)
Lowner ellipsoid. Further examples will be given all along this paper.
Remarks. (i) It is easy to see that if λ ∈ R, p, q ∈ Pn and A ∈ Sn, then p ◦ A ∈ Pn
and (1 − λ)p + λq ∈ Pn. Thus, Pn is an affine space and for every K ∈ Kn, Pn(K) is
an affine subspace of Rn. Moreover, for A, B ∈ Sn, the maps
K → (A ◦ B)(K),
are affine invariant set mappings.
(1 − λ)A(K) + λB(K) and conv[A, B](K) = conv[A(K), B(K)]
(ii) Properties (4) and (2) imply in particular that for every translation by a fixed
vector x0 and for every convex body K ∈ Kn,
and
p(K + x0) = p(K) + x0, for every p ∈ Pn
A(K + x0) = A(K) + x0, for every A ∈ Sn
(8)
(9)
(iiii) Unless p = q, it is not possible to compare two different affine invariant points
p and q via an inequality of the following type
(10)
where k · k is a norm on Rn and c > 0 a constant. Indeed, by (ii), p(K − p(K)) = 0 and
q(L − q(L)) = 0. Therefore, if (10) would hold, then
kp(K) − p(L)k ≥ c kq(K) − q(L)k,
0 = kp(K − p(K)) − p(L − p(L))k ≥ c kq(K − p(K)) − q(L − p(L))k
= c kq(K) − p(K) − q(L) + p(L)k.
Choose now for L a symmetric convex body. Then kq(K)− p(K)k = 0, or p(K) = q(K).
Remark (i) provides examples of non-proper affine invariant points: once there are
two different affine invariant points, there are affine invariant points p(K) /∈ K, i.e. non-
proper affine invariant points. An explicit example is the convex body Cn constructed
in [34], for which the centroid and the Santal´o point differ.
The next results describe some properties of affine invariant points and sets.
Proposition 1. Let p, q ∈ Pn and suppose that p is proper. For K ∈ Kn, define
φq(K) = inf {t ≥ 0 : q(K) − p(K) ∈ t (K − p(K))} .
Then φq : Kn → R+ is continuous and
(i) there exist c = c(q) > 0 such that
q(K) − p(K) ∈ c (K − p(K)) for every K ∈ Kn.
(ii) If moreover q is proper, then one can chose c ∈ (0, 1) in (i).
5
Proof. Since p(K) ∈ int(K), Rn = ∪t≥0t(K − p(K)). Therefore φq is well defined. Now
we show that φq is continuous. Suppose that Km → K in (Kn, dH ). By definition, we
have
By continuity of p and q it follows that
q(Km) − p(Km) ∈ φq(Km) (Km − p(Km)) for all m.
and thus
q(K) − p(K) ∈ lim inf
m
φq(Km) (K − p(K)) ,
φq(K) ≤ lim inf
m
φq(Km).
Since p is proper, there exists d > 0, such that Bn
2 ⊆ d(K − p(K)). Since K is bounded,
there exists D > 0 such that d(K − p(K)) ⊆ DBn
2 . Let η > 0 and fix ε = η/d > 0. Since
Km → K, and by continuity of p and q, there exists m0 > 0 such that for every m ≥ m0,
K − p(K) ⊆ Km − p(Km) + εBn
2 ⊆ Km − p(Km) + η(K − p(K))
and
q(Km) − p(Km) ∈ q(K) − p(K) + εBn
2 ⊆ (φq(K) + η)(K − p(K)).
Now we observe that, if two convex bodies A and B in Rn satisfy A ⊆ B + tA for some
0 < t < 1, then A ⊆ B/(1 − t). It then follows that for every m ≥ m0,
K − p(K) ⊆
q(Km) − p(Km) ∈
Hence
and thus
1
1 − η
φq(K) + η
1 − η
(Km − p(Km)).
(Km − p(Km)) ,
φq(K) ≥ lim sup
m
φq(Km),
2 }, which is compact in Kn. ✷
and the continuity of φq is proved. Assertions (i) and (ii) follow from the continuity of
φq. Indeed, by affine invariance, we may reduce the problem to the set {K ∈ Kn : Bn
2 ⊆
K ⊆ nBn
Lemma 1. Let p, q ∈ Pn and suppose that p is proper. Then there exists a proper
r ∈ Pn such that q is an affine combination of p and r.
Proof. By the preceding proposition, there is c > 0 such that q(K)−p(K) ∈ c(K−p(K)
for all K ∈ Kn. Put r = q−p
2c + p. Then r ∈ Pn and q = 2cr + (1 − 2c)p is an affine
combination of p and r. Since p(K) ∈ int(K),
(K + p(K)) ⊆ int(K),
for all K ∈ Kn.
(cid:3)
r(K) ∈
1
2
Analogous results to Proposition 1 for affine invariant sets are also valid. We omit
their proofs.
Proposition 2. Let A ∈ Sn, p, q ∈ Pn and suppose that p is proper. Then there exists
a constant c1 > 0 such that for every K ∈ Kn,
If moreover A is proper and p = q, one can choose c1 < 1.
A(K) − q(cid:0)K(cid:1) ⊆ c1 (K − p(K)).
6
Lemma 2. Let A ∈ Sn and p ∈ Pn be proper. Then there exists t > 0 such that
K → t(A(K) − p(K)) + p(K) = tA(K) + (1 − t)p(K)
is a proper affine invariant set mapping.
The next proposition gives a reverse inclusion for affine invariants sets. We need first
another lemma, where, as in the proposition, g denotes the center of gravity.
Lemma 3. For every D, d > 0 and n ≥ 1, there exists c > 0 such that, whenever K ∈ Kn
satisfies K ⊆ DBn
Proof. Suppose that K ∈ Kn satisfies the two assumptions. Define
2 and K ≥ d, then cBn
2 ⊆ K − g(K).
cK = sup{c ≥ 0 : cBn
2 ⊆ K − g(K)}.
Then cK > 0, and there exists x ∈ ∂K such that kx − g(K)k = cK. Since K − g(K) ⊆
n(cid:0)g(K) − K(cid:1), the length of the chord of K passing through g(K) and x is not bigger
than (n + 1)cK. Let u ∈ Sn−1 be the direction of the segment [g(K), x] and let PuK be
the orthogonal projection of K onto u⊥, the subspace orthogonal to u. Then,
d ≤ K ≤ k(g + Ru) ∩ Kk PuK ≤ (n + 1)cK Dn−1Bn−1
2
.
The second inequality follows from a result by Spingarn [50]. Thus we get a strictly
positive lower bound c for cK which depends only on n, d and D. (cid:3)
Proposition 3. Let A be an affine invariant set mapping. Then there exist c > 0 such
that
c(cid:0)K − g(K)(cid:1) ⊆ A(K) − g(cid:0)A(K)(cid:1),
for every K ∈ Kn.
Proof. We first prove that there exists d > 0 such that A(K) ≥ dK for every K ∈ Kn.
By affine invariance, it is enough to prove that
inf
{K∈Kn:Bn
2 ⊂K⊆nBn
2 }
A(K)
K
> 0.
Since K → A(K)
2 } is compact in Kn,
K
this infimum is a minimum and it is strictly positive. By Proposition 2, applied with
q = g ◦ A, there exists c > 0 such that
is continuous and since {K ∈ Kn : Bn
2 ⊆ K ⊂ nBn
Therefore,
By Lemma 3, there is c0 > 0 such that
A(K) − g(cid:0)A(K)(cid:1) ⊆ c(cid:0)K − g(K)(cid:1),
A(K) − g(cid:0)A(K)(cid:1) ⊆ 2ncBn
2 ⊆ A(K) − g(cid:0)A(K)(cid:1),
c0Bn
2 ,
for every K ∈ Kn.
for every K ∈ K.
for every K ∈ K.
Now, since K − g(K) ⊆ 2nBn
for K ∈ Kn with Bn
2 ⊆ K ⊂ nBn
2 for every K ∈ Kn with Bn
2 ⊆ K ⊂ nBn
2 , and thus for all K ∈ Kn by affine invariance. (cid:3)
2 , we get the result
7
3 Several questions by Grunbaum.
We now give the proof of Theorems 1 - 3. To do so, we first need to introduce new affine
invariant points.
3.1 The convex floating body as an affine invariant set mapping.
n+1(cid:17)n
Let K ∈ Kn and 0 ≤ δ <(cid:16) n
. For u ∈ Rn and a ∈ R, H = {x ∈ Rn : hx, ui = a} is
the hyperplane orthogonal to u and H + = {x ∈ Rn : hx, ui ≥ a} and H− = {x ∈ Rn :
hx, ui ≤ a} are the two half spaces determined by H. Then the (convex) floating body
Kδ [46] of K is the intersection of all halfspaces H + whose defining hyperplanes H cut
off a set of volume at most δK from K,
Kδ =
\
{H:H−∩K≤δK}
H +.
(11)
Clearly, K0 = K and Kδ ⊆ K for all δ ≥ 0. The condition δ < (cid:16) n
g(K) ∈ int(Kδ) 6= ∅ (see [46]). Moreover, for all invertible affine maps T , one has
n+1(cid:17)n
insures that
(T (K))δ = T (Kδ) .
(12)
To prove that K → Kδ is continuous from Kn to Kn, we need some notation. For
u ∈ Sn−1, we define aδ,K(u) to be unique real number such that
Then
voln ({x ∈ K : hx, ui ≥ aδ,K(u)}) = δ voln(K).
Kδ = \u∈Sn−1 {x ∈ K : hx, ui ≤ aδ,K(u)} .
(13)
Lemma 4. Let K ∈ Kn, u ∈ Sn−1, 0 < δ < (cid:16) n
Let a ∈ R satisfy R +∞
f (t)dt = δR +∞
−∞
f (a) ≥ δ
a
n+1(cid:17)n
and f (t) = {x ∈ K : hx, ui = t}.
f (t)dt. Then one has
n−1
n max
t∈R
f (t).
Proof. By the Brunn-Minkowski theorem (see [7, 43]), f
M = f (m) = maxt∈R f (t).
1
n−1 is concave on {f > 0}. Put
We suppose first that m < a. Let g be the affine function on R such that gn−1(m) =
f (m) and gn−1(a) = f (a). As f
n−1 is concave on {f 6= 0}, one has gn−1 ≤ f on [m, a]
and gn−1 ≥ f on {f 6= 0}\ [m, a]. Thus there exists c ≤ m and d ≥ a, such that g(c) > 0,
g(d) > 0,
1
Z d
c
gn−1(t)dt =Z +∞
−∞
f (t)dt
and
Z d
a
gn−1(t)dt =Z +∞
a
f (t)dt.
8
(c) ≥ M = f (m). Moreover, by
(a) = f (a). We replace now g1 with a new function g2 that is affine
1
1
Let g1 = g1[c,d]. Since g is non increasing on [c, d], gn−1
construction, gn−1
on its support [c′, d′], c′ ≤ a ≤ d′, and satisfies gn−1
Z d′
(c′) ≥ gn−1
(t)dt =Z +∞
(a) = f (a) and clearly gn−1
One still has gn−1
computation gives
Z d′
gn−1
2
gn−1
2
f (t)dt
−∞
and
c′
a
2
2
2
(a) = f (a), gn−1
(d′) = 0,
2
(t)dt =Z +∞
a
f (t)dt.
1
(c) ≥ M . Now, an easy
f (a) = gn−1
2
(a) = δ
n−1
n , gn−1
2
(c′) ≥ δ
n−1
n M.
We suppose next that m ≥ a. The same reasoning, with 1 − δ instead of δ, gives
n−1
n M . Since 0 < δ < 1
2 , the statement follows. (cid:3)
f (a) ≥ (1 − δ)
2 . Let
2 and η > 0. There there exists ε > 0 (depending only on r, R, n, δ) such that,
Proposition 4. Let 0 < r ≤ R < ∞ and let K ∈ Kn satisfy, rBn
0 < δ < 1
whenever a convex body L satisfies dH (K, L) ≤ ε, one has for every u ∈ Sn−1
2 ⊆ K ⊆ RBn
aδ,K(u) − η ≤ aδ,L(u) ≤ aδ,K(u) + η.
Proof. Let ρ > 0. With the hypothesis on K, we may choose ε > 0 small enough such
that whenever dH (K, L) ≤ ε, then (1 − ρ)K ⊆ L ⊆ (1 + ρ)K. Fix u ∈ Sn−1 and define
fK(t) = voln−1 ({x ∈ K : hx, ui = t}) and fL(t) = voln−1 ({x ∈ L : hx, ui = t}) .
Then aδ,K := aδ,K(u) and aδ,L := aδ,L(u) satisfy
Z +∞
aδ,K
fK(t)dt = δK
and
Z +∞
aδ,L
fL(t)dt = δL.
Let θ > 0. For ρ > 0 small enough one has,
K∆L ≤ ((1 + ρ)n − (1 − ρ)n)K ≤ θ.
For such a ρ one has also
ZR fK(t) − fL(t)dt ≤ ZR
so that
voln−1 ({x ∈ K∆L : hx, ui = t}) dt = K∆L ≤ θ,
+Z +∞
aδ,L fK(t) − fL(t)dt ≤ 2θ.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Z[aδ,K ,aδ,L]
For α > 0 given, let θ = αK2 . Then
fK(t)dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Z +∞
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aδ,K
fK(t)dt −Z +∞
aδ,L
fL(t)dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
fK(t)dt − δK(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
9
Z +∞
aδ,L
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ α δK.
1
n−1
K
on
For some β ∈ [−α, α], one has hence that aδ,L = a(1+β)δ,K. Concavity of f
{fK 6= 0} implies that
Z[aδ,K ,aδ,L]
≥ aδ,K − aδ(1+β),K min(cid:0)fK(aδ,K), fK(a(1+β)δ,K)(cid:1) .
fK(t)dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
min(cid:18)fK(aδ,K), fK(a(1+α)δ(cid:19) ≥(cid:18) min(1 + β, 1)δ(cid:19)
If M = maxt∈R fK(t), we get by Lemma 4,
n−1
n
n−1
n M.
≥(cid:0)(1 − α)δ(cid:1)
Since K ⊆ RBn
2 , we estimate M from above by
M ≤ γn = Rn−1Bn−1
2
,
which is an upper bound independent of u. It follows that if α > 0 is small enough, then
(cid:3)
aδ,K − aδ,L ≤ 2 θ γn (cid:0)(1 − α)δ(cid:1)− n−1
n ≤ η.
The next proposition shows that the map K 7→ Kδ as defined in (11), is an affine
invariant set mapping.
, the mapping K 7→ Kδ is is an affine invariant set
n+1(cid:1)n
Proposition 5. For 0 < δ <(cid:0) n
mapping from Kn to Kn .
n+1(cid:1)n
Proof. We take 0 < δ < (cid:0) n
so that int(Kδ) 6= ∅ and g(K) ∈ int(Kδ) . It is clear
that K → Kδ is an affine invariant mapping and it is clear that g(K) ∈ Kδ. We now fix
a body K ∈ Kn and we verify the continuity of the mapping K → Kδ at K. We may
suppose that 0 is the center of mass of K. For some 0 < r ≤ R < ∞, one has
rBn
2 ⊆ Kδ ⊆ K ⊆ RBn
2 .
By the choice of δ, aδ,K(u) > 0 for every u ∈ Sn−1, where aδ,K(u) is as in (13). Let
η, η′ > 0 satisfy η′ ≤ ηr ≤ η minu∈Sn−1 aδ,K(u). We use the notation of the preceding
proposition to find ε > 0 such that for any L with dH (K, L) ≤ ε, one has
aδ,K(u) − η′ ≤ aδ,L(u) ≤ aδ,K(u) + η′,
(1 − η)aδ,K(u) ≤ aδ,L(u) ≤ (1 + η)aδ,K(u),
or
whence
(1 − η)Kδ ⊆ Lδ ⊆ (1 + η)Kδ.
2 , it then follows that, given ρ > 0, for η > 0 small enough, one
Since rBn
2 ⊆ Kδ ⊆ RBn
has dH (Lδ, Kδ) ≤ ρ. (cid:3)
As a corollary, we obtain new affine invariant points.
10
n+1(cid:1)n
Corollary 1. Let 0 < δ < (cid:0) n
and let p : Kn → Rn be an affine invariant point.
Then K → p(Kδ) is also an affine invariant point. In particular, for the centroid g,
K 7→ g(K \ Kδ) is an affine invariant point.
Proof. Affine invariance follows from Remark (i) after Definition 2 and continuity from
Proposition 5. The second statement follows now from the trivial identity
g(K) = Kδ
K
g(Kδ) + K \ Kδ
K
g(K \ Kδ),
which gives
g(K \ Kδ) = K
K \ Kδ
g(K) − Kδ
K \ Kδ
g(Kδ),
as an affine combination of continuous affine invariant points. (cid:3)
The next lemma is key for many of the proofs that will follow.
Lemma 5. Let m ≥ n + 1 and for 1 ≤ i ≤ m, let vi ∈ Rn be the vertices of a polytope
P in Kn. For all ε > 0 there exists z ∈ P with kv1 − zk ≤ ε and 0 < r ≤ ε such that
2 (z, r) ⊂ P and if K = conv (Bn
Bn
2 (z, r), v2, . . . , vm), then K satisfies
(i) K ⊆ P , v2, . . . , vm are extreme points of K and dH (K, P ) ≤ ε.
(ii) For sufficiently small δ, kv1 − g (K \ Kδ)k ≤ 2ε
Proof. There exists a hyperplane H that striclty separates v1 and {v2, . . . , vm}, such
that for all x ∈ H−∩ P we have that kx− v1k < ε. Let z ∈ int(H−)∩ int(P ). Then there
exists 0 < r ≤ ε such that Bn
(i) By construction of K, v2 . . . , vm are extreme points of K. Also, for all x ∈ K ∩ H−,
one has kx − v1k < ε. Therefore K ⊆ P ⊆ K + εBn
(ii) We have for δ > 0,
2 (z, r) ⊆ H− ∩ P . Let K = conv (Bn
2 and thus dH (K, P ) ≤ ε.
2 (z, r), v2, . . . , vm) .
g (K \ Kδ) = (K ∩ H +) \ Kδ
+ (K ∩ H−) \ Kδ
K \ Kδ
K \ Kδ
g(cid:0)(cid:0)K ∩ H +(cid:1) \ Kδ(cid:1)
g(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1) .
Since g ((K ∩ H−) \ Kδ) ∈ int (K) ∩ int (H−), one has
(14)
Observe that ∂K contains a cap of ∂B(z, r), so that
kv1 − g(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1)k ≤ ε.
C =Z∂K
1
κ
n+1
K dµK > 0.
By Theorem 4, one has for δ sufficiently small,
K \ Kδ = K − Kδ ≥
C
2cn
2
n+1 .
(δK)
(15)
11
Let R = max{kxk : x ∈ P}. As the Gauss curvature is equal to 0 everywhere on the
boundary ∂ (K ∩ H +), again by Theorem 4, one has for sufficiently small δ,
As
we get
K ∩ H + −(cid:12)(cid:12)(cid:12)(cid:12)
cn
2
n+1
(δK)
(K ∩ H +) δK
K∩H+ (cid:12)(cid:12)(cid:12)(cid:12)
Cε
4R
.
≤
(cid:0)K ∩ H +(cid:1) \ Kδ ⊆(cid:0)K ∩ H +(cid:1) \(cid:0)K ∩ H +(cid:1) δK
K∩H
+
2
(cid:12)(cid:12)(cid:0)K ∩ H +(cid:1) \ Kδ(cid:12)(cid:12) ≤
Cε
4Rcn
2
n+1 .
(δK)
,
(16)
(17)
It folllows from (15) and (16) that for δ small enough one has
We get thus from (14) and (17)
(K ∩ H +) \ Kδ
K \ Kδ
ε
R
.
≤
K \ Kδ
kg (K \ Kδ) − g(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1)k
= (K ∩ H +) \ Kδ
≤ (K ∩ H +) \ Kδ
K \ Kδ
(R + R) ≤ ε
≤
ε
2R
kg(cid:0)(cid:0)K ∩ H +(cid:1) \ Kδ(cid:1) − g(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1)k
(cid:0)kg(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1)k + kg(cid:0)(cid:0)K ∩ H +(cid:1) \ Kδ(cid:1)k(cid:1)
Altogether,
kv1−g (K \ Kδ)k ≤ kv1−g(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1)k+kg (K \ Kδ)−g(cid:0)(cid:0)K ∩ H−(cid:1) \ Kδ(cid:1)k ≤ 2ε.
(cid:3)
3.2 Proof of Theorem 1: Pn is infinite dimensional.
Here, we answer in the negative Grunbaum's question whether there exists a finite basis
for Pn, i.e. affine invariant points pi ∈ Pn, 1 ≤ i ≤ l, such that every p ∈ Pn can be
written as
l
l
p =
Xi=1
αipi, with αi ∈ R and
αi = 1.
Xi=1
Recall that Pn is an affine subspace of C(Kn, Rn), the continuous functions on Kn
with values in Rn and that we denote by V Pn the subspace parallel to Pn. Thus, with
the centroid g,
The dimension of Pn is the dimension of V Pn. We introduce a norm on V Pn,
V Pn = Pn − g.
kvk =
sup
2 ⊆K⊆nBn
2 kv(K)k,
K∈Kn,Bn
for v ∈ V Pn.
12
(18)
(19)
Observe that the set {K ∈ Kn : Bn
2 } is a compact subset of (Kn, dH ).
Therefore (19) is well defined and it is a norm: v = p − g 6= 0 implies that there is C
with v(C) 6= 0. By John's theorem (e.g., [?]), there is an affine, invertible map T with
Bn
2 ⊆ T (C) ⊆ nBn
2 ⊆ K ⊆ nBn
2 . Thus,
v(T (C)) = (p − g)(T (C)) = p(T (C)) − g(T (C)) = T (p(C)) − T (g(C)).
Since T = S + x0, where S is a linear map,
v(T (C)) = S(p(C) − g(C)) = S((p − g)(C)) 6= 0.
Hence
kvk ≥ kv(T (C))k > 0.
For the proof of Theorem 1 and Theorem 3, we will make use of the following theorem
by Schutt and Werner [46]. There, µK is the usual surface measure on ∂K and for
x ∈ ∂K, κ(x) is the generalized Gauss curvature at x, which is defined µK almost
everywhere.
Theorem 4. [46] Let K be a convex body in Rn. Then, if cn = 2(cid:16)Bn−1
n+1 (cid:17)
, one has
n+1
2
cn lim
δ→0
K − Kδ
(δK)
n+1
2
=Z∂K
1
n+1 (x) dµK(x).
κ
Proof of Theorem 1. We show that the closed unit ball of V Pn is not compact. For
K ∈ Kn and δ > 0, let Kδ be the convex floating body of K. Let g be the centroid and
let gδ : Kn → Rn be the affine invariant point given by
gδ(K) = g(K \ Kδ).
The set of vectors {vδ = gδ − g: δ > 0} is bounded.
gδ(K) ∈ K
Indeed, since g(K) ∈ K and
kvδk ≤
, j ≥ 1, does not have a convergent subsequence: We show that for all
(kg(K)k + kgδ(K)k) ≤ 2n.
sup
2 ⊆K⊆nBn
K∈Kn,Bn
2
2 such that
j
The sequence v 1
N there are ℓ ≥ m ≥ N and K ∈ Kn with Bn
(K) − v 1
v 1
m
ℓ
As K we choose the union of the cylinder D = [−1, 1] × Bn−1
ball,
2
where, with e1 = (1, 0, . . . , 0) ∈ Rn,
K(h) = D ∪ C(h),
and a cap of a Euclidean
(20)
C(h) =(cid:18) h2 + 2h − 1
2h
e1 +
1 + h2
2h
Bn
2(cid:19) ∩ {x = (x1, . . . , xn) ∈ Rn : x1 ≥ 1}.
2 ⊆ K ⊆ nBn
.
1
10
(K)(cid:13)(cid:13)(cid:13) ≥
(cid:13)(cid:13)(cid:13)
As h → 0, K(h) → D and, by Corollary 1, g 1
h0 > 0 such that
g(cid:16)K(h) \ (K(h)) 1
kg 1
m
m
(K(h))k =(cid:13)(cid:13)(cid:13)
m(cid:17)(cid:13)(cid:13)(cid:13) ≤
13
(K(h)) → g 1
m
(D) = 0. Thus there exists
1
10
,
for all h ≤ h0.
Now we show that we can choose ℓ sufficiently big so that
kg 1
ℓ
(K(h))k ≥
1
2
.
(21)
We apply the same argument as in the proof of Lemma 5. Let H be the hyperplane such
that
K(h) ∩ H− = C(h)
and
K(h) ∩ H + = D.
Then as in (14),
(K(h)) = (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
g 1
ℓ
Since
we get
D \ K(h) 1
K(h) \ K(h) 1
g(cid:16)D \ K(h) 1
C(h) \ K(h) 1
K(h) \ K(h) 1
ℓ(cid:17) .
g(cid:16)C(h) \ K(h) 1
ℓ(cid:17) + (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12)
g(D \ K(h) 1
ℓ
) ∈ D
and
g(C(h) \ K(h) 1
ℓ
) ∈ C(h)
)k ≤
Therefore, by triangle inequality,
kg(D \ K(h) 1
ℓ
√2
and
kg(C(h) \ K(h) 1
ℓ k ≥ 1.
kg 1
ℓ
C(h) \ K(h) 1
K(h) \ K(h) 1
ℓ(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12)
By Theorem 4, we get as in (15), for ℓ large enough, with α(h) =R∂K(h) κ
ℓK(h)(cid:1)
(K(h))k ≥ (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
ℓK(h)(cid:1)
ℓ(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12)
Z∂K(h)
D \ K(h) 1
K(h) \ K(h) 1
n+1
K(h)dµK(h) = (cid:0) 1
√2 (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
K(h) \ K(h) 1
2cn
−
.
2
2
n+1
(22)
1
n+1
K(h)dµK(h),
α(h).
Also by Theorem 4, we get as in (16),
(cid:12)(cid:12)(cid:12)
1
κ
n+1
2cn
ℓ(cid:12)(cid:12)(cid:12) ≥ (cid:0) 1
ℓ(cid:12)(cid:12)(cid:12)
ℓ(cid:12)(cid:12)(cid:12) ≤
=(cid:12)(cid:12)(cid:12)(cid:0)K(h) ∩ H +(cid:1) \ K(h) 1
D \ K(h) 1
(cid:12)(cid:12)(cid:12)
Now we finish the proof as in Lemma 5. ✷
ε
cn
α(h) (cid:18) 1
ℓK(h)(cid:19)
2
n+1
.
3.3 Proof of Theorem 2.
It was also asked by Grunbaum [13] if for every K ∈ Kn,
Pn(K) = Fn(K),
where Fn(K) = {x ∈ Rn : T x = x, for all affine T with T K = K}. Observe that it is
clear that Pn(K) ⊆ Fn(K). We will prove that Pn(K) = Fn(K), if Pn(K) is (n − 1)-
dimensional. To do so, we, again, first need to define new affine invariant set mappings.
Actually, in the proof of Theorem 2 we show that the group of isometries of K equals
{In, S} = {T : Rn → Rn affine one to one, T K = K},
where S is reflection about a hyperplane, i.e. S : Rn → Rn is bijective and there is a
hyperplane H and a direction ξ /∈ H such that S(h + tξ) = h − tξ for all h ∈ H.
14
Lemma 6. (i) Let p ∈ Pn and let g be the centroid. For 0 < ε < 1, define
Ap,ε(K) =(cid:26)x ∈ K(cid:12)(cid:12)(cid:12)(cid:12)
hx, p((K − g(K))◦)i ≥ (1 − ε) sup
y∈Khy, p((K − g(K))◦)i(cid:27) .
Then Ap,ε : Kn → Kn is an affine invariant set map.
(ii) Let p ∈ Pn and let q ∈ Pn be proper. Then Aq,p,ε : Kn → Kn given by
Aq,p,ε(K) = {x ∈ K hx, p((K − g(K))◦)i ≥ (1 − ε)hq(K), p((K − g(K))◦)i}
is an affine invariant set map.
Observe that 0 ∈ Pn((K − g(K))◦) since 0 is the Santal´o point of (K − g(K))◦.
Therefore, Pn((K − g(K))◦) is a subspace of Rn.
Proof. Let T be an invertible, affine map and T = S + a its decomposition in a linear
map S and a translation a. Then for any convex body C that contains 0 in its interior,
Moreover,
(S(C))◦ = S∗−1(C◦).
p((T (K) − g(T (K)))◦) = p((S(K − g(K)))◦)
= p(S∗−1((K − g(K))◦)) = S∗−1(p((K − g(K))◦)).
Since S∗−1∗ = S−1,
Ap,ε(T (K)) =
=(x ∈ T (K)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(x ∈ T (K)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
hx, p((T (K) − g(T (K)))◦)i ≥ (1 − ε)
hS−1x, p((K − g(K))◦)i ≥ (1 − ε)
sup
sup
y∈T (K)hy, p((T (K) − g(T (K)))◦)i)
y∈T (K)hS−1y, p((K − g(K))◦)i)
and one verifies easily that Ap,ε(T (K)) = T (Ap,ε(K)). Please note that Ap,ε(K) is
convex, compact and nonempty. ✷
Lemma 7. Let K ∈ Kn and let P : Rn → Rn be the orthogonal projection onto Pn((K−
g(K))◦). Then the restriction of P to the subspace Pn(K − g(K)) is an isomorphismn
between Pn(K − g(K)) and Pn((K − g(K))◦).
In particular,
dim(Pn(K − g(K))) = dim(Pn((K − g(K)))◦)
Proof. On the hyperplane Pn((K − g(K))◦), P (K − g(K)) has an interior point. This
holds because otherwise, by Fubini, voln(K) = 0.
Let k = dim(Pn((K−g(K)))◦). We choose u1 ∈ Pn((K−g(K))◦). Then g(Au1,ε1 ) is
a proper affine invariant point. Now we choose u2 ∈ Pn((K − g(K))◦) that is orthogonal
to P (g(Au1,ε1 )). Then P (g(Au1,ε1 )) and P (g(Au2,ε2 )) are linearly independent.
15
Eventually,
P (g(Au1,ε1)), . . . , P (g(Auk ,εk ))
are linearly independent, and therefore
g(Au1,ε1 ), . . . , g(Auk ,εk )
are linearly independent. Therefore,
dim(Pn((K − g(K)))◦) ≤ dim(Pn(K − g(K))).
Now we interchange the roles of Pn(K − g(K)) and Pn((K − g(K))◦) and get the inverse
inequality.
Let Q denote the restriction of P to the subspace Pn(K−g(K)). g(Au1,ε1), . . . , g(Auk ,εk )
is a basis of Pn(K − g(K)) and P (g(Au1,ε1 )), . . . , P (g(Auk ,εk )) is a basis of Pn((K −
g(K)))◦. Q is a bijection between the two bases, thus Q is an isomorphism. ✷
Lemma 8. Let K ∈ Kn. Then for every point x from the relative interior of K ∩ Pn(K)
there is a proper affine invariant point q with q(K) = x.
Proof. We use the same notation as in Lemma 7 and its proof. We may assume that
g(K) = 0. Suppose that there is an interior point x of Pn(K) ∩ K in the hyperplane
Pn(K) for which there is no proper affine invariant point q with q(K) = x. The set
{p(K)p is a proper affine invariant point}
is convex. P : Rn → Rn is the orthogonal projection onto Pn(K◦). Then P (Pn(K)∩ K)
is a convex set in the hyperplane Pn(K◦). Since P is an isomorphism between the
hyperplanes Pn(K ) and Pn(K◦) we have
P (x) /∈ P ({p(K)p is a proper affine invariant point}).
Moreover, P (x) is an interior point of P (Pn(K) ∩ K). By the Hahn-Banach theorem
there is u ∈ Pn(K◦) such that for all proper affine invariant points p we have
On the other hand, there is an affine invariant point q with q(K◦) = u. Then g ◦ Au,hu,xi
is a proper affine invariant point with
hu, xi ≥ hu, P (p(K))i.
which is a contradiction. ✷
hu, xi < hu, g ◦ Aq,hu,xii,
Lemma 9. Let K ∈ Kn and suppose that dim(Pn(K)) = n − 1. Then S : Rn → Rn
with
for all y ∈ Pn(K − g(K)) and x ∈ Pn((K − g(K))◦)⊥ is a linear map such that
S(y + x) = y − x
S(K − g(K)) = K − g(K).
16
Proof. By Lemma 7, the orthogonal projection onto Pn((K − g(K))◦) restricted to
Pn(K − g(K)) is an isomorphism. Therefore,
Rn = Pn(K − g(K)) ⊕ Pn((K − g(K))◦)⊥.
By Lemma 8 for every y ∈ Pn(K − g(K)) ∩ int(K) there is a proper affine invariant
point q with y = q(K). Let u1, . . . , un−1 be an orthonormal basis in Pn((K − g(K))◦).
The map Aε : Kn → Kn defined by
Aε(K) =
n−1
\i=1
{x ∈ Khq(K), uii − ε ≤ hx, uii ≤ hq(K), uii + ε}
is an affine invariant set map. As q(K) is an interior point of Aε(K), Aε(K) ∈ Kn.
Moreover,
lim
ε→0
Aε(K) = K ∩ (q(K) + Pn((K − g(K))◦)⊥)
in the Hausdorff metric. g◦ Aε is a proper affine invariant point. Since all affine invariant
points are elements of Pn(K)
lim
ε→0
(g ◦ Aε)(K) = q(K).
On the other hand, q(K) is the midpoint of K ∩ (q(K) + Pn((K − g(K))◦)⊥). ✷
Proof of Theorem 2. Theorem 2 now follows immediately from Lemma 9. Indeed,
Lemma 9 provides a map T = S − S(g(K)) + g(K) with T (K) = K and such that for
all z ∈ Pn(K) and for all x ∈ Pn((K − g(K))◦)⊥,
T (z + x) = z − x.
Consequently, if w /∈ Pn(K), then T (w) 6= w, which means that the complement of
Pn(K) is contained in the complement of Fn(K).
Remark. As a byproduct of the preceding results, it can be proved that if K ∈ Kn
satisfies Pn(K) = Rn and if Sn(K) = {A(K) : A ∈ Sn}, then Sn(K) is dense in Kn. It
might be conjectured that for general K ∈ Kn, Sn(K) is dense in {C ∈ Kn : Fn(C) ⊆
Fn(K)}.
3.4 Proof of Theorem 3.
In this subsection we show that the set of all K such that Pn(K) = Rn, is dense in Kn
and consequently the set of all K such that Pn(K) = Fn(K) is dense in Kn. A further
corollary is that, for every k ∈ N, 0 ≤ k ≤ n, there exists a convex body Qk such that
P(Qk) is a k-dimensional affine subspace of Rn.
It is relatively easy to construct examples of convex bodies K in the plane such
that Pn(K) = R2. To do so in higher dimensions is more involved and we present a
construction in the proof of Theorem 3 below. First, we will briefly mention two examples
in the plane.
Example 1. Let S be a regular simplex in the plane and let J (S) be the ellipsoid of
maximal area inscribed in S. We show in the section below that the center j(S) of J (S)
17
2 , the Euclidean ball centered
is an affine invariant point. We can assume that J (S) = B2
at 0 with radius 1. Then e.g. S = conv(cid:0)(−1,−√3), (−1,√3), (2, 0)(cid:1).
Let 0 < λ < 1 be such that H((1 + λ)e1, e1) ∩ int(S) 6= ∅ and consider the convex
body S1 = S ∩ H +((1 + λ)e1, e1) obtained from S by cutting of a cap from S. Then
still j(S1) = 0 but the center of gravity has moved to the left of 0. Next, let γ > 0 be
such that H((1 + γ)u, u)∩ int(S1) 6= ∅, where u = (−1,√3)
and consider the convex body
S2 = S1 ∩ H +((1 + γu), u) obtained from S1 by cutting of a cap from S1. Then still
j(S2) = 0 but the center of gravity g(S2) of S2 has moved and it is different from the
Santal´o point s(S2) of S2. j(S2), g(S2) and s(S2) are three affinely independent points
of R2, hence span R2.
2
Example 2. Let S be the equilateral triangle in the plane centered at 0 of Example 1
with vertices a = (2, 0), b = (−1,√3) and c = (−1,−√3). Then, as noted in Example
1, B2
2 is the John ellipse J (S) of S. Let b1, c1 be two points on the segments [a, b] and
[a, c], such that the segment [b1, c1] does not intersect B2
2 is still the John
ellipse of the quadrangle conv (b, b1, c1, c). Now the Lowner ellipse L(T ) of the triangle
T = conv (b1, b, c) is centered at 1
3 (b1 + b + c) 6= 0, if b1 6= a. L(T ) intersects the segment
2 and thus c′ → a.
[a, c] at c and at some point c′. When b1 → a, one has L(T ) → 2B2
2, and thus for some c′′ ∈ [a, c],
So we may choose b1 such that [b1, c′] does not meet B2
[b1, c1] does not meet B2
2 for any c1 ∈ [c′, c′′]. Finally, let P (c1) be the quadrangle
P (c1) = conv (b, b1, c1, c), with c1 ∈ [c′, c′′]. Since b1, b, c and c1 are the vertices of P (c1)
and c1 ∈ L(T ), L(T ) is also the Lowner ellipsoid L(P (c1))of P (c1). Altogether,
2 . Then B2
The John ellipse of P (c1) is B2
2 which is centered at 0, so that the affine invariant
point j(P (c1)) = 0.
The Lowner ellipse of P (c1) is centered at 1
3 (b1 + b + c), so that the affine invariant
point l((P (c1)) = 1
3 (b1 + b + c) 6= 0.
An easy computation shows that the centroid of P (c1) moves on an hyperbola when
c1 varies in [c′, c′′].
So, in general, these three points are not on line. (cid:3)
Proof of Theorem 3. The set of n-dimensional polytopes is dense in (Kn, dH ). Let
P be a polytope and let η > 0 be given. Then it is enough to show that there exists a
convex body Q with dH (P, Q) < η and such that Pn(Q) = Rn.
We describe the idea of the proof. For a properly constructed convex body Q we will
construct ∆i ∈ Pn, 1 ≤ i ≤ n + 1, in such a way that the ∆i(Q) are affinely independent.
The construction of such a Q is done inductively: we first construct Q1 very near P
and such that ∆1(Q1) is near an extreme point v1 of Q1. Then we construct Q2 very
near Q1 and P and such that ∆1(Q2) is near the extreme point v1 of Q2 and ∆2(Q2) is
near an extreme point v2 6= v1 of Q2.
Let P = conv (v1, . . . , vm) be a polytope with non-empty interior and with m vertices,
m ≥ n + 1. We pick n + 1 affinely independent vertices of P . We can assume that these
are v1, . . . , vn+1. Let 0 < η1 < η
n+2 be given. By Lemma 5, there exists z1 ∈ P ,
kv1 − z1k ≤ η1, and 0 < r1 ≤ η1 such that Bn
2 (z1, r1) ⊆ P and such that
Q1 = conv (Bn
2 (z1, r1), v2, . . . , vm)
18
has v2, . . . , vm as extreme points,
and for sufficiently small δ1,
dH (Q1, P ) ≤ η1,
We let ε1 < η1 and choose an ε1-net Pε1 on ∂ (Bn
2 (z1, r1)) and put
kv1 − g (Q1 \ (Q1)δ1 )k ≤ 2η1.
P1 = conv (Pε1, v2, v3, . . . , vm) .
(23)
(24)
and ε > 0, there exists γ(K, δ, ε) such that if
Then P1 ⊆ Q1 ⊆ P and dH (P1, Q1) ≤ ε1 < η1. By Corollary 1, for a given K ∈ Kn, for
n+1(cid:1)n
a given 0 < δ <(cid:0) n
As dH (P1, Q1) ≤ ε1, we get that
then kg(K \ Kδ) − g(L \ Lδ)k < ε.
(25)
dH (K, L) < γ(K, δ, ε),
for L ∈ Kn,
kg(Q1 \ (Q1)δ1 ) − g(P1 \ (P1)δ1 )k < η1,
if we choose in addition ε1 such that ε1 < γ(Q1, δ1, η1). Thus, together with (24),
kv1 − g(P1 \ (P1)δ1 )k ≤ 3 η1.
(26)
Observe that v2, . . . , vm are extreme points of P1. Now we apply Lemma 5 to P1. Let
η2 < min{ε1, γ(P1, δ1, η1)}. By Lemma 5 there exists z2 ∈ P1, kv2 − z2k ≤ η2, and
0 < r2 ≤ η2 such that Bn
2 (z2, r2) ⊂ P1 and such that
Q2 = conv (Pε1, Bn
2 (z2, r2), v3, . . . , vm)
has v3, . . . , vm as extreme points,
and for sufficiently small δ2,
dH (Q2, P1) ≤ η2,
kv2 − g (Q2 \ (Q2)δ2 )k ≤ 2η2.
(27)
(28)
As kv1 − z1k ≤ η1 and kv2 − z2k ≤ η2, we have that dH (Q2, P ) ≤ η1. Moreover, as
dH (Q2, P1) ≤ η2 < γ(P1, δ1, η1), we get by (25) with ε = η1 and by (26) that
kv1 − g(Q2 \ (Q2)δ1 )k ≤ kv1 − g(P1 \ (P1)δ1 )k+kg(P1 \ (P1)δ1 ) − g(Q2 \ (Q2)δ1 )k ≤ 4η1.
Now we let ε2 < min{η2, γ(Q2, δ1, η1)}, choose an ε2-net Pε2 on ∂ (Bn
2 (z2, r2)) and put
P2 = conv (Pε1 ,Pε2 , v3, . . . , vm) .
Then P2 ⊆ Q2 ⊆ P and dH (P2, Q2) ≤ ε2. By (25), with ε = η2, and if we choose in
addition ε2 < η(Q2, δ2, η2), we get
and thus, together with (28),
kg(Q2 \ (Q2)δ2 ) − g(P2 \ (P2)δ2 )k < η2
kv2 − g(P2 \ (P2)δ2 )k ≤ 3 η2.
(29)
19
Please note that v3, . . . , vm are extreme points of P2. Now we apply Lemma 5 to P2.
Let η3 < min{ε2, γ(P2, δ2, η2)}. By Lemma 5 there exists z3 ∈ P2, kv3 − z3k ≤ η3, and
0 < r3 ≤ η3 such that Bn
2 (z3, r3) ⊂ P2 and such that
Q3 = conv (Pε1 ,Pε1 , Bn
2 (z3, r3), v4, . . . , vm)
has v4, . . . , vm as extreme points,
and for sufficiently small δ3,
dH (Q3, P2) ≤ η3,
kv3 − g (Q3 \ (Q3)δ3 )k ≤ 2 η3.
(30)
(31)
As kv1 − z1k ≤ η1, kv2 − z2k ≤ η2 and kv3 − z3k ≤ η3 we have that dH (Q3, P ) ≤ η1.
Moreover, as dH (Q3, P2) ≤ η3 < γ(P2, δ2, η2), we get by (25) with ε = η2 and (29) that
kv2 − g(Q3 \ (Q3)δ2 )k ≤ kv2 − g(P2 \ (P2)δ2 )k+kg(P2 \ (P2)δ2 ) − g(Q3 \ (Q3)δ2 )k ≤ 4η2.
As dH (Q2, Q3) ≤ ε2 < γ(Q2, δ1, η1), it follows from (25) with ε = η1 that
kg(Q2 \ (Q2)δ1 ) − g(Q3 \ (Q3)δ1 )k ≤ η1.
By (30), it also follows from (25) with ε = η1 that
kg(P1 \ (P1)δ1 ) − g(Q2 \ (Q2)δ1 )k ≤ η1.
This, together with (26) gives
kv1 − g(Q3 \ (Q3)δ1 )k ≤ kv1 − g(P1 \ (P1)δ1 )k + kg(P1 \ (P1)δ1 ) − g(Q2 \ (Q2)δ1 )k
+ kg(Q2 \ (Q2)δ1 ) − g(Q3 \ (Q3)δ1 )k
≤ 5η1.
We continue to obtain Q = Qn+1 and affine invariant points ∆i = g(Q \ Qδi), 1 ≤ i ≤
n + 1, such that for all i,
kvi − ∆i(Q)k ≤ (n + 2)η1 < η.
As for 1 ≤ i ≤ n + 1, the vi are affinely independant, so are the ∆i.
It remains to show that On = {K ∈ Kn : Pn(K) = Rn} is open in (Kn, dH ). Observe
that K ∈ On if and only if for some p1 . . . , pn+1 ∈ Pn (depending on K),
voln(conv(cid:0)p1(K) . . . , pn+1(K)(cid:1) > 0.
(cid:3)
Since L → vol(conv(cid:0)p1(L) . . . , pn+1(L)(cid:1) is continuous on Kn, it follows that On is open.
Corollary 2. For every k ∈ N, 0 ≤ k ≤ n, there exists a convex body Qk such that
P(Qk) is a k-dimensional affine subspace of Rn.
Proof. For k = 0, we take a centrally symmetric body. For k = n, we take the body
Q of Theorem 3. For 1 ≤ k ≤ n − 1, we take the intermediate bodies Qk constructed in
the proof of Theorem 3. (cid:3)
20
References
[1] W. Blaschke, Vorlesungen uber Differentialgeometrie II, Affine Differentialgeometrie.
Springer Verlag, Berlin, (1923).
[2] K.J. Boroczky, Stability of the Blaschke-Santal´o and the affine isoperimetric inequality
Adv. in Math. 225 (2010), 1914 -- 1928.
[3] K.J. Boroczky, E. Lutwak, D. Yang, and G. Zhang, The logarithmic Minkowski
problem, to appear in Journal of AMS.
[4] J. Bourgain and V. D. Milman, New volume ratio properties for convex symmetric bodies
in Rn , Invent. Math. 88 (1987), 319 -- 340.
[5] S. Campi and P. Gronchi, The Lp-Busemann-Petty centroid inequality, Adv. in Math.
167 (2002), 128 -- 141.
[6] R. J. Gardner, A positive answer to the Busemann-Petty problem in three dimensions,
Ann. of Math. (2) 140 (1994), 435-47.
[7] R.J. Gardner Geometric tomography. Second edition. Encyclopedia of Mathematics and
its Applications, 58. Cambridge University Press, Cambridge (2006).
[8] R. J. Gardner, The dual Brunn-Minkowski theory for bounded Borel sets: Dual affine
quermassintegrals and inequalities, Adv. Math. 216 (2007), 358-386.
[9] R. J. Gardner, A. Koldobsky, and T. Schlumprecht, An analytical solution to the
Busemann-Petty problem on sections of convex bodies, Ann. of Math. (2) 149 (1999),
691-703.
[10] R. J. Gardner and G. Zhang, Affine inequalities and radial mean bodies. Amer. J. Math.
120, no.3, (1998), 505-528.
[11] Y. Gordon, M. Meyer and S. Reisner, Zonoids with minimal volume product -- a new
proof, Proc. Amer. Math. Soc. 104 (1988), 273 -- 276.
[12] E. Grinberg and G. Zhang, Convolutions, transforms, and convex bodies, Proc. London
Math. Soc. (3) 78 (1999), 77 -- 115.
[13] B. Grunbaum, Measures of symmetry for convex sets, Proc. Sympos. Pure Math. 7, (1963),
233 -- 270.
[14] C. Haberl, Blaschke valuations, Amer. J. Math., 133, (2011), 717 -- 751.
[15] C. Haberl and F. Schuster, General Lp affine isoperimetric inequalities. J. Differential
Geometry 83 (2009), 1-26.
[16] C. Haberl, E. Lutwak, D. Yang and G. Zhang, The even Orlicz Minkowski problem,
Adv. Math. 224 (2010), 2485-2510.
[17] D. Klain, Star valuations and dual mixed volumes, Adv. Math. 121 (1996), 80-101.
[18] D. Klain, Invariant valuations on star-shaped sets, Adv. Math. 125 (1997), 95-113.
[19] G. Kuperberg, From the Mahler Conjecture to Gauss Linking Integrals, Geometric And
Functional Analysis 18 (2008), 870 -- 892.
[20] M. Ludwig, Ellipsoids and matrix valued valuations, Duke Math. J. 119 (2003), 159-188.
[21] M. Ludwig, Minkowski areas and valuations, J. Differential Geometry, 86, (2010), 133 -- 162.
21
[22] M. Ludwig and M. Reitzner, A Characterization of Affine Surface Area, Adv. in Math.
147 (1999), 138-172.
[23] M. Ludwig and M. Reitzner, A classification of SL(n) invariant valuations. Annals of
Math. 172 (2010), 1223-1271.
[24] E. Lutwak, The Brunn-Minkowski-Firey theory I : Mixed volumes and the Minkowski
problem, J. Differential Geom. 38 (1993), 131 -- 150.
[25] E. Lutwak, The Brunn-Minkowski-Firey theory II : Affine and geominimal surface areas,
Adv. Math. 118 (1996), 244-294.
[26] E. Lutwak and G. Zhang, Blaschke-Santal´o inequalities, J. Differential Geom. 47, (1997),
1 -- 16.
[27] E. Lutwak, D. Yang and G. Zhang, Lp affine isoperimetric inequalities, J. Differential
Geom. 56 (2000), 111 -- 132.
[28] E. Lutwak, D. Yang and G. Zhang, A new ellipsoid associated with convex bodies, Duke
Math. J. 104 (2000), 375 -- 390.
[29] E. Lutwak, D. Yang and G. Zhang, Sharp Affine Lp Sobolev inequalities, J. Differential
Geom. 62 (2002), 17 -- 38.
[30] E. Lutwak, D. Yang and G. Zhang, The Cramer -- Rao inequality for star bodies, Duke
Math. J. 112 (2002), 59-81.
[31] E. Lutwak, D. Yang and G. Zhang, Volume inequalities for subspaces of Lp, J. Differ-
ential Geom. 68 (2004), 159 -- 184.
[32] M. Meyer and E. Werner, The Santalo-regions of a convex body, Transactions of the
AMS 350 (1998), 4569 -- 4591.
[33] M. Meyer and E. Werner, On the p-affine surface area. Adv. Math. 152 (2000), 288 --
313.
[34] M. Meyer, C. Schutt and E. Werner, New affine measures of symmetry for convex
bodies, Adv. Math. 228, (2011), 2920 -- 2942.
[35] F. Nazarov, The Hormander proof of the Bourgain-Milman theorem, preprint, 2009.
[36] F. Nazarov, F. Petrov, D. Ryabogin and A. Zvavitch, A remark on the Mahler
conjecture: local minimality of the unit cube, Duke Mathematical J. 154, (2010), 419 -- 430.
[37] G. Paouris, Concentration of mass on convex bodies, Geometric and Functional Analysis
16, (2006), 1021 -- 1049.
[38] C. M. Petty, Isoperimetric problems, Proc. Conf. Convexty and Combinatorial Geometry
Univ. Oklahoma 1971, University of Oklahoma, (1972), 26 -- 41.
[39] S. Reisner, Zonoids with minimal volume-product, Math. Z. 192 (1986), 339 -- 346.
[40] S. Reisner, Minimal volume product in Banach spaces with a 1-unconditional basis, J.
London Math. Soc.36 (1987), 126 -- 136.
[41] S. Reisner, C. Schutt and E. Werner, Mahler's conjecture and curvature International
Mathematics Research Notices, DOI 10.1093/imrn/rnr003 (2011)
[42] J. Saint-Raymond, Sur le volume des corps convexes sym´etriques, S´eminaire d'Initiation
`a l'Analyse, 1980-1981, Universit´e PARIS VI, Paris 1981.
22
[43] R. Schneider, Convex Bodies: The Brunn-Minkowski Theory, Encyclopedia of Mathe-
matics and its Applications 44, Cambridge University Press, Cambridge (1993).
[44] F. Schuster, Crofton measures and Minkowski valuations, Duke Math. J. 154 (2010),
1 -- 30.
[45] F. Schuster and M. Weberndorfer, Volume Inequalities for Asymmetric Wulff Shapes,
J. Differential Geom., in press.
[46] C. Schutt and E. Werner, The convex floating body, Math. Scand. 66, (1990), 275 -- 290.
[47] C. Schutt and E. Werner The convex floating body of almost polygonal bodies, Geom.
Dedic. 44, (1992), 169 -- 188.
[48] C. Schutt and E. Werner, Polytopes with Vertices Chosen Randomly from the Boundary
of a Convex Body, Geom. Aspects of Funct. Analysis, Lecture Notes in Math. 1807, (2003),
241 -- 422.
[49] C. Schutt and E. Werner, Surface bodies and p-affine surface area. Adv. Math. 187
(2004), 98-145.
[50] J. E. Spingarn, An Inequality for Sections and Projections of a Convex Set, Proc. Amer.
Math. Soc., 118, No. 4, (1993), 1219 -- 1224.
[51] A. Stancu, The Discrete Planar L0-Minkowski Problem. Adv. Math. 167 (2002), 160-174.
[52] A. Stancu, On the number of solutions to the discrete two-dimensional L0-Minkowski
problem. Adv. Math. 180 (2003), 290-323.
[53] E. Werner and D. Ye, New Lp affine isoperimetric inequalities, Adv. Math. 218 (2008),
no. 3, 762-780.
[54] E. Werner and D. Ye, Inequalities for mixed p-affine surface area, Math. Ann. 347
(2010), 703-737.
[55] G. Zhang, Restricted chord projection and affine inequalities, Geom. Dedicata, 39 (1991),
213 -- 222.
[56] G. Zhang, Intersection bodies and Busemann-Petty inequalities in R4, Annals of Math.
140 (1994), 331-346.
[57] G. Zhang, A positive answer to the Busemann-Petty problem in four dimensions, Annals
of. Math. 149 (1999), 535-543.
[58] G. Zhang, New Affine Isoperimetric Inequalities, ICCM 2007, Vol. II, 239-267.
Mathieu Meyer
Universit´e de Paris Est - Marne-la-Vall´ee
Equipe d'Analyse et de Math´ematiques Appliqu´ees
Cit´e Descartes - 5, bd Descartes
Champs-sur-Marne 77454 Marne-la-Vall´ee, France
[email protected]
Carsten Schutt
Christian Albrechts Universitat
Mathematisches Seminar
23
24098 Kiel, Germany
[email protected]
Elisabeth Werner
Department of Mathematics
Case Western Reserve University
Cleveland, Ohio 44106, U. S. A.
[email protected]
Universit´e de Lille 1
UFR de Math´ematiques
59655 Villeneuve d'Ascq, France
24
|
1411.7292 | 3 | 1411 | 2016-02-27T09:52:54 | A convenient notion of compact set for generalized functions | [
"math.FA"
] | We introduce the notion of functionally compact sets into the theory of nonlinear generalized functions in the sense of Colombeau. The motivation behind our construction is to transfer, as far as possible, properties enjoyed by standard smooth functions on compact sets into the framework of generalized functions. Based on this concept, we introduce spaces of compactly supported generalized smooth functions that are close analogues to the test function spaces of distribution theory. We then develop the topological and functional analytic foundations of these spaces. | math.FA | math |
A CONVENIENT NOTION OF COMPACT SET FOR
GENERALIZED FUNCTIONS
PAOLO GIORDANO AND MICHAEL KUNZINGER
Abstract. We introduce the notion of functionally compact sets into the
theory of nonlinear generalized functions in the sense of Colombeau. The
motivation behind our construction is to transfer, as far as possible, properties
enjoyed by standard smooth functions on compact sets into the framework of
generalized functions. Based on this concept, we introduce spaces of compactly
supported generalized smooth functions that are close analogues to the test
function spaces of distribution theory. We then develop the topological and
functional analytic foundations of these spaces.
1. Introduction
A main advantage of nonlinear generalized functions in the sense of Colombeau
as compared to Schwartz distributions is the fact that they can be viewed as set-
theoretic maps on domains consisting of generalized points. This change of perspec-
tive allows to develop several branches of the theory in close analogy to classical
analysis, and thereby has become increasingly important in recent years (cf., e.g.,
[26, 4, 2, 11, 12, 27, 3, 16, 6, 18]). In particular, appropriate topologies on spaces
of nonlinear generalized functions, the so called sharp topologies (see below for the
definition), have been introduced in [28, 29] and have since been studied by many
authors. Apart from their central position in the structure theory of Colombeau
algebras, they also supply the foundation for applications in the theory of nonlinear
partial differential equations (e.g., for a suitable concept of well-posedness).
From the point of view of analysis, a key notion underlying many existence re-
sults is that of compactness. It turns out, however, that sharply compact subsets
of generalized points display certain unwanted properties: e.g., no infinite subset
of Rn is sharply compact since the trace of the sharp topology on subsets of Rn is
2010 Mathematics Subject Classification. 46F30,46A13,13J99.
Key words and phrases. Functionally compact sets, Colombeau generalized functions, generalized
smooth functions, locally convex modules.
P. Giordano has been supported by grant P25116 and P25311 of the Austrian Science Fund FWF.
M. Kunzinger has been supported by grants P23714 and P25326 of the Austrian Science Fund
FWF.
1
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
2
discrete. In fact, this is a necessary consequence of the set eR of generalized num-
bers containing actual infinitesimals, hence seems unavoidable also in alternative
approaches, cf. [17, Prop. 2.1] and [18, Thm. 25].
The importance of a convenient notion of compactness for nonlinear general-
ized functions has been recognized by several authors, most recently in [1]. The
approach we take in the present paper is to introduce an appropriate concept of
compactly supported generalized function, and then to study spaces consisting of
such functions, which, in analogy to the test function space D(Ω) in distribution
theory, we denote by GD(U ). The domain U here is a set of generalized points.
Based on Garetto's theory of locally convex eC-modules [11, 12, 14] we then de-
velop the topological and functional analytic foundations of these spaces. We find
that they are indeed close analogues of the classical spaces of test functions in that
they are countable strict inductive limits of complete metric spaces GDK(U ) (ana-
logues of DK(Ω) in distribution theory) satisfying properties paralleling those of
the classical strict (LF)-spaces D(Ω).
The plan of the paper is as follows: In the remainder of this introduction we
fix some basic notions used throughout this work. Section 2 introduces what we
call functionally compact sets, based on work by Oberguggenberger and Vernaeve
in [27]. Building on this, in Section 3 we define compactly supported generalized
smooth functions (GSF), as well as the corresponding spaces GD(U ) and GDK(U ).
We also show that every Colombeau generalized function f ∈ Gs(Ω) (in particular,
every Schwartz distribution) defines a compactly supported GSF ¯f : eR −→ eR that
coincides with f on eΩc. In order to obtain appropriate topologies on these spaces,
the basic properties of classical norms, yet take values in eR, thereby generalizing
we define so-called generalized norms in Section 4. These are maps that share
a standard alternative description of the sharp topology on generalized numbers
(cf. [4, 17]). In Sections 5 and 6 these generalized norms are employed to endow
the spaces GDK(U ) with metric topologies. In particular, in Section 5.1 we study
connections between non-Archimedean properties and Hausdorff topological vector
spaces of generalized functions, proving an impossibility theorem: there does not
exist a Hausdorff topological vector subspace of the Colombeau special algebra
which contains the Dirac delta and even a single trace of an open set of the sharp
topology. The completeness of the spaces GDK(U ) is established in Section 7. In
the final Section 8 we derive the fundamental functional analytic properties of the
space GD(U ).
1.1. Basic notions. Our main references for Colombeau's theory are [8, 9, 25, 22].
The special Colombeau algebra Gs(Ω) over an open subset Ω of Rn is defined as the
quotient E s
M (Ω)/N s(Ω), where (setting I := (0, 1] and noting that in the naturals
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
3
N = {0, 1, 2, 3 . . .} we include zero.)
M (Ω) := {(uε) ∈ C∞(Ω)I ∀K ⋐ Ω∀α ∈ Nn ∃N ∈ N : sup
x∈K ∂αuε(x) = O(ε−N )}
E s
N s(Ω) := {(uε) ∈ C∞(Ω)I ∀K ⋐ Ω∀α ∈ Nn ∀m ∈ N : sup
x∈K ∂αuε(x) = O(εm)}.
M (Ω) are called moderate, those of N s(Ω) are called negligible. Nets
Elements of E s
in E s
M (Ω) are written as (uε), and u = [uε] denotes the corresponding equivalence
class in Gs(Ω). For (uε) ∈ N s(Ω) we also write (uε) ∼ 0. We will abbreviate
'Colombeau generalized function' by CGF. Gs(−) is a fine sheaf of differential al-
gebras and there exist sheaf embeddings (based on smoothing via convolution) of
the space of Schwartz distributions D′ into Gs (cf. [22]).
Given Ω ⊆ Rn open, the space of generalized points in Ω is eΩ = ΩM / ∼, where
ΩM = {(xε) ∈ ΩI ∃N ∈ N : xε = O(ε−N )} is called the set of moderate nets
and (xε) ∼ (yε) if xε − yε = O(εm) for every m ∈ N. In the particular case Ω = R
we obtain the ring of Colombeau generalized numbers (CGN) eR = RM / ∼ (and
analogously for eC), which can also be written as eR = RM /N s, where N s is the
set of all negligible nets of real numbers (xε) ∈ RI , i.e. such that (xε) ∼ 0. eR
is an ordered ring with respect to its natural order relation: x ≤ y iff there are
representatives (xε) and (yε) such that xε ≤ yε for ε sufficiently small. We point
out that, in the present work, the notion x > y does not mean x ≥ y and x 6= y.
Rather, it is to be understood as x − y ≥ 0 and x − y invertible. By [22, 1.2.38]
and [24, Prop. 3.2] we have:
Lemma 1. Let x ∈ eR. Then the following are equivalent:
(i)
(ii) For each representative (xε) of x there exists some ε0 and some m such that
x > 0.
xε > εm for all ε < ε0.
(iii) For each representative (xε) of x there exists some ε0 such that xε > 0 for
all ε < ε0.
We shall use the notation dεm := [εm] ∈ eR for any m ∈ R. Hence x > 0 is
equivalent to x ≥ dεm for some m > 0. If P(ε) is a property of ε ∈ I, we will also
sometimes use the notation ∀0ε : P(ε) to denote ∃ε0 ∈ I ∀ε ∈ (0, ε0] : P(ε).
The space of compactly supported generalized points eΩc is defined by Ωc/ ∼,
where Ωc := {(xε) ∈ ΩI ∃K ⋐ Ω∀0ε : xε ∈ K} and ∼ is the same equivalence
relation as in the case of eΩ.
Concerning intervals, we use the following notations: [a, b] := {x ∈ eR a ≤ x ≤
b}, [a, b]R := [a, b] ∩ R. Also, for x, y ∈ eRn we write x ≈ y if x − y is infinitesimal,
i.e. if x − y ≤ r for all r ∈ R>0.
As already indicated above, the natural topology for Colombeau-type spaces is
the so-called sharp topology ([7, 29, 28, 4, 5, 23, 16]). This topology is generated
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
4
([2, 3, 17]). For Euclidean balls, we will write B E
by balls Bρ(x) = {y ∈ eRn y − x < ρ}, where − is the natural extension of
the Euclidean norm to eRn, [xε] := [xε] ∈ eR, and ρ ∈ eR>0 is strictly positive
ρ (x) = {y ∈ Rn y − x < ρ}.
On the other hand, the so-called Fermat-topology on eRn (see [17, 18]) is generated
by the balls Br(x) for x ∈ eRn and r ∈ R>0. Originally, the sharp topology was
introduced using an ultrametric as follows: The map
v : RM −→ (−∞,∞]
v((uε)) := sup{b ∈ R uε = O(εb)}.
gives a pseudovaluation on eR. Then setting − e : eR → [0,∞), ue := exp(−v(u))
provides a translation-invariant complete ultrametric
ds : eR ×eR −→ R+
ds(u, v) := u − ve
on eR, which induces the sharp topology on eR.
Garetto in [11, 12] extended the above construction to arbitrary locally convex
spaces by functorially assigning a space of CGF GE to any given locally convex space
E. In this approach, the seminorms of E are used to define pseudovaluations which
induce a generalized locally convex topology on the eC-module GE, again called
sharp topology. In the present paper, we will exclusively work with eR-modules. We
note, however, that all our constructions trivially carry over to the eC-case.
For any S ⊆ I, eS denotes the equivalence class in eR of the characteristic function
of S (cf. [4, 30]). Any eS is an idempotent, and eS + eSc = 1. Also, eS 6= 0 if and
only if 0 ∈ S. For any subset A of eRn, its interleaving (cf. [27]) is defined as
eSj aj m ∈ N,{S1, . . . , Sm} a partition of I, aj ∈ A
interl(A) :=
If (Aε) is a net of subsets of Rn then the internal set ([27, 31]) generated by (Aε)
mXj=1
.
is
and the strongly internal set ([18]) generated by (Aε) is
[Aε] =n[xε] ∈ eRn xε ∈ Aε for ε smallo ,
hAεi :=n[xε] ∈ eRn xε ∈ε Aεo .
Here, xε ∈ε Aε means that xε ∈ Aε for ε small and that the same property holds
for any representative of [xε]. The net (Aε) is called sharply bounded if there
exists some N ∈ R>0 such that for ε sufficiently small we have supx∈Aε x ≤ ε−N .
Equivalently, we have that (Aε) is sharply bounded if there exists ρ ∈ eR>0 such
that [Aε] ⊆ Bρ(0).
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
5
f : X −→ Y is a generalized smooth function (GSF)
Finally, given X ⊆ eRn and Y ⊆ eRd, then (see [18])
if there exists a net uε ∈ C∞(Ωε, Rd) defining f in the sense that X ⊆ hΩεi,
f ([xε]) = [uε(xε)] ∈ Y and (∂αuε(xε)) ∈ Rd
M for all x = [xε] ∈ X and all α ∈ Nn.
The space of GSF from X to Y is denoted by GC∞(X, Y ) (in contrast to [18],
where the notation eG(X, Y ) was used). GSF are a natural generalization of CGF
to general domains. In particular, for any Ω ⊆ Rn open, GC∞(eΩc) ≃ Gs(Ω). GSF
on subsets of eRn, endowed with the sharp topology, form a sub-category of the
category of topological spaces. In particular, they can be composed unrestrictedly.
2. A new notion of compact subset for nonlinear generalized
functions
Even though the intervals [a, b] ⊆ eR, a, b ∈ R, are neither compact in the sharp
nor in the Fermat topology (see [18, Thm. 25]), analogously to the case of smooth
functions, a GSF satisfies an extreme value theorem on such sets. In fact, we have:
Proposition 2. Let f ∈ GC∞(X,eR) be a generalized smooth function defined on
the subset X of eRn. Let ∅ 6= K = [Kε] ⊆ X be an internal set generated by a
sharply bounded net (Kε) of compact sets Kε ⋐ Rn , then
∃m, M ∈ K ∀x ∈ K : f (m) ≤ f (x) ≤ f (M ).
(2.1)
Proof. By [18, Lem. 28], f can be represented by a net uε ∈ C∞(Rn, Rd). Since K 6=
∅, for ε sufficiently small, say for ε ∈ (0, ε0], Kε is non-empty and, by assumption,
it is also compact. For all ε ∈ (0, ε0] we have
∃mε, Mε ∈ Kε ∀x ∈ Kε : uε(mε) ≤ uε(x) ≤ uε(Mε).
Since the net (Kε) is sharply bounded, both the nets (mε) and (Mε) are moderate.
Therefore m = [mε], M = [Mε] ∈ K ⊆ X. Take any x ∈ [Kε], then there exists a
representative (xε) such that xε ∈ Kε for ε small. Therefore f (m) = [uε(mε)] ≤
[uε(xε)] = f (x) ≤ f (M ).
(cid:3)
We shall use the assumptions on K and (Kε) given in this theorem to introduce
a new notion of "compact subset" which behaves better than the usual classical
notion of compactness in the sharp topology.
Definition 3. A subset K of eRn is called functionally compact, denoted by K ⋐f
eRn, if there exists a net (Kε) such that
(i) K = [Kε] ⊆ eRn
(ii)
(iii) ∀ε ∈ I : Kε ⋐ Rn
(Kε) is sharply bounded
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
6
If, in addition, K ⊆ U ⊆ eRn then we write K ⋐f U . Finally, we write [Kε] ⋐f U if
(ii), (iii) and [Kε] ⊆ U hold.
We note that in (iii) it suffices to ask that Kε be closed since it is bounded by
(ii), at least for ε small. In fact, we have:
Lemma 4. A subset K of eRn is functionally compact if and only if it is internal
and sharply bounded.
Proof. By [27, Lemma 2.4 and Cor. 2.2], every sharply bounded internal set K has
a sharply bounded representative (Kε) consisting of closed (hence compact) subsets
of Rn.
(cid:3)
We motivate the name functionally compact subset by anticipating that on this
type of subsets, GSF have properties very similar to those that ordinary smooth
functions have on standard compact sets.
Remark 5.
(i)
By [27, Prop. 2.3], any internal set K = [Kε] is closed in the sharp topology.
it is not closed.
(ii)
compact. In particular, [0, 1] = ^[0, 1]R = [[0, 1]R] is functionally compact.
In particular, the open interval (0, 1) ⊆ eR is not functionally compact since
If H ⋐ Rn is a non-empty ordinary compact set, then eH = [H] is functionally
(iii) The empty set ∅ =e∅ ⋐f eR.
(iv) By Lemma 4, eRn is not functionally compact since it is not sharply bounded.
(v) The set of compactly supported points eRc is not functionally compact because
the GSF f (x) = x does not satisfy the conclusion (2.1) of Prop. 2.
We start the study of functionally compact sets by proving suitable generaliza-
tions of theorems from classical analysis.
Theorem 6. Let K ⊆ X ⊆ eRn, f ∈ GC∞(X,eRd). Then K ⋐f eRn implies f (K) ⋐f
eRd.
Proof. Let (Kε) be as in Def. 3 and let the GSF f be defined by the net uε ∈
C∞(Rn, Rd). Let us first prove that f (K) = [uε(Kε)]. In fact, y ∈ f (K) = f ([Kε])
is equivalent to
∃(xε) ∈ Rn
M ∀0ε : xε ∈ Kε and y = [uε(xε)].
(2.2)
This necessary entails y ∈ [uε(Kε)]. Vice versa, if y ∈ [uε(Kε)], then there exists
(yε) ∈ Rd
M such that yε ∈ uε(Kε) for ε small. Hence, for each of these ε there also
exists xε ∈ Kε such that yε = uε(xε), which implies y = [uε(xε)], i.e. (2.2) since
(Kε) is sharply bounded. Clearly, uε(Kε) ⋐ Rd, so it remains to prove that the
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
7
net (uε(Kε)) is sharply bounded. If ∀ε0 ∃ε ≤ ε0 : Kε = ∅, then [Kε] = K = ∅, so
f (K) = ∅ and the conclusion is trivial. Otherwise, assume that Kε 6= ∅ for ε ≤ ε0
and proceed by contradiction assuming that
∀k ∈ N∃(εkn)n ↓ 0 ∀n∃ykn ∈ uεkn (Kεkn ) :
ykn > ε−k
kn .
(2.3)
We can write ykn = uεkn (xkn) for some xkn ∈ Kεkn . Next, set ε0 := ε00 and for
k , εk−1(cid:1) and set εk := εknk . Take any ¯xε ∈ Kε
k > 0 pick nk such that εknk < min(cid:0) 1
for each ε ≤ ε0, and set xε := xknk if ε = εk and xε := ¯xε otherwise. Then xε ∈ Kε
for ε ≤ ε0, so x = [xε] ∈ K ⊆ X and (uε(xε)) ∈ Rd
M by the definition of GSF,
which contradicts (2.3).
(cid:3)
As a corollary of this theorem and Rem. (5).(ii) we get
a = [ε−N ], b = [ε−M ] with M > N .
Corollary 7. If a, b ∈ eR and a ≤ b, then [a, b] ⋐f eR.
Let us note that a, b ∈ eR can also be infinite, e.g. a = [−ε−N ], b = [ε−M ] or
Lemma 8. Let K, H ⋐f eRn, then we have:
(i) K ∪ H ⊆ interl(K ∪ H) ⋐f eRn
If K ∪ H is internal, then it is functionally compact
If K ∩ H is internal, then it is functionally compact.
(ii)
(iii)
Proof. (i) follows from [27, Prop. 2.8] which implies K ∪ H ⊆ interl(K ∪ H) =
[Kε ∪ Hε], where the nets (Kε) and (Hε) satisfy Def. 3. Property (ii) follows from
[27, Lemma 2.7] which implies that if the union of internal sets is internal, then it
is equal to its interleaving. Property (iii) is a consequence of Lemma 4.
(cid:3)
Lemma 4 is the following:
If H ⊆ K ⋐f eRn, then also H is sharply bounded. So, another consequence of
Corollary 9. Let H ⊆ K ⋐f eRn, then H internal implies H ⋐f eRn.
Finally, in the following result we consider the product of functionally compact sets:
Proposition 10. Let K ⋐f eRn and H ⋐f eRd, then K × H ⋐f eRn+d. In particular,
if ai ≤ bi for i = 1, . . . , n, then Qn
Proof. From [27, Prop. 2.13] if follows that K × H is internal, in fact for K = [Kε],
H = [Hε], K × H = [Kε × Hε]. From this representation it immediately follows
that H × K is sharply bounded as well, so Lemma 4 gives the claim.
i=1[ai, bi] ⋐f eRn.
(cid:3)
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
8
3. Compactly supported generalized smooth functions
Our main goal in this section is to define and study an analogue within GSF
of the space DK(Ω) of smooth functions supported in a fixed compact set K ⋐ Ω.
In order to define this space, we first try to define the concept of support of a
GSF. Clearly, if ϕ ∈ D[−a,a]R(R), one would expect that ϕ should have compact
support also if we think of ϕ as a GSF. In fact, supp(ϕ) should be contained in
[−a, a] ⋐f eR. Already this basic requirement implies that if f ∈ GC∞(X, Y ), the
natural definition
S(f ) := X \[nBρ(x) ∩ X x ∈ X, ρ ∈ eR>0, fBρ(x)∩X = 0o
doesn't fit with our intuition.
Indeed, if we take the aforementioned ϕ so that
ϕ(0) = 1, and S ⊆ (0, 1] such that 0 ∈ ¯S and 0 ∈ Sc, then ϕ(eS dε−1) = eSc 6= 0
and eS dε−1 ∈ S(ϕ) \ [−a, a]. This motivates the following
Definition 11. Let X ⊆ eRn, Y ⊆ eRd and f ∈ GC∞(X, Y ), then
where here (−) denotes the relative closure in X with respect to the sharp topology.
Using this concept, we have supp(ϕ) ⊆ [−a, a]. In fact, ϕ(x) = [ϕ(xε)] > 0
supp(f ) := {x ∈ X f (x) > 0},
implies ϕ(xε) > εq for some q ∈ R>0 and for ε small, and hence xε ∈ [−a, a]R.
Remark 12.
(i)
(ii)
In the setting of Colombeau algebras, one usually defines the support of some
f ∈ Gs(Ω) as a subset of Ω ⊆ Rn, i.e., as a set of classical points, namely as
suppGs(f ) := Ω \S(cid:8)Br(x) ∩ Ω x ∈ Ω, r ∈ R>0, fBr (x)∩Ω = 0(cid:9), where the
last equality has to be understood in Gs(Br(x) ∩ Ω). Using X = eΩc as the
natural domain of any f ∈ Gs(Ω) (cf. [18, Thm. 37]), it is then immediate
that supp(f ) ∩ Ω ⊆ suppGs(f ).
Let u ∈ D′(Ω) be a Schwartz distribution and denote by ι : D′(Ω) →
GC∞(eΩc, R) a standard embedding via convolution. Then supp(ι(u)) ∩ Ω ⊆
(iii) Assume that the embedding ι : D′(Ω) → GC∞(eΩc, R) has been defined by
using a mollifier ρ ∈ S(Rn) which is identically equal to 1 in the ball B E
p (0),
p ∈ R>0. Then δ(x) = dε−n for each x ∈ Bp·dε(0) and hence Bp·dε(0) ⊆
supp(ι(δ)), whereas supp(ι(δ)) ∩ Rn = {0}.
In general, supp(f ) is not an internal set because it is not generally closed
supp(u), as follows from (i) and the fact that ι is a sheaf-morphism.
(iv)
by finite interleaving (see [27, Lem. 2.7]). Consider e.g. X with only near
standard points and f ∈ GC∞(X,eR) which is strictly positive on two disjoint
intervals. However, if X itself is closed under finite interleaving then so is
supp(f ).
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
9
If (uε) defines f ∈ GC∞(X, Y ), the internal set [supp(uε)] is not intrinsically
defined since it depends on the defining net (uε). Consider, e.g., uε(x) := ϕ(x) +
ε1/ε > 0 where ϕ ∈ C∞(R, R
In our further analysis we will repeatedly make use of the following notion:
≥0).
Definition 13. For A ⊆ eRn we call the set
the strong exterior of A.
ext(A) := {x ∈ eRn ∀a ∈ A :
x − a > 0}
This set can also be described in the following way:
Lemma 14. If A ⊆ eRn, then ext(A) = {x ∈ eRn ∀S ⊆ I : eS 6= 0 ⇒ xeS 6∈ AeS}.
Proof. ⊆: Let x = [xε] ∈ ext(A) and suppose that there exists some S ⊆ I with
eS 6= 0 and some a = [aε] ∈ A such that xeS = aeS. Then there exists some
q > 0 such that xε − aε > εq for ε small. However, xeS = aeS implies that
xε − aε = O(εq+1) for ε → 0, ε ∈ S, a contradiction.
⊇: If there exists some a = [aε] ∈ A such that x − a 6> 0 then there is a
k for all k. Letting S := {εk k ∈ N} implies
sequence εk ↓ 0 with xεk − aεk < εk
xeS = aeS.
(cid:3)
For non-trivial internal sets we have the following characterization of the strong
exterior:
Lemma 15. Let ∅ 6= [Kε] = K ⊆ eRn. Then
ext(K) = hK c
εi.
Proof. Suppose first that x ∈ hK c
εi and let S ⊆ I with eS 6= 0. Suppose that there
existed some a ∈ K with xeS = aeS. Since a ∈ K there exists a representative
(aε) of a with aε ∈ Kε for all ε. Then xeS = aeS implies that there exists a
representative (xε) of x and a sequence εk ↓ 0 in S with xεk = aεk ∈ Kεk for all k.
But this contradicts the fact that x ∈ hK c
εi.
Conversely, if x 6∈ hK c
εi then there exists a representative (xε) of x and a sequence
εm ↓ 0 with xεm ∈ Kεm for all m. Since K 6= ∅, there exists some w = [wε] ∈ K
with wε ∈ Kε for all ε. Now let
aε :=
and set S := {εm m ∈ N}. Then a = [aε] ∈ [Kε] and xeS = aeS by construction.
Thus xeS ∈ KeS, and so x 6∈ ext(K).
if ε = εm
otherwise
xεm
wε
(cid:3)
As an immediate conclusion we obtain:
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
10
Corollary 16. Let K = [Kε] = [Lε] 6= ∅. Then hK c
εi = hLc
εi.
The next result relates the support of a GSF to the exterior of certain internal
sets. To formulate it concisely, we introduce the following notations: Denote by Kf
the set of all internal ∅ 6= K ⊆ eRn with ext(K) 6= ∅ and such that there exists a
net uε ∈ C∞(Rn, Rd) that defines f and such that [uε(xε)] = 0 for all [xε] ∈ ext(K).
Also, denote by Hf the set of all the internal sets of the form K = [supp(uε)] ⊆ eRn
for some net uε ∈ C∞(Rn, Rd) that defines f and such that both K and ext(K) are
non empty.
Then we have:
Lemma 17. Let X ⊆ eRn, Y ⊆ eRd and f ∈ GC∞(X, Y ). Then
supp(f ) ⊆ X ∩ \K∈Kf
K ⊆ X ∩ \K∈Hf
Proof. Since X ∩TK∈Kf
first inclusion in (3.1), it suffices to prove that
K.
(3.1)
K is a sharply closed subset of X, in order to show the
{x ∈ X f (x) > 0} ⊆ X ∩ \K∈Kf
K.
Let x ∈ X be such that f (x) > 0, so that
∃r ∈ R>0 :
f (x) > dεr.
(3.2)
Let K = [Kε] ∈ Kf , and assume, by contradiction, that x = [xε] /∈ K. We first
prove that
(xεk ) ⊆ K c
εk ,
where r comes from (3.2). In fact, suppose to the contrary that
∃q ∈ N>r ∃(εk)k∈N ↓ 0 ∀k ∈ N : B E
εq
k
(3.3)
∀q ∈ N>r ∃εq ∀ε ≤ εq ∃y(q)
ε ∈ B E
εq (xε) : y(q)
We may assume that (εq)q∈N ↓ 0. Setting yε := y(q)
x = [yε] ∈ K, which contradicts x /∈ K.
ε
By assumption ∃z = [zε] ∈ ext(K), and hence
ε ∈ Kε.
for ε ∈ (εq+1, εq], we have
∃s ∈ N>q ∀0ε : d(zε, Kε) > εs,
where q comes from (3.3). Using (εk)k∈N from (3.3), we set xε := xεk if ε = εk
and xε := zε otherwise. Then x := [xε] ∈ ext(K). But K ∈ Kf , so there exists
a net uε ∈ C∞(Rn, Rd) that defines f and such that [uε(xε)] = 0.
In particu-
lar, uεk (xεk ) = uεk (xεk ) = O(ε2s
k ) as k → +∞, which contradicts f (x) =
[uε(xε)] > dεr > dεs.
Turning now to the second inclusion, assume that x ∈ X\TK∈Hf
exists a net (uε) that defines f , with ∅ 6= K := [supp(uε)] ⊆ eRn and ext(K) 6= ∅,
K, so that there
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
11
but such that x /∈ K. We want to show that x /∈ TK ′∈Kf
K′. But for all [yε] ∈
ext(K) = hsupp(uε)ci, we have yε /∈ supp(uε) for ε small. Hence, uε(yε) = 0 for
these ε, and this yields [uε(yε)] = 0. This proves that K ∈ Kf , but x /∈ K.
(cid:3)
Remark 18.
(i)
Using methods from Nonstandard Analysis, one can prove the converse of the
K.
first inclusion in (3.1) in the following case: If X = eRn and the sharp interior
of nx ∈ eRn f (x) = 0o is non-empty, then
supp(f ) = \K∈Kf
K and choose some z in the interior of ny ∈ eRn
To see this, let x ∈ TK∈Kf
f (y) = 0} and some q > 0 such that Bdεq (z) ⊆ny ∈ eRn f (y) = 0o. Given
a representative (vε) of f , let (χε) be a moderate net of smooth functions such
that χε vanishes on B E
εq+1 (zε).
Then setting uε := χεvε gives a new representative of f such that [{y ∈ Rn
uε(y) = 0}] has non-empty sharp interior. Let mε ∈ N, mε → ∞ as ε → 0.
Then set Kε := {y ∈ Rn uε(y) ≥ εmε} and K := [Kε]. It follows that
∅ 6= ext(K) ⊆ {y ∈ Rn f (y) = 0}, so x ∈ K. Fix any k ∈ N. Now using
nonstandard notation, letting ρ := [ε], u := [uε], and fixing a representative of
x, which we temporarily simply denote by x, the above in particular implies
εq+2 (zε) and is identically equal to 1 on Rn \ B E
∗N
∞ ⊆ {m ∈ ∗N ∃yk ∈ ∗Rn : yk − x ≤ ρk ∧ u(yk) ≥ ρm}.
By the underspill principle, therefore, there also exists some m ∈ N and some
yk ∈ ∗Rn such that yk−x ≤ ρk and u(yk) ≥ ρm. Taking equivalence classes
of these nets, it follows that there exist yk ∈ eRn such that yk − x ≤ dεk
and f (yk) > 0. Since yk → x in the sharp topology, this implies that
x ∈ supp(f ).
(ii) The assumption ext(K) 6= ∅ in the definition of Kf is essential. To illustrate
this, take ϕ as defined before Def. 11, pick any sequence εk ↓ 0 and set
Kε := {0} for ε = εk, and Kε = Rn otherwise. Then ext(K) = ∅, and
obviously supp(ϕ) 6⊆ K.
(iii) The question of whether the reverse of the last inclusion in (3.1) also holds
remains open.
In the following section we shall see that even though the notion of support intro-
duced above may not be entirely satisfactory, there nevertheless is a very convenient
notion of being compactly supported for GSF.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
12
3.1. The spaces GDK(U, Y ) and GD(U, Y ). A frequently used idea to solve prob-
lems like the previous ones comes by considering a family of GSF having "good rep-
resentatives", i.e. possessing a defining net (uε) that conforms to our intuition and
includes the examples we have in mind. In the following, we denote by (u1, . . . , ud)
the components of a function u which takes values, e.g., in Rd.
Definition 19. Let ∅ 6= K ⋐f U ⊆ eRn and Y ⊆ eRd, then we say that f is a GSF
compactly supported in K, and we write
f ∈ GDK(U, Y )
if f ∈ GC∞(U, Y ) and there exists a net (uε) such that:
(uε) defines f , where uε ∈ C∞(Rn, Rd) for all ε.
(i)
∀α ∈ Nn ∀[xε] ∈ ext(K) :
(ii)
[∂αuε(xε)] = 0.
Moreover, we set
GD(U, Y ) := [∅6=K⋐fU
GDK(U, Y ).
We will simply use the symbols GDK(U ) and GD(U ) if Y = eR.
Remark 20.
(i)
(ii)
Lemma 17 implies that if f ∈ GDK(U, Y ), then supp(f ) ⊆ K because K ∈
Kf . The converse implication for an arbitrary subset U remains an open
problem. For the case U = eRn, and if supp(f ) is not empty, see Thm. 27
It is clear that, in general, another net defining f : U −→ eRd will not nec-
essarily satisfy (ii) of Def. 19 because such a net is not bound to have any
below.
particular behavior outside of U .
(iii) Set Kε := B E
1 (0) and Lε := Kε \ B E
εi in (ii) instead of the simpler [K c
ε -mesh of points for Kε = B E
e−1/ε(0), then for the Hausdorff distance
of Kε and Lε we obtain dH(Kε, Lε) = e− 1
ε . By [27, Cor. 2.10], it fol-
lows that [Kε] = [Lε]. If we consider a net of smooth functions such that
1/2(0) = 1, uεR2\Kε = 0, uε ≥ 0, then supy∈R2\Kε ∂αuε(y) = 0 but
uεBE
supy∈R2\Lε ∂αuε(y) = 1. This motivates the use of the strongly internal set
hK c
ε ]. Analogously, we can consider as Lε
an e− 1
1 (0). Lε may also contain points in K c
ε,
but so that d(x, Kε) = e− 1
ε . This example shows clearly that we need to
be sufficiently "far" from ∂Kε to be sure that [∂αuε(xε)] = 0, i.e. at points
x = [xε] such that [d(xε, Kε)] > 0, as stated in (ii). Since ext(K) is indepen-
dent of the choice of representative of K by Cor. 16, so is Def. 19. This is
essential to prove the completeness of the spaces GDK(U ) and GD(U ).
The following result will turn out to be useful when proving results by contradic-
tion in several instances below. It permits to restrict the analysis to only two cases:
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
13
points in K or in ext(K). To state it more clearly, we say that a generalized point
[yε] joins points of the sequence (yk)k∈N at (εk)k∈N if ∀N ∈ N∃k ≥ N : yεk = yk.
Lemma 21. Let K = [Kε] ⋐f eRn, (εk)k∈N a sequence in (0, 1] which strictly
decreases to 0, and (yk)k∈N a sequence in Rn. For each k ∈ N, let xk ∈ Kεk be such
that d(yk, xk) = d(yk, Kεk ). Then either
(i)
or
(i)
∃[yε] ∈ ext(K) :
[yε] joins points of the sequence (yk)k∈N at (εk)k∈N
∃[¯yε], [¯xε] joining points of the sequences (yk)k∈N and (xk)k∈N (respectively)
at (εk)k∈N such that [¯yε] = [¯xε] and ¯xε ∈ Kε ∀ε.
Proof. We can always pick a point eε ∈ K c
ε so that e := [eε] ∈ ext(K); in fact, since
(Kε) is sharply bounded, we can find eε ∈ Rn \ Kε so that d(eε, Kε) > 1 and (eε)
is moderate. We can also take a point iε ∈ Kε for each ε because, without loss of
generality, we can assume that Kε 6= ∅ for all ε.
The first alternative in the statement is realized if
∃b ∈ R>0 ∃N ∈ N∀k ≥ N :
yk − xk > εb
k.
(3.4)
Set yε := yk if ε = εk and yε := eε otherwise. Then, if ε = εk, by (3.4) we have
d(yε, Kε) = yk − xk > εb; otherwise d(yε, Kε) = d(eε, Kε) > 1 ≥ εb. Therefore,
[yε] ∈ hK c
εi = ext(K) and, of course, [yε] joins points of the sequence (yk)k∈N at
(εk)k∈N.
Vice versa, if
∀h ∈ N∃kh > h :
ykh − xkh ≤ εh
kh,
(3.5)
then we can set ¯yε := ykh, ¯xε := xkh if ε = εkh and ¯yε := ¯xε := iε otherwise.
Then, if ε = εkh , by (3.5) we have ¯yε − ¯xε = ykh − xkh ≤ εh
= εh; otherwise
¯yε − ¯xε = iε − iε = 0, and (i) follows.
(cid:3)
kh
We use the above result to guarantee that the maximum values of any partial
derivative ∂αf are attained on K and not outside, as precisely stated in the following
Lemma 22. Let (uε) and K = [Kε] satisfy Def. 19, then
Proof. By contradiction, assume that
∀α ∈ Nn ∀i = 0, . . . , n : (cid:20) sup
y∈Rn ∂αui
ε(y)(cid:21) 6=(cid:20) sup
∃a ∈ R>0 ∃εk ց 0 ∀k ∈ N : (cid:12)(cid:12)(cid:12)(cid:12) sup
(cid:20) sup
y∈Rn ∂αui
x∈Kε ∂αui
y∈Rn vεk (y) − sup
For simplicity of notation, set vε := ∂αui
ε. Inequality (3.7) means
(3.6)
(3.7)
ε(y)(cid:21) =(cid:20) sup
ε(x)(cid:21) .
x∈Kε ∂αui
ε(x)(cid:21) .
x∈Kε vεk (x)(cid:12)(cid:12)(cid:12)(cid:12) > εa
k.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
14
Thus
∀k ∈ N∃yk ∈ Rn : εa
k + sup
x∈Kεk
vεk (x) < vεk (yk) .
(3.8)
We can hence apply Lemma 21. In the first case (i) we have y := [yε] ∈ ext(K) so
that [vε(yε)] = 0 by Def. 19 (ii). Since [yε] joins points of (yk)k∈N at (εk)k∈N, from
(3.8) and [vε(yε)] = 0 we get
εa
k ≤ εa
k + sup
x∈Kεk
vεk (x) < vεk (yk) = vεk (yεk ) < εa+1
k
,
for k sufficiently big, which gives a contradiction in the first case. In the second
case, i.e., (i) of Lemma 21, we have ¯y := [¯yε] = ¯x := [¯xε] ∈ K ⊆ U . Therefore
[vε(¯yε)] = [vε(¯xε)] by the definition of GSF. Since [¯yε] and [¯xε] join points of (yk)k∈N
and (xk)k∈N, respectively, at (εk)k∈N, for k sufficiently big we have
εa
k + sup
x∈Kεk
vεk (x) < vεk (yk) ≤ vεk (yk) − vεk (xk) + vεk (xk)
≤ εa+1
k + sup
x∈Kεk
vεk (x) ,
leading to a contradiction also in the second case.
(cid:3)
The previous result will be essential to prove that any compactly supported GSF
can be extended to the whole of eRn, and to define an eR-valued norm of f that does
not depend on K.
Using Lemma 22, we can prove that any derivative of a compactly supported GSF
is globally bounded in an appropriate sense:
Lemma 23. Let the net (uε) satisfies Def. 19, then
β ≤ α ⇒(cid:20) sup
∀α ∈ Nn ∃C ∈ eR∀β ∈ Nn :
y∈Rn ∂βuε(y)(cid:21) ≤ C.
Proof. By the extreme value property Prop. 2 we can set Ciβ :=(cid:2)supy∈Kε ∂βui
where β ∈ Nn, β ≤ α, and i = 1, . . . , n. Set C := 1 + √n maxβ≤α
C > Ciβ and property (3.6) yields(cid:2)supy∈Rn ∂βui
ε(y)(cid:3) =(cid:2)supx∈Kε ∂βui
and hence (cid:2)supy∈Rn ∂βuε(y)(cid:3) < C.
entire eRn:
Theorem 24. Let ∅ 6= K ⋐f U ⊆ eRn and f ∈ GDK(U,eRd) be defined by (uε)
which satisfies Def. 19. Then (uε) defines a GSF of the type eRn −→ eRd and there
exists one and only one ¯f ∈ GDK(eRn,eRd) such that:
Moreover, compactly supported GSF can be extended in a unique way to the
Ciβ . Then
ε(y)(cid:3),
ε(x)(cid:3) = Ciβ
(cid:3)
1≤i≤n
(i)
(ii)
¯fK = fK
¯fext(K) = 0.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
15
Moreover, this ¯f satisfies
(iii)
(i)
If U is sharply open and α ∈ Nn, then ∂α ¯fU = ∂αf
¯feRn
can be identified with a Colombeau generalized function.
c
Proof. The existence part follows by showing that for all α ∈ Nn and all [xε] ∈ eRn
M . This follows since Lemma 23 yields that for any α ∈ Nn
we have (∂αuε(xε)) ∈ Rd
there exists some (Cαε) ∈ Rn
M such that supx∈Rn ∂αuε(x) ≤ Cαε for all ε small. To
prove uniqueness, let g ∈ GDK(eRn,eRd) be such that gK = ¯fK = f and gext(K) =
¯fext(K) = 0. By contradiction, assume that ¯f (y) = [uε(yε)] 6= g(y) = [vε(yε)], for
some y = [yε] ∈ eRn, where (vε) defines g. Thus
∃a ∈ R>0 ∃εk ց 0 ∀k ∈ N :
uεk (yεk ) − vεk (yεk ) > εa
k.
(3.9)
By Lemma 21, this leaves two possibilities. In the first one, there exists a point
z = [zε] ∈ ext(K) which joins points of the sequence (yεk )k∈N at (εk)k∈N. Therefore
g(z) = [vε(zε)] = ¯f (z) = [uε(zε)] = 0, which gives a contradiction at ε = εk when
compared with (3.9). In the second one, there exists a point ¯z = [¯zε] ∈ K joining
points of the sequence (yεk )k∈N at (εk)k∈N. Once again, we have g(¯z) = [vε(¯zε)] =
¯f (¯z) = [uε(¯zε)] = f (z), in contradiction to (3.9).
Furthermore, [18, Thm. 31] implies claim (iii). Finally, (i) follows from the
(cid:3)
We also have this simple but useful result:
Thm. 24 opens the possibility to restrict our attention to compactly supported
isomorphism GC∞(eRc,eRd) ≃ Gs(R)d.
Lemma 25. Let ∅ 6= K ⋐f U ⊆ eRn, U be a sharply open set, f ∈ GDK(U,eRd) and
α ∈ Nn. Then ∂αf ∈ GDK(U,eRd).
GSF whose domain is the entire eRn, as stated in the following
Theorem 26. For ∅ 6= K ⋐f eRn and Y ⊆ eRd set
GDg(K, Y ) :=nf ∈ GC∞(eRn, Y ) fext(K) = 0o ,
where the g superscript means globally defined. Let K ⊆ U ⊆ eRn, then
(3.10)
(i)
(ii)
If f ∈ GDg(K, Y ), then fU ∈ GDK(U, Y ).
If f ∈ GDK(U,eRd), then ∃! ¯f ∈ GDg(K,eRd) : ¯fK = fK.
(iii) GDg(K, Y ) = GDK(eRn, Y ).
Proof. (i): Let (uε) be any net that defines f , so that uε ∈ C∞(Rn, Rd). Clearly
fU ∈ GC∞(U, Y ), so for all α ∈ Nn and x = [xε] ∈ ext(K) it remains to show that
[∂αuε(xε)] = 0. By Lem. 15, we get ext(K) = hK c
εi, which is a sharply open set.
So x ∈ ext(K) yields Br(x) ⊆ ext(K) for some r ∈ eR>0, and hence fBr(x) = 0.
Thereby ∂αf (x) = [∂αuε(xε)] = 0.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
16
(ii): This is exactly Thm. 24.
(iii): This follows directly from (i) by setting U = eRn, and by Def. 19 which
yields fext(K) = 0.
For the extension of property (ii) to arbitrary codomains Y ⊆ eRd (provided that
U is strongly internal) see Thm. 62 below.
(cid:3)
Note explicitly that in (3.10), instead of the more technical properties of Def. 19,
we have a concise and simpler pointwise condition. Notwithstanding this and several
other positive aspects of definition (3.10) (see the second part of Lem. 17 and the
following Thm. 27), in the present work, we prefer not to change Def. 19 in favor
of (3.10): On the one hand, Def. 19 is nearer to the classical definition of DK(Ω),
where the domain is Ω ⊆ Rn; on the other hand, for applications of these notions
to geometry, locally defined functions are a more natural setting. We can also
summarize these results by saying that compactly supported GSF have a good
notion of being compactly supported, and globally defined compactly supported
GSF are in addition well-behaved with respect to the concept of support introduced
above. This is also clearly confirmed by the following
Theorem 27. Let Y ⊆ eRd, and f ∈ GC∞(eRn, Y ) such that supp(f ) 6= ∅. Then
f ∈ GDK(eRn, Y ) if and only if supp(f ) ⊆ K.
Proof. One implication is immediate from Remark 20.(i). Conversely, if supp(f ) ⊆
K, by Thm. 26.(iii) we have to show that fext(K) = 0. Let x ∈ ext(K) but assume
that f (x) 6= 0. Pick any y ∈ ny′ ∈ eRn f (y′) > 0o, which is non-empty because
supp(f ) 6= ∅. Let (uε) be a net that defines f and let y = [yε]. Since f (x) 6= 0, there
exists S ⊆ I such that eS 6= 0 and f (x)eS > 0. Then setting z := xeS + yeSc, we
have f (z) > 0 and hence z ∈ supp(f ) ⊆ K. But then xeS = zeS ∈ KeS which, by
Lem. 14, yields x /∈ ext(K), a contradiction.
3.2. Examples of compactly supported GSF. A first class of examples comes
by considering each ϕ ∈ DK(Ω), K ⋐ Ω ⊆ Rn. Indeed, it suffices to set Kε := K
and uε(x) := ϕ(x) if x ∈ Ω and uε(x) := 0 otherwise to have that ϕ ∈ GD eK(eΩc,eR),
where we recall that eK = [K]. Therefore DK (Ω) ⊆ GD eK(eΩc,eR).
Moreover, since any given CGF can be defined by a net (uε) of maps with sharply
(cid:3)
bounded compact supports, we have the following result:
Theorem 28. Let Ω be an open subset of Rn and J = [Jε] ∈ eR be a CGN such
that limε→0+ Jε = +∞. Set Kε := {x ∈ Ω x ≤ Jε} and K := [Kε]. Then for all
f ∈ GC∞(eΩc,eRd) there exists ¯f ∈ GDK(eRn,eRd) such that ¯f eΩc
Proof. Set Uε := {x ∈ Ω x < 1
2 Jε} so that Uε ⊆ Kε for ε small. Let χε ∈
C∞(Ω, R) be such that χUε = 1 and supp(χε) ⊆ Kε. Let f ∈ GC∞(eΩc,eRd) be
= f .
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
17
represented by (vε), with vε ∈ C∞(Rn, Rd), and set uε := χε · vε. Then each uε is
compactly supported and any x = [xε] ∈ eΩc satisfies xε ∈ Uε for ε small because
limε→0+ Jε = +∞. Therefore ¯f := [uε(−)] ∈ GDK(eRn,eRd), and if xε ∈ Uε then
uε(xε) = vε(xε), so ¯f eΩc
= f .
(cid:3)
This theorem gives an infinity of non-trivial examples of compactly supported
GSF. Moreover, even though ¯f depends on the fixed infinite number J ∈ eR, every
such ¯f contains all the information of the original CGF f because ¯f eΩc
Finally, the constant function f (x) = 1 for all x ∈ eR is not compactly supported.
holds. Then choosing r large enough that dε−r ∈ eR \ K we arrive at f ( dε−r) =
In fact, by contradiction, assume that f admits (uε) and (Kε) such that Def. 19
[uε(ε−r)] = 0.
= f .
4. Generalized norms on GDK and GD
(i)
As a first step to topologizing the spaces GDK and GD we prove:
Theorem 29. Let ∅ 6= K ⋐f U ⊆ eRn, then
(ii) For all non-empty H ⋐f U , the inclusion K ⊆ H implies GDK(U,eRd) ⊆
GDK(U,eRd) is an eR-module
GDH (U,eRd).
(ii): Take f ∈ GDK(U,eRd). Since K ⊆ H, by [27, Prop. 2.8] for each repre-
sentative (Kε) of K we get the existence of a representative (Hε) of H such that
Kε ⊆ Hε for all ε. Therefore, ext(K) = hK c
εi = ext(H) and hence the
conclusion follows.
Proof. (i) is immediate from Def. 19.
εi ⊇ hH c
(cid:3)
From the extreme value property, Prop. 2, it is natural to expect that the fol-
lowing CGN could serve as generalized eR-valued norms.
Definition 30. Let ∅ 6= K ⋐f U ⊆ eRn, where U is a sharply open set. Let m ∈ N
and f ∈ GDK(U,eRd). Then
kfkm,K := max
α≤m
1≤i≤d
max(cid:0)(cid:12)(cid:12)∂αf i(Mαi)(cid:12)(cid:12) ,(cid:12)(cid:12)∂αf i(mαi)(cid:12)(cid:12)(cid:1) ∈ eR,
where mαi, Mαi ∈ K satisfy
∀x ∈ K : ∂αf i(mαi) ≤ ∂αf i(x) ≤ ∂αf i(Mαi).
The following result permits to calculate the (generalized) norm kfkm,K using
any net (vε) that defines f . In case the net (vε) satisfies Def. 19, it also permits
to prove that this norm does not depend on K, as is the case for any ordinary
compactly supported smooth function.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
18
Then we have:
Even though kfkm,K ∈ eR, using an innocuous abuse of language, in the following
we will simply call kfkm,K a norm.
Proposition 31. Under the assumptions of Def. 30, let the set K = [Kε] ⋐f eRn.
ε(x)(cid:12)(cid:12)(cid:21)
If the net (vε) defines f , then kfkm,K =(cid:20)maxα≤m
.
x∈Rn(cid:12)(cid:12)∂αui
ε(x)(cid:12)(cid:12)
If (uε) defines f and (uε) satisfies Def. 19, then
supx∈Kε(cid:12)(cid:12)∂αvi
kfkm,K =
max
α≤m
1≤i≤d
1≤i≤d
(4.1)
sup
(ii)
(i)
Proof. In proving (i) we will also prove that the norm kfkm,K is well-defined, i.e.
it does not depend on the particular choice of points mαi, Mαi as in Def. 30. As in
the proof of Prop. 2, we get the existence of ¯mαiε, ¯Mαiε ∈ Kε such that
Hence (cid:12)(cid:12)∂αvi
ε( ¯mαiε) ≤ ∂αvi
ε( ¯mαiε)(cid:12)(cid:12) ,(cid:12)(cid:12)∂αvi
max(cid:0)(cid:12)(cid:12)∂αvi
ε(x) ≤ ∂αvi
ε( ¯Mαiε)(cid:12)(cid:12)(cid:1). Thus
ε( ¯mαiε)(cid:12)(cid:12) ,(cid:12)(cid:12)∂αvi
ε( ¯Mαiε).
sup
max
α≤m
1≤i≤d
∀x ∈ Kε : ∂αvi
ε(x)(cid:12)(cid:12) ≤ max(cid:0)(cid:12)(cid:12)∂αvi
x∈Kε(cid:12)(cid:12)∂αvi
But ¯mαiε, ¯Mαiε ∈ Kε, so
x∈Kε(cid:12)(cid:12)∂αvi
ε(x)(cid:12)(cid:12)
ε(x)(cid:12)(cid:12) ≤ max
=
max
max
α≤m
1≤i≤d
α≤m
1≤i≤d
sup
ε( ¯Mαiε)(cid:12)(cid:12)(cid:1) .
ε( ¯Mαiε)(cid:12)(cid:12)(cid:1)
max(cid:0)(cid:12)(cid:12)∂αvi
=
max(cid:0)(cid:12)(cid:12)∂αf i( ¯Mαi)(cid:12)(cid:12) ,(cid:12)(cid:12)∂αf i( ¯mαi)(cid:12)(cid:12)(cid:1) .
ε( ¯mαiε)(cid:12)(cid:12) ,(cid:12)(cid:12)∂αvi
α≤m
1≤i≤d
= max
α≤m
1≤i≤d
(ii): By Lemma 22, we have that
From this, both the fact that the norm kfkm,K is well-defined and claim (i) follow.
max
x∈Kε(cid:12)(cid:12)∂αui
ε(x)(cid:12)(cid:12)
= max
α≤m
1≤i≤d
α≤m
1≤i≤d
sup
(cid:20) sup
x∈Kε(cid:12)(cid:12)∂αui
(cid:20) sup
x∈Rn(cid:12)(cid:12)∂αui
ε(x)(cid:12)(cid:12)(cid:21) =
ε(x)(cid:12)(cid:12)(cid:21) =
max
α≤m
1≤i≤d
= max
α≤m
1≤i≤d
sup
ε(x)(cid:12)(cid:12)
x∈Rn(cid:12)(cid:12)∂αui
.
Corollary 32. Let ∅ 6= K ⋐f U ⊆ eRn, where U is a sharply open set. Let ∅ 6= H ⋐f
U and m ∈ N. If f ∈ GDK(U,eRd)∩GDH (U,eRd), then kfkm,K = kfkm,H =: kfkm.
Proof. The right hand side of (4.1) does not depend on K.
(cid:3)
Another consequence of Prop. 31 is the following:
(cid:3)
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
19
Corollary 33. Let U ⊆ eRn, f ∈ GD(U,eRd) and ¯f ∈ GD(eRn,eRd) be the extension
of f defined in Thm. 24. Then for all m ∈ N, kfkm = k ¯fkm.
Our use of the term "norm" is justified by the following
(i)
(ii)
Proposition 34. Let ∅ 6= K ⋐f U ⊆ eRn, where U is a sharply open set. Let f ,
g ∈ GDK(U,eRd) and m ∈ N. Then
kfkm ≥ 0
kfkm = 0 if and only if f = 0
(iii) ∀c ∈ eR : kc · fkm = c · kfkm
kf + gkm ≤ kfkm + kgkm.
kf · gkm ≤ cm · kfkm · kgkm for some cm ∈ R>0.
(iv)
(v)
Proof. (i), (iii) and (iv) follow directly from Prop. 31, as does (v), using the Leibniz
rule. The 'if'-part of property (ii) follows from (4.1).
(cid:3)
U :
ε ), d(kε, U c
(4.2)
(cid:3)
∀K, H ⋐f U :
interl(H ∪ K) ⊆ U.
We now prove that also the space GD(U,eRd) is an eR-module, at least for certain
Proposition 35. Let U ⊆ eRn be a non empty sharply open set. Assume that
Then GD(U,eRd) is an eR-module.
Proof. Since in Thm. 29 we already proved that GDK(U,eRd) is closed with respect
to products by scalars, we only need to prove that GD(U,eRd) is closed with respect
to sum. Let f ∈ GDK(U,eRd), g ∈ GDH (U,eRd) and let (uε), (vε) satisfy Def. 19 for
f and g, respectively. Lemma 8 and [27, Prop. 2.8] imply that interl(H ∪ K) =
[Hε ∪ Kε] ⋐f eRn. But ext(H ∪ K) = hH c
ε ∩ K c
εi = ext(H) ∩
ext(K). Therefore, ∂α(uε + vε) is zero on ext(H ∪ K). By our assumption (4.2)
∅ 6= interl(H ∪ K) ⋐f U , so that f + g ∈ GD(U,eRd).
Proposition 36. Let U ⊆ eRn be a non-empty sharply open set. If U is eR-convex
Proof. Assume that U is eR-convex, i.e. xh + (1 − x)k ∈ U for all h, k ∈ U and all
x ∈ [0, 1]. Then for all H, K ⊆ U (even if we do not assume them to be functionally
compact), and all y ∈ interl(H ∪ K), we can write y = eS · h + eSc · k for some S ⊆ I
and h ∈ H, k ∈ K. Thus y = eS · h + (1− eS)· k ∈ U since eS ∈ [0, 1] and h, k ∈ U .
Now, assume that U is a strongly internal set, i.e. for some net (Uε) of subsets
of Rn, we have U = hUεi. We continue to use the notations for y as above. Since
ε ) > εq for some q ∈ R>0 and
h, k ∈ U , [18, Thm. 8] entails that d(hε, U c
In the following result, we give two general sufficient conditions for (4.2) to hold.
or U is a strongly internal set, then (4.2) holds.
εi = hH c
εi ∩ hK c
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
20
1
n ,
ε small, where h = [hε], k = [kε]. But y = eS · h + eSc · k, so for all ε small, if
ε ∈ S then yε = hε and if ε /∈ S then yε = kε. In any case, d(yε, U c
ε ) > εq, hence
y ∈ hUεi = U .
(cid:3)
Example 37. If U = (−1, 1) ∪ (2, 4) ⊆ eR, then U is a sharply open set, but it
2(cid:3)R(cid:3)
does not satisfy condition (4.2) of Prop. (35): let H :=(cid:2)(cid:2)− 1
and xε := 0 if ε ∈ h 1
n+1(cid:17) if n ∈ N>0 is even, and xε := 3 otherwise. Then
(R) ⊆ GDH (eRc),
x := [xε] ∈ interl(H ∪ K) but x /∈ U . Moreover, let ϕ ∈ D[− 1
2 , 1
(R) ⊆ GDK(eRc) be positive non-trivial functions. Then, as we showed
ψ ∈ D[ 5
2 , 7
in the proof of Prop. 35, the GSF ϕ + ψ ∈ GC∞(U,eR) is compactly supported in
2(cid:3)R ∪(cid:2) 5
2(cid:3)R(cid:3) = interl(H ∪ K) 6⊆ U . Finally, let f := ϕU and g := ψU , so
(cid:2)(cid:2)− 1
that f ∈ GDH (U,eR) and g ∈ GDK(U,eR). Rem. 20.(i) yields that f +g /∈ GDJ (U,eR)
2(cid:1)R and M of
for all J ⋐f U : Otherwise, taking suitable sub-intervals L of (cid:0)− 1
(cid:0) 5
2(cid:1)R where f resp. g do not vanish, we would have [L∪M ] ⊆ supp(f +g) ⊆ J ⊆ U
2 , 7
(here supp(f + g) is the support as in Def. 11). But the inclusion [L ∪ M ] ⊆ U is
impossible -- a counterexample can be constructed similar to the above x.
2(cid:3)R(cid:3), K :=(cid:2)(cid:2) 5
2 , 7
2 , 1
2 , 1
2 , 1
2 , 7
2 ]R
2 ]R
This example shows that an assumption like (4.2) is necessary to have the closure
of the space GD(U,eR) with respect to sum.
5. Topological structure on GDK
Using our eR-valued norms, it is now natural to define
Definition 38. Let ∅ 6= K ⋐f U ⊆ eRn, where U is a sharply open set. Let
f ∈ GDK(U,eRd), m ∈ N, ρ ∈ eR>0, then
ρ (f ) := ng ∈ GDK(U,eRd) kf − gkm < ρo. In case any confusion could
If V ⊆ GDK(U,eRd), then we say that V is a sharply open set if
Bm
arise, we will use the more precise symbol Bm
ρ (f, K) := Bm
ρ (f ).
(ii)
(i)
Moreover, we say that V is Fermat open if
∀v ∈ V ∃m ∈ N∃ρ ∈ eR>0 : Bm
∀v ∈ V ∃m ∈ N∃r ∈ R>0 : Bm
ρ (v) ⊆ V.
r (v) ⊆ V.
As in [18, Thm. 2] it follows that sharply open sets as well as Fermat open sets
form topologies on GDK(U,eRd).
On the other hand, it is also natural to view the space GDK(U,eRd) inside Garetto's
theory [11, 12] of eR-locally convex algebras.
In this section, we will realize this
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
21
comparison, proving that the space GDK(U,eR) is a Fr´echet eR-module and a topo-
logical algebra. For this purpose, we will only consider the sharp topology.
In-
deed, as we will see below, the Fermat topology is less interesting in this con-
text since it doesn't permit to prove the continuity of the product by scalars
developed in [11, 12], is that the domain U contains generalized points.
empty sharply open set. The main problem in performing this comparison, which
(r, f ) ∈ eR × GDK(U,eRd) 7→ r · f ∈ GDK(U,eRd).
In the following, we will always assume that ∅ 6= K ⋐f U ⊆ eRn, where U is a non-
doesn't permit to view our space GDK(U,eR) as a particular case of the theory
Using the valuation v on eR, it is natural to introduce the following notions:
Definition 39. Let m ∈ N and f ∈ GD(U,eR), then:
vm(f ) := v(kfkm) ∈ R
(i)
(ii) Pm(f ) := e−vm(f ).
From the properties of the valuation v and of the e-norm − e = e−v(−) on eR
Proposition 40. For each m ∈ N, we have:
(i)
(see [4]), the following result directly follows.
vm : GD(U ) −→ R ∪ {+∞} is a valuation, i.e. for all f , g ∈ GD(U ):
• vm(0) = +∞
• vm(λ · f ) ≥ v(λ) + vm(f ) ∀λ ∈ eR
• vm( dεa · f ) = v( dεa) + vm(f ) = a + vm(f ) ∀a ∈ R
• vm(f + g) ≥ min [vm(f ), vm(g)].
(ii) Pm : GD(U ) −→ R is an ultra-pseudo-norm, i.e. for all f , g ∈ GD(U ):
• Pm(f ) = 0 if and only if f = 0
• Pm(λ · f ) ≤ λe · Pm(f ) ∀λ ∈ eR
• Pm( dεa · f ) = dεae · Pm(f ) = e−a · Pm(f ) ∀a ∈ R
• Pm(f + g) ≤ max [Pm(f ),Pm(g)].
The following result states that to define the sharp topology, instead of the above
ultra-pseudo-norms we can equivalently use the countable family of generalized
norms (kfkm)m∈N.
Theorem 41.
(i)
Sum and product in GDK(U ) are continuous in the sharp topology. Therefore,
GDK(U ) is a topological eR-algebra.
subspace {(f, g) ∈ GDK(U ) × GDK(U ) ∀ m ∈ N : kfkm, kgkm < ∞}.
(ii) The product in GDK(U ) is continuous in the Fermat topology only on the
(iii) For f ∈ GDK(U ), set
Cm
r (f ) := {g ∈ GDK(U ) Pm(f − g) < r}
(r ∈ R>0, m ∈ N).
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
22
Then for each q, s ∈ R>0 we have:
(i)
(ii)
If q ≤ − log r, then Cm
r (f ) ⊆ Bm
If q ≥ − log s and s < r, then Bm
dεq (f )
dεq (f ) ⊆ Cm
r (f ).
(iv) The sharp topology on GDK(U ) is the coarsest topology such that each Pm is
continuous.
(v)
GDK(U ) is a separated locally convex topological eR-module.
Proof. (i), (ii): Continuity of the sum in the sharp (and in the Fermat) topology
follows directly from the triangle inequality Prop. 34.(iv). The continuity of the
product at (f0, g0) and property (ii) follow from Prop. 34 (v) via
kf · g − f0 · g0km ≤ cm(kf − f0km · kg − g0km + kf − f0km · kg0km+
+ kf0km · kg − g0km).
(iii): Let us first assume q ≤ − log r and g ∈ Cm
r (f ), so that Pm(f − g) < r and
vm(f − g) > − log r. This implies
∃b > − log r : max
α≤m
i≤d
sup
x∈Kε(cid:12)(cid:12)∂αui
ε(x) − ∂αvi
ε(x)(cid:12)(cid:12) = O(εb),
(5.1)
where (uε) and (vε) define f and g, respectively. Property (5.1) yields the existence
of some M > 0 such that for ε sufficiently small we obtain
sup
x∈Kε(cid:12)(cid:12)∂αui
max
α≤m
i≤d
ε(x) − ∂αvi
ε(x)(cid:12)(cid:12) + εb ≤ (M + 1) · εb < ε− log r ≤ εq.
Therefore kf − gkm + dεb ≤ dεq, so kf − gkm < dεq.
Now, let us assume q ≥ − log s, s < r, and g ∈ Bm
Therefore,
∀0ε : max
α≤m
i≤d
sup
x∈Kε(cid:12)(cid:12)∂αui
ε(x) − ∂αvi
Taking the − e-norm we get
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
sup
x∈Kε(cid:12)(cid:12)∂αui
ε(x)(cid:12)(cid:12)
ε(x) − ∂αvi
max
α≤m
i≤d
that is Pm(f − g) < r as claimed.
(iv) follows directly from (iii) and Prop. 40.(ii).
(v) follows from Prop. 34.(ii) and [11, Prop. 1.11].
dεq (f ), so that kf − gkm < dεq.
ε(x)(cid:12)(cid:12) < εq ≤ ε− log s.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)e
≤ elog s = s < r,
(cid:3)
5.1. Generalized functions and non-Archimedean properties. In this sec-
tion, we want to clarify some relationships between the classical notion of convexity,
We have seen that balls Bm
the notion of eR-convexity of [11] and the use of eR-valued norms.
GDK(U ); they are convex in the usual sense, i.e. if f , g ∈ Bm
ρ (0), ρ ∈ eR>0, define a neighborhood system of 0 for
ρ (0), t ∈ [0, 1] (in
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
23
particular if t ∈ [0, 1]R), then
ktf + (1 − t)gkm ≤ tkfkm + (1 − t)kgkm < tρ + (1 − t)ρ = ρ.
ρ (0) is also balanced: if t ∈ eR, t ≤ 1, then t·Bm
Moreover, each ball Bm
ρ (0).
However, this space is not a classical locally convex topological vector space over
the field R because of two reasons: (i) the product by scalars is not continuous with
respect to the Euclidean topology on R, (ii) Lemma 23 implies that the property
ρ (0) ⊆ Bm
∀f ∈ GDK(U )∃t ∈ eR : f ∈ t · Bm
ρ (0)
proof of Thm. 41 (ii), this is a necessary consequence of the existence of generalized
holds for t ∈ eR but it cannot be extended to t ∈ R. As we have seen in the
functions with infinite eR-valued k − km-norm.
On the other hand, even though the sets Cm
r (0) are defined using R only, i.e.
without mentioning any non-Archimedean property, they satisfy
∀f ∈ Cm
r (0)∀λ ∈ R : λ · f ∈ Cm
r (0),
(5.2)
and this is possible only because they are infinitesimal sets. In fact, we have seen
in Thm. 41 that Cm
r (0) ⊆ Bm
dεq (0) for q ≤ − log r.
More generally, a set A ⊆ GDK(U ) can be eR-balanced (see [11]), i.e.
λA ⊆ A ∀λ ∈ eR :
λe ≤ 1,
and at the same time can be thought of as "small" only in case A consists infini-
tesimal points. For example, the ball Bm
1 (0) is not. In
fact, we have
dεb (0) is eR-balanced, but Bm
Lemma 42. Suppose that A ⊆ GDK(U ) and m ∈ N are such that
A + A ⊆ A
∃r ∈ R>0 : A ⊆ Bm
r (0).
(5.3)
r (0) for all n ∈ N
6=0. Therefore, kukm < r
Then every element u ∈ A has infinitesimal norm: kukm ≈ 0.
Proof. In fact, (5.3) implies n·u ∈ A ⊆ Bm
for all n ∈ N
6=0, which proves our claim.
Let us note that condition (5.3) holds both for A which is eR-balanced or eR-
naturally induced to consider a topology on eR which contains infinitesimal neigh-
convex.
These remarks permit to show that in dealing with generalized functions, we are
borhoods (hence inducing the discrete topology on R, see [17]). This is due to
the coexistence of a continuous product by scalars and of an infinite element in
GDK(U ), as stated in the following general result. In a possible interpretation of
n
(cid:3)
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
24
topology τ , and < eR as the strict order relation < of Lemma 1.
its statement, we can think of R as R with a topology τ , eR as eR with the sharp
Theorem 43. Let (R, +R,·R, <R, τ ) and (eR, + eR,· eR, < eR, τ ) be Hausdorff topological
ordered rings such that (R, +,·, <) is a substructure of (R, +R,·R, <R), which in
turn is a substructure of (eR, + eR,· eR, < eR) and such that
Let (G, +G,·G, σ) be a Hausdorff topological R-module, and −G : G −→ eR, −R :
R −→ R be maps such that r ·G gG = rR · eR gG for all r ∈ R and all g ∈ G. As-
sume that any τ -neighborhood of 0 ∈ R contains a ball BR
η (0) := {s ∈ R sR <R η}
for some η ∈ R, η >R 0, and that there exists some ρ ∈ eR with ρ > eR 0 such that
∀r ∈ R ∀s ∈ R : r <eR s =⇒ r <R s.
ρ (0) := {g ∈ G gG < eR ρ} is σ-open. Finally, assume that
gG > eR ρ · eR M.
gG is invertible , ∀M ∈ R>0 :
∃g ∈ G :
the ball BG
(5.4)
(5.5)
Then the induced topology τ ∩ R is discrete.
Proof. Since G is a topological R-module, the product by scalars is τ×σ-continuous,
and
r ·G g = 0,
(5.6)
lim
r→0
r∈R
where g ∈ G comes from assumption (5.5). By hypothesis, the ball BG
and every τ -neighborhood of r = 0 ∈ R contains some ball BR
entails that there exists some η >R 0 such that
ρ (0) ∈ σ
η (0). Therefore (5.6)
∀r ∈ R :
rR <R η =⇒ ρ > eR r ·G gG = rR · eR gG.
(5.7)
For each s ∈ R>0 take M ∈ R>0 such that 1
from (5.5). For all r ∈ R such that rR <R η, we have
< s
< eR
rR < eR
ρ
gG
1
M
M < s, so that gG > eR ρ · eR M > eR 0
because gG is invertible in eR. Therefore rR < eR s and hence rR <R s by (5.4).
This means that r is infinitesimal in the ring R, i.e. the ball BR
ρ (0) is contained
in the monad of 0 (see e.g. [17] for the notion of monad) and so also every ball
BR
η (¯r) is contained in the monad of ¯r ∈ R. Therefore, [17, Prop. 2.1] implies the
conclusion.
(cid:3)
We can therefore say that if we want to find a space G of generalized functions
which is an ordinary Hausdorff topological vector space on R, then we cannot
define the topologies τ and σ using seminorms valued in a non-Archimedean (see
(5.5)) extension of R. This results confirms Rem. 43 of [10].
As a consequence, we have the following impossibility result:
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
25
(G, +G,·G) is a linear subspace of GDK(U ) for some ∅ 6= K ⋐f U ⊆ eRn
Corollary 44. There does not exist any real Hausdorff topological vector space
(G, +G,·G, σ) such that:
(i)
(ii) G contains the Dirac delta δ ∈ G
(iii) ∃m ∈ N∃ρ ∈ eR>0 : ρ < 1 , Bm
In particular, the Colombeau algebra Gs(Ω) does not contain any real Hausdorff
topological vector subspace G such that some Bm
ρ (0) ∩ G ∈ σ
ρ (0) ∩ G is open and δ ∈ G.
Proof. By contradiction, in Thm. 43, set R := R with the usual Euclidean topology
τ , eR := eR with the sharp topology; set gG := kgkm, where m ∈ N comes from (iii)
and we used the inclusion (i); set rR := r the usual absolute value in R. Note
also that Bm
ρ (0) ∩ G = {g ∈ G kgkm < ρ} = {g ∈ G gG < eR ρ}. If δ ∈ G,
then δG = kδkm is infinite and invertible in eR, so that Thm. 43 implies that the
Euclidean topology would be discrete. The second part of the claim follows from
Thm. 28.
(cid:3)
We can summarize Cor. 44 by saying that a real Hausdorff topological vector
space structure G for a space of generalized functions cannot contain even a single
trace Bm
ρ (0) ∩ G, ρ < 1, of a sharply open ball. This result does not contradict
[1, Prop. 4], where it is stated that the sharp topology induces on bounded sets
of the real locally convex space Ga(Ω) ⊆ Gs(Ω) a topology which is finer than the
topology σa on Ga(Ω). On the other hand, Cor. 44 implies that Bm
ρ (0)∩Ga(Ω) /∈ σa
for all ρ ∈ eR>0, ρ < 1.
6. Metric structure on GDK
In this section, we want to use [11, Thm. 1.14] to prove metrizability of GDK(U ).
However, we will apply this result using an explicit and simple countable base of
neighborhoods of the origin which consists of eR-absorbent and absolutely eR-convex
The idea is to consider only points of balls f ∈ Bm
ρ (0) whose norm kfkm is
infinitely smaller than ρ. To formally express this idea, we introduce the following
sets. In this way, we will arrive at a simpler metric.
Definition 45. Let ρ ∈ eR>0, m ∈ N and g ∈ GDK(U ). Then
ρ · kf − gkm ≈ 0(cid:27) .
ρ (g) :=(cid:26)f ∈ GDK(U )
U m
1
In case any confusion might arise, we will use the more precise symbol U m
U m
ρ (g).
ρ (g, K) :=
Proposition 46.
(i)
U m
ρ (0) is eR-absorbent and absolutely eR-convex.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
26
(ii) Both the system
nU m
ρ (v) v ∈ GDK(U ), m ∈ N, ρ ∈ eR>0, ρ ≈ 0o ,
and the system
{U n
dεn (v) v ∈ GDK(U ), n ∈ N>0}
generate the sharp topology on GDK(U ).
i.e. u
dεb ∈ U m
Proof. To see that U m
ρ (0) is eR-absorbent, by [11, Def. 1.1] we have to show that
dεb·ρ ≈ 0. But ρ is strictly
positive, so ρ ≥ dεp for some p ∈ R. Moreover, kukm ∈ eR, so kukm ≤ dεq for some
q ∈ R. Therefore
∀u ∈ GDK(U )∃a ∈ R∀b ∈ R
ρ (0), which is equivalent to (cid:13)(cid:13) u
≤a : u ∈ dεb · U m
dεb(cid:13)(cid:13)m · 1
kukm
dεb · ρ ≤ dεq−b−p,
ρ = kukm
ρ (0),
and we have dεq−b−p ≈ 0 if and only if b < q − p. This proves (6.1).
(6.1)
ρ (0) is balanced, assume λ ∈ eR with λe ≤ 1. Then v(λ) ≥ 0,
ρ (0) then
To prove that U m
so λ ≤ c for some c ∈ R>0. Therefore, if u ∈ U m
= λkukm
kλukm
ρ
ρ ≈ 0,
1
ρ (0).
ρ (0), we have
In order to prove (ii), we note that U m
so λu ∈ U m
Finally, we show eR-convexity: for all a, b ∈ R
ρ · k dεa · u + dεb · vkm ≤ dεa · kukm
ρ (v) ⊆ Bm
≈ 0 implies
ku−vkm
ρ (v) ⊆ U m√ρ(v) because ku − vkm < ρ implies
1√ρ·ku−vkm ≤ √ρ ≈ 0. Finally, if ρ ≈ 0, every U m
ρ (v)
and w ∈ Bm
ρ ·ku− vkm ≈ 0. The proof for
the second system in (ii) follows by observing that given ρ > 0, ρ ≈ 0, there exists
q ∈ N such that ρ ≥ dεq, and setting n := max(m, q) we have U n
ρ (v). (cid:3)
≥0 and all u, v ∈ U m
+ dεb · kvkm
ρ ≈ 0.
ρ (v) because ku−vkm
ρ (v) is sharply open: if u ∈ U m
< 1. Also, if ρ ≈ 0, then Bm
ρ ·kw− vkm ≤ 1
ρ ·kw− ukm + 1
dεn (v) ⊆ U m
ρ2 (u), then 1
ρ
ρ
ρ
From [11, Thm. 1.14], we have that GDK(U ) is metrizable with metric
d2(u, v) =
+∞Xn=1
2−n · minnPU n
dεn (0)(u − v), 1o .
(6.2)
Concerning (6.2) we recall (see [11]) that if A ⊆ GDK(U ) is eR-absorbent, then, for
all u ∈ GDK(U ), we define
VA(u) := sup(cid:8)b ∈ R u ∈ dεb · A(cid:9)
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
27
The following result gives a metric which is equivalent to (6.2) but is defined by a
simpler formula.
PA(u) := e−VA(u).
(6.3)
Proposition 47. Set An := U n
dεn (0) for n ∈ N>0, and let u ∈ GDK(U ). Then
(i)
VAn (u) = vn(u) − n
(ii) The map
de(u, v) =
+∞Xn=1
emin[n−vn(u−v),0]−n.
is a metric on GDK(U ) that is equivalent to d2.
Proof. Concerning (i), we note that
u ∈ dεb · An ⇒
⇒ b ≤ vn(u) − n,
u
dεb ∈ U n
dεn (0) ⇒ kukn
dεb+n ≈ 0 ⇒ v(cid:0) dε−b−n · kukn(cid:1) ≥ 0
so VAn (u) ≤ vn(u) − n. Conversely, if we had VAn (u) < vn(u) − n, we could
pick VAn (u) < b < vn(u) − n. Then as above it would follow that kukn
dεb+n ≈ 0,
contradicting the definition of VAn (u). This proves (i).
In order to prove (ii), we use (i) in (6.2): PAn(u − v) = e−vn(u−v)+n and
d2(u, v) =
=
≥
2−n · minne−vn(u−v)+n, 1o =
2−n · emin[n−vn(u−v),0] ≥
emin[n−vn(u−v),0]−n = de(u, v).
+∞Xn=1
+∞Xn=1
+∞Xn=1
But we also have e−n ≥ 2−n−1 for all n, so that de(u, v) ≥ 1
2 d2(u, v). Using
Prop. 40, it is easily checked that de is a metric, which we have just proved to be
equivalent to d2.
(cid:3)
7. Completeness of GDK
In order to prove the completeness of GDK(U ), we generalize the proof of [11,
Prop. 3.4] (based in turn on [28]) to the present context.
Theorem 48. The space GDK(U ) with the sharp topology is complete.
Proof. By Prop. 47 GDK(U ) with the sharp topology is metrizable. Hence, it
suffices to consider a Cauchy sequence (un)n∈N in this topology, i.e.,
∀q ∈ R>0 ∀i ∈ N∃N ∈ N∀m, n ≥ N : kun − umki < dεq.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
28
Setting i = q = k ∈ N>0, this implies the existence of a strictly increasing sequence
(nk)k∈N in N such that kunk+1 − unkkk < dεk. Hence picking any representatives
(unε) of un as in Def. 19 we have
(cid:20)max
α≤k
sup
x∈Rn(cid:12)(cid:12)∂αunk+1,ε(x) − ∂αunk,ε(x)(cid:12)(cid:12)(cid:21) <(cid:2)εk(cid:3)
∀k ∈ N>0.
By Lemma 1 this yields that for each k ∈ N>0 there exists an εk such that εk ց 0
and
∀ε ∈ (0, εk) : max
α≤k
sup
x∈Rn(cid:12)(cid:12)∂αunk+1,ε(x) − ∂αunk,ε(x)(cid:12)(cid:12) < εk.
Now set
(7.1)
(7.2)
hkε :=
unk+1,ε − unk,ε ∈ C∞(Rn, R)
0 ∈ C∞(Rn, R)
uε := un0ε +
∞Xk=0
hkε ∀ε ∈ I.
if ε ∈ (0, εk)
if ε ∈ [εk, 1)
Since εk ց 0, for all ε ∈ I we have ε /∈ (0, εk) for all k ≥ ¯k, with ¯k sufficiently big.
Therefore, uε = un¯k+1,ε ∈ C∞(Rn, R). In order to prove that (uε) defines a GSF of
the type U → eR, take [xε] ∈ U and α ∈ N. We claim that (∂αuε(xε)) ∈ RM . Now
if p ∈ N satisfies α ≤ p, then for any x ∈ Rn we have
∞Xk=p+1
∂αuε(x) ≤(cid:12)(cid:12)∂αunp+1,ε(x)(cid:12)(cid:12) +
∂αhkε(x) .
From (7.1) and (7.2) we get that ∂αhkε(x) ≤ εk for all k ≥ p + 1, x ∈ Rn and all
ε ∈ (0, 1]. Hence for ε ∈ (0, 1], α ≤ p and all x ∈ Rn we obtain
∂αuε(x) ≤(cid:12)(cid:12)∂αunp+1,ε(x)(cid:12)(cid:12) +
εp+1
1 − ε
.
(7.3)
Inserting x = xε and noting that (∂αunp+1,ε(xε)) ∈ RM proves our claim. Moreover,
since all (unε) satisfy Def. 19, we also conclude from (7.3) that for any α and any
[xε] ∈ ext(K) we have [∂αuε(xε)] = 0, and hence the GSF [uε(−)]U ∈ GDK(U ).
Finally, ku− unpki < dεp−1 for all p ∈ N>1 and all i ≤ p. This yields that (unk )k
tends to u in the sharp topology, and hence so does (un).
(cid:3)
8. The space GD as inductive limit of GDK
In this section, we always assume that U ⊆ eRn is a non-empty strongly internal
set. By Prop. 35 and Prop. 36, this entails that GD(U ) is an eR-module.
In order to define a natural topology on GD(U ) we will employ [11, Thm. 1.18],
which we restate here for the reader's convenience:
Theorem 49. Let G be an eR-module, (Gγ)γ∈Γ be a family of locally convex topolog-
ical eR-modules and, for each γ ∈ Γ, let iγ : Gγ −→ G be an eR-linear map. Assume
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
29
that
G = span
[γ∈Γ
iγ (Gγ)
γ (V ) is a
−→ Gγ.
(see [11] and Def. (6.3)) is the finest
and let V ∈ V if and only if V ⊆ G, V is absolutely eR-convex and i−1
neighborhood of 0 in Gγ for all γ ∈ Γ. Then each V ∈ V is eR-absorbent and the
topology τ induced by the gauges {PV }V ∈V
eR-locally convex topology on G such that iγ is continuous for all γ ∈ Γ. Endowed
with this topology, G is called the inductive limit (colimit) of the spaces (Gγ)γ∈Γ and
we write G = lim
Since GD(U ) =S∅6=K⋐fU GDK(U ), we may therefore equip it with the inductive
limit topology with respect to the inclusions ιK : GDK(U ) ֒→ GD(U ). We call the
resulting eR-locally convex topology the sharp topology on GD(U ). Hence
Henceforth we will denote the sharp topology on GDK(U ) by σK(U ) (or, for short,
by σK). Also, the inductive limit topology on GD(U ) will be denoted by σ(U ) (or
by σ). Setting
−→ GDK(U )
(∅ 6= K ⋐f U ).
GD(U ) = lim
dεn (0) :=(cid:26)f ∈ GD(U ) kfkn
dεn ≈ 0(cid:27) ,
U n
where n ∈ N>0, we obtain an eR-absorbent and absolutely eR-convex subset of GD(U )
such that i−1
dεn (0, K) ∈ σK. Therefore, these
sets generate a coarser topology than the sharp topology σ. The proof is identical
to that of Prop. 46 (i).
dεn (0) ∩ GDK(U ) = U n
dεn (0)) = U n
K (U n
From the (co-)universal property of inductive limits ([11, Prop. 1.19]) we imme-
diately conclude:
Proposition 50. Let H be a locally convex topological eR-module. For each non-
empty K ⋐f U , let TK : GDK(U ) −→ H be an eR-linear and continuous map.
Assume that TK(f ) = TH (f ) if f ∈ GDK(U )∩GDH (U ). Then there exists one and
only one map T : GD(U ) −→ H which is eR-linear and continuous and such that
T ◦ ιK = TK for all non-empty K ⋐f U .
The work [11] includes a detailed analysis of countable inductive limits G =
−→ Gn (n ∈ N), where (Gn)n∈N is increasing, G =Sn∈N Gn, and where the topology
lim
on Gn is that induced by Gn+1. Such inductive limits are called strict. As in the
case of classical function spaces like D(Ω) for Ω open in Rn, the importance of
strict inductive limits in the theory of eR-locally convex models ultimately stems
from the possibility of covering every open set Ω ⊆ Rn by a countable increasing
family of compact sets. The key point in the structure theory of strict inductive
limits as above is that a countable family (Gn)n∈N permits to define recursively a
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
30
family of neighborhoods of 0. Using the latter, one can prove that the topology on
G induces on each Gn its given topology. We will show below that similar properties
hold for GD(U ). We shall see that the assumption of U being strongly internal and
sharply open are essential for this task. To begin with, we prove a strengthening
of Thm. 29:
Proposition 51. Let H, K ⋐f U be non-empty sets, with H ⊆ K. Then GDH (U )
is a topological subspace of GDK(U ), i.e.,
σK(U )GDH (U) = σH (U ).
Proof. In this proof we will use the more precise notation Bm
Def. 38).
ρ (u, K) for balls (see
Let V ∈ σK. We claim that V ∩GDH (U ) ∈ σH . For each u ∈ V ∩GDH (U ) there
exist m ∈ N and ρ ∈ eR>0 such that Bm
ρ (u, K)
because GDH (U ) ⊆ GDK(U ) and because the norm k − km doesn't depend on H,
K. Therefore, Bm
ρ (u, K) ⊆ V . But Bm
ρ (u, H) ⊆ Bm
ρ (u, H) ⊆ V ∩ GDH (U ), which proves our claim.
ρ (u, K) u ∈ W , m ∈ N , ρ ∈ eR>0 , Bm
Conversely, if W ∈ σH , then we set
V :=[nBm
and we claim that W = V ∩GDH (U ). In fact, since W ∈ σH , for all u ∈ W we have
ρ (u, H) ⊆ W for some ρ and m. By (8.1) Bm
Bm
ρ (u, K) ⊆ V , and so u ∈ V ∩GDH (U ).
Vice versa, if u ∈ V ∩ GDH (U ), then u ∈ Bm
ρ (v, K) for some v ∈ W , m, ρ, such
that Bm
ρ (u, H) ⊆ Wo ∈ σK,
ρ (v, H) ⊆ W . So ku − vkm < ρ and hence u ∈ Bm
ρ (v, H) ⊆ W .
(8.1)
(cid:3)
We now show that the space GD(U ) can be seen as a strict inductive limit of
a countable increasing family of subspaces GDK(U ). Indeed, since U is strongly
internal, we can write U = hUεi for some net (Uε) of subsets of Rn. Since U is
non-empty, by [18, Thm. 8], fixing any x = [xε] ∈ U we obtain:
ε ) > εN , xε ≤ ε−N .
∃N ∈ N∀0ε : d(xε, U c
(8.2)
With N as in (8.2), we define
ε ) ≥ εj , x ≤ ε−j(cid:9)
Kjε : =(cid:8)x ∈ Rn d(x, U c
Kj : = [Kjε] ∀j ∈ N
≥N .
Using this notation, we have:
Theorem 52. If N ∈ N satisfies (8.2) for some x ∈ U , then
∅ 6= Kj ⋐f U for all j ∈ N
(i)
(ii) Kj ⊆ Kj+1 for all j ∈ N
≥N
(iii) U =Sj≥N Kj
(iv) For every ∅ 6= K ⋐f U = hUεi there exists some j ≥ N such that K ⊆ Kj.
≥N
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
31
(v)
GD(U ) is the strict inductive limit of the family GDKj (U ), j ≥ N :
GD(U ) = lim
−→ GDKj (U )
(j ≥ N ).
Proof. (i), (ii): It follows immediately from the definition that each (Kjε) is com-
pact and that (Kjε)ε is sharply bounded, so Kj ⋐f U . Moreover, KN 6= ∅ by (8.2),
hence Kj 6= ∅ follows by (ii), which again is immediate from the definition.
(iii): If x = [xε] ∈ U = hUεi, then d(xε, U c
ε ) > εj1 for ε small and for some
j1 ∈ N>0. Since (xε) is moderate, xε ≤ ε−j2 for ε small and some j2 ∈ N. Setting
j := max(j1, j2, N ) we hence have that x ∈ Kj.
In order to prove (iv), we need the following strengthening of [18, Thm. 11]:
Lemma 53. Let H = [Hε] ⋐f V = hVεi, then
∃j ∈ N∀[xε] ∈ [Hε]∀0ε : d(xε, V c
ε ) ≥ εj.
(8.3)
Proof of Lemma 53. Equation (8.3) expresses that for all representatives (xε) ∈
Rn
M , if xε ∈ Hε for ε small, then ∀0ε : d(xε, V c
By contradiction, assume that
ε ) ≥ εj.
M : (cid:0)∀0ε : xjε ∈ Hε(cid:1) , ∃(εjk)k ↓ 0 ∀k : d(xjεjk , V c
∀j ∈ N∃(xjε) ∈ Rn
jk.
By recursively applying this condition, we get that for all j ∈ N there exists a
moderate (xjε) and some εj ∈ (0, 1] such that xjε ∈ Hε for ε ≤ εj and
εjk ) < εj
d(xjεjk , V c
εjk ) < εj
jk,
(8.4)
where (εjk)k ↓ 0. Since (εjk)k ↓ 0, without loss of generality we can assume to have
defined recursively (εj)j so that (εj)j ↓ 0 and εj > εjkj > εj+1 for some subsequence
(kj)j ↑ +∞. Set xε := xjε ∈ Hε for ε ∈ (εj+1, εj], so that xεjkj
for all
j. Then (xε) ∈ Rn
M since H is sharply bounded and x := [xε] ∈ H ⊆ hVεi, which
entails
= xjεjkj
(8.5)
Therefore, for j ∈ N sufficiently big, (8.5) holds at ε = εjkj < 1 and j > q. Thus
d(xjεjkj
∃q ∈ R>0 ∀0ε : d(xε, V c
, which contradicts (8.4).
) = d(xεjkj
ε ) > εq.
) > εq
jkj
(cid:3)
, V c
εjkj
, V c
εjkj
> εj
jkj
Continuing the proof of (iv), if K ⋐f U is non-empty, by applying Lemma 53 we
obtain
∃j1 ∈ N∀[xε] ∈ K ∀0ε : d(xε, U c
ε ) ≥ εj1 .
On the other hand, sharp boundedness of (Kε), where K = [Kε], implies
∃j2 ∈ N∀[xε] ∈ K ∀0ε :
xε ≤ ε−j2 .
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
32
The converse inclusion follows directly from (i).
Therefore, for j := max(j1, j2, N ), we get K ⊆ Kj, hence (iv), and thereby also
GDK(U ) ⊆ GDKj (U ) (using Thm. 29). It follows that GD(U ) ⊆ Sj≥N GDKj (U ).
It remains to prove that the topology σ on GD(U ) coincides with the inductive
eR-locally convex topology generated by (cid:0)GDKj (U )(cid:1)j≥N . Let us denote the latter
topology by σ′. We have σ ⊆ σ′ by definition of the inductive topology and by
Thm. 52.(i). To see that, conversely, σ′ ⊆ σ we show that for every K ⋐f U the
inclusion (GDK(U ), σK) ֒→ (GD(U ), σ′) is continuous (see Thm. 49). Now given
any K ⋐f U , by what we have proved above there exists some j ≥ N with K ⊆ Kj.
But then Prop. 51 implies the continuity of
(GDK(U ), σK) ֒→ (GDKj (U ), σKj ) ֒→ (GD(U ), σ′)
and thereby our claim.
(cid:3)
As a consequence of this result, we can now prove a series of corollaries by
applying the general theorems of [11] concerning countable strict inductive limits.
Corollary 54. If ∅ 6= K ⋐f U = hUεi, then GDK(U ) is a topological subspace of
GD(U ), i.e., σ(U )GDK (U) = σK (U ).
Proof. According to Thm. 52 (iv) we may pick j ≥ N such that K ⊆ Kj. By [11,
Prop. 1.21], GDKj (U ) carries the trace topology of GD(U ). Since, in turn, GDK(U )
is a topological subspace of GDKj (U ) by Prop. 51, the claim follows.
Corollary 55. GD(U ) is separated.
Proof. This follows from Thm. 41 (v) and [11, Cor. 1.24].
(cid:3)
(cid:3)
Lemma 56. If ∅ 6= H, K ⋐f U and H ⊆ K, then GDH (U ) is closed in GDK(U ).
Proof. This is immediate from Prop. 51 and Thm. 48.
(cid:3)
Corollary 57. If ∅ 6= K ⋐f U , then GDK(U ) is closed in GD(U ).
Proof. This follows from Thm. 48, Cor. 55, and Cor. 54.
(cid:3)
Corollary 58. Let B ⊆ GD(U ), then B is bounded in GD(U ) if and only if there
exists a non-empty K ⋐f U such that B is bounded in GDK(U ).
Proof. ⇒: This follows from [11, Thm. 1.26] and Lemma 56.
⇐: If B is bounded in GDK(U ), then [11, Lem. 1.27] yields
∀(un)n ∈ BN ∀(λn)n ∈ eRN : λn → 0 in eR =⇒ λnun → 0 in GDK(U ).
Pick j ≥ N such that K ⊆ Kj. Since generalized norms k − km do not depend
on K, Kj, condition (8.6) holds also in GDKj (U ) ⊇ GDK(U ). Therefore, from [11,
(8.6)
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
33
Lem. 1.27] we get that B is bounded in GDKj (U ) and [11, Thm. 1.26] yields that
B is bounded in GD(U ).
(cid:3)
A similar proof applies to this corollary, which is a consequence of [11, Cor. 1.29]:
Corollary 59. Let (un)n ∈ GD(U )N, then un → 0 in GD(U ) if and only if there
exists a non-empty K ⋐f U such that un ∈ GDK(U ) and un → 0 in GDK(U ).
Finally, from [11, Thm. 1.32], Lemma 56 and Thm. 48 we obtain:
Corollary 60. GD(U ) with the sharp topology is complete.
Using Lemma 53, we can also show that any compactly supported generalized
assuming that U is a strongly internal set.
a net (uε) of smooth functions which are compactly supported in an arbitrarily
smooth function f ∈ GDK(U, Y ) on a sharply open set U ⊆ eRn is defined by
εa(0)i of K = [Kε]. We recall that in this section we are
small extension hKε + B E
Theorem 61. Let ∅ 6= K ⋐f U . Let Y ⊆ eRd, f ∈ GDK(U, Y ) and K = [Kε] ⋐f eRn.
Let j ∈ N be as in (8.3) and a ∈ R such that a ≥ j. Then there exist nets (uε),
(Hε) such that:
εa (0) for all ε.
[Hε] ⋐f U
(uε) defines f and uε ∈ DHε(Rn, Rd) for all ε
(i)
(ii)
(iii) Hε ⊆ Kε + B E
Proof. Let U = hUεi, where each Uε ⊆ Rn is an open set (cf. [18, Cor. 9]), and let
(vε) satisfy Def. 19 for f and K = [Kε]. By [18, Thm. 11], we can assume Kε ⊆ Uε
εa/2(0), and denote by χLε the characteristic function
for all ε. Let Lε := Kε + B E
1 (0)) have unit integral and set ψε := (εa/2)−nψ(2x/εa). Then
of Lε. Let ψ ∈ D(B E
M (Rn), and (χLε ∗ ψεa )Kε = 1. Set uε := (χLε ∗ ψε) · vε. Then (χLε ∗ ψε)
(ψε) ∈ E s
defines a GSF of the type eRn −→ eR and hence (uε) defines a GSF of the type
U −→ eRd. Moreover, Hε := supp(uε) ⊆ supp(χLε ∗ ψε) ⊆ Kε + B E
a ≥ j, we have [Hε] ⊆ hUεi = U .
It remains to prove that f (x) = [uε(xε)] for all x = [xε] ∈ U . Suppose this was
not the case. Then there would exist some y = [yε] ∈ U , some b > 0 and a sequence
εk ց 0 such that
εa (0). Since
uεk (yεk ) − vεk (yεk ) ≥ εb
(8.7)
for all k ∈ N. By Lemma 21, we may without loss of generality assume that either
yε ∈ Kε for all ε or that y ∈ ext(K). In the first case, vε(yε) = uε(yε) for all ε,
contradicting (8.7). In the second case,
k
Since (vε(yε))) is negligible, we again arrive at a contradiction to (8.7).
(cid:3)
vε(yε) − uε(yε) ≤ 2vε(yε).
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
34
Using this result, we can now prove the extension of property (ii) of Thm. 26 to
Assume that we have an operator I : D(Rn) −→ R with the property that if
(uε) and (vε) define f ∈ GD(U ), where uε, vε ∈ D(Rn), then [I(uε)] = [I(vε)] ∈ eR.
Then Thm. 61 permits to extend I to the whole of GD(U ).
arbitrary codomains Y ⊆ eRd:
Theorem 62. Let ∅ 6= K ⋐f U , Y ⊆ eRd and f ∈ GDK(U, Y ), then ∃! ¯f ∈
GDg(K, Y ) : ¯fK = fK.
Proof. We only have to prove that ¯f (x) ∈ Y for all x ∈ eRn. Let (uε) and (Hε) as
in Thm. 61. By Thm. 24, we have that (uε) also defines ¯f . For each ε pick any
point hε ∈ ∂Hε and set ¯xε := xε if xε ∈ Hε, and ¯xε := hε otherwise. Therefore
¯x := [¯xε] ∈ [Hε] ⊆ U . Then, if xε /∈ Hε, uε(xε) − uε(¯xε) = uε(xε) − uε(hε) = 0
because uε ∈ DHε (Rn, Rd). Thus ¯f (x) = f (¯x) ∈ Y .
(cid:3)
9. Conclusions and Further developments
The notion of functionally compact set we introduced in the present work permits
to show that compactly supported GSF are close analogues of classical compactly
supported smooth functions.
In particular, their functional analytic properties
parallel those of the test function space of distribution theory. At the same time,
for suitable K, the space GDK(eRn) contains extensions to all CGF Gs(Ω) and hence
also all Schwartz distributions.
The theory developed here opens the door to addressing several central topics in
the theory of nonlinear generalized functions from a new angle. As indicated after
Thm. 61, a direct generalization of the integral of compactly supported functions
to compactly supported GSF is feasible. An immediate application of this lies in a
theory of integration for GSF that we hope will allow to harmonize the Schwartz
view of generalized functions as functionals with that prevalent in Colombeau theory
of considering generalized functions as pointwise maps. Our approach will take
inspiration from Garetto's very fruitful duality theory of locally convex eC-modules
[12, 15, 13].
A further natural development of the present article goes in the direction of a
generalization to suitable types of asymptotic gauges (see [20, 19]) and hence to
the full and the diffeomorphism invariant Colombeau algebras.
for ODE with a GSF right hand side, or a Hahn-Banach theorem for functionals
mit to generalize results from classical analysis, like a Picard-Lindelof theorem
Moreover, one can ask whether eR-valued generalized norms in GDK(U ) per-
I : GDK(U ) −→ eR defined by diffeologically smooth functionals (see [21]) of the
type Iε : DKε(Uε) −→ R, analogously to the way a GSF is defined by a net of
smooth functions.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
35
Acknowledgment: We would like to thank H. Vernaeve for helpful discussions
and the anonymous referee for several suggestions that have led to considerable
improvements in Sec. 3.
References
[1] Aragona, J., Colombeau, J.F., Juriaans, S.O., Locally convex topological algebras of general-
ized functions: compactness and nuclearity in a nonlinear context, Trans. Amer. Math. Soc.,
to appear.
[2] Aragona, J., Fernandez, R., Juriaans, S. O., A Discontinuous Colombeau Differential Calcu-
lus. Monatsh. Math. 144 (2005), 13 -- 29.
[3] Aragona, J., Fernandez, R., Juriaans, S.O., Natural topologies on Colombeau algebras,
Topol. Methods Nonlinear Anal. 34 (2009), no. 1, 161 -- 180.
[4] Aragona, J., Juriaans, S. O., Some structural properties of the topological ring of Colombeau's
generalized numbers, Comm. Algebra 29 (2001), no. 5, 2201 -- 2230.
[5] Aragona, J., Juriaans, S.O., Oliveira, O.R.B., Scarpal´ezos, D., Algebraic and geometric theory
of the topological ring of Colombeau generalized functions, Proc. Edinb. Math. Soc. (2) 51
(2008), no. 3, 545564.
[6] Aragona, J., Fernandez, R., Juriaans, S.O., Oberguggenberger, M., Differential calculus and
integration of generalized functions over membranes. Monatsh. Math. 166 (2012), 1 -- 18.
[7] Biagioni, H.A., A Nonlinear Theory of Generalized Functions, Lecture Notes in Mathematics
1421, Springer, Berlin, 1990.
[8] Colombeau, J.F., New generalized functions and multiplication of distributions. North-
Holland, Amsterdam, 1984.
[9] Colombeau, J.F., Elementary introduction to new generalized functions. North-Holland, Am-
sterdam, 1985.
[10] Delcroix, A., Hasler, M.F., Pilipovic, S., Valmorin, V., Sequence spaces with exponent
weights. Realizations of Colombeau type algebras, Dissertationes Mathematicae 447, 1 -- 73
(2007).
[11] Garetto, C., Topological structures in Colombeau algebras: Topological C-modules and du-
ality theory. Acta Appl. Math. 88, no. 1, 81 -- 123 (2005).
[12] Garetto, C., Topological structures in Colombeau algebras:
investigation of the duals of
Gc(Ω), G(Ω) and GS (Rn), Monatsh. Math. 146 (2005), no. 3, 203 -- 226.
[13] Garetto, C., Fundamental solutions in the Colombeau framework: applications to solvability
and regularity theory, Acta. Appl. Math. 102 (2008), 281 -- 318.
[14] Garetto, C., Closed graph and open mapping theorems for topological eC-modules and appli-
cations, Math. Nachr. 282 (2009), no. 8, 1159 -- 1188.
[15] Garetto, C., Hormann, G., On duality theory and pseudodifferential techniques for Colom-
beau algebras: generalized delta functionals, kernels and wave front sets, Bull. Acad. Serbe
Cl. Sci. (2006), 31, 115 -- 136.
[16] Garetto, C., Vernaeve, H., Hilbert eC-modules: structural properties and applications to
variational problems, Trans. Amer. Math. Soc. 363 (2011), no. 4, 2047 -- 2090.
[17] Giordano, P., Kunzinger, M., New topologies on Colombeau generalized numbers and the
Fermat-Reyes theorem, J. Math. Anal. and Appl. 399 (2013) 229 -- 238.
[18] Giordano, P., Kunzinger, M., Vernaeve, H., Strongly internal sets and generalized smooth
functions, J. Math. Anal. Appl. 422 (2015) 56 -- 71.
A CONVENIENT NOTION OF COMPACT SET FOR GENERALIZED FUNCTIONS
36
[19] Giordano, P., Luperi Baglini, L., Asymptotic gauges: Generalization of Colombeau type
algebras, Math. Nachr. 1 -- 28 (2015).
[20] Giordano, P., Nigsch, E., Unifying order structures for Colombeau algebras, Math. Nachr. Vol-
ume 288, Issue 11-12, pages 1286 -- 1302, (2015).
[21] Giordano P., Wu E., Categorical frameworks for generalized functions, Arab. J. Math., Vol-
ume 4, Issue 4, 301 -- 328, 2015.
[22] Grosser, M., Kunzinger, M., Oberguggenberger, M., Steinbauer, R., Geometric theory of
generalized functions, Kluwer, Dordrecht, 2001.
[23] Mayerhofer, E., Spherical completeness of the non-Archimedean ring of Colombeau general-
ized numbers, Bull. Inst. Math. Acad. Sin. (N.S.) 2 (2007), no. 3, 769 -- 783.
[24] Mayerhofer, E., On Lorentz geometry in algebras of generalized functions, Proc. Roy. Soc. Ed-
inburgh Sect. A 138 (2008), no. 4, 843 -- 871.
[25] Oberguggenberger, M. Multiplication of Distributions and Applications to Partial Differential
Equations, volume 259 of Pitman Research Notes in Mathematics. Longman, Harlow, 1992.
[26] Oberguggenberger, M., Kunzinger, M., Characterization of Colombeau generalized functions
by their pointvalues, Math. Nachr. 203 (1999), 147 -- 157.
[27] Oberguggenberger, M., Vernaeve, H., Internal sets and internal functions in Colombeau the-
ory, J. Math. Anal. Appl. 341 (2008) 649 -- 659.
[28] Scarpal´ezos, D., Some remarks on functoriality of Colombeau's construction; topological and
microlocal aspects and applications, Int. Transf. Spec. Fct. 1998, Vol. 6, no. 1 -- 4, 295 -- 307.
[29] Scarpal´ezos, D., Colombeau's generalized functions: topological structures; microlocal prop-
erties. A simplified point of view, I. Bull. Cl. Sci. Math. Nat. Sci. Math. no. 25 (2000),
89 -- 114.
[30] Vernaeve, H., Ideals in the ring of Colombeau generalized numbers, Comm. Algebra (2010)
38(6):2199 -- 2228.
[31] Vernaeve, H. Nonstandard principles for generalized functions, J. Math. Anal. Appl., Volume
384, Issue 2, 2011, 536 -- 548.
University of Vienna, Austria
E-mail address: [email protected]
University of Vienna, Austria
E-mail address: [email protected]
|
1204.5654 | 4 | 1204 | 2012-06-13T14:18:25 | Oscillating convolution operators on the Heisenberg group | [
"math.FA"
] | In this paper, we consider oscillating convolution operotors on the Heisenberg group $H^n_a$ with respect to the norm $\rho(x,t) = \rho_1(b x, b t)$ with $\rho_1(x,t)= (|x|^4 + t^2)^{1/4}$. We obtain $L^2$ boundedness properties using the oscillatory integral estimates for degenerate phases in the Euclidean setting. Our result contains an improvement of the Lyall's result for the cases $\frac{a^2}{b^2} \geq C_{\beta}$. | math.FA | math |
OPTIMAL CONDITIONS FOR L2 BOUNDEDNESS OF STRONGLY
SINGULAR CONVOLUTION OPERATORS ON THE HEISENBERG GROUP
WOOCHEOL CHOI
Abstract. Strongly singular convolution operators TKα,β with the kernel Kα,β on the Heisen-
berg group Hn
a were introduced in [10]. For case a2 < Cβ , Laghi and Lyall [8] obtained the sharp
range for α, β for which the operator TKα,β are bounded on L2(Hn
a ). They used the classical
L2 −L2 boundedness theorem for oscillatory integral operators with non-degenerate phases. But,
if a2 ≥ Cβ , the phase function related with the operators are no more non-degenerate. However,
in this paper, we obtain the sharp range for α, β for the case a2 ≥ Cβ . To carry out this case, we
show that the canonical graph related with the phases satisfy folding type conditions and utilize
the recent developed theory for degenerate oscillatory integral operators (see [5], [11]).
Our setting is on the Heisenberg group Hn
a (a ∈ R∗) which have base manifold R2n+1 with the
group law
1. Introduction
where J is the 2n × 2n matrix
(x, t) · (y, s) = (x + y, s + t − 2axtJy)
0
−In
In
0!
(In is the n × n identity matrix). This group has the following dilation law
λ · (x, t) = (λx, λ2t).
For each kernel K, associated convolution operators are defined by
TKf (x, t) := K ∗ f (x, t) =ZHn
K(cid:0)(x, t) · (y, s)−1(cid:1) f (y, s)dydx.
We say that the operator TK is bounded on Lp(Hn) if there exist a C > 0 such that
kTKf kp ≤ Ckf kp,
for all f ∈ C∞
0 (Hn)
And a natural quasi-norm on the Heisenberg group is defined by ρ(x, t) = (x4 + t2)1/4. This
quasi-norm satisfies ρ(λ · (x, t)) = λρ(x, t). For this quasi-norm, we define the strongly singular
kernels for each α > 0
• Kα,β(x, t) = ρ(x, t)−(2n+2+α)eiρ(x,t)−β
χ(ρ(x, t)),
β > 0,
where χ is a smootho bump function in a small neighborhood of the origin. We let the operator
TKα,β be the convolution operators on the Heisenberg group Hn
a with the kernel Kα,β. This operator
was firstly introduced in [10] and there, the necessary condition is referred as α ≤ (n + 1
2 )β and
it was shown that TKα,β is bounded if α ≤ nβ. To show this, the group fourier transform was
utilized and a lengthy calculation for estimating some oscillatory integrals was needed to obtain the
2000 Mathematics Subject Classification. Primary.
1
2
WOOCHEOL CHOI
results. But, Laghi and Lyall [8] showed that we can get sharp results in the restricted case a2 < Cβ
for some Cβ > 0 only using the Hormander's L2 − L2 boundedness theorems for non-degenerate
oscillatory integral operators [1]. In this paper, we consider the cases a2 ≥ Cβ and obtain sharp
conditions using the recent theory for oscillatory integral operators with degenerate phases (See
section 2 for the details). The theory for the degenerate oscillatory integral operators have been
developed deeply, but the use of the theorem has been restricted to the X-ray transform.
Strongly singular convolution operators were studied originally in Rn. Such operators represent
some oscillating multipliers and operators of this type were first studied, using Fourier transform
techniques, in the Euclidean with ρ(x) = x by Hirschman [7] in the case d = 1 and then in higher
dimensions by Wainger [15], Fefferman [3], and Fefferman and Stein [4].
On the other hand, a similar kind of convolution operators with the kernel
with
α, β > 0 also have been studied in [12],[13]. This kernel has no sinularity near the zero, but it has
relatively small decaying property at infinity. Note that the case β = 1 corresponds to the kernel
of Bochner-Riesz means. For β 6= 1, the Lp − Lq estimates including the hardy space estimates
were well understood. The difference between two cases comes from the fact that the phase kernel
x − yβ is degenerate only if β = 1. We now consider an analogue problem on the Heisenberg
groups with the following kernels
1
xn−α eixβ
• Lα,β(x, t) = ρ(x, t)−(2n+2−α)eiρ(x,t)β
χ(ρ(x, t)−1),
β > 0.
We denote the gruop convolution operators TLα,β with the kernel Lα,β. In literatures, operators
like TKα,β are called as strongly singular operators and TLα,β as oscillating convolution operators.
In this paper, we shall find the optimal ranges of α and β where the convolution operators
associated with Kα,β ,Lα,β are bounded on L2(Hn).
For a2
b2 ≥ Cβ, the phases are no more non-degenerate. So, we need to deal with oscillatory inte-
gral operators with degenerate phases. Theory for this kind of operators have been studied largely
with considering various conditions on phase functions to give a different decaying properties, See
[6]. We will use results in [5],[11]. To utilize such theory, we should carefully investigate the folding
type for our phases. Interestingly, we have different folding types according to the values a, b and
β. Before stating our results, we recall the former results in [8], [10].
Theorems [8],[10].
(1) TKα,β is bounded on L2(Hn) if α ≤ nβ.
(2) If 0 < a2 < Cβ , then TKα,β is bounded on L2(Hn) if and only if α ≤ (n + 1/2)β. (Cβ =
β+2
2 (2β + 5 +p(2β + 5)2 − 9))
We obtain sharp results on the L2 → L2 boundedness of TKα,β for the cases a2 ≥ Cβ. Namely,
we obtain the following results:
Theorem 1.1. If a2 > Cβ , then TKα,β is bounded on L2(Hn
a2 = Cβ, then TKα,β is bounded on L2(Hn
a ) if and only if α ≤ (n + 1
4 )β.
a ) if and only if α ≤ (n + 1
3 )β. If
For the operators TLα,β , we also have the sharp ranges for the L2 → L2 boundedness except the
cases β = −1 or −2.
Theorem 1.2. TLα,β is bounded on L2 if and only if
(i) 0 < β < 1 : For a2 < Cβ, α ≤ (n + 1
2 )β, for a2 = Cβ, α ≤ (n + 1
4 )β, and for a2 > Cβ,
α ≤ (n + 1
3 )β.
OPTIMAL BOUNDS FOR OSCILLATING CONVOLUTION OPERATORS ON THE HEISENBERG GROUP
3
(ii) 1 < β < 2 : α ≤ (n + 1
(iii) 2 < β : α ≤ (n + 1
2 )β.
3 )β.
Remark 1.3. For the cases β = −1 or β = −2, we can also obtain the sharp results for some values
α, b where the phase becomes non-degenerate or of folding type 2. But, in these cases, higher order
folding type conditions than 3 appear for some α, b and the present theory for degenerate oscillatory
integral estimates does not cover these cases. The theory have been established optimally only for
the one or two folding cases ([5], [11]).
Remark 1.4. In Rn, the oscillating kernel is of the form x−γeixβ
In this case,
the different behavior for the phases xβ according to the value a is characterized only by the
with β 6= 0.
two cases where β 6= 1 or β = 1. Precisely, we have det(cid:16) ∂2
det(cid:16) ∂2
∂x∂y x − yβ(cid:17) 6= 0 for x 6= y, but
∂x∂y x − y(cid:17) = 0 for every x 6= y and this is the hardest case, which is correspond to the
Bochner-Riesz means operators. We see that for our cases, β = 1 and β = 2 present the lowest
degeneracy oscillatory operators (see 4.1) and it also hard to establish from the lack of the theory
of oscillatory integral estimates for lower folding type cases.
This paper is organized as follows. In section 2, we proceed an usual process to decompose the
kernel dyadically and reduce our problem to a local oscillatory integral estimates. In section 3, we
recall some essential things for our oscillatory integral operators with degenerate phase functions.
Finally, in section 4, we study some geometry of the canonical relation and projection maps related
with the phase functions. Then, we conclude the proof of the main theorems.
2. Dyadic decompostion and Localization
We use a standard argument to reduce our problems to some oscillatory integral estimates on
R2n+1. For euclidean space, the argument is well explained in [14]. Though the same mechanism
is easily adapted, we shall prove it because our setting is on a group.
We split the kernel Kα,β and Lα,β as
∞
Kα,β(x, t) =
Lα,β(x, t) =
η(2jρ(x, t))Kα,β(x, t)
(2.1)
η(2−jρ(x, t))Lα,β(x, t)
∞
Xj=0
Xj=0
0 (R) is a bump function supported in [1/2, 2]. and satisfiesP∞
where η ∈ C∞
r ≤ 1. We let K j
For notational convenience, we omit the index α and β from now.
We let Tjf = K j
α,β = η(2j(ρ(x, t)))Kα,β (x, t) and Tjf = Kj ∗ f and Lj
α,β ∗ f and Sjf = Lj
α,β ∗ f . Then,
j=0 η(2jr) = 1 for all 0 <
α,β = η(2−jρ(x, t))Lα,β (x, t).
Lemma 2.1. For each N ∈ N, there exists a constant CN such that
kT ∗
kS∗
j Tj ′ kL2→L2 + kTjT ∗
j Sj ′ kL2→L2 + kSjS∗
j ′ kL2→L2 ≤ CN 2−max{j,j ′}N
j ′ kL2→L2 ≤ CN 2−max{j,j ′}N
(2.2)
holds when j − j′ ≥ cβ for some sufficiently large constant cβ > 0.
Proof. The proof follows using the integration parts technique in usual ways. See [10] where the
proof for Tj is given.
(cid:3)
4
WOOCHEOL CHOI
By the Cotlar-Stein Lemma, we only need to show that
holds uniformly for j ∈ N with some constant C > 0.
kTjkL2→L2 + kSjkL2→L2 ≤ C
Let
K j
α,β(x, t) = K j
Lj
α,β(x, t) = Lj
α,β(2−j · (x, t)) = η(ρ(x, t))2j(Q+α)ρ(x, t)−Q−αei2jβ ρ(x,t)−β
α,β(2−j · (x, t)) = η(ρ(x, t))2−j(Q−α)ρ(x, t)−Q+αei2jβ ρ(x,t)β
Let fj(x, t) = f (2j · (x, t)). Then K j ∗ f (2−j · (x, t)) = 2−jQ( K j ∗ fj)(x, t).
kK j
α,β ∗ f (x, t)kL2 = 2−jQ/2kKα,β ∗ f (2−j · (x, t))kL2
≤ 2−jQ/2 · 2−jQk K j
≤ 2−jQ/2 · 2−jQk K j
= 2−jQk K j
α,βkL2→L2kf kL2.
α,β ∗ f (2j · ())(x, t)kL2
α,βkL2→L2kf (2−j·)kL2
,
.
(2.3)
(2.4)
(2.5)
(2.6)
Similarly, we also have kLj
α,β ∗ f kL2 ≤ 2jQk Lj
α,βkL2→L2kf kL2.
So, it suffice to prove suitable boundedness for Tj and Sj. Now, we further modify our operator
locally using the fact that the kernels of Tj and Sj are supported in {(x, t) : ρ(x, t) ≤ 2}. To do
a and each B(gk, 4)
contains only dn's other gl members in G.
this, we find a set of point G = {gk : k ∈ N} such that Sk∈N B(gk, 2) = Hn
We can split f =P∞
j f (x, t) =Z K j
α,β(cid:0)(x, t) · (y, s)−1(cid:1) · η(cid:0)ρ(cid:0)(x, t) · g−1
k=1 fk with each fk supported in B(gk, 2). And we let
T l,k
k (cid:1)(cid:1) η(cid:0)ρ(cid:0)(x, t) · g−1
l (cid:1)(cid:1) f (y, s)dyds
Then,
L2(B(gk,2))
flk2
L2(B(gk,2))
k Tj ∗ f k2
L2(Hn
a ) ≤
≤
=
.
.
∞
k Tj ∗ f k2
∞
Xl=1
∞
∞
k Tj ∗
Xk=1
Xk=1
Xk=1
Xk=1 Xl:ρ(gl·g−1
∞
sup
ρ(gl·g−1
k )≤2
k Tj ∗ X{l:ρ(gl·g−1
k )≤2}
flk2
L2((B(gk,2)))
k T l,k
j kL2→L2kflk2
L2
k )≤2
k T l,k
j kL2→L2kf k2
L2.
(2.7)
(2.8)
(2.9)
(2.10)
Because det (Dx,t ((x, t) · g)) = 1 for all g ∈ Hn
a ,
T l,k
j f ((x, t) · gk) =Z K j(cid:0)(x, t) · (y, s)−1(cid:1) η(ρ(x, t))η(ρ(y, s) · (gk · g−1
) . 1. If we let ψ ((x, t), (y, s)) = η(ρ(x, t))η(ρ((y, s)·(gk ·g−1
l
))f ((y, s) · gk)dyds.
))) and substitute
l
Notice that ρ(gk ·g−1
f as f (x) := f (x · gk).
l
OPTIMAL BOUNDS FOR OSCILLATING CONVOLUTION OPERATORS ON THE HEISENBERG GROUP
5
k )≤2 k T l,k
Pj,k,ρ(gl·g−1
j k will be obtained if we prove kAjkL2→L2 . 1 for
Aj f (x, t) =Z K j
α,β(cid:0)(x, t) · (y, s)−1(cid:1) ψ ((x, t), (y, s)) f (y, s)dyds
with a compactly supported smooth function ψ. Since
K j
α,β(x, t) = η(ρ(x, t))2j(Q+α)ρ(x, t)−Q−αei2jβ ρ(x,t)−β
α,β(x, t) = η(ρ(x, t))2−j(Q+α)ρ(x, t)−Q+αei2jβ ρ(x,t)β
Lj
The matters are reduced to show that kTAj kL2→L2 . 2jQ and kTBj kL2→L2 . 2−jQ Let
Aj(x, t) = 2jαµ(x, t)ei2jβ ρ(x,t)−β
Bj(x, t) = 2−jαµ(x, t)ei2jβ ρ(x,t)β
(2.11)
(2.12)
where with µ is a smooth funtion supported on ρ(x, t) ∼ 1. We define the operators LAj , LBj by
LAj f (x, t) =Z Aj(cid:0)(x, t) · (y, s)−1(cid:1) ψ ((x, t), (y, s)) f (y, s)dyds
LBj f (x, t) =Z Bj(cid:0)(x, t) · (y, s)−1(cid:1) ψ ((x, t), (y, s)) f (y, s)dyds.
To prove Theorem 1.1 and Theorem 1.2, it suffices to establish the following two theorems.
(2.13)
Theorem 2.2. The following inequalities hold uniformly for j.
If a2 > Cβ,
If a2 = Cβ,
kLAj kL2→L2 . 2j(α−(n+ 1
kLAj kL2→L2 . 2j(α−(n+ 1
3 )β),
4 )β).
Theorem 2.3. The following inequalities hold uniformly for j.
(i) − 1 < β < 0,
If a2 < Cβ,
If a2 = Cβ,
(ii) − 2 < β < −1,
(iii) β < −2,
kLBj kL2→L2 . 2j(α−(n+ 1
kLBj kL2→L2 . 2j(α−(n+ 1
2 )β),
4 )β).
kLBj kL2→L2 . 2j(α−(n+ 1
2 )β).
kLBj kL2→L2 . 2j(α−(n+ 1
3 )β).
In the next section, we shall breifly review on the theory related to the operators LAj and LBj .
And in section 4, we will expoit some required geometry related with the phase function ρ(x, t)β
to conclude the proof of Theorem 2.2 and Theorem 2.3.
3. L2 theory for oscillatory integral operators
In this section, we review on some L2 → L2 theory for oscillatory integral operators. The form
of operators we are concern is given as
T φ
λ f (x) =ZRn
eiλφ(x,y)a(x, y)f (y)dy
where φ, a ∈ C∞(Rn × Rn) and a has a compact support. We firstly state the fundamental theorem
of Hormander [1].
Theorem 3.1. Suppose that the phase function φ satisfies det(cid:16) ∂2φ
Then we have the following inequality
∂xi∂yj(cid:17) 6= 0 on the support of a.
where the implicit constant is independent of λ.
kT φ
λ kL2→L2 . λ− n
2 ,
λ ≥ 1
6
WOOCHEOL CHOI
We say that φ is non-degenerate if it satisfies the assumption of the above theorem. And we
equals to zero. This
say φ is degenerate if there is some point (x0, y0) where det(cid:16) ∂2
theorem gives the sharp decaying of kT φ
λ kL2→L2 in terms of λ. But, the phase functions of our
operators (see ) can become degenerate according to the values of a and β. For a degenerate φ,
the optimal number κφ for which the inequality kTλkL2→L2 . λ−κφ holds would be less than n
2 .
The number κφ's are related to the folding type conditions of the phase φ (See below). For phases
whose folding degrees are ≤ 3, the sharp numbers κφ were obtained in [5], [11]. We shall use the
results. In fact, the results for folding types ≤ 3 in [5] is the best known results and there are
no results for folding types > 3 except the very restricted result in [2]. Finding completely the
numbers κφ corresponding to all degenerate phases φ seems a widely open problem.
∂xi∂yj(cid:17)(cid:12)(cid:12)(cid:12)(x0,y0)
It is well-known that the decaying property is strongly related th the geometry of the canonical
relation
Cφ = {(x, φx(x, y), y, −φy(x, y)) ⊂ T ∗(Rn
x ) × T ∗(Rn
y )
(3.1)
To describe the geometry of Cφ, we need the following definition
Definition 3.2. Let M1,M2 be smooth manifolds of dimension n, and f : M1 → M2 is a smooth
map of corank ≤ 1. We let the singular set S = {P ∈ M1 : rank(Df ) < n at P }. Then we say that
f has a k− type fold at P0 for P0 ∈ S if
(1) rank(Df )P0 = n − 1,
(2) det(Df ) vanishes of k order in the null direction at P0.
Here, the null direction means the unique direction vector v such that (Dvf )P0 = 0.
Now we consider the two projection maps
πL : CΦ → T ∗(Rn
x ) and πR : CΦ → T ∗(Rn
y ).
(3.2)
Then the following theorem is proved in [11] for one folds cases and in [5] for two folds cases.
Theorem 3.3. Suppose that the projection maps πL and πR have one fold singularities, then
kTλf kL2(Rn) ≤ Cλ− (n−1)
2 − 1
3 kf kL2(Rn).
If the projection maps φL and πR have two hold singularities,
kTλf kL2(Rn) ≤ Cλ− (n−1)
2 − 1
4 kf kL2(Rn).
4. Geometry of the Canonical relation maps
In this section, we study the projection maps (3.2) associated with the phase function of our
operators in Theorem 2.2 and Theorem 2.3 to use Theorem 3.3. Recall that ρ(x, t) = (x4 + t2)1/4
and the phase function φ of the integral operators LAj and LBj is
To write the group law explicitly, we write x = (x1, x2) and y = (y1, y2) with x1, x2, y1, y2 ∈ Rn.
Let Φ(x, t) = ρ(x, t)−β for simplicity, then we have
φ(x, t, y, s) = ρ−β(cid:0)(x, t) · (y, s)−1(cid:1) .
φ(x, t, y, s) = Φ(cid:0)x1 − y1, x2 − y2, t − s − 2a(x1y2 − x2y1)(cid:1)
(4.1)
OPTIMAL BOUNDS FOR OSCILLATING CONVOLUTION OPERATORS ON THE HEISENBERG GROUP
7
Let us use the derivatives ∂x, ∂t to denote the position derivatives of Φ (i,e. ∂xj Φ = Φxj , ∂tΦ =
Φt and ∂xj (xjΦ) = xj Φxj ).
Firstly, to determine the phase function Φ is non-degenerate or not, we should calculate the
determinant of the matrix
So, we now compute the matrix H. In (4.1), by the chain-rule, we have
H =(cid:18) ∂2φ(x, t, y, s)
∂(y,s)∂(x,t) (cid:19) .
∂
∂xj
∂
∂xj+n
φ(x, t, y, s) = (∂xj + 2ayn+j∂t)Φ((x,t)·(y,s)−1)
(4.2)
φ(x, t, y, s) = (∂xj+n − 2ayj∂t)Φ((x,t)·(y,s)−1)
Let
A(y) = I
0
2aJy
1 !
Using a similar calculation, we see that
H(x, t, y, s) = A(x) (∂i∂jΦ) A(y)t + 2a∂tΦ J 0
0
= A(x)"(∂i∂jΦ) + 2a∂tΦ J 0
0! A(y)t = J
0
0
0
where the second equality holds because A(x) J
L(x, t, y, s) = "(∂i∂jΦ) + 2a∂tΦ J 0
0
A(y)t
0!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)((x,t)·(y,s)−1)
0!#(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)((x,t)·(y,s)−1)
0!#(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)((x,t)·(y,s)−1)
0!. We let
0
0
.
(4.3)
(4.4)
(4.5)
Thus, to study the matrix H, it is enough to analyze the matrix L. Morerover, we have det(A(x)) =
det(A(y)) = 1, which implies det(H(x, t, y, s)) = det(L(x, t, y, s)). In [8], the determinant of the
matrix H was calculated directly without using this observation. Here we calculate it by obtaining
the determinant of L.
We now calculate the hessian matrix of Φ. For 1 ≤ i, j ≤ 2n,
∂jΦ(x, t) = −
∂tΦ(x, t) = −
β
4
β
4
(x4 + t2)− β
4 −1(4xjx2)
(x4 + t2)− β
4 −1(2t)
(4.6)
and
∂i∂jΦ(x, t) = β(x4 + t2)− β
∂i∂tΦ(x, t) = β(β + 4)(x4 + t2)− β
4 −2x2xi ·
4 −2(cid:2)(β + 4)x4 − 2(x4 + t2)(cid:3) xixj − β(x4 + t2)− β
4 −1δij x2 (4.7)
(4.8)
t
2
t Φ(x, t) = β(β + 4)(x4 + t2)− β
∂2
4 −2 t
2
·
t
2
− β(x4 + t2)− β
4 −1 1
2
.
8
WOOCHEOL CHOI
Let D = (x2x, t
2 )t, then the above computations show that
(∂i∂jΦ) + 2a∂tΦ J 0
0
0!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(x,t)
= β(β + 4)(x4 + t2)− β
4 −2D · Dt− β(x4 + t2)− β
4 −1 x2I + atJ + 2x · xt
0
= −β(x4 + t2)− β
4 −1(E + R)
To write the matrices simply, we let B = x2I + atJ, K = x · xt and
E = B + 2K 0
2! and R = −
0
1
(β + 4)
x4 + t2 D · Dt.
(4.9)
0
1
2! (4.10)
(4.11)
Then, in (4.5) and (4.9), we have
L(x, t, y, s) = [−β(x4 + t2)− β
4 −1(E + R)](x,t)=(x,t)·(y,s)−1
(4.12)
Now we are ready to obtain the following Lemma
Lemma 4.1. We have
det H(x, t, y, s) = F ((x, t) · (y, s)−1)
where F (x, t) = ca,β(x4 + a2t2)m1 (x4 + t2)m2f (x, t) for some m1, m2, ca,β ∈ R and f (x, t) =
2(β + 1)x8 + (3(β + 2) − 2a2)x4t2 + (β + 2)a2t4.
Proof. From (4.3) and (4.12), it is enought to show that
det[−β(x4 + t2)− β
4 −1(E + R)] = F (x, t).
Moreover, from the form of given F in the theorem, we only need to compute det(E + R). We
recall the form
E + R = B + 2K 0
2! −
0
1
(β + 4)
x4 + t2 D · Dt.
For notational convenience, we always use case-letter fi to denote the i′th row of matrix capital
F . Notice that D · Dt is of rank 1 and will exploit the following convention
det(P + Q) = det(P ) +
det
m
Xj=1
(4.13)
p1
...
pj−1
qj
pj+1
...
pm
for any m × m matrix P and Q with rank Q = 1. Recall B = x2 + atJ and K = x · xt, then
direct calculations show that
det(B) = (x4 + a2t2)n
(4.14)
OPTIMAL BOUNDS FOR OSCILLATING CONVOLUTION OPERATORS ON THE HEISENBERG GROUP
9
and
=
n
Xj=1
xj det
n
Xj=1
b1
...
bj−1
kj
bj+1
...
b2n
+
n
Xj=1
xj+n det
(4.15)
xj(x2xj + xn+j at)(x4 + a2t2)n−1 +
xj+n (x2xj+n − xjat)(x4 + a2t2)n−1(4.16)
b1
...
bj+n−1
kj+n
bj+n+1
...
b2n
n
Xj=1
(4.17)
(4.18)
(4.19)
(4.20)
= (x4 + a2t2)n−1x4
Thus, from (4.4), (4.14) and (4.15), we have
det(B + 2K) = (x4 + a2t2)n + 2x4(x4 + a2t2)n−1
= (x4 + a2t2)n−1(3x4 + a2t2)
Using (4.4) again, we have
det(E + R) = det(E) +
1
2
2n
Xj=1
det
From (4.18)
=: S1 + S2 + S3
e1
...
ej−1
rj
ej+1
...
e2n
+ det
e1
...
e2n
r2n+1
S1 = det B + 2K 0
2! =
0
1
1
2
det(B + 2K) =
1
2
(x4 + a2t2)n−1(3x4 + a2t2)
Since rank K = 1, we see that
det
e1
...
ej−1
rj
ej+1
...
e2n
= det
b1 + 2k1
...
bj−1 + 2kj−1
−(β+4)x4
x4+t2 kj
bj+1 + 2kj+1
...
b2n + 2k2n
= −
(β + 4)x4
x4 + t2 det
b1
...
bj−1
kj
bj+1
...
b2n
.
So,
S2 = −
1
2(cid:18) (β + 4)x4
x4 + t2 (cid:19) x4(x4 + a2t2)n−1
10
Finally,
WOOCHEOL CHOI
S3 = det E + 2K
∗
0
− β+4
x4+t2
4! = −
t2
β + 4
x4 + t2
t2
4
= −
β + 4
x4 + t2
t2
4
(x4 + a2t2)n−1(3x4 + a2t2)
det(B + 2K)
(4.21)
Adding all these terms, we have
det (∂i∂jΦ)(x, t) + 2a∂tΦ(x, t) J 0
0!! = p(x4 + a2t2)q(x4 + t2)f (x, t)
0
where p(r) = cprm1 , q(r) = rm2 for some m1, m2, cp ∈ R and
f (x, t) = 2(β + 1)x8 + (3(β + 2) − 2a2)x4t2 + (β + 2)a2t4.
Now, we should determine when the determinant of H(x, t, y, s) can be zero for some values
(x, t, y, s) with ρ(cid:0)(x, t) · (y, s)−1(cid:1) ∼ 1. Furthermore, to determin the folding type in the degenerate
case, it will be decisive to know the shape of factorization. We take it in the following theorem
Lemma 4.2. For some nonzero constants γ, c, c1, c2, c3 with c1 6= c2 and c3 > 0 which are deter-
mined by β and a,
(cid:3)
• case 1 : β ∈ (−1, 0) ∪ (0, ∞)
· For a2
· For a2
· For a2
b2 < Cβ , f (x, t) > 0.
b2 = Cβ , f (x, t) = γ(x2 − ct2)2.
b2 > Cβ , f (x, t) = γ(x2 − c1t)(x2 + c1t)(x2 − c2t)(x2 + c2t).
• case 2 : β ∈ (−2, −1)
· f (x, t) = γ(x2 − c1t)(x2 + c1t)(x4 + c3t2).
• case 3 : β ∈ (∞, −2)
· f (x, t) < 0.
Proof. Firstly, we see that f (x, t) > 0 for 3(β + 2) − 2a2 > 0. And f (x, t) > 0 also holds if the
dicriminant
∆ = 4a4 − 4(β + 2)(2β + 5)a2 + 9(β + 2)2
is negative. This range is equal to
C−
β < a2 < C+
β
where
But,
C±
β =
β + 2
2
(cid:16)2β + 5 ±p(2β + 5)2 − 9(cid:17) .
C−
β =
(β + 2)
2
(2β + 5 −p(2β + 5)2 − 9) =
<
(β + 2)
2
(β + 2)
2
So, we compress two conditions as f (x, t) > 0 for a2 < C+
β .
(2β + 5 −p(2β + 2)(2β + 8)
(2β + 5 −p(2β + 2)2) =
2
3(β + 2)
(4.22)
(cid:3)
OPTIMAL BOUNDS FOR OSCILLATING CONVOLUTION OPERATORS ON THE HEISENBERG GROUP 11
For degenerate cases, we need to analyze the canonical relation (3.1) associated with our phase
function Φ
and the projection maps (3.2)
CΦ = {(cid:0)(x, t), Φ(x,t), (y, s), −Φ(y,s)(cid:1)} ⊂ T ∗(R2n+1) × T ∗(R2n+1)
πL : CΦ → T ∗(R2n+1) and πR : CΦ → T ∗(R2n+1).
We will check the condition (1) and (2) of Definition 3.2 to prove the following theorem
Theorem 4.3. On the hypersurface S,
(1) If β ∈ (−2, −1) or β ∈ (−1, 0) ∪ (0, ∞) and a2
b2 > Cβ, the projection maps πL and πR are
both of folding type 1.
(2) If β ∈ (−1, 0) ∪ (0, ∞) and a2
b2 = Cβ, the projection maps πL and πR are both of folding
type 2.
Let
We need the following lemma to show that rank of πL and πR drop by 1 simply.
S = {(x, t, y, s) : det H(x, t, y, s) = 0}.
Lemma 4.4. Let L1(x, t, y, s) be the first (2n) × (2n)matrix of L(x, t, y, s) and suppose that
(x, t, y, s) is contained in S. Then,
provided β 6= −4.
det L1(x, t, y, s) 6= 0
Proof. For simplicity, let (z, w) := (x, t) · (y, s)−1. Except the nonzero common facts, we only need
to check the determinant of
is nonzero for (z, w) 6= (0, 0). It can be calculated in the same way using (??) and (4.18) so that
M (z, w) =(cid:16)z2I + awJ + 2z · zt − (β + 4)
z4+w2 x · zt(cid:17)
det(M (z, w)) = (z4 + a2w2)n + (z4 + a2w2)n−1z4(cid:18)2 − (β + 4)
z4
=
(z4 + a2w2)n−1
z4 + w2
z4
z4 + w2(cid:19)
(cid:2)−(β + 1)z8 + (a2 + 3)z4w2 + a2w4(cid:3)
(4.23)
But, (z, w) is on S and satisfies
2(β + 1)z8 + (3(β + 2) − 2a2)z4w2 + (β + 2)a2w4 = 0
(4.24)
From above two equalites, we have
(3(β + 2) + 6)z4w2 + (β + 4)a2w4 = 0
So
det(M (z, w)) =
(z4 + a2w2)n−1
z4 + w2
w2
2
(β + 4)(cid:2)3z4 + a2w2(cid:3)
If w = 0, then z becomes zero in (4.24). Because (z, w) 6= (0, 0), w should be nonzero. Thus
det(M (z, w)) 6= 0.
(cid:3)
12
WOOCHEOL CHOI
proof of Theorem 4.3. We will only consider the case (1) in the theorem. The other case can be
derived similrary. It is only remained to show that at the hypersurface S, we have the folding type
conditions as in the theorem. Recall that
S = {(x, t, y, s) ∈ R(2(2n+1) F(cid:0)(x, t) · (y, s)−1(cid:1) = F (cid:0)x − y, s − t + 2axtJy(cid:1) = 0,
= {(x, t, y, s) ∈ R2(2n+1) f(cid:0)x − y, s − t + 2axtJy(cid:1) = 0,
ρ(cid:0)(x, t) · (y, s)−1(cid:1) ∼ 1}.
From Theorem 4.2, we have
ρ(cid:0)(x, t) · (y, s)−1(cid:1) ∼ 1}
(4.25)
f (x, t) = γ(x2 − c1t)(x2 + c1t)(x2 − c2t)(x2 + c2t).
We need to show that at each point P0 ∈ S, det(Df ) vanishes of 1 order in each null directions of
dπL and dπR at P0. Let P0 = (x0, t0, y0, s0) and we may assume that P0 is contained in
S1 =: {(x, t, y, s) ∈ R2(2n+1) x − y2 − c1(s − t + 2axtJy) = 0}.
We may identifiy the manifold CΦ = {(cid:0)(x, t), Φ(x,t), (y, s), −Φ(y,s)(cid:1)} with an open set in R(2n+1) ×
R(2n+1) by the diffeomorphsim φ : R(2n+1) × R(2n+1) → S given by
Let the null direction vL in R2(2n+1) of dπL at P0. It means that
φ(x, t, y, s) =(cid:0)(x, t), Φ(x,t), (y, s), −Φ(y,s)(cid:1) .
I
∂2Φ
∂(x,t)∂(x,t)
0
∂2Φ
∂(y,s)∂(x,t)! vt = 0
Thus, v is of the form v = (0, 0, z, w) with w ∈ R2n, s ∈ R and (w, s) satisfies
∂2Φ
∂(y,s)∂(x,t) zt
w! = 0
To verify det H(x, t, y, s) vanishes of order 1 in the direction vL, it suffice to show that vL is not
orthogonal to the gradient vector vg of det H(x, t, y, s) at P0. From a direct calculation, we get
the gradient vector vg as
D(x,t),(y,s)Φ(cid:0)(x, t) · (y, s)−1(cid:1)(cid:12)(cid:12)p =(cid:0)2(x − y) − 2acβ,1aJy, −cβ,1, −2(x − y) − 2acβ,1xtJ, cβ,1(cid:1)
Suppose that vL and vg are orthogonal. We are going to find a contradiction. It means that
−2(x − y) · z − 2acβ,1xtJ · z + cβ,1w = 0
From (4.3), we have
(
∂2
∂xi∂yj
Φ) = A(y)"(∂i∂jΦ) − 2a∂tρ J 0
0!# A(x)t · zt
w!
0
and
A(x)t · zt
w! =
0
. . .
0
0
0
1
0
0
· · · −2ax1
· · ·
· · ·
0
0
0
· · ·
. . .
· · · −2axn
1
0
0
...
0
1
z1
z2
...
z2n
w
1
0
0
0
2axn+1
(4.26)
=(cid:16)z1,
z2,
· · · ,
z2n, 2a(xn+1z1 + · · · + x2nzn − x1zn+1 − · · · − xnz2n) + w(cid:17)t
OPTIMAL BOUNDS FOR OSCILLATING CONVOLUTION OPERATORS ON THE HEISENBERG GROUP 13
On the other hand, from the orthogonal assumption, we have
It means that
so
Observe that
z ·(cid:0)−2(x − y) − 2cβ,1axtJ(cid:1) + w · cβ,1 = 0
2a(xn+1z1 + · · · − xnz2n) + w =
2(x − y) · z
cβ,1
A(x)t · zt
z2,
· · · ,
z2n,
2(x−y)·z
cβ,1 (cid:17)t
w! =(cid:16)z1,
x4x2
1
...
x4xnx1
x2 t
2 x1
"(∂i∂jΦ) − 2a∂tρ J 0
0!# = (β + 4)
0
· · ·
. . .
· · ·
· · ·
x4x1xn
...
x4x2
n
x2 t
2 xn
t
2
t
2
x2x1
...
x2xn
t2
4
− (x4 + t2) J
0
0
1
2! .
Here, x → x − y and t → t − s + 2axtJy = x−y2
cβ,1
following should hold
. So, if we calculate with the bottom row, the
(β + 4)"x − y2 ·
1
2
x − y2
cβ,1
Rearranging it, we obtain
(x − y) · z +
x − y4
c2
β,1
·
2
cβ,1
(x − y) · z# −
1
2
(x − y4 +
x − y4
c2
β,1
)
2
cβ,1
(x − y) · z = 0
" β + 2
2cβ,1
+
1
c3
β,1# x − y4 (x − y) · z = 0.
And Thus we have (x − y) · z = 0. So
A(x)t · z
and H1(x, t, y, s) ·(cid:16)(z1, z2, · · · , z2n)(cid:17)t
w! =(cid:16)z1, z2, · · · , z2n, 0(cid:17)t
We can also have the same conclusion for dπR with the same argument.
(cid:3)
= 0. But this is a contradiction because det H1 6= 0.
proof of Theorem 2.2 and Theorem 2.3. The theorems follow immediately when we use Theorem
3.3 based on the folding types established in Theorem 4.3.
(cid:3)
References
[1] L. Hormander, Oscillatory integrals and multipliers on F Lp, Ark. Math. 11, (1973), 1-11.
[2] Cuccagna, Scipio L2L2 estimates for averaging operators along curves with two-sided kk-fold singularities. Duke
Math. J. 89 (1997), no. 2, 203-216.
[3] C. Fefferman, Inequalities for strongly singular convolution operators, Acta Math., 124 (1970), pp. 9-36.
[4] C. Fefferman and E. M. Stein, H p spaces of several variables, Acta Math., 129 (1972), pp. 137-193.
[5] A. Greenleaf and A. Seeger, Oscillatory integral operators with low-order degeneracies. Duke Math. J. 112
(2002), no. 3, 397-420.
[6]
, Oscillatory and Fourier integral operators with degenerate canonical relations. Proceedings of the 6th
International Conference on Harmonic Analysis and Partial Differential Equations (El Escorial, 2000). Publ.
Mat. 2002, Vol. Extra, 93-141.
[7] I. I. Hirschman, Multiplier Transforms I, Duke Math. J., 26 (1956), pp. 222242.
[8] N. Laghi and N Lyall, Strongly singular integral operators associated to different quasi-norms on the Heisenberg
group. Math. Res. Lett. 14 (2007), 825-838.
14
WOOCHEOL CHOI
[9]
, Strongly singular Radon transforms on the Heisenberg group and folding singularities Pacific J. Math.
233 (2007), 403-415.
[10] N. Lyall, Strongly singular convolution operators on the Heisenberg group. Trans. Amer. Math. Soc. 359 (2007),
4467-4488 .
[11] Y. Pan and C.D. Sogge, Oscillatory integrals associated to folding canonical relations, Colloq. Math. 60/61,
(1990), 413-419.
[12] Y. Pan and G. Sampson, The complete (Lp, Lp) mapping properties for a class of oscillatory integrals. J. Fourier
Anal. Appl. 4 (1998), 93-103.
[13] P. Sjolin, Convolution with oscillating kernels. Indiana Univ. Math. J. 30 (1981), no. 1, 47-55.
[14] E. M. Stein, harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory integrals, Princeton
Unive. Press, Princeton, 1993.
[15] S. Wainger, Special trigonometric series in k-dimensions, Memoirs of the AMS 59, (1965), American Math.
Society.
School of Mathematical Sciences, Seoul National University, Seoul 151-747, Korea
E-mail address: [email protected]
|
1501.03267 | 2 | 1501 | 2016-04-21T07:15:44 | Operator Lipschitz functions on Banach spaces | [
"math.FA",
"math.OA"
] | Let $X$, $Y$ be Banach spaces and let $\mathcal{L}(X,Y)$ be the space of bounded linear operators from $X$ to $Y$. We develop the theory of double operator integrals on $\mathcal{L}(X,Y)$ and apply this theory to obtain commutator estimates of the form $\|f(B)S-Sf(A)\|_{\mathcal{L}(X,Y)}\leq \textrm{const} \|BS-SA\|_{\mathcal{L}(X,Y)}$ for a large class of functions $f$, where $A\in\mathcal{L}(X)$, $B\in \mathcal{L}(Y)$ are scalar type operators and $S\in \mathcal{L}(X,Y)$. In particular, we establish this estimate for $f(t):=|t|$ and for diagonalizable operators on $X=\ell_{p}$ and $Y=\ell_{q}$, for $p<q$ and $p=q=1$, and for $X=Y=\mathrm{c}_{0}$. We also obtain results for $p\geq q$. We also study the estimate above in the setting of Banach ideals in $\mathcal{L}(X,Y)$. The commutator estimates we derive hold for diagonalizable matrices with a constant independent of the size of the matrix. | math.FA | math |
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
ABSTRACT. Let X, Y be Banach spaces and let L(X, Y) be the space of bounded
linear operators from X to Y. We develop the theory of double operator integrals
on L(X, Y) and apply this theory to obtain commutator estimates of the form
k f (B)S − S f (A)kL(X,Y) ≤ const kBS − SAkL(X,Y)
for a large class of functions f , where A ∈ L(X), B ∈ L(Y) are scalar type opera-
tors and S ∈ L(X, Y). In particular, we establish this estimate for f (t) := t and
for diagonalizable operators on X = ℓ p and Y = ℓq, for p < q and p = q = 1, and
for X = Y = c0. We also obtain results for p ≥ q.
We study the estimate above in the setting of Banach ideals in L(X, Y). The
commutator estimates we derive hold for diagonalizable matrices with a constant
independent of the size of the matrix.
1. INTRODUCTION
Let X be a Banach space and let L(X) be the space of all bounded linear op-
erators on X. Let A, B ∈ L(X) be scalar type operators (see Definition 3.1 below)
: sp(A) ∪ sp(B) → C be a bounded Borel function, where sp(A)
on X. Let f
(resp. sp(B)) is the spectrum of the operator A (resp. B). We are interested in Lip-
schitz type estimates
(1.1)
where k · kL(X) is the uniform operator norm on the space L(X), and more gener-
ally in commutator estimates
k f (B) − f (A)kL(X) ≤ const kB − AkL(X) ,
k f (B)S − S f (A)kL(X,Y) ≤ const kBS − SAkL(X,Y)
(1.2)
for Banach spaces X and Y, scalar type operators A ∈ L(X) and B ∈ L(Y), and
S ∈ L(X, Y). This problem is well known in the special case where X = Y is a
separable Hilbert space, such as ℓ2, and A and B are normal operators on X. In
this paper we study such estimates in the Banach space setting, and specifically
for X = ℓp and Y = ℓq with p, q ∈ [1, ∞].
space H the estimate
In the special case where A, B are self-adjoint bounded operators on a Hilbert
k f (B) − f (A)kL(H) ≤ const kB − AkL(H)
(1.3)
was established by Peller [24, 26] (see also [10]) for f : R → R in the Besov class
B1
∞,1(R) (for the definition of B1
∞,1(R) see Section 3.3). This result extended a long
line of results from [6]- [7], in which the theory of double operator integration
Date: October 16, 2018.
2010 Mathematics Subject Classification. Primary 47A55, 47A56; secondary 47B47.
The first-named author is supported by NWO-grant 613.000.908 "Applications of Transference
Principles".
1
2
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
was developed to study the difference f (B) − f (A) (see also [8]). This theory was
revised and extended in various directions, including the Banach space setting, in
[11]. However, until now the results in the general setting were much weaker than
in the Hilbert space setting. In this paper we show that for scalar type operators
on Banach spaces one can obtain results matching those on Hilbert spaces.
∞,1(R).
In Corollary 4.9 below we prove that (1.1) holds when A, B ∈ L(X) are scalar
type operators with real spectrum and f ∈ B1
It is immediate from the
definition of a scalar type operator that every normal operator on H is of scalar
type. Therefore, Corollary 4.9 extends (1.3) to the Banach space setting. More
generally, (1.2) holds for f ∈ B1
∞,1(R) and for all S ∈ L(X, Y) (see Corollary 4.8.)
∞,1(R) and the results mentioned
above do not apply. Moreover, the techniques which we used to obtain (1.1) for
f ∈ B1
∞,1(R) cannot be applied to the absolute value function (see Remark 8.4).
However, the absolute value function is very important in the theory of matrix
analysis and perturbation theory (see [5, Sections VII.5 and X.2]).
If f is the absolute value function then f /∈ B1
In the case where H is an infinite-dimensional Hilbert space, it was proved by
Kato [17] that the function t 7→ t, t ∈ R does not satisfy (1.3). An earlier example
of McIntosh [22] showed the failure in general of the commutator estimate (1.2)
for this function, in the case X = Y = H. Later, it was proved by Davies [9] that
for 1 ≤ p ≤ ∞ and the Schatten von-Neumann ideal Sp with the corresponding
norm k · kS p, the estimate
kB − AkS p ≤ constkB − AkS p
holds for all A, B ∈ Sp if and only if 1 < p < ∞. Commutator estimates for
the absolute value function and different Banach ideals in L(H) have also been
studied in [14]. The proofs in [9, 11, 14] are based on Macaev's celebrated theorem
(see [16]) or on the UMD-property of the reflexive Schatten von-Neumann ideals.
However, the spaces L(X, Y) are not UMD-spaces, and therefore the techniques
used in [9, 11, 14] do not apply to them. To study (1.2) for X = ℓp and Y = ℓq,
we use completely different methods from those of [9, 11, 14]. Instead, we use the
theory of Schur multipliers on the space L(ℓp, ℓq) developed by Bennett [3], [4].
Let p, q ∈ [1, ∞] with p < q, p = q = 1 or p = q = ∞. In Section 6 we show
(see Theorem 6.7) that, for diagonalizable operators (for the definition see Section
5) A ∈ L(ℓp) and B ∈ L(ℓq), and for the absolute value function f ,
k f (B)S − S f (A)kL(ℓp,ℓq) ≤ const kBS − SAkL(ℓp,ℓq)
(1.4)
holds for all S ∈ L(ℓp, ℓq) (where ℓ∞ should be replaced by c0). In particular,
(1.5)
k f (B) − f (A)kL(ℓ1) ≤ const kB − AkL(ℓ1)
and
(1.6)
k f (B) − f (A)kL(c0) ≤ const kB − AkL(c0)
for diagonalizable operators on ℓ1 respectively c0. Therefore we show that, even
though (1.4) fails for p = q = 2, and in particular (1.1) fails for X = ℓ2 and f
the absolute value function, one can obtain commutator estimates for p < q and
Lipschitz estimates (1.1) for X = ℓ1 or X = c0.
We also obtain results for p ≥ q. In particular, for p = q = 2 we prove (see
Corollary 6.15) that for each ǫ ∈ (0, 1] there exists a constant C ≥ 0 such that
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
3
the following holds. Let A, B ∈ L(ℓ2) be compact self-adjoint operators, and let
U, V ∈ L(ℓ2) be unitaries such that
λjPj
U AU−1 =
VBV−1 =
∞
∑
j=1
µjPj,
∞
∑
j=1
and
j=1 and {µj}∞
where {λj}∞
j ∈ N, are the basis projections corresponding to the standard basis of ℓ2. Then
(1.7)
j=1 are sequences of real numbers and the Pj ∈ L(ℓ2), for
kB − AkL(ℓ2) ≤ C min(cid:16)kV(B − A)U−1kL(ℓ2,ℓ2−ǫ), kV(B − A)U−1kL(ℓ2+ǫ,ℓ2)(cid:17),
where we let the right-hand side equal infinity if V(B − A)U−1 /∈ L(ℓ2, ℓ2−ǫ) ∪
L(ℓ2+ǫ, ℓ2).
We note that the constants which appear in our results depend on the spectral
constants of A and B from Section 3.1, and in (1.4), (1.5) and (1.6) on the diagonal-
izability constants of A and B from (5.5). These quantities are independent of the
norms of A and B, and to obtain constants which do not depend on A and B in
any way one merely has to restrict to operators with a sufficiently bounded spec-
tral or diagonalizability constant. This is already done implicitly on Hilbert spaces
by considering normal operators, for which the quantities involved are equal to 1.
For example, in (1.7) the constant C does not depend on A or B in any way. Our
results therefore truly extend the known estimates on Hilbert spaces, the main dif-
ference between Hilbert spaces and general Banach spaces being that on Hilbert
spaces one has a large and easily identifiable class of operators with spectral con-
stant 1 or which are diagonalizable by an isometry.
We study the commutator estimate in (1.2) in the more general form
k f (B)S − S f (A)kI ≤ const kBS − SAkI ,
(1.8)
where I is an operator ideal in L(X, Y). For example, we prove in Corollary 4.8
that (1.3) holds for a general Banach ideal I in L(X) with the strong convex com-
pactness property (for definitions see Section 3.2), with respect to the norm k·kI .
We also present (see Theorem 7.3) an example of a Banach ideal (I, k · kI ) in
L(lp∗, ℓp), for 1 < p < ∞ and 1
p∗ = 1, namely the ideal of p-summing opera-
tors, such that any Lipschitz function f (in particular, the absolute value function)
satisfies (1.8).
p + 1
In the final section we apply our results to finite-dimensional spaces, and obtain
commutator estimates for diagonalizable matrices. Any diagonalizable matrix is
a scalar type operator, hence estimates (1.4)-(1.8) hold for diagonalizable matrices
A and B with a constant independent of the size of the matrix.
2. NOTATION AND TERMINOLOGY
All vector spaces are over the complex number field. Throughout, X and Y de-
note Banach spaces, the space of bounded linear operators from X to Y is L(X, Y),
and L(X) := L(X, X). We identify the algebraic tensor product X∗⊗ Y with the
space of finite rank operators in L(X, Y) via (x∗⊗ y)(x) := hx∗, xiy for x ∈ X,
x∗ ∈ X∗ and y ∈ Y. The spectrum of A ∈ L(X) is sp(A), and by IX ∈ L(X) we
denote the identity operator on X. Throughout the text we use the abbreviations
SOT and WOT for the strong and weak operator topology, respectively.
4
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
The Borel σ-algebra on a Borel measurable subset σ ⊆ C will be denoted by Bσ,
and B := BC. For measurable spaces (Ω1, Σ1) and (Ω2, Σ2) we denote by Σ1 ⊗ Σ2
the σ-algebra on Ω1 × Ω2 generated by all measurable rectangles σ1 × σ2 with
σ1 ∈ Σ1 and σ2 ∈ Σ2. If (Ω, Σ) is a measurable space then B(Ω, Σ) is the space of
all bounded Σ-measurable complex-valued functions on Ω, a Banach algebra with
the supremum norm
k fkB(Ω,Σ) := sup
ω∈Ω f (ω)
( f ∈ B(Ω, Σ)).
We simply write B(Ω) := B(Ω, Σ) and k fk∞ := k fkB(Ω,Σ) when little confusion
can arise.
If µ is a complex Borel measure on a measurable space (Ω, Σ) and X is a Banach
space, then a function f : Ω → X is µ-measurable if there exists a sequence of X-
valued simple functions converging to f µ-almost everywhere. For Banach spaces
: Ω → L(X, Y), we say that f is strongly measurable if
X and Y and a function f
ω 7→ f (ω)x is a µ-measurable mapping Ω → Y for each x ∈ X.
If µ is a positive measure on a measurable space (Ω, Σ) and f : Ω → [0, ∞] is a
function, we let
ZΩ
f (ω) dµ(ω) := infZΩ
g(ω) dµ(ω) ∈ [0, ∞],
The H older conjugate of p ∈ [1, ∞] is denoted by p∗ and is defined by 1
where the infimum is taken over all measurable g : Ω → [0, ∞] such that g(ω) ≥
f (ω) for ω ∈ Ω.
p∗ =
1. The indicator function of a subset σ of a set Ω is denoted by 1σ. We will often
identify functions defined on σ with their extensions to Ω by setting them equal to
zero off σ.
p + 1
3. PRELIMINARIES
3.1. Scalar type operators. In this section we summarize some of the basics of
scalar type operators, as taken from [15].
Let X be a Banach space. A spectral measure on X is a map E : B → L(X) such
that the following hold:
• E(∅) = 0 and E(C) = IX;
• E(σ1 ∩ σ2) = E(σ1)E(σ2) for all σ1, σ2 ∈ B;
• E(σ1 ∪ σ2) = E(σ1) + E(σ2) − E(σ1)E(σ2) for all σ1, σ2 ∈ B;
• E is σ-additive in the strong operator topology.
Note that these conditions imply that E is projection-valued. Moreover, by [15,
Corollary XV.2.4] there exists a constant K such that
(σ ∈ B).
kE(σ)kL(X) ≤ K
(3.1)
An operator A ∈ L(X) is a spectral operator if there exists a spectral measure E
on X such that AE(σ) = E(σ)A and sp(A, E(σ)X) ⊆ σ for all σ ∈ B, where
sp(A, E(σ)X) denotes the spectrum of A in the space E(σ)X. For a spectral opera-
tor A, we let ν(A) denote the minimal constant K occurring in (3.1) and call ν(A)
the spectral constant of A. This is well-defined since the spectral measure E asso-
ciated with A is unique, cf. [15, Corollary XV.3.8]. Moreover, E is supported on
sp(A) in the sense that E(sp(A)) = IX [15, Corollary XV.3.5]. Hence we can define
an integral with respect to E of bounded Borel measurable functions on sp(A), as
This definition is independent of the representation of f , and
follows. For f = ∑n
Borel for 1 ≤ j ≤ n, we let
(3.2)
(cid:13)(cid:13)(cid:13)(cid:13)Zsp(A)
f dE(cid:13)(cid:13)(cid:13)(cid:13)L(X)
αjE(σj).
=
n
∑
j=1
f dE :=
Zsp(A)
kxkX=kx∗kX∗ =1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
sup
j
sup
αj
≤ sup
≤ 4 k fkB(sp(A))
≤ 4ν(A)k fkB(sp(A)) ,
n
∑
j=1
αjx∗E(σj)x(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
kxkX=kx∗kX∗ =1kx∗E(·)xkvar
sup
kxkX=kx∗kX∗ =1
sup
σ⊆sp(A)x∗E(σ)x
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
5
j=1 αj1σj a finite simple function with αj ∈ C and σj ⊆ sp(A)
where kx∗E(·)xkvar is the variation norm of the measure x∗E(·)x. Since the simple
functions lie dense in B(sp(A)), for general f ∈ B(sp(A)) we can define
Zsp(A)
f dE := lim
n→∞Zsp(A)
fn dE ∈ L(X)
if { fn}∞
n=1 ⊆ B(sp(A)) is a sequence of simple functions with k fn − fk∞ → 0 as
n → ∞. This definition is independent of the choice of approximating sequence
and
(3.3)
It is straightforward to check that
(cid:13)(cid:13)(cid:13)(cid:13)Zsp(A)
f dE(cid:13)(cid:13)(cid:13)(cid:13)L(X) ≤ 4ν(A)k fkB(sp(A)) .
Zsp(A)
(α f + g) dE = αZsp(A)
f dE +Zsp(A)
f g dE =(cid:18)Zsp(A)
f dE(cid:19)(cid:18)Zsp(A)
g dE(cid:19)
Zsp(A)
g dE,
for all α ∈ C and simple f , g ∈ B(sp(A)), and approximation extends these iden-
tities to general f , g ∈ B(sp(A)). Moreover,Rsp(A) 1 dE = E(sp(A)) = IX. Hence
the map f 7→ Rsp(A) f dE is a continuous morphism B(sp(A)) → L(X) of unital
is bounded on sp(A) andRsp(A) λ dE(λ) ∈ L(X) is well defined.
Banach algebras. Since the spectrum of A is compact, the identity function λ 7→ λ
Definition 3.1. A spectral operator A ∈ L(X) with spectral measure E is a scalar
type operator if
A =Zsp(A)
λ dE(λ).
The class of scalar type operators on X is denoted by Ls(X).
For A ∈ Ls(X) with spectral measure E and f ∈ B(sp(A)) we define
(3.4)
f (A) :=Zsp(A)
f dE.
6
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
Then, as remarked above, f 7→ f (A) is a continuous morphism B(sp(A)) → L(X)
of unital Banach algebras with norm bounded by 4ν(A). Note also that
(3.5)
hx∗, f (A)xi =Zsp(A)
f (λ) dhx∗, E(λ)xi
for all f ∈ B(sp(A)), x ∈ X and x∗ ∈ X∗. Indeed, for simple functions this follows
from (3.2), and by taking limits one obtains (3.5) for general f ∈ B(sp(A)).
Finally, we note that a normal operator A on a Hilbert space H is a scalar type
operator with ν(A) = 1, and in this case (3.3) improves to
(3.6)
(cid:13)(cid:13)(cid:13)(cid:13)Zsp(A)
f dE(cid:13)(cid:13)(cid:13)(cid:13)L(H) ≤ k fkB(sp(A)) ,
as known from the Borel functional calculus for normal operators.
3.2. Spaces of operators. In this section we discuss some properties of spaces of
operators that we will need later on.
< ǫ.
First we provide a lemma about approximation by finite rank operators. Recall
that a Banach space X has the bounded approximation property if there exists M ≥ 1
such that, for each K ⊆ X compact and ǫ > 0, there exists S ∈ X∗⊗ X with
kSkL(X) ≤ M and supx∈K kSx − xkX
Lemma 3.2. Let X and Y be Banach spaces such that X is separable and either X or Y has
the bounded approximation property. Then each T ∈ L(X, Y) is the SOT-limit of a norm
bounded sequence of finite rank operators.
Proof. Fix T ∈ L(X, Y). By [21, Proposition 1.e.14] there exists a norm bounded
net(cid:8)Tj(cid:9)j∈J ⊆ X∗⊗ Y having T as its SOT-limit. It is straightfroward to see that the
strong operator topology is metrizable on bounded subsets of L(X, Y) by
d(S1, S2) :=
2−k kS1xk − S2xkkY
(S1, S2 ∈ L(X, Y)),
∞
∑
k=1
(cid:3)
where {xk}k∈N ⊆ X is a countable subset that is dense in the unit ball of X. Hence
there exists a subsequence of(cid:8)Tj(cid:9)j∈J with T as its SOT-limit.
Let X and Y be Banach spaces and let Z be a Banach space which is continu-
ously embedded in L(X, Y). Following [36] (in the case where Z is a subspace
of L(X, Y)), we say that Z has the strong convex compactness property if the follow-
ing holds. For any finite measure space (Ω, Σ, µ) and any strongly measurable
bounded f : Ω → Z, the operator T ∈ L(X, Y) defined by
(x ∈ X),
(3.7)
f (ω)x dµ(ω)
belongs to Z with kTkZ ≤ RΩk f (ω)kZ dµ(ω). By the Pettis Measurability Theo-
rem, any separable Z has this property. Indeed, if Z is separable then combining
Propositions 1.9 and 1.10 in [35] shows that any strongly measurable f : Ω → Z is
µ-measurable as a map to Z. If f is bounded as well, then (3.7) defines an element
of Z with
Tx :=ZΩ
kTkZ ≤ZΩ k f (ω)kZ dµ(ω).
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
7
It is shown in [36] and [31] that the compact and weakly compact operators have
the strong convex compactness property, but not all subspaces of L(X, Y) do.
Moreover, if N is a semifinite von Neumann algebra on a separable Hilbert space
H, with faithful normal semifinite trace τ, and F is a rearrangement invariant Ba-
nach function space with the Fatou property, then E = N ∩ F (N, τ) has the strong
convex compactness property [2, Lemma 3.5].
Now let (Ω, µ) be a finite measure space and let f
Lemma 3.3. Let X and Y be separable Banach spaces and Z a Banach space continuously
embedded in L(X, Y). If BZ := {z ∈ Z kzkZ ≤ 1} is SOT-closed in L(X, Y), then Z
has the strong convex compactness property.
Proof. The proof follows that of [2, Lemma 3.5]. First we show that BZ is a Polish
space in the strong operator topology. As in the proof of Lemma 3.2, bounded
subsets of L(X, Y) are SOT-metrizable. The finite rank operators are SOT-dense
in L(X, Y), hence L(X, Y) is SOT-separable. Therefore BZ is SOT-separable and
metrizable. By assumption, BZ is complete.
: Ω → Z be bounded and
strongly measurable. Without loss of generality, we may assume that f (Ω) ⊆ BZ
and that µ is a probability measure. For each y∗ ∈ Y∗ and x ∈ X, the mapping
BZ → [0, ∞), T 7→ hy∗, Txi is continuous. The collection of all these mappings,
for y∗ ∈ Y∗ and x ∈ X, separates the points of BZ. Moreover, ω 7→ hy∗, f (ω)xi
is a measurable mapping Ω → [0, ∞) for each y∗ ∈ Y∗ and x ∈ X. By [35, Propo-
sitions 1.9 and 1.10], f is the µ-almost everywhere SOT-limit of a sequence of BZ-
k=1. Let Tn := RΩ fn dµ ∈ BZ. By the dominated
valued simple functions { fk}∞
convergence theorem, Tn(x) → T(x) :=RΩ f (ω)x dµ(ω) as n → ∞, for all x ∈ X.
By assumption, T ∈ BZ.
Now let g : Ω → [0, ∞) be measurable such that 1 ≥ g(ω) ≥ k f (ω)kZ for
ω ∈ Ω, and define h(ω) := f (ω)
RΩ g(η) dµ(η)dµ(ω) for ω ∈ Ω.
By what we have shown above, x 7→ RΩ h(ω)x dν(ω) defines an element of BZ.
Since Tx =RΩ f (ω)x dµ(ω) =RΩ g(ω) dµ(ω)RΩ h(ω)x dν(ω), we obtain kTkZ ≤
RΩ g(ω) dµ(ω), as remained to be shown.
Remark 3.4. Note that the converse implication does not hold. Indeed, if X is a
Hilbert space (or more generally, a Banach space with the metric approximation
property) then the finite rank operators of norm less than or equal to 1 are SOT-
dense in the unit ball of L(X). Therefore the compact operators of norm less than
or equal to 1 are not SOT-closed in L(X) if X is infinite-dimensional. However,
by [36, Theorem 1.3], the space of compact operators on X has the strong convex
compactness property.
(cid:3)
g(ω) and dν(ω) :=
g(ω)
embedded in L(X, Y). We say that (I, k · kI ) is a Banach ideal in L(X, Y) if
Let X and Y be Banach spaces and I a Banach space which is continuously
• For all R ∈ L(Y), S ∈ I and T ∈ L(X), RST ∈ I with kRSTkI ≤
• X∗⊗ Y ⊆ I with kx∗⊗ ykI = kx∗kX∗kykY for all x∗ ∈ X∗ and y ∈ Y.
kRkL(Y) kSkI kTkL(X);
By Lemma 3.3 and [12, Proposition 17.21] (using that the SOT and WOT closures
of a convex set coincide), for separable X and Y, any maximal Banach ideal (for the
definition see e.g. [27]) in L(X, Y) has the strong convex compactness property.
This includes a large class of operator ideals, such as the ideal of absolutely p-
summing operators, the ideal of integral operators, etc. [12, p. 203].
8
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
3.3. Algebras of functions. In this section we discuss some algebras of functions
that will be essential in later sections.
Let σ1, σ2 ⊆ C be Borel measurable subsets and let A(σ1 × σ2) be the class of
Borel functions ϕ : σ1 × σ2 → C such that
(3.8)
ϕ(λ1, λ2) =ZΩ
a1(λ1, ω)a2(λ2, ω) dµ(ω)
for all (λ1, λ2) ∈ σ1 × σ2, where (Ω, Σ, µ) is a finite measure space and a1 ∈ B(σ1 ×
Ω, Bσ1 ⊗ Σ), a2 ∈ B(σ2 × Ω, Bσ2 ⊗ Σ). For ϕ ∈ A(σ1 × σ2) let
kϕkA(σ1×σ2) := infZΩ ka1(·, ω)kB(σ1) ka2(·, ω)kB(σ2) dµ(ω),
where the infimum runs over all possible representations1 in (3.8) (it is straightfor-
ward to show that the map ω 7→ ka1(·, ω)kB(σ1) ka2(·, ω)kB(σ2) is measurable).
Lemma 3.5. For all σ1, σ2 ⊆ C measurable, A(σ1 × σ2) is a unital Banach algebra which
is contractively included in B(σ1 × σ2).
Proof. That A(σ1 × σ2) is a vector space is straightforward, and that it is normed
algebra is shown in [28, Lemma 3] for σ1 = σ2 = R (the proof in our setting is
identical). The completeness of A(σ1 × σ2) follows by showing that an absolutely
convergent series of elements in A(σ1 × σ2) converges in A(σ1 × σ2). This is done
by considering a direct sum of the measure spaces involved.
(cid:3)
We now state sufficient conditions for a function to belong to A. The first will be
used in the proof of Proposition 5.6. Let W1,2(R) be the space of all g ∈ L2(R) with
weak derivative g′ ∈ L2(R), endowed with the norm kgkW1,2(R) := kgkL2(R) +
kg′kL2(R) for g ∈ W1,2(R).
Lemma 3.6.
(3.9)
[28, Theorem 9] Let g ∈ W1,2(R) and let
ψg(λ1, λ2) :=( g(log( λ1
λ2
0
))
if λ1, λ2 > 0
otherwise
.
Then ψg ∈ A(R2) with(cid:13)(cid:13)ψg(cid:13)(cid:13)A(R2) ≤ √2kgkW1,2(R).
The second condition involves the Besov space B1
∞,1(R). Following [25], let
{ψk}k∈Z be a sequence of Schwartz functions on R such that, for each k ∈ Z,
the Fourier transform F ψk of ψk is supported on [2k−1, 2k+1] and F ψk+1(x) =
F ψk(2x) for all x > 0, and such that ∑∞
k=−∞ F ψk(x) = 1 for all x > 0. Let ψ∗k be
defined by F ψ∗k (x) = F ψk(−x) for k ∈ Z and x ∈ R. If f is a distribution on R
such that {2k k f ∗ ψkkL∞(R)}k∈Z ∈ ℓ1(Z) and {2k(cid:13)(cid:13) f ∗ ψ∗k(cid:13)(cid:13)L∞(R)}k∈Z ∈ ℓ1(Z), then
f admits a representation
(3.10)
f = ∑
k∈Z
f ∗ ψk + ∑
k∈Z
f ∗ ψ∗k + P,
where P is a polynomial.
1This might seem problematic from a set-theoretic viewpoint. The problem can be fixed by choosing
an equivalence class of such a representation for each r ∈ R which can occur in the infimum.
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
9
We let the homogeneous Besov space B1
∞,1(R) consist of all distributions as above
for which P = 0. Then B1
∞,1(R) is a Banach space when equipped with the norm
∞,1(R) :=
k fk B1
For f ∈ B1
∞
∑
k=−∞
2k k f ∗ ψkkL∞(R) +
∞,1(R) define ψ f : C2 → C by
ψ f (λ1, λ2) :=( f (λ2)− f (λ1)
f ′(λ1)
λ2−λ1
∞
∑
k=−∞
2k k f ∗ ψ∗kkL∞(R)
( f ∈ B1
∞,1(R)).
if (λ1, λ2) ∈ R2 and λ1 6= λ2
if λ1 = λ2 ∈ R
.
Lemma 3.7. There exists a constant C ≥ 0 such that ψ f ∈ A(R2) for each f ∈ B1
∞,1(R),
with(cid:13)(cid:13)(cid:13)ψ f(cid:13)(cid:13)(cid:13)A(R2) ≤ C k fk B1
Proof. Let f ∈ B1
has a representation
∞,1(R).
∞,1(R). In [25, Theorem 2] (see also [26, p. 535]) it is shown that ψ f
ψ f (λ1, λ2) =ZΩ
a1(λ1, ω)a2(λ2, ω) dµ(ω)
for (λ1, λ2) ∈ R2, where (Ω, µ) is a measure space and a1 and a2 are measurable
functions on R × Ω such that
ZΩ ka1(·, ω)k∞ ka2(·, ω)k∞ dµ(ω) ≤ C k fk B1
∞,1(R)
for some constant C ≥ 0 independent of f . The desired conclusion now follows by
replacing a1(λ1, ω) and a2(λ2, ω) by a1(λ1,ω)
, and dµ(ω)
ka1(·,ω)k∞
by ka1(·, ω)k∞ ka2(·, ω)k∞ dµ(ω).
(cid:3)
respectively a2(λ2,ω)
ka2(·,ω)k∞
In [25, Theorem 3] Peller also states a condition on a function f on R which
is necessary in order for ϕ f ∈ A(R2) to hold, and this condition is only slightly
weaker than f ∈ B1
∞,1(R).
4. DOUBLE OPERATOR INTEGRALS AND LIPSCHITZ ESTIMATES
4.1. Double operator integrals. Fix Banach spaces X and Y, scalar type opera-
tors A ∈ Ls(X) and B ∈ Ls(Y) with spectral measures E respectively F, and
ϕ ∈ A(sp(A) × sp(B)). Let a representation (3.8) for ϕ be given, with corre-
sponding (Ω, µ) and a1 ∈ B(sp(A) × Ω), a2 ∈ B(sp(B) × Ω). For ω ∈ Ω, let
a1(A, ω) := a1(·, ω)(A) ∈ L(X) and a2(B, ω) := a2(·, ω)(B) ∈ L(Y) be defined
by the functional calculus for A respectively B from Section 3.1.
Lemma 4.1. Let S ∈ L(X, Y) have separable range. Then, for each x ∈ X, ω 7→
a2(B, ω)Sa1(A, ω)x is a weakly measurable map Ω → Y.
Proof. Fix x ∈ X. If a1 = 1σ for some σ ⊆ sp(A) × Ω then it is straightforward
to show that hx∗, a1(A, ·)xi is measurable for each x∗ ∈ X∗. As S has separable
range, Sa1(A, ·)x is µ-measurable. If a2 is an indicator function as well, the same
argument shows that a2(B, ·)y is weakly measurable for each y ∈ Y. General argu-
ments, approximating Sa1(A, ·)x by simple functions, show that a2(B, ·)Sa1(A, ·)x
is weakly measurable. By linearity this extends to simple a1 and a2, and for gen-
eral a1 and a2 let { fk}k∈N, {gk}k∈N be sequences of simple functions such that
10
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
a1 = limk→∞ fk and a2 = limk→∞ gk uniformly. Then a1(A, ω) = limk→∞ fk(A)
and a2(B, ω) = limk→∞ gk(B) in the operator norm, for each ω ∈ Ω. The desired
measurability now follows.
(cid:3)
Now suppose that Y is separable, that I is a Banach ideal in L(X, Y) and let
S ∈ L(X, Y). By (3.3),
(4.1)
ka2(B, ω)Sa1(A, ω)kI ≤ 16 ν(A)ν(B)kSkI ka1(·, ω)kB(sp(A)) ka2(·, ω)kB(sp(B))
for w ∈ Ω. Since I is continuously embedded in L(X, Y), by the Pettis Measura-
bility Theorem, Lemma 4.1 and (4.1) we can define the double operator integral
(4.2)
T A,B
ϕ
(S)x :=ZΩ
a2(B, ω)Sa1(A, ω)x dµ(ω) ∈ Y
(x ∈ X).
Throughout, we will use Tϕ for T A,B
the context.
ϕ when the operators A and B are clear from
Proposition 4.2. Let X and Y be separable Banach spaces such that X or Y has the
bounded approximation property, and let A ∈ Ls(X), B ∈ Ls(Y), and ϕ ∈ A(sp(A) ×
sp(B)). Let I be a Banach ideal in L(X, Y) with the strong convex compactness prop-
erty. Then (4.2) defines an operator T A,B
ϕ ∈ L(I) which is independent of the choice of
representation of ϕ in (3.8), with
(4.3)
(cid:13)(cid:13)(cid:13)T A,B
ϕ (cid:13)(cid:13)(cid:13)L(I ) ≤ 16 ν(A)ν(B)kϕkA(sp(A)×sp(B)) .
Proof. By (4.1) and the strong convex compactness property, Tϕ(S) ∈ L(I) for all
S ∈ I, with
(cid:13)(cid:13)Tϕ(S)(cid:13)(cid:13)I ≤ 16 ν(A)ν(B)kSkIZΩ ka1(·, ω)kB(sp(A)) ka2(·, ω)kB(sp(B)) dµ(ω).
Clearly Tϕ is linear, hence the result follows if we establish that Tϕ is independent
of the representation of ϕ. For this it suffices to let ϕ ≡ 0. Now, first consider
S = x∗⊗ y for x∗ ∈ X∗ and y ∈ Y, and let x ∈ X, y∗ ∈ Y∗ and w ∈ Ω. Recall that E
and F are the spectral measures of A and B, respectively. Then
hy∗, a2(B, ω)Sa1(A, ω)xi =Zsp(B)
=Zsp(B)
=Zsp(B)Zsp(A)
a1(λ, ω)a2(η, ω)dhx∗, E(λ)xidhy∗, F(η)yi
by (3.5). Now Fubini's Theorem and the assumption on ϕ yield Tϕ(S) = 0. By
linearity, Tϕ(S) = 0 for all S ∈ X∗ ⊗ Y. By Lemma 3.2, a general S ∈ I is the
SOT-limit of a bounded (in L(X, Y)) sequence {Sn}n∈N ⊆ X∗ ⊗ Y. The dominated
convergence theorem shows that Tϕ(S)x = limn→∞ Tϕ(Sn)x = 0 for all x ∈ X,
which implies that Tϕ is independent of the representation of ϕ and concludes the
proof.
(cid:3)
a2(η, ω) dhy∗, F(η)Sa1(A, ω)xi
a2(η, ω)hx∗, a1(A, ω)xi dhy∗, F(η)yi
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
11
Note that, if A and B are normal operators on separable Hilbert spaces X and
Y, then (4.3) improves to
(4.4)
(cid:13)(cid:13)(cid:13)T A,B
ϕ (cid:13)(cid:13)(cid:13)L(I ) ≤ kϕkA(sp(A)×sp(B)) ,
ϕ
by appealing to (3.6) instead of (3.3) in (4.1).
Remark 4.3. Let H be an infinite-dimensional separable Hilbert space and S2 the
ideal of Hilbert-Schmidt operators on H. There is a natural definition (see [8])
of a double operator integral T A,B
∈ L(S2) for all ϕ ∈ B(C2) and normal op-
erators A, B ∈ L(H), such that T A,B
if ϕ ∈ A(sp(A) × sp(B)). One
could wonder whether Proposition 4.2 can be extended to a larger class of func-
tions than A(sp(A) × sp(B)) using an extension of the definition of T A,B
in (4.2)
which coincides with T A,B
on S2. But it follows from [24, Theorem 1] that T A,B
extends to a bounded operator on I = L(H) if and only if ϕ ∈ A(sp(A) ×
sp(B)). Hence Proposition 4.2 cannot be extended to a larger function class than
A(sp(A) × sp(B)) in general. However, for specific Banach ideals, e.g. ideals
with the UMD property, results have been obtained for larger classes of func-
tions [11, 29].
= T A,B
ϕ
ϕ
ϕ
ϕ
ϕ
Remark 4.4. The assumption in Proposition 4.2 that X or Y has the bounded ap-
proximation property is only used, via Lemma 3.2, to ensure that each S ∈ I is
the SOT-limit of a bounded (in L(X, Y)) net of finite-rank operators. Clearly this
is true for general Banach spaces X and Y if I is the closure in L(X, Y) of X∗⊗ Y.
In [20] the authors consider an assumption on X and I, called condition c∗λ, which
guarantees that each S ∈ I is the SOT-limit of a bounded net of finite-rank oper-
ators. It is shown in [20] that this condition is strictly weaker than the bounded
approximation property, for certain non-trivial ideals. In the results throughout
the paper where we assume that X has the bounded approximation property, one
may assume instead that X has satisfies condition c∗λ for I for some λ ≥ 1.
4.2. Commutator and Lipschitz estimates. Let p1 :, p2 : C2 → C be the coordinate
projections p1(λ1, λ2) := λ1, p2(λ1, λ2) := λ2 for (λ1, λ2) ∈ C2. Note that f ◦
p1, f ◦ p2 ∈ A(σ1 × σ2) for all σ1, σ2 ⊆ C Borel and f ∈ B(σ1 ∪ σ2). For A and B
selfadjoint operators on a Hilbert space and I a non-commutative Lp-space, the
following lemma is [28, Lemma 3].
Lemma 4.5. Under the assumptions of Proposition 4.2, the following hold:
ϕ
algebras.
(1) The map ϕ 7→ T A,B
is a morphism A(sp(A) × sp(B)) → L(I) of unital Banach
(2) Let f ∈ B(sp(A) ∪ sp(B)) and S ∈ L(X, Y). Then Tf ◦p1(S) = S f (A) and
Tf ◦p2(S) = f (B)S. In particular, Tp1(S) = SA and Tp2(S) = BS.
Proof. The structure of the proof is the same as that of [28, Lemma 3]. Linearity
in (1) is straightforward. Fix ϕ1, ϕ2 ∈ A(sp(A) × sp(B)) with representations as
in (3.8), with corresponding measure spaces (Ωj, µj) and bounded Borel functions
a1,j ∈ B(sp(A) × Ωj) and a2,j ∈ B(sp(B) × Ωj) for j ∈ {1, 2}. Then ϕ := ϕ1 ϕ2 also
has a representation as in (3.8), with Ω = Ω1 × Ω2, µ = µ1 × µ2 the product
measure and a1 = a1,1a1,2, a2 = a2,1a2,2. By multiplicativity of the functional
12
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
calculus for A,
a1(A, (ω1, ω2)) =(cid:0)a1,1(·, ω1)a1,2(·, ω2)(cid:1)(A) = a1,1(A, ω1)a1,2(A, ω2)
for all (ω1, ω2) ∈ Ω, and similarly for a2(B, (ω1, ω2)). Applying this to (4.2) yields
Tϕ(S)x =ZΩ
=ZΩ1
= Tϕ1 (Tϕ2(S))x
a2(B, ω)Sa1(A, ω)x dµ(ω)
a2,1(B, ω1)Tϕ2 (S)a1,1(A, ω1)x dµ1(ω1)
for all S ∈ I and x ∈ X, which proves (1). Part (2) follows from (4.2) and the fact
that Tϕ is independent of the representation of ϕ.
(cid:3)
For f : sp(A) ∪ sp(B) → C define
(4.5)
ϕ f (λ1, λ2) :=
f (λ2) − f (λ1)
λ2 − λ1
for (λ1, λ2) ∈ sp(A) × sp(B) with λ1 6= λ2.
Theorem 4.6. Let X and Y be separable Banach spaces such that X or Y has the bounded
approximation property, and let I be a Banach ideal in L(X, Y) with the strong convex
compactness property. Let A ∈ Ls(X) and B ∈ Ls(Y), and let f ∈ B(sp(A) ∪ sp(B))
be such that ϕ f extends to an element of A(sp(A) × sp(B)). Then
(4.6)
for all S ∈ L(X, Y) such that BS − SA ∈ I.
In particular, if X = Y and B − A ∈ I,
k f (B)S − S f (A)kI ≤ 16 ν(A)ν(B)(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A(sp(A)×sp(B)) kBS − SAkI
k f (B) − f (A)kI ≤ 16 ν(A)ν(B)(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A(sp(A)×sp(B)) kB − AkI .
Proof. Note that (p2 − p1)ϕ f = f ◦ p2 − f ◦ p1. By Lemma 4.5,
f (B)S − S f (A) = Tf ◦p2(S) − Tf ◦p1(S) = T(p2−p1)ϕ f
(S) = Tp2 ϕ f (S) − Tp1 ϕ f (S)
= Tϕ f (Tp2(S) − Tp1 (S)) = Tϕ f (BS − SA)
for each S ∈ I. Proposition 4.2 now concludes the proof.
(cid:3)
Letting X and Y be Hilbert spaces and A and B normal operators, we gener-
alize results from [8, 28] to all Banach ideals with the strong convex compactness
property. As mentioned in Section 3.2, this includes all separable ideals and the
so-called maximal operator ideals, which in turn is a large class of ideals contain-
ing the absolutely (p, q)-summing operators, the integral operators, and more [12,
p. 203]. Note that, for normal operators, we can improve the estimate in (4.6) by
appealing to (4.4) instead of (4.3).
Corollary 4.7. Let A ∈ L(X) and B ∈ L(Y) be normal operators on separable Hilbert
spaces X and Y. Let I be a Banach ideal in L(X, Y) with the strong convex compact-
ness property, and let f ∈ B(sp(A) ∪ sp(B)) be such that ϕ f extends to an element of
A(sp(A) × sp(B)). Then
k f (B)S − S f (A)kI ≤(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A(sp(A)×sp(B)) kBS − SAkI
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
13
for all S ∈ L(X, Y) such that BS − SA ∈ I. In particular, if X = Y and B − A ∈ I,
k f (B) − f (A)kI ≤(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A(sp(A)×sp(B)) kB − AkI .
Combining Theorem 4.6 with Lemma 3.7 yields the following, a generalization
of [25, Theorem 4].
Corollary 4.8. There exists a universal constant C ≥ 0 such that the following holds.
Let X and Y be separable Banach spaces such that X or Y has the bounded approximation
property, and let I be a Banach ideal in L(X, Y) with the strong convex compactness
property. Let f ∈ B1
∞,1(R), and let A ∈ Ls(X) and B ∈ Ls(Y) be such that sp(A) ∪
sp(B) ⊆ R. Then
(4.7)
k f (B)S − S f (A)kI ≤ Cν(A)ν(B)k fk B1
∞,1(R)kBS − SAkI
for all S ∈ L(X, Y) such that BS − SA ∈ I. In particular, if X = Y and B − A ∈ I,
k f (B) − f (A)kI ≤ Cν(A)ν(B)k fk B1
∞,1(R)kB − AkI .
In the case where the Banach ideal I is the space L(X, Y) of all bounded opera-
tors from X to Y, we obtain the following corollary.
Corollary 4.9. There exists a universal constant C ≥ 0 such that the following holds. Let
X and Y be separable Banach spaces such that either X or Y has the bounded approximation
property. Let f ∈ B1
∞,1(R), and let A, B ∈ Ls(X) be such that sp(A) ∪ sp(B) ⊆ R.
Then
(4.8)
k f (B)S − S f (A)kL(X,Y) ≤ Cν(A)ν(B)k fk B1
∞,1(R)kBS − SAkL(X,Y)
for all S ∈ L(X, Y). In particular, if X = Y then
k f (B) − f (A)kL(X) ≤ Cν(A)ν(B)k fk B1
∞,1(R)kB − AkL(X) .
Remark 4.10. Corollaries 4.8 and 4.9 yield global estimates, in the sense that (4.7)
and (4.8) hold for all scalar type operators A and B with real spectrum, and the
constant in the estimate depends on A and B only through their spectral constants
ν(A) and ν(B). Local estimates follow if f ∈ B(R) is contained in the Besov class
locally. More precisely, given scalar type operators A ∈ Ls(X) and B ∈ Ls(Y)
with real spectrum, suppose there exists g ∈ B1
∞,1(R) with g(s) = f (s) for all
s ∈ sp(A) ∪ sp(B). Then (with notation as in Corollary 4.8)
(4.9)
k f (B)S − S f (A)kI ≤ Cν(A)ν(B) kgk B1
∞,1(R)kBS − SAkI
for all S ∈ L(X, Y) such that BS − SA ∈ I. This follows directly from Theorem
4.6.
5. SPACES WITH AN UNCONDITIONAL BASIS
In this section we prove some results for specific scalar type operators, namely
operators which are diagonalizable with respect to an unconditional Schauder ba-
sis. These results will be used in later sections. In this section we assume all spaces
to be infinite-dimensional, but the results and proofs carry over directly to finite-
dimensional spaces. This will be used in Section 8.1.
14
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
5.1. Diagonalizable operators. Let X be a Banach space with an unconditional
Schauder basis {ej}∞
j=1 ⊆ X. For j ∈ N, let Pj ∈ L(X) be the projection given by
Pj(x) := xjej for all x = ∑∞
Assumption 5.1. For simplicity, we assume in this section that(cid:13)(cid:13)(cid:13)∑j∈N Pj(cid:13)(cid:13)(cid:13)L(X)
= 1
for all non-empty N ⊆ N. This condition is satisfied in the examples we consider in
later sections, and simplifies the presentation. For general bases one merely gets additional
constants in the results.
k=1 xkek ∈ X.
An operator A ∈ L(X) is diagonalizable (with respect to {ej}∞
j=1) if there exists
j=1 ∈ ℓ∞ of complex numbers such that
U ∈ L(X) invertible and a sequence {λj}∞
λjPjx
U AU−1x =
(5.1)
∞
∑
j=1
(x ∈ X),
where the series converges since {ek}∞
this case A is a scalar type operator, with point spectrum equal to {λj}∞
(cid:8)λj(cid:9)∞
k=1 is unconditional, cf. [32, Lemma 16.1]. In
j=1, sp(A) =
j=1 and spectral measure E given by
E(σ) = ∑
λj∈σ
U−1PjU
(5.2)
for σ ⊆ C Borel. The set of all diagonalizable operators on X is denoted by Ld(X).
We do not explicitly mention the basis {ej}∞
j=1 with respect to which an operator
is diagonalizable, since this basis will be fixed throughout. Often we write A ∈
j=1, U) in order to identify the operator U and the sequence {λj}∞
Ld(X, {λj}∞
from above. For A ∈ Ld(X, {λj}∞
j=1, U) and f ∈ B(C), it follows from (3.4) that
j=1
(5.3)
f (A) = U−1(cid:16) ∞
∑
j=1
f (λj)Pj(cid:17)U.
Since any Banach space with a Schauder basis is separable and has the bounded
approximation property, we can apply the results from the previous section to
diagonalizable operators, and we obtain for instance the following.
Corollary 5.2. There exists a universal constant C ≥ 0 such that the following holds.
Let X and Y be Banach spaces with unconditional Schauder bases, and let I be a Banach
ideal in L(X, Y) with the strong convex compactness property. Let f ∈ B1
∞,1(R), and let
A ∈ Ld(X) and B ∈ Ld(Y) be such that sp(A) ∪ sp(B) ⊆ R. Then
k f (B)S − S f (A)kI ≤ Cν(A)ν(B)k fk B1
∞,1(R)kBS − SAkI
for all S ∈ L(X, Y) such that BS − SA ∈ I. In particular, if X = Y and B − A ∈ I,
k f (B) − f (A)kI ≤ Cν(A)ν(B)k fk B1
∞,1(R)kB − AkI .
Since this result does not apply to the absolute value function (and neither does
the more general Theorem 4.6), and because of the importance of the absolute
value function, we now study Lipschitz estimates for more general functions.
Let Y be a Banach space with an unconditional Schauder basis { fk}∞
let the projections Qk ∈ L(Y) be given by Qk(y) := yk fk for all y = ∑∞
k=1 ⊆ Y, and
l=1 yl fl ∈ Y
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
15
and k ∈ N. Let λ = {λj}∞
j=1 and µ = {µk}∞
and let ϕ : C2 → C. For n ∈ N, define Tλ,µ
ϕ,n ∈ L(L(X, Y)) by
k=1 be sequences of complex numbers,
(5.4)
Tλ,µ
ϕ,n (S) :=
ϕ f (λj, µk)QkSPj
(S ∈ L(X, Y)).
n
∑
j,k=1
Note that Tλ,µ
ϕ,n ∈ L(I) for each Banach ideal I in L(X, Y).
Let f ∈ B(C) and extend the divided difference from (4.5), given by
ϕ f (λ1, λ2) :=
f (λ2) − f (λ1)
λ2 − λ1
for (λ1, λ2) ∈ C2 with λ1 6= λ2, to a function ϕ f : C2 → C.
Lemma 5.3. Let X and Y be Banach spaces with unconditional Schauder bases, and let
I be a Banach ideal in L(X, Y). Let λ = {λj}∞
k=1 be sequences of
complex numbers, and let A ∈ Ld(X, λ, U), B ∈ Ld(Y, µ, V), f ∈ B(C) and n ∈ N.
Then
j=1 and µ = {µk}∞
k f (B)Sn − Sn f (A)kI ≤ kUkL(X)kV−1kL(Y)(cid:13)(cid:13)(cid:13)Tλ,µ
ϕ f ,n(V(BS − SA)U−1)(cid:13)(cid:13)(cid:13)I
j,k=1 V−1QkVSU−1PjU.
for all S ∈ I such that BS − SA ∈ I, where Sn := ∑n
Proof. Let S ∈ I be such that BS − SA ∈ I. For the duration of the proof write
Pj := U−1PjU ∈ L(X) and Qk := V−1QkV ∈ L(Y) for j, k ∈ N. By (5.3), and
using that PjPk = 0 and QjQk = 0 for j 6= k,
∑
i,l=1
f (B)Sn − Sn f (A) =
∑
i,l=1
∞
∑
j=1
f (µk)Qk(cid:16) n
QlSPi(cid:17) −
f (λj)(cid:16) n
QlSPi(cid:17)Pj
∞
∑
k=1
n
∑
j,k=1
n
∑
j,k=1
n
∑
j,k=1
n
∑
j,k=1
=
=
=
=
( f (µk) − f (λj))QkSPj
f (µk) − f (λj)
∑
µk6=λj
µk − λj
(µkQkSPj − λjQkSPj)
ϕ f (λj, µk)Qk(cid:16)(cid:16) ∞
∑
l=1
µl Ql(cid:17)S − S(cid:16) ∞
∑
i=1
λi Pi(cid:17)(cid:17)Pj
ϕ f (λj, µk)Qk(BS − SA)Pj
(V(BS − SA)U−1)U.
= V−1T A,B
ϕ f
where we have used that BS − SA ∈ Mn. Now use the ideal property of I to
conclude the proof.
(cid:3)
(5.5)
For a sequence λ of complex numbers and A ∈ Ld(X, λ, U ), define
KA := infnkUkL(X)kU−1kL(X)(cid:12)(cid:12)(cid:12) A ∈ Ld(X, λ, U)o .
We will call KA the diagonalizability constant of A. Using the unconditionality of the
Schauder basis of X and Assumption 5.1, one can show that KA does not depend
on the specific ordering of the sequence λ. Since the sequence λ is, up to ordering,
16
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
uniquely determined by A (it is the point spectrum of A), KA only depends on A.
Moreover, by Assumption 5.1 and (5.2), kE(σ)kL(X) ≤(cid:13)(cid:13)U−1(cid:13)(cid:13)L(X) kUkL(X) for all
σ ⊆ C Borel and U ∈ L(X) such that A ∈ Ld(X, λ, U), where E is the spectral
measure of A. Hence
(5.6)
ν(A) ≤ KA,
where ν(A) is the spectral constant of A from Section 3.1.
We now derive commutator estimates for A and B in the operator norm, under
a boundedness assumption which will be verified for specific X and Y in later
sections.
Proposition 5.4. Let X and Y be Banach spaces with unconditional Schauder bases, A ∈
Ld(X, λ, U), B ∈ Ld(Y, µ, V) and f ∈ B(C). Suppose that
< ∞.
(5.7)
C := sup
Then
n∈N(cid:13)(cid:13)(cid:13)Tλ,µ
ϕ f ,n(cid:13)(cid:13)(cid:13)L(L(X,Y))
k f (B)S − S f (A)kL(X,Y) ≤ CKAKB kBS − SAkL(X,Y)
for all S ∈ L(X, Y).
Proof. Let S ∈ L(X, Y) and let Sn ∈ L(X, Y) be as in Lemma 5.3 for n ∈ N. It
is straightforward to show that, for each x ∈ X, Snx → Sx as n → ∞. Hence
f (B)Snx − Sn f (A)x → f (B)Sx − S f (A)x as n → ∞, for each x ∈ X. Lemma 5.3
and (5.7) now yield
k f (B)S − S f (A)kL(X,Y) ≤ lim sup
n→∞ k f (B)Sn − Sn f (A)kL(X,Y)
≤ CkUkkV−1kkV(BS − SA)U−1kL(X,Y)
≤ CkUkkU−1kkVkkV−1k kBS − SAkL(X,Y).
Taking the infimum over U and V concludes the proof.
(cid:3)
Remark 5.5. Proposition 5.4 also holds for more general Banach ideals in L(X, Y).
Indeed, let I be a Banach ideal in L(X, Y) with the property that, if {Sm}∞
m=1 ⊆ I
is an I-bounded sequence which SOT-converges to some S ∈ L(X, Y) as m → ∞,
then S ∈ I with kSkI ≤ lim supm→∞ kSmkI . With notation as in Proposition 5.4,
if
C := sup
n∈N(cid:13)(cid:13)(cid:13)Tλ,µ
ϕ f ,n(cid:13)(cid:13)(cid:13)L(I )
< ∞
then
k f (B)S − S f (A)kI ≤ CKAKB kBS − SAkI
for all S ∈ L(X, Y) such that BS − SA ∈ I. This follows directly from the proof of
Proposition 5.4.
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
17
5.2. Estimates for the absolute value function. It is known that Lipschitz esti-
mates for the absolute value function are related to estimates for so-called trian-
gular truncation operators. For example, in [18] and [14] it was shown that the
boundedness of the standard triangular truncation on many operator spaces is
equivalent to Lipschitz estimates for the absolute value function. We prove that,
in our setting, triangular truncation operators are also related to Lipschitz esti-
mates for f the absolute value function. We do so by relating the assumption in
(5.7) to so-called triangular truncation operators associated to sequences. We will
then bound the norms of these operators in later sections for specific X and Y.
Let λ = {λj}∞
j=1 and µ = {µk}∞
k=1 be sequences of real numbers, and let X and
Y be as before. For n ∈ N define Tλ,µ
∆,n ∈ L(L(X, Y)) by
∑
µk≤λj
(S ∈ L(X, Y)).
Tλ,µ
∆,n (S) :=
n
∑
j,k=1
QkSPj
(5.8)
We call T A,B
∆
the triangular truncation associated to λ and µ.
For f (t) := t for t ∈ R, define ϕ f : C2 → C by
ϕ f (λ1, λ2) :=( λ1−λ2
λ1−λ2
1
(5.9)
if λ1 6= λ2
otherwise
.
The following result relates the norm of Tλ,µ
ϕ f ,n to that of Tλ,µ
∆,n .
Proposition 5.6. There exists a universal constant C ≥ 0 such that the following holds.
Let X and Y be Banach spaces with unconditional Schauder bases and let I be a Banach
ideal in L(X, Y) with the strong convex compactness property. Let λ and µ be bounded
sequences of real numbers. Let f (t) := t for t ∈ R. Then
(cid:13)(cid:13)(cid:13)Tλ,µ
ϕ f ,n(S)(cid:13)(cid:13)(cid:13)I ≤ C(cid:16)kSkI + kTλ,µ
∆,n (S)kI(cid:17)
for all n ∈ N and S ∈ I. In particular, if supn∈N kTλ,µ
∆,n (S)kL(L(X,Y))
holds.
Proof. Let n ∈ N and S ∈ I, and write λ = {λj}∞
k=1. Throughout
the proof we will only consider λj and µk for 1 ≤ j, k ≤ n, but to simplify the
presentation we will not mention this explicitly from here on. We can decompose
Tλ,µ
ϕ f ,n(S) as
j=1 and µ = {µk}∞
< ∞ then (5.7)
Tλ,µ
ϕ f ,n(S) = ∑
λk,µk≥0 QkSPj − ∑
µk
<0<λj
µk + λj
∑
λj <0<µk
µk − λj QkSPj − ∑
µk + λj
µk − λj QkSPj+
λk,µk≤0QkSPj + ∑
λk,µk=0QkSPj.
Note that some of these terms may be zero. By the ideal property of I and As-
sumption 5.1,
(5.10)
(cid:13)(cid:13)(cid:13) ∑
λj,µk≥0QkSPj(cid:13)(cid:13)(cid:13)I ≤(cid:13)(cid:13)(cid:13) ∑
µk≥0 Qk(cid:13)(cid:13)(cid:13)L(Y) kSkI(cid:13)(cid:13)(cid:13) ∑
λj≥0Pj(cid:13)(cid:13)(cid:13)L(X) ≤ kSkI .
are each bounded by kSkI .
18
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
To bound the other terms it is sufficient to show that
Similarly,(cid:13)(cid:13)(cid:13)∑λk,µk≤0 QkSPj(cid:13)(cid:13)(cid:13)I
and(cid:13)(cid:13)(cid:13)∑λk,µk=0 QkSPj(cid:13)(cid:13)(cid:13)I
µk + λj QkSPj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)I
≤ C′(cid:16) kSkI +(cid:13)(cid:13)(cid:13)Tλ,µ
µk − λj
∆,n (S)(cid:13)(cid:13)(cid:13)I(cid:17)
∑
λj,µk
>0
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
for some universal constant C′ ≥ 0. Indeed, replacing λ by −λ and µ by −µ then
yields the desired conclusion. Let
Φ(S) := ∑
λj,µk
>0
µk − λj
µk + λj QkSPj,
and define g ∈ W1,2(R) by g(t) := 2
et+1
for t ∈ R. Note that Φ(S) is equal to
∑
0<µk≤λj(cid:16)g(cid:16)log
λj
µk(cid:17) − 1(cid:17) QkSPj + ∑
0<λj<µk(cid:16)1 − g(cid:16)log
λj
µk(cid:17)(cid:17) QkSPj.
Now let ψg : C2 → C be as in (3.9), and let A := ∑∞
k=1 µkQk ∈ L(Y). Let T A,B
∑∞
ψg be as in (4.2). One can check that
(Tλ,µ
Φ(S) =T A,B
ψg
∆,n (S)Pj+
∆,n (S)) − ∑
>0Qk(S − Tλ,µ
∑
λj,µk
>0QkTλ,µ
∆,n (S))Pj − T A,B
λj,µk
ψg
(S − Tλ,µ
∆,n (S)).
j=1 λjPj ∈ L(X) and B :=
Any Banach space with a Schauder basis is separable and has the bounded ap-
proximation property, hence Lemma 3.6 and Proposition 4.2 yield
By (5.6), ν(A) = ν(B) = 1. Similarly,
.
ψg
ψg
(cid:13)(cid:13)(cid:13)T A,B
(cid:13)(cid:13)(cid:13)T A,B
(Tλ,µ
(S − Tλ,µ
∆,n (S))(cid:13)(cid:13)(cid:13)I ≤ 16√2 ν(A)ν(B)kgkW1,2(R)(cid:13)(cid:13)(cid:13)Tλ,µ
∆,n (S))(cid:13)(cid:13)(cid:13)I
∆,n (S))(cid:13)(cid:13)(cid:13)I ≤ 16√2 kgkW1,2(R)(cid:16)(cid:13)(cid:13)(cid:13)S(cid:13)(cid:13)(cid:13)I
+(cid:13)(cid:13)(cid:13)Tλ,µ
∆,n (S))(cid:13)(cid:13)(cid:13)I(cid:17).
+(cid:13)(cid:13)(cid:13) ∑
∆,n (S)(cid:13)(cid:13)(cid:13)I
+(cid:13)(cid:13)(cid:13)Tλ,µ
∆,n (S))Pj(cid:13)(cid:13)(cid:13) ≤ 2(cid:13)(cid:13)(cid:13)S(cid:13)(cid:13)(cid:13)I
+(cid:13)(cid:13)(cid:13)Tλ,µ
∆,n (S)(cid:13)(cid:13)(cid:13)I(cid:17),
∆,n (S)Pj(cid:13)(cid:13)(cid:13)I
kΦ(S)kI ≤(cid:16)2 + 32√2 kgkW1,2(R)(cid:17)(cid:16)(cid:13)(cid:13)(cid:13)S(cid:13)(cid:13)(cid:13)I
>0Qk(S − Tλ,µ
λj,µk
>0 QkTλ,µ
.
(cid:3)
(cid:13)(cid:13)(cid:13) ∑
λj,µk
as desired.
By the same arguments as in (5.10),
Combining all these estimates yields
6. THE ABSOLUTE VALUE FUNCTION ON L(ℓp, ℓq)
In this section we study the absolute value function on the space L(ℓp, ℓq). We
show that the absolute value function is operator Lipschitz on L(ℓp, ℓq) for p <
q < ∞, on L(ℓ1) and on L(c0). We also obtain results for p ≥ q.
The key idea of the proof is entirely different from the techniques used in [9],
[11], [14] and [18], which are based on a special geometric property of the reflexive
Schatten von Neumann ideals (the UMD-property), a property which L(ℓp, ℓq)
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
19
does not have. Instead, we prove our results by relating estimates for the operators
from (5.8) to the standard triangular truncation operator, defined in (6.1) below.
For this we use the theory of Schur multipliers on L(ℓp, ℓq) developed in [4]. We
then appeal to results from [3] about the boundedness of the standard triangular
truncation on L(ℓp, ℓq).
6.1. Schur multipliers. For p ∈ [1, ∞), let {ej}∞
j=1 be the standard Schauder basis
of ℓp, with the corresponding projections Pj(x) := xjej for x = ∑∞
k=1 xkek and
j ∈ N. We consider this basis and the corresponding projections on all ℓp-spaces
simultaneously, for simplicity of notation. Note that Assumption 5.1 is satisfied
for this basis. For q ∈ [1, ∞], any operator S ∈ L(ℓp, ℓq) can be represented by an
infinite matrix S = {sjk}∞
j,k=1, where sjk := (S(ek), ej) for j, k ∈ N. For an infinite
j,k=1 the product M ∗ S := {mjksjk} is the Schur product of the
matrix M = {mjk}∞
matrices M and S. The matrix M is a Schur multiplier if the mapping S 7→ M ∗ S is
a bounded operator on L(ℓp, ℓq). Throughout, we identify Schur multipliers with
their corresponding operators.
The notion of a Schur multiplier is a discrete version of a double operator in-
tegral (for details see e.g. [30, 33]). Schur multipliers on the space L(ℓp, ℓq) are
also called (p, q)-multipliers. We denote by M(p, q) the Banach space of (p, q)-
multipliers with the norm
kMk(p,q) := supn(cid:13)(cid:13)M ∗ S(cid:13)(cid:13)L(ℓp,ℓq)(cid:12)(cid:12)(cid:12)kSkL(ℓp,ℓq) ≤ 1o .
Remark 6.1. We will also consider (p, q)-multipliers M for p = ∞ and q ∈ [1, ∞].
Any operator S ∈ L(c0, ℓq) corresponds to an infinite matrix S = {sjk}∞
j,k=1, and M
is said to be a (∞, q)-multiplier if the mapping S 7→ M ∗ S is a bounded operator
on L(c0, ℓq), and we define the Banach space M(∞, q) in the obvious way. Below
we will often not explicitly distinguish the case p = ∞ from 1 ≤ p < ∞, in order
to keep the presentation simple. The reader should keep in mind that for p = ∞
the space ℓp should be replaced by c0.
Remark 6.2. It is straightforward to see that kMk(p,q) ≥ supj,k∈Nmj,k for all p, q ∈
[1, ∞] and M ∈ M(p, q).
For p, q ∈ [1, ∞] and S ∈ L(ℓp, ℓq), define
T∆(S) := ∑
(6.1)
k≤j PkSPj,
which is a well-defined element of L(ℓr, ℓs) for suitable r, s ∈ [1, ∞], by Proposition
6.3 below. The operator T∆ is the (standard) triangular truncation. This operator can
be identified with the following Schur multiplier. Let T′∆ = {t′jk}∞
j,k=1 be a matrix
given by t′jk = 1 for k ≤ j and t′jk = 0 otherwise. It is clear that T∆ extends to a
bounded linear operator on L(ℓp, ℓq) if and only if T′∆ is a (p, q)-multiplier. For n ∈
N and r, s ∈ [1, ∞] we will consider the operators T∆,n ∈ L(L(ℓp, ℓq), L(ℓr, ℓs)),
given by
T∆,n(S) := ∑
1≤k≤j≤nPkSPj
(S ∈ L(ℓp, ℓq).
The dependence of the (p, q)-norm of T∆ on the indices p and q was determined
in [3] and [19] (see also [34]), and is as follows.
20
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
Proposition 6.3. Let p, q ∈ [1, ∞]. Then the following statements hold.
(i) [3, Theorem 5.1] If p < q, 1 = p = q or p = q = ∞, then T∆ ∈ M(p, q).
(ii)
[19, Proposition 1.2] If 1 6= p ≥ q 6= ∞, then there is a constant C > 0 such
that
kT∆,nkL(L(ℓp,ℓq)) ≥ C ln n
for all n ∈ N.
(iii) [3, Theorem 5.2] If 1 6= p ≥ q 6= ∞, then for each s > q and r < p,
and T∆ : L(ℓp, ℓq) → L(ℓr, ℓq)
T∆ : L(ℓp, ℓq) → L(ℓp, ℓs)
are bounded.
bounded mapping L(ℓp, ℓq) → L(ℓr, ℓs), with
j,k=1, let eM = {emjk}∞
We will also need the following result, a generalization of [4, Theorem 4.1]. For
j,k=1 be obtained from M by repeating the
a matrix M = {mjk}∞
first column, i.e., emj1 = mj1 and emjk = mj(k−1) for j ∈ N and k ≥ 2.
Proposition 6.4. Let p, q, r, s ∈ [1, ∞] with r ≤ p. Let M = {mjk}∞
S 7→ M ∗ S is a bounded mapping L(ℓp, ℓq) → L(ℓr, ℓs). Then S 7→ eM ∗ S is also a
In particular, if M ∈ M(p, q) then eM ∈ M(p, q) with kMk(p,q) =(cid:13)(cid:13)(cid:13)eM(cid:13)(cid:13)(cid:13)(p,q)
kMkL(L(ℓp,ℓq),L(ℓr,ℓs)) =(cid:13)(cid:13)(cid:13)eM(cid:13)(cid:13)(cid:13)L(L(ℓp,ℓq),L(ℓr,ℓs))
Proof. The proof is almost identical to that of [4, Theorem 4.1], and the condition
r ≤ p is used to ensure that x1p + x2p ≤ (x1r + x2r)p/r for all x1, x2 ∈ C (with
the obvious modification for p = ∞ or r = ∞).
j,k=1 be such that
.
.
(cid:3)
Remark 6.5. By considering the transpose M′ of a matrix M, and using that M′ :
L(ℓq∗, ℓp∗ ) → L(ℓs∗, ℓr∗) with
(cid:13)(cid:13)M′(cid:13)(cid:13)L(L(ℓ
q∗,ℓ
p∗ ), L(ℓ
s∗,ℓ
r∗ )) = kMkL(L(ℓp,ℓq), L(ℓr,ℓs)) ,
as is straightforward to check, one obtains from Proposition 6.4 that for q∗ ≥
s∗, that is, for s ≤ q, row repetitions do not alter the k·kL(L(ℓp,ℓq), L(ℓr,ℓs))-norm
of a matrix. Moreover, since kSkL(ℓp,ℓq) is invariant under permutations of the
columns and rows of S ∈ L(ℓp, ℓq), rearrangement of the rows and columns of
M ∈ L(L(ℓp, ℓq), L(ℓr, ℓs)) does not change the norm kMkL(L(ℓp,ℓq), L(ℓr,ℓs)).
The following lemma is crucial to our main results.
Lemma 6.6. Let p, q, r, s ∈ [1, ∞] with r ≤ p and s ≤ q. Let λ = {λj}∞
µ = {µk}∞
k=1 be sequences of real numbers. Then
j=1 and
(cid:13)(cid:13)(cid:13)Tλ,µ
∆,n(cid:13)(cid:13)(cid:13)L(L(ℓp,ℓq), L(ℓr,ℓs)) ≤ kT∆,nkL(L(ℓp,ℓq), L(ℓr,ℓs))
for all n ∈ N.
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
21
∆,n (S) = M ∗ S for all S ∈ L(ℓp, ℓq), where M = {mjk}∞
Proof. Note that Tλ,µ
j,k=1 is
such that mjk = 1 if 1 ≤ j, k ≤ n and µk ≤ λj, and mjk = 0 otherwise. It suffices to
prove that kMkL(L(ℓp,ℓq), L(ℓr,ℓs)) ≤ kT kL(L(ℓp,ℓq), L(ℓr,ℓs)). Assume that M is non-
zero, otherwise the statement is trivial. By Remark 6.5, rearrangement of the rows
and columns of M does not change its norm. Hence, we may assume that {λj}n
j=1
and {µk}n
k=1 are decreasing. Now M has the property that, if mjk = 1, then mil = 1
for all i ≤ j and k ≤ l ≤ m2. By Proposition 6.4 and Remark 6.5, we may omit
repeated rows and columns of M, and doing this repeatedly reduces M to T∆,N for
some 1 ≤ N ≤ n. Noting that kT∆,NkL(L(ℓp,ℓq), L(ℓr,ℓs)) ≤ kT∆,nkL(L(ℓp,ℓq), L(ℓr,ℓs))
concludes the proof.
(cid:3)
6.2. The cases p < q, p = q = 1, p = q = ∞. We now combine the theory
from the previous sections to deduce our main results. Throughout this section let
f (t) := t for t ∈ R.
Theorem 6.7. Let p, q ∈ [1, ∞] with p < q, p = q = 1 or p = q = ∞. Then there exists
a constant C ≥ 0 such that the following holds (where ℓ∞ should be replaced by c0). Let
A ∈ Ld(ℓp) and B ∈ Ld(ℓq) have real spectrum. Then
k f (B)S − S f (A)kL(ℓp,ℓq) ≤ CKAKB kBS − SAkL(ℓp,ℓq)
for all S ∈ L(ℓp, ℓq).
Proof. Simply combine Propositions 5.4 and 5.6 and Lemma 6.6 with Proposition
6.3 (i).
(cid:3)
We single out the specific case in Theorem 6.7 where p = q = 1 or p = q = ∞
and S is the identity operator.
Corollary 6.8. There exists a universal constant C ≥ 0 such that
k f (B) − f (A)kL(ℓ1) ≤ CKAKB kB − AkL(ℓ1)
for all A, B ∈ Ld(ℓ1) with real spectrum.
Corollary 6.9. There exists a universal constant C ≥ 0 such that
k f (B) − f (A)kL(c0) ≤ CKAKB kB − AkL(c0)
for all A, B ∈ Ld(c0) with real spectrum.
In the case of Theorem 6.7 where p = 2 or q = 2, we can apply our results to
compact self-adjoint operators. By the spectral theorem, any compact self-adjoint
operator A ∈ L(ℓ2) has an orthonormal basis of eigenvectors, and therefore A ∈
Ld(ℓ2, λ, U) for some sequence λ of real numbers and an isometry U ∈ L(ℓ2).
Thus Theorem 6.7 yields the following corollaries.
Corollary 6.10. Let p ∈ [1, 2). Then there exists a constant C ≥ 0 such that the following
holds. Let A ∈ Ld(ℓp) and let B ∈ L(ℓ2) be compact and self-adjoint. Then
k f (B)S − S f (A)kL(ℓp,ℓ2) ≤ CKA kBS − SAkL(ℓp,ℓ2)
for all S ∈ L(ℓp, ℓ2).
22
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
Corollary 6.11. Let q ∈ (2, ∞]. Then there exists a constant C ≥ 0 such that the
following holds (where ℓ∞ should be replaced by c0). Let A ∈ L(ℓ2) be compact and
self-adjoint, and let B ∈ Ld(ℓq). Then
k f (B)S − S f (A)kL(ℓ2,ℓq) ≤ CKB kBS − SAkL(ℓ2,ℓq)
for all S ∈ L(ℓ2, ℓq).
Remark 6.12. Corollaries 6.8 and 6.9 show that the absolute value function is op-
erator Lipschitz on ℓ1 and c0, in the following sense. For fixed M ≥ 1, there exists
a constant C ≥ 0 such that
k f (B) − f (A)k ≤ C kB − Ak
for all diagonalizable operators A and B such that KA, KB ≤ M, and C is indepen-
dent of A and B.
For p < q a similar statement holds. Restricting B ∈ Ld(ℓq) and f (B) ∈ L(ℓq)
to operators from ℓp to ℓq, and letting S be the inclusion mapping ℓp ֒→ ℓq in
Theorem 6.7, one can suggestively write
k f (B) − f (A)kL(ℓp,ℓq) ≤ C kB − AkL(ℓp,ℓq) ,
for all A ∈ Ld(ℓp) and B ∈ Ld(ℓq) with KA, KB ≤ M. This also applies to Corol-
laries 6.10 and 6.11.
6.3. The case p ≥ q. We now examine the absolute value function f on L(ℓp, ℓq)
for p ≥ q, and obtain the following result.
Proposition 6.13. Let p, q ∈ (1, ∞] with p ≥ q. Then for each s < q there exists a
constant C ≥ 0 such that the following holds (where ℓ∞ should be replaced by c0). Let
A ∈ Ld(ℓp, λ, U) and B ∈ Ld(ℓq, µ, V) have real spectrum, and let S ∈ L(ℓp, ℓq) be
such that V(BS − SA)U−1 ∈ L(ℓp, ℓs). Then
k f (B)S − S f (A)kL(ℓp,ℓq) ≤ CkUkL(ℓp)kV−1kL(ℓq)kV(BS − SA)U−1kL(ℓp,ℓs).
In particular, if p = q and V(B − A)U−1 ∈ L(ℓp, ℓs), then
k f (B) − f (A)kL(ℓp) ≤ CkUkL(ℓp)kV−1kL(ℓp)kV(B − A)U−1kL(ℓp,ℓs).
Proof. Let R := V(BS − SA)U−1. With notation as in Lemma 5.3,
for each n ∈ N. Proposition 5.6, Lemma 6.6 (with p = r and with q and s in-
terchanged) and Proposition 6.3 (iii) (with q and s interchanged) yield a constant
C′ ≥ 0 such that
k f (B)Sn − Sn f (A)kL(ℓp,ℓq) ≤ kUkL(ℓp)kV−1kL(ℓq)(cid:13)(cid:13)(cid:13)Tλ,µ
(cid:13)(cid:13)(cid:13)Tλ,µ
ϕ f ,n(R)(cid:13)(cid:13)(cid:13)L(ℓp,ℓq) ≤ C′(cid:16)kRkL(ℓp,ℓq) + kRkL(ℓp,ℓs)(cid:17) .
ϕ f ,n(R)(cid:13)(cid:13)(cid:13)L(ℓp,ℓq)
Since L(ℓp, ℓs) ֒→ L(ℓp, ℓq) contractively,
k f (B)Sn − Sn f (A)kL(ℓp,ℓq) ≤ CkUkL(ℓp)kV−1kL(ℓq)kV(BS − SA)U−1kL(ℓp,ℓs)
for all n ∈ N, where C = 2C′. Finally, as in the proof of Proposition 5.4, one lets n
tend to infinity to conclude the proof.
(cid:3)
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
23
In the same way, appealing to the second part of Proposition 6.3 (iii), one de-
duces the following result.
Proposition 6.14. Let p, q ∈ [1, ∞) with p ≥ q. Then for each r > p there exists a
constant C ≥ 0 such that the following holds (where ℓ∞ should be replaced by c0). Let
A ∈ Ld(ℓp, λ, U) and B ∈ Ld(ℓq, µ, V) have real spectrum, and let S ∈ L(ℓp, ℓq) be
such that V(BS − SA)U−1 ∈ L(ℓr, ℓq). Then
k f (B)S − S f (A)kL(ℓp,ℓq) ≤ CkUkL(ℓp)kV−1kL(ℓq)kV(BS − SA)U−1kL(ℓr,ℓq).
In particular, if p = q and V(B − A)U−1 ∈ L(ℓr, ℓq), then
k f (B) − f (A)kL(ℓp) ≤ CkUkL(ℓp)kV−1kL(ℓp)kV(B − A)U−1kL(ℓr,ℓq).
We single out the case where p = q = 2. Here we write f (A) = A for a
normal operator A ∈ L(ℓ2), since then f (A) is equal to A := √A∗ A. Note also
that the following result applies in particular to compact self-adjoint operators.
For simplicity of the presentation we only consider ǫ ∈ (0, 1], it should be clear
how the result extends to other ǫ > 0.
Corollary 6.15. For each ǫ ∈ (0, 1] there exists a constant C ≥ 0 such that the follow-
ing holds. Let A ∈ Ld(ℓ2, λ, U) and B ∈ Ld(ℓ2, µ, V) be self-adjoint, with U and V
unitaries, and let S ∈ L(ℓ2). If V(BS − SA)U−1 ∈ L(ℓ2, ℓ2−ǫ), then
kBS − SAkL(ℓ2) ≤ CkV(BS − SA)U−1kL(ℓ2,ℓ2−ǫ)
and if V(BS − SA)U−1 ∈ L(ℓ2+ǫ, ℓ2) then
kBS − SAkL(ℓ2) ≤ CkV(BS − SA)U−1kL(ℓ2+ǫ,ℓ2).
In particular, if V(B − A)U−1 ∈ L(ℓ2, ℓ2−ǫ), then
kB − AkL(ℓ2) ≤ CkV(B − A)U−1kL(ℓ2,ℓ2−ǫ)
and if V(B − A)U−1 ∈ L(ℓ2+ǫ, ℓ2), then
kB − AkL(ℓ2) ≤ CkV(B − A)U−1kL(ℓ2+ǫ,ℓ2).
7. LIPSCHITZ ESTIMATES ON THE IDEAL OF p-SUMMING OPERATORS
Let H be a separable infinite-dimensional Hilbert space. It was shown in [1]
that a matrix M = {mjk}∞
j,k=1 is a Schur multiplier on the Hilbert-Schmidt class
S2 ⊂ L(H) if and only if supj,k mjk < ∞. By [23], S2 coincides with the Banach
ideal Πp(H) of all p-summing operators (see the definition below) for all 1 ≤ p <
∞. Hence a matrix M = {mjk}∞
j,k=1 is a Schur multiplier on Πp(H) if and only if
supj,k mjk < ∞. In Lemma 7.2 we show that the same statement is true for the
Banach ideal Πp(ℓp∗, ℓp) in L(ℓp∗, ℓp), for all 1 < p < ∞. As a corollary we obtain
operator Lipschitz estimates on Πp(ℓp∗, ℓp) for each Lipschitz function f on C.
Let X and Y be Banach spaces and 1 ≤ p < ∞. An operator S : X → Y is
p-absolutely summing if there exists a constant C such that for each n ∈ N and each
collection {xj}n
j=1 ⊆ X,
(7.1)
∑
Y(cid:17) 1
(cid:16) n
j=1(cid:13)(cid:13)S(xj)(cid:13)(cid:13)p
p ≤ C sup
kx∗kX∗≤1(cid:16) n
∑
j=1hx∗, xjip(cid:17) 1
p
.
24
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
Below we consider p-absolutely summing operators from ℓp∗ to ℓp. We first
The smallest such constant is denoted by πp, and Πp(X, Y) is the space of p-
absolutely summing operators from X to Y. We let Πp(X) := Πp(X, X). By Propo-
sitions 2.3, 2.4 and 2.6 in [13], (Πp(X, Y), πp(·)) is a Banach ideal in L(X, Y).
present the following result.
Lemma 7.1. Let 1 < p < ∞ and S = {sjk}∞
∞
∑
j,k=1. Then S ∈ Πp(ℓp∗, ℓp) if and only if
< ∞.
p
cp :=(cid:16) ∞
∑
j=1
k=1sjkp(cid:17) 1
In this case, πp(S) = cp.
Proof. In [13, Example 2.11] it is shown that, if cp < ∞, then S ∈ Πp(ℓp∗, ℓp) with
πp(S) ≤ cp. For the converse, let n ∈ N and let xj := ej ∈ ℓp∗ for 1 ≤ j ≤ n. By
(7.1) (with X = ℓp∗ and Y = ℓp),
(cid:16) n
∑
k=1
∞
∑
j=1sjkp(cid:17) 1
p ≤ πp(S).
Letting n tend to infinity concludes the proof.
(cid:3)
For the following corollary of Lemma 7.1, recall that a matrix M is said to be a
Schur multiplier on a subspace I ⊆ L(ℓp, ℓq) if S 7→ M ∗ S is a bounded map on
I. Recall also the definition of the standard triangular truncation T∆ from (6.1).
Corollary 7.2. Let p ∈ (1, ∞) and let M = {mjk}∞
multiplier on Πp(ℓp∗, ℓp) if and only if supj,k mjk < ∞. In this case,
j,k=1 be a matrix. Then M is a Schur
p∗ ,ℓp)) = sup
In particular, T∆ ∈ L(Πp(ℓp∗, ℓp)) with kT∆kL(Πp(ℓ
kMkL(Πp(ℓ
j,k mjk.
p∗,ℓp)) = 1.
less, T∆ is bounded on the ideal Πp(ℓp∗, ℓp) ⊂ L(ℓp∗, ℓp) for all p ∈ (1, ∞).
Observe that T∆ /∈ L(L(ℓp∗, ℓp)) for p∗ ≥ p, by Proposition 6.3 (ii). Neverthe-
For a Lipschitz function f : C → C, write
k fkLip := sup
z1,z2∈C
z16=z2
Moreover, let ϕ f : C2 → C be given by
f (z1) − f (z2)
.
z1 − z2
ϕ f (λ1, λ2) :=( λ1−λ2
λ1−λ2
0
if λ1 6= λ2
otherwise
.
We now prove our main result concerning commutator estimates on Πp(ℓp∗, ℓp).
: C → C be
Theorem 7.3. Let p ∈ (1, ∞), A ∈ Ld(ℓp∗ ) and B ∈ Ld(ℓp). Let f
Lipschitz. Then
(7.2)
for all S ∈ L(ℓp∗, ℓp) such that BS − SA ∈ Πp(ℓp∗, ℓp).
πp( f (B)S − S f (A)) ≤ KAKB k fkLip πp(BS − SA)
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
25
j=1 and µ = {µk}∞
It follows directly from (7.1) that, if {Sm}∞
Proof. Let λ = {λj}∞
k=1 be sequences such that A ∈ Ld(ℓp∗, λ, U)
and B ∈ Ld(ℓp, µ, V) for certain U ∈ L(ℓp∗) and V ∈ L(ℓp).
m=1 ⊆ Πp(ℓp∗, ℓp) is a πp-bounded se-
quence which SOT-converges to some S ∈ L(X, Y), then S ∈ Πp(ℓp∗, ℓp) with
πp(S) ≤ lim supm→∞ πp(Sm). Hence, by Remark 5.5, it suffices to prove that
supn∈N kTλ,µ
p∗,ℓp)) ≤ k fkLip, where
ϕ f ,nkL(Πp(ℓ
Tλ,µ
ϕ f ,n(S) =
n
∑
j,k=1
ϕ f (λj, µk)PkSPj
(S ∈ Πp(ℓp∗, ℓp))
for n ∈ N. Fix n ∈ N and note that Tλ,µ
M = {mjk}∞
mjk = 0 otherwise. Then
ϕ f ,n(S) = M ∗ S for S ∈ Πp(ℓp∗, ℓp), where
j,k=1 is the matrix given by mjk = ϕ f (λj, µk) for 1 ≤ j, k ≤ n, and
sup
j,k mjk ≤ sup
j,k ϕ f (λj, µk) ≤ k fkLip .
Corollary 7.2 now concludes the proof.
q + 1
Problem 7.4. Let p, q, r ∈ [1, ∞] be such that q ≥ 2 and 1
2 .
Then the Schatten class Sr coincides with the Banach ideal Πp,q(ℓ2) of (p, q)-summing
operators on ℓ2 (for the definition of (p, q)-summing operators see [13]). Hence, by [14],
the standard triangular truncation is bounded on Πp,q(ℓ2) for these indices. For which
r1, r2 ∈ [1, ∞] is the standard triangular truncation bounded on Πp,q(ℓr1, ℓr2 )? Are there
other non-trivial ideals I in L(ℓp, ℓq) such that the standard triangular truncation is
bounded on I? As shown in Theorem 7.3 (and the rest of this paper), answers to these
questions are linked to commutator estimates for diagonalizable operators.
2 and 1
q − 1
p − 1
r = 1
< 1
p
(cid:3)
8. MATRIX ESTIMATES
In this section we apply the theory developed in Sections 4, 5 and 6 to finite-
dimensional spaces, and in doing so we obtain Lipschitz estimates which are in-
dependent of the dimension of the underlying space. The results from Section 7,
in particular Theorem 7.3, also yield dimension-independent estimates on finite-
dimensional spaces, but we leave the straightforward derivation of these results
to the reader.
8.1. Finite-dimensional spaces. Let n ∈ N and let X be an n-dimensional Banach
space with basis {e1, . . . , en} ⊂ X. For 1 ≤ k ≤ n, let Pk ∈ L(X) be the projec-
tion given by Pk(x1e1 + . . . + xnen) := xkek for (x1, . . . , xn) ∈ Cn. Recall that an
operator A ∈ L(X) is diagonalizable if there exists U ∈ L(X) invertible such that
(8.1)
U AU−1 =
n
∑
k=1
λkPk
for some (λ1, . . . , λn) ∈ Cn. In this case we write A ∈ Ld(X, {λj}n
also the definition of spectral and scalar type operators from Section 3.1.
Lemma 8.1. Each A ∈ L(X) is a spectral operator. Furthermore, A is a scalar type
operator if and only if A is diagonalizable. If A ∈ Ld(X, {λj}n
j=1, U) then the spectral
j=1, U). Recall
26
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
measure E of A is given by E(U) = 0 if U ∩ sp(A) = ∅, and E({λ}) = ∑λj=λ U−1PjU
for λ ∈ sp(A).
Proof. It was already remarked in Section 5 that any diagonalizable operator is a
scalar type operator, with spectral measure as specified. By [15, Theorem XV.4.5],
an operator T ∈ L(Y) on an arbitrary Banach space Y is a spectral operator if and
only if T = S + N for a scalar type operator S ∈ L(Y) and a generalized nilpotent
operator N ∈ L(Y) such that NS = SN, and this decomposition is unique. Com-
bining this with the Jordan decomposition for matrices yields that each A ∈ L(X)
is a spectral operator. If A is a scalar type operator, then the Jordan decomposition
yields a commuting diagonalizable S and a nilpotent N such that A = S + N. By
the uniqueness of such a decomposition [15, Theorem XV.4.5], N = 0 and A = S
is diagonalizable.
(cid:3)
Let A ∈ Ld(X, {λj}n
j=1, U). As in (5.3),
f (A) = U−1(cid:16) n
∑
k=1
f (λk)Pk(cid:17)U.
L(X);
Let Y be a finite-dimensional Banach space. A norm k·k on L(X, Y) is symmetric if
• kRSTk ≤ kRkL(Y) kSk kTkL(X) for all R ∈ L(Y), S ∈ L(X, Y) and T ∈
• kx∗⊗ yk = kx∗kX∗ kykY for all x∗ ∈ X∗ and y ∈ Y.
Clearly (L(X, Y), k·k) is a Banach ideal in the sense of Section 3.2 if and only if k·k
is symmetric.
The following result extends inequalities which were known for self-adjoint
operators on finite-dimensional Hilbert spaces and unitarily invariant norms (see
e.g. [18] and [5, Chapter X]), to diagonalizable operators on finite-dimensional Ba-
nach spaces and symmetric norms. Note that, for general finite-dimensional Ba-
nach spaces X and Y, a symmetric norm on L(X, Y) is not unitarily invariant. Let
A := A(C × C) be as in Section 3.3, and for f ∈ B(C) let ϕ f (λ1, λ2) := f (λ2)− f (λ1)
for (λ1, λ2) ∈ C2 with λ1 6= λ2, as in (4.5). Recall the definition of the spectral
constants ν(A) and ν(B) from Section 3.1. The following is a direct corollary of
Theorem 4.6, since L(X, Y) has the strong convex compactness property.
Theorem 8.2. Let f ∈ B(C) be such that ϕ f extends to an element of A. Let X and Y be
finite-dimensional Banach spaces, k·k a symmetric norm on L(X, Y), and let A ∈ Ld(X)
and B ∈ Ld(Y). Then
k f (B)S − S f (A)k ≤ 16 ν(A)ν(B)(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A kBS − SAk
k f (B) − f (A)k ≤ 16 ν(A)ν(B)(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A kB − Ak .
for all S ∈ L(X, Y). In particular, if X = Y,
(8.2)
Corollary 8.3. There exists a universal constant C ≥ 0 such that the following holds. Let
X and Y be finite-dimensional Banach spaces and k·k a symmetric norm on L(X, Y). Let
f ∈ B1
∞,1(R), and let A ∈ Ld(X) and B ∈ Ld(Y) be such that sp(A) ∪ sp(B) ⊆ R.
Then
λ2−λ1
k f (B)S − S f (A)k ≤ C ν(A)ν(B) k fk B1
∞,1(R)kBS − SAk
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
27
for all S ∈ L(X, Y). In particular, if X = Y,
k f (B) − f (A)k ≤ C ν(A)ν(B)k fk B1
∞,1(R)kB − Ak .
Remark 8.4. Let σ1, σ2 ⊂ C be finite sets. Then any ϕ : σ1 × σ2 → C belongs to
A(σ1 × σ2). Indeed, one can find a representation as in (3.8) by letting Ω be finite
and solving a system of linear equations. Therefore Theorem 4.6 yields an estimate
k f (B)S − S f (A)k ≤ 16 ν(A)ν(B)(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A(sp(A)×sp(B)) kBS − SAk
as in (4.6) for all f ∈ B(C). This might lead one to think that the restriction in
Theorem 8.2 that ϕ f extends to an element of A is not really necessary. However,
for general f ∈ B(C) the norm(cid:13)(cid:13)ϕ f(cid:13)(cid:13)A(sp(A)×sp(B)) may blow up as the number of
points in sp(A) and sp(B) grows to infinity. Indeed, as remarked before, letting
f ∈ B(C) be the absolute value function and considering the operator norm, a
dimension-independent estimate as in (8.2) does not hold for all self-adjoint oper-
ators on all finite-dimensional Hilbert spaces [5, (X.25)]. Hence ϕ f does not extend
to an element of A, and one cannot expect to obtain Theorem 8.2 or Corollary 8.3
for all bounded Borel functions on C.
8.2. The absolute value function. We now apply the results for the absolute value
function from previous sections to finite-dimensional spaces. Throughout this sec-
tion, let f (t) := t for t ∈ R.
First note that Lemma 5.3 and Proposition 5.6 relate commutator estimates to
estimates for triangular truncation operators for general symmetric norms on ma-
trix spaces.
Theorem 6.7 immediately yields the following result. It shows that, although the
Lipschitz estimate
k f (B) − f (A)kL(ℓn
q ) ≤ const kB − AkL(ℓn
p,ℓn
p,ℓn
q )
does not hold with a constant independent of the dimension n for f the absolute
value function, p = q = 2 and all self-adjoint operators on ℓn
2 , one can neverthe-
less obtain such estimates for p < q, p = q = 1 or p = q = ∞ by considering
diagonalizable operators A and B for which KA, KB ≤ M, for some fixed M ≥ 1.
For A a diagonalizable operator, recall the definition of KA from (5.5).
Theorem 8.5. Let p, q ∈ [1, ∞] with p < q, p = q = 1 or p = q = ∞. Then there
exists a constant C ≥ 0 such that the following holds. Let n ∈ N and let A ∈ Ld(ℓn
p)
and B ∈ Ld(ℓn
q ) have real spectrum. Then
k f (B)S − S f (A)kL(ℓn
q ) ≤ CKAKB kBS − SAkL(ℓn
p,ℓn
p,ℓn
q )
p, ℓn
q ). In particular,
k f (B) − f (A)kL(ℓn
q ) ≤ CKAKB kB − AkL(ℓn
p,ℓn
p,ℓn
q ) .
for all S ∈ L(ℓn
For n ∈ N and p ∈ [1, ∞) let ℓn
k(x1, . . . , xn)kp :=(cid:16) n
∑
and let ℓn
∞ be Cn with the norm
k(x1, . . . , xn)k∞ := max
p denote Cn with the p-norm
j=1xjp(cid:17)1/p
1≤j≤nxj
(cid:0)(x1, . . . , xn) ∈ Cn(cid:1),
(cid:0)(x1, . . . , xn) ∈ Cn(cid:1).
28
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
Of course Corollaries 6.10 and 6.11 imply that for p = 2 or q = 2 and A or B
self-adjoint, the estimates in Theorem 8.5 simplify.
For p ≥ q, Propositions 6.13 and 6.14 yield dimension-independent estimates.
We state the estimates which follow from Proposition 6.13, the analogous result
that follows from Proposition 6.14 should be obvious.
Proposition 8.6. Let p, q ∈ (1, ∞] with p ≥ q. For each s < q there exists a constant
C ≥ 0 such that the following holds. Let n ∈ N, and let A ∈ Ld(ℓn
p, λ, U) and B ∈
Ld(ℓn
q , µ, V) have real spectrum. Then
k f (B)S − S f (A)kL(ℓn
q ) ≤ CkUkL(ℓn
p,ℓn
p)kV−1kL(ℓn
q )kV(BS − SA)UkL(ℓn
p,ℓn
s )
for all S ∈ L(ℓn
p, ℓn
q ). In particular,
k f (B) − f (A)kL(ℓn
q ) ≤ CkUkL(ℓn
p,ℓn
p)kV−1kL(ℓn
q )kV(B − A)UkL(ℓn
p,ℓn
s ).
In the case p = q = 2, Corollary 6.15 implies the following. Again we only
consider ǫ ∈ (0, 1], for simplicity, but the result extends in an obvious manner to
other ǫ > 0. We write f (A) = A = √A∗ A for a normal operator A on ℓn
Corollary 8.7. For each ǫ ∈ (0, 1] there exists a constant C ≥ 0 such that the following
holds. Let n ∈ N and let A ∈ Ld(ℓn
2 , µ, V) be self-adjoint
operators, with U and V unitaries. Then
2 , λ, U) and B ∈ Ld(ℓn
2 .
2 ) ≤ C min(kV(B − A)U−1kL(ℓn
2−ǫ), kV(B − A)U−1kL(ℓn
2 ,ℓn
(8.3)
kB − AkL(ℓn
We note that (8.3) in turn yields, for instance, the following estimate:
2 ) min(kU−1kL(ℓn
2−ǫ), kVkL(ℓn
2 ,ℓn
2 ) ≤ CkB − AkL(ℓn
kB − AkL(ℓn
2+ǫ,ℓn
2 )).
2+ǫ,ℓn
2 )).
Acknowledgements. Most of the work on this article was performed during a re-
search visit of the first-named author to the University of New South Wales. He
would like to thank UNSW and its faculty, in particular Michael Cowling, for their
wonderful hospitality. We are indebted to many colleagues in Delft and at UNSW
for their suggestions and ideas. In particular, we would like to mention Ian Doust,
Ben de Pagter, Denis Potapov and Dmitriy Zanin. We also thank Albrecht Pietsch
for useful comments and references concerning p-summing operators.
REFERENCES
[1] J. Arazy. Certain Schur-Hadamard multipliers in the spaces Cp. Proc. Amer. Math. Soc., 86(1):59 -- 64,
1982.
[2] N. A. Azamov, A. L. Carey, P. G. Dodds, and F. A. Sukochev. Operator integrals, spectral shift, and
spectral flow. Canad. J. Math., 61(2):241 -- 263, 2009.
[3] G. Bennett. Unconditional convergence and almost everywhere convergence. Z. Wahrschein-
lichkeitstheorie und Verw. Gebiete, 34(2):135 -- 155, 1976.
[4] G. Bennett. Schur multipliers. Duke Math. J., 44(3):603 -- 639, 1977.
[5] R. Bhatia. Matrix analysis, volume 169 of Graduate Texts in Mathematics. Springer-Verlag, New York,
1997.
[6] M. S. Birman and M. Z. Solomjak. Double Stieltjes operator integrals. In Probl. Math. Phys., No. I,
Spectral Theory and Wave Processes (Russian), pages 33 -- 67. Izdat. Leningrad. Univ., Leningrad, 1966.
[7] M. S. Birman and M. Z. Solomjak. Double Stieltjes operator integrals. III. In Problems of mathematical
physics, No. 6 (Russian), pages 27 -- 53. Izdat. Leningrad. Univ., Leningrad, 1973.
[8] M. Sh. Birman and M. Solomyak. Double operator integrals in a Hilbert space. Integral Equations
Operator Theory, 47(2):131 -- 168, 2003.
OPERATOR LIPSCHITZ FUNCTIONS ON BANACH SPACES
29
[9] E. B. Davies. Lipschitz continuity of functions of operators in the Schatten classes. J. London Math.
Soc. (2), 37(1):148 -- 157, 1988.
[10] B. de Pagter and F. A. Sukochev. Differentiation of operator functions in non-commutative L p-
spaces. J. Funct. Anal., 212(1):28 -- 75, 2004.
[11] B. de Pagter, H. Witvliet, and F. A. Sukochev. Double operator integrals. J. Funct. Anal., 192(1):52 --
111, 2002.
[12] A. Defant and K. Floret. Tensor norms and operator ideals, volume 176 of North-Holland Mathematics
Studies. North-Holland Publishing Co., Amsterdam, 1993.
[13] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge Studies
in Advanced Mathematics. Cambridge University Press, Cambridge, 1995.
[14] P. G. Dodds, T. K. Dodds, B. de Pagter, and F. A. Sukochev. Lipschitz continuity of the absolute
value and Riesz projections in symmetric operator spaces. J. Funct. Anal., 148(1):28 -- 69, 1997.
[15] N. Dunford and J. T. Schwartz. Linear operators. Part III: Spectral operators. Interscience Publishers
[John Wiley & Sons, Inc.], New York-London-Sydney, 1971. With the assistance of William G. Bade
and Robert G. Bartle, Pure and Applied Mathematics, Vol. VII.
[16] I. C. Gohberg and M. G. Kreın. Theory and applications of Volterra operators in Hilbert space. Translated
from the Russian by A. Feinstein. Translations of Mathematical Monographs, Vol. 24. American
Mathematical Society, Providence, R.I., 1970.
[17] T. Kato. Continuity of the map S 7→ S for linear operators. Proc. Japan Acad., 49:157 -- 160, 1973.
[18] H. Kosaki. Unitarily invariant norms under which the map A → A is Lipschitz continuous. Publ.
[19] S. Kwapie ´n and A. Pełczy ´nski. The main triangle projection in matrix spaces and its applications.
Res. Inst. Math. Sci., 28(2):299 -- 313, 1992.
Studia Math., 34:43 -- 68, 1970.
[20] S. Lassalle and P. Turco. The weak bounded approximation property for A. Online at
[21] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. I. Springer-Verlag, Berlin, 1977. Sequence
http://arxiv.org/abs/1410.5670, 2014.
spaces, Ergebnisse der Mathematik und ihrer Grenzgebiete, Vol. 92.
[22] A. McIntosh. Counterexample to a question on commutators. Proc. Amer. Math. Soc., 29:337 -- 340,
1971.
[23] A. Pełczy ´nski. A characterization of Hilbert-Schmidt operators. Studia Math., 28:355 -- 360,
1966/1967.
[24] V. V. Peller. Hankel operators in the theory of perturbations of unitary and selfadjoint operators.
Funktsional. Anal. i Prilozhen., 19(2):37 -- 51, 96, 1985.
[25] V. V. Peller. Hankel operators in the perturbation theory of unbounded selfadjoint operators. In
Analysis and partial differential equations, volume 122 of Lecture Notes in Pure and Appl. Math., pages
529 -- 544. Dekker, New York, 1990.
[26] V. V. Peller. Multiple operator integrals and higher operator derivatives. J. Funct. Anal., 233(2):515 --
544, 2006.
[27] A. Pietsch. Operator ideals, volume 20 of North-Holland Mathematical Library. North-Holland Pub-
lishing Co., Amsterdam-New York, 1980. Translated from German by the author.
[28] D. Potapov and F. Sukochev. Double operator integrals and submajorization. Math. Model. Nat.
Phenom., 5(4):317 -- 339, 2010.
[29] D. Potapov and F. Sukochev. Operator-Lipschitz functions in Schatten-von Neumann classes. Acta
Math., 207(2):375 -- 389, 2011.
[30] D. Potapov, F. Sukochev, and A. Tomskova. On the Arazy conjecture concerning Schur multipliers
on Schatten ideals. Adv. Math., 268:404 -- 422, 2015.
[31] G. Schl uchtermann. On weakly compact operators. Math. Ann., 292(2):263 -- 266, 1992.
[32] Ivan Singer. Bases in Banach spaces. I. Springer-Verlag, New York-Berlin, 1970. Die Grundlehren der
mathematischen Wissenschaften, Band 154.
[33] A. Skripka, F. Sukochev, and A. Tomskova. Multilinear operator integrals: Development and ap-
plications. preprint, 2015.
[34] F. Sukochev and A. Tomskova. (E, F)-Schur multipliers and applications. Studia Math., 216(2):111 --
129, 2013.
[35] N. N. Vakhania, V. I. Tarieladze, and S. A. Chobanyan. Probability distributions on Banach spaces,
volume 14 of Mathematics and its Applications (Soviet Series). D. Reidel Publishing Co., Dordrecht,
1987. Translated from the Russian and with a preface by Wojbor A. Woyczynski.
[36] J. Voigt. On the convex compactness property for the strong operator topology. Note Mat., 12:259 --
269, 1992.
30
JAN ROZENDAAL, FEDOR SUKOCHEV, AND ANNA TOMSKOVA
DELFT INSTITUTE OF APPLIED MATHEMATICS, MEKELWEG 4, 2628CD DELFT, THE NETHER-
LANDS
E-mail address: [email protected]
SCHOOL OF MATHEMATICS & STATISTICS, UNIVERSITY OF NSW, KENSINGTON NSW 2052 AUS-
TRALIA
E-mail address: [email protected]
E-mail address: [email protected]
|
1701.07311 | 1 | 1701 | 2017-01-25T13:55:27 | Simultaneous universality | [
"math.FA",
"math.DS"
] | In this paper, the notion of simultaneous universality is introduced, concerning operators having orbits that simultaneously approximate any given vector. This notion is related to the well known concepts of universality and disjoint universality. Several criteria are provided, and several applications to specific operators or sequences of operators are performed, mainly in the setting of sequence spaces or spaces of holomorphic functions. | math.FA | math |
SIMULTANEOUS UNIVERSALITY
L. BERNAL-GONZ ´ALEZ AND A. JUNG
Abstract. In this paper, the notion of simultaneous universality is introduced,
concerning operators having orbits that simultaneously approximate any given
vector. This notion is related to the well known concepts of universality and
disjoint universality. Several criteria are provided, and several applications to
specific operators or sequences of operators are performed, mainly in the setting
of sequence spaces or spaces of holomorphic functions.
1. Introduction
In this paper, we are concerned with the phenomenon of simultaneous appro-
ximation by the action of several operators or, more generally, by the action of
several sequences of mappings. When the existence of a dense orbit under an
operator is proved, we are speaking about universality or hypercyclicity, see below.
In many situations, it is possible to show the existence of one vector whose orbits
under two or more operators approximate any given vector. Pushing the question
quite further, we wonder under what conditions such approximation takes place
by using a common subsequence. This, together with its connection with other
kinds of joint universality, will make up the main aim of the present manuscript.
Next, we fix some related notation and terminology to be used in this work.
For a good account of concepts, results and history concerning hypercyclicity, the
reader is referred to the books [2, 21].
By N, N0, R, C, D, B(a, r), B(a, r) (a ∈ C, r > 0) we denote, respectively, the
set of positive integers, the set N ∪ {0}, the real line, the complex plane, the open
unit disk {z ∈ C :
z < 1}, the open disk with center a and radius r, and the
corresponding closed disk. Let X, Y be two Hausdorff topological spaces, and
Tn : X → Y (n = 1, 2, . . . ) be a sequence of continuous mappings. Recall that
(Tn) is said to be universal whenever there is some (Tn)-orbit which is dense in Y ,
that is, there exists an element x0 ∈ X -- called universal for (Tn) -- such that
{Tnx0 : n ∈ N} = Y.
Note that Y must be separable. We denote by U((Tn)) the set of universal ele-
ments for (Tn). When X = Y and T : X → X is a continuous self-mapping,
2010 Mathematics Subject Classification. 30E10, 47B33, 47A16, 47B38.
Key words and phrases. hypercyclic operator, composition operator, disjoint universality, si-
multaneous universality.
1
2
BERNAL AND JUNG
then T is called universal provided that the sequence (T n) of iterates of T (i.e.,
T 1 = T , T 2 = T ◦ T , T 3 = T ◦ T 2, and so on) is universal, in which case the set
U((T n)) of universal elements will be denoted by U(T ). A sequence Tn : X → Y
(n = 1, 2, . . . ) of continuous mappings is said to be densely universal if U((Tn)) is
dense in X. Birkhoff's transitivity theorem asserts that, if X is a Baire space (in
particular, if X is completely metrizable) and Y is second-countable (in particular,
if X is metrizable and separable), then (Tn) is densely universal if and only if (Tn)
is transitive (that is, given nonempty open sets U ⊂ X, V ⊂ Y , there is N ∈ N
with TN (U) ∩ V 6= ∅); if this is the case, then U((Tn)) is residual (in fact, a dense
Gδ subset) in X. If X lacks isolated points and T : X → X is universal, then
U(T ) is dense in X (so residual if X is, in addition, completely metrizable).
In the case in which X and Y are topological vector spaces over K (= R or
C) and (Tn) ⊂ L(X, Y ) := {linear continuous mappings X → Y }, the words
hypercyclic and universal are synonymous, although hypercyclic is mostly used, as
well as the alternative notation HC((Tn)) := U((Tn)) (and HC(T ) := U(T ) for
T ∈ L(X) := L(X, X) = {operators on X}). In particular, we have if X and Y
are F-spaces with Y separable, then HC((Tn)) (HC(T ), with X separable, resp.)
is residual in X as soon as (Tn) is transitive (as soon as T is hypercyclic, resp.).
Recall that an F-space is a completely metrizable topological vector space.
Assume now that X, Y are topological spaces, with X a Baire space and Y
second-countable, and that Sn : X → Y and Tn : X → Y (n ∈ N) are densely
universal sequences. Since U((Sn)), U((Tn)) are dense Gδ subsets of X, we have
that U((Sn)) ∩ U((Tn)) is also dense, so non-empty. Hence there is a common
hypercyclic element x ∈ X. So, for a given point y ∈ Y , there are sequences
{n1 < n2 < · · · } and {m1 < m2 < · · · } in N such that
Snj x → y and Tmj x → y as j → ∞.
Then the following question arises naturally:
Under what conditions on (Sn) and (Tn) one can guarantee the exis-
tence of an element x ∈ X such that, for any given y ∈ Y , there is
one sequence {n1 < n2 < · · · } ⊂ N such that
Snj x −→ y ←− Tnj x as j → ∞?
Of course, a similar question can be posed for finitely many sequences and for
finitely many single operators on X, just by considering the sequences of their
iterates in the latter case. With this in mind, the new concept of simultaneous
universality will be introduced in the next section, and compared to other related
notions existing in the literature, such as those of disjoint hypercyclicity and the
weakly mixing property. Several sufficient conditions for simultaneous universal-
ity/hypercyclicity will be provided in Section 3. Examples of finite families of
simultaneous hypercyclic operators will be furnished in sections 4 -- 6, starting with
SIMULTANEOUS UNIVERSALITY
3
multiples of an operator and ending up in the frameworks of sequence spaces and
of spaces of analytic functions on complex domains.
2. Simultaneously universal sequences
Let us define the new concept that is the matter of this paper. If p ∈ N and
Y is a nonempty set, then by ∆(Y p) we denote the diagonal of Y p = Y × · · · × Y
(p times), that is, the subset ∆(Y p) = {(y, y, . . . , y) : y ∈ Y }. If Y is a topological
space, then Y p is assumed to be endowed with the product topology.
Definition 2.1. Let p ∈ N and X, Y be Hausdorff topological spaces. Assume
that, for each j ∈ {1, . . . , p}, Tj,n : X → Y (n ∈ N) is a sequence of continuous
mappings. Consider the sequence
[T1,n, . . . , Tp,n] : x ∈ X 7−→ (T1,nx, . . . , Tp,nx) ∈ Y p
(n ∈ N).
Let also T1, . . . , Tp : X → X be continuous mappings.
(a) We say that the sequences (T1,n), . . . , (Tp,n) are simultaneously universal
(or s-universal ) whenever there exists an element x0 ∈ X -- called s-universal
for (T1,n), . . . , (Tp,n) -- satisfying
{[T1,n, . . . , Tp,n]x0 : n ∈ N} ⊃ ∆(Y p).
The set of such s-universal elements will be denoted by s-U((T1,n), . . . , (Tp,n)).
(b) The sequences (T1,n), . . . , (Tp,n) are said to be densely simultaneously uni-
versal if the set s-U((T1,n), . . . , (Tp,n)) is dense in X. And they are called
hereditarily simultaneously universal (hereditarily densely simultaneously
universal, resp.) if, for every strictly increasing sequence (nk) ⊂ N, the
sequences (T1,nk), . . . , (Tp,nk) are s-universal (densely s-universal, resp.).
(c) The mappings T1, . . . , Tp are called s-universal (densely s-universal, here-
ditarily s-universal, hereditarily densely s-universal, resp.) if the sequences
(T n
p ) are s-universal (densely s-universal, hereditarily s-universal,
hereditarily densely s-universal, resp.). The set s-U((T n
p )) of cor-
responding s-universal elements will be denoted by s-U(T1, . . . , Tp).
1 ), . . . , (T n
1 ), . . . , (T n
Remarks 2.2. 1. If Y is first-countable (in particular, if Y is metrizable), then
the s-simultaneous universality of (Tj,n)n∈N (1 ≤ j ≤ p) means the existence of
some x0 ∈ X enjoying the property that, for every y ∈ Y , there is a (strictly
increasing) sequence (nk) ⊂ N such that Tj,nkx0 → y as k → ∞ (j = 1, . . . , p).
2. In [18, Kapitel 1] the notion of relative universality on a closed subset of the
arrival space is introduced under very general assumptions. In the present paper
we study a special case of this situation (note that ∆(Y p) is closed in Y p since Y p
is Hausdorff) under more specific hypotheses.
3. According to the introduction, if X, Y are topological vector spaces and Tj,n, Tj ∈
L(X, Y ) (j = 1, . . . , p; n ∈ N), then we use the expressions "s-hypercyclic",
4
BERNAL AND JUNG
"densely s-hypercyclic" and "hereditarily densely s-hypercyclic" rather than "s-
universal", "densely s-universal" and "hereditarily densely s-universal", respec-
tively.
s-U((T1,n), . . . , (Tp,n)) and s-HC(T1, . . . , Tp) := s-U(T1, . . . , Tp) in this case.
4. For a single operator T , hypercyclicity (hereditary hypercyclicity, resp.)
equivalent to dense hypercyclicity (hereditary dense hypercyclicity, resp.).
s-HC((T1,n), . . . , (Tp,n))
denote
In
addition, we will
:=
is
5. The property of simultaneous universality of (T1,n), . . . , (Tp,n) is weaker than
the property that the sequence ([T1,n, . . . , Tp,n]) is subspace-universal for ∆(Y p),
meaning that the set {[T1,n, . . . , Tp,n]x0 : n ∈ N} ∩ ∆(Y p) is dense in ∆(Y p) for
some x0 ∈ X (see e.g. [1,22,24] for results on subspace-hypercyclicity/universality).
Before going on, we want to compare s-universality to other related concepts
defined in the literature. In 2007, B`es, Peris and the first author ([11],[4]) intro-
duced the notion of disjoint (or d-) universality (sometimes called d-hypercyclicity
in the mentioned references). Under the same assumptions and terminology as in
Definition 2.1, the sequences (T1,n), . . . , (Tp,n) are said to be d-universal whenever
the sequence [T1,n . . . , Tp,n] : X → Y p (n ∈ N) is universal, that is, whenever there
exists some x0 ∈ X such that the joint orbit {(T1,nx0, . . . , Tp,nx0) : n ∈ N} is
dense in Y p. As a matter of fact, d-universality should not be confused with the
universality of the sequence
T1,n ⊕ · · · ⊕ Tp,n : (x1, . . . , xp) ∈ X p 7−→ (T1,nx1, . . . , Tp,nxp) ∈ Y p.
Trivially, disjoint universality of (T1,n), . . . , (Tp,n) implies universality of the last
sequence as well as simultaneous universality of (T1,n), . . . , (Tp,n). Also, trivially,
s-universality implies the universality of each sequence (Tj,n)n∈N (j = 1, . . . , p)
(in particular, Y must be separable). But no other implications among these
properties hold, even considering only p = 2 and sequences of iterates of single
operators. The following examples illustrate this situation:
1. Assume that T is a hypercyclic operator on a topological vector space.
Then the operators T, T are s-hypercyclic but not d-hypercyclic.
2. In 1969, S. Rolewicz [26] proved that if c ∈ K has modulus > 1 and B
is the backward shift (xn) ∈ ℓ2 7→ (xn+1) ∈ ℓ2, then the operator cB is
hypercyclic. In particular, the operators T = 2B and S = 4B = 2T are
hypercyclic, but T, S are clearly not s-hypercyclic.
3. Since each of the operators T, S of the latter example is mixing (see the
definition at the beginning of the next section, regarding the sequences of
iterates; see also [21, p. 46]), the operator T ⊕ S is hypercyclic, but T, S
are not s-hypercyclic.
4. De la Rosa and Read [15] were able to construct a Banach space X and an
operator T ∈ L(X) such that T is hypercyclic (hence T, T are s-hypercyclic)
but T is not weakly mixing on X, meaning that T ⊕ T is not hypercyclic
on X 2.
SIMULTANEOUS UNIVERSALITY
5
While d-hypercyclic operators must be substantially different, s-hypercyclicity
allows more similarity. For instance, an operator can never be d-hypercyclic with a
scalar multiple of itself (see [11, p. 299]). Nevertheless, s-hypercyclicity is possible
in concrete situations. This will be analyzed in Section 4. Sections 5 and 6 are
devoted to more specific operators, namely backward shifts and operators on spaces
of analytic functions.
We close this section by establishing, under appropriate assumptions, the exis-
tence of large vector subspaces consisting, except for zero, of s-hypercyclic vectors.
Theorem 2.3.
(a) Let X be a topological vector space and Tj ∈ L(X) (j =
1, . . . , p). If T1, . . . , Tp are s-hypercyclic and at least one of them commutes
with the others, then s-HC(T1, . . . , Tp) contains, except for 0, a dense lin-
ear subspace of X.
(b) Let X and Y be two topological vector spaces such that Y is metrizable.
Assume that (Tj,n) ⊂ L(X, Y ) (j = 1, . . . , p) are hereditarily s-hypercyclic
sequences. Then s-HC((T1,n), . . . , (Tp,n)) contains, except for 0, an infinite
dimensional vector subspace of X.
(c) Let X and Y be two metrizable separable topological vector spaces. Assume
that (Tj,n) ⊂ L(X, Y ) (j = 1, . . . , p) are hereditarily densely s-hypercyclic
sequences. Then s-HC((T1,n), . . . , (Tp,n)) contains, except for 0, a dense
linear subspace of X.
Proof. (a) By hypothesis, there is i ∈ {1, . . . , p} such that TiTj = TjTi (j =
1, . . . , p). Therefore P (Ti)Tj = TjP (Ti) for all j and every polynomial P with
coefficients in K. Let P denote the set of such polynomials. Of course, the
operator Ti is hypercyclic. From a result by Wengenroth [29], the operator P (Ti)
has dense range as soon as P ∈ P \ {0}. Pick any x0 ∈ s-HC(T1, . . . , Tp). Let
us define M := {P (Ti)x0 : P ∈ P \ {0}}. Then M is a linear subspace of X.
It is dense because M contains the orbit {T n
i x0 : n ∈ N}, that is dense in X as
x0 ∈ HC(Ti). It remains to show that M \ {0} ⊂ s-HC(T1, . . . , Tp).
To this end, fix u ∈ M \ {0}. Then there is P ∈ P \ {0} such that u = P (Ti)x0.
It must be proved that
Z ⊃ ∆(X p),
1 u, . . . , T n
p u) : n ∈ N} = {(P (Ti)T n
where Z := {(T n
p x0) : n ∈ N},
where the last equality follows from commutativity. We know that ∆(X p) ⊂
{(T n
p x0) : n ∈ N}, ϕ := P (Ti)
and Φ : X p → X p be the mapping defined as Φ(x1, . . . , xp) := (ϕ(x1), . . . , ϕ(xp)).
Then, as ϕ is continuous, we get
p x0) : n ∈ N}. Let A := {(T n
1 x0, . . . , P (Ti)T n
1 x0, . . . , T n
1 x0, . . . , T n
Z = Φ(A) ⊃ Φ(A) ⊃ Φ(∆(X p)) = {(ϕ(x), . . . , ϕ(x)) : x ∈ X},
so Z ⊃ {(ϕ(x), . . . , ϕ(x)) : x ∈ X}. Given y ∈ X and a neighborhood U of
(y, y, . . . , y), there exists a neighborhood V of y such that U ⊃ V p. Since ϕ
6
BERNAL AND JUNG
has dense range, one can find x ∈ X with ϕ(x) ∈ V . Then (ϕ(x), . . . , ϕ(x)) ∈
In other words, (y, . . . , y) ∈ {(ϕ(x), . . . , ϕ(x)) : x ∈ X}, so (y, . . . , y) ∈ Z.
U.
Consequently, Z ⊃ ∆(X p), as required.
(b) -- (c). By mimicking the proofs of Theorems 1 -- 2 of [3] (in which the results
are given for a single sequence (Tn)), we can construct recursively a sequence
(xN )N ∈N ⊂ X and a family {(q(N, k))k∈N : N ∈ N0} of strictly increasing
subsequences of N satisfying, for all N ∈ N, the following conditions: xN ∈
GN ∩ s-HC((T1,q(N −1,k), . . . , (Tp,q(N −1,k))) and Tj,q(l,k)xN → 0 as k → ∞ for all
l ≥ N and all j ∈ {1, . . . , p}, where G0 := X and GN := X \ span {x1, . . . , xN −1}
(N ∈ N) if the assumptions of (b) hold, while {GN }N ∈N denotes any fixed open
basis of X if the assumptions of (c) hold. Then M := span {xN : N ∈ N} is the
sought-after vector subspace. The details are left as an exercise.
(cid:3)
3. s-Universality criteria
A number of workable sufficient conditions will be useful to detect s-universality.
Recall that a sequence of continuous mappings Tn : X → Y (n ∈ N) is called mix-
ing provided that,given nonempty open sets U ⊂ X, V ⊂ Y , there is N ∈ N such
that Tn(U) ∩ V 6= ∅ for all n ≥ N. The corresponding notion of simultaneous
mixing property arises naturally, as well as the one of simultaneous transitivity.
Note that Tn(U) ∩ V 6= ∅ is equivalent to U ∩ T −1
n (V ) 6= ∅.
Definition 3.1. Let p ∈ N and X, Y be Hausdorff topological spaces. Assume
that, for each j ∈ {1, . . . , p}, Tj,n : X → Y (n ∈ N) is a sequence of continuous
mappings. Let also T1, . . . , Tp : X → X be continuous mappings. We say that:
(a) The sequences (T1,n), . . . , (Tp,n) are simultaneously transitive (or s-transitive)
provided that, for every pair of nonempty open sets U ⊂ X, V ⊂ Y , there
j=1 T −1
j,N (V ) 6= ∅.
is N ∈ N such that U ∩Tp
is N ∈ N such that U ∩Tp
(b) The sequences (T1,n), . . . , (Tp,n) are simultaneously mixing (or s-mixing)
provided that, for every pair of nonempty open sets U ⊂ X, V ⊂ Y , there
j=1 T −1
j,n (V ) 6= ∅ for all n ≥ N.
(c) The mappings T1, . . . , Tp are simultaneously transitive (simultaneously mix-
p ) are s-transitive (s-mixing,
1 ), . . . , (T n
ing, resp.) whenever the sequences (T n
resp.).
j=1 T −1
j=1 T −1
Remark 3.2. Corresponding concepts of d-transitivity and d-mixing were intro-
j,n (Vj) (Vj nonempty open subsets of Y , j = 1, . . . , p)
j,n (V ). Also, most criteria given in this section have their
counterparts for the related d-properties as provided in [4] and [11]. A thorough
study of d-mixing operators is provided in [8].
duced in [11], where Tp
appears instead of Tp
Note that, contrary to the one-sequence case, the facts U ∩Tp
j=1 Tj,N (U) ∩ V 6= ∅ are not equivalent. Observe also that Tp
and Tp
j,N (V ) 6= ∅
j,n (V ) =
j=1 T −1
j=1 T −1
SIMULTANEOUS UNIVERSALITY
7
[T1,n, . . . , Tp,n]−1(V p). From the definitions, it is easy to check that the sequences
(T1,n), . . . , (Tp,n) are s-mixing if and only if, for every strictly increasing sequence
(nk) in N, the sequences (T1,nk), . . . , (Tp,nk) are s-transitive. The following propo-
sition provides what can be called the Birkhoff s-transitivity theorem.
Proposition 3.3. Under the same assumptions and terminology as in Definition
3.1, let us suppose, in addition, that X is Baire and Y is second-countable. Then
we have:
(i) The sequences (T1,n), . . . , (Tp,n) are s-transitive if and only if they are densely
s-universal. If this is the case, then the set s-U((T1,n), . . . , (Tp,n)) is residual
in X.
(ii) The sequences (T1,n), . . . , (Tp,n) are s-mixing if and only if, for every strictly
increasing sequence (nk) ⊂ N, the sequences (T1,nk), . . . , (Tp,nk) are densely
s-universal.
Proof. Part (ii) is an immediate consequence of (i). Let us prove (i). Fix a
countable open basis (Vm) of Y , as well as a point x0 ∈ X. Then x0 ∈ s-
U((T1,n), . . . , (Tp,n)) if and only if, given a nonempty open set V ⊂ Y , there is
j,n (V ). Since each V
contains some Vm and each Vm is a nonempty subset of Y , the last property is the
n ∈ N with [T1,n, . . . , Tp,n]x0 ∈ V p, that is, x0 ∈Sn∈NTp
same as x0 ∈Tm∈NSn∈NTp
j,n (Vm), which shows that
j=1 T −1
j=1 T −1
s-U((T1,n), . . . , (Tp,n)) = \m∈N [n∈N
p\j=1
T −1
j,n (Vm).
(1)
j=1 T −1
Since the Tj,n's are continuous, each setTp
are s-transitive then every set Sn∈NTp
j=1 T −1
j,n (Vm) is open. If (T1,n), . . . , (Tp,n)
j,n (Vm) (m ∈ N) is (open and) dense.
Hence their (countable) intersection, which equals s-U((T1,n), . . . , (Tp,n)) by (1), is
a dense Gδ subset (so residual) in X because X is Baire. Conversely, assume that
the set of s-universal elements is dense in X and fix a nonempty open subset V of
j,n (Vm)
j,n (V ) is also dense. But this means
j,N (V ) 6= ∅
(cid:3)
Y . Then there is m ∈ N with V ⊃ Vm. It follows from (1) thatSn∈NTp
is dense in X, so the bigger set Sn∈NTp
that, given a nonempty set U ⊂ X, there is N ∈ N such that U ∩Tp
or, in other words, the sequences (T1,n), . . . , (Tp,n) are s-transitive.
j=1 T −1
j=1 T −1
j=1 T −1
In the linear case, we state the following set of sufficient conditions, that are
inspired by the results contained in [19, Sect. 1c] and the references cited in it.
Theorem 3.4. Let X and Y be topological vector spaces such that X is Baire and
Y is metrizable and separable, and let (Tj,n)n∈N (j = 1, . . . , p) be sequences in
L(X, Y ). Assume that there are respective dense subsets X0 of X and Y0 of Y
satisfying at least one of the following conditions:
8
BERNAL AND JUNG
(A) For every pair of vectors x ∈ X0, y ∈ Y0, there exist sequences (nk) ⊂ N
and (xk) ⊂ X with xk → 0, Tj,nkx → 0 and Tj,nkxk → y (j = 1, . . . , p)
as k → ∞.
(B) For every x ∈ X0, the sequences (Tj,nx)n∈N (j = 1, . . . , p) converge in Y
to a common limit and, for every y ∈ Y0, there exist sequences (nk) ⊂ N
and (xk) ⊂ X with xk → 0 and Tj,nkxk → y (j = 1, . . . , p) as k → ∞.
(C) For every x ∈ X0, there exists a sequence (nk) ⊂ N such that the sequences
(Tj,nkx)k∈N (j = 1, . . . , p) converge in Y to a common limit and, for every
y ∈ Y0, there exists a sequence (xn) ⊂ X such that xn → 0 and Tj,nxn → y
(j = 1, . . . , p) as n → ∞.
Then (Tj,n)n∈N (j = 1, . . . , p) are densely s-hypercyclic.
Proof. According to Proposition 3.3, we should show that (Tj,n)n∈N (j = 1, . . . , p)
are s-transitive. With this aim, fix a pair of nonempty open sets U ⊂ X, V ⊂ Y .
We should exhibit an N ∈ N such that U ∩Tp
j=1 T −1
j,N (V ) 6= ∅.
Assume first that (A) holds. By density, there are x ∈ X0 and y ∈ Y0 such
that x ∈ U and y ∈ V . Define A := U − x and B := V − y. Then A and B
are open neighborhoods of 0 in X and Y respectively. Take a 0-neighborhood
C ⊂ Y satisfying C + C ⊂ B. Consider the sequences (nk) and (xk) provided
by (A). Then there is k ∈ N such that xk ∈ A, Tj,nkx ∈ C and Tj,nkxk ∈ y + C
(j = 1, . . . , p). Let u := x + xk and N := nk. We get u ∈ x + A = U and
Tj,N u = Tj,N x + Tj,N xk ∈ C + y + C ⊂ y + B = V (j = 1, . . . , p), so that
u ∈ U ∩Tp
j=1 T −1
j,N (V ).
Suppose now that (B) holds. By density, there is x ∈ X0 such that x ∈ U.
Define A := U − x, a neighborhood of 0. By hypothesis, there is z ∈ Y such that
Tj,n → z as n → ∞ (j = 1, . . . , p). Since Y0 is dense in Y , there is y ∈ Y0 with
y ∈ z + V . Let B := V − y + z, a neighborhood of 0 in Y . Take a 0-neighborhood
C ⊂ Y satisfying C + C ⊂ B. We have that Tj,n ∈ z + C (j = 1, . . . , p) for n ≥ n0,
say. Consider the sequences (nk) and (xk) provided by (B) for the vector y − z, so
that xk → 0 and Tj,nkxk → y − z (j = 1, . . . , p) as k → ∞. Choose k ∈ N so large
that nk ≥ n0, xk ∈ A and Tj,nkxk ∈ y − z + C (j = 1, . . . , p). Let u := x + xk and
N := nk. Then u ∈ x + A = U and, for every j = 1, . . . , p,
Tj,N u = Tj,N x + Tj,N xk ∈ z + C + y − z + C = y + C + C ⊂ y + B = V,
so that u ∈ U ∩Tp
similar and left as an exercise.
j=1 T −1
j,N (V ), as required. Under assumption (C), the proof is
(cid:3)
Two of the most popular criteria of hypercyclicity are the so-called blow-up/col-
lapse criterion and the hypercyclicity criterion (see [2, 20, 21]). Now, we can obtain
their respective s-versions.
Proposition 3.5. [s-Blow-up/Collapse Criterion] Let X be a Baire metrizable
separable topological vector space, and let (Tj,n)n∈N (j = 1, . . . , p) be sequences
SIMULTANEOUS UNIVERSALITY
9
in L(X). Suppose that, for every nonempty open subsets U, V of X and every
0-neighborhood W ⊂ X there is N ∈ N such that
W ∩
p\j=1
T −1
j,N (V ) 6= ∅ 6= U ∩
p\j=1
T −1
j,N (W ).
Then (Tj,n)n∈N (j = 1, . . . , p) are densely s-hypercyclic.
Proof. Fix a pair of nonempty open sets U, V ⊂ X. Choose vectors x ∈ U, y ∈ V .
It suffices to exhibit sequences sequences (nk) ⊂ N and (xk) ⊂ X with xk → x
and Tj,nkxk → y (j = 1, . . . , p), because this would entail the existence of some
(V ).
In other words, the sequences (Tj,n)n∈N (j = 1, . . . , p) would be s-transitive, hence
densely s-hypercyclic by Proposition 3.3.
k ∈ N such that xk ∈ U and Tj,nkxk ∈ V (j = 1, . . . , p), so xk ∈ U ∩Tp
j=1 T −1
j,nk
With this aim, choose a fundamental decreasing sequence (Wk) of 0-neighbor-
hoods. Then (Uk) := (x + Wk) and (Vk) := (y + Wk) are fundamental decreasing
sequences of x-neighborhoods and y-neighborhoods, respectively. By hypothesis,
for each k ∈ N, there are nk ∈ N and points x′
k ∈ Wk ∩
k. Then xk → x
k → x). Finally,
k ∈ Vk and
(cid:3)
(Vk) and x′′
as k → ∞ because x′
Tj,nkxk = Tj,nkx′
Tj,nkx′′
k → y + 0 = y (j = 1, . . . , p) because Tj,nkx′
j=1 T −1
j,nk
k → 0) and x′′
k ∈ Uk ∩Tp
k ∈ Wk (so x′
k + Tj,nkx′′
k ∈ Wk for all k ∈ N.
j=1 T −1
j,nk
Tp
k and x′′
(Wk). Let xk := x′
k such that x′
k + x′′
k ∈ Uk (so x′′
Recall that the convex hull conv(A) of a subset A of a vector space X is the
least convex subset of X containing A.
Definition 3.6. Let X be a Baire metrizable separable locally convex space,
(nk) ⊂ N be a strictly increasing sequence and Tj ∈ L(X) (j = 1, . . . , p). We
say that T1, . . . , Tp satisfy the s-hypercyclicity criterion with respect to (nk) if
there are subsets X0 ⊂ X, W0 ⊂ X p such that X0 is dense in X and
as well as mappings Rk : W0 → X (k ∈ N) such that
W0 ⊃ ∆(X p)
j → 0 pointwise on X0 as k → ∞ (j = 1, . . . , p),
(i) T nk
(ii) Rk → 0 pointwise on W0 as k → ∞ and
(iii) For every w = (w1, . . . , wp) ∈ W0 and every j ∈ {1, . . . , p} there is yj ∈
conv({w1, . . . , wp}) such that T nk
j Rkw → yj as k → ∞.
Theorem 3.7. [s-Hypercyclicity Criterion] Let X be a Baire metrizable separa-
ble locally convex space and Tj ∈ L(X) (j = 1, . . . , p). If T1, . . . , Tp satisfy the
s-hypercyclicity criterion with respect to some (nk) ⊂ N, then (T nk
p ) are
s-mixing. In particular, T1, . . . , Tp are densely s-hypercyclic.
1 ), . . . , (T nk
Proof. Let U, V ⊂ X be nonempty open sets. Then there are x0 ∈ U ∩ X0 and
y0 ∈ V . By local convexity, there is a convex open set eV with y0 ∈ eV ⊂ V . As
10
BERNAL AND JUNG
(y0, . . . , y0) ∈ ∆(X p) ⊂ W0, one can find w = (w1, · · · , wp) ∈ W0 such that wj ∈ eV
for all j = 1, . . . , p. Put
zk := x0 + Rkw (k ∈ N).
Then, due to (ii), zk → x0 + 0 = x0 ∈ U as k → ∞. Moreover, for every
j ∈ {1, . . . , p} we get thanks to (i) and (iii) that
T nk
j zk = T nk
j x0 + T nk
j Rkw −→ 0 + yj = yj as k → ∞,
there is k0 ∈ N such that, for all k ≥ k0, we have zk ∈ U and T nk
(V ) 6= ∅, as required.
where, for each j, yj ∈ conv({w1, . . . , wp}) ⊂ conv(eV ) = eV ⊂ V . Consequently,
(j = 1, . . . , p) or, in other words, zk ∈ U ∩Tp
j zk ∈ V
(cid:3)
Remarks 3.8. 1. Examples of spaces X satisfying the assumptions of Theorem
3.7 are the Fr´echet spaces, that is, the locally convex F-spaces. If local convexity
is dropped from the assumptions, then the conclusion still holds if we replace (iii)
by the (stronger) condition:
(iii') T nk
j Rkw → wj as k → ∞ for every w = (w1, . . . , wp) ∈ W0 and every
j ∈ {1, . . . , p}.
j=1 T −nk
j
2.
In [11, Proposition 2.6] the following d-hypercyclicity criterion was proved,
where X is a Fr´echet space, (nk) ⊂ N is a strictly increasing sequence and Tj ∈
L(X) (j = 1, . . . , p): Assume that there exist dense subsets X0, X1, . . . , Xp ⊂ X
and mappings Sk,j : Xj → X (k ∈ N; 1 ≤ j ≤ p) satisfying T nk
j → 0 (k → ∞)
pointwise on X0, Sk,j → 0 (k → ∞) pointwise on Xj, and T nk
j Sk,l → δj,lidXl
(k → ∞) pointwise on Xl (1 ≤ j, l ≤ p). Then (T nk
p ) are d-mixing (see
[11, Definition 2.1]). In particular, by [11, Proposition 2.3], T1, . . . , Tp are densely
d-hypercyclic.
1 ), . . . , (T nk
Now, we can obtain a disjoint hypercyclicity criterion under weaker assump-
tions. Namely, let us assume that there are dense subsets X0 ⊂ X, W0 ⊂ X p and
mappings Rk : W0 → X (k ∈ N) satisfying (i) -- (ii) of Definition 3.6 together with
(iii') of the preceding remark (it is easy to check that these assumptions are weaker
than those of the d-hypercyclicity criterion in [11]). Then (T nk
p ) are d-
mixing. Indeed, let U, V1, . . . , Vp ⊂ X be nonempty open sets. By density, there
are x0 ∈ U ∩ X0 and w = (w1, · · · , wp) ∈ W0 ∩ (V1 × · · · × Vp). Let zk := x0 + Rkw
(k ∈ N). Then zk → x0 +0 = x0 ∈ U as k → ∞. Moreover, for every j ∈ {1, . . . , p}
we get T nk
j Rkw −→ 0 + wj = wj as k → ∞. Consequently, there
is k0 ∈ N such that, for all k ≥ k0, we have zk ∈ U and T nk
j zk ∈ Vj (j = 1, . . . , p),
1 ), . . . , (T nk
j zk = T nk
j x0 + T nk
(Vj) 6= ∅, which is the d-mixing property.
that is, zk ∈ U ∩Tp
j=1 T −nk
j
3. Several sets of conditions on T1, . . . , Tp such that these operators satisfy the
s-hypercyclicity criterion with respect to a strictly increasing sequence (nk) ⊂ N
are -- as it is easy to check -- the following:
SIMULTANEOUS UNIVERSALITY
11
T nk
l=1 Sk,l → idY0 pointwise on Y0 (j = 1, . . . , p).
(a) There are dense subsets X0, Y0 ⊂ X and mappings Sk,j : Y0 → X (k ∈
j=1 Sk,j → 0 pointwise on Y0 and
N; 1 ≤ j ≤ p) such that (i) holds, Pp
j Pp
N; 1 ≤ j ≤ p) such that (i) holds, Pp
(b) There are subsets X0, X1, . . . , Xp ⊂ X in such a way that X0 is dense in
X and X1 × · · · × Xp ⊃ ∆(X p) as well as mappings Sk,j : Xj → X (k ∈
j=1 Sk,jxj → 0 for all (x1, . . . , xp) ∈
l=1 Sk,lxl) → xj for all (x1, . . . , xp) ∈ X1 ×· · ·×Xp
X1 ×· · ·×Xp, and T nk
and all j = 1, . . . , p.
j (Pp
In view of (b), we see that if T1, . . . , Tp satisfy the d-hypercyclicity criterion with
respect to (nk), then they also satisfy the s-hypercyclicity criterion with respect
to (nk).
B`es and Peris [10] have proved that satisfaction of the hypercyclicity criterion,
hereditary hypercyclicity and transitivity of self-sums are equivalent (see also [5]).
Moreover, they established a similar result for d-hypercyclicity [11, Theorem 2.7].
Now, we prove that a corresponding statement also holds for s-hypercyclicity, with
the d-hypercyclicity criterion replaced by the s-hypercyclicity criterion (Theorem
3.7), so showing that the latter is rather natural.
Proposition 3.9. Let X be a separable Fr´echet space and Tj ∈ L(X)
(j = 1, . . . , p). Consider the following statements:
(a) T1, . . . , Tp satisfy the s-hypercyclicity criterion.
(b) (T nk
(c) ⊕m
(d) T1 ⊕ T1, . . . Tp ⊕ Tp are s-transitive on X 2.
1 ), . . . , (T nk
k=1T1, · · · ⊕m
p ) are hereditarily densely s-hypercyclic for some (nk) ⊂ N.
k=1 Tp are s-transitive on X m for all m ∈ N.
Then we have:
(A) (a), (b) and (c) are equivalent.
(B) If there exists i ∈ {1, . . . , p} such that TiTj = TjTi for all j ∈ {1, . . . , p},
then (a), (b), (c) are equivalent to (d).
Proof. In the proof of (A), we follow closely the proof of Theorem 2.7 in [11], while
the proof of (B) runs similar as the proof of Theorem 2.3, (3) ⇒ (1), in [10].
(A) (a) ⇒ (b): T1, . . . , Tp satisfy the s-hypercyclicity criterion with respect to
some (nk) ⊂ N, so that they also satisfy it for any subsequence (mk) of (nk). By
Theorem 3.7, (T mk
) are s-mixing and therefore densely s-hypercyclic.
), . . . , (T mk
1
p
(b) ⇒ (c): Let m ∈ N be fixed and let ∅ 6= Ul, Vl ⊂ X open (l = 1, . . . , m). It
suffices to show that there exists N ∈ N such that
Ul ∩
p\j=1
T −N
j
(Vl) 6= ∅ for all l = 1, . . . , m.
(1)
Since (T mk
p
the sequences (T nk
), . . . , (T mk
1
) are densely s-hypercyclic for each subsequence (mk) of (nk),
p ) are s-mixing (cf. Proposition 3.3(ii)). Hence, for
1 ), . . . , (T nk
12
BERNAL AND JUNG
each l ∈ {1, . . . , m}, there exists k0(l) ∈ N such that Ul ∩ Tp
(Vl) 6= ∅ for
all k ≥ k0(l). Then (1) is satisfied simply by choosing N := max{k0(1), . . . , k0(m)}.
(c) ⇒ (a): Due to the assumption, we have:
j=1 T −nk
j
(∗) For every m ∈ N and every 2m-tuple U1, . . . , Um, V1, . . . , Vm of nonempty
open subsets of X there is N ∈ N arbitrarily large such that (1) holds.
Let (An)n∈N, (Bn)n∈N be bases of nonempty sets of the topology of X. For
n ∈ N, we write Wn := B(0, 1/n) (open d-balls, d being a translation-invariant
distance generating the topology of X) and An,0 := An, Bn,0 := Bn.
j=1 T −n1
j
j=1 T −n1
j
to (∗) (with m = 2), there is n1 ∈ N such that B1 ∩Tp
W1 ∩ Tp
Choose a nonempty open set A1,1 with diam(A1,1) < 1/2 and A1,1 ⊂ A1. Due
(W1) 6= ∅ and
(A1,1) 6= ∅. Thus, there exist a nonempty open set B1,1 with
diam(B1,1) < 1/2, B1,1 ⊂ B1 and T n1
j (B1,1) ⊂ W1 for all j = 1, . . . , p, as well as
a point w1,1 ∈ W1 with T n1
j w1,1 ∈ A1,1 for all j = 1, . . . , p. Now, for i = 1, 2,
choose Ai,3−i open, nonempty, such that diam(Ai,3−i) < 1/3, Ai,3−i ⊂ Ai,2−i and
A1,2 ∩ A2,1 = ∅. Due to (∗) (with m = 4), there is n2 ∈ N with n2 > n1 such that
(Ai,3−i) 6= ∅ for i = 1, 2. Thus,
there exist nonempty open sets Bi,3−i with diam(Bi,3−i) < 1/3, Bi,3−i ⊂ Bi,2−i
and T n2
j (Bi,3−i) ⊂ W2 (i = 1, 2) for all j = 1, . . . , p as well as points wi,3−i ∈ W2
(i = 1, 2) with T n2
(W2) 6= ∅ and W2 ∩Tp
j wi,3−i ∈ Ai,3−i (i = 1, 2) for all j = 1, . . . , p.
Bi,2−i ∩Tp
j=1 T −n2
j=1 T −n2
j
j
Continuing this process inductively, by using (∗) with m = 2k in step k, we ob-
tain a strictly increasing sequence (nk) ⊂ N, nonempty open sets Ai,k+1−i, Bi,k+1−i
with diam(Ai,k+1−i) < 1
k+1 and points wi,k+1−i ∈ Wk
(1 ≤ i ≤ k; k ∈ N) such that
k+1, diam(Bi,k+1−i) < 1
(i) Ai,k+1−i ⊂ Ai,k−i, Bi,k+1−i ⊂ Bi,k−i for all 1 ≤ i ≤ k, k ∈ N.
(ii) For each k ∈ N, the sets Ai,k+1−i, 1 ≤ i ≤ k, are pairwise disjoint.
(iii) T nk
(iv) T nk
j (Bi,k+1−i) ⊂ Wk (k ∈ N; 1 ≤ i ≤ k; 1 ≤ j ≤ p), and
j wi,k+1−i ∈ Ai,k+1−i (k ∈ N; 1 ≤ i ≤ k; 1 ≤ j ≤ p).
For each fixed i ∈ N, the sequences of closed sets (Ai,r)r∈N and (Bi,r)r∈N are
decreasing (due to (i)) with diam(Ai,r), diam(Bi,r) < 1
r+i. The completeness of X
implies the existence of points ai, bi ∈ X (i ∈ N) such that Tr∈N Ai,r = {ai} and
Tr∈N Bi,r = {bi}.
Put X0 := {bi : i ∈ N} ⊂ X and W0 := {ai : i ∈ N}p ⊂ X p. As ai ∈ Ai,1 ⊂
Ai,0 = Ai and bi ∈ Bi,1 ⊂ Bi,0 = Bi for all i ∈ N, we obtain that X0 is dense in X
and W0 is dense in X p. Due to (ii), we have that ai 6= ak whenever i 6= k (indeed,
if i < k, say, then ai ∈ Ai,k+1−i and ak ∈ Ak,1, but Ai,k+1−i ∩ Ak,1 = ∅). Hence,
SIMULTANEOUS UNIVERSALITY
13
for each k ∈ N, the function Rk : W0 → X given by
Rk(ai1, . . . , aip) =
is well defined. Altogether, we have:
pXl=1
1
p
0
wil,k+1−il
if k ≥ max
l=1,...,p
il
otherwise
pPp
pPp
• For all j = 1, . . . , p, all i ∈ N and all k ≥ i one has, due to (iii), that
j → 0 (k → ∞) pointwise
j (Bi,k+1−i) ⊂ Wk = B(0, 1/k), so T nk
T nk
j bi ∈ T nk
on X0 for every j = 1, . . . , p.
• For every (ai1, . . . , aip) ∈ W0 and every k ≥ maxl=1,...,p il, one has
l=1 wil,k+1−il → 0 (k → ∞), because wil,k+1−il ∈
Rk(ai1, . . . , aip) = 1
Wk = B(0, 1/k). Therefore Rk → 0 (k → ∞) pointwise on W0.
• For all j = 1, . . . , p, all (ai1, . . . , aip) ∈ W0 and all k ≥ maxl=1,...,p il, we get
T nk
j Rk(ai1, . . . , aip) = 1
j wil,k+1−il ∈ Ail,k+1−il
and the sequence of sets Ail,k+1−il (k ∈ N) collapses to the singleton
{ail} as k → ∞ for each l, we get T nk
l=1 ail ∈
conv({ai1, . . . , aip}) as k → ∞.
j wil,k+1−il. Since T nk
j Rk(ai1, . . . , aip) → 1
pPp
l=1 T nk
Thus, T1, . . . , Tp satisfy the s-hypercyclicity criterion with respect to (nk). The
proof of (A) is finished.
(B) Obviously, (c) always implies (d). Assume now that (d) holds and that some
Ti commutes with all Tj's. Our goal is to prove that (a) is satisfied.
Let us fix any vector (x0, y0) ∈ s-HC(T1 ⊕ T1, . . . , Tp ⊕ Tp). We claim that, for
each m ∈ N, the vector (x0, T m
i y0) is also s-hypercyclic for T1 ⊕ T1, . . . , Tp ⊕ Tp.
Indeed, as Ti is hypercyclic, it has dense range, from which one obtains, inductively,
i (X) is dense in X. Put A := X × T m
that every set T m
i (X), so that A is dense
in X 2. Given (u, v) ∈ A there is w ∈ X such that v = T m
i w. By s-hypercyclicity,
there exists (nk) ⊂ N such that T nk
j y0 → w (k → ∞) for all
j = 1, . . . , p. Hence, for all j, T nk
j x0 → u and, by commutativity together with
continuity of T m
j (T m
i y0) = T m
i w = v (k → ∞).
i
i y0) : n ∈ N} ⊃ ∆(Ap). Since
Therefore Σ := {[(T1 ⊕ T1)n, . . . , (Tp ⊕ Tp)n](x0, T m
∆(Ap) ⊃ ∆((X 2)p) and Σ is closed, we get Σ ⊃ ∆((X 2)p), which proves the
claim.
j x0 → u and T nk
j y0) −→ T m
, we get T nk
i (T nk
In particular, as y0 is hypercyclic for Ti, for each nonempty open set U ⊂ X
there exists some u ∈ U such that (x0, u) is s-hypercyclic for T1 ⊕ T1, . . . , Tp ⊕ Tp.
Thus, fixing a decreasing basis (Uk) of neighborhoods of 0 and using induction,
we can find for each k ∈ N some uk ∈ Uk and nk ∈ N with nk > nk−1 (where
n0 := 0) such that
(α) T nk
(β) T nk
j x0 ∈ Uk for all j = 1, . . . , p and
j uk ∈ x0 + Uk for all j = 1, . . . , p.
14
BERNAL AND JUNG
i x0 : n ∈ N} and W0 := X p
We define X0 := {T n
0 . Note that X0 is dense in X as
x0 is Ti-hypercyclic, so W0 is dense in X p (hence W0 ⊃ ∆(X p)). Now, observe that
no orbit of any hypercyclic vector can be finite, that is, T m
i x0 if m 6= n.
Thus, for each k ∈ N, the mapping
i x0 6= T n
Rk : (T m1
i x0, . . . , T mp
i x0) ∈ W0 7−→
1
p
·
pXl=1
T ml
i uk ∈ X
(1)
is well defined. We have:
(i) For every j = 1, . . . , p and every m ∈ N, T nk
j x0) →
T m
i 0 = 0 (k → ∞), where commutativity and continuity of Ti together
with property (α) have been used. This shows that T nk
j → 0 pointwise on
X0 for all j = 1, . . . , p.
i x0) = T m
i (T nk
j (T m
(ii) From the continuity of each T ml
it follows that T ml
from (1) that Rk → 0 pointwise on W0.
and the fact that uk ∈ Uk (hence uk → 0),
i uk → 0 (k → ∞) for every l = 1, . . . , p. Then one derives
i
(iii) For every j = 1, . . . , p and every (m1, . . . , mp) ∈ Np, it follows from (1)
that
T nk
j Rk(T m1
i x0, . . . , T mp
i x0) =
1
p
·
pXl=1
T nk
j T ml
i uk =
1
p
·
pXl=1
T ml
i T nk
j uk
−→
1
p
·
pXl=1
T ml
i x0 ∈ conv({T m1
i x0, . . . , T mp
i x0})
as k → ∞, because of (β) (which implies T nk
commutativity and continuity of each T ml
.
i
j uk → x0) together with the
This tells us that T1, . . . , Tp satisfy the s-hypercyclicity criterion, as required. (cid:3)
We raise here the question whether (d) is equivalent to (a)-(b)-(c) without as-
suming any commutativity.
4. Scalar multiples of an operator
We start by studying s-hypercyclicity of scalar multiples of one operator. We
have already pointed out that there is no chance of d-hypercyclicity in this case.
Recall that an operator T on a topological vector space X is called hereditarily
hypercyclic whenever (T nk) is universal for every strictly increasing sequence (nk) ⊂
N. It is well known -- and easy to see -- that, if X is an F-space and T ∈ L(X), then
T is hereditarily hypercyclic if and only if T is mixing.
Proposition 4.1. Let X be a topological vector space, p ∈ N, c1, . . . , cp ∈ K and
T, T1, . . . , Tp ∈ L(X). We have:
SIMULTANEOUS UNIVERSALITY
15
(a) Assume that X is metrizable and locally convex.
If T, c1T, . . . , cpT are
s-hypercyclic then the cj's are unimodular, that is, c1 = · · · = cp = 1.
(b) Suppose that X is metrizable.
If T ∈ L(X) is hereditarily hypercyclic
and the scalars cj are unimodular, then T, c1T, . . . , cpT are densely s-
hypercyclic.
Proof. (a) Assume that T, c1T, . . . , cpT are s-hypercyclic, and fix j ∈ {1, . . . , p}.
Let c := cj. Then T, cT are s-hypercyclic, so there is x0 ∈ s-HC(T, cT ). Since
X is metrizable, we can find a sequence (nk) ⊂ N such that T nkx0 → x0
and cnkT nkx0 → x0 as k → ∞. Of course, x0 6= 0. But X is locally convex,
so its topology is defined by a separating family of seminorms. Therefore there is
a continuous seminorm q on X such that q(x0) > 0. Consider the sequence of
vectors
uk := (cnk − 1)T nkx0
(k ∈ N).
q(T nk x0) → 0
On the one hand, we have uk = cnkT nkx0 − T nkx0 → x0 − x0 = 0, so q(uk) → 0 by
the continuity of q. On the other hand, we get q(uk) = cnk − 1q(T nkx0), hence
cnk − 1 = q(uk)
q(x0) = 0. Therefore cnk → 1 as k → ∞, which implies
c = 1, that is, cj = 1, as required.
(b) The result is trivial if K = R (for cj = ±1, so (cjT )2n = T 2n for all n and all
j = 1, . . . , p). The complex case K = C is more delicate. Recall that a subset
E ⊂ T := {z = 1} is said to be a Dirichlet set provided that there is a strictly
increasing sequence (nk) ⊂ N such that supz∈E znk − 1 → 0 as k → ∞. It is
well-known that every finite subset of T is Dirichlet (see [13, Theorem 8.138(a)]).
In particular, there exists (nk) ⊂ N strictly increasing such that cnk
j → 1 (j =
1, . . . , p). According to the hypothesis, we may take x0 ∈ HC((T nk)). Given
x ∈ X, there is a subsequence (mk) ⊂ (nk) such that T mkx0 → x and, of course,
cmk
j → 1 (j = 1, . . . , p) as k → ∞. Therefore, we obtain (cjT )mk x0 → x for all
j = 1, . . . , p and hence HC((T nk)) ⊂ s-HC(T, c1T, . . . , cpT ). But HC((T nk)) is
dense, so s-HC(T, c1T, . . . , cpT ) also is.
(cid:3)
Remarks 4.2. 1. In part (b) of the last proposition, hereditary hypercyclicity is
needed in order to obtain common subsequences (nk) to perform approximations.
If this is not claimed, then, by a result due to Le´on and Muller, any unimodular
multiple of a hypercyclic operator on any topological vector space is always hyper-
cyclic, even with the same set of hypercyclic vectors (see [23] and [21, pp. 339 -- 340]).
2. It is known that the d-mixing property of T1, . . . , Tp implies that c1T1, . . . , cpTp
are also d-mixing for all unimodular scalars c1, . . . , cp (cf. [7, Remark 24(i)]). How-
ever, the corresponding result in case of s-mixing operators does not hold in gen-
eral. Indeed, for a mixing operator T , the pair T, T is clearly s-mixing, but T, −T
are not s-mixing any more. To see this, assume that T, −T are s-mixing. Then
Proposition 3.3(ii) would imply that (T nk), ((−T )nk) are densely s-universal for
each strictly increasing sequence (nk) in N -- but s-universality of the sequences
16
BERNAL AND JUNG
(T 2k+1) and ((−T )2k+1) = (−T 2k+1) is clearly not possible.
In connection with
this, it is stated in [7, Remark 24(ii)] and actually proved in [28, Proposition
4.9] that in case of unimodular scalars c1, . . . , cp every d-hypercyclic vector x0 for
T1, . . . , Tp is also d-hypercyclic for c1T1, . . . , cpTp. The proof uses crucially the fact
that such a vector x0 satisfies (x0, . . . , x0) ∈ HC(T1 ⊕ · · · ⊕ Tp). Thus, it cannot
be adapted for s-hypercyclicity. Hence, we pose the question: Does the equality
s-HC(T1, . . . , Tp) = s-HC(c1T1, . . . , cpTp) hold?
3. Concerning again part (b) and regarding its proof, we may obtain a much
stronger result in the case K = C and X a Banach space. Recall that a
nonempty subset E ⊂ C is said to be perfect if it is closed and each point
of E is an accumulation point of E.
In particular, every perfect set is un-
countable.
It is well known (see [13, Theorem 8.138(b)]) that there are perfect
Dirichlet subsets of T. We have that if E ⊂ T is a perfect Dirichlet set and
T ∈ L(X) is mixing, then the uncountable family of rotations {cT : c ∈ E ∪ {1}}
is densely uniformly s-hypercyclic, in the sense that there is a dense set of vectors
x0 ∈ X satisfying the following: for every y ∈ X there is (nk) ⊂ N such that
limk→∞ supc∈E∪{1} k(cT )nkx0 − yk = 0. Indeed, we can take a sequence (mk) ⊂ N
such that supc∈E∪{1} cmk − 1 = supc∈E cmk − 1 → 0 as k → ∞. As T is
mixing, the set HC((T mk)) is dense. If x0 ∈ HC((T mk)), then there is a subse-
quence (nk) ⊂ (mk) with T mk x0 → y. The conclusion follows from the inequality
k(cT )nkx0 − yk ≤ kcnk(T nkx0 − y)k + k(cnk − 1)yk.
4. Proposition 4.1 furnishes examples of pairs of operators -- on spaces of sequences
or of holomorphic functions (see sections 5 -- 6) -- that are s-hypercyclic but not d-
hypercyclic: the multiples 2B, −2B of the backward shift B on ℓq (1 ≤ q < ∞) or
c0; D, −D on H(C) (Df := f ′); Cϕ, −Cϕ on H(G), where Cϕf := f ◦ ϕ, G ⊂ C is
a simply connected domain and ϕ is a run-away automorphism of G.
5. Backward shifts and s-hypercyclicity
In this section, we consider the sequence spaces c0 and ℓq (1 ≤ q < ∞) over
K = R or C. If a = (an)n∈N is a bounded sequence in K \ {0}, then Ba will denote
the weighted backward shift
Ba : (x0, x1, x2, . . . ) ∈ X 7→ (a1x1, a2x2, . . . ) ∈ X
on X = c0 or ℓq. The unweighted backward shift B is B = Ba, where a =
(1, 1, 1, . . . ). Salas characterized the hypercyclicity of Ba in terms of the weight
sequence a. B`es and Peris [11, Theorem 4.1] did the same for the d-hypercyclicity
of different powers of Ba. This characterization happens to hold also for s-
hypercyclicity.
Proposition 5.1. Let X = c0 or ℓq (1 ≤ q < ∞), p ≥ 2 and let r1, . . . , rp ∈ N
with r1 < r2 < · · · < rp be given. For each l ∈ {1, . . . , p}, let al = (al,n)n∈N be a
weight sequence. Then the following are equivalent:
SIMULTANEOUS UNIVERSALITY
17
a1, . . . , Brp
a1, . . . , Brp
al,j+1 ··· al,j+rlm
ap are d-hypercyclic.
ap are s-hypercyclic.
(i) Br1
(ii) Br1
(iii) For every M > 0 and every k ∈ N there is m ∈ N satisfying, for
each j ∈ {0, 1, . . . , k}, that al,j+1 · · · al,j+rlm > M (1 ≤ l ≤ p) and
as,j+(rl−rs)m+1 ··· as,j+rlm > M (1 ≤ s < l ≤ p).
a1, . . . , Brp
a1, . . . , Brp
ap satisfy the d-hypercyclicity criterion.
ap satisfy the s-hypercyclicity criterion.
(iv) Br1
(v) Br1
Proof. The equivalence of (i), (iii) and (iv) is proved in [11, Theorem 4.1]. That
(i) implies (ii) is trivial. Moreover, (ii) ⇒ (iii) is proved in fact in the proof of "(a)
⇒ (b)" of the same reference, since only the simultaneous approximation of one
vector (namely e0 + · · · + eq) is used. Finally, we clearly have (iv) ⇒(v) ⇒ (ii). (cid:3)
Remarks 5.2. 1. An analogous result about equivalence of d- and s-hypercyclicity
also works for powers of weighted bilateral shifts (see Theorem 4.7 of [11] and its
proof).
a, . . . , Bp
a ⊕ · · · ⊕ Bp
2. Corollary 4.4 in [11] also works with just s-universality, as it is a consequence
of Theorem 4.1 there. In particular, we have that Ba, B2
a are s-hypercyclic
on X if and only if Ba ⊕ B2
a is hypercyclic on X p. B`es, Martin and Peris
[7, p. 855] constructed an operator T := Ba on ℓ2 such that T is hypercyclic but
T ⊕ T 2 is not hypercyclic on ℓ2 ⊕ ℓ2, so that T, T 2 is not d-hypercyclic on ℓ2. Then
we obtain that T, T 2 are even not s-hypercyclic. According to [21, Theorem 4.8],
the mentioned T = Ba is not mixing. In [8, Sect. 3], a mixing operator T ∈ L(ℓ2)
for which T, T 2 are not d-mixing is exhibited. But the existence of a mixing T on
a separable Banach space such that T, T 2 are not d-hypercyclic is unknown so far
[8, Question 3.7].
A more delicate question arises when r1 ≤ r2 ≤ · · · ≤ rp.
In [11, Corol-
lary 4.2], the following is proved for weighted powers of the unweighted backward
shift:
if p ≥ 2 and rl ∈ N, λl ∈ K (1 ≤ l ≤ p) with r1 ≤ r2 ≤ · · · ≤ rp,
then λ1Br1, . . . , λpBrp are d-hypercyclic if and only if r1 < r2 < · · · < rp and
1 < λ1 < λ2 < · · · < λp. The following result shows that s-hypercyclicity is
possible under slightly weaker assumptions.
Proposition 5.3. Let p ≥ 2, and let rl ∈ N, λl ∈ K (1 ≤ l ≤ p) with r1 ≤ r2 ≤
· · · ≤ rp. Let A denote the set A := {j ∈ {1, . . . , p − 1} : rj = rj+1} and consider
the conditions
(i) 1 < λj for all j ∈ {1, . . . , p},
(ii) λj < λj+1 for all j ∈ {1, . . . , p − 1} \ A,
(iii) λj = λj+1 for all j ∈ A.
Then λ1Br1, . . . , λpBrp are s-hypercyclic on X = c0 or ℓq (1 ≤ q < ∞) if and only
if (i),(ii) and (iii) hold.
18
BERNAL AND JUNG
Proof. First, suppose that conditions (i),(ii) and (iii) hold. We write {1, . . . , p}\A =
{t1, . . . , td}, with d ∈ N and t1 < · · · < td. As the set {λi/λj : i, j ∈ {1, . . . , p}
with λi = λj} ⊂ T is finite, it is a Dirichlet set. Hence there exists a strictly
increasing sequence (nk) ⊂ N such that
→ 1 (k → ∞) for all i, j ∈ {1, . . . , p} with λi = λj.
(1)
(cid:18) λi
λj(cid:19)nk
Consider the set X0 of finite sequences, that is, X0 := c00 = {x = (xn) ∈ X :
exists n0 = n0(x) ∈ N such that xn = 0 for all n ≥ n0}. Then X0 is dense in
X. If we set W0 := ∆(X p
0 ) ⊃ ∆(X p) because X0 is
dense in X. Now, we set Tj := λjBrj (j = 1, . . . , p). Define, for each k ∈ N, the
mapping Rk : W0 → X as follows. If x = (x1, x2, . . . , xN , 0, 0, 0, . . . ) ∈ X0 and
w = (x, x, . . . , x), then
0 ) ⊂ X p, then W0 = ∆(X p
(0, 0, 0, . . . )
Rkw =(cid:26) (01, u1, 02, u2, . . . , 0N , uN , 0, 0, 0, . . . )
xN(cid:1) (l ≥ 1). We have:
x1, . . . , 1
nk
λ
tl
if l ≥ 2 and ul :=(cid:0) 1
λ
nk
tl
if nk ≥ N
if nk < N,
where 01 := (0, 0, . . . , 0) [rt1nk times], 0l := (0, 0, . . . , 0) [(rtl − rtl−1)nk − N times]
(a) For each j ∈ {1, . . . , p} and each x = (x1, x2, . . . , xN , 0, 0, 0, . . . ) ∈ X0,
T nk
j x = 0 as soon as rjnk > N, so T nk
j → 0 (k → ∞) pointwise on X0.
(b) For every w = (x, . . . , x) ∈ W0 as before, the definition of Rk together
with (i) yields Rkw → 0 as k → ∞.
(c) Fix w = (x, . . . , x) ∈ W0, where x = (x1, x2, . . . , xN , 0, 0, 0, . . . ). For every
j ∈ {1, . . . , p} there is exactly one l ∈ {1, . . . , d} such that λj = λtl, due
to (ii) and (iii). Finally, if nk ≥ N, we have
T nk
j Rkw =(cid:18)(cid:0) λj
λtl(cid:1)nk x1,(cid:0) λj
(cid:0) λj
λtl+1(cid:1)nk x1,(cid:0) λj
λtd(cid:1)nk x1,(cid:0) λj
(cid:0) λj
λtl(cid:1)nkx2, . . . ,(cid:0) λj
λtl+1(cid:1)nkx2, . . . ,(cid:0) λj
λtd(cid:1)nkx2, . . . ,(cid:0) λj
λtl(cid:1)nk xN , 0, 0, . . . , 0,
λtl+1(cid:1)nkxN , 0, 0, . . . , 0, . . . ,
λtd(cid:1)nk xN , 0, 0, 0, 0, . . .(cid:19).
It follows from (ii) that ( λj
λts
all ν ∈ {1, . . . , N}, while (1) entails that ( λj
λtl
ν ∈ {1, . . . , N}. Consequently, T nk
)nkxν → 0 as k → ∞ for all s ∈ {l+1, . . . , d} and
)nkxν → xν as k → ∞ for all
j Rkw → (x1, x2, . . . , xN , 0, 0, 0, . . . ) = x.
An application of the s-hypercyclicity criterion (see also Remark 3.8.1) concludes
the first part of the proof.
Now, suppose that λ1Br1, . . . , λpBrp are s-hypercyclic. Since hypercyclic opera-
tors on normed spaces have norm larger than 1, we obtain
1 < kλjBrj k = λjkBrj k = λj
SIMULTANEOUS UNIVERSALITY
19
for all j = 1, . . . , p (cf. the proof of Corollary 4.2 in [11]), i.e. condition (i) holds.
For each j ∈ {1, . . . , p − 1}\A, we have rj < rj+1. Hence, as λjBrj , λj+1Brj+1 are
s-hypercyclic, Proposition 5.1, (ii) ⇒ (iii), and the same approach as in the proof
of Corollary 4.2 in [11] yield λj < λj+1, i.e. condition (ii) holds. Finally, for
each j ∈ A, we have rj = rj+1. Hence, the s-hypercyclicity of
λjBrj , λj+1Brj+1 =
λj+1
λj
· λjBrj
implies λj+1/λj = 1 (see Proposition 4.1(a)) and thus λj = λj+1, i.e. condition
(iii) holds.
(cid:3)
For instance, the operators 2B, 3B2, −3B2, being not d-hypercyclic, are s-
hypercyclic. Further study of d-hypercyclicity of weighted unilateral and bilateral
backward shifts can be found in [9].
6. s-hypercyclicity in spaces of holomorphic functions
Let G ⊂ C be a domain, that is, a nonempty connected open subset of C. We
endow the space H(G) of all holomorphic (or analytic) functions G → C with the
topology of uniform convergence on compacta, so that H(G) becomes a separable
Fr´echet space. In this section we are concerned with s-hypercyclicity of finite sets
of operators on H(G) (or on subspaces of it) for certain domains G.
Recall that if X is a topological vector space and T ∈ L(X), then T is said
to be supercyclic provided that there exists some x0 ∈ X whose projective orbit
{λT nx0 : n ∈ N, λ ∈ K} is dense in X. If T1, . . . , Tp ∈ L(X), they are called d-
supercyclic (see [7]) if there is x0 ∈ X such that {λ[T n
p ]x0 : n ∈ N, λ ∈ K}
is dense in X p. Consistently, we say that T1, . . . , Tp are s-supercyclic whenever
{λ[T n
p ]x0 : n ∈ N, λ ∈ K} ⊃ ∆(X p).
1 , . . . , T n
1 , . . . , T n
Let LF T (D) denote the family of all linear fractional transformations ϕ(z) =
az+b
cz+d of the complex plane such that ϕ(D) ⊂ D. The subfamily Aut(D) of automor-
phisms of D consists of all onto members of LF T (D). See e.g. [27, Chapter 1] for
terminology related to these families. If ν ∈ R, then Sν denotes the weighted Hardy
space Sν = {f (z) = Pn≥0 anzn ∈ H(D) : kf k := (Pn≥0 an2(n + 1)2ν)1/2 < ∞}.
Each Sν is a Hilbert space, and the choices ν = −1/2, 0, 1/2 correspond, respecti-
vely, to the classical Bergman, Hardy and Dirichlet spaces. Thanks to the results
in [7], we obtain without effort the next two assertions.
Proposition 6.1. Let ϕ1, . . . , ϕp ∈ LF T (D) pairwise distinct. Then the following
are equivalent:
(a) Cϕ1, . . . , Cϕp are s-supercyclic on H(D).
(b) µ1Cϕ1, . . . , µpCϕp are s-mixing on H(D) for all nonzero scalars µ1, . . . , µp.
(c) Cϕ1, . . . , Cϕp are d-supercyclic on H(D).
(d) µ1Cϕ1, . . . , µpCϕp are d-mixing on H(D) for all nonzero scalars µ1, . . . , µp.
20
BERNAL AND JUNG
(e) ϕ1 . . . , ϕp have no fixed point in D, and satisfy that if any two ϕl, ϕj have
j(α) < 1 is not possible.
the same attractive fixed point α, then ϕ′
l(α) = ϕ′
Proof. The equivalence of (c), (d) and (e) is proved in [7, Theorem 4]. The im-
plications (d) ⇒ (b) ⇒ (a) are trivial. Finally, (a) ⇒ (e) is proved in fact in the
proof of Theorem 4 in [7]. Indeed, it is used there a result (Lemma 14 in [7]) as-
serting that if ϕ1, ϕ2 ∈ LF T (D) are hyperbolic and share an attractive fixed point
α with ϕ′
2(α), then Cϕ1, Cϕ2 are not d-supercyclic on H(D). But a closer
look at its proof shows that Cϕ1, Cϕ2 are in fact even not s-supercyclic; indeed, via
contradiction, only one function g is assumed to be simultaneously approximated
by projective orbits.
(cid:3)
1(α) = ϕ′
Proposition 6.2. Let ϕ1, . . . , ϕp ∈ LF T (D) pairwise distinct and let ν < 1/2.
Then the following are equivalent:
(a) Cϕ1, . . . , Cϕp are s-supercyclic on Sν.
(b) Cϕ1, . . . , Cϕp are s-mixing on Sν.
(c) Cϕ1, . . . , Cϕp are d-supercyclic on Sν.
(d) Cϕ1, . . . , Cϕp are d-mixing on Sν.
(e) Each ϕl is a parabolic automorphism or a hyperbolic map without fixed
points in D, and there are no two ϕl, ϕj having a common fixed point α
such that ϕ′
l(α) = ϕ′
j(α) < 1.
Proof. The equivalence of (c), (d) and (e) is proved in [7, Theorem 3]. The impli-
cations (d) ⇒ (b) ⇒ (a) are trivial. As for (a) ⇒ (e), observe that in the proof
of Theorem 3 in [7], only the supercyclicity of each Cϕl is necessary for the first
assertion in (e) and that the Comparison Principle [7, Proposition 8] -- that also
works for s-supercyclicity -- implies that Cϕ1, . . . , Cϕp are s-supercyclic on H(D).
Now, the second assertion of (e) follows from Proposition 6.1.
(cid:3)
Remarks 6.3. 1. Recall that if X is an F-space and T ∈ L(X) is invertible
and hypercyclic, then T −1 is also hypercyclic. Analogously as in Example 22
in [7], by combining the preceding two propositions, we obtain that there are
hyperbolic ϕ1, ϕ2 ∈ Aut(D) such that Cϕ1, Cϕ2 are d-hypercyclic (so s-hypercyclic)
= (Cϕ2)−1
on H 2(D) (the Hardy space) and on H(D), and Cϕ−1
are even not s-supercyclic on H 2(D) or H(D) (note that ϕ−1
are also
1
hyperbolic). Hence, in general, the d-hypercyclicity of T1, . . . , Tp does not imply
the s-hypercyclicity of T −1
if T1, . . . , Tp are invertible. Moreover, finitely
many composition operators generated by non-elliptic automorphisms of D may
be not s-hypercyclic on H(D) or on H 2(D).
= (Cϕ1)−1, Cϕ−1
and ϕ−1
2
1
, . . . , T −1
p
1
2
2. Further study of d-hypercyclicity of composition operators, this time on weighted
Bergman spaces on D, is performed in [30].
In 1929 Birkhoff [12] proved that the translation operator τa (a ∈ C \ {0}) given
by (τaf )(z) = f (z + a) is hypercyclic on the space H(C) of entire functions. It is
SIMULTANEOUS UNIVERSALITY
21
proved in [4, Prop. 5.5] and [11, Theorem 3.1] that if a1, . . . , ap are pairwise distinct
nonzero complex numbers, then τa1, . . . , τap are d-hypercyclic. Trivially, we obtain:
if a1, . . . , ap ∈ C \ {0}, then τa1, . . . , τap are s-hypercyclic. As the next proposition
shows, we may obtain a slight extension to weighted translation operators.
Proposition 6.4. Let p ≥ 2, and let a1, . . . , ap, λ1, . . . , λp ∈ C \ {0} such that
λj = λl for all j, l ∈ {1, . . . , p} with aj = al. Then there is a sequence (nk) ⊂ N
such that the sequences (λ1τa1)nk, . . . , (λpτap)nk are s-mixing. In particular, the
operators λ1τa1, . . . , λpτap are densely s-hypercyclic on H(C).
Proof. Select a finite sequence {j(1) < j(2) · · · < j(q)} ⊂ {1, . . . , p} satisfy-
ing that, if bl
:= aj(l) (l = 1, . . . , q), then the bl's are pairwise distinct and
{a1, . . . , ap} = {b1, . . . , bq}. Let µl := λj(l). Consider the operators Tj := λjτaj
(j = 1, . . . , p) and Sl := Tj(l) = µlτbl (l = 1, . . . , q).
Let us prove that S1, . . . , Sq are s-mixing. In fact, by following the approach
of the proof of [11, Theorem 3.1], we can prove that they are even d-mixing.
To this end, and taking into account that the sets V (h, r, ε) := {f ∈ H(C) :
f (z) − h(z) < ε for all z ∈ B(0, r)} (h ∈ H(C), ε > 0, r > 0), form a basis for
the topology of H(C), it is enough to prove that, for given h, g1, . . . , gq ∈ H(C)
and ε, r > 0, there is n0 ∈ N such that, for every n ≥ n0, there exists an entire
function f with
2r
f (z) − h(z) < ε and (Sn
l f )(z) − gl(z) < ε
bi−bl + max1≤l≤q
(z ∈ B(0, r), l = 1, . . . , q).
(1)
2r
Select n0 ∈ N with n0 > maxi6=l
bl . Then, for each n ≥ n0, the
disks B(0, r), B(nb1, r), . . . , B(nbq, r) are pairwise disjoint. Pick s > r such that
the disks B(0, s), B(nb1, s), . . . , B(nbq, s) are still pairwise disjoint. Let K :=
B(0, r) ∪ B(nb1, r) ∪ · · · ∪ B(nbq, r) and Ω := B(0, s) ∪ B(nb1, s) ∪ · · · ∪ B(nbq, s).
Note that Ω is an open set, Ω ⊃ K and K is a compact subset having connected
complement. Consider the function F : Ω → C defined by
F (z) = h(z) if z ∈ B(0, s) and F (z) := µ−n
l gl(z−nbl) if z ∈ B(nbl, s) (1 ≤ l ≤ q).
Then F ∈ H(Ω). From Runge's approximation theorem (see e.g. [16]), it follows
that there exists a polynomial f (so f ∈ H(C)) such that f (z) − F (z) < ε/(1 +
µn
l ) for all z ∈ K. But this implies that f (z) − h(z) < ε on B(0, r) and
µn
l f (z) − gl(z − nbl) < ε on B(nbl, r). Since the last inequality is equivalent to
µn
l f (z + nbl) − gl(z) < ε on B(0, r), (1) is obtained.
As the set D := {λj/λl : j, l ∈ {1, . . . , p} with aj = al} ⊂ T is finite, it is
a Dirichlet set. Then there is a strictly increasing sequence (nk) ⊂ N such that
ξnk → 1 as k → ∞, for all ξ ∈ D.
Fix a subsequence (mk) of (nk). Since S1, . . . , Sq are s-mixing, the set
in
))
)). For each ν ∈ {1, . . . , p} there is a unique l = l(ν) ∈
(see Proposition 3.3).
1 ), . . . , (Smk
1 ), . . . , (Smk
s-HC((Smk
s-HC((Smk
is dense
q
q
Fix
f
22
BERNAL AND JUNG
ν → 1, hence ξmk
{1, . . . , q} such that aν = bl, so that λν = µl. Observe that ξν := λν/µl ∈ D.
Then ξnk
ν → 1 (k → ∞) for all ν ∈ {1, . . . , p}. Given g ∈ H(C),
we can find a subsequence (pk) of (mk) with Spk
l(ν)f → g (k → ∞) uniformly
on compacta for every ν ∈ {1, . . . , p}. Since ξpk
ν → 1 for all ν, we obtain that
T pk
ν f = ξpk
l(ν)f −→ 1 · g = g (k → ∞) uniformly on compacta for every
ν = 1, . . . , p. Therefore f ∈ s-HC((T mk
)), which shows that this set
is dense. By Proposition 3.3, the sequences (T nk
p ) are s-mixing, as re-
quired.
(cid:3)
1 ), . . . , (T nk
ν Spk
), . . . , (T mk
p
1
Another important collection of operators on H(C) is that of differentiation
operators. Consider the derivative operator D : f ∈ H(C) 7→ f ′ ∈ H(C).
Its
It is shown in
hypercyclicity on H(C) was proved by MacLane in 1952 [25].
[11, Prop. 3.3] that if p ≥ 2, r1, . . . , rp ∈ N with r1 < · · · < rp and λ1, . . . , λp ∈
C \ {0}, then λ1Dr1, . . . , λpDrp are d-mixing, so densely d-hypercyclic. Concerning
s-hypercyclicity, the following proposition shows that somewhat softer assumptions
are allowed, although, similarly to the last proposition, we have not been able to
obtain the s-mixing property for the whole sequences.
Proposition 6.5. Let r1 ≤ · · · ≤ rp be positive integers and λ1, . . . , λp ∈ C \ {0},
where p ≥ 2. Suppose that λj = λl for all j, l ∈ {1, . . . , p} with rj = rl.
Then there is a sequence (nk) ⊂ N such that the sequences (λ1Dr1)nk, . . . , (λpDrp)nk
are s-mixing. In particular, the operators λ1Dr1, . . . , λpDrp are densely s-hypercyclic
on H(C).
Proof. As the set {λj/λl : j, l ∈ {1, . . . , p} with rj = rl} ⊂ T is finite, it is a
Dirichlet set. Then there is a strictly increasing sequence (nk) ⊂ N such that
(λj/λl)nk → 1 as k → ∞, for all j, l ∈ {1, . . . , p} with rj = rl. Put X0 :=
{polynomials} = span{zm : m ∈ N0} and W0 := ∆(X p
0 ). Then X0 is dense in
X := H(C) and W0 = ∆(X p
0 ) ⊃ ∆(X p). Let Tj := λjDrj (1 ≤ j ≤ p). For each
k ∈ N, define the map Rk : W0 → X via
Rk(zm, . . . , zm) :=
pXl=1
1
τ (l)
·
1
λnk
l
·
zm+rlnk
(m + 1)(m + 2) · · · (m + rlnk)
,
where τ (l) := card {i ∈ {1, . . . , p} : ri = rl} (1 ≤ l ≤ p). Then Rk is extended to
the whole W0 by linearity. We have:
(i) T nk
j zm = 0 as soon as nkrj > m, so T nk
{1, . . . , p} and all m ≥ 0. Therefore, by linearity, T nk
for all j ∈ {1, . . . , p}.
j zm → 0 as k → ∞ for all j ∈
j → 0 (k → ∞) on X0
SIMULTANEOUS UNIVERSALITY
23
(ii) Fix m ∈ N0 and a compact set K ⊂ C. There is M ∈ (0, +∞) with
K ⊂ B(0, M). Given k ∈ N, we obtain
Rk(zm, . . . , zm) ≤
sup
z∈K
≤
≤
pXl=1
pXl=1
pXl=1
1
τ (l)
·
1
λnk
l
·
M m+rlnk
(m + 1)(m + 2) · · · (m + rlnk)
1
M m+rlnk/λnk
l
τ (l)
(m + 1)(m + 2) · · · (m + nk)
M m
τ (l)
(M rl/λl)nk
nk!
→ 0 (k → ∞)
Hence, by linearity, Rk → 0 (k → ∞) pointwise on W0.
(iii) Fix m ∈ N0, j ∈ {1, . . . , p} and k ∈ N with nk > m. Let us compute
the action of T nk
j Rk on each (zm, . . . , zm). This yields three sums, the first
of them corresponding to those l ∈ {1, . . . , p} with rl < rj, that equals 0.
Therefore
T nk
j Rk(zm, . . . , zm) = 0 +
pXl=1
rl=rj
1
τ (l)
λl(cid:1)nk · zm
·(cid:0) λj
zm+(rl−rj )nk
(m + 1)(m + 2) · · · (m + (rl − rj)nk)
+
pXl=1
rl>rj
1
τ (l)
−→
1
τ (j)
·(cid:0) λj
λl(cid:1)nk ·
pXl=1
rl=rj
· zm ·
1 + 0 = zm (k → ∞)
uniformly on compacta in C, because τ (j) = τ (l) and ( λj
λl
(j, l) with rj = rl. By linearity again, we get T nk
j = 1, . . . , p and all (w, . . . , w) ∈ W0.
)nk → 1 for all
j Rk(w, . . . , w) → w for all
The conclusion now follows from Theorem 3.7 (or from Remark 3.8.1).
(cid:3)
For instance, the operators 5D, D2, −D2, eiD2, 1
10D3, −3D4 are s-hypercyclic,
but clearly not d-hypercyclic.
An extension unifying both Birkhoff's and MacLane's theorems takes place by
considering convolution operators on H(C), that is, operators commuting with
n=0 anzn ∈ H(C). Then Φ is said to be
of exponential type provided that there are positive constants A, B such that
Φ(z) ≤ A exp(Bz) for all z ∈ C. Then its associated differential operator
n=0 anf (n) (f ∈ H(C)) defines an ope-
rator on H(C). Moreover, an operator T ∈ L(H(C)) is of convolution if and only
if T = Φ(D) for some entire function Φ of exponential type. Note that D and τa
all translations τa. Let Φ(z) = P∞
Φ(D) = P∞
n=0 anDn given by Φ(D)f = P∞
24
BERNAL AND JUNG
are special cases (take Φ(z) ≡ z and Φ(z) ≡ eaz, resp.). Godefroy and Shapiro
[17] proved in 1991 that any nonscalar convolution operator is hypercyclic. If G
is any domain in C, then Φ(D) is also an operator on H(G) whenever Φ is of
subexponential type, that is, for given ε > 0 there is a constant A > 0 such that
Φ(z) ≤ A exp(εz) for all z ∈ C. We have that also Φ(D) is hypercyclic on
H(G) provided that G is simply connected (i.e. its complement with respect to
the one-point compactification C∞ of C is connected) and Φ is not constant. For
s-hypercyclicity, we present the following assertion, with which we put an end to
this introductory paper on s-universality.
Proposition 6.6. Assume that G ⊂ C is a simply connected domain and that
Φ1, . . . , Φp are entire functions of subexponential type (or just of exponential type
if G = C). Assume also that the set
is nonempty and that each set
U0 :=(cid:8)λ ∈ C : max
Ui :=(cid:8)λ ∈ C : Φi(λ) > 1 and max
1≤j≤p
Φj(λ) < 1(cid:9)
Φj(λ) ≤ Φi(λ)(cid:9) (1 ≤ i ≤ p)
1≤j≤p
has nonempty interior U 0
satisfy Φi(λ) = Φj(λ) for some λ ∈ U 0
i . Suppose, in addition, that whenever i, j ∈ {1, . . . , p}
i , there exists ζ ∈ T with Φj = ζ · Φi.
Then there is a sequence (nk) ⊂ N such that the sequences (Φ1(D))nk, . . .
. . . , (Φp(D))nk are s-mixing. In particular, the operators Φ1(D), . . . , Φp(D) are
densely s-hypercyclic on H(C).
Proof. We write eλ := exp(·λ)G for λ ∈ C. It is easy to see that the functions eλ
are linearly independent. Denote Vi := U 0
i (1 ≤ i ≤ p). As U0, V1, . . . , Vp are open
and nonempty, we obtain that X0 := span{eλ : λ ∈ U0} is dense in X := H(G)
(because G is simply connected: use Runge's approximation theorem together
with the fact that span{exp(·λ) : λ ∈ U0} is dense in H(C); see e.g. [17, Sect. 5]).
i=1 span{eλ : λ ∈ Vi} is dense in X p.
Hence W0 :=Qp
As A := {ζ ∈ T : exist l, j ∈ {1, . . . , p} with Φj = ζΦl} ⊂ T is finite, it is
a Dirichlet set; hence there is a strictly increasing sequence (nk) ⊂ N such that
ζ nk → 1 for all ζ ∈ A.
For each i ∈ {1, . . . , p}, we put Ti := Φi(D)H(G), Ei := {j ∈ {1, . . . , p} : exists
ζ ∈ T with Φj = ζΦi} and τ (i) := card(Ei). Notice that if i ∈ Ej, then Ei = Ej
(just use that T is a multiplicative group), hence τ (i) = τ (j). Given i ∈ {1, . . . , p}
and vi ∈ span{eλ : λ ∈ Vi}, there are uniquely determined scalars ci,1, . . . , ci,J(i) ∈
l=1 ci,leλi,l. For k ∈ N
C and pairwise distinct λi,1, . . . , λi,J(i) ∈ Vi such that vi =PJ(i)
we define Rk : W0 → X as
Rkw :=
pXi=1
1
τ (i)
·
J(i)Xl=1
ci,l
Φi(λi,l)nk
· eλi,l,
(1)
SIMULTANEOUS UNIVERSALITY
25
where w = (v1, . . . , vp) ∈ W0 and the vi's are as above. We have:
(i) If λ ∈ U0 and j ∈ {1, . . . , p}, then T nk
j eλ = Φj(λ)nkeλ → 0 as k → ∞,
because Φj(λ) < 1. By linearity, we get T nk
j → 0 on X0.
l=1 ci,leλi,l, as above. Since
Φi(λi,l) > 1, we get Φi(λi,l)nk → +∞ as k → ∞, for each i ∈ {1, . . . , p}
and each l = 1, . . . , J(i). From (1) one derives that Rkw → 0.
(ii) Let w = (v1, . . . , vp) ∈ W0, so that vi = PJ(i)
(iii) Again, let w = (v1, . . . , vp) ∈ W0, with vi = PJ(i)
{1, . . . , p} and k ∈ N. We compute
l=1 ci,leλi,l. Fix j ∈
T nk
j Rkw =
=
pXi=1
pXi=1
1
τ (i)
1
τ (i)
·
·
J(i)Xl=1
J(i)Xl=1
ci,l
Φi(λi,l)nk
· T nk
j eλi,l
ci,l ·(cid:18)Φj(λi,l)
Φi(λi,l)(cid:19)nk
eλi,l = Ak + Bk,
where Ak (Bk, resp.) denotes the part of the preceding sum corresponding
to those i ∈ Ej (i 6∈ Ej, resp.). If i ∈ Ej, there is ζ = ζi,j ∈ A such that
= ζ nk → 1 as k → ∞. Note that τ (i) = τ (j)
if i ∈ Ej. Therefore, on the one hand,
Φj = ζ ·Φi, so that(cid:16) Φj(λi,l)
Φi(λi,l)(cid:17)nk
J(i)Xl=1
pXi=1
Ak →
ci,l · eλi,l =
1
τ (i)
·
i∈Ej
1
τ (j)
·
pXi=1
i∈Ej
J(i)Xl=1
ci,l · eλi,l =
1
τ (j)
·
pXi=1
i∈Ej
vi.
On the other hand, if i 6∈ Ej, we have that Φj(λi,l)/Φi(λi,l) < 1 for all
l = 1, . . . , J(i) (indeed, as λi,l ∈ Vi, we have Φj(λi,l) ≤ Φi(λi,l); if we
assume Φj(λi,l) = Φi(λi,l), then there would exist ζ ∈ T with Φj = ζ · Φi,
which would yield i ∈ Ej, a contradiction). Hence (cid:0) Φj (λi,l)
Φi(λi,l)(cid:1)nk → 0, so
Bk → 0. This entails
T nk
j Rkw = Ak + Bk →
1
τ (j)
·
pXi=1
i∈Ej
vi (k → ∞),
and the last vector belongs to conv({v1, . . . , vp}) since in the last sum there
are exactly τ (j) summands.
The conclusion follows, once again, from the s-hypercyclicity criterion (Theorem
3.7).
(cid:3)
Remark 6.7. Proposition 3.4 in [11] (see also [4, Theorem 5.3]) asserts that if U0
and Wi := {λ ∈ C : Φi(λ) > 1 and maxj6=i Φj(λ) < Φi(λ)} (1 ≤ i ≤ p) are
nonempty, then Φ1(D), . . . , Φp(D) are d-mixing. If these assumptions are satisfied,
then the assumptions of Proposition 6.6 are also satisfied. Note that Proposition
6.6 includes the case Φ1 = Φ, Φj = cjΦ with cj = 1 (j = 2, . . . , p).
26
BERNAL AND JUNG
This paper does not intend to be exhaustive. Of course, many more sets of
operators or of sequences of operators may be analyzed under the point of view
of s-universality/s-hypercyclicity. For instance, consider a compact set K ⊂ C
and the Banach space (A(K), k · k∞) of continuous functions K → C that are
holomorphic on K 0. Let
TK,n : f ∈ H(D) 7→ (Snf )K ∈ A(K) (n ∈ N),
where Snf denotes the nth partial sum of the Taylor series of f around the origin.
Assume that K ⊂ C\D and that K has connected complement. Then Costakis and
Tsirivas [14, Sect. 3] have recently shown that, given any two strictly increasing
sequences (nk), (mk) in N, the sequences (TK,nk) and (TK,mk) are -- by using our
terminology -- s-universal. Even more, they have shown that
\ (cid:8)s-U((TK,nk), (TK,mk)) : K ⊂ C \ D compact, C \ K connected(cid:9)
is a residual subset of H(D).
Acknowledgements. The first author has been partially supported by Plan
Andaluz de Investigaci´on de la Junta de Andaluc´ıa FQM-127 Grant P08-FQM-
03543 and by MEC Grant MTM2015-65242-C2-1-P. The second author has been
supported by DFG-Forschungsstipendium JU 3067/1-1.
References
[1] N. Bamerni, V. Kadets, and A. Kılı¸cman, Hypercyclic operators are subspace hypercyclic,
J. Math. Anal. Appl. 435 (2016), 1812 -- 1815.
[2] F. Bayart and E. Matheron, Dynamics of Linear Operators, Cambridge University Press,
2009. Cambridge Tracts in Mathematics.
[3] L. Bernal-Gonz´alez, Densely hereditarily hypercyclic sequences and large hypercyclic mani-
folds, Proc. Amer. Math. Soc. 127 (1999), 3279 -- 3285.
[4] L. Bernal-Gonz´alez, Disjoint hypercyclic operators, Studia Math. 182 (2007), 113 -- 131.
[5] L. Bernal-Gonz´alez and K.-G. Grosse-Erdmann, The Hypercyclicity Criterion for sequences
of operators, Studia Math. 157 (2003), no. 1, 17 -- 32.
[6] J. B`es and O. Martin, Compositional disjoint hypercyclicity equals disjoint supercyclicity,
Houston J. Math. 38 (2012), 1149 -- 1163.
[7] J. B`es, O. Martin, and A. Peris, Disjoint hypercyclic linear fractional composition operators,
J. Math. Anal. Appl. 381 (2011), 843 -- 856.
[8] J. B`es, O. Martin, A. Peris, and S. Shkarin, Disjoint mixing operators, J. Funct. Anal. 263
(2012), 1283 -- 1322.
[9] J. B`es, O. Martin, and R. Sanders, Weighted shifts and disjoint hypercyclicity, J. Operator
Theory 72 (2014), 15 -- 40.
[10] J. B`es and A. Peris, Hereditarily hypercyclic operators, J. Funct. Anal. 167 (1999), no. 1,
94 -- 112.
[11]
[12] G. D. Birkhoff, D´emonstration d'un th´eor`eme ´el´ementaire sur les fonctions enti`eres, C. R.
, Disjointness in Hypercyclicity, J. Math. Anal. Appl. 336 (2007), 297 -- 315.
Acad. Sci. Paris 189 (1929), 473 -- 475.
[13] L. Bukovsk´y, The structure of the real line, Springer Basel, Warsaw, 2011.
SIMULTANEOUS UNIVERSALITY
27
[14] G. Costakis and N. Tsirivas, Doubly universal Taylor series, J. Approx. Theory 180 (2014),
21 -- 31.
[15] M. De La Rosa and C. Read, A hypercyclic operator whose direct sum T ⊕T is not hypercyclic,
J. Operator Theory 61 (2009), 369 -- 380.
[16] D. Gaier, Lectures on complex approximation, Birkhauser, Basel-London-Stuttgart, 1987.
[17] G. Godefroy and J. H. Shapiro, Operators with dense, invariant, cyclic vectors manifolds,
J. Funct. Anal. 98 (1991), no. 2, 229 -- 269.
[18] K.-G. Grosse-Erdmann, Holomorphe Monster und universelle Funktionen, Vol. 176, Mit.
Math. Sem. Giessen, 1987.
[19]
[20]
, Universal families and hypercyclic operators, Bull. Amer. Math. Soc. 36 (1999),
no. 3, 345 -- 381.
, Recent developments in hypercyclicity, Rev. R. Acad. Cienc. Ser. A Mat. 97 (2003),
no. 3, 273 -- 286.
[21] K.-G. Grosse-Erdmann and A. Peris, Linear Chaos, Springer, London, 2011.
[22] C. M. Le, On subspace-hypercyclic operators, Proc. Amer. Math. Soc. 139 (2011), 2847 -- 2852.
[23] F. Le´on-Saavedra and V. Muller, Rotations of hypercyclic and supercyclic operators, In-
tegr. equ. oper. theory 50 (2004), 385 -- 391.
[24] B. F. Madore and R. A. Mart´ınez-Avendano, Subspace hypercyclicity, J. Math. Anal. Appl.
373 (2011), 502 -- 511.
[25] G. R. MacLane, Sequences of derivatives and normal families, J. Anal. Math. 2 (1952),
no. 1, 72 -- 87.
[26] S. Rolewicz, On orbits of elements, Studia Math. 32 (1969), 17 -- 22.
[27] J. H. Shapiro, Composition Operators and Classical Function Theory, Universitext, Springer-
Verlag, New York, 1993.
[28] S. Shkarin, Universal elements for non-linear operators and their applications, J. Math.
Anal. Appl. 348 (2008), 193 -- 210.
[29] J. Wengenroth, Hypercyclic operators on nonlocally convex spaces, Proc. Amer. Math. Soc.
131 (2003), no. 6, 1759 -- 1761.
[30] L. Zhang and Z. H. Zhou, Dynamics of composition operators on weighted Bergman spaces,
Indag. Math. (N.S.) 27 (2016), no. 1, 406 -- 418.
Departamento de An´alisis Matem´atico
Facultad de Matem´aticas,
Universidad de Sevilla
Apdo. 1160, Avenida Reina Mercedes
41080 Sevilla, Spain.
E-mail address: [email protected]
Fachbereich IV Mathematik
Universitat Trier
D-54286 Trier, Germany
E-mail address: [email protected]
|
1904.03088 | 1 | 1904 | 2019-04-05T14:32:16 | Elliptic problems and holomorphic functions in Banach spaces | [
"math.FA"
] | In the first part we show that a vector-valued almost separably valued function $f$ is holomorphic (harmonic) if and only if it is dominated by an $L^1_\mathrm{loc}$ function and there exists a separating set $W\subset X'$ such that $\langle f,x'\rangle$ is holomorphic (harmonic) for all $x'\in W$. This improves a known result which requires $f$ to be locally bounded. In the second part we consider classical results in the $L^p$ theory for elliptic differential operators of second order. In the vector-valued setting these results are shown to be equivalent to the UMD property. | math.FA | math |
ELLIPTIC PROBLEMS AND HOLOMORPHIC
FUNCTIONS IN BANACH SPACES
WOLFGANG ARENDT, MANUEL BERNHARD, AND MARCEL KREUTER
Abstract. In the first part we show that a vector-valued almost
separably valued function f is holomorphic (harmonic) if and only
if it is dominated by an L1
loc function and there exists a separating
set W ⊂ X ′ such that hf, x′i is holomorphic (harmonic) for all
x′ ∈ W . This improves a known result which requires f to be
locally bounded. In the second part we consider classical results in
the Lp theory for elliptic differential operators of second order. In
the vector-valued setting these results are shown to be equivalent
to the UMD property.
1. Introduction
Let f : Ω → X, where Ω is an open subset of C (or Rd) and X is
a complex (real) Banach space. The function f is called holomorphic
(harmonic) if it is complex differentiable (twice partially differentiable
with ∆f = 0). The first part of this article is concerned with a cri-
terion for vector-valued holomorphy (harmonicity). The function f is
called weakly holomorphic (weakly harmonic) if x′ ◦ f is holomorphic
(harmonic) for all x′ ∈ X ′. We say that f is very weakly holomorphic
(very weakly harmonic) if there exists a separating subset W ⊂ X ′
such that x′ ◦ f is holomorphic (harmonic) for all x′ ∈ W .
It was shown in [11] (see also [12]) that a vector-valued function f
is holomorphic if and only if it is locally bounded and very weakly
holomorphic. This answered a question posted in [17] ten years earlier.
A very short proof was given in [4]. In [2] it was shown that a similar
approach yields the analogous result for harmonic functions. The first
part of this paper is concerned with an improvement of these results. It
is known, that very weak holomorphy alone is not sufficient [4, Theorem
1.5]. However, we will show that the boundedness assumption can be
weakened. We say that a set F of functions from Ω to X is locally L1-
bounded if there exists a function g ∈ L1
loc(Ω, R) such that kf (ξ)kX ≤
g(ξ) holds for almost all ξ ∈ Ω and for all f ∈ F . A single function
Date: April 8, 2019.
2010 Mathematics Subject Classification. Primary: 35J25; Secondary: 46B20,
31C05, 30A99.
Key words and phrases. Vector-valued holomorphic functions, harmonic func-
tions, elliptic equations, regularity, UMD spaces.
1
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
2
f : Ω → X is called locally L1-bounded if {f } is locally L1-bounded.
A net {fi}i∈I is called locally L1-bounded if {fi, i ∈ I} is locally L1-
bounded. Our result is the following.
Theorem 1.1. Let X be a Banach space and Ω ⊂ C (Ω ⊂ Rd) be
open. A function f : Ω → X is holomorphic (harmonic) if and only
if it is locally L1-bounded and very weakly holomorphic (very weakly
harmonic).
We give two proofs of this result, one of which is very short, but is
only valid if X is separable, and one that follows the approach in [4] and
[2]. The first proof will also yield a shortcut proof for the vector-valued
version of Weyl's lemma. This result will be needed in the second part
of the paper. The set L(D(Ω, R), X) of X-valued distributions on Ω
will be denoted by D′(Ω, X).
Theorem 1.2 (Weyl). Let Ω ⊂ Rd be open, X be a Banach space
and let f ∈ L1
loc(Ω, X) such that ∆f = 0 in D′(Ω, X). Then f has a
harmonic representative; that is, there exists f ∗ ∈ C ∞(Ω, X) such that
∆f ∗ = 0 and f = f ∗ almost everywhere.
Recall that ∆f = 0 in the sense of distributions means thatR f ∆ϕ =
0 for all test functions ϕ ∈ C ∞
c (Ω, R). In view of the formulation of
Theorem 1.1 we want to remark that this is equivalent to saying that
there exists a separating set W ⊂ X ′ such that ∆(x′◦f ) = 0 in D′(Ω, R)
for all x′ ∈ W .
In the second part of the paper, Sections 4 - 6, we investigate some
classical elliptic problems. Again, we ask whether the solutions with
values in a Banach space have the same regularity as in the scalar case.
Using a result by Geiss, Montgomery and Saksman [8] on homogeneous
vector-valued multipliers we prove our main technical tool, Thereom
4.3, on regularity properties of Newtonian potentials. This result will
be used to determine the domain of the Laplacian on Lp(Rd, X). One
of our main results shows that on a bounded domain Ω of class C 1,1
the following classical property characterizes UMD-spaces: Given f ∈
Lp(Ω, X) there exists a unique u ∈ W 1,p
0 (Ω, X) ∩ W 2,p(Ω, X) solving
∆u = f . More general elliptic operators are also considered in Section
6.
Parts of this work are contained in the third author's thesis [15].
2. Harmonic and Holomorphic Functions -- The Separable
Case
In this section we will give a proof of Theorem 1.1 in the case of a
separable Banach space.
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
3
Lemma 2.1. Let Ω ⊂ Rd be open and let f ∈ L1
loc(Ω, R) such that
∆f = 0 in D′(Ω, R). Furthermore let ρr be a mollifier supported in
B(0, r), r > 0, consisting of radial functions. Then ρr ∗ f = f almost
everywhere in Ωr := {ξ ∈ Ω, dist(ξ, ∂Ω) > r}.
Proof. Since ∆f = 0 distributionally there exists a harmonic represen-
tative f ∗ of f by Weyl's Lemma in the real-valued case [5, Chapter
II, §3, Proposition 1]. Since f = f ∗ almost everywhere it follows that
ρr ∗ f = ρr ∗ f ∗ everywhere in Ω. Now by [7, Chapter 2, Proof of
Theorem 6] we have that ρr ∗ f ∗ = f ∗ on Ωr from which the claim
follows.
(cid:3)
Proof of Theorem 1.1 if X is separable. (a) We start with the case that
f is very weakly harmonic. Then f is very weakly measurable, and it
follows from the Krein-Smulyan theorem (c.f. Section 3) that f is mea-
surable -- for a full proof we refer the reader to [1, Theorem 1.2] or
[14, Theorem 1.1.20]. Since f is locally L1-bounded it follows that
f ∈ L1
loc(Ω, X). Let ρr be a mollifier supported in B(0, r). Then the
function fr := ρr ∗ f is well-defined and smooth in Ωr. Lemma 2.1
shows that hfr, x′i = ρr ∗ hf, x′i = hf, x′i in Ωr for every x′ in the sepa-
rating set W ⊂ X ′ for which hf, x′i is harmonic. Since W is separating
it follows that fr = f in Ωr. In particular: f is smooth and hence --
using again that W is separating -- ∆f = 0.
(b) Now we come to the case where f is very weakly holomorphic.
Analogously to the harmonic case one sees that f is locally integrable.
Let z0 ∈ Ω and let r0 > 0 such that B(z0, r0) ⊂⊂ Ω. Since f is inte-
grable on B(z0, r0) it follows from Fubini's theorem that f is integrable
on the sphere S(z0, r) for almost all r ≤ r0. Choose such an r and
define
u(z) :=
1
2πiZw−z0=r
f (w)
z − w
dw
for all z ∈ B(z0, r). As in the scalar case one shows that u defines a
holomorphic function. Cauchy's integral formula shows that
hu(z), x′i = hf (z), x′i
for all x′ ∈ W and all z ∈ B(z0, r). Since W is separating it follows
that u = f and hence f is holomorphic.
(cid:3)
The approach used for the case where f is very weakly harmonic also
yields the
Proof of Theorem 1.2. Let ρr be a mollifier supported in B(0, r) ⊂⊂ Ω.
By assumption the function fr := ρr ∗ f is well-defined. Since f is
measurable we may assume that X is separable.
In this case there
exists a countable separating set W ⊂ X ′ [14, Proposition B.1.10].
Lemma 2.1 shows that for every x′ ∈ W there exists a negligible set
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
4
Nx′ such that hfr, x′i = hf, x′i in Ωr\Nx′. Since W is countable the set
N := Sx′∈W Nx′ is negligible. Furthermore W separates X and hence
fr = f almost everywhere in Ωr. For every x′ ∈ W the function hf, x′i
has a harmonic representative by Weyl's Lemma in the real-valued
case [5, Chapter II, §3, Proposition 1]. Since hfr, x′i is a continuous
representative of hf, x′i in Ωr it follows that hfr, x′i is the harmonic
representative of hf, x′i in Ωr. Since W is separating it follows from
Theorem 1.1 that fr is harmonic in Ωr. The claim now follows by
taking a sequence rn → 0 and defining f ∗(ξ) := frn(ξ), where ξ ∈ Ωrn.
Then f ∗ : Ω → X is well-defined, harmonic, and conincides with f
almost everywhere.
(cid:3)
We want to give a holomorphic version of Theorem 1.2 using the
distributional Cauchy-Riemann equations
D1u = D2v
D2u = −D2v.
For vector-valued functions -- which do not have a real or imaginary
part -- we make sense of these equations by saying that a function
f : Ω → X from an open subset of C into a complex Banach space
satisfies the Cauchy-Riemann equations very weakly distributionally if
there exists a separating set W ⊂ X ′ such that the functions
u := Rehf, x′i and v := Imhf, x′i
satisfy the Cauchy-Riemann equations distributionally for every x′ ∈
W . The following lemma is known, but we present a proof using our
results.
Lemma 2.2 ([10, Theorem 9]). Let Ω ⊂ C be open and let f ∈
L1
loc(Ω, C) such that f satisfies the Cauchy-Riemann equations distri-
butionally. Then f has a holomorphic representative.
Proof. It follows from the Cauchy-Riemann equations that the func-
tions u := Re f and v := Im f are harmonic in D′(Ω, R). Lemma 2.1
shows that for a radially symmetric mollifier ρr supported in B(0, r) we
have fr := ρr ∗ f = f almost everywhere in Ωr. Since fr is continuously
partially differentiable it satisfies the Cauchy-Riemann equations in the
classical sense in Ωr and thus is holomorphic. We may now define the
representative of f analogously to the proof of Theorem 1.2.
(cid:3)
Theorem 2.3. Let Ω ⊂ C be open and let X be a complex Banach
space. Suppose f ∈ L1
loc(Ω, X) satisfies the Cauchy-Riemann equations
very weakly distributionally. Then f has a holomorphic representative.
Proof. Since f is measurable we may assume that X is separable. By
[14, Theorem B.1.11] we may assume that W is countable. Let ρr be a
radially symmetric mollifier supported in B(0, r) and define fr := ρr ∗f .
Let x′ ∈ W . By Lemma 2.2 we know that hf, x′i has a holomorphic
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
5
representative and the proof tells us that in Ωr this representative is
given by ρr ∗ hf, x′i = hfr, x′i. Since W is countable, it follows that fr
is a representative of f in Ωr. Furthermore, Theorem 1.1 shows that fr
is holomorphic in Ωr. The representative is then defined as above. (cid:3)
3. Harmonic and Holomorphic Functions -- The General
Case
As announced before, we will now give a proof of Theorem 1.1 which
is valid also in non-seperable spaces. We use arguments of [4] and
[2] but add a new idea to get around with the L1
loc-hypothesis only.
We gather some results which we will need for the proof. By σd−1 we
denote the (d − 1)-dimensional Hausdorff measure in Rd and by ωd the
Hausdorff measure of the unit sphere Sd−1 ⊂ Rd.
Lemma 3.1 ([2, Lemma 2.3 and Theorem 5.2], also c.f. [13]). (a) Let
Ω ⊂ C be open and let X be a Banach space. A function f : Ω → X
is holomorphic if and only if it is weakly holomorphic. In this case,
f satisfies Cauchy's integral formula
f (z) =
1
2πiZw−z0=r0
f (w)
w − z
dw
for all z0 ∈ Ω and r0 > 0 such that z ∈ B(z0, r0) ⊂⊂ Ω.
(b) Let Ω ⊂ Rd be open and let X be a Banach space. A function
f : Ω → X is harmonic if and only if it is weakly harmonic. In
this case, f satisfies Poisson's integral formula
f (ξ) =
1
ωdr0 ZSd−1(ξ0,r0)
r2
0 − ξ − ξ02
ξ − sd
f (s) dσd−1(s)
for all ξ0 ∈ Ω and r0 > 0 such that ξ ∈ B(ξ0, r0) ⊂⊂ Ω.
Proposition 3.2. Let X be a Banach space and let {fi}i∈I be a locally
L1-bounded net of X-valued holomorphic (harmonic) functions on the
open set Ω ⊂ C (Ω ⊂ Rd). Assume that f := limi∈I fi exists pointwise
in Ω. Then f is a holomorphic (harmonic) function and f = limi∈I fi
uniformly on compact sets.
Proof. We start with the case of holomorphic functions. Let z0 ∈ Ω.
By Fubini's theorem the net {fi}i∈I is locally L1-bounded on the set
{w ∈ Ω, w − z0 = r0} for almost all r0 > 0. Fix such an r0 > 0 and
denote by Γ the set {w ∈ Ω, w − z0 = r0}. Cauchy's integral formula
yields
kfi(z1) − fi(z2)kX ≤
≤
1
2πZΓ
C
2π
kfi(w)kX(cid:12)(cid:12)(cid:12)(cid:12)
z1 − z2
(w − z1)(w − z2)(cid:12)(cid:12)(cid:12)(cid:12)
kfikL1(Γ,X)z1 − z2
dw
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
6
for all z1, z2 ∈ B(z0, r0), some constant C = C(dist(z1, Γ), dist(z2, Γ))
and all i ∈ I. Since the net {fi}i∈I is locally L1-bounded this shows
that {fi}i∈I is equicontinuous on compact subsets. Hence f = limi∈I fi
exists uniformly on compact sets and it follows that f satisfies Cauchy's
integral formula and is thus holomorphic. The case of harmonic func-
tions is treated analogously using Poisson's integral formula.
(cid:3)
Theorem 3.3 (Krein-Smulyan [16, Corollary 2.7.12]). Let X be a Ba-
nach space and let Y ⊂ X ′ be a subspace. Then Y is closed in the
weak-∗ topology if and only if Y ∩ BX ′ is weakly-∗ closed, where BX ′
denotes the closed unit ball in X ′.
Proof of Theorem 1.1 for general Banach spaces. Consider the space
Y := {x′ ∈ X ′, x′ ◦ f is holomorphic (harmonic)}.
Since W ⊂ Y it follows that Y is weak-∗ dense in X ′.
It remains
to show that Y is closed in the weak-∗ topology since then the result
follows from Lemma 3.1. By the Krein-Smulyan theorem it suffices to
show that Y ∩ BX ′ is weakly-∗ closed for every r > 0. Let {x′
i}i∈I be
a net in Y ∩ BX ′ such that x′
i ⇀∗ x′ ∈ BX ′. The net formed by the
i ◦ f is locally L1-bounded and converges pointwise to
functions fi := x′
x′◦f . By Proposition 3.2 it follows that x′◦f is holomorphic (harmonic)
and hence x′ ∈ Y .
(cid:3)
Vitali's convergence theorem is usually stated for bounded sequences
of holomorphic functions. We apply our results to show that it also
holds for locally L1-bounded sequences. Let Ω be an open and con-
nected set in C (or Rd). A subset N ⊂ Ω is called a set of uniqueness
for holomorphic (harmonic) functions if every holomorphic (harmonic)
function which vanishes on N also vanishes on Ω. It is well known that
any infinite set contained in a compact subset ofΩ is a set of uniqueness
for holomorphic functions. This does not hold for harmonic functions.
On the other hand, if the closure of N ⊂ Ω has non-empty interior,
then N is a set of uniqueness for harmonic functions.
Theorem 3.4 (Vitali). Let X be a Banach space and let fn be a locally
L1-bounded sequence of X-valued holomorphic (harmonic) functions.
Suppose that N ⊂ Ω is a set of uniqueness for holomorphic (harmonic)
functions such that limn→∞ fn exists pointwise on N. Then limn→∞ fn
exists uniformly on compact sets and defines a holomorphic (harmonic)
function.
Proof. The function
F : Ω → ℓ∞(N, X)
z 7→ (fn(z))n∈N
is holomorphic (harmonic) by Theorem 1.1. Let c(N, X) ⊂ ℓ∞(N, X)
be the closed subspace of all convergent sequences and denote by q the
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
7
quotient map ℓ∞(N, X) → ℓ∞(N, X)/c(N, X). Then q ◦ F is holomor-
phic (harmonic) and vanishes on N. Since N is a set of uniqueness we
have q ◦ F = 0, that is, F (z) is convergent for every z ∈ Ω. The claim
now follows from Proposition 3.2.
(cid:3)
4. Newtonian Potentials
With this section we start the second part of this paper on elliptic Lp
theory in Banach spaces. Our results about harmonic functions from
Section 3 will play a role in Section 6. In the remainder of the paper X
denotes a real Banach space. We recall some facts about the Newtonian
potential which can be proved analogously to the real-valued case, see
[9, Section 4.2] and [5, Chapter II, §3]. The fundamental solution for
the Laplace equation is given by
Rd\{0} ∋ ξ 7→ Φ(ξ) =( 1
2π log ξ,
d(2−d)λ(B(0,1)) ξ2−d,
1
if d = 2
if d > 2.
For f ∈ L1(Rd, X) with compact support we define the Newtonian
potential of f via
Φ ∗ f.
The Newtonian potential of f is an element of L1
loc(Rd, X) and satisfies
Poisson's equation ∆(Φ ∗ f ) = f in D′(Rd, X). Furthermore, if f is
compactly supported and Holder continuous, the Newtonian potential
of f is in C 2(Rd, X) and satisfies ∆(Φ ∗ f ) = f in the classical sense,
cf. [9, Section 4.2].
In this section we will show that certain classical Lp estimates for
the Newtonian potential on domains imply the UMD property of X.
For an overview concerning the UMD property we refer the reader to
[14, Chapter 5]. The base for our results is the following multiplier
theorem. We denote by MLp(Rd, X) the space of all scalar-valued
Lp(Rd, X) multipliers, see [14, Definition 5.3.3].
Theorem 4.1 ([8, Theorem 3.1]). Let d ≥ 2 and let X be a Ba-
nach space. Let m ∈ C ∞(Rd\{0}, R) be even, not constant and 0-
homogeneous, that is,
m(λξ) = m(ξ)
for all ξ ∈ Rd\{0} and λ > 0. Suppose that m ∈ MLp(Rd, X) for some
1 < p < ∞. Then X has the UMD property
Corollary 4.2. Let j, k ∈ {1, . . . , d}. If the second-order Riesz trans-
form RjRk (associated with the multiplier − ξiξj
ξ2 ) is bounded in Lp(Rd, X)
for some 1 < p < ∞, then X has the UMD property.
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
8
Using this corollary we may prove the main result of this section -- a
further characterization of the UMD property which will be useful in
the next section.
Theorem 4.3. Let d ≥ 2, Ω ⊂ Rd be open and non-empty and let
1 < p < ∞. Suppose that there exists a constant C > 0 and j, k ∈
{1, . . . , d} such that the estimate
kDjk(Φ ∗ f )kLp(Ω,X) ≤ Ckf kLp(Ω,X)
holds for all f ∈ C ∞
Proof. Since Φ ∈ L1
C ∞
c (Rd, X). Moreover
c (Ω, X). Then X has the UMD property.
loc(Rd, R), one has Φ ∗ f ∈ C ∞(Rd, X) for all f ∈
∆(Φ ∗ f ) = f
in the classical sense. For f : Rd → X and λ > 0 define the dilation
fλ(x) := f (λx)
of f . Consider the operator T : C ∞
c (Rd, X) → C ∞(Rd, X) given by
It is remarkable that T commutes with dilation, that is,
T f = Djk(Φ ∗ f ).
(T f )λ = T fλ
for all f ∈ C ∞
c (Rd, X). To see this we first note that
Φ(λ−1·) = λd−2Φ(·) + cd(λ)
where cd(λ) = 0 if d > 2 and c2(λ) is a constant. Consequently, for
f ∈ C ∞
c (Rd, X), λ > 0 and ξ ∈ Rd we have
(Φ ∗ fλ)(ξ) =Z Φ(ξ − η)f (λη) dη
ω
λ(cid:17) f (ω) dω
= λ−dZ Φ(cid:16)ξ −
= λ−dZ Φ(λ−1(λξ − ω)f (ω) dω
= λ−2Z Φ(λξ − ω)f (ω) dω + λ−dcd(λ)Z f (ω) dω
= λ−2(Φ ∗ f )(λξ) + λ−dcd(λ)Z f (ω) dω.
Consequently, since the second term does not depend on ξ,
(1)
Djk(Φ ∗ fλ) = (Djk(Φ ∗ f ))λ.
Next we note that for each measurable function g : Rd → X and λ > 0
we have
(2)
d
p kgλkLp(Rd,X) = kgkLp(Rd,X).
λ
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
9
Using Corollary 4.2 it remains to show that RjRk is bounded on Lp(Rd, X).
Since Ω is open we may assume that it contains (−1, 1)d. Let f ∈
C ∞
c (Rd, X). Since ∆(Φ ∗ f ) = f it follows that RjRkf = Djk(Φ ∗ f ) =
T f . Choose λ > 0 such that supp f ⊂ (−λ, λ)d. Thus fλ ∈ C ∞
c (Ω, X)
and we have RjRkf = Djk(Φ ∗ f ) by (1). Using (1) and (2) as well as
the assumption we obtain
kRjRkf kLp((−λ,λ)d,X) = λ
d
p k(RjRkf )λkLp((−1,1)d,X)
≤ λ
= λ
d
p k(Djk(Φ ∗ f ))λkLp(Ω,X)
d
p kDjk(Φ ∗ fλ)kLp(Ω,X)
d
p CkfλkLp(Ω,X)
≤ λ
= Ckf kLp(Rd,X).
Letting λ → ∞, this shows that RjRk is bounded.
(cid:3)
5. The Domain of the Laplacian on Lp(Rd, X)
Let X be a Banach space and 1 < p < ∞. The operator ∆p is defined
as the distributional Laplacian with maximal domain in Lp(Rd, X),
that is,
D(∆p) := {f ∈ Lp(Rd, X), ∆f ∈ Lp(Rd, X)}
∆pf := ∆f.
It is not difficult to see that ∆p is the generator of the Gaussian
semigroup, see Proposition 5.5 below. If X has the UMD property, the
following estimate is known.
Proposition 5.1 ([14, Proposition 5.5.4]). Let X be a Banach space
that has the UMD property and let 1 ≤ p < ∞. Suppose that u ∈
Lp(Rd, X) with ∆u ∈ Lp(Rd, X). Then for j, k ∈ {1, . . . , d} we have
Djku ∈ Lp(Rd, X) and there exists a constant C ≥ 0 such that
kDjkukLp(Rd,X) ≤ Ck∆ukLp(Rd,X).
In fact: Djku is given by the second-order Riesz transform RjRk∆u.
Using this we now show
Proposition 5.2. Let X be a Banach space which has the UMD prop-
erty and let 1 < p < ∞. Then
D(∆p) = W 2,p(Rd, X).
We will need the following lemmata for the proof.
Lemma 5.3 ([14, Lemma 5.5.5]). The space C ∞
∆p (1 ≤ p < ∞).
c (Rd, X) is a core for
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
10
Lemma 5.4 (Interpolation). Let 1 ≤ p < ∞ and let Ω ⊂ Rd be open.
For every ε > 0 there exists a constant Cε > 0 such that
k∇ukLp(Ω,X d) ≤ εkD2ukLp(Ω,X d×d) + CεkukLp(Ω,X),
for every u ∈ W 2,p
is also valid in W 2,p(Ω, X).
0 (Ω, X). If Ω has a C 1,1 boundary such an inequality
Proof. This can be proved analogously to the real-valued case [9, Theo-
rems 7.27 and 7.28]. Note that the more elegant proof [9, Exercise 7.19]
using the Rellich-Kondrachov theorem does not work in this case since
the compact embeddings obviously cannot hold in infinite dimensional
spaces.
(cid:3)
Proof of Proposition 5.2. The inclusion " ⊇ " is clear. Now let f ∈
D(∆p) and let ϕn ∈ C ∞
c (Rd, X) such that ϕn → f and ∆ϕn → ∆f . By
the estimates in Proposition 5.1 and Lemma 5.4 there exists a constant
C > 0 such that
kϕnkW 2,p(Rd,X) ≤ C(kϕnkLp(Rd,X) + k∆ϕnkLp(Rd,X)),
for all n ∈ N. This shows that ϕn is Cauchy in W 2,p(Rd, X) and hence
f ∈ W 2,p(Rd, X).
(cid:3)
We now want to show the converse of Proposition 5.2. We will need
Proposition 5.5. Let 1 ≤ p < ∞. The operator ∆p is the generator
of the strongly continuous Gaussian semigroup G on Lp(Rd, X) given
by
(G(t)f )(ξ) := (4πt)− d
2 ZRd
f (ξ − η) exp(cid:18)−
η2
4t (cid:19) dη,
where t > 0, ξ ∈ Rd and f ∈ Lp(Rd, X).
Proof. The assertion is well-known if X = R [3, Example 3.7.6]. Testing
with x′ ∈ X ′ it follows immedeately that G is a semigroup. The strong
continuity of G is also well-known [3, Lemma 1.3.3]. Let A be the
generator of G and let ∆R
p be the operator ∆p for X = R. Consider
the space
D := span{f ⊗ x, f ∈ D(∆R
p ), x ∈ X}.
Since D is dense in Lp(Rd, X) and invariant under the semigroup G
it follows that D is a core for A [6, Proposition I.1.7]. Obviously,
D ⊂ D(∆p) and ∆p coincides with A on D. Since ∆p is closed, it
follows that A ⊂ ∆p. To show the inclusion A ⊃ ∆p note that λ ∈ ρ(A)
for λ > 0. It remains to show that λ − ∆p is injective. But this follows
immediately from the real-valued case.
(cid:3)
Theorem 5.6. The Banach space X has the UMD property if and only
if
D(∆p) = W 2,p(Rd, X)
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
11
for some, equivalently all, 1 < p < ∞.
Proof. It remains to show the "if" part. Since ∆p generates a C0 semi-
group there exists µ > 0 such that µ ∈ ρ(∆p). By assumption we
have
kDjDkf kLp(Rd,X) ≤ kf kW 2,p(Rd,X) ≤ Ckµf − ∆f kLp(Rd,X),
where C = kR(µ, ∆p)kL(Lp(Rd,X),W 2,p(Rd,X)) and 1 ≤ j, k ≤ d. This holds
in particular for f ∈ C ∞
c (Rd, X). Taking the Fourier transform on both
sides yields that the function
m(ξ) :=
−4π2ξjξk
µ + 4π2ξ
(ξ ∈ Rd)
is in MLp(Rd, X). We now use a scaling argument with the same
notation as in the proof of Theorem 4.3. The transformation formula
shows that
F ±1fλ = λ−d(F ±1f )λ−1
for every f ∈ C ∞
multiplier mλ satisfies
c (Rd, X). Hence the operator Tmλ assiociated with the
Tmλf = F −1(mλF f )
= F −1((m(F f )λ−1)λ)
= F −1((mλdF fλ)λ)
= (F −1(mF fλ))λ−1
= (Tmfλ)λ−1,
from which we can estimate
kTmλf kLp(Rd,X) = k(Tmfλ)λ−1kLp(Rd,X)
= λ
d
p kTmfλkLp(Rd,X)
d
p kTmkL(Lp(Rd,X))kfλkLp(Rd,X)
≤ λ
= kTmkL(Lp(Rd,X))kf kLp(Rd,X).
By symmetry we obtain kTmλkL(Lp(Rd,X)) = kTmkL(Lp(Rd,X)). Note that
mλ → −
ξjξk
ξ2 =: m∞
pointwise and that mλ ≤ 1 for all λ > 0. By the dominated con-
vergence theorem we have Tmf → Tm∞f . Fatou's lemma shows that
m∞ ∈ MLp(Rd, X) and hence the claim follows from Theorem 4.1. (cid:3)
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
12
6. Elliptic operators on domains
In the last section we showed that the Laplacian on Lp(Rd, X) has
the maximal regularity domain W 2,p(Rd, X) if and only if X is a UMD
space. Our aim in this section is to show the analogous result for the
Dirichlet Laplacian on a bounded domain Ω of class C 1,1. In fact, we
also consider more general operators.
Let L be an elliptic operator in non-divergence form given by
L := aijDij + biDi + c,
where aij, bi, c ∈ L∞(Ω, R) and a = (aij)ij is a symmetric matrix satis-
fying
aij(·)ξiξj ≥ λξ2
almost everywhere in Ω for some fixed λ > 0 and all ξ ∈ Rd. In this
section we consider the Dirichlet problem
(Lu = f
u − ϕ ∈ W 1,p
0 (Ω, X),
where f ∈ Lp(Ω, X) and ϕ ∈ W 2,p(Ω, X) are given. We will show that
the existence of a unique solution is equivalent to the UMD property.
We first start with the sufficiency of the UMD property. For L we have
the following Lp estimate.
Theorem 6.1. Let Ω ⊂ Rd be open and bounded with a C 1,1 boundary
and let L be an elliptic operator as above. Furthermore assume that
a ∈ C(Ω, Rd×d). Let X be a Banach space which has the UMD property
and let 1 < p < ∞. Then there exists a constant C > 0 such that
kukW 2,p(Ω,X) ≤ C(kukLp(Ω,X) + kLukLp(Ω,X))
for all u ∈ W 2,p(Ω, X) ∩ W 1,p
0 (Ω, X).
Proof. Proceed as in the proof of [9, Theorem 9.13] proving the esti-
mates for the Laplacian [9, Theorem 9.9] using Proposition 5.1 and also
using the interpolation estimate in Lemma 5.4.
(cid:3)
To show existence we will need an estimate which does not depend
on kukLp(Ω,X). As in Lemma 5.4, we cannot prove this estimate analo-
gously to the real-valued case [9, Lemma 9.17] since this proof uses the
Rellich-Kondrachov theorem. We gather some information about the
real-valued case.
Theorem 6.2. (a) Let Ω ⊂ Rd be open and bounded with a C 1,1-
boundary and let L be an elliptic operator with a ∈ C(Ω, Rd×d)
and c ≤ 0. Then for every data f ∈ Lp(Ω, R) and ϕ ∈ W 2,p(Ω, R)
with 1 < p < ∞ there exists a unique u ∈ W 2,p(Ω, R) satisfying
Lu = f and u − ϕ ∈ W 1,p
0 (Ω, R).
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
13
(b) In the setting of (a) let ϕ = 0 and define
T : Lp(Ω, R) → W 2,p(Ω, R) ∩ W 1,p
0 (Ω, R)
via f 7→ u. Then −T is a positive operator, that is, T f ≤ 0
whenever f ≥ 0.
Proof. (a) is the assertion of [9, Theorem 9.15]. For the proof of (b) we
first let f ∈ C ∞
c (Ω, R)+. Then f ∈ Ld(Ω, R) and hence by uniqueness
the solution u := T f is an element of W 2,d(Ω, R) ∩ W 1,d
0 (Ω, R). Fur-
thermore it is continuous up to the boundary by Morrey's embedding
theorem. Since Ω has a C 1,1 boundary this implies that u∂Ω = 0 in the
classical sense. Suppose that u(ξ) > 0 for some ξ ∈ Ω. Then u has a
nonnegative maximum in Ω. This contradicts the maximum principle
[9, Theorem 9.6].
Now let f ≥ 0 be arbitrary. There exist nonnegative functions fn ∈
C ∞
c (Ω, R) such that fn → f in Lp(Ω, R). By the first step we know
that the solution un := T fn is non-positive. The estimate in [9, Lemma
9.17] shows that un is Cauchy and hence convergent in W 2,p(Ω, R) ∩
W 1,p
(cid:3)
0 (Ω, R). The uniqueness of the solution shows that u ≤ 0.
Proposition 6.3. In the setting of Theorem 6.1 let c ≤ 0. Then we
have the estimate
kukW 2,p(Ω,X) ≤ CkLukLp(Ω,X),
for some C > 0.
Proof. Let T be the operator in Theorem 6.2 (b) considered as a bounded
operator Lp(Ω, R) → Lp(Ω, R). The operator T can be linearly ex-
tended to finite sums of tensors of the form f ⊗x with f ∈ Lp(Ω, R) and
x ∈ X. Since −T is a positive operator there exists a unique bounded
operator T with the same norm as T mapping Lp(Ω, X) → Lp(Ω, X)
which coincides with T on finite sums of tensors [14, Theorem 2.1.3].
Note that T Lu = u and thus kukLp(Ω,X) ≤ k T kL(Lp(Ω,X))kLukLp(Ω,X).
Combined with the estimate in Theorem 6.1 this yields the result. (cid:3)
We are now in a position to prove the existence and uniqueness of
strong solutions for the Poisson problem with Dirichlet boundary data.
Theorem 6.4. Let Ω ⊂ Rd be open and bounded with a C 1,1 boundary
and let L be an elliptic operator with a ∈ C(cid:0)Ω, Rd×d(cid:1) and c ≤ 0.
Furthermore let X be a space which has the UMD property. Then for
every data f ∈ Lp(Ω, X) and ϕ ∈ W 2,p(Ω, X) with 1 < p < ∞ there
exists a unique u ∈ W 2,p(Ω, X) solving Lu = f such that u − ϕ ∈
W 1,p
0 (Ω, X).
Proof. By subtracting Lϕ from f one sees that it is enough to consider
k=1 fk ⊗ xk be a simple function with
fk ∈ Lp(Ω, R) and xk ∈ X. For the data fk the real-valued Theorem
the case ϕ = 0. Let first f = Pn
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
14
6.2 yields existence of a solution uk ∈ W 2,p(Ω, R) ∩ W 1,p
the function
0 (Ω, R). Thus
n
u :=
uk ⊗ xk
Xk=1
is a solution for f . For general f ∈ Lp(Ω, X) there exists a sequence
fn of finite sums of tensors which converges to f in Lp(Ω, X). Let un
be the solution for fn. Then by Proposition 6.3 we have
kun − umkW 2,p(Ω,X) ≤ Ckfn − fmkLp(Ω,X).
Hence un → u in W 2,p(Ω, X)∩W 1,p
follows from the Lp estimates.
0 (Ω, X) and Lu = f . The uniqueness
(cid:3)
The existence theorem has the following converse which gives again
a characterization of the UMD property by a regularity property of the
Poisson equation on domains.
Corollary 6.5. Let 1 < p < ∞, Ω ⊂ Rd be an open and bounded set
with a C 1,1 boundary and let X be a Banach space. The following are
equivalent:
(i) For every f ∈ Lp(Ω, X) there exists a unique u ∈ W 2,p(Ω, X) ∩
W 1,p
0 (Ω, X) satisfying ∆u = f .
(ii) X has the UMD property.
Proof. It remains to show the implication (i) ⇒ (ii). Let ω ⊂⊂ Ω be
nonempty with a C 1,1 boundary. For f ∈ Lp(ω, X) denote by f the
extension to Rd by 0. Recall that w := Φ ∗ f ∈ L1
loc(Rd, X) satisfies
∆w = f in D′(Ω, X).
Let u be the solution for f according to (i). Then u −w solves ∆(u −
w) = 0. Theorem 1.2 shows that u − w has a harmonic representative
which is in particular in C 2(Ω, X). Since u ∈ W 2,p(ω, X) we also have
w = u − (u − w) ∈ W 2,p(ω, X). Hence the Newtonian potential defines
a mapping from Lp(ω, X) into W 2,p(ω, X). We claim that the graph
of this mapping is closed. Let fn → f ∈ Lp(ω, X) such that Φ ∗ fn →
w ∈ W 2,p(ω, X). Then for every x′ ∈ X ′ the functions hfn, x′i, hf, x′i
and hw, x′i satisfy the analogue. R has the UMD property and thus
Theorem 5.1 shows that hw, x′i = hΦ ∗ f, x′i. Choosing x′ from a
countable separating subset of X ′ [14, Proposition B.1.10] yields the
claim. Now the closed graph theorem shows the existence of a constant
C ≥ 0 such that kΦ ∗ f kW 2,p(ω,X) ≤ Ckf kLp(ω,X) for all f ∈ Lp(ω, X).
Finally, Theorem 4.3 shows that X has the UMD property.
(cid:3)
We want to relate Corollary 6.5 to the generator of the Dirichlet
Laplacian. At first we establish an abstract result.
Lemma 6.6. Let Ω ⊂ Rd be an open set, 1 ≤ p < ∞ and let T be
a positive strongly continuous semigroup on Lp(Ω, R) with generator
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
15
A. Let X be a Banach space. Then there exists a unique strongly
continuous semigroup T on LP (Ω, X) satisfying x′◦ T (t)f = T (t)(x′◦f )
for all f ∈ Lp(Ω, X) and all x′ ∈ X ′. Denote by A its generator.
Let f, g ∈ Lp(Ω, X). Then f ∈ D( A) and Af = g if and only if
x′ ◦ f ∈ D(A) and A(x′ ◦ f ) = x′ ◦ g for all x′ ∈ X ′.
Proof. By [14, Theorem 2.1.3] there is a unique bounded operator Tt
on Lp(Ω, X) such that Tt(f ⊗ x) = Ttf ⊗ x for all f ∈ Lp(Ω, R) and
all x ∈ X. This is the same as saying that x′ ◦ Ttf = Tt(x′ ◦ f ) for
all f ∈ Lp(Ω, X) and x′ ∈ X ′. It is obvious from the first property
that T := ( Tt)t≥0 is a strongly continuous semigroup on Lp(Ω, X).
For f, g ∈ Lp(Ω, X) one has f ∈ D( A) and Af = g if and only if
Tsg ds = Ttf − f for all t > 0. Using this and the corresponding
0
assertion for A the last claim follows from the fact that the integral
commutes with functionals.
(cid:3)
R t
Now let Ω ⊂ Rd be open and bounded with a C 1,1 boundary and let
1 ≤ p < ∞. Then the operator A given by
D(A) := W 1,p
Au := ∆u
0 (Ω, R) ∩ W 2,p(Ω, R)
generates a positive strongly continuous semigroup T on Lp(Ω, R).
Consider the induced semigroup T on Lp(Ω, X), where X is a Banach
space, and denote by A its generator. Then clearly
0 (Ω, X) ∩ W 2,p(Ω, X) ⊂ D( A)
W 1,p
by the preceding lemma. The identity does not hold in general:
Corollary 6.7. Let 1 < p < ∞ and let T, A, T and A be as above.
Then
W 1,p
0 (Ω, X) ∩ W 2,p(Ω, X) = D( A)
if and only if X has the UMD property.
Proof. By [14, Theorem 2.1.3] the norms of T and T coincide. Since
kTtk ≤ Me−εt for all t > 0 the same estimate holds for T . Thus A is
invertible. Now Corollary 6.5 yields the claim.
(cid:3)
References
[1] W. Arendt. Approximation of degenerate semigroups. Taiwanese J. Math.,
5(2):279 -- 295, 2001.
[2] W. Arendt. Vector-valued holomorphic and harmonic functions. Concr. Oper.,
3:68 -- 76, 2016.
[3] W. Arendt, C. J. K. Batty, M. Hieber, and F. Neubrander. Vector-valued
Laplace transforms and Cauchy problems, volume 96 of Monographs in Math-
ematics. Birkhauser/Springer Basel AG, Basel, second edition, 2011.
[4] W. Arendt and N. Nikolski. Vector-valued holomorphic functions revisited.
Math. Z., 234(4):777 -- 805, 2000.
ELLIPTIC PROBLEMS AND HOLOMORPHIC FUNCTIONS
16
[5] R. Dautray and J.-L. Lions. Mathematical analysis and numerical methods for
science and technology. Vol. 1. Springer-Verlag, Berlin, 1990.
[6] K.-J. Engel and R. Nagel. One-parameter semigroups for linear evolution equa-
tions, volume 194 of Graduate Texts in Mathematics. Springer-Verlag, New
York, 2000. With contributions by S. Brendle, M. Campiti, T. Hahn, G. Meta-
fune, G. Nickel, D. Pallara, C. Perazzoli, A. Rhandi, S. Romanelli and R.
Schnaubelt.
[7] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in
Mathematics. American Mathematical Society, Providence, RI, second edition,
2010.
[8] S. Geiss, S. Montgomery-Smith, and E. Saksman. On singular integral and
martingale transforms. Trans. Amer. Math. Soc., 362(2):553 -- 575, 2010.
[9] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second
order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the
1998 edition.
[10] J. D. Gray and S. A. Morris. When is a function that satisfies the Cauchy-
Riemann equations analytic? Amer. Math. Monthly, 85(4):246 -- 256, 1978.
[11] K.-G. Grosse-Erdmann. The Borel-Okada Theorem Revisited. Habilitationss-
chrift, FernUniversitat Hagen, 1992.
[12] K.-G. Grosse-Erdmann. A weak criterion for vector-valued holomorphy. Math.
Proc. Cambridge Philos. Soc., 136(2):399 -- 411, 2004.
[13] A. Grothendieck. Sur certains espaces de fonctions holomorphes. I. J. Reine
Angew. Math., 192:35 -- 64, 1953.
[14] T. Hytonen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach
spaces. Vol. I. Martingales and Littlewood-Paley theory, volume 63 of Ergeb-
nisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern
Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Se-
ries. A Series of Modern Surveys in Mathematics]. Springer, Cham, 2016.
[15] M. Kreuter, Vector-valued elliptic boundary value problems on rough do-
mains, Open Access Repositorium der Universitat Ulm, Dissertation,
http://dx.doi.org/10.18725/OPARU-11852, 2019
[16] R. E. Megginson. An introduction to Banach space theory, volume 183 of Grad-
uate Texts in Mathematics. Springer-Verlag, New York, 1998.
[17] V. Wrobel. Analytic functions into Banach spaces and a new characterization
for isomorphic embeddings. Proc. Amer. Math. Soc., 85(4):539 -- 543, 1982.
Wolfgang Arendt, Institute of Applied Analysis, Ulm University,
89069 Ulm, Germany
E-mail address: [email protected]
Manuel Bernhard, Institute of Applied Analysis, Ulm University,
89069 Ulm, Germany
E-mail address: [email protected]
Marcel Kreuter, Institute of Applied Analysis, Ulm University,
89069 Ulm, Germany
E-mail address: [email protected]
|
1306.3356 | 3 | 1306 | 2013-10-09T15:05:36 | Microlocal Properties of Bisingular Operators | [
"math.FA",
"math.AP"
] | We study the microlocal properties of bisingular operators, a class of operators on the product of two compact manifolds. We define a wave front set for such operators, and analyse its properties. We compare our wave front set with the $SG$ wave front set, a global wave front set which shares with it formal similarities. | math.FA | math |
MICROLOCAL PROPERTIES OF BISINGULAR
OPERATORS
MASSIMO BORSERO AND REN´E SCHULZ
Abstract. We study the microlocal properties of bisingular operators,
a class of operators on the product of two compact manifolds. We define
a wave front set for such operators, and analyse its properties. We
compare our wave front set with the SG wave front set, a global wave
front set which shares with it formal similarities.
Introduction
Bisingular operators were originally introduced by L. Rodino in [Rod75] as
a class of operators on a product of two compact manifolds Ω1 × Ω2, defined
as linear and continuous operators A = Op(a) whose symbol satisfies, in
local product-type coordinates, the estimate
Dα1
ξ1
Dα2
ξ2
Dβ1
x1 Dβ2
x2 a(x1, x2, ξ1, ξ2) ≤ Cα1,α2,β1,β2hξ1im1−α1hξ2im2−α2.
A simple and fundamental example of a bisingular operator is the tensor
product A1 ⊗ A2 of two pseudodifferential operators, with symbols in the
Hormander class, Ai ∈ Lmi(Ωi), i = 1, 2, while more complex examples
include the vector-tensor product A1 ⊠A2 studied in [Rod75], and the double
Cauchy integral operator studied in [NR06]. To each symbols of a bisingular
operator we can associate two maps
σ1 : Ω1 × Rn1 → Lm2(Ω2)
σ2 : Ω2 × Rn2 → Lm1(Ω1),
and with these maps the bisingular calculus takes the form of a calculus
with vector valued symbols. General vector valued calculi have been deeply
studied, for example, by B. V. Fedosov, B.-W. Schulze and N. N. Tarkhanov
in [FST98].
In this paper we deal with bisingular operators whose symbols follow
Hormander-type estimates (see e.g. [Hor85]), however a global version of
bisingular calculus was defined by U. Battisti, T. Gramchev, S. Pilipovi´c
and L. Rodino in [BGPR13].
In particular, we only study operators on
compact manifolds, given explicitly in local coordinates. We also note that
'product-type' operators calculi, similar to bisingular calculus, were intro-
duced by V. S. Pilidi [Pil73], R. V. Duducava [Dud79a], [Dud79b], and, more
recently, by R. Melrose and F. Rochon [MR06]. Moreover, multisingular cal-
culi were considered by V. S. Pilidi [Pil71] and L. Rodino [Rod80].
Applications of bisingular calculus include Index Theory, see e.g. [NR06],
Analitic Number Theory, see e.g. [Bat12], and Geometric Analysis, see
2010 Mathematics Subject Classification. 35S05, 35A18, 35A27.
Key words and phrases. Bisingular operators, Microlocal analysis, Wave front set.
1
2
M. BORSERO AND R. SCHULZ
e.g. [GH13].
The aim of this paper is to study the microlocal properties of bisingular
operators. In order to do this, we define a suitable wave front set for such
operators, called the Bi-wave front set, which is the union of three compo-
nents
WFbi(u) = WF1
bi(u) ∪ WF2
bi(u) ∪ WF12
bi (u),
u ∈ D′(Ω1 × Ω2). This definition is formulated using the calculus only, and,
roughly speaking, has this interpretation: WF1
bi takes care of the singulari-
ties which, in the wave front space, lie in the axis ξ2 = 0, WF2
bi takes care
of the singularities which lie in the axis ξ1 = 0, and WF12
bi of the remaining
singularities, which include all the classical ones. Our wave front set is re-
lated to the classical Hormander wave front set WFcl (see [Hor83]) via the
following inclusion
WFclu ∩ (Ω1 × Ω2 × (Rn1 \ {0}) × (Rn2 \ {0})) ⊂ WF12
bi (u).
Moreover, we have a global regularity result
WFbi(u) = ∅ ⇔ u ∈ C∞(Ω1 × Ω2),
which implies that the bi-wave front set encompasses all the singularities.
Our main result is the following:
Proposition 0.1. Let C be a bisingular operator, u ∈ D′(Ω1 × Ω2). Then
WFbi(Cu) ⊂ WFbi(u).
This Proposition shows the our wave front set is microlocal with respect to
bisingular operators. Then, we define an appropriate notion of characteristic
set for bisingular operators, given again as a union of three components,
Charbi(C) := Char1
bi(C) ∪ Char2
bi(C) ∪ Char12
bi (C),
and with this notion we can get a microellipticity result for the 1- and 2-
components of the bi-wave front set
Proposition 0.2. Let C be a bisingular operator, u ∈ D′(Ω1 × Ω2). Then
WFi
bi(u) ⊆ Chari
bi(C) ∪ WFi
bi(Cu),
i = 1, 2.
We note strong formal similarities between the bisingular calculus and the
so called SG-calculus, introduced on Rn by H. O. Cordes [Cor95] and C.
Parenti [Par72], see also R. Melrose [Mel95], Y. Egorov and B.-W. Schulze
[ES97], and E. Schrohe [Sch87]. For this reason, our wave front set has for-
mal connections to and similar features as the global SG-wave front set, or
S -wave front set, introduced by S. Coriasco and L. Maniccia [CM03], see
also R. Melrose [Mel94] for a geometric scattering version.
The paper is organized as follows. In Section 1 we fix some notation and
briefly review the bisingular calculus.
In Section 2, following the ideas
in [Hor83] and [GS94], we study the mapping properties of bisingular oper-
ators and their microlocal properties with respect to the classical wave front
set WFcl. In Section 3 we define the bi-wave front set and state the main
results concercing microlocality and microellipticity of bisingular operators.
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
3
In Sections 4 we compare the bisingular calculus and the SG calculus, fo-
cusing on the relations and differences between the bi-wave front set and the
SG wave front set.
Acknowledgements. We are grateful to Profs. D. Bahns, U. Battisti, S.
Coriasco, L. Rodino, I. Witt for valuable advice and constructive criticism.
This work was supported by the German Research Foundation (Deutsche
Forschungsgemeinschaft, DFG) through the Institutional Strategy of the
University of Gottingen, in particular through the research training group
GRK 1493 and the Courant Research Center "Higher Order Structures in
Mathematics".
The second author is also grateful for the support received by the Studiens-
tiftung des Deutschen Volkes and the German Academic Exchange Service
(DAAD), as part of this collaboration was funded within the framework of
a "DAAD Doktorandenstipendium".
1. Preliminaries
1.1. Introduction to bisingular calculus. In this section we will recall
the main definitions and properties of bisingular symbols and bisingular
operators with homogeneous principal symbols. For the Hormander pseu-
dodifferential operators, whose tensor products provide the model example
of bisingular operators, we use the notations from [Hor85]. Ωi, i = 1, 2,
denotes an open domain of Rni.
Definition 1.1. Sm1,m2(Ω1, Ω2) is the set of C∞(Ω1 × Ω2 × Rn1 × Rn2)
functions such that, for all multiindex αi, βi and for all compact subsets
Ki ⊂⊂ Ωi, i = 1, 2, there exists a constant Cα1,α2,β1,β2,K1,K2 > 0 such that
Dα1
x2 a(x1, x2, ξ1, ξ2) ≤ Cα1,α2,β1,β2,K1,K2hξ1im1−α1hξ2im2−α2,
for all xi ∈ Ki, ξi ∈ Rni. As usual, hξi := (1 + ξ2)
2 . An element of
Sm1,m2(Ω1, Ω2) is called a symbol.
Definition 1.2. A linear operator A : C∞
a bisingular operator if it can be written in the form
0 (Rn1+n2) → C∞(Rn1+n2) is called
Dα2
ξ2
Dβ1
x1 Dβ2
ξ1
1
A(u)(x1, x2) = Op(a)[u](x1, x2)
1
(2π)n1+n2 ZRn1ZRn2
=
ei(x1·ξ1+x2·ξ2) a(x1, x2, ξ1, ξ2)u(ξ1, ξ2) dξ1dξ2,
where a ∈ Sm1,m2(Ω1, Ω2) and u denotes the Fourier transform of u.
Lm1,m2(Ω1, Ω2) denotes the set of all bisingular operators with symbol in
Sm1,m2(Ω1, Ω2). Moreover, we set
S∞,∞(Ω1, Ω2) := [m1,m2
S−∞,−∞(Ω1, Ω2) := \m1,m2
Sm1,m2(Ω1, Ω2)
Sm1,m2(Ω1, Ω2)
and by L∞,∞(Ω1, Ω2), L−∞,−∞(Ω1, Ω2) the corresponding class of operators.
The operators in L−∞,−∞(Ω1, Ω2) are called smoothing operators.
4
M. BORSERO AND R. SCHULZ
We associate to every a ∈ Sm1,m2(Ω1, Ω2) the two maps
A1 : Ω1 × Rn1 → Lm2(Ω2)
(x1, ξ1) 7→ a(x1, x2, ξ1, D2),
A2 : Ω2 × Rn2 → Lm1(Ω1)
(x2, ξ2) 7→ a(x1, x2, D1, ξ2),
and for a ∈ Sm1,m2(Ω1, Ω2), b ∈ Sp1,p2(Ω1, Ω2) we set (for fixed x1, x2 re-
spectively)
a ◦1 b(x1, x2, ξ1, D2) := (A1 ◦ B1)(x1, x2, ξ1, D2) ∈ Lm2+p2(Ω2)
a ◦2 b(x1, x2, D1, ξ2) := (A2 ◦ B2)(x1, x2, D1, ξ2) ∈ Lm1+p1(Ω1).
Definition 1.3. Let a ∈ Sm1,m2(Ω1, Ω2). Then a has a homogeneous prin-
cipal symbol if
i) there exists am1;· ∈ Sm1,m2(Ω1, Ω2) such that
am1;·(x1, x2, tξ1, ξ2) = tm1am1;·(x1, x2, ξ1, ξ2), ∀x1, x2, ξ2, ∀ξ1 > 1, t > 0,
a − ψ1(ξ1)am1;· ∈ Sm1−1,m2,
where ψ1 is an 0-excision function. Moreover, am1;·(x1, x2, ξ1, D2) ∈
Lm2
cl (Ω2), so, being a classical symbol on Ω2, it admits an asymptotic
expansion with respect to the ξ2 variable.
ii) there exists a·;m2 ∈ Sm1,m2(Ω1, Ω2) such that
a·;m2(x1, x2, ξ1, tξ2) = tm2a·;m2(x1, x2, ξ1, ξ2), ∀x1, x2, ξ1, ∀ξ2 > 1, t > 0,
a − ψ2(ξ2)a·;m2 ∈ Sm1,m2−1, ψ2 as ψ1 above.
Moreover, a·;m2(x1, x2, D1, ξ2) ∈ Lm1
cl (Ω1), so, being a classical symbol
on Ω1, it admits an asymptotic expansion with respect to the ξ1 variable.
iii) the symbols am1;· and a·;m2 have the same leading term, so there exists
am1;m2 such that
am1;· − ψ2(ξ2)am1;m2 ∈ Sm1,m2−1(Ω1, Ω2),
a·;m2 − ψ1(ξ1)am1;m2 ∈ Sm1−1,m2(Ω1, Ω2),
and
a − ψ1am1;· − ψ2a·;m2 + ψ1ψ2am1;m2 ∈ Sm1−1,m2−1(Ω1, Ω2).
The symbols which admit a full bi-homogeneous expansion in ξ1 and ξ2 are
called classical symbols, their class is denoted by Sm1,m2
(Ω1, Ω2), and the
corresponding operator class by Lm1,m2
(Ω1, Ω2).
cl
cl
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
5
The previous Definition implies that, given A ∈ Lm1,m2
define maps σ1, σ2, σ12 in this way
cl
(Ω1, Ω2), we can
σ1(A) : T ∗Ω1 \ 0 → Lm2
cl (Ω2)
(x1, ξ1) 7→ am1;·(x1, x2, ξ1, D2),
σ2(A) : T ∗Ω2 \ 0 → Lm1
cl (Ω1)
(x2, ξ2) 7→ a·;m2(x1, x2, D1, ξ2),
σ12(A) : (T ∗Ω1 \ 0) × (T ∗Ω2 \ 0) → C
(x1, x2, ξ1, ξ2) 7→ am1;m2(x1, x2, ξ1, ξ2),
such that, denoting by σ(P )(x, ξ) the principal symbol of a pseudodifferential
operator P , we have
σ(σ1(A)(x1, ξ1))(x2, ξ2) = σ(σ2(A)(x2, ξ2))(x1, ξ1)
= σ12(A)(x1, x2, ξ1, ξ2) = am1;m2(x1, x2, ξ1, ξ2).
We call the couple (σ1(A), σ2(A)) the principal symbol of A.
In the following, we only consider bisingular operators on the product of two
compact manifolds Ω1, Ω2. They are defined as above in local coordinates.
For those, there exists a notion of ellipticity, called bi-ellipticity. For more
details, see [Rod75].
Definition 1.4. Let A ∈ Lm1,m2
cl
(Ω1, Ω2). We say that A is bi-elliptic if
i) σ12(A)(v1, v2) 6= 0 for all (v1, v2) ∈ (T ∗Ω1 \ 0) × (T ∗Ω2 \ 0);
ii) σ1(A)(v1) is exactly invertible as an operator in Lm2
cl (Ω2) for all
iii) σ2(A)(v2) is exactly invertible as an operator in Lm1
cl (Ω1) for all
v1 ∈ T ∗Ω1 \ 0, with inverse in L−m2
v2 ∈ T ∗Ω2 \ 0, with inverse in L−m1
cl
cl
(Ω2);
(Ω1).
To elaborate on this definition, we give some examples.
Example 1.5. Consider the differential operator
cβ1,β2(x1, x2)Dβ1
1 Dβ2
2 ,
A = Xβ1≤m1
with C∞ coefficients. In this case
(1)
β2≤m2
(2)
σ1(A)(x1, ξ1) = Xβ1=m1
σ2(A)(x2, ξ2) = Xβ1≤m1
β2≤m2
β2=m2
cβ1,β2(x1, x2)ξβ1
1 Dβ2
2
cβ1,β2(x1, x2)Dβ1
1 ξβ2
2 .
A full bi-homogeneous expansion is given by
σi,j(A)(x1, x2, ξ1, ξ2) = Xβ1=i
cβ1,β2(x1, x2)ξβ1
1 ξβ2
2 .
condition σ12(A) =
The bi-ellipticity
σm1,m2(A)(v1, v2) 6= 0 for all (v1, v2) ∈ (T ∗Ω1 \ 0) × (T ∗Ω2 \ 0) and
given by
of A is
the
β2=j
6
M. BORSERO AND R. SCHULZ
the invertibility of the two maps (1) and (2).
We may give a global meaning to A on a product of compact manifolds
Ω1× Ω2 by taking, for example, Ωj = Tj, the nj-dimensional torus, j = 1, 2,
and xj angular coordinates on Tj.
With this in mind, it is possible to study some model cases of operators of
the form A ⊗ B. In the following Table 1.1 we mean by ΨDO the classical
pseudodifferential operators on Ω1 × Ω2, and by ΨDO-order and ΨDO-ell.
their order and ellipticity, respectively.
Operator
I ⊗ I
−∆1 ⊗ I + I ⊗ (−∆2)
−∆1 ⊗ (−∆2)
−∆1 ⊗ (−∆2 + I)
ΨDO-order ΨDO-ell. Bi-order Bi-ell.
√
×
×
×
√
√
(0, 0)
(2, 2)
(2, 2)
(2, 2)
(2, 2)
(−2,−2)
√
√
×
×
×
0
2
4
4
4
(−∆1 + I) ⊗ (−∆2 + I)
(−∆1 + I)−1 ⊗ (−∆2 + I)−1 not a ΨDO
Table 1. Some model cases of bisingular operators of tensor
product type
Theorem 1.6. Let A ∈ Lm1,m2
B ∈ L−m1,−m2
(Ω1, Ω2) such that
cl
cl
(Ω1, Ω2) be bi-elliptic. Then there exists
AB = I + K1
BA = I + K2,
where I is the identity map and K1, K2 are smoothing operators. Moreover
the principal symbol of B is (σ1(A)−1, σ2(A)−1).
From now on we will assume, for simplicity, that symbols of bisingular op-
erators have compact support in the x1, x2 variables.
Theorem 1.7. Let a ∈ Sm1,m2(Ω1, Ω2), b ∈ Sp1,p2(Ω1, Ω2). Then AB ∈
Lm1+p1,m2+p2(Ω1, Ω2), and its symbol c(x1, x2, ξ1, ξ2) has the asymptotic ex-
pansion
c ∼
∞Xj=0
cm1+p1−j,m2+p2−j
where
cm1+p1−j,m2+p2−j = c1
m1+p1−j−1,m2+p2−j + c2
+ c12
m1+p1−j,m2+p2−j
m1+p1−j,m2+p2−j−1
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
7
and
c1
c2
c12
x2 b
x2 b
m1+p1−j−1,m2+p2−j
= Xα2=j
1
α2!
∂α2
ξ2
a ◦1 Dα2
x2 b − Xα1≤j
1
α1!
∂α1
ξ1
∂α2
ξ2
a Dα1
x1 Dα2
m1+p1−j,m2+p2−j−1
= Xα1=j
1
α1!
∂α1
ξ1
a ◦2 Dα1
x1 b − Xα2≤j
1
α2!
∂α1
ξ1
∂α2
ξ2
a Dα1
x1 Dα2
m1+p1−j,m2+p2−j
1
= Xα1=α2=j
α1!α2!
∂α1
ξ1
∂α2
ξ2
a Dα1
x1 Dα2
x2 b
Corollary 1.7.1. Let a ∈ Sm1,m2(Ω1, Ω2), b ∈ Sp1,p2(Ω1, Ω2). Then the com-
mutator [A, B] := AB − BA belongs to Lm1+p1−1,m2+p2 + Lm1+p1,m2+p2−1.
Proof. By Theorem 1.7 we have as leading order terms (j = 0):
c1
m1+p1−1,m2+p2 = a ◦1 b − b ◦1 a
c2
m1+p1,m2+p2−1 = a ◦2 b − b ◦2 a
c12
m1+p1,m2+p2 = 0.
Then, expanding c1 and c2 according to the definition of ◦j, j = 1, 2, we get
cj = i{a, b}j + terms of order (mj + pj − 2),
where with {a, b}j we denote the Poisson bracket of a and b in the j-
argument. Therefore, the leading terms (up to order (m1+p1−2, m2+p2−2))
of the expansion of c can be written as
c = 0 + i({a, b}1 + {a, b}2) ∈ Sm1+p1−1,m2+p2 + Sm1+p1,m2+p2−1.
(cid:3)
Remark 1.8. This behaviour under commutators is indeed something pe-
culiar about bisingular calculus. It has the consequence that we can not use
a lot of the common "commutator tricks" in the proofs to obtain microlocal
properties.
Example 1.9. For a better understanding of this phenomenon, consider
the model case of a tensor product where A = A1 ⊗ A2 ∈ Lm1,m2, B =
B1 ⊗ B2 ∈ Lp1,p2. Then
[A, B] = [A1 ⊗ A2, B1 ⊗ B2] = A1B1 ⊗ A2B2 − B1A1 ⊗ B2A2
{z
= [A1, B1] ⊗ A2B2
}
Lm1+p1−1,m2+p2
+ B1A1 ⊗ [A2, B2]
}
Lm1+p1,m2+p2−1
{z
.
To close this section, we note that all pseudodifferential operators of order
zero or lower, in particular those corresponding to cut-offs and excision
functions, are bisingular operators.
Lemma 1.9.1. S0(Ω1 × Ω2) ⊂ S0,0(Ω1, Ω2).
8
M. BORSERO AND R. SCHULZ
Proof. Let a ∈ S0(Ω1×Ω2). Then for all pair of multiindex α = (α1, α2), β =
(β1, β2) we have
Dβ1
x1 Dβ2
x2 a(x1, x2, ξ1, ξ2) = Dα
ξ Dβ
x a(x, ξ) ≺ hξi−α
Dα1
ξ1
Dα2
ξ2
≺ hξi−α1hξi−α2 ≺ hξ1i−α1hξ2i−α2,
that is a ∈ S0,0(Ω1, Ω2).
(cid:3)
2. Classical microlocal analysis of bisingular operators
Let us first recall the notion of classical wave front set, as introduced by
Hormander [Hor83].
Definition 2.1. Let Ω ⊂ Rn open, u ∈ D′(Ω). The distribution u is microlo-
cally C∞ near (x0, ξ0) ∈ T∗Ω\ 0 if one of the following equivalent conditions
is satisfied:
(1) There exists a pseudodifferential operator A ∈ L0(Ω), non-
characteristic at (x0, ξ0), such that Au ∈ C∞(Ω),
(2) There exists a cut-off φ ∈ C∞
0 (Ω) with φ ≡ 1 in an open set contain-
ing x0 such that there exists a conic open set Γ ⊂ Rn \ 0 containing
ξ0 and constants CN , R > 0 such that ∀ N ∈ N
cφu(ξ) ≤ CNhξi−N ∀ξ ∈ Γ, ξ > R
The classical wave front set of u ∈ D′(Ω), that we denote by WFcl(u), is the
complement of the set of points where u is microlocally C∞.
It is now interesting to compare, for given operators A, the sets WFcl(Au)
and WFcl(u). An operator is microlocal, if WFcl(Au) ⊂ WFcl(u).
2.1. Mapping properties of bisingular operators. In this Section we
shall estimate the classical wave front set WFcl(Au) for a linear operator A
and a distribution u in terms of the Schwartz kernel KA of A, defined as
follows:
Definition 2.2. To each operator A : D(Ω1) → D′(Ω2) we can uniquely
associate a distribution KA, called the Schwartz kernel of A, such that for
all f ∈ D(Ω1), g ∈ D(Ω2) we have
hAf, gi = hKA, f ⊗ gi.
The Schwartz kernel of a bisingular operator with symbol a is then defined
by the oscillatory integral
(3) KA(x1, x2, y1, y2) =ZRn1+n2
ei(x1−y1,x2−y2)·(ξ1,ξ2)a(x1, x2, ξ1, ξ2)d−ξ1d−ξ2.
The Schwartz kernel Theorem states the following smoothing property:
Proposition 2.3. A linear map D′(Ω1 × Ω2) → D′(Ω1 × Ω2) is actually
a mapping to D(Ω1 × Ω2), i.e. smoothing, if and only if its distributional
kernel is in D((Ω1 × Ω2)×2).
Therefore pseudodifferential operators with symbols in S−∞ and bisingular
operators with symbol in S−∞,−∞ can be seen to be smoothing.
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
9
As the prototype of a bisingular operator is the tensor product of two pseu-
dodifferential operators, it makes sense also to define what is meant by an
operator that is smoothing in one set of variables only.
Definition 2.4. We define the space C∞(Ω1,D′(Ω2)) as all such u ∈ D′(Ω1×
Ω2) such that for each f ∈ D(Ω2), the distribution D(Ω1) ∋ u(g ⊗ ·) : g 7→
u(g ⊗ f ) is actually a smooth function.
Correspondingly, we can define C∞(Ω2,D′(Ω1)).
We can now list the mapping properties of bisingular operators on these
spaces, following the ideas in [Tr`e67].
Lemma 2.4.1. Bisingular operators map the spaces C∞(Ω1,D′(Ω2)) and
C∞(Ω2,D′(Ω1)) into themselves.
Let a ∈ S−∞,m, then the bisingular operator Op(a) maps D′(Ω1 × Ω2) to
C∞(Ω1,D′(Ω2)) and C∞(Ω2,D′(Ω1)) to C∞(Ω1 × Ω2).
let a ∈ Sm,−∞, then the bisingular operator Op(a) maps
Accordingly,
D′(Ω1 × Ω2) to C∞(Ω2,D′(Ω1)) and C∞(Ω1,D′(Ω2)) to C∞(Ω1 × Ω2).
The following Lemma (see e.g. [GS94]) indicates how the singularities of a
distribution transform under the action of a linear operator in terms of the
singularities of its kernel.
Lemma 2.4.2. Let K ⊂ D′(Ω1 × Ω2), and denote by the same letter the
corresponding operator K : C∞
WF′(K) := {(x1, x2, ξ1,−ξ2) ∈ T ∗(Ω1 × Ω2) \ 0; (x1, x2, ξ1, ξ2) ∈ WFcl(K)}
WF′
WF′
Then if u ∈ E ′(Ω2) and WFcl(u) ∩ WF′
Ω1(K) := {(x1, ξ1) ∈ T ∗Ω1 \ 0; ∃y ∈ Ω2 with (x1, y, ξ1, 0) ∈ WF′(K)}
Ω2(K) := {(x2, ξ2) ∈ T ∗Ω2 \ 0; ∃x ∈ Ω1 with (x, x2, 0, ξ2) ∈ WF′(K)}.
0 (Ω2) → D′(Ω1). Set
Ω2(K) = ∅, we have
WFcl(Ku) ⊂ WF′(K)(WFcl(u)) ∪ WF′
Ω1(K),
where we considered WF′(K) as a relation T ∗Ω2 → T ∗Ω1.
2.2. Classical microlocality properties of bisingular operators.
From the previous Lemma it is easy to derive that all pseudodifferential
operators are microlocal. In fact, if K is the kernel of a pseudodifferential
operator A on Ω = Ω1 = Ω2, then WF′
Ω2(K), WF′(K) =
{(x, x, ξ,−ξ)}, hence from Lemma 2.4.2 we get WFcl(Au) ⊂ WFcl(u). We
now study the microlocal properties of bisingular operators.
Example 2.5. Consider Ω1 = Ω2 = R. We further pick positive φ, ψ ∈
0 (R). Now define the pseudodifferential operator Tφ on C∞
C∞
Ω1(K) = ∅ = WF′
0 (R) by
Tφ(f ) := φ ∗ f.
Then the operator A := Tφ ⊗ I is a tensor product of two pseudodifferential
operators and thus a bisingular operator on R2.
Now consider the distribution u = ψ⊗ δ. It has the following wave front set:
WFcl(u) = {(x1, 0, 0, ξ2) x1 ∈ supp(ψ), ξ2 ∈ R \ 0} .
Then it is easy to see that
WFcl(Au) =(cid:8)(x1, 0, 0, ξ2) x1 ∈(cid:0) supp(ψ) + supp(φ)(cid:1), ξ2 ∈ R \ 0(cid:9) .
10
M. BORSERO AND R. SCHULZ
The example can be similarly given in local coordinates on a product of two
compact manifolds. We restricted ourselves to the Euclidean space for the
sake of comprehensibility.
The previous example shows that, in general, bisingular operators do not
have the microlocal property. As a motivation, we start with the model
case of a tensor product of two pseudodifferential operators. For that we
use the well-known estimate for the wave front set of a tensor product of
distributions (cf. e.g. [Hor83]):
Lemma 2.5.1. Let u ∈ D′(Ω1), v ∈ D′(Ω2). Then
WFcl(u ⊗ v) ⊂ WFcl(u) × WFcl(v) ∪ (supp(u) × {0}) × WFcl(v)
∪ WFcl(v) × (supp(v) × {0}).
This can be used to estimate the wave front relation for a bisingular oper-
ator given as the tensor product of two pseudodifferential operators. As a
matter of fact this extends by direct calculation using standard techniques
of integral regularization (see, e.g. [Hor83], [Shu01], [GS94]) to
Theorem 2.6. Let A ∈ Lm1,m2(Ω1, Ω2). Then we can estimate the wave
front set of the corresponding kernel as follows:
WFcl(KA) ⊂ {(x1, x2, x1, x2, ξ1, ξ2,−ξ1,−ξ2)xi ∈ Ωi, ξi ∈ Rni \ 0}
∪ {(x1, x2, x1, y2, ξ1, 0,−ξ1, 0)xi, yi ∈ Ωi, ξ1 ∈ Rn1 \ 0}
∪ {(x1, x2, y1, x2, 0, ξ2, 0,−ξ2)xi, yi ∈ Ωi, ξ2 ∈ Rn2 \ 0}.
Corollary 2.6.1. Let A ∈ Lm1,m2(Ω1, Ω2), u ∈ E ′(Ω1 × Ω2). Then
WFcl(Au) ⊂ WFcl(u) ∪ {(x1, x2, 0, ξ2);∃y1 ∈ Ω1 : (y1, x2, 0, ξ2) ∈ WFcl(u)}
∪ {(x1, x2, ξ1, 0);∃y2 ∈ Ω2 : (x1, y2, ξ1, 0) ∈ WFcl(u)}.
(cid:3)
Proof. Use Theorem 2.6 and Lemma 2.4.2.
Based on this observation we find that bisingular operators are microlo-
cal with respect to a modified version of the classical wave front set. It is
obtained by dropping the information about the precise location of the sin-
gularities with the covariable ξ1 = 0 or ξ2 = 0 in the corresponding variable.
Proposition 2.7. Let A ∈ Lm1,m2(Ω1, Ω2), u ∈ E ′(Ω1 × Ω2). Then
WFcl(Au) ⊂ ]WFcl(u)
:= {(x1, x2, ξ1, ξ2) : (x1, x2, ξ1, ξ2) ∈ WFcl(u);ξ1ξ2 6= 0}
∪ {(x1, x2, ξ1, 0);∃y2 ∈ Ω2 : (x1, y2, ξ1, 0) ∈ WFcl(u)}
∪ {(x1, x2, 0, ξ2);∃y1 ∈ Ω1 : (y1, x2, 0, ξ2) ∈ WFcl(u)}.
Proposition 2.8. Let A ∈ Lm1,m2
cl
(Ω1, Ω2) be bi-elliptic. Then
]WFcl(Au) = ]WFcl(u).
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
11
Proof. We already know that ]WFcl(Au) ⊂ ]WFcl(u).
Now, A ∈ Lm1,m2
that BA − I = K ∈ L−∞,−∞(Ω1, Ω2). Thus
(Ω1, Ω2), therefore there exists B ∈ L−m1,−m2
cl
cl
]WFcl(u) := ]WFcl((BA − K)u) ⊂ ]WFcl(BAu) ∪ ]WFcl(Ku)
(Ω1, Ω2) such
⊂ ]WFcl(Au) ∪ ∅ = ]WFcl(Au).
(cid:3)
This notion encourages to study microlocal properties of bisingular operators
with respect to this modified notion of gWF(u). However, all of the previous
results have been obtained by the study of distribution kernels. It is far more
desirable to have a notion of singularities in terms of the actual calculus.
This will be provided in the next section.
3. A wave front set in terms of bisingular operators
3.1. Microlocal properties of bisingular operators. While being the
description that naturally arises when analysing the kernels of bisingular op-
erators, the notion of wave front set used in the previous section has several
drawbacks:
it is defined in terms of the Hormander wave front set, so in
order to calculate it one first has to find that set and then "forget informa-
tion". Also, it is not defined in terms of the bisingular calculus, but indeed
with respect to the pseudodifferential one.
In the following we establish a second notion that does not have these draw-
backs.
In fact it turns out to be quite similar to the notion introduced
in [CM03] for the SG-calculus. This is not surprising, as there is a strong
similarity in the formulas arising in both calculi.
From now on, all the pseudodifferential operators will be assumed as prop-
erly supported, and all the bisingular operators to be classical.
Definition 3.1. Let u ∈ D′(Ω1 × Ω2). We define WFbi(u) ⊂ Ω1 × Ω2 ×
(Rn1+n2 \ 0) as
WFbi(u) = WF1
bi(u) ∪ WF2
bi(u) ∪ WF12
bi (u)
where
• (x1, x2, ξ1, 0) is not in WF1
characteristic at (x1, ξ1) such that
bi(u) if there exists an A ∈ L0
cl(Ω1) non-
(A ⊗ I)u ∈ C∞(Ω1,D′(Ω2)).
• (x1, x2, 0, ξ2) is not in WF2
characteristic at (x2, ξ2) such that
bi(u) if there exists an A ∈ L0
cl(Ω2) non-
(I ⊗ A)u ∈ C∞(Ω2,D′(Ω1)).
• (x1, x2, ξ1, ξ2), ξ1ξ2 6= 0, is not in WF12
L0
cl(Ωi), non-characteristic at (xi, ξi), i = 1, 2, such that
bi (u) if there exists a Ai ∈
(A1 ⊗ A2)u ∈ C∞(Ω1 × Ω2)
(A1 ⊗ I)u ∈ C∞(Ω1,D′(Ω2)),
(I ⊗ A2)u ∈ C∞(Ω2,D′(Ω1)).
(4)
(5)
(6)
(7)
(8)
12
M. BORSERO AND R. SCHULZ
Remark 3.2. Note that the conditions (7) and (8) do not follow from
(6): Take u ∈ D′(R × R), u = δ(x − 1)δ(y + 1) and ψ smooth such that
ψ ≡ 1 for x > 1/2 and ψ ≡ 0 for x ≤ 0. Then (ψ(x) ⊗ ψ(y))u = 0, as
(1,−1) /∈ supp(ψ ⊗ ψ). However, for g ∈ D(R) with g(−1) 6= 0 we have
(ψ(x) ⊗ I)u(. ⊗ g) = δ(x − 1)g(−1), which is not smooth.
Remark 3.3. For a distribution of the form u = u1 ⊗ u2, we have that
WF12
bi (u) = WFcl(u1) × WFcl(u2).
In fact we have the following inclusion result:
Lemma 3.3.1. If a point (x1, x2, ξ1, ξ2), ξ1ξ2 6= 0, is not in WF12
then it is not in WFcl(u).
bi (u),
Proof. The proof is a variant of [Hor91], Proposition 2.8. By definition there
exists A := A1 ⊗ A2 ∈ L0,0(Ω1 × Ω2), with σ12(A)(x1, x2, ξ1, ξ2) 6= 0, such
that Au ∈ C∞(Ω1 × Ω2). Now take a (ΨDO symbol) ψ(x1, x2, ξ1, ξ2) ∈
L0
cl(Ω1 × Ω2) such that
• on the support of ψ we have hξ1i . hξ2i . hξ1i
• ψ is non-characteristic at (x1, x2, ξ1, ξ2).
Then the (bi-singular) operator product
B := ψ(x1, x2, D1, D2) ◦ A(x1, x2, D1, D2)
yields a pseudodifferential operator of order 0, plus a smoothing remainder,
by virtue of the above inequality on the support of ψ and the symbol expand
in Theorem 1.7. It has the following properties:
the sense of ΨDOs) at (x1, x2, ξ1, ξ2) and of zero order.
• its principal symbol is ψ · σ12(A), and thus is non-characteristic (in
• Bu = ψ(x1, x2, D1, D2)A(x1, x2, D1, D2)u ∈ C∞
This proves the claim.
(cid:3)
Remark 3.4. This lemma asserts that in the conic region where both co-
variables are non-vanishing we can pass from bisingular to pseudodifferential
calculus by multiplying by a ΨDO. This has the consequence that the two
operator classes have similar microlocal properties (with respect to the clas-
sical wave front set) in that region.
1, y, ξ0
The following Lemma gives a similar interpretation to the remaining compo-
nents, illustrating the loss of localization of singularities already encountered
in the previous section.
Lemma 3.4.1. Let u ∈ E ′(Ω1 × Ω2), (x0
all y ∈ Ω2 we have (x0
(x0
bi(u).
Similarly, if for all x ∈ Ω1 we have (x, x0
we have (x, x0
Proof. We prove the claim for WF1
there exist a cut-off φ ∈ C∞
non-vanishing in a (conic) neighbourhood (x0
such that
If for
1, 0) /∈ WFcl(u) then for all y ∈ Ω2 we have
2) /∈ WFcl(u) then for all x ∈ Ω1
1, 0) /∈ WFcl(u). Then
0 (Ω1×Ω2) and a conic localizer ψy ∈ C∞(Rn1+n2),
1, 0) respectively,
1, 0) /∈ WF1
2, 0, ξ0
bi. Take (x0
1, y, ξ0
1, ξ0
1) ∈ Ω1 × (Rn1 \ 0).
2 ) /∈ WF2
bi(u).
1, y) and (ξ0
1, y, ξ0
2, 0, ξ0
ψy(ξ1, ξ2)F(x1,x2)7→(ξ1,ξ2){φ(x1, x2)u} ∈ S (Rn1+n2).
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
13
As this holds true for any y, and due to compactness of the support of
u, there exists a cut-off φ1 ∈ C∞
0 (Ω1) such that for some conic localizer
ψ ∈ C∞(Rn1+n2)
ψ(ξ1, ξ2)F(x1,x2)7→(ξ1,ξ2){φ(x1)u} ∈ S (Rn1+n2).
This means we can find a conic localizer ψ1 ∈ C∞(Rn1), non-vanishing
1, such that ψ(ξ1)F(x1,x2)7→(ξ1,ξ2){φ(x1)u} ∈ C∞(Rn1+n2), rapidly
around ξ0
decaying with respect to the first variable ξ1 for fixed ξ2, and polynomially
bounded everywhere, by the Paley-Wiener-Schwartz Theorem.
We define A ∈ L0
cl(Ω1) as the operator
Av(y1) = F −1
ξ17→y1
(ψ(ξ1)Fx17→ξ1{φ(x1)u}.
By the assumptions on φ1 and ψ1, A is non-characteristic in the sense of
pseudodifferential operators at (x0
1). But for any f ∈ D(Ω2) we have
ξ17→x1hψ(ξ1)F(y1,y2)7→(ξ1,ξ2){φ(y1)u}, bfi is a smooth
that [(A⊗I)u](f )(x1) = F −1
function, which means (A ⊗ I)u ∈ C∞(Ω1,D′(Ω2)).
Proposition 3.5. Let u ∈ E ′(Ω1 × Ω2). Then
1, ξ0
(cid:3)
WFbi(u) = ∅ ⇔ u ∈ C∞(Ω1 × Ω2).
bi(u) = ∅ we can find for each (x1, x2, ξ1, ξ2) an A ∈ L0
Proof. Assume WFbi(u) = ∅. Then, by virtue of Lemma 3.3.1, we have
WFcl(u) ∩ {(x1, x2, ξ1, ξ2) : ξ1ξ2 6= 0} = ∅. Thus bu is rapidly decaying on
any ray R · (ξ1, ξ2) where ξ1ξ2 6= 0.
As also WF1
cl(Ω1) non-
characteristic at (x1, ξ1) such that, by Lemma 2.4.1, for any B ∈ Lm,−∞(Ω1×
Ω2) we have B(I ⊗ A)u ∈ C∞(Ω1 × Ω2). By compactness and a parametrix
construction we can thus conclude that for all B ∈ Lm,−∞(Ω1 × Ω2) we get
Bu ∈ C∞(Ω1 × Ω2). Now pick φ ∈ C∞
0 (Ω1 × Ω2) with φ ≡ 1 on the support
of u, and define b(x1, x2, ξ1, ξ2) = φ(x1, x2)f (ξ2), with f ∈ S (Rd2). Then
S (Rd1+d2) ∋ F(Bu) = F(cid:0)b(x1, x2, D1, D2)u(cid:1) = (1 ⊗ f )bu.
As f was arbitrary and rapidly decaying, this means that bu must already
be rapidly decaying in the first variable. Repeating the argument for the
second variable proves the assertion.
(cid:3)
Using a parametrix construction, we get by the same arguments:
Proposition 3.6. Let u ∈ E ′(Ω1 × Ω2). Then
WF1
WF2
bi(u) = ∅ ⇔ u ∈ C∞(Ω1,D′(Ω2)),
bi(u) = ∅ ⇔ u ∈ C∞(Ω2,D′(Ω1)).
Remark 3.7. Note that u ∈ C∞(Ω1,D′(Ω2))TC∞(Ω2,D′(Ω1)) does not
imply that u ∈ C∞(Ω1 × Ω2). Following [Tr`e67], a counterexample is, for
instance, δ(x1−x2). The additional regularity needed such that u ∈ C∞(Ω1×
Ω2) is therefore, by Proposition 3.5, encoded in WF12
bi (u).
The bisingular wave front set admits the following properties:
Proposition 3.8 (Properties of WFbi). Let u, v ∈ D′(Ω1 × Ω2), f ∈
C∞(Ω1 × Ω2).
14
M. BORSERO AND R. SCHULZ
jointly.
• WFbi a closed set and is conic with respect to both covariables
• Let Ω1 = Ω2 = Ω. Define for f ∈ C∞(Ω× Ω) Au(x1, x2) = u(x2, x1).
Then we can define the pull-back A∗ as an endomorphism on
D′(Ω × Ω) by duality and we have (x1, x2, ξ1, ξ2) ∈ WFbi(Au) ⇔
(x2, x1, ξ2, ξ1) ∈ WFbi(u).
• WFbi(u + v) ⊂ WFbi(u) ∪ WFbi(v); WFbi(f u) ⊂ WFbi(u).
bi(u).
bi(Cu) ⊂ WF1
Remark 3.9. These properties are quite similar to the ones the SG-wave
front set of [CM03] admits. This is not very surprising, as the bisingular
calculus is formally very similar in its definition to the SG-calculus, through
which the SG-wave front set is introduced. We will explore this connection
in Section 4.
Lemma 3.9.1. Let C ∈ Lm1,m2(Ω1 × Ω2), u ∈ D′(Ω1 × Ω2). Then we have
WF1
Proof. Let (x1, x2, ξ1, 0) /∈ WF1
bi(u). By definition there exists a non-char
(at (x1, ξ1)) A ∈ L0
cl(Ω1) such that (A⊗ I)u ∈ C∞(Ω1,D′(Ω2). In particular,
by Lemma 2.4.1 we have ∀B ∈ Lm,−∞(Ω1 × Ω2) that B(A ⊗ I)u is smooth.
By the standard pseudodifferential calculus we can thus find an A′ ∈ L0(Ω1)
such that A + A′ is elliptic in the sense of pseudodifferential operators and
such that the symbol of A′ vanishes on a conic neighbourhood Γ of (x1, ξ1).
We thus have a parametrix P ∈ L0(Ω1) such that P (A + A′) = I − R with
R ∈ L−∞(Ω1).
Take H ∈ L0(Ω1) such that H is non-characteristic at (x1, ξ1) and such that
the symbol of H vanishes outside a proper subcone of Γ. Then we have:
(H ⊗ I)Cu = (H ⊗ I)C(cid:0)(P (A + A′) + R) ⊗ I(cid:1)u
= (H ⊗ I)C(P ⊗ I)(A ⊗ I)u + (H ⊗ I)C(P A′ ⊗ I)u+
+ (H ⊗ I)C(R ⊗ I)u ∈ C∞(Ω1,D′(Ω2)).
The first summand is in C∞(Ω1,D′(Ω2)) due to the definition of A, the
second as the symbol expansion given in Theorem 1.7, using the support
properties of the symbols of H and A′, gives an operator in L−∞,0. The
third one is in C∞(Ω1,D′(Ω2)) as R ∈ L−∞,0 is already a smoothing operator
in the first variable. This proves the claim.
(cid:3)
Lemma 3.9.2. Let C ∈ Lm1,m2(Ω1 × Ω2), u ∈ D′(Ω1 × Ω2). Then we have
WF12
Proof. Let (x1, x2, ξ1, ξ2) /∈ WF12
exist Ai ∈ L0
cl(Ωi), non-char at (xi, ξi), i = 1, 2, such that
bi (u). Then, by definition, we know there
bi (Cu) ⊂ WF12
bi (u).
(A1 ⊗ A2)u ∈ C∞(Ω1 × Ω2),
(A1 ⊗ I)u ∈ C∞(Ω1,D′(Ω2)),
(I ⊗ A2)u ∈ C∞(Ω2,D′(Ω1)).
By the standard pseudodifferential calculus we can thus find A′
such that Ai + A′
such that the symbol of A′
i ∈ L0(Ωi)
i is elliptic in the sense of pseudodifferential operators and
i vanishes on a conic neighborhood Γi of (xi, ξi).
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
15
1) ⊗ (A2 + A′
We then have two parametrices Pi ∈ L0(Ωi) such that
(P1 ⊗ P2)(cid:0)(A1 + A′
2)(cid:1) = I ⊗ I − R1 ⊗ I − I ⊗ R2 − R1 ⊗ R2
{z
}
with Ri ∈ L−∞(Ωi). Now pick Hi ∈ L0(Ωi) such that Hi is non-characteristic
at (xi, ξi) and such that the symbol of Hi vanishes outside a proper subcone
of Γi. Then, recalling that by the standard pseudodifferential calculus if two
operators have disjoint support their product is a smoothing operator, and
using Lemma 2.4.1,
:=R
,
(H1 ⊗ H2)Cu =
2)u
∈C∞by eq (6)
{z
{z
∈C∞by (8) and by the support of H1,A′
1
(H1 ⊗ H2)C(cid:18)(P1 ⊗ P2)(cid:0)(A1 + A′
1) ⊗ (A2 + A′
= (H1 ⊗ H2)C(P1 ⊗ P2)(A1 ⊗ A2)u
}
+ (H1 ⊗ H2)C(P1 ⊗ P2)(A1 ⊗ A′
}
1,A′
2
+ (H1 ⊗ H2)C(R1 ⊗ I)u + (H1 ⊗ H2)C(I ⊗ R2)u + (H1 ⊗ H2)CRu
}
2)(cid:1) + R1 ⊗ I + I ⊗ R2 + R(cid:19)u
1 ⊗ A2)u
}
}
+ (H1 ⊗ H2)C(P1 ⊗ P2)(A′
+ (H1 ⊗ H2)C(P1 ⊗ P2)(A′
= (H1 ⊗ H2)C(R1 ⊗ I)u + (H1 ⊗ H2)C(I ⊗ R2)u modC∞.
Now, without loss of generality, we proceed with the calculations only for
the term (H1 ⊗ H2)C(R1 ⊗ I)u. We have
∈C∞by (7) and by the support of H2,A′
2
{z
{z
∈C∞by the support of H1,H2,A′
∈C∞because R∈L−∞,−∞
1 ⊗ A′
2)u
{z
2) + R2))u
(H1 ⊗ H2)C(R1 ⊗ I)u = (H1 ⊗ H2)C(R1 ⊗ (P2(A2 + A′
= (H1 ⊗ H2)C(R1 ⊗ P2)(I ⊗ A2)u
}
+ (H1 ⊗ H2)C(R1 ⊗ P2)(I ⊗ A′
}
∈ C∞,
+ (H1 ⊗ H2)C(R1 ⊗ I)(I ⊗ R2)u
}
{z
{z
{z
∈C∞by the support of H2,A′
2
∈C∞because R∈L−∞,−∞
∈C∞by (8)
2)u
therefore (H1 ⊗ H2)Cu ∈ C∞. With similar computations, one can check
that
(H1 ⊗ I)Cu ∈ C∞(Ω1,D′(Ω2))
(I ⊗ H2)Cu ∈ C∞(Ω2,D′(Ω1))
and this proves the claim.
(cid:3)
The preceding Lemmas lead to the following proposition, which asserts that
this definition of wave front set is indeed suitable for the calculus of bisin-
gular operators:
Proposition 3.10 (Microlocality of bisingular operators). Let C ∈
Lm1,m2(Ω1 × Ω2), u ∈ D′(Ω1 × Ω2). Then we have WFbi(Cu) ⊂ WFbi(u).
16
M. BORSERO AND R. SCHULZ
3.2. Microelliptic properties of bisingular operators. From the pre-
vious proposition, we can conclude that bielliptic operators preserve the
bi-wave front set:
Corollary 3.10.1. Let A ∈ Lm1,m2(Ω1, Ω2) be bi-elliptic.
WFbi(Au) = WFbi(u).
Then
Proof. One inclusion follows directly form Proposition 3.10. The other fol-
lows proceeding like in Proposition 2.8.
(cid:3)
Next we study the microellipticity properties of bisingular operators. For
that we need a suitable definition of a characteristic set. As in Definition
1.4, Ω1 and Ω2 are considered as compact manifolds.
Definition 3.11. Let B ∈ Lm1,m2(Ω1, Ω2) and v0 = (x1, ξ1) ∈ Ω1×(Rn1\0).
We say that B is not 1-characteristic at V := π−1
2 v0 := {(x1, y, ξ1, 0) : y ∈
Ω2} if
(1) for all v ∈ V there exists an open conic neighbourhood Θ of v such
(2) σ1(B) ∈ Lm2
(Ω2) in an open
that σ12 6= 0 on Θ \ (R+v),
conic neighbourhood Γ of v0.
cl (Ω2) is invertible with inverse in L−m2
cl
bi(B) be the set of all V such that B is 1-characteristic at V .
Let Char1
We define Char2
exchanging the roles of σ1(B) and σ2(B).
Finally, we define Char12
0, where σ12(z) = 0.
Set
bi(B) accordingly for W := π−1
2 w0, w0 ∈ Ω2 × (Rn2 \ 0), by
bi (B) as the set of points z = (x1, x2, ξ1, ξ2), ξ1ξ2 6=
Charbi(B) := Char1
bi(B) ∪ Char2
bi(B) ∪ Char12
bi (B).
Remark 3.12. B is bi-elliptic iff Charbi(B) = ∅.
Remark 3.13. With this notion Definition 3.1 can also be expressed in the
form
cl(Ω1)
Charcl(A)×(Ω2×(Rn2 \0))
(A⊗I)u∈C∞(Ω1,D′(Ω2))
cl(Ω2)
(Ω1×(Rn1 \0))×Charcl(A)
(I⊗A)u∈C∞(Ω2,D′(Ω1))
WF1
bi(u) =
WF2
bi(u) =
WF12
bi (u) =
\A∈L0
\A∈L0
\Ai∈L0
cl(Ω1)
Char1
Char2
bi(A ⊗ I),
{z
}
bi(I ⊗ A),
}
{z
bi (A1 ⊗ A2).
Char12
(A1⊗A2)u∈C∞
(A1⊗I)u∈C∞(Ω1,D′(Ω2))
(I⊗A2)u∈C∞(Ω2,D′(Ω1))
With the definition of Charbi we can review in the following Table 3.2 the
model cases of Table 1.1, setting Rn
0 =
Rn1
0 × Rn2
0 .
Lemma 3.13.1. Let C ∈ Lm1,m2(Ω1, Ω2) be such that
bi(C) ∩ (Γ × Ω2 × {0}) = ∅.
0 = Rn \ 0, Ω = Ω1 × Ω2 and Rn12
Char1
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
17
Operator
I ⊗ I
−∆1 ⊗ I + I ⊗ (−∆2)
−∆1 ⊗ (−∆2)
−∆1 ⊗ (−∆2 + I)
(−∆1 + I) ⊗ (−∆2 + I)
(−∆1 + I)−1 ⊗ (−∆2 + I)−1
Char1
bi
Char2
bi
Char12
bi
0 Ω × Rn12
0
∅
∅
0 × {0} Ω × {0} × Rn2
Ω × Rn1
Ω × Rn1
0 × {0} Ω × {0} × Rn2
Ω × Rn1
0 × {0}
∅
∅
∅
∅
∅
0
∅
∅
∅
∅
∅
Table 2. Charbi for model cases of bisingular operators
Let a ∈ S0(Ω1) have support in a closed cone Γ and be non-char (in the
sense of ΨDO) in Γ0. Then there exists a H ∈ L−m1,−m2(Ω1, Ω2) such that
HC = A ⊗ I − R,
where R ∈ L−∞,0.
Proof. The requirements on the support of a mean precisely that C is el-
liptic with respect to a in the sense of [Cor95], Theorem 2.3.3. Therefore,
by the classical calculus of pseudodifferential operators, we can find a sym-
bol e ∈ S−m1,−m2 such that for all fixed (x2, ξ2) the operator E(x2, ξ2) =
e(x1, x2, D1, ξ2) is a local parametrix with respect to a namely,
E(x2, ξ2) σ2(C)(x2, ξ2) = R(x2, ξ2) + (A ⊗ 1)(x2, ξ2),
where R(x2, D2) ∈ L−∞,0(Ω1, Ω2) and E(x2, D2) ∈ L−m1,−m2(Ω1, Ω2).
Now define H as the operator with principal symbol
h = ψ1(x1, ξ1)am1;·em1;· + ψ2(x2, ξ2)a·;m2c−1
·;m2
− ψ1(x1, ξ1)ψ2(x2, ξ2)am1;m2c−1
m1;m2.
Using the calculus and Theorem 1.7, it is straightforward that H matches
the requirements.
(cid:3)
Lemma 3.13.2. Let C ∈ Lm1,m2(Ω1 × Ω2), u ∈ D′(Ω1 × Ω2). Then we have
WF1
bi(u) ⊆ Char1
bi(C) ∪ WF1
bi(Cu).
Proof. Let (x1, x2, ξ1, 0) /∈ Char1
bi(Cu). Then there exists A ∈
L0(Ω1) such that we have (A ⊗ I)HCu ∈ C∞(Ω1,D′(Ω2)), with H as in
Lemma 3.13.1, due to microlocality (Proposition 3.10) of H. Then we have
bi(C) ∪ WF1
(A2 ⊗ I)u = (A ⊗ I)(R − HC)u
u −
= (A ⊗ I)R
}
{z
∈ L−∞,0
(A ⊗ I)HCu
}
{z
∈ C∞(Ω1,D′(Ω2)) by assumption on A
∈ C∞.
Using Lemma 2.4.1 on the first summand, this proves the claim.
(cid:3)
Using the previous Lemma, we conclude that
Proposition 3.14 (Microellipticity of bisingular operators with re-
spect to the 1 and 2 components of WFbi). Let C ∈ Lm1,m2(Ω1, Ω2),
u ∈ D′(Ω1 × Ω2). Then
WFi
bi(u) ⊆ Chari
bi(C) ∪ WFi
bi(Cu),
18
i = 1, 2.
M. BORSERO AND R. SCHULZ
bi -component. This can be seen by the following example:
We do, however, not obtain full microellipticity, i.e. with respect to the
WF12
Example 3.15. Consider C = −∆ ⊗ (−∆), u = δ ⊗ 1 + 1 ⊗ δ. Then
Char12
bi (Cu) = ∅. But take
any A1 non-characteristic at (0, ξ1), ξ1 6= 0. Then (A1 ⊗ I)u = (A1δ) ⊗ 1
which is never in C∞(Ω1,D′(Ω2)), and this means that WF12
bi (u) is non-
empty, because the (7)- and, by a similar argument, (8)-requirements fail to
hold.
bi (C) = ∅ and Cu = 0 ∈ C∞(Ω1 × Ω2), i.e. WF12
Remark 3.16. The counterexample to full microellipticity could be cir-
cumvented by imposing stronger invertibility conditions in the definition of
Char12
bi . This would, however, break the characterization of the wave front
set of Remark 3.13, and lead to a loss of local information, while yielding
no interesting cases not already covered by Corollary 3.10.1.
With our definition we obtain, however, the following Lemma, which can
be regarded as a microellipticity result for the (6)-part of Definition 3.1, for
operators given by a tensor product.
Lemma 3.16.1. Let Ci ∈ Lmi(Ωi), u ∈ D′(Ω1 × Ω2). If
(x1, x2, ξ1, ξ2) /∈ Char12
bi (C1 ⊗ C2) ∪ WF12
bi ((C1 ⊗ C2)u)
there exist operators Hi ∈ L0(Ωi), non-characteristic at (xi, ξi), such that
(H1 ⊗ H2)u ∈ C∞(Ω1 × Ω2).
Proof. By the standard pseudodifferential calculus we can pick Bi ∈
L−mi(Ωi) non-characteristic at (xi, ξi). Then the product BiCi ∈ L0 is
non-characteristic at (xi, ξi). Proposition 3.10 and the definition of the bi-
wave front set guarantee us the existence of Ai ∈ L0, non-characteristic at
(xi, ξi), such that (A1 ⊗ A2)(B1C1 ⊗ B2C2)u ∈ C∞(Ω1 × Ω2). Therefore
Hi := AiBiCi fulfils the claim.
(cid:3)
4. Comparison with SG calculus
In this section we will compare bisingular calculus with SG calculus. SG
calculus is a global calculus obtained from the classical calculus by treating
the variables and covariables equivalently by imposing on the symbols similar
estimates as in bisingular calculus. These a priori formal similarities lead
to interesting similarities in the calculus and in the analysis of singularities.
However, the two calculi also differ in important aspects, as we will point
out throughout the section.
Definition 4.1. A function a(x, ξ) ∈ C∞(R2n) is called a SG symbol belong-
ing to SGµ,m(Rn) := SGµ,m iff for every α, β ∈ Zn
+ there exists a constant
Cα,β > 0 such that
x Dβ
Dα
ξ a(x, ξ) ≤ Cα,βhξiµ−βhxim−α
for every x, ξ ∈ Rn. A SG pseudodifferential operator is an operator of the
form
Au(x) :=Z eix·ξ a(x, ξ)u(ξ) d−ξ, u ∈ S ,
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
19
and the class of operators with symbols in SGµ,m is denoted by LGµ,m.
A symbol a ∈ SGm1,m2 is called SG classical if it admits a homogeneous ex-
pansion with respect to ξ, for ξ >> 1, a homogeneous expansion in x, for
x >> 1, and the two expansions satisfy certain compatibility conditions,
we refer here to [Cor95, ES97] for a precise definition of classical symbols.
We limit our attention (in this context) to classical operators, i.e. such that
their symbols are SG classical. As usual one proceeds to develop a calculus
for these operators. As every classical SG operator A is also classical pseu-
dodifferential operator, it admits a principal symbol σψ(A), homogeneous in
the first variable. In addition, by exchanging the roles of the variables and
covariables one obtains a symbol σe(A), homogeneous in the second vari-
able. The two satisfy a compatibility condition, i.e. that there is a third,
bihomogeneous symbol σψe(A), the leading term of the corresponding ex-
pansions of the ψ and e-symbols. The principal homogeneous symbol of the
operator is then the couple (σψ(A), σe(A)) which gives rise to the principal
symbol
Symp(x, ξ) = φψ(ξ)σψ(A) + φe(x)σe(A) − φψ(x)φe(ξ)σψe(A),
where φ∗ are 0-excision functions.
So far, this is very similar to the bisingular calculus, but the expansion
formula for the symbol of a product is in fact a lot simpler. The operator
compositions arising there are not present, and the composition formula is
just the one corresponding to the c12-term in Theorem 1.7.
This leads to a definition of ellipticity close to our notion of 12-ellipticity,
as no such thing as full invertibility of the symbols as operators is needed in
the parametrix construction:
Definition 4.2. A symbol a ∈ SGµ,m is SG-elliptic iff there exist constants
R, C1, C2 > 0 such that
C1hξiµhxim ≤ a(x, ξ) ≤ C2hξiµhxim
when x + ξ ≥ R.
With this notion of ellipticity, we have the Fredholm property, i.e. an SG-
elliptic operator admits a parametrix in the calculus.
Another important aspect to note is that in SG calculus we have
Proposition 4.3. Let p ∈ SGm,µ(Rn), q ∈ SGr,ν(Rn). Then the commuta-
tor [P, Q] := P Q − QP belongs to LGm+r−1,µ+ν−1(Rn).
while in the bisingular setting the analogon in fact does not hold true, see
Corollary 1.7.1 and Example 1.9 which illustrates that the difference stems
from an interaction of the Ω1 and Ω2-parts of the operators which is not
present in the SG case, where variables and covariables are independent.
The notion of SG calculus can be used to introduce a global analysis of
singularities. It turns out that this exhibits interesting similarities with the
above analysis of bisingular operators. In the following, we refer to [Cor95],
[CM03]; see also [CJT13a], [CJT13b]. First, we introduce SG characteristic
sets and the SG wave front set.
20
M. BORSERO AND R. SCHULZ
Definition 4.4 (SG characteristic sets). Let A ∈ LGm1,m2. Define the
SG characteristic set of A as
CharSG(A) = Charψ
SG(A) ∪ Chare
SG(A) ∪ Charψe
SG(A),
where
Charψ
Chare
Charψe
SG(A) = {(x, ξ) ∈ Rn × (Rn \ 0) : σψ(A)(x, ξ) = 0},
SG(A) = {(x, ξ) ∈ (Rn \ 0) × Rn : σe(A)(x, ξ) = 0},
SG(A) = {(x, ξ) ∈ (Rn \ 0) × (Rn \ 0) : σψe(A)(x, ξ) = 0}.
Definition 4.5 (SG wave front set). Let u ∈ S ′(Rn). Define the SG
wave front set of u as
WFSG(u) = WFψ
SG(u) ∪ WFψe
SG(u),
where
SG(u) ∪ WFe
SG(u) = \A∈LG0,0(Rn)
SG(u) = \A∈LG0,0(Rn)
SG(u) = \A∈LG0,0(Rn)
Au∈S (Rn)
Au∈S (Rn)
Charψ
SG(A),
Chare
SG(A),
Charψe
SG(A).
WFψ
WFe
WFψe
This notion exhibits the following properties:
Au∈S (Rn)
Proposition 4.6 (Properties of the SG wave front set). Let u, v ∈
S ′(Rn), f ∈ S (Rn). Then:
(1) WFSG(u) is a closed set and WFψ
variable x, WFe
with respect to both of them independently,
SG(u) is conical with respect to the
SG(u) with respect to the covariable ξ and WFSG(u)
(2) (x, ξ) ∈ WFSG(u) ⇔ (ξ,−x) ∈ WFSG(Fu) (Fourier Symmetry),
(3) WFSG(u + v) ⊆ WFSG(u) ∪ WFSG(v); WFSG(f u) ⊆ WFSG(u),
(4) WFSG(u) = ∅ ⇔ u ∈ S (Rn) (Global regularity).
Already we see the similarities, but also apparent differences, with the bisin-
gular notion:
• Fourier transformation (i.e. exchange of variables and covariables)
corresponds to the exchange of variables x1 and x2 in Proposition
3.8.
• The conical properties of the individual components of the wave front
sets correspond to the homogeneity properties of the corresponding
principal symbol part.
• The global (i.e.) S -regularity is of course due to the fact that SG-
calculus imposes bounds on the variables also.
Moreover, SG operators satisfy SG microlocalty and microellipticity, i.e.
(just as in the bisingular case):
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
21
Proposition 4.7 (SG Microlocality and SG Microellipticity). Let
u ∈ S ′(Rn) and C ∈ LGm,µ. Then we have the double inclusion
WFSG(Cu) ⊆ WFSG(u) ⊆ CharSG(C) ∪ WFSG(Cu).
This again stems from an individual double inclusion with respect to each
wave front set component. A capital difference lies, however, in the structure
of the wave front set. For the components of the SG wave front set have
WFSG(u) ⊂(cid:0)Rn × (Rn \ 0)(cid:1) ∪(cid:0)(Rn \ 0) × Rn(cid:1) ∪(cid:0)(Rn \ 0) × (Rn \ 0)(cid:1).
Here, the(cid:0)Rn×(Rn\0)(cid:1) ∋ (x, ξ) component corresponds exactly to singular-
ities at finite arguments x with propagation direction ξ, the (cid:0)Rn × (Rn \ 0)(cid:1)
(cid:0)(Rn \ 0) × (Rn \ 0)(cid:1) component corresponds to high oscillations present at
component yields the same interpretation in the Fourier transformed space
(growth singularities of u become differential singularities of u) and the
infinite arguments.
In the bisingular case, new phenomena are present. The 12-component has
the classical interpretation, in the sense that it includes all the 'classical' sin-
gularities (see Lemma 3.3.1), but the other components lose some amount of
localization. This is reflected in the structure of the (1- and 2-components)
of the bi-wave front set. In fact,
WFbi(u) ⊂ (Ω1 × Ω2) × (Rn1+n2 \ 0) = (Ω1 × Ω2) ×(cid:0)(Rn1 \ 0) × {0}(cid:1)
⊔ (Ω1 × Ω2) ×(cid:0){0} × (Rn2 \ 0)(cid:1)
⊔ (Ω1 × Ω2) ×(cid:0)(Rn1 \ 0) × (Rn2 \ 0)(cid:1),
where if e.g. for one x1 we have (x1, x2, 0, ξ2) present in the wave front set,
all (y, x2, 0, ξ2) are present in the wave front set. This is due to the fact that
bi-ellipticity involves true invertibility, i.e. a non-local requirement.
Another difference arises as follows: the 1 and 2-component can be under-
stood as the boundary faces of the 12-component, whereas in the SG case
the ψe-component is interpreted as the corner of the wave front space where
the e- and ψ-component meet, i.e. the roles as boundaries are interchanged,
see Figure 1.
This has the following consequence:
SG(u) and WFψe
SG(u) = (R \ 0) × (R \ 0).
Example 4.8. Consider the one dimensional case. Following here Example
in [CM03], there exists a distribution u(x) = eix2/2, x ∈ R, such that
2.7.
WFψ
SG(u) = ∅ = WFe
However, there cannot exist a distribution v ∈ E ′(Ω1×Ω2), Ω1, Ω2 ⊂ R, such
bi(v) = ∅ = WF2
that WF1
bi (v) = Ω1 × Ω2 × (R \ 0) × (R \ 0).
This is because WF12
bi (v) = Ω1 × Ω2 × (R \ 0) × (R \ 0) is an open set, and
WFbi(v) has to be closed.
Similarly v = δ ⊗ 1 has WF1
bi (v) =
∅ = WF2
SG(u) =
R × (R \ 0) but WFψe
bi(v) = ({0} × Ω2) × ((R \ 0) × {0}), WF12
bi(v), but there cannot exist a distribution u such that WFe
SG(u) = ∅, again due to closedness.
bi(v) and WF12
[Bat12]
U. Battisti, Weyl asymptotics of bisingular operators and Dirichlet divisor
problems, Math. Z. (2012), no. 272, 1365 -- 1381.
References
22
M. BORSERO AND R. SCHULZ
ξ
WFψ
SG
WFψe
SG
ξ2
WF2
bi
WF12
bi
WFe
SG
x
WF1
bi
ξ1
WFSG
WFbi
Figure 1. A schematic comparison of the components of
WFSG and WFbi
[BGPR13] U. Battisti, T. Gramchev, S. Pilipovi´c, and L. Rodino, Globally Bisingular El-
liptic Operators, Operator Theory, Pseudo-Differential Equations, and Math-
ematical Physics, Operator Theory: Advances and Applications, vol. 228,
Birkhauser, Basel, 2013.
[CJT13a] S. Coriasco, K. Johansson, and J. Toft, Global wave-front sets of Banach,
Fr´echet and Modulation space types, and pseudo-differential operators, J. of
Diff. Eq. (2013).
[CJT13b]
[CM03]
[Cor95]
, Local wave-front sets of Banach and Fr´echet types, and pseudo-
differential operators, Monatshefte fur Mathematik 169 (2013), no. 3-4, 285 --
316.
S. Coriasco and L. Maniccia, Wave Front Set at Infinity and Hyperbolic Lin-
ear Operators with Multiple Characteristics, Ann. Global Anal. Geom. (2003),
no. 24, 375 -- 400.
H. O. Cordes, The Technique of Pseudodifferential Operators, Cambridge Uni-
versity Press, 1995.
[Dud79a] R. V. Duducava, On the index of bisingular integral operators. I., Math. Nachr.
(1979), no. 91, 431 -- 460.
[Dud79b]
, On the index of bisingular integral operators. II., Math. Nachr. (1979),
[ES97]
[FST98]
[GH13]
[GS94]
[Hor83]
[Hor85]
[Hor91]
no. 92, 289 -- 307.
Y. V. Egorov and B.-W. Schulze, Pseudo-differential operators, singularities,
applications, Operator Theory: Advances and Applications, vol. 91, Birkhauser
Verlagl, Basel, 1997.
B.V. Fedosov, B.-W. Schulze, and N.N. Tarkhanov, On the index of elliptic
operators on a wedge, J. Funct. Anal. (1998), no. 156, 164 -- 209.
D. Grieser
Laplacian on Q-rank 1 Locally Symmetric Spaces, Preprint (2013),
http://arxiv.org/abs/1212.3459v2.
A. Grigis and J. Sjostrand, Microlocal Analysis for Differential Operators, An
Introduction, Cambridge University Press, 1994.
L. Hormander, The Analysis of Linear Partial Differential Operators, vol. I,
Springer-Verlag, 1983.
and E. Hunsicker, A Parametrix Construction for
the
in
, The Analysis of Linear Partial Differential Operators, vol. III,
Springer-Verlag, 1985.
, Quadratic hyperbolic operators, Microlocal Analysis and Applications
(L. Cattabriga and L. Rodino, eds.), Lecture Notes in Mathematics, vol. 1495,
Springer Berlin Heidelberg, 1991, pp. 118 -- 160.
MICROLOCAL PROPERTIES OF BISINGULAR OPERATORS
23
[Mel94]
[Mel95]
[MR06]
[NR06]
[Par72]
[Pil71]
[Pil73]
[Rod75]
R. Melrose, Spectral and scattering theory for the laplacian on asymptotically
euclidian spaces, 1994.
, Geometric scattering theory. stanford lectures, Cambridge University
Press, 1995.
R. Melrose and F. Rochon, Index in K-theory for families of fibred cusp oper-
ators, K-theory 37 (2006), pp.25 -- 104.
F. Nicola and L. Rodino, Residues and Index for Bisingular Operators, In:
C ∗-algebras and elliptic theory, Trends Math., Birkhauser (2006), pp.187 -- 202.
C. Parenti, Operatori pseudodifferenziali in Rn e applicazioni, Ann. Mat. Pura
Appl. 93 (1972), 359 -- 389.
V. S. Pilidi, Multidimensional bisingular operators, Dokl. Akad. Nauk SSSR
(1971), no. 201, 787 -- 789.
, Computation of the index of a bisingular operator., Funkcional. Anal.
i Prilozen. (1973), no. 7, 93 -- 94.
L. Rodino, A class of pseudo differential operators on the product of two man-
ifolds and applications, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (1975), no. 4,
287 -- 302.
[Rod80]
, Polysingular integral operators, Ann. Mat. Pura Appl. (1980), no. 124,
[Sch87]
59 -- 106.
E. Schrohe, Spaces of weighted symbols and weighted sobolev spaces on man-
ifolds, Pseudo-Differential Operators, Lecture Notes in Math., vol. 1256,
Springer, Berlin, 1987, pp. 360 -- 377.
[Shu01] M. A. Shubin, Pseudodifferential Operators and Spectral Theory, second ed.,
[Tr`e67]
Springer-Verlag, 2001.
F. Tr`eves, Topological Vector Spaces, Distributions, and Kernels, Academic
Press, 1967.
Dipartimento di Matematica "Giuseppe Peano", Universit`a degli Studi di
Torino, Via Carlo Alberto 10, 10123 Torino (TO), Italy.
E-mail address: [email protected]
Mathematisches Institut, Georg-August Universitat Gottingen, Bunsen-
strasse 3 -- 5, D -- 37073 Gottingen, Germany
E-mail address: [email protected]
|
1809.02593 | 1 | 1809 | 2018-09-07T17:37:21 | Cyclic row contractions and rigidity of invariant subspaces | [
"math.FA",
"math.OA"
] | It is known that pure row contractions with one-dimensional defect spaces can be classified up to unitary equivalence by compressions of the standard $d$-shift acting on the full Fock space. Upon settling for a softer relation than unitary equivalence, we relax the defect condition and simply require the row contraction to admit a cyclic vector. We show that cyclic pure row contractions can be "transformed" (in a precise technical sense) into compressions of the standard $d$-shift. Cyclic decompositions of the underlying Hilbert spaces are the natural tool to extend this fact to higher multiplicities. We show that such decompositions face multivariate obstacles of an algebraic nature. Nevertheless, some decompositions are obtained for nilpotent commuting row contractions by analyzing function theoretic rigidity properties of their invariant subspaces. | math.FA | math |
CYCLIC ROW CONTRACTIONS AND
RIGIDITY OF INVARIANT SUBSPACES
RAPHAEL CLOU ATRE AND EDWARD J. TIMKO
Abstract. It is known that pure row contractions with one-dimensional de-
fect spaces can be classified up to unitary equivalence by compressions of the
standard d-shift acting on the full Fock space. Upon settling for a softer rela-
tion than unitary equivalence, we relax the defect condition and simply require
the row contraction to admit a cyclic vector. We show that cyclic pure row
contractions can be "transformed" (in a precise technical sense) into compres-
sions of the standard d-shift. Cyclic decompositions of the underlying Hilbert
spaces are the natural tool to extend this fact to higher multiplicities. We show
that such decompositions face multivariate obstacles of an algebraic nature.
Nevertheless, some decompositions are obtained for nilpotent commuting row
contractions by analyzing function theoretic rigidity properties of their invari-
ant subspaces.
1. Introduction
The structure of Hilbert space contractions has long been recognized as deeply
connected to that of concrete multiplication operators on spaces of analytic func-
tions on the unit disc. The prototypical result in that direction is the fact that pure
contractions may be viewed as compressions to co-invariant subspaces of unilateral
shifts. This crucial fact was fleshed out significantly in the original edition of [37],
and the resulting theory has been influential since then.
A major theme in modern operator theory is the investigation of multivariate
phenomena, prompted by the discovery of curious behaviour [30],[38]. The multi-
variate counterpart to the Sz.-Nagy -- Foias theory was systematically developed by
Popescu in a series of papers, starting with [32],[33]. The setting in this case is that
of row contractions, with no commutation relation assumed between the individual
operators. The non-commutative multivariate theory runs parallel to the univariate
one, with most classical results finding a close analogue. The key insight into the
commutative world was provided by Arveson in [5] who identified the appropriate
space of analytic functions on the ball to serve as a replacement for the classical
Hardy space. This space is now known as the Drury-Arveson space and is the
centerpiece of many problems in function theory due its universal property [1]. It
should also be pointed out that the commutative world can in fact be recovered ef-
ficiently from the non-commutative one upon taking appropriate quotients, as laid
out in [21],[35].
Going back to the foundational single-variable result, it is known that if T is
a Hilbert space contraction with one-dimensional defect space (in the sense that
I − T T ∗ has rank one), then it is unitarily equivalent to a compression of the
2010 Mathematics Subject Classification. 47A13, 47A20, 47A45 (47B32, 46L52).
The first author was partially supported by an NSERC Discovery Grant.
1
2
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
standard unilateral shift on the Hardy space of the disc. In this case, the connection
with function theory is especially transparent, and tools developed in the setting
of the Hardy space can be brought to bear quite successfully to understand the
structure of T . There are corresponding results in the multivariate setting [32],[29]
that connect the study of row contractions to complex function theory in several
variables.
Unfortunately, the condition on the defect space being one-dimensional is quite
restrictive, and it is not stable under natural perturbations such as scaling or simi-
larity. It is thus desirable to replace it by a more flexible condition, such as cyclicity.
The tradeoff, then, is that we cannot expect to describe T up to unitary equiva-
lence with compressions of the standard shift anymore. This relaxation idea is the
basis for the very successful Jordan model theory of constrained contractions (also
known as C0-contractions) developed in [8]. The crucial result is that a cyclic pure
contraction can be "transformed" into a compression of the standard shift. The
multivariate analogue of this relaxation procedure seems to have been overlooked,
and providing it is one of the main motivation behind this project. It should be
noted here that there is a well-developed model theory for arbitrary row contrac-
tions that takes place in spaces of vector-valued holomorphic functions [37, Theorem
VI.2.3],[12],[31],[35]. However, the focus there is different than ours, and the two
strands of the theory do not overlap beyond the case of one-dimensional defect
spaces.
Of course, it is desirable to deal with contractions T which are not necessarily
cyclic as well. The natural approach for doing so is to decompose the underlying
Hilbert space into cyclic subspaces that are invariant for T , and then apply the
cyclic result on each piece. In [8], this is called the splitting principle, and allows
the cyclic theory to be leveraged in the general case of arbitrary multiplicity and
to culminate in a powerful classification theorem that mirrors the Jordan canonical
form for matrices. Inspired by this success, at the onset of this project another
one of our objectives was to obtain such cyclic decompositions in the multivariate
context. Surprisingly, this ambition was met with serious obstacles, as we unraveled
multivariate obstructions of a seemingly purely algebraic nature. This is an inter-
esting discovery, and shows that if a satisfactory model theory is to emerge in this
context, it will have to rely on entirely new ideas and tools. At least superficially,
this is reminiscent of what prompted the work in [7].
Although cyclic decompositions appear elusive in general, we obtain some lim-
ited partial results in this direction based on considerations of independent interest
on function theoretic rigidity properties of invariant subspaces. More precisely, if
M is an invariant subspace for a row contraction T , we are interested in describ-
ing situations where M is completely determined by the kernel of the functional
calculus associated to the restriction TM. We also explore the corresponding ques-
tion for co-invariant subspaces. Such inquiries generate significant interest in the
setting of Hilbert modules on reproducing kernel Hilbert spaces; see [14] and the
references therein. Our work on this topic initiates a multivariate exploration of
equivalence classes of invariant subspaces for constrained contractions, as studied
in [11],[9],[10],[28],[15],[16].
We now describe the organization of the paper and state our main results. To
do so, we require the following notation. For a pure commuting row contraction
T , we denote by Annc(T ) the set of multipliers ϕ on the Drury-Arveson space
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
3
with the property that ϕ(T ) = 0. More detail can be found in Section 2, which
contains background material and gathers many preliminary results that are used
throughout.
In Section 3, we investigate the class of commuting row contractions T that can
be transformed into the standard commuting d-shift. Our analysis is based on the
existence of cyclic vectors enjoying certain additional properties (Theorem 3.1).
One of the main results of the paper is the following (see Corollary 3.7).
Theorem 1.1. Let T = (T1, . . . , Td) be a pure, commuting, cyclic row contraction
on some Hilbert space. Let M denote the compression of the standard commuting
d-shift associated to Annc(T ). Then, we have that M is a quasi-affine transform of
T .
Interestingly, despite the commuting nature of the statement, our proof hinges
crucially on the non-commuting dilation results of [32].
In Section 4, we turn to the question of rigidity of invariant and co-invariant
subspaces. For general non-commuting row contractions, this problem is known
to encounter significant difficulties, so we focus on the commuting setting. We
illustrate that the rigidity witnessed in the univariate context can fail in several
variables (Example 2). Nevertheless, some amount of rigidity can be established
in various cases.
In the special case of nilpotent row contractions, our results
(Theorems 4.1 and 4.5) imply the following.
Theorem 1.2. Let T = (T1, . . . , Td) be a nilpotent commuting row contraction on
some Hilbert space H. The following statements hold.
(1) Assume that T ∗ is cyclic. Let M, N ⊂ H be invariant subspaces for T such
(2) Assume that T is cyclic. Let M ⊂ H be an invariant subspace for T such
that Annc(TM) = Annc(TN). Then, M = N.
that Annc(TM) = Annc(T ). Then, M = H.
It is relevant to mention here that it is not generally true that T ∗ is cyclic if
T is a cyclic commuting nilpotent row contraction (see Example 5). Regarding
co-invariant subspaces, we mention here the finite-dimensional version of Theorem
4.3.
Theorem 1.3. Let T = (T1, . . . , Td) be a cyclic pure commuting row contraction on
some finite-dimensional Hilbert space H. Let M, N ⊂ H be co-invariant subspaces
for T such that Annc(PMTM) = Annc(PNTN). Then, M = N.
Finally, in Section 5 we examine the existence of invariant decompositions. If
T = (T1, . . . , Td) is a d-tuple of operators on some Hilbert space H, a pair (M, N)
of non-trivial invariant subspaces for T1, . . . , Td is an invariant decomposition for
T if M ∩ N = {0} and M + N = H. Building on the results of Section 4, Theorem
5.1 establishes the existence of invariant decompositions for nilpotent commuting
row contractions under some additional conditions. In another direction, Example
3 (an adaptation of an idea from [25]) exhibits a non-cyclic nilpotent commuting
row contraction that admits no invariant decomposition with the property that
TM is cyclic. This is surprising, as such cyclic decompositions always exist for
single constrained contractions. We shed some light on these difficulties by con-
sider separating vectors. These are known to be abundant for single constrained
contractions, yet we show that they can fail to exist in several variables, even for
4
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
nilpotent commuting row contractions on finite-dimensional spaces (Example 5).
On the other hand, we have the following (Theorem 5.3).
Theorem 1.4. Let T = (T1, . . . , Td) be a nilpotent commuting row contraction on
some Hilbert space H. Set
A = C[x1, . . . , xd]/(C[x1, . . . , xd] ∩ Annc(T )).
Then, for any positive integer s ≥ dimA, the ampliation T (s) has a dense set of
separating vectors in H(s).
One of the recurring themes of the paper is that nilpotent commuting row con-
tractions exhibit rich and intricate behaviours that are somehow unique to the
multivariate world. Indeed, as we will see throughout, such row contractions give
rise to multivariate counter-examples to classical univariate theorems of interest.
This is the motivation behind our focusing on the nilpotent case in several of our
main results above. While some of our proofs can doubtlessly be adapted to the
wider class of commuting row contractions satisfying other types of polynomial
relations, we do not to pursue such generalizations in this paper.
2. Preliminaries
2.1. The Fock spaces, the Drury-Arveson space and some associated op-
erator algebras. Throughout the paper, d is a fixed positive integer. Let F+
d
denote the free semigroup on the symbols {1, . . . , d}. Let F2
d denote the full Fock
space over Cd, that is
F2
d = CΩ ⊕
(Cd)⊗n
∞Mn=1
where Ω is some distinguished unit vector. Fix an orthonormal basis {e1, . . . , ed}
of Cd. If w = k1k2 . . . ks ∈ F+
d , then we set
ew = ek1 ⊗ ek2 ⊗ . . . ⊗ eks ∈ F2
d.
The set {ew : w ∈ F+
Lk : F2
d → F2
d } is an orthonormal basis of F2
d denote the left creation operator acting as
d. For each 1 ≤ k ≤ d, we let
Lkv = ek ⊗ v,
v ∈ F2
d.
It is readily verified that the operators L1, . . . , Ld are isometries with pairwise
orthogonal ranges, so that
dXk=1
LkL∗
k ≤ I.
The weak-∗ closed unital subalgebra of B(F2
d) generated by L1, . . . , Ld is denoted
by Ld, and is usually referred to as the non-commutative analytic Toeplitz algebra
[33],[20],[21]. Likewise, the norm closed unital subalgebra of B(F2
d) generated by
L1, . . . , Ld is denote by Ad, and is usually called the non-commutative disc algebra
[33].
The subspace Fs
d ⊂ F2
d spanned by the symmetric elementary tensors is the
symmetric Fock space, and it is co-invariant for L1, . . . , Ld. It was recognized by
Arveson [5] that Fs
d can be identified with a space of analytic functions on the open
unit ball Bd ⊂ Cd. Nowadays, this space H 2
d is called the Drury-Arveson space.
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
5
For each 1 ≤ k ≤ d, we denote by xk : Cd → C the coordinate function, so
that xk(z) = zk for z = (z1, . . . , zd) ∈ Cd. We write N for the set of non-negative
integers. We use the standard multi-index notation: for
The Drury-Arveson space H 2
with powers series expansion
d is the Hilbert space of analytic functions f on Bd
we set
and
satisfying
α = (α1, . . . , αd) ∈ Nd
α = α1 + ··· + αd,
α! = α1!··· αd!
1 ··· xαd
xα = xα1
d .
f = Xα∈Nd
cαxα
α!
α1!··· αd!
Xα∈Nd
α1!··· αd!(cid:19)1/2
α!
(cid:18)
cα2 < ∞.
xα, α ∈ Nd.
An orthonormal basis for H 2
d is given by the weighted monomials
See [5] or [21] for more details. Another useful interpretation of H 2
reproducing kernel Hilbert space on Bd (see [2]) with kernel given by the formula
d is that it is the
k(z, w) =
1
,
z, w ∈ Bd.
1 − hz, wi
We denote by Md the multiplier algebra of H 2
there is a bounded linear operator Mϕ ∈ B(H 2
Mϕf = ϕf,
d . Associated to a multiplier ϕ ∈ Md
d ) defined as
f ∈ H 2
d .
Identifying each multiplier in Md with its associated multiplication operator, we
may view Md as a weak-∗ closed unital subalgebra of B(H 2
d ). In particular, Md is
an operator algebra. Under this identification, the multiplier norm of ϕ ∈ Md is
simply the operator norm of Mϕ. Every polynomial is a multiplier, and we denote
by Ad the norm closure of the polynomials in Md. We will require the following
basic fact about multipliers.
Theorem 2.1. Let ϕ ∈ Md. Fix n ∈ N and w ∈ Bd. For each α ∈ Nd with
α = n, there exists a multiplier ϕα ∈ Md with the property that
(x − w)αϕα.
∂αϕ
∂xα (w)
(x − w)α
α!
ϕ = Xα<n
+ Xα=n
Proof. Apply [24, Cor. 4.2] inductively.
(cid:3)
Next, we record useful identifications which are well known. The details of the
proof appear to be difficult to track down explicitly, so we provide them for the
reader's convenience.
Theorem 2.2. The following statements hold.
6
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
(1) There is a unitary operator U : H 2
LkFs
U Mxk U ∗ = PFs
d
d → Fs
,
d
d with the property that
1 ≤ k ≤ d.
(2) Let Wd ⊂ Ld denote the weak-∗ closure of the commutator ideal of Ld.Then,
there is a unital completely isometric and weak-∗ homeomorphic algebra
isomorphism
such that
Ψ : Ld/Wd → Md
Ψ(Lk + Wd) = U ∗(PFs
d
LkFs
d
)U = Mxk
for every 1 ≤ k ≤ d.
there is a unital completely isometric algebra isomorphism
(3) Let Cd ⊂ Ad denote the norm closure of the commutator ideal of Ad. Then,
such that
Φ : Ad/Cd → Ad
Φ(Lk + Cd) = Mxk ,
1 ≤ k ≤ d.
Proof. (1) The existence of U follows from the discussion in [5, Section 1].
(2) This follows from (1) and [21, Corollary 2.3].
(3) This argument is based on the idea behind the proof of [19, Corollary 8.3].
Let q : Ad → Ad/Cd denote the quotient map. By the Blecher -- Ruan -- Sinclair
characterization of operator algebras [13], there is a Hilbert space H and a unital
completely isometric homomorphism π : Ad/Cd → B(H). Then,
(π(q(L1)), . . . , π(q(Ld)))
is a commuting row contraction on H. In particular, invoking [5, Theorem 6.2] for
instance, we obtain a unital completely contractive homomorphism
such that σ(Mxk ) = π(q(Lk)) for every 1 ≤ k ≤ d. Next, let ι : Ad/Cd → Ld/Wd
be defined as
σ : Ad → π(Ad/Cd)
It is readily checked that Cd is the norm closure of the ideal generated by
ι(T + Cd) = T + Wd,
T ∈ Ad.
{LjLk − LkLj : 1 ≤ j, k ≤ d}
inside of Ad. Furthemore, it follows from [20, Proposition 2.4] that Wd is the weak-∗
closure of the ideal generated by
{LjLk − LkLj : 1 ≤ j, k ≤ d}
inside of Ld. Hence, Cd ⊂ Wd so that ι is a well-defined unital completely contrac-
tive homomorphism. We note that
(Ψ ◦ ι ◦ π−1 ◦ σ)(Mxk ) = Mxk
for every 1 ≤ k ≤ d, whence
Ψ ◦ ι ◦ π−1 ◦ σ
is simply the inclusion of Ad inside Md. This forces π−1 ◦ σ : Ad → Ad/Cd to be
completely isometric and in particular surjective, since its image contains Lk + Cd
for every 1 ≤ k ≤ d. Take Φ = π−1 ◦ σ.
(cid:3)
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
7
When d = 1, the space H 2
1 coincides with the usual Hardy space on the open unit
disc D. In this case, the operator algebras A1 and M1 can be identified with the disc
algebra A(D) and the algebra of bounded analytic functions H ∞(D), respectively.
It is important to note that when d > 1, however, the quantities
kMϕk
and kϕk∞ = sup
z∈Bd ϕ(z)
are in general not comparable [5],[21]. This makes the function theory in H 2
d and
Md less transparent than it is in the algebra of bounded holomorphic functions
H ∞(Bd).
2.2. Dilations to d-shifts. We recall the main features of the multivariate gener-
alization of the classical Sz.-Nagy -- Foias model theory from [37], starting with the
non-commuting setting. define the standard d-shift to be the d-tuple of operators
L = (L1, L2, . . . , Ld) acting on F2
d. More generally, a d-tuple V = (V1, . . . , Vd) of
operators on some Hilbert space H is called a d-shift if there is another Hilbert
space K and a unitary operator U : H → F2
U VkU ∗ = Lk ⊗ IK,
d ⊗ K such that
1 ≤ k ≤ d.
The cardinal number dim K is then called the multiplicity of the d-shift. Interest-
ingly, the multiplicity of a d-shift can be interpreted in a different way as the next
theorem shows. Recall that a subset Γ ⊂ H is cyclic for a d-tuple T = (T1, . . . , Td)
if
Here, given a word w = i1i2 . . . is ∈ F+
d , we use the notation
H = span{Twγ : w ∈ F+
d , γ ∈ Γ}.
Tw = Ti1 Ti2 . . . Tis.
Theorem 2.3. Let V = (V1, . . . .Vd) be a d-shift. Then, the multiplicity of V
coincides with the least cardinality of a cyclic set for V .
Proof. This follows from [35, Corollary 1.10].
(cid:3)
The importance of d-shifts lies in the fact that they can be used to model gen-
eral d-tuples of operators satisfying some natural necessary conditions, as we now
illustrate.
A d-tuple T = (T1, . . . , Td) of operators on some Hilbert space H is said to be a
row contraction if the row operator
defined as
[T1, . . . , Td]ξ =
ξ = (ξ1, . . . , ξd) ∈ H(d)
[T1, . . . , Td] : H(d) → H
dXk=1
H(d) = H ⊕ H ⊕ . . . ⊕ H.
Tkξk,
is contractive. Here, we use the standard notation
Equivalently, T is a row contraction if and only if
dXk=1
TkT ∗
k ≤ I.
Furthermore, we define the defect space of T to be
lim
kT ∗
n→∞ Xw=n
∆T = I −
wξk2 = 0,
ξ ∈ H.
k! H,
TkT ∗
dXk=1
8
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
The row contraction T is said to be pure if
and the defect of T to be
dT = dim ∆T .
A straightforward calculation reveals that d-shifts are pure row contractions. The
following result shows that in some sense they are the universal example.
Theorem 2.4. Let T = (T1, . . . , Td) be a d-tuple of operators on some Hilbert space
H. Then, the following statements are equivalent.
(i) The d-tuple T is a pure row contraction.
(ii) There is another Hilbert space K containing H along with a d-shift V =
(V1, . . . , Vd) on K such that
In this case, the multiplicity of the d-shift V is dT .
T ∗
k = V ∗
k H,
1 ≤ k ≤ d.
Proof. This is [32, Theorems 2.1 and 2.8].
(cid:3)
Of particular interest for us in this paper are row contractions T = (T1, . . . , Td)
with the property that T1, . . . , Td commute with each other. In this case, we say
that T is a commuting row contraction. While Theorem 2.4 applies just as well to
commuting row contractions, it fails to encode the commutativity. In particular,
one may want to replace the d-shift therein by some "commutative" version which
acts on the symmetric Fock space (or Drury-Arveson space), as opposed to the
full Fock space. To make this precise, recall that for each 1 ≤ k ≤ d the operator
Mxk ∈ B(H 2
d ) denotes the multiplication operator by the coordinate function xk.
A standard calculation reveals that the commuting d-tuple Mx = (Mx1, . . . , Mxd)
is a pure row contraction. We will refer to this commuting row contraction as the
standard commuting d-shift. More generally, a commuting d-tuple Z = (Z1, . . . , Zd)
of operators on some Hilbert space H is called a commuting d-shift if there is another
Hilbert space K and a unitary operator U : H → H 2
d ⊗ K such that
U ZkU ∗ = Mxk ⊗ IK,
1 ≤ k ≤ d.
The cardinal number dim K is then called the multiplicity of the commuting d-shift.
Just like their non-commuting counterparts, commuting d-shifts occupy a universal
role among pure commuting row contractions.
Theorem 2.5. Let T = (T1, . . . , Td) be a d-tuple of operators on some Hilbert space
H. Then, the following statements are equivalent.
(i) The d-tuple T is a pure commuting row contraction.
(ii) There is another Hilbert space K containing H along with a commuting d-
shift Z = (Z1, . . . , Zd) on K such that
In this case, the multiplicity of the commuting d-shift Z is dT .
k = Z ∗
T ∗
kH,
1 ≤ k ≤ d.
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
Proof. This is [5, Theorem 4.5] (alternatively, see [35, Theorem 2.1]).
9
(cid:3)
2.3. Functional calculus and constrained row contractions. In [33], Popescu
extends the classical von Neumann inequality to the setting of row contractions and
defines an appropriate functional calculus, as follows.
Theorem 2.6. Let T = (T1, . . . , Td) be a row contraction on some Hilbert space H.
Then, there is a unital completely contractive homomorphism αT : Ad → B(H) such
that αT (Lk) = Tk for every 1 ≤ k ≤ d. If T is pure, then αT extends to a unital
completely contractive and weak-∗ continuous homomorphism βT : Ld → B(H).
Proof. Combine [33, Theorems 3.6 and 3.9] with [34, Theorem 2.1].
(cid:3)
There is a version of this result adapted to the commuting setting, and the
resulting homomorphisms are naturally compatible.
Corollary 2.7. Let T = (T1, . . . , Td) be a commuting row contraction on some
Hilbert space H. Let Cd ⊂ Ad and Wd ⊂ Ld denote respectively the norm closure
and the weak-∗ closure of the commutator ideals. Let q : Ad → Ad/Cd and qw :
Ld → Ld/Wd be the corresponding quotient maps. Let Φ : Ad/Cd → Ad and
Ψ : Ld/Wd → Md be the canonical isomorphisms from Theorem 2.2. Then, the
following statements hold.
T : Ad/Cd → B(H) such that αT = α′
T ◦ Φ−1, and (1) follows.
T : Ld/Wd → B(H) such that βT = β′
(cid:3)
(1) There is a unital completely contractive homomorphism cαT : Ad → B(H)
such that cαT (Mxk ) = Tk for every 1 ≤ k ≤ d. Moreover, we have αT =
cαT ◦ Φ ◦ q.
(2) If T is pure, then cαT extends to a unital completely contractive and weak-∗
continuous homomorphism cβT : Md → B(H). Moreover, we have βT =
cβT ◦ Ψ ◦ qw.
Proof. Since T is commuting, we see that Cd ⊂ ker αT , so there is a unital com-
pletely contractive homomorphism α′
T ◦ q. Put
For (ii), we note that Wd ⊂ ker βT , so there is a unital completely contractive
homomorphism β′
T ◦ qw. A standard compact-
ness argument using the Krein-Smulyan theorem shows that β′
T is in fact weak-∗
T ◦ Ψ−1.
cαT = α′
continuous. Statement (2) follows upon setting cβT = β′
We typically make no explicit mention of the homomorphisms αT ,cαT , βT and
cβT defining the functional calculi, and for an appropriate element θ we simply write
The existence of a weak-∗ continuous extension to Md of the Ad -- functional
calculus for a commuting row contraction does not require T to be pure. Indeed,
the class of so-called absolutely continuous commuting row contractions is larger
[17],[18]. Nevertheless, for the purposes of this work we will restrict our attention
to the pure case.
Let T = (T1, . . . , Td) be a pure row contraction. Then, the non-commutative
annihilator of T is the weak-∗ closed two-sided ideal of Ld given by
Annnc(T ) = {θ ∈ Ld : θ(T ) = 0}.
We say that T is constrained if Annnc(T ) is non-trivial. If T happens to be com-
muting, then its commutative annihilator is the weak-∗ closed ideal of Md given
θ(T ).
10
by
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
It follows from Corollary 2.7 that
Annc(T ) = {θ ∈ Md : θ(T ) = 0}.
so in particular
(1)
Annnc(T ) = (Ψ ◦ qw)−1(Annc(T ))
Annc(T ) = (Ψ ◦ qw)(Annnc(T )).
In the classical case where d = 1, a constrained contraction T is usually said to be
of class C0 [8]. We feel our choice of terminology is a bit more descriptive, and it
is consistent with its use in [35]. We also mention that an absolutely continuous
contraction for which the associated H ∞ -- functional calculus has non-trivial kernel
is automatically pure, and thus constrained. To see this, combine the proof of
[8, Lemma II.1.12] with [17, Theorem 3.2]. For commuting row contractions, the
situation is more complicated: see [17, Theorem 5.4] and the discussion that follows
it. Let b ⊂ Ld be a weak-∗ closed two-sided ideal. We let
d] = span{θv : θ ∈ b, v ∈ F2
d}
[bF2
and
Then, Fb is co-invariant for Ld + L′
Ld. Define
Fb = F2
d ⊖ [bF2
d].
d, where L′
d ⊂ B(F2
d) denotes the commutant of
Lb = (PFb L1Fb , . . . , PFb L1Fb ).
It is readily verified that Lb is a pure row contraction, and it is a consequence of
the following result that Annnc(Lb) = b. Thus, the row contraction Lb provides a
basic example of a constrained row contraction.
Lemma 2.8. Let K ⊂ F2
d be a closed subspace which is co-invariant for Ld + L′
d.
Let b = Annnc(PKL1K, . . . , PKLdK). Then, b is the unique weak-∗ closed two-sided
ideal of Ld with the property that K = Fb.
Proof. It follows from [20, Theorem 2.1] that there is a weak-∗ closed two-sided
ideal c ⊂ Ld such that F2
d], whence K = Fc. It only remains to show
uniqueness. Assume that a ⊂ Ld is a weak-∗ closed two-sided ideal such that
d ⊖ K = [aF2
F2
d]. Since b was chosen to be the annihilator of (PKL1K, . . . , PKL1K),
we see that PKθK = 0 for every θ ∈ b. Using that K is co-invariant for Ld we obtain
d ⊂ F2
θF2
d ⊖ K for every θ ∈ b. We conclude that
d ⊖ K = [cF2
[bF2
d] ⊂ F2
d ⊖ K = [aF2
d].
An application of [20, Theorem 2.1] yields b ⊂ a. On the other hand, it is immediate
that a annihilates (PKL1K, . . . , PKLdK), whence a ⊂ b.
(cid:3)
We mention here that the condition that K be co-invariant for L′
d in the previous
Indeed, by [27, Theorem 3.7] there is a proper
lemma cannot simply removed.
closed subspace K ⊂ F2
d which is co-invariant for Ld and which satisfies
{θ ∈ Ld : θΩ ⊂ K⊥} = {0}
so in particular we see that
{θ ∈ Ld : θF2
d ⊂ K⊥} = {0}.
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
11
Thus
Annnc(PKL1K, . . . , PKL1K) = {0},
yet K 6= F2
d.
ideal and let
We now examine the commuting case. Let J ⊂ Md be a weak-∗ closed two-sided
HJ = H 2
d ⊖ [J H 2
d ]
In this setting, the analogue of Lemma 2.8 holds without additional conditions on
the co-invariant subspace.
Lemma 2.9. Let K ⊂ H 2
d be a closed subspace which is co-invariant for Md. Let
J = Annc(PKMx1K, . . . , PKMxdK). Then, J is the unique weak-∗ closed ideal of
Md with the property that K = HJ .
Proof. Repeat the argument used in the proof of Lemma 2.8, invoking [23, Theorem
2.4] rather than [20, Theorem 2.1].
(cid:3)
Define a pure commuting row contraction as
We record an important property of MJ .
MJ = (PHJ Mx1HJ , . . . , PHJ MxdHJ ).
Lemma 2.10. Let J ⊂ Md be a weak-∗ closed ideal. Let
b = (Ψ ◦ qw)−1(J ) ⊂ Ld.
Then, MJ is unitarily equivalent to Lb and Annc(MJ ) = J .
Proof. By virtue of Lemma 2.9 we see that Annc(MJ ) = J . Next, it follows from
Theorem 2.2 that
UJ U ∗ = {PFs
d
XFs
d
: (Ψ ◦ qw)(X) ∈ J } = {PFs
d
XFs
d
: X ∈ b}
whence
[bF2
d].
Clearly, we see that Wd ⊂ b. Invoking [20, Proposition 2.4], we find
Xv : X ∈ b, v ∈ Fs
d ] = span{PFs
d} = PFs
U [J H 2
d
d
(Fs
d)⊥ = [WdF2
d] ⊂ [bF2
d]
and thus
In particular, we find
[bF2
d] = U [J H 2
d ] ⊕ (Fs
d)⊥.
Fb = [bF2
d]⊥ = Fs
d ⊖ U [J H 2
d ] = U (H 2
d ⊖ [J H 2
d ]) = U HJ .
We conclude that MJ is unitarily equivalent to Lb via the operator U .
(cid:3)
2.4. Quasi-affine transforms. Let T = (T1, . . . , Td) and T ′ = (T ′
d) be two
d-tuples of operators acting on Hilbert spaces H and H′ respectively. We say T ′ is
a quasi-affine transform of T and write T ′ ≺ T , if there exists an injective bounded
linear operator X : H′ → H with dense range such that XT ′
k = TkX for 1 ≤ k ≤ d.
The relationship T ′ ≺ T is a common weakening of similarity that we will play a
prominent role in our work.
1, . . . , T ′
12
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
Lemma 2.11. Let H and H′ be Hilbert spaces and let X : H′ → H be a bounded
linear operator. Let T ∈ B(H) and T ′ ∈ B(H′) such that XT ′ = T X. Then, the
subspace H′ ⊖ ker X is co-invariant for T ′ and
In particular, if X has dense range, then
XPH′⊖ker X T ′H′⊖ker X = T XH′⊖ker X .
PH′⊖ker X T ′H′⊖ker X ≺ T.
Proof. Note that
whence T ′ ker X ⊂ ker X. We conclude that H′ ⊖ ker X is co-invariant for T ′.
Furthermore, we have
XT ′ ker X = T X ker X = {0}
XPH′⊖ker X T ′H′⊖ker X = XT ′H′⊖ker X = T XH′⊖ker X .
Finally, since XH′⊖ker X : H′ ⊖ ker X → H is injective and has the same range as
X, we conclude that
whenever X has dense range.
PH′⊖ker X T ′H′⊖ker X ≺ T
(cid:3)
Annihilators are well behaved under quasi-affine transforms, and we record the
following easy fact.
Lemma 2.12. Let T = (T1, . . . , Td) and T ′ = (T ′
d) be pure row contractions
on some Hilbert spaces H and H′. Let X : H′ → H be a bounded linear operator
such that XT ′
k = TkX for every 1 ≤ k ≤ d. Then, following statements hold.
1, . . . , T ′
(1) If X is injective, then Annnc(T ) ⊂ Annnc(T ′).
(2) If X has dense range, then Annnc(T ′) ⊂ Annnc(T ).
(3) If X is injective and has dense range, then Annnc(T ) = Annnc(T ′).
Proof. This easily follows from the fact that for every θ ∈ Ld we have Xθ(T ′) =
θ(T )X.
(cid:3)
3. Transforming the standard d-shift
In order to understand general pure row contractions, it is natural to turn to The-
orem 2.4. This result reduces the problem to the study of compressions of d-shifts
to co-invariant subspaces. Such a strategy is appealing, as the non-commutative
function theory of Ld that was significantly developed in [20],[21],[22] could then be
leveraged, much in the way that the function theory of H ∞(D) has helped single
operator theory blossom [37]. Unfortunately, in general the d-shift from Theorem
2.4 is of high multiplicity, and thus the link with Ld is obscured. In this section, to
circumvent this difficulty we investigate the possibility for d-shifts of multiplicity
one to be quasi-affine transforms of pure row contractions, taking inspiration from
and generalizing a known result for single operators [8, Theorem I.3.7].
We first exhibit a necessary and sufficient condition in the form of the existence of
certain cyclic vectors, the images of which satisfy some weak form of orthogonality.
Recall that given vectors ξ, η ∈ H, we denote by ξ⊗ η ∈ B(H) the rank-one operator
defined as
Theorem 3.1. Let T = (T1, . . . , Td) be a row contraction on some Hilbert space H
and let C > 0. The following statements are equivalent.
(ξ ⊗ η)h = hh, ηiξ,
h ∈ H.
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
13
(i) There is a cyclic vector ξ ∈ H for T that satisfies
Xw∈F+
d
Twξ ⊗ Twξ ≤ C2I.
(ii) There exists an operator X : F2
and XLk = TkX for 1 ≤ k ≤ d.
injective operator X : V → H with dense range such that kXk ≤ C and
d → H with dense range such that kXk ≤ C
d which is co-invariant for L and an
(iii) There exists a closed subspace V ⊂ F2
XPVLkV = TkX
for 1 ≤ k ≤ d.
Proof. Assume first that there exists a closed subspace V ⊂ F2
d which is co-invariant
for L and an injective operator X : V → H with dense range such that kXk ≤ C
and
XPVLkV = TkX
for k = 1, . . . , d. Let ξ = XPVΩ. Since Ω is obviously cyclic for L and V is co-
invariant, it is readily seen that PVΩ is cyclic for (PVL1V, . . . , PVLdV). For a
word w ∈ F+
d , we compute
Twξ = TwXPVΩ = XPVLwPVΩ
and thus conclude that ξ is cyclic for T , since X has dense range. In addition, for
w ∈ F+
d and h ∈ H we see that
hX ∗h, LwΩi = hX ∗h, PVLwΩi = hX ∗h, PVLwPVΩi
= hh, XPVLwPVΩi = hh, Twξi,
and since {LwΩ}w∈F+
d
is an orthonormal basis for F2
d, we infer that
hX ∗h, LwΩi2 = Xw∈F+
d
h(Twξ ⊗ Twξ)h, hi
hh, Twξi2
Twξ ⊗ Twξ ≤ C2I.
d
kX ∗hk2 = Xw∈F+
= Xw∈F+
Xw∈F+
d
d
whence
(2)
This shows that (iii) implies (i). Assume next that there is a cyclic vector ξ ∈ H
for T satisfying (2). Then, given h ∈ H, we see that
h(Twξ ⊗ Twξ)h, hi ≤ C2khk2.
We conclude that the vector
Xw∈F+
d
hh, Twξi2 = Xw∈F+
Xw∈F+
d
d
hh, TwξiLwΩ
14
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
belongs to F2
operator Y : H → F2
d by
d and has norm at most Ckhk. Hence, we may define a bounded linear
hh, TwξiLwΩ,
h ∈ H.
d
Y h = Xw∈F+
k h, LuΩi =*Xw∈F+
d
Plainly kY k ≤ C. Moreover, given 1 ≤ k ≤ d, a word u ∈ F+
we find
k h, TwξiLwΩ, LuΩ+ = hT ∗
hT ∗
hY T ∗
k h, Tuξi
d , and a vector h ∈ H
and
= hh, TkTuξi
kY h, LuΩi = hY h, LkLuΩi =*Xw∈F+
d
hL∗
hh, TwξiLwΩ, LkLuΩ+
= hh, TkTuξi.
We conclude that Y T ∗
Y h = 0 then
k = L∗
kY for every 1 ≤ k ≤ d. Furthermore, we note that if
0 = hY h, LwΩi = hh, Twξi
for every w ∈ F+
Thus, Y is injective. Choosing X = Y ∗ shows that (i) implies (ii).
Finally, the fact that (ii) implies (iii) follows from Lemma 2.11.
d , which implies that h = 0 since ξ is assumed to be cyclic for T .
(cid:3)
It is now a trivial matter to extract the desired necessary and sufficient condition.
Corollary 3.2. Let T = (T1, . . . , Td) be a row contraction on some Hilbert space
H. Then, the following statements are equivalent.
(i) There is a cyclic vector ξ ∈ H for T such that the operatorPw∈F+
Twξ⊗Twξ
d which is co-invariant for L and such
(ii) There exists a closed subspace V ⊂ F2
is bounded.
d
that PVLV ≺ T.
Proof. This follows immediately from Theorem 3.1.
(cid:3)
The property above that Pw∈F+
Twξ ⊗ Twξ be bounded can be seen as a re-
laxation of the condition that ξ be a wandering vector, in the sense that the set
is orthogonal. It will turn out below that this weak orthogonality con-
{Twξ}w∈Fd
dition is automatically satisfied for pure row contractions that admit any cyclic
vector (see Corollary 3.6).
+
d
Given a d-tuple T = (T1, . . . , Td) on some Hilbert space H, we defined its mul-
tiplicity to be the least cardinality µT of a cyclic set for T . We emphasize that in
the special case of d-shifts, this definition of multiplicity agrees with the one we
gave in Section 2. The following simple observation will be use implicitly. Let T ′
be a d-tuple of operators on some other Hilbert space H′ for which there exists an
operator X : H′ → H with dense range such that XT ′
k = TkX for k = 1, . . . , d. If
Γ ⊂ H′ is cyclic for T ′, then a straightforward verification reveals that XΓ ⊂ H is
cyclic for T . In particular, this shows that µT ≤ µT ′ .
In preparation for our main result, we need a technical tool.
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
15
Lemma 3.3. Let T = (T1, . . . , Td) be a pure row contraction on some Hilbert space
H. Then, there is a d-shift V = (V1, . . . , Vd) acting on some Hilbert space K with
multiplicity µT and a contractive operator X : K → H with dense range such that
TkX = XVk for 1 ≤ k ≤ d.
Proof. By Theorem 2.4, there is a Hilbert space K0 containing H along with a d-shift
S = (S1, . . . , Sd) on K0 with µS = dT and such that
1 ≤ k ≤ d.
T ∗
k = S∗
kH,
By assumption, there is a subset Γ ⊂ H of cardinality µT such that
H = span{Twγ : w ∈ F+
d , γ ∈ Γ}.
Let
d , γ ∈ Γ}
K = span{Swγ : w ∈ F+
w = Xw=n
PKSwPKS∗
and put Vk = SkK for k = 1, . . . , d. For each n ∈ N, we note that
wPK ≤ PK Xw=n
which shows at once that V is pure, and that Pd
k ≤ I. By the non-
commutative Wold -- von Neumann decomposition [32, Theorem 1.3], we infer that
V is a d-shift.
w PK
Xw=n
k=1 VkV ∗
VwV ∗
SwS∗
Next, define X = PHK : K → H. For k = 1, . . . , d, using that H and K are
respectively co-invariant and invariant for S we compute
TkX = PHSkPHK = PHSkK
= PHPKSkK = XSkK = XVk.
Hence, for w ∈ F+
d and γ ∈ Γ we find
XSwγ = XVwγ = TwXγ = Twγ
which implies X K = H, so that X has dense range. In particular, we must have
that µT ≤ µV . On the other hand, by definition of K we see that Γ is a cyclic set
for V , so that µV ≤ card(Γ) = µT .
(cid:3)
We now arrive at the last technical step before the main result of the section.
Roughly speaking, it says that up to a quasi-affine transform, the multiplicity of a
pure row contraction can be made to agree with its defect.
Lemma 3.4. Let T = (T1, . . . , Td) be a pure row contraction on some Hilbert space
H. Then, there is a pure row contraction T ′ = (T ′
d) on some other Hilbert
space H′ such that µT ′ = dT ′ = µT , along with a contractive injective operator
X : H′ → H with dense range such that TkX = XT ′
If T is
commuting, then so is T ′.
k for 1 ≤ k ≤ d.
1, . . . , T ′
Proof. By Lemma 3.3, there a d-shift V = (V1, . . . , Vd) with µV = µT acting on
some Hilbert space K and a contractive operator Y : K → H with dense range such
that TkY = Y Vk for every 1 ≤ k ≤ d. Put
and for each 1 ≤ k ≤ d let
H′ = K ⊖ ker Y
and X = Y H′
T ′
k = PH′ VkH′ .
16
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
Clearly, X is contractive, injective, and it has dense range. Moreover, it follows
from Lemma 2.11 that TkX = XT ′
(3)
Since H′ is co-invariant subspace for V , we see that T ′∗
It follows that T ′ is a pure row contraction. Using that
k for 1 ≤ k ≤ d. In particular
µT ≤ µT ′.
w = V ∗
wH′ for every w ∈ F+
d .
I −
dXk=1
′∗
k = PH′ I −
T ′
kT
k!H′
VkV ∗
dXk=1
we find dT ′ ≤ dV . But V is a d-shift, so dV = µV by Theorem 2.3 and we infer
(4)
On the other hand, since T ′ is pure we may apply Theorem 2.4 to find a Hilbert
space K′ containing H′ and a d-shift S = (S1, . . . , Sd) with dT ′ = µS and such that
′∗
kH′ for every 1 ≤ k ≤ d. Choose a subset Γ ⊂ K′ with cardinality dT ′ that
k = S∗
T
is cyclic for S. For w ∈ F+
d and γ ∈ Γ, we may use that H′ is co-invariant for S to
find
dT ′ ≤ µT .
T ′
wPH′γ = PH′SwPH′γ = PH′Swγ.
Consequently, we have that
span{T ′
wPH′ γ : w ∈ F+
d , γ ∈ Γ} = PH′(cid:16)span{Swγ : w ∈ F+
= PH′ K′ = H′
d , γ ∈ Γ}(cid:17)
so that PH′ Γ is cyclic for T ′. Hence,
(5)
µT ′ ≤ dT ′ .
Combining inequalities (3), (4) and (5) yields µT ′ = dT ′ = µT as desired.
Finally, if T is commuting, then Annnc(T ) contains the commutator ideal, so
(cid:3)
that T ′ is also commuting by Lemma 2.12.
We remark here that the preceding two proofs are faithful adaptations of the
single-variable ones found in [8, Lemma I.3.5 and Theorem I.3.7].
Interestingly
however, this approach forces us through the non-commuting multivariate world,
even if we start with a commuting row contraction.
To see why, we note that there are examples of cyclic invariant subspaces for Mx
for which the restriction has infinite defect. Take for instance the cyclic invariant
subspace M ⊂ H 2
d generated by x1, so that
M = Mα∈Nd,α1≥1
Cxα.
Then M clearly has infinite co-dimension, so we may invoke [6, Theorem F] to
conclude that the restriction MxM has infinite defect. This underlines potential
difficulties in establishing the equality dV = µV , which is used crucially in the
proof of Lemma 3.4. Fortunately, with our approach this equality follows from
Theorem 2.3 since V is known to be a d-shift, being the restriction of a d-shift to
an invariant subspace. For commuting d-shifts, the behaviour of such restrictions is
more subtle. Assume for instance that the restriction of the standard commuting d-
shift to some cyclic proper invariant subspace is another commuting d-shift. Since it
must be cyclic, this second commuting d-shift must be of multiplicity one, and thus
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
17
unitarily equivalent to the standard commuting d-shift. This is however impossible,
by [26, Corollary 5.5]. Although it is plausible that a completely different approach
could circumvent these issues in the commuting setting, the preceding remarks show
that our current proofs have an inexorable non-commutative aspect built into them.
The following is one of our main results.
Theorem 3.5. Let T = (T1, . . . , Td) be a pure cyclic row contraction on some
Hilbert space H. Then there is a contractive operator X : F2
d → H with dense range
such that XLk = TkX for k = 1, . . . , d.
d → H with dense
If T is commuting, then there is a contractive operator X ′ : H 2
range such that X ′Mxk = TkX ′ for k = 1, . . . , d.
Proof. By Lemma 3.4, we see that there is a pure row contraction T ′ = (T ′
1, . . . , T ′
d)
on some other Hilbert space H′ such that dT ′ = 1, along with a contractive injective
operator Y : H′ → H with dense range such that TkY = Y T ′
k for 1 ≤ k ≤ d. Up to
unitary equivalence, we may assume by Theorem 2.4 that H′ ⊂ F2
kH′
for every k = 1, . . . , d. Define X : F2
d → H as X = Y PH′ . For 1 ≤ k ≤ d, we
calculate using that H′ is co-invariant for L that
′∗
k = L∗
d and T
XLk = Y PH′ Lk = Y PH′ LkPH′
kPH′ = TkY PH′
= Y T ′
= TkX.
Moreover, we note that
X F2
d = Y PH′ F2
d = Y H′ = H
so that X has dense range, which establishes the first statement. For the second,
we assume in addition that T is commuting. Then,
0 = (TjTk − TkTj)X = X(LjLk − LkLj)
and thus ker X contains (LjLk − LkLj)F2
ker X contains (Fs
d → Fs
2.2, there is a unitary operator U : H 2
U Mxk U ∗ = PFs
LkFs
d for every 1 ≤ j, k ≤ d. We infer that
d)⊥ by [20, Proposition 2.4], whence X = XPFs
. By Theorem
d with the property that
1 ≤ k ≤ d.
,
d
d
d
We set X ′ = XU , which still has dense range. Using that Fs
we obtain
d is co-invariant for L
X ′Mxk = XU Mxk = XPFs
d
LkU
= XLkU = TkX ′
for k = 1, . . . , d.
(cid:3)
Comparing the previous result with Theorem 3.1, we obtain the following con-
sequences.
Corollary 3.6. Let T = (T1, . . . , Td) be a pure cyclic row contraction on some
Hilbert space H. Then, there is a vector ξ ∈ H that is cyclic for T with the additional
Twξ ⊗ Twξ ≤ I. Moreover, the following statements hold.
property thatPw∈F+
(1) If kTwξk = 1 for every w ∈ F+
(2) If d > 1 and T is commuting, then kTwξk < 1 for some w ∈ F+
d .
d , then T must be unitarily equivalent to the
standard d-shift.
d
18
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
Proof. The existence of a vector ξ with the announced property follows by com-
bining Theorems 3.1 and 3.5. For (1), assume that kTwξk = 1 for each w ∈ F+
d .
The condition that Pw∈F+
d } to be an
orthonormal set. But since ξ is cyclic, the set {Twξ : w ∈ F+
d } must in fact be an
orthonormal basis for H. It is now straightforward to construct a unitary equiva-
lence between T and the standard d-shift L. Finally, (2) is trivial consequence of
(1) since the standard d-shift is non-commuting.
(cid:3)
Twξ ⊗ Twξ ≤ I then forces {Twξ : w ∈ F+
d
We close this section by refining Theorem 3.5 to obtain another one of our main
results. See Subsection 2.3 for the definition of the d-tuple MJ .
Corollary 3.7. Let T = (T1, . . . , Td) be a pure cyclic commuting row contraction
on some Hilbert space H. Let J = Annc(T ). Then MJ is a quasi-affine transform
of T .
Proof. By Theorem 3.5, there is a contractive operator X : H 2
d → H with dense
range such that XMxk = TkX for k = 1, . . . , d. In turn, invoking Lemma 2.11, we
see that there is a closed subspace K ⊂ H 2
d which is co-invariant for Mx and such
that
It follows from Lemma 2.12 that
(PKMx1K, . . . , PKMxdK) ≺ T.
whence
Annnc(T ) = Annnc(PKMx1K, . . . , PKMxdK)
Annc(T ) = Annc(PKMx1K, . . . , PKMxdK)
by virtue of Equation (1). By Lemma 2.9, we conclude that K = HJ so that
MJ ≺ T .
(cid:3)
Given a pure cyclic row contraction T = (T1, . . . , Td) with b = Annnc(T ), we
do not claim that Lb ≺ T . The difficulty in achieving this is connected with the
discussion following Lemma 2.8. This is also related to determining whether a given
co-invariant subspace coincides with the so-called "maximal b-constrained piece"
of L as defined in [35]. We will not address these issues further in this paper.
To summarize, the basic outcome of this section is a classification of some pure
row contractions by means of compressions of the standard d-shift. By allowing
for quasi-affine transforms we capture the behaviour of a class of objects that is
more flexible than those with defect equal to one (which is required in Theorem
2.4). To illustrate this flexibility, we note for instance that if T is a cyclic pure row
contraction with defect one, then the scaled row contraction rT has defect equal
to dim H for 0 < r < 1, even though it is still pure and cyclic. Furthermore, the
cyclicity condition is preserved under similarity.
4. Rigidity of invariant subspaces
In this section, we examine the rigidity of invariant or co-invariant subspaces for
pure row contractions. More precisely, assume that T = (T1, . . . , Td) is a pure row
contraction on some Hilbert space H, and let M ⊂ H be a closed subspace which is
invariant for T . Throughout, we let TM denote the restricted d-tuple
(T1M, . . . , TdM)
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
19
and PM⊥ TM⊥ denote the compressed d-tuple
(PM⊥ T1M⊥, . . . , PM⊥ TdM⊥ ).
We aim to determine the extent to which the annihilators
Annnc(TM)
or Annnc(PM⊥ TM⊥)
determine M. We begin by revisiting a pathological example that we encountered
before.
Example 1. By [27, Theorem 3.7], there is a proper closed subspace K ⊂ F2
is co-invariant for Ld and which satisfies
{θ ∈ Ld : θF2
d ⊂ K⊥} = {0}.
d which
In particular, we see that
yet K 6= F2
d.
Annnc(PKLK) = {0} = Annnc(L)
(cid:3)
In view of this difficulty, we will focus our attention for the remainder of the
paper on the commuting setting. We start by considering a relatively simple case.
A commuting d-tuple (N1, . . . , Nd) is said to be nilpotent if for each 1 ≤ k ≤ d
there is mk ∈ N such that N mk
Theorem 4.1. Let N = (N1, . . . , Nd) be a cyclic nilpotent commuting row con-
traction on some Hilbert space H. Let M ⊂ H be an invariant subspace for N such
that Annc(NM) = Annc(N ). Then, M = H.
Proof. Let ξ ∈ H be a cyclic vector for N . We find
k = 0.
Since N is nilpotent, this implies that H is finite-dimensional and
H = span{N αξ : α ∈ Nd}.
H = {p(N )ξ : p ∈ C[x1, . . . , xd]}.
The subset J ⊂ C[z1, . . . , zd] defined as
is an ideal with the property that
J = {p ∈ C[z1, . . . , zd] : p(N )ξ ∈ M}
For convenience, we set
and
M = {p(N )ξ : p ∈ J }.
A = Annc(NM) ∩ C[x1, . . . , xd]
Observe that a polynomial p satisfies p(N ) = 0 if and only if p(N )ξ = 0. Thus,
B = Annc(N ) ∩ C[x1, . . . , xd].
A = {f ∈ C[x1, . . . , xd] : fJ ⊂ B}.
Clearly, the set {α ∈ Nd : xα /∈ B} is finite and contains (0, . . . , 0). Choose an
element β of that set with maximal length. Then, we have xkxβ ∈ B for every
1 ≤ k ≤ d.
Assume now that M is a proper subspace of H, whence ξ /∈ M. Let q be a
polynomial with non-zero constant term. Then, there is another polynomial r
such that r(N )q(N ) = I, so in particular ξ = r(N )q(N )ξ. We conclude that
20
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
q /∈ J , which shows J consists of polynomials with zero constant term. Hence,
xβJ ⊂ B, and therefore xβ ∈ A. This shows that A 6= B and in particular
Annc(NM) 6= Annc(N ).
(cid:3)
In the case of a single cyclic constrained contraction, the conclusion of the pre-
vious theorem always holds [8, Theorem III.2.13]. Unfortunately, the general mul-
tivariate situation is more complicated. In fact, the aforementioned theorem does
not even extend to the commuting bivariate case, as the following example shows.
Example 2. Let J be the weak-∗ closed ideal of M2 generated by x2. Then, the
space HJ = H 2
2 that depends only on
the variable x1 and in particular we have
2 ] consists of those functions in H 2
2 ⊖ [J H 2
The pure commuting pair
PHJ Mx2HJ = 0.
MJ = (PHJ Mx1HJ , 0)
is constrained, and by Lemma 2.10 we have Annc(MJ ) = J . Note also that 1 ∈ HJ
is a cyclic vector for MJ . Let U : H 2
1 → HJ be the linear operator defined as
(U f )(z1, z2) = f (z1),
(z1, z2) ∈ B2
for f ∈ H 2
1 . It is readily verified that U is unitary. Observe now that
U ∗(PHJ Mx1HJ )U
is the usual isometric 1-shift. Let θ ∈ M1 be a non-constant inner function and let
M = U (θH 2
1 ) ⊂ HJ . Then, M is a proper closed subspace of HJ which is invariant
for MJ . If we let R1 = PHJ Mx1M, then we see that (R1, 0) is the restriction of
MJ to M. Further, if we let U ′ = U MθU ∗ then we see that U ′ : HJ → M is
unitary with
Thus R is unitarily equivalent to MJ , and in particular it is a pure cyclic commuting
row contraction with
U ′(PHJ Mx1HJ )U ′∗ = R1.
Annnc(R) = Annnc(MJ ) and Annc(R) = Annc(MJ ),
even though M 6= HJ .
(cid:3)
We note that the zero set of the commutative annihilator has dimension one in
the preceding example. At present, we know of no multivariate counterexample to
[8, Theorem III.2.13] where this zero set has dimension zero. Theorem 4.1 may be
evidence towards the relevance of having a zero-dimensional zero set, but at the
time of this writing we do not know for sure.
In light of Example 2, we then seek natural sufficient conditions for multivariate
rigidity of invariant subspaces. The next development will gain insight by first
considering co-invariant subspaces; it is based on the following simple fact.
Lemma 4.2. Let M, N ⊂ H 2
Then, we have M = N if and only if Annc(PMMxM) = Annc(PNMxN).
Proof. This follows immediately from Lemma 2.9.
d be closed subspaces that are co-invariant for Md.
(cid:3)
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
21
Before proceeding, we introduce some notation. Let S = (S1, . . . , Sd) be a d-tuple
of operators on some Hilbert space H and let S′ = (S′
d) be another d-tuple
of operators on some other Hilbert space H′. Denote by Q(S′, S) the collection
of injective bounded linear operators X : H′ → H with dense range such that
XS′
k = SkX for every k = 1, . . . , d. Recall that if T is a cyclic pure commuting row
contraction and J = Annc(T ), then Q(MJ , T ) is non-empty by Corollary 3.7.
ing row contractions.
We now prove a rigidity result for co-invariant subspaces of cyclic pure commut-
1, . . . , S′
Theorem 4.3. Let T = (T1, . . . , Td) be a cyclic pure commuting row contraction
on some Hilbert space H. Put J = Annc(T ). Let M, N ⊂ H be closed co-invariant
subspaces for T such that
Annc(PMTM) = Annc(PNTN).
Then, X ∗M = X ∗N for every X ∈ Q(MJ , T ). In particular, M = N if X ∗M and
X ∗N are closed for some X ∈ Q(MJ , T ).
Proof. Fix X ∈ Q(MJ , T ). Define M′ = X ∗M and N′ = X ∗N, so that M′ and N′
are co-invariant for MJ , and hence for Mx. Let Z = PMXM′ : M′ → M. We have
Z ∗ = X ∗M, so Z ∗ is injective with dense range and
xM′ ).
Z ∗ ∈ Q(T ∗M, M ∗
Thus,
Likewise, the operator Y = PNXN′ is injective with dense range, and
Z ∈ Q(PM′ MxM′ , PMTM).
Y ∈ Q(PN′ MxN′ , PNTN).
We conclude from Lemma 2.12 and from the assumption on M and N that
Annc(PM′ MxM′ ) = Annc(PN′ MxN′).
By virtue of Lemma 4.2, we obtain M′ = N′, that is X ∗M = X ∗N. If X ∗M and
X ∗N are closed, then X ∗M = X ∗N and we conclude that M = N since X ∗ is
injective.
(cid:3)
The following consequence is noteworthy as it applies in particular to the nilpo-
tent setting, thus complementing Theorem 4.1.
Corollary 4.4. Let T = (T1, . . . , Td) be a cyclic pure commuting row contraction
on some finite-dimensional Hilbert space H. Let M, N ⊂ H be co-invariant subspaces
for T such that
Then, M = N.
Annc(PMTM) = Annc(PNTN).
Proof. Finite-dimensional subspaces are closed, so this follows at once from Theo-
rem 4.3.
(cid:3)
We now return to invariant subspaces and identify a sufficient condition for
rigidity.
Theorem 4.5. Let T = (T1, . . . , Td) be a pure commuting row contraction on some
Hilbert space H. Let J ⊂ Md be a weak-∗ closed ideal and let X ∈ Q(M ∗
J , T ). Let
M, N ⊂ H be closed invariant subspaces for T such that Annc(TM) = Annc(TN).
If M and N are contained in the range of X, then M = N.
22
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
Proof. If f ∈ X −1M, then for every 1 ≤ k ≤ d we obtain
X(M ∗
xkHJ )f = TkXf ∈ TkM ⊂ M.
Thus X −1M is invariant for M ∗
that X −1N is also invariant for M ∗
x . Next, let
J and hence for M ∗
x . An identical argument shows
Z = XX −1M : X −1M → M
Y = XX −1N : X −1N → N.
Because X is injective, it follows that Z and Y are injective as well. Moreover, Z
and Y are surjective since M, N are contained in the range of X. For k = 1, . . . , d,
we note that
xkX −1M) and (TkN)Y = Y (M ∗
xkX −1N).
d be the conjugate linear, anti-unitary operator such that Jxα = xα
(TkM)Z = Z(M ∗
d → H 2
Let J : H 2
for every α ∈ Nd. Then, we note that
(TkM)Z = Z((JMxk J)∗X −1M)
and (TkN)Y = Y ((JMxk J)∗X −1N)
for every k = 1, . . . , d. A standard approximation procedure yields that JMdJ =
Md and
ϕ(TM)Z = Z((JMϕJ)∗X −1M)
and ϕ(TN)Y = Y ((JMϕJ)∗X −1N)
for every ϕ ∈ Md. Observe then that ϕ ∈ Annc(TM) if and only if
PX −1M(JMϕJ)X −1M = 0
which in turn is equivalent to
JMϕJ ∈ Annc(PX −1MMxX −1M).
Thus, we find
Likewise, we infer that
Annc(PX −1MMxX −1M) = J Annc(TM)J.
By assumption, we may therefore write
Annc(PX −1NMxX −1N) = J Annc(TN)J.
Annc(PX −1MMxX −1M) = Annc(PX −1NMxX −1N).
In view of Lemma 4.2, we then know that X −1M = X −1N. Applying X on both
sides yields M = N.
(cid:3)
In general, it is not clear how to determine that Q(M ∗
J , T ) is non-empty in order
to apply the previous result. In some cases however, the existence of a cyclic vector
can be exploited.
Corollary 4.6. Let T = (T1, . . . , Td) be a nilpotent commuting row contraction
on some Hilbert space H. Assume that T ∗ is cyclic. Let M, N ⊂ H be invariant
subspaces for T such that Annc(TM) = Annc(TN). Then, M = N.
Proof. The commuting d-tuple T ∗ is also nilpotent. Applying [36, Theorem 3.8]
with k = m = 1, Q = {0} and f (Z) = Z1 + . . . + Zd (in the notation of that paper),
we see that there is an invertible operator Y ∈ B(H) such that the d-tuple
Y T ∗Y −1 = (Y T ∗
1 Y −1, . . . , Y T ∗
d Y −1)
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
23
is a nilpotent cyclic commuting row contraction. In particular, Y T ∗Y −1 is pure,
so if we put J = Annc(Y T ∗Y −1), then by virtue of Corollary 3.7, we see that
there is X ∈ Q(MJ , Y T ∗Y −1). Next, observe that H is finite-dimensional since the
nilpotent d-tuple T ∗ is cyclic, whence X is invertible and (X −1Y )∗ ∈ Q(M ∗
J , T ).
Theorem 4.5 then yields M = N.
(cid:3)
It is not clear that Theorem 4.1 could be derived from the previous result, as it
is not typically true that the adjoint of a cyclic nilpotent row contraction is still
cyclic; see Example 5 below.
5. Invariant decompositions and separating vectors
We saw in Corollary 3.7 that pure commuting row contractions can be effectively
classified using compressions of the standard commuting d-shift, provided that they
admit a cyclic vector. Inspired by the successful univariate theory developed in [8],
a natural subsequent step in this classification program would involve moving past
the setting of cyclicity using the following device.
Let T = (T1, . . . , Td) be a d-tuple of operators on some Hilbert space H and let
M, N ⊂ H be non-trivial invariant subspaces for T . We say that the pair (M, N) is
an invariant decomposition for T if M ∩ N = {0} and M + N = H. In the special
case where TM is cyclic, we say that (M, N) is a cyclic decomposition for T .
In the single-variable case, non-cyclic constrained contractions always admit
cyclic decompositions [8, Theorem III.3.1]. We start this section with an exam-
ple illustrating that cyclic decompositions are more elusive in several variables. For
this purpose, we introduce one more notion. An invariant decomposition (M, N)
will be said to be topological if the stronger condition M + N = H is satisfied. The
following is an adaptation of [25, Theorem 2.3]
Example 3. Let {en : n ∈ N} denote the canonical orthonormal basis of ℓ2(N).
Let H = ℓ2(N) ⊕ ℓ2(N). For each n ∈ N, define vectors ξn, ηn ∈ H as
ξn = en ⊕ 0,
ηn = 0 ⊕ en.
Then, {ξn, ηn : n ∈ N} is an orthonormal basis for H. We may define bounded
linear operators T1, T2 on H as
T1 =
ηn+1 ⊗ ξn,
T2 =
ηn ⊗ ξn
where the series converge in the strong operator topology of B(H). We find
1
√2
1
√2
∞Xn=1
∞Xn=1
∞Xn=1
∞Xn=1
2T1T ∗
1 =
ηn+1 ⊗ ηn+1
2T2T ∗
2 =
ηn ⊗ ηn.
and
24
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
1 and 2T2T ∗
Thus, 2T1T ∗
In particular, we see that
T1 and T2 are scalar multiples of partial isometries, and thus have closed range.
Moreover,
2 are orthogonal projections.
T1T ∗
1 + T2T ∗
2 ≤ I
so that T = (T1, T2) is a row contraction on H. Since ξn is orthogonal to ηm for
every n, m ∈ N, we see that
T 2
1 = T1T2 = T2T1 = T 2
2 = 0.
Thus, T is a nilpotent commuting row contraction. Using Theorem 2.1 it is readily
seen that Annc(T ) is the weak-∗ closed ideal of M2 generated by {x2
1, x1x2, x2
2}.
We infer that the range of T β is closed for all β ∈ N2.
Assume that M, N ⊂ H are closed invariant subspaces for T with M ∩ N = {0}
and H = M + N. Observe that
and
Furthermore, we have that
T1H = ⊕∞
n=2
Cηn
T2H = ⊕∞
n=1
Cηn.
T1H = T1M + T1N,
T2H = T2M + T2N
and
T1M ⊂ M,
T2M ⊂ M,
T1N ⊂ N,
T2N ⊂ N.
Using that M ∩ N = {0} and that T1H ⊂ T2H, we infer that T1M ⊂ T2M and
T1N ⊂ T2N. Since T1H is a co-dimension one closed subspace of T2H, we infer that
either T1M = T2M or T1N = T2N. By symmetry, we may suppose the former.
Given a subspace K ⊂ H, we set
and
νξ(K) = inf{n ∈ N : ξn /∈ K⊥}
νη(K) = inf{n ∈ N : ηn /∈ K⊥}.
Note that νξ(K) is infinite whenever ξn ∈ K⊥ for every n ∈ N. Using that T ∗
ξn and T ∗
2 ηn = ξn for every n ∈ N, we see that
1 ηn+1 =
νη(T2K) = νξ(K)
and νη(T1K) = νξ(K) + 1.
Recall that T1M = T2M, thus
νξ(M) = νη(T2M) = νη(T1M) = νξ(M) + 1
which forces νξ(M) = +∞. Therefore, ξn ∈ M⊥ for every n ∈ N. Consequently,
we find
M ⊂ ⊕∞
n=1
Cηn = T2H.
We infer that M ⊂ T2M which further implies that T2M = M. Finally, using
that T 2
2 = 0 we obtain M = {0}. This shows that T does not admit a topological
invariant decomposition.
If (M, N) is a cyclic decomposition for T , then using that T is nilpotent and that
TM is cyclic we conclude that M must be finite-dimensional. It is well-known then
that H = M + N, so that (M, N) would be a topological invariant decomposition
for T . We thus conclude that T admits no cyclic decomposition either.
(cid:3)
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
25
We mention here a classical result of Apostol and Stampfli [4, Theorem 5] which
implies that if T is a nilpotent operator on a Hilbert space H with the property that
T nH is closed for every n ∈ N, then T admits a topological cyclic decomposition.
The previous example shows that this theorem fails in several variables.
In spite of Example 3, by leveraging the work done previously on rigidity of
invariant subspaces we identify a sufficient condition that allows for an invariant
decomposition.
Theorem 5.1. Let T = (T1, . . . , Td) be a nilpotent commuting row contraction on
some Hilbert space H. Let M ⊂ H be an invariant subspace for T such that (TM)∗
has a cyclic vector and Annc(TM) = Annc(T ). Then, there is a subspace N ⊂ H
that is invariant for T such that M ∩ N = {0} and H = M + N.
Proof. By assumption, there is a vector ξ ∈ M which is cyclic for (TM)∗. Let
Put S = PKTK and observe that Annc(T ) ⊂ Annc(S). Define X : M → K as
X = PKM. Using that K is coinvariant for T and M is invariant for T , we obtain
K = span{T ∗αξ : α ∈ Nd}.
SkX = PKTkPKM = PKTkM
= PKPMTkM = XTkM
for every 1 ≤ k ≤ d, so that for α ∈ Nd we have
X ∗T ∗αK = X ∗S∗α = (TM)∗αX ∗.
X ∗ξ = PMξ = ξ
Furthemore,
whence
X ∗T ∗αξ = (TM)∗αX ∗ξ = (TM)∗αξ
for every α ∈ Nd. By choice of ξ being cyclic for (TM)∗, we conclude that X ∗ has
dense range and thus X is injective. Now, XM ⊂ K is easily seen to be invariant
for S and
so it follows from Lemma 2.12 that
(SkXM)X = XTkM,
1 ≤ k ≤ d
By assumption, we then infer that
Annc(SXM) ⊂ Annc(TM).
and thus
We conclude that
Annc(SXM) ⊂ Annc(T )
Annc(SXM) ⊂ Annc(S).
Notice now that S∗ is cyclic by the construction of K. We may thus invoke Corollary
4.6 to find XM = K, so that X ∗ is injective. Finally, define N = K⊥. We find
Annc(SXM) = Annc(S).
and
M ∩ N = M ∩ K⊥ = ker X = {0}
M + N = (M⊥ ∩ N⊥)⊥ = (M⊥ ∩ K)⊥ = (ker X ∗)⊥ = H.
(cid:3)
26
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
One shortcoming of the previous theorem is that the existence of the decompo-
sition hinges on the adjoint of TM being cyclic. This condition is a bit mysterious.
Interestingly, in the case of a single constrained contraction, this condition is in
fact equivalent to TM being cyclic by [8, Theorem III.2.3]. Unfortunately, this
convenient equivalence does not hold in the multivariate world, as we will see in
Example 5.
Given these multivariate obstacles, and in light of the univariate mechanism
[8, Theorem III.3.1] for producing cyclic decompositions, it seems worthwhile to
investigate more precisely where this mechanism fails in several variables. The rest
of the section will be devoted to this goal, through the lens of the following notion.
Let T = (T1, . . . , Td) be a constrained commuting row contraction on some
Hilbert space H. A subset Σ ⊂ H is said to be separating for T if, given any
θ ∈ Md, the condition Σ ⊂ ker θ(T ) forces θ(T ) = 0. This means that Σ is a
separating set for the operator algebra
{θ(T ) : θ ∈ Md}.
When Σ consists of only one element ξ ∈ H, we say that ξ is a separating vector for
H. In the classical one-variable setting, separating vectors are called maximal [8],
but we feel that in our context our choice of terminology is a bit more descriptive.
We start with a concrete example where a separating vector can be constructed
explicitly.
Example 4. Let T = (T1, . . . , Td) be a pure commuting row contraction on some
Hilbert space H. Assume that Annc(T ) is the weak-∗ closed ideal of Md generated
by {xn1
d } for some fixed positive integers n1, . . . , nd. Set
1 , . . . , xnd
Ω = {α ∈ Nd : xα /∈ Annc(T )}
and put
γ = (n1 − 1, . . . , nd − 1).
Since γ ∈ Ω, we may choose a non-zero vector ξ ∈ H\ ker(T γ). We claim that ξ
is separating for T . To see this, let ϕ ∈ Md with ϕ(T ) 6= 0. An application of
Theorem 2.1 with
n > (n1 − 1) + (n2 − 1) + . . . + (nd − 1)
shows that there is a polynomial p such that ϕ−p ∈ Annc(T ), whence ϕ(T ) = p(T ).
Furthermore, it is clear that p can be chosen to be of the form
p =Xα∈Ω
cαxα.
Because p(T ) 6= 0, the set S = {α ∈ Ω : cα 6= 0} is non-empty. Choose µ ∈ S with
minimal length. It is easily seen that there is β ∈ Nd such that µ + β = γ and
α + β /∈ Ω for every α ∈ S \ {µ}. Thus
T βp(T ) = cµT γ.
Since ξ /∈ ker T γ and cµ 6= 0, it follows that T βp(T )ξ 6= 0 and in particular p(T )ξ 6=
0. Therefore, ϕ(T )ξ 6= 0, and we conclude that ξ is separating for T .
(cid:3)
Unfortunately, separating vectors do not always exist, even for commuting nilpo-
tent pairs on finite-dimensional spaces.
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
27
Example 5. With respect to the usual orthonormal basis, we define linear opera-
tors on C3 as
T1 =
T2 =
1
√3
0
1
0
0
0
0 −1
0
0 ,
1
√3
0
0
0
0 0
0 1
0 0 .
Easy computations show that
T 2
1 = T1T2 = T2T1 = T 2
2 = 0
and that
T1T ∗
1 + T2T ∗
2 =
0
0
0
0 0
1 0
0 0 .
Hence, T = (T1, T2) is a nilpotent commuting row contraction. Let ϕ ∈ M2. An ap-
plication of Theorem 2.1 yields constants a, b, c ∈ C and multipliers ϕ12, ϕ11, ϕ22 ∈
M2 such that
ϕ = a + bx1 + cx2 + ϕ11x2
1 + ϕ12x1x2 + ϕ22x2
2.
Then, we find
ϕ(T1, T2) = aI + bT1 + cT2 =
a
0
b/√3 a (c − b)/√3
0
a
0
0
.
Thus, ϕ(T1, T2) = 0 if and only if a = b = c = 0. This shows that Annc(T ) is the
weak-∗ closed ideal of M2 generated by {x2
C3, and consider the polynomial p defined as
We claim that no vector v ∈ C3 is separating for T . Indeed, write v = (v1, v2, v3) ∈
1, x1x2, x2
2}.
p =(v3x1 + (v3 − v1)x2
x1
if v3 6= 0 or v1 6= 0,
otherwise.
Then p(T )v = 0 yet p(T ) 6= 0, which means that v is not separating. This es-
tablishes the claim that T has no separating vector, and thus no cyclic vector
either. Note however that the vector (0, 1, 0) is easily verified to be cyclic for the
pair T ∗ = (T ∗
2 ). Anticipating Theorem 5.3 below, we also remark that the set
{(1, 0, 0), (0, 0, 1)} is separating for T .
(cid:3)
1 , T ∗
Interestingly, separating vectors for single constrained contractions are known
to exist in abundance [8, Theorem II.3.6]. Inspection of the proofs of the results
leading up to [8, Theorem III.3.1] reveals that separating vectors play a crucial role
there. Thus, the difficulties associated to the construction of cyclic decompositions
in the multivariate context may be explained, in part, by the scarcity of separating
vectors brought to light in Example 5.
Another consequence of Example 5 is that nilpotent commuting d-tuples are not
as nicely behaved as their single variable counterparts. For instance, by [3, Theorem
1] it is known that if T is a nilpotent contraction, then there is a collection {Tλ :
λ ∈ Λ} of cyclic nilpotent contractions such that
⊕λ∈ΛTλ ≺ T.
Trivially, a cyclic vector is necessarily separating, so the following lemma implies
that such an operator T must admit a separating vector. In light of the previous
28
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
example, we conclude that a direct analogue of [3, Theorem 1] cannot hold in several
variables.
Lemma 5.2. The following statements hold.
1, . . . , T ′
(1) Let T = (T1, . . . , Td) and T ′ = (T ′
d) be constrained commuting row
contractions on some Hilbert spaces H and H′. Assume that T ′ ≺ T . If T ′
has a separating vector, then so does T .
(2) For each n ∈ N, let Tn be a constrained commuting row contraction on
some Hilbert space Hn. Assume that Tn has a separating vector for every
n ∈ N. Then, ⊕∞
n=1Tn has a separating vector as well.
Proof. For (1), let X : H′ → H be an injective operator with dense range such that
XT ′
k = TkX for every 1 ≤ k ≤ n. Let ξ ∈ H′ be a separating vector for T ′ and let
θ ∈ Md. Assume that θ(T )Xξ = 0. Then
0 = θ(T )Xξ = Xθ(T ′)ξ
and we infer that θ(T ′)ξ = 0 since X is injective. Using that ξ is separating for T ′,
we thus find that θ(T ′) = 0. Next, Lemma 2.12 and Equation (1) imply that
so that θ(T ) = 0. Hence Xξ is separating for T .
Annc(T ) = Annc(T ′)
Turning to (2), for each n ∈ N we let ξn ∈ Hn be a separating vector for Tn with
norm 1. Define ξ ∈ ⊕∞
n=1Hn as
It is readily verified that if θ ∈ Md, then
ξ =(cid:18)ξ1,
1
2
ξ2, . . . ,
1
2n ξn, . . .(cid:19) .
θ(T ) = ⊕∞
n=1θ(Tn)
and
θ(T )ξ =(cid:18)θ(T1)ξ1,
1
2
θ(T2)ξ2, . . . ,
1
2n θ(Tn)ξn, . . .(cid:19) .
The desired statement readily follows from these observations.
Our final result below offers a counterpoint to Example 5 and shows that, at
least for nilpotent commuting row contractions, the existence of separating vectors
can be insured provided that we allow for finite ampliations. Given a d-tuple of
operators T = (T1, . . . , Td) on some Hilbert space H and a positive integer n ∈ N,
we define
(cid:3)
which acts on the Hilbert space
T (n) = T ⊕ T ⊕ . . . ⊕ T
H(n) = H ⊕ H ⊕ . . . ⊕ H.
Theorem 5.3. Let N = (N1, . . . , Nd) be a nilpotent commuting row contraction
on some Hilbert space H. Set
A = C[x1, . . . , xd]/(C[x1, . . . , xd] ∩ Annc(N )).
Then, for any positive integer s ≥ dimA the ampliation N (s) has a dense set of
separating vectors in H(s).
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
29
Proof. Since N is nilpotent, there are positive integers m1, . . . , md such that N mk
0 for every 1 ≤ k ≤ d. An application of Theorem 2.1 with
k =
n > (m1 − 1) + (m2 − 1) + . . . + (md − 1)
shows that for any ϕ ∈ Md, there is a polynomial p such that ϕ(N ) = p(N ). For
this reason, a vector (ξ1, . . . , ξs) ∈ H(s) is separating for N (s) if whenever p is a
polynomial, the equalities
are equivalent to p(N ) = 0.
rather than general multipliers.
In other words, it suffices to consider polynomials
p(N )ξ1 = ··· = p(N )ξs = 0
Note that A is a finite-dimensional vector space, and put δ = dimA. For each
1 ≤ i ≤ δ, choose a non-empty open subset Ui ⊂ H. Given a polynomial p, we
denote by [p] its image in the quotient A. For every h ∈ H, set
Ah = {[p] ∈ A : p(T )h = 0}.
We claim that there are vectors ξ1 ∈ U1, . . . , ξδ ∈ Uδ such that ∩δ
i=1Aξi = {0}.
Assume otherwise. For convenience, set Aξ0 = A.
Fix a non-zero element [p1] ∈ Aξ0 . Then, p1(T ) 6= 0 so we may choose ξ1 ∈
U1 \ ker p1(T ). Thus, [p1] /∈ Aξ1 and in particular Aξ1 is a proper subspace of
Aξ0 . Suppose that for 1 ≤ i ≤ δ − 1 we have constructed ξ1 ∈ U1, . . . , ξi ∈ Ui
j=0Aξj is a proper subspace of ∩i−1
with the property that ∩i
j=0Aξj . By our standing
assumption, we see that ∩i
j=0Aξj is non-zero, so there is a non-zero element [pi+1] ∈
∩i
j=0Aξj . Correspondingly, there is a vector ξi+1 ∈ Ui+1 \ ker pi+1(T ). We see that
[pi+1] /∈ Aξi+1 , whence ∩i+1
j=0Aξj is a non-zero proper subspace of ∩i
j=0Aξj . By
induction, we obtain vectors ξ1 ∈ U1, . . . , ξδ ∈ Uδ with the property that for every
j=0Aξj is a non-zero proper subspace of ∩i−1
1 ≤ i ≤ δ, we have that ∩i
j=0Aξj . This
forces dimA ≥ δ + 1, contrary to the choice of δ. The claim is established.
Now, let p be a polynomial such that p(T )ξi = 0 for every 1 ≤ i ≤ δ. Then,
[p] ∈ ∩δ
i=0Aξi and therefore [p] = 0. This shows that U1× . . .× Uδ ⊂ H(δ) contains a
separating vector for T (δ). Since U1, . . . , Uδ were arbitrary non-empty open subsets
of H, we conclude that the set Σ ⊂ H(δ) of separating vectors for T (δ) is dense
in H(δ). Finally, for s > δ it is readily verified that every vector in Σ ⊕ H(s−δ) is
separating for T (s), so the result follows.
(cid:3)
We record the following simple reformulation.
Corollary 5.4. Let N = (N1, . . . , Nd) be a nilpotent commuting row contraction
on some Hilbert space H. Then, there is a finite set of cardinality at most δ which
is separating for T , where δ is the dimension of
C[x1, . . . , xd]/(C[x1, . . . , xd] ∩ Annc(N )).
Proof. This is an immediate consequence of Theorem 5.3.
(cid:3)
Comparing with Example 4, it is apparent that the upper bound on the car-
dinality of the separating set in the previous corollary is far from sharp. In fact,
the argument used in Example 4 can be refined and extended to provide a better
estimate. Because we do not have a concrete use for it, we only state the result and
leave the details of the proof to the interested reader. Recall that we denote by [p]
the image of a polynomial p in the quotient A.
30
RAPHA EL CLOU ATRE AND EDWARD J. TIMKO
Let N = (N1, . . . , Nd) be a nilpotent commuting row contraction
on some Hilbert space H. Let
Ωe = {α ∈ Nd : T α 6= 0 and T αTk = 0 for every 1 ≤ k ≤ d}.
If {[xα] : α ∈ Ωe} is a linearly independent subset of A, then T
admits a separating set of cardinality at most card Ωe.
The reader will note that in Example 4, we have card Ωe = 1, while for Example 5
we find card Ωe = 2.
References
[1] Jim Agler and John E. McCarthy, Complete Nevanlinna-Pick kernels, J. Funct. Anal. 175
(2000), no. 1, 111 -- 124. MR1774853
[2]
, Pick interpolation and Hilbert function spaces, Graduate Studies in Mathematics,
vol. 44, American Mathematical Society, Providence, RI, 2002. MR1882259 (2003b:47001)
[3] C. Apostol, R. G. Douglas, and C. Foia¸s, Quasi-similar models for nilpotent operators, Trans.
Amer. Math. Soc. 224 (1976), no. 2, 407 -- 415. MR0425651
[4] Constantin Apostol and Joseph Stampfli, On derivation ranges, Indiana Univ. Math. J. 25
(1976), no. 9, 857 -- 869. MR0412890
[5] William Arveson, Subalgebras of C ∗-algebras. III. Multivariable operator theory, Acta Math.
181 (1998), no. 2, 159 -- 228. MR1668582 (2000e:47013)
[6]
[7]
, The curvature invariant of a Hilbert module over C[z1, · · · , zd], J. Reine Angew.
Math. 522 (2000), 173 -- 236. MR1758582 (2003a:47013)
, The free cover of a row contraction, Doc. Math. 9 (2004), 137 -- 161 (electronic).
MR2054985 (2005e:47013)
[8] Hari Bercovici, Operator theory and arithmetic in H∞, Mathematical Surveys and Mono-
graphs, vol. 26, American Mathematical Society, Providence, RI, 1988. MR954383
(90e:47001)
[9]
, The quasisimilarity orbits of invariant subspaces, J. Funct. Anal. 95 (1991), no. 2,
344 -- 363. MR1092130 (92e:47017)
[10] Hari Bercovici and Thomas Smotzer, Quasisimilarity of invariant subspaces for uniform
Jordan operators of infinite multiplicity, J. Funct. Anal. 140 (1996), no. 1, 87 -- 99. MR1404575
[11] Hari Bercovici and Allen Tannenbaum, The invariant subspaces of a uniform Jordan operator,
J. Math. Anal. Appl. 156 (1991), no. 1, 220 -- 230. MR1102607
[12] T. Bhattacharyya, J. Eschmeier, and J. Sarkar, Characteristic function of a pure commuting
contractive tuple, Integral Equations Operator Theory 53 (2005), no. 1, 23 -- 32. MR2183594
(2006m:47012)
[13] David P. Blecher, Zhong-Jin Ruan, and Allan M. Sinclair, A characterization of operator
algebras, J. Funct. Anal. 89 (1990), no. 1, 188 -- 201. MR1040962
[14] Xiaoman Chen and Kunyu Guo, Analytic Hilbert modules, Chapman & Hall/CRC Research
Notes in Mathematics, vol. 433, Chapman & Hall/CRC, Boca Raton, FL, 2003. MR1988884
[15] Raphael Clouatre, Quasisimilarity of invariant subspaces for C0 operators with multiplicity
two, J. Operator Theory 70 (2013), no. 2, 495 -- 511. MR3138367
[16]
, Quasiaffine orbits of invariant subspaces for uniform Jordan operators, J. Funct.
Anal. 266 (2014), no. 7, 4101 -- 4114. MR3170203
[17] Raphael Clouatre and Kenneth R. Davidson, Absolute continuity for commuting row con-
tractions, J. Funct. Anal. 271 (2016), no. 3, 620 -- 641. MR3506960
[18]
, Duality, convexity and peak interpolation in the Drury-Arveson space, Adv. Math.
295 (2016), 90 -- 149. MR3488033
[19] Raphael Clouatre and Michael Hartz, Multiplier algebras of complete Nevanlinna-Pick spaces:
dilations, boundary representations and hyperrigidity, J. Funct. Anal. 274 (2018), no. 6, 1690 --
1738. MR3758546
[20] Kenneth R. Davidson and David R. Pitts, The algebraic structure of non-commutative ana-
lytic Toeplitz algebras, Math. Ann. 311 (1998), no. 2, 275 -- 303. MR1625750 (2001c:47082)
[21]
, Nevanlinna-Pick interpolation for non-commutative analytic Toeplitz algebras, In-
tegral Equations Operator Theory 31 (1998), no. 3, 321 -- 337. MR1627901 (2000g:47016)
CYCLIC ROW CONTRACTIONS AND RIGIDITY OF INVARIANT SUBSPACES
31
[22]
, Invariant subspaces and hyper-reflexivity for free semigroup algebras, Proc. London
Math. Soc. (3) 78 (1999), no. 2, 401 -- 430. MR1665248 (2000k:47005)
[23] Kenneth R. Davidson, Christopher Ramsey, and Orr Moshe Shalit, Operator algebras for
analytic varieties, Trans. Amer. Math. Soc. 367 (2015), no. 2, 1121 -- 1150. MR3280039
[24] Jim Gleason, Stefan Richter, and Carl Sundberg, On the index of invariant subspaces in
spaces of analytic functions of several complex variables, J. Reine Angew. Math. 587 (2005),
49 -- 76. MR2186975
[25] Phillip Griffith, On the decomposition of modules and generalized left uniserial rings, Math.
Ann. 184 (1969/1970), 300 -- 308. MR0257136
[26] 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. MR2052120
(2005a:47007)
[27] David W. Kribs, Factoring in non-commutative analytic Toeplitz algebras, J. Operator The-
ory 45 (2001), no. 1, 175 -- 193. MR1823067
[28] Wing Suet Li and Vladim´ır Muller, Invariant subspaces of nilpotent operators and LR-
sequences, Integral Equations Operator Theory 34 (1999), no. 2, 197 -- 226. MR1694708
(2001d:47013)
[29] V. Muller and F.-H. Vasilescu, Standard models for some commuting multioperators, Proc.
Amer. Math. Soc. 117 (1993), no. 4, 979 -- 989. MR1112498 (93e:47016)
[30] Stephen Parrott, Unitary dilations for commuting contractions, Pacific J. Math. 34 (1970),
481 -- 490. MR0268710 (42 #3607)
[31] Gelu Popescu, Characteristic functions for infinite sequences of noncommuting operators, J.
[32]
[33]
[34]
[35]
[36]
Operator Theory 22 (1989), no. 1, 51 -- 71. MR1026074 (91m:47012)
, Isometric dilations for infinite sequences of noncommuting operators, Trans. Amer.
Math. Soc. 316 (1989), no. 2, 523 -- 536. MR972704 (90c:47006)
, von Neumann inequality for (B(H)n )1, Math. Scand. 68 (1991), no. 2, 292 -- 304.
MR1129595 (92k:47073)
, Non-commutative disc algebras and their representations, Proc. Amer. Math. Soc.
124 (1996), no. 7, 2137 -- 2148. MR1343719 (96k:47077)
, Operator theory on noncommutative varieties, Indiana Univ. Math. J. 55 (2006),
no. 2, 389 -- 442. MR2225440 (2007m:47008)
, Similarity problems in noncommutative polydomains, J. Funct. Anal. 267 (2014),
no. 11, 4446 -- 4498. MR3269883
[37] B´ela Sz.-Nagy, Ciprian Foias, Hari Bercovici, and L´aszl´o K´erchy, Harmonic analysis of op-
erators on Hilbert space, enlarged, Universitext, Springer, New York, 2010. MR2760647
(2012b:47001)
[38] N. Th. Varopoulos, On an inequality of von Neumann and an application of the metric theory
of tensor products to operators theory, J. Functional Analysis 16 (1974), 83 -- 100. MR0355642
(50 #8116)
Department of Mathematics, University of Manitoba, Winnipeg, Manitoba, Canada
R3T 2N2
E-mail address: [email protected]
E-mail address: [email protected]
|
1209.2574 | 1 | 1209 | 2012-09-12T11:35:32 | On Iyengar-Type Inequalities via Quasi-Convexity and Quasi-Concavity | [
"math.FA"
] | In this paper, we obtain some new estimations of Iyengar-type inequality in which quasi-convex(quasi-concave) functions are involved. These estimations are improvements of some recently obtained estimations. Some error estimations for the trapezoidal formula are given. Applications for special means are also provided. | math.FA | math |
ON IYENGAR-TYPE INEQUALITIES VIA QUASI-CONVEXITY
AND QUASI-CONCAVITY
M. EMIN OZDEMIR⋆
Abstract. In this paper, we obtain some new estimations of Iyengar-type
inequality in which quasi-convex(quasi-concave) functions are involved. These
estimations are improvements of some recently obtained estimations. Some
error estimations for the trapezoidal formula are given. Applications for special
means are also provided.
1. Short Historical Background and Introduction
If it is necessary to bound one quantity by another, the classical inequalities
are very useful for this purpose. This first book called " Inequalities " written
by Hardy, Littlewood and Polya at cambridge University Press in 1934 represents
the first effort to systemize a rapidly expanding domain. In this sense, the second
important book " Classical and New Inequalities in Analysis " is written by D.S.
Mitrinovi´c, J.E.Pe´caric and A.M. Fink. The third book " Analytic Inequalities "
written by D.S. Mitrinovi´c, and the other book " Means and Their Inequalities "
written by Bullen, D.S. Mitrinovi´c, D.S. Vasic, P.M.
Today inequalities play a significant role for the development in all fields of
Mathematics. They have applications in a variety of applied Mathematics. For
example, convex functions are tractable in optimization because local optimality
guarantees global optimality.
In recent years a number of authors have discov-
ered new integral inequalities for convex, s−convex functions, logarithmic con-
vex functions, h−convex functions, quasi-convex functions, m−convex functions,
(α, m)−convex functions, co-ordinated convex functions, and Godunova-Levin func-
tion, P −function.
On November 22, 1881, Hermite (1822-1901) sent a letter to the Journal Math-
esis. This letter was published in Mathesis 3 (1883,p.82) and in this letter an
inequality presented which is well-known in the literature as Hermite-Hadamard
integral inequality :
(1.1)
f(cid:18) a + b
2 (cid:19) ≤
1
b − aZ b
a
f (x) dx ≤
f (a) + f (b)
2
,
where f : I ⊆ R → R is a convex function on the interval I of a real numbers and
a, b ∈ I with a < b. If the function f is concave, the inequality in (1.1) is reversed.
That is
f(cid:18) a + b
2 (cid:19) ≥
1
b − aZ b
a
f (x) dx ≥
f (a) + f (b)
2
.
2000 Mathematics Subject Classification. Primary 26D15.
Key words and phrases. Weighted Holder Inequality, Holder Inequality, Power-mean Inequal-
ity, Differentiable Function, quasi-convex Function.
⋆Corresponding Author.
1
2
M. EMIN OZDEMIR⋆
For recent results, generalizations and new inequalities related to the inequality
(1.1) see ([7]-[17]).
Then left hand side of Hermite-Hadamard inequality (LHH) can also be esti-
mated by the inequality of Iyengar.
f (a) + f (b)
2
(1.2)
where
−
f (x) dx ≤
M (b − a)
4
−
[f (b) − f (a)]2
4M (b − a)
In [3], Daniel Alexandru Ion proved the following inequalities of Iyengar type for
differentiable quasi−convex functions:
(1.3)
f (a) + f (b)
2
(b − a)
4
(sup {f ′ (a) , f ′ (b)})
where f : [a, b] → R is differentiable function on (a, b) , and f ′ is quasi−convex on
[a, b] with a < b.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
and
(1.4)
≤
f (x) − f (y)
x − y
; x 6= y(cid:27)
(cid:12)(cid:12)(cid:12)(cid:12)
a
1
b − aZ b
M = sup(cid:26)(cid:12)(cid:12)(cid:12)(cid:12)
b − aZ b
1
−
a
≤
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
b − aZ b
p (cid:16)supnf ′ (a)
−
1
a
1
f (a) + f (b)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
(b − a)
2 (p + 1)
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
p−1 , f ′ (b)
p−1
p
p
p−1o(cid:17)
where f : [a, b] → R is differentiable function on (a, b) , and f ′
p
p−1 is quasi−convex
on [a, b] with a < b.
We give some necessary definitions and mathematical preliminaries for quasi−convex
functions which are used throughout this paper.
Definition 1. (see [1]) A function f : [a, b] → R is said to be quasi−convex on
[a, b] if
(1.5)
f (λx + (1 − λ) y) ≤ max {f (x) , f (y)} ,
holds for all x, y ∈ [a, b] and λ ∈ [0, 1] .For additional results on quasi−convexity,
see [2]. Clearly, any convex function is quasi−convex function. Furthermore, there
exists quasi−convex functions which are not convex. See [3] :
1,
t2,
t ∈ [−2, −1]
t ∈ (−1, 2]
g (t) =
is not a convex function on [−2, 2] ,but it is a quasi−convex function on [−2, 2] .If
we choose g : [−2, 2] → R, g (−2) = 1, g (2) = 4 and for α = 1
2 , a = −2, b = 0,
we get g (αa + (1 − α) b) = g (−1) = 1 and αg (a) + (1 − α) g (b) = 1
2 g (−2) +
1
2 g (0) = 1
2 .Thus it is not convex but it is quasi−convex function for all α ∈
[0, 1] , g (−α2 + (1 − α) 2) ≤ max {g (−2) , g (2)} = max {1, 4} = 4.
The main purpose of this paper is to point out new estimations of the inequality
in (1.2) , but now for the class of quasi−convex functions.
In order to prove our main results we need the following lemma (see [4]).
ON IYENGAR-TYPE INEQUALITIES
3
Lemma 1. Let f : I ⊂ R → R be a twice differentiable mapping on I ◦, a, b ∈ I
with a < b and f ′′ be integrable on [a, b]. Then the following equality holds:
f (a) + f (b)
2
−
1
b − aZ b
a
f (x)dx =
(b − a)2
2
Z 1
0
t (1 − t) f ′′ (ta + (1 − t)b) dt.
The main results of this paper are given by the following theorems.
2. The Results
Theorem 1. Let f : I ◦ ⊂ [0, ∞) → R, be a twice differentiable mapping on I ◦,
such that f ′′ ∈ L[a, b], a, b ∈ I with a < b. If f ′′q is quasi−convex on [a, b] for
q > 1, then the following inequality holds:
(2.1)
≤
q
a
2
2
1
q−1
−
(b − a)2
f (a) + f (b)
f (x)dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
, f ′′(b)q(cid:9)(cid:1)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
b − aZ b
(cid:18) q − 1
2q − p − 1(cid:19)
×(cid:0)max(cid:8)f ′′(a)q
β (x , y) =Z 1
tx−1 (1 − t)y−1 dt,
1
q
0
(β (p + 1, q + 1))
1
q
x, y > 0.
where 1
p + 1
q = 1 and β ( , ) is Euler Beta Function:
Theorem 2. Let f : I ◦ ⊂ [0, ∞) → R, be a twice differentiable mapping on I ◦,
such that f ′′ ∈ L[a, b], a, b ∈ I with a < b. If f ′′q is quasi−convex on [a, b] for
q ≥ 1, then the following inequality holds:
Theorem 3. With the assumptions of Theorem 1, we obtain another
(2.2)
≤
f (a) + f (b)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
(b − a)2
(cid:18)
q
a
2
1
q−1
−
f (x)dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
b − aZ b
(q + 1) (q + 2)(cid:19)
(cid:0)max(cid:8)f ′′(a)q
f (x)dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p (cid:0)max(cid:8)f ′′(a)q
b − aZ b
(β (2, p + 1))
−
1
a
1
3. proof of main results
4
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (a) + f (b)
2
(b − a)2
21+ 1
q
≤
1
q .
, f ′′(b)q(cid:9)(cid:1)
1
q .
, f ′′(b)q(cid:9)(cid:1)
Proof of Theorem 1: Using Lemma 1 and the well known Holder's inequality
for q > 1,
≤
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (a) + f (b)
2
−
(b − a)2
2
(cid:18)Z 1
0
t
a
1
b − aZ b
q−1 dt(cid:19)
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q (cid:20)Z 1
q−1
q−p
0
tp (1 − t)q f ′′ (ta + (1 − t) b)q
1
q
,
dt(cid:21)
4
M. EMIN OZDEMIR⋆
where 1
p + 1
q = 1.
On the other hand, since f ′′q is quasi−convex on [a, b], we know that for any
t ∈ [0, 1]
Therefore, we obtain
f ′′ (ta + (1 − t) b)q ≤ max(cid:8)f ′′(a)q
, f ′′(b)q(cid:9) .
f (a) + f (b)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
(b − a)2
2
(b − a)2
2
(b − a)2
2
q−1
a
0
t
1
q−p
−
b − aZ b
(cid:18)Z 1
q−1 dt(cid:19)
(cid:18)Z 1
q−1 dt(cid:19)
2q − p − 1(cid:19)
(cid:18) q − 1
q−p
t
0
q
q−1
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q (cid:20)Z 1
q (cid:20)Z 1
tp (1 − t)q(cid:0)max(cid:8)f ′′(a)q
0
0
q−1
tp (1 − t)q f ′′ (ta + (1 − t) b)q
(β (p + 1, q + 1))
1
q (cid:0)max(cid:8)f ′′(a)q
≤
≤
=
1
q
dt(cid:21)
1
q
1
q ,
, f ′′(b)q(cid:9)(cid:1) dt(cid:21)
, f ′′(b)q(cid:9)(cid:1)
which completes the proof.
Corollary 1. In Theorem 1, if we choose M = Supx∈(a,b) f ′′(x) < ∞, we get
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (a) + f (b)
2
1
−
b − aZ b
2q − p − 1(cid:19)
M(cid:18) q − 1
a
≤
(b − a)2
2
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q−1
q
(β (p + 1, q + 1))
1
q .
Proof of Theorem 2: From Lemma 1 and the well known power-mean in-
equality we obtain
t (1 − t) f ′′ (ta + (1 − t) b) dt
f (a) + f (b)
2
−
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(b − a)2
2
(b − a)2
2
(b − a)2
2
(b − a)2
2
(b − a)2
4
a
0
0
1
1− 1
b − aZ b
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q (cid:18)Z 1
q (cid:18)Z 1
(q + 1)(q + 2)(cid:19)
Z 1
(cid:18)Z 1
tdt(cid:19)
(cid:18)Z 1
tdt(cid:19)
2(cid:19)1− 1
q (cid:18)
(cid:18) 1
(q + 1)(q + 2)(cid:19)
(cid:18)
1− 1
2
1
1
q
0
0
0
≤
≤
≤
=
=
1
q
dt(cid:19)
t (1 − t)q f ′′ (ta + (1 − t) b)q
1
q
t (1 − t)q(cid:0)max(cid:8)f ′′(a)q
(cid:0)max(cid:8)f ′′(a)q
, f ′′(b)q(cid:9)(cid:1)
(cid:0)max(cid:8)f ′′(a)q
1
q
, f ′′(b)q(cid:9)(cid:1) dt(cid:19)
, f ′′(b)q(cid:9)(cid:1)
1
q
1
q .
The proof of Theorem 2 is completed.
Corollary 2. Under the assumptions of Theorem 2,
ON IYENGAR-TYPE INEQUALITIES
5
have
2
Case i: Since limq→∞(cid:16)
<(cid:18)
(q+1)(q+2)(cid:17)
(q + 1)(q + 2)(cid:19)
1
3
2
1
q
= 1 and limq→1+(cid:16)
1
q
2
(q+1)(q+2)(cid:17)
= 1
3 , we
1
q
< 1,
q ∈ [1, ∞).
Therefore,
f (a) + f (b)
2
−
1
b − aZ b
a
(3.1) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
In (3.1),
≤
(b − a)2
4
(cid:0)max(cid:8)f ′′(a)q
, f ′′(b)q(cid:9)(cid:1)
1
q .
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
• if f ′′q is decreasing, we get
f (a) + f (b)
2
−
1
b − aZ b
a
• if f ′′q is increasing , we get
f (a) + f (b)
2
−
1
b − aZ b
a
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤
(b − a)2
4
f ′′(a) ,
≤
(b − a)2
4
f ′′(b) .
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Case ii: If we choose M = Supx∈(a,b) f ′′(x) < ∞ in (2.2), then the inequality
in (2.1) is better than the inequality in (2.2).
Proof of Theorem 3: From Lemma 1 with properties of modulus we get
(3.2)
≤
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (a) + f (b)
2
−
1
b − aZ b
a
(b − a)2
2
Z 1
0
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
t (1 − t) f ′′ (ta + (1 − t) b) dt.
Now, if we use the following weighted version of Holder's inequality [5, p. 117]:
(3.3)
ZI
(cid:12)(cid:12)(cid:12)(cid:12)
f (s)g(s)h(s)ds(cid:12)(cid:12)(cid:12)(cid:12)
≤(cid:18)ZI
f (s)p h(s)ds(cid:19)
1
p (cid:18)ZI
1
q
g(s)q h(s)ds(cid:19)
for p > 1, p−1 + q−1 = 1, h is nonnegative on I and provided all the other integrals
exist and are finite.
6
M. EMIN OZDEMIR⋆
If we rewrite the inequality (3.2) with respect to (3.3) with f ′′q is quasi−convex
on [a, b] for all t ∈ [0, 1], we get
f (a) + f (b)
2
−
1
b − aZ b
a
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
t (1 − t) f ′′ (ta + (1 − t) b) dt
(1 − t) f ′′ (ta + (1 − t) b) tdt
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
0
Z 1
Z 1
(cid:18)Z 1
0
0
(b − a)2
2
(b − a)2
2
(b − a)2
2
(b − a)2
2
(1 − t)p tdt(cid:19)
1
p (cid:18)Z 1
0
f ′′ (ta + (1 − t) b)q
1
q
tdt(cid:19)
(β (2, p + 1))
1
p max(cid:8)f ′′(a)q , f ′′(b)q(cid:9)
2
1
q
!
.
The proof of Theorem 3 is completed.
Corollary 3. In Theorem 3, if we choose M = Supx∈(a,b) f ′′(x) < ∞, we get
f (a) + f (b)
2
−
1
b − aZ b
a
(b − a)2
21+ 1
q
M (β (2, p + 1))
1
p .
Remark 1. From Theorems 1-3, we get
≤
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
b − aZ b
1
a
−
q−1
q
(β (p + 1, q + 1))
1
≤ min {v1, v2, v3}
≤
=
≤
=
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
f (a) + f (b)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18) q − 1
2q − p − 1(cid:19)
(cid:18)
(b − a)2
4
where
v1 =
(b − a)2
2
v2 =
and
q−1
q
2
(q + 1) (q + 2)(cid:19)
1
q ,
, f ′′(b)q(cid:9)(cid:1)
q ×(cid:0)max(cid:8)f ′′(a)q
, f ′′(b)q(cid:9)(cid:1)
(cid:0)max(cid:8)f ′′(a)q
1
q
v3 =
(b − a)2
21+ 1
q
(β (2, p + 1))
1
p (cid:0)max(cid:8)f ′′(a)q
, f ′′(b)q(cid:9)(cid:1)
1
q .
4. Error Estimates for the Trapezoidal Rule
Let d be a partition a = x0 < x1 < x2 < ... < xn = b of the interval [a, b] and
consider the quadrature formula
(4.1)
where
Z b
a
f (x)dx = Ti(f, d) + Ei(f, d),
i = 1, 2, ..., n − 1
T1(f, d) =
n−1
Xi=0
f (xi) + f (xi+1)
2
(xi+1 − xi)
ON IYENGAR-TYPE INEQUALITIES
7
for the Trapezoidal version and
n−1
T2(f, d) =
f(cid:18) xi + xi+1
2
(cid:19) (xi+1 − xi)
Xi=0
for the Midpoint formula and Ei(f, d) denotes the associated approximation errors.
Proposition 1. Suppose that all the assumptions of Theorem 1 are satisfied for
every division d of [a, b], we have
E(f, d) ≤
q−1
(β (p + 1, q + 1))
q
1
n−1
2q − p − 1(cid:19)
2(cid:18) q − 1
(xi+1 − xi)3(cid:0)max(cid:8)f ′′(xi)q
Xi=0
×
1
q
, f ′′(xi+1)q(cid:9)(cid:1)
1
q .
Proof. Appliying Theorem 1 on the subinterval (xi+1, xi) , i = 1, 2, ..., n − 1 of the
partition and by using the quasi−convexity of f ′′q , we obtain
Hence in (4.1), we have
a
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Z b
f (x)dx − T (f, d)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤
f (x)dx(cid:12)(cid:12)(cid:12)(cid:12)
(β (p + 1, q + 1))
1
q
1
2
xi
1
2
q−1
q
−
q .
n−1
(xi+1 − xi)2
f (xi) + f (xi+1)
(xi+1 − xi)Z xi+1
(cid:18) q − 1
2q − p − 1(cid:19)
, f ′′(xi+1)q(cid:9)(cid:1)
(cid:12)(cid:12)(cid:12)(cid:12)
×(cid:0)max(cid:8)f ′′(xi)q
= (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xi=0(cid:26)Z xi+1
Z xi+1
Xi=0(cid:12)(cid:12)(cid:12)(cid:12)
2q − p − 1(cid:19)
2(cid:18) q − 1
(xi+1 − xi)3(cid:0)max(cid:8)f ′′(xi)q
Xi=0
f (x)dx −
f (x)dx −
n−1
n−1
×
≤
≤
q−1
q
1
2
2
xi
xi
(β (p + 1, q + 1))
f (xi) + f (xi+1)
f (xi) + f (xi+1)
(xi+1 − xi)(cid:27)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(xi+1 − xi)(cid:12)(cid:12)(cid:12)(cid:12)
, f ′′(xi+1)q(cid:9)(cid:1)
q .
1
q
1
(cid:3)
Proposition 2. Suppose that all the assumptions of Theorem 2 are satisfied for
every division d of [a, b], we have
E(f, d) ≤
1
q
2
1
n−1
4(cid:18)
(q + 1)(q + 2)(cid:19)
(xi+1 − xi)3(cid:0)max(cid:8)f ′′(xi)q
Xi=0
×
, f ′′(xi+1)q(cid:9)(cid:1)
1
q .
Proof. The proof is immediate follows from Theorem 2 and by applying a similar
argument to the Proposition 1.
(cid:3)
8
M. EMIN OZDEMIR⋆
Proposition 3. Suppose that all the assumptions of Theorem 3 are satisfied for
every division d of [a, b], we have
(β (1, p + 1))
1
p
E(f, d) ≤
1
2
×
(xi+1 − xi)2 max(cid:8)f ′′(a)q , f ′′(b)q(cid:9)
2
!
1
q
.
n−1
Xi=0
Proof. The proof is immediate follows from Theorem 3 and by applying a similar
argument to the Proposition 1.
(cid:3)
5. Applications to special means
Let us consider the special means for real numbers a, b (a 6= b). We take
1. Arithmetic mean:
A(a, b) =
a + b
2
,
a, b ∈ R.
2. Logarithmic mean:
L(a, b) =
a − b
ln a − ln b
3. Generalized log-mean:
Ln(a, b) =(cid:20) bn+1 − an+1
(n + 1)(b − a)(cid:21)
a 6= b , a, b 6= 0, a, b ∈ R.
,
1
n
, n ∈ Z\ {−1, 0} , a, b ∈ R, a 6= b.
Proposition 4. Let a, b ∈ R, a < b and n ∈ N, n ≥ 2. Then we have
A(an, bn) − Ln
n(a, b) ≤
n(n − 1) (b − a)2
2
×(cid:18) q − 1
2q − p − 1(cid:19)
q−1
q
(β (p + 1, q + 1))
1
q (cid:16)maxna(n−2)q , b(n−2)qo(cid:17)
1
q
.
Proof. The assertion follows from Theorem 1 applied to the quasi−convex mapping
f (x) = xn, x ∈ R.
(cid:3)
Proposition 5. Let a, b ∈ R, a < b and n ∈ N, n ≥ 2. Then we have
A(an, bn) − Ln
n(a, b) ≤
n(n − 1) (b − a)2
4
1
×(cid:18)
2
(q + 1)(q + 2)(cid:19)
q (cid:16)maxna(n−2)q , b(n−2)qo(cid:17)
1
q
.
Proof. The assertion follows from Theorem 2 applied to the quasi−convex mapping
f (x) = xn, x ∈ R.
(cid:3)
Proposition 6. Let a, b ∈ R, a < b and n ∈ N, n ≥ 2. Then we have
A(an, bn) − Ln
n(a, b) ≤
n(n − 1) (b − a)2
21+ 1
q
× (β (2, q + 1))
1
p (cid:16)maxna(n−2)q , b(n−2)qo(cid:17)
1
q
.
Proof. The assertion follows from Theorem 3 applied to the quasi−convex mapping
f (x) = xn, x ∈ R.
(cid:3)
ON IYENGAR-TYPE INEQUALITIES
9
References
[1] J. Pe´caric, F.Proschan, and Y.L.Tong, Convex Functions, Partial Orderings and Statistical
Applications, Academic Press(1992), Inc.
[2] J. Ponstein, Seven kinds of convexity, SIAM Review 9, 115-119 (1967)
[3] D. A. Ion, Some estimates on the Hermite- Hadamard inequality through Quasi- convex
functions, Annals of University of Craiova Math. Comp.Sc. Ser A (2007), 82-87.
[4] M. Alomari, M: Darus
and
type
Hadamard
quasi−convex, RGMIA Res. Rep.Coll.12(2009) Supplement, Article
htpp://www.staff.vu.edu.au/RGMIA/v12(E).asp.
functions whose
for
S.S.Dragomir,
second
New inequalities
derivatives
absolute
of Hermite-
are
17, Online:
value
[5] S.S.Dragomir, R.P.Agarwal, N.S. Barnett, Inequalities for Beta and Gamma Functions via
some classical and new integral inequalities, Journal of Inequalities and Applications, Vol 5,
pp.103-165,2000.
[6] D.S. Mitrinovi´c, J.E.Pe´caric. A.M. Fink, Classical and new Inequalities in analysis, Kluwer
Academic Publishers, 1993, p.106,10,15.
[7] U.S. Kirmaci, M. K. Bakula, M. E. Ozdemir and J. Pe´caric, Hadamard type inequalities for
s−convex functions, Applied Mathematics and Computation, 193 (2007), p.106, 26-35.
[8] M.K. Bakula, M.E. Ozdemir and J. Pe´caric, Hadamard type inequalities for m−convex and
(α, m)−convex functions, Journal of Inequalities in Pure and Applied Mathematics, Vol. 9
(2008), Issue 4, article 96.
[9] E. Set, M.E. Ozdemir and S.S.Dragomir, On the Hermite-Hadamard inequality and other
integral inequalities involving two functions, Journal of Inequalities and Applications, 2010,
Article ID 148102,9 pages,doi.10.1155/2010/148102.
[10] M.E. Ozdemir, C¸ . Yildiz, A. O. Akdemir, On some new Hadamard type inequalities for
Coordinated quasi−convex functions, Submitted.
[11] E. Set, M.E. Ozdemir, S.S.Dragomir, On Hadamard type inequalities involving several
kinds of convexity, Journal of Inequalities and Applications, 2010, Article ID 1286845,12
pages,doi.10.1155/2010/286845.
[12] M. Avci, H. Kavurmaci, M. E. Ozdemir, New inequalities of Hermite-Hadamard type via
s−convex functions in the second sense with applications, Appl. Math. Comput., 217 (2011)
5171-5176.
[13] M. E. Ozdemir, M. Avci and H. Kavurmaci, Hermite-Hadamard type inequalities via
(α, m)−convexity, Comput. Math. Appl., 61 (2011) 2614 -- 2620.
[14] M. E. Ozdemir, M. Avci and E. Set, On some inequalities of Hermite-Hadamard type via
m−convexity, Appl. Math. Lett., 23 (2010) 1065 1070.
[15] H. Kavurmaci, M. Avci and M. E. Ozdemir, New inequalities of Hermite-Hadamard type for
convex functions with applications, Journal of Inequalities and Applications 2011, 2011:86.
[16] M.E. Ozdemir, A. O. Akdemir, On some Hadamard type inequalities for convex functions on
a rectangular box, Journal of non linear Analysis and Application, Volume 2011, Year 2011
Article ID jnaa-00101, 10pages, doi: 10.5899/2011/jnaa-00101, Research Article.
[17] A. O. Akdemir, M.E. Ozdemir and S. Varosanec, On some inequalities for h−concave func-
tions, Mathematical and Computer Modelling 55 (2012) 746-753.
⋆ATAT URK UNIVERSITY, K.K. EDUCATION FACULTY, DEPARTMENT OF MATH-
EMATICS, 25240, CAMPUS, ERZURUM, TURKEY
E-mail address: [email protected]
|
1910.11053 | 1 | 1910 | 2019-10-24T12:30:07 | Minimal passive realizations of generalized Schur functions in Pontryagin spaces | [
"math.FA"
] | Passive discrete-time systems in Pontryagin space setting are investigated. In this case the transfer functions of passive systems, or characteristic functions of contractive operator colligations, are generalized Schur functions. The existence of optimal and *-optimal minimal realizations for generalized Schur functions are proved. By using those realizations, a new definition, which covers the case of generalized Schur functions, is given for defects functions. A criterion due to D.Z. Arov and M.A. Nudelman, when all minimal passive realizations of the same Schur function are unitarily similar, is generalized to the class of generalized Schur functions. The approach used here is new; it relies completely on the theory of passive systems. | math.FA | math |
Minimal passive realizations of generalized
Schur functions in Pontryagin spaces
Lassi Lilleberg
Abstract. Passive discrete-time systems in Pontryagin space setting are
investigated. In this case the transfer functions of passive systems, or
characteristic functions of contractive operator colligations, are gener-
alized Schur functions. The existence of optimal and ∗-optimal minimal
realizations for generalized Schur functions are proved. By using those
realizations, a new definition, which covers the case of generalized Schur
functions, is given for defects functions. A criterion due to D.Z. Arov
and M.A. Nudelman, when all minimal passive realizations of the same
Schur function are unitarily similar, is generalized to the class of general-
ized Schur functions. The approach used here is new; it relies completely
on the theory of passive systems.
1. Introduction
An operator colligation Σ = (TΣ; X , U , Y; κ) consists of separable Pontryagin
spaces X (the state space), U (the incoming space), and Y (the outgoing
space) and the system operator TΣ ∈ L(X ⊕ U , X ⊕ Y), the space of bounded
U(cid:19), means the direct
operators from X ⊕ U to X ⊕ Y, where X ⊕ U , or (cid:18)X
orthogonal sum with respect to the indefinite inner product. The symbol κ
is reserved for the finite negative index of the state space. The operator TΣ
has the block representation of the form
TΣ =(cid:18)A B
C D(cid:19) :(cid:18)X
U(cid:19) →(cid:18)X
Y(cid:19) ,
(1.1)
Department of Mathematics and Statistics, University of Vaasa, P.O. Box 700,
65101 Vaasa, Finland
E-mail address: [email protected] .
Mathematics Subject Classification (2010). Primary: 47A48; Secondary:
47A56, 47B50, 93B05, 93B07, 93B28.
Keywords. Operator colligation, passive system, transfer function, defect
functions, generalized Schur class, contractive operator.
2
L. Lilleberg
where A ∈ L(X ) (the main operator), B ∈ L(U , X ) (the control operator),
C ∈ L(X , Y) (the observation operator), and D ∈ L(U , Y) (the feedthrough
operator). If needed, the colligation is written as Σ = (A, B, C, D; X , U , Y; κ).
It is always assumed in this paper that U and Y have the same negative index.
All notions of continuity and convergence are understood to be with respect
to the strong topology, which is induced by any fundamental decomposition
of the space in question.
The colligation (1.1) will be called as a system since it can be seen as a
linear discrete time system of the form
(hk+1 = Ahk + Bξk,
= Chk + Dξk,
σk
k ≥ 0,
where {hk} ⊂ X , {ξk} ⊂ U and {σk} ⊂ Y. In what follows, the "system" is
identified with the operator expression appearing in (1.1). When the system
operator TΣ in (1.1) is contractive (isometric, co-isometric, unitary), with
respect to the indefinite inner product, the corresponding system is called
passive (isometric, co-isometric, conservative). In literature, conservative sys-
tems are also called unitary systems. The transfer function of the system (1.1)
is defined by
θΣ(z) := D + zC(I − zA)−1B,
(1.2)
whenever I − zA is invertible. Especially, θΣ is defined and holomorphic in a
neighbourhood of the origin. The values θΣ(z) are bounded operators from
U to Y. Conversely, if θ is an operator valued function holomorphic in a
neighbourhood of the origin, and transfer function of the system Σ coinsides
with it, then Σ is a realization of θ. In some sources, transfer functions of the
systems are also called characteristic functions of operator colligations.
Σ of TΣ. That is, Σ∗ = (T ∗
The adjoint or dual of the system Σ is the system Σ∗ such that its system
operator is the indefinite adjoint T ∗
Σ; X , Y, U; κ). In
this paper, all the adjoints are with respect to the indefinite inner product.
For an operator valued function ϕ, the notation ϕ∗(z) is used instead of
(ϕ(z))∗ , and the function ϕ#(z) is defined to be ϕ∗(¯z). With this notation,
#(z). Since
for the transfer function θΣ∗ of Σ∗, it clearly holds θΣ∗ (z) = θΣ
contractions between Pontryagin spaces with the same negative index are
bi-contractions (cf. eg. [24, Corollary 2.5]), Σ∗ is passive whenever Σ is.
In the case where all the spaces are Hilbert spaces, it is well known; see
for instance [6, Proposition 8], that the transfer function of the passive system
is a Schur function. That is, contractive operator valued function holomorphic
in the unit disk D. In the case where U and Y are Hilbert spaces and the state
space X is a Pontryagin space, Saprikin showed in [28, Theorem 2.2] that the
transfer function of the passive system (1.1) is a generalized Schur function.
It will be proved later in Proposition 2.5 that this result holds also in the
case when all the spaces are Pontryagin spaces. The generalized Schur class
Sκ(U , Y), where U and Y are Pontryagin spaces with the same negative index,
is the set of L(U , Y)-valued functions S(z) holomorphic in a neighbourhood
(cid:0)hKS(wj, wi)fj, fiiY(cid:1)n
i,j=1
,
(1.4)
Minimal realizations
3
Ω of the origin such that the Schur kernel
KS(w, z) =
1 − S(z)S ∗(w)
1 − z ¯w
,
w, z ∈ Ω,
(1.3)
has κ negative squares (κ = 0, 1, 2, . . .). This means that for any finite set of
points w1, . . . , wn in the domain of holomorphy ρ(S) of S and set of vectors
{f1, . . . , fn} ⊂ Y, the Hermitian matrix
where h·, ·iY is the indefinite inner product of the space Y, has no more than
κ negative eigenvalues, and there exists at least one such matrix that has
exactly κ negative eigenvalues. A function S belongs to Sκ(U , Y) if and only
if S#
κ ∈ S(Y, U); see [1, Theorem 2.5.2]. The class S0(U , Y) coinsides with
the ordinary Schur class, and it is written as S(U , Y).
The direct connection between the transfer functions of passive systems
of the form (1.1) and the generalized Schur functions allows to study the prop-
erties of generalized Schur functions by using passive systems, and vice versa.
Therefore, a fundamental problem of the subject is, for a given θ ∈ Sκ(U , Y),
find a realization Σ of θ with the desired minimality or optimality properties
(observable, controllable, simple, minimal, optimal, ∗-optimal); for details,
see Theorems 2.7 and 3.5 and Lemma 2.9. The described problem is called
a realization problem. In the standard Hilbert space settings, realizations
problems, as well as other properties of passive systems, were studied, for
instance, by Ando [3], Arov [5, 6], Arov et al. [7, 8, 9, 10], de Branges and
Rovnyak [19, 20], Brodskiı [21], Helton [25] and Sz.-Nagy and Foias [30]. The
case where the state space is a Pontryagin space while incoming and outgo-
ing spaces are still Hilbert spaces, unitary systems were studied, for instance,
by Dijksma et al. [22, 23], and passive systems by Saprikin [28], Saprikin
and Arov [12], Saprikin et al. [11] and by the author in [27]. The case where
all the spaces are Pontryagin spaces, theory of isometric, co-isometric and
conservative systems is considered, for instance, in [1, 2, 24].
Especially, in [6], Arov proved the existence of so-called optimal minimal
realizations of an ordinary Schur function; for definitions, see Section 3. The
proof was based on the existence (right) defect functions. For an ordinary
Schur function S(ζ), the (right) defect function ϕ of S is, roughly speaking,
the maximal analytic minorant of I − S ∗(ζ)S(ζ). More precicely, this means
that for almost every (a.e.) ζ on the unit circle T, it holds
ϕ∗(ζ)ϕ(ζ) ≤ I − S ∗(ζ)S(ζ),
it holds
and for every other operator valued analytic function bϕ with similar property,
For the existence of defect functions, see [30, Theorem V.4.2], and for a
detailed treatise, see [16, 17, 18]. In [7], Arov et al. constructed (∗-)optimal
minimal passive systems in the Hilbert space setting without using defect
functions. The construction can be done by taking an appropriate restriction
bϕ∗(ζ)bϕ(ζ) ≤ ϕ∗(ζ)ϕ(ζ).
4
L. Lilleberg
of some system. In the indefinite setting, if one uses a suitable definition of
optimality, a similar method as was used by Arov et al. still produces a (∗-
)optimal minimal passive system. In Pontryagin state space case, this was
proved by Saprikin [28]. It will be shown in Theorem 3.5 that the same result
still holds in the case where all the spaces are Pontryagin spaces.
r
The study of the class of generalized Schur functions Sκ(U , Y) was con-
tinued in [11, 12], in the case where U and Y are Hilbert spaces and the state
space is a Pontryagin space. In [12], Saprikin and Arov used the right Kreın-
Langer factorization of the form S = SrB−1
for S ∈ Sκ(U , Y), and proved
that the existence of the optimal minimal realization of S is equivalent to
the existence of the right defect function of Sr. However, they did not define
the defect functions for the generalized Schur functions. This was done by
the author in [27] by using the Kreın-Langer factorizations. With the defini-
tion given therein, the main results of [4] were generalized to the Pontryagin
state space setting. The main subjects of [27] include some continuation of
the study of products of systems and the stability properties of passive sys-
tems, subjects treated earlier by Saprikin et al. in [11]. In the present paper,
it will be shown that a concept of defect functions can be defined in the
case where all the spaces are Pontryagin spaces. The key idea here is to use
optimal minimal passive realizations and conservative embeddings. By using
such a definition, it is shown that one can generalize and improve some of
the main results from [4], using different proofs than those given in [4] or
[27], see Theorem 4.8. Furthermore, in Theorem 4.10, the main results from
[9, 10] concerning the criterion when all the minimal realizations of a Schur
function are unitarily similar, is generalized to the present indefinite setting.
The proof will be carried out entirely by using the theory of passive systems,
without applying Hardy space theory or the theory of Hankel operators as in
the proof provided in [10].
The paper is organized as follows. In Section 2 basic facts of linear
systems, Julia operators, dilations and embeddings are recalled. As a prepa-
ration, it is shown in Proposition 2.3 that an arbitrary passive linear system
has a conservative dilation. Moreover, Lemma 2.9 gives some usefull rep-
resentations and restrictions of passive systems. That lemma will be used
extensively later on in this paper.
In Section 3, the existence and basic properties of (∗-)optimal minimal
realizations are established. The main result of this section is Theorem 3.5.
The generalized defect functions are introduced in Section 4. In partic-
ularly, Theorem 4.10 in this section can be seen as the main result of the
paper.
Minimal realizations
5
2. Linear systems, dilations and embeddings
Let Σ = (TΣ; X , U , Y; κ) be a linear system as in (1.1). The following sub-
spaces
X c := span {ran AnB : n = 0, 1, . . .}
X o := span {ran A∗nC ∗ : n = 0, 1, . . .}
X s := span {ran AnB, ran A∗mC ∗ : n, m = 0, 1, . . .},
(2.1)
(2.2)
(2.3)
are called, respectively, controllable, observable and simple subspaces. The
system is said to be controllable (observable, simple) if X c = X (X o =
X , X s = X ) and minimal if it is both controllable and observable. When
Ω ∋ 0 is some symmetric neighbourhood of the origin, that is, ¯z ∈ Ω when-
ever z ∈ Ω, then also
X c = span {ran (I − zA)−1B, z ∈ Ω}
X o = span {ran (I − zA∗)−1C ∗, z ∈ Ω}
X s = span {ran (I − zA)−1B, ran (I − wA∗)−1C ∗, z, w ∈ Ω}
(2.4)
(2.5)
(2.6)
The system (1.1) can be expanded to a larger system without changing
the transfer function. It can be done by using the so-called Julia operator,
see (2.8) below. For a proof of the following theorem and more details about
the Julia operators, see [24]. The basic information about the indefinite inner
product spaces and their operators can be recalled from [13, 15, 24].
Theorem 2.1. Suppose that X1 and X2 are Pontryagin spaces with the same
negative index, and let A : X1 → X2 be a contraction. Then there exist Hilbert
spaces DA and DA∗ , linear operators DA : DA → X1, DA∗ : DA∗ → X2 with
zero kernels and a linear operator L : DA → DA∗ such that it holds
I − A∗A = DAD∗
A,
I − AA∗ = DA∗D∗
A∗ ,
(2.7)
and the operator
UA :=(cid:18) A DA∗
A −L∗(cid:19) :(cid:18) X1
D∗
DA(cid:19)
DA∗(cid:19) →(cid:18) X2
(2.8)
is unitary. Moreover, DA, DA∗ and UA are unique up to unitary equivalence.
Remark 2.2. The operator DA from Theorem 2.1 is called a defect operator
of A.
A dilation of a system Σ = (A, B, C, D; X , U , Y; κ) is any system of the
form bΣ = (bA, bB, bC, D; bX , U , Y; κ), where
bX = D ⊕ X ⊕ D∗,
bAD ⊂ D,
bA∗D∗ ⊂ D∗,
bCD = {0},
bB∗D∗ = {0}.
(2.9)
6
L. Lilleberg
The spaces D and D∗ are required to be Hilbert spaces. The system operator
B1
B
(2.10)
(2.11)
(2.12)
D
X
D∗
0
0
0
0
D
X
D∗
B1
B
D
A11 A12 A13
A A23
0
A33
A11 A12 A13
A A23
0
A33
Y
,
U
→
bC =(cid:0)0 C C1(cid:1) .
TbΣ of bΣ is of the form
TbΣ =
0
:
(cid:0)0 C C1(cid:1)
0 ,
,
bB =
bA =
The system Σ is called a restriction of bΣ. Recall that subspace N of
inner product of H. Since X clearly is a regular subspace of bX , there exists
the unique orthogonal projection PX from bX to X . Let bA↾X be the restriction
of bA to the subspace X . Then, the system Σ can be represented as
the Pontryagin space H is regular if it is itself a Pontryagin space with the
inherited inner product of h·, ·iH. The subspace N is regular precicely when
N ⊥ is regular, where ⊥ refers to orthogonality with respect to the indefinite
Σ = (PX bA↾X , PXbB,bC↾X , D; PX bX , U , Y; κ).
Dilations and restrictions are denoted by
bΣ = dilX →bX Σ, Σ = resbX →XbΣ,
mostly without subscripts when the corresponding state spaces are clear. A
calculation show that the transfer functions of the original system and its
dilation coincide. Moreover, if Σ is passive, then is any retriction of it. The
following proposition shows that a passive system has a conservative dilation.
For the Hilbert space case, this result is from [5], and for the Pontryagin state
space case, see [28].
Proposition 2.3. Let Σ = (A, B, C, D; X , U , Y; κ) be a passive system. Then
Proof. Since the system operator T of Σ is a contraction, by Theorem 2.1,
there exists a Julia operator
there exists a conservative dilation bΣ = (bA, bB,bC, D; bX , U , Y; κ) of Σ.
(cid:18) T DT ∗
DT (cid:19) ⇐⇒
T −L∗(cid:19) :(cid:18)X ⊕ U
DT ∗ (cid:19) →(cid:18)X ⊕ Y
(cid:18)PX DT ∗
PY DT ∗(cid:19)
−L∗ :(cid:18)X
→(cid:18)X
U(cid:19)DT ∗
Y(cid:19)DT
(cid:18)A B
C D(cid:19)
(cid:0)D∗
T ↾U(cid:1)
with properties introduced in Theorem 2.1. Denote EH(h) = hh, hiH for a
vector h in an inner product space H. Then, for x ∈ X , u ∈ U and f ∈ DT ∗ ,
one has
T ↾X D∗
(2.13)
D∗
EX (x) + EU (u) + ED
T ∗ (f ) = EX (Ax + Bu + PX DT ∗ f )
+ EY (Cx + Du + PY DT ∗ f )
(2.14)
+ EDT (D∗
T ↾X x + D∗
T ↾U u − L∗f ).
Minimal realizations
7
Also, for given x′ ∈ X , y ∈ Y and f ′ ∈ DT , there exists an unique triplet
x ∈ X , u ∈ U and f ∈ DT ∗ such that
A
B
C
D
T ↾X D∗
D∗
T ↾U
DT ,
ℓ2
+(DT ∗ ) :=
x
u
PX DT ∗
PY DT ∗
−L∗
f =
∞M1
bX := ℓ2
DT ∗ ,
x′
y
f ′
(2.15)
−(DT ) ⊕ X ⊕ ℓ2
+(DT ∗ ),
Define
ℓ2
−(DT ) :=
−1M−∞
ℓ2
where L denotes an orthogonal sum of Hilbert spaces. Since ℓ2
+(DT ∗ ) are Hilbert spaces, the space bX clearly is a Pontryagin space with the
negative index κ. For u ∈ U and (. . . f−1, x, f1 . . .) ∈ bX , where the underlined
element belongs to X , define the operators
−(DT ) and
T ↾X x − L∗f1, Ax + PX DT ∗ f1, f2, f3, . . .),
T ↾U u, Bu, 0, . . .),
(. . . , f−1, D∗
bA(. . . , f−2, f−1, x, f1, f2, f3, . . .) :=
bBu := (. . . , 0, D∗
bC(. . . , f−1, x, f1, . . .) := Cx + PY DT ∗f1.
tion of Σ. Moreover, calculations by using the identities (2.14) and (2.15)
A calculation shows that the system bΣ := (bA, bB,bC, D; bX , U , Y; κ) is a dila-
show that the system operator TbΣ is unitary, and therefore the system bΣ is
conservative.
(cid:3)
It is possible that D = {0} or D∗ = {0} in (2.9). In those cases, the zero
space and the corresponding row and column will be left out in (2.10). In
particular, if the system Σ with the system operator T as in (1.1) is isometric
(co-isometric), then DT = 0 (DT ∗ = 0), and proceeding as in the proof of
form (2.10) such that D = {0} (D∗ = {0}). Moreover, for an arbitrary pas-
sive system as in (1.1), it is possible to construct an isometric (co-isometric)
Proposition 2.3, it is possible to construct a conservative dilation bΣ of the
dilation bΣ of the form (2.10) such that D∗ = {0} (D = {0}).
There is also an another way to expand the system (1.1), and it is called
an embedding. In this expansion, the state space and the main operator will
not change. The embedding of the system (1.1) is any system determined by
the system operator
TeΣ = A eB
eC eD! :(cid:18)X
eU(cid:19) →(cid:18)X
eY(cid:19) ⇐⇒
D21 D22(cid:19) :
U ′(cid:19) →
Y ′(cid:19) ,
(cid:0)B B1(cid:1)
X(cid:18) U
X(cid:18) Y
C1(cid:19) (cid:18) D D12
(cid:18) C
A
(2.16)
8
L. Lilleberg
where U ′ and Y ′ are Hilbert spaces. The transfer function of the embedded
system is
C1(cid:19) (IX − zA)−1(cid:0)B B1(cid:1)
θeΣ(z) =(cid:18) D D12
D21 D22(cid:19) + z(cid:18) C
D21 + zC1(IX − zA)−1B D22 + zC1(IX − zA)−1B1(cid:19)
=(cid:18) D + zC(IX − zA)−1B
=(cid:18) θΣ(z)
θ22(z)(cid:19) ,
D12 + zC(IX − zA)−1B1
θ12(z)
θ21(z)
(2.17)
where θΣ is the transfer function of the original system. The embedded sys-
tems will be needed in Section 4.
It will be proved in Proposition 2.5 below that the transfer function of
any passive system (1.1) is a generalized Schur function with index not larger
than the negative index of the state space. For a special case where incoming
and outcoming spaces are Hilbert spaces, this result is due to [28, Theorem
2.2]. The proof of the general case follows the lines of Saprikin's proof of the
special case.
Lemma 2.4. Let Σ = (A, B, C, D; X , U , Y; κ) be a passive system with the
transfer function θ. Denote the system operator of Σ as T. If
DT =(cid:18)DT,1
DT,2(cid:19) : DT →(cid:18)X
U(cid:19) DT ∗ =(cid:18)DT ∗
DT ∗
,2(cid:19) : DT ∗ →(cid:18)X
Y(cid:19) ,
,1
(2.18)
are defect operators of T and T ∗, respectively, then the identities
IY − θ(z)θ∗(w) = (1 − z ¯w)G(z)G∗(w) + ψ(z)ψ∗(w),
IU − θ∗(w)θ(z) = (1 − z ¯w)F ∗(w)F (z) + ϕ∗(w)ϕ(z),
(2.19)
(2.20)
with
G(z) = C(IX − zA)−1,
F (z) = (IX − zA)−1B,
ψ(z) = DT ∗
,2 + zC(IX − zA)−1DT ∗
1
ϕ(z) = D∗
T,2 + zD∗
T,1(IX − zA)−1B,
,
(2.21)
(2.22)
hold for every z and w in a sufficiently small symmetric neighbourhood of the
origin.
Proof. By applying the results from [1, Theorem 1.2.4] and the identities in
(2.7), one deduces that for every z and w in a sufficiently small symmetric
Minimal realizations
9
neighbourhood of the origin, it holds
IY
(cid:19)
IY(cid:1)(cid:18)(IX − ¯wA∗)−1C ∗
(cid:19)
IY
IY − θ(z)θ∗(w) =(cid:0)C(IX − zA)−1
−(cid:0)zC(IX − zA)−1
=(cid:0)C(IX − zA)−1
−(cid:0)zC(IX − zA)−1
−(cid:0)zC(IX − zA)−1
IY(cid:1) T T ∗(cid:18) ¯w(IX − ¯wA∗)−1C ∗
IY(cid:1)(cid:18)(IX − ¯wA∗)−1C ∗
IY(cid:1) (I − DT ∗ D∗
IY(cid:1) ¯w(cid:16)IX − DT ∗
= C(IX − zA)−1(IX − ¯wA∗)−1C ∗ + IY
= (1 − z ¯w)C(IX − zA)−1(IX − ¯wA∗)−1C ∗
− ¯wDT ∗
(cid:19)
D∗
T ∗
,1
T ∗)(cid:18) ¯w(IX − ¯wA∗)−1C ∗
IY
IY
,1
,2
(cid:19)
,1(cid:17)(IX − ¯wA∗)−1C ∗ −DT ∗
D∗
T ∗
(IX − ¯wA∗)−1C ∗ + IY −DT ∗
,1
,2
,2!
T,2
D∗
D∗
T ∗
+ z ¯wC(IX − zA)−1DT ∗
,1
D∗
T ∗
,1
(IX − ¯wA∗)−1C ∗ + zC(IX − zA)−1DT ∗
,1
D∗
T,2
+ ¯wDT ∗
,2
D∗
T ∗
,1
(IX − ¯wA∗)−1C ∗ + DT ∗
,2
D∗
T ∗
,2
= (1 − z ¯w)G(z)G∗(w) + ψ(z)ψ∗(w).
Similar calculations show that (2.20) holds also, and the proof is complete.
(cid:3)
Note that if Σ in Lemma 2.4 is isometric (co-isometric), then DT = 0
(DT ∗ = 0) and therefore ϕ ≡ 0 (ψ ≡ 0).
Proposition 2.5. If Σ = (A, B, C, D; X , U , Y; κ) is a passive system, the the
transfer function θ of Σ belongs to Sκ′(U , Y), where κ′ ≤ κ.
Proof. Denote the system operator of Σ as T. By Lemma 2.4, the kernel Kθ
defined as in (1.3) has a representation
Kθ(w, z) = G(z)G∗(w) + (1 − z ¯w)−1ψ(z)ψ∗(w),
(2.23)
where G(z) and ψ(z) are defined as in (2.21). Since the negative index of X is
κ and the negative index of the Hilbert space DT ∗ is zero, it follows from [1,
Lemma 1.1.1.], that for any finite set of points w1, . . . , wn in the domain of
holomorphy of θ and the set of vectors {y1, . . . , yn} ⊂ Y, the Gram matrices
(cid:0)hG∗(wj)yj, G∗(wi)yiiX(cid:1)n
,
i,j=1
(cid:16)hψ∗(wj)yj, ψ∗(wi)yiiD
T ∗(cid:17)n
,
i,j=1
have, respectively, at most κ and zero negative eigenvalues. The kernel (1 −
z ¯w)−1 has no negative square, since it is the reproducing kernel of the clas-
sical Hardy space H 2(D). The Schur product theorem shows that the ker-
nel (1 − z ¯w)−1ψ(z)ψ∗(w) has no negative square. Then it follows from [1,
Theorem 1.5.5] that the kernel Kθ has at most κ negative square. That is,
θ ∈ Sκ′(U , Y), where κ′ ≤ κ, and the proof is complete.
(cid:3)
10
L. Lilleberg
Definition 2.6. A passive realization Σ of a genaralized Schur function θ ∈
Sκ(U , Y) is called κ-admissible if the negative index of the state space of Σ
coinsides with the negative index κ of θ.
In what follows, this paper deals mostly with the κ-admissible realiza-
tions. It will turn out that the κ-admissible realizations of θ ∈ Sκ(U , Y)
are well behaved is some sense; they have many similar propeties than the
standard passive Hilbert space systems.
The following realizations theorem is well known, see [1, Theorems 2.2.1,
2.2.2 and 2.3.1].
Theorem 2.7. For a generalized Schur function θ ∈ Sκ(U , Y) there exist re-
alizations Σk = (Tk; Xk, U , Y; κ), k = 1, 2, 3, of θ such that
(i) Σ1 is observable co-isometric;
(ii) Σ2 is controllable isometric;
(iii) Σ3 is simple conservative.
Conversely, if the system Σ has some of the properties (i) -- (iii), then θΣ ∈
Sκ(U , Y), where κ is the negative index of the state space of Σ.
Recall that a Hilbert subspace of the Pontryagin space X is a regular
subspace such that its negative index is zero. Conversely, anti-Hilbert sub-
space is a regular subspace such that its positive index is zero. When U and
Y happens to be Hilbet spaces, the transfer function θ of the passive system
Σ = (TΣ; X , U , Y; κ) belongs to class Sκ(U , Y) (with κ = ind−X ) if and only
if (X s)⊥ is a Hilbert subspace [27, Lemma 3.2]. In the case when U and Y are
Pontryagin spaces with the same negative index, the transfer function θ of the
isometric (co-isometric, conservative) system Σ = (TΣ; X , U , Y; κ) belongs to
class Sκ(U , Y) if and only if (X c)⊥ ((X o)⊥,(X s)⊥) is a Hilbert subspace [1,
Theorem 2.1.2]. For a passive system, one has the following result.
Proposition 2.8. For a passive realization Σ = (A, B, C, D; X , U , Y; κ) of
θ ∈ Sκ(U , Y), spaces X c, X o and X s are regular and their orthogonal com-
plements are Hilbert subspaces.
Proof. Let Ω be a symmetric neighbourhood of the origin such that (I−zA)−1
and (I − zA∗)−1 exist for every z ∈ Ω. Represent the kernel Kθ as in (2.23).
Since Kθ has κ negative square, a similar argument used in the proof of 2.5
shows that the kernel K1(z, w) = G(z)G∗(w), where G(z) = C(I −zA)−1, has
κ negative square. It follows now from [1, Lemma 1.1.1'] that span{ran (I −
wA∗)−1C ∗, w ∈ Ω} contains a κ-dimensional maximal anti-Hilbert subspace
Xκ. Then, Xκ ⊕ (Xκ)⊥ = X is a fundamental decomposition of X . Especially,
(Xκ)⊥ is a Hilbert subspace of X . But
(cid:0)span{ran (I − wA∗)−1C ∗, w ∈ Ω}(cid:1)⊥
= (X o)⊥ ⊂ (Xκ)⊥,
which implies that (X o)⊥ is a Hilbert subspace, and therefore its orthocom-
plement X o is regular.
By duality argument, the space X c is a regular subspace and the space
(X c)⊥ is a Hilbert subspace. It easily follows from (2.1) -- (2.3) that (X s)⊥ =
Minimal realizations
11
(X c)⊥ ∩ (X o)⊥, and therefore (X s)⊥ is also a Hilbert subspace and X s is
regular.
(cid:3)
It follows from the Proposition 2.8 above that the state space X of a
κ-admissible realization Σ of θ ∈ Sκ(U , Y) can be decombosed to the control-
lable, observable and simple parts. Using this fact, the lemma below, which
will be used extensively, can be proved.
Lemma 2.9. Let Σ = (A, B, C, D; X , U , Y; κ) be a passive system such that
the spaces (X o)⊥, (X c)⊥ and (X s)⊥ are Hilbert subspaces of X . Then the
system operator T of Σ has the following representations
U
D
D
U
U
0
0
(cid:18)(X o)⊥
X o (cid:19)
Bo(cid:19)
0 Ao(cid:19) (cid:18)B1
(cid:18)A1 A2
:
T =
(cid:0)0 Co(cid:1)
(cid:18)(X c)⊥
Bc(cid:19)
A4 Ac(cid:19) (cid:18) 0
X c (cid:19)
(cid:18)A3
:
T =
(cid:0)C1 Cc(cid:1)
(cid:18)(X s)⊥
Bs(cid:19)
X s (cid:19)
0 As(cid:19) (cid:18) 0
(cid:18)A5
:
T =
(cid:0)0 Cs(cid:1)
12 A′′
13
A′′ A′′
23
A′′
0
33
(cid:0)0 C ′ C ′
1(cid:1)
12 A′
13
A′ A′
A′
0
A′′
11 A′′
0
0
A′
11 A′
0
0
B′′
1
B′′
0
D
B′
1
B′
0
D
(cid:0)0 C ′′ C ′′
1(cid:1)
D
23
33
:
:
T =
T =
Y
Y
Y
(cid:18)(X o)⊥
X o (cid:19)
→
(cid:18)(X c)⊥
X c (cid:19)
→
(cid:18)(X s)⊥
X s (cid:19)
→
→
→
(X o)⊥
PX o X c
PX c X o
(X c)⊥
U
U
X o ∩ (X c)⊥
X c ∩ (X o)⊥
X c ∩ (X o)⊥
(2.24)
(2.25)
(2.26)
(2.27)
(2.28)
(X o)⊥
PX o X c
X o ∩ (X c)⊥
Y
PX c X o
(X c)⊥
Y
The restrictions
Σo = (Ao, Bo, Co, D; X o, U , Y; κ)
Σc = (Ac, Bc, Cc, D; X c, U , Y; κ)
Σs = (As, Bs, Cs, D; X s, U , Y; κ)
Σ′ = (A′, B′, C ′, D; PX oX c, U , Y; κ)
Σ′′ = (A′′, B′′, C ′′, D; PX cX o, U , Y; κ)
(2.29)
(2.30)
(2.31)
(2.32)
(2.33)
of Σ are passive, and Σo is observable, Σc is controllable, Σs is simple, and
Σ′ and Σ′′ are minimal. For any n ∈ N0 and any z in a sufficiently small
12
L. Lilleberg
symmetric neighbourhood of the origin, it holds
AnB = An
c Bc = An
s Bs,
(I − zA)−1B = (I − zAs)−1Bs = (I − zAc)−1Bc,
A∗nC ∗ = A∗
o
nC ∗
o = A∗
s
(I − zA∗)−1C ∗ = (I − zA∗
s)−1C ∗
nC ∗
s ,
s = (I − zA∗
c)−1C ∗
c .
(2.34)
(2.35)
(2.36)
(2.37)
Moreover, if Σ is co-isometric (isometric), then so are Σo and Σs (Σc and
Σs).
Proof. Since (X o)⊥, (X c)⊥ and (X s)⊥ are Hilbert spaces, the spaces X o, X c
and X s are regular subspaces with the negative index κ. It follows from the
identities (2.1) -- (2.3) that
(X o)⊥, (X s)⊥ are A-invariant,
(X c)⊥, (X s)⊥ are A∗-invariant,
ran C ∗ ⊂ X o ⊂ X s,
ran B ⊂ X c ⊂ X s,
,
(2.38)
and the representations (2.24) -- (2.26) follow. That is, Σo, Σc and Σs are re-
strictions of the passive system Σ, ans therefore they are passive.
Let TΣk be the system operator of Σk where k = o, c, s, and let x ∈
X k ⊕ U and x ∈ X k ⊕ Y. Calculation show that
TΣk x = T x,
T ∗
Σk x = T ∗x,
k = c, s,
k = o, s.
It follows from the equations above that if Σ is co-isometric (isometric), then
so are Σo and Σs (Σc and Σs).
Suppose x ∈ X o such that CoAn
o x = 0 for every n = 0, 1, 2, . . .. Then
CAnx =(cid:0)0 Co(cid:1)(cid:18)A1 A2
0 Ao(cid:19)n(cid:18)0
x(cid:19) = C0An
0 x = 0,
and the identity (2.2) implies that x ∈ X o ∩ (X o)⊥ = {0}. Thus x = 0, and
it can be deduced that Σo is observable. Similar arguments show that Σc is
controllable and Σs is simple, the details will be omitted.
Let u ∈ U , and n ∈ N0. Then, by (2.25) and (2.26),
0
AnBu =(cid:18)A3
AnBu =(cid:18)A5
A4 Ac(cid:19)n(cid:18) 0
0 As(cid:19)n(cid:18) 0
0
An
Bc(cid:19) =(cid:18) 0
Bs(cid:19) =(cid:18) 0
c Bcu(cid:19) = An
s Bsu(cid:19) = An
An
c Bcu
s Bsu,
and (2.34) holds. By Neumann series,
(I − zA)−1B =
znAnB
∞Xn=0
Minimal realizations
13
holds for all z in a sufficiently small symmetric neighbourhood of the origin,
and (2.35) follows now from (2.34). The equalities (2.36) and (2.37) can be
deduced similarly.
Since the orthocomplements (X o)⊥ and (X c)⊥ are Hilbert subspaces, it
follows from [28, Lemma 3.1] that PX oX c and PX cX o are regular subspaces,
and
X o ∩ (PX oX c)⊥ = X o ∩ (X c)⊥,
X c ∩ (PX cX o)⊥ = X c ∩ (X o)⊥.
Since (X o)⊥ ⊂ (PX oX c)⊥, (X c)⊥ ⊂ (PX cX o)⊥ and all the spaces are regular,
simple calculations show that
(PX oX c)⊥ = (X o)⊥ ⊕ (X o ∩ (PX o X c)⊥),
(PX cX o)⊥ = (X c)⊥ ⊕ (X c ∩ (PX cX o)⊥).
Therefore,
X = PX oX c ⊕ (PX oX c)⊥ = (X o)⊥ ⊕ PX oX c ⊕ (X o ∩ (PX oX c)⊥)
= (X o)⊥ ⊕ PX oX c ⊕ (X o ∩ (X c)⊥),
and similarly, X = (X c ∩ (X o)⊥) ⊕ PX cX o ⊕ (X c)⊥. Since (X o ∩ (X c)⊥
and X c ∩ (X o)⊥ are also Hilbert spaces, the spaces PX oX c and PX cX o are
Pontryagin spaces with the negative index κ. By considering the properties
in (2.38), the representations (2.27) and (2.28) follow now easily. That is, Σ′
and Σ′′ are restrictions of Σ, and therefore passive.
Denote X ′ := PX oX c. Represent the system operator T of Σ as in (2.27).
Then
PX ′AnB = PX ′
A′
11 A′
0
0
12 A′
13
A′ A′
23
A′
0
33
n
B′
1
B′
0 =
A′nB′
0
0 = A′nB′,
and similarly A′∗nC
X ′c = span {ran A′nB′ : n = 0, 1, . . .} = span {ran PX ′AnB : n = 0, 1, . . .}
′∗ = PX ′A∗nC ∗. Therefore,
= PX ′span {ran AnB : n = 0, 1, . . .} = PX ′X c = PX ′PX oX c = PX ′X ′
= X ′,
and similarly X ′o = PX ′X o = X ′, which implies that Σ′ is minimal. A similar
argument shows that Σ′′ is minimal, and the proof is complete.
(cid:3)
Note that in particular, Lemma 2.9 implies the existence of a minimal
passive realization of θ ∈ Sκ(U , Y).
Definition 2.10. The restrictions Σo, Σc, Σs, Σ′, and Σ′′ in Lemma 2.9 are
called, respectively, the observable, the controllable, the simple (or proper),
the first minimal and the second minimal restrictions of Σ.
The first minimal and the second minimal restrictions will be considered
later in Sections 3 and 4.
14
L. Lilleberg
Two realizations
Σ1 = (A1, B1, C1, D1; X1, U , Y; κ1),
Σ2 = (A2, B2, C2, D2; X2, U , Y; κ2)
of the same function θ ∈ Sκ(U , Y) are called unitarily similar if D1 = D2 and
there exists a unitary operator U : X1 → X2 such that
A1 = U −1A2U, B1 = U −1B2, C1 = C2U.
(2.39)
In that case, it easily follows that κ1 = κ2. Unitary similarity preserves dy-
namical properties of the system and also the spectral properties of the main
operator. If two realizations of θ ∈ Sκ(U , Y) both have the same property
(i), (ii) or (iii) of Theorem 2.7, then they are unitarily similar [1, Theorem
2.1.3].
The realizations Σ1 and Σ2 above are said to be weakly similar if D1 =
D2 and there exists an injective closed densely defined possible unbounded
linear operator Z : X1 → X2 with the dense range such that
ZA1x = A2Zx, C1x = C2Zx,
x ∈ D(Z),
and ZB1 = B2,
(2.40)
where D(Z) is the domain of Z. In Hilbert state space case, a result of Helton
[25] and Arov [5] states that two minimal passive realizations of θ ∈ S(U , Y)
are weakly similar. However, weak similarity preserves neither dynamical
properties of the system nor the spectral properties of its main operator.
Helton's and Arov's statement holds also in case where all the spaces
are indefinite. This result is stated for reference purposes. Similar argument
as Hilbert space case can be applied, definiteness of the inner product play no
role. For a proof of special cases, see [14, Theorem 7.1.3], [29, p. 702] and [27,
Theorem 2.5]. Note that the realizations are not assumed to be κ-admissible
or passive.
Proposition 2.11. Two minimal realizations of θ ∈ Sκ(U , Y) are weakly sim-
ilar.
3. Optimal minimal systems
For κ-admissible realizations of θ ∈ Sκ(U , Y), where U and Y are Pontryagin
spaces with the same negative index, one can form the similar theory of
optimal minimal passive systems as represented in the standard Hilbert space
case in [7] and the Pontryagin state space case in [28]. Techniques, definitions
and notations to be used here are similar to what appears in those papers.
Denote EX (x) = hx, xiX for a vector x in an inner product space X .
The same notation has been used in the proof of Proposition 2.3. Following
[7, 12, 28], a passive realization Σ = (A, B, C, D; X , U , Y; κ) of θ ∈ Sκ(U , Y)
is called optimal if for any passive realization Σ′ = (A′, B′, C ′, D′; X ′, U , Y; κ)
of θ, the inequality
n ∈ N0,
uk ∈ U ,
(3.1)
EX nXk=0
AkBuk! ≤ EX ′ nXk=0
A′kB′uk! ,
Minimal realizations
15
holds. On the other hand, the system Σ is called *-optimal if it is observable
and
EX nXk=0
AkBuk! ≥ EX ′ nXk=0
A′kB′uk! ,
n ∈ N0,
uk ∈ U ,
(3.2)
holds for every observable passive realization Σ′ of θ. The requirement for
observability must be included for avoiding trivialities, since otherwise every
isometric realization of θ would be ∗-optimal; see Lemma 3.3 below and [7,
Proposition 3.5 and example on page 144].
In the definition of optimality, the requirement that the considered re-
alizations are κ-admissible is essential, as the example below shows.
Example 3.1. Let
Σ = (A, B, C, D; X , U , Y; κ), Σ′ = (A′, B′, C ′, D′; X ′, U , Y; κ′),
where κ < κ′, be passive realization of θ ∈ Sκ(U , Y). Suppose that (3.1)
holds. By Lemma 2.9, if (3.1) holds for Σ, it holds also for the controllable
restriction Σc = (Ac, Bc, Cc, D′; X c, U , Y; κ) of Σ. For any vector x of the
form
x =
MXn=0
define
An
c Bcun,
{un} ⊂ U ,
M ∈ N0,
Rx =
A′nB′un.
MXn=0
It is easy to deduce that R is a linear relation. Moreover, since Σc is control-
lable by Lemma 2.9, R is densely defined. Since (3.1) holds, R is contractive.
It follows now from [1, Theorem 1.4.2] that R can be extended to be every-
where defined contractive linear operator. Since ind−X c = κ < κ′ = ind−X ′,
it follows from [24, Theorem 2.4] that linear operator from X c to X ′ cannot
be contractive, and hence (3.1) cannot hold.
It will be shown in Theorem 3.5 below that an optimal (∗-optimal) min-
imal realization exists, and it can be constructed by taking the first (second)
minimal restriction, introduced in Definition 2.10, of simple conservative re-
alizations. More lemmas will be needed before that.
Lemma 3.2. Let Σ = (A, B, C, D; X , U , Y; κ) is a passive realization of θ ∈
Sκ(U , Y), and let Σs = (As, Bs, Cs, D; X s, U , Y; κ) be the restriction of Σ to
the simple subspace. Then, the first (second) minimal restrictions of Σ and
Σs coinside.
Proof. Only the proof of the statement conserning about the second minimal
restrictions is provided, since the other case is similar. To make notation less
cumbersome, write X s = Xp, where p refers to proper part. By Lemma 2.9,
16
L. Lilleberg
the equalities (2.34) and (2.36) hold, and it easily follows that
X o = X o
p ,
X c = X c
p
(X o)⊥ = (X s)⊥ ⊕ (X o
p )⊥,
(X c)⊥ = (X s)⊥ ⊕ (X c
p )⊥,
where orthogonal complements (X o
the space Xp. Therefore,
p )⊥ and (X c
p )⊥ are taken with respect to
PX cX o = PX c
p X o
p ⊂ X s = Xp,
and consequently,
PPX c
p
X o
p
Ap↾PX c
p
X o
p
= PPX c X oA↾X s↾PX c X o = PPX c X o A↾PX c X o,
PPX c
p
X o
p
Bp = PPX c X oB,
Cp↾PX c
p
X o
p
= C↾PX c X o ,
which shows that the second minimal restrictions of Σ and Σs co-inside. (cid:3)
To prove the (∗-)optimality of a system, the following lemma is helpful.
Lemma 3.3. Let
Σ = (A, B, C, D; X , U , Y,κ),
Σ′ = (A′, B′, C ′, D; X ′, U , Y; κ),
bΣ = (bA, bB,bC, D; bX , U , Y, κ),
be realizations of θ ∈ Sκ(U , Y) such that Σ is passive, bΣ is a passive dilation
of Σ and Σ′ is the first minimal restriction of bΣ. Then
EX ′ nXk=0
A′kB′uk! ≤ EX nXk=0
AkBuk! ,
n ∈ N0,
uk ∈ U .
(3.3)
Moreover, for any isometric realization Σ1 = (A1, B1, C1, D; X1, U , Y, κ) of
θ, it holds
n ∈ N0,
uk ∈ U .
(3.4)
1 B1uk! ,
Note that Proposition 2.3 quarantees the existence of a passive dilation
AkBuk! ≤ EX1 nXk=0
EX nXk=0
bΣ of Σ with the properties discribed above.
Proof. SincebΣ is a dilation of Σ, the system operator TbΣ has a representation
bC D! =
,
Y
TbΣ = bA bB
U
→
A11 A12 A13
A A23
0
A33
:
B1
B
0
D
D
X
D∗
D
X
D∗
(3.5)
Ak
0
0
(cid:0)0 C1 C(cid:1)
Minimal realizations
17
33
X1
X ′
X3
B′
1
B′
D
X1
X ′
X3
Y
can also be represented as
A′
11 A′
0
0
12 A′
13
A′ A′
23
A′
0
0
:
X3 are Hilbert spaces, and X ′ is a Pontryagin space with the negative index
κ. Let n ∈ N0 and {uk}n
where D and D∗ are Hilbert spaces. On the other hand, by Lemma 2.9, bΣ
TbΣ =
,
where X1 = (bX o)⊥, X ′ = PbX o bX c and X3 = bX o ∩ (bX c)⊥. The spaces X1 and
EX ′ nXk=0
U
→
1(cid:1)
(cid:0)0 C ′ C ′
k=0 ⊂ U . Since X3 ⊂ (bX c)⊥, it holds
nXk=0 bAkbBuk!
= E bX nXk=0 bAkbBuk! − EbX PX1
nXk=0 bAkbBuk!
− E bX PX3
= E bX nXk=0 bAkbBuk! − EbX PX1
nXk=0 bAkbBuk!
nXk=0 bAkbBuk! .
A′kB′uk! = E bX PX ′
With D and D∗ as in (3.5), the identities in (2.9) hold. Therefore, it follows
(3.6)
A similar calculation as above yields then
from the identities (2.1) and (2.2) that D∗ ⊂ (bX c)⊥ and D ⊂ (bX o)⊥ = X1.
EX nXk=0
nXk=0 bAkbBuk! .
Since X1 is a Hilbert space, the inclusion D ⊂ X1 implies
(3.7)
AkBuk! = E bX nXk=0 bAkbBuk! − EbX PD
EbX PD
EX ′ nXk=0
nXk=0 bAkbBuk! ≤ E bX PX1
A′kB′uk! ≤ EX nXk=0
nXk=0 bAkbBuk! .
AkBuk! ,
It follows now from the equations (3.6) and (3.7) that
and the inequality (3.3) is proved.
(3.7) that
Assume that bΣ is isometric. Since D is a Hilbert space, it follows from
EX nXk=0
AkBuk! ≤ EbX nXk=0 bAkbBuk! .
(3.8)
18
L. Lilleberg
0
U
Y
c Bc. Therefore
By Lemma 2.9, the system operator of bΣ can be represented as
(bX c)⊥
(bX c)⊥
→
bX c !
D :
bX c !
,
Bc(cid:19)
A4 Ac(cid:19) (cid:18) 0
TbΣ =(cid:18)A3
(cid:0)C1 Cc(cid:1)
where the restriction Σc = (Ac, Bc, Cc, D; bX c, U , Y, κ) is controllable isomet-
ric, and for every n = 0, 1, 2, . . . , it holds bAnbB = An
kBcuk! .
EbX nXk=0 bAkbBuk! = E bX c nXk=0
1 nXk=0
kB1uk! = EX c
EX1 nXk=0
kB0uk! = EX c
1 nXk=0
E bX c nXk=0
Similar argument show that if Σc
1 , U , Y, κ) is the restric-
tion of the isometric system Σ1 = (A1, B1, C1, D; X1, U , Y, κ) to the control-
lable subspace X c
1uk! .
1uk! .
1 is controllable isometric and it holds
1 are unitarily similar, and therefore
1 = (Ac
1, Bc
1, C c
1, D; X c
But Σ0 and Σc
A1
A0
1 , then Σc
Ac
(3.9)
Ac
1
kBc
(3.10)
Ac
1
kBc
(3.11)
(cid:3)
By combining (3.8) -- (3.11), the inequality (3.4) follows.
Remark 3.4. It follows from the inequality (3.4) of Lemma 3.3 that if there
exists an observable isometric realization of θ ∈ Sκ(U , Y), then it is ∗-optimal.
In the standard Hilbert space case, results of Arov [6] show that there
exist optimal minimal realizations of a Schur function. The construction was
based on the existence of the defect functions, see Section 4. Arov et. all
provided new geometric proofs of these results in [7]. Saprikin used those
new proofs and generalized Arov's results to Pontryagin state space case in
[28]. It will be proved next that Arov's results holds in the case when all
spaces are Pontryagin spaces. The geometric proofs in [7] can still be applied
in the precent setting with few appropriate changes.
Theorem 3.5. Let θ ∈ Sκ(U , Y), where U and Y are Pontryagin spaces with
the same negative index. Then:
(i) The first minimal restriction of a simple conservative realization of θ is
optimal minimal;
(ii) The minimal passive system Σ∗ is optimal if and only if the dual system
Σ is *-optimal minimal;
(iii) The second minimal restriction of a simple conservative realization of θ
is *-optimal minimal;
(iv) Optimal (*-optimal) minimal systems are unique up to unitary similar-
ity, and every optimal (*-optimal) minimal realization of θ is the first
minimal restriction (second minimal restriction) of some simple conser-
vative realization of θ.
Minimal realizations
19
Proof. (i) Let Σ′ = (A′, B′, C ′, D; X ′, U , Y; κ) be the first minimal restriction
n ∈ N0,
uk ∈ U .
(3.12)
AkBuk! ,
By Lemma 3.2, it can be assumed that Σ is the first minimal restriction of
A′kB′uk! ≤ EX nXk=0
Sκ(U , Y). Let Σ = (A, B, C, D; X , U , Y; κ) be the first minimal restriction of
some conservative realization of θ such that its state space has negative index
κ. To prove that Σ′ is optimal, Lemma 3.3 shows that it is enough to prove
of a simple conservative realization bΣ′ = (bA′, bB′,bC ′, D; bX ′, U , Y; κ) of θ ∈
EX ′ nXk=0
some simple conservative realization bΣ = (bA, bB, bC, D; bX , U , Y; κ) of θ. Since
bΣ and bΣ′ are both simple conservative, they are unitarily similar, so there
exists a unitary operator U : bX → bX ′ such that
bB = U −1bB′,
bA = U −1bA′U,
bC = bC ′U.
′o = U bX o, bX
′o)⊥ = U (bX o)⊥,
′c = U bX c, (bX
Easy calculations shows that bX
′c)⊥ = U (bX c)⊥ and P bX
′c = U PbX o bX c. In particular,
(bX
′o bX
PX = PPcX o bX c = U −1PP
′ o bX
cX
′c U = U −1PX ′U,
which implies
A = PX bA↾X = U −1PX ′bA′U ↾X = (U ↾X )−1PX ′bA′↾X ′U ↾X = (U ↾X )−1A′U ↾X
B = (U ↾X )−1B′,
C = C ′U ↾X .
It follows that Σ and Σ′ are unitarily similar and the corresponding unitary
operator is U0 = U ↾X . Then
EX nXk=0
AkBuk! = EX U −1
0
A′kB′uk! = EX ′ nXk=0
nXk=0
A′kB′uk! ,
Then (3.12) holds, and Σ′ is an optimal minimal system.
(ii) Let Σ∗ = (A∗, C ∗, B∗, D∗; X , Y, U; κ) be an optimal minimal passive
realization of θ# ∈ Sκ(Y, U). Then Σ = (A, B, C, D; X , U , Y; κ) is a minimal
passive realization of θ ∈ Sκ(U , Y). Consider an arbitrary observable passive
′∗ =
realization Σ′ = (A′, B′, C ′, D; X ′, U , Y; κ) of θ ∈ Sκ(U , Y). Then Σ
′∗, D∗; X ′, Y, U; κ) is a controllable passive realization of θ#. For
(A
a vector of the form
′∗, B
′∗, C
x′ =
nXk=0
′∗k
A
′∗yk,
C
n ∈ N0,
yk ∈ Y,
define
Sx′ =
nXk=0
(A∗)kC ∗yk.
20
L. Lilleberg
′∗ is controllable and Σ∗ is optimal, the domain of S is dense, and it
Since Σ
holds
EX (Sx) = EX nXk=0
(A∗)kC ∗yk! ≤ EX ′ nXk=0
∗k
′
A
′
C
∗yk! = EX ′(x).
That is, S is a contractive linear relation with the dense domain. Then [1,
Theorem 1.4.4] shows that the closure of S, which is still denoted as S, is
contractive everywhere defined linear operator from X ′ → X . Since X ′ and
X are Pontryagin spaces with the same negative index, S ∗ : X → X ′, is
contractive as well. The transfer functions of the Σ and Σ′ coincide, and
′∗ =
therefore CAmB = C ′A′kB′ for every m ∈ N0. By definition, S(A
(A∗)mC ∗, or what is the same thing, C ′A′mS ∗ = CAm, for every m ∈ N0.
Then also
′∗)mC
C ′A′m+k
B′ = CAmAkB = C ′A′m
S ∗AkB for m, k ≥ 0.
This implies A′kB′ = S ∗AkB since the system Σ′ is observable. It follows
now that
S ∗ nXk=0
AkBuk! =
B′uk! = EX ′ S ∗ nXk=0
A′k
B′uk.
nXk=0
AkBuk!! ≤ EX nXk=0
AkBuk! ,
Therefore,
EX ′ nXk=0
A′k
since S ∗ is contractive. This proves that Σ is ∗-optimal.
Suppose then that Σ = (A, B, C, D; X , U , Y; κ) is minimal passive ∗-
optimal realization of θ ∈ Sκ(U , Y). Then Σ∗ is a minimal passive realiza-
tion of θ# ∈ Sκ(Y, U). To prove the optimality of Σ∗, it suffices to con-
′∗ =
sider all the minimal passive realizations of θ#; see Lemma 3.3. Let Σ
′∗, D∗; X ′, Y, U; κ) be a minimal passive realization of θ#. Then
(A
Σ′ is a minimal passive realization of θ. Since Σ is ∗-optimal, the inequality
′∗, B
′∗, C
EX nXk=0
AkBuk! ≥ EX ′ nXk=0
A′kB′uk! ,
k=0 A′kB′uk for x =Pn
holds. Define Kx =Pn
k=0 AkBuk. Using similar tech-
niques as above, K can be extended to be a contractive operator from X → X ′
such that
n ∈ N0,
uk ∈ U ,
K ∗(A
′∗)kC
′∗ = (A∗)kC ∗.
Since K ∗ is contractive,
EX nXk=0
A∗kC ∗yk! = EX K ∗
′∗k
A
C
nXk=0
′∗yk! ≤ EX ′ nXk=0
′∗k
A
C
′∗yk! ,
for {yk} ⊂ Y. This shows that Σ∗ is optimal.
(iii) Let Σ be a simple conservative realization of θ ∈ Sκ(U , Y). Then
Σ∗ is a simple conservative realization of θ#, and the first minimal restriction
Minimal realizations
21
Σ∗′ of Σ∗ is optimal minimal by the part (i). By using the representations
(2.27) and (2.28) from Lemma 2.9, it is easy to deduce that the dual system
of Σ∗′ is the second minimal restriction Σ′′ of Σ, and it follows from the part
(ii) that Σ′′ is ∗-optimal.
(iv) Only the proofs of the claims considering optimal minimal realiza-
tions will be given, since the claims considering ∗-optimal minimal realizations
can be proved analogously. Let
Σj = (Aj , Bj, Cj, D; Xj, U , Y; κ),
j = 1, 2,
be optimal minimal realizations of θ ∈ Sκ(U , Y). In a sufficiently small neigh-
bourhood of the origin, the transfer functions θΣ1 and θΣ2 of the systems Σ1
and Σ2 have the Neumann series and they coincide, so C1Ak
2 B2
for k = 0, 1, 2, . . . Since Σ1 is controllable, vectors of the form
1 B1 = C2Ak
x =
NXk=0
Ak
1 B1uk,
uk ∈ U ,
are dense in X1. Define
U x =
Ak
2 B2uk.
(3.13)
NXk=0
Because Σ2 is controllable as well, vectors of the form U x are dense in X2.
Since Σ1 and Σ2 both are optimal realizations, EX1(x) = EX1(U x), and
therefore U is an isometric linear relation with the dense domain and the
dense range. It follows now from [1, 1.4.2] that the closure of U is a unitary
operator, which is still denoted as U. Then, trivially B1 = U −1B2. For vector
x in (3.13), it holds
U A1x = U
Ak+1
1 B1uk =
Ak+1
2 B2uk = A2U x.
NXk=0
NXk=0
It follows that U A1x = A2U x holds in a dense set, and therefore by conti-
nuity, everywhere. Thus A1 = U −1A2U. Moreover, for k = 0, 1, 2, . . . , one
concludes
C1Ak
1 B1 = C2Ak
2 B2 = C2U Ak
1B1.
Since spank∈N0 Ak
1 B1 is dense in X1, it must be C1 = C2U. It has been shown
that the unitary operator U has all the properties of (2.39), and therefore Σ1
and Σ2 are unitarily similar.
Suppose then that Σ = (A, B, C, D; X , U , Y; κ) is an optimal minimal
be represented as
realization of θ. Let bΣ0 = (bA0, bB0,bC0, D; bX0, U , Y; κ) be some simple conser-
vative realization of θ. Lemma 2.9 shows that the system operator of bΣ can
,
0
:
→
=
A′
11 A′
0
0
12 A′
13
A′ A′
23
A′
0
X1
X ′
X2
Y
X1
X ′
X2
U
B′
1
B′
D
TbΣ0
33
(cid:0)0 C ′ C ′
1(cid:1)
22
L. Lilleberg
A′
13
U −1A′
A′
11 A′
12U
0
A
0
0
follows from part (i) that Σ′ is optimal minimal, and moreover, as proved
above, unitarily similar with Σ. Therefore, there exists a unitary operator
U : X → X ′ such that A = U −1A′U, B = U −1B′ and C = C ′U. Define
where X1 = (bX o)⊥, X ′ = PbX o bX c and X2 = bX o ∩ (bX c)⊥. Then the system
Σ′ = (A′, B′, C ′, D; X ′, U , Y; κ) is the first minimal restriction of bΣ, and it
TbΣ = bA bB
bC D!
=
,
Y
:
and let bΣ be the system corresponding the system operator TbΣ. Easy calcu-
lations show that bΣ and bΣ0 are unitarily similar and
→
I :
is the corresponding unitary operator. Therefore bΣ is a simple conservative
system. Now bU maps PX oX c to PX ′o X ′c, and bU X ′ = U X ′ = X . It follows
that Σ is the first minimal restriction of bΣ.
(cid:0)0 C C ′
1(cid:1)
bU =
→
0
I
0 U 0
0
B′
1
B
0
D
X1
X
X2
U
23
A′
33
X1
X
X2
0
0
X1
X
X2
X1
X ′
X2
(cid:3)
4. Generalized defect functions
If U and Y are Hilbert spaces, it is well known that S ∈ S(U , Y) is holomor-
phic in the unit disk and it has non-tangential contractive strong limit values
almost everywhere (a.e.) on the unit circle T. Therefore, S can be extended
to L∞(U , Y) function, that is, the class of weakly measurable a.e. defined
and essentially bounded L(U , Y)-valued functions on T. Then it follows from
[30, Theorem V.4.2] that there exist a Hilbert space K and an outer function
ϕS ∈ S(U , K) such that
ϕ∗
S(ζ)ϕS (ζ) ≤ I − S ∗(ζ)S(ζ)
(4.1)
same property, then
a.e. on T, and if a function bϕ ∈ S(U ,bK), where bK is a Hilbert space, has this
(4.2)
S(ζ)ϕS (ζ)
a.e. on T. The function ϕS is called the right defect function of S. For the
notions of the outer functions, ∗-outer functions, inner functions and ∗-inner
functions, see [30, Chapter V]. From [30, Theorem V.4.2] it is also easy to
deduce that there exists a Hilbert space H and a ∗-outer function ψS ∈
S(H, Y) such that
bϕ∗(ζ)bϕ(ζ) ≤ ϕ∗
ψS(ζ)ψ∗
S(ζ) ≤ I − S(ζ)S ∗(ζ)
(4.3)
Minimal realizations
23
a.e. ζ ∈ T and if a Schur function bψ ∈ S(bH, Y) has this same property, then
ψS(ζ)ψ∗
(4.4)
The function ψS is called the left defect function of S. Both ϕS and ψS are
unique up to a unitary constant.
S(ζ) ≤ bψ(ζ)bψ∗(ζ).
The theory of the defect functions is considered, for instance, in [16,
17, 18]. Various connections of defect functions and passive realizations can
be found in [4, 9, 10]. The definition of the defect functions was generalized
for functions S ∈ Sκ(U , Y) in [27] by using the Kreın-Langer factorizations
and the fact that all functions in Sκ(U , Y) have also contractive strong limit
values a.e. on T. If U and Y are Pontryagin spaces such that their negative
index is not zero, the defect functions cannot be defined similarly as in the
Hilbert space settings, since S ∈ Sκ(U , Y) may not be extendable to the unit
circle. In the Hilbert state space case, Arov and Saprikin showed in [12] that
for a function S = SrB−1
is the right Kreın-Langer
factorization of S, the existence of the optimal minimal realization of S is
connected with the existence of the right defect function of Sr. In general,
similar connections exist with certain functions constructed by embedded
systems, and those function are called defect functions; this is the approach
taken here.
r ∈ Sκ(U , Y), where SrB−1
r
Suppose that Σ = (A, B, C, D; X , U , Y; κ) is a passive realization of
θ ∈ Sκ(U , Y). Denote the system operator of Σ by T. Theorem 2.1 shows that
there exists a Julia operator of T. By using the same notation as in (2.13),
A
one can form the Julia embeddingeΣ of the system Σ. Then the corresponding
system operator eT is a Julia operator of T, and it is of the form
(cid:0)B DT ∗
,1(cid:1)
TeΣ =
T,2 −L∗(cid:19) :
DT(cid:19) ,
X(cid:18) Y
(cid:18) C
T,1(cid:19) (cid:18) D DT ∗
DT ∗ =(cid:18)DT ∗
,2(cid:19), DT =(cid:18)DT,1
DT,2(cid:19), DT ∗ D∗
DT ∗(cid:19) →
X(cid:18) U
T ∗ = IX −T T ∗, DT D∗
T = IX −T ∗T,
where
(4.5)
DT ∗
D∗
D∗
,2
,1
such that DT and DT ∗ have zero kernels. The transfer function of the Julia
embedding is
θeΣ(z) =(cid:18) D + zC(I − zA)−1B
T,2 + zD∗
D∗
T,1(I − zA)−1B −L∗ + zD∗
DT ∗
,2 + zC(I − zA)−1DT ∗
T,1(I − zA)−1DT ∗
,1
,1(cid:19)
(4.6)
(4.7)
(4.8)
=(cid:18)θ(z) ψ(z)
ϕ(z) χ(z)(cid:19) .
Lemma 2.4 shows that the identities
I − θ(z)θ∗(w) = (1 − z ¯w)G(z)G∗(w) + ψ(z)ψ∗(w),
I − θ∗(w)θ(z) = (1 − z ¯w)F ∗(w)F (z) + ϕ∗(w)ϕ(z)
where
G(z) = C(IX − zA)−1, F (z) = (IX − zA)−1B,
24
L. Lilleberg
holds for every z and w in a sufficiently small symmetric neighbourhood
of the origin. If the negative index of U and Y is denoted as κ1, a similar
argument as in the proof of Proposition 2.5 shows that the kernels Kϕ(w, z)
and Kψ(w, z) defined as in (1.3) have at most κ + κ1 negative squares.
Definition 4.1. Let U and Y be Pontryagin spaces with the same negative
index. Let Σ = (A, B, C, D; X , U , Y; κ) be an optimal minimal passive real-
as in (4.5). Then the function ϕ in (4.6) is defined to be the right defect
function ϕθ of θ.
ization of θ ∈ Sκ(U , Y), and let eΣ be the Julia embedding of it, represented
sive realization of θ ∈ Sκ(U , Y), and let eΣ be the Julia embedding of it,
represented as in (4.5). Then the function ψ in (4.6) is defined to be the left
defect function ψθ of θ.
Moreover, let Σ = (A, B, C, D; X , U , Y; κ) be a ∗-optimal minimal pas-
Remark 4.2. Since optimal (∗-optimal) minimal realizations are unitarily sim-
ilar by Theorem 3.5, and Julia operators for contractive operator are essen-
tially unique by Theorem 2.1, it can be deduced that the defect functions
are essentially uniquely defined by θ ∈ Sκ(U , Y). The definition above is also
slightly different from the one given in [27] for functions in the class Sκ(U , Y),
where U and Y are Hilbert spaces.
The right defect function of θ ∈ Sκ(U , Y) and the left defect function of
θ# are closely related to each other.
Lemma 4.3. For θ ∈ Sκ(U , Y), it holds ϕ#
θ = ψθ# and ψ#
θ = ϕθ#
Proof. Let Σ = (A, B, C, D; X , U , Y; κ) be an optimal (∗-optimal) minimal
realization of θ. Denote the system operator of Σ as T, and the Julia operator
TeΣ of T as in (4.5). By Theorem 3.5, the system Σ∗ is ∗-optimal (optimal)
minimal, and a calculation shows that T ∗
eΣ
the results follow means of (4.6).
is the Julia operator of T ∗. Now
(cid:3)
In the Hilbert space settings, S ∈ S(U , Y) has factorizations of the form
S = SiSo = S∗oS∗i,
where Si ∈ S(Y, Y) is inner, So ∈ S(U , Y) is outer, S∗o ∈ S(U , Y) is ∗-outer
and S∗i ∈ S(Y, Y) is ∗-inner [30, p. 204]. The next proposition shows that
for an ordinary Schur function θ ∈ S(U , Y), the outer factor of ϕθ and the ∗-
outer factor of ψθ defined above coincide essentially with the usual definition
of defect functions.
Proposition 4.4. Let θ ∈ Sκ(U , Y), where U and Y are Hilbert spaces. Then
ϕ∗
θ(ζ)ϕθ(ζ) ≤ I − θ∗(ζ)θ(ζ)
(4.9)
Hilbert space and κ′ does not depend on κ, has this same property, then
a.e. on T, and if a generalized Schur function bϕ ∈ Sκ′ (U ,bK), where bK is a
(4.10)
θ(ζ)ϕθ(ζ),
bϕ∗(ζ)bϕ(ζ) ≤ ϕ∗
Minimal realizations
25
a.e. on T. If κ = 0, denote the inner and outer factors of ϕθ as ϕθi and
ϕθo , respectively. Then, ϕθi is an isometric constant, and if ϕ′ is an outer
function with properties (4.1) and (4.2), then it holds U ϕθo = ϕ′, where U is
a unitary operator.
Moreover,
ψθ(ζ)ψ∗
θ (ζ) ≤ I − θ(ζ)θ∗(ζ)
a.e. ζ ∈ T and if a generalized Schur function bψ ∈ Sκ′ (bH, Y), where bK is a
Hilbert space and κ′ does not depend on κ, has this same property, then
ψθ(ζ)ψ∗
θ (ζ) ≤ bψ(ζ)bψ∗(ζ)
a.e. ζ ∈ T.If κ = 0, denote the ∗-inner and ∗-outer factors of ψθ as ψθ∗i
and ψθ∗o , respectively. Then, ψθ∗i is a co-isometric constant, and if ψ′ is a
∗-outer function with properties (4.3) and (4.4), then it holds ψθ∗ o U ′ = ψ′,
where U ′ is a unitary operator.
Proof. Let
(4.11)
be an optimal minimal realization of θ. Denote the system operator of Σ as
Σ = (A, B, C, D; X , U , Y; κ)
A
D∗
B
D∗
T, the Julia operator TeΣ of T as in (4.5) and the function ϕ = ϕθ as in (4.6).
Since TeΣ is unitary, the operator
TΣ′ =
(cid:18) C
T,1(cid:19) (cid:18) D
Σ′ =(cid:18)A, B,(cid:18) C
T,2(cid:19) :(cid:18)X
DT(cid:19) .
U(cid:19) →
X(cid:18) Y
T,1(cid:19) ,(cid:18) D
DT(cid:19) ; κ(cid:19)
T,2(cid:19) ; X , U ,(cid:18) Y
must be isometric, and therefore the system
(4.12)
D∗
D∗
for a.e. ζ ∈ T.
θ(ζ)(cid:1)(cid:18) θ(ζ)
ϕθ(cid:19) . Since Σ′ is an embedding
is an isometric realization of the function (cid:18) θ
of the minimal system Σ, the system Σ′ is also minimal. It follows from
Theorem 2.7 that (cid:18) θ
ϕθ(cid:19) ∈ Sκ (U , Y ⊕ DT ) . Since contractive boundary values
of generalized Schur functions exist for a.e. ζ ∈ T, it holds
ϕθ(ζ)(cid:19) ≤ I ⇐⇒ ϕ∗
(cid:0)θ∗(ζ) ϕ∗
Suppose that a function bϕ ∈ Sκ′(U ,bK), where bK is a Hilbert space, has
for a.e. ζ ∈ T. Since the function bϕ has the left Kreın -- Langer factorization
of the form bϕ = B−1
bϕ bϕl, where bϕl is an ordinary Schur function, it holds
θ(ζ)ϕθ(ζ) ≤ I − θ∗(ζ)θ(ζ)
the property
for a.e. ζ ∈ T. Then the function
(4.13)
bϕ∗(ζ)bϕ(ζ) ≤ I − θ∗(ζ)θ(ζ)
bϕ∗(ζ)bϕ(ζ) = bϕ∗
l (ζ)bϕl(ζ)
bϕl(cid:19) ,
θ =(cid:18) θ
26
L. Lilleberg
belongs to the Schur class Sκ(cid:16)U , Y ⊕ bK(cid:17) , and it has a controllable isometric
realization Σ with the system operator
That is,
A1
B1
D2(cid:19) :(cid:18)X1
T Σ =
(cid:18)C1
C2(cid:19) (cid:18)D1
D2(cid:19) + z(cid:18)C1
θ(z) =(cid:18) θ(z)
bϕl(z)(cid:19) =(cid:18)D1
=(cid:18)D1 + zC1(I − zA1)−1B1
D2 + zC2(I − zA1)−1B1(cid:19) .
U(cid:19) →
bK(cid:19) .
X1(cid:18)Y
C2(cid:19) (I − zA1)−1B1
It follows that
Σ1 = (A1, B1, C1, D1; X1, U , Y; κ)
(4.14)
is a realization of θ, and since Σ is isometric and bK is a Hilbert space, the
system Σ1 is passive. Since T Σ is isometric, the defect operator DT Σ of T Σ is
zero, and it follows from Lemma 2.4 that
1)−1)(I − zA1)−1B1
(4.15)
whenever the expressions are meaningful. By combining the identities (4.8)
and (4.15) for optimal minimal realization Σ, one gets
1 (I − zA∗
l (z)bϕl(z)
I − θ∗(z)θ(z) = I − θ∗(z)θ(z) −bϕ∗
(cid:0)1 − z2(cid:1)B∗
=(cid:0)1 − z2(cid:1) B∗
1)−1(I − zA1)−1B1 +bϕ∗
l (z)bϕl(z)
=(cid:0)1 − z2(cid:1) B∗(I − zA∗)−1(I − zA)−1B + ϕθ
1 (I − zA∗
for every z in a sufficiently small symmetric neighbourhood Ω of the origin.
Since the system Σ is optimal, if follows by using Neumann series that
(cid:10)B∗(I − zA∗)−1(I − zA)−1Bu, u(cid:11) = EX(cid:0)(I − zA)−1Bu(cid:1) = EX ∞Xn=0
1 B1uzn!
1)−1(I − zA1)−1B1u, u(cid:11)
for every z ∈ Ω and for every u ∈ U . Then it follows from (4.16) that
1 (I − zA∗
An
AnBuzn!
∗(z)ϕθ(z)
(4.16)
≤ EX1 ∞Xn=0
=(cid:10)B∗
bϕ∗
l (z)bϕl(z) ≤ ϕθ
l (ζ)bϕl(ζ) = bϕ∗(ζ)bϕ(ζ) ≤ ϕθ
bϕ∗
∗(z)ϕθ(z),
z ∈ Ω.
∗(ζ)ϕθ(ζ)
(4.17)
By continuity,
for a.e. ζ ∈ T.
Next suppose that κ = 0. By combining (4.2) and (4.17), it can be
deduced that
′∗(ζ)ϕ′(ζ) = ϕθ
ϕ
∗(ζ)ϕθ(ζ) = ϕθo
∗(ζ)ϕθi
∗(ζ)ϕθi (ζ)ϕθo (ζ) = ϕθo
∗(ζ)ϕθo (ζ)
Minimal realizations
27
for a.e. ζ ∈ T. Then it follows from [30, Proposition V.4.1] that ϕ′ = U ϕθo ,
where U is a unitary operator. If one puts an outer function bϕl = ϕθo = U −1ϕ′
in (4.13), the construction of an optimal minimal system used in the proof
of [6, Theorem 7] shows that the associated system Σ1 in (4.14) is optimal.
Since Σ in (4.11) is also optimal, for every z ∈ D, it holds
B∗(I − zA∗)−1(I − zA)−1B = B∗
1 (I − zA∗
1)−1(I − zA1)−1B1,
and then by (4.16)
kϕθi(z)ϕθo(z)uk = kϕθo(z)uk
for every z ∈ D and every u ∈ U . The outer function ϕθo (z) has a dense
range for every z ∈ D [30, Proposition V.2.4]. This implies that ϕθi (z) is an
isometry for every z ∈ D, and arguing as in the proof of [30, Proposition
V.2.1] one deduces that ϕθi is an isometric constant. The claims involving ϕθ
are proved.
The claims involving ψθ follow now directly by applying Lemma 4.3. (cid:3)
Lemma 4.5. Let
Σ0 = (A0, B0, C0, D; X0, U , Y; κ), Σ = (A, B, C, D; X , U , Y; κ)
be passive realizations of θ ∈ Sκ(U , Y) such that Σ0 is optimal. If for every
z and w in a sufficiently small symmetric neighbourhood Ω of the origin the
equality
B∗(I − wA∗)−1(I − zA)−1B = B∗
0 (I − wA∗
0)−1(I − zA0)−1B0
(4.18)
holds, then Σ is optimal.
Proof. It follows from Lemma 2.9 that the system operator TΣ of Σ can be
represented as in (2.25), the restriction Σc = (Ac, Bc, Cc, D; X c, U , Y; κ) of Σ
to the controllable subspace X c is controllable passive, and (2.34) and (2.35)
hold.
Let
x =
MXj=1
Aj
cBcuj, M ∈ N,
{uj}M
j=1 ⊂ U .
(4.19)
For the vectors of the form (4.19), define
Rx =
MXj=1
Aj
0B0uj.
(4.20)
Since Σc is controllable, the domain of R is dense. Moreover, Σ0 is optimal,
and therefore EX0 (Rx) ≤ EX c (x) . That is, R is contractive, and it follows
from [1, Theorem 1.4.2] that the closure of R is everywhere defined contractive
linear operator. It is still denoted by R. Since
(I − zAc)−1Bc =
∞Xn=0
znAn
c Bc,
(I − zA0)−1B0 =
znAn
0 B0,
∞Xn=0
28
L. Lilleberg
holds for every z in a sufficiently small symmetric neighbourhood Ω of the
origin, it follows by continuity that
for every z ∈ Ω and u ∈ U . Then
for all M ∈ N, {zj}M
imply now
(IX0 − zjA0)−1B0uj,
MXj=1
j=1 ⊂ U . Equalities (2.35) and (4.18)
j=1 ⊂ Ω, and {uj}M
R(cid:0)(I − zAc)−1Bcu(cid:1) = (I − zA0)−1B0u
(I − zjAc)−1Bcuj =
R
MXj=1
(I − zjAc)−1Bcuj
EX c
MXj=1
MXj=1
MXk=1(cid:10)B∗
MXk=1(cid:10)B∗
MXj=1
(I − zjA0)−1B0uj
= EX0
MXj=1
(I − zjAc)−1Bcuj .
= EX0R
MXj=1
0 (I − zkA∗
c (I − zkA∗
=
=
c )−1(I − zjAc)−1Bcuj, uk(cid:11)U
0)−1(I − zjA0)−1B0uj, uk(cid:11)U
.
This implies that R is isometric in span{ran (I − zA1)−1B1, z ∈ Ω}, which is
a dense set, since Σ1 is controllable. Since R is bounded, it is now isometric
everywhere, and identity (4.20) implies that Σc is optimal. Then it follows
from (2.34) that Σ is optimal, and the proof is complete.
(cid:3)
The main results of [4, Theorem 1.1] were generalized to the Pontryagin
state space setting in [27, Theorem 4.4]. By using Definition 4.1, it can be
shown that parts of this result, as well as [10, Theorem 1], hold also in the
case when all the spaces are indefinite. Moreover, certain parts of [4, Theorem
1.1], [10, Theorem 1] and [27, Theorem 4.4] can be improved. Before stating
these results, some lemmas are needed.
Lemma 4.6. Let θ ∈ Sκ(U , Y). Then the following statements are equivalent:
(i) all κ-admissible minimal passive realizations of θ are unitarily similar;
(ii) there exists a minimal passive realization of θ such that it is both optimal
and ∗-optimal;
(iii) all κ-admissible minimal passive realizations of θ are both optimal and
∗-optimal.
Minimal realizations
29
Proof. (i) ⇒ (iii). Suppose (i). Let
Σ1 = (A1, B1, C1, D; X1, U , Y; κ), Σ2 = (A2, B2, C2, D; X2, U , Y; κ)
be, respectively, minimal passive and optimal (∗-optimal) minimal passive
realizations of θ. Let U be the unitary operator from X1 to X2 with the
properties described in (2.39). An easy calculation shows that
EX2 nXk=0
2 B2uk! = EX1 U
Ak
Ak
1 B1uk! = EX1 nXk=0
nXk=0
1 B1uk!
Ak
for every u ∈ U and for every n = 0, 1, 2, . . . which implies that Σ1 is actually
optimal (∗-optimal), and therefore (iii) holds.
(iii) ⇒ (ii). The claim (iii) trivially implies (ii).
(ii) ⇒ (i). Suppose (ii). Let the systems
Σ1 = (A1, B1, C1, D; X1, U , Y; κ), Σ2 = (A2, B2, C2, D; X2, U , Y; κ)
be minimal passive realizations of θ such that Σ1 is optimal and ∗-optimal. Let
Z be the weak similarity mapping from X1 to X2 with the properties described
1 B1uk
2 B2uk. Recall
also here the construction of Z in the proof of [27, Theorem 2.5]. Since Σ1 is
both optimal and ∗-optimal,
in (2.40). It follows from (2.40) that all elements of the formPn
1 B1uk(cid:1) =Pn
belongs to the domain of Z, and Z(cid:0)Pn
1 B1uk! = EX1 nXk=0
EX2 nXk=0
nXk=0
2 B2uk! = EX2 Z
1 B1uk! .
k=0 Ak
k=0 Ak
k=0 Ak
Ak
Ak
Ak
Then it follows from [1, Theorem 1.4.2] that the operator Z has a unitary
extension, and the properties in (2.39) follow by continuity. Therefore Σ1
and Σ2 are unitarily similar. Since unitary similarity clearly is a transitive
property, (i) holds, and the proof is complete.
(cid:3)
Lemma 4.7. If the system Σ = (A, B, C, D; X , U , Y; κ) is an optimal passive
realization of θ ∈ Sκ(U , Y), then X c ⊂ X o.
Proof. According to Proposition 2.8, the spaces X o and (X o)⊥ are regular
subspaces and (X o)⊥ is a Hilbert space. It follows from Lemma 2.9 that
the system operator T of Σ can be represented as in (2.24), and the restric-
tion Σo = (Ao, Bo, Co, D; X o, U , Y; κ) of Σ to the observable subspace X o is
observable passive realization of θ. For n = 0, 1, 2, . . ., it holds
An =(cid:18)An
1
0
f (n)
An
0 (cid:19) ,
!=
P(X o)⊥(cid:16)PN
PX o(cid:16)PN
n=0 AnBun(cid:17) .
n=0 AnBun(cid:17)
where f (n) is an operator depending on n. Then for any N ∈ N0 and any
{un}N
n=0 ⊂ U , it holds
AnBun = PN
NXn=0
n=0 (An
1 B1un + f (n)Boun)
n=0 An
o Boun
PN
But since Σ is optimal and (X o)⊥ is a Hilbert space, one deduces that
AnBun! =E(X o)⊥ P(X o)⊥ NXn=0
P(X o)⊥ NXn=0
AnBun!! + EX o NXn=0
AnBun! = 0.
o Boun! .
An
30
L. Lilleberg
This implies
EX NXn=0
That is, span{AnB : n = 0, 1, . . .} ⊂ X o and since X o is closed, also
span{AnB : n = 0, 1, . . .} = X c ⊂ X o.
(cid:3)
The next Theorem contains promised extensions for some results of [4].
In particular, the fact that statements (I)(b), (II)(b) and (III)(b) implies the
other statements, respectively, in parts (I), (II) and (III), is new also in the
Hilbert space setting.
Theorem 4.8. Let θ ∈ Sκ(U , Y), where U and Y are Pontryagin spaces with
the same negative index.
(I) The following statements are equivalent:
(a) ϕθ ≡ 0;
(b) all κ-admissible controllable passive realizations of θ are minimal
isometric;
(c) there exists an observable conservative realization of θ;
(d) all simple conservative realization of θ are observable;
(e) all observable co-isometric realizations of θ are conservative.
(II) The following statements are equivalent:
(a) ψθ ≡ 0;
(b) all κ-admissible observable passive realization of θ are minimal co-
isometric;
(c) there exists a controllable conservative realization of θ;
(d) all simple conservative realization of θ are controllable;
(e) all controllable isometric realizations of θ are conservative.
(III) The following statements are equivalent:
(a) ϕθ ≡ 0 and ψθ ≡ 0;
(b) all κ-admissible simple passive realization of θ are minimal con-
servative;
(d) there exists a minimal conservative realization of θ.
Proof. (I) (a) ⇒ (b). Suppose (a). Let the systems
Σ = (A, B, C, D; X , U , Y; κ), Σ0 = (A0, B0, C0, D; X0, U , Y; κ)
be, respectively, a controllable passive and an optimal minimal passive real-
izations of θ. Represent the Julia embeddings of Σ and Σ0 as in (4.5). Then,
(4.8) holds for Σ. Since ϕθ ≡ 0, if follows from the definition of ϕθ that
I − θ∗(w)θ(z) = (1 − z ¯w)B∗
0 (I − ¯wA∗
0)−1(I − zA0)−1B0
Minimal realizations
31
holds for every z and w in a sufficiently small symmetric neighbourhood
Ω of the origin. Since Σ0 is optimal, by considering the Neuman series of
(I − zA0)−1B0 and (I − zA0)−1B0, one deduces that
B∗
0 (I − ¯zA∗
0)−1(I − zA0)−1B0 ≤ B∗(I − ¯zA∗)−1(I − zA)−1B,
z ∈ Ω.
Then it holds ϕ∗(z)ϕ(z) ≤= 0 for every z ∈ Ω. But since ϕ(z) is an operator
whose range belongs to the Hilbert space DT , this implies
ϕ(z) = D∗
T,2 + zD∗
T,1(I − zA)−1B = 0,
z ∈ Ω.
T,2 = 0. Since Σ is controllable, span{(I − zA)−1B; z ∈ Ω}
It follows that D∗
is dense in X , and therefore also D∗
T,1 = 0. Then DT = 0, so T is isometric,
and Σ is a controllable isometric system. In particular, if Σ is chosen to be
minimal passive; for the existence, see Lemma 2.9, the previous argument
shows that Σ is a minimal isometric realization of θ. Since all controllable
isometric realizations of θ are unitarily similar, they are now also minimal,
and (b) holds.
(b) ⇒ (c). Suppose (b). Let Σ′ = (A′, B′, C ′, D; X ′, U , Y; κ) be an opti-
mal minimal passive realization of θ. The existence of Σ′ follows from The-
orem 3.5 (i). By assumption, Σ′ is isometric. It follows from Theorem 3.5
(iv) that Σ′ is the first minimal restriction of the simple conservative system
Σ = (A, B, C, D; X , U , Y; κ). By Lemma 2.9, the system operator TΣ of Σ
can be represented as in (2.27), where now X ′ = PX oX c. Since the system
operator
TΣ′ =(cid:18)A′ B′
C ′ D(cid:19) :(cid:18)X ′
U(cid:19) →(cid:18)X ′
Y(cid:19) ,
of Σ′ is isometric and TΣ is unitary, an easy calculation using the fact that
the range space (X o)⊥ is a Hilbert space shows that B′
12 = 0 in
(2.27). But then for every x ∈ (X o)⊥ and every n = 0, 1, 2, . . . ,
1 = 0 and A′
That is, (X o)⊥ ⊂ (X c)⊥ and therefore X c ⊂ X o. Since Σ is simple, this
implies now X o = X . Then Σ is observable, and (c) holds.
(c) ⇒ (a). Suppose (c). Let Σ = (A, B, C, D; X , U , Y; κ) be an observable
conservative realization of θ. By Lemma 2.9, Σ can be represented as in (2.27).
The first minimal restriction (2.32) of Σ is an optimal minimal realization of
θ by Theorem 3.5 (i). But since Σ is observable, X o = X and (X o)⊥ = {0}.
It follows that the reprentations (2.25) and (2.27) coinsides. That is, the first
minimal restriction Σ′ is just a restriction to the controllable subspace of Σ.
By Lemma 2.9, Σ′ is now isometric. Thus if one constructs a Julia operator
of TΣ′ as in (2.13), DTΣ′ = 0, and then it follows from the definition of ϕθ
and (4.6) that ϕθ ≡ 0, and (a) holds.
The equivalences of the statements (c), (d) and (e) follow easily from the
facts that all observable co-isometric realizations of θ are unitarily similar, all
simple conservative realization of θ are unitarily similar and unitary similarity
B∗A∗nx =(cid:0)0 B′∗
0(cid:1)
∗
A′
11
0
A′
13
∗
0
A′
0
∗ A′
23
0
0
∗ A′
33
∗
n
x
0
0 = 0.
32
L. Lilleberg
preserves the structural properties of the system and system operator. The
part (I) is proven.
(II) The proof is analogous to the proof of the part (I), and the details
are omitted.
(III) (a) ⇒ (b). Suppose (a). By combining the parts (I) and (II), it fol-
lows that all controllable or observable passive realizations of θ are minimal
conservative. Consider a simple passive realization Σ = (A.B, C, D; X , U , Y; κ)
of θ. It follows from Lemma 2.9 that the contractive system operator T of
Σ can be represented as in (2.24), where the restriction Σo in (2.29) is ob-
servable passive, and therefore now minimal conservative. Then the system
operator
of Σo is unitary. Let x ∈ X o. Then, by contractivity of T and unitarity of
TΣo
Co D(cid:19) :(cid:18)X o
TΣo =(cid:18)Ao Bo
U (cid:19) →(cid:18)X o
Y(cid:19)
0 = E
Cox = E (A2x) + E(cid:18)(cid:18)A0x
Cox(cid:19)(cid:19)
= E (T x) ≤ E(x) = E(TΣo x) = E(cid:18)(cid:18)A0x
Cox(cid:19)(cid:19) .
A2x
Aox
0
x
E
A1 A2 B1
0 Ao Bo
0 Co D
Since A2x ∈ (X o)⊥ and (X o)⊥ is a Hilbert space, it follows that A2 = 0. If
one chooses u ∈ U , a similar argument as above shows that B1 = 0. Then for
any n ∈ N, it holds
0
AnB =(cid:18)A1
A∗nC ∗ =(cid:18)A∗
0 Ao(cid:19)n(cid:18) 0
o(cid:19)n(cid:18) 0
0
1
0 A∗
Bo(cid:19) =(cid:18) 0
o(cid:19) =(cid:18) 0
o Bo(cid:19) ,
o(cid:19) .
An
A∗n
o C ∗
C ∗
This is only possible if (X o)⊥ = 0, since Σ is simple. But then the systems
Σ0 and Σ coincide, so the system Σ is minimal conservative, and (b) holds.
Now (b) trivially implies (c), and the fact that (c) implies (a) follows
(cid:3)
by combining the parts (I) and (II). The proof is complete.
Remark 4.9. If U and Y are Hilbert spaces, it follows from [27, Lemma 3.2]
that simple passive realizations of θ ∈ Sκ(U , Y) are κ-admissible. There-
fore, in that case it is not necessary to assume the considered systems to
be κ-admissible in Lemma 4.6 and Theorems 4.8 and 4.10, since the other
assumptions already guarantee it. However, if U and Y are Pontryagin spaces
with the same negative index, it is not known that are all simple passive, or
even all minimal passive, realizations of θ ∈ Sκ(U , Y) κ-admissible.
If ϕθ ≡ 0 (ψθ ≡ 0), then Theorem 4.8 shows that all κ-admissible mini-
mal passive realizations of θ ∈ Sκ(U , Y) are minimal isometric (co-isometric).
In particular, they are controllable isometric (observable coisometric), and it
follows from Theorem 2.7 that they are unitarily similar. This situation can
Minimal realizations
33
occur also when the defect functions do not vanish identically. In what follows,
the range of ϕθ and the domain of ψθ will be denoted, respectively, by Dϕθ
and Dψθ . In the Hilbert space settings, it is well known [17, 18] that for a stan-
dard Schur function θ ∈ S(U , Y), there exists a function χθ ∈ L∞(Dψθ , Dϕθ )
such that the function
Θ(ζ) :=(cid:18) θ(ζ)
ϕθ(ζ) χθ(ζ)(cid:19)
ψθ(ζ)
(4.21)
has contractive values for a.e. ζ ∈ T. Under certain normalizing conditions
for the functions ϕθ and ψθ, the function χθ is unique. In general, χθ may has
negative Fourier coefficients and therefore it is not a Schur function. In that
case the function Θ in (4.21) is not a Schur function either. However, Arov
and Nudelmann showed in [9, 10] that Θ is a Schur function if and only if
all minimal passive realizations of θ are unitarily similar. This result will be
generalized to the indefinite settings in the following theorem. The proof uses
optimal and ∗-optimal realizations as in [9, 10], but it is more elementary.
Theorem 4.10. Let θ ∈ Sκ(U , Y), where U and Y are Pontryagin spaces with
the same negative index, and let ϕθ and ψθ be defect functions of θ. Then all
κ-admissible minimal passive realizations of θ are unitarily similar if and only
if there exist an L(Dψθ , Dϕθ )-valued function χθ analytic in a neighbourhood
of the origin such that
Θ =(cid:18) θ
ϕθ χθ(cid:19) ∈ Sκ(cid:18)(cid:18) U
Dψθ(cid:19) ,(cid:18) Y
Dϕθ(cid:19)(cid:19)
ψθ
(4.22)
Proof. Suppose that all κ-admissible minimal passive realizations of θ ∈
Sκ(U , Y) are unitarily similar. Then it follows from Lemma 4.6 that every
κ-admissible minimal passive realization is optimal and ∗-optimal. Take any
κ-admissible minimal passive realization Σ of θ and consider its Julia em-
bedding as in (4.5). Then the transfer function (4.6) of the Julia embedding
belongs to the class Sκ (U ⊕ DT ∗ , Y ⊕ DT ) , and since Σ is both optimal and
∗-optimal, the upper right corner and lower left corner of (4.6) are defect
functions of θ. Choose χθ = χ in (4.6), and the necessity is proven.
Suppose then that there exists an L(Dψθ , Dϕθ )-valued function χθ such
that Θ in (4.22) belongs to the class Sκ (U ⊕ Dψθ , Y ⊕ Dϕθ ) . It suffices to
show that there exists minimal passive realization Σ of θ such that it is both
optimal and ∗-optimal; see Lemma 4.6. Let
be a simple conservative realization of Θ ∈ Sκ (U ⊕ Dψθ , Y ⊕ Dϕθ ) . Then the
system operator TΘ of ΣΘ can be represented as
ΣΘ = (A, eB,eC, eD; X , U ⊕ Dψθ , Y ⊕ Dϕθ ; κ)
Dψθ(cid:19) →
Dϕθ(cid:19) .
X(cid:18) U
X(cid:18) Y
D21 D22(cid:19) :
(cid:0)B B1(cid:1)
TΘ =
C1(cid:19) (cid:18) D D12
(cid:18) C
A
34
L. Lilleberg
In a sufficiently small symmetric neighbourhood Ω of the origin, it holds
Θ(z) =(cid:18) θ(z) ψθ(z)
ϕθ(z) χθ(z)(cid:19)
D21 + zC1 + (I − zA)−1B D22 + zC1(I − zA)−1B1(cid:19) .
=(cid:18) D + zC(I − zA)−1B
D12 + zC(I − zA)−1B1
The spaces Dϕθ and Dψθ are Hilbert spaces, and therefore it follows that Σ =
(A, B, C, D; X , U , Y; κ) is a passive realization of θ. Since ΣΘ is conservative,
Lemma 2.4 shows that
I − Θ(z)Θ∗(w)
I − Θ∗(w)Θ(z)
B∗
−θ(z)ϕ∗
IDϕθ
− ϕθ(z)ϕ∗
θ (w)
θ (w)
θ(w)(cid:19)
θ(w) − ψθ(z)χ∗
θ(w)
θ(w) − χθ(z)χ∗
−ϕθ(z)θ∗(w) − χθ(z)ψ∗
C(I − zA)−1(I − ¯wA∗)−1C ∗
1
C1(I − zA)−1(I − ¯wA∗)−1C ∗ C1(I − zA)−1(I − ¯wA∗)−1C ∗
=(cid:18)IY − θ(z)θ∗(w) − ψθ(z)ψ∗
= (1 − ¯wz)eC(I − zA)−1(I − ¯wA∗)−1eC ∗
= (1 − ¯wz)(cid:18) C(I − zA)−1(I − ¯wA∗)−1C ∗
θ(w)χθ(z)(cid:19)
=(cid:18)IU − θ∗(w)θ(z) − ϕ∗
1 (I − ¯wA∗)−1(I − zA)−1B1(cid:19) .
= (1 − ¯wz)(cid:18)B∗(I − ¯wA∗)−1(I − zA)−1B B∗(I − ¯wA∗)−1(I − zA)−1B1
1 (I − ¯wA∗)−1(I − zA)−1B B∗
− ψ∗
θ (w)ψθ(z) − χ∗
−ψ∗
θ (w)θ(z) − χ∗
θ(w)ϕθ(z)
−θ∗(w)ψθ(z) − ϕ∗
θ(w)χθ(z)
θ(w)ϕθ(z)
IDψθ
1(cid:19)
That is,
IY − θ(z)θ∗(w) = (1 − ¯wz)C(I − zA)−1(I − ¯wA∗)−1C ∗ + ψθ(z)ψ∗
θ (w),
(4.23)
IU − θ∗(w)θ(z) = (1 − ¯wz)B∗(I − ¯wA∗)−1(I − zA)−1B + ϕ∗
θ(w)ϕθ(z).
(4.24)
An easy calculation and Lemma 4.3 show that the equation (4.23) is equiva-
lent to
IY − θ#∗
(w)θ#(z) = (1 − ¯wz)C(I − ¯wA)−1(I − zA∗)−1C ∗ + ϕθ#
∗(w)ϕθ# (z).
(4.25)
Let
Σ′ = (A′, B′, C ′, D; X ′, U , Y; κ),
Σ′′ = (A′′, B′′, C ′′, D; X ′′, U , Y; κ)
be, respectively, an optimal minimal and a ∗-optimal minimal realizations of
θ. It follows from Theorem 3.5 (ii) that Σ′′∗ = (A′′ ∗, C ′′ ∗, B′′∗, D; X ′′, U , Y; κ)
is an optimal minimal realization of θ#. Then, by the definition of ϕθ and
ϕθ#, it holds
IU − θ∗(w)θ(z) = (1 − wz)B′∗(I − ¯wA′∗)−1(I − zA′)−1B′ + ϕ∗
θ (w)ϕθ(z)
IY − θ# ∗
(w)θ#(z) = (1 − wz)C ′′(I − ¯wA′′)−1(I − zA′′∗)−1C ′′∗ + ϕθ#
∗(w)ϕθ# (z).
Minimal realizations
35
It follows that
B∗(I − ¯wA∗)−1(I − zA)−1B = B′∗(I − ¯wA′∗)−1(I − zA′)−1B′,
C(I − ¯wA)−1(I − zA∗)−1C ∗ = C ′′(I − ¯wA′′)−1(I − zA′′∗)−1C ′′∗.
By using Lemma 4.5, it can be deduced that Σ and Σ∗ are optimal systems.
Then it follows from Lemma 4.7 that X c = X o and therefore X s = X c = X o.
By Lemma 2.9, the restriction Σs = (As, Bs, Cs, D; X s, U , Y; κ) of Σ to the
s Bs and A∗nC ∗ =
simple subspace X s is simple, and it holds AnB = An
A∗
s also are optimal systems. More-
s
over, they are minimal since X s = X c = X o. It follows now from Theorem
3.5 (ii) that Σs is also ∗-optimal, and the proof is complete.
(cid:3)
s for every n ∈ N0. That is, Σs and Σ∗
nC ∗
I wish to thank Seppo Hassi for helpful discussions while
Acknowledgements
preparing this paper.
References
[1] D. Alpay, A. Dijksma, J. Rovnyak, and H. S. V. de Snoo, Schur functions,
operator colligations, and Pontryagin spaces, Oper. Theory Adv. Appl., 96,
Birkhauser Verlag, Basel-Boston, 1997.
[2] D. Alpay, T.Y. Azizov, A. Dijksma and J. Rovnyak, Colligations in Pontrya-
gin Spaces with a Symmetric Characteristic Function, Linear Operators and
Matrices, Oper. Theory Adv. Appl., 130, Birkhauser, Basel, 2002.
[3] T. Ando, De Branges spaces and analytic operator functions, Division of
Applied Mathematics, Research Institute of Applied Electricity, Hokkaido
University, Sapporo, Japan, 1990.
[4] Yu. M. Arlinskiı, S. Hassi and H.S.V de Snoo, Parametrization of contractive
block operator matrices and passive discrete-time systems, Complex Anal.
Oper. Theory 1 (2007), no. 2, 211 -- 233.
[5] D. Z. Arov, Passive linear steady-state dynamical systems, Sibirsk. Mat. Zh.
20 (1979), no. 2, 211 -- 228 (Russian); English transl. in Siberian Math. J. 20
(1979), no. 2, 149 -- 162.
[6] D. Z. Arov, Stable dissipative linear stationary dynamical scattering sys-
tems, J. Operator Theory 2 (1979), no. 1, 95 -- 126 (Russian); English transl.
in Oper. Theory Adv. Appl., 134, Interpolation theory, systems theory and
related topics (Tel Aviv/Rehovot, 1999), 99 -- 136, Birkhauser, Basel, 2002.
[7] D. Z. Arov, M. A. Kaashoek, and D. P. Pik, Minimal and optimal linear
discrete time-invariant dissipative scattering systems. Integr. Equat. Oper.
Theory, 29 (1997), 127 -- 154.
[8] D. Z. Arov, M. A. Kaashoek, and D. P. Pik, The Kalman-Yakubovich-Popov
inequality for discrete time systems of infinite dimension, J. Operator Theory
55 (2006), no. 2, 393 -- 438.
[9] D. Z. Arov and M. A. Nudel'man, A criterion for the unitary similarity of
minimal passive systems of scattering with a given transfer function, Ukraın.
Mat. Zh. 52 (2000), no. 2, 147 -- 156 (Russian); English transl. in Ukrainian
Math. J. 52 (2000), no. 2, 161 -- 172.
36
L. Lilleberg
[10] D.Z. Arov and M. A. Nudel'man, Conditions for the similarity of all minimal
passive realizations of a given transfer function (scattering and resistance
matrices), Mat. Sb. 193 (2002), no. 6, 3 -- 24 (Russian); English transl. in Sb.
Math. 193 (2002), no. 5-6, 791 -- 810.
[11] D. Z. Arov, J. Rovnyak and S. M. Saprikin, Linear passive stationary scat-
tering systems with Pontryagin state spaces, Math. Nachr. 279 (2006), no.
13 -- 14, 1396 -- 1424.
[12] D. Z. Arov and S. M. Saprikin, Maximal solutions for embedding problem
for a generalized Shur function and optimal dissipative scattering systems
with Pontryagin state spaces, Methods Funct. Anal. Topology 7 (2001), no.
4, 69 -- 80.
[13] T. Ya. Azizov and I. S. Iokhvidov, Foundations of the theory of linear opera-
tors in spaces with indefinite metric, Nauka, Moscow, 1986; English transl.,
John Wiley & Sons Ltd., Chichester, 1989.
[14] H. Bart, , I. Z. Gohberg, M. A. Kaashoek and A. C. M. Ran, Factorization
of matrix and operator functions: the state space method, Oper. Theory Adv.
Appl., 178, Linear Operators and Linear Systems, Birkhauser Verlag, Basel,
2008.
[15] J. Bogn´ar, Indefinite inner product spaces, Ergebnisse der Mathematik und
ihrer Grenzgebiete, Band 78. Springer-Verlag, New York-Heidelberg, 1974.
[16] S. S. Boiko, V. K. Dubovoj, B. Fritzsche and B. Kirstein, Models of con-
tractions constructed from the defect function of their characteristic func-
tion. Operator theory, system theory and related topics (Beer-Sheva/Rehovot,
1997), 6787, Oper. Theory Adv. Appl., 123, Birkhauser, Basel, 2001.
[17] S. S. Boiko, V. K. Dubovoj, Defect functions of holomorphic contractive op-
erator functions and the scattering suboperator through the internal chan-
nels of a system. Part I, Complex Anal. Oper. Theory 5 (2011), no. 1, 157196.
[18] S. S. Boiko, V. K. Dubovoj, A. Ya. Kheifets, Defect functions of holomorphic
contractive operator functions and the scattering suboperator through the
internal channels of a system: Part II. Complex Anal. Oper. Theory 8 (2014),
no. 5, 9911036.
[19] L. de Branges and J. Rovnyak, Square Summable Power Series, Holt, Rine-
hart and Winston, New-York, 1966.
[20] L. de Branges and J. Rovnyak, Appendix on square summable power series,
Canonical models in quantum scattering theory, Perturbation Theory and its
Applications in Quantum Mechanics (Proc. Adv. Sem. Math. Res. Center,
U.S. Army, Theoret. Chem. Inst., Univ. of Wisconsin, Madison, Wis., 1965),
pp. 295 -- 392, Wiley, New York, 1966.
[21] M.S. Brodskiı, Unitary operator colligations and their characteristic func-
tions, Uspekhi Mat. Nauk 33 (1978), no. 4(202), 141 -- 168, 256 (Russian);
English transl. in Russian Math. Surveys 33 (1978), no. 4, 159 -- 191.
[22] A. Dijksma, H. Langer and H. S. V. de Snoo, Characteristic functions of uni-
tary operator colligations in πκ-spaces, Operator theory and systems (Am-
sterdam, 1985), 125 -- 194, Oper. Theory Adv. Appl., 19, Birkhauser, Basel,
1986.
Minimal realizations
37
[23] A. Dijksma, H. Langer and H. S. V. de Snoo, Unitary colligations in Πκ-
spaces, characteristic functions and Straus extensions, Pacific J. Math. 125
(1986), no. 2, 347 -- 362.
[24] M. A. Dritschel and J. Rovnyak, Operators on indefinite inner product
spaces, Lectures on operator theory and its applications (Waterloo, ON,
1994), 141 -- 232, Fields Inst. Monogr., 3, Amer. Math. Soc., Providence, RI,
1996.
[25] J. W. Helton, Discrete time systems, operator models, and scattering theory,
J. Functional Analysis 16 (1974), 15 -- 38.
[26] M. G. Kreın and H. Langer, Uber die verallgemeinerten Resolventen und
die charakteristische Funktion eines isometrischen Operators im Raume Πκ
(German), Hilbert space operators and operator algebras (Proc. Internat.
Conf., Tihany, 1970), pp. 353 -- 399, Colloq. Math. Soc. J´anos Bolyai, 5.
North-Holland, Amsterdam, 1972.
[27] L. Lilleberg, Isometric discrete-time systems with Pontryagin state space,
Complex Anal. Oper. Theory (2019). https://doi.org/10.1007/s11785-019-
00930-1
[28] S. M. Saprikin, The theory of linear discrete time-invariant dissipative scat-
tering systems with state πκ-spaces, Zap. Nauchn. Sem. S.-Peterburg. Ot-
del. Mat. Inst. Steklov. (POMI) 282 (2001), Issled. po Lineın. Oper. i Teor.
Funkts. 29, 192 -- 215, 281 (Russian); English transl. in J. Math. Sci. (N. Y.)
120 (2004), no. 5, 1752 -- 1765.
[29] O. J. Staffans, Well-posed linear systems, Encyclopedia of Mathematics and
its Applications, 103, Cambridge University Press, Cambridge, 2005.
[30] B. Sz.-Nagy and C. Foias, Harmonic analysis of operators on Hilbert space,
North-Holland, New York, 1970.
|
1510.08801 | 1 | 1510 | 2015-10-29T17:57:31 | Riemann integrability versus weak continuity | [
"math.FA"
] | In this paper we focus on the relation between Riemann integrability and weak continuity. A Banach space $X$ is said to have the weak Lebesgue property if every Riemann integrable function from $[0,1]$ into $X$ is weakly continuous almost everywhere. We prove that the weak Lebesgue property is stable under $\ell_1$-sums and obtain new examples of Banach spaces with and without this property. Furthermore, we characterize Dunford-Pettis operators in terms of Riemann integrability and provide a quantitative result about the size of the set of $\tau$-continuous non Riemann integrable functions, with $\tau$ a locally convex topology weaker than the norm topology. | math.FA | math |
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
GONZALO MART´INEZ-CERVANTES
Abstract. In this paper we focus on the relation between Riemann integrabil-
ity and weak continuity. A Banach space X is said to have the weak Lebesgue
property if every Riemann integrable function from [0, 1] into X is weakly
continuous almost everywhere. We prove that the weak Lebesgue property
is stable under ℓ1-sums and obtain new examples of Banach spaces with and
without this property. Furthermore, we characterize Dunford-Pettis operators
in terms of Riemann integrability and provide a quantitative result about the
size of the set of τ -continuous non Riemann integrable functions, with τ a
locally convex topology weaker than the norm topology.
1. Introduction
The study of the relation between Riemann integrability and continuity on Ba-
nach spaces started on 1927, when Graves showed in [13] the existence of a vector-
valued Riemann integrable function not continuous almost everywhere (a.e. for
short). Thus, the following problem arises:
Given a Banach space X, determine necessary and sufficient conditions for
the Riemann integrability of a function f : [0, 1] → X.
A Banach space X for which every Riemann integrable function f : [0, 1] → X is
continuous a.e. is said to have the Lebesgue property (LP for short). All classical
infinite-dimensional Banach spaces except ℓ1 do not have the LP. For more details
on this topic, we refer the reader to [12], [6], [24], [14] and [19].
Regarding weak continuity, Alexiewicz and Orlicz constructed in 1951 a Riemann
integrable function which is not weakly continuous a.e. [2]. A Banach space X is said
to have the weak Lebesgue property (WLP for short) if every Riemann integrable
function f : [0, 1] → X is weakly continuous a.e. This property was introduced in
[27]. Every Banach space with separable dual has the WLP and the example of
[2] shows that C([0, 1]) does not have the WLP. Other spaces with the WLP, such
as L1([0, 1]), can be found in [5] and [28]. In this paper we focus on the relation
between Riemann integrability and weak continuity. In Section 2 we present new
results on the WLP. In particular, we prove that the James tree space JT does not
have the WLP (Theorem 2.3) and we study when ℓp(Γ) and c0(Γ) have the WLP in
the nonseparable case (Theorem 2.8). Moreover, we prove that the WLP is stable
2010 Mathematics Subject Classification. 46G10, 28B05, 03E10.
Key words and phrases. Riemann integral, Lebesgue property, weak Lebesgue property, Ba-
nach space.
This work was supported by the research project 19275/PI/14 funded by Fundaci´on S´eneca -
Agencia de Ciencia y Tecnolog´ıa de la Regi´on de Murcia within the framework of PCTIRM 2011-
2014. This work was also supported by Ministerio de Econom´ıa y Competitividad and FEDER
(project MTM2014-54182-P).
1
2
GONZALO MART´INEZ-CERVANTES
under ℓ1-sums (Theorem 2.13) and we apply this result to obtain that C(K)∗ has
the WLP for every compact space K in the class M S (Corollary 2.16).
Alexiewicz and Orlicz also provided in their paper an example of a weakly con-
tinuous non Riemann integrable function. V. Kadets proved in [15] that a Banach
space X has the Schur property if and only if every weakly continuous function
f : [0, 1] → X is Riemann integrable. Wang and Yang extended this result in [29]
to arbitrary locally convex topologies weaker than the norm topology. In the last
section of this paper we give an operator theoretic form of these results that, in
particular, provides a positive answer to a question posed by Sofi in [26].
Terminology and Preliminaries. All Banach spaces are assumed to be real.
In what follows, X ∗ denotes the dual of a Banach space X. The weak and weak∗
topologies of X and X ∗ will be denoted by ω and ω∗ respectively. By an operator we
mean a linear continuous mapping between Banach spaces. The Lebesgue measure
in R is denoted by µ. The interior of an interval I will be denoted by Int(I). The
density character dens(T ) of a topological space T is the minimal cardinality of a
dense subset.
N
a f (t)dt.
A partition of the interval [a, b] ⊂ R is a finite collection of non-overlapping closed
subintervals covering [a, b]. A tagged partition of the interval [a, b] is a partition
{[ti−1, ti] : 1 ≤ i ≤ N} of [a, b] together with a set of points {si : 1 ≤ i ≤ N} that
satisfy si ∈ (ti−1, ti) for each i. Let P = {(si, [ti−1, ti]) : 1 ≤ i ≤ N} be a tagged
partition of [a, b]. For every function f : [a, b] → X, we denote by f (P) the Riemann
Pi=1
(ti − ti−1)f (si). The norm of P is kPk := max{ti − ti−1 : 1 ≤ i ≤ N}. We
sum
say that a function f : [a, b] → X is Riemann integrable, with integral x ∈ X, if for
every ε > 0 there is δ > 0 such that kf (P) − xk < ε for all tagged partitions P of
[a, b] with norm less than δ. In this case, we write x =R b
Theorem 1.1 ([12]). Let f : [0, 1] → X. The following statements are equivalent:
The following criterion will be used for proving the existence of the Riemann
integral of certain functions:
(1) The function f is Riemann integrable.
(2) For each ε > 0 there exists a partition Pε of [0, 1] with kf (P1)− f (P2)k < ε
for all tagged partitions P1 and P2 of [0, 1] that have the same intervals as
Pε.
(3) There is x ∈ X such that for every ε > 0 there exists a partition Pε of [0, 1]
such that kf (P)− xk < ε whenever P is a tagged partition of [0, 1] with the
same intervals as Pε.
We will also be concerned about cardinality. Throughout this paper, c denotes
the cardinality of the continuum and cov(M) denotes the smallest cardinal such
that there exist cov(M) nowhere dense sets in [0, 1] whose union is the interval [0, 1].
This cardinal coincides with the smallest cardinal such that there exist cov(M)
closed sets in [0, 1] with Lebesgue measure zero whose union does not have Lebesgue
measure zero (see [4, Theorem 2.6.14]).
(εn)∞
A set A ⊂ R is said to be strongly null if for every sequence of positive reals
n=1 there exists a sequence of intervals (In)∞
n=1 such that µ(In) < εn for every
n ∈ N and A ⊂Sn∈N In. We will be interested in the following result:
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
3
Theorem 1.2 ([22]). A set A ⊂ R is strongly null if and only if for every closed
set F with Lebesgue measure zero, the set A + F = {a + z : a ∈ A and z ∈ F} has
Lebesgue measure zero.
We will denote by non(SN ) the smallest cardinal of a non strongly null set.
We have that ℵ1 ≤ cov(M) ≤ non(SN ) ≤ c and, under Martin's axiom, and
therefore under the Continuum Hypothesis, non(SN ) = cov(M) = c. Furthermore,
if b = c then non(SN ) = cov(M). However, there exist models of ZFC satisfying
cov(M) < non(SN ). For further references and results on this subject we refer the
reader to [3].
2. The weak Lebesgue property
It is known that every Banach space with separable dual has the WLP [28]. Next
theorem gives a generalization in terms of cov(M).
Theorem 2.1. Every Banach space X such that dens(X ∗) < cov(M) has the
WLP.
Proof. Let D = {x∗
i }i∈I be a dense subset in X ∗ with I < cov(M) and take
f : [0, 1] → X a Riemann integrable function. Then every function x∗
i f is Riemann
integrable. Let Ei be the set of points of discontinuity of x∗
i f for every i ∈ I.
Each Ei is a countable union of closed sets with measure zero, so E =Si∈I Ei has
measure zero since I < cov(M). We claim that f is weakly continuous at every
point of Ec. Let x∗ ∈ X ∗ and let M be an upper bound for {kf (t)k : t ∈ [0, 1]}.
Fix ε > 0 and t ∈ Ec. Then, there exists x∗
3M . Since
t /∈ Ei, there exists a neighbourhood U of t such that x∗
3 for
every t′ ∈ U . Thus,
x∗f (t) − x∗f (t′) ≤ x∗f (t) − x∗
for every t′ ∈ U .
Corollary 2.2. Every Banach space with separable dual has the WLP.
i f (t′) − x∗f (t′) < ε
i ∈ D such that kx∗
i − x∗k < ε
i f (t) + x∗
i f (t) − x∗
i f (t′) + x∗
i f (t) − x∗
i f (t′) < ε
(cid:3)
The space ℓ1 has the WLP because it has the LP. Since every asymptotic ℓ1
space has the LP [19], the space ΛT defined by Odell in [21] is a separable Banach
space with nonseparable dual such that it does not contain an isomorphic copy of
ℓ1 but it has the WLP (it is asymptotic ℓ1). On the other hand, the James tree
space JT (see [1, Section 13.4]) is a separable Banach space with nonseparable dual
such that it does not contain an isomorphic copy of ℓ1 and it does not have the
WLP:
Theorem 2.3. The James tree space does not have the WLP.
Proof. We represent the dyadic tree by
T = {(n, k) : n = 0, 1, 2, . . . and k = 1, 2, . . . , 2n}.
A node (n, k) ∈ T has two immediate successors (n+1, 2k−1) and (n+1, 2k). Then,
a segment of T is a finite sequence {p1, . . . , pm} such that pj+1 is an immediate
successor of pj for every j = 1, 2, . . . , m − 1. The James tree space JT is the
completion of c00(T ) with the norm
l
kxk = supvuuut
Xj=1
2
X(n,k)∈Sj
x(n, k)
< ∞,
4
GONZALO MART´INEZ-CERVANTES
where the supremum is taken over all l ∈ N and all sets of pairwise disjoint segments
S1, S2, . . . , Sl. Let {e(n,k)}(n,k)∈T be the canonical basis of JT , i.e. e(n,k) is the
characteristic function of (n, k) ∈ T . Define f : [0, 1] → JT as follows:
f (t) =( e(n−1,k)
0
if t = 2k−1
in any other case.
2n with n ∈ N and k = 1, 2, . . . , 2n−1
We claim that f is Riemann integrable. Fix N ∈ N and let {I1, I2, . . . , I2N −1} be a
family of closed disjoint intervals of [0, 1] with
µ(In) ≤
1
2N and
n
2N ∈ Int(In) for each 1 ≤ n ≤ 2N − 1.
X1≤n≤2N −1
Let J1, J2, . . . , J2N be the closed disjoint intervals of [0, 1] determined by
Then, µ(Jn) ≤ 1
n for every an ∈ R and every
tn ∈ Jn due to the definition of the norm in JT . Thus, any tagged partition PN
1, t2, . . . , t′
with intervals J1, I1, J2, . . . , I2N −1, J2N and points t1, t′
2N −1, t2N satisfies
n=1 a2
Int(In).
[0, 1] \ [1≤n≤2N −1
n=1 anf (tn)k ≤qP2N
µ(Jn)f (t2n−1)(cid:13)(cid:13)(cid:13)(cid:13)
2N ≤vuut
Xn=1
2N and kP2N
kf (PN )k ≤(cid:13)(cid:13)(cid:13)(cid:13)
Xn=1
≤vuut
Xn=1
µ(Jn)2 +
2N
2N
2N
1
+
2N −1
Xn=1
µ(In) ≤
1
22N +
1
2N ≤
2
√2N
.
Hence, kf (PN )k N→∞−−−−→ 0 and f is Riemann integrable with integral zero.
We show that f is not weakly continuous at any irrational point t ∈ [0, 1]. Fix
2nj (cid:17)∞
a irrational point t ∈ [0, 1]. There exists a sequence of dyadic points (cid:16) 2kj −1
converging to t with (nj − 1, kj)∞
j=1 a sequence in T such that (nj+1 − 1, kj+1) is
an immediate successor of (nj − 1, kj) for every j ∈ N. Then, P∞
(nj −1,kj ) is a
functional in JT ∗, so the sequence f ( 2kj −1
2nj ) = e(nj −1,kj ) is not weakly null and f
is not weakly continuous at t.
(cid:3)
j=1 e∗
j=1
Corollary 2.4 ([2]). C([0, 1]) does not have the WLP.
Proof. Since every subspace of a Banach space with the WLP has the WLP and
every separable Banach space is isometrically isomorphic to a subspace of C([0, 1]),
it follows from the previous theorem and the separability of JT that C([0, 1]) does
not have the WLP.
(cid:3)
Corollary 2.5. Let K be a compact Hausdorff space.
(1) If K is metrizable, then C(K) has the WLP if and only if K is countable.
(2) If C(K) has the WLP then K is scattered. The converse is not true since
c0(c) does not have the WLP (Theorem 2.8) and it is isomorphic to a C(K)
space with K scattered.
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
5
Proof. If K is a countable compact metric space, then C(K)∗ is separable [10,
Theorem 14.24], so C(K) has the WLP (Theorem 2.1).
If K is an uncountable
compact metric space, then C(K) is isomorphic to C([0, 1]) [1, Theorem 4.4.8], so
C(K) does not have the WLP (Corollary 2.4). Finally, if K is not scattered, then
C(K) has a subspace isomorphic to C([0, 1]) (see the proof of [10, Theorem 14.26]),
so C(K) does not have the WLP.
Remark 2.6. Let {Xi}i∈Γ be a family of Banach spaces. Define X := (Li∈Γ Xi)ℓp
with 1 < p < ∞ or X := (Li∈Γ Xi)c0 . If f : [0, 1] → X is a bounded function,
then its set of points of weak discontinuity is E = Si∈Γ Ei, where each Ei is the
set of points of weak discontinuity of fi and fi is the i'th cordinate of f . Thus,
the countable ℓp-sum or c0-sum of Banach spaces with the WLP has the WLP. We
cannot extend this result to uncountable ℓp-sums or c0-sums even when Xi = R for
every i ∈ Γ (Theorem 2.8).
(cid:3)
Now, we study the WLP for the spaces of the form c0(κ) and ℓp(κ) with κ a
cardinal.
Theorem 2.7. For any cardinal κ and any 1 < p < ∞, c0(κ) has the WLP if and
only if ℓp(κ) has the WLP.
α has not Lebesgue measure zero.
Proof. If ℓp(κ) does not have the WLP, then there exists a Riemann integrable
function f : [0, 1] → ℓp(κ) which is not weakly continuous a.e. If I : ℓp(κ) → c0(κ)
is the canonical inclusion, then the function I ◦ f is weakly continuous at a point
t ∈ [0, 1] if and only if f is weakly continuous at t by Remark 2.6. Therefore, I ◦ f is
not weakly continuous a.e. Since I is an operator, I ◦ f is also Riemann integrable.
Thus, c0(κ) does not have the WLP.
To prove the other implication, suppose c0(κ) does not have the WLP. Then,
there exists a Riemann integrable function f : [0, 1] → c0(κ) which is not weakly
continuous a.e. Let fα be the α'th cordinate of f for every α ∈ κ and En
α be the
set of points where fα has oscillation strictly bigger than 1
n for every n ∈ N. Note
that each En
α has Lebesgue measure zero. Since f is not weakly continuous a.e.,
Sα<κ(cid:0)Sn∈N En
α(cid:1) has not Lebesgue measure zero, so there exists n ∈ N such that
Sα<κ En
β(cid:17) for every α ∈ κ \ {0}. The sets
Fα are pairwise disjoint. Let χFα : [0, 1] → {0, 1} be the characteristic function
of Fα for every α < κ and g : [0, 1] → c0(κ) the function defined by the formula
g(t) = Pα<κ χFα (t)eα for every t ∈ [0, 1], where {eα}α<κ is the canonical basis of
X.
Notice that g is not weakly continuous a.e. since each χFα is not continuous at
any point of Fα (because µ(Fα) = 0) and Sα<κ Fα = Sα<κ En
α is not Lebesgue
null. We claim that g is Riemann integrable. Let ε > 0. Since f is Riemann
integrable, there exists a partition Pε of [0, 1] such that kf (P1)− f (P2)k < ε
n for all
tagged partitions P1 and P2 of [0, 1] that have the same intervals as Pε. For every
α < κ and any tagged partitions P1 and P2 of [0, 1] that have the same intervals as
Pε,
α \(cid:16)Sβ<α En
0 and Fα := En
Set F0 := En
χFα(P1) − χFα(P2) ≤
N
Xi=1
µ(Ii) ≤ nfα(P ′
1) − fα(P ′
2) ≤ nkf (P ′
1) − f (P ′
2)k < ε
µ [Int(Ii)∩Fα6=∅
µ(Ii)eφ(si)(cid:13)(cid:13)(cid:13)(cid:13)
kh(P)k =(cid:13)(cid:13)(cid:13)(cid:13)Xsi∈F
Ii(cid:19)p(cid:19)
µ(cid:18) [φ(si)=α
=(cid:18)Xα<κ
(1)
p (cid:18)Xα<κ
µ(cid:18) [φ(si)=α
≤ ε
=(cid:13)(cid:13)(cid:13)(cid:13)Xα<κ
µ(cid:18) [φ(si)=α
Ii(cid:19)(cid:19)
p−1
1
p
=(cid:18)Xα<κ
=
Ii(cid:19)eα(cid:13)(cid:13)(cid:13)(cid:13)
µ(cid:18) [φ(si)=α
µ(cid:18) [φ(si)=α
Ii(cid:19)p−1
≤ ε
1
p
p
p−1
1
p
Ii(cid:19)(cid:19)
≤
(cid:3)
6
GONZALO MART´INEZ-CERVANTES
1 and P ′
for suitable tagged partitions P ′
2 of [0, 1] with the same intervals as Pε, where
I1, I2, . . . , IN are the intervals of Pε whose interior has non-empty intersection with
En
α.
Therefore, g is Riemann integrable. Let h : [0, 1] → ℓp(κ) be the function given
by the formula h(t) = Pα<κ χFα(t)eα. Since the sets Fα are pairwise disjoint,
the function h is well-defined. Moreover, h is not weakly continuous a.e. because
I ◦ h = g. Set F =Sα<κ Fα and φ : F → κ such that φ(t) = α if t ∈ Fα. We claim
that h is Riemann integrable with integral zero. Let ε > 0 and Pε = {I1, I2, . . . , IM}
be a partition of [0, 1] such that kg(P ′)k < ε for any tagged partition P ′ of [0, 1]
with the same intervals as Pε. Notice that
(1)
Ii! < ε for every α < κ.
Thus, for any tagged partition P = {(si, Ii)}M
i=1 the following inequalities hold:
Therefore, h is Riemann integrable with Riemann integral zero.
The LP is separably determined [24]. Nevertheless, it follows from the following
theorem that the WLP is not separably determined, since every separable infinite-
dimensional subspace of ℓ2(κ) is isomorphic to ℓ2 (which has separable dual).
Theorem 2.8. Let κ be a cardinal and X = c0(κ) or X = ℓp(κ) with 1 < p < ∞.
(1) If κ < cov(M) then X has the WLP.
(2) If κ ≥ non(SN ) then X does not have the WLP.
Proof. It is enough to prove the result when X = c0(κ) due to Theorem 2.7. Since
c0(κ)∗ = ℓ1(κ) has density character κ, it follows from Theorem 2.1 that c0(κ) has
the WLP if κ < cov(M).
Suppose non(SN ) ≤ κ ≤ c. Due to Theorem 1.2, there exist a closed Lebesgue
null set F and a set E = {xα}α<κ in R such that E + F does not have Lebesgue
measure zero. Without loss of generality, we may assume that E, F ⊂ [0, 1] and
(E + F )∩[0, 1] does not have Lebesgue measure zero. Set F0 := (x0 + F )∩[0, 1] and
Fα := ((xα + F ) ∩ [0, 1])\(cid:16)Sβ<α Fβ(cid:17) for every 0 < α < κ. Let χFα : [0, 1] → {0, 1}
be the characteristic function of Fα for every α < κ and f : [0, 1] → X the function
defined by the formula f (t) = Pα<κ χFα (t)eα for every t ∈ [0, 1], where {eα}α<κ
is the canonical basis of c0(κ).
Since the sets Fα are pairwise disjoint, the function f is well-defined. Each χFα
is not continuous at Fα, since Fα cannot contain an interval of [0, 1]. Therefore, f
is not weakly continuous a.e. because Sα<κ Fα = (E + F ) ∩ [0, 1] does not have
Lebesgue measure zero.
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
7
We claim that f is Riemann integrable. For every α < κ and every tagged
i=1 we have
partition P = {(si, Ii)}N
Xi=1
χFα(P) =
N
µ(Ii)χFα (si) ≤
µ(Ii − xα)χF (si − xα) = χF (P ′)
N
Xi=1
for a suitable tagged partition P ′ with kPk = kP ′k. Since F ⊂ [0, 1] is a closed
Lebesgue measure zero set, the characteristic function χF is Riemann integrable
due to Lebesgue's Theorem. Then, for every ε > 0 there exists δ > 0 such that
χF (P) < ε for every tagged partition P with kPk < δ. Therefore, for every ε > 0
there exists δ > 0 such that χFα(P) < ε for all tagged partitions P with kPk < δ and
for every α < κ. Thus, f is Riemann integrable since kf (P)k = supα<κ χFα(P) < ε
for every tagged partition P of [0, 1] with kPk < δ.
(cid:3)
The facts that the countable ℓ1-sum of spaces with the WLP has the WLP
(Theorem 2.11) and that L1(λ) has the WLP if dens(L1(λ)) < cov(M) (Theorem
2.12) will be a consequence of the following lemma.
Lemma 2.9. Let (Ω, Σ, λ) be a probability space and P = {PA : A ∈ Σ} a family
of operators on a Banach space X such that
A ∩ B = ∅;
(1) PA + PΩ\A = PΩ = idX for every A ∈ Σ;
(2) kPA(x)k ≤ kxk for every x ∈ X and every A ∈ Σ;
(3) kPA(x)k+kPB(x′)k ≤ max{kx+x′k,kx−x′k} for every x, x′ ∈ X whenever
(4) limλ(A)→0 kPA(x)k = 0 for every x ∈ X.
Let f : [0, 1] → X be a Riemann integrable function. Then there is a measurable
set E ⊆ [0, 1] with µ(E) = 1 such that, for every sequence (tn)∞
n=1 in [0, 1] converg-
ing to some t ∈ E, the set {f (tn) : n ∈ N} is P-uniformly integrable, in the sense
that
Proof. The proof is similar to that of [5, Lemma 2.3] and [28, Lemma 3]. Fix β > 0
and denote by Eβ the set of points t ∈ [0, 1] such that for every δ > 0 there exist
t′ ∈ [0, 1] with t′ − t < δ and a set A ∈ Σ with λ(A) < δ such that
lim
λ(A)→0
sup
n∈N(cid:13)(cid:13)PA(f (tn))(cid:13)(cid:13) = 0.
kPA(f (t) − f (t′))k > β.
Let µ∗ be the Lebesgue outer measure in [0, 1]. We show that µ∗(Eβ) = 0 with a
proof by contradiction. Suppose µ∗(Eβ) > 0. Since f is Riemann integrable, we
can choose a partition P = {J1, . . . , Jm} of [0, 1] such that
< βµ∗(Eβ)
(2)
m
µ(Jj)(f (ξj ) − f (ξ′
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
j))(cid:13)(cid:13)(cid:13)(cid:13)
for all choices ξj , ξ′
where Ij = Int(Jj) for each j = 1, . . . , m. Thus,
j ∈ Jj, 1 ≤ j ≤ m. Let S = {j ∈ {1, . . . , m} : Ij ∩ Eβ 6= ∅},
(3)
µ∗(Ij ∩ Eβ) = µ∗(Eβ).
Xj∈S
It is not restrictive to suppose S = {1, . . . , n} for some 1 ≤ n ≤ m.
8
GONZALO MART´INEZ-CERVANTES
k
Because of the definition of Eβ and I1, there exist points t1 ∈ I1 ∩ Eβ and
1 ∈ I1 such that kf (t1) − f (t′
t′
1))k > β for some A ∈ Σ, hence
kµ(I1)(f (t1) − f (t′
1))k > βµ(I1).
Fix 1 ≤ k < n and assume that we have already chosen points tj, t′
j ∈ Ij for all
1 ≤ j ≤ k with the property that
1)k ≥ kPA(f (t1) − f (t′
> β(cid:18) k
Xj=1
Define x :=
µ(Ij)(cid:19).
µ(Ij )(f (tj) − f (t′
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
µ(Ij )(f (tj) − f (t′
j))(cid:13)(cid:13)(cid:13)(cid:13)
α := kxk − β(cid:18) k
µ(Ij )(cid:19) > 0.
Xj=1
Due to (4), we can choose δ > 0 such that kPA(x)k < α whenever A ∈ Σ satisfies
λ(A) < δ. Take tk+1, t′
k+1 ∈ Ik+1 and a set A ∈ Σ with λ(A) < δ such that
kPA(f (tk+1) − f (t′
k+1))k > β, so y := µ(Ik+1)(f (tk+1) − f (t′
j)) ∈ X and
k+1)) satisfies
Pj=1
k
kPA(y)k > βµ(Ik+1).
By the choice of A, (1) and (3), we also have (interchanging the role of tk+1 and
t′
k+1 if necessary)
k+1
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
µ(Ij)(f (tj ) − f (t′
j))(cid:13)(cid:13)(cid:13)(cid:13)
≥ kPA(y)k + kPAc(x)k ≥ kPA(y)k + kxk − kPA(x)k >
k
k+1
n
µ(Ij ).
Xj=1
> βµ(Ik+1) + α + β
Thus, there exist tj, t′
µ(Ij )(f (tj) − f (t′
Xj=1
µ(Ij ) − kPA(x)k > β
j ∈ Ij for all 1 ≤ j ≤ n such that
µ(Ij)(cid:19) (3)
Xj=1
j))(cid:13)(cid:13)(cid:13)(cid:13)
Therefore, E := [0, 1] \Sn∈N E 1
m ∈ N. Since t /∈ E 1
t′ − t < δm and every set A ∈ Σ with λ(A) < δm,
kPA(f (t) − f (t′))k ≤
> β(cid:18) n
Xj=1
is measurable with µ(E) = 1. Fix t ∈ E and
, there exists δm > 0 such that for every t′ ∈ [0, 1] with
which contradicts the inequality (2). So we can conclude that µ∗(Eβ) = 0.
≥ βµ∗(Eβ),
1
m
.
(cid:13)(cid:13)(cid:13)(cid:13)
m
n
Thus, for every m ∈ N, every sequence (tn)∞
with λ(A) < δm,
n=1 converging to t and every A ∈ Σ
1
m
Now the conclusion follows from (4).
kPA(f (tn))k ≤ kPA(f (t))k +
for n big enough depending only on m.
(cid:3)
Let {Xi}i∈Γ be a family of Banach spaces. We denote by πj : (Li∈Γ Xi) → Xj
the canonical projection onto Xj for each j ∈ Γ.
2.11 and 2.12:
We will need the following property of ℓ1-sums and the space L1(λ) for Theorems
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
9
Proof. The second part is essentially Lemma 2 of [28]. The proof of the first part is
x, y ∈ (Li∈Γ Xi)ℓ1 and any disjoint sets A, B ⊂ Γ.
disjoint sets A, B ∈ Σ.
Lemma 2.10. Let (Ω, Σ, λ) be a probability space and {Xi}i∈Γ a family of Banach
spaces. Then:
(1) max{kx + yk,kx − yk} ≥ Pi∈A kπi(x)k +Pi∈B kπi(y)k for every vectors
(2) max{kf + gk,kf − gk} ≥RA fdλ +RB gdλ for any f, g ∈ L1(λ) and any
analogous and we include it for the sake of completeness. Let x, y ∈ (Li∈Γ Xi)ℓ1,
A, B ⊂ Γ be disjoint sets and ε > 0. Without loss of generality, we may assume that
A = {an : n ∈ N} and B = {bn : n ∈ N} are countable subsets. Consider the func-
i )ℓ∞ defined by x∗(u) = Pi∈A x∗
tionals x∗, y∗ ∈ (Li∈Γ Xi)∗
i (πi(u))
i (πi(u)) for every u ∈ (Li∈Γ Xi)ℓ1, where each x∗
and y∗(u) = Pi∈B y∗
i ∈ X ∗
i
satisfies kx∗
i k = 1, x∗
i k = ky∗
i (πi(y)) = kπi(y)k if
i = bn. Then, since A, B are disjoint, kx∗ + y∗k = kx∗ − y∗k = 1. Therefore,
kx + yk + kx − yk ≥ hx + y, x∗ + y∗i + hx − y, x∗ − y∗i = 2hx, x∗i + 2hy, y∗i =
ℓ1 = (Li∈Γ X ∗
i (πi(x)) = kπi(x)k if i = an and y∗
i , y∗
= 2(cid:18)Xi∈A
x∗
i (πi(x)) +Xi∈B
y∗
i (πi(y))(cid:19) = 2(cid:18)Xi∈A
kπi(x)k +Xi∈B
kπi(y)k(cid:19),
so max{kx + yk,kx − yk} ≥Pi∈A kπi(x)k +Pi∈B kπi(y)k.
Theorem 2.11. Let {Xi}i∈N be Banach spaces with the WLP. Then the space
X := (Li∈N Xi)ℓ1 has the WLP.
Proof. We are going to apply Lemma 2.9. Take Ω := N, Σ := P(N) the power set
of N, λ(A) :=Pn∈A 2−n and P = {PA : A ∈ Σ} with
if i ∈ A
if i /∈ A
πi(PA(x)) =(πi(x)
0
(cid:3)
for every A ∈ Σ and every x ∈ X. Property (3) of Lemma 2.9 is Lemma 2.10(1)
and property (4) holds because if λ(A) < 1
2n , then A ⊂ {n, n + 1, . . .}, so
kPA(x)k =Xi∈A
kπi(x)k ≤Xi≥n
kπi(x)k
for every x ∈ X. Therefore, we can apply Lemma 2.9, so there exists a measurable
set E ⊂ [0, 1] with µ(E) = 1 such that for every sequence (tn)∞
n=1 in [0, 1] converging
to some t ∈ E the set {f (tn) : n ∈ N} is P-uniformly integrable. We can assume
that, for each i ∈ N, the map t 7→ πi(f (t)) is weakly continuous at each point of E
because each Xi has the WLP.
n=1 in X converges weakly to x ∈ X
It is a well known fact that a sequence (xn)∞
if and only if it satisfies the following two conditions:
(i) πi(xn) → πi(x) weakly in Xi for every i ∈ N;
(ii) for every ε > 0 there is a finite set J ⊆ N such that supn∈N kPN\J (xn)k ≤ ε.
Since P-uniform integrability is equivalent to (ii), it follows that f is weakly
(cid:3)
continuous at each point of E.
A similar idea to that of Theorem 2.11 let us prove the following theorem, which
improves [28, Theorem 5] and [5, Proposition 2.10].
Theorem 2.12. Let (Ω, Σ, λ) be a probability space.
10
GONZALO MART´INEZ-CERVANTES
(1) If dens(L1(λ)) < cov(M) then L1(λ) has the WLP.
(2) If dens(L1(λ)) ≥ non(SN ) then L1(λ) does not have the WLP.
Proof. Fix a Riemann integrable function f : [0, 1] → L1(λ). Take PA(x) := xχA
for every A ∈ Σ and every x ∈ L1(λ). The family of operators {PA : A ∈ Σ} fulfills
the requirements of Lemma 2.9 (bear in mind Lemma 2.10). Then P-uniform
integrability is the usual uniform integrability and therefore a set is bounded and
P-uniformly integrable if and only if it is relatively weakly compact due to Dunford's
Theorem (see [1, Theorem 5.2.9]). Lemma 2.9 ensures that there exist a measurable
n=1 in [0, 1] converging
set E ⊂ [0, 1] with µ(E) = 1 such that for every sequence (tn)∞
to some t ∈ E, the set {f (tn) : n ∈ N} is relatively weakly compact.
Let C ⊂ Σ be a dense family of λ-measurable sets, i.e. such that
inf
C∈C
λ(A △ C) = 0 for every A ∈ Σ.
Let (hn)∞
n=1 be a relatively weakly compact sequence in L1(λ) and h ∈ L1(λ).
Since C is a dense family of λ-measurable sets, if RC hn dµ → RC h dµ for every
C ∈ C, then h = ω-lim hn.
Suppose dens(L1(λ)) < cov(M). Then C can be taken such that C < cov(M).
Therefore, we can assume that, for each C ∈ C, the Riemann integrable map
t 7→RC f (t) dλ is continuous at each point of E. Then, for every sequence (tn)∞
in [0, 1] converging to a point t ∈ E, we have f (t) = ω-lim f (tn).
Now suppose ν = dens(L1(λ)) ≥ non(SN ). Due to Maharam's Theorem (see
[16, p. 127, Theorem 9]), L1(λ) contains an isometric copy of L1(µν), where µν
is the usual product probability measure on {0, 1}ν. Since L1(µν ) contains an
isomorphic copy of ℓ2(ν) (see [16, p. 128, Theorem 12]) and ℓ2(ν) does not have the
WLP (Theorem 2.8), we conclude that L1(λ) does not have the WLP.
(cid:3)
n=1
Theorem 2.11 can be extended to arbitrary ℓ1-sums:
Theorem 2.13. The arbitrary ℓ1-sum of a family of Banach spaces with the WLP
has the WLP.
Proof. The proof uses some ideas of [18]. Let f : [0, 1] → X := (Li∈Γ Xi)ℓ1 be
a Riemann integrable function, where {Xi}i∈Γ is a family of Banach spaces with
the WLP. For each J ⊂ Γ, we denote by PJ : X → X the function defined by
πi (PJ (x)) = πi(x) if i ∈ J and πi (PJ (x)) = 0 in any other case. Let (rn)∞
n=1 be
an enumeration of the rational numbers in [0, 1] and fix a countable set L ⊂ Γ such
that PL(f (rn)) = f (rn) for every n ∈ N. Then, f = (f − PLf ) + PLf . Since PLf
is Riemann integrable and takes values in the space
XL := {x ∈ X : πi(x) = 0 for each i /∈ L},
which is isomorphic to a countable ℓ1-sum of spaces with the WLP, by Theorem
2.11 PLf is weakly continuous almost everywhere.
0 f (t)dt = 0 and that f is null over a dense set.
Let
Therefore, we can assume that R 1
for each n ∈ N and each subset J ⊂ Γ. If J1 ⊂ J2 ⊂ Γ, then AJ2
Claim: For every n ∈ N there exists a countable set J ⊂ Γ with µ(cid:16)AJ
Suppose this is not the case. Then, there exist n ∈ N and δ > 0 with µ(cid:16)AJ
AJ
n := {t ∈ [0, 1] : kPJ c (f (t))k ≥
n ⊂ AJ1
n .
1
n}
n(cid:17) = 0.
n(cid:17) > δ
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
11
N
<
(4)
δ
n
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
for all choices ξj, ξ′
for every countable subset J ⊂ Γ (if for every m ∈ N we can take a countable
n(cid:17) = 0). Let
set Jm ⊂ Γ with µ(cid:16)AJm
P = {I1, I2, . . . , IN} be a partition of [0, 1] such that
m , then J = Sm∈N Jm verifies µ(cid:16)AJ
n (cid:17) < 1
j))(cid:13)(cid:13)(cid:13)(cid:13)
µ(Ij)(f (ξj ) − f (ξ′
j ∈ Ij, 1 ≤ j ≤ N.
n(cid:17) = µ(cid:16)AJ
Let J ⊂ Γ be a countable subset. Since
null over a dense set, we can suppose that there exist ξ1 ∈ Int(I1)∩ AJ
such that kµ(I1)(f (ξ1) − f (ξ′
(4) we have µ(I1) < δ < PN
Int(I2) ∩ AJ1
kµ(I1)(f (ξ1) − f (ξ′
n(cid:17) > δ and f is
Pj=1
1 ∈ I1
1))k ≥ 1
1). By
n (cid:17) and so it is not restrictive to suppose
j=1 µ(cid:16)Ij ∩ AJ1
2 ∈ I2 such that
1)) + µ(I2)(f (ξ2) − f (ξ′
n and ξ′
n µ(I1). Let J1 = supp f (ξ1) ∪ supp f (ξ′
n 6= ∅. Thus, due to Lemma 2.10, we can choose ξ2, ξ′
µ(cid:16)Ij ∩ AJ
(µ(I1) + µ(I2)).
2))k ≥
1
n
N
Fix 1 ≤ k < N and assume that we have already chosen points ξj , ξ′
1 ≤ j ≤ k with the property that
j ∈ Ij for all
N
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
µ(Ij)(f (ξj ) − f (ξ′
δ
n
.
≥
j))(cid:13)(cid:13)(cid:13)(cid:13)
But this is a contradiction with (4). Therefore, the Claim is proved.
n (cid:17) = 0. Fix
Thus, for every n ∈ N there exists a countable set Jn such that µ(cid:16)AJn
J :=Sn∈N Jn. Theorem 2.11 guarantees the existence of a set F ⊂ [0, 1] of measure
one such that PJ (f ) is weakly continuous at every point of F . Let E = F \Sn∈N AJ
n.
Then, µ(E) = 1, f = PJ (f ) + PJ c(f ), PJ (f ) is weakly continuous at each point of
E and PJ c (f ) is norm continuous at each point of E (if tn → t ∈ E, then, for every
m ∈ N, tn /∈ AJ
Corollary 2.14 ([24, 20]). ℓ1(κ) has the LP for any cardinal κ.
m for n big enough so kPJ c(f )(tn)k < 1
m ).
(cid:3)
Proof. Since ℓ1(κ) has the Schur property, ℓ1(κ) has the LP if and only it has the
WLP. Therefore, the conclusion follows from Theorem 2.13.
(cid:3)
µ(Ij )(f (ξj ) − f (ξ′
≥
k
1
n
Xj=1
µ(Ij )
.
j ))(cid:13)(cid:13)(cid:13)(cid:13)
j=1 supp f (ξj) ∪ supp f (ξ′
j), which is countable. By (4) we have
j=1 µ(cid:16)Ij ∩ AJk
n (cid:17), hence it is not restrictive to suppose that
k+1 ∈ Ik+1 such
k+1
n 6= ∅ and therefore that there exist points ξk+1, ξ′
µ(Ij )
.
µ(Ij )(f (ξj ) − f (ξ′
n
≥
k+1
1
j ))(cid:13)(cid:13)(cid:13)(cid:13)
µ(Ij ) = 1 > δ, it follows that there exist ξj, ξ′
Xj=1
j ∈ Ij for every 1 ≤ j ≤ N
k
k
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
Set Jk := Sk
µ(Ij ) < δ < PN
Pj=1
Int(Ik+1) ∩ AJk
that
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
N
Since
Pj=1
such that
12
GONZALO MART´INEZ-CERVANTES
As an application of 2.13 we also obtain the following result:
Corollary 2.15. Let K be a compact Hausdorff space. Then, C(K)∗ has the WLP
if dens(L1(λ)) < cov(M) for every regular Borel probability λ on K.
Proof. For every compact Hausdorff space K, the Banach space C(K)∗ is isometric
to a ℓ1-sum of L1(λ) spaces, where each λ is a regular Borel probability measure on
K (see the proof of [1, Proposition 4.3.8]). Thus, C(K)∗ has the WLP if each space
L1(λ) has the WLP, due to Theorem 2.13. Hence, the result follows from Theorem
2.12.
(cid:3)
Corollary 2.16. If K is a compact Hausdorff space in the class M S (i.e. L1(λ)
is separable for every regular Borel probability on K), then C(K)∗ has the WLP.
Some classes of compact spaces in the class M S are metric compacta, Eber-
lein compacta, Radon-Nikod´ym compacta, Rosenthal compacta and scattered com-
pacta. For more details on this class, we refer the reader to [8], [17] and [25].
The LP is a three-space property, i.e. if X is a Banach space and Y is a subspace
of X such that Y and X/Y have the LP, then X has the LP [24, Proposition 1.19].
This result follows from Michael's Selection Theorem. However, as far as we are
concerned, it is not known whether the WLP is a three-space property. We have a
positive result in the following case:
Theorem 2.17. Let X be a Banach space and Y a subspace of X. If Y is reflexive,
dens(Y ) < cov(M) and X/Y has the WLP, then X has the WLP.
Proof. Let Q : X → X/Y be the quotient operator and φ : X/Y → X be a norm-
norm continuous map such that Qφ = 1X/Y given by Michael's Selection Theorem
(see [10, Section 7.6]). Let f : [0, 1] → X be a Riemann integrable function. Then,
since Qf is Riemann integrable and X/Y has the WLP, there exists a set E ⊂ [0, 1]
with µ(E) = 1 such that Qf is weakly continuous at every point of E. Set
(5) C = {x ∈ X : ∃ (tn)∞
n=1 converging to some t ∈ E with x = ω- lim f (tn)}.
Let {x∗
First we are going to see that dens(C) < cov(M). Let x ∈ C and (tn)∞
n=1 as in
(5). Then Qx = ω-lim Qf (tn) = Qf (t). Therefore, x = φ(Qx) + (x − φ(Qx)) with
φ(Qx) ∈ φ(Qf (E)) and x− φ(Qx) ∈ Y . Notice that φ(Qf (E)) is separable because
of the ω-separability of Qf (E) and Mazur's Lemma. Thus, C ⊂ φ(Qf (E)) + Y
satisfies dens(C) < cov(M).
α}α∈Γ ⊂ X ∗ be a set separating points of C with Γ < cov(M). Set
E0 ⊂ E with µ(E0) = 1 such that x∗
α ◦ f is continuous at every point of E0 for
every α ∈ Γ. Notice that this can be done because the set of discontinuity points
of each x∗
α ◦ f is an Fσ Lebesgue null set and Γ < cov(M). We claim that f
is weakly continuous at each point of E0. Let t ∈ E0 and (tn)∞
n=1 be a sequence
converging to t. Since Qf (t) = ω-lim Qf (tn), the set {Qf (tn) : n ∈ N} is relatively
weakly compact in X/Y . From the reflexivity of Y , it follows that Q is a Tauberian
operator, so {f (tn) : n ∈ N} is relatively weakly compact in X (see [11, Theorem
2.1.5 and Corollary 2.2.5]). Therefore, it is enough to prove the uniqueness of the
limit of the subsequences of (f (tn))∞
f (tnk ). Then, x, f (t) ∈ C
and x∗
n=1. Let x = ω-lim
k
α(x) = lim
k
x∗
α(f (tnk )) = x∗
α(f (t)) for every α ∈ Γ, so x = f (t).
(cid:3)
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
13
3. Weak continuity does not imply integrability
It is not true that every weakly continuous function is Riemann integrable [2].
In fact, V. Kadets proved the following theorem:
Theorem 3.1 ([15]). If X is a Banach space without the Schur property, then there
is a weakly continuous function f : [0, 1] → X which is not Riemann integrable.
The proof of the previous theorem together with Josefson-Nissenzweig Theorem
(see [7, Chapter XII]) gives the following corollary:
Corollary 3.2. Given an infinite-dimensional Banach space X, there always exists
a weak* continuous function f : [0, 1] → X ∗ which is not Riemann integrable.
In [29], Wang and Yang extend the previous result to a general locally convex
topology weaker than the norm topology. In this section, we generalize these results
in Theorem 3.4.
Following the terminology used in [9], we say that a subset M of a Banach space
is spaceable if M ∪ {0} contains a closed infinite-dimensional subspace.
We start with the definitions of τ -Dunford-Pettis operator and the τ -Schur prop-
erty, that coincide with the classical definitions of Dunford-Pettis or completely
continuous operator and the Schur property when τ is the weak topology.
Definition 3.3. Let X and Y be Banach spaces and τ a locally convex topology
on X weaker than the norm topology. An operator T : X → Y is said to be τ -
Dunford-Pettis (τ -DP for short) if it carries bounded τ -null sequences to norm null
sequences. A Banach space X is said to have the τ -Schur property if the identity
operator I : X → X is τ -DP.
Theorem 3.4. Let X and Y be Banach spaces and τ be a locally convex topology
on X weaker than the norm topology. If T : X → Y is an operator which is not
τ -DP, then the family of all bounded τ -continuous functions f : [0, 1] → X such that
T f is not Riemann integrable is spaceable in ℓ∞([0, 1], X), the space of all bounded
functions from [0, 1] to X with the supremum norm.
Proof. The proof uses ideas from [15]. Since T is not τ -DP, we can take a bounded
sequence (xn)∞
n=1 that is τ -convergent to zero such that kT xnk = 1 for all n ∈ N.
Let K ⊂ [0, 1] be a copy of the Cantor set constructed by removing from [0, 1] an
2 , . . . I n
1 in the middle of [0, 1] and removing open intervals I n
open interval I 1
2n
from the middles of the remaining intervals in each step. Suppose that the removed
intervals are so small that µ(K) > 2
3 . Let Ca([0, 1]) be the closed subspace of
C([0, 1]) consisting of all continuous functions g : [0, 1] → R antisymmetric with
respect to the axe x = 1
2 and with g(0) = g(1) = 0. For every g ∈ Ca([0, 1]) and
every open interval I = (a, b) in [0, 1], we define the functions gI : [0, 1] → R and
fg : [0, 1] → X as follows
1 , I n
The function φ : Ca([0, 1]) → ℓ∞([0, 1], X) given by the formula φ(g) := fg for
every g ∈ Ca([0, 1]) is a linear map and satisfies kφ(g)k = (supn kxnk)kgk for every
g( t−a
b−a )
gI (t) =( 0
fg(t) =( 0
if t /∈ (a, b),
if t ∈ [a, b].
if t ∈ K,
if t ∈ I n
k .
gI n
k
(t)xn
14
GONZALO MART´INEZ-CERVANTES
g ∈ Ca([0, 1]). Therefore, φ is a multiple of an isometry. Thus, V := φ(Ca([0, 1]) is
an infinite-dimensional closed subspace of ℓ∞([0, 1], X).
We are going to check that each function fg 6= 0 is τ -continuous but T fg is not
τ−→ 0, fg is
Riemann integrable. Since g is continuous, g(0) = g(1) = 0 and xn
τ -continuous. Suppose T fg is Riemann integrable. Then,
y∗(cid:18)Z 1
0
T fg(t)dt(cid:19) =Z 1
0
y∗T fg(t)dt =Xk,n
y∗(T xn)ZI n
k
gI n
k
(t)dt = 0
N
for each y∗ ∈ Y ∗. The only possible value for the Riemann integral of T fg is 0
due to the above equality. Choose a partition P = {J1, J2, . . . , JN} of [0, 1]. Let
Int Jj ∩ K 6= ∅}. We can take m ∈ N such that if j ∈ A
A = {j : 1 ≤ j ≤ N,
then Jj contains some interval I m
k . Hence, if j ∈ A, there is tj ∈ Jj such that
fg(tj) = kgkxm.
If j /∈ A, then we pick any tj ∈ Int Jj. From the inequality
Pj∈A µ(Jj) ≥ µ(K) > 2
µ(Jj)T fg(tj)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)
Xj=1
kgkµ(Jj)T xm(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)(cid:13)Xj /∈A
1
3kgk.
Then, T fg is Riemann integrable if and only if g = 0 if and only if fg = 0. (cid:3)
µ(Jj)T fg(tj)(cid:13)(cid:13)(cid:13)(cid:13)
≥(cid:13)(cid:13)(cid:13)(cid:13)Xj∈A
t∈[0,1]kT fg(t)k =
sup
µ(Jj)T fg(tj ) +Xj /∈A
2
3kgk −
=(cid:13)(cid:13)(cid:13)(cid:13)Xj∈A
µ(Jj)T fg(tj)(cid:13)(cid:13)(cid:13)(cid:13)
>
1
3
3 , we deduce
≥
The next corollary gives an affirmative answer to a question posed by Sofi in
[26].
Corollary 3.5. Given an infinite-dimensional Banach space X, the set of all weak*
continuous functions f : [0, 1] → X ∗ which are not Riemann integrable is spaceable
in ℓ∞([0, 1], X ∗).
Proof. X ∗ is not ω∗-Schur for any infinite-dimensional Banach space X due to the
Josefson-Nissenzweig Theorem. Thus, the conclusion follows from Theorem 3.4. (cid:3)
Given a Banach space X, a function f : [0, 1] → X is said to be scalarly Riemann
integrable if every composition x∗f with x∗ ∈ X ∗ is Riemann integrable.
We can also characterize Dunford-Pettis operators thanks to Theorem 3.4. The
equivalence (1) ⇔ (3) in the following corollary was mentioned without proof in
[23].
Corollary 3.6. Let X and Y be Banach spaces and T : X → Y be an operator.
The following statements are equivalent:
(1) T is Dunford-Pettis.
(2) T f is Riemann integrable for every ω-continuous function f : [0, 1] → X.
(3) T f is Riemann integrable for every scalarly Riemann integrable function
f : [0, 1] → X.
Proof. (2) ⇒ (1) is a consequence of Theorem 3.4. Since every ω-continuous func-
tion f : [0, 1] → X is scalarly Riemann integrable, (3) implies (2). Therefore,
it remains to prove (1) ⇒ (3). Suppose T is Dunford-Pettis and fix (Pn)∞
n=1 a
sequence of tagged partitions of [0, 1] with kPnk n−→ 0. Let f : [0, 1] → X be
a scalarly Riemann integrable function. Then, x∗f (Pn) n−→ R 1
0 x∗f (t)dt for every
RIEMANN INTEGRABILITY VERSUS WEAK CONTINUITY
15
x∗ ∈ X ∗. Thus, f (Pn) is a ω-Cauchy sequence in X, so T f (Pn) is norm convergent
to some y ∈ Y . The limit y does not depend on the sequence of tagged partitions,
nk n−→ 0, then
n)∞
since if (P ′
n=1 is any other sequence of tagged partitions with kP ′
n)k n−→ 0.
f (Pn)−f (P ′
n) is weakly null and this in turn implies that kT f (Pn)−T f (P ′
Thus, T f is Riemann integrable.
(cid:3)
Acknowledgements
I would like to thank Antonio Avil´es and Jos´e Rodr´ıguez the useful suggestions
and the help provided in some proofs of this paper.
References
[1] F. Albiac and N. J. Kalton. Topics in Banach Space Theory. Graduate texts in Mathematics.
Springer, New York, London, 2006.
[2] A. Alexiewicz and W. Orlicz. Remarks on Riemann-integration of vector-valued functions.
Studia Math., 12:125 -- 132, 1951.
[3] T. Bartoszynski and H. Judah. Set Theory: On the Structure of the Real Line. Ak Peters
Series. Taylor & Francis, 1995.
[4] T. Bartoszynski and S. Shelah. Closed measure zero sets. Ann. Pure Appl. Logic, 58(2):93 --
110, 1992.
[5] J. M. Calabuig, J. Rodr´ıguez, and E. A. S´anchez-P´erez. Weak continuity of Riemann inte-
grable functions in Lebesgue-Bochner spaces. Acta Math. Sin. (Engl. Ser.), 26(2):241 -- 248,
2010.
[6] G. C. da Rocha Filho. Integral de Riemann Vetoral e Geometria de Espa¸cos de Banach. PhD
thesis, Universidade de Sao Paulo, 1979.
[7] J. Diestel. Sequences and series in Banach spaces. Graduate texts in mathematics. Springer-
Verlag, 1984.
[8] M. Dzamonja and K. Kunen. Properties of the class of measure separable compact spaces.
Fundam. Math., 147(3):261 -- 277, 1995.
[9] P. Enflo, V. I. Gurariy, and J. B. Seoane-Sep´ulveda. Some results and open questions on
spaceability in function spaces. Trans. Am. Math. Soc., 366(2):611 -- 625, 2014.
[10] M. Fabian, P. Habala, P. H´ajek, V. Montesinos, and V. Zizler. Banach Space Theory: The
Basis for Linear and Nonlinear Analysis. CMS Books in Mathematics. Springer New York,
2011.
[11] M. Gonz´alez and A. Mart´ınez-Abej´on. Tauberian Operators. Operator Theory: Advances and
Applications. Birkhauser Basel, 2010.
[12] R. Gordon. Riemann Integration in Banach Spaces. Rocky Mountain J. Math., 21(3):923 -- 949,
1991.
[13] L. M. Graves. Riemann integration and Taylor's theorem in general analysis. Trans. Amer.
Math. Soc., 29(1):163 -- 177, 1927.
[14] R. Haydon. Darboux integrability and separability of types in stable Banach spaces. S´emin.
analyse fonctionnelle, Paris 1983-84, Publ. Math. Univ. Paris VII 20, 95 -- 115, 1984.
[15] V. M. Kadets. On the Riemann integrability of weakly continuous functions. Quaest. Math.,
17(1):33 -- 35, 1994.
[16] H. E. Lacey. The Isometric Theory of Classical Banach Spaces. Grundlehren der mathema-
tischen Wissenschaften. Springer Berlin Heidelberg, 2012.
[17] W. Marciszewski and G. Plebanek. On measures on Rosenthal compacta. J. Math. Anal.
Appl., 385(1):185 -- 193, 2012.
[18] F. Miraglia and G. C. da Rocha Filho. The measurability of Riemann integrable functions
with values in Banach spaces and applications. Sao Paulo Ime/usp, 1984.
[19] K. M. Naralenkov. Asymptotic Structure of Banach Spaces and Riemann Integration. Real
Anal. Exchange, 33(1):113 -- 126, 2007.
[20] A. S. Nemirovski, M. Ju. Ochan, and R. Redjouani. Conditions for the Riemann integrability
of functions with values in a Banach space. Mosc. Univ. Math. Bull., 27(3-4):124 -- 126, 1973.
[21] E. Odell. A nonseparable Banach space not containing a subsymmetric basic sequence. Israel
J. Math., 52(1 -- 2):97 -- 109, 1985.
16
GONZALO MART´INEZ-CERVANTES
[22] J. Pawlikowski. A characterization of strong measure zero sets. Israel J. Math., 93(1):171 -- 183,
1996.
[23] A. Pelczynski and G. C. da Rocha Filho. Operadores de Darboux. Seminario Brasileiro de
Analise, 12, Sao Jose dos Campos, 1980. Trabalhos Apresentados. Rio de Janeiro, SBM, 1980.
[24] M. Pizzotti. Darboux-integrabilidade e mensurabilidade de fun¸coes Riemann-integr´aveis
definidas em compactos. PhD thesis, Universidade de Sao Paulo, 1989.
[25] G. Plebanek and D. Sobota. Countable tightness in the spaces of regular probability measures.
Fundam. Math., 229(2):159 -- 169, 2015.
[26] M. A. Sofi. Weaker forms of continuity and vector-valued Riemann integration. Colloq. Math.,
129(1):1 -- 6, 2012.
[27] C. Wang. On the weak property of Lebesgue of Banach spaces. Journal of Nanjing University
Mathematical Biquarterly (English Series), 13(2):150 -- 155, 1996.
[28] C. Wang and K. Wan. On the Weak Property of Lebesgue of L1(Ω, Σ, µ). Rocky Mountain
J. Math., 31(2):697 -- 703, 2001.
[29] C. Wang and Z. Yang. Some Topological Properties of Banach Spaces and Riemann Integra-
tion. Rocky Mountain J. Math., 30(1):393 -- 400, 2000.
Departamento de Matem´aticas, Facultad de Matem´aticas, Universidad de Murcia,
30100 Espinardo (Murcia), Spain
E-mail address: [email protected]
|
1512.08741 | 2 | 1512 | 2016-09-30T14:33:08 | Ideal structures in vector-valued polynomial spaces | [
"math.FA"
] | This paper is concerned with the study of geometric structures in spaces of polynomials. More precisely, we discuss for $E$ and $F$ Banach spaces, whether the class of weakly continuous on bounded sets $n$-homogeneous polynomials, $\mathcal P_w(^n E, F)$, is an HB-subspace or an $M(1,C)$-ideal in the space of continuous $n$-homogeneous polynomials, $\mathcal P(^n E, F)$. We establish sufficient conditions under which the problem can be positively solved. Some examples are given. We also study when some ideal structures pass from $\mathcal P_w(^n E, F)$ as an ideal in $\mathcal P(^n E, F)$ to the range space $F$ as an ideal in its bidual $F^{**}$. | math.FA | math | IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL
SPACES
VER ´ONICA DIMANT1, SILVIA LASSALLE1, ´ANGELES PRIETO2
Abstract. This paper is concerned with the study of geometric structures in
spaces of polynomials. More precisely, we discuss for E and F Banach spaces,
whether the class of n-homogeneous polynomials, Pw(nE, F ), which are weakly
continuous on bounded sets, is an HB -subspace or an M (1, C)-ideal in the space
of continuous n-homogeneous polynomials, P(nE, F ). We establish sufficient
conditions under which the problem can be positively solved. Some examples
are given. We also study when some ideal structures pass from Pw(nE, F ) as
an ideal in P(nE, F ) to the range space F as an ideal in its bidual F ∗∗.
6
1
0
2
p
e
S
0
3
]
.
A
F
h
t
a
m
[
2
v
1
4
7
8
0
.
2
1
5
1
:
v
i
X
r
a
1. Introduction
Let X be a (real or complex) Banach space and let J be a closed subspace of
X. According to the Hahn -- Banach theorem, every continuous linear functional
g ∈ J ∗ has an extension f ∈ X ∗, with the same norm. A long standing problem is
to determine when every functional on J has a unique norm-preserving extension
to X. This question is closely related to geometric properties of both spaces
which, in many cases, imply the existence of a norm-one projection on X ∗ whose
kernel is J ⊥ := {x∗ ∈ X ∗ : x∗(y) = 0, for all y ∈ J}, the annihilator of J. When
there exists such a projection J is said to be an ideal in X. A canonical example
of this fact is that X is always an ideal in its bidual X ∗∗.
The notion of M-ideal, introduced by Alfsen and Effros and widely studied
in the book by Harmand, Werner and Werner [18] is one of these geometric
properties ensuring unique Hahn -- Banach extensions. Recall that J is an M -
ideal in X if it is an ideal in X with associated projection q such that for each
f ∈ X ∗ one has
kf k = kqf k + kf − qf k.
The fact that J is an M-ideal in X has a strong impact on both J and X, and
sometimes seems to be too restrictive. So, we will be interested in studying some
weaker properties among those implying unique norm-preserving extensions.
Recall that a closed subspace J is HB -smooth in X if every element in J ∗ has
a unique norm-preserving extension to an element in X ∗. A closed subspace J is
strongly HB -smooth in X if there exists a linear projection q on X ∗ whose kernel
is J ⊥ such that for each f ∈ X ∗ with f 6= qf one has
kqf k < kf k.
2010 Mathematics Subject Classification. Primary 46G25; Secondary 47H60, 46B04, 47L22.
Key words and phrases. HB -subspaces, homogeneous polynomials, weakly continuous on
bounded sets polynomials.
1
2
V. DIMANT, S. LASSALLE, A. PRIETO
The interplay between uniqueness of the extension and strong HB -smoothness
was clarified by Oja [22]. Namely, the uniqueness of the extensions and being an
ideal are independent notions for a subspace J, strong HB -smoothness implies
both, and if J is an HB -smooth ideal in X then J is strongly HB -smooth in X.
A particular case of HB -smoothness is the notion of HB -subspace, introduced
by Hennefeld [19]. A closed subspace J is an HB-subspace of X if there exists a
projection q on X ∗ whose kernel is J ⊥ such that for each f ∈ X ∗ with f 6= qf
one has
kqf k < kf k
and
kf − qf k ≤ kf k.
Finally, given C ∈ (0, 1], a closed subspace J is an M(1, C)-ideal in X if J is
an ideal of X with associated projection q on X ∗ such that for each f ∈ X ∗ one
has
kqf k + Ckf − qf k ≤ kf k.
The last inequality is called the M(1, C)-inequality. Note that when C = 1, the
notion of M-ideal is covered and to be M(1, C)-ideal immediately implies strong
HB -smoothness. However, the notions of M(1, C)-ideal and HB -subspace are
independent. On the one hand, Cabello and Nieto [6, Example 3.7] showed that
if X is a nonreflexive separable M-ideal in its bidual, then ℓp(X) as a subspace
of its bidual, 1 < p < ∞, is an HB -subspace that cannot be renormed to be an
M(1, C)-ideal for any 0 < C < 1. On the other hand, Cabello, Nieto and Oja [7,
Example 4.3] showed that for any 0 < C < 1, there is a renorming of c0, c0 due
to Johnson and Wolfe, such that the space of compact operators on c0, K(c0) is
an M(1, C)-ideal in the space of all continuous operators L(c0) without being an
HB -subspace.
Several authors have been interested in this kind of properties for arbitrary
subspaces of Banach spaces, and also for distinguished particular cases. The space
of compact operators K(E, F ) between Banach spaces E and F as a subspace of
the space of all continuous linear operators L(E, F ) received special interest (see,
for example, [6, 7, 19, 21, 22, 23, 24, 26]). The strongest of the abovementioned
properties is the one of being an M-ideal. All other properties which are more
flexible still allow us to deal with uniqueness of Hahn -- Banach extensions.
Here, we will be concerned with P(nE, F ), the space of continuous n-homogen-
eous polynomials between Banach spaces E and F . In the polynomial context,
the space of compact mappings is usually replaced by Pw(nE, F ) the subspace of
homogeneous polynomials which are weakly continuous on bounded sets. Recall
that a polynomial P ∈ P(nE, F ) is in Pw(nE, F ) if it maps bounded weakly
convergent nets into convergent nets. Note that we could have considered poly-
nomials in P(nE, F ) mapping bounded sets into relatively compact sets, which
are called compact polynomials. For linear operators to be compact and to be
weakly continuous on bounded sets are equivalent notions. For n-homogeneous
polynomials with n > 1, every polynomial in Pw(nE, F ) is compact (as can be
derived from results in [3] and [4]), but the converse might not be true. Every
scalar-valued continuous polynomial is compact but it is not necessarily weakly
continuous on bounded sets. The prototypical example of this situation is given
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
3
by P (x) = Pk x2
k, for all x = (xk)k ∈ ℓ2. Therefore, we will focus our atten-
tion on determining the presence of ideal structures for Pw(nE, F ) as a subspace
of P(nE, F ). To be more precise, our main concern is to study the notion of
HB -subspace in the polynomial setting.
Some previous results in this direction can be found in [12], where the problem
of determining when Pw(nE) is an M-ideal in P(nE) was considered. A vector-
valued approach of the same question was treated in [14]. Note that the searching
of ideal structures for Pw(nE, F ) as a subspace of P(nE, F ) makes sense when
the spaces Pw(nE, F ) and P(nE, F ) do not coincide. The equality Pw(nE, F ) =
P(nE, F ) is a long standing nontrivial problem, considered for instance in [1, 5,
16, 17].
The plan of the paper is as follows. First, we review the notation and the basic
facts that will be used in Sections 3 and 4. Then, in Section 3, we investigate
sufficient conditions under which the subspace Pw(nE, F ) enjoys an additional
geometric structure inside P(nE, F ) and we exhibit some particular examples.
In the last section, we study some ideal structures for the range space F as a
subspace of F ∗∗, when they are fulfilled by Pw(nE, F ) as a subspace of P(nE, F ).
2. Notation and basic facts
Before proceeding, we fix some notation. Every time we write E or F we will
be considering Banach spaces over the real or complex field, K. The closed unit
ball of E will be denoted by BE and the unit sphere by SE. As usual, E∗ and
E∗∗ stand for the dual and bidual of E, respectively. The space of linear bounded
operators from E to F will be denoted by L(E, F ) (and L(E) when E = F );
its subspace of compact mappings will be denoted by K(E, F ) (K(E) in the case
E = F ).
A function P : E → F is an n-homogeneous polynomial if there exists a
(unique) symmetric n-linear form A : E × · · · × E
→ F such that
n
{z
}
P (x) = A(x, . . . , x),
for all x ∈ E. The space of all continuous n-homogeneous polynomials from E to
F , P(nE, F ), endowed with the supremum norm
kP k = sup{kP (x)k : x ∈ BE},
is a Banach space.
L(b⊗
Every polynomial P in P(nE; F ) can be associated with a linear operator in
n,s
πs E; F ), where πs is the symmetric projective tensor norm. We will identify
P with its linearization without further mention. Even though this identification
preserves the norm, there is no Hahn -- Banach Theorem for homogeneous poly-
nomials of degree 2 or greater. However, Aron and Berner [2] and Davie and
Gamelin [9] showed that for every P ∈ P(nE, F ) there is a norm-preserving ex-
tension of P to P ∈ P(nE∗∗, F ∗∗) such that P (x) = P (x) for all x ∈ E. The
construction of this canonical extension is based on the Arens extension of the
4
V. DIMANT, S. LASSALLE, A. PRIETO
symmetric mapping A associated to the polynomial P . To obtain the Arens ex-
tension, we simply extend by weak-star continuity, one variable at a time, the n
variables of A. This process depends on the order that the variables are extended
and the final result might not be a symmetric mapping. However, the n! possible
extensions coincide on the diagonal and P is well defined. For the particular case
in which P belongs to Pw(nE, F ), the range of P is also in F (as can be derived
from [3] and [8, Proposition 2.5]). This fact will be used repeatedly in Section 4.
In this paper, we will present several results in which at least one of the spaces
involved enjoys the metric compact approximation property. Recall that a Ba-
nach space E has the metric compact approximation property if there is a net of
compact operators (Kα) on E such that Kα → IdE pointwise and supα kKαk ≤ 1.
Usually, the net (Kα) is called a metric compact approximation of the identity.
If in addition K ∗
α → IdE ∗ pointwise, the net (Kα) is called a shrinking met-
ric compact approximation of the identity. As usual, K α denotes the operator
IdE − Kα. For dual spaces we have the following intermediate property. We say
that E∗ has a metric compact approximation of the identity with adjoint operators
if there exists a net (Kα) ⊂ K(E) such that K ∗
α converges to IdE ∗ pointwise and
supα kKαk ≤ 1. These notions are closely related with ideal structures on Banach
spaces. For instance, [21, Theorem 1.1] asserts that the following conditions are
equivalent:
(i) F has the metric compact approximation property
(ii) K(E, F ) is an ideal in L(E, F ) for every Banach space E.
So, it is natural to expect that the metric compact approximation property
shows up when describing Pw(nE, F ) as an ideal in P(nE, F ).
One further ingredient will appear in our discussion. In [10], Delpech obtained
an appropriate connection between the moduli of asymptotic uniform smoothness
and convexity and weak sequential continuity of polynomials. In [13], Dimant,
Gonzalo and Jaramillo followed his approach to obtain results on compactness
or weak-sequential continuity of multilinear mappings. Here, we will impose
restrictions on the growth of the moduli of the underlying spaces E or F to ensure
that Pw(nE, F ) enjoys an appropriate property in P(nE, F ) (see Theorem 3.4 and
Proposition 3.13). Some definitions are in order.
For an infinite dimensional Banach space E, the modulus of asymptotic point-
wise smoothness is defined for kxk = 1 and t > 0 by
ρE(t; x) =
inf
dim(E/H)<∞
sup
kx + hk − 1,
h∈H,khk≤t
and the modulus of asymptotic uniform smoothness is defined for t > 0 by
ρE(t) = sup
kxk=1
ρE(t; x).
The space E is asymptotically uniformly smooth if limt→0
= 0.
ρE(t)
t
For an infinite dimensional Banach space, the modulus of asymptotic pointwise
convexity is defined for kxk = 1 and t > 0 by
δE(t; x) =
sup
dim(E/H)<∞
inf
h∈H,khk≥t
kx + hk − 1,
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
5
and the modulus of asymptotic uniform convexity is defined for t > 0 by
δE(t) = inf
kxk=1
δE(t; x).
The space E is asymptotically uniformly convex if δE(t) > 0, for every 0 < t ≤ 1.
Finally, E has modulus of asymptotic uniform convexity of power p if there exists
C > 0 such that δE(t) ≥ Ctp, for all 0 < t ≤ 1.
We refer to [15] for the necessary background on polynomials on Banach spaces.
3. Sufficient conditions
When working with polynomials, the lack of linearity provides, in many cases,
difficulties that can be overcome not without certain detours. The value of n for
which Pw(nE, F ) has the chance to be a nontrivial M-ideal in P(nE, F ) cannot
be chosen arbitrarily. In fact, in the scalar-valued case [12] it was proved that,
whenever P(mE) \ Pw(mE) 6= ∅ for some m, there exists a unique value n, called
the critical degree, for which Pw(nE) can be a non-trivial M-ideal in P(nE). The
critical degree of E is defined as
cd(E) := min{k ∈ N : Pw(kE) 6= P(kE)}.
In the vector-valued case, the critical degree is defined by analogy [14] as
cd(E, F ) := min{k ∈ N : Pw(kE, F ) 6= P(kE, F )},
and the problem of whether Pw(nE, F ) is an M-ideal in P(nE, F ) is worth being
studied only for polynomials of degree n with cd(E, F ) ≤ n ≤ cd(E). Although
we are interested in studying ideal structures which are more flexible than to
be an M-ideal, in order to show positive results we could not get rid of some
restrictions on the degree of homogeneity. We start with a lemma which, under
certain conditions on n, gives a version of Johnson's projection [20, Lemma 1.1]
for the polynomial case.
Lemma 3.1. Let E, F be Banach spaces and let n < cd(E). Suppose that F
has the metric compact approximation property. Then Pw(nE, F ) is an ideal in
P(nE, F ).
Proof. Let (Lβ) ⊂ K(F ) be a metric compact approximation of the identity. As
(Lβ) is bounded, there exists a subnet, still denoted by (Lβ), that converges w∗
to some L0 ∈ K(F )∗∗. Define Λ : P(nE, F )∗ → P(nE, F )∗ by
Λ(f )(P ) = lim
β
f (Lβ ◦ P ).
(3.1)
Note that Λ is well defined. In fact, if P ∈ P(nE, F ) and τP : K(F ) → P(nE, F )
is the composition operator, τP (K) = K ◦ P , its transpose τ ∗
P satisfies that
τ ∗
P (f ) ∈ K(F )∗, for any f ∈ P(nE, F )∗. So
lim
β
f (Lβ ◦ P ) = lim
β
τ ∗
P (f )(Lβ) = L0(τ ∗
P (f )).
It is clear that Λ is linear and kΛk ≤ 1.
It is also a projection: since any
P ∈ Pw(nE, F ) is compact and (Lβ) converges to the identity on compact sets
6
V. DIMANT, S. LASSALLE, A. PRIETO
we see that limβ Lβ ◦ P = P . Thus,
Λ(f )(P ) = lim
β
f (Lβ ◦ P ) = f (P ),
for P ∈ Pw(nE, F ). Now, by [14, Lemma 1.8], as n < cd(E), the net of polyno-
mials (Lβ ◦ Q) belongs to Pw(nE, F ), for every Q ∈ P(nE, F ). Hence,
f (Lβ ◦ Q) = Λ(f )(Q),
Λ(Λ(f ))(Q) = lim
β
Λ(f )(Lβ ◦ Q) = lim
β
and Λ2 = Λ. Finally, it is easy to check that ker Λ = Pw(nE, F )⊥. Then, Λ is a
norm one projection on P(nE, F )∗ with ker Λ = Pw(nE, F )⊥.
(cid:3)
Remark 3.2. Every time Pw(nE, F ) is an ideal in P(nE, F ) with associated pro-
jection Λ, we have the decomposition
P(nE, F )∗ = Pw(nE, F )∗ ⊕ Pw(nE, F )⊥,
and any f ∈ P(nE, F )∗ has a unique representation such that
f = g + h, with g = Λ(f ) ∈ Pw(nE, F )∗,
Now, if F has a metric compact approximation of the identity (Lβ) ⊂ K(F ),
we may (and will) suppose that (Lβ) is w∗-convergent in K(F )∗∗ and that the
projection Λ is defined as in (3.1), Λ(f )(P ) = limβ f (Lβ ◦ P ). Then, with Lβ =
IdF − Lβ, we have the following facts that were already used:
h = f − g ∈ Pw(nE, F )⊥.
(3.2)
• limβ Λ(f )(Lβ ◦ P ) = 0, for all f ∈ P(nE, F )∗ and all P ∈ P(nE, F ).
• limβ Lβ ◦ Q = Q, for all Q ∈ Pw(nE, F ).
Indeed, the first assertion follows by [14, Lemma 1.8]. For the second one, note
that Lβ converges uniformly to the identity on compact sets.
Finally, note that since kΛk ≤ 1, we automatically have kgk ≤ kf k.
If in
addition kLβk ≤ 1, we also obtain khk ≤ kf k since h(P ) = (f − Λf )(P ) =
limβ f (Lβ ◦ P ) for all P ∈ P(nE, F ).
As mentioned before, Delpech used the moduli of asymptotic uniform convex-
ity and smoothness of a Banach space to obtain properties of weak sequential
continuity of polynomials. Dimant, Gonzalo and Jaramillo [13] showed the con-
nection between these moduli and compactness or weak sequential continuity of
multilinear mappings. The moduli play their role when dealing with Pw(nE, F )
as an HB -subspace of P(nE, F ). We present a refinement of [10, Lemma 10.3],
that will be used in the sequel.
Lemma 3.3. Let E be an infinite dimensional Banach space and let (wα) ⊂ E
be a weakly null bounded net.
(a) If x ∈ BE, then limαkx + wαk ≤ 1 + ρE(limαkwαk).
(b) If (xα) ⊂ BE is a net contained in a compact set, then
limαkxα + wαk ≤ 1 + ρE(limαkwαk).
Proof. To prove (a) first note that [10, Lemma 10.3] remains valid if we consider
weakly null nets instead of weakly null sequences. That is, limαkx + wαk ≤
1 + ρE(limαkwαk), for any x ∈ SE. Now, fix a nonzero x ∈ BE and consider each
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
7
x + wα as a convex combination of
x
kxk
+ wα and
−x
kxk
+ wα. Applying the above
inequality to
±x
kxk
we get
limα(cid:13)(cid:13)(cid:13)(cid:13)
±x
kxk
+ wα(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1 + ρE(limαkwαk),
and the statement follows.
Now, suppose that (b) does not hold. Then, we may find subnets (xβ), (wβ),
and x0 ∈ BE, so that limβ xβ = x0 and limβkxβ + wβk > 1 + ρE(limαkwαk). As
ρE is increasing, limβkxβ + wβk > 1 + ρE(limβkwβk). Note that for any subnet
(βi) such that limi kxβi + wβik exists, so too does the limit limi kx0 + wβik, and
both coincide. This implies that limβkx0 + wβk = limβkxβ + wβk. It follows that
limβkx0 + wβk > 1 + ρE(limβkwβk), which contradicts (a).
(cid:3)
We are ready to describe Pw(nE, F ) as an HB -subspace of P(nE, F ), under
certain conditions on F and n.
Theorem 3.4. Let E be a Banach space, and let n < cd(E). Let F be an infinite
dimensional Banach space with a shrinking metric compact approximation of the
identity (Lβ) ⊂ K(F ) such that supβ kLβk ≤ 1 and suppose that F is asymptoti-
cally uniformly smooth. Then, Pw(nE, F ) is an HB-subspace of P(nE, F ).
Proof. Consider the projection Λ, given in (3.1), under which Pw(nE, F ) is an
ideal in P(nE, F ). For any f ∈ P(nE, F )∗, write f = g + h as in (3.2). Then,
as commented in Remark 3.2, kgk ≤ kf k and khk ≤ kf k. In order to finish, we
have to prove that kgk < kf k, for h 6= 0.
Fix ε > 0 and take P ∈ P(nE, F ) and Q ∈ Pw(nE, F ) such that
kP k = kQk = 1,
h(P ) > khk − ε
and g(Q) > kgk − ε.
Since limβ Lβ ◦ Q = Q, we can choose β0 satisfying g(Lβ0 ◦ Q) > kgk − 2ε.
Change, if necessary, Q to λQ (with λ = 1) to obtain g(Lβ0 ◦ Q) > kgk − 2ε.
For t > 0, consider Lβ0 ◦ Q + tLβ ◦ P and take a net (xβ) ⊂ BE with
limβkLβ0 ◦ Q + tLβ ◦ P kP(nE,F ) = limβk(Lβ0 ◦ Q)(xβ) + t(Lβ ◦ P )(xβ)kF .
Now, note that ((Lβ ◦ P )(xβ)) is weakly null. Indeed, (P (xβ)) is bounded and for
any y∗ ∈ F ∗, limβ(Lβ)∗y∗ = 0. Then,
lim
β
h(Lβ ◦ P )(xβ), y∗i = lim
β
hP (xβ), (Lβ)∗y∗i = 0
for any y∗ ∈ F ∗. On the other hand, the compact set Lβ0(BF ) ⊂ BF contains the
net ((Lβ0 ◦ Q)(xβ)). Therefore, Lemma 3.3 can be applied to get
limβkLβ0 ◦ Q + tLβ ◦ P kP(nE,F ) ≤ 1 + ρF (limβkt(Lβ ◦ P )(xβ)kF ) ≤ 1 + ρF (t).
As observed in Remark 3.2, limβ g(Lβ ◦ P ) = 0. Then, g(Lβ ◦ P ) < ε, for
β ≥ β1. Also, h(Lβ ◦ P ) > khk − ε, since Lβ ◦ P belongs to Pw(nE, F ) and
h(P ) = h(Lβ ◦ P ).
8
V. DIMANT, S. LASSALLE, A. PRIETO
Combining the previous estimates, we conclude for t > 0 and β ≥ β1 that
(kgk − 2ε) − tε + t(khk − ε) < g(Lβ0 ◦ Q) − tg(Lβ ◦ P ) + th(Lβ ◦ P )
≤ f (Lβ0 ◦ Q + tLβ ◦ P ).
Then, for t > 0 and ε > 0,
kgk + tkhk − 2ε(1 + t) ≤ limβf (Lβ0 ◦ Q + tLβ ◦ P )
≤ kf k limβkLβ0 ◦ Q + tLβ ◦ P k
≤ kf k(1 + ρF (t)).
Thus, kgk + tkhk ≤ kf k(1 + ρF (t)). Now, suppose that kgk = kf k, then tkhk ≤
kf kρF (t), for t > 0. Since F is asymptotically smooth, limt→0
h = 0, which completes the proof.
= 0 and
(cid:3)
ρF (t)
t
Our next result gives another set of sufficient conditions, also related with the
notion of smoothness, under which Pw(nE, F ) is an HB -subspace in P(nE, F ).
It is reminiscent of [23, Theorem 1]. The proof is similar to that of the above
theorem and we omit it.
Theorem 3.5. Let E, F be Banach spaces, and let n < cd(E). Suppose that
there exists a metric compact approximation of the identity (Lβ) ⊂ K(F ) such
that supβ kLβk ≤ 1 and for any ε > 0 there exist µ > 0 and β0 so that
sup
kLβy + µLβzk ≤ 1 + εµ , for all β ≥ β0.
(3.3)
kyk,kzk≤1
Then, Pw(nE, F ) is an HB-subspace of P(nE, F ).
Following [19, Definition 3.6] we say that a Schauder basis of a Banach space is
uniformly smooth if for every ε > 0 there exists δ > 0 such that kx+yk+kx−yk <
2 + εkyk, whenever x and y have disjoint supports with respect to the basis,
kxk = 1 and kyk < δ. Note that by using convex combinations, the definition
can be restated for x with kxk ≤ 1. The next result should be compared with
[19, Theorem 3.7].
Corollary 3.6. Let E, F be Banach spaces, and let n < cd(E). Suppose that F
has a uniformly smooth 1-unconditional basis. Then Pw(nE, F ) is an HB-subspace
of P(nE, F ).
Proof. Let ΠN be the natural projection on F onto the subspace generated by
the first N elements of the 1-unconditional basis. Then, (ΠN ) is a metric com-
pact approximation of the identity and satisfies supN kΠN k ≤ 1. Also, (3.3) in
Theorem 3.5 holds. Indeed, take ε > 0 and consider µ = δ as in the definition
of uniform smoothness of the basis. For any y, z ∈ BF and N ∈ N, being the
basis 1-unconditional, we have kΠN y + µΠN zk ≤ 1 + µε/2. Thus, an immediate
application of Theorem 3.5 gives the result.
(cid:3)
The above corollary can be applied to show some examples of HB -subspaces of
polynomials where the spaces ℓp and the Lorentz sequence spaces d(w, p) appear.
Recall that, for 1 < p < ∞, both ℓp and d(w, p) have uniformly smooth 1-
unconditional bases. Also, the critical degree of ℓp is the integer number satisfying
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
9
p ≤ cd(ℓp) < p + 1 and cd(ℓp, ℓq) is the integer satisfying p
q + 1.
For the case of cd(ℓp, d(w, q)), a restatement of (I) and (II) in [14, p. 705] reads
as
q ≤ cd(ℓp, ℓq) < p
cd(ℓp, d(w, q)) = max{k ∈ N : k <
+ 1 and w 6∈ ℓ(
p
(k−1)q )∗}.
p
q
Example 3.7. Let 1 < p, q < +∞.
(a) Let E be a Banach space, and let n < cd(E). Then,
• Pw(nE, ℓq) is an HB -subspace of P(nE, ℓq).
• Pw(nE, d(w, q)) is an HB -subspace of P(nE, d(w, q)).
(b) Let cd(ℓp, ℓq) < n < cd(ℓp). Then, Pw(nℓp, ℓq) is an HB -subspace but not an
M-ideal in P(nℓp, ℓq).
(c) Let cd(ℓp, d(w, q)) < n < cd(ℓp). Then, Pw(nℓp, d(w, q)) is an HB -subspace but
not an M-ideal in P(nℓp, d(w, q)). The same result holds for n = cd(ℓp, d(w, q))
for the case cd(ℓp, d(w, q)) < p
q .
The statements about not being M-ideals in the previous examples are proved
in [14, Theorems 3.2 and 3.9].
Now, we consider conditions satisfied by the domain space E so that we also
have geometric structures in P(nE; F ). Similarly to what happens in Lemma 3.1,
we will describe Pw(nE, F ) as an ideal of P(nE, F ) whenever E∗ has a metric
compact approximation of the identity with adjoint operators. Here, no restric-
tions on the degree of the polynomials are imposed.
Lemma 3.8. Let E, F be Banach spaces such that E∗ has a metric compact
approximation of the identity with adjoint operators. Then, Pw(nE, F ) is an
ideal of P(nE, F ) for all n ∈ N.
Proof. Let (Kα) ⊂ K(E) be a net satisfying limα K ∗
αx∗ = x∗, for all x∗ ∈ E∗
and supα kKαk ≤ 1. Without loss of generality, we may assume that (Kα) is
weak∗-convergent in K(E)∗∗. Therefore, as in Lemma 3.1, the mapping
Λ : P(nE, F )∗ → P(nE, F )∗
given by
Λ(f )(P ) = lim
α
f (P ◦ Kα),
(3.4)
is well defined. It is clear that Λ is linear and kΛk ≤ 1. By [14, Lemma 2.1],
limα kP − P ◦ Kαk = 0 for every P ∈ Pw(nE, F ). Then, Λ(f )(P ) = f (P ), for all
P ∈ Pw(nE, F ). Furthermore, Λ is a projection: for all Q ∈ P(nE, F ), (Q ◦ Kα)
belongs to Pw(nE, F ). Thus, for all Q ∈ P(nE, F ),
Λ(Λ(f ))(Q) = lim
α
Λ(f )(Q ◦ Kα) = lim
α
f (Q ◦ Kα) = Λ(f )(Q),
and Λ2 = Λ. It is easy to check that ker Λ = Pw(nE, F )⊥ and the result follows.
(cid:3)
The next result gives a sufficient condition to obtain the dual space Pw(nE, F )∗
as a quotient. We denote by π the projective tensor norm. Recall that P denotes
the canonical extension of P in P(nE, F ) to P(nE∗∗, F ∗∗).
10
V. DIMANT, S. LASSALLE, A. PRIETO
Proposition 3.9. Let E, F be Banach spaces such that Pw(nE, F ) does not con-
n,s
elementary tensor u ⊗ y∗ by j(u ⊗ y∗)(P ) = y∗(P (u)), is a quotient mapping.
tain ℓ1. Then, the application j : b⊗
Proof. Take v ∈ b⊗
with ui ∈ b⊗
j(Xi
n,s
πs E∗∗ and y∗
ui ⊗ y∗
i )(P ) = Xi
πs E∗∗b⊗πF ∗ → Pw(nE, F )∗, given on any
πs E∗∗b⊗πF ∗. For each representation of v of the formPi ui⊗y∗
i ∈ F ∗ for all i, we have
kuikky∗
i k.
y∗
i (P (ui)) ≤ kP kXi
n,s
i ,
So, j is continuous and kjk = 1. Using Haydon's characterization of spaces not
containing ℓ1, we may write the unit ball of Pw(nE, F )∗ as the closed convex
hull of its extreme points. Now, by [14, Proposition 1.2], with ez(P ) = P (z) for
z ∈ E∗∗, we obtain
BPw(nE,F )∗ = Γ(ExtPw(nE,F )∗) ⊂ Γ(ez ⊗ y∗ : z ∈ SE ∗∗, y∗ ∈ SF ∗)
⊂ j(B b⊗
n,s
πs E ∗∗ b⊗πF ∗) ⊂ BPw(nE,F )∗.
Then, all the inclusions are (actually) equalities and j is a quotient mapping. (cid:3)
In the next result we show that the natural hypothesis on E and F guarantee
that Pw(nE, F ) does not contain ℓ1, and the above proposition can be applied.
We will appeal to the result by Stegall which asserts that if a Banach space E
has a separable subspace whose dual is nonseparable, then E∗ lacks the Radon --
Nikod´ym property, see for instance [11, Theorem VII.2.6].
Proposition 3.10. Let E, F be Banach spaces such that E∗∗ and F ∗ have the
Radon -- Nikod´ym property. Then, Pw(nE, F )∗ has the Radon -- Nikod´ym property
and hence Pw(nE, F ) does not contain ℓ1 for all n ∈ N.
Proof. For any P ∈ Pw(nE, F ) we consider its associated symmetric multilinear
map A and define TP ∈ L(E, Pw(n−1E, F )) as the operator given by TP (x)(x) =
A(x, x, . . . , x). By [3, Theorem 2.9], TP is a well-defined compact operator and
kP k ≤ kTP k ≤ ekP k. Then, the mapping Φ : Pw(nE, F ) → K(E, Pw(n−1E, F ))
given by Φ(P ) = TP is an isomorphism with its image. As the Radon -- Nikod´ym
property is preserved by isomorphisms, induction on n and [25, Theorem 1.9]
yield the result.
(cid:3)
Lemma 3.11. Let E, F be Banach spaces such that Pw(nE, F ) does not contain ℓ1
and such that n < cd(E). Suppose that E∗ has a metric compact approximation of
the identity with adjoint operators given by (Kα) ⊂ K(E). Then, limα P ◦ K α = 0
in the topology σ(P(nE, F ), Pw(nE, F )∗), for any P ∈ P(nE, F ).
n,s
fined by j(u ⊗ y∗)(P ) = y∗(P (u)) is a quotient mapping. We will show for
i that limαhu, P ◦ K αi = 0. So, as for any
n,s
i=1 wi ⊗ wi ⊗ · · · ⊗ wi ⊗ y∗
Proof. By Proposition 3.9, the application j : b⊗
u = PM
h ∈ Pw(nE, F )∗ there exists u ∈ b⊗
K αi = PM
i ) = PM
πs E∗∗b⊗πF ∗ → Pw(nE, F )∗, de-
πs E∗∗b⊗πF ∗ so that j(u) = h and u can be
approximated by such u's, the result follows. Take u as above. Then hu, P ◦
i ). As n < cd(E), P is
w∗ − w∗ continuous and, since limα(K α)∗∗(wi) = 0 in the w∗-topology, we obtain
limαhu, P ◦ K αi = 0, which completes the proof.
(cid:3)
i=1 P ((K α)∗∗(wi))(y∗
i=1 P ◦ K α(wi)(y∗
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
11
Proposition 1.4 in [13] provides an appropriate equivalence of asymptotic uni-
form convexity of power p. With a slight modification of its proof we drop the
hypothesis of separability and obtain a refinement, analogous to condition (c) in
Lemma 3.3, as follows.
Lemma 3.12. Let E be an infinite dimensional Banach space and let 1 < p < ∞.
The following statements are equivalent.
(a) E has modulus of asymptotic uniform convexity of power p.
(b) There exists a constant C > 0 such that for every x ∈ SE and every bounded
weakly null net (wα) in E, we have
limαkx + wαkp ≥ 1 + C limαkwαkp.
(c) There exists a constant C > 0 such that for every net (xα) in a compact set
of BE and every bounded weakly null net (wα) in E, we have
limαkxα + wαkp ≥ limαkxαkp + C kwαkp.
Proposition 3.13. Let E, F be Banach spaces such that E∗∗ and F ∗ enjoy the
Radon -- Nikod´ym property. Suppose that E has modulus of asymptotic uniform
convexity of power n = cd(E, F ), with n < cd(E), and suppose that E has a
shrinking metric compact approximation of the identity (Kα) ⊂ K(E) satisfying
supα kK αk ≤ 1 and limαkIdE − 2Kαk ≤ 1.
Then, there exists C > 0 such that Pw(nE, F ) is an M(1, C)-ideal of P(nE, F ).
Furthermore, Pw(nE, F ) is an HB-subspace of P(nE, F ).
Proof. We proceed as in Theorem 3.4. Consider Λ as in (3.4), under which
Pw(nE, F ) is an ideal in P(nE, F ) and write f ∈ P(nE, F )∗, f = g + h as
in (3.2), where kgk = kΛ(f )k ≤ kf k.
Consider ε > 0 and take P ∈ P(nE, F ), Q ∈ Pw(nE, F ) with
kP k = kQk = 1,
h(P ) ≥ khk − ε
and g(Q) ≥ kgk − ε.
For any α, the polynomial P − P ◦ K α is weakly continuous on bounded sets
at 0 (see, for instance, the proof of [12, Proposition 2.2]). Since n = cd(E, F ),
the net (P − P ◦ K α)α is in Pw(nE, F ). Then, with h ∈ Pw(nE, F )⊥, h(P ) =
h(P ◦ K α) ≥ khk − ε, for all α. Also, as limα Q ◦ Kα = Q, there exists α0 so that
g(Q◦Kα) > kgk−2ε and kIdE −2Kα0k < 1+ε, for all α ≥ α0. Changing Q◦Kα0
to λQ◦ Kα0 with λ = 1, if necessary, we may assume that g(Q◦ Kα0) > kgk −2ε.
Now, with C > 0 (to be fixed later) we have
kf k kQ ◦ Kα0 + CP ◦ K αk ≥ f (Q ◦ Kα0 + CP ◦ K α)
≥ g(Q ◦ Kα0) + Ch(P ◦ K α) − Cg(P ◦ K α)
≥ kgk + Ckhk − (2 + C)ε − Cg(P ◦ K α).
By Lemma 3.11, we may find α1 ≥ α0 so that g(P ◦ K α) < ε for α ≥ α1. Thus,
we obtain
kf k kQ ◦ Kα0 + CP ◦ K αk ≥ kgk + Ckhk − 2ε(1 + C),
for all α ≥ α1.
Now, take (xα) ⊂ BE such that limαkQ ◦ Kα0 + CP ◦ K αk = limαkQ(Kα0xα) +
CP (K αxα)k and note that (Kα0xα) is contained in a compact subset of BE and
12
V. DIMANT, S. LASSALLE, A. PRIETO
that (K αxα) is weakly null. Since E has modulus of asymptotic convexity of
power n we apply Lemma 3.12, with C > 0 as in item (c), and get
limαkQ ◦ Kα0 + CP ◦ K αk ≤ limαkQkkKα0xαkn + CkP kkK αxαkn
= limαkKα0xαkn + CkK αxαkn
≤ limαkKα0xα + K αxαkn
≤ limαkKα0 + K αkn
≤ kIdE − 2Kα0kn,
where the last inequality, being standard, can be found for instance in [18, p.
300]. Then, we may find α2 > α1 so that
kQKα0 + CP K α2k < 1 + ε,
and therefore,
kgk + Ckhk − 2ε(1 + C) ≤ kf k(1 + ε).
Since ε > 0 is arbitrary, kgk + Ckhk ≤ kf k and Pw(nE, F ) is an M(1, C)-ideal
in P(nE, F ).
To prove that Pw(nE, F ) is also an HB -subspace of P(nE, F ) note that for
h 6= 0, kgk < kgk + Ckhk ≤ kf k. On the other hand, for α > α1,
kf k ≥ f (P ◦ K α) ≥ h(P ◦ K α) − g(P ◦ K α) ≥ khk − 2ε,
implying that kf k ≥ khk, which completes the proof.
(cid:3)
4. Ideal structures inherited by the range space
Our purpose in this section is to give sufficient conditions on the spaces E and
F under which those geometric properties enjoyed by Pw(nE, F ) as a subspace
of P(nE, F ) are inherited by the range space F as a subspace of F ∗∗. We start
with HB -smoothness presenting an extension to the polynomial setting of [24,
Theorem 7]. Our proof also follows their ideas.
Proposition 4.1. Let E, F be Banach spaces such that there exists a surjection
ρ : E → F . If Pw(nE, F ) is HB-smooth in P(nE, F ) for some n ∈ N, then F is
HB-smooth in F ∗∗.
Proof. Denote by NA(E∗∗) the subset of norm-attaining elements in E∗∗ and
consider
A := {ρ∗∗(x∗∗) : x∗∗ 6= 0,
x∗∗ ∈ NA(E∗∗)}.
As ρ is surjective, ρ∗∗ is also surjective and, by the Bishop-Phelps theorem, A =
F ∗∗. By [24, Theorem 1 (c)] it is enough to show that for every ρ∗∗(x∗∗) ∈ A
and any sequence (yk) ⊂ F with ky1k < 1, kyk+1 − ykk < 1 there are y ∈ F and
k0 ∈ N so that kρ∗∗x∗∗ − y ± yk0k < k0.
0) = kx∗∗
0 (x∗
0 k and define,
0(x)nyk ∈ Pw(nE, F ). It is
Fix x∗∗
0
6= 0 in NA(E∗∗), take x∗
0 ∈ SE ∗ such that x∗∗
for k ∈ N, the n-homogeneous polynomial Pk(x) = x∗
clear that
kP1k < 1 and kPk+1 − Pkk ≤ kyk+1 − ykk < 1,
for all k ∈ N.
Now, define Q(x) = ρ(x) x∗
Q given by Q(x∗∗) = ρ∗∗(x∗∗) x∗∗(x∗
0(x)n−1kx∗∗
0)n−1kx∗∗
0 k.
0 k and consider its Aron -- Berner extension
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
13
By [24, Theorem 1 (a)] there exist R ∈ Pw(nE, F ) and k0 ∈ N with kQ −
R ± Pk0k < k0. As the Aron -- Berner extension preserves the norm, we obtain
0 kn. The result follows by taking y =
kQ(x∗∗
0 ) − R(x∗∗
R(x∗∗
0 )
0 kn .
kx∗∗
As an immediate consequence we have the following result.
0 )k < k0kx∗∗
0 ) ± P k0(x∗∗
(cid:3)
Corollary 4.2. Let E be a Banach space. If Pw(nE, E) is HB-smooth in P(nE, E)
for some n ∈ N, then E is HB-smooth in E∗∗.
Note that the above corollary says that Pw(nℓ1, ℓ1) is not HB -smooth in P(nℓ1, ℓ1)
for any n ∈ N.
Now we address the notion of HB -subspace, when the range space F is a quo-
tient of the space E. The following technical result, inspired by [26, Proposition
2.3], will be useful.
Lemma 4.3. Let J be an ideal in the Banach space E under the projection
q . Suppose that J is HB-smooth and λ ∈ (0, 2]. The following statements are
equivalent.
(i) kIdE ∗ − λqk ≤ 1.
(ii) For each x ∈ BE there exists a net (yα) ⊂ J such that limα yα = x in the
σ(E, J ∗)-topology and limαkx − λyαk ≤ 1.
(iii) For each x ∈ BE and ε > 0 there exists a net (yα) ⊂ J such that limα yα = x
in the σ(E, J ∗)-topology and limαkx − λyαk ≤ 1 + ε.
Proof. To prove that (i) implies (ii), consider the set of indices A := {α =
(N, M, ε) : N ∈ FIN(E∗∗), M ∈ FIN(E∗), ε > 0}, where FIN denotes the set
of all finite dimensional subspaces, with the usual order. By the Principle of
Local Reflexivity, for any α ∈ A there exists Tα : N → E so that
• hTαx∗∗, x∗i = hx∗∗, x∗i, for x∗∗ ∈ N, x∗ ∈ M,
• kTαk ≤ 1 + ε,
• TαN ∩E = IdE.
Fix x ∈ BE and consider yα = Tα(q
y∗ ∈ J ∗ and ε > 0, then if α ≥ ({q
∗x) ∈ E, defined for α large enough. Fix
∗x}, {y∗}, ε), as J ∗ is the range of q, we have
hy∗, yαi = hy∗, Tα(q
∗x)i = hy∗, q
∗xi = hqy∗, xi = hy∗, xi,
and limα yα = x in the σ(E, J ∗)-topology. On the other hand, also for α large
enough,
kx − λyαk = kTα(x − λq
∗x)k ≤ (1 + ε)kx − λq
∗xk ≤
≤ (1 + ε)kIdE ∗∗ − λq
∗k = (1 + ε)kIdE ∗ − λqk ≤ 1 + ε.
Then, limαkx − λyαk ≤ 1 and (ii) follows.
Clearly, (ii) implies (iii). To prove that (iii) implies (i), fix ε > 0, x ∈ BE
and choose (yα) ⊂ J satisfying (iii). For each x∗ ∈ BE ∗, as limα x∗(yα) =
limα qx∗(yα) = qx∗(x), we have that
x∗(x) − λqx∗(x) = lim
α
x∗(x) − λx∗(yα) ≤ kx∗k limαkx − λyαk ≤ 1 + ε.
14
V. DIMANT, S. LASSALLE, A. PRIETO
Since ε is arbitrary, the implication follows.
(cid:3)
Proposition 4.4. Let E, F be Banach spaces such that there exists a quotient
If Pw(nE, F ) is an HB-subspace of P(nE, F ), for some
mapping ρ : E → F .
n ∈ N, then F is an HB-subspace of F ∗∗.
Proof. First, note that F is an ideal in its bidual and call q the associated pro-
jection. By Proposition 4.1, F is HB -smooth in F ∗∗. Then, by [22, Theorem], F
is strongly HB -smooth in F ∗∗ and we only have to show that kf − qf k ≤ kf k for
all f ∈ F ∗∗∗. In order to do so, we will see that condition (iii) in Lemma 4.3 is
satisfied for λ = 1.
Take y∗∗ ∈ BF ∗∗ and ε > 0. Choose w∗ ∈ SF ∗ so that y∗∗(w∗) > 0 and
(cid:18) ky∗∗k
y∗∗(w∗)(cid:19)n
< 1 + ε. Now, with µ = y∗∗(w∗) define P ∈ P(nE, F ) by P (x) =
µρ(x)(ρ∗w∗)n−1(x), for each x ∈ E. By Lemma 4.3, there exists a net (Pα) ⊂
Pw(nE, F ) converging to P in the σ(P(nE, F ), Pw(nE, F )∗)-topology and such
that limαkP − Pαk ≤ 1. As ρ is a quotient mapping, ρ∗ is an isometry and we
may find x∗∗ ∈ E∗∗ with ρ∗∗(x∗∗) = y∗∗ and kx∗∗k = ky∗∗k. Since the Aron --
Berner extension of each Pα has range in F , we define yα = P α(x∗∗/µ) ∈ F , for
all α.
Note that each x∗∗ ⊗ y∗ ∈ E∗∗ ⊗ F ∗ acts in a natural way as an element
of P(nE, F )∗ and, therefore, as an element of Pw(nE, F )∗. Then, as P (z) =
µρ∗∗(z)zn−1(ρ∗w∗), for z ∈ E∗∗, we have P (x∗∗/µ) = y∗∗ and
y∗∗(y∗) = (P (x∗∗/µ))(y∗) = lim
α
Thus, yα → y∗∗ in the w∗-topology. Also,
y∗(Pα(x∗∗/µ)) = lim
α
y∗(yα).
limαky∗∗ − yαk = limαkP (x∗∗/µ) − P α(x∗∗/µ)k
≤ limαkP − Pαkkx∗∗/µkn
≤ (ky∗∗k/µ)n < 1 + ε.
Another application of Lemma 4.3 gives that kIdF ∗∗∗ − qk ≤ 1 and, therefore, F
is an HB -subspace of F ∗∗.
(cid:3)
Finally, we focus on M(1, C)-ideal structures. Recall that if J is an ideal in a
Banach space E satisfying the M(1, C)-inequality the following condition holds:
For any m ∈ N, y1, y2, . . . , ym ∈ BJ , x ∈ BE, and ε > 0, there is z ∈ J such that
kyi + Cx − zk ≤ 1 + ε, for 1 ≤ i ≤ m, [7, Lemma 2.2]. In fact, when dealing with
E = J ∗∗, it is true that being an M(1, C)-ideal is equivalent to an appropriate
2-ball property of type (1, C). Namely, we have the following equivalence, which
can be proved with the arguments appearing in the proof of [7, Lemma 2.3].
Lemma 4.5. Let E be a Banach space, and letC ∈ (0, 1]. The following state-
ments are equivalent.
(i) E is an M(1, C)-ideal of E∗∗.
(ii) For all x ∈ SE, x∗∗ ∈ SE ∗∗ and ε > 0, there exists x0 ∈ E with k ± x +
Cx∗∗ − x0k < 1 + ε.
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
15
We use the above characterization to give an analogous statement to Proposi-
tion 4.4 in the case of M(1, C)-ideals.
Proposition 4.6. Let E, F be Banach spaces such that there exists a quotient
mapping ρ : E → F . If Pw(nE, F ) is an M(1, C)-ideal of P(nE, F ), for some
n ∈ N and C > 0, then F is an M(1, C)-ideal of F ∗∗.
Proof. Let us prove that condition (ii) in Lemma 4.5 is satisfied. Fix y ∈ SF ,
(1 − δ)n−1 < 1+ε and y∗ ∈ SF ∗
y∗∗ ∈ SF ∗∗ and ε > 0. Choose δ > 0 such that δ +
with y∗∗(y∗) ≥ 1 − δ. Define, for x ∈ E, P ∈ P(nE, F ) and Q ∈ Pw(nE, F ) by
1 + δ
P (x) = (ρ∗y∗)n−1(x)ρ(x)
and Q(x) = (ρ∗y∗)n(x)y,
with kP k, kQk ≤ kρkn = 1. Due to [7, Lemma 2.2], there exists R ∈ Pw(nE, F )
so that k ± Q + CP − Rk ≤ 1 + δ. As ρ∗ is an isometry, there is x∗∗ ∈ SX ∗∗ with
ρ∗∗(x∗∗) = y∗∗. Extending by Aron -- Berner, for z ∈ E∗∗,
P (z) = z(ρ∗y∗)n−1ρ∗∗(z),
and Q(z) = z(ρ∗y∗)n y.
Then, with µ = y∗∗(y∗), P (x∗∗) = µn−1y∗∗ and Q(x∗∗) = µny,
k ± µn y + Cµn−1 y∗∗ − R(x∗∗)k = k ± Q(x∗∗) + CP (x∗∗) − R(x∗∗)k ≤ 1 + δ.
As R ∈ Pw(nE, F ), the range of R is also in F and we may take y0 = R(x∗∗)/µn−1.
Thus,
k ± µy + Cy∗∗ − y0k ≤
1 + δ
(1 − δ)n−1 .
Finally, k ± y + Cy∗∗ − y0k ≤ ky − µyk + k ± µy + Cy∗∗ − y0k < 1 + ε, and the
result follows.
(cid:3)
In the above proposition, the case C = 1 corresponds to the structure of an
M-ideal. The corollary follows directly and seems to be new in this context.
Corollary 4.7. Let E, F be Banach spaces such that there exists a quotient map-
ping ρ : E → F . If Pw(nE, F ) is an M -ideal of P(nE, F ), for some n ∈ N, then
F is an M -ideal of F ∗∗.
As we have already noted in Corollary 4.2, it is now immediate to derive ver-
sions of Proposition 4.4, Proposition 4.6 and Corollary 4.7 for the case E = F .
Acknowledgement. The work of this paper was initiated while the third author
was visiting the Departamento de Matem´atica, Universidad de San Andr´es. She
is greatly indebted to Profs. Dimant and Lassalle for the invitation during her
sabbatical, and for the warm hospitality extended by them and their families.
Work partially supported by PAI-UdeSA 2013, CONICET-PIP 0624, ANPCyT
PICT 1456, UBACyT X038 and MINECO grant MTM 2012-3431.
16
V. DIMANT, S. LASSALLE, A. PRIETO
References
1. R. Alencar, K. Floret. Weak-strong continuity of multilinear mappings and the Pelczynski-
Pitt theorem. J. Math. Anal. Appl. 206 (1997), no. 2, 532 -- 546.
2. R. Aron, P. Berner. A Hahn -- Banach extension theorem for analytic mappings. Bull. Math.
Soc. France 106 (1978), 3 -- 24.
3. R. Aron, C. Herv´es, M. Valdivia. Weakly continuous mappings on Banach spaces. J. Funct.
Anal. 52 (1983), 189 -- 204.
4. R. Aron, J. B. Prolla. Polynomial approximation of differentiable functions on Banach
spaces. J. Reine Angew. Math. 313 (1980), 195 -- 216.
5. C. Boyd, R. Ryan. Bounded weak continuity of homogeneous polynomials at the origin.
Arch. Math. (Basel) 71 (1998), no. 3, 211 -- 218.
6. J. C. Cabello, E. Nieto. On properties of M -ideals. Rocky Mountain J. Math. 28 (1998),
no. 1, 61-93.
7. J. C. Cabello, E. Nieto, E. Oja. On ideals of compact operators satisfying the M (r, s)-
inequality. J. Math. Anal. Appl. 220 (1998), 334 -- 348.
8. D. Carando, S. Lassalle. E ′ and its relation with vector-valued functions on E. Ark. Mat.
42 (2004), no. 2, 283 -- 300.
9. A. Davie, T. Gamelin. A theorem on polynomial-star approximation. Proc. Amer. Math.
Soc. 106 (1989), no. 2, 351 -- 356.
10. S. Delpech. Approximations hold´eriennes de fonctions entre espaces d'Orlicz. Modules
asymptotiques uniformes. Doctoral thesis. Universit´e de Bordeaux, 2005.
11. J. Diestel, J. J. Uhl. Vector measures. Mathematical Surveys 15. American Mathematical
Society, Providence, R. I., 1977.
12. V. Dimant. M -ideals of homogeneous polynomials. Studia Math. 202 (2011), 81 -- 104.
13. V. Dimant, R. Gonzalo, J. A. Jaramillo. Asymptotic structure, ℓp-estimates of sequences,
and compactness of multilinear mappings. J. Math. Anal. Appl. 359 (2009), 680 -- 693.
14. V. Dimant, S. Lassalle. M -structures in vector-valued polynomial spaces. J. Convex Anal.
19 (2012), no. 3, 685 -- 711.
15. S. Dineen. Complex analysis on infinite-dimensional spaces. Springer Monographs in Math-
ematics. Springer-Verlag London, Ltd., London, 1999.
16. M. Gonz´alez, J. Guti´errez. The polynomial property (V). Arch. Math. (Basel) 75 (2000),
no. 4, 299 -- 306.
17. R. Gonzalo, J. A. Jaramillo. Compact polynomials between Banach spaces. Proc. Roy. Irish
Acad. Sect. A 95 (1995), no. 2, 213 -- 226.
18. P. Harmand, D. Werner, W. Werner. M -ideals in Banach spaces and Banach algebras.
Lecture Notes in Mathematics, 1547. Springer-Verlag, Berlin, 1993.
19. J. Hennefeld. M -ideals, HB-subspaces, and compact operators. Indiana Univ. Math. J. 28
(1979), no. 6, 927 -- 934.
20. J. Johnson. Remarks on Banach spaces of compact operators. J. Funct. Anal. 32 (1979),
304 -- 311.
21. V. Lima, A. Lima. Ideals of operators and the metric approximation property. J. Funct.
Anal. 210 (2004), no. 1, 148 -- 170.
22. E. Oja. Strong uniqueness of the extension of linear continuous functionals according to the
Hahn -- Banach theorem. (Russian) Mat. Zametki 43 (1988), no. 2, 237 -- 246, 302; translation
in Math. Notes 43 (1988), no. 1 -- 2, 134 -- 139.
23. E. Oja. Isometric properties of the subspace of compact operators in the space of continuous
linear operators. (Russian) Mat. Zametki 45 (1989), no. 6, 61 -- 65, 111; translation in Math.
Notes 45 (1989), no. 5 -- 6, 472 -- 475.
24. E. Oja, M. Poldvere. On subspaces of Banach spaces where every functional has a unique
norm-preserving extension. Studia Math. 117 (1996), no. 3, 289 -- 306.
25. W. M. Ruess, C. P. Stegall. Extreme points in duals of operator spaces. Math. Ann. 261
(1982), no 4, 535 -- 546.
IDEAL STRUCTURES IN VECTOR-VALUED POLYNOMIAL SPACES
17
26. D. Werner. M -ideals and the "basic inequality". J. Approx. Theory 76 (1994), no. 1, 21 -- 30.
1Departamento de Matem´atica, Universidad de San Andr´es, Vito Dumas 284,
(B1644BID) Victoria, Buenos Aires, Argentina and CONICET.
E-mail address: [email protected]; [email protected]
2Departamento de An´alisis Matem´atico, Facultad de CC. Matem´aticas, Uni-
versidad Complutense de Madrid, Plaza de Ciencias, 3, 28040 Madrid, Spain.
E-mail address: [email protected]
|
1807.10790 | 2 | 1807 | 2018-08-26T06:06:33 | Interpolation of weighted Sobolev spaces | [
"math.FA"
] | In this work we present a newly developed study of the interpolation of weighted Sobolev spaces by the complex method. We show that in some cases, one can obtain an analogue of the famous Stein-Weiss theorem for weighted $L^{p}$ spaces. We consider an example which gives some indication that this may not be possible in all cases. Our results apply in cases which cannot be treated by methods in earlier papers about interpolation of weighted Sobolev spaces. They include, for example, a proof that $\left[W^{1,p}(\mathbb{R}^{d},\omega_{0}),W^{1,p}(\mathbb{R}^{d},\omega_{1})\right]_{\theta}=W^{1,p}(\mathbb{R}^{d},\omega_{0}^{1-\theta}\omega_{1}^{\theta})$ whenever $\omega_{0}$ and $\omega_{1}$ are continuous and their quotient is the exponential of a Lipschitz function. We also mention some possible applications of such interpolation in the study of convergence in evolution equations. | math.FA | math |
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
MICHAEL CWIKEL AND AMIT EINAV
ABSTRACT. In this work we present a newly developed study of the interpo-
lation of weighted Sobolev spaces by the complex method. We show that
in some cases, one can obtain an analogue of the famous Stein-Weiss the-
orem for weighted Lp spaces. We consider an example which gives some
indication that this may not be possible in all cases. Our results apply in
cases which cannot be treated by methods in earlier papers about interpo-
lation of weighted Sobolev spaces. They include, for example, a proof that
hW 1,p³Rd ,ω0´ ,W 1,p³Rd ,ω1´iθ = W 1,p³Rd ,ω1−θ
0 ωθ
are continuous and their quotient is the exponential of a Lipschitz function.
We also mention some possible applications of such interpolation in the study
of convergence in evolution equations.
1´ whenever ω0 and ω1
CONTENTS
Introduction
1.
1.1. Goals and aims
1.2. The setting of the problem
1.3. Main results
1.4. Structure of the paper
Acknowledgement
2. Preliminaries: Interpolation spaces and the complex interpolation
method
3. The simpler inclusion
4. Towards the difficult inclusion
5. Approximation theorems
6. Proofs of the main theorem, and of its main consequences
7. More about the space W p (θ, rω)
7.1. Can W p (θ, rω) be smaller than W 1,p (ωθ)?
7.2. A simpler subspace of£W 1,p (U , ω0),W 1,p (U , ω1)¤θ
8. Final Remarks
8.1. Equivalence of weights
8.2. Possible applications to evolution equations
8.3. Future research and open problems
Appendix A. Additional Proofs
2
2
5
8
11
11
11
14
18
30
39
40
41
43
45
45
45
47
47
The first named author's work was supported by the Technion V.P.R. Fund, and by the Fund for
Promotion of Research at the Technion. The second named author was supported by the Austrian
Science Fund (FWF) grant M 2104-N32.
1
2
MICHAEL CWIKEL AND AMIT EINAV
A.1. An alternative characterization of the space W
A.2. From locally Lipschitz to globally Lipschitz - A proof of Fact (a)
A.3. A Stein-Weiss like theorem in the case where rω is not Lipschitz
A.4. A comment about homogeneous weighted Sobolev spaces of
1,p
l oc (U )
univariate functions
References
48
49
50
51
52
1. INTRODUCTION
1.1. Goals and aims. Consider two weighted Sobolev spaces, W s0,p0 (U , ω0) and
W s1,p1(U , ω1) on an open subset U of Rd . (Precise definitions of these spaces and
of other notions, which are mentioned only in general terms in this subsection,
will be given later). A natural question one can ask oneself is the following:
Question 1. If T is a linear operator which is bounded on both W s0,p0 (U , ω0) and
W s1,p1(U , ω1), does this imply that T is also bounded on various other weighted
Sobolev spaces W s,p (U , ω)?
For certain choices of U , weight functions ω0, ω1, and s0, s1, p0 and p1 the
answer is known to be affirmative. In most cases, the affirmation of Question
1 is obtained with the explicit, or implicit, help of Alberto Calderón's complex
interpolation spaces. As in many (but not all) papers dealing with these inter-
polation spaces, we will use Calderón's notation [A0, A1]θ for them. Thus, so far,
the affirmation of Question 1 comes hand in hand with the affirmation of the
following question:
Question 2. Does the complex interpolation space [W s0,p0 (U , ω0),W s1,p1(U , ω1)]θ
coincide with W s,p (U , ω) for some s, p and ω?
Our goal in this work is to investigate Question 2 further, and provide some
new affirmative answers to it - thereby also finding more cases where Question 1
has an affirmative answer. We consider that our results can be potentially useful
in the study of asymptotic behaviour of solutions of certain evolution equations.
Before continuing, let us recall two previously treated cases of Questions 1
and 2:
The Stein-Weiss Theorem: Suppose that s0 = s0 = 0, i.e., suppose that our two
weighted Sobolev spaces are simply weighted Lp spaces. Then a celebrated re-
sult, often referred to as the Stein-Weiss Theorem, which goes back to the paper
[20] of Eli Stein and Stein's joint paper [21] with Guido Weiss, gives a positive
answer to Question 1 for all p0 and p1 in [1,∞] and all choices of positive mea-
surable functions ω0 and ω1 on U (and even in a more general setting, when
(U , d x) is replaced by an arbitrary measure space). More explicitly, in this case,
the operator T is bounded on W 0,p (U , ω) whenever p and ω satisfy
(1.1)
1
p =
1− θ
p0 +
θ
p1
,
and ω
1
1−θ
p0
0 ω
θ
p1
1 .
p = ω
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
3
for some θ ∈ (0,1). This result was obtained before Calderón developed his the-
ory of complex interpolation spaces, but in the light of that theory the Stein-
Weiss Theorem can also be expressed by the formula
£W 0,p0 (U , ω0),W 0,p1 (U , ω1)¤θ = W 0,p (U , ω)
(1.2)
which holds isometrically for each θ ∈ (0,1) and for p and ω as in (1.1). (When
p0 = p1 = ∞ the weight ω of course cannot be determined from (1.1) but in that
case (1.2) holds for all choices of ω for a trivial reason related to the discussion
below in Remarks 1.1 and 2.8.)
A result of Jörgen Löfström: In his paper [14], among various other results, Löf-
ström gives an affirmative answer to Question 2 in the setting where s0 and s1
1/p0
can be non-zero. However, the set U must be all of Rd and the powers ω
and
0
1/p1
of the given weight functions are required to satisfy a special condition
ω
1
which he calls "polynomial regularity". With these conditions fulfilled, Löfström
is able to use the Fourier transform and the Mihlin multiplier theorem and to
define the weighted Sobolev spaces W s0,p0¡Rd , ω0¢ and W s1,p1¡Rd , ω1¢ via prop-
erties of the Fourier transforms of their elements. In this setting Löfström proves
that
(1.3)
hW s0,p0³Rd , ω0´,W s1,p1³Rd , ω1´iθ = W sθ,p³Rd , ω´
to within equivalence of norms, for each p0 and p1 ∈ (1,∞) and each θ ∈ (0,1)
where sθ = (1−θ)s0 +θs1 and p and ω are given by (1.1). Taking into account the
particular different notation and formulation of definitions used by Löfström,
one can see that this result is expressed by the formula (5.3) in Theorem 4 on
page 208 of [14]. An interesting advantage of Löfström's method is that he can
extend his definition of weighted Sobolev spaces to the case where their orders
of smoothness are not necessarily integers, and in fact he proves his formula
(1.3) in this more general setting.
Stein-Weiss formula (1.2).
Of course, in the special case where s0 = s1 = 0, the formula (1.3) becomes the
In this paper, in contrast to Löfström's work, we shall only consider the case
s0 = s1 = 1,
p0 = p1 = p ∈ [1,∞).
However, making these restrictions gives us much more flexibility in our choices
of the set U and the weight functions ω0 and ω1 than is available in [14]. We can
also include the case where p = 1. For a relatively large class of weight functions,
and for U = Rd , we obtain that
(1.4)
to within equivalence of norms, for all p ∈ [1,∞) and θ ∈ (0,1), where here, and
also in fact throughout this paper, ωθ is the weight function obtained, for each θ,
from whichever weight functions ω0 and ω1 are currently under consideration,
by the formula
£W 1,p (U , ω0),W 1,p (U , ω1)¤θ = W 1,p (U , ωθ)
(1.5)
ωθ = ω1−θ
0 ωθ
1.
4
MICHAEL CWIKEL AND AMIT EINAV
Remark 1.1. When p = ∞, a case which we have not formally included in our
discussion, the formula (1.4) is an uninteresting triviality (see Remark 2.8 below)
which holds isometrically for all choices of the weight functions ω0 and ω1 (as
can be justified later by part (iv) of Theorem 2.7).
Here is one easily formulated special case of our results, which gives some
indication of the considerable flexibility we have in our choices of the weight
functions ω0 and ω1: When the set U is chosen to be all of Rd , we can prove
that (1.4) holds whenever ω0/ω1 is the exponential of a Lipschitz function and
ω0 and ω1 are continuous. In fact we can make do here with a much weaker
condition than continuity which is satisfied by various functions which are not
even equivalent to continuous functions.
In fact (1.4) is valid not only for U = Rd , but also for a certain collection of
other open sets which satisfy properties which we shall specify below. Moreover,
(U , ω) of functions in W 1,p (U , ω) that
if we consider the natural subspaces W
can be approximated by a natural class of compactly supported functions, an
analogue of (1.4) for these spaces is valid for an even larger collection of open
sets U .
1,p
0
Our strategy for obtaining (1.4) will be to show that, under somewhat milder
conditions on ω0 and ω1 than those imposed in Löfström's theorem, and for ap-
propriate choices of U , the interpolation space£W 1,p (U , ω0),W 1,p (U , ω1)¤θ sat-
isfies the continuous inclusions
W p (U , θ, rω) ⊂£W 1,p (U , ω0),W 1,p (U , ω1)¤θ ⊂ W 1,p(U , ωθ)
for a certain space W p (U , θ, rω) which is the intersection of W 1,p (U , ωθ) with an
appropriate weighted Lp space. These inclusions may also be of independent
interest. We can then deduce that (1.4) holds under some explicitly formulated
conditions (see Theorem 1.22), which ensure that W 1,p (U , ωθ) and W p (U , θ, rω)
coincide.
Interestingly enough, we will also present a somewhat intricate example where
these spaces do not coincide, but we cannot exclude the possibility that (1.4)
might still hold for that example.
We are now ready, in the following subsections, to give more explicit and more
detailed formulations of our results and of the definitions of the notions which
play roles in them.
Remark 1.2. There are a number of other papers and books which contain other
interpolation results involving weighted Sobolev spaces (and also sometimes
other related spaces such as weighted Besov spaces and/or weighted Triebel-
Lizorkin spaces) but the results in them do not quite correspond to the kind of
results that we are seeking here. Some of them deal with the real method, rather
than the complex method of interpolation. Some deal with interpolation be-
tween weighted Sobolev spaces and weighted Lp spaces. Also, in some of them,
the definitions of weighted Sobolev spaces differ in some non-trivial way from
the definition that we have chosen to use here. We can refer, for example, to
[5, 7, 8, 9, 10, 15, 16, 17, 18, 19, 22]. Among the many other papers and some
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
5
books which deal with other aspects of weighted Sobolev spaces and some of
their applications we can mention, for example, the works [11, 12, 13] of Kufner,
Kufner-Opic, and Kufner-Sändig.
1.2. The setting of the problem. In what is to follow, d will always be a positive
integer, p will always lie in the interval [1,∞], and U will always denote a non-
empty open subset of Rd . For most of our results p will be restricted to the range
[1,∞). The letter ω, with or without a subscript, will always denote a weight
function on U , i.e. a Lebesgue measurable function ω : U → (0,∞).
Often the weight functions which are of interest in various applications are con-
tinuous. However, for our results here which establish the formula (1.4) in cer-
tain cases, we will not need the weight functions ω0 and ω1 to be continuous,
nor even equivalent to continuous functions. Instead we will require them to
satisfy a rather weaker condition which we are now about to define. (We will
however need ω0/ω1 to have some smoothness properties.)
We will be considering weight functions ω : U → (0,∞) which, like continuous
functions, satisfy
(1.6)
0 < inf
x∈K
ω(x) ≤ sup
x∈K
ω(x) < ∞,
for every compact set K ⊂ U .
Consequently, it will sometimes be convenient to use the notation
(1.7)
m(K , ω) := inf
x∈K
ω(x),
and M(K , ω) := sup
x∈K
ω(x).
Definition 1.3. A weight function ω : U → (0,∞) on an open subset U of Rd
which satisfies (1.6) will be said to satisfy the compact boundedness condition on
U .
Remark 1.4. A weight function ω satisfying this condition can be very far from
being continuous. As we already intimated above, it can, for example, have the
property that, for every choice of a pair of positive constants C1 and C2 and of
a continuous function f , the inequality C1 f (x) ≤ ω(x) ≤ C2 f (x) cannot hold
for all x ∈ U , not even for almost all x ∈ U . As an example of this, let U =
(0,∞) and let ω be the weight function on U defined by ω(x) =P∞n=0 nχ(2n,2n+1]+
P∞n=0 n2χ(2n,+1,2n+2].
Remark 1.5. We will have occasions to use the obvious fact that, if ω0 and ω1 are
weight functions which satisfy the compact boundedness condition on some set
U , then the same is true for every power, product and quotient of these weights,
and in particular for ω1−θ
0 ωθ
1 and ω0/ω1.
In the literature there are two standard ways of defining weighted Lp spaces
with respect to a given weight ω. In one of them the weighted space consists of
all functions f for which ω f is in the (unweighted) Lp space. But here we shall
choose the second way, which proceeds by replacing the ambient underlying
measure, in our case Lebesgue measure d x, with the weighted measure ω(x)d x.
In fact we shall need to use weighted Lp spaces of vector valued functions.
6
MICHAEL CWIKEL AND AMIT EINAV
Definition 1.6. Given a weight function ω on U and p ∈ [1,∞), and m ∈ N, we
define the Banach space
Lp¡U , ω, Cm¢ =nφ : U → Cm φ is measurable and °°φ°°Lp (U,ω,Cm ) < ∞o
where
°°φ°°Lp (U,ω,Cm ) =Ã m
Xi=1U¯¯φi (x)¯¯
p
ω(x)d x!
1
p
,
for each φ =¡φ1,..., φm¢.
We will sometimes use the abbreviated notation Lp (U , ω) for Lp(U , ω, C), i.e.
when m = 1. We will also use the notation Lp (U , Cm ) for the "unweighted" case,
where ω is identically 1.
As is customary in such definitions, there is a natural "limiting" extension to
the case where p = ∞, which in our context takes the following form:
Definition 1.7. For U , ω and m as above we define the Banach space
L∞¡U , ω, Cm¢ =nφ : U → Cm φ is measurable and °°φ°°L∞(U,ω,Cm ) < ∞o
where
°°φ°°L∞(U,ω,Cm) = esssup
x∈Uµ max
i∈{1,2,...,m}¯¯φi (x)¯¯¶
for each φ =¡φ1,..., φm¢.
Remark 1.8. The consequence of the choice that we made in formulating Defi-
nition 1.6 is that here the weighted L∞ space in fact does not depend on ω and
coincides isometrically with the unweighted space L∞(U , Cm ) which we shall
occasionally use below. Had we chosen the first way of defining weighted Lp
spaces we would have obtained a more "interesting" space here.
We can now introduce the main spaces that we wish to study.
Definition 1.9. Given an open set U of Rd , a weight function ω on U and p ∈
[1,∞], we define
W 1,p (U , ω) =nφ : U →C φ has a weak gradient ∇φ on U and
¡φ,∇φ¢ ∈ Lp³U , ω, Cd+1´o,
where the weak gradient ∇φ (i.e. the vector of its weak first order partial deriva-
tives) is defined in the usual way via the theory of distributions, i.e. for every
ψ ∈ C∞c (U ) it satisfies
(1.8)
U ∇ψ(x)φ(x)d x = −U
ψ(x)∇φ(x)d x.
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
7
The norm1 which we will define on W 1,p (U , ω) is given by
°°φ°°W 1,p (U,ω) =°°¡φ,∇φ¢°°Lp(U,ω,Cd+1) .
Remark 1.10. The requirement that φ and ∇φ satisfy the equation (1.8) in which
the integrals are evidently understood to be Lebesgue integrals, forces both of
these functions to be locally integrable on U (with respect to (unweighted) Lebesgue
measure).
Remark 1.11. In the case where the weight function ω is identically equal to 1 the
above weighted Sobolev spaces coincide with their original and more frequently
studied "unweighted" analogues. We shall occasionally need to use these un-
weighted spaces here and (analogously to what we did for unweighted Lp spaces)
we shall use the same notation as above for them, but with ω simply omitted.
Note (cf. Remark 1.8) that, for every weight function ω, W 1,∞(U , ω) coincides
isometrically with the unweighted Sobolev space W 1,∞(U ).
The study of weighted Sobolev spaces is certainly not new. The definition
which we have chosen to use for them here is only one of a number of various
different definitions chosen by authors of the many papers and several books
(already mentioned above in §1.1) which deal with weighted Sobolev spaces and
in some cases also give or indicate various applications of them to other topics
in analysis.
Consider the normed space W k,p(Ω, σ) introduced on page 11 of [11] and also,
with slightly different notation, on page 537 of [12]. In the case where Ω = U and
k = 1 and all the elements of the vector σ = {σα : α ≤ 1} are chosen to be ω this
space coincides isometrically with our space W 1,p (U , ω). (In [11] it is assumed
that p ∈ [1,∞) and that the open set Ω is also connected.)
For most of our purposes in this paper we need the weighted Sobolev spaces
with which we are working to be complete. Accordingly, most of our results
here will be formulated subject to the "abstract" requirement that the weighted
Sobolev spaces being considered in them are complete. When seeking to apply
these results we can keep in mind that there are various "concrete" conditions on
a weight function ω which are sufficient to ensure that W 1,p (U , ω) is complete.
One of these is a condition which appears in Theorem 1.11 of [12, pp. 540 -- 541]
in the case where p > 1. More relevantly for us, and as we shall show below in
Lemma 5.2, W 1,p(U , ω) is complete whenever ω satisfies the compact bounded-
ness condition. This is a condition which will be imposed anyway, also for other
purposes, on the weight functions appearing in our main theorem 1.16, and in
Theorems 1.20, 1.21 and 1.22 which follow from it.
Remark 1.12. We have not attempted to find examples of p, U and ω for which
W 1,p(U , ω) is not complete. However Example 1.12 on pp. 541 -- 543 of [12] shows
explicitly that a certain variant of W 1,p (U , ω), where different weight functions
1Of course in this definition, and in Definitions 1.6 and 1.7, we permit ourselves the usual
abuse of terminology where the word "function" really means an equivalence class of functions
which are equal to each other almost everywhere.
8
MICHAEL CWIKEL AND AMIT EINAV
are used for the weighted Lp norms of the function and of its weak first order
partial derivatives, can fail to be complete.
In some of our results we will also be requiring the weighted Sobolev spaces
W 1,p(U , ω) appearing in them to have the property that each of their elements
can be approximated by a sequence of compactly supported C∞ or Lipschitz
functions. Here too there are concrete conditions on U and ω which suffice to
imply such properties. The interested reader can find more information about
such issues in [11] and [12].
1.3. Main results. As we've stated in §1.1, our main goal in this work is to find
settings in which an analogue of the Stein-Weiss theorem holds for weighted
Sobolev spaces, or for suitable substitutes for these spaces.
In what follows, we will denote by Li pc (U ), or Li pc for short, the space of all Lip-
schitz functions ϕ : U → C with compact support. To avoid any ambiguity here,
we should point out that ϕ ∈ Li pc(U ) means that the closure of©x ∈ U : ϕ(x) 6= 0ª
in Rd is a compact subset, not only of Rd , but also of U .
Definition 1.13. Let ω be a weight function such that W 1,p (U , ω) is a Banach
(U , ω) to be the sub-
space, containing Li pc (U ). Define the Banach space W
space of W 1,p (U , ω) equipped with the same norm as that of W 1,p (U , ω) which
is the closure of Li pc (U ) with respect to that norm.
1,p
0
1,p
0
(U , ω) applicable.
In particular we should mention that if ω satisfies the compact boundedness
condition on U then, as we shall show below in Lemma 5.2, W 1,p (U , ω) has
the two above mentioned properties required to make the above definition of
W
As certain steps in the proof of our main theorem will show, Li pc is a very natu-
ral space to consider for our purposes. In addition, for many choices of the open
set U , the subspace C∞c (U ) of Li pc(U ) can be shown to have the same closure
in W 1,p (U , ω) norm as that of Li pc (U ).
We will also sometimes need to consider functions which satisfy the following
variant of the Lipschitz condition.
Definition 1.14. A function φ : U → C is said to be locally Lipschitz if, for each
compact subset K of U , the restriction of φ to K is Lipschitz.
Remark 1.15. It is easy to verify that φ : U → C is locally Lipschitz if and only if
its restriction to the set V is Lipschitz whenever V is an open set whose closure
is a compact subset of U .
Let us now state our main theorem:
Theorem 1.16. Let ω0 and ω1 be weight functions on U that satisfy the com-
pact boundedness condition (1.6). Assume furthermore that the function rω : U →
(0,∞) defined by
(1.9)
rω(x) :=
ω0(x)
ω1(x)
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
9
is locally Lipschitz on U . For each p ∈ [1,∞) and θ ∈ (0,1), let W p (U , θ, rω) be the
space
with the norm
W p (U , θ, rω) =©φ : U → C φ ∈ W 1,p (U , ωθ), φ¯¯∇log (rω)¯¯ ∈ Lp (U , ωθ)ª
ωθ(x)d x¶
°°φ°°W p (U,θ,rω) =µ°°φ°°
W 1,p (U,ωθ) +U¯¯φ(x)¯¯
p¯¯∇log (rω(x))¯¯
0 (U , θ, rω) to be the closure of Li pc (U ) in W p (U , θ, rω) with respect
1
p
,
p
p
p
(1.10)
and define W
to this norm. Then, for every p ∈ [1,∞) and θ ∈ (0,1), we have that
£W 1,p (U , ω0) ,W 1,p (U , ω1)¤θ ⊂ W 1,p (U , ωθ),
(1.11)
0 (U , θ, rω) ⊂hW
(U , ω1)iθ
(U , ω0),W
(1.12)
1,p
0
1,p
0
and
W
p
.
p
0 (U , θ, rω),
Moreover, there exists a positive constant Cp which depends only on p, such that,
for every φ ∈ W
(1.13)
°°φ°°W 1,p (U,ωθ) ≤°°φ°°[W 1,p (U,ω0),W 1,p (U,ω1)]θ ≤ Cp°°φ°°W p (U,θ,rω)
Remark 1.17. The exact value of the constant in (1.13) interests us rather less
than the fact that it is independent of our choices of weight functions and θ,
and depends only on p. But we also note that the arguments that will eventually
combine to give us a proof of Theorem 1.16 will also show that the said constant
can be chosen to be
(1.14)
Cp = min
β>0
2e β max(1,
e β
pp2βe).
Remark 1.18. From the definition of [A0, A1]θ it is obvious (cf. part (iii) of Theo-
rem 2.7) that
1,p
0
(U , ω0),W
1,p
0
hW
(U , ω1)iθ ⊂£W 1,p (U , ω0) ,W 1,p (U , ω1)¤θ .
We can thus combine this inclusion together with (1.11) and (1.12) to obtain one
chain of inclusions.
p can assume the value 0, it
will be helpful for us to permit ourselves the following slight abuse of notation:
Although the non-negative function ωθ¯¯∇log (rω)¯¯
Definition 1.19. Let Lp¡U , ωθ¯¯∇log (rω)¯¯
tions φ : U → C for which the semi-norm, which we shall denote by
p¢ denote the space of measurable func-
(1.15)
is finite.
°°φ°°Lp(U,ωθ∇ log(rω)p) =µRd¯¯φ(x)¯¯
p
p¯¯∇log (rω(x))¯¯
1
p
,
ωθ(x)d x¶
10
MICHAEL CWIKEL AND AMIT EINAV
While Theorem 1.16 is quite general, it might also be desirable to have a "cleaner"
1,p
0
version of it which deals only with spaces of the kind "W 1,p " or only with spaces
of the kind "W
" rather than a mixture of these two kinds. Such versions can
indeed be obtained when suitable additional conditions are imposed on U and/or
on the weight functions ω0 and ω1. The following two theorems give instances
of this.
Theorem 1.20. Let U , ω0 and ω1 be as in Theorem 1.16. Assume furthermore
(U , ω1)
that, for some p ∈ [1,∞), whenever φ is an element of W
there exists a sequence©φnªn∈N of functions in Li pc(U ) which converges to φ in
W 1,p (U , ω0) or in W 1,p (U , ω1) and is a bounded sequence in the other space.
Then, as well as the inclusions (1.11) and (1.12), we also have that
(U , ω0)∩ W
1,p
0
1,p
0
(1.16)
W
p
0 (U , θ, rω) ⊂hW
1,p
0
(U , ω0) ,W
1,p
0
(U , ω1)iθ ⊂ W
1,p
0
(U , ωθ)
for that same value of p and for all θ ∈ (0,1).
Theorem 1.21. Let U , ω0 and ω1 be as in Theorem 1.16. Assume furthermore
that, for some p ∈ [1,∞),
(1.17)
W
and
(1.18)
W
1,p
0
¡U , ω j¢ = W 1,p¡U , ω j¢ ,
0 (U , θ, rω,) = W p (U , θ, rω),
p
j = 0,1,
θ ∈ (0,1).
Then, as well as the inclusions (1.11) and (1.12), we also have that
(1.19)
W p (U , θ, rω) ⊂£W 1,p (U , ω0) ,W 1,p (U , ω1)¤θ ⊂ W 1,p (U , ωθ)
for that same value of p and for all θ ∈ (0,1).
In particular if U = Rd and if ω0 and ω1 satisfy the conditions imposed in The-
orem 1.16, then (1.17) and (1.18) both hold for every p ∈ [1,∞) and therefore so
does (1.19).
With the help of Theorems 1.20 and 1.21 we are able to obtain the follow-
ing theorem which identifies new conditions which suffice for a Stein-Weiss like
theorem to hold for pairs of W 1,p spaces:
Theorem 1.22. Let U be a non-empty open subset of Rd and let g : U → R be a
Lipschitz function. Let ω0 and ω1 be weight functions on U such that at least one
of them satisfies the compact boundedness condition and
ω0(x) = ω1(x)e g (x).
(i) Then the conditions of Theorem 1.16 are satisfied.
(ii) If U = Rd then
(1.20)
hW 1,p³Rd , ω0´,W 1,p³Rd , ω1´iθ = W 1,p³Rd , ωθ´
to within equivalence of norms, for every p ∈ [1,∞) and θ ∈ (0,1).
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
11
(iii) If the open set U and the above two weight functions ω0 and ω1 satisfy the
conditions of Theorem 1.20 then
1,p
0
(U , ω0) ,W
1,p
0
hW
(U , ω1)iθ = W
1,p
0
(U , ωθ)
(iv) If the number p, the open set U and the above two weight functions ω0 and
to within equivalence of norms, for every p ∈ [1,∞) and θ ∈ (0,1).
ω1 satisfy the conditions of Theorem 1.21 then
£W 1,p (U , ω0) ,W 1,p (U , ω1)¤θ = W 1,p (U , ωθ)
to within equivalence of norms, for that value of p and for every θ ∈ (0,1).
1.4. Structure of the paper. In §2 we will recall some basic definitions and facts
about interpolation spaces, and in particular Calderón's complex interpolation
spaces.
In §3 we will show how a generalisation of the Stein-Weiss theorem
p
1,p
0
easily implies the inclusion of£W 1,p (U , ω0) ,W 1,p (U , ω1)¤θ in W 1,p (U , ωθ) and
0 (U , θ, rω) in£W 1,p (U , ω0),W 1,p (U , ω1)¤θ, which will eventually yield our main
we shall also consider the analogous inclusion for the W
spaces. §4 is the
main section of this work. In it we will discuss the more difficult inclusion of
W
result. With the help of some approximation theorems which we will present in
§5, we will prove our main theorem and several interesting consequences of it in
§6. In §7 we will discuss further properties of W p (U , θ, rω), such as the fact that
it can sometimes happen that W p (U , θ, rω) is strictly smaller than W 1,p (U , ωθ).
In the final section §8 we will share some final thoughts about this, and future,
work. These will include a discussion of potential applications to rates of con-
vergence in evolution equations.
There is also an appendix at the end of the paper. It contains some auxiliary
proofs which we felt should not hinder the flow of our presentation in the pre-
ceding sections.
Acknowledgement. A. Einav would like to thank Anton Arnold and Tobias Wöhrer
for stimulating discussions about using interpolation theory to estimate rates of
convergence to equilibrium in Fokker-Planck equations, which led him to start
investigating the main problem presented in this work.
2. PRELIMINARIES: INTERPOLATION SPACES AND THE COMPLEX INTERPOLATION
METHOD
This short section has been included for the convenience of those readers
who may not be familiar with the theory of interpolation spaces, in particular of
those which are generated by the complex interpolation method (sometimes re-
ferred to simply as the complex method). We briefly recall some basic definitions
and some basic properties of these interpolation spaces. For more information
about this subject we refer the reader to Calderón's fundamental paper [3] or to
[2] or many other excellent textbooks. The notations and definitions which we
shall use are fairly standard.2
2But note that there are some differences between the notation used in [3] and in [2].
12
MICHAEL CWIKEL AND AMIT EINAV
The basic setting of interpolation theory is a pair of normed vector spaces A0
and A1 which have the property that there exists a Hausdorff topological vector
space A satisfying
A0 ⊂ A ,
A1 ⊂ A ,
where the inclusions of A0 and A1 in A are linear and continuous. If A0 and
A1 are also Banach spaces, then it is customary to say that (A0, A1) is a Banach
couple, and also to sometimes use the abbreviated notation A instead of (A0, A1).
Definition 2.1. Let A = (A0, A1) be a Banach couple. A normed vector space A is
called an intermediate space with respect to (A0, A1) if
A0 ∩ A1 ⊂ A ⊂ A0 + A1
with continuous inclusions, where the norms on A0 ∩ A1 and A0 + A1 are given
by
and
kakA0∩A1 = max¡kakA0 ,kakA1¢
kakA0+A1 = inf©ka0kA0 +ka1kA1 a0 + a1 = aª.
The notation T : (A0, A1) → (A0, A1) (or T : A → A for short) is taken to mean that
T is a linear operator which maps A0+ A1 into itself and also satisfies T (A0) ⊂ A0
and T (A1) ⊂ A1 and is bounded from A0 into A0 and from A1 into A1.
An intermediate space A is called an interpolation space with respect to (A0, A1)
if it has the additional property that every linear operator T : (A0, A1) → (A0, A1)
is also a bounded map of A into itself.
We often need to consider a more elaborate version of the previous definition,
for operators which map between the spaces of two possibly different Banach
couples:
Definition 2.2. Let A = (A0, A1) and B = (B0, B1) be Banach couples. The nota-
tion T : (A0, A1) → (B0, B1) (or T : A → B for short) is taken to mean that T is a
linear operator which maps A0 + A1 into B0 + B1 and also satisfies T (A0) ⊂ B0
and T (A1) ⊂ B1 and is bounded from A0 into B0 and from A1 into B1. Let A
be an intermediate space with respect to (A0, A1) and let B be an intermedi-
ate space with respect to (B0, B1). Then we say that A and B are relative in-
terpolation spaces with respect to (A0, A1) and (B0, B1) if every linear operator
T : (A0, A1) → (B0, B1) is a bounded map of A into B .
There are many methods for creating interpolation spaces with respect to a
Banach couple and relative interpolation spaces with respect to a pair of Ba-
nach couples. In our work here we will only concern ourselves with the so-called
complex method, which was introduced and developed by Calderón in [3], and
which we will now describe.
In what follows we will let S denote the infinite strip {z ∈ C 0 ≤ Rez ≤ 1} in the
complex plane.
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
13
Definition 2.3. Let A = (A0, A1) be a Banach couple of complex Banach spaces.
We define F (A0, A1), also denoted by F³A´, to be the space
F³A´ =½ f : S →A0 + A1 f is continuous and bounded on S and analytic on int (S)
t → f ( j + i t ) is a continuous bounded map of R into A j , for j = 0,1¾,
t∈R°° f (i t )°°A0
,sup
t∈R°° f (1+ i t )°°A1¶.
and define a norm on F³A´ by
°° f°°F³A´ = maxµsup
It can very readily be shown (using e.g., obvious very minor modifications of the
argument in paragraph 22 of [3, p. 129]) that F³A´ with the above norm is a
Banach space, and we will make use of this fact.
For each given θ ∈ [0,1] we define [A0, A1]θ to be the linear space of all elements
of a ∈ A0 + A1 for which there exists f ∈ F³A´ such that a = f (θ). This space,
when equipped with the norm
kak[A0,A1]θ = inf½°° f°°F³A´ f ∈ F³A´, f (θ) = a¾,
is a Banach space.
The basic and very useful interpolation property of the spaces defined by the
previous definition is expressed by the following theorem of Calderón. (See §4
of [3, p. 115], also e.g., Theorem 4.1.2 of [2, p. 88].)
Theorem 2.4. Let A = (A0, A1) and B = (B0, B1) be Banach couples of complex Ba-
nach spaces. Then [A0, A1]θ and [B0, B1]θ are relative interpolation spaces with
respect to A and B for each θ ∈ (0,1). Furthermore, for each linear operator T :
A → B , the norms of T as a operator from A0 to B0, from A1 to B1 and from
[A0, A1]θ to [B0, B1]θ satisfy the inequality
kTkB([A0,A1]θ,[B0,B1]θ) ≤ kTk1−θ
B(A0,B0)kTkθ
B(A1,B1) .
Remark 2.5. The above definition, describing the complex method for creating
interpolation spaces, indicates why one may think of such spaces as 'in between'
spaces. The original spaces A0 and A1 are somehow identified with elements of
F³A´ when restricted to Rez = 0 or Rez = 1. The θ−interpolation space [A0, A1]θ
is identified with elements of F³A´ restricted to the intermediate line Rez = θ.
Remark 2.6. Our definition here of the space F³A´, which is often, but not al-
that we have not required that limt→∞°° f ( j + i t )°°A j = 0 for j ∈ {0,1}. However it
ways, the definition chosen to appear in papers about the complex interpolation
method is in fact slightly different from the original definition appearing in [3] in
is well known (implicit already in [3] and mentioned e.g. in [4, p. 1007]) and very
14
MICHAEL CWIKEL AND AMIT EINAV
easy to check (by multiplying functions in F (A) by e β(z−θ)2
for arbitrarily small
positive β) that this difference does not effect the definition of [A0, A1]θ nor its
norm.
Here are some more properties of complex interpolation spaces which will be
relevant for our purposes:
and
Theorem 2.7. Let (A0, A1) be a Banach couple of complex Banach spaces and let
θ be a number in (0,1). Then
(i) A0 ∩ A1 is a dense subspace of [A0, A1]θ,
(ii) each element a of A0 ∩ A1 satisfies the inequality
A0 kakθ
.
(2.1)
(iii) Suppose, furthermore, that, for j ∈ {0,1}, B j is a closed subspace of A j equipped
kak[A0,A1]θ ≤ kak1−θ
with the same norm as A j . Then (B0, B1) is a Banach couple, and [B0, B1]θ ⊂
[A0, A1]θ, and each b ∈ [B0, B1]θ satisfies
A1
kbk[A0,A1]θ ≤ kbk[B0,B1]θ .
(iv) If A0 = A1 with equality of norms, then [A0, A1]θ = A0 = A1 with equality of
norms.
For the statement and proof of (i) we refer to page 116 and then pages 132 -- 133
of [3], or to Theorem 4.2.2 and Lemma 4.2.3 of [2, pp. 91 -- 92]. The rather obvi-
ous, immediate and well known proof of (ii) proceeds via the function f (z) =
a/kak1−z
which is a norm one element of F (A0, A1). Part (iii) is an im-
mediate consequence of Definition 2.3. Part (iv) is an obvious consequence of
Definition 2.3 and part (i).
A0 kakz
A1
Remark 2.8. We note that part (iv) of this theorem combined with Remark 1.11,
justifies the claim made in Remark 1.1.
With the basics of interpolation theory and the complex method in hand, we
can start proving our main theorem.
3. THE SIMPLER INCLUSION
Before beginning our systematic study of the interpolation spaces
we of course need to be sure that
£W 1,p (U , ω0),W 1,p (U , ω1)¤θ
¡W 1,p (U , ω0),W 1,p (U , ω1)¢ is a Banach couple,
(3.1)
when we either assume that W 1,p (U , ω0) and W 1,p (U , ω1) are both Banach spaces
or impose explicit conditions on U , ω0 and ω1 which guarantee that. In fact (3.1)
holds under such assumptions, because both of the spaces W 1,p (U , ω0) and
W 1,p (U , ω1) are continuously embedded in the Banach space Lp (U , ω) when ω
is chosen to be the weight function ω = min {ω0, ω1}.
Our main theorem, Theorem 1.16, establishes two inclusions. The reader has
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
15
probably already noticed that one of them, (1.11), is actually an obvious and
almost immediate consequence of the Stein-Weiss theorem, or of a slightly gen-
eralised version of it due to Calderón. Nevertheless, we will provide an explicit
proof of that inclusion in this section. We will also consider the case where this
inclusion can be generalised to the W
Here is the version that we need of Calderón's generalization of the Stein-Weiss
theorem.
spaces.
1,p
0
Theorem 3.1. The formula
£Lp¡U , ω0, Cm¢ , Lp¡U , ω1, Cm¢¤θ = Lp¡U , ωθ, Cm¢ ,
(3.2)
holds with equality of norms for every m ∈ N, every p ∈ [1,∞), every θ ∈ (0,1),
every open subset U of Rd , and all pairs of weight functions ω0 and ω1 on U ,
when ωθ is defined by (1.5).
Remark 3.2. In fact the formula (3.2) is a special case of the isometric formula
£Lp0 (U , ω0, Cm), Lp1¡U , ω1, Cm¢¤θ = Lpθ¡U , ωθ,pθ , Cm¢
(3.3)
which holds for every p0 and p1 in [1,∞) and each θ ∈ (0,1) when we take pθ
and ωθ,pθ to be the number p and the function ω specified by (1.1). This can be
proved by a straightforward variant of the proof of Theorem 2 in [20]. It can also
be seen to be a very special case of a result3 in [3].
We shall now use this theorem to easily show the aforementioned inclusion,
as well as the associated norm inequality.
Theorem 3.3. Let U be an open subset of Rd and suppose that ω0 and ω1 are
weight functions which both satisfy the compact boundedness condition on U .
Then, for each p ∈ [1,∞), we obtain that¡W 1,p (U , ω0),W 1,p (U , ω1)¢ is a Banach
couple and that, for each θ ∈ (0,1), its complex interpolation spaces satisfy
and
(3.4)
£W 1,p (U , ω0),W 1,p (U , ω1)¤θ ⊂ W 1,p (U , ωθ) ,
°°φ°°W 1,p (U,ωθ) ≤°°φ°°[W 1,p (U,ω0),W 1,p (U,ω1)]θ
,
Proof. In view of Lemma 5.2, W 1,p (U , ω0) and W 1,p (U , ω1) are both Banach
for every φ ∈£W 1,p (U , ω0),W 1,p (U , ω1)¤θ.
spaces, and therefore also, as explained at the beginning of this section,¡W 1,p (U , ω0) ,W 1,p (U , ω1)¢
is a Banach couple.
Define a linear operator
T : W 1,p (U , ω0)+ W 1,p (U , ω1) → Lp³U , ω0, Cd+1´+ Lp³U , ω1Cd+1´
3It is a special case of the isometric formula [X0(B0), X1(B1)]θ = X 1−θ
0
X θ
stated as the second part of the result labelled "i)" on p. 125 of [3], holds for all choices of the
Banach couple (B0, B1) whenever X0 and X1 are Banach lattices of measurable functions on the
same underlying σ-finite measure space and X 1−θ
X θ
1 has a certain property. That property holds
in our case in view of the Lebesgue Dominated Convergence Theorem.
0
1 ¡[B0, B1]θ¢ which, as
16
by
MICHAEL CWIKEL AND AMIT EINAV
T φ :=¡φ,∇φ¢ .
By the definition of W 1,p (U , ω) we have that T φ is a norm 1 map from W 1,p¡U , ω j¢
into Lp¡U , ω j , Cd+1¢ for j = 0,1. Using the standard interpolation theorem, The-
orem 2.4, and Theorem 3.1, we conclude that
T :£W 1,p (U , ω0),W 1,p (U , ω1)¤θ → Lp³U , ωθ, Cd+1´
is bounded with norm less than, or equal to, 1. Thus, if φ ∈£W 1,p (U , ω0),W 1,p (U , ω1)¤θ
then¡φ,∇φ¢ ∈ Lp¡U , ωθ, Cd+1¢, i.e. φ ∈ W 1,p (U , ωθ). Moreover,
°°φ°°W 1,p (U,ωθ) =°°¡φ,∇φ¢°°Lp(U,ωθ,Cd+1) =°°T φ°°Lp(U,ωθ,Cd+1)
=°°T φ°°[Lp(U,ω0,Cd+1),Lp(U,ω1,Cd+1)]θ ≤°°φ°°[W 1,p (U,ω0),W 1,p (U.ω1)]θ
This concludes the proof of this theorem. Thus it also establishes (1.11) and the
left side of the norm inequality (1.13) in the statement of Theorem 1.16.
(cid:3)
.
Remark 3.4. It should perhaps also be pointed out that the proof of Theorem 3.3
does not require the weight functions ω0 and ω1 to satisfy the compact bound-
edness condition. In fact it suffices if they have any other property which en-
sures that W 1,p (U , ω0) and W 1,p (U , ω1) are both Banach spaces. (For example,
some such properties are considered in the latter part of Lemma 5.2.) Essentially
the same proof as for Theorem 3.3 also shows that, for any p0 and p1 in [1,∞),
if W 1,p0 (U , ω0) and W 1,p1 (U , ω1) are both Banach spaces then they satisfy the
norm one inclusion
£W 1,p0 (U , ω0),W 1,p1 (U , ω1)¤θ ⊂ W 1,p (U , ω)
for each θ ∈ (0,1), where p and ω are as in (1.1). This is an immediate conse-
quence of the isometry (3.3). We have not yet considered under what conditions
we can use some version of our approach to obtain the reverse of this inclusion
when p0 6= p1. Such a result would complement the particular case considered
in Formula (5.3) of Theorem 4 of [14, p. 208] in which U = Rd and ω0, ω1 and ωθ
are all required to be so-called polynomially regular weight functions.
We still have one more task in this section, which is to consider under what
conditions we can replace the spaces W 1,p (U , ω) by their subspaces W
(U , ω)
in the inclusion (1.11). We do this in the following theorem which will lead us
eventually to Theorem 1.20.
1,p
0
1,p
0
(U , ω0)∩ W
Theorem 3.5. Let U be an open subset of Rd and suppose that ω0 and ω1 are
weight functions which both satisfy the compact boundedness condition on U .
Suppose further that, for some value of p ∈ [1,∞), whenever φ is an element of
(U , ω1), then there exists a sequence©φnªn∈N of functions in
W
Li pc(U ) which converges to φ in W 1,p (U , ω0) or W 1,p (U , ω1) and is a bounded
sequence in the other space. Then, for that value of p and for each θ ∈ (0,1),
hW
(U , ω1)iθ ⊂ W
(U , ω0) ,W
(U , ωθ)
1,p
0
1,p
0
1,p
0
1,p
0
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
17
and, furthermore,
(3.5)
for each φ ∈hW
1,p
0
1,p
0
°°φ°°W
(U , ω0),W
(U,ωθ) ≤°°φ°°hW
(U , ω1)iθ
1,p
0
.
1,p
0
(U,ω0),W
1,p
0
(U,ω1)iθ
Remark 3.6. We suspect that, given a pair of weight functions ω0 and ω1 which
are known to both satisfy the compact boundedness condition, quite mild ad-
ditional conditions might suffice to ensure that they also satisfy the second hy-
pothesis required in Theorem 3.5. Perhaps it is even not necessary to impose any
additional conditions for this to happen. Pending further investigation of this,
we simply point out that this second hypothesis is obviously satisfied whenever
one of the weight functions ω0 and ω1 is dominated by some constant multiple
of the other. In that case there even exists a sequence©φnªn∈N in Li pc (U ) which
converges to φ in both W 1,p (U , ω0) and W 1,p (U , ω1).
Proof of Theorem 3.5. As pointed out in Remark 1.5, the fact that ω0 and ω1 both
satisfy the compact boundedness condition immediately implies that ωθ also
satisfies this condition for each θ ∈ (0,1). Consequently, by Lemma 5.2, W 1,p (U , ω)
is a Banach space which contains Li pc (U ) whenever ω is any one of the weight
functions ω0, ω1 or ωθ. This enables us to define W
(U , ω) in accordance with
Definition 1.13 for each of these choices of ω. Then the relevant definitions im-
1,p
0
mediately imply (cf. part (iii) of Theorem 2.7) that³W
a Banach couple, and that
1,p
0
(U , ω0),W
1,p
0
(U , ω1)´ is
and that
(3.6)
1,p
0
1,p
0
(U , ω0),W
hW
°°φ°°[W 1,p (U,ω0),W 1,p (U,ω1)]θ ≤°°φ°°hW
(U , ω1)iθ ⊂£W 1,p (U , ω0),W 1,p (U , ω1)¤θ
(U,ω1)iθ
(U,ω0),W
1,p
0
1,p
0
.
for each φ ∈hW
1,p
0
(U , ω0),W
1,p
0
By definition, Li pc(U ) is contained in W
fore also in W
that
1,p
0
1,p
0
(U , ω1)iθ
.
1,p
0
(U , ω0)∩ W
Li pc(U ) ⊂hW
1,p
0
(U , ω0),W
1,p
0
(U , ω1)iθ
.
(U , ω1) and there-
(U , ω1). This implies (cf. part (i) of Theorem 2.7)
(U , ω0) and in W
1,p
0
1,p
0
In view of this inclusion and since we also know that Li pc(U ) is also contained
in W
(U , ωθ), we are now ready to proceed to our next step of the proof, which
will be to show that the inequality (3.5) holds for every φ ∈ Li pc(U ). By defini-
(U,ω) =°°φ°°W 1,p (U,ω) for such functions φ and this, combined with
the inequality (3.4) of Theorem 3.3 and (3.6), indeed accomplishes this step.
In view of the standard fact recalled in part (i) of Theorem 2.7, we can assert that
. Therefore,
W
tion °°φ°°W
(U , ω0),W
1,p
0
1,p
0
(U , ω0)∩ W
1,p
0
(U , ω1) is dense inhW
1,p
0
since W
we show that each element φ in W
(U , ωθ) is a Banach space, the proof of this theorem will be complete if
(U , ωθ)
(U , ω1) is also in W
1,p
0
1,p
0
(U , ω0)∩W
1,p
0
1,p
0
(U , ω1)iθ
1,p
0
18
MICHAEL CWIKEL AND AMIT EINAV
and satisfies the inequality (3.5).
(U , ω1). We keep in mind
Let φ be an arbitrary element of W
that Theorem 3.3 combined with parts (i) and (iii) of Theorem 2.7 guarantee
(U , ω0)∩ W
1,p
0
1,p
0
with the properties guaranteed by the hypotheses of the present theorem, so
and
that φ must also be in W 1,p(U , ωθ). Now let©φnªn∈N be a sequence in Li pc(U )
that at least one of the two numerical sequences n°°φn − φ°°W 1,p (U,ω0)on∈N
n°°φn − φ°°W 1,p (U,ω1)on∈N
converges to 0 and both are bounded. We now use the
result about the complex interpolation method which is recalled in part (ii) and
inequality (2.1) of Theorem 2.7. In our current context the inequality (2.1) be-
comes
°°φn − φ°°hW
1,p
0
(U,ω0),W
1,p
0
which gives us that
(3.7)
(U,ω1)iθ ≤°°φn − φ°°
=°°φn − φ°°
lim
n→0°°φ− φn°°hW
1,p
0
(U,ω0),W
1,p
0
θ
W
1,p
0
(U,ω1)
θ
W 1,p (U,ω1)
1−θ
W
1,p
0
1−θ
(U,ω0) ·°°φn − φ°°
W 1,p (U,ω0) ·°°φn − φ°°
(U,ω1)iθ = 0,
and therefore also (cf. part (iii) of Theorem 2.7) that limn→0°°φ− φn°°[W 1,p (U,ω0),W 1,p (U,ω1)]θ =
0. This in turn implies, by Theorem 3.3 that
(3.8)
lim
n→0°°φ− φn°°W 1,p (U,ωθ) = 0
and therefore φ is not only an element of W 1,p (U , ωθ) but also of W
So we can deduce from (3.8) and Definition 1.13 that
1,p
0
(U , ωθ).
1,p
0
°°φ°°W
(U,ωθ) =°°φ°°W 1,p (U,ωθ) = lim
Since an earlier step of this proof gives us that each φn satisfies (3.5), the above
(U,ω1)iθ
n→0°°φn°°W 1,p (U,ωθ) = lim
limit must be less than or equal to the limit limn→∞°°φn°°hW
which exists and equals°°φ°°hW
(U , ω1), as well as be-
(U , ωθ), also satisfies (3.5). As already explained above, this suffices
(cid:3)
1,p
0
1,p
shown that our arbitrary element φ of W
0
ing in W
to complete the proof of our theorem.
(U,ω1)iθ
(U , ω0)∩ W
n→0°°φn°°W
in view of (3.7). Thus we have
(U,ωθ) .
1,p
0
(U,ω0),W
(U,ω0),W
1,p
0
1,p
0
1,p
0
1,p
0
1,p
0
In the next section we will make the significant step towards showing the
other inclusion and inequality of our main theorem. This, however, will not be
as simple.
4. TOWARDS THE DIFFICULT INCLUSION
W
1,p
0
The goal of this section is to consider the relationship between the two spaces
p
. We do not find it surprising that
in general we cannot obtain an analogous relationship between the "full" spaces
0 (U , θ, rω) andhW
W p (U , θ, rω) and£W 1,p (U , ω0),W 1,p (U , ω1)¤θ, and can only deal with their above
mentioned smaller but "significant" respective subspaces. However, as we have
(U , ω1)iθ
(U , ω0) ,W
1,p
0
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
19
mentioned in §1, in many cases these smaller spaces are either everything, or
"big" enough to warrant a theorem on their own.
We begin with a definition which specifies some conditions which may or may
not be satisfied by a given pair of weight functions ω0 and ω1. As we shall see, for
most of the results of this section, we will require our weight functions to satisfy
these conditions.
Definition 4.1. We say that a pair of weight functions (ω0, ω1) is a (θ, p)−admissible
pair of weight functions on U (or, in short, a (θ, p)−admissible pair) for some
θ ∈ (0,1) and p ∈ [1,∞] if the following holds:
(i) W 1,p (U , ω0), W 1,p (U , ω1) and W 1,p (U , ωθ) are Banach spaces that contain
Li pc (U ).
(ii) The function log (rω(x)) = log³ ω0(x)
ω1(x)´ is locally Lipschitz on U .
Remark 4.2. In view of the inequality ω1−θ
1 ≤ max {ω0, ω1} it is easy to see that
if W 1,p(U , ω0) and W 1,p (U , ω1) both contain Li pc(U ), then in fact it follows au-
tomatically that W 1,p (U , ωθ) also contains Li pc(U ) for every θ ∈ (0,1).
0 ωθ
From this point onward, till the end of this section, we will assume that we
are dealing with a given fixed open set U and often we will not mention U in the
notation that we use.
Remark 4.3. It is simple to observe that if (ω0, ω1) forms a (θ, p)−admissible pair
for some θ ∈ (0,1) and p ∈ [1,∞), then W p (θ, rω) also contains Li pc. (Relevant
definitions, i.e. of W p (θ, rω) and of rω, appear in the statement of Theorem 1.16.)
Indeed, given any φ ∈ Li pc, we first note that it is contained in W 1,p (ωθ), in view
of condition (i) of Definition 4.1. Then the other requirement for the member-
ship of φ in W p (θ, rω) follows immediately from condition (ii) of the same def-
inition, which ensures that the auxiliary function φ(x)¯¯∇log (rω(x))¯¯ is continu-
ous. Since that function also has compact support, it is obviously an element of
Lp (ωθ) as required.
Remark 4.4. In most of our main results in this paper we require ω0 and ω1 to
both satisfy the compact boundedness condition. In view of Remark 1.5, Lemma
5.2 and Lemma 5.6, this implies that (i) holds for all θ ∈ (0,1) and p ∈ [1,∞).
From this we see that Definition 4.1, and the results in this section which use
it, are somewhat more general than they need to be for obtaining those main
results. But this greater generality may be useful in future investigations where
more general weight functions are considered.
Our main goal in this section is to prove the following result:
Theorem 4.5. Let (ω0, ω1) form a (θ, p)−admissible pair for some θ ∈ (0,1) and
p ∈ [1,∞). Then, we have that
(4.1)
W
p
(ω0),W
0 (θ, rω) ⊂hW
1,p
0
Moreover, there exists a positive constant Cp which depends only on p, such that
(4.2)
°°φ°°[W 1,p (ω0),W 1,p (ω1)]θ ≤ Cp°°φ°°W p (θ,rω)
1,p
0
(ω1)iθ
for every φ ∈ W
p
0 (θ, rω).
20
MICHAEL CWIKEL AND AMIT EINAV
The proof of Theorem 4.5 has two main parts and will occupy the rest of this
section. In the first part, for each given φ ∈ Li pc, we will construct an element
f φ of F³W
(ω1)´ which satisfies f φ(θ) = φ and also the norm esti-
mates necessary for showing that φ satisfies (4.2). Then, in the second part, we
will extend this construction to elements φ which are in the closure of Li pc in
W p (θ, rω).
(ω0) ,W
1,p
0
1,p
0
Definition 4.6. Given θ ∈ (0,1), an arbitrary function φ ∈ Li pc and a number
β > 0, we define the function
(z−θ)
(4.3)
log(rω(x))+β(z−θ)2
φ(x),
φ
p
f
β,θ(x, z) = e
x ∈ U , z ∈ S.
φ
(Recall, cf. (1.9) that rω denotes the ratio ω0/ω1.) Thus f
β,θ denotes the map
from S into the space of measurable functions on U , which is such that the im-
age f
β,θ(z) of each z ∈ S is the function of x defined by (4.3).
φ
φ
The reader might be surprised by the inclusion of another parameter β > 0 in
the formula (4.3). This parameter appears in the exponential factor e β(z−θ)2
in
φ
β,θ, and we need that factor in order to control the size of the
the formula for f
β,θ(x, z) as z tends to ∞ on the strip S. (The same function e β(z−θ)2
has a quite
f
similar role elsewhere, as mentioned in Remark 2.6.) As will be evident in the
course of the proofs of the results presented in this section, β could be chosen
arbitrarily, but we can also note that there is one choice of β which will minimize
the value that we can obtain for the above mentioned constant.
For convenience, we will often drop the subscripts of β and θ and accordingly
write f φ to denote the function f
We begin our preparations for achieving the first of the above mentioned main
parts of our proof by recalling a rather standard definition and mentioning an
important theorem which we will use in several places.
φ
β,θ.
1,p
Definition 4.7. For each p ∈ [1,∞] and each open subset U of Rd , the space
l oc (U ) consists of all measurable functions f : U → C which have a weak gra-
W
dient ∇ f in U and for which, for every compact subset K of U , the functions χK f
and χK ∇ f are elements, respectively, of the (unweighted) spaces Lp (U ) and of
Lp (U , Cd ).
We remark that it is quite straightforward, but perhaps a little tedious, to
1,p
l oc (U ) if and only if, for every
check that a function f : U → C is an element of W
open subset V of U whose closure is a compact subset of U , the restriction of f
to V is an element of W 1,p (V ). We refer the interested reader to the appendix for
the proof of this fact.
Theorem 4.8 (Rademacher's Theorem). Let U be an open set. Then,
(1) a function φ : U → C is locally Lipschitz if and only if φ ∈ W 1,∞l oc (U ).
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
21
(2) Moreover, if φ is locally Lipschitz on U , then it is differentiable almost
everywhere4 and its gradient equals its weak gradient almost everywhere5.
The proof of this theorem can be obtained from material in Chapter 5 of [6]
(subchapter 5.8.2 b., pages 279-281). Part (1), as pointed out in the remark on p.
280 of [6], can be obtained by an easy adaptation of the arguments of the proof of
Theorem 4 on pp. 279 -- 280. Part (2) is the case p = ∞ of Theorem 5 ([6, pp. 280 --
281]) combined with Theorem 6 ([6, p. 281]). The alternative characterisation of
W
1,p
l oc (U ) mentioned just after Definition 4.7 is relevant here.
Remark 4.9. When we use Theorem 4.8, it is sometimes helpful to also keep in
mind the very obvious fact that, at every point x ∈ U where the pointwise gradi-
is the Lipschitz constant of the restriction of φ to a bounded open set containing
x.
ent ∇φ(x) of a locally Lipschitz function φ exists, it satisfies¯¯∇φ(x)¯¯ ≤ L where L
Lemma 4.10. Let (ω0, ω1) form a (θ, p)−admissible pair for some θ ∈ (0,1) and
p ∈ [1,∞), and let φ ∈ Li pc. Then, for each fixed z ∈ S, we have that f φ(x, z) ∈
Li pc and consequently
f φ(·, z) ∈ W
(ω1) ⊂ W
(ω0)∩ W
(ω0)+ W
(ω1).
(4.4)
1,p
0
1,p
0
1,p
0
1,p
0
Before we begin proving this lemma, it is convenient to present a few simple
facts that we will use in the proof of this lemma, as well as in subsequent proofs:
(a) If the function ψ : U → C is locally Lipschitz and has compact support, then
it is Lipschitz on all of U . This rather standard result is perhaps slightly less
obvious than may seem at first glance. For the sake of completeness, we
include its quite straightforward and short proof in the appendix.
(b) Every Lipschitz function is bounded on every compact set. Therefore every
locally Lipschitz function also has this property.
(c) The pointwise product of two bounded Lipschitz functions is itself also a
bounded Lipschitz function. So, by (b), the pointwise product of two locally
Lipschitz functions is also locally Lipschitz.
(d) The composition of two Lipschitz functions is a Lipschitz function. There-
fore the composition of two locally Lipschitz functions is a locally Lipschitz
function.
(e) A function mapping R to C whose real and imaginary parts have bounded
derivatives on every bounded interval is a locally Lipschitz function.
p
(z−θ)
t+β(z−θ)2
Proof of Lemma 4.10. Let us define g1 : U → R by g1(x) = log (rω(x)) and define
. Then f φ(x, z) = φ(x)g2¡g1(x)¢ for all x ∈ U .
g2 : R → C by g2(t ) = e
Since φ has compact support so does φ(x)g2¡g1(x)¢. So, in view of fact (a) it
suffices to show that φ(x)g2¡g1(x)¢ is locally Lipschitz, which we will now do
4When we say "almost everywhere" we always mean with respect to d−dimensional Lebesgue
5This equality is to be understood in the following way: each component of the pointwise
gradient belongs to the equivalence class of the corresponding component of the weak gradient.
measure.
22
MICHAEL CWIKEL AND AMIT EINAV
(cid:3)
1,p
0
1,p
0
lemma.
(ω0),W
(ω1)´.
with the help of the other above mentioned obvious facts.
We first use (e) to show that g2 is locally Lipschitz. The (θ, p)−admissibility of
(ω0, ω1) implies that g1 is also locally Lipschitz. So (d) gives us that g2 ◦ g1 is
locally Lipschitz. Finally, since φ is obviously locally Lipschitz, (c) completes the
proof that φ(x)g2¡g1(x)¢ is locally Lipschitz and therefore also the proof of this
Lemma 4.11. Let (ω0, ω1) form a (θ, p)−admissible pair for some θ ∈ (0,1) and
p ∈ [1,∞), and let φ ∈ Li pc. Then the function f φ is an element of the space
F³W
Proof. For reasons which the reader can quite possibly guess and which anyway
will become apparent later, we begin this proof by establishing some properties
of the functions¡log (rω(x))¢k and¡log (rω(x))¢k
φ(x) for each non-negative inte-
ger k. We first note that the (θ, p)−admissibility of (ω0, ω1), together with facts
(d) and (e) that were stated after Lemma 4.10, ensure that, for each such k, the
function¡log (rω(x))¢k is locally Lipschitz. Consequently, using fact (c) and then
fact a, we obtain that¡log (rω(·))¢k
φ(·) is in Li pc. This in turn implies, again be-
cause (ω0, ω1) is (θ, p)−admissible, that this same function is also an element of
W
Our next task will be to estimate the norm of this function in each of these
spaces.
The fact that φ has compact support ensures the existence of an open set Vφ
whose closure is a compact subset of U such that at every point x ∈ U \ Vφ we
have that φ(x) = 0 and also that the pointwise gradient ∇φ(x) exists and also
equals 0.
We observe that, for j ∈ {0,1},
(ω0) and of W
(ω1).
1,p
0
1,p
0
φ(·)°°°
°°°¡log (rω(·))¢k
x∈Vφ¯¯log (rω(x))¯¯!pk
≤Ãsup
p
p
p¯¯log (rω(x))¯¯
Lp(ω j ) = Λpk°°φ°°
Lp(ω j ) =U¯¯φ(x)¯¯
°°φ°°
x∈Vφ¯¯log(rω(x))¯¯
Λ := sup
pk
ω j (x)d x
p
Lp(ω j )
,
where Λ denotes the quantity
which is finite, as the supremum of a Lipschitz function on a bounded set.
Next we need to make some more preparations for performing a somewhat more
elaborate calculation to estimate the size of°°°∇³¡log (rω(·))¢k
We note the following:
• For any a, b ∈ R we have that
φ(·)´°°°
(4.5)
(a+b)p ≤ 2p¡ap +bp¢,
p ≥ 0.
p
.
Lp(ω j )
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
23
• By part (2) of Theorem 4.8 we know that the pointwise gradients∇log (rω(x))
and ∇φ(x) of the locally Lipschitz functions log (rω(x)) and φ(x) both ex-
ist for all x in a certain subset U# of U which is such that U \U# has mea-
sure 0. We shall let Ξ denote the supremum
(4.6)
Ξ := sup
x∈Vφ∩U#¯¯∇log (rω(x))¯¯ ,
which is obviously finite (cf. Remark 4.9).
p
1
p
p
Lp(ω j )¶
φ(x).
another application of Rademacher's theorem tells us that the pointwise
statements hold for the pointwise and weak gradients of the functions φ
• Furthermore it is also obvious that the pointwise gradient of¡log (rω(x))¢k
exists and equals k¡log (rω(x))¢k−1
∇log (rω(x)) for every x ∈ U#. As al-
ready observed above, the function¡log (rω(x))¢k is locally Lipschitz. So
gradient of¡log(rω(x))¢k on U# is also its weak gradient on U . Analogous
and¡log (rω(x))¢k
Having made these preparations, we can now see that, for j ∈ {0,1} and for each
k ∈ N,
φ(·)´°°°
°°°∇³¡log (rω(·))¢k
+2p k pVφ¯¯φ(x)¯¯
Lp(ω j ) ≤ 2pVφ¯¯∇φ(x)¯¯
p¯¯∇log (rω(x))¯¯
≤ 2p Λp(k−1)³Λp°°∇φ°°
p¯¯log (rω(x))¯¯
Lp(ω j ) + k p Ξp°°φ°°
p¯¯log (rω(x))¯¯
Lp(ω j )´
ω j (x)d x
ω j (x)d x
p(k−1)
pk
p
W 1,p(ω j )
p
W 1,p(ω j )
φ(·) ∈ W
and for k = 0 we obviously obtain that
≤ 2p Λp(k−1)¡Λp + k p Ξp¢°°φ°°
°°°∇³¡log (rω(·))¢k
φ(·)´°°°
Lp(ω j ) ≤°°φ°°
We already noted above that ¡log (rω(·))¢k
can conclude furthermore, from the preceding estimates, that, for each k ∈ N,
0 (ω j ) =µ°°°¡log (rω(·))¢k
φ(·)°°°W
°°°¡log(rω(·))¢k
≤³Λpk°°φ°°
Lp(ω j ) + 2p Λp(k−1)¡Λp + k p Ξp¢°°φ°°
≤³2p Λp(k−1)¡2−p Λp + Λp + k p Ξp¢°°φ°°
W 1,p(ω j )´
≤ 2Λk−1¡2Λp + k p Ξp¢1/p°°φ°°W 1,p (ω j ) .
φ(·)°°°W
We estimate this last expression, using (4.5) again, but this time with 1
of p, and thus obtain that
Lp(ω j ) +°°°∇³¡log (rω(·))¢k
p Λ+ k Ξ´°°φ°°W 1,p (ω j ) .
°°°¡log (rω(·))¢k
0 (ω j ) ≤ 21+ 1
p Λk−1³2
W 1,p(ω j )´
φ(·)°°°
(ω j ) for j = 0,1. Now we
φ(·)´°°°
p in place
p
1
p
p
1
p
1,p
0
p
p
p
p
,
.
1,p
p
(4.7)
1,p
1
24
MICHAEL CWIKEL AND AMIT EINAV
The analogue of this inequality for k = 0 is just the trivial fact that
For each N ∈ N we define the function f
N : S → Li pc by setting
1,p
0
°°φ°°W
N (x, z) = e β(z−θ)2 N
φ
φ
(ω j ) =°°φ°°W 1,p (ω j ) .
Xk=0¡log (rω(x))¢k (z − θ)k
p k k!
φ(x).
(4.8)
f
In view of (4.7) we have that
Here, analogously to the convention adopted in Definition 4.6, it must be under-
φ
stood that f
N denotes the map from S into the space of measurable functions
on U , which is such that the image f
N (z) of each z ∈ S is the function of x de-
φ(·) is known to be
(ω0). Also, a simple
fined by (4.8). For each k ∈ N∪ {0}, the function¡log (rω(·))¢k
1,p
0
1,p
0
φ
(ω0) ∩ W
an element of Li pc and therefore also of W
calculation shows that the entire function
z 7→ e β(z−θ)2
(z − θ)k
1,p
0
is bounded on S. (Later we will do this calculation in more detail, to find an
explicit bound for the modulus of this function on S.) From these two facts it
φ
is obvious that f
N indeed maps S into Li pc and, furthermore, that it is an ele-
ment of F (A0, A1), where, throughout this proof, we will take (A0, A1) to be the
(ω0) and letting
Banach couple obtained by setting A0 = A1 = W
(ω0)∩ W
A0 and A1 be normed by°°ψ°°A0 =°°ψ°°A1 = maxn°°ψ°°W
next step will be to use the inequalities (4.7) to show thatn f
Xk=M¡log (rω(x))¢k (z − θ)k
sequence in F (A0, A1). For that it will clearly suffice to consider the quantity
1,p
0
is a Cauchy
(ω1)o. Our
for all integers M and N which satisfy 0 ≤ M ≤ N and to show that E (M, N ) tends
to 0 when M and N tend to ∞.
We see that
p k k!
1,p
0
1,p
0
N
N
p k k!
e β(j+i t−θ)2
E (M, N )≤
e β(z−θ)2 N
e β(z−θ)2¡log (rω(x))¢k (z − θ)k
E (M, N ) :=°°°°°
Xk=M°°°°°
t∈R ¯¯¯
¡ j + i t − θ¢k¯¯¯
φ(x)°°°A0 =°°°¡log (rω(x))¢k
°°°¡log (rω(x))¢k
≤ 21+ 1
p Λ+ k Ξ´ max
p Λk−1³2
°°φ°°A0 =°°φ°°Ai = max
°°°¡log (rω(x))¢k
φ(x)°°°A1
j∈{0,1}°°φ°°W 1,p(ω j )
j∈{0,1}°°φ°°W 1,p(ω j )
Xk=M
max
j=0,1
p k k!
sup
1
=
φ
(ω0) ,°°ψ°°W
NoN∈N
φ(x)°°°°°F (A0,A1)
φ(x)°°°°°F (A0,A1)
φ(x)°°°A j
.
for all k ∈ N. The analogue of this for k = 0 is simply and obviously
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
25
We also have, for each k ∈ N, that
sup
(4.9)
≤ sup
t≥0
max
j=0,1
t∈R¯¯¯
e β(1−t 2)¡1+ t 2¢
e β(j+i t−θ)2
¡ j + i t − θ¢k¯¯¯
e−β(1+t 2)¡1+ t 2¢
k
2
k
2 = e 2β sup
t≥0
e−βy 2
y k,
= e 2β sup
y≥1
and a simple computation shows that
(4.10)
y k e−βy 2
max
y≥0
k
2
=µ k
2βe¶
for each k > 0.
These preceding inequalities show that
E (M, N )≤ max
where, for each k ∈ N,
21+ 1
γk =
p Λk−1³2
and for k = 0, since max j=0,1 supt∈R¯¯¯
1
Xk=M
j∈{0,1}°°φ°°W 1,p(ω j )
p Λ+ k Ξ´
µ k
2βe¶
e β(j+i t−θ)2¯¯¯ ≤ e β,
γ0 = e β.
p k k!
N
γk
k
2
e 2β,
For each k ∈ N we have that
1
1
k
2
k
,
·
1
1
Λ
Λ
1
1
1
=
γk+1
γk =
2 pk + 1
p2βe ·µ k + 1
k ¶
·µ1+
k¶
pk + 1
p Λ+ (k + 1)Ξ´
p(k + 1) ·³2
p Λ+ k Ξ´
³2
p Λ+ (k + 1)Ξ´
pp2βe ·³2
p Λ+ k Ξ´
³2
γk+1
γk = 0. This shows that the positive term
from which it is clear that limk→∞
series P∞k=0 γk is convergent and this implies that E (M, N ) indeed has the be-
haviour required to show thatn f
is a Cauchy sequence in F (A0, A1). There-
fore n f
converges in F (A0, A1) norm to an element g ∈ F (A0, A1). In
our setting, where A0 and A1 are the same space with the same norm, each ele-
ment h ∈ F (A0, A1) is a continuous bounded A0 valued function on S which is
analytic in the interior of S. Therefore an obvious extension of the Phragmen-
Lindelöf theorem to Banach space valued analytic functions shows that
NoN∈N
NoN∈N
φ
φ
kh(z)kA0 ≤ sup
ζ∈∂Skh(ζ)kA0 = khkF (A0,A1)
26
MICHAEL CWIKEL AND AMIT EINAV
for each z ∈ S. In particular, this means that, for each fixed z ∈ S, g (z) ∈ A0 and
(4.11)
lim
f
N→∞°°°
φ
N (z)− g (z)°°° A0 = 0.
φ
lows from (4.11) that, for each fixed z ∈ S the sequencen f
Since g (z) is an element of A0 and therefore a measurable function, or rather
an equivalence class of measurable functions on U , it will be convenient to let
g (x, z) denote the value of g (z) at each (or almost every) point x of U . It fol-
converges
to g (z) also in Lp (ω) when ω is either one of the weight functions ω0 and ω1. By
standard results, a norm convergent sequence in any (weighted or unweighted)
Lp space must have a subsequence which converges pointwise at almost every
point of the underlying measure space. So there exists a subset Uz of U pos-
sibly depending on z, and an unbounded subsequence {Nk}k∈N of positive in-
tegers, also possibly depending on z, such that U \ Uz has measure zero and
g (x, z) = limk→∞ f
But, from the formula (4.8) we see that the "whole" sequencenf
verges pointwise for every x ∈ U and that its pointwise limit is
N (x, z)oN∈N
(x, z) for all x ∈ Uz .
N (z)oN∈N
con-
φ
Nk
φ
(z−θ)
p
e
log(rω(x))+β(z−θ)2
φ(x) = f
φ
β,θ(x, z).
This shows that f φ(x, z) = g (x, z) for every x ∈ Uz which means that f φ(z) and
(ω1) for each z ∈ S. Consequently
g (z) are the same element of W
f φ coincides with g and is therefore an element of F (A0, A1). Since A0 is con-
tinuously embedded in W
(ω1)
(ω0)∩ W
1,p
0
1,p
0
it follows that F (A0, A1) is continuously embedded in F³W
and this completes our proof of Lemma 4.11.
(ω0) and A1 is continuously embedded in W
(ω0) ,W
1,p
0
1,p
0
1,p
0
1,p
0
Remark 4.12. Note that the fact f φ(z) and g (z) are the same element of W
W
(ω1) for each z ∈ S provides an alternative proof of the inclusion (4.4).
1,p
0
1,p
0
1,p
0
(ω0),W
4.11, via its proof, also provides the means to obtain a expression which is an
As well as showing that f φ is an element of F³W
upper bound for°° f φ°°F³W
(ω1)´, Lemma
(ω1)´. But that very complicated expression6
is not useful for our needs here. Instead we need the upper bound which will be
provided by the next result:
Proposition 4.13. Let (ω0, ω1) form a (θ, p)−admissible pair for some θ ∈ (0,1)
and p ∈ [1,∞), and let φ be an element of Li pc.
Then the element f
(ω0),W
1,p
0
1,p
0
(ω0),W
1,p
0
1,p
0
(cid:3)
(ω1)´
(ω0)∩
1,p
0
(4.12)
φ
φ
β,θ of F³W
β,θ°°°F³W
(ω0),W
1,p
0
f
°°°
(ω1)´ ≤ 2e β max(1,
1,p
0
(ω1)´ defined by (4.3) satisfies
pp2βe)°°φ°°W p (θ,rω) .
e β
6It depends in quite complicated ways on φ, β, p, ω0 and ω1 including dependence on ex-
pressions involving the set Vφ and the numbers Λ and Ξ which themselves have complicated
dependence on φ, ω0 and ω1.
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
27
Consequently, we also obtain that
(4.13)
1,p
0
(ω0),W
1,p
0
°°φ°°hW
(ω1)iθ ≤ Cp°°φ°°W p (θ,rω) for every φ ∈ Li pc,
for a positive constant Cp depending only on p, which can be chosen to be de-
fined as in (1.14).
φ
Remark 4.14. In the statement of Proposition 4.13 we have used the notation
β,θ instead of its abbreviated variant f φ, and we will sometimes also do this in
f
the following proof of this proposition, in some places where it is necessary to
emphasise the dependence of this element on the parameter β.
Proof of Proposition 4.13. We first notice that
¯¯ f φ(x, z)¯¯ = rω(x)
Rez−θ
p e β((Rez−θ)2−(Imz)2)¯¯φ(x)¯¯ .
Then we recall that, as shown just before (4.6) in the proof of Lemma 4.11, there
exists a set U# (which of course depends on our given function φ ∈ Li pc) which
is almost all of U and consists of all points x for which the pointwise gradients
of log (rω(x)) and of φ(x) both exist. It obviously follows that, at every point x in
this same set U# and for each fixed z ∈ S, the pointwise gradient with respect to
x of f φ(x, z) exists and equals
(z−θ)
p
∇x f φ(x, z) = e
and therefore also
log(rω(x))+β(z−θ)2µ∇x φ(x)+
p e β((Rez−θ)2−(Imz)2)¯¯¯¯
Rez−θ
z − θ
p
φ(x)∇x log (rω(x))¶,
φ(x)∇x log (rω(x))¯¯¯¯
z − θ
p
.
∇xφ(x)+
¯¯∇x f φ(x, z)¯¯ = rω(x)
Since φ ∈ Li pc, Lemma 4.10 ensures that we also have f φ (·, z) ∈ Li pc. Therefore,
by (2) of Theorem 4.8, the weak gradients of each of the functions φ and f φ(·, z)
exist and coincide almost everywhere with their respective pointwise gradients.
We will need to use this fact because the norms appearing in (4.12) are defined
via weak gradients, rather than pointwise gradients.
We will also need to use the fact that, for all x ∈ U and j ∈ {0,1},
ω j (x) = ωθ(x).
rω(x) j−θω j (x) =µ ω0(x)
ω1(x)¶ j−θ
We compute
p
ω j (x)d x
e
rω(x) j−θe
p
Lp(ω j ) =U
pβ³( j−θ)2
°° f φ(·, j + i t )°°
=U
−t 2´¯¯φ(x)¯¯
p
Lp (ωθ) .
Then we also see that the weak gradients of f φ(·, z) and of φ satisfy
z − θ
°°∇x f φ(·, j + i t )°°
p
Lp(ω j ) =U
∇x φ(x)+
pβ³( j−θ)2
−t 2´¯¯φ(x)¯¯
ωθ(x)d x ≤ e pβ°°φ°°
pβ³( j−θ)2
rω(x) j−θe
p
p
−t 2´¯¯¯¯
p
ω j (x)d x
φ(x)∇x log (rω(x))¯¯¯¯
28
MICHAEL CWIKEL AND AMIT EINAV
=U
e
∇x φ(x)+
z − θ
p
which, by (4.5), does not exceed
(4.14)
pβ³( j−θ)2
−t 2´¯¯¯¯
−t 2´
¯¯∇x φ(x)¯¯
p
ωθ(x)d x
φ(x)∇x log (rω(x))¯¯¯¯
+ t 2´
p
2
+³¡ j − θ¢2
p
e
p p
pβ³( j−θ)2
¯¯φ(x)¯¯
2pU
In turn, in order to estimate this expression we use the fact that¯¯j − θ¯¯ ≤ 1, for
one of its terms. For the other term we recall and use the calculations of (4.9)
and (4.10) which together, when k = 1, give us that
≤µ 1
2βe¶
p¯¯∇x log (rω(x))¯¯
β³( j−θ)2
e 2β.
e
1
2
1
2
−t 2´³¡ j − θ¢2
These considerations enable us to deduce that the expression (4.14) is domi-
nated by
+ t 2´
p
ωθ(x)d x.
2p e pβ°°∇x φ°°
p
Lp (ωθ) + 2p
We conclude that
p
1,p
0
1,p
0
(ω0),W
°° f φ°°F³W
(ω1)´ = sup
≤
Lp (ωθ) + 2p e pβ°°∇x φ°°
e pβ°°φ°°
≤
2p e pβ°°φ°°
W 1,p (ωθ) + 2p
p
e 2pβ
p
p
.
1,p
0
¡2βe¢
Lp(ωθ∇ log(rω)p)
2 p p °°φ°°
t∈R³°° f φ(·, i t )°°W
(ω0) ,°° f φ(·,1+ i t )°°W
Lp (ωθ) + 2p
2 p p °°φ°°
¡2βe¢
Lp(ωθ∇ log(rω)p)
2 p p °°φ°°
(ω1)´
Lp(ωθ∇ log(rω)p)
e 2pβ
e 2pβ
1,p
0
1
p
p
p
p
.
p
p
¡2βe¢
1
p
Recalling the notation introduced in (1.10), we see that the preceeding ex-
pression is dominated by
2p e pβ,2p
max
with°°φ°°hW
1,p
0
(ω0),W
This establishes (4.12). Since f
1
p
p
e 2pβ
2 p p
°°φ°°W p (θ,rω) = 2e β max(1,
¡2βe¢
β,θ(x, θ) = φ(x), we conclude that φ ∈hW
β,θ°°°F³W
(ω1)iθ ≤°°°
(ω1)´
(ω0),W
1,p
0
1,p
0
1,p
0
φ
φ
f
e β
pp2βe)°°φ°°W p (θ,rω) .
. Combining this inequality
(ω0) ,W
1,p
0
with (4.12) and taking the infimum over all values of β in the interval (0,∞), (an
infimum which is obviously attained) establishes (4.13) when Cp is the constant
defined in (1.14). This completes the proof of the proposition.
(cid:3)
We are finally ready to present the proof of Theorem 4.5. In fact nearly all of
the tools needed for the proof have already been developed by the preceding
results of this section. It remains only to use an approximation argument to
1,p
0
(ω1)iθ
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
29
extend the result which we already have obtained for functions in Li pc to all
functions in W
p
0 (θ, rω).
Proof of Theorem 4.5. Let φ be an arbitrary function in the space W
view of the definition of this space, there exists a sequence of functions©φnªn∈N ∈
Li pc such that
p
0 (θ, rω). In
0.
and since, in view of (4.13),
°°φ− φn°°W p (θ,rω) −→n→∞
(ω1)iθ
(ω1)iθ ≤ Cp°°φn − φm°°W p (θ,rω) −→n,m→∞
(ω1)iθ
(ω0),W
1,p
0
(ω0),W
1,p
0
1,p
0
1,p
0
1,p
0
0,
. AshW
(ω1)iθ
1,p
0
(ω0),W
1,p
0
such
(ω1)iθ
Since Li pc ⊂hW
1,p
0
(ω0) ,W
1,p
0
°°φn − φm°°hW
1,p
0
(ω0),W
we find that©φnªn∈N is a Cauchy sequence inhW
is a Banach space we conclude that there exists g ∈hW
(ω1)iθ −→n→∞
(ω0),W
that
1,p
0
1,p
0
°°φn − g°°hW
0.
The relevant definitions immediately imply7 (cf. part (iii) of Theorem 2.7) that
(ω0),W
1,p
0
1,p
0
hW
(ω1)iθ ⊂£W 1,p (ω0),W 1,p (ω1)¤θ ,
and
k·k[W 1,p (ω0),W 1,p (ω1)]θ ≤ k·khW
According to Theorem 3.3, g ∈ W 1,p (ωθ), and
(ω1)iθ
°°φn − g°°W 1,p (ωθ) ≤°°φn − g°°[W 1,p (ω0),W 1,p (ω1)]θ
(ω0),W
(ω0),W
1,p
0
1,p
0
.
≤°°φn − g°°hW
1,p
0
1,p
0
(ω1)iθ −→n→∞
0.
p
Therefore ©φnªn∈N converges to g in W 1,p (ωθ). However, since ©φnªn∈N con-
0 (θ, rω) it must also converge to φ in W 1,p (ωθ). As W 1,p (ωθ) is
verges to φ in W
a Banach space due to the (θ, p)−admissibility of the pair (ω0, ω1), we conclude
that g = φ. Thus φ is an element of hW
and©φnªn∈N con-
verges to it in the norm of this space.
To complete this proof it remains only to observe that inequality (4.2) follows
immediately from the facts that
(ω1)iθ
(ω0),W
1,p
0
1,p
0
lim
n→∞°°φn°°hW
1,p
0
(ω0),W
1,p
0
(ω1)iθ =°°φ°°hW
1,p
0
(ω0),W
1,p
0
(ω1)iθ
and
together with (4.13).
lim
n→∞°°φn°°W p (θ,rω) =°°φ°°W p (θ,rω) ,
(cid:3)
7In fact we already took note of and used this simple fact while proving Theorem 3.5.
30
MICHAEL CWIKEL AND AMIT EINAV
5. APPROXIMATION THEOREMS
Theorems 3.3 and 4.5 are the principal ingredients that we require for proving
our main theorem, Theorem 1.16, as well as for most of the proof of Theorem
1.21. As further preparation for these tasks, we require some approximation re-
sults that will allow us to show that the conditions imposed in Theorems 1.16
and 1.21 in fact also suffice to ensure that Li pc¡Rd¢ and even also C∞c ¡Rd¢ are
dense in W 1,p¡Rd , ω¢ and also in W p¡Rd , θ, rω¢.
These approximation results are the focus of this section, with the following as
our main goal:
Theorem 5.1. Let ω0 and ω1 be weight functions that satisfy the compact bound-
edness condition on Rd . Assume in addition that the function rω(x) defined in
(1.9) is locally Lipschitz on Rd . Then, for each p ∈ [1,∞),
(5.1)
and
(5.2)
1,p
0
W 1,p³Rd , ω j´ = W
W p³Rd , θ, rω´ = W
³Rd , ω j´ for j ∈ {0,1}
0 ³Rd , θ, rω´ for each θ ∈ (0,1).
p
The proof of Theorem 5.1 relies on several approximation results, which we
will establish via a series of lemmas. We start by showing that the space W 1,p (U , ω)
has two important properties whenever ω satisfies the compact boundedness
condition.
Lemma 5.2. Let U be an open subset of Rd , let p ∈ [1.∞) and let ω be a weight
function which satisfies the compact boundedness condition on U . Then
(i) W 1,p (U , ω) is a Banach space,
and
(ii) W 1,p (U , ω) is contained in W
1,p
l oc (U ).
In fact these two properties also hold if our hypothesis on the weight function ω is
replaced by the weaker requirement that
(5.3)
inf
x∈K
ω(x) > 0 for every compact subset K of U .
Furthermore, in order to obtain that (i) holds, it suffices to assume that ω only
satisfies an even weaker condition, namely that
(5.4)
1
ω1/p
χK ∈ Lp′ (U ) for every compact subset K of U .
where, as usual, 1/p + 1/p′ = 1.
Remark 5.3. For p = ∞ the space W 1,∞(U , ω) has both of the properties (i) and
(ii) in the statement of Lemma 5.2 for all choices of the weight function ω. This
is simply because, as already pointed out in Remark 1.11, W 1,∞(U , ω) coincides
isometrically with the unweighted space W 1,∞(U ), for which (ii) is a triviality
and (i) is a standard result. (Cf. e.g., Theorem 2 of [6, p. 249].)
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
31
Remark 5.4. In fact, one can also deduce that W 1,p (ω) is a Banach space for
p ∈ (1,∞) from Theorem 1.11 of [12]. For p > 1 the condition (5.4) is just a refor-
mulation of the condition ω ∈ Bp(Ω) introduced in Definition 1.4 on page 538 of
[12] with Ω in the role of our set U . So part (i) of this lemma and our proof of
it are very closely related to the statement and proof of Theorem 1.11 of [12, pp.
540 -- 541]. Here, unlike in that theorem, we also deal with the case where p = 1
and we can observe that the condition (5.4), when p = 1, would be the natural
substitute for the condition ω ∈ Bp(Ω) of [12] and in fact is exactly the same as
(5.3).
Proof of Lemma 5.2. One should keep in mind that we are adopting the frequently
used convention that, for each point a of Cd , a denotes the ℓ2 norm of a. Since
the ℓ2 and ℓp norms are equivalent on Cd we have, for every f ∈ Lp (U , ω, Cd ),
that
(5.5) r°° f°°Lp (U,ω,Cd ) ≤°°¯¯ f¯¯°°Lp (U,ω) =µU¯¯ f (x)¯¯
≤ R°°f°°Lp (U,ω,Cd )
Using Hölder's inequality and then (5.5), for each f ∈ Lp¡U , ω, Cd¢, for each
compact subset K of U and for each g ∈ L∞(U ) which vanishes on U \ K we
obtain that
for some constants r and R depending only on d and p.
ω(x)d x¶1/p
p
χK
χK
χK
f (x)g (x)d x¯¯¯¯
Then, by quite similar reasoning, for K as above, but when the function f is in
U
¯¯¯¯
(5.6)
ω1/p°°°°Lp′ (U)
ω1/p°°°°Lp′ (U)
ω1/p°°°°Lp′ (U)
≤U¯¯ f (x)¯¯¯¯g (x)¯¯ d x ≤°°g°°L∞(U)U¯¯f (x)¯¯ χK (x)d x
≤°°g°°L∞(U)°°¯¯ f¯¯ ω1/p°°Lp (U)°°°°
=°°g°°L∞(U)°°¯¯ f¯¯°°Lp (U,ω)°°°°
≤ R°°g°°L∞(U)°° f°°Lp (U,ω,Cd )°°°°
≤U¯¯ f (x)¯¯¯¯g (x)¯¯ d x ≤
pd°° f°°L∞(U,ω)°°g°°Lp (U,ω)°°°°
After these preparations we turn to showing that (i) holds: Since ω is measur-
L∞¡U , Cd¢ and vanishes on U \K and when g is an arbitrary element of Lp (U , ω),
pd°° f°°L∞(U,ω)U¯¯g (x)¯¯ χK (x)d x
¯¯¯¯
able and strictly positive and since °°φ°°Lp (U,ω) ≤°°φ°°W 1,p (U,ω), it is obvious that
k·kW 1,p (ω) is a norm. Now let©φnªn∈N be an arbitrary Cauchy sequence in W 1,p (U , ω).
This obviously implies that©φnªn∈N is also a Cauchy sequence in Lp (U , ω) and
that©∇φnªn∈N is a Cauchy sequence in Lp (U , ω, Cd ). Since Lp (U , ω) and Lp (U , ω, Cd )
are both Banach spaces, these sequences ©φnªn∈N and©∇φnªn∈N converge re-
spectively, in the norms of these spaces, to an element φ ∈ Lp (U , ω) and an ele-
ment ξ ∈ Lp (U , ω, Cd ). Therefore, in order to show that W 1,p (U , ω) is complete,
f (x)g (x)d x¯¯¯¯
.
χK
ω1/p°°°°Lp′ (U)
≤
we have that
U
(5.7)
32
MICHAEL CWIKEL AND AMIT EINAV
it remains only to show that ξ is the weak gradient of φ on U .
Let ψ be an arbitrary function in C∞c (U ) and let Kψ denote a compact subset of
U on which ψ is supported. We claim that
(5.8)
and also that
(5.9)
lim
n→∞U
ψ(x)¡∇φn(x)− ξ(x)¢ d x = 0.
n→∞U ∇ψ(x)¡φ(x)− φn(x)¢ d x = 0
lim
χKψ
ω1/p°°°Lp′ (U)
We shall obtain these two formulae with the help of the inequalities (5.6) and
(5.7). We shall choose K in both of these inequalities to be Kψ. If we require
ω to satisfy the compact boundedness condition or if we merely make do with
requiring one of the weaker variants (5.3) and (5.4) of this condition, we will ob-
is finite. We use this fact and substitute f (x) =
this gives us (5.8). Then we substitute f (x) = ∇ψ(x) and g (x) = φ(x)− φn(x) in
Using (5.8) and then (5.9), we see that
tain that the quantity°°°
∇φn(x)− ξ(x) and g (x) = ψ(x) in (5.6). Since limn→∞°°∇φn − ξ°°Lp(U,ω,Cd ) = 0,
(5.7). Since limn→∞°°φ− φn°°Lp (U,ω) = 0, this gives us (5.9).
n→∞U ∇ψ(x)φn(x)d x
U ∇ψ(x)φ(x)d x = lim
n→∞U
ψ(x)∇φn(x)d x
= − lim
= −U
ψ(x)ξ(x)d x
for each ψ ∈ C∞c (U ). This shows that ξ is indeed the weak gradient of φ on U and
completes the proof of part (i) of the lemma, also when ω only satisfies (5.3) or
(5.4).
We now turn our attention to part (ii). Here our hypothesis on ω is either that it
satisfies the compact boundedness condition, or merely one "half" of that con-
dition, namely (5.3).
For each compact subset K of U and each φ ∈ W 1,p (U , ω) we have, recalling the
notation introduced in (1.7), that
Lp(U,ω,Cd )´
°°χK φ°°
Lp (U) +°°χK ∇φ°°
Lp(U,Cd ) ≤
1
1
p
p
p
m(K , ω)p ³°°χK φ°°
m(K , ω)p ³°°φ°°
m(K , ω)p °°φ°°
1
≤
=
p
p
Lp (U,ω) +°°χK∇φ°°
Lp (U,ω) +°°∇φ°°
Lp(U,ω,Cd )´
p
p
W 1,p (U,ω) .
This establishes the inclusion claimed in part (ii), and therefore completes the
proof of the lemma.
(cid:3)
Remark 5.5. In the formulations of several of the following results of this section
we will find it convenient to require the relevant weight function to satisfy the
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
33
compact boundedness condition, even though, just as for the preceding lemma,
some of these results hold under weaker hypotheses.
Lemma 5.6. Let ω be a weight function that satisfies the compact boundedness
condition on an open set U . Then Li pc (U ) ⊂ W 1,p (U , ω) for each p ∈ [1,∞].
Proof. Recalling Rademacher's theorem, Theorem 4.8, we know that if φ ∈ Li pc (U )
then it is in W 1,∞l oc (U ), and therefore has a weak gradient ∇φ on U . Obviously (cf.
Remark 4.9) ∇φ ∈ L∞(U , Cd ) and°°∇φ°°L∞(U,Cd ) is bounded by the Lipschitz con-
stant of φ. Also φ, like every other compactly supported Lipschitz function, is
bounded. This completes the proof in the case where p = ∞.
To deal with the case p ∈ [1,∞), we note that φ and ∇φ both vanish almost ev-
erywhere on U \suppφ, and that the Lebesgue measure of suppφ is finite. Then,
using the notation of (1.7), and once more the boundedness of φ we conclude
that
p
p
p
p
and
U¯¯φ(x)¯¯
U¯¯∇φ(x)¯¯
ω(x)d x ≤°°φ°°
ω(x)d x ≤°°∇φ°°
L∞(U)¯¯suppφ¯¯ M¡suppφ, ω¢ < ∞8
L∞(U,Cd )¯¯suppφ¯¯ M¡suppφ, ω¢ < ∞.
This shows that φ ∈ W 1,p (U , ω) also for p ∈ [1,∞), completing the proof.
Lemma 5.7. Let ω be a weight function that satisfies the compact boundedness
condition on an open set U . Then, given p ∈ [1,∞), and functions φ ∈ W 1,p (U , ω)
and ξ : U → C such that ξ is Lipschitz on U , we have that their pointwise product
φξ has a weak gradient which can be expressed in terms of the weak gradients of
φ and ξ by the formula
(cid:3)
∇¡φξ¢(x) = ξ(x)∇φ(x)+ φ(x)∇ξ(x)
for almost every x ∈ U .
1,p
Proof. Much as in the proof of the previous lemma, we know, due to Rademacher's
Theorem and Remark 4.9, that ξ ∈ W 1,∞l oc (U ) (and in fact ξ ∈ W 1,∞ (U )), and con-
sequently ξ ∈ W 1,r
l oc (U ) for every r ≥ 1. We also note that, due to Lemma 5.2,
l oc (U ). Let ψ be an arbitrary function in C∞c (U ) and let Kψ = suppψ. Stan-
φ ∈ W
dard approximation theorems which are used in the theory of Sobolev spaces
(see for example Theorem 1 of [6, pp. 250-251] and its proof, which is based on
Theorem 6 of [6, pp. 630-631]) enable us to find an open set V containing Kψ,
whose closure is compact and contained in U , and functions ξn : U → C whose
restrictions to V are in C∞ (V ) and which satisfy
(5.10)
and
(5.11)
lim
n→∞V ξn(x)− ξ(x)q d x = 0
lim
n→∞V ∇ξn(x)−∇ξ(x)q d x = 0.
8Here (and also elsewhere) we use the standard notation E for the d−dimensional Lebesgue
measure of any measurable subset E of Rd .
34
MICHAEL CWIKEL AND AMIT EINAV
where q is the Hölder conjugate of p. Since ψ(x)ξn(x) ∈ C∞c (U ) for each n ∈ N
we see that
U
φ(x)ξn(x)∇ψ(x)d x =U
φ(x)∇¡ψ(x)ξn(x)¢ d x −U
φ(x)ψ(x)∇ξn(x)d x
= −U
ψ(x)¡ξn(x)∇φ(x)+ φ(x)∇ξn(x)¢ d x.
1,p
We recall that ψ and ∇ψ both vanish on U \ V . Therefore, in all the integrals
in the preceding calculation, we can replace the domain of integration U by V .
l oc (U ) (due to (ii) of Lemma 5.2), we also observe that φ∇ψ and ψ∇φ
Since φ ∈ W
are both elements of Lp (U , Cd ) and ψφ ∈ Lp (U ). So we can use (5.10), (5.11) and
Hölder's inequality to pass to the limit in the preceding calculation and thus to
obtain that
U
φ(x)ξ(x)∇ψ(x)d x = −U
ψ(x)¡ξ(x)∇φ(x)+ φ(x)∇ξ(x)¢ d x.
Since this holds for every ψ ∈ C∞c (U ), the proof is complete.
(cid:3)
From this point onward, till the end of this section, we will only deal with the
case where U = Rd , and, accordingly, we shall usually remove any mention of
the underlying set (namely Rd ) from our notation.
Note that in the following lemma we will be permitting a function ρ which seems
to be playing the role of a weight function to possibly also take the value 0. So,
analogously to the conventions adopted in Definition 1.19 and in (1.15), it may
happen that Lp (ρ) and k·kLp (ρ) have to be understood to be, respectively, a semi-
normed space and the semi-norm on that space.
Lemma 5.8. Let ρ : Rd → [0,∞) and φ : Rd → C be arbitrary measurable functions
which satisfy
for some p ∈ [1,∞). For each n ∈ N define the function ξn : Rd → [0,∞) by
p
Rd¯¯φ(x)¯¯
ξn(x) =
1
2− xn
0
ρ(x)d x < ∞,
x < n
n < x < 2n
x ≥ 2n
(5.12)
Then,
the d -dimensional ℓp norm of its weak gradient satisfies
(i) ξn ∈ Li pc¡Rd¢ with Lipschitz constant 1
k∇ξn(x)kℓp ≤
(5.13)
1
p
d
n
.
n , and at almost every point x ∈ Rd
(ii) φξn ∈ Lp¡ρ¢ and
(5.14)
Rd¯¯φ(x)ξn(x)− φ(x)¯¯
p
ρ(x)d x −→n→∞
0.
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
35
Proof. For part (i) the fact that ξn is in Li pc¡Rd¢ with Lipschitz constant 1
n is
immediate. It is also obvious that, at each point x ∈ Rd such that x 6= n and
x 6= 2n, the pointwise gradient ∇ξn(x) exists and equals either 0 or − x
nx
This gives us (5.13) since every x ∈ Rd satisfies
.
1
kxkℓp ≤ d
p kxkℓ∞ .
We also have
For part (ii) we note that ¯¯φ(x)ξn(x)¯¯ ≤¯¯φ(x)¯¯ which ensures that φξn ∈ Lp(ρ).
ρ ∈ L1 and since φ(x)ξn(x)
for all n ∈ N and for almost every x ∈ Rd . Since¯¯φ¯¯
converges pointwise to φ(x) we obtain (5.14) as a consequence of the Domi-
nated Convergence Theorem.
(cid:3)
1− ξn(x)p ≤¯¯φ(x)¯¯
¯¯φ(x)ξn(x)− φ(x)¯¯
=¯¯φ(x)¯¯
p
p
p
p
Lemma 5.9. Let ν be a weight function that satisfies the compact boundedness
condition and let ξn be as in (5.12). Then, for each p ∈ [1,∞) and for every φ ∈
W 1,p (ν), we have that φξn ∈ W 1,p (ν) and
(5.15)
0.
Proof. Our hypothesis on ν enables us to use Lemma 5.7 and so we see that φξn
has a weak gradient on Rd , which is given by
°°φ− φξn°°W 1,p (ν) −→n→∞
ξn(x)∇φ(x)+ φ(x)∇ξn(x).
Since ξn and ∇ξn are bounded, for each fixed n, and since φ ∈ W 1,p (ν) we see
that φξn ∈ Lp (ν) and ∇¡φξn¢ ∈ Lp¡ν, Cd¢. This implies that φξn ∈ W 1,p (ν). More-
over, with the help of (5.13), we have that
°°∇φ−∇¡φξn¢°°Lp(ν,Cd) ≤°°∇φ− ξn∇φ°°Lp(ν,Cd) +°°φ∇ξn°°Lp(ν,Cd )
d
1
p
1
p
≤µRd 1− ξn(x)p¯¯∇φ(x)¯¯
p
ℓp ν(x)d x¶
+
n °°φ°°Lp (ν) .
So, as in the proof of part (ii) of Lemma 5.8, the Dominated Convergence Theo-
rem obviously provides what is needed to show that
And of course Lemma 5.8 also gives us that
lim
n→∞°°∇φ−∇¡φξn¢°°Lp(ν,Cd) = 0.
These two conditions show that (5.15) holds, and thus complete the proof of the
present lemma.
(cid:3)
lim
n→∞°°φ− φξn°°Lp (ν) = 0.
A very useful feature of Lemmas 5.8 and 5.9 is the fact that the approximating
sequence in the relevant semi-normed weighted Lp space or weighted Sobolev
space is always constructed in the same way, regardless of which p or weight
function we use! This fact immediately proves the following obvious extension
of these two lemmas which will be essential for us when we seek to approxi-
mate functions in the normed space W p (θ, rω), precisely because its norm is the
36
MICHAEL CWIKEL AND AMIT EINAV
sum of the (semi-)norms of two differently weighted spaces. As in Lemma 5.8
and Definition 1.19, here again we will encounter non-negative functions acting
essentially as weights. This time there will be a finite collection of possibly dif-
ferent non-negative functions ρβ and semi-normed spaces Lqβ(ρβ) associated
with them.
Lemma 5.10. Let ν1,..., ν j be weight functions that satisfy the compact bound-
edness condition, and let ρ1,... , ρk be arbitrary non-negative measurable func-
tions. Given j + k exponents p1,..., p j , q1,...., qk in [1,∞), we let X be the linear
space X =³T j
β=1 Lqβ¡ρβ¢´ equipped with the norm
α=1 W 1,pα (να)´∩³Tk
k
j
°°φ°°X :=
Xα=1°°φ°°W 1,pα (να) +
Xβ=1°°φ°°L
qβ(ρβ) .
For each n ∈ N, let ξn be the function defined by (5.12). Then, for every φ ∈ X , the
function φξn is also in X and
lim
n→∞°°φ− φξn°°X = 0.
Now that we know how to approximate functions of interest to us by com-
pactly supported functions, our next step will be to approximate those approxi-
mating functions in turn by smoother functions, not merely Lipschitz functions
lowing standard result:
but even elements of C∞c ¡Rd¢. An essential ingredient for doing this is the fol-
Lemma 5.11. Let η ∈ C∞c ¡Rd¢ be a given non-negative function which is sup-
ported in B1(0) and satisfies ´Rd η(x)d x = 1. For p ∈ [1,∞) let φ be a given func-
tion in Lp¡Rd¢ that is compactly supported. Define:
η¡n(x − y)¢ φ(y)d y = ηn ∗ φ(x),
φn(x) := ndRd
where ηn(x) = nd η(nx). Then
(i) φn ∈ C∞c ¡Rd¢ and is supported in Dn =©x ∈ Rd dist¡x, suppφ¢ ≤ 1
nª.
(ii) φn converges to φ in Lp¡Rd¢.
(iii) If in addition φ ∈ W 1,p¡Rd¢ we have that φn ∈ W 1,p¡Rd¢ and
lim
n→∞°°φ− φn°°W 1,p(Rd ) = 0.
We refer to standard textbooks for the standard proof 9 of this result.
We will now obtain a variant of Lemma 5.10 as our next step towards proving
our principal result in this section, Theorem 5.1. This variant will be stated in
much greater generality than we need for our immediate purposes here - but
that comes at no extra cost in the proof. That greater generality may well turn
out to be useful in future extensions of this research.
9For example, it is a straightforward variant of the proof of Theorem 6 of [6, p. 630]. See rele-
vant material on pp. 629-631 of [6] and also, for example, on pp. 148-149 of [23].
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
37
Theorem 5.12. Let X be the normed space introduced in Lemma 5.10 and sup-
pose that each of the weight functions ν1,..., ν j appearing in its definition sat-
isfies the compact boundedness condition. Suppose also that each of the non-
negative measurable functions ρ1,..., ρk appearing in its definition is bounded
from above on every compact set and that
(5.16)
max©q1,...., qkª ≤ max©p1,..., p jª .
Then C∞c (Rd ) is a dense subspace of X .
Proof. The inclusion C∞c (Rd ) ⊂ X follows immediately from the fact that all the
functions ν1,..., ν j and ρ1,..., ρk are bounded above on every compact set.
Let φ be an arbitrary compactly supported element of X and let φn = ηn ∗ φ for
each n ∈ N, where the functions ηn are as specified in Lemma 5.11. Then, by
Lemma 5.11, all of the functions φn are in C∞c (Rd ) and they and their gradients
vanish on the complement of the compact set D1 =©x ∈ Rd : dist(x,suppφ) ≤ 2ª.
Furthermore, our hypotheses ensure that the supremum
x∈D1½
M := sup
max
α∈{1,..., j },β∈{1,...,k}©να(x), ρβ(x)ª¾
is finite. This enables us to use the inequalities
(5.17)
and
(5.18)
1
qβ (ρβ) ≤ M
°°φ− φn°°L
°°φ− φn°°W 1,pα (να) ≤ M
qβ (Rd )
qβ °°φ− φn°°L
pα °°φ− φn°°W 1,pα (Rd )
1
which hold for all n ∈ N, α ∈ {1,..., j } and β ∈ {1,..., k}.
The fact that infx∈D1 να(x) > 0 ensures that φ ∈ W 1,pα¡Rd¢ and therefore also φ ∈
Lpα(Rd ) for each α ∈©1,..., jª. Consequently, in view of (5.16), Hölder's inequality
and the fact that the support of φ is compact, we also have that φ ∈ Lqβ(Rd ) for
each β ∈ {1,.., k}. The membership of φ in all of these unweighted Lp spaces
permits us to use Lemma 5.11 to obtain that the right sides of (5.17) and (5.18)
tend to 0 as n tends to ∞ for all α ∈ {1,..., j } and β ∈ {1,..., k}. Consequently we
also have that
lim
n→∞°°φ− φn°°X = 0.
This shows that every compactly supported element in X can be approximated
in X norm by a sequence of functions in C∞c ¡Rd¢. Since we already know from
Lemma 5.10 that every element in X can be approximated in X norm by a se-
quence of compactly supported functions in X , this completes the proof of the
theorem.
(cid:3)
We are now ready to prove Theorem 5.1:
Proof of Theorem 5.1. We shall in fact prove a stronger result, namely that C∞c (Rd )
is dense in each of the spaces W 1,p¡Rd , ω0¢, W 1,p¡Rd , ω1¢ and W p¡Rd , θ, rω¢.
38
MICHAEL CWIKEL AND AMIT EINAV
This of course will imply (5.1) and (5.2), since C∞c ¡Rd¢ ⊂ Li pc¡Rd¢. We begin by
applying Theorem 5.12 in the special case where
j = k = 1, p1 = q1 = p, ν1 = ωm, ρ1 = 0
and where m = 0 or m = 1. In this case the theorem shows that C∞c ¡Rd¢ is dense
in X = W 1,p¡Rd , ωm¢. This establishes (5.1).
It remains to prove (5.2). We start by recalling (cf. Remark 1.5) that if ω0 and ω1
satisfy the compact boundedness condition, then obviously so do ωθ = ω1−θ
0 ωθ
1
and rω. As rω is locally Lipschitz, we use Rademacher's theorem to conclude that
it is differentiable almost everywhere with essentially bounded gradient on each
compact set. Consequently, due to the strict positivity of rω, the function log rω
is also differentiable almost everywhere and its gradient, given by
∇log (rω(x)) = ∇rω(x)
rω(x)
is also essentially bounded on each compact set.
We summarise the above:
,
• ωθ satisfies the compact boundedness condition.
• ωθ¯¯∇log (rω)¯¯
ρ1 = ωθ¯¯∇log rω¯¯
p is essentially bounded from above on every compact set.
These two properties are exactly what we need to enable us to apply Theorem
5.12 once more, where this time we choose j = k = 1, p1 = q1 = p, ν1 = ωθ and
p . These choices make X coincide with W p (θ, ωθ) to within
equivalence of norms, and so the density of C∞c (Rd ) in X , which is again guar-
anteed by Theorem 5.12, establishes (5.2) and therefore completes the proof of
the theorem.
(cid:3)
In all of our results in this section from Lemma 5.8 onwards we have been
dealing only with the case where the underlying open set U is all of Rd . But
let us conclude this section with the following remark about the possibility of
extending our approximation theorems to the case where U is a proper open
subset of Rd :
Remark 5.13. As the reader may have noticed from the various steps leading to
the proof of Theorem 5.12, the two main ingredients which we used to show the
theorem were the fact that we can approximate a function in W 1,p¡Rd , ω¢ by a
compactly supported function and Lemma 5.11. As a more general version of
Lemma 5.11 can be shown to be valid for any open set U ⊂ Rd instead of Rd (see
e.g. [6] pp. 629-631), we see that if the open set U and the weight function ω
are such that any φ ∈ W 1,p (U , ω) can be approximated by compactly supported
functions, then we will have that
W
1,p
0
(U , ω) = W 1,p (U , ω) ,
and also that C∞c (U ) is dense in W 1,p (U , ω) and in W p (U , θ, rω).
This raises the question of what properties of U and ω might suffice to ensure
that compactly supported functions are dense in W 1,p (U , ω). It would seem that
this density property will hold if ω(x) is required to tend to 0 sufficiently rapidly
p
W
0 (U , θ, rω) = W p (U , θ, rω),
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
39
when x approaches the boundary of U . Thus it might be natural to consider
this question for weight functions which depend solely on the distance to the
boundary, such as those which appear in the weighted Sobolev spaces discussed
on pages 17-20 of [11].
6. PROOFS OF THE MAIN THEOREM, AND OF ITS MAIN CONSEQUENCES
In this short section we pool all the resources that we have developed in Sec-
tions §3, §4 and §5 to prove our main theorem, Theorem 1.16 and its main con-
sequences: Theorems 1.20, 1.21 and 1.22.
Proof of Theorem 1.16. Using Theorem 3.3 we obtain the inclusion (1.11) and
one side of the inequality (1.13). To complete the proof it remains to establish
the inclusion (1.12) and the other side of (1.13). We start doing this by observ-
ing that, since ω0, ω1, and therefore obviously also ωθ, each satisfy the compact
boundedness condition (cf. Remark 1.5), we can apply Lemma 5.2 and Lemma
5.6 to obtain that
(1) For every p ∈ [1,∞), each of the spaces W 1,p (ω0), W 1,p (ω1) and also
W 1,p (ωθ), for every θ ∈ (0,1), is a Banach space that contains Li pc.
One of the hypotheses of Theorem 1.16 is that rω is locally Lipschitz on U . The
strict positivity of rω therefore implies that
(2) The function log rω is locally Lipschitz on U .
The above two conditions (1) and (2) tell us (cf. Definition 4.1) that the pair
(ω0, ω1) is¡θ, p¢−admissible for every θ ∈ (0,1) and p ∈ [1,∞). This permits us to
invoke Theorem 4.5, which applies to any open subset U of Rd and whose con-
clusions (4.1) and (4.2) are exactly the above mentioned inclusion and inequality
which are required to complete the proof of Theorem 1.16.
(cid:3)
Proof of Theorem 1.20. The first of the two inclusions in (1.16) has already been
established as part of the proof of Theorem 1.16. The additional condition im-
posed in Theorem 1.20 for some p ∈ [1,∞) is exactly what is required to enable
Theorem 3.5 to be applied in order to establish the second inclusion in (1.16) for
that value of p and for all θ ∈ (0,1).
(cid:3)
Proof of Theorem 1.21. The proof of Theorem 1.16 shows that the inclusions (1.11)
and (1.12) hold also here. So, to prove the first part of Theorem 1.21 we simply
substitute (1.17) and (1.18) in (1.11) and (1.12) to obtain (1.19). To prove the sec-
ond part of the theorem, in which it is assumed that U = Rd , we simply apply
Theorem 5.1 to give us that (1.17) and (1.18) both hold for all p ∈ [1,∞).
Proof of Theorem 1.22. Since g (x) is continuous and the exponential function
is continuous and strictly positive, we conclude that if ω1 (or ω0) satisfies the
compact boundedness condition, then so does ω0 (or ω1). Next, we note that
(cid:3)
which implies, as g (x) is Lipschitz and the exponential function is C 1 on R, that
rω(x) is locally Lipschitz. Thus we have shown that the conditions of Theorem
rω(x) = e g (x),
40
MICHAEL CWIKEL AND AMIT EINAV
1.16 are satisfied, as claimed in part (i) of Theorem 1.22. We next show part (ii),
namely that (1.20) holds to within equivalence of norms when U = Rd . Theorem
1.16 (or even simply Theorem 3.3) gives us "half" of (1.20), i.e. the inclusion "⊂"
and the corresponding norm inequality. So we only need to show that
W 1,p³Rd , ωθ´ ⊂hW 1,p³Rd , ω0´ ,W 1,p³Rd , ω1´iθ
and to obtain the reverse norm inequality corresponding to this reverse inclu-
sion. In view of (1.12) and of Theorem 1.16 and Theorem 5.1 we can in turn
reduce this task to showing that the inclusion
W 1,p³Rd , ωθ´ ⊂ W p³Rd , θ, rω´
holds and is continuous.
For any given element φ ∈ W 1,p (ωθ) we see that
(6.1)
(6.2)
This implies that φ is automatically in W p (θ, rω) and
p
p
°°φ°°
Lp(ωθ∇ log(rω)p) =U¯¯φ(x)¯¯
L∞°°φ°°
W 1,p (ωθ) +°°φ°°
L∞¢
≤°°∇g°°
°°φ°°W p (θ,rω) =³°°φ°°
≤¡1+°°∇g°°
p
1
p
p
ωθ(x)d x
p¯¯∇g (x)¯¯
p
Lp (ωθ) .
1
p
p
Lp(ωθ∇ log(rω)p)´
p °°φ°°W 1,p (ωθ) .
This establishes (6.1) and the required associated norm inequality and thus takes
care of part (ii). (In fact (6.2) also shows that the three norms appearing in (1.13)
are equivalent to each other.) Finally we establish the remaining parts (iii) and
(iv) of Theorem 1.22 by using exactly the same considerations, together with
Theorems 1.20 and 1.21.
(cid:3)
In the last two sections of the paper we will discuss additional topics related
to our study, including an example of when W p (θ, rω) is not W 1,p (ωθ), and some
potential applications and open problems.
7. MORE ABOUT THE SPACE W p (θ, rω)
In this section we will discuss two more properties of the space W p (θ, rω).
Namely, we will answer the following two questions:
Question 3. Is it possible for W p (θ, rω) to not be W 1,p (ωθ)?
Question 4. Can we find a smaller, yet "big enough", subspace of W p (θ, rω) that
will allow us to only consider the weight ωθ?
Our answers will give us more information about the space£W 1,p (U , ω0),W 1,p (U , ω1)¤θ.
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
41
7.1. Can W p (θ, rω) be smaller than W 1,p (ωθ)? We consider it very unlikely that
the Stein-Weiss like formula (1.4) could hold for all choices of the set U and
weight functions ω0 and ω1. Here we shall begin some attempts towards finding
a counterexample.
One might expect that when U has a very complicated structure or when some
or all of the hypotheses of Theorem 1.16 do not hold, this could prevent (1.4)
from holding. But the examples which we are about to present hint that maybe
(1.4) can fail even when U is simply all of Rd and the above mentioned hypothe-
ses do hold.
From this point onwards we will take U to be Rd and remove almost all mention
of the underlying set from our notation.
As can be seen from (1.19) in Theorem 1.21, a sufficient condition for obtaining
a Stein-Weiss like theorem when U = Rd is that
W p (θ, rω) = W 1,p (ωθ)
and indeed (since obviously W p (θ, rω) ⊂ W 1,p (ωθ)) this condition is effectively
what we have used in the proof of Theorem 1.22. (Cf. (6.1).) So a natural first
step in searching for the above mentioned counterexample is to look for pairs of
weight functions for which
(7.1)
W p (θ, rω) ⊂6=
W 1,p (ωθ) .
We shall describe such pairs. In fact, for each choice of θ ∈ (0,1), we will find
such a pair which satisfies the hypotheses of Theorem 1.16. But we should em-
phasize that at this stage we do not know whether these weight functions, which
satisfy (7.1), will also turn out to provide a counterexample to (1.4). To simplify
the discussion we shall make do with presenting explicit formulas for our weight
functions only in the case p = 1. However it is not difficult to modify our exam-
ples so that they apply when p ∈ (1,∞).
Let us start by stating the relevant properties that a function, say φ, which is in
W 1,p (ωθ) but not in W p (θ, rω) must have: Since φ ∈ W 1,p (ωθ) we have that φ
has a weak gradient on Rd and
(7.2)
On the other hand, since φ 6∈ W p (θ, rω) we must have that
(7.3)
p
Rd¯¯φ(x)¯¯
p
ωθ(x)d x < ∞, Rd¯¯∇φ(x)¯¯
Rd¯¯φ(x)¯¯
p¯¯∇log (rω(x))¯¯
ωθ(x)d x = ∞.
p
ωθ(x)d x < ∞.
We shall simply choose φ ≡ 1. Then the gradient condition in (7.2) is automat-
ically satisfied, and conditions (7.2) and (7.3) become the following two condi-
tions on ω0 and ω1:
(7.4)
ωθ(x)d x = ∞.
Let us simplify things further by also choosing ω0 ≡ 1. In this case
∇logµ ω0(x)
ω1(x)¶¯¯¯¯
ωθ ∈ L1³Rd´,
Rd¯¯¯¯
p
ωθ(x) = ω1(x)θ,
42
MICHAEL CWIKEL AND AMIT EINAV
and we can see that showing that there is a weight function ωθ which satisfies
(7.4) in this case is equivalent to finding a weight function ω (which will play the
role of ωθ) such that:
(7.5)
Rd
ω(x)d x < ∞,
Rd ∇ω(x)p ω(x)1−p d x = ∞.
It is clear that finding such a function ω will in fact show that (7.4) can be sat-
isfied for every choice of θ ∈ (0,1) and for the particular value of p appearing in
both (7.5) and (7.4).
At this stage we shall specialise to the case p = 1. In this case (7.5) is simply the
statement that ω is a function in L1¡Rd¢ whose gradient is not in L1¡Rd , Cd¢. For
various obvious examples of integrable functions ω which have some monotone
decay at infinity (for example, like the reciprocal of a polynomial, or of an expo-
nential) it turns out that (7.5) does not hold. However, as we shall now see, if
we modify such a function by adding a term which oscillates rapidly near in-
finity and is suitably bounded, then the resulting function ω does satisfy (7.5):
Consider the strictly positive C 1 function
(7.6)
=
ω(x) =
1
1
¡1+x2¢αµ1+ sin³xβ´+
¡1+x2¢αµ1+
1+x2¶+
1+x2¶
sin¡xβ¢
¡1+x2¢α .
1
1
We shall now show that one can easily choose positive numbers α and β for
which ω will satisfy (7.5) for p = 1. In fact, a similar computation shows that for
any p ∈ (1,∞), one can find an α and a β for which (7.6) will satisfy (7.5).
Clearly, if α > d
2 then ω(x) ∈ L1¡Rd¢ since
ω(x) ≤
3
Next, as
and
and
∇ω(x) =
2x
¡1+x2¢α+2
x
1
−
−2α
¡1+x2¢α .
1+x2¶ x −
βxβ−2 cos¡xβ¢
¡1+x2¢α
1+x2¶ x¯¯¯¯¯
≤
¡1+x2¢α+1
¡1+x2¢α ,
≤
¡1+x2¢α+1µ1+
2αsin¡xβ¢
¡1+x2¢α+1 x +
¡1+x2¢α+1µ1+
¯¯¯¯¯
¡1+x2¢α+2¯¯¯¯¯
¯¯¯¯¯
¡1+x2¢α+1 x¯¯¯¯¯
sin¡xβ¢
≤
2x
1
1
1
1
¯¯¯¯¯
1
¡1+x2¢α ,
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
43
We now compute
we see that if α > d
2 then ∇ω ∈ L1¡Rd , Cd¢ if and only if
x¯¯¯¯¯
Iα,β :=Rd¯¯¯¯¯
xβ−2 cos¡xβ¢
¡1+x2¢α
xβ−1¯¯cos¡xβ¢¯¯
Iα,β =Rd
¡1+x2¢α
r β+d−2¯¯cos¡r β¢¯¯
Sd−1¯¯
d r ≥¯¯
2α ∞
¡1+ r 2¢α
Sd−1¯¯
=y=r β ¯¯
2αβ ∞
If, for example, we choose α = d and β = d + 1, then this last integral is infinite.
Since we then also have α > d
2 as required in the previous part of this discussion,
this completes our proof that (7.1) can hold for p = 1 and for every θ ∈ (0,1) for
suitable choices of weight functions.
r β+d−2−2α¯¯¯
¯¯cos¡y¢¯¯ d r
cos³r β´¯¯¯
=¯¯¯
Sd−1¯¯¯
d−1−2α
β
y
1
∞
0
d x < ∞.
d x
1
d r
Remark 7.1. As we already mentioned, there exist other examples of pairs of
weight functions which satisfy (7.1) for p ∈ (1,∞). Some of them nevertheless
also satisfy (1.4). We refer the reader to the appendix where we give an example
of a pair of weight functions (even in C∞¡Rd¢) which in fact satisfies both (7.1)
and (1.4) for every p ∈ [1,∞) and θ ∈ (0,1).
7.2. A simpler subspace of£W 1,p (U , ω0),W 1,p (U , ω1)¤θ. The main result of this
paper shows that, under relatively mild conditions, the interpolation space
£W 1,p (U , ω0),W 1,p (U , ω1)¤θ
lies between W 1,p (U , ωθ) and a smaller subspace W p (U , θ, rω), which can be
expressed as W 1,p (U , ωθ)∩ X (with an appropriate norm) where X is the semi-
normed weighted space Lp¡U , ωθ¯¯∇log rω¯¯
For some applications, such as those which we will shortly discuss in 8.2, it will
be convenient to find a more simply defined space X , which is embedded con-
tinuously in X . Such a space will therefore yield the chain of inclusions
p¢ (see Definition 1.19).
W 1,p (U , ωθ)∩ X ⊂£W 1,p (U , ω0),W 1,p (U , ω1)¤θ ⊂ W 1,p (U , ωθ).
A simple X to choose, if possible, would be a standard weighted Lebesgue space
with respect to the weight function ωθ alone. We shall now discuss one possible
way which can enable us to choose an X of this kind, however with an expo-
nent larger than p. It requires us to impose one more condition on our weight
functions.
Definition 7.2. Let ω0 and ω1 be weight functions on an open subset U of Rd and
let ωθ and rω be as defined throughout this paper (by (1.5) and (1.9)). Suppose
44
MICHAEL CWIKEL AND AMIT EINAV
(7.7)
that log rω has a weak gradient on U . For each q ∈ [1,∞) and each θ ∈ [0,1] let
M¡U , θ, rω, q¢ be the quantity
1
q
M¡U , θ, rω, q¢ :=µU¯¯∇log (rω(x))¯¯
q
ωθ(x)d x¶
.
(Note that here θ may also equal 0 or 1.) We will often use the abbreviated nota-
tion M¡θ, q¢ for M¡U , θ, rω, q¢.
Lemma 7.3. Let U , ω0, ω1, ωθ and rω have the properties stated in Definition
7.2. Suppose, furthermore, that there exist q ∈ (p,∞) and θ ∈ (0,1) for which
M¡θ, q¢ < ∞. Then, in terms of the notation of Definition 1.19, we have that
qp
and
(7.8)
L
q−p (U , ωθ) ⊂ Lp¡U , ωθ¯¯∇log (rω)¯¯
k·kLp(U,ωθ∇ log(rω)p) ≤ M¡θ, q¢k·kL
p¢
qp
q−p (U,ωθ)
for these values of q and θ.
Proof. Using Hölder's inequality with the exponent q
p we find that
U¯¯φ(x)¯¯
≤µU¯¯∇log (rω(x))¯¯
q
p¯¯∇log (rω(x))¯¯
ωθ(x)d x¶
p
q µU¯¯φ(x)¯¯
p
ωθ(x)d x
This proves both the inclusion and (7.8).
pq
q−p ωθ(x)d x¶
q−p
q
.
(cid:3)
With this in hand we can deduce the following:
Theorem 7.4. Let ω0 and ω1 be a pair of weight functions on U that satisfy the
compact boundedness condition. Assume furthermore that
is locally Lipschitz on U , and that U , ω0 and ω1 are such that the conditions of
rω(x) =
ω0(x)
ω1(x)
Theorem 1.21 hold. Then, if M¡θ, q¢, defined by (7.7), is finite for some q > p, we
have that
W 1,p (U , ωθ)∩ L
qp
q−p (U , ωθ) ⊂£W 1,p (U , ω0),W 1,p (U , ω1)¤θ .
Moreover, there exists a constant Cp > 0 which depends only on p, such that for
any φ ∈ W 1,p (U , ωθ)∩ L
q−p (U , ωθ),
qp
(7.9)
°°φ°°[W 1,p (U,ω0),W 1,p (U,ω1)]θ
p
W 1,p (U,ωθ) + M¡θ, q¢p°°φ°°
p
L
≤ Cpµ°°φ°°
1
p
q−p (U,ωθ)¶
qp
Proof. This theorem is an immediate consequence of Theorem 1.21 and Lemma
7.3. The inequality (7.9) follows immediately from (1.13) of Theorem 1.16 and
(7.8).
(cid:3)
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
45
Obviously, a modification of the above theorem can be made to accommo-
date the more general case of Theorem 1.16 when we start considering appro-
priate closures.
We conclude this subsection, and section, with the following remark:
Remark 7.5. A simple application of Hölder's inequality shows that if M¡0, q¢
and M¡1, q¢ are finite then
< ∞ for every θ ∈ (0,1).
M¡θ, q¢ ≤ M¡0, q¢1−θ
M¡1, q¢θ
8. FINAL REMARKS
In this last section of our work, we reflect on possible generalisations of our
main results via notions of equivalence of weight functions, and possible appli-
cations to evolution equations. Then, finally, we discuss some further possibili-
ties for research related to the issue of interpolation of weighted Sobolev spaces.
8.1. Equivalence of weights. As our main results of this work depend strongly
on the underlying weights of our respective spaces, one natural question can
arise:
Question 5. Can one change the weight under consideration to a 'better one' (i.e.
suitable for our theorems) without changing the underlying Sobolev space?
The answer to that is in the affirmative - in the obvious case when one deals
with equivalent weights.
Definition 8.1. Adopting standard terminology, we say that two strictly positive
measurable functions ω and η are equivalent, if there exists a constant A > 0
such that
A−1η(x) ≤ ω(x) ≤ Aη(x).
It is immediate to check that if ω0 and ω1 satisfy the compact boundedness
condition, and are equivalent, respectively, to η0 and η1, then the latter func-
tions also satisfy the same condition. Moreover, W 1,p (ω) and W 1,p¡η¢ are iso-
morphic to each other for equivalent weights. (Remark 1.4 has some relevance
in this context.)
Accordingly, one can immediately state and almost immediately prove more
general versions of our main theorem and its main consequences, in which the
weight functions ω0 and ω1 are required merely to be equivalent to other weight
functions which satisfy the requirements in the current statements of those the-
orems.
8.2. Possible applications to evolution equations. Many linear evolution equa-
tions of the form
(8.1)
give rise to an evolution semigroup, {T (t )}t≥0, which acts smoothly on several
spaces simultaneously. Moreover, in fields such as parabolic PDEs and kinetic
theory, one frequently encounters evolution semigroups whose action on the
∂t f (t , x) = L ( f )(t , x)
46
MICHAEL CWIKEL AND AMIT EINAV
initial datum of the problem produces functions which have increased smooth-
ness, decay and integrability (hypercontractivity). One good example of semi-
groups with such behaviour are the so-called Markov semigroups. We refer the
reader to [1] for more information on the subject.
In many cases, one is interested in the decay, or convergence to equilibrium as t
tends to ∞, of the solution at time t of (8.1) with respect to an appropriate norm,
such as k·kW 1,p(Rd ,ω).
Under suitable hypotheses, Theorem 1.21, together with observations that we
made in §7.2, allows us to interpolate rates of decay/convergence between two
different weighted Sobolev spaces to an intermediate space of the same kind.
More explicitly, if we know that for a given semigroup {T (t )}t≥0 the solution to
(8.1) decays or converges to an equilibrium at rates that are given by the norms
of T (t ) on appropriate subspaces of both W 1,p (U , ω0) and W 1,p (U , ω1), then it
may be possible to deduce analogous information about the norm of T (t ) on a
suitable subspace of W 1,p (U , ωθ), for θ ∈ (0,1).
We shall now give one example of a setting where this can be done. Here we as-
sume that the norms of T (t ) decay to zero. The underlying open set for all of the
spaces is to be understood to be Rd .
Theorem 8.2. Let (ω0, ω1) be a pair of weight functions that satisfy the compact
boundedness condition on Rd , and such that rω, defined by (1.9), is locally Lips-
chitz on Rd . Consider a semigroup of operators {T (t )}t>0 such that, for some fixed
p ∈ [1,∞),
(8.2)
T (t ) :¡W 1,p (ω0),W 1,p (ω1)¢ →¡W 1,p (ω0),W 1,p (ω1)¢ for each t > 0.
Assume furthermore that there exist q ∈ (p,∞), θ ∈ (0,1), t0 > 0 and a class of
initial data G such that for every g ∈ G
T (t0)g ∈ W 1,p (ωθ)∩ L
qp
q−p (ωθ).
Suppose furthermore that M¡θ, q¢, defined in (7.7), is finite for these values of θ
and q. Then we have, for every g ∈ G and t > t0, that T (t )g ∈ W 1,p (ωθ) and also
that
°°T (t )g°°W 1,p (ωθ) ≤ C kT (t − t0)k1−θ
µ°°T (t0)g°°W 1,p (ωθ) + M¡θ, q¢°°T (t0)g°°L
where the constant C depends only on p.
B(W 1,p (ω0))kT (t − t0)kθ
q−p (ωθ)¶
qp
B(W 1,p (ω1))
Proof. We start by noting that, due to Theorem 7.4, for each g ∈ G we have that
T (t0)g ∈£W 1,p (ω0),W 1,p (ω1)¤θ .
Recalling Theorem 3.3 and the semigroup property
T (t ) = T (t − t0)T (t0)
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
47
which holds for every t ≥ t0, we see that for each g ∈ G and t > t0 we have that
T (t )g ∈£W 1,p (ω0),W 1,p (ω1)¤θ ⊂ W 1,p (ωθ). Furthermore, using this and (3.4)
together with (8.2) and Theorem 2.4, we obtain that
°°T (t )g°°W 1,p (ωθ) ≤°°T (t )g°°[W 1,p (ω0),W 1,p (ω1)]θ =°°T (t − t0)¡T (t0)g¢°°[W 1,p (ω0),W 1,p (ω1)]θ
≤ kT (t − t0)k1−θ
B(W 1,p (ω0))kT (t − t0)kθ
.
B(W 1,p (ω1))°°T (t0)g°°[W 1,p (ω0),W 1,p (ω1)]θ
Using (7.9) yields the inequality required to complete the proof of this theorem.
(cid:3)
8.3. Future research and open problems. We feel excited about future possi-
bilities in continuing to try and understand whether or not one can obtain a
Stein-Weiss like theorem for weighted Sobolev spaces under more general hy-
potheses on the underlying set and the weight functions. In particular, some of
the problems which we think are worth investigating are the following:
Question. Can one find an open set U and weights ω0 and ω1 on it for which
W p (U , θ, rω) ⊂6=£W 1,p (U , ω0),W 1,p (U , ω1)¤θ?
(U , ω) = W 1,p (U , ω)
Question. Can one find necessary and sufficient conditions under which
0 (U , θ, rω) = W p (U , θ, rω)?
and W
1,p
0
W
p
Let us mention that pages 550-554 of [12] contain a discussion of the closure
of C∞c (U ) (there denoted by C∞0 (U )) in various weighted Sobolev spaces. A dif-
ferent but somewhat related topic is considered on pages 43-49 of [11].
Question. Can one obtain similar results to Theorems 1.16, 1.20, 1.21 and 1.22
for weighted homogeneous Sobolev spaces W 1,p (U , ω) (i.e. where we only con-
sider the norm of the weak gradient)?
(The answer is evidently yes for each p ∈ [1,∞) in the very special case where
d = 1 and U = R and both of the weight functions satisfy the compact bounded-
ness condition. In that case the maps φ 7→ φ′ and its inverse ψ 7→´ x
0 ψ(t )d t give
us an isometric identification between the two couples
¡ W 1,p (R, ω0), W 1,p (R, ω1)¢ and ¡Lp (R, ω0), Lp (R, ω1)¢
and so we can apply the "classical" Stein-Weiss theorem. A more detailed expla-
nation can be found in the appendix.)
We are confident that more can be done (and asked), and will be done in the
next few years.
APPENDIX A. ADDITIONAL PROOFS
In this appendix we include some additional proofs that we felt would hinder
the flow of the main body of the paper.
48
MICHAEL CWIKEL AND AMIT EINAV
1,p
l oc (U ). In this part of the
A.1. An alternative characterization of the space W
appendix we will provide some rather straightforward arguments which can be
used to justify the following claim, made immediately after Definition 4.7:
1,p
l oc (U ) if and only if, for every open subset V of U whose clo-
Claim A.1. f ∈ W
sure is a compact subset of U , the restriction of f to V is an element of W 1,p (V ).
In order to prove this claim, we require the following notation: For a given
open set U , and each n ∈ N we define the open set
Un = {x ∈ U dist(x, ∂U ) > 1/n}∩ Bn(0)
(where Bn(0) is the open ball of radius n in Rd centred at the origin). The closure
of Un is contained in U and in Bn(0) and is therefore a compact subset of U .
Proof of Claim A.1. Suppose that the function f : U → C has the property that,
for each n ∈ N, its restriction fn := f¯¯Un
to Un has a weak gradient on Un. Let
us denote this weak gradient by gn. We do not yet need to suppose also that
fn ∈ W 1,p (Un) but later, when we do, we will also be able to assert that gn ∈
Lp (Un, Cd ).
Due to the uniqueness of the weak gradient, and the fact that Un ⊂ Un1 ⊂ U for
all n1 > n, we see that if n1 > n we have that gn1Un = gn. For each n ∈ N let
vn : U → Cd be the function which coincides with gn on Un and equals ~0 on
U \Un. Thus we have that
(A.1)
vn1(x) = gn1(x) = gn1(x)¯¯Un = gn(x) = vn(x)
for all x ∈ Un and all n1 > n and all n ∈ N.
This shows that, for each fixed n ∈ N, the pointwise limit limk→∞ vk(x) exists for
all x ∈ Un. Since U =Sn∈NUn this in turn implies that this same pointwise limit
limk→∞ vk(x) exists for all x ∈ U and so we can define a new function g : U → Cd
by setting
g (x) := lim
k→∞
vk(x)
for all x ∈ U .
In view of (A.1) we have that, for each n ∈ N,
(A.2)
g (x) = gn(x) for all x ∈ Un
We will now show that f has a weak gradient on U and that g is that gradient.
Given an arbitrary φ ∈ C∞c (U ) let K be a compact set contained in U such that φ
and therefore also ∇φ both vanish on U \K . There exists n ∈ N such that K ⊂ Un,
and therefore the restriction of φ to Un is in C∞c (Un). We also have that
(A.3)
φ(x) = 0 and ∇φ(x) =~0 for all x ∈ U \Un.
Consequently, using (A.3), then the definitions of fn and gn and then (A.2) and
then (A.3) again, we obtain that
U
f ∇φd x =Un
= −Un
f ∇φd x =Un
φg d x = −U
φg d x.
fn∇φd x = −Un
φgnd x
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
49
Since the preceding argument holds for every φ ∈ C∞c (U ), this shows that g in-
deed is the (necessarily unique) weak gradient of f on U .
Suppose now that f has the property that, for every open subset V of U such that
V is a compact subset of U , the restriction of f to V is an element of W 1,p (V ).
Then the preceding argument shows that f has a weak gradient ∇ f on U which
equals the function g obtained above. Furthermore, the functions fn and gn de-
fined as above now satisfy fn ∈ W 1,p (Un) and gn ∈ Lp (Un, Cd ) and they coincide
with the restrictions of f and of g respectively to the set Un . Thus we have χUn f ∈
Lp (U ) and χUn∇ f ∈ Lp (U , Cd ). Let K be an arbitrary compact subset of U . For
some n ∈ N we have that K ⊂ Un and so χK f ∈ Lp (U ) and χK∇ f ∈ Lp (U , Cd ).
This shows that f ∈ W
We now prove the reverse implication.
Suppose that f ∈ W
an arbitrary open subset of U whose closure K := V is a compact subset of U .
l oc (U ) and that ∇ f denotes its weak gradient on U . Let V be
1,p
l oc (U ).
1,p
1,p
Let g = f¯¯V and let ∇g denote the Cd -valued function defined only on V which
1,p
l oc (U ) implies that χK f and
is the weak gradient of g on V . The fact that f ∈ W
χK∇ f are elements, respectively of Lp (U ) and Lp (U , Cd ). The restriction of χK f
to V must of course equal g . Furthermore, by the definition of weak gradients
and by their uniqueness, the restriction to V of χK ∇ f must equal ∇g . The fact
l oc (U ) implies that χK f and χK ∇ f are elements, respectively of Lp (U )
that f ∈ W
and Lp (U , Cd ). Therefore g and ∇g are in Lp (V ) and Lp (V, Cd ) respectively.
In other words g ∈ W 1,p(V ), and this completes our proof of Claim A.1.
A.2. From locally Lipschitz to globally Lipschitz - A proof of Fact (a). In this
part of the appendix we will prove the following claim, which was formulated as
the fact (a) just after the statement of Lemma 4.10:
(cid:3)
Claim A.2. If the function ψ : U → C is locally Lipschitz and has compact sup-
port, then it is Lipschitz on all of U .
Proof. Suppose that ψ : U → C is locally Lipschitz and has compact support.
Then there exists a compact subset K of U such that ψ vanishes on U \ K . The
number δ := dist(K , Rd \U ) is strictly positive and we let K1 be the set
K1 =½x ∈ Rd : dist(x, K ) ≤
δ
2¾.
Obviously K1 ⊂ U . The compactness of K implies that K1 is closed and bounded
and therefore also compact. Our hypotheses on ψ and the compactness of K1
ensure that the quantities
and
C1 := sup½¯¯ψ(x)− ψ(y)¯¯
¯¯x − y¯¯
: x, y ∈ K1, x 6= y¾
C2 := sup©¯¯ψ(x)¯¯ : x ∈ K1ª
50
MICHAEL CWIKEL AND AMIT EINAV
are both finite. To complete the proof we will show that
(A.4)
¯¯ψ(x)− ψ(y)¯¯ ≤¯¯x − y¯¯ max½C1,
2C2
δ ¾
for every x and y in U . The left side of (A.4) equals 0 whenever x and y are both
in U \ K1 and it is bounded by C1¯¯x − y¯¯ whenever x and y are both in K1. Thus
we only need to consider what happens when x ∈ K1 and y ∈ U \ K1.
• If x ∉ K then we again have¯¯ψ(x)− ψ(y)¯¯ = 0.
• If x ∈ K then¯¯x − y¯¯ > δ/2 and so
¯¯ψ(x)− ψ(y)¯¯ =¯¯ψ(x)¯¯ ≤ C2 =
2C2
δ ·
δ
2 ≤
2C2
δ ¯¯x − y¯¯.
We conclude that (A.4) indeed holds for all x and y in U and our proof is com-
plete.
(cid:3)
A.3. A Stein-Weiss like theorem in the case where rω is not Lipschitz. In this
subsection of the appendix we will focus on showing that there exist weight
functions, ω0 and ω1, for which one can obtain a Stein-Weiss like theorem, (i.e.
for which (1.4) is valid) even though rω, defined in (1.9), is not Lipschitz. This
will show that Theorem 1.22 gives a sufficient yet not necessary condition for
obtaining such a result.
Our starting point for finding such an example is the following simple lemma:
Lemma A.3. Let ρ0 and ρ1 be weight functions on Rd which satisfy the hypotheses
of Theorem 1.22. (In particular this implies that the function log³ ρ0
on Rd .) Let ω1 be a weight function on Rd which satisfies
cρ1(x) ≤ ω1(x) ≤ C ρ1(x) for all x ∈ Rd
(A.5)
for two positive constants c and C . If we set ω0 = ρ0, then we have that
ρ1´ is Lipschitz
hW 1,p³Rd , ω0´ ,W 1,p³Rd , ω1´iθ = W 1,p³Rd , ωθ´
to within equivalence of norms.
Proof. This is an immediate consequence of Theorem 1.22 combined with the
observations made in Subsection 8.1.
(cid:3)
Lemma A.3 gives us a possible approach for constructing our desired exam-
ple: We start with a pair of weights that satisfy our Stein-Weiss like theorem,
Theorem 1.22, and modify one of the weights in a way that keeps the new weight
equivalent to the original weight, but violates the Lipschitz condition on rω. We
shall do that now.
Consider the weight functions ρ0 ≡ 1 and ρ1(x) = e−p1+x2
and ρ1 both satisfy the compact boundedness condition and that
. It is obvious that ρ0
logµ ρ0(x)
ρ1(x)¶ =p1+x2
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
51
is Lipschitz on Rd . Thus, we conclude that ρ0 and ρ1 satisfy the conditions of
Theorem 1.22.
Now, let ω1 be the function
ω1(x) = e−p1+x2−sin³ex2´.
Clearly ω1 satisfies (A.5) with c = 1/e and C = e.
Since rω = ω0/ω1 = 1/ω1 we see that
∇log (rω(x)) = ∇³p1+x2 + sin³ex2´´ =
x
p1+x2 + 2xex2
cos³ex2´.
The above is unbounded on Rd , which gives us the example we needed.
In fact, one can do more with this example. One can even use it to show that this
pair of weight functions has the property that:
W p¡Rd , θ, rω¢ is strictly smaller than W 1,p¡Rd , ωθ¢
for every p ∈ [1,∞) and θ ∈ (0,1).
We leave the proof of this claim to the reader.
This example also leads us naturally to ask:
Question. Can one find weight functions ω0 and ω1 on Rd , that are not equiv-
alent to weights that satisfy the conditions of Theorem 1.22 but for which we
nevertheless have (1.4)?
We'd like to conclude this subsection by noting that a theme common to this
example and the one presented in Subsection 7.1 is that they both use weight
functions which are the product of a "well behaved" function with a highly os-
cillatory but bounded function.
A.4. A comment about homogeneous weighted Sobolev spaces of univariate
functions. In this last subsection we will show that, when the open set U is R,
one can easily obtain a Stein-Weiss like theorem for the homogeneous weighted
Sobolev spaces, W 1,p(R, ω), as a direct consequence of the regular Stein-Weiss
theorem (as we claimed at the end of Subsection 8.3).
We start with a definition
Definition A.4. For each weight function ω : R → (0,∞) and each p ∈ [1,∞] let
W 1,p(R, ω) be the space of equivalence classes modulo constants of measurable
functions f : R → C which have a weak derivative f ′ in Lp (R, ω). This space is
normed by°° f°° W 1,p (R,ω) =°° f ′°°Lp (R,ω).
Suppose that ω satisfies the compact boundedness condition, or at least that
it satisfies the weaker condition (5.4) which we used in Lemma 5.2. Then every
function g in Lp (R, ω) is locally integrable and therefore the function
G(x) := x
0
g (t )d t
is defined for all x ∈ R and is absolutely continuous on every bounded inter-
val. Moreover, its pointwise derivative G′ exists almost everywhere and coin-
cides with g almost everywhere. The fact that g is also the weak derivative of G
52
MICHAEL CWIKEL AND AMIT EINAV
is immediate due to the fact that we are on R and G is absolutely continuous.
From all the above we see that the map T defined by
T g (x) = x
0
g (t )d t
is a continuous linear map from Lp (R, ω) onto W 1,p (R, ω). The inverse mapping
of W 1,p(R, ω) onto Lp (R, ω) is of course simply the derivative map D defined by
D f = f ′. It is also clear that T and D are both isometries. Since Lp (R, ω) is
complete, we conclude that W 1,p (R, ω) is complete for all weight functions ω
which satisfy (5.4).
Now, given p0 and p1 in [1,∞) and weight functions ω0 and ω1 on R which
both satisfy the compact boundedness condition (as we saw, we can also make
do with somewhat weaker assumptions), we see that¡ W 1,p0 (R, ω1), W 1,p1 (R, ω1)¢
is a Banach couple. We use Theorem 2.4 to obtain that the operators T and D
satisfy
and
T :£Lp0 (R, ω0), Lp1 (R, ω1)¤θ →£ W 1,p0 (R, ω1), W 1,p1 (R, ω1)¤θ
D :£ W 1,p0 (R, ω1), W 1,p1 (R, ω1)¤θ →£Lp0 (R, ω0), Lp1 (R, ω1)¤θ ,
with norms that do not exceed 1.
Since the weight function ωθ,pθ defined by (1.1) also has the compact bounded-
ness property, we see that T and D are also isometries, respectively, of Lpθ¡R, ωθ,pθ¢
onto W 1,pθ¡R, ωθ,pθ¢ and of W 1,pθ¡R, ωθ,pθ¢ onto Lpθ¡R, ωθ,pθ¢. Combining these
facts with the formula (3.3) shows that
with equality of norms.
£ W 1,p0 (R, ω1), W 1,p1 (R, ω1)¤θ = W 1,pθ¡R, ωθ,pθ¢,
REFERENCES
[1] D. Bakry, I. Gentil, and M. Ledoux. Analysis and Geometry of Markov Diffusion Operators, vol-
ume 348 of Grundlehren der mathematischen Wissenschaften. Springer International Pub-
lishing, Cham, 2014.
[2] J. Bergh and J. Löfström, Interpolation Spaces, an introduction. Grundlehren der Mathema-
tischen Wissenschaften, No. 223. Springer-Verlag, Berlin-New York, 1976.
[3] A. P. Calderón, Intermediate spaces and interpolation, the complex method. Studia Math., 24
(1964), 113 -- 190. http://matwbn.icm.edu.pl/ksiazki/sm/sm24/sm24110.pdf
[4] M. Cwikel, Complex interpolation, a discrete definition and reiteration, Indiana Univ. Math.
J. 27 (1978), 1005 -- 1009.
[5] D. E. Edmunds and H. Triebel, Function spaces, entropy numbers, differential operators. Cam-
bridge Univ. Press, Cambridge, 1996.
[6] C. L. Evans, Partial differential equations. Second edition. Graduate Studies in Mathematics,
19. American Mathematical Society, Providence, RI, 2010. xxii+749 pp.
[7] P. Grisvard, Espaces intermédiaires entre espaces de Sobolev avec poids. (French) Ann. Scuola
Norm. Sup. Pisa. 17 (1963), 255 -- 296.
[8] D. D. Haroske and H. Triebel, Entropy numbers in weighted function spaces and eigenvalue
distributions of some degenerate pseudodifferential operators. I, Math. Nachr. 167 (1994), 131 --
156.
INTERPOLATION OF WEIGHTED SOBOLEV SPACES
53
[9] D. D. Haroske and H. Triebel, Entropy numbers in weighted function spaces and eigenvalue
distributions of some degenerate pseudodifferential operators. II, Math. Nachr. 168 (1994),
109 -- 137.
[10] D. D. Haroske and H. Triebel, Wavelet bases and entropy numbers in weighted function spaces,
Math. Nachr. 278 (2005), 108 -- 132.
[11] A. Kufner, Weighted Sobolev spaces. Translated from the Czech. A Wiley-Interscience Publi-
cation. John Wiley & Sons, Inc., New York, 1985. 116 pp. ISBN: 0-471-90367-1.
[12] A. Kufner and B. Opic, How to define reasonably weighted Sobolev spaces. Comment. Math.
Univ. Carolin. 25 (1984), 537 -- 554.
[13] A. Kufner and A-M Sändig, Some applications of weighted Sobolev spaces, Teubner Texts in
Mathematics, 100. B. G. Teubner Verlagsgesellschaft, Leipzig, 1987. 268 pp.
[14] J. Löfström, Interpolation of weighted spaces of differentiable functions on Rd . Ann. Mat. Pura.
Appl. 132 (1982), 189 -- 214.
[15] S. G. Pyatkov, Interpolation of some function spaces and indefinite Sturm-Liouville problems,
(English summary) Differential and integral operators (Regensburg, 1995), 179 -- 200, Oper.
Theory Adv. Appl., 102, Birkhäuser, Basel, 1998.
[16] S. G. Pyatkov. Interpolation of weighted Sobolev spaces. Siberian Adv. Math. 10 (2000), no. 3,
83 -- 132.
[17] S. G. Pyatkov, Interpolation of weighted Sobolev spaces. (Russian. Russian summary) Mat. Tr.
4 (2001), no. 1 122 -- 173.
[18] S. G. Pyatkov, Interpolation of Sobolev spaces and indefinite elliptic spectral problems, (Eng-
lish summary) Recent advances in operator theory in Hilbert and Krein spaces, 265 -- 290,
Oper. Theory Adv. Appl., 198, Birkhäuser Verlag, Basel, 2010.
[19] H.-J. Schmeisser and H. Triebel, Topics in Fourier analysis and function spaces, Leipzig: Geest
& Portig, 1987; Chichester; Wiley, 1987.
[20] E. Stein, Interpolation of linear operators. Trans. Amer. Math. Soc. 83 (1956), 482 -- 492.
[21] E. M. Stein and G. Weiss, Interpolation of operators with change of measures, Trans. Amer.
Math. Soc. 87 (1958), 159 -- 172.
[22] H. Triebel, Interpolation theory, function spaces, differential operators, North Holland Pub-
lishing Company 1978.
[23] R. Wheeden and A. Zygmund, Measure and integral. An introduction to real analysis. Pure
and Applied Mathematics, Vol. 43. Marcel Dekker, Inc., New York-Basel, 1977.
DEPARTMENT OF MATHEMATICS, TECHNION - ISRAEL INSTITUTE OF TECHNOLOGY, HAIFA 32000,
ISRAEL
E-mail address: [email protected]
INSTITUT FÜR ANALYSIS UND SCIENTIFIC COMPUTING, TECHNISCHE UNIVERSITÄT WIEN, WIED-
NER HAUPTSTRASSE 8-10 A-1040 VIENNA, AUSTRIA
E-mail address: [email protected]
|
1106.3351 | 1 | 1106 | 2011-06-16T21:15:59 | The spine of a Fourier-Stieltjes algebra: corrigenda | [
"math.FA",
"math.OA"
] | Some unfortunate errors from our paper math/0505591 are corrected. | math.FA | math |
THE SPINE OF A FOURIER-STIELTJES ALGEBRA:
CORRIGENDA
MONICA ILIE AND NICO SPRONK
It has come to the authors' attention that there are several errors in our
paper [3]. Fortunately, these errors are correctable. Some of these errors
carry on to modest, though mainly cosmetic, errors in follow-up papers [5]
and [4].
We are grateful to El¸cim Elgun for pointing out the gap in the proof of
[3, Theorem 2.2] and the error in the proof of [3, Theorem 5.1]. We are
also grateful to Pekka Salmi for pointing out the flaw in the statement of [3,
Theorem 4.2] and suggesting its correct form.
We appeal to [3] for pertinent notation and terminology.
1. On non-quotient locally precompact topologies
The following should replace [3, Theorem 2.2].
Theorem 1.1. Let τ ∈ T (G). Then the following hold.
(i) There exists a unique τnq in Tnq(G) such that τ is a quotient of τnq.
(ii) If G is abelian, then τnq = τ ∨ τap.
In [4, Theorem 2.2] it is claimed that τnq = τ ∨ τap for general locally
compact groups G and τ in T (G). This is false. For example let G =
SL2(R). We have that, on G, τap is the trivial topology ε = {∅, G}. Now
if q : G → G/{−I, I} is the quotient map and τ = q−1(τG/{−I,I}) is the
coarsest topology making q continuous, then we have that τnq = τG whereas
τ ∨ τap = τ ( τG.
may be used to show that s 7→ (cid:0)ητ1
The proof of part (ii) proceeds exactly as does the proof of [3, Theorem
2.2]. In particular, in the second paragraph of that proof, [3, Lemma 2.3]
τ1 is a bicontin-
uous isomorphism, since Gτ1, being abelian, is maximally almost periodic.
The proof of part (i), however, demands more care. We fix τ0 in T (G)
ap) : Gτ1 → Gτ × Gap
τ (s), ητ1
and let
(1.1)
Qτ0 = {τ ∈ T (G) : τ0 is a quotient of τ }.
Date: October 6, 2018.
2000 Mathematics Subject Classification. Primary 43A30; Secondary 43A60, 43A07,
46L07, 22B05. Key words and phrases. Locally precompact topology, spine.
We like to thank NSERC for partial funding of our work.
1
2
MONICA ILIE AND NICO SPRONK
Lemma 1.2. (i) If τ1, τ2 ∈ Qτ0 then τ1 ∨ τ2 ∈ Qτ0 as well.
(ii) If τ1, τ2 ∈ T (G) satisfy τ0 ⊆ τ1 ⊆ τ2 and τ2 ∈ Qτ0 , then τ1 is a
quotient of τ2.
Proof. (i) For j = 1, 2 we let K j
τ0 . Our assumptions provide
that for j = 1, 2, K j
∼= Gτ0 . We identify Gτ1∨τ2
as a subgroup of Gτ1 × Gτ2 and define K = Gτ1∨τ2 ∩ (K 1
0 ). We let
q : Gτ1∨τ2 → Gτ1∨τ2/K denote the quotient map. We obtain the following
commutative diagram
0 = ker ητj
0 is compact and Gτj /K j
0 × K 2
0
Gτ1∨τ2
q
Gτ1∨τ2/K
/ Gτ1 × Gτ2
ητ1
τ0 ×ητ2
τ0
+WWWWWWWWWWWWWWWWWWWWWW
/ Gτ1 × Gτ2 /(K 1
0 × K 2
0 ) ∼= Gτ1 /K 1
0 × Gτ2/K 2
0
∼ Gτ0 × Gτ0.
Identifying Gτ0 with the diagonal subgroup of Gτ0 × Gτ0, this diagram shows
that the map ητ1∨τ2
τ0 Gτ1 ∨τ2 is a proper map.
τ0 × ητ2
= ητ1
τ0
(ii) We recall that our assumptions give the following commuting diagram
(1.2)
Gτ2
ητ2
τ1
H
H
H
H
ητ2
τ0
H
H
H
H
#H
Gτ2 /K 2
0
/ Gτ1
ητ1
τ0
∼ Gτ0
τ0 is compact. Then K 2
0 = ker ητ2
where K 2
0 , and is thus
compact. We let q2
1 be the quotient map and τ =
(q2
τ1 is injective.
The commuting diagram (1.2) and the first isomorphism theorem give the
commuting diagram
), so that Gτ ∼= Gτ2/K 1
1 : Gτ2 → Gτ2 /K 2
1 , τ ⊇ τ1 and ητ
1 ◦ητ2)−1(τGτ2 /K 2
1 = ker ητ2
τ1 ⊂ K 2
1
(1.3)
Gτ2
ητ2
τ
/ Gτ
ητ
τ1
/ Gτ1
R
R
R
R
R
R
F
F
F
q
F
F
R
R
R
R
F
F
(R
R
R
R
R
ητ2
τ0
"F
ητ1
τ0
Gτ /K ∼ Gτ0
where K = K 2
0 /K 2
1 and q : Gτ → Gτ /K is the quotient map.
Since ητ
τ1 is injective, it suffices to prove that it is open for ητ2
τ1 to be a
quotient map, i.e. we obtain that τ1 = τ where ητ2
is a quotient map. To this
τ
end, let U ⊂ Gτ be relatively compact open set. Then U K is also relatively
compact and open. Hence ητ
τ1 (U K), and
furthermore
τ1(U K) is closed and equal to ητ
(1.4)
ητ
τ1 U K : U K → ητ
τ1(U K) is a homeomorphism.
The commuting diagram (1.3) tells us that
ητ
τ1(U K) = (ητ1
τ0 )−1(cid:0)q(U )(cid:1)
/
+
/
#
/
(
/
"
/
THE SPINE
which is open in Gτ1. Hence by (1.4),
ητ
τ1U K : U K → ητ
τ1 (U K)
is a homeomorphism onto an open subset so ητ
τ1(U ) is open.
3
(cid:3)
Proof of Theorem 1.1 (i). We first note that Qτ , as defined in (1.1),
is a directed system:
if τ1, τ2 ∈ Qτ then τ1 ∨ τ2 ∈ Qτ by Lemma 1.2 (i).
Moreover, it follows Lemma 1.2 (ii) that if τ1, τ2 ∈ Qτ with τ1 ⊆ τ2, then
ητ2
τ1 is a proper map. Hence the inverse mapping system
τ1 : τ ′ ∈ Qτ , τ1 ⊂ τ2 in Qτ }
{Gτ ′, ητ2
gives rise to the projective limit
Gτ ′ : ητ2
τ1 (sτ2) = sτ1 if τ1 ⊆ τ2 in Qτ
Gτ ′ : ητ ′
τ (sτ ′) = sτ for τ ′ in Qτ
GQτ = lim
←−
τ ′∈Qτ
Gτ ′ =
=
(sτ ′) ∈ Y
τ ′∈Qτ
(sτ ′) ∈ Y
τ ′∈Qτ
which is locally compact by [3, Proposition 2.1]. We let τnq denote the coars-
est topology which makes the map s 7→ (cid:0)ητ ′(s)(cid:1) : G → GQτ , continuous.
We now show that τnq ∈ Qτ . First observe that ητnq
∼= GQτ → Gτ
We have that τnq ⊇ τ ′ for every τ ′ in Qτ .
: Gτnq
τ
is given by the map (sτ ′) 7→ sτ . With this identification we have that
ker ητnq
τ = lim
←−
τ ′∈Qτ
ker ητ ′
τ
and is thus compact. Moreover ητnq
τ
is open, since for any basic open set
V =
Y
τ ′∈Qτ (cid:31){τ ′
1,...,τ ′
n}
Gτ ′ ×
nY
j=1
Uτ ′
j
⊂ Y
τ ′∈Qτ
Gτ ′
where each Uτ ′
j
is open in Gτ ′
j
, we have
ητnq
τ
(V ∩ GQτ ) =
n\
j=1
τ ′
j
η
τ (Uτ ′
j
)
) is open by assumption.
where each η
τ ′
j
τ (Uτ ′
j
Finally, if there were τ1 in T (G) of which τnq is a quotient, then by the
first isomorphism theorem we would have that τ1 ∈ Qτ . Hence τ1 ⊆ τnq.
Thus τnq is a non-quotient topology, and the unique such one of which τ is
a quotient.
(cid:3)
For any locally compact group G for which τnq = τ ∨τap for any τ in T (G),
we obtained in [3, Section 2.4] that Tnq(G) = T (G) ∨ τap and is thus an ideal
in, and hence a subsemilattice of, the semilattice (T (G), ∨). We note that
for any τ for which Gτ is maximally almost periodic, we have τnq = τ ∨ τap.
4
MONICA ILIE AND NICO SPRONK
Unfortunately, it is not clear whether Tnq(G) is a subsemilattice of T (G), in
general. However the following is immediate.
Corollary 1.3. If τ1, τ2 ∈ Tnq(G) define
τ1 ∨τ2 = (τ1 ∨ τ2)nq.
Then (Tnq(G), ∨) is a quotient semilattice of (T (G), ∨).
The only inobvious aspect of this corollary is the associativity of ∨. We
observe that τ1 ∨ (τ2 ∨τ3) admits τ1 ∨ τ2 ∨ τ3 as a quotient, and hence, by
Lemma 1.2 (ii), is itself a quotient of (τ1 ∨ τ2 ∨ τ3)nq. Symmetrically, the
same is true of (τ1 ∨τ2) ∨ τ3. Hence
τ1 ∨(τ2 ∨τ3) = (τ1 ∨ τ2 ∨ τ3)nq = (τ1 ∨τ2)∨τ3.
Unless it can be shown that τ1 ∨τ2 = τ1 ∨ τ2 for all τ1, τ2 ∈ Tnq(G), some
changes have to be made to the exposition in [3, Section 4], where ∨ must
always be replaced by, or understood to be, ∨. Fortunately, this change does
not appear to affect any of the results or proofs in this section, or any later
part of the paper, in more than a cosmetic manner.
2. Topology of the spine compactification
As shown in [3, Theorem 4.1], with appropriate notational changes as
suggested by Corollary 1.3, the spectrum of the algebra A∗(G) is given by
G∗ = G
S∈HD(G)
GS
where each GS is a projective limit over a hereditary directed subset S of
(Tnq(G), ∨). The statement of [3, Theorem 4.2] is flawed, and should be
replaced with the following.
Theorem 2.1. The topology on G∗ is given as follows: for any s0 in G∗,
say s0 ∈ GS0 for some S0 in HD(G), a neighbourhood base at s0 is formed
by the sets
U (Vτ ; Wτ1, . . . , Wτn ) =(cid:8) s ∈ G∗ :s ∈ GS for some S ⊇ Sτ in HD(G)
with sτ ∈ Vτ , and sτj ∈ Wτj if
S ⊇ Sτj , for j = 1, . . . , n (cid:9)
where τ ∈ S0, τ1, . . . , τn ∈ Tnq(G) \ S0, Vτ is an open neighbourhood of s0,τ
in Gτ , and each Wτj is a cocompact subset of Gτj .
Pekka Salmi has pointed out to us that the error in the description of
[3, Theorem 4.2], implies that all groups GS , for S ∈ HD(G) are locally
compact. This is false, as is implicit in [3, Section 6.3], or is shown in [1,
Theorem 2].
Unfortunately, the proof of [3, Theorem 4.2] requires slight modification.
THE SPINE
5
Proof of Theorem 2.1. We should first observe that the family of sets
described above indeed is a base for a topology.
It is straightforward to
check that for τ, τ ′ ∈ S0 and τ1, . . . , τn, τ ′
m in Tnq(G) \ S0 that
1, . . . , τ ′
m)
, . . . , Wτ ′
U (Vτ ; Wτ1 , . . . , Wτn ) ∩ U (Vτ ′; Wτ ′
1
⊃ U (cid:16)(ητ ∨τ ′
τ
)−1(Vτ ) ∩ (ητ ∨τ ′
τ ′
)−1(Vτ ′); Wτ1 , . . . , Wτn , Wτ ′
1
, . . . , Wτ ′
n(cid:17)
m
for neighbourhoods Vτ of s0,τ , Vτ ′ of s0,τ ′, and cocompact subsets Wτ1, . . . , Wτ ′
of Gτ1 , . . . , Gτ ′
m , respectively.
We note that for a net (si)i∈I in G∗, the following are equivalent:
(i) si → s0 in G∗ with the topology described above;
(ii) for each τ0 ∈ S0 there is i0 such that i ≥ i0 implies that si ∈ GSi
for some Si ⊃ Sτ0 and limi≥i0 si,τ = s0,τ ; and for each τ 6∈ S0 for which
Iτ = {i : si ∈ Si for some Si ⊃ Sτ } is admits no maximal element in I, and
for any co-compact Wτ ⊂ Gτ , there is iτ in I for which si,τ ∈ Wτ if i ∈ Iτ
and i ≥ iτ ;
(iii) χsi → χs0 weak* in A∗(G)∗.
The equivalence of (i) and (ii) is clear. If we write u in A∗(G) as u =
Pτ ∈Tnq(G) uτ as in [3, (4.2)], and suppose (ii), above, then
χsi(u) = X
uτ (si,τ ) + X
uτ (si,τ ) → X
uτ (s0,τ ) = χs0(u)
τ ∈S0\Si
τ ∈Si\S0
τ ∈S0
where si ∈ Si for each i; which shows (iii). Likewise, selecting u = uτ , a
repeat of the computation above shows that (iii) implies (ii).
(cid:3)
We remark that [3, Corollary 4.3] remains unchanged. The neighbourhood
in the proof of part (ii), therein, should be changed to
U (Vτ0) = {s ∈ G∗ : s ∈ GS ′ for some S ′ ⊃ Sτ0 and Sτ0 ∈ Vτ0}.
Furthermore, a modest change must be made in the proof of [3, Proposition
4.6 (ii)].
If a net (ei) of idempotents converges to s in GS , then for all
τ ∈ S, there is an iτ for which Si ⊃ Sτ , where ei ∈ GSi, for i ≥ iτ , and
limi≥iτ ei,τ = sτ ; and for τ in Tnq(G) \ S, there is iτ for which Si 6⊃ Sτ for
i ≥ iτ , which may be seen by observing that otherwise such ei,τ must be
within the cocompact set Wτ = Gτ \ {ητ (e)}.
3. On abelian groups
The fourth paragraph of the proof of [3, Theorem 5.1] contains an error
bicontinuous is false, for obviously only its inverse is continuous. Fortunately
the claim which that paragraph is attempting to establish, namely that if
in its claim that the map from cGτ to the diagonal subgroup of cGτ × bGd is
τ in T (G) is such that cGτ is the group bG but with a finer locally compact
: dGτ ′ → bG is continuous and injective. The
in T (G), the dual map cητ ′
assumptions on τ above imply that bητ is also surjective. Since the map
group topology τ , then τ ∈ Tnq(G), remains true. Recall that for any τ ′
6
MONICA ILIE AND NICO SPRONK
ητnq
τ
example. Since ητ = ητnq
dητnq
: Gτnq → Gτ is proper, its dual map dητnq
is continuous, bijective and open, hence τ = cτnq. Thus by Pontryagin
◦ητnq we have bητ = dητnq
τ
duality τ = τnq.
◦dητnq
τ
τ
τ
is open; see [2, (23.24)(d)], for
. Thus it follows that
References
[1] C.F. Dunkl, D.E. Ramirez. Locally compact subgroups of the spectrum of the measure
algebra II Semigroup Forum, 3:267 -- 269, 1971.
[2] E. Hewitt and K.A. Ross. Abstract Harmonic Analysis I Second Ed., Grundlehern
der mathematics Wissenshaften 115, Springer, New York, 1979.
[3] M. Ilie and N. Spronk. The spine of a Fourier-Stieltjes algebra Proc. Lond. Math. Soc.
(3), 94:273 -- 301, 2007.
[4] M. Ilie and N. Spronk. The algebra generated by idempotents in a Fourier-Stieltjes
algebra Houston J. Math., 33:1131 -- 1145, 2007.
[5] V. Runde and N. Spronk. Operator amenability of Fourier-Stieltjes algebras. II Bull.
Lond. Math. Soc., 39:194 -- 202, 2007.
Monica Ilie, Department of Mathematical Sciences, Lakehead
University, 955 Oliver Road, Thunder Bay, ON, P7B 5E1, Canada
[email protected]
Department of Pure Mathematics, University of Waterloo,
[email protected]
Waterloo, ON, N2L 3G1, Canada
|
1903.00370 | 1 | 1903 | 2019-03-01T15:28:50 | Compact operators between lattice normed spaces | [
"math.FA"
] | In this paper we continue the study of compact-like operators in lattice normed spaces started recently by Aydin, Emelyanov, Erkur\c{s}un \"Ozcand and Marabeh. We show among others, that every p-compact operator between lattice normed spaces is p-bounded. The paper contains answers of almost all questions asked by these authors. | math.FA | math |
On compact operators between lattice normed
spaces
Youssef Azouzi and Mohamed Amine Ben Amor
Research Laboratory of Algebra, Topology, Arithmetic, and Order
Department of Mathematics
Faculty of Mathematical, Physical and Natural Sciences of Tunis
Tunis-El Manar University, 2092-El Manar, Tunisia
Abstract
In this paper we continue the study of compact-like operators in
lattice normed spaces started recently by Aydin, Emelyanov, Erkur¸sun
Ozcand and Marabeh. We show among others, that every p-compact
operator between lattice normed spaces is p-bounded. The paper con-
tains answers of almost all questions asked by these authors.
1
Introduction
In [3], the authors introduced a new notion of compact operators in Lattice-
normed spaces and studied some of their properties. These operators act on
spaces equipped with vector valued norms taking their values in some vector
lattices. Recall that an operator from a normed space X to a normed space
Y is said to be compact if the image of every norm bounded sequence (xn)
in X has a norm convergent subsequence. This notion has been generalized
in the setting of lattice normed spaces giving rise to two new notions: se-
quentially p-compactness and p-compactness (p referred to the vector valued
norm). Notice that these notions coincide in the classical case of Banach
spaces. In general setting with vector lattice valued norms boundedness and
convergence are considered with respect to these 'norms'. Also as notions
of relatively uniform convergence and almost order boundedness have been
1
generalized, new properties for the operator are considered like semicompact-
ness. Recall that if (E, p, V ) and (F, q, W ) are Lattice-normed spaces and
T is a linear operator from E to F , then T is said to be p-compact (respec-
tively, rp-compact) if for every p-bounded net (xα) in E, there is a subnet
(cid:0)T xϕ(β)(cid:1) that p-converges (respectively, rp-converges) to some y ∈ F . The
operator is said to be sequentially p-compact if nets and subnets are replaced
by sequences and subequences above. In this paper we prove some new re-
sults in this direction. Namely we show that every p-compact operator is
p-bounded. As a consequence we get that every rp-compact is p-bounded.
Also we give an example of sequentially p-compact operator which fails to
be p-bounded. As a consequence we deduce that a sequentially p-compact
need not be p-compact. In fact these two notions are totally independent.
Example of p-compact operators that fail to be sequentially p-compact is
given. As mentioned above the study of p-compact operators between lattice
normed spaces was started in [3]. That paper contains several new results
but also some open questions. Almost all these questions will be answered
in our paper.
2 Preliminaries
The goal of this section is to introduce some basic definitions and facts. For
general informations on vector lattices, Banach spaces and lattice-normed
spaces, the reader is referred to the classical monographs [1] and [6].
Consider a vector space E and a real Archimedean vector lattice V . A
map p : E → V is called a vector norm if it satisfies the following axioms:
1) p(x) ≥ 0; p(x) = 0 ⇔ x = 0; (x ∈ E).
2) p(x1 + x2) ≤ p(x1) + p(x2); (x1, x2 ∈ E).
3) p(λx) = λp(x); (λ ∈ R, x ∈ E).
A triple (E, p, V ) is a lattice-normed space if p(.) is a V -valued vector
norm in the vector space E. When the space E is itself a vector lattice the
triple (E, p, V ) is called a lattice-normed vector lattice. A set M ⊂ E is called
p-bounded if p (M) ⊂ [−e, e] for some e ∈ V+. A subset M of a lattice-normed
vector lattice (E, p, V ) is called p-almost order bounded if, for any w ∈ V+,
there is xw ∈ E+ such that p((x − xw)+) = p(x − xw ∧ x) ≤ w for any
x ∈ M.
2
p
Let (xα)α∈∆ be a net in a lattice-normed space (E, p, V ). We say that
(xα)α∈∆ is p-convergent to an element x ∈ E and write xα
−→ x, if there
exists a decreasing net (eγ)γ∈Γ in V such that inf γ∈Γ(eγ) = 0 and for every
γ ∈ Γ there is an index α(γ) ∈ ∆ such that p(v − vα) ≤ eγ for all α ≥ α(γ).
Notice that if V is Dedekind complete, the dominating net (eγ) may be chosen
over the same index set as the original net. We say that (xα) is p-unbounded
convergent to x (or for short, up-convergent to x) if xα − x ∧ u
−→ 0 for all
u ∈ V+. It is said to be relatively uniformly p-convergent to x ∈ X (written
rp
as, xα
−→ x) if there is e ∈ E+ such that for any ε > 0, there is αε satisfying
p(xα − x) ≤ εe for all α ≥ αε.
p
When E = V and p is the absolute value in E, the p-convergence is the
order convergence, the up-convergence is the unbounded order convergence,
and the rp-convergence is the relatively uniformly convergence. We refer to
[5] and [4] for the basic facts about nets in topological spaces and vector
lattices respectively. We will use [6, 8] as unique source for unexplained
terminology in Lattice-Normed Spaces. Since the most part of this paper is
devoted to answer several open questions in [3], the reader must have that
paper handy, from which we recall some definitions.
Definition 1 Let X, Y be two lattice-normed spaces and T ∈ L(X, Y ).
Then
1. T is called p-compact if, for any p-bounded net (xα) in X, there is a
p
subnet xαβ such that T xαβ
−→ y in Y for some y ∈ Y .
2. T is called sequentially p-compact if, for any p-bounded sequence xn in
−→ y in Y for some
X, there is a subsequence (xnk ) such that T xnk
y ∈ Y .
p
3. T is called p-semicompact if, for any p-bounded set A in X, the set
T (A) is p-almost order bounded in Y .
3
p-compact operators are p-bounded
It is well known that compact operators between Banach spaces are bounded.
This result remains valid for general situation of p-compact operators as it
will be shown in our first result, which answers positively Question 2 in [3].
3
Theorem 2 Every p-compact operator between two Lattice-normed spaces is
p-bounded.
Proof. Assume, by contradiction, that there exists a p-compact operator
T : (E, p, V ) −→ (F, q, W ) which is not p-bounded. Then there exists a
p-bounded subset A of E such that T (A) is not q-bounded. So, for every
u ∈ W + there exists some xu ∈ A satisfying q(T (xu)) 6≤ u. Since the
net (xu)u∈W + is p-bounded there is a subnet (cid:0)yv = xϕ(v)(cid:1)v∈Γ and an element
f ∈ F such that (T yv)
tail, which means that for some v0 in Γ and some w ∈ W + we have,
q
−→ f . It follows that the net (T yv) has a q-bounded
q(cid:0)T xϕ(v)(cid:1) ≤ w,
for v ≥ v0.
(1)
Pick v1 in Γ such that ϕ(v) ≥ w for all v ≥ v1. It follows that for v ≥ v0 ∨ v1,
we have q(T xϕ(v)) 6≤ ϕ(v) and so
q(T xϕ(v)) 6≤ w,
which is a contradiction with 1. and the proof comes to its end.
The following lemma, which connects unbounded order convergence with
pointwise convergence, is a known fact, although a quick proof is included
for the sake of completeness.
Lemma 3 Let E = RX be the Riesz space of all real-valued functions defined
on a nonempty set X. The following statements are equivalent:
(i) The net (fα)α∈A is uo-convergent in E.
(ii) for every x ∈ X, the net (fα (x))α∈A is convergent in R.
uo−→ f in the Dedekind complete Riesz
Proof. (i) =⇒ (ii) Assume that fα
space E. Then there is a net (gα)α∈A which decreases to 0 and for some α0
we have
fα − f ∧ 1 ≤ gα for all α ≥ α0.
(2)
Since (gα (x)) decreases to 0 for every x ∈ X, it follows easily from 2 that
fα (x) − f (x) converges to 0, as desired.
(ii) =⇒ (i) Assume now that fα converges simply to some f ∈ E and let
h ∈ E+. Define a net (gα) by putting
gα (x) = sup
β≥α
(fβ − f ∧ h) (x) , x ∈ X.
4
it is clear that gα decreases to 0 and fα − f ∧ h ≤ gα. This shows that
fα
uo−→ f and we are done.
Consider the Riesz space F of all bounded real valued functions defined
on the real line with countable support and denote by E the direct sum
R1 ⊕ F, where 1 denotes the constant function taking the value 1. This
example will be of great interest for us. The following lemma establishes
some of its properties. Recall that a vector sublattice Y of a vector lattice X
is said to be regular if every subset in Y having a supremum in Y has also
a supremum in X and these suprema coincide. For more information about
this notion and nice characterizations of it via unbounded order convergence
the reader is referred to [4].
Lemma 4 The space E introduced above has the following properties.
(i) E is a regular vector sublattice of RR.
(ii) E is Dedekind σ-complete but not Dedekind complete.
Proof. (i) It is clear that E is a vector sublattice of RR. To show that it is
regular assume that (gα)α∈A is a net in E satisfying gα ↓ 0 in E. Let g = inf
gα
α
in RR and x ∈ R. Then h = g (x) 1{x} ∈ E and 0 ≤ h ≤ gα for all α. This
implies that h = 0 and then g (x) = 0. Hence g = 0 and the regularity is
proved.
(ii) Let (gn) be an order bounded sequence in E and write gn = λn + fn,
with λn ∈ R and fn ∈ F. Let Ω be the union of the supports of fn, then Ω is
countable. Let g be the supremum of (gn) in RR, that is,
g (a) = sup gn (a) , for all a ∈ R.
It will be sufficient to show that g ∈ E. To this end observe that g (x) =
α := sup αn for all x ∈ R\Ω. Now put f = (g − α) 1Ω. Then f ∈ F and
g = α + f ∈ E as required. Next we show that E is not Dedekind complete.
Consider the net (gx)x∈[0,1] in E defined by gx = x1{x}. It is a bounded net
in E and its supremum in RR does not belong to E. As E is regular in RR
this net can not have a supremum in E.
Remark 5 Consider the following operator:
T : L1 [0, 1] −→ c0;
f 7−→ T f = (cid:18)Z 1
0
f (t) sin ntdt(cid:19)n≥1
.
5
It is mentioned in [1], that T is not order bounded; it is perhaps more conve-
nient to consider the same operator defined on L1 [0, 2π] . In this case if we
define un by u (t) = sin nt for t ∈ [0, 2π] , then un ≤ 1, however (T un) = (en)
is not bounded in c0, where (en) denotes the standard basis of c0. This state-
ment implies also that T is not sequentially order compact. Because (en) has
no order bounded subsequence, it follows that (T un) can not admit an order
convergent subsequence. So the statement made in [3] that T is p-bounded is
not correct.
The above example is presented in [3] to show that sequentially p-compact
operators need not be p-bounded. Although the operator given in that ex-
ample fails to be sequentially p-compact, the assertion that sequentially p-
compact operators need not be p-bounded is true. This will be shown in our
next example.
Example 6 Consider the Riesz spaces E and F defined just before Lemma
4 and let T be the projection defined on E with range F and kernel R1.
We claim that T is sequentially order compact, but not order bounded. Let
(fn) be an order bounded sequence in E. Then fn ≤ λ for some real λ > 0
and for all n. Write fn = gn + λn with λn real and gn ∈ F and observe
that gn ≤ 2λ for all n. We have also gn ≤ 2λ1A ∈ F where A is the
union of the supports of gn, n = 1, 2, ... A standard diagonal process yields
a subsequence (gkn) of (gn) which converges pointwise on A and then on R
since all functions gnkvanish on R\A. Hence (gkn) is uo-convergent in RR.
As (gkn) is order bounded this implies that (gkn) is order convergent in RR.
gkn belongs to F, which shows that (gkn) is order
Observe moreover that sup
p≥n
convergent in F. The fact that T is not order bounded is more obvious: it is
clear that the image of the net (cid:0)1{x}(cid:1)x∈[0,1] by T is not order bounded in F.
As an immediate consequence of Theorem 2 and Example 6 we deduce
that sequentially p-compactness does not imply p-compactness. At this stage
one might expect that the converse is true. Does p-compactness imply se-
quentially p-compactness? This is an open question left in [3]. Unfortunately
the answer is again negative.
Example 7 Let X be the set of all strictly increasing maps from N to N and
E = RX be the space of all real-valued functions defined on X, equipped with
the product topology.
6
1. First we will prove that the identity map, I, is a p-compact operator on
the lattice-normed space (E, , E). To this aim, pick a p-bounded net
(fα)α∈A in E, that is, fα ≤ f for some f ∈ E+ and for every α ∈ A.
It follows that
fα ∈ Yx∈X
[−f (x), f (x)].
topology, is compact by Tychonoff 's Theorem. Thus (fα) has a con-
Notice that the space Qx∈X[−f (x), f (x)], equipped with the product
vergent subnet (gβ)β∈B in Qx∈X[−f (x), f (x)] to some g. This means
that
gβ(x) −→ g(x) for all x ∈ X.
According to Lemma 3, gβ is uo-convergent to g in E. Since bounded
o−→ g. This
uo-convergent nets are order convergent, we have that gβ
proves that I is a p-compact operator.
2. We prove now that I is not sequentially p-compact. Let (ϕn) be a
sequence in {−1, 1}X which has no convergent subsequence (see Ex-
ample 3.3.22 in [7]). This sequence is order bounded in E and every
subsequence (ψn) of (ϕn) does not converges in {−1, 1}X, that is, for
some x ∈ X, ψn (x) diverges. According to Lemma 3, (ψn) is not uo-
convergent in E. Since (ψn) is order bounded it does not converge in
order. This finishes the proof.
In classical theory of Banach spaces the identity map is compact if and
only if the space is finite-dimensional. In contrast of this the situation is not
clear in general case. We already have seen an example of infinite-dimensional
space on which the identity map is p-compact. This question has been in-
vestigated in [3] where the authors showed that IL1[0,1] fails to be compact
however, Iℓ1 is p-compact. In the next example we show that IL∞[0,1] is not
p-compact, answering a question asked in [3].
Example 8 The identity operator I on the lattice normed space (L∞[0, 1], . , L∞[0, 1])
is neither p-compact nor sequentially p-compact. To this end, consider the
sequence of Rademacher function given by :
rn
:
[0, 1] −→ R
t
7−→ sgn (sin(2nπt))
7
for all n ∈ N, which is order bounded since rn = 1. Suppose now that (rn) has
o→ r for some r ∈ L∞ [0, 1] .
an order convergent subnet (rnα)α∈Γ . Then rnα
0 rnαrnβ dµ = 0. On the other hand (cid:0)rnαrnβ(cid:1)β
Let α ∈ A. For every β > α, R 1
converges in order to rnαr in L∞ [0, 1] and then in L1 [0, 1] . Since the integral
is order continuous, we deduce that
Z 1
0
rnαrdµ = 0.
This equality holds for every α ∈ A, and a similar argument leads to
Z 1
0
r2dµ = 0,
which is a contradiction since r = 1, and the claim is now proved.
4 Semicompact operators
The notion of semicompact operators has been introduced by Zaanen in [9]
and extended in the framework of lattice normed spaces in [3].
Let (X, p, E) be a lattice normed space and (Y, q, F ) be an lattice normed
vector lattice. A linear operator T : X → Y is called p-semicompact if it
maps p-bounded sets in X to p-almost order bounded sets in Y . We recall
that a subset B of Y is said to be p-almost order bounded if for any w ∈ F+,
there is yw ∈ Y such that
q((y − yw)+) = q(y − yw ∧ y) ≤ w for all y ∈ B.
Semicompact operators from Banach spaces to Banach lattices fail, in gen-
eral, to be compact (see [1]). This yields trivially that p-semicompactness
does not imply p-compactness. However, the converse is true in the classical
case as has been shown in Theorem 5.71 in [1]. And one can expect to extend
this result in general situation. This is already the subject of Question 4 in
[3]. Unfortunately the answer is again negative. Before stating our coun-
terexample let us recall that every order bounded operator from a vector
lattice E to a Dedekind complete vector lattice F has a modulus ([1]).
Example 9 Let E be a Dedekind complete Banach lattice with order contin-
uous norm and T be a norm-compact operator in L(E) such that T has no
8
modulus, and therefore T can not be order bounded. For the existence of such
operator we refer the reader to the Krengel's example in [1, p 277.]. Consider
now the following lattice-normed vector spaces (E, k.k, R) and (E, p, R2) ,
where p(x) = (cid:18)kxk
0 (cid:19) for all x ∈ E. It is straightforward to prove that
T is again p-compact operator and we claim that T is not p-semicompact.
To this end we we will argue by contradiction and we assume that T is p-
semicompact. Fix an element u ∈ E+ and let w = (cid:18)0
1(cid:19), then there exists
zw such that p ((T (x) − zw)+) ≤ w for all x ∈ [−u, u], which means that
(T (x) − zw)+ = 0. Noting that this occurs for x and −x we see that
T (x) ≤ zw for all x ∈ [−u, u] .
This shows that T is order bounded, a contradiction. and our proof comes to
an end.
A slight modification of the proof of Example 9 leads to a more general
result. The proof of it will be left for the reader.
Proposition 10 Let (E, p, V ) be a lattice normed space and (F, q, W ) a lat-
tice normed vector lattice. We assume that q (F )d is not trivial. Then every
semicompact operator T : (E, p, V ) −→ (F, q, W ) is p-bounded as an operator
from (E, p, V ) to (F, . , F ) .
5
rp-compact operators
As every rp-compact operator between lattice-normed spaces is p-compact,
the following result is an immediate consequence of Theorem 2.
Theorem 11 Let (E, p, V ) and (F, q, W ) be lattice-normed spaces and T be
in L(E, F ). If T is rp-compact then T is p-bounded.
In the following example we will prove that sequentially p-compact oper-
ators need not be rp-compact.
Example 12 Let E be the Riesz space defined above. We claim that the
identity operator I : E → E is sequentially p-compact but fails to be rp-
compact. Let (xn) be a bounded sequence in E, that is, xn ≤ x for some
9
x ∈ E+. Write x = α + f, and xn = αn + fn where α ∈ R+ and f ∈ F and
αn ∈ R, fn ∈ F for n = 1, 2, ... It is easily seen that αn ≤ α, fn ≤ x+α. By
a standard diagonal argument there exists a subsequence such that fϕ(n) (a)
converges for every a ∈ R and αϕ(n) converges in R.T his shows that xn
uo−→ x
converges pointwise on R and its limit is clearly in E. By Lemma 3, xn
o−→ x in RR as it is an order bounded sequence. Now using
in RR and then xn
Lemma 27 in [2] and Lemma 4 we deduce that x ∈ E. On the other hand, let
F be the collection of finite subsets of R+ ordered by inclusion and consider
the net (gA)A∈F where gα = 1α. Then (gα) is order bounded in E but has no
convergent subnet. Since gα ↑ 1R+ in RR and E is regular in RR, it follows
that (gα) is not order convergent in E and so are all its subnets.
References
[1] C.D. Aliprantis and O. Burkinshaw, Positive Operators, Springer, 2006.
[2] Y. Azouzi, Completeness for vector lattices, J. Math. Anal. Appl. 472
(2019) 216 -- 230
[3] A. Aydın, E.Yu. Emelyanov, N. Erkur¸sun Ozcand, M.A.A. Marabeh,
Indag. Math. (N.S.) 29 (2018) 633 -- 656.
[4] N. Gao, V.G. Troitsky, F. Xanthos, Uo-convergence and its applications
to Ces´aro means in Banach lattices, Isr. J. Math. 220 (2) (2017) 649 -- 689.
[5] J. L. Kelley, General topology. Courier Dover Publications, 2017.
[6] A.G. Kusraev, Dominated Operators, in: Mathematics and its Applica-
tions, vol. 519, Kluwer Academic Publishers, Dordrecht, 2000.
[7] V. Runde, A Taste of Topology. Springer, Berlin (2005)
[8] B.Z. Vulikh, Introduction To the Theory of Partially Ordered Spaces,
Wolters-Noordhoff Scientifc Publications, Ltd., Groningen, 1967.
[9] A. C. Zaanen, Riesz spaces II, North-Holland, Amsterdam, 1983.
10
|
1712.08450 | 1 | 1712 | 2017-12-22T14:04:24 | On the weighted fractional Poincare-type inequalities | [
"math.FA"
] | Weighted fractional Poincar\'e-type inequalities are proved on John domains whenever the weights defined on the domain are depending on the distance to the boundary and to an arbitrary compact set in the boundary of the domain. | math.FA | math |
ON THE WEIGHTED FRACTIONAL POINCAR´E-TYPE
INEQUALITIES
RITVA HURRI-SYRJ ANEN AND FERNANDO L ´OPEZ-GARC´IA
Abstract. Weighted fractional Poincar´e-type inequalities are proved on John do-
mains whenever the weights defined on the domain depend on the distance to the
boundary and to an arbitrary compact set in the boundary of the domain.
1. Introduction
In this article we study a version of the classical fractional Poincar´e-type inequality
where the domain in the double integral in the Gagliardo seminorm is replaced by a
smaller one:(cid:18)(cid:90)
(cid:19)1/p ≤ C
(cid:18)(cid:90)
(cid:90)
u(x) − uΩpdx
Ω
Ω
B(x,τ d(x))
(cid:19)1/p
u(x) − u(y)p
x − yn+sp dydx
.
(1.1)
The parameter τ in the double integral belongs to (0, 1) and d(x) denotes the distance
from x to ∂Ω. The inequality (1.1) was introduced in [4]. It is well-known that the
fractional classical Poincar´e inequality is valid for any bounded domain, while this new
version (1.1) depends on the geometry of the domain. In [4] it was proved that the
inequality (1.1) is valid on John domains and, hence, in particular on Lipschitz domains.
An example of a domain where the inequality (1.1) is not valid was also given. We
refer the reader to [5] and [2] where the fractional Sobolev-Poincar´e versions of (1.1)
are considered. For a weighted version of (1.1) where weights are power functions to
the boundary we refer to [3].
The main result of our paper is the following theorem where the distance to an
arbitrary set of the boundary has been added as a weight.
Theorem 1.1. Let Ω in Rn be a bounded John domain and 1 < p < ∞. Given
a compact set F in ∂Ω, and the parameters β ≥ 0 and s, τ ∈ (0, 1), there exists a
constant C such that
(cid:18)(cid:90)
Ω
u(x) − uΩ,ωpdpβ
F (x)dx
(cid:19)1/p
(cid:18)(cid:90)
(cid:90)
(cid:19)1/p
(1.2)
for all functions u ∈ Lp(Ω, d(x)pβ), where d(x) and dF (x) denote the distance from x
to ∂Ω and F respectively, and uΩ,ω is the weighted average
u(x) − u(y)p
x − yn+sp dps(x)dpβ
(cid:82)
F (x)dydx
≤ C
Ω u(z)dpβ
F (z)dz.
Ω
B(x,τ d(x))
1
dpβ
F (Ω)
Date: May 24, 2021.
2010 Mathematics Subject Classification. Primary: 46E35 ; Secondary: 26D10.
Key words and phrases. Fractional Poincar´e inequalities, Hardy-type operator, Tree covering,
Weights.
1
2
In addition, the constant C in (1.2) can be written as
C = Cn,p,β τ s−nK n+β,
where K is the geometric constant introduced in (5.1).
We would like to emphasize two points in this result: The first one is that no extra
conditions are required for the compact set F in ∂Ω. The second point is that the
estimate shows how the constant depends on the given τ and a certain geometric
condition of the domain.
Some of the essential auxiliary parts for the proofs for weighted inequalities are
from [7] and [8] where a useful decomposition technique was introduced by the second
author. Our work was stimulated by the papers of Augusto C. Ponce, [10], [11], [12],
where more general fractional Poincar´e inequalities for functions defined on Lipschitz
domains were investigated.
The paper is organized as follows: In Section 2, we introduce some definitions and
preliminary results. In Section 3, we show how to use decompositions of functions to
extend the validity of certain inequalities on "simple domains", such as cubes, to more
complex ones. We are interested in extending the results from cubes to John domains.
In Section 4, we apply the results obtained in the previous section to estimate the
constant in the unweighted version of (1.2) on cubes. Especially we are interested in
how the constant depends on τ . This result is auxiliary of our main theorem but it
might be of independent interest. In Section 5, we show the validity of the weighted
fractional Poincare inequality studied in this paper with the estimate of the constant
and a generalization to the type of inequalities considered by Ponce.
2. Notation and preliminary results
Throughout the paper Ω in Rn is a bounded domain with n ≥ 2, 1 < p < ∞, and
1 < q < ∞ with 1
q = 1, unless otherwise stated. Moreover, given η : Ω → R a
weight (i.e., a positive measurable function) and 1 ≤ r ≤ ∞, we denote by Lr(Ω, η)
the space of Lebesgue measurable functions u : Ω → R equipped with the norm
p + 1
(cid:18)(cid:90)
(cid:19)1/r
(cid:107)u(cid:107)Lr(Ω,η) :=
u(x)rη(x) dx
if 1 ≤ r < ∞, and
Ω
(cid:107)u(cid:107)L∞(Ω,η) := ess sup
x∈Ω
u(x)η(x).
Finally, given a set A we denote by χA(x) its characteristic function.
Definition 2.1. Let C be the space of constant functions from Rn to R and {Ut}t∈Γ
a collection of open subsets of Ω that covers Ω except for a set of Lebesgue measure
zero; Γ is an index set. It also satisfies the additional requirement that for each t ∈ Γ
the set Ut intersects a finite number of Us with s ∈ Γ. This collection {Ut}t∈Γ is called
we say that a collection of functions {gt}t∈Γ in L1(Ω) is a C-orthogonal decomposition
of g subordinate to {Ut}t∈Γ if the following three properties are satisfied:
an open covering of Ω. Given g ∈ L1(Ω) orthogonal to C (i.e.,(cid:82) g ϕ = 0 for all ϕ ∈ C),
(1) g =(cid:80)
(3) (cid:82)
gt ϕ = 0, for all ϕ ∈ C and t ∈ Γ.
t∈Γ gt.
(2) supp(gt) ⊂ Ut.
Ut
We also refer to this collection of functions by a C-decomposition. We say that {gt}t∈Γ
is a finite C-decomposition if gt (cid:54)≡ 0 only for a finite number of t ∈ Γ.
distance to the space of constant functions C by replacing its left hand side by
Inequality (1.2), and similar Poincar´e type inequalities, can be written in terms of a
3
(cid:18)(cid:90)
inf
α∈C
Ω
(cid:19)1/p
.
u(x) − αpdpβ
F (x)dx
The technique used in this paper may also be considered when the distance to other
vector spaces V are involved, in which case, a V-orthogonal decomposition of func-
tions is required. We direct the reader to [9] where a generalized version of the Korn
inequality is studied by using decomposition of functions.
Let us denote by G = (V, E) a graph with vertices V and edges E. Graphs in this
paper have neither multiple edges nor loops and the number of vertices in V is at most
countable.
A rooted tree (or simply a tree) is a connected graph G in which any two vertices
are connected by exactly one simple path, and a root is simply a distinguished vertex
a ∈ V . Moreover, if G = (V, E) is a rooted tree with a root a, it is possible to define
a partial order "(cid:22)" in V as follows: s (cid:22) t if and only if the unique path connecting t
with the root a passes through s. The height or level of any t ∈ V is the number of
vertices in {s ∈ V : s (cid:22) t with s (cid:54)= t}. The parent of a vertex t ∈ V is the vertex s
satisfying that s (cid:22) t and its height is one unit smaller than the height of t. We denote
the parent of t by tp. It can be seen that each t ∈ V different from the root has a
unique parent, but several elements in V could have the same parent. Note that two
vertices are connected by an edge (adjacent vertices) if one is the parent of the other.
Definition 2.2. Let Ω be in Rn be a bounded domain. We say that an open covering
{Ut}t∈Γ is a tree covering of Ω if it also satisfies the properties:
t∈Γ χUt(x) ≤ N χΩ(x), for almost every x ∈ Ω, where N ≥ 1.
(1) χΩ(x) ≤(cid:80)
(2) Γ is the set of vertices of a rooted tree (Γ, E) with a root a.
(3) There is a collection {Bt}t(cid:54)=a of pairwise disjoint open cubes with Bt ⊆ Ut∩ Utp.
Definition 2.3. Given a tree covering {Ut}t∈Γ of Ω we define the following Hardy-type
operator T on L1-functions:
g,
(2.1)
(2.2)
T g(x) :=
(cid:88)
a(cid:54)=t∈Γ
(cid:90)
Wt
χt(x)
Wt
(cid:91)
s(cid:23)t
Us ,
where
Wt :=
and χt is the characteristic function of Bt for all t (cid:54)= a.
We may refer to Wt by the shadow of Ut.
Note that the definition of T is based on the a-priori choice of a tree covering {Ut}t∈Γ
of Ω. Thus, whenever T is mentioned in this paper there is a tree covering {Ut}t∈Γ of
Ω explicitly or implicitly associated to it.
The following fundamental result was proved in [8, Theorem 4.4], which shows the
existence of a C−decomposition of functions subordinate to a tree covering of the
domain.
g ∈ L1(Ω) such that(cid:82)
Theorem 2.4. Let Ω in Rn be a bounded domain with a tree covering {Ut}t∈Γ. Given
Ω gϕ = 0, for all ϕ ∈ C, and supp(g)∩ Us (cid:54)= ∅ for a finite number
of s ∈ Γ, there exists a C-decompositions {gt}t∈Γ of g subordinate to {Ut}t∈Γ (refer to
Definition 2.1).
Moreover, let t ∈ Γ. If x ∈ Bs where s = t or sp = t then
4
gt(x) ≤ g(x) +
Ws
Bs T g(x),
(2.3)
where Wt denotes the shadow of Ut defined in (2.2). Otherwise
(2.4)
Remark 2.5. The C-decomposition stated in Theorem 2.4 is finite. This fact is not in
the statement of [8, Theorem 4.4] but it is easily deduced from its proof.
gt(x) ≤ g(x).
In the next lemma, the continuity of the operator T is shown. We refer the reader
to [7, Lemma 3.1] for its proof.
Lemma 2.6. The operator T : Lq(Ω) → Lq(Ω) defined in (2.1) is continuous for any
1 < q ≤ ∞. Moreover, its norm is bounded by
(cid:19)1/q
(cid:18) qN
q − 1
.
(cid:107)T(cid:107)Lq→Lq ≤ 2
Here N is the overlapping constant from Definition 2.2.
If q = ∞, the previous inequality means (cid:107)T(cid:107)L∞→L∞ ≤ 2. Actually, for being T
an averaging operator, it can be easily observed that (cid:107)T(cid:107)L∞→L∞ = 1, but it does not
affect our work. Notice that Lq(Ω, ω−q) ⊂ L1(Ω) if the weight ω : Ω → R>0 satisfies
that ωp ∈ L1(Ω). Then, the operator T introduced in Definition 2.3 for functions in
L1(Ω) is well-defined in Lq(Ω, ω−q).
Lemma 2.7. Let Ω in Rn be a bounded domain, {Ut}t∈Γ a tree covering of Ω and
ω : Ω → R a weight which satisfies ωp ∈ L1(Ω). If ω satisfies that
(2.5)
for all a (cid:54)= t ∈ Γ, then the Hardy-type operator T defined in (2.1) and subordinate to
{Ut}t∈Γ is continuous from Lq(Ω, ω−q) to itself. Moreover, its norm for 1 < q < ∞ is
bounded by
ess sup
y∈Wt
ω(x),
ω(y) ≤ C2 ess inf
x∈Bt
(cid:19)1/q
(cid:18) qN
q − 1
C2,
(cid:107)T(cid:107)L→L ≤ 2
where L denotes Lq(Ω, ω−q), and N is the overlapping constant from Definition 2.2.
Proof. Given g ∈ Lq(Ω, ω−q) we have
(cid:90)
(cid:90)
(cid:90)
Ω
Ω
Ω
=
=
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:88)
T g(x)qω−q(x) dx
ω−q(x)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:88)
a(cid:54)=t∈Γ
a(cid:54)=t∈Γ
χt(x)
Wt
Wt ω−1(x)
χt(x)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q
dx
g(y) dy
g(y)ω−1(y) ω(y) dy
(cid:90)
(cid:90)
Wt
Wt
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q
dx.
Now, condition (2.5) implies that ω(y) ≤ C2ω(x) for almost every x ∈ Bt and y ∈ Wt.
5
Thus,
≤
(cid:90)
(cid:90)
Ω
Ω
= C q
2
= C q
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q
dx
g(y)ω−1(y) dy
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q
g(y)ω−1(y) dy
dx
χt(x)
a(cid:54)=t∈Γ
T g(x)qω−q(x) dx
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:88)
Wt ω−1(x) C2 ω(x)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:88)
(cid:90)
(cid:90)
(cid:90)
(cid:12)(cid:12)T (gω−1)(cid:12)(cid:12)q dx.
(cid:18)
χt(x)
Wt
a(cid:54)=t∈Γ
Wt
Ω
Ω
(cid:90)
Wt
(cid:19)
Finally, gω−1 belongs to Lq(Ω) and T is continuous from Lq(Ω) to itself; we refer to
Lemma 2.6, hence(cid:90)
T g(x)qω−q(x) dx ≤
2q qN
q − 1
Ω
2 (cid:107)g(cid:107)q
C q
Lq(Ω,ω−q).
(cid:3)
(cid:19)1/p
3. A decomposition and Fractional Poincar´e inequalities
(cid:82)
Let Ω in Rn be an arbitrary bounded domain and {Ut}t∈Γ an open covering of Ω.
The weight ω : Ω → R>0 satisfies that ωp ∈ L1(Ω). In addition, uΩ denotes the average
1Ω
Ω u(z)dz. For weighted spaces of functions, uΩ,ω represents the weighted average
Now, given a bounded domain U in Rn and a nonnegative measurable function
µ : U × U → R we introduce the following Poincar´e type inequality
1
ω(Ω)
(cid:82)
Ω u(z)ω(z)dz, where ω(Ω) :=(cid:82)
(cid:18)(cid:90)
c∈R(cid:107)u − c(cid:107)Lp(U,ωp) ≤ C
inf
(cid:90)
Ω ω(z)dz.
u(x) − u(y)pµ(x, y) dydx
(3.6)
where u ∈ Lp(U, ωp). Notice that the right hand side in this inequality might be
infinite. The validity of (3.6) depends on U , p, µ and ω. The function µ(x, y) might
be zero, however, ω(x) is strictly positive almost everywhere in Ω.
U
U
,
Let us mention three examples.
Examples 3.1.
(1) The weighted fractional Poincar´e inequality with µ(x, y) =
x−yn+sp , where s ∈ (0, 1) , is the classical fractional Poincar´e inequality which
1
is clearly valid for any arbitrary bounded domain.
(2) If µ(x, y) = χBx (y)
x−yn+sp , where Bx is the ball centered at x with radius τ d(x) for
s, τ ∈ (0, 1), then the inequality represents a more recently studied fractional
Poincar´e inequality whose validity depends on the geometry of the domain (re-
fer to [4] for details).
(3) Finally, µ(x, y) = ρ(x−y)
x−yp , where ρ is a certain nonnegative radial function, is
another inequality which has also been studied recently (refer to [10] for details).
6
Inequality (3.6) deals with an estimation of the distance to C of an arbitrary function
u in Lp(Ω, ωp). The local-to-global argument used in this paper to study this Poincar´e
type inequalities is based on the fact that Lp(Ω, ωp) is the dual space of Lq(Ω, ω−q)
and the existence of decompositions of functions in Lq(Ω, ω−q) orthogonal to C. Let us
properly define this set and a subspace:
W := {g ∈ Lq(Ω, ω−q) :
W0 := {g ∈ W : supp(g) intersects a finite number of Ut}.
(3.8)
The integrability of ωp implies that Lq(Ω, ω−q) ⊂ L1(Ω), then W and W0 are well-
defined. Following Remark 2.5, the C-decomposition of functions in W0 stated in
Theorem 2.4 is finite, which is not valid in general for functions in W. This property
verified by the functions in W0 simplifies the proof of Lemma 3.3, which motivates the
definition of this space.
gϕ = 0 for all ϕ ∈ C}
(cid:90)
(3.7)
Now, we introduce the spaces
W ⊕ ωpC = {g + αωp / g ∈ W and α ∈ C}
S := W0 ⊕ ωpC = {g + αωp / g ∈ W0 and α ∈ C}.
(3.9)
It is not difficult to observe that Lq(Ω, ω−q) = W ⊕ ωpC and S is a subspace of
Lq(Ω, ω−q). The following lemma, which was proved in [8, Lemma 3.1], states that S
is also dense in Lq(Ω, ω−q) and uses in its proof the requirement that says that for each
t ∈ Γ the set Ut intersects a finite number of Us with s ∈ Γ.
Lemma 3.2. The space S is dense in Lq(Ω, ω−q). Moreover, if g + αωp is an element
in S then
(cid:107)g(cid:107)Lq(Ω,ω−q) ≤ 2(cid:107)g + αωp(cid:107)Lq(Ω,ω−q).
Lemma 3.3. If there exists an open covering {Ut}t∈Γ of Ω such that (3.6) is valid on
Ut for all t ∈ Γ, with a uniform constant C1, and there exists a finite C-orthogonal
decomposition of any function g in W0 subordinate to {Ut}t∈Γ, with the estimate
0(cid:107)g(cid:107)q
Lq(Ω,ω−q),
(cid:88)
t∈Γ
(cid:107)gt(cid:107)q
(cid:32)(cid:88)
Lq(Ut,ω−q) ≤ C q
(cid:90)
(cid:90)
t∈Γ
Ut
Ut
then, there exists a constant C such that
(cid:107)u − uΩ,ω(cid:107)Lp(Ω,ωp) ≤ C
u(x) − u(y)pµ(x, y) dydx
(cid:33)1/p
(3.10)
is valid for any u ∈ Lp(Ω, ωp). Moreover, the constant C = 2C0C1 holds in (3.10).
Proof. Without loss of generality we can assume that uΩ,ω = 0. We estimate the
norm on the left hand side of the inequality by duality. Thus, let g + ωpψ be an
arbitrary function in S, we refer to Lemma 3.2. Then, by using the finite C-orthogonal
decomposition of g we conclude that
(cid:90)
Ω
u(g + αωp) =
=
ug =
u
Ω
ugt =
gt
(cid:90)
(u − ct)gt.
(3.11)
(cid:90)
(cid:90)
(cid:88)
Ω
t∈Γ
(cid:90)
(cid:88)
(cid:88)
t∈Γ
t∈Γ
Notice that the identity in the second line is valid for any t ∈ Γ and ct ∈ R.
Ut
Ut
Next, by using the Holder inequality in (3.11), the fact that (3.6) is valid on Ut with
a uniform constant C1 and, finally, the Holder inequality over the sum, we obtain
7
(cid:90)
Ω
≤ C1
≤ C1
Ut
Ut
t∈Γ
u(g + αωp) ≤(cid:88)
(cid:18)(cid:90)
(cid:90)
(cid:88)
(cid:32)(cid:88)
(cid:90)
(cid:90)
(cid:32)(cid:88)
(cid:90)
(cid:32)(cid:88)
(cid:90)
(cid:90)
(cid:90)
t∈Γ
t∈Γ
Ut
Ut
Ut
≤ C0C1
≤ 2C0C1
t∈Γ
Ut
Ut
inf
c∈R(cid:107)u − c(cid:107)Lp(Ut,ωp)(cid:107)gt(cid:107)Lq(Ut,ω−q)
(cid:19)1/p (cid:107)gt(cid:107)Lq(Ut,ω−q)
t∈Γ
u(x) − u(y)pµ(x, y) dydx
(cid:33)1/p(cid:32)(cid:88)
(cid:33)1/p
(cid:33)1/p
u(x) − u(y)pµ(x, y) dydx
u(x) − u(y)pµ(x, y) dydx
t∈Γ
(cid:107)g(cid:107)Lq(U,ω−q)
(cid:107)gt(cid:107)q
Ut
(cid:33)1/q
Lq(Ut,ω−q)
u(x) − u(y)pµ(x, y) dydx
(cid:107)g + αωp(cid:107)Lq(U,ω−q).
Finally, as S is dense in Lq(Ω, ω−q), by taking the supremum over all the functions
g + αωp in S with (cid:107)g + αωp(cid:107)Lq(Ω,ω−q) ≤ 1 we prove the result.
(cid:3)
4. On fractional Poincar´e inequalities on cubes
In this section, we use the results stated in the previous two sections to show a
certain fractional Poincar´e inequality on an arbitrary cube Q. Thus, in order to show
the existence of the C-decomposition, which is used later to apply Lemma 3.3, we define
a tree covering {Ut}t∈Γ of Q. This covering is only used in this section and for cubes. In
the following section, we work with a different bounded domain, an arbitrary bounded
John domain, which requires a different covering. However, let us warn the reader that
we will keep the notation {Ut}t∈Γ used in Section 3.
The validity of the local inequality stated in the following proposition is well-known.
We refer the reader to [3] for its proof.
Proposition 4.1. The fractional Poincar´e inequality
c∈R(cid:107)u(x) − c(cid:107)Lp(U ) ≤
inf
(cid:18)diam (U )n+sp
(cid:90)
(cid:90)
U
u(y) − u(x)p
y − xn+sp dydx
U
U
(cid:19)1/p
holds for any bounded domain U in Rn and 1 ≤ p < ∞.
The following proposition is a special case of [4, Lemma 2.2]. In the present paper,
we give a different proof which let us estimate the dependance of the constant with
respect to τ .
Proposition 4.2. Let Q in Rn be a cube with side length l(Q) = L, 1 < p < ∞ and
τ ∈ (0, 1). Then, the following inequality holds
(cid:18)(cid:90)
(cid:90)
u(y) − u(x)p
y − xn+sp dydx
Q∩B(x,τ L)
Q
(cid:19)1/p
,
c∈R(cid:107)u(x) − c(cid:107)Lp(Q) ≤ Cn,p τ s−nLs
inf
where Cn,p depends only on n and p.
Proof. This result follows from Lemma 3.3 on the cube Q, where µ(x, y) =
x−yn+sp
and ω ≡ 1. So, let us start by defining an appropriate tree covering of Q to obtain,
via Theorem 2.4 and Remark 2.5, a finite C-decomposition of any functions in W0. Let
and {At}t∈Γ the regular partition of Q with
m ∈ N be such that
mn open cubes. The side length of each cube is l(At) = L
m. In the example shown in
Figure 1, m = 4 and the index set Γ has 16 elements.
τ < m ≤ 1 +
n+3
τ
√
√
n+3
1
8
Figure 1. A tree covering of Q
The tree covering of Q that we are looking for will be defined by enlarging the sets in
the covering {At}t∈Γ in an appropriate way but keeping the tree structure of Γ, which
is introduced in the following lines. Indeed, we pick a cube Aa, whose index will be the
root, and inductively define a tree structure in Γ such that the unique chain connecting
t with a is associated to a chain of cubes connecting Qt with Qa, with minimal number
of cubes, such that two consecutive cubes share a n − 1 dimensional face. In Figure 1,
the cube Aa is in the lower left corner and the tree structure is represented using black
arrows that"descend" to the root. Now that Γ has a tree structure, we define the tree
covering {Ut}t∈Γ of Q with the rectangles Ut := (At ∪ Atp)◦ if t (cid:54)= a and Ua := Aa. In
order to have a better understanding of the construction, notice that Ut ∩ Utp = Atp
for all t (cid:54)= a. Moreover, the index set Γ in the example with its tree structure has 7
levels, from level 0 to level 6 (refer to page 3 for definitions), with only one index of
level 6, whose rectangle Ut appears in Figure 1 in a different color.
Now, let us define the collection {Bt}t(cid:54)=a of pairwise disjoint open cubes Bt ⊆ Ut∩Utp
or equivalently Bt ⊆ Atp. Given t (cid:54)= a, we split Atp into 3n cubes with the same size.
The open set Bt is the cube in the regular partition of Atp whose closure intersects the
n − 1 dimensional face Atp in the intersection (At ∩ Atp). There are 3n−1 cubes with
that property but we pick Bt to be the one which does not share any part of any other
n − 1 dimensional face of Atp.
The cubes in {Bt}t(cid:54)=a have side length equal to L
3m and are represented in Figure
1 by the 15 grey gradient small cubes. By its construction, it is easy to check that
{Bt}t(cid:54)=a is a collection of pairwise disjoint open cubes Bt ⊆ Ut ∩ Utp, hence, {Ut}t∈Γ is
a tree covering of Q with N = 2n (it could also be less).
By Theorem 2.4, there is a finite C-decomposition of functions {gt}t∈Γ subordinate
to {Ut}t∈Γ which satisfies (2.3) and (2.4). Moreover, it can be seen that
Bs ≤ Q
Ws
Bs = (3m)n,
for all s ∈ Γ, thus,
for all t ∈ Γ and x ∈ Ut. Next, using the continuity of T stated in Lemma 2.6 and
some straightforward calculations we conclude
gt(x) ≤ g(x) + (3m)nT g(x),
9
(cid:18)
(cid:19)
(cid:107)gt(cid:107)q
Lq(Ut) ≤ 2q−1N
1 + (3m)nq2q qN
q − 1
(cid:107)g(cid:107)q
Lq(Q)
(cid:88)
t∈Γ
Hence, we have a finite C-decomposition of any function in W0 subordinate to {Ut}t∈Γ
with the constant in the estimate equal to
≤ 22q+2n2q
q − 1
≤ 22q+23nqn2q
q − 1
(cid:18)22q+23nqn2q
q − 1
(3m)nq (cid:107)g(cid:107)q
√
1 +
(cid:16)
(cid:19)1/q (cid:16)
1 +
√
C0 =
(cid:17)nq
τ−nq (cid:107)g(cid:107)q
Lq(Q).
Lq(Q)
n + 3
(cid:17)n
√
n + 3
τ−n.
n+3
Now, from Proposition 4.1 and using that m >
τ
(3.6) is valid on each Ut with an uniform constant
, we can conclude that inequality
Thus, using Lemma 3.3 we can claim that
(cid:107)u − uQ(cid:107)Lp(Q) ≤ 2C0C1
(cid:33)1/p
.
u(x) − u(y)p
x − yn+sp dydx
Finally, diam(Ut) ≤ √
the control on the overlapping of the tree covering given by N = 2n, it follows that
n + 3 L
m ≤ τ L, thus Ut ⊂ B(x, τ L) for any x ∈ Ut, thus, using
C1 = (n + 3)n/2p(τ L)s.
(cid:90)
(cid:90)
Ut
Ut
(cid:90)
t∈Γ
(cid:32)(cid:88)
(cid:18)(cid:90)
(cid:19)1/q (cid:16)
Q∩B(x,τ L)
Q
√
n + 3
1 +
(cid:17)n
(cid:19)1/p
u(x) − u(y)p
x − yn+sp dydx
(n + 3)n/2p(2n)1/p.
,
(4.1)
(cid:3)
(cid:107)u − uQ(cid:107)Lp(Q) ≤ Cn,p τ−n(τ L)s
where
Cn,p = 2
(cid:18)22q+23nqn2q
q − 1
5. On fractional Poincar´e inequalities on John domains
In this section, we apply the results obtained in the previous sections on an arbitrary
bounded John domain Ω. Its definition is recalled below. The weight ω(x) is defined
as dF (x)β, where dF (x) denotes the distance from x to an arbitrary compact set F in
∂Ω and β ≥ 0. In the particular case where F = ∂Ω, d∂Ω(x) is simply denoted as d(x).
Notice that ωp belongs to L1(Ω) for being Ω bounded and β nonnegative.
A Whitney decomposition of Ω is a collection {Qt}t∈Γ of closed pairwise disjoint
dyadic cubes, which verifies
(1) Ω =(cid:83)
t∈Γ Qt.
(2) diam(Qt) ≤ dist(Qt, ∂Ω) ≤ 4diam(Qt).
(3) 1
4diam(Qs) ≤ diam(Qt) ≤ 4diam(Qs), if Qs ∩ Qt (cid:54)= ∅.
10
Here, dist(Qt, ∂Ω) is the Euclidean distance between Qt and the boundary of Ω, denoted
by ∂Ω. The diameter of the cube Qt is denoted by diam(Qt) and the side length is
written as (cid:96)(Qt).
Two different cubes Qs and Qt with Qs ∩ Qt (cid:54)= ∅ are called neighbors. This kind of
covering exists for any proper open set in Rn (refer to [13, VI 1] for details). Moreover,
each cube Qt has less than or equal to 12n neighbors. And, if we fix 0 < < 1
4 and
define (1 + )Qt as the cube with the same center as Qt and side length (1 + ) times
the side length of Qt, then (1 + )Qt touches (1 + )Qs if and only if Qt and Qs are
neighbors.
Given a Whitney decomposition {Qt}t∈Γ of Ω we refer by an expanded Whitney
decomposition of Ω to the collection of open cubes {Q∗
t}t∈Γ defined by
Observe that this collection of cubes satisfies that
Q∗
t :=
Q◦
t .
9
8
χΩ(x) ≤ 12n(cid:88)
t∈Γ
(x) ≤ (12n)2χΩ(x)
χQ∗
t
for all x ∈ Rn.
We recall the definition of a bounded John domain. A bounded domain Ω in Rn is a
John domain with constants a and b, 0 < a ≤ b < ∞, if there is a point x0 in Ω such
that for each point x in Ω there exists a rectifiable curve γx in Ω, parametrized by its
arc length written as length(γx), such that
dist(γx(t), ∂Ω) ≥
a
length(γx)
for all t ∈ [0, length(γx)]
t
and
length(γx) ≤ b.
Examples of John domains are convex domains, uniform domains, and also domains
with slits, for example B2(0, 1)\[0, 1). The John property fails in domains with zero
angle outward spikes. John domains were introduced by Fritz John in [6] and they
were renamed by O. Martio and J. Sarvas as John domains later.
There are other equivalent definitions of John domains.
In these notes, we are
interested in a definition of the style of Boman chain condition (see [1]) in terms of
Whitney decompositions and trees. This equivalent definition is introduced in [8].
Definition 5.1. A bounded domain Ω in Rn is a John domain if for any Whitney
decomposition {Qt}t∈Γ, there exists a constant K > 1 and a tree structure of Γ, with
a root a, that satisfies
Qs ⊆ KQt,
(5.1)
for any s, t ∈ Γ with s (cid:23) t. In other words, the shadow of Qt written as Wt is contained
in KQt; refer to (2.2). Moreover, the intersection of the cubes associated to adjacent
indices, Qt and Qtp, is an n − 1 dimensional face of one of these cubes.
Now, given a Whitney decomposition {Qt}t∈Γ of a bounded John domain Ω in Rn,
with constant K in the sense of (5.1), we define the tree covering {Ut}t∈Γ of expanded
Whitney cubes such that
Ut := Q∗
t .
(5.2)
11
The overlapping is bounded by N = 12n. Now, each open cube Bt in the collection
{Bt}t(cid:54)=a shares the center with the n− 1 dimensional face Qt ∩ Qtp and has side length
lt
64, where lt is the side length of Qt. It follows from the third condition in the Whitney
decomposition, and some calculations, that this collection is pairwise disjoint and
Moreover, it can be seen that
Bt ⊂ Q∗
t ∩ Q∗
tp = Ut ∩ Utp.
Wt
Bt ≤ (K 9
8lt)n
( lt
64)n
= 72n K n,
(5.3)
for all t ∈ Γ, with t (cid:54)= a.
Lemma 5.2. Let Ω in Rn be a John domain with the constant K in the sense of (5.1),
F in ∂Ω a compact set and dF (x) the distance from x to F . Then,
dF (y) ≤ 3K
sup
y∈Wt
for all t ∈ Γ.
√
n inf
x∈Bt
dF (x),
A similar inequality is also valid if we consider the weight dβ
F (x) with a nonnegative
power of the distance to F . Thus, this lemma implies, via Lemma 2.7, the continuity
−qβ
of the operator T from Lq(Ω, d
) to itself with an estimation of its constant. Then,
there exists a C-decomposition with a weighted estimate for a certain weight.
F
Proof. Given t ∈ Γ, with t (cid:54)= a, x ∈ Bt and y ∈ Wt := ∪s(cid:23)tUs, we have to prove that
dF (y) ≤ 3KdF (x). Notice that d(x) ≤ dF (x) for all x ∈ Ω. Moreover, Qs ⊆ KQt for
all s (cid:23) t, then Wt ⊆ KUt. In addition,
dF (y) ≤ y − x + dF (x) ≤ diam(Wt) + dF (x)
≤ Kdiam(Ut) + dF (x)
= K 9
8 diam(Qt) + dF (x).
Finally, using the second property stated in the Whitney decomposition it follows that
3Qt ⊂ Ω. Then, as
dist(Q∗
t , ∂Ω) ≥ dist(Q∗
t , (3Qt)c) ≥ 15
16 lt,
doing some calculations we can assert that
n dist(Q∗
diam(Qt) ≤ 16
√
15
t , ∂Ω) ≤ 16
9
√
n dist(Q∗
t , ∂Ω).
Thus,
dF (y) ≤ 2K
≤ 2K
√
√
n dist(Q∗
t , ∂Ω) + dF (x)
√
n d(x) + dF (x) ≤ 2K
n dF (x) + dF (x).
(cid:3)
Now we are able to prove Theorem 1.1 and also to give the dependence of the constant
C on the given value of τ and the constant K from (5.1).
Proof of Theorem 1.1. This result follows from Lemma 3.3 with the tree covering {Ut}t∈Γ
of Ω defined in (5.2), ω(x) := dβ
F (x) and
12
µ(x, y) :=
dps(x) dpβ
F (x) χB(x,τ d(x))(y)
x − yn+sp
.
(5.4)
Notice that ωp belongs to L1(Ω), the condition assumed at the beginning of Section 3.
The validity of (3.6) on a cube Ut, with a uniform constant C1, follows from Proposition
4.2. Indeed, by using the fact that Ut is an expanded Whitney cube by a factor 9/8
and F ⊆ ∂Ω, it follows that
F (x) ≤ 2β inf
dβ
x∈Ut
sup
x∈Ut
dβ
F (x).
Thus, we have
c∈R(cid:107)u(x) − c(cid:107)Lp(Ut,dpβ
inf
F )
≤Cn,p τ s−nLs
t 2β
(cid:18)(cid:90)
(cid:90)
Ut
Ut
u(x) − u(y)p
x − yn+sp dpβ
F (x)χB(x,τ Lt)(y) dydx
(cid:19)1/p
,
where Lt is the side length of Ut and Cn,p is the constant in (4.1). Now, observe that
Lt ≤ d(x) for all x ∈ Ut. Indeed, if x ∈ Qt then
√
Lt = 9
8 lt <
n lt = diam(Qt) ≤ dist(Qt, ∂Ω) ≤ d(x),
where lt is the side length of Qt. Now, if x ∈ Ut \ Qt then
n lt ≤ dist(Qt, ∂Ω) ≤ dist(Ut, ∂Ω) + 1
√
√
n lt,
16
√
hence, 15
16
n lt ≤ dist(Ut, ∂Ω) and
8 lt < 15
Lt = 9
16
√
n lt ≤ dist(Ut, ∂Ω) ≤ d(x).
Then, the validity of Lt ≤ d(x) for all x ∈ Ut implies (3.6) for all Ut, where µ(x, y)
is the function defined in (5.4), and the uniform constant
(5.5)
Next, by Theorem 2.4, there is a finite C-decomposition of functions {gt}t∈Γ subor-
dinate to {Ut}t∈Γ of any function g in W0 which satisfies (2.3) and (2.4). Moreover,
using (5.3), it can be seen that
C1 = Cn,pτ s−n2β.
gt(x) ≤ g(x) + (72K)nT g(x),
for all t ∈ Γ and x ∈ Ut.
√
Now, ω(x) := dβ
F (x) fulfills the hypothesis of Lemma 2.7 where the constant in
n)β. The last assertion uses Lemma 5.2. Thus, the operator T is
(2.5) is C2 = (3K
−qβ
continuous from L := Lq(Ω, d
F
(cid:18) qN
) to itself with the norm
√
n)β.
(3K
(cid:19)1/q
q − 1
(cid:107)T(cid:107)L→L ≤ 2
13
(cid:33)(cid:41)
(x)
Hence,(cid:88)
−qβ
F
)
(cid:90)
(cid:33)
(x)
g(x)q d
−qβ
F
(cid:107)gt(cid:107)q
t∈Γ
≤ 2q−1
≤ 2q−1N
≤ 2q−1N
Lq(Ut,d
(cid:40)(cid:32)(cid:88)
(cid:26)(cid:90)
(cid:26)
t∈Γ
Ω
Ut
g(x)q d
(cid:18) q
1 + (72K)qn2q
−qβ
F
(cid:18) qN
(cid:19)
(cid:18) q
(cid:19)
q − 1
√
(3K
q − 1
≤ 4qN 2(72K)qn
q − 1
√
= 4q 122n 72qn (3
n)qβ
+ (72K)qn
(cid:90)
(cid:19)
(x) dx + (72K)qn
√
n)qβ
(3K
Ω
n)qβ(cid:107)g(cid:107)q
Lq(Ω,d
−qβ
F
)
K q(n+β)(cid:107)g(cid:107)q
(cid:90)
Ut
(cid:32)(cid:88)
(cid:27)
t∈Γ
(cid:107)g(cid:107)q
T g(x)q d
−qβ
F
(cid:27)
T g(x)q d
−qβ
F
(x) dx
Lq(Ω,d
−qβ
F
)
.
−qβ
F
)
(cid:19)1/q
Lq(Ω,d
(cid:18) q
q − 1
Therefore, we have a C-decomposition subordinate to {Ut}t∈Γ with constant
√
C0 = 4(12)2n/q(72)n(3
n)β
K n+β.
(5.6)
Finally, inequality (3.10) and the control on the overlapping of the tree covering by
(cid:3)
N = 12n implies (1.2).
Remark 5.3. Notice that the proof of Theorem 1.1 provides an explicit constant
C = 2C0C1 for inequality (1.2), where C0 and C1 are described respectively in (5.6)
and (5.5).
The next result, similar to Proposition 4.1, follows from the Holder inequality (equiv-
alently, from Minkowski's integral inequality).
Proposition 5.4. Let ρ : Rn \ {0} → R be a positive radial Lebesgue measurable
function which is increasing with respect to the radius. Then, the fractional Poincar´e
type inequality
(cid:107)u(x) − uU(cid:107)Lp(U ) ≤ diam (U )n/p ρ(diam(U ))
U1/p
u(y) − u(x)p
y − xn(ρy − x)p dydx
(cid:18)(cid:90)
(cid:90)
U
U
holds for any bounded domain U in Rn and 1 < p < ∞, where uU := 1U
Proof.(cid:90)
(cid:90)
(cid:90)
(cid:90)
(cid:12)(cid:12)(cid:12)(cid:12) 1
(cid:90)
u(x) − u(y)dy
≤ diam(U )n{ρ(diam(U ))}p
U
U
U
(cid:12)(cid:12)(cid:12)(cid:12)p
(cid:90)
U
(cid:90)
dx ≤ 1
U
u(x) − u(y)p
U
U
u(x) − u(y)pdydx
x − yn(ρx − y)p dydx.
U
U
u(x) − uUpdx =
U
(cid:3)
Remark 5.5. If ρ(x) = xs, with s ∈ (0, 1), we recover the classical fractional Poincar´e
inequality.
(cid:19)1/p
(cid:82)
(5.7)
U u(y)dy.
14
We generalize the fractional Poincar´e inequality stated in Theorem 1.1 by replacing
the fractional derivatives given by the power functions xs, with 0 < s < 1, by general
increasing and positive radial functions ρx.
Theorem 5.6. Let Ω in Rn be a bounded John domain and 1 < p < ∞. Given an
arbitrary compact set F in ∂Ω, a parameter β ≥ 0 and a positive radial Lebesgue
measurable function ρ : Rn \{0} → R increasing with respect to the radius, there exists
a constant C such that
u(x) − uΩ,ωpdpβ
(cid:19)1/p
(cid:18)(cid:90)
F (x)dx
Ω
(cid:19)1/p
(cid:18)(cid:90)
(cid:90)
≤ C
u(x) − u(y)p
(5.8)
for all function u ∈ Lp(Ω, d(x)pβ). We denote by d(x) and dF (x) the distance from x
to ∂Ω and F respectively, and by uΩ,ω the weighted average
x − yn(ρx − y)p [ρ(2d(x))]pdpβ
(cid:82)
F (x) dydx
Ω∩B(x,d(x))
Ω
Ω u(z)dpβ
F (z)dz.
1
dpβ
F (Ω)
In addition, the constant C in (5.8) can be written as
C = Cn,p,β K n+β,
where K is the geometric constant introduced in (5.1).
Proof. This proof mimics the one of Theorem 1.1 with Proposition 5.4 instead of Propo-
sition 4.2. Indeed, we will use again the tree covering {Ut}t∈Γ of Ω defined in (5.2) and
the weight ω(x) = dβ
F (x), however, in this case µ(x, y) is defined as
µ(x, y) :=
[ρ(2d(x))]pdpβ
F (x)
x − yn(ρx − y)p .
We only have to show that (3.6) is verified on Ut, for all t, with uniform constant. This
fact follows from (5.7) by using the inequality diam(Ut) ≤ 2d(x) for all x ∈ Ut.
(cid:3)
Acknowledgements
This research was initiated when the second author visited the University of Helsinki
during the Summer of 2016. This author gratefully acknowledges Professor Michel
Lapidus from University of California Riverside for his financial support for the ex-
penses of the trip.
References
[1] S. Buckley, P. Koskela, and G. Lu, Boman equals John, Proceedings of the 16th Rolf
Nevanlinna Colloquium (1995), 91 -- 99.
[2] B. Dyda, L. Ihnatsyeva, and A. Vahakangas, On improved fractional Sobolev-Poincar´e
inequalities, Ark. Mat., 54 (2) (2016), 437 -- 454.
[3] I. Drelichman, and R. G. Dur´an, Improved Poincar´e inequalities in fractional Sobolev spaces,
arXiv: 1705.04227v1 (2017).
[4] R. Hurri-Syrjanen, and A. V. Vahakangas, On fractional Poincar´e inequalities, J. Anal.
Math. 120 (2013), 85 -- 104.
[5] R. Hurri-Syrjanen, and A. V. Vahakangas, Fractional Sobolev-Poincar´e and fractional
Hardy inequalities, Mathematika 61 (2015), 385 -- 401.
[6] F. John, Rotation and strain Comm. Pure Appl. Math. 14 (1961), 391 -- 413.
[7] F. L´opez-Garc´ıa, A decomposition technique for integrable functions with applications to the
divergence problem, J. Math. Anal. Appl. 418 (2014), 79 -- 99.
15
[8] F. L´opez-Garc´ıa, Weighted Korn inequality on John domains, Studia Math. 241 (2018),
17 -- 39.
[9] F. L´opez-Garc´ıa, Weighted generalized Korn inequality on John domains, submitted for
publication (2017).
[10] Augusto C. Ponce, A variant of of Poincar´e' s inequality, C. R. Acad. Sci. Paris, Ser. I
337 (2003), 253 -- 257.
[11] Augusto C. Ponce, A new approach to Sobolev spaces and connections to Γ-convergence,
Calc. Var. 19 (2004), 229 -- 255.
[12] Augusto C. Ponce, An estimate in the spirit of Poincar´e's inequality, J. Eur. Math. Soc.
6 (2004), 1 -- 15.
[13] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton Univ.
Press, Princeton, NJ, 1970.
Department of Mathematics and Statistics, FI-00014 University of Helsinki, Finland
E-mail address: [email protected]
Department of Mathematics and Statistics, California State Polytechnic Univer-
sity Pomona, 3801 West Temple Avenue, Pomona, CA (91768), US
E-mail address: [email protected]
|
1702.08671 | 1 | 1702 | 2017-02-28T07:09:30 | On The Absolute Value of The Product and the Sum of Linear Operators | [
"math.FA"
] | Let $A,B\in B(H)$. In the present paper, we establish simple and interesting facts on when we have $|A||B|=|B||A|$, $|AB|=|A||B|$, $|A\pm B|\leq |A|+|B|$, $||A|-|B||\leq |A\pm B|$ and $\||A|-|B|\|\leq \|A\pm B\|$, where $|\cdot|$ denotes the absolute value (or modulus) of an operator. The results give some other interesting consequences. | math.FA | math |
ON THE ABSOLUTE VALUE OF THE PRODUCT AND
THE SUM OF LINEAR OPERATORS
MOHAMMED HICHEM MORTAD
Abstract. Let A, B ∈ B(H). In the present paper, we establish simple
and interesting facts on when we have AB = BA, AB = AB,
A ± B ≤ A + B, A − B ≤ A ± B and kA − Bk ≤ kA ± Bk,
where · denotes the absolute value (or modulus) of an operator. The
results give some other interesting consequences.
1. Introduction
Let H be a complex Hilbert space and let A, B ∈ B(H). We say that A
is positive, and we write A ≥ 0, if < Ax, x >≥ 0 for all x ∈ H. Since H is
a complex Hilbert space, a positive operator is clearly self-adjoint. We say
that A ≥ B if they are both self-adjoint and A − B ≥ 0. Recall also that if
A ≥ 0, then there is a unique positive operator B such that B 2 = A. We call
it the (positive) square root of A and we denote it by √A (or A
2 ). Next, we
gather basic results on square roots of sums and products.
1
Lemma 1.1. Let A, B ∈ B(H) be such that AB = BA and A, B ≥ 0. Then
• AB ≥ 0.
• √AB = √A√B.
• √A + B ≤ √A + √B.
The unique positive square root of the positive operator A∗A is commonly
known as the absolute value (or modulus) of A. We denote it by A, that
is, A = √A∗A. Notice that kAk = k A k always holds.
We usually warn students to be careful with this notation as it may mislead
them to think that e.g.
A = A∗, A + B ≤ A + B or AB = AB
would hold. Counterexamples are easily found in the setting of 2 by 2 ma-
trices. Notice that A is normal if and only if A = A∗.
Also, a priori if A, B are arbitrary, then there is no reason why we should
expect AB = BA to hold (for instance, just think of positive operators).
2010 Mathematics Subject Classification. Primary 47A63, Secondary 47A62, 47B15,
47B20.
Key words and phrases. Absolute Value. Triangle Inequality. Normal, Hyponormal, Self-
adjoint and Positive Operators. Commutativity. Fuglede Theorem.
1
2
M. H. MORTAD
Even when AB = BA, the equality AB = BA need not hold. For
example, let
A = (cid:18) 1 1
0 1 (cid:19) and B = (cid:18) 0 1
0 0 (cid:19) .
Then as we can easily verify:
AB = (cid:18) 0 1
0 0 (cid:19) = BA
whereas AB 6= BA.
terexample (cf. Proposition 2.1).
Observe that we have purposely avoided normal operators in our coun-
The main aim of this paper is to investigate when relations of the types
• AB = BA;
• AB = AB;
• A + B ≤ A + B;
• A − B ≤ A + B;
• k A − B k ≤ kA + Bk;
hold. It turns out that normality and sometimes hyponormality plus commu-
tativity are sufficient for these relations to hold. This comes to corroborate
the resemblance to complex numbers which is already known to many. Notice
also that commutativity is not unnatural as we already have it in (C,×).
The idea here is to start from scratch, and use as basic results as possible to
make the paper accessible to a wide audience. We note that for example, we
have wittingly avoided the use of the spectral theorem of normal operators.
Therefore, most of the results here can be taught at elementary courses in
Operator Theory.
It is worth noticing that there is a big amount of papers which have dealt
with inequalities involving absolute values and/or norms of operators. The
literature is so rich that we rather refer readers to books which have gathered
most of these results. For example, see [1], [4] and [10].
Finally, we assume the reader is familiar with other basic results on Op-
erator Theory. A well established reference is [2]. We do recall two crucial
results though.
Theorem 1.2. (Löwner-Heinz Inequality, see [8] for a simple proof ) If
A ≥ B ≥ 0, then Aα ≥ Bα for any α ∈ [0, 1].
Remark. It is known to readers that A ≥ B ≥ 0 implies that A2 ≥ B 2 when
AB = BA.
Since we will be dealing with sums and products of commuting normal
operators, the use of the celebrated Fuglede-Putnam theorem is inevitable.
The following lemma will be used below without further notice.
Lemma 1.3. Let A, B ∈ B(H) where A is normal. Then, we have
AB = BA ⇐⇒ A∗B = BA∗ ⇐⇒ AB∗ = B∗A ⇐⇒ A∗B∗ = B∗A∗.
ABSOLUTE VALUE OF THE PRODUCT AND THE SUM OF OPERATORS
3
2. Main Results: Absolute Value and Products
We start with the following:
Proposition 2.1. Let A, B ∈ B(H) be such that AB = BA. If A is normal,
then AB = BA.
Remark. The preceding result was proved in [9] by assuming that both A
and B are normal.
Proof. Since AB = BA and A is normal, we have A∗B = BA∗. We then
clearly have from the previous two relations:
AB = BA =⇒ A∗AB = A∗BA = BA∗A.
Hence
AB = BA.
Since A is self-adjoint, the previous equality gives (by taking adjoints)
AB∗ = B∗A. Hence
B∗AB = AB∗B =⇒ B∗BA = AB∗B =⇒ BA = AB,
as required.
(cid:3)
We have already observed above that in general AB 6= AB. The
following result is somewhat inspired by a one in [5].
Theorem 2.2. Let A, B ∈ B(H) be self-adjoint such that AB is normal.
Then
AB = AB.
Remark. It was noted in [6] that if S, T are two non-commuting self-adjoint
operators, then the inequality ST ≤ ST never holds. So, in our result
the normality of the product transforms the non valid inequality into a true
full equality.
Remark. Notice that AB being a normal product of two self-adjoint operators
does not necessarily imply that AB is self-adjoint, i.e. we do not necessarily
have AB = BA. If, however, we impose further that A ≥ 0 (or B ≥ 0), then
AB becomes self-adjoint. See e.g. [7].
Proof. Since A and B are self-adjoint, we may write
B(AB) = BAB = (AB)∗B.
Since AB and (AB)∗ are normal, the Fuglede-Putnam theorem gives
B(AB)∗ = (AB)∗∗B or merely B 2A = AB 2.
Consequently, B 2A2 = AB 2A = A2B 2.
On the other hand, we easily see that
AB2 = (AB)∗AB = AB(AB)∗ = AB 2A = A2B 2
4
M. H. MORTAD
and so
as required.
AB = √A2B 2 = √A2√B 2 = AB,
(cid:3)
Since AB is self-adjoint, we have:
Corollary 2.3. Let A, B ∈ B(H) be self-adjoint such that AB is normal.
Then AB is self-adjoint, i.e. AB = BA. Moreover, if we also assume
that A, B ≥ 0, then AB ≥ 0.
The assumptions of the previous theorem cannot just be dropped. We
give a counterexample for each hypothesis.
• Let
A = (cid:18) 2
0
1 0 (cid:19) .
0 −1 (cid:19) and B = (cid:18) 0 1
AB = (cid:18) 0
−1 0 (cid:19)
2
Then each of A and B is self-adjoint but AB is not normal for
We can easily check that
AB = (cid:18) 1 0
0 2 (cid:19) , A = (cid:18) 2 0
0 1 (cid:19) and B = (cid:18) 1 0
0 1 (cid:19) ,
i.e.
• Let
AB 6= AB.
A = (cid:18) 0 1
2 0 (cid:19) and B = (cid:18) 0 2
1 0 (cid:19) .
Then, neither A nor B is normal. Their product AB is, however,
self-adjoint (hence normal!) because
AB = (cid:18) 1 0
0 4 (cid:19)
Next, we have
A = (cid:18) 2 0
0 1 (cid:19) , B = (cid:18) 1 0
0 2 (cid:19) and AB = (cid:18) 1 0
0 2 (cid:19) .
Accordingly,
AB = (cid:18) 1 0
0 2 (cid:19) 6= (cid:18) 2 0
0 2 (cid:19) = AB.
An akin result to Theorem 2.2 is:
Theorem 2.4. Let A, B ∈ B(H) be such that AB = BA. If A is normal,
then
AB = AB.
ABSOLUTE VALUE OF THE PRODUCT AND THE SUM OF OPERATORS
5
Remark. It was also noted in [6] that if S, T are two commuting normal
operators, then the inequality ST ≤ ST holds. So, our result here is
stronger.
Proof. Since AB = BA and A is normal, we get A∗B = BA∗ or AB∗ = B∗A.
Hence
AB2 = (AB)∗AB = B∗A∗AB = A∗B∗AB = A∗AB∗B.
By Proposition 2.1, A∗AB∗B = B∗BA∗A. Consequently,
AB = √A∗AB∗B = √A∗A√B∗B = AB,
as required.
(cid:3)
Before generalizing the previous result, we give some direct consequences.
The first one is a funny application.
Corollary 2.5. Let A, B ∈ B(H) be such that AB = BA. If A and B are
normal, then
AB = A∗B = AB∗ = A∗B∗ = B∗A∗ = B∗A = BA∗ = BA.
Proof. Since A and B are normal, A = A∗ and B = B∗. As AB = BA,
then AB = BA for A (or B!) is normal. Now, apply Theorem 2.4 to
each of the eight products.
(cid:3)
Corollary 2.6. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and B is invertible, then
AB−1 = AB−1.
Proof. Since AB = BA and B is invertible, we have AB−1 = B−1A. Theo-
rem 2.4 does the remaining job.
(cid:3)
Corollary 2.7. Let A ∈ B(H) be normal and invertible. Then
A−1 = A−1.
Proof. It is clear that
So, the self-adjoint A is right invertible and so it is invertible (cf. [3]) and:
I = AA−1 = AA−1.
A−1 = A−1.
Theorem 2.4 may be generalized as follows:
Proposition 2.8. Let (Ai)i=1,··· ,n be a family of pairwise commuting ele-
ments of B(H). If all (Ai)i=1,··· ,n but one are normal, then
A1A2 ··· An−1An = A1A2··· An−1An.
(cid:3)
6
M. H. MORTAD
Proof. If A1, A2,··· , An−1 are normal, then just apply the preceding theorem
by using a proof by induction. Otherwise, just use commutativity to push
the non normal factor to the right as many times as possible until it will be
i=1 Ai". Then proceed as just
(cid:3)
the last factor on the right of the product "Qn
It is simple to see that A2 = A2 does not hold in general. For instance,
indicated three lines above.
let
A = (cid:18) 0 2
1 0 (cid:19) .
Then
A2 = (cid:18) √2
0
0 √2 (cid:19) 6= A2 = (cid:18) 2 0
0 1 (cid:19) .
But for normal A, things are better.
Corollary 2.9. Let A ∈ B(H) be normal and invertible. Let n ∈ Z. Then
An = An.
Proof. The case n ≥ 0 follows from Proposition 2.8. The case n < 0 follows
from Proposition 2.8 and Corollary 2.7.
(cid:3)
3. Main Results: Absolute Value and Sums
versions.
We now turn to the triangle inequality w.r.t. · . We have two different
Before all else, we state a result (perhaps known to many) which will be
called on later. Its proof relies on the following yet simpler result.
Lemma 3.1. If A ∈ B(H) is anti-symmetric, i.e. A∗ = −A, then A2 ≤ 0.
Lemma 3.2. Let T ∈ B(H) be hyponormal (i.e. T T ∗ ≤ T ∗T , that is,
kT ∗xk ≤ kT xk for all x ∈ H). Then
T + T ∗
ReT =
2
√T ∗T = T.
≤
Proof. It is clear that T − T ∗ is anti-symmetric and so by Lemma 3.1:
(T − T ∗)2 ≤ 0. Hence
(T − T ∗)2 ≤ 0 ⇐⇒ T 2 + T ∗2
− T T ∗ − T ∗T ≤ 0.
But, T is hyponormal and so −T T ∗ − T ∗T ≥ −2T ∗T . So,
T 2 + T ∗2
− 2T ∗T ≤ 0
+ T T ∗ + T ∗T ≤ T 2 + T ∗2
(T + T ∗)2 ≤ 4T ∗T or T + T ∗ ≤ 2T
+ 2T ∗T ≤ 4T ∗T.
or
Therefore,
T 2 + T ∗2
by Theorem 1.2.
ABSOLUTE VALUE OF THE PRODUCT AND THE SUM OF OPERATORS
7
Remembering that S− = 1
2 (S − S) ≥ 0 whenever S is self-adjoint, we
conclude that
as required.
T + T ∗ ≤ T + T ∗ ≤ 2T = 2√T ∗T ,
(cid:3)
The following fairly simple result is also useful to us.
Lemma 3.3. Let A, B ∈ B(H) such that A is normal and B is hyponormal.
If AB = BA, then A∗B is hyponormal.
Proof. Let x ∈ H. As AB = BA, by the normality of A and the hyponor-
mality of B we have
k(A∗B)∗xk = kB∗Axk ≤ kBAxk = kABxk = kA∗Bxk,
establishing the hyponormality of A∗B.
(cid:3)
Here is the first version of the triangle inequality.
Theorem 3.4. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and B is hyponormal, then the following triangle inequality holds:
A + B ≤ A + B.
Proof. Since A is normal and AB = BA, we know from Proposition 2.1 that
AB = BA. Hence
A + B2 ≤ (A + B)2 ⇐⇒ (A + B)∗(A + B) ≤ A∗A + B∗B + 2√A∗A√B∗B
We already know from above that √A∗A√B∗B = √A∗AB∗B. So, to
⇐⇒ A∗B + B∗A ≤ 2√A∗A√B∗B.
prove the desired triangle inequality, we are only required to prove
A∗B + B∗A ≤ 2√A∗AB∗B.
But
the following holds:
A∗AB∗B = AA∗B∗B = AB∗A∗B = B∗AA∗B.
If we set T = A∗B, then are done with the proof if we come to show that
T + T ∗ ≤ 2√T ∗T .
But this is just Lemma 3.2 once we show that A∗B is hyponormal. This is
in effect the case as A∗B is hyponormal by Lemma 3.3.
Therefore, under the assumptions of our theorem we have shown that
A + B2 ≤ (A + B)2.
Hence, by Theorem 1.2, we have ended up with
A + B ≤ A + B,
and this is precisely what we wanted to prove.
(cid:3)
8
M. H. MORTAD
Remark. The foregoing result need not hold if commutativity is dropped
even if A and B are self-adjoint. The reader may check this easily via the
following example:
A = (cid:18) −1
1
1 −1 (cid:19) and B = (cid:18) 2 0
0 0 (cid:19) .
Corollary 3.5. Let T ∈ B(H) be normal. Then
T ≤ ReT + ImT.
Proof. Write T = ReT + iImT where ReT and ImT are commuting self-
adjoint operators. Then apply Theorem 3.4.
(cid:3)
We have another simple consequence of Theorem 3.4.
Corollary 3.6. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and B is hyponormal, then the following triangle inequality holds:
A − B ≤ A + B.
Proof. Since AB = BA, we know that A(−B) = (−B)A. Also −B is hy-
ponormal. Then, apply Theorem 3.4.
(cid:3)
Theorem 3.4 may be generalized to a finite sum of operators. Before, recall
that the sum of two commuting normal operators remains normal. This too
may be generalized (the proof by induction is omitted).
Proposition 3.7. Let (Ai)i=1,··· ,n be a family of normal pairwise commuting
elements of B(H). Then A1 + A2 + ··· + An is normal.
We are ready for the promised generalization of Theorem 3.4 whose proof
is again a proof by induction.
Corollary 3.8. Let (Ai)i=1,··· ,n be a family of pairwise commuting elements
of B(H). If all (Ai)i=1,··· ,n are normal except one which is assumed to be
hyponormal, then
A1 + A2 + ··· + An ≤ A1 + A2 + ··· + An.
It is known that an inequality of the type kA − Bk ≤ kA ± Bk is not
true in general even if A and B are self-adjoint.
Proposition 3.9. Let A, B ∈ B(H) be such that AB = BA. If A and B
are normal, then the following inequality holds:
kA − Bk ≤ kA + Bk.
Probably the following lemma has been noted elsewhere but we state it
here anyway with a proof.
Lemma 3.10. Let S, T ∈ B(H) be self-adjoint where S ≥ 0. If −S ≤ T ≤ S,
then kTk ≤ kSk.
ABSOLUTE VALUE OF THE PRODUCT AND THE SUM OF OPERATORS
9
Proof. By assumption, for all x ∈ H
− < Sx, x >≤< T x, x >≤< Sx, x > or merely < T x, x > ≤< Sx, x > .
Therefore,
kTk = sup
kxk=1 < T x, x > ≤ sup
kxk=1
as desired.
Let us prove Proposition 3.9.
< Sx, x >= kSk,
(cid:3)
Proof. Since A and B are commuting normal operators, we know that A + B
too is normal. Since A + B commutes with B, by Corollary 3.6 we have
A = A + B − B ≤ A + B + B =⇒ A − B ≤ A + B.
Similarly, as A + B commutes with A, we get
Whence
B − A ≤ A + B.
−A + B ≤ A − B ≤ A + B.
By Lemma 3.10 (and remembering that kTk = k T k for T ∈ B(H)), we
obtain
as required.
kA − Bk ≤ k A + B k = kA + Bk,
(cid:3)
Remark. In the previous proposition, if B is only hyponormal, then at the
moment we are only sure that:
B − A ≤ A + B
because we can only prove that A + B is hyponormal. We will remedy this
little problem shortly.
Corollary 3.11. Let A, B ∈ B(H) be such that AB = BA. If A and B are
normal, then the following inequality holds:
kA − Bk ≤ kA − Bk.
Proposition 3.9 can be improved as it is a particular case of the following
remarkable result:
Proposition 3.12. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and B is hyponormal, then the following inequality holds:
A − B ≤ A − B.
Proof. We easily see as AB = BA that
A − B2 ≤ A − B2 ⇐⇒ A2 + B2 − 2AB ≤ A2 + B2 − A∗B − B∗A
⇐⇒ A∗B + B∗A ≤ 2√A∗A√B∗B
⇐⇒ A∗B + B∗A ≤ 2√B∗AA∗B.
10
M. H. MORTAD
But, this is always true in virtue of Lemma 3.2 as A∗B is hyponormal.
Therefore, we have shown
A glance at Theorem 1.2 finally gives
A − B2 ≤ A − B2.
A − B ≤ A − B.
(cid:3)
Corollary 3.13. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and B is hyponormal, then the following inequality holds:
Proof. Since B is hyponormal, so is −B. The rest is obvious.
A − B ≤ A + B.
Here is the improvement of Proposition 3.9:
(cid:3)
Corollary 3.14. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and B is hyponormal, then the following inequality holds:
Proof. By Proposition 3.9 and Corollary 3.13, we know that
kA − Bk ≤ kA ± Bk.
Then, calling on Lemma 3.10 yields
A − B ≤ A ± B.
kA − Bk = k A − B k ≤ k A ± B k = kA ± Bk.
(cid:3)
If we want to drop commutativity in Theorem 3.4, then this is at the cost
of adding an extra condition. Also, we only have to assume that one of the
two operators is normal.
Theorem 3.15. Let A, B ∈ B(H) be such that AB = BA. If A is normal
and A∗B + B∗A ≤ 0, then
A + B ≤ A + B.
A + B = √A∗A + A∗B + B∗A + B∗B ≤ √A∗A + B∗B.
Since AB = BA and A is normal, Proposition 2.1 implies that AB =
BA or A2B2 = B2A2. Finally, Lemma 1.1 does the remaining job,
i.e. it gives us
and this completes the proof.
A + B ≤ A + B
(cid:3)
Proof. Clearly,
As A∗B + B∗A ≤ 0, then
By Theorem 1.2, we have
(A + B)∗(A + B) = A∗A + A∗B + B∗A + B∗B.
A∗A + A∗B + B∗A + B∗B ≤ A∗A + B∗B.
ABSOLUTE VALUE OF THE PRODUCT AND THE SUM OF OPERATORS
11
acknowledgement
The author wishes to thank Mr. S. Dehimi for a discussion which led to
a slight improvement of the result of Theorem 3.4.
References
[1] R. Bhatia, Matrix analysis, Graduate Texts in Mathematics, 169. Springer-Verlag,
New York, 1997.
[2] J. B. Conway, A course in functional analysis, Graduate Texts in Mathematics, 96.
Springer, 1990 (2nd edition).
[3] S. Dehimi, M. H. Mortad, Right (or left) invertibility of bounded and un-
(submitted).
bounded operators and applications to the spectrum of products,
https://arxiv.org/pdf/1505.02719v1.pdf.
[4] T. Furuta, Invitation to linear operators: From matrices to bounded linear operators
on a Hilbert Space, Taylor & Francis, Ltd., London, 2001.
[5] K. Gustafson, M. H. Mortad, Conditions implying commutativity of unbounded self-
adjoint operators and related topics, J. Operator Theory, 76/1 (2016) 159-169.
[6] R. Harte, The triangle inequality in C ∗ algebras, Filomat No. 20, part 2 (2006), 51-53.
[7] M. H. Mortad, An application of the Putnam-Fuglede theorem to normal products of
selfadjoint operators, Proc. Amer. Math. Soc. 131 (2003) 3135-3141.
[8] G. K. Pedersen, Some operator monotone functions, Proc. Amer. Math. Soc., 36
(1972) 309-310.
[9] R. Zeng, Young's inequality in compact operators-the case of equality, JIPAM. J. In-
equal. Pure Appl. Math., 6/4 (2005), Article 110, 10 pp.
[10] X. Zhan, Matrix inequalities, Lecture Notes in Mathematics, 1790. Springer-Verlag,
Berlin, 2002.
Department of Mathematics, University of Oran 1, Ahmed Ben Bella, B.P.
1524, El Menouar, Oran 31000, Algeria.
Mailing address:
Pr Mohammed Hichem Mortad
BP 7085 Seddikia Oran
31013
Algeria
E-mail address: [email protected], [email protected].
|
1711.04122 | 2 | 1711 | 2018-01-29T21:22:52 | Bohr's equivalence relation in the space of Besicovitch almost periodic functions | [
"math.FA"
] | Based on Bohr's equivalence relation which was established for general Dirichlet series, in this paper we introduce a new equivalence relation on the space of almost periodic functions in the sense of Besicovitch, $B(\mathbb{R},\mathbb{C})$, defined in terms of polynomial approximations. From this, we show that in an important subspace $B^2(\mathbb{R},\mathbb{C})\subset B(\mathbb{R},\mathbb{C})$, where Parseval's equality and Riesz-Fischer theorem holds, its equivalence classes are sequentially compact and the family of translates of a function belonging to this subspace is dense in its own class. | math.FA | math | Noname manuscript No.
(will be inserted by the editor)
Bohr's equivalence relation in the space of Besicovitch
almost periodic functions
J.M. Sepulcre · T. Vidal
Received: date / Accepted: date
Abstract Based on Bohr's equivalence relation which was established for gen-
eral Dirichlet series, in this paper we introduce a new equivalence relation on
the space of almost periodic functions in the sense of Besicovitch, B(R, C),
defined in terms of polynomial approximations. From this, we show that in
an important subspace B2(R, C) ⊂ B(R, C), where Parseval's equality and
Riesz-Fischer theorem holds, its equivalence classes are sequentially compact
and the family of translates of a function belonging to this subspace is dense
in its own class.
Keywords Almost periodic functions · Besicovitch almost periodic functions ·
Bochner's theorem · Exponential sums · Fourier series
Mathematics Subject Classification (2010) 42A75 · 42A16 · 42B05 ·
46xx · 42Axx · 30B50 · 30Bxx
1 Introduction
The class of almost periodic functions, whose theory was created and developed
in its main features by H. Bohr during the 1920's, is the class of continuous
functions possessing a certain structural property, which is a generalization of
pure periodicity. This theory opened a way to study a wide class of trigono-
metric series of the general type and even exponential series. In this context,
we can cite, among others, the papers [3,4,5,6,8,10].
Let f (t) be a real or complex function of an unrestricted real variable t. The
notion of almost periodicity given by Bohr involves the fact that f (t) must be
continuous, and for every ε > 0 there corresponds a number l = l(ε) > 0 such
J.M. Sepulcre · T. Vidal
Department of Mathematics
University of Alicante
03080-Alicante, Spain
E-mail: [email protected], [email protected]
8
1
0
2
n
a
J
9
2
]
.
A
F
h
t
a
m
[
2
v
2
2
1
4
0
.
1
1
7
1
:
v
i
X
r
a
2
J.M. Sepulcre, T. Vidal
that each interval of length l contains a number τ satisfying f (t+τ )−f (t) ≤ ε
for all t. As in [8], we will denote as AP (R, C) the space of almost periodic
functions in the sense of this definition (Bohr's condition). A very important
result of this theory is the approximation theorem according to which the
class of almost periodic functions AP (R, C) coincides with the class of limit
functions of uniformly convergent sequences of trigonometric polynomials of
the type
a1eiλ1t + . . . + aneiλnt
(1)
with arbitrary real exponents λj and arbitrary complex coefficients aj. More-
over, S. Bochner observed that Bohr's notion of almost periodicity of a function
f is equivalent to the relative compactness, in the sense of uniform convergence,
of the family of its translates {f (t + h)}, h ∈ R.
In the course of time, some variants and extensions of Bohr's concept have
been introduced, most notably by A. S. Besicovitch, W. Stepanov and H.
Weyl. We refer the reader to the papers by Besicovitch [3, Chapter II], Bohr
and Foelner [7], Corduneanu [8], and by Andres, Bersani and Grande [1] and
the references therein for a comprehensive treatment of this subject.
In particular, A.S. Besicovitch enlarged the class of almost periodic func-
tions by considering the convergence of sequences of functions in a more general
sense than uniform convergence. In this way, the Besicovitch spaces Bp(R, C),
1 ≤ p < ∞, are obtained by the completion of the trigonometric polynomials
of the form (1) with respect to the seminorms
lim sup
l→∞
(2l)−1Z l
−l f (t)pdt!1/p
.
This topology is certainly weaker than that of the uniform convergence. In
particular, the space B1(R, C) is denoted by B(R, C) and contains AP (R, C),
B2(R, C) and all variants of almost periodic functions which were mentioned
above.
Moreover, for any function f ∈ B(R, C) there exists the mean value
M (f ) = lim
l→∞
(2l)−1Z l
−l
f (t)dt
(2)
and, at most, a countable set of values of λk ∈ R such that ak = a(f, λk) =
M (f (t)e−λkt) 6= 0. Thus the series Pk≥1 akeiλkt is called the Fourier series
of f [8, Section 4.2]. Also, λk and ak are called the Fourier exponents and
coefficients of the function f , respectively.
In the case that f ∈ B2(R, C), with Pk≥1 akeiλkt the Fourier series of f ,
it is accomplished the Parseval's equality [3, p. 109]
ak2 = M (f (t)2) < ∞.
Xk≥1
In this respect, if f1, f2 ∈ B2(R, C), then f1 and f2 have the same Fourier
series if and only if M (f1(t) − f2(t)2) = 0. That is, two functions satisfying
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
3
this condition belong to the same class of equivalence defined in terms of the
Fourier series. This equivalence relation is inherent to the classes Bp(R, C)
and it is different from the generalization of Bohr's equivalence of Definition
2 which is the main tool of this paper.
Besicovitch's generalization is interesting because, for this extension, the
analogue of the Riesz-Fischer theorem is also valid, that is to say, any trigono-
B2(R, C) almost periodic function [3, p. 110] (in this sense, B2(R, C) is also
called AP2(R, C) in [8]). This is not the case for some Stepanov or Weyl func-
metric series Pn≥1 aneiλnt with Pn≥1 an2 finite is the Fourier series of a
tions [10]. As a consequence of the above, a Fourier series Pn≥1 aneiλnt so
thatPn≥1 an2 < ∞ represents an equivalence class of functions in B2(R, C).
In this paper we extend Bohr's equivalence relation to the Fourier series
associated with the Besicovitch almost periodic functions, and hence to the
Besicovitch almost periodic functions too. In this way, in view of the analogue
of the Riesz-Fischer theorem and with respect to the topology of B2(R, C),
the main result of our paper shows that, fixed an almost periodic function in
B2(R, C), the limit points of the set of its translates are precisely the functions
which are equivalent to it (see Theorem 2 in this paper). This means that the
Bochner-type property, which is satisfied for the Besicovitch classes of almost
periodic functions defined as above in terms of polynomials approximations
(see [1, Definition 5.5, Definition 5.17 and Theorem 5.34] or [8, Section 3.4,
p. 65]), is now refined for B2(R, C) in the sense that we show that the condi-
tion of almost periodicity in the Besicovitch sense implies that every sequence
of translates has a subsequence that converges in B2(R, C) to an equivalent
function.
2 Preliminary definitions and results on exponential sums
We shall refer to the expressions of the type
P1(p)eλ1p + . . . + Pj(p)eλj p + . . .
as exponential sums, where the frequencies λj are complex numbers and the
Pj(p) are polynomials in p. In this paper we are going to consider some func-
tions which are associated with a concrete subclass of these exponential sums,
where the parameter p will be changed by t in the real case. In this way, as in
[11], we take the following definition.
Definition 1 Let Λ = {λ1, λ2, . . . , λj, . . .} be an arbitrary countable set of
distinct real numbers, which we will call a set of exponents or frequencies. We
will say that an exponential sum is in the class SΛ if it is a formal series of
type
ajeλj p, aj ∈ C, λj ∈ Λ.
Xj≥1
(3)
We next introduce an equivalence relation on the classes SΛ.
4
J.M. Sepulcre, T. Vidal
Definition 2 Given an arbitrary countable set Λ = {λ1, λ2, . . . , λj , . . .} of
distinct real numbers, consider A1(p) and A2(p) two exponential sums in the
class SΛ, say A1(p) =Pj≥1 ajeλj p and A2(p) =Pj≥1 bjeλj p. We will say that
∼A2) if for each integer
A1 is equivalent to A2 (in that case, we will write A1
value n ≥ 1, with n ≤ ♯Λ, there exists a Q-linear map ψn : Vn → R, where Vn
is the Q-vector space generated by {λ1, λ2, . . . , λn}, such that
∗
bj = ajeiψn(λj ), j = 1, . . . , n.
∗
equivalence relation.
It is plain that the relation
∼ considered in the foregoing definition is an
Let GΛ = {g1, g2, . . . , gk, . . .} be a basis of the vector space over the ra-
tionals generated by a set Λ of exponents, which implies that GΛ is linearly
independent over the rationals and each λj is expressible as a finite linear
combination of terms of GΛ, say
λj =
qj
Xk=1
rj,kgk, for some rj,k ∈ Q.
(4)
By abuse of notation, we will say that GΛ is a basis for Λ. Moreover, we will
say that GΛ is an integral basis for Λ when rj,k ∈ Z for any j, k. By taking
this into account, the equivalence relation introduced in Definition 2 can be
characterized in terms of a basis for Λ (see [11, Proposition 1']).
∗
∗
∗
Proposition 1 Given Λ = {λ1, λ2, . . . , λj, . . .} a set of exponents, consider
A1(p) and A2(p) two exponential sums in the class SΛ, say A1(p) =Pj≥1 ajeλj p
and A2(p) = Pj≥1 bjeλj p. Fixed a basis GΛ for Λ, for each j ≥ 1 let rj be
∼A2 if and only
the vector of rational components satisfying (4). Then A1
if for each integer value n ≥ 1, with n ≤ ♯Λ, there exists a vector xn =
(xn,1, xn,2, . . . , xn,k, . . .) ∈ R♯GΛ such that bj = aje<rj ,xn>i for j = 1, 2, . . . , n.
Furthermore, if GΛ is an integral basis for Λ then A1
∼A2 if and only if there
exists x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈ R♯GΛ such that bj = aje<rj ,x0>i for
every j ≥ 1.
Proof For each integer value n ≥ 1, let Vn be the Q-vector space generated by
{λ1, . . . , λn}, V the Q-vector space generated by Λ, and GΛ = {g1, g2, . . . , gk, . . .}
∼A2, by Definition 2 for each integer value n ≥ 1, with
a basis of V . If A1
n ≤ ♯Λ, there exists a Q-linear map ψn : Vn → R such that bj = ajeiψn(λj ), j =
ij
1, 2 . . . , n. Hence bj = ajei P
k=1 rj,kψn(gk), j = 1, 2 . . . , n or, equivalently,
bj = ajei<rj ,xn>, j = 1, 2 . . . , n, with xn := (ψn(g1), ψn(g2), . . .). Conversely,
suppose the existence, for each integer value n ≥ 1, of a vector of the form xn =
(xn,1, xn,2, . . . , xn,k, . . .) ∈ R♯GΛ such that bj = aje<rj ,xn>i, j = 1, 2 . . . , n.
Thus a Q-linear map ψn : Vn → R can be defined from ψn(gk) := xn,k, k ≥ 1.
Therefore ψn(λj) = Pij
k=1 rj,kψ(gk) = < rj, xn >, j = 1, 2 . . . , n, and the
result follows.
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
5
∗
Now, suppose that GΛ is an integral basis for Λ and A1
∼A2. Thus, by
above, for each fixed integer value n ≥ 1, let xn = (xn,1, xn,2, . . .) ∈ R♯GΛ be a
vector such that bj = ajei<rj ,xn>, j = 1, 2 . . . , n. Since each component of rj
is an integer number, without loss of generality, we can take xn ∈ [0, 2π)♯GΛ
as the unique vector in [0, 2π)♯GΛ satisfying the above equalities, where we
assume xn,k = 0 for any k such that rj,k = 0 for j = 1, . . . , n. Therefore, under
this assumption, if m > n then xm,k = xn,k for any k so that xn,k 6= 0. In this
way, we can construct a vector x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈ [0, 2π)♯GΛ such
that bj = aje<rj,x0>i for every j ≥ 1. Indeed, if r1,k 6= 0 then the component
x0,k is chosen as x1,k, and if r1,k = 0 then each component x0,k is defined as
xn+1,k where rj,k = 0 for j = 1, . . . , n and rn+1,k 6= 0. Conversely, if there
exists x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈ R♯GΛ such that bj = aje<rj ,x0>i for
every j ≥ 1, then it is clear that A1
∼A2 under Definition 2.
∗
Λ then λj =Pij
Λ, and so on. In this way, if λj ∈ G∗
In terms of the natural basis, we next prove another characterization which
On the other hand, we will say that GΛ is the natural basis for Λ, and we
will denote it as G∗
Λ, when it is constituted by elements in Λ. That is, firstly if
λ1 6= 0 then g1 := λ1 ∈ G∗
Λ. Secondly, if {λ1, λ2} are Q-rationally independent,
then g2 := λ2 ∈ G∗
Λ. Otherwise, if {λ1, λ3} are Q-rationally independent, then
g2 := λ3 ∈ G∗
Λ then rj,mj = 1 and rj,k = 0
for k 6= mj, where mj is such that gmj = λj. In fact, each element in G∗
Λ
is of the form gmj for j such that λj is Q-linear independent of the previous
elements in the basis. Furthermore, if λj /∈ G∗
k=1 rj,kgk, where
{g1, g2, . . . , gij} ⊂ {λ1, λ2, . . . , λj−1}.
will be used later.
Proposition 2 Given Λ = {λ1, λ2, . . . , λj, . . .} a set of exponents, consider
A1(p) and A2(p) two exponential sums in the class SΛ, say A1(p) =Pj≥1 ajeλj p
and A2(p) =Pj≥1 bjeλj p. Fixed the natural basis G∗
Λ = {g1, g2, . . . , gk, . . .} for
Λ, for each j ≥ 1 let rj ∈ R♯G∗
Λ be the vector of rational components verifying
∼A2 if and only if there exists x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈
(4). Then A1
[0, 2π)♯G∗
Λ such that for each j = 1, 2, . . . it is satisfied bj = aje<rj ,x0+pj >i for
some pj = (2πnj,1, 2πnj,2, . . .) ∈ R♯G∗
Λ} and
Proof Suppose that A1
In = {1, 2, . . . , k, . . . , n : λk ∈ G∗
Λ}. Let j ∈ I, then rj,mj = 1 and rj,k = 0
for k 6= mj, where mj is such that gmj = λj. Thus, by Proposition 1, let
xj = (xj,1, xj,2, . . .) ∈ R♯G∗
∼A2. Consider I = {1, 2, . . . , k, . . . : λk ∈ G∗
Λ, with nj,k ∈ Z.
Λ be a vector such that
∗
∗
ij
k=1 rj,kxj,k = ajeirj,mj xj,mj = ajeixj,mj .
bj = ajei<rj ,xj> = ajei P
Define x0 = (x0,1, x0,2, . . .) ∈ R♯G∗
Λ = R♯I as x0,mj := xj,mj for j ∈ I. Thus, by
taking pj = (0, 0, . . .), the result trivially holds for those j's such that λj ∈ G∗
Λ,
i.e. for j ∈ I. Now, let j be such that λj /∈ G∗
Λ, i.e. j /∈ I. By Proposition 1,
let xj = (xj,1, xj,2, . . .) ∈ R♯G∗
Λ be a vector such that
bp = apei<rp,xj> = apei P
ij
k=1 rp,kxj,k , p = 1, 2, . . . , j.
6
J.M. Sepulcre, T. Vidal
Note that if p = 1, 2, . . . , j is such that λp ∈ G∗
bp = apeirp,mp xj,mp ,
Λ, then
which necessarily implies that rp,mp xj,mp = rp,mp xp,mp + 2πnp, i.e. xj,mp =
xp,mp + 2πnj,p for some nj,p ∈ Z. Hence
bj = ajei<rj ,xj > = ajei P
ij
k=1 rj,kxj,k = aje
i Pp∈Ij−1
rj,mp xj,mp =
i Pp∈Ij−1
rj,mp (xp,mp +2πnj,p)
aje
= ajei<rj ,x0+pj >,
Λ without loss of generality.
where pj = (2πnj,1, 2πnj,2, . . . , 0, 0, . . .). Moreover, by changing conveniently
the vectors pj, we can take x0 ∈ [0, 2π)♯G∗
Conversely, suppose the existence of x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈ R♯G∗
satisfying bj = aje<rj ,x0+pj>i for some pj = (2πnj,1, 2πnj,2, . . .) ∈ R♯G∗
Λ,
with nj,k ∈ Z. Let rj,k = pj,k
with pj,k and qj,k coprime integer numbers,
and define qn,k := lcm(q1,k, q2,k, . . . , qn,k) for each k = 1, 2, . . .. Thus, for any
integer number n ≥ 1, take xn = x0+mn, where mn,k = 2πp1,kp2,k ··· pn,kqn,k,
k = 1, 2, . . .. Therefore, it is satisfied bj = aje<rj,xn>i for each j = 1, 2, . . . , n,
which implies that A1
qj,k
Λ
∗
∼A2.
As corollary, we can formulate the following result.
Corollary 1 Given Λ = {λ1, λ2, . . . , λj, . . .} a set of exponents, consider A1(p)
and A2(p) two exponential sums in the class SΛ, say A1(p) = Pj≥1 ajeλj p
and A2(p) = Pj≥1 bjeλj p. Fixed a basis GΛ = {g1, g2, . . . , gk, . . .} for Λ,
for each j ≥ 1 let rj ∈ R♯GΛ be the vector of rational components verifying
∼A2 if and only if there exists x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈
(4). Then A1
[0, 2π)♯GΛ such that for each j = 1, 2, . . . it is satisfied bj = aje<rj,x0+qj >i
for some qj ∈ R♯GΛ which are of the form qj = T · pt
j, where T is the
change of basis matrix, with respect to the natural basis, and pj is of the
form (2πnj,1, 2πnj,2, . . . , 2πnj,k, . . .), nj,k ∈ Z.
∗
In particular, note that the coefficients of equivalent exponential sums have
the same modulus.
3 The finite exponential sums of the classes PR,Λ
This section is focused on the following classes of finite exponential sums.
Definition 3 Let Λ = {λ1, . . . , λn} be a set of n ≥ 1 distinct real numbers.
We will say that a function f : R 7→ C is in the class PR,Λ if it is of the form
(5)
f (t) = a1eiλ1t + . . . + aneiλnt, aj ∈ C, λj ∈ Λ, j = 1, . . . , n.
The functions f (t) of type (5) are also called trigonometric polynomials.
Note that Definition 2 can be particularized to the classes PR,Λ. Further-
more, if Λ is finite it is clear that it is always possible to find an integral basis
for Λ. In this context, we next prove the following important result.
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
7
Theorem 1 Given Λ = {λ1, λ2, . . . , λn} a set of n ≥ 1 exponents, let a1eiλ1t+
. . . + aneiλnt and b1eiλ1t + . . . + bneiλnt be two equivalent functions in the class
PR,Λ. Fixed d > 0 and ε > 0, there exists τ > d such that
n
Xj=1
ajeiλj τ − bj < ε.
Proof Let GΛ = {g1, . . . , gm}, for a certain m ≥ 1, be linearly independent
over the rationals so that each λj ∈ Λ is expressible as a linear combination
of its terms, say
λj =
m
Xk=1
rj,kgk, for some rj,k =
pj,k
qj,k ∈ Q, j = 1, 2, . . . , n.
(6)
Consider ε > 0, q := lcm(qj,k : j = 1, . . . , n, k = 1, . . . , m), r := max{rj,k :
j = 1, . . . , n, k = 1, . . . , m} > 0 and a := max{aj : j = 1, 2, . . . , n} > 0. Since
a1eiλ1t + . . . + aneiλnt and b1eiλ1t + . . . + bneiλnt are equivalent, Proposition
1 assures the existence of a vector of real numbers x0 = (x0,1, x0,2, . . . , x0,m)
such that
bj = aje<rj ,x0>i = ajei Pm
k=1 rj,kx0,k , j = 1, 2, . . . , n.
(7)
Now, as the numbers ck = gk
2πq , k = 1, 2, . . . , m, are rationally independent,
we next apply Kronecker's theorem [9, p.382] with the following choice: ck,
ε1 =
2πq , k = 1, 2, . . . , m. In this manner we assure
the existence of a real number τ > d > 0 and integer numbers e1, e2, . . . , em
such that
a·m·n·r·E > 0 and dk = x0,k
ε
that is
τ ck − ek − dk =(cid:12)(cid:12)(cid:12)(cid:12)
τ gk
2πq − ek −
< ε1,
x0,k
2πq(cid:12)(cid:12)(cid:12)(cid:12)
τ gk = 2πqek + x0,k + ηk, with ηk < ε1.
Therefore, from (6) and (7), with t ∈ R, we have
(8)
=
ajeiλj τ − ajei Pm
n
Xj=1(cid:12)(cid:12)(cid:12)
n
ajeiλj τ − bj =
Xj=1
aj(cid:12)(cid:12)(cid:12)
eiτ λj − ei Pm
n
n
Xj=1
k=1 rj,kx0,k(cid:12)(cid:12)(cid:12) ≤ a
eiτ Pm
k=1 rj,kgk − ei Pm
n
eiτ λj − ei Pm
Xj=1(cid:12)(cid:12)(cid:12)
k=1 rj,kx0,k(cid:12)(cid:12)(cid:12)
,
a
Xj=1(cid:12)(cid:12)(cid:12)
k=1 rj,kx0,k(cid:12)(cid:12)(cid:12) ≤
k=1 rj,kx0,k(cid:12)(cid:12)(cid:12)
which, from (8), is equal to
a
n
Xj=1(cid:12)(cid:12)(cid:12)
ei Pm
k=1(rj,k2πqek+rj,kx0,k+rj,kηk) − ei Pm
=
k=1 rj,kx0,k(cid:12)(cid:12)(cid:12)
8
J.M. Sepulcre, T. Vidal
a
n
Xj=1(cid:12)(cid:12)(cid:12)
anr
ei Pm
k=1 rj,kηk − 1(cid:12)(cid:12)(cid:12) ≤ a
Xk=1
Xk=1
ηk < anr
m
m
n
Xj=1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ε
≤
m
Xk=1
rj,kηk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= ε.
a · m · n · r
As an immediate consequence of Theorem 1, we obtain the following corol-
lary (compare with [11, Corollary 3]).
Corollary 2 Given Λ = {λ1, λ2, . . . , λn} a finite set of exponents, let f1(t) =
Pn
j=1 ajeiλj t and f2(t) =Pn
j=1 bjeiλj t be two equivalent functions in the class
PR,Λ. Fixed ε > 0, there exists a relatively dense set of real numbers τ such
that
f1(t + τ ) − f2(t) < ε ∀t ∈ R.
Proof Fixed τ > 0, note that for any t ∈ R it is accomplished that
f1(t + τ ) − f2(t) ≤
n
Xj=1
ajeiλj (t+τ ) − bjeiλj t =
n
Xj=1
ajeiλj τ − bj.
Thus, by Theorem 1 and given d > 0, there exists τ1 > d such that
f1(t + τ1) − f2(t) < ε/2 ∀t ∈ R.
(9)
Moreover, since f1(t) is almost periodic, there exists a real number l = l(ε)
such that every interval of length l contains at least one translation number
τ , associated with ε, satisfying
f1(t + τ ) − f1(t) ≤ ε/2 for all t ∈ R.
(10)
Consequently, from (9) and (10) we deduce the existence of a relatively dense
set of real numbers τ such that any t ∈ R satisfies
f1(t + τ + τ1) − f2(t) ≤ f1(t + τ1 + τ ) − f1(t + τ1) + f1(t + τ1) − f2(t) < ε.
This proves the result.
It was proved in [11, Proposition 2] that, with respect to the topology
of uniform convergence, the equivalence classes in PR,Λ/
∼ are sequentially
compact. We can analogously prove that this property is also true with respect
to the topology of B2(R, C) (see the proof in [11, Proposition 2]).
∗
Proposition 3 Let Λ be a finite set of exponents and G an equivalence class
in PR,Λ/
∼. Thus G is sequentially compact.
∗
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
9
4 Besicovitch almost periodic functions in terms of an equivalence
relation
For our purposes, we next focus our attention on the Besicovitch space B(R, C),
whose functions are obtained by the completion of the trigonometric polynomi-
−l f (t)dt(cid:17) (see for example
als with respect to the seminorm lim supl→∞(cid:16) 1
2lR l
[8, Section 3.4]). In particular, the space of functions B(R, C) contains those
of the space of the almost periodic functions AP (R, C) and those functions of
B2(R, C). We recall that every function in B(R, C) is associated with a real
its Fourier series.
exponential sum with real frequencies of the formPj≥1 ajeiλj t, which is called
Definition 4 Let Λ = {λ1, λ2, . . . , λj, . . .} be an arbitrary countable set of
distinct real numbers. We will say that a function f : R → C is in the class
FB2,Λ if it is an almost periodic function in B2(R, C) whose associated Fourier
series is of the form
(11)
ajeiλj t, aj ∈ C, λj ∈ Λ.
Xj≥1
It is worth noting that, in general, when we write that a function f is in
B(R, C) we do not have in mind the function f itself, it does represent a whole
class of equivalent functions according to the relation f1 ≃ f2 if and only if
lim sup
l→∞ 1
2l Z l
−l f (t) − g(t)2dt! = 0.
In terms of Definition 2, we can define an equivalence relation on the func-
tions in B(R, C), in particular on the classes FB2,Λ. More specifically, we es-
tablish the following definition.
Definition 5 Given Λ = {λ1, λ2, . . . , λj, . . .} a set of exponents, let f1 and f2
denote two equivalence classes of B(R, C)/ ≃ whose associated Fourier series
are given by
Xj≥1
ajeiλj t and Xj≥1
bjeiλj t, aj, bj ∈ C, λj ∈ Λ.
We will say that f1 is equivalent to f2 if for each integer value n ≥ 1 there
exists a Q-linear map ψn : Vn → R, where Vn is the Q-vector space generated
by {λ1, λ2, . . . , λn}, such that
bj = ajeiψn(λj ), j = 1, . . . , n.
In that case, we will write f1
∗
∼f2.
The next important lemma allows us to prove that if a function f2 is
equivalent (in the sense of Definition 5) to a function f1 belonging to the space
B2(R, C), then f2 also belongs to B2(R, C). This is clearly a consequence of
Riesz-Fischer theorem [3, p. 110].
10
J.M. Sepulcre, T. Vidal
Lemma 1 Let f1(t) ∈ B2(R, C) be an almost periodic function whose Fourier
series is given by Pj≥1 ajeiλj t, aj ∈ C, where {λ1, . . . , λj, . . .} is a set of dis-
tinct exponents. Consider bj ∈ C such that Pj≥1 bjeiλj t and Pj≥1 ajeiλj t are
equivalent. Then Pj≥1 bjeiλj t is the Fourier series associated with an almost
periodic function f2(t) ∈ B2(R, C) so that f1
Proof Take Λ = {λ1, . . . , λj , . . .}. By the hypothesis, f1 ∈ FB2,Λ ⊂ B2(R, C)
is determined by the series Pj≥1 ajeiλj t, aj ∈ C, λj ∈ Λ. Moreover, since
Pj≥1 ajeiλj t ∗
∼Pj≥1 bjeiλj t, we deduce from Corollary 1 that bj = aj for
j ≥ 1 and hence
∼f2.
∗
aj2 < ∞.
Xj≥1
bj2 =Xj≥1
By Riesz-Fischer theorem [3, p. 110], there exists a function f2 ∈ B2(R, C)
such that the values bn are the Fourier coefficients of f2.
As it was said before, it is worth noting that a Fourier seriesPn≥1 aneiλnt,
such that Pn≥1 an2 < ∞, represents an equivalence class (according to the
relation ≃) of functions in B2(R, C) (not a single function). In fact, as we
pointed out in introduction, since two almost periodic functions in the Besi-
covitch sense are connected in B2(R, C) when they have the same Fourier
series ([3, p. 148] or [8, Section 4.2]), we immediately deduce from the results
above the following corollary.
Corollary 3 Let f1(t) and f2(t) be two equivalent functions in B(R, C). If
f1(t) ∈ B2(R, C), then f2(t) ∈ B2(R, C).
The following result is concerned with the concept of convergence in B2(R, C)
which is certainly weaker than the uniform convergence. Under this topology,
∼ are closed. In fact, more
we next show that the equivalence classes of FB2,Λ/
specifically, they are sequentially compact.
∗
∗
∼. Thus G is sequentially compact.
Proposition 4 Let Λ be a set of exponents and G an equivalence class in
FB2,Λ/
∼. For
Proof Let {fl}l≥1 be a sequence in an equivalence class G in FB2,Λ/
each l = 1, 2, . . ., suppose that the Fourier series which is associated with fl(t)
is given by
∗
al,jeiλj t with al,j ∈ C, λj ∈ Λ.
Xj≥1
Fixed a basis GΛ = {g1, g2, . . . , gk, . . .} for Λ, let rj = (rj,1, rj,2, . . .) be the
vector satisfying < rj, g >= λj for each j ≥ 1, where g = (g1, g2, . . . , gk, . . .).
Since f1
∼fl for each l = 1, 2, . . ., we deduce from Proposition 1 that for each
integer value n ≥ 1 there exists xl,n = (xl,n,1, xl,n,2, . . .) ∈ R♯GΛ such that
∗
al,j = a1,jei<rj ,xl,n>, j = 1, 2 . . . , n.
(12)
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
11
Given l ≥ 1, let Pl,k(t) =Pj≥1 pj,kal,jeiλj t, k = 1, 2, . . ., be the Bochner-Fej´er
polynomials which converge to fl with respect to the topology of B2(R, C)
(and converge formally to its Fourier series on R) [3, p. 105, Theorem II]. It
is worth noting that for each k only a finite number of the factors pj,k dif-
fer from zero, and these factors pj,k do not depend on l [3, p. 48]. Thus,
by taking into account (12), it is clear that {Pl,1(t)}l≥1 is a sequence of
equivalent trigonometric polynomials and, by Proposition 3, there exists a
subsequence {Plm,1,1(t)}m≥1 ⊂ {Pl,1(t)}l≥1 convergent to a certain P1(t) =
Pj≥1 pj,1ajeiλj t ∈ PR,Λ1 , where Λ1 = {λj ∈ Λ : pj,1 6= 0}, which is in the
same equivalence class as P1,1(t). Furthermore, by Proposition 1, this means
that there exists x(1)
0,2, . . .) ∈ Rm1 such that
0 = (x(1)
0,1, x(1)
pj,1aj = pj,1a1,jei<rj ,x
(1)
0 >, j = 1, 2 . . . , with λj ∈ Λ1,
where m1 is the number of elements of any basis for Λ1. Equivalently
aj = a1,jei<rj ,x
(1)
0 >, j = 1, 2 . . . , with λj ∈ Λ1.
Analogously, from the sequence {Plm,1,2(t)}m≥1, we can draw a subsequence
{Plm,2,2(t)}m≥1 ⊂ {Plm,1,2(t)}m≥1 convergent to a certain
P2(t) =Xj≥1
pj,2ajeiλj t ∈ PR,Λ2 ,
where Λ2 = {λj ∈ Λ : pj,2 6= 0}∪ Λ1, which is in the same equivalence class as
P1,2(t). This implies that there exists x(2)
0,2, . . .) ∈ Rm2 such that
0 = (x(2)
0,1, x(2)
aj = a1,jei<rj ,x
(2)
0 >, j = 1, 2 . . . , with λj ∈ Λ2,
where m2 is the number of elements of any basis for Λ2. In general, for each k =
2, 3, . . ., we can extract a subsequence {Plm,k,k(t)}m≥1 ⊂ {Plm,k−1,k(t)}m≥1
convergent to a certain
Pk(t) =Xj≥1
pj,kajeiλj t ∈ PR,Λk ,
where Λk = {λj ∈ Λ : pj,k 6= 0} ∪ Λk−1, which is in the same equivalence
0 = (x(k)
class as P1,k(t) and hence there exists x(k)
0,2, . . .) ∈ Rmk (mk is the
number of elements of any basis for Λk) such that
0,1, x(k)
aj = a1,jei<rj ,x
(k)
0 >, j = 1, 2 . . . , with λj ∈ Λk.
(13)
So we get by induction a sequence {Pk(t)}k≥1 of trigonometric polynomials
which converges formally to the series
ajeiλj t, λj ∈ Λ,
Xj≥1
(14)
12
J.M. Sepulcre, T. Vidal
and, since (13) is satisfied for any k = 1, 2, . . ., we can construct, for each
integer value n ≥ 1, a vector x0,n ∈ R♯GΛ such that
aj = a1,jei<rj ,x0,n>, j = 1, 2 . . . , n with λj ∈ Λ.
Hence the series (14) is equivalent toPj≥1 a1,jeiλj t and, by Lemma 1, it is the
Fourier series associated with an almost periodic function h(t) ∈ B2(R, C) such
that h
∼f1. Consequently, {Pk(t)}k≥1 converges with respect to the topology
of B2(R, C) to h(t) ∈ G and we can extract a subsequence of {fl(t)}l≥1 which
also converges in B2(R, C) to h(t).
∗
As a consequence of Proposition 4, in the topology of B2(R, C), we next
show that the family of translates of a function f ∈ FB2,Λ is closed on its
equivalence class of FB2,Λ/
Corollary 4 Let Λ be a set of exponents and f ∈ FB2,Λ. Thus the limit points
of the set of functions Tf = {fτ (t) := f (t + τ ) : τ ∈ R} are functions which
are equivalent to f .
∼.
∗
Proof Since it is plain that the functions included in Tf = {fτ (p) := f (t + τ ) :
τ ∈ R} are in the same equivalence class of f (see in [8, Section 4.2] the
Fourier series of the translates of a function in the Besicovitch spaces), the
result follows easily from Proposition 4.
Now Corollary 4 can be improved with the following result. Indeed, we
next prove that, fixed a function f ∈ FB2,Λ, the limit points of the set of
the translates Tf = {f (t + τ ) : τ ∈ R} of f are precisely the almost periodic
functions which are equivalent to f .
∗
∼
Theorem 2 Let Λ be a set of exponents, G an equivalence class in FB2,Λ/
and f ∈ G. Thus the set of functions Tf = {fτ (t) := f (t + τ ) : τ ∈ R} is dense
in G.
Proof Let f (t) be a function in the class FB2,Λ. We know by Corollary 4 that
the limit points of the set of functions Tf = {fτ (t) := f (t + τ ) : τ ∈ R}
are functions in B2(R, C) which are equivalent to f . We next demonstrate
that any function h(t) which is equivalent to f (t) is also a limit point of Tf .
If ♯Λ < ∞, given εn = 1
n , n ∈ N, Corollary 2 assures the existence of an
increasing sequence {τn}n≥1 of positive real numbers such that any n ∈ N
verifies
f (t + τn) − h(t)2 < εn ∀t ∈ R.
Hence M (fτn(t)− h(t)2) → 0 as n goes to ∞ (see (2) for the definition of the
mean value M (f )), and the result holds for the case ♯Λ < ∞. Consider ♯Λ = ∞
and let Pj≥1 ajeiλj t and Pj≥1 bjeiλj t be the Fourier series of f ∈ FB2,Λ and
∼f , respectively. Take ε1 = Pj>1 aj2 > 0, then Theorem 1 assures the
h
existence of τ1 > 0 such that
∗
(cid:12)(cid:12)(cid:12)
a1eiλ1(t+τ1) − b1eiλ1t(cid:12)(cid:12)(cid:12)
< √ε1 ∀t ∈ R,
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
13
which implies
< ε1.
(15)
2
(cid:12)(cid:12)a1eiλ1τ1 − b1(cid:12)(cid:12)
Thus, from (15) and aj = bj for any j ≥ 1 (Corollary 1), we have that
Xj≥1
(aj + bj)2 =
ajeiλ1τ1 − bj2 < ε1 +Xj>1
ajeiλj τ1 − bj2 ≤ ε1 +Xj>1
Consequently,
aj2 = 5ε1.
ε1 + 4Xj>1
M (fτ1(t) − h(t)2) < 5ε1.
Similarly, take ε2 = Pj>2 aj2 > 0, then Theorem 1 assures the existence of
τ2 > τ1 such that
2
which implies
< √ε2,
Xj=1(cid:12)(cid:12)(cid:12)
ajeiλj (t+τ2) − bjeiλj t(cid:12)(cid:12)(cid:12)
Xj=1(cid:12)(cid:12)ajeiλj τ2 − bj(cid:12)(cid:12)
2
2
< ε2.
(16)
Therefore, from (16) and aj = bj for any j ≥ 1, we have
Xj≥1
ajeiλ1τ2 − bj2 = a1eiλ1τ2 − b12 + a2eiλ1τ2 − b22 +Xj>2
ajeiλj τ2 − bj2 ≤
(a1eiλ1τ2 − b1 + a2eiλ1τ2 − b2)2 +Xj>2
(aj + bj)2 = ε2 + 4Xj>2
M (fτ2(t) − h(t)2) < 5ε2.
≤ ε2 +Xj>2
ajeiλj τ2 − bj2 ≤
aj2 = 5ε2.
Consequently,
In general, by repeating this process, we can construct an increasing sequence
{τn}n≥1 such that each τn satisfies that
which implies
n
Xj=1(cid:12)(cid:12)(cid:12)
ajeiλj (t+τn) − bjeiλj t(cid:12)(cid:12)(cid:12)
Xj=1(cid:12)(cid:12)ajeiλj τn − bj(cid:12)(cid:12)
n
2
< √εn,
< εn.
(17)
14
J.M. Sepulcre, T. Vidal
with εn =Pj>n aj2. Thus, from (17) we have
ajeiλ1τn − bj2 =
ajeiλj τn − bj2 ≤
n
M (fτn(t) − h(t)2) =Xj≥1
ajeiλj τn − bj2 +Xj>n
Xj=1
ajeiλj τn − bj
2
+Xj>n
n
Xj=1
≤ εn +Xj>n
ajeiλj τn − bj2 ≤
(aj + bj)2 = εn + 4Xj>n
aj2 = 5εn.
Note that Pj≥1 aj2 < ∞, then Pj>n aj2 tends to 0 when n goes to ∞.
Consequently, the sequence of functions {f (t + τn)}n≥1 converges in B2(R, C)
to h(t), and the result holds.
Corollary 5 Let f ∈ B2(R, C) and f1
∼f . There exists an increasing un-
bounded sequence {τn}n≥1 of positive numbers such that the sequence of func-
tions {f (t + τn)}n≥1 converges in B2(R, C) to f1(t). In fact, given ε > 0 there
exists a satisfactorily uniform set of positive numbers τ such that
∗
M (f (t + τ ) − f1(t)2) < ε.
Proof Let f ∈ B2(R, C), then f ∈ FB2,Λ for some set Λ of exponents. Let G
be the equivalence class in FB2,Λ/
∼f . Thus, by
Theorem 2 (see also its proof), there exists an increasing unbounded sequence
{τn}n≥1 of positive numbers such that the sequence of functions {f (t+τn)}n≥1
converges in B2(R, C) to f1(t). Equivalently, given ε > 0 there exists n0 ∈ N
such that
∼ so that f ∈ G and let f1
∗
∗
M (f (t + δn) − f1(t)2) < ε/2 ∀n ≥ n0.
Moreover, since f (t) is almost periodic in the sense of Besicovitch, there exist
a set S = {τk} ⊂ R and l = l(ε) > 0 such that the ratio of the maximum
number of elements of S included in an interval (a, a + l) to the minimum
number is less than 2 and satisfy
M (f (t + τk) − f (t)2) < ε/2.
Hence any τk satisfies
M (f (t + δn + τk) − f1(t)2) ≤ M (f (t + δn + τk) − f (t + δn)2)+
+ M (f (t + δn) − f1(t)2) < ε ∀n ≥ n0,
which proves the result.
Bohr's equivalence relation in the space of Besicovitch almost periodic functions
15
It is known that the almost periodic functions in the Besicovitch spaces
Bp(R, C), 1 ≤ p < ∞, satisfy the Bochner-type property consisting of the
relative compactness of the set {f (t + τ )}, τ ∈ R, associated with an arbitrary
function f ∈ Bp(R, C) (see [1, Theorem 5.34] or [8, Section 3.4]). As an im-
portant consequence of Theorem 2, we next refine this property for the case
of B2(R, C) in the sense that we show that the condition of almost periodic-
ity, in the sense of Besicovitch, of a function f (t) implies that every sequence
{f (t + τn)}, τn ∈ R, of translates of f has a subsequence that converges with
the topology of B2(R, C) to a function which is equivalent to f .
Corollary 6 If f ∈ B2(R, C), then the compact closure of its set of translates
coincides with its equivalence class.
∗
Proof First of all, we recall that any function f ∈ B(R, C) has an associated
Fourier series. Let f ∈ B2(R, C), then f ∈ FB2,Λ for some set Λ of exponents.
Now, let G be the equivalence class in FB2,Λ/
∼ so that f ∈ G. By Theorem
2, all the limit points of the translates of f are exponential sums which are
included in G and, in fact, the compact closure of the set of the translates of
f coincides with G.
Remark 1 Given Λ = {λ1, λ2, . . . , λj , . . .} a set of exponents, consider A1(p)
and A2(p) two exponential sums in the class SΛ, say A1(p) =Pj≥1 ajeλj p and
A2(p) = Pj≥1 bjeλj p. Let V be the Q-vector space generated by Λ. We will
say that A1 is B-equivalent to A2 if there exists a Q-linear map ψn : V → R
such that
bj = ajeiψn(λj ), j = 1, 2, . . . .
It is easy to prove that, fixed a basis GΛ for Λ, A1 is B-equivalent to A2
if and only if there exists x0 = (x0,1, x0,2, . . . , x0,k, . . .) ∈ R♯GΛ such that
bj = aje<rj,x0>i for every j ≥ 1, where the rj's are the vectors of rational
components verifying (4).
From this and Proposition 1, it is worth noting that Definition 2 and defi-
nition of B-equivalence are equivalent in the case that it is possible to obtain
an integral basis for the set of exponents Λ. Consequently, all the results of
this paper which can be formulated in terms of an integral basis are also valid
under the B-equivalence. (in particular, those related to the finite exponential
sums in Section 3).
References
1. J. Andres, A.M. Bersani, R.F. Grande, Rendiconti di Matematica, Serie VII, 26 (2006),
121-188.
2. T.M. Apostol, Modular functions and Dirichlet series in number theory, Springer-Verlag,
New York, 1990.
3. A.S. Besicovitch, Almost periodic functions, Dover, New York, 1954.
4. S. Bochner, A new approach to almost periodicity, Proc. Nat. Acad. Sci., 48 (1962),
2039-2043.
5. H. Bohr, Almost periodic functions, Chelsea, New York, 1951.
16
J.M. Sepulcre, T. Vidal
6. H. Bohr, Contribution to the theory of almost periodic functions, Det Kgl. danske Viden-
skabernes Selskab. Matematisk-fisiske meddelelser. Bd. XX. Nr. 18, Copenhague, 1943.
7. H. Bohr, E. Foelner, On some types of Functional Spaces, Acta Math., 76 (1945), 31-155.
8. C. Corduneanu, Almost Periodic Oscillations and Waves, Springer, New York, 2009.
9. G.H. Hardy, E.M. Wright, An Introduction to the Theory of Numbers, Oxford Science,
Oxford, 1979.
10. B. Jessen, Some aspects of the theory of almost periodic functions, in Proc. Internat.
Congress Mathematicians Amsterdam, 1954, Vol. 1, North-Holland, 1954, pp. 304–351.
11. J.M. Sepulcre, T. Vidal, Almost Periodic Functions in terms of Bohr's Equivalence
Relation, Ramanujan J., DOI: 10.1007/s11139-017-9950-1, 2017. Corrigendum sent to
the journal. See also arXiv: 1801.08035 [math.CV].
|
1711.04470 | 1 | 1711 | 2017-11-13T08:47:17 | New variations of power increasing sequences | [
"math.FA"
] | The aim of this paper is to generalize a main theorem concerning weighted mean summability to absolute matrix summability which plays a vital role in summability theory and applications to the other sciences by using quasi-$f$-power sequences. | math.FA | math |
New variations of power increasing sequences
S¸ebnem Yıldız
Abstract. The aim of this paper is to generalize a main theorem concerning
weighted mean summability to absolute matrix summability which plays a
vital role in summability theory and applications to the other sciences by
using quasi-f -power sequences.
1. Introduction
Definition 1.1. [1] A positive sequence (bn) is said to be an almost increas-
ing sequence if there exists a positive increasing sequence (cn) and two positive
constants M and N such that M cn 6 bn 6 N cn.
Definition 1.2. [19] A positive sequence X = (Xn) is said to be quasi-f -power
increasing sequence if there exists a constant K = K(X, f ) > 1 such that KfnXn >
fmXm for all n > m > 1, where f = {fn(σ, β)} =(cid:8)nσ(logn)β, β > 0, 0 < σ < 1(cid:9).
Definition 1.3. The sequence (λn) is said to be of bounded variation, denoted
∞
by (λn) ∈ BV, if
∆λn < ∞. If we take β = 0, then we have a quasi-σ-power
Pn=1
increasing sequence. Every almost increasing sequence is a quasi-σ-power increasing
sequence for any non-negative σ, but the converse is not true for σ > 0 (see [13]).
For any sequence (λn) we write that ∆2λn = ∆λn − ∆λn+1 and ∆λn = λn − λn+1.
n and tα
n
we denote the nth Ces`aro means of order α, with α > −1, of the sequence (sn) and
(nan), respectively, that is (see [8])
Let P an be a given infinite series with the partial sums (sn). By uα
(1.1)
where
(1.2)
uα
n =
1
Aα
n
n
Xv=0
Aα−1
n−vsv
and tα
n =
1
Aα
n
n
Xv=0
Aα−1
n−vvav,
Aα
n =
(α + 1)(α + 2)...(α + n)
n!
= O(nα),
Aα
−n = 0 for n > 0.
2010 Mathematics Subject Classification. 26D15; 42A24; 40F05; 40G99.
Key words and phrases. Summability factors, absolute matrix summability, Fourier series,
infinite series, Holder inequality, Minkowski inequality.
1
2
YILDIZ
k > 1, if
Definition 1.4. [10],[12] The series P an is said to be summable C, αk,
(1.3)
nk−1uα
n − uα
n−1k =
tα
nk < ∞.
1
n
∞
Xn=1
∞
Xn=1
Xv=0
If we take α = 1, then C, αk summability reduces to C, 1k summability.
Let (pn) be a sequence of positive real numbers such that
n
(1.4)
Pn =
pv → ∞ as n → ∞,
(P−i = p−i = 0,
i > 1).
The sequence-to-sequence transformation
(1.5)
tn =
1
Pn
pvsv
n
Xv=0
defines the sequence (tn) of the Riesz mean or simply the ( ¯N , pn) mean of the
sequence (sn) generated by the sequence of coefficients (pn) (see [11]).
Definition 1.5. [2] The series P an is said to be summable (cid:12)(cid:12)
∞
¯N , pn(cid:12)(cid:12)k, k > 1,
if
(1.6)
Xn=1(cid:18) Pn
pn(cid:19)k−1
tn − tn−1 k< ∞.
summability is the same as C, 1k (resp. ¯N , pn ) summability.
In the special case when pn = 1 for all values of n (resp. k = 1), (cid:12)(cid:12)
The following theorems are known dealing with the (cid:12)(cid:12)
¯N , pn(cid:12)(cid:12)k
¯N , pn(cid:12)(cid:12)k summability fac-
Theorem 2.1. [14] Let (Xn) be a almost increasing sequence. If the sequences
2. The Known Results
tors of infinite series.
(Xn), (λn), and (pn) satisfy the conditions
m
Xn=1
(2.1)
(2.2)
(2.3)
(2.4)
(2.5)
λmXm = O(1) as m → ∞,
nXn∆2λn = O(1) as m → ∞,
Pn
n
= O(Pm)
m
Xn=1
m
pn
Pn
m
Xn=1
Xn=1
tnk = O(Xm) as m → ∞,
tnk
n
= O(Xm) as m → ∞,
then the series P anλn is summable (cid:12)(cid:12)
¯N , pn(cid:12)(cid:12)k, k > 1.
NEW VARIATIONS OF POWER INCREASING SEQUENCES
3
Theorem 2.2. [6] Let (Xn) be a quasi-σ-power increasing sequence.
If the
sequences (Xn), (λn) and (pn) satisfy the conditions (2.1)-(2.3), and
(2.6)
(2.7)
pn
Pn
tnk
X k−1
n
= O(Xm) as m → ∞,
tnk
nX k−1
n
= O(Xm) as m → ∞,
m
m
Xn=1
Xn=1
then the series P anλn is summable (cid:12)(cid:12)
¯N , pn(cid:12)(cid:12)k, k > 1.
creasing sequence instead of a quasi-σ-power increasing sequence.
Later on, Bor has proved the following theorem by taking quasi-f-power in-
Theorem 2.3. [7] Let (Xn) be a quasi-f -power increasing sequence.
If the
sequences (Xn), (λn) and (pn) satisfy all the conditions of Theorem 2.2, then the
series P anλn is summable (cid:12)(cid:12)
¯N , pn(cid:12)(cid:12)k, k > 1.
Let A = (anv) be a normal matrix, i.e., a lower triangular matrix of nonzero
diagonal entries. Then A defines the sequence-to-sequence transformation, mapping
the sequence s = (sn) to As = (An(s)), where
(2.8)
An(s) =
anvsv, n = 0, 1, ...
n
Xv=0
if
Definition 2.1. [18] The series P an is said to be summable A, pnk, k > 1,
(2.9)
where
(2.10)
∞
Xn=1(cid:18) Pn
pn(cid:19)k−1
k
< ∞,
(cid:12)(cid:12)
¯∆An(s)(cid:12)(cid:12)
¯∆An(s) = An(s) − An−1(s).
If we take pn = 1 for all values of n, then we have Ak summability (see [20]).
And also if we take anv = pv
Pn
we take anv = pv
Pn
to C, 1k summability (see [10]).
, then we have (cid:12)(cid:12)
¯N , pn(cid:12)(cid:12)k summability. Furthermore, if
and pn = 1 for all values of n, then A, pnk summability reduces
3. The Main Results
The Fourier series play an important role in many areas of applied mathematics
and mechanics. Recently some papers have been done concerning absolute matrix
summability of infinite series and Fourier series (see [3]-[5], [15]-[17], [21]-[23]).
The aim of this paper is to generalize Theorem 2.3 for A, pnk summability method
for these series by taking quasi-f-power increasing sequence instead of a quasi-σ-
power increasing sequence.
4
YILDIZ
Given a normal matrix A = (anv), we associate two lower semimatrices ¯A = (¯anv)
and A = (anv) as follows:
(3.1)
and
(3.2)
¯anv =
n
Xi=v
ani, n, v = 0, 1, ...
a00 = ¯a00 = a00,
anv = ¯anv − ¯an−1,v, n = 1, 2, ...
It may be noted that ¯A and A are the well-known matrices of series-to-sequence
and series-to-series transformations, respectively. Then, we have
(3.3)
and
(3.4)
An(s) =
anvsv =
n
Xv=0
¯anvav
n
Xv=0
¯∆An(s) =
anvav.
n
Xv=0
Using this notation we have the following theorem.
Theorem 3.1. Let (Xn) be a quasi-f -power increasing sequence. Let k > 1
and A = (anv) be a positive normal matrix such that
(3.5)
(3.6)
(3.7)
(3.8)
ano = 1, n = 0, 1, ...,
an−1,v > anv, for n > v + 1,
Pn(cid:19)
ann = O(cid:18) pn
an,v+1 = O(ann).
1
v
n−1
Xv=1
If the sequences (Xn), (λn) and (pn) satisfy all the conditions of Theorem 2.3, then
the series P anλn is summable A, pnk, k > 1.
It may be remarked that if we take A = ( ¯N , pn), the conditions (3.5)-(3.7) are
satisfied automatically and the condition (3.8) is satisfied by the condition (2.3).
We need the following lemmas for the proof of our theorem.
Lemma 3.1. [3] Under the conditions of Theorem 2.1 we have that
(3.9)
(3.10)
nXn∆λn = O(1) as n → ∞,
Xn∆λn < ∞.
∞
Xn=1
NEW VARIATIONS OF POWER INCREASING SEQUENCES
5
Proof of Theorem 3.1
Let Xn be a be a quasi-f -power increasing sequence and (In) denotes the A-
transform of the series P∞
n=1 anλn. Then, we have
¯∆In =
n
Xv=1
anvavλv.
Applying Abel's transformation to this sum, we have that
anvavλv
v
v
=
n−1
Xv=1
∆(
anvλv
v
)
v
Xr=1
rar +
annλn
n
rar
n
Xr=1
∆(
anvλv
v
)(v + 1)tv + annλn
n + 1
n
tn
¯∆In =
=
=
n
n−1
Xv=1
Xv=1
Xv=1
n−1
¯∆anvλvtv
v + 1
v
+
n−1
Xv=1
an,v+1∆λvtv
v + 1
v
+
n−1
Xv=1
an,v+1λv+1
tv
v
+ annλntn
n + 1
n
= In,1 + In,2 + In,3 + In,4.
To complete the proof of Theorem 3.1, by Minkowski's inequality, it is sufficient to
show that
(3.11)
∞
Xn=1(cid:18) Pn
pn(cid:19)k−1
In,r k< ∞,
for
r = 1, 2, 3, 4.
First, by applying Holder's inequality with indices k and k′, where k > 1 and
1
k + 1
k′ = 1, we have that
m+1
Xn=2(cid:18) Pn
= O(1)
m+1
v + 1
pn(cid:19)k−1
Xn=2(cid:18) Pn
In,1 k6
¯∆anv(cid:12)(cid:12) λvtv)k
pn(cid:19)k−1(n−1
Xn=2(cid:18) Pn
Xv=1
(cid:12)(cid:12)
¯∆anv(cid:12)(cid:12))k−1
¯∆anv(cid:12)(cid:12) λvktvk ×(n−1
pn(cid:19)k−1 n−1
Xv=1(cid:12)(cid:12)
Xv=1(cid:12)(cid:12)
m+1
v
,
using
∆anv = anv − an,v+1 = ¯anv − ¯an−1,v − ¯an,v+1 + ¯an−1,v+1 = anv − an−1,v,
and from (3.5) and (3.6) we have
¯∆anv =
n−1
Xv=1
=
n−1
Xv=1
Xv=0
n
Xv=1
Xv=0
n−1
n−1
anv − an−1,v =
(an−1,v − anv)
an−1,v − an−1,0 −
anv + an0 + ann
= 1 − an−1,0 − 1 + an0 + ann 6 ann,
6
YILDIZ
n=v+1 ¯∆anv 6 avv we have,
pn(cid:19)k−1
ak−1
nn (n−1
Xv=1
¯∆anvλvktvk)
In,1 k= O(1)
λvk−1λvtvk
m+1
Xn=2(cid:18) Pn
Xn=v+1
m+1
¯∆anv
m−1
λvtvkavv = O(1)
∆λv
Xv=1
arr
trk
X k−1
r
v
Xr=1
+ O(1)λm
avv
tvk
X k−1
v
m
Xv=1
∆λvXv + O(1)λmXm = O(1) as m → ∞,
by virtue of the hypotheses of Theorem 3.1 and Lemma 3.1. Also, we have that
m
m+1
= O(1)
and using Pm+1
pn(cid:19)k−1
Xn=2(cid:18) Pn
Xv=1
Xv=1
Xv=1
= O(1)
= O(1)
m−1
1
m
v
X k−1
m+1
Xn=2(cid:18) Pn
= O(1)
= O(1)
= O(1)
m+1
m+1
m+1
pn(cid:19)k−1
Xn=2(cid:18) Pn
Xn=2(cid:18) Pn
Xn=2(cid:18) Pn
Xv=1
Xv=1
Xv=1
Xv=1
m−1
m−1
m−1
m
an,v+1∆λvtv)k
an,v+1∆λvtv
v
Xv
v + 1
pn(cid:19)k−1(n−1
Xv=1
Xv)k
tvk) ×(n−1
Xv=1
1
X k
v
an,v+1∆λvXv
an,v+1∆λvXv
m+1
In,2 k6
Xn=2(cid:18) Pn
pn(cid:19)k−1(n−1
Xv=1
pn(cid:19)k−1(n−1
Xv=1
nn (n−1
pn(cid:19)k−1
Xv=1
Xn=v+1
X k−1
ak−1
tvk
1
v
m+1
1
v
1
X k
v
an,v+1∆λvXv)k−1
∆λvXv)k−1
tvk) ×(m−1
Xv=1
Xv=1
Xr=1
trk
rX k−1
vX k−1
v∆λv
tvk
1
m
m
v
r
= O(1)
v∆λv
an,v+1 = O(1)
= O(1)
∆(v∆λv)
trk
rX k−1
r
+ O(1)m∆λm
v
Xr=1
= O(1)
∆(v∆λv)Xv + O(1)m∆λmXm
= O(1)
vXv∆2λv + O(1)
= O(1)
as m → ∞,
Xv∆λv + O(1)m∆λmXm
m−1
Xv=1
Xn=1(cid:18) Pn
pn(cid:19)k−1
In,4k = O(1)
= O(1)
pn(cid:19)k−1
1
X k−1
n
Xn=1(cid:18) Pn
Xn=1
ann
m
ak
nnλnktnk = O(1)
ak−1
nn annλnk−1λntnk
Xn=1(cid:18) Pn
pn(cid:19)k−1
λntnk = O(1) as m → ∞,
NEW VARIATIONS OF POWER INCREASING SEQUENCES
7
by virtue of the hypotheses of Theorem 3.1 and Lemma 3.1. Furthermore, as in
In,1, we have
m+1
Xn=2(cid:18) Pn
= O(1)
= O(1)
= O(1)
an,v+1λv+1
m+1
m+1
In,3 k6
pn(cid:19)k−1
Xn=2(cid:18) Pn
Xn=2(cid:18) Pn
Xv=1
tvk
pn(cid:19)k−1(n−1
Xn=2(cid:18) Pn
Xv=1
v ) ×(n−1
pn(cid:19)k−1(n−1
an,v+1λv+1k tvk
Xv=1
Xv=1
pn(cid:19)k−1
Xv=1
Xn=v+1
λv+1λv+1k−1 tvk
v
an,v+1 = O(1)
Xv=1
X k−1
λv+1
ak−1
nn
m+1
m+1
n−1
1
1
v
m
m
v
X k−1
v
1
v
tv
v )k
an,v+1)k−1
Xv=1
m
λv+1
tvk
v
an,v+1 = O(1)
= O(1) as m → ∞,
tvk
1
v
X k−1
v
λv+1
m+1
Xn=v+1
an,v+1
by virtue of the hypotheses of Theorem 3.1 and Lemma 3.1. Again, as in In,1, we
have that
m
m
m
by virtue of hypotheses of the Theorem 3.1 and Lemma 3.1. This completes the
proof of Theorem 3.1.
4. An application of absolute matrix summability to Fourier series
Let f be a periodic function with period 2π and integrable (L) over (−π, π).
Without any loss of generality the constant term in the Fourier series of f can be
taken to be zero, so that
(4.1)
where
a0 =
We write
1
π Z π
−π
∞
f (t) ∼
(ancosnt + bnsinnt) =
Xn=1
Cn(t).
∞
Xn=1
f (t)dt,
an =
1
π Z π
−π
f (t)cos(nt)dt,
bn =
1
π Z π
−π
f (t)sin(nt)dt.
(4.2)
(4.3)
φα(t) =
φ(t) =
1
2
{f (x + t) + f (x − t)} ,
(t − u)α−1φ(u) du,
(α > 0).
α
tα Z t
0
It is well known that if φ(t) ∈ BV(0, π), then tn(x) = O(1), where tn(x) is the
(C, 1) mean of the sequence (nCn(x)) (see [9]).
8
YILDIZ
Using this fact, Bor has obtained the following main result dealing with the trigono-
metric Fourier series.
Theorem 4.1. [6] Let (Xn) be a quasi-σ-power increasing sequence. If φ1(t) ∈
BV(0, π), and the sequences (pn), (λn), and (Xn) satisfy the conditions of Theorem
Theorem 4.2. [7] Let (Xn) be a quasi-f -power increasing sequence. If φ1(t) ∈
BV(0, π), and the sequences (pn), (λn), and (Xn) satisfy the conditions of Theorem
2.3, then the series P Cn(x)λn is summable ¯N , pnk, k > 1.
2.3, then the series P Cn(x)λn is summable ¯N , pnk, k > 1.
We now apply the above theorems to the weighted mean in which A = (anv)
when 0 6 v 6 n, where Pn = p0 + p1 + ... + pn. Therefore,
is defined as anv = pv
Pn
it is well known that
¯anv =
Pn − Pv−1
Pn
and an,v+1 =
pnPv
PnPn−1
.
So, one can easily verify that the conditions of Theorem 3.1 reduce to those of
Theorem 2.3 and also we can obtain new results dealing with absolute matrix
summability of Fourier series in the following manner.
Theorem 4.3. Let A be a positive normal matrix satisfying the conditions of
Theorem 3.1. Let (Xn) be a quasi-σ-power increasing. If φ1(t) ∈ BV(0, π), and
the sequences (pn), (λn), and (Xn) satisfy the conditions of Theorem 3.1, then the
series P Cn(x)λn is summable A, pnk, k > 1.
Theorem 4.4. Let A be a positive normal matrix satisfying the conditions of
Theorem 3.1. Let (Xn) be a quasi-f -power increasing sequence. If φ1(t) ∈ BV(0, π),
and the sequences (pn), (λn), and (Xn) satisfy the conditions of Theorem 3.1, then
the series P Cn(x)λn is summable A, pnk, k > 1.
5. APPLICATIONS
n−v/Aα
We may now ask whether there are some examples other than weighted mean
methods of matrices A that satisfy the hypotheses of Theorem 3.1. For example,
apply Theorem 3.1 to the Ces`aro method of order α with 0 < α 6 1 in which A is
given by anv = Aα−1
n, and by applying Theorem 3.1, Theorem 4.3 and Theorem
4.4 to weighted mean so, the following results can be easily verified.
1. If we take anv = pv
Pn
have Theorem 2.3, Theorem 4.1 and Theorem 4.2.
2. If we take β = 0 and anv = pv
Pn
Theorem 2.2 and Theorem 4.1.
3. If we take pn = 1 for all values of n in Theorem 3.1, Theorem 4.3 and Theorem
4.4, then we have a new result dealing with Ak summability.
4. If we take anv = pv
Pn
and Theorem 4.4, then we have a new result concerning C, 1k summability.
and pn = 1 for all values of n in Theorem 3.1, Theorem 4.3
in Theorem 3.1, Theorem 4.3 and Theorem 4.4, then we
in Theorem 3.1 and Theorem 4.4, then we have
NEW VARIATIONS OF POWER INCREASING SEQUENCES
9
References
[1] Bari, N. K., Steckin, S.B, Best approximation and differential properties of
two conjugate functions. Trudy. Moskov. Mat. Obsc. (in Russian) 5, 483-522
(1956)
[2] Bor, H., On two summability methods. Math. Proc. Cambridge Philos Soc.
97, 147-149 (1985)
[3] Bor, H., Quasi-monotone and almost increasing sequences and their new appli-
cations. Abstr. Appl. Anal. Art. ID 793548, 6 PP.(2012)
[4] Bor, H., On absolute weighted mean summability of infinite series and Fourier
series. Filomat 30, 2803-2807 (2016)
[5] Bor, H., Some new results on absolute Riesz summablity of infinite series and
Fourier series. Positivity 20, 3 599-605 (2016)
[6] Bor, H., An Application of power increasing sequences to infinite series and
Fourier series. Filomat 31,6 1543-1547 (2017)
[7] Bor, H., Absolute weighted arithmetic mean summability factors of infinite
series and trigonometric Fourier series. Filomat 31, 15 4963-4968 (2017)
[8] Ces`aro, E., Sur la multiplication des s`eries. Bull. Sci. Math. 14, 114-120 (1890)
[9] Chen, K. K., Functions of bounded variation and the cesaro means of Fourier
series. Acad. Sin. Sci. Record 1 283-289 (1945)
[10] Flett, T. M., On an extension of absolute summability and some theorems of
Littlewood and Paley. Proc. Lond. Math. Soc. 7, 113-141 (1957)
[11] Hardy, G. H., Divergent Series. Clarendon Press, Oxford (1949)
[12] Kogbetliantz, E., Sur l`es series absolument sommables par la methode des
moyennes arithmetiques. Bull. Sci. Math. 49, 234-256 (1925)
[13] Leindler, L., A new application of quasi power increasing sequences. Publ.
Math. Debrecen 58, 791-796 (2001)
[14] Mazhar, S. M., Absolute summability factors of infinite series. Kyungpook
Math. J. 39, 67-73 (1999)
[15] Ozarslan, H. S., Yıldız, S¸., A new study on the absolute summability factors
of Fourier series. J. Math. Anal. 7 31-36 (2016)
10
YILDIZ
[16] Ozarslan, H. S., Yıldız, S¸., On the local property of summability of factored
Fourier series, Int. J. Pure Math. 3, 1-5, (2016).
[17] Sarıgol, M. A., On the local properties of factored Fourier series. Appl. Math.
Comp. 216 3386-3390 (2010)
[18] Sulaiman, W. T., Inclusion theorems for absolute matrix summability methods
of an infinite series. IV. Indian J. Pure Appl. Math. 34, 11 1547-1557 (2003)
[19] Sulaiman, W. T., Extension on absolute summability factors of infinite series.
J. Math. Anal. Appl. 322, 1224-1230 (2006)
[20] Tanovic-Miller, N., On strong summability. Glas. Mat. Ser III 14 , (34) 87-97
(1979)
[21] Yıldız, S¸., On absolute matrix summability factors of infinite series and Fourier
series. GU J. Sci. 30, 1 363-370 (2017)
[22] Yıldız, S¸., On Riesz summability factors Fourier series. Trans. A. Razmadze
Math. Inst. http://dx.doi.org/10.1016/j.trmi.2017.06.003. (2017)
[23] Yıldız, S¸., On Application of Matrix Summability to Fourier Series,Math.
Methods Appl. Sci., DOI: 10.1002/mma.4635, (2017)
Department of Mathematics, Ahi Evran University, Kırs¸ehir, Turkey
E-mail address: [email protected]; [email protected]
|
1103.1005 | 4 | 1103 | 2011-09-24T17:23:16 | On the reproducing kernel of a Pontryagin space of vector valued polynomials | [
"math.FA",
"math.AC"
] | We give necessary and sufficient conditions under which the reproducing kernel of a Pontryagin space of $d \times 1$ vector polynomials is determined by a generalized Nevanlinna pair of $d \times d$ matrix polynomials. | math.FA | math |
ON THE REPRODUCING KERNEL OF A PONTRYAGIN SPACE
OF VECTOR VALUED POLYNOMIALS
B. ´CURGUS AND A. DIJKSMA
Abstract. We give necessary and sufficient conditions under which the repro-
ducing kernel of a Pontryagin space of d × 1 vector polynomials is determined
by a generalized Nevanlinna pair of d × d matrix polynomials.
1. Introduction
1.1. By Baire's category theorem, the Pontryagin space B in the title is neces-
sarily finite dimensional (see Remark 2.1 below) and hence is a reproducing kernel
space. Indeed, if (cid:0)B, [· ,· ]B(cid:1) is an n-dimensional Pontryagin space of d × 1 vec-
tor polynomials and if B(z) is a d × n matrix polynomial whose columns Bk(z),
k ∈ {1, . . . , n}, form a basis of B, then the reproducing kernel of B is the d × d
matrix polynomial in z and w∗ given by
K(z, w) = B(z)G−1B(w)∗,
z, w ∈ C,
where G is the n × n Gram matrix associated with B(z), that is,
G = [gjk]n
j,k=1,
gjk = [Bk, Bj]B,
j, k ∈ {1, . . . , n}
(see [3, Example 2.1.8] and the remark following it). The reproducing kernel of a
reproducing kernel space is unique but can often be written in various ways. In
this paper we give necessary and sufficient conditions under which K(z, w) above
is a polynomial Nevanlinna kernel. This means that it can be written in the form
K(z, w) = KM,N (z, w) :=
M (z)N (w)∗ − N (z)M (w)∗
where M (z) and N (z) are d × d matrix polynomials such that
z − w∗
, z, w ∈ C, z 6= w∗,
for all
z ∈ C
and
M (z)N (z∗)∗ − N (z)M (z∗)∗ = 0
rank(cid:2)M (z) N (z)(cid:3) = d
(1.1)
If, in addition, the equality in (1.1) holds for all z ∈ C, then the Nevanlinna kernel
KM,N (z, w) is called a full Nevanlinna kernel.
The following theorem is the main result in this paper. It is proved in Section 4.
for at least one
z ∈ C.
Date: June 7, 2018.
2000 Mathematics Subject Classification. 46C20, 46E22, 47A06.
Key words and phrases. Polynomials, generalized Nevanlinna pair, Pontryagin space, repro-
ducing kernel, Smith normal form, Forney indices, symmetric operator, defect number, self-adjoint
extension, Q-function.
1
2
B. ´CURGUS AND A. DIJKSMA
Theorem 1.1. Let B be a (finite dimensional) Pontryagin space of d × 1 vec-
tor polynomials. Denote by SB the operator of multiplication by the independent
variable in B and by Eα the operator of evaluation at a point α ∈ C. Then the re-
producing kernel of B is a polynomial Nevanlinna kernel if and only if the following
two conditions hold:
(A) The operator SB is symmetric in B.
(B) For some α ∈ C we have ran(cid:0)SB − α(cid:1) = B ∩ ker Eα.
In this case the reproducing Nevanlinna kernel is full if and only if the equality in
(B) holds for all α ∈ C.
We think Theorem 1.1 is new, possibly even in the positive definite case, that
is, the case where the space B is a reproducing kernel Hilbert space of vector
polynomials. In that case B in the theorem is a special case of L. de Branges' Hilbert
spaces of entire functions. For scalar functions, see [12]; for vector functions, see [13]
and [14]. In particular, [14, Theorems 1-3] are closely related to Theorem 1.1. For
results on the indefinite scalar case we refer to the series of papers on Pontryagin
spaces of entire functions by M. Kaltenback and H. Woracek. More specifically, [26,
Theorem 5.3] is closely related to Theorem 1.1 with d = 1, [27, Proposition 2.8]
can be used to obtain a scalar version of Theorem 1.2 below, and [26, Lemma 6.4]
is linked with Theorem 5.1 in Section 5. The emphasis in this paper is on vector
polynomials and an indefinite setting.
In the proof of Theorem 1.1 we use the following result which shows that the
1.2.
condition (B) in Theorem 1.1 completely determines the structure of B as a linear
space. We believe Theorem 1.2 is also new, but closely related to results around
[20, Proposition 2.3]. For the proof of Theorem 1.2 we refer to Section 3.
Theorem 1.2. Let B be a finite dimensional linear space of d×1 vector polynomials
and let α ∈ C. The equality
(1.2)
holds if and only if there exist nonnegative integers µ1, . . . , µd and a d × d matrix
polynomial W (z) with det W (α) 6= 0 such that the space B consists of all vector
where pj(z) runs through all scalar
polynomials of degree strictly less than µj, j ∈ {1, . . . , d}. The matrix W (z) can be
chosen such that
polynomials of the form W (z)(cid:2)p1(z)··· pd(z)(cid:3)⊤
ran(cid:0)SB − α(cid:1) = B ∩ ker Eα
(cid:8)α ∈ C : det W (α) 6= 0(cid:9) =(cid:8)α ∈ C : ran(cid:0)SB − α(cid:1) = B ∩ ker Eα(cid:9).
(1.3)
It follows that the dimension of B in Theorem 1.2 is µ1 +··· + µd. If the conditions
(A) and (B) of Theorem 1.1 hold, then the numbers µ1, . . . , µd are the Forney indices
Nevanlinna kernel of B. Moreover, the defect numbers of SB coincide with the
of the block matrix polynomial (cid:2)M (z) N (z)(cid:3) corresponding to the reproducing
cardinality of the set (cid:8)j ∈ {1, . . . , d} : µj > 0(cid:9), see Remarks 4.1 and 4.3. This
from the block matrix polynomial(cid:2)M (z) N (z)(cid:3).
offers a direct way of determining the dimension of the reproducing kernel space
B with reproducing Nevanlinna kernel KM,N (z, w) and the defect numbers of SB
In the scalar case (d = 1) the space B in the above theorems is analogous to the
so-called Szego space, in the Hilbert space setting defined and studied in [34, 35]
and in the Pontryagin space setting in [1]. In the literature there are many papers
characterizing special forms of the reproducing kernel of a reproducing kernel space.
ON THE REPRODUCING KERNEL
3
Of those related to a reproducing kernel Pontryagin space we mention [7, Section 6]
and [2]. We refer to the references in these papers for papers dealing with the
Hilbert space case. The characterizations in these works are often in terms of a
special identity to be satisfied by the difference-quotient operator on the space. In
some cases, such as in [2, Theorem 4.1] and [12, Problems 51, Theorem 23] the
invertibility of K(z, z) for some values of z plays a role in proving the asserted
representation of the kernel K(z, w). We give in Section 6 some examples where
detK(z, w) = 0 for all z, w ∈ C, see Example 6.6 and Example 6.7.
1.3. A pair {M (z), N (z)} of d× d matrix functions M (z) and N (z) is called a gen-
eralized Nevanlinna pair if the functions are meromorphic on C\ R, the intersection
of the domains of holomorphy hol(M ) of M (z) and hol(N ) of N (z) is symmetric
with respect to the real axis,
(1.4)
(1.5)
M (z)N (z∗)∗ − N (z)M (z∗)∗ = 0
rank(cid:2)M (z) N (z)(cid:3) = d for at least one
for all
and the Nevanlinna kernel
z ∈ hol(M ) ∩ hol(N ),
z ∈ hol(M ) ∩ hol(N ),
M (z)N (w)∗−N (z)M (w)∗
(1.6) KM,N (z, w) :=
, z, w ∈ hol(M ) ∩ hol(N ), z 6= w∗,
has a finite number of negative squares. Here, by a finite number of negative
squares we mean that the set of numbers of negative eigenvalues counted according
to multiplicity of the self-adjoint matrices of the form
z − w∗
with
j KM,N (zj, zi)xi(cid:3)n
(cid:2) x∗
i,j=1
n ∈ N, xi ∈ Cd, zi ∈ hol(M ) ∩ hol(N ), zi 6= z∗
j , i, j ∈ {1, . . . , n}
has a maximum. If this maximum is κ, then we say that the pair and the kernel have
κ negative squares. If κ = 0 the adjective "generalized" is omitted; in that case the
matrix functions are holomorphic at least on C\ R. The number of positive squares
is defined in the same way. The pair and kernel are called full if the equality in (1.5)
holds for all z ∈ hol(M )∩ hol(N ). If a (generalized) Nevanlinna pair {M (z), N (z)}
is such that N (z) = Id, the d × d identity matrix, then it is identified with its first
entry M (z) and M (z) is a (generalized) Nevanlinna function.
Nevanlinna pairs and generalized Nevanlinna pairs have been used in interpola-
tion and moment problems (see [30], [4, 5] and [8]), the description of generalized
resolvents (see [28]) and in the theory of boundary value problems with eigenvalue
dependent boundary conditions (see [17, 18], [15] and [9]). Theorem 1.1 arose in our
study [10] of an eigenvalue problem for an ordinary differential operator in a Hilbert
space with boundary conditions which depend polynomially on the eigenvalue pa-
rameter. In that paper we linearize the original problem by extending the Hilbert
space with a finite dimensional Pontryagin space of d × 1 vector polynomials. This
paper concerns the structure of such spaces.
1.4. The Nevanlinna pair in a Nevanlinna kernel is not unique (see the paragraph
before Example 6.7) and if {M (z), N (z)} is a pair that determines the kernel, then
the polynomial matrix N (z) may be such that detN (z) = 0 for all z ∈ C.
In
Section 5 we prove that one can always choose the pair so that detN (z) 6≡ 0 and
4
B. ´CURGUS AND A. DIJKSMA
the rational generalized Nevanlinna matrix function N (z)−1M (z) is essentially a Q-
function of the symmetric operator SB. We show that every self-adjoint extension
of SB with nonempty resolvent set gives rise to a reproducing Nevanlinna kernel for
the space B. The proof of Theorem 1.1 given in Section 4 is geometric, the proof
of the first if statement in Theorem 1.1 given in Section 5 is analytic. The last two
examples in Section 6, Example 6.6 and Example 6.7, also serve to show that this
analytic proof is constructive. In Section 6 we present three corollaries of Theorem
1.1 and four examples.
In Section 2 we fix the notation related to vector and matrix polynomials and
we recall the Smith normal form and the Forney indices of a matrix polynomial.
Moreover, we prove some lemmas on the structure of a degenerate subspace of
a finite dimensional Pontryagin space, the defect numbers of a simple symmetric
relation in such a space and on polynomial Hermitian kernels. Although most
proofs in this paper are based on methods from linear algebra, in the sequel we
assume that the reader is familiar with (i) Pontryagin spaces and (multi-valued)
operators on such spaces such as symmetric and self-adjoint relations (as in [24],
[19] and [11]), (ii) generalized Nevanlinna matrix functions (as in [30, 31]) and (iii)
reproducing kernel Pontryagin spaces (as in [6, Chapter 1] and [3, Chapter 7]).
The notion of a Q-function of a simple symmetric operator in a Pontryagin space
is recalled in Section 5.
Acknowledgements. The authors thank Prof. Marius van der Put for many
discussions about Theorem 3.4 and another proof of it and Prof. Daniel Alpay for
acquainting them with his unpublished note [1] which among other things lead to
Remark 2.1 and pointing out the references [34, 35]. We also thank the referee for
useful comments.
2. Notation and basic objects
2.1. The symbols N, R, and C denote the sets of positive integers, real numbers
and complex numbers. For d ∈ N the vector space of all d × 1 vectors is written as
Cd and Id stands for the d × d identity matrix. The k-th row of Id will be denoted
by ed,k. For k ∈ {1, . . . , n} the subspace of Cd spanned by ed,1, . . . , ed,k will be
called a top coordinate subspace of Cd; it will be denoted by Cd
k. The corresponding
d× d projection matrix is denoted by Pd,k. We consider Cd
0 = {0} a top coordinate
subspace spanned by the empty set.
By Cd[z] we denote the vector space over C of all polynomials with coefficients
in Cd. The space Cd is identified with the subspace of all constant polynomials in
Cd[z]. If d = 1 we simply write C[z] and C. For f ∈ Cd[z] \ {0} with
Matrix polynomials are written as B(z), M (z), N (z), . . ., that is, with their argu-
ment z; we use the bold face P(z), S(z), . . ., for d × 2d matrix polynomials. Vector
polynomials are sometimes written with and sometimes without their argument.
The Fraktur alphabet A, B, C, H, . . . is used to denote vector subspaces of Cd[z].
One exception to this is that L will be used for a subspace of C2d[z]. An inner
and for the zero polynomial 0 we define
f (z) = a0 + a1z + ··· + anzn
deg f = max(cid:8)k ∈ {0, . . . , n} : ak 6= 0(cid:9) and
deg 0 = −∞.
ON THE REPRODUCING KERNEL
5
product on B is denoted by [· ,· ]B. In a vector space, the symbol ⊕ denotes the
direct sum of subspaces.
Remark 2.1. A Banach space with a countable Hamel basis is separable and
hence, by [32], it is finite dimensional. Since (cid:8)zn : n ∈ {0} ∪ N(cid:9) is a countable
Hamel basis of C[z], the space Cd[z] and all its subspaces also have countable
Hamel bases. Therefore any Pontryagin subspace of Cd[z] is finite dimensional. In
spite of this fact, to emphasize the finite dimensionality, we continue to speak of
finite dimensional Pontryagin subspaces of Cd[z].
We introduce some special subspaces of Cd[z]. Let n ∈ {0} ∪ N. The symbol
Cd[z]<n stands for the set of all f ∈ Cd[z] such that deg f < n.
In particular,
Cd[z]<1 = Cd and Cd[z]<0 = {0}. A subspace C of Cd[z] is called canonical if there
exist nonnegative integers µk, k ∈ {1, . . . , d}, such that
C =
dMk=1(cid:0)C[z]<µk(cid:1)ed,k
=(cid:8)[p1(z)··· pd(z)]⊤ : pk(z) ∈ C[z], degpk < µk, k ∈ {1, . . . , d}(cid:9).
The numbers µ1, . . . , µd will be called the degrees of C. Without loss of generality
we can assume that they are ordered: µ1 ≥ ··· ≥ µd ≥ 0. Then a canonical
subspace is uniquely determined by its degrees. Clearly, the dimension of C is the
sum of its degrees.
Next we introduce some useful operators on Cd[z]. By Pd,k, k ∈ {1, . . . , d}, we
denote the natural extension of Pd,k to Cd[z], by S : Cd[z] → Cd[z] the operator
of multiplication by the independent variable, that is,
(Sf )(z) = zf (z),
f ∈ Cd[z],
and by Eα : Cd[z] → Cd the evaluation operator at the point α ∈ C:
Eα(f ) = f (α),
f ∈ Cd[z].
It follows from the fundamental theorem of algebra that
ran(cid:0)S − α(cid:1) = ker Eα.
(2.1)
A wide class of operators on Cd[z] is induced by d× d matrix polynomials. If M (z)
is such a polynomial we define the operator M : Cd[z] → Cd[z] by
(cid:0)M f(cid:1)(z) = M (z)f (z),
f ∈ Cd[z].
Clearly, M S = SM . A square matrix polynomial is unimodular if its determinant is
identically equal to a nonzero constant. If M (z) is a unimodular matrix polynomial
we will call M a unimodular operator. In this case M is a bijection and its inverse
is also a unimodular operator.
2.2.
a Smith normal form representation (see for example [22, Satz 6.3] or [25]):
In the sequel we use that any nonzero d × n matrix polynomial B(z) admits
(2.2)
B(z) = U (z)"D(z) 0
0# V (z),
0
where U (z) is a d × d unimodular matrix polynomial, V (z) is an n × n unimod-
ular matrix polynomial and the matrix in the middle is a d × n matrix in which,
for some l ∈ {1, . . . , min{d, n}}, D(z) is a diagonal l × l matrix polynomial with
6
B. ´CURGUS AND A. DIJKSMA
monic diagonal entries: D(z) = diag(cid:0)b1(z), . . . , bl(z)(cid:1) such that bi(z) is divisible by
bi+1(z), i ∈ {1, . . . , l − 1}. Notice that rank B(α) = l if and only if b1(α) 6= 0. If
for some z ∈ C the rank of B(z) is d (n, respectively), then l = d (l = n) and the
zero block row (column) in the matrix in the middle of the right hand side in (2.2)
is not present.
Remark 2.2. The matrix in the middle of the right hand side in (2.2) is uniquely
determined by B(z). In this paper B(z) often is a matrix polynomial whose columns
form a basis of a subspace B of Cd[z]. Then for any d × n matrix polynomial
B1(z) whose columns also form a basis of B, the middle term of its Smith normal
form is identical to that of B(z). Thus, the number l and the monic polynomials
bj(z), j ∈ {1, . . . , l}, above are uniquely determined by the subspace B of Cd[z].
2.3. Let S(z) be a d × 2d polynomial matrix. For j ∈ {1, . . . , d} let σj be the
degree of the j-th row of S(z). By definition, a degree of a row is the degree of its
transpose. Define S∞, the internal degree and the external degree of S(z) by:
S∞ = lim
z→∞
z
−σ1
.
.
.
0
· · ·
.
.
.
· · ·
0
.
.
.
−σd
z
S(z),
extdeg S(z) = σ1 + ··· + σd,
intdeg S(z) = max{ deg m(z) : m(z) is a d × d minor of S(z)}.
and
For a proof of the following theorem we refer to [33].
Theorem 2.3. Let P(z) be a d × 2d matrix polynomial with rankP(z) = d for all
z ∈ C. Let S(z) be a matrix polynomial in the family
(2.3)
(cid:8)U (z)P(z) : U (z) unimodular(cid:9).
The following statements are equivalent:
rankS∞ = d.
(b)
(c) extdeg S(z) = intdeg S(z).
(d) S(z∗)∗ has the "predictable degree property":
(a) extdeg S(z) = min(cid:8)extdeg U (z)P(z) : U (z) unimodular(cid:9).
For every u(z) =(cid:2)u1(z) ··· ud(z)(cid:3)⊤
∈ Cd[z] we have
deg(cid:0)S(z∗)∗u(z)(cid:1) = max(cid:8)σj + deg uj(z), j ∈ {1, . . . , d}(cid:9).
A matrix polynomial S(z) in the family (2.3) satisfying the conditions (a) -- (d) is
called row reduced. The multiset {σ1, . . . , σd} of row degrees for each row reduced
matrix in the family (2.3) is the same. Its elements are called the Forney indices
of any of the matrices in the family (2.3), in particular of P(z). We extend this
definition to the case where the d × 2d matrix polynomial P(z) has full rank for
some z ∈ C. For that we use the following lemma which is a standard tool in system
theory, see for example [21].
Lemma 2.4. Let P(z) be a d × 2d matrix polynomial with rankP(z) = d for some
z ∈ C. Then P(z) admits the factorization:
(2.4)
where G(z) is a d × d matrix polynomial with det G(z) 6≡ 0 and T(z) is a d × 2d
matrix polynomial with rankT(z) = d for all z ∈ C. This factorization is essentially
P(z) = G(z)T(z)
z ∈ C,
for all
ON THE REPRODUCING KERNEL
7
unique, meaning that if also P(z) = G1(z)T1(z) for all z ∈ C, where G1(z) and
T1(z) have the same properties as G(z) and T(z), then for some unimodular d × d
matrix polynomial E(z): G1(z) = G(z)E(z)−1 and T1(z) = E(z)T(z), z ∈ C.
The Forney indices of P(z) in the lemma are by definition the Forney indices
of the matrix polynomial T(z) in the factorization (2.4). By the second part of
the lemma, this definition is independent of the choice of the matrix G(z) in this
factorization.
For convenience of the reader we give a proof of Lemma 2.4 based on the Smith
normal form of a matrix polynomial.
Proof of Lemma 2.4. Let P(z) have the Smith normal form (2.2). The assumptions
imply that l = d and that the matrix in the middle of (2.2) is equal to(cid:2)D(z) 0(cid:3).
0(cid:3) V (z). Then the factorization (2.4)
Set G(z) = U (z)D(z) and T(z) = (cid:2)Id
holds and G(z) and T(z) have the properties mentioned in the lemma. To prove
uniqueness we use the fact that, since T(z) and T1(z) have full rank for all z ∈ C,
they have right inverses, see [25]. These are 2d × d matrix polynomials S(z) and
S1(z) such that T(z)S(z) = Id and T1(z)S1(z) = Id for all z ∈ C. Define the
matrix polynomials E(z) = T1(z)S(z) and F (z) = T(z)S1(z). Then the equality
G(z)T(z) = G1(z)T1(z) implies E(z) = G1(z)−1G(z) and F (z) = G(z)−1G1(z),
hence E(z)F (z) = Id for all but finitely many z ∈ C. By continuity the last equality
holds for all z ∈ C, hence E(z) is unimodular and has the stated properties.
(cid:3)
2.4. The next two lemmas concern finite dimensional Pontryagin spaces. By the
positive (negative) index of a Pontryagin space K we mean the dimension of a
maximal positive (negative) subspace of K; evidently, the dimension of K is equal
to the sum of the indices.
Lemma 2.5. Let K be a Pontryagin space with positive and negative index equal to
n. Let L be a subspace of K with dimL = 2n − τ . If L contains a maximal neutral
subspace of K, then L⊥ is the isotropic part of L and L/L⊥ is a Pontryagin space
with positive and negative index equal to n − τ .
Proof. Let N be a maximal neutral subspace contained in L. Since N ⊥ = N , the
inclusion N ⊆ L, yields L⊥ ⊂ N ⊂ L. Therefore, L⊥ is the isotropic part of L and
dimL⊥ = τ . Let L = L⊥ +L− +L+ be a pseudo-fundamental decomposition of L.
Since N is a neutral subspace of L, we have n = dimN ≤ τ + dim L±. Therefore
2n − τ = dimL = τ + dimL− + dimL+ ≥ τ + n − τ + n − τ = 2n − τ.
(cid:3)
This proves that dimL− = dimL+ = n − τ .
Recall that a symmetric relation S in a Pontryagin space K is simple if S has no
non-real eigenvalues and K = span{ker(S∗ − z) : z ∈ C \ R}. Below mul S∗ stands
for the multi-valued part of the adjoint S∗ of S: mul S∗ = {g ∈ K : {0, g} ∈ S∗}.
Lemma 2.6. Let S be a simple symmetric relation in a finite dimensional Pontrya-
gin space of dimension n. Then the spaces mul S∗, ker S∗, and S∗ ∩ zI, z ∈ C, have
the same dimension d′, say. In particular, the defect numbers of S are both equal
to d′. Furthermore, dim ranS = dim S = dim domS = n − d′ and dim S∗ = n + d′.
Proof. First notice that by [9, Proposition 2.4] S is an operator and S has no
eigenvalues. The following statements are equivalent:
8
B. ´CURGUS AND A. DIJKSMA
(a) dim(cid:0)mul S∗(cid:1) = d′.
(b) codim(cid:0)dom S(cid:1) = d′.
(c) codim(cid:0)ran(S − z∗)(cid:1) = d′ for all z ∈ C.
(d) dim(cid:0)S∗ ∩ zI(cid:1) = d′ for all z ∈ C.
that d′ = dim(cid:0)ker S∗(cid:1). The equalities n − d′ = dim domS = dim S = dim ranS
The relation (dom S)⊥ = mul S∗ implies the equivalence (a)⇔(b). The equivalence
(b)⇔(c) follows from the fact that S − z∗ is one-to-one. By taking the orthogonal
complements we obtain the equivalence (c)⇔(d). Notice that (d) with z = 0 implies
follow from (b) and the fact that S is an injective operator. Since dim S∗ = 2n −
dim S the last equality follows.
(cid:3)
2.5. A d× d matrix function K(z, w) will be called a polynomial Hermitian kernel
if it is a polynomial of two variables z and w∗ and K(z, w)∗ = K(w, z), z, w ∈ C.
This implies that the degree of K(z, w) as a polynomial in z equals the degree of
K(z, w) as a polynomial in w∗. If we denote this common degree by p − 1, then
K(z, w) can be expanded as
(2.5)
K(z, w) =
p−1Xk=0
Ajkzjw∗k,
z, w ∈ C,
where Ajk, j, k ∈ {0, . . . , p − 1}, are d × d matrices. Since K(z, w) is a Hermitian
kernel, the dp × dp block matrix
p−1Xj=0
A =
A00
...
Ap−1,0
A0,p−1
···
. . .
··· Ap−1,p−1
...
(2.6)
is self-adjoint. It also follows that the number of negative squares of K(z, w) equals
the number of negative eigenvalues of A and the number of positive squares of
K(z, w) equals the number of positive eigenvalues of A. The dimension of the
reproducing kernel space corresponding to K(z, w) is the rank of A. These obser-
vations are used in the proof of the following lemma.
Lemma 2.7. Let K(z, w) be a d× d matrix polynomial Hermitian kernel of degree
p − 1. For q ∈ N set
Lq(z, w) = i (zq − w∗q)K(z, w),
z, w ∈ C.
If q ≥ p, then the positive and the negative index of the reproducing kernel Pon-
tryagin space with kernel Lq(z, w) are equal and coincide with the dimension of the
reproducing kernel Pontryagin space with kernel K(z, w).
Proof. Write K(z, w) in the form (2.5) and denote by A the matrix (2.6). We
calculate the coefficients of the matrix polynomial Lq(z, w) for q ≥ p:
Lq(z, w) = izq
Ajkzjw∗k − iw∗q
p−1Xk=0
p−1Xj=0
p−1Xk=0
p−1Xj=0
=
iAjkzq+jw∗k +
Ajkzjw∗k
p−1Xk=0
(−i)Ajkzjw∗(q+k)
p−1Xj=0
p−1Xk=0
p−1Xj=0
ON THE REPRODUCING KERNEL
9
=
=
q+p−1Xj=0
q+p−1Xj=0
p−1Xj=0
iA(j−q)kzjw∗k +
q+p−1Xk=0
p−1Xk=0
q+p−1Xk=0 (cid:0)iA(j−q)k − iAj(k−q)(cid:1)zjw∗k,
(−i)Aj(k−q)zjw∗k
where we set Ajk = 0 whenever j < 0 or k < 0 or j > p − 1 or k > p − 1. In other
words, the 2d(p + q) × 2d(p + q) self-adjoint matrix formed by the coefficients of
Lq(z, w) is given by
where the 0 in the center is a d(q − p) × d(q − p) matrix. With
0 −iA
0
0
0
0
iA 0
0
Idp
0
iIdp
0
Id(q−p)
0
iIdp
0
Idp
B =
√2
E∗BE = E∗
E =
1
0 −iA
0
0
0
0
iA 0
0 E =
A 0
0
0
0
0 −A .
0
0
we have EE∗ = Id(q+p) and
Therefore the rank of B is twice the rank of A. Moreover, B has equal numbers
of positive and negative eigenvalues. Since the positive and negative index of the
reproducing kernel Pontryagin space with kernel Lq(z, w) coincide with the number
of positive and negative eigenvalues of B the lemma is proved.
(cid:3)
A polynomial reproducing Nevanlinna kernel introduced in the Introduction is a
polynomial Hermitian kernel. Since in the proof of Theorem 1.1 the polynomials in a
Nevanlinna pair never appear separate we adopt the following equivalent definition
of a polynomial Nevanlinna kernel: A d × d matrix function K(z, w) is called a
polynomial Nevanlinna kernel if it can be represented as
(2.7)
P(z)Q−1P(w)∗ = i (z − w∗)K(z, w)
for all
z, w ∈ C,
where Q is a 2d × 2d self-adjoint matrix with d positive and d negative eigenvalues
and P(z) is a d × 2d matrix polynomial such that P(z) has rank d for some z ∈ C.
With
(2.8)
Q = Q1 :=" 0 −iId
0 #
iId
and P(z) =(cid:2)M (z) N (z)(cid:3)
the definition in the Introduction is obtained from the new one. The assumptions
on Q imply that there exists a constant invertible matrix T such that Q = T Q1T ∗.
Now, if we write P(z)T = (cid:2)M (z) N (z)(cid:3), we have K(z, w) = KM,N (z, w). Since
P(z) is a polynomial, the condition that rank P(z) = d for some z ∈ C implies that
rank P(z) = d for all but finitely many z ∈ C. A polynomial Nevanlinna kernel will
be called a full Nevanlinna kernel if P(z) can be chosen such that rank P(z) = d
for all z ∈ C.
10
B. ´CURGUS AND A. DIJKSMA
3. Proof of Theorem 1.2
3.1. Let B be a vector subspace of Cd[z]. By SB we denote the range restriction
of S to B, that is,
In graph notation this means:
(cid:0)SBf(cid:1)(z) = zf (z), f ∈ dom SB.
dom SB = B ∩ S−1B,
SB =(cid:8){f, g} : f, g ∈ B, g(z) = zf (z) for all z ∈ C(cid:9).
By (2.1), for α ∈ C we have
(3.1)
ran(cid:0)SB − α(cid:1) =(cid:0)S − α(cid:1)(cid:0)B ∩ S−1B(cid:1) ⊆ ran(cid:0)S − α(cid:1) ∩ B = B ∩ ker Eα.
The reverse inclusion is equivalent to the implication
f ∈ B, α ∈ C, f (α) = 0 ⇒ f (z) = (z − α)g(z) for some g ∈ dom SB.
In some cases this implication does not hold. For example, it does not hold for any
α ∈ C in the space B ⊂ C2[z] given by
B =(cid:26)(cid:20) a0 + a2z2
b0 + b1z (cid:21) : a0, a2, b0, b1 ∈ C(cid:27) .
z − α(cid:3)⊤
which is 0 at z = α, but B does not
Indeed, B contains (cid:2)z2 − α2
contain(cid:2)z + α 1(cid:3)⊤
. That the implication, or equivalently, equality in (3.1) holds,
is characterized in terms of canonical subspaces of Cd[z] in Theorem 1.2 in the
Introduction. This section is devoted to the proof of this theorem.
Let B(z) be a d × n matrix polynomial whose columns form a basis for B,
n = dim B. Then, as will be shown in the proof of Theorem 1.2, the sets in (1.3)
are equal to (cid:8)α ∈ C : b1(α) 6= 0(cid:9), where b1(z) is the scalar polynomial in the
Smith normal form (2.2) of B(z). We will first prove Theorem 1.2 for the case
where the sets in (1.3) are equal to C, see Theorem 3.4 below. In this case W (z) is
unimodular. The proof of Theorem 3.4 is based on the following three lemmas.
Lemma 3.1. Let B be a finite dimensional subspace of Cd[z] such that
ran(cid:0)SB − α(cid:1) = B ∩ ker Eα
for all α ∈ C.
If dom SB ⊆ B′ ⊆ B, then ran(cid:0)SB′ − α(cid:1) = B′ ∩ ker Eα for all α ∈ C.
Proof. Let f ∈ B′ ∩ ker Eα. Then f ∈ B ∩ ker Eα = ran(cid:0)SB − α(cid:1), that is,
g ∈ dom SB′ and f = (SB′ − α)g. This proves B′ ∩ ker Eα ⊆ ran(cid:0)SB′ − α(cid:1). Since
f = Sg− αg for some g ∈ domSB ⊆ B′. From f, g ∈ B′ we infer g, Sg ∈ B′. Hence
the reverse inclusion is obvious, the lemma is proved.
(cid:3)
Lemma 3.2. Let B be an n-dimensional subspace of Cd[z]. Then
(3.2)
if and only if there exists a unimodular operator W such that B = W Cd
is a top coordinate subspace of Cd.
B ∩ ker Eα = {0}
for all α ∈ C
n, where Cd
n
Proof. If n = 0, the statements are trivial with W (z) = Id. From now on we assume
n ≥ 1. If B(z) is any d × n matrix polynomial whose columns form a basis of B,
then, clearly,
(3.3)
(cid:8)α ∈ C : rank B(α) = n(cid:9) =(cid:8)α ∈ C : B ∩ ker Eα = {0}(cid:9).
ON THE REPRODUCING KERNEL
11
Assume (3.2). Let B(z) be a d × n matrix polynomial whose columns form a
basis of B. By (3.3), for all α ∈ C the rank of B(α) is n and n ≤ d. Hence B(z)
V (z), where U (z)
and V (z) are unimodular. Define
admits the Smith normal form (see (2.2)): B(z) = U (z)(cid:2)In
Id−n# .
W (z) = U (z)"V (z)
(3.4)
0
0
0(cid:3)⊤
it follows that B = W Cd
n. This proves the only if statement.
polynomial W (z) such that B = W Cd
Then W (z) is a unimodular d×d matrix polynomial and from B(z) = W (z)(cid:2)In 0(cid:3)⊤
Cd. Then the columns of B(z) = W (z)(cid:2)In 0(cid:3)⊤
To prove the if statement, assume that there exists a d × d unimodular matrix
n is a top coordinate subspace of
form a basis of B and the rank of
(cid:3)
B(α) is n for all α ∈ C. The equality (3.2) follows from (3.3).
Lemma 3.3. Let B be an n-dimensional subspace of Cd[z] and let C be a canonical
subspace of Cd[z] with degrees µ1 ≥ ··· ≥ µd ≥ 0 of which k are positive. Assume
C + SC ⊆ B and
(3.5)
n, where Cd
B ∩ ker Eα ⊆ C + SC
for all α ∈ C.
Then there exists a unimodular operator W which acts as the identity on C + SC
and is such that
(3.6)
where m = n − (µ1 + ··· + µd)(≥ 0).
B = W(cid:0)Cd
m + C + SC(cid:1),
Notice that Cd
m + C + SC is a canonical subspace. If m ≤ k, then Cd
m + C + SC
coincides with C + SC and W = Id.
Proof. If C = {0} the statement follows from Lemma 3.2. From now on we assume
C 6= {0}. Then µ1 > 0, consequently k ∈ {1, . . . , d} and Cd
k ⊆ C. We consider two
cases: k = d and k < d.
(i) Assume k = d. Then Cd ⊆ C ⊆ B. Let f ∈ B.
It can be written as
f (z) = f (0)+zh(z) = f (0)+(Sh)(z). Then Sh = f −f (0) ∈ B. Since(cid:0)Sh(cid:1)(0) = 0,
by (3.5) we get Sh ∈ B∩ ker E0 ⊆ C + SC, which implies f = f (0) + Sh ∈ C + SC.
That is, B = C + SC. In this case m = d and with W = Id the lemma is proved.
(ii) Assume k < d. If C + SC = B, then (3.6) holds with W = Id and m = k,
implying that Cd
m ⊆ C. From now on we assume that C + SC is a proper subspace
of B. Recall that Pd,k is the coordinate projection. A trivial, but important
observation is
(3.7)
Let α ∈ C be arbitrary and let f ∈ B be such that (Id − Pd,k)f (α) = 0. By (3.7),
there exists a p ∈ C + SC such that p(α) = Pd,kf (α), hence
for all α ∈ C.
k = ran Pd,k
Eα(cid:0)C + SC(cid:1) = Cd
(f − p)(α) = (Id − Pd,k)f (α) + Pd,kf (α) − p(α) = 0,
that is, f − p ∈ ker Eα. Since also f − p ∈ B, (3.5) implies f − p ∈ C + SC. Thus
both p and f − p belong to C + SC, implying that f ∈ C + SC. We have proved the
implication:
(3.8)
f ∈ B, α ∈ C and (Id − Pd,k)f (α) = 0 ⇒ f ∈ C + SC.
12
B. ´CURGUS AND A. DIJKSMA
Let L0 be a subspace of B be such that
(cid:0)C + SC(cid:1) ∩ L0 = {0}
The dimension of L0 is
and
B =(cid:0)C + SC(cid:1) ⊕ L0.
Let B0(z) be a d× j matrix polynomial whose columns form a basis of L0. Decom-
pose B0(z) as
j = n −(cid:0)µ1 + ··· + µd + k(cid:1) ≥ 1.
B0,b(z)# ,
B0(z) ="B0,t(z)
where B0,t(z) is a k × j matrix polynomial and B0,b(z) a (d − k) × j matrix poly-
nomial. We will prove that
for all α ∈ C.
rank B0,b(α) = rank(Id − Pd,k)B0(α) = j
(Id − Pd,k)f (α) = 0. By (3.8), f ∈ (cid:0)C + SC(cid:1) ∩ L0, consequently f = 0, that is,
(3.9)
The first equality is trivial. To prove the second let α ∈ C be arbitrary and x ∈ Cj
be such that (Id − Pd,k)B0(α)x = 0. Set f (z) = B0(z)x. Then f ∈ L0 and
B0(z)x = 0 for all z ∈ C. Since the columns of B0(z) form a basis of L0, this implies
x = 0. This proves (3.9). Hence j ≤ d− k. If j = d− k, the (d− k)× (d− k) matrix
polynomial Wb(z) := B0,b(z) is unimodular. If j < d − k we can extend B0,b(z) to
a unimodular (d − k) × (d − k) matrix polynomial (also denoted by) Wb(z) with
det Wb(α) 6= 0 in the same way as the matrix B(z) was extended to W (z) in (3.4)
by the means of the Smith normal form. In both cases the first j columns of Wb(z)
are the columns of B0,b(z).
Let Wt(z) be the k × (d − k) matrix obtained from the k × j matrix B0,t(z) by
adding d − k − j zero columns on the right. Define the d × d matrix polynomial
W (z) by
W (z) ="
Ik
0(d−k)×k Wb(z)# .
Wt(z)
Then W (z) is unimodular and W (z)ed,k+l, l = 1, . . . , j, are the columns of the
matrix B0(z). The operator W acts as the identity on C+ SC and W Cd
k + L0,
where
m = Cd
m = k + j = k + n −(cid:0)µ1 + ··· + µd + k(cid:1) = n −(cid:0)µ1 + ··· + µd(cid:1).
m + C + SC(cid:1) = Cd
k + L0 + C + SC = B.
Theorem 3.4. Let B be a finite dimensional subspace of Cd[z]. The equality
Hence W(cid:0)Cd
(3.10)
holds if and only if there exist a d × d unimodular matrix polynomial W (z) and a
canonical subspace C of Cd[z] such that B = W C.
ran(cid:0)SB − α(cid:1) = B ∩ ker Eα
for all α ∈ C
(cid:3)
Proof. We first prove the if statement. To prove (3.10) it suffices to show that
B ∩ ker Eα ⊆ ran(SB − α).
Let f ∈ B ∩ ker Eα. Then f (α) = 0 and f = W g for some g ∈ C. Since W is
unimodular, g(α) = 0. Since C is canonical, the polynomial g(z)/(z − α) belongs
We prove the only if statement by induction on the dimension of B. Assume
(3.10). The theorem is obviously true if dim B = 0. Lemma 3.2 implies that it
to C. Therefore f (z)/(z − α) = W (z)(cid:0)g(z)/(z − α)(cid:1) ∈ B, hence f ∈ ran(SB − α).
ON THE REPRODUCING KERNEL
13
is true if dim B = 1 for then B ∩ Eα = {0}. Let n ∈ N and state the inductive
hypothesis:
If A is a subspace of Cd[z] with dim A < n and such that
for all α ∈ C,
ran(cid:0)SA − α(cid:1) = A ∩ ker Eα
(3.11)
then there exists a unimodular d × d matrix polynomial operator F (z) such that
F A is a canonical subspace of Cd[z].
Let B be a finite dimensional subspace of Cd[z] such that (3.10) holds and
dim B = n. Then A = dom SB is a proper subspace of B. Therefore dim A < n.
If A = {0}, then B ∩ ker Eα = ran(cid:0)SB − α(cid:1) = {0} and the theorem follows from
Lemma 3.2. Now we assume A 6= {0}. By Lemma 3.1 the subspace A satisfies (3.11).
By the inductive hypothesis there exists a unimodular matrix polynomial F (z) such
that D := F A is a canonical subspace of Cd[z]. Since F and S commute we have
D = F A = F dom SB = dom SF B, hence D + SD ⊆ U B. To apply Lemma 3.3 to
F B we need to verify (3.5). Let f ∈ B be such that (F f )(α) = 0. Then f (α) = 0
and, by (3.10), there exists a g ∈ dom SB = A such that f = SBg − αg ∈ A + SA.
Therefore, F f ∈ D+SD, which verifies (3.5). Lemma 3.3 applied to F B yields that
there exists a unimodular operator U such that U −1F B is a canonical subspace of
Cd[z]. This proves the theorem with W = F −1U
(cid:3)
3.2. The following lemma will be used to deduce Theorem 1.2 from Theorem 3.4.
Lemma 3.5. Let B be an n-dimensional subspace of Cd[z] and let B(z) be a d× n
matrix polynomial whose columns form a basis of B. Let l be the size of the square
diagonal matrix in the Smith normal form (2.2) of B(z). Then
(cid:8)α ∈ C : ran(SB − α) = B ∩ ker Eα(cid:9) =(cid:8)α ∈ C : rank B(α) = l(cid:9)
if and only if the set on the left hand side is nonempty. In this case dim ran SB =
dim B − l.
Proof. The only if statement follows from the fact that the set on the right hand
side in (3.12) is nonempty. Before proving the if statement we show
(3.12)
(3.13)
For all α ∈ C we have ran(SB − α) ⊆ B ∩ ker Eα, and hence
dim ran SB ≤ dim B − l.
dim ran SB = dim ran(SB − α) ≤ dim(B ∩ ker Eα) = dim B − rank B(α).
Consequently, l = maxα∈C rank B(α) ≤ dim B − dim ran SB. This proves (3.13).
of (3.12). Then equality holds in (3.13). Indeed, this follows from
To prove the if statement assume that α0 ∈ C is in the set on the left hand side
dim B − l ≥ dim ran SB
= dim ran(SB − α0)
= dim(cid:0)B ∩ ker Eα0(cid:1)
= dim B − rank B(α0)
≥ dim B − l.
This proves the last statement in the lemma. Now the equality (3.12) follows from
the following sequence of equivalences which hold for all α ∈ C:
rank B(α) = l ⇔ dim(B ∩ ker Eα) = dim B − l
⇔ dim(B ∩ ker Eα) = dim ran SB
14
B. ´CURGUS AND A. DIJKSMA
⇔ dim(B ∩ ker Eα) = dim ran(SB − α)
⇔ ran(SB − α) = B ∩ ker Eα.
(cid:3)
Proof of Theorem 1.2. We first prove the if statement.
It suffices to prove the
inclusion B ∩ ker Eα ⊆ ran(SB − α), as the reverse inclusion always holds. Let
f ∈ B ∩ ker Eα. Then f (α) = 0 and f = W g with g ∈ C. Since W (α) is
invertible, g(α) = 0. As C is canonical, the polynomial h(z) = g(z)/(z − α) belongs
to C. Therefore W (z)h(z) ∈ B and (x − α)W (z)h(z) = f (z), which implies f ∈
ran(SB − α).
To prove the only if statement, assume that (1.2) holds for α = α0. Let B(z) be
a d × n matrix polynomial whose columns form a basis of B. Let
B(z) = U (z)"D(z) 0
0# V (z)
0
be the Smith normal form (2.2) of B(z) where D(z) is an l× l diagonal matrix with
nonzero diagonal entries. Now define the space B1 ⊂ Cd[z] as the span over C of
the columns of
Set
B1(z) = U (z)"Il
F (z) = U (z)"D(z)
0
0
0
0# V (z).
Id−l# U (z)−1.
0
Then B = F B1 and detF (α0) 6= 0. Moreover, since
{α ∈ C : detF (α) 6= 0} = {α ∈ C : rankB(α) = l}
and by Lemma 3.5, (1.3) holds for F (z). From detF (α0) 6= 0 it follows that
ran(SB − α0) = B ∩ ker Eα0 ⇒ ran(SB1 − α0) = B1 ∩ ker Eα0 .
Since rank B1(α) = l for all α ∈ C, Lemma 3.5 implies that
ran(SB1 − α) = B1 ∩ ker Eα
for all α ∈ C.
By Theorem 3.4, there exists a unimodular matrix U (z) such that C = U −1B1 is
a canonical subspace of Cd[z], hence B = W C with W = F U . Finally, (1.3) holds,
because U is unimodular and F (z) satisfies (1.3).
(cid:3)
3.3. Theorem 1.2 can also be formulated in terms of matrix polynomials:
Theorem 3.6. Let B be an n-dimensional subspace of Cd[z], n ≥ 1. Let B(z) be
a d × n matrix polynomial whose columns form a basis of B. Let b1(z) and l be
as in the Smith normal form (2.2) of B(z). Then l + dim dom SB = dim B if and
only if there exist
(a) a d×d matrix polynomial W (z) whose determinant has the same zeros as b1(z),
(b) nonnegative integers m and δ0 ≥ δ1 ≥ ··· ≥ δm with δ0 + ··· + δm = n and
(c) an invertible n × n constant matrix T
such that
(3.14)
B(z) = W (z)hPδ0 Pδ1 z ··· Pδm zmi T for all
z ∈ C,
where Pδ stands for the d × δ matrix: Pδ =(cid:2)Iδ
.
0(cid:3)⊤
ON THE REPRODUCING KERNEL
15
Proof. For all α ∈ C we have ker(SB − α) ⊆ B ∩ kerEα. For all α ∈ C with
b1(α) 6= 0 we have
l + dim dom SB = dim ran(cid:0)EαB(cid:1) + dim ran(SB − α)
≤ dim ran(cid:0)EαB(cid:1) + dim(B ∩ kerEα)
= dim B
and equality holds if and only if ker(SB − α) = B ∩ kerEα.
To prove the only if statement, assume l + dim dom SB = dim B. Then we can
apply Theorem 1.2: There exist a matrix polynomial W (z) satisfying (a) and a
canonical subspace C of Cd[z] such that B = W C. Let µ1 ≥ ··· ≥ µd be the
degrees of C. Since n ≥ 1, we have µ1 ≥ 1. set m = µ1 − 1 and
δj = #{i ∈ {1, . . . , d} : µi > j},
j ∈ {0, . . . , m}.
Then the equality in (b) holds. Since the columns of the matrix(cid:2)Pδ0
form a basis for C, there exists a matrix T satisfying (c) such that (3.14) holds.
··· Pδmzm(cid:3)
To prove the if statement, we note that (a)-(c) and (3.14) imply that B = W C
with C as above, and hence Theorem 1.2 can be applied and together with the if
and only if statement at the beginning of the proof yield that l + dim dom SB =
dim B.
(cid:3)
Remark 3.7. Denote by (Sj)B the range restriction of Sj to B. Then in item (b)
of Theorem 3.6: m is the nonnegative integer with
and
{0} = dom(Sm+1)B ( dom(Sm)B
Moreover, if we set δ−1 = d, then the numbers
δj = dim dom(Sj)B − dim dom(Sj+1)B,
µk = 1 + max(cid:8)j ∈ {−1, 0, . . . , m} : δj ≥ k(cid:9),
(3.15)
j ∈ {0, . . . , m}.
k ∈ {1, . . . , d},
are the degrees of the canonical space W −1B. In the next section we will see that
if B ⊂ Cd[z] is a Pontryagin space which satisfies the conditions (A) and (B) of
Theorem 1.1, then the numbers (3.15) are the Forney indices of a matrix polynomial
P(z) in a representation (2.7) of the Nevanlinna reproducing kernel K(z, w) of B;
see Remark 4.3.
4. Proof of Theorem 1.1
4.1. We divide the proof of Theorem 1.1 in two parts. In the first part we prove
the if statements and in the second part we prove the only if statements. In the
first part we will need characterizations of the defect numbers of the operator SB
of multiplication by the independent variable in the Pontryagin space B which are
collected in the following remark.
Remark 4.1. Clearly, SB has no eigenvalues and for any subset Ω of C containing
more than d × max{degf : f ∈ B} elements we have ∩w∈Ω ran(SB − w∗) = {0}
or, equivalently,
B = span(cid:8)ker(S∗
B − w) : w ∈ Ω(cid:9).
Now assume (A) of Theorem 1.1. Then, by the above observations, SB is a simple
symmetric operator and hence its defect numbers coincide and are equal to the
codimension of ranSB, see Lemma 2.6. It follows from Lemma 3.5 that the defect
16
B. ´CURGUS AND A. DIJKSMA
numbers of SB are also equal to the integer l introduced in Remark 2.2. Hence
l ∈ {1, . . . , min{d, n}}, where n = dim B. Now also assume (B) of Theorem 1.1.
Then l can be characterized in a different way. Indeed, by Theorem 1.2, there exist
a canonical subspace C ⊆ Cd[z] with degrees µ1 ≥ ··· ≥ µd ≥ 0 and a d × d matrix
polynomial W (z) with det W (α) 6= 0 such that B = W C. Since, by Lemma 2.6, we
have n − l = dimdomSB = dimranSB and since multiplication by z and by W (z)
commute, l is uniquely determined by the inequalities:
(4.1)
µ1 ≥ ··· ≥ µl ≥ 1
and µl+1 = ··· = µd = 0.
Proof of the if statements in Theorem 1.1. Assume (A) and (B). We show that B
has a reproducing Nevanlinna kernel in steps (i) -- (iv). In step (v) we prove the last
if statement in the theorem.
(i) By Theorem 1.2 there exist a canonical subspace C ⊆ Cd[z] with degrees µ1 ≥
··· ≥ µd ≥ 0 and a d × d matrix polynomial W (z) with det W (α) 6= 0 such that
B = W C. Then, by Remark 4.1, the defect numbers of the symmetric operator SB
are both equal to l, where l is determined by the inequalities (4.1). It follows that
the elements of B are of the form:
(4.2)
f (z) ∈ B ⇒ f (z) = W (z)(cid:20)x(z)
0 (cid:21) ,
where x(z) is an l × 1 vector polynomial and 0 denotes the zero vector of size
(d − l) × 1. Let n = dim B and let B(z) be a d × n matrix polynomial whose
columns form a basis of B. Let G be the corresponding Gram matrix and write the
reproducing kernel K(z, w) of B as K(z, w) = B(z)G−1B(w)∗, z, w ∈ C. By (B),
this representation implies that for each w ∈ C which belongs to the set in (3.12)
the columns of K(z, w) span an l-dimensional subspace of B, in formula:
(4.3) dim(cid:8)K(· , w)x : x ∈ Cd(cid:9) = l whenever w ∈(cid:8)α ∈ C : rank B(α) = l(cid:9).
(ii) In the following we use graph notation in the space B⊕ B. The operator SB is
identified with its graph in B⊕ B and its adjoint S∗
B is the orthogonal complement
of SB in B ⊕ B equipped with the Lagrange inner product
(cid:2)(cid:2){f, g},{p, q}(cid:3)(cid:3) = −i(cid:0)[g, p]B − [f, q]B(cid:1),
Let w ∈ C, x ∈ Cd and {f, Sf} ∈ SB be arbitrary. Then
{f, g},{p, q} ∈ B ⊕ B.
= 0
= −i(cid:0)x∗(Sf )(w) − wx∗f (w)(cid:1)
B ∩ (w∗I) for all w ∈ C. Ac-
cording to the definition of defect number (see [9, p.369]) and by (4.3), it follows
(cid:2)(cid:2){f, Sf},{K(·, w)x, w∗K(·, w)x}(cid:3)(cid:3) = −i(cid:0)(cid:2)Sf, K(·, w)x(cid:3)B −(cid:2)f, w∗K(·, w)x(cid:3)B(cid:1)
and hence (cid:8){K(·, w)x, w∗K(·, w)x} : x ∈ Cd(cid:9) ⊆ S∗
that for all w ∈(cid:8)α ∈ C \ R : rank B(α) = l(cid:9)
(cid:8){K(·, w)x, w∗K(·, w)x} : x ∈ Cd(cid:9) = S∗
B ∩ (w∗I),
L0 := span(cid:8){K(·, w)x, w∗K(·, w)x} : w ∈ C, x ∈ Cd(cid:9)
because for such w's both sets have dimension l. Consider the subspace
(4.4)
(4.5)
ON THE REPRODUCING KERNEL
17
of S∗
B. Since SB has no eigenvalues, the generalized von Neumann formula given
in [9, Theorem 3.7] implies that for d + 1 distinct points w0, . . . , wd from the set
(cid:8)α ∈ C : Imα > 0, rank B(α) = l(cid:9) we have
B ∩ (w∗
B = SB + S∗
S∗
0I) +
dXj=1
S∗
B ∩ (wj I).
Combined with (4.4) and (4.5) this yields SB + L0 ⊆ S∗
(4.6)
B = SB + L0.
S∗
B ⊆ SB + L0, and hence
z, w ∈ C.
(iii) Let(cid:0)B1, [· ,· ]B1(cid:1) be the reproducing kernel Pontryagin space whose kernel is
L1(z, w) = i (z − w∗)K(z, w),
B, [[· ,· ]](cid:1) → B1 defined by T(cid:0){f, g}(cid:1) = Sf − g, {f, g} ∈ S∗
the operator T :(cid:0)S∗
We claim that its positive and negative index are l. To prove the claim we consider
B,
and show that it is a partial isometry onto B1 with null space kerT = SB. The last
equality is easy to verify. That ran T = B1 follows from (4.6) as it implies (with
w ∈ C and x ∈ Cd):
That T is isometric follows from (4.6), the symmetry of SB and the equalities (with
w, v ∈ C and x, y ∈ Cd):
(cid:0)T(cid:0){K(·, w)x, w∗K(·, w)x}(cid:1)(cid:1)(z) = (z − w∗)K(z, w)x = −iL1(z, w)x.
(cid:2)(cid:2)(cid:8)K(·, w)x,w∗K(·, w)x(cid:9),(cid:8)K(·, v)y, v∗K(·, v)y(cid:9)(cid:3)(cid:3)
= y∗L1(v, w)x
= −i(cid:16)(cid:2)w∗K(·, w)x, K(·, v)y(cid:3)B −(cid:2)K(·, w)x, v∗K(·, v)y(cid:3)B(cid:17)
= i(cid:0)v − w∗(cid:1)y∗K(v, w)x
=(cid:2)−iL1(·, w)x,−iL1(·, v)y(cid:3)B1
=hT(cid:0){K(·, w)x, w∗K(·, w)x}(cid:1), T(cid:0){K(·, v)y, v∗K(·, v)y}(cid:1)iB1
.
The claim now follows because(cid:0)S∗
B/SB, [[· ,· ]](cid:1) is a Pontryagin space with positive
and negative index l (see [9, Theorem 2.3(c)]) and T establishes a unitary mapping
between this space and B1.
(iv) Let B1(z) be a d × 2l matrix polynomial whose columns form a basis of B1,
and let Q1 be the corresponding 2l × 2l Gram matrix. Then Q1 is self-adjoint and,
by the claim proved in (iii), has l positive and l negative eigenvalues. Let B2(z) be
the d × 2(d − l) matrix polynomial defined by
B2(z) = W (z)(cid:20) 0
Id−l
0
Id−l(cid:21) ,
where the zero matrices are of size l × (d − l). Define the d× 2d matrix polynomial
P(z) by P(z) =(cid:2)B1(z) B2(z)(cid:3) and the 2d × 2d block diagonal matrix Q by
0
iId−l
Q ="Q1
0 Q2# , where Q2 ="
0
0 # .
−iId−l
Then Q is self-adjoint and has d positive and d negative eigenvalues. We claim that
(I) rankP(z) = d for some z ∈ C, and
18
B. ´CURGUS AND A. DIJKSMA
(II) P(z)Q−1P(w)∗ = i(z − w∗)K(z, w) for all z, w ∈ C.
We prove (I): The inclusion B1 = T (S∗
exists an l × 2l matrix polynomial X(z) such that
B) ⊆ B + SB and (4.2) imply that there
B1(z) = W (z)(cid:20)X(z)
0 (cid:21) ,
where now 0 stands for the (d−l)×2l zero matrix. The complex number α satisfying
(B) belongs to the sets in (3.12) and (1.3) and hence
rankX(α) = rankB1(α)
= dimEαB1
= dimspan(cid:8)L1(α, w)x : w ∈ C, x ∈ Cd(cid:9)
= dimspan(cid:8)i(α − w∗)K(α, w)x : w ∈ C, x ∈ Cd(cid:9)
= dimspan(cid:8)K(α, w)y : w ∈ C, y ∈ Cd(cid:9)
= dimEαB
= rankB(α)
= l.
The equality
P(z) = W (z)(cid:20)X(z)
0
0
Id−l
0
Id−l(cid:21)
implies that rankP(α) = d. This proves (I). We prove (II):
P(z)Q−1P(w)∗ = B1(z)Q−1
1 B1(w)∗ + B2(z)Q−1
2 B2(w)∗
= L1(z, w) + W (z)(cid:20) 0
= i(z − w∗)K(z, w).
Id−l
0
Id−l(cid:21)(cid:20)
0
−iId−l
iId−l
0 (cid:21)(cid:20)0
0
Id−l
Id−l(cid:21) W (w)∗
Items (I) and (II) show that K(z, w) is a polynomial Nevanlinna kernel for B. This
completes the proof of the if statement.
(v) If (B) holds for all α ∈ C, then, by Theorem 3.4, W (z) is unimodular and the
proof of (I) shows that then rankP(z) = d for all z ∈ C.
4.2.
lemma.
In the proof of the only if statements in Theorem 1.1 we use the following
(cid:3)
Lemma 4.2. Let Q be a self-adjoint 2d × 2d matrix with d positive and d negative
eigenvalues. Let P(z) be a d × 2d matrix polynomial such that
(a) P(z)Q−1P(z∗)∗ = 0 for all z ∈ C,
(b) rankP(z) = d for all z ∈ C, and
(c) P(z) is row reduced and has row degrees σ1, . . . , σd, assumed ordered so that
σ1 ≥ ··· ≥ σd and σ1 = degP(z) =: p.
Equip C2d[z]<p with the inner product
[f, g]Q =
p−1Xj=0
b∗
p−1−j Q−1aj,
f (z) =
ajzj, g(z) =
p−1Xj=0
p−1Xj=0
bjzj, aj, bj ∈ C2d,
ON THE REPRODUCING KERNEL
19
and consider the following subspace of C2d[z]<p :
zp−1−kw∗kP(w)∗x : w ∈ C, x ∈ Cd).
Lp = span(p−1Xk=0
p =nf (z) ∈ C2d[z]<p : f (z) = P(z∗)∗u(z) with u(z) ∈ Cd[z]o.
Then the orthogonal complement of Lp in (C2d[z]<p, [· ,· ]Q) is
(4.7)
L⊥
p is a Pontryagin space with positive and
It is the isotropic part of Lp and Lp/L⊥
negative index σ1 + ··· + σd.
Proof. For an element f (z) = Pp−1
hold:
j=0 ajzj ∈ C2d[z]<p the following equivalences
f (z) ∈ L⊥
p ⇔ p−1Xk=0
w∗ka∗
k!Q−1P(w)∗ = 0 for all w ∈ C,
⇔ P(z)Q−1f (z) = 0
⇔ f (z) = P(z∗)∗uz
for all z ∈ C,
for some uz ∈ Cd and all z ∈ C.
The last equivalence follows from (a) and (b). To prove that the vector uz depends
polynomially on z we use that the Smith normal form (2.2) of P(z) is given by:
P(z) = U (z)(cid:2)Id 0(cid:3)V (z), where U (z) and V (z) are unimodular matrices. Then
f (z) = P(z∗)∗uz = V (z∗)∗(cid:20)Id
0(cid:3) U (z∗)−∗f (z)
0(cid:21) U (z∗)∗uz ⇒ uz = V (z∗)−∗(cid:2)Id
and the right hand side belongs to Cd[z]. This proves (4.7).
Since P(z∗)∗ has full rank for every z ∈ C, it acts as an injection on Cd[z],
dimL⊥
p = dim(cid:8)u(z) ∈ Cd[z] : deg(cid:0)P(z∗)∗u(z)(cid:1) < p(cid:9).
The number on the right hand side can be expressed in terms of the Forney indices
of P(z). Indeed, since P(z) is row reduced, it has the "predictable degree property"
(see Theorem 2.3):
therefore
(4.8)
Consequently, the space on the right hand side of in (4.8) equals
deg(cid:0)P(z∗)∗u(z)(cid:1) = max(cid:8)σj + deg uj(z) : j ∈ {1, . . . , d}(cid:9).
(cid:8)u(z) ∈ Cd[z] : deg uj(z) < p − σj, j ∈ {1, . . . , d}(cid:9),
p = dp +(cid:0)σ1 + ··· + σd(cid:1).
dim Lp = dim C2d[z]<p − dim L⊥
whose dimension is dp −(cid:0)σ1 + ··· + σd(cid:1). Hence dim L⊥
p = dp −(cid:0)σ1 + ··· + σd(cid:1) and
To prove the last two statements of the lemma we apply Lemma 2.5 with n = dp
and τ = dp− (σ1 +··· + σd). The assumptions about Q in the lemma readily imply
that C2d[z]<p is a 2dp-dimensional Pontryagin space with negative index dp.
It
remains to construct a maximal neutral subspace of C2d[z]<p which is contained in
Lp. We begin with the subspace H = ranH, where the operator H : Cd[z] → C2d[z]
u(z). For example, if P(z)
is written as:
maps u(z) ∈ Cd[z] into the polynomial part of P(cid:0)1/z∗(cid:1)∗
P(z) = P0 + zP1 + ··· + zpPp,
20
B. ´CURGUS AND A. DIJKSMA
then for k ∈ {0} ∪ N and x ∈ Cd
H(cid:0)zkx(cid:1) =
zjP ∗
k−j x
if k < p,
kXj=0
zk−p
zjP ∗
p−jx
if
p ≤ k.
pXj=0
{0} ∪ N and x, y ∈ Cd we have
These formulas imply that H is neutral in (cid:0)C2d[z]<p, [· ,· ]Q(cid:1). Indeed, for k, m ∈
hH(cid:0)zkx(cid:1), H(cid:0)zmy(cid:1)iQ
Xj+k=n
and the last expression equals 0 because the assumption (a) is equivalent to
k−jx = Xi+j=p−1−m+k
k = 0 for all n ∈ {0, . . . , 2p}.
min{k,m}Xj=0
y∗Pp−1−m+j Q−1P ∗
y∗PiQ−1P ∗
j x
Pj Q−1P ∗
i,j∈{0,...,p}
=
j,k∈{0,...,p}
(4.9)
From
Since, by (b), P0 = P(0) has full rank, H is degree preserving and hence injective.
Therefore, dim H = dp = (1/2) dim(cid:0)C2d[z]<p(cid:1) and H is maximal neutral.
subspace of(cid:0)C2d[z]<p, [· ,· ]Q(cid:1). The proof of the lemma is complete if we show that
Define the mapping R : C2d[z]<p → C2d[z]<p by (Rf )(z) = zp−1f (1/z). Then
R is unitary with respect to [· ,· ]Q and hence N := RH is also a maximal neutral
N ⊆ Lp. For that we consider the polynomials of the form
2p−1Xk=0
zkw∗kx, w ∈ C, x ∈ Cd.
z∗(cid:19)∗ 2p−1Xk=0
zkw∗kx! =
P(cid:18) 1
j−1Xk=0
2p−1Xk=0
H 2p−1Xk=0
zkw∗kx! =
zkw∗k! P ∗
2p−1Xk=0
2p−1−jXk=p
2p−1Xj=0 1
p−1Xk=0
zkw∗kP(w)∗x + higher order terms.
w∗k
zj−k + w∗j
Since the space Cd[z]<2p is spanned by polynomials in (4.9), each element of Cd[z]<p
is also a sum of polynomials in (4.9). As H is degree preserving, the polynomials
in H = H(cid:0)Cd[z]<p(cid:1) have degrees < p and therefore they have the form (4.10) with
zero higher order terms. Thus
p−1Xk=0
zkw∗(k+j)
zkw∗k =
zkw∗k +
we obtain
(4.10)
1
zj
and
zj
j x
H ⊂ span(p−1Xk=0
zkw∗kP(w)∗x : w ∈ C, x ∈ Cd)
and N = RH ⊆ Lp. This proves Lemma 4.2.
(cid:3)
Lp(z, w) = i (zp − w∗p)K(z, w) = p−1Xk=0
Then
Bp = span(P(z)Q−1
and
(cid:2)P(z)Q−1f, P(z)Q−1g(cid:3)Bp
where
z, w ∈ C.
zp−1−kw∗k!P(z)Q−1P(w)∗,
zp−1−kw∗kP(w)∗x : w ∈ C, x ∈ Cd)
p−1Xk=0
p−1Xk=0(cid:16)v∗(p−1−k)P(v)∗y(cid:17)∗
p−1Xk=0
zp−1−kv∗kP(v)∗y.
g(z) =
=
Q−1(cid:16)w∗kP(w)∗x(cid:17),
f (z) =
zp−1−kw∗kP(w)∗x,
p−1Xk=0
ON THE REPRODUCING KERNEL
21
Proof of the only if statements in Theorem 1.1. Assume that the reproducing ker-
nel of B is a polynomial Nevanlinna kernel K(z, w):
i (z − w∗)K(z, w) = P(z)Q−1P(w)∗
(4.11)
where Q is a self-adjoint 2d × 2d matrix with d positive and d negative eigenvalues
and P(z) is a d × 2d matrix polynomial with rankP(z) = d for some z ∈ C. Note
that (4.11) implies (a) of Lemma 4.2:
z, w ∈ C,
for all
(4.12)
P(z)Q−1P(z∗)∗ = 0 for all z ∈ C.
We prove (A) and (B) in the steps (i) -- (iv), in step (v) we prove the last only if
statement in the theorem.
(i) In this step we prove (B) under the assumption that (b) and (c) of Lemma 4.2
hold. Denote by Bp the reproducing kernel Pontryagin space with kernel
Comparing this inner product with the one defined in Lemma 4.2, we find that
P(z)Q−1 considered as a multiplication operator maps Lp ⊂ C2d[z]<p isometri-
cally onto Bp and its null space is L⊥
p (see the second of the three equivalences
in the beginning of the proof of Lemma 4.2). Hence, dim Bp = 2(σ1 + ··· + σd)
and the positive and the negative index of Bp equal σ1 + ··· + σd. According to
Lemma 2.7, we have dimB = σ1 +··· + σd. The space B is spanned by the columns
of K(z, w), w ∈ C and for j ∈ {1, . . . , d} the degree of the j-th row of K(z, w) as
Since both spaces have dimension σ1 + ··· + σd, equality prevails:
a polynomial in z is equal to max(cid:8)0, σj − 1(cid:9). Therefore B ⊆Ld
j=1(cid:0)C[z]<σj(cid:1)ed,k.
(4.13)
B0 =
This implies (B).
dMj=1(cid:0)C[z]<σj(cid:1)ed,k.
(ii) In this step we prove (A) under the assumption that (b) and (c) of Lemma 4.2
hold. Set
M (z) =
z−σ1
...
0
···
. . .
···
0
...
z−σd
.
22
B. ´CURGUS AND A. DIJKSMA
Then P∞ = limz→∞ M (z)P(z) and by (4.12) we have
(4.14)
P∞Q−1P∗
∞ = lim
z→∞
M (z)P(z)Q−1P(z∗)∗D(z∗)∗ = 0.
∞ is
Since P∞ has full rank, (4.14) implies that the linear span of the columns of P∗
a maximal neutral subspace of(cid:0)C2d, [· ,· ]Q(cid:1) and this span coincides with the null
P(z)a ∈ B ⇔ P∞a = 0.
space of P∞Q−1. We claim that for a ∈ C2d
(4.15)
To prove the claim assume first that P(z)a ∈ B. From (4.13) we see that the
degree of the j-th entry of the vector polynomial P(z)a is strictly less than σj, j ∈
notice that by the definition of P∞ the row degrees of the matrix polynomial
P0(z) = P(z) − M (z)−1P∞ are strictly less than σj, j ∈ {1. . . . , d}. By (4.13) we
have that P0(z)a ∈ B for all a ∈ C2d. Now assume P∞a = 0. Then
{1, . . . , d}. Hence P∞a = limz→∞ M (z)(cid:0)P(z)a(cid:1) = 0. As to the converse, first
P(z)a = P0(z)a + M (z)−1P∞a = P0(z)a ∈ B.
This completes the proof of (4.15).
Consider f ∈ B. Since B is finite dimensional it can be written as
(4.16)
f (z) =
K(z, wi)xi, m ∈ N, wi ∈ C, xi ∈ Cd, i ∈ {1, . . . , m}.
The next sequence of equivalences follows from (4.15) and the observation after
(4.14):
mXi=1
f ∈ dom SB ⇔ Sf ∈ B
(cid:2)Sf, g(cid:3)B −(cid:2)f, Sg(cid:3)B =
nXj=1
mXi=1
vjy∗
mXi=1
j K(vj, wi)xi −
mXi=1
nXj=1
wix∗
∗
i K(wi, vj)yj
i )K(z, wi)xi ∈ B
⇔
mXi=1
(z − w∗
⇔ P(z)Q−1 mXi=1
⇔ P∞Q−1 mXi=1
⇔
mXi=1
P(wi)∗xi = P∗
P(wi)∗xi! ∈ B
P(wi)∗xi! = 0
∞x for some x ∈ Cd.
Let f ∈ B be given by (4.16) and let g ∈ B be of the form
g(z) =
K(z, vj)yj, n ∈ N, vj ∈ C, yj ∈ Cd, j ∈ {1, . . . , n}.
Assume that f, g ∈ dom SB. Then there exist x, y ∈ Cd such that
P(vi)∗yi = P∗
∞y
P(wi)∗xi = P∗
∞x and
and using the reproducing kernel property of K(z, w) we have
nXj=1
mXi=1
ON THE REPRODUCING KERNEL
23
mXi=1
w∗
i y∗
j K(vj, wi)xi
nXj=1
j K(vj, wi)xi
=
=
vjy∗
j K(vj, wi)xi −
nXj=1
mXi=1
mXi=1(cid:0)vj − w∗
nXj=1
i(cid:1)y∗
mXi=1
nXj=1
j P(vj )!Q−1 mXi=1
= −i nXj=1
= −i
y∗
= −i y∗P∞Q−1P∗
= 0.
∞x
y∗
j P(vj)Q−1P(wi)∗xi
P(wi)∗xi!
This proves that SB is symmetric.
(iii) In this step we only assume (b) of Lemma 4.2: rankP(z) = d for all z ∈ C.
Then there is a unimodular d×d matrix polynomial U (z) such that S(z) = U (z)P(z)
is row reduced with ordered row degrees σ1 ≥ ··· ≥ σd. Then U is an isometry
from B onto the reproducing kernel Pontryagin space C with kernel
(4.17)
− i
According to what has already been proved in (i)
S(z)Q−1S(w)∗
.
z − w∗
U B = C =
dMj=1(cid:0)C[z]<σj(cid:1)ed,k.
Thus B = U −1C and, by Theorem 3.4, (B) holds for all α ∈ C. According to part
(ii) of this proof, SUB is symmetric, hence SB = U −1SUBU is also symmetric, that
is, (A) holds.
(iv) Finally we prove that (A) and (B) hold if rankP(z) = d for some z ∈ C as
in the beginning of this proof. In that case there exist a d × d matrix polynomial
G(z) with detG(z) 6≡ 0 and a d × 2d matrix polynomial S(z) with rankS(z) = d
for all z ∈ C such that P(z) = G(z)S(z) for all z ∈ C, see Lemma 2.4. If by A we
denote the reproducing kernel space with Nevanlinna kernel (4.17), then, by what
has been proved in (iii), the operator SA is symmetric and for almost all α ∈ C we
have ran(SA − α) = A ∩ ker Eα. Now (A) and (B) follow since the multiplication
operator G corresponding to G(z) is an isomorphism from A onto B.
(v) The last only if statement in the theorem follows from step (iii) above and
Theorem 3.4.
(cid:3)
Remark 4.3. Assume that B ⊂ Cd[z] is a Pontryagin space which satisfies the
conditions (A) and (B) of Theorem 1.1. Then, by Theorem 1.1, there is a gen-
eralized Nevanlinna pair {M (z), N (z)} such that the d × 2d matrix polynomial
P(z) = [M (z)N (z)] provides a representation (2.7), with Q given by (2.8), for the
Nevanlinna reproducing kernel K(z, w) of B. In addition, by Theorem 1.2 there
is a canonical subspace C such that B = W C for some d × d matrix polynomial
W (z) with detW (z) 6≡ 0. The proof of Theorem 1.1 and Lemma 2.4 show that the
multiset of the Forney indices of P(z) coincides with the multiset of the degrees of
24
B. ´CURGUS AND A. DIJKSMA
C. This implies that the Forney indices are independent of the Nevanlinna repre-
sentation (2.7) of the kernel K(z, w). In the special case when the defect numbers
of SB are equal to d this fact can also be proved directly by using [23, Theorem 1.3].
In view of Remark 4.1, this remark substantiates the observations about the Forney
indices and the defect numbers after Theorem 1.2 in the Introduction.
5. Q-functions
5.1. Let M (z) be a generalized Nevanlinna d × d matrix function and denote by
L(M ) the reproducing kernel Pontryagin space with reproducing kernel KM (z, w) =
KM,Id(z, w). By [16, Theorem 2.1], the operator S in L(M ) of multiplication by
the independent variable is a simple symmetric operator with equal defect numbers
and its adjoint is given by
S∗ = span(cid:8){KM (· , w∗)x, wKM (· , w∗)x} : x ∈ Cd, w ∈ hol(M )(cid:9)
=(cid:8){f, g} ∈ L(M )2 : ∃ x, y ∈ Cd such that g(z) − zf (z) ≡ x − M (z)y(cid:9).
It follows that for all w ∈ hol(M )
ker(S∗ − w) =(cid:8)KM (· , w∗)x : x ∈ Cd(cid:9) = ran E∗
w,
where Ew is considered as a mapping Ew : L(M ) → Cd. Taking orthogonal com-
plements we see that
ran(S − α) = L(M ) ∩ ker Eα, α ∈ hol(M ).
Thus (A) and (B) of Theorem 1.1 hold. Moreover, [16, Theorem 2.1] and its proof
imply that there is a constant invertible d × d matrix T such that
0(cid:21) ,
T M (z)T ∗ = M0 +(cid:20)cM (z) 0
0
Clearly, Q(z) depends on the choice of the pair {A, Γz} and if this choice has to be
mentioned explicitly we shall say that Q(z) is a Q-function for S associated with the
where M0 is a constant self-adjoint d × d matrix and, if the defect numbers of S
a Q-function for S. The theorem below concerns a converse implication. But first
we recall the notion of a Q-function.
are denoted by l, cM (z) is a generalized Nevanlinna l × l matrix function which is
Let S be a simple symmetric operator in a Pontryagin space K with defect
numbers equal to l. Let A be a self-adjoint extension of S in K with a nonempty
resolvent set ρ(A). Let µ ∈ ρ(A) \ R and define a function Γµ : Cl → K such that
it is a linear bijection from Cl onto ker(S∗ − µ). Finally, for z ∈ ρ(A) define the
defect mappings Γz : Cl → K by
Γz =(cid:0)I + (z − µ)(A − z)−1(cid:1) Γµ,
K = span{Γzc : z ∈ ρ(A) ∩ (C \ R), c ∈ Cl}
Then Γz is a bijection from Cl onto ker(S∗ − z),
(5.1)
and, by the resolvent identity, Γ∗
by definition an l × l matrix function that satisfies the equation
(5.2)
Q(z) − Q(w)∗
z ∈ ρ(A).
wΓz = Γ∗
= Γ∗
wΓz,
z, w ∈ ρ(A).
z − w∗
z∗ Γw∗, w, z ∈ ρ(A). A Q-function for S is
ON THE REPRODUCING KERNEL
25
pair {A, Γz}. Q(z) is uniquely determined up to an additive constant self-adjoint
d × d matrix Q0:
Q(z) = Q0 − iImµ Γ∗
µΓµ + (z − µ∗)Γ∗
µΓz, Q0 = Q∗
0.
From (5.1) and the defining relation (5.2) it follows that Q(z) is a generalized
Nevanlinna l × l matrix function with κ negative squares where κ is the negative
index of the Pontryagin space K; in particular Q(z)∗ = Q(z∗). Q-functions in an
indefinite setting were introduced and studied by M.G. Krein and H. Langer in
[28, 29].
5.2. The following theorem shows that the Nevanlinna pair {M (z), N (z)} of ma-
trix polynomials M (z) and N (z) in Theorem 1.1 can be chosen such that detN (z) 6≡
0 and such that N (z)−1M (z) is essentially the Q-function for SB. As before, by
L(Q) we denote the reproducing kernel space with reproducing kernel given by
(5.2).
Theorem 5.1. Let B be a finite dimensional Pontryagin subspace of Cd[z] for
which the conditions (A) and (B) of Theorem 1.1 hold. Denote by l ∈ {1, . . . d}
the equal defect numbers of the symmetric operator SB. Let Q(z) be an l × l
matrix Q-function for SB. Then there is a d × d matrix polynomial N (z) with
detN (z) 6≡ 0 such that M (z) = N (z)diag(Q(z), 0) is a d× d matrix polynomial and
B = N (L(Q) ⊕ {0}). In particular, {M (z), N (z)} is a Nevanlinna pair of matrix
polynomials and KM,N (z, w) is the reproducing kernel of B.
Proof. Assume (A) and (B) of Theorem 1.1. By Theorem 1.2 there is a d × d
matrix function W (z) with detW (z) 6≡ 0 such that B = W C, where C is a canonical
subspace of Cd[z]. By Remark 4.1 the defect numbers of the symmetric operator
SB are both equal to l with l ≤ d. We consider two cases: l = d and l < d.
(i) l = d. Let Q(z) be the Q-function for SB associated with the pair {A, Γz},
where A is a self-adjoint extension of SB and the defect mappings Γz are defined
above with l = d. Since SB is simple, the mapping
f 7→ g with f (z) = Γ∗
z∗ Γw∗ x,
g(z) = (Γw∗x) (z),
x ∈ Cd, w ∈ ρ(A),
can be extended by linearity to a unitary mapping U from L(Q) onto B. That U
is isometric follows from
(cid:2)Γ∗
wΓz x, Γ∗
vΓz y(cid:3)L(Q) = y∗Γ∗
wΓv x = y∗Γ∗
v∗ Γw∗ x = [Γw∗x, Γv∗ y]B,
x, y ∈ Cd.
We claim that U is the operator of multiplication by a d× d matrix function. To
prove the claim we use the equality B = W C. Since the defect numbers of SB are
equal to d, the degrees of C are all ≥ 1 (see Remark 4.1) and hence the d columns
of W (z) belong to B and are linearly independent over C. We denote by Γ∗
z∗ W the
d × d matrix function defined by
z∗ W )x = Γ∗
z∗ (W x),
(Γ∗
x ∈ Cd, z ∈ ρ(A).
We show that its inverse exists for z ∈ Ω := ρ(A) ∩ {z ∈ C : detW (z) 6= 0}.
Suppose there is an x ∈ Cd such that (Γ∗
z∗ W ) x = 0. Then for all y ∈ Cd
0 =(cid:2)(Γ∗
z∗ W ) x, y(cid:3)Cd =(cid:2)Γ∗
z∗(W x), y(cid:3)Cd =(cid:2)W x, Γz∗ y(cid:3)B,
hence W x ∈ ker(S∗
follows that x = 0. This proves that (Γ∗
B − z∗)⊥ = ran(SB − z) = B ∩ Ez. That is, W (z)x = 0 and it
z∗ W )−1 is well defined for all z ∈ Ω. We set
26
B. ´CURGUS AND A. DIJKSMA
N (z) = W (z)(Γ∗
with multiplication by N (z) if we have proved that
z∗ W )−1. Clearly, detN (z) 6≡ 0. We have shown that U coincides
N (z)Γ∗
z∗Γw∗ x = (Γw∗x) (z),
x ∈ Cd, w ∈ ρ(A), z ∈ Ω,
or, equivalently, that with y(z, w, x) := W (z)−1 (Γw∗x) (z) ∈ Cd
Γ∗
z∗ Γw∗ x = Γ∗
z∗ (W y(z, w, x)) ,
x ∈ Cd, w ∈ ρ(A), z ∈ Ω.
But this equality holds, since Γ∗
z∗ (ran(SB − z)) = {0} and
Γw∗ x − W y(z, w, x) ∈ B ∩ Ez = ran(SB − z).
This completes the proof of the claim that U is multiplication by N (z). It follows
from [6, Theorem 1.5.7] and its proof that the formula for the kernel K(z, w) of B
is given by
K(z, w) = N (z)
Q(z) − Q(w)∗
N (w)∗
z − w∗
and hence B = NL(Q).
It remains to show that M (z) = N (z)Q(z) and N (z) are matrix polynomials.
Since the elements of the space B are polynomials, the matrix function z 7→ K(z, w)
is a matrix polynomial, hence the matrix function
M (z) − N (z)Q(w)∗ = N (z)Q(z) − N (z)Q(w)∗ = (z − w∗)K(z, w)N (w)−∗
is a matrix polynomial in z. Thus if N (z) is a matrix polynomial, then so is M (z).
It remains to show that N (z) is a polynomial. For this we note that the above
formula implies that for x ∈ Cd
N (z)
(z − µ∗)K(z, µ)N (µ)−∗ − (z − w∗)K(z, w)N (w)−∗
Q(µ∗) − Q(w)∗
x =
x.
µ∗ − w∗
µ∗ − w∗
The right hand side is a matrix polynomial in z and hence it follows from the
equality that N (z) is a matrix polynomial if we can show that
To prove this equality we argue by contradiction and suppose it is not true. Then
there is a nonzero vector x ∈ Cd orthogonal to the set on the right hand side, that
is,
µΓw∗ x : w ∈ ρ(A) ∩ (C \ R), x ∈ Cd(cid:9).
Cd = span(cid:8)Γ∗
(cid:2)Γw∗ y, Γµx(cid:3)B = 0, w ∈ ρ(A) ∩ (C \ R), y ∈ Cd.
Since SB is simple and Γµ is injective, we find that Γµx = 0 and that x = 0, which
contradicts the choice of the nonzero vector x.
(ii) l < d. Then C = C1 ⊕ {0}, where C1 is a canonical subspace of Cl[z] of which
the degrees are all ≥ 1. Using the relation B = W (C1 ⊕ {0}) we equip C1 with
an indefinite inner product that makes W an isomorphism. Then SB and SC1 are
isomorphic: SC1 = W SBW −1, hence SC1 is symmetric and has defect numbers
equal to l. Thus (A) holds and it is not difficult to verify that also (B) holds on
C1. Finally, since Q(z) is the Q-function for SB associated with the pair {A, Γz},
Q(z) is the Q-function for SC1 associated with the pair {W −1AW, W −1Γz}. This
all shows that we may apply part (i) of this proof (with W (z) = Il): There exists
an l× l matrix polynomial N1(z) with detN1(z) 6≡ 0 such that N1(z)Q(z) is an l× l
matrix polynomial and C1 = N1L(Q). It follows that if
N (z) = W (z) diag(N1(z), Id−l),
ON THE REPRODUCING KERNEL
27
then detN (z) 6≡ 0, N (z) diag(Q(z), 0) is a d × d matrix polynomial and B =
N (L(Q) ⊕ {0}).
(cid:3)
6. Corollaries and examples
In the next corollary we extend Theorem 1.1 to finite dimensional Pontryagin
spaces of rational vector functions. A rational Nevanlinna kernel is a kernel of the
form KM,N (z, w) as in (1.6), in which M (z) and N (z) are rational d × d matrix
functions satisfying (1.4) and (1.5).
Corollary 6.1. Let B be a finite dimensional Pontryagin space of rational d × 1
vector functions and let Ω ⊂ C be the finite set of all the poles of the functions in B.
Denote by SB the operator of multiplication by the independent variable in B and
by Eα the operator of evaluation at a point α ∈ C. Then the reproducing kernel of
B is a rational Nevanlinna kernel if and only if the following two conditions hold:
(a) The operator SB is symmetric in B.
(b) For some α ∈ C \ Ω we have ran(cid:0)SB − α(cid:1) = B ∩ ker Eα.
Proof. Assume (a) and (b). Let q(z) be the monic scalar polynomial of minimal
degree such that B′ := {q(z)f (z) : f ∈ B} consists of polynomials. Equip B′
with the Pontryagin space inner product that makes the mapping q : B → B′ of
multiplication by q(z) a unitary mapping. Then items (A) and (B) of Theorem 1.1
hold for B′. Hence B′ has a polynomial reproducing Nevanlinna kernel KM,N (z, w).
It follows that B has reproducing kernel KM/q,N/q(z, w).
Now assume KM,N (z, w) is a rational reproducing Nevanlinna kernel of B. Let
r(z) be a polynomial such that r(z)M (z) and r(z)N (z) are polynomials and hence
form a polynomial Nevanlinna pair {r(z)M (z), r(z)N (z)}. Then KrM,rN (z, w) is a
polynomial reproducing Nevanlinna kernel of the space B′′ := {r(z)f (z) : f (z) ∈
B} equipped with the inner product that makes multiplication by r(z) an isomor-
phism from B onto B′′. Since the elements of B′′ are polynomials, we can apply
Theorem 1.1 to conclude that items (A) and (B) hold for the space B′′. Since
multiplication by r(z) and by z commute, (A) implies (a). By Theorem 1.2 and
(1.3), the equality
(6.1)
holds for all but finitely many α ∈ C. Choose α ∈ C \ Ω such that (6.1) is valid.
Then for this α item (b) holds.
ran(cid:0)SB′′ − α(cid:1) = B′′ ∩ ker Eα
(cid:3)
Corollary 6.2. Let (B, [· , · ]B) be a finite dimensional Pontryagin subspace of
Cd[z] whose reproducing kernel is a Nevanlinna kernel determined by a generalized
Nevanlinna pair. Let J be a fundamental symmetry on B. Then the Hilbert space
(B, [J · , · ]B) has a reproducing Nevanlinna kernel determined by a Nevanlinna pair
if and only if SB is symmetric in this space.
The corollary follows from Theorem 1.1, because condition (B) is independent
of the topology on B.
Example 6.3. Consider the subspace B of C2[z] spanned by the columns of the
matrix
B(z) =(cid:20)1
0
0
1(cid:21)
z
0
z2
0
28
B. ´CURGUS AND A. DIJKSMA
and equipped with the inner product [· , · ]B so that
G =
1
0
0
0
0
1
0
0
0 0
0 0
0 1
1 0
and J =
.
is the Gram matrix associated with B(z): G = [B, B]B. The spectral decomposition
of G is G = U JU ∗ with unitary matrix
0 √2
0 0
−1 1
0
0
0 0 √2
0
0
1 1
0
−1 0 0
0
1 0
0 1
0
0
0 0
1
√2
0
0
0
1
U =
It follows that (B, [· , · ]B) is a Pontryagin space with positive index 3 and negative
index 1. The equality [BU, BU ]B = J defines a fundamental decomposition of B
with corresponding fundamental symmetry J determined by J BU = BU J. In the
Hilbert space inner product [· , · ]J := [J · , · ]B we have [BU, BU ]J = J 2 = In
and hence [B, B]J = In. The operator SB is symmetric in the Pontryagin space
(B, [· , · ]B), but not in the Hilbert space (B, [· , · ]J ). Since B is a canonical
subspace of C2[z], Theorem 1.2 implies that Theorem 1.1 (B) holds in B. Hence,
according to Theorem 1.1, the Pontryagin space (B, [· , · ]B) has a reproducing
Nevanlinna kernel, whereas the reproducing Hilbert space (B, [· , · ]J ) does not
have a reproducing Nevanlinna kernel.
(cid:3)
Corollary 6.4. Let B be a finite dimensional Pontryagin subspace of Cd[z] whose
reproducing kernel is a Nevanlinna kernel. Let B0 be a Pontryagin subspace of B.
Then the reproducing kernel of B0 is a Nevanlinna kernel if and only if for some
α ∈ C we have ran(cid:0)SB0 − α(cid:1) = B0 ∩ ker Eα.
The corollary follows from Theorem 1.1, because the hypothesis implies that
SB0, being a subset of SB, is symmetric in B0, that is, that (A) holds for SB0.
Example 6.5. Consider the Hilbert subspace
B0 = span(cid:26)(cid:20)1
0(cid:21) ,(cid:20)z2
0(cid:21)(cid:27)
of the space B in Example 6.3. Then for arbitrary α ∈ C we have
ran(cid:0)SB0 − α(cid:1) = {0} and B0 ∩ ker Eα = span(cid:26)(cid:20)z2 − α2
Thus the condition ran(cid:0)SB0 −α(cid:1) = B0∩ker Eα does not hold for any α ∈ C. Hence
Corollary 6.4 implies that the reproducing kernel of B0, which is calculated to be
(cid:21)(cid:27) .
0
K(z, w) ="1 + z2w∗2
0
0
0# ,
z, w ∈ C,
is not a Nevanlinna kernel. This fact can be verified using [23, Theorem 1.3]. First
observe that for all z, w ∈ C we have
(z − w∗)K(z, w) = M (z)N (w)∗ − N (z)M (w)∗ =(cid:2)M (z) N (z)(cid:3)" N (w)∗
−M (w)∗# ,
ON THE REPRODUCING KERNEL
29
where
N (z) =(cid:20)1
0
z2
0(cid:21)
and M (z) = zN (z),
z ∈ C.
Now [23, Theorem 1.3 and Section 4] imply that for any 2 × 2 matrix polynomials
M1(z) and N1(z) such that
(6.2)
there exists a 4 × 4 invertible matrix S such that
(z − w∗)K(z, w) =(cid:2)M1(z) N1(z)(cid:3)" N1(w)∗
−M1(w)∗# ,
(cid:2)M1(z) N1(z)(cid:3) =(cid:2)M (z) N (z)(cid:3)S,
z ∈ C.
z, w ∈ C,
Hence (6.2) yields that rank(cid:2)M1(z) N1(z)(cid:3) = 1 for all z ∈ C. Consequently
K(z, w) is not a Nevanlinna kernel. Using the same results from [23] one can also
show that the scalar reproducing kernel K(z, w) = 1 + z2w∗2 of the Hilbert space
with orthonormal basis {1, z2} is not a Nevanlinna kernel.
✷
We end the paper with two examples in which det K(z, w) ≡ 0. These examples
also show that the proof of Theorem 5.1 is constructive.
Example 6.6. Consider the space B with reproducing kernel
K(z, w) =
−1
0
0 −w∗
0
0
−1 −z
0 .
We show that, even though detK(z, w) ≡ 0, the kernel is a Nevanlinna kernel. We
follow the proof of the first part of Theorem 5.1 and construct two Nevanlinna pairs
that determine K(z, w). The space B is spanned by the columns of the 3× 4 matrix
polynomial
It follows that B is a canonical subspace of C3[z] with degrees 1, 1 and 2. The
Gram matrix associated with B(z) is given by
hence B is a Pontryagin space with positive and negative index 2. The operator of
multiplication by z on B is given by
It is easy to see that (A) and (B) of Theorem 1.1 are satisfied. The defect numbers
of SB are both equal to 3, see Remark 4.1. The Q-function of SB associated with
the self-adjoint extension
0
0
0
0
1
−I
1 0
0 1
0 0
B(z) =
z .
G = [B, B]B =(cid:20) 0 −I
0 (cid:21) ,
: a ∈ C
SB =
, B
B
: a, b, c, d ∈ C
A =
, B
B
a
0
b
0
0
0
a
0
0
0
0
a
0
c
0
d
.
30
B. ´CURGUS AND A. DIJKSMA
is
of SB (which is multi-valued and has a nonempty resolvent set) and the defect
mappings Γz : C3 → ker(S∗ − z) defined by
Γz =(cid:0)I + (z − i)(A − z)−1(cid:1) Γi = B
Q(z) = Q0 +
0
1/z
−iz
1/z
0
0
iz
0
i/z
i
0
0
0
0
i/z
0
0
0
0
1
0.
0 , Q0 = Q∗
z2 and Y (z) =
0
0
0 −iz
0
0 −1
0
−iz
0
0
X(z) = Y (z)Q0 +
−i
0
iz
0
0 −i
We find that X(z) = Y (z)Q(z) and Y (z) = (Γ∗
linna pair of matrix polynomials:
z∗ I)−1 form a full generalized Nevan-
such that K(z, w) = KX,Y (z, w) and (cid:2)X(z) Y (z)(cid:3) is row reduced with Forney
indices 1, 1 and 2, which is in accordance with Remark 4.3.
The Q-function associated with the self-adjoint operator extension A of SB and
defect mappings Γz defined by
AB = B
0
0
0
0
1 0
0 0
0 0
0 1
0
0
0
0
, Γz = B
−1/z2
−1/z
0
0
0
0
i/z
−(z − i)/z2
0
0
0
i/z
, Q1 = Q∗
1.
is given by
0
0
0
z2
−i
z
z2
i
z
iz2 + z + i
Q(z) = Q1 +
−iz2 + z − i
Again we find that M (z) = N (z)Q(z) and N (z) = (Γ∗
M (z) = N (z)Q1 +
0 −iz2 + z − i −iz , N (z) =
(cid:2)M (z) N (z)(cid:3) is row reduced with Forney indices 1, 1 and 2.
0
0
0
0
z
1
0
0
which form a full generalized Nevanlinna pair such that K(z, w) = KM,N (z, w) and
(cid:3)
Two generalized Nevanlinna pairs {X(z), Y (z)} and {M (z), N (z)} of d×d matrix
polynomials define the same Nevanlinna kernel if one is a J-unitary transformation
of the other, that is, if they are connected via the formulas
z∗ I)−1 are matrix polynomials:
0 −iz
0
0
z2
0
z + i
−iz
0
M (z) = X(z)A + Y (z)C, N (z) = X(z)B + Y (z)D,
where A, B, C and D are constant d × d matrices such that if we set
U =(cid:20)A B
C D(cid:21)
and J =(cid:20) 0
−iId
iId
0(cid:21) ,
ON THE REPRODUCING KERNEL
31
then U is J-unitary: U JU ∗ = J. The pairs {X(z), Y (z)} and {M (z), N (z)} in Ex-
ample 6.6 (with arbitrary constant self-adjoint matrices Q0 and Q1) are connected
via a J-unitary transformation. The following example shows that the converse of
the foregoing statement does not hold.
Example 6.7. Consider the Nevanlinnna pair {X(z), Y (z)} given by
X(z) = diag(z, 0, z2),
Y (z) = diag(0, z, z).
it is spanned by B(z) = (cid:2)0
Then the space B with reproducing kernel KX,Y (z, w) = diag(0, 0, zw∗) is a 1-
dimensional Hilbert space:
and the corre-
sponding Gram matrix is G = [B, B]B = 1. Note that detX(z) ≡ 0, detY (z) ≡ 0
that the pair {X(z), Y (z)} can be replaced by a Nevanlinna pair {M (z), N (z)} such
that detN (z) 6≡ 0.
Since P(z) does not have full rank for all z ∈ C, to calculate the Forney indices
we must first apply Lemma 2.4. We write P(z) as P(z) = G(z)T(z) with
and P(z) :=(cid:2)X(z) Y (z)(cid:3) does not have full rank at z = 0: P(0) = 0. We show
0 z(cid:3)⊤
G(z) =
0
0
z
0 z
z
0
0
0 , T(z) =
0
0
1
0 z
0
0
0
0 0
0 0
0
1
0
1
0
0 .
Since T(z) has full rank for all z ∈ C and is row reduced, the Forney indices of
P(z) are those of T(z) and they are µ1 = 1, µ2 = µ3 = 0. This fits in well with the
observations after Theorem 1.2 indicating that dim B = 1 and the defect numbers
of the symmetric operator SB = {{0, 0}} are both equal to 1. We follow part (ii)
of the proof of Theorem 5.1 and write B = W (C1 ⊕ {0}) with
W (z) =
0
0
z
0
1
0
1
0
0
and C1 = C. We make the multiplication operator W an isometry when C is
equipped with the Euclidean inner product. Then SC1 = {{0, 0}} is symmetric. The
defect subspaces ker(S∗
C1 − z) all coincide with C and A is a self-adjoint extension
of SC1 if and only if A = Am, the operator of multiplication by m, m ∈ R, or
and relations have a non-empty resolvent set. Choose µ ∈ C \ R, γ ∈ C \ {0} and
define Γµ : C → ker(S∗
C1 − µ) = C by Γµx = γx, x ∈ C. Then the Q-function q(z)
of SC1 associated with {A, Γz} is given by
A = Arel =(cid:8){0, c} : c ∈ C(cid:9). Since C1 is a Hilbert space, all self-adjoint operators
where q0 is an arbitrary real number, and
q(z) = q0 +
z∗ 1)−1 =
(Γ∗
m − µ2γ2
m − z
γ2z
if A = Am,
if A = Arel,
m − z
γ∗(m − µ∗)
1
γ∗
if A = Am,
if A = Arel.
32
B. ´CURGUS AND A. DIJKSMA
We find that KX,Y (z, w) = KM,N (z, w) with matrix polynomials
M (z) = N (z)
q(z) 0 0
0 0
0
0
0 0 and N (z) = W (z)
(Γ∗
z∗ 1)−1
0
0
0 0
1 0
0 1
.
It is easy to see that the Nevalinna pairs {X(z), Y (z)} and {M (z), N (z)} are not
related via a J-unitary transformation.
(cid:3)
References
[1] D. Alpay, handwritten manuscript (French), unpublished.
[2] D. Alpay, A structure theorem for reproducing kernel Pontryagin spaces, J. Comput.
Appl. Math. 99 (1998) 413 -- 422.
[3] D. Alpay, The Schur algorithm, reproducing kernel spaces and system theory, SMF/AMS
Texts and Monographs 5, Panorama et Synth`eses 6, 1998.
[4] D. Alpay, P. Bruinsma, A. Dijksma, H.S.V. de Snoo, Interpolation problems, extensions
of symmetric operators and reproducing kernel spaces I, Oper. Theory Adv. Appl. 50,
Birkhauser, Basel, 1991, pp. 35 -- 82.
[5] D. Alpay, P. Bruinsma, A. Dijksma, H.S.V. de Snoo, Interpolation problems, extensions of
symmetric operators and reproducing kernel spaces II, Integral Equations Operator Theory
14 (1991) 465 -- 500; missing Section 3, Integral Equations Operator Theory 15 (1991) 378 -- 388.
[6] D. Alpay, A. Dijksma, J. Rovnyak, H.S.V. de Snoo, Schur functions, operator colligations,
and reproducing kernel Pontryagin spaces, Oper. Theory Adv. Appl. 96, Birkhauser, Basel,
1997.
[7] D. Alpay, H. Dym, On applications of reproducing kernel spaces to the Schur algorithm and
rational J unitary factorization, Oper. Theory Adv. Appl. 18, Birkhauser, Basel, 1986, pp.
89 -- 159.
[8] A. Amirshadyan, V. Derkach, Interpolation in generalized Nevanlinna and Stieltjes classes,
J. Operator Theory 42 (1999) 145 -- 188.
[9] T.Ya. Azizov, B. ´Curgus, A. Dijksma, Standard symmetric operators in Pontryagin spaces:
A generalized von Neuman formula and minimality of boundary coefficients, J. Funct. Anal.
198 (2003) 361 -- 412.
[10] T.Ya. Azizov, B. ´Curgus, A. Dijksma, Eigenvalue problems with boundary conditions de-
pending polynomially on the eigenparameter (tentative title), in preparation.
[11] J. Behrndt, V.A. Derkach, S. Hassi, H.S.V. de Snoo, A realization theorem for generalized
Nevanlinna families, to appear in Oper. Matrices.
[12] L. de Branges, Hilbert spaces of entire functions, Prentice Hall, Englewood Cliffs, 1968.
[13] L. de Branges, The expansion theorem for Hilbert spaces of entire functions, in: Entire
Functions and Related Parts of Analysis (Proc. Sympos. Pure Math., La Jolla, Calif., 1966),
Amer. Math. Soc., Providence, 1968, pp. 79 -- 148.
[14] L. de Branges, J. Rovnyak, Canonical models in quantum scattering theory, in: Perturbation
Theory and its Applications in Quantum Mechanics (Proc. Adv. Sem. Math. Res. Center,
U.S. Army, Theoret. Chem. Inst., Univ. of Wisconsin, Madison, Wis., 1965) Wiley, New York,
1966, pp. 295 -- 392.
[15] B. ´Curgus, A. Dijksma, T.T. Read, The linearization of boundary eigenvalue problems and
reproducing kernel Hilbert spaces, Linear Algebra Appl. 329 (2001) 97 -- 136.
[16] A. Dijksma, H. Langer, A. Luger, Y. Shondin, Minimal realizations of scalar generalized
Nevanlinna functions related to their basic factorization, Oper. Theory Adv. Appl. 154,
Birkhauser, Basel, 2004, pp. 69 -- 90.
[17] A. Dijksma, H. Langer, H.S.V. de Snoo, Selfadjoint Πκ-extensions of symmetric subspaces:
an abstract approach to boundary problems with spectral parameter in the boundary condi-
tions, Integral Equations Operator Theory 7 (1984) 459 -- 515; Addendum, Integral Equations
Operator Theory 7 (1984) 905.
[18] A. Dijksma, H. Langer, H.S.V. de Snoo, Hamiltonian systems with eigenvalue depending
boundary conditions, Oper. Theory Adv. Appl. 35, Birkhauser, Basel, 1988, pp. 37 -- 83.
[19] A. Dijksma, H.S.V. de Snoo, Symmetric and selfadjoint relations in Krein spaces I, Oper.
Theory Adv. Appl. 24, Birkhauser, Basel, 1987, pp. 145 -- 166.
ON THE REPRODUCING KERNEL
33
[20] H. Dym, D. Volok, Zero distribution of matrix ploynomials, Linear Algebra Appl. 425 (2007)
714 -- 738.
[21] P. Fuhrmann, On duality in some problems of geometric control. Acta Appl. Math. 91 (2006)
207 -- 251.
[22] F.R. Gantmacher, Matrizentheorie, VEB Deutscher Verlag der Wissenschaften, Berlin, 1986.
[23] I. Gohberg, T. Shalom, On Bezoutians of nonsquare matrix polynomials and inversion of
matrices with nonsquare blocks, Linear Algebra Appl. 137-138 (1990) 249 -- 323.
[24] I.S. Iohvidov, M.G. Krein, H. Langer, Introduction to the spectral theory of operators in
spaces with an indefinite metric, Akademie-Verlag, Berlin, 1982.
[25] T. Kailath, Linear systems, Prentice-Hall, Englewood Cliffs, 1980.
[26] M. Kaltenback, H. Woracek, Pontryagin spaces of entire functions I, Integral Equations Op-
erator Theory 33 (1999) 34 -- 64.
[27] M. Kaltenback, H. Woracek, Pontryagin spaces of entire functions V, Acta Sci. Math. (Szeged)
77 (2011) 179 -- 292.
[28] M.G. Krein, H. Langer, Defect subspaces and generalized resolvents of an Hermitian operator
in the space Πκ, Funct. Anal. Appl. 5 (1971) 136 -- 146; ibid 5 (1971) 217 -- 228.
[29] M.G. Krein, H. Langer, Uber die Q-Funktion eines π-hermiteschen Operators im Raume Πκ,
Acta Sci. Math. (Szeged) 34 (1973) 191 -- 230.
[30] M.G. Krein, H. Langer, Uber einige Fortsetzungsprobleme, die eng mit der Theorie her-
mitescher Operatoren im Raume Πκ zusammenhangen. I. Einige Funktionenklassen und ihre
Darstellungen, Math. Nachr. 77 (1977) 187 -- 236.
[31] M.G. Krein, H. Langer, Some propositions on analytic matrix functions to the theory of
operators in the space Πκ, Acta. Sci. Math. (Szeged) 43 (1981) 181 -- 205.
[32] H. Elton Lacey, The Hamel dimension of any infinite dimensional separable Banach space
is c, Amer. Math. Monthly 80 (3) (1973) 298.
[33] R. J. McEliece, The algebraic theory of convolutional codes. Handbook of coding theory, Vol.
I, II, North-Holland, Amsterdam, 1998, pp. 1065 -- 1138.
[34] Xian-Jin Li, The Riemann hypothesis for polynomials orthogonal on the unit circle, Math.
Nachr. 166 (1994) 229 -- 258.
[35] Xian-Jin Li, On Reproducing Kernel Hilbert Spaces of Polynomials, Math. Nachr. 186 (1997)
115 -- 148.
Department of Mathematics, Western Washington University, Bellingham, WA 98225,
USA
E-mail address: [email protected]
Johann Bernoulli Institute of Mathematics and Computer Science, University of
Groningen, P.O. Box 407, 9700 AK Groningen, The Netherlands
E-mail address: [email protected]
|
1609.01083 | 1 | 1609 | 2016-09-05T10:26:42 | Approaching bilinear multipliers via a functional calculus | [
"math.FA"
] | We propose a framework for bilinear multiplier operators defined via the (bivariate) spectral theorem. Under this framework we prove Coifman-Meyer type multiplier theorems and fractional Leibniz rules. Our theory applies to bilinear multipliers associated with the discrete Laplacian on $\mathbb{Z}^d,$ general bi-radial bilinear Dunkl multipliers, and to bilinear multipliers associated with the Jacobi expansions. | math.FA | math | APPROACHING BILINEAR MULTIPLIERS VIA A FUNCTIONAL
CALCULUS
BŁAŻEJ WRÓBEL
Abstract. We propose a framework for bilinear multiplier operators defined via the (bi-
variate) spectral theorem. Under this framework we prove Coifman-Meyer type multiplier
theorems and fractional Leibniz rules. Our theory applies to bilinear multipliers associated
with the discrete Laplacian on Zd, general bi-radial bilinear Dunkl multipliers, and to bilinear
multipliers associated with the Jacobi expansions.
6
1
0
2
p
e
S
5
]
.
A
F
h
t
a
m
[
1
v
3
8
0
1
0
.
9
0
6
1
:
v
i
X
r
a
The theory of spectral multipliers is now a well established and vast branch of linear harmonic
analysis. Its origins lie in trying to extend the Fourier multiplier operators on R given by
1. Introduction
f 7→
1
2π ZR
m(ξ) f (ξ)eixξ dξ,
x ∈ R,
to other settings. Here m is a bounded function on R while f (ξ) = RR f (x)e−ixξ dx, ξ ∈ R.
For a self-adjoint operator L its spectral multipliers are the operators m(L) defined by the
spectral theorem. In the Fourier case L is merely i d
dx . As in the Fourier case the boundedness
of m(L) on L2 is equivalent with the boundedness of m. The main task in the theory of spectral
multipliers is to extend the boundedness of m(L) to Lp, for some 1 < p < ∞, p 6= 2.
The bilinear multipliers for the Fourier transform are the operators
(1.1)
Fm(f1, f2)(x) =
m(ξ1, ξ2) f1(ξ1) f2(ξ2)eix(ξ1+ξ2) dξ,
x ∈ R,
1
4π2 ZZR2
with m : R2 → C being a bounded function. As far as we know, in the bilinear case, there
has been no systematic approach to extend the operators Fm outside of the Fourier transform
setting. The main idea behind the creation of this paper is to provide a theory for bilinear mul-
tipliers defined by the (bivariate) spectral theorem that parallels the correspondence between
the linear Fourier multipliers and spectral multipliers.
Our starting point is the observation that (1.1) may be rephrased as
Fm(f1, f2)(x) = m(i∂1, i∂2)(f1 ⊗ f2)(x, x),
x ∈ R;
here ∂1, ∂2 denote the partial derivatives, while m(i∂1, i∂2) is defined by the bi-variate spectral
theorem. Note that ∂1 = ∂ ⊗ I and ∂2 = I ⊗ ∂, where ∂ denotes the derivative on R, while I
is the identity operator. We investigate the possibility of replacing i∂1 and i∂2 by some other
operators L1 = L ⊗ I and L2 = I ⊗ L. The bilinear multipliers we consider are of the form
(1.2)
Bm(f1, f2)(x) = m(L1, L2)(f1 ⊗ f2)(x, x),
x ∈ X.
2010 Mathematics Subject Classification. 42B15, 47B38, 26D10.
Key words and phrases. bilinear multiplier, joint spectral theorem, fractional Leibniz rule.
1
2
BŁAŻEJ WRÓBEL
Here L is a self-adjoint non-negative operator on L2(X, ν), and m(L1, L2) is defined by the
bi-variate spectral theorem. We also assume that L is injective on its domain, and that the
contractivity condition (CT) (see p. 4) and the well definiteness condition (WD) (see p. 5) are
satisfied. These assumptions should be regarded as technical ones. The main assumptions on
L that are in force in this paper are the existence of a Mikhlin-Hörmander functional calculus
(MH), see p. 4, together with a product formula for the spectral multipliers of L, see (PF) on
p. 6.
There are two main goals of our paper. Firstly, we would like to prove Coifman-Meyer type
multiplier theorems outside of the Fourier transform setting. Secondly, we would like to apply
these results to obtain fractional Leibniz rules.
The classical Coifman-Meyer multiplier theorem [7] says that the Mikhlin-Hörmander con-
dition supξ∈R2 ξα1+α2∂αm(ξ) ≤ Cα, α ∈ N2, implies the boundedness of Fm from Lp1 × Lp2
to Lp, 1/p = 1/p1 + 1/p2, p1 > 1, p2 > 1, p > 1/2. This was proved by Coifman and Meyer for
p > 1, while for p > 1/2 it is due to Grafakos and Torres [12] and Kenig and Stein [15]. There
are also Coifman-Meyer type multiplier theorems which are known in settings other than the
Fourier transform. For bilinear multipliers on the torus a theorem of Coifman-Meyer type may
be deduced from Fan and Sato [9, Theorems 1-3]. Similarly, for bilinear multipliers on the
integers such a theorem follows from Blasco [5, Theorem 3.4]. Next, in the product Dunkl
setting a Coifman-Meyer type multiplier theorem was proved by Amri, Gasmi, and Sifi [3].
The main result of this paper is the following generalized Coifman-Meyer type theorem.
Theorem (Theorem 2.3). Let m : (0,∞)2 → C satisfy the Hörmander's condition
λα1+α2 ∂αm(λ) ≤ Cα,
λ ∈ (0,∞)2,
for sufficiently many multi-indices α ∈ N2. Then Bm given by (1.2) is bounded from Lp1(X) ×
Lp2(X) to Lp(X), where 1/p1 + 1/p2 = 1/p, with p1, p2, p > 1.
Theorem 2.3 is formally stated and proved in Section 2. The main difficulty in obtaining the
theorem lies in finding an appropriate proof of the classical Coifman-Meyer multiplier theorem,
which is prone to modifications towards our setting. The proof we present in Section 2 follows
the scheme by Muscalu and Schlag [18, pp. 67-71]. An important ingredient in our proof is a
spectrally defined Littlewood-Paley theory. For this method to work the assumption (PF) is
very useful. It might be interesting to try to replace (PF) with a less rigid condition.
An application of Theorem 2.3 provides Coifman-Meyer type multiplier results for bilinear
multipliers given by (1.2) in three cases different than the Fourier transform setting. In Theo-
rem 3.1 we treat bilinear multipliers for L being the discrete Laplacian on Zd. This is close to
[5, Theorem 3.4], however our results here are of a different kind. In Theorem 4.1 we consider
bi-radial bilinear Dunkl multipliers, here L is the general Dunkl Laplacian. In Corollary 4.2 we
also reprove [3, Theorem 4.1]. Finally, in Theorem 5.1 we give a Coifman-Meyer type multiplier
result for Jacobi trigonometric polynomials, here L is the Jacobi operator.
The second main goal of this paper is to obtain fractional Leibniz rules for operators different
from the Laplacian. The fractional Leibniz rule states that, if ∆Rd is the Laplacian on Rd,
then for each s ≥ 0 and 1/p = 1/p1 + 1/p2, p1, p2 > 1, p > 1/2, we have,
k(−∆Rd)s(f g)kp . k(−∆Rd)s(f )kp1kgkp2 + k(−∆Rd)s(g)kp2kfkp1.
The proof of this inequality can be found in Grafakos and Ou [11], see also Bourgain and Li [6]
for the endpoint case. The fractional Leibniz rule is also known as the Kato-Ponce inequality, as
Kato and Ponce studied a similar estimate [13] (see also [14]). Generalizations of Kato-Ponce or
similar inequalities were considered by many authors. For example Muscalu, Pipher, Tao, and
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
3
Thiele [17] extended this inequality by admitting partial fractional derivatives in R2, Bernicot,
Maldonado, Moen, and Naibo [4] proved the Kato-Ponce inequality in weighted Lebesgue
spaces, while Frey [10] obtained a fractional Leibniz rule for general operators satisfying Davies-
Gaffney estimates and p1 = p = 2, p2 = ∞
In the present paper we obtain fractional Leibniz rules of the form
kLs(f g)kp .s kLs(f )kp1kgkp2 + kLs(g)kp2kfkp1,
where s > 0 and 1/p1 +1/p2 = 1/p, with p1, p2, p > 1, in two other settings. In Corollary 3.2 we
prove a fractional Leibniz rule for L being the discrete Laplacian on Zd, while in Corollary 4.3
we justify a fractional Leibniz rule when L is the Dunkl Laplacian in the product setting. The
proofs of these fractional Leibniz rules rely on two properties of L. Firstly, we need appropriate
Coifman-Meyer type multiplier results; these are Theorems 3.1 and 4.1 and are deduced from
Theorem 2.3. Secondly, we require the existence of certain operators related to L that satisfy
(or almost satisfy) an integer order Leibniz rule. As we do not know such an operator in the
Jacobi setting we do not provide a fractional Leibniz rule there.
The article is organized as follows. In Section 2 we provide a general Coifman-Meyer type
multiplier result, see Theorem 2.3. This is then a basis to establish Coifman-Meyer type
multiplier results in various cases. In Section 3 we apply Theorem 2.3 for the discrete Laplacian
on Zd, see Theorem 3.1. As a consequence, in Corollary 3.2 we also obtain a fractional Leibniz
rule. Next, in Section 4 we deduce from Theorem 2.3 a Coifman-Meyer multiplier theorem for
general bi-radial Dunkl multipliers, see Theorem 4.1. From this result we obtain a fractional
Leibniz rule for the Dunkl Laplacian in the product case, see Corollary 4.3. Finally, in Section 5,
using Theorem 2.3 we prove a bilinear multiplier theorem for Jacobi trigonometric polynomial
expansions.
It is straightforward to extend the result of this paper to the multilinear setting. However,
to keep the presentation simple, we decided to limit ourselves to the bilinear case.
Troughout the paper we use the variable constant convention, where C, Cp, Cs, etc. may
denote different constants that may change even in the same chain of inequalities. We write
X . Y, whenever X ≤ CY, with C being independent of significant quantities. Similarly, by
X ≈ Y we mean that C −1Y ≤ X ≤ CY. By S(Rd) we denote the space of Schwartz functions.
The symbols Z and N denote the sets of integers and non-negative integers, respectively. For a
multi-index α ∈ N2 by α we denote its length α1 + α2. Throughout the paper, for a function
ψ : [0,∞) → C we set
ψk(λ) = ψ(2−kλ),
λ ∈ [0,∞).
2. General bilinear multipliers
We say that a function µ : (0,∞) → C satisfies the (one-dimensional) Mikhlin-Hörmander
condition of order ρ ∈ N if it is differentiable up to order ρ and
(2.1)
sup
kµkM H(ρ) := sup
j≤ρ
λ∈(0,∞)λj
dj
dλj µ(λ) < ∞.
Similarly, we say that m : (0,∞)2 → C satisfies the (two-dimensional) Mikhlin-Hörmander
condition of order s ∈ N, if the partial derivatives ∂αm exist for multi-indices α ≤ s and
(2.2)
kmkM H(s) := sup
α≤s
sup
λ∈(0,∞)2 λα∂αm(λ1, λ2) < ∞.
4
BŁAŻEJ WRÓBEL
Consider a non-negative self-adjoint operator L on L2(X, ν) with domain Dom(L). Here
(X, ν) is a σ-finite measure space with ν being a Borel measure. Throughout the paper we
assume that L generates a symmetric contraction semigroup, namely
ke−tLfkLp(X,ν) ≤ kfkLp(X,ν),
f ∈ Lp(X, ν) ∩ L2(X, ν),
(CT)
and that L is injective on Dom(L). Then, for µ : (0,∞) → C, the spectral theorem allows us
to define the multiplier operator µ(L) =R(0,∞) µ(λ)dE(λ) on the domain
Dom(µ(L)) =(cid:26)f ∈ L2(X, ν) : Z(0,∞) µ(λ)2 dEf,f (λ) < ∞(cid:27).
Here E is the spectral measure of L, while Ef,f is the complex measure defined by Ef,f (·) =
hE(·)f, fiL2(X,ν).
We shall need the following assumption on L;
(MH)
L has a Mikhlin-Hörmander functional calculus of a finite order ρ > 0. More pre-
cisely, every function µ that satisfies (2.1) gives rise to an operator µ(L) which is
bounded on all Lp(X, ν), 1 < p < ∞, and
kµ(L)kLp(X,ν)→Lp(X,ν) ≤ CpkµkM H(ρ).
Note that if L = (−∆R)1/2 then (MH) follows from the Mikhlin-Hörmander multiplier theorem.
There are two consequence of (MH) which will be needed later. The first of them is well
known and follows from Khintchine's inequality.
Proposition 2.1. Let ψ : [0,∞) → C be a function supported in [ε, ε−1], for some ε > 0, and
assume that ψ ∈ C ρ([0,∞)). Then the square function
is bounded on Lp(X, ν), p > 1, and
f 7→ Sψ(f ) =(cid:0)Xk∈Z
ψk(L)f2(cid:1)1/2
(2.3)
kSψ(f )kLp(X,ν) ≤ Cε kψkC ρ([0,∞))kfkLp(X,ν).
The second of the required consequences is proved in [23, Corollary 3.2].
Proposition 2.2. Let ϕ : [0,∞) → C be compactly supported, and assume that ϕ ∈ C α([0,∞))
for some α > ρ + 2. Then the maximal operator
is bounded on Lp(X, ν), p > 1, and
f 7→ Mϕ(f ) = sup
k∈Z ϕk(L)f
(2.4)
kMϕ(f )kLp(X,ν) ≤ kϕkC ρ+2([0,∞))kfkLp(X,ν).
To simplify the proof of our main Theorem 2.3 we will need an auxiliary subspace of L2(X, ν).
Namely, consider the spaces
(2.5) A2 = {g ∈ L2(X, ν) : g = E(ε,ε−1)g, for some ε > 0} and A = A2 ∩ \1<p<∞
Then, (MH) implies that A is dense in Lp(X, ν) for 1 < p < ∞.
For the convenience of the reader we shall justify this statement. Let ψ : [0,∞) → C be a
smooth function which is supported in [1/2, 2] and such that Pk∈Z ψk(λ) = 1, λ > 0. Then,
for each N ∈ N and f ∈ L1(X, ν) ∩ L∞(X, ν) the partial sum SN f =PN
k=−N ψk(L)f belongs
Lp(X, ν).
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
5
to A by (MH). We claim that SN f → f in Lp(X, ν). To see this we take 1 < r < ∞ if p ≤ 2
or r > p if p > 2. Then we observe that kSN fkLr(X,ν) is uniformly bounded in N (this follows
from (MH)) and that SN f → f in L2(X, ν) (this follows from the spectral theorem, since
E{0} = 0 by the injectivity of L). Therefore, the log-convexity of Lp norms proves the claim.
Finally, a density argument together with the fact that kSN fkLp(X,ν) is uniformly bounded in
N shows that A is dense in Lp(X, ν) and finishes our task.
Besides being dense in Lp(X, ν) the space A has the nice property that each f ∈ A satisfies
f = PN (f )
k=−N (f ) ψk(L)f, where N (f ) is a fixed integer depending on f and ψ is the function
from the previous paragraph. This allows us to deal easily with some rather delicate questions
on convergence in the proof of Theorem 2.3.
We proceed to define formally the bilinear multipliers studied in this paper. To do this we
will need the operators L1 = L ⊗ I and L2 = I ⊗ L. These may be regarded as non-negative
self-adjoint operators on L2(X × X, ν ⊗ ν), see [21, Theorem 7.23] and [25, Proposition A.2.2].
Moreover, the spectral measure of L1 is EL ⊗ I, while the spectral measure of L2 is I ⊗ EL.
Thus, the operators L1 and L2 commute strongly and the bivariate spectral theorem, see e.g.
[21, Theorem 5.21], allows us to consider multiplier operators
m(L1, L2) =Z(0,∞)2
m(λ) dE⊗(λ)
on the domain
Dom(m(L1, L2)) =(cid:26)F ∈ L2(X × X, ν ⊗ ν) : Z(0,∞) m(λ)2 dE⊗
F,F (λ) < ∞(cid:27).
Here m : [0,∞)2 → C is a Borel measurable function, E⊗ = EL ⊗ EL is the joint spectral mea-
sure of (L1, L2), while E⊗
In the most general form the bilinear multiplier operators studied in the paper are given by
F,F is the complex measure defined by (E⊗)F,F (·) = hE⊗(·)F, FiL2(X×X,ν⊗ν).
(2.6)
Bm(f1, f2)(x) = m(L1, L2)(f1 ⊗ f2)(x, x) = ZZ(0,∞)2
m(λ) d(Ef1 ⊗ Ef2)! (x, x),
where L1 = L ⊗ I and L2 = I ⊗ L. Since the diagonal {(x, x) : x ∈ X} may be of measure 0 in
(X × X, ν ⊗ ν) the equation (2.6) is not formal. In order to make it rigorous we assume that:
(WD)
if f1, f2 ∈ A and m : (0,∞)2 → C is bounded, then
• m(L1, L2)(f1 ⊗ f2) has a continuous representative on X × X
• km(L1, L2)(f1 ⊗ f2)kL∞((X×X),ν⊗ν) ≤ Cf1,f2kmkL∞((0,∞)2).
Thus, restricting m(L1, L2)(f1 ⊗ f2)(x1, x2) to the diagonal, we have a formal definition of
Bm(f1, f2), for f1, f2 ∈ A. For instance, if L = (−∆R)1/2, then the operator Bm is closely
related to the bilinear multiplier for the Fourier transform, namely,
Bm(f1, f2)(x) =ZZR2
m(ξ1,ξ2) f (ξ1) f (ξ2)eix(ξ1+ξ2) dξ.
If m is bounded and f1, f2 ∈ A then m(ξ1, ξ2) f (ξ1) f (ξ2) ∈ L1(Rd), and thus Bm(f1, f2)(x) is
well defined (in fact continuous) by the Lebesgue dominated convergence theorem.
6
BŁAŻEJ WRÓBEL
We need one more assumption to prove the main theorem. Namely, we require that:
(PF)
there is b > 0 with the following property: if ϕ and ψ are bounded smooth functions
such that supp ϕk ⊆ [0, 2k−b] and supp ψk ⊆ [2k−2, 2k+2], k ∈ Z, then
ϕk(L)(f1) · ψk(L)(f2) = ψk(L)[ϕk(L)(f1) · ψk(L)(f2)],
for f1, f2 ∈ A,
where ψk is a smooth function which is bounded by 1, equals 1 on [2k−3−b, 2k+3+b]
and vanishes outside [2k−5−b, 2k+5+b].
We remark that, since f1, f2 ∈ A, the function g = ϕk(L)(f1)·ψk(L)(f2) belongs to L2(X, ν), so
that an application of ψk(L) to g is legitimate. Note that when L = (−∆R)1/2 the formula (PF)
can be easily deduced by using the convolution structure on the frequency space associated
with Fourier multipliers.
In what follows we often abbreviate Lp := Lp(X, ν) and k · kp := k · kLp. Let p, p1, p2 > 1.
We say that a bilinear operator B is bounded from Lp1 × Lp2 to Lp if
f1, f2 ∈ A.
kB(f1, f2)kp ≤ Ckf1kp1kf2kp2,
Note that in this case B has a unique bounded extension from Lp1 × Lp2 to Lp.
The main result of this paper is a Coifman-Meyer type general bilinear multiplier theorem.
Theorem 2.3. Let L be a non-negative self-adjoint operator on L2(X, ν), which is injective on
its domain and satisfies (CT), (MH), (WD), and (PF). Assume that m : (0,∞)2 → C satisfies
the Mikhlin-Hörmander condition (2.2) of an order s > 2ρ + 4. Then the bilinear multiplier
operator Bm, given by (2.6), is bounded from Lp1 × Lp2 to Lp, where 1/p1 + 1/p2 = 1/p, and
p1, p2, p > 1. Moreover, for such p, p1, p2, there is C = C(p1, p2, p, s) such that
(2.7)
kBm(f1, f2)kp ≤ C kmkM H(s) kf1kp1kf2kp2.
Proof. Let ψ be a smooth function supported in [1/2, 2] and such that Pk ψk ≡ 1. We set
F = f1 ⊗ f2 : X × X → C and split
Bm(f1, f2)(x) = Xk1,k2∈Z
= Xk1−k2≤b+2
. . . + Xk1>k2+b+2
. . . + Xk2>k1+b+2
[ψk1(L1)ψk2(L2)m(L1, L2)](F )(x, x)
. . . := T1 + T2 + T3.
There is no issue of convergence here as for f1, f2 ∈ A each of the sums defining T1, T2, and T3
is finite.
We estimate separately each of the operators Ti, i = 1, 2, 3, starting with T1. This is the
easiest part, in fact here the assumption (PF) is redundant.
For k ∈ Z set
mk(λ1, λ2) = ψk(λ1) Xk2 : k−k2≤b+2
ψk2(λ2)m(λ) = ψk(λ1)φk(λ2)m(λ),
with φ(λ2) =Pj≤b+2 ψj(λ2), so that supp φ ⊆ [2−b−3, 2b+3], and
supp ψ ⊗ φ ⊆ [2−1, 21] × [2−b−3, 2b+3].
Let ψ be another smooth function, which vanishes outside [2−b−4, 2b+4] and equals 1 on
[2−b−3, 2b+3]. Then
mk(λ1, λ2) = [ ψk(λ1) ψk(λ2)]ψk(λ1)φk(λ2)m(λ),
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
7
Moreover, supp mk ⊆ [2k−b−4, 2k+b+4]2, and, consequently, Mk(λ) := mk(2kλ) is supported
in [−2b+4, 2b+4]2 := [−a, a]2. Thus, Mk can be expanded into a double Fourier series inside
[−a, a]2, i.e.,
cn,keπin1λ1/aeπin2λ2/a,
λ ∈ [−a, a]2,
with the Fourier coefficients
1
Mk(λ) = Xn1,n2∈Z
cn,k =
4a2 ZZ[−a,a]2
[ψ ⊗ φ]m(2kξ) eπin1ξ1/aeπin2ξ2/a dλ.
Now, using integration by parts, together with the assumption (2.2), and the fact that ψ ⊗ φ
is compactly supported away from 0, we we obtain the uniform in k ∈ Z bound
(2.8)
cn,k ≤ C kmkM H(s) (1 + n)−s,
n ∈ Z2.
We remark that here, in order to conclude (2.8), it is perfectly enough to assume the Marcinkiewicz
'product' condition
instead of (2.2).
Dγm(λ) ≤ Cλ1γ1λ2γ2,
Coming back to mk we now write, for λ ∈ [2k−b−4, 2k+b+4]2,
ψk(λ1)φk(λ2)m(λ) = Xn∈Z2
cn,k e2πin12−kλ1/ae2πin22−kλ2/a.
Thus, mk can be expressed as
cn,k[ ψk(λ1)e(2π/a)in12−kλ1][ ψk(λ2)e(2π/a)in22−kλ2]
mk(λ1, λ2) = Xn∈Z2
:= Xn∈Z2
cn,kψn1
k (λ1)ψn2
k (λ2).
By (2.8) and the bivariate spectral theorem we have that
cn,k [ ψk(L1)e(2π/a)in12−kL1(f1)](x1)[ ψk(L2)e(2π/a)in22−kL2](f2)(x2),
mk(L1, L2)(F )(x1, x2) = Xn∈Z2
for a.e. x1, x2 ∈ X; here we have convergence in L2(X × X, ν ⊗ ν). Moreover, (2.8) and the
assumption (WD) imply that the above sum converges also pointwise (and gives a continuous
function on X × X).
Consequently, for x ∈ X we have
T1(f1, f2)(x) =Xk∈Z
mk(L1, L2)(F )(x, x) = Xn∈Z2 Xk∈Z
k (L)(f1)(x) · ψn2
k (L)(f2)(x),
cn,k ψn1
where we have used the fact that the sum in k is finite when f1, f2 ∈ A. Now Schwarz's in-
equality (first inequality below), and Hölder's inequality together with (2.8) (second inequality
below), lead to the estimate
(2.9)
kT1(f1, f2)kp ≤ Xn∈Z2
. kmkM H(s) Xn∈Z2
sup
k∈Z cn,k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xk∈Z
(1 + n)−s(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xk∈Z
ψn2
k (L)(f2)2(cid:1)1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
k (L)(f1)2(cid:1)1/2(cid:0)Xk∈Z
ψn1
k (L)(f1)2(cid:1)1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xk∈Z
k (L)(f2)2(cid:1)1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p2
ψn1
ψn2
.
8
BŁAŻEJ WRÓBEL
Thus, taking into account the presence of the modulations e2πinj2−kλj/a in the definition of
ψnj
k , j = 1, 2, and using Proposition 2.1 we obtain
(cid:18)Xk∈Z
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ψnj
k (L)(fj)2(cid:19)1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)pj
. (1 + nj)ρ kfjkpj .
However, since we have the rapidly decaying factor in (2.9), if s > 2ρ + 4, we arrive at the
desired bound
Now we pass to estimating T2 and T3. Since the proofs are mutatis mutandis the same, we
kT1(f1, f2)kp . kmkM H(s) kf1kp1kf2kp2.
treat only the former operator. Setting ϕ =Pj<−b−2 ψj we rewrite T1 as
[ψk1(L1)ψk2(L2)m(L1, L2)](F )(x, x)
T2(f1, f2)(x) = Xk1>k2+b+2
[ψk1(L1)(cid:18) Xk2<k1−b−2
[ψk(L1)ϕk(L2)m(L1, L2)](f1 ⊗ f2)(x, x),
ψ(2−k2 L2)(cid:19)m(L1, L2)](f1 ⊗ f2)(x, x)
=Xk1
=Xk
where ϕ(λ2) = Pk2<−b−2 ψk2(λ2). Then clearly supp ϕ ⊆ [0, 2−b−1]. Recall that in the above
decomposition of T2 all the appearing sums in k, k1, and k2, are in fact finite since f1, f2 ∈ A.
Set mk := ψkϕkm and note that mk is supported in [2k−1, 2k+1]× [0, 2k−b−1], this is because
supp ψ ⊗ ϕ ⊆ [2−1, 21] × [0, 2−b−1].
Similarly to the case of T1 we expand the function Mk = mk(2kλ) in a Fourier series. Namely,
let ψ be a smooth function vanishing outside [2−2, 22] and equal to 1 on [2−1, 21], and let ϕ be
a smooth function vanishing outside [0, 2−b] and equal to 1 on [0, 2−b−1]. Then
mk(λ1, λ2) = [ ψk(λ1) ϕk(λ2)]ψk(λ1)ϕk(λ2)m(λ),
Moreover, supp mk ⊆ [2k−1, 2k+1] × [0, 2k−b−1], and, consequently, Mk(λ) = mk(2kλ) is sup-
ported in [−2, 2]2. Hence, Mk can be expanded into a double Fourier series inside [−2, 2]2, i.e.,
for λ ∈ [−2, 2]2,
π
π
cn,ke
2 in1λ1e
2 in2λ2,
with the Fourier coefficients
cn,k =
[ψ ⊗ ϕ]m(2kξ) e
π
2 in1ξ1e
π
2 in2ξ2 dξ.
Mk(λ) = Xn1,n2∈Z
16ZZ[−2,2]2
1
As with T1, we now use integration by parts, together with the assumption (2.2). Here it is
important that we assume the stronger Mikhlin-Hörmander condition instead of merely the
Mikhlin-Marcinkiewicz condition. Indeed, from integration by parts we obtain, for arbitrary β
cn,k = O((1 + n)−β)ZZ[−2,2]2
dβ
dξβ ([ψ ⊗ ϕ]m(2kξ)) e
π
2 in1ξ1e
π
2 in2ξ2 dξ.
However, as ψ ⊗ ϕ does not vanish for λ2 close to zero, in order to conclude that the above
integral is uniformly bounded we do need (2.2). In summary we proved that (2.8) holds also
in this case.
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
9
Coming back to mk we now write, for λ ∈ [2k−2, 2k+2] × [0, 2k−b]
ψk(λ1)ϕk(λ2)m(λ) = Xn∈Z2
cn,k e
π
2 in12−kλ1e
π
2 in22−kλ2.
Thus, mk, k ∈ Z, can be expressed as
mk(λ1, λ2) =Xn
:= Xn∈Z
cn,kψn1
k (λ1)ϕn2
k (λ2).
cn,k[ ψk(λ1)e
π
2 in12−kλ1][ ϕk(λ2)e
π
2 in22−kλ2]
With the aid of (WD) and (2.8), arguing as on p. 7 we see that
mk(L1, L2)(F )(x, x) = Xn∈Z2
cn,k ψn1
k (L)(f1)(x) · ϕn2
k (L)(f2)(x),
where the series on the right converges pointwise to a continuous function on X.
Summarizing the above, we have just decomposed
T2(f1, f2)(x) =Xk∈Z
mk(L1, L2)(F )(x, x) = Xn∈Z2 Xk∈Z
cn,k ψn1
k (L)(f1)(x) · ϕn2
k (L)(f2)(x).
Now, let ψ be a real-valued smooth function equal to 1 on [2−3−b, 23+b] and vanishing outside
[2−5−b, 25+b]. Since, for each n = (n1, n2) ∈ Z2, the function ϕn2
is supported in [0, 2k−b], and
the function ψn1
k
is supported in [2k−2, 2k+2], using the assumption (PF) we have
k
cn,k ψk(L)[ψn1
k (L)(f1) · ϕn2
k (L)(f2)](x).
Hence, if h is a function in Lq, 1/p + 1/q = 1, then we obtain
T2(f1, f2)(x) = Xn∈Z2 Xk∈Z
T2(f1, f2)(x)h(x) dν(x) =ZX Xn∈Z2 Xk∈Z
ZX
and, consequently,
cn,k ψn1
k (L)(f1)(x)· ϕn2
k (L)(f2)(x) ψk(L)(h)(x) dν(x),
(2.10)
sup
k∈Z
ZX
T2(f1, f2)(x)h(x) dν ≤ Xn∈Z2
k (L)(f1)2(cid:19)1/2
×ZX(cid:20)(cid:18)Xk∈Z
ψn1
(1 + n)−s(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xk∈Z
. kmkM H(s) Xn∈Z2
k (L)(f1)2(cid:19)1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p1
k (L)(f2)(cid:13)(cid:13)(cid:13)(cid:13)p2
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18)Xk
(cid:13)(cid:13)(cid:13)(cid:13)
k ϕn2
ψn1
sup
where we used Proposition 2.1 with ψ in the second inequality above. Similarly to the estimate
for T1, applying Propositions 2.1 and 2.2 leads to
ϕn2
k∈Z cn,k×
sup
k (L)(f2)(cid:21)(cid:18)Xk∈Z
k (L)(f1)2(cid:1)1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p1(cid:13)(cid:13)(cid:13)(cid:13)
ψn1
dν
ψk(L)(h)2(cid:19)1/2
k∈Z ϕn2
sup
,
k (L)(f2)(cid:13)(cid:13)(cid:13)(cid:13)p2
. (1 + n1)ρkf1kp1
(cf. (2.3))
. (1 + n2)ρ+2kf2kp2
(cf. (2.4)).
10
BŁAŻEJ WRÓBEL
Finally, the rapidly decaying factor in (2.10) gives, for s > 2ρ + 4, the desired bound
The proof of Theorem 2.3 is thus completed.
kT2(f1, f2)kp . kmkM H(s) kf1kp1kf2kp2.
(cid:3)
3. Bilinear multipliers on Zd
In the present section we formalize Theorem 2.3 for bilinear multiplier operators on Zd. We
also prove a fractional Leibniz rule for the discrete Laplacian.
Let ej = (0, . . . , 1, . . . , 0) ∈ Zd be the j-th coordinate vector. Consider the discrete Laplacian
on Zd, given by
∆Zd(f )(n) = 2d f (n) −
d
d
Xj=1
(f (n + ej) + f (n − ej)) = 2dIf (n) −
Xj=1
(f ∗ δej + f ∗ δ−ej ).
The multilinear operators (2.6) for the discrete Laplacian are defined via Fourier analysis on
Zd. Namely, let Td ≡ (−1/2, 1/2]d be the d-dimensional torus, let
ξ ∈ Td
f (k)e2πin·ξ,
FZd(f )(ξ) = Xn∈Zd
be the Fourier transform on Zd, and define
Sin2(ξ) = 4
d
Xj=1
sin2(πξj),
ξ ∈ Td.
Then, since
the formula (2.6) takes the form
FZd(∆Zd(f ))(ξ) = Sin2(ξ)FZd (f )(ξ),
ξ ∈ Td,
Bm(f1, f2)(n) := m((−∆Zd)1/2 ⊗ I, I ⊗ (−∆Zd)1/2)(f1 ⊗ f2)(n, n)
(3.1)
=ZTdZTd
m( Sin(ξ1), Sin(ξ2))FZd (f1)(ξ1)FZd(f2)(ξ2)e−2πin(ξ1+ξ2) dξ,
where n ∈ Zd. Note that the space A2 from (2.5) in this case is given by
A2 = {g ∈ L2(Td) : FZd(g)(ξ) = 0 for some ε > 0 and all ξ < ε.}
Throughout this section we denote by Lp the space lp(Zd) equipped with the counting
measure. Using Theorem 2.3 we prove the following Coifman-Meyer multiplier theorem for the
discrete Laplacian.
Theorem 3.1. Assume that m satisfies Hörmander's condition (2.2) of order s > d + 4.
Then the bilinear multiplier operator given by (3.1) is bounded from Lp1 × Lp2 to Lp, where
1/p1 + 1/p2 = 1/p, and p1, p2, p > 1. Moreover, the bound (2.7) holds.
Proof. It is well known that L = (−∆Zd)1/2 is injective on L2 and satisfies (CT). Moreover, it
also satisfies (WD) since for f1, f2 ∈ A we have FZd(f1)(ξ1)FZd(f2)(ξ2) ∈ L1(Td × Td). From
[1, Theorem 1.1] it follows that −∆Zd has a Mikhlin-Hörmander functional calculus (of order
[d/2] + 1). Then, clearly, the same is true for (−∆Zd)1/2. Hence, (MH) has been justified.
To apply Theorem 2.3 it remains to show that L = (−∆Zd)1/2 satisfies (PF). We prove
2 log2 d. Since the spectrum of (−∆Zd)1/2 is contained in [0, 2√d], we have
it with b = 7 + 1
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
11
2 log2 d. Hence, it suffices to show (PF) for k ≤ 2 + 1
ψk((−∆Zd)1/2) ≡ 0, if k > 2 + 1
2 log2 d.
Using elementary Fourier analysis on Zd we see that to prove (PF) it is enough to show that
ψk ◦ Sin = 1
((ψk ◦ Sin)FZd(f1)) ∗Td ((ϕk ◦ Sin))FZd (f2)),
where ψk, ψk, and ϕk are the functions from (PF). In other words that we are left with proving
that if Sin(ξ) < 2k−3−b or Sin(ξ) > 2k+3+b, then
(3.2)
ψk( Sin(ξ − η))FZd (f1)(ξ − η) · ϕk( Sin(η))FZd (f2)(η) dη = 0.
on the support of
ZTd
The formula
(3.3)
leads to sin π(ξj) ≤ sin π(ξj − ηj) + sin πηj, j = 1, . . . , d, and, consequently,
sin π(t − s) = sin πt cos πs − sin πs cos πt,
√d( Sin(ξ − η) + Sin(η)),
Sin(ξ) ≤
η ∈ Td.
s, t ∈ T,
From the above it follows that if Sin(ξ) > 2k+3+b, then for every η ∈ Td the integrand in
(3.2) vanishes.
It remains to show that also Sin(ξ) < 2k−3−b forces (3.2). We argue by contradiction
assuming that Sin(ξ) < 2k−3−b yet the integral in (3.2) is non-zero. Then, for some η ∈ Td,
we must have ψk( Sin(ξ − η)) ϕk( Sin(η)) 6= 0, which implies that
(3.4)
Note that since k ≤ 2 + 1
2 log2 d, the integral in (3.2) runs over Sin(η) ≤ 2k−b ≤ 2−1,
and, consequently, we consider only those η satisfying cos πηj > √3/2 > 1/2, for every
j = 1, . . . , d. Now, using (3.3) (with t − s = ξj, s = −ηj) we obtain
2k−1 ≤ Sin(ξ − η) ≤ 2k+1
Sin(η) ≤ 2k−b.
and
sin πξj ≥ sin π(ξj − ηj) cos πηj − cos π(ξj − ηj) sin πηj
≥
1
2 sin π(ξj − ηj) − sin πηj.
Summing the above estimate in j and using Schwarz inequality we arrive at
d
d
1
2
Xj=1
sin πξj ≥
√d Sin(ξ) ≥
Xj=1
Now, since Sin(ξ) < 2k−3−b, using (3.4) we arrive at
1
√d
sin π(ξj − ηj) −
Xj=1
2k−b−3 > Sin(ξ) >
which is a contradiction.
√d2k−b(cid:1) =
1
√d(cid:0)2k−1 −
d
sin πηj ≥
1
2 Sin(ξ − η) −
√d Sin(η).
(2k−1 − 2k−7) >
1
√d
2k−2 = 2k−b+5,
(cid:3)
As a corollary of Theorem 3.1 we prove a fractional Leibniz rule for the discrete Laplacian
on Zd. For Re(z) ≥ 0 and h ∈ L2 the complex derivative (−∆Zd)zh is given by
FZd((−∆Zd)zh)(ξ) = Sin ξ2zFZd(h)(ξ),
ξ ∈ Td.
This coincides with taking the n-th composition of (−∆Zd) when z = n is a non-negative
integer. Clearly, (−∆Zd)z is bounded on L2. Moreover, when z = s ∈ R, s ≥ 0, then (−∆Zd)s
is also bounded on all Lp, 1 ≤ p ≤ ∞. To see this we just use the Taylor series expansion
of the function xs = (1 − (1 − x))s, with x replaced by (−∆Zd)/(4d). This is legitimate since
(−∆Zd)/(4d) is a contraction on all Lp spaces. Our fractional Leibniz rule is the following.
12
BŁAŻEJ WRÓBEL
k(−∆Zd)s(f g)kp . k(−∆Zd)sfkp1kgkp2 + k(−∆Zd)sgkp2kfkp1,
Corollary 3.2. Let 1/p = 1/p1 + 1/p2, with p, p1, p2 > 1. Then, for every s > 0,
(3.5)
where f, g ∈ A.
Remark 1. Note that if f, g ∈ A then f g ∈ L2, hence (−∆Zd)s(f g) makes sense.
Remark 2. Since (−∆Zd)s is bounded on all Lp spaces, 1 ≤ p ≤ ∞, a version of (3.5) without
the Laplacians on the right hand side is obvious. This is in contrast with the fractional Leibniz
rule on Rd.
In the proof of the corollary we shall need two lemmata. The first of them follows from the
lp(Z) boundedness of the discrete Hilbert transform.
Lemma 3.3. The one-dimensional linear multiplier operator
H(f )(n) =Z 1/2
is bounded on all lp(Z) spaces, 1 < p < ∞.
0 FZ(f )(x)e2πiξn dξ,
The second of the lemmata is the following.
n ∈ Z
Lemma 3.4. Let d = 1. Assume that ϕ : (0,∞)2 → C is a bounded function that satisfies the
Mikhlin-Hörmander condition (MH) of order 6. Then, for Re(z) ≥ 0 we have
(3.6)
(−∆Z)z(Bϕ(f, g))(n)
=ZZT2
ϕ(2 sin πξ1, 2 sin πξ2)2 sin π(ξ1 + ξ2)2ze2πi(ξ1+ξ2)n FZ(f )(ξ1)FZ(g)(ξ2) dξ,
where f, g ∈ A, and n ∈ Z.
Proof. From Theorem 3.1 and the assumptions on ϕ it follows that Bϕ(f, g) ∈ ℓ2(Z). Thus,
the left hand side of (3.6) makes sense as a function on ℓ2(Z). Moreover, a continuity argument
shows that it suffices to demonstrate (3.6) for Re(z) > 0.
Thus, for P being a polynomial we obtain
ϕ(ξ1, ξ2) (4 sin2 π(ξ1 + ξ2))ke2πi(ξ1+ξ2)n FZ(f )(ξ1)FZ(g)(ξ2) dξ.
Set ϕ(ξ1, ξ2) = ϕ(2 sin πξ1, 2 sin πξ2). Since −∆Z(e2πit·)(n) = 4(sin2 πt)e2πitn, for t ∈ T
and n ∈ Z, we deduce that (−∆Z)k(e2πit·)(n) = 22k sin πt2ke2πitn, k ∈ N. Hence, for k, n ∈ N,
we have
(−∆Z)k(Bϕ(f, g))(n) =ZZT2
P (−∆Z)(Bϕ(f, g))(n) =ZZT2
where n ∈ Z.
tions in place of polynomials. In particular, taking λ 7→ λz, Re(z) > 0, we obtain (3.6).
Finally, a density argument shows that the above formula remains true for continuous func-
(cid:3)
ϕ(ξ1, ξ2) P (4 sin2 π(ξ1 + ξ2))e2πi(ξ1+ξ2)n FZ(f )(ξ1)FZ(g)(ξ2) dξ,
We proceed to the proof of the corollary.
Proof of Corollary 3.2. We claim that it is enough to prove the corollary in dimension d = 1.
Indeed, fix s > 0 and assume that (3.5) is true in this case. Let ∆Z be the one dimensional
discrete Laplacian on Z. Define Lj := −∆Z ⊗ I(j), j = 1, . . . , d, to be the one-dimensional
discrete Laplacian acting on the j-th variable, so that, clearly, −∆Zd = Pd
j=1 Lj. Since each
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
13
Lj generates a symmetric contraction semigroup, using e.g. the multivariate multiplier theorem
[24, Corollary 3.2] we see that the operator
is bounded on Lp, p > 1. In other words, we have the bound
(X Lj)s(X Ls
j)−1
k(−∆Zd)s(f g)kp . k
d
Xj=1
Ls
j(f g)kp ≤
d
Xj=1
kLs
j(f g)kp.
Since the multiplier Ls
Corollary 3.2]) in order to conclude the proof of our claim it is thus enough to show that
j(−∆Zd)−s is bounded on all Lp, p > 1, (this again follows from [24,
(3.7)
for every j = 1, . . . , d.
kLs
j(f g)kp . kLs
jfkp1kgkp2 + kLs
jgkp2kfkp1,
For notational simplicity we justify (3.7) only for j = 1, the proofs for other j are analogous.
For a sequence h : Zd → C denote hn(k) := h(k, n), k ∈ Z, n ∈ Zd−1. Clearly, we have
(f g)n(·) = fn(·)gn(·). Then, using (3.5) in the dimension d = 1 (first inequality below), together
with the simple fact that (a + b)p ≈ ap + bp (second and last inequalities below), and Hölder's
inequality with exponents p1/p, p2/p > 1 (third inequality below) we obtain
1(f g)kp = Xn∈Zd−1 kLs
kLs
. Xn∈Zd−1(cid:0)kLs
. Xn∈Zd−1 kLs
.(cid:0) Xn∈Zd−1 kLs
+(cid:0) Xn∈Zd−1 kLs
.(cid:0)kLs
1(fn(·)gn(·))kp
1(gn)klp1 (Z)kfnklp2 (Z)(cid:1)p
1(gn)kp
lp1 (Z)kgnkp
lp1 (Z)(cid:1)p/p1(cid:0) Xn∈Zd−1 kgnkp2
lp2 (Z)(cid:1)p/p2(cid:0) Xn∈Zd−1 kfnkp1
lp1 (Z)kfnkp
lp2 (Z)(cid:1)p/p2
lp1 (Z)(cid:1)p/p1 = kLs
1((f g)n(·))kp
lp(Z) = Xn∈Zd−1 kLs
1(fn)klp1 (Z)kgnklp2 (Z) + kLs
1(fn)kp
1(fn)kp1
1(gn)kp2
lp2 (Z) + kLs
1(g)kp2kfkp1(cid:1)p.
1(f )kp1kgkp2 + kLs
1(f )kp
p1kgkp
p2 + kLs
Hence, (3.7) is proved.
lp2 (Z)
lp(Z)
1(g)kp
p2kfkp
p1
Having justified the claim we now focus on proving (3.5) for d = 1. Till the end of the proof
of the corollary we work on Z and write lp and k · kp for lp(Z) and k · klp(Z), respectively.
Let η0 and η1 be smooth functions satisfying supp η0 ⊆ [0, 1/4], supp η1 ⊆ [1/8, 10] and η0 +
η1 = 1 on [0, 4]. For a function h ∈ A we set h0 = η0((−∆Z)1/2)(h) and h1 = η1((−∆Z)1/2)(h),
so that h = h0 + h1. From [1, Theorem 1.1] it follows that, for each fixed s > 0 the mul-
tiplier (−∆Z)−sη1(−∆Z) is bounded on all lp, 1 < p < ∞. Moreover, h0, h1 ∈ A. Since
h1 = (−∆Z)−sη1(−∆Z)[(−∆Z)s(h)], we thus have the estimate
kh1kp . k(−∆Z)shkp.
Hence, using the boundedness of (−∆Z)s and Hölder's inequality we obtain
k(−∆Z)s(fi1gi2)kp . kfi1kp1kgi2kp2 . k(−∆Z)sfkp1kgkp2 + k(−∆Z)sgkp1kfkp2,
14
BŁAŻEJ WRÓBEL
for i1, i2 ∈ {0, 1} not both equal to 0. In summary, to finish the proof it is enough to demon-
strate that
k(−∆Z)s(f0g0)kp . k(−∆Z)sfkp1kgkp2 + k(−∆Z)sgkp1kfkp2.
Clearly, FZ(f0)(x) = η0( sin πx)FZ(f )(x) and FZ(g)(y) = η0( sin πy)FZ(g0)(y). Hence,
denoting I = [0, 1/2) and using Lemma 3.4 together with (3.3) we now write
(−∆Z)s(f0g0)(n)
= 22sZTZT
= 22s Xǫ∈{−1,1}2Zǫ1IZǫ2I sin πξ1q1 − sin2 πξ2 + sin πξ2q1 − sin2 πξ12sη0( sin πξ1)η0( sin πξ2)
[ sin πξ1 cos πξ2 + sin πξ1 cos πξ22sη0( sin πξ1)η0( sin πξ2)]
× e2πi(ξ1+ξ2)n FZ(f )(ξ1)FZ(g)(ξ2) dξ
× e2πi(ξ1+ξ2)n FZ(f )(ξ1)FZ(g)(ξ2) dξ := Xǫ∈{−1,1}2
Tǫ(f, g)(n),
n ∈ Z.
Thus, in order to finish the proof it is enough to show that, for ǫ ∈ {−1, 1}2 it holds
(3.8)
kTǫ(f, g)kp . k(−∆Z)sfkp1kgkp2 + k(−∆Z)sgkp2kfkp1.
It is enough to justify (3.8) only for T1,1 and T1,−1 as the proofs for T−1,1 and T−1,−1 are
symmetric. In what follows we let φ be a function in C ∞([0,∞)) supported in [0, 1/4] and
such that φ(t) + φ(t−1) = 1. Note that then φ(λ2/λ1) satisfies Hörmander's condition (2.2) of
arbitrary order.
Let (η⊗
0 )(λ) = η0(λ1)η0(λ2), λ ∈ [0,∞)2. To justify (3.8) for T1,1 we set
φ(λ2/λ1)(η⊗
2/4)1/2 + λ2(1 − λ2
1/4)1/22s
0 )(λ),
1,1(λ) = λ1(1 − λ2
ms
1,1(λ) = λ1(1 − λ2
ms
2/4)1/2 + λ2(1 − λ2
1/4)1/22s
φ(λ1/λ2)(η⊗
0 )(λ).
λ2s
1
λ2s
2
λ2z
1
λ2z
2
Then, using (3.1) (in the case d = 1) we rewrite T1,1 as
T1,1(f, g) = Bms
1,1(H(−∆Z)sf, Hg) + B ms
1,1(Hf, H(−∆Z)sg).
In view of Lemma 3.3, to demonstrate (3.8) it suffices to show that
kBms
1,1 (f, g)kp + kB ms
1,1 (f, g)kp ≤ Ckfkp1kgkp2 .
This, however, follows directly from Theorem 3.1, since, for each s > 0, the multipliers ms
and ms
1,1, satisfy Hörmander's condition (2.2) of arbitrary order.
1,1,
Finally, we prove (3.8) for T1,−1. For Re(z) ≥ 0 we set
1,−1(λ) = λ1(1 − λ2
mz
1,−1(λ) = λ1(1 − λ2
mz
2/4)1/2 − λ2(1 − λ2
1/4)1/22z
2/4)1/2 − λ2(1 − λ2
1/4)1/22z
φ(λ2/λ1)(η⊗
0 )(λ),
φ(λ1/λ2)(η⊗
0 )(λ).
Then using (3.1) (in the case d = 1) we rewrite T1,−1 as
T1,−1(f, g) = Bms
1,−1(H(−∆Z)sf, (I − H)g) + B ms
1,−1 (Hf, (I − H)(−∆Z)sg).
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
15
Note that A is preserved by (−∆Z)s. Thus, by Lemma 3.3, to demonstrate (3.8) it is enough
to prove, for f, g ∈ A, the bounds
(3.9)
kBms
kB ms
1,−1 (Hf, (I − H)g)kp ≤ CkHfkp1k(I − H)gkp2 ,
1,−1 (Hf, (I − H)g)kp ≤ CkHfkp1k(I − H)gkp2 .
We focus only on the first estimate, the reasoning for the second being analogous. We are
going to apply Stein's complex interpolation theorem [22] for each fixed f ∈ A. The argument
used here takes ideas from the proof of [16, Theorem 1.4]. For further reference we note that
the formula
(3.10)
Bmz
1,−1 (Hf, (I − H)g)(n) =Z 1/2
φ(cid:18) sin πξ2
sin πξ1(cid:19)η0( sin πξ1)η0( sin πξ2)
× sin πξ1p1 − sin2 πξ2 − sin πξ2p1 − sin2 πξ12z
sin πξ12z
e2πi(ξ1+ξ2)n FZ(f )(ξ1)FZ(g)(ξ2) dξ;
Z 0
−1/2
0
makes sense not only for f, g ∈ A but more generally, for f, g ∈ ℓ2.
Mikhlin-Hörmander condition (2.2) of order 8. Thus, Theorem 3.1 (with d = 1) gives
Let n be an even integer larger than 8. Then the multipliers mn+iv
1,−1 , v ∈ R, satisfy the
kBmn+iv
1,−1
Now, Lemma 3.4 applied to ϕ(λ) = φ(λ2/λ1)η⊗
(Hf, (I − H)g)kp ≤ C(1 + v)8kHfkp1k(I − H)gkp2 ,
0 (λ), λ ∈ (0,∞)2, implies
v ∈ R.
Bmiv
1,−1
(Hf, (I − H)g) = (−∆Z)iv(cid:2)Bφ(λ2/λ1)η⊗
0
(H(−∆Z)−ivf, (I − H)g)(cid:3).
By [1, Theorem 1.1] we have k(−∆Z)ivkℓq→ℓq ≤ Cq(1 + v)4, 1 < q < ∞. Hence, Theorem 3.1
applied to the multiplier φ(λ1/λ2)η⊗
0 produces
kBmiv
1,−1
(Hf, (I − H)g)kp ≤ C(1 + v)8kHfkp1k(I − H)gkp2,
v ∈ R.
By (3.10), for fixed f ∈ A, the family {Bmz
1,−1 (Hf, (I − H)g)}Re(z)>0 consists of analytic
operators. This family has admissible growth, more precisely, for each finitely supported g, h
we have
Consequently, an application of Stein's complex interpolation theorem is permitted and leads
to the first inequality in (3.9). The proof of the corollary is thus finished.
(cid:3)
(cid:12)(cid:12)hBmz
1,−1 (Hf, (I − H)g), hil2(Z)(cid:12)(cid:12) ≤ Cf,g,h,
Re(z) ≤ s.
4. Bilinear radial multipliers for the generic Dunkl transform
Here we apply Theorem 2.3 for bilinear multiplier operators associated with the generic
Dunkl transform. In the case when the underlying group of reflections is isomorphic to Z2 we
also prove a fractional Leibniz rule.
Let R be a root system in Rd and G the associated reflection group (see [19, Chapter 2]).
Let σα(x) denote the reflection of x in the hyper-plane orthogonal to α ∈ Rd and let κ be a
nonnegative, G invariant function on R. The differential-difference (rational) Dunkl operators,
are defined as
δjf (x) = ∂jf (x) + Xα∈R+
αjκ(α)
f (x) − f (σα(x))
hα, xi
,
j = 1, . . . , d.
16
BŁAŻEJ WRÓBEL
Here f is a Schwartz function, R+ is a fixed positive subsystem of R and hx, yi = Pd
j=1 xjyj
is the standard inner product. The fundamental property of the operators δj is that, similarly
to the usual partial derivatives (which appear when we take κ ≡ 0), they commute,
i.e.
δlδj = δjδl, l, j = 1, . . . , d. The operators δj are also symmetric on L2 = L2(Rd, w(x)dx), with
i=1 hα, xi2κ(α) . Moreover they leave S(Rd) invariant. Additionally the
w(x) = wκ(x) := Qd
Leibniz rule
(4.1)
δj(f1f2)(x) = δj(f1)(x)f2(x) + δj(f1)(x)f2(x),
x ∈ Rd,
holds under the extra assumption that one of the functions f1, f2 is invariant under G.
The easiest case of Dunkl operators arrises when G ∼ Zd
reflections through the coordinate axes. In this case
2. In other words G consists of
δjf (x) = ∂jf (x) + κj
f (x) − f (σj(x))
xj
,
j = 1, . . . , d,
where κj ≥ 0, while σj(x) denotes the reflection of x in the hyperplane orthogonal to the j-th
coordinate vector. In this case the weight wκ(x) takes the product form wκ(x) =Qd
j=1 wκj (xj),
x ∈ Rd.
transform. It is defined by
In the (general) Dunkl setting there is an analogue of the Fourier transform, called the Dunkl
Df (ξ) = cκZRd
E(−iξ, x)f (x)wκ(x) dx
where E(z, w) = Eκ(z, w) = Eκ(w, z) is the so called Dunkl kernel. A defining property of
this kernel is the equation
(4.2)
The operator D has properties similar to the Fourier transform. Namely, we have the Plancherel
formula
δj,x(Eκ(iξ, x)) = iξjEκ(iξ, x),
x ∈ Rd.
(4.3)
ZRd
f (x)g(x) w(x) dx = cκ ZRd D(f )(ξ)D(h)(ξ) w(ξ) dξ,
and the inversion formula,
(4.4)
f (x) = D2f (−x) = cZRd D(f )(ξ)E(iξ, x) w(ξ) dξ,
f ∈ S(Rd).
Additionally, the Dunkl transform diagonalizes simultaneously the Dunkl operators δi, i.e.
(4.5)
Dδjf = iξjD.
δjDf = −D(ixjf ),
The Dunkl Laplacian is given by ∆κ =Pd
the operator −∆κ may be formally defined as a non-negative self-adjoint operator on L2(Rd, w).
The same is true for L := (−∆κ)1/2. Then, for a bounded function µ the spectral multiplier
µ(L) is uniquely determined on S(Rd) by
(4.6)
D(∆κf )(ξ) = −ξ2 D(f )(ξ),
i . Using the identity
ξ ∈ Rd,
i=1 δ2
D(µ(L)f )(ξ) = µ(ξ)D(f )(ξ)
ξ ∈ Rd.
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
17
Consider now L1 := L ⊗ I and L2 = I ⊗ L. Analogously to the case of bilinear Fourier
multipliers the formula (2.6) can given by the Dunkl transform. Namely, for a bounded function
m : [0,∞)2 → C we have
Bm(f1, f2)(x)
(4.7)
=ZRdZRd
m(ξ1,ξ2)D(f )(ξ1)D(g)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ1dξ2.
The above formula is valid pointwise e.g. for Schwartz functions f1 and f2 on Rd. We observe
that in this section the space A2 from (2.5) is
(4.8)
Thus, by (4.5) the Dunkl derivatives δj, j = 1, . . . , d, preserve A2.
lution. For x, y ∈ Rd The Dunkl translation is defined by
A2 = {g ∈ L2(Rd, wκ) : there is ε > 0 such that D(g)(ξ) = 0 for ξ 6∈ [ε, ε−1]}.
In this section we will heavily rely on the concepts of Dunkl translation and Dunkl convo-
τ yf (x) = cκZRd D(f )(ξ)E(iξ, x)E(iξ, y) w(ξ) dξ.
The inversion formula (4.4) and the properties of the Dunkl kernel imply
For f, g ∈ A the Dunkl convolution is
D(τ yf )(ξ) = E(−iξ, y)D(f )(ξ).
f ∗κ g(x) =ZRd
f (y) τxg(y) w(y) dy,
The first result of this section is the following Coifman-Meyer type theorem. In what follows
where g(x) = g(−x). It is known that the Dunkl transform turns this convolution into multi-
plication, i.e.
(4.9) D(f ∗κ g)(x) = D(f )(x)D(g)(x),
f, g ∈ A.
we set λκ = (d− 1)/2 +Pα∈R+ κ(α) and for brevity write Lp := Lp(Rd, wκ) and k·kp = k·kLp.
Theorem 4.1. Assume that m satisfies the Mikhlin-Hörmander condition (2.2) of an order
s > 2λκ + 6. Then the bilinear multiplier operator given by (4.7) is bounded from Lp1 × Lp2 to
Lp, where 1/p1 + 1/p2 = 1/p, and p1, p2, p > 1. Moreover, the bound (2.7) holds.
[D(f ) ∗κ D(g)](x) = D(f g)(x),
Proof. We are going to apply Theorem 2.3. In order to do so we need to check that its as-
sumptions are satisfied for the operator L = (−∆κ)1/2. To see that L is injective on its domain
we merely note that wκ(ξ) dξ is absolutely continuous with respect to Lebesgue measure. The
contractivity condition (CT) follows from [19, Theorem 4.8] and the subordination method.
The assumption (WD) is straightforward from (4.7) and the Lebesgue dominated convergence
theorem, while (MH) was proved by Dai and Wang [8, Theorem 4.1] (with arbitrary ρ > λκ+1).
Thus we are left with verifying the property (PF), which we prove with b = 2. This will
be deduced by using the convolution structure associated with Dunkl operators. Let ϕk and
ψk, be smooth functions such that supp ϕk ⊆ [0, 2k−2] and supp ψk ⊆ [2k−1, 2k+1]. Let ψk be
a smooth function equal 1 on [2k−5, 2k+5] and vanishing outside of [2k−7, 2k+7]. Taking the
Dunkl transform of the both sides of (PF) and using (4.6) we see that our task is equivalent
to proving the formula
D(ϕk(L)(f1)ψk(L)(f2)) = ψk(ξ)D(ϕk(L)(f1)ψk(L)(f2)),
ξ ∈ Rd.
18
BŁAŻEJ WRÓBEL
Denote gj = D(fj), j = 1, 2. By (4.9) and (4.6) the equation above is exactly
[(ϕk( · )g1) ∗κ (ψk( · )g2)](ξ) = ψk(ξ)[(ϕk( · )g1) ∗κ (ψk( · )g2)](ξ),
By definition of ψ to prove the last formula it is enough to show that
ξ ∈ Rd.
supp[h1 ∗κ h2] ⊆ [2k−5, 2k+5],
(4.10)
for any functions h1 supported in B(0, 2k−2) and h2 supported in B(0, 2k+1)\ B(0, 2k−1). Take
ξ 6∈ [2k−5, 2k+5] and y ∈ B(0, 2k−2). We claim that τ ξ h2(y) = 0. This implies (4.10).
Till the end of the proof we thus focus on proving the claim. Let γξ,y be the distribution
given by γξ,y(f ) = (τ ξf )(y), f ∈ S(Rd). In [2, Theorem 5.1] Amri, Anker, and Sifi proved that
γξ,y is supported in the spherical shell
Sξ,y :=(cid:8)z ∈ Rd : ξ − y ≤ z ≤ ξ + y(cid:9).
Therefore, if we prove that supp h2 ∩ Sξ,y = ∅, then τ ξh2(y) = 0. Recall that we have ξ 6∈
[2k−5, 2k+5] and y ∈ B(0, 2k−2). Take z ∈ Sξ,y and consider two possibilities, either ξ < 2k−5
or ξ > 2k+5. In the first case we obtain z ≤ 2k−5 + 2k−2 < 2k−1, while in the second
z ≥ ξ − y ≥ 2k+5 − 2k−2 > 2k+1. Thus, in both the cases z 6∈ supp h2, and the proof of
(PF) is completed.
(cid:3)
Theorem 4.1 is quite far from a general bilinear Dunkl multiplier theorem, i.e. when the
multiplier function m is not necessarily radial in each of its variables. However, in the case
d = 1 (and G ∼ Z2), Theorem 4.1 implies [3, Theorem 4.1] by Amri, Gasmi, and Sifi. We
slightly abuse the notation and, for ϕ : R2 → C, f1, f2 ∈ A, and x ∈ R, define
(4.11)
ϕ(ξ)D(f1)(ξ1)D(f2)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2) dξ.
Bϕ(f1, f2)(x) =ZRZR
This will cause no confusion with (4.7), as till the end of the present section we only use Bϕ
given by (4.11).
Corollary 4.2 (Theorem 4.1 of [3]). Let G ∼ Z2. Assume that ϕ : R2 → C satisfies the
Mikhlin-Hörmander condition on R2 of an order s > 2λκ + 6, namely
ξ∈R2 ξα∂αϕ(ξ1, ξ2) < ∞.
kϕkM H(R2,s) := sup
(4.12)
sup
α≤s
Then the bilinear multiplier operator given by (4.11) is bounded from Lp1 × Lp2 to Lp, where
1/p1 + 1/p2 = 1/p, and p1, p2, p > 1.
Remark. When κ = 0 we recover the Coifman-Meyer multiplier theorem in the Fourier trans-
form setting.
Proof of Corollary 4.2 (sketch). Let Π(f )(x) = D−1(χξ>0)D(f )(ξ))(x) be the projection onto
the positive Dunkl frequencies. The corollary can be deduced from the boundedness of Π on
all Lp spaces 1 < p < ∞.
(cid:3)
For Re z ≥ 0, let (−∆κ)z be the complex Dunkl derivative
D[(−∆κ)z(h)](ξ) = ξ2zD(h)(ξ),
ξ ∈ Rd.
The natural L2 domain of this operator is
DomL2((−∆κ)z) = {h ∈ L2 : ξ2 Re zD(h)(ξ) ∈ L2}.
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
19
By Plancherel's formula for the Dunkl transform (−∆κ)z(h) ∈ L2 for h ∈ A. The second main
result of this section is the following fractional Leibniz rule for (−∆κ)s, in the case G ∼ Zd
2.
Corollary 4.3. Let G ∼ Zd
s > 0, we have
2 and take 1/p = 1/p1 + 1/p2, with p, p1, p2 > 1. Then, for any
k(−∆κ)s(f g)kp . k(−∆κ)s(f )kp1kgkp2 + kfkp1k(−∆κ)s(g)kp2 ,
where f, g ∈ A and at least one of the functions f or g is invariant by G.
Before proving the fractional Leibniz rule we need a lemma which is an analogue of Lemma
3.4. Its proof is similar, however a bit more technical. Therefore we give more details.
Lemma 4.4. Take d = 1 and let G ∼ Z2. Assume that at least one of the functions f, g ∈ A
is G-invariant. Take Re(z) ≥ 0 and let ϕ : R2 → C be a bounded function that satisfies the
Mikhlin-Hörmander condition (4.11) of order s > 2λκ + 6. Then
(−∆κ)z(Bϕ(f, g))(x) =ZZR2
for almost all x ∈ Rd.
Remark. It is not obvious why Bϕ(f, g) ∈ DomL2((−∆κ)z). This is explained in the proof of
the lemma.
ϕ(ξ)ξ1+ξ22z D(f )(ξ1)D(g)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ,
Proof. Since the argument is symmetric in f and g we assume that f is G-invariant. Denote
EG(iξ1, x) = G−1Pg∈G E(iξ1, gx), and observe that EG is G-invariant in x. Then, since both
f and D(f ) are G-invariant our task reduces to proving that
(4.13)
(−∆κ)z(Bϕ(f, g))(x) =ZRZR
ϕ(ξ)ξ1+ξ22z D(f )(ξ1)D(g)(ξ2) EG(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ,
for almost all x ∈ Rd.
For z = n ∈ N this formula is a direct computation, and follows from the Leibniz rule.
Indeed, by (4.1) and (4.2) we have
δ(Bϕ(f, g))(x) =ZZR2
=ZZR2
ϕ(ξ)D(f )(ξ1)D(g)(ξ2) i(ξ1 + ξ2) EG(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ,
ϕ(ξ)D(f )(ξ1)D(g)(ξ2) δx[EG(iξ1, x)E(iξ2, x)] w(ξ1)w(ξ2)dξ
the interchange of differentiation and integration being allowed since f, g ∈ A. Iterating the
above equality 2n times we obtain (4.13) for z = n.
We remark that (4.13) for z ∈ N also explains why does (−∆κ)z(Bϕ(f, g)) make sense
for general Re(z) ≥ 0.
Indeed, let n be an integer larger than Re(z). Then, to prove that
Bϕ(f, g) ∈ DomL2((−∆κ)z) it is enough to show that Bϕ(f, g) ∈ DomL2((−∆κ)n). Now, using
(4.13) for z = n, together with the binomial formula and (4.5), we arrive at
2n
(−∆κ)n(Bϕ(f, g))(x)
Xj=0(cid:18)2n
j (cid:19)ZRZR
=
ϕ(ξ)D(δj f )(ξ1)D(δ2n−j g)(ξ2) EG(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ,
20
BŁAŻEJ WRÓBEL
with δ being the Dunkl operator on R. Since f, g belong to A2 the same is true for δj f and
δ2n−j g. Thus, an application of Corollary 4.2 proves that Bϕ(f, g) ∈ DomL2((−∆κ)n), as
desired.
We come back to demonstrating (4.13) for general Re(z) ≥ 0. Note first that by a continuity
Tz(f, g)(x) =ZZR2
ϕ(ξ)ξ1 + ξ22z D(f )(ξ1)D(g)(ξ1) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ.
argument it suffices to consider Re(z) > 0. Denoting
our task is reduced to proving that
(4.14)
for h ∈ A2∩S(R) (recall that A2 is given by (4.8)). This is enough because A2∩S(R) is dense
in L2. From (4.13) for z ∈ N we deduce that for any polynomial P it holds
h(−∆κ)z(Bϕ(f, g)), hiL2 = hTz(f, g), hiL2 ,
(4.15)
P (−∆κ)(Bϕ(f, g))(x)
=ZZR2
ϕ(ξ)P (ξ1 + ξ22)D(f )(ξ1)D(g)(ξ2) EG(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ.
For brevity we denote by T P (f, g)(x) the right hand side of (4.15). Note that D(f ), D(g), and
D(h) are supported in [−N, N ] for some large N. Let {Pr(t)}r∈N, be a sequence of polynomials
that converges uniformly to tz on [0, 4N 2]. Then, (4.3), (4.5), and (4.15) imply
(4.16)
Pr(ζ2)D(Bϕ(f, g))(ζ)D(¯h)(ζ) w(ζ) dζ = hPr(−∆κ)(Bϕ(f, g)), hiL2
ZR
= hT Pr (f, g), hiL2 .
Now, since suppD(¯h) ⊆ [−N, N ] and D(Bϕ(f, g))D(¯h) ∈ L1, the dominated convergence
theorem shows that the left hand side of (4.16) converges to h(−∆κ)z(Bϕ(f, g)), hiL2 as r →
∞. Similarly, since D(f ) and D(g) are supported in [−N, N ] the expression T Pr (f, g)(x) is
uniformly bounded in r ∈ N and x ∈ R and converges to Tz(f, g)(x) as r → ∞. As h ∈ S(R) the
dominated convergence theorem implies limr→∞hT Pr (f, g), hiL2 = hTz(f, g), hiL2 . Therefore,
(4.14) is justified and hence, also (4.13). This completes the proof of Lemma 4.4.
(cid:3)
We now pass to the proof of Corollary 4.3.
Proof. By repeating the argument from the beginning of the proof of Corollary 3.2 (with sums
replaced by integrals) our task is reduced to d = 1. We devote the present paragraph to a brief
justification of this statement Here we need the fact that for s ≥ 0 and Lj = −δ2
j , the operators
(Lj)s(−∆κ)−s as well as (−∆κ)s(Pd
j=1(Lj)s)−1, are bounded on all Lp, 1 < p < ∞. This is
true by e.g. [24, Corollary 3.2], since in the product setting each Lj, j = 1, . . . , d, generates a
symmetric contraction semigroup. Then we are left with showing that
(4.17)
kLs
j(f g)kp . kLs
jfkp1kgkp2 + kLs
jgkp1kfkp2
cf. (3.7). The proof of (4.17) is similar to that of (3.7), thus we give a sketch when j = 1. For
t ∈ R and x ∈ Rd−1, consider the auxiliary functions fx(t) = f ((t, x)) and gx(t) = g((t, x)).
Then, setting w(1)
i=2 wκi(x), we write
κ (x) =Qd
1(f g)kp =ZRd−1 kLs
kLs
1(fx(·)gx(·))kp
Lp(R,wκ1 ) w(1)
κ (x) dx.
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
21
From this point on we repeat the steps in the proof of (3.7). Namely, we apply the fractional
Leibniz rule for d = 1 and Hölder's inequality (for integrals). We omit the details here. From
now on we focus on proving Corollary 4.3 for d = 1.
Setting
Let φ be a function in C ∞([0,∞)) supported in [0, 1/4] and such that φ(t) + φ(t−1) = 1.
T1(f, g)(x) =ZZR2
T2(f, g)(x) =ZZR2
and using Lemma 4.4 with ϕ ≡ 1 we rewrite
φ(ξ2/ξ1)ξ1 + ξ22s D(f )(ξ1)D(g)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ,
φ(ξ1/ξ2)ξ1 + ξ22s D(f )(ξ1)D(g)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ.
(−∆κ)s(f g) = T1(f, g) + T2(f, g).
From now on the proof resembles that of Corollary 3.2 (in fact it is even easier). We need to
prove, for f, g ∈ A, the estimate
kT1(f, g)kp ≤ Ck(−∆κ)sfkp1kgkp2 ,
kT2(f, g)kp ≤ Ckfkp1k(−∆κ)sgkp2.
We focus only on the first inequality, as the proof of the second is analogous. For Re(z) ≥ 0
we set
mz(ξ1, ξ2) = ξ1 + ξ22z
ξ12z
φ(ξ2/ξ1),
ξ ∈ R2,
so that T1(f, g) = Bms((−∆κ)sf, g). Since A is preserved under (−∆κ)s our task is reduced to
showing that, for s > 0 it holds
(4.18)
kBms (f, g)kp ≤ Ckfkp1kgkp2 ,
f, g ∈ A
As in Section 3 we are going to apply Stein's complex interpolation theorem. To do this we
need to extend Bmz (f, g) outside of A × A, by allowing g to be a simple function. This may
be achieved by a limiting process. Namely, instead of mz we consider mz
. Then,
ε = mze−εξ2
(4.19)
Bms
ε(f, g)(x)
e−εξ2
φ(ξ2/ξ1) ξ1 + ξ22z
:=ZRdZRd
ξ12z D(f )(ξ1)D(g)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ
converges pointwise to Bms(f, g) as ε → 0+, whenever f, g ∈ A. Therefore, by Fatou's Lemma,
to prove (4.18) for Bms it is enough to prove it for each Bms
ε (f, g)kp ≤ Ckfkp1kgkp2 ,
ε, ε > 0, as long as
kBms
where C is independent of ε. The gain is that now (4.19) is well defined for g ∈ L2, in particular
it is valid for simple functions.
, j = 1, 2, v ∈ R, satisfy Hörmander's condition
Let n > 2λκ + 6. Then the multipliers mn+iv
(4.12) of order 2λκ + 6. Thus, using Corollary 4.2 (with d = 1) we obtain
ε
kBmn+iv
ε
(f, g)kp ≤ Cn(1 + v)2λκ+2kf1kp1kf2kp2,
v ∈ R.
Now, Lemma 4.4 applied to ϕ(ξ) = φ(ξ2/ξ1)e−ε(ξ2) implies
Bmiv
ε
(f, g) = (−∆κ)iv(cid:2)Bϕ((−∆κ)−ivf, g)(cid:3).
22
BŁAŻEJ WRÓBEL
Thus, using [8, Theorem 4.1] followed by Corollary 4.2 (for the multiplier φ(ξ2/ξ1)e−ε(ξ2))
we obtain
kBmiv (f, g)kp ≤ C(1 + v)2λκ+2kf1kp1kf2kp2,
v ∈ R.
By definition
Bmz
ε (f, g)
=ZZR2
ξ1 + ξ2z
ξ1z
φ(ξ2/ξ1)e−εξ2
D(f )(ξ1)D(g)(ξ2) E(iξ1, x)E(iξ2, x) w(ξ1)w(ξ2)dξ1dξ2.
Hence, for fixed f1 ∈ A the family {Bmz (f, g)}Re(z)>0 consists of analytic operators. This
family has admissible growth, more precisely, for each simple function h we have
Consequently, using Stein's complex interpolation theorem is permitted and leads to (4.18).
The proof of the corollary is thus finished.
(cid:3)
(cid:12)(cid:12)hBmz ((−∆κ)zf, g), hiL2(cid:12)(cid:12) ≤ Cf,g,h,s,
Re(z) ≤ s.
5. Bilinear multipliers for Jacobi trigonometric polynomials
In this section we give a bilinear multiplier theorem for expansions in terms of Jacobi trigono-
metric polynomials. Contrary to the previous sections we do not prove a fractional Leibniz
rule here. The reason for this is that there is no natural first order operator in the Jacobi
setting that satisfies a Leibniz-type rule of integer order.
Let α, β > −1/2 be fixed, and let P α,β
n
α, β. For n ∈ N and −1 < x < 1 these are given by the Rodrigues formula
be the one-dimensional Jacobi polynomials of type
P α,β
n (x) =
(−1)k
2nn!
(1 − x)−α(1 + x)−β dn
dxnh(1 − x)α+k(1 + x)β+ki.
We now substitute x = cos θ, θ ∈ [0, π], and consider the trigonometric Jacobi polynomials
P α,β
n (cos θ). This is an orthogonal and complete system in L2(dµα,β), where
θ
θ
dµα,β(θ) =(cid:16) sin
(cos θ), where kP α,β
2(cid:17)2β+1
2(cid:17)2α+1(cid:16) cos
n (cos ·)k2 = (cα,β
dθ.
Throughout this chapter we abbreviate Lp := Lp([0, π], µα,β ) and k·kp := k·kLp . Now, setting
Pn(θ) = P α,β
n )−1 we obtain a complete
orthonormal system in L2. Each P α,β
is an eigenfunction of the differential operator
n (θ) = cα,β
n P α,β
n
k
J = J α,β = −
d2
dθ2 −
α − β + (α + β + 1) cos θ
sin θ
with the corresponding eigenvalue being (n+ α+β+1
observe that γ > 0.
2
In this setting the spectral multipliers of J 1/2 are given by
µ(J 1/2)f = Xn∈N
µ(cid:0)n + γ(cid:1)hf,PkiL2 Pk.
d
dθ
+(cid:16) α + β + 1
2
(cid:17)2
;
)2. In what follows we set γ = (α+β +1)/2;
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
23
If µ : R+ → C is bounded, then µ(J 1/2) is a bounded operator on L2. In this section the
formula (2.6) defining bilinear multipliers becomes
(5.1)
Bm(f1, f2)(θ) = m(J 1/2 ⊗ I, I ⊗ J 1/2)(x, x)
= Xn1∈N,n2∈N
m(cid:16)n1 + γ, n2 + γ(cid:17)hf1,Pn1ihf2,Pn2i Pn1(θ)Pn2(θ).
The space A from (2.5) coincides with the linear span of {Pn}n∈N. We prove the following
Coifman-Meyer type multiplier theorem.
Theorem 5.1. Assume that m satisfies Hörmander's condition (2.2) of order s > 4(α+β)+15.
Then the bilinear multiplier operator given by (5.1) is bounded from Lp1 × Lp2 to Lp, where
1/p1 + 1/p2 = 1/p, and p1, p2, p > 1. Moreover, the bound (2.7) is valid.
Remark. The theorem implies a Coifman-Meyer type multiplier result for bilinear multipliers
associated with the modified Hankel transform. This follows from a transference results of
Sato [20].
Proof. Once again the proof hinges on Theorem 2.3. We need to verify that L = J 1/2 satisfies
its assumptions. The injectivity condition is clear since 0 is not an eigenvalue of J 1/2. The
contractivity assumption (CT) can be inferred from the formula
e−tJ (f ◦ cos)(θ) = e−t(α+β+1)2/4T α,β
t
f (cos θ)
relating the semigroup e−tJ with the semigroup T α,β
from the Jacobi polynomial setting, as
T α,β
is well known to be Markovian. The condition (WD) is straightforward, since A is the
t
linear span of Jacobi trigonometric polynomials. The Mikhlin-Hörmander functional calculus
(MH) for J 1/2 (with ρ = 2α + 2β + 13/2) was obtained in [26, Corollary 4.3].
It remains to show (PF). Here we need the following identity
t
(5.2)
Pn1(θ)Pn2(θ) =
j=n1+n2
Xj=n1−n2
cn1,n2(j)Pj (θ).
The above is well known to hold for general orthogonal polynomials on an interval contained
in R, hence also for Pj as they are merely a reparametrisation of the Jacobi polynomials.
We prove that (PF) holds with b = 3. Take f, g ∈ A. Then
f2 = Xn2∈N
c2
n2Pn2,
f1 = Xn1∈N
c1
n1Pn1,
where all but a finite number of c1
n, c2
n vanish. Denote
Since ϕk and ψk are supported in [0, 2k−3] and [2k−1, 2k+1], respectively, we have
Ra,b = {n ∈ N : 2a − γ ≤ n ≤ 2b − γ}.
whereas
ϕk(L)(f1) = Xn1∈N : n1+γ≤2k−3
ψk(L)(f2) = Xn2∈Rk−1,k+1
c1
n1 ϕk(n1 + γ)Pn1
c2
n2 ψk(n2 + γ)Pn2.
24
BŁAŻEJ WRÓBEL
Now, if n1 + γ ≤ 2k−3 and 2k−1 ≤ n2 + γ ≤ 2k+1, then we must also have
n1 − n2 ≥ 2k−1 − 2k−3 ≥ 2k−2
and n1 + n2 ≤ 2k−3 − γ + 2k+1 − γ ≤ 2k+2 − 2γ.
Since γ > 0, we see that if n1− n2 ≤ n ≤ n1 + n2, then 2k−2 ≤ n + γ ≤ 2k+2. Consequently, in
view of (5.2), the operator ψk(L) leaves invariant each product Pn1·Pn2 , hence, also ϕk(L)(f1)·
ψk(L)(f2). The proof of (PF) is thus completed.
(cid:3)
Acknowledgments. I thank prof. Krzysztof Stempak for suggesting the idea to combine
joint spectral multipliers with multilinear multipliers, prof. Christoph Thiele for a discussion
on the Coifman-Meyer multiplier theorem and useful remarks during the preparation of the
paper, prof. Camil Muscalu for a discussion on multilinear multipliers, prof. Herbert Koch,
prof. Fulvio Ricci, and dr Gian Maria Dall'Ara for discussions on the fractional Leibniz rule,
dr hab. Wojciech Młotkowski for bringing to my attention the formula (5.2), and dr Jotsaroop
Kaur for discussions on Hermite polynomials.
Part of the research presented in this paper was carried over while the author was Assegnista
di ricerca at the Università di Milano-Bicocca. The research was supported by Italian PRIN
2010 "Real and complex manifolds: geometry, topology and harmonic analysis"; Polish funds
for sciences, National Science Centre (NCN), Poland, Research Project 2014/15/D/ST1/00405;
and by the Foundation for Polish Science START Scholarship.
References
[1] G. Alexopoulos, Spectral multipliers on discrete groups, Bull. Lond. Math. Soc. (4) 33 (2001), 417 -- 424.
[2] B. Amri, J. P. Anker, and M. Sifi, Three results in Dunkl analysis, Colloquium Math. (1) 118 (2010), pp.
299-312.
[3] B. Amri, A. Gasmi, M. Sifi, Linear and Bilinear Multiplier Operators for the Dunkl Transform, Mediterr.
J. Math. (4) 7 (2010), 503 -- 521.
[4] F. Bernicot, D. Maldonado, K. Moen, and V. Naibo, Bilinear Sobolev-Poincaré inequalities and Leibniz-type
rules, J. Geom. Anal. (2) 24 (2014), pp. 1144-1180.
[5] O. Blasco, Bilinear multipliers and transference, Int. J. Math. Math. Sci. (4) 2005, pp. 545-554.
[6] J. Bourgain, D. Li, On an endpoint Kato-Ponce inequality, Diff. Int. Eq. (11/12) 27 (2014), pp. 1037-1072.
[7] R. Coifman, Y. Meyer, Nonlinear harmonic analysis, operator theory, and PDE in Beijing Lectures in
Harmonic Analysis (Beijing, 1984), Annals of Mathematics Studies 112, Princeton Univ. Press, Princeton,
NJ, 1986, pp. 3 -- 45.
[8] F. Dai, H. Wang, A transference theorem for the Dunkl transform and its applications, J. Funct. Anal. (12)
258 (2010), pp. 4052-4074.
[9] D. Fan, S. Sato, Transference on certain multilinear multiplier operators, J. Austral. Math. Soc., 70 (2001),
pp.37-55.
[10] D. Frey, Paraproducts via H ∞-functional calculus, Rev. Math. Iberoam. (2) 29 (2013), 635-663.
[11] L. Grafakos, S. Oh, The Kato-Ponce inequality, Commun. Part. Diff. Eq. (6) 39 (2014), pp. 1128-1157.
[12] L. Grafakos, R. Torres, Multilinear Calderón-Zygmund theory, Adv. Math. 165 (2002), pp. 124-164.
[13] T. Kato, G. Ponce, Commutator estimates and the Euler and Navier-Stokes equations, Commun. Pure
Appl. Math., 41 (1988), pp. 891-907.
[14] C. Kenig, G. Ponce, and L. Vega, Well-posedness and scattering results for the generalized Korteweg-de
Vries equation via the contraction principle, Comm. Pure Appl. Math., 46 (1993), pp. 527-620.
[15] C. Kenig, E. M. Stein, Multilinear estimates and fractional integrals, Math. Res. Lett. 6 (1999) pp. 1-15.
[16] A. Gulisashvili, M. A. Kon, Exact Smoothing Properties of Schrödinger Semigroups, Amer. J. Math. (6)
118 (1996), pp. 1215-1248.
[17] C. Muscalu, J. Pipher, T. Tao, and C. Thiele, Bi-parameter paraproducts, Acta Math. 193 (2004), pp.
269-296.
[18] C. Muscalu, W. Schlag, Classical and multilinear harmonic analysis, Vol II, Cambridge Studies in advanced
mathematics 138, 2013.
BILINEAR MULTIPLIERS VIA A FUNCTIONAL CALCULUS
25
[19] M. Rösler, Dunkl operators:
theory and applications, in Orthogonal Polynomials and Special Functions
(Leuven, 2002), Lecture Notes in Math., Vol. 1817, Springer, Berlin, 2003, 93Ű135.
[20] E. Sato, Transference of bilinear operators between Jacobi series and Hankel transforms, Taiwanese J.
Math. (4) 15 (2011), pp. 1561-1573.
[21] K. Schmüdgen, Unbounded Self-adjoint Operators on Hilbert Space, Grad. Texts in Math. 265 (2012).
[22] E. Stein, Interpolation of Linear Operators, Trans. Amer. Math. Soc. (2) 83 (1956), pp. 482-492.
[23] B. Wróbel, On the consequences of a Mihlin-Hörmander functional calculus: maximal and square function
estimates, preprint 2015, arXiv:1507.08114.
[24] B. Wróbel, Joint spectral multipliers for mixed systems of operators, J. Fourier Anal. Appl. (2016) online
first, DOI:10.1007/s00041-016-9469-7, pp. 1-43.
[25] B. Wróbel, Multivariate spectral multipliers, PhD thesis, Scuola Normale Superiore, Pisa and Uniwersytet
Wrocławski (2014), http://arxiv.org/abs/1407.2393.
[26] B. Wróbel, Multivariate spectral multipliers for tensor product orthogonal expansions, Monatsh. Math. 168
(2012), 124 -- 149.
Błażej Wróbel: Mathematical Institute, Universität Bonn, Endenicher Allee 60, D -- 53115
Bonn, Germany
& Instytut Matematyczny, Uniwersytet Wrocławski, pl. Grunwaldzki 2/4, 50-384 Wrocław,
Poland
[email protected]
E-mail address:
[email protected]
|
1005.0391 | 1 | 1005 | 2010-05-03T20:03:30 | On the higher rank numerical range of the shift | [
"math.FA",
"math.CV",
"math.OA"
] | For any n-by-n complex matrix T and any $1\leqslant k\leqslant n$, let $\Lambda_{k}(T)$ the set of all $\lambda\in \C$ such that $PTP=\lambda P$ for some rank-k orthogonal projection $P$ be its higher rank-k numerical range. It is shown that if $\bbS$ is the n-dimensional shift on ${\C}^{n}$ then its rank-k numerical range is the circular disc centred in zero and with radius $\cos\dfrac{k\pi}{n+1}$ if $1<k\leqslant[\frac{n+1}{2}]$ and the empty set if $[\frac{n+1}{2}]<k\leqslant n$, where $[x] $ denote the integer part of $x$. This extends and rafines previous results of U. Haagerup, P. de la Harpe \cite{Haagerup} on the classical numerical range of the n-dimensional shift on${\C}^{n}$. An interesting result for higher rank-$k$ numerical range of nilpotent operator is also established. | math.FA | math |
ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT
OPERATOR
HAYKEL GAAYA
centred in zero and with radius cos
Abstract. For any n-by-n complex matrix T and any 1 6 k 6 n, let Λk(T )
the set of all λ ∈ CI such that P T P = λP for some rank-k orthogonal pro-
jection P be its higher rank-k numerical range. It is shown that if Sn is the
n-dimensional shift on CI n then its rank-k numerical range is the circular disc
2 (cid:3) and the empty
set if (cid:2) n+1
2 (cid:3) < k 6 n, where [x] denote the integer part of x. This extends
and rafines previous results of U. Haagerup, P. de la Harpe [8] on the classical
numerical range of the n-dimensional shift on CI n. An interesting result for
higher rank-k numerical range of nilpotent operator is also established.
if 1 < k 6 (cid:2) n+1
kπ
n + 1
1. Introduction
Let H be a complex separable Hilbert space and B(H) the collection of all
bounded linear operator on H. The numerical range of an operators T in B(H) is
the subset
W (T ) = {< T x, x >∈ CI ; x ∈ H, kxk6 1}
of the plane, where < ., . > denotes the inner product in H and the numerical range
of T is defined by
We denote by S the unilateral shift acting on the Hardy space HI 2 of the square
ω2(T ) = sup {z; z ∈ W (T )} .
summable analytic functions.
S : HI 2 → HI 2
7→ zf (z)
f
Beurling's theorem implies that the non zero invariant subspaces of S are of the
forme φ HI 2, where φ is some inner function . Let S(φ) denote the compression of
S to the space H(φ) = HI 2 ⊖ φ HI 2 :
S(φ)f (z) = P (zf (z)),
where P denotes the ortogonal projection from HI 2 onto H(φ). The space H(φ)
is a finite-dimensional exactly when φ is a finite Blaschke product. The numerical
radius and numerical range of the model operator S(φ) seems to be important and
have many applications. In [1], Badea and Cassier showed that there is relation-
ship between numerical radius of S(φ) and Taylor coefficients of positive rational
functions on the torus and more recently in [6], the author gave an extension of this
result. However the evaluation of the numerical radius of S(φ) under an explicit
form is always an open problem. The reader may consult [6] for an estimate of S(φ)
2000 Mathematics Subject Classification. 47A12, 47B35.
Key words and phrases. Operator theory, Numerical radius, Numerical range, higher rank
numerical range, Eigenvalues, Toeplitz forms.
1
2
HAYKEL GAAYA
where φ is a finite Blashke product with unique zero. In the particular case where
φ(z) = zn, S(φ) is unitarily equivalent to Sn where
Sn =
0
1
. . .
. . .
. . .
1
0
.
n+1o
In [8]; it is proved that W (Sn) is the closed disc Dn = nz ∈ CI ; z6 cos π
and ω2(Sn) = cos π
n+1 and more general
Theorem 1.1 ([8]). Let T be an operator on H such that T n = 0 for some n ≥ 2.
One has:
ω2(T ) 6 kT kcos
π
n + 1
and ω2(T ) = kT kcos π
n+1 when T is unitarily equivalent to kT kSn.
In this mathematical note, we extend this result to the higher rank-k numerical
range of the shift. The notion of the the higher rank-k numerical range of T ∈ B(H)
is introduced in [4] and it's denoted by:
Λk(T ) = {λ ∈ CI : P T P = λP for some rank-k orthogonal projection P } ,
The introduction of this notion was motivated by a problem in quantum error
correction; see [5]. If P is a rank-1 orthogonal projection then P = x ⊗ x for some
x ∈ CI n and P T P =< T x, x > P . Then when k = 1, this concept is reduces to the
classical numerical range W (T ), which is well known to be convex by the Toeplitz-
Hausdorff theorem; for exemple see [10] for a simple proof. In [2], it's conjectured
that Λk(T ) is convex, and reduced the convexity problem to the problem of showing
that 0 ∈ Λk(T ′) where
T ′ =(cid:18) Ik X
Y −Ik (cid:19)
for arbitrary X, Y ∈ Mk (the algebra of k × k complex matrix). They further
reduced this problem to the existance of a Hermitian matrix H satisfying the matrix
equation
(1.1)
Ik + M H + HM ∗ − HRH = H
for arbitrary M ∈ Mk and a positive definite R ∈ Mk. In [16], H. Woerdeman
proved that equation (1.1) is equivalent to Ricatti equation:
(1.2)
HRH − H(M ∗ − Ik/2) − (M − Ik/2)H − Ik = 0k,
and using the theory of Ricatti equations (see [9], Theorem 4), the equation (1.2)
is solvable which prove the convexity of Λk(T ). In [4], the authors showed that if
dimH < ∞ and T ∈ B(H) is a Hermitian matrix with eigenvalues λ1 6 λ2 · · · 6 λn
then the rank-k nuemrical range Λk(T ) coincides with [λk, λn+1−k] which is a non-
degenerate closed interval if λk < λn+1−k, a singleton set if λk = λn+1−k and an
empty set if λk > λn+1−k. In [13], the authors proved that if dimH = n
Λk(T ) = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθT + e−iθT ∗(cid:1)(cid:9) ,
ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR
3
for 1 6 k 6 n, where λk(H) denote the kth largest eigenvalue of the hermitian
matrix H ∈ Mn. This result establishes that if dimH = n and T ∈ B(H) is a
normal matrix with eigenvalues λ1, . . . , λn then
Λk(T ) =
\
16j1<···<jn−k+16n
conv(cid:8)λj1 , . . . , λjn+1−k(cid:9) .
We close this section by the following properties wich are easly checked. The
reader may consult [2],[3],[4],[5],[7] and [11].
P1. For any a and b ∈ CI , Λk(aT + bI) = aΛk(T ) + b.
P2. Λk(T ∗) = Λk(T ).
P3. Λk(T ⊕ S) ⊇ Λk(T ) ∪ Λk(S).
P4. For any unitary U ∈ B(H), Λk(U ∗T U ) = Λk(T ).
P5. If T0 is a compression of T on a subspace H0 of H such that dimH0 ≥ k,
then Λk(T0) ⊆ Λk(T ).
P6. W (T ) ⊇ Λ2(T ) ⊇ Λ3(T ) ⊇ . . . .
Some results from [1] will be also developed in this context in a forthcoming paper.
2. main theorem
In the following theorem we give the higher rank-k numerical range of the n-
dimensional shift on CI n.
Theorem 2.1. For any n ≥ 2 and 1 6 k 6 n, Λk(Sn) coincides with the circular
disc {z ∈ CI : z6 cos
kπ
n + 1
} if 1 6 k 6(cid:2) n+1
2 (cid:3) and the empty set if(cid:2) n+1
2 (cid:3) < k 6 n.
Proof. First observe that
Λk(Sn) = \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθSn + e−iθS ∗
n(cid:1)(cid:9)
n(cid:1)(cid:27)
λk(cid:0)eiθSn + e−iθS ∗
n(cid:1)(cid:27)
λk(cid:0)eiθSn + e−iθS ∗
= \θ∈[0,2π[(cid:26)µ ∈ CI : Re(eiθµ) 6
eiθ(cid:26)z ∈ CI : Re(z) 6
= \θ∈[0,2π[
1
2
1
2
(2.1)
On the other hand, we have
eiθSn + e−iθS ∗
n =
0
eiθ
0
...
0
0
e−iθ
0
eiθ
...
0
0
0
e−iθ
0
...
0
0
. . .
. . .
. . .
. . .
. . .
. . .
0
0
0
...
0
eiθ
0 . . .
0 . . .
0
...
e−iθ
0
.
Note that eiθSn + e−iθS ∗
n is a Toeplitz matrix associated to the Toeplitz form
fθ(t) = 2 cos(θ + t).
4
HAYKEL GAAYA
The eigenvalues satisfy the caracteristic equation
∆n(λ) = Det(cid:0)eiθSn + e−iθS ∗
n(cid: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)
−λ e−iθ
eiθ −λ
eiθ
0
...
...
0
0
0
0
0
0
0
...
0 . . .
0 . . .
. . .
. . .
. . .
0
...
. . .
. . . −λ e−iθ
eiθ −λ
. . .
e−iθ
−λ
...
0
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)
Expanding this determinant, we obtain the recurrence relation
∆n(λ) = −λ∆n−1 − ∆n−2, n = 2, 3, 4, . . . ,
This recurrence relation holds also for n = 1 provided we put ∆0 = 1 and ∆−1 = 0.
In order to find an explicit representation of ∆n(λ), we write convenently
λ = 2 cos(θ + t) = fθ(t)
and form the caracteristic equation
ρ2 = −λρ − 1 = −2ρ cos(θ + t) − 1
with the roots −ei(θ+t) and −e−i(θ+t) so that
∆n(2 cos(θ + t)) = (−1)n(Aein(θ+t) + Be−in(θ+t))
where the constants A and B can be determined from the cases n = −1 and n = 0.
Thus
∆n(2 cos(θ + t)) = (−1)n sin((n + 1)(θ + t))
sin(θ + t).
This yields the eigenvalues
λν = 2 cos(
νπ
n + 1
), ν = 1, 2, . . . n.
This implies of course that
Λk(Sn) = \θ∈[0,2π[
eiθ(cid:26)z ∈ CI : Re(z) 6 cos(
kπ
n + 1
)(cid:27)
Thus Λk(Sn) is the intersection of closed half planes. We note that cos(
positive if and only if k 6(cid:2) n+1
2 (cid:3).
Case 1.If k 6(cid:2) n+1
Case 2. If k >(cid:2) n+1
Λk(Sn) ⊆ (cid:26)z ∈ CI : Re(z) 6 cos(
= ∅.
2 (cid:3) In this case Λk(Sn) is circular disc {z ∈ CI : z6 cos
2 (cid:3), then
kπ
n + 1
)(cid:27)\ eiπ(cid:26)z ∈ CI : Re(z) 6 cos(
kπ
n + 1
) is
kπ
n + 1
}.
kπ
n + 1
)(cid:27)
(cid:3)
This completes the proof.
Theorem 2.2. For any integer k ≥ 1
Λk(S) = D(0, 1)
ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR
5
Proof. Let a fixed k ≥ 1, we have D(0, 1) ⊆ Λk(S) which is du to (P.5) and
theorem (2.1). Now let λ in Λk(S) then there exists a rank-k orthogonal projection
P such that P SP = λP . Let denote by Uθ the unitary operator on HI 2 defined
by Uθ(f )(z) = f (ze−iθ), then if we denote by Q the rank-k orthogonal projection
θ we can easly check that QSQ = λeiθ which implies that Λk(S) is a circular
UθP U ∗
disc centred in 0. On the other hand if 1 ∈ Λk(S) then 1 ∈ W (S) and there exists
a unitary f ∈ HI 2 such that < Sf, f >= 1 wich implies that 1 is un eigenvalue for
S which is absurd.
(cid:3)
On the sequel of this paper, let denote by
ρ(k, r) =(cid:26) k/r
[k/r] + 1 unless
if k/r is is integer
where k and r are arbitrary numbers.
Lemma 2.3. For a fixed n ≥ 1 and r ≥ 1, let denote by λ1 > · · · > λn; n real
numbers and (λ′
p)16p6nr a finite sequence defined by:
1 = · · · = λ′
r = λ1, . . . , λ′
(n−1)r+1 = · · · = λ′
λ′
nr = λn.
Then for each 1 6 k 6 nr, the kth largest term of (λ′
t)16t6nr is λρ(k,r).
Proof. The claim is obvious in the case where r = 1. We may assume r ≥ 2. We
prove the result by induction on k. If k = 1, then the largest term is λ1 = λρ(1,r).
So the result hold for k = 1. Assume that k > 1, and the reslut is valid for the mth
largest term of (λ′
t)16t6nr whenever m < k.
Case 1. Suppose that ρ(k − 1, r) = k−1
k − 1 = pr. By induction assumption, we have λρ(k−1,r) = λ′
that the kth largest term of (λ′
r , then there exists 1 6 p 6 n − 1 such that
pr = λp, which implies
t)16t6nr is
λ′
pr+1 = λp+1 = λ k−1
r +1 = λ[ k
r ]+1 = λρ(k,r).
Case 2. Suppose that ρ(k − 1, r) = [ k−1
r ] + 1, then there exist 1 6 q 6 n − 1 and
1 6 s 6 r − 1 such that k − 1 = qr + s. First, note that ρ(k − 1, r) = ρ(k, r).
On the other hand, by induction assumption, we have λρ(k−1,r) = λ′
qr+s = λq+1.
Consequently the kth largest term of (λ′
t)16t6nr is
λ′
qr+s+1 = λq+1 = λρ(k−1,r) = λρ(k,r).
The proof is now complete.
(cid:3)
Let DT = (IN − T ∗T )1/2 be the defect operator of T and DT the closed range
of DT . Let denote by r = dimDT .
Theorem 2.4. Consider T ∈ B(H) such that kT k6 1 and T n = 0. Then Λk(T )
is contained in the circular disc {z ∈ CI : z6 cos( ρ(k,r)π
n+1 )} if 1 6 ρ(k, r) 6 [ n+1
2 ]
and empty if ρ(k, r) > [ n+1
2 ].
Proof. If T is a contaction with T n = 0, then T can be viewed as a compression of
n acting on the Hilbert space DT ⊗ CI n. Consider the isometry H → DT ⊗ CI n,
Ir ⊗S ∗
V (x) =Xn
t=1
DT T t−1x ⊗ et
6
HAYKEL GAAYA
where {el}n
l=1 is the canonical basis of CI n. Note that
V T x =Xn
t=1
It follows that
DT T tx ⊗ et =Xn−1
t=1
DT T tx ⊗ et = (Ir ⊗ S ∗
n)V x.
T = V ∗(Ir ⊗ S ∗
n)V
and from (P.5)
(2.2)
Λk(T ) = Λk(V ∗(Ir ⊗ S ∗
n)V ) ⊆ Λk(Ir ⊗ S ∗
n), for any 1 6 k 6 nr.
Now,
Λk(Ir ⊗ S ∗
n)
n) + e−iθ(Ir ⊗ S ∗
n)∗(cid:1)(cid:9)
n) + e−iθ(Ir ⊗ Sn)(cid:1)(cid:9)
= \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθ(Ir ⊗ S ∗
= \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)eiθ(Ir ⊗ S ∗
= \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)Ir ⊗ (eiθSn + e−iθS ∗
n)(cid:1)(cid:9)
= \θ∈[0,2π[(cid:8)µ ∈ CI : eiθµ + e−iθµ 6 λk(cid:0)⊕r
n)(cid:1)(cid:9)
= \θ∈[0,2π[
eiθ(cid:26)z ∈ CI : Re(z) 6 cos(
i (eiθSn + e−iθS ∗
ρ(k, r)π
n + 1
) (cid:27)
where the last equality is due to the lemma (2.2) and theorem (2.1). Thus
Λk(Ir ⊗ S ∗
n) =( D(0, cos( ρ(k,r)π
n+1 ))
∅
if 1 6 ρ(k, r) 6 [ n+1
2 ]
if [ n+1
2 ] < ρ(k, r) 6 n
Therefore,
if 1 6 k 6 nr, (2.2) implies that Λk(T ) ⊆ D(0, cos( ρ(k,r)π
and empty if [ n+1
2 ] < ρ(k, r) 6 n. Finally, if k > nr, Λk(T ) = ∅ from (P6).
n+1 )) if 1 6 ρ(k, r) 6 [ n+1
2 ]
(cid:3)
Corollary 2.5 (U. Haagerup, P. de la Harpe,[8]). Consider T ∈ B(H) such that
kT k6 1 and T n = 0. Then we have ω2(T ) 6 cos( π
n+1 ).
Proof. T = V ∗(Ir ⊗ S ∗
n)V where V : H → DT ⊗ CI n,
Now
V (x) =Xn
t=1
DT T t−1x ⊗ et.
W (T ) = Λ1(T ) = Λ1(V ∗(Ir ⊗ Sn)V ) ⊆ Λ1(Ir ⊗ Sn) = D(0, cos
π
n + 1
).
(cid:3)
Acknowledgements: The author would like to express his gratitude to Gilles Cassier
for his help and his good advices.
ON THE HIGHER RANK NUMERICAL RANGE OF THE SHIFT OPERATOR
7
References
[1] C. Badea and G. Cassier, Constrained von neumann inequalities, Adv. Math. 166 (2002), no.
2, 260 -- 297.
[2] M.-D. Choi, M. Giesinger, J. A. Holbrook, and D. W. Kribs, Geometry of higher-rank nu-
merical ranges, Linear and Multilinear Algebra 56 (2008), 53-64.
[3] M.-D. Choi, J. A. Holbrook, D.W. Kribs, and K. Zyczkowski, Higher-rank numerical ranges
of unitary and normal matrices, Operators and Matrices 1 (2007), 409-426.
[4] M.-D. Choi, D. W. Kribs, and K. Zyczkowski, Higher-rank numerical ranges and compression
problems, Linear Algebra Appl. 418 (2006), 828-839.
[5] M.-D. Choi, D. W. Kribs, and K. Zyczkowski, Quantum error correcting codes from the
compression formalism, Rep. Math. Phys. 58 (2006), 77-91.
[6] H. Gaaya, On the numerical radius of the truncated adjoint Shift. to appear.
[7] H.L. Gau, C.K. Li, P.Y. Wu, Higher-rank numerical ranges and dilations, J. Operator Theory,
in press.
[8] U.Haagerup and P. dela Harpe, The numerical radius of a nilpotent operator on a Hilbert
space, Proc. Amer. Math. Soc. 115(1992), 371 -- 379.
[9] P. Lancaster and L. Rodman, Algebraic Riccati equations, Oxford Science Publications, The
Clarendon Press, Oxford University Press, New York, 1995.
[10] C.-K. Li, A simple proof of the elliptical range theorem, Proc. Amer. Math. Soc. 124 (1996),
1985-1986.
[11] C.-K. Li, Y. T. Poon and N.-S. Sze, Condition for the higher rank numerical range to be
non-empty, Linear and Multilinear Algebra, to appear.
[12] C.-K. Li, Y. T. Poon and N.-S. Sze, Higher rank numerical ranges and low rank perturbations
of quantum channels, preprint. http://arxiv.org/abs/0710.2898.
[13] C.-K. Li, N.-S. Sze, Canonical forms, higher rank numerical ranges, totally isotropic sub-
spaces, and matrix equations, Proc. Amer. Math. Soc. Volume 136, Number 9, September
2008, Pages 3013 -- 3023 .
[14] C. Pop, On a result of Haagerup and de la Harpe. Rev. Roumaine Math. Pures Appl. 43
(1998), no. 9-10, 869 -- 871.
[15] G. W. Stewart and J.-G. Sun, Matrix Perturbation Theory, Academic Press, New York, 1990.
[16] H. Woerdeman, The higher rank numerical range is convex, Linear and Multilinear Algebra
56 (2008), 65-67.
‡.Institute Camille Jordan, Office 107 University of Lyon1, 43 Bd November 11,
1918, 69622-Villeurbanne, France.
E-mail address: ‡[email protected]
|
1306.2423 | 1 | 1306 | 2013-06-11T05:31:18 | Numerical Radii for Tensor Products of Matrices | [
"math.FA"
] | For $n$-by-$n$ and $m$-by-$m$ complex matrices $A$ and $B$, it is known that the inequality $w(A\otimes B)\le\|A\|w(B)$ holds, where $w(\cdot)$ and $\|\cdot\|$ denote, respectively, the numerical radius and the operator norm of a matrix. In this paper, we consider when this becomes an equality. We show that (1) if $\|A\|=1$ and $w(A\otimes B)=w(B)$, then either $A$ has a unitary part or $A$ is completely nonunitary and the numerical range $W(B)$ of $B$ is a circular disc centered at the origin, (2) if $\|A\|=\|A^k\|=1$ for some $k$, $1\le k<\infty$, then $w(A)\ge\cos(\pi/(k+2))$, and, moreover, the equality holds if and only if $A$ is unitarily similar to the direct sum of the $(k+1)$-by-$(k+1)$ Jordan block $J_{k+1}$ and a matrix $B$ with $w(B)\le\cos(\pi/(k+2))$, and (3) if $B$ is a nonnegative matrix with its real part (permutationally) irreducible, then $w(A\otimes B)=\|A\|w(B)$ if and only if either $p_A=\infty$ or $n_B\le p_A<\infty$ and $B$ is permutationally similar to a block-shift matrix \[[
{array}{cccc}
0 & B_1 & &
& 0 & \ddots &
& & \ddots & B_k
& & & 0
{array}
]\] with $k=n_B$, where $p_A=\sup\{\ell\ge 1: \|A^{\ell}\|=\|A\|^{\ell}\}$ and $n_B=\sup\{\ell\ge 1 : B^{\ell}\neq 0\}$. | math.FA | math |
Numerical Radii for Tensor Products of Matrices
Hwa-Long Gaua∗, Kuo-Zhong Wangb, Pei Yuan Wub
aDepartment of Mathematics, National Central University, Chung-Li 32001, Taiwan
bDepartment of Applied Mathematics, National Chiao Tung University, Hsinchu 30010,
Taiwan
Abstract. For n-by-n and m-by-m complex matrices A and B, it is known that the
inequality w(A ⊗ B) ≤ kAkw(B) holds, where w(·) and k · k denote, respectively, the
numerical radius and the operator norm of a matrix. In this paper, we consider when
this becomes an equality. We show that (1) if kAk = 1 and w(A ⊗ B) = w(B), then
either A has a unitary part or A is completely nonunitary and the numerical range
W (B) of B is a circular disc centered at the origin, (2) if kAk = kAkk = 1 for some
k, 1 ≤ k < ∞, then w(A) ≥ cos(π/(k + 2)), and, moreover, the equality holds if
and only if A is unitarily similar to the direct sum of the (k + 1)-by-(k + 1) Jordan
block Jk+1 and a matrix B with w(B) ≤ cos(π/(k + 2)), and (3) if B is a nonnegative
matrix with its real part (permutationally) irreducible, then w(A ⊗ B) = kAkw(B)
if and only if either pA = ∞ or nB ≤ pA < ∞ and B is permutationally similar to a
block-shift matrix
0 B1
0
. . .
. . . Bk
0
with k = nB, where pA = sup{ℓ ≥ 1 : kAℓk = kAkℓ} and nB = sup{ℓ ≥ 1 : Bℓ 6= 0}.
Keywords: numerical range; numerical radius; tensor product; Sn-matrix; nonneg-
ative matrix
AMS Subject Classifications: 15A60; 15A69; 15B48
∗Corresponding author. Email: [email protected]
1
Introduction and Preliminaries
For any n-by-n complex matrix A, its numerical range W (A) is, by definition, the
subset {hAx, xi : x ∈ Cn,kxk = 1} of the complex plane C, where h·,·i and k·k denote
the standard inner product and its associated norm in Cn, respectively. The numerical
radius w(A) of A is max{z : z ∈ W (A)}. It is known that W (A) is a nonempty
compact convex subset of C, and w(A) satisfies kAk/2 ≤ w(A) ≤ kAk, where kAk
denotes the usual operator norm of A. For other properties of the numerical range
and numerical radius, the reader may consult [7], [9, Chapter 22] or [12, Chapter 1].
The tensor product (or Kronecker product) A⊗B of an n-by-n matrix A = [aij]n
i,j=1
and an m-by-m matrix B is the (mn)-by-(mn) matrix
.
a11B · · · a1nB
...
...
an1B · · · annB
It is known that A ⊗ B and B ⊗ A are unitarily similar and kA ⊗ Bk = kAk · kBk.
Other properties of the tensor product can be found in [12, Chapter 4].
The main concern of this paper is the relations between the numerical radius of A⊗
B and those of A and B. For one direction, we have w(A⊗B) ≤ min{kAkw(B),kBkw(A)}.
This can be proven by using the unitary dilation of contractions, as to be done below.
On the other hand, we also have w(A ⊗ B) ≥ w(A)w(B). We are interested in when
these become equalities. In the present paper, we obtain various conditions, neces-
sary or sufficient, for w(A ⊗ B) = kAkw(B) to hold. The discussions on the equality
w(A ⊗ B) = w(A)w(B) will be the subject of a subsequent paper of ours.
For the ease of exposition, we introduce two indices for an n-by-n matrix A: the
power norm index pA and nilpotency index nA of A. They are defined, respectively,
by
pA = sup{k ≥ 1 : kAkk = kAkk}
2
and
sup{k ≥ 1 : Ak 6= 0n}
0
if A 6= 0n,
if A = 0n,
nA =
where 0n denotes the n-by-n zero matrix.
We start in Section 2 by proving that if kAk = 1 and w(A ⊗ B) = w(B), then
either A has a unitary part or A is completely nonunitary and W (B) is a circular
disc centered at the origin (Theorem 2.2). The proof depends on the dilation of A
to a direct sum of Sℓ-matrices with ℓ ≤ n, the Poncelet property of the numerical
ranges of matrices of the latter class, and Anderson's theorem on the circular disc
numerical range. As a by-product, we obtain a lower bound for w(A) when A satisfies
kAk = kAkk = 1 for some k, 1 ≤ k < n: w(A) ≥ cos(π/(k+2)), and determine exactly
when this bound is attained: this is the case if and only if A is unitarily similar to
Jk+1 ⊕ B, where Jk+1 is the (k + 1)-by-(k + 1) Jordan block
0 1
0
. . .
. . . 1
0
and B is a finite matrix with w(B) ≤ cos(π/(k + 2)) (Theorem 2.10). This generalizes
the classical result of Willams and Crimmins [17] for k = 1. We conclude Section
2 with a result on nilpotent contractions, namely, we prove that if A is an n-by-n
matrix with kAk = 1, then a necessary and sufficient condition for pA = nA < ∞
to hold is that A be unitarily similar to a direct sum Jk+1 ⊕ B, where k = pA and
Bk+1 = 0 (Theorem 2.13).
Finally, in Section 3, we consider B to be a nonnegative matrix with Re B (=
(B + B∗)/2) (permutationally) irreducible. We obtain in Theorem 3.1 a complete
characterization for w(A ⊗ B) = kAkw(B), namely, this is the case if and only if
either pA = ∞ or nB ≤ pA < ∞ and B is permutationally similar to a block-shift
3
matrix of the form
with k = nB.
0 B1
0
. . .
. . . Bk
0
As was mentioned before, the inequality w(A ⊗ B) ≤ kAkw(B) for n-by-n and
m-by-m matrices A and B is known. It is a consequence of [10, Theorem 3.4] because
A ⊗ B is the product of A ⊗ Im and In ⊗ B, and the latter two matrices doubly
commute, that is, A ⊗ Im commutes with both In ⊗ B and its adjoint In ⊗ B∗. Here
we give a simple proof based on the unitary dilation of contractions.
Proposition 1.1. If A and B are n-by-n and m-by-m matrices, respectively, then
w(A ⊗ B) ≤ min{kAkw(B),kBkw(A)}.
Proof. We need only prove that w(A⊗B) ≤ kAkw(B), and may assume that kAk = 1.
Then the (2n)-by-(2n) matrix
U =
A
(In − A∗A)1/2
(In − AA∗)1/2
−A∗
is unitary. Let U be unitarily similar to the diagonal matrix diag (u1, . . . , u2n), where
uj = 1 for all j. Then
w(A ⊗ B) ≤ w(U ⊗ B) = w(
2n
Xj=1
⊕ujB) = max
j
w(ujB) = w(B) = kAkw(B).
We conclude this section with some basic properties of the indices pA and nA of a
matrix A.
Proposition 1.2. Let A be an n-by-n matrix. Then
(a) 1 ≤ pA ≤ n − 1 or pA = ∞,
4
(b) pA = n − 1 if and only if A is a nonzero multiple of a Sn-matrix, and
(c) the following conditions are equivalent:
(1) pA = ∞,
(2) kAk = ρ(A),
(3) kAk = w(A),
and if kAk = 1, then the above are also equivalent to
(4) A has a unitary part.
Here ρ(A) denotes the spectral radius of A, that is, ρ(A) is the maximum modulus
of eigenvalues of A.
Recall that an n-by-n matrix A is of class Sn if it is a contraction (kAk ≤ 1),
its eigenvalues are all in D ≡ {z ∈ C : z < 1}, and rank (In − A∗A) = 1. Any
contraction A is unitarily similar to the direct sum of a unitary matrix U, called the
unitary part of A, and a completely nonunitary contraction A′, called the c.n.u. part
of A. The latter means that A′ is not unitarily similar to any direct sum with a
unitary summand.
Proof of Proposition 1.2. (a) was obtained by Pt´ak in 1960 (cf.
[15, Theorem 2.1])
and (b) was proven in [4, Theorem 3.1]. As for (c), the implication (1) ⇒ (2) is by [9,
Problem 88], (2) ⇒ (3) by the known inequalities ρ(A) ≤ w(A) ≤ kAk, (3) ⇒ (2) by
[9, Problem 218 (b)], and (2) ⇒ (1) by the inequalities ρ(A) ≤ kAkk1/k ≤ kAk for all
k ≥ 1. If kAk = ρ(A) = 1, then, letting λ be an eigenvalue of A with λ = 1, we have
the unitary similarity of A and a matrix of the form
. Since kAk = λ = 1
implies that B = 0, A is unitarily similar to [λ]⊕ C and thus has a unitary part. This
proves (2) ⇒ (4). That (4) ⇒ (2) is trivial.
0 C
λ B
Proposition 1.3. Let A be an n-by-n matrix. Then
5
(a) 0 ≤ nA ≤ n − 1 or nA = ∞,
(b) nA = n − 1 if and only if A is similar to the n-by-n Jordan block Jn,
(c) nA = ∞ if and only if A is not nilpotent, and
(d) pA ≤ nA for A 6= 0n.
We omit its easy proofs.
In the following, we use σ(A) to denote the spectrum of A, that is, σ(A) is the
set of eigenvalues of A. An n-by-n matrix A is a dilation of an m-by-m matrix B (or
B is a compression of A) if there is an n-by-m matrix V such that B = V ∗AV and
V ∗V = Im. This is equivalent to A being unitarily similar to a matrix of the form
B ∗
∗
∗
.
2 Contractions
We start with a simple condition which yields the equality w(A ⊗ B) = kAkw(B).
Lemma 2.1. If A is an n-by-n matrix with pA = ∞, then w(A ⊗ B) = kAkw(B)
for any m-by-m matrix B. In particular, this is the case for A a contraction with a
unitary part.
Proof. Since pA = ∞ implies, by Proposition 1.2 (c), that kAk = w(A). If λ is a
number in W (A) with λ = w(A), then λ = kAk. Since A is unitarily similar to a
matrix of the form
.
It follows that kAkw(B) = w(λB) ≤ w(A ⊗ B). On the other hand, we also have
w(A ⊗ B) ≤ kAkw(B) by Proposition 1.1. Thus w(A ⊗ B) = kAkw(B) holds.
, we have the unitary similarity of A ⊗ B and
λB ∗
∗
∗
λ ∗
∗
∗
The next theorem is one of the main results of this section. It gives a necessary
condition for the equality w(A ⊗ B) = kAkw(B).
6
Theorem 2.2. Let A and B be n-by-n and m-by-m matrices, respectively. If kAk = 1
and w(A ⊗ B) = w(B), then either A has a unitary part or A is c.n.u. and W (B) is
a circular disc centered at the origin.
We first prove this for the case when A is an Sn-matrix. The numerical ranges of
such matrices are known to have the Poncelet property, namely, if A is of class Sn,
then, for any point λ on the unit circle ∂D, there is a unique (up to unitary similarity)
(n + 1)-by-(n + 1) unitary dilation U of A such that λ is an eigenvalue of U and each
edge of the (n + 1)-gon ∂W (U) intersects W (A) at exactly one point (cf. [2, Theorem
2.1 and Lemma 2.2]).
Lemma 2.3. Let A be an Sn-matrix and B an m-by-m matrix. If w(A⊗ B) = w(B),
then W (B) is a circular disc centered at the origin.
Proof. Let U1, . . . , Um+1 be (n + 1)-by-(n + 1) unitary dilations of A with σ(Ui) ∩
σ(Uj) = ∅ for all i and j, 1 ≤ i 6= j ≤ m + 1. We may assume that Uj =
diag (λ1j, . . . , λn+1,j) for each j, where λij = 1 for all i and j. Let Vj be an (n + 1)-
by-n matrix such that A = V ∗
j Vj = In for each j. Since kAk = 1 and
j UjVj and V ∗
w(A ⊗ λB) = w(A ⊗ B) = w(B) = w(λB)
for any λ, λ = 1, we may further assume that w(B) is in W (A ⊗ B). Let x be a
unit vector in Cn ⊗ Cm such that h(A ⊗ B)x, xi = w(B). We decompose (Vj ⊗ Im)x
as y1j ⊕ · · · ⊕ yn+1,j with yij, 1 ≤ i ≤ n + 1, in Cm for each j. Then
w(B) = h(A ⊗ B)x, xi
n+1
=
= h(Uj ⊗ B)(Vj ⊗ Im)x, (Vj ⊗ Im)xi
= h(λ1jB ⊕ · · · ⊕ λn+1,jB)(y1j ⊕ · · · ⊕ yn+1,j), y1j ⊕ · · · ⊕ yn+1,ji
Xi=1
hλijByij, yiji
Xi=1
hByij, yiji.
≤
n+1
7
Letting ηij = hB(yij/kyijk), yij/kyijki for each yij 6= 0, we obtain
w(B) = Xyij 6=0
λijkyijk2ηij ≤ Xyij 6=0
kyijk2ηij ≤ Xyij 6=0
kyijk2w(B) = w(B)
since
kyijk2 = k(Vj ⊗ Im)xk2 = kxk2 = 1.
Xi
Thus we have equalities throughout the above sequence, which yields that w(B) =
λijηij for yij 6= 0. SincePi kyijk2 = 1, this must hold for at least one i, say, ij. Hence
λijjw(B) = ηijj is in ∂W (B) for each j. Note that such λij jw(B)'s, 1 ≤ j ≤ m + 1,
are distinct from each other by our assumption on the disjointness of the spectra of
the Uj's. This shows that the boundary of W (B) and the circle z = w(B) intersect
at at least m + 1 points. Since W (B) is contained in {z ∈ C : z ≤ w(B)}, we apply
[3, Theorem] or [20]) to infer that W (B) = {z ∈ C : z ≤
Anderson's theorem (cf.
w(B)}.
Proof of Theorem 2.2. We assume that A is c.n.u. Then A can be dilated to the direct
sum A′⊕· · ·⊕ A′ of rank (In− A∗A) many copies of some Sℓ-matrix A′ with ℓ ≤ n (cf.
[18, Theorem 1.4] or [21, Lemma 3 (a)]). Hence A⊗ B dilates to (A′⊕· · ·⊕ A′)⊗ B =
(A′ ⊗ B) ⊕ · · · ⊕ (A′ ⊗ B). We have
w(B) = w(A⊗ B) ≤ w((A′ ⊗ B)⊕· · ·⊕ (A′ ⊗ B)) = w(A′ ⊗ B) ≤ kA′kw(B) = w(B).
Thus w(A′ ⊗ B) = w(B). It follows from Lemma 2.3 that W (B) is a circular disc
centered at the origin.
An easy consequence of Theorem 2.2 is that the converse of Lemma 2.1 is also
true.
Corollary 2.4. For an n-by-n matrix A, the equality w(A ⊗ B) = kAkw(B) holds
for all matrices B if and only if pA = ∞.
8
Proof. For the necessity, assume that kAk = 1 and let B be any matrix with its
numerical range not a circular disc centered at the origin. Theorem 2.2 yields that A
has a unitary part. Then pA = ∞ follows immediately.
In Theorem 2.2, if B is the Jordan block Jm, then we have the following charac-
terizations for w(A ⊗ B) = kAkw(B).
Theorem 2.5. Let A be an n-by-n matrix with kAk = 1. Then the following condi-
tions are equivalent:
(a) W (A ⊗ Jm) = W (Jm),
(b) w(A ⊗ Jm) = w(Jm),
(c) A ⊗ Jm is unitarily similar to Jm ⊕ B for some matrix B with w(B) ≤ w(Jm),
and
(d) kAm−1k = 1.
If, in addition, n = m, then the above conditions are also equivalent to
(e) either A has a unitary part or A is of class Sn, and
(f) pA = ∞ or n − 1.
Note that W (Jm) = {z ∈ C : z ≤ cos(π/(m + 1))} (cf. [8, Proposition 1]).
Proof of Theorem 2.5. The implication (a) ⇒ (b) is trivial. To prove (b) ⇒ (c), note
that (A⊗Jm)m = Am⊗J m
m = 0nm and kA⊗Jmk = kAkkJmk = 1. If x is a unit vector
in Cn ⊗ Cm such that h(A ⊗ Jm)x, xi = w(A ⊗ Jm), then w(A ⊗ Jm) = w(Jm) =
cos(π/(m + 1)) implies that the subspace K of Cn ⊗ Cm generated by the vectors
x, (A⊗ Jm)x, . . . , (A⊗ Jm)m−1x is reducing for A⊗ Jm, and the restriction of A⊗ Jm
[8, Theorem 1 (2)]). Hence A ⊗ Jm is unitarily
to K is unitarily similar to Jm (cf.
9
similar to Jm ⊕ B, where B is the restriction of A ⊗ Jm to K ⊥. We obviously have
w(B) ≤ w(A ⊗ Jm) = w(Jm).
For (c) ⇒ (d), note that Am−1 ⊗ J m−1
m
is unitarily similar to J m−1
m ⊕ Bm−1 under
(c). Hence
kAm−1k = kAm−1 ⊗ J m−1
m k = kJ m−1
m ⊕ Bm−1k = max{kJ m−1
m k,kBm−1k} = 1.
To prove (d) ⇒ (c), let x be a unit vector in Cn such that kAm−1xk = 1. Then
kAm−jxk = 1 for all j, 1 ≤ j ≤ m. Let {e1, . . . , em} be the standard basis for Cm,
let xj = Am−jx ⊗ ej, 1 ≤ j ≤ m, and let K be the subspace of Cn ⊗ Cm generated
by x1, . . . , xm. Then (A ⊗ Jm)x1 = 0 and (A ⊗ Jm)xj = xj−1 for 2 ≤ j ≤ m. Since
{x1, . . . , xm} is an orthonormal basis of K, this shows that (A ⊗ Jm)K ⊆ K and the
restriction of A ⊗ Jm to K is unitarily similar to Jm. On the other hand, it follows
from kA ⊗ Jmk = kAkkJmk = 1 and
(A ⊗ Jm)∗xm = (A∗ ⊗ J ∗
m)(x ⊗ em) = (A∗x) ⊗ (J ∗
mem) = (A∗x) ⊗ 0 = 0
that K is reducing for A⊗ Jm, and hence A⊗ Jm is unitarily similar to Jm⊕ B, where
B is the restriction of A ⊗ Jm to K ⊥. Obviously, we have
w(B) ≤ w(A ⊗ Jm) ≤ kAkw(Jm) = w(Jm).
To prove (c) ⇒ (a), note that the unitary similarity of Jm and eiθJm for all real
θ implies the same for A ⊗ Jm and eiθ(A ⊗ Jm). Thus W (A ⊗ Jm) is a circular
disc centered at the origin. (c) implies that w(A ⊗ Jm) = w(Jm), which means that
the radii of the two circular discs W (A ⊗ Jm) and W (Jm) are equal. Therefore,
W (A ⊗ Jm) = W (Jm) holds.
Now assume that n = m and that kAn−1k = 1. If kAnk = 1, then pA = ∞ and
hence A has a unitary part by Proposition 1.2 (a) and (c). On the other hand, if
kAnk < 1, then A is of class Sn by [4, Theorem 3.1]. This shows that (d) ⇒ (e).
Next, if (e) is true, then pA = ∞ or n− 1 depending on whether A has a unitary part
10
or A is of class Sn (cf.
[4, Theorem 3.1] for the latter). This proves (f). Finally, if
pA = ∞, then kAkk = 1 for all k ≥ 1, and, in particular, kAn−1k = 1. On the other
hand, if pA = n − 1, then kAn−1k = kAkn−1 = 1. This proves (f) ⇒ (d).
The next proposition gives a characterization of w(A⊗ B) = kAkw(B) when B is
of class Sm.
Proposition 2.6. Let A be an n-by-n matrix with kAk = 1, and B be an Sm-matrix.
Then w(A ⊗ B) = w(B) if and only if either A has a unitary part or A is c.n.u.,
kAm−1k = 1 and B is unitarily similar to Jm.
Its proof depends on a special property of Sn-matrices. The following lemma is
from [19, Lemma 5]. Here we give a shorter geometric proof.
Lemma 2.7. Let A be an Sn-matrix. Then W (A) is a circular disc centered at the
origin if and only if A is unitarily similar to Jn.
Proof. If W (A) is as asserted, then the Poncelet property of W (A) says that it is
circumscribed by (n + 1)-gons with vertices on the unit circle. As the circular disc
{z ∈ C : z ≤ cos(π/(n + 1))}(= W (Jn)) is circumscribed by any regular (n + 1)-gon
on the unit circle, if the radius of W (A) is not equal to cos(π/(n + 1)), then we infer
from a geometrical consideration that W (A) cannot have the Poncelet property. Thus
W (A) must equal W (Jn). The unitary similarity of A and Jn then follows from [2,
Theorem 3.2]. The converse is trivial.
Proof of Proposition 2.6. If w(A ⊗ B) = w(B), then, by Theorem 2.2, either A has a
unitary part or A is c.n.u. and W (B) is a circular disc centered at the origin. In the
latter case, Lemma 2.7 yields the unitary similarity of B and Jm, and then Theorem
2.5 gives kAm−1k = 1. The converse also follows from Theorem 2.5.
11
Note that, under the conditions of Proposition 2.6, if A is c.n.u., then we auto-
matically have m ≤ n. This is because if, otherwise, m > n, then kAm−1k = 1 yields,
by Proposition 1.2 (a) and (c), that A has a unitary part.
A specific example of the results obtained so far is in the next proposition.
Proposition 2.8. Let n and m be positive integers. Then W (Jn ⊗ Jm) = W (Jℓ),
where ℓ = min{n, m}, and thus w(Jn ⊗ Jm) = min{w(Jn), w(Jm)}.
Proof. Assume that m ≤ n. Since the principal submatrix of Jn ⊗ Jm formed by
its rows and columns numbered 1, m + 2, 2m + 3, . . ., and (m − 1)m + m is Jm, we
have that Jn ⊗ Jm is a dilation of Jm. Thus w(Jm) ≤ w(Jn ⊗ Jm). The reversed
inequality w(Jn ⊗ Jm) ≤ kJnkw(Jm) = w(Jm) is by Proposition 1.1. Therefore,
w(Jn ⊗ Jm) = w(Jm) holds. As was seen in the proof of (c) ⇒ (a) in Theorem 2.5,
W (Jn ⊗ Jm) is a circular disc centered at the origin. Thus the equality of w(Jn ⊗ Jm)
and w(Jm) implies that of W (Jn ⊗ Jm) and W (Jm).
Besides Sn-matrices, another generalization of the Jordan blocks is the companion
matrices. Recall that a companion matrix is one of the form
0
1
0
1
·
−an −an−1
·
·
·
·
·
·
·
0
1
· −a2 −a1
,
12
whose characteristic and minimal polynomials are both equal to zn +Pn
The numerical ranges of such matrices have been studied in [5, 6, 1].
j=1 ajzn−j.
Proposition 2.9. Let A be an n-by-n (n ≥ 2) companion matrix. Then the following
conditions are equivalent:
(a) w(A ⊗ A) = kAkw(A),
(b) A is unitary, A = Jn, or A is unitarily similar to a direct sum [aωj
n]⊕ B, where
a > 1, ωn = ei(2π/n), 0 ≤ j ≤ n − 1, and B is an Sn−1-matrix with eigenvalues
(1/a)ωk
n, 0 ≤ k ≤ n − 1 and k 6= j, and
(c) pA = nA = ∞ or n − 1.
Proof. To prove (a) ⇒ (b), let A′ = A/kAk. Then (a) gives w(A′ ⊗ A′) = w(A′).
By Theorem 2.2, either A′ has a unitary part or it is c.n.u. with numerical range
a circular disc centered at the origin. In the former case, either A is normal or is
unitarily similar to a matrix of the form [aωj
n] ⊕ B, where a = kAk ≥ 1 and B is of
size n − 1 with eigenvalues (1/a)ωk
[5, Theorem 1.1
and Corollary 1.3]). If A is normal or a = 1, then A is unitary by [5, Corollary 1.2].
Hence we may assume that a > 1. Thus the eigenvalues of B are all contained in D.
Moreover, by [1, Theorem 2.1], we have rank (In−1 − B∗B) = 1. These two together
imply, by way of the singular value decomposition of B, that kBk = 1. Hence B is
of class Sn−1. On the other hand, if it is the latter case, then W (A) is also a circular
n, 0 ≤ k ≤ n − 1 and k 6= j (cf.
disc centered at the origin. Therefore, A = Jn by [5, Theorem 2.9]. This proves (b).
For (b) ⇒ (c), if A is unitary (resp., A = Jn), then, obviously, pA = nA = ∞
(resp., pA = nA = n − 1). On the other hand, if A is unitarily similar to the asserted
[aωj
n] ⊕ B, then kAk = max{a,kBk} = a = ρ(A). Thus pA = nA = ∞ by
Proposition 1.2 (c) and 1.3.
Finally, for (c) ⇒ (a), if pA = nA = ∞, then (a) is a consequence of Lemma 2.1.
On the other hand, if pA = nA = n − 1, then An = 0n. This implies that A = Jn and
thus (a) holds by Proposition 2.8.
13
The next theorem is a consequence of Theorem 2.5. It gives a lower bound, in
terms of pA, for w(A) when A is an n-by-n matrix with kAk = 1.
Theorem 2.10. If A is an n-by-n matrix with kAk = kAkk = 1 for some k ≥ 1,
then w(A) ≥ cos(π/(k + 2)). Moreover, in this case, the following conditions are
equivalent:
(a) w(A) = cos(π/(k + 2)),
(b) A is unitarily similar to Jk+1 ⊕ B, where B is a finite matrix with w(B) ≤
cos(π/(k + 2)), and
(c) W (A) = {z ∈ C : z ≤ cos(π/(k + 2))}.
For the proof of (a) ⇒ (b), we need the following lemma.
Lemma 2.11. Let
A =
0
a1
0
. . .
. . .
an−2
0
an−1
a
and B =
0
a1
0
. . .
. . .
an−2
0
be n-by-n and (n− 1)-by-(n− 1) matrices, respectively, where n ≥ 2 and aj is nonzero
for all j. Then w(A) > w(B).
Proof. We prove this by induction on n. If n = 2, then A =
in which case we obviously have w(A) > 0 = w(B). Assume now that the assertion
is true for the matrix A of size at most n − 1 (n ≥ 3), and let A and B be of the
above form. By considering eiθA for a suitable real θ instead of A, we may assume
and B = [0],
0 a1
0
a
14
that w(A) equals the largest eigenvalue of Re A. Let
C =
0
a1
0
. . .
. . .
an−3
0
,
and let p(z), q(z) and r(z) be the characteristic polynomials of Re A, Re B and Re C,
respectively. We expand the determinant of
z
−a1/2
−a1/2
z
. . .
. . .
. . .
. . .
. . .
z
−an−1/2
−an−1/2 z − Re a
by minors on its last row to obtain p(z) = (z − Re a)q(z) − (an−12/4)r(z). Let
α, β and γ be the largest eigenvalues of Re A, Re B and Re C, respectively. Then
α = w(A), β = w(B) and γ = w(C). Since Re B (resp., Re C) is a principal submatrix
of Re A (resp., Re B), we have β ≤ α (resp., γ ≤ β). Assume that α = β. Then the
above equation yields
0 = p(α) = (α − Re a)q(β) −
1
4an−12γ(β) = −
1
4an−12γ(β).
Since an−1 6= 0 and β is larger than or equal to all eigenvalues of Re C, we infer from
γ(β) = 0 that β = γ or w(B) = w(C). This contradicts our induction hypothesis for
B and C. Hence we must have α > β or w(A) > w(B).
Proof of Theorem 2.10. By Theorem 2.5, the assumption kAk = kAkk = 1 implies
that w(A ⊗ Jk+1) = w(Jk+1). Hence
w(A) = kJk+1kw(A) ≥ w(A ⊗ Jk+1) = w(Jk+1) = cos
π
k + 2
as asserted.
15
We now prove the equivalence of (a), (b) and (c). The implications (b) ⇒ (c)
and (c) ⇒ (a) are trivial. To prove (a) ⇒ (b), let x be a unit vector in Cn such that
kAkxk = 1. Then kAjxk = 1 for all j, 0 ≤ j ≤ k. We now check that Ak+1x = 0.
Assuming otherwise that kAk+1xk > 0, let ut = [ut1 . . . ut,k+2]T in Ck+2 ⊗ Cn, where
Ak+1x
if j = 1,
sin
(j − 1)π
k + 2
Ak−j+2x
if j = 2, . . . , k + 2
√1 − t2
kAk+1xk
tq 2
utj =
v ≡r 2
k + 2
for any t, 0 < t < 1. Note that
k + 2(cid:20)sin
π
k + 2
sin
2π
k + 2
. . .
sin
(k + 1)π
k + 2 (cid:21)T
is a unit vector in Ck+1 with hJk+1v, vi = cos(π/(k + 2)) (cf. [8, Proposition 1 (3)]).
Hence kutk = ((1 − t2) + t2kvk2)1/2 = 1, and
h(Jk+2 ⊗ A)ut, uti = t√1 − t2r 2
Xj=1
= t√1 − t2r 2
= t√1 − t2r 2
π
k + 2kAk+1xk
jπ
(j + 1)π
k + 2 kAk−j+1xk2
π
k + 2kAk+1xk + t2hJk+1v, vi
π
k + 2kAk+1xk + t2 cos
k + 2
k
k + 2
k + 2
k + 2
k + 2
k + 2
+t2
sin
sin
sin
π
.
2
sin
sin
To reach a contradiction, we need to find some t0, 0 < t0 < 1, such that h(Jk+2 ⊗
A)ut0, ut0i > cos(π/(k + 2)). This is the same as
t0q1 − t2
0r 2
k + 2
sin
π
k + 2kAk+1xk > (1 − t2
0) cos
π
k + 2
or
t0
>rk + 2
2
cot π
k+2
kAk+1xk
.
p1 − t2
0
Since limt→1− t/√1 − t2 = ∞, the existence of such a t0 is guaranteed. On the other
hand, we also have
h(Jk+2 ⊗ A)ut0, ut0i ≤ w(Jk+2 ⊗ A) ≤ kJk+2kw(A) = w(A) = cos
π
k + 2
,
16
hence a contradiction. Thus we must have Ak+1x = 0. Let K be the subspace of Cn
generated by x, Ax, . . . , Akx. Then AK ⊆ K. If A′ is the restriction of A to K, then
A′k+1 = 0 and kA′jxk = kAjxk = 1 for all j, 0 ≤ j ≤ k. Hence kA′jk = 1 for all
such j's. Together with A′k+1 = 0, this says that pA′ = k and thus dim K = k + 1
by Proposition 1.2 (a). Therefore, A′ is unitarily similar to a matrix of the form
i,j=1 with aij = 0 for all i ≥ j. Since 1 = kA′kk = a12 · · · ak,k+1, we infer that
[aij]k+1
a12 = · · · = ak,k+1 = 1, and thus all the other aij's are zero. Therefore, [aij]k+1
i,j=1,
and hence A′, is unitarily similar to Jk+1. Then A is unitarily similar to a matrix of
the form
Jk+1
0
0
b1 · · · bn−k−1
∗
c1
. . .
∗
cn−k−1
.
To show that all the bj's are zero, we appeal to Lemma 2.11.
1 ≤ j ≤ n − k − 1, consider the (k + 2)-by-(k + 2) matrix
Indeed, for each j,
Aj =
0
...
0
bj
cj
Jk+1
0
.
If bj 6= 0, then w(Aj) > w(Jk+1) = cos(π/(k + 2)) by Lemma 2.11, which contradicts
w(Aj) ≤ w(A) = cos(π/(k + 2)). This proves (a) ⇒ (b).
Theorem 2.10 generalizes the classical result of Williams and Crimmins [17] for
k = 1. The following corollary is for k = n − 1. Part of it has been proven in [19]:
the equivalence of (b) and (c) is in [19, Theorem 1] and that of (b) and (d) in [19, p.
352].
17
Corollary 2.12. The following conditions are equivalent for an n-by-n matrix A with
kAk = 1:
(a) kAn−1k = 1 and w(A) = cos(π/(n + 1)),
(b) A is unitarily similar to Jn,
(c) W (A) = {z ∈ C : z ≤ cos(π/(n + 1))},
(d) kAn−1k = 1 and An = 0n, and
(e) pA = nA = n − 1.
Proof. The equivalence of (a) and (b) is by Theorem 2.10. The other implications
are either in [19] or trivial.
Note that, in the preceding corollary, the conditions that kAk = 1 and w(A) =
cos(π/(n + 1)) for an n-by-n matrix A are not sufficient to guarantee that A be
unitarily similar to Jn. One example is A = Jn−1 ⊕ [cos(π/(n + 1))].
We end this section with a characterization of matrices A satisfying pA = nA.
This is related to the previous results.
Theorem 2.13. Let A be an n-by-n matrix with kAk = 1. Then
(a) A satisfies pA = nA (≤ ∞) if and only if either it has a unitary part or is
unitarily similar to a direct sum Jk+1 ⊕ B, where k = pA < ∞ and Bk+1 =
0n−k−1, and
(b) if pA = nA (≤ ∞), then w(A ⊗ A) = w(A) holds, but not conversely.
Proof. (a) For the necessity, we may assume, in view of Proposition 1.2 (c), that
k ≡ pA = nA < ∞ and prove that A is unitarily similar to the asserted direct sum.
18
Since Ak+1 = 0n, A is unitarily similar to a block matrix A′ of the form [Aij]k+1
Aij = 0 for 1 ≤ j ≤ i ≤ k + 1. Hence
0 · · ·
0
0 Qk
i=1 Ai,i+1
i,j=1 with
0
...
0
.
A′k =
. . .
Since kA′kk = kAkk = kAkk = 1, we have kQk
i=1 Ai,i+1k = 1. Let x be a unit vector
such that k(Qk
i=1 Ai,i+1)xk = 1. Then k(Qk
i=j Ai,i+1)xk = 1 for all j, 1 ≤ j ≤ k.
Let {e1, . . . , ek+1} be the standard basis for Ck+1, and let xj = ej ⊗ (Qk
i=j Ai,i+1)x if
1 ≤ j ≤ k, and xk+1 = ek+1 ⊗ x. Then x1, . . . , xk+1 are orthonormal vectors in Cn,
and A′x1 = 0 and A′xj = xj−1 for 2 ≤ j ≤ k + 1. Thus if K is the subspace generated
by x1, . . . , xk+1, then dim K = k + 1, A′K ⊆ K, and the restriction of A′ to K is
unitarily similar to Jk+1. We infer from kA′k = 1 and A′∗xk+1 = 0 that K reduces
A′, and thus A′ is unitarily similar to Jk+1 ⊕ B with Bk+1 = 0.
For the converse, if A has a unitary part, then pA = nA = ∞ by Proposition
1.2 (c). On the other hand, if A is unitarily similar to Jk+1 ⊕ B with the asserted
properties, then Ak+1 = 0 implies that pA ≤ nA ≤ k. But
kAkk = kJ k
k+1 ⊕ Bkk = max{kJ k
k+1k,kBkk} = 1 = kAkk
and kAk+1k = 0 < 1 = kAkk+1 together yield pA = nA = k.
(b) If A has a unitary part, then w(A⊗A) = w(A) by Proposition 2.1. On the other
hand, if A is unitarily similar to Jk+1 ⊕ B as in (a), then A⊗ A is unitarily similar to
(Jk+1⊗Jk+1)⊕(Jk+1⊗B)⊕(B⊗Jk+1)⊕(B⊗B). Note that w(Jk+1⊗Jk+1) = w(Jk+1)
by Proposition 2.8, and
(1)
w(Jk+1 ⊗ B) = w(B ⊗ Jk+1) ≤ kJk+1kw(B) = w(B)
by Proposition 1.1. Since Bk+1 = 0 and kBk ≤ 1, [21, Lemma 3 (a)] implies that B
can be dilated to the direct sum of rank (I − B∗B) copies of Jm for some m ≤ k + 1.
19
Thus w(B) ≤ w(Jm) ≤ w(Jk+1). Combined with (1), this yields w(Jk+1 ⊗ B) ≤
w(Jk+1). Also,
w(B ⊗ B) ≤ kBkw(B) ≤ w(B) ≤ w(Jk+1).
Therefore,
w(A ⊗ B) = max{w(Jk+1 ⊗ Jk+1), w(Jk+1 ⊗ B), w(B ⊗ B)}
= w(Jk+1)
= max{w(Jk+1), w(B)}
= w(A).
That w(A ⊗ A) = w(A) does not imply pA = nA is seen by A = J2 ⊕ [a], where
0 < a ≤ 1/2, in which case, kAk = 1 and w(A ⊗ A) = w(A) = 1/2, but pA = 1 and
nA = ∞.
The final result of this section is conditions for a matrix A with pA = nA so that
it be unitarily similar to a block-shift matrix
(2)
A′ =
with kA1 · · · Akk = kAk.
0 A1
0
. . .
. . . Ak
0
Proposition 2.14. Let A be an n-by-n matrix with pA = nA ≡ k < ∞. If either
(a) k = 1, n − 2 or n − 1, or (b) n = 2, 3, 4 or 5, then A is unitarily similar to the
block-shift matrix A′ in (2) with kA1 · · · Akk = kAk.
Proof. We may assume that kAk = 1.
(a) If k = nA = 1, then A2 = 0n. Hence A is unitarily similar to a block-shift
matrix of the form
0 A1
0
0
with kA1k = kAk.
20
If k = pA = nA = n − 1 (resp., n − 2), then Theorem 2.13 (a) implies that A is
unitarily similar to Jn (resp., Jn−1 ⊕ [0]). The latter matrix plays the role of A′ with
k = n − 1 (resp., n − 2) and A1 = · · · = An−1 = [1] (resp., A1 = · · · = An−3 = [1] and
An−2 = [1 0]).
(b) In light of (a), we need only prove for n = 5 and k = 2. Invoking Theorem
0
2.13 to obtain the unitary similarity of A and J3⊕
matrix is permutationally similar to a block-shift matrix A′ with k = 2, A1 =
and A2 =
0
. We obviously have kA1A2k = k
0
k = 1 = kAk.
0
1
b
0
, where b ≤ 1. The latter
b
1
0
0
We remark that the preceding proposition fails for n = 6 and k = 2. Here is an
example. Let A = J3 ⊕ B, where
1
B = b
0 1 1
0 0 1
0 0 0
with b = q2/(3 + √5). Then kA2k = 1 = kAk2 and A3 = 06. This shows that
pA = nA = 2. Since w(B) = 2b > √2/2 = w(J3) and w(B) is not a circular disc
centered at the origin (cf. [13, Theorem 4.1 (2)]), we infer that nor is W (A) (= the
convex hull of W (J3) ∪ W (B)). This implies that A cannot be unitarily similar to a
block-shift matrix.
3 Nonnegative Matrices
Recall that a matrix A = [aij]n
i,j=1 is nonnegative (resp., positive), denoted by A < 0
(resp., A ≻ 0), if aij ≥ 0 (resp., aij > 0) for all i and j. Two n-by-n matrices A and
B are permutationally similar if there is an n-by-n permutation matrix P (one with
21
each row and column has exactly one 1 and all other entries 0) such that P T AP = B.
A is said to be (permutationally) reducible if either A is the 1-by-1 zero matrix or
n ≥ 2 and it is permutationally similar to a matrix of the form
and D are square matrices; otherwise, it is (permutationally) irreducible. It is known
, where B
B C
0 D
that if A is nonnegative with Re A irreducible, then it is permutationally similar to a
block-shift matrix if and only if its numerical range is a circular disc centered at the
[16, Theorem 1 (a)⇔(r)]). Other properties of nonnegative matrices can
origin (cf.
be found in [11, Section 6.2 and Chapter 8].
The main result of this section is the following theorem, which essentially gener-
alizes Theorem 2.5.
Theorem 3.1. Let A be an n-by-n matrix and B an m-by-m nonnegative matrix with
Re B irreducible. Then the following conditions are equivalent:
(a) w(A ⊗ B) = kAkw(B),
(b) either pA = ∞ or nB ≤ pA < ∞ and W (B) is a circular disc centered at the
origin, and
(c) either pA = ∞ or nB ≤ pA < ∞ and B is permutationally similar to a block-
shift matrix
0 B1
0
. . .
. . . Bk
0
with k = nB.
For its proof, we need the following two lemmas.
Lemma 3.2. Let A = [aij]n
i,j=1 be a nonnegative matrix. Then the following hold:
22
(a) The index nA is finite if and only if there is no sequence of indices i0, i1, . . . , ik−1, ik (k ≥
1) with i0 = ik such that ai0i1, . . . , aik−1ik are all nonzero.
In particular, we
have nA = sup{k ≥ 1 : there are distinct ij, 0 ≤ j ≤ k, such that aij ij+1 6=
0 for all j}.
(b) nA = ∞ if and only if there is a k ≥ 1 such that some diagonal entry of Ak is
nonzero.
(c) If aii 6= 0 for some i, 1 ≤ i ≤ n, then nA = ∞.
(d) If A is irreducible, then nA = ∞.
(e) If A is the block-shift matrix
0n1 A1
0n2
. . .
. . . Ak
0nk+1
on Cn = Cn1 ⊕ · · · ⊕ Cnk+1
and Re A is irreducible, then k = nA.
Proof. (a) Assume first that the indices i0, i1, . . . , ik−1, ik = i0 (k ≥ 1) are such that
ai0i1, . . . , aik−1ik 6= 0.
[11, Theorem 6.2.16] says that this is the case if and only if
(Ak)i0i0, the (i0, i0)-entry of Ak, is nonzero. Hence Ak 6= 0n. Similarly, considering
the sequence i0, . . . , ik, i1, . . . , ik, . . . , i1, . . . , ik of ℓk + 1 indices for any ℓ ≥ 1, we also
obtain Aℓk 6= 0n. It follows that nA = ∞. Conversely, assume that nA = ∞. Then
Ak 6= 0n for some k ≥ n.
[11, Theorem 6.2.16] yields that, for some i and j, there
are indices i0 = i, i1, . . . , ik−1, ik = j such that ai0i1, . . . , aik−1ik are all nonzero. By
the pigeonhole principle, we infer that is = it for some s and t, 0 ≤ s < t ≤ k. Then
is, . . . , it are such that is = it and aisis+1, . . . , ait−1it 6= 0. This proves the converse.
The expression for nA is an easy consequence of [11, Theorem 6.2.16] and the above
arguments. So are (b) and (c).
23
(d) Note that the irreducibility of A is equivalent to the existence, for every distinct
pair i and j, of indices i0 = i, i1, . . . , ik−1, ik = j (k ≥ 1) such that ai0i1, . . . , aik−1ik
are all nonzero. Combining such indices from i to j with those from j to i yields one
from i to i with the corresponding entries nonzero. Thus nA = ∞ by [11, Theorem
6.2.16] and (b).
(e) Since Ak+1 = 0n, we have nA ≤ k. If nA < k, then Ak = 0n, which implies that
A1 · · · Ak = 0. If there are any nonzero ai0i1, ai1i2, . . . , aik−1ik , where (Pℓ
j=1 nj) + 1 ≤
iℓ ≤ Pℓ+1
j=1 nj for 0 ≤ ℓ ≤ k, then the (i0, nk+1 − (n − ik))-entry of A1 · · · Ak, being
larger than or equal to Qk−1
j=0 aijij+1, is nonzero, which contradicts the zeroness of
the product A1 · · · Ak. Thus no such nonzero sequence exists. This results in the
reducibility of Re A, a contradiction. Hence we must have nA = k.
We remark that the conditions in the preceding lemma can all be expressed equiv-
alently in terms of the directed graph associated with the matrix A (cf. [11, Section
6.2]).
Lemma 3.3. Let A and B be n-by-n and m-by-m matrices, respectively. If B is
unitarily similar to a block-shift matrix
(3)
on Cm = Cm1 ⊕ · · · ⊕ Cmk+1
0m1 B1
0m2
. . .
. . . Bk
0mk+1
with k ≤ pA ≤ ∞, then w(A ⊗ B) = kAkw(B).
Proof. We may assume that kAk = 1 and B is equal to the block-shift matrix (3).
Since k ≤ pA ≤ ∞, we have kAkk = kAkk = 1. Let x be a unit vector in Cn such that
kAkxk = 1, and let y = [y1 . . . yk+1]T , where yj is in Cmj , 1 ≤ j ≤ k + 1, be a unit
vector in Cm such that hBy, yi = w(B). Let u = [y1⊗Akx y2⊗Ak−1x . . . yk+1⊗x]T .
24
Then u is a vector in Cm ⊗ Cn with
kuk = (
= (
k+1
k+1
Xj=1
Xj=1
kyj ⊗ Ak−j+1xk2)1/2 = (
k+1
Xj=1
kyjk2kAk−j+1xk2)1/2
kyjk2)1/2 = kyk = 1.
Moreover, we have
h(B ⊗ A)u, ui
*
0m1n B1 ⊗ A
0m2n
=
. . .
. . . Bk ⊗ A
0mk+1n
y1 ⊗ Akx
y2 ⊗ Ak−1x
...
yk+1 ⊗ x
,
y1 ⊗ Akx
y2 ⊗ Ak−1x
...
yk+1 ⊗ x
+
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
k
k
Xj=1
Xj=1
Xj=1
k
=
=
h(Bjyj+1) ⊗ (Ak−j+1x), yj ⊗ (Ak−j+1x)i
hBjyj+1, yjikAk−j+1xk2
hBjyj+1, yji
=
= hBy, yi = w(B).
This shows that w(B) ≤ w(B ⊗ A) = w(A ⊗ B). But w(A ⊗ B) ≤ kAkw(B) = w(B)
always holds by Proposition 1.1. Hence w(A ⊗ B) = w(B) as asserted.
We are now ready to prove Theorem 3.1.
Proof of Theorem 3.1. For (a) ⇒ (b), We assume that kAk = 1 and A is c.n.u. In view
of Theorem 2.2 and Proposition 1.2 (c), we need only check that w(A ⊗ B) = w(B)
implies nB ≤ pA (< ∞). Let B = [bij]m
i,j=1, and let x be a unit vector in Cm ⊗ Cn
such that w(B ⊗ A) = h(B ⊗ A)x, xi. If x = [x1 . . . xm]T , where xj is in Cn for
25
1 ≤ j ≤ m, then
w(B) = w(B ⊗ A) = h[bijA]x, xi
(4)
(5)
(6)
bijhAxj, xii
bijkAxjkkxik
bijkxjkkxik
≤Xi,j
≤Xi,j
≤ kAkXi,j
≤ hBx′, x′i
≤ w(B),
where x′ = [kx1k . . . kxmk]T is a unit vector in Cm. This shows that the above
inequalities are equalities throughout. Since B < 0 and Re B is irreducible, there is a
unique unit vector y in Cm with y ≻ 0 such that hBy, yi = w(B) (cf. [14, Proposition
3.3]). The equality in (6) yields that x′ = y and thus xj 6= 0 for all j. Also, the
equalities in (4) and (5) imply that hAxj, xii = kAxjkkxik = kxjkkxik for all those
bij's with bij > 0. Thus Axj = λijxi for some λij satisfying λij = kxjk/kxik.
Assume first that k ≡ nB < ∞. Thus Bk 6= 0m. By Lemma 3.2 (a), there are
distinct indices i0, . . . , ik such that bi0i1, . . . , bik−1ik > 0. It thus follows from above
that Axij = λij−1ij xij−1 for 1 ≤ j ≤ k. Hence Akxik = (Qk
)kxi0k = kxikk,
kAkxikk = (
k
Yj=1
kxijk
kxij−1k
j=1 λij−1ij )xi0. Since
we obtain kAkk = 1 or pA ≥ k = nB. On the other hand, if nB = ∞, then the same
arguments as above with k arbitrarily large yield that pA = ∞, which contradicts our
assumption that A is c.n.u. This proves (a) ⇒ (b).
That (b) ⇔ (c) is a consequence of [16, Theorem 1 (a)⇔(r)], and (c) ⇒ (a) is by
Lemma 3.2 (e) and Lemma 3.3.
Note that, in Theorem 3.1, the implication (a) ⇒ (b) or (a) ⇒ (c) is no longer
true if B is nonnegative but without the irreducibility of Re B. One example is
26
A = B = J2 ⊕ [a], where 0 < a ≤ 1/2 (cf. the end of the proof of Theorem 2.13 (b)).
The next example shows that the same can be said if B is not nonnegative but Re B
is irreducible.
Example 3.4. Let A = J3 and
B =
0 −√2
1
0
0
0
1
0 √2/2
.
Then Re B is easily seen to be irreducible. We now show that W (B) = D. This is
seen via [13, Corollary 2.5] by letting u = 0 and λ = √2/2 therein and checking that
tr (B∗B2) = tr
0 0
0
0 0
1
0 0 √2/4
√2
4
=
= λλ2
and tr (B∗B) = 9/2 ≥ 5λ2, where tr (·) denotes the trace of a matrix. We next prove
that 1 is an eigenvalue of Re (A ⊗ B). Indeed, since
we need to check that
Re (A ⊗ B) =
03 B 03
B∗
03 B
03 B∗ 03
1
2
,
det
2I3 −B 03
−B∗
2I3 −B
03 −B∗
2I3
= 0.
By a repeated use of the Schur decomposition, the above determinant is seen to be
27
equal to
2I3 −B
−B∗
2I3
−B∗
03
1
2
(
I3) [−B 03]
det (2I3) det
−B∗
−
2I3 − (1/2)B∗B −B
2I3
−1 −√2/2
√2/2
1
= 8 det
= 8 det (4I3 − B∗B − BB∗)
= 8 det
1
−1
−√2/2 √2/2
1
= 0
as required. Since W (A ⊗ B) is a circular disc centered at the origin (by the unitary
similarity of A ⊗ B and eiθ(A ⊗ B) for all real θ) and w(A ⊗ B) ≤ kAkw(B) = 1, we
infer from 1 ∈ σ(Re (A⊗ B)) that W (A⊗ B) = D. Hence w(A⊗ B) = 1 = kAkw(B).
But, obviously, we have nB = ∞ and pA = 2.
(cid:3)
The next corollary gives a more concrete equivalent condition, in terms of block-
shift matrices, for w(A ⊗ B) = kAkw(B) when A = B < 0 and Re B is irreducible.
Corollary 3.5. Let A be an n-by-n nonnegative matrix with Re A irreducible. Then
the following conditions are equivalent:
(a) w(A ⊗ A) = kAkw(A),
(b) pA = nA (≤ ∞), and
(c) either A is unitarily similar to [a]⊕ A′ with a ≥ kA′k, or A is permutationally
similar to a block-shift matrix
A′′ =
0 A1
0
28
. . .
. . . Ak
0
with kA1 · · · Akk = kAk.
Proof. We may assume that kAk = 1. The implication (a) ⇒ (b) is by Theorem 3.1
and Proposition 1.3 (d). For (b) ⇒ (c), if pA = nA = ∞, then A has a unitary part
by Proposition 1.2 (c), and hence A is unitarily similar to [a]⊕ A′ with a = 1 ≥ kA′k
as asserted. On the other hand, if pA = nA < ∞, then w(A⊗ A) = w(A) by Theorem
2.13 (b). Hence Theorem 2.2 implies that W (A) is a circular disc centered at the
origin. For a nonnegative A with Re A irreducible, this is equivalent to A being
permutationally similar to the block-shift matrix A′′ (cf.
[16, Theorem 1 (a)⇔(r)]).
As nA′′ = k by Lemma 3.2 (e), we also have pA = k. Thus kAkk = kAkk = 1,
which yields that kA1 · · · Akk = 1 = kAk as required. Finally, for (c) ⇒ (a), if A is
unitarily similar to [a] ⊕ A′ with a ≥ kA′k, then w(A ⊗ A) = w(A) by Lemma 2.1.
On the other hand, if A is permutationally similar to the block-shift matrix A′′ with
kA1 · · · Akk = 1, then
kAkk = kA′′kk = kA1 · · · Akk = 1 = kAkk.
Thus pA ≥ k = nA. The equality w(A ⊗ A) = w(A) then follows from Theorem
3.1.
Corollary 3.6. Let A = [aij]n
i,j=1, where aij ≥ 0 for all i and j, aij = 0 for i ≥ j,
and ai,i+1 > 0 for all i. Then the following conditions are equivalent:
(a) w(A ⊗ A) = kAkw(A),
(b) pA = nA = n − 1, and
(c) a12 = · · · = an−1,n and aij = 0 for all other pairs of i and j.
Proof. In this case, A is nonnegative, Re A is irreducible and nA = n − 1. Con-
sequently, Corollary 3.5 yields the equivalence of (a), (b) and the condition (c')
that A is permutationally similar to a block-shift matrix A′′ as in Corollary 3.5
29
(c). Since k = nA′′ = nA by Lemma 3.2 (e), A′′ is necessarily equal to A with
a12 · · · an−1,n = kAk and aij = 0 for all other pairs of i and j. The norm condition
above yields that a12 = · · · = an−1,n = kAk. Thus (c') is the same as (c), and we have
the equivalence of (a), (b) and (c).
Acknowledgements
This research was partially supported by the National Science Council of the Re-
public of China under projects NSC-101-2115-M-008-006, NSC-101-2115-M-009-001
and NSC-101-2115-M-009-004 of the respective authors. P. Y. Wu was also supported
by the MOE-ATU. This paper was presented by him at the 4th International Confer-
ence on Matrix Analysis and Applications in Konya, Turkey. He thanks the organizers
for their works with the conference.
30
References
[1] H.-L. Gau, Numerical ranges of reducible companion matrices, Linear Algebra
Appl. 432 (2010), pp. 1310 -- 1321.
[2] H.-L. Gau and P. Y. Wu, Numerical range of S(φ), Linear Multilinear Algebra
45 (1998), pp. 49 -- 73.
[3] H.-L. Gau and P. Y. Wu, Condition for the numerical range to contain an elliptic
disc, Linear Algebra Appl. 364 (2003), pp. 213 -- 222.
[4] H.-L. Gau and P. Y. Wu, Finite Blaschke products of contractions, Linear Algebra
Appl. 368 (2003), pp. 359 -- 370.
[5] H.-L. Gau and P. Y. Wu, Companion matrices: reducibility, numerical ranges
and similarity to contractions, Linear Algebra Appl. 383 (2004), pp. 127 -- 142.
[6] H.-L. Gau and P. Y. Wu, Numerical ranges of companion matrices, Linear Al-
gebra Appl. 421 (2007), pp. 202 -- 218.
[7] K. Gustafson and D. K. M. Rao, Numerical Range. The Field of Values of Linear
Operators and Matrices, Springer, New York, 1997.
[8] U. Haagerup and P. de la Harpe, The numerical radius of a nilpotent operator
on a Hilbert space, Proc. Amer. Math. Soc. 115 (1992), pp. 371 -- 379.
[9] P. R. Halmos, A Hilbert Space Problem Book, 2nd ed., Springer, New York, 1982.
[10] J. A. R. Holbrook, Multiplicative properties of the numerical radius in operator
theory, J. Reine Angew. Math. 237 (1969), pp. 166 -- 174.
[11] R. A. Horn and C. R. Johnson, Matrix Analysis, Cambridge University Press,
Cambridge, 1985.
[12] R. A. Horn and C. R. Johnson, Topics in Matrix Analysis, Cambridge University
Press, Cambridge, 1991.
31
[13] D. S. Keeler, L. Rodman and I. M. Spitkovsky, The numerical range of 3 × 3
matrices, Linear Algebra Appl. 252 (1997), pp. 115 -- 139.
[14] C.-K. Li, B.-S. Tam and P. Y. Wu, The numerical range of a nonnegative matrix,
Linear Algebra Appl. 350 (2002), pp. 1 -- 23.
[15] V. Pt´ak, Lyapunov equations and Gram matrices, Linear Algebra Appl. 49
(1983), pp. 33 -- 55.
[16] B.-S. Tam and S. Yang, On matrices whose numerical ranges have circular or
weak circular symmetry, Linear Algebra Appl. 302/303 (1999), pp. 193 -- 221.
[17] J. P. Williams and T. Crimmins, On the numerical radius of a linear operator,
Amer. Math. Monthly 74 (1967), pp. 832 -- 833.
[18] P. Y. Wu, Unitary dilations and numerical ranges, J. Operator Theory 38 (1997),
pp. 25 -- 42.
[19] P. Y. Wu, A numerical range characterization of Jordan blocks, Linear Multilin-
ear Algebra 43 (1998), pp. 351 -- 361.
[20] P. Y. Wu, Numerical ranges as circular discs, Applied Math. Lett. 24 (2011),
pp. 2115 -- 2117.
[21] P. Y. Wu, H.-L. Gau and M.-C. Tsai, Numerical radius inequality for C0 con-
tractions, Linear Algebra Appl. 430 (2009), pp. 1509 -- 1516.
32
|
1804.01022 | 1 | 1804 | 2018-04-03T15:05:49 | Green's function of the problem of bounded solutions in the case of a block triangular coefficient | [
"math.FA",
"math.CA",
"math.DS",
"math.OA",
"math.SP"
] | It is known that the equation $x'(t)=Ax(t)+f(t)$, where $A$ is a bounded linear operator, has a unique bounded solution $x$ for any bounded continuous free term~$f$ if and only if the spectrum of the coefficient $A$ does not intersect the imaginary axis. The solution can be represented in the form \begin{equation*} x(t)=\int_{-\infty}^{\infty}\mathcal G(s)f(t-s)\,ds. \end{equation*} The kernel $\mathcal G$ is called Green's function. In this paper, the case when $A$ admits a representation by a block triangular operator matrix is considered. It is shown that the blocks of $\mathcal G$ are sums of special convolutions of Green's functions of diagonal blocks of $A$. | math.FA | math |
GREEN'S FUNCTION OF THE PROBLEM OF BOUNDED SOLUTIONS
IN THE CASE OF A BLOCK TRIANGULAR COEFFICIENT
V.G. KURBATOV AND I.V. KURBATOVA
Abstract. It is known that the equation x′(t) = Ax(t) + f (t), where A is a bounded linear
operator, has a unique bounded solution x for any bounded continuous free term f if and only
if the spectrum of the coefficient A does not intersect the imaginary axis. The solution can be
represented in the form
G(s)f (t − s) ds.
x(t) = Z ∞
−∞
The kernel G is called Green's function. In this paper, the case when A admits a representation
by a block triangular operator matrix is considered. It is shown that the blocks of G are sums
of special convolutions of Green's functions of diagonal blocks of A.
Introduction
We consider the equation
(1)
where A is a linear bounded operator acting in a Banach space X. We assume that f is
continuous.
x′(t) − Ax(t) = f (t),
The bounded solutions problem is the problem of finding a bounded solution x that
corresponds to a bounded free term f . The bounded solutions problem is closely connected
with the problem of the exponential dichotomy of solutions. For the discussion of the
bounded solutions problem from different points of view and related questions, see [1, 2,
4, 9, 10, 20, 26, 27, 38, 42, 44, 46] and the references therein.
It is well known (see Theorem 5) that equation (1) has a unique bounded solution x
for any bounded continuous free term f if and only if the spectrum of the coefficient A is
disjoint from the imaginary axis. In this case, the solution can be represented in the form
The kernel G is called Green's function.
x(t) =Z ∞
−∞
G(s)f (t − s) ds.
In this paper, we consider the case when A admits a representation in the form of a
block triangular matrix (4). The simplest 2×2 matrix representation of the coefficient A is
naturally induced by the decomposition of the space X into the direct sum X−⊕X+ of two
spectral subspaces related to the parts of σ(A) that lie in the left and right complex half-
planes. This matrix representation is diagonal, but it can be 'bad' in the sense that the
corresponding projectors have large norms; in such a case it may be convenient to replace
one of the subspaces by the orthogonal (or close to orthogonal) complement of the other; as
a result one will arrive at a triangular matrix representation of A. Similarly, the spectrum
of A may be divided into clusters; so, it is again natural to use a diagonal or triangular
matrix representation; the phenomenon of clusterization is discussed, e.g., in [21, lecture
12], [11, 37]. Representation by triangular operator matrices is also natural for causal
1991 Mathematics Subject Classification. 47A60; 47A80; 34B27; 34B40; 34D09.
Key words and phrases. bounded solutions problem; Green's function; divided difference with operator argu-
ments; block matrix; causal operator.
1
2
V.G. KURBATOV AND I.V. KURBATOVA
operators; in their turn, causal operators are widely used in control theory [13, 16, 53] and
functional differential equations [34, 35, 36]. For other aspects of the theory of triangular
operator matrices, see [7, 8, 19, 22, 23, 28, 31, 33, 45] and the references therein.
The main results of this paper are Theorems 19 and 23; see also Theorem 5. From these
theorems, it follows that Green's function is also induced by a triangular matrix and its
blocks can be represented as the sums of special convolutions of Green's functions of the
diagonal blocks of A; see Example 2.
Similar representations and related formulas for the fundamental solution of equa-
tion (1) were proposed, discussed, and applied by many authors [6, 12, 14, 17, 24, 32,
41, 43, 47, 50, 51, 52]; such formulas are widely used in numerical methods and other
applications. We repeat some of these results in this paper (i) for the convenience of their
comparison with our results connected with Green's function, and because (ii) we propose
a new proof for them, (iii) and discuss the infinite-dimensional case, which requires some
additional considerations in the proof; see Section 3.
The paper is organized as follows. In Section 1, we recall the definition of an analytic
function with an operator argument. In Section 2, we describe the representation of the
fundamental solution of initial value problem and Green's function of bounded solutions
problem in the form of the analytic functions exp±,t and gt, respectively, of the coeffi-
cient A. In Section 3, we discuss the subalgebra of operators induced by block triangular
matrices. This subalgebra is not full, which leads to some technical difficulties in the sub-
sequent presentation. In Section 4, we describe (Theorem 10) a representation of blocks of
an analytic function f of a triangular matrix via contour integrals. In Section 5, the main
terms of the formula from Theorem 10 are represented as divided differences of f with
operator arguments (Theorem 17). In Section 6, we show that divided differences of exp±,t
and gt can be represented as convolutions with respect to the variable t of functions of one
variable (Theorem 22). In Section 7, we describe a representation of divided differences of
exp±,t and gt with operator arguments (Theorem 23). The combination of Theorems 19
and 23 allows one to represent the blocks of the fundamental solution of initial value
problem and Green's function of the bounded solutions problem as special convolutions
of the functions exp±,t and gt applied to the diagonal blocks of A (Examples 1 and 2).
1. Functions of operators
Let X and Y be non-zero complex Banach spaces. We denote by B(X, Y ) the set of all
bounded linear operators A : X → Y . If X = Y , we use the brief notation B(X). The
symbol 1 = 1X stands for the identity operator from B(X).
Let B be a non-zero complex Banach algebra [5, 29, 48] with the unit 1 (unital algebra).
The main example of a unital Banach algebra is the algebra B(X); another important
example is the algebra of all n × n matrices, n ∈ N.
A subset R of an algebra B is called a subalgebra if A + B, λA, AB ∈ R for all A, B ∈ R
and λ ∈ C. If the unit 1 of an algebra B belongs to its subalgebra R, then R is called a
subalgebra with a unit or a unital subalgebra.
A unital subalgebra R of a unital algebra B is called [5, ch. 1, § 3.6] full if it possesses the
property: if for B ∈ R there exists B−1 ∈ B such that BB−1 = B−1B = 1, then B−1 ∈ R.
Below (Remark 1) we will see that the subalgebra of all block triangular matrices is not
always full.
Let B be a (nonzero) unital algebra and A ∈ B. The set of all λ ∈ C such that the
element λ1 − A is not invertible is called the spectrum of the element A (in the algebra B)
and is denoted by the symbol σ(A) or σB(A). The complement ρ(A) = ρB(A) = C \ σ(A)
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
3
is called the resolvent set of A. The function Rλ = (λ1 − A)−1 is called the resolvent of
the element A. The spectral radius r(A) = rB(A) is the radius of the smallest closed circle
in C with center at 0 that contains σ(A).
Proposition 1 ([5, Chap. 1, Sec. 4, Theorem 3], [48, Theorem 10.18]). Let R be a closed
unital subalgebra of a unital algebra B. Then the spectrum σR(A) of an element A ∈ R in
the algebra R is the union of the spectrum σ(A) of A in the algebra B and (possibly empty)
collection of bounded connected components of the resolvent set ρ(A). In particular, the
spectral radii rR(A) and rB(A) coincide.
Let A ∈ B and let U ⊆ C be an open set that contains the spectrum σ(A). The set U
must not be connected. Let f : U → C be an analytic function. The function f of the
element A is defined [29, ch. V, § 1], [10, p. 17] by the formula
f (A) =
f (λ)(λ1 − A)−1 dλ,
(2)
1
2πiZΓ
where the contour Γ surrounds the set σB(A) in the counterclockwise direction and the
function f is analytic inside Γ.
Proposition 2 ([29, Theorem 5.2.5], [48, Theorem 10.27]). The mapping f 7→ f (A)
preserves algebraic operations, i. e.,
(f + g)(A) = f (A) + g(A),
(αf )(A) = αf (A),
(f g)(A) = f (A)g(A),
where f + g, αf , and f g are defined pointwise.
Corollary 3. For the function rλ0(λ) = 1
λ0−λ , λ0 ∈ ρ(A), we have
rλ0(A) = (λ01 − A)−1.
Proof. The proof follows from Proposition 2.
(cid:3)
2. The differential equation with a constant coefficient
In this Section, we describe three analytic functions that are closely related to the
representation of solutions of linear differential equations with constant coefficients.
For λ ∈ C and t ∈ R, we consider the functions
exp+, t(λ) =(eλt,
exp−, t(λ) =(0,
0,
−eλt,
if t > 0,
if t < 0,
if t > 0,
if t < 0,
gt(λ) =(exp−, t(λ),
exp+, t(λ),
if Re λ > 0,
if Re λ < 0.
These functions are undefined for t = 0. The function gt is also undefined for Re λ = 0.
For any fixed t 6= 0, all three functions are analytic on their domains.
Let X be a Banach space and A ∈ B(X). We consider the differential equation
(3)
We recall two well-known theorems. The first theorem shows that exp±, (·)(A) are
x′(t) = Ax(t) + f (t),
t ∈ R.
fundamental solutions of the initial value problems.
4
V.G. KURBATOV AND I.V. KURBATOVA
Theorem 4 ([10, ch. 1, § 4], [26, ch. IV, Corollary 2.1]). Let f : R → X be a continuous
function. The solution of the initial value problem
x′(t) = Ax(t) + f (t),
x(0) = 0
t > 0,
is the function
x(t) =Z t
0
exp+, s(A) f (t − s) ds,
t > 0.
The solution of the initial value problem
is the function
x′(t) = Ax(t) + f (t),
x(0) = 0
t < 0,
x(t) =Z 0
t
exp−, s(A) f (t − s) ds,
t < 0.
The function t 7→ exp+, t(A) is called [26] the fundamental solution for equation (3).
Now we turn to the bounded solutions problem, i.e. the problem of seeking bounded
solution x : R → X under the assumption that the free term f : R → X is a bounded
function.
Theorem 5 ([10, Theorem 4.1, p. 81]). Let A ∈ B(X). Equation (3) has a unique bounded
on R solution x for any bounded continuous function f if and only if the spectrum σ(A)
of A does not intersect the imaginary axis. This solution admits the representation
where
x(t) =Z ∞
−∞
G(s)f (t − s) ds,
G(t) = gt(A),
t 6= 0.
The function G is called [10] the Green's function of the bounded solutions problem for
equation (3).
3. Causal spectrum of a block triangular matrix
Let a Banach space X be represented as the direct sum of its closed nonzero subspaces
Xi, i = 1, . . . , n:
This means that every x ∈ X can be uniquely represented in the form
X = X1 ⊕ X2 ⊕ · · · ⊕ Xn.
x = x1 + x2 + · · · + xn,
where xi ∈ Xi, i = 1, . . . , n. It is easy to prove that the norm on X is equivalent to the
norm
We denote by M = M(X1, X2, . . . , Xn) the set of all operator matrices
kxk = kx1k + kx2k + · · · + kxnk.
{ Tij ∈ B(Xj, Xi) : i, j = 1, . . . , n }.
We endow M with the norm k{ Tij }k = maxjPn
i=1kTijk. It is easy to show that M is
a unital Banach algebra with respect to the usual matrix multiplication, and the Banach
algebra M is isomorphic (not isometrically) to the algebra B(X). As usual, we do not
distinguish very carefully matrices and operators induced by them.
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
5
We denote by M+ = M+(X1, X2, . . . , Xn) the set of all lower triangular matrices
.
(4)
A11
A21
. . .
0
A22
. . .
An−1,1 An−1,2
An,1
An,2
0
0
. . .
. . .
. . .
. . .
. . . An−1,n−1
. . . An,n−1 An,n
0
0
. . .
0
We denote by B+(X) the class of operators induces by M+. Clearly, M+ is a closed
subalgebra of the algebra M. Therefore, B+(X) is a closed subalgebra of the algebra
B(X). We call operators from the class B+(X) causal in analogy with a similar class
of operators in the control theory [13, 16, 53] and in the theory of functional differential
equations [34, 35, 36], see also the references therein. Namely, if one interprets the indices
i = 1, . . . , n as successive instants of time, then the triangularity of a matrix A means
that the value (Ax)i of the 'output' Ax at any instant i may depend only on values xj of
the 'input' x at the previous instants j ≤ i.
Remark 1. The subalgebra M+ (and consequently, the subalgebra B+(X)) may be not
full if the space X is infinite-dimensional. We give a corresponding example. Let X
be the space Lp(R), 1 ≤ p ≤ ∞. We represent X = Lp(R) as Lp(−∞, 0] ⊕ Lp[0, ∞),
where Lp(−∞, 0] and Lp[0, ∞) are the subspaces of functions that are equal to zero out-
is induced by a lower triangular matrix (thus it is causal), but the inverse operator
side (−∞, 0] and [0, ∞) respectively. Clearly, the operator of delay (cid:0)Sx(cid:1)(t) = x(t − 1)
(cid:0)S−1x(cid:1)(t) = x(t + 1) is induced by an upper triangular matrix (thus S−1 is not causal).
Consequently, in contrast to the finite-dimensional case, the (ordinary) spectrum of a
triangular matrix may be not the union of the spectra of its diagonal blocks, see Propo-
sition 7. See a more detailed discussion of this phenomenon in [25].
If an operator T ∈ B+(X) is invertible and the inverse operator belongs to B+(X),
we say that T is causally invertible. We call the spectrum of T ∈ B+(X) in the algebra
B+(X) the causal spectrum and denote it by σ+(T ). Clearly,
σ(T ) ⊆ σ+(T ).
We denote by ρ+(T ) the causal resolvent set C \ σ+(T ). The same terminology and
notation will be used for matrices M ∈ M+.
We recall that an open set D ⊆ C is called simply-connected if any simple closed curve
in D can be shrunk continuously to a point.
Proposition 6. Let the domain D ⊆ C of an analytic function f be simply-connected
(examples of such functions are exp±, t and gt). Let T ∈ B+(X). Then σ+(T ) ⊂ D
provided σ(T ) ⊂ D. Thus the function f (T ) of a causal operator T is defined in algebras
B(X) and B+(X) simultaneously.
Proof. A possible difficulty can occur when the spectrum σ(T ) is contained in the domain
D of the definition of f , but σ+(T ) * D. Therefore the resolvent (λ1−A)−1 in integral (2)
is defined in B(X), but it may not exist in B+(X).
By Proposition 1, the causal spectrum σ+(T ) is the union of the ordinary spectrum σ(T )
and (possibly) some bounded components of the resolvent set ρ(A). Since the domain D
of f is simply-connected, bounded components of the resolvent set ρ(A) are contained in
the domain D, provided the spectrum σ(T ) itself is contained in the domain D.
(cid:3)
6
V.G. KURBATOV AND I.V. KURBATOVA
Proposition 7 ([34], [36, Proposition 2.1.7]). The causal spectrum of a lower triangular
matrix { Tij } (and the causal spectrum of the corresponding operator) is the union of the
(ordinary) spectra σ(Tii) of the diagonal blocks Tii.
Proof. It suffices to prove that a lower triangular matrix has a lower triangular inverse if
and only if all diagonal blocks Tii are invertible.
Let the lower triangular matrix { Bij } be the inverse of the lower triangular ma-
trix { Tij }. Then it follows from the matrix multiplication rule that Bii are inverses
of Tii.
Conversely, let the diagonal blocks Tii be invertible. Then from the Gaussian elimination
(cid:3)
algorithm it easily follows that the inverse matrix exists and is triangular.
4. Functions of block triangular matrices
Theorem 8. Let a causal matrix
be causally invertible. Then the elements of the inverse matrix
T =
B =
T1,1
T2,1
. . .
0
T2,2
. . .
Tn−1,1 Tn−1,2
Tn,1
Tn,2
0
0
. . .
. . .
. . .
. . .
. . . Tn−1,n−1
. . .
Tn,n−1
0
0
. . .
0
Tn,n
B1,1
B2,1
. . .
0
B2,2
. . .
Bn−1,1 Bn−1,2
Bn,1
Bn,2
0
0
. . .
. . .
. . .
. . .
. . . Bn−1,n−1
. . . Bn,n−1 Bn,n
0
0
. . .
0
have the form
Bi,j = Xi=i1>i2···>im=j
In particular,
(−1)m+1T −1
i1,i1
Ti1,i2T −1
i2,i2
Ti2,i3 . . . Tim−1,imT −1
im,im
,
i ≥ j.
Bi,i = T −1
i,i ,
Bi+1,i = −T −1
Bi+2,i = −T −1
i+1,i+1Ti+1,iT −1
i,i ,
i+2,i+2Ti+2,iT −1
i,i + T −1
i+2,i+2Ti+2,i+1T −1
i+1,i+1Ti+1,iT −1
i,i .
Proof. Let us verify that T B = 1. Clearly, (T B)ii = 1. We calculate, for example,
(T B)n,1:
(T B)n,1 = Tn,1T −1
11 − Tn,2T −1
33 T31T −1
11 + T −1
33 T32T −1
22 T21T −1
11 ) + . . .
22 T21T −1
(−1)m+1T −1
11 + Tn,3(−T −1
n,nTn,i2T −1
i2,i2
Ti2,i3 . . . Tim−1,1T −1
1,1
22 T21T −1
11 + Tn,3(−T −1
33 T31T −1
11 + T −1
33 T32T −1
22 T21T −1
11 ) + . . .
(−1)m+1Tn,i2T −1
i2,i2Ti2,i3 . . . Tim−1,1T −1
1,1 = 0.
= Tn,1T −1
+ Tn,n Xn=i1>i2···>im=1
+ Xn=i1>i2···>im=1
11 − Tn,2T −1
In a similar way one establishes that BT = 1.
(cid:3)
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
7
Theorem 9. Let A ∈ M+. Then the causal resolvent (i.e., the resolvent (λ1 − A)−1 at
the points λ ∈ ρ+(A)) of A has the form
(λ1 − A)−1 =
R1,1
R2,1
. . .
0
R2,2
. . .
Rn−1,1 Rn−1,2
Rn,2
Rn,1
0
0
. . .
. . .
. . .
. . .
. . . Rn−1,n−1
. . . Rn,n−1 Rn,n
0
0
. . .
0
,
where Rij for i ≥ j are defined by the formula
Rij = Xi=i1>i2···>im=j
In particular,
(λ1 − Ai1,i1)−1Ai1,i2(λ1 − Ai2,i2)−1Ai2,i3 . . . Aim−1,im(λ1 − Aim,im)−1.
Rii = (λ1 − Aii)−1,
Ri+1,i = (λ1 − Ai+1,i+1)−1Ai+1,i(λ1 − Aii)−1,
Ri+2,i = (λ1 − Ai+2,i+2)−1Ai+2,i(λ1 − Aii)−1
+ (λ1 − Ai+2,i+2)−1Ai+2,i+1(λ1 − Ai+1,i+1)−1Ai+1,i(λ1 − Aii)−1.
Proof. The proof follows from Theorem 8. The sign (−1)m+1 disappears because (λ1 −
A)ij = −Aij for i > j.
(cid:3)
Theorem 10. Let a function f be analytic in a neighborhood of the causal spectrum σ+(A)
of a matrix A ∈ M+. Then the matrix F = f (A) has the form
F =
0
F2,2
. . .
0
0
. . .
. . .
. . .
. . .
. . . Fn−1,n−1
. . .
Fn,n−1
Fn−1,1 Fn−1,2
Fn,1
Fn,2
0
0
. . .
0
Fn,n
,
F1,1
F2,1
. . .
2πiZΓ
1
where Fij for i ≥ j are defined by the formula
Fij = Xi=i1>i2···>im=j
× Aim−1,im(λ1 − Aim,im)−1 dλ,
f (λ)(λ1 − Ai1,i1)−1Ai1,i2(λ1 − Ai2,i2)−1Ai2,i3 . . .
where Γ surrounds the causal spectrum σ+(A) of the matrix A. In particular,
Fii =ZΓ
Fi+1,i =ZΓ
Fi+2,i =ZΓ
+ZΓ
f (λ)(λ1 − Aii)−1 dλ,
f (λ)(λ1 − Ai+1,i+1)−1Ai+1,i(λ1 − Aii)−1 dλ,
f (λ)(λ1 − Ai+2,i+2)−1Ai+2,i(λ1 − Aii)−1 dλ
f (λ)(λ1 − Ai+2,i+2)−1Ai+2,i+1(λ1 − Ai+1,i+1)−1Ai+1,i(λ1 − Aii)−1 dλ.
Proof. Substituting into formula (2) the representation of (λ1 − A)−1 from Theorem 9,
we obtain the desired result.
(cid:3)
8
V.G. KURBATOV AND I.V. KURBATOVA
Remark 2. Let a matrix A ∈ M+ has only two non-zero diagonals:
.
0
A1,1
A2,1 A2,2
. . .
. . .
0
0
0
0
0
0
. . .
. . .
. . .
. . .
. . . An−1,n−1
. . . An,n−1 An,n
0
0
. . .
0
Let a function f be analytic in a neighborhood of the causal spectrum σ+(A) of the
matrix A. Then it follows from Theorem 10 that the elements Fij for i ≥ j of the matrix
F = f (A) consist of exactly one summand:
Fij =
1
2πiZΓ
f (λ)(λ1 − Ai,i)−1Ai,i+1(λ1 − Ai+1,i+1)−1Ai+1,i+2 . . . Aj−1,j(λ1 − Aj,j)−1 dλ.
For the function f = exp+, t this phenomenon was described in [6] and applied in [24].
Corollary 11. Let the domain D ⊆ C of an analytic function f be simply-connected
(examples of such functions are exp±, t and gt). Then the conclusion of Theorem 10 is
true if the function f is analytic in a neighborhood of the ordinary spectrum σ(A) of the
matrix A ∈ M+.
Proof. The proof follows from Proposition 6.
(cid:3)
Remark 3. For scalar matrices, Theorem 10 goes back to [47]. For matrices consisting of
finite-dimensional blocks, it was published in [12, Theorem 2]. More precisely, in [12] it
was considered only the case of polynomial functions f ; but from the case of polynomials,
it follows the case of general analytic functions, since (if A is a scalar matrix) any analytic
function can be replaced by its interpolating polynomial.
5. Divided differences
Let µ1, µ2, . . . , µn be given complex numbers (some of them may coincide with others)
called points of interpolation. Let a complex-valued function f be defined and analytic in
a neighborhood of these points. Divided differences of the function f with respect to the
points µ1, µ2, . . . , µn are defined (see, e.g., [18, 30]) by the recurrent relations
f [0](µi) = f (µi),
f [1](µi, µi+1) =
f [0](µi+1) − f [0](µi)
µi+1 − µi
,
f [m](µi, . . . , µi+m) =
f [m−1](µi+1, . . . , µi+m) − f [m−1](µi, . . . , µi+m−1)
µi+m − µi
(5)
.
In these formulas, if the denominator vanishes, then the quotient is understood as the
derivative with respect to one of the arguments of the previous divided difference (the
naturalness of this agreement can be derived by continuity from Corollary 13).
Proposition 12 ([18, ch. 1, formula (54)]). Let a function f be analytic in an open
neighbourhood of the points of interpolation µ1, µ2, . . . , µn. Then divided differences
admit the representation
f [n−1](µ1, . . . , µn) =
1
2πiZΓ
f (z)
Ω(z)
dz,
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
9
where the contour Γ encloses all the points of interpolation and
n
Corollary 13. Divided differences are differentiable functions of their arguments.
Ω(z) =
(z − µk).
Yk=1
Proof. The proof follows from Proposition 12.
Corollary 14. If D ⊆ C is the domain of an analytic function f , then f [n−1] is defined
in Dn.
(cid:3)
Proof. The proof follows from Proposition 12.
Corollary 15. Divided differences f [n−1](µ1, . . . , µn) are symmetric function, i.e., they
do not depend on the order of their arguments µ1, . . . , µn.
(cid:3)
Proof. The proof follows from Proposition 12.
(cid:3)
Proposition 16 ([18, ch. 1, formula (48)], [30, p. 19, formula (1)]). Let the points of
interpolation µ1, . . . , µn be distinct. Then divided differences admit the representation
f [n−1](µ1, . . . , µn) =
n
Xj=1
Proof. The proof follows from Proposition 12.
f (µj)
n
.
(µj − µk)
Qk=1
k6=j
(cid:3)
For a function f analytic in a neighborhood of the causal spectrum of a matrix A ∈ M+,
we denote by f [m−1](A; i1, i2, . . . , im) the summands from Theorem 10:
f [m−1](A; i1, i2, . . . , im) =
1
2πiZΓ
f (λ)(λ1 − Ai1,i1)−1Ai1,i2(λ1 − Ai2,i2)−1Ai2,i3 . . .
× Aim−1,im(λ1 − Aim,im)−1 dλ,
(6)
where Γ surrounds the causal spectrum σ+(A) of the matrix A.
Theorem 17. Let a function f be analytic in a neighborhood of the causal spectrum
σ+(A) of a matrix A ∈ M+. Then
f [m−1](A; i1, i2, . . . , im) =
1
(2πi)m ZΓi1
· · ·ZΓim
f [m−1](λi1, . . . , λim)(λi11 − Ai1,i1)−1
(7)
× Ai1,i2(λi21 − Ai2,i2)−1Ai2,i3 . . . Aim−1,im(λim1 − Aim,im)−1 dλi1 . . . dλim,
where Γik surrounds the spectrum of Aik,ik .
Proof. Since f is analytic in a neighborhood of σ+(A) =Sn
we may assume without loss of generality that Γik in (7) surrounds the whole Sn
and, moreover, Γik surrounds Γik+1, see Figure 1.
i=1 σ(Aii) (see Proposition 7),
i=1 σ(Aii)
Since the contours Γik are disjoint, the points λi1, . . . , λim in the integrand of (7)
are distinct. Therefore we can substitute the representation of divided differences from
Proposition 16 into definition (7):
f [m−1](A; i1, i2, . . . , im) =
1
(2πi)m ZΓi1
· · ·ZΓim
m
Xj=1
f (λij )
m
(λij − λik)
(8)
× (λi11 − Ai1,i1)−1Ai1,i2 . . . Aim−1,im(λim1 − Aim,im)−1 dλi1 . . . dλim.
Qk=1
k6=j
10
V.G. KURBATOV AND I.V. KURBATOVA
✬
✫
✬
⑦⑦⑦✬
✫
✫
σ+(A)
✻Γi3
✩
✪
✻Γi2
✩
✩
✻Γi1
✪
✪
f
Figure 1. The choice of the contours Γik in the proof of Theorem 17. The
symbol f means the localization of singularities of f
Let us begin with the first summand. We have
f (λi1)(λi11 − Ai1,i1)−1Ai1,i2
1
2πiZΓii1
×(cid:20). . .
×(cid:20) 1
×(cid:20) 1
1
2πiZΓim−2
2πiZΓim−1
2πiZΓim
1
1
λi1 − λim−2
× (λim−21 − Aim−2,im−2)−1Aim−2,im−1
1
λi1 − λim−1
(λim−11 − Aim−1,im−1)−1Aim−1,im
λi1 − λim
(λim1 − Aim,im)−1 dλim(cid:21) dλim−1(cid:21) dλim−2 . . .(cid:21) dλi1.
By Corollary 3, for the internal integral, we have
1
2πiZΓim
1
λi1 − λim
(λim1 − Aim,im)−1 dλim = (λi11 − Aim,im)−1.
Now we can calculate the next internal integral (again using Corollary 3):
1
2πiZΓim−1
=(cid:20) 1
2πiZΓim−1
1
λi1 − λim−1
(λim−11 − Aim−1,im−1)−1Aim−1,im(λi11 − Aim,im)−1 dλim−1
1
λi1 − λim−1
(λim−11 − Aim−1,im−1)−1 dλim−1(cid:21)Aim−1,im(λi11 − Aim,im)−1
= (λi11 − Aim−1,im−1)−1Aim−1,im(λi11 − Aim,im)−1.
And so on. Finally, we arrive at the representation (for the first summand in (8))
1
2πiZΓi1
f (λi1)(λi11 − Ai1,i1)−1Ai1,i2(λi11 − Ai2,i2)−1Ai2,i3 . . . Aim−1,im(λi11 − Aim,im)−1 dλi1,
which coincides with formula (7).
Next we show that the other summands in (8) are zero. Let us consider, for example,
the second summand
1
(2πi)m ZΓi1
· · ·ZΓim
f (λi2)
m
(λi2 − λik )
Qk=1
k6=2
× (λi11 − Ai1,i1)−1Ai1,i2 . . . Aim−1,im(λim1 − Aim,im)−1 dλi1 . . . dλim.
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
11
Proceeding as above (i.e. successively calculating integrals over all variables except λi1),
at the final stage, we arrive at the integral
1
2πiZΓi1
1
λi2 − λi1
(λi11 − Ai1,i1)−1 dλi1.
We notice that the singularity of the function λi1 7→ 1
lies inside the contour Γi1. Hence the integrand λi1 7→ 1
outside the contour Γi1 and decreases at infinity as
zero.
λi2 −λi1
1
λ2
i1
λi2 −λi1
(i.e., the point λi2 ∈ Γi2)
(λi11 − Ai1,i1)−1 is analytic
. Therefore the integral equals
(cid:3)
Remark 4. For the case of the first divided difference, a more detailed discussion of
formula (7) can be found in [39, Theorem 41], see also the references therein.
Corollary 18. Let a function f be analytic in a neighborhood of the causal spectrum
σ+(A) of a matrix A ∈ M+. Then the matrix F = f (A) has the form
where Fij, i ≥ j, admits the representation
F11
F21
. . .
0
F22
. . .
Fn−1,1 Fn−1,2
Fn,1
Fn,2
F =
Fij = Xi=i1>i2···>im=j
0
0
. . .
. . .
. . .
. . .
. . . Fn−1,n−1
. . .
Fn,n−1
0
0
. . .
0
Fn,n
,
f [m−1](A; i1, i2, . . . , im).
Proof. The proof follows from Theorems 10 and 17.
(cid:3)
Theorem 19. Let A ∈ M+. Then
E±,1,1
E±,2,1
where E±,i,j, i ≥ j, admits the representation
0
E±,2,2
. . .
. . .
exp±, t(A) =
E±,i,j = Xi=i1>i2···>im=j
E±,n−1,1 E±,n−1,2
E±,n,2
E±,n,1
0
0
. . .
. . .
. . .
. . .
. . . E±,n−1,n−1
. . . E±,n,n−1 E±,n,n
0
0
. . .
0
,
exp[m−1]
±, t
(A; i1, i2, . . . , im);
and (provided the spectrum σ(A) does not intersect the imaginary axis)
gt(A) =
0
G2,2
. . .
Gn−1,1 Gn−1,2
Gn,2
Gn,1
G1,1
G2,1
. . .
Gi,j = Xi=i1>i2···>im=j
0
0
. . .
. . .
. . .
. . .
. . . Gn−1,n−1
. . . Gn,n−1 Gn,n
0
0
. . .
0
,
g[m−1]
t
(A; i1, i2, . . . , im).
where Gi,j, i ≥ j, admits the representation
Proof. The proof follows from Corollary 11 and Theorem 17.
(cid:3)
12
V.G. KURBATOV AND I.V. KURBATOVA
6. Divided differences of the functions exp±, t and gt
Lemma 20. Let the points of interpolation λj be distinct. Then the divided differences
of the function rλ(ν) = 1
λ−ν admit the representation
Proof. The proof is by induction on n. For n = 1 we have
r[n−1]
λ
(λ1, . . . , λn) =
j=1(λ − λj)
.
1
Qn
r[1]
λ (λ1, λ2) =
λ−λ1
1
− 1
λ1 − λ2
λ−λ2
=
1
(λ − λ1)(λ − λ2)
.
Assuming that the formula holds for n − 2, we prove it for n − 1. We have
r[n−1]
λ
(λ1, . . . , λn) =
r[n−1]
λ
(λ1, . . . , λn−2, λn−1) − r[n−1]
λ
λn−1 − λn
(λ1, . . . , λn−2, λn)
We recall [15, 40, 49] that the bilateral (or two-sided) Laplace transform of a function
f : R → C is the function
The value (cid:0)Bf(cid:1)(λ) at the point λ ∈ C is defined if the integral converges absolutely. If f
equals zero on (−∞, 0) (as the function exp+, (·)), this definition takes the form
1
1
Qn−2
j=1 (λ−λj )
λ−λn−1
−
1
1
Qn−2
j=1 (λ−λj )
λ−λn
λn−1 − λn
1
1
λ−λn−1
− 1
λ−λn
λn−1 − λn
=
=
=
. (cid:3)
1
j=1(λ − λj)
j=1 (λ − λj)
Qn−2
Qn
(cid:0)Bf(cid:1)(λ) =Z ∞
−∞
e−λtf (t) dt.
(cid:0)Bf(cid:1)(λ) =Z ∞
0
e−λtf (t) dt.
(cid:0)Bf(cid:1)(λ) =Z 0
−∞
e−λtf (t) dt.
Usually in this case, the integral converges absolutely for Re λ sufficiently large.
If f
equals zero on (0, ∞) (as the function exp−, (·)), the definition of the bilateral Laplace
transform takes the form
In this case, we assume that the integral converges absolutely for Re λ sufficiently small.
We recall the following statement.
Lemma 21. Let λ0 ∈ C.
(a) The bilateral Laplace transform of the function t 7→ exp+, t(λ0) is the function
(cid:0)B exp+, (·)(λ0)(cid:1)(λ) =
1
λ − λ0
,
Re λ > Re λ0.
(b) The bilateral Laplace transform of the function t 7→ exp−, t(λ0) is the function
(cid:0)B exp−, (·)(λ0)(cid:1)(λ) =
1
λ − λ0
,
Re λ < Re λ0.
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
13
(c) The bilateral Laplace transform of the function t 7→ gt(λ0), Re λ0 6= 0, is the
function (the complete domain of the function of B g(·)(λ0) is Re λ > Re λ0 if
Re λ0 < 0 and is Re λ < Re λ0 if Re λ0 > 0)
(cid:0)B g(·)(λ0)(cid:1)(λ) =
1
λ − λ0
,
Re λ < Re λ0.
Proof. Assertion (a) is widely known [40, pp. 300, 305]. The proofs of all assertions are
reduced to straightforward calculations. The proof of assertion (c) can also be obtained
from the definition of gt and (a) and (b).
(cid:3)
We recall [48, ch. 6] that the convolution of two summable functions f, g : R → C is
the function
If f (t) = 0 and g(t) = 0 for t < 0, then this formula takes the form
(cid:0)f ∗ g(cid:1)(t) =Z ∞
−∞
f (s)g(t − s) ds.
If f (t) = 0 and g(t) = 0 for t > 0, then the definition of convolution takes the form
(cid:0)f ∗ g(cid:1)(t) =(R t
(cid:0)f ∗ g(cid:1)(t) =(0
R 0
0 f (s)g(t − s) ds
0
for t > 0,
for t < 0.
t f (s)g(t − s) ds
for t > 0,
for t < 0.
Theorem 22.
(a) The divided differences of the function t 7→ exp+, t admit the representation
exp[n−1]
+, (·) (λ1, . . . , λn) = exp+, (·)(λ1) ∗ · · · ∗ exp+, (·)(λn).
For example, for t > 0, we have
exp[1]
+, t(λ1, λ2) =Z t
+, t(λ1, λ2, λ3) =Z t
0 Z r
0
0
exp[2]
exp+, s(λ1) exp+, t−s(λ2) ds,
exp+, s(λ1) exp+, r−s(λ2) exp+, t−r(λ3) ds dr.
(b) The divided differences of the function t 7→ exp−, t admit the representation
exp[n−1]
−, (·) (λ1, . . . , λn) = exp−, (·)(λ1) ∗ · · · ∗ exp−, (·)(λn).
For example, for t < 0, we have
exp[1]
−, t(λ1, λ2) =Z 0
−, t(λ1, λ2, λ3) =Z 0
t Z 0
r
t
exp[2]
exp−, s(λ1) exp−, t−s(λ2) ds,
exp−, s(λ1) exp−, r−s(λ2) exp−, t−r(λ3) ds dr.
(c) The divided differences of the function t 7→ gt admit the representation
g[n−1]
(·)
(λ1, . . . , λn) = g(·)(λ1) ∗ · · · ∗ g(·)(λn),
Re λ1, . . . , Re λn 6= 0.
14
V.G. KURBATOV AND I.V. KURBATOVA
For example, for t 6= 0, we have
gs(λ1)gt−s(λ2) ds,
g[1]
t (λ1, λ2) =Z ∞
t (λ1, λ2, λ3) =Z ∞
−∞
g[2]
−∞Z ∞
−∞
gs(λ1)gr−s(λ2)gt−r(λ3) ds dr.
Remark 5. Assertion (a) is established in [51]. Assertion (b) is proved in a similar way.
Assertion (c) is proved in [20, p. 53] for other aims. For completeness, we give here an
independent proof of (c).
Proof. Suppose that the points of interpolation λj are distinct. By Proposition 16, we
have
g[n−1]
t
(λ1, . . . , λn) =
Xj=1
n
gt(λj)
n
.
(λj − λk)
Qk=1
k6=j
Form this representation and Lemma 21, it easily follows that the bilateral Laplace trans-
form of the function t 7→ g[n−1]
(λ1, . . . , λn) is
t
(·)
(cid:0)B g[n−1]
(λ1, . . . , λn)(cid:1)(λ) =
1
λ − λj
n
Xj=1
1
n
.
(λj − λk)
Qk=1
k6=j
By Proposition 16, the last expression is the (n − 1)-th divided difference of the function
rλ(ν) = 1
λ−ν . By Lemma 20,
r[n−1]
λ
(λ1, . . . , λn) =
1
j=1(λ − λj)
.
Qn
We apply the inverse bilatiral Laplace transform to tht function λ 7→ r[n−1]
(λ1, . . . , λn).
Clearly, the restriction of the bilatiral Laplace transform to the imaginary axis is the
Fourier transform. The Fourier transform maps the convolution of functions to the prod-
uct of their images [40, p. 337], which implies assertion (c).
λ
The case of coinciding points of interpolation λj follows from continuity.
(cid:3)
7. The divided differences exp[m]
t with operator arguments
In this Section, we apply previous results to the calculation of the functions exp[m]
g[m]
t with operator arguments.
Theorem 23. Let A ∈ M+. Then for t > 0, we have
±, t and g[m]
±, t and
exp[m−1]
+, t
(A; i1, i2, . . . , im−1, im) =Z t
0 Z sm−1
0
· · ·Z s2
0
exp+, s1(Ai1,i1)Ai1,i2 exp+, s2−s1(Ai2,i2)
× Ai2,i3 . . . exp+, sm−1−sm−2(Aim−1,im−1)Aim−1,im exp+, t−sm−1(Aim,im) ds1 . . . dsm−1;
for t < 0, we have
exp[m−1]
−, t
(A; i1, i2, . . . , im−1, im) =Z 0
t Z 0
sm−1
· · ·Z 0
s2
× Ai2,i3 . . . exp−, sm−1−sm−2(Aim−1,im−1)Aim−1,im exp−, t−sm−1(Aim,im) ds1 . . . dsm−1;
exp−, s1(Ai1,i1)Ai1,i2 exp−, s2−s1(Ai2,i2)
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
15
and (if the spectrum σ(A) is disjoint from the imaginary axis) for t 6= 0, we have
g[m−1]
t
(A; i1, i2, . . . , im−1, im) =Z ∞
−∞
· · ·Z ∞
−∞
gs1(Ai1,i1)Ai1,i2 gs2−s1(Ai2,i2)
× Ai2,i3 . . . gsm−1−sm−2(Aim−1,im−1)Aim−1,im gt−sm−1(Aim,im) ds1 . . . dsm−1.
Proof. For simplicity of notation, we prove only the formula
g[2]
t (A; 3, 2, 1) =Z ∞
−∞Z ∞
−∞
By Theorem 17, we have
gs1(A33)A32 gs2−s1(A22)A21 gt−s2(A11) ds1 ds2.
g[2]
t (A; 3, 2, 1) =
1
(2πi)3 ZΓ1ZΓ2ZΓ3
g[2]
t (λ1, λ2, λ3)
× (λ31 − A33)−1A32(λ21 − A22)−1A21(λ11 − A11)−1 dλ1 dλ2 dλ3.
By Theorem 22, we have
g[2]
t (λ1, λ2, λ3) =Z ∞
−∞Z ∞
−∞
gs(λ1) gr−s(λ2) gt−r(λ3) ds dr.
Substituting the latter formula into the former one and performing the integration over
λ1, λ2, and λ3, we arrive at the desired formula.
(cid:3)
Combining Theorems 19 and 23, we obtain the following examples.
Example 1. Let A be the block matrix
Then for t > 0, we have
0
0
A11
0
A21 A22
A31 A32 A33
A =
.
exp+, t(A) =
where
exp+, t(A11)
exp[1]
+, t(A; 2, 1)
+, t(A; 3, 1) + exp[2]
exp[1]
0
exp+, t(A22)
0
0
+, t(A; 3, 2, 1) exp[1]
+, t(A; 3, 2) exp+, t(A33)
(9)
(10)
(11)
,
exp[1]
+, t(A; i1, i2) =Z t
+, t(A; 3, 2, 1) =Z t
0 Z r
0
0
exp[2]
exp+, s(Ai,1i1)Ai1,i2 exp+, t−s(Ai2,i2) ds,
exp+, s(A33)A32 exp+, r−s(A22)A21 exp+, t−r(A11) ds dr.
Remark 6. Integral (11) was first obtained in [3, ch. 10, § 14, formula (5)] in a different
context. Formula (10) for a triangular block matrix (with blocks consisting of scalars) of
the size less than or equal to 4 × 4 was appeared in [51, theorem 1] and for a triangular
block matrix of any size in [6].
Example 2. Let A be the block matrix (9), whose spectrum is disjoint from the imaginary
axis. Then for t 6= 0, we have
gt(A) =
gt(A11)
g[1]
t (A; 2, 1)
g[1]
t (A; 3, 1) + g[2]
t (A; 3, 2, 1) g[1]
0
gt(A22)
t (A; 3, 2) gt(A33)
0
0
,
16
where
V.G. KURBATOV AND I.V. KURBATOVA
gs(Ai1,i1)Ai1,i2gt−s(Ai2,i2) ds,
g[1]
t (A; i1, i2) =Z ∞
t (A; 3, 2, 1) =Z ∞
−∞
−∞Z ∞
−∞
g[2]
gs(A33)A32gr−s(A22)A21gt−r(A11) ds dr.
Acknowledgments
The first author was supported by the Ministry of Education and Science of the Russian
Federation under state order No. 3.1761.2017/4.6. The second author was supported by
the Russian Foundation for Basic Research under research projects No. 16-01-00197.
References
[1] R. R. Akhmerov and V. G. Kurbatov, Exponential dichotomy and stability of neutral type equations, J.
Differential Equations 76 (1988), no. 1, 1 -- 25. MR 964610
[2] A. G. Baskakov, Estimates for the Green's function and parameters of exponential dichotomy of a hyperbolic
operator semigroup and linear relations, Mat. Sb. 206 (2015), no. 8, 23 -- 62, (in Russian); English translation
in Sb. Math., 206 (2015), no. 8, 1049 -- 1086. MR 3438589
[3] R. Bellman, Introduction to matrix analysis, McGraw-Hill Book Co., New York -- Toronto -- London, 1960.
MR 0122820
[4] A. A. Boichuk and A. A. Pokutnii, Bounded solutions of linear differential equations in a Banach space,
Nonlinear Oscillations 9 (2006), no. 1, 3 -- 14. MR 2369770
[5] N. Bourbaki, ´El´ements de math´ematique. Fascicule XXXII. Th´eories spectrales. Chapitre I: Alg`ebres norm´ees.
Chapitre II: Groupes localement compacts commutatifs, Actualit´es Scientifiques et Industrielles, No. 1332,
Hermann, Paris, 1967, (in French). MR 0213871
[6] F. Carbonell, J. C. J´ımenez, and L. M. Pedroso, Computing multiple integrals involving matrix exponentials,
J. Comput. Appl. Math. 213 (2008), no. 1, 300 -- 305. MR 2382698
[7] M. Ceballos, J. N´unez, and A. F. Tenorio, Complete triangular structures and Lie algebras, Int. J. Comput.
Math. 88 (2011), no. 9, 1839 -- 1851. MR 2810866
[8] Wai-Shun Cheung, Lie derivations of triangular algebras, Linear and Multilinear Algebra 51 (2003), no. 3,
299 -- 310. MR 1995661
[9] C. Chicone and Y. Latushkin, Evolution semigroups in dynamical systems and differential equations,
Mathematical Surveys and Monographs, vol. 70, American Mathematical Society, Providence, RI, 1999.
MR 1707332
[10] Ju. L. Daleckiı and M. G. Kreın, Stability of solutions of differential equations in Banach space, Translations
of Mathematical Monographs, vol. 43, American Mathematical Society, Providence, RI, 1974. MR 0352639
[11] Ph. I. Davies and N. J. Higham, A Schur-Parlett algorithm for computing matrix functions, SIAM J. Matrix
Anal. Appl. 25 (2003), no. 2, 464 -- 485 (electronic). MR 2047429
[12] Ch. Davis, Explicit functional calculus, Linear Algebra and Appl. 6 (1973), 193 -- 199. MR 0327792
[13] C.A. Desoer and M. Vidyasagar, Feedback systems:
input-output properties, Academic Press, New York --
London, 1975. MR 0490289
[14] L. Dieci and A. Papini, Pad´e approximation for the exponential of a block triangular matrix, Linear Algebra
Appl. 308 (2000), no. 1-3, 183 -- 202. MR 1751139
[15] M. A. Evgrafov, Analiticheskie funktsii [Analytic functions], Izdat. "Nauka", Moscow, 1965, (in Russian).
MR 0188404
[16] A. Feintuch and R. Saeks, System theory: A Hilbert space approach, Pure and Applied Mathematics, vol.
102, Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York -- London, 1982. MR 663906
[17] Richard P. Feynman, An operator calculus having applications in quantum electrodynamics, Physical Rev.
(2) 84 (1951), 108 -- 128. MR 0044379
[18] A. O. Gel′fond, Calculus of finite differences, International Monographs on Advanced Mathematics and
Physics, Hindustan Publishing Corp., Delhi, 1971, Translation of the third Russian edition. MR 0342890
[19] H. Ghahramani, Zero product determined triangular algebras, Linear and Multilinear Algebra 61 (2013),
no. 6, 741 -- 757. MR 3005653
[20] S. K. Godunov, Ordinary differential equations with constant coefficient, Translations of Mathematical Mono-
graphs, vol. 169, American Mathematical Society, Providence, RI, 1997, Translated from the 1994 Russian
original by Tamara Rozhkovskaya. MR 1465434
[21]
, Lectures on modern aspects of linear algebra, University series, vol. 12, Science book, Novosibirsk,
Russia, 2002, (in Russian).
GREEN'S FUNCTION IN THE CASE OF A TRIANGULAR COEFFICIENT
17
[22] I. Gohberg, S. Goldberg, and M. A. Kaashoek, Classes of linear operators. Vol. II, Operator Theory: Ad-
vances and Applications, vol. 63, Birkhauser Verlag, Basel, 1993. MR 1246332
[23] I. Gohberg and M. G. Kreın, Theory and applications of Volterra operators in Hilbert space, Translations of
Mathematical Monographs, vol. 24, American Mathematical Society, Providence, RI, 1970. MR 0264447
[24] D. L. Goodwin and I. Kuprov, Auxiliary matrix formalism for interaction representation transformations,
optimal control, and spin relaxation theories, The Journal of Chemical Physics 143 (2015), 084113 -- 1 -- 084113 --
7.
[25] Jin Kyu Han, Hong Youl Lee, and Woo Young Lee, Invertible completions of 2 × 2 upper triangular operator
matrices, Proc. Amer. Math. Soc. 128 (2000), no. 1, 119 -- 123. MR 1618686
[26] Ph. Hartman, Ordinary differential equations, second ed., Classics in Applied Mathematics, Society for
Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2002. MR 1929104
[27] D. Henry, Geometric theory of semilinear parabolic equations, Lecture Notes in Mathematics, vol. 840,
Springer-Verlag, Berlin -- New York, 1981. MR 610244
[28] N. J. Higham, A survey of condition number estimation for triangular matrices, SIAM Rev. 29 (1987), no. 4,
575 -- 596. MR 917696
[29] E. Hille and R. S. Phillips, Functional analysis and semi-groups, American Mathematical Society Colloquium
Publications, vol. 31, Amer. Math. Soc., Providence, Rhode Island, 1957. MR 0089373
[30] Ch. Jordan, Calculus of finite differences, third ed., Chelsea Publishing Co., New York, 1965. MR 0183987
[31] R. V. Kadison and I. M. Singer, Triangular operator algebras. Fundamentals and hyperreducible theory, Amer.
J. Math. 82 (1960), 227 -- 259. MR 0121675
[32] C. S. Kenney and A. J. Laub, A Schur-Fr´echet algorithm for computing the logarithm and exponential of a
matrix, SIAM J. Matrix Anal. Appl. 19 (1998), no. 3, 640 -- 663 (electronic). MR 1611163
[33] D. Kressner, R. Luce, and F. Statti, Incremental computation of block triangular matrix exponentials with
application to option pricing, arXiv:1703.00182 (2017).
[34] V. G. Kurbatov, Linear functional-differential equations of neutral type and the retarded spectrum, Sibirskii
Matematicheskii Zhurnal 16 (1975), no. 3, 538 -- 550, (in Russian); English translation in Siberian Mathemat-
ical Journal, 16 (1975), no. 3, 412 -- 422. MR 0402226
[35]
[36]
, On the stability of functional-diffferential equations, Differencial′nye Uravnenija 17 (1981), no. 6,
963 -- 972, (in Russian); English translation in Differential Equations, 17 (1981), no. 6, 611 -- 618. MR 620094
, Functional differential operators and equations, Mathematics and its Applications, vol. 473, Kluwer
Academic Publishers, Dordrecht, 1999. MR 1702280
[37] V. G. Kurbatov and I. V. Kurbatova, Computation of a function of a matrix with close eigenvalues by means
of the Newton interpolating polynomial, Linear and Multilinear Algebra 64 (2016), no. 2, 111 -- 122 (English).
MR 3434507
[38]
, Computation of Green's function of the bounded solutions problem, Computational Methods in Ap-
plied Mathematics (Published Online 2017-10-12, DOI: 10.1515/cmam-2017-0042).
[39] V. G. Kurbatov, I. V. Kurbatova, and M. N. Oreshina, Analytic functional calculus for two operators,
arXiv:1604.07393v (2016).
[40] W. R. LePage, Complex variables and the Laplace transform for engineers, Dover Publications, Inc., New
York, 1980, Corrected reprint of the 1961 original. MR 616824
[41] M. Lutzky, Parameter differentiation of exponential operators and the Baker -- Campbell -- Hausdorff formula,
J. Mathematical Phys. 9 (1968), 1125 -- 1128.
[42] J. L. Massera and J. J. Schaffer, Linear differential equations and function spaces, Pure and Applied Math-
ematics, Vol. 21, Academic Press, New York -- London, 1966. MR 0212324
[43] I. Najfeld and T. F. Havel, Derivatives of the matrix exponential and their computation, Adv. in Appl. Math.
16 (1995), no. 3, 321 -- 375. MR 1342832
[44] A. A. Pankov, Bounded and almost periodic solutions of nonlinear operator differential equations, Mathemat-
ics and its Applications (Soviet Series), vol. 55, Kluwer Academic Publishers, Dordrecht, 1990, Translated
from the 1985 Russian edition. MR 1120781
[45] B. N. Parlett, A recurrence among the elements of functions of triangular matrices, Linear Algebra and Appl.
14 (1976), no. 2, 117 -- 121. MR 0448846
[46] A. V. Pechkurov, Bisectorial operator pencils and the problem of bounded solutions, Izv. Vyssh. Uchebn.
Zaved. Mat. (2012), no. 3, 31 -- 41, (in Russian); English translation in Russian Math. (Iz. VUZ), 56 (2012),
no. 3, 26 -- 35. MR 3076516
[47] P. C. Rosenbloom, Bounds on functions of matrices, Amer. Math. Monthly 74 (1967), 920 -- 926. MR 0222102
[48] W. Rudin, Functional analysis, first ed., McGraw-Hill Series in Higher Mathematics, McGraw-Hill Book Co.,
New York -- Dusseldorf -- Johannesburg, 1973. MR 0365062
[49] B. van der Pol and H. Bremmer, Operational calculus. Based on the two-sided Laplace integral, third ed.,
Chelsea Publishing Co., New York, 1987. MR 904873
[50] Ch. F. Van Loan, The sensitivity of the matrix exponential, SIAM J. Numer. Anal. 14 (1977), no. 6, 971 -- 981.
MR 0468137
18
[51]
V.G. KURBATOV AND I.V. KURBATOVA
, Computing integrals involving the matrix exponential, IEEE Trans. Automat. Control 23 (1978),
no. 3, 395 -- 404. MR 0494865
[52] R. M. Wilcox, Exponential operators and parameter differentiation in quantum physics, J. Mathematical
Phys. 8 (1967), 962 -- 982. MR 0234689
[53] J. C. Willems, The analysis of feedback systems, The MIT Press, Cambridge, 1971.
E-mail address: [email protected]
Department of Mathematical Physics, Voronezh State University, 1, Universitetskaya Square,
Voronezh 394018, Russia
E-mail address: la [email protected]
Department of Software Development and Information Systems Administration, Voronezh State
University, 1, Universitetskaya Square, Voronezh 394018, Russia
|
1606.02125 | 1 | 1606 | 2016-06-07T12:57:25 | Uncertainty Principles of Ingham and Paley-Wiener on Semisimple Lie Groups | [
"math.FA"
] | Classical results due to Ingham and Paley-Wiener characterize the existence of nonzero functions supported on certain subsets of the real line in terms of the pointwise decay of the Fourier transforms. Viewing these results as uncertainty principles for Fourier transforms, we prove certain analogues of these results on connected, noncompact, semisimple Lie groups with finite center. We also use these results to show unique continuation property of solutions to the initial value problem for time-dependent Schr\"odinger equations on Riemmanian symmetric spaces of noncompact type. | math.FA | math |
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON
SEMISIMPLE LIE GROUPS
MITHUN BHOWMIK AND SUPARNA SEN
Abstract. Classical results due to Ingham and Paley-Wiener characterize the existence of
nonzero functions supported on certain subsets of the real line in terms of the pointwise decay
of the Fourier transforms. Viewing these results as uncertainty principles for Fourier transforms,
we prove certain analogues of these results on connected, noncompact, semisimple Lie groups
with finite center. We also use these results to show unique continuation property of solutions to
the initial value problem for time-dependent Schrodinger equations on Riemmanian symmetric
spaces of noncompact type.
1. Introduction
It is a well known fact in harmonic analysis that if the Fourier transform of an integrable
function on R is very rapidly decreasing then the function can not be compactly supported
unless it vanishes identically. A manifestation of this fact is as follows: Let f ∈ L1(R) and a > 0
be such that
bf (ξ) ≤ Ce−aξ,
for all ξ ∈ R.
If f is compactly supported then f is identically zero. This holds due to the fact that the
very rapid decay of the Fourier transform imposes real analyticity on the function. In fact, f
extends to a holomorphic function on an open subset of C. This initial observation motivates
one to endeavour for a more optimal decay of the Fourier transform bf for such a conclusion.
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 be compactly supported without being identically zero? A more precise
question could be: is there a nonzero integrable compactly supported function f on R with its
Fourier transform satisfying
(1.1)
log ξ ,
for large ξ?
bf (ξ) ≤ Ce− ξ
The answer to the above question is in the negative and follows from classical results due to
Paley-Wiener ([28], Theorem II; [27], P. 16, Theorem XII) and Ingham [21] (see also Theorem
2010 Mathematics Subject Classification. Primary 22E30; Secondary 22E46, 43A80.
Key words and phrases. uncertainty principle, semisimple Lie group, symmetric space, Schrodinger equation.
This work was supported by Indian Statistical Institute, India (Research fellowship to Mithun Bhowmik) and
Department of Science and Technology, India (INSPIRE Faculty Award to Suparna Sen).
1
2
MITHUN BHOWMIK AND SUPARNA SEN
1.1). These results may be viewed as instances of uncertainty principles in harmonic analysis.
We refer the reader to [14, 16, 35] for the literature on uncertainty principles.
Despite being results of same genre the treatment of Paley-Wiener and Ingham mentioned
above have intriguing differences: while Paley-Wiener's method is complex analytic, Ingham
relies on the notion of quasianalytic functions and the celebrated Denjoy-Carleman theorem
([30], Theorem 19.11). We would like to point out that the result of Paley and Wiener leads to
a proof of the Denjoy-Carleman theorem. In some sense the result of Ingham is stronger than
that of Paley and Wiener. In fact, a close examination of the proof of Ingham's result (as given
in [21]) reveals that if bf satisfies (1.1) and f vanishes on an open set then f actually vanishes
identically. Different versions of the result of Ingham was later proved by Levinson and Beurling
[24, 25, 23]. All these results mentioned above are obtained only for the circle group and real
line. Only very recently we have obtained the following analogues of the results of Paley-Wiener
and Ingham in the context of n-dimensional Euclidean spaces (see Theorem 2.3 and Theorem
2.2 of [2] respectively). For f ∈ L1(Rn), we shall define its Fourier transform bf by
f (x) e−ix·ξ dx,
for ξ ∈ Rn.
Theorem 1.1. Let ψ : [0, ∞) → [0, ∞) be a locally integrable function and f ∈ L1(Rn) be such
that
(1.2)
Let
(1.3)
bf (ξ) =ZRn
bf (ξ) ≤ Ce−ψ(kξk),
I =Z ∞
0
for all ξ ∈ Rn.
ψ(r)
1 + r2 dr.
(a) If f ∈ Cc(Rn) and I = ∞ then f is identically zero on Rn. Conversely, if I < ∞ and ψ
is non-decreasing then there exists a nonzero function f ∈ C ∞
c (Rn) satisfying (1.2).
(b) Let ψ(r) = rθ(r), for r ∈ [0, ∞), where θ : [0, ∞) → [0, ∞) is a decreasing function
with limr→∞ θ(r) = 0. If f vanishes on an open set in Rn and I = ∞ then f is zero.
c (Rn) satisfying (1.2).
Conversely, if I < ∞ then there exists a nonzero function f ∈ C ∞
Part (a) of the above theorem corresponds to the result of Paley-Wiener and part (b) corresponds
to that of Ingham.
Viewing Rn as a noncompact Riemannian symmetric space one can naturally ask about
analogues of the above results for Riemannian symmetric spaces of noncompact type or more
generally for connected, noncompact, semisimple Lie groups with finite centre. In these cases nice
parametrization of relevant representations are available and consequently the natural domain of
the Fourier transform turns out to be R or Rn. This enables us to formulate analogous questions
for these spaces. In this paper we will first prove an analogue of the result of Paley-Wiener for
connected, noncompact semisimple Lie groups with finite centre and arbitrary rank (Theorem
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
3
3.5). The main ingredient here is Lemma 3.3 whose proof was inspired by a related result of
Hirschmann [20].
However, an analogue of Ingham's result poses much greater difficulty. This is due to the
unavailability of a suitable analogue of the Denjoy-Carleman theorem for Riemannian manifolds.
Certain analogues of the Denjoy-Carleman theorem for Riemannian manifolds were proved by
Bochner and Taylor [4, 5] but these results don't seem to be suitable for our purpose. We
therefore restrict ourselves to bi-K-invariant functions on complex semisimple Lie groups and
obtain an analogue of Ingham's theorem (Theorem 3.9). The main idea here is to use the explicit
description of elementary spherical functions and reduce matters to Euclidean spaces.
Our next set of results deal with Riemannian symmetric spaces of noncompact type. Our aim
here is to relate the results of Paley-Wiener and Ingham to the problem of unique continuation
of solution to the Schrodinger equation in the spirit of [12, 13, 7, 29]. These results present
some pleasant surprise in the sense that the expected results demand more decay than that is
required in Theorem 3.5 and Theorem 3.9 (see the beginning of §4.2.1 and Remark 4.4). This
can be attributed to the exponential volume growth of the invariant measure on the symmetric
space. Taking this into account we could obtain unique continuation properties of the solution
to the Schrodinger equation on symmetric spaces (Theorem 4.3, Theorem 4.5).
This paper is organized as follows: In the next section we describe the required preliminaries
on connected, noncompact, real semisimple Lie groups with finite centre and the associated
symmetric spaces. In Section 3 we shall prove an analogue of Theorem 1.1 (a) for f ∈ Cc(G)
where G is a connected, noncompact, real semisimple Lie group with finite centre. We also prove
an analogue of Theorem 1.1 (b) on a noncompact, complex semisimple Lie group. In the last
section we consider unique continuation properties of solutions to the initial value problem for
time-dependent Schrodinger equation on Riemannian symmetric spaces of the noncompact type.
First we prove such a result for bi-K-invariant initial value on noncompact, complex semisimple
Lie groups related to Theorem 1.1 (b). We also prove such a result on Riemannian symmetric
spaces of the noncompact type related to Theorem 1.1 (a).
2. Notation and Preliminaries
In this section, we shall discuss the preliminaries and notation related to the noncompact
semisimple Lie groups and the associated symmetric spaces. These are standard and can be
found, for example, in [15, 17, 18, 19, 22]. To make the article self-contained, we shall gather
only those results which will be used throughout this paper.
We shall use the following notation in this paper: Cc(X) denotes the set of compactly sup-
ported continuous functions on X, C ∞
c (X) denotes the set of compactly supported smooth
functions on X and C denotes a constant whose value may vary. For x, y ∈ Rn, we shall use kxk
to denote the Euclidean norm of the vector x and x · y to denote the Euclidean inner product
4
MITHUN BHOWMIK AND SUPARNA SEN
of the vectors x and y. We shall also denote the Euclidean norm in Cn by k · k. ℑ(λ) denotes
the imaginary part of λ ∈ Cn.
Let G be a connected, noncompact, real semisimple Lie group with finite centre and K be
a fixed maximal compact subgroup of G. Let g and k denote the Lie algebras of G and K
respectively. Suppose that B is the Cartan Killing form on g and g = k ⊕ p is a Cartan
decomposition of g. It is known that B restricted to p is positive definite, thus it gives an inner
product and hence a norm k · kB on p. Fix a maximal abelian subspace a of p. If the dimension
of a is l, then G is said to be of real rank l. We can identify a with Rl endowed with the inner
product induced from p. Let R denote the set of nonzero roots for the adjoint action of a on g
and W denote the Weyl group corresponding to R. Fix a Weyl chamber a+ of a and let R+ be
the corresponding set of positive roots.
Let A = exp a and A+ = exp a+. If A+ denotes the closure of A+ in G, then one gets the
polar decomposition G = KA+K, that is, each g ∈ G can be uniquely written as
g = k1ak2,
for k1, k2 ∈ K and a ∈ A+.
Using the polar decomposition k · kB on G is defined by
kgkB = kk1ak2kB = k log akB,
where log a denotes the unique element in a such that exp(log a) = a. Let gα denote the root
space corresponding to α ∈ R with mα = dim gα. The Haar measure dg on G relative to the
polar decomposition is given by dg = J(a) dk1 da dk2, that is, for any suitable function f on G,
we have
f (k1ak2) J(a) dk1 da dk2,
f (g)dg =ZKZA+ZK
ZG
J(a) = Yα∈R+(cid:16)eα(log a) − e−α(log a)(cid:17)mα
where dk is the normalised Haar measure on K, da is the Lebesgue measure on A ∼= Rl and
,
for a ∈ A+.
We define the half-sum of the elements of R+ counted with their multiplicities by ρ given by
ρ =
1
2 Xα∈R+
mαα.
Thus we get the trivial estimate
(2.1)
J(a) ≤ Ce2ρ(log a),
for a ∈ A+.
For an element λ in the dual a∗ of a, let Hλ be the unique element in a such that
λ(H) = B(H, Hλ),
for all H ∈ a.
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
5
The map λ 7→ Hλ identifies a∗ with a, and we use it to define the dual inner product on a∗, also
denoted B(·, ·), by the formula
B(λ, µ) = B(Hλ, Hµ),
for λ, µ ∈ a∗.
The elements of the Weyl group W are orthogonal transformations of a∗ which correspond to
orthogonal transformations of a by the formula
(2.2)
sHλ = Hsλ,
for s ∈ W and λ ∈ a∗.
The bilinear extension of B(·, ·) to a∗
by B(·, ·).
C, the dual of the complexification aC of a, is also denoted
We now describe the principal series representations of G. The Iwasawa decomposition G =
KAN gives rise to the projection mappings κ : G → K, H : G → a and η : G → N such that
g = κ(g) exp H(g)η(g).
dimensional Hilbert space on which ξ is realised. We define the Hilbert space Hξ by
Let M be the centraliser of A in K. Given ξ in the unitary dual cM of M let H be the finite
kφ(k)k2dk < ∞(cid:27) ,
(cid:26)φ : K → H measurable, φ(km) = ξ(m−1)φ(k) for k ∈ K, m ∈ M and ZK
where k · k denotes the norm on H induced from the inner product h·, ·i on H. The Hilbert space
Hξ is equipped with the inner product
hφ, ψiξ =ZK
hφ(k), ψ(k)i dk.
C we have a representation πξ,λ acting on the Hilbert space Hξ given by
It is known that πξ,λ is unitary if λ ∈ a∗ and πξ,λ, πξ,µ are unitarily equivalent if and only if
(πξ,λ(g)φ) (k) = e(iλ−ρ)H(g−1k)φ(cid:0)κ(g−1k)(cid:1) ,
For ξ ∈ cM and λ ∈ a∗
λ = sµ where s ∈ W . Moreover, given ξ ∈ cM there exists a dense open subset Oξ ⊂ a∗ such
that for λ ∈ Oξ, πξ,λ is irreducible. We now define the group Fourier transform of f ∈ L1(G)
given by the operator valued integral
for g ∈ G, k ∈ K, φ ∈ Hξ.
πξ,λ(f ) =ZG
f (g)πξ,λ(g)dg,
for (ξ, λ) ∈ cM × a∗.
For f ∈ L1∩L2(G), πξ,λ(f ) is a Hilbert-Schmidt operator and we shall denote its Hilbert-Schmidt
norm by kπξ,λ(f )kHS.
If ξ is the trivial representation in cM , then we denote πξ,λ by πλ and the set of representations
{πλ}λ∈a∗ , realized on the Hilbert space L2(K/M ), are called the class one principal series repre-
sentations of G. We observe that πλK are given by left translations on L2(K/M ), in particular,
the K-fixed vectors are given by constant functions. The right-K-invariant functions on G can
be viewed as functions on the symmetric space X = G/K and vice-versa. For the harmonic
6
MITHUN BHOWMIK AND SUPARNA SEN
analysis of a function f on X, only the class one principal series representations πλ are relevant
and πλ(f ) is completely determined by πλ(f )e0 where e0 denotes the constant function 1 on
K/M . In this case, the group theoretic Fourier transform can be reinterpreted as the Helgason-
Fourier transform on the symmetric space X, as introduced by Helgason. For a sufficiently nice
C × K/M given by
function f on X, its Helgason-Fourier transform ef is a function defined on a∗
f (g)e(iλ−ρ)H(g−1k)dg,
(2.3)
C, kM ∈ K/M whenever this integral exists (see [18], Ch. III, §1). Thus it follows that
for λ ∈ a∗
(2.4)
ZK/M
(πλ(f )e0)(kM ) = ef (λ, kM ) =ZG
ef (λ, kM )2dk = kπλ(f )e0k2
+ZK/M
ZX
f (x)2dx =Za∗
The Plancherel theorem states that the Helgason Fourier transform extends to an isometry of
L2(X) onto L2(a∗
+ × K/M, c(λ)−2dk dλ) where c(λ) is Harish-Chandra's c-function, that is,
L2(K/M ) = kπλ(f )k2
HS ,
for λ ∈ a∗.
ef (λ, kM )2c(λ)−2dk dλ.
Using the polar decomposition of G we may view a bi-K-invariant function f on G as that
depending only on its values on A+, or by using the inverse exponential map we may also view
f as a function on a solely determined by its values on a+. Henceforth we shall denote the set of
bi-K-invariant functions in L1(G) by L1(K\G/K). If f ∈ L1(K\G/K) then πλ(f ) is determined
by hπλ(f )e0, e0i and the spherical transform ef (λ) of f is defined by
f (g)φλ(g)dg = hπλ(f )e0, e0i,
for λ ∈ a∗,
where φλ is the elementary spherical function corresponding to λ ∈ a∗ given by
(2.5)
e(iλ−ρ)H(g−1k)dk,
for g ∈ G.
ef (λ) =ZG
φλ(g) = hπλ(g)e0, e0i =ZK
It is clear that the elementary spherical function φλ is a bi-K-invariant function on G for any
λ ∈ a∗. The following estimates regarding φλ are well known (see [15], §4.6).
(2.6)
e−ρ(log a) ≤ φ0(a) ≤ C(1 + k log akB)me−ρ(log a),
for a ∈ A+, some m, C > 0,
(2.7)
0 < φiλ(a) ≤ eλ+(log a)φ0(a),
for a ∈ A+, λ ∈ a∗,
where λ+ is the element in the fundamental Weyl chamber corresponding to λ. In this case
(2.8)
kπλ(f )kHS = hπλ(f )e0, e0i = ef (λ),
for λ ∈ a∗.
Apart from Fourier transform, we will also need the notion of Radon transform on Riemannian
symmetric space X. In the following, we shall describe the properties of Radon transform which
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
7
will be used in this paper. By identifying the space G/M N of horocycles on X with (K/M ) × A,
we define the Radon transform of a sufficiently regular function f : X → C by
(2.9)
Rf (kM, a) = eρ(log a)ZN
f (kan · o) dn
for all kM ∈ K/M , a ∈ A whenever this integral exists (see [18], P. 220, see also [31]) and
o = eK denotes the identity coset in X. It is known that the Helgason-Fourier transform of a
sufficiently regular function f is the Euclidean Fourier transform of the Radon transform (see
[18], P. 219) given by
(2.10)
ef (λ, kM ) = FA(Rf (kM, ·))(λ),
where FA denotes the Euclidean Fourier transform of h ∈ L1(A) defined by
for almost all kM ∈ K/M and all λ ∈ a∗,
(FAh)(λ) =ZA
h(a)e−iλ(log a)da,
for λ ∈ a∗ ∼= Rl.
It is well known that Radon transform is injective on L1(X) (see [18], Ch. II, Theorem 3.2).
For f ∈ L1(K\G/K), we define the Abel transform by the following integral
Af (a) = eρ(log a)ZN
f (an · o) dn,
for a ∈ A,
(see [18], P. 381). Note that the Radon transform of a bi-K-invariant function reduces to the
Abel transform. It follows from (2.10) that the spherical transform of f ∈ L1(K\G/K) is the
Euclidean Fourier transform of the Abel transform Af given by
(2.11)
ef (λ) = FA(Af )(λ),
for λ ∈ a∗ ∼= Rl.
This property of Abel transform is crucial for reducing some questions on bi-K-invariant func-
tions defined on a semisimple Lie group G to related questions on Rl. Moreover, it is known
that the Abel transform A induces a bijection between C ∞
c (K\G/K) and the set of W -invariant
functions in C ∞
c (A) (see [18], Ch. IV, Theorem 4.1).
3. Uncertainty Principles
In this section we shall present analogues of the results of Paley-Wiener and Ingham, namely
Theorem 1.1 (a) and Theorem 1.1 (b) on connected, noncompact, real semisimple Lie groups
with finite centre and noncompact, complex semisimple Lie groups respectively.
3.1. Real Semisimple Lie Groups. To prove the first theorem we shall prove a lemma on
entire functions on Cn. The proof of this lemma for the one-variable case is given in [3]. Since
the paper is yet to be accepted, we reproduce the proof here for the sake of completeness. To
prove this lemma, we shall need two results. The first one is regarding upper semicontinuous
functions.
8
MITHUN BHOWMIK AND SUPARNA SEN
Theorem 3.1 ([8], P. 218, Theorem 3.6). Let (X, d) be a metric space, v : X → [−∞, ∞)
be upper semi-continuous and v ≤ M < ∞ on X. Then there exists a decreasing sequence
of uniformly continuous functions {fn} on X such that fn ≤ M and for every x ∈ X, fn(x)
decreases to v(x).
The second result we need is an analogue of the maximum modulus principle on unbounded
domain for subharmonic functions. We briefly recall the definition of subharmonic functions.
Let D be an open subset of C. A function u : D → [−∞, ∞) is called subharmonic if u is upper
semi-continuous and satisfies the local submean inequality, that is, for any w ∈ D there exists
ρ > 0 such that for all r ∈ (0, ρ) the following holds
u(w) ≤
1
2πZ 2π
0
u(w + reit)dt.
It is well known that if f is a holomorphic function then g(z) = log(f (z)) is a subharmonic
function (see [30], P. 336).
Theorem 3.2 ([6], P. 224, Theorem 7.15). Let Ω be a region (not necessarily bounded) and
u : Ω → R be a subharmonic function which is bounded above. Let A be a proper, countable subset
of the boundary ∂Ω of Ω and M a finite constant such that lim
u(z) ≤ M for all ξ ∈ ∂Ω r A.
z→ξ
Then u ≤ M throughout Ω.
We shall now present the required lemma on entire functions.
In the following we shall
interpret a radial function on Rn as an even function on R.
Lemma 3.3. Let f be an entire function on Cn and ψ be a non-negative, radial, locally integrable
function on Rn such that for positive constants a and C, we have
(3.1)
(3.2)
If
f (z) ≤ Ceakzk,
f (x) ≤ Ce−ψ(x),
for all z ∈ Cn,
for all x ∈ Rn.
ZR
ψ(r)
1 + r2 dr = ∞,
then f is identically zero on Cn.
Proof. First we shall prove the result for n = 1. We consider the function
g(z) =
1
C
eiazf (z),
for all z ∈ C,
and observe that g is an entire function. We want to apply Phragm´en-Lindelof theorem to show
that for all z in the closed upper half plane H = {z ∈ C : ℑz ≥ 0}
(3.3)
g(z) ≤ 1,
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
9
(see [32], P. 124, Theorem 3.4). Let Q1 = {z = x + iy ∈ C : x > 0, y > 0}. It follows from the
estimate (3.1) that
g(iy) =
1
C
e−ayf (iy) ≤ e−ayeay = 1,
for all y > 0.
It is also immediate from (3.2) that
g(x) =
1
C
f (x) ≤ e−ψ(x) ≤ 1,
for x ∈ R.
In particular, g is bounded by 1 on the positive real and positive imaginary axes. As g satisfies
the estimate (3.1) we can apply the Phragm´en-Lindelof theorem to the sector Q1 to obtain (3.3).
A similar argument for the quadrant Q2 = {z = x + iy ∈ C : x < 0, y > 0} proves the estimate
(3.3) for all z ∈ H. Since g is an entire function, log g is a subharmonic function on C and
(3.4)
log g(z) ≤ 0,
for all z ∈ H.
Now, we apply Theorem 3.1 for X = H, v = log g, and M = 0. Then there exists a decreasing
sequence of uniformly continuous functions fn on H such that fn ≤ 0 and fn(z) decreases to
v(z) for every z ∈ H. We define
vn(z) = max{fn(z), −n},
for z ∈ H, n ∈ N.
It is clear that {vn} is a decreasing sequence of continuous functions. As fn takes only negative
values it follows that vn(z) ∈ [−n, 0] for all z ∈ H. In particular, vn is bounded for each n ∈ N.
We now claim that {vn} converges pointwise to v on H. We first assume that v(z) = −∞. Then
{fn(z)} and {−n} both converge to −∞ and hence so does {vn(z)}. Now assume that v(z) is
finite. In this case fn(z) ∈ (v(z) − 1, v(z) + 1) for all large n and hence vn(z) = fn(z) for all large
n ∈ N. It follows that {vn(z)} converges to v(z) for all z ∈ H. Let Un be the Poisson integral of
the restriction of the function vn on R given by
Since vn ∈ L∞(R), the above integral exists and defines a harmonic function on the open upper
half plane H = {z ∈ C : ℑz > 0}. Moreover, since vn is continuous, we can extend Un to H as a
continuous function by letting Un(x) = vn(x) for x ∈ R ([33], P. 47, Theorem 2.1(b)). We now
define
(3.5)
Un(x + iy) =
yvn(t)
y2 + (x − t)2 dt,
for x ∈ R, y > 0.
1
πZR
(3.6)
Vn(z) = log g(z) − Un(z),
for z ∈ H.
Since Un is harmonic, we get that Vn is subharmonic on H. Since
vn(t) ≥ −n,
for all t ∈ R,
it follows from the definition of Un given in (3.5) that
Un(z) ≥ −n,
for z ∈ H.
10
MITHUN BHOWMIK AND SUPARNA SEN
Using (3.4) we get that
In particular, Vn is bounded above for each n ∈ N. Since
Vn(z) ≤ n,
for all z ∈ H.
v(z) = log g(z) ≤ vn(z),
for all z ∈ H,
it follows that
lim
y→0
Vn(x + iy) = log g(x) − vn(x) ≤ 0,
for all x ∈ R.
We now apply Theorem 3.2 for Ω = H, u = Vn and A = φ, the empty set to conclude that
Vn(z) ≤ 0,
for all z ∈ H.
It follows from (3.6) that
log g(x + iy) ≤
1
πZR
yvn(t)
y2 + (x − t)2 dt,
for all y > 0, x ∈ R.
Since {vn} is a decreasing sequence, by using monotone convergence theorem and taking limit
as n → ∞ in the inequality above we get
log g(x + iy) ≤
1
πZR
y log g(t)
y2 + (x − t)2 dt.
The estimate (3.2) now implies that
log g(x + iy) ≤ −
1
πZR
yψ(t)
y2 + (x − t)2 dt ≤ −Cx,yZR
ψ(t)
1 + t2 dt = −∞,
where Cx,y is a positive constant which depends on x and y. So, for each x ∈ R and y positive,
it follows that g(x + iy) = 0. As f is an entire function it follows that f (z) = 0 for all z ∈ C.
This proves the lemma for n = 1.
Now, we shall prove the case n > 1. Let ξ ∈ Rn−1 be fixed and we define
g(z) = f (ξ, z),
for z ∈ C,
which is clearly an entire function on C. Using (3.1) we have for any z ∈ C
g(z) = f (ξ, z) ≤ Ceak(ξ,z)k = Ceak(ξ,0)+(0,z)k ≤ Cξeaz.
Using (3.2) we get that for any x ∈ R
g(x) = f (ξ, x) ≤ e−ψ(ξ,x) = e−ψξ(x),
where the function ψξ on R is defined as
ψξ(x) = ψ(ξ, x),
for x ∈ R.
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
11
Since ψ is radial, ψξ is an even function on R. Moreover using the radiality of ψ and the change
of variable kξk2 + x2 = r2 we have
ZR
ψξ(x)
1 + x2 dx = ZR
= 2Z ∞
≥ 2Z ∞
kξk
kξk
= ∞.
ψ(pkξk2 + x2)
1 + x2
dx
ψ(r)
1 + r2 − kξk2
ψ(r)
1 + r2 dr
dr
r
pr2 − kξk2
Applying the lemma for the case n = 1 on g, we get that g is zero on C. So it follows that
f (ξ, z) is zero for all z ∈ C. Since this is true for all ξ ∈ Rn−1 we obtain that f is identically
zero on Rn. Since f is an entire function on Cn which vanishes on Rn, we can conclude that f
vanishes on Cn.
(cid:3)
(3.7)
We shall need few other facts to prove our first main theorem. Let G be a connected, non-
compact, semisimple Lie group with finite centre. We shall continue to assume the notation
j : j ∈ N} be an orthonormal basis of Hξ
introduced in the previous section. For ξ ∈ cM let {eξ
consisting of K-finite vectors. For g ∈ G, λ ∈ a∗ and m, n ∈ N we define
Φm,n
ξ,λ (g) =Dπξ,λ(g)eξ
m, eξ
nEξ
=ZK
e(iλ−ρ)H(g−1k)Deξ
m(cid:0)κ(g−1k)(cid:1) , eξ
n(k)E dk.
m, eξ
The basis vectors eξ
n ∈ Hξ being K-finite, actually belong to C ∞(K, H) and hence are
bounded as functions into H. Therefore it follows that for each g ∈ G the integral defining
Φm,n
ξ,λ (g) is continuous and the
function λ 7→ Φm,n
∼= Cl. From (3.7) and (2.5) we
get the following easy estimate
ξ,λ (g) extends as an entire function of λ ∈ a∗
C. Moreover the function g 7→ Φm,n
ξ,λ (g) makes sense for λ ∈ a∗
C
Using the bi-K-invariance of φiℑ(λ) and (2.7) we get that
e−(ℑ(λ)+ρ)H(g−1k)dk = Cφiℑ(λ)(g),
ξ,λ (g)(cid:12)(cid:12)(cid:12) ≤ CZK
(cid:12)(cid:12)(cid:12)Φm,n
(cid:12)(cid:12)(cid:12)Φm,n
ξ,λ (g)(cid:12)(cid:12)(cid:12) ≤ Ceℑ(λ)+(log a)φ0(a) ≤ Cekℑ(λ)kB k log akB φ0(a),
for g ∈ G.
for g ∈ G,
where g = k1ak2; k1, k2 ∈ k and a ∈ A+.
(3.8)
We have already defined πξ,λ for all λ ∈ a∗
C and noted that πξ,λ is a unitary operator on
the Hilbert space Hξ for λ ∈ a∗. In general for λ ∈ a∗
C \ a∗, the representation πξ,λ may not
be unitary. However for f ∈ Cc(G), it is easy to see that πξ,λ(f ) makes sense for all λ ∈ a∗
C
and ξ ∈ cM . We shall also need the following lemma which is essentially proved in [10] using
Harish-Chandra's subquotient theorem.
12
MITHUN BHOWMIK AND SUPARNA SEN
Lemma 3.4. Let f ∈ Cc(G). If πξ,λ(f ) = 0 for all λ ∈ a∗
Now we shall present the first main result, which is an analogue of Theorem 1.1 (a).
Theorem 3.5. Let ψ : [0, ∞) → [0, ∞) be a locally integrable function such that
C and ξ ∈ cM then f is zero.
I =Z ∞
0
ψ(r)
1 + r2 dr.
(a) Suppose f ∈ Cc(G) satisfies the estimate
kπξ,λ(f )kHS ≤ Cξe−ψ(kλkB ),
(3.9)
for ξ ∈ cM , λ ∈ a∗ ∼= Rl,
where Cξ is a positive constant depending on ξ. If I = ∞ then f is zero on G.
(b) If I is finite and ψ is nondecreasing then there exists a nontrivial f ∈ C ∞
c (K\G/K)
satisfying the estimate (3.9).
Proof. First we shall prove (a). For fixed ξ ∈ cM and m, n ∈ N we define
f (g) Φm,n
ξ,λ (g)dg,
F m,n
ξ
m, eξ
(3.10)
for λ ∈ a∗.
(λ) =Dπξ,λ(f )eξ
nEξ
=ZG
C, it follows that F m,n
Since f ∈ Cc(G) and Φm,n
of λ ∈ a∗
is enough to prove that for each fixed m, n ∈ N, F m,n
decomposition, (3.10) and (3.8), it follows that
ξ,λ (g) is a continuous function of g ∈ G as well as an entire function
C. From Lemma 3.4, it
C. Using polar
extends as an entire function of λ ∈ a∗
(λ) is zero for all λ ∈ a∗
ξ
ξ
(3.11) F m,n
ξ
(λ) ≤ZKZa+ZK
f (k1(exp H)k2) ekℑ(λ)kB kHkB φ0(exp H) J(exp H) dk1 dH dk2.
Since f ∈ Cc(G), it follows from (3.11), (2.6) and (2.1) that there exists a constant γ > 0 such
that
(3.12)
F m,n
ξ
(λ) ≤ Cf eγkℑ(λ)kB ≤ Cf eγkλkB ,
for all λ ∈ a∗
C
∼= Cl.
Now by (3.9) and the fact that
we get that
(3.13)
m, eξ
(cid:12)(cid:12)(cid:12)hπξ,λ(f )eξ
niξ(cid:12)(cid:12)(cid:12) ≤ kπξ,λ(f )kHS ,
for λ ∈ a∗,
F m,n
ξ
(λ) ≤ Cξe−ψ(kλkB ),
for all λ ∈ a∗ ∼= Rl.
ξ
Since all norms are equivalent on finite dimensional spaces, (3.12), (3.13) and Lemma 3.3 implies
that F m,n
vanishes on a∗
C
3.4 that f vanishes on G.
∼= Cl. By varying m, n over N and ξ over cM it now follows from Lemma
Now we shall prove (b). Since I is finite and ψ is nondecreasing, by Theorem 1.1 (a) there
c (Rl) satisfying (1.2). Since f0 is radial on Rl, it can
c (K\G/K) such that
exists a nontrivial radial function f0 ∈ C ∞
be thought of as a W -invariant function on A ∼= Rl. So there exists f ∈ C ∞
Af = f0. Using (2.8) and (2.11) it follows that f ∈ C ∞
c (K\G/K) satisfies the estimate (3.9).
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
13
(cid:3)
Remark 3.6. If we think of a function on X = G/K as a right-K-invariant function on G then
the following theorem is, in view of the relation (2.4), an easy corollary of Theorem 3.5.
Theorem 3.7. Let ψ and I be as in Theorem 3.5.
(a) Let f ∈ Cc(X) and the Helgason Fourier transform ef of f satisfies the estimate
for λ ∈ a∗, kM ∈ K/M.
(3.14)
If I = ∞ then f = 0.
ef (λ, kM ) ≤ Ce−ψ(kλkB ),
(b) If I is finite and ψ is nondecreasing then there exists a nontrivial K-invariant f ∈ C ∞
c (X)
satisfying the estimate (3.14).
3.2. Complex Semisimple Lie Groups. We shall prove an analogue of Theorem 1.1 (b) for
bi-K-invariant functions on a noncompact, complex semisimple Lie group using the explicit
expression of the elementary spherical functions φλ available in this case. First we shall recall
the following result proved in [2] which is needed in our proof.
Lemma 3.8. Let P be a polynomial on Rn and θ : Rn → [0, ∞) be a decreasing radial function
with limkξk→∞ θ(ξ) = 0 and
(a) Let f ∈ L1(Rn) be a nontrivial function satisfying the estimate
θ(ξ)
kξkn dξ.
I =Zkξk≥1
bf (ξ) ≤ CP (ξ)e−kξkθ(ξ),
for all ξ ∈ Rn.
If f vanishes on a nonempty open set then I is finite.
(b) If I is finite then there exists a nontrivial f ∈ C ∞
c (Rn) satisfying (3.15).
If G is noncompact, complex semisimple, it is known that the Haar measure on A correspond-
ing to the polar decomposition is given by φ(H)2dH for H ∈ a where φ(H) is defined by the
formula
(3.15)
(3.16)
(3.17)
φ(H) = Xs∈W
det(s) esρ(H),
for H ∈ a,
(see [9], P. 907 and P. 910). It is also known that on a noncompact, complex semisimple Lie
group, the elementary spherical functions are given by the expression
φλ(H) = c(λ)Ps∈W det(s)e−isλ(H)
φ(H)
,
for H ∈ a, λ ∈ a∗,
and the function λ 7→ c(λ)−1, for λ ∈ a∗, is of polynomial growth ([19], P. 432, Theorem 5.7).
Thus for f ∈ L1(K\G/K) the spherical transform ef of f is given by the integral
f (H)φλ(H)φ(H)2dH,
for λ ∈ a∗.
(3.18)
ef (λ) =Za
14
MITHUN BHOWMIK AND SUPARNA SEN
Theorem 3.9. Let G be a connected, noncompact, complex semisimple Lie group and θ :
[0, ∞) → [0, ∞) be a decreasing function with limr→∞ θ(r) = 0 and
(a) Let f ∈ L1(K\G/K) satisfy the estimate
I =Z ∞
1
θ(r)
r
dr.
(3.19)
ef (λ) ≤ Ce−kλkB θ(kλkB ),
for λ ∈ a∗ ∼= Rl.
If f vanishes on a nonempty open set in G and I = ∞ then f = 0.
(b) If I is finite then there exists a nontrivial f ∈ C ∞
c (K\G/K) satisfying the estimate
(3.19).
Proof. First we shall prove (a). Using (3.17) and (2.2) the spherical transform of f can be
written as
ef (λ) = c(λ)Xs∈W
= c(λ)Xs∈W
= c(λ)Xs∈W
where the function g on a is defined as
det(s)Za
det(s)Za
det sbg(sHλ),
e−isλ(H)f (H)φ(H)dH
e−iB(H,sHλ)f (H)φ(H)dH
for λ ∈ a∗,
g(H) = f (H)φ(H),
for H ∈ a,
andbg denotes the Euclidean Fourier transform of g. As f ∈ L1(K\G/K) it follows that
Consequently bg is well defined. Since f is bi-K-invariant and φ is odd under the action of the
g(H)dH =Za
Weyl group, that is,
f (H)φ(H)dH < ∞.
Za
φ(sH) = det(s)φ(H),
for H ∈ a, s ∈ W,
we get that
g(sH) = det(s)g(H),
for H ∈ a, s ∈ W.
It follows thatbg is also odd under the action of the Weyl group, that is,
for λ ∈ a∗, s ∈ W.
Consequently the spherical transform of f can be written as
bg(sHλ) = det(s)bg(Hλ),
ef (λ) = c(λ)W bg(Hλ),
for λ ∈ a∗,
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
15
where W denotes the number of elements in the Weyl group W . From the estimate of ef given
in (3.19) it follows that
bg(Hλ) ≤ Cc(λ)−1e−θ(kλkB )kλkB ,
for λ ∈ a∗ ∼= Rl.
Since c(λ)−1 is of polynomial growth, using the equivalence of all norms on finite dimensional
spaces we can apply Lemma 3.8 to the function g on a to get that g is zero. Hence it follows
that f is zero.
To prove (b), first we note that by Theorem 1.1 (b) we get a nontrivial radial function f0 ∈
c (Rl) satisfying (1.2). Then we proceed as in Theorem 3.5 (b) to construct f ∈ C ∞
c (K\G/K)
(cid:3)
C ∞
satisfying the estimate (3.19).
4. Unique Continuation Property of Solutions to the Schrodinger Equation
We consider the initial value problem for the time-dependent Schrodinger equation on Rn
given by
∂u
∂t
(x, t) − i∆u(x, t) = 0,
for (x, t) ∈ Rn × R,
u(x, 0) = f (x),
for x ∈ Rn.
Our aim is to obtain sufficient conditions on the behaviour of the solution u at two different
times t = 0 and t = t0 which guarantee that u ≡ 0 is the unique solution of the above equation.
It has recently been observed that uncertainty principles can be used to obtain such sufficient
conditions. We refer the reader to [13] and the references therein for results in this regard. These
results were further generalized in the context of noncommutative groups in [7, 29, 1, 26, 2, 3]. In
this section we wish to relate the theorems of Ingham and Paley-Wiener to the above mentioned
problem in the context of symmetric spaces. We first deduce one such result for the damped
Schrodinger Equation on Rn.
4.1. Euclidean Space. We consider the initial value problem for the time-dependent damped
Schrodinger Equation on Rn given by
(4.1)
∂u
∂t
(x, t) − i(∆ − c)u(x, t) = 0,
u(x, 0) = f (x),
for (x, t) ∈ Rn × R,
for x ∈ Rn,
where ∆ denotes the Laplacian on Rn, c ∈ R is the damping parameter and f ∈ L1(Rn). It
follows that the unique solution ut(x) = u(x, t) of (4.1) is characterized via the Euclidean Fourier
transform of ut given by the equation
(4.2)
but(ξ) = e−it(kξk2+c)bf (ξ),
for ξ ∈ Rn.
16
MITHUN BHOWMIK AND SUPARNA SEN
Theorem 4.1. Let u be the solution to the equation (4.1) given by (4.2), with initial value
f ∈ Cc(Rn). Suppose there exists t0 > 0 such that
(4.3)
u(x, t0) ≤ Ce−ψ(kxk),
for x ∈ Rn,
where ψ : [0, ∞) → [0, ∞) is a locally integrable function. If
Z ∞
1
ψ(r)
1 + r2 dt = ∞,
then u is zero.
Proof. The solution ut(x) = u(x, t) of (4.1) is given by
ut(x) = γc,t ∗ f (x),
for x ∈ Rn, t ∈ R,
where the kernel γc,t is defined by
(4.4)
γc,t(x) = (4πt)−n/2e−icte−iπsign(t)n/4eikxk2/4t,
for x ∈ Rn, t ∈ R,
so that its Fourier transform is
cγc,t(ξ) = e−i(kξk2+c)t,
for ξ ∈ Rn, t ∈ R,
(see [29], P. 869). Here sign(t) = t/t denotes the sign of t. Using the expression for γc,t in
(4.4) we can express the solution u(x, t0) at time t = t0 > 0 as
u(x, t0) = (4πt0)−n/2e−ict0e−iπn/4ei kxk2
(4.5)
4t0 bh(cid:18) x
2t0(cid:19)
where the function h on Rn is defined by
(4.6)
h(y) = ei kyk2
4t0 f (y),
for y ∈ Rn.
Using (4.3) and (4.5) and applying Theorem 1.1 (a) to the function h ∈ Cc(Rn) we get that h is
zero. Hence f is zero and so is u.
(cid:3)
Remark 4.2. It is easy to see from the relations (4.5) and (4.6) that the result corresponding to
Theorem 1.1 (b) can also be proved similarly if we assume the corresponding conditions on the
function f and the solution u(·, t0).
4.2. Riemannian Symmetric Space. Let X = G/K be a Riemannian symmetric space of
noncompact type where G is a noncompact, connected semisimple Lie group with finite center
and K is a maximal compact subgroup of G. We have a G-invariant Riemannian metric on X
induced by the Killing form B restricted to p and we can form a Laplace-Beltrami operator
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
17
∆ using this metric (see [19]). We consider the initial value problem for the time-dependent
Schrodinger equation on X given by
(4.7)
∂u
∂t
(x, t) − i∆u(x, t) = 0,
for (x, t) ∈ X × R,
u(x, 0) = f (x),
for x ∈ X.
For g ∈ G, we define σ(g) = d(gK, K) where d is the canonical distance function on X = G/K
coming from the Riemannian metric on X induced by the Killing form B. The function σ is
bi-K-invariant and continuous. If g = k1 exp Hk2 with g ∈ G, H ∈ a+ and k1, k2 ∈ K, then
σ(g) = σ(exp H) = kHkB.
4.2.1. Complex Semisimple Lie Groups. We shall first prove a unique continuation property of
solution to the Schrodinger equation (4.7) in the context of complex semisimple Lie groups with
bi-K-invariant initial data corresponding to Theorem 1.1 (b). In this case, the solution u(·, t)
is also bi-K-invariant and can be considered as a function on a. It turns out that the situation
here is slightly different from that of Euclidean spaces. The reason is roughly speaking the
exponential growth of the G-invariant measure on G/K. Consequently, to prove this unique
continuation property in the context of complex semisimple Lie groups we need to impose more
decay on the solution u(x, t). We will also show that the unique continuation property does not
hold in the absence of such extra decay (see Remark 4.4). Our method of proof will follow that
in [7] and will use some of their notation and calculations.
Theorem 4.3. Let θ : [0, ∞) → [0, ∞) be a decreasing function with limr→∞ θ(r) = 0 and u be
a solution to the equation (4.7) with initial value f ∈ L1(K\G/K). Suppose there exists t0 > 0
such that
(4.8)
u(H, t0) ≤ C φ0(H) e−kHkB θ(kHkB ),
for H ∈ a.
If f vanishes on an open set in a ∼= Rl and
(4.9)
then u is identically zero.
I =Z ∞
1
θ(r)
r
dr = ∞
Proof. It is proved in [7] that the solution u(H, t) of (4.7) at t = t0 can be written as
(4.10)
u(H, t0)φ(H) = CW 2t−l/2
0
−i(cid:18)t0kρk2
B −
e
kHk2
B
4t0 (cid:19)
where the function g on a is defined by
bgf(cid:18) H
2t0(cid:19) ,
for H ∈ a,
(4.11)
gf (H) = ei
kHk2
B
4t0 f (H)φ(H),
for H ∈ a,
18
MITHUN BHOWMIK AND SUPARNA SEN
where φ is given by (3.16). Since f vanishes on an open set in a ∼= Rl, g also vanishes on the
same open set in a. From the estimate of the Jacobian J given in (2.1) we get that
(4.12)
φ(H) =pJ(exp H) ≤ Ceρ(H),
for H ∈ a.
It follows from (4.8), (4.10), (4.12) and (2.6) that for H ∈ a
(4.13)
2t0(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) ≤ Cφ0(H)e−kHkB θ(kHkB )eρ(H) ≤ C(1 + kHkB)me−kHkB θ(kHkB ).
(cid:12)(cid:12)(cid:12)(cid:12)bgf(cid:18) H
As gf vanishes on an open set and bgf satisfies (4.13) it follows from Lemma 3.8 that gf is zero.
Therefore f is zero and hence so is u.
(cid:3)
Remark 4.4. We give an example to show that if 0 ≤ α < 1 then
u(H, t0) ≤ C φ0(H)α e−kHkB θ(kHkB ),
for H ∈ a ∼= R,
with θ as in Theorem 4.3 satisfying I = ∞ does not imply that u ≡ 0. We consider the group
G = SL(2, C) (see [19], P. 433; [34], P. 313). In this case
A =( eH
0
0
e−H! : H ∈ R) and a =(AH = H 0
0 −H! : H ∈ R) .
It is known that the Jacobian of the Haar measure is sinh2 2H so that we have
and the distance function is given by
φ(H) = sinh 2H,
for H ∈ R,
kHkB = 4H,
for H ∈ R.
The elementary spherical functions are given by
sin λH
φλ(H) = C
,
for λ, H ∈ R,
λ sinh 2H
φ0(H) = C
H
sinh 2H
,
for H ∈ R.
We fix 0 ≤ α < 1 and t0 > 0. We will give an example of a non-zero initial value f ∈ Cc(K\G/K)
such that the corresponding solution u to the Schrodinger equation (4.7) satisfies
(4.14)
u(H, t0) ≤ C φ0(H)α e−kHkB θ(kHkB ),
for H ∈ a ∼= R,
where θ is as in Theorem 4.3 satisfying (4.9). We consider η ∈ R such that 0 < η < 1 − α and
define β = 1 − α − η. For any h ∈ C ∞
c (R) we have
(4.15)
for all H ∈ R,
bh(H) ≤ CeβH,
We consider such a non-zero function h which is supported in [β′, β] ⊂ [−β, β] ⊂ R where
0 < β′ < β. Since θ is a function decreasing to zero, there exists M1 > 1 such that
(4.16)
θ(4H) <
η
4
,
for H > M1,
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
19
and there exists M2 > 1 such that
(4.17)
eH ≤ C sinh 2H,
for H > M2.
From (4.15), (4.16) and (4.17) it follows that for M = max{M1, M2} > 1 and H > M
≤ CHα(sinh 2H)1−αe−4Hθ(4H)
bh(H) ≤ Ce(1−α)He−ηH
sinh 2H(cid:19)α
= C(cid:18) H
(sinh 2H)e−4Hθ(4H)
(4.18)
= Cφ(H)φ0(H)αe−4Hθ(4H).
We define a function f ∈ C ∞
c (R) by
(4.19)
f (H) =
1
2t0
e−4i H2
t0 φ(H)−1h(cid:18) H
2t0(cid:19) ,
for H ∈ R.
This is well defined since φ vanishes only at 0 and h is supported away from zero. We consider
the solution u(H, t) of the system (4.7) with the initial value f . It follows from (4.11) and (4.19)
that
So by (4.10)
Hence by (4.18) we get that
1
2t0
gf (H) =
h(cid:18) H
2t0(cid:19) ,
u(H, t0)φ(H) = Cbh(H),
for H ∈ R.
for H ∈ R.
u(H, t0)φ(H) ≤ Cφ(H)φ0(H)αe−4Hθ(4H),
for H > M.
Since by continuity the function
H 7→ u(H, t0)e4Hθ(4H)(φ0(H))−α
is bounded on the compact set [0, M ] it follows that the solution u satisfies (4.14) with non-zero
initial data f ∈ C ∞
c (R).
4.2.2. Symmetric Space of Noncompact Type. We shall now prove a unique continuation prop-
erty of solution to the Schrodinger equation (4.7) in the context of Riemannian symmetric spaces
of noncompact type using Theorem 4.1. As in the case of complex semisimple Lie groups, here
also we need to impose extra decay on the solution u(x, t) and using a similar example we will
show that such decay is necessary (see Remark 1). However, here we shall prove our result by
reducing the problem to the Euclidean case via the Radon transform. It can be seen that for the
initial data f ∈ L2(X) there exists unique solution u(·, t) ∈ L2(X) satisfying the Schrodinger
equation (4.7) in the sense of distributions. However the Radon transform is not defined in gen-
eral for functions in L2(X). So we need to consider our initial data f in the L2-Schwartz space
20
MITHUN BHOWMIK AND SUPARNA SEN
S(X) consisting of smooth rapidly decreasing functions on X, on which the Radon transform
happens to be defined. We shall start with the definition of S(X).
Let DL(G) and DR(G) denote the algebra of left invariant and that of right invariant differen-
tial operators on G respectively. The L2-Schwartz space S(G) is defined as the space of smooth
functions f on G such that for each D ∈ DL(G), E ∈ DR(G) and l ∈ N
(4.20)
(1 + σ(g))lφ0(g)−1(DEf )(g) < ∞.
sup
g∈G
The L2-Schwartz space S(X) is then defined as the space of f ∈ S(G) which are right invariant
under K (see [18], P. 214).
For λ ∈ a∗
C and kM ∈ K/M we define the function eλ,k : X → C by
eλ,k(x) = e(iλ−ρ)H(x−1k),
for x = gK ∈ X, g ∈ G.
It is known that these functions eλ,k appearing in the definition of the Helgason-Fourier transform
(2.3) are eigenfunctions of the Laplace-Beltrami operator ∆ given by
∆eλ,k = −(kλk2
B + kρk2
B)eλ,k,
for λ ∈ a∗
C, kM ∈ K/M,
(see [18], P. 99 and [19], Ch. II). It follows that for f ∈ L2(X) the unique solution u(·, t) ∈ L2(X)
satisfying the Schrodinger equation (4.7) is characterized by
(4.21)
for kM ∈ K/M and λ ∈ a∗.
eut(λ, kM ) = e−it(kλk2
B +kρk2
B)ef (λ, kM ),
However we are interested in the special case where the initial data f ∈ S(X). In this case
it turns out that the solution ut(·) = u(t, ·) ∈ S(X) (see [29], P. 872; [11], Theorem 4.1.1).
It follows that that the analysis of solutions of the Schrodinger equation on Rn carries out to
X when the Fourier transform on Rn is replaced by the Helgason-Fourier transform. For any
f ∈ S(X) the Radon transform Rf is well defined (see [18], P. 218, Theorem 1.15). From (2.10)
and (4.21) it follows that for λ ∈ a∗ and fixed kM ∈ K/M we have
(4.22)
FA((Rut)(kM, ·)(λ) = e−it(kλk2
B +kρk2
B)FA((Rf )(kM, ·)(λ).
We shall recall few facts which will be needed. Firstly, the function σ satisfies the following
inequality
(4.23)
σ(an) ≥ σ(a),
for all a ∈ A, n ∈ N
(see [15], Lemma 6.2.7). Moreover, there exists a non-negative integer m ≥ 0 such that for some
constant C0 > 0,
(4.24)
φ0(an) (1 + σ(an))−mdn ≤ C0,
for all a ∈ A,
eρ log aZN
(see [15], P. 264, Theorem 6.2.3). We shall now present the unique continuation property for
solutions to the Schrodinger equation (4.7) for symmetric spaces.
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
21
Theorem 4.5. Let u(x, t) be a solution to the equation (4.7) with initial value f ∈ C ∞
Suppose there exists t0 > 0 such that
c (X).
(4.25)
u(x, t0) ≤ C φ0(x) e−ψ(σ(x)),
for x ∈ X,
where ψ : [0, ∞) → [0, ∞) is a non-decreasing function. If
then f is zero on X.
I =Z ∞
1
ψ(r)
1 + r2 dr = ∞,
Proof. Since f ∈ C ∞
function on G, from the definition of L2-Schwartz space (4.20) we get that for each l ∈ N
c (X) ⊂ S(X), it follows that ut ∈ S(X). Viewing ut as a right K-invariant
(4.26)
ut(g) ≤ Cφ0(g)(1 + σ(g))−2l,
for g ∈ G.
By multiplying the inequalities in (4.25) and (4.26) it follows that for each l ∈ N
(4.27)
u(x, t0) ≤ C φ0(x) (1 + σ(x))−le− 1
2 ψ(σ(x)),
for x ∈ X.
Now, from the expression of ut given in (4.2) and the relation (4.22) it follows that for fixed
kM ∈ K/M , (Rut)(kM, ·) is a solution to the system (4.1) with initial value (Rf )(kM, ·) and
damping parameter kρk2
B. We wish to apply Theorem 4.1 to this particular solution. Let us fix
kM ∈ K/M . Since f ∈ C ∞
c (A). We need to prove
that (Rut)(kM, ·) satisfies the estimate (4.3) at time t = t0. Indeed, for a ∈ A and l = m using
(2.9) and (4.27) we get that
c (X) it is easy to see that Rf (kM, ·) ∈ C ∞
(Rut0 )(kM, a) = eρ log aZN
≤ C eρ log aZN
ut0 (kan · o) dn
φ0(an) (1 + σ(an))−me− 1
2 ψ(σ(an)) dn.
Since ψ is non-decreasing using (4.23) it follows that
(4.28)
ψ(σ(an)) ≥ ψ(σ(a)),
for all a ∈ A, n ∈ N.
Using (4.28) and (4.24) we get that
(Rut0 )(kM, a) ≤ Ce− 1
2 ψ(σ(a)),
for all a ∈ A.
Using the equivalence of norms in finite dimensional spaces, we can invoke Theorem 4.1 to get
that (Rf )(kM, ·) is zero on A. Since this is true for all kM ∈ K/M we get that Rf is zero on
K/M × A. By the injectivity of Radon transform ([18], P. 220, Corollary 1.6) we conclude that
f is zero. Hence so is u.
(cid:3)
22
MITHUN BHOWMIK AND SUPARNA SEN
Remark 4.6.
(1) We give an example to show that if 0 ≤ α < 1 then there exists a non-
decreasing function ψ (depending on α) and a non-zero initial value f ∈ Cc(K\G/K)
such that the corresponding solution u to the Schrodinger equation (4.7) satisfies
(4.29)
u(H, t0) ≤ C φ0(H)α e−ψ(kHkB ),
for H ∈ a ∼= R,
where ψ is as in Theorem 4.5 with I = ∞. We again consider the group G = SL(2, C)
and as in Remark 4.4 we get h ∈ C ∞
c (R) supported in [β′, β] ⊂ [−β, β] with 0 < β′ < β
such that
where β = 1 − α − η and 0 < η < 1 − α. We now define
bh(H) ≤ CeβH,
for all H ∈ R,
(4.30)
ψ(H) = ηH,
for H ∈ R.
Using (4.17) and (4.30) in (4.18) we get that for H > M2
bh(H) ≤ Cφ(H)φ0(H)αe−ψ(H).
c (R) by (4.19) we conclude as in Remark 4.4 that u satisfies (4.29) with
Defining f ∈ C ∞
non-zero initial data f ∈ C ∞
c (R).
(2) It seems to be an interesting problem to see whether Theorem 3.9 and Theorem 4.3 can
be generalized to general Riemannian symmetric spaces of noncompact type.
Acknowledgement. We would like to thank Swagato K. Ray for suggesting this problem
and for the many useful discussions during the course of this work. We also thank Rudra P
Sarkar for his valuable comments regarding this work.
References
[1] Ben Saıd, S.; Thangavelu, S.; Dogga, V. N. Uniqueness of solutions to Schrodinger equations on H-type
groups, J. Aust. Math. Soc. 95 (2013) no. 3, 297-314. MR3164504
[2] Bhowmik, M.; Ray, S. K.; Sen, S. Around Uncertainty Principles of Ingham-type on Rn, Tn and two step
nilpotent Lie Groups, preprint, arXiv:1605.09616
[3] Bhowmik, M.; Sen, S. An Uncertainty Principle of Paley and Wiener on Euclidean Motion Group, preprint,
arXiv:1606.01704
[4] Bochner, S. Quasi-analytic functions, Laplace operator, positive kernels, Ann. of Math. (2) 51, (1950). 68-91.
MR0032708 (11,334g)
[5] Bochner, S.; Taylor, A. E. Some Theorems on Quasi-Analyticity for Functions of Several Variables, Amer.
J. Math. 61 (1939), no. 2, 303-329. MR1507378
[6] Burckel, Robert B. An introduction to classical complex analysis, Pure and Applied Mathematics, 82, Aca-
demic Press, Inc [Harcourt Brace Jovanovich, Publishers], New York-London, 1979. MR0555733 (81d:30001)
[7] Chanillo, S. Uniqueness of solutions to Schrodinger equations on complex semi-simple Lie groups, Proc.
Indian Acad. Sci. Math. Sci. 117 (2007), no. 3, 325-331. MR2352052 (2008h:22010)
[8] Conway, John B. Functions of one complex variable II, Graduate Texts in Mathematics, 159, Springer-Verlag,
New York, 1995. MR1344449 (96i:30001)
UNCERTAINTY PRINCIPLES OF INGHAM AND PALEY-WIENER ON SEMISIMPLE LIE GROUPS
23
[9] Cowling, M.; Nevo, A. Uniform estimates for spherical functions on complex semisimple Lie groups, Geom.
Funct. Anal. 11 (2001), no. 5, 900-932. MR1873133 (2002k:43005)
[10] Cowling, M.; Sitaram, A.; Sundari, M. Hardy's uncertainty principle on semisimple groups, Pacific J. Math.
192 (2000), no. 2, 293-296. MR1744570 (2001c:22007)
[11] Eguchi, M. Asymptotic expansions of Eisenstein integrals and Fourier transform on symmetric spaces, J.
Funct. Anal. 34 (1979), no. 2, 167-216. MR0552702 (81e:43022)
[12] Escauriaza, L.; Kenig, C. E.; Ponce, G.; Vega, L. On uniqueness properties of solutions of Schrodinger
equations, Comm. Partial Differential Equations 31 (2006), no. 10-12, 18111823. MR2273975 (2009a:35198)
[13] Escauriaza, L.; Kenig, C. E.; Ponce, G.; Vega, L. Uniqueness properties of solutions to Schrodinger equations,
Bull. Amer. Math. Soc. (N.S.) 49 (2012), no. 3, 415-442. MR2917065
[14] Folland, G, B.; Sitaram, A. The uncertainty principle: A mathematical survey, J. Four. Anal. Appl. 3 (1997),
no. 3, 207-238. MR1448337 (98f:42006)
[15] Gangolli, R.; Varadarajan V. S. Harmonic Analysis of Spherical Functions on Real Reductive Groups, Results
in Mathematics and Related Areas, 101, Springer-Verlag, Berlin, 1988. MR954385 (89m:22015)
[16] Havin, V.; Joricke, B. The uncertainty principle in harmonic analysis, Ergebnisse der Mathematik und ihrer
Grenzgebiete, 3, Folge, 28, Berlin, Springer-Verlag, 1994. MR1303780 (96c:42001)
[17] Helgason, S. Differential geometry, Lie groups, and symmetric spaces, Graduate Studies in Mathematics, 34,
American Mathematical Society, Providence, RI, 2001. MR1834454 (2002b:53081)
[18] Helgason, S. Geometric Analysis on Symmetric Spaces, Mathematical Surveys and Monographs 39. American
Mathematical Society, Providence, RI, 1994. MR1280714 (96h:43009)
[19] Helgason, S. Groups and geometric analysis, Integral geometry, invariant differential operators, and spherical
functions, Mathematical Surveys and Monographs, 83. American Mathematical Society, Providence, RI, 2000.
MR1790156 (2001h:22001)
[20] 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)
[21] Ingham, A. E. A Note on Fourier Transforms, J. London Math. Soc. S1-9 no. 1, 29. MR1574706
[22] Knapp, A., Representation theory of semisimple groups, An overview based on examples, Princeton Mathe-
matical Series, 36. Princeton University Press, Princeton, NJ, 1986. MR855239 (87j:22022)
[23] Koosis, P. The logarithmic integral I (Corrected reprint of the 1988 original) Cambridge Studies in Advanced
Mathematics, 12. Cambridge University Press, Cambridge, 1998. xviii+606 pp. MR1670244 (99j:30001)
[24] Levinson, N. Gap and Density Theorems American Mathematical Society Colloquium Publications, v. 26.
American Mathematical Society, New York, 1940. MR0003208 (2,180d)
[25] Levinson, N. On a Class of Non-Vanishing Functions Proc. London Math. Soc. S2-41 no. 5, 393. MR1576177
[26] Ludwig, J.; Muller, D. Uniqueness of solutions to Schrodinger equations on 2-step nilpotent Lie groups, Proc.
Amer. Math. Soc. 142 (2014) no. 6, 2101-2118. MR3182028
[27] Paley, R. E. A. C.; Wiener, N. Fourier transforms in the complex domain, American Mathematical Society
Colloquium Publications, 19. American Mathematical Society, Providence, RI, 1987. MR1451142 (98a:01023)
[28] 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, 348-355. MR1501688
[29] Pasquale, A.; Sundari, M. Uncertainty principles for the Schrodinger equation on Riemannian symmetric
spaces of the noncompact type, Ann. Inst. Fourier (Grenoble) 62 (2012), no. 3, 859-886. MR3013810
[30] Rudin, W. Real and Complex Analysis, McGraw-Hill Book Co., New York, 1987. MR924157 (88k:00002)
[31] Sarkar, R. P.; Sengupta, J. Beurling's theorem and characterization of heat kernel for Riemannian symmetric
spaces of noncompact type, Canad. Math. Bull. 50 (2007), no. 2, 291-312. MR2317450 (2008e:43014)
24
MITHUN BHOWMIK AND SUPARNA SEN
[32] Stein, E. M.; Shakarchi, R. Complex analysis, Princeton Lectures in Analysis, II. Princeton University Press,
Princeton, NJ, 2003. MR1976398 (2004d:30002)
[33] Stein, E. M.; Weiss, G. Introduction to Fourier analysis on Euclidean spaces, Princeton Mathematical Series,
No. 32. Princeton University Press, Princeton, N.J., 1971. MR0304972 (46 # 4102)
[34] Terras, A. Harmonic analysis on symmetric spaces and applications II, Springer-Verlag, Berlin, 1988.
MR0955271 (89k:22017)
[35] Thangavelu, S. An Introduction to the Uncertainty Principle. Hardy's theorem on Lie groups, Progress in
Mathematics, 217. Birkhauser Boston, Inc., Boston, MA, 2004. MR2008480 (2004j:43007)
[36] Thangavelu, S. On Theorems of Hardy, Gelfand-Shilov and Beurling for Semisimple Lie Groups, Publ. Res.
Inst. Math. Sci. 40 (2004), no. 2, 311-344. MR2049638 (2005e:43014)
Stat-Math Unit, Indian Statistical Institute, 203 B. T. Road, Kolkata - 700108, India.
E-mail address: [email protected], [email protected]
|
1604.05485 | 4 | 1604 | 2017-04-19T07:02:51 | Contractions with Polynomial Characteristic Functions II. Analytic Approach | [
"math.FA",
"math.OA"
] | The simplest and most natural examples of completely nonunitary contractions on separable complex Hilbert spaces which have polynomial characteristic functions are the nilpotent operators. The main purpose of this paper is to prove the following theorem: Let $T$ be a completely nonunitary contraction on a Hilbert space $\mathcal{H}$. If the characteristic function $\Theta_T$ of $T$ is a polynomial of degree $m$, then there exist a Hilbert space $\mathcal{M}$, a nilpotent operator $N$ of order $m$, a coisometry $V_1 \in \mathcal{L}(\overline{ran} (I - N N^*) \oplus \mathcal{M}, \overline{ran} (I - T T^*))$, and an isometry $V_2 \in \mathcal{L}(\overline{ran} (I - T^* T), \overline{ran} (I - N^* N) \oplus \mathcal{M})$, such that \[ \Theta_T = V_1 \begin{bmatrix} \Theta_N & 0 0 & I_{\mathcal{M}} \end{bmatrix} V_2. \] | math.FA | math |
CONTRACTIONS WITH POLYNOMIAL CHARACTERISTIC FUNCTIONS
II. ANALYTIC APPROACH
CIPRIAN FOIAS, CARL PEARCY, AND JAYDEB SARKAR
This paper is dedicated to our lifelong friend Ron Douglas on the occasion of his upcoming 80th birthday
Abstract. The simplest and most natural examples of completely nonunitary contractions
on separable complex Hilbert spaces which have polynomial characteristic functions are the
nilpotent operators. The main purpose of this paper is to prove the following theorem: Let T
be a completely nonunitary contraction on a Hilbert space H. If the characteristic function ΘT
of T is a polynomial of degree m, then there exist a Hilbert space M, a nilpotent operator
N of order m, a coisometry V1 ∈ L(ran(I − N N ∗) ⊕ M, ran(I − T T ∗)), and an isometry
V2 ∈ L(ran(I − T ∗T ), ran(I − N ∗N ) ⊕ M), such that
ΘT = V1(cid:20)ΘN
0
0
IM(cid:21) V2.
1. Introduction
This is a sequel to our paper [4], where we identified the structure of the completely
nonunitary contractions on a Hilbert space that have a polynomial characteristic function.
Namely, we proved that the characteristic function ΘT of a completely nonunitary contraction
T on a separable, infinite dimensional, complex Hilbert space H is a polynomial if and only if
there exist three closed subspaces H1, H0, H−1 of H with H = H1 ⊕H0 ⊕H−1, a pure isometry
S in L(H1), a nilpotent N in L(H0), and a pure co-isometry C in L(H−1), such that T has
the matrix representation
T =
S ∗
∗
0 N ∗
0
0 C
.
Moreover, the multiplicities of S and C, in other words, dim ker S∗ and dim ker C are
unitary invariants of T , and the nilpotent operator is uniquely determined by T up to a
quasi-similarity. For earlier results on contractions with constant characteristic functions see
[1], [9] and [10].
Recall that a pure isometry is a unilateral shift of some multiplicity and a pure coisometry
is the adjoint of a pure isometry. Recall also that a contraction T on a Hilbert space H, (i.e.,
kT hk ≤ khk for all h in H) is completely nonunitary (c.n.u.) if there is no nontrivial reducing
subspace M of H for T such that T M is unitary.
2000 Mathematics Subject Classification. 47A45, 47A20, 47A48, 47A56.
Key words and phrases. Characteristic function, model, nilpotent operators, operator valued polynomials.
1
2
FOIAS, PEARCY, AND SARKAR
In this paper we shall adopt a second approach to prove the theorem stated in the abstract,
based essentially on new factorizations of characteristic functions of upper triangular 3 × 3
block contractions (see Theorem 2.3).
Before we continue we recall the notion of the characteristic function of a contraction.
Consider a contraction T on a Hilbert space H. The defect operators DT and DT ∗ and the
defect spaces DT and DT ∗ of T are defined by
DT = (IH − T ∗T )
1
2 ,
and
DT ∗ = (IH − T T ∗)
1
2 ,
and
DT = RanDT ,
and
DT ∗ = RanDT ∗,
respectively. Then the characteristic function of the contraction T is the L(DT , DT ∗)-valued
contractive analytic function defined by
ΘT (z) = [−T + zDT ∗(IH − zT ∗)−1DT ]DT
(z ∈ D).
In particular, ΘT is a L(DT , DT ∗)-valued bounded analytic function on D (see [6]). Moreover,
the characteristic function ΘT is purely contractive, that is,
kΘT (0)ηk < kηk
(η ∈ DT , η 6= 0).
Let Θ : D → L(M, M∗) and Ψ : D → L(N , N∗) be two operator valued analytic functions
on D. We say that Θ and Ψ coincide and write Θ ∼= Ψ if there exist two unitary operators
τ : M → N and τ∗ : M∗ → N∗ such that
Θ(z) = τ −1
(z ∈ D),
∗ Ψ(z)τ
or, equivalently, for all z ∈ D the following diagram commutes:
M
Θ(z)
−−−→ M∗
τy
N
τ∗y
Ψ(z)
−−−→ N∗
The characteristic function is a complete unitary invariant in the following sense (see [6],
Theorem 3.4): Two c.n.u. contractions T on H and R on K are unitarily equivalent (that is,
there is a unitary operator U from H to K such that T = U ∗RU) if and only if
ΘT ∼= ΘR.
Moreover, for a given L(E , E∗)-valued purely contractive analytic function Θ defined on D,
there exists a c.n.u. contraction T on some Hilbert space, explicitly determined by Θ, such
that ΘT coincides with Θ.
Contractive operator valued analytic functions play an important role in operator theory
and serve as a bridge between operator theory and function theory in terms of systems theory
and interpolation theory (cf. [2], [6], [8]).
The class of nilpotent contractions yields a natural set of examples of operators that have
polynomial characteristic functions. Indeed, let N be a contraction and a nilpotent operator
CONTRACTIONS WITH POLYNOMIAL CHARACTERISTIC FUNCTIONS
3
of order m, m ≥ 1, that is, kNk ≤ 1, N m = 0, and N m−1 6= 0. The characteristic function
ΘN of N is given by
ΘN (z) = [−N + zDN ∗(IH − zN ∗)−1DN ]DN
= [−N +
= [−N +
∞
Xp=0
Xp=0
m−1
zp+1DN ∗N ∗pDN ]DN
zp+1DN ∗N ∗pDN ]DN ,
for all z ∈ D. Therefore ΘN is a polynomial in z of degree at most m with operator coefficients.
From this viewpoint, it is important to understand, up to unitary equivalance, the analytic
structure of polynomial characteristic functions of contractions. The main goal of the present
paper is to address this issue. More specifically, in Theorem 3.2 we prove: If the characteristic
function ΘT of a c.n.u. contraction T is a polynomial of degree m, then there exist a Hilbert
space M, a nilpotent operator N of order m, a co-isometry V1 ∈ L(DN ∗ ⊕ M, DT ∗), and an
isometry V2 ∈ L(DT , DN ⊕ M), such that
ΘT = V1(cid:20)ΘN
0
0
IM(cid:21) V2.
Along the way we prove the following factorization result for characteristic functions (see
Theorem 2.3): Let H1, H0, H−1 be three Hilbert spaces and set H = H1 ⊕ H0 ⊕ H−1. Let
T =
S ∗
∗
0 N ∗
0 C
0
be any contraction on H with the above matricial form. Then the characteristic function ΘT
of T and
(cid:20)ΘC
0
0
IE1(cid:21) U1(cid:20)ΘN
0
0
IM(cid:21) U2(cid:20)ΘS
0
0
IE2(cid:21) .
coincide, where E1, E2 and M are Hilbert spaces, and U1 ∈ L(DN ∗ ⊕ M, DC ⊕ E1) and
U2 ∈ L(DS ∗ ⊕ E2, DN ⊕ M) are unitary operators.
Our results rely on the upper triangular representation of operators with polynomial char-
acteristic functions (see Theorem 3.1) and a factorization of characteristic functions of upper
triangular 2 × 2 block contractions due to Sz.-Nagy and the first author (see Theorem 2.2).
The rest of this paper is organized as follows:. In Section 2, we give the factorization of
the characteristic function of an upper triangular 3 × 3 block contraction on Hilbert space.
Our main result is given in Section 3, and provides a complete analytic characterization of
polynomial characteristic functions for c.n.u. contractions on Hilbert space.
2. Factorizations of characteristic functions
We start by recalling some known facts about upper triangular 2 × 2 block contractions,
since they will be frequently used in what follows.
4
FOIAS, PEARCY, AND SARKAR
The first is a classification of 2 × 2 block contractions. This is the content of Theorem 1 in
[7] (also see Chapter IV, Lemma 2.1 in [3]).
Theorem 2.1. Let H1 and H2 be Hilbert spaces and let T = (cid:20)T1 X
0 T2(cid:21) be a bounded linear
operator on H1 ⊕ H2. Then T is a contraction if and only if T1 and T2 are contractions and
for some contraction Γ from DT2 to DT ∗
1 .
X = DT ∗
1 ΓDT2,
The second key tool used in our development is the factorization of characteristic functions
of 2 × 2 block contractions (see Theorem 2 in [7]).
Theorem 2.2. Let H1 and H2 be Hilbert spaces, let T = (cid:20)T1 X
0 T2(cid:21) be a contraction on
H1 ⊕ H2, and let X = DT ∗
unitary operators τ ∈ L(DT , DT1 ⊕ DΓ) and τ∗ ∈ L(DT ∗, DT ∗
1 ΓDT2 for some contraction Γ ∈ L(DT ∗
2 ⊕ DΓ∗) such that
1 , DT2). Then there exist
ΘT (z) = τ −1
∗ (cid:20)ΘT2(z)
0
0
IDΓ∗(cid:21) J[Γ](cid:20)ΘT1(z)
0
0
IDΓ(cid:21) τ
(z ∈ D),
where
J[Γ] = (cid:20) Γ∗ DΓ
DΓ∗ −Γ(cid:21) ∈ L(DT ∗
1 ⊕ DΓ, DT2 ⊕ DΓ∗).
Recall that if A is a contraction from H to K then
(2.1)
J[A] = (cid:20) A∗ DA
DA∗ −A(cid:21)
is a unitary operator from K ⊕ DA to H ⊕ DA∗ (see Halmos [5]).
We are now ready to prove our first factorization result.
Theorem 2.3. Let H1, H0, H−1 be Hilbert spaces, and let H = H1 ⊕ H0 ⊕ H−1. Let
T =
S ∗
∗
0 N ∗
0 C
0
,
be a contraction on H. Then there exist three Hilbert spaces E1, E2 and M and two unitary
operators U1 ∈ L(DN ∗ ⊕ M, DC ⊕ E1) and U2 ∈ L(DS ∗ ⊕ E2, DN ⊕ M) such that
ΘT ∼= (cid:20)ΘC
0
Proof. Set
0
IE1(cid:21) U1(cid:20)ΘN
0
0
IM(cid:21) U2(cid:20)ΘS
0
0
IE2(cid:21) .
K1 = H1 ⊕ H0,
T = (cid:20)T1 X1
0 C (cid:21) ∈ L(K1 ⊕ H−1),
CONTRACTIONS WITH POLYNOMIAL CHARACTERISTIC FUNCTIONS
5
and
T1 = (cid:20)S X
0 N(cid:21) ∈ L(H1 ⊕ H0),
where X1 ∈ L(H−1, K1) and X ∈ L(H0, H1). Theorem 2.1 implies that there exist contrac-
tions Γ1 ∈ L(DC, DT ∗
1 ) and Γ ∈ L(DN , DS ∗) such that
1 Γ1DC,
X1 = DT ∗
and
By Theorem 2.2 there exist unitary operators
X = DS ∗ΓDN .
(2.2)
and
(2.3)
such that
and
for all z ∈ D, where
and
τ1 : DT → DT1 ⊕ DΓ1,
τ1∗ : DT ∗ → DC ∗ ⊕ DΓ∗
1 ,
τ : DT1 → DS ⊕ DΓ,
τ∗ : DT ∗
1 → DN ∗ ⊕ DΓ∗,
ΘT (z) = τ −1
0
1∗ (cid:20)ΘC(z)
∗ (cid:20)ΘN (z)
0
IDΓ∗
0
0
1(cid:21) J[Γ1](cid:20)ΘT1(z)
IDΓ∗(cid:21) J[Γ](cid:20)ΘS(z)
0
0
0
IDΓ1(cid:21) τ1,
IDΓ(cid:21) τ,
0
ΘT1(z) = τ −1
J[Γ1] = (cid:20) Γ∗
DΓ∗
1 DΓ1
1 −Γ1(cid:21) ∈ L(DT ∗
1 ⊕ DΓ1, DC ⊕ DΓ∗
1),
J[Γ] = (cid:20) Γ∗ DΓ
DΓ∗ −Γ(cid:21) ∈ L(DS ∗ ⊕ DΓ, DN ⊕ DΓ∗),
are unitary operators (see (2.1)). Now setting
ΦS(z) = (cid:20)ΘS(z)
0
for all z ∈ D, we get
0
IDΓ(cid:21) , ΦN (z) = (cid:20)ΘN (z)
0
0
IDΓ∗(cid:21) , and ΦC(z) = (cid:20)ΘC(z)
0
0
1(cid:21) ,
IDΓ∗
ΘT (z) = τ −1
0
IDΓ1(cid:21) τ1
0
1∗ ΦC(z)J[Γ1](cid:20)ΘT1(z)
1∗ ΦC(z)J[Γ1](cid:20)τ −1
1∗ ΦC(z)(cid:16)J[Γ1](cid:20)τ −1
1∗ ΦC(z)U1(cid:20)ΦN (z)J[Γ]ΦS(z)τ
∗
0
0
0
∗ ΦN (z)J[Γ]ΦS(z)τ
0
IDΓ1(cid:21) τ1
IDΓ1(cid:21)(cid:17)(cid:20)ΦN (z)J[Γ]ΦS (z)τ
0
0
0
IDΓ1(cid:21) τ1,
= τ −1
= τ −1
= τ −1
0
IDΓ1(cid:21) τ1
for all z ∈ D, where U1 ∈ L((DN ∗ ⊕ DΓ∗) ⊕ DΓ1, DC ⊕ DΓ∗
1) is the unitary operator defined by
6
FOIAS, PEARCY, AND SARKAR
Hence
U1 = J[Γ1](cid:20)τ −1
∗
0
0
IDΓ1(cid:21) .
ΘT (z) = τ −1
0
1∗ ΦC(z)U1(cid:20)ΦN (z)J[Γ]ΦS(z)τ
IDΓ1(cid:21)(cid:20)J[Γ]
1∗ ΦC(z)U1(cid:20)ΦN (z)
IDΓ∗ ⊕DΓ1(cid:21)(cid:20)J[Γ]
1∗ ΦC(z)U1(cid:20)ΘN (z)
0
0
0
0
0
0
= τ −1
= τ −1
0
IDΓ1(cid:21) τ1
IDΓ1(cid:21)(cid:20)ΦS(z)
0
0
0
IDΓ1(cid:21)(cid:20)τ
0
0 IDΓ1(cid:21) τ1
IDΓ⊕DΓ1(cid:21)(cid:20)τ
0
0
0 IDΓ1(cid:21) τ1,
0
IDΓ1(cid:21)(cid:20)ΘS(z)
0
for all z ∈ D. Let U2 ∈ L((DS ∗ ⊕DΓ)⊕DΓ1 , (DN ⊕DΓ∗)⊕DΓ1) and τ1 ∈ L(DT , (DS ⊕DΓ)⊕DΓ1)
be unitary operators defined by
U2 = (cid:20)J[Γ]
0
0
IDΓ1(cid:21) ,
and
respectively. Hence we obtain
τ1 = (cid:20)τ
0 IDΓ1(cid:21) τ1,
0
(2.4)
ΘT (z) = τ −1
1∗ (cid:16)(cid:20)ΘC(z)
0
0
1(cid:21) U1(cid:20)ΘN (z)
0
IDΓ∗
0
IDΓ∗ ⊕DΓ1(cid:21) U2(cid:20)ΘS(z)
0
0
IDΓ⊕DΓ1(cid:21)(cid:17)τ1,
for all z ∈ D, and therefore
ΘT ∼= (cid:20)ΘC
0
0
1(cid:21) U1(cid:20)ΘN
0
IDΓ∗
0
IDΓ∗ ⊕DΓ1(cid:21) U2(cid:20)ΘS
0
0
IDΓ⊕DΓ1(cid:21) ,
holds. Setting E1 = DΓ∗
proof of the theorem.
1, M = DΓ∗ ⊕ DΓ1 and E2 = DΓ ⊕ DΓ1 in the above, we conclude the
Of particular interest is the case when S and C ∗ are pure isometries.
Corollary 2.4. With the hypotheses of Theorem 2.3, let also assume that S and C ∗ are
pure isometries. Then there exist a Hilbert space M, a co-isometry V1 ∈ L(DN ∗ ⊕ M, DT ∗),
and an isometry V2 ∈ L(DT , DN ⊕ M), such that
ΘT = V1(cid:20)ΘN
0
0
IM(cid:21) V2.
Proof. Notice that since DC ∗ = {0H−1} and DS = {0H1}, the characteristic functions ΘC :
D → L(DC, DC ∗) of C and ΘS : D → L(DS, DS ∗) of S are identically zero, that is,
0C := ΘC ≡ 0 : DC → {0H−1}, and 0S := ΘS ≡ 0 : {0H1} → DS ∗.
Furthermore, the unitary operators in (2.2) and (2.3) become
(2.5)
τ1 : DT → DT1 ⊕ DΓ1,
τ1∗ : DT ∗ → {0H−1} ⊕ DΓ∗
1 ,
CONTRACTIONS WITH POLYNOMIAL CHARACTERISTIC FUNCTIONS
7
and
(2.6)
τ : DT1 → {0H1} ⊕ DΓ,
τ∗ : DT ∗
1 → DN ∗ ⊕ DΓ∗.
This along with (2.4) yields
ΘT = τ −1
0
1∗ (cid:20)0C
= V1(cid:20)ΘN
0
0
1(cid:21) U1(cid:20)ΘN
0
IDΓ∗
0
IDΓ∗ ⊕DΓ1(cid:21) U2(cid:20)0S
0
0
IDΓ⊕DΓ1(cid:21) τ1
0
IM(cid:21) V2,
M = DΓ∗ ⊕ DΓ1,
V1 = τ −1
1∗ (cid:20)0C
0
0
1(cid:21) U1 ∈ L((DN ∗ ⊕ DΓ∗) ⊕ DΓ1, DT ∗)
IDΓ∗
V2 = U2(cid:20)0S
0
0
IDΓ⊕DΓ1(cid:21) τ1 ∈ L(DT , (DN ⊕ DΓ∗) ⊕ DΓ1).
where
and
Now using 0C0∗
readily see that V1V ∗
1 = IDT ∗ and V ∗
2 V2 = IDT . This completes the proof of the corollary.
C = IDC∗ = I{0H−1 } and 0∗
S0S = IDS = I{0H1 } along with (2.5) and (2.6) we
3. Polynomial characteristic functions
For the readers convenience, we first state the main result of [4].
Theorem 3.1. Let T be a c.n.u. contraction on a Hilbert space H. Then the characteristic
function ΘT of T is a polynomial of degree m if and only if there exist three closed subspaces
H1, H0, H−1 of H with H = H1⊕H0⊕H−1, a pure isometry S in L(H1), a nilpotent N of order
m in L(H0), and a pure co-isometry C in L(H−1), such that T has the matrix representation
T =
S ∗
∗
0 N ∗
0 C
0
.
We are now ready for the main theorem on analytic description of contractions which have
polynomial characteristic functions.
Theorem 3.2. Let T be a c.n.u. contraction on a Hilbert space H.
If the characteris-
tic function ΘT of T is a polynomial of degree m, then there exist a Hilbert space M, a
nilpotent operator N of order m, a co-isometry V1 ∈ L(DN ∗ ⊕ M, DT ∗), and an isometry
V2 ∈ L(DT , DN ⊕ M), such that
ΘT = V1(cid:20)ΘN
0
0
IM(cid:21) V2.
8
FOIAS, PEARCY, AND SARKAR
Proof. Let T be a c.n.u. contraction such that the characteristic function ΘT of T is a
polynomial of degree m. According to Theorem 3.1 there exist closed subspaces H1, H0, H−1
of H such that H = H1 ⊕ H0 ⊕ H−1 and such that with respect to that decomposition, T
admits the matrix representation
T =
S ∗
∗
0 N ∗
0 C
0
,
where S ∈ L(H1) is a pure isometry, N ∈ L(H0) is a nilpotent operator of order m, and
C ∈ L(H−1) is a pure coisometry. The result now follows from Corollary 2.4.
Remark 3.3. The converse of the above theorem is not true in full generality: Let M be an
infinite dimensional separable Hilbert space, and let T be a c.n.u. contraction with infinite
dimensional defect spaces (for example, one can consider T = S ⊕ S∗ on H 2
E(D),
where E is an infinite dimensional Hilbert space, H 2
E(D) is the E-valued Hardy space, and S
is the shift operator on H 2
E(D)). Let N be a nilpotent operator of order m and let
E(D) ⊕ H 2
be an isometry, where
V2 = (cid:20)V21
V22(cid:21) : DT → DN ⊕ M,
kV21ηk = kV22ηk
(η ∈ DT ).
Also, let V1 : DN ∗ ⊕ M → DT ∗ be a coisometry with ker V1 = DN ∗. If
ΘT = V1(cid:20)ΘN
0
0
IM(cid:21) V2,
then ΘT is a polynomial of degree 0.
However, it is easy to see that the following weak converse of Theorem 3.2 is true: Let T be
a c.n.u. contraction on a Hilbert space H. Let
ΘT = V1(cid:20)ΘN
0
0
IM(cid:21) V2,
for some Hilbert space M, nilpotent operator N of order m, a co-isometry V1 ∈ L(DN ∗ ⊕
M, DT ∗), and an isometry V2 ∈ L(DT , DN ⊕ M). Then the characteristic function ΘT of T
is a polynomial of degree less than or equal to m.
It is important to note that the conclusion of Theorem 3.2 depends explicitly on the de-
composition of T as used in the proof of Theorem 2.3. With the same setting as in Theorem
3.2, below we will show that the same conclusion holds for the following decomposition of T :
T = (cid:20)S X−1
0 T−1(cid:21) = (cid:20)S DS ∗Γ−1DT−1
T−1
0
(cid:21) ∈ L(H0 ⊕ K−1),
where K−1 = H0 ⊕ H−1,
T−1 = (cid:20)N X
0 C(cid:21) = (cid:20)N DN ∗ΓDC
C
0
(cid:21) ∈ L(H1 ⊕ H0),
CONTRACTIONS WITH POLYNOMIAL CHARACTERISTIC FUNCTIONS
9
and X−1 = DS ∗Γ−1DT1, X = DN ∗ΓDC, and Γ−1 in L(DT−1, DS ∗) and Γ in L(DC, DN ∗) are a
pair of contractions. In this case, again by Theorem 2.2, we have
(3.1)
and
(3.2)
where
(3.3)
and
(3.4)
ΘT = τ −1
−1∗(cid:20)ΘT−1
0
0
IDΓ∗
−1(cid:21) J[Γ−1](cid:20)0S
0
ΘT−1 = τ −1
∗ (cid:20)0C
0
0
IDΓ∗(cid:21) J[Γ](cid:20)ΘN
0
0
IDΓ−1(cid:21) τ−1,
0
IDΓ(cid:21) τ,
τ−1 : DT → {0H1} ⊕ DΓ−1,
τ−1∗ : DT ∗ → DT ∗
−1 ⊕ DΓ∗
−1,
τ : DT−1 → DN ⊕ DΓ,
τ∗ : DT ∗
−1 → {0H−1} ⊕ DΓ∗,
are unitary operators. Moreover
J[Γ−1] = (cid:20) Γ∗
DΓ∗
−1 DΓ−1
−1 −Γ−1(cid:21) ∈ L(DS ∗ ⊕ DΓ−1, DT−1 ⊕ DΓ∗
−1),
and
By setting
J[Γ] = (cid:20) Γ∗ DΓ
DΓ∗ −Γ(cid:21) ∈ L(DN ∗ ⊕ DΓ, DC ⊕ DΓ∗).
Ψ0 = (cid:20)0C
0
and using (3.1) and (3.2) we obtain
0
IDΓ∗(cid:21) and ΨN = (cid:20)ΘN
0
0
IDΓ(cid:21) ,
ΘT = τ −1
0
IDΓ∗
−1(cid:21) J[Γ−1](cid:20)0S
0
0
IDΓ∗
0
IDΓ−1(cid:21) τ−1
−1(cid:21) J[Γ−1](cid:20)0S
−1(cid:21)(cid:20)J[Γ]
0
IDΓ∗
0
0
0
IDΓ∗
0
IDΓ−1(cid:21) τ−1
−1(cid:21)(cid:20)ΨN
0
0
IDΓ∗
−1(cid:21)(cid:20)τ
0
0 IDΓ∗
−1(cid:21) J[Γ−1](cid:20)0S
0
0
IDΓ−1(cid:21) τ−1
0
= τ −1
= τ −1
−1∗(cid:20)ΘT−1
−1∗(cid:20)τ −1
−1∗(cid:20)τ −1
= V1(cid:20)ΨN
= V1(cid:20)ΘN
∗
0
0
0
∗ Ψ0J[Γ]ΨN τ
0
0
0
IDΓ∗
−1(cid:21)(cid:20)Ψ0
−1(cid:21) V2
0
IDΓ∗
0
−1(cid:21) V2,
IDΓ⊕DΓ∗
10
where
and
Hence
FOIAS, PEARCY, AND SARKAR
V1 = τ −1
∗
0
−1∗(cid:20)τ −1
−1∗(cid:20)τ −1
∗
0
= τ −1
0
IDΓ∗
0
IDΓ∗
0
−1(cid:21)(cid:20)Ψ0
−1(cid:21)(cid:20)0C
0
0
IDΓ∗
−1(cid:21)(cid:20)J[Γ]
0
0
IDΓ∗
−1(cid:21)
0
−1(cid:21)(cid:20)J[Γ]
0
0
IDΓ∗
−1(cid:21) ,
IDΓ∗ ⊕DΓ∗
V2 = (cid:20)τ
0
0 IDΓ∗
−1(cid:21) J[Γ−1](cid:20)0S
0
0
IDΓ−1(cid:21) τ−1.
0
ΘT = V1(cid:20)ΘN
= V1(cid:20)ΘN
0
−1(cid:21) V2
,
0
IDΓ⊕DΓ∗
0
I M(cid:21) V2
where M = DΓ ⊕ DΓ∗
isometric operators, that is, V1 V ∗
−1. Finally, by virtue of (3.3) and (3.4), we have that V ∗
V2 = IDT . Yet, we do not know if
1 = IDT ∗ and V ∗
2
1 and V2 are
where M is as in Theorem 3.2.
dim M = dim M,
Acknowledgement: The authors are grateful to the referee for pointing out that the formula-
tion of Theorem 3.2 in the submitted version was inappropriate. The example in Remark 3.3
is due to the referee.
The third author is supported in part by NBHM (National Board of Higher Mathematics,
India) grant NBHM/R.P.64/2014. The third author is also grateful for hospitality of Texas
A&M University, USA, during July 2016 and July-August 2015.
References
[1] B. Bagchi and G. Misra, Constant characteristic functions and homogeneous operators, J. Operator
Theory 37 (1997), 51 -- 65.
[2] L. de Branges and J. Rovnyak, Canonical models in quantum scattering theory, Perturbation Theory and
its Applications in Quantum Mechanics, C.H. Wilcox (ed.), 295-392. Wiley, New York, 1966.
[3] C. Foias and A. Frazho, The commutant lifting approach to interpolation problems, Operator Theory:
Advances and Applications, 44. Birkhauser Verlag, Basel, 1990.
[4] C. Foias and J. Sarkar, Contractions with polynomial characteristic functions I. Geometric approach,
Transaction of American Math. Society, 364 (2012), 4127 -- 4153.
[5] P. Halmos, A Hilbert space problem book, Graduate Texts in Mathematics, 19, Springer-Verlag, New
York-Berlin, 1982.
[6] B. Sz.-Nagy and C. Foias, Harmonic Analysis of Operators on Hilbert Space, North Holland, Amsterdam,
1970.
[7] B. Sz.-Nagy and C. Foias, Forme triangulaire d'une contraction et factorisation de la fonction car-
act´eristique, Acta Sci. Math. (Szeged) 28 (1967) 201-212.
CONTRACTIONS WITH POLYNOMIAL CHARACTERISTIC FUNCTIONS
11
[8] N. Nikolskii and V. Vasyunin, Notes on two function models, in: The Bieberbach Conjecture, West
Lafayette, IN, 1985, in: Math. Surveys Monogr., vol. 21, Amer. Math. Soc., Providence, RI, 1986, pp.
113-141.
[9] R. Teodorescu, Fonctions caract´eristiques constantes, Acta Sci. Math. (Szeged) 38 (1976), no. 1-2, 183 --
185.
[10] P. Y. Wu, Contractions with constant characteristic function are reflexive, J. London Math. Soc. (2) 29
(1984), 533 -- 544.
(Ciprian Foias) Department of Mathematics, Texas A&M University, College Station, Texas
77843, USA
(Carl Pearcy) Department of Mathematics, Texas A&M University, College Station, Texas
77843, USA
E-mail address: [email protected]
(Jaydeb Sarkar) Indian Statistical Institute, Statistics and Mathematics Unit, 8th Mile,
Mysore Road, Bangalore, 560059, India
E-mail address: [email protected], [email protected]
|
1003.3382 | 1 | 1003 | 2010-03-17T14:41:58 | Arens Regularity And 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 AND ABDOLHAMID RIAZI
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]
Department of Mathematics, Amirkabir University of Technology, Tehran, Iran
Email address: [email protected]
|
1509.03644 | 1 | 1509 | 2015-09-11T20:23:26 | Fundamental function for Grand Lebesgue Spaces | [
"math.FA"
] | We investigate in this short article the fundamental function for the so-called Grand Lebesgue Spaces (GLS) and show in particular a one-to-one and mutually continuous accordance between its fundamental and generating function. | math.FA | math |
FUNDAMENTAL FUNCTION FOR
GRAND LEBESGUE SPACES.
E.Ostrovsky, L.Sirota
Department of Mathematic, Bar-Ilan University, Ramat Gan, 52900, Israel,
e-mails: [email protected], [email protected]
Abstract.
We investigate in this short article the fundamental function for the so-called
Grand Lebesgue Spaces (GLS) and show in particular a one-to-one and mutually
continuous accordance between its fundamental and generating function.
Key words and phrases: Young-Orlicz function, ordinary and Grand Lebesgue
Spaces (GLS); Orlicz, GLS norms, rearrangement invariant spaces, fundamental and
generating function, Young-Fenchel, or Legendre transform, theorem of Fenchel-
Moraux, inverse function, Exponential Orlicz function (EOF) and Spaces (EOS).
Mathematics Subject Classification (2000): primary 60G17;
secondary 60E07;
60G70.
1 Notations. Statement of problem.
A. A triplet (X, B, µ), where X = {x} is arbitrary set, B is non-trivial certain
sigma-algebra of subsets X and µ is probabilistic: µ(X) = 1 diffuse non-negative
completely additive measure defined on the B.
The non-probabilistic case µ(X) = ∞ will be consider further.
Recall that the measure µ is said to be diffuse iff for arbitrary measurable set
A1 ∈ B with positive measure: µ(A1) > 0 there exists it subset A2 ⊂ A1 such that
µ(A2) = µ(A1)/2.
We denote as usually for any arbitrary measurable function f : X → R
f p = (cid:20)ZX
f (x)p µ(dx)(cid:21)1/p
, p ≥ 1;
Lp = {f, f p < ∞}.
B. The so-called Grand Lebesgue Space (GLS) Gψ with norm · Gψ is defined
( not only in this article) as follows:
Gψ = {f, f Gψ < ∞}, f Gψ def= sup
p≥1 " f p
ψ(p)# .
(1.1)
Here ψ = ψ(p), 1 ≤ p < ∞ is some continuous strictly increasing function such
that limp→∞ ψ(p) = ∞.
1
The detail investigation of this spaces (and more general spaces) see in [14], [19].
See also [5], [6], [8], [9], [10] etc.
The case when in (1.1) supremum is calculated over finite interval is investigated
in [14], [19], [20]:
Gbψ = {f, f Gbψ < ∞}, f Gbψ def= sup
1≤p<b" f p
ψ(p)# , b = const > 1,
(1.2)
but in (1.2) ψ = ψ(p) is continuous function in the semi-open interval 1 ≤ p < b
such that inf p∈(1,b) ψ(p) > 0.
We will denote
or simple b = b(ψ) := supp ψ(·), including the case b = ∞.
[1, b) := supp ψ(·),
Definition 1.1. The function = ψ(p) which appeared in (1.1) and (1.2), will
be named as generating function for the correspondent Banach space Gψ.
An used further example:
ψ(β,b)(p) = (b − p)−β, 1 ≤ p < b, β = const > 0; b = const > 1,
Gβ,b(p) := Gbψ(β,b)(p).
C. We denote as ordinary for any measurable set A, A ∈ B it indicator function
by I(A) = IA(ω).
D. The Grand Lebesgue Spaces {Gψ} are rearrangement invariant in the
[2], chapter 1. Therefore, its fundamental function
classical definition, see e.g.
φG(ψ)(δ), δ ≥ 0 is correctly defined in the considered case as follows:
φG(ψ)(δ) def= sup
p∈supp ψ" δ1/p
ψ(p)# ,
φG(ψ)(1) =
1
inf p∈supp ψ ψ(p)
.
see [2], chapters 2 and 5.
For instance,
Note also
(1.3)
(1.3a)
φG(C·ψ) = φG(ψ)/C, C = const > 0.
(1.3b)
This notion play a very important role in the functional analysis, [2], [22], [23];
in the theory of interpolation of operators, [2], [5], [7], in the theory of probability
[13], [15], [16], [17]; in the theory of Partial Differential equations [7], [9]; in the
2
theory of martingales [20]; in the theory of approximation, in the theory of random
processes etc.
E. Let g = g(p), p ∈ (a, b), 1 ≤ a < b ≤ ∞ be some numerical valued continuous
strictly increasing (or decreasing) function. The inverse function will be denoted
by g(−1)(z), g(a) ≤ z ≤ g(b), in contradistinction to the usually notation g−1(p) =
1/g(p).
F. The Young-Fenchel, or Legendre transform g∗(q) for the function g = g(p)
one can to define
g∗(q) def= sup
p∈supp g
(p q − g(p)).
(1.4)
Our goal in this short report is to establish a one-to-one and mutually
continuous connection between the fundamental and generating functions
for the Grand Lebesgue Spaces.
In some previous works: [6], [12], chapter 8; [14], [19], [18], [22] these function
was evaluated and applied in many practical cases.
2 Main result.
Problem A. Let the generating function ψ be a given: ψ ∈ GΨ(a, b), 1 ≤
a < b ≤ ∞. Find the fundamental function for the correspondent Grand
Lebesgue Space Gψ.
Suppose the function
p →
p
ψ(p)
, p ∈ (a, b)
is strictly increasing; and define therefore the function
ν(p) = νψ(p) = " p
ψ(p)#(−1)
, p ∈ supp ψ,
(2.1)
and ν(p) = +∞ otherwise.
Introduce also the following Young-Orlicz function
and define finally
N(u) = Nψ(u) := exp(cid:16)ν ∗
ψ(u)(cid:17) − exp(cid:16)ν ∗
ψ(0)(cid:17) ,
θ(δ) = θψ(δ) def=
1
N (−1)
ψ
(1/δ)
, δ > 0.
(2.2)
(2.3)
Proposition 2.1. We propose under formulated above conditions, for instance,
µ(X) = 1, diffuseness of the measure µ, and in the case when b = ∞
3
φG(ψ)(δ) = θψ(δ), δ > 0.
(2.4)
Remark 2.1. The equality (2.4) is more convenient than source definition (1.3).
In particular, it allows for a relatively simple inversion.
Proof is very simple; it based on the computation of the fundamental function
for Orlicz spaces, see the book of Krasnosel'skii M.A. and Rutickii Ya.B. [11], chapter
3; see also the classical monographs [26], [27].
In detail, it is proved in particular in the articles [14], [18], [19] that under our
conditions the Grand Lebesgue Space Gψ coincides with certain Orlicz space over
source probability triplet (X, B, µ) relative the Young-Orlicz function Nψ(u).
We deduce reducing considered case to the well-known calculation of fundamen-
tal function for Orlicz space, [11], chapter 3
φG(ψ)(δ) =
Q.E.D.
1
N (−1)
ψ
(1/δ)
= θψ(δ), δ > 0,
(2.4)
An inverse problem B. Let the fundamental function φGψ(δ) = φ(δ) be
a given. Find the correspondent generating function ψ(p).
A first restrictions: the function φ = φ(δ) is strictly increasing and continuous;
in particular φ(0+) = φ(0) = 0.
We find from the equality (2.4)
or equivalently
Nψ(1/δ) = 1
φ(δ)!(−1)
,
Nψ(z) = 1
φ(δ)!(−1)
.δ:=1/z. .
(2.5)
(2.5a)
A second restriction: the function
V ∗(z) = ln(C + Nψ(z)), z ≥ 0,
where V ∗(0) = ln C, is continuous and upward convex.
It follows immediately from (2.6) by virtue of theorem of Fenchel-Moraux
Since
we derive finally
V (z) = {ln(C + Nψ(z))}∗ , z ≥ 0.
V (p) = " p
ψ(p)#(−1)
,
(2.6)
(2.7)
Proposition 2.2. We conclude under formulated in this pilcrow conditions
4
ψ(p) =
p
V (−1)(p)
.
(2.8)
3 The case of infinite measure.
The case when µ(X) = ∞ is more complicated.
Recall first of all definition and some facts about the so-called Exponential Orlicz
Spaces (EOS), see for example [21].
Let N = N(u) be an N − Young-Orlicz's function, i.e. downward convex,
even, continuous function differentiable for all sufficiently greatest values u, u ≥
u0, u0 = const > 0, strongly increasing along the right semi-axis and such that
N(u) = 0 ⇔ u = 0; u → ∞ ⇒ dN(u)/du → ∞. We can say that N(·)
is an exponential Orlicz function, briefly, N(·) ∈ EOF, if N(u) has a form of a
continuous differentiable strongly increasing downward convex function W = W (u)
in the domain [2, ∞] such that u → ∞ ⇒ W /(u) → ∞ and
N(u) = N(W, u) = exp(W (log u)), u ≥ e2.
For the values u ∈ [−e2, e2] we define N(W, u) arbitrarily, but so that the function
N(W, u) is even, continuous, convex, strictly increasing along the right semi-axis
and so that N(u) = 0 ⇔ u = 0. We denote the correspondent Orlicz space on
(X, µ) with a measure µ and with N − function of the form N(W, u) as L(N) =
EOS(W ); EOS = ∪W {EOS(W )} (exponential Orlicz's space).
For example, let m = const > 0, r = const ∈ R1,
Nm,r(u) = exphum (cid:16)log−mr(C1(r) + u)(cid:17)i − 1,
C1(r) = e, r ≤ 0; C1(r) = exp(r), r > 0. Then Nm,r(·) ∈ EOS. In the case r = 0
we can write Nm = Nm,0.
Recall that the Orlicz's norm on the arbitrarily measurable space (X, A, µ)
f L(N) = f L(N, X, µ) can be calculated by the following formula (see, for
example, [11], p. 66; [26], p. 73 )
f L(N) = inf
v>0(cid:26)v−1(cid:18)1 +ZX
N(vf (x)) µ(dx)(cid:19)(cid:27) .
Let α be arbitrary number, α = const ≥ 1, and N(·) ∈ EOS(W ) for some
W = W (·). For such a function N = N(W, u) we denote by N (α)(W ; u) = N (α)(u)
a new Young-Orlicz's function N (α)(u) such that
N (α)(u) = C1 uα,
N (α)(u) = C3 + C4u,
u ∈ [0, C2];
u ∈ (C2, C5];
N (α)(u) = N(u),
u > C5, 0 < C2 < C5 < ∞,
(3.1)
C1,2,3,4,5 = C1,2,3,4,5(α, N(·)).
5
In the case of α = m(j +1), m > 0, j = 0, 1, 2, . . . the function N (α)
to the following Trudinger's function:
m (u) is equivalent
N (α)
m (u) ∼ N (α)
[m] (u) = exp (um) −
uml/l!.
j
Xl=0
This method is described in [28], p. 42-47. These Orlicz spaces are applicable to
the theory of non-linear partial differential equations.
We denote hereinafter generally by Ck = Ck(·), k = 1, 2, . . . some positive finite
essentially constructive constants, and by C, C0 non-essentially constants, also con-
structive. We proved the existence of constants C1,2,3,4,5 = C1,2,3,4,5(α, N(·)) such
that N (α) is a new exponential N Orlicz's function in [21]. We denote classical
absolute constants by the symbols Kj.
Now we introduce some new Grand Lebesgue Spaces. Let ψ = ψ(p), p ≥ α, α =
const ≥ 1 be a continuous positive ψ(α) > 0 finite strictly increasing function such
that the function p → p log ψ(p) is downward convex, and
We denote the set of all these functions by Ψ; Ψ = {ψ}. A particular case
lim
p→∞
ψ(p) = ∞.
where
ψ(p) = ψ(W ; p) = exp(W ∗(p)/p),
W ∗(p) = sup
z≥α
(pz − W (z))
is so-called Young-Fenchel, or Legendre transform of W (·). It follows from the the-
orem of Fenchel-Moraux that in this case
W (p) = [p log ψ(W ; p)]∗ , p ≥ p0 = const ≥ 2,
and, consequently, for all ψ(·) ∈ Ψ we introduce a correspondent Young-Orlicz N −
function by the equality:
N([ψ]) = N([ψ], u) = exp {[p log ψ(p)]∗ (log u)} , u ≥ e2.
Definition 3.1. We introduce for such arbitrary function ψ(·) ∈ Ψ the so-
called G(α; ψ) norms and correspondent Banach GLS space G(α; ψ) as a set of all
measurable (complex) functions with finite norms:
f G(α; ψ) = sup
p≥α
(f p/ψ(p)).
(3.2)
For instance, ψ(p) may be ψ(p) = ψm(p) = p1/m, m = const > 0; in this case,
we can write G(α, ψm) = G(α, m) and
f G(α, m) = sup
p≥α(cid:16)f p p−1/m(cid:17) .
6
Theorem A, see [21]. Let the measure µ be diffuse, µ(X) = ∞, α = const ≥ 1
and ψ ∈ Ψ. We assert that the norms Orlicz-Luxemburg norm · L(N (α), [ψ]) and
Grand Lebesgue Space norm · G(α, ψ), α ≥ 1 are equivalent.
Arguing similarly to the second section, we obtain the following result.
Proposition 3.1. We propose under conditions of theorem A
φG(ψ)(δ) ≍ θψ(δ), δ > 0.
Remark 3.1. Note that this case δ ∈ (0, ∞), in contradiction to the proposition
2.1, where it in naturally to take δ ∈ (0, 1).
4 Concluding remarks. Open problems.
It is interest by our opinion to investigate the notion of fundamental function
and also its relation with generating function for the so-called mixed, or equally
anisotropic Grand Lebesgue Spaces.
Recall that the definition of mixed, or equivalently anisotropic ordinary Lebesgue
Spaces appeared at first in the article [1] and was investigated in detail in the classical
books [3], [4].
The anisotropic Grand Lebesgue Spaces as a slight generalization of Lp spaces
arises in turn in [23] with the correspondent fundamental function; in the preprint
[24] both this notions was applied in the operator's theory.
References
[1] Benedek A. and Panzone R. The space L(p) with mixed norm. Duke Math.
[2] Bennet C., Sharpley R. Interpolation of operators. Orlando, Academic
Press Inc., (1988).
[3] Besov O.V., Ilin V.P., Nikolskii S.M. Integral representation of functions
and imbedding theorems. Vol.1; Scripta Series in Math., V.H.Winston and Sons,
(1979), New York, Toronto, Ontario, London.
[4] Besov O.V., Ilin V.P., Nikolskii S.M. Integral representation of functions
and imbedding theorems. Vol.2; Scripta Series in Math., V.H.Winston and Sons,
(1980), New York, Toronto, Ontario, London.
[5] Capone C., Fiorenza A., Krbec M. On the Extrapolation Blowups in the
Lp Scale. Collectanea Mathematica, 48, 2, (1998), 71-88.
[6] Fiorenza A. Duality and reflexivity in grand Lebesgue spaces. Collectanea
Mathematica (electronic version), 51, 2, (2000), 131-148.
7
[7] Fiorenza A., and Karadzhov G.E. Grand and small Lebesgue spaces and
their analogs. Consiglio Nationale Delle Ricerche, Instituto per le Applicazioni
del Calcoto Mauro Picine, Sezione di Napoli, Rapporto tecnico n. 272/03,
(2005).
[8] Iwaniec T., and Sbordone C. On the integrability of the Jacobian under
minimal hypotheses. Arch. Rat.Mech. Anal., 119, (1992), 129143.
[9] Iwaniec T., P. Koskela P., and Onninen J. Mapping of finite distortion:
Monotonicity and Continuity. Invent. Math. 144 (2001), 507-531.
[10] Kozachenko Yu. V., Ostrovsky E.I. (1985). The Banach Spaces of ran-
dom Variables of subgaussian type. Theory of Probab. and Math. Stat. (in
Russian). Kiev, KSU, 32, 43-57.
[11] Krasnosel'skii, M.A.; Rutickii, Ya.B. (1961). Convex Functions and Or-
licz Spaces. Groningen: P.Noordhoff Ltd
[12] A. Kufner, O. John and S. Fucik. Function Spaces. Noordhoff Interna-
tional Publishingr, Leyden, 1977.
[13] Ledoux M., Talagrand M. (1991) Probability in Banach Spaces. Springer,
Berlin, MR 1102015.
[14] Liflyand E., Ostrovsky E., Sirota L. Structural Properties of Bilateral
Grand Lebesgue Spaces. Turk. Journal of Math., 34, (2010), 207-219.
[15] Ostrovsky E., Rogover E. Exact exponential Bounds for the random field
Maximum Distribution via the majorizing Measures (Generic Chaining). arX-
iv:0802.0349v1 [math.PR] 4 Feb 2008.
[16] Ostrovsky E.I. (1999). Exponential estimations for random Fields and its
applications. (in Russian). Moscow-Obninsk, OINPE.
[17] Ostrovsky E.I. (1994.) Exponential Bounds in the Law of Iterated Logarithm
in Banach Space. Math. Notes, 56, 5, p. 98-107.
[18] Ostrovsky E., Sirota L. Fouier Transforms in exponential rearrangement
invaiant Spaces.
arXiv:040639v1 [math.FA], 20 Jun 2004.
[19] Ostrovsky E. and Sirota L. Moment Banach spaces: theory and applica-
tions. HIAT Journal of Science and Engineering, C, Volume 4, Issues 1-2, pp.
233-262, (2007).
[20] Ostrovsky E. and Sirota L. Moment and Tail Inequalities for polynomial
Martimgales. The case of heavy tails.
arXiv: 1112.2768v1 [math.PR] 13 Dez 2011.
[21] Ostrovsky E. and Sirota L. Fourier transforms in exponential rearrange-
ment invariant spaces.
arXiv:math/0406391v1 [math.FA] 20 Jun 2004
8
[22] Ostrovsky E. Exponential Orlicz Spaces: New Norms and Applications.
arXiv:math/0406534v1 [math.FA] 25 Jul 2004
[23] Ostrovsky E. and Sirota L. Exact norm estimates for multivariate di-
lation operators between two Bilateral Weight Grand Lebesgue Spaces. arX-
iv:1503.05235v1 [math.FA] 17 Mar 2015
[24] Ostrovsky E. and Sirota L. Central Limit Theorem and exponential tail
estimation in mixed (anisotropic) Grand Lebesgue Spaces. arXiv:1308.5606v1
[math.PR] 26 Aug 2013
[25] Pizier G. Condition d entropic assupant la continuite de certains processus et
applications a lanalyse harmonique. Seminaire d analyse fonctionalle. (1980),
Exp.13, p. 23-34.
[26] Rao M.M., Ren Z.D. Theory of Orlicz Spaces. Marcel Dekker Inc., 1991.
New York, Basel, Hong Kong.
[27] Rao M.M., Ren Z.D. Applications of Orlicz Spaces. Marcel Dekker Inc., 2002.
New York, Basel, Hong Kong.
[28] Taylor M.E. Partial Differential Equations.
v.3, Nonlinear Equations.
Springer Verlag, 1991. Berlin, Heidelberg, New York.
[29] Talagrand M. (1996). Majorizing measure: The generic chaining. Ann.
Probab., 24 1049-1103. MR1825156
9
|
1602.08790 | 2 | 1602 | 2016-05-14T15:33:04 | Product operators on mixed norm spaces | [
"math.FA"
] | Inequalities for product operators on mixed norm Lebesgue spaces and permuted mixed norm Lebesgue spaces are established. They depend only on inequalities for the factors and on the Lebesgue indices involved. Inequalities for the bivariate Laplace transform are given to illustrate the method. Also, an elementary proof is presented for an $n$-variable Young's inequality in mixed norm spaces. | math.FA | math |
PRODUCT OPERATORS ON MIXED NORM SPACES
WAYNE GREY AND GORD SINNAMON
Abstract. Inequalities for product operators on mixed norm Lebesgue spaces
and permuted mixed norm Lebesgue spaces are established. They depend only
on inequalities for the factors and on the Lebesgue indices involved. Inequal-
ities for the bivariate Laplace transform are given to illustrate the method.
Also, an elementary proof is presented for an n-variable Young's inequality in
mixed norm spaces.
1. Introduction
The techniques used to study embeddings of mixed norm spaces in [6] and [8]
can be extended to work with operators other than the identity. Here we begin this
work by considering a select class of operators. Mixed norm spaces are spaces of
multivariable functions in which the norm takes advantage of the product structure
in the domain. They have a long informal history but were first named and formally
studied by Benedek and Panzone in [4]. Permuted mixed norms only arise when
studying more than a single space, since they occur when the order in which the
factor norms are taken is different in different spaces. The importance of permuted
mixed norms was noted in [5] and [11, §XI.1]. They also appear in [3], a study
of the Hardy operator in mixed-norm Lebesgue spaces, and in the papers [10] and
[2], in which partial integral operators are considered between mixed-norm Banach
function spaces, with Lebegue and Orlicz norms as special cases. A systematic study
of continuous inclusions between mixed-norm Lebesgue spaces was undertaken in
the first author's thesis, [6].
It will be convenient to introduce the two-variable mixed norm spaces needed
in Section 2 first, postponing the introduction of n-variable spaces until Section
3. Let (T1, λ1) and (T2, λ2) be σ-finite measure spaces and let L0
denote the
collection of (λ1 × λ2)-measurable functions on T1 × T2. In particular, either or
both of λ1 and λ2 may be "counting" measure. Thus, the definitions of this section
and the results of the next apply to weighted or unweighted sums as special cases.
λ1×λ2
Fix indices p1, p2 ∈ (0, ∞). For any f ∈ L0
λ1×λ2 ,
kf kL(p1,p2)
λ1×λ2
= ZT2(cid:18)ZT1
f (t1, t2)p1 dλ1(t1)(cid:19)p2/p1
dλ2(t2)!1/p2
.
The first variable of the function f is always in the λ1 measure space, and the order
of the indices and measures indicates which variable is the "inner" one. So for any
2010 Mathematics Subject Classification. Primary 46E30, Secondary 44A35, 26D15.
Key words and phrases. mixed norm, permuted mixed norm, product operator, Young's in-
equality, bivariate Laplace transform.
Supported by the Natural Sciences and Engineering Research Council of Canada.
1
2
WAYNE GREY AND GORD SINNAMON
f ∈ L0
λ1×λ2
,
kf kL(p2,p1)
λ2×λ1
= ZT1(cid:18)ZT2
f (t1, t2)p2 dλ2(t2)(cid:19)p1/p2
dλ1(t1)!1/p1
.
Although these are genuine norms only when p1 ≥ 1 and p2 ≥ 1 we will refer to
them as mixed norms even when some indices are less than 1. Also, the results
we present will often extend to the case when one or both of the indices is infinite,
indicating that the supremum norm is to be taken in that factor.
The product operators that we consider will be introduced in Section 2. As
mentioned above, we restrict our attention to the two-variable case. Although the
extension from bivariate operators to multivariate operators may seem straightfor-
ward, the delicate arguments needed in the embedding case in [6] and [8] and the
advanced techniques introduced in [7] put this extension beyond the scope of the
present article.
Convolution operators have some product structure but not enough to make
them product operators in general. Convolution is considered in Section 3, where
it leads to an elementary proof of a multivariate mixed norm Young's inequality.
The Hardy operators considered in [3] are product operators (and also convolu-
tion operators) with the added advantage that weighted norm inequalities for the
factors have been characterized. The results may be compared with Theorem 2.1,
below. Hardy operators are not the only operators with these useful properties, see
for example, the Riemann-Liouville operators studied in [13].
We will make frequent use of Minkowski's (integral) inequality in the following
mixed norm form: If 0 < p1 ≤ p2 < ∞, then,
≤ kf kL(p2,p1)
kf kL(p1,p2)
λ1×λ2
λ2×λ1
,
f ∈ L+
λ1×λ2
.
For easy recognition, we will enclose the function in square brackets when applying
Minkowski's inequality in integral estimates. It becomes,
ZT2(cid:18)ZT1
[f (t1, t2)]p1 dλ1(t1)(cid:19)p2/p1
≤ ZT1(cid:18)ZT2
dλ2(t2)!1/p2
[f (t1, t2)]p2 dλ2(t2)(cid:19)p1/p2
dλ1(t1)!1/p1
.
2. Product operators
Let (T1, λ1), (T2, λ2), (X1, µ1), and (X2, µ2) be σ-finite measure spaces. Let
denote the collection of non-negative (λ1 × λ2)-measurable functions.
→ L+
λ1×λ2
µ1×µ2 will be called a product operator provided it
L+
λ1×λ2
An operator K : L+
can be expressed in the form,
Kf (x1, x2) =ZZT1×T2
k1(x1, t1)k2(x2, t2)f (t1, t2) d(λ1 × λ2)(t1, t2),
where kj is a non-negative (µj × λj )-measurable function for j = 1, 2. In this case
we define Kj : L+
λj
→ L+
µj by
Kjf (x) =ZTj
kj(x, t)f (t) dλj (t),
j = 1, 2.
PRODUCT OPERATORS ON MIXED NORM SPACES
3
Note that, by Tonelli's theorem,
Kf (x1, x2) =ZT1
=ZT2
k1(x1, t1)ZT2
k2(x2, t2)ZT1
k2(x2, t2)f (t1, t2) dλ2(t2) dλ1(t1)
k1(x1, t1)f (t1, t2) dλ1(t1) dλ2(t2).
Suppose p1, p2, r1, and r2 are positive and let Cj be the least constant, finite or
infinite, such that
kKjf kL
rj
µj
≤ Cjkf kL
,
pj
λj
f ∈ L+
λj
.
Theorem 2.1. Fix positive indices p1, p2, r1, and r2. Suppose K is a product
operator, with kj , Kj, Cj as above for j = 1, 2.
(a) If r1 ≥ 1 then K satisfies the mixed norm inequality,
kKf kL(r1,r2 )
µ1×µ2
≤ C1C2kf kL(p1,p2)
λ1×λ2
,
f ∈ L+
λ1×λ2
.
(b) If r1 ≥ 1, r2 ≥ 1 and min(p1, r1) ≤ max(p2, r2) then K satisfies the permuted
mixed norm inequality,
kKf kL(r1,r2 )
µ1×µ2
≤ C1C2kf kL(p2,p1)
λ2×λ1
,
f ∈ L+
λ1×λ2
.
Proof. In a slight abuse of notation we write,
K1f (x1, t2) =ZT1
K2f (t1, x2) =ZT2
k1(x1, t1)f (t1, t2) dλ1(t1) and
k2(x2, t2)f (t1, t2) dλ2(t2).
To prove part (a), use the hypothesis r1 ≥ 1 to apply Minkowski's inequality
and then invoke the definitions of C1 and C2. This yields,
Kf (x1, x2)r1 dµ1(x1)(cid:19)r2/r1
dµ2(x2)!1/r2
ZX2(cid:18)ZX1
= ZX2(cid:18)ZX1(cid:18)ZT2hk2(x2, t2)K1f (x1, t2)i dλ2(t2)(cid:19)r1
≤ ZX2 ZT2(cid:18)ZX1hk2(x2, t2)K1f (x1, t2)ir1
= ZX2 ZT2
≤ C1 ZX2 ZT2
≤ C1C2 ZT2(cid:18)ZT1
dµ1(x1)(cid:19)1/r1
K1f (x1, t2)r1 dµ1(x1)(cid:19)1/r1
f (t1, t2)p1 dλ1(t1)(cid:19)1/p1
dλ2(t2)!1/p2
k2(x2, t2)(cid:18)ZX1
k2(x2, t2)(cid:18)ZT1
f (t1, t2)p1 dλ1(t1)(cid:19)p2/p1
.
1/r2
dµ1(x1)(cid:19)r2/r1
dλ2(t2)!r2
dλ2(t2)!r2
dλ2(t2)!r2
dµ2(x2)!1/r2
dµ2(x2)!
dµ2(x2)!
dµ2(x2)!
1/r2
1/r2
Part (b) will be done in four cases: They arise from the observation that the
condition min(p1, r1) ≤ max(p2, r2) is satisfied if and only if one of, p1 ≤ p2, r1 ≤ r2,
4
WAYNE GREY AND GORD SINNAMON
p1 ≤ r2, or r1 ≤ p2 holds.
inequality, yields
If p1 ≤ p2 then part (a), followed by Minkowski's
kKf kL(r1,r2)
µ1×µ2
≤ C1C2kf kL(p1,p2)
λ1×λ2
≤ C1C2kf kL(p2,p1)
λ2×λ1
,
f ∈ L+
λ1×λ2
.
If r1 ≤ r2 then we begin by using Minkowski's inequality, and follow with part (a)
to get
kKf kL(r1,r2 )
µ1×µ2
≤ kKf kL(r2,r1 )
µ2×µ1
≤ C2C1kf kL(p2,p1)
λ2×λ1
,
f ∈ L+
λ1×λ2
.
If p1 ≤ r2 then we apply the definition of C1 followed by Minkowski's inequality
and then the definition of C2 to get
Kf (x1, x2)r1 dµ1(x1)(cid:19)r2/r1
dµ2(x2)!1/r2
k1(x1, t1)K2f (t1, x2) dλ1(t1)(cid:19)r1
ZX2(cid:18)ZX1
= ZX2(cid:18)ZX1(cid:18)ZT1
≤ C1 ZX2(cid:18)ZT1hK2f (t1, x2)ip1
≤ C1 ZT1(cid:18)ZX2hK2f (t1, x2)ir2
≤ C1C2 ZT1(cid:18)ZT2
dλ1(t1)(cid:19)r2/p1
dµ2(x2)(cid:19)p1/r2
f (t1, t2)p2 dλ2(t2)(cid:19)p1/p2
dµ2(x2)!1/r2
dλ1(t1)!1/p1
dλ1(t1)!1/p1
.
dµ1(x1)(cid:19)r2/r1
dµ2(x2)!1/r2
If r1 ≤ p2 the process is somewhat lengthy, using the definitions of C1 and C2
as well as three applications of Minkowski's inequality. First, apply Minkowski's
inequality with r1 ≥ 1, followed by the definition of C2 to get
Kf (x1, x2)r1 dµ1(x1)(cid:19)r2/r1
dµ2(x2)!1/r2
ZX2(cid:18)ZX1
= ZX2(cid:18)ZX1(cid:18)ZT2hk2(x2, t2)K1f (x1, t2)i dλ2(t2)(cid:19)r1
≤ ZX2 ZT2(cid:18)ZX1hk2(x2, t2)K1f (x1, t2)ir1
= ZX2 ZT2
k2(x2, t2)(cid:18)ZX1
≤ C2 ZT2(cid:18)ZX1hK1f (x1, t2)ir1
dµ1(x1)(cid:19)1/r1
K1f (x1, t2)r1 dµ1(x1)(cid:19)1/r1
dµ1(x1)(cid:19)p2/r1
dλ2(t2)!1/p2
.
dµ1(x1)(cid:19)r2/r1
dλ2(t2)!r2
dλ2(t2)!r2
dµ2(x2)!1/r2
dµ2(x2)!
dµ2(x2)!
1/r2
1/r2
PRODUCT OPERATORS ON MIXED NORM SPACES
5
Now Minkowski's inequality with p2 ≥ r1 shows that the last expression is no larger
than
C2 ZX1(cid:18)ZT2hK1f (x1, t2)ip2
dλ2(t2)(cid:19)r1/p2
dµ1(x1)!1/r1
.
After expanding K1f (x1, t2) in this expression we may apply Minkowski's inequality
with p2 ≥ 1 to get
C2 ZX1(cid:18)ZT2(cid:18)ZT1hk1(x1, t1)f (t1, t2)i dλ1(t1)(cid:19)p2
≤ C2 ZX1 ZT1(cid:18)ZT2hk1(x1, t1)f (t1, t2)ip2
= C2 ZX1 ZT1
≤ C1C2 ZT1(cid:18)ZT2
k1(x1, t1)(cid:18)ZT2
f (t1, t2)p2 dλ2(t2)(cid:19)p1/p2
dλ2(t2)(cid:19)1/p2
f (t1, t2)p2 dλ2(t2)(cid:19)1/p2
dλ1(t1)!1/p1
,
dλ2(t2)(cid:19)r1/p2
dλ1(t1)!r1
dλ1(t1)!r1
dµ1(x1)!1/r1
dµ1(x1)!
dµ1(x1)!
1/r1
1/r1
where the last inequality uses the definition of C1. These estimates complete the
proof of the fourth and last case.
(cid:3)
The index conditions in Theorem 2.1(b) are somewhat stronger than needed.
It is enough to assume that one (or more) of the following four conditions holds:
p1 ≤ r2; 1 ≤ r1 ≤ p2; 1 ≤ r1 and p1 ≤ p2; or 1 ≤ r2 and r1 ≤ r2. The statement
of the theorem remains valid when some indices are infinite, provided the index
conditions are met, but the proofs would have to be modified to accommodate
occurrences of the supremum norm.
The inequalities of Theorem 2.1 are stated for non-negative functions only but
it is routine to extend the operator K to all functions for which the right hand side
is finite, in a way that preserves the norm inequalities.
Also, the results of Theorem 2.1 may be seen to hold for a more general class of
positive operators. Instead of supposing that the factors K1 and K2 are integral
operators with non-negative kernels, it would suffice to assume that they map
positive functions to positive functions and possess formal adjoints. See [12, Lemma
2.4] for properties of such operators and [9, Section 4] for connections with positive
integral operators. One advantage of such an extension is that the identity operator
is not an integral operator (in general) but it does have a formal adjoint.
Corollary 2.2. Suppose p1, p2 ∈ (1, 2]. The bivariate Laplace transform L2, de-
fined by
L2f (x1, x2) =Z ∞
0 Z ∞
0
satisfies the permuted mixed norm inequality
e−x1t1−x2t2 f (t1, t2) dt1 dt2,
x1 > 0, x2 > 0,
kL2f kL(p′
1 = 1/p2 + 1/p′
1
) ≤ C(p1)C(p2)kf kL(p2,p1) .
,p′
2
Here 1/p1 + 1/p′
Laplace transform as a map from Lp to Lp′
.
2 = 1, and C(p) is the norm of the single-variable
6
WAYNE GREY AND GORD SINNAMON
Proof. The bivariate Laplace transform is a product operator, in the above sense,
whose factors are both the single-variable Laplace transform, L , given by
L f (x) =Z ∞
0
e−xtf (t) dt,
x > 0.
It is well known that L is bounded as a map from Lp to Lp′
when 1 < p ≤ 2. Since p′
the second statement of Theorem 2.1 gives the conclusion.
2 ≥ 1, and min(p1, p′
1 ≥ 1, p′
1) = p1 ≤ 2 ≤ p′
, i.e. C(p) is finite,
2)
(cid:3)
2 = max(p2, p′
Note that if 1 < p2 < p1 ≤ 2, the permuted mixed norm inequality given above
can be used to refine the unpermuted one given by Theorem 2.1(a) in two different
ways. Corollary 2.2 and Minkowski's inequality give
kL2f kL(p′
1
,p′
2
) ≤ C(p1)C(p2)kf kL(p2,p1) ≤ C(p1)C(p2)kf kL(p1,p2) ,
and also,
kL2f k
L(p′
1,p′
2) ≤ kL2f k
L(p′
2,p′
1) ≤ C(p1)C(p2)kf kL(p1,p2) .
Using Theorem 2.1 it is easy to generate additional permuted mixed norm in-
equalities, by beginning with known single-variable inequalities. See, for example,
[1, Corollary 1] for more single-variable inequalities involving the Laplace transform.
3. Young's inequality
Fix a positive integer n and let L+ be the collection of non-negative Lebesgue
measurable functions on Rn. For P = (p1, . . . , pn) ∈ [1, ∞)n define
kf kLP =
ZR
. . . ZR(cid:18)ZR
f (t1, . . . , tn)p1 dt1(cid:19)p2/p1
dt2!p3/p2
1/pn
.
. . . dtn
The convolution of two (real-valued) functions on Rn is defined by,
f ∗ g(x) =ZRn
whenever the integral exists.
f (x − t)g(t) dt,
For a fixed function g the map f 7→ f ∗ g has a kind of product structure. But,
even in the two-variable case, it is not a product operator of the sort considered
in the previous section unless g factors as g(t1, t2) = g1(t1)g2(t2). This will keep
us from establishing permuted mixed norm inequalities. Nevertheless, exploiting
the existing product structure provides an elementary proof of a n-variable mixed
norm Young's inequality.
Recall the single-variable Young's inequality: If p, q, r ∈ [1, ∞] satisfy 1/p+1/q =
1/r + 1, f ∈ Lp and g ∈ Lq, then f ∗ g is well defined and kf ∗ gkLr ≤ kf kLpkgkLq.
Theorem 3.1. Suppose P = (p1, . . . , pn), Q = (q1, . . . , qn), and R = (r1, . . . rn)
satisfy,
1
pj
+
1
qj
=
1
rj
+ 1,
j = 1, . . . , n.
If f ∈ LP and g ∈ LQ then f ∗ g is well defined, and
kf ∗ gkLR ≤ kf kLP kgkLQ.
PRODUCT OPERATORS ON MIXED NORM SPACES
7
Proof. We begin by proving the inequality when f and g are non-negative, so that
f ∗ g is well-defined as a function taking values in [0, ∞].
As the first step in a recursive argument, let f1 = f and g1 = g. Define x and
t so that (x1, x2, . . . , xn) = (x1, x) and (t1, t2, . . . , tn) = (t1, t). Also, let dt denote
dt2, . . . dtn. Then Young's inequality gives, for each x and t,
(cid:18)ZR(cid:18)ZR
f1(x1 − t1, x − t)g1(t1, t) dt1(cid:19)r1
dx1(cid:19)
1/r1
≤(cid:18)ZR
f1(y1, x − t)p1 dy1(cid:19)1/p1(cid:18)ZR
g1(t1, t)q1 dt1(cid:19)1/q1
.
So, applying Minkowski's inequality, we get
f1 ∗ g1(x1, x)r1 dx1(cid:19)1/r1
(cid:18)ZR
=(cid:18)ZR(cid:18)ZRn−1(cid:20)ZR
≤ZRn−1(cid:18)ZR(cid:20)ZR
≤ZRn−1(cid:18)ZR
f1(x1 − t1, x − t)g1(t1, t) dt1(cid:21) dt(cid:19)r1
f1(x1 − t1, x − t)g1(t1, t) dt1(cid:21)r1
dx1(cid:19)
dx1(cid:19)
1/r1
dt
1/r1
f1(y1, x − t)p1 dy1(cid:19)1/p1(cid:18)ZR
g1(t1, t)q1 dt1(cid:19)1/q1
dt.
Now define f2, g2 : Rn−1 → [0, ∞] by
We have just shown that,
Applying the above argument to f2 and g2 gives
f1(y1, y2, . . . , yn)p1 dy1(cid:19)1/p1
g1(t1, t2, . . . , tn)q1 dt1(cid:19)1/q1
.
f2(y2, . . . , yn) =(cid:18)ZR
g2(t2, . . . , tn) =(cid:18)ZR
f1 ∗ g1(x1, x2, . . . , xn)r1 dx1(cid:19)1/r1
f2 ∗ g2(x2, x3, . . . , xn)r2 dx2(cid:19)1/r2
f3(y3, . . . , yn) =(cid:18)ZR
g3(t3, . . . , tn) =(cid:18)ZR
fn−1 ∗ gn−1(xn−1, xn)rn−1 dxn−1(cid:19)1/rn−1
f2(y2, y3, . . . , yn)p2 dy2(cid:19)1/p2
g2(t2, t3, . . . , tn)q2 dt2(cid:19)1/q2
.
and
(3.1)
(cid:18)ZR
(cid:18)ZR
(3.2)
where
and
(3.3)
(cid:18)ZR
≤ f2 ∗ g2(x2, . . . , xn).
≤ f3 ∗ g3(x3, . . . , xn),
We continue in this way until reaching
≤ fn ∗ gn(xn),
8
where
and
WAYNE GREY AND GORD SINNAMON
fn(yn) =(cid:18)ZR
gn(tn) =(cid:18)ZR
fn−1(yn−1, yn)pn−1 dyn−1(cid:19)1/pn−1
gn−1(tn−1, tn)qn−1 dtn−1(cid:19)1/qn−1
.
Then Young's inequality gives,
(3.4)
(cid:18)ZR
fn ∗ gn(xn)rn dxn(cid:19)1/rn
≤(cid:18)ZR
fn(tn)pn dtn(cid:19)1/pn(cid:18)ZR
gn(tn)qn dtn(cid:19)1/qn
.
Recursively applying the definitions of fj and gj shows that the right hand side of
(3.4) is just
kf kLP kgkLQ.
The convolution estimates (3.1),(3.2),...,(3.3) concatenate to give a lower bound for
the left hand side of (3.4) and we have
kf ∗ gkLR ≤ kf kLP kgkLQ.
Now we drop the assumption of positivity on f ∈ LP and g ∈ LQ. For bounded,
integrable functions the convolution exists and is finite everywhere. And it is routine
to express f and g as pointwise limits of bounded, integrable functions fk and gk,
respectively, satisfying fk ≤ f and gk ≤ g. Since f ∈ LP and g ∈ LQ
we have shown f ∗ g ∈ LR and hence f ∗ g(x) < ∞ almost everywhere. So
for almost every x the dominated convergence theorem proves that f ∗ g(x) exists.
Also,
kf ∗ gkLR ≤ kf ∗ gkLR ≤ kf kLP kgkLQ = kf kLP kgkLQ.
This completes the proof.
(cid:3)
Once again, the statement of the theorem remains valid when some indices are
infinite, but the proof would have to be modified to accommodate occurrences of
the supremum norm.
The same straightforward procedure may be applied to prove a mixed norm
Young's inequality over any finite product of locally compact unimodular groups.
References
[1] Kenneth F. Andersen, Weighted inequalities for convolutions, Proc. Amer. Math. Soc. 123
(1995), no. 4, 1129–1136, DOI 10.2307/2160710.
[2] J. Appell, A. S. Kalitvin, and P. P. Zabreıko, Partial integral operators in Orlicz spaces with
mixed norm, Colloq. Math. 78 (1998), no. 2, 293–306.
[3] Jurgen Appell and Alois Kufner, On the two-dimensional Hardy operator in Lebesgue spaces
with mixed norms, Analysis 15 (1995), no. 1, 91–98, DOI 10.1524/anly.1995.15.1.91.
[4] A. Benedek and R. Panzone, The space Lp, with mixed norm, Duke Math. J. 28 (1961),
301–324.
[5] John J. F. Fournier, Mixed norms and rearrangements: Sobolev's inequality and Littlewood's
inequality, Ann. Mat. Pura Appl. (4) 148 (1987), 51–76, DOI 10.1007/BF01774283.
[6] Wayne Grey, Inclusions Among Mixed-Norm Lebesgue Spaces, University of Western Ontario,
London, Canada, 2015. Ph.D. Thesis.
[7]
, Mixed-norm estimates and symmetric geometric means, To appear in Positivity.
arXiv:1507.08327 [math.FA] (July 29, 2015).
[8] Wayne Grey and Gord Sinnamon, The inclusion problem for mixed-norm spaces, Trans.
Amer. Math. Soc., posted on January 26, 2016, DOI 10.1090/tran6665, (to appear in print).
PRODUCT OPERATORS ON MIXED NORM SPACES
9
[9] Ralph Howard and Anton R. Schep, Norms of positive operators on Lp-spaces, Proc. Amer.
Math. Soc. 109 (1990), no. 1, 135–146, DOI 10.2307/2048373.
[10] A. S. Kalitvin and P. P. Zabrejko, On the theory of partial integral operators, J. Integral
Equations Appl. 3 (1991), no. 3, 351–382, DOI 10.1216/jiea/1181075630.
[11] L. V. Kantorovich and G. P. Akilov, Functional analysis, 2nd ed., Pergamon Press, Oxford-
Elmsford, N.Y., 1982. Translated from the Russian by Howard L. Silcock.
[12] Gord Sinnamon, Weighted inequalities for positive operators, Math. Inequal. Appl. 8 (2005),
no. 3, 419–440, DOI 10.7153/mia-08-39.
[13] V. D. Stepanov, Two-weight estimates for Riemann-Liouville integrals, Izv. Akad. Nauk
SSSR Ser. Mat. 54 (1990), no. 3, 645–656 (Russian); English transl., Math. USSR-Izv. 36
(1991), no. 3, 669–681.
Department of Mathematics, University of Western Ontario, London, Canada
E-mail address: [email protected]
Department of Mathematics, University of Western Ontario, London, Canada
E-mail address: [email protected]
|
1909.00996 | 1 | 1909 | 2019-09-03T08:22:56 | Which Topologies induced by order convergences | [
"math.FA"
] | In this paper, we will study on some topologies induced by order convergences in a vector lattice. We will investigate the relationships of them. | math.FA | math |
WHICH TOPOLOGIES INDUCED BY ORDER
CONVERGENCES
KAZEM HAGHNEJAD AZAR∗
Abstract. In this paper, we will study on some topologies induced by order
convergences in a vector lattice. We will investigate the relationships of them.
1. Introduction
Recall that a net (xα)α∈A in a Riesz space E is order convergent to x ∈ E,
o−→ x whenever there exists another net (yβ)β∈B in E such that
denoted by xα
yβ ↓ 0 and that for every β ∈ B, there exists α0 ∈ A such that xα − x ≤ yβ for
all α ≥ α0. If there exists a net (yα)α∈A (with the same index set) in a Riesz space
o−→ x. Conversely,
E such that yα ↓ 0 and xα − x ≤ yα for each α ∈ A, then xα
if E is a Dedekind complete Riesz space and (xα)α∈A is order bounded, then
o−→ x in E implies that there exists a net (yα)α∈A (with the same index set)
xα
such that yα ↓ 0 and xα − x ≤ yα for each α ∈ A. For sequences in a Riesz
o−→ x if and only if there exists a sequence (yn) such that yn ↓ 0 and
space E, xn
xn − x ≤ yn for each n ∈ N (cf. [1, P.17 and P.18]).
We adopt [2] as standard reference for basic notions on Riesz spaces and Banach
lattices. Recall that a real vector space E (with elements x,y,...) is called an
ordered vector space if E is partially ordered in such a manner that the vector
space structure and order structure are compatible, that is to say, x ≤ y implies
x + z ≤ y + z for every z ∈ E and x ≥ y implies αx ≥ αy for every α ≥ 0 in R.
A Riesz space E is an order vector space in which sup(x, y) ( it is customary to
write sometimes x ∨ y instead of sup(x, y) and x ∧ y instead of inf(x, y) ) exists
for every x, y ∈ E. Let E be a Riesz space, for each x, y ∈ E with x ≤ y, the set
[x, y] = {z ∈ E : x ≤ z ≤ y} is called an order interval. A subset of E is said to
be order bounded if it is included in some order interval. A Riesz space is said to
be Dedekind complete (resp. σ-Dedekind complete) if every order bounded above
subset (resp. countable subset) has a supremum. A subset A of a Riesz space E is
said to be solid if it follows from y ≤ x whit x ∈ A and y ∈ E that y ∈ A . An
order ideal of E is a solid subspace. A band of E is an order closed order ideal.
A Banach lattice E is a Banach space (E, k.k) such that E is a Riesz space and
its norm satisfies the following property: for each x, y ∈ E such that x ≤ y,
we have kxk ≤ kyk. A Banach lattice E has order continuous norm if kxαk → 0
Copyright 2016 by the Tusi Mathematical Research Group.
Date: Received: xxxxxx; Revised: yyyyyy; Accepted: zzzzzz.
∗Corresponding author.
2010 Mathematics Subject Classification. Primary 46B42; Secondary 47B60.
Key words and phrases. Riesz space, unbounded order convergence, uo-continuous operator,
uσo-continuous operator.
1
2
K. HAGHNEJAD AZAR
for every decreasing net (xα)α with inf α xα = 0. A vector x > 0 in a Riesz space
E is called an atom if Ex = {y ∈ E : ∃λ > 0,
y ≤ λx}, the ideal generated
by x, is one-dimensional if and only if u, v ∈ [0, x] with u ∧ v = 0 implies u = 0
or v = 0. A Riesz space E is said to be atomic if the linear span of all atoms is
order dense in E if and only if it is the band generated by its atoms. For example
c, c0, ℓp(1 ≤ p ≤ ∞) are atomic Banach lattices and C[0, 1], L1[0, 1] are atomless
Banach lattices. Let E, F be Riesz spaces. An operator T : E → F is said to
be order bounded if it maps each order bounded subset of E into order bounded
subset of F . The collection of all order bounded operators from a Riesz space
E into a Riesz space F will be denoted by Lb(E, F ). The collection of all order
bounded linear functionals on a Riesz space E will be denoted by E ∼, that is
E ∼ = Lb(E, R) . A functional on a Riesz space is order continuous (resp. σ-order
continuous) if it maps order null nets (resp. sequences) to order null nets (resp.
sequences). The collection of all order continuous (resp. σ-order continuous)
linear functionals on a Riesz space E will be denoted by E ∼
c ). For
unexplained terminology and facts on Banach lattices and positive operators, we
refer the reader to the excellent book of [2].
n (resp. E ∼
2. Order Topology
Let E be a vector lattice. A subset A of a E is said to be quasi-order closed
whenever for every (xα) ⊆ A with xα ↑ x or xα ↓ x implies x ∈ A. We observe
that a solid subset A ⊆ E is a quasi-order closed if and only if A is order closed.
θ ⊆ E is called order open if and only if E \ θ is quasi-order closed. Now consider
the following topologies:
(1) First topology is called quasi-order topology which we define as follows.
τo = {θ ⊆ E : E \ θ is quasi-order closed}
It is clear, τo is a topology for E.
(2) Assume that τe be a topology for E with following basis
{(a, b) : a, b ∈ E and a < b}.
We call this topology as order topology.
In the following proposition, we show that (E, τo) and (E, τe) are both vector
topologies.
Proposition 2.1. Let E be a Dedekind complete vector lattice. Then τo and τe
both are vector topology.
Proof. Obvious that τe is a vector topology. We only show that τo is vector
topology. First, we prove that the operation x → tx for each t ∈ R is continuous.
Let θ ⊂ E be an order open subset of E, then we must show that tθ is an order
open subset of E for each t ∈ R. Since θ is order open, it follows that θc = F
is quasi-order closed. Put (tθ)c = G and (xα) ⊆ G with xα ↑ x. Then we have
xα /∈ tθ iff t−1xα /∈ θ iff t−1xα ∈ F for each α and since t−1xα ↑ t−1x, follows that
t−1x ∈ F , implies that x ∈ tF . Then we have t−1x ∈ F iff t−1x /∈ θ iff x /∈ θ
iff x ∈ G, which follows that G is quasi order closed, and so tθ is an order open
subset of E.
WHICH TOPOLOGIES INDUCED BY ORDER CONVERGENCES
3
Now we show that the operation (x, y) → x + y is continuous. Set θ1 and θ2
order open subsets of E, we show that θ1 + θ2 is an order open subset of E. Let
a ∈ θ1. First we prove that a + θ2 is an order open subset of E. Put θc
2 = F
and (a + θ2)c = G. We show that G is quasi-order closed. Let (xα) ⊆ G and
xα ↑ x in G. Then we have xα ∈ G iff xα /∈ (a + θ2) iff (xα − a) /∈ θ2. Since
(xα − a) ↑ (x − a), follows that (x − a) ∈ F , and so (x − a) /∈ θ2 iff x /∈ (a + θ2)
iff x ∈ G. Thus G is quasi-order closed, and so a + θ2 is an order open subset of
E. Now by θ1 + θ2 = Sa∈θ1
(a + θ2), the proof follows.
(cid:3)
Lemma 2.2. Let E be a Dedekind complete vector lattice and τo be a order
topology for E. Then for each c ∈ E and neighborhood Uc of c, there are a, b ∈ E
such that c ∈ (a, b) ⊂ Uc.
Proof. let c ∈ E and Uc be an neighbourhood of c in order topology. First we show
that there is a ∈ E such that (a, c) ⊂ Uc. By contradiction, let (a, c) ∩ U c
c 6= ∅.
Then for each a < c there is ca ∈ (a, c) ∩ U c
c . It follows that
sup{ca : ca ∈ (a, c) ∩ U c
c } = c.
For each a < b < c, we can set ca < cb. It follows that for each a < c, there exists
cα(a) ∈ (a, c) ∩ U c
c , which is not possible.
Thus there is a < x such that (a, c) ⊂ Uc. In the similar way there is a c < b
such that (c, b) ⊂ Uc and proof follows.
(cid:3)
c with cα(a) ↑ c. It follows that c ∈ U c
The preceding lemma shows that τo ⊆ τe, but as following example, in general
two topologies not coincide.
Example 2.3. Consider E = ℓ∞ and e1 = (1, 0, 0, 0...). Then (−e1, e1) is member
of τe, but is not belong to τo. Consider xn ∈ ℓ∞ which first n terms are zero and
others are 1. Obviously xn ↓ 0, but xn /∈ (−e1, e1) for each n. This example
shows that the sequence (xn) is order convergent to zero, but is not topological
convergence to zero. On the other hand, since (−e1, e1) /∈ τo, two topologies not
coincide.
Theorem 2.4. Let E be a Dedekind complete vector lattice with topology τe and
(xα) ⊂ E. If xα
τe−→ 0, then (xα) is order convergence to zero.
τe−→ x, there exists
Proof. Assume that a, b ∈ E with x ∈ (a, b) ⊆ E. Since xα
α(a,b) such that xα ∈ (a, b) for each α ≥ α(a,b). Put yα(a,b) = b − a. On the other
hands, (α(a,b))x∈(a,b) is a directed set with the following order relation
α(a,b) ≤ α(c,d) iff (c, d) ⊆ (a, b).
It follows that
Thus xα
o−→ x.
xα − x = (xα ∨ x) − (xα ∧ x) ≤ b − a = yα(a,b) ↓ 0.
(cid:3)
By Example 2.3, the converse of Theorem 2.4 in general not holds.
4
K. HAGHNEJAD AZAR
Proposition 2.5. Let E be a Dedekind complete vector lattice and τo be an order
topology for E. If B is an ideal and quasi-order closed subset of E, then B is a
band in E.
o−→ x, we show that x ∈ B. Obversely sup{xα ∧ x} =
Proof. Let (xα) ⊆ B and xα
x. Set yβ = (Wα6β xα) ∧ x, then yβ ↑ x. Since (yβ) ⊆ B and B is quasi-order
closed, follows that x ∈ B and the result follows.
(cid:3)
References
1. Y. A. Abramovich, C. D. Aliprantis, An Invitation to Operator Theory, Graduate Studies
in Mathematics, vol. 50, American Mathematical Society, Providence, RI, 2002.
2. C.D. Aliprantis, and O. Burkinshaw, Positive Operators, Springer, Berlin, 2006.
3. R. Demarr, Partially ordered linear spaces and locally convex linear topological spaces, Illi-
nois J. Math., 8, 1964, 601-606.
Department of Mathematics, University of Mohaghegh Ardabili, Ardabil, Iran.
E-mail address: [email protected]
|
1010.2342 | 3 | 1010 | 2012-10-25T03:06:37 | Fourier transform and rigidity of certain distributions | [
"math.FA"
] | Let $E$ be a finite dimensional vector space over a local field, and $F$ be its dual. For a closed subset $X$ of $E$, and $Y$ of $F$, consider the space $D^{-\xi}(E;X,Y)$ of tempered distributions on $E$ whose support are contained in $X$ and support of whose Fourier transform are contained in $Y$. We show that $D^{-\xi}(E;X,Y)$ possesses a certain rigidity property, for $X$, $Y$ which are some finite unions of affine subspaces. | math.FA | math |
FOURIER TRANSFORM AND RIGIDITY OF CERTAIN
DISTRIBUTIONS
BINYONG SUN AND CHEN-BO ZHU
Abstract. Let E be a finite dimensional vector space over a local field, and
F be its dual. For a closed subset X of E, and Y of F , consider the space
D−ξ(E; X, Y ) of tempered distributions on E whose support are contained in
X and support of whose Fourier transform are contained in Y . We show that
D−ξ(E; X, Y ) possesses a certain rigidity property, for X, Y which are some
finite unions of affine subspaces.
1. Introduction and main result
One of the most fundamental results in Euclidean harmonic analysis is the un-
certainty principle. As a meta-theorem, it states that a nonzero function and its
Fourier transform cannot both be sharply localized. There are various concrete for-
malisations of this principle, most famously the Heisenberg uncertainty principle.
For a lively and insightful discussion of this topic, see [Tao].
We shall consider distributions ([Sch]). Fix a finite dimensional vector space
E and its dual F . For a closed subset X of E, and Y of F , consider the space
D−ξ(E; X, Y ) of tempered distributions on E whose support are contained in X
and support of whose Fourier transform are contained in Y . The general thrust of
the current note is to examine to what extent one is able to separate the support
and support of the Fourier transform for distributions in D−ξ(E; ∪X , ∪Y), where
X and Y are finite sets of affine subspaces. This is the meaning of rigidity in the
title, which the authors consider as another instance of the uncertainty principle.
One key observation, which surely has been noted by others before us, is the
general importance of relative position of the pair (X, Y ), now assumed to be
affine subspaces. As it turns out, there will be three different circumstances, which
we respectively call thin, perfect, thick (Equation (2)). As an indication of the
relevance of these concepts, we have: (a) D−ξ(E; X, Y ) = 0 if (X, Y ) is a thin pair;
(b) D−ξ(E; X, Y ) can be explicitly described in terms of a countable linear basis
if (X, Y ) is a perfect pair. This is similar to the classical result of L. Schwartz
on the structure of distributions supported on a single point; (c) If (X, Y ) is a
thick pair, then D−ξ(E; X, Y ) contains (in a non-canonical fashion) the space of
2000 Mathematics Subject Classification. 42B35, 46F05 (Primary).
Key words and phrases. Fourier transform, tempered distributions, affine subspaces, rigidity.
1
2
BINYONG SUN AND CHEN-BO ZHU
tempered distributions on a nonzero subspace of E and is thus not rigid in any
reasonable sense.
Our main result (Theorem A in this section) is that D−ξ(E; ∪X , ∪Y) possesses
the afore-mentioned rigidity property as long as there is no pair (X, Y ) ∈ X × Y
which is thick. Very roughly the idea goes as follows: By applying a good multiplier
operator (a suitable function which vanishes on a part of the support of the Fourier
transform), one may cut off that part of the support in the Fourier transform
side. At the distribution side, the process will generally yield a distribution with
additional support. If no thick pairs are involved, then this process can be carried
out in such a way that the additional support is very much controlled. We will
have more to say on the detailed strategy later.
The result of this note was motivated by certain representation-theoretic issues
arising from the proof of archimedean multiplicity-one theorems [SZ]. More specifi-
cally the rigidity statement of Theorem A allows us to establish the semi-simplicity
and non-negativity of an Euler vector field, crucial in certain descent step involving
the so-called distinguished nilpotent orbits. For applications to representation the-
ory of algebraic groups over non-archimedean local fields, we will prove our results
over an arbitrary local field, rather than over R or C. We note that in Euclidean
harmonic analysis and integral geometry, it is of substantial interest to calculate
the Fourier (or Radon) transform of distributions with support in an algebraic set.
We refer the interested reader to the classical book "Generalized functions" by
Gel'fand and Shilov [GS, vol. 1 and 4]. We would also like to point out that for
a real quadratic space E and X = Y the null cone, the space D−ξ(E; X, Y ) is of
special interest for a variety of reasons. Some related topics are explored in [HT,
Chapter 4].
We now introduce some necessary notation for this note.
Let k be an arbitrary local field, and let ψ : k → C× be a fixed nontrivial unitary
character. Let E be a finite dimensional k-vector space. Denote by
C ς (E) ⊂ C−ξ(E), D ς(E) ⊂ D−ξ(E)
the (complex) spaces of Schwartz functions, tempered generalized functions, Schwartz
densities, and tempered distributions on E, respectively. Thus C−ξ(E) (resp.,
D−ξ(E)) is the dual of D ς(E) (resp., C ς(E)).
Let F be a finite-dimensional k-vector space which is dual to E, i.e., a non-
degenerate bilinear map
is given. The Fourier transform
h , i : E × F → k
D ς(F ) → C ς(E)
ω 7→ ω
FOURIER TRANSFORM AND DISTRIBUTIONS
3
is the linear isomorphism given by
Dually for every D ∈ D−ξ(E), its Fourier transform bD ∈ C−ξ(F ) is given by
For every closed subset X of E, and Y of F , denote
ψ(hx, yi) ω(y), x ∈ E.
ω(x) :=ZF
bD(ω) := D(ω), ω ∈ D ς (F ).
D−ξ(E; X) := {D ∈ D−ξ(E) supp(D) ⊂ X},
and
(1)
Let X be an affine subspace of E, and Y an affine subspace of F . Denote
D−ξ(E; X, Y ) := {D ∈ D−ξ(E; X) supp(bD) ⊂ Y }.
L(X) := {u − v u, v ∈ X}
the subspace associated to X, and likewise L(Y ) for Y . We say that the pair
(X, Y ) is thick, perfect or thin according as
(2)
L(X)⊥ ( L(Y ), L(X)⊥ = L(Y ),
or L(X)⊥ * L(Y ),
or what is the same,
L(Y )⊥ ( L(X), L(Y )⊥ = L(X),
or L(Y )⊥ * L(X),
respectively.
Example: Take F to be a non-degenerate quadratic space. Suppose that F0 is a
non-degenerate nonzero subspace of F and
(F0)⊥ = F + ⊕ F −
is a decomposition into totally isotropic subspaces F + and F −. Then the pairs
(F +, F +), (F + ⊕ F0, F +), and (F + ⊕ F0, F + ⊕ F0) are thin, perfect and thick,
respectively.
Remark: If (X, Y ) is a perfect pair, then
D−ξ(E; X, Y ) =(cid:26) C µX,y0,
(C[X] µX,y0) ⊗ D−ξ(L′; {0}),
if k is nonarchimedean,
if k is archimedean.
Here y0 ∈ Y is an arbitrary element, µX,y0 ∈ D−ξ(X) is the product of the func-
tion ψ(h·, −y0i) with a fixed L(X)-invariant positive Borel measure on X. In the
archimedean case, L′ is an arbitrary subspace of E such that L(X) ⊕ L′ = E.
Through addition we have a decomposition E = X × L′. The space C[X] is then
the algebra of (complex valued) polynomial functions on X, viewed as a real affine
space.
4
BINYONG SUN AND CHEN-BO ZHU
We now state the main result of this note.
Theorem A. Let X be a finite set of affine subspaces of E, and Y a finite set of
affine subspaces of F . Assume that there is no pair (X, Y ) ∈ X × Y which is thick.
Then
D−ξ(E; ∪X , ∪Y) =
D−ξ(E; X, Y ).
M
(X,Y )∈X ×Y that is perfect
Remarks: (i) Theorem A asserts in particular that D−ξ(E; ∪X , ∪Y) = 0, if every
(ii) In the archimedean case, D−ξ(E; X, Y ) is a
pair (X, Y ) ∈ X × Y is thin.
module for the Weyl algebra of E (consisting of (complex) polynomial coefficient
differential operators on E). We note that for a perfect pair (X, Y ), the Weyl
algebra module D−ξ(E; X, Y ) is irreducible.
The strategy to prove Theorem A is to control support and it goes as follows.
For every vector u ∈ E, define the following function on F :
(3)
φu := ψ(hu, ·i).
Take a distribution D ∈ D−ξ(E; ∪X , ∪Y). For any Y ∈ Y, pick a nonzero
uY ∈ L(Y )⊥. The function φuY takes a constant value on Y , which we denote by
cY . Thus φuY − cY will vanish on Y , and multiplying a high power of φuY − cY will
cut Y out of the support of the Fourier transform of D. The result is the Fourier
transform of a new distribution which is a linear combination of D and translates
of D by multiples of uY . Doing this consecutively for different Y 's in Y will thus
yield a distribution which is a linear combination of D and translates of D by
elements of the lattice in L(Y )⊥ generated by uY 's, and which has a significantly
reduced support for its Fourier transform. If X ∈ X is thin with respect to some
Y 's, then one could arrange the uY 's (c.f. Lemma 2.1) so that the lattice generated
by uY 's is in a favorable position relative to X, resulting in an excellent control on
the support of the new distribution.
Here are some words on the organization of this note. In Section 2, we show
that if X ∈ X has the property that (X, Y ) is thin for every Y ∈ Y, then X in
fact does not appear in the support of any D ∈ D−ξ(E; ∪X , ∪Y). This is a form
of the uncertainty principle. In Section 3, we show that the rigidity property as
claimed in Theorem A holds in the special case, when X (resp. Y) is a finite set of
translations of a subspace X0 of E (resp. a subspace Y0 of F ). Section 4 is devoted
to the general case. We prove an inductive step in Proposition 4.1. Theorem A
will then follow quickly from the results of Sections 2 and 3, and the induction
result just alluded to.
FOURIER TRANSFORM AND DISTRIBUTIONS
5
Acknowledgements. The authors thank the referee for helpful comments. Biny-
ong Sun is supported by NSFC grants 10801126 and 10931006. Chen-Bo Zhu is
supported by MOE2010-T2-2-113.
2. Thin pairs and elimination of support
Let X be a finite set of affine subspaces of E, and let Y be a finite set of affine
subspaces of F , as in the Introduction. Assume that both X and Y are nonempty.
Note that an integer may also be considered as an element of k. For any family
a = {aY }Y ∈Y ∈ ZY , denote [a] the corresponding element of kY .
Lemma 2.1. Assume that X1 ∈ X and that (X1, Y ) is thin for every Y ∈ Y. Let
x1 ∈ X1 \ ∪(X \ {X1}). Then there is a family {uY ∈ L(Y )⊥}Y ∈Y of vectors in E
with the following property: for every a = {aY }Y ∈Y ∈ ZY with [a] 6= 0, we have
Proof. For every X ∈ X and every a = {aY }Y ∈Y ∈ ZY with [a] 6= 0, put
XY ∈Y
SX,a :=({uY } ∈ YY ∈Y
aY uY /∈ ∪X − x1.
L(Y )⊥ XY ∈Y
aY uY ∈ X − x1) .
If X 6= X1, then 0 /∈ SX,a, and SX,a is a proper affine subspace of QY ∈Y L(Y )⊥.
If X = X1, then SX,a is a subspace of QY ∈Y L(Y )⊥, and is proper due to the
measure zero set ofQY ∈Y L(Y )⊥, and so is the (countable) union ∪X∈X , [a]6=0SX,a.
hypothesis that (X1, Y ) is thin for every Y ∈ Y.
We finish the proof by taking a vector in
In any case each SX,a is a
L(Y )⊥) \ (∪X∈X , [a]6=0SX,a).
(YY ∈Y
(cid:3)
For every vector u ∈ E, denote by Tu : D−ξ(E) → D−ξ(E) the push forward of
the translation by u, and write
(Tuf )(ω) := f (Tuω),
f ∈ C−ξ(E), ω ∈ D ς (E).
Similar notation applies for v ∈ F .
The following is a form of the uncertainty principle.
Proposition 2.2. Assume that X1 ∈ X and that (X1, Y ) is thin for every Y ∈ Y.
Then we have
D−ξ(E; ∪X , ∪Y) = D−ξ(E; ∪(X \ {X1}), ∪Y).
Consequently we have
D−ξ(E; ∪X , ∪Y) = D−ξ(E; ∪X ′, ∪Y ′),
YY ∈Y
(φuY − cY )mbD = 0,
(TuY − cY )mD = 0,
YY ∈Y
Xa={aY }∈{0,1,··· ,m}Y
ca := YY ∈Y(cid:18) m
ua :=XY ∈Y
caTuaD = 0,
aY(cid:19)(−cY )m−aY ,
aY uY .
or what is the same,
(4)
where
and
6
where
BINYONG SUN AND CHEN-BO ZHU
X ′ := {X ∈ X (X, Y ) is not thin for some Y ∈ Y},
Y ′ := {Y ∈ Y (X, Y ) is not thin for some X ∈ X }.
Proof. Let x1 ∈ X1 \ ∪(X \ {X1}) and {uY ∈ L(Y )⊥}Y ∈Y be as in Lemma 2.1. Let
φuY be as in (3), and cY be its common value on Y , as in the Introduction.
Let D ∈ D−ξ(E; ∪X , ∪Y). Then
where m = 1 if k is nonarchimedean, and m is a sufficiently large positive integer
if k is archimedean. The above equality is equivalent to
The choice of {uY } ensures that −ua /∈ ∪X − x1 whenever [a] is nonzero. Let
U be an open neighborhood of 0 in E, small enough so that
(−ua + U) ∩ (∪X − x1) = ∅,
for all a ∈ {0, 1, · · · , m}Y such that [a] 6= 0.
Since D is supported in ∪X , this implies that
(TuaD)x1+U = 0,
for all a ∈ {0, 1, · · · , m}Y such that [a] 6= 0.
Together with (4), this implies that Dx1+U = 0. Since x1 is arbitrary, we conclude
that D is supported in ∪(X \ {X1}).
(cid:3)
3. A special case
Proposition 3.1. If X is a finite set of translations of a subspace X0 of E, and
Y is a finite set of translations of a subspace Y0 of F , then
(5)
D−ξ(E; ∪X , ∪Y) = M(X,Y )∈X ×Y
D−ξ(E; X, Y ).
FOURIER TRANSFORM AND DISTRIBUTIONS
7
Proof. Assume that X ⊥
sides of (5) are 0 by Proposition 2.2, and there is nothing to prove.
0 ⊆ Y0, i.e., (X0, Y0) is not a thin pair. Otherwise both
Let D ∈ D−ξ(E; ∪X , ∪Y). For every X ∈ X , denote by DX ∈ D−ξ(E; X) the
distribution which coincides with D on a neighborhood of X. Then
D = XX∈X
DX.
Note that ∪X is a disjoint union, by our assumption on X .
Assume that k is archimedean. Let P be a real polynomial function whose zero
locus is ∪Y. Then there is a positive integer k such that P kbD = 0, or equivalently,
(eP )kD = 0, where eP is a certain constant coefficient differential operator, acting
on the space D−ξ(E). Therefore for all X ∈ X , (eP )kDX = 0, which implies that
DX ∈ D−ξ(E; X, ∪Y). This proves that
D−ξ(E; X, ∪Y).
(6)
D−ξ(E; ∪X , ∪Y) = MX∈X
Now assume that k is nonarchimedean. Fix X1 ∈ X . For any X ∈ X \ {X1},
choose a vector vX ∈ X ⊥
0 such that
ψ(hX, vXi) 6= ψ(hX1, vXi).
Here ψ(hX, vXi) stands for ψ(hu, vXi), which is independent of u ∈ X, and
ψ(hX1, vXi) is defined similarly. Then we have that
YX∈X \{X1}
(φvX − ψ(hX, vXi)) D = YX∈X \{X1}
(This is not true when k is archimedean.)
(ψ(hX1, vXi − ψ(hX, vXi)) DX1.
The Fourier transform of the left hand side of the above equality is
YX∈X \{X1}
(TvX − ψ(hX, vXi))bD.
Since vX ∈ X ⊥
0 ⊆ Y0, and since ∪Y is invariant under translation by elements of
Y0, the above generalized function is again supported in ∪Y. Therefore the Fourier
transform of DX1 is also supported in ∪Y. This proves (6) in the nonarchimedean
case.
Applying (6) to the pair Y and {X}, we have that
(7)
D−ξ(E; X, ∪Y) =MY ∈Y
D−ξ(E; X, Y ).
We finish the proof by combining (6) and (7).
(cid:3)
8
BINYONG SUN AND CHEN-BO ZHU
4. The general case
As before, let X be a finite set of affine subspaces of E, and let Y be a finite set
of affine subspaces of F . We start with the following
Proposition 4.1. Let X0 be a subspace of E and Y0 a subspace of F . Write
X0 := {X ∈ X L(X) = X0} and Y0 := {Y ∈ Y L(Y ) = Y0}.
Assume that (X ′, Y0) and (X0, Y ′) are thin for all
X ′ ∈ X ′ := X \ X0 and all Y ′ ∈ Y ′ := Y \ Y0.
Then
(8)
D−ξ(E; ∪X , ∪Y) = D−ξ(E; ∪X0, ∪Y0) ⊕ D−ξ(E; ∪X ′, ∪Y ′).
Proof. Proposition 2.2 implies that the right hand side of (8) is a direct sum.
Let {uY ′ ∈ L(Y ′)⊥}Y ′∈Y ′ be a family of vectors in E such that for every family
a = {aY ′}Y ′∈Y ′ ∈ ZY ′ with [a] 6= 0, we have
XY ′∈Y ′
aY ′uY ′ /∈ [X1,X2∈X0
(X1 − X2).
The existence of such a family is proved along the same line as that of Lemma 2.1.
Let φuY ′ be as in (3), and cY ′ be its common value on Y ′, as in the Introduction.
Let D ∈ D−ξ(E; ∪X , ∪Y). We take m to be a sufficiently larger positive integer
if k is archimedean and m = 1 if k is non-archimedean. Then the generalized
function
is supported in ∪Y0. It is the Fourier transform of the distribution
YY ′∈Y ′
D′ := YY ′∈Y ′
(φuY ′ − cY ′)m!bD
(TuY ′ − cY ′)m! D.
D′ =
Xa={aY ′ }∈{0,1,··· ,m}Y′
caTuaD,
ca := YY ′∈Y ′(cid:18) m
ua := XY ′∈Y ′
aY ′(cid:19)(−cY ′)m−aY ′ ,
aY ′uY ′.
By expansion, we have
(9)
where
and
FOURIER TRANSFORM AND DISTRIBUTIONS
9
For every set Z of affine subspaces of E, we put
eZ :=nua + Z a ∈ {0, 1, · · · , m}Y ′
, Z ∈ Zo .
since (X ′, Y0) is thin for all X ′ ∈ X ′, Proposition 2.2 implies that it is supported
Then D′ is clearly supported in ∪eX . Its Fourier transform is supported in ∪Y0, and
in ∪eX0.
Now the choice of {uY ′} ensures that
(a disjoint union),
fX0 = X0 ⊔ X1
where
X1 :=nua + X a ∈ {0, 1, · · · , m}Y ′
, [a] 6= 0, X ∈ X0o .
By Proposition 3.1, we may write
(10)
where D0 ∈ D−ξ(E; ∪X0, ∪Y0) and D1 ∈ D−ξ(E; ∪X1, ∪Y0).
D′ = D0 + D1,
The disjointness of X0 and X1 allows us to choose an open neighborhood U of 0 in
E, small enough so that
Let
and
(11)
x0 ∈ ∪X0 \ ∪fX ′.
(x0 + U) ∩ (∪fX ′) = ∅,
(x0 + U) ∩ (∪X1) = ∅.
For all nonzero [a], we thus have
and therefore
(x0 + U − ua) ∩ (∪X ) = ∅
(TuaD)x0+U = 0.
Then (9) implies that
D′x0+U = c0Dx0+U , with c0 = YY ′∈Y ′
(−cY ′)m,
and (10) and (11) implies that
(D′ − D0)x0+U = D1x0+U = 0.
We thus conclude from the last two equalities that D − D0/c0 vanishes on x0 + U.
Since x0 is arbitrary, we see that D − D0/c0 is supported in
where
(∪X ) \ (∪X0 \ ∪fX ′) ⊂ (∪X ′) ∪ (∪(X0 ∧fX ′)).
X0 ∧fX ′ = {X0 ∩ X ′ X0 ∈ X0, X ′ ∈fX ′}.
10
BINYONG SUN AND CHEN-BO ZHU
see that D − D0/c0 actually belongs to
Since every pair in (X0 ∧fX ′) × Y is thin, and every pair in X ′ × Y0 is thin, we
D−ξ(E; (∪X ′) ∪ (∪(X0 ∧fX ′)), ∪Y) = D−ξ(E; ∪X ′, ∪Y) = D−ξ(E; ∪X ′, ∪Y ′),
by two applications of Proposition 2.2. This finishes the proof of the current
proposition.
(cid:3)
We are now ready to prove Theorem A. Assume that no pair in X × Y is thick.
Put
For every L ∈ L, put
L := {L(X) (X, Y ) ∈ X × Y is perfect}.
XL := {X ∈ X L(X) = L} and YL⊥ := {Y ∈ Y L(Y ) = L⊥}.
Then we have
D−ξ(E; ∪X , ∪Y)
(by Proposition 2.2)
(by Proposition 4.1)
= D−ξ(E; ∪(∪L∈LXL), ∪(∪L∈LYL⊥))
D−ξ(E; ∪XL, ∪YL⊥)
= ML∈L
ML∈L, (X,Y )∈XL×YL⊥
=
D−ξ(E; X, Y ))
(by Proposition 3.1)
D−ξ(E; X, Y ).
=
M
(X,Y )∈X ×Y that is perfect
This finishes the proof of Theorem A.
References
[GS] I.M. Gel'fand and G.E. Shilov, Generalized functions, Vol. 1–5, Academic Press, 1966–1968.
[HT] R. Howe and E.C. Tan, Non-abelian harmonic analysis: applications of SL(2, R), Univer-
sitext, Springer Verlag, 1992.
[Sch] L. Schwartz, Theorie des distributions, Hermann, 1966.
[SZ] B. Sun and C.-B. Zhu, Multiplicity one theorems: the Archimedean case, Annals Math.
175, (2012), no. 1, 23–44.
[Tao] T. Tao, The uncertainty principle, http://terrytao.wordpress.com/2010/06/25/the-uncertainty-principle/.
Academy of Mathematics and Systems Science, Chinese Academy of Sciences,
Beijing, 100190, P.R. China
E-mail address: [email protected]
Department of Mathematics, National University of Singapore, Block S17, 10
Lower Kent Ridge Road, Singapore 119076
E-mail address: [email protected]
|
1804.08518 | 1 | 1804 | 2018-04-23T15:44:47 | The Bezout equation on the right half plane in a Wiener space setting | [
"math.FA"
] | This paper deals with the Bezout equation $G(s)X(s)=I_m$, $\Re s \geq 0$, in the Wiener space of analytic matrix-valued functions on the right half plane. In particular, $G$ is an $m\times p$ matrix-valued analytic Wiener function, where $p\geq m$, and the solution $X$ is required to be an analytic Wiener function of size $p\times m$. The set of all solutions is described explicitly in terms of a $p\times p$ matrix-valued analytic Wiener function $Y$, which has an inverse in the analytic Wiener space, and an associated inner function $\Theta$ defined by $Y$ and the value of $G$ at infinity. Among the solutions, one is identified that minimizes the $H^2$-norm. A Wiener space version of Tolokonnikov's lemma plays an important role in the proofs. The results presented are natural analogs of those obtained for the discrete case in [11]. | math.FA | math |
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A
WIENER SPACE SETTING
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
Abstract. This paper deals with the B´ezout equation G(s)X(s) = Im, ℜs ≥
0, in the Wiener space of analytic matrix-valued functions on the right half
plane. In particular, G is an m × p matrix-valued analytic Wiener function,
where p ≥ m, and the solution X is required to be an analytic Wiener function
of size p × m. The set of all solutions is described explicitly in terms of a p × p
matrix-valued analytic Wiener function Y , which has an inverse in the analytic
Wiener space, and an associated inner function Θ defined by Y and the value
of G at infinity. Among the solutions, one is identified that minimizes the H 2-
norm. A Wiener space version of Tolokonnikov's lemma plays an important
role in the proofs. The results presented are natural analogs of those obtained
for the discrete case in [11].
1. Introduction and main results
In this paper we deal with the B´ezout equation G(s)X(s) = Im on the closed
right half plane ℜs ≥ 0, assuming that the given function G is of the form
G(s) = D +Z ∞
0
e−stg(t) dt
(ℜs ≥ 0),
g ∈ L1
where
m×p(R+) ∩ L2
(1.1)
In particular, G belongs to the analytic Wiener space W m×p
solutions X ∈ W p×m
(1.2)
X(s) = DX +Z ∞
e−stx(t) dt
, that is,
+
+
0
(ℜs ≥ 0), where x ∈ L1
m×p(R+).
m×p(R+).
. We are interested in
Throughout p ≥ m. We refer to the final paragraph of this introduction for a
further explanation of the notation.
With G given by (1.1) we associate the Wiener-Hopf operator TG mapping
p(R+) into L2
m(R+) which is defined by
L2
(1.3)
(TGh)(t) = Dh(t) +Z ∞
0
g(t − τ )h(τ )dτ,
t ≥ 0 (h ∈ L2
m(R+)).
2010 Mathematics Subject Classification. Primary 47A56; Secondary 47A57, 47B35, 46E40,
46E15.
Key words and phrases. B´ezout equation, corona problem, Wiener space on the line,
matrix-valued functions, minimal norm solutions, Tolokonnikov's lemma.
The third author gratefully thanks the mathematics department of North-West University,
Potchefstroom campus, South Africa, for the hospitality and support during his visit from Sep-
tember 21 -- October 15, 2015.
This work is based on the research supported in part by the National Research Foundation of
South Africa (Grant Number 93406).
1
2
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
For X as in (1.2) we define the Wiener-Hopf operator TX mapping L2
L2
p(R+) in a similar way, replacing D by DX and g by x. If the B´ezout equation
m(R+) into
(1.4)
G(s)X(s) = Im, ℜs ≥ 0.
has a solution X as in (1.2), then (using the analyticity of G and X) the theory
of Wiener-Hopf operators (see [7, Section XII.2] or [2, Section 9]) tells us that
TGTX = TGX = I, where I stands for the identity operator on L2
m(R+). Thus for
the B´ezout equation (1.4) to be solvable the operator TG must be surjective or,
equivalently, TGT ∗
G must be strictly positive. We shall see that this condition is
also sufficient.
To state our main results, we assume that TGT ∗
G is strictly positive. Then D =
G(∞) is surjective, and hence DD∗ is strictly positive too. We introduce two
matrices D+ and E, of sizes p× m and p× (p− m), respectively, and a p × p matrix
function Y in W p×p
+ , as follows:
(i) D+ = D∗(DD∗)−1, where D = G(∞);
(ii) E is an isometry mapping Cp−m into Cp such that Im E = Ker D;
(iii) Y is the p × p matrix function given by
Y (s) = Ip −Z ∞
0
e−sty(t)dt, ℜs ≥ 0, where y = T ∗
G(TGT ∗
G)−1g.
From the definitions of D+ and E it follows that the p × p matrix (cid:2)D+ E(cid:3) is
non-singular. In fact
(1.5)
(1.6)
(1.8)
(cid:20) D
E ∗(cid:21)(cid:2)D+ E(cid:3) =(cid:20)Im
0
0
Ip−m(cid:21) .
As we shall see (Proposition 2.1 in Section 2 below), the fact that the given function
g ∈ L1
m×p(R+) ∩ L2
In
particular, Y ∈ W p×p
(1.7)
m×p(R+) implies that a similar result holds true for y.
+ . In what follows Ξ and Θ are the functions defined by
0
Ξ(s) =(cid:16)Ip −Z ∞
Θ(s) =(cid:16)Ip −Z ∞
+ , we have Ξ ∈ W p×m
+
0
e−sty(t)dt(cid:17)D+ = Y (s)D+, ℜs > 0;
e−sty(t)dt(cid:17)E = Y (s)E, , ℜs > 0.
+
+
and Θ ∈ W p×(p−m)
. Finally, recall that a
is inner whenever Ω(s) is an isometry
Since Y ∈ W p×p
function Ω in the analytic Wiener space W k×r
for each s ∈ iR. We now state our main results.
Theorem 1.1. Let G be the m × p matrix-valued function given by (1.1). Then
the equation G(s)X(s) = Im, ℜs > 0, has a solution X ∈ W p×m
if and only if
TG is right invertible. In that case the function Ξ defined by (1.7) is a particular
solution and the set of all solutions X ∈ W p×m
(1.9)
is given by
X(s) = Ξ(s) + Θ(s)Z(s), ℜs > 0,
+
+
where Ξ and Θ are defined by (1.7) and (1.8), respectively, and the free parameter
. Moreover, the function Θ belongs to
and is inner. Furthermore, the solution Ξ is the minimal H 2 solution
Z is an arbitrary function in W (p−m)×m
W p×(p−m)
+
+
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING3
in the following sense
kX(·)uk2
H 2
p
(1.10)
= kΞ(·)uk2
H 2
p
+ kZ(·)uk2
H 2
p−m
,
where u ∈ Cm and Z ∈ fW (p−m)×m
+
.
In the above theorem, for any positive integer k, H 2
k = H 2
space of Ck-valued functions on the right half plane given by H 2
where J is the unitary operator defined by
1
√2πF : L2
k(R) → L2
k(iR)
(1.11)
J =
k (iR) is the Hardy
k (iR) = J L2
k(R+),
k(R) onto L2
k(iR). Moreover, Z ∈
+
means that
with F being the Fourier transform mapping L2
fW (p−m)×m
Z(s) = DZ +Z ∞
(ℜs ≥ 0), where
DZ is a (p − m) × p matrix and z ∈ L1
e−stz(t) dt
0
(p−m)×p(R+).
See the final part of this introduction for further information about the used nota-
tion, in particular, see (1.14) for the definition of the Fourier transform F .
setting. The result emphasizes the central role of the function Y .
The second theorem is a variant of the Tolokonnikov lemma [20] in the present
(p−m)×p(R+) ∩ L2
G is strictly positive, and let Y be the matrix function
+ , det Y (s) 6= 0 whenever
+ . Furthermore, the p× p matrix function
Theorem 1.2. Assume TGT ∗
defined by (1.5). Then Y belongs to the Wiener space W p×p
ℜs ≥ 0, and hence Y is invertible in W p×p
(1.12)
(cid:20) G(s)
E ∗Y (s)−1(cid:21) , ℜs ≥ 0,
is invertible in the Wiener algebra W p×p
(1.13)
(cid:20) G(s)
E ∗Y (s)−1(cid:21)−1
+
and its inverse is given by
= Y (s)(cid:2)D+ E(cid:3) =(cid:2)Ξ(s) Θ(s)(cid:3) , ℜs ≥ 0.
The literature on the B´ezout equation and the related corona problem is exten-
sive, starting with Carleson's corona theorem [3] (for the case when m = 1) and
Fuhrmann's extension to the matrix-valued case [6], both in a H ∞ setting. The
topic has beautiful connections with operator theory (see the books [14], [16], [17],
[18], and the more recent papers [21], [22], [23]). Rational matrix equations of the
form (1.4) play an important role in solving systems and control theory problems,
in particularly, in problems involving coprime factorization, see, e.g., [24, Section
4.1], [10, Section A.2], [25, Chapter 21]. For more recent work see [12] and [13],
and [15, page 3] where it is proved that the scalar analytic Wiener algebra is a
pre-B´ezout ring. For matrix polynomials, the equation (1.4) is closely related to
the Sylvester resultant; see, e.g., Section 3 in [9] and the references in that paper.
The present paper is inspired by [5] and [11]. The paper [5] deals with equation
(1.4) assuming the matrix function G to be a stable rational matrix function, and
the solutions are required to be stable rational matrix functions as well. The com-
ment in the final paragraph of [5, Section 2] was the starting point for our analysis.
The paper [11] deals with the discrete case (when the right half plane is replaced by
the open unit disc). Theorems 1.1 and 1.2 are the continuous analogue of Theorem
4
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
1.1 in [11]. The absence of an explicit formula for the function Y −1 in the present
setting makes the proofs more complicated than those in [11].
The paper consists of five sections, including the present introduction and an
appendix. Section 2, which deals with the right invertibility of the operator TG,
has an auxiliary character. Theorem 1.2 is proved in Section 3, and Theorem 1.1
in Section 4. The Appendix, Section A, contains a number of auxiliary results
involving the Lebesgue space L1(R) ∩ L2(R) and its vector-valued counterpart,
which are collected together simply for the convenience of the reader and contains
no significantly new material.
Notation and terminology. We conclude this section with some notation and termi-
nology. Throughout, a linear map A : Cr → Ck is identified with the k×r matrix of
A relative to the standard orthonormal bases in Cr and Ck. The space of all k × r
k×r(R). As usual bf denotes the
matrices with entries in L1(R) will be denoted by L1
Fourier transform of f ∈ L1
(1.14)
k×r(R), that is,
e−stf (t)dt,
s ∈ iR.
bf (s) = (F f )(s) =Z ∞
−∞
Note that bf is continuous on the extended imaginary axis iR∪ {± i∞}, and is zero
at ± i∞ by the Riemann-Lebesgue lemma. By W k×r we denote the Wiener space
consisting of all k × r matrix functions F on the imaginary axis of the form
(1.15)
F (s) = DF + bf (s), s ∈ iR, where f ∈ L1
k×r(R) and
DF is a constant matrix.
Since bf is continuous on the extended imaginary axis and is zero at ±i∞, the
function F given by (1.15) is also continuous on the extended imaginary axis and
the constant matrix DF is equal to the value of F at infinity. We write W k×r
for the
space of all F of the form (1.15) with the additional property that f has its support
in R+ = [0,∞), that is, f is equal to zero on (−∞, 0). Any function F ∈ W k×r
is
analytic and bounded on the open right half plane. Thus any F ∈ W k×r
+ is a matrix-
valued H ∞ function. Finally, by W k×r
−,0 we denote the Wiener space consisting of
all F of the form (1.15) with the additional property that DF = 0 and f has its
support in (−∞, 0]. Thus we have the following direct sum decomposition:
(1.16)
+
+
W k×r = W k×r
+
+W k×r
−,0 .
+
if the function f in (1.15) belongs to L1
We write F ∈ fW k×r
Similarly, F ∈ fW k×r
Let F ∈ W k×r be given by (1.15). With F we associate the Wiener-Hopf op-
k(R+). This operator (see [7, Section XII.2]) is
k×r(R−) and DF = 0.
−,0 if f ∈ L1
k×r(R+) ∩ L2
k×r(R−) ∩ L2
r(R+) into L2
k×r(R+).
erator TF mapping L2
defined by
(1.17)
(TF h)(t) = DF h(t) +Z ∞
f (t − τ )h(τ )dτ,
t ≥ 0 (h ∈ L2
r(R+)).
The orthogonal complement of H 2
will be denoted by K 2
belongs to H 2
k(iR). If F ∈ fW k×r
k (iR). Similarly, F (·)u belongs to K 2
k(R+), with J as in (1.11), in L2
k(iR)
+ , then for each u ∈ Cr the function F (·)u
k(iR) if F ∈ fW k×r
−,0 .
0
k (iR) = J L2
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING5
Finally, for f ∈ L1
function in L1
k×m(R), see [19, Section 7.13], given by
r×m(R) the convolution product f ⋆ g is the
k×r(R) and g ∈ L1
(f ⋆ g)(t) =Z ∞
−∞
(1.18)
f (t − τ )g(τ ) dτ
a.e. on R.
2. Right invertibility of TG
In this section G ∈ W m×p
know that the B´ezout equation (1.4) having a solution X in W p×m
TG is right invertible or, equivalently, TGT ∗
containing formula (1.4).
, where G is given by (1.1) and p ≥ m. We already
implies that
G is strictly positive; see the paragraph
+
+
In this section we present an auxiliary result that will be used to prove our main
theorems. For this purpose we need the m × m matrix-valued function R on the
imaginary axis defined by R(s) = G(s)G(s)∗, s ∈ iR. It follows that R ∈ W m×m.
By TR we denote the corresponding Wiener-Hopf operator acting on L2
m(R+). Thus
(TRf )(t) = DD∗f (t) +Z ∞
r(t − τ )f (τ ) dτ,
with r(t) = Dg∗(t) + g(t)D∗ +Z ∞
0
−∞
0 ≤ t < ∞,
g(t − τ )g∗(τ ) dτ,
t ∈ R.
Here g∗(t) = g(−t)∗ for t ∈ R. It is well-known (see, e.g., formula (24) in Section
XII.2 of [7]) that
(2.1)
Here HG is the Hankel operator mapping L2
is,
TR = TGT ∗
G + HGH ∗
G.
(2.2)
(HGf )(t) =Z ∞
0
g(t + τ )f (τ )dτ,
f ∈ L2
p(R+).
p(R+) into L2
m(R+) defined by G, that
We shall prove the following proposition. For the case when G is a rational
matrix function, the first part (of the "if and only if" part) of the proposition is
covered by Lemma 2.3 in [5]. The proof given in [5] can also be used in the present
setting. For the sake of completeness we include a proof of the first part.
Proposition 2.1. Let G be given by (1.1). Then the operator TG is right invertible
if and only if TR and I − H ∗
R HG are both invertible operators. In that case the
inverse of TGT ∗
G is given by
(TGT ∗
G)−1 = T −1
R HG)−1H ∗
R + T −1
GT −1
R .
GT −1
GT −1
(2.3)
R HG(I − H ∗
Furthermore,
m(R+) ∩ L2
(a) (TGT ∗
(b) the function y defined by y = T ∗
G)−1 maps L1
m(R+) in a one-to-one way onto itself;
p×p(R+)∩L2
G(TGT ∗
G)−1g belongs to L1
+ .
p×p(R+),
in particular, the function Y given by (1.5) is in fW p×p
Proof. We split the proof into four parts. In the first part we assume that TG
is right invertible, and we show that TR and I − H ∗
R HG are both invertible
operators and that the inverse of TGT ∗
G is given by (2.3). The second part deals
with the reverse implication. Items (a) and (b) are proved in the last two parts.
Part 1. Assume TG is right invertible. Then the operator TGT ∗
According to (2.1) we have TR ≥ TGT ∗
G is strictly positive.
G, and hence TR is also strictly positive.
GT −1
6
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
In particular, TR is invertible. Rewriting (2.1) as TGT ∗
multiplying the latter identity from the left and from the right by T
that
G = TR − HGH ∗
−1/2
R
G, and
shows
(2.4)
T
−1/2
R
GT
−1/2
R
−1/2
R HGH ∗
GT
−1/2
R
.
TGT ∗
−1/2
R
= I − T
R HG = (cid:0)H ∗
−1/2
R HGH ∗
GT
is strictly positive which shows that H ∗
GT −1
is a
Hence I − T
strict contraction. But then H ∗
is also a strict
contraction, and thus the operator I − H ∗
R HG is strictly positive. In particular,
I − H ∗
G, a usual Schur
complement type of argument (see, e.g., Section 2.2 in [1]), including the well-
known inversion formula
(cid:1)∗
G = TR − HGH ∗
R HG is invertible. Finally, since TGT ∗
(cid:1)(cid:0)H ∗
GT −1
GT −1
−1/2
R
−1/2
R
GT
GT
GT
−1/2
R
(A − BC)−1 = A−1 + A−1B(I − CA−1B)−1CA−1,
GT
GT −1
G)−1 is given by (2.3).
G = TR − HGH ∗
then shows that (TGT ∗
Part 2. In this part we assume that TR and I − H ∗
R HG are both invertible
operators, and we show that TG is right invertible. According to (2.1) the operator
TR is positive. Since we assume TR to be invertible, we conclude that TR is strictly
positive. Rewriting (2.1) as TGT ∗
G, and multiplying the latter
we obtain the identity (2.4).
identity from the left and from the right by T
−1/2
R HGH ∗
is a contraction.
Hence I−T
GT −1
But then H ∗
is also a contraction, and thus the
GT −1
operator I − H ∗
R HG is invertible.
−1/2
It follows that I − H ∗
R HG is a strict
−1/2
. This implies
contraction. But then the same holds true for T
R
is strictly positive, and (2.4) shows that TG is right
that I − T
invertible.
Part 3. In this part we prove item (a). Observe that g ∈ L1
TG maps L1
R HG is strictly positive, and hence T
GT
is positive which shows that H ∗
−1/2
R
R HG is positive. By assumption I − H ∗
R HG =(cid:0)H ∗
m×p(R) implies that
−1/2
R HGH ∗
−1/2
R HGH ∗
(cid:1)(cid:0)H ∗
p(R+) into L1
p×m(R) and
GT −1
GT −1
−1/2
R
−1/2
R
−1/2
R
−1/2
R
−1/2
R
(cid:1)∗
GT
m(R+) ∩ L2
m(R+). Since g∗ ∈ L1
GT
GT
GT
p(R+) ∩ L2
(T ∗
Gf )(t) = D∗f (t) +Z ∞
m(R+)∩ L2
g∗(t − τ )f (τ ) dτ,
0
m(R+) into L1
0 ≤ t < ∞,
p(R+). Thus TGT ∗
G maps
m(R+) into itself. We have to show that the same holds true for its
p(R+)∩ L2
G maps L1
the operator T ∗
L1
m(R+) ∩ L2
inverse. To do this we apply Lemmas A.3 and A.4.
Lemma A.3 tells us that T −1
m(R+)∩ L2
itself. This allows us to apply Lemma A.4 with
R maps L1
m(R+) in a one-to-one way onto
Q = T −1
R , H = HG and
H = H ∗
G.
Recall that HG is a Hankel operator, see (2.2), and H ∗
in fact
G is also a Hankel operator,
(H ∗
Gf )(t) =Z ∞
0
g(t + τ )∗f (τ ) dτ,
0 ≤ t < ∞.
Since I − H ∗
L1
p(R+)∩ L2
its inverse (I − H ∗
m(R+)∩L2
maps L1
m(R+) ∩ L2
into L1
R HG is invertible, Lemma A.4 then shows that I − H ∗
GT −1
R HG maps
p(R+) in a one-to-one way onto itself, and hence the same holds true for
GT −1
R
p(R+)
m(R+) ∩
R HG)−1. To complete the proof of item (a) note that H ∗
p(R+)∩L2
GT −1
m(R+) into L1
m(R+). But then (2.3) shows that (TGT ∗
p(R+), and T −1
R HG maps L1
G)−1 maps L1
p(R+)∩L2
GT −1
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING7
m(R+) ∩ L2
m(R+) ∩ L2
G)−1 is one-to-one on L1
m(R+) into itself. To see that (TGT ∗
m(R+) and
m(R+) onto itself, one can follow the same argumentation as in
L2
maps L1
the last part of the proof of Lemma A.3.
Part 4. In this part we prove item (b). Since g belongs to L1
G)−1g also belongs to L1
item (a) tells us that f := (TGT ∗
We already have seen (in the first paragraph of the previous part) that T ∗
L1
m(R+) ∩ L2
L1
p×p(R+) ∩ L2
m×p(R+),
m×p(R+).
G maps
Gf belongs to
(cid:3)
m(R+) into L1
p×p(R+), as desired.
m×p(R+)∩ L2
m×p(R+) ∩ L2
It follows that y = T ∗
p(R+) ∩ L2
p(R+).
3. The functions Y and Θ, and proof of Theorem 1.2
We begin with three lemmas involving the functions Y and Θ defined by (1.5)
and (1.8), respectively. From Proposition 2.1, item (b), and (1.8) we know that
Y ∈ W p×p
Lemma 3.1. Assume that TG is right invertible, and let Y ∈ W p×p
defined by (1.5). Then
; see also the paragraph preceding Theorem 1.1.
and Θ ∈ W p×(p−m)
be the function
+
+
+
(3.1)
G(s)Y (s) = D,
ℜs > 0.
G)−1g = g. Since the functions
Proof. To prove (3.1) note that TGy = TGT ∗
g and y both have their support in R+, the identity TGy = g can be rewritten as
Dy + g ⋆ y = g, where ⋆ is the convolution product of matrix-valued functions with
entries in L1(R); see (1.18). Thus
G(TGT ∗
(3.2)
Dy(t) + (g ⋆ y)(t) = Dy(t) +Z ∞
−∞
g(t − τ )y(τ )dt = g(t),
t ∈ R.
Next use that the Fourier transform of a convolution product is just the product
of the Fourier transforms of the functions in the convolution product. Thus taking
Fourier transforms in (3.2) yields Dby +bgby =bg. The latter identity can be rewritten
as Gby =bg. Hence, using the definition of Y in (1.5), we obtain
G(s)Y (s) = G(s)(cid:16)Ip −by(s)(cid:17) = G(s) −bg(s) = D.
This proves (3.1).
(cid:3)
+
Lemma 3.2. Assume that TG is right invertible. Then the function Θ defined by
(1.8) belongs to W p×(p−m)
and is an inner function, that is, Θ(s) is an isometry
for each s ∈ iR and at infinity.
Proof. We already know that Θ ∈ W p×(p−m)
G)−1g as in (1.5), and put f = (TGT ∗
y = T ∗
Proposition 2.1 (b), and y = T ∗
. To prove that Θ is inner, let
G)−1g. Thus f ∈ L1
m×m(R+), by
Gf . The latter can be rewritten as
G(TGT ∗
+
g∗(t − τ )f (τ ) dτ,
t ≥ 0.
y(t) = D∗f (t) +Z ∞
y(t) = D∗f (t) +Z ∞
−∞
0
Note that g∗(t) = g(−t)∗, and hence g∗ has its support in (−∞, 0]. Therefore
(3.3)
g∗(t − τ )f (τ ) dτ,
t ∈ R.
Put
(3.4)
ρ(t) =(cid:26)
0
when t ≥ 0,
(g∗ ⋆ f )(t) when t < 0.
8
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
Using the definition of the convolution product ⋆, see (1.18), we can rewrite (3.3)
as
Taking Fourier transforms we obtain
y(t) = D∗f (t) + (g∗ ⋆ f )(t) − ρ(t)
t ∈ R.
by(s) = D∗bf (s) + bg∗(s)bf (s) −bρ(s) = G(s)∗bf (s) −bρ(s),
s ∈ iR.
Hence, we have
s ∈ iR.
E ∗Y (s)∗Y (s)E =
Y (s) = I − G(s)∗bf (s) +bρ(s),
(3.5)
Now let us compute Θ(s)∗Θ(s) = E ∗Y (s)∗Y (s)E for s ∈ iR. We have
= E ∗Y (s)∗E − E ∗Y (s)∗G(s)∗bf (s)E + E ∗Y (s)∗bρ(s)E
= E ∗Y (s)∗E + E ∗Y (s)∗bρ(s)E
= E ∗E − E ∗by(s)∗E + E ∗Y (s)∗bρ(s)E
= Ip−m − Ω(s).
−,0 , and thus Ω belongs to W (p−m)×(p−m)
Here Ω(s) = E ∗by(s)∗E−E ∗Y (s)∗bρ(s)E. Note that the functionsby(·)∗ and Y (·)∗bρ(·)
belong to W p×p
. On the other hand, the
function E ∗Y (·)∗Y (·)E is hermitian on the imaginary axis, and hence the same is
true for Ω. But for any positive integer k we have
(because G(s)Y (s)E = 0 by (3.1) and DE = 0)
−,0
W k×k
−,0 ∩ (W k×k
−,0 )∗ = {0}.
+
be the function
Thus Ω is identically zero, and thus Θ(s)∗Θ(s) = E ∗Y (s)∗Y (s)E = Ip−m for any
s ∈ iR. Moreover, Θ(∞)∗Θ(∞) = E ∗E = I. This proves that Θ is inner.
(cid:3)
Lemma 3.3. Assume that TG is right invertible, and let Y ∈ W p×p
defined by (1.5). Then Y is invertible in W p×p
+ .
Proof. Fix s ∈ iR, and assume u ∈ Cp such that Y (s)u = 0. Then G(s)Y (s) = D
implies that Du = 0. By definition of E, u = Ev for some v ∈ Cp−m. Next use
Θ(s) = Y (s)E. It follows that Θ(s)v = Y (s)Ev = Y (s)u = 0. However, Θ(s) is an
isometry, by Lemma 3.2. So v = 0, and hence u = 0. We see that det Y (s) 6= 0.
Also Y (∞) = Ip. We conclude that TY is a Fredholm operator; see [7, Theorem
XII.3.1].
Next we prove that Ker TY = {0}. Take h ∈ Ker TY . Then TY h = 0, and hence
Y (s)bh(s) = 0 for each s ∈ iR. But det Y (s) 6= 0 for each s ∈ iR. Hence bh = 0, and
We want to prove that TY is invertible. Given the results of the preceding first
two paragraphs it suffices to show that ind TY = 0. This will be done in the next
step by an approximation argument, using the fact, from [5], that we know the
result is true for rational matrix functions.
therefore h = 0.
Let g be as in (1.1). Note that g is the limit in L1 of a sequence g1, g2, . . . such
of [7]. Since TG is right invertible, TGn will also be right invertible for n sufficiently
)−1gn.
large. In fact, TGnT ∗
that Gn(s) = D+cgn(s) is a stable rational matrix function; cf., Part (v) on page 229
Then yn → y in the L1-norm. Put Yn(s) = I −cyn(s). Then TYn → TY in operator
G in operator norm. Put yn = T ∗
Gn
Gn → TGT ∗
(TGnT ∗
Gn
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING9
norm. For n sufficiently large the operator TYn is invertible (see the paragraph
preceding Theorem 1.2 in [5] and formula (2.17) in [5]). In particular, the Fredholm
index of TYn is zero. But ind TY = limn→∞ ind TYn = 0. Thus TY is invertible, and
hence Y is invertible in W p×p
+ .
(cid:3)
Proof of Theorem 1.2. From Lemma 3.3 we know that Y ∈ W p×p
and that Y
is invertible in W p×p
+ . Thus we only have to prove the second part of the theorem.
Since Y is invertible in W p×p
+ , the p × p matrix function given by (1.12) is well-
defined and belongs to W p×p
+ . Furthermore, from (1.6) we know that the p × p
matrix (cid:2)D+ E(cid:3) is invertible. Hence the function defined by the right hand side
of (1.13) belongs to W p×p
and is invertible in W p×p
(1.6) we see that
(cid:20) G(s)
E ∗Y (s)−1(cid:21) Y (s)(cid:2)D+ E(cid:3) =(cid:20)G(s)Y (s)
+ . Using (3.1) and the identity
E ∗
+
+
(cid:21)(cid:2)D+ E(cid:3)
Ip−m(cid:21) , ℜs ≥ 0.
0
=(cid:20) D
E ∗(cid:21)(cid:2)D+ E(cid:3) =(cid:20)Im
0
This proves the first identity (1.13). The second identity is an immediate conse-
quence of the definitions of Ξ and Θ in (1.7) and (1.8), respectively.
(cid:3)
4. Proof of Theorem 1.1
We begin with a lemma concerning the functions Ξ and Θ.
Lemma 4.1. Assume that TG is right invertible, and let Ξ and Θ be the functions
defined by (1.7) and (1.8), respectively. Then
(4.1)
Ker TG = TΘL2
p−m(R+) and Θ∗Ξ ∈ fW (p−m)×m
0,−
.
Proof. We split the proof into two parts.
Part 1. In this part we prove the inclusion of (4.1). Take s ∈ iR. Note that in
Proposition 3.2 it was shown for s ∈ iR that
Y (s) = I − G(s)∗bf (s) +bρ(s),
G)−1g and ρ is defined by (3.4); see (3.5). From (1.7) and (1.8) we
where f = (TGT ∗
then see that
Θ(s)∗Ξ(s) = E ∗Y (s)∗Y (s)D+ = E ∗Y (s)∗(cid:16)I − G(s)∗bf (s) +bρ(s)(cid:17)D+.
Now use that G(s)Y (s)E = DE = 0, and hence E ∗Y (s)∗G(s)∗ = 0. The latter
identity and the fact that E ∗D+ = 0 and Y = I −by imply that
Θ(s)∗Ξ(s) = −E ∗by(s)∗D+ + E ∗bρ(s)D+ − E ∗by(s)∗bρ(s)D+
= −A(s) + B(s) − C(s).
p×p(R+) ∩ L2
and by(∞) = 0, that is, by ∈ fW p×p
(4.2)
From item (b) in Proposition 4.1 we know that y ∈ L1
thus by ∈ fW p×p
A(·) := E ∗by(·)∗D+ ∈ fW (p−m)×p
0,+ . It follows that
(4.3)
0,−
+
.
Recall that ρ is given by (3.4) with f = (TGT ∗
to L1
G)−1g. Since the function g belongs
m×p(R+), item (b) in Proposition 2.1 tells us that the same holds
m×p(R+)∩ L2
p×p(R+), and
10
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
.
0,−
p×p(R) ∩ L2
true for f . It follows that g∗ ⋆ f ∈ L1
ρ ∈ L1
(4.4)
p×p(R−) ∩ L2
p×p(R−). We conclude that
B(·) := E ∗bρ(·)D+ ∈ fW (p−m)×p
Finally, note that y∗(t) = y(−t)∗ for t ∈ R and (cid:0)by(s)(cid:1)∗
by(s)∗bρ(s) =(cid:16)\y∗ ⋆ ρ(cid:17)(s),
s ∈ iR,
y∗(t − τ )ρ(τ ) dτ =Z 0
p×p(R−) ∩ L2
Since both y∗ and ρ belong to L1
Section 2 in [4]) that the same holds true for y∗ ⋆ ρ. But then
(y∗ ⋆ ρ)(t) =Z ∞
and
−∞
−∞
p×p(R). The latter implies that
= by∗(s) for s ∈ iR. Thus
y∗(t − τ )ρ(τ ) dτ.
p×p(R−), it is well known (see, e.g.,
(4.5)
C(·) := E ∗by(·)∗bρ(·)D+ ∈ fW (p−m)×p
0,−
.
Part 2. In this part we prove the identity of (4.1). Using (3.1) we see that
From (4.3), (4.4), (4.5) and (4.2) it follows that Θ∗Ξ ∈ fW (p−m)×m
s ∈ iR.
G(s)Θ(s) = G(s)Y (s)E = DE = 0,
0,−
.
0
This implies that TGTΘ = 0, and hence Im TΘ ⊂ Ker TG. To prove the reverse
inclusion, take h ∈ Ker TG. Thus h ∈ L2
It follows that
G(s)bh(s) = 0 for ℜs > 0. Put H(s) =bh(s). Then H(·) belongs to H 2
m(iR). Next
we apply Theorem 1.2. Using the identities in (1.13) we see that
p(R+) and TGh = 0.
=(cid:2)Ξ(s) Θ(s)(cid:3)(cid:20)
H(s) =(cid:2)Ξ(s) Θ(s)(cid:3)(cid:20) G(s)
E ∗Y (s)−1(cid:21) H(s)
E ∗Y (s)−1H(s)(cid:21) = Θ(s)E ∗Y (s)−1H(s).
Hence bh(s) = Θ(s)Ψ(s), where Ψ(s) = E ∗Y (s)−1bh(s). Since h ∈ L2
Y (·)−1 is a matrix function with H ∞ entries, we conclude that Ψ ∈ H 2
hence Ψ = bu for some u ∈ L2
bh(s) = Θ(s)bu(s). This shows that h = TΘu, and hence Ker TG ⊂ Im TΘ.
W (p−m)×m
Part 1. In this part we show that the equation G(s)X(s) = Im, ℜs > 0, has a
solution X ∈ W p×m
if and only if TG is right invertible. Furthermore, we show
that in that case the function Ξ defined by (1.7) is a particular solution. From the
one but last sentence of the paragraph containing (1.4) we know that it suffices to
prove the "if part" only. Therefore, in what follows we assume that TG is right
invertible. Since Ξ(s) = Y (s)D+ and Y ∈ W p×p
. Moreover,
using the identity (4.1) we have
p−m(R+). The identity bh(s) = Θ(s)Ψ(s) then yields
Proof of Theorem 1.1. From Lemma 3.2 we know that Θ is an inner function in
. The proof of the other statements is split into three parts.
+ , we have Ξ ∈ W p×m
p(R+) and
p−m, and
(cid:3)
+
+
+
G(s)Ξ(s) = G(s)Y (s)D+ = DD+ = Im.
Thus Ξ is a particular solution.
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING11
Part 2. This second part deals with the description of all in solutions in W p×m
.
Let Z be an arbitrary function in W p×m
be defined by (1.9).
Then
, and let X ∈ W p×m
+
+
+
G(s)X(s) = G(s)Ξ(s) + G(s)Θ(s)Z(s) = Im + G(s)Θ(s)Z(s), ℜs ≥ 0.
Recall that G(s)Θ(s) = G(s)Y (s)E = DE = 0. Thus G(s)X(s) = Im, ℜs ≥ 0, and
thus X is a solution.
be a solution of the equation
and G(s)H(s) = 0. Using the
To prove the converse implication, let X ∈ W p×m
G(s)X(s) = Im. Put H = X − Ξ. Then H ∈ W p×m
identities in (1.13), we obtain
E ∗Y (s)−1(cid:21) H(s)
E ∗Y (s)−1H(s)(cid:21) = Θ(s)E ∗Y (s)−1H(s).
H(s) =(cid:2)Ξ(s) Θ(s)(cid:3)(cid:20) G(s)
=(cid:2)Ξ(s) Θ(s)(cid:3)(cid:20)
0
+
+
Thus H(s) = Θ(s)Z(s), where Z(s) = E ∗Y (s)−1H(s). Since Y is invertible in
W p×p
. Together with the fact that H ∈ W p×m
+ , the function Y (·)−1 is in W p×p
,
this yields Z ∈ W (p−m)×m
Part 3. In this part we prove the identity (1.10). Assume Z ∈ fW (p−m)×m, and
let X be the function defined by (1.9). Fix u ∈ Cm. Then Z(·)u ∈ H 2
(4.6)
. It follows X has the desired representation (1.9).
m(iR), and
+
+
+
p−m(iR).
Θ(·)Z(·)u = MΘZ(·)u ∈ H 2
Here MΘ is the operator of multiplication by Θ(·) mapping H 2
Furthermore, since MΘ is an isometry, we also see that
m(iR) into H 2
p−m(iR).
(4.7)
kZ(·)uk = kΘ(·)Z(·)uk.
p×p(R+) implies that Y ∈ fW p×p
p×p(R+) ∩ L2
The fact that y ∈ L1
+ . But then
Ξ(s) = Y (s)D+ yields Ξ(·)u ∈ H 2
m(iR). Using the identity (1.9) we conclude that
Ξ(·)u also belongs to H 2
m(iR). It follows that all norms in (1.10) are well defined,
and in order to prove the identity (1.10) it suffices that to show that in H 2
m(iR) the
function Θ(·)Z(·)u is orthogonal to the function Ξ(·)v for any v ∈ Cm. The latter
fact follows from the inclusion in the second part of (4.1). Indeed, this inclusion
tells us that M ∗
ΘΞ(·)v = 0, and hence
hΞ(·)v, Θ(·)Z(·)uiH 2
m (iR) = hΞ(·)v, MΘZ(·)uiH 2
ΘΞ(·)v, Z(·)uiH 2
= hM ∗
m(iR)
m (iR) = 0.
This completes the proof.
(cid:3)
Appendix A. The Lebesgue space L1(R) ∩ L2(R)
The material in this section is standard and is presented for the convenience
of the reader. Throughout we deal with the Lebesgue spaces of complex-valued
functions on the real line L1(R) and L2(R), their vector-valued counterparts L1
m(R)
and L2
m(R). The
m(R), and the intersection of the latter two spaces: L1
m(R) ∩ L2
12
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
norms on these spaces are given by
for f ∈ L2(R),
for f ∈ L1(R),
for f = (f1, . . . , fm)⊤ ∈ L1
−∞ f (t) dt
−∞ f (t)2 dt(cid:17)1/2
kfik2
kfk1 =Z ∞
kfk2 =(cid:16)Z ∞
1(cid:17)1/2
kfk1 =(cid:16) mXi=1
2(cid:17)1/2
kfk2 =(cid:16) mXi=1
kfk0 = max{kfk1,kfk2}
m×p(R). Thus k is an m× p matrix function of which the (i, j)-th entry
kij ∈ L1(R). For each ϕ ∈ L1
p(R) the convolution products k ⋆ ϕ
and k ⋆ ψ, see (1.18), are well defined, k ⋆ ϕ belongs to L1
m(R) and k ⋆ ψ belongs to
L2
m(R). In particular, if f ∈ L1
m(R).
It follows that for a given k ∈ L1
m×p(R) the convolution product induces linear
maps from the space L1
p(R) into L1
m(R), from the space L2
m(R), and
from the space L1
m(R) ∩ L2
m(R). The resulting operators will
be denoted by K1, K2 and K0, respectively. The proof of the following lemma is
standard (see, e.g., page 216 in [6]) and therefore it is omitted.
for f = (f1, . . . , fm)⊤ ∈ L2
for f ∈ L1
m(R).
p(R) and ψ ∈ L2
p(R) ∩ L2
p(R), then k ⋆ f belongs to L1
Let k ∈ L1
p(R) ∩ L2
p(R) into L1
m(R),
m(R),
m(R) ∩ L2
m(R) ∩ L2
p(R) into L2
kfik2
Lemma A.1. The operators K1, K2 and K0 are bounded linear operators, and
(A1)
kKνk ≤ κ (ν = 1, 2, 0), where κ =(cid:16) mXi=1
pXj=1
kkijk2
1(cid:17)1/2
.
With k ∈ L1
Hankel operator H defined by
m×p(R+) we also associate the Wiener-Hopf operator W and the
(W f )(t) =Z ∞
(Hf )(t) =Z ∞
0
0
k(t − τ )f (τ ) dτ,
0 ≤ t < ∞,
k(t + τ )f (τ ) dτ,
0 ≤ t < ∞.
Using the classical relation between the convolution operator defined by k and the
operators W and H (see, e.g., Section XII.2 in [7]) it is easy to see that W and H
map the space L1
m(R+), and the space
L1
p(R+) ∩ L2
m(R+). We denote the resulting operators by
W1, W2, W0, and H1, H2, H0, respectively. Lemma A.1 shows that these operators
are bounded and
m(R+), the space L2
m(R+) ∩ L2
p(R+) into L1
p(R+) into L1
p(R+) into L2
(A2)
kWνk ≤ κ and kHνk ≤ κ (ν = 1, 2, 0), where κ =(cid:16) mXi=1
pXj=1
kkijk2
1(cid:17)1/2
.
Furthermore, using the line of reasoning in Lemma XX.2.4 in [7], we have the
following corollary.
Corollary A.2. The Hankel operators H1, H2, and H0 are the limit in operator
norm of finite rank operators, and hence compact.
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING13
Next we present an auxiliary result that is used in the proof of Proposition 2.1.
Put
(A3)
R(s) = DR +Z ∞
−∞
e−str(t) dt where
r ∈ L1
m×m(R).
By TR we denote the Wiener-Hopf operator on L2
m(R+) defined by R, that is,
(A4)
(TRf )(t) = DRf (t) +Z ∞
0
r(t − τ )f (τ ) dτ,
0 ≤ t < ∞.
As we know from the first paragraph of this section, the fact that r ∈ L1
implies that TR maps L1
m(R+) into itself.
m×m(R)
m(R+) ∩ L2
Lemma A.3. If TR is invertible as an operator on L2
space L1
m(R+) in a one-to-one way onto itself.
m(R+) ∩ L2
m(R+), then T −1
R maps the
Proof. Since TR is invertible, R admits a canonical factorization (see Section
XXX.10 in [8]), and hence we can write T −1
R = LU , where L and U are Wiener-
Hopf operators on L2
m(R+),
(A5)
(A6)
(Lf )(t) = DLf (t) +Z t
(U f )(t) = DU f (t) +Z ∞
0
t
ℓ(t − τ )f (t) dt,
0 ≤ t < ∞,
u(t − τ )f (t) dt,
0 ≤ t < ∞.
R maps L1
m(R+)∩ L2
m(R+) ∩ L2
m(R+) and f = T −1
m(R+) into L2
p(R+) and from L2
m(R+) into itself. Hence T −1
m(R+) ∩ L2
m(R+) ∩ L2
Here ℓ and u both belong to L1
in R−. The fact that both ℓ and u belong to L1
U map L1
T −1
R is one-to-one on L2
f ∈ L1
R g. This shows that T −1
T −1
m×m(R), with support of ℓ in R+ and support of u
m×m(R) implies that both L and
R has the same property. Since
m(R+). For
R TRf =
m(R+). (cid:3)
m(R+), it is also one-to-one on L1
m(R+)∩ L2
m(R+) onto L1
m(R+), respectively. Let Q be any operator on L2
m(R+), we have g = TRf ∈ L1
m(R+) ∩ L2
p×m(R+) and k ∈ L1
Lemma A.4. Let k ∈ L1
corresponding Hankel operators acting from L2
into L2
m(R+) mapping L1
m(R+) into itself, and assume that the restricted operator Q0 acting on L1
L2
m(R+) is bounded. If the operator I − HQH is invertible on L2
L2
HQH maps the space L1
m×p(R+), and let H and H be the
p(R+)
p(R+) ∩
m(R+)∩
m(R+), then I −
m(R+) in a one-to-one way onto itself.
p(R+) ∩ Lp
m(R+) ∩ L2
Proof. We know that H maps L1
m(R+). Fur-
thermore the same holds true for H with the role of p and m interchanged. Hence
our hypothesis on Q implies that I − HQH maps the space L1
m(R+)
into itself. Let M0 be the corresponding restricted operator. We have to prove
that M0 is invertible. Note that Corollary A.2 implies that M0 is equal to the
identity operator minus a compact operator, and hence M0 is a Fredholm operator
of index zero. Therefore, in order to prove that M0 is invertible, it suffices to show
that Ker M0 consists of the zero element only. Assume not. Then there exists a
non-zero f in L1
m(R+) such that M0f = 0. The fact that f belongs to
m(R+) shows that 0 = M0f = (I − HQH)f . But I − HQH
m(R+)∩ L2
L1
is assumed to be invertible. Hence f must be zero. Thus M0 is invertible.
m(R+) ⊂ L2
m(R+) ∩ L2
m(R+) ∩ L2
2(R+) into L1
m(R+) ∩ L2
(cid:3)
14
G.J. GROENEWALD, S. TER HORST, AND M.A. KAASHOEK
Acknowledgement
This work is based on the research supported in part by the National Research
Foundation. Any opinion, finding and conclusion or recommendation expressed in
this material is that of the authors and the NRF does not accept any liability in
this regard.
References
[1] H. Bart, I. Gohberg, M.A. Kaashoek, and A.C.M. Ran, Factorization of matrix and operator
functions: the state space method, Oper. Theory Adv. Appl. 178, Birkhauser Verlag, Basel,
2008.
[2] A. Bottcher and B. Silbermann, Analysis of Toeplitz operators, 2nd Edition, Springer Mono-
graphs in Mathematics, Springer Verlag, Berlin Heidelberg, 2006.
[3] L. Carleson, Interpolation by bounded analytic functions and the corona problem, Ann. Math.
76 (1962), 547 -- 559.
[4] H. Dym and I. Gohberg, On an extension problem, generalized Fourier analysis, and an
entropy formula, Integr. Equ. Oper. Theory 3 (1980), 144 -- 215.
[5] A.E. Frazho, M.A. Kaashoek, and A.C.M. Ran, Rational Matrix Solutions of a Bezout Type
Equation on the Half-plane, in: Advances in Structured Operator Theory and Related Areas.
The Leonid Lerer Anniversary Volume, Oper. Theory Adv. Appl. 237 (2012), 145 -- 160.
[6] P. Fuhrmann, On the corona theorem and its applications to spectral problems in Hilbert
space, Trans. Amer. Math. Soc. 132 (1968), 55 -- 66.
[7] I. Gohberg, S. Goldberg, and M.A. Kaashoek, Classes of Linear Operators, Volume I, Oper.
Theory Adv. Appl. 49, Birkhauser Verlag, Basel, 1990.
[8] I. Gohberg, S. Goldberg, and M.A. Kaashoek, Classes of Linear Operators, Volume II, Oper.
Theory Adv. Appl. 63, Birkhauser Verlag, Basel, 1993.
[9] I. Gohberg, M.A. Kaashoek, and L. Lerer, The resultant for regular matrix polynomials and
quasi commutativity, Indiana Univ. Math. J., 57 (2008), 2783 -- 2813.
[10] M. Green and D.J.N. Limebeer, Linear Robust Control, Prentice Hall, Englewood Cliffs, NJ,
1995.
[11] G.J. Groenewald, S. ter Horst and M.A. Kaashoek, The Bezout-Corona problem revisited:
Wiener space setting, Complex Anal. Oper. Theory 10 (2016), 115 -- 139.
[12] G. Gu and E.F. Badran, Optimal design for channel equalization via the filterbank approach,
IEEE Trans. Signal Proc. 52 (2004), 536 -- 545.
[13] G. Gu and L. Li, Worst-case design for optimal channel equalization in filterbank transceivers,
IEEE Trans. Signal Proc. 51 (2003), 2424 -- 2435.
[14] J.W. Helton, Operator Theory, analytic functions, matrices and electrical engineering, Re-
gional Conference Series in Mathematics 68, Amer. Math. Soc., Providence, RI, 1987.
[15] R. Mortini and A. Sasane, On the pre-B´ezout property of Wiener algebras on the disc and
the half-plane, New Zealand J. Math. 38 (2008), 45 -- 55.
[16] N.K. Nikol'skii, Treatise on the shift operator, Grundlehren 273, Springer Verlag, Berlin
1986.
[17] N.K. Nikol'skii, Operators, Functions and Systems, Math. Surveys Monographs 92, Amer.
Math. Soc., Providence, RI, 2002.
[18] V.V. Peller, Hankel Operators and their Applications, Springer Monographs in Mathematics,
Springer, 2003.
[19] W. Rudin, Real and Complex Analysis, McGraw-Hill, New York, 1966.
[20] V.A. Tolokonnikov, Estimates in Carleson's corona theorem. Ideals of the algebra H∞, the
problem of Szekefalvi-Nagy, Zap. Naucn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI)
113 (1981), 178 -- 198 (Russian).
[21] S. Treil, Lower bounds in the matrix corona theorem and the codimension one conjecture,
GAFA 14 (2004), 1118 -- 1133.
[22] S. Treil and B.D. Wick, The matrix-valued H p corona problem in the disk and polydisk, J.
Funct. Anal. 226 (2005), 138 -- 172.
[23] T.T. Trent and X. Zhang, A matricial corona theorem, Proc. Amer. Math. Soc. 134 (2006),
2549 -- 2558.
THE B´EZOUT EQUATION ON THE RIGHT HALF PLANE IN A WIENER SPACE SETTING15
[24] M. Vidyasagar, Control system synthesis: a factorization approach, The MIT Press, Cam-
bridge, MA, 1985.
[25] K. Zhou with J.C. Doyle and K. Glover, Robust and optimal control, Prentice Hall, NJ, 1996.
G.J. Groenewald, Department of Mathematics, Unit for BMI, North-West Univer-
sity, Private Bag X6001-209, Potchefstroom 2520, South Africa
E-mail address: [email protected]
S. ter Horst, Department of Mathematics, Unit for BMI, North-West University,
Private Bag X6001-209, Potchefstroom 2520, South Africa
E-mail address: [email protected]
M.A. Kaashoek, Department of Mathematics, VU University Amsterdam, De Boele-
laan 1081a, 1081 HV Amsterdam, The Netherlands
E-mail address: [email protected]
|
1810.07624 | 1 | 1810 | 2018-10-17T15:38:11 | Best proximity point results for almost contraction and application to nonlinear differential equation | [
"math.FA"
] | Brinde [Approximating fixed points of weak contractions using the Picard itration, Nonlinear Anal. Forum 9 (2004), 43-53] introduced almost contraction mappings and proved Banach contraction principle for such mappings. The aim of this paper is to introduce the notion of multivalued almost $\Theta$- contraction mappings and present some best proximity point results for this new class of mappings. As applications, best proximity point and fixed point results for weak single valued $\Theta$-contraction mappings are obtained. An example is presented to support the results presented herein. An application to a nonlinear differential equation is also provided. | math.FA | math |
Best proximity point results for almost contraction
and application to nonlinear differential equation
Azhar Hussain ∗Mujahid Abbas †, Muhammad Adeel‡Tanzeela Kanwal§.
Abstract
Brinde [Approximating fixed points of weak contractions using the Picard
itration, Nonlinear Anal. Forum 9 (2004), 43-53] introduced almost contrac-
tion mappings and proved Banach contraction principle for such mappings.
The aim of this paper is to introduce the notion of multivalued almost Θ-
contraction mappings and present some best proximity point results for this
new class of mappings. As applications, best proximity point and fixed point
results for weak single valued Θ-contraction mappings are obtained. An ex-
ample is presented to support the results presented herein. An application
to a nonlinear differential equation is also provided.
Mathematics Subject Classification 2010: 55M20, 47H10
Keywords: Almost contraction, Θ-contraction, best proximity points.
1
Introduction and preliminaries
The following concept was introduced by Berinde as 'weak contraction' in
[9]. But in [10], Berinde renamed 'weak contraction' as 'almost contraction'
which is appropriate.
∗Department of Mathematics, University of Sargodha, Sargodha-40100, Pakistan.
Email: [email protected],
†Department of Mathematics, Government College University, Lahore 54000, Pakistan.
Email: [email protected]
‡Department of Mathematics, University of Sargodha, Sargodha-40100, Pakistan.
Email: [email protected],
§Department of Mathematics, University of Sargodha, Sargodha-40100, Pakistan.
Email: [email protected]
1
Definition 1.1. Let (X, d) be a metric space. A mapping F : X → X
is called almost contraction or (δ, L)-contraction if there exist a constant
δ ∈ (0, 1) and some L ≥ 0 such that for any x, y ∈ X, we have
d(F x, F y) ≤ δ.d(x, y) + Ld(y, F x),
(1)
Von Neumann [28] considered fixed points of multivalued mappings in
the study of game theory. Indeed, the fixed point results for multivalued
mappings play a significant role in study of control theory and in solving
many problems of economics and game theory.
Nadler [25] used the concept of the Hausdorff metric to obtain fixed
points of multivalued contraction mappings and obtained the Banach con-
traction principle as a special case.
Here, we recall that a Hausdorff metric H induced by a metric d on a set X
is given by
H(A, B) = max{sup
x∈A
d(x, B), sup
y∈B
d(y, A)}
(2)
for every A, B ∈ CB(X), where CB(X) is the collection of the closed and
bounded subsets of X.
M. Berinde and V. Berinde [11] introduced the notion of multivalued
almost contraction as follows:
Let (X, d) be a metric space. A mapping F : X → CB(X) is called
almost contraction if there exist two constants δ ∈ (0, 1) and L ≥ 0 such
that for any x, y ∈ X, we have
H(F x, F y) ≤ δd(x, y) + LD(y, F x).
(3)
Berinde [11] proved Nadler's fixed point theorem in ( [25]):
Theorem 1.1. Let (X, d) be a complete metric space and F : X → CB(X)
a almost contraction. Then F has a fixed point..
Jleli et al. [22] defined Θ-contraction mapping as follows:
A mapping F : X → X is called Θ-contraction if for any x, y ∈ X
Θ(d(F x, F y)) ≤ [Θ(d(x, y))]k
(4)
where, k ∈ (0, 1) and Θ : (0, ∞) → (1, ∞) is a mapping which satisfies the
following conditions
(Θ1) Θ is nondecreasing;
2
(Θ2) for each sequence {αn} ⊆ R+,
0;
lim
n→∞
Θ(αn) = 1 if and only if lim
n→∞
(αn) =
(Θ3) there exist 0 < k < 1 and l ∈ (0, ∞) such that lim
α→0+
Θ(α)−1
αk = l;
Denote
Ω = {Θ : (0, ∞) → (1, ∞) : Θ satisf ies Θ1 − Θ3}.
(5)
Theorem 1.2. ( [22]) Let (X, d) be a complete metric space and F : X → X
a Θ-contraction, then F has a unique fixed point.
Hancer et al. [20] introduced the notion of multi-valued Θ-contraction
mapping as follows:
Let (X, d) be a complete metric space and T : X → CB(X) a multi-
valued mapping. Suppose that there exists Θ ∈ Ω and 0 < k < 1 such
that
Θ(H(T x, T y)) ≤ [Θ(d(x, y))]k
(6)
for any x, y ∈ X provided that H(T x, T y) > 0, where CB(X) is a collection
of all nonempty closed and bounded subsets of X.
Theorem 1.3. Let (X, d) be a complete metric space and F : X → K(X)
a multi-valued Θ-contraction, then F has a fixed point.
Let A and B be two nonempty subsets of a metric space (X, d) and
F : A → CB(B). A point x∗ ∈ A is called a best proximity point of F if
D(x∗, F x∗) = inf{d(x∗, y) : y ∈ F x∗} = dist(A, B),
where
dist(A, B) = inf{d(a, b) : a ∈ A, b ∈ B}.
If A ∩ B 6= φ, then x∗ is a fixed point of F. If A ∩ B = φ, then D(x, F x) > 0
for all x ∈ A and F has no fixed point.
Consider the following optimization problem:
min{D(x, F x) : x ∈ A}.
(7)
It is then important to study necessary conditions so that the above mini-
mization problem has at least one solution.
Since
d(A, B) ≤ D(x, F x)
(8)
3
for all x ∈ A. Hence the optimal solution to the problem
min{D(x, F x) : x ∈ A}
(9)
for which the value d(A, B) is attained is indeed a best proximity point of
multivalued mapping F.
For more results in this direction, we refer to [1, 2, 4, 5, 7, 8, 13, 14, 17, 18,
21, 24, 29, 30] and references mentioned therein.
Let A and B two nonempty subsets of X. Denote
A0 = {a ∈ A : d(a, b) = d(A, B) for some b ∈ B}
B0 = {b ∈ B : d(a, b) = d(A, B) for some a ∈ A}.
Definition 1.2.
the pair (A, B) has the P -property if
[27] Let (X, d) be a metric space and A0 6= φ, we say that
d(x1, y1) = d(A, B)
d(x2, y2) = d(A, B) (cid:27) implies that d(x1, x2) = d(y1, y2),
(10)
where x1, x2 ∈ A and y1, y2 ∈ B.
Definition 1.3.
the pair (A, B) has the weak P -property if
[31] Let (X, d) be a metric space and A0 6= φ, we say that
d(x1, y1) = d(A, B)
d(x2, y2) = d(A, B) (cid:27) implies that d(x1, x2) ≤ d(y1, y2),
(11)
where x1, x2 ∈ A and y1, y2 ∈ B.
Definition 1.4.
[12] Let (X, d) be a metric space, A, B two subsets of X
and α : A × A → [0, ∞). A mapping F : A → 2B\{φ} is called α-proximal
admissible if
α(x1, x2) ≥ 1,
d(u1, y1) = d(A, B),
d(u2, y2) = d(A, B)
implies that α(u1, u2) ≥ 1
(12)
for all x1, x2, u1, u2 ∈ A, y1 ∈ F x1 and y2 ∈ F x2.
Definition 1.5. [12] Let F : X → CB(Y ) be a multi-valued mapping, where
(X, d1), (Y, d2) are two metric spaces. A mapping F is said to be continuous
at x ∈ X if H(F x, F xn) → 0 whenever d1(x, xn) → 0 as n → ∞.
The aim of this paper is to obtain some best proximity point results
for multivalued almost Θ-contraction mappings. We also present some best
proximity point and fixed point results for single valued mappings. More-
over, an example to prove the validity and application to nonlinear differen-
tial equation for the usability of our results is presented. Our results extend,
unify and generalize the comparable results in the literature.
4
2 Best proximity points of multivalued almost Θ-
contraction
We begin with the following definition:
Definition 2.1. Let A, B be two nonempty subsets of a metric space (X, d)
and α : A × A → [0, ∞). Let Θ ∈ Ω be a continuous function. A multivalued
mapping F : A → 2B\{φ} is called almost Θ-contraction if for any x, y ∈ A,
we have
α(x, y)Θ[H(F x, F y)] ≤ [Θ(d(x, y) + λD(y, F x)))]k
(13)
where k ∈ (0, 1) and λ ≥ 0.
Theorem 2.1. Let (X, d) be a complete metric space and A, B nonempty
closed subsets of X such that A0 6= φ. Suppose that F : A → K(B) is a
continuous mapping such that
(i) F x ⊆ B0 for each x ∈ A0 and (A, B) satisfies the weak P -property;
(ii) F is α-proximal admissible mapping;
(iii) there exists x0, x1 ∈ A0 and y0 ∈ F x0 ⊆ B0 such that
d(x1, y0) = d(A, B) and α(x0, x1) ≥ 1;
(iv) F is multivalued almost Θ-contraction.
Then F has a best proximity point in A.
Proof. Let x0, x1 be two given points in A0 and y0 ∈ F x0 ⊆ B0 such
that d(x1, y0) = d(A, B) and α(x0, x1) ≥ 1. If y0 ∈ F x1, then d(A, B) ≤
D(x1, F x1) ≤ d(x1, y0) = d(A, B) implies that D(x1, F x1) = d(A, B) and
x1 is a best proximity point of F . If y0 /∈ F x1 then,
0 < D(y0, F x1) ≤ H(F x0, F x1).
Since F (x1) ∈ K(B), we can choose y1 ∈ F x1 such that
1 < Θ[d(y0, y1)] ≤ Θ[H(F x0, F x1)].
As F is multivalued almost Θ-contraction mapping, we have
1 < Θ[d(y0, y1)] ≤ α(x0, x1)Θ[H(F x0, F x1)]
≤ [Θ(d(x0, x1) + λD(x1, F x0))]k
= [Θ(d(x0, x1))]k.
(14)
5
Since y1 ∈ F x1 ⊆ B0, there exists x2 ∈ A0 such that d(x2, y1) = d(A, B)
and α(x1, x2) ≥ 1. By weak P -property of the pair (A, B) we obtain that
d(x2, x1) ≤ d(y0, y1). If y1 ∈ F x2, then x2 is a best proximity point of F . If
y1 /∈ F x2, then
D(y1, F x2) ≤ H(F x1, F x2).
We now choose y2 ∈ F x2 such that
1 < Θ[d(y1, y2)] ≤ Θ[H(F x1, F x2)]
≤ α(x1, x2)Θ[H(F x1, F x2)]
≤ [Θ(d(x1, x2) + λD(x2, F x1))]k
= [Θ(d(x1, x2))]k.
(15)
Continuing this process, we can obtain two sequences {xn} and {yn} in
A0 ⊆ A and B0 ⊆ B, respectively such that yn ∈ F xn and it satisfies
d(xn+1, yn) = d(A, B) with α(xn, xn+1) ≥ 1
where n = 0, 1, 2, .... Also,
1 < Θ[d(yn, yn+1)] ≤ Θ[H(F xn, F xn+1)]
≤ α(xn, xn+1)Θ[H(F xn, F xn+1)]
≤ [Θ(d(xn, xn+1) + λD(xn, F xn+1))]k
= [Θ(d(xn, xn+1))]k.
implies that
Since
and
1 < Θ[d(yn, yn+1)] ≤ (Θ(d(xn, xn+1)))k.
d(xn+1, yn) = d(A, B)
d(xn, yn−1) = d(A, B)
for all n ≥ 1, it follows by the weak P -property of the pair (A, B) that
d(xn, xn+1) ≤ d(yn−1, yn)
(16)
(17)
(18)
(19)
(20)
for all n ∈ N. Now by repeated application of (17), (20) and the monotone
6
property of Θ, we have
1 < Θ[d(xn, xn+1)] ≤ Θ(d(yn−1, yn)) ≤ Θ(H(F xn−1, F xn))
≤ α(xn−1, xn)Θ(H(F xn−1, F xn))
≤ [Θ(d(xn−1, xn) + λD(xn, F xn−1))]k
= (Θ(d(xn−1, xn)))k ≤ (Θ(d(yn−2, yn−1)))k
≤ (Θ(H(F xn−2, F xn−1)))k
≤ (α(xn−2, xn−1)Θ(H(F xn−2, F xn−1))k
≤ [Θ(d(xn−2, xn−1) + λD(xn−1, F xn−2))]k2
= (Θ(d(xn−2, xn−1)))k2
.
.
.
≤ (Θ(d(x0, x1)))kn
.
(21)
Θ(d(xn, xn+1)) = 1 and hence
for all n ∈ N ∪ {0}. This shows that lim
n→∞
limn→∞ d(xn, xn+1) = 0. From (Θ3), there exist 0 < r < 1 and 0 < l ≤ ∞
such that
lim
n→∞
Θ(d(xn, xn+1)) − 1
[d(xn, xn+1)]r
= l.
(22)
Assume that l < ∞ and β = l/2. From the definition of the limit there
exists n0 ∈ N such that
Θ(d(xn, xn+1)) − 1
[d(xn, xn+1)]r
(cid:12)(cid:12)(cid:12)(cid:12)
which implies that
≤ B, for all n ≥ n0
− l(cid:12)(cid:12)(cid:12)(cid:12)
Θ(d(xn, xn+1) − 1
[d(xn, xn+1)]r ≥ l − β = β for all n ≥ n0.
Hence
n[d(xn, xn+1)]r ≤ nα[Θ(d(xn, xn+1) − 1] for all n ≥ n0,
where α = 1/β. Assume that l = ∞. Let β > 0 be a given real number.
From the definition of the limit there exists n0 ∈ N such that
Θ(d(xn, xn+1) − 1
[d(xn, xn+1)]r ≥ β for all n ≥ n0
7
implies that
n[d(xn, xn+1)]r ≤ nα[Θ(d(xn, xn+1) − 1] for all n ≥ n0,
where α = 1/β. Hence, in all cases there exist α > 0 and n0 ∈ N such that
n[d(xn, xn+1)]r ≤ nα[Θ(d(xn, xn+1) − 1] for all n ≥ n0.
From (21), we have
n[d(xn, xn+1)]r ≤ nα[Θ(d(xn, xn+1) − 1]
for all n ≥ n0.
On taking the limit as n → ∞ on both sides of the above inequality, we have
lim
n→∞
n[d(xn, xn+1)]r = 0.
(23)
It follows from (23) that there exists n1 ∈ N such that
n[d(xn, xn+1)]r ≤ 1 for all n > n1.
This implies that
d(xn, xn+1) ≤
1
n1/r
for all n > n1.
Now, for m > n > n1, we have
d(xn, xm) ≤
m−1
Xi=n
d(xi, xi+1) ≤
m−1
Xi=n
1
1
r
i
.
Since 0 < r < 1,P∞
converges. Therefore d(xn, xm) → 0 as m, n → ∞.
This shows that {xn} and {yn} are Cauchy sequences in A and B, respec-
tively. Next, we assume that there exists elements u, v ∈ A such that
i=n
1
1
i
r
xn → u and yn → v as n → ∞.
Taking limit as n → ∞ in (18), we obtain that
d(u, v) = d(A, B).
(24)
Now , we claim that v ∈ T u. Since yn ∈ F xn, we have
D(yn, F u) ≤ H(F xn, F u).
8
Taking limit as n → ∞ on both sides above sides of above inequality, we have
D(v, F u) = lim
n→∞
D(yn, F u) ≤ lim
n→∞
H(F xn, F u) = 0.
As F u ∈ K(B), D(v, F u) = 0 implies that v ∈ F u. By (24), we have
D(u, F u) ≤ d(u, v) = dist(A, B) ≤ D(u, F u),
which implies that D(u, F u) = dist(A, B) and hence u is a best proximity
point of F in A.
Remark 2.1. In the next theorem, we replace the continuity assumption on
F with the following condition:
If {xn} is a sequence in A such that α(xn, xn+1) ≥ 1 for all n and
xn → x ∈ A as n → ∞, then there exists a subsequence {xn(k)} of {xn}
such that α(xn(k), x) ≥ 1 for all k. If the above condition is satisfied then
we say that the set A satisfies α− subsequential property.
Theorem 2.2. Let (X, d) be a complete metric space and (A, B) a pair
of nonempty closed subsets of X such that A0 6= φ. Let F : A → K(B)
be a multivalued mapping such that conditions (i)-(iv) of Theorem 2.1 are
satisfied. Then F has a best proximity point in A provided that A satisfies
α− subsequential property.
Proof. Following Arguments similar to those in the proof of Theorem 2.1,
we obtain two sequences {xn} and {yn} in A and B, respectively such that
α(xn, xn+1) ≥ 1,
xn → u ∈ A and yn → v ∈ B as n → ∞
and
d(u, v) = dist(A, B)
(25)
(26)
(27)
By given assumption, there exists a subsequence {xn(k)} of {xn} such that
α(xn(k), u) ≥ 1 for all k.
Since yn(k) ∈ F xn(k) for all k ≥ 1, applying
condition (iv) of Theorem 2.1, we obtain that
1 < Θ(D(yn(k), F u)) ≤ Θ(H(F xn(k), F u))
≤ α(xn(k), u)Θ(H(F xn(k), F u))
= (Θ(d(xn(k), u)))k.
(28)
On taking limit as k → ∞ in (28) and using the continuity of Θ, we have
Θ(D(v, F u) = 1. Therefore, by (Θ2) we obtain that D(v, F u) = 0. As
shown in the proof of Theorem 2.1, we have D(v, F u) = dist(A, B) and
hence u is a best proximity point of F in A.
9
Remark 2.2. To obtain the uniqueness of the best proximity point of mul-
tivalued almost Θ-contraction mappings, we propose the following H condi-
tion:
H : for any best proximity points x1, x2 of mapping F, we have
α(x1, x2) ≥ 1.
Theorem 2.3. Let A and B be two nonempty closed subsets of a com-
plete metric space (X, d) such that A0 6= φ and F : A → K(B) continuous
multivalued mapping satisfying the conditions of Theorem 2.1 (respectively
in Theorem 2.2). Then the mapping F has a unique best proximity point
provided that it satisfies the condition H.
Proof. Let x1, x2 be two best proximity points of F such that x1 6= x2,
then by the given hypothesis H we have α(x1, x2) ≥ 1 and D(x1, F x1) =
D(A, B) = D(x2, F x2). Since F x1 and F x2 are compact sets, there exists
elements y0 ∈ F x1 and y1 ∈ F x2 such that
d(x1, y0) = dist(A, B) and d(x2, y1) = dist(A, B).
As F satisfies the weak P -property, we have
d(x1, x2) ≤ d(y0, y1),
Since F is multivalued almost Θ-contraction mapping, we obtain that
Θ(d(x1, x2)) ≤ Θ(d(y0, y1)) ≤ Θ(H(F x1, F x2))
≤ α(x1, x2)Θ(H(F x1, F x2))
≤ [Θ(d(x1, x2) + λD(x2, F x1))]k
= (Θ(d(x1, x2)))k
< Θ(d(x1, x2)),
a contradiction. Hence d(x1, x2) = 0, and x1 = x2.
If the pair (A, B) satisfies the weak P -property, then it satisfies the P -
property, we have the following corollaries:
Corollary 2.1. Let (X, d) be a complete metric space and (A, B) a pair of
nonempty closed subsets of X such that A0 6= φ. Suppose that a continuous
mapping F : A → K(B) satisfies the following properties:
(i) F x ⊆ B0 for each x ∈ A0 and (A, B) satisfies the P -property;
10
(ii) F is multivalued α-proximal admissible mapping;
(iii) there exists x0, x1 ∈ A0 and y0 ∈ F x0 ⊆ B0 such that d(x1, y0) =
d(A, B) and α(x0, x1) ≥ 1;
(iv) F is multivalued almost Θ-contraction.
Then F has a best proximity point in A.
Corollary 2.2. Let (X, d) be a complete metric space and (A, B) a pair
of nonempty closed subsets of X such that A0 6= φ.
let F : A → K(B)
be a multi-valued mapping such that conditions (i)-(iv) of Corollary 2.1 are
satisfied. Then F has a best proximity point in A provided that A has α−
subsequential property.
Now we give an example to support Theorem 2.1.
Example 2.1. Let X = R2 be a usual metric space. Let
A = {(−2, 2), (2, 2), (0, 4)}
(29)
and
B = {(−8, γ) : γ ∈ [−8, 0]} ∪ {(8, γ) : γ ∈ [−8, 0]} ∪ {(β, −8) : β ∈ (−8, 8)}.
(30)
Then d(A, B) = 8, A0 = {(−2, 2), (2, 2)} and B0 = {(−8, 0), (8, 0)}.
Define the mapping F : A → K(B) by
{(−8, 0)}
{(8, 0)}
{(β, −8) : β ∈ (−8, 8)}
if x = (−2, 2)
if x = (2, 2)
if x = (0, 4).
F x =
and α : A × A → [0, ∞) by
α((x, y), (u, v)) =
11
10
.
(31)
Clearly, F (A0) ⊆ B0. For (−2, 2), (2, 2) ∈ A and (−8, 0), (8, 0) ∈ B, we have
(cid:26) d((−2, 2), (−8, 0)) = d(A, B) = 8,
d((2, 2), (8, 0)) = d(A, B) = 8.
Note that
d((−2, 2), (2, 2)) < d((−8, 0), (8, 0)).
(32)
11
that is, the pair (A, B) has weak P -property. Also, F is α-proximal admis-
sible mapping. Now we show that F is multivalued almost Θ-contraction
where Θ : (0, ∞) → (1, ∞) is given by Θ(t) = 5t.
Note that
α((−2, 2), (2, 2))Θ[H(F (−2, 2), F (2, 2))] =
11
10
(516)
(33)
and
If we take k ∈ ( 717
[Θ(d((−2, 2), (2, 2)) + λD((2, 2), (−8, 0)))]k =(cid:0) 528(cid:1)k
1250 , 1) and λ = 2 in (34), we have
Similarly,
11
10
(516) <(cid:0) 528(cid:1)k
.
.
(34)
(35)
α(x, y)Θ[H(F x, F y)] ≤ [Θ(d(x, y) + λD(y, F x)))]k
(36)
holds for the remaining pairs. Hence all the conditions of Theorem 2.1 are
satisfied. Moreover, (−2, 2), (2, 2) are best proximity points of F in A.
Remark 2.3. Note that mapping F in the above example does not hold
for the case of Nadler [25] as well as for Hancer et al. [20]. For if, take
x = (−2, 2), y = (2, 2) ∈ A, we have
Θ(H(T x, T y)) = 516 > 54 > (54)k = [Θ(d(x, y))]k
for k ∈ (0, 1). Also
H(T x, T y) = 16 > 4 = d(x, y) > αd(x, y)
for α ∈ (0, 1).
Remark 2.4. In above example 2.1, the pair (A, B) does not satisfy the
P -property and hence the Corollary 2.1 is not applicable in this case.
3 Application to single valued mappings
In this section, we obtain some best proximity point results for single-
valued mappings as applications of our results obtain in section 2.
12
Definition 3.1.
X, a nonself mapping F : A → B is called α-proximal admissible if
[22] Let (X, d) be a metric space and A, B two subsets of
α(x1, x2) ≥ 1,
d(u1, F x1) = d(A, B),
d(u2, F x2) = d(A, B).
implies α(u1, u2) ≥ 1
(37)
for all x1, x2, u1, u2 ∈ A where α : A × A → [0, ∞).
Definition 3.2. Let α : A × A → [0, ∞) and Θ : (0, ∞) → (1, ∞) a nonde-
creasing and continuous function. A mapping F : A → B is called almost
Θ-contraction if for any x, y ∈ A, we have
α(x, y)Θ(d(F x, F y)) ≤ [Θ(d(x, y) + λd(y, F x)))]k
(38)
where, k ∈ (0, 1) and λ ≥ 0.
Theorem 3.1. Let (X, d) be a complete metric space and (A, B) a pair of
nonempty closed subsets of X such that A0 is nonempty. If F : A → B is a
continuous mapping such that
(i) F (A0) ⊆ B0 and (A, B) satisfies the weak P -property;
(ii) F is α-proximal admissible mapping;
(iii) there exists x0, x1 ∈ A0 such that d(x1, F x0) = d(A, B) and α(x0, x1) ≥
1;
(iv) F is almost Θ-contraction.
Then F has a best proximity point in A.
Proof. As for every x ∈ X, {x} is compact in X. Define a multivalued map-
ping T : A → K(B) by T x = {F x} for x ∈ A. The continuity of F implies
that T is continuous. Now F (A0) ⊆ B0 implies that T x = {F x} ⊆ B0 for
each x ∈ A0. If x1, x2, v1, v2 ∈ A, y1 ∈ T x1 = {F x1} and y2 ∈ T x2 = {F x2}
are such that
α(x1, x2) ≥ 1, d(v1, y1) = dist(A, B) and d(v2, y2) = dist(A, B).
(39)
That is,
α(x1, x2) ≥ 1, d(v1, F x1) = dist(A, B) and d(v2, F x2) = dist(A, B). (40)
13
Then we have α(v1, v2) ≥ 1 as F is α-proximal admissible mapping. Hence
T is α-proximal admissible mapping.
Suppose there exist x0, x1 ∈ A0 such that d(x1, F x0) = dist(A, B) and
α(x0, x1) ≥ 1. Let y0 ∈ T x0 = {F x0} ⊆ B0. Then d(x1, F x0) = dist(A, B)
gives that d(x1, y0) = dist(A, B). By condition (iii), there exist x0, x1 ∈ A0
and y0 ∈ T x0 ⊆ B0 such that d(x1, y0) = dist(A, B) and α(x0, x1) ≥ 1.
Since F is almost Θ-contraction, we have
α(x, y)Θ(H(T x, T y)) = α(x, y)Θ[d(F x, F y)] ≤ [Θ(d(x, y) + λD(y, F x)))]k,
(41)
for any x, y ∈ A which implies that T is multivalued almost Θ-contraction.
Thus, all the conditions of Theorem 2.1 are satisfied and hence T has a best
proximity point x∗ in A. Thus we have D(x∗, T x∗) = dist(A, B) and hence
d(x∗, F x∗) = dist(A, B), that is x∗ is a best proximity point of F in A.
Theorem 3.2. Let (X, d) be a complete metric space and (A, B) a pair of
nonempty closed subsets of X such that A0 is nonempty. If F : A → B is
a mapping such that conditions (i)-(iv) of Theorem 3.1 are satisfied. Then
F has a best proximity point in A provided that A satisfies α-subsequential
property.
Proof. Let T : A → K(B) be as given in proof of Theorem 3.1. Following
arguments similar to those in the proof of Theorem 3.1, we obtain that
(i) T x ⊆ B0 for each x0 ∈ A0;
(ii) T is multi-valued α-proximal admissible mapping;
(iii) there exists x0, x1 ∈ A0 and y0 ∈ T x0 ⊆ B0 such that d(x1, y0) =
d(A, B) and α(x0, x1) ≥ 1;
(iv) T is multivalued almost Θ-contraction.
Thus, all the conditions of Theorem 2.2 are satisfied and hence T has a
best proximity point x∗ in A, that is,
D(x∗, T x∗) = dist(A, B)
Consequently, d(x∗, F x∗) = dist(A, B), and x∗ is a best proximity point of
F in A.
Corollary 3.1. Let (X, d) be a complete metric space and (A, B) a pair of
nonempty closed subsets of X such that A0 is nonempty. If F : A → B is a
continuous mapping such that
14
(i) F (A0) ⊆ B0 and (A, B) satisfies the P -property;
(ii) F is α-proximal admissible mapping;
(iii) there exists x0, x1 ∈ A0 such that d(x1, T x0) = d(A, B) and α(x0, x1) ≥
1;
(iv) F is almost Θ-contraction.
Then F has a best proximity point in A.
Proof. Replace the condition of weak P-property with P-property in Theo-
rem 3.1.
Corollary 3.2. Let (X, d) be a complete metric space and (A, B) be a pair
of nonempty closed subsets of X such that A0 is nonempty. If F : A → B is
a mapping such that conditions (i)-(iv) of Corollary 3.1 are satisfied. Then
F has a best proximity point in A provided that A satisfies α− subsequential
property.
Proof. Replace the condition of weak P-property with P-property in Theo-
rem 3.2.
4 Fixed point results for single and multi-valued
mappings
In this section, fixed points of singlevalued and multivalued almost Θ-
contraction mappings are obtained.
Taking A = B = X in Theorem 2.1 (Theorem 2.2), we obtain corre-
sponding fixed point results for almost Θ-contraction mappings.
Theorem 4.1. Let (X, d) be a complete metric space. If F : X → K(X) is
a continuous mapping satisfying
(i) F is α admissible mapping;
(ii) there exists x0 ∈ X such that α(x0, F x0) ≥ 1;
(iii) F is multivalued almost Θ-contraction.
Then F has a fixed point in X.
15
Theorem 4.2. Let (X, d) be a complete metric space. Let F : X → K(X)
be a multi-valued mapping such that conditions (i)-(iii) of Theorem 4.1 are
satisfied. Then F has a fixed point in X provided that X satisfies α− sub-
sequential property.
Taking A = B = X in Theorem 3.1 (in Theorem 3.2), we obtain the
corresponding fixed point results of almost Θ-contraction mappings.
Theorem 4.3. Let (X, d) be a complete metric space. Let F : X → X be a
continuous mapping satisfying
(i) F is α admissible mapping;
(ii) there exists x0 ∈ X such that α(x0, F x0) ≥ 1;
(iii) F is almost Θ-contraction.
Then F has a fixed point in X.
Theorem 4.4. Let (X, d) be a complete metric space. Let F : X → X be a
multi-valued mapping such that conditions (i)-(iii) of Theorem 4.1 are satis-
fied. Then F has a fixed point in X provided that X has a α−subsequential
property.
Remark 4.1. In Theorem 4.1 (respectively in 4.3)
(i). If we take α(x, y) = 1, we obtain the main results of Durmaz [16]
and Altun [3].
(ii). Taking λ = 0 and α(x, y) = 1, we obtain the main result of Hancer
et al. [20] and Jelli [22] respectively.
(iii). Taking α(x, y) = 1 and Θ(t) = et, we obtain the main result of
Berinde [11] and [9].
(iv). Taking α(x, y) = 1, λ = 0 and Θ(t) = et, we obtain the main result
of Nadler [25] and Banach [6].
5 Application to Nonlinear Differential Equations
Let C([0, 1]) be the set of all continuous functions defined on [0, 1] and
d : C([0, 1]) × C([0, 1]) → R be the metric defined by
d(x, y) = x − y∞ = max
t∈[0,1]
x(t) − y(t).
(42)
It is known that (C([0, 1]), d) is a complete metric space.
16
Let us consider the two-point boundary value problem of the second-
order differential equation:
− d2x
dt2 = f (t, x(t))
x(0) = x(1) = 0
t ∈ [0, 1];
(cid:27)
where f : [0, 1] × R → R is a continuous mapping.
The Green function associated with (43) is defined by
G(t, s) =(cid:26) t(1 − s)
s(1 − t)
if 0 ≤ t ≤ s ≤ 1,
if 0 ≤ s ≤ t ≤ 1.
Let φ : R × R → R be a given function.
Assume that the following conditions hold:
(43)
(44)
(i) f (t, a) − f (t, b) ≤ max
a,b∈R
φ(a, b) ≥ 0;
a − b for all t ∈ [0, 1] and a, b ∈ R with
(ii) there exists x0 ∈ C[0, 1] such that φ(x0(t), F x0(t)) ≥ 0 for all t ∈ [0, 1]
where F : C[0, 1] → C[0, 1];
(iii) for each t ∈ [0, 1] and x, y ∈ C[0, 1], φ(x(t), y(t)) ≥ 0 implies φ(F x(t), F y(t)) ≥
0;
(iv) for each t ∈ [0, 1], if {xn} is a sequence in C[0, 1] such that xn → x in
C[0, 1] and φ(xn(t), xn+1(t)) ≥ 0 for all n ∈ N , then φ(xn(t), x(t)) ≥ 0
for all n ∈ N .
We now prove the existence of a solution of the second order differential
equation (43).
Theorem 5.1. Under the assumptions (i)-(iv), (43) has a solution in C 2([0, 1]).
Proof. It is well known that x ∈ C 2([0, 1]) is a solution of (43) is equivalent
to x ∈ C([0, 1]) is a solution of the integral equation
x(t) =Z 1
0
G(t, s)f (s, x(s))ds, t ∈ [0, 1].
(45)
Let F : C[0, 1] → C[0, 1] be a mapping defined by
F x(t) =Z 1
0
G(t, s)f (s, x(s))ds.
(46)
17
Suppose that x, y ∈ C([0, 1]) such that φ(x(t), y(t)) ≥ 0 for all t ∈ [0, 1]. By
applying (i), we obtain that
0
F u(x) − F v(x) = Z 1
≤ Z 1
≤ Z 1
≤ Z 1
0
0
0
G(t, s)f (s, x(s))ds −Z 1
0
G(t, s)f (s, y(s))ds
G(t, s)[f (s, x(s)) − f (s, y(s))]ds
G(t, s)f (s, x(s)) − f (s, y(s))ds
G(t, s) · (max x(s) − y(s))ds
≤ x − y∞ · sup
t∈[0,1] Z 1
0
G(t, s)ds!.
Since
1
R0
G(t, s)ds = −(t2/2) + (t/2), for all t ∈ [0, 1], we have
sup
t∈[0,1] Z 1
0
G(t, s)ds! =
1
8
.
It follows that
F x − F y∞ ≤
1
8
(x − y∞
Taking exponential on the both sides, we have
eF x−F y∞ ≤ e
1
8 (x−y∞)
= [e(x−y∞)]
1
8 ,
(47)
(48)
for all x, y ∈ C[0, 1]. Now consider a function Θ : (0, ∞) → (1, ∞) by
Θ(t) = et. Define
α(x, y) =(cid:26) 1 if φ(x(t), y(t)) ≥ 0, t ∈ [0, 1],
otherwise.
0
Then from (48) with k = 1
8 , we obtain that
α(x, y)Θ(F x − F y∞) ≤ [Θ(d(x, y))]k ≤ [Θ(d(x, y) + λd(y, F x))]k.
Therefore the mapping F is multivalued almost Θ-contraction.
18
From (ii) there exists x0 ∈ C[0, 1] such that α(x0, F x0) ≥ 1. Next, for
any x, y ∈ C[0, 1] with α(x, y) ≥ 1, we have
φ(x(t), y(t)) ≥ 0
for all t ∈ [0, 1]
⇒ φ(F x(t), F y(t)) ≥ 0
for all t ∈ [0, 1]
⇒ α(F x, F y) ≥ 1,
and hence F is α-admissible. It follows from Theorem 4.3 that F has a fixed
point x in C([0, 1]) which in turns is the solution of (43).
6 Conclusion
This paper is concerned with the existence and uniqueness of the best
proximity point results for Brinde type contractive conditions via auxiliary
function Θ ∈ Ω in the framework of complete metric spaces. Also, some
fixed point results as a special cases of our best proximity point results in
the relevant contractive conditions are studied. Moreover, the correspond-
ing fixed point results are obtained. An example is discussed to show the
significance of the investigation of this paper. An application to a nonlinear
differential equation is presented to illustrate the usability of the new theory.
Acknowledgments
This paper was funded by the University of Sargodha, Sargodha funded
research project No. UOS/ORIC/2016/54. The first and third author,
therefore, acknowledges with thanks UOS for financial support.
References
[1] A. Abkar, M. Gabeleh, The existence of best proximity points for multi-
valued non-selfmappings, Rev. R. Acad. Cienc. Exactas F´ıs. Nat., Ser.
A Mat., RACSAM, 107 (2013), 319 -- 325.
[2] A. Abkar, M. Gabeleh, Global optimal solutions of noncyclic mappings
in metric spaces, J. Optim. Theory Appl., 153 (2012), 298 -- 305.
[3] I. Altun, H.A. Hnacer, G. Minak, On a general class of weakly picard
operators, Miskolc Mathematical Notes, Vol. 16 (2015), No. 1, pp. 25 --
32.
19
[4] M.A. Al-Thagafi, N. Shahzad, Convergence and existence results for
best proximity points, Nonlinear Anal., Theory Methods Appl.,70
(2009), 3665 -- 3671.
[5] A. Amini-Harandi, Best proximity points for proximal generalized con-
tractions in metric spaces, Optim. Lett., 7 (2013), 913 -- 921.
[6] S. Banach, Sur les op´erations dans les ensembles abstraits et leur appli-
cation aux equations int´egrales, Fundamenta Mathematicae, 3, 133 --
181, (1922).
[7] S.S. Basha, Best proximity points: Global optimal approximate solu-
tions, J. Glob. Optim., 49 (2011), 15 -- 21.
[8] S.S. Basha, P. Veeramani, Best proximity pair theorems for multifunc-
tions with open fibres, J. Approx. Theory, 103 (2000), 119 -- 129.
[9] V. Berinde, Approximating fixed points of weak contractions using the
Picard itration, Nonlinear Anal. Forum 9 (2004), 43 -- 53.
[10] V. Berinde, General constructive fixed point theorems for ´Ciri´c-type
almost contractions in metric spaces, Carpathian J. Math. 24 (2008),
10 -- 19.
[11] M. Berinde, V. Berinde, On a general class of multivalued weakly picard
mappings, J. Math. Anal. 326 (2007), 772 -- 782.
[12] BS. Choudhurya, P. Maitya, N. Metiya, Best proximity point results in
set-valued analysis, Nonlinear Analysis: Modelling and Control, 21(3)
2016, 293 -- 305.
[13] B.S. Choudhury, P. Maity, P. Konar, A global optimality result using
nonself mappings, Opsearch, 51 (2014), 312 -- 320.
[14] R.C. Dimri, P. Semwal, Best proximity results for multivalued map-
pings, Int. J. Math. Anal., 7 (2013), 1355 -- 1362.
[15] G. Durmaz, Some theorems for a new type of multivalued contractive
maps on metric space, Turkish Journal of Mathematics, 41 (2017),
1092 -- 1100.
[16] G. Durmaz, I. Altun, A new perspective for multivalued weakly picard
operators, Publications De L'institut Math´ematique, 101(115) (2017),
197 -- 204.
20
[17] A.A. Eldred, J. Anuradha, P. Veeramani, On equivalence of general-
ized multivalued contractions and Nadler's fixed point theorem, J. Math.
Anal. Appl., 336 (2007), 751 -- 757.
[18] A.A. Eldred, P. Veeramani, Existence and convergence of best proximity
points, J. Math. Anal. Appl., 323 (2006), 1001 -- 1006.
[19] M. Gabeleh, Best proximity points: Global minimization of multivalued
non-self mappings, Optim Lett., 8 (2014) 1101 -- 1112.
[20] H.A. Hancer, G. Mmak, I. Altun, On a broad category of multivalued
weakly Picard operators, Fixed Point Theory 18 (2017), 229 -- 236.
[21] A. Hussain, M. Adeel, T. Kanwal, N. Sultana, Set valued contraction
of Suzuki-Edelstein-Wardowski type and best proximity point results,
Bulletin of Mathematical Analysis and Applications, 10(2) (2018), 53-
67.
[22] M. Jleli, B. Samet, A new generalization of the Banach contraction
principle, J. Inequal. Appl., 2014(2014), 8 pages.
[23] Z. Kadelburg, S. Radenovi, A note on some recent best proximity point
results for non-self mappings, Gulf J. Math., 1 (2013), 36 -- 41.
[24] A. Latif, M. Abbas, A. Husain, Coincidence best proximity point of Fg-
weak contractive mappings in partially ordered metric spaces, Journal
of Nonlinear Science and Application, 9 (2016), 2448 -- 2457.
[25] S. B. Nadler Jr, Multivalued contraction mappings, Pacific J. Math.
30 (1969), 475 -- 488.
[26] E. Nazari, Best proximity points for generalized muktivalued contrac-
tions in metric spaces, Miskolc Mathematical Notes, 16 (2015), 1055 --
1062.
[27] V. Sankar Raj, A best proximity point theorem for weakly contractive
non-self-mappings, Nonlinear Anal., Theory Methods Appl., 74 (2011),
4804 -- 4808.
[28] J. Von Neuman, Uber ein okonomisches Gleichungssystem und eine
Verallgemeinerung des Brouwerschen Fixpunktsatzes, Ergebn. Math.
Kolloq. 8 (1937), 73-83.
21
[29] K. Wlodarczyk, R. Plebaniak, A. Banach, Best proximity points for
cyclic and noncyclic setvalued relatively quasi-asymptotic contractions
in uniform spaces, Nonlinear Anal., Theory Methods Appl., 70 (2009),
3332 -- 3341.
[30] K. Wlodarczyk, R. Plebaniak, C. Obczynski, Convergence theorems, best
approximation and best proximity for set-valued dynamic systems of rel-
atively quasi-asymptotic contractions in cone uniform spaces, Nonlinear
Anal., Theory Methods Appl., 72 (2010), 794 -- 805.
[31] J. Zhang, Y. Su, Q. Cheng, A note on 'A best proximity point theorem
for Geraghty-contractions', Fixed Point Theory Appl., 2013:99 (2013).
22
|
1302.0517 | 1 | 1302 | 2013-02-03T18:01:35 | On the optimal multilinear Bohnenblust--Hille constants | [
"math.FA"
] | The upper estimates for the optimal constants of the multilinear Bohnenblust--Hille inequality obtained in [J. Funct. Anal. 264 (2013), 429--463] are here improved to: {0.1cm}
{enumerate} For real scalars: $K_{n}\leq\sqrt{2}(n-1)^{0.526322}$. For complex scalars: $K_{n}\leq\frac{2}{\sqrt{\pi}}(n-1)^{0.304975}$.{enumerate} {0.1cm} \noindent We also obtain sharper estimates for higher values of $n$. For instance, \[ K_{n}<1.30379(n-1) ^{0.526322}\] for real scalars and $n>2^{8}$ and \[ K_{n}<0.99137(n-1) ^{0.304975}\] for complex scalars and $n > 2^{15}.$ | math.FA | math |
ON THE OPTIMAL MULTILINEAR BOHNENBLUST -- HILLE CONSTANTS
D. NU NEZ-ALARC ´ON*, D. PELLEGRINO**, J.B. SEOANE-SEP ´ULVEDA***,
AND D. M. SERRANO-RODR´IGUEZ*
Abstract. The upper estimates for the optimal constants of the multilinear Bohnenblust -- Hille
inequality obtained in [2] are here improved to:
(1) For real scalars: Kn ≤
(2) For complex scalars: Kn ≤ 2
We also obtain sharper estimates for higher values of n. For instance,
√2 (n − 1)0.526322 .
√π (n − 1)0.304975.
for real scalars and n > 28 and
for complex scalars and n > 215.
Kn < 1.30379 (n − 1)0.526322
Kn < 0.99137 (n − 1)0.304975
1. Preliminaries. First estimates
Let K be the real or complex scalar field. The multilinear Bohnenblust -- Hille inequality asserts
that for every positive integer n ≥ 1 there exists a sequence of positive scalars (Cn)∞n=1 in [1,∞)
such that
N
Xi1,...,in=1(cid:12)(cid:12)U (ei1 , . . . , ein)(cid:12)(cid:12)
n+1
2n
2n
n+1
≤ Cn
sup
z1,...,zn∈DN U (z1, . . . , zn)
for all n-linear forms U : KN × ··· × KN → K and every positive integer N , where (ei)N
i=1 denotes
the canonical basis of KN and DN represents the open unit polydisc in KN . In the last years a
considerable effort was spent in the searching of optimal values for the constants Cn; for details and
the state-of-the-art we refer to [2]. The notation and terminology used in this note are the same as
those from [2], where it is proved that the optimal multilinear Bohnenblust -- Hille constants (Kn)∞n=2
satisfy
(1.1)
and
Kn < 1.65 (n − 1)0.526322 + 0.13 (real scalars)
Kn < 1.41 (n − 1)0.304975 − 0.04 (complex scalars).
The proof of the above estimates is achieved by following a series of technical steps. In the case of
real scalars, using some previous lemmata, it is observed that the sequence
(cid:0)√2(cid:1)n−1
DM n
2
DM n+1
2
if n = 1, 2
if n is even, and
if n is odd
Mn =
*Supported by Capes.
**Supported by CNPq Grant 301237/2009-3.
***Supported by grant MTM2012-34341.
1
2
D. NU NEZ ET AL.
satisfies the multilinear Bohnenblust -- Hille inequality, where D = e
. Then, using a "uniform
approximation" argument, the estimate (1.1) is achieved. In this section we remark that this final
step of the proof, i.e., the uniform approximation argument, can be dropped and a quite simple
argument provides even better constants. In fact, from [2] we know that, for all k ≥ 1 and n ≥ 2,
we have
1
1−
2 γ
√2
Mn = √2Dk−1 whenever n ∈ Bk = {2k−1 + 1, . . . , 2k}.
Thus, k − 1 ≤ log2 (n − 1) and, hence,
≤
Using a similar argument (for complex scalars) it follows that
√2Dlog2(n−1) = √2 (n − 1)
Mn ≤
√2 (n − 1)0.526322 .
log2 e
1−
1
2 γ
√2 !
Mn ≤
2
√π
(n − 1)
log2(cid:16)e
1
2 −
1
2 γ(cid:17) ≤
2
√π
(n − 1)0.304975
for the complex scalar field. Of course, the other estimates of [2] related to the above results can
be straightforwardly improved by using these new estimates.
2. Sharper estimates for big values of n
In this section we improve our previous estimates for large values of n.
2.1. Real case. If (Cn)∞n=1 denotes the sequence in [2, (4.3)], if we fix any k0, it is obvious that
1
with D = e1−
2 γ
√2
such that
Then
and
Jn =
Cn
DJ n
2
2 (cid:17)
D(cid:16)J n−1
n+1
2n
n−1
2n (cid:16)J n+1
2 (cid:17)
if n ≤ 2k0 ,
if n > 2k0 is even, and
if n > 2k0 is odd
, satisfies the multilinear Bohnenblust -- Hille inequality. For n > 2k0, let k1 > k0 be
2k1−1 + 1 ≤ n ≤ 2k1 .
k1 − k0 ≤ log2(cid:18) n − 1
2k0−1(cid:19)
Kn ≤ J2k1 = Dk1−k0 C2k0
log2(cid:16) n−1
2k0−1(cid:17)
≤ C2k0 D
C2k0
Dk0−1 (n − 1)log2 D .
=
We thus have
Kn ≤
From [3, Theorem 3.1] we know that
C2k0
Dk0−1 (n − 1)
log2 e
1−
1
2 γ
√2 ! .
(2.1)
C2k0 ≤ 4Dk0−4
ON THE OPTIMAL MULTILINEAR BOHNENBLUST -- HILLE CONSTANTS
3
whenever k0 ≥ 4. Thus,
Kn ≤
4
√2 (cid:19)3 (n − 1)
(cid:18) e1−
1
2 γ
log2 e
1−
1
2 γ
√2 ! < 1.338887 (n − 1)0.526322 .
Summarizing, we have:
Theorem 2.1. If n > 16, then
Kn ≤
(cid:18) e
log2 e
1−
1
2 γ
√2 ! .
4
1−
1
2 γ
√2 (cid:19)3 (n − 1)
Numerically,
(2.2)
Kn < 1.338887 (n − 1)0.526322 .
If we use the exact value of C2k0 instead of estimate (2.1) we can improve (2.2) as n grows. For
example,
n > 26 ⇒ Kn < 1.310883 (n − 1)0.526322
n > 27 ⇒ Kn < 1.306156 (n − 1)0.526322
n > 28 ⇒ Kn < 1.303787 (n − 1)0.526322 .
2.2. Complex case. Let (Cn)∞n=1 denote the sequence in [1, Theorem 2.3]. If we fix any k0, and
as the authors did in [2], we can show that
Jn =
Cn
DJ n
2
2 (cid:17)
D(cid:16)J n−1
n+1
2n
n−1
2 (cid:17)
2n (cid:16)J n+1
if n ≤ 2k0,
if n > 2k0 is even, and
if n > 2k0 is odd,
1
1
with D = e
the real case we obtain
2−
2 γ, satisfies the multilinear Bohnenblust -- Hille inequality. For n > 2k0, by mimicking
Thus, using the values of C2k from [1] we have
Kn ≤
C2k0
Dk0−1 (n − 1)
log2(cid:16)e
1
2 −
1
2 γ(cid:17) .
n > 23 ⇒ Kn < 1.02960973695 (n − 1)0.304975 ,
n > 24 ⇒ Kn < 1.01089344604 (n − 1)0.304975 ,
n > 25 ⇒ Kn < 1.00123230777 (n − 1)0.304975 ,
n > 26 ⇒ Kn < 0.99632125476 (n − 1)0.304975 , . . .
n > 214 ⇒ Kn < 0.99137409768 (n − 1)0.304975 ,
n > 215 ⇒ Kn < 0.99136434217 (n − 1)0.304975 , . . .
n > 225 ⇒ Kn < 0.99135459597 (n − 1)0.304975 , . . .
n > 250 ⇒ Kn < 0.99135458644 (n − 1)0.304975 .
4
D. NU NEZ ET AL.
References
[1] D. Nunez-Alarc´on, D. Pellegrino, J.B. Seoane-Sep´ulveda, On the Bohnenblust -- Hille inequality and a variant of
Littlewoods 4/3 inequality, J. Funct. Anal. 264 (2013), 326 -- 336.
[2] D. Nunez-Alarc´on, D. Pellegrino, J.B. Seoane-Sep´ulveda, D.M. Serrano-Rodr´ıguez, There exist multilinear
Bohnenblust -- Hille constants (Cn)∞n=1 with limn→∞
(Cn+1 − Cn) = 0, J. Funct. Anal. 264 (2013), 429 -- 463.
[3] D.M. Serrano-Rodr´ıguez,
Improving the closed formula for subpolynomial constants in the multilinear
Bohnenblust -- Hille inequalities, Linear Alg. Appl., in press.
Departamento de Matem´atica,
Universidade Federal da Para´ıba,
58.051-900 - Joao Pessoa, Brazil.
E-mail address: [email protected]
Departamento de Matem´atica,
Universidade Federal da Para´ıba,
58.051-900 - Joao Pessoa, Brazil.
E-mail address: [email protected] and [email protected]
Departamento de An´alisis Matem´atico,
Facultad de Ciencias Matem´aticas,
Plaza de Ciencias 3,
Universidad Complutense de Madrid,
Madrid, 28040, Spain.
E-mail address: [email protected]
Departamento de Matem´atica,
Universidade Federal da Para´ıba,
58.051-900 - Joao Pessoa, Brazil.
E-mail address: [email protected]
|
1703.02909 | 1 | 1703 | 2017-03-08T16:48:00 | Weak separation properties for closed subgroups of locally compact groups | [
"math.FA",
"math.OA"
] | Three separation properties for a closed subgroup $H$ of a locally compact group $G$ are studied: (1) the existence of a bounded approximate indicator for $H$, (2) the existence of a completely bounded invariant projection of $VN\left(G\right)$ onto $VN_{H}\left(G\right)$, and (3) the approximability of the characteristic function $\chi_{H}$ by functions in $M_{cb}A\left(G\right)$ with respect to the weak$^{*}$ topology of $M_{cb}A\left(G_{d}\right)$. We show that the $H$-separation property of Kaniuth and Lau is characterized by the existence of certain bounded approximate indicators for $H$ and that a discretized analogue of the $H$-separation property is equivalent to (3). Moreover, we give a related characterization of amenability of $H$ in terms of any group $G$ containing $H$ as a closed subgroup. The weak amenability of $G$ or that $G_{d}$ satisfies the approximation property, in combination with the existence of a natural projection (in the sense of Lau and \"Ulger), are shown to suffice to conclude (3). Several consequences of (2) involving the cb-multiplier completion of $A\left(G\right)$ are given. Finally, a convolution technique for averaging over the closed subgroup $H$ is developed and used to weaken a condition for the existence of a bounded approximate indicator for $H$. | math.FA | math |
WEAK SEPARATION PROPERTIES FOR CLOSED
SUBGROUPS OF LOCALLY COMPACT GROUPS
ZSOLT TANKO
Abstract. Three separation properties for a closed subgroup H of a
locally compact group G are studied: (1) the existence of a bounded
approximate indicator for H, (2) the existence of a completely bounded
invariant projection V N (G) → V NH (G), and (3) the approximability
of the characteristic function χH by functions in McbA (G) with respect
to the weak∗ topology of McbA (Gd). We show that the H-separation
property of Kaniuth and Lau is characterized by the existence of certain
bounded approximate indicators for H and that a discretized analogue
of the H-separation property is equivalent to (3). Moreover, we give
a related characterization of amenability of H in terms of any group
G containing H as a closed subgroup. The weak amenability of G or
that Gd satisfies the approximation property, in combination with the
existence of a natural projection (in the sense of Lau and Ülger), are
shown to suffice to conclude (3). Several consequences of (2) involving
the cb-multiplier completion of A (G) are given. Finally, a convolution
technique for averaging over the closed subgroup H is developed and
used to weaken a condition for the existence of a bounded approximate
indicator for H.
1. Introduction
Our objective is to study connections between various forms of amenability
for a locally compact group G and certain separation properties for closed
subgroups, and moreover to establish relationships between these separa-
tion properties. Following the influential work of Ruan [34], much work
on the homology of the Fourier algebra A (G) as a completely contractive
Banach algebra has affirmed this as the appropriate category in which to
consider A (G) and the related algebras of abstract harmonic analysis (e.g.
[1, 9, 17, 18]). Motivated by the success of this perspective, we focus on
completely bounded projections, operator amenability, and the completely
bounded multiplier algebra of A (G). We consider the following separation
properties for a closed subgroup H of G:
(1) The existence of a bounded approximate indicator for H.
(2) The existence of a completely bounded A (G)-bimodule projection of
V N (G) onto V NH (G).
2010 Mathematics Subject Classification. Primary 43A15, Secondary 43A22, 43A30,
46L07.
Key words and phrases. locally compact group, group von Neumann algebra, Fourier
algebra, completely bounded multiplier, invariant projection.
1
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
2
(3) When the characteristic function χH may be approximated by func-
tions in B (G) or McbA (G) in the weak∗ topology on the correspond-
ing algebra of the discretized group.
Condition (3) is also considered for subsets of G that are not necessarily
closed subgroups.
Bounded approximate indicators for closed subgroups were introduced in
[1] as a means of obtaining invariant projections. They have subsequently
been shown to have an intimate connection with homological properties
of A (G) and its completion in the cb-multipliers Acb (G) [9]. A result of
Granirer and Leinert [19, Theorem B2] yields bounded approximate indi-
cators in B (G) from weaker conditions than given in Definition 2.2. This
useful tool is unavailable for nets in McbA (G) and Section 6 develops a
convolution technique relative to the closed subgroup H that recovers the
weakened condition in the cb-multiplier setting.
The existence of invariant projections has been studied by several authors
in connection with other separation properties and the existence of approx-
imate identities for ideals in A (G) [1, 9, 10, 15, 17, 24]. In particular, in
[24] it is shown that if G has the H-separation property, then an invariant
projection V N (G) → V NH (G) exists. We show in Section 4 that, in fact,
the H-separation property is equivalent to the existence of a bounded ap-
proximate indicator for H consisting of positive definite functions that are
identically one on H. An analogue of the H-separation property is more-
over shown to characterize the B (G)-approximability of χH. Condition (3)
was first studied in [1], where it was claimed that a bounded approximate
indicator for H exists whenever χH is B (G)-approximable. This argument
was later found to contain a gap [2]. We give examples in Section 3 showing
that the cb-multiplier analogue is false.
Conditions (1) to (3) are related to amenability properties of G and to
homological properties of A (G), and it is this connection that is the main
focus of the present article. We show in Section 4 that H is already amenable
when χH is A (G)-approximable for any locally compact group G contain-
ing H as a closed subgroup. In the case that G is amenable, the algebra
A (G) has a bounded approximate identity and [15, Proposition 6.4] then as-
serts that an invariant projection V N (G) → V NH (G) exists exactly when
the ideal IA(G) (H) in A (G) has a bounded approximate identity. By [17],
the latter occurs for every closed subgroup of G and it follows that an ap-
proximate indicator exists for every closed subgroup, since (1G − eα)α is an
approximate indicator for H when (eα)α is a bounded approximate identity
for IA(G) (H). Thus all closed subgroups of an amenable locally compact
group are separated in the strongest sense that we consider. For generic
locally compact groups the situation is more complicated, although some
strong connections are known to hold in general. For example, using the
identity A (G) b⊗A (G) = A (G × G), it is routine to show that an approx-
imate indicator for the diagonal G∆ in A (G × G) (and bounded there) is
exactly a bounded approximate diagonal for A (G), the existence of which
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
3
characterizes amenability of G [34]. Moreover, the existence of an invariant
projection V N (G × G) → V NG∆ (G × G) characterizes the operator biflat-
ness of A (G), a weaker homological condition than operator amenability.
Contractive operator biflatness, which asks that this invariant projection
to be a complete contraction, was recently shown to be equivalent to the
existence of a contractive approximate indicator for G∆ in B (G × G) [9].
A bounded approximate indicator for H always yields the approximability
of χH in the corresponding algebra, however it is unclear when the latter
follows from the existence of an invariant projection V N (G) → V NH (G)
alone.
In Section 3 we show that if H satisfies certain weak forms of
amenability, then the existence of a bounded map V N (G) → V NH (G) sat-
isfying λ (s) 7→ χH (s) λ (s) -- a weaker condition than the existence of an
invariant projection onto V NH (G) -- implies the McbA (G)-approximability
of χH. We give an example in which the former condition fails while the
latter holds. Establishing relations amongst the conditions (1) to (3) in the
presence of amenability type conditions on H or G is the second goal of this
article.
2. Preliminaries
For a locally compact group G, the following algebras were defined by
Eymard in [13], who established the basic properties we outline below. The
space of coefficient functions of strongly continuous unitary representations
of G,
(2.1) B (G) = {hπ (·) ξηi : π : G → B (H) is a representation, ξ, η ∈ H} ,
forms a commutative completely contractive Banach algebra, the Fourier --
Stieltjes algebra of G, under pointwise multiplication and norm
kukB(G) = inf {kξk kηk : u = hπ (·) ξηi with π, ξ, η as in (2.1)} .
The operator space structure on B (G) arises from its identification with
the dual space of the universal enveloping C ∗-algebra of L1 (G) -- the group
C ∗-algebra C ∗ (G) of G -- via the duality
hu, π (f )iB(G),C ∗(G) =ZG
f u
(cid:16)u ∈ B (G) , f ∈ L1 (G)(cid:17) .
We refer to [12] for the theory of operator spaces and completely contractive
Banach algebras. The positive definite functions in B (G) are those that
correspond to positive functionals on C ∗ (G) and are denoted P (G). For
u ∈ P (G) we have kukB(G) = kukL∞(G) = u (e). The adjoint on B (G) is
given by
u∗ (s) = u (s−1)
(u ∈ B (G) , s ∈ G)
and the self-adjoint functions in B (G) correspond to the self-adjoint bounded
functionals on C ∗ (G). Consequently, given u ∈ B (G) self-adjoint, there ex-
ist u± ∈ P (G) such that u = u+ − u− and kukB(G) = ku+kB(G) + ku−kB(G).
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
4
The Fourier algebra of G is the closed ideal A (G) of B (G) given by the
coefficients of the left regular representation λ : G → B (L2 (G)), which is
defined by
λ (s) ξ (t) = ξ(cid:16)s−1t(cid:17)
(cid:16)s, t ∈ G, ξ ∈ L2 (G)(cid:17) .
When necessary, we denote this representation of G by λG. The Fourier
algebra coincides with the closure of the compactly supported functions in
B (G), is closed under the adjoint, and is regular in the sense that for any
K ⊂ G compact and U ⊃ K open, there is a function v ∈ A (G) with
v (K) = 1 and v (G \ U) = 0. The group von Neumann algebra of G is the
weak operator topology closure V N (G) of spanλ (G) in B (L2 (G)) and is
identified with the dual of the Fourier algebra via
hλ (s) , uiV N (G),A(G) = u (s)
(s ∈ G, u ∈ A (G)) .
Given an algebra A of functions on G and a subset E of G, we denote by
IA (E) the ideal of functions in A vanishing on E. For a closed subgroup H
of G, the annihilator IA(G) (H)⊥ coincides with the von Neumann algebra
V NH (G) generated by λG (H) [39, Theorem 6], which is identified with
V N (H) via the normal ∗-isomorphism defined by λH (h) 7→ λG (h) for h ∈
H. The preadjoint of this normal ∗-isomorphism is the restriction map rH :
A (G) → A (H), which is thus a complete quotient. The closed subgroups
of G are sets of spectral synthesis for A (G) [20], meaning that the ideal
IA(G) (H) of A (G) coincides with the closure of the functions in A (G) that
have compact support disjoint from H.
The completely bounded (cb-) multiplier algebra McbA (G) of A (G) is the
algebra of functions m on G for which mA (G) ⊂ A (G) and the map Mm :
V N (G) → V N (G) is completely bounded, where Mm is the adjoint of the
multiplication map u 7→ mu on A (G). Such functions are continuous and
bounded, and form a completely contractive Banach algebra under pointwise
multiplication and norm kmkMcbA(G) = kMmkcb. The cb-multipliers of A (G)
admit the following representation theorem [22].
Theorem 2.1. (Gilbert's representation theorem) Let G be a locally com-
pact group. A function m on G is in McbA (G) if and only if there exists a
Hilbert space H and bounded continuous maps P, Q : G → H such that
m(cid:16)s−1t(cid:17) = hP (t) Q (s)i
(s, t ∈ G) .
The norm kmkMcbA(G) is the infimum of the quantities kP k∞ kQk∞ taken
over all such maps P and Q and Hilbert spaces H.
It follows from Gilbert's representation theorem that B (G) ⊂ McbA (G)
and that k·kMcbA(G) ≤ k·kB(G) on B (G). The restriction rH : McbA (G) →
McbA (H) is a well defined complete contraction [11, Proposition 1.12]. The
norm closure of A (G) in the potentially smaller cb-multiplier norm is de-
noted Acb (G) and forms a completely contractive Banach subalgebra of the
cb-multipliers with spectrum G [18, Proposition 2.2].
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
5
Since cb-multipliers of A (G) lie in L∞ (G), we may consider L1 (G) as a
subspace of the dual of McbA (G). Taking the completion of L1 (G) with
respect to the norm given, for f ∈ L1 (G), by
kf kQ(G) = sup(cid:26)(cid:12)(cid:12)(cid:12)(cid:12)ZG
f m(cid:12)(cid:12)(cid:12)(cid:12) : m ∈ McbA (G) with kmkMcbA(G) ≤ 1(cid:27)
yields a predual Q (G) for McbA (G) [11, Proposition 1.10]. With this pre-
dual, the cb-multipliers McbA (G) form a completely contractive dual Ba-
nach algebra in the sense of Runde [35]. It follows from Theorem 2.1 that
k·kL∞(G) ≤ k·kMcbA(G) and consequently k·kQ(G) ≤ k·kL1(G) on L1 (G). We let
Cc (G) and Cb (G) denote respectively the continuous compactly supported
and continuous bounded functions of G. We have
A (G) ⊂ Acb (G) and A (G) ⊂ B (G) ⊂ McbA (G) ⊂ Cb (G) .
The second containment is strict unless G is compact. An unpublished
result of Ruan asserts that A (G) is closed in McbA (G) exactly when G is
amenable, so that that the first and third containments are strict unless G
is amenable. For u ∈ A (G) and v ∈ B (G),
kuk∞ ≤ kukAcb(G) = kukMcbA(G) ≤ kukA(G) = kukB(G) ,
kvk∞ ≤ kvkMcbA(G) ≤ kvkB(G) ,
so the first, third, and fourth inclusions are in general contractive while the
second is isometric.
The locally compact group G equipped with the discrete topology is de-
noted Gd. The inclusions B (G) ⊂ B (Gd) and McbA (G) ⊂ McbA (Gd) are
complete isometries ([13] and [38, Corollary 6.3], respectively). For each
s ∈ G the point mass δs ∈ ℓ1 (Gd) is contained in C ∗ (Gd) and in Q (Gd)
as the evaluation functional at s, from which it follows that convergence in
the weak∗ topology of B (Gd) or McbA (Gd) implies pointwise convergence.
It will be important for us that, on bounded sets, the converse holds (see
[13] regarding B (Gd) and [18, Lemma 2.6] or the useful Appendix A of [27]
regarding McbA (Gd)).
The separation properties for closed subgroups that we discuss are defined
as follows.
Definition 2.2. Let G be a locally compact group and H a closed subgroup.
(1) A bounded approximate indicator for H is a bounded net (mα)α in
McbA (G) satisfying
(a) kurH (mα) − ukA(H) → 0 for all u ∈ A (H), and
(b) kumαkA(G) → 0 for all u ∈ IA(G) (H).
If (mα)α is in B (G) and is also bounded there, then we refer to a
bounded approximate indicator in B (G).
(2) A completely bounded projection V N (G) → V NH (G) is invariant
if it is an A (G)-bimodule map.
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
6
(3) Given a norm closed subalgebra A of B (G) or McbA (G) and E ⊂ G,
the characteristic function χE is called A-approximable if χE is in the
weak∗ closure of A in B (Gd) or McbA (Gd), respectively.
The operator amenability of the Fourier algebra asserts the existence of a
bounded approximate diagonal in the operator space projective tensor prod-
uct A (G) b⊗A (G). This is a bounded net (dα)α in the tensor product which
satisfies the norm convergence
u · dα − dα · u → 0 and ∆ (dα) u → u
(u ∈ A (G)) ,
where ∆ : A (G) b⊗A (G) → A (G) is the completely bounded linearization
of the multiplication and the A (G) action on the tensor product is given by
u·(v ⊗ w) = uv⊗w and (v ⊗ w)·u = v⊗wu, for u, v, w ∈ A (G). The product
map ∆ is a complete quotient and the operator biflatness of A (G) asserts the
existence of a completely bounded A (G)-bimodule left inverse to its adjoint
∆∗ : V N (G) → V N (G) ⊗V N (G), where the A (G) action on V N (G) is the
dual action. The identification A (G) b⊗A (G) = A (G × G) (see [12, Theo-
rem 7.2.4]) yields ker ∆ = IA(G×G) (G∆), from which it follows that such a left
inverse is exactly an invariant projection V N (G × G) → V NG∆ (G × G). A
thorough account of the homological conditions we consider is given in [36].
Leptin's classical result states that amenability of a locally compact group
G is characterized by the existence of a bounded approximate identity in
A (G) [29]. The locally compact group G is called weakly amenable when
Acb (G) has a bounded approximate identity. This weaker notion was intro-
duced by de Cannière and Haagerup [11], who showed that the free group on
two generators is weakly amenable. The weak amenability of G is equivalent
to the assertion that every cb-multiplier is the weak∗ limit of a bounded net
in A (G). When A (G) is merely weak∗ dense in McbA (G), the group G is
said to have the approximation property (see [21]).
3. Approximability of characteristic functions
In this section, we investigate when characteristic functions of subsets of
a locally compact group G are approximable. For a closed subgroup H of
G, the assertion that χH is approximable may be viewed as a very weak
form of subgroup separation. In [1], the discretized Fourier -- Stieltjes algebra
Bd (G) is defined to be the weak∗ closure of B (G) in B (Gd). We make the
analogous definition for the cb-multipliers of G.
Definition 3.1. The discretized cb-multiplier algebra M d
cbA (G) of a locally
compact group G is the weak∗ closure of McbA (G) in McbA (Gd). The pred-
ual Q (Gd) /M d
cb (G)
denote the weak∗ closures of A (G) in B (Gd) and in McbA (Gd), respectively.
cbA (G) is denoted Qd (G). Let Ad (G) and Ad
cb (G)⊥ of M d
Given E ⊂ G, for χE to be approximable,
it must already be that
χE ∈ McbA (Gd), and the subsets of G for which this occurs are not well
understood when Gd is not amenable. In the amenable case, the algebras
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
7
McbA (Gd) and B (Gd) coincide [31] and the Cohen-Host idempotent theo-
rem provides a complete description of the subsets of G with characteristic
function in B (Gd). For discrete groups G, determining the approximable
characteristic functions is exactly the problem of determining the sets with
characteristic function in the cb-multipliers. When G is moreover weakly
amenable, Corollary 5.4 of [9] together with Corollary 3.5 of [18] implies
that such sets E are exactly those for which the ideal IA(G) (E) has a cb-
multiplier bounded approximate identity.
Example 3.2. Let G be a locally compact group and H a closed subgroup
for which a bounded approximate indicator (mα)α exists. We show that χH
is approximable. If s ∈ H, then we may find u ∈ A (H) with u (s) = 1, in
which case
mα (s) − 1 ≤ kurH (mα) − ukL∞(H) ≤ kurH (mα) − ukA(H) → 0.
If s ∈ G \ H, then [13, Lemme 3.2] asserts that we may find w ∈ IA(G) (H)
with w (s) = 1, and
mα (s) ≤ kwmαkL∞(G) ≤ kwmαkA(G) → 0.
Since McbA (G) is contained in McbA (Gd) and weak∗ and pointwise conver-
gence coincide on bounded subsets of McbA (Gd), it follows that (mα)α has
weak∗ limit χH in McbA (Gd).
The algebra M d
cbA (G) is a weak∗ closed subalgebra of McbA (Gd) and thus
has separately weak∗ continuous multiplication. If we impose a rather weak
condition on the discrete group Gd, then every function in McbA (Gd) may
be approximated in the weak∗ topology by functions in A (G).
Proposition 3.3. Let G be a locally compact group. The inclusion
Acb (Gd) ⊂ Ad
cb (G) always holds. Consequently, if Gd has the approxima-
tion property, then McbA (Gd) = Ad
cb (G).
Proof. By Proposition 1 of [4] we have A (Gd) ⊂ Ad (G) and, because weak∗
convergence in B (Gd) implies weak∗ convergence in McbA (Gd), it follows
that A (Gd) ⊂ Ad
cb (G). Taking norm closures in McbA (Gd) yields Acb (Gd) ⊂
cb (G). If (eα)α is a net in A (Gd) converging weak∗ to 1G in McbA (Gd)
Ad
w∗
→ m in
and if m ∈ McbA (Gd), then meα ∈ A (Gd) for each α and meα
McbA (Gd), implying that m ∈ Ad
(cid:3)
cb (G).
In Section 3 of [1], that the Fourier -- Stieltjes algebra is the dual of a C ∗-
algebra is noted to imply that the unique weak∗ continuous extension of the
inclusion B (G) ⊂ Bd (G) is a quotient map B (G)∗∗ → Bd (G). We provide
a more concrete construction of the analogous canonical map McbA (G)∗∗ →
M d
cbA (G). Let ιd :
McbA (G) → McbA (Gd) and κQ : Q (Gd) → Q (Gd)∗∗ denote the inclusion
maps. By the bipolar theorem M d
cbA (G) that exploits the relation between McbA (G) and M d
cbA (G)⊥ = McbA (G)⊥ and we have
κQ (McbA (G)⊥) ⊂ McbA (G)⊥ = im (ιd)⊥ = ker (ι∗
d) ,
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
8
which together imply that the composition
κQ→ Q (Gd)∗∗ ι∗
Q (Gd)
d→ Q (G)∗∗
induces a map Qd (G) → Q (G)∗∗. Denote the adjoint of this induced map
by τ : McbA (G)∗∗ → M d
cbA (G). It is straightforward to verify that τ extends
inclusion McbA (G) ⊂ M d
cbA (G), so that τ (m) (s) = m (s) for m ∈ McbA (G)
and s ∈ G. It follows that if a net (mα)α in McbA (G) converges weak∗ to
ω ∈ McbA (G)∗∗, then
τ (ω) (s) = lim
α
mα (s)
(s ∈ G) ,
so that the map τ extracts pointwise limits from nets in McbA (G) that are
weak∗ convergent in the bidual. Moreover, the range of τ consists of exactly
those functions in M d
cbA (G) that are limits of bounded nets in McbA (G).
Proposition 3.4. Let G be a locally compact group and E ⊂ G. If there
is a bounded map Ψ : V N (G) → V N (G) satisfying Ψ (λ (s)) = χE (s) λ (s)
for all s ∈ G, then χEA (G) ⊂ Ad
cb (G). If, moreover, χE ∈ McbA (Gd) and
1G is Acb (G)-approximable, then χE is Acb (G)-approximable.
Proof. Let κA : A (G) → A (G)∗∗ and ιA : A (G) → McbA (G) be the inclu-
sions and let σ denote the composition
A (G) κA−→ A (G)∗∗ Ψ∗
−→ A (G)∗∗ ι∗∗
A−→ McbA (G)∗∗
τ−→ M d
cbA (G) .
For u ∈ A (G) and s ∈ G, with δs denoting the point mass at s in ℓ1 (Gd) ⊂
Q (Gd) ⊂ Qd (G),
σ (u) (s) = hσ (u) , δsiM d
cbA(G),Qd(G)
dκQ (δs)iMcbA(G)∗∗,McbA(G)∗
A Ψ∗κA (u) , ι∗
= hι∗∗
= hΨ∗κA (u) , λ (s)iA(G)∗∗,V N (G)
= hΨ (λ (s)) , uiV N (G),A(G)
= χE (s) u (s) .
cbA (G), it maps A (G) into Ad
Thus χEu = σ (u) ∈ M d
inclusion of McbA (G) into M d
together with the weak∗ continuity of τ ι∗∗
in Ad
then that χE ∈ McbA (Gd) implies χEeα
multiplication in McbA (Gd). Since χEeα ∈ Ad
Acb (G)-approximable.
cbA (G) for all u ∈ A (G). Since τ extends the
cb (G), which
A implies that σ in fact has range
cb (G). If (eα)α is a net in A (G) converging weak∗ to 1G in McbA (Gd),
w∗
→ χE, by weak∗ continuity of
cb (G), we conclude that χE is
(cid:3)
For a subgroup H of a locally compact group G,
it is straight for-
ward to verify that χH is a positive definite function on Gd, so that
χH ∈ B (Gd) ⊂ McbA (Gd) and the second part of Proposition 3.4 is ap-
plicable to characteristic functions of subgroups.
It is shown in Lemma
3.8 below that, when χHA (G) ⊂ Ad
cb (G), we need only require 1H to be
Acb (H)-approximable to deduce that χH is Acb (G)-approximable.
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
9
In [30], Lau and Ülger define a projection Ψ on V N (G) to be natural if
Ψ (λ (s)) = χE (s) λ (s) for some subset E of G. We may interpret Propo-
sition 3.4 as imposing restrictions on which subsets of G can arise from a
natural projection.
Example 3.5. Let G be a locally compact group and H a closed subgroup.
We show that an invariant projection Ψ : V N (G) → V NH (G) is natural.
It is clear that Ψ (λ (s)) = λ (s) for s ∈ H. Let s ∈ G \ H and let Tα ∈
spanλ (H) converge weak∗ to Ψ (λ (s)) in V N (G). If u ∈ A (G) with u (s) =
1 and uH = 0, then
hu · λ (s) , vi = hλ (s) , vui = v (s) u (s) = v (s) = hλ (s) , vi
(v ∈ A (G)) ,
so u · λ (s) = λ (s). If S =Pj αjλ (sj) ∈ spanλ (H), then
hu · S, vi =Pj αjv (sj) u (sj) = 0
so that u · S = 0. Thus 0 = u · Tα
and Ψ (λ (s)) = 0.
(v ∈ A (G)) ,
w∗
→ u · Ψ (λ (s)) = Ψ (u · λ (s)) = Ψ (λ (s))
Let G be a locally compact group. A bounded net (eα)α in A (G) or Acb (G)
is called a ∆-weak bounded approximate identity if it converges pointwise to
1G. This notion was introduced in [23] and shown in [30] to be closely
related to the existence of natural projections. Reasoning as in Example
3.2 shows that a bounded approximate identity for Acb (G) is a ∆-weak one,
so that Acb (G) has ∆-weak bounded approximate identity whenever G is
weakly amenable. The function 1G is Acb (G)-approximable when Acb (G)
has a ∆-weak bounded approximate identity.
The construction of the map τ above and the proof of Proposition 3.4 may
be carried out with McbA (G) replaced by B (G), but, to conclude that χH
is B (G)-approximable using this result, we require 1G to be in the weak∗
closure of A (G) in B (Gd). The proof of Theorem 4.6 shows that the canoni-
cal map A (G)∗∗ → Ad (G) extending inclusion A (G) ⊂ Ad (G) is surjective,
so that 1G is then the weak∗ limit of a bounded net, which is then a ∆-
weak bounded approximate identity for A (G), implying that G is already
amenable [25, Theorem 5.1]. It is the availability of ∆-weak bounded ap-
proximate identities in Acb (G) for a larger class of groups -- containing at
least the weakly amenable groups -- that is responsible for the utility of
Proposition 3.4. Whether the existence of a ∆-weak bounded approximate
identity for Acb (G) implies weak amenability of G appears to be an open
question.
For a locally compact group G and closed subgroup H, let
rH : McbA (Gd) → McbA (Hd) and eH : McbA (Hd) → McbA (Gd)
denote respectively the restriction map and the extension by zero. For a
◦
function f on H let
f denote its extension by zero to G. The restriction rH
is a complete quotient and extension eH a complete isometry ([38, Corollary
6.3] or [37, Proposition 4.1]), and it is clear that rHeH = idMcbA(Hd) and
eHrH = MχH , the multiplication by χH.
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
10
Lemma 3.6. Let G be a locally compact group and H a closed subgroup.
The maps rH and eH are weak∗ continuous.
Proof. If f =Pn
hr∗
H (f ) , mi =
j=1 αjδxj ∈ Cc (Hd), then
◦
f ∈ Cc (Gd) and
αjm (xj) =(cid:28) ◦
f , m(cid:29)
(m ∈ McbA (Gd)) ,
nXj=1
showing that r∗
follows that rH is weak∗ continuous.
H (Cc (Hd)) ⊂ Q (Gd). Since Cc (Hd) is dense in Q (Hd), it
Now if f =Pn
H (f ) , mi =D ◦
he∗
j=1 αjδxj ∈ Cc (Gd), then, for m ∈ McbA (Hd),
m, fE =
αjχH (xj) m (xj) =* nXj=1
nXj=1
αjχH (xj) δxj , m+ ,
j=1 αjχH (xj) δxj ∈ Cc (Hd). Therefore e∗
the claim follows by density, as above.
H (Cc (Gd)) ⊂ Q (Hd) and
(cid:3)
and so Pn
Lemma 3.7. Let G be a locally compact group and H a closed subgroup.
The restriction rH maps M d
cbA (H). In addition, the following
are equivalent:
cbA (G) into M d
cb (G).
(1) χH A (G) ⊂ Ad
(2) eH(cid:16)Ad
(3) Ad
cb (H)(cid:17) ⊂ Ad
cb (G) = IAd
cb (G).
cb(G) (G \ H) ⊕ IAd
cb(G) (H) (algebraic direct sum).
Proof. Since the restriction of a cb-multiplier of G to the closed subgroup
H yields a cb-multiplier of H [11, Proposition 1.12], the first claim follows
from weak∗ continuity of rH.
(1) implies (2): If χHA (G) ⊂ Ad
cb (G), then, because A (H) = rH (A (G)),
eH (A (H)) = eH (rH (A (G))) = χHA (G) ⊂ Ad
cb (G)
and (2) follows by weak∗ continuity of eH.
(2) implies (3): If eH(cid:16)Ad
cb (H)(cid:17) ⊂ Ad
weak∗ continuity of rH implies rH (m) ∈ Ad
eHrH (m) ∈ Ad
m = χHm + χG\H m ∈ IAd
have trivial intersection.
cb(G) (G \ H) + IAd
cb (G), whence χG\H m = m − χHm ∈ Ad
cb (G), then given m ∈ Ad
cb (G), the
cb (H) and it follows that χHm =
cb (G) and therefore
cb(G) (H). These ideals clearly
(3)implies (1): Write m ∈ A (G) as m = m1 + m2 for m1 ∈ IAd
and m2 ∈ IAd
cb(G) (H), in which case χH m = m1 ∈ Ad
cb (G).
cb(G) (G \ H)
(cid:3)
When the equivalent conditions of Lemma 3.7 hold, condition (2) implies
cb(G) (G \ H) and, because eH is isometric, condition
that eH(cid:16)Ad
cb (H)(cid:17) = IAd
(3) asserts that Ad
cb (G) = Ad
cb (H) ⊕ IAd
cb(G) (H).
Lemma 3.8. Let G be a locally compact group and H a closed subgroup.
If χHA (G) ⊂ Ad
cb (G) and 1H is Acb (H)-approximable, then χH is Acb (G)-
approximable.
Proof. It follows from Lemma 3.7(2) that χH = eH (1H) ∈ Ad
cb (G).
(cid:3)
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
11
Combining the results of this section, we obtain the following.
Theorem 3.9. The characteristic function of a closed subgroup H of a
locally compact group G is Acb (G)-approximable when either of the following
conditions is satisfied:
(1) Gd has the approximation property.
(2) There is a bounded map Ψ : V N (G) → V N (G) such that Ψ (λ (s)) =
χH (s) λ (s) for s ∈ G, which is satisfied if Ψ is a natural or invari-
ant projection onto V NH (G), and 1H is Acb (H)-approximable, which
occurs when H is weakly amenable or Hd has the approximation prop-
erty.
Proof. (1) When Gd has the approximation property, Proposition 3.3 as-
serts that Ad
cb (G) = McbA (Gd). Since McbA (Gd) contains the characteristic
functions of all subgroups of G, we have χH ∈ Ad
cb (G).
(2) If Ψ : V N (G) → V NH (G) is a bounded map such that Ψ (λ (s)) =
χH (s) λ (s) for s ∈ G and 1H is Acb (H)-approximable, then χHA (G) ⊂
Ad
cb (G) by Proposition 3.4. Lemma 3.8 then asserts that χH is Acb (G)-
approximable. Example 3.5 shows that invariant projections V N (G) →
V NH (G) are natural, and the latter are by definition bounded map satis-
fying the condition of (2). It was noted above that weak amenability of H
implies 1H ∈ Ad
cb (H) when Hd
has the approximation property.
(cid:3)
cb (H), while Proposition 3.3 implies 1H ∈ Ad
Theorem 3.7 of [1] claims that a bounded approximate indicator in B (G)
for a closed subgroup H of the locally compact group G exists whenever
χH is B (G)-approximable. The argument establishing this result was found
to contain an error [2] and it is not known whether the claim holds. The
following examples show that χH may be McbA (G)-approximable even when
no invariant projection onto V NH (G) exists, in which case no bounded
approximate indicator for H exists, either, by Proposition 5.1.
Example 3.10. The locally compact group G = SL (2, R) contains H = F2
as a closed subgroup. It has recently been shown that Gd is weakly amenable
[28], so that χH is Acb (G)-approximable by Proposition 3.3. Since G is con-
nected, its group von Neumann algebra is injective [6, Corollary 6.9(c)], so
there exists a completely bounded projection B (L2 (G)) → V N (G). If a
completely bounded projection V N (G) → V NH (G) existed, then composi-
tion would yield a completely bounded projection B (L2 (G)) → V NH (G),
implying that V NH (G) = V N (H) is an injective von Neumann algebra [5].
It would follow that the discrete group H is amenable [32, (2.35)], which is
false.
Example 3.11. Let G = SL (2, R) and consider the diagonal subgroup G∆
of G × G. The Fourier algebra A (G) is not operator biflat by Corollary 3.7
of [9], meaning that no invariant projection V N (G × G) → V NG∆ (G × G)
exists. But the weak amenability of Gd, noted in the preceding example,
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
12
implies the weak amenability of (G × G)d, so that χG∆ is Acb (G × G)-
approximable, again by Proposition 3.3.
4. The discretized H-separation property
In this section, we characterize the approximability of the characteristic
function of a closed subgroup H of a locally compact group G in the spirit
of the H-separation property of Kaniuth and Lau. For a closed subgroup H
of G, let PH (G) denote the norm closed convex set {u ∈ P (G) : u (H) = 1}.
Definition 4.1. ([24]) A locally compact group G is said to have the H-
separation property for a closed subgroup H if, for each s ∈ G \ H, there
exists u ∈ PH (G) such that u (s) 6= 1.
It is routine to verify that G has the H-separation property for any open,
compact, or normal subgroup H, and it was shown by Forrest [16] that if
G is a SIN group, then G has the H-separation property for every closed
subgroup H. In [24], a fixed point argument is used to show that an invariant
projection V N (G) → V NH (G) exists when the locally compact group G
has the H-separation property (it is noted in Proposition 5.1 below that the
projections arising this way are completely positive, in particular completely
bounded). In fact, the following stronger result holds.
Proposition 4.2. Let G be a locally compact group and H a closed subgroup.
Then G has the H-separation property if and only if there exists a bounded
approximate indicator for H in PH (G).
Proof. Suppose that G has the H-separation property. The proof of [24,
Proposition 3.1] constructs an invariant projection P : V N (G) → V NH (G)
that is the weak∗ operator topology limit of a net (Muα)α, where uα ∈ PH (G)
and Muα : V N (G) → V N (G) is the adjoint of the multiplication map
u 7→ uαu on A (G). Let rH : A (G) → A (H) be the restriction map, which
is a surjection satisfying r∗
H (V N (H)) ⊂ V NH (G). Given u ∈ A (H), let
u ∈ A (G) with rH (u) = u, so that for T ∈ V N (H),
hT, urH (uα)i = hT, rH (uuα)i
H (T )) , ui
H (T )) , ui
= hMuα (r∗
→ hP (r∗
= hr∗
= hT, ui .
H (T ) , ui
If w ∈ IA(G) (H), then
hT, wuαi = hMuα (T ) , wi → hP (T ) , wi = 0
(T ∈ V N (G))
since P (T ) ∈ V NH (G) = IA(G) (H)⊥. Therefore urH (uα) → u weakly in
A (H) for all u ∈ A (H) and wuα → 0 weakly in A (G) for all w ∈ IA(G) (H).
Passing to convex combinations yields a bounded approximate indicator for
H which remains in the convex set PH (G).
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
13
Conversely, if (uα)α is a bounded approximate indicator for H in PH (G),
then, given s ∈ G \ H, choose w ∈ IA(G) (H) with w (s) = 1, in which case
uα (s) = uα (s) w (s) ≤ kuαwkL∞(G) ≤ kuαwkA(G) → 0
implies uα (s) 6= 1 eventually.
(cid:3)
For a closed subgroup H of a locally compact group G, we now show that
a weaker form of the H-separation property, replacing the algebra B (G)
with Bd (G), characterizes when χH is B (G)-approximable.
Definition 4.3. Let G be a locally compact group and H a closed subgroup.
The group G is said to have the discretized H-separation property if, for any
s ∈ G \ H, there exists u ∈ Bd (G) ∩ PH (Gd) such that u (s) 6= 1.
Proposition 4.4. Let G be a locally compact group and H a closed subgroup.
Then G has the discretized H-separation property if and only if χH is B (G)-
approximable.
s kB(Gd) = un
sH = 1 and u0
s (e) = 1 and thus has a weak∗ cluster point u0
Proof. Suppose that G has the discretized H-separation property and for
each s ∈ G \ H let us ∈ Bd (G) ∩ PH (Gd) with us (s) 6= 1. Replac-
ing us by 1
2 (1G + us), which remains in Bd (G) ∩ PH (Gd), we may assume
s )n≥1 is in Bd (G) ∩ PH (Gd) with
that us (s) < 1. Then the sequence (un
kun
s in the unit
s (s) ≤ lim supn us (s)n = 0, so
ball of Bd (G). Then u0
u0
s (s) = 0. Let F be the collection of finite subsets of G and for each
s. Ordering F by inclusion, we have uF H = 1 and
ptw
→ χH and by bounded-
w∗
→ χH in B (Gd), whence χH ∈ Bd (G). The converse is clear, given
(cid:3)
uF (s) = 0 eventually for each s ∈ G \ H, so that uF
ness uF
that the characteristic function of a subgroup is always in PH (Gd).
F ∈ F let uF = Qs∈F u0
When the locally compact group G is second countable, the H-separation
property may also be characterized in terms of a single function on G.
Theorem 4.5. Let G be a second countable locally compact group. For a
closed subgroup H, the following are equivalent:
(1) G has the H-separation property.
(2) There is u ∈ B (G) of norm one with {s ∈ G : u (s) = 1} = H.
(3) There is u ∈ P (G) with {s ∈ G : u (s) = 1} = H.
Proof. (1) implies (2): For s ∈ G \ H, let us ∈ PH (G) with us (s) 6= 1 and
choose an open neighborhood Us of s with 1 /∈ us (Us). Then (Us)s∈G\H is an
open cover of G \ H, so has a countable subcover (Usn)n≥1 by σ-compactness
closed convex set PH (G), so that kukB(G) = u (e) = 1. Given s ∈ G \ H,
choose n such that s ∈ Usn, so usn (s) 6= 1. Since kusnkL∞(G) = 1, we have
of the open set G \ H in G. The function u = Pn≥1 2−nusn is in the norm
Reusn (s) < 1, implying that Reu (s) = Pn≥1 2−nReusn (s) < 1 and hence
that u (s) 6= 1.
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
14
(2) implies (3): Replacing u by 1
2 (u + u∗), where u∗ (s) = u (s−1), we
obtain function of B (G)-norm one (because the B (G) norm dominates the
L∞ (G) norm) for which we may write u = u+ − u− with u± ∈ P (G)
satisfying kukB(G) = ku+kB(G) + ku−kB(G). Then
u+ (e)−u− (e) = u (e) = 1 = kukB(G) =(cid:13)(cid:13)(cid:13)u+(cid:13)(cid:13)(cid:13)B(G)
and consequently ku−kB(G) = u− (e) = 0, so that u = u+ ∈ PH (G).
+(cid:13)(cid:13)(cid:13)u−(cid:13)(cid:13)(cid:13)B(G)
= u+ (e)+u− (e)
(3) implies (1): This is clear.
(cid:3)
The amenability of H is known to imply the existence of an invariant
projection V N (G) → V NH (G) for any locally compact group G containing
H as a closed subgroup [8, Corollary 3.7] (see also [10]). We now show that
the former condition may be characterized in terms of a separation property
relative to any such G.
Theorem 4.6. A locally compact group H is amenable if and only if χH
is A (G)-approximable for some (equivalently, any) locally compact group G
containing H as a closed subgroup.
Proof. Fix a locally compact group G that contains H as a closed subgroup.
Suppose that H is amenable. Let (eα)α be a bounded approximate identity
for A (H) and let Ψ : V N (G) → V NH (G) be an an invariant projection.
Example 3.5 shows that Ψ (λG (s)) = χH (s) λG (s) for all s ∈ G and the
ptw
→ 1H. Recall that the adjoint of
argument of Example 3.2 shows that eα
the restriction rH : A (G) → A (H) is a ∗-isomorphism τ = r∗
H : V N (H) →
V NH (G) taking λH (s) to λG (s) for all s ∈ H. The composition
V N (H)∗ (τ ∗)−1
→ V NH (G)∗ Ψ∗
→ V N (G)∗
satisfies, for s ∈ G,
DΨ∗ (τ ∗)−1 (eα) , λG (s)E = D(τ ∗)−1 (eα) , χH (s) λG (s)E
= Deα, τ −1 (χH (s) λG (s))E
=
heα, λH (s)i ,
0,
s ∈ H
s /∈ H
→ χH (s) .
Let E be a weak∗ cluster point of the bounded net (cid:16)Ψ∗ (τ ∗)−1 (eα)(cid:17)α
in
V N (G)∗, so that hE, λG (s)i = χH (s) for all s ∈ G by the above computa-
tion. Letting (uα)α be a bounded net in A (G) converging weak∗ to E, we
have uα
w∗
→ χH in B (Gd) by boundedness.
Conversely, suppose that (uα)α is a net in A (G) such that uα
w∗
→ χH in
B (Gd). Analogous to the proof of Lemma 3.6, the restriction rH : B (Gd) →
B (Hd) is weak∗ continuous, so that rH (uα) w∗
→ 1H and 1H is in the weak∗ clo-
sure of A (H) in B (Hd). Viewing λH as a representation of Hd, the universal
ptw
→ χH and therefore uα
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
15
property of C ∗ (Hd) yields a quotient ∗-homomorphism C ∗ (Hd) → C ∗
δ (H) :
ωHd (s) 7→ λH (s), where ωHd denotes the universal representation of Hd and
C ∗
δ (H) the C ∗-algebra generated by λH (H) in B (L2 (H)), a subalgebra
of V N (H). Composing with the inclusion, we obtain a ∗-homomorphism
Ψ : C ∗ (Hd) → V N (H). The adjoint Ψ∗ : A (H)∗∗ → B (Hd) is the
weak∗ continuous extension of the inclusion A (H) ⊂ B (Hd), so that
ker Ψ = A (H)⊥ (preannihilator taken with respect to the B (Hd) -- C ∗ (Hd)
duality). Then A (H)⊥ = Ad (H)⊥ is an ideal in C ∗ (Hd) and Ψ drops to
an injective ∗-homomorphism C ∗ (Hd) /Ad (H)⊥ → V N (H). This injec-
tive ∗-homomorphism is isometric, so that its adjoint A (H)∗∗ → Ad (H)
is a quotient map and given ǫ > 0, the function 1H is the weak∗ limit in
B (Hd) of a net (vα)α in A (H) of bound 1 + ǫ. The adjoint on B (Hd) is
a weak∗ continuous isometry preserving A (H), so that we may replace vα
by 1
α ∈ P (H) ∩ A (H) with
α kB(H). Passing to subnets, let v± be weak∗ limits
kvαkB(H) = kv+
of (v±
α) and assume that vα = v+
α kB(H) + kv−
2 (vα + v∗
α )α in B (Hd). Then
α − v−
α for v±
1 = lim
α
vα (e) = lim
α
v+
α (e) − lim
α
v−
α (e) = v+ (e) − v− (e)
and so
v− (e) = v+ (e) − 1 = lim
α
Let ωǫ be a weak∗ cluster point of (v+
V N (H) satisfying, for s ∈ H,
v+
α (e) − 1 = lim
− 1 ≤ ǫ.
α )α in A (H)∗∗, so that ωǫ is a state on
α(cid:13)(cid:13)(cid:13)B(H)
α (cid:13)(cid:13)(cid:13)v+
vα (s)(cid:12)(cid:12)(cid:12)(cid:12)
α (e) = v− (e) ≤ ǫ.
v−
α
hωǫ, λH (s)i − 1 = (cid:12)(cid:12)(cid:12)(cid:12)lim
α (s)(cid:12)(cid:12)(cid:12) ≤ lim
α (cid:12)(cid:12)(cid:12)v−
v+
α (s) − lim
α
= lim
α
Letting ω be a weak∗ cluster point of the states (ωǫ)ǫ>0 on V N (H), we have
hω, λH (s)i = 1 for all s ∈ H, and any extension of ω to a state on B (L2 (H))
still takes the value 1 on the unitaries λH (s) for s ∈ H. The amenability of
H follows: by a Cauchy-Schwarz argument, any extension of ω to a state on
B (L2 (H)) is invariant under the conjugation action of the unitaries λH (s)
for s ∈ H, whence λH is an amenable representation of H (see [3]).
(cid:3)
Corollary 4.7. Let G be a locally compact group and H a closed subgroup.
Then H is amenable if and only if G has the discretized H-separation prop-
erty witnessed by functions in Ad (G) ∩ PH (Gd).
Proof. In the argument establishing Proposition 4.4, substituting 1
for the function 1
if and only if χH is A (G)-approximable.
s + us)
2 (1G + us) yields a proof that G has the desired property
(cid:3)
2 (u2
5. Invariant projections and bounded approximate indicators
In this section we establish some consequences of the existence of a
bounded approximate indicator for a closed subgroup of a locally compact
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
16
group. We first provide the well known argument that this stronger separa-
tion property indeed yields invariant projections. For a commutative com-
pletely contractive Banach algebra A, let CBA (A∗) denote the completely
bounded A-bimodule maps on A∗. This space has compact unit ball when
given the weak∗ operator topology, which is determined by the seminorms
Ψ 7→ hΨ (ϕ) , ai
(ϕ ∈ A∗, a ∈ A) .
Proposition 5.1. Let G be a locally compact group and H a closed subgroup.
If there is a bounded approximate indicator for H, then there is a completely
bounded invariant projection V N (G) → V NH (G). If there is a bounded
approximate indicator for H consisting of positive definite functions, then
there is a completely positive invariant projection V N (G) → V NH (G).
Proof. Let (mα)α a bounded approximate indicator for H, so that the net of
multiplication maps (Mmα)α in CBA(G) (V N (G)) is then bounded and thus
has a weak∗ operator topology cluster point Ψ ∈ CB (V N (G)). Passing to
a subnet if necessary, we may assume that Ψ is the limit of this net. For
u, v ∈ A (G) and T ∈ V N (G),
hΨ (v · T ) , ui = lim
α
hT, mαuvi = hΨ (T ) , uvi = hv · Ψ (T ) , ui ,
showing that Ψ is invariant. Given T ∈ V NH (G), so that T = r∗
some S ∈ V N (H), we have for u ∈ A (G) that
H (S) for
hΨ (T ) , ui = lim
α
= lim
α
hS, rH (uuα)i
hS, rH (u) rH (uα)i
= hS, rH (u)i
= hT, ui ,
whence Ψ is the identity on V NH (G). For T ∈ V N (G) and u ∈ IA(G) (H)
we have hΨ (T ) , ui = limα hS, uuαi = 0, so that Ψ maps into IA(G) (H)⊥ =
V NH (G) and is thus a projection onto V NH (G).
If the functions mα are in P (G), then the maps Mmα are completely
positive [11, Proposition 4.2] and by [33, Theorem 7.4] their weak∗ operator
topology cluster point Ψ is also completely positive.
(cid:3)
From the preceding we obtain an analogous result for Acb (G), at least
when G is a weakly amenable locally compact group.
Proposition 5.2. Let G be a weakly amenable locally compact group and H
a closed subgroup. If there is a bounded approximate indicator for H, then
there is a completely bounded invariant projection Acb (G)∗ → IAcb(G) (H)⊥.
Proof. Let (mα)α an approximate indicator for H. Since A (G) is an ideal
in McbA (G), so too is its closure Acb (G), so that multiplication by mα is a
completely bounded Acb (G)-module map on Acb (G). Denote its adjoint by
Mmα. Passing to a subnet, we may assume that (Mmα)α has a weak∗ operator
topology limit Ψ ∈ CB (Acb (G)∗), and passing to a further subnet we may
assume that the net of maps (Mmα)α in CBA(G) (V N (G)) also has a weak∗
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
17
operator topology limit ΨA, which is an invariant projection V N (G) →
V NH (G) by the argument of Proposition 5.1. For u, v ∈ Acb (G) and T ∈
Acb (G)∗, we have
hΨ (v · T ) , ui = lim
α
hT, mαuvi = hΨ (T ) , uvi = hv · Ψ (T ) , ui ,
so Ψ is invariant. Let ι : A (G) → Acb (G) be the inclusion. If T ∈ Acb (G)∗
and u ∈ A (G), then
hΨAι∗ (T ) , ui = lim
α
= lim
α
hι∗ (T ) , umαi
hT, ι (u) mαi
= hΨ (T ) , ι (u)i
= hι∗Ψ (T ) , ui
and ΨAι∗ = ι∗Ψ by density of A (G) in Acb (G). It follows that
ι∗Ψ2 = ΨAι∗Ψ = Ψ2
which, together with injectivity of ι∗,
IAcb(G) (H)⊥, then ι∗ (T ) ∈ IA(G) (H)⊥ and
Aι∗ = ΨAι∗ = ι∗Ψ,
implies that Ψ2 = Ψ.
If T ∈
hΨ (T ) , ι (u)i = hΨAι∗ (T ) , ui = hι∗ (T ) , ui = hT, ι (u)i
(u ∈ A (G)) ,
whence Ψ (T ) = T , again by density of A (G). Therefore IAcb(G) (H)⊥
is contained in the range of Ψ. Finally, for any T ∈ Acb (G)∗,
if u ∈
IA(G) (H), then hΨ (T ) , ι (u)i = hΨAι∗ (T ) , ui = 0, so Ψ (T ) ∈ IA(G) (H)⊥ =
(cid:18)IA(G) (H)Acb(G)(cid:19)⊥
. That Acb (G) has bounded approximate identity im-
plies every set of synthesis for A (G) is one for Acb (G) [18, Proposition
3.1] and, because compactly supported functions in Acb (G) are in A (G),
that H is of spectral synthesis for Acb (G) is exactly the assertion that
IA(G) (H)Acb(G) = IAcb(G) (H). Therefore Ψ has range in IAcb(G) (H)⊥.
(cid:3)
Note that the arguments of the preceding two propositions yield projec-
tions of completely bounded norm at most the bound on an approximate
indicator for the subgroup. Let ΛG denotes the Cowling -- Haagerup constant,
that is, the infimum of bounds on approximate identities for Acb (G).
Corollary 5.3. Let G be a weakly amenable locally compact group and H a
closed subgroup for which an approximate indicator of bound C exists. The
following hold:
(1) IAcb(G) (H) has an approximate identity of bound (1 + C) ΛG.
(2) An approximate indicator for H of bound 1 + (1 + C) ΛG exists that
takes the value one on H.
(3) An approximate indicator for H of bound CΛG exists in Acb (G).
Proof. (1) The argument of Proposition 5.2 yields an invariant projection
Acb (G)∗ → IAcb(G) (H)⊥ of norm at most C and, because the Banach alge-
bra Acb (G) has an approximate identity of bound ΛG, it follows from [15,
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
18
Proposition 6.4] and its proof that the ideal IAcb(G) (H) has an approximate
identity of bound (1 + C) ΛG.
(2) If (eα)α is an approximate identity for IAcb(G) (H) of bound (1 + C) ΛG,
then (1G − eα)α is an approximate indicator for H with the claimed norm
bound.
(3) Let
(eα)α∈A be a bounded approximate identity for Acb (G),
let (mβ)β∈B a bounded approximate indicator for H, and for γ =
(cid:16)β, (αβ ′)β ′∈B(cid:17) ∈ B × AB set uγ = mβeαβ , which is in the ideal Acb (G)
of McbA (G). Giving B × AB the product order,
w ∈ IA(G) (H) we have the norm limits
for u ∈ A (H) and
lim
γ∈B×AB
rH (uγ) u = lim
β
lim
α
rH (mβeα) u = lim
β
rH (mβ) u = u
and
lim
γ∈B×AB
uγw = lim
β
lim
α
mβeαw = lim
β
mβw = 0
by [26, p. 69], hence (uγ)γ∈B×AB is a bounded approximate indicator for H
of norm bound supα keαkMcbA(G) supβ kmβkMcbA(G).
(cid:3)
6. Convergence of cb-multipliers and averaging over closed
subgroups
Fix a locally compact group G and a closed subgroup H. It is folklore
that the convergence properties of nets of cb-multipliers can be improved by
convolving them with probability measures in Cc (G). For example, Knudby
recently recorded the following, the second part of which originates in an
argument of Cowling and Haagerup [7, Proposition 1.1]. If a net (mα)α of
functions on a topological space converges uniformly on compact sets to a
function m, we write mα
Theorem 6.1. ([27, Lemma B.2]) Let (mα)α be a bounded net in McbA (G),
m ∈ McbA (G), and let f ∈ Cc (G) be such that f ≥ 0 and RG f = 1.
Convolution on the left with f is a contraction on McbA (G) and the following
hold:
ucs→ m.
(1) If mα
(2) If mα
A (G).
w∗
→ m in McbA (G), then f ∗ mα
ucs→ m, then k(f ∗ mα) u − (f ∗ m) ukA(G) → 0 for all u ∈
ucs→ f ∗ m.
In this section, we develop an analogue of the convolution technique rel-
ative to a closed subgroup. Fix a function f ∈ Cc (H) such that f ≥ 0 and
RH f = 1. For any function f on G and s, t ∈ G, let sf (t) = f (st).
Definition 6.2. For u ∈ Cb (G), define a function Ωf (u) on G by the formula
We will show that Ωf defines a bounded map on McbA (G). For a Hilbert
space H, let Cc (G, H) and Cb (G, H) denote the continuous functions G → H
that are compactly supported and bounded, respectively.
Ωf (u) (s) =ZH
f (h) u(cid:16)h−1s(cid:17) dh
(s ∈ G) .
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
19
If u ∈ Cc (G, H) then for
Lemma 6.3. Let H be a Hilbert space.
any ǫ > 0 there is an open neighborhood U of the identity e such that
supt∈G ku (st) − u (t)k < ǫ for all s ∈ U.
Proof. The standard proof in the case that H = C, for example [14, Propo-
sition 2.6], works for any Hilbert space.
(cid:3)
If u ∈ Cb (G, H), s0 ∈ G,
Lemma 6.4. Let H be a Hilbert space.
and ǫ > 0, then there is an open neighborhood U of s0 in G such that
suph∈H kf (h) u (sh) − f (h) u (s0h)k < ǫ for all s ∈ U.
Proof. If u = 0, then the claim trivially holds, so assume u 6= 0. Since H
is closed in G, the function f extends to a continuous compactly supported
function f ′ on G. Assume that s0 = e. Since f ′u is compactly supported,
Lemma 6.3 yields an open neighborhood U of e such that
kf ′u (st) − f ′u (t)k <
sup
t∈G
ǫ
2
and sup
t∈G
f ′ (st) − f ′ (t) <
ǫ
2 kuk∞
for all s ∈ U. Then
sup
h∈H
kf (h) u (sh) − f (h) u (h)k ≤ sup
t∈G
≤ sup
t∈G
kf ′ (t) u (st) − f ′ (t) u (t)k
(kf ′ (t) u (st) − f ′u (st)k +
kf ′u (st) − f ′ (t) u (t)k)
ǫ
f ′ (st) − f ′ (t) +
2
< kuk∞ sup
t∈G
< ǫ,
for all s ∈ U. For s0 6= e, the above argument with u replaced by s0u yields
a neighborhood U of e and s0U is then the desired neighborhood of s0. (cid:3)
Proposition 6.5. If u ∈ McbA (G),
kΩf (u)kMcbA(G) ≤ kukMcbA(G) and rHΩf (u) = f ∗ rH (u).
then Ωf (u) ∈ McbA (G) with
Proof. Let u ∈ McbA (G) and apply Gilbert's theorem to obtain a Hilbert
space H and functions P, Q ∈ Cb (G, H) such that u (s−1t) = hP (t) Q (s)i
for all s, t ∈ G. Then, for s, t ∈ G,
f (h) Q (sh) dh(cid:29).
Ωf (u)(cid:16)s−1t(cid:17) =ZH
We show that q (s) =RH f (h) Q (sh) dh defines a bounded continuous func-
f (h) u(cid:16)h−1s−1t(cid:17) dh =(cid:28)P (t)(cid:12)(cid:12)(cid:12)(cid:12)ZH
tion on G, from which it will follow that Ωf (u) is in McbA (G), again by
Gilbert's theorem. Define Q′ : G → L1 (H, H) by Q′ (s) = f (sQ), which
maps into L1 (H, H) since f has compact support. Set K = suppf and let
K denote the Haar measure of K, which is nonzero and finite by continuity
of the nonzero, compactly supported function f . Given s0 ∈ G and ǫ > 0,
Lemma 6.4 yields an open neighborhood U of s0 in G such that
kQ′ (s) − Q′ (s0)kL∞(H,H) = sup
h∈H
kf (h) Q (sh) − f (h) Q (s0h)k <
ǫ
K
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
20
for all s ∈ U. Since Q′ (s) is supported in K for every s ∈ G, it follows that
kQ′ (s) − Q′ (s0)kL1(H,H) = kχK (Q′ (s) − Q′ (s0))kL1(H,H)
≤ kχKkL1(H,H) kQ′ (s) − Q′ (s0)kL∞(H,H)
< ǫ
for all s ∈ U. Thus Q′ is continuous and so too is q, the latter being the
Using that f is nonnegative with mass one, if s ∈ G, then kq (s)k ≤
composition of Q′ with the bounded map L1 (H, H) → H : g 7→RH g.
RH f (h) kQ (sh)k dh ≤ kQk∞, so q is bounded with kqk∞ ≤ kQk∞.
By the norm characterization of Gilbert's theorem, kΩf (u)kMcbA(G) ≤
kP k∞ kqk∞ ≤ kP k∞ kQk∞ and since P , Q, and H are an arbitrary rep-
resentation of u, we conclude that kΩf (u)kMcbA(G) ≤ kukMcbA(G). Finally,
we have for s ∈ H that
rHΩf (u) (s) = ZH
= ZH
f (h) u(cid:16)h−1s(cid:17) dh
f (h) rH (u)(cid:16)h−1s(cid:17) dh
= f ∗ rh (u) (s) .
(cid:3)
Theorem 6.6. Let (mα)α be a bounded net in McbA (G) and let m ∈
McbA (H). The following hold:
(1) If rH (mα) w∗
(2) If rH (mα) ucs→ m, then krHΩf (mα) u − (f ∗ m) ukA(H) → 0 for all
→ m in McbA (H), then rHΩf (mα) ucs→ f ∗ m.
u ∈ A (H).
Proof. These follow immediately from Theorem 6.1 and Proposition 6.5. (cid:3)
In our applications, the preceding theorem will be applied with m = 1H,
which is fixed under convolution with f on the left. We list some additional
properties that the map Ωf enjoys.
(1) An argument very similar to that establishing Proposition 6.5 shows
that Ωf (u) is bounded and continuous for any bounded continuous
function u on G.
(2) If u = hπ (·) ξηi is a coefficient of the unitary representation π of G,
then
Ωf (u) (s) = ZH
f (h) hπ (s) ξπ (h) ηi dh
so Ωf (u)
∈
= (cid:28)π (s) ξ(cid:12)(cid:12)(cid:12)(cid:12)(cid:18)ZH
B (G),
and
f (h) π (h) dh(cid:19) η(cid:29) ,
from kRH f (h) π (h) dhk
RH f (h) kπ (h)k dh = 1 it follows that kΩf (u)kB(G) ≤ kukB(G).
Thus Ωf restricts to a contraction on B (G) and moreover restricts
to a contraction on A (G), since Ωf (u) is a coefficient of the same
representation as u.
≤
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
21
(3) An argument similar to that establishing the weak∗ continuity of the
map Φf in the proof of [21, Lemma 1.16] shows that Ωf is weak∗
continuous on McbA (G) with preadjoint mapping g ∈ L1 (G) to the
L1 (G) function s 7→RH f (h) g (hs) dh.
Say that a net (mα)α of functions on a topological space X converges locally
le→ 0 if for any compact subset
eventually to zero on A ⊂ X and write mα
K of A there is an index α0 such that mαK = 0 for all α ≥ α0.
Proposition 6.7. If (mα)α is a bounded net in McbA (G) and m′
Ωf (mα), then the net (m′
α)α has the same norm bound as (mα)α and
α =
(1) if rH (mα) ucs→ 1H, then ku · rH (m′
α) − ukA(H) → 0 for all u ∈ A (H),
and
(2) if mα
le→ 0 on G \ H, then m′
α
le→ 0 on G \ H.
If the bounded net (mα)α satisfies the hypotheses of both (1) and (2), then
(m′
α)α is a bounded approximate indicator for H.
Proof. The claim regarding norm bounds holds since the map Ωf of Section
6 is a contraction on McbA (G).
(1) If rH (mα) ucs→ 1H, then, since restriction is a contraction from McbA (G)
into McbA (H), the net (rH (mα))α is bounded and (1) follows from Theorem
6.6.
(2) Suppose that mα
le→ 0 on G \ H. Let K ⊂ G \ H be compact
and choose α0 such that α ≥ α0 implies mα = 0 on the compact set
(supp (f ))−1 K. For α ≥ α0, if s ∈ K and h ∈ H, then f (h) mα (h−1s) = 0
since either h /∈ supp (f ) or h−1s ∈ (supp (f ))−1 K, implying that m′
α (s) =
RH f (h) mα (h−1s) dh = 0. Therefore m′
αu = 0 eventually by (2), so certainly kum′
If (mα)α satisfies the hypotheses of both (1) and (2), then (m′
α)α satisfies
the first condition of Definition 2.2. If u ∈ IA(G) (H) has compact support,
then m′
αkA(G) → 0. Using that
H is of spectral synthesis in A (G), if u ∈ IA(G) (H) is arbitrary, then given
ǫ > 0 choose u0 ∈ IA(G) (H) of compact support with ku − u0kA(G) < ǫ. For
sufficiently large α,
α = 0 on K, for all α ≥ α0.
kum′
αkA(G) ≤ ku0m′
αkA(G) + ku0m′
α − um′
αkA(G) < ǫ km′
αkMcbA(G) ,
and thus kum′
αkA(G) → 0 by boundedness of (m′
α)α.
(cid:3)
Proposition 6.7 allows one to obtain approximate indicators consisting of
cb-multipliers by verifying the same conditions that yielded approximate
indicators in [1].
References
[1] O. Y. Aristov, V. Runde, N. Spronk, Operator biflatness of the Fourier algebra and
approximate indicators for subgroups. J. Funct. Anal. 209(2) (2004), 367 -- 387.
[2] O. Y. Aristov, V. Runde, N. Spronk, Z. Tanko, Corrigendum to: "Operator biflatness
of the Fourier algebra and approximate indicators for subgroups" [J. Funct. Anal.
209(2) (2004) 367 -- 387]. J. Funct. Anal. 270(6) (2016), 2381 -- 2382.
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
22
[3] M. B. Bekka, Amenable unitary representations of locally compact groups. Invent.
Math. 100(2) (1990), 383 -- 401.
[4] M. B. Bekka, E. Kaniuth, A. T. Lau, G. Schlichting, On C ∗-algebras associated with
locally compact groups. Proc. Amer. Math. Soc. 124(10) (1996), 3151 -- 3158.
[5] E. Christensen, A. M. Sinclair, On von Neumann algebras which are complemented
subspaces of B (H). J. Funct. Anal. 122 (1994), 91 -- 102.
[6] A. Connes, Classification of injective factors. Cases II1, II∞, III∞, λ 6= 1. Ann. of
Math. (2) 104(1) (1976), 73 -- 115.
[7] M. Cowling, U. Haagerup, Completely bounded multipliers of the Fourier algebra of
a simple Lie group of real rank one. Invent. Math. 96(3) (1989), 507 -- 549.
[8] J. Crann, On hereditary properties of quantum group amenability. arXiv:1603.03842
[math.OA]
[9] J. Crann, Z. Tanko, On the operator homology of the Fourier algebra and its cb-
multiplier completion. arXiv:1602.05259 [math.OA]
[10] A. Derighetti, Conditional expectations on CVp (G). Applications. J. Funct. Anal.
247(1) (2007), 231 -- 251.
[11] J. de Cannière, U. Haagerup, Multipliers of the Fourier algebras of some simple Lie
groups and their discrete subgroups. Amer. J. Math. 107(2) (1985), 455 -- 500.
[12] E. G. Effros, Z.-J. Ruan, Operator Spaces. London Mathematical Society Mono-
graphs, New Series 23, Oxford University Press, New York, 2000.
[13] P. Eymard, L'algèbre de Fourier d'un groupe localement compact. Bull. Soc. Math.
France 92(1964), 181 -- 236.
[14] G. B. Folland, A course in abstract harmonic analysis. Studies in Advanced Mathe-
matics. CRC Press, Boca Raton, Florida, 1995.
[15] B. Forrest, Amenability and bounded approximate identities in ideals of A (G). Illinois
J. Math. 34(1) (1990), 1 -- 25.
[16] B. Forrest, Amenability and ideals in A (G). J. Austral. Math. Soc. Ser. A 53(2)
(1992), 143 -- 155.
[17] B. Forrest, E. Kaniuth, A. T. Lau, N. Spronk, Ideals with bounded approximate
identities in Fourier algebras. J. Funct. Anal. 203(1) (2003), 286 -- 304.
[18] B. Forrest, V. Runde, N. Spronk, Operator amenability of the Fourier algebra in the
cb-multiplier norm. Canad. J. Math. 59(5) (2007), 966 -- 980.
[19] E. E. Granirer, M. Leinert, On some topologies which coincide on the unit sphere of
the Fourier -- Stieltjes algebra B(G) and of the measure algebra M (G). Rocky Moun-
tain J. Math. 11(1981), 459 -- 472.
[20] C. Herz, Harmonic synthesis for subgroups. Ann. Inst. Fourier (Grenoble) 23(3)
(1973), 91 -- 123.
[21] U. Haagerup, J. Kraus, Approximation properties for group C ∗-algebras and group
von Neumann algebras. Trans. Amer. Math. Soc. 344(2) (1994), 667 -- 699.
[22] P. Jolissaint, A characterization of completely bounded multipliers of Fourier algebras.
Colloq. Math. 63(2) (1992), 311 -- 313.
[23] C. A. Jones, C. D. Lahr, Weak and norm approximate identities are different, Pacific
J. Math. 72 (1977), no. 1, 99 -- 104.
[24] E. Kaniuth, A. T. Lau, A separation property of positive definite functions on locally
compact groups and applications to Fourier algebras. J. Funct. Anal. 175(1) (2000),
89 -- 110.
[25] E. Kaniuth, A. Ülger, The Bochner-Schoenberg-Eberlein property for commutative
Banach algebras, especially Fourier and Fourier-Stieltjes algebras. Trans. Amer.
Math. Soc. 362(8) (2010), 4331 -- 4356.
[26] J. L. Kelley, General Topology. Graduate Textbooks
in Mathematics 27,
Springer -- Verlag, New York -- Berlin -- Heidelberg, 1955.
[27] S. Knudby, The weak Haagerup property. Trans. Amer. Math. Soc. 368(5) (2016),
3469 -- 3508.
WEAK SEPARATION PROPERTIES FOR CLOSED SUBGROUPS
23
[28] S. Knudby, K. Li, Approximation properties of simple Lie groups made discrete. J.
Lie Theory 25(4) (2015), 985 -- 1001.
[29] H. Leptin, Sur l'algèbre de Fourier d'un groupe localement compact. C. R. Acad. Sci.
Paris Sér. A-B 266 (1968), A1180 -- A1182.
[30] A. T. Lau, A. Ülger, Characterization of closed ideals with bounded approximate
identities in commutative Banach algebras, complemented subspaces of the group
von Neumann algebras and applications. Trans. Amer. Math. Soc. 366(8) (2014),
4151 -- 4171.
[31] V. Losert, Properties of the Fourier algebra that are equivalent to amenability. Proc.
Amer. Math. Soc. 92(3) (1984), 347 -- 354.
[32] A. L. T. Paterson, Amenability. Mathematical Surveys and Monographs, 29. Amer-
ican Mathematical Society, Providence, RI, 1988.
[33] V. Paulsen, Completely bounded maps and operator algebras. Cambridge Studies in
Advanced Mathematics, 78. Cambridge University Press, Cambridge, 2002.
[34] Z.-J. Ruan, The operator amenability of A (G). Amer. J. Math. 117 (1995),
1449 -- 1474.
[35] V. Runde, Amenability for dual Banach algebras. Studia Math. 148 (2001), 47 -- 66.
[36] V. Runde, Lectures on amenability. Lecture Notes in Mathematics, 1774. Springer-
Verlag, Berlin, 2002.
[37] A. P. Stan, On idempotents of completely bounded multipliers of the Fourier algebra
A (G). Indiana Univ. Math. J. 58(2) (2009), 523 -- 535.
[38] N. Spronk, Measurable Schur multipliers and completely bounded multipliers of the
Fourier algebras. Proc. London Math. Soc. (3) 89(1) (2004), 161 -- 192.
[39] M. Takesaki, N. Tatsuuma, Duality and subgroups. Ann. of Math. 93(2) (1971)
344 -- 364.
Department of Pure Mathematics, University of Waterloo, Waterloo,
Ontario, N2L 3G1, Canada
E-mail address: [email protected]
|
1809.06583 | 1 | 1809 | 2018-09-18T08:25:57 | Carleson measures and Toeplitz operators on small Bergman spaces on the ball | [
"math.FA"
] | We study the Carleson measures and the Toeplitz operators on the class of so-called small weighted Bergman spaces, introduced recently by Seip. A characterization of Carleson measures is obtained which extends Seip's results from the unit disc of $\mathbb C$ to the unit ball of $\mathbb C^n$. We use this characterization to give necessary and sufficient conditions for the boundedness and compactness of Toeplitz operators. Finally, we study the Schatten $p$ classes membership of Toeplitz operators for $1<p<\infty$. | math.FA | math |
CARLESON MEASURES AND TOEPLITZ OPERATORS ON
SMALL BERGMAN SPACES ON THE BALL
VAN AN LE
Abstract. We study the Carleson measures and the Toeplitz operators
on the class of so-called small weighted Bergman spaces, introduced
recently by Seip. A characterization of Carleson measures is obtained
which extends Seip's results from the unit disc of C to the unit ball of Cn.
We use this characterization to give necessary and sufficient conditions
for the boundedness and compactness of Toeplitz operators. Finally,
we study the Schatten p classes membership of Toeplitz operators for
1 < p < ∞.
1. Introduction
Let Cn denote the n−dimensional complex Euclidean space, Bn = {z ∈
Cn : z < 1} be the unit ball and Sn = {z ∈ Cn : z = 1} be the unit sphere
in Cn. Denote by H(Bn) the space of all holomorphic functions on the unit
ball Bn. Let dv be the normalized volume measure on Bn. The normalized
surface measure on Sn will be denoted by dσ.
Let ρ be a positive continuous and integrable function on [0, 1). We extend
it to Bn by ρ(z) = ρ(z), and call such ρ a weight function. The weighted
Bergman space A2
ρ is the space of functions f in H(Bn) such that
ρ =ZBnf (z)2ρ(z)dv(z) < ∞.
kfk2
Note that A2
endowed with the inner product
ρ is a closed subspace of L2(Bn, ρdv) and hence is a Hilbert space
hf, giρ =ZBn
f (z)g(z)ρ(z)dv(z),
f, g ∈ A2
ρ.
When ρ(r) = (1 − r2)α, α > −1, we obtain the standard Bergman spaces
A2
α.
We impose a normalization condition on ρ:
Consider the points rk ∈ [0, 1) determined by the relation
0
Z 1
Z 1
rk
x2n−1ρ(x)dx = 1.
ρ(x)dx = 2−k.
2010 Mathematics Subject Classification. 30H20; 47B35.
Key words and phrases. Bergman spaces, Carleson measures, Toeplitz operators, Schat-
ten classes.
1
CARLESON MEASURES AND TOEPLITZ OPERATORS
2
Denote by S the class of weights ρ such that
inf
k
1 − rk
1 − rk+1
> 1.
(1.1)
Since the function r 7→RSnf (rξ)2dσ(ξ) is non-decreasing, we also have
the equivalent norm
kfk2
ρ ≍
2−kZSnf (rkξ)2dσ(ξ),
∞Xk=1
f ∈ A2
ρ.
(1.2)
The class S was introduced by Kristian Seip in [13]. It is easy to see that
the functions
and
belong to S.
ρ(x) = (1 − x)−β,
1
ρ(x) = (1 − x)−1(cid:18)log
1 − x(cid:19)−α
0 < β < 1,
,
1 < α < ∞,
Bergman spaces A2
properties of Toeplitz operators on these spaces.
In this paper we prove a characterization of Carleson measure for weighted
ρ, with ρ ∈ S. This result is then used to study spectral
Let µ be a finite positive Borel measure on Bn. We say that µ is a Carleson
measure for a Hilbert space X of analytic functions in Bn if there exists a
positive constant C such that
ZBnf (z)2dµ(z) ≤ Ckfk2
X,
f ∈ X.
It is clear that µ is a Carleson measure for A2
and the identity operator Id : A2
constant of µ, denoted by Cµ(A2
Suppose that µ is a Carleson measure for A2
Carleson measure for A2
is,
ρ ⊂ L2(Bn, dµ)
ρ → L2(Bn, dµ) is bounded. The Carleson
ρ), is the norm of this identity operator Id.
ρ. We say that µ is a vanishing
ρ if the above identity operator Id is compact. That
ρ if and only if A2
k→∞ZBnfk(z)2dµ(z) = 0
lim
whenever {fk} is a bounded sequence in A2
on compact subsets of Bn.
ρ which converges to 0 uniformly
The concept of a Carleson measure was first introduced by L. Carleson
[2, 3] in order to study interpolating sequences and the corona problem on
the algebra H ∞ of all bounded analytic functions on the unit disk. It quickly
became a powerful tool for the study of function spaces and operators acting
on them. The Carleson measures on Bergman spaces were studied by Hast-
ings [4], and later on by Luecking [6], and many others. Recently, Pau and
Zhao [8] gave a characterization for Carleson measures and vanishing Car-
leson measures on the unit ball by using the products of functions in weighted
Bergman spaces. In [9], Pel´aez and Rattya gave a description of Carleson
measures for A2
r ρ(t)dt is
either equivalent to 1 or tends to ∞, and in [10] they then got a criterion
for A2
(1 − r)ρ(r)R 1
ρ on unit disk when ρ is such that
ρ(s)ds.
ρ on unit disk when ρ ∈ bD, which meansR 1
r ρ(s)ds .R 1
r+1
2
1
CARLESON MEASURES AND TOEPLITZ OPERATORS
3
In [13], Seip gave a characterization of Carleson measures for A2
ρ with
ρ ∈ S in the case n = 1. One of our main results, Theorem 2.1, extends this
result to the case n > 1.
ρ with
Given a function ϕ ∈ L∞(Bn), the Toeplitz operator Tϕ on A2
symbol ϕ is defined by
Tϕf = P (ϕf ),
f ∈ A2
ρ,
where P : L2(Bn, ρdv) → A2
the integral representation of P , we can write Tϕ as
ρ is the orthogonal projection onto A2
ρ. Using
Tϕf (z) =ZBn
Kρ(z, w)f (w)ϕ(w)ρ(w)dv(w),
z ∈ Bn,
where Kρ(z, w) is the reproducing kernel for A2
ρ. The Toeplitz operators
can also be defined for unbounded symbols or for finite measures on Bn. In
fact, given a finite positive Borel measure µ on Bn, the Toeplitz operator
Tµ : A2
ρ is defined as follows:
ρ → A2
Note that
Tµf (z) =ZBn
hTµf, giρ =ZBn
Kρ(z, w)f (w)dµ(w),
z ∈ Bn.
f (z)g(z)dµ(z),
f, g ∈ A2
ρ.
ρ, with ρ ∈ S.
The Toeplitz operators acting on various spaces of holomorphic functions
have been extensively studied by many authors, and the theory is especially
well understood in the case of Hardy spaces or standard Bergman spaces
(see [14], [15] and the references therein). Luecking [7] was the first to study
Toeplitz operators on Bergman spaces with measures as symbols, and some
interesting results about Toeplitz acting on large Bergman spaces were ob-
tained by Lin and Rochberg [5]. In this paper, we will study the boundedness
and compactness of Tµ on A2
Next we study when our Toeplitz operators belong to the Schatten class.
We refer to [15, Chapter 1] for a brief account on the Schatten classes. A
description for the standard Bergman spaces on the unit disk was given (see
[15, Chapter 7]), and a description for the case of large Bergman spaces on
the disk was obtained in 2015 by H. Arroussi, I. Park, and J. Pau [1]. In
2016, Pel´aez and Rattya [11] gave an interesting characterization for the
case of small Bergman spaces on unit disk, where the weight ρ ∈ bD. Note
that S $ bD, but(cid:8)A2
that ρ∗(r) . ρ(r) for r ∈ (0, 1), where
1
We introduce a subclass S∗ of weights in S determined by the condition
ρ : ρ ∈ S(cid:9) =nA2
ρ∗(r) =
ρ(t)dt.
For example, the weights
ρ(x) = (1 − x)−β(cid:18)log
0 < β < 1, α ∈ R
ρ : ρ ∈ bDo.
1 − rZ 1
1 − x(cid:19)α
1
,
r
CARLESON MEASURES AND TOEPLITZ OPERATORS
4
belong to S∗, but the weights
ρ(x) = (1 − x)−1(cid:18)log
1
1 − x(cid:19)α
1 − x(cid:19)−1(cid:18)log log
1
ρ(x) = (1 − x)−1(cid:18)log
,
α < −1,
1
1 − x(cid:19)α
,
α < −1,
do not belong to S∗.
For weights ρ in S∗, we obtain a characterization of the symbols of the
Toeplitz operators in the Schatten classes Sp. In [12], Pel´aez, Rattya and
Sierra gave a characterization for the case of dimension n = 1 when the
weight is regular, that is ρ∗(r) ≍ ρ(r). As an easy observation, our result
is equivalent to their result when n = 1. We point out that our approach
is completely different from that of [12], which does not seem to work in
higher dimensions. On the other hand, for regular weights ρ in S \ S∗, this
characterization fails. A counterexample was given in [12].
The paper is organized as follows: The main results are stated in Section
2 and their proofs are given in Section 3 -- 5.
2. Main results
Throughout this text, we use the following notation. For every nonnega-
tive integer k, set
Ωk = {z ∈ Bn : rk ≤ z < rk+1} ,
and let µk be the measure defined by µk = χΩk µ whenever a nonnegative
Borel measure µ on Bn is given. The notation U (z) . V (z) (or equivalently
V (z) & U (z)) means that there is a positive constant C such that U (z) ≤
CV (z) holds for all z in the set in question, which may be a space of functions
or a set of numbers. If both U (z) . V (z) and V (z) . U (z), then we write
U (z) ≍ V (z).
Our results are following:
Theorem 2.1. Let ρ ∈ S, and let µ be a finite positive Borel measure on
Bn. Then
(i) µ is a Carleson measure for A2
ρ if and only if each µk is a Carleson
measure for the Hardy space H 2 with Carleson constant Cµk (H 2) .
2−k, k ≥ 0.
ρ if and only if
(ii) µ is a vanishing Carleson measure for A2
2kCµk (H 2) = 0.
lim
k→∞
Theorem 2.1 (i) for the case n = 1 was obtained by Seip in [13].
Theorem 2.2. Let ρ ∈ S, and let µ be a finite positive Borel measure on
Bn. Then
(i) The Toeplitz operator Tµ is bounded on A2
ρ if and only if µ is a
Carleson measure for A2
ρ.
(ii) The Toeplitz operator Tµ is compact on A2
ρ if and only if µ is a
vanishing Carleson measure for A2
ρ.
CARLESON MEASURES AND TOEPLITZ OPERATORS
5
Given z ∈ Bn and 0 < α < 1, we consider the Bergman metric ball
where β(z, w) is the Bergman metric given by
E(z, α) = {w ∈ Bn : β(z, w) < α} ,
β(z, w) =
1
2
log
1 + ϕz(w)
1 − ϕz(w)
,
z, w ∈ Bn.
Here, ϕz is the Mobius transformation on Bn that interchanges 0 and z.
We know that E(0, α) is actually a Euclidean ball of radius R = tanh α,
centered at the origin, and
E(z, α) = ϕz(cid:0)E(0, α)(cid:1).
more details.
Moreover, for fixed α, v(cid:0)E(z, α)(cid:1) ≍ (1 − z)n+1. See [14, Chapter 1] for
For a measure µ on Bn and α > 0, we define the function bµα by
Denote by fTµ the Berezin transform of Tµ, and set
2kµ(cid:0)E(z, α)(cid:1)
bµα(z) =
(1 − z)n
2kρ(z)dv(z)
z ∈ Ωk.
dλρ(z) =
,
(1 − z)n ,
z ∈ Ωk.
Theorem 2.3. Let ρ be in S∗, µ be a finite positive Borel measure and
1 < p < ∞. The following conditions are equivalent:
(a) The Toeplitz operator Tµ is in the Schatten class Sp.
(b) The function fTµ is in Lp(Bn, dλρ).
(c) The function bµα is in Lp(Bn, dλρ) for sufficiently small α > 0.
Given a ∈ Bn \ {0} and r > 0. Let δ(a) =p2(1 − a). Define Q(a, r) ⊂
Bn and O(a, r) ⊂ Sn as follows:
3. Proof of Theorem 2.1
Q(a, r) = {z ∈ Bn :p1 − ha/a, zi < r},
O(a, r) = {ζ ∈ Sn :p1 − ha/a, ζi < r}.
For simplicity of notation, we write Qa instead of Q(cid:0)a, δ(a)(cid:1), Oa instead of
O(cid:0)a, δ(a)(cid:1).
We recall a well known characterization of Carleson measures for the
Hardy space (see [14]): A positive Borel measure µ on Bn is a Carleson
measure for H 2 if and only if µ(Qa) . (1 − a)n for all a ∈ Bn \ {0}.
Furthermore, Cµ(H 2) ≍ supa∈Bn\{0} µ(Qa)(1 − a)−n.
We use the following covering lemma from [14, Lemma 4.7].
Lemma 3.1. Suppose N is a natural number, al ∈ Bn \ {0} , 1 ≤ l ≤ N ,
E =
Oal .
N[l=1
There exists a subsequence {li}, 1 ≤ i ≤ M , such that
(a) Oali
, 1 ≤ i ≤ M , are disjoint.
CARLESON MEASURES AND TOEPLITZ OPERATORS
6
(b) O(cid:0)ali, 3δ(ali )(cid:1), 1 ≤ i ≤ M , cover E.
Lemma 3.2. Let µ be a finite positive measure on Bn. Then µk is a Car-
leson measure for H 2 if and only if µk(Qa) . (1 − a)n for all a ∈ Ωk.
Furthermore, Cµk (H 2) ≍ supa∈Ωk (1 − a)−nµk(Qa).
Proof. Let a ∈ Bn \ {0}. Then a ∈ Ωl for some l ≥ 1.
If l > k, then
µk(Qa) = 0 and there is nothing to prove. When a ∈ Ωl, l ≤ k, we can cover
Qa \ rkBn by a finite family Λ of Qal with al ∈ Ωk−1. Denote by Λ0 the
sub-family of Λ obtained by Lemma 3.1. Then
µk(Qa) = µk(Qa \ rkBn) ≤ Xl∈Λ0
µk(cid:16)Q(cid:0)al, 3δ(al)(cid:1)(cid:17).
Finally,
σ(Oal) = σ(cid:16)[l∈Λ0
Since al ∈ Ωk−1, we have µk(cid:16)Q(cid:0)al, 3δ(al)(cid:1)(cid:17) . (1 − al)n ≍ σ(Oal ). Hence
µk(Qa) . Xl∈Λ0
σ(cid:16)[l∈Λ0
Oal(cid:17).
Oal(cid:17) . σ(Oa) ≍ (1 − a)n.
Therefore µk(Qa) . (1 − a)n. This completes the proof.
3.1. Proof of Part (i). (⇐=) Since µk are Carleson measures for H 2 with
Carleson constants . 2−k, the same holds for H 2 on the smaller ball rk+2Bn.
This means that
(cid:3)
ZΩkf (z)2dµ(z) . 2−kZSnf (rk+2ξ)2dσ(ξ)
for an arbitrary function f in A2
all k ≥ 1 we get
ρ and for all k. Summing this estimate over
ZBnf (z)2dµ(z) .
∞Xk=1
2−kZSnf (rk+2ξ)2dσ(ξ) ≍ kfk2
ρ.
(=⇒) We just need to check that µk(Qa) . 2−k(1 − a)n when a is in
Ωk, k ≥ 0. We use the test function
with large γ. By (1.2), we have
kfak2
ρ ≍
1 − ha, rjξi2γ dσ(ξ)
1
fa(z) = (1 − ha, zi)−γ
2−jZSn
∞Xj=1
∞Xj=1
(1 − rja)2γ−n .
ρ ≍ 2−k(1 − a)−2γ+n.
kfak2
2−j
≍
Since a ∈ Ωk, by Property (1.1) we obtain
(3.1)
(3.2)
CARLESON MEASURES AND TOEPLITZ OPERATORS
7
On the other hand, for every z in Qa we have
1 − ha, zi = (1 − a) + a(1 − ha/a, zi)
≤ (1 − a) + a1 − ha/a, zi
< (1 − a) + 2a(1 − a)
≤ 3(1 − a).
fa(z) & (1 − a)−γ,
z ∈ Qa.
(3.3)
Hence,
Thus,
ZBnfa(z)2dµ(z) & (1 − a)−2γµ(Qa ∩ Ωk).
Since µ is a Carleson measure for A2
ρ, we get
µ(Qa ∩ Ωk) . 2−k(1 − a)n.
This implies that µk is a Carleson measure for Hardy space H 2 with Carleson
constant Cµk (H 2) . 2−k.
3.2. Proof of Part (ii). Suppose that µ is a vanishing Carleson measure
for A2
ρ. Given a in Ωk, consider the function fa defined by (3.1). By (3.2),
kfak2
ρ ≍ 2−k(1 − a)−2γ+n. Set
ha(z) =
(1 − ha, zi)−γ
(3.4)
✷
.
2−k/2(1 − a)−γ+n/2
z ∈ Qa.
ρ and ha tends to 0 uniformly
Then khak2
ρ ≍ 1 and by (3.3),
ha(z)2 &
2k
(1 − a)n ,
Since µ is a vanishing Carleson measure for A2
on compact subsets of the unit ball as a → 1, we have
lim
a→1ZBnha(z)2dµ(z) = 0.
2kµk(Qa ∩ Ωk)
(1 − a)n → 0 as k → ∞. Hence,
Thus, sup
a∈Ωk
Conversely, let µr = µBn\rBn
(i) of Theorem 2.1 implies that
, where rBn = {z ∈ Bn : z < r}. Then part
lim
k→∞
2kCµk (H 2) = 0.
where
ZBnh(z)2dµr(z) ≤ Crkhk2
k:rk>r Cµk (H 2),
and
ρ,
Cr = sup
h ∈ A2
ρ,
lim
r→1
Cr = 0.
(3.5)
Let {fk} be a bounded sequence in A2
ρ converging uniformly to 0 on compact
subsets of Bn. Let ε > 0. By (3.5), there exists r0 ∈ (0, 1) such that Cr < ε
for all r ≥ r0. Moreover, by the uniform convergence on compact subsets,
CARLESON MEASURES AND TOEPLITZ OPERATORS
8
we may choose k0 ∈ N such that fk(z)2 < ε for all k ≥ k0 and z ∈ r0Bn. It
follows that
ZBnfk(z)2dµ(z) =Zr0Bnfk(z)2dµ(z) +ZBn\r0Bnfk(z)2dµ(z)
< εµ(r0Bn) +ZBnfk(z)2dµr0(z)
≤ εµ(r0Bn) + Cr0kfkk2
≤ εC,
k ≥ k0,
ρ
for some positive constant C. Hence, µ is a vanishing Carleson measure for
A2
ρ.
✷
4. Proof of Theorem 2.2
4.1. Proof of Part (i). (=⇒) Given a in Ωk, we define ha by (3.4). Then
khak2
ρ ≍ 1 and ha(z)2 & 2k(1 − a)−n,
z ∈ Qa.
Consider the function
fTµ(a) = hTµha, haiρ =ZBnha2dµ(z).
Since Tµ is bounded, A := supa∈BnfTµ(a) < ∞. Then
A ≥ZBnha(z)2dµ(z) ≥ZBnha(z)2dµk(z)
≥ZQaha(z)2dµk(z) & 2k(1 − a)−nµk(Qa).
(4.1)
(4.2)
Hence, µk(Qa) . 2−k(1 − a)n for every a ∈ Ωk. By Theorem 2.1 and
Lemma 3.2, µ is Carleson measure for A2
ρ.
and
ρ,
ZBnf (z)2dµ(z) ≤ Ckfk2
ZBng(z)2dµ(z) ≤ Ckgk2
ρ.
(⇐=) For every f, g ∈ A2
ρ we have
hTµf, giρ =ZBn
Then by Cauchy -- Schwarz inequality, we get
f (z)g(z)dµ(z).
hTµf, giρ ≤ZBnf (z)g(z)dµ(z)
≤(cid:18)ZBnf (z)2dµ(z)(cid:19) 1
2(cid:18)ZBng(z)2dµ(z)(cid:19) 1
2
.
Since µ is a Carleson measure for A2
that
ρ, there exists a positive constant C such
CARLESON MEASURES AND TOEPLITZ OPERATORS
Hence,
hTµf, giρ ≤ Ckfkρkgkρ
Thus, Tµ is bounded on A2
ρ.
for all f, g ∈ A2
ρ.
4.2. Proof of part (ii). We need the following auxiliary results.
Proposition 4.1. Suppose that f ∈ A2
ρ with ρ ∈ S. Then
(1 − z)nkfk2
ρ,
z ∈ Ωk, k ≥ 0,
f (z)2 ≤
C2k
where C is a positive constant independent of k and z.
9
✷
(4.3)
Proof. Let z ∈ Ωk. Applying [14, Corollary 4.5] to the function g(z) =
f (rk+2z) at the point
, we obtain
z
rk+2
f (z)2 ≤ZSnf (rk+2ζ)2 (1 − z/rk+22)n
By (1.1), 1−hz/rk+2, ζi ≥ 1−hz/rk+2, ζi ≥ 1− zζ
for z ∈ Ωk, ζ ∈ Sn. Thus,
f (z)2 .ZSnf (rk+2ζ)2 (1 − z2)n
rk+2
1 − hz/rk+2, ζi2n dσ(ζ).
= 1−z/rk+2 & 1−z
(1 + z)n
(1 − z)2n dσ(ζ)
(1 − z)nZSnf (rk+2ζ)2dσ(ζ)
(1 − z)n 2−kZSnf (rk+2ζ)2dσ(ζ)
2−jZSnf (rj+2ζ)2dσ(ζ)
2k
2k
∞Xj=1
(1 − z)n
(1 − z)nkfk2
2k
ρ,
≤
.
≤
.
(cid:3)
with constants independent of k and z.
Corollary 4.2. A sequence of functions {fk} ⊂ A2
ρ if and only if it is bounded in A2
A2
compact subset of Bn.
Proof of part (ii) of Theorem 2.2. Suppose that Tµ is compact on A2
ρ converges to 0 weakly in
ρ and converges to 0 uniformly on each
ρ. We
ρ ≍ 1 and ha converges
uniformly to 0 on compact subsets of Bn as a → 1. Since Tµ is compact,
define ha, a ∈ Bn by (3.4) andfTµ by (4.1). Then khak2
fTµ(a) → 0 as a → 1. By (4.2) this implies that
2kµk(Qa)
(1 − a)n → 0 as k → ∞.
sup
a∈Ωk
Hence,
lim
k→∞
2kCµk (H 2) = 0.
By part (ii) of Theorem 2.1, µ is a vanishing Carleson measure for A2
ρ.
CARLESON MEASURES AND TOEPLITZ OPERATORS
10
Conversely, assume that µ is a vanishing Carleson measure for A2
ρ. For
every h ∈ A2
ρ we have
kTµhkρ = sup
hTµh, giρ.
g∈A2
ρ
kgkρ≤1
Furthermore,
hTµh, giρ =(cid:12)(cid:12)(cid:12)(cid:12)ZBn
h(z)g(z)dµ(z)(cid:12)(cid:12)(cid:12)(cid:12) ≤ZBnh(z)g(z)dµ(z)
≤(cid:18)ZBnh(z)2dµ(z)(cid:19)1/2(cid:18)ZBng(z)2dµ(z)(cid:19)1/2
.(cid:18)ZBnh(z)2dµ(z)(cid:19)1/2
kgkρ.
The last inequality follows from the fact that µ is a Carleson measure for
A2
ρ. Therefore,
kTµhkρ .(cid:18)ZBnh(z)2dµ(z)(cid:19)1/2
,
h ∈ A2
ρ.
Now, let {fk} ⊂ A2
subsets of Bn. Since µ is a vanishing Carleson measure for A2
ρ,
ρ be bounded and converge uniformly to 0 on compact
k→∞ZBnfk(z)2dµ(z) = 0.
lim
It follows that kTµfkkρ → 0 and hence Tµ is compact.
5. Proof of Theorem 2.3
Proposition 5.1. Let Kρ(z, w) be the reproducing kernel of A2
ρ.
(a) Let k ≥ 1, z ∈ Ωk. Then
Kρ(z, z) ≍
2k
(1 − z)n .
(b) There exists α = α(ρ) > 0 such that for every z ∈ Bn,
Kρ(z, w)2 ≍ Kρ(z, z)Kρ(w, w)
whenever w ∈ E(z, α).
(cid:3)
(5.1)
(5.2)
Proof. (a) Fix k ≥ 1. Given z ∈ Ωk, let Lz be the point evaluation at z on
A2
ρ. It is well known that
By Proposition 4.1,
Kρ(z, z) = kLzk2.
Furthermore, choosing hz by (3.4), we have khzkρ ≍ 1 and
kLzk2 .
2k
(1 − z)n .
hz(z)2 &
2k
(1 − z)n .
CARLESON MEASURES AND TOEPLITZ OPERATORS
11
kLzk2 &
2k
(1 − z)n .
Hence,
Thus
Kρ(z, z) ≍
(b) It is well known that
2k
(1 − z)n ,
z ∈ Ωk.
for all z, w ∈ Bn. For any fixed z0 ∈ Ωk, consider the subspace A2
as
ρ(z0) defined
Denote by Lz0 the one-dimensional subspace spanned by the function
A2
Kρ(z, w)2 ≤ Kρ(z, z)Kρ(w, w)
ρ : f (z0) = 0(cid:9) .
ρ(z0) =(cid:8)f ∈ A2
pKρ(z0, z0)
kρ,z0(z) =
Kρ(z, z0)
.
Then we have the orthogonal decomposition
A2
ρ = A2
ρ(z0) ⊕ Lz0.
Hence Kρ(z, w) = Kρ,z0(z, w) + kρ,z0(w)kρ,z0(z), where Kρ,z0 is the repro-
ducing kernel of A2
ρ(z0). Therefore,
Kρ(z0, w) = kρ,z0(w)kρ,z0(z0)
and
Kρ(w, w) = Kρ,z0(w, w) + kρ,z0(w)2.
We are going to prove that there exist α > 0 such that
Kρ,z0(w, w) <
1
2
Kρ(w, w),
w ∈ E(z0, α).
(5.3)
(5.4)
By (1.1), there exists α1 > 0 such that E(z0, α) ⊂ Ωk−1 ∪ Ωk ∪ Ωk+1, 0 <
α < α1. Hence, for every f ∈ A2
ρ(z0) such that kfkρ = 1, by Proposition 4.1
we have
f (w)2 .
2k
(1 − w)n ≍
2k
(1 − z0)n
as
whenever w ∈ E(z0, α). Since E(z0, α) = ϕz0(cid:0)E(0, α)(cid:1), we can rewrite (5.5)
whenever η ∈ E(0, α). Note that f (z0) = f(cid:0)ϕz0(0)(cid:1) = 0. Therefore, by the
f(cid:0)ϕz0(η)(cid:1)2 .
Schwarz lemma, we get
(1 − z0)n
(5.6)
2k
(5.5)
whenever η ∈ E(0, α). This implies that there is a constant C > 0 such that
f(cid:0)ϕz0(η)(cid:1)2 . η2
f(cid:0)ϕz0(η)(cid:1)2 ≤ Cη2
2k
(1 − z0)n ≍ η2
2k
(1 − ϕz0 (η))n
2k
(1 − ϕz0(η))n ,
η ∈ E(0, α).
CARLESON MEASURES AND TOEPLITZ OPERATORS
12
Therefore, we can choose α so small that
f(cid:0)ϕz0(η)(cid:1)2 <
This proves (5.4).
1
2
Kρ(cid:0)ϕz0(η), ϕz0 (η)(cid:1),
η ∈ E(0, α).
Now, from (5.3) and (5.4), we obtain that kρ,z0(w)2 >
ever w ∈ E(z0, α). This means that
1
Kρ(w, z0)2 >
2
Kρ(z0, z0)Kρ(w, w)
whenever w ∈ E(z0, α), which completes the proof.
Lemma 5.2. Let T be a positive operator on A2
transform of T , defined by
1
2
Kρ(w, w) when-
(cid:3)
ρ, and let eT be the Berezin
eT (z) = hT kz, kziρ,
z ∈ Bn.
if z ∈ Ωk.
Here, dλρ(z) =
2kρ(z)dv(z)
(1 − z)n
(a) Let 0 < p ≤ 1. If eT ∈ Lp(Bn, dλρ), then T is in Sp.
(b) Let p ≥ 1. If T is in Sp, then eT ∈ Lp(Bn, dλρ).
Proof. Note that dλρ(z) ≍ K(z, z)ρ(z)dv(z) = kKzk2ρ(z)dv(z).
The proof is similar to the proof of [1, Lemma 4.2]. The positive operator
T is in Sp if and only if T p is in the trace class S1. Fix an orthonormal basis
ρ. Since T p is positive, it is in S1 if and only ifPkhT pek, ekiρ < ∞.
{ek} of A2
Let U = √T p. By Fubini's theorem, the reproducing property of Kz, and
Parseval's identity, we have
Xk
hT pek, ekiρ =Xk
kU ekk2
ρ =Xk ZBnU ek(z)2ρ(z)dv(z)
U ek(z)2! ρ(z)dv(z)
hU ek, Kziρ2! ρ(z)dv(z)
hek, U Kziρ2! ρ(z)dv(z)
=ZBn Xk
=ZBn Xk
=ZBn Xk
=ZBnkU Kzk2
=ZBnhT pKz, Kziρρ(z)dv(z)
=ZBnhT pkz, kziρkKzk2
≍ZBnhT pkz, kziρdλρ(z).
ρρ(z)dv(z)
ρρ(z)dv(z)
CARLESON MEASURES AND TOEPLITZ OPERATORS
13
Hence, both (a) and (b) are the consequences of the well known inequalities
(see [15, Proposition 1.31])
hT pkz, kziρ ≤ hT kz, kzip
hT pkz, kziρ ≥ hT kz, kzip
ρ =(cid:0)eT (z)(cid:1)p,
ρ =(cid:0)eT (z)(cid:1)p,
(cid:3)
Lemma 5.3. Let ρ ∈ S∗ and z ∈ Ωk. Then there exists α0 > 0 such that
for every α ∈ (0, α0) we have
0 < p ≤ 1,
p ≥ 1.
f (z)2 .
2k
(1 − z)nZE(z,α)f (w)2ρ(w)dv(w)
for all f ∈ H(Bn).
Proof. Let z ∈ Ωk. For each f ∈ H(Bn), by the subharmonicity of the
function w 7→ f (w)2 and the estimate v(cid:0)E(z, α)(cid:1) ≍ (1 − z)n+1, we have
1
It is easy to see that 1 − z ≍ 1 − w for w ∈ E(z, α). Hence,
dv(w)
1
f (z)2 .
f (z)2 .
=
(1 − z)n+1ZE(z,α)f (w)2dv(w).
(1 − z)nZE(z,α)f (w)2
(1 − z)nZE(z,α)f (w)2 2−k
1 − w
1 − w
2k
1
dv(w).
(5.7)
By (1.1), for small α0 we have E(z, α0) ⊂ Ωk−1 ∪ Ωk ∪ Ωk+1. Therefore,
for every α ∈ (0, α0), we have rk−1 < w < rk+2 for w ∈ E(z, α). Since
w ρ(t)dt for every w ∈ E(z, α), α ∈
rk+2
(0, α0). Plugging this into (5.7) and using that ρ∗(w) . ρ(w), we get
ρ(t)dt = 2−k−2, we obtain 2−k .R 1
R 1
f (z)2 .
.
2k
(1 − z)nZE(z,α)f (w)2ρ∗(w)dv(w)
(1 − z)nZE(z,α)f (w)2ρ(w)dv(w).
2k
fTµ(z) =ZBnkz(w)2dµ(w) = kKzk−2
ρ ZE(z,α)Kz(w)2dµ(w)
ρdµ(w) ≍bµα(z).
≥ kKzk−2
≍ZE(z,α)kKwk2
Since fTµ is in Lp(Bn, dλρ), bµα is also in Lp(Bn, dλρ).
This completes the proof.
(cid:3)
Proof of Theorem 2.3. (a) ⇒ (b). This follows from Lemma 5.2 (b).
(b) ⇒ (c). By Proposition 5.1 (b), for sufficiently small α > 0, we have
ρkKwk2
ρ,
Then by Proposition 5.1 (a), we get
Kz(w)2 ≍ kKzk2
w ∈ E(z, α), z ∈ Bn.
ρ ZBnKz(w)2dµ(w)
where 1
p + 1
q = 1. Thus, (5.8) implies that
Xl
hTµel, elip
,
≤(cid:18)ZBnel(w)2bµα(w)pρ(w)dv(w)(cid:19)1/p
×(cid:18)ZBnel(w)2ρ(w)dv(w)(cid:19)1/q
=(cid:18)ZBnel(w)2bµα(w)pρ(w)dv(w)(cid:19)1/p
ρ .ZBn Xl
el(w)2!bµα(w)pρ(w)dv(w)
=ZBnkKwk2
ρbµα(w)pρ(w)dv(w)
≍ZBnbµα(w)pdλρ(w) < ∞.
CARLESON MEASURES AND TOEPLITZ OPERATORS
14
(c) ⇒ (a). For every orthonormal basis {el} of A2
ρ, we have
By Lemma 5.3,
el(z)2 .
2k
ρ =Xl (cid:18)ZBnel(z)2dµ(z)(cid:19)p
Xl
hTµel, elip
(1 − z)nZE(z,α)el(w)2ρ(w)dv(w),
ZBnel(z)2dµ(z) .ZBnel(w)2bµα(w)ρ(w)dv(w)
By Fubini's theorem and Holder's inequality, we have
.
(5.8)
z ∈ Ωk.
This proves (a).
(cid:3)
Remark 5.4. Let 1 < p < ∞.
spaces, Arroussi, Park and Pau proved in [1, Theorem 4.6] that
In the case of large weighted Bergman
µ(cid:0)B(z, ε)(cid:1)
(1 − z)2n is in the corresponding weighted Lp,
Tµ ∈ Sp ⇐⇒eµε(z) =
where B(z, ε) is the Euclidean ball with center z and radius ε(1−z). When
the dimension n = 1, we can see that eµε is in Lp if and only if bµε is in Lp.
However, for n > 1, this equivalence is not true anymore.
Let us verify this. Choose zk ∈ Bn such that zk tend to 1 sufficiently
rapidly as k → ∞. Consider
µ =
ckχB(zk,ε)
and
µ∗ =
∞Xk=1
∞Xk=1
ckχB(zk,3ε),
where ck > 0 will be chosen later. We have
µ .eµε . µ∗
CARLESON MEASURES AND TOEPLITZ OPERATORS
15
and
Hence
∞Xk=1
ck
ck
∞Xk=1
v(cid:0)B(zk, ε)(cid:1)
v(cid:0)E(zk, ε)(cid:1) χE(zk,3ε).
cp
v(cid:0)B(zk, ε)(cid:1)
v(cid:0)E(zk, ε)(cid:1) χE(zk,ε) .bµε .
eµε ∈ Lp ⇐⇒
bµε ∈ Lp ⇐⇒
kv(cid:0)B(zk, ε)(cid:1)
cp
∞Xk=1
∞Xk=1
k(cid:0)v(cid:0)B(zk, ε)(cid:1)(cid:1)p(cid:0)v(cid:0)E(zk, ε)(cid:1)(cid:1)1−p
kv(cid:0)B(zk, ε)(cid:1) < ∞,
k (cid:0)v(cid:0)B(zk, ε)(cid:1)(cid:1)p
(cid:0)v(cid:0)E(zk, ε)(cid:1)(cid:1)p−1 < ∞.
v(cid:0)E(zk, ε)(cid:1)!p−1
= v(cid:0)B(zk, ε)(cid:1)
as k → ∞, we can choose ck such that bµε ∈ Lp but eµε /∈ Lp. On the other
hand, one can easily see that eµε ∈ Lp implies bµε ∈ Lp.
Acknowledgments.
I am deeply grateful to my advisors Professors
Alexander Borichev and El Hassan Youssfi for their help and many sug-
gestions during the preparation of this paper.
≍ (1 − zk)(n−1)(p−1) −→ 0
and
Since
cp
cp
References
[1] H. Arroussi, I. Park and J. Pau, Schatten class Toeplitz operators acting on large
weighted Bergman spaces, Studia Mathematica 229 (2015), no. 3 , 203 -- 221.
[2] L. Carleson, An interpolation problem for bounded analytic functions, Amer. J. Math.
80 (1958), 921 -- 930.
[3] L. Carleson, Interpolations by bounded analytic functions and the corona problem, Ann.
of Math. 76 (1962), 547 -- 559.
[4] W. Hastings, A Carleson measure theorem for Bergman spaces, Proceedings of the
American Mathematical Society 52 (1975), 237 -- 241.
[5] P. Lin and R. Rochberg, Trace ideal criteria for Toeplitz and Hankel operators on the
weighted Bergman spaces with exponential type weights, Pacific Journal of Mathematics
173 (1996), no. 1, 127 -- 146.
[6] D. Luecking, A technique for characterizing Carleson measures on Bergman spaces,
Proceedings of the American Mathematical Society 87 (1983), 656 -- 660.
[7] D. Luecking, Trace ideal criteria for Toeplitz operators, Journal of Functional Analysis
73 (1987), 345 -- 368.
[8] J. Pau and R. Zhao, Carleson measures and Teoplitz operators for weighted Bergman
spaces on the unit ball, Michigan Math. J. 64 (2015), no. 4, 759 -- 796.
[9] J. A. Pel´aez and J. Rattya, Weighted Bergman spaces induced by rapidly increasing
weights, Mem. Amer. Math. Soc. 227 (2014), no. 1066, 124 pp.
[10] J. A. Pel´aez and J. Rattya, Embedding theorems for Bergman spaces via harmonic
analysis, Mathematische Annalen 362 (2015), no. 1, 205 -- 239.
[11] J. A. Pel´aez and J. Rattya, Trace class criteria for Toeplitz and composition operators
on small Bergman spaces, Advances in Mathematics 293 (2016), 606-643.
[12] J. A. Pel´aez, J. Rattya and K. Sierra, Berezin Transform and Toeplitz Operators on
Bergman Spaces Induced by Regular Weights, The Journal of Geometric Analysis 28
(2018), 656-687.
CARLESON MEASURES AND TOEPLITZ OPERATORS
16
[13] K. Seip, Interpolation and sampling in small Bergman spaces, Collectanea Mathe-
matica 64 (2013), no. 1, 61 -- 72.
[14] K. Zhu, Spaces of Holomorphic Functions in the Unit Ball, Graduate Texts in Math-
ematics, vol. 226, Springer-Verlag, New York, 2005.
[15] K. Zhu, Operator Theory in Function Spaces, 2nd ed., Mathematical surveys and
monographs, vol. 138, American Mathematical Society, Providence, RI, 2007.
Aix -- Marseille University, CNRS, Centrale Marseille, I2M, Marseille,
France
University of Quynhon, Department of Mathematics, 170 An Duong Vuong,
Quy Nhon, Vietnam
E-mail address: [email protected]
|
1202.3963 | 1 | 1202 | 2012-02-17T16:55:14 | On the numerical radius of the truncated adjoint Shift | [
"math.FA"
] | A celebrated thorem of Fejer (1915) asserts that for a given positive trigonometric polynomial
$\sum_{j=-n+1}^{n-1}c_{j}e^{ijt}$, we have $\lvert c_{1}\lvert\leqslant c_{0}\cos\frac{\pi}{n+1}$. A more recent inequality due to U. Haagerup and P. de la Harpe asserts that, for any contraction $T$ such that $T^{n}=0$, for some $n\geq2$, the inequality $\omega_{2}(T)\leqslant\cos\frac{\pi}{n+1}$ holds, and $\omega_{2}(T)=\cos\frac{\pi}{n+1}$ when T is unitarily equivalent to the extremal operator ${S}^{\ast}_{n}={\bbs}_{\lvert{\C}^{n}}={\bbs}_{\lvert Ker (u_{n}(\bbs))}$ where $u_{n}(z)=z^{n}$ and $\bbs$ is the adjoint of the shift operator on the Hilbert space of all square summable sequences. Apparently there is no relationship between them. In this mathematical note, we show that there is a connection between Taylor coefficients of positive rational functions on the torus and numerical radius of the extremal operator $\bbs(\phi)=\bbs_{\lvert Ker(\phi(\bbs))}$ for a precise inner function $\phi$. This result completes a line of investigation begun in 2002 by C. Badea and G. Cassier \cite{Cassier}. An upper and lower bound of the numerical radius of $\bbs(\phi)$ are given where $\phi$ is a finite Blashke product with unique zero. | math.FA | math |
ON THE NUMERICAL RADIUS OF THE TRUNCATED
ADJOINT SHIFT
HAYKEL GAAYA
j=−n+1 cj eijt, we have c16 c0 cos π
n+1 holds, and ω2(T ) = cos π
CI n = S∗
Abstract. A celebrated thorem of Fejer (1915) asserts that for a given pos-
itive trigonometric polynomial Pn−1
n+1 . A
more recent inequality due to U. Haagerup and P. de la Harpe [9] asserts
that, for any contraction T such that T n = 0, for some n ≥ 2, the inequality
ω2(T ) 6 cos π
n+1 when T is unitarily equivalent
to the extremal operator S∗
Ker(un(S ∗)) where un(z) = zn
and S∗ is the adjoint of the shift operator on the Hilbert space of all square
summable sequences. Apparently there is no relationship between them. In
this mathematical note, we show that there is a connection between Taylor
coefficients of positive rational functions on the torus and numerical radius of
the extremal operator S∗(φ) = S∗
Ker(φ(S ∗)) for a precise inner function φ.
This result completes a line of investigation begun in 2002 by C. Badea and
G. Cassier [1]. An upper and lower bound of the numerical radius of S∗(φ)
are given where φ is a finite Blashke product with unique zero.
n = S∗
1. Introduction
Let H be a complex separable Hilbert space and B(H) the collection of all
bounded linear operators on H. The numerical range of an operators T in B(H) is
the subset
W (T ) = {< T x, x >∈ CI ; x ∈ H,kxk= 1}
of the plane, where < ., . > denotes the inner product in H and the numerical radius
of T is defined by
Re(T ) is the self-adjoint operator defined by
ω2(T ) = sup{z; z ∈ W (T )} .
Re(T ) =
1
2
(T + T ∗).
We denote by [x] the integer part of x and by S the unilateral shift acting on the
Hardy space HI 2 of the square summable analytic functions and by S∗ its adjoint:
S : HI 2 → HI 2
7→ zf (z)
f
S∗
: HI 2 →
7→
f
HI 2
f (z) − f (0)
z
.
2000 Mathematics Subject Classification. 47A12, 47B35.
Key words and phrases. Numerical radius, Numerical range, Truncated shift, Eigenvalues,
Toeplitz forms, Inequalities for positive trigonometric polynomials.
1
2
HAYKEL GAAYA
Beurling's theorem implies that the non zero invariant subspaces of S are of the
forme φ HI 2, where φ is some inner function . Let S(φ) denote the compression of
S to the space H(φ) = HI 2 ⊖ φ HI 2 :
S(φ)f (z) = P (zf (z)),
where P denotes the ortogonal projection from HI 2 onto H(φ). We denote by S∗(φ)
the adjoint of S(φ):
S∗(φ) = S(φ)∗ = S∗
H(φ) = S∗
Ker(φ(S)∗) .
The model operator S(φ) has many properties and it was studied intensively
in the 1960s and '70s. For exemple, it has norm 1 (for dim H(φ) > 1) and it is
cyclic. The function φ is the minimal function of S(φ) meaning that φ(S(φ)) = 0
and φ divides any function ψ in H ∞ with ψ(S(φ)) = 0. The space H(φ) is finite-
dimensional exactly when φ is a finite Blaschke product :
φ(z) =
Yj=1
In this case the polynomial p(z) = Qn
j=1(z − αj) is both the minimal and char-
acteristic polynomial of S(φ) and (αj )16j6n are its eigenvalues. In particular, if
φ(z) = zn then S(φ) is unitarily equivalent to Sn where
.
n
z − αj
1 − αjz
0
1
. . .
. . .
. . .
1
0
.
Sn =
n+1o and
In 1992 U. Haagerup and P. de la Harpe proved that W (Sn) is the disc
Dn =nz ∈ CI ;z6 cos π
(1.1)
ω2(Sn) = cos
π
n + 1
and more generally a natural connection between Fejer's inequality and the numer-
ical radius of a nilpotent matrix was established by Haagerup and de la Harpe.
They proved, using solely elementary methods (positive definite kernels) that:
Theorem 1.1 ([9]). Let T be an operator on H such that T n = 0 for some n ≥ 2.
One has:
ω2(T ) 6 kTkcos
n + 1
π
and ω2(T ) = kTkcos π
n+1 when T is unitarily equivalent to kTkSn.
The reader may consult [10] chapter 22 for properties of numerical ranges
of operators in general, [11] chapter 1 for those of finite dimensional operators and
particulary [6] for the geometric properties of the numerical range of S(φ).
Organisation of the paper:
In [1], C. Badea and G. Cassier showed:
ON THE NUMERICAL RADIUS OF THE TRUNCATED ADJOINT SHIFT
3
Theorem 1.2. ([1]) Let F = P/Q be a rational function with no principal part
(d◦P < d◦Q) which is positive on the torus. Then the Taylor coefficient ck of order
k satisfies the following inequalitiy
where R = S∗
Ker(Q(S ∗)).
ck6 c0 ω2(Rk),
In Section 2, our main theorem is the Theorem 1.2. We give an extension of the
result of C. Badea and G. Cassier for Taylor coefficients of all rational functions
which are positive on the torus. We make no extra assumptions about P and Q.
We do not, for exemple require them to obey any degree restrictions, they need
only be coprime. The theorem has many applications and will explain how we can
easily recover the remarkable Egerváry and Szász inequality. See Corollary 2.3.
Toeplitz matrices are found in several areas of mathematics and its applications
such as complex and harmonic analysis. The KMS Toeplitz matrix
Kn(α) = (αr−s)n
r,s=1
associated with the Poisson kernel introduced by Kac, Murdokh and Szegö [12]
is of particular interst in these areas. In order to formulate our problem we first
review in Section 3.1 some of the known results on the spectra of these matrices
and we give a better upper bound of λ(n)
is the largest eigenvalue of the
KMS matrix. In Section 3.2, our main theorem is the Theorem 3.7 which gives un
upper and lower bound of the numerical radius of the truncated shift S∗(φ) where
1 , where λ(n)
1
1 − αz(cid:19)n
φ(z) = (cid:18) z − α
with 0 6 α < 1 is a finite Blashke product with unique zero.
Our preoccupation with this particular case 0 6 α < 1 to the exclusion to any α
in the disc is explained by the fact that the numerical radius is independent with
the argument of α. This formula is expressed in terms of eigenvalues of the KMS
matrices and provides an easy proof for the Haagerup and de la Harpe result (1.1).
2. Main Theorem
There are many classical inequalities for coefficients of positive trigonometric
polynomials. The next result shows the links between the numerical radius of the
truncated adjoint shift and the Taylor coefficients of rational functions positive on
the torus.
Theorem 2.1. Let F = P/Q be a rational function which is positive on the torus,
where P and Q are coprime. Denote by
and
p
q
φ(z) =
1 − αj z(cid:19)mj
1 − βjz(cid:19)dj
Yj=1(cid:18) z − αj
Yj=1(cid:18) z − βj
j=1 mj and d = Pq
ψ(z) =
the respectively finite Blashke products formed by the nonzero roots of P and Q in
j=1 dj . Then the Taylor coefficient ck
the open disc, let m = Pp
of order k of F satisfies the following inequality:
ck6 c0 ω2(S∗k(ϕ)), where ϕ(z) = zmax(0,m−d+1)ψ(z).
4
HAYKEL GAAYA
Lemma 2.2 ([1] lemma 3.2). Let u be a inner function and let f be a positive
function in the subspace u HI 1
0 of L1(TI ). Then there exists a function h in H(u) =
HI 2 ⊖ u HI 2 such that f = h2.
Proof. First, note that by continuity we may assume that F is strictly positive on
the torus. Let F = P/Q and assume that F (z) > 0 for every z ∈ TI . Now, let
z(cid:19).
G (z) = F(cid:18) 1
We see that G is analytic, except a finite set of complex numbers. Since F is real
on the torus, we have G(eit) = F (eit) = F (eit) for every t ∈ RI and the analytic
extension principle implies that
F (z) =
P (z)
Q (z)
= G(z) =
,
P (cid:0) 1
z(cid:1)Q(cid:0) 1
z(cid:1)
z(cid:19)Q (z)
P (z) Q(cid:18) 1
z(cid:19) = P(cid:18) 1
α(cid:1) = 0. Then P can be written as
necessarily P(cid:0) 1
except for a finite set in CI .
(2.1) implies that if P (α) = 0, with α 6= 0 then
P (z) = c1 zm0 (z − α1)m1 ... (z − αp)mp (1 − α1z)m1 ... (1 − αpz)mp
with a constant c1. With the same argument as before, Q can be written as
Q (z) = c2 zd0 (z − β1)d1 ... (z − βq)dq (cid:0)1 − β1z(cid:1)d1 ...(cid:0)1 − βqz(cid:1)dq
with a constant c2. Since P and Q are coprime, we must have m0 = 0 or d0 = 0.
thus
(2.1)
Then
where P1 (z) =Qp
P1(cid:0)eit(cid:1)
Q1 (eit)2
j=1 (z − βj)dj therefore we have
F(cid:0)eit(cid:1) = F(cid:0)eit(cid:1) = c
j=1 (z − αj)mj and Q1 (z) =Qq
j=1(cid:0)eit − αj(cid:1)mj(cid:0)e−it − αj(cid:1)mj
F (cid:0)eit(cid:1) = c Qp
j=1 (eit − βj)dj(cid:0)e−it − βj(cid:1)dj
Qq
with a constant c. Let m = m1 + ··· + mp, d = d1 + ··· + dq and ϕ(z) = zrψ(z)
where r = max(0, m − d + 1). Now,
q
= c eirt
j=1(cid:0)eit − αj(cid:1)mj(cid:0)e−it − αj(cid:1)mj
ϕ(cid:0)eit(cid:1) F (cid:0)eit(cid:1) = c eirtψ(cid:0)eit(cid:1)Qp
j=1 (eit − βj)dj(cid:0)e−it − βj(cid:1)dj
Qq
j=1(cid:0)eit − αj(cid:1)mj(cid:0)1 − αjeit(cid:1)mj e−imt
1 − βjeit(cid:19)dj Qp
j=1 (eit − βj)dj(cid:0)1 − βjeit(cid:1)dj e−idt
Qq
j=1(cid:0)eit − αj(cid:1)mj(cid:0)1 − αjeit(cid:1)mj
j=1(cid:0)1 − βjeit(cid:1)2dj
Qq
j=1(cid:0)eit − αj(cid:1)mj(cid:0)1 − αjeit(cid:1)mj
j=1(cid:0)1 − βjeit(cid:1)2dj
Qq
Yj=1(cid:18) eit − βj
= c ei(d−m)teirt Qp
= c eit max(d−m,1) Qp
ON THE NUMERICAL RADIUS OF THE TRUNCATED ADJOINT SHIFT
5
0. It follows from Lemma 2.2 that we have F = f2
which implies that ϕF ∈ HI 1
with a suitable f ∈ H(ϕ). Then for any integer k, we get
ck = < F, eikt > = < f f , eikt >
= < f e−ikt, f >
= < (S∗(ϕ))kf, f >
= < S∗k(ϕ)f, f > .
Therefore
ck ≤ kfk2
2 ω2(S∗k(ϕ)) = kFk1 ω2(S∗k(ϕ)) = c0 ω2(S∗k(ϕ)).
The proof is now complete.
(cid:3)
Corollary 2.3 (Egerváry and Szász [4]). Let F (eit) =Pn−1
trigonometric polynomial (n ≥ 2). Then
k (cid:3) + 2! f or 1 6 k 6 n − 1.
(cid:2) n−1
ck6 c0 cos
π
j=−n+1 cjeijt be a positive
Proof. We have :
F (eit) = c−n+1ei(−n+1)t + ··· + c0 + ··· + cn−1ei(n−1)t
= e(−n+1)it(cid:16)c−n+1 + ··· + c0ei(n−1)t + ··· + cn−1e2i(n−1)t(cid:17)
=
P (eit)
Q(eit)
where P (eit) = c−n+1 + ··· + c0ei(n−1)t + ··· + cn−1e2i(n−1)t and Q(eit) = ei(n−1)t.
In this case we have m = n − 1, d = 0 and ϕ(z) = zn. Then Theorem 2.1 implies
that
ck ≤ c0 ω2(S∗
n
k).
But generally S∗
n
dimension, the largest dimension being s(k, n) + 1 where s(k, n) = [ n−1
ω2(S∗
n
k is unitarily equivalent to an orthogonal sum of shifts of smaller
k ]. Therefore
. The same computation follows from [8].
k) = ω2(S∗
s(k,n)+1) = cos
π
s(k, n) + 2
Finally, this implies that
π
s(k, n) + 2
ck 6 c0 cos
= c0 cos
π
k (cid:3) + 2! .
(cid:2) n−1
(cid:3)
Remark 2.1. The bound for c1 is due to Fejer (1915).
6
HAYKEL GAAYA
3. The numerical radius of the shift compression
3.1. Preliminaries. The spectral decomposition of the KMS matrix
Kn(α) =
1
α
...
αn−1
α
. . .
. . .
···
··· αn−1
...
. . .
. . .
α
α
1
= (αr−s)n
r,s=1 ; 0 6 α < 1
is very well understood in the computational sense. For this reason, these matrices
are often used as test matrices. It's shown in [7] page 69 -- 72 that Kn(α) is a Toeplitz
matrix associated with the Poisson kernel Pα(eit) =
1 − α2
1 − αeit2 and its eigenvalues
are:
where t(n)
k
are the solutions of
λ(n)
k = Pα(eit
(n)
k ) , 1 ≤ k ≤ n
(3.1)
pn(cos t) =
sin(n + 1)t − 2α sin nt + α2 sin(n − 1)t
sin t
= 0.
The expression pn(cos t) is a polynomial of degree n in cos t and it has n real distinct
zeros cos t(n)
k
for 1 6 k 6 n where :
1 < t(n)
0 < t(n)
2 < t(n)
3 < ··· < t(n)
n < π .
This implies that
1 + α
1 − α
> λ(n)
1 > λ(n)
2 > λ(n)
3 > ··· > λ(n)
n >
1 − α
1 + α
.
The evaluation of the zeros t(n)
k
it is easy to show that they are separated by the quantities xk =
in explicit terms seems to be out of end. However,
, 1 ≤ k ≤ n.
n + 1
kπ
Indeed, for 1 6 k 6 n
and
pn(cos xk) = (−1)k2α(1 − α cos xk)
Also we see by direct substitution that the latter equation holds for k = 0, so that
sgn pn(cos xk) = (−1)k .
0 < t(n)
1 6 x1 < t(n)
2 6 x2 < ··· < t(n)
n 6 xn < π .
Remark 3.1. In the case where α = 0 in (3.1) we have t(n)
k = xk.
In the next proposition we give a better lower and upper bound for tn
1 .
Proposition 3.1. For each integer n ≥ 2;
(3.2)
arccos(α) 6 t(n)
1 6 arccos(α) .
2
n + 1
Proof. First, we note that
(3.3)
pn(t) = 2
Since for all 0 < t 6 π
sin t(cid:16)sin (n+1)t
n+1 , we have
2 − α sin (n−1)t
2 (cid:17)(cid:16)cos (n+1)t
2 (cid:17) .
2 − α cos (n−1)t
(3.4)
0 < (n−1)t
2 < (n+1)t
2
6 π
2
ON THE NUMERICAL RADIUS OF THE TRUNCATED ADJOINT SHIFT
7
this implies that
then t(n)
1
is zero of
(3.5)
α sin (n−1)t
2 < sin (n−1)t
2 < sin (n+1)t
2
cos (n+1)t
.
2
2 = α cos (n−1)t
n+1 arccos(α) then (n+1)t
which contradicts(3.5).
2
(n)
1
< arccos(α) and
Now if we suppose that t(n)
cos (n+1)t(n)
Hence t(n)
> α ≥ α cos (n−1)t(n)
n+1 arccos(α) holds.
2
2
1 ≥ 2
1
1
1 < 2
From (3.5), we have
cos t cos (n−1)t
2 − sin t sin (n−1)t
2 = α cos (n−1)t
2
which implies that
while from (3.4), sin (n−1)t
positive, which completes the proof.
2
(cos t − α) cos (n−1)t
and cos (n−1)t
2
2 = sin t sin (n−1)t
2
are both positive, therefore cos t − α is
(cid:3)
Remark 3.2. For a fixed 0 6 α < 1, since Pα(eit) is positive and decreasing on
the interval [0, π], then it is easy to obtain the sharp lower and upper bound of the
largest eigenvalues of Kn(α):
1 6 λ(n)
1 6
1 − 2α cos(cid:16) 2
1 − α2
n+1 arccos(α)(cid:17) + α2
.
Note that λ(n)
is also the numerical radius of Kn(α). This is due to the fact that
the norm and numerical radius of a symmetric matrix coincides with its largest
eigenvalue.
1
For 0 6 α < 1, we denote by
1 − α)t
1 + α2
; −α(cos t(n)
1 − 2α cos t(n)
n + α2) ;
n − α)
1 − 2α
1 + α2
1 + 2α3
1 + α2
, −(1 + α2) cos t(n)
1 − 2α cos t(n)
n + 2α
n + α2 )
, −(1 + α2) cos t(n)
1 − 2α cos t(n)
n + 2α
n + α2 ) .
and
sn(α) = max( α(cos t(n)
1 − 2α cos t(n)
mn(α) = max((1 + α2) cos t(n)
1 − 2α cos t(n)
Mn(α) = max( (1 − 3α2) cos t(n)
1 − 2α cos t(n)
Jn(α) =
0
...
...
0
α
. . .
. . .
Proposition 3.2. For each integer n ≥ 2 and 0 6 α < 1, let
··· αn−1
...
. . .
. . .
. . .
α
0
;
then we have
ω2(Jn(α)) = ω2(Re(Jn(α))) = sn(α).
8
HAYKEL GAAYA
Proof. First observe that
ω2(Re(Jn(α))) =
=
=
=
=
sup
,kuk=1< Re(Jn(α))u, u >
αl−mulum(cid:12)(cid:12)(cid:12)
αl−mulum
αl−mulum
1
2
1
2
u=(u0,··· ,un−1)∈ CI n
1
2
sup
sup
Pn−1
Pn−1
l=0 ul2=1(cid:12)(cid:12)(cid:12) X1≤m6=l≤n−1
l=0 ul2=1 X1≤m6=l≤n−1
2 X1≤m<l≤n−1
l=0 ul2=1(cid:12)(cid:12)(cid:12) X1≤m<l≤n−1
l=0 ul2=1
Pn−1
sup
sup
Pn−1
αl−mulum(cid:12)(cid:12)(cid:12)
α(cos t − α)
1 − 2α cos t + α2 = g(t).
= ω2(Jn(α)).
We note that Re(Jn(α)) is the Toeplitz matrix associated with the Toeplitz form
1
2
(Pα(eit) − 1) =
To complete the proof of the proposition, we can easily observe that if a and b
are arbitrary real number and f (x) a Toeplitz form with γn
k as eigenvalues then
the eigenvalues of a + bf (x) will be a + bγn
k . This shows that the eigenvalues of
Re(Jn(α)) are
λ′(n)
k =
1
2
(λ(n)
k − 1) =
k − α)
α(cos t(n)
1 − 2α cos t(n)
1 + α2
, 1 ≤ k ≤ n.
Now, since g(t) is decreasing on the interval [0, π] and Re(Jn(α) is symmetric then
ω2(Jn(α)) = ω2(Re(Jn(α)))
k
; 1 6 k 6 no
= maxnλ′(n)
= maxnλ′(n)
n o
;−λ′(n)
= max( α(cos t(n)
1 − α)
1 − 2α cos t(n)
1
1 + α2
= sn(α)
(3.6)
; −α(cos t(n)
1 − 2α cos t(n)
n + α2)
n − α)
where (3.6) is due to the fact that cos t(n)
3.1.
n
is nonpositive and by using Proposition
(cid:3)
Corollary 3.3. For 0 6 α < 1, we have
ω2(J2(α)) =
α
2
.
Proof. This result is known, but it is interesting to notice that this result can also
easily be obtained by using Proposition 3.2. Indeed
p2(t) =
sin(3t) − 2α sin(2t) + α2 sin t
sin t
= 4 cos2 t − 4α cos t + α2 − 1
Therefore, we obtain: cos t(2)
then we have λ′(2)
1 =
2 = −
α
2
α + 1
2
and λ′(2)
1 =
α
2
. Now since T r(J2(α)) = 0
. This completes the proof.
(cid:3)
ON THE NUMERICAL RADIUS OF THE TRUNCATED ADJOINT SHIFT
9
Corollary 3.4. For 0 6 α < 1, we have
ω2(J3(α)) =
α(√α2 + 8 − 3α)
4 + 2α2 − 2α√α2 + 8
.
Proof. We have
p3(t) =
2
sin t(cid:0) sin(2t) − α sin t(cid:1)(cid:0) cos(2t) − α cos t(cid:1)
α + √α2 + 8
, cos t(3)
this implies that cos t(3)
then
1 =
4
λ′(3)
1 =
1 − α)
1 + α2
α(cos t(3)
1 − 2α cos t(3)
= −α(cos t(3)
1 − 2α cos t(3)
3 − α)
3 + α2
=
=
and
λ′(3)
3
The proof is complete.
3 =
α
2
2 =
and cos t(3)
α(√α2 + 8 − 3α)
4 + 2α2 − 2α√α2 + 8
α(√α2 + 8 + 3α)
4 + 2α2 + 2α√α2 + 8
α − √α2 + 8
4
.
(cid:3)
Corollary 3.5. For each integer n ≥ 4 and α 6qcos 2π
n+1 , we have
ω2(Jn(α)) =
1 − α)
α(cos t(n)
1 − 2α cos t(n)
1 + α2
.
Proof. It follows from the hypothesis of Corollary 3.5 that
t(n)
1 ≤
(3.7)
then
and
π
n + 1 ≤ arccosr 1 + α2
2 ≤ arccos(α) 6 t(n)
n 6
nπ
n + 1
λ′(n)
1 =
1 − α)
α(cos t(n)
1 − 2α cos t(n)
1 + α2
λ′(n)
n = −λ′(n)
n = −α(cos t(n)
1 − 2α cos t(n)
n − α)
n + α2
In view of the inequality (3.7), g(arccos α) = 0 and the fact that g is decreasing in
[0, π], it suffices to prove that g( π
g(
π
n + 1
) − g(
nπ
n + 1
) =
=
=
n+1 ) ≥ g( nπ
α(cos π
1 − 2α cos π
n+1 ). We have
n+1 − α)
n+1 + α2 −
2α2(2 cos2 π
α(cos π
n+1 + α)
n+1 + α2
1 + 2α cos π
n+1 − α2 − 1)
n+1 + α2)(1 + 2α cos π
2α2(cos 2π
n+1 + α2)(1 + 2α cos π
n+1 − α2)
n+1 + α2)
n+1 + α2)
(1 − 2α cos π
(1 − 2α cos π
≥ 0.
The result follows.
(cid:3)
10
HAYKEL GAAYA
3.2. The numerical radius of the compressed shift. In this section, we will
focus on the particular case where
First, we notice some properties for the general case where φ is a finite Blashke
product : φ(z) = Qn
evaluation functional kλ ∈ HI 2 by the requirement that f (λ) =< f, kλ >. Thus
. For each λ in the unit disc, we define the
j=1
.
1 − αz(cid:19)n
φ(z) =(cid:18) z − α
z − αj
1 − αjz
kλ(z) =
1
1 − λz
and {e1, . . . , en} the collection of functions of H(φ) defined as follows :
and
e1(z) =(cid:0)1 − α12(cid:1)
ek(z) =(cid:0)1 − αk2(cid:1)
for any k = 2, ..., n.
1
2
1
k−1
1 − α1z
Yj=1
1
2
1
1 − αkz
z − αj
1 − αjz
It is known that {e1, . . . , en} is an orthonormal basis of H(φ) and with respect to
this basis the matrix of S∗(φ) is given by [alk], where
αl
σlσl+1
σlσkQk−1
0
j=l+1(−αj)
if l = k
if k = l + 1
if k > l + 1
unless
alk =
1
2 , for each 1 6 k 6 n.
0
Indeed, for k > l + 1, we have
and σk =(cid:0)1 − αk2(cid:1)
< S∗(φ)ek, el > = σkσlZ 2π
= σkσlZ 2π
= σkσlZ 2π
Yj=l+1
= σkσl
k−1
0
0
(−αj).
1
e−iθ
1 − αkeiθ
1 − αle−iθ
1
e−iθ
1 − αkeiθ
1 − αle−iθ
1
1
1 − αkeiθ
1 − αleiθ
dθ
2π
k−1
k−1
eiθ − αj
1 − αjeiθ
Yj=l+1
eiθ − αj
1 − αjeiθ
Yj=l
eiθ − αl
1 − αleiθ
Yj=l+1
k−1
eiθ − αj
1 − αjeiθ
dθ
2π
dθ
2π
Using the same scheme as before,we prove easily that < S∗(φ)ek+1, ek >= σkσk+1
and < S∗(ek), el >= 0 if k < l.
Finally
< S∗(φ)ek, ek > = σ2
kZ 2π
0
1
e−iθ
1 − αkeiθ
z
1 − αkz
1 − αke−iθ
>
dθ
2π
= σ2
k < Kαk ,
= αk
ON THE NUMERICAL RADIUS OF THE TRUNCATED ADJOINT SHIFT
11
In the sequel of the paper, φ denotes the finite Blashke product with unique zero
α:
S∗(φ) gets the following matricial representation:
.
1 − αz(cid:19)n
φ(z) =(cid:18) z − α
···
σ(−α)n−2
σ
. . .
. . .
σ −σα ···
. . .
α
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
α
0
α
0
...
...
...
0
...
...
−σα
σ
α
where σ = 1 − α2.
Proposition 3.6. For 0 6 α < 1, we have
ω2(Re(S∗(φ))) = mn(α).
Proof. First, notice that where α = 0, then
Re(S∗(φ)) =
1
2
1
0
1
0
. . .
0
1
0
1
. . .
0
0
1
0
. . .
. . .
. . .
. . .
. . .
. . .
.
0
1
0
0
. . .
kπ
n + 1
In this case the eigenvalues are cos
, for k = 1, . . . , n. For the proof there
are many references, we refer the reader for example to [7] page 67 or [2] page 35,
therefore ω2(Re(S∗(φ))) = cos
= mn(0). Then we can limit our study to the
case α 6= 0. Now, notice that
n + 1
π
Re(S∗(φ)) =
1 − α2
2α
−
2α2
1 − α2
α
...
αn−1
α
. . .
. . .
···
···
. . .
. . .
α −
αn−1
...
α
2α2
1 − α2
.
Here Re(S∗(φ)) is the Toeplitz matrix associated with the Toeplitz form:
1 − α2
2α
(Pα(eit) − 1 −
2α2
1 − α2 ) =
(1 + α2) cos t − 2α
1 − 2α cos t + α2 = h(t).
Since h(t) is monotonic on [0, π], thus with the same argument that in the proof of
Proposition 3.2, we may assume that:
ω2(Re(S∗(φ))) = maxn (1+α2) cos tn
1−2α cos tn
= mn(α).
1 −2α
1 +α2
, −(1+α2) cos tn
1−2α cos tn
n+2α
n+α2 o
This completes the proof.
(cid:3)
12
HAYKEL GAAYA
1 − αz(cid:19)n
Theorem 3.7. Let φ(z) =(cid:18) z − α
with α ∈ CI and α< 1.
(1) The numerical radius of S∗(φ) is independent from the argument of α and
for 0 6 α < 1 the numerical range of S∗(φ) is symmetric with respect to
the real axis.
(2) For n ≥ 2 , we have
mn(α) 6 ω2(S∗(φ)) 6 Mn(α).
Proof. For α 6= 0, and for t = arg(α), we have:
ω2(S∗(φ)) =
=
=
=
=
(3.8)
sup
kuk2=1 < S∗(φ)u, u >
1 − α2
Pn−1
Pn−1
Pn−1
sup
sup
α −
l=0 ul2=1(cid:12)(cid:12)(cid:12)
l=0 ul2=1(cid:12)(cid:12)(cid:12)α −
l=0 ul2=1(cid:12)(cid:12)(cid:12)α −
l=0 vl2=1(cid:12)(cid:12)(cid:12)α −
sup
sup
Pn−1
α
1 − α2
α
1 − α2
α
1 − α2
α
X1≤m<l≤n−1
X1≤m<l≤n−1
X1≤m<l≤n−1
X1≤m<l≤n−1
(−α)l−m ulum(cid:12)(cid:12)(cid:12)
(−α)l−m ulum(cid:12)(cid:12)(cid:12)
(−α)l−m eiltuleimtum(cid:12)(cid:12)(cid:12)
(−α)l−m vlvm(cid:12)(cid:12)(cid:12)
.
The last equality implies that the numerical radius of S∗(φ) is independant from
the argument of α. Hence we can suppose that 0 < α < 1. Now assume that z is
in W (S∗(φ)), then there is u = (u0, . . . , un−1) a unit vector in CI n such that
z = < S∗(φ)u, u >
= α −
1 − α2
α
X1≤m<l≤n−1
(−α)l−m ulum
and
z = α −
1 − α2
α
X1≤m<l≤n−1
(−α)l−m umul
= < S∗(φ)u, u > .
This implies that z is in W (S∗(φ)) and (1) holds.
As remarked before, we can restrict our study to the case where 0 < α < 1 and
from (3.8) we have:
ω2(S∗(φ)) ≤ α +
1 − α2
α
1 − α2
= α +
α
= Mn(α).
Pn−1
sup
l=0 ul2=1(cid:12)(cid:12)(cid:12) X1≤m<l≤n−1
αl−mulum(cid:12)(cid:12)(cid:12)
ω2(Jn(α))
On the other hand, it is obvious to note that ω2(ReT ) 6 ω2(T ) for each bounded
operator T , an application of the Proposition 3.6 completes the proof of (2).
(cid:3)
ON THE NUMERICAL RADIUS OF THE TRUNCATED ADJOINT SHIFT
13
Corollary 3.8. For n = 2 and 0 6 α < 1; we have:
1 + 2α − α2
ω2(S∗(φ)) =
2
.
Corollary 3.9. For each integer n ≥ 2;
ω2(S∗
n) = cos
π
n + 1
.
In the last two corollaries, the results are due to the fact that both quantities mn
and Mn coincide.
Acknowledgements: The author would like to express his gratitude to Gilles Cassier
for his help and his good advices and Alfonso Montes Rodriguez as well for advices
in the translation of this work..
References
[1] C. Badea and G. Cassier, Constrained von Neumann inequalities, Adv. Math. 166 (2002), no.
2, 260 -- 297.
[2] A. Böttcher and S. Grudsky, Spectral properties of banded Toeplitz matrices. Society for
Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2005.
[3] G. Cassier, I. Chalendar, and B. Chevreau, New examples of contractions illustring member-
ship and non membership in The classe An,m, Acta Sci. Math (Szeged) 64 (1998), 701 -- 731.
[4] E.V. Egerváry and O. Szász, Einige Extremalprobleme in Bereiche der trigonometrischen
Polynomen, Math. Z. 27 (1928), 641-652.
[5] H. L. Gau and P. Y. Wu, Numerical range and Poncelet property, Taiwanese J. Math. 7
(2003), no. 2, 173 -- 193.
[6] H. L. Gau and P. Y. Wu, Numerical range of S(φ). Linear and Multilinear Algebra 45 (1998),
no. 1, 49 -- 73.
[7] U. Grenander and G. Szegö, Toeplitz forms and their applications. California Monographs in
Mathematical Sciences University of California Press, Berkeley-Los Angeles 1958.
[8] K.E. Gustafson, D.K.M. Rao: Numerical Range, Springer, New York, 1997.
[9] U. Haagerup and P. de la Harpe, The numerical radius of a nilpotent operator on a Hilbert
space, Proc. Amer. Math. Soc. 115(1992), 371 -- 379.
[10] P. R. Halmos, A Hilbert space problem book. Second edition. Graduate Texts in Mathematics,
19. Encyclopedia of Mathematics and its Applications, 17. Springer-Verlag, New York-Berlin,
1982.
[11] R. A. Horn and C. R. Johnson, Topics in matrix analysis. Cambridge University Press,
Cambridge, 1991.
[12] M. Kac, W. L. Murdock and G. Szegö, On the eigenvalues of certain Hermitian forms. J.
Rational Mech. Anal. 2, (1953). 767 -- 800.
[13] B. Sz.-Nagy and C. Foias, Harmonic analysis of operator on Hilbert space, North-Holland
Publishing Co., Amesterdam-London; Americain Elsevier Publishing Co., Inc., New York;
Akadémiai Kiado, Budapest, 1970.
[14] W. F. Trench, Interlacement of the even and odd spectra of real symmetric Toeplitz matrices.
Linear Algebra Appl. 195 (1993), 59 -- 68.
‡.Institute Camille Jordan, Office 107 University of Lyon1, 43 Bd November 11,
1918, 69622-Villeurbanne, France.
E-mail address: ‡[email protected]
|
1001.1516 | 1 | 1001 | 2010-01-10T13:46:35 | Continuous Shearlet Tight Frames | [
"math.FA"
] | Based on the shearlet transform we present a general construction of continuous tight frames for $L^2(\mathbb{R}^2)$ from any sufficiently smooth function with anisotropic moments. This includes for example compactly supported systems, piecewise polynomial systems, or both. From our earlier results it follows that these systems enjoy the same desirable approximation properties for directional data as the previous bandlimited and very specific constructions due to Kutyniok and Labate. We also show that the representation formulas we derive are in a sense optimal for the shearlet transform. | math.FA | math |
Continuous Shearlet Tight Frames
Philipp Grohs∗
Abstract
Based on the shearlet transform we present a general construction of
continuous tight frames for L2(R2) from any sufficiently smooth function
with anisotropic moments. This includes for example compactly sup-
ported systems, piecewise polynomial systems, or both. From our earlier
results in [5] it follows that these systems enjoy the same desirable ap-
proximation properties for directional data as the previous bandlimited
and very specific constructions due to Kutyniok and Labate; [8]. We also
show that the representation formulas we derive are in a sense optimal for
the shearlet transform.
Contents
1 Introduction
1.1 Shearlets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2 The construction
Ingredients
2.1
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2 Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3 Concluding Remarks
4 Acknowledgments
1
Introduction
1
2
3
4
7
9
10
The purpose of this short paper is to give a construction of a system of bivariate
functions which has the following desirable properties:
directionality. The geometry of the set of singularities of a tempered
distribution f can be accurately described in terms of the interaction be-
tween f and the elements of the system.
tightness. The system forms a tight frame of L2(R2).
locality. The representation is local. By this we mean that the represen-
tation can also be interpreted as a representation with respect to a non
tight frame and its dual frame such that both of these frames only consist
of compactly supported functions.
∗TU Wien,
strasse
http://www.dmg.tuwien.ac.at/grohs, phone: +43 58801 11318.
8-10,
Institute of Discrete Mathematics and Geometry, Wiedner Haupt-
1040 Wien, Austria.
[email protected], web:
Email:
1
The construction is based on the shearlet transform which has been introduced
in [9] and has become popular in Computational Harmonic Analysis for its
ability to sparsely represent bivariate functions. A related important result
is that the coefficients of a general tempered distribution with respect to this
transform exactly characterize the Wavefront Set (see e.g. [7] for the definition)
of this distribution [8].
Notation. We shall use the symbol · indiscriminately for the absolute value
on R, R2, C and C2. We usually denote vectors in R2 by x, t, ξ, ω and their
elements by x1, x2, t1, t2, . . . .
In general it should always be clear to which
space a variable belongs. The symbol k · k is reserved for various function space
and operator norms. For two vectors s, t ∈ R2 we denote by st their Euclidean
inner product. We use the symbol f . g for two functions f, g if there exists
a constant C such that f (x) ≤ Cg(x) for large values of x. We will often
speak of frames. By this we mean continuous frames as defined in [1]. We
define f (ω) = R f (x) exp(2πiωx)dx to be the Fourier transform for a function
f ∈ L1 ∩ L2 and continuously extend this notion to tempered distributions.
1.1 Shearlets
We start by defining what a shearlet is and what the shearlet transform is:
Definition 1.1. A function ψ ∈ L2(R2) is called shearlet if it possesses M ≥ 1
vanishing moments in x1-direction, meaning that
ZR2
ψ(ω)2
ω12M dω < ∞.
Let f ∈ L2(R2). The shearlet transform of f with respect to a shearlet ψ maps
f to
where
SHψf (a, s, t) := hf, ψasti,
ψast(x) := ψ(cid:0)
x1 − t1 − s(x2 − t2)
a
,
x2 − t2
a1/2 (cid:1).
The following reproduction formula holds [4]:
Theorem 1.2. For all f ∈ L2(R2) and ψ a shearlet
f (x) = ZR2 ZRZR+ SHψ(a, s, t)ψast(x)a−3dadsdt,
(1)
where equality is understood in a weak sense.
The shearlet transform captures local, scale- and directional information via
the parameters t, a, s respectively. A significant drawback of this representation
is the fact that the directional parameter runs over the non compact set R. Also
it is easy to see that the distribution of directions becomes infinitely dense as
s grows. These problems led to the construction of shearlets on the cone [8].
The idea is to restrict the shear parameter s to a compact interval. Since this
only allows to caption a certain subset of all possible directions, the function
f is split into f = P f + P νf , where P is the frequency projection onto the
cone with slope ≤ 1 and only P f is analyzed with the shearlet ψ while P νf
2
is analyzed with ψν(x1, x2) := ψ(x2, x1). For very specific choices of ψ Labate
and Kutyniok proved a representation formula
2 = ZR2 hf, TtWi2dt +ZR2 Z 2
kfk2
−2Z 1
0 SHψP f (a, s, t)2a−3dadsdt
+ZR2 Z 2
−2Z 1
0 SHψν P νf (a, s, t)2a−3dadsdt,
(2)
Tt denoting the translation operator by t. The main weakness of this decompo-
sition is its lack of locality: indeed, first of all the need to perform the frequency
projection P to f destroys any locality. But also the functions ψ which have
been considered in [8] are very specific bandlimited functions which do not
have compact support. As a matter of fact no useful local representation via
compactly supported functions which is able to capture directional smoothness
properties has been found to date. The present paper aims at providing a step
towards finding such a representation using two crucial observations: First, in
[5] we were able to show that the description of directional smoothness via the
decay rate of the shearlet coefficients for a → 0 essentially works for any func-
tion which is sufficiently smooth and has sufficiently many vanishing moments
in the first direction, hence also for compactly supported functions. The second
observation is that actually a full frequency projection P is not necessary to ar-
rive at a useful representation similar to (2). Instead of the operator P we will
use a 'localized frequency projection' which is given by Fourier multiplication
with a function p0 to be defined later.
Our main result Theorem 2.4 will prove a representation formula similar to
(2) where ψ is allowed to be compactly supported and the frequency projection
is replaced with a local variant. We also show that this local variant of the
frequency projection is the best we can do -- without it, no useful continuous
tight frames (or continuous frames with a structured dual) can be constructed
within the scope of the shearlet transform. This is shown in Theorem 2.6.
2 The construction
The goal of this section is to derive a representation formula
kfk2
2 =
0 hf, q0 ∗ ψasti2a−3dadsdt
1
Cψ(cid:0)ZR2 Z 2
−2Z 1
−2Z 1
0 hf, q1 ∗ ψν
+ ZR2 Z 2
+ ZR2 hf, Ttϕi2dt(cid:1)
asti2a−3dadsdt
(3)
for L2(R2)-functions f and with some localized frequency projections q0, q1 and
a window function ϕ to be defined later. In order to guarantee that the shearlet-
part contains the high-frequency part of f and the rest contains low frequencies,
it is necessary to ensure that the window function ϕ in this formula is sufficiently
smooth.
3
2.1
Ingredients
Here we first state the definitions and assumptions that we use in the construc-
tion. Then we collect some auxiliary results which we later combine to prove our
main results Theorems 2.4 and 2.6. The large part of the results will concern
the construction of a useful window function and to ensure its smoothness. We
say that a bivariate function f has Fourier decay of order Li in the i-th variable
(i ∈ {1, 2}) if
f (ξ) . ξi−Li.
We start with a shearlet ψ which has M vanishing moments in x1-direction
and Fourier decay of order L1 in the first variable. It is clear from the definition
θ with θ ∈ L2(R2). We assume that
θ has Fourier decay of order L2 in the second variable so that the following
relation holds:
of vanishing moments that ψ = (cid:0) ∂
∂x1(cid:1)M
We also set
and
2M − 1/2 > L2 > M ≥ 1.
N := 2 min(L2 − M, L1),
Cψ := ZR2
ψ(ω)2
ω12 dω,
∆ψ(ξ) := Z 2
−2Z 1
0 ψ(aξ1, a1/2(ξ2 − sξ1))2a−3/2dads.
We define functions ϕ0, ϕ1 via
(4)
(5)
(6)
(7)
(8)
ϕ0(ξ)2 = Cψ − ∆ψ(ξ)
and
ϕ1(ξ)2 = Cψ − ∆ψν (ξ),
where
ψν (x1, x2) := ψ(x2, x1).
We write χC for the indicator function of the cone C = {ξ = (ξ1, ξ2) ∈ R2 :
ξ2 ≤ ξ1}. Finally, we pick a smooth and compactly supported bump function
Φ with Φ(0) = 1 and define functions p0, p1 via
Clearly, p0 and p1 are both compactly supported tempered distributions.
p0 = Φ ∗ χC
and p1 = 1 − p0.
Lemma 2.1. We have
p0(ξ) . ξ−N for ξ1/ξ2 ≥ 3/2
p1(ξ) . ξ−N for ξ2/ξ1 ≥ 3/2
(9)
(10)
Proof. Assume that ξ = te with e a unit vector with e1/e2 ≥ 3/2 and t > 0.
There exists a uniform δ > 0 such that for all η with η < δt we have ξ− η ∈ Cc,
and hence χC(ξ − η) = 0. It follows that we can write
p0(ξ) = ZR2
χC(ξ − η) Φ(η)dη ≤ Zη>δt Φ(η)dη
. t−N = ξ−N
4
if Φ is sufficiently smooth. On the other hand, let ξ = te with e a unit vector
with e2/e1 ≥ 3/2 and t > 0. There exists a uniform δ > 0 such that for all η
with η < δt we have ξ − η ∈ C and hence χC(ξ − η) = 1. Now we can estimate
p1(ξ) = 1 − p0(ξ) = ZR2
Φ(η)(1 − χC(ξ − η))dη
= Zη>δt
Φ(η)(1 − χC(ξ − η))dη . t−N = ξ−N
again for Φ smooth. Note that in the first equality we have used that Φ(0) = 1.
This proves the statement.
Lemma 2.2. We have
ϕ0(ξ)2 . ξ−N for ξ1/ξ2 ≤ 3/2
ϕ1(ξ)2 . ξ−N for ξ2/ξ1 ≤ 3/2
(11)
Proof. This follows from [5, Lemma 4.7]. For the convenience of the reader we
present a proof here as well. We only prove the assertion for ϕ0 since the proof
of the corresponding statement for ϕ1 is the same. By definition we have
ϕ0(ξ)2 = (cid:0)Za∈R, s>2 ψ(cid:0)aξ1,√a(ξ2 − sξ1)(cid:1)2a−3/2dads
+Za>1, s<2 ψ(cid:0)aξ1,√a(ξ2 − sξ1)(cid:1)2a−3/2dads(cid:1).
We start by estimating the second integral using the Fourier decay in the first
variable:
Za>1, s<2 ψ(cid:0)aξ1,√a(ξ2 − sξ1)(cid:1)2a−3/2dads . 4Za>1
(aξ1)−2L1 a−3/2da
. ξ1−2L1 . ξ−2L1
for all ξ in the cone with slope 3/2. To estimate the other term we need the
θ(ξ)
moment condition and the decay in the second variable. We write ψ(ξ) = ξM
1
and ξ = (ξ1, rξ1), r ≤ 3/2. We start with the high frequency part:
Za<1, s>2 ψ(cid:0)aξ1,√a(ξ2 − sξ1)(cid:1)2a−3/2da
= Za<1, s>2 aξ12Mθ(cid:0)aξ1,√a(ξ2 − sξ1)(cid:1)2a−3/2dads
. Za<1, s>2 aξ12M√a(ξ2 − sξ1)−2L2a−3/2dads
= Za<1, s>2
a2M−L2−3/2ξ12M−2L2r − s−2L2dads
we now use that r − s is always strictly away from zero. By assumption
L2 = 2M − 1/2 − ε for some ε > 0. Hence we can estimate further
. . . = ξ1−2(L2−M)Za<1, s>2
a−1+εr − s−2L2 dads . ξ−2(L2−M).
5
The low-frequency part can simply be estimated as follows:
Za>1, s>2 ψ(cid:0)aξ1,√a(ξ2 − sξ1)(cid:1)2a−3/2dads
a−3/2−L2r − s−2L2dads
. ξ1−2L2 Za>1, s>2
. ξ−2L2.
Putting these estimates together proves the statement.
Lemma 2.3. Assume that ψ is compactly supported with support in the ball
BA := {ξ : ξ ≤ A}. Then the tempered distributions ( ϕi2)∨, i = 0, 1 are both
of compact support with support in the ball 2p3 + √5BA.
Proof. Since the inverse Fourier transform of the constant function Cψ is a
Dirac, this follows if we can establish that the functions ∆ψ and ∆ψν are Fourier
transforms of distributions of compact support. We show this only for ∆ψ, the
In what follows we will use the notation f−(x) :=
other case being similar.
f (−x) for a function f . Since ψ has M anisotropic moments we can write ψ =
∂x1(cid:1)M
(cid:0) ∂
θ for some θ ∈ L2(R2) with the same support as ψ. Consider the function
Θ(ξ) := θ(ξ)2. It is easy to see that this is the Fourier transform of the so-called
Autocorrellation function θ∗ θ− of θ which is compactly supported with support
in B2A. Therefore, by the theorem of Paley-Wiener, the function Θ possesses
an analytic extension to C2 which we will henceforth call F . Furthermore, by
the same theorem F is of exponential type:
F (ζ) . (1 + ζ)K exp(2Aℑζ),
(12)
where ζ ∈ C2 and K ∈ N. We now consider the analytic extension Γ of the
function ∆ψ which is given by
Γ(ζ) = Z 2
−2Z 1
0
(aζ1)2M F (aζ1, a1/2(ζ2 − sζ1))a−3/2dads,
ζ ∈ C2.
a
0
Since M ≥ 1 by (4) the above integral is locally integrable which implies that
Γ is actually an entire function. It is also of exponential type: Writing Mas :=
(cid:18)
−sa1/2 a1/2 (cid:19) a short computation reveals that
kMaskL2(C2)→L2(C2) ≤ a1/2(cid:0)1 +
)1/2(cid:1)1/2
=: a1/2C(s).
+ (s2 +
s2
2
s2
4
Now we estimate
∆ψ(ζ) = Z 2
−2Z 1
0
(aζ1)2M F (Masζ)a−3/2dads . ζ2M (1 + Masζ)K exp(Masℑζ)
(1 + ζ)2M+K exp(2AC(s)ℑζ) . (1 + ζ)2M+K exp(2AC(2)ℑζ).
.
sup
s∈[−2,2]
By the Theorem of Paley-Wiener-Schwartz [7] it follows that ∆ψ is of compact
support with support in B2√3+√5A
.
6
2.2 Main Result
We are now ready to prove our main result, the local representation formula in
Theorem 2.4 which is similar to (2) only with local frequency projections (given
by convolution with p0, p1) and with possibly compactly supported shearlets.
In addition, in Theorem 2.6 we show that in a way this is the simplest repre-
sentation that one can achieve with shearlets -- the local frequency projections
are necessary in order to wind up with useful systems.
From now on we shall assume that 0 ≤ Φ(ξ) ≤ 1 for all ξ ∈ R2 and define
qi(ξ) := pi(ξ)1/2.
Furthermore, we define
ϕ(ξ) := (p0(ξ)ϕ0(ξ)2 + p1(ξ)ϕ1(ξ)2)1/2.
Observe that by the positivity assumption above the radicands in the previous
definitions are nonnegative and real.
Theorem 2.4. We have the representation formulas
kfk2
2 =
0 hf, q0 ∗ ψasti2a−3dadsdt
1
Cψ(cid:0)ZR2 Z 2
−2Z 1
−2Z 1
0 hf, q1 ∗ ψν
+ ZR2 Z 2
+ ZR2 hf, Ttϕi2dt(cid:1)
asti2a−3dadsdt
and
f = f high + f low :=
+ ZR2 Z 2
−2Z 1
0 hf, ψν
0 hf, ψastip0 ∗ ψasta−3dadsdt
1
−2Z 1
Cψ(cid:0)ZR2 Z 2
astip1 ∗ ψasta−3dadsdt(cid:1)
+
1
Cψ(cid:0)ZR2hf, Ttϕ0ip0 ∗ Ttϕ0dt +ZR2hf, Ttϕ1ip1 ∗ Ttϕ1dt(cid:1).
The function f low satisfies
(f low)∧(ξ) . ξ−N
(13)
(14)
(15)
for any f . We have the following locality property: If ψ is compactly supported
in BA and Φ has support in BB, then f low(t) only depends on f restricted to
t + B2√3+√5A+B
Proof. We first prove (14). Taking the Fourier transform of the right hand side
yields
.
1
Cψ(cid:0)p0(ξ)(∆ψ(ξ) + ϕ0(ξ)2) + p1(ξ)(∆ψν (ξ) + ϕ1(ξ)2)(cid:1) f (ξ) = f (ξ).
7
The proof of (13) is similar. We prove the equation (15): Again taking the
Fourier transform of f low gives
1
Cψ(cid:0)p0(ξ) ϕ0(ξ)2 + p1(ξ) ϕ1(ξ)2(cid:1) f (ξ).
Now, the desired estimate follows from Lemmas 2.1 and 2.2. The last statement
follows from the observation that f low can be written as
f low =
1
Cψ
f ∗(cid:0)p0 ∗ ( ϕ02)∨ + p1 ∗ ( ϕ12)∨(cid:1)
together with Lemma 2.3.
Remark 2.5. While we could show that the functions ϕ−i ∗ ϕi and pi, i = 1, 2
are compactly supported, it would be desirable to show that also the functions
ϕi, qi, i = 1, 2 have compact support. We currently do not know how to do this.
By a result of Boas and Kac [2], in the univariate case there always exists for
any compactly supported function g with positive Fourier transform a compactly
supported function f with f ∗ f− = g. However, in dimensions ≥ 2 this holds
no longer true and things become considerably more difficult.
The previous theorem for the first time gives a completely local represen-
tation for square integrable functions which also allows to handle directional
phenomena efficiently: by the results of [5] it follows that the decay rate of the
coefficients hf, ψasti for a → 0 accurately describes the microlocal smoothness
of f at t in the direction with slope s.
It is interesting to ask if the frequency projections given by convolution with
p0, p1 are really necessary, or in other words if it is possible to construct tight
frame systems (Ttϕ)t∈R2 ∪ (ψast)a∈[0,1],s∈[−2,2],t∈R2 ∪ (ψν
ast)a∈[0,1],s∈[−2,2],t∈R2 for
L2(R2). We show that this is actually impossible, meaning that in a sense
Theorem 2.4 is the best we can do.
We remark that for any shearlet ψ, the constant Cψ can also be computed
as
Cψ = Z ∞
−∞Z ∞
0
ψ(aξ1, a1/2(ξ2 − sξ1))2a−3/2dads,
as a short computation reveals. This fact is related to the inherent group struc-
ture of the shearlet transform [4].
Theorem 2.6. Assume that ψ = (cid:0) ∂
θ is a shearlet with M ≥ 1 vanishing
moments in x1-direction such that either M > 1 or M = 1 and θ ∈ L∞(R2).
Furthermore we assume and L1 > 0, L2 > M with L1, L2 defined as in Lemma
2.2. Then there does not exist a window function ϕ such that
∂x1(cid:1)M
ϕ(ξ) = 0
lim
ξ→∞
and such that the system (Ttϕ)t∈R2∪(ψast)a∈[0,1],s∈[−2,2],t∈R2∪(ψν
constitutes a tight frame for L2(R2), which means that a representation formula
ast)a∈[0,1],s∈[−2,2],t∈R2
kfk2
2 =
1
C(cid:0)ZR2 Z 2
−2Z 1
−2Z 1
0 hf, ψν
+ ZR2 Z 2
0 hf, ψasti2a−3dadsdt
asti2a−3dadsdt +ZR2 hf, Ttϕi2dt(cid:1)
(16)
8
holds for all f ∈ L2(R2) and some constant C.
Proof. In terms of the Fourier transform, (16) translates to
∆ψ(ξ) + ∆ψν (ξ) + ϕ(ξ)2 = C.
If we assume that ϕ has Fourier decay limξ→∞ ϕ(ξ) = 0 this would imply that
lim
ξ→∞(cid:0)∆ψ(ξ) + ∆ψν (ξ)(cid:1) = C.
Lemma 2.2 implies that limξ→∞ ∆ψ(ξ) = limξ→∞ ∆ψν (ξ) = Cψ for 2
ξ2 ≤
3
2 . It follows that C = 2Cψ. On the other hand, using the moment condition
θ for M = 1 and θ ∈ L∞(R2), we have the following estimate for ∆ψ
ψ = ξM
1
and ξ in the strip Sδ := {ξ : ξ1 ≤ δ}:
−2Z 1
a2kθk∞a−3/2dads ≤ 8kθk∞δ2.
3 ≤ ξ1
∆ψ(ξ) ≤ δ2Z 2
0
If we assume that M > 1, we can write ψ(ξ) = ξM−1
shearlet. We get a similar estimate as above:
1
µ(ξ) where µ is still a
∆ψ(ξ) ≤ δ2(M−1)Z 2
0 µ(aξ1, a1/2(ξ2 − sξ1))2a−3/2dads ≤ Cµδ2(M−1).
At any rate, by choosing δ appropriately small, this implies that for ξ ∈ Sδ we
have
−2Z 1
C = lim
ξ→∞(cid:0)∆ψ(ξ) + ∆ψν (ξ)(cid:1) < C,
which gives a contradiction.
It is easy to extend this argument to show that there do not exist shearlet
frames such that there exists a dual frame which also has the structure of a
shearlet system, see [1] for more information on frames in general and [6] for
more information on shearlet frames in particular.
3 Concluding Remarks
In future work we would like to address the problem of constructing continuous
tight frames for the so-called 'Hart Smith Transform' [3, 11] where the shear
operation is replaced with a rotation. We think that in this case the results
might become simpler. One reason for this is that in this case no (smoothed)
projection onto a frequency cone is needed.
In view of constructing discrete
tight frames we think that a simple discretization of a continuous tight frame
will not work for non-bandlimited shearlets. The approach that we are currently
pursuing in this direction is to construct so-called Shearlet MRA's via specific
scaling functions and to try to generalize the 'unitary extension principle' of
Ron and Shen [10] to the shearlet setting [6].
9
4 Acknowledgments
The research for this paper has been carried out while the author was working
at the Center for Geometric Modeling and Scientific Visualization at KAUST,
Saudi Arabia.
References
[1] S. T. Ali, J. P. Antoine, and J. P. Gazeau. Continuous frames in Hilbert
space. Annals of Physics, 222:1 -- 37, 1993.
[2] R. Boas Jr. and M. Kac.
Inequalities for Fourier transforms of positive
functions. Duke Mathematical Journal, 12(1):189 -- 206, 1945.
[3] E. J. Candes and D. L. Donoho. Continuous curvelet transform: I. resolu-
tion of the wavefront set. Applied and Computational Harmonic Analysis,
19:162 -- 197, 2003.
[4] S. Dahlke, G. Kutyniok, P. Maass, C. Sagiv, H.-G. Stark, and G. Teschke.
The uncertainty principle associated with the continuous shearlet trans-
form. International Journal of Wavelets, Multiresolution and Information
Processing, 6:157 -- 181, 2008.
[5] P. Grohs.
set.
Continuous shearlet frames and resolution of the wave-
from
Technical
front
http://www.dmg.tuwien.ac.at/grohs/papers/shres.pdf.
report, KAUST, 2009.
available
[6] P. Grohs. Refinable functions for composite dilation systems.
2009.
manuscript in preparation.
[7] L. Hormander. The Analysis of linear Partial Differential Operators.
Springer, 1983.
[8] G. Kutyniok and D. Labate. Resolution of the wavefront set using con-
tinuous shearlets. Transactions of the American Mathematical Society,
361:2719 -- 2754, 2009.
[9] D. Labate, G. Kutyniok, W.-Q. Lim, and G. Weiss. Sparse multidimen-
In Wavelets XI (San Diego, CA,
sional representation using shearlets.
2005), 254-262, SPIE Proc. 5914, SPIE, Bellingham, WA, 2005.
[10] A. Ron and Z. Shen. Affine systems in L2(Rd): The analysis of the analysis
operator. Journal of Functional Analysis, 148(2):408 -- 447, 1997.
[11] H. F. Smith. A Hardy space for Fourier integral operators. Journal of
Geometic Analysis, 8:629 -- 653, 1998.
10
|
1004.2103 | 1 | 1004 | 2010-04-13T03:48:04 | On dominant contractions and a generalization of the zero-two law | [
"math.FA"
] | Zaharopol proved the following result: let $T,S:L^1(X,{\cf},\m)\to L^1(X,{\cf},\m)$ be two positive contractions such that $T\leq S$. If $\|S-T\|<1$ then $\|S^n-T^n\|<1$ for all $n\in\bn$. In the present paper we generalize this result to multi-parameter contractions acting on $L^1$. As an application of that result we prove a generalization of the "zero-two" law. | math.FA | math | ON DOMINANT CONTRACTIONS AND A
GENERALIZATION OF THE ZERO-TWO LAW
FARRUKH MUKHAMEDOV
Abstract. Zaharopol proved the following result: let T, S : L1(X,F , µ) →
L1(X,F , µ) be two positive contractions such that T ≤ S. If kS − Tk < 1
then kS n − T nk < 1 for all n ∈ N. In the present paper we generalize this
result to multi-parameter contractions acting on L1. As an application of
that result we prove a generalization of the "zero-two" law.
Keywords: dominant contraction, positive operator, "zero-two" law.
AMS Subject Classification: 47A35, 17C65, 46L70, 46L52, 28D05.
0
1
0
2
r
p
A
3
1
]
.
A
F
h
t
a
m
[
1
v
3
0
1
2
.
4
0
0
1
:
v
i
X
r
a
1. Introduction
Let (X,F , µ) be a measure space with a positive σ-additive measure µ. In
what follows for the sake of shortness by L1 we denote the usual L1(X,F , µ)
space associated with (X,F , µ). A linear operator T : L1 → L1 is called a
positive contraction if T f ≥ 0 whenever f ≥ 0 and kTk ≤ 1.
In [9] it was proved so called "zero-two" law for positive contractions of
L1-spaces:
Theorem 1.1. Let T : L1 → L1 be a positive contraction. If for some m ∈
N ∪ {0} one has kT m+1 − T mk < 2, then
n→∞kT n+1 − T nk = 0.
lim
In [2] it was proved a "zero-two" law for Markov processes, which allowed
to study random walks on locally compact groups. Other extensions and gen-
eralizations of the formulated law have been investigated by many authors
[7, 4, 5].
Using certain properties of L1-spaces Zaharopol [10] by means of the follow-
ing theorem reproved Theorem 1.1.
Theorem 1.2. Let T, S : L1 → L1 be two positive contractions such that
T ≤ S. If kS − Tk < 1 then kSn − T nk < 1 for all n ∈ N
In the paper we provide an example (see Example 2) for which the formu-
lated theorem 1.2 can not be applied. Therefore, we prove a generalization of
Theorem 1.2 for multi-parameter contractions acting on L1. As a consequence
1
2
FARRUKH MUKHAMEDOV
of that result we shall provide a generalization of the "zero-two" law. Similar
generalization has been considered in [5].
2. Dominant operators
Let T, S : L1 → L1 be two positive contractions. We write T ≤ S if S − T
is a positive operator. In this case we have
(2.1)
for every x ≥ 0. Moreover, for positive operator T : L1 → L1 one can prove
the following equality
kSx − T xk = kSxk − kT xk,
(2.2)
kxk=1kT xk = sup
The main result of this section is the following
kTk = sup
kxk=1,x≥0kT xk.
2 k < 1. Then kS1Sn
Theorem 2.1. Let T1, T2, S1, S2 : L1 → L1 be positive contractions such that
Ti ≤ Si, i = 1, 2 and S1S2 = S2S1. If there is an n0 ∈ N such that kS1Sn0
2 −
T1T n0
2 − T1T n
Proof. Let us assume that kS1Sn
denote
2 k < 1 for every n ≥ n0.
2 − T1T n
m = min{n ∈ N : kS1Sn0+n
2 k = 1 for some n > n0. Therefore,
It is clear that m ≥ 1. The inequalities T1 ≤ S1, T2 ≤ S2 imply that S1Sn0+n
−
T1T n0+n
is a positive operator. Then according to (2.2) there exists a sequence
{xn} ∈ L1 such that xn ≥ 0, kxnk = 1,∀n ∈ N and
(2.3)
− T1T n0+n
k = 1}.
2
2
2
2
Positivity of S1Sn0+n
2
n→∞k(S1Sn0+n
lim
− T1T n0+n
2
2
2
)xnk = 1.
− T1T n0+n
and xn ≥ 0 together with (2.1) imply that
k(S1Sn0+n
(2.4)
for every n ∈ N. It then follows from (2.3),(2.4) that
)xnk = kS1Sn0+m
− T1T n0+n
2
2
2
xnk − kT1T n0+m
2
xnk
(2.5)
(2.6)
2
n→∞kS1Sn0+m
lim
n→∞kT1T n0+m
lim
2
xnk = 1,
xnk = 0.
Thanks to the contractivity of S, Z and S1S2 = S2S1 one gets
kS1Sn0+m
2
xnk = kS2(S1Sn0+m−1
2
xn)k ≤ kS1Sn0+m−1
2
xnk ≤ kSm
2 xnk
which with (2.5) yields
(2.7)
n→∞kS1Sn0+m−1
lim
2
xnk = 1,
n→∞kSm
lim
2 xnk = 1.
Moreover, the contractivity of Si, Ti (i = 1, 2) implies that kT1T n0+m−1
xnk ≤
2 T mxnk ≤ 1 for every n ∈ N. Therefore, we may
2 xnk ≤ 1 and kS1Sn0
1, kT m
2
ON DOMINANT CONTRACTIONS
choose a subsequence {yk} of {xn} such that the sequences {kT1T n0+m−1
{kT m
(2.8)
2 T mxkk} converge. Put
2 ykk}, {kS1Sn0
α = lim
2
3
ykk},
(2.9)
(2.10)
β = lim
2
ykk,
2 T mykk,
k→∞kT1T n0+m−1
k→∞kS1Sn0
k→∞kT m
2 ykk.
− T1T n0+m−1
γ = lim
2
The inequality kS1Sn0+m−1
Hence we may choose a subsequence {zk} of {yk} such that kT1T n0+m−1
for all k ∈ N.
and hence γ > 0.
2 zkk together with (2.8), (2.10) we find α ≤ γ,
k < 1 with (2.7) implies that α > 0.
zkk 6= 0
zkk ≤ kT m
2
2
2
From kT1T n0+m−1
Using (2.1) one gets
kS1Sn0
2 T m
2
2 zkk = kS1Sn0+m
= kS1Sn0+m
≥ kS1Sn0+m
= kS1Sn0+m
2
2
2
2
zk − (S1Sn0+m
zkk − kS1Sn0+m
zkk − kSm
zkk − kSm
zk − S1Sn0
2 T m
2 zk)k
zk − S1Sn0
2 T m
2 zkk
2 zk − T m
2 zkk
2 zkk + kT m
2 zkk
2
(2.11)
Due to (2.5),(2.7) we have
k→∞kS1Sn0+m
lim
2
zkk − kSm
2 zkk = 0;
which with (2.11) implies that
k→∞kS1ZSn0
lim
2 T m
2 zkk ≥ lim
k→∞kT m
2 zkk,
therefore, β ≥ γ.
On the other hand, by kS1Sn0
Now set
2 T m
2 zkk one gets γ ≥ β, hence γ = β.
2 zkk ≤ kT m
T m
2 zk
kT m
2 zkk
uk =
, k ∈ N.
Then using the equality γ = β and (2.6) one has
kS1Sn0
2 T m
kT m
2 zkk
2 ukk = lim
k→∞
2 zkk
k→∞kS1Sn0
lim
k→∞kT1T n0
lim
2 ukk = lim
k→∞
So, owing to (2.1) and positivity of S1Sn0
2 , we get
= 1,
= 0.
kT1T n0+mzkk
kT m
2 zkk
2 − T1T n0
2 )zkk = 1.
k→∞k(S1Sn0
lim
2 − T1T n0
Since kukk = 1, uk ≥ 0,∀k ∈ N from (2.2) one finds kS1Sn0
which is a contradiction. This completes the proof.
2 − T1T n0
2 k = 1,
(cid:3)
4
FARRUKH MUKHAMEDOV
Corollary 2.2. Let Z, T, S : L1 → L1 be positive contractions such that T ≤ S
and ZS = SZ. If there is an n0 ∈ N such that kZ(Sn0 − T n0)k < 1. Then
kZ(Sn − T n)k < 1 for every n ≥ n0.
Assume that Z = Id. If n0 = 1, then from Corollary 2.2 we immediately
get the Zaharopol's result (see Theorem 1.2). If n0 > 1 then we obtain a main
result of [8].
Let us provide an example of Z, S, T positive contractions for which state-
ment of Corollary 2.2 is satisfied.
Example 1. Consider R2 with a norm kxk = x1 +x2, where x = (x1, x2).
An order in R2 is defined as usual, namely x ≥ 0 if and only if x1 ≥ 0, x2 ≥ 0.
Now define mappings Z : R2 → R2,T : R2 → R2 and S : R2 → R2, respectively,
by
(2.12)
(2.13)
Z(x1, x2) = (ux1 + vx2, ux2),
S(x1, x2) = (cid:18)x1 + x2
2
,
x2
2 (cid:19),
T (x1, x2) = (λx2, 0).
(2.14)
The positivity of Z,S and T implies that u, v, λ ≥ 0. It is easy to check that
T ≤ S holds if and only if 2λ ≤ 1.
One can see that
kZk = sup
kxk=1
x≥0
kZxk = max
x1+x2=1
x1,x2≥0 {ux1 + (u + v)x2}
0≤x2≤1{u + vx2}
= max
= u + v
Hence, contractivity of Z implies that u + v = 1. Similarly, we find that
kSk = 1 and kTk = λ. From (2.12) and (2.13) one gets that ZS = SZ.
By means of (2.12),(2.13),(2.14) one finds Similarly, one gets
kZ(S − T )k = sup
kxk=1
x≥0
kZ(S − T )xk = max
(2.15)
=
2(cid:0)ux1 + x2 + ux2 − 2λux2(cid:1)(cid:27)
x1+x2=1
x1 ,x2≥0 (cid:26) 1
1 + u(1 − 2λ)
.
2
The condition 2λ ≤ 1 yields that kZ(S − T )k < 1. Consequently, Corollary
2.2 implies kZ(Sn − T n)k < 1 for all n ∈ N.
Now let us formulate a multi-parametric version of Theorem 1.1.
Theorem 2.3. Let Ti, Si : L1 → L1, i = 1, . . . , N be positive contractions such
that Ti ≤ Si with
(2.16)
for every i, j = 1, . . . , N.
TiTj = TjTi, SiSj = SjSi
ON DOMINANT CONTRACTIONS
5
If there are ni,0 ∈ N, i = 1, . . . , N such that
N − T n1,0
(2.17)
· · · SnN,0
kSn1,0
1
1
· · · T nN,0
N k < 1.
Then
(2.18)
kSm1
1
· · · SmN
N − T m1
1
· · · T mN
N k < 1
for all mi ≥ ni,0, i = 1, . . . , N .
Proof. Let us fix the first N − 1 operators in (2.17), i.e.
denote
for a moment we
(2.19)
SN −1 = Sn1,0
1
· · · SnN−1,0
N −1
TN −1 = T n1,0
1
· · · T nN−1,0
N −1
,
then (2.17) can be written as follows
kSN −1SnN,0
N − TN −1T nN,0
N k < 1.
After applying Theorem 2.1 to the last inequality we find
(2.20)
kSN −1SmN
N − TN −1T mN
N k < 1
for all mN ≥ nN,0. Now taking into account (2.19) and (2.16) we rewrite (2.20)
as follows
(2.21)
N Sn1,0
kSmN
1
· · · SnN−1,0
N −1 − T mN
N T n1,0
1
· · · T nN−1,0
N −1 k < 1.
Now again applying the same idea as above to (2.21) we get
N Sn1,0
kSmN
1
· · · SmN−1
N −1 − T mN
N T n1,0
1
· · · T mN−1
N −1 k < 1,
for all mN −1 ≥ nN −1,0, mN ≥ nN,0. Hence, continuing this procedure N − 2
times we obtain the desired inequality.
(cid:3)
Remark 3.1. It should be noted the following:
(i) Since the dual of L1 is L∞ then due to the duality theory the proved
Theorems 2.1 and 2.3 holds true if we replace L1-space with L∞.
(ii) Unfortunately, that the proved theorems and its corollaries are not
longer true if one replaces L1-space by an Lp-space, 1 < p < ∞. Indeed,
consider X = {1, 2}, F = P({1, 2}) and the measure µ is given by
µ({1}) = µ({2}) = 1/2. In this case, Lp is isomorphic to the Banach
lattice R2 (here an order is defined as usual, namely x ≥ 0 if and only
if x1 ≥ 0, x2 ≥ 0) with the norm kxkp = (cid:0)x1p + x2p(cid:1)1/p/2, where
x = (x1, x2). Define two operators by
S(x1, x2) = (cid:18)x1 + x2
2
,
x1 + x2
2 (cid:19), T (x1, x2) = (cid:18)0,
x1
2 (cid:19)
Then it is shown (see [10]) that kS − Tk < 1, but kS2 − T 2k = 1.
6
FARRUKH MUKHAMEDOV
(iii) It would be better to note that certain ergodic properties of dominant
positive operators has been studied in [3]. In general, a monograph [6]
is devoted to dominant operators.
Let us give another example, for which conditions of Theorem 1.2 does not
hold, but Theorem 2.1 can be applied.
Example 2. Let us consider R2 as in Example 1. Now define mappings
T : R2 → R2 and S : R2 → R2 as follows
(2.22)
x1 +
S(x1, x2) = (cid:18)1
T (x1, x2) = (cid:18)1
2
4
x2, 0(cid:19).
1
3
x2,
1
2
x1 +
1
3
x2(cid:19),
(2.23)
(2.24)
(2.25)
It is clear that S and T are positive and T ≤ S.
One can see that kSk = 1, kTk = 1/4. From (2.22),(2.23) one gets
kS − Tk = sup
kxk=1
x≥0
k(S − T )xk = max
0≤x1≤1(cid:26) 7x1 + 5
12 (cid:27) = 1
kS2 − T 2k = sup
kxk=1
x≥0
k(S2 − T 2)xk = max
0≤x1≤1(cid:26) 5x1 + 10
18 (cid:27) =
15
18
Consequently, we have positive contractions T and S with S ≥ T such that
kS − Tk = 1,kS2 − T 2k < 1. This shows that the condition of Theorem 1.2 is
not satisfied, but due to Corollary 2.2 with Z = id we have kSn − T nk < 1 for
all n ≥ 2. Therefore the proved Theorem 2.2 is an extension of the Zaharopol's
result.
3. A generalization of the zero-two law
In this section we are going to prove a generalization of the zero-two law for
positive contractions on L1. Before formulate the main result we prove some
auxiliary facts.
First note that for any x, y ∈ L1 one defines
(3.1)
x ∧ y =
(x + y − x − y).
1
2
It is well known (see [1]) that for any mapping S of L1 one can define its
modulus by
(3.2)
Sx = sup{Sy :
y ≤ x}, x ∈ L1, x ≥ 0.
Hence, similarly to (3.1) for given two mappings S, T of L1 we define
(3.3)
(S ∧ T )x =
1
2
(Sx + T x − S − Tx), x ∈ L1.
ON DOMINANT CONTRACTIONS
7
A linear operator Z : L1 → L1 is called a lattice homomorphism whenever
(3.4)
Z(x ∨ y) = Zx ∨ Zy
holds for all x, y ∈ L1. One can see that such an operator is positive. Note
that such homomorphisms were studied in [1].
Recall that a net {xα} in L1 is order convergent to x, denoted xα →o x
whenever there exists another net {yα} with the same index set satisfying
xα − x ≤ yα ↓ 0. An operator T : L1 → L1 is said to be order continuous, if
xα →o 0 implies T xα →o 0.
Lemma 3.1. Let S, T be positive contractions of L1, and Z be an order
continuous lattice homomorphism of L1. Then one has
(3.5)
Moreover, we have
(3.6)
ZS − T = Z(S − T ).
Z(S ∧ T ) = ZS ∧ ZT.
Proof. From (3.2) we find that
ZS − Tx = Z(sup{(S − T )y :
= sup{Z(S − T )y :
= Z(S − T )x,
y ≤ x})
y ≤ x})
(3.7)
for every x ∈ L1, x ≥ 0.
The equality (3.3) yields that
(3.8)
ZS − T = ZS + ZT − 2Z(S ∧ T ),
Z(S − T ) = ZS + ZT − 2(ZS ∧ ZT ),
which with (3.7) imply that
Z(S ∧ T ) = ZS ∧ ZT.
(cid:3)
In what follows, an order continuous lattice homomorphism Z : L1 → L1
Now we have the following
with kZk ≤ 1, is called a lattice contraction.
Lemma 3.2. Let Z be a lattice contraction and T be a positive contraction of
L1 such that ZT = T Z. If for some m ∈ N ∪ {0}, k ∈ N one has kZ(T m+k −
T m)k < 2, then kZ(T m+k − T m+k ∧ T m)k < 1.
8
FARRUKH MUKHAMEDOV
Proof. According to the assumption there is δ > 0 such that kZ(T m+k−T m)k =
2(1 − δ). Let us suppose that kZ(T m+k − T m+k ∧ T m)k = 1. Then thanks to
(2.2) there exists x ∈ L1 with x ≥ 0, kxk = 1 such that
δ
kZ(T m+k − T m+k ∧ T m)xk > 1 −
4
,
which with (2.1) implies that kZT m+kxk > 1 − δ/4 and kZ(T m+k ∧ T m)xk <
δ/4. The commutativity T and Z yields that kZT mxk > 1 − δ/4.
Now using (3.8) and (3.6) one finds
(cid:13)(cid:13)Z(T m+k − T m)x(cid:13)(cid:13) = kZT m+kxk + kZT mxk − 2kZ(T m+k ∧ T m)xk
δ
4 − 2 ·
δ
4
δ
4
> 1 −
= 2(cid:18)1 −
+ 1 −
2(cid:19).
δ
This with the equality
(cid:13)(cid:13)Z(T m+k − T m)(cid:13)(cid:13) = kZ(T m+k − T m)k,
contradicts to kZ(T m+k − T m)k = 2(1 − δ/2).
Lemma 3.3. Let Z be a lattice contraction and T be a positive contraction of
L1 such that ZT = T Z. If for some m ∈ N ∪ {0}, k ∈ N one has kZ(T m+k −
T m+k ∧ T m)k < 1, then for any ε > 0 there are d, n0 ∈ N such that
(cid:3)
kZ d(T n+k − T n)k < ε
for all n ≥ n0
Proof. It is known that (see [11], p. 310) for any contraction T on L1 there is
γ > 0 such that
(3.9)
≤
γ
√ℓ
.
2 (cid:19)ℓ(cid:13)(cid:13)(cid:13)(cid:13)
− T(cid:18)I + T
2 (cid:19)ℓ(cid:13)(cid:13)(cid:13)(cid:13)
− T k(cid:18) I + T
≤
Let ε > 0 and fix ℓ ∈ N such that kγ/√ℓ < ε/4.
Then for given k ∈ N, using (3.9) one easily finds that
kγ
√ℓ
2 (cid:19)ℓ
(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18) I + T
2 (cid:19)ℓ
(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18)I + T
.
(3.10)
Then according to Corollary 2.2 from the assumption of the lemma we have
(3.11)
(cid:13)(cid:13)Z(T ℓ(m+k) − (T m+k ∧ T m)ℓ)(cid:13)(cid:13) < 1.
ON DOMINANT CONTRACTIONS
9
Hence,
2 (cid:19)ℓ
(cid:13)(cid:13)(cid:13)(cid:13)
Z(cid:18)T ℓ(m+k) − (cid:18)I + T
Z(cid:18)T ℓ(m+k) −
(T m+k ∧ T m)ℓ(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
ℓ
C i
=
ℓ
C i
ℓ
1
2ℓ
Xi=0
ℓT i(T m+k ∧ T m)ℓ(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
2ℓ (cid:13)(cid:13)Z(T ℓ(m+k) − T i(T m+k ∧ T m)ℓ)(cid:13)(cid:13)
Xi=0
= (cid:13)(cid:13)(cid:13)(cid:13)
Xi=0
2ℓ(cid:13)(cid:13)Z(T ℓ(m+k) − (T m+k ∧ T m)ℓ)(cid:13)(cid:13) +
1
2ℓ +
C i
ℓ
2ℓ = 1.
C i
ℓ
2ℓ
≤
≤
<
1
ℓ
(3.12)
Define
ℓ
Xi=1
2 (cid:19)ℓ
Qℓ := T ℓ(m+k) −(cid:18)I + T
(T m+k ∧ T m)ℓ
and put V (1)
ℓ = (T m+k ∧ T m)ℓ. Then one can see that
2 (cid:19)ℓ
T ℓ(m+k) = (cid:18)I + T
V (1)
ℓ + Qℓ.
Now for every d ∈ N, define
V (d+1)
ℓ
= T ℓ(m+k)V (d)
ℓ + V (1)
ℓ Qd
ℓ .
Then by induction one can establish [11] that
(3.13)
2 (cid:19)ℓ
T dℓ(m+k) = (cid:18)I + T
V (d)
ℓ + Qd
ℓ
for every d ∈ N.
Due to Proposition 2.1 [10] one has
(3.14)
kV (d)
ℓ k ≤ 2
for all d ∈ N.
Now from (3.12) we find kZQℓk < 1, therefore there exists d ∈ N such that
k(ZQℓ)dk < ε/4. So, commutativity Z and T implies that ZQℓ = QℓZ, which
yields that kZ dQd
ℓk < ε/4.
10
FARRUKH MUKHAMEDOV
Put n0 = dℓ(m + k), then from (3.13) with (3.10),(3.14) we get
kZ d(T n0+k − T n0)k = (cid:13)(cid:13)(cid:13)(cid:13)
≤ (cid:13)(cid:13)(cid:13)(cid:13)
+Z d(T kQd
2 (cid:19)ℓ
Z d(cid:18)T k(cid:18)I + T
ℓ )(cid:13)(cid:13)(cid:13)(cid:13)
ℓ − Qd
2 (cid:19)ℓ
(cid:18)T k(cid:18) I + T
−(cid:18) I + T
+kZ dQd
ℓ (T − 1)k
+ 2 ·
< ε.
kγ
√ℓ
ε
4
≤ 2 ·
−(cid:18)I + T
2 (cid:19)ℓ(cid:19)V (d)
ℓ
2 (cid:19)ℓ(cid:19)V (d)
ℓ (cid:13)(cid:13)(cid:13)(cid:13)
Take any n ≥ n0, then from the last inequality one finds
kZ d(T n+k − T n)k = kT n−n0Z d(T n0+k − T n0)k ≤ kZ d(T n0+k − T n0)k < ε
which completes the proof.
(cid:3)
Now we are ready to formulate the main result of this section.
Theorem 3.4. Let Z, T be two positive contractions of L1 such that T Z = ZT .
If for some m ∈ N ∪ {0}, k ∈ N one has kZ(T m+k − T m)k < 2, then for any
ε > 0 there are d, n0 ∈ N such that
kZ d(T n+k − T n)k < ε
for all n ≥ n0
The proof of this theorem immediately follows from Lemmas 3.2 and 3.3.
Remark. Note that if we take as Z = I,k = 1 then we obtain Theorem 1.1
as a corollary of Theorem 3.4.
References
[1] C.D. Aliprantis, O. Burkinshaw, Positive operators, Springer, 2006.
[2] Y. Derriennic, Lois "z´ero ou deux" pour les processes de Markov. Applications aux
marches al´eatoires. Ann. Inst. H. Poincar´e Sec. B 12(1976), 111 -- 129.
[3] E.Ya. Eemel'yanov, Non-spectral asymptotic analysis of oneparameter operator semi-
groups. In: Operator Theory: Advances and Applications, 173. Basel. Birkhauser Ver-
lag. 2007.
[4] S. R. Foguel, On the zero-two law, Israel J. Math. 10(1971), 275 -- 280.
[5] S. R. Foguel, A generalized 0-2 law, Israel J. Math. 45(1983), 219 -- 224.
[6] A.G. Kusraev, Dominated operators. In: Mathematics and its Applications, V. 519.
Dordrecht: Kluwer Academic Publishers 2000.
[7] M. Lin, On the 0 -- 2 law for conservative Markov operators, Z. Wahrscheinlichkeitstheor.
Verw. Geb. 61 (1982), 513 -- 525.
[8] F. Mukhamedov, S. Temir, H. Akin, A note on dominant contractions of Jordan alge-
bras, Turk. J. Math. 34,(2010), 85 -- 93.
[9] D. Ornstein, L. Sucheston, An operator theorem on L1 convergence to zero with appli-
cations to Markov operators, Ann. Math. Statist. 41(1970) 1631 -- 1639.
ON DOMINANT CONTRACTIONS
11
[10] R. Zaharopol, On the 'zero-two' law for positive contraction, Proc. Edin. Math. Soc.
32,(1989), 363 -- 370.
[11] R. Zaharopol, The modulus of a regular linear operator and the 'zero-two' law in Lp-
spaces (1 < p < +∞, p 6= 2), Jour. Funct. Anal. 68,(1986), 300 -- 312.
Farrukh Mukhamedov, Department of Computational & Theoretical Sci-
ences, Faculty of Science, International Islamic University Malaysia, P.O.
Box, 141, 25710, Kuantan, Pahang, Malaysia
E-mail address: [email protected] farrukh [email protected]
|
1609.08612 | 1 | 1609 | 2016-09-27T15:40:05 | Representations of $p$-convolution algebras on $L^q$-spaces | [
"math.FA",
"math.OA"
] | For a nontrivial locally compact group $G$, and $p\in [1,\infty)$, consider the Banach algebras of $p$-pseudofunctions, $p$-pseudomeasures, $p$-convolvers, and the full group $L^p$-operator algebra. We show that these Banach algebras are operator algebras if and only if $p=2$. More generally, we show that for $q\in [1,\infty)$, these Banach algebras can be represented on an $L^q$-space if and only if one of the following holds: (a) $p=2$ and $G$ is abelian; or (b) $|\frac 1p - \frac 12|=|\frac 1q - \frac 12|$. This result can be interpreted as follows: for $p,q\in [1,\infty)$, the $L^p$- and $L^q$-representation theories of a group are incomparable, except in the trivial cases when they are equivalent.
As an application, we show that, for distinct $p,q\in [1,\infty)$, if the $L^p$ and $L^q$ crossed products of a topological dynamical system are isomorphic, then $\frac 1p + \frac 1q=1$. In order to prove this, we study the following relevant aspects of $L^p$-crossed products: existence of approximate identities, duality with respect to $p$, and existence of canonical isometric maps from group algebras into their multiplier algebras. | math.FA | math | REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON
Lq-SPACES
EUSEBIO GARDELLA AND HANNES THIEL
Abstract. For a nontrivial locally compact group G, consider the Banach
algebras of p-pseudofunctions, p-pseudomeasures, p-convolvers, and the full
group Lp-operator algebra. We show that these Banach algebras are operator
algebras if and only if p = 2. More generally, we show that for q ∈ [1, ∞),
these Banach algebras can be represented on an Lq-space if and only if one
(cid:12)
of the following holds: (a) p = 2 and G is abelian; or (b)
.
(cid:12)
(cid:12)
This result can be interpreted as follows: for p, q ∈ [1, ∞), the Lp- and Lq-
representation theories of a group are incomparable, except in the trivial cases
when they are equivalent.
=
(cid:12)
(cid:12)
(cid:12)
(cid:12)
(cid:12)
(cid:12)
(cid:12)
(cid:12)
(cid:12)
1
2
1
2
1
q
−
1
p
−
1
p
+ 1
As an application, we show that, for distinct p, q ∈ [1, ∞), if the Lp and
Lq crossed products of a topological dynamical system are isomorphic, then
= 1. In order to prove this, we study the following relevant aspects of
Lp-crossed products: existence of approximate identities, duality with respect
to p, and existence of canonical isometric maps from group algebras into their
multiplier algebras.
q
1
v
2
1
6
8
0
.
9
0
6
1
:
v
i
X
r
a
Contents
1.
Introduction
2. Preliminaries
3. F p(Z) is not isomorphic to F q(Z)
4. Nonrepresentability on Lq-spaces
5. An application to crossed products
References
2
4
6
10
20
30
Date: 17 July 2018.
2010 Mathematics Subject Classification. Primary: 47L10, 43A15, Secondary: 43A65, 46E30.
Key words and phrases. Locally compact group, algebra of p-pseudofunctions, algebra of p-
pseudomeasures, contractive approximate identity, multiplier algebra, group amenability, crossed
product.
The first named author was partially supported by the D. K. Harrison Prize from the University
of Oregon and by a Postdoctoral Research Fellowship from the Humboldt Foundation. The second
named author was partially supported by the Deutsche Forschungsgemeinschaft (SFB 878). Part
of this work was completed while the authors were taking part in the Research Program Classifi-
cation of operator algebras, complexity, rigidity and dynamics, held at the Institut Mittag-Leffler,
between January and April of 2016. We would like to thank the staff and organizers, and Søren
Eilers in particular, for the hospitality and financial support.
1
2
EUSEBIO GARDELLA AND HANNES THIEL
1. Introduction
double commutant of C ∗
λ(G) is the algebra F p
We say that a Banach algebra is an operator algebra if it admits an isometric
representation as bounded operators on a Hilbert space. Associated to any locally
compact group G there are three fundamentally important operator algebras: its
reduced group C ∗-algebra C ∗
λ(G), its full group C ∗-algebra C ∗(G), and its group von
Neumann algebra L(G). These are, respectively, the Banach algebra generated by
the left regular representation of G on L2(G); the universal C ∗-algebra with respect
to unitary representations of G on Hilbert spaces; and the weak∗ closure (also called
λ(G) in B(L2(G)). (We canonically identify B(L2(G)) with
ultraweak closure) of C ∗
the dual of the projective tensor product L2(G)b⊗L2(G).) Equivalently, L(G) is the
λ(G) in B(L2(G)).
These operator algebras admit generalizations to representations of G on Lp-
spaces, for p ∈ [1,∞). The analog of C ∗
λ (G) of p-pseudofunc-
tions on G, introduced by Herz in [Her73] and originally denoted by P Fp(G). The
analog of C ∗(G) is the full group Lp-operator algebra F p(G), defined by Phillips in
[Phi13]. Finally, the von Neumann algebra L(G) has two analogs, at least for p 6= 1:
the algebra P Mp(G) of p-pseudomeasures, which is the weak∗ closure of F p
λ (G) in
B(Lp(G)) (where we canonically identify B(Lp(G)) with the dual of the projective
tensor product Lp(G)b⊗Lp(G)∗); and the algebra CVp(G) of p-convolvers, which is
the double commutant of F p
λ (G) in B(Lp(G)) (it is also the commutant of the right
regular representation). Both P Mp(G) and CVp(G) were introduced in [Her73].
These objects, and related ones, have been studied by a number of authors in
the last three decades. For instance, see [Cow98], [NR09], [Run05], [DS13], and
the more recent papers [Phi13], [Phi14], [GT15b], and [GT14]. In [GT16a], it is
shown that there is a certain quotient of F p(Z) which cannot be represented on
an Lp-space (in fact, on any Lq-space for q ∈ [1,∞)), thus answering a 20-year-old
question of Le Merdy.
Despite the advances in the area, some basic questions remain open. One im-
portant open problem is whether P Mp(G) = CVp(G) for all p ∈ (1,∞) and for all
locally compact groups G. This is known to be true when p = 2 (both algebras
agree with L(G)), essentially by the double commutant theorem. It is immediate
that P Mp(G) ⊆ CVp(G) in general, while Herz showed in [Her73] that equality
holds for all p if G is amenable, a result that was later generalized by Cowling in
[Cow98] to groups with the approximation property.
A less studied problem is the following. By universality of F p(G), there is a
canonical contractive homomorphism κp : F p(G) → F p
λ (G) with dense range. For
p = 2, this map is known to be a quotient map, and for p = 1 it is an isomorphism
regardless of G. On the other hand, we do not know if κp is also a quotient map
for all other values of p. In fact, we do not even know whether κp is surjective. If
this map is not necessarily surjective, can it be injective without the group being
amenable? (By Theorem 3.7 in [GT15b], G is amenable if and only if κp is bijective
for some (equivalently, for all) p ∈ (1,∞). This result was independently obtained
by Phillips in [Phi13] and [Phi14], using different methods.) In this case, it would
be interesting to describe precisely for what groups (and Holder exponents) the
map κp is injective but not surjective.
Questions of the nature described above would in principle be easier to tackle if
the objects considered had a better understood structure, as is the case for operator
algebras. Despite the fact that the Banach algebras F p(G), F p
λ (G), P Mp(G) and
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
3
CVp(G) have natural representations as operators on an Lp-space, this by itself
does not rule out having isometric representations on Hilbert spaces as well; see,
for example, [BLM95]. It is therefore not a priori clear whether the Lp-analogs of
group operator algebras can be isometrically represented on Hilbert spaces.
In this paper, we settle this question negatively. Indeed, we show in Theorem 4.12
that for a nontrivial locally compact group G, and for p ∈ [1,∞) \ {2}, none of the
algebras F p(G), F p
λ (G), P Mp(G), or CVp(G) can be isometrically represented on a
Hilbert space. This result generalizes Theorem 2.2 in [NR09], where Neufang and
Runde assume that G is amenable and has a closed infinite abelian subgroup. More
generally, for p, q ∈ [1,∞) we show that the algebras F p
λ (G), F p(G), P Mp(G), or
CVp(G) can be isometrically represented on an Lq-space if and only if one of the
following holds:
(1) p = 2 and G is abelian; or
(2) (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12). (This is equivalent to either p = q or 1
p + 1
q = 1.)
This result can be interpreted as asserting that the Lp- and Lq-representation
theories of a nontrivial group are incomparable, whenever they are not "obviously"
equivalent. As a consequence, it follows that if there is an isometric Banach algebra
isomorphism F p
λ (G) (or between full group algebras, pseudomeasures or
convolvers) for distinct p, q ∈ [1,∞), then 1
q = 1; see Corollary 4.13. The
converse also holds; see Theorem 4.11.
λ (G) ∼= F q
p + 1
(cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12); see Corollary 5.11. Since we do not know whether an isomor-
As an application, we show that, for p, q ∈ [1,∞), if the Lp- and Lq-crossed
products of a topological dynamical system are isometrically isomorphic, then
p − 1
phism F p(G, X, α) → F q(G, X, α) (or F p
λ (G, X, α)) must necessar-
ily respect the group action α, Corollary 4.13 is not enough to obtain the conclusion.
This means that even if we are only interested in isomorphisms of crossed prod-
ucts, we are forced to consider arbitrary representations of p-convolution algebras
on Lq-spaces. In order to obtain these results, we need to develop the theory of
Lp-crossed products further, and we do so by exploring the following fundamental
aspects: existence of approximate identities, duality with respect to p, and existence
of canonical isometric maps from group algebras into their multiplier algebras.
λ (G, X, α) → F q
We give an outline of the proof of our main result (Theorem 4.12). Let E be
λ (G) → B(E) be an isometric representation. (Similar
an Lq-space and let ϕ : F p
arguments apply for the algebras F p(G), P Mp(G), or CVp(G).)
Step 1. The case q = 2 is treated separately, since in this case one can show that
F p
λ (G) is a C ∗-algebra; see Theorem 4.6. So assume that q 6= 2.
Step 2. Using results in [GT16c], we may assume that ϕ is non-degenerate.
Step 3. Functoriality properties of F p
λ with respect to subgroups ([GT14]) allow us
to further reduce the problem to the case where G is a cyclic group (finite
or infinite); see Lemma 4.10.
Step 4. For cyclic G, we need to know that ϕ(F p
λ (G)) is isometrically isomorphic
to F q
λ (G). This requires non-trivial results from [GT15a] on spectral con-
figurations. (For instance, we use the fact that F q
λ (Z) is the unique Lq-
operator algebra generated by an invertible isometry whose Gelfand trans-
form is not surjective.) We conclude that there is an isometric isomorphism
λ (G) ∼= F q
F p
λ (G).
4
EUSEBIO GARDELLA AND HANNES THIEL
Step 5. When G = Z, we show that (cid:12)(cid:12)(cid:12) 1
mentary computations.
p − 1
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12) in Theorem 3.6 using ele-
Step 6. When G = Zn, we use the existence, for 1 ≤ p ≤ q ≤ 2, of a canonical,
contractive map γp,q : F p(G) → F q(G) with dense range (see Theorem 2.5),
to show that the Gelfand transform of F p
λ (G) is an isometric isomorphism,
thus reducing the problem to the case in Step 1 above.
one is only interested in the case (cid:12)(cid:12)(cid:12) 1
We wish to point out that the use of spectral configurations can be avoided if
Indeed, this situation can
be entirely dealt with the maps γp,q (see Proposition 4.1 or Theorem 4.14 for a
similar argument, but in a different context). Also, the fact that we can assume
representations to be non-degenerate (Step 2) is by no means obvious, and the
paper [GT16c] grew out of our attempts to prove this.
2(cid:12)(cid:12)(cid:12) > (cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12).
p − 1
q − 1
n
p + 1
1.1. Notation. We take N = {1, 2, . . .}. For n ∈ N and p ∈ [1,∞), we write ℓp
in place of ℓp({1, . . . , n}), and we write ℓp in place of ℓp(Z). For a Banach space
E, we write B(E) for the Banach algebra of bounded linear operators on E. For
p ∈ (1,∞), we denote by p′ its conjugate (Holder) exponent, which is determined
by the identity 1
p′ = 1. Consistently, for a Banach space E, we denote its dual
space by E′, and for a linear map π : E → F between Banach spaces E and F , we
denote by π′ : F ′ → E′ its transpose map.
Locally compact groups are assumed to be Hausdorff, and will always be im-
plicitly endowed with a (fixed) left Haar measure, which will be chosen to be the
counting measure whenever the group is discrete. The left Haar measure of a locally
compact group G will be denoted by µ, and we will denote by ν the right Haar
measure on G determined by ν(U ) = µ(U −1) for all measurable sets U ⊆ G. The
modular function of G will be denoted by ∆ : G → R+. We will repeatedly use the
following identities:
ZG
valid for all f ∈ L1(G, µ).
∆(t−1)f (t−1) dµ(t) =ZG
f (t−1) dν(t) =ZG
f (t) dµ(t) =ZG
∆(t)f (t) dν(t),
2. Preliminaries
In this section, we recall and collect the necessary definitions and theorems that
will be used throughout the paper. The only thing in this section that is really new
is Proposition 2.3.
We begin by defining the main objects of study of this work.
Definition 2.1. Let G be a locally compact group, and let p ∈ [1,∞). Denote by
Repp(G) the class of all contractive representations of L1(G) on Lp-spaces. The
full group Lp-operator algebra of G, denoted F p(G), is the completion of L1(G) in
the norm given by
kfkF p(G) = sup(cid:8)kπ(f )k : π ∈ Repp(G)(cid:9)
for f ∈ L1(G).
Denote by λp : L1(G) → B(Lp(G)) the left regular representation, which is given
by λp(f )ξ = f ∗ ξ for f ∈ L1(G) and ξ ∈ Lp(G). The algebra of p-pseudofunctions
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
5
on G (sometimes also called reduced group Lp-operator algebra of G), here denoted
F p
λ (G), is the completion of L1(G) in the norm
for f ∈ L1(G).
kfkF p
λ (G) = kλp(f )kB(Lp(G))
The algebra F p(G) has been defined in [Phi13] and [GT15b] as the completion
of L1(G) with respect to non-degenerate, contractive representations on Lp-spaces.
In the proposition below, we show that this distinction is irrelevant. We recall the
following, which is a particular case of a result from [GT16c].
Theorem 2.2. ([GT16c]). Let A be a Banach algebra with a left contractive ap-
proximate identity, let E be a reflexive Banach space, and let ϕ : A → B(E) be
a contractive homomorphism. Denote by E0 the essential subspace of ϕ, this is,
E0 = span ϕ(A)E. Then there exists a contractive idempotent e ∈ B(E) satisfying
e(E) = E0.
Proposition 2.3. Let p ∈ [1,∞) and let G be a locally compact group. If E is
an Lp-space and π : L1(G) → B(E) is a contractive representation, then there exist
an Lp-space F and a contractive, non-degenerate representation ϕ : L1(G) → B(F )
such that
for all f ∈ L1(G). In particular, Definition 2.1 agrees with the definitions given in
[Phi13] and [GT15b].
kπ(f )k ≤ kϕ(f )k
Proof. We treat the case p = 1 first. In this case, we may take ϕ to be the left regular
representation λ1. Indeed, let π : L1(G) → B(E) be a contractive representation
on an L1-space, and let (aj )j∈J be a contractive approximate identity for L1(G).
Then
kπ(f )k ≤ kfk1 = lim
j∈J kf ∗ ajk1 = lim
j∈J kλ1(f )ajk1 ≤ kλ1(f )k,
for all f ∈ L1(G), as desired.
Assume now that p > 1. Let π : L1(G) → B(E) be a contractive representation
on an Lp-space. Since E is reflexive and L1(G) has a contractive approximate
identity, it follows from Theorem 2.2 that there exists a contractive projection e ∈
B(E) such that e(E) is the essential subspace E0 of π. Then E0 is an Lp-space by
Theorem 6 in [Tza69]. Let ϕ : L1(G) → B(E0) be the restriction of π. Then ϕ is
a non-degenerate, contractive representation, and it is clear that kπ(f )k = kϕ(f )k
for all f ∈ L1(G).
(cid:3)
The following duality principle was established in Proposition 2.18 of [GT15b].
Proposition 2.4. Let G be a locally compact group, let p ∈ (1,∞), and let p′ be its
conjugate exponent. Then the inversion map on G extends to a canonical isometric
isomorphism F p(G) ∼= F p′
(G).
The next result is a combination of Theorem 2.30 and Corollary 3.20 in [GT15b].
The case 2 ≤ q ≤ p < ∞ is obtained using duality, so we omit it.
Theorem 2.5. Let G be a locally compact group.
identity map on L1(G) extends to a contractive homomorphism
If 1 ≤ p ≤ q ≤ 2, then the
γp,q : F p(G) → F q(G)
6
EUSEBIO GARDELLA AND HANNES THIEL
with dense range. In particular, kfkF q(G) ≤ kfkF p(G) for every f ∈ L1(G).
If, moreover, G is amenable, then:
(a) γp,q is injective.
(b) γp,q is surjective if and only if G is finite.
Remark 2.6. When p = 2, we have F p(G) = C ∗(G), the full group C ∗-algebra,
and F p
In particular, when G is
λ(G), the reduced group C ∗-algebra.
λ (G) = C ∗
abelian, then γp,2 : F p(G) → C ∗(G) ∼= C0(bG) is the Gelfand transform.
It follows from universality of the norm of F p(G) that the identity map on
L1(G) extends to a contractive homomorphism κp : F p(G) → F p
λ (G) with dense
range. The following is part of Theorem 3.7 in [GT15b]. This result has been
independently obtained by Phillips, and will appear in [Phi14].
Theorem 2.7. Let G be a locally compact group, and let p ∈ (1,∞). The following
are equivalent:
(1) The group G is amenable.
(2) The canonical map κp : F p(G) → F p
In view of the above result, we will identify F p(G) and F p
λ (G) is an (isometric) isomorphism.
λ (G) in a canonical
manner whenever G is amenable.
3. F p(Z) is not isomorphic to F q(Z)
The main step in the proof of Theorem 4.12 is to use the functoriality properties
studied in [GT14] to reduce the statement to the case where G is a cyclic group
(finite or infinite). While the case of a finite cyclic group will be dealt with using
spectral configurations introduced in [GT15a], the case of the infinite cyclic group
has to be dealt with separately, and we do so in this section. Our goal here is to
prove the following: for p, q ∈ [1,∞), there is an isometric isomorphism F p(Z) ∼=
F q(Z) as Banach algebras if and only if (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12) (which is equivalent
to p and q being either equal or Holder conjugate); see Theorem 3.6. In order to
prove this, we will reduce to the problem of ruling out the possibility of a canonical
isomorphism for the cyclic group Z2. In this case, the norm of a certain element
can be explicitly computed, yielding the result; see Proposition 3.3.
It should be pointed out that the methods of this section do not seem to gen-
eralize to other finite cyclic groups besides Z2. We will postpone dealing with
such groups until the next section; see Theorem 4.12, and specifically the proof of
Claim 2 in it. The approach we take here for the case of Z2 has the advantage of
being completely elementary and it does not depend on results from other works
(in particular, we do not need here anything about spectral configurations from
[GT15a]).
We begin by looking at the group Lp-operator algebra of a finite cyclic group,
which by amenability coincides with the algebra of p-pseudofunctions (this is the
presentation we actually use).
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
7
Example 3.1. Let n ∈ N and let p ∈ [1,∞). Then Lp-group algebra F p(Zn) is
the Banach subalgebra of B(ℓp
n) generated by the cyclic shift sn of order n
sn =
0
1
0
. . .
1
0
. . .
. . .
0
1
.
(The algebra B(ℓp
n) is just Mn with the Lp-operator norm.) It is easy to check that
F p(Zn) is algebraically isomorphic to Cn, but the canonical embedding F p(Zn) →
Mn is not as diagonal matrices.
In fact, computing the norm of an element in
F p(Zn) is challenging for p other than from 1 and 2, essentially because computing
p-norms of matrices that are not diagonal is difficult. Indeed, let ωn = e
n , and
set
2πi
If ξ = (ξ0, . . . , ξn−1) ∈ Cn ∼= F p(Zn), then its norm in F p(Zn) is
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1
ωn
ω2
n
...
1
1
1
...
1 ωn−1
n
1
ω2
n
ω4
n
...
ω2(n−1)
n
n
n
1
···
ωn−1
···
··· ω2(n−1)
. . .
··· ω(n−1)2
...
n
ξ0
ξ1
. . .
ξn−1
u−1
n
.
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)B(ℓp
.
n)
un =
1
√n
kξkF p(Zn) =
un
The matrix un is unitary (its inverse is the transpose of its conjugate), and hence
kξkF 2(Zn) = kξk∞. Moreover, if 1 ≤ p ≤ q ≤ 2, then k · kF q(Zn) ≤ k · kF p(Zn) by
Theorem 2.5. In particular, the norm k·kF p(Zn) always dominates the norm k·k∞.
The automorphism group of F p(Zn) is not easy to describe when p 6= 2, since
not every permutation of the coordinates of Cn is isometric. The next proposition
asserts that the cyclic shift on Cn is isometric.
Proposition 3.2. Let n ∈ N and let p ∈ [1,∞). Denote by τ : Cn → Cn the cyclic
forward shift, this is,
τ (ξ0, . . . , ξn−1) = (ξn−1, ξ0, . . . , ξn−2)
for all (ξ0, . . . , ξn−1) ∈ Cn. Then τ induces an isometric isomorphism F p(Zn) →
F p(Zn).
Proof. We follow the notation from Example 3.1, except that we write u in place
of un, and we write s in place of sn.
For ξ ∈ Cn, let d(ξ) denote the diagonal n × n matrix with d(ξ)j,j = ξj for
j = 0, . . . , n − 1. Denote by ρ : Cn → Mn the unital homomorphism given by
ρ(ξ) = ud(ξ)u−1 for ξ ∈ Cn. Then
kξkF p(Zn) = kρ(ξ)kB(ℓp
n) = kud(ξ)u−1kB(ℓp
n)
8
EUSEBIO GARDELLA AND HANNES THIEL
Set ω = (1, ω1
for all ξ ∈ Cn.
Given ξ ∈ Cn, one checks that
n, . . . , ωn−1
n
) ∈ Cn, and denote by ω its (coordinate-wise) conjugate.
d(τ (ξ)) = sd(ξ)s−1, us = d(ω)u,
and s−1u−1 = u−1d(ω).
It follows that
kτ (ξ)kF p(Zn) = kud(τ (ξ))u−1kB(ℓp
= kusd(ξ)s−1u−1kB(ℓp
= kd(ω)ud(ξ)u−1d(ω)kB(ℓp
n).
n, we conclude that
n)
n)
Since d(ω) and d(ω) are isometries of ℓp
kτ (ξ)kF p(Zn) = kd(ω)uτ (ξ)u−1d(ω)kB(ℓp
n) = kud(ξ)u−1kB(ℓp
n) = kξkF p(Zn),
as desired.
(cid:3)
The fact that F p(Z2) is isometrically isomorphic to F q(Z2) only in the "obvious"
cases can be proved directly by computing the norm of a special element, as we
show below.
Proposition 3.3. Let p, q ∈ [1,∞). Then F p(Z2) is isometrically isomorphic, as
Banach algebras, to F q(Z2) if and only if (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12).
Proof. The "if" implication follows from Proposition 2.4. We proceed to show the
"only if" implication. Given t ∈ [1,∞), we claim that
2.
k(1, i)kF t(Z2) = 2 1
By Proposition 2.4, the quantity on the left-hand side remains unchanged if one
replaces t with its conjugate exponent. Since the same holds for the quantity on
the right-hand side, it follows that it is enough to prove the claim for t ∈ [1, 2].
Let a be the matrix
Define a continuous function δ : [1, 2] → R by δ(t) = k(1, i)kF t(Z2) for t ∈ [1, 2].
− 1
t
a =
1
2(cid:18)1 + i
1 − i
1 + i(cid:19) .
1 − i
Then δ(t) = kakB(ℓt
to compute, and we have δ(1) = 2
satisfy
2) for all t ∈ [1, 2]. The values of δ at t = 1 and t = 2 are easy
2 and δ(2) = 1. Fix t ∈ (1, 2) and let θ ∈ (0, 1)
1
t
1 − θ
1
θ
2
=
+
.
1
Using the Riesz-Thorin Interpolation Theorem in [1, 2], we conclude that
δ(t) ≤ δ(1)1−θ · δ(2)θ = 2
1
2 ( 2
t
−1) · 1 = 2
1
t
− 1
2 .
For the converse inequality, fix t ∈ [1, 2] and consider the vector ξ = ( 1
0 ) ∈ ℓt
2.
1−i ). We have
2 ( 1+i
Then kξkt = 1 and ax = 1
1 − i(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)t
2(cid:18)1 + i
We conclude that
(cid:13)(cid:13)(cid:13)(cid:13)
1
=
1
2
(1 + it + 1 − it)
1
t = 2( 1
t
− 1
2 ).
δ(t) = kakB(ℓt
2) ≥ kaxkt
kξkt
= 2( 1
t
− 1
2 ).
This shows that δ(t) = 2( 1
t
− 1
2 ) for all t ∈ [1, 2], and the claim is proved.
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
9
Now let p, q ∈ [1,∞) and let ψ : F p(Z2) → F q(Z2) be an isometric isomorphism.
Since ψ is an algebra isomorphism, we must have either ψ(x, y) = (x, y) or ψ(x, y) =
(y, x) for all (x, y) ∈ C2. By Proposition 3.2, the flip (x, y) 7→ (y, x) is an isometric
isomorphism of F q(Z2), so we may assume that ψ is the identity map on C2. It
follows that k(1, i)kF p(Z2) = k(1, i)kF q(Z2), so (cid:12)(cid:12)(cid:12) 1
complete.
p − 1
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12) and the proof is
(cid:3)
Remark 3.4. Adopt the notation from the proof above. Then the function δ
attains the upper bound given by the Riesz-Thorin Interpolation Theorem, which
is a rare situation. This fortunate coincidence makes the argument possible, but it
is not clear to us how to generalize these computations to other cyclic groups, or
even to Z3. However, knowing the result just for Z2 is enough to prove Theorem 3.6.
The following is probably standard, but we have not been able to find a reference
in the literature. Accordingly, we prove it here.
Proposition 3.5. Let f : S1 → S1 be a homeomorphism. Then there exists ζ ∈ S1
such that f (−ζ) = −f (ζ).
Proof. Regard f as a homeomorphism h : [0, 2π] → [0, 2π] with either h(0) = 0 (if
f is orientation preserving) or h(0) = 2π (if f is orientation reversing). Without
loss of generality, assume that h(0) = 0, and hence that h is strictly increasing with
h(2π) = 2π. Let g : [0, π] → R be given by
g(t) =
h(t + π) − h(t)
π
for t ∈ [0, π]. We need to find s ∈ [0, π] with g(s) = 1. If h(π) = π, then g(π) = 1
and we are done.
If h(π) > π, we have g(0) > 1, while g(π) < 1. Likewise, if
h(π) < π, then g(0) < 1 and g(π) > 1. In either case, the conclusion then follows
from the Mean Value Theorem.
(cid:3)
We are now ready to prove that for p, q ∈ [1,∞), the algebras F p(Z) and F q(Z)
are (abstractly) isometrically isomorphic only in the trivial case when (cid:12)(cid:12)(cid:12) 1
(cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12) =
2(cid:12)(cid:12)(cid:12). This result should be compared with part (2) of Corollary 3.20 in [GT15b],
q − 1
here reproduced as Theorem 2.5, where only the canonical homomorphism is con-
sidered. The strategy will be to use Proposition 3.5 to reduce to the case when the
group is Z2, which is Proposition 3.3. The fact that the spectrum of F p(Z) is the
circle is crucial in our proof, and we do not know how to generalize these methods
to directly deal with, for example, Z2.
Theorem 3.6. Let p, q ∈ [1,∞). Then F p(Z) and F q(Z) are isometrically isomor-
p − 1
phic (as Banach algebras) if and only if (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12).
Proof. The "if" implication follows from Proposition 2.4. Let us show the converse.
By Proposition 3.13 in [GT15b], the maximal ideal spaces of F p(Z) and F q(Z) are
canonically homeomorphic to S1. We let Γp : F p(Z) → C(S1) denote the Gelfand
transform, which maps the canonical generating invertible isometry u ∈ F p(Z),
associated to 1 ∈ Z, to the canonical inclusion ι : S1 → C.
Let ϕ : F p(Z) → F q(Z) be an isometric isomorphism. Then ϕ induces a homeo-
morphism f : S1 → S1 that maps z ∈ S1 to the unique point f (z) ∈ S1 satisfying
evz ◦ ϕ = evf (z) : F p(Z) → C.
10
EUSEBIO GARDELLA AND HANNES THIEL
Use Proposition 3.5 to choose ζ0 ∈ S1 such that f (−ζ0) = −f (ζ0). Denote
by πp : F p(Z) → F p(Z2) and πq : F p(Z) → F q(Z2) the canonical homomorphisms
associated with the surjective map Z → Z2, which are quotient maps by The-
orem 2.5 in [GT14]. Let ωζ0 : F p(Z) → F p(Z) be the isometric isomorphism
induced by multiplying the canonical generator u ∈ F p(Z) by ζ0. Analogously,
let ωf (ζ0) : F q(Z) → F q(Z) be the isometric isomorphism induced by multiplying
u ∈ F q(Z) by f (ζ0). Then the following diagram commutes:
C(S1)
f ∗
/ C(S1)
Γp
Γq
F p(Z)
ωζ0
F p(Z)
ϕ /
F q(Z)
ωf (ζ0 )
/ F q(Z)
πp
F p(Z2)
/❴❴❴❴❴❴❴❴❴❴❴❴❴❴
bψ
πq
F q(Z2).
Define an isometric isomorphism ψ : F p(Z) → F q(Z) by ψ = ωf (ζ0) ◦ ϕ ◦ ω−1
.
It follows that ψ induces an isometric
One checks that ψ(ker(πp)) = ker(πq).
isomorphism bψ : F p(Z2) → F q(Z2). By Proposition 3.3, this implies that(cid:12)(cid:12)(cid:12) 1
(cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12), as desired.
4. Nonrepresentability on Lq-spaces
q − 1
p − 1
ζ0
2(cid:12)(cid:12)(cid:12) =
(cid:3)
The goal of this section is to show that for a nontrivial locally compact group
G and for p, q ∈ [1,∞) distinct, the Banach algebras F p(G), F p
λ (G), P Mp(G) and
CVp(G) can be represented on an Lq-space only in the trivial cases, namely if
either p = 2 and G is abelian; or if p and q are Holder conjugate; see Theorem 4.12.
The main step in the proof is to use the functoriality properties of these objects,
studied in [GT14], to reduce the statement to the case where G is a cyclic group
(finite or infinite). The case of a finite cyclic group will be dealt with using spectral
configurations as in [GT15a], as well as the canonical maps γp,q : F p(G) → F q(G)
constructed in Theorem 3.7 in [GT15b].
Proposition 4.1 below is our first preparatory result on representability of full
group Lp-operator algebras on Lq-spaces. We need some notation first. Let G
be a locally compact group, and denote by ∆ : G → R its modular function. For
f ∈ L1(G), let f ♯ : G → C be given by f ♯(s) = ∆(s−1)f (s−1) for all s ∈ G. It is
easy to check that the map ♯ : L1(G) → L1(G) is an anti-multiplicative isometric
linear map of order two.
Proposition 4.1. Let G be a locally compact group, and let p, q ∈ [1,∞) satisfy
p − 1
space.
2(cid:12)(cid:12)(cid:12). Suppose that F p(G) is isometrically representable on an Lq-
2(cid:12)(cid:12)(cid:12) > (cid:12)(cid:12)(cid:12) 1
(1) If p, q ∈ [1, 2] or p, q ∈ [2,∞), then the identity map on L1(G) extends to
an isometric isomorphism F p(G) ∼= F q(G).
(2) If p ∈ [1, 2] and q ∈ [2,∞), or if q ∈ [1, 2] and p ∈ [2,∞), then the map
♯ on L1(G) induces, when composed with the transpose map B(Lp(G)) →
B(Lp′
(G)), an isometric isomorphism F p(G) ∼= F q(G).
q − 1
(cid:12)(cid:12)(cid:12) 1
/
o
o
/
O
O
O
O
/
/
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
11
Proof. (1). The result is trivial when p = q. Suppose first that 1 ≤ p < q ≤ 2. Sup-
pose that there exist an Lq-space E and an isometric representation ϕ : F p(G) →
B(E). Let ιp : L1(G) → F p(G) be the canonical contractive inclusion with dense
range. Then
ψ = ϕ ◦ ιp : L1(G) → B(E)
is a contractive representation of L1(G) on an Lq-space.
Let f ∈ L1(G). Using Theorem 2.5 at the first step, we deduce that
kfkF q(G) ≤ kfkF p(G) = kψ(f )kB(F ) ≤ kfkF q(G).
Hence kfkF q(G) = kfkF p(G). Since f ∈ L1(G) is arbitrary, it follows that the
identity on L1(G) extends to an isometric isomorphism F p(G) → F q(G), as desired.
The case 2 ≤ q < p < ∞ follows from duality; see Proposition 2.4.
(2). We can assume, without loss of generality, that 1 ≤ p ≤ 2 ≤ q < ∞.
Denote by q′ ∈ [1, 2] the Holder conjugate exponent of q. By Proposition 2.4, the
map ♯ : L1(G) → L1(G), when composed with the transpose map B(Lq(G)) →
B(Lq′
(G). (The details
are in the proof of Lemma 2.16 of [GT15b], the main point being that given a
representation of G on an Lq-space, its dual representation is a representation of Gop
on Lq′
which induces the same norm on L1(G). One composes this isometric anti-
isomorphism with the map ♯ to get a multiplicative (isometric) isomorphism.) Since
the identity map on L1(G) extends to an isometric isomorphism F p(G) → F q′
(G)
by part (1), the result follows.
(cid:3)
(G)), extends to an isometric isomorphism F q(G) → F q′
Besides F p(G) and F p
λ (G), the other two Banach algebras we will be concerned
with, at least when p > 1, are the algebra P Mp(G) of p-pseudomeasures, and the
algebra CVp(G) of p-convolvers. These are, respectively, the ultraweak closure, and
the bicommutant, of F p
λ (G) in B(Lp(G)).
Algebras of pseudomeasures and of convolvers on groups have been thoroughly
studied since their inception by Herz in the early 70's; see [Her73]. It is clear that
P Mp(G) ⊆ CVp(G), and it is conjectured that they are equal for every locally
compact group G and every Holder exponent p ∈ (1,∞). The conjecture is known
to be true if p = 2, or if G is amenable ([Her73]), or, more generally, if G has the
approximation property ([Cow98]).
Our next goal is showing that these convolution algebras are never operator
algebras when p 6= 2. We first need two results about C ∗-algebras which are
interesting in their own right. The first one is well-known, and it follows, from
example, from Theorem 10 of [Bon54].
Theorem 4.2. Let A be a C ∗-algebra, let B be a Banach algebra, and let ϕ : A → B
be a contractive, injective homomorphism. Then ϕ is isometric.
Remark 4.3. If in the theorem above ϕ is not assumed to be injective, then the
conclusion is that it is a quotient map. On the other hand, we must assume that ϕ
is contractive, and not merely continuous, for the result to hold; counterexamples
are easy to construct. The result also fails for not necessarily self-adjoint operator
algebras, even for uniform algebras.
The next fact about C ∗-algebras is proved in [GT16c], and had not been noticed
before, at least not in this generality. The main difficulty is dealing with non-
degenerate representations, for which Theorem 2.2 is essential.
12
EUSEBIO GARDELLA AND HANNES THIEL
Theorem 4.4. ([GT16c]) Let A be a C ∗-algebra. Then the following are equivalent:
(1) A can be isometrically represented on an Lp-space, for some p ∈ [1,∞)\{2};
(2) A can be isometrically represented on an Lp-space, for all p ∈ [1,∞);
(3) A is commutative (and hence A ∼= C0(X) for X = Max(A)).
We also need to recall the following, whose proof can be found, for example,
in [GT14]. For a Banach algebra A, we denote its left (respectively, right, two-
sided) multiplier algebra by ML(A) (respectively, MR(A) and M (A)), and we write
ιL : A → ML(A) (respectively, ιR : A → MR(A) and ι : A → M (A)) for the canon-
ical inclusion. When confusion may arise, we write the product in M (A) with a
dot.
Theorem 4.5. Let A be a Banach algebra with a left (respectively, right, two-sided)
contractive approximate identity. Let X be a Banach space and let ϕ : A → B(X) be
a non-degenerate contractive representation. Then there exists a unique unital con-
tractive homomorphism ψL : ML(A) → B(X) (respectively, ψR : MR(A) → B(X)
and ψ : M (A) → B(X)) satisfying ψL ◦ ιL = ϕ (respectively, ψR ◦ ιR = ϕ and
ψ ◦ ι = ϕ). Moreover, the map ψL is given by
for all m ∈ ML(A), for all a ∈ A and for all ξ ∈ X.
ψL(m)(ϕ(a)(ξ)) = ϕ(m · a)(ξ)
The following theorem generalizes a result of Neufang and Runde, Theorem 2.2
in [NR09], where the authors assumed that the group in question was amenable
and had a closed infinite abelian subgroup. Observe that for G = Z, Theorem 4.6
follows from the main result of [GT16a], since there it is shown that for p 6= 2,
there exists a semisimple quotient AV of F p(Z) that is not representable on Lq
for any 1 ≤ q < ∞. Indeed, if F p(Z) were an operator algebra, so would be any
of its quotients. Recall that a unital commutative, semisimple operator algebra is
isometrically isomorphic to a closed subalgebra of C(X) for a compact Hausdorff
space X (that is, it is a uniform algebra). Hence, if F p(Z) were an operator algebra,
then AV would be a uniform algebra, hence isometrically representable on an Lq-
space for any q ∈ [1,∞).
Theorem 4.6 will be used in the proof of Theorem 4.12, which gives a much more
general result. In the proof below, for a locally compact group G and p ∈ [1,∞),
and to emphasize the role played by G, we will denote by γG
p,2 : F p(G) → C ∗(G)
the map from Theorem 2.5.
Theorem 4.6. Let G be a locally compact group and let p ∈ [1,∞). Then one
of F p(G), F p
λ (G), P Mp(G) or CVp(G) is an operator algebra if and only if either
p = 2 or G is the trivial group.
Proof. The "if" implication is obvious if G is the trivial group, since the associated
Banach algebras are all C, while the statement is clear if p = 2.
For the "only if" implication, it is clear that if either P Mp(G) or CVp(G) is an
operator algebra, then so is F p
λ (G), since there are isometric inclusions
F p
λ (G) ⊆ P Mp(G) ⊆ CVp(G).
Assume now that F p(G) is an operator algebra and that p 6= 2. There is a canon-
ical identification of F p(G) with C ∗(G), in view of Proposition 4.1 and Remark 2.6.
Let κp : F p(G) = C ∗(G) → F p
λ (G) denote the canonical contractive homomorphism
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
13
with dense range. It is well-known that the quotient C ∗(G)/ ker(κp) is a C ∗-algebra.
The induced map
cκp : C ∗(G)/ ker(κp) → F p
λ (G)
It is therefore enough to show the statement assuming that F p
metrically isomorphic to C ∗(G)/ ker(κp). In particular, F p
hence an operator algebra itself.
is an injective, contractive homomorphism. Now, Theorem 4.2 shows that cκp is
isometric. Since κp, and hence bκp, has dense range, it follows that F p
algebra. Let H be a Hilbert space and let ϕ : F p
representation.
Claim: F p
λ (G) is an operator
λ (G) → B(H) be an isometric
λ (G) is a C ∗-algebra. Note that A has a (two-sided) contractive
approximate identity, since so does L1(G) and there is a contractive homomorphism
ι : L1(G) → F p
λ (G) is iso-
λ (G) is a C ∗-algebra, and
λ (G) with dense range. Moreover, since
λ (G), η ∈ H}
{ϕ(a)η : a ∈ F p
is itself a Hilbert space, we may assume that the representation ϕ is non-degenerate.
It follows from Theorem 4.5 that the algebra M (F p
λ (G)) of multipliers
on F p
λ (G), is unitally representable on H.
By Corollary 2.5 in [GT16b], there is a canonical isometric identification of the
λ (G)) ⊆ B(F p
multiplier algebra M (F p
λ (G)) ⊆ B(F p
λ (G)), with the Banach algebra
C(F p
λ (G)) = {x ∈ B(Lp(G)) : xa, ax ∈ F p
λ (G) for all a ∈ F p
λ (G)} ⊆ B(Lp(G))
λ (G). Denote by ψ : C(F p
of centralizers of F p
isometric representation.
λ (G)) → B(H) the resulting unital,
There is an obvious identification of G with a subgroup of the invertible isome-
tries of C(F p
λ (G)), given by letting a group element g ∈ G act on Lp(G) as
the convolution operator with respect to the point mass measure δg. Now, for
g ∈ G, set ug = ψ(δg). Then ug is an invertible isometry on H, that is, a uni-
tary operator. Moreover, the map u : G → U(H), given by g 7→ ug, is easily
seen to be a strongly-continuous unitary representation of G on H. The inte-
grated form ρu : L1(G) → B(H) of u is therefore a contractive, non-degenerate
∗-homomorphism. Whence the subalgebra ρu(L1(G)) ⊆ B(H) is closed under the
adjoint operation. Moreover, it is clear that the following diagram commutes:
L1(G)
ι
❍
❍
❍
ρu
❍
❍
❍
❍
❍
F p
λ (G)
ϕ /
$❍
/ B(H).
It follows from Theorem 4.4 that F p
We conclude that ϕ(F p
B(H), that is, a C ∗-algebra. The claim is proved.
mutative, and in particular F q
[GT15b].
λ (G)), which equals ρu(L1(G)), is a closed ∗-subalgebra of
λ (G) is commutative. Thus G is itself com-
λ (G) = F q(G) for all q ∈ [1,∞), by Theorem 3.7 in
p,2 : F p
λ(G) is an isometric isomorphism by Proposition 4.1.
p,2 is surjective implies that G is finite, by Theorem 2.5. Using that
The fact that γG
γG
p,2 is isometric, we will show that G must be the trivial group.
λ (G) → C ∗
The map γG
Using finiteness of G,
let g ∈ G be an element with maximum order. Set
n = ord(g) ≥ 1, and let j : Zn ֒→ G be the group homomorphism determined
$
14
EUSEBIO GARDELLA AND HANNES THIEL
by j(1) = g. By Proposition 2.3 in [GT14], there are natural isometric embeddings
jp : F p
λ(G). Naturality of the maps involved
implies that the following diagram is commutative:
λ (Zn) ֒→ F p
λ(Zn) ֒→ C ∗
λ (G) and j2 : C ∗
F p
λ (Zn)
jp
F p
λ (G)
γ
Zn
p,2
γG
p,2
C ∗
λ(Zn)
/ C ∗
λ(G).
j2
In particular, γ Zn
really just the identity on Cn.)
λ (Zn) → C ∗
p,2 : F p
λ(Zn) is an isometric isomorphism. (This map is
2πi
Set ω = e
n ∈ S1. Using that p 6= 2 together with Proposition 2.8 in [GT15a]
(see also the comments above it), we conclude that if x ∈ F p
λ (Zn) is an invertible
isometry, then there exist ζ ∈ S1 and k ∈ {0, . . . , n − 1} such that, under the
algebraic identification of F p
λ (Zn) with Cn, we have
x =(cid:16)ζ, ζωk, . . . , ζωk(n−1)(cid:17) .
In particular, if n > 1, then not every element in (S1)n ⊆ Cn has norm one in
F p
λ(Zn), we must have n = 1. By the
λ (Zn). Since this certainly is the case in C ∗
choice of n, we conclude that G must be the trivial group, and the proof of the
theorem is finished.
(cid:3)
In contrast with Theorem 4.6, we point out that some Lp-operator group al-
gebras are contractively and isomorphically representable on Hilbert spaces. For
example, for any finite group G, abelian or not, and for any p ∈ [1,∞), the map
γp,2 : F p(G) → C ∗(G) from Theorem 2.5, is a contractive isomorphism.
For a locally compact group G, we review the definitions of the Banach al-
gebras M p(G) and M p
λ(G) from [GT14]. Recall that M 1(G) is the (unital) Ba-
nach algebra of finite complex Radon measures on G, and it can be identified
with the multiplier algebra of L1(G).
In particular, observe that the left regu-
lar representation λp : L1(G) → B(Lp(G)) can be extended to a representation
λp : M 1(G) → B(Lp(G)).
Definition 4.7. Let G be a locally compact group, and let p ∈ [1,∞). Define
M p(G) to be the completion of M 1(G) in the norm given by
kµkM p(G) = sup(cid:8)kπ(µ)k : π : M 1(G) → B(Lp(ν)) is a contractive homomorphism(cid:9)
for µ ∈ M 1(G).
λ(G) is the completion of M 1(G) in the norm
The algebra M p
kµkM p
λ (G) = kλp(µ)kB(Lp(G))
for µ ∈ M 1(G).
Remark 4.8. For p ∈ [1,∞), one can show that there are natural (isometric)
inclusions
and
λ(G) ⊆ M (F p
Somewhat less easy is the existence of an inclusion M p
see [GT14].
F p(G) ⊆ M p(G) ⊆ M (F p(G))
F p
λ (G) ⊆ M p
λ (G)).
λ(G) ⊆ CVp(G) for p ∈ (1,∞);
/
/
/
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
15
The following is a particular case of a result in [GT14].
Proposition 4.9. Let p ∈ [1,∞), let G be a locally compact group, and let H ⊆ G
be an amenable subgroup. Then the inclusion H ֒→ G induces canonical isometric
unital homomorphisms
M p
λ(H) → M p
λ(G)
and M p(H) → M p(G).
λ (G), M p
λ (H) → B(X).
The way in which the above proposition will be used is through the following
lemma. We state it and prove it in greater generality than needed here for use
elsewhere.
Lemma 4.10. Let p ∈ [1,∞) and let G be a non-trivial locally compact group.
Denote by Ap(G) any of the following algebras: F p(G), M p(G), F p
λ(G),
If Ap(G) is isometri-
P Mp(G) or CVp(G) (the last two in the case p > 1).
cally and non-degenerately representable on a Banach space X, then there exist
a cyclic subgroup H (finite or infinite) of G, and a unital isometric representation
ψ : F p
Proof. Choose a non-trivial element g ∈ G, and denote by H the (not necessarily
closed) cyclic subgroup of G generated by g. Since either F p(G) or F p
λ (G) is
isometrically a subalgebra of Ap(G) (see Remark 4.8), we may assume, without
loss of generality, that Ap(G) is either F p
λ (G) or F p(G). Denote by Bp(G) either
M p
λ (G)) or M p(G) (if Ap(G) = F p(G)). By Proposition 4.9,
and since H is amenable and discrete, the inclusion of H ⊆ G induces a canonical
isometric embedding F p
Let X be a Banach space and let ϕ : Ap(G) → B(X) be an isometric and non-
degenerate representation. Note that Ap(G) has a contractive approximate identity,
since so does L1(G) and there is a contractive homomorphism L1(G) → Ap(G)
with dense image. By Theorem 4.5, ϕ can be extended to an isometric unital
representation ψ : M (Ap(G)) → B(X). Since there are inclusions
λ(G) (if Ap(G) = F p
λ (H) ֒→ Bp(G).
F p
λ (H) ⊆ Bp(G) ⊆ M (Ap(G)),
λ (H) is the desired isometric unital representation of F p
λ (H)
(cid:3)
the restriction of ψ to F p
on X.
The next theorem will be needed in the proof of Theorem 4.12. Recall that the
algebra P Mp(G) of p-pseudomeasures is the closure of F p
λ (G) in B(Lp(G)) with
respect to the weak∗ topology (also called the ultraweak topology) induced by the
(G) given by the
pairing determined on simple tensors by
(canonical) identification of B(Lp(G)) with the dual of Lp(G)b⊗Lp′
ha, ξ ⊗ ηiB(Lp(G)),Lp(G) b⊗Lp′ (G) = haξ, ηiLp(G),Lp′ (G)
(G).
for all a ∈ B(Lp(G)), all ξ ∈ Lp(G) and all η ∈ Lp′
We denote by Gop the opposite group of G. With µ and ν as in subsection 1.1,
the inversion map ι : (G, µ) → (Gop, ν) is in fact a measure-preserving group iso-
morphism. In particular, the p-convolution algebras associated to G are canonically
isometrically isomorphic to those associated to Gop.
Theorem 4.11. Let G be a locally compact group, and let p ∈ (1,∞). Then there
are canonical isometric isomorphisms
F p(G) ∼= F p′
(G), M p(G) ∼= M p′
(G), F p
λ (G) ∼= F p′
λ (G),
16
EUSEBIO GARDELLA AND HANNES THIEL
M p
λ(G) ∼= M p′
λ (G), P Mp(G) ∼= P Mp′(G) and CVp(G) ∼= CVp′ (G).
Proof. For F p(G), this is Proposition 2.18 in [GT15b] (here recalled as Proposition 2.4),
while the proof for M p(G) is analogous, using M 1(G) instead of L1(G).
(Gop), M p
For the reduced versions, we will show that there are canonical isometric iso-
morphisms F p(G) ∼= F p′
λ (Gop), P Mp(G) ∼= P Mp′(Gop) and
CVp(G) ∼= CVp′ (Gop), since, by the remarks preceding this lemma, this implies the
result.
From now on, and to minimize confusion, we will write (G, µ) and (Gop, ν)
instead of G and Gop, to state explicitly which is the left Haar measure in each
case.
λ(G) ∼= M p′
Define an isometric anti-isomorphism φ : L1(G, µ) → L1(Gop, ν) by φ(f )(s) =
∆(s)f (s) for all f ∈ L1(G, µ) and all s ∈ G. Observe that the assignment ξ 7→
ξ ◦ ι defines an isometric isomorphism Lp′
(Gop, ν). We denote by
ψ : B(Lp(G, µ)) → B(Lp′
(Gop, ν)) the map given by
(G, µ) ∼= Lp′
ψ(a)(ξ)(s) = [a′(ξ ◦ ι)](s−1)
for all a ∈ B(Lp(G, µ)), for all ξ ∈ Lp′
(Gop, µ) and for all s ∈ G. (In other words,
ψ(a) is the transpose of a, once Lp′
(G, µ) is identified with Lp′
(Gop, ν) via the
assignment ξ 7→ ξ ◦ ι.) It is easy to check that ψ is an isometric anti-isomorphism.
We claim that the following diagram commutes:
L1(G, µ)
λG
p
φ
ψ
L1(Gop, ν)
λGop
p′
/ B(Lp′
(Gop, ν)).
B(Lp(G, µ))
Let f ∈ L1(G), let ξ ∈ Lp′
(Gop, ν) and let s ∈ G be given. Recall (see, for
example, Proposition 3.5 in [GT15b]) that the transpose of λG
p′ (φ(f ) ◦ ι).
(The function φ(f ) ◦ ι ∈ L1(G, µ) is denoted by f ♯ in [GT15b].) Using this at the
second step, we get
p (f ) is λG
ψ(λG
p (f ))(ξ)(s) = [λG
p (f )′(ξ ◦ ι)](s−1)
∆(t−1)f (t−1)ξ(st) dµ(t)
= [(φ(f ) ◦ ι) ∗ (ξ ◦ ι)](s−1)
=ZG
=ZG
f (t)ξ(st−1) dµ(t).
On the other hand,
λGop
p′
which proves the claim.
(φ(f ))(ξ)(s) = (φ(f ) ∗op ξ)(s)
=ZG
=ZG
∆(t)f (t)ξ(st−1) dν(t)
f (t)ξ(st−1) dµ(t),
/
/
/
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
17
It follows that there is a canonical isometric isomorphism F p
λ (Gop), im-
plemented by ψ. By taking double commutants, we deduce that ψ also implements
an isomorphism between CVp(G) and CVp′ (G). Extending the above diagram to
M 1(G, µ) and M 1(Gop, ν), we conclude that there is also a canonical isometric
isomorphism M p
λ (G) ∼= F p′
λ(G) ∼= M p′
λ (Gop).
To show the result for the algebras of pseudomeasures, we need to identify the ul-
(Gop, ν)). In order to do this, observe
traweak topologies on B(Lp(G, µ)) and B(Lp′
that the canonical isometric identification
(G, µ) ∼= Lp′
Lp(G, µ)b⊗Lp′
(G, µ)b⊗Lp(G, µ)
induces the isometric isomorphism ψ : B(Lp(G, µ)) → B(Lp′
(G, µ)). Commutativity
of the above diagram then implies that φ extends to an isometric isomorphism
between the ultraweak closures of L1(G, µ) and L1(Gop, ν), that is, to an isometric
isomorphism P Mp(G) ∼= P Mp′ (Gop).
(cid:3)
The folowing is one of the main results of this paper. It determines precisely when
one of the p-convolution algebras considered in the literature can be represented
on an Lq-space, for some q ∈ [1,∞). It can be interpreted as stating that the Lp-
and Lq-representation theories of a nontrivial group, are incomparable whenever
they are not "obviously" equivalent. We point out that this represents a significant
generalization of Theorem 2.2 in [NR09], where the authors only deal with the case
q = 2, and moreover assume that G is amenable and has a closed infinite abelian
subgroup.
The proof of Theorem 4.12 uses machinery from a number of other works. The
reader is referred to the third page in the introduction for an overview of the proof.
Theorem 4.12. Let G be a locally compact group, and let p, q ∈ [1,∞) with q >
1. Then one (or all) of F p(G), M p(G), F p
λ(G), P Mp(G) or CVp(G) is
isometrically representable on an Lq-space if and only if one of the following holds:
λ (G), M p
(1) G is the trivial group;
2 − 1
q(cid:12)(cid:12)(cid:12); or
p(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
2 − 1
(3) p = 2 and G is abelian.
(2) we have (cid:12)(cid:12)(cid:12) 1
space for any q ∈ [1,∞). On the other hand, the identity (cid:12)(cid:12)(cid:12) 1
Proof. We begin with the "if" implication. When G is the trivial group, all the
Banach algebras in the statement are C, which is clearly representable on an Lq-
equivalent to p and q being either equal or conjugate. The case p = q is trivial.
If p and q are conjugate, then Theorem 4.11 shows that all of the algebras in the
statement are (canonically) representable on Lq(G). Finally, if p = 2 and G is
abelian, then all of the Banach algebras in the statement are commutative C ∗-
algebras, and hence have the form C0(X) for some locally compact Hausdorff space
X. It is then an easy exercise to check that for any q ∈ [1,∞) and for any such
space X, there exist an Lq-space and an isometric representation of C0(X) on it.
λ(G) ⊆ CVp(G), F p
We turn to the "only if" implication. Since there are isometric inclusions
F p
λ (G) ⊆ M p
λ (G) ⊆ P Mp(G) ⊆ CVp(G) and F p(G) ⊆ M p(G)
(see Remark 4.8), we may assume that either F p
λ (G) or F p(G) is representable on
an Lq-space. By Proposition 2.4, and without loss of generality, we can assume
that p, q ∈ [1, 2]. Suppose that p = 2 and p 6= q. Then F p
λ(G) and
λ (G) = C ∗
2 − 1
p(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
2 − 1
q(cid:12)(cid:12)(cid:12) is
18
EUSEBIO GARDELLA AND HANNES THIEL
F p(G) = C ∗(G) are C ∗-algebras. By Theorem 4.4, these must be abelian C ∗-
algebras, whence G itself must be abelian.
Since the arguments for F p
λ (G) and F p(G) are very similar, we will carry them
out together until we have to distinguish the two cases. We therefore denote by
Ap(G) either F p
λ (G) or F p(G).
Now suppose that p 6= 2. Then q 6= 2 by Theorem 4.6. Let E be an Lq-space
and let ϕ : Ap(G) → B(E) be an isometric representation. Note that Ap(G) has
a contractive approximate identity, since so does L1(G) and there is a contractive
homomorphism L1(G) → Ap(G) with dense image. By Theorem 2.2, there exists a
contractive idempotent e ∈ B(E) such that e(E) is the essential subspace E0 of ϕ.
Then E0 is an Lq-space by Theorem 6 in [Tza69]. Denote by ϕ0 : Ap(G) → B(E0)
the restriction of ϕ. Then ϕ0 is a non-degenerate isometric representation. By
Lemma 4.10, there exist a cyclic subgroup H of G, and an isometric unital repre-
sentation ψ : F p
λ (H)) is isometrically isomorphic
to F q
λ (H) → B(E0). We claim that ψ(F p
λ (H). We divide the proof into two cases.
Case 1: H ∼= Z. Observe that ψ(F p
λ (Z)) and F q
λ (Z) are both Banach algebras
generated by an invertible isometry of an Lq-space and its inverse. Since q 6= 2,
by part (2) of Corollary 5.21 in [GT15a], F q
λ (Z) is the unique, up to (isomet-
ric) isomorphism, Banach algebra generated by an invertible isometry of an Lq-
space and its inverse whose Gelfand transform is not an isomorphism (isometric or
not). Since p 6= 2, the Gelfand transform of F p
λ (Z) is not an isomorphism, so the
same is true for ψ(F p
λ (Z)). It thus follows that there is an isometric isomorphism
ψ(F p
λ (H)) ∼= F q
Case 2: H ∼= Zr for r ≥ 2. By replacing H with a subgroup, we may assume
that r is prime. Denote by v ∈ F p
λ (Zr) the canonical invertible isometry generating
F p
λ (Zr), and set w = ψ(v), which is an invertible isometry of E0. Then
λ (H), as desired. This proves the first case.
sp(w) = sp(v) = {ζ ∈ S1 : ζ r = 1}.
Thus, the only admissible spectral configurations σ = (σn)n∈N for w are
• σ0 = sp(w) and σn = ∅ for n ≥ 1; or
• σ0 = sp(w) = σr and σn = ∅ for n ≥ 1 with n 6= r.
In the first case, there is an isometric isomorphism between F q(w, w−1) =
ψ(F p
λ (Zr)) and (Cr,k · k∞). This would then contradict part (5) of Theorem 3.5
in [GT15a], since p 6= 2. In the second case, by the definition of the norm k · kσ,q
on F q(σ) ∼= F q(w, w−1), there is an isometric isomorphism F q
λ (Zr)).
This finishes the proof of the claim. For later use, we stress the fact that, in the
case H = Zr, the isomorphism ψ(F p
λ (H) is in fact canonical, in the sense
that the spectral configurations of v and ψ(v) agree, and thus the isomorphism is
induced by the identity on ℓ1(H).
λ (Zr) → ψ(F p
λ (H)) ∼= F q
λ (H) ∼=
We deduce from the claim that there is an isometric isomorphism F p
λ (H). For H ∼= Z, Theorem 3.6 implies that p = q. For H ∼= Zr, we use the fact
F q
that the isomorphism can be chosen to be canonical to prove that p must equal q.
So suppose that p 6= q and F p
λ (Zr) canonically. Without loss of gener-
ality, we may assume that 1 ≤ p < q < 2. We claim that the map γp,2 : F p
λ (Zr) →
λ(Zr) from Theorem 2.5 is an isometric isomorphism. In view of Theorem 3.7 in
C ∗
λ (Zr) ∼= F q
for all f ∈ ℓ1(Zr).
θ ∈ (0, 1) satisfy 1
Let f ∈ ℓ1(Zr). Then kλ2(f )kB(ℓ2(Zr)) ≤ kλp(f )kB(ℓp(Zr)) by Theorem 2.5. Let
2 . By the Riesz-Thorin interpolation theorem, we have
p + 1−θ
q = θ
kλq(f )kB(ℓq(Zr)) ≤ kλ2(f )kθ
B(ℓ2(Zr))kλp(f )k1−θ
B(ℓp(Zr)).
Since kλp(f )kB(ℓp(Zr)) = kλq(f )kB(ℓq(Zr)), we conclude that
B(ℓ2(Zr)),
B(ℓp(Zr)) ≤ kλ2(f )kθ
kλp(f )kθ
Since γp,2 is an isometric isomorphism, F p(Zr) is an operator algebra. The result
(cid:3)
and hence kλp(f )kB(ℓp(Zr )) ≤ kλ2(f )kB(ℓ2(Zr)), as desired. This proves the claim.
now follows from Theorem 4.6.
Corollary 4.13. Let G be a locally compact group, and let p, q ∈ [1,∞). Then the
following are equivalent:
(1) There is an isometric isomorphism F p(G) ∼= F q(G);
(2) There is an isometric isomorphism F p
λ (G);
(3) When p > 1, there is an isometric isomorphism P Mp(G) ∼= P Mq(G);
(4) When p > 1, there is an isometric isomorphism CVp(G) ∼= CVq(G);
λ (G) ∼= F q
(5) (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12).
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
19
[GT15b], this is equivalent to showing that
kλp(f )kB(ℓp(Zr)) = kλ2(f )kB(ℓ2(Zr))
λ (G) ∼= F q
F p
λ (H)
In connection with [GT16b], we mention here that for the algebras of pseudofunc-
tions, the result above can be improved to moreover allow isomorphisms between
algebras associated to different groups, as follows: for locally compact groups G
and H, and for p, q ∈ [1,∞) not both equal to 2, there is an isometric isomorphism
if and only if (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12) and G is isomorphic to H. (For discrete groups,
λ .)
the same result holds with P Mp or CVp instead of F p
A variant of the techniques used to prove Theorem 4.12 can be used to rule out
representability on certain QSLp-spaces. (Recall that a Banach space E is a QSLp-
space if it is isometrically isomorphic to a subspace of a quotient of an Lp-space.
Group representations on QSLp-spaces are studied, for example, in [NR09].) We
are not able to use spectral configurations as in Theorem 4.12, since we do not
have a description of the Banach algebra generated by an invertible isometry of
a QSLp-space. Instead, we will use the fact (see Theorem 3.7 in [GT15b]) that
F p
λ (Zn) and F p
λ (Z) are universal with respect to representations on QSLp-spaces,
together with the maps γp,q, to obtain the result.
The theorem below is only stated for p, q ∈ [1, 2], while the other exponents can
be handled using duality.
Theorem 4.14. Let G be a nontrivial locally compact group, and let p, q ∈ [1, 2].
λ(G), P Mp(G) and CVp(G)
(1) If q ≤ p, then all of F p(G), M p(G), F p
(2) If p < q, then none of F p(G), M p(G), F p
can be isometrically represented on a QSLq-space.
λ (G), M p
λ(G), P Mp(G) and
λ (G), M p
CVp(G) can be isometrically represented on a QSLq-space.
20
EUSEBIO GARDELLA AND HANNES THIEL
Proof. If q ≤ p, then every Lp-space is isomorphic to a closed subspace of an Lq-
space. In particular, every Lp-space is a QSLq-space, and the result is immediate.
Suppose that p < q. As in the proof of Theorem 4.12, it is enough to show that
λ (G) are not representable on a QSLq-space. For convenience, we will
λ(G),
F p(G) and F p
denote by Ap(G) one of these algebras, and by Bp(G) either M p(G) or M p
depending on which one Ap(G) is denoting.
Let E be a QSLq-space and let ϕ : Ap(G) → B(E) be an isometric representation.
Since a subspace of a QSLq-space is obviously a QSLq-space, we can assume that ϕ
is non-degenerate. By Lemma 4.10, there are a cyclic subgroup H of G and a unital,
isometric representation ψ : F p
λ (H) → B(E). Unlike in the proof of Theorem 4.12,
we cannot really conclude that ψ(F p
λ (H) directly. Instead,
and using the notation from [GT15b], for f ∈ ℓ1(H), we have (explanations below):
λ (H)) is isomorphic to F q
kfkLp = kλp(f )kB(Lp(H)) = kψ(f )kB(E) ≤ kfkQSLq = kfkLq ≤ kfkLp.
The first step is Theorem 3.7 in [GT15b] for F p(H) (the group H is amenable); the
second one is the fact that ψ is isometric; the third one is the definition of the norm
k·kQSLq; the fourth one is the fact (Theorem 3.7 in [GT15b]) that F p
QS(H) = F p(H);
and the last one is Theorem 2.5.
It follows that the map γH
λ (H) is an isometric isomorphism. In
particular, F p
λ (H) is representable on an Lq-space, which contradicts Theorem 4.12.
The contradiction implies that Ap(G) cannot be isometrically represented on a
QSLq-space, as desired.
(cid:3)
p,q : F p
λ (H) → F q
We close this section with a question. Theorem 4.12 asserts that a group C ∗-
algebra can be represented on an Lp-space, for some p 6= 2, if and only if the
group is commutative. More generally, Theorem 4.4 shows that a C ∗-algebra can
be represented on an Lp-space, for some p 6= 2, if and only if it is commutative. This
may well be a particular case of a more general fact ruling out representability of
certain Lp-operator algebras on Lq-spaces for two different, nonconjugate, Holder
exponents p and q. On the other hand, the examples in [BLM95] show that there are
operator algebras, which are not C ∗-algebras, that can be represented on Lp-spaces
for p 6= 2. Therefore, we suggest:
Question 4.15. Let A be a unital Banach algebra, and suppose that the set
{v ∈ A : v is invertible and kvk = kv−1k = 1}
has dense linear span in A. (This guarantees that A is a C ∗-algebra if it is an
operator algebra.) Let p, q ∈ [1, 2] with p 6= q. Suppose that A can be isometrically
represented on an Lp-space and on an Lq-space. Is A commutative? Does it follow
that A ∼= C(X) for some compact Hausdorff space X?
5. An application to crossed products
Let X be a locally compact Hausdorff space, let G be a locally compact group,
and let α : G → Homeo(X) be a topological action. For p ∈ [1,∞), N. Christopher
Phillips defined in [Phi13] the full and the reduced Lp-operator crossed products
F p(G, X, α) and F p
λ (G, X, α) of the topological dynamical system (G, X, α), gener-
alizing the well established constructions in C ∗-algebras, which are the case p = 2.
In Question 8.2 of [Phi13], Phillips asked whether, for a topological dynamical
system (G, X, α) and for distinct p, q ∈ [1,∞), there are any non-zero continuous
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
21
homomorphisms
F p(G, X, α) → F q(G, X, α) or F p
λ (G, X, α) → F q
λ (G, X, α).
While Theorem 2.5 shows that there may in general exist such homomorphisms
(even contractive ones with dense range), in Corollary 5.11 we show that there is an
isometric isomorphism F p(G, X, α) → F q(G, X, α) or F p
λ (Gop, X, α)
λ (G, X, α) → F q
if and only if (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
q − 1
2(cid:12)(cid:12)(cid:12).
For the convenience of the reader, and since our notation is somewhat different,
we recall below the definitions of the crossed products (for Lp-operator algebras
other than C0(X)). Recall that for p ∈ [1,∞), we say that a Banach algebra A is
an Lp-operator algebra if it can be isometrically represented on an Lp-space. The
group of isometric automorphisms of A is denoted Aut(A), and is always endowed
with the strong topology.
The object we define next is the analog of the group algebra L1(G). All integrals
are taken with respect to a fixed left Haar measure.
Definition 5.1. Fix p ∈ [1,∞). Let α : G → Aut(A) be an action of a locally
compact group G on a Banach algebra A. Denote by L1(G, A, α) the Banach algebra
completion of the space of continuous compactly supported functions G → A with
respect to the L1-norm, with product given by twisted convolution, that is,
(a ∗ b)(g) =ZG
a(h)αh(b(h−1g)) dh
for a, b ∈ Cc(G, A, α) ⊆ L1(G, A, α) and g ∈ G.
Next, we define (regular) covariant representations and the associated (reduced)
crossed product.
Definition 5.2. Adopt the notation from the previous definition. A (contractive)
covariant representation of (G, A, α) on an Lp-space E is a pair (π, v) consisting of
a nondegenerate contractive homomorphism π : A → B(E) and an isometric group
representation v : G → B(E), satisfying the covariance condition
vgπ(a)vg−1 = π(αg(a))
π(a(g))vg(ξ) dg
for all g ∈ G and for all a ∈ A. Given such a covariant representation, its integrated
form is the nondegenerate contractive homomorphism π ⋊ v : L1(G, A, α) → B(E)
given by
(π ⋊ v)(a)(ξ) =ZG
for all a ∈ L1(G, A, α) and for all ξ ∈ E.
space E0, its associated regular covariant representation is the pair (π, λE0
Lp(G) ⊗p E0 ∼= Lp(G, E0) given by
Given a contractive nondegenerate representation π0 : A → B(E0) on an Lp-
p ) on
π(a)(ξ)(g) = π0(cid:0)αg−1 (a)(cid:1) (ξ(g))
and (λE0
for all a ∈ A, for all ξ ∈ Lp(G, E0), and for all g, h ∈ G.
Denote by Repp(G, A, α) the class of all contractive covariant representations of
(G, A, α) on Lp-spaces, and by RegRepp(G, A, α) the subclass of Repp(G, A, α) con-
sisting of the regular covariant representations. The full crossed product F p(G, A, α)
p )g(ξ)(h) = ξ(g−1h)
22
EUSEBIO GARDELLA AND HANNES THIEL
and the reduced crossed product F p
L1(G, A, α) in the following norms:
λ (G, A, α) are defined as the completions of
kakF p(G,A,α) = sup{k(π ⋊ v)(a)k : (π, v) ∈ Repp(G, A, α)}
and
kakF p
λ (G,A,α) = sup(cid:8)(cid:13)(cid:13)(π ⋊ λE0
p )(a)(cid:13)(cid:13) : (π, λE0
p ) ∈ RegRepp(G, A, α)(cid:9) ,
for all a ∈ L1(G, A, α).
Remark 5.3. In the context of the above definition, if A is an Lp-operator algebra,
then so will be F p(G, A, α) and F p
λ (G, A, α).
Lp-operator crossed products generalize group Lp-operator algebras, since for the
λ (G).
one point space X = {∗} we have F p(G,∗, id) = F p(G) and F p
Remark 5.4. Most results in [Phi13] assume that the algebra A is separable and
that the group is second countable, and conclude that a number of Banach algebras
can be represented on σ-finite Lp-spaces; see also Remark 1.18 in [Phi13]. (The
purpose of using σ-finite measure spaces is to apply Lamperti's theorem [Lam58].)
However, the theory can be developed without these countability assumptions, and
in fact the arguments in [Phi13] go through in general, except that one gets an Lp-
operator algebra that is not necessarily representable on a σ-finite Lp-space. This,
in particular, applies to Remark 4.6 and Proposition 4.8 in [Phi13].
λ (G,∗, id) = F p
We note that while Definition 5.2 assumes covariant representations to be non-
degenerate, this assumption is unnecessary whenever A has a (left or right) contrac-
tive approximate identity, as can be shown with essentially the same argument as in
Proposition 2.3. In fact, our first preparatory result guarantees the existence of an
approximate identity for full and reduced crossed products whenever the underlying
algebra has one.
Theorem 5.5. Let p ∈ [1,∞), let A be an Lp-operator algebra, let G be a lo-
cally compact group, and let α : G → Aut(A) be an action. Consider the following
statements.
(1) A has a left (right, two-sided) contractive approximate identity;
(2) F p(G, A, α) has a left (right, two-sided) contractive approximate identity;
(3) F p
λ (G, A, α) has a left (right, two-sided) contractive approximate identity.
Then (1) implies (2), and (2) implies (3).
Finally, if G is discrete, then also (3) implies (1), so they are all equivalent in
this case.
Proof. We prove the result for left contractive approximate identities; the proof for
right contractive approximate identities is analogous.
(1) implies (2). Let (aµ)µ∈Λ be a contractive approximate identity for A. Set
U = {U ⊆ G open, such that e ∈ U and U is compact}.
Given U ∈ U, let fU : G → [0,∞) be a continuous positive function with support
contained in U satisfying RG
fU (g)dg = 1. Define a partial order on U by setting
U ≤ V if V ⊆ U , and give U × Λ the partial order given by (U, µ) ≤ (V, ν) if
and only if U ≤ V and µ ≤ ν. Recall that there exists a canonical contractive
homomorphism ι : L1(G, A, α) → F p(G, A, α) with dense range. Write k · k1 for
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
23
the norm on both L1(G) and L1(G, A, α), and k · k for the norm on both A and
F p(G, A, α).
For (U, µ) ∈ U × Λ, set x(U,µ) = fU aµ (pointwise product), which is an element
in Cc(G, A, α) ⊆ L1(G, A, α). It is clear that
kι(x(U,µ))k ≤ kfU aµk1 = kfUk1kaµk ≤ 1
left contractive approximate identity for F p(G, A, α), it is enough to show that
for all (U, µ) ∈ U × Λ. Thus, in order to show that (cid:0)ι(x(U,µ))(cid:1)(U,µ)∈U ×Λ is a
(cid:0)x(U,µ)(cid:1)(U,µ)∈U ×Λ is a left contractive approximate identity for L1(G, A, α). For
this, it is enough to consider functions in L1(G) · A ⊆ L1(G, A, α), since these span
a dense subalgebra.
Let f ∈ L1(G) and a ∈ A be given. We claim that (cid:13)(cid:13)x(U,µ) ∗ (f a) − f a(cid:13)(cid:13) → 0 as
(U, µ) → ∞. If either f or a is zero, then so is their product and there is nothing to
show. Without loss of generality, we may assume that kfk1 = kak = 1. Let ε > 0.
We make the following choices:
(1) Using continuity of α, choose an open set U1 ⊆ G containing the unit of G
such that kαg(a) − ak < ε/3 for all g ∈ U1. Since G is locally compact, we
may assume that U1 has compact closure, so that U1 ∈ U.
(2) Using continuity of the left regular representation on L1(G), choose an open
set U2 ⊆ G containing the unit of G such thatRG(cid:12)(cid:12)f (g−1h) − f (h)(cid:12)(cid:12) dg < ε/3
for all g ∈ U2. Again, we may assume that U2 ∈ U.
fU (g)kαg(a) − akdg + kfUk1kfk1kaµa − ak
(3) Choose µ0 ∈ Λ such that kaµa − ak < ε/3 for all µ ≥ µ0.
Set U0 = U1 ∩ U2 ∈ U. Given (U, µ) ≥ (U0, µ0), we have
(cid:13)(cid:13)x(U,µ) ∗ (f a) − f a(cid:13)(cid:13) =ZG(cid:13)(cid:13)(cid:13)(cid:13)
x(U,µ)(g)αg((f a)(g−1h))dg − f (h)a(cid:13)(cid:13)(cid:13)(cid:13) dh
ZG
=ZG(cid:13)(cid:13)(cid:13)(cid:13)
fU (g)aµf (g−1h)αg(a)dg − f (h)a(cid:13)(cid:13)(cid:13)(cid:13) dh
ZG
=ZG(cid:13)(cid:13)(cid:13)(cid:13)
fU (g)(cid:0)f (g−1h)aµαg(a) − f (h)a(cid:1) dg(cid:13)(cid:13)(cid:13)(cid:13) dh
ZG
≤ZGZG
fU (g)(cid:13)(cid:13)f (g−1h)aµαg(a) − f (h)a(cid:13)(cid:13) dgdh
≤ZGZG
fU (g)(cid:13)(cid:13)f (g−1h)aµαg(a) − f (h)aµαg(a)(cid:13)(cid:13) dgdh
+ZGZG
+ZGZG
≤ kaµkkαg(a)kZGZU
+ kaµkkfk1ZU
≤ ε/3 + ε/3 + ε/3 = ε.
fU (g)(cid:12)(cid:12)f (g−1h) − f (h)(cid:12)(cid:12) dhdg
fU (g)kf (h)aµαg(a) − f (h)aµak dgdh
fU (g)kf (h)aµa − f (h)ak dgdh
We conclude that F p(G, A, α) has a left contractive approximate identity.
24
EUSEBIO GARDELLA AND HANNES THIEL
(2) implies (3). This is immediate since there is a contractive homomorphism
λ (G, A, α) with dense range; see Lemma 3.13 in [Phi13].
κp : F p(G, A, α) → F p
Assume now that G is discrete, and let us prove that (3) implies (1). De-
note by E : F p
λ (G, A, α) → A the faithful conditional expectation constructed in
Proposition 4.8 of [Phi13] (observe that separability is not necessary; see comments
above). Let (xµ)µ∈Λ be a contractive left approximate identity for F p
λ (G, A, α), and
set yµ = E(xµ) ∈ A for µ ∈ Λ. Using Remark 4.6 in [Phi13] (again, separability is
not needed there), we identify A with a subalgebra of F p
λ (G, A, α). For a ∈ A, we
have
µ∈Λkyµa − ak = lim
lim
µ∈Λ kE(xµ)a − ak = lim
µ∈Λ kE(xµa − a)k ≤ lim
µ∈Λ kxµa − ak = 0.
It follows that (yµ)µ∈Λ is an left approximate identity for A, and it is clear that
kyµk ≤ 1 for all µ ∈ Λ.
(cid:3)
Even when G is discrete, the crossed products F p(G, A, α) and F p
λ (G, A, α) from
the proposition above may not contain F p(G) and F p
λ (G) canonically; this happens
only when A is unital. In general, however, they are canonically subalgebras of the
multiplier algebras of the crossed products, as we show below. Recall our convention
that the product in multiplier algebras is written with a dot.
Theorem 5.6. Let p ∈ [1,∞), let A be an Lp-operator algebra with a left con-
tractive approximate identity, let G be a locally compact group, and let α : G →
Aut(A) be an action by isometric isomorphisms. For g ∈ G, define a linear map
Lg : L1(G, A, α) → L1(G, A, α) by
Lg(a)(h) = αg(a(g−1h))
for all a ∈ L1(G, A, α) and all h ∈ G. Then the assignment g 7→ Lg induces natural
contractive homomorphisms
ιG
L : F p(G) → ML(F p(G, A, α)) and ιG
L,λ : F p
λ (G) → ML(F p
λ (G, A, α)).
When A has a right contractive approximate identity, for g ∈ G, the linear maps
Rg : L1(G, A, α) → L1(G, A, α), given by Rg(a)(h) = a(hg) for all a ∈ L1(G, A, α)
and all h ∈ H, define natural contractive homomoprhisms
ιG
R : F p(G) → MR(F p(G, A, α)) and ιG
R,λ : F p
λ (G) → MR(F p
λ (G, A, α)).
Finally, when A has a two-sided contractive approximate identity, then the above
maps define natural contractive homomorphisms
ιG : F p(G) → M (F p(G, A, α)) and ιG
λ : F p
λ (G) → M (F p
λ (G, A, α)).
Moreover, the maps ιG
L,λ, ιG
R,λ and ιG
λ are isometric.
Proof. We prove the theorem only for left approximate identities and left multiplier
algebras, but analogous proofs apply to the right and two-sided versions.
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
25
We claim that Lg is a left multiplier on L1(G, A, α), that is, Lg(a∗ b) = Lg(a)∗ b
for all a, b ∈ L1(G, A, α). Given a, b ∈ L1(G, A, α) and h ∈ G, we have
Lg(a ∗ b)(h) = αg(cid:18)ZG
a(k)αk(cid:0)b(k−1g−1h)(cid:1) dk(cid:19)
αg(a(k))αgk(cid:0)b((gk)−1h)(cid:1) dk
αg(a(g−1t))αt(cid:0)b(t−1h)(cid:1) dt
=ZG
=ZG
= (Lg(a) ∗ b)(h),
as desired. It is clear that g 7→ Lg defines a strongly continuous group homomor-
phism L : G → ML(L1(G, A, α)).
Claim 1: The homomorphism L extends to strongly continuous representations,
ιG and ιG
λ (G, A, α) by isometric left multipliers.
Let g ∈ G be given. To show that Lg extends to isometric automorphisms ιG
λ )g of F p(G, A, α) and F p
g and
λ (G, A, α), it is enough to show that for every covariant
λ , of G on F p(G, A, α) and F p
(ιG
representation (π, v) of (G, A, α) on an Lp-space, one has
k(π ⋊ v)(a)k = k(π ⋊ v)(Lg(a))k
for all a ∈ L1(G, A, α). To prove this, let (π, v) be a covariant representation on
an Lp-space E, and let a ∈ L1(G, A, α). For ξ ∈ E, we use the covariance identity
π(αg(x)) = vgπ(x)vg−1 at the third step, to get
(π ⋊ v)(Lg(a))(ξ) =ZG
=ZG
=ZG
π(Lg(a)(h))vh(ξ) dh
π(αg(a(g−1h)))vh(ξ) dh
vgπ(a(g−1h))vg−1h(ξ) dh
= vg(π ⋊ v)(a)(ξ).
The fact that ϕg and (ιG
Since vg is an isometry, it follows that k(π ⋊ v)(a)k = k(π ⋊ v)(Lg(a))k, as desired.
L,λ)g are left multipliers follows immediately from the
fact that Lg is a left multiplier, using an ε/3 argument. Finally, it is routine to
check that the assignments g 7→ (ιG
L,λ)g are strongly continuous
actions of G. The claim is proved.
L )g and g 7→ (ιG
Since L1(G) is universal with respect to strongly continuous isometric actions
L : L1(G) → ML(F p(G, A, α)) and
λ (G, A, α)), which admit explicit descriptions as follows. For
of G, there are contractive homomorphisms ιG
L,λ : L1(G) → ML(F p
ιG
f ∈ L1(G) and a ∈ L1(G, A, α), we have
L (f ) · a)(g) =ZG
f (h)Lh(a)(g) dh =ZG
(ιG
f (h)αh(a(h−1g)) dh,
for all g ∈ G. On the other hand, let κ : F p(G, A, α) → F p
λ (G, A, α) denote the
canonical contractive homomorphism with dense range. Since κ is the identity
on L1(G, A, α), by Theorem 5.5 it maps a contractive left approximate identity of
F p(G, A, α) to a contractive left approximate identity of F p
λ (G, A, α). Thus there
λ (G, A, α)), and we
exists a unique unital extension eκ : ML(F p(G, A, α)) → ML(F p
have ιG
L .
L,λ =eκ ◦ ιG
26
EUSEBIO GARDELLA AND HANNES THIEL
It remains to show that ιG
L,λ is isometric when L1(G) is endowed with the
norm of F p
λ (G). We begin with some general observations. Since F p(G, A, α) and
F p
λ (G, A, α) have left contractive approximate identities by Theorem 5.5, it follows
from Theorem 4.5 that any contractive, nondegenerate representation of any these
algebras extends to a contractive, unital representation of its left multiplier alge-
bra. Let (π, v) be a covariant representation of (G, A, α) on an Lp-space E. By
applying Theorem 2.2, we can assume that π (and hence π ⋊ v) is nondegenerate.
We write ^π ⋊ v : ML(F p(G, A, α)) → B(E) for the extension of π ⋊ v to the left
multiplier algebra, and similarly with ^π ⋊ v : ML(F p
λ (G, A, α)) → B(E) if (π, v) is
a regular covariant representation. By a slight abuse of notation, we denote also by
v : L1(G) → B(E) the integrated form of v, which is given by v(f )ξ =RG
f (g)vg(ξ)dg
for f ∈ L1(G) and ξ ∈ E.
Let f ∈ L1(G), let a ∈ L1(G, A, α), and let ξ ∈ E. Then
L (f ) · a](ξ)
L (f ) · a)(g)(cid:3) vg(ξ) dg
f (h)αh(a(h−1g)) dh(cid:21) vg(ξ) dg
π(cid:2)αh(a(h−1g))(cid:3) vg(ξ) dgdh
L (f )) = v(f ) for all f ∈ L1(G).
L (f )) [(π ⋊ v)(a)ξ] = (π ⋊ v)[ιG
Claim 2: We have ^π ⋊ v(ιG
(^π ⋊ v)(ιG
π(cid:2)(ιG
π(cid:20)ZG
f (h)ZG
f (h)ZG
f (h)ZG
=ZG
=ZG
=ZG
=ZG
=ZG
=ZG
π [αh(a(k))] vhk(ξ) dkdh
vhπ [a(k)] vk(ξ) dkdh
f (h)vh ((π ⋊ v)(a)ξ) dh
= v(f )((π ⋊ v)(a)ξ),
and the claim is proved.
Now suppose that (π, v) is a regular covariant representation, so that there exists
p ; see Definition 5.2. Let f ∈ L1(G), and observe
an Lp-space E0 such that v = λE0
that
kλE0
p (f )k = kλp(f ) ⊗ idE0k = kλp(f )k.
Using Claim 2 at the second step, and the above identity at the third step, we get
kιG
p ) ∈ RegRepp(G, A, α)}
L,λ(f )kM(F p
λ (G,A,α)) = sup{k(π ⋊ λE0
L (f ))k : (π, λE0
p )(ιG
p (f )k : (π, λE0
p ) ∈ RegRepp(G, A, α)}
We conclude that ιG
proof.
L,λ : F p
λ (G, A, α)) is isometric. This finishes the
(cid:3)
When p = 1, and regardless of whether G is amenable or not, there is a canonical
λ (G) (see Proposition 2.11 in [GT15b]), and hence the
identification F 1(G) = F 1
= sup{kλE0
= kλp(f )k
= kfkF p
λ (G).
λ (G) → ML(F p
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
27
L , ιG
R and ιG from Theorem 5.6 are isometric (because they agree with the
maps ιG
ones defined on F 1
λ (G)). However, when p > 1, it is not in general true that the
maps defined on F p(G) are isometric, or even injective with closed range, as we
explain in the following example.
Example 5.7. Let G be a discrete group, and let Lt denote the action of G
on c0(G) by left translation. By Theorem 4.3 in [GT15b], the canonical map
F p(G, G, Lt) → F p
λ (G, G, Lt) is an isometric isomorphism, regardless of G. De-
note by κp : F p(G) → F p
λ (G) the canonical contractive map with dense range, and
recall (Theorem 2.7) that κp is a (not necessarily isometric) isomorphism if and
only if G is amenable. Naturality of the maps involved implies that the following
diagram of unital homomorphisms commutes:
F p(G)
κp
F p
λ (G)
ιG
ιG
λ
/ M (F p(G, G, Lt))
∼=
/ M (F p
λ (G, G, Lt)),
where the vertical map on the right is the canonical one. Now, if ιG were injective
and had closed range, that is, if it were an isomorphism onto its range, then it
would follow that κp is an isomorphism, and hence that G is amenable.
Since an identical reasoning applies to left or right multiplier algebras, we con-
L , ιG
R
clude that for any non-amenable group G, and for the action Lt, the maps ιG
and ιG are not isomorphisms onto their ranges.
Let α : G → Aut(A) be an action of a locally compact group G on a Banach
algebra A. We denote by Aop the opposite Banach algebra, and write αop : G →
Aut(Aop) for the action given by αop
g = αg for all g ∈ G. When A is abelian, then
clearly A = Aop and α = αop.
We note that in L1(G, Aop, αop), convolution is performed using opposite product
on A, which we denote by ·op to minimize confusion. Recall that ∆ : G → R+
denotes the modular function (see subsection 1.1).
Proposition 5.8. Let p ∈ (1,∞), let A be an Lp-operator algebra, let G be a locally
compact group, and let α : G → Aut(A) be a continuous action. Then the map
θ : L1(G, A, α) → L1(G, Aop, αop)
given by θ(f )(s) = ∆(s−1)αs(f (s−1)), for all f ∈ L1(G, A, α) and all s ∈ G, is an
isometric anti-isomorphism, which moreover extends to isometric anti-isomorphisms
F p(G, A, α) ∼= F p′
(G, Aop, αop) and F p
λ (G, X, α) ∼= F p′
λ (G, Aop, αop).
In particular, F p(G, A, α) and F p
Lp′
-spaces.
λ (G, A, α) are anti-isometrically representable on
Proof. Be begin by showing that the map θ from the statement is an isometric
anti-isomorphism.
That θ is isometric follows from the definition of ∆. For f, g ∈ L1(G, A, α) and
s ∈ G, we have
θ(f ∗ g)(s) = ∆(s−1)αs(f ∗ g)(s−1) = ∆(s−1)αs(cid:18)ZG
f (t)αt(g(t−1s−1)) dt(cid:19) .
/
/
28
EUSEBIO GARDELLA AND HANNES THIEL
On the other hand, in the next computation we set s−1t = k at the fourth step to
get
(θ(g) ∗ θ(f )) (s) =ZG
=ZG
= ∆(s−1)αs(cid:18)ZG
= ∆(s−1)αs(cid:18)ZG
t (cid:0)θ(f )(t−1s)(cid:1) dt
θ(g)(t) ·op αop
αt(∆(st−1)αt−1s(f (s−1t))∆(t−1)αt(g(t−1)) dt
f (s−1t)αs−1t(g(t−1)) dt(cid:19)
f (k)αk(g(k−1s−1)) dk(cid:19) ,
which proves the claim.
Denote by ι : G → G the inversion map, which is anti-multiplicative. Let π and
u be representations of A and G, respectively, on an Lp-space E. By abuse of no-
tation, we denote by π′ : A → B(E′) and u′ : G → B(E′) the homomorphisms given
by π′(a) = π(a)′ and u′
g = (ug)′ for all a ∈ A and g ∈ G. Then π′ and u′ ◦ ι are
representations of Aop and G, respectively, on the Lp′
-space E′ (and conversely, by
reflexivity). Moreover, for g ∈ G and a ∈ A, the identity ugπ(a)ug−1 = π(αg(a))
is equivalent to u′
g = π′(αg(a)). It follows that (π, u) is a covariant rep-
resentation for (G, A, α) if and only if (π′, u′ ◦ ι) is a covariant representation for
(G, Aop, αop). This shows that the assignment (π, u) 7→ (π′, u′ ◦ ι) induces a nat-
ural bijection between the classes Repp(G, A, α) and Repp′ (G, Aop, αop). By the
definition of the norm on the full crossed product, we conclude that
g−1 π′(a)u′
kfkF p(G,A,α) = kθ(f )kF p′ (G,Aop,αop).
for all f ∈ L1(G, A, α). This proves the statement for full crossed products.
The case of reduced crossed products follows similarly: the above bijection re-
stricts to a bijection between RegRepp(G, A, α) and RegRepp′ (G, Aop, αop), since
for an Lp-space E0, the transpose of the representation λE0
(cid:3)
p
is λE ′
0
p′ ◦ ι.
Note that if a Banach algebra B is isometrically isomorphic to its opposite, then
any Banach algebra completion of B is also isomorphic to its opposite. We will
use this (trivial) observation in the next corollary, with B = L1(G, A, α) and the
completions being the full and reduced crossed products.
Corollary 5.9. Adopt the notation of Proposition 5.8, and suppose that A is
abelian. Then L1(G, A, α) is canonically isometrically isomorphic to L1(G, A, α)op,
and moreover there are natural isometric isomorphisms
F p(G, A, α) ∼= F p′
(G, A, α)
and F p
λ (G, A, α) ∼= F p′
λ (G, A, α).
Proof. Observe that A is also an Lp′
-operator algebra, since it is abelian. By
the first part of Proposition 5.8, and since A is abelian, the map θ is a natural
isometric isomorphism L1(G, A, α) ∼= L1(G, A, α)op. Upon taking completions with
respect to all covariant representations of (G, A, α), we conclude that F p(G, A, α)
is isometrically isomorphic to its opposite algebra. Composing this isomorphism
with the isomorphism F p(G, A, α)op ∼= F p′
(G, A, α) given by Proposition 5.8, we
obtain the desired isometric isomorphism for full crossed products.
The case of reduced crossed products is identical: one completes with respect to
(cid:3)
regular covariant representations instead.
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
29
The following is the main result of this section. When X is the one point
space, we recover Theorem 4.12. We point out that we do not know how to prove
Theorem 5.10 directly without first obtaining some form of Theorem 4.12, and that
Corollary 4.13 is not strong enough to deduce Theorem 5.10 from it.
q − 1
q − 1
p − 1
p − 1
Theorem 5.10. Let X be a locally compact Hausdorff space, let G be a nontrivial
locally compact group, and let α : G → Homeo(X) be a topological action. Given
p, q ∈ [1,∞) with q > 1, the Banach algebras F p(G, X, α) and F p
λ (G, X, α) can be
isometrically represented on an Lq-space if and only if one of the following holds:
λ (G, X, α) and F 2(G, X, α) are both
(2) G is abelian, p = 2 and the action α is trivial.
dual group of G. Then one easily checks that F 2
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12); or
(1) (cid:12)(cid:12)(cid:12) 1
Proof. If (cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12), then the conclusion follows from Corollary 5.9. Sup-
pose that p = 2, that G is abelian and that the action is trivial. Denote by bG the
isometrically isomorphic to C0(X × bG), so they are representable on an Lq-space,
for any q ∈ [1,∞). This proves the "if" implication.
Let us show the converse. We treat the case of reduced crossed products first.
Let p, q ∈ [1,∞) with q > 1, let E be an Lq-space and let ϕ : F p
λ (G, X, α) → B(E)
be an isomeric isomorphism. Since C0(X) has a (two-sided) contractive approx-
imate identity, so does F p
λ (G, X, α) by Theorem 5.5. Hence, upon restricting to
its essential subspace and using Theorem 2.2, we may assume that ϕ is non-
λ (G, X, α)) → B(E) be the extension of ϕ provided by
Theorem 4.5. By Theorem 5.6, there exists a canonical isometric homomorphism
λ : F p
ιG
λ is an isometric repre-
sentation of F p
q − 1
or p = 2 and G is abelian. Assuming the latter, F 2(G, X, α) is a C ∗-algebra.
Suppose, without loss of generality, that q 6= 2. Then F 2(G, X, α), and thus its
multiplier algebra M (F 2(G, X, α)), must be abelian by Theorem 4.4. For g ∈ G,
let ug ∈ M (F 2(G, X, α)) denote the canonical unitary implementing αg. For
a ∈ C0(X) ⊆ M (F 2(G, X, α)), we have
λ (G) on an Lq-space. By Theorem 4.12, either (cid:12)(cid:12)(cid:12) 1
λ (G, X, α)). We conclude that eϕ ◦ ιG
degenerate. Let eϕ : M (F p
λ (G) → M (F p
2(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12) 1
p − 1
2(cid:12)(cid:12)(cid:12)
αg(a) = ugau∗
g = a.
It follows that α is trivial, as desired. This shows the statement for F p
λ (G, X, α).
We prove the statement for full crossed products now. Let p, q ∈ [1,∞) with
q > 1, let F be an Lq-space and let ψ : F p(G, X, α) → B(F ) be an isomeric iso-
morphism. As before, we may assume that ψ is nondegenerate, and we denote by
eψ : M (F p(G, X, α)) → B(F ) its unital extension. Let g ∈ G \ {1}, and denote by
H ≤ G the (not necessarily closed) cyclic subgroup of G generated by g. Then
H is amenable and there is a commutative diagram of contractive homomorphisms
(explanations follow below)
F p(H)
ιH,G
M (F p(G))
❘
❘
❘
❘
❘
❘
❘
❘
ιH,G
λ
❘
❘
❘
❘
(❘
M (F p
λ (G))
f
ιG
f
ιG
λ
/ M (F p(G, X, α))
eκ
/ M (F p
λ (G, X, α)).
/
/
(
/
/
30
EUSEBIO GARDELLA AND HANNES THIEL
In the diagram above, ιH,G and ιH,G
λ
are the canonical isometric inclusions provided
λ are the canonical unital extensions of
map κ from full to reduced crossed product. Since ιG
by Remark 4.8 and Proposition 4.9;fιG andfιG
the maps constructed in Theorem 5.6; andeκ is the unital extension of the canonical
isfιG
λ . By commutativity of the diagram,eκ◦fιG ◦ ιH,G is an isometric representation
of F p(H) on the Lq-space F . By Theorem 4.12, we must have either (cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12) =
p − 1
2(cid:12)(cid:12)(cid:12), or p = 2. Assume that p = 2 and q 6= 2. Then F p(G, X, α) is a C ∗-algebra,
(cid:12)(cid:12)(cid:12) 1
q − 1
so it must be abelian by Theorem 4.4. Since C ∗
algebra M (F 2
argument used before shows that α must be trivial. This finishes the proof.
λ(G) embeds into the abelian C ∗-
λ (G, X, α)), the group G itself must be abelian. Finally, the same
(cid:3)
λ is isometric by Theorem 5.6, so
Finally, the following corollary asserts that the Lp-crossed products obtained
from topological dynamical systems, for varying p, are pairwise non-isometrically
isomorphic, except for conjugate exponents.
Corollary 5.11. Let X be a locally compact Hausdorff space, let G be a locally
compact group, and let α : G → Homeo(X) be a topological action. Given p, q ∈
[1,∞), the following conditions are equivalent:
(1) F p(G, X, α) is isometrically isomorphic to F q(G, X, α);
(2) F p
λ (G, X, α);
λ (G, X, α) is isometrically isomorphic to F q
p − 1
q − 1
(3) (cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) 1
2(cid:12)(cid:12)(cid:12).
References
[BLM95] D. Blecher and C. Le Merdy, On quotients of function algebras and operator algebra
structures on ℓp, J. Operator Theory 34 (1995), no. 2, 315–346.
[Bon54] F. Bonsall, A minimal property of the norm in some Banach algebras, J. London Math.
Soc. 29(1954), 156–164.
[Cow98] M. Cowling, The predual of the space of convolutors on a locally compact group, Bull.
Austral. Math. Soc. 57 (1998), 409–414.
[Dal00] G. Dales, Banach Algebras and Automatic Continuity, London Mathematical Society
Monographs, New Series, no. 24, The Clarendon Press, Oxford University Press, New
York, 2000.
[DS13] M. Daws and N. Spronk, The approximation property implies that convolvers are pseudo-
measures, Preprint. (arXiv:1308.1073 [math.FA]), 2013.
[GT14]
, Functoriality of group algebras acting on Lp-spaces, preprint, arXiv:1408.6137,
2014.
[GT15a] E. Gardella and H. Thiel, Banach algebras generated by an invertible isometry of an
Lp-space. J. Funct. Anal. 269 (2015), 1796–1839.
[GT15b]
, Group algebras acting on Lp-spaces, J. Fourier Anal. Appl. 21 (2015), 1310–
1343.
[GT16a]
, Quotients of Banach algebras acting on Lp-spaces, Adv. Math. 296 (2016),
85–92.
[GT16b]
[GT16c]
, Isomorphisms of algebras of convolution operators, in preparation, 2016.
, Extensions of representations of Banach algebras to biduals, in preparation,
2016.
[Her73] C. Herz, Harmonic synthesis for subgroups, Ann. Inst. Fourier (Grenoble) 23 (1973),
91–123.
[Lam58] J. Lamperti, On the isometries of certain function-spaces, Pacific J. Math. 8 (1958),
459–466.
[NR09] M. Neufang and V. Runde, Column and row operator spaces over QSLp-spaces and their
use in abstract harmonic analysis, preprint, arxiv:0711.2057, 2007.
REPRESENTATIONS OF p-CONVOLUTION ALGEBRAS ON Lq-SPACES
31
[Phi13] N. C. Phillips, Crossed products of Lp operator algebras and the K-theory of Cuntz alge-
bras on Lp spaces, preprint, arXiv:1309.6406, 2013.
[Phi14]
, Multiplicative domain for Lp operator algebras, in preparation. Draft of May
2014.
[Run05] V. Runde, Representations of locally compact groups on QSLp-spaces and a p-analog of
the Fourier-Stieltjes algebra, Pacific J. Math. 221 (2005), 379–397.
[Spa12] P. Spain, Representations of C ∗-algebras in dual & right dual Banach algebras. Houston
J. Math. 41 (2015), no. 1, 231–263.
[Tza69] L. Tzafriri, Remarks on contractive projections in Lp-spaces. Israel J. Math. 7 1969 9–15.
Eusebio Gardella Mathematisches Institut, Fachbereich Mathematik und Informatik
der Universitat Munster, Einsteinstrasse 62, 48149 Munster, Germany.
E-mail address: [email protected]
URL: www.math.uni-muenster.de/u/gardella/
Hannes Thiel Mathematisches Institut, Fachbereich Mathematik und Informatik der
Universitat Munster, Einsteinstrasse 62, 48149 Munster, Germany.
E-mail address: [email protected]
URL: www.math.uni-muenster.de/u/hannes.thiel/
|
1206.1764 | 2 | 1206 | 2012-06-26T16:44:22 | Constructive Analysis in Infinitely many variables | [
"math.FA",
"math-ph",
"math-ph"
] | In this paper we investigate the foundations for analysis in infinitely-many (independent) variables. We give a topological approach to the construction of the regular $\s$-finite Kirtadze-Pantsulaia measure on $\R^\iy$ (the usual completion of the Yamasaki-Kharazishvili measure), which is an infinite dimensional version of the classical method of constructing Lebesgue measure on $\R^n$ (see \cite{YA1}, \cite{KH} and \cite{KP2}). First we show that von Neumann's theory of infinite tensor product Hilbert spaces already implies that a natural version of Lebesgue measure must exist on $\R^{\iy}$. Using this insight, we define the canonical version of $L^2[\R^{\iy}, \la_{\iy}]$, which allows us to construct Lebesgue measure on $\R^{\iy}$ and analogues of Lebesgue and Gaussian measure for every separable Banach space with a Schauder basis. When $\mcH$ is a Hilbert space and $\la_{\mcH}$ is Lebesgue measure restricted to $\mcH$, we define sums and products of unbounded operators and the Gaussian density for $L^2[\mcH, \la_{\mcH}]$. We show that the Fourier transform induces two different versions of the Pontryagin duality theory. An interesting new result is that the character group changes on infinite dimensional spaces when the Fourier transform is treated as an operator. Since our construction provides a complete $\s$-finite measure space, the abstract version of Fubini's theorem allows us to extend Young's inequality to every separable Banach space with a Schauder basis. We also give constructive examples of partial differential operators in infinitely many variables and briefly discuss the famous partial differential equation derived by Phillip Duncan Thompson \cite{PDT}, on infinite-dimensional phase space to represent an ensemble of randomly forced two-dimensional viscous flows. | math.FA | math | CONSTRUCTIVE ANALYSIS IN INFINITELY MANY
VARIABLES
TEPPER L. GILL, G. R. PANTSULAIA, AND W. W. ZACHARY*
2
1
0
2
n
u
J
6
2
]
.
A
F
h
t
a
m
[
2
v
4
6
7
1
.
6
0
2
1
:
v
i
X
r
a
1991 Mathematics Subject Classification. Primary (45) Secondary(46) .
Key words and phrases. infinite-dimensional Lebesgue measure, Gaussian measure,
Fourier transforms, Banach space, Pontryagin duality theory, partial differential operators.
*deceased.
1
2
GILL, PANTSULAIA, AND ZACHARY
Abstract. In this paper we investigate the foundations for analysis in
infinitely-many (independent) variables. We give a topological approach
to the construction of the regular σ-finite Kirtadze-Pantsulaia measure
on R∞ (the usual completion of the Yamasaki-Kharazishvili measure),
which is an infinite dimensional version of the classical method of con-
structing Lebesgue measure on Rn (see [YA1], [KH] and [KP2]). First
we show that von Neumann's theory of infinite tensor product Hilbert
spaces already implies that a natural version of Lebesgue measure must
exist on R∞. Using this insight, we define the canonical version of
L2[R∞, λ∞], which allows us to construct Lebesgue measure on R∞ and
analogues of Lebesgue and Gaussian measure for every separable Ba-
nach space with a Schauder basis. When H is a Hilbert space and λH
is Lebesgue measure restricted to H, we define sums and products of
unbounded operators and the Gaussian density for L2[H, λH]. We show
that the Fourier transform induces two different versions of the Pon-
tryagin duality theory. An interesting new result is that the character
group changes on infinite dimensional spaces when the Fourier transform
is treated as an operator. Since our construction provides a complete
σ-finite measure space, the abstract version of Fubini's theorem allows
us to extend Young's inequality to every separable Banach space with
a Schauder basis. We also give constructive examples of partial dif-
ferential operators in infinitely many variables and briefly discuss the
famous partial differential equation derived by Phillip Duncan Thomp-
son [PDT], on infinite-dimensional phase space to represent an ensemble
of randomly forced two-dimensional viscous flows.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
Contents
Introduction
Historical Background
Purpose
Summary
1. Why λ∞ Must Exist
2.
Lebesgue Measure on R∞I
2.1. The Construction
2.2. The Extension to R∞I
2.3. Separable Banach Spaces
2.4. Translations
2.5. Gaussian measure
2.6. Rotational Invariance
Discussion
3. Operators
3.1. Bounded Operators on H2
⊗
3.2. Unbounded Operators on H2
⊗
4. Function Spaces
4.1. L1-Theory
5. Fourier Transform Theory
Background
3
4
6
10
10
11
14
14
18
25
28
29
32
35
35
35
38
43
44
46
46
4
GILL, PANTSULAIA, AND ZACHARY
5.1. Pontryagin Duality Theory I
5.2. L2-Theory
5.3. Pontryagin Duality Theory II
5.4. Lp-Theory
6. Partial Differential Operators (Examples)
6.1. Discussion
7. Conclusion
Acknowledgments
References
47
49
50
52
56
60
61
61
61
Introduction
On finite-dimensional space it is useful to think of Lebesgue measure in
terms of geometric objects (e.g.,volume, surface area, etc.). Thus, it is nat-
ural to expect that this measure will leave these objects invariant under
translations and rotations, so that rotational and translational invariance is
an intrinsic property of Lebesgue measure. However, we then find ourselves
disappointed when we try to use this property to help define Lebesgue mea-
sure on R∞. A more fundamental problem is that the natural Borel algebra
for R∞, B[R∞], does not allow an outer measure (since the measure of any
open set is infinite).
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
5
The lack of any definitive understanding of the cause for this lack of invari-
ance on R∞ has led some researchers to believe that it is not possible to have
a reasonable version of Lebesgue measure on R∞ (see, for example, DaPrato
[DP] or Bakhtin and Mattingly [BM]). In many applications, the study of
infinite dimensional analysis is restricted to separable Hilbert spaces, using
Gaussian measure as a replacement for (the supposed nonexistent) Lebesgue
measure. In some cases the Hilbert space structure arises as a natural state
space for the modeling of systems. In other cases, both the Hilbert spaces
and probability measures are imposed for mathematical convenience and
are physically artificial and limiting. However, all reasonable models of in-
finite dimensional (physical) systems require some functional constraint on
the effects of all but a finite number of variables. Thus, what is needed, in
general, is the imposition of constraints on the functions while preserving
the modeling freedom associated with infinitely-many independent variables
(in some well-defined sense). Any attempt to solve this problem necessarily
implies a theory of Lebsegue measure on R∞.
Even if a reasonable theory of Lebsegue measure on R∞ exists, this is
not sufficient to make it useful in engineering and science.
In addition,
all the tools developed for finite-dimensional analysis, differential operators,
Fourier transforms, etc are also required. Furthermore, researchers need
operational control over the convergence properties of these tools. In par-
ticular, one must be able to approximate an infinite-dimensional problem as
6
GILL, PANTSULAIA, AND ZACHARY
a natural limit of the finite-dimensional case in a manner that lends itself
to computational implementation. This implies that a useful approach also
has a well-developed theory of convergence for infinite sums and products
of unbounded linear operators.
Historical Background
Research into the general problem of Lebesgue measure on infinite-
dimensional vector spaces and R∞ in particular, has a long and varied past,
with participants living in a number of different countries, during times
when scientific communication was constrained by war, isolation and/or na-
tional competition. These conditions allowed quite a bit of misinformation
and folklore to grow up around the subject, so that even experts may have
a limited view of the history. Our own experience suggest that at least a
brief survey of some important events is in order. (We do not claim com-
pleteness and apologize in advance if we fail to mention equally important
contributions.)
Early studies in infinite dimensional analysis focused on the foundations
of probability theory and had a broad base of participation. However, the
major inputs were made by researchers in Poland, Russia, and France, with
later contributions from the US. The first important advance of the general
theory was made in 1933 when Haar [HA] proved the following theorem:
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
7
Theorem 0.1. On every locally compact abelian group G there exists a non-
negative regular measure m (Haar measure) on G, which is not identically
zero and is translation invariant. That is, m(A+x) = m(A) for every x ∈ G
and every Borel set A in G.
This theorem stimulated interest in the subject and von Neumann [VN1]
proved that it is the only locally finite left-invariant Borel measure on the
group (uniqueness up to a mulitplicative constant). Weil [WE] developed an
axiomatic approach to the subject, made a number of important refinements
and, proved the "Inverse Weil theorem" (in moderm terms):
Theorem 0.2. If G is a (separable) topological group and m is a transla-
tion invariant Borel measure on G, then it is always possible to define an
equivalent locally compact topology on G.
In 1946, Oxtoby [OX] initiated the study of translation-invariant Borel
measures on Polish groups (i.e., complete separable metric groups). In this
paper, Oxtoby provides a proof of the following result which he attributes
to Ulam:
Theorem 0.3. Let G be any complete separable metric group which is not
locally compact, and let m be any left-invariant Borel measure in G. Then ev-
ery neighborhood contains an uncountable number of disjoint mutually con-
gruent sets of equal finite positive measure.
Stated another way, he proved that
8
GILL, PANTSULAIA, AND ZACHARY
Theorem 0.4. There always exists a left-invariant Borel measure on any
Polish group which assigns positive finite measure to at least one set and
vanishes on singletons. However, a locally finite measure is possible if and
only if the group is locally compact.
(In 1967, Vershik [V] proved a related result for probability measures.)
Apparently uninformed of Oxtoby's work, In 1959 Sudakov [SU] indepen-
dently proved a special case of Theorem 0.4: If R∞ is regarded as a linear
topological space, then there does not exist a σ-finite translation-invariant
Borel measure for R∞. In 1964, Elliott and Morse [EM] developed a general
theory of translation invariant product measures (non-σ-finite) and, in 1965,
C. C. Moore [MO] initiated the study of measures that are translation in-
variant with respect to vectors in R∞0 (i.e., the set of sequences that are zero
except for a finite number of terms). This work was extended and refined
by Hill [HI] in 1971.
Motivated by Kakutani's work on infinite product measures [KA], a num-
ber of young Japanese researchers entered the field. In 1973, Hamachi [HA]
made major improvements on Hill's work which, indirectly suggested the
problem of identifying the largest group T, of admissible translations in the
sense of invariance for any σ-finite Borel measure µ on R∞ which assigns
2 , 1
2 ]ℵo and is metrically transitivity with respect to
the value of one to [− 1
R∞0 (equivalently, for each A with µ(A) > 0, there is a sequence (hk) ∈ R∞0
such that µ(R∞ \ ∪∞k=1(A + hk)) = 0).
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
9
Yamasaki [YA1] solved this problem in 1980. Unaware of the Yamasaki's
proof,
Kharazishvili independently solved the same problem in 1984. In 1991 Kir-
tadze and Pantsulaia [KP1] provided yet another solution (see also Pantsu-
laia [PA]). Finally, In 2007, Kirtadze and Pantsulaia proved that, if µ is the
completion of the measure µ, then: (see [KP2])
Theorem 0.5. The measure µ is the unique regular σ-finite measure on R∞
(uniqueness up to a mulitplicative constant), which is assigns the value one
2 , 1
to the set [− 1
has the metrically transitivity property with respect to ℓ1.
2 ]ℵo, is invariant under translations from the group ℓ1 and
In the mean time, in 1991 Baker [BA1], unaware of the Elliott-Morse
measures, dropped the requirement that the measure be σ-finite and con-
structed a translation invariant measure, ν, on R∞ (see also Baker (2004),
[BA2]). In 1992, Ritter and Hewitt [RH] constructed a translation invariant
measure related to that of Elliott Morse.
Starting in 2007, A. M. Vershik (see [V1], [V2], [V3] and references con-
tained therein) started an investigation of an infinite-dimensional analogue
of Lebesgue measure that is constructed in a different manner than that
studied in the previous papers. Roughly stated, he considers the weak limit
as n → ∞ of invariant measures on certain homogeneous spaces (hypersur-
faces of high dimension) of the Cartan subgroup of the Lie groups SL(n, R)
(i.e., the subgroups of diagonal matrices with unit determinant). Vershik's
10
GILL, PANTSULAIA, AND ZACHARY
measure is also unique and invariant under the multiplicative group of pos-
itive functions, suggesting that a logarithmic transformation may lead to a
version of the measure in this paper. (The paper of Vandev [VA] should also
be consulted.)
Purpose. The purpose of this paper is to show that a minor change in
the way we represent R∞ makes it possible to construct a σ-finite regular
version of Lebesgue measure using basic methods of measure theory from Rn.
Since the measure is regular, it turns out to be the Kirtadze and Pantsulaia
[KP1] measure, which is unique (see Theorem 0.5). Using our approach, we
construct an analogue of both Lebesgue and Gaussian measure (countably
additive) on every (classical) separable Banach space with a Schauder basis.
The version of Gaussian measure constructed is also rotationally invariant
(a property not shared by Wiener measure). This approach also allows us
to satisfy all the requirements of a useful infinite dimensional theory.
Summary. In the first section, we show how von Neumann's infinite ten-
sor product Hilbert space theory implies that a natural version of Lebesgue
measure must exist on R∞ and points to a possible approach. In the first
part of Section 2, we show that a slight change in thinking about the cause
for problems with unbounded measures on R∞ makes the construction of
Lebesgue measure not only possible, but no more difficult then the same
construction on Rn. (We denote it by R∞I , for reasons that are discussed in
this section.) We also provide natural analogues of Lebesgue and Gaussian
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
11
measure for every separable Banach space with a Schauder basis and show
that ℓ1 is the maximal translation invariant subspace. In the last part of
Section 2, we show that ℓ2 is the maximal rotation invariant subspace. In
Section 3, we study the convergence properties of infinite sums and products
of bounded and unbounded linear operators. In Section 4, we investigate
some of the function spaces over R∞I and in Section 5, we discuss Fourier
transforms and Pontryagin duality theory for Banach spaces. A major result
is that there are two different extensions of the Pontrjagin Duality theory
for infinite dimensional spaces. In this section, we also show that our the-
ory allows us to extend Young's inequality to ever separable Banach space
with a Schauder basis. In Section 6, we give some constructive examples
of partial differential operators in infinitely many variables. This allows us
to briefly discuss the famous partial differential equation derived by Phillip
Duncan Thompson [PDT], on infinite-dimensional phase space to represent
an ensemble of randomly forced two-dimensional viscous flows.
1. Why λ∞ Must Exist
In order to see that some reasonable version of Lebesgue measure must
exist, we need to review von Neumann's infinite tensor product Hilbert
space theory [VN2]. To do this, we first define infinite products of complex
numbers. (There are a number of other possibilities, see [GU] and [PA],
pg. 272-274.) In order to avoid trivialities, we always assume that, in any
product, all terms are nonzero.
12
GILL, PANTSULAIA, AND ZACHARY
Definition 1.1. If {zi} is a sequence of complex numbers indexed by i ∈ N
(the natural numbers),
every ε > 0, there is a finite set J(ε) such that, for all finite sets
(1) We say that the product Qi∈N zi is convergent with limit z if, for
J ⊂ N, with J(ε) ⊂ J, we have (cid:12)(cid:12)Qi∈J zi − z(cid:12)(cid:12) < ε.
(2) We say that the product Qi∈N zi is quasi-convergent if Qi∈N zi is
convergent. (If the product is quasi-convergent, but not convergent,
we assign it the value zero.)
We note that
(1.1)
0 <(cid:12)(cid:12)(cid:12)Yi∈N
zi(cid:12)(cid:12)(cid:12)
< ∞ if and only if Xi∈N 1 − zi < ∞.
Let Hi = L2[R, λ] for each i ∈ N and let H2
tensor product of von Neumann. To see what this object looks like:
⊗ = ⊗∞i=1L2[R, λ] be the infinite
Definition 1.2. Let g = ⊗i∈N
gi and h = ⊗i∈N
hi be in H2
⊗.
(1) We say that g is strongly equivalent to h (g ≡s h) if and only if
Pi∈N 1 − hgi, hiii < ∞ .
(2) We say that g is weakly equivalent to h (g ≡w h) if and only if
Pi∈N 1 − hgi, hiii < ∞.
Proofs of the following may be found in von Neumann [VN2] (see also
[GZ], [GZ1]).
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
13
Lemma 1.3. We have g ≡w h if and only if there exist zi, zi = 1, such
that ⊗i∈N
zigi ≡s ⊗i∈N
hi.
Theorem 1.4. The relations defined above are equivalence relations on
⊗, which decomposes H2
H2
spaces).
⊗ into disjoint equivalence classes (orthogonal sub-
Definition 1.5. For g = ⊗i∈N
⊗(g) to be the closed
subspace generated by the span of all h ≡s g and we call it the strong partial
tensor product space generated by the vector g. (von Neumann called it an
⊗, we define H2
gi ∈ H2
incomplete tensor product space.)
Theorem 1.6. For the partial tensor product spaces, we have the following:
(1) If hi 6= gi occurs for at most a finite number of i, then h = ⊗i∈N
hi ≡s
gi.
g = ⊗i∈N
(2) The space H2
⊗(g) is the closure of the linear span of h = ⊗i∈N
hi such
that hi 6= gi occurs for at most a finite number of i.
(3) If g = ⊗i∈Ngi and h = ⊗i∈Nhi are in different equivalence classes of
H2
⊗, then (g, h)⊗ =Qi∈N hgi, hiii = 0.
⊗(g)w = ⊕h≡wg (cid:2)H2
(4) H2
(5) For each g, H2
(6) For each g, H2
⊗(g)s is a separable Hilbert space.
⊗(g)w is not a separable Hilbert space.
⊗(h)s(cid:3) .
It follows from (6) that H2
⊗ = ⊗∞i=1L2[R, λ] is not a separable Hilbert
space.
14
GILL, PANTSULAIA, AND ZACHARY
From (5), we see that it is reasonable to define L2[R∞, λ∞] = H2
⊗(h)s, for
some h = ⊗∞i=1hi. This definition is ambiguous, but, in most applications,
the particular version does not matter. To remove the ambiguity, we should
identify a canonical version of h = ⊗∞i=1hi. Any reasonable version of λ∞
should satisfy λ∞(I0) = 1, where I = [−1
2 ] and I0 = ×∞i=1I.
2 , 1
Definition 1.7. If χI is the indicator function for I and hi = χI , we set h =
⊗∞i=1hi. We define the canonical version of L2[R∞, λ∞] = L2[R∞, λ∞](h)s.
2. Lebesgue Measure on R∞I
2.1. The Construction. We now have the problem of identifying the mea-
sure space associated with L2[R∞, λ∞](h)s. In the historical approach to the
construction of infinite products of measures {µk, k ∈ N} on R∞, the cho-
sen topology defines open sets to be the (cartesian) product of an arbitrary
finite number of open sets in R, while the remaining infinite number are
copies of R (cylindrical sets). The success of Kolmogorov's work on the
foundations of probability theory naturally led to the condition that µk(R)
be finite for all but a finite number of k (see [KO]). Thus, any attempt
to construct Lebesgue measure via this approach starts out a failure in the
beginning. However, Kolmogorov's approach is not the only way to induce
a total measure of one for the spaces under consideration.
Our definition of the canonical version of L2[R∞, λ∞] offers another ap-
proach. To see how, consider a simple extension of the theory on R. Let
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
15
2 ] and define RI = R × I1, where I1 = ∞×i=2
2 , 1
I = [− 1
Borel σ-algebra for R, let B(RI ) be the Borel σ-algebra for RI. For each
I.
If B(R) is the
set A ∈ B(R) with λ(A) < ∞, let AI be the corresponding set in B(RI ),
AI = A × I1. We define λ∞(AI ) by:
λ∞(AI ) = λ(A) ×
λ(I) = λ(A).
∞
Yi=2
We can construct a theory of Lebesgue measure on RI that completely par-
allels that on R. This suggests that we use Lebesgue measure and replace
the (tail end of the) infinite product of copies of R by infinite products of
copies of I. The purpose of this section is to provide such a construction.
Since we will be studying unbounded measures, for consistency, we use the
following conventions: 0 · ∞ = 0 and 0 · ∞∞ = ∞.
Recall that R∞ is the set of all x = (x1, x2,···), where xi ∈ R. This is a
linear space which is not a Banach space. However, it is a complete metric
space with metric given by:
d(x, y) =X∞
n=1
1
2n
xn − yn
1 + xn − yn
.
Remark 2.1. R∞ is a special case of a Polish space, which Banach called
a Fr´echet space i.e., a Polish space with a translation invariant metric (see
Banach [BA]). The topology generated by d(·,·) is generally known as the
Tychonoff topology.
For each n, define Rn
I = Rn × In, where In = ∞×i=n+1
I.
16
GILL, PANTSULAIA, AND ZACHARY
Definition 2.2. If An = A × In, Bn = B × In are any sets in Rn
define:
I , then we
(1) An ∪ Bn = A ∪ B × In,
(2) An ∩ Bn = A ∩ B × In, and
(3) Bc
n = Bc × In.
In order to avoid confusion, we always assume that I0 = ×∞i=1I ⊂ R1
I via the following class of open sets:
can now define the topology for Rn
I . We
On = {Un : Un = U × In, U open in Rn} .
2.1.1. Definition of R∞I . It is easy to see that Rn
I ⊂ Rn+1
I
. Since this is an
increasing sequence, we can define R′∞I by:
R′∞I = limn→∞Rn
I = ∞∪k=1
Rk
I .
Let τ1 be the topology on R′∞I = X1 induced by the class of open sets O
defined by:
O =
∞
[n=1
On =
∞
[n=1
{Un : Un = U × In, U open in Rn},
and let τ2 be topology on R∞ \ R′∞I = X2 induced by the metric d2, for
which d2(x, y) = 1, x 6= y and d2(x, y) = 0, x = y, for all x, y ∈ X2.
Definition 2.3. We define (R∞I , τ ) to be the sum (X1, τ1) and (X2, τ2), so
that every open set in (R∞I , τ ) is union of two disjoint sets G1 ∪ G2, where
G1 is open in (X1, τ1) and G2 is open in (X2, τ2).
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
17
It now follows from the above construction that R∞I = R∞ as sets. (How-
ever, they are not equal as topological spaces.) The following result shows
that convergence in the τ -topology always implies convergence in the Ty-
chonoff topology.
Theorem 2.4. If yk converges to x in the τ -topology, then yk converges to
x in the Tychonoff topology.
Proof. Case 1. If x ∈ R∞ \ R′∞I
k > N . Indeed, for a neighborhood of diameter 1
then there is N such that yk = x for all
2 about x, there is a N
such that d2(x, yk) < 1/2 for all k > N . This means that yk = x for k > N
({z : d2(x, z) < 1/2} only contains x), so that yk converges to x in the
Tychonoff topology.
Case 2. If x ∈ R′∞I and yk converges to x, then for any neighborhood
Un ⊂ On, there is N such that or all k > N, yk ∈ Un. This means that
yk ∈ R′∞I
for k > N , so that yk converges to x in the Tychonoff topology. (cid:3)
2.1.2. Definition of B(R∞I ). In a similar manner, if B(Rn
I ) is the Borel
σ-algebra for Rn
I (i.e., the smallest σ-algebra generated by the On), then
B(Rn
I ) ⊂ B(Rn+1
I
), so we can define B′(R∞I ) by:
B′(R∞I ) = limn→∞B(Rn
I ) = ∞∪k=1
B(Rk
I ).
If P(·) denotes a powerset of a set (i.e., P(A) = {X : X ⊆ A}), let B(R∞I )
be the smallest σ-algebra containing B′(R∞I ) ∪ P(R∞I \ ∪∞n=1Rn
(It is
I ).
obvious that the class B(R∞I ) coincides with Borel σ-algebra generated by
18
GILL, PANTSULAIA, AND ZACHARY
the τ -topology on R∞.) From our definition of B(R∞I ) we see that B(R∞) ⊂
B(R∞I ) and the containment is proper.
Theorem 2.5. λ∞(·) is a measure on B(Rn
Lebesgue measure on Rn.
I ), equivalent to n-dimensional
Ai ∈ B(Rn
Proof. If A = ∞×i=1
λ∞(A) =Q∞i=1 λ(Ai) always converges. Furthermore,
I ), then λ(Ai) = 1 for i > n so that the series
(2.1)
0 < λ∞(A) =Y∞
i=1
Since sets of the type A =
n
i=1
λ(Ai) = λn(
λ(Ai) =Yn
Ai generate B(Rn), we see that λ∞(·), re-
×i=1
×i=1
Ai).
n
stricted to Rn
I , is equivalent to λn(·).
(cid:3)
Corollary 2.6. The measure λ∞(·) is both translationally and rotationally
invariant on (Rn
I , B[Rn
I ]).
2.2. The Extension to R∞I . It is not obvious that λ∞(·) can be extended
to a countably additive measure on B(R∞I ).
Definition 2.7. Let
I ) ⊂ B(R∞I ) : n ∈ N, K is compact and 0 < λ∞(Kn) < ∞},
Kni, N ∈ N; Kni ∈ ∆0 and λ∞(Knl ∩ Knm) = 0, l 6= m}.
i=1
∆0 = {Kn = K×In ∈ B(Rn
∆ = {PN =[N
Definition 2.8. If PN ∈ ∆, we define
λ∞(PN ) =XN
λ∞(Kni).
i=1
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
19
Since PN ∈ B(Rn
next result follows:
I ) for some n, and λ∞(·) is a measure on B(Rn
I ), the
Lemma 2.9. If PN1, PN2 ∈ ∆ then:
(1) If PN1 ⊂ PN2, then λ∞(PN1) ≤ λ∞(PN2 ).
(2) If λ∞(PN1 ∩ PN2) = 0, then λ∞(PN1 ∪ PN2) = λ∞(PN2) + λ∞(PN2).
Definition 2.10. If G ⊂ R∞I
is any open set, we define:
λ∞(G) = lim
N→∞
sup{λ∞(PN ) : PN ∈ ∆, PN ⊂ G, } .
Theorem 2.11. If O is the class of open sets in B(R∞I ), we have:
(1) λ∞(R∞I ) = ∞.
(2) If G1, G2 ∈ O, G1 ⊂ G2, then λ∞(G1) ≤ λ∞(G2).
(3) If {Gk} ⊂ O, then
λ∞([∞
Gk) ≤X∞
λ∞(Gk).
k=1
k=1
(4) If the Gk are disjoint, then
λ∞([∞
k=1
Gk) =X∞
k=1
λ∞(Gk).
Proof. The proof of (1) is standard. To prove (2), observe that
{PN : PN ⊂ G1} ⊂(cid:8)P ′N : P ′N ⊂ G2(cid:9) ,
so that λ∞(G1) ≤ λ∞(G2). To prove (3), let PN ⊂ S∞k=1 Gk. Since PN
is compact, there is a finite number of the Gk which cover PN , so that
k=1 Gk. Now, for each Gk, there is a PNk ⊂ Gk. Furthermore, as
k=1 PNk . Since there is
I ), we may also assume that λ∞(PNl ∩ PNm) =
20
GILL, PANTSULAIA, AND ZACHARY
PN ⊂ SL
PN is arbitrary, we can assume that PN = P ′N =SL
an n such that all PNk ∈ B(Rn
0, l 6= m. We now have that
Xk=1
λ∞(PNk ) 6
λ∞(PN ) =
λ∞(Gk) 6
L
L
Xk=1
∞
Xk=1
λ∞(Gk).
It follows that
λ∞([∞
k=1
Gk) ≤X∞
k=1
λ∞(Gk).
If the Gk are disjoint, observe that if PN ⊂ P ′M ,
Xk=1
λ∞(P ′M ) ≥ λ∞(PN ) =
L
λ∞(PNk ).
It follows that
λ∞([∞
k=1
Gk) ≥XL
k=1
λ∞(Gk).
This is true for all L so that this, combined with (3), gives our result. (cid:3)
If F is an arbitrary compact set in B(R∞I ), we define
(2.2)
λ∞(F ) = inf {λ∞(G) : F ⊂ G, G open} .
Remark 2.12. At this point we see the power of B(R∞I ). Unlike B(R∞),
equation (2.2) is well-defined for B(R∞I ) because it has a sufficient number
of open sets of finite measure.
Theorem 2.13. Equation (2.2) is consistent with Definition 2.6 and the
results of Theorem 2.11.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
21
Definition 2.14. Let A be an arbitrary set in R∞I .
(1) The outer measure (on R∞I ) is defined by:
λ∗
∞(A) = inf {λ∞(G) : A ⊂ G, G open} .
We let L0 be the class of all A with λ∗∞(A) < ∞.
(2) If A ∈ L0, we define the inner measure of A by
λ∞,(∗)(A) = sup{λ∞(F ) : F ⊂ A, F compact} .
(3) We say that A is a bounded measurable set if λ∗∞(A) = λ∞,(∗)(A),
and define the measure of A, λ∞(A), by λ∞(A) = λ∗∞A).
Theorem 2.15. Let A, B and {Ak} be arbitrary sets in R∞I with finite
outer measure.
(1) λ∞,(∗)(A) ≤ λ∗∞(A).
(2) If A ⊂ B then λ∗∞(A) ≤ λ∗∞(B) and λ∞,(∗)(A) ≤ λ∞,(∗)(B).
(3) λ∗∞(S∞k=1 Ak) ≤P∞k=1 λ∗∞(Ak).
(4) If the {Ak} are disjoint, λ∞,(∗)(S∞k=1 Ak) ≥P∞k=1 λ∞,(∗)(Ak).
Proof. The proofs of (1) and (2) are straightforward. To prove (3), let ε > 0
be given. Then, for each k, there exists an open set Gk such that Ak ⊂ Gk
and λ∞(Gk) < λ∗∞(Ak) + ε2−k. Since (S∞k=1 Ak) ⊂ (S∞k=1 Gk), we have
λ∗
∞(cid:16)[∞
k=1
k=1
Ak(cid:17) 6 λ∞(cid:16)[∞
<X∞
Gk(cid:17) 6X∞
∞(Ak) + ε2−k] =X∞
[λ∗
k=1
λ∞(Gk)
k=1
λ∗
∞(Ak) + ε.
k=1
Since ε is arbitrary, we are done.
22
GILL, PANTSULAIA, AND ZACHARY
To prove (4), let F1, F2, . . . , FN be compact subsets of A1, A2, . . . , AN ,
respectively. Since the Ak are disjoint,
λ∞,(∗)(cid:16)[∞
k=1
Ak(cid:17) > λ∞(cid:18)[N
k=1
Fk(cid:19) =XN
k=1
λ∞(Fk).
Thus,
λ∞,(∗)(cid:16)[∞
k=1
Ak(cid:17) ≥XN
k=1
λ∞,(∗)(Ak).
Since N is arbitrary, we are done.
(cid:3)
The next two important theorems follow from the last one.
Theorem 2.16. (Regularity) If A has finite measure, then for every ε > 0
there exist a compact set F and an open set G such that F ⊂ A ⊂ G, with
λ∞(G \ F ) < ε.
Proof. Let ε > 0 be given. Since A has finite measure, it follows from our
definitions of λ∞,(∗) and λ∗∞ that there is a compact set F ⊂ A and an open
set G ⊃ A such that
λ∞(G) < λ∗
∞(A) + ∈2
and λ∞(F ) > λ∞,(∗)(A) − ∈2 .
Since λ∞(G) = λ∞(F ) + λ∞(G \ F ), we have:
λ∞(G \ F ) = λ∞(G) − λ∞(F ) < (λ∞(A) + ε
2 ) − (λ∞(A) − ε
2 ) = ε.
(cid:3)
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
23
Theorem 2.17. (Countable Additivity) If the family {Ak} consists of dis-
joint sets with bounded measure and A = S∞k=1 Ak, with λ∗∞(A) < ∞. then
λ∞(A) =P∞k=1 λ∞(Ak).
Proof. Since λ∗∞(A) < ∞, we have:
∞(Ak) =X∞
λ∞,(∗)(Ak) 6 λ∞,(∗)(A) 6 λ∗
∞(A) 6X∞
∞(A).
λ∗
λ∗
k=1
k=1
It follows that λ∞(A) = λ∗∞(A) = λ∞,(∗)(A), so that
Ak(cid:17) =X∞
λ∞(A) = λ∞(cid:16)[∞
k=1
k=1
λ∞(Ak).
(cid:3)
Definition 2.18. Let A be an arbitrary set in R∞I . We say that A is mea-
surable if A ∩ M ∈ L0 for all M ∈ L0.
by:
In this case, we define λ∞(A)
λ∞(A) = sup{λ∞(A ∩ M ) : M ⊂ L0} .
We let L∞I be the class of all measurable sets A.
Proofs of the following results are standard (see Jones [J], pages 48-52).
Theorem 2.19. Let A and {Ak} be arbitrary sets in L∞I .
(1) If λ∗∞(A) < ∞, then A ∈ L0 if and only if A ∈ L∞I . In this case,
λ∞(A) = λ∗∞(A).
(2) L∞I
is closed under countable unions, countable intersections, differ-
ences and complements.
24
GILL, PANTSULAIA, AND ZACHARY
(3)
k=1
λ∞([∞
(4) If {Ak} are disjoint,
λ∞([∞
k=1
Ak) ≤X∞
k=1
λ∞(Ak).
Ak) =X∞
k=1
λ∞(Ak).
(5) If Ak ⊂ Ak+1 for all k, then
λ∞([∞
k=1
Ak) = lim
k→∞
λ∞(Ak).
(6) If Ak+1 ⊂ Ak for all k and λ∞(A1) < ∞, then
λ∞(\∞
k=1
Ak) = lim
k→∞
λ∞(Ak).
We end this section with an important result that relates Borel sets to
L∞I -measurable sets (Lebesgue).
Theorem 2.20. Let A be a L∞I -measurable set. Then there exists a Borel
set F and a set N with λ∞(N ) = 0 such that A = F ∪ N .
Thus, we see that λ∞(·) is a regular countably additive σ-finite Borel
measure on R∞I = R∞ (as sets). More important is the fact that the de-
velopment is no more difficult than the corresponding theory for Lebesgue
measure on Rn.
Throughout the remainder of the paper we will also use B[R∞I ] for its
completion L∞I when convenient. This should cause no confusion since the
given context will always be clear.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
25
2.3. Separable Banach Spaces. In order to see what other advantages
our construction of (R∞I , B[R∞I ], λ∞(·)) offers, in this section we study sep-
arable Banach spaces. Let B be any separable Banach space.
Recall that (see Diestel [DI], page 32):
Definition 2.21. A sequence (un) is called a Schauder basis for B if
kunkB = 1 and, for each f ∈ B, there is a unique sequence (an) of scalars
such that
f = limn→∞Xn
k=1
akuk.
Definition 2.22. A sequence (vn) is called an absolutely convergent
Schauder basis for B if P∞n=1 kvnkB < ∞ and, for each f ∈ B, there is
a unique sequence (bn) of scalars such that
f = limn→∞Xn
k=1
bkvk.
Lemma 2.23. Let (un) be a Schauder basis for B, then there exists an
absolutely convergent Schauder basis for B.
Proof. Let (vn) = ( un
2n ). Then
∞
Xn=1
kvnkB =
kunkB
2n =
∞
Xn=1
∞
Xn=1
1
2n = 1 < ∞.
To see that (vn) is a Schauder basis for B, let f ∈ B. By definition, there is
a unique sequence (an) of scalars such that
f = limn→∞Xn
k=1
akuk.
26
GILL, PANTSULAIA, AND ZACHARY
If we take the sequence (bn) = (2nan), then
limn→∞Xn
k=1
bkvk = limn→∞Xn
k=1
akuk = f.
(cid:3)
It is known that most of the natural separable Banach spaces, and all that
have any use for applications in analysis, have a Schauder basis. In partic-
ular, it is easy to see from the definition of a Schauder basis that, for any
sequence (an) ∈ R∞I
representing a function f ∈ B, we have limn→∞ an = 0.
It follows that every separable Banach space (with a Schauder basis) is iso-
morphic to a subspace of R∞I .
Let BI be the set of all sequences (an) for which limn→∞Pn
in B. Define
k=1 akuk exists
Lemma 2.24. An operator
n (cid:13)(cid:13)(cid:13)Xn
k=1
k(an)kBI = sup
.
akuk(cid:13)(cid:13)(cid:13)B
T : (B, · B) → (BI , · BI ),
defined by T (f ) = (ak) for f = limn→∞Pn
from B onto BI .
k=1 akuk ∈ B, is an isomorphism
Let B be a separable Banach space with a Schauder basis and let BI =
T [B]. If B(BI ) = BI ∩ B[R∞I ], we define the σ−algebra generated on B, and
associated with B(BI ) by:
BI [B] =(cid:8)T −1(A) A ∈ B[BI ](cid:9) =: T −1 {B [BI ]} .
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
27
Note that, just as B[R∞] ⊂ B[R∞I ], we also have B[B] ⊂ BI [B] (with the
containment proper).
Theorem 2.25. Let A ∈ BI (B) and set λB(A) = λ∞[T (A)]. Let λB be the
completion of λB, then λB is a non-zero σ-finite Borel measure on B.
Proof. Let {vk} be an absolutely convergent Schauder basis. We first prove
that, for any L > 0 and any sequence (ak) ∈ [−L, L]ℵo, the function f =
P∞k=1 akvk ∈ B. We then prove that λB is nonzero.
Part 1
Let L be given. Since (vn) is an absolutely convergent Schauder basis,
given ε > 0 we can choose N such that P∞k=N vk < ǫ
N ≤ m ≤ n, we have
L . It follows that, for
n
6
Xk=m
Thus, the sequence {fn}, defined by fn =Pn
in B. Since B is a Banach space, the sequence converges.
kvkk < ε.
n
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=m
akvk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
k=1 akvk, is a Cauchy sequence
Part 2
To prove that λB is nonzero, it suffices to show that λB(cid:2)T −1 (I0)(cid:3) 6= 0,
where (I0) = [− 1
R∞I , so that B = T −1(I0) ∈ BI (B). Thus,
2 ]ℵo. First, we note that T is an injective linear map into
2 , 1
λB(B) = λ∞(cid:2)T (cid:0)T −1(I0)(cid:1)(cid:3) = λ∞(I0) = 1.
(cid:3)
28
GILL, PANTSULAIA, AND ZACHARY
2.4. Translations. In the theorem below, we will provide a new proof that
ℓ1 is the largest (dense) group of admissible translations for R∞I , so neces-
sarily ℓ1 is the largest group of admissible translations for every separable
Banach space B.
Recall that h(x) = ⊗∞k=1hk(xk), where hk(xk) = 1, for xk ∈ [− 1
2 , 1
2 ]. It
follows from dν = hdλ∞, that ν is absolutely continuous with respect to λ∞.
Thus, ν is equivalent to λ∞. Let Tλ∞ be the set of admissible translations
for R∞I
(i.e., λ∞[A − x] = λ∞[A] for all A ∈ B[R∞I ] and x ∈ Tλ∞).
Theorem 2.26. If A ∈ B[R∞I ] then λ∞[A − x] = λ∞[A] if and only if
Tλ∞ = ℓ1.
Proof. Suppose that x ∈ ℓ1. Since ν ∼ λ∞, we have that Tν = Tλ∞ (see
Yamasaki [YA1]). Thus, it suffices to prove that ν[A − x] = ν[A]. By
Kakutani's Theorem ([KA], see also [HHK] pg. 116), ν[A − x] ∼ ν[A] if and
only if
∞
Yk=1Z ∞
−∞phk (yk) hk (yk − xk)dλ(yk) > 0.
(2.3)
Now,
1
2 ,
1
2 ]∩[−
Z ∞
−∞phk (yk) hk (yk − xk)dλ(yk) =Z[−
dλ(yk) = (1−xk)+,
where r+ = max(0, r). Since x ∈ ℓ1, Q∞k=n (1 − xk)+ > 0 for n large
enough. Thus, equation (2.3) will be satisfied for every x ∈ ℓ1, so that
ℓ1 ⊂ Tν.
1
2 +xk,
1
2 +xk]
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
29
Now, suppose that x ∈ Tλ∞, so that λ∞[A − x] = λ∞[A] for all A ∈
B[R∞I ]. Thus, for A ∈ B[Rn
I ], we have
λ∞ [A − x] = λn [An − xn] ·
∞
Yk=n+1
λ(cid:8)(cid:2)− 1
2 , 1
λ(cid:8)(cid:2)− 1
2(cid:3) ∩(cid:2)− 1
2 , 1
2(cid:3) ∩(cid:2)− 1
2 − xk, 1
= λn [An] ·
∞
Yk=n+1
2 − xk, 1
2 − xk(cid:3)(cid:9)
2 − xk(cid:3)(cid:9) = λn [An] ·
∞
Yk=n+1
(1 − xk)+.
2 , 1
k=1[− 1
If An = In = ×n
that P∞k=1 xk < ∞, so that x ∈ ℓ1.
2 ], we have 1 = lim
n→∞
∞
Qk=n+1
(1 − xk)+. It follows
(cid:3)
In closing, we note that, since λ∞ is complete and regular, it is metrically
It follows from Theorem 0.5 that λ∞ is
transitivity with respect to R∞0 .
unique (this comment also applies to λB).
2.5. Gaussian measure. If we replace Lebesgue measure by the infinite
product Gaussian measure, µ∞, on R∞, we get countable additivity but
lose rotational invariance. Furthermore, the µ∞ measure of l2 is zero. On
the other hand, another approach is to use the standard projection method
onto finite dimensional subspaces to construct a probability measure directly
on l2.
In this case, we recover rotational invariance but not translation
invariance (and lose countable additivity). The resolution of this problem
led to the development of the Wiener measure [WSRM] and this is where
we are today. A nice discussion of this and related issues can be found in
Dunford and Schwartz [DS] (see pg. 402).
30
GILL, PANTSULAIA, AND ZACHARY
We now turn to take a look at infinite product Gaussian measure from
our new perspective. The canonical Gaussian measure on R is defined by:
dµ(x) =
1
√2π
exp(−x2
2 ) dλ(x).
Recall that µ∞ = ⊗∞k=1µ is countably additive on R∞, but its measure of
ℓ2 is zero. If we introduce a scaled version of Gaussian measure on R∞I , we
can resolve this difficulty. We seek a family of variances {σ2
k} such that
µB(B) = ∞⊗k=1
µk(T [B] = 1,
where µk is a linear Gaussian measure on R with parameters (0, σk) for
k ∈ N and µB is defined by:
µB(B) = ∞⊗k=1
µk(T [B]),
for any Borel subset B of B.
Lemma 2.27. Let (cid:8)σ2
k(cid:9) be a family of variances such that
∞
Xk=1
σ2
k < ∞,
then µB(cid:0)T −1([−L, L]ℵo )(cid:1) > 0 for every positive number L.
Proof. Let {Xk} be the family of independent Gaussian random variables
defined on some common probability space, (Ω, B, P [·]), with law µk. If
X = (X1, X2, . . . ), then
P hnω ∈ Ω X(ω) ∈ [−L, L]ℵooi = P h\∞
=Y∞
P [{ω ∈ Ω Xk(ω) 6 L]}] >Y∞
k=1
k=1 {ω ∈ Ω Xk(ω) ∈ [−L, L]}i
k=1(cid:18)1 −
L2(cid:19),
σ2
k
by Chebyshev's inequality.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
31
Clearly the product is positive. We are done since B = T −1([−L, L]ℵo) ∈
B(B) and
µB (B) = (⊗∞k=1µk)(cid:16)T [T −1([−L, L]ℵo )](cid:17) = P hnω ∈ Ω X(ω) ∈ [−L, L]ℵooi .
(cid:3)
Theorem 2.28. If the family of variances (cid:8)σ2
k(cid:9) satisfies the stronger con-
dition
(2.4)
∞
Xk=1
σ2
k
xk
< ∞
for some sequence (xk) ∈ ℓ1, then µB([B]) = 1.
Proof. By definition, if f ∈ B and (un) is a Schauder basis for B, then
there is a sequence of scalars (ak) such that f = limn→∞Pn
k=1 akuk. Since
T (f ) = (ak),
k(an)kBI = sup
n
n (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
6" ∞
Xk=1
ak# ,
akuk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)B
so that, if (an) ∈ ℓ1, then (an) ∈ T (B) = BI.
Suppose that there is a sequence (xk) ∈ ℓ1 such that such that the in-
equality (2.3) is satisfied. As in Lemma 2.24, by Chebyshev's inequality and
inequality (2.3) we have
µB(cid:26)T −1(cid:18) ∞×k=1h−xk1/2 , xk1/2i(cid:19)(cid:27) > 0.
32
GILL, PANTSULAIA, AND ZACHARY
If An = Rn × (×∞k=n+1[−xk1/2 ,xk1/2]), then An ⊂ An+1 and An ⊆ BI for
all natural n. Thus, we have
µB[T −1(An)]
µB[T −1(BI )] ≥ lim
n→∞
µk([−xk1/2 ,xk1/2]) ≥ lim
n→∞
= lim
n→∞
∞
Yk=n+1
∞
Yk=n+1(cid:18)1 −
σ2
k
xk(cid:19) = 1.
(cid:3)
Definition 2.29. We call µB a scaled version of Gaussian measure for B.
Theorem 2.30. The measure µB is a countably additive version of Gaussian
measure on B.
In particular, observe that we obtain a countably additive version of
Gaussian measure for both ℓ2 and C0[0, 1] (the continuous functions x(t)
on [0, 1] with x(0) = 0).
2.6. Rotational Invariance. In this section we study rotational invariance
on subspaces of (R∞I , BI [R∞] λ∞). First, we need a little more information
about Gaussian measures on vector spaces. (See Yamasaki [YA], pg. 151,
for a proof of the next Theorem).
Let F be a a real vector space, let F a be its algebraic dual space, and
let BF be the smallest σ-algebra such that L(x) is measurable for each
functional L ∈ F a and all x ∈ F.
Theorem 2.31. If µ is a measure on (F a, BF ), then the following are
equivalent.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
33
(1) The Fourier transform of µ, µ, is of the form:
µ(x) = exp(cid:8)− 1
2 hx, xi(cid:9) ,
for some inner product on F.
(2) For every x ∈ F, the distribution of L(x) is a one-dimensional
Gaussian measure.
In this general setting, a measure µ is said to be Gaussian on (F a, BF )
if it satisfies either of the above conditions.
Example 2.32. Let F = R∞0 , the set of sequences that are zero except for a
finite number of terms and let h·, ·i be the inner product on R∞0 . It is easy
to show that the corresponding measure on F a = R∞ (satisfying either (1)
or (2) above) is the infinite product Gaussian measure.
To understand the importance of this example, let (an) be any sequence
of positive numbers and let
(2.5)
Ha =( x ∈ R∞
a2
nx2
n < ∞) .
∞
Xn=1
The proof of the following is due to Yamasaki ([YA], pg. 153).
Lemma 2.33. If a ∈ ℓ2, µ[Ha] = 1, and if a /∈ ℓ2, µ[Ha] = 0.
Now, let us note that the standard one-dimensional Gaussian density,
which is normally written as fX(x) = [√2π]−1exp{− 1
2 x2}, may also be
written as fX(x) = exp{−π x2} with no factors of √2π if we scale x →
34
GILL, PANTSULAIA, AND ZACHARY
. With this convention, we can write the infinite dimensional version for
x√2π
L2[H, λH] as the derivative of the Gaussian distribution µH with respect to
the Lebesgue measure on H:
(2.6)
f (x) = exp{−π x2
H} =
dµH(x)
dλH(x)
.
This shows that, with the appropriate definition of Lebesgue measure, there
is a corresponding density for a Gaussian distribution on Hilbert space.
Remark 2.34. In the general case (see DePrato [DP]), when Q is a (positive
definite) trace-class operator and x is a Gaussian random variable with mean
m and covariance Q, we can write equation (2.6) as:
f (x) = [det Q]−1/2 exp(cid:8)−π(cid:10)Q−1(x − m), (x − m)(cid:11)H(cid:9)
dµH(x)
dλH(x)
.
Definition 2.35. A rotation on H is a bijective isometry U : H → H.
It is well-known that µH is invariant under rotations over (H, BH) (see
Yamasaki [YA], pg. 163).
Theorem 2.36. The measure, λH, is invariant under rotations and R =:
(T −1(ℓ2)) is dense in H and the maximal rotation invariance subspace for
λH.
Proof. Let any measurable set A ∈ BH. If U is any rotation on H, then
µH(U A) = µH(A) and U x2
H. It follows from equation (2.6) that
λH(U A) = λH(A).
H = x2
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
35
It follows from R∞0 ⊂ R ⊂ H, that R is dense, and from Lemma 2.33 that
R is maximal.
(cid:3)
Discussion. In this section, we have shown that what appears to be a
minor change in the way we represent R∞ makes it possible to define an
analogue of both Lebesgue and Gaussian measure (countably additive) on
every (classical) separable Banach space with a Schauder basis. Further-
more, our version of Gaussian measure is rotationally invariant, a property
not shared by Wiener measure. (What is more important, we have obtained
our core results using basic methods of Lebesgue measure theory from Rn.)
3. Operators
This section provides the background to understand the relationship be-
⊗ (which is nonseparable), and their restriction
⊗(h). We also obtain general conditions that allow us to define infinite
tween operators defined on H2
to H2
sums and products of linear operators on H2
⊗(h) for a given h.
3.1. Bounded Operators on H2
⊗. In this section we review the class of
bounded operators on H2
⊗ and their relationship to those on each Hi. Many
of the results are originally due to von Neumann [VN2]. However, the proofs
are new or simplified versions (some from the literature).
36
GILL, PANTSULAIA, AND ZACHARY
Let L[H2
⊗] be the set of bounded operators on H2
⊗. For each fixed i0 ∈ N
i
N
N
i ) =
Ai0gk
⊗i∈Ngk
and Ai0 ∈ L(Hi0), define Ai0 ∈ L(H2
⊗) by:
Xk=1
k=1 ⊗i∈Ngk
i0 ⊗ (⊗i6=i0gk
i )
Xk=1
Ai0(
for PN
in H2
⊗ and N finite but arbitrary. Extending to all of
H2
⊗ produces an isometric isomorphism of L[Hi0] into L[H2
⊗], which we
denote by L[H(i0)], so that the relationship L[Hi] ↔ L[H(i)] is an isometric
isomorphism of algebras. Let L#[H2
⊗] be the uniform closure of the algebra
It is clear that L#[H2
generated by {L[H(i)], i ∈ N}.
⊗]. von
Neumann has shown that the inclusion becomes equality if and only if N is
⊗] ⊂ L[H2
⊗] clearly consists of all
⊗ that are generated directly from the family {L[H(i)], i ∈
replaced by a finite set. On the other hand, L#[H2
operators on H2
N} by algebraic and topological processes.
g denote the projection from H2
⊗(g)s, and let Pw
⊗ onto H2
g denote
Let Ps
the projection from H2
⊗ onto H2
⊗(g)w.
Theorem 3.1. If T ∈ L#(H2
⊗), then Ps
gT = TPs
g and Pw
g T = TPw
g .
Proof. The weak case follows from the strong case, so we prove that Ps
gT =
TPs
i , with gk
i = gi for all
k=1 ⊗i∈Ngk
g. Since vectors of the form G = PL
⊗(g)s; it suffices to show that Tf ∈
but a finite number of i, are dense in H2
H2
⊗(g)s. Now, T ∈ L#(H2
⊗) implies that there exists a sequence of operators
Tn such that kT − Tnk⊗ → 0 as n → ∞, where each Tn is of the form:
Tn = PNn
k =
k a complex scalar, Nn < ∞, and each T n
k , with an
k=1 an
k T n
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
37
⊗i∈N\Mk Ii for some finite set of i-values Mk, where Ii is the identity
ki
⊗i∈Mk T n
operator on Hi. Hence,
Tnf =XL
l=1XNn
k=1
kigl
an
k ⊗i∈Mk T n
i ⊗i∈N\Mk gl
i.
i ⊗i∈N\Mk gl
⊗(g)s for each n, so that Tn ∈ L[H2
Now, it is easy to see that, for each l, ⊗i∈MkT n
kigl
It follows that Tnf ∈ H2
L[H2
gT = TPs
Ps
g.
⊗(g)s] is a norm closed algebra, T ∈ L[H2
i ≡s ⊗i∈Ngi.
⊗(g)s]. Since
⊗(g)s] and it follows that
(cid:3)
Let zi ∈ C,
zi = 1, and define U [z] by: U [z] ⊗i∈N gi = ⊗i∈Nzigi.
Theorem 3.2. The operator U [z] has a unique extension to a unitary op-
erator on H2
⊗, which we also denote by U [z], such that:
⊗(g)w, so that Pw
⊗(g)w → H2
is quasi-convergent but not convergent,
g U [z] = U [z]Pw
g .
then U [z]
:
(1) U [z] : H2
(2) If Qν zν
⊗(g)s → H2
H2
⊗(h)s, for some h ∈ H2
⊗(g)w with g⊥h.
(3) U [z] : H2
⊗(g)s → H2
⊗(g)s if and only if Qi zi converges and U [z] =
⊗. This implies
(Qi zi)I⊗, where I⊗ is the identity operator on H2
that Ps
gU [z] = U [z]Ps
g.
Proof. For (1), let h = PN
trary and 1 6 k 6 N . Then
k=1 ⊗i∈Nhk
i , where ⊗i∈Nhk
i ≡w ⊗i∈Ngi, N is arbi-
U∗[z]U [z]h =
N
Xk=1
⊗i∈Nz∗i zihk
i = h = U [z]U∗[z]h.
38
GILL, PANTSULAIA, AND ZACHARY
Thus, we see that U [z] is a unitary operator, and since h of the above
k=1 ⊗i∈Nhk
g U [z] = U [z]Pw
⊗(g)w if PN
k=1 ⊗i∈Nzihk
⊗(g)w → H2
⊗. By def-
⊗(g)w, so that
g . To prove (2), use The-
ν ∈ H2
i ∈ H2
⊗(g)w and Pw
form are dense, U [z] extends to a unitary operator on H2
inition, PN
U [z] : H2
orem 1.6 (3) and (4) to note that Qi zi = 0 and ⊗i∈Nhk
that ⊗i∈Nzihk
0 < Qi zi < ∞, then U [z] = [(Qi zi)I⊗], so that U [z] : H2
Now suppose that U [z] : H2
and so Qi zi must converge. Therefore, U [z]h = [(Qi zi)I⊗]h and Ps
i ≡s ⊗i∈Ngi imply
⊗(g)s. To prove (3), note that, if
⊗(g)s.
i ≡s ⊗i∈Ngi
gU [z] =
⊗(g)s → H2
⊗(g)s, then ⊗i∈Nzihk
i ∈ H2
⊗(f )s with H2
⊗(f )s⊥H2
⊗(g)s → H2
U [z]Ps
g.
It is easy to see that, for each fixed i ∈ N, A(i) ∈ L[H(i)] commutes with
any Ps
g, Pw
g or U [z], where g and z are arbitrary.
(cid:3)
Theorem 3.3. Every T ∈ L#[H2
where g and z are arbitrary.
⊗] commutes with all Ps
g, Pw
g and U [z],
Proof. Let L be the set of all Ps
g, Pw
g or U [z], with g and z arbitrary. From
the above observation, we see that all Ai ∈ L[H(i)], i ∈ N, commute with
L and hence belong to its commutator L′. Since L′ is a closed algebra, this
implies that L#[H2
⊗] ⊆ L′ so that all T ∈ L#[H2
⊗] commute with L.
(cid:3)
3.2. Unbounded Operators on H2
⊗. In this section, we consider a re-
stricted class of unbounded operators and the notion of a strong convergence
vector introduced by Reed [RE].
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
39
For each i ∈ N, let Ai be a closed densely defined linear operator on
Hi, with domain D(Ai), and let Ai be its extension to H2
⊗, with domain
D(Ai) ⊃ D(Ai) = D(Ai) ⊗ (⊗k6=iHk). The next theorem follows directly
from the definition of the tensor product of semigroups.
Theorem 3.4. Let Ai, 1 6 i 6 n, be generators of a family of C0-
6 Mieωit. Then Sn(t) = ⊗i=1,n Si(t),
semigroups Si(t) on Hi with kSi(t)kHi
defined on ⊗i=1,nHi, has a unique extension (also denoted by Sn(t)) to all
of H2
k=1 ⊗i∈Ngk
l ∈ D(Al), 1 6 l 6 n,
the infinitesimal generator for Sn(t) satisfies:
⊗, such that, for all vectors PK
i with gk
An" K
Xk=1
⊗i∈Ngk
i# =
n
K
Xl=1
Xk=1
Algk
l (⊗i6=l
i∈Ngk
i ).
Definition 3.5. Let {Ai}, i ∈ N, be a family of closed densely defined linear
operators on Hi and let gi ∈ D(Ai) (respectively, fi ∈ D(Ai)), with kgikH =
1 (respectively, kfikH = 1), for all i ∈ N.
(1) We say that g = ⊗i∈Ngi is a strong convergence sum (scs)-vector for
i∈Ngi) exists.
(2) We say that f = ⊗i∈Nfi is a strong convergence product (scp)-vector
k=1 Akg =P∞k=1 Akgk(⊗i6=k
the family {Ai} if s - lim
n→∞ Pn
n→∞ Qn
for the family {Ai} if s - lim
k=1 Akf = ⊗i∈NAifi exists.
Let Dg be the linear span of {χ = ⊗i∈Nχi, χi ∈ D(Ai)}, with χi = gi
(and let Df be the linear span of {η = ⊗i∈Nηi, ηi ∈ D(Ai)}, with ηi = fi)
⊗(g)s
for all i > L, where L is arbitrary but finite. Clearly, Dg is dense in H2
(Dη is dense in H2
If there is a possible chance for confusion, we
⊗(f )s).
40
GILL, PANTSULAIA, AND ZACHARY
let As, respectively Ap, denote the closure of P∞k=1 Ak on H2
tively Q∞k=1 Ak on H2
⊗(g)s (respec-
⊗(f )s)
are natural spaces for the study of infinite sums or products of unbounded
⊗(f )s). It follows that H2
⊗(g)s (respectively H2
operators. (The notion of a strong convergence sum vector first appeared in
Reed [RE].)
Definition 3.6. We call H2
space ) for the family {Ai}.
⊗(g)s an RS-space (respectively, H2
⊗(f )s an RP-
Let {Uk(t)} be a set of unitary groups on {Hk}. It is easy to see that
⊗. However, we know from
Theorem 3.2 (2), that it need not be reduced on any partial tensor product
U (t) = ⊗∞k=1Uk(t) is a unitary group on H2
subspace. The following results are due to Streit [ST] and Reed [RE], as
indicated.
Theorem 3.7. (Streit) Suppose {Ak} is a set of selfadjoint linear operators
⊗(g)s, with corresponding unitary groups {Uk(t)}. If U (t) =
on the space H2
⊗∞k=1Uk(t), then Ps
⊗(g)s) and
⊗(g)s if and only if, for
U (t) is a strongly continuous unitary group on H2
each c > 0, the following three conditions are satisfied:
g (i.e., U (t) is reduced on H2
gU (t) = U (t)Ps
(1) P∞k=1 hAkEk[−c, c]gk, gki < ∞,
(2) P∞k=1(cid:12)(cid:12)(cid:10)A2
(3) P∞k=1 h(Ik − Ek[−c, c]gk, gki < ∞,
kEk[−c, c]gk, gk(cid:11)(cid:12)(cid:12),
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
41
where Ek[−c, c] are the spectral projectors of Ak and, in this case, U (t) =
s − limn→∞
⊗n
k=1Uk(t).
Corollary 3.8. Conditions 1-3 are satisfied if and only if there exists a
strong convergence vector g = ⊗∞k=1gk for the family {Ak} such that gk ∈
D(Ak) and
k=1 kAkgkk2 < ∞.
X∞
k=1 hAkgk, gki < ∞, X∞
⊗(g)s and U (t) is a strongly
Theorem 3.9. (Reed) U (t) is reduced on H2
continuous unitary group on H2
⊗(g)s if and only if g = ⊗∞k=1g is a strong
convergence vector for the family {Ak} and P∞k=1 hAkgk, gki < ∞. If each
case, A, the closure of P∞k=1 Ak, is the generator of U (t).
Ak is positive, the statement is true without the above condition. In either
The next result strengthens and extends Reed's theorem to contraction
semigroups (i.e., the positivity requirement above can be dropped).
Theorem 3.10. Let {Sk(t)} be a family of strongly continuous contraction
semigroups with generators {Ak} defined on {Hk}, and let g = ⊗∞k=1gk be
a strong convergence vector for the family {Ak}. Then S(t) = ⊗∞k=1Sk(t)
is reduced on H2
⊗(g)s and is a strongly continuous contraction semigroup.
If S(t) = ⊗∞k=1Sk(t) is reduced on H2
⊗(g)s and is a strongly continuous
⊗(g)s, then there exists a strong convergence
contraction semigroup on H2
vector f = ⊗∞k=1fk ∈ H2
⊗(g)s for the family {Ak}.
42
GILL, PANTSULAIA, AND ZACHARY
Proof. Let g = ⊗∞k=1gk be a strong convergence vector for the family
{Ak}. Without loss, we can assume that kgkk = 1.
Let Sn(t) =
⊗n
k=1Sk(t) ⊗(⊗∞k=n+1Ik) and observe that Sn(t) is a contraction semigroup
⊗(g)s for all finite n . Furthermore, its generator is the closure of
on H2
An =Pn
k=1 Ak, where Ak = Ak ⊗(⊗∞i6=kIi). If n and m are arbitrary, then
[Sn(t) − Sm(t)] g =Z 1
= tZ 1
d
dλ {Sn[λt]Sm[(1 − λ)t]} gdλ
0
Sn[λt]Sm[(1 − λ)t] [An − Am] gdλ,
0
where we have used the fact that, if two semigroups commute, then their
corresponding generators also commute. It follows that:
k[Sn(t) − Sm(t)] gk 6 tk[An − Am] gk .
Since g = ⊗∞k=1g is a strong convergence vector for the family {Ak}, it fol-
lows that
s−limn→∞ Sn(t) = S(t) exists on a dense set in H2
⊗(g)s and the convergence
is uniform on bounded t intervals. It follows that S(t) extends to a bounded
⊗(g)s. To see that the closure of S(t) must be a con-
linear operator on H2
traction, for any ε > 0, choose n so large that k[Sn(t) − S(t)] gk⊗ < εkgk⊗.
It follows that
kS(t)gk⊗
6 kSn(t)gk⊗ + k[Sn(t) − S(t)] gk⊗ < kgk⊗ (1 + ε).
Thus, S(t) is a contraction operator on H2
is a C0-semigroup.
⊗(g)s. It is easy to check that it
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
43
Now suppose that S(t) = ⊗∞k=1Sk(t) is a strongly continuous contraction
semigroup which is reduced on H2
⊗(g)s. It follows that the generator A of
S(t) is m-dissipative, and hence defined on a dense domain D(A) in H2
⊗(g)s
with S′(t)f = S(t)Af = AS(t)f for all f ∈ D(A). Since any such f is of the
form f =P∞l=1 ⊗∞k=1f l
k is in D(A). A simple computation
shows that Af l =P∞k=1 Akf l, so that any f l is a strong convergence vector
for the family {Ak}.
k, each f l = ⊗∞k=1f l
(cid:3)
It is easy to see that, in the second part of the theorem, we cannot require
that g = ⊗∞k=1gk itself be a strong convergence vector for the family {Ak}
since it need not be in the domain of A. For example, g1 /∈ D(A1), while
gk ∈ D(Ak), k 6= 1.
4. Function Spaces
Let χIn be the indicator (or characteristic) function of In = ×∞k=n+1I. If
we let L(Rn) represent the class of measurable functions on Rn, then for each
measurable function fn ∈ L(Rn) we identify f ∈ L(Rn
I ) by f = fn ⊗ χIn.
Definition 4.1. A real-valued function f defined on the measure space
(R∞I , B[R∞I ], λ∞) is said to be measurable if f−1(A) ∈ B[R∞I ] for every
A ∈ B[R].
In this section we develop those aspects of function space theory that
will be of use later. We note that all the standard theorems for Lebesgue
measure apply. (The proofs are the same as for integration on Rn.)
44
GILL, PANTSULAIA, AND ZACHARY
4.1. L1-Theory. Let L1[Rn
I ] be the class of integrable functions on Rn
I .
Since L1(Rn
I ) ⊂ L1(Rn+1
I
), we define L1[R′∞I ] =S∞n=1 L1(Rn
I ) and let L1[R∞I ]
be the norm closure of L1[R′∞I ]. It follows that every function in L1[R∞I ] is
the limit of a sequence of functions in L1[Rnk
I ], for some sequence {nk} ⊂ N.
Let Cc(Rn
I ) be the class of continuous functions on Rn
I which vanish out-
side compact sets. We define Cc(R∞I ) to be the closure of S∞n=1 Cc(Rn
I ) =
Cc(R′∞I ) in the sup norm. Thus, for any f ∈ Cc(R∞I ), there always exists a
sequence of functions {fnk} ∈ Cc(Rnk
I ) such that fnk → f , for some sequence
{nk} ⊂ N. We define C0(R∞I ), the functions that vanish at ∞, in the same
manner.
Lemma 4.2. If f ∈ Cc(R∞I ) or C0(R∞I ), then f is continuous.
Proof. Let f (x) ∈ Cc(R∞I ) and let {xn n = 1, 2, . . . } be any sequence in Rn
such that xn → x as n → ∞. If ε > 0 is given, choose K1 so that for k ≥ K1
and fk ∈ Cc(R∞I ), fk(xn) − f (xn) < ε
3 . Then choose K2 so that for k ≥
K2, fk(x) − f (x) < ε
3 . Choose N so that for n ≥ N, fk(xn) − fk(x) < ε
3 .
If n ≥ N and k ≥ max{K1, K2}, we have:
I
f (xn) − f (x) ≤ fk(xn) − f (xn) + fk(x) − fk(xn) + fk(x) − f (x) < ε.
The same proof applies to C0(R∞I )
(cid:3)
Theorem 4.3. Cc(R∞I ) is dense in L1(R∞I ).
Proof. We prove this result in the standard manner, by reducing the proof to
positive simple functions and then to one characteristic function and finally
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
45
using the approximation theorem to approximate a measurable set which
contains a closed set and is contained in an open set.
= 0 for all f ∈ L1 (by the
DCT), we can prove the result for functions that vanish outside a compact
First note that, since limk→∞(cid:13)(cid:13)f χBI (0,k) − f(cid:13)(cid:13)1
set. In this case, as f = f+ − f−, we need only consider positive f . How-
ever, this function can be approximated by simple functions in S+. Since
each simple function is a finite sum of characteristic functions (of bounded
measurable sets) multiplied by finite constants, it follows that we need only
show that we can approximate the characteristic function of a bounded mea-
surable set by a continuous function which vanishes outside a compact set.
Let ε > 0 be given and let g = χA, where A is any bounded measurable set.
By the regularity of λ∞, there exists an open set O and a compact set H
with H ⊂ A ⊂ O and λ∞(O \ H) < ε.
Let {Vn} be the class of open intervals with rational end points. For each
n ∈ N, let Fn ⊂ g−1[Vn] and Gn ⊂ (O \ g−1[Vn]) be compact sets, such that
λ∞[(O \ Fn ∪ Gn)] < ε
2n . If H = ∩∞n=1[Fn ∪ Gn], then λ∞(O \ H) < ε.
If x ∈ H, there is an n such that f (x) ∈ Vn and x ∈ Gc
n ∩
H] ⊂ Vn. It follows that g restricted to H is continuous and λ∞(A \ H) ≤
λ∞(O \ H) < ε.
n, so that g[Gc
(cid:3)
In a similar fashion we can define the Lp spaces, 1 < p < ∞. We should
note that, each space is defined relative to the family of indicator functions
46
GILL, PANTSULAIA, AND ZACHARY
for I. Thus, each space is the canonical one for that particular class of
spaces.
5. Fourier Transform Theory
In this section, we study the implications of Lebesgue measure on R∞ for
the Fourier transform and discuss two different extensions of the Pontrjagin
Duality theory for Banach spaces.
Background. Let G be a locally compact abelian (LCA) group (c.f., Rn).
The following is a restatement of Theorem 0.1 (see Rudin [RU1]).
Theorem 5.1. If G is a LCA group and B(G) is the Borel σ-algebra of
subsets of G, then there is a non-negative regular translation invariant mea-
sure µ (i.e., µ(g + A) = µ(A), A ∈ B(G). The (Haar) measure µ is unique
up to multiplication by a constant.
Definition 5.2. A complex valued function α : G → C on a LCA group is
called a character on G provided that α is a homomorphism and α(g) = 1
for all g ∈ G.
The set of all continuous characters of G defines a new group G, called
the dual group of G and (α1 + α2)(g) = α1(g) · α2(g). If we define a map
G, by γg(α) = α(g), then the following theorem was proven by
γ : G →
Pontryagin:
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
47
Theorem 5.3. (Pontryagin Duality Theorem) If G is a LCA-group, then
the mapping γ : G →
G is an isomorphism of topological groups.
Thus, Pontrjagin Duality identifies those groups that are the character
groups of their character groups. If the group is not locally compact The-
orem 5.1 does not hold (e.g., there is no Haar measure). However Kaplan
[KA1] has shown that the class of topological abelian groups for which the
Pontrjagin Duality holds is closed under the operation of taking infinite
products of groups. This result immediately implies that this class is larger
than the class of locally compact abelian groups because the infinite product
of locally compact groups (for example, R∞) may be non-locally compact
(see also [KA2]).
5.1. Pontryagin Duality Theory I. In this section, we treat the Fourier
transform as an operator. As will be seen, this approach has the advantage
of being constructive. It also provides us with some insight into the problem
that arises when we look at analysis on infinite dimensional spaces.
We define F on L1[R, λ] by
g(x) = F(g)(x) =ZR
exp{−2πixy}g(y)dy.
It is easy to check that F−1 is defined by
g(y) = F−1(g)(y) =ZR
exp{2πiyx}g(x)dx.
48
GILL, PANTSULAIA, AND ZACHARY
This representation is more convenient for the infinite-dimensional case, be-
cause we have no factors of √2π to worry about.
It is possible to define F as a mapping on L1[Rn
I , λ] to C0[Rn
I , λ] for all n
as one fixed linear operator. However, in the case of Hilbert spaces,Theorem
3.2(2) requires that we clearly specify our canonical domain and range space.
The same is also true for L1[Rn
I , λ](h) and C0[Rn
I , λ](h) (see [GZ]). Since
h = ⊗∞k=1χI (xk), an easy calculation shows that h = ⊗∞k=1
I ](h) by
we can define F(fn)(x), mapping L1[Rn
I ](h) into C0[Rn
sin(πxk)
πxk
. Thus,
(5.1)
F(fn))(x) = ⊗n
k=1Fk(f (n)) ⊗∞k=n+1
hk(xk).
Theorem 5.4. The operator F extends to a bounded linear mapping of
L1[R∞I ](h) into C0[R∞I ](h).
Proof. Since
L1[Rn
I ](h) =
lim
n→∞
∞
[n=1
L1[Rn
I ](h) = L1[R′∞I ](h)
and L1[R∞I ](h) is the closure of L1[R′∞I ](h) in the L1 - norm, it follows that
F is a bounded linear mapping of L1[R′∞I ](h) into C0[R∞I ](h).
Supposed that {fn} ⊂ L1[R′∞I ](h), converges to f ∈ L1[R∞I ](h). Thus,
the sequence is Cauchy, so that kfn − fmk1 → 0 as m, n → ∞. It follows
that
F (fn(x) − fm(x)) 6ZR∞
I
fn(y) − fm(y) dλ∞(y) = kfn − fmk1 ,
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
49
so that F (fn(x) − fm(x))
L1[R′∞I ](h) is dense in L1[R∞I ](h), it follows that F has a bounded extension,
is a Cauchy sequence in C0[R∞I ](h).
Since
mapping L1[R∞I ](h) into C0[R∞I ](h).
(cid:3)
5.2. L2-Theory. In the case of L2, the Fourier transform is an isometric
isomorphism from L2[Rn] onto L2[Rn].
Theorem 5.5. The operator F is an isometric isomorphism of L2[R∞I ](h)
onto L2[R∞I ](h).
Proof. Let f ∈ L2[R∞I ](h). By construction, there exists a sequence of
functions {fk ∈ L2[Rnk
k→∞kf − fkk2 (h) = 0. Fur-
thermore, since the sequence converges, it is a Cauchy sequence. Hence,
I ], nk ∈ N} such that lim
given ε > 0, there exists a N (ε) such that m, n ≥ N (ε) implies that
kfm − fnk2 (h) < ε. Since F is an isometry, kF(fm) − F(fn)k2 (h) < ε,
so that the sequence F(fk) is also a Cauchy sequence in L2[R∞I ](h). Thus,
there is a f ∈ L2[R∞I ](h) with lim
F(f ) = f . It is easy to see that f is unique.
(h) = 0, and we can define
f − F(fk)(cid:13)(cid:13)(cid:13)2
k→∞(cid:13)(cid:13)(cid:13)
(cid:3)
We can also prove a version of Theorems 5.4 and 5.5 for every separable
Banach space (with a basis). Fix B and for each n, let Bn
is clear that Bn
, so that B is the norm closure of
following have the same proofs as Theorems 5.4 and 5.5.
I ⊂ Bn+1
I
I . It
I = B ∩ Rn
n→∞Bn
lim
I . The
Theorem 5.6. The operator F extends to a bounded linear mapping of
L1[B](h) into C0[B](h).
50
GILL, PANTSULAIA, AND ZACHARY
Theorem 5.7. The operator F is an isometric isomorphism from L2[B](h)
onto L2[B](h).
Theorems 5.4 - 5.7 show that ⊗∞i=1
hi is a strong (product) convergence
vector for the Fourier transform operator F. In the L2-theory, we know that
L2[R∞I ](h) and L2[R∞I ](h) are orthogonal subspaces of H2
⊗. Thus, in this
approach, the natural interpretation is that the Fourier transform induces a
Pontryagin duality like theory that does not depend on the group structure of
R∞I , but depends on the pairing of different function spaces. This approach
is direct, constructive and applies to all separable Banach spaces (with a
basis). Thus, the group structure of the underlying measure space plays no
role.
5.3. Pontryagin Duality Theory II. In this section, we show that the
standard form of Pontryagin duality theory is also possible, using the un-
derlying measure space group structure. It is constructive but restrictive, in
that, it does not apply to every separable Banach space with a basis.
Let B be any uniformly convex separable Banach space UCB over the
reals, so that B = B∗∗ (second dual). The next theorem follows from our
theory of Lebesgue measure on Banach spaces.
Theorem 5.8. If λB is our version of Lebesgue measure on B, then B and
B∗ are also duals as character groups (i.e., B∗ = B).
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
51
Proof. If we consider the restriction to L2[B, λB], we can define F directly
by:
(5.2)
[F(f )](x∗) = f (x∗) =ZB
exp{−2πihy, x∗i}f (y)dλB(y),
where hy, x∗i is the pairing between B and B∗. From Plancherel's Theorem,
we have:
=(cid:16) f , f(cid:17)2
= (f, f )2 = kfk2
2 .
It follows that B and B∗ are duals as character groups and
2
2
(cid:13)(cid:13)(cid:13)
f(cid:13)(cid:13)(cid:13)
f (y) =ZB∗
exp{2πihy, x∗i} f (x∗)dλB∗ (x∗).
(cid:3)
I , we can represent fn directly as a mapping from
I = BI ∩ Rn
If Bn
I , λB] → L2[B∗,n
L2[Bn
, λB∗], by
[F(fn)](x∗) = fn(x∗) =ZB
I
exp{−2πihy, x∗in}fn(y)dλB(y),
where hy, x∗in is the restricted pairing of y and x∗ to Bn
tively. It follows that equations 5.1 and 5.2 provide two distinct definitions
I and B∗,n
respec-
I
of the Fourier transform for B. Thus, in this approach the group structure
of the underlying measure space changes.
It is clear that representation for f (x∗) also applies if we use L1[B, λB],
but in this case f (x∗) ∈ C0[B∗].
If we define y(·) mapping B → C, by y(x) = exp{−2πihy, x∗i}, then
y(x) is a character of B. Furthermore, it is easy to see that (y1 + y2)(x) =
52
GILL, PANTSULAIA, AND ZACHARY
y1(x)·y2(x). We now have the extension of the Pontryagin Duality Theorem
to all UCB (with a basis).
Theorem 5.9. If B is a UCB, then the mapping γx : B →
γx(y) = y(x), is an isomorphism of topological groups.
B, defined by
In case B = H, is a Hilbert space, we can replace equation (5.2) by
(5.3)
f (x) = F[f ](x) =ZH
exp{−2πihx, yiH}f (y)dλH(y),
so that H is self-dual (as expected). From equation (5.3), we also get the
expected result that:
Fhexp{−π x2
H}i = exp{−π x2
H}.
In closing, we observe that by Theorem 2.31 (see Example 2.32), if we
use Gaussian measure on R∞, the dual character groups are R∞ and R∞ =
R∞0 . From this we see that probability measures on (R∞, B[R∞]) induce a
different character theory compared to that induced by λ∞ on (R∞I , B[R∞I ]).
5.4. Lp-Theory. We can obtain Lp[R∞I ] as in the construction of L1[R∞I ].
In this section we want to show the power of our approach to measure
theory by establishing a version of Young's Theorem for every separable
Banach space with a Schauder basis:
Theorem 5.10. (Young) Let p, q, r ∈ [1,∞] with
1
r
=
1
p
+
1
q − 1.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
53
If f ∈ Lp[R∞I ] and g ∈ Lq[R∞I ], then the convolution of f and g, f ∗ g, exists
(a.s.), belongs to Lr[R∞I ] and
kf ∗ gkr 6 kfkp kgkq .
Corollary 5.11. Let B be a separable Banach space with a Schauder basis
and let p, q, r ∈ [1,∞] with
1
r
=
1
p
+
1
q − 1.
If f ∈ Lp[B] and g ∈ Lq[B], then the convolution of f and g, f ∗ g, exists
(a.s.), belongs to Lr[B] and
kf ∗ gkr 6 kfkp kgkq .
In order to prove Theorem 5.10, we first need the appropriate version of
Fubini's Theorem. Since (R∞I , L∞I , λ∞) is a complete σ-finite measure space,
a proof of the following may be found in Royden [RO] (see Theorems 19 and
20, pgs. 269-270):
Theorem 5.12. (Fubini) If f ∈ L1[R∞I × R∞I ], then
(1) for almost all x ∈ R∞I
the function fx defined by fx(y) = f (x, y) ∈
L1[R∞I ](y):
(2) for almost all y ∈ R∞I
the function fy defined by fy(x) = f (x, y) ∈
L1[R∞I ](x):
f (x, y)dλ∞(y) ∈ L[R∞I ](x);
f (x, y)dλ∞(x) ∈ L[R∞I ](y);
I
(3) RR∞
(4) RR∞
I
54
GILL, PANTSULAIA, AND ZACHARY
(5)
f (x, y)d (λ∞ ⊗ λ∞) (x, y)
I ×R∞
I
ZR∞
=ZR∞
I "ZR∞
I
f (x, y)dλ∞(y)# dλ∞(x) =ZR∞
I "ZR∞
I
f (x, y)dλ∞(x)# dλ∞(y).
Theorem 5.13. Let f, g ∈ L1[R∞I ], then (f ∗ g)(x) exists (a.s.); that is
f (y)g(x − y) ∈ L1[R∞I ]. In addition, f ∗ g ∈ L1[R∞I ] and
kf ∗ gk1 6 kfk1 kgk1 .
Proof. First, it is easy to see that f (y)g(x − y) is a measurable function
on R∞I . (There is no change from the case of Rn.) We can apply Fubini's
theorem to get that:
I
ZR∞
=ZR∞
I
(f ∗ g) (x)dλ∞(x)
dλ∞(x)"ZR∞
=ZR∞
I
I
f (y)g(x − y)dλ∞(y)# =ZR∞
f (y)dλ∞(y) ·ZR∞
g(x)dλ∞(x).
I
I
dλ∞(y)"ZR∞
I
f (y)g(x − y)dλ∞(x)#
It follows from the last equality that kf ∗ gk1 6 kfk1 kgk1.
(cid:3)
5.4.1. Proof of Young's Theorem.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
55
Proof. First, assume that f and g are nonnegative and kfkp = kgkq = 1.
Let 1
p . Now note that
q and 1
q′ = 1 − 1
p′ = 1 − 1
1
1
q′
r
1
p′
+
1
r
+
(cid:16)1 −
(cid:16)1 −
1
p
=
+(cid:18)1 −
r(cid:17) q′ = p(cid:18) 1
p −
r(cid:17) p′ = q(cid:18) 1
q −
q(cid:19) +(cid:18)1 −
r(cid:19) q′ = p(cid:18)1 −
r(cid:19) p′ = q(cid:18)1 −
1
1
q
1
1
p(cid:19) = 1;
q(cid:19) q′ = p;
p(cid:19) p′ = q.
1
If we use Holder's inequality (for three functions), we can write (f ∗ g)(x)
as:
(f ∗ g) (x) =ZR∞
6"ZR∞
I hf (y)p/rg(x − y)q/rihf (y)1−p/rg(x − y)1−q/ri dλ∞(y)
"ZR∞
f (y)pg(x − y)qdλ∞(y)#1/r"ZR∞
dλ∞(y)#1/q′
f (y)(1−p/r)q′
I
I
I
g(x − y)(1−q/r)p′
dλ∞(y)#1/p′
.
This last inequality shows that
I
⇒
f (y)pg(x − y)qdλ∞(y)#1/r
(f ∗ g) (x) 6"ZR∞
f (y)pg(x − y)qdλ∞(y)# ⇒ (f ∗ g)r (x) 6 (f p ∗ gq) (x).
(f ∗ g)r (x) 6"ZR∞
From Theorem 5.13, we have k(f ∗ g)rk1 6 kf pk1 kgqk1 = 1. In the general
case, we know that f ∗ g exists (a.e.), so that f (y)g(x − y) ∈ L1[R∞I ].
But then, f (y)g(x − y) ∈ L1[R∞I ].
(cid:3)
I
In closing we note that, Beckner [BE] and Brascamp-Lieb [BL] have
shown that on Rn we can write Young's inequality as kf ∗ gkr 6
(Cp,q,r;n)n kfkp kgkq, where Cp,q,r;n ≤ 1 is sharp. We conjecture that 1 is
the sharp constant for R∞I .
56
GILL, PANTSULAIA, AND ZACHARY
6. Partial Differential Operators (Examples)
In this section, we give examples of strong product and sum vectors for
differential operators that have found interest in infinite dimensional analy-
sis.
Definition 6.1. For x ∈ R, 0 ≤ y < ∞ and 1 < a < ∞ define ¯g(x, y), ¯h(x)
by:
¯g(x, y) = exp(cid:8)−yaeiax(cid:9) ,
¯h(x) =
¯g(x, y)dy, x ∈ [− π
otherwise .
2a , π
2a ],
Z ∞
0
0
The following properties of ¯g are easy to check:
(1)
(2)
(3)
so that
∂¯g(x, y)
∂x
= −iayaeiax¯g(x, y),
∂¯g(x, y)
∂y
= −aya−1eiax¯g(x, y),
iy
∂¯g(x, y)
∂y
=
∂¯g(x, y)
∂x
.
It is also easy to see that ¯h(x) is in L1[R] for x ∈ [− π
2a , π
2a ] and,
(6.1)
d¯h(x)
dx
=Z ∞
0
∂¯g(x, y)
∂x
dy =Z ∞
0
iy
∂¯g(x, y)
∂y
dy.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
57
Integration by parts in the last expression of equation (6.1) shows that
¯h′(x) = −i¯h(x), so that ¯h(x) = ¯h(0)e−ix for x ∈ [− π
R ∞0 exp{−ya}dy, an additional integration by parts shows that ¯h(0) = Γ( 1
2a ]. Since ¯h(0) =
2a , π
a +
1).
Let a = π
2a , π
1−ε , ¯h(x) = ¯hε(x), x ∈ [− π
c expn ε2
2x2−ε2o ,
(6.2)
fε(x) =
0,
2a ], where 0 < ε ≪ 1, and define
x < ε/2,
x > ε/2,
where c is the standard normalizing constant. We now define ξ(x) = (¯h ∗
fε)(x), so that spt(ξ) = [− 1
2 ] = Iε. Thus, ξ(x) = 0, x /∈ Iε and otherwise,
2 , 1
ξ(x) =Z ∞
−∞
¯h[x − z]fε(z)dz = e−ixZ ∞
−∞
eizfε(z)dz = αεe−ix.
It follows from this that:
α−1
ε ξ(ix) =
ex, x ∈ Iε
0, x /∈ Iε
Define λε = λ and,
I ε = ×∞k=1Iε, I ε
n = ×∞k=n+1 and, λε
∞ = ⊗∞k=1λε.
Example 6.2. In this example, we let hk(xk) = α−1
ε ξ(ixk),
for each
k ∈ N so that Dkhk = hk , for xk ∈ I. Let L2
D∞ = Q∞k=1 Dk and fn ∈ L2
D∞fn(x) = Dnfn(x) = Qn
∞].
I ] ∩ D(D∞), we can define D∞ on Rn
l=1 Dlf(n)(x) ⊗(cid:0)⊗∞l=n+1hl(cid:1) , (a.s). This opera-
tor is well-defined and has a closed densely defined extension to L2
I ] = L2[Rn
I , λε
ε[Rn
ε[Rn
I by
If
ε[R∞I ](h),
where h = ⊗∞k=1hk. Thus, h is a strong product vector for D∞.
58
GILL, PANTSULAIA, AND ZACHARY
The operator D∞ is required if we want to obtain the probability den-
sity for a distribution function. (Note, by construction the density can be
approximated from below by densities in a finite number of variables.)
In the following example, we construct a general elliptic operator on
L2
ε[R∞I ].
Example 6.3. If ∇ = (D1, D2,···) and σk : R∞I → R is a bounded analytic
function for each k ∈ N, then let σ(x) = (σ1(x), σ2(x),···). We assume that
∞
∞
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
σj(x)σk(x) +
Xj,k=1
Xk=1
We can now define ∆∞ by:
bk(x)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)2
< ∞, where bk(x) =X∞
j=1
σj(x)Djσk(x).
∞
∞
Xk=1
σj(x)σk(x)Dj Dk +
For the same version of L2
∆∞ = (σ(x) · ∇)2 =
Xj,k=1
ε[R∞I ] as in the last example, if gn ∈ L2[Rn
I ] ∩
D(∆∞) and cn(x) = P∞j=n+1 σj(x) [Djσk(x)], then ∆∞ is defined on Rn
bk(x)Dk.
I
and
∆∞gn(x) =
n
Xj,k=1
σj(x)σk(x)Dj Dkgn(x) +
bkDkψ(x) + cn(x)gn(x).
n
Xk=1
In order to obtain the same equation with cn(x) = 0, we use the version
of L2
ε[R∞I ] defined with hk = ξI(0), k ≥ n + 1. In this case, for gn(x) ∈
ε[R∞I ] ∩ D(∆∞), we see that
L2
∆∞gn(x) =
n
Xj,k=1
σj(x)σk(x)Dj Dkgn(x) +
bk(x)Dkgn(x).
n
Xk=1
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
59
In either case, ∆∞ is well-defined for each n and has a closed densely defined
ε[R∞I ] and gn(x) → g(x) implies that limn→∞∆∞gn(x) =
extension to L2
∆∞g(x).
It follows that different versions of L2[R∞I ] offer advantages for particular
types of differential operators. (For other approaches, see [BK], [GZ] and
[UM].)
The following special cases have appeared in the literature (all can be
obtained from the last example):
(1) The natural infinite dimensional Laplacian:
A = ∆∞ =X∞
i=1
i .
∂2(cid:14)∂x2
(2) The nonterminating diffusion generator in infinitely many variables
(also known as the Ornstein-Uhlenbeck operator):
A = 1
2 ∆∞ − Bx · ∇∞ = 1
2X∞
i=1
∂2(cid:14)∂x2
i −X∞
i=1
bixi∂/∂xi.
The infinite dimensional Laplacian of Umemura [UM]:
A =
∞
Xi=1(cid:18) ∂2
i −
∂x2
xi
c2
∂
∂xi(cid:19).
Berezanskii and Kondratyev ([BK], pages 520-521) have also discussed op-
erators analogous to (2) and (3).
60
GILL, PANTSULAIA, AND ZACHARY
6.1. Discussion. In a very interesting paper, Phillip Duncan Thompson
[PDT] used the amplitudes of a set of orthogonal modes as the co-ordinates
in an infinite-dimensional phase space. This allowed him to derive the prob-
ability distribution for an ensemble of randomly forced two-dimensional vis-
cous flows as the solution of the continuity equation for the phase flow. He
obtained the following equation for the probability density:
∂ρ
∂t
+
∞
Xk=1
Mk(x)
∂ρ
∂xk − ν
∞
Xk=1
(6.3)
where
∂
∂xk (cid:2)ρα2
kxk(cid:3) −
∂2ρ
∂x2
k
∞
Xk=1
= 0,
Mk(x) =
∞
Xi=1
∞
Xj=1
α2
j βijk
αiαjαk
(µiµjµk)
1
2
µk
xixj.
The coefficients βijk vanishes if any two indices are equal, is invariant under
cyclic permutation of indices and reverses sign under non-cyclic permutation
of indices, while the coefficients αi and µi are positive constants, determined
by the problem. Thompson imposed the natural condition
(6.4)
∞
Z−∞
···
∞
Z−∞
ρ(x, t)
∞
Yk=1
dxk = 1.
At that time, he ran into the obvious mathematical criticism because equa-
tion (6.4) was meaningless at the time. He also derived the equilibrium
density
(6.5)
ρ0(x) = C exp(− 1
2 ν
kx2
α2
k) .
∞
Xk=1
The results in section 2.5, see also equation (2.5), along with those in section
4.4, show that Thomson's paper was prescient.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
61
7. Conclusion
In this paper we provided a reasonable version of Lebesgue measure on
R∞, which together with the standard Gaussian measure on R∞, have al-
lowed us to construct natural analogues of Lebesgue and Gaussian measure
for every separable Banach space with a Schauder basis. We have extended
the Fourier transform to L1[R∞, λ∞], L2[R∞, λ∞], defined sums and prod-
ucts of unbounded operators, and presented a few constructive examples of
partial differential operators in infinitely many variables.
Acknowledgments. This work could not have been written without the
generous help and critical remarks of Professor Frank Jones. We thank Pro-
fessor Anatoly Vershik for appraising us of recent work by the Russian school
and correcting some of our historical remarks. We would also like to thank
an anonymous referee for corrections that have improve our presentation
and for suggesting that we reconsider our approach to the Fourier transform
and discuss its relationship to the Pontryagin duality theory, which led to a
complete revision and extension of Section 5.
References
[BA] S. Banach Th´eorie des Op´erations lin´eaires, Monografj Matematyczn, Vol. 1, War-
saw, (1932).
[BA1] R. Baker "Lebesque measure" on R∞, Proc. Amer. Math. Soc. 113 (1991), 1023-
1029.
62
GILL, PANTSULAIA, AND ZACHARY
[BA2] R. Baker "Lebesque measure" on R∞, II, Proc. Amer. Math. Soc. 132 (2004),
2577-2591.
[BE] W. Beckner Inequalities in Fourier Analysis, Ann. of Math. 102 (1975), 159-182.
[BK] Yu. M. Berezanskii and Yu. G. Kondratyev, Spectral methods in infinite dimensional
analysis, Naukova Dumka, Kiev, (Russian) (1988).
[BL] H. J. Brascamp and E. H. Lieb Best constants in Young's inequality, its converse,
and its generalization to more than three functions, Ad. in Math. 20 (1976), 151-173.
[BM] Y. Bakhtin and J. C. Mattingly, Malliavin calculus for infinite-dimensional systems
with additive noise, arXiv:math.PR/0610754v1 (2006).
[DI] J. Diestel , Sequences and Series in Banach Spaces, Grad. Texts in Math. Springer-
Verlag, New York, (1984).
[DP] G. DaPrato, An Introduction to Infinite-Dimensional Analysis, Springer Berlin,
(2006).
[DS] N. Dunford and J. T. Schwartz, Linear Operators Part I: General Theory, Wiley
Classics edition, Wiley Interscience (1988).
[EM] E. O. Elliott and A. P. Morse, General product measures, Trans. Amer. Math. Soc.
110, (1964) 245-283.
[GU] A. Guichardet, Symmetric Hilbert Spaces and Related Topics, Lectures Notes in
Mathematics, No. 261, Springer-Verlag, New York, (1969).
[GZ] T. L. Gill and W. W. Zachary, Banach spaces of von Neumann type, Georgian
International Journal of Science Technology 3 (2011), 1-35.
[GZ1] T. L. Gill and W. W. Zachary, Feynman operator calculus: The constructive theory,
Expositiones Mathematicae 29 (2011), 165-203.
[HA] A. Haar, Der Massbegriff in der Theorie der kontinuierlichen Gruppe , Ann. Math.,
34 (1933), 147-169.
[HHK] H. H. Kuo , Gaussian Measures in Banach Spaces, Lecture Notes in Mathematics
463, Springer, New York (1975).
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
63
[HI] D. Hill, σ-finite invariant measures on infinite product spaces, Trans. Amer. Math.
Soc. 153, (1971) 347-370.
[J] F. Jones , Lebesgue Integration on Euclidean Space, Revised Edition, Jones and
Bartlett Publishers, Boston (2001).
[KA] S. Kakutani, On equivalence of infinite product measures , Ann. Math., 49 (1948),
214-224.
[KA1] S. Kaplan Extensions of Pontjagin duality I: Infinite products, Duke Math. J. 15
(1948), 649-659.
[KA2] S. Kaplan Extensions of Pontjagin duality II: Direct and inverse sequences, Duke
Math. J. 17 (1950), 419-435.
[KH] A.B. Kharazishvili On invariant measures in the Hilbert space, Bull. Acad. Sci.
Georgian SSR, 114 (1) (1984),4148 (in Russian).
[KK] K. Kodaira and S. Kakutani, A nonseparable translation-invariant extension of
Lebesgue measure space, Ann. Math., 52 (1950), 574-579.
[KO] A. N. Kolmogorov, Grundbegriffe der Wahrscheinlichkeitsrechnung, Springer-Verlag,
Vienna, (1933).
[KP1] A. P. Kirtadze and G. R. Pantsulaia, Invariant measures in the space RN , (in
Russian) Soobshch. Akad. Nauk Gruzii 141 (1991), 273-276.
[KP2] A. P. Kirtadze and G. R. Pantsulaia, Lebesgue nonmeasurable sets and the unique-
ness of invariant measures in infinite-dimensional vector spaces, Proc. A. Razmadze
Math. Inst. 143 (2007), 95-101.
[MO] C. C. Moore, Invariant measures on product spaces, Proc. 5th Berkeley Sym. Math.
Stat. & Prob. (Berkeley, 1965) 2 Part 2, 447-459.
[OX] J. C. Oxtoby, Invariant measures in groups which are not locally compact, Trans.
Amer. Math. Soc. 60, (1946) 215237.
[PA] G. Pantsulaia,
Invariant and Quasiinvariant Measures in Infinite-Dimensional
Topological Vector Spaces , Nova Science Publishers, New York, (2007).
64
GILL, PANTSULAIA, AND ZACHARY
[PDT] P. D. Thompson, Some exact statistics of two-dimensional viscous flow with random
forcing, Journal of Fluid Mechanics 55 (1972), 711-717.
[RE] M. C. Reed, On self-adjointness in infinite tensor product spaces, Journal of Func-
tional Analysis 5 (1970), 94-124.
[RH] G. E. Ritter and E. Hewitt, Elliott-Morse measures and Kakutanis dichotomy the-
orem, Math. Zeitschrift 211, (1992) 247263.
[RO] H. L. Royden, Real Analysis, (2nd Ed.) Macmillan Press, New York, (1968).
[RU] W. Rudin, Functional Analysis, McGraw-Hill Press, New York, (1973).
[RU1] W. Rudin, Fourier Analysis on Groups, John Wiley & Sons, New York, (1990).
[ST] L. Streit, Test function spaces for direct product representations, Commun. Math.
Phys. 4 (1967), 22-31.
[SU] V.N. Sudakov, Linear sets with quasi-invariant measure, Dokl.Akad.Nauk SSSR;
127 (1959), 524-525 (in Russian).
[T] P. D. Thompson Some exact statistics of two-dimensional viscous flow with random
forcing, J. Fluid Mech. 55 (1972), 711-717.
[UM] Y. Umemura, On the infinite dimensional Laplacian operator, J. Math. Kyoto Univ.
4 (1964/1965), 477-492.
[V] A. M. Vershik, Duality in the theory of measure in linear spaces, (English translation):
Sov. Math. Dokl. 7, (1967) 1210-1214.
[V1] A. M. Vershik, Does there exist the Lebesgue measure in the infinite-dimensional
space?, Proceedings of the Steklov Institute of Mathematics, 259 (2007), 248-272.
[V2] A. M. Vershik, The behavior of Laplace transform of the invariant measure on the
hyperspace of high dimension, J. Fixed Point Theory Appl., 3 (2008), 317-329.
[V3] D. L. Vandev, Invariant measures for the continual Cartan subgroup, J. Functional
Analysis, 255 (2008), 2661-2682.
CONSTRUCTIVE ANALYSIS IN INFINITELY MANY VARIABLES
65
[VA] A. M. Vershik, Action of groups on an infinite product measure, (in Russian) Pa-
pers presented at the Fifth Balkan Mathematical Congress (Belgrade, 1974). Math.
Balkanica 4 (1974), 643647.
[VN1] J. von Neumann, The uniqueness of Haar's measure, Rec. Math. (Mat. Sbornik)
N.S., 1 (1936), 721-734.
[VN2] J. von Neumann, On infinite direct products, Compositio Mathematica, 6 (1938),
1-77.
[WE] A. Weil, L'int´egration dans les groupes topologiques et ses applications, Actualit´es
Scientifiques et Industrielles, no. 869, Paris, 1940.
[WSRM] N. Wiener, A. Siegel, W. Rankin and W. T. Martin, Differential Space, Quantum
Systems, and Prediction, M. I. T. Press, Cambridge, MA, (1966).
[YA] Y. Yamasaki, Measures on infinite-dimensional spaces, World Scientific, Singapore,
(1985).
[YA1] Y. Yamasaki, Translationally invariant measure on the infinite-dimensional vector
space, Publ. Res. Inst. Math. Sci. 16 (3) (1980), 693720.
(Tepper L. Gill) Department of Mathematics, Physics and E&CE, Howard
University, Washington DC 20059, USA, E-mail : [email protected]
(Gogi R. Pantsulaia) Department of Mathematics, Georgian Technical Uni-
versity, Tbilisi 0175, Georgia;
I. Vekua Institute of Applied Mathematics,
Tbilisi State University, Tbilisi 0143, Georgia E-mail : g:pantsulaia@gtu:ge
(Woodford W. Zachary) Department of Mathematics and E&CE, Howard Uni-
versity, Washington DC 20059, USA, E-mail : [email protected]
|
1503.06750 | 3 | 1503 | 2015-04-06T09:44:40 | $C^{*}$ algebra and inverse chaos | [
"math.FA"
] | If an invertible linear dynamical systems is Li-York chaotic or other chaotic, what's about it's inverse dynamics? what's about it's adjoint dynamics? With this unresolved but basic problems, this paper will give a criterion for Lebesgue operator on separable Hilbert space. Also we give a criterion for the adjoint multiplier of Cowen-Douglas functions on $2$-th Hardy space. Last we give some chaos about scalars perturbation of operator and some examples of invertible bounded linear operator such that $T$ is chaotic but $T^{-1}$ is not. | math.FA | math |
C ∗ algebra and inverse chaos
Luo Lvlin∗ and Hou Bingzhe†
January 24, 2015
Abstract:
If an invertible linear dynamical systems is Li-York chaotic or
other chaotic, what's about it's inverse dynamics? what's about it's adjoint
dynamics? With this unresolved but basic problems, this paper will give a
criterion for Lebesgue operator on separable Hilbert space. Also we give a
criterion for the adjoint multiplier of Cowen-Douglas functions on 2-th Hardy
space. Last we give some chaos about scalars perturbation of operator and
some examples of invertible bounded linear operator such that T is chaotic
but T −1 is not.
Keywords:
inverse, chaos, Hardy space, rooter function, Cowen-Douglas
function, Spectrum, C∗ algebra, Lebesgue operator.
1.Introduction
∗E-mail:[email protected] and Adrress:Department of Mathematics,Jilin Univer-
sity, ChangChun,130012,P. R. China.
†E-mail:[email protected] and Adrress:Department of Mathematics,Jilin University,
ChangChun,130012,P. R. China.
1
The ideas of chaos in connection with a map was introduced by Li T.Y.and
his teacher Yorke,J.A.[22], after that there are various definitions of what it
means for a map to be chaotic and there is a series of papers on Topological
Dynamics and Ergodic Theory about chaos, such as [27][11][17][13][23].
Following Topological Dynamics,Linear Dynamics is also a rapidly evolv-
ing branch of functional analysis,which was probably born in 1982 with
the Toronto Ph.D.thesis of C.Kitai [3].
It has become rather popular be-
cause of the efforts of many mathematicians, for the seminal paper [6] by
G.Godefroy and J.H.Shapiro,the notes [12] by J.H.Shapiro,the authoritative
survey [20] by K.-G.Grosse-Erdmann,,and finally for the book [4] by F.Bayart
and E.Matheron,the book [19] by K.-G.Grosse-Erdmann and A.Peris.
For finite-dimension linear space,the authors have made a topologically
conjugate classification about Jordan blocks in [18][24]. So for the eigenvalues
λ 6= 1, the operators of Jordan block are not Li-Yorke chaotic.With [4] we
know that a Jordan block is not supercyclic when its eigenvalues λ = 1,
and following a easy discussion it is not Li-Yorke chaotic too.
So the definition of Li-Yorke chaos should be valid only on infinite-
dimension Frechet space or Banach space such that in this paper the Hilbert
space is infinite-dimensional. Because a finite-dimensional linear operator
could be regard as a compact operator on some Banach spaces or some Hilbert
spaces, we can get the same conclusion from [25]P12 or the Theorem 7 of
[10].
For a Frechet space X,let L(X) denote the set of all bounded linear
2
operators on X. Let B denote a Banach space and let H denote a Hilbert
space. If T ∈ L(B), then define σ(T ) = {λ ∈ C; T − λ is not invertible} and
define rσ(T ) = sup{λ; λ ∈ σ(T )}.
Definition 1. Let T ∈ L(B),if there exists x ∈ B satisfies:
(1) lim
n→∞
T n(x)k > 0; and (2) lim
n→∞
kT n(x)k = 0.
Then we say that T is Li-Yorke chaotic,and named x is a Li-Yorke chaotic
point of T ,where x ∈ B, n ∈ N.
Define a distributional function F n
x (τ ) = 1
n ♯{0 ≤ i ≤ n : kT n(x)k < τ },
where T ∈ L(B), x ∈ B, n ∈ N. And define
Fx(τ ) = lim inf
n→∞
F n
x (τ ); and F ∗x (τ ) = lim sup
n→∞
F n
x (τ ).
Definition 2. Let T ∈ L(B),if there exists x ∈ B and
(1) If Fx(τ ) = 0, ∃τ > 0, and F ∗x (ǫ) = 1, ∀ǫ > 0, then we say that T is
distributional chaotic or I-distributionally chaotic .
(2) If F ∗x (ǫ) > Fx(τ ), ∀τ > 0, and F ∗x (ǫ) = 1, ∀ǫ > 0, then we say that T
is II-distributionally chaotic.
(3) If F ∗x (ǫ) > Fx(τ ), ∀τ > 0,then we say that T is III-distributionally
chaotic.
Definition 3 ([25]). Let X is an arbitrary infinite-dimensional separable
Frechet space, T ∈ L(X),If there exists a subset X0 of X satisfies:
(1) For any x ∈ X0, {T nx}∞n=1 has a subsequence converging to 0;
3
(2) There is a bounded sequence {an}∞n=1 in span(X0) such that the se-
quence {T nan}∞n=1 is unbounded.
Then we say T satisfies the Li-Yorke Chaos Criterion.
Theorem 1 ([25]). Let X is an arbitrary infinite-dimensional separable
Frechet space, If T ∈ L(X),then the following assertions are equivalent.
(i) T is Li-Yorke chaotic;
(ii) T satisfies the Li-Yorke Chaos Criterion.
Lemma 1 ([29]). There are no hypercyclic operators on a finite-dimensional
space X 6= 0.
Example 1 ([4]P8). Let φ ∈ H∞(D) and let Mφ : H2(D) → H2(D) be the
associated multiplication operator. The adjoint multiplier M∗φ is hypercyclic
if and only if φ is non-constant and φ(D)T T 6= ∅.
Theorem 2 ([5]). Let X is an arbitrary separable Frechet space, T ∈ L(X).The
following assertions are equivalent.
(i) T is hypercyclic.
(ii) T is topologically transitive; that is,for each pair of non-empty open
sets U, V ⊆ X, there exists n ∈ N such that T n(U)T V 6= ∅.
Theorem 3 ([4]). Let X is a topological vector space,T is a bounded linear
4
operator on X.Let
Λ1(T ) , span( Sλ=1,n∈N
Λ+(T ) , span(Λ1(T )S Sλ>1,n∈N
Λ−(T ) , span(Λ1(T )S Sλ<1,n∈N
ker(T − λ)N T ran(T − λ)N );
ker(T − λ)N );
ker(T − λ)N ).
If Λ+(T ) and Λ−(T ) are both dense in X,then T is mixing.
2.From Polar Decomposition to functional calculus
The Polar Decomposition Theorem [15]P15 on Hilbert space is a useful
theorem, especially for invertible bounded linear operator. We give some
properties of C∗ algebra generated by normal operator.
Let H be a separable Hilbert space over C and let X be a compact subset
of C. Let C(X) denote the linear space of all continuous functions on the
compact space X, let P(x) denote the set of all polynomials on X and let T
be an invertible bounded linear operator on H. By the Polar Decomposition
Theorem [15]P15 we get T = UT , where U is an unitary operator and
T 2 = T ∗T . Let A(T ) denote the C∗ algebra generated by the positive
operator T and 1.
Lemma 2. Let 0 /∈ X be a compact subset of C. If P(x) is dense in C(X),
then P( 1
x ) is also dense in C(X).
Proof. By the property of polynomials we know that P( 1
x ) is a algebraic
closed subalgebra of C(X) and we get:
5
(1) 1 ∈ P( 1
x );
(2) P( 1
x ) separate the points of X;
(3) If p( 1
x) ∈ P(x),then ¯p( 1
x ) ∈ P(x).
By the Stone-Weierstrass Theorem [14]P145 we get the conclusion.
Lemma 3. Let X ⊆ R+. If P(x) is dense in C(X), then P(x2) is also
dense in C(X).
Proof. For X ⊆ R+, x 6= y ⇐⇒ x2 6= y2. By Lemma 2 we get the conclusion.
By the GNS construction [14]P250 for the C∗ algebra A(T ), we get the
following decomposition.
Lemma 4. Let T be an invertible bounded linear operator on the separable
Hilbert space H over C, A(T ) is the complex C∗ algebra generated by T
and 1. There is a sequence of nonzero A(T )-invariant subspace. H1, H2, · · ·
such that:
(1) H = H1 L H2 L · · · ;
(2) For every Hi, there is a A(T )-cyclic vector ξi such that Hi = A(T )ξi
and T Hi = Hi = T −1Hi.
Proof. By [30]P54 we get (1),and T Hi ⊆ Hi,that is Hi ⊆ T −1Hi; by
Lemma 2 we get T −1Hi ⊆ Hi. Hence we get T Hi = Hi = T −1Hi.
6
For ∀n ∈ N,T n is invertible when T is invertible. By the Polar Decom-
position Theorem [15]P15 T n = UnT n, where Un is unitary operator and
T n2 = T ∗nT n, we get the following conclusion.
Lemma 5. Let T be an invertible bounded linear operator on the separable
Hilbert space H over C, let A(T k) be the complex C∗ algebra generated by
T k and 1 and let HT k
invariant subspace, there is a decomposition H = Li
Given a proper permutation of HT k
and T −1HT k
k be a sequence of non-zero A(T k)-
k ∈ H, i, k ∈ N.
i = HT (k+1)
, we get T ∗HT k
and HT (k+1)
= A(T k)ξi
i = HT (k+1)
HT k
, ξi
.
j
i
i
i
i
i
Proof. By Lemma 3,it is enough to prove the conclusion on P(T k2)ξi
k. For
any given ξi
HT (k+1)
.
j
k ∈ H = Lj
HT (k+1)
j
, there is a unique j ∈ N such that ξi
k ∈
(1) Because T is invertible, for any given ξi
k, there is an unique ηi ∈
k. For ∀p ∈ P(x2), we get p(T (k+1)2)ηi =
HT (k+1)
s
such that ηi = T −1ξi
T ∗p(T k2)ξi
k. Hence we get
HT (k+1)
s
= P(T (k+1)2)ξs
k+1 ⊇ T ∗P(T k2)ξi
k = T ∗HT k
i
.
(2) Similarly,for any given ξs
k+1, there is an unique ηr ∈ HT k
r
such that
ηr = T ξs
k+1. For ∀p ∈ P(x2),we get
p(T k2)ηr = p(T k2)T ξs
k+1 = T ∗−1p(T (k+1)2)ξs
k+1.
Hence we get
HT k
r = P(T k2)ξr
k ⊇ T ∗−1P(T (k+1)2)ξs
k+1 = T ∗−1HT (k+1)
s
.
7
Let i = r,by (1)(2) we get T ∗−1HT (k+1)
s
⊆ HT k
i ⊆ T ∗−1HT (k+1)
j
.
Fixed the order of HT k
i = HT (k+1)
.
i
i
T ∗HT k
, by a proper permutation of HT (k+1)
j
we get
By Lemma 2 and T is invertible,we get
(3) For any given ξi
k, there is an unique ηi ∈ HT (k+1)
i
such that ηi = T ∗ξi
k.
For ∀p ∈ P(x−2), we get p(T (k+1)−2)ηi = T −1p(T k−2)ξi
k. Hence we get
HT (k+1)
i
= P(T (k+1)−2)ξi
k+1 ⊇ T −1P(T k−2)ξi
k = T −1HT k
i
.
(4) For any given ξi
k+1,there is an unique ηi ∈ HT k
i
such that ηi =
T ∗−1ξi
k+1. For ∀p ∈ P(x−2),we get
T −1p(T k−2)ηi = T −1p(T k−2)T ∗−1ξi
k+1 = p(T (k+1)−2)ξi
k+1.
Hence we get
T −1HT k
i = T −1P(T k−2)ξi
k ⊇ P(T (k+1)−2)ξi
k+1 = HT (k+1)
i
.
By (3)(4) we get T −1HT k
i ⊆ HT (k+1)
i
⊆ T −1HT k
i
.
that is,T −1HT k
i = HT (k+1)
i
.
Let ξ ∈ H is a A(T )-cyclic vector such that A(T )ξ is dense in H.
Because of σT
6= ∅, on C(σ(T )) define the non-zero linear functional
ρT:ρT(f ) =< f (T )ξ, ξ >, ∀f ∈ C(σ(T )). Then ρT is a positive lin-
ear functional, by [30]P54 and the Riesz-Markov Theorem, on C(σ(T )) we
8
get that there exists an unique finite positive Borel measure µT such that
f (z) dµT(z) =< f (T )ξ, ξ >, ∀f ∈ C(σ(T )).
Rσ(T)
Theorem 4. Let T be an invertible bounded linear operator on the separable
Hilbert space H over C, there is ξ ∈ H such that A(T )ξ = H. For any given
n ∈ N, let A(T n) be the complex C∗ algebra generated by T n and 1 and let
ξn be a A(T n)-cyclic vector such that A(T n)ξn = H. Then:
(1) For any given ξn,there is an unique positive linear functional
Rσ(T n)
f (z) dµT n(z) =< f (T n)ξn, ξn >, ∀f ∈ L2(σ(T n)).
(2) For any given ξn,there is an unique finite positive complete Borel
measure µT n such that L2(σ(T n)) is isomorphic to H.
Proof. Because T is invertible, by Lemma 5 we get that if there is a A(T )-
cyclic vector ξ, then there is a A(T n)-cyclic vector ξn.
(1):For any given ξn, define the linear functional, ρT n(f ) =< f (T n)ξn, ξn >,
by [30]P54 and the Riesz-Markov Theorem we get that on C(σ(T n)) there
is an unique finite positive Borel measure µT n such that
Rσ(T n)
f (z) dµT n(z) =< f (T n)ξn, ξn >, ∀f ∈ C(σ(T n)).
Moreover we can complete the Borel measure µT n on σ(T n), also us-
ing µT n to denote the complete Borel measure, By [26] we know that the
complete Borel measure is uniquely.
9
For ∀f ∈ L2(σ(T n)),because of
ρT n(f 2) = ρT n( ¯f f ) =< f (T n)∗f (T n)ξn, ξn >= kf (T n)ξnk ≥ 0.
we get that ρT n is a positive linear functional,hence (1) is right.
(2) we know that C(σ(T n)) is a subspace of L2(σ(T n)) such that C(σ(T n))
is dense in L2(σ(T n)).
For any f, g ∈ C(σ(T n)) we get
< f (T n)ξn, g(T n)ξn >H=< g(T n)∗f (T n)ξn, ξn >
= ρT n(¯gf ) = Rσ(T n)
f (z)¯g(z) dµT n(z) =< f, g >L2(σ(T n)) .
Therefor U0 : C(σ(T n)) → H, f (z) → f (T n)ξn is a surjection isometry
from C(σ(T n)) to A(T n)ξn, also C(σ(T n)) and A(T n)ξn is a dense sub-
space of L2(σ(T n)) and H,respectively. Because of U0 a closable operator,it
closed extension U : L2(σ(T n)) → H, f (z) → f (T n)ξn is a unitary opera-
tor. Hence for any given ξn, U is the unique unitary operator induced by the
unique finite positive complete Borel measure µT n such that L2(σ(T n)) is
isomorphic to H.
By the Polar Decomposition Theorem [15]P15, we get U∗T ∗T U = T T ∗
and U∗T −2U = T −2 when T = UT . In fact,when T is invertible, we can
choose an specially unitary operator such that T −1 and T −1 are unitary
equivalent. We give the following unitary equivalent by Theorem 4.
Theorem 5. Let T be an invertible bounded linear operator on the separable
Hilbert space H over C and let A(T ) be the complex algebra generated by
10
T and 1. There is σ(T −1) = σ(T −1) and we get that T −1 and T −1 are
unitary equivalent by the unitary operator F H
xx∗, more over the unitary opera-
xx∗ is induced by an almost everywhere non-zero function pφT,where
tor F H
pφT ∈ L∞(σ(T ), µT). That is,d µT −1 = φTd µT−1.
Proof. By Lemma 4,lose no generally,let ξT is a A(T )-cyclic vector such
that H = A(T )ξT.
(1) Define the function Fx−1 : P(x) → P(x−1), Fx−1(f (x)) = f (x−1), it is
easy to find that Fx−1 is linear.Because of
f (z−1)dµT(z) =< f (T −1)ξ, ξ > = Rσ(T−1)
Rσ(T)
We get dµT−1(z) = z2dµT(z).Hence
f (z)dµT−1(z).
Fx−1(f (x)) ¯Fx−1(f (x))dµT−1(x)
f (x−1) ¯f (x−1)dµT−1(x)
kFx−1(f (x))kL2(σ(T−1),µT −1 )
= Rσ(T−1)
= Rσ(T−1)
= Rσ(T)
≤ sup σ(T )2 Rσ(T)
≤ sup σ(T )2kf (z)kL2(σ(T),µT ).
z2f (z) ¯f (z)dµT(z)
f (z) ¯f (z)dµT(x)
So we get kFx−1k ≤ sup σ(T )2, by the Banach Inverse Mapping Theo-
rem [14]P91 we get that Fx−1 is an invertible bounded linear operator from
11
L2(σ(T ), µT) to L2(σ(T −1), µT−1).
Define the operator F H
x−1 : A(T )ξ → A(T −1)ξ, F (f (T )ξ) = f (T −1)ξ.
By Lemma 2 and [30]P55, we get that F H
x−1 is a bounded linear operator
on the Hilbert space A(T )ξ = H, kF H
x−1k ≤ sup σ(T )2.Moreover we get
H
T
−−−−−−−−→
H
F H
x−1 ↓
H
−−−−−−−−−−→
T −1
↓ F H
x−1
H
(2) Define the function Fxx∗ : P(xx∗) → P(x∗x), Fxx∗(f (x, x∗)) = f (x∗x).
By [8] we get that Fxx∗ is a linear algebraic isomorphic.
For any x ∈ σ(T −1), by Lemma 3 we get that P(x2) is dense in
C(x), C(x) is dense in L2(σ(T −1), µT −1). Hence P(x2) is dense in
L2(σ(T −1), µT −1).
With a similarly discussion, for any y ∈ σ(T −1),we get that P(y2) is
dense in L2(σ(T −1), µT−1).
So Fxx∗ is an invertible bounded linear operator from L2(σ(T −1), µT−1)
to L2(σ(T −1), µT −1). Therefor Fxx∗ ◦ Fx−1 is an invertible bounded linear
operator from L2(σ(T ), µT) to L2(σ(T −1), µT −1).
Because of λ ∈ σ(T ∗T ) ⇐⇒ 1
λ ∈ σ(T ∗−1T −1),
we get that λ ∈ σ(T ) ⇐⇒ 1
λ ∈ σ(T −1), that is,σ(T −1) = σ(T −1).
For any pn ∈ P(σ(T −1)) ⊆ A(σ(T −1)), because of T ∗−1pn(T −1) =
12
pn(T −1)T ∗−1, by [15]P60 we get that P(T −1) and P(T −1) are unitary
equivalent. Hence there is an unitary operator U ∈ B(H) such that UP(T −1) =
P(T −1)U, that is,UA(T −1) = A(T −1)U and UA(T −1)ξT = A(T −1)UξT.
If let ξT −1 = UξT,then ξT −1 is a A(T −1)-cyclic vector and A(T −1)ξT −1 =
H.Because of
f ( 1
z )dµT(z).
f (z)dµT−1(z) =< f (T −1)ξT, ξT > = Rσ(T)
f (z)dµT −1(z) =< f (T −1)ξT −1, ξT −1 >.
Rσ(T−1)
Rσ(T −1)
We get [dµT −1] = [dµT−1],that is, dµT −1 and dµT−1 are mutually ab-
solutely continuous, by [14]IX.3.6Theorem and (1) we get that dµT −1 =
φT( 1
z )dµT−1 = z2φT(z)dµT, where φT(z) 6= 0, a.e. and φT(z) ∈
L∞(σ(T ), µT). So we get
Fxx∗ ◦ Fx−1(f (x))Fxx∗ ◦ Fx−1(f (x))dµT −1(x)
f (x−1) ¯f (x−1)dµT −1(x)
kFxx∗ ◦ Fx−1(f (x))kL2(σ(T −1),µT −1)
= Rσ(T −1)
= Rσ(T −1)
= Rσ(T−1)
= Rσ(T−1)
= kqφT( 1
φT( 1
x )f (x−1) ¯f (x−1)dµT−1(x)
Fx−1(pφT(x)f (x))Fx−1(pφT(x) ¯f (x))dµT−1(x)
x)Fx−1(f (x))kL2(σ(T−1),µT −1 )
Hence Fxx∗ is an unitary operator that is induced by Uf ( 1
x) = qφT( 1
x )f ( 1
x).
13
Define F H
xx∗:
A(T −2)ξT → A(T −2)ξT −1,
xx∗(f (T −2)ξT) = f (T −2)ξT −1.
F H
Therefor F H
By Lemma 4 and [8] we get A(T −1)ξ = H = A(T −1)ξ. That is,F H
xx∗ is an unitary operator from A(T −1)ξT to A(T −1)ξT −1.
xx∗ is an
unitary operator and we get
H
T −1
−−−−−−−−−−→
H
F H
xx∗ ↓
H
−−−−−−−−−−→
T −1
↓ F H
xx∗
H
So T −1 and T −1 are unitary equivalent by F H
xx∗ is induced by the function qφT( 1
F H
Corollary 1. Let T be an invertible bounded linear operator on the separable
xx∗, the unitary operator
x ).
Hilbert space H over C and let A(T ) be the complex algebra generated by
T and 1. There is σ(T ) = σ(T ∗) and we get that T and T ∗ are unitary
equivalent by the unitary operator F H
xx∗ is
induced by an almost everywhere non-zero function pφT,where pφT ∈
L∞(σ(T ), µT). That is,d µT ∗ = φTd µT.
xx∗, more over the unitary operator F H
3.The chaos between T and T ∗−1 for Lebesgue operator
For the example of singular integral in mathematical analysis, we know
that is independent the convergence or the divergence of the weighted integral
between x and x−1, however some times that indeed dependent for a special
weighted function. For T is an invertible bounded operator on the separable
Hilbert space H over C, we get 0 /∈ σ(T n) ⊆ R+.
In the view of the
14
singular integral in mathematical analysis and by Theorem 4, we get that
T and T ∗−1 should not be convergence or divergence at the same time for
T is an invertible bounded operator. Anyway,they should be convergence or
divergence at the same time for some special operators.
Therefor we define the Lebesgue operator and prove that T and T ∗−1 are
Li-Yorke chaotic at the same time for T is a Lebesgue operator. Then we
give an example that T is a Lebesgue operator,but not is a normal operator.
Let dx be the Lebesgue measure on L2(R+), by Theorem 4 we get that
dµT n is the complete Borel measure and L2(σ(T n)) is a Hilbert space. If
∃N > 0, for ∀n ≥ N, n ∈ N,dµT n is absolutely continuity with respect to dx,
by the Radon-Nikodym Theorem [14]P380 there is fn ∈ L1(R+) such that
dµT n = fn(x) dx.
Definition 4. Let T be an invertible bounded linear operator on the separable
Hilbert space H over C, moreover if T satisfies the following assertions:
(1) If ∃N > 0,for ∀n ≥ N, n ∈ N
dµT n = fn(x) dx,
x2fn(x) = fn(x−1),
fn ∈ L1(R+).
0 < x ≤ 1.
(2) If ∃N > 0, for ∀n ≥ N, n ∈ N,there is a A(T n)-cyclic vector ξn. And
for any given 0 6= x ∈ H and for any given 0 6= gn(t) ∈ L2(σ(T n)), there is
an unique 0 6= y ∈ H such that y = gn(T n−1)ξn when x = gn(T n)ξn.
Then we say that T is a Lebesgue operator, let LLeb(H) denote the set of
all Lebesgue operators.
15
Theorem 6. Let T be a Lebesgue operator on the separable Hilbert space H
over C, then T is Li-Yorke chaotic if and only if T ∗−1 is.
Proof. We prove the conclusion by two parts.
(1) Let H be A(T )-cyclic, that is,there is a vector ξ such that A(T )ξ =
H. By Lemma 5 we get that if there is a A(T )-cyclic vector ξ, then there
is also a A(T n)-cyclic vector ξn.
Let x0 be a Li-Yorke chaotic point of T , by Theorem 4 and the Polar
Decomposition Theorem [15]P15 and by the define of Lebesgue operator,we
get that for enough large n ∈ N,there are gn(x) ∈ L2(σ(T n)), fn(x) ∈
L2(R+) and y0 ∈ H such that x0 = gn(T n)ξn,y0 = gn(T n−1)ξn and dµT n =
fn(x) dx.
kT nx0k
=< T n∗T nx0, x0 >
=< T n2gn(T n)ξn, gn(T n)ξn >
=< gn(T n)∗T n2gn(T n)ξn, ξn >
x2gn(x)¯g(x) dµT n(x)
x2gn(x)2fn(x) dx
0 x2gn(x)2fn(x) dx + R +∞
0 x2gn(x)2fn(x) dx + R 1
0 gn(x)2fn(x−1) dx + R 1
1
x−2gn(x−1)2fn(x) dx + R 1
x−2gn(x−1)2fn(x) dx
0
0
= Rσ(T n)
= R +∞
= R 1
= R 1
, R 1
= R +∞
= R +∞
1
x2gn(x)2fn(x) dx
0 x−4gn(x−1)2fn(x−1) dx
0 x−2gn(x−1)2fn(x) dx
0 x−2gn(x−1)2fn(x) dx
16
= Rσ(T n)
x−2gn(x−1)¯gn(x−1) dµT n(x)
=< gn(T n)∗T n−2gn(T n−1)ξn, ξn >
=< T n−2gn(T n−1)ξn, gn(T n−1)ξn >
=< T n−2y0, y0 >
=< T −nT −n∗y0, y0 >
= kT ∗−ny0k.
Where , following the define of fn(x).
(2) If H is not A(T )-cyclic, by Lemma 5 we get that for ∀n ∈ N, there
k ∈ H, i, k ∈ N, where HT k
is a decomposition H = Li
is a sequence of A(T k)-invariant subspace, and do (1) for HT k
HT k
, ξi
i
i = A(T k)ξi
k
.
i
By (1)(2) we get that T is Li-Yorke chaotic if and only if T ∗−1 is Li-Yorke
chaotic.
Corollary 2. Let T be a Lebesgue operator on the separable Hilbert space H
over C, then T is I-distributionally chaotic (or II-distributionally chaotic or
III-distributionally chaotic) if and only if T ∗−1 is I-distributionally chaotic
(or II-distributionally chaotic or III-distributionally chaotic).
Theorem 7. There is an invertible bounded linear operator T on the sep-
arable Hilbert space H over C, T is Lebesgue operator but not is a normal
operator.
Proof. Let 0 < a < b < +∞, then L2([a, b]) is a separable Hilbert space over
R, because any separable Hilbert space over R can be expanded to a separable
Hilbert space over C, it is enough to prove the conclusion on L2([a, b]). We
17
prove the conclusion by six parts:
(1) Let 0 < a < 1 < b < +∞,M = {[a, b−a
2 ], [ b−a
2 , b]}. Construct measure
preserving transformation on [a, b].
There is a Borel algebra ξ(M) generated by M, define Φ : [a, b] → [a, b],
Φ([a, b−a
2 ]. Then Φ is an invertible measure
2 ]) = [ b−a
2 , b], Φ([ b−a
2 , b]) = [a, b−a
preserving transformation on the Borel algebra ξ(M).
By [27]P63,UΦ 6= 1 is a unitary operator induced by Φ, where UΦ is the
composition UΦh = h ◦ Φ, ∀h ∈ L2([a, b]) on L2([a, b]).
(2) Define Mxh = xh on L2([a, b]),then Mx is an invertible positive oper-
ator.
(3) For f (x) = ln x
x , x > 0, define dµ = f (x)d x. then f (x) is continuous
and f (x) > 0, a.e.x ∈ [a, b], hence d µ that is absolutely continuous with
respect to d x is finite positive complete Borel measure, therefor L2([a, b], d µ)
a separable Hilbert space over R. Moreover L2([a, b]) and L2([a, b], d µ) are
unitary equivalent.
(4) Let T = UΦMx,we get T ∗T = UΦT T ∗U∗Φ and UΦ 6= 1. Because
of UΦMx 6= MxUΦ and UΦMx2 6= UΦMx2, we get that T is not a normal
operator and σ(T ) = [a, b].
(5) The operator T = UΦMx on L2([a, b]) is corresponding to the operator
T ′ on L2([a, b], d µ),T ′ is invertible bounded linear operator and is not a
normal operator and σ(T ′) = [a, b].
18
(6) From R b
1
n−1
n
we get that fn(t) is continuous and almost everywhere positive, hence fn(t)d t
a xnf (x)d x = R bn
d t, let fn(t) = 1
n I[an,bn]f (t
an tf (t
1
n−1
n
nt
t
,
1
n )
1
n )
is a finite positive complete Borel measure.
For any E ⊆ R+ define IE = 1 when x ∈ E else IE = 0,so IE is the
identity function on E.
(i) fn(t−1) = 1
n I[an,bn]f (t−
1
n )
1
n−1
n
−
t
= 1
nI[an,bn]t ln t
1
n
= 1
n I[an,bn] ln t
−
t
1
n
1
n−1
n
−
t
−
1
n
(i) t2fn(t) = 1
n I[an,bn]f (t
1
n ) t2
n−1
n
t
= 1
nI[an,bn]t ln t
1
n
= 1
n I[an,bn] ln t
1
n
t
1
n
t2
n−1
n
t
By (i)(ii) we get x2fn(x) = fn(x−1).
From σ(T ′n) = [an, bn] and R bn
t2fn(t)dt, let
d µT ′n = fn(t)d t,then d µT ′n is the finite positive complete Borel measure.
d t = R +∞
an t2f (t
1
n−1
n
nt
0
1
n )
For any given 0 6= h(x) ∈ L2([a, b]) we get 0 6= h(x−1) ∈ L2([a, b]). I[a,b]
is a A(M n
x )-cyclic vector of the multiplication M n
x = Mxn, and I[an,bn] is
a A(T ′n)-cyclic vector of T ′n, By Definition 4 we get that T ′ is Lebesgus
operator but not is a normal operator.
Corollary 3. There is an invertible bounded linear operator T on the sepa-
rable Hilbert space H over C, T is a Lebesgue operator and also is a positive
operator.
19
By the cyclic representation of C∗ algebra and the GNS construction, also
by the functional calculus of invertible bounded linear operator, we could
study the operator by the integral on R. This way neither change Li-Yorke
chaotic nor the computing, but by the singular integral in mathematical
analysis on the theoretical level we should find that there is a invertible
bounded linear operator T that is Li-Yorke chaotic but T −1 is not Li-Yorke
chaotic.
4. Cowen-Douglas function on Hardy space
r<1 R π
For D = {z ∈ C, z < 1}, if g is an complex analytic function on D and
−π g(reiθ)2 dθ < +∞, then we denote g ∈ H2(D), hence H2(D)
2π , especially
H2(D) is denoted as a Hardy space. By the completion theory of complex
is a Hilbert space with the norm kgk2
−π g(reiθ)2 dθ
there is sup
H2 = sup
r<1 R π
analytic functions, the Hardy space H2(D) is a special Hilbert space that is
relatively easy not only for theoretical but also for computing, and we should
give some properties about adjoint multiplier operators on H2(D).
Any given complex analytic function g has a Taylor expansion g(z) =
anzn, so g ∈ H2(D) and
+∞
a2
Pn=0
n < +∞ are naturally isomorphic. For
T = ∂D, if L2(T) denoted the closed span of all Taylor expansions of functions
+∞
Pn=0
in L2(T), then H2(T) is a closed subspace of L2(T). From the naturally
isomorphic of H2(D) and H2(T) by the properties of analytic function, we
denote H2(T) also as a Hardy space.
By [14]P6 we get that any Cauchy sequence with the norm k • kH2 on
20
H2(D) is a uniformly Cauchy sequence on any closed disk in D, in particu-
lar,we get that the point evaluations f → f (z) are continuous linear func-
tional on H2(D), by the Riesz Representation Theorem[14]P13,for any g(s) ∈
H2(D), there is a unique fz(s) ∈ H2(D) such that g(z) =< g(s), fz(s) >, so
define that fz is a reproducing kernel at z.
Let H∞(D) denote the set of all bounded complex analytic function on D,
for any given φ ∈ H∞(D), it is easy to get that kφk∞ = sup{φ(z); z < 1} is
a norm on φ ∈ H∞(D). for any given g ∈ H2(D), the multiplication operator
Mφ(g) = φg associated with φ on H2(D) is a bounded linear operator, and by
the norm on H2(D) we get kMφ(g)k ≤ kφk∞kgkH2. If we denote H∞(T) as
the closed span of all Taylor expansions of functions in L∞(T), then H2(T) is
a closed subspace of L2(T) and H∞(D) and H∞(T) are naturally isomorphic
by the properties of complex analytic functions [7]P55P97 and the Dirichlet
Problem [7]P103.
There are more properties about Hardy space in [4]P7, [16]P48, [21]P39,
[28]P133 and [30]P106.
Definition 5. For a connected open subset Ω of C,n ∈ N, let Bn(Ω) denotes
the set of all bounded linear operator T on H that satisfies:
(a) Ω ∈ σ(T ) = {ω ∈ C : T − ωnot invertible};
(b) ran(T − ω) = H for ω ∈ Ω;
(c) W kerω∈Ω(T − ω) = H;
(d) dim ker(T − ω) = n for ω ∈ Ω.
21
If T ∈ Bn(Ω), then say that T is a Cowen-Douglas operator.
Theorem 8 ([2]). For a connected open subset Ω of C,T ∈ Bn(Ω),we get
(1) If ΩT T 6= ∅,then T is Devaney chaotic.
(2) If ΩT T 6= ∅,then T is distributionally chaotic.
(3) If ΩT T 6= ∅,then T strong mixing.
Definition 6 ([28]P141). Let P(z) be the set of all polynomials about z,where
z ∈ T. Define a function h(z) ∈ H2(T) is an outer function if cl[h(z)P(z)] =
H2(T).
Lemma 6 ([28]P141). A function h(z) ∈ H∞(T) is invertible on the Banach
algebra H∞(T), if and only if h(z) ∈ L∞(T) and h(z) is an outer function.
Theorem 9 ([16]P81). Let P(z) be the set of all polynomials about z,where
z ∈ D. then h(z) ∈ H2(D) is an outer function if and only if P(z)h(z) =
{p(z)h(z); p ∈ P(z)} is dense in H2(D).
Let φ is a non-constant complex analytic function on D, for any given
z0 ∈ D, by [7]P29 we get that there exists δz0 > 0, exists kz0 ∈ N, when
z − z0 < δz0, there is
φ(z) − φ(z0) = (z − z0)kz0 hz0(z)
where hz0(z) is complex analytic on a neighbourhood of z0 and hz0(z0) 6=
0.
22
Definition 7. Let φ is a non-constant complex analytic function on D, for
any given z0 ∈ D, there exists δz0 > 0 such that
φ(z) − φ(z0)z−z0<δz0
= pnz0
(z)hz0(z)z−z0<δz0
,
hz0(z) is complex analytic on a neighbourhood of z0 and hz0(z0) 6= 0,
(z) is a nz0-th polynomial and the nz0-th coefficient is equivalent 1. By
pnz0
the Analytic Continuation Theorem [7]P28, we get that there is a unique
complex analytic function hz0(z) on D such that φ(z)−φ(z0) = pnz0
then define hz0(z) is a rooter function of φ at the point z0. If for any given
(z)hz0(z),
z0 ∈ D, the rooter function hz0(z) has non-zero point but the roots of pnz0
are all in D and nz0 ∈ N is a constant on D that is equivalent m, then define
(z)
φ is a m-folder complex analytic function on D.
Definition 8. Let φ(z) ∈ H∞(D), n ∈ N, Mφ is the multiplication by φ on
H2(D). if the adjoint multiplier M∗φ ∈ Bn( ¯φ(D)), then define φ is a Cowen-
Douglas function.
By Definition 8 we get that any constant complex analytic function is not
a Cowen-Douglas function
Theorem 10. Let φ(z) ∈ H∞(D) be a m-folder complex analytic function,
Mφ is the multiplication by φ on H2(D). If for any given z0 ∈ D, the rooter
functions of φ at z0 is a outer function, then φ is a Cowen-Douglas function,
that is,the adjoint multiplier M∗φ ∈ Bm( ¯φ(D)).
Proof. By the definition of m-folder complex analytic function Definition 7
we get that φ is not a constant complex analytic function. For any given
23
z ∈ D, let fz ∈ H2(D) be the reproducing kernel at z. We confirm that M∗φ
is valid the conditions of Definition 5 one by one.
(1) For any given z ∈ D, fz is an eigenvector of M∗φ with associated
eigenvalue λ = ¯φ(z).Because for any g ∈ H2(D) we get
< g, M∗φ(fz) >H2=< φg, (fz) >H2= φ(z)f (z) =< g, ¯φ(z)fz >H2
By the Riesz Representation Theorem[14]P13 of bounded linear func-
tional in the form of inner product on Hilbert space, we get M∗φ(fz) =
¯φ(z)fz = λfz, that is,fz is an eigenvector of M∗φ with associated eigenvalueλ =
¯φ(z).
(2) For any given ¯λ ∈ φ(D), because of 0 6= φ ∈ H∞(D), we get that
the multiplication operator Mφ − λ is injection by the properties of complex
analysis, hence ker(Mφ − λ) = 0. Because of H2(D) = ker(Mφ − λ)⊥ =
cl[ran(M∗φ − ¯λ)], we get that ran(M∗φ − ¯λ) is a second category space. By
[14]P305 we get that M∗φ − ¯λ is a closed operator, also by [14]P93 or [31]P97
we get ran(M∗φ − ¯λ) = H2(D).
(3) Suppose that span{fz; z ∈ φ(D)} = span{ 1
1−¯zs ; z ∈ φ(D)} is not
dense in H2(D), By the definition of reproducing kernel fz and because of
0 6= φ ∈ H∞(D), we get that there exists 0 6= g ∈ H2(D),for any given z ∈ D,
we have
0 =< g, ¯φ(z)fz >H2= φ(z)g(z) =< φ(z)g(z), fz >H2 .
So we get g = 0 by the Analytic Continuation Theorem [7]P28,that is a
24
contradiction for g 6= 0. Therefor we get that span{fz; z ∈ φ(D)} is dense in
H2(D), that is,W ker¯λ∈φ(D)(M∗φ − ¯λ) = H2(D).
(4) By Definition 7 and the conditions of this theorem, for any given
λ ∈ φ(D), there exists z0 ∈ D, exists m-th polynomial pm(z) and outer
function h(z) such that
φ(z) − λ = φ(z) − φ(z0) = pm(z)h(z),
We give dim ker(M∗φ(z) − ¯λ) = m by the following (i)(ii)(iii) assertions.
(i) Let the roots of pm(z) are z0, z1, · · · , zm−1, then there exists decompo-
sition pm(z) = (z −z0)(z −z1) · · · (z −zm−1), and denote pm,z0z1···zm−1(z) is the
decomposition of pm(z) by the permutation of z0, z1, · · · , zm−1, the following
to get dim ker M∗pm,z0 z1···zm−1
= m.
By the Taylor expansions of functions in H2(T), we get there is a naturally
isomorphic
Fs : H2(D) → H2(D − s), Fs(g(z)) → g(z + s), s ∈ C.
It is easy to get that G = {Fs; s ∈ C} is a abelian group by the composite
operation ◦, hence for 0 ≤ n ≤ m − 1,there is
H2((D))
Fzn ↓
H2(D − zn)
−−−−−−−−−−−−−−−−→
Mz−zn
−−−−−−−−−−−−−−−−→
M ′
z
H2((D))
↓ Fzn
H2(D − zn)
Let T is the backward shift operator on the Hilbert space L2(N), that is,
T (x1, x2, · · · ) = (x2, x3, · · · ). With the naturally isomorphic between H2(D−
25
zn) and H2(∂(D − zn)), M ′∗z
H2(∂(D − zn)), that is,M ′∗z
is a surjection and dim ker M∗z−zn = 1, where 0 ≤ n ≤ m − 1.
is equivalent the backward shift operator T on
is a surjection and dim ker M ′∗z = 1, hence M∗z−zn
By the composition of Fzm−1 ◦ Fzm−2 ◦ · · · ◦ Fz0, M ′∗pm,z0z1···zm−1
is equivalent
that is,M ′∗pm,z0 z1···zm−1
T m.
hence M∗pm,z0z1···zm−1
is a surjection and dim ker M ′∗pm,z0 z1···zm−1
= m,
is a surjection and dim ker M∗pm,z0 z1···zm−1
= m.
(ii) Because H∞ is a abelian Banach algebra, Mpm is independent to the
permutation of 1-th factors of pm(z), that is,M∗pm is independent to the 1-th
factors multiplication of pm(z) = (z − z0)(z − z1) · · · (z − zm−1).
Because G = {Fs; s ∈ C} is a abelian group by composition operation ◦,
for 0 ≤ n ≤ m − 1, Fzm−1 ◦ Fzm−2 ◦ · · · ◦ Fz0 is independent to the permutation
of composition. Hence M∗pm is a surjection and
dim ker M∗pm = dim ker M∗pm,z0z1···zm−1
= m.
(iii) By Definition 6 and Theorem 9, also by [14]P93 or [31]P97 and
by [14]P305 we get that the multiplication operator Mh is surjection that
associated with the outer function h. Hence we get
ker M∗h(z) = (ranMh(z))⊥ = (H2(D))⊥ = 0.
Because there exists decomposition M∗pm(z)h(z) = M∗h(z)M∗pm(z) on H2(D),
we get
dim ker(M∗φ − ¯λ) = dim ker M∗pm(z)h(z) = dim ker M∗pm(z) = m.
26
By (1)(2)(3)(4) we get the adjoint multiplier operator M∗φ ∈ Bm( ¯φ(D)).
By Theorem 10 and Lemma 6, we get
Corollary 4. Let φ ∈ H∞(D) is a m-folder complex analytic function, for
any given z0 ∈ D, if the rooter function of φ at z0 is invertible in the Banach
algebra H∞(D), then φ is a Cowen-Douglas function. Especially,for any given
n ∈ D,if a and b are both non-zero complex, then a + bzn ∈ H∞(D) is a
Cowen-Douglas function.
The following gives some properties about the adjoint multiplier of Cowen-
Douglas functions.
Theorem 11. If φ ∈ H∞(D) is a Cowen-Douglas function, Mφ is the mul-
tiplication by φ on H2(D), Then the following assertions are equivalent
(1) M∗φ is Devaney chaotic;
(2) M∗φ is distributionally chaotic;
(3) M∗φ is strong mixing;
(4) M∗φ is Li-Yorke chaotic;
(5) M∗φ is hypercyclic;
(6) φ(D)T T 6= ∅.
27
Proof. By Example 1 we get that M∗φ is hypercyclic if and only if φ is non-
constant and φ(D)T T 6= ∅, hence (6) is equivalent to (5).
First to get that (6) imply (1)(2)(3)(4).
Because φ ∈ H2(D) is a Cowen-Douglas function, by Definition 8, M∗φ ∈
Bn( ¯φ(D)).
By Theorem 8 we get that if φ(D)T T 6= ∅, then (1)(2)(3) is valid. On
Banach spaces Devaney chaotic, distributionally chaotic and strong mixing
imply Li-Yorke chaotic, respectively.Hence (4) is valid.Because ¯φ(D)T T 6= ∅
and φ(D)T T 6= ∅ are mutually equivalent, (6) imply (1)(2)(3)(4).
Then to get that (1)(2)(3)(4) imply (6).By (1)(2)(3) imply (4),respectively,
it is enough to get that (4) imply (6).
If M∗φ is Li-Yorke chaotic, then we get that φ is non-constant and by
φ k → ∞, hence kMφk = kM∗φk > 1, that
[9]Theorem3.5 we get
kM∗n
sup
n→+∞
is,sup
z∈D
that inf
z∈D
φ(z) > 1. Moreover,we also have inf
z∈D
φ ∈ H∞ and kM∗1
φ(z) ≥ 1 then 1
0 6= x ∈ H2(D) we get kM∗n
φ xk ≥
contradiction to M∗φ is Li-Yorke chaotic.
φ(z) < 1, Indeed,if we assume
k = kM 1
φ
1
kM ∗−n
φ
k
kxk ≥ 1
kM ∗
φ
k ≤ 1. Hence for any
φ kn kxk ≥ kxk. It is a
1
Therefor that M∗φ is Li-Yorke chaotic imply inf
z∈D
φ(z) < 1 < sup
z∈D
φ(z),
By the properties of a simple connectedness argument of complex analytic
functions we get φ(D)T T 6= ∅. Hence we get (1)(2)(3)(4) both imply (6).
Corollary 5. If φ ∈ H∞(D) is a invertible Cowen-Douglas function in the
Banach algebra H∞(D), and let Mφ be the multiplication by φ on H2(D).
28
Then M∗φ is Devaney chaotic or distributionally chaotic or strong mixing or
Li-Yorke chaotic if and only if M∗−1
is.
φ
Proof. Because of T = (T −1)−1, it is enough to prove that M∗φ is Devaney
chaotic or distributionally chaotic or strong mixing or Li-Yorke chaotic imply
M∗−1
φ
is.
By Definition 8 we get M∗φ ∈ Bn( ¯φ(D)), with a simple computing we get
φ ∈ Bn( 1
¯φ (D)).
M∗−1
If M∗φ is Devaney chaotic or distributionally chaotic or strong mixing or
Li-Yorke chaotic, by Theoem 11 we get φ(D)T T 6= ∅, and by the properties
of complex analytic functions we get 1
φ (D)T T 6= ∅.
Because of M∗−1
φ ∈ B1( 1
¯φ (D)) and by Theorem 11 we get M∗−1
φ
is Devaney
chaotic or distributionally chaotic or strong mixing or Li-Yorke chaotic.
5. The chaos of scalars perturbation of an operator
We now study some properties about scalars perturbation of an operator
inspired by [10] and [1] that research some properties about the compact
perturbation of scalar operator.
Definition 9. Let λ ∈ C,T ∈ L(H).Define
(i) Let SLY (T ) denote the set such that λI + T is Li-Yorke chaotic for
every λ ∈ SLY (T ).
29
(ii) Let SDC(T ) denote the set such that λI + T is distributionally chaotic
for every λ ∈ SDC(T ).
(iii) Let SDV (T ) denote the set such that λI + T is Devaney chaotic for
every λ ∈ SDV (T ).
(iv) Let SH (T ) denote the set such that λI + T is hypercyclic for every
λ ∈ SH (T ).
By Definition 9 we get SLY (λI + T ) = λ + SLY (T ), SDC(λI + T ) =
λ + SDC(T ), SDV (λI + T ) = λ + SDV (T ) and SH(λI + T ) = λ + SH(T ).
Lemma 7. Let T ∈ L(H) be a normal operator,then SLY (T ) = ∅.
Proof. Because T is a normal operator,λI + T is a normal operator,too. by
[10] we get SLY (T ) = ∅.
Lemma 8. There is a quasinilpotent compact operator T ∈ L(H) such that
SLY (T ) = T is closed and SLY (T ∗) = ∅,where T = ∂D, D = {z ∈ C, z < 1}.
Proof. Let {en}n∈N be a orthonormal basis of L2(N) and let T be a weighted
n }+∞n=1 such that Sω(e0) =
backward shift operator with weight sequence {ωn = 1
0, Sω(en) = ωnen−1, where 0 < ωn < M < +∞,∀n > 0.
By the Spectral Radius formula[14]P197 rσ(T ) = lim
n→+∞
σ(T ) = {0}, hence σ(λI + T ) = λ.
kT nk 1
n we get
(1) If λ < 1, we can select ε > 0 such that λ + ε < q < 1. Then
30
∀0 6= x ∈ H we get
lim
n→∞
k(λI + T )nxk ≤ lim
n→∞
k(λI + T )nkkxk ≤ lim
n→∞
(λ + ε)nkxk = 0.
(2) If λ > 1,because of σ((λI + T )−1) = 1
λ,then we get
lim
n→∞
k(λI + T )nxk ≥ lim
n→∞
1
k(λI + T )−nk
kxk ≥ lim
n→∞
1
k(λI + T )−1kn kxk ≥ kxk 6= 0.
(3) By Theorem 3 we get that if λ = 1, then λI + T is mixing. Mixing
imply Li-Yorke chaotic.
(4) Because of σ(T ) = σ(T ∗), by (1)(2) we get that if λ 6= 1,then λ + T ∗
is not Li-Yorke chaotic.
(5) If λ = 1,from the view of infinite matrix λI + T ∗ is lower triangu-
lar matrix, then with a simple computing,for any 0 6= x ∈ L2(N), we get
kλI + S∗ωxk > 0. Hence λI + T ∗ is not Li-Yorke chaotic.
lim
n→∞
By (1)(2)(3)(4)(5) we get that SLY (T ) = T is closed and SLY (T ∗) = ∅.
Lemma 9. Let T be the backward shift operator on L2(N), T (x1, x2, · · · )
= (x2, x3, · · · ). Then SLY (T ) = SDC(T ) = SDV (T ) = SH(T ) = 2D \ {0},
SLY (2T ) = SDC(2T ) = SDV (2T ) = SH(2T ) = 3D, Hence SLY (T ) and
SLY (2T ) are open sets.
Proof. By [14]P209 we get σ(T ) = clD and σ(2T ) = cl2D, by Definition σ(T )
we get σ(λI + T ) = λ + clD. Because of the method to prove the conclusion
is similarly for T and 2T , we only to prove the conclusion for T .
31
By the naturally isomorphic between H2(T) and H2(D). Let L2(N) =
H2(T), by the definition of T we get (λI + T )∗ is the multiplication operator
Mf by f (z) = ¯λ + z on the Hardy space H2(T). By the Dirichlet Problem
[7]P103 we get that f (z) is associated with the complex analytic function
φ(z) = ¯λ + z ∈ H∞(D) determined by the boundary condition φ(z)T = f (z).
By Corollary 4 we get that φ is a Cowen-Douglas function. Therefor
by the natural isomorphic between H2(T) and H2(D), λI + T is naturally
equivalent to the operator M∗φ on H2(D).
By Theorem 11 we get that M∗φ is hypercyclic or Devaney chaotic or
distributionally chaotic or Li-Yorke chaotic if and only if φ(D)T T 6= ∅.
Because of σ(λI +T ) = σ(¯λI +T ∗), we get σ(λI +T ) = σ(M∗φ) = σ(Mφ) ⊇
φ(D), hence SLY (T ) = SDC(T ) = SDV (T ) = SH(T ) = 2D \ {0} is an open
set.
Therefor we can get
Corollary 6. Let T be the backward shift operator on L2(N), T (x1, x2, · · · ) =
(x2, x3, · · · ). For λ 6= 0, a 6= 0, n ∈ N, if λ+aT n is a invertible bounded linear
operator, then λ+aT n is strong mixing or Devaney chaotic or distributionally
chaotic or Li-Yorke chaotic if and only if (λ + aT n)−1 is.
Theorem 12. There is T ∈ L(H), SLY (T ) is neither open nor closed.
Proof. Let T1, T2 ∈ L(H), because T1 or T2 is Li-Yorke chaotic if and only if
T1 L T2 is, we get SLY (T1 L T2) = SLY (T1)S SLY (T2).
32
By Lemma 8,Lemma 9 and Definition 9 we get the conclusion.
6. Examples that T is chaotic but T −1 is not
In the last we give some examples to confirm the theory giving by func-
tional calculus on the begin that T is chaotic but T −1 is not.
Example 2. Let {en}n∈N be a orthonormal basis of L2(N) and let Sω be a
backward shift operator on L2(N) with weight sequence ω = {ωn}n≥1 such
that Sω(e0) = 0, Sω(en) = ωnen−1,where 0 < ωn < M < +∞, ∀n > 0.
(1) If λ = 1,then λI +Sω is Li-Yorke chaotic, but λI +S∗ω and (λI +S∗ω)−1
are not Li-Yorke chaotic.
(2) Let (λI + Sω)n = Un(λI + Sω)n is the polar decomposition of (λI + Sω)n,
{Un}∞n=1 is not a constant sequence.
Proof. (1) By Theorem 3, we get that for λ = 1,λI +Sω is mixing and mixing
imply Li-Yorke chaotic, hence λI + Sω is Li-Yorke chaotic. From the view of
infinite matrix,λI + S∗ω and (λI + S∗ω)−1 are lower triangular matrix, with a
simple computing,for any 0 6= x ∈ L2(N), we get lim
kλI + S∗ωxk > 0,and
n→∞
k(λI + S∗ω)−1xk > 0. Hence λI + S∗ω and (λI + S∗ω)−1 are not Li-Yorke
lim
n→∞
chaotic.
(2) Let (λI + Sω)n = Un(λI + Sω)n is the polar decomposition of (λI + Sω)n.
If {Un}∞n=1 is a constant sequence, then by Theorem 4 we get that (λI +S∗ω)−1
is Li-Yorke chaotic. A contradiction. Hence {Un}∞n=1 is not a constant se-
quence.
33
Theorem 13 ([10]). For any ε > 0, there is a small compact operator Kε ∈
L(H) and kKεk < ε such that I + Kε is distributionally chaotic.
In [10], I+Kε =
(Ij+Kj) is distributionally chaotic. where H =
+∞
Lj=1
+∞
Lj=1
Hj,
nj = 2mj,Hj is the nj-dimension subspace of H. On Hj define:
0 2εj
. . .
Sj =
. . .
. . . 2εj
0
nj×nj
, Kj =
−εj 2εj
. . .
. . .
. . .
2εj
−εj
nj×nj
.
Ij + Kj =
1 − εj 2εj
. . .
. . .
. . .
2εj
1 − εj
nj×nj
= (1 − εj)Ij + Sj.
We can construct a invertible bounded linear operator I + Kε in the same
way that is Li-Yorke chaotic, but (I + Kε)−1 is not.
Example 3. There is a invertible bounded linear operator I + Kε on H =
L2(N) such that I + Kε is Li-Yorke chaotic, but (I + Kε)−1,(I + Kε)∗−1 and
(I + Kε)∗ are not Li-Yorke chaotic.
Proof. Let {ei}∞i=1 is a orthonormal basis of H = L2(N) and Let H =
+∞
Lj=1
Hj,
j ∈ N, where Hj = span{ei}, 1 + j(j−1)
2 ≤ i ≤ j(j+1)
2
, Hj is j-dimension
subspace of H. For any given positive sequence {εj}∞1 such that εj → 0 and
(1 + εj)j → +∞, on Hj define:
sup
j→∞
34
0 2εj
. . .
Sj =
. . .
. . . 2εj
0
j×j
, Kj =
−εj 2εj
. . .
. . .
. . .
2εj
−εj
j×j
Ij + Kj =
1 − εj 2εj
. . .
. . .
. . .
2εj
1 − εj
j×j
= (1 − εj)Ij + Sj
First to prove that (I + Kε) is Li-Yorke chaotic.
.
.
Let I+Kε =
(I+Kj), fj = 1√j (11, · · · , 1j) and fj,n = 1√j (11, · · · , 1n, 0, · · · , 0) ∈
+∞
Lj=1
Hj. for any 1 ≤ n ≤ j we get
k(I + Kε)n(fj)k
= k(Ij + Kj)n(fj)k
= k((1 − εj)Ij + Sj)n(fj)k
= k
≥ k
n
n
Pk=0
Pk=0
C k
n(1 − εj)kSn−k
j
fjk
C k
n(1 − εj)k(2εj)n−k(11, · · · , 1n, 0, · · · , 0)k
= (1 + εj)nkfj,nk.
Hence we get
35
(a) lim
j→∞
k(I + Kε)j(fj)k ≥ lim
n→∞
kfjk(1 + εj)j = +∞.
Because of rσ(I + Kε) < 1,we get
(b) lim
n→∞
k(I + Kε)n(fj)k = 0.
By (a)(b) and by Definition 3 we get that λI + Kε satisfies the Li-Yorke
Chaos Criterion, by Theorem 1 we get that λI + Kε is Li-Yorke chaotic.
Then to prove that (I + Kε)−1 is not Li-Yorke chaotic. For convenience
we define n?m by induction on m for any given n ∈ N.
For any given j ∈ N,define:
(1) j? = 1 + 2 + · · · + j,
(2) If defined j?n , then define j?n+1 = 1?n + 2?n + · · · + j?n.
Let A =
+∞
Lj=1
Aj,where Aj = (Ij + Kj)−1.Because of Aj(Ij + Kj) = Aj(Ij +
Kj) = Ij, we get A(I + Kε) = (I + Kε)A =
+∞
Lj=1
Ij = I. By the Banach Inverse
Mapping Theorem [14]P91 we get that A = (I + Kε)−1 is a bounded linear
operator, that is,A = (I + Kε)−1.Hence we get
Aj =
1
1 − εj
(−2)j−1εj−1
j
−2εj
1
j×j
1 −2εj
. . .
. . .
. . .
36
.
A2
j =
1
(1 − εj)2
1 2 · (−2)εj
j · (−2)j−1εj−1
j
. . .
. . .
. . .
2 · (−2)εj
1
j×j
A3
j =
1
(1 − εj)3
1 (1 + 2)(−2)εj
j? · (−2)j−1εj−1
j
. . .
. . .
. . .
2?(−2)εj
1
j×j
For m ≥ 3, m ∈ N,If defined
Am
j =
1
(1 − εj)m
1 2?m−2(−2)εj
j?m−2(−2)j−1εj−1
j
. . .
. . .
. . .
2?m−2(−2)εj
1
j×j
Then define
.
.
.
37
A(m+1)
j
= AAm =
1
(1 − εj)(m+1)
1 2?m−1(−2)εj
j?m−1(−2)j−1εj−1
j
. . .
. . .
. . .
2?m−1(−2)εj
1
j×j
.
.
For any given 0 6= x0 = (x1, x2, · · · , ) ∈ H, Let
yj = (x(1+ j(j−1)
y′j = (x
2
);
), · · · , x j(j+1)
, · · · , x( j(j+1)
2
(2+
j(j−1)
)
2
2 −1));
zj,m = x(1+ j(j−1)
2
) +
j
Pk=2
k?(m−2)(−2)k−1εk
j .
Following a brilliant idea of Zermelo,we shall give the conclusion by in-
duction.
(1) If y1 6= 0,then we get
lim
n→∞
kAnx0k = lim
n→∞
+∞
Pj=1
kAn
j yjk ≥ lim
n→∞
kAn
1 y1k = lim
n→∞
x1
(1−ε1)n = +∞.
(2) If y1 = 0,but y2 = (x2, x3) 6= 0.
(i) If x3 6= 0, by (1) we get
kAnx0k ≥ lim
n→∞
lim
n→∞
(ii) If x3 = 0 and ε2 > 1
x3
(1−ε1)n → +∞.
2?(n−2) ,because of y2 = (x2, x3) 6= 0,we get
lim
n→∞
kAnx0k = lim
n→∞
+∞
Pj=1
kAn
j yjk ≥ lim
n→∞
kAn
2 y2k
38
= lim
n→∞
1
(1−ε1)np(x2
2 + (2?n−2(−2)ε2x3)2) ≥ lim
n→∞
x2
(1−ε1)n = +∞.
(3) Assume for k ≤ m−1, there is lim
n→∞
Then for k = m and ym 6= 0 we get
kAnx0k → +∞ for Ak and yk 6= 0.
(i) If xm? 6= 0, by (1) we get
lim
n→∞
kAnx0k ≥ lim
n→∞
xm?
(1−ε1)n = +∞.
(ii) If xm? = 0 and εm >
1
m?(n−2) ,because of ym 6= 0,we get
lim
n→∞
= lim
n→∞
kAnx0k = lim
n→∞
(1−ε1)n qz2
1
+∞
Pj=1
kAn
j yjk ≥ lim
n→∞
kAn
mymk
m,n + kAn
m−1y′mk2.
If y′m 6= 0,by the induction hypothesis we get
kAn
kAnx0k ≥ lim
n→∞
lim
n→∞
If y′m = 0,because of ym 6= 0 we get x(1+ m(m−1)
m−1y′mk = +∞;
2
) 6= 0. by (1) we get
lim
n→∞
kAnx0k ≥ lim
n→∞
kzm,nk
(1−ε1)n = lim
n→∞
x
)
m(m−1)
(1+
2
(1−ε1)n = +∞.
Therefor for k = m and ym 6= 0, we get lim
n→∞
induction we get that for any m ∈ N and ym 6= 0,there is lim
n→∞
kAnx0k → +∞, by the
kAnx0k → +∞.
From (1)(2)(3) and H =
+∞
Lj=0
Hj, we get that for any given 0 6= x0 =
(x1, x2, · · · , ) ∈ H we can find m ∈ N such that ym 6= 0. Hence for any given
0 6= x0 = (x1, x2, · · · , ) ∈ H we get lim
n→∞
is not Li-Yorke chaotic. From the view of infinite matrix, (I + Kε)∗ and
kAnx0k = +∞. Therefor (I + Kε)−1
39
(I + Kε)∗−1 are lower triangular matrix, for any 0 6= x ∈ H,with a simple
computing we get lim
n→∞
(I + Kε)∗ and (I + Kε)∗−1 are not Li-Yorke chaotic.
k(I + Kε)∗nxk > 0,
k(I + Kε)∗−nxk > 0. Hence
lim
n→∞
Corollary 7. There is a invertible bounded linear operator I + Kε on H =
L2(N) such that I + Kε is distributionally chaotic, but (I + Kε)−1,(I + Kε)∗−1
and (I + Kε)∗ are not distributionally chaotic.
Proof. By the construction of Theorem 13, it is only to give the conclusion
by induction on {ki}∞i=1 as Example 3.
Theorem 14. There is T ∈ L(H), SLY (T ) = SDC(T ) = ω is an open arc of
T = {λ = 1; λ ∈ C}, and for ∀λ ∈ ω,we get that (λ + T )∗,(λ + T )∗−1 and
(λ + T )−1 are not Li-Yorke chaotic.
Proof. As Example 3,give the same H =
sequence {εj}∞1 such that εj → 0 and sup
j→∞
Hj,Sj and Kj, give positive
+∞
Lj=1
i + εjj → +∞,where i ∈ C.
Let λI + Kε =
+∞
Lj=1
(λI + Kj), so σ(λI + Kε) = {λ − εj; j ∈ N}.
(i) If λ < 1,because of εj → 0, we get that ∃N > 0 when n > N,
λ − εj < 1. With the introduction of this paper we get that Li-Yorke
chaos is valid only on infinite Hilbert space. Loss no generally,for any j ∈
N,let λ − εj < 1, so rσ(λI+Kε) < 1. Hence for any 0 6= x ∈ H there is
k(I + Kε)n(x)k = 0.
lim
n→∞
(ii) If λ > 1 or λ ∈ [ π
2 , 3π
2 ], because of εj > 0, for j ∈ N we get
40
λ − εj > 1,
1
λ−εj
< 1, and σ(λI + Kε)−1 = { 1
λ−εj
; j ∈ N}. By Example 3,
for any given x 6= 0 there is ym 6= 0, m ∈ N,hence we get
k(λI + Kε)nx0k
k(λIm + Km)nymk
1
kymk
k(λIm+Km)−nk
k(λIm+Km)−1kn kymk
1
lim
n→∞
≥ lim
n→∞
≥ lim
n→∞
≥ lim
n→∞
≥ kymk > 0.
(iii) For ∀λ ∈ (− π
2 , π
j > N, we get λ − εj < 1. Let H
2 ), because of εj → 0, there exists N > 0, when
Hj ,
Hj, (λI + Kε)′ = (λI + Kε) L
′ = Lj>N
j>N
then for fj = 1√j (11, · · · , 1j) and fj,n = 1√j (11, · · · , 1n, 0, · · · , 0) ∈ H
that when 1 ≤ n ≤ j,there is
′, we get
k(λI + Kε)n(fj)k
= k(λIj + Kj)n(fj)k
= k((λ − εj)Ij + Sj)n(fj)k
≥ k
n
Pk=0
C k
n(λ − εj)k(2εj)n−k(11, · · · , 1n, 0, · · · , 0)k
= λ + εjnkfj,nk.
By (i)(ii),if λ 6= 1 or λ ∈ [ π
2 , 3π
2 ], λI + Kε is not Li-Yorke chaotic.
By (iii) and by the property of the triangle,if λ ∈ (− π
2 , π
get λ + εj > i + εj and rσ((λI + Kε)′) < 1. Hence we get
2 ) and j > N, we
k(λI + Kε)′n(fj)k = 0, and
lim
n→∞
lim
j→∞
k(I + Kε)′j(fj)k ≥ lim
j→∞
kfjkλ + εjj ≥ lim
j→∞
i + εjj = +∞.
41
By Definition 3 we get that λI +Kε satisfies the Li-Yorke Chaos Criterion,
by Theorem 1 we get that λI + Kε is Li-Yorke chaotic.
Using the same proof of Example 3 we get that (λI + Kε)∗,(λI + Kε)∗−1
and (λI + Kε)−1 are not Li-Yorke chaotic.
Using Corollary 7 and Theorem 13 we get that λI + Kε is distributionally
chaotic,but (λI +Kε)∗,(λI +Kε)∗−1 and (λI +Kε)−1 are not Li-Yorke chaotic.
Conjecture 1. For any given m ∈ N, there exists m-folder complex analytic
function φ(z) ∈ H∞(D) such that φ is not a Cowen-Douglas function.
Question 1. Gives the equivalent characterization of a m-folder complex
analytic function; Gives the equivalent characterization of a rooter function;
Gives the equivalent characterization of a Cowen-Douglas function. Gives
the relations between them.
Question 2. Let Mφ is the multiplication operator of the Cowen-Douglas
function φ(z) ∈ H∞(D) on the Hardy space H2(D), then is M∗φ a Lebesgue
operator? If not and if they have relations ,gives the relations between them.
References
[1] Bermdez, Bonilla, Martnez-Gimnez and Peiris. Li-Yorke and distributionally
chaotic operators. J.Math.Anal.Appl.,(373)2011:1-83-93.
[2] B.Hou,
P.Cui
and Y.Cao. Chaos
for Cowen-Douglas
operators.
Pro.Amer.Math.Soc,138(2010),926-936.
42
[3] C.Kitai. Invariant closed sets for linear operators. Ph.D.thesis, University of
Toronto, Toronto, 1982.
[4] F.Bayart and E.Matheron. Dynamics of Linear Operators, Cambridge Uni-
versity Press, 2009.
[5] G.D.Birkhoff. Surface transformations and their dynamical applications. Acta
Mathmatica, 1922:43-1-119.
[6] G.Godefroy. Renorming of Banach spaces. In Handbook of the Geometry of
Banach Spaces. volume 1,pp.781-835, North Holland, 2003.
[7] Henri Cartan(Translated by Yu Jiarong). Theorie elementaire des fonctions
analytuques d'une ou plusieurs variables complexes. Higer Education Press,
Peking, 2008.
[8] Hua
Loo-kang.
On
the
automorphisms
of
a
sfield.
Proc.Nat.Acad.Sci.U.S.A.,1949:35-386-389.
[9] Hou B, Liao G, Cao Y. . Dynamics of shift operators. Houston Journal of
Mathematics, 2012: 38(4)-1225-1239
[10] Hou Bingzhe,Tian Geng and Shi Luoyi. Some Dynamical Properties For Lin-
ear Operators. Illinois Journal of Mathematics,Fall 2009.
[11] Iwanik,A.. Independent sets of transitive points. Dynamical Systems and Er-
godic Theory. vol.23,Banach Center publications,1989,pp277-282.
[12] J.H.Shapiro. Notes on the dynamics of linear operators. Available at the au-
thor'web page 2001.
[13] John Milnor. Dynamics in one complex variable. Third Edition. Princeton
University Press, 2006.
43
[14] John B.Conway. A Course in Functional Analysis. Second Edition,Springer-
Verlag New York, 1990.
[15] John B.Conway. A Course in Operator Theory. American Mathmatical Soci-
ety, 2000.
[16] John B.Garnett. Bounded Analytic Functions. Revised First Edition.
Springer-Verlag, 2007.
[17] J.R.Browm. Ergodic Theory and Topological Dynamics. Academic Press,New
York,1976.
[18] J.W.Robbin. Topological Conjungacy and Structural Stability for discrete
Dynamical Systems. Bulletin of the American Mathematical Society, Vol-
ume78, 1972.
[19] K.-G.Grosse-Erdmann and A.Peris Manguillot. Linear Chaos. Springer, Lon-
don, 2011.
[20] K.-G.Grosse-Erdmann. Universal
families and hypercyclic vectors. Bull.
Amer. Math. Soc.,36(3):345-381,1999.
[21] Kenneth Hoffman. Banach Spaces of Analytic Functions. Prentice-Hall, Inc.,
Englewood Cliffs, N.J., 1962.
[22] Li T.Y. and Yorke J.A.. Period three implies chaos. Amer.Math.Monthly
82(1975), 985C992.
[23] M.Shub. Global Stability of Dynamical Systems. Springer-Verlag New York,
1987.
[24] N.H.Kuiper and J.W.Robbin. Topological Classification of Linear Endomor-
phisms. Inventiones math.19,83-106, Springer-Verlag, 1973.
44
[25] Nilson C.Bernardes Jr., Antonio Bonilla, Vladimr Mller and A.peiris. Li-
Yorke chaos in linear dynamics. Academy of Sciences Czech Republic,
Preprint No.22-2012.
[26] Paul R.Halmos. Measure Theory. Springer-Verlag New York, 1974.
[27] Peter Walters. An Introduction to Ergodic Theory. Springer-Verlag New
Yorke, 1982.
[28] Ronald G.Douglas. Banach Algebra Techniques in Operator Theory. Second
Edition. Springer-Verlag New York, 1998.
[29] S.Rolewicz. On orbits of elements. Studia Math., 1969:32-17-22.
[30] William Arveson. A Short Course on Spectral Theory. Springer Sci-
ence+Businee Media, LLC, 2002.
[31] Zhang Gongqing,Lin Yuanqu. Lecture notes of functional analysis(Volume 1).
Peking University Press, Peking, 2006.
45
|
1812.11064 | 1 | 1812 | 2018-12-23T04:48:30 | Extension of differentiable local mappings on linear topological spaces | [
"math.FA"
] | Usually, for extension of local maps, one uses multiplication by so called bump functions. However, majority of infinite-dimensional linear topological spaces do not have smooth bump functions. Therefore, in \cite{BR} we suggested a new approach for Banach spaces, based on the composition with locally identical maps. In the present work we discuss a possibility of generalization of this method for arbitrary spaces and applications of this theory. | math.FA | math |
Extension of differentiable local mappings on linear
topological spaces
Genrich Belitskii and Victoria Rayskin
Abstract. Usually, for extension of local maps, one uses multiplication by
so called bump functions. However, majority of infinite-dimensional linear
topological spaces do not have smooth bump functions. Therefore, in [5] we
suggested a new approach for Banach spaces, based on the composition with
locally identical maps. In the present work we discuss a possibility of gener-
alization of this method for arbitrary spaces and applications of this theory.
Bump functions and Blid maps and Topological spaces and Map extension and
Linearization
1. Introduction
Let X and Y be linear topological spaces. In this work, we will discuss the
differentiable local maps f : X → Y and the possibility of differentiable extension
of the maps. The map's extension is usually not unique and can be studied in
the context of the equivalence class of f , i.e. a germ [f ]. Recall that a germ [f ]
at x ∈ X is the equivalence class of local maps, such that any pair of the class
members coincides on some neighborhood of x. Each element of the class is called
a representative of a germ. Occasionally, we denote germ [f ] as f . In the future,
without loss of generality, we will assume that x = 0. We are interested in the
question of existence of a global representative of the germ.
Recall that f : U ⊂ X → Y is differentiable at a point x ∈ U if
f (x + h) = f (x) + A · h + r(h)
with r(h) = o(h) and with the continuous operator A : X → Y .
Each meaning of "smallness" o(h) defines a corresponding specific notion of r(h)
and consequently a specific notion of differentiability. The value of the derivative is
A (usually denoted by f ′(x)) and it is independent of the definition of r(h). Various
definitions of differentiability are discussed in the works of F. and R. Nevanlinna
([8]), H. R. Fischer ([9]), H. H. Keller ([7]), E. inz ([2]), E. Binz, H. H. Keller
([3] ) and E. Binz, W. Meier - Solfrian ([4]). The main ideas of differentiation in
abstract spaces were developed and clarified in the sequence of works by Averbuh,
V. I. and Smoljanov, O. G ([1]), and later by M. Schechter in [12]. We remind all
these definitions and connections between them in Section 2. Thus, the question
investigated in this work is the following.
Question 1.1. Let f be a germ at 0, differentiable in some a priori specified
meaning. Does there exist a global differentiable (in the same sense) representative
of the germ?
The classical method of a smooth extension of a locally defined map f is mul-
tiplication by the smooth bump function. Recall that a smooth (differentiable)
1
2
GENRICH BELITSKII AND VICTORIA RAYSKIN
bump function is a real-valued function ranging between 0 and 1 with bounded
support, which is equal to 1 in a neighborhood 0. Because it is known that very
few spaces have smooth bump functions, we introduce1 here the new method based
on the idea of considering the composition of f with a locally identical map. Be-
cause we use composition in this construction, we will discuss only those notions of
differentiability that satisfy the Chain Rule of differentiation.
In Section 3 we show how locally defined maps with the help of our new method
can be extended to the entire space. We present examples of the extension con-
structed for the specific Banach spaces (Section 4) and for the specific non-Banach
linear topological spaces (Section 5).
In Section 6 we discuss applications of this theory to the problem of conjugation
with linear map. We show that for the construction of differentiable conjugation,
the assumption of the existence of smooth bump function is not necessary, and con-
sequently the corresponding conjecture stated in the paper of W. Zhang, K. Lu and
W. Zhang "Differentiability of the Conjugacy in the Hartman-Grobman Theorem"
([14]) is incorrect.
2. Background Definitions
Let us recall three definitions of differentiation on linear topological spaces,
which we use in this work. The reader can also find these definitions in [12]
1. Bounded differentiability
Definition 2.1. The map f : X → Y is bounded-differentiable at
x ∈ X, if for every bounded subset S ⊂ X and every h ∈ S and t ∈ R
r(th)/t → 0
uniformly in h as t → 0.
2. Compact (Hadamard) differentiability
Definition 2.2. The map f : X → Y is compact (Hadamard) differ-
entiable at x ∈ X, if
f (x + tnhn) − f (x) = tnAh + o(tn)
as tn → 0, and hn → h.
3. If both X and Y are Banach spaces with the norms .1 and .2 respec-
tively, then Fr´echet differentiation is well defined.
Definition 2.3. The map f is Fr´echet differentiable at 0 if
lim
h→0
r(h)2/h1 = 0.
Thus, as discussed in [1], bounded differentiability implies the compact one. The
latter, in turn, is equivalent to Hadamard differentiability (see [12]). If both X and
Y are Banach then bounded and Fr´echet differentiability coincide. For these types
of differentiability many important rules, including the Chain Rule, hold.
The compact (Hadamard) differentiability is the weakest for which the Chain
Rule is satisfied (for instance, for the Gateaux derivative the Chain Rule does not
hold).
1Earlier this idea was introduced in [5] for Banach spaces and for differentiability in Fr´echet
sense.
EXTENSION OF DIFFERENTIABLE LOCAL MAPPINGS ON LINEAR TOPOLOGICAL SPACES3
3. The Main Proposition
As explained above, we need to define differentiability, which satisfies the Chain
Rule. Thus, we assume that the Cain Rule holds for the next definition and for the
later discussion.
In our work [5] we defined the notion of the blid map for Banach spaces, which
stands for Bounded Local Identity map. In this paper we generalize this idea for
linear topological spaces.
Definition 3.1. A space X satisfies blid-differentiable property if for every
neighborhood U ⊂ X of 0 there is a differentiable map H defined on X, locally
coinciding with the identity map, such that H(X) ⊂ U .
Let us recall that a neighborhoods base of zero is a system B = {Vα} of
neighborhoods of 0, such that for any neighborhood U ⊂ X of 0 there exists some
Vβ ∈ B, Vβ ⊂ U .
Therefore, if there is a neighborhoods base B such that for every Vα from B
there exists local identity Hα, Hα(X) ⊂ Vα, then X satisfies the blid-property.
Proposition 3.2 (The Main Proposition). If X satisfies blid-differentiable
property, then every differentiable germ [f ] : X → Y has a global differentiable
representative.
Proof. Let f be a local representative of the germ defined on a neighborhood
U ⊂ X of zero. Let H : X → X be a differentiable local identity map such that
H(X) ⊂ U . Then the map
is a global representative of the germ as we need.
(cid:3)
F (x) = f (H(x)), x ∈ X
4. Banach Spaces
In this section we will consider a Banach space X and a general linear topolog-
ical space Y . In [5] we introduced the following definition:
Definition 4.1. A differentiable blid map for a space X is a global Bounded
Local Identity differentiable map H : X → X.
Lemma 4.2. If there exists differentiable blid map, then X satisfies differen-
tiable blid-property.
Proof. It is convenient to chose the balls Bc = {x ∈ X : x < c} to be the
base of neighborhoods in the Banach space X. Let H be a blid map, such that
H(x) < N . Then
Hc(x) =
c
N
H(
N
c
x)
is the blid map as well and its image is inside of Bc.
(cid:3)
The next statement immediately follows from the Lemma 4.2 and the Main
Proposition 3.2.
Corollary 4.3. If the space X admits differentiable blid map, then every
differentiable germ at 0 ∈ X into Y has a global differentiable representative.
Note, if Y is a Banach space as well, then this result is proved in [5] for
differentiability in Fr´echet sense.
Corollary 4.4. If the space X has a differentiable bump function h : X → R,
then it satisfies differentiable blid-property, and every differentiable germ at 0 ∈ X
has a global differentiable representative.
4
GENRICH BELITSKII AND VICTORIA RAYSKIN
Proof. The map
H(x) = h(x)x
is a differentiable blid map for X.
(cid:3)
4.1. The Spaces of Bounded Continuous Functions. In this section we
will consider the space X = C(T ) of bounded continuous functions on a topological
space T with the norm . = supT ..
In the space X we will construct a blid map in the following way.
H(x)(t) = h(x(t))x(t)
where h is a smooth bump function on the real line. This blid map is Fr´echet-
(consequently bounded-, consequently compact-) differentiable.
Corollary 4.5. Any bounded- (consequently compact-) differentiable germ at
0 ∈ C(T ) has a global (in the corresponding sense) differentiable representative.
4.2. The Spaces of Smooth Functions. In this section we will consider
X = Cq[0, 1], 0 < q < ∞, equipped with the norm xq = max0≤j≤q supt∈[0,1] x(j)(t).
Recall that on the real line all types of differentiability coincide, and consequently
can be viewed as Fr´echet differentiability. We will see that X has a differentiable
blid map, and consequently satisfies differentiable blid property.
Lemma 4.6. The space Cq[0, 1] admits Fr´echet-differentiable blid map.
Proof. The map
H(x)(t) =
tj
j!
h(x(j)(0))x(j)(0) +Z t
0
dt1Z t1
0
dt2...Z tq−1
0
h(cid:16)x(q)(s)(cid:17) x(q)(s) ds,
q−1
Xj=0
where h is a C∞ bump function on R, represents a bounded differentiable blid
map.
(cid:3)
Corollary 4.7. Any bounded- (compact-) differentiable germ at 0 ∈ Cq[0, 1]
has a global (in the corresponding sense) differentiable representative.
5. Metric Spaces
Let X be a metric space with a metric d(x, y). Here we consider germs of maps
from X into an arbitrary linear topological space Y . Although instead of Fr´echet
differentiation (which is not defined for general metric spaces) we use bounded
and compact (Hadamar) differentiation. The neighborhoods base B can be chosen
as a collection {Bc}c = {x ∈ X : d(x, 0) < c}c. Then the space X satisfies
differentiable-blid property if for every c there exists a differentiable, local identity
map Hc : X → X such that d(Hc(x), 0) < c for all x, i.e., Hc(X) ⊂ Bc.
In particular, if topology on X is defined by countable collection of norms xk,
then the metric can be written as
d(x, y) :=
1
2k ·
∞
Xk=0
x − yk
x − yk + 1
.
Lemma 5.1. Suppose for every k = 0, 1, ... there exists a global differentiable
local identity map Hk such that
Then X satisfies the differentiable blid property.
Hk(x)k < ∞.
sup
x
EXTENSION OF DIFFERENTIABLE LOCAL MAPPINGS ON LINEAR TOPOLOGICAL SPACES5
Proof. For a given c > 0 choose any
(5.1)
k > 1 − ln c/ ln 2
and let Hk be such that
Set
Hk(x)k < N, x ∈ X.
Hc(x) =
c
4N
Hk(cid:18) 4N
c
x(cid:19) .
Then inequality 5.1 and the fact that xj is monotonically increasing with j imply
that
d(Hc(x), 0) < c,
i.e. Hc(X) ∈ Bc.
(cid:3)
5.1. The Space of Smooth Functions on the Real Line. The space X =
Cq(R) (0 ≤ q < ∞) of all smooth functions on R is endowed with the collection of
norms
xk = max
t∈[−k,k]
max
l≤q
x(l)(t).
Lemma 5.2. The space X possesses the bounded- (consequently compact-) dif-
ferentiable blid property.
Proof. Let h(u) be a C∞-bump function on R, which equals to 1 in a neigh-
borhood of 0 and such that a = supu∈R h(u)u < ∞. Then
H(x)(t) =( h(x(t))x(t), q = 0
j=0
tj
Pq−1
j! h(x(j)(0))x(j)(0) +R t
0 dt1R t1
0 dt2...R tq−1
0
is differentiable local identity map, and
H(x)k < aek, k = 0, 1, ..., x ∈ X.
ha(cid:0)x(q)(s)(cid:1) x(q)(s) ds, q ≥ 1
(cid:3)
Corollary 5.3. Every bounded- (consequently compact-) differentiable germ
at 0 ∈ C(R) has a global differentiable (in the corresponding sense) representative.
5.2. The Space of Infinitely Differentiable Functions on a Closed In-
terval. The space X = C∞[0, 1] is endowed with the collection of norms
xk = max
j≤k
max
t∈[0,1]
x(j)(t).
Lemma 5.4. The space X possesses the bounded- (consequently compact-) dif-
ferentiable property.
Proof. Let h(u) be the same bump function on R as above. Then
is differentiable local identity map, and
H0(x)(t) = h(x(t))x(t)
H0(x)0 < a.
Further, let k > 0. Then
Hk(x)(t) =
k−1
Xj=0
tj
j!
h(x(j)(0))x(j)(0) +Z t
0
dt1Z t1
0
dt2...Z tk−1
0
h(cid:16)x(k)(s)(cid:17) x(k)(s) ds.
is differentiable local identity map, and
Hk(x)k < aek, k = 0, 1, ..., x ∈ X.
(cid:3)
6
GENRICH BELITSKII AND VICTORIA RAYSKIN
Corollary 5.5. Every bounded- (consequently compact-) differentiable germ
at 0 ∈ C∞[0, 1] has a global representative.
5.3. The Space of Infinitely Differentiable Functions on the Real
Line. The space X = C∞(R) is endowed with the collection of norms
xk = max
j≤k
max
t∈[−k,k]
x(j)(t), k = 0, 1, 2, ...
Lemma 5.6. The space X possesses the bounded- (consequently compact-) dif-
ferentiable property.
Proof. Let h(u) be the same bump function on R as above. Then
is differentiable local identity map, and
H0(x)(t) = h(x(t))x(t)
H0(x)0 < a.
Further, let k > 0. Then
Hk(x)(t) =
tp
j!
h(x(p)(0))x(p)(0) +Z t
0
dt1Z t1
0
dt2...Z tk−1
0
h(cid:16)x(k)(s)(cid:17) x(k)(s) ds.
k−1
Xp=0
is differentiable local identity map, and
Hk(x)k < aek, k = 0, 1, ..., x ∈ X.
(cid:3)
Corollary 5.7. Every bounded- (consequently compact-) differentiable germ
at 0 ∈ C∞(R) has a global representative.
In conclusion, we would like to pose the following
Question 5.8. Which linear topological spaces have differentiable blid prop-
erty?
This question was not answered even for Banach spaces.
6. Applications
Local linearization and normal forms are convenient simplification of complex
dynamics. In this section we discuss differentiable linearization on Banach spaces.
For a diffeomorphism F with a fixed point 0, we would like to find a smooth
transformation Φ defined in a neighborhood of 0 such that Φ ◦ F ◦ Φ−1 has a
simplified (polynomial) form called the normal form. If Φ ◦ F ◦ Φ−1 = DF = Λ is
linear, the conjugation is called linearization. There are two major questions in this
area of research: how to increase smoothness of the conjugation Φ, and whether it
is sufficient to assume low smoothness of the diffeomorphism F .
Hartman and Grobman independently showed that if Λ is hyperbolic, then for a
diffeomorphism F there exists a local homeomorphism Φ such that Φ◦ F ◦ Φ−1 = Λ.
Different proofs were given by Pugh [10]. A higher regularity of Φ has been an active
area of research.
The first attempt to answer the question of differentiability of Φ at the fixed
point 0 under hyperbolicity assumption was made in [48], but an error was found
and discussed in [11]. Later, in [6], Guysinsky, Hasselblatt and Rayskin presented
correct proof. However, it was restricted to F ∈ C∞ (or more precisely, it was
restricted to F ∈ Ck, where k is defined by complicated expression). It was con-
jectured in the paper that the result is correct for F ∈ C2, as it was announced in
[13].
EXTENSION OF DIFFERENTIABLE LOCAL MAPPINGS ON LINEAR TOPOLOGICAL SPACES7
Zhang, Lu and Zhang ([14], Theorem 7.1) showed that for a Banach space
diffeomorphism F with a hyperbolic fixed point and α-Holder DF , the local con-
jugating homeomorphism Φ is differentiable at the fixed point. Moreover,
Φ(x) = x + O(x1+β ) and Φ−1(x) = x + O(x1+β )
as x → 0, for certain β ∈ (0, α].
There are two additional assumptions in this theorem. The first one is the
spectral band width inequality. The authors explain that this inequality is sharp
if the spectrum has at most one connected component inside of the unit circle in
X, and at most one connected component outside of the unit circle in X. For
the precise formulation of the spectral band width condition we refer the reader to
the paper [14]. It is important (and it is pointed out in [14]) that this is not a
non-resonance condition. The latter is required for generic linearization of higher
smoothness.
The second assumption is the assumption that the Banach space must possess
smooth bump functions. It is conjectured in the paper that the second assumption
is a necessary condition.
In this section we explain that this conjecture is not correct (see Theorem 6.1).
The bump function condition can be replaced with the less restrictive blid map
condition. Blid maps allow to reformulate Theorem 7.1 in the following way:
Theorem 6.1. Let X be a Banach space possessing a differentiable blid map
with bounded derivative. Suppose F : X → X is a diffeomorphism with a hyperbolic
fixed point, DF is α-Holder, and the spectral band width condition is satisfied.
Then, there exists local linearizing homeomorphism Φ which is differentiable at the
fixed point. Moreover,
Φ(x) = x + O(x1+β ) and Φ−1(x) = x + O(x1+β )
as x → 0, for certain β ∈ (0, α].
In particular, we have the following
Corollary 6.2. Let X = Cq[0, 1]. Suppose F : X → X is a diffeomorphism
with a hyperbolic fixed point, DF is α-Holder, and the spectral band width condition
is satisfied. Then, the local conjugating homeomorphism Φ is differentiable at the
fixed point. Moreover,
Φ(x) = x + O(x1+β ) and Φ−1(x) = x + O(x1+β )
as x → 0, for certain β ∈ (0, α].
Below we justify Theorem 6.1
Proof. Zhang, Lu and Zhang showed that for the conclusion of their Theorem
7.1 it is enough to satisfy the inequalities 1 and 2 (see 6.2 below), which are called
condition (7.6) in their paper.
In order to apply the blid maps instead of bump functions to the inequalities
6.2, it is sufficient to construct a bounded blid map, which has only first-order
bounded derivative. I.e., let blid map H(x) : X → X be as follows:
(6.1)
1. H(x) = x for x < 1
2. H ∈ C1 and H (j)(x) ≤ cj, j = 0, 1.
The condition (7.6) of [14] is:
(6.2)
1.
2.
supx∈X DF (x) − Λ ≤ δη
supx∈V \O {DF (x) − Λ/xα} = M < ∞
8
GENRICH BELITSKII AND VICTORIA RAYSKIN
Let DF − Λ = f . Define
f (x) := f (δH(x/δ))
We will show that if f satisfies (7.6), then so does f .
sup
x∈X
D f (x) ≤ sup
x∈X
Df (x) · sup
x∈X
DH(x) ≤ δη · c1.
Thus, the first inequality of (7.6) holds for f . For the second inequality we have
the following estimate:
D f (x)
xα ≤
Df (δH(x/δ))
δH(x/δ)α
·(cid:18) δH(x/δ)
x
(cid:19)α
.
The second multiple is bounded, because for small x (say, x/δ < ǫ for some
ǫ > 0) we have
while for x/δ ≥ ǫ
δH(x/δ)
x
< c1 + o(1),
δH(x/δ)
x
< c0/ǫ.
I.e., δH(x/δ)
x
is less than some constant m. Then,
D f (x)
xα ≤
sup
x∈V \O
sup
0<x<δc0
{Df (x)/xα} · sup
x∈X
DH(x) · mα
=
sup
0<x<δc0
{Df (x)/xα} c1 · mα.
This quantity is bounded by M c1mα if δ is sufficiently small.
(cid:3)
Other applications in the area of local analysis on Banach spaces possessing
blid maps can be found in [5].
A generalization of Theorem 6.1 might be possible for the case of linear topo-
logical spaces (e.g., space of smooth functions), which posses differentiable blid
property.
References
[1] Averbuh, V. I.; Smoljanov, O. G, The theory of differentiation in linear topological spaces.
Russian Mathematical Surveys. 1967. V. 22, #6, 201 -- 260.
[2] E. inz, Ein Differenzierbarkeitsbegriff in limitierten Vektorraumen, Comment. Math. Helv. 41
(1966), 137 -- 156.
[3] E. Binz, H. H. Keller, Funktionenraume in der Kategorie der Limesraume, Ann. Acad. Sci.
Fennicae, ser. A, 383 (1966), 1 -- 21.
[4] E. Binz, W. Meier - Solfrian, Zur Differentialrechnung in limitierten Vektorraumen, Comment.
Math. Helv. 42, #4 (1967), 285 -- 296.
[5] G. Belitskii, V. Rayskin, A New Method of Extension of Local Maps of Banach Spaces. Ap-
plications and Examples
[6] M. Guysinsky, B. Hasselblatt, and V. Rayskin, Differentiability of the Hartman-Grobman
linearization, Discrete Contin. Dyn. Syst. 9 (2003), no. 4, 979984.
[7] H. H. Keller, Raume stetiger multilinearen Abbildungen als Limesraume, Math. Ann. 159
(1965), 259 -- 270.
[8] F. und R. Nevanlinna, Absolute Analysis, Springer Verl., 1959.
[9] H. R. Fischer, Limesraume, Math. Ann. 137 (1959), 269 -- 303.
[10] C. C. Pugh, On a theorem of P. Hartman, Amer. J. Math. 91 (1969), 363367.
[11] V. Rayskin α-Holder linearization J. Differential Equations 147 (1998), no. 2, 271284.
[12] M. Schechter, Differentiation in Abstract Spaces, Journal of Differentiable Equations, 55
(1984) 330 -- 345.
[13] S. van Strien, Smooth linearization of hyperbolic fixed points without resonance conditions,
J. Differential Equations 85 (1990), no. 1, 6690.
EXTENSION OF DIFFERENTIABLE LOCAL MAPPINGS ON LINEAR TOPOLOGICAL SPACES9
[14] W. Zhang, K. Lu and W. Zhang Dfifferentiability of the Conjugacy in the Harman-Grobman
Theorem, Transactions of the American Mathematical Society, 369, Number 7 (2017), 4995-5030.
E-mail address: [email protected]
E-mail address: [email protected]
|
1105.0467 | 1 | 1105 | 2011-05-03T02:01:15 | A note on the boundedness of Riesz transform for some subelliptic operators | [
"math.FA",
"math.AP",
"math.DG"
] | Let $\M$ be a smooth connected non-compact manifold endowed with a smooth measure $\mu$ and a smooth locally subelliptic diffusion operator $L$ satisfying $L1=0$, and which is symmetric with respect to $\mu$. We show that if $L$ satisfies, with a non negative curvature parameter $\rho_1$, the generalized curvature inequality in \eqref{CD} below, then the Riesz transform is bounded in $L^p (\bM)$ for every $p>1$, that is \[\| \sqrt{\Gamma((-L)^{-1/2}f)}\|_p \le C_p \| f \|_p, \quad f \in C^\infty_0(\bM), \] where $\Gamma$ is the \textit{carr\'e du champ} associated to $L$. Our results apply in particular to all Sasakian manifolds whose horizontal Tanaka-Webster Ricci curvature is nonnegative, all Carnot groups with step two, and wide subclasses of principal bundles over Riemannian manifolds whose Ricci curvature is nonnegative. | math.FA | math |
A NOTE ON THE BOUNDEDNESS OF RIESZ TRANSFORM FOR SOME
SUBELLIPTIC OPERATORS
FABRICE BAUDOIN AND NICOLA GAROFALO
Abstract. Let M be a smooth connected non-compact manifold endowed with a smooth mea-
sure µ and a smooth locally subelliptic diffusion operator L satisfying L1 = 0, and which is
symmetric with respect to µ. We show that if L satisfies, with a non negative curvature param-
eter ρ1, the generalized curvature inequality in (2.9) below, then the Riesz transform is bounded
in Lp(M) for every p > 1, that is
(cid:13)(cid:13)(cid:13)(cid:13)
qΓ((−L)−1/2f )(cid:13)(cid:13)(cid:13)(cid:13)p
≤ Cpkf kp,
f ∈ C∞
0 (M),
where Γ is the carr´e du champ associated to L. Our results apply in particular to all Sasakian
manifolds whose horizontal Tanaka-Webster Ricci curvature is nonnegative, all Carnot groups
with step two, and wide subclasses of principal bundles over Riemannian manifolds whose Ricci
curvature is nonnegative.
Contents
1.
Introduction
2. Background
2.1. Assumptions
2.2. Some known results
3. Riesz transform
3.1. The case 1 < p ≤ 2
3.2. The case p > 2
4. Pointwise gradient estimates of the heat kernel
References
1. Introduction
1
5
5
7
8
9
9
14
15
A central result in the analysis of Rn is the Lp continuity of singular integrals in the range
1 < p < ∞. One basic consequence of this result is the Lp boundedness of the Riesz transforms
Rj = ∂
(−∆)−1/2, j = 1, ..., n, with their vector-valued counterpart
∂xj
R = (R1, ...,Rn) = ∇(−∆)−1/2,
see [S]. In [Str] Strichartz asked the question whether such Lp continuity of the Riesz transform
could be extended to non-compact Riemannian manifolds under suitable assumptions on the
latter. In this context the analogue of the vector-valued Riesz transform is the operator
where we have denoted by ∆ the Laplacian on M in its realization as a positive self-adjoint
operator on L2(M). Strichartz's question is important for the purpose of developing analysis on
R = ∇∆−1/2,
First author supported in part by NSF Grant DMS-0907326.
Second author supported in part by NSF Grant DMS-1001317.
1
2
FABRICE BAUDOIN AND NICOLA GAROFALO
manifolds. To explain this point let us indicate by (·,·) the inner product in L2(M), and with
∆1/2 the positive self-adjoint square root of ∆. Then one has the equality
(∆f, f ) = (∆1/2f, ∆1/2f ).
This immediately gives
k∇fk2 = k∆1/2fk2,
which in turn allows to identify the first-order Sobolev subspaces of L2(M) obtained by com-
0 (M) with respect to the seminorms k∇fk2 and k∆1/2fk2. Let us also notice in
pletion of C ∞
passing that the latter equality can be reformulated in terms of R as follows
kRfk2 = kfk2.
However, when 1 < p < ∞ and p 6= 2, a similar identification of the two Sobolev spaces of order
0 (M) with respect to the seminorms k∇fkp and k∆1/2fkp is
one obtained by completion of C ∞
no longer such a simple matter. It is a well-known fact that an estimate such as
(1.1)
Apk∆1/2fkp ≤ k∇fkp ≤ Bpk∆1/2fkp,
f ∈ C ∞
0 (M),
1
′
p
= 1.
would suffice for such identification. It is also known that the validity of the right-hand inequality
in (1.1) for a certain 1 < p < ∞ implies that of the left-hand inequality in Lp′
(M), where
p + 1
Now the right-hand inequality in (1.1) is equivalent to the Lp continuity of the Riesz operator
R. It is then clear that (1.1) is true for all 1 < p < ∞ if
(1.2)
Rfp ≤ Cpfp,
f ∈ C ∞
0 (M),
for 1 < p < ∞, and this clarifies the relevance of the question raised by Strichartz.
An interesting result due to Bakry [B] states that if the Ricci curvature of M is bounded from
below by a non negative constant then (1.2), and therefore (1.1) hold for every 1 < p < ∞. The
purpose of the present note is to extend this result to a sub-Riemannian framework by using
the generalized curvature-dimension inequality recently introduced by the authors in [BG1].
This extension has been recently become possible thanks to a combination of the theory devel-
oped in the two papers [BG1], [BBG], with the remarkable results in [CD], [ACDH]. The latter
two works have established that (1.2) does hold in the range 1 < p < ∞ for complete, non-
compact Riemannian manifolds satisfying suitable general assumptions which will be discussed
below. In [CD] the authors have proved that (1.2) is true when 1 < p ≤ 2. In the paper [ACDH]
the authors have established (1.2) in the remaining range 2 ≤ p < ∞. The essential new contri-
bution of the present note is to verify that such general assumptions are verified (in a non-trivial
manner) for a general class of locally subelliptic operators satisfying on a given smooth manifold
M the generalized curvature-dimension inequality CD(ρ1, ρ2, κ, d) in (2.9) below, with curvature
parameter ρ1 ≥ 0 (this in the Riemannian case corresponds to Ric≥ 0). Once this is done, the
Lp continuity of an appropriately defined Riesz operator will follow by the general real variable
methods developed in [CD], [ACDH].
To state the main result in this paper we assume that M be a C ∞ connected, non-compact
manifold endowed with a smooth measure µ. Throughout the paper, the notation Lp(M),
1 ≤ p ≤ ∞, indicates the space of p-summable functions on M with respect to the measure µ.
We assume that on M a second-order diffusion operator L with real coefficients is given. We also
suppose that L be locally subelliptic, non-positive, and that it satisfy the assumptions listed in
Section 2. There is a natural notion of (square of the length of the)"gradient" associated with
RIESZ TRANSFORMS
3
L, namely
Γ(f ) =
1
2{L(f 2) − 2f Lf},
and a canonical distance d, see (2.4) below. We assume throughout that the metric space (M, d)
be complete. We also suppose that M be endowed with another bilinear differential form ΓZ ,
see (2.6) below, and that Γ and ΓZ satisfy all the hypothesis in Section 2 below. From our
perspective, the most significant assumption is the so-called generalized curvature-dimension
inequality CD(ρ1, ρ2, κ, d) in (2.9) below, which we now recall for the reader's convenience:
There exist constants ρ1 ≥ 0, ρ2 > 0, κ ≥ 0, and d ≥ 2 such that the inequality
(1.3)
Γ2(f ) + νΓZ
κ
(Lf )2 +(cid:16)ρ1 −
ν(cid:17) Γ(f ) + ρ2ΓZ(f )
2 (f ) ≥
1
d
hold for every f ∈ C ∞(M) and every ν > 0, where Γ2 and ΓZ
below.
2 are defined by (2.7) and (2.8)
The assumption (1.3) constitutes a sub-Riemannian generalization of the classical curvature-
diemension inequality CD(ρ, n)
Γ2(f ) ≥
1
n
(∆f )2 + ρΓ(f ),
which, as a consequence of the well-known Bochner's identity, is known to hold on any n-
dimensional Riemannian manifold satisfying Ric≥ ρ.
The parameter ρ1 in CD(ρ1, ρ2, κ, d) has the meaning of a lower bound on a sub-Riemannian Ricci
tensor, see [BG1] for extensive details. Throughout the present paper the assumption ρ1 ≥ 0
will be in force. The semigroup Pt = etL is a strongly continuous semigroup of contraction
operators on Lp(M) for 1 ≤ p ≤ ∞. We denote by p(x, y, t) = p(y, x, t) the positive heat kernel
on M associated with the semigroup Pt. Given f ∈ C ∞
0 (M), the function
u(x, t) = Ptf (x) =ZM
f (y)p(x, y, t)dµ(y),
is a solution of the equation Lu − ut = 0 in M × (0,∞), corresponding to the initial datum
u(x, 0) = f (x), x ∈ M.
We now recall that, in the general framework described above, in the paper [BG1] we proved a
generalized Li-Yau type inequality for solutions of the heat equation on M of the form u = Ptf .
From such inequality, we were able to derive several basic facts, among which the following
off-diagonal Gaussian upper bound: For any 0 < ε < 1 there exists a constant C(ρ2, κ, d, ε) > 0,
which tends to ∞ as ε → 0+, such that for every x, y ∈ M and t > 0 one has
(1.4)
C(d, κ, ρ2, ε)
d(x, y)2
p(x, y, t) ≤
V (x,√t)
1
2 V (y,√t)
1
2
exp(cid:18)−
(4 + ε)t(cid:19) .
Hereafter in this paper we adopt the notation
V (x, r) = µ(B(x, r)),
where for x ∈ M and r > 0 we have let B(x, r) = {y ∈ M d(y, x) < r}.
In the paper [BBG] we further developed the program initiated in [BG1] and were able to obtain
the following basic result: There exists a constant Cd > 0 depending only on ρ1, ρ2, κ, d, such
that for every x ∈ M and r > 0 one has
(1.5)
V (x, 2r) ≤ CdV (x, r).
4
FABRICE BAUDOIN AND NICOLA GAROFALO
For a purely analytical proof of (1.5) in the Riemannian setting we refer the reader to the paper
[BG2].
Now in their work [CD] the authors proved that the two results (1.4) and (1.5) are enough to
establish the weak-(1, 1) continuity of the Riesz transforms for the space of homogeneous type
(M, d). Since from integration by parts, and from the identity (Lf, f ) = k(−L)1/2fk2 the strong
L2 continuity of the Riesz transform
kqΓ((−L)1/2)f )k2 = kfk2
trivially follows, by the Marcinckiewicz interpolation theorem we thus obtain the following result.
Theorem 1.1. Let 1 < p ≤ 2. There is a constant Cp > 0 such that for every f ∈ C ∞
(1.6)
0 (M),
qΓ((−L)−1/2f )(cid:13)(cid:13)(cid:13)(cid:13)Lp(M) ≤ CpkfkLp(M).
(cid:13)(cid:13)(cid:13)(cid:13)
Theorem 1.1 provides the Lp continuity of the absolute value of the Riesz operator
T f =qΓ((−L)−1/2f ),
within the range 1 < p ≤ 2. We emphasize is that T is a sublinear operator. For the remaining
range 2 ≤ p < ∞ we appeal to the real variable theory developed in the work [ACDH]. We
recall the salient ingredients of the general approach in that paper:
1) etL1 = 1 (stochastic completeness);
2) global doubling condition;
3) global Poincar´e inequality;
4) Caccioppoli type inequalities;
5) Gaffney type estimates;
6) bounds for √tpΓ(etL).
As for 1) the stochastic completeness in our framework follows as a special case of Theorem 3.5
in [BG1], see also [Mu] for an extension of such result. Regarding 2) we have already discussed
(1.5). As for 3), we mention that in [BBG] it was proved that there exists Cp > 0, depending
only on ρ1, ρ2, κ, d, such that for every x ∈ M and r > 0 one has
(1.7)
Γ(f )dµ,
ZB(x,r) f − fB2dµ ≤ Cpr2ZB(x,r)
for every f ∈ C 1(B(x, r)). Thus 3. is available to us.
We are thus missing ingredients 4), 5) and 6) In this note we establish these results, see Corollary
3.5, Lemmas 3.6, 3.7 and 3.10, and Theorem 4.1 below. This allows us to close the circle and,
by using the work [ACDH], obtain the following result.
Theorem 1.2. Let 2 ≤ p < ∞. There is a constant Cp > 0 such that for every f ∈ C ∞
(1.8)
0 (M),
By combining Theorems 1.1 and 1.2 we obtain the following result.
qΓ((−L)−1/2f )(cid:13)(cid:13)(cid:13)(cid:13)Lp(M) ≤ CpkfkLp(M).
(cid:13)(cid:13)(cid:13)(cid:13)
Theorem 1.3. Let 1 < p < ∞. There exist constants Ap, Bp > 0 such that
f ∈ C ∞
(1.9)
Apk(−L)1/2fkp ≤ kpΓ(f )kp ≤ Bpk(−L)1/2fkp,
0 (M),
RIESZ TRANSFORMS
5
The results in this paper establish the continuity of the Riesz transform and the equivalence
of the Sobolev spaces defined by completion of C ∞
0 (M) with respect to the two seminorms in
(3.2) for the various classes of sub-Riemannian manifolds which are encompassed by the general
framework of [BG1]. While we refer the reader to that source for a detailed discussion of the
examples, here we confine ourselves to mention the following basic result which is a corollary of
our work.
Theorem 1.4. Let (M, θ) be a CR manifold with real dimension 2n + 1 and vanishing Tanaka-
Webster torsion, i.e., a Sasakian manifold. If there exists ρ1 ≥ 0 such that for every x ∈ M the
Tanaka-Webster Ricci tensor satisfies the bound
for every horizontal vector v ∈ Hx, then given any 1 < p < ∞ the Riesz transform associated
with a sub-Laplacian on M is continuous on Lp(M).
Ricx(v, v) ≥ ρ1v2,
In connection with Theorem 1.4 we mention that it was proved in [BG1] that in the framework
of Theorem 1.4 the generalized curvature-dimension inequality CD(ρ1, ρ2, κ, d) does hold with
ρ1 ≥ 0. Thus these manifolds fall within the scope of the assumptions in Section 2.
In closing, we mention some known partial results related to those in the present paper. In [LV]
the boundedness of the Riesz transforms was proved on every stratified nilpotent Lie group. In
[A] this result was generalized to Lie groups of polynomial growth.
Acknowledgment: The second named author would like to thank Steve Hofmann for several
helpful discussions.
2. Background
2.1. Assumptions. Hereafter in this paper, M will be a C ∞ connected and non-compact mani-
fold endowed with a smooth measure µ. Throughout the paper, the notation Lp(M), 1 ≤ p ≤ ∞,
indicates the space of p-summable functions on M with respect to the measure µ.
We assume that on M a second-order diffusion operator L with real coefficients is given. We also
suppose that L be locally subelliptic (for the relevant definition and properties of such operators
see [FSC] and [JSC]), and that it satisfy:
1) L1 = 0;
2) RM f Lgdµ =RM gLf dµ;
3) RM f Lf dµ ≤ 0,
for every f, g ∈ C ∞
There is a natural gradient (or rather, a natural square of the length of a gradient) canonically
associated with L, and it is given by the quadratic functional Γ(f ) = Γ(f, f ), where
0 (M).
(2.1)
Γ(f, g) =
1
2
(L(f g) − f Lg − gLf ),
f, g ∈ C ∞(M).
The functional Γ(f ) is known as le carr´e du champ. Notice that Γ(1) = 0. Furthermore, using
the results in [PS], locally in the neighborhood of every point x ∈ M we can write
(2.2)
X ∗
i Xi,
L = −
m
Xi=1
where the vector fields Xi are Lipschitz continuous (such representation is not unique, but this
fact is of no consequence for us. We note for further reference that the number m of vector fields
entering in the local representation (2.2) is bounded above by the dimension of the manifold
6
FABRICE BAUDOIN AND NICOLA GAROFALO
M). Therefore, for any x ∈ M there exists an open neighborhood Ux such that in Ux we have
for any f ∈ C ∞(M)
(2.3)
Γ(f ) =
(Xif )2.
m
Xi=1
This shows that Γ(f ) ≥ 0 and it actually only involves differentiation of order one. Furthermore,
the value of Γ(f )(x) does not depend on the particular representation (2.2) of L. With the
operator L we can also associate a canonical distance:
d(x, y) = sup{f (x) − f (y) f ∈ C ∞(M),kΓ(f )k∞ ≤ 1} ,
(2.4)
where for a function g on M we have let g∞ = ess sup
A tangent vector v ∈ TxM is called subunit for L at x if v = Pm
i ≤ 1,
see [FP]. It turns out that the notion of subunit vector for L at x does not depend on the local
representation (2.2) of L. A Lipschitz path γ : [0, T ] → M is called subunit for L if γ′(t) is
subunit for L at γ(t) for a.e. t ∈ [0, T ]. We then define the subunit length of γ as ℓs(γ) = T .
Given x, y ∈ M, we indicate with
i=1 aiXi(x), with Pm
x, y ∈ M,
M g.
i=1 a2
S(x, y) = {γ : [0, T ] → M γ is subunit for L, γ(0) = x, γ(T ) = y}.
In this paper we assume that
Under such assumption it is easy to verify that
S(x, y) 6= ∅,
for every x, y ∈ M.
(2.5)
ds(x, y) = inf{ℓs(γ) γ ∈ S(x, y)},
defines a true distance on M. Furthermore, thanks to Lemma 5.43 in [CKS] we know that
hence we can work indifferently with either one of the distances d or ds.
d(x, y) = ds(x, y),
x, y ∈ M,
In addition to the differential form (2.1), we assume that M be endowed with another smooth
bilinear differential form, indicated with ΓZ , satisfying for f, g ∈ C ∞(M)
(2.6)
and ΓZ (f ) = ΓZ(f, f ) ≥ 0. Given the first-order bilinear forms Γ and ΓZ on M, we now introduce
the following second-order differential forms:
ΓZ(f g, h) = f ΓZ(g, h) + gΓZ (f, h),
1
(2.7)
(2.8)
Γ2(f, g) =
2(cid:2)LΓ(f, g) − Γ(f, Lg) − Γ(g, Lf )(cid:3),
2(cid:2)LΓZ(f, g) − ΓZ (f, Lg) − ΓZ(g, Lf )(cid:3).
2 ≡ 0 as well. As for Γ and ΓZ , we will use the notations
2 (f, g) =
Observe that if ΓZ ≡ 0, then ΓZ
Γ2(f ) = Γ2(f, f ), ΓZ
2 (f, f ).
We make the following assumptions that will be in force throughout the paper:
2 (f ) = ΓZ
ΓZ
1
(H.1) There exists an increasing sequence hk ∈ C ∞
0 (M) such that hk ր 1 on M, and
Γ(hk)∞ + ΓZ (hk)∞ → 0, as k → ∞.
(H.2) For any f ∈ C ∞(M) one has
Γ(f, ΓZ(f )) = ΓZ(f, Γ(f )).
RIESZ TRANSFORMS
7
(H.3) The generalized curvature-dimension inequality CD (ρ1, ρ2, κ, d) be satisfied with ρ1 ≥ 0,
that is: There exist constants ρ1 ≥ 0, ρ2 > 0, κ ≥ 0, and d ≥ 2 such that the inequality
(2.9)
Γ2(f ) + νΓZ
2 (f ) ≥
1
d
κ
(Lf )2 +(cid:16)ρ1 −
ν(cid:17) Γ(f ) + ρ2ΓZ(f )
hold for every f ∈ C ∞(M) and every ν > 0, where Γ2 and ΓZ
2 are defined by 2.7 and 2.8.
For example, the assumptions (H.1)-(H.3) are satisfied in all Carnot groups of step two, and in all
complete Sasakian manifolds whose horizontal Tanaka-Webster Ricci curvature is non negative.
For further examples, including a wide class of bundles over Riemannian manifolds we refer the
reader to [BG1].
In this framework:
• L is essentially self-adjoint on C ∞
0 (M), so that by using the spectral theorem for the
Friedrichs extension of L in the Hilbert space L2(M), we may construct a strongly
continuous contraction semigroup (Pt)t≥0 in L2(M) whose infinitesimal generator is L;
• By hypoellipticity of L, (Pt)t≥0 admits a heat kernel, that is: There is a smooth function
p(t, x, y), t ∈ (0,∞), x, y ∈ M, such that for every f ∈ L2(M) and x ∈ M ,
Ptf (x) =ZM
p(t, x, y)f (y)dµ(y).
Moreover, the heat kernel satisfies the two following conditions:
(i) (Symmetry) p(t, x, y) = p(t, y, x);
(ii) (Chapman-Kolmogorov relation) p(t + s, x, y) =RM p(t, x, z)p(s, z, y)dµ(z);
• The semigroup (Pt)t≥0 is a sub-Markovian semigroup: If 0 ≤ f ≤ 1 is a function in
L2(M), then 0 ≤ Ptf ≤ 1;
• By the Riesz-Thorin interpolation theorem, (Pt)t≥0 defines a contraction semigroup on
Lp(M), 1 ≤ p ≤ ∞.
2.2. Some known results. In this section we collect some results that, under the above listed
assumptions, were proved in the works [BG1] and [BBG]. Such results constitute the backbone
of the present paper.
The first basic result is that the manifold M is stochastically complete with respect to L, i.e.,
for every t > 0
(2.10)
Pt1 = etL1 = 1.
We recall that this result is equivalent to the uniqueness of the bounded solution of the Cauchy
problem
(Lu − ut = 0, M × (0,∞),
x ∈ M,
u(x, 0) = ϕ(x),
with bounded initial datum ϕ. The property (2.10) is a corollary of Theorem 3.5 in [BG1], see
also [Mu] for an extension of such result.
Another basic result is the following Gaussian upper bound that was proved in [BG1].
Theorem 2.1. For any 0 < ε < 1 there exists a constant C(ε) = C(d, κ, ρ2, ε) > 0, which tends
to ∞ as ε → 0+, such that for every x, y ∈ M and t > 0 one has
d(x, y)2
(4 + ε)t(cid:19) .
exp(cid:18)−
C(ε)
V (x,√t)
p(x, y, t) ≤
In the paper [BBG] it has been proved that the metric measure space (M, d, µ) satisfies the
global volume doubling property and that, furthermore, the L2 Poincar´e inequality is satisfied
on balls. More precisely, we have the following result.
8
FABRICE BAUDOIN AND NICOLA GAROFALO
Theorem 2.2. There exist constants Cd, Cp > 0, depending only on ρ1, ρ2, κ, d, for which one
has for every x ∈ M and every r > 0:
(2.11)
(2.12)
V (x, 2r) ≤ Cd V (x, r);
ZB(x,r) f − fB2dµ ≤ Cpr2ZB(x,r)
Γ(f )dµ,
for every f ∈ C 1(B(x, r)).
We list for future use the following well-known consequence of the doubling condition (2.11).
Corollary 2.3. With Cd as in (2.11) define
Q = log2 Cd.
Then, for every x ∈ M and any 0 < r < R < ∞ one has
r(cid:19)Q
V (x, R) ≤ Cd(cid:18) R
(2.13)
+ 1(cid:19)Q
In particular, if y, z ∈ M and t > 0 we have
(2.14)
V (y,√t) ≤ Cd(cid:18) d(y, z)
√t
V (x, r).
V (z,√t),
and also for any given α > 0, there exists a constant C > 0 depending on Cd and α, such that
(2.15)
ZM
exp(cid:18)−α
d(y, z)2
t
(cid:19) dµ(z) ≤ CV (y,√t).
Proof. The proof of (2.13) is standard and we omit it. As for (2.14) it is enough to apply (2.13)
with R = d(y, z) + √t and r = √t, to obtain
V (y,√t) ≤ V (z, d(y, z) + √t) ≤ Cd(cid:18) d(y, z) + √t
√t
(cid:19)
Q
V (z,√t).
Finally, (2.15) easily follows by covering M with dyadic rings B(y, 2k+1√t) \ B(y, 2k√t), and
(cid:3)
then using (2.11).
Our objective in this section is proving that the Riesz transform associated to L is bounded in
Lp(M). 1
3. Riesz transform
Theorem 3.1. Let for every 1 < p < ∞. There is a constant Cp > 0 such that for every
f ∈ Lp(M),
(3.1)
As a consequence of Theorem 3.1, we obtain the following result.
(cid:13)(cid:13)(cid:13)(cid:13)
qΓ((−L)−1/2f )(cid:13)(cid:13)(cid:13)(cid:13)Lp(M) ≤ CpkfkLp(M).
Apk(−L)1/2fkp ≤ kpΓ(f )kp ≤ Bpk(−L)1/2fkp,
Theorem 3.2. Let 1 < p < ∞. There exist constants Ap, Bp > 0 such that
f ∈ C ∞
(3.2)
1In the case where µ(M) < ∞ one has to consider the space Lp
[ACDH]; this modification will be implicit in the text.
0 (M).
0(M) of functions in Lp(M) with mean 0, see
RIESZ TRANSFORMS
9
3.1. The case 1 < p ≤ 2. Following our discussion in the introduction, the boundedness of
the Riesz transform on Lp(M) for 1 < p ≤ 2 follows by combining Theorem 4.1 and (2.11)
in Theorem 2.2 above with Theorem 1.1 in [CD]. We note explicitly that (2.11) implies that
(M, d, µ) is a space of homogeneous type according to [CW], see also [C], and so all tools of real
analysis are available. From these results the weak-(1, 1) continuity of the Riesz transform
µ({x ∈ M qΓ((−L)−1/2f )(x)) > λ}) ≤
C
λ kfkL1(M), λ > 0,
can be established as in [CD]. Then, for the range 1 < p ≤ 2 the inequality (3.1) follows by
applying Marcinckiewicz real interpolation theorem. The reader should notice that the operator
T =qΓ((−L)−1/2f ))
is a sublinear operator, i.e., T (f + g)(x) ≤ T f (x) + T g(x) for every f, g and a.e. x ∈ M.
3.2. The case p > 2. Following the general method in the proof of Theorem 3.1 in [ACDH], to
establish the boundedness of the Riesz transform when p > 2, we need to the ingredients listed
as 1)-6) in the introduction. As it was mentioned there the items which are at this point missing
are Caccioppoli and Gaffney type estimates, as well as bounds for √tpΓ(etL). This section is
devoted to filling this gap. We begin with establishing the former type of result.
3.2.1. Caccioppoli type estimates. In what follows we prove an a priori inequality of Caccioppoli
type for the heat semigroup Pt.
Proposition 3.3. Let f ∈ C ∞
0 (M). For t ≥ 0 we have
1 + 2κ
ρ2
Γ(Ptf ) + ρ2tΓZ (Ptf ) ≤
(Pt(f 2) − (Ptf )2).
2t
Proof. Let us fix T > 0. Given a function f ∈ C0(M), for 0 ≤ t ≤ T we introduce the functionals
φ1(x, t) = Γ(PT −tf )(x),
φ2(x, t) = ΓZ (PT −tf )(x),
which are defined on M × [0, T ]. It is is easy to check that, with Γ2 and ΓZ
and (2.8), we have
2 defined as in (2.7)
and
Lφ1 +
Lφ2 +
∂φ1
∂t
∂φ2
∂t
= 2Γ2(PT −tf ).
= 2ΓZ
2 (PT −tf ),
see also [BG1]. Consider now the function
φ(x, t) = a(t)φ1(x, t) + b(t)φ2(x, t)
= a(t)Γ(PT −tf )(x) + b(t)ΓZ (PT −tf )(x),
where a and b are two nonnegative functions that will be chosen later. At this point we observe
that, since by hypothesis ρ1 ≥ 0, the generalized curvature-dimension CD(ρ1, ρ2, κ, d) in (2.9)
trivially implies the generalized curvature-dimension inequality CD(0, ρ2, κ,∞). Applying the
latter with the choice ν = b
a we thus obtain
Lφ +
∂φ
∂t
= a′Γ(PT −tf ) + b′(PT −tf )ΓZ(PT −tf ) + 2aΓ2(PT −tf ) + 2b(PT −tf )ΓZ
2 (PT −tf )
≥(cid:18)a′ − 2κ
a2
b (cid:19) Γ(PT −tf ) + (b′ + 2ρ2a)ΓZ (PT −tf ).
10
FABRICE BAUDOIN AND NICOLA GAROFALO
Let us now chose
a(t) =
1
ρ2
(T − t),
b(t) = (T − t)2.
With this choice we have on [0, T ],
and
We find then
Lφ +
a2
b
b′ + 2ρ2a ≡ 0,
1
ρ2 −
2(cid:19) Γ(PT −tf ).
1
ρ2 −
a′ − 2κ
∂t ≥(cid:18)−
= −
2κ
ρ2
2
2κ
ρ2
∂φ
.
From a comparison theorem for parabolic partial differential equations (see for instance p.52 in
[F] or Proposition 3.2 in [BG1]) we deduce
(3.3)
PT (φ(·, T ))(x) ≥ φ(x, 0) −(cid:18) 1
ρ2
+
2κ
ρ2
2(cid:19)Z T
0
Pt(Γ(PT −tf ))dt.
To conclude, we consider the functional
A straightforward computation shows that
Ψ(t) =
1
2
Pt(cid:0)(PT −tf )2)(cid:1) .
This gives
Ψ′(t) =
=
1
2
1
2
∂
∂t
(PT −tf )(cid:19)
Pt(cid:0)L(PT −tf )2)(cid:1) + Pt(cid:18)PT −tf
Pt(cid:18)L(PT −tf )2) − 2PT −tf L(PT −tf )(cid:19) = Pt (Γ(PT −tf )) .
2(cid:0)PT (f 2) − (PT f )2(cid:1) .
Pt(Γ(PT −tf )dt = Ψ(T ) − Ψ(0) =
1
Z T
0
Replacing this information in (3.3), along with the identities
φ(x, 0) =
1
ρ2
T Γ(PT f ) + T 2ΓZ (PT f ),
φ(x, T ) = 0,
we reach the desired conclusion.
(cid:3)
As a corollary of Proposition 3.3 and of the Lp continuity of Pt, for 1 ≤ p ≤ ∞, we obtain the
following Caccioppoli type estimates.
Corollary 3.4. For any f ∈ C ∞
0 (M), 2 ≤ p ≤ ∞ we have
kpΓ(Ptf )kp ≤s1 + 2κ
2t
ρ2
kfkp.
In what follows we will also need the following result.
Corollary 3.5. Let n be the dimension of M. With C =q (2κ+ρ2)n
2ρ2
C
√t
(3.4)
.
This estimate implies that for any f ∈ L∞(M),
kpΓ(Ptf )k∞ ≤
(3.5)
sup
x∈MZMpΓ(p(·, y, t))(x)dµ(y) ≤
C
√tkfk∞.
we have
RIESZ TRANSFORMS
11
Proof. For x ∈ M fixed, let then Ux be a sufficiently small neighborhood of x in which we can
write L as in (2.2), where the vector fields Xi are Lipschitz continuous. For f ∈ C ∞
0 (M), we
have
(cid:12)(cid:12)(cid:12)(cid:12)
ZM
Xip(·, y, t)(x)f (y)dµ(y)(cid:12)(cid:12)(cid:12)(cid:12)
Let ε > 0 and take now
= XiPtf (x) ≤ kpΓ(Ptf )k∞ ≤s 1 + 2κ
2t
ρ2
kfk∞
f (y) = hk(y)
Xip(·, y, t)(x)
Xip(·, y, t)(x) + ε
,
where 0 ≤ hk ≤ 1 is an increasing sequence in C ∞
Xip(·, y, t)(x)2
Xip(·, y, t)(x) + ε
ZM
hk(y)
dµ(y) ≤s 1 + 2κ
2t
ρ2
0 (M), converging to 1. We obtain
By the monotone convergence theorem and by Fatou's theorem we deduce, first letting k → ∞
and then ε → 0, that
ZM Xip(·, y, t)(x)dµ(y) ≤s 1 + 2κ
2t
ρ2
.
The estimate (3.4) follows immediately from the latter inequality.
(cid:3)
3.2.2. Gaffney-type estimates. We now turn to the second main ingredient which are Gaffney
type estimates. In what follows we indicate with E, F ⊂ M two closed subsets. We need the
following results.
Lemma 3.6. For every two closed sets E, F ⊂ M, and any f ∈ L∞(M) supported in E, one
has
PtfL2(F ) ≤ e− d(E,F )2
4t
fL2(E).
Proof. Let us suppose d(E, F ) > 0, otherwise the conclusion follows trivially from the L2 con-
tinuity of Pt. As a first step we let ψ denote a function in Lip(M), such that Γ(ψ) ≤ 1 a.e. on
M. With α > 0 and φ = e−αψ, we note that
(3.6)
Let f ∈ L∞(M), and consider
Γ(φ) = α2φ2Γ(ψ) ≤ α2φ2.
y(t) = φPtf2
L2(M).
Denoting with < ·,· > the inner product in L2(M), we have
y′(t) = 2 < L(Ptf ), φ2Ptf >= −2ZM
Γ(Ptf, φ2Ptf )dµ
= −2ZM
≤ −2ZM
φ2Γ(Ptf )dµ − 4ZM
εZM
φ2Γ(Ptf )dµ +
2
φPtf Γ(Ptf, φ)dµ
φ2Γ(Ptf )dµ + 2εα2ZM
φ2(Ptf )2dµ,
where to estimate the last term we have used (3.6). Choosing ε = 1 we conclude
and, upon integrating this inequality, we obtain
(3.7)
φPtfL2(M) ≤ eα2tφfL2(M).
y′(t) ≤ 2α2y(t),
12
FABRICE BAUDOIN AND NICOLA GAROFALO
We now want to show that (3.7) yields the desired conclusion. Suppose that suppf ⊂ E, then
we argue as follows. We take ψ(x) = d(x, F ), and for any α > 0 we let φ = e−αψ. Since φ ≡ 1
on F , from (3.7) we obtain
PtfL2(F ) ≤ φPtfL2(M) ≤ eα2tφfL2(M) = eα2tφfL2(E).
Now, for any x ∈ E we have ψ(x) ≥ d(E, F ), and thus φ ≤ e−αd(E,F ) on E. This gives
PtfL2(F ) ≤ eα2t−αd(E,F )fL2(E).
By choosing α = d(E,F )
2t > 0 we reach the desired conclusion.
(cid:3)
For ω > 0 sufficiently small denote
Sω = {z = reiθ ∈ C 0 < r < ∞, θ <
π
2
+ ω}.
Since L generates an analytic semigroup ezL in a sector Sω, we have
Pt =
1
2πiZΓδ
etζ R(ζ; L)dζ,
π
where Γδ ⊂ Sω is the path composed of the two rays reiθ and re−iθ, with 0 < r < ∞ and
2 < θ < π
2 + δ, 0 < δ < ω, and R(ζ; L) is the resolvent of L. Using the same argument as in
Lemma 3.6 it is easy to see that the Gaffney estimate (3.7) continues to be valid for Pzf with
z ∈ Sδ for any fixed 0 < δ < ω. Similarly to what was done above this leads to an estimate of
the type
PzfL2(F ) ≤ eα2ℜz−αd(E,F )fL2(E),
(3.8)
for any f ∈ L∞(M) which is supported in E, and any α > 0.
Lemma 3.7. There exists a constant C > 0 such that for every two closed sets E, F ⊂ M, and
any f ∈ L∞(M) supported in E, one has
z ∈ Sδ,
tLPtfL2(F ) ≤ Ce− d(E,F )2
6t
fL2(E).
Proof. Consider the path
Γt = {z = t(cid:18)1 +
eiθ
2 (cid:19) ∈ C 0 ≤ θ ≤ 2π} ⊂ Sω.
Using the analyticity of Pt in the sector Sω we can write
t
∂Pt
∂t
=
t
2πiZΓt
ezL
(z − t)2 dz =
Using (3.8), this gives
1
π Z 2π
0
t(cid:16)1+ eiθ
2 (cid:17)L
e−iθe
dθ.
tLPtfL2(F ) ≤
t(cid:16)1+ eiθ
2 (cid:17)fL2(F )
)−αd(E,F )fL2(E)
0
1
P
π Z 2π
≤ 2eα2t(1+ cos θ
≤ 2e
2
3
2 α2t−αd(E,F )fL2(E)
Choosing α = d(E,F )
3t
yields the desired conclusion.
(cid:3)
RIESZ TRANSFORMS
13
Since the distance function y → d(x, y) is obviously a Lipschitz continuous function on M (with
respect to d itself), in view of (2.2) and (2.3), and of a Rademacher type theorem for Lipschitz
vector fields we can construct Lipschitz continuous cut-off functions on metric balls, see for
instance Theorem 1.5 in [GN]. We collect this fact in the following lemma.
Lemma 3.8. Let 0 < s < t < ∞. There exists a constant C > 0 such that for every B(x, s) ⊂
B(x, t) ⊂ M there exists a function ϕ ∈ Lip(M), with 0 ≤ ϕ ≤ 1, ϕ ≡ 1 on B(x, s), and
supp ϕ ⊂ B(x, t), for which
pΓ(ϕ) ≤
C
t − s
.
Corollary 3.9. Given a closed set F ⊂ M, consider the open set Fε = {x ∈ M d(x, F ) < ε}.
There exists a function ϕ ∈ Lip(M) such that 0 ≤ ϕ ≤ 1, ϕ ≡ 1 on Fε/2, and ϕ ≡ 0 in M \ Fε,
and for which
Proof. It follows from Lemma 3.8 by a standard partition of unity argument.
C
ε
.
pΓ(ϕ) ≤
(cid:3)
We are now in a position to establish the third Gaffney type estimate which we will need.
Lemma 3.10. There exist constants C ≥ 0 and α > 0 such that for any two closed sets
E, F ⊂ M, and every function f ∈ L∞(M) supported in E, one has
kfkL2(E).
Proof. We adapt the argument on p. 930 in [ACDH]. If d(E, F ) ≤ √t there is nothing to prove.
We can thus assume that d(E, F ) > √t. With ε = d(E,F )
function ϕ ∈ Lip(M) supported in Fε as in Corollary 3.9. We have
tZM
√tkpΓ(Ptf )kL2(F ) ≤ Ce−α d(E,F )2
, consider the set Fε, and pick a
ϕ2Γ(Ptf )dµ =
t
3
t
ϕ2(cid:2)L(cid:0)(Ptf )2(cid:1) − 2Ptf L(Ptf )(cid:3) dµ
ϕ2Ptf LPtf dµ − 2tZM
2ZM
= −tZM
≤ tLPtfL2(Fε)PtfL2(M) + 2(cid:18)tZM
ϕPtf Γ(ϕ, Ptf )dµ
We now use Lemma 3.7 and the L2 continuity of Pt to obtain for some α > 0
f2
tLPtfL2(Fε)PtfL2(M) ≤ Ce− d(E,Fε)2
f2
6t
Recalling that √t < d(E, F ), and using the support property of Γ(ϕ), and the estimate Γ(ϕ) ≤
Cd(E, F )−2 from Corollary 3.9, we conclude
L2(E).
Γ(ϕ)(Ptf )2dµ(cid:19)1/2
.
ϕ2Γ(Ptf )dµ(cid:19)1/2(cid:18)tZM
L2(E) ≤ Ce−α d(E,F )2
t
(cid:18)√tZM
Γ(ϕ)(Ptf )2dµ(cid:19)1/2
(Ptf )2dµ(cid:19)1/2
≤ C(cid:18)ZFε
≤ Ce− d(E,Fε)2
4t
fL2(E) ≤ Ce−α d(E,F )2
t
for an appropriate α > 0. We conclude
tZM
ϕ2Γ(Ptf )dµ ≤ Ce−α d(E,F )2
f2
A trivial estimate allows to conclude that
t
L2(E) + Ce−α d(E,F )2
t
ϕ2Γ(Ptf )dµ ≤ Ce−α d(E,F )2
t
tZM
fL2(E)(cid:18)tZM
f2
L2(E).
fL2(E),
ϕ2Γ(Ptf )dµ(cid:19)1/2
.
14
Since
FABRICE BAUDOIN AND NICOLA GAROFALO
we have reached the desired conclusion.
tkpΓ(Ptf )k2
L2(F ) ≤ tZM
ϕ2Γ(Ptf )dµ,
(cid:3)
3.2.3. The completion of the proof Theorem 3.1 in the range 2 ≤ p < ∞. We are now in a
position to prove Theorem 3.1 in the range 2 ≤ p < ∞. As mentioned above, all we need to do
at this point is combine Theorem 2.2, Corollary 3.5 and Lemmas 3.6, 3.7 and 3.10 with the real
variable results in [ACDH].
4. Pointwise gradient estimates of the heat kernel
In this section we investigate the validity of pointwise gradient estimates of the heat kernel.
Besides being interesting in their own right, such estimates are also connected with Theorem 3.1
in the range 2 ≤ p < ∞. For a detailed discussion of this aspect we refer the reader to [ACDH].
Theorem 4.1. There exists a constant C = C(d, κ, ρ2) > 0 such that for every x, y ∈ M and
t > 0 one has
pΓ (p(·, y, t))(x) ≤
C
√tV (y,√t)
.
Proof. We fix a point x ∈ M and s > 0 and begin with the observation that if we consider the
function
fx,s(z) = p(x, z, s),
then thanks to Theorem 4.1 there exists a constant C > 0 (independent of x ∈ M and s > 0)
such that fx,s ∈ L∞(M) and
kfx,skL∞(M) ≤
V (x,√s)
.
C
We next observe that, given points x, y ∈ M and t > 0, then we can write
p(x, y, t) = P t
2
(fy, t
2
)(x).
For x ∈ M fixed, let then Ux be a sufficiently small neighborhood of x in which we can write L
as in (2.2). We thus have
2
2
(fy, t
Xip(·, y, t)(x) = Xi(cid:16)P t
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
rΓ(cid:16)P t
√tV (y,pt/2)
(fy, t
≤
C
2
2
)(cid:17) (x) ≤rΓ(cid:16)P t
)(cid:17)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L∞(M)
√t(cid:13)(cid:13)(cid:13)
≤
C
,
2
(fy, t
2
fy, t
)(cid:17)(x)
2(cid:13)(cid:13)(cid:13)L∞(M)
where in the last inequality we have used the L∞ Caccioppoli inequality in Corollary 3.5. This
estimate finally gives
pΓ (p(·, y, t))(x) =vuut
d
Xi=1
(Xip(·, y, t)(x))2 ≤
C
√tV (y,pt/2)
The proof is completed by an application of (2.11) in Theorem 2.2.
.
(cid:3)
RIESZ TRANSFORMS
15
Theorem 4.2. For every ε > 0 there exists a constant C(ε) = C(d, κ, ρ2, ε) > 0 such that for
every x, y ∈ M and t > 0 one has
pΓ (p(·, y, t))(x) ≤
C(ε)
√tV (y,√t)
exp(cid:18)−
d(x, y)2
4(1 + ε)t(cid:19) .
Proof. It suffices to appeal to Theorem 4.11 in [CS]. In such result the authors, by a beautiful
use of Phragm´en-Lindelof theory, prove that the combination of (2.13) in Corollary 2.3, and of
the estimates
and
p(x, x, t) ≤
pΓ (p(·, y, t))(x) ≤
√tV (y,√t)(cid:18)1 +
C
pΓ (p(·, y, t))(x) ≤
C
V (x,√t)
,
C
√tV (y,√t)
,
allows to improve the estimate in Theorem 4.1 into the following one
d(x, y)2
4t (cid:19)1+3Q
exp(cid:18)−
d(x, y)2
4t (cid:19) ,
where Q is as (2.13). From the latter estimate the desired conclusion follows.
(cid:3)
References
[ACDH] P. Auscher, T. Coulhon, X.T. Duong, S. Hofmann, Riesz transform on manifolds and heat kernel regu-
larity Ann. Sci. ´Ecole Norm. Sup. (4) 37 (2004), no. 6, 911-957.
[A] G. Alexopoulos, An application of homogenization theory to harmonic analysis: Harnack inequalities and
Riesz transforms on Lie groups of polynomial growth. Canad. J. Math. 44 (1992), no. 4, 691 -- 727.
[B] D. Bakry, ´Etude des transformations de Riesz dans les vari´et´es riemanniennes `a courbure de Ricci minor´ee.
(French) [A study of Riesz transforms in Riemannian manifolds with minorized Ricci curvature] S´eminaire
de Probabilit´es, XXI, 137 -- 172, Lecture Notes in Math., 1247, Springer, Berlin, 1987.
[BG1] F. Baudoin & N. Garofalo, Curvature-dimension inequalities and Ricci lower bounds for sub-Riemannian
manifolds with transverse symmetries, Arxiv preprint 1101.3590, submitted paper (2009).
[BG2]
, Perelman's entropy and doubling property on Riemannian manifolds, J. Geom. Anal., to appear.
[BBG] F. Baudoin. M. Bonnefont & N. Garofalo, A sub-Riemannian curvature-dimension inequality, volume
doubling property and the Poincar´e inequality, arXiv:1007.1600, submitted, (2010).
[CKS] E. Carlen, S. Kusuoka & D. Stroock, Upper bounds for symmetric Markov transition functions, Ann. Inst.
H. Poincar´e Probab. Statist. 23 (1987), no. 2, suppl., 245 -- 287.
[C] M. Christ, A T (b) theorem with remarks on analytic capacity and the Cauchy integral, Colloq. Math.
60/61 (1990), no. 2, 601628.
[CW] R. R. Coifman & G. Weiss, Analyse harmonique non-commutative sur certains espaces homog`enes. (French)
´Etude de certaines int´egrales singuli`eres, Lecture Notes in Mathematics, Vol. 242. Springer-Verlag, Berlin-
New York, 1971. v+160 pp.
[CD] T. Coulhon, X. T. Duong, Riesz transforms for 1 ≤ p ≤ 2, Trans. Amer. Math. Soc., 351 3 (1999),
1151-1169.
[CS] T. Coulhon, A. Sikora, Gaussian heat kernel upper bounds via the Phragmn-Lindelf theorem. Proc. Lond.
Math. Soc. (3) 96 (2008), no. 2, 507 -- 544.
[FP] C. Fefferman & D. H. Phong, Subelliptic eigenvalue problems, Conference on harmonic analysis in honor of
Antoni Zygmund, Vol. I, II (Chicago, Ill., 1981), 590 -- 606, Wadsworth Math. Ser., Wadsworth, Belmont, CA,
1983.
[FSC] C. L. Fefferman & A. S´anchez-Calle, Fundamental solutions for second order subelliptic operators, Ann. of
Math. (2) 124 (1986), no. 2, 247 -- 272.
[F] A. Friedman, Partial differential equations of parabolic type, Dover, 2008.
[GN] N. Garofalo & D.-M. Nhieu, Lipschitz continuity, global smooth approximations and extension theorems for
Sobolev functions in Carnot-Carathodory spaces, J. Anal. Math. 74 (1998), 67-97.
[JSC] D. Jerison & A. S´anchez-Calle, Subelliptic second order differential operators, Lecture. Notes in Math., 1277
(1987), pp. 46-77.
16
FABRICE BAUDOIN AND NICOLA GAROFALO
[LV] Lohou´e, N., Varopoulos, N.: Remarques sur les transform´ees de Riesz sur les groupes nilpotents. C.R.A.S.
Paris, 301, 11 (1985) 559-560.
[Mu] I. Munive, Generalized curvature-dimension inequalities, stochastic completeness and volume growth,
preprint, 2011.
[PS] R. S. Phillips & L. Sarason, Elliptic-parabolic equations of the second order, J. Math. Mech. 17 1967/1968,
891-917.
[S] E. M. Stein, Singular integrals and differentiability properties of functions, Princeton Univ. Press, 1970.
[Str] R. Strichartz, Analysis of the Laplacian on the complete Riemannian manifold, J. Funct. Anal. 52 (1983),
48-79.
Department of Mathematics, Purdue University, West Lafayette, IN 47907
E-mail address, Fabrice Baudoin: [email protected]
Department of Mathematics, Purdue University, West Lafayette, IN 47907
E-mail address, Nicola Garofalo: [email protected]
|
1708.09549 | 1 | 1708 | 2017-08-31T03:29:09 | Necessary and sufficient conditions for boundedness of commutators of bilinear Hardy-Littlewood maximal function | [
"math.FA"
] | Let $\mathcal{M}$ be the bilinear Hardy-Littlewood maximal function and $\vec{b}=(b,b)$ be a collection of locally integrable functions. In this paper, the authors establish characterizations of the weighted {\rm BMO} space in terms of several different commutators of bilinear Hardy-Littlewood maximal function, respectively; these commutators include the maximal iterated commutator $\mathcal{M}_{\Pi \vec{b}}$, the maximal linear commutator $\mathcal{M}_{\Sigma\vec{b}}$, the iterated commutator $[\Pi \vec{b},\mathcal{M}]$ and the linear commutator $[\Sigma \vec{b},\mathcal{M}]$. | math.FA | math |
Appl. Math. J. Chinese Univ.
2013, 28(*): ***-***
Necessary and sufficient conditions for boundedness of
commutators of bilinear Hardy-Littlewood maximal
function
Wang Ding-huai and Zhou Jiang∗
Abstract. Let M be the bilinear Hardy-Littlewood maximal function and ~b = (b, b) be a
collection of locally integrable functions. In this paper, the authors establish characterizations of
the weighted BMO space in terms of several different commutators of bilinear Hardy-Littlewood
maximal function, respectively; these commutators include the maximal iterated commutator
Σ~b, the iterated commutator [Π~b, M] and the linear
Π~b, the maximal linear commutator M
M
commutator [Σ~b, M].
§1
Introduction
A locally integrable function f is said to belong to BMO space if there exists a constant
C > 0 such that for any cube Q ⊂ Rn,
1
Q ZQ
f (x) − fQdx ≤ C,
where fQ = 1
Q RQ f (x)dx and the minimal constant C is defined by kf k∗.
There are a number of classical results that demonstrate BMO functions are the right
collections to do harmonic analysis on the boundedness of commutators. A well known result
of Coifman, Rochberg and Weiss [7] states that the commutator
[b, T ](f ) = bT (f ) − T (bf )
is bounded on some Lp, 1 < p < ∞, if and only if b ∈ BMO, where T be the classical Calder´on-
Zygmund operator. Chanillo [5] proved that if b ∈ BMO, the commutator
is bounded from Lp to Lq with 1 < p < n/α and 1/q = 1/p − α/n, where Iα be a fractional
[b, Iα]f (x) = b(x)Iαf (x) − Iα(bf )(x)
Received: 2016-**-**.
MR Subject Classification: 42B20, 42B25, 42B35.
Keywords: BMO function, Characterization, Commutator, Hardy-Littlewood maximal function.
Digital Object Identifier(DOI): 10.1007/s11766-013-****-*.
Supported by the National Natural Science Foundation of China (11661075)
∗Corresponding author
2
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
integral operator. Moreover, if n − α is even, the reverse is also valid. A complete characteriza-
tion of BMO via the commutator [b, Iα] was shown by Ding [8]. During the past thirty years,
the theory was then extended and generalized to several directions. For instance, Bloom [3] in-
vestigated the characterization of BMO spaces in the weighted setting. In 1991, Garc´ıa-Cuerva,
Harboure, Segovia and Torrea [11] showed that the maximal commutator
Mb(f )(x) = sup
Q∋x
1
Q ZQ
b(x) − b(y)f (y)dy
is bounded on Lp, 1 < p < ∞, if and only if b ∈ BMO. In 2000, Bastero, Milman and Ruiz [1]
studied the necessary and sufficient conditions for the boundedness of [b, M ] on Lp spaces when
1 < p < ∞. They showed that the commutator of Hardy-Littlewood maximal operator
[b, M ](f )(x) = b(x)M (f )(x) − M (bf )(x)
is bounded on Lp, 1 < p < ∞, if and only if b ∈ BMO with b− ∈ L∞, where b−(x) =
− min{b(x), 0}. In 2014, Zhang [24] considered the characterization of BMO via the commutator
of the fractional maximal function on variable exponent Lebesgue spaces.
In the multilinear setting, the boundedness of commutators has been extensively studied
already, as in P´erez and Torres' [16], Tangs [19], Lerner, Ombrosi, P´erez, Torres, and Trujillo-
Gonz´alezs [12] and Chen and Xues [6], and P´erez, Pradolini, Torres, and Trujillo-Gonz´alezs [15].
Specially, Chaffee and Torres [4], Wang, Pan and Jiang [20] and Zhang [23] contributed the
theory of characterization of BMO spaces by considering the linear commutator of Multilinear
operators, respectively. In this paper, we will extend Zhang's result to weighted case and we
replace the linear commutators by iterated commutators.
Our main results as follows.
Theorem 1.1. Let 1 < p1, p2 < ∞,~b = (b, b), 1/p = 1/p1 + 1/p2 and ω ∈ A1. Then the
following are equivalent,
(A1) b ∈ BMO(ω);
(A2) MΣ~b is bounded from Lp1 (ω) × Lp2 (ω) to Lp(ω1−p);
(A3) MΠ~b is bounded from Lp1(ω) × Lp2(ω) to Lp(ω1−2p).
Theorem 1.2. Let 1 < p1, p2 < ∞,~b = (b, b), 1/p = 1/p1 + 1/p2 and ω ∈ A1. Then the
following are equivalent,
(B1) b ∈ BMO(ω) and b−/ω ∈ L∞;
(B2) [Σ~b, M] is bounded from Lp1 (ω) × Lp2 (ω) to Lp(ω1−p);
(B3) [Π~b, M] is bounded from Lp1(ω) × Lp2(ω) to Lp(ω1−2p).
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
3
§2 Some preliminaries and notations
In 2009, Lerner, Ombrosi, P´erez, Torres and Trujillo-Gonz´alez [12] introduced the following
multilinear maximal function that adapts to the multilinear Calder´on-Zygmund theory. In this
paper, we only consider the bilinear case. A similar argument also works for the multilinear
cases.
Definition 2.1. For a collection of locally integrable functions ~f = (f1, f2), the bilinear maxi-
mal function M is defined by
M( ~f )(x) = sup
Q∋x
2
Yi=1
1
Q ZQ
fi(yi)dyi.
We now give the definitions of the maximal commutators and the commutators related to
the bilinear maximal function M.
Definition 2.2. For two collections of locally integrable functions ~f = (f1, f2) and ~b = (b1, b2),
the maximal linear commutator MΣ~b is defined by
Xi=1
MΣ~b( ~f )(x) =
( ~f )(x),
2
M(i)
bi
where
M(i)
bi
( ~f )(x) = sup
Q∋x
1
Q2 ZQZQ
bi(x) − bi(yi)
2
Yj=1
fj(yj)dy1dy2.
The maximal iterated commutator MΠ~b is defined by
MΠ~b( ~f )(x) = sup
Q∋x
1
Q2 ZQZQ
2
Yi=1
The linear commutator of M is defined by
bi(x) − bi(yi)fi(yi)dy1dy2.
[Σ~b, M]( ~f )(x) = [b1, M](1)( ~f )(x) + [b2, M](2)( ~f )(x),
[b1, M](1)( ~f )(x) = b1(x)M( ~f )(x) − M(b1f1, f2)(x)
where
and
[b2, M](2)( ~f )(x) = b2(x)M( ~f )(x) − M(f1, b2f2)(x).
The iterated commutator of M is defined by
[Π~b, M]( ~f )(x) = b1(x)b2(x)M( ~f )(x) − b1(x)M(f1, b2f2)(x)
−b2(x)M(b1f1, f2)(x) + M(b1f1, b2f2)(x).
We now recall the definition of Ap weight introduced by Muckenhoupt [13].
Definition 2.3. For 1 < p < ∞ and a nonnegative locally integrable function ω on Rn, ω is in
the Muckenhoupt Ap class if it satisfies the condition
sup
Q (cid:18) 1
Q ZQ
ω(x)dx(cid:19)(cid:18) 1
Q ZQ
ω(x)− 1
p−1 dx(cid:19)p−1
< ∞.
4
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
And a weight function ω belongs to the class A1 if there exists C > 0 such that for every cube
Q,
1
Q ZQ
ω(x)dx ≤ C ess inf
x∈Q
ω(x).
We write A∞ = S1≤p<∞ Ap.
Definition 2.4. Let 1 ≤ p < ∞. Given a a nonnegative locally integrable function ω, the
weighted BMO space BMOp(ω) is defined be the set of all functions f ∈ L1
loc(Rn) such that
kf kBMOp(w) := sup
Q (cid:18) 1
w(Q) ZQ
f (y) − fQpω(y)1−pdy(cid:19)1/p
< ∞,
BMO1(ω) = BMO(ω) simple.
where the supremum is taken over all cubes Q ⊂ Rn and ω(Q) = RQ ω(x)dx. We write
Remark For 1 ≤ p < ∞ and ω ∈ A1, Garc´ıa-Cuerva [10] proved that BMO(ω) = BMOp(ω)
with equivalence of the corresponding norms.
Standard real analysis tools as the weighted maximal function Mω(f ), the sharp maximal
function M ♯(f ) carries over to this context, namely,
Mω(f )(x) = sup
Q∋x
1
ω(Q) ZQ
f (y)ω(y)dy;
M ♯(f )(x) = sup
Q∋x
inf
c
1
Q ZQ
f (y) − cdy ≈ sup
Q∋x
1
Q ZQ
f (y) − fQdy.
A variant of weighted maximal function and sharp maximal operator Mω,s(f )(x) = (cid:0)Mω(f s)(cid:1)1/s
, which will become the main tool in our scheme.
and M ♯
δ (f )(x) = (cid:0)M ♯(f δ)(x)(cid:1)1/δ
§3 Main lemmas
To prove Theorem 1.1 and Theorem 1.2, we need the following results.
Lemma 3.1. Let ω ∈ A1, ~b = (b, b) and b ∈ BMO(ω). Then
M ♯
1
3(cid:0)MΠ~b( ~f )(cid:1)(x) . kbk2
BMO(ω)ω(x)2M (M( ~f )(x))
+kbk2
BMO(ω)ω(x)2
2
Yi=1
Mω,s(fi)(x)
+
2
Xi=1
kbkBMO(ω)ω(x)M 1
2
(M(i)
b ( ~f ))(x),
for any 1 < s < ∞ and bounded compact supported functions f1, f2.
Proof. First of all, we give the definition of the following auxiliary maximal function, which has
been studied in [17] and [18] for the linear case. Let ϕ(x) ≥ 0 be a smooth function such that
ϕǫ(t) = ǫ−2nϕ( t
ǫ ), ϕ′(t) . t−1 and χ[0,1](t) ≤ ϕ(t) ≤ χ[0,2](t).
Let
Φ(f1, f2)(x) = sup
ǫ>0ZRn ZRn
ϕǫ(x − y1 + x − y2)
2
Yi=1
fi(yi)dy1dy2,
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
5
and
ΦΠ~b(f1, f2)(x) = sup
ǫ>0 ZRn ZRn
ϕǫ(x − y1 + x − y2)
2
Yi=1
b(x) − b(yi)fi(yi)dy1dy2.
We first show that
Φ(f1, f2)(x) ≈ M(f1, f2)(x).
In fact, let Bǫ = {y ∈ Rn : x − y ≤ ǫ}. It is easy to see that
The bounded compact supported condition of ϕ gives
B ǫ
2 × B ǫ
2 ⊂ (cid:8)(y1, y2) : x − y1 + x − y2 ≤ ǫ(cid:9) ⊂ Bǫ × Bǫ.
ǫ>0ZRn ZRn
ϕǫ(x − y1 + x − y2)f1(y1)f2(y2)dy1dy2
Φ(f1, f2)(x) = sup
≤ sup
ǫ>0
1
ǫ2n ZBǫ ZBǫ
ϕ(cid:16) x − y1 + x − y2
ǫ
(cid:17)f1(y1)f2(y2)dy1dy2
. M(f1, f2)(x)
and
Φ(f1, f2)(x) ≥ sup
ǫ>0
1
ǫ2n ZB ǫ
2
ZB ǫ
2
ϕ(cid:16) x − y1 + x − y2
ǫ
(cid:17)f1(y1)f2(y2)dy1dy2
& M(f1, f2)(x).
We can also obtain that ΦΠ~b(f1, f2)(x) ≈ MΠ~b(f1, f2)(x).
Now, we shall estimate the sharp maximal function of the auxiliary maximal function. Let
Q be a cube and x ∈ Q. Then, for any z ∈ Q we have
(cid:12)(cid:12)ΦΠ~b(f1, f2)(z) − cQ(cid:12)(cid:12) ≤ (cid:12)(cid:12)b(z) − bQ2Φ(f1, f2)(z)
+(cid:12)(cid:12)b(z) − bQΦ(f1, (b − bQ)f2)(z)(cid:12)(cid:12)
+(cid:12)(cid:12)b(z) − bQΦ((b − bQ)f1, f2)(z)(cid:12)(cid:12)
+(cid:12)(cid:12)Φ((b − bQ)f1, (b − bQ)f2)(z) − cQ(cid:12)(cid:12)
1 (z) + AQ
2 (z) + AQ
3 (z) + AQ
i will be defined later.
4 (z),
ΦΠ~b(f1, f2)(z) − cQδ(cid:12)(cid:12)(cid:12)
(cid:19)1/δ
. (cid:18) 1
Q ZQ(cid:12)(cid:12)(cid:12)
Xj=1
Aj,
4
.
1 , (b − bQ)f ∞
Therefore,
where cQ = (cid:0)Φ((b − bQ)f ∞
(cid:18) 1
Q ZQ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ΦΠ~b(f1, f2)(z)(cid:12)(cid:12)
δ
− cQδ(cid:12)(cid:12)(cid:12)
=: AQ
2 )(cid:1)Q and f ∞
dz(cid:19)1/δ
where Aj = (cid:0) 1
Q RQ(cid:0)AQ
Let us consider first the term A1. By averaging AQ
, j = 1, 2, 3, 4 and taking δ = 1/3.
j (z)(cid:1)δ
dz(cid:1)1/δ
A1 = (cid:18) 1
Q ZQ(cid:16)(cid:12)(cid:12)b(z) − bQ2Φ(f1, f2)(z)(cid:17)1/3
1 over Q, we get
dz(cid:19)3
. kbk2
BMO(ω)
ω(Q)2
Q2
·
1
Q ZQ
M(f1, f2)(z)dz
. kbk2
BMO(ω)ω(x)2M (M(f1, f2))(x).
6
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
Let us consider next the term A2. We write
AQ
2 (z) = (cid:12)(cid:12)b(z) − bQΦ(f1, b − bQf2)(z)
≤ (cid:12)(cid:12)b(z) − bQΦ(f1,(cid:0)b(z) − bQ + b(z) − b(cid:1)f2)(z)
≤ (cid:12)(cid:12)b(z) − bQ2M(f1, f2)(z) +(cid:12)(cid:12)b(z) − bQM(2)
21(z) + AQ
=: AQ
22(z).
b (f1, f2)(z)
For AQ
21(z), the fact that Φ(f1, f2)(z) . M(f1, f2)(z) gives
A21
Q ZQ(cid:0)AQ
21(z)(cid:1)δ
:= (cid:18) 1
. (cid:16) ω(Q)
. kbk2
kbkBMO(ω)(cid:17)2 1
Q
BMO(ω)ω(x)2M (M(f1, f2))(x).
M(f1, f2)(z)dz
dz(cid:19)1/δ
Q ZQ
For AQ
22(z),
A22
:= (cid:18) 1
Q ZQ(cid:0)AQ
. ω(x)kbkBMO(ω)
dz(cid:19)1/δ
22(z)(cid:1)δ
Q2(cid:18)ZQ(cid:12)(cid:12)M(2)
1
. ω(x)kbkBMO(ω)M1/2(M(2)
b (f1, f2))(x).
1/2(cid:19)2
b (f1, f2)(z)(cid:12)(cid:12)
The same process also follows that
A3 . kbk2
BMO(ω)ω(x)2M (M(f1, f2))(x) + ω(x)kbkBMO(ω)M1/2(M(1)
b (f1, f2))(x).
To estimate A4, we split fj to fj = f 0
j + f ∞
j = fjχ2Q. We write
j with f 0
1 , (b − bQ)f 0
AQ
4 ≤ (cid:12)(cid:12)Φ((b − bQ)f 0
+(cid:12)(cid:12)Φ(((b − bQ)f 0
+(cid:12)(cid:12)Φ((b − bQ)f ∞
+(cid:12)(cid:12)Φ((b − bQ)f ∞
41(z) + AQ
=: AQ
1 , (b − bQ)f ∞
1 , (b − bQ)f 0
1 , (b − bQ)f ∞
2 )(z)(cid:12)(cid:12)
2 )(z)(cid:12)(cid:12)
2 )(z)(cid:12)(cid:12)
2 )(z) − cQ(cid:12)(cid:12)
44(z).
42(z) + AQ
43(z) + AQ
Then
A4 ≤ (cid:18) 1
4
AQ
Q ZQ(cid:0)
Xj=1
(cid:18) 1
Q ZQ(cid:0)AQ
Xj=1
4j (z)(cid:1)δ
4j(z)(cid:1)δ
4
dz(cid:19)1/δ
dz(cid:19)1/δ
.
.
4
Xj=1
A4j .
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
7
By Kolmogorov inequality and the fact that M is bounded from L1 × L1 to L1/2,∞, we have
1 , (b − bQ)f 0
2 )kL1/2,∞
1 , (b − bQ)f 0
2 )kL1/2,∞
b(yi) − bQfi(yi)dyi
C
Q2 kΦ((b − bQ)f 0
C
Q2 kM((b − bQ)f 0
C
Q2
2
Z2Q
(cid:18)Z2Q
Yi=1
Yi=1
2
C
Q2
A41 ≤
≤
≤
≤
.
2
Yi=1
kbkBMOs′ (ω)ω(x)Mω,s(fi)(x).
b(yi) − bQs′
ω1−s′
(yi)dyi(cid:19)1/s′
(cid:18)Z2Q
fi(yi)sω(yi)dyi(cid:19)1/s
For A42, it is easy to see that
ϕǫ(z − y1 + z − y2) .
then
A42 .
1
Q ZQZ2QZRn\2Q
1
(cid:0)z − y1 + z − y2(cid:1)2n ,
b(y1) − bQf1(y1)b(y2) − bQf2(y2)
(cid:0)z − y1 + z − y2(cid:1)2n
b(y1) − bQf1(y1)b(y2) − bQf2(y2)
dy1dy2dz
dy1dy2dz
. ZQZ2QZRn\2Q
.
1
Q Z2Q
(cid:0)z − y1 + z − y2(cid:1)2n
b(y1) − bQf1(y1)dy1ZQZRn\2Q
2kQ Z2kQ
2−kn
∞
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)
Xk=1
Xk=1
∞
2−kn
2kQ
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)
b(y2) − bQf2(y2)dy2
b(y2) − bQf2(y2)
z − y22n
dy2dz
×(cid:20)Z2kQ
b(y1) − m2kQ(b)f2(y2)dy2 +Z2kQ
m2kQ(b) − bQf2(y2)dy2(cid:21)
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)
∞
Xk=1
2−kn
2kQ
×(cid:20)kbkBMOs′ (ω)ω(x)Mω,s(f2)(x) + kkbkBMO(ω)ω(x)M (f2)(x)(cid:21)
.
2
Yi=1
kbkBMOs′ (ω)ω(x)Mω,s(fi)(x).
Similarly, for A43, we have
A43 .
2
Yi=1
kbkBMOs′ (ω)ω(x)Mω,s(fi)(x).
8
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
For z − z′ ≤ 1
2 max{z − y1, z − y2},
(cid:12)(cid:12)ϕǫ(z − y1 + z − y2) − ϕǫ(z′ − y1 + z′ − y2)(cid:12)(cid:12) .
Therefore,
1 , (b(z) − b)f ∞
z − z′
(cid:0)z − y1 + z − y2(cid:1)2n+1 .
2 )(z′)(cid:12)(cid:12)
ϕǫ(z − y1 + z − y2) − ϕǫ(z′ − y1 + z′ − y2)(cid:12)(cid:12)(cid:12)
2 )(z) − Φ((b(z) − b)f ∞
1 , (b(z) − b)f ∞
b(yi) − bQfi(yi)dyi
b(yi) − bQfi(yi)dyi
b(yi) − bQfi(yi)dy1dy2
2
2
×
. sup
(cid:12)(cid:12)Φ((b(z) − b)f ∞
ǫ>0 ZRn\2QZRn\2Q(cid:12)(cid:12)(cid:12)
Yi=1
ZRn\2Q
Yi=1
Yi=1
Yi=1
2kQ Z2kQ
Xk=1
−2knǫi
.
.
2
∞
.
2
z − z′ǫi
z − yin+ǫi
kbkBMOs′ (ω)ω(x)Mω,s(fi)(x),
where ǫ1, ǫ2 > 0 with ǫ1 + ǫ2 = 1.
Collecting our estimates, we have shown that
M ♯
1
3(cid:0)MΠ~b( ~f )(cid:1)(x) . kbk2
BMO(ω)ω(x)2M (M( ~f )(x))
+kbk2
BMO(ω)ω(x)2
2
Yi=1
Mω,s(fi)(x)
for any 1 < s < ∞ and bounded compact supported functions f1, f2.
+
2
Xi=1
kbkBMO(ω)ω(x)M 1
2
(M(i)
b ( ~f ))(x),
Lemma 3.2. Let ω ∈ A1, ~b = (b, b) and b ∈ BMO(ω). Then there exist a constant C such that
M ♯
1
2(cid:0)M(1)
b ( ~f )(cid:1)(x) . kbkBMO(ω)ω(x)M (M( ~f )(x))
+kbkBMO(ω)ω(x)Mω,s(f1)(x)M (f2)(x),
for any 1 < s < ∞ and bounded compact supported functions f1, f2.
Proof. Let Q be a cube and x ∈ Q. Then, for z ∈ Q we have
(cid:12)(cid:12)Φ(1)
b (f1, f2)(z) − cQ(cid:12)(cid:12) ≤ (cid:12)(cid:12)b(z) − bQΦ(f1, f2)(z)
+(cid:12)(cid:12)Φ((b − bQ)f1, f2)(z) − cQ(cid:12)(cid:12)
1 (z) + BQ
2 (z).
=: BQ
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
9
Therefore,
where Bj = (cid:0) 1
Q RQ(cid:0)BQ
j (z)(cid:1)δ
Let us consider first the term B1. By averaging BQ
1 over Q, we get
dz(cid:19)2
2
.
1/2
1/2
Bj,
b (f1, f2)(z)(cid:12)(cid:12)
− cQ1/2(cid:12)(cid:12)(cid:12)
dz(cid:19)2
(cid:18) 1
Q ZQ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Φ(1)
. (cid:18) 1
Q ZQ(cid:12)(cid:12)ΦΠ~b(f1, f2)(z) − cQ(cid:12)(cid:12)
Xj=1
dz(cid:1)1/δ
Q ZQ(cid:16)(cid:12)(cid:12)b(z) − bQΦ(f1, f2)(z)(cid:17)1/2
, j = 1, 2.
. kbkBMO(ω)
M(f1, f2)(z)dz
ω(Q)
Q
·
1
Q ZQ
. kbkBMO(ω)ω(x)M (M(f1, f2))(x).
dz(cid:19)2
B1 = (cid:18) 1
Let us consider next the term B2. We split fj to fj = f 0
j with f 0
j = fjχ2Q. We write
BQ
1 , f 0
2 ≤ (cid:12)(cid:12)Φ((b − bQ)f 0
+(cid:12)(cid:12)Φ((b − bQ)f ∞
21(z) + BQ
=: BQ
2 )(z)(cid:12)(cid:12) +(cid:12)(cid:12)Φ(((b − bQ)f 0
2 )(z)(cid:12)(cid:12) +(cid:12)(cid:12)Φ((b − bQ)f ∞
23(z) + BQ
24(z).
22(z) + BQ
1 , f 0
1 , f ∞
2 )(z)(cid:12)(cid:12)
2 )(z) − cQ(cid:12)(cid:12)
j + f ∞
1 , f ∞
By Kolmogorov inequality and the fact that M is bounded from L1 × L1 to L1/2,∞, we have
B21 ≤
.
.
C
Q2 kΦ((b − bQ)f 0
1
Q2 kM((b − bQ)f 0
1
Q2 Z2Q
1 , f 0
2 )kL1/2,∞
1 , f 0
2 )kL1/2,∞
b(y1) − bQf1(y1)dy1Z2Q
f2(y2)dy2
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)M (f2)(x).
For B22,
B22 .
.
.
1
1
b(y1) − bQf1(y1)f2(y2)
b(y1) − bQf1(y1)f2(y2)
Q ZQZ2QZRn\2Q
Q ZQZ2QZRn\2Q
Q Z2Q
(cid:0)z − y1 + z − y2(cid:1)2n
(cid:0)z − y1 + z − y2(cid:1)2n
b(y1) − bQf1(y1)dy1ZQZRn\2Q
2kQ Z2kQ
2−kn
∞
1
Xk=1
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)M (f2)(x).
dy1dy2dz
dy1dy2dz
f2(y2)
z − y22n dy2dz
f2(y2)dy2
10
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
For B23, we have
B23 .
.
.
1
1
Q ZQZRn\2QZ2Q
Q ZQZ2QZRn\2Q
Q ZQZRn\2Q
1
b(y1) − bQf1(y1)f2(y2)
b(y1) − bQf1(y1)f2(y2)
(cid:0)z − y1 + z − y2(cid:1)2n
(cid:0)z − y1 + z − y2(cid:1)2n
dy1dzZ2Q
z − y12n
b(y1) − bQf1(y1)
dy1dy2dz
dy1dy2dz
f2(y2)dy2
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)M (f2)(x).
Concerning the last estimate for B24. For any z′ ∈ Q and y1, y2 ∈ Rn\2Q, we have
1 , f ∞
2 )(z′)(cid:12)(cid:12)
ϕǫ(z − y1 + z − y2) − ϕǫ(z′ − y1 + z′ − y2)(cid:12)(cid:12)(cid:12)
. sup
1 , f ∞
2 )(z) − Φ((b − bQ)f ∞
×b(y1) − bQf1(y1)f2(y2)dy1dy2
(cid:12)(cid:12)Φ((b − bQ)f ∞
ǫ>0ZRn\2QZRn\2Q(cid:12)(cid:12)(cid:12)
. ZRn\2Q
Xk=2
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)M (f2)(x).
2kQ Z2kQ
dy1ZRn\2Q
b(y1) − bQf1(y1)dy1
b(y1) − bQf1(y1)
z − y12n+ǫ1
2−knǫ1
.
∞
f2(y2)
z − y2ǫ2
dy2
∞
Xi=2
2−knǫ2
2kQ Z2iQ
f2(y2)dy2
where ǫ1, ǫ2 > 0 with ǫ1 + ǫ2 = 1. Taking the mean over Q for z and z′ respectively, we obtain
B24 .
.
1
1 , f ∞
Q ZQ(cid:12)(cid:12)Φ((b − bQ)f ∞
Q ZQ
Q ZQ(cid:12)(cid:12)Φ((b − bQ)f ∞
1
1
2 )(z) − cQ(cid:12)(cid:12)dz
. kbkBMOs′ (ω)ω(x)Mω,s(f1)(x)M (f2)(x).
1 , f ∞
2 )(z) − Φ((b − bQ)f ∞
1 , f ∞
2 )(z′)(cid:12)(cid:12)dzdz′
Collecting our estimates, we have shown that
M ♯
1
2(cid:0)M(1)
b ( ~f )(cid:1)(x) . kbkBMO(ω)ω(x)M (M( ~f )(x))
+kbkBMO(ω)ω(x)Mω,s(f1)(x)M (f2)(x),
for any 1 < s < ∞ and bounded compact supported functions f1, f2.
Similarly, we have
Lemma 3.3. Let ω ∈ A1, ~b = (b, b) and b ∈ BMO(ω). Then there exist a constant C such that
M ♯
1
2(cid:0)M(2)
b ( ~f )(cid:1)(x) . kbkBMO(ω)ω(x)M (M( ~f )(x))
+kbkBMO(ω)ω(x)Mω,s(f2)(x)M (f1)(x),
for any 1 < s < ∞ and bounded compact supported functions f1, f2.
Lemma 3.4. Let ω ∈ A1 and 0 < p < ∞. Then ω1−p ∈ A∞.
Proof. If 0 < p ≤ 1, then 1 − p ∈ [0, 1). It is easy to see that ω1−p ∈ A1 ⊂ A∞.
If 1 < p < ∞, if follows from ω ∈ A1 ⊂ Ap that ω1−p ∈ Ap′ ⊂ A∞.
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
11
Lemma 3.5. Let ω ∈ A1, 1 < s < p1, p2 < ∞ and 1/p = 1/p1 + 1/p2. Then both M( ~f ) and
Q2
i=1 Mω,s(fi) are bounded from Lp1(ω) × Lp2(ω) to Lp(ω).
1 < s < p1, p2 < ∞, it is easy to obtain that both M( ~f ) and Q2
Proof. From the fact that M (f )(x) . Mω,s(f )(x) and Mω,s(f )(x) is bounded on Lp(ω) for
i=1 Mω,s(fi) are bounded from
Lp1 (ω) × Lp2 (ω) to Lp(ω).
The following relationships between Mδ and M ♯ to be used is a version of the classical ones
due to Fefferman and Stein [9].
Lemma 3.6. Let 0 < p, δ < ∞ and ω ∈ A∞. There exist a positive C such that
ZRn
(Mδf (x))pω(x)dx ≤ CZRn
(M ♯
δ f (x))pω(x)dx,
for any smooth function f for which the left-hand side is finite.
Lemma 3.7. Let Q0 be any fixed cube and b be a locally integral function. Then, for any
x ∈ Q0, we get
(1)
(2)
(3)
M(χQ0 , χQ0)(x) ≡ 1;
M(bχQ0, χQ0 )(x) = M(χQ0 , bχQ0)(x) = MQ0 (b)(x);
where MQ0(b)(x) = supQ0⊃Q∋x
1
M(bχQ0, bχQ0)(x) = M2
Q RQ b(y)dy.
Q0 (b)(x),
Proof. We only give the proof of (3) and the proof of (1),(2) are similar. For any x ∈ Q0, we
have
M 2
Q0(b)(x) = (cid:16) sup
Q0⊃Q∋x
1
sup
1
Q ZQ
Q ZQ
=
Q0⊃Q∋x
b(y)dy(cid:17)2
b(y1)χQ0 (y1)dy1 ·
1
Q ZQ
b(y2)χQ0 (y2)dy2
≤ M(bχQ0, bχQ0 )(x).
On the other hand, for any cube Q ⊂ Rn, we can construct a cube Q1 such that
and Q1 ≤ Q. Therefore,
Q0 ⊃ Q1 ⊃ Q0 ∩ Q ∋ x
1
Q ZQ∩Q0
b(y)dy ≤
1
Q1 ZQ1
b(y)dy ≤ MQ0(b)(x).
Thus,
then (3) is proved.
M(bχQ0 , bχQ0)(x) = sup
Q∋x(cid:16) 1
Q ZQ
b(y)χQ0 (y)dy(cid:17)2
≤ M 2
Q0(b)(x),
§4 Proofs of Theorem 1.1 and Theorem 1.2
Proof of Theorem 1.1. (A1) ⇒ (A2): It is enough to prove Theorem 1.1 for f1, f2 being
bounded functions with compact support. We observe that to use the Fefferman-Stein inequal-
ity, one needs to verify that certain terms in the left-hand side of the inequalities are finite.
12
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
Applying a similar argument as in [12, pp.32-33], the boundedness properties of M and Fatou's
lemma, one gets the desired result.
Since Lemma 3.4 and ω ∈ A1, then ω1−p ∈ A∞. By Lemma 3.2 and Lemma 3.3 with
1 < s < min{p1, p2}, from a standard argument that we can obtain
kMΣ~b( ~f )kLp(ω1−p) . kM 1
2(cid:0)MΣ~b( ~f )(cid:1)kLp(ω1−p) . kM ♯
. kbkBMO(ω)(cid:18)(cid:13)(cid:13)M(cid:0)M( ~f )(cid:1)(cid:13)(cid:13)Lp(ω) +(cid:13)(cid:13)
2(cid:0)MΣ~b( ~f )(cid:1)kLp(ω1−p)
Mω,s(fi)(cid:13)(cid:13)Lp(ω)(cid:19)
Yi=1
2
1
2
. kbkBMO(ω)
kfikLpi (ω).
Yi=1
(A2) ⇒ (A1): Let Q be any fixed cube. Suppose that MΣ~b is bounded from Lp1 (ω)×Lp2(ω)
into Lp(ω1−p), then
kMΣ~b(χQ, χQ)kLp(ω1−p) . kχQkLp1 (ω)kχQkLp2 (ω) . ω(Q)
1
p ,
which implies that
b(x) − bQdx
2
ω(Q) ZQ
ω(Q) ZQ
1
≤
Q−2ZQZQ
b(x) − b(y1)χQ(y1)χQ(y2)dy1dy2dx
+
1
ω(Q) ZQ
Q−2ZQZQ
b(x) − b(y2)χQ(y1)χQ(y2)dy1dy2dx
.
.
1
MΣ~b(χQ, χQ)(x)dx
ω(Q) ZQ
ω(Q)(cid:16)ZQ(cid:12)(cid:12)M~b(χQ, χQ)(x)(cid:12)(cid:12)
1
1
.
ω(Q)1/p kMΣ~b(χQ, χQ)kLp(ω1−p)
. kMΣ~bkLp1 (ω)×Lp2 (ω)→Lp(ω1−p).
p
ω(x)1−pdx(cid:17)1/p(cid:16)ZQ
ω(x)dx(cid:17)1/p′
Thus showing that b ∈ BMO(ω).
(A1) ⇒ (A3): Since ω ∈ A1, Lemma 3.4 implies that ω1−2p ∈ A∞. From Lemma 3.1,
Lemma 3.2 and Lemma 3.3 with 1 < s < min{p1, p2}, we get
kMΠ~b( ~f )kLp(ω1−2p) . kM 1
. kbk2
3(cid:0)MΠ~b( ~f )(cid:1)kLp(ω1−2p) . kM ♯
BMO(ω)(cid:18)(cid:13)(cid:13)M(cid:0)M( ~f )(cid:1)(cid:13)(cid:13)Lp(ω) +(cid:13)(cid:13)
Xi=1
kbkBMO(ω)(cid:13)(cid:13)M 1
(M(i)
2
2
2
+
. kbk2
BMO(ω)
kfikLpi (ω).
Yi=1
1
2
3(cid:0)MΠ~b( ~f )(cid:1)kLp(ω1−2p)
Mω,s(fi)(cid:13)(cid:13)Lp(ω)(cid:19)
Yi=1
b ( ~f ))(x)(cid:13)(cid:13)Lp(ω1−p)
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
13
(A3) ⇒ (A1): By MΠ~b is bounded from Lp1 (ω) × Lp2 (ω) into Lp(ω1−2p), we get
1
1
b(x) − bQ2ω(x)−1dx
ω(x)−1Q−2ZQZQ
MΠ~b(χQ, χQ)(x)ω(x)−1dx
ω(Q) ZQ
ω(Q) ZQ
ω(Q) ZQ
ω(Q)(cid:16)ZQ(cid:12)(cid:12)MΠ~b(χQ, χQ)(x)(cid:12)(cid:12)
1
1
1
.
.
.
.
ω(Q)1/p kMΠ~b(χQ, χQ)kLp(ω1−2p)
. kMΠ~bkLp1 (ω)×Lp2 (ω)→Lp(ω1−2p).
b(x) − b(y1)b(x) − b(y2)dy1dy2dx
p
ω(x)1−2pdx(cid:17)1/p(cid:16)ZQ
ω(x)dx(cid:17)1/p′
Thus we complete the proof of Theorem 1.1.
Proof of Theorem 1.2. (B1) ⇒ B2): By the definition of M( ~f ), we have
M (bf1, f2)(x) = M (bf1, f2)(x), M (f1, bf2)(x) = M (f1, bf2)(x).
Then
(cid:12)(cid:12)[b, M](1)( ~f )(x) − [b, M](1)( ~f )(x)(cid:12)(cid:12)
. (cid:12)(cid:12)(cid:12)
b(x)M( ~f )(x) − M(bf1, f2)(x) − b(x)M( ~f )(x) + M(bf1, f2)(x)(cid:12)(cid:12)(cid:12)
. b−(x)M( ~f )(x).
Similarly, we also have (cid:12)(cid:12)[b, M](2)( ~f )(x) − [b, M](2)( ~f )(x)(cid:12)(cid:12) . b−(x)M( ~f )(x). Since (cid:12)(cid:12)a − c(cid:12)(cid:12) ≤
a − c for any real numbers a and c, there holds
for i = 1, 2. This shows that
(cid:12)(cid:12)[b, M](i)(f1, f2)(x)(cid:12)(cid:12) ≤ M(i)
b (f1, f2)(x),
Applying (4) and Theorem 1.1 we have
(cid:12)(cid:12)Σ~b, M]( ~f )(x)(cid:12)(cid:12) . MΣ~b( ~f )(x) + b−(x)M( ~f )(x).
(4)
(cid:13)(cid:13)[Σ~b, M]( ~f )(x)(cid:13)(cid:13)Lp(ω1−p) . (cid:13)(cid:13)MΣ~b( ~f )(cid:13)(cid:13)Lp(ω1−p) +(cid:13)(cid:13)b−M( ~f )(cid:13)(cid:13)Lp(ω1−p)
. (cid:0)kb−/ωkL∞ + kbkBMO(ω)(cid:1)kf1kLp1 (ω)kf2kLp2 (ω).
Therefore, b ∈ BMO(ω) with b−/ω ∈ L∞ implies that [Σ~b, M] is bounded from Lp1(ω)× Lp2(ω)
to Lp(ω1−p).
(B2) ⇒ (B1): Let Q0 be any fixed cube. By Lemma 3.7, for any x ∈ Q0,
b(x) = b(x)M(χQ0 , χQ0)(x),
MQ0 (b)(x) = M(bχQ0, χQ0 )(x) = M(χQ0 , bχQ0)(x),
14
Then,
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
=
2
2
ω(Q0) ZQ0
ω(Q0) ZQ0
ω(Q0) ZQ0
1
=
b(x) − MQ0(b)(x)dx
b(x)M(χQ0 , χQ0)(x) − M(bχQ0 , χQ0)(x)dx
b(x)M(χQ0 , χQ0)(x) − M(bχQ0 , χQ0)(x)
+b(x)M(χQ0 , χQ0 )(x) − M(χQ0, bχQ0 )(x)dx
1
.
.
.
1
ω(Q0) ZQ0 (cid:12)(cid:12)[Σ~b, M](χQ0, χQ0 )(x)(cid:12)(cid:12)dx
ω(Q0)(cid:18)ZQ0 (cid:12)(cid:12)(cid:12)
[Σ~b, M](χQ0, χQ0 )(x)(cid:12)(cid:12)(cid:12)
ω(Q0)1/p(cid:13)(cid:13)[Σ~b, M](χQ0 , χQ0)(cid:13)(cid:13)Lp(ω1−p)
1
p
. k[Σ~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−p).
ω(x)1−pdx(cid:19)1/p
·(cid:18)ZQ0
ω(x)dx(cid:19)1/p′
Now, we have all the ingredients to prove b ∈ BMO(ω) and b−/ω ∈ L∞.
1
ω(Q0) ZQ0
b(x) − bQ0dx .
1
ω(Q0) ZQ0
b(x) − MQ0(b)(x)dx
. k[Σ~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−p).
which implies that b ∈ BMO(ω).
In order to show show that b−/ω ∈ L∞, observe that for any x ∈ Q0, MQ0 (b)(x) ≥ b(x).
Therefore,
which gives
0 ≤ b−(x) . MQ0 (b)(x) − b+(x) + b−(x) = MQ0 (b)(x) − b(x),
1
Q0 ZQ0
b−(x)
ω(x)
dx .
.
1
Q0 ZQ0
Q0 ZQ0
1
b−(x)dx ·
1
inf x∈Q0 ω(x)
b(x) − MQ0 (b)(x)dx ·
Q0
ω(Q0)
. k[Σ~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−p),
this yields that
Thus, the boundedness of b−/ω follows from Lebesgue's differentiation theorem.
(b−/ω)Q0 . k[Σ~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−p).
(B1) ⇒ (B3): Let ~B = (b, b) and ~B = (b, b). Then
b(x)b(x)M( ~f )(x) − b(x)M(f1, bf2)(x)
(cid:12)(cid:12)(cid:12)
[Π~b, M](f1, f2)(x) − [Π ~B, M](f1, f2)(x)(cid:12)(cid:12)(cid:12)
. (cid:12)(cid:12)(cid:12)
−b(x)b(x)M( ~f )(x) + b(x)M(f1, bf2)(x)(cid:12)(cid:12)(cid:12)
. b−(x)(cid:12)(cid:12)[b, M](2)(f1, f2)(x)(cid:12)(cid:12).
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
15
Similarly, we also have
Noting that
which yields that
b(x)b(x)M( ~f )(x) − b(x)M(bf1, f2)(x)
(cid:12)(cid:12)(cid:12)
[Π~B, M](f1, f2)(x) − [Π ~B, M](f1, f2)(x)(cid:12)(cid:12)(cid:12)
. (cid:12)(cid:12)(cid:12)
−b(x)b(x)M( ~f )(x) + b(x)M(bf1, f2)(x)(cid:12)(cid:12)(cid:12)
. b−(x)(cid:12)(cid:12)[b, M](1)(f1, f2)(x)(cid:12)(cid:12).
(cid:12)(cid:12)[Π~B, M]( ~f )(x)(cid:12)(cid:12) ≤ MΠ~b( ~f )(x),
It follows from Theorem 1.1 and b−/ω ∈ L∞ that
(cid:12)(cid:12)[Π~b, M]( ~f )(x)(cid:12)(cid:12) . MΠ~b( ~f )(x) + b−(x)MΣ~b( ~f )(x) + (b−(x))2M( ~f )(x).
(cid:13)(cid:13)[Π~b, M]( ~f )(x)(cid:13)(cid:13)Lp(ω1−2p)
. (cid:13)(cid:13)MΠ~b( ~f )(cid:13)(cid:13)Lp(ω1−2p) +(cid:13)(cid:13)b−MΣ~b( ~f )(cid:13)(cid:13)Lp(ω1−2p) +(cid:13)(cid:13)(b−)2M( ~f )(cid:13)(cid:13)Lp(ω1−2p)
BMO(ω)kf1kLp1 (ω)kf2kLp2 (ω) + kb−/ωkL∞(cid:13)(cid:13)MΣ~b( ~f )(cid:13)(cid:13)Lp(ω1−p)
. kbk2
+kb−/ωk2
L∞kM( ~f )kLp(ω)
. (cid:0)kb−/ωkL∞ + kbkBMO(ω)(cid:1)2
this leads to our results.
kf1kLp1 (ω)kf2kLp2 (ω),
(B3) ⇒ (B1): Let Q0 be any fixed cube. By Lemma 3.5, for any x ∈ Q0,
b(x)2 = b(x)2M(χQ0 , χQ0 )(x),
b(x)MQ0 (b)(x) = b(x)M(bχQ0 , χQ0 )(x) = b(x)M(χQ0 , bχQ0)(x),
M 2
Q0 (b)(x) = M(bχQ0 , bχQ0)(x).
Then,
1
1
1
1
=
b(x) − MQ0(b)(x)2ω(x)−1dx
ω(Q0) ZQ0
ω(Q0) ZQ0 (cid:16)b(x)2 − 2b(x)MQ0 (b)(x) + M 2
ω(Q0) ZQ0
[Π~b, M](χQ0, χQ0 )(x)ω(x)−1dx
ω(Q0)(cid:18)ZQ0 (cid:12)(cid:12)(cid:12)
[Π~b, M](χQ0 , χQ0 )(x)(cid:12)(cid:12)(cid:12)
ω(Q0)1/p(cid:13)(cid:13)[Π~b, M](χQ0 , χQ0 )(cid:13)(cid:13)Lp(ω1−2p)
. k[Π~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−2p).
=
1
=
.
Q0(b)(x)(cid:17)ω(x)−1dx
p
ω(x)1−2pdx(cid:19)1/p
·(cid:18)ZQ0
ω(x)dx(cid:19)1/p′
16
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
Now, we have all the ingredients to prove b ∈ BMO(ω) and b−/ω ∈ L∞.
1
ω(Q0) ZQ0
ω(Q0) ZQ0
1
.
b(x) − bQ0 2ω(x)−1dx
b(x) − MQ0(b)(x)2ω(x)−1dx
+
1
ω(Q0) ZQ0
bQ0 − MQ0(b)(x)2ω(x)−1dx
1
.
b(x) − MQ0(b)(x)2ω(x)−1dx
ω(Q0) ZQ0
. k[Π~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−2p).
which implies that b ∈ BMO2(ω); that is, b ∈ BMO(ω).
For any x ∈ Q0, we have
b−(x)
ω(x)
Q0 ZQ0
1
dx .
.
1
1
Q0 ZQ0
Q0 ZQ0
ω(Q0) ZQ0
b−(x)dx ·
1
inf x∈Q0 ω(x)
b(x) − MQ0(b)(x)dx ·
Q0
ω(Q0)
. (cid:18) 1
. k[Π~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−2p),
b(x) − MQ0 (b)(x)2ω(x)−1dx(cid:19)1/2
which yields
Thus, the boundedness of b−/ω follows from Lebesgue's differentiation theorem.
(b−/ω)Q0 . k[Π~b, M]kLp1 (ω)×Lp2 (ω)→Lp(ω1−2p).
The proof of Theorem 1.2 is complete.
§5 Acknowledgments
The research was supported by National Natural Science Foundation of China (Grant
NO.11661075).
References
[1] J Bastero, M Milman, F J Ruiz. Commutators for the maximal and sharp functions, Proc.
Amer. Math. Soc., 2000(128): 3329-3334.
[2] ´A B´enyi, W Dami´an, K Moen and R H Torres. Compactness properties of commutators
of bilinear fractional integrals, Math. Z. (2015), doi: 10.1007/s00209-015-1437-4.
[3] S Bloom. A commutator theorem and weighted BMO, Trans. Amer. Math. Soc., 1985(292):
103-122.
[4] L Chaffee, R H Torres. Characterization of Compactness of the Commutators of Bilinear
Fractional Integral Operators, Potential Anal., 2015(43): 481-494.
Wang Ding-huai and Zhou Jiang.
Commutators of bilinear Hardy-Littlewood maximal function
17
[5] S Chanillo. A note on commutators, Indiana Univ. Math. J., 1982(31): 7-16.
[6] X Chen, Q Xue. Weighted estimates for a class of multilinear fractional type operators,
J. Math. Anal. Appl. (2010(362): 355-373.
[7] R Coifman, R Rochberg, G Weiss. Factorization theorems for Hardy spaces in several
variables, Ann. of Math, 1976(103): 611-635.
[8] Y Ding. A characterization of BMO via commutators for some operators, Northeast.
Math. J., 1997(13): 422-432.
[9] C Fefferman, E M Stein. H p spaces of several variables Acta Math., 1972(129): 137-193.
[10] J Garc´ıa-Cuerva. Weighted H p spaces, Dissertationes Math. 1979(162).
[11] J Garc´ıa-Cuerva, E Harboure, C Segovia, J L Torrea. Weighted norm inequalities for
commutators of strongly singular integrals, Indiana Univ. Math. J., 1991(40): 1397-1420.
[12] A K Lerner, S Ombrosi, C P´erez, R H Torres, R Trujillo-Gonz´alez. New maximal functions
and multiple weights for the multilinear Calder´on-Zygmund theory, Adv. Math., 2009(220)
1222-1264.
[13] B Muckenhoupt. Weighted norm inequalities for the Hardy maximal function, Trans.
Amer. Math. Soc., 1972(165): 207-226.
[14] C P´erez. Endpoint estimates for commutators of singular integral operators, J. Funct.
Anal., (1995(128): 163-185.
[15] C P´erez, G Pradolini, R H Torres, R Trujillo-Gonz´alez. End-points estimates for iterated
commutators of multilinear singular integrals, Bull. London Math. Soc., 2014(46): 26-42.
[16] C. P´erez and R.H. Torres, Sharp maximal function estimates for multilinear singular
integrals, Contemp. Math. 320 (2003) 323-331. 1045.42011 MR1979948
[17] C Segovia, J L Torrea. Weighted inequalities for commutators of fractional and singular
integrals, Publicacions Matem`atuques, (1991(35): 209-235.
[18] C Segovia, J L Torrea. Higher order commutators for vector-valued Calder´on-Zygmund
operators, Trans. Amer. Math. Soc., 1993(336): 537-556.
[19] L Tang. Weighted estimates for vector-valued commutators of multilinear operators, Proc.
Roy. Soc. Edinburgh Sect. A, (2008(138): 897-922.
[20] S B Wang, J B Pan, Y S Jiang. Necessary and sufficient conditions for boundedness of
commutators of multilinear fractional integral operators, Acta Math. Sin., 2015(35): 1106-
1114. (in Chinese)
18
Appl. Math. J. Chinese Univ.
Vol. 28, No. *
[21] M Wilson. Weighted Littlewood-Paley Theory and Exponential-Square Integrability, Lec-
ture Notes in Math vol. 1924. Springer, Berlin (2008).
[22] M M Rao, Z D Ren. Theory of Orlicz Spaces, Marcel Dekker, New York (1991).
[23] P Zhang, Multiple Weighted Estimates for Commutators of Multilinear Maximal Function,
Acta Math. Sin. (Engl. Ser.), 2015(31): 973-994.
[24] P Zhang, Commutators of the fractional maximal function on variable exponent Lebesgue
spaces, Czech. Math. J., 2014(64): 183-197.
College of Mathematics and System Sciences, Xinjiang University, Urumqi 830046, Republic
of China
Email:[email protected]; [email protected].
|
1003.5066 | 4 | 1003 | 2012-06-28T06:46:31 | A Bernstein-type inequality for rational functions in weighted Bergman spaces | [
"math.FA"
] | Given $n\geq1$ and $r\in[0, 1),$ we consider the set $\mathcal{R}_{n, r}$ of rational functions having at most $n$ poles all outside of $\frac{1}{r}\mathbb{D},$ were $\mathbb{D}$ is the unit disc of the complex plane. We give an asymptotically sharp Bernstein-type inequality for functions in $\mathcal{R}_{n, r}\:$ (as n tends to infinity and r tends to 1-) in weighted Bergman spaces with "polynomially" decreasing weights. We also prove that this result can not be extended to weighted Bergman spaces with "super-polynomially" decreasing weights. | math.FA | math |
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL
FUNCTIONS IN WEIGHTED BERGMAN SPACES
ANTON BARANOV AND RACHID ZAROUF
Abstract. Given n ≥ 1 and r ∈ [0, 1), we consider the set Rn, r of rational
functions having at most n poles all outside of 1
D, were D is the unit disc of the
complex plane. We give an asymptotically sharp Bernstein-type inequality for
functions in Rn, r in weighted Bergman spaces with "polynomially" decreasing
weights. We also prove that this result can not be extended to weighted
Bergman spaces with "super-polynomially" decreasing weights.
r
1. Introduction
Estimates of the norms of derivatives for polynomials and rational functions (in
different functional spaces) is a classical topic of complex analysis (see surveys by
A.A. Gonchar [10], V.N. Rusak [16], and P. Borwein and T. Erd´elyi [3, Chapter
7]). Such inequalities have applications in many domains of analysis; to mention
just some of them: 1) matrix analysis and in operator theory (see "Kreiss Matrix
Theorem" [12, 17] or [19, 18] for resolvent estimates of power bounded matrices), 2)
inverse theorems of rational approximation (see [4, 15, 14]), 3) effective Nevanlinna --
Pick interpolation problems (see [23, 22]).
Rn, r =(cid:26) p
q
Here, we present Bernstein-type inequalities for rational functions f of degree n
with poles in {z :
z > 1}, involving Hardy norms and weighted Bergman norms.
Let Pn be the complex space of polynomials of degree less or equal to n ≥ 1. Let D =
{z ∈ C : z < 1} be the unit disc of the complex plane and D = {z ∈ C : z ≤ 1}
its closure. Given r ∈ [0, 1), we define
: p, q ∈ Pn, d◦p < d◦q, q(ζ) 6= 0 ζ <
r(cid:27) ,
(where d◦p denotes the degree of p ∈ Pn), the set of all rational functions in D of
degree less or equal than n ≥ 1, having at most n poles all outside of 1
D. Notice
that for r = 0, we get Rn, 0 = Pn−1.
1.1. Definitions of Hardy spaces and radial weighted Bergman spaces.
We denote by Hol (D) the space of all holomorphic functions on D. From now on,
if f ∈ Hol (D) then for every ρ ∈ (0, 1) we define
ξ ∈
fρ : ξ 7→ f (ρξ) ,
1
ρ
D.
1
r
We consider the two following scales of Banach spaces X ⊂ Hol (D) :
2000 Mathematics Subject Classification. Primary 32A36, 26A33; Secondary 26C15, 41A10.
Key words and phrases. Rational function, Bernstein-type inequality, weighted Bergman norm.
The first author is supported by the Chebyshev Laboratory (St.Petersburg State University)
under RF Government grant 11.G34.31.0026 and by RFBR grant 12-01-00434.
1
2
ANTON BARANOV AND RACHID ZAROUF
a. The Hardy spaces H p = H p(D), 1 ≤ p ≤ ∞ :
H p =(cid:26)f ∈ Hol (D) : kfkp
Hp = sup
0≤ρ<1ZT fρ(ξ)p dm(ξ) < ∞(cid:27) ,
where m stands for the normalized Lebesgue measure on T = {z ∈ C : z = 1}. As
usual,we denote by H ∞ the space of all bounded analytic functions in D.
a (w), 1 ≤ p < ∞ (where "a" means
b. The radial weighted Bergman spaces Lp
analytic),
0
Lp
Lp
a (w) =(cid:26)f ∈ Hol (D) : kfkp
a(w) =Z 1
where the weight w satisfies w ≥ 0 and R 1
weights w(ρ) = wβ(ρ) = (1 − ρ)β, β > −1, we have Lp
A being the normalized area measure on D.
ρw (ρ)ZT fρ (ζ)p dm(ζ)dρ < ∞(cid:27) ,
a(cid:0)(1−z)βdA(z)(cid:1),
0 w(ρ)dρ < ∞. For the classical power
For general properties of these spaces we refer to [11, 24].
From now on, for two positive functions a and b, we say that a is dominated by
b, denoted by a . b, if there is a constant c > 0 such that a ≤ cb; and we say that
a and b are comparable, denoted by a ≍ b, if both a . b and b . a.
1.2. Statement of the problem and known results. By Bernstein-type in-
equalities for rational functions one usually understands the inequalities of the
form
a (wβ) = Lp
(1.1)
kf ′kX ≤ φX, Y (n)kfkY ,
f ∈ Rn,
where Rn is the set of all proper rational functions of degree at most n with the
poles in {z > 1}, X and Y are some normed spaces of functions analytic in the
unit disc, and φ is some increasing (often polynomially growing) function. Thus,
for a given pair of the function spaces X and Y , the question is to determine the
dependence on n for the norm of the differentiation operator (Rn,k · kX ) to Y .
Bernstein-type inequalities of E.P. Dolzhenko [5] and A.A. Pekarskii [14] are of this
form; e.g., it is shown in [5] that
kfkH 1
1 ≤ c1nkfk∞,
kfkB
1/2
2,2 ≤ c2n1/2kfk∞,
f ∈ Rn,
1 is the Hardy -- Sobolev space, and B1/2
where H 1
2,2 is the Besov (or Dirichlet) space,
see the definition in Section 3. Let us also mention that this problem is a part of
a more general one given by G. Lorentz in a letter sent to T. Erd´elyi in 1988 (see
[9]).
Looking at (1.1), we notice that for some choices of X and Y , we have φX, Y (n) =
+∞ for every n = 1, 2, . . . . Indeed, it may happen for instance when the poles of
our function f are allowed to be arbitrary close to the torus T : we can observe
this phenomenon for example in the special case X = Y = H p, 1 ≤ p ≤ +∞ but
also when X = Y = Lp
a(w), 1 ≤ p ≤ +∞. This observation leads us to come back
on the problem in (1.1) and to state it more generally : that is replacing Rn by
Rn, r (for any fixed r ∈ [0, 1)) and φX, Y (n) by φX, Y (n, r) so that to focus on this
phenomenon of "natural dependence on the parameter r". For most of the classical
cases already studied by others (for instance E. P. Dolzhenko [5], A. A. Pekarskii
[14], V.V. Peller [15]) the spaces X and Y are such that supr∈(0, 1) φX, Y (n, r) <
+∞: in this case we can set φX, Y (n) = supr∈(0, 1) φX, Y (n, r). As a consequence,
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL FUNCTIONS
3
if supr∈(0, 1) φX, Y (n, r) = +∞, it may be of interest to search (as a continuation of
the investigations of the second author [20, 21]) for the "best possible" φX, Y (n, r)
in an asymptotically sense, that is to say as n → ∞ and r → 1−. This question
has already been answered for the case X = Y = H p, 1 ≤ p ≤ +∞ by K. M.
Dyakonov [6] see (1.2) below. In this paper, we answer the same question for the
case X = Y = Lp
a(w), 1 ≤ p ≤ +∞. Let us give a general formulation of our
problem for the special case X = Y for which we set Cn, r(X) = φX, Y (n, r) : given
a Banach space X of holomorphic functions in D, we are searching for the best
possible constant Cn, r(X) such that
kf ′kX ≤ Cn, r(X)kfkX ,
f ∈ Rn, r.
For the case where X = H p is a Hardy space, an estimate which gives a correct
order of growth for Cn, r(X) was obtained by K.M. Dyakonov [6] (as a very special
case of more general results): for any p ∈ [1,∞] there exist positive constants Ap
and Bp such that
(1.2)
Ap
n
1 − r ≤ Cn, r(H p) ≤ Bp
n
1 − r
for all n ≥ 1 and r ∈ [0, 1). More precisely, the upper estimate for p ∈ (1, +∞)
is treated in [6, Theorem 1], the case p = 1, in [6, Corollary 1], and the case
p = +∞ (known much earlier) is given in [3, Theorem 7.1.7]. The below estimate
follows trivially when applying the differentiation operator to the test function
f (z) = (1 − rz)−n.
any r ∈ (0, 1) there exists the limit
For the case p = 2 an asymptotically sharp result was obtained later in [20]: for
lim
n→∞
Cn, r(cid:0)H 2(cid:1)
n
=
1 + r
1 − r
.
Related results about Bernstein-type inequalities in a more general setting of the
so-called model or star invariant subspaces may be found in [8, Theorems 10,11],
[7, Theorem 1], and [1, 2].
1.3. Main results. We obtain estimates for the derivatives of rational functions
with respect to weighted Bergman norms.
It turns out that there is an essen-
tial difference between slowly (polynomially) decreasing weights and fast (super-
polynomially) decreasing weights. In the first case we have a two-sided estimate
analogous to (1.2), while in the second case only the above estimate remains true.
Let us give the precise definitions. Recall that w is always an integrable nonnegative
function on (0, 1).
Definition 1.1. (Polynomially decreasing weights) The weight w is said to be
γ-polynomially decreasing if there exists γ > 0 such that
is increasing on [r0, 1) for some 0 ≤ r0 < 1. We say that w is polynomially
decreasing if it is γ-polynomially decreasing for some γ > 0.
ρ 7→ (1 − ρ)−γw(ρ),
Definition 1.2. (Super-polynomially decreasing weights) The weight w is said to
be super-polynomially decreasing if for any γ > 0 there exists r(γ) ∈ (0, 1) such
that the function
decreases on the interval [r(γ), 1).
ρ 7→ (1 − ρ)−γw(ρ),
4
ANTON BARANOV AND RACHID ZAROUF
Typical example of the weights from the first class are given by w(r) = (1 − r)β,
β > −1, or w(r) = (1 − r)β ( log(1 − r) + 1)γ, β > −1, γ ∈ R. The weights
Our first result may be considered as an analogue of Dyakonov's theorem for the
w(r) = exp(cid:0) − c(1 − r)−γ(cid:1), c > 0, γ > 0 are super-polynomially decreasing.
radial weighted Bergman spaces.
Theorem 1.3. Let 1 ≤ p < ∞ and let w be an integrable nonnegative function on
[0, 1). Then there exists a positive constant K depending only on p (but not on the
weight w) such that
(1.3)
Cn, r (Lp
a (w)) ≤ K
n
1 − r
for all r ∈ [0, 1) and n ≥ 1. Moreover, if we fix r ∈ (0, 1) and let n tend to infinity,
then we have
(1.4)
eKr
1 − r ≤ lim inf
n→∞
a(w))
Cn,r(Lp
n
≤ lim sup
n→∞
a(w))
Cn,r(Lp
n
≤
K
1 − r
,
where eK is, as K, a positive constant depending only on p.
The next theorem shows that for the polynomially decreasing weights the quan-
a(w)) admits a below estimate of the same form.
tity Cn, r(Lp
Theorem 1.4. If w is γ-polynomially decreasing, then there exists a positive con-
stant K ′ depending only on w and p such that
(1.5)
K ′ n
1 − r ≤ Cn, r (Lp
a (w)) ≤ K
n
1 − r
,
where K is defined in (1.3) and where the left-hand side inequality of (1.5) holds
for all r ∈ [0, 1) and n ≥ γ+3
p + 1. In particular, (1.5) holds for the classical weights
w(ρ) = wβ (ρ) = (1 − ρ)β ρ, β > −1.
The polynomial decrease is essential and provides a sharp bound for the validity
of the uniform estimate (1.5) for all possible values of n and r. Namely, if the
weight is super-polynomially decreasing, then (1.5) will fail along some sequence of
radii.
Theorem 1.5. Suppose that w is super-polynomially decreasing. Then there exists
a sequence rn → 1− such that for any p,
Crn,n(Lp
a(w))
n
= o(cid:18) 1
1 − rn(cid:19),
n → ∞.
Acknowledgements. The authors are deeply grateful to Nikolai Nikolski,
Alexander Borichev, and Evgueny Doubtsov for many helpful discussions, con-
structive comments and precious remarks which definitively helped to improve the
manuscript.
2. Proofs of Theorems 1.3 and 1.4
Proof of Theorem 1.3. First, we notice that for any 0 ≤ α < 1,
(2.1)
kfkp
Lp
a(w) ≍ZZu∈Cα
ρf (ρξ)p w(ρ)dm(ζ)dρ
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL FUNCTIONS
5
for all f ∈ Lp
a(w), where Cα = {z : α < z < 1}. Let f ∈ Rn, r with r ∈ [0, 1)
and n ≥ 1. Using (2.1) with α = 1
2 we get
1
Lp
ρ w(ρ)
kf ′kp
a(w) ≍ ZZρξ∈C1/2 f ′(ρξ)p w (ρ) dm(ζ)dρ
= Z 1
ρp(cid:16)(cid:13)(cid:13)(cid:13)(fρ)
′(cid:13)(cid:13)(cid:13)
Hp(cid:17) dρ.
Hp(cid:17) dρ ≤ (2Cn, r (H p))pZ 1
ρp(cid:16)(cid:13)(cid:13)(cid:13)(fρ)
′(cid:13)(cid:13)(cid:13)
≍ (Cn, r (H p))p kfkp
a(w) .
Lp
1
1
2
1
2
p
p
ρw (ρ)(cid:0)kfρkHp(cid:1)p
dρ
Now using the fact that fρ ∈ Rn, ρr ⊂ Rn, r for every ρ ∈ (0, 1), we get
Z 1
1
2
ρw(ρ)
In particular, using the right-hand side inequality of (1.2), we get
Cn, r (Lp
a (w)) ≤ Kp
n
1 − r
,
for all p ∈ [1, ∞), and β ∈ (−1, ∞), where Kp is a constant depending on p only.
Now, let us prove (1.4). Let
1
Lp
fn(z) =
kfnkp
and, analogously,
(1 − rz)n ∈ Rn,r,
and D = {z ∈ D : 1 − rz ≤ 21 − r}. We claim that
a(w) ∼ZD fn(z)pw(z)dA(z),
a(w) ∼ZD f ′
nkp
ZD\D fn(z)pw(z)dA(z) ≤
Indeed, by a very rough estimate
n(z)pw(z)dA(z),
kf ′
Lp
C1
2pn(1 − r)pn ,
n → ∞,
n → ∞.
where C1 > 0 depends only on w. On the other hand, if we put D = {z ∈ D :
1 − rz ≤ 3
21 − r}, then
Since r (thus D and D) are fixed we see that
w(z)dA(z).
1
ZD fn(z)pw(z)dA(z) ≥
2pn(1 − r)pn = o(cid:18)
a(w) ∼ZD f ′
nkp
kf ′
kfnkp
a(w)
Lp
Lp
1
(3/2)pn(1 − r)pnZ D
w(z)dA(z)(cid:19),
1
(3/2)pn(1 − r)pnZ D
n(z)pw(z)dA(z).ZD fn(z)pw(z)dA(z).
n → ∞.
Thus,
6
ANTON BARANOV AND RACHID ZAROUF
Obviously,
ZD f ′
n(z)pw(z)dA(z) =ZD
nprp
1
nprp
1 − rzpn+p w(z)dA(z)
2p(1 − r)pZD
1 − rzpn w(z)dA(z)
2p(1 − r)pZD fn(z)pw(z)dA(z).
nprp
≥
=
Thus,
lim inf
n→∞
kf ′
nkLp
nkfnkLp
a(w)
a(w) ≥
r
2(1 − r)
.
(cid:3)
For the proof of Theorem 1.4 we will need two lemmas.
Lemma 2.1. Let r ∈ [0, 1) and t ≥ 0. We set
I(t, r) =ZT 1 − rξ−t dm(ξ) and ϕr(t) =ZT 1 + rξt dm(ξ).
Then,
I(t, r) =
1
(1 − r2)t−1 ϕr(t − 2)
for every t ≥ 2, and t 7→ ϕr(t) is an increasing function on [0, +∞) for every
r ∈ [0, 1). Moreover, both
r 7→ ϕr(t − 2) and r 7→ I(t, r),
are increasing on [0, 1), for all t ≥ 0.
Proof. Indeed, supposing that t ≥ 2, we can write
I(t, r) =
1
1 − r2ZT b′
1
r(ξ)
1 − rξt−2 dm(ξ),
where br(z) = r−z
in the above integral we get
1−rz ). Using the fact that br ◦ br(z) = z and changing the variable
I(t, r) =
=
=
1
1 − rbr ◦ br(ξ)t−2 dm(ξ)
r(ξ)
1 − rbr(ξ)t−2 dm(ξ)
1
1
1
1
1 − r2ZT b′
1 − r2ZT
(1 − r2)t−1 ϕr(t − 2),
exp(cid:18) t
1−rz . Now,
= 1−r2
2
ln(cid:0)1 + r2 − 2r cos s(cid:1)(cid:19) ds,
0
1−rz
since 1 − rbr(z) = 1−rz−r(r−z)
ϕr(t) =Z 2π
ln(cid:0)1 + r2 + 2r cos s(cid:1) exp(cid:18) t
4Z 2π
exp(cid:18) t
4Z 2π
(cid:2)ln(cid:0)1 + r2 − 2r cos s(cid:1)(cid:3)2
r (t) =
r(t) =
and
ϕ′′
ϕ′
1
1
2
0
0
2
ln(cid:0)1 + r2 + 2r cos s(cid:1)(cid:19) ds,
ln(cid:0)1 + r2 − 2r cos s(cid:1)(cid:19) ds ≥ 0,
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL FUNCTIONS
7
for every t ≥ 0, r ∈ [0, 1). Thus, ϕr is a convex fonction on [0, ∞) and ϕ′
increasing on [0, ∞) for all r ∈ [0, 1). Moreover,
r is
Thus,
ϕ′
1
r(0) =
4Z 2π
ϕ′
r(t) ≥ ϕ′
0
ln(cid:0)1 + r2 − 2r cos s(cid:1) ds = 0.
r(0) = 0, ∀t ∈ [0, ∞), r ∈ [0, 1),
and so ϕr is increasing on [0, ∞). The fact that
r 7→ I(t, r),
is increasing on [0, 1) for all t ≥ 0 is obvious since
2
I(t, r) =(cid:13)(cid:13)(cid:13)(cid:13)
1
(1 − rz)t/2(cid:13)(cid:13)(cid:13)(cid:13)
H 2
=Xk≥0
a2
k(t)r2k,
where ak(t) is the kth Taylor coefficient of (1 − z)−t/2. The same reasoning gives
that r 7→ ϕr(t) is increasing on [0, 1).
(cid:3)
Lemma 2.2. If for some r0 ∈ [0, 1) and γ > 0 the function w(ρ)
on [r0, 1), then
(1−ρ2)γ is increasing
Z 1
r
ρw(ρ)I(t, rρ)dρ ≍Z 1
r0
ρw(ρ)I(t, rρ)dρ,
for all t such that t ≥ γ + 3 and for all r ≥ r0, with constants independent on t.
Proof. Clearly,
Moreover,
r
r0
ρw(ρ)I(t, rρ)dρ ≥Z 1
Z 1
ρw(ρ)I(t, rρ)dρ =Z 1
Z 1
ρw(ρ)I(t, rρ)dρ = Z r
Z r
r0
r0
r
and applying Lemma 2.1,
ρw(ρ)I(t, rρ)dρ,
r ∈ [r0, 1).
≤
≤
r0
r0
ρw(ρ)I(t, rρ)dρ,
ρw(ρ)
(1 − ρ2)γ
(1 − r2)γ Z r
w(r)
(1 − r2)γ ϕr2(t)Z 1
ρw(ρ)I(t, rρ)dρ +Z r
(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 ϕrρ(t)dρ
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 ϕrρ(t)dρ
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 dρ,
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 ϕrρ(t)dρ
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 dρ.
w(ρ)
(1 − ρ2)γ
(1 − r2)γ ϕr2 (t)Z 1
w(r)
w(r)
≥
r0
r0
r
because u 7→ ϕu(t) is increasing for all t > 0. For the same reason,
Z 1
r
ρw(ρ)
1
(1 − (rρ)2)t−1 ϕrρ(t)dρ = Z 1
r
8
ANTON BARANOV AND RACHID ZAROUF
Now note that
Z r
r0
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 dρ .Z 1
r
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 dρ,
r ∈ [r0, 1),
with constants independent on t ≥ γ + 3. Indeed, this estimate holds for t = γ + 3,
and, hence, by monotonicity of the function ρ 7→ (1 − (ρr)2)−1, for all t ≥ γ + 3.
Thus, using Lemma 2.1 and the fact that the function (1−ρ)−γw(ρ) is increasing
on [r0, 1), we obtain
Z r
r0
ρw(ρ)I(t, rρ)dρ ≤
r0
w(r)
w(r)
(1 − r2)γ ϕr2 (t − 2)Z r
(1 − r2)γ ϕr2 (t − 2)Z 1
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 dρ
ρ(cid:0)1 − ρ2(cid:1)γ
(1 − (rρ)2)t−1 dρ
(1 − (rρ)2)t−1 ϕrρ(t)dρ,
ρw(ρ)
1
r
≤ κ1
≤ κ2Z 1
r
(where κ1, κ2 are positive constants which do not depend on t), which completes
the proof.
(cid:3)
Proof of Theorem 1.4. We need to prove only the lower bound, the upper bound is
already proved in Theorem 1.3. Let us prove the minoration with the test function
f (z) =
(1−rz)n . Using (2.1) with α = r0, we need to show that
1
Lp
a(w)
nprp =Z 1
kf ′kp
r0
ρw(ρ)I(pn + p, rρ)dρ
C
(1 − r)pZ 1
r0
≥
ρw(ρ)I(pn, rρ)dρ =
C
(1 − r)p kfkp
a(w).
Lp
Since r ∈ [r0, 1) and n ≥ γ+3
this means that
p , by Lemma 2.2 applied with t = pn + p and t = pn
Z 1
r
ρw(ρ)I(pn + p, rρ)dρ ≥
C
(1 − r)pZ 1
r
ρw(ρ)I(pn, rρ)dρ.
By Lemma 2.1, this is equivalent to the estimate
Z 1
r
ρw(ρ)
ϕrρ(pn + p − 2)
(1 − (rρ)2)pn+p−1 dρ
≥
The last statement is obvious since
C
(1 − r)pZ 1
r
ρw(ρ)
ϕrρ(pn − 2)
(1 − (rρ)2)pn−1 dρ.
where bλ(z) = λ−z
1−λz
, λ ∈ D. Define the model space KBσ by
bλ,
Bσ = Yλ∈σ
KBσ =(cid:0)BσH 2(cid:1)⊥
and ek(z) =(cid:18)k−1Yj=1
= H 2 ⊖ BσH 2.
bλj (z)(cid:19) (1 − λk)1/2
1 − λkz
,
k ∈ [2, n],
Consider the family (ek)1≤k≤n in KBσ (known as Malmquist basis, see [13, p. 117]),
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL FUNCTIONS
9
Z 1
r
ρw(ρ)
ϕrρ(pn + p − 2)
(1 − (rρ)2)pn+p−1 dρ
(1 − r2)pZ 1
≥
1
r
ρw(ρ)
ϕrρ(pn + p − 2)
(1 − (rρ)2)pn−1 dρ
(1 − r2)pZ 1
1
r
ϕrρ(pn − 2)
(1 − (rρ)2)pn−1 dρ,
where the last inequality is due to the fact that t 7→ ϕu(t) is increasing for all
0 ≤ u < 1.
(cid:3)
ρw(ρ)
≥
3. The case of super-polynomially decreasing weights. Proof of
Theorem 1.5:
For the proof of Theorem 1.5 we will need a definition from the theory of model
subspaces of the Hardy space. For a finite subset σ of D with card σ = n, consider
the finite Blaschke product
e1(z) =
(1 − λ1)1/2
1 − λ1z
The family (ek)1≤k≤n associated with σ is an orthonormal basis of the n-dimensional
space KBσ .
In what follows we denote by Lp
a(w, sD) and by H p(sD), s > 0, the weighted
Bergman space and the Hardy space in the disc sD = {z : z < s}, respectively. If
w ≡ 1, we write simply Lp
Lemma 3.1. Let n ≥ 1, r, s ∈ [0, 1) and p ∈ [1, +∞]. We set
a(sD) and we write Lp
a if s = 1.
Mp, s(n, r) = supnf (ξ) : ξ ∈ D, f ∈ Rn, r, kfkLp
a(sD) ≤ 1o.
Then
(3.1)
Mp, 2
3
(n, r) ≤ d
cn
(1 − r)n+b ,
where d > 0, b > 0, c > 1 are some absolute positive constants (may be, depending
on p).
3 , but for every s ∈ (0, 1), with
Remark 3.2. Lemma 3.1 is valid not only for s = 2
constants d > 0, b > 0, c > 1 depending both on s and p.
Proof. For every f ∈ Rn, r and ξ ∈ D, we have
f 2
(cid:12)(cid:12)(cid:12)(cid:12)f(cid:18) 1
2
ξ(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)f 2
3(cid:18) 3
4
ξ(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)ZD
(u)(cid:16)k 3
4 ξ(u)(cid:17)2
3
dA(u)(cid:12)(cid:12)(cid:12)(cid:12) ,
10
ANTON BARANOV AND RACHID ZAROUF
where kλ(z) = 1
is the normalized area measure on D. Applying Holder's inequality we obtain
is the standard Cauchy kernel associated with λ ∈ D, and A
1−λz
,
ξ ∈ T,
where p′ is such that 1
(cid:12)(cid:12)(cid:12)(cid:12)f(cid:18) 1
2
3
a
a( 2
p + 1
kfkLp
p′ = 1. Now, note that
D)(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)k 3
4 ξ(cid:17)2(cid:13)(cid:13)(cid:13)(cid:13)Lp′
a(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)k 3
ξ(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:13)(cid:13)(cid:13)f 2
=(cid:18) 3
2(cid:19)1/p
3(cid:13)(cid:13)(cid:13)Lp
(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)k 3
4 ξ(cid:17)2(cid:13)(cid:13)(cid:13)(cid:13)H∞
4 ξ(cid:17)2(cid:13)(cid:13)(cid:13)(cid:13)Lp′
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)k 3
4(cid:19)2
=(cid:18) 1
1 − 3
2(cid:19)1/p
D) ≤ 16(cid:18) 3
D) ≤ 1, we obtain
D) ≤ kfkH∞( 1
kfkLp
a( 2
a( 1
a
2
2
3
≤ 24,
4 ξ(cid:17)2(cid:13)(cid:13)(cid:13)(cid:13)Lp′
a
= 16.
Finally, supposing kfkLp
which gives
(3.2)
Mp, 2
3
(n, r) ≤ 24M2, 1
2
(n, r).
It remains to obtain a suitable upper bound for M2, 1
2
(n, r). Let us prove that
(3.3)
M2, 1
2
(n, r) ≤ 2√n(cid:18) 2
1 − r(cid:19)n+ 1
2
.
2 ∈ Rn, 1
2 r ⊂ Rn, r. If {1/λ1, . . . , 1/λn} is the set of
For every f ∈ Rn, r, we have f 1
the poles of f (thus, λj ≤ r, j = 1, . . . , n), then f ∈ KBσ with σ = {λ1, . . . , λn} ⊂
rD), whereas the set {2/λ1, . . . , 2/λn} is the set of the poles of the function f 1
and
D. Hence, there exist α1, . . . , αn ∈ C
f 1
such that
2 ∈ KBσ′ with σ′ = (cid:8) 1
2 λ1, . . . , 1
2
2
2 λn(cid:9) ⊂ r
nXk=1
f 1
=
2
αkek,
(3.4)
andPn
in C. Thus,
on D, where (ek)n
k=1 is the Malmquist basis associated with the set σ′. Since both f 1
k=1 αkek are meromorphic in C the equality (3.4) is in fact valid everywhere
2
λj
f (ξ) =
2 − 2ξ
nXk=1
and by the Cauchy -- Schwarz inequality,
αk
1 − λj ξ
k−1Yj=1
αk2(cid:17)1/2 nXk=1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18)k−1Yj=1
f (ξ) ≤(cid:16) nXk=1
4 − ξ(cid:1)
2(cid:0) λ
1 − λ
4 ξ
for any ξ ∈ D. Now, if λ ∈ rD and ξ ∈ D,
1 − λ
4 ξ
1 − λξ
2 − 2ξ
1 − λξ
(3.5)
=
λ
(cid:16)1 − 1
4 λk2(cid:17)1/2
1 − λjξ
,
ξ ∈ D,
λj
2 − 2ξ
1 − λjξ
1 − λjξ(cid:19)(cid:0)1 − 1
(ξ)(cid:18)1 +
4λk2(cid:1)1/2
4(1 − λξ)(cid:19),
3λ
4
= 2b λ
2!1/2
.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL FUNCTIONS
11
22(k−1)(cid:18) 1
1 − r(cid:19)2(k−1)
2
1 − r
.
which gives
2
λ
1
1
λj
≤
(3.6)
3r
4
We get
2 − 2ξ
2 − 2ξ
1 − λjξ
(cid:18)k−1Yj=1
nXk=1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
4λk2(cid:1)1/2
(cid:12)(cid:12)(cid:12)(cid:12)
1 − λξ(cid:12)(cid:12)(cid:12)(cid:12) ≤ 2(cid:18)1 +
1 − λj ξ(cid:19)(cid:0)1 − 1
1 − r(cid:19) =
4 − r
2(1 − r) ≤
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXk=1
4(cid:18) 2
1 − r(cid:19)2n+1
2(cid:13)(cid:13)(cid:13)H 2
=(cid:13)(cid:13)(cid:13)f 1
αk2(cid:17)1/2
For any function ϕ(z) =Pk≥0bϕ(k)zk in H 2, one has
√k + 1 bϕ(k) ≤ kϕkL2
(1 − r)2
1
Now we first notice that
≤
.
.
a kϕkB
1/2
2, 2
,
H 2 =Xk≥0
kϕk2
kϕk2
1/2
2, 2
B
We now use the upper bound of [21, Theorem A, (4)]: for ϕ ∈ Rn, ρ one has
√k + 1
(cid:16) nXk=1
bϕ(k)
= kϕ′k2
≤ (2 + r)
≤
4n
L2
a
H 2
+ kϕk2
n
1 − rkϕk2
H 2 + kϕk2
H 2
which gives
In particular, with ϕ = f 1
(3.7)
that is,
2
a
.
H 2 ,
we get ϕ ∈ Rn, 1
1 − ρ kϕkL2
2 r and
1 − rkϕk2
kϕkH 2 ≤ 2r n
2(cid:13)(cid:13)(cid:13)H 2 ≤ 2r n
a ≤ 2√2n(cid:13)(cid:13)(cid:13)f 1
(cid:13)(cid:13)(cid:13)f 1
2(cid:13)(cid:13)(cid:13)L2
2(cid:13)(cid:13)(cid:13)L2
1 − r/2(cid:13)(cid:13)(cid:13)f 1
2(cid:13)(cid:13)(cid:13)H 2(cid:18) 1
2(cid:18) 2
1 − r(cid:19)n+ 1
(1 − r)2n+1(cid:19) 1
2√2n(cid:13)(cid:13)(cid:13)f 1
2(cid:13)(cid:13)(cid:13)L2
f (ξ) ≤(cid:13)(cid:13)(cid:13)f 1
√2n(cid:18) 2
1 − r(cid:19)n+ 1
f (ξ) ≤
kfkL2
22n+1
≤
4
1
.
a
2
2
2
,
a
We conclude from (3.5), (3.6) and (3.7) that for any ξ ∈ D,
a( 1
D) ,
Combining (3.2) and (3.3) and choosing d = 48, b = 1
ξ ∈ D.
Taking the supremum over ξ ∈ D and f ∈ Rn, r we obtain (3.3).
2n√n ≤ cn for any n ≥ 1, we complete the proof and obtain (3.1).
Proof of Theorem 1.5. Take r ∈ (0, 1) and R ∈ (0, r) and let us represent the norm
kf ′kp
a(w) of a function f ∈ Rn, r as I1 + I2,
Lp
2 and c > 2 such that
(cid:3)
2
I1 =Z R
0 k(fρ)′kp
pw(ρ)dρ,
I2 =Z 1
R k(fρ)′kp
pw(ρ)dρ.
where the last inequality follows from Lemma 3.1. Note that
p, 2/3(n, r)w(ρ)dρ
R
3
3
Lp
a( 2
M p
∞w(ρ)dρ
I2 ≤ C3(cid:16) n
≤ C3kfkp
≤ C3kfkp
1 − r(cid:17)pZ 1
R kfρkp
1 − r(cid:17)pZ 1
D)(cid:16) n
1 − r(cid:17)p
D)(cid:16) n
kfkp
D) ≤ (w(2/3))−1kfkp
I2 ≤ C4(cid:16) n
1 − r(cid:17)p
cpn
cpn
a( 2
Lp
a( 2
Lp
3
(1 − r)pn+pb w(R),
a(w).
Lp
(1 − r)pn+pb w(R)kfkp
a(w).
Lp
Hence,
12
ANTON BARANOV AND RACHID ZAROUF
Here and everywhere below in this proof, Ci, i = 1, . . . , 5, are positive constants,
depending, may be, only on p and w (but not on n and r). By (1.2), we have for
the first integral
I1 ≤ C1(cid:16) n
1 − R(cid:17)pZ R
0 kfρkp
pw(ρ)dρ ≤ C2(cid:16) n
1 − R(cid:17)p
a(w).
Lp
kfkp
(n, r)kfρkLp
D). Ap-
Note that fρ ∈ Rn, ρr ⊂ Rn, r, and, thus, kfρk∞ ≤ Mp, 2
plying (1.2) once again together with an obvious inequality kfρkp ≤ kfρk∞, we
get
a( 2
3
3
Now, choose a positive increasing sequence (γn)n∈N such that n = o (γn), as n →
+∞. For any n we fix r◦
n, 1).
Now for a fixed n take r, R so that r◦
1 − R = (1 − r)1/2,
n such that the function w(r)(1− r)−γn decreases on [r◦
n = (1 − r)1/4.
n < R < r < 1 and
1 − r◦
Hence, using the fact that w is bounded on [r◦
1, 1), we obtain
We have
w(R) ≤ w(r◦
n)
I2 ≤ C4(cid:16) n
1 − r(cid:17)p
kfkp
Let us show that for sufficiently large n,
cpn (1 − r)γn/4
(1 − R)γn
n)γn
(1 − r◦
= w(r◦
n)(1 − r)γn/4.
Lp
a(w) · cpn (1 − r)γn/4
(1 − r)pn+pb .
(1 − r)pn+pb → 0,
r → 1 − .
Indeed, choosing r so that c < (1 − r)−1, we get
cpn (1 − r)γn/4
(1 − r)pn+pb ≤ (1 − r)
= o(cid:18) 1
I 1/p
2
nkfkLp
a(w)
1 − rn(cid:19),
n → ∞.
since n = o(γn), n → ∞. Hence, there exists a sequence (rn), rn → 1−, such that
γn
4 −2pn−pb → 0,
r → 1−,
The corresponding estimate for I1 is obvious since 1 − Rn = (1 − rn)1/2.
(cid:3)
A BERNSTEIN-TYPE INEQUALITY FOR RATIONAL FUNCTIONS
13
References
1. A.D. Baranov, Bernstein-type inequalities for shift-coinvariant subspaces and their
applications to Carleson embeddings, J. Funct. Anal. 223 (2005), no. 1, 116 -- 146.
2. A.D. Baranov, Embeddings of model subspaces of the Hardy space: compactness and
Schatten -- von Neumann ideals, Izvestia RAN Ser. Matem. 73 (2009), no. 6, 3 -- 28;
English transl.: Izv. Math. 73 (2009), no. 6, 1077 -- 1100.
3. P. Borwein and T. Erd´elyi, Polynomials and Polynomial Inequalities, Springer, New
York, 1995.
4. V.I. Danchenko, An integral estimate for the derivative of a rational function, Izv.
Akad. Nauk SSSR Ser. Mat. 43 (1979), no. 2, 277 -- 293; English transl. Math. USSR
Izv., 14 (1980), no. 2, 257 -- 273.
5. E.P. Dolzhenko, Rational approximations and boundary properties of analytic func-
tions, Mat. Sb. (N.S.), 69(111):4 (1966), 497-524.
6. K.M. Dyakonov, Kernels of Toeplitz operators, smooth functions, and Bernstein-type
inequalities, Zap. Nauchn. Semin. S. Peterburg. Otdel. Mat. Inst. Steklov. (POMI)
201 (992), 5 -- 21; English transl.: J. Math. Sci. 78 (1996), 131 -- 141.
7. K.M. Dyakonov, Differentiation in star-invariant subspaces I. Boundedness and com-
pactness, J. Funct. Anal. 192 (2002), 364 -- 386.
8. K.M. Dyakonov, Smooth functions in the range of a Hankel operator, Indiana Univ.
Math. J. 43 (1994), 805 -- 838.
9. T. Erd´elyi, George Lorentz and inequalities in approximation, Algebra i Analiz 21
(2009), no. 3, 1 -- 57.
10. A.A. Gonchar, Degree of approximation by rational fractions and properties of func-
tions, Proc. Internat. Congr. Math. (Moscow, 1966), Mir, Moscow, 1968, 329 -- 356;
English transl.: Amer. Math. Soc. Transl. (2) 91 (1970).
11. H. Hedenmalm, B. Korenblum, and K. Zhu, Theory of Bergman spaces, Graduate
Texts in Mathematics, 199, Springer-Verlag, New York, 2000.
12. R.J. Leveque, L.N. Trefethen, On the resolvent condition in the Kreiss matrix Theo-
rem, BIT 24 (1984), 584 -- 591.
13. N. Nikolski, Treatise on the Shift Operator, Springer-Verlag, Berlin, 1986.
14. A.A. Pekarskii, Inequalities of Bernstein type for derivatives of rational functions,
and inverse theorems of rational approximation, Mat. Sb. 124(166) (1984), 571-588;
English transl.: Math. USSR-Sb. 52 (1985), no. 2, 557 -- 574.
15. V.V. Peller, Hankel operators of class Sp and their applications (rational approxima-
tions, Gaussian processes, the problem of majorizing operators), Mat. Sb. 113(155)
(1980), no. 4, 538 -- 581; English transl.: Math. USSR Sb. 41 (1982), no. 4, 443 -- 479.
16. V.N. Rusak, Rational Functions as Approximation Apparatus, Izdat. Beloruss. Gos.
Univ., Minsk, 1979 (Russian).
17. M.N. Spijker, On a conjecture by LeVeque and Trefethen related to the Kreiss matrix
theorem, BIT 31 (1991), 551 -- 555.
18. R. Zarouf, Sharpening a result by E.B. Davies and B. Simon, C. R. Acad. Sci. Paris
347 (2009), no. 16, 939 -- 942.
19. R. Zarouf, Analogs of the Kreiss resolvent condition for power bounded matrices,
Actes des journ´ees du GDR "Analyse Fonctionnelle Harmonique et Applications",
Metz, 2010.
20. R. Zarouf, Asymptotic sharpness of a Bernstein-type inequality for rational functions
in H 2, Algebra i Analiz 23 (2011), no. 2, 147 -- 161.
21. R. Zarouf, Application of a Bernstein type inequality to rational interpolation in the
Dirichlet space, Zap. Nauchn. Semin. S. Peterburg. Otdel. Mat. Inst. Steklov. (POMI)
39:101 -- 112, (2011).
22. R. Zarouf, Effective H∞ interpolation constrained by Hardy and Bergman weighted
norms, Ann. Funct. Anal., 2(2):59 -- 74 (2011).
23. R. Zarouf, Effective H∞ interpolation, to appear in Houston J. Math.
24. K. Zhu, Operator theory in function spaces, Monographs and Textbooks in Pure and
Applied Mathematics, 139. Marcel Dekker, Inc., New York, 1990.
14
ANTON BARANOV AND RACHID ZAROUF
Department of Mathematics and Mechanics, Saint Petersburg State University, 28,
Universitetski pr., St. Petersburg, 198504, Russia
E-mail address: [email protected]
CMI-LATP, UMR 6632, Universit´e de Provence, 39, rue F.-Joliot-Curie, 13453 Mar-
seille cedex 13, France
E-mail address: [email protected]
|
1003.2224 | 1 | 1003 | 2010-03-10T23:09:52 | A characterization of separable conjugate spaces | [
"math.FA"
] | Necessary and sufficient conditions for a separable Banach space to be a dual space are proved. Some applications are discussed | math.FA | math |
A CHARACTERIZATION OF SEPARABLE CONJUGATE
BANACH SPACES
By STEFANO ROSSI
Abstract The following elegant characterization of dual Banach spaces will
be proved: a separable Banach space X is a dual space if and only exists a
norm closed subspace M ⊂ X∗ with (Dixmier) characteristic equal to 1, whose
functionals are norm-attaining.
Contents
1 Introduction
2 Notations and preliminaries
3 A general result
4 The separable case
5 Applications
1
Introduction
1
2
2
4
7
Among all Banach spaces, dual spaces have some additional properties, which
assure a more far-reaching treatment of their structure. Probably the most
useful property is represented by weak*-compactness of the unit ball of such
spaces. Giving necessary and sufficient conditions for a Banach space X to be
a dual space is a long standing problem; it goes back essentially to J. Dixmier
[3]. Many theorems concerning this intriguing area are known; these results
usually require the compactness property for X1, the unit ball of X, with re-
spect to suitable locally convex topology. Elegant as theoretical results, these
theorems, however, do not provide any practical tool to solve the question in
the applications, because compactness property is very difficult to be checked
in infinite-dimensional contexts.
In spite of this, reflexive Banach spaces (which are indeed dual spaces with
unique predual) are well understood thanks to a celebrate thorem by R. C.
James [4], which states that a Banach space X is reflexive if and only if every
continuous linear functional attains its norm on X1.
The result quoted above can be regarded as a straightforward application of the
following powerful characterization of weak compactness, which is due as well
to James:
1
Theorem 1 (James, [6]). Let Y be a complete locally convex space and C ⊂ Y
a bounded, σ(Y, Y∗) closed subset The following are equivalent:
1. C is a weakly compact subset.
2. Every ϕ ∈ Y∗ attains its sup on C,i.e. for every ϕ ∈ Y∗ there exists c ∈ C
such that ϕ(c) = supy∈C ϕ(y).
The proof of the theorem announced in the abstract essentially depends on
James ’ theorem 1, where completeness assumption unfortunately cannot be
removed, as James himself has shown in [5].
2 Notations and preliminaries
If X is a Banach space, X1 will stand for its unit ball, that is
X1 = {x ∈ X : kxk ≤ 1}
The dual (conjugate) space of X will be denoted by X∗. A Banach space X
is a dual space if there exists a Banach space Y such that Y∗ ∼= X (isometric
isomorphism); in this case Y is said to be a predual.
Accordig to Dixmier, we define the characteristic of a subspace M ⊂ X∗ as the
the real number given by
c(M)
.
= inf
x6=0(cid:26) sup
ϕ∈M1
ϕ(x)
kxk (cid:27)
When c(M) = 1, we say that M is a determinant subspace of the conjugate
space; a standard application of the bipolar theorem shows that M is determi-
nant iff M1 is weakly* dense in X∗
1.
Finally, we recall that a continuous linear functional ϕ ∈ X∗
1 is said to be norm-
attaining if there exist x ∈ X1 such that ϕ(x) = kϕk. A subspace M ⊂ X∗ is
norm-attainig if each ϕ ∈ M is a norm-attaining linear functional.
3 A general result
This section is devoted to present a general result for the existence of preduals.
In what follows, we will need some tools from the general theory of locally convex
space, in particular some basic facts belonging to the theory of dual pair. For an
extensive treatment see, for istance, [8]. For completeness’ sake, here we recall
that, given a dual pair (X, Y), the Mackey topology on X is the locally convex
topology described by the base of seminorms pK given by
pK(x) = supy∈Khx, yi
where K is a σ(Y, X)-compact, (circled) convex subset of Y. The Mackey
topology is often denoted by τ (X, Y). Analogously, one can define the Mackey
topology τ (y, X) on Y.
2
We begin our analysis by observing that any conjugate space X∗ is a complete
locally convex space with the Mackey topology τ (X∗, X). This can be proved,
for instance, with the aid Grothendieck completeness theorem or as an easy
application of the lemma 2 from section 4 of the present paper. A second
observation needed is that any predual of a Banach space M should be sought
as a suitable subspace of the conjugate space X∗. More precisely, one has the
following:
Lemma 1. Let X and M be Banach spaces.
If the isometric isomorphism
M∗ ∼= X holds, then there is an isometric injection i : M → X∗ such that i(M)
is a determinant subspace, whose functionals are norm-attaining.
Proof. Let Φ : X → M∗ be an isometric isomorphism. Taking the (Banach)
adjoint of Φ, we get an isometric isomorphism Φ∗ : M∗∗ → X∗. Let i : M → X∗
the linear isometry given by the composition Φ∗ ◦ j, where j is the canonical
injection of M into its second conjugate space M∗∗. It only remains to prove
that i(M) has the required properties.
Let x ∈ M, then we have
kxk = kΦ(x)k = sup
m∈M1
hΦ(x), mi = sup
m∈M1
hi(m), xi
The last equality says that i(M) is determinant.
In order to prove that i(M) is norm attaining, observe that, given any m ∈ M,
one has ki(m)k = kmk = ϕ(m), for some ϕ ∈ M∗
1 by virtue of the Hanh-
Banach theorem. Now, if x be is the unique element of X such that Φ(x) = m,
we have ki(m)k = hi(m), xi. This concludes the proof, since kxk = kΦ(x)k =
kmk ≤ 1.
Under suitable additional requests the previous lemma can be reverted to
get existence of a predual, as indicated in the following result (due to Dixmier
[3]), which can be regarded as a partial convers of Alaoglu theorem. For the
completeness’ sake, we provide a different prove based on the Krein-Smulian
theorem.
Theorem 2. Let M ⊂ X∗ be a norm closed determinant subspace, such that
X1 is σ(X, M)- compact. Then M∗ ∼= X, through the isometric isomorphism
Φ : X → M∗ given by hΦ(x), ϕi = ϕ(x) for all ϕ ∈ M∗ and for all x ∈ X.
Proof. The linear map Φ : X → M∗ described in the statement is surely isomet-
ric, since M is determinant. Let N ⊂ M∗ be its range. N is weakly* dense in
M∗, since it is a total subspace, i.e. N⊥ = 0 (in the duality (M∗, M)). To get
the conclusion, it is enough to observe that N is weakly* closed by virtue of the
Krein-Smulian theorem (here is the point werw completeness of M is needed).
In fact N1 = Φ(X1) is weakly* compact as the continuous image of a compact
set (Φ is apparently a σ(X, M) − σ(M∗, M) continuous map).
Actually the theorem remains true under the milder assumption that M is
just a total subspace, see [7], anyway Theorem 2 is enough to state and prove
the following general characterization of dual Banach space:
Theorem 3. Let X be a Banach space. The following conditions are equivalent:
3
1. X is a dual space.
2. There exists a norm closed subspace M ⊂ X∗ which is determinant, norm-
attaining and such that X is complete with respect to the Mackey topology
τ (X, M).
Moreover, each subspace as in 2. is canonically1 a predual of X .
Proof. The implication 1. ⇒ 2. has been already proved. Conversely, if 2.
holds, we can equip X with the Mackey topology τ (X, M). By the Mackey-
Arens theorem, the Mackey dual of X is M. The unit ball X1 ⊂ X is clearly
Mackey-bounded since it is even norm bounded; moreover X1 is σ(X, M)-closed.
In fact, thanking to the fact that M is determinant, we can rewrite X1 as
{x ∈ X : ϕ(x) ≤ 1 for all ϕ ∈ M1}
so it is a σ(X, M)- closed subset as intersection of σ(X, M)-closed sets. By
the assumption of completeness, James’theorem applies, so we get σ(X, M)-
compactness of X1. Hence M is canonically a predual of X.
Remark 1. Unfortunately completeness hypothesi is not omissible, as the fol-
lowing counterexample shows. Let us consider the Banach space l1[0, 1] of (real)
function f defined on [0, 1] such that Px∈[0,1] f (x) < ∞, endowed with the cor-
responding k · k1- norm. Its conjugate space is clearly represented by B[0, 1],
the space of all bounded function over [0, 1] with the sup norm, acting by
hg, f i = Px∈[0,1] g(x)f (x) for all g ∈ B[0, 1], f ∈ l1[0, 1]. The norm closed
subspace of continuous C[0, 1] ⊂ B[0, 1] is determinant and norm-attaining.
However it is not a predual, since C[0, 1]∗ ∼= M([0, 1]) (the space of all finite
Borel measures on [0,1]), while the inclusion2 l1[0, 1] ⊂ M([0, 1]) is proper. Ac-
tually the trouble is that l1[0, 1] is not complete under the Mackey topology
τ (l1[0, 1], C[0, 1]), as pointed out in [1].
Although elegant, the previous theorem is not completely satisfactory, be-
cause to handle completeness notions in the general context of locally convex
spaces is rather difficult; however the result can be considerably strengthened
in the separable case, as it is shown in the next section.
4 The separable case
In this section we devote our analysis to separable Banach spaces X, the rea-
son being that separability assumption allows us to prove the following useful
completeness result:
Theorem 4. Let X be a separable Banach space and M ⊂ X∗ a determinant
subspace. X is a complete locally convex space with respect to the Mackey
topology τ (X, M).
1 M∗ ∼= X through the isometric isomorphism Φ : X → M∗ given by the restriction on M
of the canonical injection j : X → X∗∗, i.e. hΦ(x), ϕi = ϕ(x) for each ϕ ∈ M.
2One think f ∈ l1[0, 1] as the Borel measure Px∈[0,1] f (x)δx.
4
The proof of theorem 4 relies on the following two lemmas:
Lemma 2. Let X be topological vector space. If {fα : α ∈ I} ⊂ X ∗ is a net of
continuous linear functionals converging to f uniformly over the compact subset
of X, then f is sequentially-continuous.
Proof. It is a standard ε
3 -argument. Let {xn : n ∈ N} ⊂ X be a sequence
.
convergin to x ∈ X. The set K
= {xn : n ∈ N} ∪ {x} is compact, thus fα
converges to f uniformly on K. This means that, given ε > 0, there exists
α0 ∈ I such that
fα0 (y) − f (y) <
sup
y∈K
By the triangle inequality we get
f (xn) − f (x) ≤ ε
ε
3
for all n ≥ Nε, where Nε is any natural number such that fα0(xn)−fα0(x) ≤ ε
3
for all n ≥ Nε.
The second lemma deals with the Mazur property of M with respect the
σ(M, X)-topology. More precisely, we have the following statement:
Lemma 3. Let X be a separable Banach space. If M ⊂ X ∗ is a determinant
subspace, then M equipped with the σ(M, X)-topology is a Mazur space, that is
every sequantially-continuous linear form f : M → C is continuous with respect
to the σ(M, X)-topology.
Proof. Since X is separable, the weak* topology on X1 is metrizable as well its
restriction to M1. Let f : M → C be a σ(M, X) sequentially-continuous linear
form. The restriction f ↾M1 is thus a σ(M, X)- (uniformly) continuous function.
Since M is determinant, M1 is weakly* dense in X∗
1, so f ↾M1 extends to a
weakly* continuous function g : X∗
1 → C, which is apparently the restricton to
1 of a linear form G defined on the whole X∗. By the Krein-Smulian theorem
X∗
(see, for instance, [2]), this linear functional is weakly*-continuous, that is there
is a unique x ∈ X such that G(ϕ) = ϕ(x) for all ϕ ∈ X∗. This concludes the
proof.
Finally we can prove theorem 4.
Proof. Let {xα : α ∈ I} ⊂ X be a τ (X, M)- Cauchy net. This means that, given
any3 σ(M, X)-compact subset K ⊂ M and any ε > 0, we have
ϕ(xα − xβ) < ε
sup
ϕ∈K
(1)
for all α, β ∈ I such that α, β (cid:23) α0 for some α0 ∈ I. In particular, fixed ϕ ∈ M,
{ϕ(xα) : α ∈ I} is a numerical Cauchy net, so we can define a linear form on
.
= limα ϕ(xα). Thanks to inequality 1, the limit is uniform
M, say G, by G(ϕ)
3One should note that if K ⊂ M is σ(M, X)-compact set, then conv(K)
σ(M,X)
is yet
σ(M, X) compact thank to Alaoglu theorem.
5
over the σ(M, X)-compact subset of M. By lemma 2, G is sequentially σ(M, X)-
continuous, so it continuous in this topology thanks to the lemma 3. This means
that there is a unique x ∈ X such that G(ϕ) = ϕ(x) for all ϕ ∈ M. Taking the
limit (with respect to β ∈ I) of the inequality 1, one easily get xα → x in the
τ (M, X)-topology. This concludes our proof.
Finally, we can state our main theorem, as announced in the abstract:
Theorem 5. Let X be a separable Banach space. Every determinant norm
closes subspace M ⊂ X∗, whose functional are norm-attaining, is canonically a
predual of X.
Proof. Theorem 3 of the previous secrion applies, since X is complete with
respect the Mackey topology τ (X, M), by virtue of completeness theorem 4.
Observation 1. The assumption that M is norm closed cannot be dropped.
Let us consider, for instance, the separable Banach space C[0, 1] of continuous
function on the compact interval [0, 1]. Let us consider the subspace N ⊂
C[0, 1]∗, algebraically spanned by the set {δx : x ∈ [0, 1]}, where δx is the Dirac
measure concentrated on x ∈ [0, 1].
It is not difficult to check that N is a
determinant subspace, whose functionals are norm-attaining. Nevertheless N
cannot be a predual of C[0, 1], since the last space has not preduals at all, as it
is well known. Observe that N is not norm-attaining.
The following corollaries are almost immediate.
Corollary 1. Let X be a separable Banach space. If M ⊂ X∗ is a norm closed
subspace as in theorem 5, then it is a maximal subspace with respect the property
to be norm closed and norm-attaining.
Proof. If N ⊇ M is norm closed and norm-attaining, then it is (canonically) a
predual of X as well as M, hence N = M, since no proper inclusion relationship
among preduals are allowed, thanks to the theorems of Alaoglu-Banach and
Krein-Smulian.
In particular we get the following non-trivial result:
Corollary 2. If X is a (separable) Banach space, such that X∗ is separable,
then j(X) ⊂ X∗∗ is a closed maximal norm-attaining subspace.
As concrete applications of the corollaries stated above, we get, for instance,
that c0 ⊂ l∞ is a closed subspace, which is maximal with respect to the property
to be norm-attaining (on l1), since c0 is a (not unique) predual of l1.
The same is true for K(H) ⊂ B(H), where K(H) is the norm closed ideal of all
compact operators on the separable Hilbert space H, thought as a predual of S1,
the nuclear (trace-class) operators on H.
6
5 Applications
To appreciate theorem 5, we illustrate two nice applications of it, simplifying
the proofs of two classical results.
The first is the Riesz-Frechet representation theorem for the dual of a (separa-
ble) Hilbert space, the second one is the Riesz representation theorem for the
dual spaces of the Lebesgue spaces Lp (p > 1) in the separable case.
Theorem 6 (Riesz-Frechet). Let H be a separable Hilbert space, with inner
product (·, ·). If ϕ ∈ H∗, then there is a unique y ∈ H such that ϕ(x) = (x, y)
for each x ∈ H.
Proof. There is no loss of generality in the assumption that H is a real Hilber
space; in this case the map Φ : H → H∗, given by hΦ(y), xi = (x, y), is lin-
ear and isometric. Note that Φ(H) ⊂ H∗ is norm closed, determinant and
norm-attaining. This means that Φ(H) is canonically a predual of H, i.e. Φ is
surjective.
The application to duality theory for Lp spaces is more interesting, since the
usual proofs requires rather involved argument invoking Clarkson inequality or
Radon-Nikodym theorem. We will consider a measure space (X, S, µ), with the
property that S is countably generated: this assures that the corresponding
Lebesgue spaces Lp(X, µ) are separable for all p ≥ 1 (for the exception of the
value p = ∞ ). Given p > 1, we denote by q > 1 its conjugate exponent, that
is the real number q such that 1
p + 1
q = 1. With these notations, the classical
Riesz theorem can be stated as follows
Theorem 7. If p > 1 is a real number, then the map Φ : Lq(X, µ) → Lp(X, µ)∗
given by hΦ(f ), gi = RX f gdµ for all f ∈ Lq(X, µ), g ∈ Lp(X, µ) , is an isometric
Proof. Let Ψ : Lp(X, µ) → Lq(X, µ)∗ be the linear map give by
isomorphism .
hΨ(f ), gi
.
= ZX
f gdµ for all f ∈ Lp(X, µ), g ∈ Lq(Xµ)
By Holder inequality one has kψ(f )k ≤ kf kp for all f ∈ Lp(X, µ). Given
f ∈ Lp(X, µ), let us consider the measurable function g given by
g(x) = (cid:26) f (x)p−2f (x)
0 if f (x) = 0
p
p, it follows that hΨ(f ), hi = kf kp, where h
It is easy to check that g ∈ Lq(X, µ) and that kgkq = kf kp
. Since
one has hΨ(f ), gi = kf kp
∈
Lq(X, µ)1. By virtue of the last equality, we conclude that kΨ(f )k = kf kp
and the norm of the functional is attained on the function h. This means that
Ψ(Lp(X, µ)) is a closed subspace of Lq(X, µ)∗, whose functionals are norm-
attaining. A similar argument as above also shows that Ψ(Lp(X, µ)) is deter-
minant. By theorem 5, we argue that Ψ(Lp(X, µ)) is canonically a predual of
Lq(X, µ),i.e. the isometric isomorphism Lp(X, µ)∗ ∼= Lq(X, µ) holds as in the
statement.
p
q = kf kp−1
.
= g
kgkq
7
Remark 2. Our proof of Riesz representation theorem only depend on Holder
inequality.
I wish to thank S. Doplicher for some useful discussions during the peparation
of the present paper.
References
[1] J. Bonet, B. Cascales, Non complete Mackey topology on Banach spaces,
Bull. Australian Matth. Soc
[2] , J. B. Conway, A Course in Functional Analysis, Springer, GraduatTexts
in Mathematics, 1989 (2nd edition).
[3] J. Dixmier, Sur un theoreme de Banach, Duke Math. J. 15, 1057-1071, 1948.
[4] R. C. James, Characterization of reflexivity, Studia Math. 23, 205-216, 1964.
[5] R. C. James, A counterexample for a sup theorem in normed space, Isreael
J. Math 9, 511-512, 1971.
[6] R. C. James, Reflexivity and the sup of linear functionals, Israel J. Math 13,
289-300, 1972.
[7] S. Kaijser, A note on dual Banach spaces, Math. Scand. 41, 325-330, 1977.
[8] A. Wilansky, Modern Methods in Topological Vector Spaces, McGraw-Hill
International Book Company, 1978.
DIP. MAT. CASTELNUOVO, UNIV. DI ROMA LA SAPIENZA, ROME, ITALY
E-mail address: [email protected]
8
|
1710.00702 | 1 | 1710 | 2017-10-02T14:52:30 | p-Riesz bases in quasi shift invariant spaces | [
"math.FA"
] | Let $ 1\leq p< \infty$ and let $\psi\in L^{p}(\R^d)$. We study $p-$Riesz bases of quasi shift invariant spaces $V^p(\psi;Y)$. | math.FA | math |
p-Riesz bases in quasi shift invariant spaces
Laura De Carli, Pierluigi Vellucci
Abstract. Let 1 ≤ p < ∞ and let ψ ∈ Lp(Rd). We study p−Riesz
bases of quasi shift invariant spaces V p(ψ; Y ).
1. Introduction
Let 1 ≤ p < ∞ and let ψ ∈ Lp(Rd). We consider the shift invariant
space V p(ψ) = Span{τkψ}k∈Zd , where τsf (x) = f (x + s) is the translation
and "bar" denotes the closure in Lp(Rd). Shift-invariant spaces appear nat-
urally in signal theory and in other branches of applied sciences. In [2] [13]
and in the recent preprint [14] quasi-shift invariant spaces of functions are
considered. Given X = {xj}j∈Zd, a countable and discrete1 subset of Rd
and a function ψ ∈ Lp(Rd), we let
(1.1)
V p(ψ; X) = Span{τxj ψ}.
Thus, V p(ψ) = V p(ψ; Zd). Quasi-shift invariant spaces are also called
Spline-type spaces in [9], [10] [11], [23].
Following [2], [7], we say that the translates {τxj ψ}j∈Zd form a p-Riesz
basis in V p(ψ; X) if there exist constants A, B > 0 such that, for every
finite set of coefficients ~d = {dj} ⊂ C,
(1.2)
Ak ~dkℓp ≤ kXj
dj τxj ψkp ≤ Bk ~dkℓp.
Here and throughout the paper, we have let kf kp = (cid:0)RRd f (x)pdx(cid:1) 1
k~ckℓp = (Pj∈Zd cjp)
p and
p . If x = (x1, ..., xd), y = (y1, ..., yd) ∈ Rd, we will
1
2 . We will also let
often let x · y = x1y1 + ... + xdyd and x2 = (x · x)
x∞ = sup1≤j≤d xj.
1
1 A countable set X ⊂ Rd is discrete if for every xj ∈ X there exists δj > 0 such that
xj − xk2 > δj for every k 6= j.
1
2
LAURA DE CARLI, PIERLUIGI VELLUCCI
If (1.2) holds, then it is possible to prove that
(1.3)
2V p(ψ; X) = {f = Xk∈Zd
dkτxk ψ(x), ~d ∈ ℓp }
and the sequence {dk}k∈Zd is uniquely determined by f .
p−Riesz bases allow a stable reconstruction of functions in V p(ψ; X);
when X = Zd and B = {τj ψ}j∈Zd is a p−Riesz basis of V p(ψ), the coefficient
dj in (1.3) can be expressed in an unique way in terms of the functions in
the dual basis of B. See [30], [5] and also [1] for explicit reconstruction
formulas.
for every y ∈ [− 1
2 , 1
When ψ has compact support, it is known (see e.g. [1, Prop. 1.1], [16],
[24]) that B is a p−Riesz basis in V p(ψ) if and only ifPm∈Zd ψ(y +m)2 6= 0
We have denoted with ψ(y) =RRd e2πix·yf (x)dx the Fourier transform of
2 )d and every m ∈ Zd. See also Lemma 5 in Section 2.
ψ. The proof of the aforementioned result relies on the lattice structure of
Zd and on standard Fourier analysis technique and does not easily generalize
to other sets of translations.
Let ψ ∈ Lp(Rd), 1 ≤ p < ∞, and let X = {xj}j∈Zd be a discrete set of
Rd. It is natural to consider the following problem:
Problem 1. Let BX = {τxj ψ}j∈Zd be a p−Riesz basis for V p(ψ; X); can
we find δ > 0 so that, for every Y = {yj}j∈Zd ⊂ Rd with supj yj − xj2 < δ,
the set BY = {τyj ψ}j∈Zd is a p−Riesz basis for V p(ψ; X)?
This problem cannot be solved if ψ has compact support. For example,
let ψ(x) = rect(x) be the characteristic function of the interval [− 1
2 ) and
let X = Z; let Y = {yn}n∈Z be such that y0 = δ > 0 and yn = n when
n 6= 0. All functions in V p(rect; Y ) vanish in the interval [− 1
2 + δ] and
so V p(rect; Y ) 6= V p(rect).
2 , − 1
2 , 1
We prove in Section 3 that Problem 1 can be solved when p = 2 and ψ
is band-limited, i.e., when the Fourier transform of ψ has compact support.
See also Section 5 for more remarks and comments on problem 1
We are concerned with the following problem:
Problem 2. With the notation of Problem 1: let BX be a p−Riesz basis
for V p(ψ; X) and let Y = {yn}n∈Zd that satisfies supn yn − xn2 < δ; is BY
a p−Riesz basis for V p(ψ; Y ) whenever δ is sufficiently small?
It is proved in [10] that Problem 2 has always solution when X is rel-
atively separated, i.e., when X = X1 ∪ ... ∪ Xk, with Xj = {xj,n}n∈Zd and
inf n6=m xj,n − xj,m2 > 0.
In Section 2 we prove the following theorem.
Theorem 1. Suppose that that ψ is in the Sobolev space W 1,p(Rd), with
1 < p < ∞, and that {τxj ψ}j∈Zd is a p−Riesz basis of V p(ψ; X). For
2 A proof of this ientity was kindly provided to us by K. Hamm
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
3
every j ∈ Zd there exists δj > 0 such that {τyj ψ}j∈Zd is a p−Riesz basis of
V p(ψ; Y ) whenever xj − yj2 < δj.
We recall that W 1,p(D) is the space of Lp(D) functions whose partial
(D) is the closure
distributional derivatives are also in Lp(D) and that W 1,p
of C ∞
0 (D) in W 1,p(D).
When X is not relatively separated the δj's in Theorem 1 may not have a
positive lower bound, but we can still solve Problem 2 in the cases considered
in Theorems 2 and 3 below.
0
Theorem 2. Assume that ψ ∈ L1(Rd) ∩ L2(Rd) satisfies
(1.4) 0 < c = Xk∈Zd
inf
x∈[0,1)d
ψ(x + k)2 < Xk∈Zd
sup
ψ(x + k)2 = C < ∞.
x∈[0,1)d
Then, Problem 2 can be solved when p = 2 and {e2πixn·x}n∈Zd is a Riesz
basis in L2([0, 1)d).
We
recall
that
measurable functions f
(cid:16)Pn∈Zd supx∈[0,1)d f (x + n)q
2(cid:17) 1
be defined in a similar manner.
the amalgam space W (L∞, ℓq)
set of
: Rd → C for which f W (L∞, ℓq) =
the
is
q < ∞. The amalgam space W (Lr, ℓq) can
The assumption (1.4) implies that ψ is in the amalgam space W (L∞, ℓ2).
From Theorem 2 follows that if ψ satisfies (1.4), then {τyk ψ}k∈Zd is a
4 . See the remark after
2−Riesz basis of V 2(ψ, Y ) whenever k − yk∞ < 1
the proof of Theorem 2 in Section 4.
Exponential Riesz bases in L2(0, 1) are completely understood and clas-
sified [20]. To the best of our knowledge, no such characterization exists for
exponential bases on L2((0, 1)d) when d > 1.
For our next theorem we consider ψ in the Sobolev space W 1,p
denote with ∂j ψ = ∂ψ
∂xj
we let ∇ψ = (∂1ψ, ..., ∂dψ) be the gradient of ψ. Let Y = {yk}k∈Zd and
(Rd); we
the partial derivative (in distribution sense) of ψ and
0
L = sup
k∈Zd
yk − k2 < ∞.
We prove the following
Theorem 3. Let D = (a1, b1) × ... × (ad, bd) be a bounded rectangle in
(D), with 1 ≤ p < ∞, and let {τkψ}k∈Zd be a p−Riesz
Rd. Let ψ ∈ W 1,p
basis of V p(ψ) with frame constants 0 < A ≤ B < ∞. If
0
(1.5)
C = L
dXj=1
(1 + [bj − aj + L])p−1k∂j ψkp
p < A,
the set {τyk ψ}k∈Zd is a p−Riesz basis of V p(ψ; Y ) with constants B + C and
A − C.
4
LAURA DE CARLI, PIERLUIGI VELLUCCI
The proofs of Theorems 2 and 3 are in Section 3.
Our Theorem 3 can be compared to [10, Theorem 3.5]. In this theorem
it is assumed that ∇(ψ) is in the amalgam space W (L∞, ℓ1), and that
inf x∈[0,1)dPk∈Zd ∇ψ(x + k)2 > 0.
In the aforementioned theorem is proved that {τyk ψ}k∈Zd is a Riesz basis
of V 2(ψ; Y ) if C ′ = L2(1 + 2L)2d∇ψ2
W (L∞, ℓ1) < A. Generalizations to
functions for which ∇(ψ) is in the amalgam space W (Lq, ℓ1), with q > d
are also possible (see Remark 3.2 in [10]).
Our Theorem 3 reduces to [10, Theorem 3.5] when p > d and ψ has
compact support. For example, when ψ has support in [0, 1)d, the norm in
W (Lp, ℓ1) reduces to ∇ψp. The constant C in Theorem 3 may be smaller
than C ′ defined above when the support of ψ is small.
Theorem 3 does not apply when ψ = rect or when ψ is a step function;
For J ≥ 1, we let SJ = ns(t) =Pj≤J sj rect(t − j), sj ∈ Ro . We let
p−1 be the dual exponent of p. The following theorem is proved in
p′ = p
Section 4.2.
Theorem 4. Assume that g ∈ SJ and that {τkg}k∈Z is a p−Riesz basis
for V p(g), with frame constants A and B. If
the sequence {τyk g}k∈Z is a Riesz basis for V p(g; Y ).
2pLJ kgkp
p′ < A ,
Acknowledgement. We are grateful to the anonymous referee of this paper
for her/his thorough reading of our manuscript and for providing suggestions
that have improved the quality of our work.
We also wish to thank K. Hamm for providing a proof of the identity
(1.3) for p 6= 2.
2. Preliminaries
2.1. Notation. We denote with hf, gi = RRd f (x)¯g(x)dx and kf k2 =
phf, f i the standard inner product and norm in L2(Rd). For a given p ∈ Rd
and δ > 0, we let B(p, δ) = {x ∈ Rd : x − p2 < δ}.
2 , 1
We let rect(x) = χ[− 1
2 )(x) be the characteristic function of the interval
2 , 1
[− 1
2 ) and βs = rect(s+1)(x) = rect ∗... ∗ rect(x) be the s + 1−times iterated
convolution of rect. The function βs(x), a piecewise polynomial function of
degree s, is a B-spline of order s. See [27], where the B-splines were first
introduced, and [21], [32] and the references cited there.
2.2. p− Riesz bases. Recall that a Schauder basis in a separable
Banach space V is a linearly independent set B = {vj}j∈Z such that:
span(B) = V , and there exists a sequence of bounded linear functions
fj : X → C (the functional coefficients of the basis) such that x =Pj fj(x)vj
for every x ∈ V .
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
5
Following [6], [7] and [2], we say that B is a p−Riesz basis of V , with
1 ≤ p < ∞, if Span(B) = V , if every series Pn anvn converges in V when
~a = (an)n∈Z ∈ ℓp and if there exist constants A, B > 0 such that, for every
finite sequence of coefficients {dj }j∈Z ⊂ C, we have
Ak ~dkℓp ≤ kXj
dj vjkp ≤ Bk ~dkℓp.
Every p−Riesz basis is a Schauder basis. As mentioned in the introduction,
when V = V p(ψ) and ψ has compact support, the functional coefficients of
the basis {τkψ}k∈Z of V p(ψ) can be written in terms of the dual functions
of the basis.
The following results are well known (see e.g. [1, Prop. 1.1], [16], [24]).
Lemma 5. a) Let ψ ∈ Lp
0(Rd). The set B = {τkψ}k∈Zd is a p−Riesz
basis in V p(ψ) if and only if
(2.1)
ψ(y + m)2 6= 0
for every y ∈ [− 1
2 , 1
2 )d.
Xm∈Zd
b) If ψ ∈ W (L∞, ℓ1) is continuous and if B is Riesz basis in V 2(ψ), then
B is a p-Riesz basis in V p(ψ) for every 1 ≤ p < ∞.
Proof. For the convenience of the reader we prove that if ψ ∈ L2
0(Rd),
B is a Riesz basis of V 2(ψ) with constants 0 < A ≤ B < ∞ if and only if
the following inequality holds for every y ∈ Q = [− 1
2 , 1
2 )d.
(2.2)
A ≤ Xm∈Zd
ψ(y + m)2 ≤ B.
We can verify (using e.g. the Poisson summation formula) that the function
in (2.1) is continuous in Q, and so (2.2) is equivalent to (2.1).
Let {ck}k∈Zd ⊂ C be a finite set of coefficients such that Pk ck2 = 1.
The Fourier transform of f =Pk∈Zd ckτkψ is
cke2πiy·k = ψ(y)M (y).
f (y) = ψ(y) Xk∈Zd
and by Plancherel's theorem kf k2
2 = k f k2
2 = Xm∈Zd
M (y)2 Xm∈Zd
Zm+[− 1
, 1
2 )d
ψ(y + m)2dy.
2
Xm∈Zd
RQ M (y)2g(y)dy andRQ M (y)2dy =Pk ck2 = 1 follows that A ≤ f 2
2 ≤ BPk ck2 and the above consid-
ψ(y + m)2M (y)2dy =ZQ
ZQ
Let g = Pm∈Zd ψ(y + m)2;
Conversely, from APk ck2 ≤ kf k2
erations, follows that
2 =
2 ≤
from f 2
(2.2) holds,
B.
if
f (y)2dy =
(2.3)
AkM 2kL1(Q) ≤ZQ
M (y)2g(y)dy ≤ BkM 2kL1(Q).
6
LAURA DE CARLI, PIERLUIGI VELLUCCI
Every non-negative h ∈ L1(Q) can be written as h = M 2, with M ∈ L2(Q).
The dual of L1(Q) is L∞(Q) and so kgkL∞(Q) = supkhkL1(Q)=1RQ f (y)g(y)dy.
From (2.3) follows that A ≤ kgkL∞(Q) ≤ B as required.
We will use the following Paley-Wiener type result.
(cid:3)
Lemma 6. Let X, Y ⊂ Rd be countable and discrete. Suppose that
{τxj ψ}j∈Zd is a p−Riesz basis of V p(ψ; X) with constants A ≤ B. If the
inequality
Xj
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
aj(τxj ψ − τyj ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
p
≤ CXn
anp
holds for all finite sequences {an}n∈Zd ∈ C with a constant C < A, the
sequence {τyj ψ}j∈Zd is a p−Riesz basis of V p(ψ; Y ) with constants B + C
and A − C.
Xj
Proof. Assume that Pn anp = 1; we have:
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
aj τyj ψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
aj(τxj ψ − τyj ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
aj τxj ψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
Xj
Xj
aj(τxj ψ − τyj ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
ajτxj ψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
ajτyj ψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
−(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj
Xj
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≥(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj
and
≤ C + B
≥ A − C.
(cid:3)
p′ = p
Proof of Theorem 1 . Assume p ∈ (1, ∞) and Pj ajp = 1. Let
constants such that Pj δj p′
p−1 is the dual exponent of p and let {δj}j∈Zd be a sequence of positive
< ∞, We recall that, when 1 < p < ∞, a
function f ∈ Lp(Rd) is in the Sobolev space W 1,p(Rd) if and only if there is
a constant c > 0 that depends on f but not on δ, such that
(2.4)
ωp(δ, f ) = sup
t<δ
τtf − f p ≤ cδ
for every δ > 0. Furthermore, one can choose c = k∇f kp. See e.g. Prop.
9.3 in [4]. By (2.4) and Holder' s inequality,
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
aj(τxj ψ − τyj ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xj
≤Xj
p
ajp
≤ c
Xj
Xj
1
aj kτxj ψ − τyj ψkp ≤Xj
δj p′
δjp′
= c
Xj
1
p′
1
p′
.
aj δj
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
7
We can chose the δj so small that c(cid:16)Pj δjp′(cid:17) 1
complete the proof.
p′
< A and use Lemma 6 to
(cid:3)
3. Problem 1 (p = 2)
In this section we prove that Problem 1 can be solved when p = 2 and
ψ has compact support.
Theorem 7. Let ψ ∈ L2(Rd). Assume that ψ has compact support and
that there exist constants c, C > 0 such that
c ≤ inf
x∈Rd
ψ(x) ≤ sup
x∈Rd
ψ(x) ≤ C.
Let {τxj ψ}j∈Zd be a Riesz basis in V 2(ψ, X). There exists δ > 0 such that
if Y = {yj}j∈Zd ⊂ Rd satisfies supj xj − yj < δ, then also {τxj ψ}j∈Zd is a
Riesz basis of V 2(ψ).
Proof. Let D = supp( ψ). When p = 2, Plancherel theorem implies
that the set {τxj ψ}j∈Zd is a Riesz basis in V 2(ψ) if and only if the set
{e2πixj ·x}j∈Zd is a Riesz basis on L2(Rd, ψ dx). Our assumptions on ψ
imply that the norm on L2(Rd, ψ dx) is equivalent to the norm on L2(D)
and that {e2πixj ·x}j∈Zd is an exponential Riesz basis on L2(D). Exponential
Riesz bases on bounded domains of Rd are stable under small perturbations
(see [22] and also Section 2.3 in [18]); we can find δ > 0 such that, if
Y = {yj}j∈Zd ⊂ Rd satisfies supj xj − yj < δ, then also {e2πiyj ·x}j∈Zd is a
Riesz basis on L2(D) and hence also in L2(Rd, ψ dx).
(cid:3)
Example. Let d = 1 and let ψ = sinc(x) = sin(πx)
; the Fourier transform
of τkψ(x) = sinc(x − k) is e2πikx rect(x) = e2πikxχ[− 1
2 )(x), and so V 2(ψ)
2 , 1
is isometrically isomorphic to Span{e2πijx rect(x)}j∈Zd = L2(− 1
2 ). By
Kadec's theorem ([17], [33]) if Y = {yn}n∈Zd ⊂ R is such that supn yn −
n ≤ δ < 1
2 ) and
therefore, the set {sinc(x − yn)}n∈Zd is a Riesz basis for V 2(sinc). Thus,
V 2(sinc; Y ) = V 2(sinc).
4 , the set {e2πiynx}n∈Zd is still a Riesz basis of L2(− 1
2 , 1
πx
2 , 1
Things are not so clear when p 6= 2. For example, the trigonometric
system B = {e2πinx}n∈Zd is a Schauder basis in Lp(− 1
2 ) when 1 < p < ∞,
but it is not a p−Riesz basis and the previous example cannot be generalized
2 , 1
in an obvious way. Stability results for the Schauder basis B in Lp(− 1
2 )
are proved in [25] and in [26].
2 , 1
4. Problem 2
In this section we prove Theorems 2 and 3. Let X = {xn}n∈Zd and
B = {e2πix·xn}n∈Zd. We first prove the following
8
LAURA DE CARLI, PIERLUIGI VELLUCCI
Lemma 8. Let ψ ∈ L2(Rd) ∩ L1(Rd) be as in (1.4); if B is a Riesz
basis in L2([0, 1)d) with constants A1 and B1 then {τxn ψ} is a Riesz basis
of V 2(ψ, X)} with constants A = A1c and B = B1C.
Proof. For k ∈ Zd, we let ck = inf x∈(0,1]d ψ(x + k)2 and Ck =
supx∈(0,1]d ψ(x + k)2. Let {dj} be a finite set of complex coefficient such
that Pj dj2 = 1. Since B is a Riesz basis in L2((0, 1]d), for every given
k ∈ Zd we have that
≤ B1.
2
L2((0,1]d)
From this inequality follows at once that
A1 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ckA1 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn
A1c ≤ Xk∈Zd
dne−2πixn·k e2πixn·y(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn
dne−2πixn·ke2πixn·y ψ(. − k)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
With c =Pk∈Zd ck and C =Pk∈Zd Ck = ψ2
dne2πixn·(.−k) ψ(. − k)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn
In view of Pk∈Zd kg(. − k)kL2((0,1]d) = g2, we obtain
dne2πixn·y ψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
A1c ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xn
≤ B1C.
2
2
2
2
≤ CkB1.
L2((0,1]d)
W (L∞, ℓ2), we have
≤ B1C.
L2((0,1]d)
By Plancherel's theorem, the latter is equivalent to A1c ≤ kPn dnτxnψk2 ≤
B1C and so {τxk ψ}k∈Zd is a Riesz basis of V 2(ψ, X), as required.
(cid:3)
Proof of Theorem 2. Let B = {e2πix·xn}n∈Zd be a Riesz basis in
L2([0, 1)d); it is proved in [22] (see also Section 2.3 in [18]) that we can
find δ > 0 such that, if Y = {yj}j∈Zd ⊂ Rd satisfies supj xj − yj2 < δ, then
also {e2πiyj ·x}j∈Zd is a Riesz basis in L2([0, 1)d). By Lemma 8, {τynψ} is a
Riesz basis of V 2(ψ, Y ).
(cid:3)
Remark. When Y = {yk}k∈Zd is such that supk∈Zd k − yk∞ < 1
4 , by the
multi-dimensional generalization of Kadec's theorem proved in [31] we have
that {e2πiyj ·x}j∈Zd is a Riesz basis in L2([0, 1)d) and by Lemma 8, {τyn ψ}n∈Zd
is a Riesz basis of V 2(ψ, Y )}.
4.1. Proof of Theorem 3. In order to prove Theorem 3 we need some
preliminary result: first, we prove the following
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
9
Lemma 9. Let (a, b) ⊂ R, with a < b < ∞, and let 1 ≤ p < ∞. Let
ψ ∈ Lp
0(a, b). For every finite set of coefficients {αj } ⊂ C, we have that
Xk
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αkτkψ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
p
≤ kψkp
p([b − a] + 1)p−1Xk
αkp
where [ ] denotes the integer part.
Proof. For simplicity we let a = 0. When b ≤ 1 the supports of the
τkψ's are disjoint and so kf kp
b > 1 the supports of the τkψ overlap, and there are at most [b] + 1 of
such supports that intersect at each point. By the elementary inequality
(x1 + · · · + xm)p ≤ mp−1 (xp
m) (which is valid when the xj are
non-negative) we have that
1 + · · · + xp
pPk αkp. When
p = kPk αkτkψkp
p = kψkp
f (t)p = Xk
akτkψ(t)p ≤ ([b] + 1)p−1Xk
pPk αkp as required.
akpτkψ(t)p
and so kf kp
p ≤ ([b] + 1)p−1kψkp
(cid:3)
Let Y = {yk}k∈Zd be a discrete subset of Rd. Assume that L =
≤ L(1 + [b − a + L])p−1kψ′kp
p
where ψ′(t) denotes the distributional derivative of ψ. Assume first that
k ψ′(t + x)dx and that
finite set of coefficients {αj} ⊂ C such that Pk αkp = 1, we have that
p
0
(D). Then, for every
(1 + [bj − aj + L])p−1k∂j ψkp
p.
supk∈Zd yk − k2 < ∞. We prove the following
p
p
p
≤ L
(4.2)
(4.1)
dXj=1
Proof. When d = 1 and D = (a, b), we prove that
j=1[aj , bj] and let ψ ∈ W 1,p
Lemma 10. Let D =Qd
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αk(τkψ − τyk ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
αk(τkψ − τykψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αk(τkψ − τyk ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αkZ t+k+L
Xk
where we have let g(t) = R t+L
yk > k. Observing that ψ(t + yk) − ψ(t + k) = R yk
αkZ t+yk
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
ψ′(x)dx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
k − yk ≤ L, we have that
ψ′(x)dx.
t+k
t+k
p
p
p
p
p
p
ψ′(x)dx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αkτkg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
p
It is easy to verify that g(t) is
supported in the interval [a − L, b]. Indeed, ψ′ is supported in [a, b] and so
g(t) ≡ 0 whenever t + [0, L] ∩ [a, b] = ∅. Thus, g(t) ≡ 0 when t + L < a or
t > b, or: g(t) ≡ 0 when t ∈ R − [a − L, b], as required.
t
10
LAURA DE CARLI, PIERLUIGI VELLUCCI
We apply a change of variables and Minkowsky's integral inequality; we
gather
≤ (1 + [b − a + L])p−1kgkp
p .
By Lemma 9
(4.3) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
αk(τkψ − τyk ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kgkp =(cid:13)(cid:13)(cid:13)(cid:13)
Z t+L
t
p
p
p
p
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αkτkg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
ψ′(x)dx(cid:13)(cid:13)(cid:13)(cid:13)p
=(cid:13)(cid:13)(cid:13)(cid:13)
Z L
≤ Lkψ′kp
0
ψ′(x + t)dx(cid:13)(cid:13)(cid:13)(cid:13)p
t
[a, b + L].
When d = 2 we can let yk = (yk,1, yk,2) and k = (k1, k2) and write
which together with the inequality (4.3) concludes the proof of (4.2). When
yk < k the proof if similar, but the function g(t) defined above should
ψ′(x)dx, a function supported in the interval
be replaced by g(t) = R t−L
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αk(τkψ − τyk ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
Xk
+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αk(τ(k1,k2)ψ − τ(yk,1,k2)ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
αk(τ(yk,1,k2)ψ − τ(yk,1,yk,2)ψ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
Xk
αk(τk2 ψ2 − τyk,2ψ2)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
+(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
αk(τk1ψ1 − τyk,1ψ1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)p
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
Xk
where we have let ψ1 = τ(0,k2)ψ and ψ2 = τ(yk1
applied to ψ1 and ψ2, yields (4.1). The case d > 2 is similar.
,0)ψ. The inequality (4.2),
Proof of Theorem 3. Follows from Lemmas 6 and 10.
(cid:3)
(cid:3)
4.2. rect and step functions. Since Sobolev spaces W 1,p(R) do not
contain discontinuous functions, we cannot apply Theorem 3 when ψ is a
step function.
Let ψ = rect; it is immediate to verify that, for every 1 ≤ p < ∞, the set
{τj rect}j∈Z is a p−Riesz basis of V p(rect) with frame constants A = B = 1.
Throughout this section we let Y = {yk}k∈Z ⊂ R, with L = supk∈Zd yk − k
and we assume 1 ≤ p < ∞.
Lemma 11 below is an easy generalization of Lemma 10 in [8].
Lemma 11. Assume 0 ≤ L < 1. For every finite set of coefficients
{αk}n∈Zd ⊂ C we have that
(4.4)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
αk(rect(t − k) − rect(t − yk))(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
p
< 2pLXk
αkp.
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
11
Proof. Assume Pk αkp = 1. Let
(4.5)
f (t) =Xk
αk (rect(t − k) − rect(t − yk)) =Xk
αkχIk,
where Ij denotes the support of rect(t − j) − rect(t − yj). When yj 6= j, Ij
is union of two intervals that we denote with I +
j . When yj > j, we
let
j and I −
I −
j = (j −
1
2
, yj −
1
2
),
I +
j = (j +
1
2
, yj +
1
2
).
We use (improperly) the same notation to denote I +
yj < j.
j and I −
j also when
h there is at most another interval I ±
Since we have assumed yh − h ≤ L < 1, for every given interval
J = I ±
k that overlap with J; thus,
for every t ∈ R, the sum in (4.5) has at most 2 terms. By the ele-
mentary inequality (x1 + · · · + xm)p ≤ mp−1 (xp
m) we have that
p ≤ 2p−1 supk Ik = 2p−1(2L) = 2p L
(cid:3)
f (t)p ≤ 2p−1Pk αkpχIk (t), and kf kp
and the proof of the Lemma is concluded.
1 + · · · + xp
Lemma 11 and Lemma 6 yield the following
Theorem 12. With the notation of Lemma 11, the set {τyk rect}k∈Z is
a p−Riesz basis in V p(rect; Y ) if 2pL < 1.
Corollary 13. Let ψ0 ∈ L1(R) and let ψ = rect ∗ψ0. Suppose that
{τkψ}k∈Z is a p−Riesz basis for V p(ψ). For every finite set of coefficients
{αk}n∈Z ⊂ C with Pk αkp = 1, we have that
αk(ψ(t − k) − ψ(t − yk))(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
p
p
< 2pLkψ0kp
1
and the set {ψ(t − yk)}k∈Z is a p−Riesz basis for for V p(ψ; Y ) whenever
(4.6)
2pLkψ0kp
1 < A.
Remark. If ψ0(x) 6= 0 for every x ∈ R, then the set {τkψ}k∈Z is a p−Riesz
basis for V p(ψ).
0 whenever x ∈ [− 1
Indeed, Pm∈Z drect(y + m)2 = Pm∈Z sinc(x + k)2 6=
2 ) and so also Pm∈Z ψ(x + k)2 = Pm∈Z ψ0(x +
k)drect(x + k)2 6= 0; by Lemma 5 the set {τkψ}k∈Z is a p−Riesz basis for
V p(ψ).
2 , 1
proof of Corollary 13. Let
αk (ψ(t − k) − ψ(t − yk)) ,
F (t) =Xk
f (y) =Xk∈Z
αk (rect(y − k)−rect(y − xk)
12
LAURA DE CARLI, PIERLUIGI VELLUCCI
and we show that kF kp
p ≤ 2pLkψ0kp
1. We gather
F (t) =Z ∞
−∞
ψ0(t − y)Xk∈Z
αk (rect(y − k)−rect(y − xk)) dy
= ψ0 ∗ f (t).
Thus, by Young's inequality and Lemma 11,
kF kp
p ≤ kψ0kp
1kf kp
p ≤ 2pLkψ0kp
1
and the proof of the corollary is concluded.
(cid:3)
Let βm(x) = rect(m+1) be the B-spline of order m > 1. We recall
that βm is supported in the interval [− m+1
2 ] and βm(x) ∈ W 1,p(R)
whenever m ≥ 1. It is easy to verify by induction on m that kβmkp ≤ 1 and
mkp ≤ 2. It is known that {τkβm}k∈Z is a Riesz basis of V 2(βm) whose
kβ′
Riesz constants A(m) and B(m) are explicitly evaluated in [19]. See also
[28]. By the observations after Lemma 5, {τkβm}k∈Z is a p−Riesz basis of
V p(βm) with constants Ap(m) > 0 and Bp(m) < ∞.
2 , m+1
We prove the following
Corollary 14. Assume that L < 2−pAp(m). Then, the set {τyk βm}k∈Z
is a p−Riesz basis of V p(ψ, Y ).
Proof. We apply Corollary 13 with ψ0 = βm−1.
(cid:3)
Remark. We could have used Theorem 3 to prove Corollary 14, but we
would have obtained a lower upper bound for L (namely, L < Ap(m)
2(2+m)p−1 ).
proof of Theorem 4. Let g(t) = Pj≤J sj rect(t − j).
{αk}n∈Zd ⊂ C be a finite set of coefficients such that Pk αkp = 1. Let
Let
f (t) =Xk
αk (g (t − k) − g (t − xk)) .
As in previous theorems, we find conditions on L for which kf kp
p < A.
We have
f (t) = Xj≤J
= Xj≤J
sjXk
αk (rect(t − j − k) − rect(t − j − xk))
sjfj(t).
p
dt
ZR
αk (rect(t − j − k) − rect(t − j − xk))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
fj(t)pdt =ZR
Xk
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
αk(cid:0)rect(t′ − k) − rect(t′ − xk)(cid:1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=ZR
Xk
αk(rect(t − k) − rect(t − yk)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk
dt′
p
p
p
.
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
13
By Minkowski and Holder inequalities, and noting that Pj≤J sjq = kgkq
it follows that
q,
kf kp ≤ Xj≤J
(4.7)
sjkfjkp ≤
sjp′
Xj≤J
= kgkp′
Xj≤J
1
p′
Xj≤J
p
1
p
kfjkp
,
1
p
kfjkp
p
With the change of variables t − j = t′ in the integral below, we obtain
From Lemma 11, follows that the integral above is ≤ 2pL. We gather:
kf kp
p ≤ 2pLJ kgkp
p′ < A, and by Lemma 6
Theorem 4 follows.
(cid:3)
p′ . By assumption 2pLJ kgkp
5. Remarks and open problems
We have discussed Problem 1 when p = 2 and the Fourier transform of
the window function ψ has compact support. When ψ is not band-limited,
Plancherel's theorem implies that the set {τxj ψ}j∈Zd is a Riesz basis in
V 2(ψ, X) if and only if the set V = {e2πixj ·x ψ}j∈Zd is a Riesz sequence
in L2(Rd), and hence a Riesz basis in V = Span(V)). By a theorem of
[33, Theorem 11]) for every j ∈ Zd there
Krein-Milman-Rutman (see e.g.
exists ǫj > 0 such that every set of functions {gj }j∈Zd ⊂ V is a Riesz
basis of V whenever gj − eπixj ·x ψ2 < ǫj. We can find δj > 0 such that
(eπixj ·x − eπiyj·x) ψ2 < ǫj whenever xj − yj2 < δj, but we do not know
whether the δj's have a lower bound or not.
For functions ψ in Lp(Rd) for every p ∈ [1, ∞) it would be interesting
to prove conditions that would ensure that a q-Riesz basis in V q(ψ, X) for
some q is automatically a p-Riesz basis in V p(ψ, X) for all p. Lemma 5 (b)
shows that, for certain ψ, if the set {τkψ}k∈Zd is a 2-Riesz basis of V 2(ψ),
it is also a p-Riesz in V p(ψ) but the method of proof of this result does not
generalize well to other sets of translations. Results in [3] and [29] may help
generalize Lemma 5.
14
LAURA DE CARLI, PIERLUIGI VELLUCCI
It would also be interesting to define and investigate p-Riesz bases in
quasi-shift invariant spaces V p(ψ, X) when 0 < p < 1. Wavelet in Lp with
0 < p < 1 have been considered in [12]. We feel that the results contained in
Section 3 of [12] may help the understanding of V p(ψ, X) when 0 < p < 1.
References
[1] A. Aldroubi, Q. Sun, Connection between p−frames and p−Riesz bases in locally finite
SIS of Lp(R) Proceedings of SPIE - The International Society for Optical Engineering,
February 1970
[2] A. Aldroubi, Q. Sun, W. Tang, p-Frames and shift invariant subspaces ofLp , J. Fourier
Anal. Appl. 7 (2001) 1-21.
[3] A. Aldroubi, A. Baskakov, I. Krishtal, Slanted matrices, Banach frames, and sampling,
J. Funct. Anal. 255 (2008) 1667 -- 1691.
[4] H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations,
Springer Verlag 2011.
[5] A. Ben-Artzi, A. Ron, On the Integer Translates of a Compactly Supported Function:
Dual Bases and Linear Projectors, SIAM J. Math. Anal., 21(6), 15501562.
[6] P. Casazza, O. Christensen, D.T. Stoeva, Frame expansions in separable Banach spaces,
J. Math. Anal. Appl. 307 (2005) 710 -- 723.
[7] O. Christensen, D.T. Stoeva, p-Frames in separable Banach spaces, Adv. Comput.
Math. 18 (2003) 117-126.
[8] L. De Carli, P. Vellucci, Stability theorems for the n-order hold model, arXiv:1605.01706
(2016) (submitted)
[9] H. Feichtinger, Spline-type spaces in Gabor analysis. Wavelet analysis (Hong Kong,
2001), 100122, Ser. Anal., 1, World Sci. Publ., River Edge, NJ, 2002.
[10] H. Feichtinger, U. Molter, J.L.Romero, Perturbation techniques in irregular spline-
type spaces, World Scientific Publishing Co. Inc., Int. J. Wavelets Multiresolut. Inf.
Process, 6 (2) (2008) 249277.
[11] H. Feichtinger, H.; D.M. Onchi, Constructive reconstruction from irregular sampling
in multi-window spline-type spaces. Progress in analysis and its applications, 257265,
World Sci. Publ., Hackensack, NJ, 2010.
[12] G. Garrigos; R. Hochmuth; A. Tabacco, Wavelet characterizations for anisotropic
Besov spaces with 0 < p < 1 Proc. Edinburgh Math. Soc. 47 (2004) 573-59.
[13] K. Grochenig, J. Stockler, Gabor Frames and Totally Positive Functions, Duke Math.
J. 162(6) (2013), 1003 -- 1031.
[14] K. Hamm, J. Ledford, On the structure and interpolation properties of quasi-invariant
shift spaces, arxiv:1703.01533 (2017)
[15] C. Heil, A basis theory primer, Appl. Num. Harm. Analysis, Birkhauser 2011.
[16] R.Q. Jia, C. A. Micchelli, Using the refinement equation for the construction of pre-
wavelets II: power of two, In "Curve and Surface" (P. J. Laurent, A. Le Mehaute and
L. L. Schumaker eds.), Academic Press, New York 1991, pp. 209 -- 246.
[17] M.I. Kadec, The exact value of the Paley-Wiener constant, Soviet Math. Dokl., 5
(1964), 559 -- 561.
[18] G. Kozma, S. Nitzan, Combining Riesz bases, Inv. Math. 199 (1) (2015), , pp 267-285
[19] E.V. Mischenko, Determination of Riesz bounds for the spline basis with the help of
trigonometric polynomials. Sib. Math. J. 51(4), 660-666 (2010)
[20] B. Pavlov, Basicity of an exponential system and Muckenhoupt's condition, Soviet
Math. Dokl. 20 (1979) 655 -- 659.
[21] Prautzsch, H., Boehm, W., Paluszny, M., Bezier and B-Spline Techniques, Springer
Science and Business Media (2002).
[22] R. Paley and N. Wiener, Fourier transforms in the complex domain. Amer. Math.
Soc. Colloquium Publications, vol. 19; Amer. Math. Soc., New York, 1934.
P-RIESZ BASES IN QUASI SHIFT INVARIANT SPACES
15
[23] J.L. Romero, Explicit localization estimates for spline-type spaces. Sampl. Theory
Signal Image Process. 8 (2009), no. 3, 249-259.
[24] A. Ron, A necessary and sufficient condition for the linear independence of the integer
translates of a compactly supported distribution, Constructive Approximation 5 (1),
297-308.
[25] A. M. Sedletskii, Izv. Vyssh. Uchebn. Zaved., Mat., No. 7, 85-91 (1973).
[26] A. M. Sedletskii, Equivalence of the trigonometric system and its perturbations in
Lp(π, π), Doklady Mathematics. Vol. 94. No. 1. Pleiades Publishing, 2016.
[27] I. J. Schonberg, Cardinal interpolation and spline functions. J. Approx. teory 2 (2)
(1969), pp. 167 -- 206.
[28] A. Antony Selvan, R. Radha, Sampling and Reconstruction in Shift Invariant Spaces
of B-Spline Functions Acta Appl. Math. DOI 10.1007/s10440-016-0053-6 (2016)
[29] C. E. Shin, Q. Sun, Stability of localized operators, J. Funct. Anal. 256 (2009) 2417 --
2439.
[30] Q. Sun, Stability of the Shifts of Global Supported Distributions J.Math. Analysis and
Appl. 261 (1)(2001) 113 -- 125.
[31] W. Sun, X. Zhou, On the stability of multivariate trigonometric systems. J. Math
Anal. Appl. 235 (1999), 159 -- 167.
[32] M. Unser, A. Aldroubi, M. Eden, B-spline signal processing. I. Theory, Signal Pro-
cessing, IEEE Transactions on 41 (2), 821 -- 833.
[33] R. M. Young, An introduction to nonharmonic Fourier series, Academic Press, 2001
Laura De Carli: Florida International Univ., Univ. Park, Miami (FL)
E-mail address: [email protected]
Pierluigi Vellucci: Dept. of Economics, Roma Tre University, via Silvio
D'Amico 77, 00145 Rome, Italy.
E-mail address: [email protected]
|
1002.3583 | 2 | 1002 | 2011-07-25T14:11:34 | Zone diagrams in compact subsets of uniformly convex normed spaces | [
"math.FA",
"cs.CG",
"math.GN"
] | A zone diagram is a relatively new concept which has emerged in computational geometry and is related to Voronoi diagrams. Formally, it is a fixed point of a certain mapping, and neither its uniqueness nor its existence are obvious in advance. It has been studied by several authors, starting with T. Asano, J. Matousek and T. Tokuyama, who considered the Euclidean plane with singleton sites, and proved the existence and uniqueness of zone diagrams there. In the present paper we prove the existence of zone diagrams with respect to finitely many pairwise disjoint compact sites contained in a compact and convex subset of a uniformly convex normed space, provided that either the sites or the convex subset satisfy a certain mild condition. The proof is based on the Schauder fixed point theorem, the Curtis-Schori theorem regarding the Hilbert cube, and on recent results concerning the characterization of Voronoi cells as a collection of line segments and their geometric stability with respect to small changes of the corresponding sites. Along the way we obtain the continuity of the Dom mapping as well as interesting and apparently new properties of Voronoi cells. | math.FA | math |
ZONE DIAGRAMS IN COMPACT SUBSETS OF UNIFORMLY
CONVEX NORMED SPACES
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
Abstract. A zone diagram is a relatively new concept which has emerged in
computational geometry and is related to Voronoi diagrams. Formally, it is a
fixed point of a certain mapping, and neither its uniqueness nor its existence
are obvious in advance. It has been studied by several authors, starting with T.
Asano, J. Matousek and T. Tokuyama, who considered the Euclidean plane with
singleton sites, and proved the existence and uniqueness of zone diagrams there.
In the present paper we prove the existence of zone diagrams with respect to
finitely many pairwise disjoint compact sites contained in a compact and convex
subset of a uniformly convex normed space, provided that either the sites or
the convex subset satisfy a certain mild condition. The proof is based on the
Schauder fixed point theorem, the Curtis-Schori theorem regarding the Hilbert
cube, and on recent results concerning the characterization of Voronoi cells as
a collection of line segments and their geometric stability with respect to small
changes of the corresponding sites. Along the way we obtain the continuity
of the Dom mapping as well as interesting and apparently new properties of
Voronoi cells.
1. Introduction
A zone diagram is a relatively new concept related to geometry and fixed point
theory. In order to understand it better, consider first the more familiar concept
of a Voronoi diagram. In a Voronoi diagram we start with a set X, a distance
function d, and a collection of subsets (Pk)k∈K in X (called the sites or the gen-
erators), and with each site Pk we associate the Voronoi cell Rk, that is, the set
of all x ∈ X the distance of which to Pk is not greater than its distance to the
union of the other sites Pj, j 6= k. On the other hand, in a zone diagram we
associate with each site Pk the set Rk of all x ∈ X the distance of which to Pk
is not greater than its distance to the union of the other sets Rj, j 6= k. Figures
1 and 2 show the Voronoi and zone diagrams, respectively, corresponding to the
same ten singleton sites in the Euclidean plane.
At first sight, it seems that the definition of a zone diagram is circular, because
the definition of each Rk depends on Rk itself via the definition of the other cells
Rj, j 6= k. On second thought, we see that, in fact, a zone diagram is defined to be
a fixed point of a certain mapping (called the Dom mapping), that is, a solution
of a certain equation. While the Voronoi diagram is explicitly defined, so its
existence (and uniqueness) are obvious, neither the existence nor the uniqueness
of a zone diagram are obvious in advance. As a result, in addition to the problem
of finding algorithms for computing zone diagrams, we are faced with the more
1
2
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
Figure 1. A Voronoi diagram
of 10 point sites in a square in
(R2, ℓ2).
Figure 2. A zone diagram of
the same 10 points as in Figure
1.
fundamental problem of establishing their existence (and uniqueness) in various
settings, and with the problem of reaching a better understanding of this concept.
The concept of a zone diagram was first defined and studied by T. Asano, J.
Matousek and T. Tokuyama [2, 3] (see also [4]), in the case where (X, d) was
the Euclidean plane, each site Pk was a single point, and all these (finitely many)
points were different. They proved the existence and uniqueness of a zone diagram
in this case, and also suggested a natural iterative algorithm for approximating
it. Their proofs rely heavily on the above setting. Several other papers related to
zone diagrams in the plane have been published, e.g., those of T. Asano and D.
Kirkpatrick [1], of J. Chun, Y. Okada and T. Tokuyama [6], and recently of S. C.
de Biasi, B. Kalantari and I. Kalantari [9].
Shortly after [3], the authors of [22] considered general sites in abstract spaces,
called m-spaces, in which X is an arbitrary nonempty set and the "distance"
function should only satisfy the condition d(x, x) ≤ d(x, y) ∀x, y ∈ X and can
take any value in the interval [−∞, ∞]. They introduced the concept of a double
zone diagram, and using it and the Knaster-Tarski fixed point theorem, proved the
existence of a zone diagram with respect to any two sites in X. They also showed
that in general the zone diagram is not unique. In a recent work by K. Imai, A.
Kawamura, J. Matousek, Y. Muramatsu and T. Tokuyama [14], the existence and
uniqueness of the zone diagram with respect to any number of general positively
separated sites in the n-dimensional Euclidean space Rn was announced. The
proof is based on results from [22] and on an elegant geometric argument specific
to Euclidean spaces. Very recently some of these authors have generalized this
result to finite dimensional normed spaces which are both strictly convex and
smooth [15].
In the present paper we prove the existence of zone diagrams with respect to
finitely many pairwise disjoint compact sites contained in a compact and convex
subset of a (possibly infinite dimensional) uniformly convex space, provided that
ZONE DIAGRAMS
3
either the sites or the convex subset satisfy a certain mild condition. This mild
condition holds, for instance, if either the convex subset X has a strictly convex
boundary or the sites are contained in the interior of X relative to the affine hull
spanned by X (and X is arbitrary). The proof is based on the Schauder fixed
point theorem, the Curtis-Schori theorem regarding the Hilbert cube, and on re-
cent results concerning the characterization of Voronoi cells as a collection of line
segments and their geometric stability with respect to small changes of the corre-
sponding sites (see Sections 3 and 4). Along the way we obtain the continuity of
the Dom mapping (Proposition 6.1) in a general setting and interesting properties
of Voronoi cells, namely Lemma 5.1 and Lemma 5.2. Although Voronoi diagrams
have been the subject of extensive research during the last decades [5, 18, 12],
this research has been mainly focused on Euclidean finite dimensional spaces (in
many cases just R2 or R3), and it seems that these lemmata are new even for R2
with a non-Euclidean norm.
It may be of interest to compare our main existence result with the recent
existence result described in [15]. On the one hand, our result is weaker than that
result, since we only prove the existence of a zone diagram in a compact and convex
set, while in [15] uniqueness is also proved and the setting is the whole space Rn. In
addition, it seems that some of the arguments in [15], although only formulated
for finitely many compact sites, can be extended to infinitely many, positively
separated closed sites. On the other hand, our result is stronger in the sense
that we allow infinite dimensional spaces and we do not require the smoothness
of the norm. As a matter of fact, the counterexamples mentioned in [15] show
that uniqueness does not necessarily hold if the norm is not smooth.
In any
case, the strategies used for proving these two results are completely different: in
[15] the authors use the existence of double zone diagrams (based on the Knaster-
Tarski fixed point theorem) and several geometric arguments, and here we use the
Schauder fixed point theorem, the Curtis-Schori theorem regarding the Hilbert
cube, and several general results about Voronoi cells in uniformly convex normed
spaces and elsewhere.
The structure of the paper is as follows:
in Section 2 we present the basic
definitions and notation. In Section 3 we provide the outline of the proof of the
main result.
In Section 4 we formulate several claims which are needed in the
proof. The proof itself is given in Sections 5, 6 and 7. We conclude the paper
with Section 8, which contains open problems and two pictures of zone diagrams.
We note that although the setting in the main result is a compact and convex
subset of a uniformly convex normed space, many of the auxiliary results actually
hold in a more general setting with essentially the same proofs, and therefore we
formulate and prove them there.
2. Definitions and Notation
In this section we present our notation and basic definitions. We consider a
closed and convex set X 6= ∅ in some normed space (eX, · ), real or complex,
4
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
finite or infinite dimensional. The induced metric is d(x, y) = x − y. We assume
that X is not a singleton, for otherwise everything is trivial.
The notation θ = 1 will always mean a unit vector θ. Given p ∈ X we set
Θp = {θ : θ = 1, p + tθ ∈ X for some t > 0}. This is the set of all directions
such that rays emanating from p in these directions intersect X not only at p.
We denote by [p, x] the closed line segment connecting p and x, i.e., the set
{p + t(x − p) : t ∈ [0, 1]}. We denote by B(x, r) the open ball of center x and
radius r. The notation Y ≈ Z means that the topological spaces Y and Z are
homeomorphic.
Definition 2.1. Given two nonempty sets P, A ⊆ X, the dominance region
dom(P, A) of P with respect to A is the set of all x ∈ X the distance of which to
P is not greater than its distance to A, that is,
dom(P, A) = {x ∈ X : d(x, P ) ≤ d(x, A)}.
Here d(x, A) = inf{d(x, a) : a ∈ A}.
Definition 2.2. Let K be a set of at least 2 elements (indices), possibly infinite.
Given a tuple (Pk)k∈K of nonempty subsets Pk ⊆ X, called the generators or the
sites, the Voronoi diagram induced by this tuple is the tuple (Rk)k∈K of nonempty
subsets Rk ⊆ X, such that for all k ∈ K,
Rk = dom(Pk,[j6=k
Pj) = {x ∈ X : d(x, Pk) ≤ d(x, Pj) ∀j 6= k, j ∈ K}.
In other words, each Rk, called a Voronoi cell, is the set of all x ∈ X the distance
of which to Pk is not greater than its distance to the union of the other Pj, j 6= k.
Definition 2.3. Let (X, d) be a metric space and let K be a set of at least 2
elements (indices), possibly infinite. Given a tuple (Pk)k∈K of nonempty subsets
Pk ⊆ X, a zone diagram with respect to that tuple is a tuple R = (Rk)k∈K of
nonempty subsets Rk ⊆ X such that
Rk = dom(Pk, Sj6=k
Rj) ∀k ∈ K.
In other words, if we define Xk = {C : Pk ⊆ C ⊆ X}, then a zone diagram is a
fixed point of the mapping Dom : Qk∈K
Xk → Qk∈K
Dom(R) = (dom(Pk, Sj6=k
(1)
Xk, defined by
Rj))k∈K.
For example, let P1 = {p1} and P2 = {p2} be two different sets in the Euclidean
plane, or more generally, in a Hilbert space X. In this case the corresponding
Voronoi diagram consists of two half-spaces: dom(P1, P2) is the half-space con-
taining P1 and determined by the hyperplane passing through the middle of the
line segment [p1, p2] and perpendicular to it, and dom(P2, P1) is the other half-
space. The zone diagram of (P1, P2) exists by the results of [3] (in the case of the
Euclidean plane), or [22] (in the general case), but it is not clear how to describe
the zone diagram explicitly.
ZONE DIAGRAMS
5
We now recall the definition of strictly and uniformly convex spaces.
Definition 2.4. A normed space (eX, · ) is said to be strictly convex if for all
x, y ∈ eX satisfying x = y = 1 and x 6= y, the inequality (x + y)/2 < 1 holds.
(eX, · ) is said to be uniformly convex if for any ǫ ∈ (0, 2], there exists δ ∈ (0, 1]
such that for all x, y ∈ eX, if x = y = 1 and x − y ≥ ǫ, then (x + y)/2 ≤ 1 − δ.
Roughly speaking, if the space is uniformly convex, then for any ǫ > 0, there
exists a uniform positive lower bound on how deep the midpoint between any two
unit vectors must penetrate the unit ball, assuming the distance between them is
at least ǫ. In general normed spaces the penetration is not necessarily positive,
since the unit ball may contain line segments. The plane R2 endowed with the
max norm · ∞ is a typical example of this. A uniformly convex space is always
strictly convex, and if it is also finite dimensional, then the converse is true too.
The n-dimensional Euclidean space Rn, or more generally, inner product spaces,
the sequence spaces ℓp, the Lebesgue spaces Lp(Ω), p ∈ (1, ∞), and a uniformly
convex product of a finite number of uniformly convex spaces, are all examples
of uniformly convex spaces. See [7] and, for instance, [16] and [11] for more
information regarding uniformly convex spaces.
We now recall three definitions of a topological character.
Definition 2.5. The Hilbert cube I ∞ is the set I ∞ =Q∞
n=1[0, 1/n] as a topological
space the topology of which is induced by the ℓ2 norm, or, equivalently, by the
product topology.
Definition 2.6. A topological space X is said to be locally (path) connected if
for any x ∈ X and any open set U containing x, there exists a (path) connected
open set V ⊆ U such that x ∈ V . X is said to be weakly locally (path) connected,
or (path) connected im kleinen, if for any x ∈ X and any open set U containing
x, there exists a (path) connected set C ⊆ U and an open set V ⊆ C such that
x ∈ V .
Definition 2.7. Let (X, d) be a metric space. Given two nonempty sets A1, A2 ⊆
X, the Hausdorff distance between them is defined by
D(A1, A2) = max{ sup
a1∈A1
d(a1, A2), sup
a2∈A2
d(a2, A1)}.
Recall that the Hausdorff distance is different from the usual distance between
two sets which is defined by
d(A1, A2) = inf{d(a1, a2) : a1 ∈ A1, a2 ∈ A2}.
Recall also that if (X, d) is compact and we consider the set of all its nonempty
closed subsets, then this space is a compact metric space with the Hausdorff
distance as its metric [13].
We end this section with a definition which is pertinent to the formulation of
It is followed by a brief
our main result and some of our auxiliary assertions.
discussion.
6
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
Definition 2.8. Let X be a closed and convex subset of a normed space. Let
p ∈ X. Let θ ∈ Θp. Let L(θ) ∈ (0, ∞] be the length of the line segment generated
from the intersection of X and the ray emanating from p in the direction of θ.
The point p is said to have the emanation property (or to satisfy the emanation
condition) in the direction of θ if for each ǫ > 0 there exists β > 0 such that for
any φ ∈ Θp, if φ − θ < β, then the intersection of X and the ray emanating
from p in the direction of φ is a line segment of length at least L(θ) − ǫ. In other
words, L(φ) ≥ L(θ) − ǫ. The point p is said to have the emanation property if it
has the emanation property in the direction of every θ ∈ Θp. A subset C of X is
said to have the emanation property if each p ∈ C has the emanation property.
The following examples illustrate the emanation property. In the first four the
emanation property holds, and in the last one it does not hold.
Example 2.9. X is any bounded closed convex set and p ∈ X is an arbitrary
point in the interior of X relative to the affine hull spanned by X.
Example 2.10. The boundary of the bounded closed and convex X is strictly
convex (if a 6= b are two points in the boundary, then the open line segment
(a, b) is contained in the interior of X relative to the affine hull spanned by X)
and p ∈ X is arbitrary. Any ball in a strictly convex space has a strictly convex
boundary.
Example 2.11. X is a cube (of any finite dimension) and p ∈ X is arbitrary.
Example 2.12. X is a closed linear subspace and p ∈ X is arbitrary.
Example 2.13. This example shows that the emanation condition does not hold
in general. Consider the Hilbert space ℓ2. Let (en)∞
n=1 be the standard basis. Let
y1 = e1 and for each n > 1 let yn = e1/2 + en/n. Let A = {−e1}. Let X be
the closed convex hull generated by AS{yn : n = 1, 2, . . .}. Let p = 0 and let
Indeed, limn→∞ θn = θ1 but L(θn) = p0.25 + 1/n2 < 0.99 =
θn = yn/yn for each n. Then p does not satisfy the emanation condition in the
direction of θ1.
L(θ1) − 0.01 for each n > 1. The subset X is in fact compact since the sequence
(yn)∞
n=1 converges.
For more details about the emanation property, see [21].
3. Outline of the proof of the main result
In this section we outline the proof of our main result, Theorem 7.2. It states
that there exists a zone diagram with respect to finitely many pairwise disjoint
compact sites in a compact and convex subset of a uniformly convex normed space,
provided that either the sites or the convex subset satisfy a certain mild condition.
The idea of the proof is to find a certain space Y homeomorphic to the Hilbert
cube I ∞ (Y ≈ I ∞ for short) such that Dom(Y ) ⊆ Y and Dom is continuous on Y .
Now, if h : I ∞ → Y is a homeomorphism, then f = h−1 ◦ Dom ◦ h : I ∞ → I ∞ is
a continuous mapping which maps a compact and convex subset of ℓ2 into itself,
so the Schauder fixed point theorem [23] (see also [10, p. 119] and Theorem 3.1
ZONE DIAGRAMS
7
below) ensures that f has a fixed point q ∈ I ∞. By taking R = h(q), we see that
R is a fixed point of Dom, that is, R is a zone diagram. In order to apply this
idea, one has to find the set Y , to prove that it is homeomorphic to I ∞, and to
prove the continuity of Dom on Y . It has turned out that even in the case of
singleton sites in a square in the Euclidean plane the proof is not obvious (the
main difficulty is to prove the continuity of Dom), and, in fact, such a proof has
never been published.
The above strategy was suggested by the first author, and was briefly mentioned
in [3, p. 1188]. The space Y was taken to beQk∈K Yk, where K was finite, Yk was
{C : Pk ⊆ C ⊆ Qk and C is closed} and Qk was the intersection of the Voronoi
cell of Pk with X (the square). Since each site Pk is taken to be a singleton, it
follows that each Qk is actually convex, so, in particular, it is a connected and
locally connected compact metric space. Since, in addition, Pk 6= Qk, it follows
from the theorem of D. Curtis and R. Schori [8, Theorem 5.2] stated below (see
also [13, p. 91]) that Yk, as a metric space endowed with the Hausdorff metric, is
homeomorphic to I ∞. The topology on Y is the product topology, induced by the
k) : k ∈ K}, so Y ,
as a finite product of spaces homeomorphic to I ∞, is also homeomorphic to I ∞.
uniform Hausdorff metric eD((Sk)k∈K, (S′
k)k∈K) = max{D(Sk, S′
Theorem 3.1. (Schauder) Let X be a nonempty convex and compact subset of
a normed space. If f : X → X is continuous, then it has a fixed point.
Theorem 3.2. (Curtis-Schori) Let X be a Peano continuum, that is, a con-
nected and locally connected compact metric space, and let P ⊆ X, P 6= X be
closed and nonempty. Let 2X
P = {C : P ⊆ C ⊆ X, C is closed}, endowed with the
Hausdorff metric. Then 2X
P ≈ I ∞.
In the general case, the application of the above strategy, and, in particular,
the verification of the hypotheses of Theorem 3.2, are not a simple task, and
they require several additional tools related to dominance regions, such as their
characterization as unions of line segments, and their stability with respect to
small perturbations of the relevant sets. These results will be stated in the next
section. They have recently been established in [19, 21], and their proofs can be
found there. Using these results, we first prove the existence of a zone diagram
with respect to finite sites, and then, approximating compact sets by finite subsets
of them and applying a continuity argument, we extend this existence result to
any compact sites with the emanation property.
4. Several technical tools
In this section we either prove or cite several technical claims needed for estab-
lishing the main result; see [21] and [19] for those proofs not included here.
The following theorem is a new representation theorem for dominance regions.
Theorem 4.1. Let X be a closed and convex subset of a normed space. Let
P, A ⊆ X be nonempty. Suppose that the distance between x and P is attained
8
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
for all x ∈ X. Then dom(P, A) is a union of line segments starting at the points
of P . More precisely, given p ∈ P and θ = 1, let
(2)
T (θ, p) = sup{t ∈ [0, ∞) : p + tθ ∈ X and d(p + tθ, p) ≤ d(p + tθ, A)}.
Then
dom(P, A) = [p∈P [θ=1
[p, p + T (θ, p)θ].
When T (θ, p) = ∞, the notation [p, p+T (θ, p)θ] means the ray {p+tθ : t ∈ [0, ∞)}.
The proof of Theorem 4.1 is based on the following simple observation, which
will also be needed for a different purpose later (see the proof of Lemma 5.2).
Lemma 4.2. Let (eX, · ) be a normed space, and let ∅ 6= A ⊆ eX. Suppose that
y, p ∈ eX satisfy d(y, p) ≤ d(y, A). Then d(x, p) ≤ d(x, A) for any x ∈ [p, y].
The next two theorems describe a continuity property of dominance regions
and the mapping T defined in (2) in uniformly convex normed spaces. We note
that condition (3) below expresses the fact that the set A is "well distributed in
X". It obviously holds when X is bounded, but it may also hold even when X
is not bounded, as in the case where X = Rn and A is the lattice of points with
integer coordinates.
Theorem 4.3. Let X be a closed and convex subset of a uniformly convex normed
space. Then, under certain conditions, the mapping dom(·, ·) has a uniform con-
tinuity property with respect to the Hausdorff distance. More precisely, assume
that P and A are nonempty with d(P, A) > 0. Suppose that
∃ρ ∈ (0, ∞) such that ∀x ∈ X the open ball B(x, ρ) intersects A.
(3)
Then for each ǫ ∈ (0, d(P, A)/6) there exists ∆ > 0 such that if D(A, A′) < ∆,
D(P, P ′) < ∆, and the distances between any point x ∈ X and both P and P ′ are
attained, then the inequality D(dom(P, A), dom(P ′, A′)) < ǫ holds.
Theorem 4.4. Let X be a closed and convex subset of a uniformly convex normed
space. Let p ∈ X and suppose that p has the emanation property. Then the
mapping T (·, p) has a certain continuity property. More precisely, let A ⊂ X
be nonempty such that (3) holds and that d(p, A) > 0. Then for each ǫ ∈
(0, d(p, A)/6) and each θ ∈ Θp there exists ∆ > 0 such that for each φ ∈ Θp,
if θ − φ < ∆, then T (θ, p) − T (φ, p) ≤ ǫ. In addition, the range of the mapping
T is bounded by ρ.
We remark in passing that in the proofs of Theorems 4.3 and 4.4, the uni-
form convexity of the space enters, inter alia, through Clarkson's strong triangle
inequality [7, Theorem 3].
The following lemma will be needed for proving the local connectedness of
certain dominance regions (Lemma 5.1 and Lemma 5.2). Note that the set Bp is
not necessarily the unit ball (or a ball of some closed affine hull), as in the case
where X is the Hilbert cube I ∞ in ℓ2, or more generally, a compact and convex
subset of an infinite dimensional normed space.
ZONE DIAGRAMS
9
Lemma 4.5. Let X be a convex set in a normed space and let p ∈ X. Define the
set Bp = {rθ : r ∈ [0, 1], θ ∈ Θp}. Then Bp is convex.
Proof. Let b1, b2 ∈ Bp. Then bi = riθi for some ri ∈ [0, 1] and θi = 1. Let
b = λ1b1 + λ2b2 for some λ1, λ2 ∈ [0, 1] with λ1 + λ2 = 1. Suppose that b 6= 0, for
otherwise b ∈ Bp. Then b = rθ with r = λ1b1 + λ2b2 and θ = (λ1b1 + λ2b2)/r.
Clearly, 0 ≤ r ≤ r1λ1 + r2λ2 ≤ 1, and θ = 1, so it remains to prove that
p + tθ ∈ X for some t > 0.
By the definition of θi, there is ti > 0 such that yi := p + tiθi ∈ X. We can
assume that λi 6= 0 and ri 6= 0 for i ∈ {1, 2}, for otherwise θ = θj for j 6= i and
p + tjθ ∈ X. Let β1 = λ1r1t2/(λ1r1t2 + λ2r2t1), β2 = λ2r2t1/(λ2r2t1 + λ1r1t2)
and t = (β1t1 + β2t2)r/(λ1r1 + λ2r2). These values were obtained by equating the
coefficients of the θi in the equation p + tθ = β1y1 + β2y2. It is easy to check that
βi ∈ (0, 1), β1 + β2 = 1, t > 0 and p + tθ = β1y1 + β2y2. Thus p + tθ ∈ X because
X is convex.
(cid:3)
The next two lemmata are probably known, at least in a version close in its spirit
to their formulation (see, e.g., [17, p. 162, exercise 6] and [13, Proposition 10.7,
pp. 82-83]), but we include their proof for the sake of completeness.
Lemma 4.6. If Z is a topological space which is weakly locally (path) connected,
then it is locally (path) connected.
Proof. The proof is based on the fact that Z is locally (path) connected if and
only if for every open set U of Z each (path) connected component of U is open in
Z [17, p. 161]. Given an open set U, let x ∈ U. Let Cx be the (path) connected
component of x in U. Since Z is weakly locally (path) connected, there exist a
(path) connected set C ⊆ U and an open set V ⊆ C such that x ∈ V . By the
maximality of Cx, we have C ⊆ Cx, and hence V ⊆ Cx. Thus x belongs to the
interior of Cx, and in the same way all the other points of Cx are in its interior.
Hence Cx is open in X and the same is true for all other (path) components of
U.
(cid:3)
Lemma 4.7. Let (Z, d) be a metric space and suppose that Z =Sm
i=1 Zi for some
locally (path) connected closed subsets Zi of Z. Then Z is locally (path) connected.
Proof. It suffices to prove the assertion for m = 2; the general case follows by
induction. Let x ∈ Z. Then x ∈ Zi for some i. Let U be an open neighborhood
of Z with x ∈ U. Assume first that x /∈ Zj for j 6= i. Then the intersection
U ∩ B(x, d(x, Zj)/2) ⊆ Zi contains an open (path) connected subset V of Zi with
x ∈ V , since Zi is locally (path) connected. Note that V is also (path) connected
with respect to the topology of Z. Since d(V, Zj) > 0, it follows that V is actually
open in Z, so x has an open (path) connected neighborhood contained in U in
this case.
Assume now that x ∈ Z1 ∩ Z2. Then U ∩ Zi is an open neighborhood of x in
Zi, so for each i ∈ {1, 2}, there is an open (path) connected subset Vi ⊆ U ∩ Zi
of Zi with x ∈ Vi, because Zi is locally (path) connected. The union V1 ∪ V2 is a
(path) connected subset of Z which is contained in U, but it is not clear whether
10
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
it is open in Z. However, by definition, Vi = V ′
and
i ∩ Zi for some open set V ′
i of Z,
x ∈ V ′
2 ∩ Z2) ⊆ V1 ∪ V2 ⊆ U.
1 ∩ V ′
2 = (V ′
1 ∩ V ′
2 ∩ Z1) ∪ (V ′
1 ∩ V ′
Since x is arbitrary, this proves that Z is weakly locally (path) connected, so by
Lemma 4.6 it is locally (path) connected.
(cid:3)
The following lemma will be useful for proving the main result of Section 5.
Lemma 4.8. Let X be a convex set in a normed space, and suppose that P, A ⊂ X
satisfy r := d(P, A) > 0. Let B(P, r/2) := {x ∈ X : d(x, P ) < r/2}. Then
B(P, r/2) ⊆ {x ∈ X : d(x, P ) < d(x, A)} and d(x, P ) ≥ r/2 for any x ∈ X
satisfying x ∈ dom(A, P ).
Proof. The set B(P, r/2) is contained in {x ∈ X : d(x, P ) < d(x, A)} because if
x ∈ B(P, r/2) and a ∈ A, then r ≤ d(a, P ) ≤ d(a, x) + d(x, P ) < d(a, x) + r/2, so
d(x, P ) < r/2 ≤ d(x, A).
Now let x ∈ dom(A, P ). Since d(x, A) ≤ d(x, P ), this point does not belong to
B(P, r/2) by the above paragraph. Let p ∈ P be arbitrary, and consider the line
segment [x, p]. It intersects the boundary of B(P, r/2) (otherwise the connected
space [x, p] would have a decomposition as a union of two disjoint open sets) at
some point y, and since B(P, r/2) is open, it follows that d(y, P ) ≥ r/2. Hence
d(x, p) = d(x, y) + d(y, p) ≥ r/2, i.e., d(x, P ) ≥ r/2, as claimed.
(cid:3)
We finish with the following lemma, the proof of which is a simple consequence
of the definition.
Lemma 4.9. The equality dom(Sn
i=1 Pi, A) = Sn
nonempty subsets P1, . . . , Pn and A of the metric space X.
i=1 dom(Pi, A) holds for any
5. The space homeomorphic to the Hilbert cube
The main result of this section is Proposition 5.4 below. It is based on Lemma 5.2
which shows that dom(p, A) is locally path connected whenever d(p, A) > 0, and
on Lemma 5.3 which generalizes this result to dom(P, A) for any finite set P . As
Lemma 4.2 shows, dom(p, A) is star-shaped. However, this property by itself is
not sufficient for concluding that it is locally (path) connected, since there are
simple examples of star-shaped sets in R2 which are not locally connected.
In
fact, a result of T. Zamfirescu [24, Theorem 4] shows that a large class (in the
sense of Baire category) of star-shaped sets in Rn are not locally connected.
As a result, the local path connectedness of dom(p, A) must be proved. It has
turned out that the proof, given in Lemma 5.2, is somewhat technical, and in
a special but important case, namely Lemma 5.1 below, a simpler proof can be
presented. This special case is of interest in itself, and its proof also casts some
light on the strategy for proving Lemma 5.2. The condition on p and s described
in Lemma 5.1 holds, for instance, when p is in the interior of X relative to the
closed affine hull spanned by it. See also the short discussion before Lemma 4.5
regarding the set Bp.
ZONE DIAGRAMS
11
Figure 3. An example related
to Lemma 5.1 in a square in
(R2, ℓ7). Here p = pk, A is the
collection of the other singleton
sites pj, j 6= k, and dom(p, A) is
the Voronoi cell of pk. The set
Bp is simply the unit ball.
4. An
the
conclusion
example
Figure
of
where
Lemma 5.1 fails if the uni-
form convexity assumption is
removed. Here X is a square
in (R2, ℓ∞), p = (0, 0), and
A = {(2, 0), (−2, 0), (0, −2)}.
The green region is dom(p, A).
Lemma 5.1. Let X be a closed and convex subset of a uniformly convex normed
space. Let p ∈ X. Suppose that there exists some s > 0 such that p + sθ ∈ X for
each θ ∈ Θp. Suppose also that p has the emanation property. Let ∅ 6= A ⊂ X
be such that d(p, A) > 0. Suppose also that condition (3) holds. Then dom(p, A)
is homeomorphic to the set Bp = {rθ : r ∈ [0, 1], θ ∈ Θp}, and, in particular,
dom(p, A) is path connected and locally path connected.
Proof. Let f : Bp → dom(p, A) be defined by f (rθ) = p + rT (θ, p)θ, i.e., f (0) = p
and f (x) = p + T (x/x, p)x for x 6= 0, where T (θ, p) is defined in (2). Let
ǫ = min{d(p, A)/2, s/2}. Let θ ∈ Θp and t ∈ [0, ǫ] be arbitrary. Then for
x = p + tθ we have x ∈ X because t ≤ s and X is convex. In addition, 2ǫ ≤
d(p, A) ≤ d(p, x) + d(x, A) ≤ ǫ + d(x, A), so d(x, p) ≤ ǫ ≤ d(x, A) and hence
x ∈ dom(p, A). Thus T (θ, p) ≥ ǫ by the definition of T . From Theorem 4.4 it
follows that T is bounded, so f is well defined. By Theorem 4.1, f is onto, and it
is one-to-one by a direct calculation. If y = p + rT (θ, p)θ, then θ = (y − p)/y − p
and r = y − p/T (θ, p), so the inverse function is defined by
f −1(y) =(cid:26)
y−p
T ((y−p)/y−p,p) y 6= p
y = p.
0
It now follows from Theorem 4.4 that both f and f −1 are continuous. Since
Bp is convex by Lemma 4.5, it is path connected and locally path connected, so
dom(p, A) is both path connected and locally path connected, as claimed.
(cid:3)
12
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
An example related to Lemma 5.1 is given in Figure 3. This may be somewhat
surprising, but in general the conclusion of Lemma 5.1 is not true, and Figure 4
presents a counterexample in the non-uniformly convex space (R2, ℓ∞) with two
simple sets.
An examination of the above proof suggests that problems can appear if no
positive number s satisfies the condition in the formulation of Lemma 5.1, because
in this case the function f −1 may not be continuous. Such a phenomenon occurs
in infinite dimensional spaces, for instance, when X is the Hilbert cube I ∞ in ℓ2,
but cannot happen if p is in the interior of X relative to the closed affine hull
spanned by it. In order to replace Lemma 5.1 one has to show directly by other
arguments that dom(p, A) is connected and locally connected. This will be done
in Lemma 5.2 below.
Lemma 5.2. Let X be a closed and convex set in a uniformly convex normed
space. Let p ∈ X and ∅ 6= A ⊂ X and suppose that d(p, A) > 0. Suppose also
that condition (3) holds and that p has the emanation property. Then dom(p, A)
is path connected and locally path connected.
Proof. By Lemma 4.2, any two points x1, x2 ∈ dom(p, A) can be connected via
p, so dom(p, A) is path connected.
In order to prove that dom(p, A) is locally
path connected it suffices by Lemma 4.6 to show that it is weakly locally path
connected. Let x0 ∈ dom(p, A).
If x0 = p, then let V = B(x0, r) ∩ X for
r = d(p, A)/3. Given y ∈ V , the triangle inequality shows that 3r = d(p, A) ≤
d(p, y) + d(y, A) ≤ r + d(y, A). Hence d(y, p) ≤ r < 2r ≤ d(y, A) and y ∈
dom(p, A), so V ⊆ dom(p, A). Since V is convex, x0 = p has a path connected
open neighborhood.
Now suppose that x0 6= p, and for any ǫ > 0 let U = B(x0, ǫ) ∩ dom(p, A) be
any neighborhood of x0. Let θ0 = (x0 − p)/x0 − p. Then θ0 = 1, and since x0 =
p+x0 −pθ0 ∈ dom(p, A), the definition of T (θ0, p) (in (2)) implies that d(x0, p) ≤
T (θ0, p). Hence x′
0 := (x0 −p)/T ((x0 −p)/x0 −p, p) = (x0 −p/T (θ0, p))θ0 is well
defined, different from 0 and belongs to Bp = {rθ : r ∈ [0, 1], θ = 1, p + tθ ∈
X for some t > 0}. In addition, x′ > 0 for each x′ ∈ B(x′
0/2) ∩ Bp.
0, x′
Since T (θ0, p) > 0, the continuity of T (·, p) implies that there exists δ2 ∈ (0, 1)
such that if θ ∈ Θp and θ − θ0 < δ2, then T (θ, p) > 0. Let δ3 = δ2x0 − p/2.
0, x′
Define f : B(x′
0 < δ1, then f (x′) − f (x′
T (·, p) is continuous (Theorem 4.4) and since f (x′
continuous, so for the given ǫ > 0, there exists some δ1 ∈ (0, x′
x′ − x′
Bp. Since U ′
continuity of f implies that f (U ′
containing x0 = f (x′
dom(p, A), it suffices to show that f (U ′
0/2) ∩ Bp → dom(p, A) by f (x′) = p + T (x′/x′, p)x′. Since
0) = x0, it follows that f is
0/2) such that if
0, δ1) ∩
0, the
1) is a path connected set contained in U and
0). In order to establish the weak (path) connectedness of
1) contains an open neighborhood of x0.
1 = B(x′
1 is a path connected (in fact, convex) open set containing x′
0) < ǫ. Hence f (U ′
1) ⊂ U for U ′
Given x ∈ B(x0, δ3) ∩ dom(p, A), we have
ZONE DIAGRAMS
13
(4)
x − p
x − p
−
x0 − p(x − p) − x − p(x0 − p)
x − px0 − p
=
(cid:12)(cid:12)(cid:12)(cid:12)
x0 − p
x0 − p(cid:12)(cid:12)(cid:12)(cid:12) =
(x0 − p − x − p)(x − p) + x − p((x − p) − (x0 − p))
x − px0 − p
≤
2x0 − x
x0 − p
< δ2.
Thus T ((x − p)/x − p, p) > 0 for each x ∈ B(x0, δ3) ∩ dom(p, A). For each
such x, let g(x) = (x − p)/T ((x − p)/x − p, p). Since g is continuous, for δ1
from the definition of U ′
1 there exists δ4 ∈ (0, δ3) such that if x − x0 < δ4,
then g(x) − g(x0) < δ1. But g(x0) = x′
1 for each x in the set
V := B(x0, δ4) ∩ dom(p, A). In addition, by the definition of f and g we have
x = f (g(x)) for each x ∈ V (just let θ = (x − p)/x − p and x′ = g(x), and
note that x′/x′ = θ). Hence V is an open neighborhood of x0 which is contained
in f (U ′
1), so dom(p, A) is weakly locally path connected and hence locally path
connected by Lemma 4.6.
(cid:3)
0 and g(x) ∈ U ′
Lemma 5.3. Let X be a closed and convex subset of a uniformly convex normed
space. Let P = {p1, . . . , pm} ⊂ X and assume that ∅ 6= A ⊂ X satisfies d(P, A) >
0. Suppose also that condition (3) holds and that the subset P has the emanation
property. Then dom(P, A) is locally path connected.
Sm
Proof. This assertion is a simple consequence of the facts that dom(P, A) =
i=1 dom(pi, A)) (by Lemma 4.9), that dom(pi, A) is locally path connected (by
Lemma 5.2) and closed for each i, and the fact that the metric space dom(P, A)
is a finite union of locally path connected closed sets and hence locally path con-
nected by Lemma 4.7. Note that the topology of dom(pi, A), which is induced by
the norm of the space, coincides with its topology as a subspace of dom(P, A),
and hence we can apply Lemma 4.7.
(cid:3)
Proposition 5.4. Let X be a compact and convex subset of a uniformly convex
normed space, and let (Pk)k∈K be a finite tuple of finite sets in X which are
pairwise disjoint. Suppose that for each k ∈ K the site Pk has the emanation
{C : C is closed and Pk ⊆ C ⊆ Qk}, endowed with the Hausdorff metric. Let
property. For each k ∈ K, let Ak = Sj6=k Pj and Qk = dom(Pk, Ak). Let Yk =
Y =Qk∈K Yk, endowed with the uniform Hausdorff metric. Then Y ≈ I ∞.
Proof. Given k ∈ K and p ∈ Pk, let Cp be the path connected component of p in
Qk, and let Ep = {q ∈ Pk : q ∈ Cp}. Since d(Pk, Ak) > 0 for each k by assumption,
we have d(q, Ak) > 0 for any q ∈ Ep, so dom(q, Ak) is path connected by Lemma
5.2. Thus dom(q, Ak) ⊆ Cp for each q ∈ Ep, and hence dom(Ep, Ak) ⊆ Cp by
Lemma 4.9. On the other hand, if x ∈ Cp, then x ∈ dom(p′, Ak) for some p′ ∈ Pk,
dom(p, Ak) by Lemma 4.9. Lemma 5.2 implies that there is
a path between p′ and x, and since x ∈ Cp, there is a path between x and p. Thus
p′ ∈ Ep, so x ∈ dom(p′, Ak) ⊆ dom(Ep, Ak) and hence dom(Ep, Ak) = Cp.
because Qk =Sp∈Pk
Since for each p1, p2 ∈ Pk, either Ep1 = Ep2 or Ep1 ∩ Ep2 = ∅, it follows that
{Ep : p ∈ Pk} = {Ek,1, . . . , Ek,mk} for some disjoint sets Ek,l, l = 1, . . . , mk,
14
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
the union of which is Pk. Hence the path connected components of Qk are
dom(Ek,1, Ak), . . . , dom(Ek,mk, Ak), and since they are closed and disjoint in the
compact set X, they are, in fact, compact and positively separated. Since Ek,l ⊆
Pk, we have d(Ek,l, Ak) ≥ d(Pk, Ak) > 0 for each k and each l. Hence, by Lemma
5.3, the sets dom(Ek,l, Ak) are also locally path connected.
By Lemma 4.8, we have Ek,l & dom(Ek,l, Ak) for each l = 1, . . . , mk and
each k, because there are points x with d(x, Ek,l) = r/4 for r = d(Ek,l, Ak)
and each such point is in dom(Ek,l, Ak), but not in Ek,l. Consequently, Theo-
rem 3.2 implies that 2dom(Ek,l,Ak)
≈ I ∞. Using the fact that the sets dom(Ek,1, Pk),
dom(Ek,2, Pk),. . . , dom(Ek,mk, Pk) are positively separated (and also the fact that
l=1(C ∩dom(Ek,l, Ak))
as a disjoint union of closed sets contained in the components dom(Ek,l, Ak)), we
can easily verify that Yk with the Hausdorff metric is homeomorphic to the fi-
, endowed with the uniform Hausdorff metric.
(cid:3)
any closed set C in Yk has the unique decomposition C =Smk
nite product spaceQmk
Therefore Yk ≈ (I ∞)mk ≈ I ∞, and hence Y ≈ I ∞.
l=1 2dom(Ek,l,Ak)
Ek,l
Ek,l
6. Continuity of the Dom mapping
Proposition 6.1. Let X be a convex subset of a uniformly convex normed space,
and let (Pk)k∈K be a tuple of nonempty and positively separated sets in X, that
is, inf{d(Pk, Pj) : j, k ∈ K, j 6= k} > 0. Assume that condition (3) holds (with the
that the distance between each x ∈ X and each Pk, k ∈ K, is attained. For each
same ρ) for each k ∈ K with Pk and Ak =Sj6=k Pj instead of P and A. Suppose
k ∈ K let Qk = dom(Pk,Sj6=k Pj) and Yk = {C : C is closed and Pk ⊆ C ⊆ Qk}.
Let Y = Qk∈K Yk, endowed with the uniform Hausdorff metric eD defined by
eD((Sk)k∈K, (S′
k) : k ∈ K}. Then Dom maps Y into itself
and is uniformly continuous there.
k)k∈K) = sup{D(Sk, S′
Proof. First, note that for any nonempty subsets P, A, B of X, if A ⊆ B, then
dom(P, B) ⊆ dom(P, A). Now let S = (Sk)k∈K ∈ Y be given. By the defini-
tion of Yk, we have Pk ⊆ Sk for each k ∈ K, so Wk := dom(Pk,Sj6=k Sj) ⊆
dom(Pk,Sj6=k Pj) = Qk by the definition of Qk. Since the k-th component of
W := Dom(S) is the closed subset Wk, we conclude that (Pk)k∈K ⊆ Dom(S) ⊆
(Qk)k∈K, and hence Dom maps Y into itself.
j) ≤ sup{D(Sj, S′
correspond to ǫ in Theorem 4.3. Let S = (Sk)k∈K, S′ = (S′
Now let ǫ > 0 satisfy 12ǫ ≤ inf{d(Pk, Pj) : j, k ∈ K, j 6= k}, and let ∆
k)k∈K be any two tuples
k) < ∆ for each
j) ≤ ∆ for each k ∈ K, because
in Y satisfying eD(S, S′) < ∆. As a result, we also have D(Sk, S′
k ∈ K. This implies that D(Sj6=k Sj,Sj6=k S′
D(Sj6=k Sj,Sj6=k S′
If x ∈ Sj6=k Sj, then x ∈ Sj for some j 6= k, so x ∈ Qj =
dom(Pj,Si6=j Pi) ⊆ dom(Pj, Pk). Hence d(x, Pk) ≥ 6ǫ by Lemma 4.8. Thus
d(Pk,Sj6=k Sj) ≥ 6ǫ for each k ∈ K. Hence all the conditions of Theorem 4.3
are satisfied (here P = Pk = P ′, A = Sj6=k Sj, A′ = Sj6=k S′
(3) of Theorem 4.3 for A follows from condition (3) for Ak = Sj6=k Pj because
j and condition
j) : j 6= k}.
Fix k ∈ K.
ZONE DIAGRAMS
15
Pj ⊆ Sj). Thus D(dom(Pk,Sj6=k Sj), dom(Pk,Sj6=k S′
the definition of Dom and eD, we have eD(Dom(S), Dom(S′)) ≤ ǫ, so Dom is indeed
k)) < ǫ for each k ∈ K. By
uniformly continuous on Y .
(cid:3)
7. The main result
Proposition 7.1. Let X be a compact and convex subset of a uniformly convex
normed space, and let (Pk)k∈K be a finite tuple of finite sets in X which are
pairwise disjoint. Suppose that for each k ∈ K, the site Pk has the emanation
property. Then there exists a zone diagram with respect to these sites.
Proof. Given the finite sites (Pk)k∈K, let Y be as in Proposition 6.1. Since Dom
maps Y into itself and is continuous there by Propostion 6.1, and since Y is
homeomorphic to the Hilbert cube by Proposition 5.4, we deduce the existence
of a zone diagram from the Schauder fixed point theorem, as explained in the
beginning of Section 3.
(cid:3)
Theorem 7.2. Let X be a compact and convex subset of a uniformly convex
normed space, and let (Pk)k∈K be a finite tuple of compact sets in X which are
pairwise disjoint. Suppose that for each k ∈ K, the site Pk has the emanation
property. Then there exists a zone diagram with respect to these sites.
Proof. Let (Pk)k∈K be the given finite tuple of pairwise disjoint compact sites.
Let r = min{d(Pk, Pj) : j 6= k}. By the compactness of the sites, for each positive
integer m and for each k ∈ K, there exists a finite subset Pm,k of Pk such that
D(Pm,k, Pk) < 1/m. Note that d(Pm,k, Pm,j) ≥ r whenever j 6= k. For each m,
let Rm = (Rm,k)k∈K be a zone diagram corresponding to the tuple (Pm,k)k∈K, the
existence of which is guaranteed by Proposition 7.1. In other words,
(5)
Rm,k = dom(Pm,k,[j6=k
Rm,j) ∀k ∈ K.
Since the space of all nonempty compact subsets of X (endowed with the Haus-
dorff metric) is compact [13], and since K is finite, we can find a convergent
subsequence of the finite sites and the components of the corresponding zone di-
agrams. Therefore for some subsequence (ml)∞
l=1 of positive integers the sequence
(Rml,k)∞
l=1 converges to
l=1 converges to some set Rk, and the sequence (Pml,k)∞
Pk by the definition of Pm,k. Hence Sj6=k Rml,j converges to Sj6=k Rj, because
D(Sj6=k Rml,j,Sj6=k Rj) ≤ max{D(Rml,j, Rj) : j 6= k}.
and x ∈Sj6=k Rml,j, then x ∈ Rml,j for some j 6= k. Therefore
Let ǫ′ = r/12. Then 12ǫ′ ≤ min{d(Pml,k, Pml,j) : j 6= k}. Let k ∈ K. If j 6= k
x ∈ dom(Pml,j,[i6=j
Rml,i) ⊆ dom(Pml,j,[i6=j
Pml,i) ⊆ dom(Pml,j, Pml,k).
Hence d(x, Pml,k) ≥ 6ǫ′ by Lemma 4.8, and, as a result, d(Pml,k,Sj6=k Rml,j) ≥
6ǫ′ for each k ∈ K and l. Thus d(Pk,Sj6=k Rj) ≥ 6ǫ′ and hence the positive
distance condition in Theorem 4.3 holds. From (5) and the continuity of dom(·, ·)
16
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
(Theorem 4.3) we conclude that R = (Rk)k∈K is a zone diagram with respect to
(Pk)k∈K.
(cid:3)
The hypotheses of Theorem 7.2 are satisfied, in particular, when all the compact
sites are contained in the interior of a closed ball (or another compact and convex
set) in Rn.
8. Concluding remarks and open problems
In this short section we describe several interesting questions and directions for
further investigation.
An interesting problem is whether the dominance region dom(P, A) is locally
connected for general positively separated sets P and A.
If so, then this will
suggest an alternative method for proving the existence of a zone diagram with
respect to compact sites. Second, it would be interesting to extend the existence
result further, say to all normed spaces, without any compactness requirement
on the sites and the subset X. The question of uniqueness is interesting too. It
seems that the uniform convexity assumption on the norm is not sufficient even
in the plane with two singleton sites, as shown in [15], but perhaps, following [15],
uniform convexity combined with uniform smoothness of the norm will imply
uniqueness in the infinite dimensional case too. It would also be of interest to
establish our main theorem (Theorem 7.2) and auxiliary results (e.g., Lemmata
5.1 and 5.2) without imposing the condition of the emanation property on the
points.
Another interesting and natural problem is how to approximate a zone diagram
in the setting described in this paper. It turns out that there is a way to do it,
and a key point in carrying out this task is to use the algorithm for computing
Voronoi diagrams of general sites in general spaces described in [19, 20]. For the
sake of completeness, we include two pictures of zone diagrams in the plane with
two different norms. The proof of Theorem 7.2 shows that one can approximate
the given compact sites by finite subsets of them and then the corresponding
zone diagram approximates the real one. Unfortunately, no error estimates are
obtained in the proof, and it would indeed be of interest to find such estimates.
Acknowledgements
We thank Akitoshi Kawamura, Jir´ı Matousek, and Takeshi Tokuyama for pro-
viding a copy of their recent paper [15] before posting it on the arXiv. The first
author was supported by Grant FWF-P19643-N18. The third author was par-
tially supported by the Israel Science Foundation (Grant 647/07), the Fund for
the Promotion of Research at the Technion (Grant 2001893), and by the Technion
President's Research Fund (Grant 2007842). All the authors thank Orr Shalit and
the referee for their helpful suggestions and corrections. We are very grateful to
Eva Goldman for kindly redrawing the figures so expertly.
ZONE DIAGRAMS
17
Figure 5. A zone diagram of
5 sites, each with 4 points, in a
square in (R2, ℓ6).
Figure 6. A zone diagram of 3
sites in a square in (R2, ℓp), p =
3.14159, each with 3 points.
References
1. T. Asano and D. Kirkpatrick, Distance trisector curves in regular convex distance met-
rics, Proceedings of the 3rd International Symposium on Voronoi Diagrams in Science and
Engineering (ISVD 2006), pp. 8 -- 17.
2. T. Asano, J. Matousek, and T. Tokuyama, The distance trisector curve, Adv. Math. 212
(2007), 338 -- 360, a preliminary version in STOC 2006, pp. 336-343.
3.
, Zone diagrams: Existence, uniqueness, and algorithmic challenge, SIAM J. Com-
put. 37 (2007), no. 4, 1182 -- 1198, a preliminary version in SODA 2007, pp. 756-765.
4. T. Asano and T. Tokuyama, Drawing equally-spaced curves between two points, Abstracts of
the 14th Annual Workshop on Computational Geometry with a Focus on Open Problems,
2004, pp. 24 -- 25.
5. F. Aurenhammer, Voronoi diagrams - a survey of a fundamental geometric data structure,
ACM Computing Surveys, vol. 3, 1991, pp. 345 -- 405.
6. J. Chun, Y. Okada, and T. Tokuyama, Distance trisector of segments and zone diagram of
segments in a plane, Proceedings of the 4th International Symposium on Voronoi Diagrams
in Science and Engineering (ISVD 2007), pp. 66 -- 73.
7. J. A. Clarkson, Uniformly convex spaces, Trans. Amer. Math. Soc. 40 (1936), 396 -- 414.
8. D. W. Curtis and R. M. Schori, Hyperspaces of Peano continua are Hilbert cubes, Fund.
Math. 101 (1978), no. 1, 19 -- 38.
9. S. C. de Biasi, B. Kalantari, and I. Kalantari, Maximal zone diagrams and their computa-
tion, Proceedings of the 7th international symposium on Voronoi Diagrams in Science and
Engineering (ISVD 2010), pp. 171 -- 180.
10. J. Dugundji and A. Granas, Fixed point theory, Springer Monographs in Mathematics,
Springer-Verlag, New York, 2003.
11. K. Goebel and S. Reich, Uniform convexity, hyperbolic geometry, and nonexpansive map-
pings, Monographs and Textbooks in Pure and Applied Mathematics, vol. 83, Marcel Dekker
Inc., New York, 1984.
12. C. Gold, The Voronoi Web Site, 2008, http://www.voronoi.com/wiki/index.php?title=Main
Page.
18
EVA KOPECK ´A, DANIEL REEM, AND SIMEON REICH
13. A. Illanes and S. B. Nadler Jr, Hyperspaces. Fundamentals and recent advances, Monographs
and Textbooks in Pure and Applied Mathematics, vol. 216, Marcel Dekker Inc., New York,
1999.
14. K. Imai, A. Kawamura, J. Matousek, Y. Muramatsu, and T. Tokuyama, Distance k-
sectors and zone diagrams, Extended abstract in EuroCG 2009, pp. 191 -- 194. Available
at http://www.cs.toronto.edu/∼kawamura/publ/090317.
15. A. Kawamura, J. Matousek, and T. Tokuyama, Zone diagrams in Euclidean spaces and
in other normed spaces, Mathematische Annalen, to appear. Preliminary versions in SoCG
2010, pp. 216-221, arXiv 0912.3016 (2009).
16. J. Lindenstrauss and L. Tzafriri, Classical Banach spaces, II: Function spaces, Springer,
Berlin, 1979.
17. J. Munkres, Topology, second ed., Prentice Hall, Upper Saddle River, NJ, 2000.
18. A. Okabe, B. Boots, K. Sugihara, and S. N. Chiu, Spatial tessellations: concepts and ap-
plications of Voronoi diagrams, second ed., Wiley Series in Probability and Statistics, John
Wiley & Sons Ltd., Chichester, 2000, with a foreword by D. G. Kendall.
19. D. Reem, An algorithm for computing Voronoi diagrams of general generators in general
normed spaces, Proceedings of the sixth international symposium on Voronoi diagrams in
science and engineering (ISVD 2009), pp. 144 -- 152.
20.
21.
, Voronoi and zone diagrams, Ph.D. thesis, The Technion, Haifa, 2010.
, The geometric stability of Voronoi diagrams with respect to small changes of the
sites, (2011), Complete version in arXiv 1103.4125, Extended abstract in SoCG 2011, pp.
254-263.
22. D. Reem and S. Reich, Zone and double zone diagrams in abstract spaces, Colloquium
Mathematicum 115 (2009), 129 -- 145, arXiv 0708.2668 (2007).
23. J. Schauder, Der Fixpunktsatz in Funktionalraumen, Studia Math. 2 (1930), 171 -- 180.
24. T. Zamfirescu, Typical starshaped sets, Aequ. Math. 36 (1988), 188 -- 200.
Institute of Mathematics, Czech Academy of Sciences, Zitn´a 25, CZ-11567
Prague, Czech Republic, and Institut fur Analysis, Johannes Kepler Univer-
sitat, Altenbergerstrasse 69, A-4040 Linz, Austria.
E-mail address: [email protected]
The Technion - Israel Institute of Technology, 32000 Haifa, Israel. Current
address (July 2011): Department of Mathematics, University of Haifa, Mount
Carmel, 31905 Haifa, Israel.
E-mail address: [email protected]
The Technion - Israel Institute of Technology, 32000 Haifa, Israel
E-mail address: [email protected]
|
1406.6712 | 2 | 1406 | 2015-06-19T12:57:30 | Stability of low-rank matrix recovery and its connections to Banach space geometry | [
"math.FA",
"cs.IT",
"cs.IT"
] | There are well-known relationships between compressed sensing and the geometry of the finite-dimensional $\ell_p$ spaces. A result of Kashin and Temlyakov can be described as a characterization of the stability of the recovery of sparse vectors via $\ell_1$-minimization in terms of the Gelfand widths of certain identity mappings between finite-dimensional $\ell_1$ and $\ell_2$ spaces, whereas a more recent result of Foucart, Pajor, Rauhut and Ullrich proves an analogous relationship even for $\ell_p$ spaces with $p < 1$. In this paper we prove what we call matrix or noncommutative versions of these results: we characterize the stability of low-rank matrix recovery via Schatten $p$-(quasi-)norm minimization in terms of the Gelfand widths of certain identity mappings between finite-dimensional Schatten $p$-spaces. | math.FA | math |
STABILITY OF LOW-RANK MATRIX RECOVERY AND ITS
CONNECTIONS TO BANACH SPACE GEOMETRY
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
Abstract. There are well-known relationships between compressed sensing and the ge-
ometry of the finite-dimensional ℓp spaces. A result of Kashin and Temlyakov [20] can
be described as a characterization of the stability of the recovery of sparse vectors via ℓ1-
minimization in terms of the Gelfand widths of certain identity mappings between finite-
dimensional ℓ1 and ℓ2 spaces, whereas a more recent result of Foucart, Pajor, Rauhut and
Ullrich [16] proves an analogous relationship even for ℓp spaces with p < 1. In this paper
we prove what we call matrix or noncommutative versions of these results: we charac-
terize the stability of low-rank matrix recovery via Schatten p-(quasi-)norm minimization
in terms of the Gelfand widths of certain identity mappings between finite-dimensional
Schatten p-spaces.
1. Introduction
A mathematical problem that appears often in real-world situations is the following:
we wish to recover a high-dimensional vector x ∈ RN from a measurement Ax where
A : RN → Rm is a linear map and m is smaller than N . As stated the problem of course
cannot be solved, but that changes if we have the additional condition that the unknown
vector x is sparse, i.e. it has a small number of non-zero coordinates. This is the subject
matter of compressed sensing, a very active area of research with numerous applications;
the book [17] is a recent comprehensive reference. Formally, this sparse recovery problem
can be stated as
(1.1)
minkxk0
subject to Ax = y,
where k·k0 represents the number of nonzero coordinates of a vector. This is an NP-hard
[27] and non-convex problem, so we are interested in conditions (especially on the map A)
that would allow us to solve an easier problem and still arrive to the right solution. In that
spirit, a basic technique in compressed sensing is that of ℓ1-minimization: if the vector x is
sparse enough, then minimizing kx′kℓ1 over all vectors x′ so that Ax′ = Ax actually allows
us to recover x. Formally, instead of problem (1.1) we are considering its convex relaxation
(1.2)
minkxkℓ1
subject to Ax = y.
Aditionally, one can consider the analogous problem of ℓp-minimization.
In practice the unknown vectors are not necessarily sparse, but are close to sparse ones.
Thus for any method of recovery it is of utmost importance to investigate its stability, that
is, having a control on the distance between the original vector and its reconstruction in
terms of the distance from the original vector to the sparse vectors. It turns out that the
1
2
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
stability of sparse vector recovery through ℓp-minimization has connections to the Banach-
space geometry of finite-dimensional ℓp-spaces. More generally, it is known that there are
connections between recovery -- in particular the compressed sensing model -- and geometric
quantities called Gelfand widths, see e.g. [28, 10, 9, 20].
In many practical situations, there is extra structure in the space of unknown vectors. A
good example is the famous matrix completion problem (also known as the Netflix problem),
where the unknown is a matrix and the measurement map gives us a subset of its entries.
In this case sparsity gets replaced by the more natural condition of having low rank, and
the last few years have witnessed an explosion of work in this area. In what follows, MN
will denote the space of N × N real-valued matrices. We now consider a linear operator
A : MN → Rm, and a fixed vector y ∈ Rm. The low-rank recovery problem can thus be
stated as the problem of finding the solution to
(1.3)
min rank(X)
subject to AX = y.
This is again an NP-hard problem, so once again we would like to replace it by another one
which is simpler to solve but has the same solution.
In noncommutative functional analysis the Schatten p-spaces are usually considered to be
the counterparts of the classical ℓp spaces (recall that the Schatten p-norm of a matrix X is
the ℓp-norm of its vector of singular values), so from that point of view it is natural to wonder
whether the Schatten p-norm minimization approach can work in the matrix context. We
would like to consider operators A for which the previous problem is equivalent to
(1.4)
minkXkS1
subject to AX = y.
Where kXkSp denotes the Schatten p-norm of the matrix X ∈ MN . This has already been
studied in several situations of interest, with the idea going back to the Ph.D. thesis of M.
Fazel [13]. Schatten 1-norm (also known as nuclear norm) minimization in the particular
case of the matrix completion problem was studied by Cand`es and Recht [4] (and later
on Cand`es and Tao [7] gave optimality results quantifying the minimum number of entries
needed to recover a matrix of low rank exactly by any method whatsoever, and showed
that nuclear norm minimization is nearly optimal). Plenty of concepts from the classical
theory of compressed sensing have found matrix counterparts: Cand`es and Recht [4] use
the idea of coherence; Recht, Fazel and Parrilo [32] used the matrix version of the restricted
isometry property [32]; whereas both Recht, Xu and Hassibi [33] and Fornasier, Rauhut
and Ward [14] consider null-space conditions; the spherical section property was used by
Dvijotham and Fazel [11] and Oymak, Mohan, Fazel and Hassibi [29].
One thing that does not appear to have been explicitly studied in the matrix context is
the aforementioned relationship to Gelfand widths. Recall that the Gelfand k-width of a
subset K of a normed space E is defined as
dk(K, E) := inf(cid:26) sup
x∈K∩LkxkE : L subspace of E with codim(L) ≤ k(cid:27) .
A closely related concept that is more commonly used in Banach space theory is that of a
Gelfand number : if T : X → Y is a linear operator between normed spaces, its k-Gelfand
STABILITY OF LOW-RANK MATRIX RECOVERY
3
number is defined by
ck(T ) := inf( sup
x∈L,kxk≤1 kT xk : L subspace of X with codim(L) < k) .
The speed of convergence to zero of the sequence of Gelfand numbers (ck(T ))∞
k=1 is a measure
of the compactness of the operator T , and is an example of a sequence of s-numbers; see
[30, 22] for more details. In the cases under consideration in this paper the concepts of
Gelfand numbers and Gelfand widths actually coincide (up to a small shift in the index), so
we will freely use them both depending on the particular context. It should be mentioned
that there is a general concept of Gelfand width for a linear map that is not always the
same as the corresponding Gelfand number (see [31, Sec. 6.2.6] for the details), but both
concepts do coincide in nice situations (see [12]).
The work of Kashin and Temlyakov [20] made more precise the already-known connection
between compressed sensing and the Kashin-Garnaev-Gluskin [19, 18] result that calculates
the m-Gelfand numbers of the identity map from ℓN
1 to ℓN
2 , namely
cm+1(id : ℓN
1 → ℓN
2 ) ≤ Cr1 + log(N/m)
m
.
In a nutshell, the main result of Kashin and Temlyakov shows that the stability of sparse
recovery via ℓ1-minimization is equivalent to the kernel of the measurement map being a
"good" subspace where the Gelfand number of a certain order is achieved. This idea was
taken further by Foucart, Pajor, Rauhut and Ullrich [16], who used compressed sensing
ideas to calculate the Gelfand numbers of identity maps from ℓN
for 0 < p ≤ 1,
p < q ≤ 2.
In this paper, we prove matrix versions of the aforementioned theorems, relating the
stability of low-rank matrix recovery to the Gelfand numbers of identity maps between
finite-dimensional Schatten p-spaces. As far as we know the only part of our results that
is already written down in the literature is the following analogue of the Kashin-Garnaev-
Gluskin result due to Carl and Defant [8, p. 252], namely the calculation of the m-Gelfand
numbers of the identity map from SN
p to ℓN
q
1 to SN
cm(id : SN
1 → SN
2 : for 1 ≤ m ≤ N 2,
m(cid:27)1/2
2 ) ≍ min(cid:26)1,
N
.
Here and in the rest of the paper, the symbol ≍ means that the quantities on the left and
the right are equivalent up to universal constants. If we want to emphasize the dependance
of the constants on some parameters, those will appear as subindices of the equivalence
symbol (≍p,q, for example).
The rest of this paper is organized as follows. In section 2 we introduce our notation and
state several known results that will be needed in the sequel. In section 3 we show the first
relationships between the stability of low-rank matrix recovery and the geometry of Banach
spaces, by proving a matrix version of the Kashin-Temlyakov theorem. Section 4 contains
a technical result, a matrix version of the main theorem from [15] that gives conditions
on the measurement map A that guarantee the stability of the Schatten p-minimization
scheme. A very similar theorem was recently obtained independently by Liu, Huang and
Chen [25], though our proof is different and we require a weaker hypothesis. In the final
4
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
section, the technical result from Section 4 is used to calculate the Gelfand numbers of the
identity maps from SN
for 0 < p ≤ 1, p < q ≤ 2 in the spirit of the work of Foucart,
Pajor, Rauhut and Ullrich.
p to SN
q
2. Notation and preliminaries
to rectangular ones. For p > 0 we will denote by SN
Schatten p-quasi-norm, given by
In this paper we will only consider square matrices, but all the results can be adapted
p the space of N × N matrices with the
σip(cid:17)1/p
kXkSp =(cid:16)
N
,
Xi=1
i=1 is the vector of singular values of the matrix X. Similarly, SN
where (σi)N
the space of N × N matrices with the weak-Schatten-p-quasi-norm given by
p,∞ will denote
k1/pσ∗
k
i=1 is the non-increasing rearrangement of (σi)N
i=1. For any quasi-normed space
kXkSp,∞
= max
1≤k≤N
where (σ∗
X, BX will denote its unit ball.
i )N
We will need to consider the best s-rank approximation error in the Schatten p-quasi-
norm,
It is well known that the infimum is actually attained at the s-spectral truncation Y = X[s]
(that is, keeping only the s largest singular values in the singular value decomposition).
ρs(X)Sp := inf(cid:8)kX − Y kSp : rank(Y ) ≤ s(cid:9).
Given a linear map A : MN → Rm and a vector y ∈ Rm, for 0 < p ≤ 1 we will denote by
∆p(y) a solution to
minimize kZkSp
subject to AZ = y.
That is, ∆p is the Schatten p-quasi-norm minimization reconstruction map. The map ∆p of
course depends on the measuring map A, but for simplicity we do not make this dependence
explicit in the notation.
2.1. The Restricted Isometry Property. The Restricted Isometry Property (RIP) for
a linear map A : RN → Rm was introduced by Cand`es and Tao [5], and quickly became
a key concept in the analysis of sparse recovery via ℓp-norm minimization. The s-order
restricted isometry constant of such a map is the smallest δ > 0 such that for every vector
x ∈ RN of sparsity at most s,
(1 − δ)kxk2
ℓ2 ≤ kAxk2
ℓ2 ≤ (1 + δ)kxk2
ℓ2 .
The importance of the RIP stems from the fact that small restricted isometry constants
imply exact recovery via ℓp-quasi-norm minimization for 0 < p ≤ 1, and it should be noted
that it is well known that random choices of the matrix A give small RIP constants of order
s, as long as m is at least of the order of s ln(eN/s) [6, 1, 26].
The version of the RIP for matrix recovery was introduced by Recht, Fazel and Parrilo
[32], and is as follows: a linear map A : MN → Rm is said to have the Restricted Isometry
Property of rank s with constant δ > 0 if for every matrix Z ∈ MN of rank at most s,
(1 − δ)kZk2
S2 ≤ kAZk2
ℓ2 ≤ (1 + δ)kZk2
S2 .
STABILITY OF LOW-RANK MATRIX RECOVERY
5
The best such constant is denoted by δs(A).
result follows from [3, Thm. 2.3], and will be very important for us in the sequel.
Just as in the vector case, random constructions give small RIP constants. The next
Theorem 2.1. Given a prescribed δ ∈ (0, 1), there is a constant Cδ such that if the entries
of the map A (seen a matrix with respect to the canonical bases in MN and Rm) are indepen-
dent gaussians with mean zero and variance 1/m, then with positive (even overwhelming)
probability δs(A) ≤ δ holds provided that
(2.1)
m ≥ CδsN.
3. A noncommutative Kashin-Temlyakov theorem
We will prove a matrix version of the Kashin-Temlyakov characterization of the stability
of sparse recovery via ℓ1-norm minimization in terms of widths. To this end, we define
three properties modeled after the ones studied in [20].
Definition 3.1. Let N 2 > m and A : MN → Rm a linear operator. We say that A has a:
(a) Matrix Strong Compressed Sensing Property (MSCSP) if for any X ∈ MN we have
kX − ∆1(AX)kS2 ≤ Cs−1/2ρs(X)S1
for s ≍ m/N .
(b) Matrix Weak Compressed Sensing Property (MWCSP) if for any X ∈ MN we have
kX − ∆1(AX)kS2 ≤ Cs−1/2 kXkS1
for s ≍ m/N .
(c) Matrix Width Property (MWP) if for any X ∈ ker(A),
kXkS2 ≤ C(N/m)−1/2 kXkS1 .
(3.1)
Notice that the MSCSP is a weakening of condition (i) in [29, Lemma 8], since we are only
considering X ′ = ∆1(AX). Also, the name of the MWP comes from its clear relationship
to the definition of the Gelfand numbers/widths. The following theorem is a matrix version
of the Kashin-Temlyakov theorem [20, Thm. 2.2]:
Theorem 3.2. For a linear operator A : MN → Rm, the MSCSP, MWCSP and MWP are
equivalent (up to a change in the constants).
Proof. The MSCSP trivially implies the MWCSP, since ρs(X)S1 ≤ kX − 0kS1 = kXk.
Assume that A has the MWSCSP. Given X ∈ ker(A), note that AX = 0 = A0, so clearly
0 = ∆1(0) = ∆1(AX) and thus from the MWSCSP we have kXkS2 ≤ Cs−1/2 kXkS1,
giving the MWP. Assume now that we have the MWP, that is, that equation (3.1) holds. If
s < 1
4 C −2N/m, from [29, Thm. 2] (which is a matrix version of [20, Thm. 2.1]) we obtain
kX − ∆1(AX)kS1 ≤ C ′ρs(X)S1
for C ′ = 2(1− 2pC 2sm/N )−1. Since X − ∆1(AX) ∈ ker A, the previous equation together
with (3.1) imply the MSCSP.
(cid:3)
6
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
The aforementioned Kashin-Temlyakov theorem says, in a nutshell, that the stability of
sparse-vector recovery via ℓ1-minimization has limits imposed by the geometry of Banach
spaces encoded in the appropriate Gelfand widths. In the previous proposition, we showed a
similar relationship relating the stability of low-rank recovery via nuclear norm minimization
with some other Gelfand widths. As in the vector case, following [17, Cor. 10.6], there is
a relationship between the geometry of SN
1 and the stability of compressed sensing by any
method. See Theorem 5.5 below for the precise statement.
4. Stability of low-rank matrix recovery through Schatten p quasi-norm
minimization
In this technical section we prove a general result (a matrix version of the main theorem
in [15]) that gives RIP-style conditions on the measuring map A that guarantee the stability
of the Schatten p-norm minimization scheme. For that we will need some notation: Let
αs, βs ≥ 0 be the best constants in the inequalities
αs kZkS2 ≤ kAZkℓ2 ≤ βs kZkS2 ,
rank(Z) ≤ s.
The results will be stated in terms of a quantity invariant under the change A ← cA,
namely
Note that this constant is related to the RIP constant, in fact
γ2s :=
β2
2s
2s ≥ 1.
α2
γ2s =
1 + δ2s
1 − δ2s
.
Unlike in the rest of the paper, we will consider the more general situation of approximate
recovery when measurements are moderately flawed, namely the problem
kAZ − ykℓ2 ≤ β2s · θ.
minimize kZkSp
subject to
(Pp,θ)
For simplicity, we will write (Pp) instead of (Pp,0). Note that by a compactness argument, a
solution of (Pp,θ) exists for any 0 < p ≤ 1 and any θ ≥ 0. The following theorem is a matrix
version of [15, Thm. 3.1]. It gives conditions (in the spirit of the RIP) that guarantee not
only the stability but also the robustness (that is, resistance to errors in the measurements)
of the Schatten p-quasi-norm-minimization for low-rank matrix recovery.
Theorem 4.1. Given 0 < p ≤ 1, if for some integer t ≥ s
s(cid:19)1/p−1/2
(4.1)
γ2t − 1 < 4(√2 − 1)(cid:18) t
then a solution X ∗ of (Pp,θ) approximates the original matrix X with errors
(4.2)
(4.3)
kX − X ∗kSp ≤ C1ρs(X)Sp + D1 · s1/p−1/2 · θ,
kX − X ∗kS2 ≤ C2
ρs(X)Sp
t1/p−1/2
+ D2 · θ,
where the constants C1, C2, D1 and D2 depend only on p, γ2t and the ratio s/t.
STABILITY OF LOW-RANK MATRIX RECOVERY
7
Proof. We will need to recall some properties of the Sp-quasi-norm. Namely, for any ma-
trices U and V ,
(4.4)
kUkS1 ≤ kUkSp ,
kUkSp ≤ N 1/p−1/2 kUkS2 ,
STEP 1: Consequence of the assumption on γ2t.
kU + V kp
Sp ≤ kUkp
Sp
+ kV kp
Sp
.
We will consider certain matrix decompositions similar to the ones in [21]. Consider the
singular value decomposition of X, given by
X = U diag(λi(X))V T
where U , V are unitary matrices and λ(X) = (λ1(X), . . . , λN (X)) are the singular values
of X arranged in decreasing order. For any matrix Z ∈ MN , we will consider a block
decomposition of Z with respect to X as follows: let U T ZV have the block form
U T ZV =(cid:18)Z11 Z12
Z21 Z22(cid:19)
where Z11, Z12, Z21, Z22 are of sizes s × s, s × (N − s), (N − s) × s, (N − s) × (N − s),
respectively. We then decompose Z as Z = Z(s) + Z c
(s) where
Z21
Z(s) = U(cid:18)Z11 Z12
0 (cid:19) V T
(s) = U(cid:18)0
0 Z22(cid:19) V T
Z c
0
Furthermore, we now consider the singular value decomposition of Z22 given by
Z22 = P diag(λ(Z22))QT
with P and Q being (N − s) × (N − s) unitary matrices, and λ(Z22) is the vector of the
N − s singular values of Z22 arranged in decreasing order. We decompose λ(Z22) as a sum
of vectors ZTi, each of sparsity at most t, where T1 corresponds to the locations of the t
largest entries of λ(Z22), T2 to the locations of the next t largest entries, and so on. For
i ≥ 1 we now define
ZTi = U(cid:18)0
0 P diag(λTi(Z22))QT(cid:19) V T ,
0
and denote ZT0 := Z(s).
We first observe that
kZT0k2
S2 + kZT1k2
S2 = kZT0 + ZT1k2
S2 ≤
1
2t kA(ZT0 + ZT1)k2
α2
ℓ2
1
α2
1
α2
=
(4.5)
2t(cid:10)A(Z − ZT2 − ZT3 − ··· ),A(ZT0 + ZT1)(cid:11)
2t(cid:10)AZ,A(ZT0 + ZT1)(cid:11) +
Let us renormalize the vectors −ZTk and ZT0 so that their S2-norms equal one by setting
and Y0 := ZT0/kZT0kS2. We then obtain, using the polarization
2t Xk≥2(cid:2)(cid:10)A(−ZTk ),AZT0(cid:11) +(cid:10)A(−ZTk ),AZT1(cid:11)(cid:3)
Yk := −ZTk /kZTkkS2
1
α2
=
8
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
identity
(cid:10)A(−ZTk ),AZT0(cid:11)
kZTkkS2 kZT0kS2
1
1
= hAYk,AY0i =
4(cid:2)β2
2t kYk + Y0k2
≤
4(cid:2)kA(Yk + Y0)k2
2t kYk − Y0k2
S2 − α2
ℓ2 − kA(Yk − Y0)k2
ℓ2(cid:3)
2t − α2
[β2
S2(cid:3) =
2t].
1
2
An analogous argument with T1 in place of T0 allows us to conclude
β2
2t − α2
2t
2
(4.6)
On the other hand, we have
(cid:10)A(−ZTk ),AZT0(cid:11) +(cid:10)A(−ZTk ),AZT1(cid:11) ≤
kZTkkS2(cid:2)kZT0kS2 + kZT1kS2(cid:3).
(4.7) (cid:10)AZ,A(ZT0 +ZT1)(cid:11) ≤ kAZkℓ2·kA(ZT0 + ZT1)kℓ2 ≤ kAZkℓ2·β2t(cid:2)kZT0kS2 +kZT1kS2(cid:3).
(cid:2)kZT0kS2 + kZT1kS2(cid:3).
Substituting the inequalities (4.6) and (4.7) into (4.5) we have
γ2t − 1
2 Xk≥2
S2 ≤
γ2t
β2t kAZkℓ2 +
kZTkkS2
S2 + kZT1k2
kZT0k2
inequality is
If we set c := kAZkℓ2 · γ2t/β2t, d := (γ2t − 1)/2 and Σ = Pk≥2 kZTkkS2
S2 − (c + dΣ)kZT1kS2 ≤ 0,
S2 − (c + dΣ)kZT0kS2 + kZT1k2
kZT0k2
, the previous
or equivalently,
c + dΣ
(cid:20)kZT0kS2 −
(cid:21)2
by getting rid of the second squared term, this easily implies
1 + √2
+(cid:20)kZT1kS2 −
(cid:21)2
c + dΣ
c + dΣ
2
2
(4.8)
(c + dΣ)2
2
.
≤
c + dΣ
√2
2
=
+
kZT0kS2 ≤
kZT0kSp ≤ s1/p−1/2 kZT0kS2 ≤ s1/p−1/2 1 + √2
2
(c + dΣ).
By Holder's inequality (see (4.4)) we get
2
p
Sp
(c + dΣ).
. Therefore,
(4.9)
We now proceed to bound Σ. For k ≥ 2, let η, η′ be singular values of ZTk , ZTk−1,
respectively. By definition, we must have η ≤ η′. Raising to the p-th power and averaging
, and hence η2 ≤ t−2/p(cid:13)(cid:13)ZTk−1(cid:13)(cid:13)
over all singular values of ZTk−1, ηp ≤ t−1(cid:13)(cid:13)ZTk−1(cid:13)(cid:13)
.
Adding over all singular values of ZTk and taking the square root, this yields kZTkkS2 ≤
t1/2−1/p(cid:13)(cid:13)ZTk−1(cid:13)(cid:13)Sp
kZTkkSp ≤ t1/2−1/p"Xk≥1
Σ =Xk≥2
2β2t · kAZkℓ2 · s1/p−1/2 + µ ·(cid:13)(cid:13)(cid:13)
kZTkkS2 ≤ t1/2−1/pXk≥1
Combining the above inequality with (4.9), we obtain
= t1/2−1/p(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)Z(s)(cid:13)(cid:13)Sp ≤
(s)(cid:13)(cid:13)(cid:13)Sp
(s)(cid:13)(cid:13)(cid:13)Sp
Sp#1/p
kZTkkp
(4.10)
2
Sp
Z c
Z c
λ
.
.
STABILITY OF LOW-RANK MATRIX RECOVERY
9
where the constants λ and µ are given by
λ := (1 + √2)γ2t and µ :=
1
4
(1 + √2)(γ2t − 1)(cid:16) s
t(cid:17)1/p−1/2
.
Note that the assumption on γ2t translates into the inequality µ < 1.
STEP 2: From now on let Z := X − X ∗.
Because X ∗ is a minimizer of (Pp,θ), we have
(4.11)
kX ∗kp
Sp ≤ kXkp
Sp
.
From [21, Lemma 2.2], whenever B, C ∈ MN satisfy BT C = 0 and BC T = 0 one has
kB + Ckp
(4.12)
. In particular, note that
X c
Z c
Sp
Sp
p
p
p
.
+(cid:13)(cid:13)(cid:13)
Sp
(s)(cid:13)(cid:13)(cid:13)
and (cid:13)(cid:13)(cid:13)
X(s) − Z c
Sp
(s)(cid:13)(cid:13)(cid:13)
p
Sp
=(cid:13)(cid:13)X(s)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)
Sp
(s)(cid:13)(cid:13)(cid:13)
From the p-triangle inequality (see (4.4)), since
= kBkp
kXkp
Sp
p
Sp
Sp
+ kCkp
=(cid:13)(cid:13)X(s)(cid:13)(cid:13)
X(s) − Z c
(s) = X − Z − X c
(s) + Z(s) = X ∗ − X c
(s) + Z(s),
Together with (4.11) and both equalities in (4.12), this yields
we get
p
p
Sp
Z c
p
Sp
(cid:13)(cid:13)(cid:13)
X(s) − Z c
Sp ≤ kX ∗kp
+(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)
Sp ≤(cid:13)(cid:13)X(s)(cid:13)(cid:13)
After a cancellation and noticing that (cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
STEP 3: Error estimates.
Sp ≤ 2ρs(X)p
(cid:13)(cid:13)X(s)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)
(4.13)
X c
p
Sp
Z c
Sp
Sp
p
p
We first note the bound
p
Sp
.
p
X c
Sp
(s)(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
Sp(cid:13)(cid:13)(cid:13)
p
+(cid:13)(cid:13)Z(s)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
Sp
p
X c
X c
= ρs(X)p
Sp
, we obtain
p
Sp
.
+(cid:13)(cid:13)Z(s)(cid:13)(cid:13)
p
Sp
.
+(cid:13)(cid:13)Z(s)(cid:13)(cid:13)
kAZkℓ2 = kAX − AX ∗kℓ2 ≤ kAX − ykℓ2 + ky − AX ∗kℓ2 ≤ 2β2s · θ.
For the Sp-error, we combine the estimates in (4.10) and (4.13) to obtain
Sp ≤ 2ρs(X)p
As a consequence of µ < 1, we have
Z c
Sp
(cid:13)(cid:13)(cid:13)
p
Sp
.
Z c
+ λp · s1−p/2 · θp + µp ·(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
1 − µp · s1−p/2 · θp.
λp
+
Sp
p
Sp ≤
2
1 − µp ρs(X)p
p
(s)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
Z c
10
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
Using the estimate (4.10) once again, we can derive that
p
p
2
Z c
p
Sp
+(cid:13)(cid:13)(cid:13)
(s)(cid:13)(cid:13)(cid:13)
kZkp
Sp ≤(cid:13)(cid:13)Z(s)(cid:13)(cid:13)
1 − µp (1 + µp)ρs(X)p
≤
≤ 21−1/p"
(1 − µp)1/p
21/p
Sp ≤ (1 + µq) ·(cid:13)(cid:13)(cid:13)
2λp
+
Sp
Z c
(s)(cid:13)(cid:13)(cid:13)
1 − µp · s1−p/2 · θp
(1 + µp)1/pρs(X)Sp +
+ λp · s1−p/2 · θp
Sp
21/pλ
(1 − µp)1/p · s1/p−1/2 · θ#p
where we have used the inequality (ap + bp)1/p ≤ 21/p−1(a + b) for a, b ≥ 0. The desired
estimate (4.2) follows with
C1 :=
22/p−1(1 + µp)1/p
(1 − µp)1/p
, D1 :=
22/p−1λ
(1 − µp)1/p
.
1/2
S2
kZTkk2
For the S2-error, let us observe that the bound in (4.8) also holds if we replace kZT0kS2 by
kZT1kS2, and hence
kZkS2 =
Xk≥0
where ν = (λ + 1 − √2)/2. We also have that
Σ ≤ t1/2−1/p(cid:13)(cid:13)(cid:13)
kZTkkS2 ≤ (1 + √2) · (c + dΣ) + Σ ≤ ν · Σ + 2λ · θ,
1 − µp · s1−p/2 · θp(cid:21)1/p
≤Xk≥0
(s)(cid:13)(cid:13)(cid:13)Sp ≤ t1/2−1/p(cid:20)
≤ t1/2−1/p21/p−1"
1 − µp ρs(X)p
(1 − µp)1/p
21/p
Z c
λp
+
Sp
2
and hence we conclude that
kZkS2 ≤ νt1/2−1/p21/p−1"
21/p
(1 − µp)1/p
This gives the estimate (4.3) with
ρs(X)Sp +
C2 =
22/p−2(λ + 1 − √2)
(1 − µp)1/p
,
D2 =
λ
ρs(X)Sp +
(1 − µp)1/p · s1/p−1/2 · θ# ,
(1 − µp)1/p · s1/p−1/2 · θ# + 2λθ.
21/p−2λ(λ + 1 − √2)
+ 2λ.
λ
(1 − µp)1/p
(cid:3)
As consequences of Theorem 4.1, we obtain two corollaries that are matrix versions of
the ones in [15]. The first one corresponds to the case of exact recovery.
Corollary 4.2. Given 0 < p ≤ 1, if
γ2t − 1 < 4(√2 − 1)(cid:18) t
s(cid:19)1/p−1/2
for some integer t ≥ s,
then every rank s matrix is exactly and stably recovered by solving (Pp).
The second one deals with the special case of nuclear norm minimization.
STABILITY OF LOW-RANK MATRIX RECOVERY
11
Corollary 4.3. Under the assumption that γ2s < 4√2 − 3 ≈ 2.6569, every rank s matrix
is exactly and stably recovered by solving (P1).
This last Corollary is clearly related to existing results on the RIP, it corresponds to
the condition δ2s < 2(3 − √2)/7 ≈ 0.4531. Note that for the specific case of p = 1 this
is in fact δ2s < 1/√2 ≈ 0.7071. Another recent result similar to our Theorem 4.1 is [25,
condition is not the best possible: a very recent result of Cai and Zhang [2] shows that the
optimal condition to have exact recovery of rank s matrices via nuclear norm minimization
Thm. 6] (which in turn generalizes results of Lee and Bresler [24, 23]), where they get a
conclusion of the same form as (4.3) but requiring a stronger hypothesis. Finally, note also
that Theorem 2.1 guarantees the existence of maps A satisfying the hypothesis of Theorem
4.1.
5. The Gelfand widths of Sp-balls for 0 < p ≤ 1
In this section we calculate the Gelfand numbers
p → SN
q )
cm(id : SN
for 0 < p ≤ 1, p < q ≤ 2. This can be considered as a noncommutative version of the
results from [16], where they use compressed sensing ideas to calculate the corresponding
Gelfand numbers
Inspired by their approach, our proof is based on low-rank matrix recovery ideas.
cm(id : ℓN
p → ℓN
q ).
Our main result is the following (compare to [16, Thm. 1.1]).
Theorem 5.1. For 0 < p ≤ 1 and p < q ≤ 2, if 1 ≤ m < N 2, then
m(cid:27)1/p−1/q
m(cid:27)1/p−1/q
q ) ≍p,q min(cid:26)1,
q ) ≍p,q min(cid:26)1,
and, if p < 1,
Sp,∞ , SN
dm(BN
dm(BN
Sp, SN
N
N
.
.
Before the proof, let us go through some preliminaries. Recall that it is classical to show
that, for q > p,
(5.1)
(5.2)
ρs(X)Sq ≤
ρs(X)Sq ≤
1
s1/p−1/q kXkSp ,
Dp,q
s1/p−1/q kXkSp,∞
,
Dp,q := (q/p − 1)−1/q.
5.1. Lower bounds. In this section we prove a result that will easily imply the desired
lower bounds in Theorem 5.1. It is a matrix version of [16, Thm. 2.1] and, just like in their
result, we note that the restriction q ≤ 2 is not imposed here.
Proposition 5.2. For 0 < p ≤ 1 and p < q ≤ ∞, there exists a constant cp,q > 0 such that
dm(BN
Sp, SN
q ) ≥ cp,q min(cid:26)1,
N
m(cid:27)1/p−1/q
12
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
Proof. With c = (1/2)2/p−1/q and µ = min{1, N/2m}, we are going to prove that
We proceed by contradiction, assuming that dm(BN
existence of a linear map A : MN → Rm such that for all V ∈ ker(A) \ {0},
q ) < cµ1/p−1/q. This implies the
dm(BN
Sp, SN
q ) ≥ cµ1/p−1/q.
Sp, SN
kV kSq < cµ1/p−1/q kV kSp .
For a fixed V ∈ ker(A) \ {0}, in view of the inequalities kV kSp ≤ N 1/p−1/q kV kSq and
c ≤ (1/2)1/p−1/q , we derive 1 < (µN/2)1/p−1/q , so 1 ≤ 1/µ < N/2. We then define
s := ⌊1/µ⌋ ≥ 1, so 2s < N and
1
2µ
< s ≤
1
µ
.
Now for V ∈ ker(A) \ {0},
(cid:13)(cid:13)V[2s](cid:13)(cid:13)Sp ≤ (2s)1/p−1/q(cid:13)(cid:13)V[2s](cid:13)(cid:13)Sq ≤ (2s)1/p−1/q kV kSq < c(2sµ)1/p−1/q kV kSp ≤
and therefore, using that kV kp
, we conclude
p
Sp
p
Sp
Sp
1
21/p kV kSp
p
=(cid:13)(cid:13)V[2s](cid:13)(cid:13)
(cid:13)(cid:13)V[2s](cid:13)(cid:13)
+(cid:13)(cid:13)V − V[2s](cid:13)(cid:13)
Sp ≤(cid:13)(cid:13)V − V[2s](cid:13)(cid:13)
p
Sp
.
This means that A satisfies the sufficient conditions in [29, Thm. 3], which implies that
Schatten p-quasinorm minimization gives exact recovery of rank s matrices. By well-known
arguments (see, for example, the discussion after the statement of theorem 2.3 in [3]), this
gives
m ≥ N s > N
1
2µ ≥
N
2
1
µ ≥
N
2
2m
N
= m,
a blatant contradiction.
(cid:3)
5.2. Upper bounds. In this subsection we establish a result from which the desired upper
bounds in Theorem 5.1 will follow easily. The proof relies on low-rank matrix recovery
methods, and the reader will notice similarities with the proof of Theorem 4.1. It should
be mentioned that the bound for the case p ≥ 1 follows easily from the result of Carl and
Defant [8] mentioned in the introduction together with an interpolation argument, but the
bound for the case p < 1 is new (as far as the authors know). Our result is a matrix version
of [16, Thm. 3.2], but the essence of the argument can be traced back to Donoho [10, Thm.
9]. As in the case of the result of [16], it is interesting to note that even when p < 1,
an optimal reconstruction map ∆ for the realization of the number Em(BN
q ) can be
chosen to be the S1-minimization mapping, at least when q ≥ 1.
Theorem 5.3. For 0 < p < 1 and p < q ≤ 2, there exists a linear map A : MN → Rm
such that, with r = min{1, q},
, SN
Sp,∞
sup
X∈BN
p,∞
kX − ∆r(AX)kSq ≤ Cp,q minn1,
where Cp,q > 0 is a constant that depends only on p and q.
N
mo1/p−1/q
,
STABILITY OF LOW-RANK MATRIX RECOVERY
13
Proof. Let C be the constant in (2.1) relative to the RIP associated with δ = 1/3, say.
Case 1: m ≥ CN .
We define s ≥ 1 as the largest integer smaller than m
m
.
(5.3)
CN
Let t = 2s. It is then possible to find a linear map A : MN → Rm with δt(A) ≤ δ. In
particular, we have δs(A) ≤ δ. Now, given Z := X − ∆r(AX) ∈ kerA, we decompose Z
into matrices ZT1, ZT2, ZT3, . . . of rank at most s by taking the s largest singular values of
Z for ZT1, then the next s largest ones for ZT2 and so on.
CN , so that
< s ≤
2CN
m
This easily implies (cid:0)kZTkk2
S2
(5.4)
kZTkkS2 ≤
Using the r-triangle inequality, we have
,
1
r
Sr
/s(cid:1)1/r, i.e.,
k ≥ 2.
/s(cid:1)1/2 ≤(cid:0)(cid:13)(cid:13)ZTk−1(cid:13)(cid:13)
s1/r−1/2 (cid:13)(cid:13)ZTk−1(cid:13)(cid:13)Sr
Sq ≤Xk≥1(cid:0)s1/q−1/2 kZTkkS2(cid:1)r ≤Xk≥1(cid:16) s1/q−1/2
√1 − δ kAZTkkℓ2(cid:17)r
.
r
Sq
kZTkkr
≤Xk≥1
Sq =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kZkr
Xk≥1
The fact that Z ∈ ker A implies AZT1 = −Pk≥2 AZTk . It follows that
Xk≥2
s1/q−1/2(cid:17)r
ZTk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Sq ≤ (cid:16) s1/q−1/2
√1 − δ (cid:17)r(cid:16)Xk≥2
≤ 2(cid:16) s1/q−1/2
√1 − δ (cid:17)r
Xk≥2
+(cid:16) s1/q−1/2
√1 − δ (cid:17)r
ℓ2 ≤ 2(cid:16)r1 + δ
1 − δ
kAZTkkℓ2(cid:17)r
kAZTkkr
kZkr
We then derive, using the inequality (5.4),
kZkr
Sq ≤ 2(cid:16)r 1 + δ
1 − δ
1
s1/r−1/q(cid:17)r
Xk≥1
kZTkkr
Sr
.
In view of the choice δ = 1/3 and of (5.3), we deduce
kAZTkkr
ℓ2
kZTkkr
S2
.
Xk≥2
(5.5)
kX − ∆r(AX)kSq ≤ 21/r√2(cid:16) 2CN
m (cid:17)1/r−1/q
kX − ∆r(AX)kSr .
Moreover, in view of δ2s ≤ 1/3 and of Theorems 4.1 and 2.1, there exists a constant C1 > 0
such that
(5.6)
kX − ∆r(AX)kSr ≤ (C1)1/rρs(x)Sr .
Finally, using (5.2) and (5.3), we have
(5.7)
m (cid:17)1/p−1/r
Putting (5.5), (5.6), and (5.7) together, we obtain, for any x ∈ BN
s1/p−1/r ≤ Dp,r(cid:16) 2CN
ρs(X)Sr ≤
Dp,r
.
p,∞,
kX − ∆r(AX)kSq ≤ 21/r√2C 1/r
1 Dp,r(cid:16) 2CN
m (cid:17)1/p−1/q
.
14
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
Case 2: m ≤ CN .
We simply choose the map A as the zero map. Then, for any X ∈ BN
kX − ∆r(AX)kSq = kXkSq ≤ Dp,q kXkp,∞ ≤ Dp,q,
Sp,∞
for some constant Dp,q > 0.
This completes the proof.
, we have
(cid:3)
Remark 5.4. When p = 1, the same proof but using inequality (5.1) instead of (5.2) gives
the following: for 1 < q ≤ 2, there exists a linear map A : MN → Rm such that,
sup
X∈BN
1
kX − ∆1(AX)kSq ≤ Cq minn1,
where Cq > 0 is a constant that depends only on q.
N
mo1−1/q
,
5.3. Proof of theorem 5.1.
Proof. First, an observation. As in the vector case, the simple inclusion BN
implies
Sp ⊆ BN
Sp,∞
dm(BN
Sp, SN
q ) ≤ dm(BN
Sp, SN
Sp,∞, SN
q ) and upper bounds for dm(BN
q ),
hence it suffices to show lower bounds for dm(BN
q ).
The lower bounds follow immediately from Proposition 5.2. When 0 < p < 1, the upper
bounds follow from Theorem 5.3. For p = 1, the upper bound when 1 ≤ m ≤ N follows
from the trivial inequality kXkSq ≤ kXkS1, whereas when N ≤ m ≤ N 2 it follows from
Remark 5.4.
, SN
Sp,∞
(cid:3)
5.4. Relation to compressive widths. As promised after the proof of our matrix version
of the Kashin-Temlyakov theorem, the relationship between the Banach space geometry
of the finite-dimensional Schatten p-classes and matrix recovery goes beyond the norm
minimization scheme. Below we use the notation from [17, Sec. 10.1]: the quantities Em and
Em
ada measure the worst-case reconstruction errors of optimal measurement/reconstruction
schemes in the nonadaptive and adaptive settings, respectively.
Theorem 5.5. For 0 < p ≤ 1 and p < q ≤ 2, if 1 ≤ m < N 2 then the adaptive and
nonadaptive compressive widths satisfy
Em
ada(BSN
p
, SN
Proof. Since −BSN
dm(BSN
p
p
p
= BSN
q ) ≤ Em
, SN
But now, since dm(BSN
the proof.
p
, SN
q ) ≍p,q min(cid:26)1,
N
m(cid:27)1/p−1/q
.
, [17, Thm. 10.4] implies
q ) ≤ 21/pdm(BSN
, SN
q )
, SN
p
q ) ≍p,q Em(BSN
+ BSN
and BSN
p
p
ada(BSN
p
, SN
p
p ⊆ 21/pBSN
q ) ≤ Em(BSN
p → SN
p
, SN
q ) = cm+1(id : SN
q ), an appeal to Theorem 5.1 finishes
(cid:3)
In the ℓp case the lower estimate is of particular importance in compressed sensing, since
it allows one to prove lower bounds for the number of measurements required to stably
recover s-sparse vectors in RN .
In the matrix case, that is no longer the case. Trying
it only gives that (under certain conditions), the minimum number of measurements m
STABILITY OF LOW-RANK MATRIX RECOVERY
15
required to stably recover rank s matrices in MN is ≥ CN s, which is not an improvement
over the information-theoretical limit. The reason behind this is that, unlike in the ℓp case,
there are compressed sensing algorithms (including norm minimization) that give stability
with a number of measurements of that order [3].
Acknowledgements
We would like to thank Rachel Ward for suggesting the reference [15], and also thank
the Workshop in Analysis in Probability at Texas A&M University. The first author was
partially supported by NSF grant DMS-1400588.
References
[1] Richard Baraniuk, Mark Davenport, Ronald DeVore, and Michael Wakin. A simple proof of the re-
stricted isometry property for random matrices. Constr. Approx., 28(3):253 -- 263, 2008.
[2] T. Tony Cai and Anru Zhang. Sparse representation of a polytope and recovery of sparse signals and
low-rank matrices. IEEE Trans. Inform. Theory, 60(1):122 -- 132, 2014.
[3] Emmanuel J. Cand`es and Yaniv Plan. Tight oracle inequalities for low-rank matrix recovery from a
minimal number of noisy random measurements. IEEE Trans. Inform. Theory, 57(4):2342 -- 2359, 2011.
[4] Emmanuel J. Cand`es and Benjamin Recht. Exact matrix completion via convex optimization. Found.
Comput. Math., 9(6):717 -- 772, 2009.
[5] Emmanuel J. Cand`es and Terence Tao. Decoding by linear programming. IEEE Trans. Inform. Theory,
51(12):4203 -- 4215, 2005.
[6] Emmanuel J. Candes and Terence Tao. Near-optimal signal recovery from random projections: universal
encoding strategies? IEEE Trans. Inform. Theory, 52(12):5406 -- 5425, 2006.
[7] Emmanuel J. Cand`es and Terence Tao. The power of convex relaxation: near-optimal matrix comple-
tion. IEEE Trans. Inform. Theory, 56(5):2053 -- 2080, 2010.
[8] Bernd Carl and Andreas Defant. Asymptotic estimates for approximation quantities of tensor product
identities. J. Approx. Theory, 88(2):228 -- 256, 1997.
[9] Albert Cohen, Wolfgang Dahmen, and Ronald DeVore. Compressed sensing and best k-term approxi-
mation. J. Amer. Math. Soc., 22(1):211 -- 231, 2009.
[10] David L. Donoho. Compressed sensing. IEEE Trans. Inform. Theory, 52(4):1289 -- 1306, 2006.
[11] K. Dvijotham and Maryam Fazel. A nullspace analysis of the nuclear norm heuristic for rank minimiza-
tion. In Proc. of ICASSP 2010, Dallas, TX, March 2010.
[12] David E. Edmunds and Jan Lang. Gelfand numbers and widths. J. Approx. Theory, 166:78 -- 84, 2013.
[13] Maryam Fazel. Matrix rank minimization with applications. PhD thesis, Stanford University, 2002.
[14] Massimo Fornasier, Holger Rauhut, and Rachel Ward. Low-rank matrix recovery via iteratively
reweighted least squares minimization. SIAM J. Optim., 21(4):1614 -- 1640, 2011.
[15] Simon Foucart and Ming-Jun Lai. Sparsest solutions of underdetermined linear systems via lq-
minimization for 0 < q ≤ 1. Appl. Comput. Harmon. Anal., 26(3):395 -- 407, 2009.
[16] Simon Foucart, Alain Pajor, Holger Rauhut, and Tino Ullrich. The Gelfand widths of ℓp-balls for
0 < p ≤ 1. J. Complexity, 26(6):629 -- 640, 2010.
[17] Simon Foucart and Holger Rauhut. A mathematical introduction to compressive sensing. Applied and
Numerical Harmonic Analysis. Birkhauser/Springer, New York, 2013.
[18] A. Yu. Garnaev and E. D. Gluskin. The widths of a Euclidean ball. Dokl. Akad. Nauk SSSR,
277(5):1048 -- 1052, 1984.
[19] B. S. Kashin. The widths of certain finite-dimensional sets and classes of smooth functions. Izv. Akad.
Nauk SSSR Ser. Mat., 41(2):334 -- 351, 478, 1977.
[20] B. S. Kashin and V. N. Temlyakov. A remark on the problem of compressed sensing. Mat. Zametki,
82(6):829 -- 837, 2007.
[21] Lingchen Kong and Naihua Xiu. Exact low-rank matrix recovery via nonconvex schatten p-
minimization. Asia-Pacific Journal of Operational Research, pages 1340010 -- 1 -- 1340010 -- 13, 2013.
16
JAVIER ALEJANDRO CH ´AVEZ-DOM´INGUEZ AND DENKA KUTZAROVA
[22] Hermann Konig. Eigenvalues of operators and applications. In Handbook of the geometry of Banach
spaces, Vol. I, pages 941 -- 974. North-Holland, Amsterdam, 2001.
[23] Kiryung Lee and Yoram Bresler. Guaranteed minimum rank approximation from linear observations
by nuclear norm minimization with an ellipsoidal constraint. ArXiv preprint, page arXiv:0903.4742.
[24] Kiryung Lee and Yoram Bresler. ADMiRA: atomic decomposition for minimum rank approximation.
IEEE Trans. Inform. Theory, 56(9):4402 -- 4416, 2010.
[25] Lu Liu, Wei Huang, and Di-Rong Chen. Exact minimum rank approximation via Schatten p-norm
minimization. J. Comput. Appl. Math., 267:218 -- 227, 2014.
[26] Shahar Mendelson, Alain Pajor, and Nicole Tomczak-Jaegermann. Uniform uncertainty principle for
Bernoulli and subgaussian ensembles. Constr. Approx., 28(3):277 -- 289, 2008.
[27] B. K. Natarajan. Sparse approximate solutions to linear systems. SIAM J. Comput., 24(2):227 -- 234,
1995.
[28] Erich Novak. Optimal recovery and n-widths for convex classes of functions. J. Approx. Theory,
80(3):390 -- 408, 1995.
[29] Samet Oymak, Karthik Mohan, Maryam Fazel, and Babak Hassibi. A simplified approach to recov-
ery conditions for low rank matrices. In 2011 IEEE International Symposium on Information Theory
Proceedings, pages 2318 -- 2322, Piscataway, NJ, 2011. IEEE.
[30] Albrecht Pietsch. Eigenvalues and s-numbers, volume 13 of Cambridge Studies in Advanced Mathemat-
ics. Cambridge University Press, Cambridge, 1987.
[31] Albrecht Pietsch. History of Banach spaces and linear operators. Birkhauser Boston, Inc., Boston, MA,
2007.
[32] Benjamin Recht, Maryam Fazel, and Pablo A. Parrilo. Guaranteed minimum-rank solutions of linear
matrix equations via nuclear norm minimization. SIAM Rev., 52(3):471 -- 501, 2010.
[33] Benjamin Recht, Weiyu Xu, and Babak Hassibi. Null space conditions and thresholds for rank mini-
mization. Math. Program., 127(1, Ser. B):175 -- 202, 2011.
Department of Mathematics, University of Texas at Austin, 2515 Speedway Stop C1200,
Austin, TX 78712-1202.
Current address:
Instituto de Ciencias Matem´aticas, CSIC-UAM-UC3M-UCM, C/Nicol´as Cabrera,
n◦ 13-15, Campus de Cantoblanco, UAM, 28049, Madrid, Spain.
E-mail address: [email protected]
Institute of Mathematics, Bulgarian Academy of Sciences, Sofia, Bulgaria.
Current address: Department of Mathematics, University of Illinois at Urbana-Champaign, 1409 W.
Green Street Urbana, IL 61801.
E-mail address: [email protected]
|
1107.0275 | 1 | 1107 | 2011-07-01T17:12:39 | Multiresolution analysis for Markov Interval Maps | [
"math.FA",
"math.OA"
] | We set up a multiresolution analysis on fractal sets derived from limit sets of Markov Interval Maps. For this we consider the $\mathbb{Z}$-convolution of a non-atomic measure supported on the limit set of such systems and give a thorough investigation of the space of square integrable functions with respect to this measure. We define an abstract multiresolution analysis, prove the existence of mother wavelets, and then apply these abstract results to Markov Interval Maps. Even though, in our setting the corresponding scaling operators are in general not unitary we are able to give a complete description of the multiresolution analysis in terms of multiwavelets. | math.FA | math |
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL
MAPS
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
Abstract. We set up a multiresolution analysis on fractal sets derived from
limit sets of Markov Interval Maps. For this we consider the Z-convolution of
a non-atomic measure supported on the limit set of such systems and give a
thorough investigation of the space of square integrable functions with respect
to this measure. We define an abstract multiresolution analysis, prove the
existence of mother wavelets, and then apply these abstract results to Markov
Interval Maps. Even though, in our setting the corresponding scaling operators
are in general not unitary we are able to give a complete description of the
multiresolution analysis in terms of multiwavelets.
Contents
Introduction and main results
1.
2. Markov Interval Maps
3. Abstract Multiresolution analysis
3.1. Proof of Theorem 1.6
3.2. Abstract multiplicative Multiresolution analysis
3.3. Translation completeness
4. Applications to Markov Interval Maps
4.1. Multiresolution Analysis for MIM
4.2. Mother wavelets for MIM
4.3. MRA for a Markov measures
4.4. Examples
5. Operator algebra
References
1
9
11
13
16
18
20
20
25
25
29
29
31
1. Introduction and main results
The main aim of this paper is to construct a wavelet basis on limit sets of Markov
Interval Maps (MIM) in the unit interval translated by Z. In this way we extend
the results in [DJ06, BK10], where wavelet bases with respect to singular measures
were provided. For this we first prove that a MIM gives rise to a multiresolution
analysis (MRA) and study the particular case where the underlying measure is
Markovian. This MRA is then reformulated in an abstract way allowing us to
prove the existence of a wavelet basis in this abstract setting.
In the case of a fractal given by an iterated function system (IFS) on [0, 1] there
are several approaches to construct a wavelet basis on the L2-space with respect
to a suitable singular measure which is supported on a so-called enlarged fractal.
The enlarged fractal is derived from the original fractal by first mapping scaled
copies of it to each gap interval and then by taking the union of translats by Z
Date: July 1, 2011.
Key words and phrases. multiwavelets, multiresolution analysis, Markov interval maps.
1
2
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
defining a dense set in R.
In [DJ06] the authors construct a wavelets basis for
fractals on self-similar Cantor sets, i.e. sets that are given by affine IFS with the
same scaling factor 1/N, N ≥ 2, for all p ≤ N branches. They consider the
L2-space with respect to µ, the δ-dimensional Hausdorff measure restricted to the
enlarged fractal, where δ denotes the dimension of the Cantor set. In this situation
the analysis depends on the two unitary operators U and T , where U denotes the
pf (N·) and T denotes the translation operator
scaling operator given by U f :=
given by T f := f (· − 1) for f ∈ L2(µ). Furthermore, a natural choice for a father
wavelet ϕ is the characteristic function on the original fractal. The authors show
that for a family of closed subspaces (Vj)j∈Z of L2(µ) the following six conditions
are satisfied.
√
• cl(cid:83)
• (cid:84)
• ··· ⊂ V−2 ⊂ V−1 ⊂ V0 ⊂ V1 ⊂ V2 ⊂ ··· ,
j∈Z Vj = L2(µ),
j∈Z Vj = {0},
• Vj+1 = U Vj, j ∈ Z,
• {T nϕ : n ∈ Z} is an orthonormal basis in V0,
• U−1T U = T N.
These observations allow the authors to construct a wavelets basis for L2(µ) explic-
itly in terms of certain filter functions.
In [BK10] we generalize this approach by allowing conformal IFS satisfying the
open set condition on [0, 1]. We choose the measure of maximal entropy supported
on the fractal and this measure is extended to a measure µ supported on the enlarged
fractal in R. Then similarly as in [DJ06] we construct the wavelet basis via MRA
in terms of the unitary scaling operator U and the unitary translation operator T .
Again via filter functions the mother wavelets ψi, i ∈ {1, . . . , N − 1} are defined
such that(cid:8)U nT kψi : n, k ∈ Z, i ∈ {1, . . . , N − 1}(cid:9) provides an orthonormal basis of
L2(µ).
(cid:83)
by X :=(cid:84)∞
n=0 F −nI, where I :=(cid:83)
Here, our aim is to extend the construction of wavelet bases with respect to
fractal measures to the construction of wavelet bases on the by Z translated limit
set of a Markov Interval Map (MIM). A Markov Interval Map consists of a family
(Bi)N−1
i=0 of closed subintervals in [0, 1] with disjoint interior, and a function F :
i∈N Bi → [0, 1], such that FBi is expanding and C 1, i ∈ N := {0, . . . , N − 1}
and such that F (Bi) ∩ Bj (cid:54)= ∅ implies Bj ⊂ F (Bi). Its (fractal) limit set is given
i∈N Bi. By considering its inverse branches τi :=
−1, i ∈ N, we obtain a Graph Directed Markov System (see [MU03]) with
(FBi)
incidence matrix A = (Aij)i,j∈N, where Aij = 1 if F (Bi) ⊃ Bj and 0 otherwise.
For the precise definition see Definition 2.1 and for an explicit example of an MIM
see Example 1.2 where we consider the β-transformation. The limit set X is -- up to
a countable set where it is finite-to-one -- homeomorphic to the topological Markov
chain ΣA := {ω = (ω0, ω1, . . . ) ∈ N
: Aωiωi+1 = 1∀i ≥ 0}. For the definition of
Given a Markov measure(cid:101)ν on the shift space ΣA with a probability vector (pi)i∈N
the canonical coding map π from ΣA to X see (2.1).
and stochastic matrix (πij)i,j∈N, we consider the probability measure ν :=(cid:101)ν ◦ π−1,
to which we also refer as ν a Markov measure. The Z-convolution (by translations)
of ν is given by
N
(cid:88)
k∈Z
νZ :=
ν(· − k).
Similar to the construction in [BK10] we introduce the scaling operator
(1.1)
U f (x) :=
· 1[ji](x − k) · f (τ−1
j
(x − k) + j + N k)
(cid:88)
(cid:88)
(cid:88)
(cid:114) pi
k∈Z
j∈N
i∈N
pjπji
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
3
Figure 1.1. The graph of the β-transform.
and the translation operator
T f (x) := f (x − 1)
(1.2)
for f ∈ L2(νZ) and x ∈ R, where [ji] ⊂ R, i, j ∈ N, denotes a cylinder set (see
Section 2). It is important to note that in contrast to the construction of the scaling
operator for IFS the operator U is in general not unitary. Nevertheless, we have
the following properties.
Proposition 1.1. Let (ϕi)i∈N denote a family of father wavelets given by ϕi :=
1[i], i ∈ N. The translation operator T and the scaling operator U satisfy
−1
(cid:112)ν([i])
(2) ϕi = U(cid:80)
the following properties.
√
πijT iϕj, i ∈ N,
(1) T U = U T N,
j∈N
(3) (cid:104)T kϕiT lϕj(cid:105) = δ(k,i),(l,j), k, l ∈ Z, i, j ∈ N,
(4) U U∗ = id,
(5) U∗U = id if and only if Aij = 1 for all i, j ∈ N .
For an explicit formula of U∗ see (4.2). As an example for this setting we consider
the β-transformation.
√
5
denote the golden mean. Then
Example 1.2 (β-Transformation). Let β := 1+
the β-transformation is given by F : [0, 1] → [0, 1], x (cid:55)→ βx mod 1 (see Figure 1.1
2
for the graph of F ). This map can be considered as a MIM as follows.
In this
β , x ∈ [0, 1], and
case we have X := [0, 1] and the inverse branches are τ0(x) := x
β , x ∈ [0, β − 1]. We may choose the two intervals B0 := [0, β − 1]
(cid:19)
τ1(x) := x+1
and B1 := [β − 1, 1] and the corresponding transition matrix is then given by
A :=
(cid:18) 1 1
.
1 0
From [Rén57, Par60] we know that there exists an invariant measure ν absolutely
continuous to the Lebesgue measure restricted to [0, 1] with density h given by
(cid:40) 5+3
√
√
10
5
5
5+
10
h(x) :=
√
for 0 ≤ x <
for
5−1
2
2 ≤ x < 1.
5−1
√
,
4
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
Figure 1.2. The graph of U(cid:0)id[0,1]
(cid:1).
The measure ν can be represented on ΣA by a stationary Markov measure with the
stochastic matrix
and probability vector p :=
is then given for x ∈ R by
. The scaling operator U acting on L2 (νZ)
(cid:19)
2 − β
0
(cid:18) β − 1
(cid:17)
1
5
5
k∈Z
Π :=
, β−1√
(cid:88)
(cid:16) β√
(cid:16)(cid:112)β1[0,β−2)(x − k) + 1[β−2,β−1)(x − k)
(cid:17) · f (β(x − k) + 2k) .
5/β(cid:1)1/2
+ β · 1[β−1,1)(x − k)
1[0,β−1) and ϕ1 =(cid:0)√
(cid:17)
(cid:16)(cid:112)β1[0,β−2)(x) + 1[β−2,β−1)(x) + β · 1[β−1,1)(x)
βx.
U f (x) =
U(cid:0)id[0,1]
(cid:1) x =
For the father wavelets we may choose ϕ0 =(cid:0)√
5β(cid:1)1/2
1[β−1,1).
We illustrate the action of U in Figure 1.2 where U is applied to the identity map
id[0,1] : x (cid:55)→ x, restricted to [0, 1], that is for x ∈ [0, 1] we have
consider a family of scaling operators (cid:0)U (n)(cid:1)
We further generalize our construction by considering non-atomic probability
measures ν on X which we do not assume to be Markovian.
In this case it is
natural to consider more than just one scaling operator U. More precisely, we
n∈Z which allow us to construct an
orthonormal wavelet basis. For this we define U (0) := id and for f ∈ L2(νZ) and
n ∈ N we let
(cid:88)
(cid:88)
(cid:88)
k∈Z
ω∈Σn
A
j∈N
·f
(cid:115)
(cid:16)
1[ωj](x − k)
νZ([j])
νZ([ωj])
τ−1
ω (x − k) +(cid:80)n−1
i=0 ωn−1−iN i + N nk
(cid:17)
U (n)f (x)
:=
(1.3)
5
(cid:33)
(cid:33)
(cid:33)
ωn−1−iN i − N nk
ωn−1−iN i − N nk
+ k
.
It is straight forward to verify that if the measure ν is Markovian, then we have
U (n) = U n for n ∈ N0 and U (−n) = (U∗)n, n ∈ N. More details are provided
n∈Z and T satisfy the following
relations.
Proposition 1.3. Let (ϕj)j∈N denote the family of father wavelets given by ϕi =
1[i], i ∈ N. The translation operator T and the family of scaling
(νZ([i]))
−1/2
A
i=0
i=0
·f
τω
:=
1[j]
(cid:115)
k∈Z
a
j∈N
ω∈Σn
(1.4)
(cid:32)
(cid:88)
(cid:88)
(cid:88)
(cid:88)
νZ([ωj])
νZ([j])
and
U (−n)f (x)
(cid:32)
x − n−1(cid:88)
(cid:32)
x − n−1(cid:88)
in Section 4.3. Furthermore, the operators (cid:0)U (n)(cid:1)
operators(cid:0)U (n)(cid:1)
(3) ϕi = U (1)T i(cid:80)
The properties of (cid:0)U (n)(cid:1)
(1) Let(cid:0)U (n)(cid:1)
(1) T U (n) = U (n)T N n, n ∈ N,
(2) U (−n)T ϕj = T N n
j∈N
(cid:113) νZ([ij])
νZ([i]) ϕj, i ∈ N,
U (−n)ϕj, n ∈ N, j ∈ N,
n∈Z satisfy the following.
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
(4) (cid:104)U (n)T kϕiU (n)T lϕj(cid:105) = δ(k,i),(l,j), n, k, l ∈ Z, i, j ∈ N,
(5) U (n)U (−n) = id, n ∈ N,
(6) if U (n)T kϕj (cid:54)= 0, then U (−n)U (n)T kϕj = T kϕj, n ∈ N, k ∈ Z, j ∈ N.
n∈Z and T lead us to the following abstract definition
of a multiresolution analysis which involves more than one father wavelet. In the
literature these functions are sometimes called multiwavelets (cf. [Alp93]).
Definition 1.4 (Abstract MRA). Let µ be a non-atomic measure on (R,B).
(cid:16)
µ,(cid:0)U (n)(cid:1)
n∈Z ,T(cid:17)
n∈Z and T be bounded, linear operators on L2(µ) such that T is
unitary and U (0) = id. We say
allows a two-sided mul-
tiresolution analysis (two-sided MRA) if there exists a family {Vj : j ∈ Z} of
closed subspaces of L2 (µ) and a family of functions (called father wavelets)
ϕj ∈ L2 (µ), j ∈ N, N ∈ N, with compact support, such that the following
conditions are satisfied.
(a) ··· ⊂ V−2 ⊂ V−1 ⊂ V0 ⊂ V1 ⊂ V2 ⊂ . . . ,
(b) cl(cid:83)
(c) (cid:84)
(d) (cid:0)U (n)(cid:8)T kϕj : k ∈ Z, j ∈ N(cid:9)(cid:1)\{0}, n ∈ Z, is an orthonormal basis of
(e) U (n)(cid:8)T kϕi : k ∈ N n, i ∈ N(cid:9) ⊂ spanU (n+1)(cid:8)T kϕi : k ∈ N n+1, i ∈ N(cid:9),
n ∈ N0 and U (−n) {ϕi : i ∈ N} ⊂ spanU (−n+1)(cid:8)T kϕi : i ∈ N , k ∈ N(cid:9),
j∈Z Vj = L2(µ),
j∈Z Vj = {0},
Vn,
n ∈ N,
(2) Let (cid:0)U (n)(cid:1)
(cid:16)
n∈N0
,T(cid:17)
µ,(cid:0)U (n)(cid:1)
(f) T U (n)V0 = U (n)T N nV0 and U (−n)T V0 = T N nU (−n)V0, n ∈ N.
and T be linear operators on L2(µ), T unitary, and let
U (0) = id. We say
allows a one-sided multiresolution
analysis ( one-sided MRA) if there exists a family (Vj : j ∈ N0) of closed
subspaces of L2 (µ) and a family of functions (called father wavelets) ϕj ∈
L2 (µ), j ∈ N, N ∈ N, with compact support, such that the following
conditions are satisfied.
(a) V0 ⊂ V1 ⊂ V2 ⊂ . . . ,
Vj = L2(µ),
(b) cl(cid:83)
n∈N0
j∈N0
introduced above.
Theorem 1.5. Let(cid:0)U (n)(cid:1)
(cid:16)
νZ,(cid:0)U (n)(cid:1)
(cid:17)
6
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
(c) (cid:0)U (n)(cid:8)T kϕj : k ∈ Z, j ∈ N(cid:9)(cid:1)\{0}, n ∈ N0, is an orthonormal basis
(d) U (n)(cid:8)T kϕi : k ∈ N n, i ∈ N(cid:9) ⊂ spanU (n+1)(cid:8)T kϕi : k ∈ N n+1, i ∈ N(cid:9),
of Vn,
n ∈ N,
(e) T U (n)V0 = U (n)T N nV0, n ∈ N.
Our next theorem shows that the abstract MRA holds in particular for MIM as
n∈N0
lows a one-sided MRA, where the father wavelets are set to be ϕi := (νZ([i]))
i ∈ N.
n∈N0
be given as in (1.3). Then
, T
−1/2
al-
1[i],
(cid:17)
basis.
For the abstract MRA we show that there always exists an orthonormal wavelet
bounded, linear operators on L2(µ) and T be a unitary operator on L2(µ).
Theorem 1.6. Let µ be a non-atomic measure on R, (cid:0)U (n)(cid:1)
(cid:16)
µ,(cid:0)U (n)(cid:1)
with dn ≥ qn, d−n ≥ q−n, and two families of mother wavelets(cid:0)ψn,l : l ∈ dn − qn
(cid:16)
n∈Z be a family of
If
allows a two-sided MRA with father wavelets ϕj, j ∈ N, then
there exist for every n ∈ N0 numbers dn ∈ N n+2, d−n ∈ N 2, qn ∈ N n+1, q−n ∈ N,
, n ∈ N0, such that the following set of functions defines an
n∈Z ,T(cid:17)
ψ−n,l : l ∈ d−n − q−n
orthonormal basis for L2(µ)
(cid:8)T kψn,l : n ∈ N0, l ∈ dn − qn, k ∈ Z(cid:9)
∪(cid:110)T N nkψ−n,l : n ∈ N, l ∈ d−n − q−n, k ∈ Z(cid:111)
(cid:1) and show
(cid:9), and respectively for n < 0
Remark 1.7. We give a precise construction for the family of mother wavelets ψn,l
in Section 3.1. More precisely, for each n ∈ Z we consider the linear subspaces
Wn := Vn+1 (cid:9) Vn, where the closed subspaces Vn of L2(µ) are given in Defi-
nition 1.4 (1d), and the finite family of functions (cid:0)ψn,l : l ∈ dn − qn
that that for n ≥ 0 (cid:8)T kψn,l : k ∈ Z, l ∈ dn − qn
(cid:110)T Nnkψn,l : k ∈ Z, l ∈ d−n − q−n
, defines an orthonormal basis of Wn.
(cid:1),
(cid:111)
.
Note that for IFS the mother wavelets are typically constructed in terms of so-
called filter functions. We will see in Section 5 that an analog construction is still
possible if the measure ν is Markovian.
sponding result for the one-sided MRA.
An immediate consequence of the proof of Theorem 1.6 is the following corre-
Corollary 1.8. Let µ be a non-atomic measure on R, (cid:0)U (n)(cid:1)
wavelets(cid:0)ψn,l : l ∈ dn − qn
µ,(cid:0)U (n)(cid:1)
(cid:1), n ∈ N0, such that the following set of functions defines
bounded, linear operators on L2(µ) and T a unitary operator on L2(µ). If
allows a one-sided MRA with the father wavelets ϕj, j ∈ N, then there exists for
every n ∈ N0 numbers dn ∈ N n+2, qn ∈ N n+1 with dn ≥ qn and a family of mother
(cid:8)T kψn,l : n ∈ N0, l ∈ dn − qn, k ∈ Z(cid:9) ∪(cid:8)T kϕi : k ∈ Z, n ∈ N(cid:9) .
an orthonormal basis for L2(µ)
a family of
n∈N0
(cid:16)
The construction for a MIM with an underlying Markov measure ν belongs to a
specific class. In this class the scaling operators U (n) can be represented multiplica-
tively. In our general framework we say that
is multiplicative if
there exists a linear, bounded operator U on L2 (µ) such that for n ∈ N0 U (n) = U n
µ,(cid:0)U (n)(cid:1)
n∈N0
,T(cid:17)
(cid:16)
,T(cid:17)
n∈N0
(cid:16)
µ,(cid:0)U (n)(cid:1)
,T(cid:17)
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
7
and U (−n) = (U∗)n. The results concerning the mother wavelets simplify in this
case as a consequence of the following lemma.
Lemma 1.9. Let us assume that
allows a two-sided MRA with
the closed subspaces Vn of L2(µ) from Definition 1.4 (1d) and set Wn := Vn+1(cid:9) Vn,
n ∈ Z.
n∈N0
• If U (n) = U n for n ∈ N then Wn = U nW0, n ∈ N.
• If U (−n) = (U∗)n for n ∈ N0 then W−n = (U∗)n−1 W−1.
Thus, we only have to find appropriate mother wavelets for W0 and W−1 and ob-
tain a wavelet basis then by applying repeatedly U. More precisely, this observation
allows us to derive the following corollary from the Theorem 1.6.
is multiplicative, then there exists an or-
(cid:41)
(cid:111)
,
Corollary 1.10. If
n∈N0
thonormal basis of L2(µ) of the form
(cid:40)
(cid:16)
µ,(cid:0)U (n)(cid:1)
,T(cid:17)
n−1(cid:88)
i=0
A, k =
∪(cid:110)
U nT kψl : n ∈ N0, ω ∈ Σn
ωiN i + N nm, m ∈ Z, l ∈ d1 − N
(U∗)n T kψ−,l : n ∈ N0, k ∈ Z, l ∈ d−1 − N
where the functions ψl, l ∈ d1 − N, and ψ−,l, l ∈ d−1 − N, are given explicitly in
Remark 3.6.
The above corollary applied to Example 1.2 leads to the following construction.
Example (Example 1.2 (continued)). The mother wavelet is
(cid:16)√
ψ =
5(2 − β)
(cid:17)1/2
1[0,(β−1)2) −(cid:16)√
5
(cid:17)1/2
1[(β−1)2,β−1)
and so a basis is given by
(cid:8)T kϕ1 : k ∈ 2Z + 1(cid:9) ∪(cid:8)U nT kψ : k ∈ Dn, n ∈ N(cid:9) ∪(cid:8)(U∗)n T kψ : k ∈ Z, n ∈ N(cid:9) ,
.
kj2j + 2nl : (kj)j∈n ∈ {0, 1}n , kj · kj−1 = 0, j ∈ n − 1, l ∈ Z
n−1(cid:88)
where
Dn :=
j=0
The proof that this indeed defines a orthonormal basis will be postponed to Section
4.4.
−1/2
(cid:16)
(cid:17)
n∈Z , T
νZ,(cid:0)U (n)(cid:1)
In the case of a MRA for a MIM with Markov measure ν we have in particular
that U (n) = U n and U (−n) = (U∗)nand we even obtain a stronger correspondence
between Markov measures for MIM and a two-sided MRA.
Theorem 1.11. We have that
the father wavelets ϕi := (νZ([i]))
Markovian.
allows a two-sided MRA with
1[i], i ∈ N, if and only if the measure ν is
appart from being multiplicative, that is we have ϕj ∈ span U(cid:8)T jϕi : i ∈ N(cid:9) for
In the case of ν being a Markov measure we even have a stronger property
each j ∈ N. We call a MRA with this property translation complete. We further
investigate multiplicative MRA which are translation complete in Section 3.2. In
this situation we derive a 0-1-valued transition matrix A given by Aij = 0 if and
only if UT iϕj = 0 and show that for MIM the matix coincides with the incidence
matrix. This observation is used to construct the mother wavelets in a simpler way
by considering for each father wavelet a unitary matrix to obtain coefficients for the
8
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
corresponding mother wavelets. We will use this approach to construct the mother
wavelets for MIM.
We would like to point out some interesting connections to C∗-algebras of Cuntz-
Krieger type, [KSS07]. We start by further considering the scaling operator U for
the MRA in the setting of a MIM with the incidence matrix A and Markov measure
ν. We can also write the operator U in a different way using the representation of
a Cuntz-Krieger-algebra. For this we consider the partial isometries Si given for
i ∈ N, f ∈ L2(ν), x ∈ supp(ν) by
Sif (x) = (ν([i]))
−1/2
1[i](x)f (τ−1
i
(x)).
It has been shown in [KSS07] that this gives a representation of the Cuntz-Krieger-
algebra OA by bounded operators acting on L2(ν), that is the Si, i ∈ N, are partial
isometries and satisfy
(cid:88)
(cid:88)
j∈N
i∈N
S∗
i Si =
1 =
AijSjS∗
j ,
SiS∗
i .
The scaling operator U acting on L2 (νZ) can then alternatively be written in terms
of the partial isometries as
(cid:88)
(cid:88)
(cid:88)
(cid:114) pi
k∈Z
j∈N
i∈N
πji
U =
T kSj 1[i]T −(j+N k),
where we notice that Sj 1[i], j, i ∈ N, acts on L2 (νZ). We can also write U∗ in
terms of the partial isometries Si, i ∈ N. In this way we obtain
j T −k.
T j+N k1[i]S∗
(cid:114) πji
(cid:88)
(cid:88)
(cid:88)
U∗ =
k∈Z
j∈N
i∈N
pi
The spaces Vn, n ∈ N0, can also be written in terms of the isometries Si, i ∈ N,
that is for n ∈ N a basis of Vn is given by
(cid:26)(cid:114) pi
πωn−1i
T lSωϕi : l ∈ Z, ω ∈ Σn
A, i ∈ N
.
(cid:27)
Let us finish this section by commenting on some known results in the literature
connected to the results in here. Up to our knowledge there are at least two
further approaches to construct a wavelet basis on the limit sets of MIM, namely
[MP09, KS10] and there is one approach for the specific case of a β-transformation
given in [GP96]. In [MP09] Marcolli and Paolucci consider the limit set X of a MIM
N for i ∈ N
inside the unit interval consisting of the inverse branches τi(x) = x+i
with some transition rule encoded in a matrix A. This limit set can be associated
with a Cantor set inside the unit interval. The Cantor set is then equipped with the
Hausdorff measure of the appropriate dimension δ. If all transitions were allowed,
the limit set would coincide with a usual Cantor set given by an affine iterated
function system. They then use the representation of the Cuntz-Krieger-algebra
OA, where A is the transition matrix, for the construction of the orthonormal
(cid:1) and not a multiresolution analysis. Their proofs
system of wavelets on L2(cid:0)H δX
mainly rely on results in [Bod07, Jon98]. Finally, Marcolli and Paolucci give a
possible application where they adapt the construction of a wavelet basis to graph
wavelets for finite graphs with no sinks, which can be associated to Cuntz-Krieger-
algebras. These graph wavelets are a useful tool for spatial network traffic analysis,
compare [MP09, CK03].
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
9
In [KS10] the authors construct a Haar basis analogous to the wavelet basis
construction in [DJ06] for the middle third Cantor set for a one-sided topologically
exact subshift of finite type and with respect to a Gibbs measure µφ for a Hölder
continuous potential φ. The construction is then used to obtain a spectral triple in
the framework of non-commutative geometry.
The construction of wavelet basis in different spaces than L2(λ), where λ is the
Lebesgue measure on R, may lead to a further understanding of non-commutative
geometry in the sense that we can obtain a "Fourier" or wavelet basis for quasi
lattices or quasi crystals.
As an essential non-linear example for the construction of a wavelet basis on
limit sets of MIM one can take the limit set of a Kleinian group together with the
measure of maximal entropy or the Patterson-Sullivan measure, compare Example
2.3.
5
√
As an example we apply the construction to a β-transformation, where β denotes
, compare Example 1.2. In this way we obtain a
the golden mean, i.e. β = 1+
2
wavelet basis for L2 (νZ) , where ν is the invariant measure for this transformation,
compare [Rén57, Par60]. This measure is absolutely continuous with respect to the
Lebesgue measure. In [GP96], Gazeau and Patera construct a similar basis to ours
for the β-transformation with respect to the Lebesgue measure on R. They use
instead of a translation by the group Z a translation by so called β-integers which
consider the β-adic expansion and are obtained by a so-called greedy algorithm.
There are some common features between our construction and the one in [GP96]
like both give characteristic functions on intervals depending on powers of β. But
since we consider different measures we have different coefficients.
n∈Z and give a proof of Theorem 1.6.
operators (cid:0)U (n)(cid:1)
In Section 4 we start with a family of operators (cid:0)U (n)(cid:1)
The paper is organized as follows. In Section 2 we provide some basic definitions
and introduce MIM. In Section 3 we elaborate the abstract MRA for families of
In Section 3.2 we then
consider the special case of multiplicative systems. In Section 3.3 we prove how
the condition of translation completeness simplifies the construction of the mother
wavelets. The rest of this paper is devoted to the special case of a MRA for MIM.
n∈Z acting on L2 (νZ) for
an arbitrary non-atomic probability measure ν on the limit set of an MIM in the
unit interval translated by Z and show that a one-sided MRA is always satisfied.
If on the other hand a two-sided MRA holds, we then prove that the measure ν
is necessarily Markovian. The construction of the mother wavelets will be given
explicitly. In Section 4.3 we give an explicit construct of the wavelet basis if the
measure ν is Markovian.
Finally, in Section 5 we show how low-pass filters and high-pass filters can be em-
ployed to construct mother wavelets for multiwavelets for MIM with an underlying
Markov measure.
2. Markov Interval Maps
In this section we give some basic definitions and notations. We consider fractals
given as limit sets of one-dimensional Markov Interval Maps.
fine I :=(cid:83)
a Markov Interval Map and its limit set is defined as X :=(cid:84)∞
(cid:17)
Definition 2.1. Let (Bi)i∈N be closed intervals in [0, 1] with disjoint interior. De-
i∈N Bi and F : I → [0, 1] exanding and C 1 on each Bi, i ∈ N, such that
(Bi)i∈N , F
if F (Bi)∩ Bj (cid:54)= ∅ then Bj ⊂ F (Bi) for i, j ∈ N. We call the system
(cid:16)
n=0 F −nI.
Remark 2.2.
10
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
(a) A fundamental do-
main of
the action of
(cid:104)g, h(cid:105) on the Poincaré disc
model.
(b) The correspond-
ing
Bowen-Series
map.
Figure 2.1. Example of a Fuchsian group.
function system (IFS).
(1) If F (Bi) = [0, 1] for each i ∈ N, then (X, F ) corresponds to an iterated
−1, i ∈ N. The family (τi)i∈N
(2) We define the inverse branches τi := (FBi)
is called a one-dimensional graph directed Markov system (GDMS) with
the incidence matrix A = (Aij)N−1
(cid:40)
and it follows that F (Bi) =(cid:83)
Aij :=
1,
0,
i,j=0 which is obtained by
if Bj ⊂ F (Bi)
else,
j∈N : Aij =1 Bj.
Example 2.3. An example is a convex, co-compact Kleinian group, as an example
consider Figure 2.1a. The limit set can be considered as the limit set of the Bowen-
Series map, which gives rise to a Markov Interval Map, compare Figure 2.1b. The
limit set is the set that is obtained by successive application of these four maps,
where the composition of gi and g−1
i are forbidden. A typical measure to be studied
would be the measure of maximal entropy or the conformal measure (of maximal
dimension).
Next we consider the corresponding shift space. Consider the alphabet N =
{0, . . . , N − 1}. The limit set X is then homeomorphic (mod ν) to the set of all
admissible words
ΣA := {ω = (ω0, ω1, . . . ) ∈ N
N
The homeomorphism is given, for x ∈ X, by
: Aωiωi+1 = 1∀i ≥ 0}.
π : ΣA → X
ω (cid:55)→ limn→∞ τω0 ◦ ··· ◦ τωn (x),
(2.1)
which is independent of the particular choice of x ∈ X.
Remark 2.4. Furthermore, we define the cylinder sets for ω0, . . . , ωk ∈ N, k ∈ N0,
as
[ω0 . . . ωk] := {(ω(cid:48)
0, ω(cid:48)
1, . . . ) ∈ ΣA : ωi = ω(cid:48)
i, i ∈ {0, . . . , k}} .
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
11
If for some i ∈ {0, . . . , k − 1} Aωiωi+1 = 0 then [ω0 . . . ωk] = ∅.
Then the sets Bi and F (Bi) for i ∈ N are homeomorphic (mod ν) to the following
sets in the shift space:
π−1 (Bi) = [i]
π−1 (F (Bi)) = {ω = (ω0, ω1, . . . ) ∈ ΣA : Aiω0 = 1}.
and
The dynamic of F is conjugated to the shift dynamic σ : ΣA → ΣA, σ (ω0, ω1, . . . ) =
(ω1, ω2, . . . ) and consequently, the functions τi correspond to the inverse branches
of the shift function, i.e. τi ◦ π (ω0, ω1, . . . ) = π (i, ω0, ω1, . . . ), for ω ∈ π−1 (F (Bi)),
i ∈ N.
A := (cid:8)ω = (ω0, . . . , ωn−1) ∈ N n : Aωiωi+1 = 1 for all i ∈ {0, . . . , n − 1}(cid:9)
Furthermore, let us fix the following notation.
• Σn
• Σ∗
• For ω ∈ Σn
• For ω ∈ Σn
defines the set of admissible words of length n ∈ N.
A stands for all finite words, i.e. Σ∗
A.
n≥1 Σn
A we define τω := τω0 ◦ τω1 ◦ ··· ◦ τωn−1.
A, τ ∈ Σm
A we define their concatenation
ωτ := (ω0, . . . , ωn−1, τ0, . . . , τm−1)
A =(cid:83)
which is an element of Σn+m
whenever Aωn−1τ0 = 1.
A
As a measure on X we could consider for instance the pull-back under π of Gibbs
measures on ΣA (for definitions see e.g. [KS10]).
Now we define the appropriate space for which we want to construct a wavelet
Definition 2.5. Let(cid:101)ν be a probability measure on (ΣA,B) and ν =(cid:101)ν◦π−1. Define
basis.
the enlarged fractal by
(cid:91)
(cid:88)
k∈Z
k∈Z
R =
X + k
νZ(B) =
ν(B − k),
and define the Z-convolution νZ of the measure ν for a Borel set B by
which clearly is an invariant measure under Z-translation.
Remark 2.6. One example is the space L2 (ΣA, µφ), where ΣA denotes a one-sided
topologically exact subshift of finite type. An important class of measures on ΣA
are given by invariant Gibbs measure with respect to a Hölder continuous potential
φ ∈ C (ΣA, R), denoted by µφ, compare [KS10]. µφ corresponds to the measure (cid:101)ν
in Definition 2.5.
In the following we use the convention 0−1 · 1∅ = 0. For simplicity we let
[ω0, . . . , ωn−1] also denote the sets τω0 ◦ ··· ◦ τωn−1 (X) using the identification by
to a measure(cid:101)ν on ΣA by ν =(cid:101)ν ◦ π−1 and νZ denotes the measure obtained from ν
π. Furthermore, in Section 4 the measure ν supported on [0, 1] always corresponds
by Z-convolution.
3. Abstract Multiresolution analysis
In this section we give a proof of Theorem 1.6. To do so we first construct
mother wavelets explicitly and in the next step we prove that these give indeed
which allows a
an orthonormal basis.
two-sided MRA.
In this section we fix
µ,(cid:0)U (n)(cid:1)
n∈Z ,T(cid:17)
(cid:16)
For the construction of an ONB we cannot define the mother wavelets in terms
of filter functions due to the fact that we have more than one father wavelet.
12
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
Remark 3.1. If we have the usual setting from the literature, compare e.g. [Dau92],
then we have a multiplicative MRA with a unitary operator U and the operator T
given in (1.2). In this case there is only one father wavelet ϕ and there exists a
k∈Z akzk, ak ∈ C, such
that ϕ = Um0(T )ϕ. For the construction of the mother wavelets we look for N − 1
k ∈ C, j ∈ N\{0},
where N ∈ N is connected to the scaling since it indicates on which interval [0, N ]
the unit interval is mapped when the operator U is applied to 1[0,1]. The high-pass
filters are chosen, such that the matrix
1√
N
so-called low-pass filter m0 : T → T of the form m0(z) =(cid:80)
high-pass filters mj : T → T of the form mj : z (cid:55)→(cid:80)
(cid:0)mj(ρlz)(cid:1)
kzk, bj
M (z) :=
k∈Z bj
j,l∈N ,
where ρ = e2πi/N, is unitary for almost all z ∈ T. In terms of these high-pass filters
the mother wavelets are defined as ψj = Umj(T )ϕ, j ∈ N\{0}.
We first notice that for n ∈ N
(cid:26)(cid:18)(cid:106) k
(cid:107)
(cid:19)
(cid:27)
{(l, j) ∈ N n × N} =
(3.1)
where (m)N := m mod N and (cid:98)x(cid:99) = maxk∈Z,k≤x(k) is the largest integer not
exceeding x.
, (k)N
N
,
Clearly from the definition of the MRA, Definition 1.4 (1e), we have the following:
(1) If for n ∈ N0, k ∈ N n+1, U (n)T (cid:98) k
N (cid:99)ϕ(k)N (cid:54)= 0 , there exists uniquely
: k ∈ N n+1
m∈N n+2 ∈ CN n+2 such that
m
and
U (n)T (cid:98) k
(cid:1)
N (cid:99)ϕ(k)N = U (n+1)(cid:80)
determined(cid:0)an,k
(cid:0)an,k
U (−n)ϕi = U (−n+1)(cid:80)
(cid:0)bn,i
m∈N 2 ∈ CN 2 such that
m = 0, m ∈ N n+2,
m = 0, m ∈ N 2,
(cid:0)bn,i
and
(cid:1)
m
(3.2)
(3.3)
m∈N n+2 an,k
if U (n+1)T (cid:98) m
m T (cid:98) m
N (cid:99)ϕ(m)N
N (cid:99)ϕ(m)N = 0(cid:1) .
m∈N 2 bn,i
if U (−n+1)T (cid:98) m
m T (cid:98) m
N (cid:99)ϕ(m)N
N (cid:99)ϕ(m)N = 0(cid:1) .
(2) If U (−n)ϕi (cid:54)= 0, n ∈ N, i ∈ N, there exists uniquely determined coefficients
m
m
(cid:1)
m∈N 2, i ∈ N, n ∈ N.
(1) For fixed n ∈ N0, define
Remark 3.2. We only consider U−(n)ϕi, since U (−n)T kϕi = T N nkU (−n)ϕi by (1f)
of Definition 1.4.
Lemma 3.3. The following holds for the coefficients (cid:0)an,k
(cid:1)
n ∈ N0, and(cid:0)bn,i
m∈N n+2, k ∈ N n+1,
(cid:111)
N (cid:99)ϕ(m)N (cid:54)= 0
(cid:111)
m∈N n+2, k ∈ Qn, are orthonormal.
(2) For fixed n ∈ N, define
m ∈ N n+1 : U (n)T (cid:98) m
Then the vectors vk =(cid:0)an,k
(cid:110)
(cid:1)
Then the vectors wi =(cid:0)bn,i
m∈N 2, i ∈ Q−n, are orthonormal.
m ∈ N : U (−n)ϕm (cid:54)= 0
Q−n :=
Qn :=
(3.5)
(3.4)
(cid:110)
(cid:1)
m
m
.
.
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
13
Proof. ad (1): For fixed n ∈ N0, let k, l ∈ Qn, then
N (cid:99)ϕ(k)NU (n)T (cid:98) l
m T (cid:98) m
an,k
N (cid:99)ϕ(l)N(cid:105)
N (cid:99)ϕ(m)N
(cid:12)(cid:12)U (n+1) (cid:88)
m∈N n+2
δ(k,l) = (cid:104)U (n)T (cid:98) k
=(cid:10)U (n+1) (cid:88)
(cid:88)
m∈N n+2
m an,l
an,k
m .
=
m∈N n+2
ad (2): Follows analogously to (1).
(cid:11)
m T (cid:98) m
an,l
N (cid:99)ϕ(m)N
(cid:3)
3.1. Proof of Theorem 1.6. The aim is to prove the existence of a basis as
given in Theorem 1.6. For this we divide the proof in two parts. First we construct
coefficients such that the functions ψn,k given in (3.6) and (3.7) give an orthonormal
basis. In the second part we verify that these functions give indeed an orthonormal
basis. We prove these parts first for n ∈ N0 and then for n ∈ Z, n < 0. We define
the mother wavelets for each scale n ∈ Z such that we obtain with their translates
a basis for Wn = Vn+1 (cid:9) Vn, where Vn is given in Definition 1.4 (1d). Define for
n ∈ N0
m (cid:54)= 0 for some k ∈ Qn
m (cid:54)= 0 for some k ∈ Q−n
(cid:9) ,
(cid:9) ,
(3.6)
(3.7)
and dn := card Dn, d−n := card D−n.
• The mother wavelets for the subspaces Wn, n ∈ N0, of the MRA shall have
the form for k ∈ dn − qn with qn := card Qn
m T (cid:98) m
cn,k
N (cid:99)ϕ(m)N ,
m∈N n+2
m ∈ C are given in (3.8).
where the coefficients cn,k
• For the negative index subspaces W−n, n ∈ N, of L2(µ) we define the
mother wavelets in terms of the coefficients of the matrix in (3.9) for n ∈ N
and k ∈ d−n − q−n, where q−n := card Q−n, as
m T (cid:98) m
c−n,k
ψ−n,k := U (−n+1) (cid:88)
N (cid:99)ϕ(m)N .
m∈N 2
(cid:1)
matrix.
m ∈ C, c−n,k
The coefficients cn,k
Schmidt process.
m ∈ C are determined in the following via the Gram-
For the definition of the basis we fix n ∈ N0 and we construct an orthonormal
. This is a (qn × dn)-
Now we consider dn − qn vectors ei, i ∈ dn − qn, of length dn which are linearly
, k ∈ Qn. Via the Gram-Schmidt process
(cid:1)
, i ∈ dn − qn, of length dn which
to
m∈Dn
m = 0 if m ∈ N n+2\Dn and we define a matrix Cn :=
k∈Qn,m∈N n+2 of
basis for Cdn in the following way. Consider(cid:0)an,k
independent of the vectors (cid:0)an,k
(cid:1)
we obtain dn − qn orthonormal vectors(cid:0)cn,i
are orthonormal to (cid:0)an,k
(cid:0)cn,k
(cid:1)
, k ∈ Qn. We extend the vectors (cid:0)cn,i
k∈dn−qn,m∈N n+2 of size (dn − qn) × N n+2 and An := (cid:0)an,k
some of length N n+2 by cn,i
k∈Qn,m∈Dn
m
m∈Dn
m
m∈Dn
m
m∈Dn
(cid:1)
(cid:1)
(cid:1)
m
size qn × N n+2. So we obtain a matrix of size dn × N n+2 by
m
m
m
Dn :=(cid:8)m ∈ N n+2 : an,k
D−n :=(cid:8)m ∈ N 2 : bn,k
ψn,k := U (n+1) (cid:88)
(3.8)
Mn :=
Now we turn to the construction of the coefficients for ψ−n,k, n ∈ N, in (3.7).
For each −n, n ∈ N, we define an orthonormal basis of Cd−n in the following way.
(cid:18) AnCn
(cid:19)
.
m
14
(cid:1)
k∈Q−n,m∈D−n
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
Consider(cid:0)bn,k
(cid:1)
q−n vectors which are linearly independent of(cid:0)bn,k
process we obtain d−n − q−n orthonormal vectors(cid:0)c−n,j
length d−n which are orthonormal to(cid:0)bn,i
(cid:1)
vectors(cid:0)c−n,i
Now we define Dn := (cid:0)c−n,i
. This is a (q−n × d−n)-matrix. Now we consider d−n−
. Via the Gram-Schmidt
m∈D−n
, j ∈ d−n − q−n, of
m∈D−n
m
. In the last step we extend the
m = 0 if m ∈ N 2\D−n.
k∈Q−n,m∈N 2 such
to some of length N 2 by defining c−n,i
i∈d−n−q−n,m∈N 2 and Bn := (cid:0)bn,k
m∈D−n
m∈D−n
(cid:1)
(cid:1)
(cid:1)
(cid:1)
m
m
m
m
m
that
(3.9)
(cid:102)Mn :=
(cid:19)
(cid:18) BnDn
is a matrix of size d−n × N 2.
In the next step we show that we obtain indeed an orthonormal basis with these
mother wavelets given in (3.6) and (3.7). First we prove this for n ∈ N0. Recall
n∈N0
k=0 Wk ⊕ V0. Now we show
(cid:1) = L2(µ) since
that Wn = Vn+1 (cid:9) Vn for n ∈ N0. Consequently, cl(cid:0)(cid:83)
for every n ∈ N0 it follows iteratively that Vn+1 =(cid:76)n
that for fixed n ∈ N we have that(cid:8)T lψn,k : k ∈ dn − qn, l ∈ Z(cid:9) is an orthonormal
Wn ∪ V0
basis of Wn. First we show the orthonormality.
To show the orthonormality of T rψn,k and T sψn,l, r, s ∈ Z, k, l ∈ dn − qn,
it is sufficient to consider T rψn,k and ψn,l since the operator T is unitary. The
orthonormality follows then from
(cid:69)
m U (n+1)T (cid:98) m
cn,l
N (cid:99)ϕ(m)N
(cid:69)
N (cid:99)+N n+1rϕ(m)N
N (cid:99)+N n+1rϕ(m)N
(cid:12)(cid:12)(cid:12) (cid:88)
(cid:68)T N n+1r+(cid:98) m
m∈N n+2
N (cid:99)ϕ(m)N
(cid:12)(cid:12)(cid:12) (cid:88)
(cid:12)(cid:12)(cid:12)T (cid:98) s
m∈N n+2
m T (cid:98) m
cn,l
N (cid:99)ϕ(m)N
(cid:69)
N (cid:99)ϕ(s)N
· δ(N n+1r+(cid:98) m
N (cid:99),(cid:98) s
N (cid:99)),((m)N ,(s)N )
(cid:104)T rψn,kψn,l(cid:105) =
=
m∈N n+2
(cid:68) (cid:88)
m U (n+1)T (cid:98) m
(cid:68) (cid:88)
cn,k
m T (cid:98) m
(cid:88)
(cid:88)
cn,k
(cid:88)
(cid:88)
= δr,0 · (cid:88)
cn,k
m cn,l
s
cn,k
m cn,l
s
m∈N n+2
m∈N n+2
m∈N n+2
s∈N n+2
s∈N n+2
cn,k
m cn,l
m
=
=
m∈N n+2
= δr,0 · δk,l.
In the next step we show that Vn+1 = Vn ⊕ Wn. We consider a basis element
N (cid:99)ϕ(k)N , k ∈ N n+2, and show that it is a linear
N (cid:99)ϕ(l)N and ψn,m, l ∈ N n+1, m ∈ dn − qn. It is
N (cid:99)ϕ(k)N = 0
N (cid:99)ϕ(k)N (cid:54)= 0, k ∈ N n+2, it can be written as the
of Vn+1 of the form U (n+1)T (cid:98) k
combination of functions U (n)T (cid:98) l
sufficient to consider only k ∈ N n+2 by Definition 1.4 (1e). If U (n+1)T (cid:98) k
it is obvious satisfied. If U (n+1)T (cid:98) k
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
15
=U (n+1)
following linear combination:
U (n+1)T (cid:98) k
N (cid:99)ϕ(k)N
(cid:88)
(cid:88)
m∈N n+2
an,l
k
l∈Qn
(cid:88)
(cid:124)
(cid:88)
l∈Qn
=U (n+1)
(cid:88)
(cid:88)
l∈Qn
=
=
k U (n)T (cid:98) l
an,l
m∈N n+2
N (cid:99)ϕ(l)N +
k U (n)T (cid:98) l
an,l
N (cid:99)ϕ(l)N +
an,l
k an,l
m
m +
l∈dn−qn
cn,l
k cn,l
(cid:88)
(cid:125)
(cid:88)
k U (n+1) (cid:88)
(cid:123)(cid:122)
m T (cid:98) m
N (cid:99)ϕ(m)N +
(cid:88)
an,l
(cid:88)
l∈dn−qn
cn,l
=δk,m
l∈dn−qn
m∈N n+2
cn,l
k ψn,l.
T (cid:98) m
N (cid:99)ϕ(m)N
(cid:88)
cn,l
k
m∈N n+2
m T (cid:98) m
cn,l
N (cid:99)ϕ(m)N
m T (cid:98) m
cn,l
N (cid:99)ϕ(m)N
l∈dn−qn
l∈Qn
If we consider T lψn,k and T rψm,s for n, m ∈ N, n (cid:54)= m, l, r ∈ Z, k ∈ dn − qn,
s ∈ dm − qm, the orthonormality follows from T lψn,k ∈ Wn, T rψm,s ∈ Wm and by
the definition of Wn, Wm.
ogous results. We show that for fixed n ∈ N(cid:110)T N nkψ−n,l : l ∈ d−n − q−n, k ∈ Z(cid:111)
Now we consider the closed subspaces Vn of L2(µ) with n < 0 and show the anal-
is an orthonormal basis of W−n = V−n+1 (cid:9) V−n. First we show that any function
U (−n+1)ϕj can be written as a linear combination of functions U (−n)ϕi and ψ−n,l,
i ∈ N, l ∈ d−n − q−n. This linear combination is precisely
U (−n+1)ϕj
(cid:68) (cid:88)
(cid:88)
m∈N 2
=
m∈N 2
= δl,k.
c−n,l
m c−n,k
m
=U (−n+1)
n,i
j bn,i
m +
b
−n,l
c
j
c−n,l
m
N (cid:99)ϕ(m)N
(cid:124)
(cid:88)
(cid:88)
(cid:88)
m∈N 2
i∈Q−n
=U (−n+1)
(cid:88)
(cid:88)
i∈Q−n
i∈Q−n
=
=
n,i
j
b
i∈Q−n
j U (−n)ϕi +
n,i
b
j U (−n)ϕi +
n,i
b
(cid:88)
=δj,m
l∈d−n−q−n
(cid:123)(cid:122)
N (cid:99)ϕ(m)N +
(cid:88)
U (−n+1) (cid:88)
m∈N 2
−n,l
j
m∈N 2
−n,l
j
c
ψ−n,l.
m∈N 2
(cid:88)
m T (cid:98) m
bn,i
(cid:88)
(cid:88)
l∈d−n−q−n
c
l∈d−n−q−n
T (cid:98) m
(cid:125)
(cid:88)
−n,l
c
j
m T (cid:98) m
c−n,l
N (cid:99)ϕ(m)N
l∈d−n−q−n
m T (cid:98) m
c−n,l
N (cid:99)ϕ(m)N
We have to show the orthonormality only for ψ−n,l and ψ−n,k since T is a unitary
operator and U (−n+1)T kϕj are mapped to orthonormal functions. For ψ−n,l and
ψ−n,k, l, k ∈ d−n − q−n, the orthonormality follows from
(cid:104)ψ−n,lψ−n,k(cid:105) =
m U (−n+1)T (cid:98) m
c−n,k
m U (−n+1)T (cid:98) m
c−n,l
N (cid:99)ϕ(m)N
N (cid:99)ϕ(m)N
(cid:69)
(cid:12)(cid:12)(cid:12) (cid:88)
m∈N 2
16
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
Furthermore, it follows that L2(µ) =(cid:76)
cl(cid:0)(cid:83)
(cid:1) = L2(µ). Consequently, we have that
(cid:8)T lψn,k : n ∈ Z, k ∈ dn − qn, l ∈ Z(cid:9)
Wn ∪ V0
n∈N0
k∈Z Wk, since we have shown before that
is an ONB of L2(µ).
Remark 3.4. For the proof of Corollary 1.8 we have to consider the first part of the
proof of Theorem 1.6 and show the orthonormality between ψn,k and ϕi in addition,
which follows from the construction of the mother wavelets.
two-sided multiplicative MRA.
3.2. Abstract multiplicative Multiresolution analysis. In this section we
want to consider how the general results simplify if we impose the extra condi-
tion of a multiplicative MRA.
Recall from the introduction that in the case of Definition 1.4, we say that we
have a multiplicative MRA if there exists an operator U such that U (n) = U n for all
n ∈ N and U (−n) = (U∗)n, n ∈ N. We then say(cid:0)µ, ((U)n , (U∗)n)n∈N0
,T(cid:1) allows a
Proof of Lemma 1.9. Recall that (cid:8)T lψ0,k : k ∈ d0 − N , l ∈ Z(cid:9) is an orthonormal
basis of W0. We have ψ0,k =(cid:80)
N (cid:99)ϕ(m)N and we show that for fixed
n ∈ N, U nW0 = Wn. First it follows that U nT mψ0,k ∈ Wn ⊂ Vn+1, n ∈ N, m ∈ Z,
since
The key observation is contained in Lemma 1.9 which we prove first.
m UT (cid:98) m
m∈N 2 c0,k
(cid:88)
l∈N 2
l U n+1T (cid:98) l
c0,k
N (cid:99)+N mϕ(l)N
U nT mψ0,k =
and
(cid:104)U nT mψ0,kU nT rϕi(cid:105)
l UT (cid:98) l
c0,k
=(cid:104)U nT m (cid:88)
=(cid:104)U n+1T N m (cid:88)
=δm,r · (cid:88)
l∈N 2
c0,k
l a0,i
l∈N 2
l = 0.
l T (cid:98) l
c0,k
l∈N 2
N (cid:99)ϕ(l)NU nT r (cid:88)
N (cid:99)ϕ(l)NU n+1T N r (cid:88)
j∈N 2
j∈N 2
j UT (cid:98) j
a0,i
N (cid:99)ϕ(j)N(cid:105)
j T (cid:98) j
a0,i
N (cid:99)ϕ(j)N(cid:105)
Consequently, U nW0 ⊂ Wn. Now consider U n+1T mϕj ∈ Vn+1, m ∈ Z, j ∈ N, and
we show that this can be written as a linear combination of functions U nT lϕi and
U nT rψ0,k, l, r ∈ Z, i ∈ N, k ∈ d0 − N, by considering the scalar product. First
we recall that UT (cid:98) k
k ψ0,l for k ∈ N 2 from the
proof of Theorem 1.6, and hence for k ∈ N 2
N (cid:99)ϕ(k)N =(cid:80)
k ϕi +(cid:80)
l∈d0−N c0,l
i∈N a0,i
1 = (cid:104)UT (cid:98) k
a0,i
k ϕi +
k ψ0,l(cid:105)
c0,l
N (cid:99)ϕ(k)N(cid:88)
(cid:88)
i∈N
(cid:88)
i∈N
=
a0,i
k a0,i
k +
c0,l
k c0,l
k .
l∈d0−N
(cid:88)
l∈d0−N
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
17
It follows that for m ∈ N n+2 with m = k + N 2k1, k ∈ N 2, k1 ∈ N n, we have
N (cid:99) = (cid:98) k
(cid:98) m
k ϕi + U nT k1 (cid:88)
(cid:88)
l∈d0−N
N (cid:99)ϕ(l)N(cid:105)
k ψ0,l(cid:105)
c0,l
i UT (cid:98) i
c0,l
N (cid:99)ϕ(i)N(cid:105)
l UT (cid:98) l
(cid:88)
a0,i
i∈N 2
N (cid:99)ϕ(l)N(cid:105)
=(cid:104)U n+1T (cid:98) k
N (cid:99) + N k1 and (m)N = (k)N, and so
(cid:104)U n+1T (cid:98) k
a0,i
N (cid:99)+N k1 ϕ(k)NU nT k1(cid:88)
N (cid:99)+N k1 ϕ(k)NU nT k1(cid:88)
N (cid:99)+N k1ϕ(k)NU nT k1 (cid:88)
(cid:88)
l∈d0−N
l T (cid:98) l
(cid:88)
a0,i
N (cid:99)+N k1ϕ(k)NT N k1(cid:88)
N (cid:99)+N k1 ϕ(k)NT N k1 (cid:88)
+ (cid:104)U n+1T (cid:98) k
l∈N 2
c0,l
k
l∈N 2
c0,l
k
+ (cid:104)T (cid:98) k
=(cid:104)T (cid:98) k
a0,i
k
a0,i
k
i∈N
i∈N
i∈N
l∈d0−N
i∈N 2
=
a0,i
k a0,i
k +
c0,l
k c0,l
k
(cid:88)
l∈d0−N
(cid:88)
i∈N
i T (cid:98) i
c0,l
N (cid:99)ϕ(i)N(cid:105)
=1.
Now we notice that we can write any element k ∈ Z as k = k0 + N n+2l for some
k0 ∈ N n+2 and l ∈ Z. Consequently, with UT NV0 = T UV0 we obtain the general
result for U n+1T kϕj, k ∈ Z, j ∈ N.
To obtain W−n = (U∗)n W−1, W−1 = V0 (cid:9) V−1, n ∈ N, we can proceed as above.
N (cid:99)ϕ(l)N
−1,l
j ψ−1,l.
First we have from the proof of Theorem 1.6 that ψ−1,k =(cid:80)
and that ϕj, j ∈ N, can be represented as ϕj =(cid:80)
0,j U∗ϕi+(cid:80)
−1,k
T (cid:98) l
l∈N 2 c
l
l∈d−1−N c
i∈N b
−1,i
With these observations we obtain as above that
(cid:104)(U∗)n T mϕj (U∗)n−1 T rψ−1,k(cid:105) = 0
and
(cid:104)(U∗)n−1 T kϕj (U∗)n(cid:88)
i∈N
b
j ϕi + (U∗)n−1 (cid:88)
−1,i
l∈d−1−N
−1,l
j ψ−1,l(cid:105) = 1.
c
(cid:3)
Remark 3.5. If we have UU∗ = id, then W0 = U (W−1). Notice that U is not
necessarily injective on W−1.
Now we turn to the mother wavelets.
Remark 3.6. If U (n) = U n, U (−n) = (U∗)n, then we only consider the mother
wavelets for k ∈ d0 − N. So
ψ0,k = U (cid:88)
l∈N 2
l T (cid:98) l
c0,k
N (cid:99)ϕ(l)N ,
where the coefficients are from (3.2) and we define ψk := ψ0,k.
For the negative indexed part of the construction we write for k ∈ d−1 − N
(cid:88)
l∈N 2
ψ−,k =
−1,k
c
l
T (cid:98) l
N (cid:99)ϕ(l)N .
and under condition (3.10) this simplifies to
U nT kϕj =
N (cid:99)ϕ(l)N
(cid:88)
l U n+1T N k+(cid:98) l
a0,j
(cid:88)
l∈N 2
i∈N
:= a0,j
l
U nT kϕj =
i U n+1T N k+jϕi.
a0,j
18
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
3.3. Translation completeness. In the following we assume a stronger condition
than (1e) of Definition 1.4, namely that the father wavelets are translation complete,
i.e. for j ∈ N
(3.10)
ϕj ∈ spanU(cid:8)T jϕi : i ∈ N(cid:9) .
This condition implies that there exist complex numbers a0,j
, i ∈ N, such that
i
(cid:88)
i∈N
ϕj =
i UT jϕi.
a0,j
We would like to point out that this condition is also satisfied for the particular
case of MIM where the father wavelets ϕj, j ∈ N , are chosen to be the scaled
characteristic functions on the cylinder sets [j] (see Section 4).
We have that (3.2) takes the following form for k ∈ Z, j ∈ N, n ∈ N,
j
j
. We now show that condition (3.10)
To simplify the notation we set aj
l
allows us to simplify the construction of the mother wavelets.
Lemma 3.7. Under condition (3.10) one possible choice of the matrix M0 in (3.8)
has a block structure consisting of N blocks.
(cid:0)ak
(cid:1)
Proof. Define Qk := {j ∈ N : UT kϕj (cid:54)= 0} and qk := card Qk for each k ∈ N. Then
(cid:1)
(cid:0)ak
j∈Qk is a vector of length qk and we choose qk−1 linearly independent vectors to
l ∈ qk\{0}, orthonormal to(cid:0)ak
,
j∈Qk of length qk. Via the Gram-Schmidt process we obtain vectors
j∈Qk
j = 0 if j ∈ N\Qk we extend
(cid:16)
(cid:17)
is a
The matrix (cid:99)M0 = (hij)i∈q1,j∈N 2 given with the blocks Mk, k ∈ N, by for k = 0
the vectors
j∈Qk
matrix of size qk × N.
j∈Qk. By setting ck,l
. Then Mk :=
(cid:0)ak
(cid:17)
l∈qk\{0},j∈N
(cid:1)
(cid:17)
(cid:16)
(cid:16)
(cid:17)
(cid:16)
ck,l
j
ck,l
j
j
j∈N
ck,l
j
to
ck,l
j
j∈N
(cid:1)
j
(hij)i∈q0,j∈N = M0,
for k ∈ N\{0}
(hij)i∈(cid:80)k
l=0 ql\(cid:80)k−1
l=0 ql,j∈(k+1)N\kN = Mk
and otherwise zero satisfies the conditions imposed on M0 in (3.8), i.e.
restrict the columns to those in D1, it is unitary, and (cid:100)M0 is of size q1 × N 2 since
(cid:80)
k∈N qk = q1. We notice that (cid:99)M0 is ordered in a different way than M0, since
the rows(cid:0)ak
i=0 qi\(cid:80)k−1
(cid:1)
simpler form for k = 0, l ∈ q0\{0} and for k ∈ N\{0}, l ∈(cid:80)k
(1) If U (n) = U n, U (−n) = (U∗)n and (3.10), the mother wavelets take the
i=0 qi,
j∈Qk are not grouped in M0.
Remark 3.8.
if we
(cid:3)
j
as
(cid:88)
j∈N
ψl =
j UT kϕj,
ck,l
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
19
where the coefficients are as constructed in Lemma 3.7. For negative in-
dexed part we define for k ∈ d−1 − N
−1,k
c
l
N (cid:99)ϕ(l)N .
ψ−,k =
T (cid:98) l
(2) In the case of (3.10), or the slightly weaker statement
(3.11)
U (n)T (cid:98) k
N (cid:99)ϕ(k)N =
i U (n+1)T N(cid:98) k
an,k
N (cid:99)+(j)N ϕi
(cid:88)
(cid:88)
l∈N 2
i∈N
we can obtain the coefficients for the mother wavelets by constructing for
each k ∈ N n with U (n)T kϕj (cid:54)
(cid:54)= 0 for at least one j ∈ N a matrix of size
qn,k × qn,k, where qn,k = card{j ∈ N : U (n)T kϕj (cid:54)
(cid:54)= 0} instead of one
unitary matrix of size dn × dn. In this way we need at most N n matrices
on the scale n ∈ N.
i
Now we turn to a correspondence to the construction of a wavelet basis for MIM.
The next proposition shows how the incidence matrix of MIM plays a role in the
MRA.
holds that a0,j
k ∈ Z, U nT kϕj (cid:54)= 0 if and only if for all i = 0, . . . , n − 2, UT ki+1ϕki (cid:54)
(cid:54)= 0 if and only if UT jϕi (cid:54)= 0, i, j ∈ N, then we have for n ∈ N,
(cid:54)= 0 and
Proposition 3.9. In the case of U (n) =(cid:0)U (1)(cid:1)n, n ∈ N0, (3.10) and if it further
UT k0 ϕj (cid:54)= 0, where k =(cid:80)n−1
iteratively. Notice that U 2T k0+N k1ϕj = UT k1(cid:0)UT k0ϕj
UT k1ϕk0 = UT k1(cid:88)
i UT k0ϕi = U 2T N k1+k0(cid:88)
Proof. We prove this for k = k0 + N k1, k0, k1 ∈ N. The general result follows
U 2T k0+N k1 ϕj (cid:54)= 0 it follows that UT k0ϕj (cid:54)= 0. Besides we have that
(cid:1). Consequently, from
i=0 kiN i + lN n, ki ∈ N , i ∈ n, and l ∈ Z.
i ϕi (cid:54)= 0
ak0
ak0
i∈N
i∈N
if U 2T k0+N k1 ϕj (cid:54)= 0.
If we assume that UT k1ϕk0 (cid:54)= 0 and UT k0ϕj (cid:54)= 0 then
U 2T N k1+k0 ϕj = UT k1UT k0 ϕj
(cid:16)
(cid:16)
ak0
j
ak0
j
=
=
ϕk0 − (cid:88)
(cid:17)−1 UT k1
UT k1 ϕk0 − (cid:88)
(cid:17)−1
i∈N\{j}
i∈N\{j}
i UT k0ϕi
ak0
i U 2T N k1+k0ϕi
ak0
(cid:54)= 0,
since UT k1ϕk0 −(cid:80)
c(cid:0)U (1)(cid:1)n T kϕj for all n ∈ N, k ∈ N n and j ∈ N and (3.10), where c may depend
(cid:3)
Remark 3.10. We can show the same result if for some c ∈ R, we have U (n)T kϕj =
i U 2T N k1+k0 ϕi = ak0
j UT N k1+k0ϕj (cid:54)= 0.
i∈N\{j} ak0
on n, k, j.
Under the conditions of Proposition 3.9 we can give a N × N matrix A, which
coincides with the incidence matrix in the case of MIM given by A = (Aij)i,j∈N
with
(cid:40)
Aij :=
if UT iϕj = 0,
else.
0,
1,
20
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
4. Applications to Markov Interval Maps
4.1. Multiresolution Analysis for MIM. Now we apply the results of Section
3 to Markov Interval Maps. More precisely, we construct a wavelet basis on the
L2-space of a limit set of a Markov Interval Map translated by Z with respect to
a measure. First we consider the case where we do not have any relation between
the measures of νZ ([ij]) and νZ ([i]), νZ ([j]). In this case we cannot define only
one operator U, but on each scale n ∈ Z we consider a different operator U (n).
n∈Z and T are defined in (1.3), (1.4) and (1.2) respectively.
(1) Notice that in general we have U (1)U (1) (cid:54)= U (2) since the multiplicative
νZ([kij]) for U (2) on the cylinder
νZ([k)
νZ([kl]) for U (1)U (1) and
νZ([ij])
Remark 4.1.
Consequently, we obtain a family of operators(cid:0)U (n)(cid:1)
The operators(cid:0)U (n)(cid:1)
(cid:113) νZ([j])
(cid:113) νZ([j])
(3) The operators(cid:0)U (n)(cid:1)
constant
sets may differ.
(2) The operator T is unitary.
n∈Z.
U (n)f ∈ L2(νZ).
Define the N father wavelets as ϕi := (µ([i]))
Remark 4.2. Notice that for ω ∈ Σn
N nl, l ∈ Z, we have
(cid:40)
n∈Z are well defined, namely for f ∈ L2(νZ) we have
−1/2
A, j ∈ N and k ∈ Z with k =(cid:80)n−1
1[i] for i ∈ N.
i=0 ωn−1−iN i +
−1/2(cid:88)
j∈N
ϕi(x)
=
j∈N
= (νZ([i]))
(cid:115)
(cid:88)
(cid:115)
(cid:88)
=U (1)T i(cid:88)
j∈N
=
j∈N
νZ([ij])
νZ([i])
(cid:115)
1[ij](x)
(cid:115)
(cid:88)
k∈N
νZ([ij])
νZ([i])
ϕj(x).
νZ([ij])
νZ([i])
νZ([j])
νZ([ij])
· (µ([j]))
−1/2
1[ij](x)
νZ([k])
νZ([ik])
1[ik](x) · ϕj
(cid:0)τ−1
(x)(cid:1)
i
(4.1)
U (n)T kϕj =
Now we turn to the proof of the properties of(cid:0)U (n)(cid:1)
−1/2 T l1[ωj],
0,
(νZ([ωj]))
if Aωn−1j = 0,
else.
n∈Z and T stated in Propo-
sition 1.3.
Proof of Proposition 1.3. ad (1): Let n ∈ N, f ∈ L2(νZ), x ∈ R, then
T U (n)f (x)
νZ([j])
νZ([ωj])
1[ωj](x − 1 − k) · f
ω (x − 1 − k) +
τ−1
ωn−1−iN i + N nk
(cid:32)
(cid:32)
n−1(cid:88)
n−1(cid:88)
i=0
i=0
ωn−1−iN i + N nl − N n
(cid:33)
(cid:33)
νZ([j])
νZ([ωj])
1[ωj](x − l) · f
ω (x − l) +
τ−1
(cid:88)
(cid:88)
k∈Z
(cid:88)
(cid:88)
ω∈Σn
A
=
=
(cid:115)
(cid:115)
(cid:88)
(cid:88)
j∈N
j∈N
l∈Z
ω∈Σn
=U (n)T N n
A
f (x).
ad (3): Let i ∈ N, x ∈ R, then
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
21
(cid:33)
ωn−1−iN i
(cid:115)
νZ([ωj])
νZ([l])
ϕj
(cid:32)
x − n−1(cid:88)
(cid:32)
x − n−1(cid:88)
i=0
i=0
νZ([ωj])
νZ([l])
ϕj
ωn−1−iN i − N nk
(cid:33)
.
ad (2): Notice that for n ∈ N, l ∈ N, k ∈ Z,
j∈N
and
A:ω0=l
ω∈Σn
(cid:88)
U (−n)ϕl(x) =
(cid:88)
(cid:115)
ad (4): Let n ∈ N and k = (cid:80)n−1
l =(cid:80)n−1
i=0 ωn−1−iN i + N nl1, ω ∈ Σn
U (−n)T kϕl(x) =
(cid:88)
(cid:88)
ω∈Σn
A:ω0=l
j∈N
i, j ∈ N then
Consequently, T N nkU (−n)ϕj = U (−n)T kϕj for all k ∈ Z, n ∈ N, j ∈ N.
i=0 ωn−1−iN i + N nk1, ω ∈ Σn
A, k1 ∈ Z and
A, l1 ∈ Z and Aωn−1i = 1 and Aωn−1j = 1 for
(cid:104)U (n)T kϕiU (n)T lϕj(cid:105) = (cid:104)(νZ([ωi]))
−1/2 T k1 1[ωi] (νZ([ωj]))
−1/2 T l1 1[ωj](cid:105)
= δk1,l1δ(ω,i),(ω,j).
Otherwise, we have U (n)T kϕi = 0 or U (n)T lϕj = 0.
Furthermore for n ∈ N, k, j ∈ Z, i, m ∈ N, we have
=
ω∈Σn
(cid:104)U (−n)T kϕiU (−n)T lϕm(cid:105)
νZ([ωj])
νZ([i])
(cid:115)
(cid:68) (cid:88)
(cid:88)
(cid:115)
(cid:88)
(cid:88)
=δk,l · δi,m · (cid:88)
(cid:88)
A:ω0=m
ω∈Σn
A:ω0=i
j∈N
j∈N
νZ([ωj])
νZ([m])
T
ω∈Σn
A:ω0=i
j∈N
=δ(k,i),(l,m),
νZ([ωj])
νZ([i])
(cid:80)n−1
i=0 ωn−1−iN i+N nkϕj
(cid:80)n−1
i=0 ωn−1−iN i+N nlϕj
T
(cid:69)
where we used in the second equality that (cid:104)T kϕjT lϕi(cid:105) = δ(k,j),(l,i).
(cid:115)
ad (5): Let n ∈ N, f ∈ L2(νZ), x ∈ R, then
U (n)U (−n)f (x)
(cid:88)
l∈Z
f
(cid:88)
A
r∈N
νZ([r])
(cid:88)
(cid:88)
(cid:88)
(cid:88)
(cid:88)
νZ([(cid:101)ωr])
1[(cid:101)ωr](x − l)
(cid:32)
(cid:101)ω∈Σn
n−1(cid:88)
(cid:101)ωn−1−iN i + N nl − n−1(cid:88)
τ−1(cid:101)ω (x − l) +
(cid:32)
(cid:32)
(cid:101)ωn−1−iN i + N nl − n−1(cid:88)
n−1(cid:88)
τ−1(cid:101)ω (x − l) +
(cid:88)
(cid:88)
ω∈Σn
j∈N
k∈Z
τω
1[ωj](x − k) · f (x)
i=0
i=0
i=0
i=0
A
=
=
1[j]
k∈Z
ω∈Σn
A
j∈N
(cid:115)
νZ([ωj])
νZ([j])
(cid:33)
(cid:33)
(cid:33)
+ k
ωn−1−iN i − N nk
ωn−1−iN i − N nk
=f (x),
where we used in the third equality that i = r, ω =(cid:101)ω and k = l since otherwise it
is zero.
(cid:17)
(cid:33)
(cid:3)
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
22
l ∈ Z, with k =(cid:80)n−1
ad (6): For n ∈ N, k ∈ Z, j ∈ N, x ∈ R, with U (n)T kϕj (cid:54)= 0, there is ω ∈ Σn
A,
i=0 ωn−1−iN i + N nl and so
U (−n)U (n)T kϕj(x) = U (−n)(cid:16)
= T N nlU (−n)(cid:16)
(cid:80)n−1
(νZ([ωj]))
(cid:17)
−1/2 T l1[ωj](x)
(νZ([ωj]))
−1/2
1[ωj](x)
−1/2
i=0 ωn−1−iN i
(νZ([j]))
1[j](x)
= T N nlT
= T kϕj(x).
(cid:88)
(cid:88)
k∈Z
ωj∈Σn+1
A
1[j]
(cid:32)
x − n−1(cid:88)
i=0
Remark 4.3. We further notice that for n ∈ N, x ∈ R, we have f ∈ L2(νZ),
U (−n)U (n)f (x) =
ωn−1−iN i − N nk
· f (x),
and consequently, in general we do not have U (−n)U (n) = id.
Now we can turn to the proof of Theorem 1.5.
(cid:110)
(cid:111)
.
Proof of Theorem 1.5. We show the properties (2a) to (2e) of Definition 1.4 with
the father wavelets ϕi = (ν([i])
1[i], i ∈ N.
−1/2
We define the closed subspaces of L2(νZ) for j ∈ N as
V0 := cl span(cid:8)T kϕi : k ∈ Z, i ∈ N(cid:9) ,
Vj := cl span
U (n)T kϕj =
i=0 ωn−1−iN i+
ad (2d): We notice that for ω ∈ Σn
ad (2c): By the definition of Vj we obviously have that(cid:8)U (j)T kϕi : k ∈ Z, i ∈ N(cid:9)
U (j)T kϕi : k ∈ Z, i ∈ N
spans Vj, j ∈ Z. The orthonormality follows from Proposition 1.3 (4).
N nl, l ∈ Z, we have
A, j ∈ N and k ∈ Z with k =(cid:80)n−1
(cid:115)
(cid:88)
A so that k = (cid:80)n−1
i=0 ωn−1−iN i + N nl, l ∈ Z, then
ad (2a): Notice that for n ∈ N, k ∈ Z and i ∈ N we obtain with Proposition 1.3
(cid:115)
U (n)T kϕi = U (n)T kU (1)T i(cid:88)
(cid:115)
= U (n)U (1)T N k+i(cid:88)
If there is not such an ω ∈ Σn
U (n)T kϕj = 0.
νZ([ωji])
νZ([ωj])
U (n+1)T N k+jϕi.
(2) and (3) that
νZ([i])
νZ([ij])
j∈N
i∈N
ϕj
νZ([i])
νZ([ij])
ϕj.
j∈N
So it follows that Vn ⊂ Vn+1 by Remark 4.1 (1).
ad (2b): First we notice that X is either totally disconnected or we can consider
X as an interval in [0, 1]. Furthermore, every characteristic function on a cylinder
[ω] ⊂ ΣA can be obtained by U (n)T kϕj, n ∈ N0, k ∈ Z, j ∈ N. Thus, we are left
to show that(cid:8)T k1[ω] : k ∈ Z, ω ∈ Σ∗
If X is totally disconnected, it follows by the Stone-Weierstrass Theorem that
(cid:9) is dense in L2(νZ).
(cid:9)
(cid:8)T k1[ω] : k ∈ Z, ω ∈ Σ∗
A
A
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
in L2(νZ) and so cl span(cid:8)T k1[ω] : k ∈ Z, ω ∈ Σ∗
is dense in C(R, C), see e.g. [KSS07]. Besides it is well known that C(R, C) is dense
If X = [a, b], notice that every interval I ⊂ [0, 1] can be approximated by τω(X),
ω ∈ Σ∗
A, generates B, thus every element A ∈ B can be
approximated by elements of {τω(X) : ω ∈ Σ∗
A}. Consequently, every elementary
function can be approximated by functions 1τω(X) and so all functions in L2(νZ)
can be approximated by elements of
A. Hence τω(X), ω ∈ Σ∗
(cid:9) = L2(νZ).
23
A
(cid:8)T k1[ω] : ω ∈ Σ∗
A, k ∈ Z(cid:9) =
(cid:110)
Consequently, cl(cid:83)
ad (2e): This follows from Proposition 1.3 (1) and (2).
k∈N Vn = L2(νZ) .
(cid:111)
.
(cid:3)
U (n)T lϕi : n ∈ N0, l ∈ Z, i ∈ N
Next we prove the forward direction of Theorem 1.11. The backward direction
will be shown in Section 4.3.
Proof of Theorem 1.11 "=⇒". We assume that
sided MRA with the father wavelets ϕi = (νZ([i])
holds by (2d) of Definition 1.4 that for n ∈ N
(cid:17)
U (−n) {ϕi : i ∈ N} ⊂ span U (−n+1)(cid:8)T kϕi : i ∈ N , k ∈ N(cid:9) .
νZ,(cid:0)U (n)(cid:1)
n∈Z , T
(cid:16)
−1/2
allows a two-
1[i]. Then in particular, it
We further notice that for n ∈ N, k, i ∈ N,
U (−n)ϕk =
ω∈Σn
A:ω0=k
(cid:88)
(cid:88)
(cid:115)
(cid:88)
(cid:115)
j∈N
(cid:88)
ω∈Σn−1
A :ω0=i
j∈N
and
U (−n+1)T kϕi =
(cid:80)n−1
l=0 ωn−1−lN l
ϕj
νZ([ωj])
νZ([k])
T
(cid:80)n−2
l=0 ωn−2−lN l+N n−1kϕj.
νZ([ωj])
νZ([i])
T
From the precise from of U (−n)ϕk and U (−n+1)T mϕi, n ∈ N, k, m, i ∈ N, it follows
that (cid:104)U (−n)ϕkU (−n+1)T mϕi(cid:105) (cid:54)= 0 only if m = k since
(cid:115)
(cid:88)
(cid:104)U (−n)ϕkU (−n+1)T mϕi(cid:105)
νZ([ωj])
νZ([k])
(cid:88)
j∈N
A:ω0=k
ω∈Σn
=(cid:104) (cid:88)
(cid:88)
(cid:88)
(cid:80)n−1
(cid:88)
ω∈Σn
A:ω0=k
(cid:104)T
ω∈Σn−1
=
=δm,k
A :ω0=i
νZ([ωj])
νZ([i])
T
j∈N
(cid:115)
(cid:88)
(cid:101)ω∈Σn−1
A :(cid:101)ω0=i
(cid:115)
ϕj1T
(cid:88)
(cid:88)
j1∈N
l=0 ωn−1−lN l
(cid:80)n−1
l=0 ωn−1−lN l
ϕj
T
(cid:80)n−2
l=0 ωn−2−lN l+N n−1mϕj(cid:105)
(cid:115)
(cid:88)
νZ([(cid:101)ωj2])
(cid:80)n−2
l=0 (cid:101)ωn−2−lN l+N n−1mϕj2(cid:105)
νZ([ωj1])
νZ([k])
(cid:115)
νZ([i])
j2∈N
(cid:115)
j∈N
ω∈Σn−1
A :ω0=i
νZ([k])
νZ([kωj])
νZ([ωj])
νZ([i])
,
where we used in the third equality the property of Proposition 1.3 (4), namely
(cid:104)T kϕj1T lϕj2(cid:105) = δ(k,j1),(l,j2) for any k, l ∈ Z and j1, j2 ∈ N.
T
l=0 ωn−1−lN l
(cid:80)n−1
(cid:80)n−2
l=0 ωn−2−lN l+N n−1kϕj
ϕj
(cid:80)n−2
l=0 ωn−2−lN l+N n−1kϕj.
T
νZ([kωj])
νZ([k])
(cid:115)
νZ([ωj])
νZ([k])
νZ([kωj])
νZ([k])
T
(cid:115)
(cid:88)
j∈N
j∈N
A:ω0=k
ω∈Σn
(cid:88)
(cid:88)
(cid:115)
(cid:88)
(cid:88)
(cid:88)
(cid:88)
(cid:113) νZ([kωj])
ω∈Σn−1
ω∈Σn−1
j∈N
i∈N
A
=
=
A :ω0=i
νZ([kωj]) = νZ([ωj])
24
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
As a consequence of (2c), (2d) of Definition 1.4 and the observation above it
∈ CN such that
follows that for every n ∈ N, k ∈ N there exist unique
αn,k
(cid:16)
(cid:17)
i
i∈N
i U (−n+1)T kϕi
αn,k
(cid:88)
(cid:88)
αn,k
i
ω∈Σn−1
A :ω0=i
j∈N
(cid:115)
(cid:80)n−2
l=0 ωn−2−lN l+N n−1kϕj.
νZ([ωj])
νZ([i])
T
On the other hand, from the precise form of U (−n)ϕk it follows that
U (−n)ϕk =
=
(cid:88)
(cid:88)
i∈N
i∈N
U (−n)ϕk =
(cid:113) νZ([ωj])
By comparing the coefficients it follows that for every ω ∈ Σn−1
αn,k
νZ([k]) . Consequently, αn,k
νZ([i]) =
i
A , ω0 = i, we have
(cid:16)
(cid:17)2 νZ([k])
i ∈ R+ and
αn,k
i
.
νZ([i])
Now it remains to be shown that cn,k
with ω0 = i and k ∈ N it follows that
i
νZ([kω]) =
νZ([kωj]) =
(cid:88)
j∈N
are independent of n ∈ N. For n ∈ N, ω ∈ Σn
A
(cid:88)
(cid:16)
(cid:17)2
αn,k
i
νZ([k])
(cid:16)
αn,k
i
j∈N
νZ([i])
νZ([ωj])
(cid:17)2
A with ω0 = i as ω =(cid:101)ωωn−1 for a suitable
νZ([k])
.
νZ([i])
= νZ([ω])
On the other hand we can write ω ∈ Σn
(cid:101)ω ∈ Σn−1
A , (cid:101)ω0 = i, and so
(cid:16)
(cid:17)2
αn−1,k
i
νZ([k])
νZ([i])
νZ([kω]) = νZ([k(cid:101)ωωn−1]) = νZ([(cid:101)ωωn−1])
(cid:16)
(cid:17)2
αn−1,k
i
νZ([k])
.
νZ([i])
= νZ([ω])
i
i
i = αm,k
= αn,k
i for αn,k
:= (cid:0)αk
and so αn,k
.
for all n, m ∈ N, k, i ∈ N. In the following
Thus, αn−1,k
(cid:1)2
we write αk
νZ([k])/νZ([i]) for k, i ∈ N, then we have νZ([kωj]) =
κk,ω0νZ([ωj]) for all ω ∈ Σ∗
A, j, k ∈ N. From this property we conclude the Markov
(cid:88)
relation since for any k, i ∈ N
Define κk,i
(cid:88)
i
i
i
ν([ki]) =
ν([kij]) =
κk,iν([ij]) = κk,iν([i])
j∈N
j∈N
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
25
and so
Define πki := κk,iν([i])/ν([k]) =(cid:0)αk
κk,iν([i])
(cid:1)2, then πki is a incidence probability. Conse-
ν([k])
.
ν([ki]) = ν([k])
quently, we have that if a two-sided MRA holds then the measure ν is Markovian.
(cid:3)
The reversed implication will be shown in Section 4.3.
i
4.2. Mother wavelets for MIM. In this section we are in the case of Remark 3.8
(2) and so we consider for each father wavelet ϕi, i ∈ N, a matrix of coefficients;
more precisely on each scale we have to consider for each element of the alphabet
for
N a matrix of coefficients. We slightly change the notation from cn,k,l
ω ∈ Σn
A, since the information about n and k are coded; n is given by the length of
A we need a matrix of size qωn × qωn, where qωn = card{j ∈ N :
∈ C, j ∈ N, k ∈ qωn\{0}, such that the
a word and k =(cid:80)n−1
For ω ∈ Σn+1
i=0 ωn−1−iN i.
Aωnj = 1}. First we determine cω,k
(qωn × qωn)-matrix
to cω,l
j
j
where Dωn = {j ∈ N : Aωnj = 1} is unitary. This is done as explained above via
the Gram-Schmidt process.
i=0 ωn−iN i the basis functions as: for l ∈ qωn\{0}
We define for ω ∈ Σn+1
,
j∈Dωn
k∈qωn\{0},j∈Dωn
Mω :=
j
j
(cid:17)
(cid:16)(cid:112)νZ([ωj])
(cid:17)
Aωnjcω,k
(cid:16)
A , k =(cid:80)n
ψω,l = U (n)T k(cid:88)
(cid:88)
j∈N
Aωnjcω,k
j
ψω,k =
j∈N
These functions can be written differently for ω ∈ Σn+1
A , k ∈ qωn\{0}, as
Aωnjcω,l
j ϕj.
· (νZ([ωj]))
−1/2 · 1[ωj].
From Theorem 1.6 and Theorem 1.5 the following corollary follows.
Corollary 4.4. An orthonormal basis for L2(νZ) is given by
(cid:8)T lψω,k : l ∈ Z, ω ∈ Σ∗
cl span(cid:8)T lψω,k : l ∈ Z, ω ∈ Σn
A, k ∈ {1, . . . , qωω−1 − 1}(cid:9) ∪(cid:8)T lϕj : l ∈ Z, j ∈ N(cid:9) .
A, k ∈ {1, . . . , qωn−1 − 1}(cid:9) = Vn (cid:9) Vn−1.
Remark 4.5. In fact, the proofs of Theorem 1.6 and Theorem 1.5 show that we
have for n ∈ N
4.3. MRA for a Markov measures. In this section we construct a wavelet basis
on the limit set translated by Z where the underlying measure ν is Markovian. For
this fix a probability vector p = (p0, p1, . . . , pN−1) and a (N × N ) stochastic matrix
Π = (πjk)j,k∈N such that for ω ∈ Σn
A we have
n−2(cid:89)
ν([ω]) = pω0
πωiωi+1.
i=0
Furthermore, we have that πjk = 0 if Ajk = 0.
This is a special case of the one in the last section. Therefore, we omit some
proofs here and mainly state the results, so that the differences become clear.
In this construction we only have to define one operator U since we obtain U (n)
by U n, i.e. by iteration of U. Another main difference is that we do not need one
matrix for every ω ∈ Σ∗
A to obtain the mother wavelets, but we only need matrices
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
26
for ω ∈ Σ1
1.9.
A = N. So we need not more than N 2 matrices. This follows from Lemma
The setting is as defined in Section 2. Set U := U (1)and so it takes the form in
(1.1). By the Markov property we have νZ([i])
and hence one easily verifies
that U (n) = U n. Also notice that U is not unitary unless we have that Aij = 1 for
all i, j ∈ N.
νZ([ji]) = pi
pj πji
Now we turn to the form of U∗.
Lemma 4.6. U∗ has the form
(4.2) U∗f (x) =
(cid:88)
(cid:88)
(cid:88)
k∈Z
j∈N
i∈N
(cid:114) pjπji
pi
· 1[i](x − j − N k) · f (τj(x − j − N k) + k).
Remark 4.7. Notice that U∗ = U (−1) and (U∗)n = U (−n).
Proof. To prove that U∗ has the form above we use the Z-translation invariance
of the measure νZ and the fact that dνZ◦τj
on [i]. We obtain this Radon-
Nikodym derivative since for a cylinder set [ω], ω ∈ Σ∗
A, we have
dνZ = pj πji
pi
νZ (τj([ω])) = pjπjω0
πωiωi+1
n(cid:89)
i=0
and νZ([ω]) = pω0
i=0 πωiωi+1.
Consequently, we obtain that for f, g ∈ L2(νZ)
(cid:104)U fg(cid:105)
j∈N
i∈N
(cid:81)n
(cid:114) pi
(cid:88)
(cid:88)
(cid:114) pi
(cid:88)
(cid:88)
(cid:114) pi
(cid:88)
(cid:88)
(cid:114) pi
(cid:88)
(cid:88)
(cid:88)
(cid:88)
(cid:88)
i∈N
i∈N
i∈N
j∈N
j∈N
j∈N
pjπji
pjπji
pjπji
k∈Z
(cid:88)
(cid:88)
(cid:88)
(cid:88)
k∈Z
k∈Z
k∈Z
=
=
=
=
=
f (x)
k∈Z
j∈N
i∈N
=(cid:104)fU∗g(cid:105),
with U∗g as in (4.2).
· 1[ji](x − k) · f (τ−1
j
(x − k) + j + N k)g(x)dνZ(x)
· 1[ji](x) · f (τ−1
j
(x) + j + N k)g(x + k)dνZ(x)
· 1[ji](τj(x)) · f ((x) + j + N k)g(τj(x) + k)dνZ(τj(x))
pjπji
(cid:114) pjπji
pi
· 1[ji](τj(x)) · f ((x) + j + N k)g(τj(x) + k) · pjπji
pi
· dνZ(x)
· 1[i](x − j − N k) · g(τj(x − j − N k) + k)dνZ(x)
(cid:3)
Now we turn to the definition of the father wavelets which we use in the MRA.
Define the N father wavelets as ϕi = (νZ([i]))
Remark 4.8. Notice that the family of father wavelets (ϕi)i∈N is orthonormal by
definition.
1[i] for i ∈ N.
−1/2
Now we turn to the properties of the operators U and T given in Proposition
1.1.
Proof of Proposition 1.1. We have that (1), (2), (3) and (4) follow directly from
Proposition 1.3 since it is a special case of U (n) in the section above.
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
27
(cid:88)
(cid:88)
ad (5): This proof is analogous to the one of (4) or Proposition 1.3 (5). We
obtain for f ∈ L2(νZ)
k∈Z(cid:80)
(cid:80)
j∈N
all i, j ∈ N.
U∗U f (x) =
Aji1[i](x − j − N k) · f (x).
(cid:80)
i∈N Aji1[i](x − j − N k) = 1 for all x ∈ R if and only if Aji = 1 for
(cid:3)
j∈N
k∈Z
Now we turn to the proof of the backward direction of Theorem 1.11. Some of
the properties follow directly from the proof of Theorem 1.5.
Proof of Theorem 1.11 "⇐=". We show the properties (1a) to (1f) of Definition 1.4.
The property (1b) follows from Theorem 1.5.
ad (1e): For n ∈ N0 it follows directly from Theorem 1.5. For n ∈ Z, n < 0,
x ∈ R, k ∈ N, it follows by
n
i∈N
ϕk(x)
(cid:88)
(cid:118)(cid:117)(cid:117)(cid:116)n−2(cid:89)
(cid:88)
(cid:32) (cid:88)
x −
n−2(cid:88)
n
A :ω0=k
√
n−1
A
ω∈Σ
πk,j
l=1
(U∗)
=
=
=
ω∈Σ
(cid:88)
j∈N
ϕi
(cid:88)
j∈N
ωn−2−lN l − kNn−1
√
l=0
πk,j (U∗)
n−1 T kϕj(x).
x −
n−1(cid:88)
l=0
ωn−1−lN l
πωl,ωl+1 · πk,ω1πωn−1,iϕi
(cid:88)
(cid:118)(cid:117)(cid:117)(cid:116)n−3(cid:89)
i∈N
:ω0=j
l=1
πωl,ωl+1 · πj,ω1πωn−2,i
(cid:33)
(cid:88)
j∈N
ad (1a): For n ∈ N0 it follows directly from Theorem 1.5. For n ∈ Z, n < 0,
k ∈ N, it follows from
(U∗)
n
ϕk =
√
πk,j (U∗)
n−1 T kϕj.
ad (1c): We have that(cid:84)
grows in n ∈ N. More precisely, for j ∈ N
νZ (supp ((U∗)n ϕj)) =
Consequently, {0} =(cid:84)
n∈Z Vn = {0} , because the support of (U∗)n ϕj, j ∈ N,
(cid:88)
: ω0 = j, ωn = i(cid:9)(cid:1) .
νZ ([i])(cid:0)card(cid:8)ω ∈ Σn+1
j∈Z Vj since any function f ∈(cid:84)
j∈Z Vj must be constant for
i∈N
A
every n ∈ N on supp ((U∗)n ϕj), for j ∈ N.
ad (1d): This property follows directly from the definition of the spaces Vj and
Proposition 1.3 (4) with the observation that U (n) = U n and U (−n) = (U∗)n,
n ∈ N0.
(cid:3)
ad (1f): This property follows from Proposition 1.1 (4) and (5).
Remark 4.9. Now we give some remarks concerning the father wavelets.
(1) The relation for the functions ϕi, i ∈ N, can also be written as
(cid:88)
l∈N
(cid:0)U T lϕj
(cid:1)
(ϕj)t
j∈N =
Ml
j∈N ,
(cid:40)√
28
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
(cid:40)
where the Ml are (N × N )-matrices with (Ml)n,k =
n, k ∈ N.
l ∈ Z, i.e. k is in the N-adic expansion. Then we obtain
(2) Notice that for k ∈ Z we can write k = a0 + N l, where a0 ∈ N and some
0,
else,
πlk, n = l,
for
0,
(pa0 · πa0j)
U T kϕj =
(3) Notice that in (cid:8)U nT kϕi : n ∈ N, k ∈ Z, i ∈ N(cid:9) some functions are con-
in the N-adic expansion, k =(cid:80)n−1
stantly zero. These functions are precisely those where for k ∈ Z written
j=0 kn−1−jN i + lN n, kj ∈ N, l ∈ Z, either
Akj kj+1 = 0 for some j ∈ {0, . . . , n − 2} or Akn−1i = 0.
−1/2 T l1[a0j],
if Aa0j = 0,
else.
4.3.1. Mother wavelets for Markov measures. The construction of the mother wavelets
simplifies in this setting because we only have to consider mother wavelets for one
scale and obtain the other by iterative application of the operators U and T by
Lemma 1.9. The mother wavelets are constructed via N matrices as given in
Lemma 3.7 and so the mother wavelets are defined for k ∈ N and l ∈ qk\{0},
by
ψk,l = U T k(cid:88)
j∈N
Akjck,l
j ϕj
(3) Alternatively we can define the mother wavelets as the elements of the
k∈N qk ≤ N 2. In the case
of N 2 mother wavelets we are back in the case of fractals given by an IFS.
for coefficients ck,l
Remark 4.10.
j ∈ C as in Lemma 3.7.
√
l=1 AkiAkjck,l
(1) The number of mother wavelets we obtain is(cid:80)
(2) Notice that(cid:80)qk−1
vector(cid:0)ψk,l(cid:1)t
(cid:88)
(cid:18)(cid:16)
(cid:118)(cid:117)(cid:117)(cid:116)pω0
(cid:17)
n−2(cid:89)
(νZ([ωj]))1/2 =
l∈{1,...,qk−1} =
l∈qk\{0},j∈N
(cid:88)
Akjck,l
j
i ck,l
πi(i+1)
Akjck,i
j
j +
√
πki
πkj = δi,j.
(cid:19)(cid:0)U T kϕj
(cid:1)t
j∈N .
√
Akjck,i
j
πωn−1j = 0,
(4) Here we can see that we only need mother wavelets for W0 since
i=1
j∈N
j∈N
which was the crucial condition in the case of the last section.
Corollary 4.11.(cid:8)U nT mψk,l : n ∈ N0, m ∈ Dn,k, k ∈ N , l ∈ qk\{0}(cid:9)
∪(cid:8)(U∗)n T mψk,l : n ∈ N, m ∈ Z, k ∈ N , l ∈ qk\{0}(cid:9)
∪(cid:8)(U∗)n T kϕj : n ∈ N, k ∈ NZ + l, j, l ∈ N , Ajl = 0(cid:9)
gives an ONB for L2(νZ), where
(cid:40)
n−1(cid:88)
m ∈ Z : m =
(cid:41)
and Aω0k = 1, l ∈ Z
i=0
.
Dn,k =
Remark 4.12.
ωn−1−iN i + N nl, ωi ∈ N , (ω0, . . . , ωn−1) ∈ Σn
A
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
29
(1) Because of U W−1 = W0 we only have to add those functions T kϕj, k ∈ Z,
j ∈ N, with U T kϕj = 0 to the basis of U∗ (W0) to obtain a basis of W−1.
(2) Notice that
ψk,l = U T k(cid:88)
(cid:88)
i∈N
Akick,l
i
=
Akick,l
i ϕi
· (pk · πki)
−1/2 · 1[ki].
(cid:1)
i∈N
matrix are considered, i.e. the IFS has the form(cid:0)τi(x) = x+i
4.4. Examples. In the construction of [MP09] only Cantor sets with incidence
i∈N, and there exists a
incidence matrix A. The limit set has then the Hausdorff dimension δ = dimH (X) =
log N , where r(A) is the spectral radius of A. So we consider the δ-dimensional
log r(A)
Hausdorff measure µ restricted to the by Z translated set X.
It follows that
pj = µ([j]) and πij = N−2δpj
. Consequently, in this case we can rewrite our
conditions for obtaining the coefficients of the mother wavelets in a simpler way.
pi
N
More precisely, for k ∈ N instead of(cid:88)
(cid:88)
we obtain the condition
j∈N
√
πkj = 0
Akjck,i
j
√
pj = 0.
Akjck,i
j
j∈N
Although the basis in [MP09] is only given in terms of the representation of a
Cuntz-Krieger algebra we can now give a scaling operator U in the sense of (1.1)
for this case. More precisely, we obtain
1[j](x − k) · f (τ−1
j
(x − k) + j + N k).
U f (x) = N δ(cid:88)
(cid:88)
k∈Z
j∈N
Proof of Example 1.2: We clearly have that the β-transformation belongs to
the class of Markov measures. Consequently, we have that (µ, U, T ) allows a MRA.
We can construct the mother wavelets along the lines of Section 4.3. Since we have
that in this case d0 = 2 and d1 = 1 we only have to construct coefficients for ϕ0
to obtain the mother wavelets. These coefficients are given in the following matrix
which is unitary:
Thus, the mother wavelet is ψ = U(cid:0)√
√
(cid:18) √
(cid:19)
√
2 − β
β − 1
2 − β −√
β − 1
2 − βϕ0 − √
.
β − 1ϕ1
(cid:1). To obtain the basis
we further notice that U T ϕ1 = 0 and so we have to keep T kϕ1, k ∈ 2Z + 1 in the
basis.
5. Operator algebra
In the case of one father wavelet we obtain a so-called low-pass filter function
and high-pass filter functions, in terms of which the mother wavelets are given. Via
these filter functions we obtain a representation of the Cuntz algebra ON, where N
is the number of filter functions. In the case of multiwavelets we can obtain weaker
relations. Here we restrict to the case of MIM with underlying Markov measure as
treated in Section 4.3. These results are in correspondence to results in [BFMP10].
The relations for the father and the mother wavelets can be written in the fol-
For the following we introduce for z ∈ T := {ω ∈ C : ω = 1} the low-pass filter
lowing way:
H(z) =(cid:0)√
πklzk(cid:1)
l,k∈N
30
and for each k ∈ N and z ∈ T the high-pass filter
JANA BOHNSTENGEL AND MARC KESSEBÖHMER
(cid:16)
l zk(cid:17)
With these definitions we obtain the following immediate lemma.
Gk(z) =
Aklck,j
j∈qk\{0},l∈N
.
j∈N, then φ = U H(T )φ and let ψk =(cid:0)ψk,j(cid:1)t
Lemma 5.1. Let φ = (ϕj)t
j∈qk\{0} for
k ∈ N, then ψk = U Gk(T )φ, where the operators U and T are applied to evey entry
in the vector.
Remark 5.2. It follows that for z ∈ T
(cid:88)
j∈N
√
πkjπljzl−k
k,l∈N
H(z)H t(z) =
and for k ∈ N, z ∈ T,
These filter functions lead us to the definitions of certain "isometries".
Definition 5.3. For z ∈ T and f = (f0, . . . , fN−1), fj ∈ L2(T, λ), define
and for k ∈ N, z ∈ T,
SH f (z) =
Gk(z)Gt
k(z) = I.
√
N H t(z)f(cid:0)zN(cid:1)
k(z)f(cid:0)zN(cid:1) .
SGk f (z) = Gt
For these "isometries" we have the following properties.
Proposition 5.4. The following relations hold:
(1) S∗
(2) S∗
(3) S∗
(4) S∗
H SH = I,
SGk = I, k ∈ N,
Gk
H SGk = 0 and S∗
SGj = 0, i, j ∈ N, i (cid:54)= j .
Gk
Gi
SH = 0, k ∈ N,
Remark 5.5. Realize that for z ∈ T, f = (f0, . . . , fN−1), fj ∈ L2(T, λ),
and for k ∈ N, z ∈ T, f = (f0, . . . , fN−1), fj ∈ L2(T, λ),
S∗
H f (z) =
1√
N
H(ω)f (ω)
S∗
Gk
f (z) =
1
N
Gk(ω)f (ω).
(cid:88)
(cid:88)
ωN =z
ωN =z
Proof. ad (1): Let z ∈ T, f = (f0, . . . , fN−1), fj ∈ L2(T), then
S∗
H SH f (z) =
=
H(ω)H t(ω)f (ωN )
H(ω)H t(ω)f (z) = f (z)
ad (2): Let k ∈ N, z ∈ T, f = (f0, . . . , fN−1), fj ∈ L2(T), then
S∗
SGk f (z) =
k(ω)f (z) = f (z)
ad (3): Let k ∈ N, z ∈ T, f = (f0, . . . , fN−1), fj ∈ L2(T), then
GK
Gk(ω)Gt
S∗
H SGk f (z) =
H(ω)Gt
k(ω)f (z) = 0,
ωN =z
ωN =z
(cid:88)
(cid:88)
(cid:88)
(cid:88)
ωN =z
1
N
1
N
ωN =z
since(cid:80)
MULTIRESOLUTION ANALYSIS FOR MARKOV INTERVAL MAPS
31
ωN =z H(ω)Gt
Gk
k(ω) = 0 by summing up the roots of unity.
For S∗
ad (4): Let i, j ∈ N, i (cid:54)= j, z ∈ T, f = (f0, . . . , fN−1), fj ∈ L2(T), then
SH we use that Gk(ω)H t(ω) = 0 by the choice of the coefficients ck,l
j .
S∗
Gi
SGj f (z) =
1
N
by summing up the roots of unity.
Gi(ω)Gt
j(ω)f (z) = 0,
ωN =z
(cid:3)
(cid:88)
Here we have seen that in contrast to the filter functions for a usual MRA
with one father wavelet and a unitary scaling operator U, we do not obtain a
representation of a Cuntz algebra since we do not neccessarily have that SH S∗
H +
= I. So we only obtain weaker relations between these filter
(cid:80)
k∈N\{0} SGk S∗
GK
functions.
References
[Alp93]
Bradley K. Alpert, A class of bases in L2 for the sparse representation of integral
operators, SIAM J. Math. Anal. 24 (1993), no. 1, 246 -- 262. MR 1199538 (93k:65104)
[BFMP10] Lawrence W. Baggett, Veronika Furst, Kathy D. Merrill, and Judith A. Packer, Clas-
sification of generalized multiresolution analyses, J. Funct. Anal. 258 (2010), no. 12,
4210 -- 4228. MR 2609543
Jana Bohnstengel and Marc Kesseböhmer, Wavelets for iterated function systems, J.
Funct. Anal. 259 (2010), no. 3, 583 -- 601. MR 2644098
[BK10]
[Bod07] Mats Bodin, Wavelets and Besov spaces on Mauldin-Williams fractals, Real Anal.
[CK03]
[Dau92]
[DJ06]
[GP96]
[Jon98]
[KS10]
Exchange 32 (2006/07), no. 1, 119 -- 143. MR 2329226 (2008h:42063)
Mark Crovella and Eric Kolaczyk, Graph wavelets for spatial traffic analysis, Proceed-
ings of IEEE Infocom, April 2003.
Ingrid Daubechies, Ten lectures on wavelets, CBMS-NSF Regional Conference Series in
Applied Mathematics, vol. 61, Society for Industrial and Applied Mathematics (SIAM),
Philadelphia, PA, 1992. MR 1162107 (93e:42045)
Dorin E. Dutkay and Palle E. T. Jorgensen, Wavelets on fractals, Rev. Mat. Iberoam.
22 (2006), no. 1, 131 -- 180. MR 2268116 (2008h:42071)
Jean-Pierre Gazeau and Jiri Patera, Tau-wavelets of Haar, J. Phys. A 29 (1996),
no. 15, 4549 -- 4559. MR 1413218 (97f:42054)
Alf Jonsson, Wavelets on fractals and Besov spaces, Journal of Fourier Analysis and
Applications 4 (1998), 329 -- 340, 10.1007/BF02476031.
Marc Kesseböhmer and Tony Samuel, Spectral metric spaces for Gibbs measures,
arXiv:1012.5152 (2010).
[MP09]
[KSS07] Marc Kesseböhmer, Manuel Stadlbauer, and Bernd Stratmann, Lyapunov spectra for
KMS states on Cuntz - Krieger algebras, Mathematische Zeitschrift 256 (2007), 871 --
893, 10.1007/s00209-007-0110-y.
Matilde Marcolli and Anna Paolucci, Cuntz - Krieger algebras and wavelets on fractals,
Complex Analysis and Operator Theory (2009), 1 -- 41, 10.1007/s11785-009-0044-y.
Daniel Mauldin and Mariusz Urbański, Graph directed Markov systems, Cambridge
Tracts in Mathematics, vol. 148, Cambridge University Press, Cambridge, 2003, Ge-
ometry and dynamics of limit sets. MR 2003772 (2006e:37036)
[MU03]
[Par60] William Parry, On the β-expansions of real numbers, Acta Math. Acad. Sci. Hungar.
[Rén57]
11 (1960), 401 -- 416. MR 0142719 (26 #288)
Alfréd Rényi, Representations for real numbers and their ergodic properties, Acta
Math. Acad. Sci. Hungar 8 (1957), 477 -- 493. MR 0097374 (20 #3843)
Fachbereich 3 - Mathematik und Informatik, Universität Bremen, Bibliothekstrasse
1, 28359 Bremen, Germany
E-mail address: [email protected], [email protected]
|
1502.05214 | 3 | 1502 | 2016-01-28T03:56:13 | Weak amenability of Fourier algebras and local synthesis of the anti-diagonal | [
"math.FA",
"math.OA"
] | We show that for a connected Lie group $G$, its Fourier algebra $A(G)$ is weakly amenable only if $G$ is abelian. Our main new idea is to show that weak amenability of $A(G)$ implies that the anti-diagonal, $\check{\Delta}_G=\{(g,g^{-1}):g\in G\}$, is a set of local synthesis for $A(G\times G)$. We then show that this cannot happen if $G$ is non-abelian. We conclude for a locally compact group $G$, that $A(G)$ can be weakly amenable only if it contains no closed connected non-abelian Lie subgroups. In particular, for a Lie group $G$, $A(G)$ is weakly amenable if and only if its connected component of the identity $G_e$ is abelian. | math.FA | math |
WEAK AMENABILITY OF FOURIER ALGEBRAS AND LOCAL
SYNTHESIS OF THE ANTI-DIAGONAL
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
Abstract. We show that for a connected Lie group G, its Fourier algebra
A(G) is weakly amenable only if G is abelian. Our main new idea is to show
that weak amenability of A(G) implies that the anti-diagonal, ∆G = {(g, g−1) :
g ∈ G}, is a set of local synthesis for A(G × G). We then show that this
cannot happen if G is non-abelian. We conclude for a locally compact group
G, that A(G) can be weakly amenable only if it contains no closed connected
non-abelian Lie subgroups. In particular, for a Lie group G, A(G) is weakly
amenable if and only if its connected component of the identity Ge is abelian.
0.1. Background. Questions on the nature of bounded derivations on (commuta-
tive) Banach algebras A have been around for a long time, in particular vanishing
of bounded Hochschild cohomologies H 1(A,M) for certain Banach A-modules M.
See, for example, [SiWe, Kam]. Johnson systematized many of these questions in
[Joh1].
In particular, he showed that for a locally compact group G, its group
algebra is amenable (i.e. H 1(L1(G),M∗) = {0} for each dual module M∗) if and
only if G is an amenable group. He also started the problem of determining when
H 1(L1(G), L1(G)∗) = {0}.
For a commutative Banach algebra A, Bade, Curtis and Dales ([BCD]) intro-
duced the concept of weak amenability, which is defined as having H 1(A,M) = {0}
for all symmetric Banach modules. They observed that this is equivalent to having
H 1(A,A∗) = {0}. There is an interesting universal module also exhibited by Runde
([Run]). The above observation of [BCD], leads us to refer to any Banach algebra
B as weakly amenable if H 1(B,B∗) = {0}. Weak amenability was established for
all L1(G) by Johnson ([Joh2]).
The Fourier algebras, A(G), as defined by Eymard ([Eym]), are dual objects
to the group algebras L1(G) in a sense which generalizes Pontryagin duality. It
was long expected that the amenability properties enjoyed by group algebras would
also extend to Fourier algebras. Hence it was a surprise when Johnson ([Joh3])
showed that A(G) fails to be weakly amenable for any compact simple Lie group,
in particular for G = SO(3). This motivated Ruan ([Rua]) to consider the operator
Date: July 24, 2018.
2000 Mathematics Subject Classification. Primary 43A30; Secondary 43A45, 43A80, 22E15,
22D35, 46H25, 46J40. Key words and phrases. Fourier algebra, weak amenability, local synthesis.
The first named author was supported by the Basic Science Research Program through the
National Research Foundation of Korea (NRF), grant NRF-2015R1A2A2A01006882. The second
named author was supported by Institut ´Elie Cartan de Lorraine. The third named author was
supported by NSERC Grant 366066-2014, and the Professeur Invit´e program at Institut ´Elie
Cartan de Lorraine. The fourth named was supported by NSERC Grant 312515-2010, and the
Korean Brain Pool Program.
1
2
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
space structure A(G) inherits by virtue of being the predual of a von Neuman
algebra. He proved that A(G) is operator amenable if and only if G is amenable.
Operator weak amenability for general A(G) was determined by Spronk ([Spr]) and,
independently, by Samei ([Sam1]). The question of amenability of A(G) was settled
by Forrest and Runde ([FoRu]):
it happens exactly when G is virtually abelian.
They also showed that if the connected component Ge is abelian, then A(G) is
weakly amenable. The following is suggested.
Question 0.1. If A(G) is weakly amenable, then must Ge be abelian?
Much progress has been made in answering this question. Building on work
of Plymen ([Ply]) -- which was written to answer a question in [Joh3] -- Forrest,
Samei and Spronk ([FSS1]) showed that A(G) is not weakly amenable whenever G
contains a non-abelian connected compact subgroup. Exciting recent progress was
made by Choi and Ghandehari. In [ChGh1] they show for the affine motion group,
and hence any simply connected semisimple Lie group, and also for the reduced
Heisenberg group, that the Fourier algebra is not weakly amenable. In [ChGh2]
they used completely different techniques to show the same for Heisenberg groups.
Our main theorem generalizes all of these results.
Let us briefly review the history of ideas around spectral synthesis. All concepts
below will be defined in Section 1.1. The study of sets of spectral synthesis, or,
for us, simply "synthesis", for A(G), especially for abelian G, has a long history.
See, for example, the historical notes in [HeRoII, §42]. Herz ([Her]) appears to
have been the first author to consider local synthesis for general G. There, he
proved this property is enjoyed by closed subgroups. Herz's work has inspired,
in part, the injection theorem of Lohou´e ([Loh]) and has motived aspects of the
work of Ludwig and Turowska ([LuTu1]). Thanks to the existence of a bounded
approximate identity consisting of compactly supported functions ([Lep]), sets of
local synthesis are sets of synthesis when G is amenable. Weak synthesis has its
origins in work of Varopoulos ([Var]) on spheres in Rn (n ≥ 3), was used in Kirsch
and Muller ([KiMu]), and was formalized by Warner ([War]). The first explicit
mention of what we call smooth synthesis is by Muller ([Mul]), which was applied
to certain manifolds in Rn. The first use of this concept for non-abelian G is due to
Ludwig and Turowska ([LuTu2]). Following their work, Park and Samei ([PaSa])
showed that for a connected Lie group G, the anti diagonal ∆G is a set of smooth
synthesis, and also of weak synthesis for A(G × G).
For compact G, building on work of Grønbaek ([Grø]), the article [Joh3] used the
failure of a weak form of spectral synthesis of the diagonal ∆G = {(g, g) : g ∈ G}
for the projective tensor product algebra A(G) ⊗A(G), to obtain the failure of weak
amenability. The local synthesis of ∆G for general locally compact G was used
by Samei ([Sam2]) to study a property which implies weak amenability of A(G).
Returning to compact groups, the ideas of [Joh3] were formalized and capitalized
upon in [FSS2]. These were used in [FSS1] to show that for compact G, A(G) is
weakly amenable exactly when the anti-diagonal ∆G = {(g, g−1) : g ∈ G} is a set of
synthesis for A(G × G). For groups containing open subgroups products of abelian
groups and compact groups, i.e. G ⊇ H ∼= A × K, this last result was extended to
local synthesis. We recall that A(G) ⊗A(G) 6= A(G × G), generally ([Los]). Hence
we do not expect obvious connections between sets of local synthesis for these two
algebras.
WEAK AMENABILITY OF A(G)
3
0.2. Structure. The starting point for the present investigation lies in the afore-
mentioned results of [PaSa]. Let G be a connected Lie group. We use the fact
that ∆G is simultaneously of weak and smooth synthesis for A(G × G), along with
the characterization of weak amenability of [Run], to show that for a connected
Lie group, weak amenability of A(G) implies that ∆G is a set of local synthesis for
A(G × G). The techniques rely intrinsically on the Lie theory, especially having
finite dimension for G. We see no way, at present, to extend them to arbitrary
connected locally compact groups.
In Section 1, we show for any connected Lie G that that weak amenability of
A(G) implies local synthesis of ∆G for A(G× G) -- a property we shall hereafter call
"local synthesis for G× G". Also in that section we discuss our two main functorial
properties satisfied by local synthesis of ∆G for G × G for locally comapct G:
the restriction to a closed connected Lie subgroup and an injection theorem with
quotients by discrete normal subgroups. In section 2 we give a criterion for testing
local synthesis of the anti-diagonal, and we show for five (classes of) low-dimensional
Lie groups that this criterion is satisfied.
In Section 3 we tie the investigation
together by noting that any non-abelian connected Lie group contains one of the
five aforementioned groups or its simply connected covering group, and use the
functorial properties to draw our conclusion for connected G. We can thus answer
Question 0.1 affirmatively for all Lie groups.
0.3. Basic notation. Let G be a locally compact group. The following spaces will
be used in this note: the space Cc(G) of compactly supported continuous functions;
and the Lp-spaces with respect the left Haar measure, Lp(G), p = 1, 2,∞.
We follow Eymard ([Eym]) for all definitions and concepts around the Fourier al-
gebra A(G). We recall that A(G) consists of all matrix coefficients u(g) = hλ(g)ξηi
where λ : G → U(L2(G)) is the left regular representation, and ξ, η ∈ L2(G). Fur-
thermore, the norm is given by kukA = inf{kξk2 kηk2 : u = hλ(·)ξηi as above}.
The bounded linear dual is given by the group von Neuman algebra V N (G) =
λ(G)′′ ⊂ B(L2(G)). If u = hλ(·)ξηi ∈ A(G) and T ∈ V N (G) we write T (u) =
hT ξηi. We shall denote the operator norm of an element S of V N (G) by kSkV N .
We recall that the maps u 7→ u (u(g) = u(g−1)) and maps of left and right transla-
tion are all isometric automorphisms of A(G). We always let A(G) ⊗A(G) denote
the projective tensor product. We shall have no need for completely bounded maps,
and will make no use of the operator projective tensor product, except, of course,
implicitly, when we discuss A(G × G).
We shall define more specialized notions in situ as their necessities arise.
1. Weak amenability and local synthesis for the anti-diagonal
1.1. Weak amenability implies local synthesis for the anti-diagonal. Let
G be a locally compact group. We let Ac(G) denote the subalgebra of elements u
of A(G) for which supp u = {g ∈ G : u(g) 6= 0} is compact. It is well known that
the Tauberian condition holds: Ac(G) is dense in A(G); and that A(G) is regular
on G: given compact K and a neighbourhood U of K, there is u in Ac(G) for which
uK = 1 and supp u ⊂ U . See [Eym, (3.38) and (3.2)].
4
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
Given a closed subset E of G we let
IG(E) = {u ∈ A(G) : uE = 0}
I 0
G(E) = {u ∈ Ac(G) : supp u ∩ E = ∅}, and
JG(E) = IG(E) ∩ Ac(G).
G(E) ⊆ JG(E) ⊆ IG(E). We say that E is a set of
It is evident that I 0
• spectral synthesis for G if I 0
• local synthesis for G if I 0
G(E) = IG(E); and
G(E) = JG(E).
For a linear subspace S ⊂ A(G) we let S(d) = span{ud : u ∈ S}. We say that E is
a set of
• weak synthesis for G if there is d in N for which IG(E)(d) ⊆ I 0
• local weak synthesis for G if there is d in N for which JG(E)(d) ⊆ I 0
We now let G be a connected Lie group and let D(G) denote the space of com-
If K is a compact subset of G we let
pactly supported smooth functions on G.
DK(G) = {u ∈ D(G) : supp u ⊂ K}, which is a Fr´echet space. For example, we fix
a basis β = (X1, . . . , Xd) for the Lie algebra g of G and for u, v in DK(G) set
G(E); and
G(E).
(1.1)
∂X u(g) =
, for X in g
ρK
β,n(u) = X1≤i1,...,in≤d(cid:13)(cid:13)∂Xi1 . . . ∂Xin u(cid:13)(cid:13)∞
and ρK
β (u, v) =
d
dt
u(g exp(tX))(cid:12)(cid:12)(cid:12)(cid:12)t=0
∞Xn=0
2n + ρK
ρK
β,n(u − v)
.
β,n(u − v)
Then given a compact neighbourhood K of the identity, D(G) = S∞
n=1 DK n (G)
is an inductive limit of Fr´echet spaces, a so called LF-space. See, for example,
[Tre]. It follows from [Eym, (3.26)] or [LuTu2, Lemma 3.3 and Remark 4.2], that
D(G) ⊂ Ac(G), and each inclusion DK(G) ֒→ A(G) is continuous. Furthermore,
since D(G) is dense in L2(G), D(G) ⊇ D(G) ∗ D(G) and {u : u ∈ D(G)} = D(G),
D(G) is dense in A(G).
For a closed subset E of G we let
J D
G (E) = D(G) ∩ IG(E)
where closure is in A(G). Since D(G) is dense in A(G), and D(G)J D
we have that J D
have inclusions I 0
G (E),
G (E) is an ideal in A(G). Since A(G) is regular and Tauberian we
G(E) ⊆ J D
G (E) ⊆ J D
G (E) ⊆ JG(E) ⊆ IG(E). We say that E is a set of
• smooth synthesis for G if J D
• local smooth synthesis for G if J D
G (E) = IG(E); and
G (E) = JG(E).
The projective tensor product E ⊗F of two locally convex spaces E and F is the
completion of the algebraic tensor product E ⊗F in the final topology with respect
to the embedding E ×F ֒→ E ⊗F . The standard proof of the following, for G = Rd,
uses techniques of Fourier analysis on Schwarz class functions. Hence we need to
take a little care to see that it holds in our setting, though the proof is standard.
Lemma 1.1. Let G be a connected Lie group and K and M be compact subsets of
G. Then DK(G) ⊗DM (G) ∼= DK×M (G × G) linearly and homeomorphically.
WEAK AMENABILITY OF A(G)
5
Proof. Let us first assume that there is a neighbourhood U of K for which there is a
diffeomorphism ϕ : U → U ′ ⊂ Rd. Then the map u 7→ u ◦ ϕ : DK (G) → Dϕ(K)(Rd)
is a linear homeomorphism. The same fact holds for M , with a neighbourhood V of
M and a diffeomorphism ψ : V → V ′ ⊂ Rn. Thus we get linear homeomorphisms
DK(G) ⊗DM (G) ∼= Dϕ(K)(Rd) ⊗Dψ(M)(Rd) ∼= Dϕ(K)×ψ(M)(R2d) ∼= DK×M (G × G)
where the middle identification is provided by [Tre, Theorem 51.6], and the last
one by the map w 7→ w ◦ (ϕ−1 × ψ−1).
i=1 of K and {Vj}n
j=1 of M , such that
i=1,{vj}n
each member is diffeomorphic to an open subset of Rn. Let {ui}m
j=1 in D(G)
be smooth partitions of unity, subordinate to the respective covers. Then we obtain,
(G),
Generally, there are finite open covers {Ui}m
whose inverse is mere inclusion. Then we obtain linear homeomorphisms
for example, a linear homeomorphism u 7→Pn
(G) ∼=
mMi=1
DM ∩Vj
DK∩Ui
as above. Hence we obtain linear homeomorphisms
i=1 uui : DK(G) →Pn
mMi=1
DK∩Ui×M ∩Vj
nMj=1
i=1 DK∩Ui
(G × G)
(G)! ⊗
nMj=1
nXj=1
mXi=1
nXj=1
mXi=1
∼=
DK(G) ⊗DM (G) ∼=
DK∩Ui
(G) ⊗DM ∩Vj
(G)
DK∩Ui×M ∩Vj
(G × G) ∼= DK×M (G × G)
by using both injectivity and projectivity of tensor product of these nuclear spaces.
(cid:3)
We let ∆G = {(g, g−1) : g ∈ G} ⊂ G × G.
Lemma 1.2. Let G be a connected Lie group. Then
(Ac(G) ⊗ Ac(G)) ∩ JG×G( ∆G)
is dense in JG×G( ∆G).
Proof. For simplicity, we let A = A(G × G). Let u ∈ JG×G( ∆G). It was shown in
[PaSa, Theorem 7] that ∆G is a set of local smooth synthesis, so we may assume
that u ∈ D(G × G). Hence we can find a compact subset M of G for which
u ∈ DM ×M (G × G). We can further assume that M is symmetric: M −1 = M .
Lemma 1.1 provides that u ∈ DM (G) ⊗DM (G).
Fix ε > 0. Let ρ be any invariant metric on DM ×M (G × G) which comes form
the Fr´echet structure. There is a δ1 > 0 for which
for w ∈ DM ×M (G × G) ⊂ A(G × G). Furthermore, it is straightforward to check
that the map
ρ(w, 0) < δ1 ⇒ kwkA < ε
Λ : DM ×M (G × G) → DM (G) given by Λw(g) = w(g, g−1)
is continuous since ∆G, qua submanifold of G × G, is diffeomorphic to G. If we fix
vM in Ac(G) for which vMM = 1, there is δ2 > 0 for which
ε
ρ(w, 0) < δ2 ⇒ kΛ(w)kA(G) <
.
kvMkA(G)
6
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
Thus, if δ = min{δ1, δ2}, then given u as in the paragraph above, there is a v in
the algebraic tensor product DM (G)⊗DM (G) for which ρ(u, v) < δ, and hence our
choice of δ entails that
ku − vkA < ε and kΛ(u − v)kA(G) <
ε
kvMkA(G)
.
Now let w = v− Λ(v)⊗ vM which is an element of Ac(G)⊗ Ac(G). Since M = M −1
G×G( ∆G),
it is easy to see that w ∆G = 0. Moreover, notice that Λ(u) = 0, as u ∈ J D
so
ku − wkA ≤ ku − vkA + kΛ(u − v) ⊗ vMkA < 2ε
since kΛ(u − v) ⊗ vMkA ≤ kΛ(u − v)kA(G) kvMkA(G).
(cid:3)
We let A(G)♯ denote the unitization of A(G) and m♯ : A(G)♯ ⊗A(G)♯ → A(G)♯
and m : A(G) ⊗A(G) → A(G) be the continuous linearization of the respective
multiplication maps. Since A(G) is Tauberian and regular, A(G)2 = A(G). Hence
It follows from [Grø, Theorem 3.2] that A(G) is weakly amenable if and only if
(1.2)
(ker m)2 = A(G) ⊗A(G) · ker m♯.
Above, and in what follows, we use for a subspace S of an algebra A, the notation
Sd = span{u1 . . . ud : u1, . . . , ud ∈ S}.
Theorem 1.3. Let G be a connected Lie group. If A(G) is weakly amenable, then
∆G is a set of local synthesis for A(G × G).
Proof. We let m : A(G) ⊗A(G) → A(G) be given on elementary tensors by m(u ⊗
v) = uv, and likewise define m♯. Since v 7→ v is an isometry on A(G), if A(G) is
weakly amenable then (1.2) implies that
(ker m)2 = A(G) ⊗A(G) · ker m♯.
(1.3)
For simplicity, we write J = JG×G( ∆G), below. Let u ∈ J. We wish to approximate
G×G( ∆G). We let ι : A(G) ⊗A(G) → A(G × G) be the linear
u by elements from I 0
contraction which embeds A(G) ⊗ A(G) into A(G × G). We have that
[Ac(G) ⊗ Ac(G)] ∩ J ⊂ ι(Ac(G) ⊗ Ac(G) · ker m♯) ⊆ ι(ker m)2 ∩ Ac(G × G) ⊆ J 2
where the first inclusion follows from regularity of A(G), the second inclusion is
provided by (1.3), and the third inclusion follows from regularity of A(G × G). By
Lemma 1.2, [Ac(G)⊗ Ac(G)]∩ J is dense in J. We thus conclude that J = J 2, and
induction shows that J = J m for any m in N.
The identity 4uv = (u + v)2 − (u − v)2 shows that J 2 = J (2), where the latter
notation was used in the definition of weak synthesis, above. By induction we see
that J 2m
= J (2m) for any m in N. We conclude that J = J (2m) for any m in N.
It is shown in [PaSa, Theorem 7], that ∆G is a set of local weak synthesis,
G×G( ∆G). Hence we see that for some m,
i.e. if n ≥ dim(G)/2, then J (n) ⊂ I 0
J = J (2m) = I 0
G×G( ∆G), and local synthesis is established.
(cid:3)
WEAK AMENABILITY OF A(G)
7
1.2. Functorial properties for local synthesis of the anti-diagonal. Our
ultimate goal is to show that a non-abelian connected Lie group does not allow
local synthesis of the anti-diagonal. The following will allow us to reduce our
calculations to certain computable cases.
We shall make use of the well-known result that for any locally compact group
G, and closed subgroup H, the restriction map RH : A(G) → A(H) is a quotient
map. See [Her], [McM, (4.21)], [Ars, (3.23)] or [DeDe]. We even have RH (Ac(G)) =
Ac(H).
Theorem 1.4. Let G be a locally compact group and H is a closed connected Lie
subgroup. If ∆G is of local synthesis for G × G, then ∆H is of local synthesis for
H × H.
Proof. We first claim that
RH×H (JG×G( ∆G)) = JH×H ( ∆H ).
i, v′
i in Ac(G) for which RH u′
It is clear that RH×H (JG×G( ∆G)) ⊆ JH×H ( ∆H ). To see the converse inclusion, let
u ∈ JH×H ( ∆H ). Since H is a connected Lie group, Lemma 1.2 allows us to assume
that u ∈ Ac(H) ⊗ Ac(H), hence u =Pn
i=1 ui ⊗ vi. The restriction theorem assures
that there are elements u′
i = vi for each i.
Hence if w in Ac(G) satisfies that wS = 1 where S =Sn
and, sincePn
i) ∈ IG×G( ∆G) ∩ Ac(G × G)
iv′
i − w ⊗ u′
i=1 ui(h)vi(h) = u(h−1, h) = 0 for h in H, we have
i=1 supp(ui), then
i = ui, RH v′
nXi=1
(u′
i ⊗ v′
u′ =
RH×H (u′) =
(ui ⊗ vi − RH (w) ⊗ uivi) = u.
nXi=1
It then follows that u ∈ RH×H (JG×G( ∆G)).
It is evident that
RH×H (I 0
G×G( ∆G)) ⊆ I 0
H×H ( ∆H ),
Our assumption that JG×G( ∆G) = I 0
paragraph, shows that I 0
H×H ( ∆H ) = JH×H ( ∆H ).
so RH×H(cid:16)I 0
G×G( ∆G)(cid:17) ⊆ I 0
H×H ( ∆H ).
G×G( ∆G), coupled with the result of the prior
(cid:3)
The proof of the next functorial property is more general, though its proof is a
little more involved. We shall use the following fact about local synthesis. This is
related to well-known facts about spectral synthesis in regular function algebras;
see, for example, expositions in [ReSt, Kan]. If G is a locally compact group and
E is a closed subset of G, we have for u in Ac(G) that
G(E) ⇔
u is "locally in I 0
bourhood Ug of g, and a ug in I 0
G(E)", i.e. for every g in E there is a neigh-
u ∈ I 0
G(E), for which uUg = ugUg .
Indeed, we need to only prove sufficiency.
In this case, take such a collection
U1 = Ug1 , . . . , Un = Ugn of open sets covering supp(u)∩E, and associated functions
k=1 Uk with Un+1∩ E = ∅.
Any partition of unity v1, . . . , vn+1 of supp(u) subordinate to U1, . . . , Un+1, satisfies
u1, . . . , un. Let Un+1 be a neighbourhood of supp(u)\Sn
uvk = ukvk, for each k = 1, . . . , n, so u =Pn
k=1 ukvk + uvn+1 ∈ I 0
G(E).
8
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
The following holds for any regular Banach function algebra. To avoid introduc-
ing new notation, we state it only for a Fourier algebra.
Proposition 1.5. Let G be a locally compact group and E be a closed subset which
admits a partition, E = Fi∈I Ei, for which each Ei is closed in G and relatively
open in E. Then E is of local synthesis for G if and only if each Ei is of local
synthesis.
Proof. (⇒) Let F be a subset of E which is closed in G and relatively open in E; in
particular we may consider F = Ei for a fixed i. Let u ∈ JG(F )∩ Ac(G). For every
g in F , our assumptions on F allows us to choose a compact neighbourhood U of g
for which U ∩ E = U ∩ F . We find v in Ac(G) for which supp(v)∩ E = supp(v)∩ F
and vU = 1. We have that u = uv on U and
uv ∈ JG(E) = I 0
G(E) ⊆ I 0
G(F ).
Thus u is locally in I 0
G(F ), whence in JG(F ).
(⇐) Let u ∈ JG(E) ∩ Ac(G). Since each Ei is of local synthesis, we have
u ∈ JG(E) ⊆ JG(Ei) = I 0
G(Ei)
G(E), thus u ∈ I 0
G(E). Hence u is locally in I 0
i.e. u = limn→∞ un, where each un ∈ I 0
G(Ei). Fix g in Ei and, again, our as-
sumption of relative openness of Ei in E allows us to take a compact neighbour-
hood U of g for which U ∩ E = U ∩ Ei. If v is any element of Ac(G) for which
supp(v)∩ E = supp(v)∩ Ei and vU = 1, then u = uv on U , and uv = limn→∞ unv,
where each unv ∈ I 0
Theorem 1.6. Let G be a locally compact group and Γ a discrete normal subgroup.
Then ∆G is of local synthesis for G × G if and only if ∆G/Γ is of local synthesis
for G/Γ × G/Γ.
Proof. The injection theorem of [Loh] tells us that ∆G/Γ is of local synthesis for
G/Γ× G/Γ if and only if q−1( ∆G/Γ), where q : G× G → G/Γ× G/Γ is the quotient
map, is a set of local synthesis for G × G. Hence we wish to verify that
∆G is of local synthesis ⇔ q−1( ∆G/Γ) is of local synthesis
(1.4)
each for G × G. Hence we wish to examine the structure of the latter set.
We first observe that since gΓ = Γg and (Γg)−1 = g−1Γ in G/Γ, we have
G(E).
(cid:3)
q−1( ∆G/Γ) = {(γg, g−1γ′−1) : g ∈ G, γ, γ′ ∈ Γ}.
Hence if we define an action of Γ × Γ on G × G by (γ, γ′) · (g, g′) = (γg, g′γ′−1),
then q−1( ∆G/Γ) = (Γ × Γ) · ∆G, the orbit of the set ∆G under this action. It is
easy to check that
i.e. if (γg, g−1γ′−1) = (h, h−1) then γg = (g−1γ′−1)−1; and hence
(γ, γ′) · ∆G = ∆G ⇔ γ′ = γ
Let us see that
(γ, γ′) · ∆G = (λ, λ′) · ∆G ⇔ λ−1γ = λ′−1γ′ ⇔ λ′λ−1 = γ′γ−1.
Thus q−1( ∆G/Γ) = F(γ,γ ′)∆Γ∈(Γ×Γ)/∆Γ(γ, γ′) · ∆G, where ∆Γ = {(γ, γ) : γ ∈ Γ}.
the individual fibres (λ, λ′) · ∆G are relatively open in q−1( ∆G/Γ).
(1.5)
Suppose that (λh, h−1λ′−1) is approached by a net of elements (γngn, g−1
n γ′
n
Then the net of elements (λ−1γngnh−1, hg−1
−1).
−1λ′) would approach (e, e), which
n γ′
n
WEAK AMENABILITY OF A(G)
9
implies that λ−1γnγ′
n
thus γnγ′
n
−1λ′ approaches e, and hence is ultimately e, by discreteness;
−1 is ultimately λλ′−1, i.e. the net is ultimately in the fibre (λ, λ′) · ∆G.
(⇒) If ∆G is of local synthesis for G × G, then so too is each fibre (γ, γ′)· ∆G =
(γ, e) ∆G(e, γ′−1). We then appeal to Proposition 1.5 and (1.5) to obtain necessity
in (1.4).
(⇐) The anti-diagonal ∆G is an open fibre of q−1( ∆G/Γ), thanks to (1.5). Then
Proposition 1.5 delivers sufficiency in (1.4).
(cid:3)
In practice we will use this last result with a central discrete subgroup. Centrality
seems to offer no meaningful simplification to the proof, however.
2. Failure of local synthesis for the anti-diagonal
2.1. The basic strategy. We wish to show that for certain given 2 or 3-dimensional
non-abelian connected Lie groups G, that its anti-diagonal fails to be a set of local
synthesis for G×G . We shall exploit the density of the space of test functions D(G)
in A(G), as outlined in Section 1.1. This implies that any T in V N (G) ∼= A(G)∗
may be understood as a distribution, i.e. T ∈ D(G)∗.
The following informs all of the choices made through the rest of this section.
Lemma 2.1. Let G be a connected Lie group with Lie algebra g and X ∈ g with
X 6= 0, hence (X, 0) is an element of the Lie algebra g × g of G × G. Let 0 6= v ∈
L1(G). If there exists S = SX,v in V N (G × G) for which
∂(X,0)u(g, g−1)v(g) dg
S(u) =ZG
Let us see that S ∈ I 0
G×G( ∆G)⊥. Let u ∈ I 0
G×G( ∆G) ( JG×G( ∆G).
whenever u ∈ D(G × G), then ∆G is not a set of local synthesis for G × G.
Proof. We will verify that S ∈ I 0
annihilator of the subspace K. This will show that I 0
G×G( ∆G)⊥ \ JG×G( ∆G)⊥, where K ⊥ denotes the
G×G( ∆G), and ε > 0. By the
regularity of the algebra D(G × G), there is w in D(G × G) such that wsupp(u) = 1
and supp(w) ∩ ∆G = ∅. Since I 0
G×G( ∆G), there is u′ in D(G × G) ∩
IG×G( ∆G) such that ku − u′kA < ε/ kwkA. But then, since u = wu, we have that
G×G( ∆G), it is obvious that ∂(X,0)(wu′) ∆G = 0, so
ku − wu′kA < ε. As wu′ ∈ I 0
S(wu′) = 0. Hence S(u) = S(u − wu′) ≤ kSkV N ε. As ε > 0 may be chosen
arbitrarily, S(u) = 0.
Let us now see that S 6∈ JG×G( ∆G)⊥. We consider general x, y, z in D(G), and
let w = xy ⊗ z − x ⊗ yz, which is an element of D(G × G) ∩ JG×G( ∆G). But then
S(w) =ZG
x(g)∂X y(g)z(g−1)v(g) dg.
[∂X (xy)(g)−∂X x(g)y(g)]z(g−1)v(g) dg = −ZG
G×G( ∆G) ⊆ J D
We may choose x, y, z for which S(w) 6= 0.
(cid:3)
We shall require disintegration of the left regular representation of G into irre-
ducible components. For this purpose and to introduce notation, we summarize
the Plancherel theorem of [Tat]. Our presentation is influenced by [Fol, (7.50)]. We
purposely restrict the description, to fit our needs. We let bG denote the space of
(equivalence classes of) irreducible representations of G, and accept the standard
10
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
abuse of notation where we conflate an equivalence class with one of its represen-
tatives. If u ∈ Cc(G) we let its Fourier transform be given at π in bG by
u(g)π(g) dg ∈ B(Hπ), i.e. hu(π)ξηi =ZG
u(π) =ZG
u(g)hπ(g)ξηi dg
where ξ, η ∈ Hπ, and Hπ is the space on which π acts.
Proposition 2.2. Suppose G is a connected Lie group for which the kernel of the
modular function, K = ker ∆, is type I, and G acts on bK regularly in the sense that
there is a Borel cross-section for the space of orbits of G on bK. Then bG contains
• a dense Borel subset S(bG) of elements, each of which is the induced repre-
• a Borel parametrization b 7→ πb : B → S(bG) and a Borel measure µ on B;
• for each b in B, a positive operator δb on Hπb, which satisfies
sentation from some closed subgroup of ker ∆;
and
πb(g)δbπb(g−1) =
1
δb
∆(g)
for which the choice of the triple ((δb)b∈B, (πb)b∈B, µ) is unique, up to measure
equivalence of ((πb)b∈B, µ), and such that the Plancherel transform on Cc(G), given
by
extends to a unitary identifying
b
)b∈B
u 7→ (bu(πb)δ1/2
B Hπb ⊗2 Hπb dµ(b).
L2(G) ∼=Z ⊕
Each Hilbertian tensor product Hπb ⊗2 Hπb is identified with the space of Hilbert-
Schmidt operators on Hπb , above.
Hence the left regular representation λ of G admits disintegration up to unitary
equivalence, and quasi-equivalence respectively, as
(2.1)
λ ∼=Z ⊕
B
πb ⊗ IHπb
dµ(b)
and λ ≃Z ⊕
B
πb dµ(b).
In (2.1) we need to concern ourselves with only the equivalence class of µ in the
relation of mutual absolute continuity of measures on B. Furthermore, the left
regular representation of G × G may now be represented by the quasi-equivalence
(2.2)
πb × πb′ d(µ × µ)(b, b′)
λ × λ ≃Z ⊕
B×B
where we use Kroenecker products of representations. Hence if v ∈ L1(G), then the
the operator field
operator on A(G× G) given by Ev(u) =RG u(g, g−1)v(g) dg may be represented by
(2.3)
Ev(b, b′) =ZG
v(g)πb(g) ⊗ πb′ (g−1) dg
for b, b′ in B.
If G is unimodular, we set each δb = IHπb
. When G is not unimodular, and πb
is induced from a character χ of abelian subgroup H of ker ∆, then Hπb may be
identified with a completion of Fπb = {f ∈ C(G) : f (gh) = χ(h)f (g)}, and δb with
multiplication on the latter space by ∆(·)−1; compare with the description in [Fol,
(7.49)].
WEAK AMENABILITY OF A(G)
11
We also have for any g in G, the Fourier inversion formula of [Tat, Corollary 2]:
(2.5)
(2.6)
Tr(πb(g)bw(πb)δb) dµ(b).
By density of span λ(G) in V N (G), and of each span πb(G) in B(Hπb) (Schur's
lemma), we also have for any T in V N (G) the duality formula
for w = hλ(·)uvi, where u, v ∈ Cc(G), we have
(2.4)
w(g) =ZB
Tr(πb(g−1) w(πb)δb) dµ(b) =ZB
T (w) =ZB
Tr(T (b)bw(πb)δb) dµ(b)
where T ≃ (T (b))b∈B ∈ L∞(B, µ;B(Hπb)).
Let G be a connected Lie group and π ∈ bG. It is well known that for any X in
the Lie algebra g of G, and any π in bG, there is a dense subspace HX
π of vectors ξ
for which
dπ(X)ξ = lim
h→0
[π(exp(hX)) − I]ξ
1
h
exists. For example HX
π = span{u(π)ξ : u ∈ D(G), ξ ∈ Hπ}. Thus dπ(X) is
an (unbounded) operator on Hπ. Given the parameterization b 7→ πb above, it can
be checked that, as the limit of a measurable field of operators, (dπb(X))b∈B is also
measurable.
π ⊇ HD
Lemma 2.3. Suppose that G is a connected Lie group with Lie algebra g, and X
in g is such that, for each b in B, there is a subspace Fb of HX
πb which is dense
in Hπb , and for which δ−1
πb . If T in V N (G) satisfies that (S(b))b∈B =
(T (b)dπb(X))b∈B is a bounded field of operators, then for u in D(G) we have that
b Fb ⊆ HX
S(u) = T (∂X u).
∂X u(g−1)πb(g)δbξ dg
b Fb then we have for u in D(G) that
Proof. Recall that as in (1.1), the symbol ∂X denotes a derivative on the right.
First we fix π. If ξ ∈ δ−1
\(∂X u)∨(πb)δbξ =ZG
h(cid:20)ZG
u(g−1 exp(hX))πb(g)δbξ dg −ZG
h(cid:20)ZG
u(g−1)πb(exp(hX)g)δbξ dg −ZG
[πb(exp(hX) − I](cid:19)ZG
u(g−1)πb(g)δbξ dg(cid:21)
u(g−1)πb(g)δbξ dg(cid:21)
u(g−1)πb(g)δb dgξ
= lim
h→∞
= lim
h→∞
1
1
1
h
h→∞
=(cid:18) lim
= dπb(X)bu(πb)δbξ
where in the second through fourth lines, the limit is understood in the weak sense,
and we may use dominated convergence theorem on associated scalar integrals.
Since δ−1
b Fb is dense in Hπb , the computation above, coupled with (2.5), tells us
that
S(u) =ZB
=ZB
Tr(T (b)dπb(X)bu(πb)δb) dµ(b)
Tr(T (b) \(∂X u)∨(πb)δb) dµ(b) = T (∂Xu).
(cid:3)
12
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
We now will embark on using Lemma 2.3 on the group G× G and operator fields
with fibres d(πb × πb′ )(X, 0) = dπb(X) ⊗ IHb′ (X ∈ g) to verify the conditions of
Lemma 2.1.
As a first illustration, let us apply these methods to the special unitary group
SU(2). We note that ∆SU(2) is a set of non-synthesis is known (see [FSS1]). We
recall the well known fact that SU(2) admits as its Lie algebra su(2), which may
be identified with 2× 2-complex matrices which are skew-Hermitian and trace zero.
each πn acts on a space of dimension n + 1, and if g ∼= diag(z, ¯z) (∼= is similarity by
conjugation in SU(2)) then πn(g) ∼= diag(zn, zn−2, . . . , z−n).
Proposition 2.4. Let X be any non-zero element of su(2). Then there is an
S = SX,1 in V N (SU(2)) for which
The representation theory of SU(2) is well-known: cSU(2) = {πn : n = 0, 1, 2, . . .},
S(u) =ZG
∂(X,0)u(g, g−1) dg
for u in D(SU(2) × SU(2)).
Proof. Since X ∗ = −X and TrX = 0, there is x in R for which we have equivalence
X ∼= diag(ix,−ix), hence exp(hX) ∼= diag(eihx, e−ihx) in SU(2). It is immediate
that
dπn(X) ∼= diag(inx, i(n − 2)x, . . . ,−inx).
The Schur orthogonality relations immediately give, in the notation of (2.3) with
v = 1, that
E1(n, n′) =ZSU(2)
πn(g) ⊗ πn′ (g)∗ dg =( 1
n+1 Tn
0
if n = n′
otherwise
where Tn is an (n + 1)2 × (n + 1)2 permutation matrix with respect to some basis.
It is then obvious that
for each n, n′. We appeal to Lemma 2.3.
kE1(n, n′)(dπn(X) ⊗ I)k ≤ x
(cid:3)
2.2. Two unimodular groups. We let E = C⋊T be the 3-dimensional Euclidean
motion group with multiplication and inversion given by
(x, z)(x′, z′) = (x + zx′, zz′)
and (x, z)−1 = (−¯zx, ¯z).
This group is unimodular with Haar integral the same as that on C × T.
The following data are well known; see, for example [Sug, IV]. The additive
group C = R2 admits a real inner product (x, x′) 7→ x· x′ = Re x Re x′ + Im x Im x′.
For each a in C\{0} we obtain an irreducible unitary representation πa by inducing
from the character χa (χa(x) = e−ix·a). Then πa ∼= πb if and only if a = b. Hence
we parameterize this family by r in (0,∞). Each representation is given
πr(x, z)ξ(w) = e−ix·(rw)ξ(¯zw).
Then the disintegration formula (2.1) takes the form
πr : E → U(L2(T)),
λ ≃Z ⊕
(0,∞)
πr r dr ≃Z ⊕
(0,∞)
πr dr
where the middle formula is with respect to the group's Plancherel measure which
is mutually absolutely equivalent to Lebesgue measure on (0,∞).
WEAK AMENABILITY OF A(G)
13
We recall that E has Lie algebra
e = hT, X1, X2 : [T, X1] = X2, [T, X2] = −X1, [X1, X2] = 0i
where exp(hX1) = (h, 1), exp(hX2) = (ih, 1) and exp(hT ) = (0, eih).
We also consider the Sobolev-type space
H 2,1(R2) = {v : R2 → C v, ∂2
xv ∈ L1(R2), for all x ∈ R2}
dt v(y + tx)(cid:12)(cid:12)t=0, and the second order derivatives may be consid-
where ∂xv(y) = d
ered in the distributional sense.
Theorem 2.5. Let v ∈ H 2,1(C) = H 2,1(R2), and v in L1(E) be given by v(x, z) =
v(x). Let X ∈ span{X1, X2}. Then there is an S = SX,v in V N (E) for which
S(u) =ZE
∂(X,0)u(g, g−1)v(g) dg
for u in D(E × E).
Proof. First, for ξ in L2(T), (2.3) provides that
Ev(r, r′)ξ(w, w′) =ZTZC
=ZT
v(x)e−ix·(rw−r′zw′)ξ(¯zw, zw′) dx dz
2πbv(rw − r′zw′)ξ(¯zw, zw′) dz
U ∗Ev(r, r′)U ξ(w, w′) = Ev(r, r′)U ξ(w, ¯ww′)
L2(T2) given by U ξ(w, w′) = ξ(w, ww′), which has adjoint given by U ∗ξ(w, w′) =
ξ(w, ¯ww′). We have
where bv is the Fourier transform of v on C = R2. Consider the unitary U on
= 2πZTbv(rw − r′z ¯ww′)ξ(¯zw, w′) dz = 2πZTbv(rw − r′zw′)ξ(¯z, w′) dz
The fact that we choose v from H 2,1(C) allows that there is a constant Cv, such
that for any y in C we have
Cv
1 + y2 .
bv(y) ≤
In fact we could choose Cv = kbvk∞ +kcLvk∞, where L = ∂2
x1 + ∂2
x2 is the Laplacian
on R2 ∼= C. Hence we estimate
kU ∗Ev(r, r′)U ξ(w, w′)k2
≤ (2πCv)2ZTZT(cid:20)ZT
≤ (2πCv)2(cid:20)ZT
dz
1 + r − r′z2(cid:21)2
kξk2
2
2 ≤ZT2(cid:20)2πCvZT
1 + r − r′zw′ ¯w2(cid:21)(cid:20)ZT
dz
ξ(¯z, w′)
1 + rw − r′zw′2 dz(cid:21)2
d(w, w′)
ξ(¯z, w′)2 dz
1 + r − r′w′z ¯w2(cid:21) dw dw′
where we have used the Cauchy-Schwarz inequality in the second line, and Tonelli's
theorem, in the third. Hence
(2.7)
kEv(r, r′)k = kU ∗Ev(r, r′)Uk ≤ZT
2πCv dz
1 + r − r′z2 for r, r′ > 0.
14
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
Now we consider for each r > 0 the operator dπr(X).
(x, 1) = exp X we have
If ξ ∈ L2(T) then for
dπr(X)ξ(w) = lim
h→0
1
h
[e−i(hx)·(rw) − 1]ξ(w) = −irx · wξ(w)
where convergence is uniform in w. Hence by (2.6) dπr(X) is the multiplication
operator by w 7→ −irx· w. In particular dπr(X) is bounded with kdπr(X)k ≤ rx.
Combining with (2.7) we see that
kEv(r, r′)(dπr(X) ⊗ I)k ≤ZT
2πCvr dz
1 + r − r′z2 =Z 2π
0
Cvr dt
1 + r2 + r′2 − 2rr′ cos t
.
However, using either methods of complex analysis, or the table of integrals [GrRy,
2.553-3], we obtain that the latter integral is equal to the first expression in the
elementary estimate
Cvr
Cvr
2q(1 + r2 + r′2)2 − 4(rr′)2 ≤
2q1 + 2(r2 + r′2)
which is clearly uniformly bounded in r and r′. Hence, by Lemma 2.3, we are
done.
(cid:3)
, x + x′)
Now we consider the reduced Heisenberg group Hr = (R × T) ⋊ R with multipli-
cation and inversion given by
(y, z, x)(y′, z′, x′) = (y + y′, zz′eixy ′
and (y, z, x)−1 = (−y, ¯zeixy,−x).
We identify the centre of Hr with T. The group is unimodular, and its Haar
integral is the same as that on the product group R × T × R. All of the infinite-
dimensional irreducible representations are known to be obtained by inducing from
the characters χ0,n on the normal subgroup R× T, χ0,n(y, z) = ¯zn, for n ∈ Z\ {0}.
This follows, for example, from [KaTa, 4.38] and the fact that Hr is a quotient of
the usual Heisenberg group. For each n in Z \ {0}, the representation is given by
πn : Hr → L2(R),
The left regular representation admits a decomposition
πn(y, z, x)ξ(t) = ¯zne−intyξ(t − x).
λ ≃Mn∈Z
πn
(2.8)
where π0 : Hr → L2(R2) is, effectively, the left regular representation of Hr/T.
Indeed, if we let Hn = {ξ ∈ L2(Hr) : ξ(g(0, z, 0)) = ¯znξ(g) for a.e. g in Hr}, then
λn = λ(·)Hn has λn(0, z, 0) = ¯znI, and is hence quasi-equivalent to πn by the
Stone-von Neumann theorem (see, for example, [Fol, (6.49)]).
We note that the Lie algebra of Hr is given by
h = hX, Y, Z : [X, Y ] = Z, [Z, X] = 0 = [Y, Z]i
where exp(hX) = (0, 1, h), exp(hY ) = (h, 1, 0) and exp(hZ) = (0, eih, 0).
Theorem 2.6. Let v1, v2 in L1(R) be so that v1 is (essentially) bounded and v2 ∈
L1(R), and set v(y, z, x) = v1(x)v2(y). Then there is S = SZ,v in V N (Hr) for
which
S(u) =ZHr
∂Z u(g, g−1)v(g) dg
for u ∈ D(Hr × Hr).
WEAK AMENABILITY OF A(G)
15
Proof. We allow (2.8) to substitute for (2.1) and we obtain a likewise decomposition
for λ × λ. Thus, using the appropriate analogue of (2.3) we compute for n, n′ in
Z \ {0} that for ξ in L2(R2) ∼= L2(R) ⊗2 L2(R) that
Ev(n, n′)ξ(t, t′) =ZRZTZR
v1(x)v2(y)¯zne−intyzeixy
= δn,n′ZRZR
= √2πδn,n′ZR
v1(x)v2(y)e−in(x+t−t′)yξ(t − x, t′ + x) dy dx
v1(x)v2(n(x + t − t′))ξ(t − x, t′ + x) dx
ein′t′yξ(t − x, t′ + x) dy dz dx
where δn,n′ is the Kroenecker delta symbol. Likewise Ev(n, 0) = 0 for n in Z \ {0}.
We consider the unitary on L2(R) given by U ξ(t, t′) = ξ(t, t′ + t). We have for n in
Z \ {0} that
n′
U ∗Ev(n, n)U ξ(t, t′) = Ev(n, n)U ξ(t, t′ − t)
Hence we compute
kU ∗Ev(n, n)U ξk2
v1(x)v2(n(x + 2t − t′))ξ(t − x, t′) dx
v1(t − x)v2(n(−x + 3t − t′))ξ(x, t′) dx
= √2πZR
= √2πZR
2π(cid:20)ZR v1(t − x)v2(n(−x + 3t − t′))ξ(x, t′) dx(cid:21)2
×(cid:20)ZR v1(t − x)v2(n(−x + 3t − t′))ξ(x, t′)2 dx(cid:21) dt dt′
∞ kv2k2
2 ≤ZRZR
≤ 2πZRZR(cid:20)ZR v1(t − x)v2(n(−x + 3t − t′)) dx(cid:21)
dt dt′
≤ 2π kv1k2
1
3n2 kξk2
2
where we have used the Cauchy-Schwarz inequality for the second inequality and
Tonelli's theorem, a Holder inequality and a change of variables for the third. In
summary
(2.9)
kEv(n, n′)k ≤ δn,n′
It is trivial to see that
√2π
√3n kv1k∞ kv2k1 , for (n, n′) in (Z \ {0}) × Z.
Combining with (2.9) we see that
dπn(Z) = −inI for n in Z.
sup
(n,n′)∈Z2 kE(n, n′)(dπn(Z) ⊗ I)k < ∞
and we may hence appeal to Lemma 2.3.
(cid:3)
2.3. Two non-unimodular groups. We shall consider a class of groups we call
Gr´elaud's groups. Fix a parameter θ > 0 and for s in R let
τ (s) = exp s(cid:20)1 −θ
1 (cid:21) = es(s), where (s) =(cid:20)cos sθ − sin sθ
cos sθ (cid:21) .
sin sθ
θ
16
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
We now let Gθ = R2 ⋊τ R with multiplication given by
(x, s)(x′, s′) = (x + τ (s)x′, s + s′), hence (x, s)−1 = (−τ (−s)x,−s).
Notice that we have det τ (s) = e2s from which we get left Haar integral, for u in
Cc(Gθ), and modular function
u(x, s)e−2s ds dx and ∆(x, s) = e2s.
ZG
u(x, s) d(x, s) =ZR2ZR
πy : Gθ → U(L2(R)),
If y ∈ R2, with associated character χy(x) = e−ix·y on R2, we get induced repre-
sentation
πy(x, s) = e−i(τ (−t)x)·yξ(t − s) = e−ie−tx·((t)y)ξ(t − s).
Notice that if y′ = (t′)y, then πy ∼= πy ′ via the intertwiner ρ(t′), where ρ is the left
regular representation on L2(R). Hence we may parameterize these representations
by the unit sphere S1. Furthermore, by [KaTa, 7.35], the family {πy}y∈S1 is a dense,
To learn the disintegration of the left regular representation (2.1), we will have
to obtain the Plancherel formula for this group. Since we do not know of a reference
for this, we compute it ourselves. We use the notation of Proposition 2.2.
compact subset of bGθ.
Proposition 2.7. The Plancherel decomposition of L2(Gθ) is given by
L2(Gθ) ∼=Z ⊕
S1
L2(R)y ⊗2 L2(R)y dν(y)
where each L2(R)y is a copy of L2(R), ν is the unique rotationally invariant prob-
ability measure on S1 and we have for ξ in L2(R)y, δyξ(t) = ∆(t)−1ξ(t) = e−2tξ(t)
for a.e. t. Hence we obtain
λ ≃Z ⊕
S1
πy dν(y)
Proof. We will simply verify that the choices above satisfy Proposition 2.2. Let for
y in S1 and u in Cc(Gθ) and ξ in L2(R)
u(πy)δ1/2
y
u(g)πy(g)(δ1/2
y
ξ)(t) dg
ξ(t) =ZG
=ZRZR2
=ZR
u(x, s)e−ie−tx·((t)y)es−tξ(t − s) dx ds
2πu1(e−t(t)y, t − s)e−sξ(s) ds
where u1 is the partial Fourier transform in the R2-variable. We note the well-
known fact that the Hilbert-Schmidt norm ku(πy)k2 is given in terms of the kernel
function as(cid:0)RRRR 2πu1(e−t(t)y, s + t)2 e−2sds dt(cid:1)1/2
ZS1 ku(πy)δyk2
2 dν(y) =ZS1ZRZR 2πu1(e−t(t)y, t − s)2 e−2sds dt dν(y)
=ZS1ZRZR 2πu1(e−ty, s)2 e2(t−s)ds dt dν(y)
=ZR2ZR 2πu1(z, s)2 e−2sds dz =ZRZR2 u(x, s)2 dx e−2sds = kuk2
. We then see that
2 .
WEAK AMENABILITY OF A(G)
17
where we have used the invariance of the chosen measures on S1 and R, then an
obvious change of variables in R2, and finally the Plancherel formula in R2.
(cid:3)
We note that Gθ admits the Lie algebra
gθ = hT, X1, X2 : [T, X1] = X1 − θX2, [T, X2] = θX1 + X2, [X1, X2] = 0i
where exp(hXj) = (hej, 0) for j = 1, 2 ((e1, e2) is the standard basis for R2) and
exp(hT ) = (0, h).
Theorem 2.8. Let v1 in L1(R2) be so that v1 ∈ Ac(R2), let v2 ∈ Cc(R), and then
let v(x, s) = v1(x)v2(s). Let X ∈ span{X1, X2}. Then there is an S = SX,v in
V N (Gθ) for which
S(u) =ZGθ
∂(X,0)u(g, g−1)v(g) dg
for u in D(Gθ × Gθ).
Proof. Our assumptions on v1 and v2 ensure that v is integrable. We use (2.3) to
compute for ξ in L2(R2) ∼= L2(R) ⊗2 L2(R) that
Ev(y, y′)ξ(t, t′)
v1(x)v2(s)e−i(τ (−t)x)·y+(−τ (−t′−s)x)·y ′
ξ(t − s, t′ + s) dx e−2sds
2πv1(e−t(t)y − e−t′−s(t′ + s)y′)v2(s)e−2sξ(t − s, t′ + s) ds.
=ZRZR2
=ZR
= 2πZR
= 2πZR
Letting, now, U ξ(t, t′) = ξ(t, t′ + t) we compute
U ∗Ev(y, y′)U ξ(t, t′) = Ev(y, y′)U ξ(t, t′ − t)
(2.10)
(2.11)
v1(e−t(t)y − e−t′+t−s(t′ − t + s)y′)v2(s)e−2sξ(t − s, t′) ds
v1(e−t(t)y − e−t′+s(t′ + s)y′)v2(t − s)e2(s−t)ξ(s, t′) ds
Let us now compute dπy(X). Let (x, 0) = exp X, and for ξ ∈ Cc(R) we have
[e−ie−t(hx)·((t)y) − 1]ξ(t) = −ie−tx · ((t)y)ξ(t)
dπy(X)ξ(t) = lim
h→0
1
h
where convergence is uniform on compact sets. Hence dπy(X) admits Fy = Cc(R) in
its domain, and by (2.6), and is given by pointwise multiplication by t 7→ −ie−tx ·
((t)y). Notice that δ−1
y Fy ⊂ Fy. Hence we may appeal to Lemma 2.3 and it
suffices to see that the field of operators
(Ev(y, y′)(dπy(X) ⊗ I))y,y ′∈S1 is bounded
(2.12)
to gain our conclusion. It is clear that U ∗(dπy(X) ⊗ I)U = dπy(X) ⊗ I, and it
suffices to work with the field (U ∗Ev(y, y′)U (dπy(X) ⊗ I))y,y ′∈S1 . We combine
(2.10) and (2.11) to see that for ξ in L2(R2) we have
U ∗Ev(y, y′)U (dπy(X) ⊗ I)ξ(t, t′)
= 2πZR
= −2πiZR
v1(e−t(t)y − e−t′+s(t′ + s)y′)v2(t − s)e2(s−t)(dπy(X) ⊗ I)ξ(s, t′) ds
v1(e−t(t)y − e−t′+s(t′ + s)y′)v2(t − s)es−2t(x · ((s)y))ξ(s, t′) ds.
18
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
Our assumptions on v1 allow us to find a non-increasing ϕ in Cc([0,∞)) for which
Let us observe that e−t(t)y − e−t′+s(t′ + s)y′ ≥ e−t − e−t′+s. Now with an
application of Cauchy-Schwarz inequality we obtain
2πv1(y) ≤ ϕ(y).
2
kU ∗Ev(y, y′)U (dπy(X) ⊗ I)ξk2
≤ x2ZR2(cid:20)ZR
≤ x2ZR2(cid:20)ZR
×(cid:20)ZR
ϕ(e−t − e−t′+s)v2(t − s)es−2tξ(s, t′) ds(cid:21)2
ϕ(e−t − e−t′+s)e−tv2(t − s)es−t ds(cid:21)
ϕ(e−t − e−t′+s)e−tv2(t − s)es−tξ(s, t′)2 ds(cid:21) d(t, t′).
d(t, t′)
Since e−t ≤ e−t − e−t′+s + e−t′+s we have
ZR
ϕ(e−t − e−t′+s)e−tv2(t − s)es−t ds
≤ZR
ϕ(e−t − e−t′+s)e−t − e−t′+sv2(t − s)es−t ds
+ZR
ϕ(e−t − e−t′+s)e−t′+sv2(t − s)es−t ds
≤ kϕιk∞ kδ−1/2v2k1 + kϕk1 kδ−1/2v2k∞
where ι(r) = r for r ≥ 0 and δv2(s) = e2sv2(s). It then follows that
kU ∗Ev(y, y′)U (dπy(X) ⊗ I)ξk2
2
≤ x2 (kϕιk∞ kδv2k1 + kϕk1 kδv2k∞)
×ZRZR(cid:20)ZR
ϕ(e−t − e−t′+s)e−tv2(t − s)es−t dt(cid:21) ξ(s, t′)2 ds dt′
≤ x2(cid:16)kϕιk∞ kδ−1/2v2k1 + kϕk1 kδ−1/2v2k∞(cid:17)kϕk1 kδ−1/2v2k∞ kξk2
2 .
Hence we have verified (2.12).
(cid:3)
We now consider the real affine motion group, also known as the "ax + b group",
F = R ⋊ R with product and inverse given by
The group has left Haar integral and modular function given by
(b, a)(b′, a′) = (b + eab′, a + a′) and (b, a)−1 = (−e−ab,−a).
u(b, a)d(b, a) =ZRZR
ZF
and ∆(b, a) = ea.
u(b, a) e−ada db
It is extremely well known, see, for example [Kha] or [Fol, 7.6-2], that the only
inequivalent infinite-dimensional irreducible representations are given by π±, which
are gained by inducing characters of positive/negative index from the normal sub-
group B = {(b, 0); B ∈ R} ∼= R. Explicitly, for ξ in L2(R) we have
π±(b, a)ξ(t) = e∓ie−tbξ(t − a).
WEAK AMENABILITY OF A(G)
19
Furthermore, we get a quasi-equivalence
(2.13)
λ ≃ π+ ⊕ π−.
We recall that the Lie algebra for F is given by
where exp(hX) = (0, h) and exp(hY ) = (h, 0).
f = hX, Y : [X, Y ] = Y i
v1(b)v2(a). Then there is S = SY,v in V N (F ) such that
Theorem 2.9. Let v1 in L1(R) be sobv1 ∈ Ac(R), let v2 ∈ Cc(R), and set v(b, a) =
S(u) =ZF
∂(Y,0)u(g, g−1)v(g) dg
for u in D(F × F ).
Proof. The details of this proof are similar to those in the proof of Theorem 2.8.
Indeed, we obtain for σ, σ′ in {±} and ξ in L2(R2) that
v1(σe−t − σ′e−a−t′
)v2(a)e−aξ(t − a, t′ + a) da
and if U ξ(t, t′) = ξ(t, t′ + t) we have
Ev(σ, σ′)ξ(t, t′) = √2πZR
U ∗E(σ, σ′)U ξ(t, t′) = √2πZR
v1(σe−t − σ′ea−t′
)v2(t − a)ea−tξ(a, t′) da.
We compute for ξ in F± = Cc(R) the derivative
dπ±(Y )ξ(t) = lim
g→0
1
h
[e∓ie−th − I]ξ(t) = ∓ie−tξ(t).
The operators from the Plancherel formula (Proposition 2.2) are given by δ±ξ(t) =
e−tξ(t), and we see that δ±F± ⊆ F±. Hence we may appeal to Lemma 2.3, and it
suffices to see that each of the operators
Ev(σ, σ′)(dπσ ⊗ I) is bounded for each σ, σ′ in {±}.
(2.14)
Since U ∗(dπσ ⊗ I)U = dπσ ⊗ I it suffices compute the norms of the operators
U ∗Ev(σ, σ′)U (dπσ ⊗ I). We first observe that
U ∗Ev(σ, σ′)U (dπσ ⊗ I)ξ(t, t′) = √2πiσZR
We then note that e−t ≤ σe−t − σ′ea−t′
v1(σe−t − σ′ea−t′
+ ea−t′
and observe that
)v2(t − a)e−tξ(a, t′) da.
ZR v1(σe−t−σ′ea−t′
)v2(t − a)e−t da
≤ZR v1(σe−t − σ′ea−t′
+ZR v1(σe−t − σ′ea−t′
≤ kv1αk∞ kv2k1 + kv1k1 kv2k∞ .
)σe−t − σ′ea−t′
v2(t − a) da
)ea−t′
v2(t − a) da
20
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
where α(t) = t for t in R. Hence by our usual technique we see that
kU ∗Ev(σ, σ′)U (dπσ ⊗ I)ξk2
2
≤ 2π (kv1αk∞ kv2k1 + kv1k1 kv2k∞)
×ZRZRZR v1(σe−t − σ′ea−t′
≤ 2π (kv1αk∞ kv2k1 + kv1k1 kv2k∞)kv1k1 kv2k∞ kξk2
2 .
)e−tv2(t − a)ξ(a, t′)2 dt da dt′
Hence (2.14) is verified.
(cid:3)
2.4. Remarks. We note that the failure of weak amenability of A(G), for G either
F or Hr, is shown in [ChGh1]; but this does not automatically imply failure of local
synthesis.
The one aspect in common in the strategies employed for each of the five (classes
of) groups G above is that the Lie derivative is always taken from a direction which
is trivial in any abelian quotient. We suspect that to do otherwise would entail
that for an abelian quotient, G/N , we would find that ∆G/N would be a set of
non-synthesis, which would contradict Theorem 1.3 as A(G/N ) ∼= L1([G/N ) is
amenable.
Our choice of Lie derivatives X, in forming operators SX,v, shares a property
in common with the choices made in [Joh3] for SO(3), and [ChGh1, ChGh2] for
F , Hr and the Heisenberg group H. Of course, there is an enormous gulf between
exploiting these for a failure of spectral synthesis calculation and showing that these
derivatives may be used to build non-trivial elements of H 1(A(G), V N (G)). It is
plausible that for each of our four semi-direct product groups G = N ⋊ A, above,
with the associated Lie algebra n ⋊ a, that the space of bounded derivations from
A(G) to V N (G) is isomorphic to the Lie algebra n.
3. Weak amenability of Fourier algebras
The following is well-known. We include a proof for convenience of non-specialists.
Proposition 3.1. (i) Let g be a non-abelian real Lie algebra. Then g contains
either su(2) or a non-abelian solvable algebra m.
(ii) Let m be a non-abelian solvable real Lie algebra. Then m contains one of the
following Lie algebras:
f = hX, Y : [X, Y ] = Y i (affine motion)
e = hT, X1, X2 : [T, X1] = X2, [T, X2] = −X1, [X1, X2] = 0i (Euclidean motion)
gθ = hT, X1, X2 : [T, X1] = X1 − θX2, [T, X2] = θX1 + X2, [X1, X2] = 0i, θ > 0,
(Gr´elaud), or
h = hX, Y, Z : [X, Y ] = Z, [X, Z] = 0 = [Y, Z]i (Heisenberg).
Proof. Every concept and fact from Lie theory which is used in this proof is well-
known and may be found in [HiNe], for example.
(i) Let r denote the solvable radical ideal of g. If r is non-abelian, we set m = r.
If r is abelian, we use the Levi-decomposition g = s + r, where s is semisimple. For
s we have the Iwasawa decomposition s = k + a + n. If k is non-abelian, then k
contains su(2) and we are done. Otherwise m = a + n is non-abelian and solvable;
WEAK AMENABILITY OF A(G)
21
indeed, see the construction in [HiNe, §13.3]. [In [ChGh1, Proposition 5.3] a further
refinement shows that m ⊇ f, in this case.]
(ii) We let m ) m′ ) ··· ) m(d−1) ) {0} be the derived series, so m(d−2) is non
abelian. For simplicity we assume m = m(d−2); thus we now have derived series
m ) m′ ) m′′ = {0}.
Suppose, first there is an element S of m for which ad S : m → m admits a
non-zero eigenvalue a + ib, a, b ∈ R. If b = 0, then there is an eigenvector Y for
ad S, i.e. [S, Y ] = aY . Put X = 1
a S and we get f = span{X, Y }. Otherwise, we
appeal to the real Jordan decomposition of ad S, which allows us to find elements
Y1, Y2 of m for which
[S, Y1] = aY1 − bY2
and [S, Y2] = bY1 + aY2.
a S and Xj = Yj (j = 1, 2) if ab > 0; and let T = 1
If a = 0, we let T = 1
b S and we notice immediately that Y1 = X1, Y2 = X2 ∈ m′,
so [Y1, Y2] = 0. Thus we get e = span{T, Y1, Y2}. Finally, if ab 6= 0, we let θ = b/a.
Let T = 1
b S, X1 = Y2, X2 = Y1
if ab < 0. It is straightforward to check that X1, X2 ∈ m′ so [X1, X2] = 0. Hence,
gθ = span{T, X1, X2}, in this case.
If no S in m has that ad S admits a non-zero eigenvalue, then each ad S is
nilpotent, and hence by Engel's theorem m is nilpotent. We consider the central
series m ) C(m) ) ··· ) Cn(m) ) {0}. Any Z in Cn(m) is hence central and we
further consider such Z for which there are X, Y in m (one of which is in Cn−1(m))
for which Z = [X, Y ]. Then h = span{X, Y, Z}.
Theorem 3.2. Let G be a connected Lie group. Then the following are equivalent:
(cid:3)
(i) A(G) is weakly amenable,
(ii) the anti-diagonal ∆G is of local synthesis for G × G, and
(iii) G is abelian.
we may restrict ourselves to the case of simply connected groups.
Proof. The result (i) ⇒ (ii) is Theorem 1.3, while (iii) ⇒ (i) is well-known, i.e.
A(G) ∼= L1(bG), in this case. Hence it remains to prove that (ii) ⇒ (iii). We will
assume that G is non-abelian, and show that ∆G is not of local synthesis for G× G.
All Lie theoretic concepts and nomenclature, used in this proof, may be found
in [HiNe]. We will show that when G is non-abelian, then ∆G fails to be of local
synthesis for G × G. Thanks to Theorem 1.6, and the fact that any G ∼= eG/Γ for
a simply connected connected Lie group eG and a discrete central subgroup Γ of eG,
Let G be simply connected with Lie algebra g. The Levi decomposition g = s + r
begets respective integral subgroups S and R of G, with S semisimple and R normal,
and hence, essentially by the smooth splitting theorem, each is closed and simply
connected. Given a Lie subalgebra m of g, in particular one of the algebras arising
from Proposition 3.1, above, we let M be the integral subgroup generated by m.
If m = su(2), so m ⊂ s, then M ∼= SU(2)/C, where C is a subgroup of the centre
{±1}, and hence M is closed.
If m ⊂ s and is non-abelian and solvable, then
M ⊆ AN (Iwasawa decomposition of S); or if m ⊆ r, then M ⊆ R. Any integral
subgroup of a simply connected solvable group is closed and simply connected.
Thus, corresponding to each of the solvable Lie algebras f, e, gθ or h, as obtained in
Proposition 3.1, above, we see that M is closed and isomorphic to one of the affine
motion group F , the simply connected cover eE of the Euclidean motion group E,
G´elaud's group Gθ, or the Heisenberg group H.
22
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
Proposition 2.4, Theorem 2.5, Theorem 2.6, Theorem 2.8, and Theorem 2.9, and
then Lemma 2.1, tells us that ∆H fails to be a set of (local) synthesis for H × H,
for each of H = SU(2), E, Hr, Gθ and F . Then, Theorem 1.6 tells us the same for
H = SU(2)/C, eE and H. Hence applying Theorem 1.4 to the subgroup M of G,
above, we obtain (ii).
Let G be a locally compact group. The following is an immediate consequence
of the restriction theorem for closed subgroups H, i.e. RH (A(G)) = A(H) (see the
comments before Theorem 1.4); and the fact that any quotient of a commutative
weakly amenable Banach algebra is again weakly amenable.
(cid:3)
Corollary 3.3. Let G be a locally compact group. If A(G) is weakly amenable,
then any connected Lie subgroup of G is abelian.
Example 3.4. Let us consider an example of a non-abelian solvable connected
group which contains no non-abelian closed connected Lie subgroups. We consider
motion group.
We let Rap denote the almost periodic compactification of the real line, and for
each r in R we let χr : Rap → T be the unique continuous character extending
and is isomorphic to the discretized reals, Rd.
If K is a closed connected Lie
a group related to the simply connected covering group, eE = C⋊R, of the Euclidean
t 7→ eitr : R → T. We observe that the dual group is given by dRap = {χr : r ∈ R}
subgroup of Rap, then K ∼= Tn, where n = 0, 1, 2, . . . , so bK ∼= Zn. On the other
hand bK is a quotient of Rd, and hence is either trivial or divisible. Thus K = {1},
and we conclude that Rap contains no non-trivial closed Lie subgroups.
Now let E = C ⋊ Rap with multiplication given by
(x, ζ)(x′, ζ ′) = (x + χ1(ζ)x′, ζζ ′).
We may think of this group as a partial compactification of eE along its quotient
subgroup. We note that q(x, ζ) = (x, χ1(ζ)) defines a quotient homomomorphism
from E onto E. For notation convenience we identify Rap ∼= {0} ⋊ Rap in E, and
T ∼= {0} ⋊ T in E, and we identify C with C ⋊ {1} in either E, or in E, as should
be clear by context. Notice that qC is a homeomorphism.
We now show that E admits no non-abelian closed connected Lie subgroups. We
first note that the proper Lie subalgebras of e are either one-dimensional, the two-
dimensional ideal n = RX1 + RX2, or {0}. Since every closed connected subgroup
of E corresponds to a Lie subalgebra of e, we find that the one-dimensional closed
connected subgroups are of the form
MS+X = exp(R(S + X)) = {((z − 1)x, z) : z ∈ T}
and LX = exp(RX) = {(tx, 1) : t ∈ R}
where X ∈ n with exp X = (x, 1), while the only two-dimensional closed connected
subgroup is C. Taking q−1(M ) for each subgroup M of E, listed above, we get
M S+X = {((χ1(ζ) − 1)x, ζ) : ζ ∈ Rap}, LX ⋊ ker χ1 and C ⋊ ker χ1.
Let H be a proper closed connected subgroup of E and H0 = H ∩ C. Notice that
in E, q(H0) = q(H) ∩ C ∼= H0. If H0 = C then for (x, ζ) in H we have (0, ζ) =
(−x, 1)(x, ζ) ∈ H ∩ Rap, so the image of H under the quotient map (x, ζ) 7→ ζ
is closed, and hence either the trivial group {1} or is non-Lie. If H0 ( C, then
WEAK AMENABILITY OF A(G)
23
q(H) ∩ C ∼= H0 ( C, and it follows that H is a subgroup of one of the abelian
groups M S+X , LX ⋊ ker χ1 or Rap, and hence is abelian.
At present, we are aware of no method for determining if A(E) is weakly amenable.
Acknowledgement. We are grateful to Søren Knudby for pointing out a flaw
in an earlier version of the proof of Theorem 3.2, and for guidance in correcting
it. We also thank Yemon Choi for pointing out an error in a proposition which
purported to reduce the problem to the case of pro-solvable Ge. We realized there
was a serious gap in the proof and have deleted the proposition from the final
version of the paper.
References
[Ars] G. Arsac. Sur l'espace de Banach engendr´e par les coefficients d'une repr´esentation unitaire.
Publ. D´ep. Math. (Lyon) 13 (1976), no. 2, 1 -- 101.
[BCD] W.G. Bade, P. Curtis and H.G. Dales. Amenability and weak amenability for Beurling
and Lipschitz algebras. Proc. London Math. Soc. (3) 55 (1987), no. 2, 359 -- 377.
[ChGh1] Y. Choi and M. Ghandehari. Weak and cyclic amenability for Fourier algebras of con-
nected Lie groups. J. Funct. Anal. 266 (2014), no. 11, 6501 -- 6530.
[ChGh2] Y. Choi and M. Ghandehari. Weak amenability for Fourier algebras of 1-connected nilpo-
tent Lie groups, pre-print, 20 pages. See arXiv:1405.6403.
[DeDe] J. Delaporte and A. Dehrighetti. On Herz' extension theorem. Boll. Un. Mat. Ital. A (7)
6 (1992), no. 2, 245 -- 247.
[Eym] P. Eymard. L'alg`ebre de Fourier d'un groupe localement compact. Bull. Soc. Math. France
92 (1964), 181 -- 236.
[Fol] G.B. Folland. A course in abstract harmonic analysis. CRC Press, Boca Raton, FL, 1995.
[FoRu] B.E. Forrest and V. Runde. Amenability and weak amenability of the Fourier algebra.
Math. Z. 250 (2005), no. 4, 731 -- 744.
[FSS1] B.E. Forrest, E. Samei and N. Spronk. Weak amenability of Fourier algebras on compact
groups. Indiana Univ. Math. J. 58 (2009), no. 3, 1379 -- 1393.
[FSS2] B.E. Forrest, E. Samei and N. Spronk. Convolutions on compact groups and Fourier alge-
bras of coset spaces. Studia Math. 196 (2010), no. 3, 223 -- 249.
[GrRy] I.S. Gradshteyn and I.M. Ryzhik. Table of integrals, series, and products. Translated from
the Russian. Translation edited and with a preface by Alan Jeffrey and Daniel Zwillinger.
Seventh edition. Elsevier/Academic Press, Amsterdam, 2007.
[Grø] N. Groenbaek. A characterization of weakly amenable Banach algebras. Studia Math. 94
(1989), no. 2, 149 -- 162.
[Her] C. Herz. Harmonic synthesis for subgroups. Ann. Inst. Fourier (Grenoble) 23 (1973), no. 3,
91 -- 123.
[HeRoII] E. Hewitt and K.A. Ross. Abstract harmonic analysis. Vol. II: Structure and analysis
for compact groups. Analysis on locally compact Abelian groups. Springer-Verlag, New York-
Berlin, 1970.
[HiNe] J. Hilgert and K.H. Neeb. Structure and geometry of Lie groups. Springer, New York,
2012.
[Joh1] B.E. Johnson. Cohomology in Banach algebras. Memoirs of the American Mathematical
Society, No. 127. American Mathematical Society, Providence, R.I., 1972.
[Joh2] B.E. Johnson. Weak amenability of group algebras. Bull. London Math. Soc. 23 (1991),
no. 3, 281 -- 284.
[Joh3] B.E. Johnson. Non-amenability of the Fourier algebra of a compact group. J. London Math.
Soc. (2) 50 (1994), no. 2, 361 -- 374.
[Kam] H. Kamowitz. Cohomology groups of commutative Banach algebras. Trans. Amer. Math.
Soc. 102 (1962), 352 -- 372.
[Kan] E. Kaniuth. A course in commutative Banach algebras. Springer, New York, 2009.
[KaTa] E. Kaniuth and K.F. Taylor. Induced representations of locally compact groups. Cam-
bridge University Press, Cambridge, 2013.
[Kha] I. Khalil. Sur l'analyse harmonique du groupe affine de la droite. Studia Math. 51 (1974),
139 -- 167.
24
HUN HEE LEE, JEAN LUDWIG, EBRAHIM SAMEI AND NICO SPRONK
[KiMu] W. Kirsch and D. Muller. On the synthesis problem for orbits of Lie groups in Rn. Ark.
Mat. 18 (1980), no. 2, 145 -- 155.
[Lep] H. Leptin. Sur l'alg`ebre de Fourier d'un groupe localement compact. C. R. Acad. Sci. Paris
S´er. A-B 266 (1968), A1180 -- A1182.
[Loh] N. Lohoue. Remarques sur les ensembles de synth`ese des alg'ebres de groupe localement
compact. J. Funct. Anal. 13 (1973), 185 -- 194.
[Los] V. Losert. On tensor products of Fourier algebras. Arch. Math. (Basel) 43 (1984), no. 4,
370 -- 372.
[LuTu1] J. Ludwig and L. Turowska. On the connection between sets of operator synthesis and sets
of spectral synthesis for locally compact groups. J. Funct. Anal. 233 (2006), no. 1, 206 -- 227.
[LuTu2] J. Ludwig and L. Turowska. Growth and smooth spectral synthesis in the Fourier algebras
of Lie groups. Studia Math. 176 (2006), no. 2, 139 -- 158.
[McM] J.R. McMullen. Extensions of positive-definite functions. Memoirs of the American Math-
ematical Society, No. 117. American Mathematical Society, Providence, R. I., 1972.
[Mul] D. Muller. On the spectral synthesis problem for hypersurfaces of RN . J. Funct. Anal. 47
(1982), no. 2, 247 -- 280.
[PaSa] B.D. Park and E. Samei. Smooth and weak synthesis of the anti-diagonal in Fourier alge-
bras of Lie groups. J. Lie Theory 19 (2009), no. 2, 275 -- 290.
[Ply] R.J. Plymen. Fourier algebra of a compact Lie group, unpublished. See ArXiv math/0104018.
[ReSt] H. Reiter and J. Stegeman. Classical harmonic analysis and locally compact groups. Second
edition. The Clarendon Press, Oxford University Press, New York, 2000.
[Rua] Z.-J. Ruan. The operator amenability of A(G). Amer. J. Math. 117 (1995), no. 6, 1449 -- 1474.
[Run] V. Runde. A functorial approach to weak amenability for commutative Banach algebras.
Glasgow Math. J. 34 (1992), no. 2, 241 -- 251.
[Sam1] E. Samei. Bounded and completely bounded local derivations from certain commutative
semisimple Banach algebras. Proc. Amer. Math. Soc. 133 (2005), no. 1, 229 -- 238
[Sam2] E. Samei. Hyper-Tauberian algebras and weak amenability of Fig`a-Talamanca -- Herz alge-
bras. J. Funct. Anal. 231 (2006), no. 1, 195 -- 220.
[SiWe] I.M. Singer and J. Wermer. Derivations on commutative normed algebras. Math. Ann.
129, (1955).
[Spr] N. Spronk. Operator weak amenability of the Fourier algebra. Proc. Amer. Math. Soc. 130
(2002), no. 12, 3609 -- 3617.
[Sug] M. Sugiura. Unitary representations and harmonic analysis. An introduction. Second edi-
tion. North-Holland Publishing Co., Amsterdam; Kodansha, Ltd., Tokyo, 1990.
[Tat] N. Tatsuuma. Plancherel formula for non-unimodular locally compact groups. J. Math.
Kyoto Univ. 12 (1972), 179 -- 261.
[Tre] F. Treves. Topological vector spaces, distributions and kernels. Academic Press, New York-
London, 1967.
[Var] N.Th. Varopoulos. Spectral synthesis on spheres. Proc. Cambridge Philos. Soc. 62 (1966),
379?387.
[War] C.R. Warner. Weak spectral synthesis. Proc. Amer. Math. Soc. 99 (1987), no. 2, 244?248.
Addresses:
Department of Mathematical Sciences and Research Institute of Math-
ematics, Seoul National University, Gwanak-ro 1, Gwanak-gu, Seoul
08826, Republic of Korea
Institut ´Elie Cartan de Lorraine, Universit´e de Lorraine -- Metz,
Batiment A, Ile du Saulcy, F-57045 Metz, France
Department of Mathematics and Statistics, University of Saskatchewan,
Room 142 McLean Hall, 106 Wiggins Road Saskatoon, SK, S7N 5E6,
Canada
Department of Pure Mathematics, University of Waterloo, Waterloo,
ON, N2L 3G1, Canada
WEAK AMENABILITY OF A(G)
25
Email-adresses:
[email protected]
[email protected]
[email protected]
[email protected]
|
1107.1700 | 1 | 1107 | 2011-07-08T19:04:59 | Wavelet analysis on adeles and pseudo-differential operators | [
"math.FA"
] | This paper is devoted to wavelet analysis on adele ring $\bA$ and the theory of pseudo-differential operators. We develop the technique which gives the possibility to generalize finite-dimensional results of wavelet analysis to the case of adeles $\bA$ by using infinite tensor products of Hilbert spaces. The adele ring is roughly speaking a subring of the direct product of all possible ($p$-adic and Archimedean) completions $\bQ_p$ of the field of rational numbers $\bQ$ with some conditions at infinity. Using our technique, we prove that $L^2(\bA)=\otimes_{e,p\in\{\infty,2,3,5,...}}L^2({\bQ}_{p})$ is the infinite tensor product of the spaces $L^2({\bQ}_{p})$ with a stabilization $e=(e_p)_p$, where $e_p(x)=\Omega(|x|_p)\in L^2({\bQ}_{p})$, and $\Omega$ is a characteristic function of the unit interval $[0,\,1]$, $\bQ_p$ is the field of $p$-adic numbers, $p=2,3,5,...$; $\bQ_\infty=\bR$. This description allows us to construct an infinite family of Haar wavelet bases on $L^2(\bA)$ which can be obtained by shifts and multi-delations. The adelic multiresolution analysis (MRA) in $L^2(\bA)$ is also constructed. In the framework of this MRA another infinite family of Haar wavelet bases is constructed. We introduce the adelic Lizorkin spaces of test functions and distributions and give the characterization of these spaces in terms of wavelet functions. One class of pseudo-differential operators (including the fractional operator) is studied on the Lizorkin spaces. A criterion for an adelic wavelet function to be an eigenfunction for a pseudo-differential operator is derived. We prove that any wavelet function is an eigenfunction of the fractional operator. These results allow one to create the necessary prerequisites for intensive using of adelic wavelet bases and pseudo-differential operators in applications. | math.FA | math |
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL
OPERATORS
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Abstract. This paper is devoted to wavelet analysis on adele ring A and the theory of
pseudo-differential operators. We develop the technique which gives the possibility to
generalize finite-dimensional results of wavelet analysis to the case of adeles A by using
infinite tensor products of Hilbert spaces. The adele ring is roughly speaking a subring
of the direct product of all possible (p-adic and Archimedean) completions Qp of the
field of rational numbers Q with some conditions at infinity. Using our technique, we
prove that L2(A) = ⊗e,p∈{∞,2,3,5,... }L2(Qp) is the infinite tensor product of the spaces
L2(Qp) with a stabilization e = (ep)p, where ep(x) = Ω(xp) ∈ L2(Qp), and Ω is a
characteristic function of the unit interval [0, 1], Qp is the field of p-adic numbers, p =
2, 3, 5, . . . ; Q∞ = R. This description allows us to construct an infinite family of Haar
wavelet bases on L2(A) which can be obtained by shifts and multi-delations. The adelic
multiresolution analysis (MRA) in L2(A) is also constructed. In the framework of this
MRA another infinite family of Haar wavelet bases is constructed. We introduce the
adelic Lizorkin spaces of test functions and distributions and give the characterization
of these spaces in terms of wavelet functions. One class of pseudo-differential operators
(including the fractional operator) is studied on the Lizorkin spaces. A criterion for
an adelic wavelet function to be an eigenfunction for a pseudo-differential operator
is derived. We prove that any wavelet function is an eigenfunction of the fractional
operator. These results allow one to create the necessary prerequisites for intensive
using of adelic wavelet bases and pseudo-differential operators in applications.
1. Introduction
1.1. p-Adic and adelic analysis. During a few hundred years theoretical physics has
been developed on the basis of real and, later, complex numbers. The p-adic numbers
were described by K. Hensel in 1897 to transfer the ideas and techniques of power
series methods to number theory. According to the well-known Ostrovsky theorem, any
nontrivial valuation on the field Q of rational numbers is equivalent either to the real
valuation · or to one of the p-adic valuations · p. The corresponding completions
of Q give the fields R or Qp. The theory of p-adic numbers has already penetrated
intensively into several areas of mathematics and its applications. In the last 20 years
the field of p-adic numbers Qp (as well as its algebraic extensions, including the field of
complex p-adic numbers Cp) has been intensively used in theoretical and mathematical
physics, p-adic string theory, gravity and cosmology, the theory of stochastic differential
equations over the field of p-adic numbers, Feynman path integration over p-adics, the
Date:
2000 Mathematics Subject Classification. 11F85, 42C40, 47G30; Secondary 26A33, 46F10.
Key words and phrases. Adeles, wavelets, multiresolution analysis, pseudo-differential operators,
fractional operator.
The first and third authors (A. K. and V. S.) were supported by the grant of the Profile Mathematical
Modeling and System Collaboration of Vaxjo University (Sweden). The third author (V. S.) was
supported in part by Grant 09-01-00162 of Russian Foundation for Basic Research. The second (A. K.)
and third (V. S.) were supported by the DFG Projects.
1
2
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
theory of p-adic valued probabilities and dynamical systems, in theory of disordered
systems (spin glasses) (see [8], [26], [27], [30], [38], [40], [64] and the references therein).
Applications were, however, not only restricted to physics. p-Adic models were also
proposed in psychology, cognitive and social sciences, and, e.g., in biology, image analysis
(see [27], [28]).
These applications induced and stimulated a development of new branches of p-adic
analysis, in particular, the theory of p-adic wavelets. Recall that nowadays wavelets are
applied in a lot of branches of modern mathematics and engineering area. The first real
wavelet basis (3.5), (3.6) was introduced by Haar in 1910. However, for almost a century
nobody could find another wavelet function (a function whose shifts and delations form
an orthogonal basis). Only in the early nineties a method for a more general construc-
tion of the wavelet functions appeared [46], [47], [50], [51]. This method is based on
the notion of multiresolution analysis (MRA in the sequel). In the p-adic setting the
situation was the following. In 2002 S. V. Kozyrev [39] constructed a compactly sup-
ported p-adic wavelet basis (3.21) in L2(Qp), which is an analog of the real Haar basis
(3.5), (3.6). J. J. Benedetto and R. L. Benedetto [10], R. L. Benedetto [11] suggested a
method for constructing wavelet bases on locally compact abelian groups with compact
open subgroups. This method is applicable for the p-adic setting. It is based on a theory
of wavelet sets and only allows the construction of wavelet functions whose Fourier trans-
forms are the characteristic functions of some sets (see [10, Proposition 5.1.]). Moreover,
these authors doubted that the development of the MRA approach is possible. In spite
of the above opinions and arguments [10], [11], in [57], the p-adic MRA theory in L2(Qp)
was developed and new p-adic wavelet bases were constructed. Some important results
in p-adic wavelet theory were obtained in [1], [2], [34], [35]. It turned out that the the-
ory of p-adic wavelets plays an important role in the study of p-adic pseudo-differential
operators and equations [3], [4], [5], [33], [39], [40], [57]. This theory gives a powerful
technique to deal with p-adic pseudo-differential equations. Recall that on complex-
valued functions defined on Qp, the operation of differentiation is not defined . As a
result, a large number of p-adic models use pseudo-differential equations instead of dif-
ferential equations. The p-adic multidimensional fractional operator Dα was introduced
by M. Taibleson [58] (see also [59]) in the space of distributions D′(Qn
p ). The spectral
theory of this fractional operator was developed by V.S. Vladimirov in [62], in particular,
explicit formulas for the eigenfunctions of this operator were constructed (see also [64]).
In [63] (see also [64]) V.S. Vladimirov constructed the spectral theory of the Schrodinger-
type operator Dα + V (x), which was further developed by A.N. Kochubei [36], [37] (see
also [38]).
The adele ring A is some subring of the direct product of all possible (p-adic and
Archimedean) completions Qp of the field of rational numbers Q. The group of ideles
was introduced by Chevalley in 1936 [16] as a part of his program to formulate class
field theory so that it worked for infinite-degree extensions. The adeles were introduced
by Weil in the late 1930s as an additive analogue of ideles. The ring A is often used
in advanced parts of number theory (for example, see [13], [15], [17], [48], [49], [65]).
For example, Tate's proof (see in [15, J. T. Tate, pp.305 -- 347]) of the functional equa-
tion Z(s) = Z(1 − s), where Z(s) = π−s/2Γ(s/2)ζ(s) for the Riemann zeta func-
ns , ℜ(s) > 1, is based on the Fourier analysis on adeles A and
ideles (see also [17, 4.7]). Recently the theory of adeles has been successfully ap-
plied in various parts of contemporary mathematical and theoretical physics. Namely,
tion ζ(s) = Pn∈N
1
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
3
there are adelic constructions, e.g., in statistical mechanics, stochastics, string the-
ory, quantum cosmology, and quantum mechanics (see [9], [14], [19] -- [21], [25], [38,
5.8] [26], [27], [60], [61], [64], [66], and the references therein).
However it should be noted that in contrast to the real and p-adic analysis, the
adelic analysis practically (in particular, the adelic wavelet theory and the theory of
adelic pseudo-differential operators and equations) has not been developed so far. In
particular, the necessity of the development the wavelet theory on adeles was mentioned
in [10].
1.2. Contents of the paper. In this paper some problems of the adelic harmonic
analysis are studied. The points to be considered are, first, the theory of adelic Haar
wavelets; and second, the theory of simplest adelic pseudo-differential operators in con-
nection with wavelets.
In Sec. 2, some facts from the p-adic and adelic analysis are given. In Sec. 3, we recall
definitions of the real MRA and present some results on p-adic MRA and wavelet bases
from the papers [35], [57].
In Sec. 4, the theory of infinite tensor products of Hilbert spaces [52], [12] is used to
generalize finite-dimensional results to the case of adeles. We recall the constructions
of infinite tensor product of Hilbert spaces He = ⊗e,n∈NHn, the complete von Neumann
product of infinitely many Hilbert spaces. The space He can be obtained as the closure
of the union of some subspaces. We observe certain stability of the space He with respect
to the variation of the corresponding subspaces (see Lemma 4.2). In Subsec. 4.4, using
Lemma 4.2 for the special stabilizing sequence e = (ep)p, ep(xp) = φp(xp) = Ω(xpp)
(here Ω(t) is the characteristic function of the segment [0, 1] ⊂ R) we show that L2(A)
coincides with the infinite tensor product of the Hilbert spaces L2(Qp) over all possible
completions of the field Q: L2(A) = ⊗e,pL2(Qp) (see Lemma 4.3).
In Sec. 5, we apply the above scheme to construct adelic wavelet bases (5.5) on
L2(A, dx) generated by the tensor product of one-dimensional Haar wavelet bases (3.5),
(3.22), (3.25). We would like to stress that to construct adelic wavelet bases, we need
the p-adic wavelet bases that contain functions φp(xp) = Ω(xpp). Thus, instead of
the Haar wavelet basis (3.20) we will use modified Haar basis (3.22). According to [1],
there are no orthogonal p-adis MRA based wavelet bases except for those described in
Theorem 3.1 and Corollary 3.1. Thus, formula (5.5) gives all adelic Haar wavelet bases
generated by the tensor product of one-dimensional Haar wavelet bases. Here infinite
tensor product depends on the special stabilization e = (ep)p, where ep(xp) = Ω(xpp).
Recall that in [20, Sec.4, formula (4.1)], a basis in the space L2(A) associated with
eigenfunctions of harmonic oscillators was constructed (see Remark 5.1). To construct
adelic wavelet bases, we use the standard Haar measures on Qp and on the adele ring A.
For measures on Qp different from Haar measure, see [3, Appendix D], based on [42].
In Sec. 6, using the idea of constructing separable multidimensional MRA by means
of the tensor product of one-dimensional MRAs suggested by Y. Meyer [50] and S. Mal-
lat [46], [47], we construct adelic wavelet bases in L2(A). In a general situation we show
how using some system of closed subspaces V (k)
, j ∈ Z, in a Hilbert spaces Hn, n ∈ N,
with properties (a) -- (c) of the MRA (see Definitions 3.1, 3.2) one can construct various
systems of subspaces with the same properties in the infinite tensor product of the spaces
Hn. In Theorem 6.2 adelic separable MRA is constructed. The refinable function of this
MRA given by (6.14) is an infinite product refinable functions of all one-dimensional
j
4
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
MRAs. In the framework of this MRA an infinite family of adelic Haar wavelet bases
(6.25) are constructed.
In Sec. 7, we introduce the adelic Lizorkin spaces of test functions Φ(A) and distri-
butions Φ′(A). These spaces are constructed by using the original real Lizorkin spaces
introduced in [43] -- [45] and the p-adic Lizorkin spaces introduced in [4] (see also [3,
Ch. 7]). A basic motivation for using Lizorkin spaces rather than the Schwartz and
Bruhat -- Schwartz spaces of distributions S ′(R) and D′(Qp) is due to the fact that the
latter spaces are not invariant under the fractional operators. Next, in Sec. 8, the char-
acterization of the adelic Lizorkin spaces in terms of wavelets is given. Namely, it is
proved that any test function from Φ(A) can be represented in the form of a finite
combination of adelic wavelet functions (5.6), and any distribution from Φ′(A) can be
represented as an infinite linear combination of adelic wavelet functions (5.6) (in [7],
assertions of these types were stated for ultrametric Lizorkin spaces).
In the framework of our constructions the following three facts seem to have the
same deep reason: (1) functions φ(p)(xp) = Ω(xpp) are stabilization functions in the
adelic Bruhat -- Schwartz space (see Definition 2.3), (2) we use the stabilization sequence
e = (ep)p where ep(xp) = Ω(xpp) in proving the fact that L2(A) = ⊗e,pL2(Qp), (3) under
the projection of the space L2(Qp) onto L2(Zp) some elements of wavelet basis (3.20) and
Kozyrev's wavelet basis (3.21) are transformed into functions which are proportional to
the same function Ω(xpp) (see Propositions 3.1 -- 3.3).
In Sec. 9, by Definition (9.17), (9.18) a class of pseudo-differential operators on the
adelic Lizorkin spaces is introduced. The fractional operators Dbγ, bγ ∈ C∞ (see Def-
inition (9.9), (9.10), (9.13)), and Dγ, γ ∈ C (see Definition (9.14), (9.15)) belong to
this class. We prove that the Lizorkin spaces of test functions Φ(A) and distributions
Φ′(A) are invariant under the above-mentioned pseudo-differential operators. More-
over, a family of fractional operators on the space of distributions Φ′(A) forms an abelian
group. Thus the Lizorkin spaces constitute "natural" domains of definition for this class
of pseudo-differential operators. Note, that in [31] the fractional operator was considered
in L2(A). In Subsec. 9.4, the spectral theory of adelic pseudo-differential operators is
developed. By Theorem 9.1, we derive a criterion for an adelic wavelet function to be an
eigenfunction for a pseudo-differential operator. It is proved that any wavelet function
is an eigenfunction of a fractional operator. Thus the adelic wavelet analysis is closely
connected with the spectral analysis of pseudo-differential operators. Using results of
Sec. 9, similarly to the p-adic case, one can develop the "variable separation method"
(an analog of the classical Fourier method) to reduce solving adelic pseudo-differential
equations to solving ordinary differential equations with respect to the real variable t
(for details, see [3, Ch. 10], [6]).
2. Preliminary results
2.1. p-Adic numbers. We shall systematically use the notation and results from the
book [64]. Let N, Z, Q, R, C be the sets of positive integers, integers, rational, real, and
complex numbers, respectively.
According to the well-known Ostrovsky theorem, any nontrivial valuation on the field
of rational numbers Q is equivalent either to the real valuation · or to one of the p-
adic valuations · p. This p-adic norm · p is defined as follows: if an arbitrary rational
number x 6= 0 is represented as x = pγ m
n , where γ = γ(x) ∈ Z and the integers m, n are
not divisible by p, then
xp = p−γ,
0p = 0.
(2.1)
x 6= 0,
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
5
The norm · p is non-Archimedean and satisfies the strong triangle inequality x + yp ≤
max(xp,yp). The completion of Q with respect to the usual absolute value · gives
the field of real numbers R. The field Qp of p-adic numbers is defined as the completion
of the field of rational numbers Q with respect to the norm · p. Next we will denote
x∞ = x, Q∞ = R and Z∞ = Z. By
(2.2)
VQ = {∞, 2, 3, 5, . . .}
we denote the set of indices for all valuations on the field Q.
Any p-adic number x ∈ Qp, x 6= 0, is represented in the canonical form
(2.3)
x =
where γ = γ(x) ∈ Z, xk ∈ Fp = {0, 1, . . . , p − 1}, xγ 6= 0, γ ≤ k < ∞. The series is
convergent in the p-adic norm · p, and one has xp = p−γ. The fractional part of the
number x ∈ Qp (given by (2.3)) is defined as follows
(2.4)
if γ(x) ≥ 0 or x = 0,
xγpγ + · · · + x−1p−1,
if γ(x) < 0.
{x}p =(cid:26) 0,
The set Q×
p = Qp \ {0} is the multiplicative group of the field Qp. p-Adic numbers
Zp = {x ∈ Qp : xp ≤ 1} are called integer p-adic numbers. In view of (2.3), Zp consists
of p-adic numbers
xkpk,
∞Xk=γ
xkpk.
∞Xk=0
(2.5)
x =
Zp is a subring of the ring Qp. The multiplicative group of the ring Zp is the set
Z×
p =(cid:8)x ∈ Zp : xp = 1(cid:9) =nx ∈ Zp : x =
∞Xk=0
xkpk,
x0 6= 0o.
Denote by Bγ(a) = {x ∈ Qp : x − ap ≤ pγ} the ball of radius pγ with the center
at the point a ∈ Qp and by Sγ(a) = {x ∈ Qp : x − ap = pγ} = Bγ(a) \ Bγ−1(a) the
corresponding sphere, γ ∈ Z. For a = 0 we set Bγ = Bγ(0) and Sγ = Sγ(0).
2.2. p-Adic distributions. A complex-valued function f defined on Qp is called locally-
constant if for any x ∈ Qp there exists an integer l(x) ∈ Z such that
f (x + y) = f (x),
y ∈ Bl(x).
Let E(Qp) and D(Qp) be the linear spaces of locally-constant C-valued functions
on Qp and locally-constant C-valued compactly supported functions (so-called Bruhat --
Schwartz test functions), respectively [64, VI.1.,2.]. If ϕ ∈ D(Qp), then according to
Lemma 1 from [64, VI.1.], there exists l ∈ Z, such that
ϕ(x + y) = ϕ(x),
y ∈ Bl,
x ∈ Qp.
The largest of such numbers l = l(ϕ) is called the parameter of constancy of the function
ϕ. Let us denote by Dl
N (Qp) from
D(Qp) having supports in the ball BN and with parameters of constancy ≥ l. The
following embedding holds:
N (Qp) ⊂ Dl′
Dl
N (Qp) the finite-dimensional space of test functions Dl
N ′(Qp), N ≤ N ′,
l ≥ l′.
6
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
N→∞
indDN (Qp), where DN (Qp) = lim
We have D(Qp) = lim
N (Qp) (see [64, VI.2]).
These representations give us the inductive limit topology on the corresponding spaces.
Denote by D′(Qp) the set of all linear functionals (Bruhat -- Schwartz distributions) on
The Fourier transform of ϕ ∈ D(Qp) is defined by the formula
D(Qp) [64, VI.3.].
l→−∞
indDl
bϕ(ξ) = F [ϕ](ξ) =ZQp
χp(ξx)ϕ(x) dx,
ξ ∈ Qp,
where dx is the Haar measure on Qp such thatRxp≤1 dx = 1, and
χp(x) = e2πi{x}p
(2.6)
is the additive character on Qp (see [64, III.1.]), {x}p is the fractional part (2.4) of the
number x ∈ Qp. The Fourier transform is a linear isomorphism D(Qp) onto D(Qp) [59,
III,(3.2)], [64, VII.2.]. Moreover,
ϕ ∈ Dl
iff F [ϕ] ∈ D−N
−l (Qp).
N (Qp)
(2.7)
The Fourier transform of a distribution f ∈ D′(Qp) is the distribution bf = F [f ] defined
by the relation
hF [f ], ϕi = hf, F [ϕ]i,
∀ ϕ ∈ D(Qp).
Here and in the sequel hf, ϕi denotes the action of a distribution f on a test function ϕ.
If f ∈ D′(Qp), a ∈ Q×
p , b ∈ Qp, then [64, VII,(3.3)]:
(2.8)
F [f (ax + b)](ξ) = a−1
According to [64, IV,(3.1)],
p χp(cid:16) −
b
a
a(cid:17),
ξ(cid:17)F [f (x)](cid:16) ξ
x, ξ ∈ Qp.
F [Ω(p−k · p)](x) = pkΩ(pkxp),
(2.9)
where Ω(t) is the characteristic function of the segment [0, 1] ⊂ R. In particular,
(2.10)
x ∈ Qp,
k ∈ Z,
F [Ω(ξp)](x) = Ω(xp).
2.3. Adeles. We use the notation and results from [24, Ch. III,§1,2] and [17].
Definition 2.1. The adeles of Q are
and the ideles of Q are
A = AQ =(cid:8)(xp)p∈VQ ∈ Yp∈VQ
J = JQ =(cid:8)(xp)p∈VQ ∈ Yp∈VQ
Qp : xp ∈ Zp for almost all p 6= ∞(cid:9),
p for almost all p 6= ∞(cid:9),
p : xp ∈ Z×
Q×
where VQ = {∞, 2, 3, 5, . . .} is the set of indices.
The adele ring A is the restricted direct product of R and Qp for all p = 2, 3, . . . with
respect to the integer rings Zp (see [17]). If componentwise operations of addition and
multiplication are introduced, A is a ring of adeles and J is a multiplicative group. The
additive group of the ring A is called the adelic group.
There is a natural imbedding Q 7→ A given by
Q ∋ r 7→ (r, r, . . . , r, . . . ) ∈ A.
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
7
Indeed, any constant sequence (r, r, . . . , r, . . . ) is an adele since r ∈ Zp for any p not
dividing the denominator of r. The adeles (ideles) of the form (r, r, . . . , r, . . . ), where
r ∈ Q, are called principal adeles (respectively, ideles).
To define a topology on A, we show that A is a union of locally compact groups. Note
Qp is not a locally compact group.
that the direct productQp∈VQ
Definition 2.2. Let S be a finite subset of VQ such that ∞ ∈ S ⊂ VQ. Define the
S-adeles of Q as
AS = AQ,S =Yp∈S
Qp ×Yp6∈S
Zp.
For an arbitrary S, the space AS of S-adeles is locally compact (in the Tikhonov
product topology) as the product of a finite product of locally compact spaces Qp,
p ∈ S by an infinite product of compact spaces Zp, p 6∈ S. Since we have A =SS AS
[17, Theorem 2.15], the adele group A is locally compact (i.e. any neighborhood of a
point contains a compact neighborhood of this point). A fundamental system of open
neighborhoods of zero in A is the following set (see [17, Theorem 2.17])
(2.11)
Yp∈S
Up ×Yp6∈S
Zp,
where S is a finite subset of VQ that contains ∞ and Up is an open set in Qp containing
0 ∈ Qp for p ∈ S.
infinite sequences
The non-Archimedean part eA of the adele ring A is defined (see [31]) as the set of
x′ = (x2, . . . xp, . . . ), where xp ∈ Qp,
p = 2, 3, . . . ,
and there exists a prime number P = P (x′) such that xp ∈ Zp for p ≥ P .
p → xp, n → ∞, for any p ∈ VQ, where VQ is defined by (2.2);
The sequence of adeles {x(n), n ∈ N} converges to the adele x (x(n) → x, n → ∞) if
(1) x(n)
(2) there exists N such that x(n)
Since the adele group A is a locally compact commutative group, it possesses the
Haar measure which will be denoted by dx, where x = (x∞, x2, . . . , xp, . . . ). The Haar
measure dx can be expressed in terms of the measures dxp on the groups Qp as follows:
p − xp ∈ Zp for all n ≥ N.
(2.12)
where
dx = dx∞ dx2 · · · dxp · · · ,
Z 1
0
ZZp
dx∞ = 1,
dxp = 1,
p ∈ VQ.
Here formula (2.12) is understood in the following sense: if
f (x) = f∞(x∞)f2(x2)· · · fp(xp)
is a cylindrical function, then
ZA
f (x) dx =ZR
f∞(x∞) dx∞ZQ2
f2(x2) dx2 · · ·ZQp
fp(xp) dxp.
Any additive character χ(x) on the adelic ring has the form
χ(x) = χ0(ax),
x = (x∞, x2, . . . xp, . . . ) ∈ A,
8
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
for some a = (a∞, a2, . . . ap, . . . ) ∈ A (see [24, Ch. III,§1.5,(3)]). Here
(2.13)
χp(ap),
a ∈ A,
χ0(a) = Yp∈VQ
where χ∞(a∞) = e2πia∞, and χp(ap) is defined by (2.6), p = 2, 3, . . . . It is clear that
for any a ∈ A there exists a prime number P = P (a) such that χp(ap) = 1 for p ≥ P ,
i.e., in fact, the product (2.13) is finite. In other words, χ0(a) = exp(cid:0)2πiσ(a)(cid:1), where
σ(a) =Pp∈VQ
Definition 2.3. (see [24, Ch. III,§2.1]) Let S(R) be the real Schwartz space of tempered
test functions. The space of Bruhat -- Schwartz adelic test functions S(A) consists of finite
linear combinations of elementary functions of the form
ap (mod 1).
(2.14)
ϕ(x) = Yp∈VQ
ϕp(xp),
x ∈ A,
where the factors ϕp(xp) are such that:
(i) ϕ∞(x∞) ∈ S(R);
(ii) ϕp(xp) ∈ D(Qp), p = 2, 3, . . . ;
(iii) there exists P = P (ϕ) such that ϕp(xp) = Ω(xpp) for all p > P (the number
P (ϕ) is called the parameter of finiteness of an elementary function ϕ).
In view of condition (iii) the space of test functions S(A) admits a natural represen-
tation
(2.15)
S(A) = lim
m∈VQ\∞
indS [m](A),
where S [m](A) is the subspace of all test functions with the parameter of finiteness m,
m ∈ VQ \ ∞. The representation (2.15) equips the space S(A) with the inductive limit
topology.
The Bruhat -- Schwartz adelic test functions are continuous on A. The Bruhat -- Schwartz
space of adelic distributions S ′(A) was studied in [31], [22].
The spaces of test functions and distributions connected with the non-Archimedean
part of adeles eA we denote by S(eA) and S ′(eA), correspondingly (see [31]).
The Fourier transform of ϕ ∈ S(A) is defined by the formula
χ0(ξx)ϕ(x) dx,
ξ ∈ A,
where χ0 is defined by (2.13). It is clear that F [S(A)] = S(A). If f ∈ S ′(A), then
def
= ZA
bϕ(ξ) = F [ϕ](ξ)
hF [f ], ϕi def
= hf, F [ϕ]i,
∀ ϕ ∈ S(A),
and F [f ] ∈ S ′(A).
For f, g ∈ L2(A) we have
(2.16)
where
(f, g) = (F [f ], F [g]),
f2 = F [f ]2,
(f, g) =ZA
f (x)g(x) dx
is the scalar product of the functions f and g in L2(A).
It is clear that the space S(A) is dense in L2(A).
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
9
3. Real and p-adic MRAs and wavelet bases
3.1. Real MRA.. Now we recall the definitions of the real and p-adic multiresolution
analysis.
Definition 3.1. (for example, see [53, 1.2]) A collection of closed spaces Vj ⊂ L2(R),
j ∈ Z, is called a multiresolution analysis (MRA) in L2(R) if the following axioms hold
(a) Vj ⊂ Vj+1 for all j ∈ Z;
Vj is dense in L2(R);
Vj = {0};
(b) Sj∈Z
(c) Tj∈Z
(3.4)
(3.5)
where
(3.6)
(d) f (·) ∈ Vj ⇐⇒ f (2·) ∈ Vj+1 for all j ∈ Z;
(e) there exists a function φ ∈ V0 such that the system {φ(· − n), n ∈ Z} is an
orthonormal basis for V0.
The function φ from axiom (e) is called refinable or scaling. One also says that a
MRA is generated by its scaling function. The function φ is a solution of a special kind
of functional equations which are called refinement equations. Their solutions are called
refinable functions.
According to the standard MRA-scheme (see, e.g., [53, §1.3]) for each j, we define
(wavelet spaces)
(3.1)
It is easy to see that
Wj = Vj+1 ⊖ Vj,
j ∈ Z.
f ∈ Wj ⇐⇒ f (2·) ∈ Wj+1,
for all
j ∈ Z
and Wj ⊥ Wk, j 6= k. Taking into account axioms (b) and (c), we obtain
(3.2)
L2(R) =Mj∈Z
Wj = V0 ⊕(cid:16)Mj∈Z+
Wj(cid:17),
where Z+ = {0} ∪ N.
It is well known that the Haar refinement equation has the following form
(3.3)
φH(t) = φH(2t) + φH(2t − 1),
t ∈ R.
Its solution (the characteristic function χ[0,1](t) of the unit interval [0, 1])
φH(t) = χ[0,1](t) =(cid:26) 1,
0,
t ∈ [0, 1],
t /∈ [0, 1],
,
t ∈ R,
generates the Haar MRA. In the framework of the Haar MRA one can construct the
well-known Haar wavelet basis in L2(R)
ψH
jn(t) = 2j/2ψH(cid:0)2jt − n(cid:1),
ψH(t) =
0 ≤ t < 1
2,
2 ≤ t < 1,
t 6∈ [0, 1),
1,
−1,
0,
1
t ∈ R,
j ∈ Z, n ∈ Z,
= χ[0, 1
2 )(t) − χ[ 1
2 ,1)(t),
t ∈ R,
is called the Haar wavelet function (whose dyadic shifts and delations form the Haar
basis (3.5)).
10
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Using the second decomposition (3.2), instead of the Haar wavelet basis (3.5) in L2(R),
(3.7)
one can consider the following wavelet basis
φH(t − n) ∈ V0,
jn(t) = 2j/2ψH(cid:0)2jt − n(cid:1) ∈ Wj,
ψH
n ∈ Z,
j ∈ Z+, n ∈ Z,
t ∈ R,
where the wavelet function ψH is given by (3.6). Basis (3.7) will be called modified Haar
basis.
3.2. p-adic MRA.. In the p-adic case a "natural" set of shifts for Qp is the following
(see [39] and [57]):
(3.8)
Ip = {a ∈ Qp : {a}p = a} = {a =
a−γ
pγ + · · · +
a−1
p
: aj ∈ Fp, −γ ≤ j ≤ −1}.
In [57], similarly to Definition 3.1, the following definition was introduced.
Definition 3.2. ([57]) A collection of closed spaces Vj ⊂ L2(Qp), j ∈ Z, is called a
multiresolution analysis (MRA) in L2(Qp) if the following axioms hold
(a) Vj ⊂ Vj+1 for all j ∈ Z;
Vj is dense in L2(Qp);
Vj = {0};
(b) Sj∈Z
(c) Tj∈Z
(d) f (·) ∈ Vj ⇐⇒ f (p−1·) ∈ Vj+1 for all j ∈ Z;
(e) there exists a function φ ∈ V0 such that the system {φ(· − a), a ∈ Ip} is an
orthonormal basis for V0.
It follows immediately from axioms (d) and (e) that the functions pj/2φ(p−j · −a),
a ∈ Ip, form an orthonormal basis for Vj, j ∈ Z.
According to the standard scheme (see, e.g., [53, §1.3]) for the construction of MRA-
based wavelets, for each j, we define a space Wj (wavelet space) as the orthogonal
complement of Vj in Vj+1, i.e.,
(3.9)
It is not difficult to see that
Vj+1 = Vj ⊕ Wj,
j ∈ Z.
f ∈ Wj ⇐⇒ f (p−1·) ∈ Wj+1,
(3.10)
and Wj ⊥ Wk, j 6= k. Taking into account axioms (b) and (c), we obtain
(3.11)
j ∈ Z
for all
In view of (3.9) and axiom (a), we have
(3.12)
L2(Qp) =Mj∈Z
Vj = V0 ⊕(cid:16) M0≤k≤j−1
Wj = V0 ⊕(cid:16)Mj∈Z+
Wj(cid:17).
Wk(cid:17) j ∈ N.
If we now find a finite number of functions ψν ∈ W0, ν ∈ A, such that the system
{ψν(x − a) : a ∈ Ip, ν ∈ A} forms an orthonormal basis for W0, then, due to (3.10),
(3.11), the system
{pj/2ψν(p−j · −a) : a ∈ Ip, j ∈ Z, ν ∈ A},
is an orthonormal basis for L2(Qp). Such functions ψν, ν ∈ A, are called wavelet
functions and the corresponding basis is called a wavelet basis.
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
11
In [32], the following conjecture was proposed: to construct a p-adic analog of the
real Haar MRA, we can use the following refinement equation
(3.13)
φ(x) =
p−1Xr=0
φ(cid:16)1
p
x −
r
p(cid:17),
x ∈ Qp,
whose solution (a refinable function) φ is the characteristic function Ω(cid:0)xp(cid:1) of the unit
disc. The above refinement equation is natural and reflects the fact that the unit disc
B0 = {x : xp ≤ 1} is the union of p mutually disjoint discs of radius p−1:
B0 =
p−1[r=0
B−1(r), where B−1(r) =nx :(cid:12)(cid:12)x − r(cid:12)(cid:12)p ≤
1
po,
r ∈ Fp.
This geometric fact is the result of the ultrametric structure of the p-adic field Qp.
The term p-adic Haar MRA is connected with the fact that for p = 2 the equation
(3.13) is a 2-adic refinement equation
φ(x) = φ(cid:16)1
2
x(cid:17) + φ(cid:16)1
2
x −
1
2(cid:17),
x ∈ Q2,
which is a direct analog of the refinement equation (3.3) generating the Haar MRA and
the Haar wavelet basis (3.5), (3.6) in the real case.
In [57], using the above relation (3.13) as a refinement equation, the Haar MRA was
constructed for p = 2 (for the case p 6= 2, see [3, 8.4]). Here
(3.14)
Vj = span(cid:8)pj/2φ(cid:0)p−j · −a(cid:1) : a ∈ Ip(cid:9),
where φ(x) = Ω(cid:0)xp(cid:1). In contrast to the Haar MRA in L2(R) (which generates only
one wavelet basis (3.5), (3.6)), in the p-adic setting there exist infinity many different
Haar orthogonal bases for L2(Qp) generated by the same MRA. Explicit formulas for
generating p-adic wavelet functions were obtained in [57] for p = 2 and later in [35] for
p 6= 2. Let us recall these results.
Theorem 3.1. ( [35]) The set of all compactly supported wavelet functions is given by
j ∈ Z,
(3.15)
ψµ(x) =
(here supp ψµ ⊂ Bs(0), s ≥ 0), where
(3.16)
ψ(0)
αµ
ν;kψ(0)
ps−1Xk=0
p−1Xν=1
ν (x) = χp(cid:16) ν
p
ν (cid:16)x −
x(cid:17)Ω(cid:0)xp(cid:1),
are Kozyrev's wavelet functions, s = 0, 1, 2, . . . , and
k
ps(cid:17), µ ∈ F×
p = {1, 2, . . . , p − 1},
ν ∈ F×
p ,
x ∈ Qp,
αµ
ν;k =
(3.17)
=
m=0 e−2πi
−p−sPps−1
−p−2sPps−1
m=0Pps−1
− ν
p +m
ps
kσµmuµµ,
n=0 e−2πi
− ν
p +m
ps
k
1−e
2πi
1−e
2πi
µ−ν
p
µ−ν
p +m−n
ps
σνmuνµ,
µ = ν,
µ 6= ν,
σµm = 1, uµν are entries of an arbitrary unitary (p− 1)× (p− 1) matrix U; µ, ν ∈ F×
p ;
k = 0, 1, . . . , ps − 1.
12
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Corollary 3.1. ( [57]) Let p = 2. For every s = 0, 1, 2, . . . the function
(3.18)
ψ(x) =
2s−1Xk=0
αkψ(0)(cid:16)x −
k
2s(cid:17),
is a compactly supported wavelet function (supp ψ ⊂ Bs(0)) for the Haar MRA if and
only if
(3.19)
αk = 2−s
γre−iπ 2r−1
2s k,
2s−1Xr=0
where the wavelet function ψ(0) := ψ(0)
1
such that γr = 1; k, r = 0, 1, . . . , 2s − 1.
is given by (3.16), γr ∈ C is an arbitrary constant
According to the general wavelet theory, all dilations and shifts of wavelet functions
(3.18), (3.19) or (3.15), (3.17) form a p-adic orthonormal Haar wavelet basis in L2(Qp):
(3.20)
ψk;ja(x) = pj/2ψk(p−jx − a),
k ∈ F×
p , j ∈ Z, a ∈ Ip.
In particular, Kozyrev's wavelet basis in L2(Qp) generated by wavelet functions (3.16)
is the following
ψ(0)
k;ja(x) = pj/2ψ(0)
k (p−jx − a)
The wavelet functions (3.16) can be expressed in terms of the refinable function φ(x) =
(3.21)
Ω(cid:0)xp(cid:1):
= pj/2χp(cid:16) k
p
ψ(0)
k (x) =
(p−jx − a)(cid:17)Ω(cid:0)p−jx − ap(cid:1),
p−1Xr=0
p φ(cid:16)1
p(cid:17),
e2πi kr
x −
p
r
k ∈ F×
p ,
x ∈ Qp.
k ∈ F×
p , j ∈ Z, a ∈ Ip.
It was proved in [1] that there are no orthogonal MRA based wavelet bases except for
those described in Theorem 3.1 and Corollary 3.1.
To construct adelic wavelet bases, we will need the p-adic wavelet bases that contain
the function φ(x). Therefore, taking into account the second decomposition (3.11),
instead of the p-adic wavelet basis (3.20) in L2(Qp), we consider the following wavelet
basis
(3.22)
φ(x − a) ∈ V0,
ψk;ja(x) = pj/2ψk(p−jx − a) ∈ Wj,
a ∈ Ip,
k ∈ F×
p , j ∈ Z+, a ∈ Ip,
p , are given by (3.15) -- (3.17). In particular, we
where the wavelet functions ψk, k ∈ F×
have the following wavelet basis
(3.23)
ψ(0)
k;ja(x) = pj/2χp(cid:16)k
p
φ(x − a) ∈ V0,
(p−jx − a)(cid:17)Ω(cid:0)p−jx − ap(cid:1) ∈ Wj, k ∈ F×
The fact that bases (3.22) and (3.23) contain the refinable function φ(x) = Ω(cid:0)xp(cid:1) plays
a crucial role in our construction (see below Sec. 5). Bases (3.22), (3.23) will be called
modified Haar type bases.
Taking into account the second decomposition (3.11), we obtain the following state-
ment.
a ∈ Ip,
p , j ∈ Z+, a ∈ Ip.
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
13
p):
(3.24)
and I j
p
where dx0
Proposition 3.1. The restriction of basis (3.23) on Zp constitutes an orthonormal basis
in L2(Zp, dx0
, k ∈ F×
p , a ∈ I j
p, j ∈ Z+,
p = {a ∈ Ip : pja ∈ Zp}.
Proof. It is easy to see that the restriction of Kozyrev's wavelets (3.21) is the following
φ(x) = Ω(cid:0)xp(cid:1),
k;ja(x) = pj/2ψk(p−jx − a)Ω(cid:0)p−jx − ap(cid:1)(cid:12)(cid:12)Zp
eψ(0)
p = dxp(cid:12)(cid:12)Zp
(p−jx − a)(cid:17)Ω(cid:0)p−jx − ap(cid:1)(cid:12)(cid:12)(cid:12)Zp
k;ja(x) = pj/2χp(cid:16)k
eψ(0)
p (p−jx − a)(cid:1)Ω(cid:0)p−jx − ap(cid:1),
pj/2χp(cid:0) k
=
k;ja(x)(cid:12)(cid:12)Zp 6= 0 only if p−jx−ap ≤ 1. Let a = a−γp−γ +· · ·+a−1p−1
For j ∈ Z+, we have eψ(0)
and x =P∞
p−jx − a =(cid:0)x0p−j + · · · + xj−1p−1(cid:1) −(cid:0)a−γp−γ + · · · + a−1p−1(cid:1) +
k ∈ Fp, j ∈ Z+, a ∈ I j
p,
j ≤ −1, a = 0,
j ≤ −1, a 6= 0.
k=0 xkpk. Then we have
Since p−jx − a ∈ Zp, the latter relation implies that x0p−j + · · · + xj−1p−1 = a−γp−γ +
· · · + a−1p−1. Hence pja ∈ Zp.
pj/2φ(x),
0,
∞Xk=j
xkpk−j.
Let us calculate the scalar product
p
k′
p
(p−j ′
χp(cid:16)k
(p−jx − a)(cid:17)χp(cid:16) −
x − a′p(cid:1) dx,
k′;j ′a′(cid:1)
(cid:0)eψ(0)
k;ja,eψ(0)
= p(j+j ′)/2ZZp
x − a)(cid:17)×
×Ω(cid:0)p−jx − ap(cid:1)Ω(cid:0)p−j ′
k, k′ ∈ Fp, j, j′ ∈ Z+, a, a′ ∈ Ip.
Here (cid:0)eψ(0)
k′;j ′a′(cid:1) 6= 0 only if p−jx − ap ≤ 1 and p−j ′x − a′p ≤ 1. Just as above,
k;ja,eψ(0)
p , we have supp Ω(cid:0)p−jx − ap(cid:1), supp Ω(cid:0)p−j ′x − a′p(cid:1) ∈ Zp. Now, taking into
k′;j ′a′(cid:1) = δkk′δjj ′δaa′, where δss′ is the Kronecker symbol.
(cid:0)eψ(0)
k;ja,eψ(0)
we conclude that a, a′ ∈ Ip are such that pja ∈ Zp, pj ′a′ ∈ Zp. Thus for a ∈ I j
p,
a′ ∈ I j ′
account orthonormality of wavelet functions ψ(0)
Using Proposition 3.1 and Theorem 3.1, one can prove the following statement.
k′;j ′a′ in Qp, one can conclude that
(cid:3)
k;ja, ψ(0)
Proposition 3.2. The restriction of the basis (3.22) to Zp constitutes in L2(Zp, dx0
p)
the orthonormal basis
(3.25)
φ(x) = Ω(cid:0)xp(cid:1),
eψk;ja(x) = pj/2ψk(p−jx − a)(cid:12)(cid:12)Zp
p , are given by (3.15) -- (3.17), and the set I (j)
where the wavelet functions ψk, k ∈ F×
defined in Remark 3.1.
Remark 3.1. We note that the set I (j)
a ∈ I (j)
p are nonzero, is finite and its description is similar to the description of the set
I j
p, though more complicated. For other wavelet bases, we will have a similar situation.
p , for which the restricted functions eψk;ja(x),
is
p
, k ∈ F×
p , a ∈ I (j)
p , j ∈ Z+,
14
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Consider the orthogonal projection
P [0] : L2(Qp, dxp) 7→ L2(Zp, dx0
p),
L2(Qp, dxp) ∋ f (x) 7→ (P [0]f )(x) = f (x)Ω(cid:0)xp(cid:1) ∈ L2(Zp, dx0
In our case, the projections of modified Haar bases (3.22) and (3.23) to the subspace
L2(Zp, dx0
p is the restriction of the Haar measure
dxp on Qp to Zp. We note that the projections of some basis elements become zeros.
p) are also bases in L2(Zp, dx0
p), where dx0
p).
Proposition 3.3. The restriction of the p-adic Haar MRA (given by formula (3.14))
to the space L2(Zp) consists of the spaces
(3.28)
where Vj(cid:12)(cid:12)Zp
eVj = Vj(cid:12)(cid:12)Zp
= span(cid:8)pj/2φ(cid:0)p−j · −a(cid:1), x ∈ Zp : a ∈ Ip(cid:9),
= eV0 = span(cid:8)φ(·)(cid:9) for all j ≤ −1, φ(x) = Ω(cid:0)xp(cid:1), and
j ∈ Z+,
Here according to (3.9) -- (3.11),
(3.26)
given by
(3.27)
(3.29)
where
eV0 ⊂ eV1 ⊂ eV2 ⊂ · · · .
L2(Zp) = eV0M(cid:16)Mj∈Z+fWj(cid:17),
p , a ∈ Ip(cid:9),
: k ∈ F×
fWj = eVj+1 ⊖eVj
= span(cid:8)pj/2ψk(p−jx − a)(cid:12)(cid:12)Zp
and the wavelet functions ψk, k ∈ F×
use the wavelet functions (3.23).
p , are given by (3.15) -- (3.17). In particular, we can
j ∈ Z+,
4. Basis in the space L2(A, dx)
4.1. Infinite tensor product of Hilbert spaces. We recall [52] (see also [12, Ch.1,
Hn of Hilbert spaces Hn, n ∈ N.
§2.3]) the definition of infinite tensor product He = Ne,n∈N
Fix the sequence
(4.1)
e = (e(n))∞
n=1, e(n) ∈ Hn, ke(n)kHn = 1,
called a stabilizing sequence. Denote by E the set of all stabilizing sequences. Fix an
orthonormal basis (o.n.b) (e(n)
1 , n ∈ N.
Let Λ be the set of multi-indices α = (αn)n∈N, αn ∈ N, n ∈ N such that αn = 1 for
sufficiently big n depending on α.
By definition, the o.n.b. of the space ⊗e,n∈NHn with the stabilizing sequence e consists
k )k∈N in the Hilbert space Hn such that e(n) = e(n)
of all vectors (eα)α∈Λ of the following form
(4.2)
eα = e(1)
α1 ⊗ e(2)
α2 ⊗ · · · ⊗ e(n)
αn ⊗ e(n+1) ⊗ · · · , α ∈ Λ
where αn = 1 for sufficiently big n depending on α. An arbitrary element f of the space
He has the following form:
(4.3)
fαeα, with kfk2
f2
α < ∞
f =Xα∈Λ
He =Xα∈Λ
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
15
and the scalar product (f, g)He of two vectors f, g ∈ He has the following form:
(4.4)
fαgα.
(f, g)He =Xα∈Λ
It will often be convenient for us (see [12, Ch.1,§ 2.3]) to represent the set Λ as the
union of disjoint sets, each consisting of "finite" sequences. Namely, for an α ∈ Λ let
ν(α) denote the minimal m = 1, 2, . . . such that αm+1 = αm+2 = · · · = 1. Let
(4.5)
Λn = {α ∈ Λ : ν(α) = n}, n ∈ N.
Obviously, ΛnT Λm = ∅ (n 6= m) and Λ =S∞
n=1 Λn.
Example 4.1. ([12, Ch.1, § 2.3, example 1]) Let L2(Xk, µk), k ∈ N be the space of
square integrable complex functions on the measurable space Xk with a probability
measure µk. Choose the stabilizing sequence e = (e(k))∞
k=1, where e(k)(x) ≡ 1, x ∈ Xk,
k ∈ N. In this case we have
Theorem 4.1. The following two spaces are isomorphic:
(4.6)
Oe, k∈N
L2(Xk, µk) ∼= L2(cid:16)Yk∈N
Xk,⊗k∈Nµk(cid:17).
Example 4.2. If in the previous example the first m measures µk are not necessarily
probability, i.e., µk(Xk) = ∞, then we get following statement.
Theorem 4.2. The following two spaces are isomorphic:
(4.7)
Xk,⊗m
L2(cid:16) mYk=1
k=1µk(cid:17) ⊗(cid:16) ∞Oe, k=m+1
L2(Xk, µk)(cid:17) ∼= L2(cid:16)Yk∈N
Xk,⊗k∈Nµk(cid:17).
Remark 4.1. Let us consider in the infinite tensor product space He = Ne,n∈N
Hn the
infinite tensor product A = ⊗n∈NAn of bounded operators An acting in the space
Hn, n ∈ N. By definition, an operator A acts on the total set of well-defined vectors
f = ⊗n∈Nfn in the space He (see Lemma 4.1) in the following way:
Af := ⊗n∈NAnfn = A1f1 ⊗ · · · ⊗ Amfm ⊗ · · · ,
(4.8)
if the latter expression is well-defined in He. In the sequel we shall use the projections
P [m] : He → H (m), where the subspaces H [m] ⊂ He are defined on the total family of
vectors f = ⊗n∈Nfn as follows:
He ∋ f = ⊗n∈Nfn 7→ P [m]f := ⊗m
n=1fn ⊗ e(m+1) ⊗ e(m+2) ⊗ · · · ∈ H [m],
and
H [m] =
Hn ⊗ e(m+1) ⊗ e(m+2) ⊗ · · ·
mOn=1
We shall denote by A(m) the projections P [m]AP [m] of the operator A in He onto the
space H [m], i.e. define A(m) := P [m]AP [m]. It is clear that
A(m) =
mOn=1
An ⊗On>m
Idn,
where Idn is the identity operator in Hn. We have the strong convergence A(m) →
A as m → ∞ on a suitable set of vectors f = ⊗n∈Nfn.
It is sufficient to estimate
k(A − A(m))fkHe. For details, see [12, Ch. I, §2.7.].
16
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
As usual, we denote by AX the restriction of an operator A acting in a Hilbert space
H to the invariant subspace X ⊂ H.
4.2. Complete von Neumann product of infinitely many Hilbert spaces. The
complete von Neumann tensor product H = ⊗k∈NHk of Hilbert spaces Hk, k ∈ N is by
definition the orthogonal sum of the spaces He ([52], see also [12, Ch.1,§ 2.10])
(4.9)
H = Me∈E/∼
To be more precise, fix the space He = Ne,n∈N
He
over all possible equivalence classes E/∼ of stabilizing sequences e.
Hn. We define the vector f = ⊗n∈Nf (n),
where f (n) ∈ Hn, as the week limit (if it exists) in He of the vectors
f [m] = f (1) ⊗ · · · ⊗ f (m) ⊗ e(m+1) ⊗ e(m+2) ⊗ · · ·
(4.10)
as m → ∞. Since the set span {eα : α ∈ Λ} is dense in He, the week limit of the
vectors f [m] exists if and only if: 1) the norms kf [m]kHe are uniformly bounded with
respect to m = 1, 2, . . . , and 2) limm→∞(f [m], eα)He exists for each α ∈ Λ. The following
statements are proved in [12, Ch.I,§ 2.10].
Lemma 4.1. The strong limit of vectors (4.10) exists in He, as m → ∞, if and only
n=q(f (n), e(n))Hn (q = 1, 2, . . . ) converge to finite
if the product Q∞
numbers, and we haveQ∞
limm→∞ f [m] exists if and only if for some q = 1, 2, . . . the product Q∞
Corollary 4.1. If f (n) in (4.10) are taken to be unit vectors, then the strong limit
n=q(f (n), e(n))Hn
converges to a finite nonzero number.
Definition 4.1. (see [12, Ch. I,§ 2.10, Theorem 2.9]) Consider the set E of all stabi-
lizing sequences e = (e(n))∞
n=1 of the form (4.1). A stabilizing sequence l ∈ E is said to
be equivalent to a stabilizing sequence e ∈ E (l ∼ e), if each strong limit
n=1 kf (n)kHn = 0 whenQ∞
n=1 kf (n)kHn and Q∞
n=q(f (n), e(n))Hn = 0 for each q.
⊗ l(2)′
⊗ · · · = lim
⊗ e(m+2) ⊗ e(m+2) ⊗ · · · ,
exists in He, where (l(n)′)∞
k=1, "diluted" by the vectors e(n), i.e.
each l(n)′ is equal either to l(n) or to e(n). The relation ∼ is an equivalence relation and
we denote by E/∼ the set of all equivalent classes of E.
⊗ · · · ⊗ l(m)′
k=1 is the sequence (l(n))∞
l(1)′
l(1)′
m→∞
Theorem 4.3. Two infinite tensor products He = Ne,n∈N
Hn and Hl = Nl,n∈N
responding to two equivalent stabilizing sequences e = (e(n))∞
isomorphic. For e 6∼ l, the spaces He and Hl are orthogonal.
4.3. On some subspaces of the infinite tensor products. Let Xn be some sub-
spaces in the Hilbert spaces Hn, n ∈ N and let (e(n)
k )k∈Z be an orthonormal basis in Hn
such that the orthonormal basis in Xn is (e(n)
k )k∈N, and let (e(n))n∈N be the stabilizing
sequence e(n) = e(n)
n=1 and l = (l(n))∞
n=1 are
Hn cor-
(4.11)
1 , n ∈ N. Note that
e(n) = e(n)
Consider two spaces He and Hl
1 ∈ Xn ⊂ Hn, n ∈ N.
e(X), l ∈ N, where
Hn = H1 ⊗ H2 ⊗ · · · ⊗ Hl ⊗ Hl+1 ⊗ Hl+2 ⊗ · · · ,
He = Oe,n∈N
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
17
(4.12)
Hl
e(X) =
lOk=1
Hk ⊗
∞Oe,k=l+1
Xk = H1 ⊗ H2 ⊗ · · · ⊗ Hl ⊗ Xl+1 ⊗ · · · .
In the particular case Xn = Ce(n) we have
Hl
e(X) = H1 ⊗ H2 ⊗ · · · ⊗ Hl ⊗ Ce(l+1) ⊗ Ce(l+2) ⊗ · · · ,
l ∈ N.
Lemma 4.2. For arbitrary subspaces Xn in Hn and a stabilizing sequence (e(n))n∈N with
properties (4.11), we have
Proof. The basis in the space He is the following (see (4.2)):
(4.14)
eα = e(1)
α1 ⊗ e(2)
α2 ⊗ · · · ⊗ e(n)
where Λn is defined by (4.5), and the basis in the space Hl
el
α = e(1)
α1 ⊗ e(2)
α2 ⊗ · · · ⊗ e(l)
αl ⊗ e(l+1)
Hl
e(X).
He =[l≥1
αn ⊗ e(n+1) ⊗ · · · , α ∈ Λ = [n∈N
αl+k ⊗ e(l+k+1) ⊗ · · · ,
αl+1 ⊗ · · · ⊗ e(l+k)
α ∈ Λl = [k∈N∪{0}
e(X) is
Λn,
Λl,k,
(4.13)
(4.15)
where
(4.16)
Λl,k = {α ∈ Λ : αi ∈ Z, 1 ≤ i ≤ l, αi ∈ N, l + 1 ≤ i ≤ l + k, αl+k+i = 1, i > 1}.
Obviously He ⊇ Hl
e(X) for all l ∈ N, hence He ⊇Sl≥1 Hl
We show that He ⊆ Sl≥1 Hl
form (4.14) are contained in the family of vectors el
take in (4.15) l = n and k = 0, we obtain all vectors eα of the form (4.14).
e(X). It is sufficient to show that all vectors eα of the
α of the form (4.15). Indeed, if we
(cid:3)
e(X).
4.4. Infinite tensor product of Hilbert spaces L2(Qp, dxp). Using the results of
the previous section we can define the infinite tensor product
He(Q) = Oe, p∈VQ
L2(Qp, dxp)
of Hilbert spaces L2(Qp, dxp) with an arbitrary stabilizing sequence e = (e(p))p∈VQ e(p) ∈
L2(Qp, dxp).
Let {e(∞)
αp }αp∈Ip be arbitrary orthonormal bases in L2(Q∞, dx∞) and
L2(Qp, dxp), respectively, where I∞ and Ip are the corresponding multi-indices. Then
the orthonormal basis in the space He(Q) is the following:
(4.17)
α∞ }α∞∈I∞ and {e(p)
α∞ ⊗Op≤m
eα = e(∞)
e(p), α = (α∞, α2, α3, . . . ) ∈ Λ,
αp ⊗Om<p
e(p)
where (e(p))p∈VQ, e(p) ∈ L2(Qp, dxp), is some stabilizing sequence and Λ is defined below.
Definition 4.2. Define Λ as the set of multi-indices α = (αp)p∈VQ, αp ∈ Ip such that
e(p)
αp = e(p) for sufficiently big p depending on α.
18
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Fix the stabilizing sequence of the form
(4.18)
e = (e(p))p∈VQ
Using Lemma 4.2, for the stabilizing sequence (4.18), we obtain the following descrip-
e(p)(xp) = φp(xp) = Ω(cid:0)xpp(cid:1) ∈ L2(Zp, dx0
p),
p ∈ VQ.
tion of the space L2(A, dx).
Lemma 4.3. We have
(4.19)
L2(A, dx) = He(Q) = Oe, p∈VQ
L2(Qp, dxp).
L2(Ap, µp) =Nq≤p
L2(Qq) ⊗Nq>p
Proof. It is sufficient to use Lemma 4.2 and set Xp = L2(Zp), Hp = L2(Qp), Hp
e =
p ∈ VQ \ ∞ (see (4.27)), µp is the restriction on Ap of the Haar measure dx on A.
4.5. Some remarks about the stabilizations. Consider the space (4.16)
L2(Zq), p ∈ VQ, where Ap = Q∞ ×Qp′≤p Qp′ ×Qp′>p Zp′,
He(Q) = Oe, p∈VQ
L2(Qp, dxp),
(cid:3)
(4.20)
where the stabilizing sequence e = (e(p))p∈VQ has the special form e(p)(xp) = φp(xp) ∈
L2(Qp, dxp). We show that the sequence e and M j
e for j ∈ Z are not equivalent (see
VQ
Definition 4.1), where
(4.21)
M j
VQ
e := (M j
p φ(p))p∈VQ
and M j
p φ(p)(xp) := p−j/2φ(p)(pjxp).
Indeed, we have
(φ(p), M j
p φ(p))L2(Qp) =ZQp
Set c∞ = (φ(∞), M j
∞φ(∞))L2(Qp∞), then we get
∞ (e, M j
c−1
VQ
(φ(p), M j
e) = Yp∈VQ\∞
pj/2,
φ(p)(xp)p−j/2φ(p)(pjxp)dx =(cid:26) p−j/2,
p φ(p))L2(Qp) =
(cid:16)Qp∈VQ\∞ p(cid:17)−j/2
(cid:16)Qp∈VQ\∞ p(cid:17)j/2
j ≥ 0,
j ≤ −1.
,
,
j ≥ 0,
j ≤ −1,
e) = 0 hence the sequences e
Thus in both cases the product is divergent, i.e., (e, M j
VQ
and M j
VQ
e are not equivalent!
αp in the space L2(Qp), i.e., e = (ψ(p)
A similar argument holds if we take as the stabilizing sequence some elements of
αp )p∈VQ. In this case we immediately
αp )p∈VQ and
Kozurev's basis ψ(p)
get (ψ(p)
M j
VQ
The above considerations force us to consider the following space
αp )L2(Qp) = 0 for all j ∈ Z. Hence the sequences e = (ψ(p)
αp )p∈VQ are not equivalent.
αp , M j
e = (M j
p ψ(p)
p ψ(p)
to be sure that the operator M j
(4.22)
(4.23)
(4.24)
He,Z(Q) = ⊕j∈ZHM j
VQ is well-defined. Further we set
e(Q)
VQ
e(Q)
VQ
Wk = HM k
k=−∞Wk = ⊕j
Vj = ⊕j
k=−∞HM k
VQ
e(Q).
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
19
Theorem 4.4. The collection of closed spaces Vj ⊂ He,Z(Q), j ∈ Z, defined by (4.24)
is a multiresolution analysis in He,Z(Q), i.e., the following properties hold:
(a) Vj ⊂ Vj+1 for all j ∈ Z;
Vj is dense in He,Z(Q);
Vj = {0};
(b) Sj∈Z
(c) Tj∈Z
(d) f (·) ∈ Vj ⇐⇒ f (M j
(e) there exists a basis (ei)i∈I in the space W0 such that (M k
VQ·) ∈ Vj+1 for all j ∈ Z;
VQei)i∈I is a basis in the
space Wk, k ∈ Z.
4.6. Complete von Neumann product of Hilbert spaces L2(Qp, dxp). It is nat-
ural to consider also the complete von Neumann product (see (4.25)) of Hilbert spaces
L2(Qp, dxp), p ∈ VQ
(4.25)
H(Q) = Me∈E/∼
He(Q),
where He(Q) is defined by (4.20). Perhaps this space could be useful in further devel-
opment of analysis on the adele space A. The space He,Z(Q) defined by (4.22) certainly,
contains the space L2(A) and is roughly speaking an infinite direct sum of non isomor-
phic copies of the space similar to L2(A).
4.7. Basis on L2(A, dx). We can construct a basis in L2(A, dx) using description (4.19)
of the space L2(A, dx) in Lemma 4.3. Without using this description we can proceed as
follows.
Ap (see (4.27)), then we
realize L2(Ap, µp) as the infinite tensor product of Hilbert spaces (see (4.30)) and use
the following considerations.
We present A as the union of some subgroups Ap: A =Sp∈VQ
Xn, X =Sn∈N Xn. Set µn = µXn, νn = µXn\Xn−1, n ≥ 2,
L2(X, µ) = ⊕n∈NL2(Xn \ Xn−1, νn) = [n∈N
Let (X, µ) be some measurable space and let X be the union of the measurable sets
ν1 = µ1 and X0 = ∅. In
this case we have
For any prime p we denote by p− and p+ the previous and the following primes.
L2(Xn, µn).
(4.26)
Define the subgroup Ap of the group A as follows:
p be the restriction of the Haar measure dxp on Qp to the subgroup Zp, let
dµp be the restriction to the subgroup Ap of the measure dx on A, and let dνp be the
restriction of the measure dx to the subset Ap \ Ap−. Then we have
(4.28)
dµp = dx∞ ⊗(cid:16) ⊗p′≤p dxp′(cid:17) ⊗(cid:16) ⊗p′>p dx0
p′(cid:17).
−, νp′)(cid:17) = [p∈VQ
L2(A, dx) = L2(Ap, µp) ⊕(cid:16)Mp′>p
\ Ap′
L2(Ap′
Lemma 4.4. For an arbitrary prime p we have the following description:
(4.29)
L2(Ap, µp).
and using (4.26) we get.
Ap = Q∞ ×Yp′≤p
Qp′ ×Yp′>p
Zp′,
p ∈ VQ.
(4.27)
We have A =Sp∈VQ
Let dx0
Ap.
20
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
To construct a basis in the space L2(A, dx) using the latter description, it is sufficient
to construct a basis in the space L2(Ap, µp). The latter space L2(Ap, µp) is the infinite
tensor product of the spaces L2(Qp, dxp) and L2(Zp, dx0
p):
(4.30)
L2(Ap, µp) = L2(Q∞, dx∞)⊗(cid:16)Op′≤p
L2(Qp′, dxp′)(cid:17)⊗(cid:16) Oe,p′>p
L2(Zp′, dx0
p′)(cid:17).
Remark 4.2. Since the measures dx0
p on Zp are probability measures, decomposition
(4.30) allows us to use an explicit description of the basis in the infinite tensor product
⊗k∈NL2(Xk, µk) of Hilbert spaces L2(Xk, µk) with probability measures µk, k ∈ N (for
example, see [12]) in order to construct a basis in the spaces L2(Ap, µp) and L2(A, dx).
Suppose that {e(∞)
α∞ }α∞∈I∞, and {e(p)
αp}αp∈Ip are arbitrary orthonormal bases in the
spaces L2(Q∞) and L2(Qp), respectively, I∞ and Ip are the corresponding indices. Fix
the stabilizing sequence (e(p))p∈VQ, where e(p) is some element of the basis {e(p)
αp}αp∈Ip
for all p ∈ VQ such that e(p)(x) ∈ L2(Zp, dx0
p). We can construct a basis in the space
L2(Ap, µp) for all p ∈ VQ using decomposition (4.30). In such a way we can construct a
basis in L2(A, dx) using (4.29). As it was mentioned before, using Lemma 4.3 and the
basis in the infinite tensor product we obtain.
Theorem 4.5. The vectors
eα = e(∞)
(4.31)
α∞ ⊗Oq≤m
e(q)
αq ⊗Om<q
e(q), α ∈ Λ = [m∈VQ
Λm,
form the orthonormal basis in the space L2(A, dx), where Λ is the set of multi-indices
α = (αp)p∈VQ, such that e(p)
αp = e(p) for sufficiently big p depending on α (see (4.2)),
Λp = {α ∈ Λ : νa(α) = p}, p ∈ VQ, and νa(α) denote the minimal p ∈ VQ such that
e(q)
αq = e(q) for q > p.
Corollary 4.2. The vectors
(4.32)
eeα =Oq≤m
e(q)
αq ⊗Om<q
e(q), α ∈eΛ,
form an orthonormal basis in the space L2(eA, dx), where eΛ is the set of multi-indices
αp = e(p) for sufficiently big p depending on α, i.e.,
α = (αp)p∈VQ\{∞} such that ψ(p)
eΛ = [m∈VQ\∞eΛm and eΛm = {α ∈eΛ : νa(α) = m}, m ∈ VQ \ ∞.
(4.33)
5. Adelic wavelet bases generated by tensor product of
one-dimensional wavelet bases
To construct an adelic wavelet basis, we apply the above scheme from Subsec. 4.7
to the one-dimensional bases (3.5), (3.22), (3.25). In particular, one can use the Haar
wavelet bases (3.5), (3.23), (3.24). Let
j∞a∞(x∞) = ψH
j∞a∞(x∞), α∞ = (j∞, a∞) ∈ I∞ := (Z, Z);
α∞ (x∞) = ψ(∞)
ψ(∞)
(5.1)
be the real Haar wavelet basis (3.5) in L2(R),
(5.2)
αp (xp) = ψ(p)
ψ(p)
αp (xp) = ψ(p)
ψ(p)
kp;jpap(xp), αp = (kp, jp, ap) ∈ I+
p , Z+, Ip),
ap (xp) = φ(p)(xp − ap), αp = (0, 0, ap) ≡ ap ∈ Ip;
p := (F×
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
21
be the p-adic modified Haar wavelet basis (3.22) in L2(Qp), where Ip is defined by (3.8),
and let
(5.3)
αp (xp) = ψ(p)
ψ(p)
αp (xp) = ψ(p)
ψ(p)
kp; jpap(xp), αp = (kp, jp, ap) ∈ I+
0 (xp) = φ(p)(xp), αp = (0, 0, 0) = 0;
p = (F×
p , Z+, Ip),
be the p-adic Haar wavelet basis (3.25) in L2(Zp), which is the restriction of the basis
(5.2) to Zp. We recall that the restrictions of some basis elements (5.2) are equal to
zero (see Remark 3.1 and Proposition 3.1). Here and in what follows, for a stabilizing
sequence we take (4.18).
Hence we have the following bases in L2(Qp) and in L2(Zp), respectively,
p [ Ip;
p [{0} ⊂ Ip,
p
(0)
p
(5.4)
(5.5)
:= I+
αp }αp∈I
ψ(p)
αp , αp ∈ Ip = I+
ψ(p)
αp , αp ∈ I(0)
where {ψ(p)
αp }αp∈Ip on L2(Zp). Now, using Lemma 4.3,
we obtain the following orthonormal wavelet basis in the space L2(A, dx) (see (4.31)):
Ψα = Ψ(bk,bj,ba)(x) = ψ(∞)
is the projection of {ψ(p)
α∞ ⊗Oq≤m
eΨα = eΨ(bk,bj,ba)(x) = O2≤q≤m
φ(q), α ∈ Λ = [m∈VQ
and the orthonormal wavelet basis in the space L2(eA, dx) (see (4.32))
φ(q), α ∈eΛ,
with the stabilization φ = (φ(p))p∈VQ\∞. Herebk = (kp)p,bj = (jp)p,ba = (ap)p (see (5.2)).
Remark 5.1. We note that in [20, Sec.4, formula (4.1)] the basis in the space L2(A, dx)
is constructed as follows
αq ⊗Om<q
αq ⊗Om<q
(5.6)
ψ(q)
ψ(q)
Λm.
ψα,β = ψ∞
n0(x∞) Yp=2,3,...
ψαp,βp(xp),
where ψ∞
n0(x∞) and ψαp,βp(xp) are orthonormal eigenfunctions (of harmonic oscillators)
in the real and p-adic cases, respectively, where ψαp,βp(xp) = Ω(cid:0)xpp(cid:1) for almost all p .
Let us introduce the adelic set of shifts IA and dilations ZA:
(5.7)
IA =(cid:8)ba = (a∞, a2, . . . , ap, . . . ) : a∞ ∈ I∞ = Z, ap ∈ Ip,
ZA =(cid:8)bj = (j∞, j2, . . . , jp, . . . ) : j∞ ∈ Z, jp ∈ Z,
Set m(ba) = min{p : aq = 0 for all q > p} and m(bj ) = min{p : jq = 0 for all q > p}.
On the space L2(A) we define the operators of shifts Tba,ba ∈ ZA, and multi-dilation
bj,bj ∈ ZA which are defined as the infinite tensor product of one-dimensional operators
there exists n depending onba such that ap = 0 for all p > n(cid:9).
there exists n depending onbj such that jp = 0 for all p > n(cid:9).
M
(4.21):
(5.8)
(5.9)
M
bj = M −j∞
∞ ⊗ O2≤q≤m
M jq
q ⊗Oq>m
Idq,
22
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
where m = m(bj ) and Idq is the identity operator on L2(Qq). We suppose that the
p in the spaces L2(Q∞) and L2(Qp) (on the functions f (∞) ∈
operators M j
L2(Q∞) and f (p) ∈ L2(Qp)) act as follows
∞ and M j
(M j
∞f (∞))(x∞) = 2−j/2f (∞)(2−jx∞),
(M j
p f (p))(x) = p−j/2f (p)(pjxp).
Let f (x) = ⊗q∈VQfq(xq) ∈ L2(A), where f (∞) ∈ L2(Q∞), f (q) ∈ L2(Qq), q ≥ 2, and
f (q) = φ(q) for almost all q > m. Then
(cid:0)Tbaf(cid:1)(x)
bjf(cid:1)(x)
(cid:0)M
(5.10)
(5.11)
(cid:0)M
bjTbaf(cid:1)(x)
def
= f (x −ba) = f∞(x∞ − a∞) ⊗O2≤q
fq(xq − aq),
def
= 2−j∞/2f (∞)(2−j∞x∞)⊗
⊗(cid:16) O2≤q≤m
q−jq/2f (q)(qjqxq)(cid:17) ⊗(cid:16)Om<q
f (q)(xq)(cid:17),
def
= 2−j∞/2f (∞)(2−j∞x∞ − a∞)⊗
f (q)(xq)(cid:17).
Now one can obtain the adelic wavelet basis functions Ψα given in (5.5) by all adelic
(5.12)
⊗(cid:16) O2≤q≤m
shifts and dilations of wavelet functions
q−jq/2f (q)(qjqxq − aq)(cid:17) ⊗(cid:16)Om<q
In the latter relation, we assume that m = m(bj ) = m(ba ).
(kq, 0 0)(xq) ⊗Om<q
bjTbaΨ(bk,0,0)(cid:1)(x),
(0,0)(x∞) ⊗Oq≤m
Ψ(bk,0,0) := ψ(∞)
(5.13)
Namely,
ψ(q)
φ(q)(xq).
Ψα(x) =: Ψ(bk,bj,ba)(x) =(cid:0)M
where bk = (0, k2, . . . , km, 0, 0, . . . ), kq ∈ F×
q ; ba = (a∞, a2, . . . , am, 0, 0, . . . ), a∞ ∈ Z,
aq ∈ Iq;bj = (j∞, j2, . . . , jm, 0, 0, . . . ), j∞ ∈ Z, jq ∈ Z+; m ∈ VQ \∞. We stress that here
m(ba) = m(bj) = m(bk) = m.
The wavelet systems (5.5) described above is of considerable interest and can be
useful in various situations. Nevertheless, they do not possess all the advantages of
one-dimensional wavelet bases, in particular, the localization property, which is of great
value for applications. In the one-dimensional case, the real Haar basis functions with
large indices j∞ ∈ Z and the p-adic Haar basis functions with large indices jp ∈ Z+, p =
2, 3, . . . , have small supports. Thus the support of the multidimensional basis function
may be large along one or several directions and small along some other directions. To
avoid these drawbacks, we shall use a different approach.
Its main idea is using the
tensor product of the MRAs generating these bases instead of the tensor product of
available wavelet bases.
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
23
6. Separable adelic MRA generated by tensor product of
one-dimensional MRAs
Using the idea of constructing separable multidimensional MRA by means of the ten-
sor product of one-dimensional MRAs (suggested by Y. Meyer [50] and S. Mallat [46], [47]
(see, e.g., [53, §2.1]), we construct wavelet bases for the space L2(A, dx).
We start with some general facts. Let we have a collection of closed subspaces V (k)
,
j ∈ Z, in a Hilbert space Hk, k = 1, 2, . . . , m, having the properties (a), (b) and (c) of
the Definition 3.1: (a) V (k)
,
j+1 for all j ∈ Z, k = 1, 2, . . . , m; (b) Hk =Sj∈Z V (k)
V (k)
j = {0}. We use the following notations for the finite tensor product:
j ⊂ V (k)
j
j
(6.1)
H [m] =
Hk, V (k)
j+1 = V (k)
,
j ∈ Z,
k = 1, 2, . . . , m,
mOk=1
V [m]
j+1 =
j
j ⊕ W (k)
mOk=1
V (k)
j+1 =
mOk=1
(V (k)
j ⊕ W (k)
j
) = V [m]
j ⊕ W [m]
j
,
V [m]
j =
mOk=1
V (k)
j
,
W [m]
j =
(i1,i2,...,im;j)∈{1,2}m\{1,1,...,1}
M
i2,j ⊗ · · · ⊗ Z (m)
im,j,
W(i1,i2,...,im;j) = Z (1)
i1,j ⊗ Z (2)
W(i1,i2,...,im;j),
Z (k)
1,j = V (k)
j
, Z (k)
2,j = W (k)
j
, k = 1, 2, . . . , m.
(c) Tj∈Z
where
(6.2)
and
(6.3)
where
(6.4)
(6.5)
For the infinite tensor product we keep the similar notations:
H = He = Oe,n∈N
V [m]
j+1 =
mOk=1
Hn, H [m] = ⊗m
k=1HkO e(m+1) ⊗ e(m+2) ⊗ · · · , m ∈ N.
V (k)
j+1 ⊗ e(m+1) ⊗ e(m+2) ⊗ · · ·
=
mOk=1
j ⊕ W (k)
(V (k)
j
) ⊗ e(m+1) ⊗ e(m+2) ⊗ · · · = V [m]
j ⊕ W [m]
j
,
mOk=1
V [m]
j =
W [m]
j =
V (k)
j ⊗ e(m+1) ⊗ e(m+2) ⊗ · · · ,
M
i2,j ⊗ · · · ⊗ Z (m)
W(i1,i2,...,im;j)
(i1,i2,...,im)∈{1,2}m\{1,1,...,1}
W(i1,i2,...,im;j) = Z (1)
i1,j ⊗ Z (2)
im,j ⊗ e(m+1) ⊗ e(m+2) ⊗ · · · .
Define the space Vj as an appropriate limit of the spaces V [m]
j
(6.6)
Vj = span(V [m]
j
: m ∈ N).
24
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
j }j∈Z of the space Hn, n ∈ N satisfy the following
Then the subspaces Vj, j ∈ Z of the space He = ⊗e,n∈NHn defined by (6.6) satisfy the
(aH) Vj ⊂ Vj+1 for all j ∈ Z;
j
following properties:
j = {0} in Hn.
j+1 for all j ∈ Z and n ∈ N;
is dense in Hn;
Theorem 6.1. Let subspaces {V (n)
properties:
j ⊂ V (n)
(a) V (n)
(b)Sj∈Z V (n)
(c)Tj∈Z V (n)
(bH)Sj∈Z Vj is dense in He = ⊗e,n∈NHn;
(cH)Tj∈Z Vj = {0}.
j+1 for all n ∈ N, we conclude that V [m]
: m ∈ N) ⊂ span(V [m]
j+1 : m ∈ N) = Vj+1.
(bH) It is clear that for any fixed m ∈ N the spaceSj∈Z V [m]
mOk=1
Hk ⊗ e(m+1) ⊗ e(m+2) ⊗ · · ·
Proof. (aH) Since V (n)
j ⊂ V (n)
(6.4)), hence Vj = span(V [m]
H [m] =
j
j
j ⊂ V [m]
j+1, m ∈ N (see
is dense in the space
hence all vectors eα, α ∈ Λm of the basis (4.14) are contained in ∪j∈ZV [m]
. To finish
the proof, we note that all elements of the basis eα, α ∈ Λ in He = ⊗e,n∈NHn are given
by (4.14), where Λ = ∪m∈NΛm.
j
(cH) 1. Let us suppose that for any m ∈ N, we have
(6.7)
2. In this case we get
∩j∈Z V [m]
j = {0}.
(6.8)
∩j∈Z Vj = ∩j∈Zspan(V [m]
j
1. To prove (6.7), we shall use (c). We set W (n)
: m ∈ N) ⊂ span(∩j∈ZV [m]
j+1 ⊖ V (n)
j = V (n)
j
j
: m ∈ N) = {0}.
for j ∈ Z and n ∈ N.
Then we get
(6.9)
Hn = ⊕k∈ZW (n)
k
,
and V (n)
j = ⊕j−1
k=−∞W (n)
k
.
k
Let (e(n)
αn )αn∈I (n)
be an orthonormal basis in Wk, then by construction, (e(n)
αn )αn∈I (n) is an
orthonormal basis in Hn, where I (n) =Sk∈Z I (n)
k . Define Λ(I) (see Definition (4.2)) as
the set of multi-indices α = (αn)n∈N, αn ∈ I (n) such that e(n)
αn = e(n) for sufficiently big
n depending on α, here e = (e(n))n∈N is a stabilizing sequence. Using (6.9) and (6.4),
we get
V [m]
j =(cid:16) mOk=1
V (k)
j (cid:17) ⊗ e(m+1) ⊗ e(m+2) ⊗ · · ·
so the basis in the space V [m]
j
=
mOk=1(cid:16) j−1Mn=−∞
W (k)
n (cid:17) ⊗ e(m+1) ⊗ e(m+2) ⊗ · · · ,
can be chosen as eα, α ∈ Λm,j(I), where
α1 ⊗ e(2)
αm ⊗ e(m+1) ⊗ · · · ,
α2 ⊗ · · · ⊗ e(m)
eα = e(1)
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
25
Λm,j(I) =(cid:16)α ∈ Λ(I) : α = (αn)n∈N,
Note thatTj∈Z Λm,j(I) = limj→−∞(cid:0)Sj−1
(compare also with Lemma 6.1 below, where Λ1,j(I) = (−∞, j − 1] ∩ Z). Indeed, for
f ∈ V [m]
)(cid:17).
(cid:1) = ∅, this imply (6.7)
(α1, . . . , αm) ∈ (
j−1[k=−∞
j−1[k=−∞
j we have
I (1)
k , . . . ,
I (m)
k
f = Xα∈Λm,j (I)
then
kfk2 = lim
If f ∈Tj∈Z V [m]
j
2. To prove (6.8), using (6.4) we get
cα2 < ∞.
k
k=−∞ I (1)
k=−∞ I (m)
k , . . . ,Sj−1
cαeα with kfk2 = Xα∈Λm,j (I)
j→−∞ Xα∈Λm,j (I)
j ⊗ e(m+1) ⊗ e(m+2) ⊗ · · · , m ∈ N,
cα2 =Xα∈∅
cα2 = 0.
V [m]
j = ⊗m
V [1]
j = V (1)
V [2]
j = V (1)
V [m]
j = V (1)
k=1V (k)
j ⊗ e(2) ⊗ · · · ⊗ e(m) ⊗ e(m+1) ⊗ · · · ,
j ⊗ V (2)
j ⊗ · · · ⊗ e(m) ⊗ e(m+1) ⊗ · · · ,
j ⊗ · · · ⊗ V (m)
j ⊗ V (2)
j ⊗ e(m+1) ⊗ · · · .
(e) = span(e(k), V (k)
) for k − 1 ∈ N, then we get
Denote by V (k)
j
j
j
span(V [1]
, V [3]
, V [2]
, V [2]
j ) = V (1)
j ) = V (1)
, k = 2, . . . , m) = V (1)
j
j
span(V [1]
span(V [k]
j
Since V (n)
j
Vj = span(V [k]
So
j
j
(e) = span(e(n), V (n)
: k ∈ N) = V (1)
Vj =\j∈Z
⊂\j∈Z(cid:0)V (1)
span(V [k]
j ⊗ Oe,n≥2
j
\j∈Z
j
j
j
j
j ⊗ V (2)
j ⊗ V (2)
j ⊗ V (2)
) ⊂ V (n)
j ⊗ Oe,n≥2
(e) ⊗ e(3) ⊗ · · · ⊗ e(m) ⊗ e(m+1) ⊗ · · · ,
(e) ⊗ V (3)
(e) ⊗ · · · ⊗ e(m) ⊗ e(m+1) ⊗ · · · ,
(e) ⊗ · · · ⊗ V (m)
(e) ⊗ V (3)
for j ≤ 0, we conclude that
j ⊗ Oe,n≥2
V (n)
V (n)
j
j
j Oe,n≥2
(e) ⊂ V (1)
(e) = V (1)
0
j
j
(e) ⊗ e(m+1) ⊗ · · · .
V (n)
0
: k ∈ N) ⊂
V (n)
0 (cid:1) =\j∈Z(cid:0)V (1)
j (cid:1) ⊗ Oe,n≥2
V (n)
0 = {0}.
.
(cid:3)
Remark 6.1. We can give another definition of the spaces Vj, j ∈ Z, namely, for fixed
n ∈ N denote by Vj(n), j ∈ Z the following family of spaces
: m ∈ N, m ≥ n).
(6.10)
Vj(n) = span(V [m]
For this system of spaces, an analog of Theorem 6.1 holds. Moreover, in the proof of
(cH) we have the following formula (see the end of the proof of the Theorem 6.1):
j
\j∈Z
Vj(n) ⊂\j∈Z(cid:16) nOk=1
V (k)
j (cid:17) ⊗ Oe,k>n
V (k)
0 = {0}.
26
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
We would like to present the following elementary lemma.
Lemma 6.1. Let we have the decomposition of the Hilbert space H = ⊕k∈ZWk and let
k=−∞Wk, thenTj∈Z Vj =
(eα)α∈Ik be an orthonormal basis in the space Wk. Set Vj := ⊕j−1
{0}.
Proof. By construction, (eα)α∈J is an orthonormal basis in H, where J =Sk∈Z Jk. We
have for any f ∈ H
(6.11)
f =Xα∈J
cαeα =Xk∈Z Xα∈Jk
If f ∈ Vn = ⊕n−1
k=−∞Wk for all n ∈ Z, using (6.11), we conclude that
α! < ∞.
c2
cαeα! and kfk2 =Xk∈Z Xα∈Jk
n−1Xk=−∞ Xα∈Jk
cα2! = 0.
(cid:3)
kfk2 = lim
n→−∞
Let {V (ν)
j }j∈Z be one-dimensional MRA (see Definitions 3.1, 3.2), and let φ(ν) be its
refinable function, ν ∈ VQ. We introduce the spaces (see (6.4))
(6.12)
V [m]
V (ν)
j ⊗ φ(m+) ⊗ · · · ⊂ L2(A, dx), m ∈ VQ,
j = Oν∈{∞,2,3,...,m}
j ∈ Z.
Now we introduce in L2(A, dx) the delation operator M j, which is defined by its
projection M j
(m) on any subspace V [m]
j
(for details, see Remark 4.1):
(6.13)
M j
(m) = M −j
∞ ⊗ O2≤q≤m
M j
q ⊗Oq>m
Idq, m ∈ VQ \ ∞,
(m)f and
(m)Tba)f similarly to (5.11) and (5.12), respectively, where IA is given by (5.7). As
where Idq is the identity operator in L2(Qq). For f ∈ L2(A) we define M j
(M −j
before, we assume that m = m(ba ).
Theorem 6.2. Let {V (ν)
and let φ(ν) be its refinable function, ν ∈ VQ. Let
(6.14)
j }j∈Z be the one-dimensional MRA (see Definitions 3.1, 3.2),
Φ(x) = φ(∞)(x∞) ⊗ φ(2)(x2) ⊗ · · · ⊗ φ(p)(xp) ⊗ · · · ,
x ∈ A,
and
(6.15)
Then
(6.16)
Vj = span(V [m]
j
: m ∈ VQ)
= span{(cid:0)M −j
(m)TbaΦ(cid:1)(·) : ba ∈ IA, m(ba) = m ∈ VQ},
Vj = Oe;ν∈VQ
V (ν)
j ⊂ L2(A, dx),
j ∈ Z+,
j ∈ Z.
(where e = (φ(ν))ν∈VQ\∞ is the stabilization sequence (4.18) and the subspaces (6.15)
satisfy the following properties:
(a) Vj ⊂ Vj+1 for all j ∈ Z;
(b) ∪j∈ZVj is dense in L2(A, dx);
(c) ∩j∈ZVj = {0};
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
27
0
j ⇐⇒(cid:0)M −1
(m)f(cid:1)(·) ∈ V [m]
j+1 for all j ∈ Z, for any m ∈ VQ;
basis for V0.
Proof. 1. For each ν ∈ VQ, the system of functions {φ(ν)(·− aν)}aν ∈Iν is an orthonormal
basis of the space V (ν)
(see axioms (e) in Definitions 3.1, 3.2). Using the construction of
(d) f (·) ∈ V [m]
(e) the function Φ ∈ V0 is such that the system {Φ(x−ba) : ba ∈ IA} is an orthonormal
Sec. 4 and taking into account that any shiftba = (a∞, a2, . . . , ap, . . . ) ∈ IA is finite and
φ(ν) ∈ V (ν)
that the system {Φ(x −ba) : ba ∈ IA} is an orthonormal basis in V0 and
V0 = Oe; ν∈VQ
Using definition (5.12) we get for m(ba) = m.
(cid:0)M −j
(m)TbaΦ(cid:1)(x) = 2j/2φ(∞)(2jx∞ − a∞)⊗
0 ⊂ L2(Qν, dxν) is a stabilization sequence (4.18) in L2(A, dx), we conclude
V (ν)
0 ⊂ L2(A, dx).
(6.17)
νj/2φ(ν)(ν−jxν − aν)(cid:17) ⊗ φ(m+)(xm+) ⊗ · · · ∈ V [m]
⊗(cid:16) mOν=2
for anyba ∈ IA, m ∈ VQ.
Since {νj/2φ(ν)(ν−jxν − aν) : aν ∈ Iν} is the basis in V (ν)
j ∈ N, using (6.15) and (6.17), we conclude that
j
j ⊂ Vj
and φ(ν) ∈ V (ν)
0 ⊂ V (ν)
j
for
Vj = Oe; ν∈VQ
(6.15), (6.13) imply that f ∈ V [m]
0
Taking into account axioms (d) in Definitions 3.1, 3.2, one can see that Definitions
if and only if M −j
(m)f (·) ∈ V [m]
j
, j ∈ N, m ∈ VQ.
Thus representation (6.16) and properties (e) and (d) hold.
2. Due to [12, Ch. I, §2.2, Theorem 2.2], we immediately obtain that property (a)
holds for j ∈ N. For any j ∈ Z the property (a) holds by Theorem 6.1. Here we
give an independent proof. According to (6.15), V0 = span{Φ(cid:0)x −ba(cid:1) : ba ∈ IA}, where
Φ(cid:0)x −ba(cid:1) = φ(∞)(x∞ − a∞) ⊗ φ(2)(x2 − a2) ⊗ · · · ⊗ φ(p)(xp − ap) ⊗ φ(p+)(xp+) ⊗ · · · .
Applying the refinement equation (3.13) to any factor of the above product and taking
into account that the numbers aν
ν ∈ Iν for all aν ∈ Iν and r = 1, 2, . . . , ν − 1,
ν = 2, 3, . . . , we conclude that V0 ⊂ V1. By Definition (6.15) of the spaces Vj and
property (d), this yields property (a).
3. The property (b) holds by Theorem 6.1. Here we give an independent proof. It is
clear that any element g ∈ L2(A, dx) can be approximated by a finite linear combination
of the basis elements Ψα(x) = Ψ(bk;bj ba)(x) of L2(A, dx) given by (5.5). At the same time
any element Ψα(x) = Ψ(bk;bj ba)(x) is a finite product of the one-dimensional basis elements
(5.1) -- (5.4):
ν + r
ν , aν
V (ν)
j ⊂ L2(A, dx),
j ∈ N.
α∞ (x∞) = ψ(∞)
ψ(∞)
αq (xq) = ψ(q)
ψ(q)
αq (xq) = ψ(q)
ψ(q)
j∞a∞(x∞) ∈ L2(Q∞),
(kq;jqaq)(xq) ∈ L2(Qq),
(kq;jqaq)(xq) ∈ L2(Zq),
αp ∈ Ip = (F×
p = (F×
αp ∈ I(0)
α∞ ∈ (Z, Z),
p , Z+, Ip)[ Ip,
p , Z+, Ip)[{0}.
In turn, it follows from the completeness property for each MRA {V (ν)
j }j∈Z (see axioms
(b) in Definitions 3.1, 3.2) that any element f∞ ∈ L2(Q∞) and fν ∈ L2(Qν) can be
28
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
approximated by a finite linear combination of the functions 2j/2φ(∞)(2j · −a∞), a∞ ∈
I∞, and νj/2φ(ν)(ν−j · −aν), aν ∈ Iν, respectively, for sufficiently large j (j ∈ Z+),
ν ∈ {2, 3, . . . , p}. Hence any basis element Ψα(x) = Ψ(bk;bj ba)(x) ∈ L2(A, dx) can be
approximated by an element of the space Vj for some sufficiently large j (j ∈ Z+). This
implies that the element g can be also approximated by an element of the space Vj for
some j, i.e., the collection of all the spaces Vj, j ∈ Z, is dense in L2(A, dx).
4. The property (c) holds by Theorem 6.1. Here we give an independent proof. Let
us prove that the intersection of all Vj does not contain a nonzero element. Assume
that there exists an element g ∈ L2(A, dx) such that g ∈ Vj for all j ∈ Z and kgk 6= 0.
According to Sec. 5, the wavelet functions Ψα = Ψ(bk;bj ba) ∈ L2(A, dx) given by (5.5) form
an orthonormal basis in L2(A, dx), then g can be represented as g =Pα∈Λ cαΨα ∈ Vj for
all j ∈ Z, where the coefficients cα are independent of j. Here α = (αp)p∈VQ = (bk;bj ba),
Consider a basis element Ψα = Ψ(bk;bj ba),bj = (j∞, j2, . . . , jm, 0, 0, . . . ). Let us take the
index jα = min(0, j∞, j2, j3, . . . , jm). For any wavelet function given by the first formula
in (5.2), we have ψ(p)
, p = 2, 3, . . . . In
view of (3.1), (3.2) we have ψ(∞)
for j < j∞. Consequently, the basis
element Ψα = Ψ(bk;bj ba) ⊥ Vj, j < jα, i.e., g ⊥ Ψα. In such a way we prove that g ⊥ Ψα
for all indices α ∈ Λ. The fact that the system {Ψα : α ∈ Λ} is an orthonormal basis in
L2(A, dx) contradicts the assumption kgk 6= 0.
p , Z+, Ip)S Ip, p = 2, 3, . . . .
kp;jpap ⊥ V (p)
α∞ = ψ(∞)
α∞ ∈ (Z, Z), αp ∈ Ip = (F×
, jp ∈ Z+, and V (p)
j∞ ⊥ V (∞)
j ⊂ V (p)
αp = ψ(p)
The collection of the spaces Vj, j ∈ Z+, satisfying conditions (a) -- (e) of Theorem 6.2
will be called the adelic separable MRA in the space L2(A). The function (6.14) is a
refinable function of this MRA.
(cid:3)
0
0
j
Next, following the standard finite-dimensional scheme (see [18, Ch. 10.1], [53, §2.1]),
we define the wavelet space Wj as the orthogonal complement of Vj in Vj+1:
(6.18)
Wj = Vj+1 ⊖ Vj,
j ∈ Z+.
It follows from properties (a) and (b) of Theorem 6.2 that (cf. (3.2) and (3.11))
Since according to (3.1), (3.2), (3.9) -- (3.11), we have V (ν)
Z (ν)
1,j = V (ν)
obtain
, setting
, using (6.19), (6.20), and taking into account (6.17), we
j ⊕ W (ν)
j+1 = V (ν)
2,j = W (ν)
and Z (ν)
j
j
j
Vj+1 = Oν∈VQ
V (ν)
j+1 = Oν∈VQ(cid:0)V (ν)
= Vj ⊕(cid:16)Mp∈VQ
j (cid:1)
j ⊕ W (ν)
M
(i∞,i2,...,ip)∈{1,2}♯(p)+1\{1,1,...,1}
W(i∞,i2,...,ip),j(cid:17),
j ∈ Z+,
L2(A) = V0 ⊕Mj∈Z+
Wj.
Let us present some evident formulas in ⊗n
k=1Hk. Namely, we have
(6.19)
and similarly,
(6.20)
(a11 ⊕ a12) ⊗ · · · ⊗ (an1 ⊕ an2) = M(i1,...,in)∈{1,2}n
k=1ank) = M(i1,...,in)∈{1,...,m}n
(⊕m
k=1a1k) ⊗ · · · ⊗ (⊕m
a1i1 ⊗ · · · ⊗ anin,
a1,i1 ⊗ · · · ⊗ an,in,
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
29
where we set ♯(p) = ♯{2, 3, . . . , p} (♯(A) is the number of elements of the set A), and
(6.21)
= W [p]
(i∞,i2,...,ip),j = Z (∞)
i∞,j ⊗ Z (2)
i2,j ⊗ · · · ⊗ Z (p)
ip,j ⊗ φ(p+) ⊗ · · · ,
W [p]
bip,j
bip = (i∞, i2, . . . , ip) ∈ {1, 2}♯(p)+1 \ {1, 1, . . . , 1}. Thus the space Wj given by (6.18) is a
direct sum of infinite number of subspaces (6.21).
Thus we have
It is clear that an orthonormal basis for the space
bip=(i∞,i2,...,ip)∈{1,2}♯(p)+1\{1,1,...,1}
Mbip∈{1,2}♯(p)+1\{1,1,...,1}
W [p]
bip,j(cid:17).
W [p]
(i∞,i2,...,ip),0
(6.22)
where
L2(A) = V0 ⊕Mj∈Z+
W0 =Mp∈VQ
Wj = Mj∈Z+(cid:16)Mp∈VQ
M
i∞,0 ⊗ Z (2)
W [p]
bip,0
= W [p]
(i∞,i2,...,ip),0 = Z (∞)
i2,0 ⊗ · · · ⊗ Z (p)
is formed by the shifts (with respect toba ∈ IA) of the functions
i2 ⊗ · · · ⊗ ϑ(p)
:= Ψbk,(i∞,i2,...,ip) = ϑ(∞)
i∞ ⊗ ϑ(2)
(6.23)
Ψbk,bip
ip ⊗ φ(p+) ⊗ · · · ,
ip,0 ⊗ φ(p+) ⊗ · · · ,
kp−
kp−
1 = ψ(ν)
1 = ψH , ϑ(∞)
2 = φH (see (3.5), (3.4)); ϑ(ν)
According to (6.22), an orthonormal Haar wavelet basis in L2(A) is formed by all
where all factors in this product starting from a certain prime p constitute a stabilization
sequence (φ(ν))ν>p; ϑ(∞)
kν , kν ∈ F×
ν ,
ϑ(ν)
(i∞, i2, . . . , ip) ∈ {1, 2}♯(p)+1 \ {1, 1, . . . , 1}.
2 = φ(ν), (see (3.22), (3.25) or (3.23), (3.24)), ν = 2, 3, . . . , p; p = 2, 3, . . . ; bip =
shiftsba ∈ IA of the refinable function (6.14)
φ(∞)(x∞)φ(2)(x2)· · · φ(p−)(xp−)φ(p)(xp)φ(p+)(xp+)· · · ,
and all shiftsba ∈ IA and dilations j ∈ Z+ of wavelet functions (6.23):
ψ(∞)(x∞)ψ(2)
k2 (x2)· · · ψ(p−)
kp (xp)φ(p+)(xp+)· · · ,
ψ(∞)(x∞)ψ(2)
k2 (x2)· · · ψ(p−)
(xp+)· · · ,
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
φ(∞)(x∞)ψ(2)
k2 (x2)· · · ψ(p−)
(xp+)· · · ,
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
ψ(∞)(x∞)φ(2)(x2)· · · φ(p−)(xp−)φ(p)(xp)φ(p+)(xp+)· · · ,
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
φ(∞)(x∞)φ(2)(x2)· · · φ(p−)(xp−)ψ(p)
kp (xp)φ(p+)(xp+)· · · ,
(xp−)ψ(p)
(xp−)φ(p)(xp)φ(p+,0)
kp (xp)ψ(p+,0)
(xp−)ψ(p)
(6.24)
kp+
kp+
kp−
(6.25)
Thus, this wavelet basis has the form
where p = 2, 3, . . . ; the real Haar wavelet function ψ(∞)(x∞) and refinable function
φ(∞)(x∞) = Ω(x∞) are given by (3.6) and (3.4), x∞ ∈ Q∞; the r-adic Haar wavelet
functions ψ(r)
kr (xr), kr ∈ F×, are given by formulas (3.15) -- (3.17), the refinable function
φ(r)(xr) = Ω(cid:0)xrr(cid:1), xr ∈ Qr; r = 2, 3, . . . .
wherebip = (i∞, i2, . . . , ip) ∈ {1, 2}♯(p)+1 \ {1, 1, . . . , 1}, bk = (k2, . . . , kp), ks ∈ F×
2, 3, . . . , p; j ∈ Z+;ba = (a∞, a2, . . . , ap) ∈ IA, p ∈ VQ. We stress that here p(ba) = p(j) =
p(bk) = p.
TbaΦ, Ψbk,bip;j ba := (2 · 2 · 3 · · · p)j/2(cid:0)M −j
(p) Tba(cid:1) Ψbk,bip
s , s =
,
30
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Due to Theorem 3.1 and [1], by formula (6.25) all Haar bases generated by adelic
Haar MRA are described.
Corollary 6.1. Let (see (6.23))
(6.26)
i2 ⊗ · · · ⊗ ϑ(m)
im ⊗ φ(m+) ⊗ · · · .
p
eΨbk,bi′
(x′) := eΨbk,(i2,...,im) = ϑ(2)
The vectors (see (6.25))
Tba′Φ,
(6.27)
p;j ba′ := (2 · 3 · · · p)j/2(cid:0)M −j
eΨbk,bi′
form the orthonormal basis in the space L2(eA), wherebi′
{1, . . . , 1}, bk = (k2, . . . , kp), ks ∈ F×
s , s = 2, 3, . . . , m; j ∈ Z+; ba′ = (a2, . . . , ap) ∈ IA,
p ∈ VQ \ ∞. We stress that here p(ba′) = p(j) = p(bk′) = p.
p = (i2, . . . , ip) ∈ {1, 2}♯(p−1)+1 \
(p) Tba(cid:1) Ψbk,bi′,
7. Lizorkin spaces on adeles
7.1. Real and p-adic Lizorkin spaces. Recall some facts from [43] -- [45], [54, 2.], [55,
§25.1]. Consider the following space
Ψ(R) = {ψ(ξ) ∈ S(R) : ψ(j)(0) = 0, j = 0, 1, 2, . . .}, .
where S(R) is the real Schwartz space of tempered test functions. The space of functions
Φ(R) = {φ : φ = F [ψ], ψ ∈ Ψ(R)}.
is called the real Lizorkin space of test functions. The Lizorkin space can be equipped
with the topology of the space S(R), which makes Φ a complete space [54, 2.2.], [55,
§25.1.]. Since the Fourier transform is a linear isomorphism S(R) onto S(R), this space
admits the following characterization: φ ∈ Φ(R) if and only if φ ∈ S(R) is orthogonal
to polynomials, i.e.,
(7.1)
ZR
xnφ(x) dx = 0, n = 0, 1, 2, . . . .
The space Φ′(R) is called the real Lizorkin space of distributions.
According to [4], [3, Ch. 7] the p-adic Lizorkin space of test functions is defined as
Φ(Qp) = {φ : φ = F [ψ], ψ ∈ Ψ(Qp)}, where Ψ(Qp) = {ψ(ξ) ∈ D(Qp) : ψ(0) = 0}. The
space Φ(Qp) equipped with the topology of the space D(Qp) is a complete space.
Lemma 7.1. ( [4]) We have (a) φ ∈ Φ(Qp) iff φ ∈ D(Qp) and
(7.2)
φ(x) dx = 0;
ZQp
N (Qp) ∩ Φ(Qp)
(b) φ ∈ Dl
The topological dual of the space Φ(Qp) is the space of p-adic Lizorkin distributions
Φ′(Qp). The space Φ′(Qp) can be obtained from D′(Qp) by "sifting out" constants
(see [4]).
iff ψ = F −1[φ] ∈ D−N
−l (Qp) ∩ Ψ(Qp).
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
31
7.2. Adelic Lizorkin spaces. Consider the subspace Ψ(A) of the space of test func-
tions S(A) (see Subsec. 2.3) consisting of finite linear combinations of elementary func-
ηp(ξp) (where VQ is defined by (2.2)) with properties (i)− (iii)
of Definition 2.3, where ηp(ξp) = Ω(ξpp) for all p > P ) and such that
tions (2.14) η(ξ) =Qp∈VQ
= 0,
s = 0, 1, 2, . . . ,
(iv)
ds
dξs
∞
η(ξ∞, ξ2, ξ3, . . . , ξr−, ξr, ξr+, . . . , ξP , ξP+, . . . )(cid:12)(cid:12)(cid:12)ξr=0
for all r ∈ {∞, 2, 3, . . . , P}, where P = P (η) is the parameter of finiteness of an elemen-
tary function η.
ηp(ξp) = Ω(ξpp) for all p > P ) and such that
Let Ψ(eA) be the space of test functions S(eA) connected with the non-Archimedean
part of adeles eA (see Subsec. 2.3) consisting of finite linear combinations of elemen-
tary functions η(ξ′) =Qp∈VQ\∞ ηp(ξp) with properties (i) − (iii) of Definition 2.3 (here
(iv∗) η(ξ2, ξ3, . . . , ξr−, ξr, ξr+, . . . , ξP , ξP+, . . . )(cid:12)(cid:12)ξr=0 = 0,
Φ(A) =(cid:8)ζ : ζ = F [η], η ∈ Ψ(A)(cid:9) and Φ(eA) =(cid:8)ζ : ζ = F [η], η ∈ Ψ(eA)(cid:9)
where P = P (η) is the parameter of finiteness of an elementary function η.
are called the adelic Lizorkin spaces of test functions.
r ∈ {2, 3, . . . , P},
Definition 7.1. The spaces
Φ(A) can be equipped with the topology of the space S(A) (see (2.15)). Since the
Fourier transform is a linear isomorphism S(A) onto S(A) (see [24, Ch. III,§2.2]), we
have Φ(A) ⊂ S(A).
The Lizorkin spaces Φ(A) and Φ(eA) admit the following characterizations.
Lemma 7.2. (1) Any elementary function
ζ(x) = ζ∞(x∞) Yp∈VQ\∞
ζp(xp) ∈ Φ(A)
(here ζp(xp) = Ω(xp) for all p > P ) if and only if ζ ∈ S(A) and
(7.3)
xs
rζ(x∞, x2, x3, . . . , xr−, xr, xr+, . . . , xP , xP+, . . . ) dxr = 0,
ZQr
where if r = ∞, then s = 0, 1, 2, . . . , and if r = 2, 3, . . . , P , then s = 0.
(2) Any elementary function
eζ(x′) = Yp∈VQ\∞
ζp(xp) ∈ Φ(eA)
(7.4)
if and only if eζ ∈ S(eA) and
Remark 7.1. In view of condition (iii), similarly to (2.15), the spaces Φ(A) and Φ(eA)
ZQreζ(x2, x3, . . . , xr−, xr, xr+, . . . , xP , xP+, . . . ) dxr = 0,
Φ(eA) = lim
ind Φ[m](eA),
admit a representation in the form
r = 2, 3, . . . , P.
ind Φ[m](A),
m∈VQ\∞
Φ(A) = lim
m∈VQ\∞
(7.5)
32
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
functions with the parameter of finiteness m, m ∈ VQ. The representation (7.5) equip
where Φ[m](A) ⊂ Φ(A) and Φ[m](eA) ⊂ Φ(eA) are subspaces of the corresponding test
the spaces Φ(A) and Φ(eA) with the inductive limit topology.
The spaces Φ′(A) and Φ′(eA) are called the adelic Lizorkin spaces of distributions.
We define the Fourier transform of distributions f ∈ Φ′(A) and g ∈ Ψ′(A) by the
relations
hF [f ], ηi = hf, F [η]i,
hF [g], ζi = hg, F [ζ]i,
∀ η ∈ Ψ(A),
∀ ζ ∈ Φ(A).
(7.6)
(7.7)
By definition, F [Φ(A)] = Ψ(A), F [Ψ(A)] = Φ(A), i.e., (7.6) give well defined objects.
Moreover, we have F [Φ′(A)] = Ψ′(A), F [Ψ′(A)] = Φ′(A).
7.3. Adelic Lizorkin spaces and wavelets. It is well known that p-adic Haar wavelet
functions (3.20) from L2(Qp) satisfy the condition (see [3, Ch. 8])
ZQp
ψk;ja(x) dx = 0,
k ∈ F×
p , j ∈ Z, a ∈ Ip,
where ψk;ja(x) = pj/2ψk(p−jx − a) and wavelet functions ψk, k ∈ F×
Theorem 3.1.
p , are given by
In view of the formula (see [3, Ch. 8] or [39])
p-adic Haar wavelet functions (3.24) from L2(Zp) satisfy the following condition:
a = 0, j ≤ −1,
otherwise,
k ∈ F×
p , j ∈ Z+, a ∈ I j
p,
0,
(cid:0)Ω(xp), ψk;ja(x)(cid:1) =(cid:26) pj/2,
ZZp eψ(0)
k;ja(x) dx =(cid:0)Ω(xp), ψk;ja(x)(cid:1) = 0,
ZZp eψk;ja(x) dx = 0,
is described in Remark 3.1.
(7.8)
(7.9)
where I (j)
p
where I j
functions (3.25) from L2(Zp) we also have
p is defined in Proposition 3.1. It is easy to see that for the p-adic Haar wavelet
k ∈ F×
p , j ∈ Z+, a ∈ I (j)
p ,
φ(x) dx = 1, and Lemma 7.2, we
conclude that the adelic wavelet functions (5.5) and (6.25) belong to the Lizorkin space
Thus, taking into account relations (7.7) -- (7.9),RQp
Φ(eA) only if products (5.6) and (6.27) do not contain the functions φ(p)(xp − ap) as
factors for p ≤ m.
8. Characterization of the adelic Lizorkin spaces in terms of wavelets
To construct adelic wavelet bases in L2(A), we used the real Haar basis (3.5) in L2(R),
jn(t) dt = 0 does not hold for n ∈ N, it
is clear that ψH
jn /∈ Φ(R). Therefore, in contrast to the p-adic case (see [3, 8.14]), the
characterization of the adelic Lizorkin spaces in terms of wavelets is possible only for
jn(t) dt = 0. SinceRR tnψH
For simplicity here and in what follows we will suppose that the adelic wavelets (5.6)
are constructed by using the one-dimensional wavelet basis (3.23) and its restriction
(3.24) to Zp.
for which we haveRR ψH
the space Φ(eA).
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
33
(8.1)
Lemma 8.1. Any test functioneζ ∈ Φ(eA) can be represented in the form of a finite sum
where eΛ is the set of indices (4.33); cα are constants; eΨα(x′) are wavelet functions (5.6)
eζ(x′) =Xα∈ eΛ
cαeΨα(x′),
eΨα = eΨ(bk;bj ba) = O2≤q≤m
x′ = (x2, x3, . . . ) ∈eA,
αq ⊗Om<q
φ(q), α ∈eΛ,
ψ(q)
αq (xq) = φ(q)(xq − aq), ap ∈ Ip, (see (5.2), (5.3)) for
which do not contain the factors ψ(q)
q ≤ m.
Proof. By definition of the space Φ(eA) (see Subsec. 7.2), it is sufficient to prove this
lemma for the case of an elementary function eζ(x′) =Qp∈VQ\∞ ζp(xp), where ζp(xp) =
φ(p)(xp) = Ω(xpp) for all p > P , and P = P (eζ) is the parameter of finiteness of an
elementary functioneζ. It is clear that eζ ∈ L2(eA). Then (8.1) holds. We will prove that
According to Subsec. 7.3, eΨα(x′) ∈ Φ(eA) only if the product (5.6) does not contain
only the finite number of cα 6= 0.
functions φ(p)(xp − ap) as factors for p ≤ m. Now we need to calculate the coefficients
(8.2)
cα = c(bk;bj ba) =(cid:0)eζ(x′),eΨα(x′)(cid:1) = Yp∈VQ\∞(cid:0)ζq(xq), ψ(q)
p =
q , Z+, Iq), are given by (3.23), (3.24), (5.2), (5.3), and among them there are no
kq;jqaq (xq), αq = (kq, jq, aq) ∈ I+
where one-dimensional wavelet functions ψ(q)
(F×
wavelet functions φ(q)(xq − aq).
αq = ψ(q)
αq (xq)(cid:1),
1. Suppose that m > P . Then we have
i.e., cα = 0.
2. Let m < P . Consider the part of product (8.2)
Ym<q≤P(cid:0)ζq(xq), φ(q)(xq)(cid:1) = Ym<q≤P(cid:0)ζ 0
q (xq), φ(q)(xq)(cid:1),
where ζ 0
support in Zq (see Sec. 7.1). Using the Parseval-Steklov theorem, we obtain
q (xq) := ζq(xq)(cid:12)(cid:12)Zq ∈ Φ(Zq) and Φ(Zq) is the Lizorkin space of test functions with
Ym<q≤P(cid:0)ζq(xq), φ(q)(xq)(cid:1) = Ym<q≤P(cid:0)F [ζ 0
q (xq) ∈ Φ(Zq) belongs to one of the spaces Dlq
q ](ξq), F [φ(q)(xq)](ξq)(cid:1).
It is clear that ζ 0
isfies condition (7.2). Then in view of (2.7) and Lemma 7.1, F [ζ 0
and supp F [ζ 0
Nq(Zq), lq ≤ Nq ≤ 0, and sat-
−lq (Zq),
q ] ⊂ B−lq \B−Nq, 0 ≤ −Nq ≤ −lq. At the same time, F [φ(q)](ξq) = φ(q)(ξq).
q ] ∈ Ψ(Qq)∩D−Nq
cα = c(bk;bj ba) =(cid:0)eζ(x′),eΨα(x′)(cid:1) = Y2≤q≤P(cid:0)ζq(xq), ψ(q)
αq (xq)(cid:1)×
αq (xq)(cid:1)Ym<q(cid:0)φ(q)(xq), φ(q)(xq)(cid:1).
× YP <q≤m(cid:0)φ(q)(xq), ψ(q)
ZZq
αq (xq) dxq = 0,
ψ(q)
for P < q ≤ m,
According to (7.7) and (7.8), (7.9), we have
34
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
3. Let m = P . In this case, using the Parseval-Steklov theorem, one can rewrite the
q ](ξq), F [φ(q)(xq)](ξq)(cid:1) = 0,
Thus supp F [ζ 0
m < q ≤ P . That is in this case cα = 0.
product (8.2) in the form
q ] ∩ supp φ(q) = ∅ and, consequently, (cid:0)F [ζ 0
cα = c(bk;bj ba) =(cid:0)eζ(x′),eΨα(x′)(cid:1)
= Y2≤q≤m(cid:0)ζq(xq), ψ(q)
αq (xq)(cid:1) = Y2≤q≤m(cid:0)F [ζq](ξq), F [ψ(q)
αq ](ξq)(cid:1).
(2.7), any function ζq(xq) ∈ Φ(Qq) belongs to one of the spaces Dlq
condition (7.2), i.e., F [ζq] ∈ Ψ(Qq)∩D−Nq
Let ζq ∈ Φ(Qq) for 2 ≤ q ≤ p. According to Definition 7.1, Lemma 7.1, and relation
Nq (Qq) and satisfies
−lq (Qq), and supp F [ζq] ⊂ B−lq\B−Nq , 2 ≤ q ≤ p.
For the wavelet function (5.2),
αq (xq) = ψ(q)
ψ(q)
kq; jqaq (xq)
F [ψ(q)
q
(q−jqxq − aq)(cid:17)Ω(cid:0)q−jqxq − aqq(cid:1),
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17),
= qjq/2χq(cid:16)kq
kq; jqaq ](ξq) = q−jq/2χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
(cid:0)F [ζq](ξq), F [ψ(q)
kq
q
xq ∈ Qq,
jq ∈ Z+.
(8.3)
we have
(8.4)
Taking into account that supp F [ζq] ⊂ B−lq \ B−Nq and using formula (8.4), one can
conclude that
(8.5)
only if q−Nq ≤ ξqq ≤ q−lq and kq
ξq = q−jq(cid:0)ηq − kq
Now we consider the scalar product(cid:0)F [ζq](ξq), F [ψ(q)
αq ](ξq)(cid:1) 6= 0,
q(cid:12)(cid:12)q = qjq+1, one can see that the product (8.5)
αq ](ξq)(cid:1), where ζq ∈ Φ(Zq) and
q(cid:1) and ξqq = qjq(cid:12)(cid:12)ηq − kq
q + qjqξq = ηq ∈ Zq for any 2 ≤ q ≤ p. Since
is nonzero only for finite number of indices jq such that
xq ∈ Zq,
is a wavelet function given by (5.3), p < q ≤ m. Using the identity [64, VII.1], [39]
j ≤ j′,
kq; jqaq (xq) = qjq/2χq(cid:16)kq
xp − a′
q−Nq < ξqq = qjq+1 ≤ q−lq ,
αq (xq) = ψ(q)
ψ(q)
2 ≤ q ≤ p.
(8.6) Ω(cid:0)pjxp − app(cid:1)Ω(cid:0)pj ′
by explicit calculation we obtain
(q−jqxq − aq)(cid:17)Ω(cid:0)q−jqxq − aqq(cid:1),
pp(cid:1) = Ω(cid:0)pjxp − app(cid:1)Ω(cid:0)pj ′−jap − a′
pp(cid:1),
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17),
Next, using (8.7) and repeating the above calculation almost word for word, we find
that
jq ∈ Z+.
kq
q
(8.7)
F [ψ(q)
αq(cid:12)(cid:12)Zq
](ξq) = q−jq/2Ω(qjqaqq)χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
](ξq)(cid:1) 6= 0,
wavelet functions eΨα(x′), in the product (8.2) we have m = P .
(8.8)
only for finite number of indices jq, p < q ≤ m.
αq(cid:12)(cid:12)Zq
(cid:0)F [ζq](ξq), F [ψ(q)
Thus cα = c(bk;bj ba) 6= 0 for finite number of indices jq, 2 ≤ q ≤ m. Moreover, for all
q
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
35
Now we consider the product (8.2) again
cα = c(bk;bj ba) =(cid:0)eζ(x′),eΨα(x′)(cid:1) = Y2≤q≤m(cid:0)ζq(xq), ψ(q)
αq (xq)(cid:1),
where according to the above calculation, m = P . Now let calculate the expression
αq (xq) is given by (8.3). Using identity (8.6) we have
αq (xq)(cid:1). Here the function ζq(xq) ∈ Φ(Qq) belongs to one of the spaces
(cid:0)ζq(xq), ψ(q)
Dlq
Nq (Qq) and ψ(q)
Ω(cid:0)q−jqxq − aqq(cid:1)Ω(cid:0)qNqxqq(cid:1) =( Ω(cid:0)q−jqxq − aqq(cid:1)Ω(cid:0)qN +jqaqq(cid:1), −jq ≤ Nq,
Ω(cid:0)qNqxqq(cid:1)Ω(cid:0)aqq(cid:1),
αq (xq)(cid:1) 6= 0 at least for qN +jqaq ∈ Zq (−jq ≤ Nq) or
This relation implies that(cid:0)ζq(xq), ψ(q)
aq = 0 (−jq > Nq). Since the set of indices jq is finite, the set of the above indices aq is
also finite. Thus products (8.5), (8.8) are nonzero only for finite number of aq ∈ Iq. (cid:3)
Corollary 8.1. Any test function ζ ∈ Φ(A) can be represented in the form of a finite
sum
−jq > Nq.
ζ(x) =Xα∈ eΛ
cα(x∞)eΨα(x′),
x = (x∞, x′) ∈ A,
(8.9)
(8.10)
where eΛ is the set of indices (4.33); cα(x∞) ∈ Φ(R) are some Lizorkin test functions;
eΨα(x′) = eΨ(bk;bj ba)(x′) are wavelet functions (5.6) which do not contain factors ψ(q)
φ(q)(xq − aq), ap ∈ Ip, (see (5.2), (5.3)) for q ≤ m.
αq (xq) =
Using standard results from the book [56], we obtain the following assertion.
Proposition 8.1. Any distribution f ∈ Φ′(A) can be realized in the form of an infinite
sum:
f (x) =Xα∈ eΛ
bα(x∞)eΨα(x′),
x = (x∞, x′) ∈ A,
(5.6).
where bα(x∞) ∈ Φ′(R) are some real Lizorkin distributions; eΨα(x′) are wavelet functions
Here any distribution f ∈ Φ′(A) is associated with representation (8.10), where the
coefficients
(8.11)
bα(x∞)
And vice versa, taking into account Corollary 8.1 and the orthonormality of the wavelet
basis (5.6), any infinite sum (8.10) is associated with the distribution f ∈ Φ′(A), whose
action on a test function ζ(x) ∈ Φ(A) is defined as
def
= (cid:10)f (x∞, x′),eΨα(x′)(cid:11), α ∈eΛ.
cα(x∞)eΨα(x′)E
bβ(x∞)eΨβ(x′),Xα∈ eΛ
=Xα∈ eΛ(cid:10)bα(x∞), cα(x∞)(cid:11),
(cid:10)f, ζ(cid:11) =DXβ∈ eΛ
(8.12)
where the latter sum is finite.
36
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Proposition 8.2. Any distribution ef ∈ Φ′(eA) can be realized as an infinite sum of the
form
(8.13)
ef (x′) =Xp Xα∈ eΛ
x′ ∈eA,
where bα are constants; eΨα(x′) are wavelet functions (5.6).
bαeΨα(x′),
wavelet functions (6.27) instead of wavelet functions (5.6).
It is clear that in Lemma 8.1, Corollary 8.1, and Propositions 8.1, 8.2 one can use
9. Pseudo-differential operators on adeles
9.1. Real and p-adic fractional operators. Let us introduce a distribution from
S ′(R) (we keep the notation · = · ∞.)
(9.1)
s = 0, 1, 2, . . . ,
, α 6= −2s, α 6= 1 + 2s,
α (x) = xα−1
κ(∞)
γ1(α)
x ∈ R,
called the Riesz kernel , where xα is a homogeneous distribution of degree α defined
in [54, Lemma 2.9.], [55, (25.19)], [23, Ch.I,§3.9.], and γ1(α) = 2απ
. The Riesz
One can define the Riesz kernel (9.1) in the real Lizorkin space of distributions Φ′(R)
kernel is an entire function of the complex variable α.
1
2 Γ( α
2 )
Γ( 1−α
2 )
(see [45], [54, Lemma 2.13.], [55, Lemma 25.2.]):
xα−1
=
(9.2)
κ(∞)
α (x) =
xα−1Γ( 1−α
2 )
2α√πΓ( α
2 )
(−1)s+1x2s log x
(−1)sδ(2s)(x),
π(2s)!
2Γ(α) cos( πα
2 )
,
α 6= −2s, α 6= 1 + 2s,
,
α = 1 + 2s,
α = −2s, s ∈ Z+.
According to [23, Ch.II,§3.3.,(2)], [54, Lemma 2.13.], [55, Lemma 25.2.],
(9.3)
F [κ(∞)
α (x)](ξ) = x−α,
x ∈ R.
Define the real Riesz fractional operator Dα
∞ on the Lizorkin space Φ(R) as a convo-
lution
(9.4)
where κ(∞)
α
(9.5)
(cid:0)Dα
∞ϕ(cid:1)(x) =(cid:0)κ(∞)
−α ∗ ϕ(cid:1)(x) =(cid:10)κ(∞)
is given by (9.2). It is clear that [55, (25.2)]
−α (·), ϕ(x − ·)(cid:11), ϕ ∈ Φ(R),
(cid:0)Dα
∞ϕ(cid:1)(x) = F −1[ξαF [ϕ](ξ)](x), ϕ ∈ Φ(R).
Lemma 9.1. The Lizorkin spaces of test functions Φ(R) and distributions Φ′(R) are
invariant under the Riesz fractional operator (9.5) and Dα
∞(Φ′(R)) = Φ′(R).
The p-adic fractional operator Dα
in [58], [59, III.4.]. In [4], the fractional operator Dα
of distributions Φ′(Qp) for all α ∈ C by the following relations:
(9.6)
p was introduced on the space of distributions D′(Qp)
p was defined in the Lizorkin space
f ∈ Φ′(Qp), α ∈ C.
Representation (9.6) can be rewritten as a convolution
(cid:0)Dα
p f(cid:1)(x) = F −1(cid:2) · α
(cid:0)Dα
p f(cid:1)(x) =(cid:0)κ(p)
−α ∗ f(cid:1)(x) = hκ(p)
p F [f ](·)(cid:3)(x),
(9.7)
−α(·), f (x − ·)i,
f ∈ Φ′(Qp), α ∈ C,
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
37
where (see [4])
(9.8)
is the Riesz kernel, and
κ(p)
α (x) =
p
xα−1
Γp(α) ,
δ(x),
log p log xp,
− 1−p−1
= ZQp xα−1
def
Γp(α)
p χp(x) dx =
1 − pα−1
1 − p−α
α 6= 0, 1,
α = 0,
α = 1,
x ∈ Qp,
is the p-adic Γ-function (see [59, III,Theorem (4.2)], [64, VIII,(4.4)]).
Lemma 9.2. ([4]) The Lizorkin spaces of test functions Φ(Qp) and distributions Φ′(Qp)
are invariant under the fractional operator (9.6) and Dα
p (Φ′(Qp)) = Φ′(Qp).
9.2. Adelic fractional operators on the Lizorkin spaces. Let us introduce on the
space Φ(A) the adelic fractional operator Dbγ of order bγ = (γ∞, γ2, . . . , γp, . . . ) ∈ C∞,
which is defined by its projection on any subspace of test functions Φ[m](A) ⊂ Φ(A) (for
details, see Remark 4.1):
Dbγ(cid:12)(cid:12)Φ[m](A)
def
= Dbγ
(m) = Op∈{∞,2,3,...,m}
Dγp
p ⊗Op>m
Idp, m ∈ VQ,
∞ and Dγp
where Dγ∞
the identity operator in Φ(Qp). Here the fractional operator Dbγ
product of one-dimensional operators.
p are defined by (9.4), (9.5) and (9.6), (9.7), respectively, Idp is
(m) is an infinite tensor
(9.9)
If
(9.10)
where
(9.11)
ζ(x) = ζ∞(x∞) Yp∈VQ\∞
ζp(xp) ∈ Φ[m](A)
is an elementary function, i.e., ζp(xp) = φ(p)(ξp) = Ω(xpp) for all p > m(ζ), then taking
into account Definition (9.9) and (9.5), (9.6), we obtain
(cid:0)Dbγ(cid:12)(cid:12)Φ[m](A)ζ(cid:1)(x) =(cid:0)Dbγ
(m)ζ(cid:1)(x) = F −1(cid:2) · bγ
(m) = Yp∈{∞,2,3,...,m}
p Yp>m
ξbγ
ξpγp
φ(p)(ξp),
(m)F [ζ](·)(cid:3)(x),
ξ = (ξ∞, ξ2, . . . ξp, . . . ) ∈ A,
is a symbol of the operator Dbγ
(m), m ∈ VQ.
It is clear that ξbγ
(m) for all bγ = (γ∞, γ2, . . . , γp, . . . ) ∈ C∞ such that for all p > m,
γp = 0 is a multiplier in the space of test functions Ψ[m](Qp), m ∈ VQ.
According to (9.4), (9.7), (2.10), relation (9.10) can be rewritten as
(cid:0)Dbγ
(m)ζ(cid:1)(x) =(cid:0)κ−bγ;(m) ∗ ζ(cid:1)(x) =(cid:10)κ−bγ;(m)(·), ζ(x − ·)(cid:11),
ζ ∈ Φ(A),
(9.12)
where
κbγ;(m)(ξ) = Yp∈{∞,2,3,...,m}
κ(p)
γp (ξp)Yp>m
φ(p)(ξp)
is the adelic Riesz kernel, and κ(∞)
γ∞ and κ(p)
γp are given by (9.2) and (9.8), respectively.
For f ∈ Φ′(A) we define the distribution Dbγf by the relation
(9.13)
hDbγf, ζi def
= hf, Dbγ
(m)ζi,
∀ ζ ∈ Φ[m](A), m ∈ VQ \ ∞.
38
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
It is easy to see that Lemmas 9.1, 9.2 imply the following statements.
Lemma 9.3. The Lizorkin spaces of test functions Φ(A) and distributions Φ′(A) are
invariant under the fractional operator Dbγ
(m)(Φ′(A)) = Φ′(A).
(m) and Dbγ
Proposition 9.1. The family of operators {Dbγ :bγ ∈ C∞} on the space of distributions
Φ′(A) forms an abelian group: if f ∈ Φ′(A) then
bβDbγf = Dbγ+ bβf,
bβf = D
DbγD
DbγD−bγf = f,
bγ, bβ ∈ C∞.
For the case bγ = (γ∞, γ2, . . . , γp, . . . ) ∈ C∞ such that γ∞ = γ2 = · · · = γp = · · · =
γ ∈ C we shall write Dγ instead of Dbγ. Similarly to (9.9), the adelic fractional operator
Dγ of order γ ∈ C∞ is defined by its projection on any subspace Φ[m](A) ⊂ Φ(A) (see
(7.5)):
(9.14)
Dγ(cid:12)(cid:12)Φ[m](A)
def
= Dγ
(m) = Op∈{∞,2,3,...,m}
Dγ
p ⊗Op>m
Idp, m ∈ VQ,
p are defined by (9.4), (9.5) and (9.6), (9.7), respectively, Idp is the
where Dγ
identity operator in Φ(Qp). If
∞ and Dγ
ζ(x) = ζ∞(x∞) Yp∈VQ\∞
ζp(xp) ∈ Φ[m](A)
is an elementary function, i.e., ζp(xp) = φ(p)(ξp) = Ω(xpp) for all p > m(ζ), then taking
into account Definition (9.14) and (9.5), (9.6), we obtain
(9.15)
where
(9.16)
(cid:0)Dγ(cid:12)(cid:12)Φ[m](A)ζ(cid:1)(x) =(cid:0)Dγ
(m)ζ(cid:1)(x) = F −1(cid:2) · γ
p Yp>m
(m) = Yp∈{∞,2,3,...,m}
ξγ
ξpγ
φ(p)(ξp),
(m)F [ζ](·)(cid:3)(x),
ξ = (ξ∞, ξ2, . . . ξp, . . . ) ∈ A.
is the symbol of the operator Dγ
(m), m ∈ VQ.
9.3. One class of adelic pseudo-differential operators. On the adelic Lizorkin
space of distributions Φ′(A) we introduce a pseudo-differential operators A with the
symbol
Ap(ξp),
ξ ∈ A,
A(ξ) = Yp∈VQ
where A∞(ξ∞) ∈ C ∞(R \ {0}) and(cid:12)(cid:12) ds
hood of ξ∞ = 0; Ap(ξp) ∈ E(Qp \ {0}), p = 2, 3, . . . .
of test functions Φ[m](A) ⊂ Φ(A) (see (7.5)):
(9.17)
dξs
def
= A(m) = F −1 A(m) F, m ∈ VQ,
∞A∞(ξ∞)(cid:12)(cid:12) ≤ Msξ∞−ms (ms ≥ 0) in a neighbor-
The adelic pseudo-differential operator A is defined by its projection on any subspace
where
(9.18)
A(m)(ξ) = A∞(ξ∞) Yp∈{2,3,...,m}
Ap(ξp)Yp>m
φ(p)(ξp),
ξ ∈ A,
A(cid:12)(cid:12)Φ[m](A)
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
39
is the symbol of the pseudo-differential operator A(m). If
ζ(x) = ζ∞(x∞) Yp∈VQ\∞
ζp(xp) ∈ Φ[m](A)
is an elementary function, i.e., ζp(xp) = φ(p)(ξp) = Ω(xpp) for all p > m(ζ), then taking
into account Definition (9.17), (9.18), we obtain
(9.19)
A(cid:12)(cid:12)Φ[m](A)ζ = A(m)ζ =(cid:0)F −1 A(m) F(cid:1)ζ,
k=1 ζ r ∈
Φ[m](A), where ζ r are elementary functions, the operator A is defined by the relation
where the symbol A(m) is defined by (9.18). For any test function ζ = Pr
(A(cid:12)(cid:12)Φ[m](A)ζ)(x) = (A(m)ζ)(x) =
rXk=1
(A(m)ζ r)(x).
Now we define a conjugate pseudo-differential operator AT on Φ(A). For any function
ζ ∈ Φ[m](A):
(AT(cid:12)(cid:12)Φ[m](A)ζ)(x)
def
= (AT
(m)ζ)(x) = F −1(cid:2)A(m)(−ξ) F [ζ](ξ)(cid:3)(x).
Then the operator A in the Lizorkin space of distributions Φ′(A) is defined in the usual
way: for f ∈ Φ′(A) we have
(9.20)
∀ ζ ∈ Φ[m](A), m ∈ VQ \ ∞.
hAf, ζi = hf, AT
(m)ζi,
It follows from the latter relation and (2.8) that
(9.21)
Af =(cid:0)F −1 A F(cid:1)f,
f ∈ Φ′(A).
Lemma 9.4. The Lizorkin spaces Φ(A) and Φ′(A) are invariant under the operator
(9.21).
Proof. It is sufficient to prove this lemma for elementary functions. Let ζ ∈ Φ[m](A) be
an elementary function. Since the function A(m)(ξ) defined by (9.18) is a multiplier in the
space Ψ[m](Qp), m ∈ VQ, both functions F [ζ](ξ) and A(m)(ξ)F [ζ](ξ) belong to Ψ[m](A)
and, consequently, Aζ ∈ Φ[m](A), m ∈ VQ. Thus the pseudo-differential operator A (see
(9.17), (9.19)) is well defined and the Lizorkin space Φ(A) is invariant under its action.
Therefore, if f ∈ Φ′(A), then according to (9.20), (9.21), Af = (F −1 A F )f ∈ Φ′(A),
i.e., the Lizorkin space of distributions Φ′(A) is invariant under A.
(cid:3)
The fractional operators (9.10) and (9.15) are particular cases of operator (9.19).
Similarly to the above constructions, we introduce the corresponding operators on the
Lizorkin spaces Φ(eA) and Φ′(eA) : the pseudo-differential operator
(9.22)
A0ζ =(cid:0)F −1 A0) F(cid:1)ζ,
ζ ∈ Φ(eA),
with the symbol
(9.23)
Ap(ξp),
A0(ξ′) = Yp∈VQ\∞
(cid:0)Dbγ
0 ζ(cid:1)(x′) = F −1(cid:2) · bγ
ξ′ = (ξ2, . . . ξp, . . . ) ∈eA;
0F [ζ](·)(cid:3)(x′),
ζ ∈ Φ(eA),
the fractional operator
(9.24)
40
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
and the fractional operator
with the symbol
(9.25)
(9.26)
with the symbol
(9.27)
ξpγp
p ,
0 = Yp∈VQ\∞
ξ′bγ
0 ζ(cid:1)(x′) = F −1(cid:2) · γ
(cid:0)Dγ
0 = Yp∈VQ\∞
ξ′γ
ξpγ
p,
ξ′ = (ξ2, . . . ξp, . . . ) ∈eA;
0F [ζ](·)(cid:3)(x′),
ζ ∈ Φ(eA),
ξ′ = (ξ2, . . . ξp, . . . ) ∈eA.
9.4. On the eigenfunctions of adelic pseudo-differential operators.
Theorem 9.1. Let A0 be a pseudo-differential operator (9.22) with symbol (9.23), and
eΨα(x′) = eΨ(bk;bj ba)(x′) = O2≤q≤m
ψ(q)
αq (xq) ⊗Om<q
be wavelet functions (5.6) which do not contain factors ψ(q)
φ(q)(xq), α ∈eΛ,
x′ ∈eA,
αq (xq) = φ(q)(xq − aq), ap ∈ Ip,
(9.28)
(see (5.2), (5.3)) for q ≤ m; eΛ is the indexes set (4.33). Then eΨα is an eigenfunction
of the operator A0 if and only if
Yq∈{2,3,...,m}
Aq(cid:0)q−jq(−q−1kq + ηq)(cid:1) = Yq∈{2,3,...,m}
Aq(cid:0) − q−jq−1kq(cid:1),
holds for all ηq ∈ Zq, q = 2, 3, . . . , m. The corresponding eigenvalue is the following
jq ∈ Z+,
def
Proof. In view of (9.23) and (2.10), we have
A0(cid:12)(cid:12)Φ[m](eA)eΨα(x′)
λ =Qq∈{2,3,...,m} Aq(cid:0) − qjq−1kq(cid:1), i.e.,
= A0(m)eΨα(x′)
=(cid:16) Yq∈{2,3,...,m}
Aq(cid:0) − qjq−1kq(cid:1)(cid:17)eΨα(x′), α ∈eΛ,
A0(cid:12)(cid:12)Φ[m](eA)eΨα(x′)
= A0(m)eΨα(x′) = F −1(cid:2)A0(m)(ξ′)F [eΨα(ξ′)(cid:3)(cid:3)(x′)
αq(cid:3)(cid:3)(xq)Yq>m
= Y2≤q≤p
F −1(cid:2)AqF [ψ(q)
kq; jqaq (xq) = qjq/2χq(cid:0) kq
where ψ(q)
(5.3)), and xq ∈ Qq, 2 ≤ q ≤ p; xq ∈ Zq, p < q ≤ m.
αq(cid:3)(cid:3)(xq) Yp<q≤m
F −1(cid:2)AqF [ψ(q)
If condition (9.28) holds, then using formulas (8.4), (8.7), we obtain
αq (xq) = ψ(q)
def
q (q−jqxq − aq)(cid:1)Ω(cid:0)q−jqxq − aqq(cid:1) (see (5.2),
x′ ∈eA.
φ(q)(xq),
A0(cid:12)(cid:12)Φ[m](eA)eΨα(x′) = A0(m)eΨα(x′)
= Y2≤q≤p
× Yp<q≤m
q−jq/2F −1hAq(ξq)χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
q−jq/2F −1hAq(ξq)Ω(qjqaqq)χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
kq
q
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17)i(xq)×
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17)i(xq)×
×Yq>m
φ(q)(xq).
kq
q
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
41
kq
q
Making the change of variables ξq = q−jq(ηq − q−1kq) and using formulas (2.8), (2.10),
we have for 2 ≤ q ≤ p
kq
q
αq (xq),
xq ∈ Zq.
kq
q
αq (xq)
q
αq (xq),
xq ∈ Qq.
In the same way, we obtain for p < q ≤ m
χq(cid:0)(−q−1kq + ηq)(−q−jqxq + aq)(cid:1)Ω(ηqq) dηq
The latter expression is nonzero only if qjqaq ∈ Zq and xq − qjq aqq ≤ q−jq. This implies
that xq ∈ Zq. Thus,
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17)i(xq)
= Aq(cid:0) − qjq−1kq(cid:1)ψ(q)
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17)i(xq)
(q−jqxq − aq)(cid:17)Ω(cid:0)q−jqxq − aqq(cid:1).
q−jq/2F −1hAq(ξq)χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
= Aq(cid:0) − qjq−1kq(cid:1)qjq/2ZQq
q−jq/2F −1hAq(ξq)Ω(qjqaqq)χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
= Aq(cid:0) − qjq−1kq(cid:1)Ω(qjqaqq)ψ(q)
= Aq(cid:0) − qjq−1kq(cid:1)Ω(qjqaqq)qjq/2χq(cid:16)kq
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17)i(xq)
q−jq/2F −1hAq(ξq)Ω(qjqaqq)χq(cid:0)qjqaqξq(cid:1)Ω(cid:16)(cid:12)(cid:12)(cid:12)
= Aq(cid:0) − qjq−1kq(cid:1)ψ(q)
Consequently, A0eΨα(x′) = A0(m)eΨα(x′) = λeΨα(x′), where
Aq(cid:0) − qjq−1kq(cid:1),
x′ ∈eA.
Conversely, if A0(cid:12)(cid:12)Φ[m](eA)eΨα(x′) = A0(m)eΨα(x′) = λeΨα(x′), then taking the Fourier trans-
+ qjqξq(cid:12)(cid:12)(cid:12)q(cid:17) = 0,
q(cid:1)q−jq for 2 ≤ q ≤ m. Thus,
q + qjqξq = ηq, ηq ∈ Zq, then ξq =(cid:0)ηq − kq
λ = A0(m)(cid:16)(cid:0)η2 − 2−1k2(cid:1)2−j2, . . . ,(cid:0)ηm − m−1km(cid:1)m−jm(cid:17),
In particular, λ = A0(m)(cid:0)2−j2−1k2, . . . , m−jm−1km(cid:1) and consequently (9.28) holds.
p Qq>m φ(q)(ξq),
ξ′ = (ξ2, . . . ξp, . . . ) ∈ eA.
A0(m)(ξ′) = ξ′bγ
is easy to see that the symbol A0(m)(ξ′) = ξ′bγ
Yq∈{2,3,...,m}
form of both left- and right-hand sides of this identity and using formulas (8.4), (8.7),
we obtain
According to (9.24), (9.25), the adelic fractional operator Dbγ
0(m) satisfies condition (9.28):
0(m) = Qq∈{2,3,...,m} ξqγq
Aq(cid:0)q−jq(−q−1kq + ηq)(cid:1) = Yq∈{2,3,...,m}(cid:12)(cid:12)(cid:12)q−jq(cid:0) − q−1kq + ηq(cid:1)(cid:12)(cid:12)(cid:12)
= Yq∈{2,3,...,m}
qjq(cid:12)(cid:12) − q−1kq + ηq(cid:12)(cid:12)γq
= Yq∈{2,3,...,m}
q = Yq∈{2,3,...,m}
Aq(cid:0) − q−jq−1kq(cid:1),
∀ ηq ∈ Zq,
qγq(jq+1)
(cid:0)A0(ξ′) − λ(cid:1) Y2≤q≤m
0(m) has the symbol
It
λ = Yq∈{2,3,...,m}
∀ ηq ∈ Zq, 2 ≤ q ≤ m.
γq
q
2 ≤ q ≤ m.
If now kq
kq
q
Ω(cid:16)(cid:12)(cid:12)(cid:12)
ξ′ ∈eA.
(cid:3)
42
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
Thus according to Theorem 9.1, we have
Corollary 9.1. The wavelet function (5.6)
which does not contain the factors ψ(q)
eigenfunction of the adelic fractional operator Dbγ
0 :
αq (xq) = φ(q)(xq − aq), ap ∈ Ip for q ≤ m, is an
Corollary 9.2. The wavelet function (5.6)
(9.29)
Dbγ
eΨα(x′) = eΨ(bk;bj ba)(x′) = O2≤q≤m
0(cid:12)(cid:12)Φ[m](eA)eΨα(x′) = Dbγ
eΨα(x′) = eΨ(bk;bj ba)(x′) = O2≤q≤m
0(cid:12)(cid:12)Φ[m](eA)eΨα(x′) = Dγ
Dγ
(9.30)
ψ(q)
αq (xq) ⊗Om<q
φ(q)(xq),
ψ(q)
φ(q)(xq),
0(m)eΨα(x′) =(cid:16) Yq∈{2,3,...,m}
αq (xq) ⊗Om<q
0(m)eΨα(x′) =(cid:16) Yq∈{2,3,...,m}
Acknowledgements
Λm.
Λm.
x′ ∈eA, α ∈ Λ = [m∈VQ\∞
qγq(jq+1)(cid:17)eΨα(x′),
x′ ∈eA.
x′ ∈eA, α ∈ Λ = [m∈VQ\∞
qjq+1(cid:17)γ
eΨα(x′),
x′ ∈eA.
which does not contain the factors ψ(q)
eigenfunction of the adelic fractional operator Dγ
0 :
αq (xq) = φ(q)(xq − aq), ap ∈ Ip, for q ≤ m, is an
The authors are greatly indebted to A.N. Kochubei, V.I. Polischook, O.G. Smolyanov
for fruitful discussions.
References
[1] S. Albeverio, S. Evdokimov, M. Skopina, p-Adic multiresolution analysis and wavelet frames, Journal
of Fourier Analysis and Applications, 16, no. 5, (2010), 693 -- 714.
[2] S. Albeverio, S. Evdokimov, M. Skopina, p-Adic non-orthogonal wavelet bases, Proc. Steklov Inst.
Math., v. 265, Moscow, 2009, 1 -- 12.
[3] S. Albeverio, A. Yu. Khrennikov, V. M. Shelkovich, Theory of p-adic distributions: Linear and
nonolinear models, London Mathematical Society Lecture Note Series (No. 370), Cambridge Uni-
versity Press, Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo, Delhi,
Dubai, Tokyo, 2010.
[4] S. Albeverio, A. Yu. Khrennikov, V. M. Shelkovich, Harmonic analysis in the p-adic Lizorkin spaces:
fractional operators, pseudo-differential equations, p-adic wavelets, Tauberian theorems, Journal of
Fourier Analysis and Applications, Vol. 12, Issue 4, (2006), 393 -- 425.
[5] S. Albeverio, A. Yu. Khrennikov, and V. M. Shelkovich, p-Adic semi-linear evolutionary pseudo-
differential equations in the Lizorkin space Dokl. Ross. Akad. Nauk, 415, no. 3, (2007), 295 -- 299.
English transl. in Russian Doklady Mathematics, 76, no. 1, (2007), 539 -- 543.
[6] S. Albeverio, A. Yu. Khrennikov, V. M. Shelkovich, The Cauchy problems for evolutionary pseudo-
differential equations over p-adic field and the wavelet theory, Journal of Mathematical Analysis
and Applications, 375, (2011), 82 -- 98.
[7] S. Albeverio, S. V. Kozyrev, Multidimensional ultrametric pseudodifferential equations, Proc.
Steklov Inst. Math., v. 265, Moscow, 2009, 13 -- 29.
[8] V. Anashin and A. Yu. Khrennikov, Applied algebraic dynamics, De Gruyter, Berlin, New York,
2009.
[9] I. Ya. Aref′eva, B. G. Dragovic, and I. V. Volovich On the adelic string amplitudes, Phys. Lett. B,
209, no. 4, (1998), 445 -- 450.
[10] J. J. Benedetto, and R. L. Benedetto, A wavelet theory for local fields and related groups, The
Journal of Geometric Analysis, 3, (2004), 423 -- 456.
[11] R. L. Benedetto, Examples of wavelets for local fields, Wavelets, Frames, and operator Theory,
(College Park, MD, 2003), Am. Math. Soc., Providence, RI, (2004), 27 -- 47.
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
43
[12] Yu.M. Berezanskii, Selfadjoint operators in spaces of functions of infinitely many variables, Trans-
lated from the Russian by H. H. McFaden. Translation edited by Ben Silver. Translations of Math-
ematical Monographs, 63. American Mathematical Society, Providence, RI, 1986.
[13] A. Borel, Some finiteness properties of adele groups over number fields, Publications
math´ematiques de l'I.H.´E.S., tome 16, (1963), 5 -- 30.
[14] L. Brekke and P. G. O. Freund, p-Adic numbers in physics, Phys. Rep., 233, 1, (1993).
[15] J. W. S. Cassels and A. Frohlich, Algebraic Number Theory, Academic Press, 1967.
[16] C. Chevalley, G´en´eralisation de la th´eorie du corps de classes pour les extensions infinies, J. Math.
Pures Appl., 15, (1936), 359 -- 371.
[17] K. Conrad, The
classical
and
adelic
point
of
view in number
theory,
(2010),
http://www.mccme.ru/ium/postscript/f10/conrad-lecturenotes.pdf
[18] I. Daubechies, Ten Lectures on wavelets, CBMS-NSR Series in Appl. Math., SIAM, 1992.
[19] B. Dragovich, On Generalized Functions in Adelic Quantum Mechanics, Integral Transform. Spec.
Funct. 6, (1998), 197 -- 203.
[20] B. Dragovich, Adelic model of harmonic oscillator, (Russian), Teoret. Mat. Fiz., 101, (1994),
349 -- 359; translation in Theoret. and Math. Phys., 101, no. 3, (1994), 1404 -- 1412.
[21] B. Dragovich, p-Adic and adelic quantum mechanics, Proc. V.A. Steklov Inst. Math. 245, (2004),
72 -- 85.
[22] B. Dragovich, Ya. V. Radyno, and A. Khrennikov, Distributions on adeles, Journal of Mathematical
Sciences, 142, no. 3, (2007), 2105 -- 2112.
[23] I. M. Gel'fand and G. E. Shilov, Generalized Functions, vol. 1, Properties and Operations. New
York, Acad. Press, 1964.
[24] I. M. Gel′fand, M. I. Graev, and I. I. Piatetskii-Shapiro, Generalized functions. vol 6: Represen-
tation theory and automorphic functions. Nauka, Moscow, 1966. Translated from the Russian by
K.A. Hirsch, Published in 1990, Academic Press (Boston).
[25] W. Karwowski, M. Vilela, Hierarchical structures and asymmetric stochastic processes on p-adics
and adeles, Journal of Mathematical Physics, Volume 35, Issue 9, (1994), 4637 -- 4650.
[26] A. Khrennikov, p-Adic valued distributions in mathematical physics, Kluwer Academic Publ., Dor-
drecht, 1994.
[27] A. Khrennikov, Non-archimedean analysis: quantum paradoxes, dynamical systems and biological
models, Kluwer Academic Publ., Dordrecht, 1997.
[28] A. Yu. Khrennikov, Information dynamics in cognitive, psychological, social and anomalous phe-
nomena, Kluwer Academic Publ., Dordrecht, 2004.
[29] A. Yu. Khrennikov, and S. V. Kozyrev, Wavelets on ultrametric spaces, Applied and Computational
Harmonic Analysis, 19, (2005), 61 -- 76.
[30] A. Yu. Khrennikov and M. Nilsson, p-Adic Deterministic and Random Dynamics, Kluwer, Dor-
drecht, 2004.
[31] A. Khrennikov, Ya. V. Radyno, On adelic analog of Laplacian, Proceedings of Jangjeon Mathe-
matical Society, 6, no. 1, (2003), 1 -- 18.
[32] A.Yu. Khrennikov, V.M.
Shelkovich,
p-Adic multidimensional wavelets
application
http://arxiv.org/abs/math-ph/0612049
p-adic
to
pseudo-differential
operators,
(2006),
Preprint
at
and
the
their
url:
[33] A. Yu. Khrennikov, V. M. Shelkovich, Non-Haar p-adic wavelets and their application to pseudo-
differential operators and equations, Applied and Computational Harmonic Analysis, 28, (2010),
1 -- 23.
[34] A.Yu. Khrennikov, V.M. Shelkovich, M. Skopina, p-Adic refinable functions and MRA-based
wavelets, Journal of Approximation Theory, 161, (2009), 226 -- 238.
[35] A.Yu. Khrennikov, V.M. Shelkovich, M. Skopina, p-Adic orthogonal wavelet bases, p-Adic Numbers,
Ultrametric Analysis and Applications, 1, no. 2, (2009), 145 -- 156.
[36] A. N. Kochubei, Parabolic equations over the field of p-adic numbers, Izv. Akad. Nauk SSSR Ser.
Mat., 55, no. 6 (1991), 1312 -- 1330. English transl. in Math. USSR Izv., 39, (1992), 1263 -- 1280
[37] A. N. Kochubei, Schrodinger-type operator over the p-adic number field, Teoret. Mat. Fiz., 86,
(1991), no. 3, 323 -- 333; translation in Theoret. and Math. Phys., 86, (1991), no. 3, 221 -- 228.
[38] A. N. Kochubei, Pseudo-differential equations and stochastics over non-archimedean fields, Marcel
Dekker. Inc. New York, Basel, 2001.
44
A. YU. KHRENNIKOV, A. V. KOSYAK, AND V. M. SHELKOVICH
[39] S. V. Kozyrev, Wavelet analysis as a p-adic spectral analysis, Izvestia Ross. Akademii Nauk, Seria
Math., 66 no. 2, 149 -- 158; English transl. in Izvestiya: Mathematics, 66 no. 2, 367 -- 376.
[40] S. V. Kozyrev, Methods and applications of ultrametric and p-adic analysis:
from wavelet
Issue 12. Steklov Inst. Math., Moscow, 2008.
theory to biophysics, Sovrem. Probl. Mat.
http://www.mi.ras.ru/spm/pdf/012.pdf.
[41] S. V. Kozyrev, p-Adic pseudodifferential operators and p-adic wavelets, Theor. Math. Physics, 138,
no. 3, 1 -- 42; English transl. in Theoretical and Mathematical Physics, 138, no. 3, 322 -- 332.
[42] A.V. Kosyak, Regular, quasiregular and induces representations of infinite-dimensional groups,
EMS, 453 p. (in preparation).
[43] P. I. Lizorkin, Generalized Liouville differentiation and the functional spaces Lp
r(En). Imbedding
theorems, (Russian) Mat. Sb. (N.S.) 60(102), (1963), 325 -- 353.
[44] P. I. Lizorkin, Generalized Liouville differentiation and the multiplier method in the theory of
imbeddings of classes of differentiable functions, (Russian) Trudy Mat. Inst. Steklov. Vol. 105, 1969
89 -- 167.
[45] P. I. Lizorkin, Operators connected with fractional differentiation, and classes of differentiable
functions, (Russian) Studies in the theory of differentiable functions of several variables and its
applications, IV. Trudy Mat. Inst. Steklov. Vol. 117, (1972), 212 -- 243.
[46] S. Mallat, Multiresolution representation and wavelets, Ph. D. Thesis, University of Pennsylvania,
Philadelphia, PA. 1988.
[47] S. Mallat, An efficient image representation for multiscale analysis, In: Proc. of Machine Vision
Conference, Lake Taho. 1987.
[48] Ju. I. Manin, Non-Archimedean integration and p-adic Jacquet-Langlands L-functions, (Russian)
Uspehi Mat. Nauk, 31(187), no. 1, (1976), 5 -- 54. Russian Mathematical Surveys, 31:1, (1976), 5 -- 57
[49] Yu. I. Manin and A. A. Panchishkin, Introduction to modern number theory fundamental problems,
Ideas and theories, 2nd Edition, Spinger-Verlag, Berlin Heidelberg, 2005.
[50] Y. Meyer, Ondelettes and fonctions splines, S´eminaire EDP. Paris. December 1986.
[51] Y. Meyer, Principe d'incertitude, bases hilbertiennes et alg`ebres d'op´erateurs, S´eminaire N. Bour-
baki, 1985 -- 1986, exp. no. 662, p.209 -- 223.
[52] J. von Neumann, On infinite direct products, Compositio mathematica, Vol. 15, no. 1, (1939),
1 -- 77.
[53] I. Novikov, V. Protasov, and M. Skopina, Wavelet Theory, Translations of Mathematical Mono-
graphs, vol. 239, AMS, 2011.
[54] S. G. Samko, Hypersingular integrals and their applications. Taylor & Francis, New York, 2002.
[55] S. G. Samko, A. A. Kilbas, and O. I. Marichev, Fractional Integrals and Derivatives and Some of
Their Applications. Minsk, Nauka i Tekhnika, 1987 (in Russian).
[56] H. H. Schaefer, Topological vector spaces, The Macmillan Co., New York; Collier-Macmillan Ltd.,
London, 1966.
[57] V. M. Shelkovich, M. Skopina, p-Adic Haar multiresolution analysis and pseudo-differential oper-
ators, Journal of Fourier Analysis and Applications, Vol. 15, Issue 3, (2009), 366 -- 393.
[58] M. H. Taibleson, Harmonic analysis on n-dimensional vector spaces over local fields. I. Basic
results on fractional integration, Math. Annalen, 176, (1968), 191 -- 207.
[59] M. H. Taibleson, Fourier analysis on local fields, Princeton University Press, Princeton, 1975.
[60] V. S. Vladimirov, and T. M. Zapuzhak, Adelic formulas for string amplitudes in fields of algebraic
numbers, Lett. Math. Phys., 37, (1996), 232 -- 242.
[61] V. S. Vladimirov, Adelic Formulas for Gamma and Beta Functions of Completions of Algebraic
Number Fields and Their Applications to String Amplitudes, Izv. Ross. Akad. Nauk, Ser. Mat., 60,
(1996), no. 1, 63 -- 86 (in Russian); English translation in Izvestiya: Mathematics, 60, (1996), no. 1,
67 -- 90.
[62] V. S. Vladimirov, p-Adic analysis and p-adic quantum mechanics, Ann. of the NY Ac. Sci.: Sym-
posium in Frontiers of Math., 1988.
[63] V. S. Vladimirov, On the spectrum of some pseudo-differential operators over p-adic number field,
Algebra and analysis, 2, no. 6, 1990, 107 -- 124.
[64] V. S. Vladimirov, I. V. Volovich, and E. I. Zelenov, p-Adic analysis and mathematical physics.
World Scientific Publishing, River Edge, NJ, 1994.
[65] A. Weil, Adeles and algebraic groups, Progress in Mathematics, vol. 23, Birkhauser Boston, Mass.,
1982, With appendices by M. Demazure and Takashi Ono.
WAVELET ANALYSIS ON ADELES AND PSEUDO-DIFFERENTIAL OPERATORS
45
[66] K. Yasuda, Markov Processes on the Adeles and Representations of Euler Products, J. Theoret.
Probab. 23, (2010), 748 -- 769.
International Center for Mathematical Modelling in Physics and Cognitive Sciences
MSI, Vaxjo University, SE-351 95, Vaxjo, Sweden.
E-mail address: [email protected]
Institute of Mathematics, Ukrainian National Academy of Sciences, 3 Tereshchenkivs'ka,
Kyiv, 01601, Ukraine.
E-mail address: [email protected]
Department of Mathematics, St.-Petersburg State Architecture and Civil Engineer-
ing University, 2 Krasnoarmeiskaya 4, 190005, St. Petersburg, Russia.
E-mail address: [email protected]
|
1207.2025 | 1 | 1207 | 2012-07-09T12:54:41 | Infinitely divisible metrics and curvature inequalities for operators in the Cowen-Douglas class | [
"math.FA"
] | The curvature $\mathcal K_T(w)$ of a contraction $T$ in the Cowen-Douglas class $B_1(\mathbb D)$ is bounded above by the curvature $\mathcal K_{S^*}(w)$ of the backward shift operator. However, in general, an operator satisfying the curvature inequality need not be contractive. In this note, we characterize a slightly smaller class of contractions using a stronger form of the curvature inequality. Along the way, we find conditions on the metric of the holomorphic Hermitian vector bundle $E_T$ corresponding to the operator $T$ in the Cowen-Douglas class $B_1(\mathbb D)$ which ensures negative definiteness of the curvature function. We obtain a generalization for commuting tuples of operators in the class $B_1(\Omega)$, for a bounded domain $\Omega$ in $\mathbb C^m$. | math.FA | math | INFINITELY DIVISIBLE METRICS AND CURVATURE INEQUALITIES
FOR OPERATORS IN THE COWEN-DOUGLAS CLASS
SHIBANANDA BISWAS, DINESH KUMAR KESHARI, AND GADADHAR MISRA
Abstract. The curvature KT (w) of a contraction T in the Cowen-Douglas class B1(D) is
bounded above by the curvature KS ∗(w) of the backward shift operator. However, in gen-
eral, an operator satisfying the curvature inequality need not be contractive. In this note, we
characterize a slightly smaller class of contractions using a stronger form of the curvature in-
equality. Along the way, we find conditions on the metric of the holomorphic Hermitian vector
bundle ET corresponding to the operator T in the Cowen-Douglas class B1(D) which ensures
negative definiteness of the curvature function. We obtain a generalization for commuting tuples
of operators in the class B1(Ω), for a bounded domain Ω in Cm.
2
1
0
2
l
u
J
9
]
Let H be a complex separable Hilbert space and L(H) denote the collection of bounded linear
operators on H. The following important class of operators was introduced in [3].
1. Introduction
Definition 1.1. For a connected open subset Ω of C and a positive integer n, let
.
A
F
h
t
a
m
[
1
v
5
2
0
2
.
7
0
2
1
:
v
i
X
r
a
Bn(Ω) = (cid:8) T ∈ L(H) Ω ⊂ σ(T ),
ran (T − w) = H for w ∈ Ω,
ker(T − w) = H,
_w∈Ω
dim ker(T − w) = n for w ∈ Ω(cid:9),
where σ(T ) denotes the spectrum of the operator T .
We recall (cf.
[3]) that an operator T in the class Bn(Ω) defines a holomorphic Hermitian
vector bundle ET in a natural manner. It is the sub-bundle of the trivial bundle Ω × H defined
by
ET = {(w, x) ∈ Ω × H : x ∈ ker(T − w)}
with the natural projection map π : ET → Ω, π(w, x) = w. It is shown in [3, Proposition 1.12]
that the mapping w −→ ker(T − w) defines a rank n holomorphic Hermitian vector bundle ET
over Ω for T ∈ Bn(Ω). In [3], it was also shown that the equivalence class of the holomorphic
Hermitian vector bundle ET and the unitary equivalence class of the operator T determine each
other.
Theorem 1.2. The operators T and eT in Bn(Ω) are unitarily equivalent if and only if the
corresponding holomorphic Hermitian vector bundles ET and EeT
2000 Mathematics Subject Classification. 47B32, 47B35.
Key words and phrases. Cowen-Douglas class, curvature inequality, holomorphic Hermitian vector bundle, ker-
are equivalent.
nel function, infinite divisibility.
The work of S. Biswas was supported by a postdoctoral fellowship funded by the Skirball Foundation via the
Center for Advanced Studies in Mathematics at Ben-Gurion University of the Negev. The work of D.K. Keshari
was supported in the form of a CSIR Research Fellowship at the Indian Institute of Science. The work of G. Misra
was supported in part by UGC - SAP and by a grant from the Department of Science and Technology.
1
2
BISWAS, KESHARI, AND MISRA
In general, it is not easy to decide if two holomorphic Hermitian vector bundles are equivalent
except when the rank of the bundle is 1. In this case, the curvature
K(w) = −
∂2 log k γ(w) k2
∂w∂w
,
of the line bundle E, defined with respect to a non-zero holomorphic section γ of E, is a complete
invariant. The definition of the curvature is independent of the choice of the section γ: If γ0
is another holomorphic section of E, then γ0 = φγ for some non-zero holomorphic function
φ defined on an open subset Ω0 of Ω, consequently the harmonicity of log φ, completes the
verification.
Let T ∈ B1(Ω). Fix w ∈ Ω and let γ be a holomorphic section of the line bundle ET . From
[3, Lemma 1.22], it follows that the vectors γ(w) and ∂γ(w) from a basis of ker(T − w)2. Let
NT (w) = T ker(T −w)2 and {γ1(w), γ2(w)} be the basis obtained by applying Gram-Schmidt ortho-
normalization to the vectors γ(w) and ∂γ(w). The linear transformation NT (w) has the matrix
representation
w (cid:19) ,
NT (w) =(cid:18)w hT (w)
0
where hT (w) =(cid:0) − KT (w)(cid:1)− 1
2 , with respect to the orthonormal basis {γ1(w), γ2(w)}.
The curvature KT (w) of an operator T in B1(Ω) is negative. To see this, recall that the
curvature may also be expressed (cf.[3, page - 195]) in the form
(1.1)
KT (w) = −
kγ(w)k2kγ′(w)k2 − hγ′(w), γ(w)i2
kγ(w)k4
.
Applying Cauchy - Schwarz inequality, we see that the numerator is positive.
Let {e0, e1} be an orthonormal set of vectors. Suppose N is a nilpotent linear transformation
defined by the rule
e1 → a e0, e0 → 0, a ∈ C.
Then a determines the unitary equivalence class of N .
0
hT (w) > 0 since we have shown that the curvature KT (w) is negative. Hence the curvature
KT (w) is an invariant for the operator T . The non-trivial converse of this statement follows from
0 (cid:17) of the operator T in B1(Ω) is nilpotent. Now,
The localization NT (w) − wI2 = (cid:16) 0 hT (w)
Theorem 1.2. Thus the operators T and eT in B1(Ω) are unitarily equivalent if and only if NT (w)
is unitarily equivalent to NeT (w) for w in Ω.
Note that if T ∈ B1(D) is a contraction, that is, kT k ≤ 1, then NT (w) is a contraction for each
w ∈ D. Observe that ( a c
0 b ) is a contraction if and only if a ≤ 1 and c2 ≤ (1 − a2)(1 − b2).
(1−w2)2 , w ∈ D. The adjoint S∗ of the unilateral
Thus kNT (w)k ≤ 1 if and only if KT (w) ≤ −
shift operator S is in B1(D). It is easy to see that γS∗(w) = (1, w, . . . , wn, . . .) ∈ ℓ2
+, w ∈ D, is
a holomorphic section for the corresponding holomorphic Hermitian line bundle ES∗. The norm
kγS∗(w)k2 of the section γS∗ is (1 − w2)−1 and hence the curvature KS∗(w) of the operator S∗
is given by the formula −
(1−w2)2 , w ∈ D. We have therefore proved:
1
1
Proposition 1.3. If T is a contractive operator in B1(D), then the curvature of T is bounded
above by the curvature of the backward shift operator S∗.
We think of the operator S∗ as an extremal operator within the class of contractions in
B1(D). This is a special case of the curvature inequality proved in [8]. The curvature inequality
is equivalent to contractivity of the operators NT (w), w ∈ Ω, while the contractivity of the
operator T is global in nature. So, it is natural to expect that the validity of the inequality
Indeed, J. Agler had
KT (w) ≤ −
(1−w2)2 , w ∈ D, need not force T to be a contraction.
1
INFINITE DIVISIBILITY AND CURVATURE INEQUALITY
3
communicated the existence of an operator T , kT k > 1, in B1(D) with KT (w) ≤ KS∗(w) (cf. [8,
Note added in proof]) to G. Misra. Unfortunately, there is a printing error in [8, Note added in
proof], which should be corrected by reversing the inequality sign. No explicit example has been
written down of this phenomenon. We provide such an example here.
The main point of this note is to investigate additional conditions on the curvature, apart from
the inequality we have discussed above, which will ensure contractivity. We give an alternative
proof the curvature inequality. A stronger inequality becomes apparent from this proof. It is
this stronger inequality which, as we will show below, admits a converse.
An operator T in the class B1(Ω), as is well-known (cf. [3, pp. 194 ]), is unitarily equivalent
to the adjoint M ∗ of the multiplication operator M by the co-ordinate function on some Hilbert
space HK of holomorphic functions on Ω∗ := {z ∈ C : ¯z ∈ Ω} possessing a reproducing kernel
K.
The kernel K is a complex valued function defined on Ω∗ × Ω∗ which is holomorphic in the first
variable and anti-holomorphic in the second. In consequence, the map ¯w → K(·, w), w ∈ Ω∗, is
holomorphic on Ω. We have K(z, w) = K(w, z) making it Hermitian. It is positive definite in
the sense that the n × n matrix
(cid:0)(cid:0) K(wi, wj)(cid:1)(cid:1)n
i,j=1
is positive definite for every subset {w1, . . . , wn} of Ω∗, n ∈ N. Finally, the kernel K reproduces
the value of functions in HK, that is, for any fixed w ∈ Ω∗, the holomorphic function K(·, w)
belongs to HK and
f (w) = hf, K(·, w)i, f ∈ HK, w ∈ Ω∗.
The correspondence between the operator T in B1(Ω) and the operator M ∗ on the Hilbert
space HK is easy to describe (cf.
[3, pp. 194 ]). Let γ be a non-zero holomorphic section (for
bounded domain in C, by Grauert's Theorem, a global section exists) for the operator T acting on
the Hilbert space H. Consider the map Γ : H → O(Ω∗), where O(Ω∗) is the space of holomorphic
functions on Ω∗, defined by Γ(x)(z) = hx, γ(¯z)i, z ∈ Ω∗. Transplant the inner product from H
on the range of Γ. The map Γ is now unitary from H onto the completion of ran Γ. Define K
to be the function K(z, w) = Γ(cid:0)γ( ¯w)(cid:1)(z) = hγ( ¯w), γ(¯z)i, z, w ∈ Ω∗. Set Kw(·) := K(·, w). Thus
Kw is the function Γ(cid:0)γ( ¯w)(cid:1). It is then easily verified that K has the reproducing property, that
is,
hΓ(x)(z), K(z, w)iran Γ = h(cid:0)hx, γ(¯z)i(cid:1),(cid:0)hγ( ¯w), γ(¯z)(cid:1)iran Γ
= hΓx, Γ(γ( ¯w))iran Γ = hx, γ( ¯w)iH
= Γ(x)(w), x ∈ H, w ∈ Ω∗.
It follows that kKw(·)k2 = K(w, w), w ∈ Ω∗. Also, Kw(·) is an eigenvector for the operator
Γ T Γ∗ with eigenvalue ¯w in Ω:
Γ T Γ∗(Kw(·)) = Γ T Γ∗(cid:0)Γ(γ( ¯w))(cid:1)
= Γ T γ( ¯w)
= Γ ¯w γ( ¯w)
= ¯w Kw(·), w ∈ Ω∗.
Since the linear span of the vectors {Kw : w ∈ Ω∗} is dense in HK, it follows that Γ T Γ∗ is
the adjoint M ∗ of the multiplication operator M on HK . We therefore assume, without loss of
generality, that an operator T in B1(Ω) has been realized as the adjoint M ∗ of the multiplication
operator M on some Hilbert space HK of holomorphic functions on Ω∗ possessing a reproducing
kernel K.
4
BISWAS, KESHARI, AND MISRA
Remark 1.4. The contractivity of the adjoint M ∗ of the multiplication operator M on some
reproducing kernel Hilbert space HK is equivalent to the requirement that K ‡(z, w) := (1 −
z ¯w)K(z, w) is positive definite on D (cf.
[1, Corollary 2.37] and [7, Lemma 1]). Suppose that
the operator M ∗ is in B1(D). Here is a second proof of the curvature inequality:
We have
∂2
∂w∂ ¯w
log K(w, w) =
∂2
∂w∂ ¯w
log
1
(1 − w2)
+
∂2
∂w∂ ¯w
log K ‡(w, w), w ∈ D,
which we rewrite as
KM ∗(w) = KS∗(w) −
∂2
∂w∂ ¯w
log K ‡(w, w), w ∈ D.
Recalling that
we conclude that
∂2
∂w∂ ¯w log K ‡(w, w) must be positive (see (1.1)) as long as K ‡ is positive definite,
KM ∗(w) ≤ KS∗(w), w ∈ D.
The fibre at ¯w of the holomorphic bundle EM ∗ for M ∗ in B1(Ω) is the one-dimensional kernel
at ¯w of the operator M ∗ spanned by Kw(·), w ∈ Ω∗. In general, there is no obvious way to
define an inner product between the two vectors Kw(·) and ( ∂
∂ ¯w Kw)(·). However since these
[5, Lemma 4.3]), in our special case, there is a
vectors belong to the same Hilbert space (cf.
natural inner product defined between them. This ensures, via the Cauchy-Schwarz inequality,
the negativity of the curvature KT . The reproducing kernel function K of the Hilbert space
HK encodes the mutual inner products of the vectors {Kw(·) : w ∈ Ω∗}. The Cauchy-Schwarz
inequality, in turn, is just the positivity of the Gramian of the two vectors Kw(·) and ( ∂
∂ ¯w Kw)(·),
w ∈ Ω∗. The positive definiteness of K is a much stronger positivity requirement involving all
the derivatives of the holomorphic section Kw(·) defined on Ω∗. We exploit this to show that the
curvature function (cid:0) ∂2
∂z∂ ¯w log K(cid:1)(z, w) is actually negative definite not just negative, whenever
K t is assumed to be positive definite for all t > 0.
We now construct an example of an operator which is not contractive but its curvature is
dominated by the curvature of the backward shift. Expanding the function K(z, w) = 8+8z ¯w−z2 ¯w2
in z ¯w, we see that it has the form 8 + 16z ¯w + 15 z2 ¯w2
1−z ¯w . Therefore, it defines a positive definite
kernel on the unit disk D. The monomials
15 for
n ≥ 2 forms an orthonormal basis in the corresponding Hilbert space HK . The multiplication
zn+1
operator M maps zn
kzn+1k . Hence it corresponds to a weighted shift operator W with
15 , 1, 1, . . .}. Evidently, it is not a contraction. (This is the same as
saying that the function K ‡(z, w) = 8 + 8z ¯w − z2 ¯w2 is not positive definite.) The operator W is
similar to the forward shift S. Since the class B1(D) is invariant under similarity and S ∈ B1(D),
it follows that W is in it as well. However,
the weight sequence {q 1
zn
kznk with k1k2 = 1
16 and kznk2 = 1
kznk to kzn+1k
8 , kzk2 = 1
kznk
2 ,q 16
1−z ¯w
−
∂2
∂w∂ ¯w
log K ‡(w, w) = −
8(8 − 4w2 − w4)
(8 + 8w2 − w4)2 , w ∈ D,
is negative for w < 1. Hence we have shown that KM ∗(w) = − ∂2
w ∈ D, although M ∗ is not a contraction.
∂w∂ ¯w log K(w, w) ≤ KS∗(w),
This is not an isolated example, it is easy to modify this example to produce a family of
examples parameterized by a real parameter.
In the following section, we discuss the case of a commuting tuple T = (T1, . . . , Tm) of operators
in B1(Ω), Ω ⊆ Cm, m ≥ 1. Even in this case, as before, it is possible to associate a holomorphic
Hermitian bundle ET to the operator tuple T such that the equivalence class of the commuting
m-tuple T determines the equivalence class of the bundle ET and conversely. We show that the
INFINITE DIVISIBILITY AND CURVATURE INEQUALITY
5
co-efficient matrix KT (w) of the curvature (1, 1) form KT of the holomorphic Hermitian vector
bundle ET is negative definite for each w ∈ Ω. The negativity of the curvature provides an
alternative proof of the curvature inequality given in [7].
In the third section, we show that the curvature KT is negative definite, that is,(cid:0)(cid:0)KT (wi, wj)(cid:1)(cid:1)
is negative-definite for all finite subsets {w1, . . . , wn} of Ω if we impose the additional condition
of "infinite divisibility" on the reproducing kernel K. The infinite divisibility of the kernel K
requires K t to be positive-definite for all t > 0.
In the final section, we give several applications of the positive definiteness of the curvature
function to contractivity of operators in the Cowen-Douglas class.
2. Negativity of the curvature in general
Let Ω be a bounded domain in Cm. Let T = (T1, . . . , Tm) be a m-tuple of commuting operators
on a separable complex Hilbert space H. For x ∈ H, let DT : H → H ⊕ . . . ⊕ H be the operator
defined by DT (x) = (T1x, . . . , Tmx). For w = (w1, . . . , wm) ∈ Ω, let T − w denote the operator
tuple (T1 − w1, . . . , Tm − wm). The joint kernel of T − w is ∩m
j=1 ker(Tj − wj), which is also
the kernel of the operator DT −w. Following [5], we say that the commuting tuple T belongs to
the Cowen-Douglas class Bn(Ω) if ran DT −w is closed, dim ker DT −w = n for all w in Ω, and
the span of {ker DT −w : w ∈ Ω} is dense in H. The class of the corresponding holomorphic
Hermitian vector bundle
ET = {(w, x) ∈ Ω × H : x ∈ ker DT −w}
determines the class of the operator tuple T . As before, if n = 1, then the curvature of ET (cf.
[4, 5]) determines the unitary equivalence class of T . If γ is a non-zero holomorphic section of the
holomorphic Hermitian line bundle ET defined on some open subset Ω0 ⊆ Ω, then the curvature
of the line bundle ET is the (1, 1) form
∂2 log kγ(w)k2
∂wi∂ ¯wj
dwi ∧ d ¯wj, w ∈ Ω0,
∂2 log k γ(w) k2
∂wi∂ ¯wj
(cid:1)(cid:1)m
i,j=1, w ∈ Ω0,
defined on Ω0. Let
(2.1)
KT (w) = −
mXi,j=1
KT (w) =(cid:0)(cid:0) −
denotes the curvature matrix. In general, for a holomorphic Hermitian vector bundle, there are
two well-known notions of positivity due to Nakano and Griffiths (cf.
[6, page - 338]). These
two notions coincide in the case of a line bundle, and one talks of positive line bundle in an
unambiguous manner. The following Proposition shows that the line bundle corresponding to a
commuting tuple of operators in B1(Ω) is negative.
Proposition 2.1. For an operator T in B1(Ω∗), the matrix KT (w) is negative definite for each
w ∈ Ω∗.
1 , . . . , M ∗
[5]), it follows that T can be realized as M ∗ =
First Proof. Fix w0 ∈ Ω. As before (cf.
(M ∗
m) where Mi is the multiplication operator by the co-ordinate function zi on the
Hilbert space HK of holomorphic functions on Ω0 ⊆ Ω, w0 ∈ Ω0, possessing a reproducing kernel
K with K(w, w) 6= 0 for w ∈ Ω0. The function
K0(z, w) = K(w0, w0)
1
2 ϕ(z)−1K(z, w)ϕ(w)−1K(w0, w0)
1
2
is defined on some open neighborhood U × U of (w0, w0), where U is the open set on which
K(z, w0) is non-zero and ϕ(z) = K(z, w0) is holomorphic on U . The kernel K0 is said to be
normalized at w0([5]). The operator of multiplication by the holomorphic function ϕ−1 then
defines a unitary operator from the Hilbert space HK determined by the kernel function K to
6
BISWAS, KESHARI, AND MISRA
the Hilbert space HK0 determined by the normalized kernel function K0. This unitary operator
intertwines the two multiplication operators on HK and HK0 respectively. Thus KM ∗(w0) is
equal to the curvature KM (0) ∗(w0) [5, Lemma 3.9], where M (0) is the m-tuple of multiplication
operator by the co-ordinate function zi on the Hilbert space HK0. Let
aIJ (z − w0)I ( ¯w − ¯w0)J , z, w ∈ U, I, J ∈ Zm
+ ,
be the power series expansion of K0 around the point (w0, w0). Since K0(z, w0) = 1, we have
that a00 = 1 and aI0 = 0 for all I with I > 0. Similarly, K0(w0, z) = K0(z, w0) shows that
a0J = 0 for all J with J > 0. Also note that if
bIJ (z − w0)I ( ¯w − ¯w0)J , z, w ∈ U, I, J ∈ Zm
+ ,
then b00 = 1 and bI0 = 0 = b0J for all I, J with I, J > 0. Since γ(w) = K0(·, ¯w), w ∈ U ∗ :=
∗ over U ∗, we have
K0(z, w) =XI,J
K0(z, w)−1 =XI,J
{¯z : z ∈ U } is a section of the holomorphic Hermitian line bundle EfM
∂2 log k γ(w) k2
(cid:12)(cid:12)w=w0
K0( ¯w, ¯w))(cid:12)(cid:12)w=w0
(K0( ¯w, ¯w)−1 ∂
∂wi
bIJ ( ¯w − ¯w0)I (w − w0)J )( XI≥1,J ≥0
{(1 + XI,J ≥1
∂wi∂ ¯wj
∂
∂ ¯wj
∂
∂ ¯wj
= aεj εi
=
=
aIJ +εi
(Ji+1)( ¯w − ¯w0)I (w − w0)J )}(cid:12)(cid:12)w=w0
where εi is the standard unit vector in Cm with 1 at the i-th co-ordinate and 0 elsewhere. On
the other hand, we have
aεj εi =
∂2K0( ¯w, ¯w)
∂wi∂ ¯wj
∂
∂wi
K0(·, ¯w),
∂
∂wj
Thus for any complex constants α1, . . . , αm,
−
mXi,j=1
αi ¯αj
∂2 log k γ(w) k2
∂wi∂ ¯wj
= −k
αi
∂
∂wi
mXi=1
= h
(cid:12)(cid:12)w=w0
(cid:12)(cid:12)w=w0
K0(·, ¯w)i(cid:12)(cid:12)w=w0
K0(·, ¯w)k2(cid:12)(cid:12)w=w0
.
≤ 0.
This completes the proof.
Second Proof. We show that −KT (w) is the Gramian of a set of n vectors which is explicitly
exhibited below. These vectors are
ei(w) = Kw ⊗
∂
∂ ¯wi
Kw −
∂
∂ ¯wi
Kw ⊗ Kw, 1 ≤ i ≤ n,
in HK ⊗ HK . Then
hei(w), ej (w)i = hKw ⊗
= 2(K(w, w)
Thus
∂
∂ ¯wi
Kw −
∂
∂ ¯wi
∂2K(w, w)
∂wi∂ ¯wj
Kw ⊗ Kw, Kw ⊗
−
∂
∂wi
K(w, w)
∂ ¯wj
∂
∂ ¯wj
∂
Kw −
∂
∂ ¯wj
Kw ⊗ Kwi
K(w, w)).
∂2 log k γ(w) k2
∂wi∂ ¯wj
(cid:12)(cid:12)w=w0
=
=
K(w, w) ∂2 K(w,w)
∂wi∂ ¯wj
K(w, w) ∂
∂ ¯wj
− ∂
∂wi
K(w, w)2
K(w, w)
hei(w0), ej (w0)i
2K(w0, w0)2
.
(cid:12)(cid:12)w=w0
INFINITE DIVISIBILITY AND CURVATURE INEQUALITY
This completes the proof.
7
(cid:3)
A commuting tuple of operators T = (T1, . . . , Tm) is said to be a row contraction ifPm
I. The following characterization of row contractions is well known (cf. [7, Corollary 2]).
i=1 TiT ∗
i ≤
Lemma 2.2. Let Bm be the unit ball in Cm and M = (M1, . . . , Mm) be m-tuples of multiplication
operator on reproducing kernel Hilbert space with reproducing kernel K. Then M is a row
contraction if and only if (1 − hz, wi)K(z, w) is positive definite.
1
Let R∗
m be the adjoint of the joint weighted shift operator on the Drury-Arveson space H 2
m.
This is the commuting tuple (M ∗
m which is determined by the reproducing kernel
1−hz,wi , z = (z1, . . . , zm), w = (w1, . . . , wm) ∈ Bm. As in Remark 1.4, using Proposition 2.1 and
Lemma 2.2, we obtain a version of curvature inequality for the multi-variate case. It appeared
earlier in [7] with a different proof.
1 , . . . , M ∗
m) on H 2
Corollary 2.3. If T = (T1, . . . , Tm) is a row contraction in B1(Bm), then KR∗
positive for each w in the unit ball Bm.
m(w) − KT (w) is
3. Infinite divisibility and curvature inequality
Starting with a positive definite kernel K on a bounded domain Ω in Cm, it is possible to
construct several new positive definite kernel functions. For instance, if K is positive definite
then the kernel K n, n ∈ N, is also positive definite. Indeed, a positive definite kernel K is said to
be infinitely divisible if for all t > 0, the kernel K t is also positive definite. While the Bergman
kernel for the Euclidean ball is easily seen to be infinitely divisible, it is not infinitely divisible
for the unit ball (with respect to the operator norm) of the n × n matrices. We give the details
for n = 2 in the final Section of this note. The following Lemma shows that if K is positive
Lemma 3.1. For any bounded domain Ω in Cm, if K defines a positive definite kernel on Ω,
definite then the matrix valued kernel(cid:0)(cid:0)
then −K(z, w) =(cid:0)(cid:0)
K(z, w)(cid:1)(cid:1)m
∂zi ∂ ¯wj
i,j=1
∂2
∂2
∂zi ∂ ¯wj
K(z, w)(cid:1)(cid:1)m
i,j=1 is positive definite as well.
is a positive definite kernel on Ω.
Proof. Let ξi = (ξi(1), . . . , ξi(m)), 1 ≤ i ≤ m, be vectors in Cm and u1, . . . , un be an arbitrary
set of n points in Ω. Since ¯∂iKw belongs to HK, as shown in [5], it follows that
h−K(ui, uj)ξj, ξiiCm =
nXi,j
=
∂2
∂wk ∂ ¯wl
K(cid:1)(ui, uj)ξj(l)ξi(k)
h ∂
∂ ¯wl
Kuj , ∂
∂ ¯wk
KuiiHK ξj(l)ξi(k)
nXi,j
mXk,l (cid:0)
mXk,l
nXi,j
nXi
mXk
= k
≥ 0
ξi(k) ∂
∂ ¯wk
Kuik2
HK
This completes the proof.
(cid:3)
Remark 3.2. Even in the case of one variable, the proof of the Lemma given is interesting. In
fact, this motivates the proof of the main theorem (Theorem 3.6) in one direction. In particular,
it says that if K is a positive definite kernel on a bounded domain Ω ⊂ C, then (cid:0) ∂2
let K(z, w) = P∞
is also a positive definite kernel on Ω. Without loss of generality, assume that 0 is in Ω and
m,n amnzm ¯wn be the power series expansion of K around 0. It is shown in [5,
∂z ∂ ¯w K(cid:1)(z, w)
8
BISWAS, KESHARI, AND MISRA
Lemma 4.1 and 4.3] that the positivity of the kernel K is equivalent to the positivity of the
matrix of Taylor co-efficients of K at 0, namely,
a00 a01 a02
a10 a11 a12
...
...
an0 an1 an2
...
· · ·
· · ·
. . .
· · ·
a0n
a1n
...
ann
(m + 1)(n + 1)a(m+1)(n+1)zm ¯wn.
Hn(0; K) :=
∞Xm,n=0
for each n ∈ Z+. The function ∂2
∂z ∂ ¯w K(z, w) admits the expansion
Therefore, for n ∈ N,
Hn−1(0;
∂2
∂z ∂ ¯w
Clearly, for n ∈ N, we have
na1n
2na2n
...
a11
2a21
...
2a12
4a22
...
· · ·
· · ·
. . .
· · · n2ann
K) =
∂z ∂ ¯w K)(cid:17) = D(cid:0)Hn(0; K)(cid:1)D,
nan1 2nan2
.
01×n
0n×1 Hn−1(0; ∂2
(cid:16) 01×1
where D : Cn+1 → Cn+1 is the linear map which is diagonal and is determined by the sequence
∂2
∂z ∂ ¯w K) is positive definite for all n ∈ Z+.
{0, 1, . . . , k, . . . , n}. It therefore follows that Hn(0;
Consequently,
∂2
∂z ∂ ¯w K is a positive definite kernel.
The following Lemma encodes a way to extract scalar valued positive definite kernel from the
matrix valued one.
Lemma 3.3. If K is a n×n matrix valued positive definite kernel on a bounded domain Ω ⊂ Cm,
then for every ζ ∈ Cn, hK(z, w)ζ, ζiCn is also a positive definite kernel on Ω.
Proof. Let Kζ(z, w) = hK(z, w)ζ, ζiCn . Let u1, . . . , ul be l points in Ω and αi, 1 ≤ i ≤ l, be
scalars in C. From [5], it follows that
αiKζ(ui, uj)¯αj =
lXi,j
lXi,j
lXj
= k
≥ 0
αi ¯αjhK(·, uj)ζ, K(·, ui)ζiHK
¯αjK(·, uj)ζk2
HK
This completes the proof.
(cid:3)
Definition 3.4. Let G be a real analytic function of w, ¯w for w in some open connected subset Ω
in the first variable and anti-holomorphic in the second and restricts to G on the diagonal set
{(w, w) : w ∈ Ω}, that is, G(w, w) = G(w, w), w ∈ Ω. If the function G is also positive definite,
of Cn. Polarizing G, we obtain a (unique) new function eG defined on Ω × Ω which is holomorphic
that is, the n × n matrix(cid:0)(cid:0)eG(wi, ¯wj)(cid:1)(cid:1) is positive definite for all finite subsets {w1, . . . , wn} of Ω,
then we say that G is a positive definite function on Ω.
INFINITE DIVISIBILITY AND CURVATURE INEQUALITY
9
The curvature K of a line bundle is a real analytic function. We have shown that −K(w),
w ∈ Ω ⊂ Cm, is positive definite. However, the following example shows that −K(w) need
convention that the positive definiteness of the real analytic function −K(w) is the same as the
not be a positive definite function, that is, −eK(w) need not be positive definite! We adopt the
positive definiteness of the Hermitian function −eK(w).
Example 3.5. Let K(z, w) = 1 +P∞
∞Xi=1
i=1 aizi ¯wi be a positive definite kernel on the unit disc D.
The kernel K then admits a power series expansion some small neighborhood of 0. Consequently,
we have
log K(z, w) = log(1 +
aizi ¯wi)
=
aizi ¯wi −
∞Xi=1
= a1z ¯w + (a2 − a2
1
i=1 aizi ¯wi)2
(P∞
3
2 )z2 ¯w2 + (a3 − a1a2 + a3
(P∞
+
2
i=1 aizi ¯wi)3
− · · ·
1
3 )z3 ¯w3 + . . .
It follows that
(cid:0) ∂2
∂z ∂ ¯w log K(cid:1)(z, w) = a1 + 4(a2 − a2
Thus if we choose 0 < ai, i ∈ N, such that a2 < a2
that
∂2
∂z ∂ ¯w log K is not positive definite.
1
2 )z ¯w + 9(a3 − a1a2 + a3
1
3 )z2 ¯w2 + . . .
1
2 , then from [5, Lemma 4.1 and 4.3], it follows
However we note, for instance, that if K is the function 1 + z ¯w + 1
K t(z, w) = 1 + tz ¯w +
t(2t − 1)
4
z2 ¯w2 + · · ·
is not positive definite for t < 1
2 .
4 z2 ¯w2 +P∞
i=3 zi ¯wi, then
It is therefore natural to ask if assuming that K is infinitely divisible is both necessary and
sufficient for positive definiteness of the curvature function −K. The following Theorem provides
an affirmative answer.
Theorem 3.6. Let Ω be a domain in Cm and K be a positive real analytic function on Ω ×
If K is infinitely divisible then there exist a domain Ω0 ⊆ Ω such that the curvature
Ω.
is positive definite function on Ω0. Conversely, if the function
∂2
and a infinitely divisible kernel K on Ω0 × Ω0 such that K(w, w) = K(w, w), for all w ∈ Ω0.
is a positive definite on Ω, then there exist a neighborhood Ω0 ⊆ Ω of w0
matrix (cid:0)(cid:0)
(cid:0)(cid:0)
∂wi ∂ ¯wj
∂2
∂wi ∂ ¯wj
log K(cid:1)(cid:1)m
log K(cid:1)(cid:1)m
i,j=1
i,j=1
Proof. For each t > 0, assume that K t is positive definite on Ω. This is the same as the positive
definiteness of exp t log K, t > 0. Clearly t−1(exp t log K − 1) is conditionally positive definite
(An Hermitian kernel L is said to be conditionally positive definite if for every n ∈ N and for
i=1 αi = 0, the inequality
i,j=1 αi ¯αjL(wi, wj) ≥ 0 holds). By letting t tend to 0, it follows that log K is conditionally
every choice n points w1, . . . , wn and complex scalars α1, . . . , αn withPn
Pn
positive definite. Hence at an arbitrary point in Ω, in particular at w0, the kernel
Lw0(z, w) = log K(z, w) − log K(z, w0) − log K(w0, w) + log K(w0, w0)
∂2
∂wi ∂ ¯wj
is positive definite. This is essentially the Lemma 1.7 in [9]. From Lemma 3.1, it follows that the
w0 such that log K(z, w0) is holomorphic on Ω0. Hence from the equation above, the curvature
Lw0(cid:1)(cid:1) is positive definite on Ω. Note that there exist a neighborhood Ω0 ⊆ Ω of
log K(cid:1)(cid:1) is positive definite on Ω0. This proves the Theorem in the forward
matrix (cid:0)(cid:0)
matrix (cid:0)(cid:0)
direction.
∂wi ∂ ¯wj
∂2
10
BISWAS, KESHARI, AND MISRA
For the other direction, without loss of generality, assume that w0 = 0. Let K(z, w) be the
function obtained by polarizing the real analytic m × m matrix valued function
∂2
∂wi∂ ¯wj
(cid:0)(cid:0)
log K(w, w)(cid:1)(cid:1)m
i,j=1
defined on some bounded domain Ω in Cm. Suppose that log K has the power series expansion
m , ¯wJ =
P aIJ zI ¯wJ , where the sum is over all multi-indices I, J of length m and zI = zi1
1 · · · ¯wjm
1 · · · zim
m . Then
¯wj1
K(z, w) =XI,J
aIJ(cid:0)(cid:0)AIJ (k, ℓ)zI−ǫk ¯wJ−ǫℓ(cid:1)(cid:1),
where AIJ (k, ℓ) =(cid:0)(cid:0)ik jℓ(cid:1)(cid:1)m
and the sum is again over all multi-indices I, J of size m. Clearly,
AIJ can be written as the product D(I) Em D(J), where D(I) and D(J) be the m × m diagonal
matrices with (i1, . . . , im) and (j1, . . . , jm) on their diagonal respectively, and Em is the m × m
matrix all of whose entries are 1.
k,ℓ=1
Let D(z) be the holomorphic function on Ω taking values in the m × m diagonal matrices
which has zi in the (i, i) position for z := (z1, z2, . . . , zm) ∈ Ω. If the function K is assumed to
be positive definite then
eK(z, w) := D(z) K(z, w) D( ¯w) =XI,J
aIJ D(I) Em D(J)zI ¯wJ
is positive definite on Ω0.
Let Λ(I) = {k : 1 ≤ k ≤ m and ik 6= 0}. Consider the m × m matrix E(I, J) defined below:
E(I, J)ij =(1 if i ∈ Λ(I) and j ∈ Λ(J),
0 otherwise.
Note that if Λ(I) = Λ(J) = {1, . . . , m}, then E(I, J) = Em. Consider the function on Ω0 × Ω0,
defined by
beK(z, w) = XI,J6=0
aIJ
E(I,J)
c(I)c(J) zI ¯wJ ,
kernel on Ω0. To facilitate the proof, we need to fix some notations.
where c(I) denotes the cardinality of the set Λ(I). We will prove that beK is a positive definite
Let δ be a multi-index of size m. Also let p(δ) =Qm
j=1(δj + 1) which is the number of multi-
indices I ≤ δ, that is, il ≤ δl, 1 ≤ l ≤ m. As par the notation in [5], given a function L on
a domain U × U which is holomorphic in the first variable and antiholomorphic in the second,
let Hδ(w0; L) be the p(δ) × p(δ) matrix whose (I, J)-entry is ∂I ¯∂J L(w0,w0)
, 0 ≤ I, J ≤ δ. For
convenience, one uses the colexicographic order to write down the matrix, that is, I ≤c J if and
only if (im < jm) or (im = jm and im−1 < jm−1) or · · · or (im = jm and . . . i2 = j2 and i1 < j1)
or I = J.
I!J!
Let D(I)♯ be the diagonal matrix with the diagonal entry D(I)♯
ℓ ℓ equal to 1
iℓ
or 0 according
as iℓ is non-zero or zero. Using this notation, we have
D(I)♯D(I) Em D(J)D(J)♯ = E(I, J).
Let Aδ be the block diagonal matrix, written in the colexicographic ordering, of the form
Therefore, in this setup, for any multi-index δ, we have
if I = J(6= 0)
otherwise.
D(I)♯
c(I)
0
(Aδ)IJ =
Hδ(0;beK) = AδHδ(0;eK)A∗
δ.
INFINITE DIVISIBILITY AND CURVATURE INEQUALITY
11
Clearly Hδ(0;beK) is positive definite since Hδ(0;eK) is so by [5, Lemma 4.1]. Thus from [5, Lemma
4.3], it follows that beK is a positive definite kernel.
Let K0 be the scalar function on Ω0 × Ω0 defined by
where the sum is over all pairs (I, J) excluding those of the form (I, 0) and (0, J). From Lemma
3.3, it follows that the function K0 is positive definite since it is of the form hbeK(z, w)1, 1i,
1 = (1, . . . , 1). It is evident that
K0(z, w) :=X aIJ zI ¯wJ ,
∂2
∂wi∂ ¯wj
(cid:16)(cid:16)(cid:0)
K0(cid:1)(w, w)(cid:17)(cid:17) = K(w, w),
that is,
∂2
∂wi ∂ ¯wj
(log K − K0(w, w)) = 0, 1 ≤ i, j ≤ m, w ∈ Ω0.
Therefore, log K − K0 is a real pluriharmonic function on Ω0 and hence there exist a holomor-
phic function φ such that
log K(w, w) − K0(w, w) = (ℜφ)(w) :=
φ(w) + φ(w)
2
.
(Alternatively, since log K is real analytic, it follows that
XI,J
aIJ wI ¯wJ = XI,J
¯aIJ wJ ¯wI
Equating coefficients, we get aIJ = ¯aJ I for all multi-indices I, J.
aI0 = a0I for all multi-indices I. The power series
(a00/2) +XI
aI0zI
In particular, we have
defines a holomorphic function ψ on Ω0 such that log K(w, w) − K0(w, w) = ψ(w) + ψ(w).)
Thus
K(w, w) = exp( φ(w)
2 ) exp(K0(w, w)) exp( φ(w)
2 ), w ∈ Ω0.
Let K : Ω0 × Ω0 → C be the function defined by the rule
K(z, w) = exp( φ(z)
2 ) exp(K0(z, w)) exp( φ(w)
2 ).
For t > 0, we then have
K t(z, w) = exp(t φ(z)
2 ) exp(tK0(z, w)) exp(t φ(w)
2 ), z, w ∈ Ω0.
By construction K(w, w) = K(w, w), w ∈ Ω0. Since K0 is a positive definite kernel as shown
above, it follows from [9, Lemma 1.6] that exp(tK0) is a positive definite kernel for all t > 0 and
therefore K t is a positive definite on Ω0 for all t > 0 completing the proof of the converse.
(cid:3)
4. Applications
Let M ∗ be the adjoint of the commuting tuple of multiplication operators acting on the Hilbert
space HK ⊆ O(Ω). Fix a positive definite kernel K on Ω. Let us say that M is infinitely divisible
with respect to K if K(z, w)−1K(z, w) is infinitely divisible in some open subset Ω0 of Ω. As an
immediate application of Theorem 3.6, we obtain :
12
BISWAS, KESHARI, AND MISRA
Theorem 4.1. Assume that the adjoint M ∗ of the multiplication operator M on the reproducing
kernel Hilbert space HK belongs to B1(Ω). The function (cid:0)(cid:0)
positive definite, if and only if the multiplication operator M is infinitely divisible with respect
to K.
log(cid:0)K(w, w)−1K(w, w)(cid:1)(cid:1)(cid:1) is
∂2
∂wi ∂ ¯wj
If K is a positive definite kernel on D such that (1 − z ¯w)K(z, w) is infinitely divisible then we
say that M ∗ on HK is a infinitely divisible contraction.
Here is an example showing that a contraction need not be infinitely divisible. Take
K(z, w) = (1 − z ¯w)−1(cid:0)1 + z ¯w + 1
4 z2 ¯w2 +
∞Xn=3
zn ¯wn(cid:1)
= 1 + 2z ¯w +
(n + 1
4 )zn ¯wn.
∞Xn=2
Clearly K defines the positive definite kernel on D. Since (1−z ¯w)K(z, w) is also positive definite,
it follows that the adjoint of multiplication operator M ∗ on HK is contractive. But
(cid:0)(1 − z ¯w)K(z, w)(cid:1)t = 1 + tz ¯w +
t(2t − 1)
4
z2 ¯w2 + · · ·
2 as was pointed out earlier. Hence M ∗ is not infinitely divisible
is not positive definite for t < 1
contraction on HK .
The following Corollary is a characterization of infinitely divisible contractions in the Cowen-
Douglas class B1(D) completing the study begun in [8]. Here, for two real analytic functions G1
and G2 on a domain Ω ⊂ Cm, G1(w) (cid:22) G2(w), w ∈ Ω, means G2 − G1 is a positive definite
function on Ω.
Corollary 4.2. Let K be a positive definite kernel on the open unit disc D. Assume that the
adjoint M ∗ of the multiplication operator M on the reproducing kernel Hilbert space HK belongs
to B1(D). The function
∂w ∂ ¯w log(cid:0)(1 − w2)K(w, w)(cid:1) is positive definite, or equivalently
KT (w) (cid:22) KS∗(w), w ∈ D,
∂2
if and only if the multiplication operator M is an infinitely divisible contraction.
Proof. Recall Theorem 3.6, which says that the positive definiteness of
∂2
∂w ∂ ¯w
log(cid:0)(1 − w2)K(w, w)(cid:1)
positive definite for all t ≥ 0.
is equivalent to infinite divisibility of the kernel (1 − z ¯w)K(z, w), that is, (cid:0)(1 − z ¯w)K(z, w)(cid:1)t is
contraction if (1 − hz, wi)K(z, w) is infinitely divisible, that is,(cid:0)(1 − hz, wi)K(z, w)(cid:1)t is positive
We say that a commuting tuple of multiplication operators M is an infinitely divisible row
Recall that R∗
m. The following theorem is a characterization of infinitely divisible row contractions.
m is the adjoint of the joint weighted shift operator on the Drury-Arveson space
definite for all t > 0.
H 2
(cid:3)
Corollary 4.3. Let K be a positive definite kernel on the Euclidean ball Bm. Assume that the
adjoint M ∗ of the multiplication operator M on the reproducing kernel Hilbert space HK belongs
, w ∈ Bm, is positive definite,
∂2
to B1(Bm). The function (cid:0)(cid:0)
or equivalently
∂wi ∂ ¯wj
log(cid:0)(1 − hw, wi)K(w, w)(cid:1)(cid:1)(cid:1)m
i,j=1
KM ∗ (w) (cid:22) KR∗
m(w), w ∈ Bm,
if and only if the multiplication operator M is an infinitely divisible row contraction.
INFINITE DIVISIBILITY AND CURVATURE INEQUALITY
13
We give one last example, namely that of the polydisc Dm. In this case, we say a commuting
tuple M of multiplication by the co-ordinate functions acting on the Hilbert space HK is infinitely
i=1(1 − zi ¯wi)−1, z, w ∈ Dm, is positive definite
for all t > 0. Every commuting tuple of contractions M ∗ need not be infinitely divisible. Let
Sm be the commuting m - tuple of the joint weighted shift defined on the Hardy space H 2(Dm).
divisible if(cid:0)S(z, w)K(z, w)(cid:1)t, where S(z, w) :=Qm
Corollary 4.4. Let K be a positive definite kernel on the polydisc Dm. Assume that the adjoint
M ∗ of the multiplication operator M on the reproducing kernel Hilbert space HK belongs to
, w ∈ Dm, is positive definite, or
∂2
B1(Dm). The function (cid:0)(cid:0)
equivalently
∂wi ∂ ¯wj
log(cid:0)S(z, w)K(w, w)(cid:1)(cid:1)(cid:1)m
i,j=1
KM ∗(w) (cid:22) KS∗
m(w), w ∈ Dm,
if and only if the multiplication operator M is an infinitely divisible m-tuple of contractions.
For a second application of these ideas, assume that K is a positive definite kernel on Dm with
the property:
Ki(z, w) = (1 − zi ¯wi)mK(z, w), 1 ≤ i ≤ m,
is infinitely divisible. Then
It now follows that
Ki(z, w).
mYi=1
Ki(z, w)(cid:1) 1
K m(z, w) =(cid:0) mYi=1
(1 − zi ¯wi)(cid:1)−m
K(z, w) = S(z, w)(cid:0) mYi=1
i=1 Ki(z, w)(cid:1) 1
m .
Let M be the commuting tuple of multiplication operators on the Hilbert space HK , which is
contractive since K admits the Szego kernel S as a factor. Clearly, the infinite divisibility of
m is positive definite. As pointed out in [7], in
Ki, 1 ≤ i ≤ m, implies that (cid:0)Qm
consequence, for any polynomial p in m - variables,
p(M1, . . . , Mm) = PSp(Sm)S,
where S is the invariant subspace of functions vanishing on the diagonal of the Hilbert space
H 2
m ⊗ HK ⊆ O(Dm × Dm) and PS is the projection onto the subspace S.
A basic question raised in the paper of Cowen and Douglas [3, Section 4] is the determination
of non-degenerate holomorphic curves in the Grassmannian. Clearly, a necessary condition for
this is the positive definiteness of the curvature function. Thus we have the following corollary
to Theorem 3.6.
Corollary 4.5. Let E be a holomorphic Hermitian vector bundle of rank 1 over a bounded
domain Ω ⊂ Cm. If the curvature K of E is negative definite, then there exists a Hilbert space
H and a holomorphic map γ : Ω0 → H, Ω0 open in Ω, such that E is isomorphic to Eγ , where
Eγ is the pullback, by the holomorphic map γ : Ω0 → Gr(1, H), of the tautological bundle defined
over Gr(1, H). Moreover, the real analytic function hγ(z), γ(w)i defined on Ω0 × Ω0 is infinitely
divisible.
We finish with an amusing application of the Lemma 1.7 in [9] which is a key ingredient in the
proof of Theorem 3.6. Let K be the function on the unit ball B2×2 (with respect to the operator
norm) of 2 × 2 matrices, given by the formula K(Z, W ) := det(I − ZW ∗)−1, Z, W ∈ B2×2.
The kernel K is normalized at 0 by definition. For δ = (1, 0, 0, 3), the matrix
(cid:0)(cid:0) ∂α ¯∂β log K(0, 0)
α!β!
(cid:1)(cid:1)0≤α,β≤δ
14
BISWAS, KESHARI, AND MISRA
¯∂1 ¯∂3
4
4 log K(0,0)
3!3!
is diagonal with ∂1∂3
= −1 < 0 (in fact for δ ≤ 3, the corresponding matrix is
diagonal with non-negative entries). Here, δ ≥ µ if and only if δi ≥ µi for all i ∈ {1, . . . , m}
and the matrix is written with respect to the colexicographic ordering. From [5, Lemma 4.1 and
4.3], it follows that log K is not positive definite. Hence Theorem 3.6 shows that the function
det(I − ZW ∗)−t cannot be positive definite for all t > 0. Of course, a lot more is known. Indeed,
the set
{0 < t : K(Z, W )t is positive definite}
is explicitly determined in [2, Corollary 4.6].
References
1. J. Agler and J. E. McCarthy, Pick interpolation and Hilbert function spaces, Graduate Studies in Mathematics,
vol. 44, American Mathematical Society, Providence, RI, 2002. MR 1882259 (2003b:47001)
2. J. Arazy, A survey of invariant Hilbert spaces of analytic functions on bounded symmetric domains, Multivari-
able operator theory (Seattle, WA, 1993), Contemp. Math., vol. 185, Amer. Math. Soc., Providence, RI, 1995,
pp. 7 -- 65. MR 1332053 (96e:46034)
3. M. J. Cowen and R. G. Douglas, Complex geometry and operator theory, Acta Math. 141 (1978), no. 3-4,
187 -- 261. MR MR501368 (80f:47012)
4.
, Operators possessing an open set of eigenvalues, Functions, series, operators, Vol. I, II (Budapest,
1980), Colloq. Math. Soc. János Bolyai, vol. 35, North-Holland, Amsterdam, 1983, pp. 323 -- 341. MR MR751007
(85k:47033)
5. R. E. Curto and N. Salinas, Generalized Bergman kernels and the Cowen-Douglas theory, Amer. J. Math. 106
(1984), no. 2, 447 -- 488. MR MR737780 (85e:47042)
6. J. P Demailly, Complex analytic and differential geometry, http://www-fourier.ujf-grenoble.fr/ de-
mailly/manuscripts/agbook.pdf.
7. R. G. Douglas, G. Misra, and J. Sarkar, Contractive Hilbert modules and their dilations, to appear, Israel J.
Math.
8. G. Misra, Curvature inequalities and extremal properties of bundle shifts, J. Operator Theory 11 (1984), no. 2,
305 -- 317. MR 749164 (86h:47057)
9. K. R. Parthasarathy and K. Schmidt, Positive definite kernels, continuous tensor products, and central limit the-
orems of probability theory, Lecture Notes in Mathematics, Vol. 272, Springer-Verlag, Berlin, 1972. MR 0622034
(58 #29849)
(Biswas) Department of Mathematics, Ben Gurion University of the Negev, Be'er Sheva 84105,
Israel
E-mail address: [email protected]
(Keshari) Department of Mathematics, Indian Institute of Science, Bangalore 560012, India
E-mail address: [email protected]
(Misra) Department of Mathematics, Indian Institute of Science, Bangalore 560012, India
E-mail address: [email protected]
|
1202.4858 | 2 | 1202 | 2012-02-26T14:05:47 | Completeness of eigenfunctions of Sturm-Liouville problems with discontinuities at three points | [
"math.FA",
"math.CA"
] | In this work, we study discontinuous Sturm-Liouville type problems with eigenparameter dependent boundary condition and transmission conditions at three interior points. A self-adjoint linear operator A is defined in a suitable Hilbert space H such that the eigenvalues of such a problem coincide with those of A. We show that the eigenvalues of the problem are analytically simple, and the eigenfunctions of A are complete in H. | math.FA | math |
Completeness of eigenfunctions of Sturm-Liouville
problems with discontinuities at three points
Erdogan S¸en
Department of Mathematics, Faculty of Science and Letters, Namık Kemal University, 59030
Tekirdag, Turkey.
e-mail: [email protected]
MSC (2010): 34L10, 47E05
Keywords : Sturm-Liouville problems; eigenparameter-dependent bound-
ary conditions; transmission conditions; completeness; eigenvalues; eigenfunc-
tions.
ABSTRACT.
In this work, we study discontinuous Sturm-Liouville type problems with
eigenparameter dependent boundary condition and transmission conditions at
three interior points. A self-adjoint linear operator A is defined in a suitable
Hilbert space H such that the eigenvalues of such a problem coincide with those
of A. We show that the eigenvalues of the problem are analytically simple, and
the eigenfunctions of A are complete in H.
1 Introduction
It is well-known that many topics in mathematical physics require the investiga-
tion of eigenvalues and eigenfunctions of Sturm-Liouville type boundary value
problems. In recent years, more and more researches are interested in the dis-
continuous Sturm-Liouville problems with eigenparameter-dependent boundary
conditions (see [1 − 4]). The literature on this subject is voluminous and we
refer to [5 − 10]. Various physics applications of this kind problem are found
in many literatures, including some boundary value problem with transmission
conditions that arise in the theory of heat and mass transfer (see [2, 10, 11]). The
study of the structure of the solution in the matching region of the layer with
the basis solution in the plate leads to consideration of an eigenvalue problem
for a second order differential operator with piecewise continuous coefficients
and transmission conditions [12].
Sturm-Liouville problems with transmission conditions have been studied
by many authors (see [3, 4, 8, 13]). Adjoint and self-adjoint boundary value
problems with interface conditions have been stutied in [14, 15]. Such problems
with point interactions are also studied in [16].
In this study, we also deal with the class of problems (1) − (9), by means of
1
a combination of the methods [5],[13],[17] and[18]. In Section 1, a self-adjoint
linear operator A is defined in a suitable Hilbert space H such that the eigen-
values of the problem (1) − (9) coincide with those of A. In Section 2, we prove
that the eigenvalues of the problem (1) − (9) are analytically simple. In Section
3, we prove that the eigenfunctions of A are complete in H. Note that each
eigenfunction of the original problem and a real number.
In this study, we consider a discontinuous eigenvalue problem which consists
of Sturm-Liouville equation
τ u :=(cid:16)−p (x) u
′
(x)(cid:17)′
+ q(x)u (x) = λu (x)
(1)
on I = [−1, h1) ∪ (h1, h2) ∪ (h2, h3) ∪ (h3, 1] , where p (x) = p2
1 for x ∈ [−1, h1);
p (x) = p2
4, for
x ∈ (h3, 1] ; p1, p2, p3, p4 are nonzero real constants, q (x) ∈ L1 (I, R) and λ ∈ C
is the eigenparameter; with the boundary condition
3, for x ∈ (h2, h3) and p (x) = p2
2, for x ∈ (h1, h2) , p (x) = p2
τ1u := α1u (−1) + α2u
′
(−1) = 0,
the eigenparameter-dependent boundary condition
τ2u := λhβ
′
1u (1) − β
′
′
2u
(1)i +hβ1u (1) − β2u
′
(1)i = 0,
and the transmission conditions
τ3u := u (h1 + 0) − α3u (h1 − 0) − β3u
′
(h1 − 0) = 0,
′
τ4u := u
(h1 + 0) − α4u (h1 − 0) − β4u
′
(h1 − 0) = 0,
τ5u := u (h2 + 0) − α5u (h2 − 0) − β5u
′
(h2 − 0) = 0,
′
τ6u := u
(h2 + 0) − α6u (h2 − 0) − β6u
′
(h2 − 0) = 0,
τ7u := u (h3 + 0) − α7u (h3 − 0) − β7u
′
(h3 − 0) = 0,
′
τ8u := u
(h3 + 0) − α8u (h3 − 0) − β8u
′
(h3 − 0) = 0,
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
where the coefficients αi, βi and β
out this paper, we assume that
θ =(cid:12)(cid:12)(cid:12)(cid:12)
and α1 + α2 6= 0.
′
α5 β5
j(cid:0)i = 1, 8, j = 1, 2(cid:1) are real numbers. Through-
α6 β6 (cid:12)(cid:12)(cid:12)(cid:12) > 0, ξ =(cid:12)(cid:12)(cid:12)(cid:12)
2 β2 (cid:12)(cid:12)(cid:12)(cid:12) > 0,
α8 β8 (cid:12)(cid:12)(cid:12)(cid:12) > 0,
α7 β7
1 β1
α3 β3
α4 β4 (cid:12)(cid:12)(cid:12)(cid:12) > 0, γ =(cid:12)(cid:12)(cid:12)(cid:12)
ρ =(cid:12)(cid:12)(cid:12)(cid:12)
′
′
β
β
2
2 Operator formulation
The relation between a symmetric linear operator A defined in a suitable Hilbert
space H and the problem (1) − (9) has been introduced in [13]. Here, we repeat
the definition and prove that the operator A is self-adjoint, not only symmetric.
We define the inner product in L2 (I) as
1
p2
1
h1Z
−1
f1g1 +
1
p2
2θ
h2Z
f2g2 +
h1
1
p2
3θγ
1Z
h2
f3g3 +
1
p2
4θγξ
1Z
h2
f4g4, ∀f, g ∈ L2 (I) ,
hf, gi1 =
where
x ∈ [−1, h1),
limx→h1+0 f (x),
limx→h1−0 f (x), x = h1,
f1 (x) :=(cid:26) f (x),
f3 (x) :=
It is easy to verify that (cid:0)L2 (I) , h·, ·i1(cid:1) is a Hilbert space. For simplicity, it is
f2 (x) :=
f4 (x) :=(cid:26) limx→h3+0 f (x),
limx→h3−0 f (x),
limx→h2−0 f (x),
limx→h2+0 f (x),
denoted by H1.
x ∈ (h2, h3) ,
x = h3,
x = h2,
f (x) ,
f (x),
f (x),
x = h1,
x ∈ (h1, h2) ,
x = h2,
x = h3,
x ∈ (h3, 1]
The inner product in H := H1 ⊕ C is defined by
hF, Gi = hf, gi1 +
1
ρθγξ
hk
for F = (f (x) , h), G = (g (x) , k) ∈ H, where f, g ∈ H1, h, k ∈ C.
We define the operator A in H as follows:
D (A) :=n(f (x) , h) ∈ H f1, f
3 ∈ ACloc ((h2, h3)) , f4, f
′
′
f3, f
′
1 ∈ ACloc ((−1, h1)) , f2, f
′
2 ∈ ACloc ((h1, h2)) ,
4 ∈ ACloc ((h3, 1)) , τ f ∈ H1, τ1f = τ3f = τ4f = τ5f = τ6f,
h = β
′
1f (1) − β
′
′
2f
(1) ,o
AF =(cid:16)τ f, −(cid:16)β1f (1) − β2f
′
(1)(cid:17)(cid:17) for F =(cid:16)f, β
′
1f (1) − β
′
2f (1)(cid:17) ∈ D (A).
Note that by our assumption on q (x) and Theorem 3.2 in [19], for each
2 are continuous on
4 are continuous on [h3, 1].
3 are continuous on [h2, h3] and f4, f
1 are continuous on [−1, h1], f2, f
(f, h) ∈ D (A), f1, f
[h1, h2] , f3, f
For simplicity, for (f, h) ∈ D (A) , set
′
′
′
′
N (f ) = β1f (1) − β2f
′
′
(1) , N
(f ) = β
′
1f (1) − β
′
′
2f
(1) .
So, we can study the problem (1)−(9) in H by considering the operator equation
AF = λF . Obviously, we have
3
Lemma 1.1 The eigenvalues of the boundary value problem (1) − (9) co-
incide with those of A, and its eigenfunctions are the first components of the
corresponding eigenfunctions of A.
Lemma 1.2 The domain D (A) is dense in H.
Proof. Suppose that F ∈ H is orthogonal to all G ∈ D (A) with respect to the
0 denote the
inner product h·, ·i, where F = (f (x) , h), G = (g (x) , k). Let eC∞
set of functions
φ (x) =
ϕ1 (x) , x ∈ [−1, h1) ,
ϕ2 (x) , x ∈ (h1, h2) ,
ϕ3 (x) , x ∈ (h2, h3) ,
ϕ4 (x) , x ∈ (h3, 1] ,
0 ⊕0 ⊂ D (A) (0 ∈ C), any U = (u (x) , 0) ∈ eC∞
0 (h1, h2) , ϕ3 (x) ∈ C∞
0 (h2, h3) and
0 ⊕0
0 [−1, h1), ϕ2 (x) ∈ C∞
where ϕ1 (x) ∈ C∞
ϕ4 (x) ∈ C∞
is orthogonal to F , namely,
0 (h3, 1]. Since eC∞
hF, U i =
1
p2
1
h1Z
−1
f (x) u (x)dx +
1
p2
2θ
h2Z
f (x) u (x)dx +
h1
1
p2
3θγ
h3Z
f (x) u (x)dx
h2
+
1
p2
4θγξ
1Z
h3
f (x) u (x)dx = hf, ui1 .
This implies that f (x) is orthogonal to eC∞
hF, Gi = 1
ρθγξ hk = 0. Thus, h = 0 since k = N
So, F = (0, 0). Therefore, D (A) is dense in H.
0
in H1 and hence vanishes. So,
(g) can be chosen arbitrarily.
′
Theorem 1.1. The linear operator A is self-adjoint in H.
Proof. For all F, G ∈ D (A), (2) implies that f (−1) g
and direct calculations using (4) and (5) then yield that
′
(−1)−f
′
(−1) g (−1) = 0,
hAF, Gi = hF, AGi + W (f, g; h1 − 0) − W (f, g; −1) +
1
θ
W (f, g; h2 − 0) −
1
θ
W (f, g; h1 + 0) +
1
θγ
W (f, g; 1) −
1
θγξ
= hF, AGi ,
1
θγξ
W (f, g; h3 + 0) −
1
θγ
1
W (f, g; h3 − 0) −
W (f, g; h2 + 0) +
ρθγξ (cid:16)N (f ) N ′ (g) − N
′
(f ) N (g)(cid:17)
where (f, g; x) denotes the Wronskians f (x) g
metric.
′
′
(x) − f
(x) g (x). So, A is sym-
It remains to show that if hAF, W i = hF, U i for all F = (f, N ′ (f )) ∈
D (A), then W ∈ D (A) and AW = U , where W = (w (x) , h) and U =
(u (x) , k), i.e., (i) w1, w
3 ∈
ACloc ((h2, h3)) w4, w
(w) =
(1) ; (iii) τ1w = τ3w = τ4w = τ5w = τ6w = τ7w = τ8w = 0; (iv)
1w (1) − β
β
u (x) = τ w; (v) k = −N (w) = −β1w (1) + β2w
4 ∈ ACloc ((h3, 1)) and τ w ∈ H1; (ii) h = N
1 ∈ ACloc ((−1, h1)), w2, w
2 ∈ ACloc ((h1, h2)), w3, w
(1).
2w
′
′
′
′
′
′
′
′
′
4
0 ⊕ 0 ⊂ D (A), we obtain
(τ f ) wdx +
1
p2
3θγ
h3Z
h2
(τ f ) wdx +
1
p2
4θγξ
1Z
h3
(τ f ) wdx
For all F ∈ eC∞
h1Z
(τ f ) wdx +
1
p2
1
−1
1
p2
2θ
h1
h2Z
h2Z
h1
=
1
p2
1
h1Z
−1
f udx +
1
p2
2θ
f udx +
1
p2
3θγ
h3Z
f udx +
h2
1
p2
4θγξ
1Z
h3
f udx,
namely, hτ f, wi1 = hf, ui1. Hence, by standart Sturm-Liouville theory, (i) and
(iv) hold. By (iv), the equation hAF, W i = hF, U i, ∀F ∈ D (A), becomes
1
p2
1
h1Z
−1
(τ f ) wdx +
1
p2
2θ
h2Z
h1
(τ f ) wdx +
1
p2
3θγ
h3Z
h2
(τ f ) wdx +
1
p2
4θγξ
1Z
h3
(τ f ) wdx −
N (f ) h
ρθγ
1
p2
1
h1Z
−1
f (τ w) dx +
=
So,
h2Z
f (τ w) dx +
1
p2
2θ
h1
1
p2
3θγ
h3Z
f (τ w) dx +
h2
1
p2
4θγξ
1Z
h3
f (τ w) dx +
′
N
(f ) k
ρθγξ
.
hτ f, wi1 = hf, τ wi1 +
′
N
(f ) k + N (f ) h
ρθγξ
.
However,
hτ f, wi1 =
1
p2
1
−1
h1Z
(cid:16)−p2
h3Z
(cid:16)−p2
h1Z
h2
−1
1
p2
3θγ
=
1
p2
1
′′
1f
′′
3f
+ q (x) f(cid:17) wdx +
+ q (x) f(cid:17) wdx +
1
p2
2θ
1
p2
4θγξ
f (τ w) dx +
h2Z
f (τ w) dx +
1
p2
2θ
h1
′′
h1
2f
h2Z
(cid:16)−p2
1Z
(cid:16)−p2
h3Z
1
p2
3θγ
h3
h2
4f
′′
+ q (x) f(cid:17) wdx+
+ q (x) f(cid:17) wdx
1Z
1
p2
4θγξ
h3
f (τ w) dx +
f (τ w) dx
+W (f, w; h1 − 0) − W (f, w; −1) +
1
θ
W (f, w; h2 − 0) −
1
θ
W (f, w; h1 + 0) +
1
θγ
W (f, w; h3 − 0) −
1
θγ
W (f, w; h2 + 0) +
1
θγξ
W (f, w; 1) −
1
θγξ
W (f, w; h3 + 0)
= hf, τ wi1 + W (f, w; h1 − 0) − W (f, w; −1) +
W (f, w; h2 − 0) −
1
θ
1
θγξ
W (f, w; 1) −
1
θγξ
1
θ
W (f, w; h1 + 0)
W (f, w; h3 + 0) .
+
1
θγ
W (f, w; h3 − 0) −
1
θγ
W (f, w; h2 + 0) +
5
Hence,
′
N
(f ) k + N (f ) h
ρθγξ
= W (f, w; h1 − 0) − W (f, w; −1) +
1
θ
W (f, w; h2 − 0)
1
θ
W (f, w; h1 + 0) +
1
θγ
W (f, w; h3 − 0) −
1
θγ
W (f, w; h2 + 0)
−
+
−
+
−
−f
−f
W (f, w; 1) −
W (f, w; h2 + 0)
1
θγξ
=(cid:16)f (h1 − 0) w
1
θγξ
′
′
1
1
(−1) w (−1)(cid:17) +
θ (cid:16)f (h1 + 0) w
θγ (cid:16)f (h3 − 0) w
θγ (cid:16)f (h2 + 0) w
(1) w (1)(cid:17) −
1
′
′
′
′
(h1 − 0) − f
′
1
θ (cid:16)f (h2 − 0) w
(h1 + 0) − f
′
(h3 − 0) − f
(h2 + 0) − f
1
θγξ (cid:16)f (h3 + 0) w
(h1 − 0) w (h1 − 0)(cid:17) −(cid:16)f (−1) w
′
(−1)
(h2 − 0) w (h2 − 0)(cid:17)
′
(h2 − 0) − f
′
′
(h1 + 0) w (h1 + 0)(cid:17)
(h3 − 0) w (h3 − 0)(cid:17)
(h2 + 0) w (h2 + 0)(cid:17) +
′
′
(h3 + 0) − f
′
′
1
θγξ (cid:16)f (1) w
(h3 + 0) w (h3 + 0)(cid:17) .
(1)
(8)
′
′
′
By Naimark's Patching Lemma [20], there is an F ∈ D (A) such that f (−1) =
(−1) = f (h1 − 0) = f
(h1 + 0) = f (h2 − 0) =
f
(h3 − 0) = f (h3 + 0) =
(h2 − 0) = f (h2 + 0) = f
f
(h3 + 0) = 0, f (1) = β
(f ) = 0. Thus,
f
(1). Namely, (ii) holds. Similarly, one
from (8) we obtain h = β
proves (v).
(h1 − 0) = f (h1 + 0) = f
(h2 + 0) = f (h3 − 0) = f
2 and f
1w (1) − β
1. For such an F , N
(1) = β
2w
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
(h3 − 0) = α7, f
(−1) = 0. Let F ∈ D (A) satisfies f (1) = f
(h1 − 0) = f (h2 − 0) = f
(−1) = −α1. N
It remains to show that (iii) holds. Choose F ∈ D (A) so that f (1) =
(1) = f (h1 − 0) = f
(h2 − 0) = f (h3 − 0) =
(f ) = N (f ) = 0. From
(h3 − 0) = 0, f (−1) = α2 and f
(1) =
(−1) = f (h1 + 0) = f (h2 + 0) = f (h3 + 0) = 0, f (h1 − 0) =
(h2 − 0) =
(h3 + 0) = ξ.
(f ) = 0. By (8), we have w (h1 + 0) = α3w (h1 − 0) +
(h2 − 0) and w (h3 + 0) =
(1) =
(h3 + 0) = 0, f (h1 − 0) =
(h2 − 0) = −α6, f (h3 − 0) = β8,
(h3 − 0) = −α8 f (h1 + 0) = θ, f (h2 + 0) = γ and f (h3 + 0) = ξ. Then
f
f
(8) , we get α1w (−1) + α2w
f (−1) = f
−β3, f (h2 − 0) = −β5, f (h3 − 0) = −β7, f
α5, f
Then N (f ) = N
β3w
α7w (h3 − 0)+ β7w
f (−1) = f
β4, f
f
N (f ) = N
w
β8w
(h3 − 0) . Finally, choose F ∈ D (A) so that f (1) = f
(h1 − 0) , w (h2 + 0) = α5w (h2 − 0) + β5w
(h1 − 0) = −α4, f (h2 − 0) = β6, f
(h2 + 0) = α6w (h2 − 0) + β6w
(h1 + 0) = α4w (h1 − 0)+β4w
(f ) = 0. From (8) , we obtain w
(h3 + 0) = α8w (h3 − 0) +
(h2 + 0) = γ and f
(h1 − 0) = α3, f
(h1 + 0) = θ, f
(h2 − 0) and w
(h1 + 0) = f
(h2 + 0) = f
(−1) = f
(h3 − 0) .
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
′
(h1 − 0),
Corollary 1.1 The eigenvalues of (1) − (9) are real, and if λ1 and λ2 are
two different eigenvalues of (1) − (9), then the corresponding eigenfunctions
6
f (x) and g (x) are orthogonal in the sense of
f g +
f g +
1
p2
2θ
h2Z
h1
−1
1
p2
1
h1Z
ρθγξ (cid:16)β
1
′
1f (1) − β
′
′
2f
1g (1) − β
′
′
2g
1
p2
3θγ
h3Z
(1)(cid:17)(cid:16)β
h2
′
f g +
1
p2
4θγξ
h3
f g+
1Z
(1)(cid:17) = 0.
3 Simplicity of eigenvalues
We consider the initial-value problem
′′
(cid:26) −p2
1u
(x) + q(x)u (x) = λu (x) ,
x ∈ [−1, h1) ,
u (−1) = α2, u
′
(−1) = −α1.
In terms of existence and uniqueness in ordinary differential equation theory, the
initial-value problem has a unique solution ϕ1 (x, λ) for every λ ∈ C. Similarly,
the initial-value problem
has a unique solution ϕ2 (x, λ) for every λ ∈ C.The initial-value problem
′′
−p2
2u
(x) + q(x)u (x) = λu (x) ,
u (h1) = α3ϕ1 (h1, λ) + β3ϕ
(h1) = α4ϕ1 (h1, λ) + β4ϕ
u
′
′
1 (h1, λ) ,
1 (h1, λ)
′
x ∈ (h1, h2) ,
′′
−p2
3u
(x) + q(x)u (x) = λu (x) ,
u (h2) = α5ϕ2 (h2, λ) + β5ϕ
(h2) = α6ϕ2 (h2, λ) + β6ϕ
u
′
′
2 (h2, λ) ,
2 (h2, λ)
′
x ∈ (h2, h3) ,
has a unique solution ϕ3 (x, λ) for every λ ∈ C. Similarly, the initial-value
problem
′′
−p2
4u
(x) + q(x)u (x) = λu (x) ,
u (h3) = α7ϕ2 (h2, λ) + β5ϕ
(h3) = α8ϕ2 (h2, λ) + β6ϕ
u
′
′
2 (h2, λ) ,
2 (h2, λ)
′
x ∈ (h3, 1] ,
has a unique solution ϕ4 (x, λ) for every λ ∈ C. For each given x ∈ [−1, h1) ,
ϕ1 (x, λ) is an entire function of λ; for every x ∈ (h1, h2) , ϕ2 (x, λ) is an entire
function of λ; for every x ∈ (h2, h3) , ϕ3 (x, λ) is an entire function of λ and for
every x ∈ (h3, 1] , ϕ4 (x, λ) is an entire function of λ.
Now we define a function φ (x, λ) on x ∈ [−1, h1) ∪ (h1, h2) ∪ (h2, h3) ∪ (h3, 1]
by
φ (x, λ) =
ϕ1 (x, λ) , x ∈ [−1, h1) ,
ϕ2 (x, λ) , x ∈ (h1, h2) ,
ϕ3 (x, λ) , x ∈ (h2, h3) ,
ϕ4 (x, λ) , x ∈ (h3, 1] .
7
Obviously φ (x, λ) satisfies (1) , (2) and (4)−(9) . Similarly, we define the function
χ (x, λ) =
which satisfies (1) , (3) − (9) .
χ1 (x, λ) , x ∈ [−1, h1) ,
χ2 (x, λ) , x ∈ (h1, h2) ,
χ3 (x, λ) , x ∈ (h2, h3) ,
χ4 (x, λ) , x ∈ (h3, 1] ,
The Wronskian W (ϕi (x, λ) , χi (x, λ)) (i = 1, 2, 3, 4) are independent of the
variable x. Let wi (λ) = W (ϕi (x, λ) , χi (x, λ)) and w (λ) = w1 (λ) , and then
we obtain w2 (λ) = θw (λ) , w3 (λ) = θγw (λ) and w4 (λ) = θγξw (λ) .
Lemma 2.1 [10] The eigenvalues of the problem (1) − (9) coincide with the
zeros of the entire function w (λ).
Definition 2.1 The analytic multiplicity of an eigenvalue λ of (1) − (9)
is its order as a root of the characteristic equation w (λ) = 0. The geometric
multiplicity of an eigenvalue is the dimension of its eigenspace, i.e., the number
of its linearly independent eigenfunctions.
′
For convenience, set φ = φ (x, λ) , χ1λ = ∂χ1
Theorem 2.1 The eigenvalues of (1) − (9) are analytically simple.
∂λ , etc.
1λ = ∂χ
∂λ , χ
1
′
Proof. Let λ = s + it, where s, t ∈ R and i2 = −1. We differentiate the
equation τ χ = λχ with respect to λ and have
By integration by parts, we get
τ χλ = λχλ + χ.
hτ χλ, φi1 − hχλ, τ φi1 =(cid:16)χ1λϕ1
θγ (cid:16)χ3λϕ3
3λϕ3(cid:17) h3
h2 +
− χ
1
′
′
1
θγξ (cid:16)χ4λϕ4
′
− χ
′
1λϕ1(cid:17) h1
−1 +
1
θ (cid:16)χ2λϕ2
′
− χ
′
4λϕ4(cid:17) 1
h3 .
′
− χ
′
2λϕ2(cid:17) h2
h1 +
(9)
Substituting τ χλ = λχλ + χ and τ φ = λφ into the left side of (9) , we have
λ hχλ, φi1 + hχ, φi1 − hχλ, λφi1 = hχ, φi1 + 2it hχλ, φi1 .
Moreover,
+
′
(cid:16)χ1λϕ1
θγ (cid:16)χ3λϕ3
1
′
− χ
− χ
−1 +
′
1λϕ1(cid:17) h1
3λϕ3(cid:17) h3
′
h2 +
− χ
′
1
θ (cid:16)χ2λϕ2
θγξ (cid:16)χ4λϕ4
1
′
′
h1
2λϕ2(cid:17) h2
4λϕ4(cid:17) 1
′
− χ
h3=
8
χ1λ (h1, λ) ϕ1
′
′
χ
1λ (−1, λ) ϕ1 (−1, λ) +
′
(h1, λ) − χ
1λ (h1, λ) ϕ1 (h1, λ) − χ1λ (−1, λ) ϕ1
1
′
′
(−1, λ) +
(h2, λ) − χ
2λ (h2, λ) ϕ2 (h2, λ)
′
θ (cid:16)χ2λ (h2, λ) ϕ2
2λ (h1, λ) ϕ2 (h1, λ)(cid:17) +
′
−χ2λ (h1, λ) ϕ2
′
(h1, λ) + χ
−χ
′
3λ (h3, λ) ϕ3 (h3, λ) − χ3λ (h2, λ) ϕ3
′
(h2, λ) + χ
′
1
θγ (cid:16)χ3λ (h3, λ) ϕ3
3λ (h2, λ) ϕ3 (h2, λ)(cid:17) +
′
(h3, λ)
(1, λ) − χ
′
4λ (1, λ) ϕ4 (1, λ) − χ4λ (h3, λ) ϕ4
′
(h3, λ)
1
θγξ (cid:16)χ4λ (1, λ) ϕ4
′
+ χ
′
4λ (h3, λ) ϕ4 (h3, λ)
′
= α1χ1λ (−1, λ) + α2χ
′
′
1
χ
+
θ (cid:16)χ2λ (h2, λ) ϕ2
2λ (h1, λ) ϕ2 (h1, λ)(cid:17) +
θγ (cid:16)χ3λ (h2, λ) ϕ3
θγξ (cid:16)β
(1, λ) − β
2ϕ4
−
1
1
′
′
′
1λ (−1, λ) + χ1λ (h1, λ) ϕ1
′
(h1, λ) − χ
′
1λ (h1, λ) ϕ1 (h1, λ)
(h2, λ) − χ
′
2λ (h2, λ) ϕ2 (h2, λ) − χ2λ (h1, λ) ϕ2
′
(h1, λ) +
1
θγ (cid:16)χ3λ (h3, λ) ϕ3
′
(h3, λ) − χ3λ (h3, λ) ϕ3 (h3, λ)(cid:17)
(h2, λ) − χ
′
1ϕ4 (1, λ)(cid:17) −
′
′
1
3λ (h2, λ) ϕ3 (h2, λ)(cid:17) +
θγξ (cid:16)χ4λ (h3, λ) ϕ4
θγξ (cid:16)β
2ϕ4
1
′
′
′
(1, λ) − β
′
4λ (h3, λ) ϕ4 (h3, λ)(cid:17)
(h3, λ) − χ
′
1ϕ4 (1, λ)(cid:17) .
= α1χ1λ (−1, λ) + α2χ
1λ (−1, λ) +
Note that
′
w
(λ) = α2χ
′
1λ (−1, λ) + α1χ1λ (−1, λ)
Therefore, (9) becomes
′
w
(λ) = hχ, φi1 + 2it hχλ, φi1 −
1
θγξ (cid:16)β
′
′
2ϕ4
(1, λ) − β
′
1ϕ4 (1, λ)(cid:17) .
(10)
Now we consider the simplicity of the eigenvalues of (1)−(9). Let µ be arbitrary
zero of w (λ). By Corollary 1.1, µ is real. SInce
w (µ) =(cid:12)(cid:12)(cid:12)(cid:12)
ϕ1 (x, µ) χ1 (x, µ)
ϕ
1 (x, µ) χ
′
′
1 (x, µ) (cid:12)(cid:12)(cid:12)(cid:12) = 0,
we have ϕ1 (x, µ) = c1χ1 (x, µ) (c1 6= 0) , ϕ2 (x, µ) = c2χ2 (x, µ) (c2 6= 0) , ϕ3 (x, µ) =
c3χ3 (x, µ) (c3 6= 0) and ϕ4 (x, µ) = c4χ4 (x, µ) (c4 6= 0) where c1, c2, c3, c4 ∈ C.
9
From
′
ϕ
ϕ2 (h1, µ) = c1(cid:16)α3χ1 (h1, µ) + β3χ
2 (h1, µ) = c1(cid:16)α4χ1 (h1, µ) + β4χ
ϕ3 (h2, µ) = c2(cid:16)α5χ2 (h2, µ) + β5χ
3 (h2, µ) = c2(cid:16)α6χ2 (h2, µ) + β6χ
ϕ4 (h3, µ) = c3(cid:16)α7χ3 (h3, µ) + β7χ
4 (h3, µ) = c3(cid:16)α8χ3 (h3, µ) + β8χ
ϕ
ϕ
′
′
′
′
′
′
′
2 (h1, µ) ,
1 (h1, µ)(cid:17) = c1χ2 (h1, µ) ,
1 (h1, µ)(cid:17) = c1χ
2 (h2, µ)(cid:17) = c2χ3 (h2, µ) ,
2 (h2, µ)(cid:17) = c2χ
3 (h3, µ)(cid:17) = c3χ4 (h3, µ) ,
3 (h3, µ)(cid:17) = c3χ
3 (h2, µ) ,
4 (h3, µ) ,
′
′
′
′
we get c1 = c2 = c3 = c4 6= 0. Thus, simple calculations using (10) and the
initial values of χ4 at x = 1 give
′
w
p2
1
(µ) = c1
h1Z
1
1Z
1
p2
4θγξ
+
−1
h3
1
p2
2θ
h1
h2Z
.
χ1 (x, µ)2 dx +
χ4 (x, µ)2 dx +
ρ
θγξ
χ2 (x, µ)2 dx +
1
p2
3θγ
h3Z
h2
χ3 (x, µ)2 dx
Note that ρ > 0, θ > 0, γ > 0, ξ > 0 and c1 6= 0, so w
analytic multiplicity of µ is one. By Lemma 2.1, the proof is completed.
Theorem 2.2 All eigenvalues of (1) − (9) are geometrically simple.
(µ) 6= 0. Hence, the
′
Proof. If f and g are two eigenfunctions for an eigenvalue λ∗ of (1) − (9), then
(2) implies that f (−1) = Kg (−1) and f
(−1) for some constant
K ∈ R. By the uniqueness theorem for solutions of ordinary differential equation
and the transmission conditions (4)−(9) , we have that f = Kg on [−1, 1] . Thus
the geometric multiplicity of λ∗ is one.
(−1) = Kg
′
′
4 Completeness of eigenfunctions
Theorem 3.1 The operator A has only point spectrum, i.e., σ (A) = σρ (A) .
Proof. It suffices to prove that if η is not an eigenvalue of A, then η ∈ ρ (A) .
Since A is self-adjoint, we only consider a real η. We investigate the equation
(A − η) Y = F ∈ H, where F = (f, h) .
10
Let us consider the initial-value problem
τ y − ηy = f, x ∈ I,
α1y (−1) + α2y
′
(−1) = 0,
′
′
′
y (h1 + 0) = α3y (h1 − 0) + β3y
(h1 + 0) = α4y (h1 − 0) + β4y
y
y (h2 + 0) = α5y (h2 − 0) + β5y
(h2 + 0) = α6y (h2 − 0) + β6y
y
y (h3 + 0) = α7y (h3 − 0) + β7y
(h3 + 0) = α8y (h3 − 0) + β8y
y
′
′
′
′
′
′
(h1 − 0) ,
(h1 − 0) ,
(h2 − 0) ,
(h2 − 0) ,
(h3 − 0) ,
(h3 − 0) ,
(11)
Let u (x) be the solution of the equation τ u − ηu = 0 satisfying
u (−1) = α2, u
′
(−1) = −α1,
u (h1 + 0) = α3u (h1 − 0) + β3u
′
u
(h1 + 0) = α4u (h1 − 0) + β4u
u (h2 + 0) = α5u (h2 − 0) + β5u
′
u
(h2 + 0) = α6u (h2 − 0) + β6u
u (h3 + 0) = α7u (h3 − 0) + β7u
′
u
(h3 + 0) = α8u (h3 − 0) + β8u
′
′
′
′
′
′
(h1 − 0) ,
(h1 − 0) ,
(h2 − 0) ,
(h2 − 0) ,
(h3 − 0) ,
(h3 − 0) .
In fact,
u (x) =
u1 (x) , x ∈ [−1, h1) ,
u2 (x) , x ∈ (h1, h2) ,
u3 (x) , x ∈ (h2, h3) ,
u4 (x) , x ∈ (h3, 1] ,
where u1 (x) is the unique solution of the initial-value problem
(cid:26) −p2
′′
+ q(x)u = ηu,
1u
u (−1) = α2, u
x ∈ [−1, h1) ,
′
(−1) = −α1;
u2 (x) is the unique solution of the problem
u3 (x) is the unique solution of the problem
′′
−p2
+ q(x)u = ηu,
2u
u2 (h1) = α3u1 (h1) + β3u
2 (h1) = α4u1 (h1) + β4u
u
′
′
1 (h1) ,
1 (h1) ;
′
x ∈ (h1, h2) ,
′′
−p2
+ q(x)u = ηu,
3u
u3 (h2) = α5u2 (h2) + β5u
3 (h2) = α6u2 (h2) + β6u
u
′
′
2 (h2) ,
2 (h2) .
′
x ∈ (h2, h3) ,
11
and u4 (x) is the unique solution of the problem
Let
′′
−p2
4u
+ q(x)u = ηu,
x ∈ (h3, 1] ,
u3 (h3) = α7u3 (h3) + β7u
3 (h3) = α8u3 (h3) + β8u
u
′
′
3 (h3) ,
3 (h3) .
′
w (x) =
w1 (x) , x ∈ [−1, h1) ,
w2 (x) , x ∈ (h1, h2) ,
w3 (x) , x ∈ (h2, h3) ,
w4 (x) , x ∈ (h3, 1] ,
be a solution of τ w − ηw = f satisfying
α1w (−1) + α2w
′
(−1) = 0,
′
w (h1 + 0) = α3w (h1 − 0) + β3w
(h1 + 0) = α4w (h1 − 0) + β4w
w
w (h2 + 0) = α5w (h2 − 0) + β5w
(h2 + 0) = α6w (h2 − 0) + β6w
w
w (h3 + 0) = α7w (h3 − 0) + β7w
(h3 + 0) = α8w (h3 − 0) + β8w
w
′
′
′
′
′
′
′
′
(h1 − 0) ,
(h1 − 0) ,
(h2 − 0) ,
(h2 − 0) ,
(h3 − 0) ,
(h3 − 0) .
Then, (11) has the general solution
y (x) =
du1 + w1, x ∈ [−1, h1) ,
du2 + w2, x ∈ (h1, h2) ,
du3 + w3, x ∈ (h2, h3) ,
du4 + w4, x ∈ (h3, 1] ,
where d ∈ C.
Since η is not an eigenvalue of the problem (1) − (7) , we have
ηhβ
′
1u2 (1) − β
′
2u
′
2(1)i +hβ1u2 (1) − β2u
′
2(1)i 6= 0.
The second component of (A − η) Y = F involves the equation
−N (y) − ηN
′
(y) = h,
namely,
Substituting (12) into (14), we get
h−β1y (1) + β2y
(cid:16)β2u
′
′
(1)i − ηhβ
′
1y (1) − β
′
′
2y
(1)i = h.
1u2 (1)(cid:17) d
′
′
′
2 (1) − β1u2 (1) + ηβ
′
2u
′
2 (1) − ηβ
= h + β1w2 (1) − β2w
2 (1) + ηβ
1w2 (1) − ηβ
2w
2 (1)
′
′
(12)
(13)
(14)
In view of (13), we know that d is uniquely solvable. Therefore, y is uniquely
determined.
12
The above arguments show that (A − ηI)−1 is defined on all of H, where I
is identity matrix. We obtain that (A − ηI)−1 is bounded by Theorem 1.1 and
the Closed Graph Theorem. Thus, η ∈ ρ (A) . Therefore, σ (A) = σρ (A) .
Lemma 3.1 [10] The eigenvalues of the boundary value problem (1) − (9)
are bounded below, and they are countably infinite and can cluster only at ∞.
For every δ ∈ R \ σρ (A) , we have the following immediate conclusion.
Lemma 3.2 Let λ be an eigenvalue of A−δI, and V a corresponding eigen-
function. Then, λ−1 is an eigenvalue of (A − δI)−1 , and V is a corresponding
eigenfunction. The converse is also true.
On the other hand, if µ is an eigenvalue of A and U is a corresponding
eigenfunction, then µ − δ is an eigenvalue of A − δI, and U is a corresponding
eigenfunction. The converse is also true. Accordingly, the discussion about
the completeness of the eigenfunctions of A is equivalent to considering the
corresponding property of (A − δI)−1.
By Lemma 1.1, Lemma 3.1 and Corollary 1.1, we suppose that {λn; n ∈ N}
is the real sequence of eigenvalues of A, then {λn − δ; n ∈ N} is the sequence
of eigenvalues of A − δI. We may assume that
λ1 − δ ≤ λ2 − δ ≤ ... ≤ λn − δ ≤ ... → ∞.
Let {µn; n ∈ N} be the sequence of eigenvalues of (A − δI)−1 . Then µn =
(λn − δ)−1 and
µ1 ≥ µ2 ≥ ... ≥ µn ≥ ... → 0.
Note that 0 is not an eigenvalue of (A − δI)−1 .
Theorem 3.2 The operator A has compact resolvents, i.e, for each δ ∈
R \ σρ (A) , (A − δI)−1 is compact on H.
Proof. Let {µ1, µ2, ...} be the eigenvalues of (A − δI)−1 , and {P1, P2, ...} the
orthogonal projections of finite rank onto the corresponding eigenspaces. Since
{µ1, µ2, ...} is a bounded sequence and all Pn′s are mutually orthogonal, we
n=1 µnPn is strongly convergent to the bounded operator (A − δI)−1 ,
n=1 µnPn. Because for every α > 0, the number of µn′s
satisfying µn > α is finite, and all Pn′s are of finite rank, we obtain that
(A − δI)−1 is compact.
have P∞
i.e., (A − δI)−1 = P∞
In terms of the above statements and the spectral theorem for compact
operators, we obtain the following theorem.
Theorem 3.3 The eigenfunctions of the problem (1) − (9) , augmented to
become eigenfunctions of A, are complete in H, i.e., if we let
nΦn =(cid:16)φn (x) , N
′
(φn)(cid:17) ; n ∈ No
be a maximum set of orthonormal eigenfunctions of A, where {φn (x) ; n ∈ N}
are eigenfunctions of (1) − (9) , then for all F ∈ H, F =P∞
n=1 hF, Φni Φn.
13
References
[1] C. T. Fulton, Two-point boundary value problems with eigenvalue parameter
contained in the boundary conditions. Proc. Royal Soc. Edinburgh 77A (1977),
293-308.
[2] A. V. Likov, Y. A. Mikhalilov, The Theory of Heat and Mass Transfer, Qosener-
gaizdat, 1963. (In Russian).
[3] E. S¸en, A. Bayramov, Calculation of eigenvalues and eigenfunctions of a discontin-
uous boundary value problem with retarded argument which contains a spectral
parameter in the boundary condition, Mathematical and Computer Modelling,
54 (2011) 3090-3097.
[4] E. S¸en, A. Bayramov, On calculation of eigenvalues and eigenfunctions of a
Sturm-Liouville type problem with retarded argument which contains a spectral
parameter in the boundary condition, Journal of Inequalities and Applications,
Vol. 2011 (1) (2011) 1-9.
[5] P. A. Binding, P. J. Browne, B. A. Watson, Sturm-Liouville problems with bound-
ary conditions rationally dependent on the eigenparameter, II, J. Comput. Appl.
Math. 148 (2002), 147-168.
[6] P. A. Binding, Patrick. J. Browne, Oscillation theory for indefinite Sturm-
Liouville problems with eigenparameter dependent boundary conditions. Proc.
Royal Soc. Edinburg 127A (1997), 1123-1136.
[7] P. A. Binding, R. Hryniv, H. Langer, Elliptic eigenvalue problems with eigen-
parameter dependent boundary conditions. J. Differential Equations 174 (2001),
30-54.
[8] M. Demirci, Z. Akdogan, O. Sh. Mukhtarov, Asymptotic behavior of eigenvalues
and eigenfunctions of one discontinuous boundary-value problem. International
J. Computational Cognition 2(3) (2004), 101-113.
[9] D. B. Hinton, An expansion theorem for an eigenvalue problem with eigenvalue
parameter in the boundary condition. Quart. J. Math. (Oxford) 30 (1979), 33-42.
[10] M. Kadakal, O.Sh. Mukhtarov, Sturm-Liouville problems with discontinuities at
two points, Comput. Math. Appl.,54 (2007) 1367-1379.
[11] O. Sh. Mukhtarov, S. Yakubov, Problems for differential equations with trans-
mission conditions, Applicable Anal. 81 (2002), 1033-1064.
[12] I. Titeux, Y. Yakubov, Completeness of root functions for thermal conduction in
a strip with piecewise continuous coefficients, Math. Models Methods Appl. Sc.
7 (1997), 1035-1050.
[13] O. Sh. Mukhtarov, M. Kadakal, Spectral properties of one Sturm-Liouville type
problem with discontinuous weight, (Russian) Sibirsk. Math. Zh., 46 (4) (2005),
681-694.
14
[14] Y. S. Li, J. Sun, Z. Wang, On the complete description of self-adjoint boundary
conditions of the Schrodinger operator with a δ (x)or δ
(x)interaction, Sympo-
sium of the Fifth Conference of Mathematics Society of Inner Mongolia, Inner
Mongolia Univ. Press, 1995, 27-30.
′
[15] A. Zettl, Adjoint and self-adjoint boundary value problems with interface condi-
tions, SIAM J. Appl. Math. 16 (1968), 851-859.
[16] F. Gesztesy, W. Kirsch, One-dimensional Schrodinger operators with interactions
on a discerete set, J. Reine Angew. Math. 362 (1985), 28-50.
[17] A. Wang, J. Sun, P. Gao, Completeness of Eigenfunctions of Sturm-Liouville
Problems with Transmission Conditions, J. Spectral Math. Appl. (2006)
[18] A. Wang, J. Sun, X. Hao, S. Yao, Completeness of Eigenfunctions of Sturm-
Liouville Problems with Transmission Conditions, Methods and Application of
Analysis 16 (3) (2009), 299-312.
[19] J. Weidmann, Spectral Theory of Ordinary Differential Operators, Lecture Notes
in Math. 1258, Springer-Verlag, Berline, 1987.
[20] M. A. Naimark, Linear Differential Operators, part II. Harrap, London, 1968.
15
|
1601.00612 | 1 | 1601 | 2015-12-30T16:48:29 | A note on the superadditive and the subadditive transformations of aggregation functions | [
"math.FA"
] | We expand the theoretical background of the recently introduced superadditive and subadditive transformations of aggregation functions $A$. Necessary and sufficient conditions ensuring that a transformation of a proper aggregation function is again proper are deeply studied and exemplified. Relationships between these transformations are also studied. | math.FA | math |
A note on the superadditive and the subadditive
transformations of aggregation functions
Alexandra Siposov´a
Slovak University of Technology
Faculty of Civil Engineering
Department of Mathematics and Descriptive Geometry
Radlinsk´eho 11, 810 05 Bratislava
Slovakia
[email protected]
Abstract
We expand the theoretical background of the recently introduced superadditive
and subadditive transformations of aggregation functions A. Necessary and sufficient
conditions ensuring that a transformation of a proper aggregation function is again
proper are deeply studied and exemplified. Relationships between these transforma-
tions are also studied.
Keywords: aggregation function, subadditive transformation, superadditive transfor-
mation
1
Introduction
Motivated by applications in economics, subadditive and superadditive transformations of
aggregation functions on R+ = [0,∞[ have been recently introduced in [4]. Formally, both
these transformations can be introduced on the improper real interval [0,∞].
Definition 1 A mapping A : [0,∞]n → [0,∞] is called an (n-ary) aggregation function
if A(0, . . . , 0) = 0 and A is increasing in each coordinate. Further, A is called a proper
(n-ary) aggregation function if it satisfies the following two additional constraints:
(i) A(x) ∈ ]0,∞[ for some x ∈ ]0,∞[n,
(ii) A(x) < ∞ for all x ∈ [0,∞[n.
Though for real applications we only need proper aggregation functions (in fact, their
restriction to the domain [0,∞[n), a broader framework of all (n-ary) aggregation function
is of advantage in a formal description of our results, making formulations and expressions
more transparent. Observe that our framework is broader than the concept of aggregation
functions on [0,∞] as introduced in [1], [3], which does not cover Sugeno integral based
1
A∗(x) = inf { k(cid:88)
A∗(x) = sup { (cid:96)(cid:88)
k(cid:88)
(cid:96)(cid:88)
aggregation functions, for example. We denote the class of all n-ary aggregation function
by An, and the class of all n-ary proper aggregation function by Pn.
The next definition was motivated by optimization tasks treated in linear programming
area and related areas [2], as well as by recently introduced concepts of concave [5] and
convex [6] integrals.
Definition 2 For every A ∈ An the subadditive transformation A∗ : [0,∞]n → [0,∞] of
A is given by
(1)
Similarly, for every A ∈ An the superadditive transformation A∗ : [0,∞]n → [0,∞] of A is
defined by
i=1
i=1
A(y(i))
y(i) ≥ x}
A(y(j))
y(j) ≤ x} .
(2)
j=1
j=1
n, where K∗
Observe that the transformation (1) was originally introduced in [4] for A ∈ Kn∗ , where
Kn∗ is the class of all n-ary proper aggregation functions (restricted to [0,∞[n) such that
also A∗ is proper, that is, A∗ ∈ Pn. Similarly, A∗ given by (2) was originally introduced
in [4] only for A ∈ K∗
n is the class of all A ∈ Pn (restricted to [0,∞[n), so that
A∗ ∈ Pn as well.
Theorem 2 in [4] gives a necessary and sufficient condition ensuring that a function
A ∈ Pn has also the property that A ∈ K∗
n. We develop this result, giving an equivalent
condition. Moreover, we also characterize all the functions A ∈ Pn such that A ∈ Kn∗ . Our
approach is based on a deep study of transformations (1) and (2) on unary aggregation
functions that belong to P1. Our approach allows to show that for any A ∈ Pn we have
the inequality (A∗)∗ ≤ (A∗)∗.
1 and K1∗ are
completely described, showing that the properties in a neighbourhood of 0 are important
for characterization of elements of these classes.
In Section 3, necessary and sufficient
conditions for a function A ∈ Pn to belong to K∗
n, or to Kn∗ , are given. Section 4 is devoted
to the study of relationships of transformations (A∗)∗ and (A∗)∗. Finally, some concluding
remarks are added.
In the next section, the classes K∗
The paper is organized as follows.
2 The one-dimensional case
We begin with basic results which show how the values of the subadditive and superadditive
transformations of one-dimensional aggregation functions depend on the behavior of the
functions near zero.
2
Theorem 1 Let h be an unary aggregation function on [0,∞] with lim inf t→0+ h(t)/t = a
and lim supt→0+ h(t)/t = b, where 0 ≤ a ≤ b ≤ ∞. Then, for every x ∈ ]0,∞[ we have
h∗(x) ≤ ax and h∗(x) ≥ bx.
Proof. Let x > 0. By definitions of h∗ and h∗, for every positive integer n we have
h∗(x) ≤ nh(x/n) ≤ h∗(x), that is,
h∗(x) ≤ x · h( x
n)
x
n
≤ h∗(x) .
(3)
Since h is increasing, for every t such that
n we have
x
n+1 ≤ t ≤ x
≤ h( x
n)
x
≤ h(t)
t
.
n+1
h( x
n+1)
x
n
Applying the limits inferior and superior to these inequalities as t → 0+ and n → ∞ (with
n → 1) shows that
n+1
lim inf
n→∞
h( x
n)
x
n
≤ lim inf
t→0+
h(t)
t
and
lim sup
n→∞
h( x
n)
x
n
≥ lim sup
t→0+
h(t)
t
.
Combining (3) with (4) now gives
h∗(x) ≤ x · lim inf
t→0+
h(t)
t
= ax and h∗(x) ≥ x · lim sup
t→0+
h(t)
t
= bx
for every x > 0, which completes the proof.
(4)
2
h(t)
t and lim supt→0+
h(t)
t
The values of lim inf t→0+
can be interpreted as the ‘lower’ and
‘upper’ slope of h at the point x = 0. The previous result may therefore be interpreted by
saying that the values of h∗ and h∗ are to a large extent influenced by the values of the
lower and upper slopes of h at 0.
Corollary 1 Suppose that h is an unary aggregation function on [0,∞] such that the
(cid:48)
derivative h
(0+) exists and is equal to c ∈ [0,∞[.
(1) If h is convex on [0,∞[, then h∗(x) = cx and h∗(x) = h(x) for every x ≥ 0.
(2) If h is concave on [0,∞[, then h∗(x) = h(x) and h∗(x) = cx for every x > 0.
Proof. For (1) it is sufficient to realize that h(x) ≥ ax for every x ≥ 0, the claim then
follows from Theorem 1 regarding h∗ and from [4] regarding h∗. The proof of (2) is similar
2
and therefore omitted.
3
Figure 1: A schematic drawing of a function h from the proof of Corollary 2.
Corollary 2 For any real a, b such that 0 < a < b < ∞ there is an infinite number of
smooth unary aggregation functions h on [0,∞] such that h∗(x) = ax and h∗(x) = bx for
each x ∈ [0,∞[.
Proof. Let q be a positive real number such that q < a/b < 1; note that bq2j < aq2j−1
for every positive integer j. Results from calculus now imply the existence of infinitely
many smooth increasing functions h(x) defined on [0,∞[ such that ax ≤ h(x) ≤ bx for
every x ∈ [0, +∞[, h(q2j−1) = aq2j−1 and h(q2j) = bq2j for every positive integer j.
Since ax ≤ h(x) ≤ bx for x ∈ [0,∞[, we obviously have ax ≤ h∗(x) and h∗(x) ≤ bx for
every x ≥ 0. But we also have lim inf t→0+ h(t)/t = a and lim supt→0+ h(t)/t = b, because of
the values of h at points in the sequences (q2j−1)∞
j=1, respectively. By Theorem
1 we have h∗(x) ≤ ax and h∗(x) ≥ bx for each x ≥ 0, completing the proof.
2
j=1 and (q2j)∞
Observe that the functions h from Corollary 2 have the property that (h∗)∗(x) = ax <
bx = (h∗)∗(x) for all x > 0.
Corollary 3 There is an infinite number of smooth aggregation functions h on [0,∞] such
that h∗(x) = 0 for every x < ∞ and h∗(x) = ∞ for every x > 0.
Proof. For every positive integer k let xk = 2−2k. For x ≥ 0 let f (x) = x5/4 and g(x) =
x3/4; a straightforward calculation shows that g(x2j) < f (x2j−1) for every positive integer
j. By known results from calculus there exists an infinite number of smooth increasing
functions h on [0,∞[ such that h(x2j−1) = f (x2j−1) and h(x2j) = g(x2j) for all positive
integers j.
Since for our function h we have lim inf t→0+ h(t)/t = 0 and lim supt→0+ h(t)/t = +∞
j=1, the result is again
2
due to the values of h at points in the sequences (x2j−1)∞
a consequence of Theorem 1.
j=1 and (x2j)∞
4
bxaxqqqq2j - 12j 2j + 12j + 20.51.01.551015202530Figure 2: A schematic drawing of a function h from the proof of Corollary 3.
The functions h from Corollary 3 have even a more striking property that (h∗)∗(x) = 0
for every x ≥ 0 while (h∗)∗(x) = ∞ for all x > 0.
To conclude this section we underscore the fundamental role of Theorem 1 by the
following complete characterization of one-dimensional degeneracies.
Theorem 2 Let h be a one-dimensional aggregation function on [0,∞[. The following
conditions are equivalent:
(a) There exists an x > 0 for which h∗(x) = ∞.
(b) h∗(x) = ∞ for every x > 0.
(c) lim supt→0+
(d) sup{ h(t)
t ∈]0, x]} = ∞ for some x > 0.
Similarly, the following statements are equivalent:
t = ∞.
h(t)
t
(a’) There exists an x > 0 for which h∗(x) = 0.
(b’) h∗(x) = 0 for every x ≥ 0.
(c’) lim inf t→0+
(d’) inf{ h(t)
t ∈]0, x]} = 0 for some x > 0.
h(t)
t = 0.
t
Proof. Clearly, the statements (c) and (d) are equivalent, and so are (c’) and (d’). By
Theorem 1, (c) implies (b) and (c’) implies (b’). Trivially, (b) implies (a) and (b’) implies
(a’), and so one only has to prove that (a) implies (d) and (a’) implies (d’).
To show that (a) implies (d), we prove the contrapositive. Assume that sup{ h(t)
t ∈
]0, x]} = b < ∞ for every x > 0. This means that h(t) ≤ bt for every t ∈]0, x]. Thus,
i=1 xi ≤ x we have
for every n-tuple (x1, x2, . . . , xn) of positive real numbers such that(cid:80)n
(cid:80)n
i=1 h(xi) ≤ b(cid:80)n
i=1 xi ≤ bx. It follows that h∗(x) ≤ bx < ∞ for every x > 0.
t
5
gfxxxx2j - 12j 2j + 12j + 20.10.20.30.40.50.61234Similarly, to show that (a’) implies (d’) we again proceed by proving the contrapositive.
Suppose that inf{ h(t)
t ∈]0, x]} = b > 0 for every x > 0. This means that h(t) ≥ bt for
every t ∈]0, x]. Thus, for every n-tuple (x1, x2, . . . , xn) of real numbers from ]0, x] such
i=1 xi ≥ bx. It follows that h∗(x) ≥ bx > 0 for
i=1 xi ≥ x we have(cid:80)n
i=1 h(xi) ≥ b(cid:80)n
that(cid:80)n
t
every x > 0. This completes the proof.
2
3 The multidimensional case
Based on the results in the one-dimensional case proved in Section 2 we continue by ex-
hibiting examples of aggregation functions A with the property that the values of (A∗)∗ are
smaller than the values of (A∗)∗ for all non-zero vectors x ∈ [0,∞[n. The method is based
on the observation that the values of A∗ and A∗ on the entire space [0,∞[n are influenced by
the behavior of the one-dimensional diagonal function A(x, . . . , x) in an arbitrarily small
neighbourhood of zero. We present details only regarding extensions of Corollary 3 to
arbitrary dimensions.
Let µx be the arithmetic mean of the entries in x, that is, if x = (x1, x2, . . . , xm), then
µx = (x1 + x2 + . . . + xm)/m.
Theorem 3 There are infinitely many aggregation functions A defined on [0,∞]n such
that (A∗)∗(x) = 0 for all x ∈ [0,∞]n while (A∗)∗(x) = ∞ for all non-zero x ∈ [0,∞[n.
Proof. It is sufficient to take any function h from Corollary 3 and define A for every
x ∈ [0,∞[n by A(x) = h(µx). Letting x = µx, for any non-zero x we have, exactly as in
the proof of Theorem 1,
h(t)
A∗(x) ≤ nh(x/n) ≤ x · lim inf
t→0+
t
A∗(x) ≥ nh(x/n) ≥ x · lim sup
t→0+
h(t)
t
.
, and
The claim now follows from Corollary 3.
2
Clearly, one can use numerous other compositions of functions h from Corollary 3 with
simple aggregation functions (such as weighted average, geometric means, etc.) to provide
examples for Theorem 3.
We also prove different sufficient conditions for the values of A∗ and A∗ to exhibit the
extreme behavior described in Theorem 3.
Proposition 1 Let A be an aggregation function on [0,∞]n and let A(x) = A(x, x, . . . , x)
for every x ≥ 0.
6
(1) If there exists a divergent series (cid:80)∞
the series(cid:80)∞
(2) If there exists a convergent series(cid:80)∞
the series(cid:80)∞
j=1
j=1
A(aj) converges, then A∗(x) = 0 for every x ∈ [0,∞[n.
j=1 aj with decreasing positive terms such that
A(aj) diverges, then A∗(x) = +∞ for every non-zero x ∈ [0,∞]n.
j=1 aj with decreasing positive terms such that
2
Proof. Let h(x) = A(x, x, . . . , x) for x ≥ 0. We show that the assumption of (1) implies
j aj would imply divergence of (cid:80)
divergence of (cid:80)
h(aj) ≤ (c + )aj for all but finitely many j’s. But then convergence of(cid:80)
convergence of(cid:80)
that lim inf j→∞ h(aj)/aj = 0 and (2) implies that lim supj→∞ h(aj)/aj = +∞.
Indeed, suppose that lim inf j→∞ h(aj)/aj = c > 0. This means that for every > 0
we have h(aj) ≥ (c − )aj for all but a finite number of positive integers j. But then,
j h(aj), contrary to the assumption of
(1). Similarly, if lim supj→∞ h(aj)/aj = c < +∞, then for every > 0 we would have
j aj would imply
This shows that lim inf t→0+ h(t)/t = 0 in the case (1) and lim supt→0+ h(t)/t = +∞ in
j h(aj), a contradiction.
the case (2). The result now follows from Theorems 1 and 3.
or f (x) = x/ ln2(x), and take for (cid:80)∞
an arbitrarily small λ > 0, or g(x) = x ln2(x), and take for(cid:80)∞
As examples, one can take the function f (x) = x1+λ for some arbitrarily small λ > 0,
j=1 aj the harmonic series to construct aggregation
functions A such that A(x) = f (x) on an arbitrarily small interval (0, δ); the result of (1)
then gives A∗(x) = 0 on [0,∞[n. Similarly, one can take the function g(x) = x1−λ for
j=1 aj the series with aj =
g−1(1/j) to obtain aggregation functions A such that A(x) = g(x) on an arbitrarily small
interval (0, δ); the result of (2) then shows that A∗(x) = +∞ for all non-zero x ∈ [0,∞[n.
n and Kn∗
by looking at the one-dimensional case. For an aggregation function A on [0,∞]n let A be
defined on [0,∞] by letting A(x) = A(x, x, . . . , x) and, for every i ∈ {1, 2, . . . , n} let Ai be
defined on [0,∞] by Ai(x) = A(xei), where ei is the i-th unit vector.
Theorem 4 Let A be an aggregation function on [0,∞]n. Then,
With the help of the results of Section 2 we can also decide membership in K∗
(i) A[0,∞[n∈ K∗
(ii) A[0,∞[n∈ Kn∗ if and only if Ai[0,∞[∈ K1∗ for some i ∈ {1, 2, . . . , n}.
n if and only if A[0,∞[∈ K∗
1, and
.
Proof. (i): We only need to show that A∗(x) = ∞ for some non-zero vector x ∈ [0,∞[n
if and only if ( A)∗(x) = ∞ for some x > 0. For the direct implication, by monotonicity
it is sufficient to take, for a given non-zero vector x, the value x equal to the maximum
of the coordinates of x; the reverse implication follows by taking, for a given x > 0, the
vector x = (x, x, . . . , x).
(ii): Here we only need to show that A∗(x) = 0 for some non-zero vector x ∈ [0,∞[n
if and only if (Ai)∗(x) = 0 for some x > 0 and some i ∈ {1, 2, . . . , n}. For the direct
7
implication it is sufficient to take an i ∈ {1, 2, . . . , n} for which the i-th coordinate of x
has a non-zero value x; by monotonicity we have (Ai)∗(x) = 0. The reverse implication
2
follows by simply taking x = xei.
As an example of this result consider the aggregation function A defined for any x ∈
It follows
[0,∞]3 by taking A(x) to be the median of the coordinates of the vector x.
(cid:51) but it does not belong to K(cid:51)
immediately from Theorem 4 that A[0,∞[3∈ K∗
∗ .
4 Comparing (A∗)∗ with (A∗)∗
In this section we prove that the values of (A∗)∗ can never be larger than the values of
(A∗)∗ for an aggregation function A defined on [0,∞]n. The proof strongly depends on the
fact that A is defined for non-zero vectors of [0,∞]n that are arbitrarily close to the zero
vector.
Theorem 5 Let A :
[0,∞]n → [0,∞] be an aggregation function. Then (A∗)∗ ≤ (A∗)∗.
Proof. For our given aggregation function A and for every i ∈ {1, . . . , n} let Ai :
[0,∞[→ [0,∞[ be the marginal function of A considered before the proof of Theorem 4,
given by Ai(x) = A(xei) = A(0, ..., x, ..., 0), where x appears only in the i-th coordinate.
It is not difficult to check that
A∗(x1, ..., xn) ≤ n(cid:88)
(A∗)∗(x1, ..., xn) ≤ n(cid:88)
(Ai)∗(xi) and A∗(x1, ..., xn) ≥ n(cid:88)
((Ai)∗)∗(xi) and (A∗)∗(x1, ..., xn) ≥ n(cid:88)
i=1
i=1
It follows that
Let now
i=1
i=1
ai = lim inf
t→0+
Ai(t)
t
and bi = lim sup
t→0+
Ai(t)
t
.
By Theorem 1 from Section 2 we have
(Ai)∗(x) ≤ aix ≤ bix ≤ (Ai)∗(x)
(Ai)∗(xi) .
((Ai)∗)∗(xi) .
which implies that
((Ai)∗)∗(x) ≤ aix ≤ bix ≤ ((Ai)∗)∗(x) .
Consequently, for every x = (x1, x2, . . . , xn) ∈ [0,∞[n we have
(A∗)∗(x) ≤ n(cid:88)
((Ai)∗)∗(xi) ≤ n(cid:88)
aixi ≤ n(cid:88)
bixi ≤ n(cid:88)
((Ai)∗)∗(xi) = (A∗)∗(x) ,
i=1
i=1
i=1
i=1
8
which completes the proof.
2
We also include an observation about the other extreme, that is, when (A∗)∗ = (A∗)∗.
A characterization of aggregation functions A on [0,∞]n for which this equality holds
is still an open problem. Examples of such non-linear functions can be constructed, for
instance, using Corollary 1 in conjunction with Theorem 3. However, if (A∗)∗ = (A∗)∗
we can at least say that the result is an additive function, i.e., a function D satisfying
D(x + y) = D(x) + D(y) for every x, y ∈ [0,∞[n. We formulate and prove the result in a
slightly larger generality.
Corollary 4 Let B and C be aggregation functions on [0,∞]n such that B∗(x) = C∗(x)
for every x ∈ [0,∞[n. Then, the function D = B∗ = C∗ is additive, D(x) =
wixi for
some w = (w1, ..., wn) ∈ [0,∞]n.
n(cid:80)
i=1
Proof. By the results of [4], B∗ is sub-additive and C∗ is super-additive; it follows that
2
D is both sub- and super-additive and hence additive.
n(cid:80)
i=1
Example 1 (i): Define A ∈ An by A(x1, . . . , xn) = max(x1, . . . , xn). Then A ∈ Pn and
A = A∗, A∗(x1, . . . , xn) =
xi, and (A∗)∗ = (A∗)∗ = A∗.
(ii): Consider B, C ∈ An given by B(x1, . . . , xn) = ln(Πn
exi) − n. Then B∗ = B, C∗ = C and B∗ = C∗ is given by B∗(x1, . . . , xn) =
i=1(1+xi)) and C(x1, . . . , xn) =
n(cid:80)
xi.
n(cid:80)
(
i=1
i=1
5 Concluding remarks
Our main aim was to characterize proper aggregation functions A with the property that A∗
(and, similarly, A∗) is proper as well. In the course of our investigation we obtained a num-
ber of related results on the behavior of the subadditive and superadditive transformations
A∗ and A∗. Regarding iterations of these transformations we proved that (A∗)∗ ≤ (A∗)∗
and that these two functions may be arbitrarily far from each other.
In a nutshell and in a somewhat more abstract setting, we clarified constraints for aggre-
gation functions admitting superadditive/subadditive transformations without a complete
loss of information. The transformations considered are related to optimization problems,
for example to production functions in economical problems. For a deeper discussion we
recommend [4].
Acknowledgement. The author acknowledges support from the projects VEGA 1/0420/15
and APVV 0013/14.
9
References
[1] Beliakov, G., Pradera, A., Calvo, T.: Aggregation Functions: A Guide for Prac-
titioners. Springer, Berlin, 2007 ISBN: 978-3-540-73720-9 (Print) 978-3-540-73721-6
(Online)
[2] Denardo, Eric V.: Linear Programming and Generalizations. Springer, Berlin, 2011,
ISBN 978-1-4419-6491-5
[3] Grabisch, M., Marichal, J.- L., Mesiar, R., Pap, E.: Aggregation Functions (Encyklo-
pedia of Mathematics and its Applications). Cambridge University Press, 2009
[4] Greco, S., Mesiar, R., Rindone, F., Sipeky, L.: The superadditive and the subaddi-
tive transformations of integrals and aggregation functions. Fuzzy Sets and Systems,
DOI:10.1016/j.fss.2015.08.006
[5] Lehrer, E.: A new integral for capacities., Economic Theory, Springer, 2009, vol. 39
(1), p. 157-176
[6] Mesiar, R., Li, J., Pap, E.: Superdecomposition integrals. Fuzzy Sets and Systems 259
(2015) p. 3-11
10
|
1807.06692 | 1 | 1807 | 2018-07-17T22:24:03 | A characterization of superreflexivity through embeddings of lamplighter groups | [
"math.FA",
"math.GR",
"math.MG"
] | We prove that finite lamplighter groups $\{\mathbb{Z}_2\wr\mathbb{Z}_n\}_{n\ge 2}$ with a standard set of generators
embed with uniformly bounded distortions into any non-superreflexive Banach space, and therefore form a set of test-spaces for superreflexivity. Our proof is inspired by the well known identification of Cayley graphs of infinite lamplighter groups with the horocyclic product of trees. We cover $\mathbb{Z}_2\wr\mathbb{Z}_n$ by three sets with a structure similar to a horocyclic product of trees, which enables us to construct well-controlled embeddings. | math.FA | math |
A characterization of superreflexivity through
embeddings of lamplighter groups
Mikhail I. Ostrovskii and Beata Randrianantoanina
July 19, 2018
Abstract
We prove that finite lamplighter groups {Z2 ≀ Zn}n≥2 with a standard
set of generators embed with uniformly bounded distortions into any non-
superreflexive Banach space, and therefore form a set of test-spaces for super-
reflexivity. Our proof is inspired by the well known identification of Cayley
graphs of infinite lamplighter groups with the horocyclic product of trees. We
cover Z2 ≀ Zn by three sets with a structure similar to a horocyclic product of
trees, which enables us to construct well-controlled embeddings.
2010 Mathematics Subject Classification. Primary: 46B85; Secondary: 05C12,
20F65, 30L05.
Keywords. distortion of a bilipschitz embedding, horocyclic product of trees, lamplighter
group, Lipschitz map, metric embedding, Ribe program, superreflexivity, word metric
1
Introduction
One of the important directions in metric geometry is to find purely metric charac-
terizations of interesting classes of Banach spaces. For classes of spaces determined
by finite-dimensional subspaces, this direction is a part of the Ribe program which
was described by Bourgain [4] who proved the first metric characterization of su-
perreflexivity. See [16] for more information on the Ribe program. The goal of this
paper is to prove that lamplighter groups are test spaces for superreflexivity.
Definition 1.1 ([21]). Let P be a class of Banach spaces and let T = {Tα}α∈A be a
set of metric spaces. We say that T is a set of test spaces for P if the following two
conditions are equivalent: (1) X /∈ P. (2) The spaces {Tα}α∈A admit bilipschitz
embeddings into X with uniformly bounded distortions.
Several different sets of test-spaces for superreflexivity are known: (1) binary
trees [4, 14, 2, 9], (2) binary diamond and Laakso graphs [8, 15], (3) multibranching
1
diamond and Laakso graphs [24]. See [23] for a survey on this matter written in
2014.
In this paper we add one more item to this list: (4) the set of Cayley graphs of
finite lamplighter groups. Moreover, we observe that the Cayley graph of an infinite
lamplighter group also is a test space for superreflexivity.
Some of the characterizations (1)–(3) are independent in the sense that the cor-
responding families of test spaces do not admit bilipschitz embeddings into each
other with uniformly bounded distortions. In non-obvious cases this was shown in
[22, 11] for finite binary trees and diamond graphs, and in [18] for diamond and
Laakso graphs.
Lamplighter groups are a very interesting class of groups which has been a rich
source of important examples in geometric group theory. In 2008, Naor and Peres
[17, Section 4] proved that the finite lamplighter groups Z2 ≀ Zn, with metric defined
as a word length with respect to natural sets of generators, are embeddable into L1
with uniformly bounded distortions. The first goal of this paper is to strengthen this
result and to prove their embeddability into an arbitrary nonsuperreflexive Banach
space with uniformly bounded distortions. As a consequence we get a new metric
characterization of superreflexivity, see Corollary 1.7.
We consider the following special case of the general wreath product construction
(see, for example, [6, p. 214] for the general definition).
Definition 1.2 ([5, p. 129]). Let H and L be two groups and LH be the set of all
L-valued finitely supported functions on H. Then the wreath product L≀H is defined
as the set LH × H equipped with the multiplication
((xh)h∈H, g) · ((yh)h∈H, k) := ((xh · ygh)h∈H, gk).
Our main interest will be the wreath product Z2 ≀ Zn, for n ∈ N, n ≥ 2, which
we denote by Gn, this group is called the lamplighter group, see [27] for a nice
introduction including an explanation for the name.
We identify ZZn
with {j ∈ Zn : xj = 1}, the group operation on ZZn
From now on we will abuse notation and treat an element x ∈ ZZn
Zn.
2 with the family of all subsets of Zn by identifying x = (xk)k∈Zn
is the symmetric difference.
2 as a subset of
2
Considering an element (x, k) ∈ Gn, we call x the set of positions where the lamp
is on, and its complement - the set of positions where the lamp is off. The number
k ∈ Zn is called the location of the lamplighter.
It is easy to see that the elements a = ({0}, 0) and t = (∅, 1) generate Gn.
Observe that multiplication by a = ({0}, 0) on the right is the act of changing the
lamp at the current location of the lamplighter and multiplication by t = (∅, 1) on
the right is the act of the lamplighter moving one position to the next lamp in the
'positive' direction around the circle.
We consider the metric ρ on Gn defined as the metric of the left-invariant Cayley
graph with respect to the set of generators S = {t, ta}. This means that x is adjacent
2
to y if and only if x = ys or y = xs (i.e x = ys−1) for one of the generators s ∈ S.
Observe that the generator ta acts by first moving one step in the 'positive' direction,
and then changing the state of the lamp at the final location of the lamplighter. See
Section 2.2 for a more detailed description and an equivalent formula (2.5) for the
metric.
The first main result of this paper is
Theorem 1.3. For any nonsuperreflexive Banach space X the Cayley graphs of
Z2 ≀ Zn (n ≥ 2) corresponding to the set S = {t, ta} admit embeddings into X with
uniformly bounded distortions.
Remark 1.4. In [17], when proving embeddability into L1, Naor and Peres considered
Z2 ≀ Zn with the set of generators equal to {t, a} instead of {t, ta} as we do. However,
it is easy to see that the metrics induced by these two generating sets are bilipschitz
equivalent to each other with a distortion 4.
In [17] Naor and Peres constructed two embeddings into L1, one based on ir-
reducible representations of finite lamplighter groups, and the other motivated by
what they refer to as a 'direct geometric reasoning'. Our embedding technique is
quite different from either of the embeddings in [17]. Our approach is inspired by the
description of the Cayley graph of the infinite lamplighter group Zq ≀ Z, for q ∈ N,
q ≥ 2, as a horocyclic product of two trees, introduced by Bartholdi and Woess
[1, 28], and by the analysis of the metric structure of horocyclic products of trees
by Stein and Taback in [26] who proved, among other results,
Theorem 1.5. [26, Corollary 10] For any q ∈ N, q ≥ 2, the Cayley graph of Zq ≀ Z
with the set of generators {t, ta, . . . , taq−1} admits an embedding into an ℓ1-sum of
two q-branching trees with distortion bounded by 4.
We refer the reader to [27] and [29] for very nice presentations of horocyclic
products of trees and their applications to lamplighter groups. We give a more
detailed overview of our method in Section 2.1.
Remark 1.6. It follows from Theorem 1.5, [4], and [19] that for all q ∈ N, q ≥ 2, the
Cayley graph of Zq ≀ Z admits an embedding into any non-superreflexive space with
distortion independent of q, as was shown in a similar case of hyperbolic groups in
[22, Section 2] (cf. also Section 2.4 below).
The bilipschitz embeddability of the infinite group Z2 ≀ Z into any non-superre-
flexive space can also be derived from Theorem 1.3 and results of [4, 19].
In [13] it is proved that there exists a constant c > 0 so that for every n ∈ N a
complete binary tree of depth cn embeds with constant distortion into Gn. In Sec-
tion 3 below we show an alternative simple proof of this fact using our construction.
As a consequence, and by Theorem 1.3, Remark 1.6, and [4], we obtain
3
Corollary 1.7. The sequence of Cayley graphs for {Z2 ≀ Zn}n≥2 with respect to
the set of generators S = {t, ta} is a set of test-spaces for superreflexivity. The
Cayley graph of Z2 ≀ Z with respect to any finite set of generators is a test space for
superreflexivity.
Remark 1.8. We note that Theorem 1.5 is valid for Zq ≀ Z for all q ∈ N, q ≥ 2. As
we elaborate in Section 2.1, our proof of Theorem 1.3 is inspired by the methods of
Theorem 1.5, even though we do not apply its conclusion for our argument.
In our statement and proof of Theorem 1.3, for greater clarity of the presentation,
we focused our attention on Z2 ≀ Zn, but it only requires straightforward adjustments
of the proof to obtain the same conclusion as in Theorem 1.3 for Zq ≀ Zn with the set
of generators {t, ta, . . . , taq−1}, for all q ∈ N, q ≥ 2. The only difference is that one
needs to use an ℓ∞-sum of two q-branching trees in place of binary trees. Similarly
as in [26], the value of q affects the branching of the trees, but not the number of
the summands. Since for all q, a q-branching tree embeds almost isometrically into
a binary tree, and since we always use an ℓ∞-sum of two trees, the uniform bound
on distortions does not depend on q ∈ N.
Since, for all q ∈ N, q ≥ 2, the Cayley graphs of Zq ≀ Zn and Zq ≀ Z with the
set of generators {t, ta, . . . , taq−1} isometrically contain the Cayley graphs of Z2 ≀ Zn
and Z2 ≀ Z with generators {t, ta}, respectively, it follows from Corollary 1.7 and the
adjusted version of Theorem 1.3, that for any q ∈ N, q ≥ 2, the sequence of Cayley
graphs of Zq ≀ Zn with the set of generators {t, ta, . . . , taq−1} is a set of test-spaces
for superreflexivity. Similarly, for any q ∈ N, q ≥ 2, the Cayley graph of Zq ≀ Z, with
respect to any finite generating set, is a test space for superreflexivity.
2 Proof of Theorem 1.3
2.1 Outline of the proof
To simplify notation, we assume that n is divisible by 6. It is clear that the same
arguments work in general, but the formulas will be somewhat more complicated.
We cover the Cayley graph of Zn with respect to the generating set {±1} by
three overlapping paths P1, P2, P3 of length 2
3 n each in such a way that each pair of
points in Zn belongs to at least one of the paths (paths that exclude three mutually
disjoint thirds of the cycle Zn work).
We consider the following three subsets of Gn for i = 1, 2, 3,
Pi,n
def= {(x, k) ∈ Gn x ⊆ Zn, k ∈ Pi}.
(2.1)
We equip Pi,n with the metric inherited from the Cayley graph of Gn with respect
to S = {t, ta}. We note that the union of the three sets Pi,n covers Gn.
Our approach is inspired by the well-known description of infinite lamplighter
groups Z2 ≀Z as a horocyclic product of two infinite trees [1, 28] and by the bilipschitz
4
embedding of the word metric on Z2 ≀ Z into an ℓ1-sum of two trees [26] (see Theo-
rem 1.5 above). Our first goal is to cover Z2 ≀ Zn by three sets with a structure similar
to a horocyclic product of trees, which will enable us to construct well-controlled
embeddings.
The sets Pi,n are defined so that for all locations k, l ∈ Pi, the length of the
path that is contained in Pi and connects k and l is at most twice the length of the
shortest path from k to l in Zn. For this reason, the metric structure of the sets Pi,n
sufficiently resembles the metric structure of a subset of Z2 ≀ Zn and we are able to
construct sets Wn, which we think of as analogs of horocyclic products of trees, and
which are bilipschitz equivalent to the sets Pi,n. The sets Wn are defined as specific
subsets of the ℓ∞-sum of two trees of depth n. To identify each element of Pi,n with
an element of the Cartesian product Tn × Tn of two trees of depth n, we first mark
a vertex v0 in Zn which is the midpoint of the complement of the path Pi, that is v0
is at the distance at least n/6 from any element k ∈ Pi. Modelling our description
on the identification of Z2 ≀ Zn with the horocyclic product of trees (cf. [28, p. 419]),
for every (x, k) ∈ Pi,n we describe the set x as the union of x ∩ Ik,+ and x ∩ Ik,−,
where Ik,+, Ik,− are two disjoint arcs in Zn both with the endpoints v0 and k. Each
set x ∩ Ik,+ and x ∩ Ik,− is encoded by a sequence of 0s and 1s of length equal to
the number of vertices in Ik,+ and Ik,−, respectively. This naturally encodes each
element (x, k) ∈ Pi,n by two elements of a binary tree whose levels (in a tree) add
up to n. We verify in Section 2.2 that this encoding is metrically faithful on Pi,n,
that is, we construct bilipschitz embeddings
ϕi,n : Pi,n → Tn ⊕∞ Tn,
with uniformly bounded distortions, see Section 2.2 for details. This completes the
first and main step of our proof.
The next step of our proof is a routine application of the well-known theory
of Lipschitz retracts to conclude that, for each i ∈ {1, 2, 3}, the bilipschitz map
ϕi,n : Pi,n → Tn ⊕∞ Tn, constructed in the first step, can be extended to a Lipschitz
map ¯ϕi,n from the entire Gn into an ℓ∞-sum of two metric trees of depth n, see
Section 2.3 for details.
In the final step of our proof we define the map Φn from Gn into an ℓ∞-sum of
six metric trees of depth n by
Φn(x, k) def= ( ¯ϕ1,n(x, k), ¯ϕ2,n(x, k), ¯ϕ3,n(x, k)).
Clearly, the maps Φn are Lipschitz with the same Lipschitz constants as those
of the maps ¯ϕi,n, for i = 1, 2, 3. Since the paths P1, P2, P3 were chosen in such a
way that for any two elements (x, k), (y, l) ∈ Gn, there exists i ∈ {1, 2, 3} so that
(x, k), (y, l) ∈ Pi,n, it follows that the map Φn is co-Lipschitz with same constant
as the map ϕi,n (in our construction all maps {ϕi,n}3
i=1 have the same co-Lipschitz
constant), see Section 2.4.
5
Hence to finish the proof of Theorem 1.3, it is enough to verify that for all n ∈ N,
the ℓ∞-sum of six metric trees of depth n embeds into any non-superreflexive Banach
space X with uniformly bounded distortions. This follows readily by known tech-
niques and results on bilipschitz embeddability of trees into any non-superreflexive
Banach space, and on extension of bilipschitz embeddings into Banach spaces from
vertex sets to graphs to the corresponding 1-dimensional complexes. In fact it even
suffices to prove the existence of Lipschitz maps that satisfy slightly weaker require-
ments, see Section 2.4 for details.
2.2 Step 1
We define the paths P1, P2 and P3, to be arcs of lengths 2
[ 1
6 n, 5
by 6).
3 n in Zn with endpoints
2n], respectively (recall that we assumed that n is divisible
6n], [ 1
2n, 1
6 n] and [ 5
6n, 1
Let θn : Zn → Zn be the 'rotation' of Zn by an arc of length n/3, i.e. for each
3 , where the addition is in the sense of Zn. The rotation θn
k ∈ Zn, θn(k) = k + n
induces the isometry θn of Gn onto itself, defined by
θn(x, k) = (θn(x), θn(k)),
where, as usual, θn(x) denotes the image of the set x under the action of θn.
Note, that the paths P1, P2, P3, satisfy P2 = θn(P1) and P3 = θ2
n(P1), and the
metric spaces {Pi,n}3
i=1, defined by (2.1), are isometric to each other, specifically
P2,n = θn(P1,n),
P3,n = θn
2
(P1,n).
(2.2)
Let Tn be a binary tree of depth n, that is, Tn is the graph whose vertices are
labelled by sequences of 0s and 1s of lengths ≤ n with the usual graph distance dT .
We consider the ℓ∞-sum Tn ⊕∞ Tn defined as the Cartesian product Tn ×Tn endowed
with the metric
d∞((A1, A2), (B1, B2)) def= max{dT (A1, B1), dT (A2, B2)}.
We define
Wn
def= (cid:26)(A1, A2) ∈ Tn ⊕∞ Tn :
A1 + A2 = n,
A1, A2 ∈(cid:20)1
6
n,
5
6
n(cid:21)(cid:27) ,
where A denotes the length of the sequence A ∈ Tn.
It will be convenient to also use the distance d1 on Wn which is 2-equivalent with
d∞.
d1((A1, A2), (B1, B2)) def= dT (A1, B1) + dT (A2, B2).
We will show that there exist bijections ϕ1,n from {(P1,n, d∞)}n onto {(Wn, d∞)}n
which have uniformly bounded distortions.
6
We define a map ϕ1,n : P1,n → Wn as follows: for any (x, k) ∈ P1,n, the element
ϕ1,n(x, k) def= (A1, A2), where A1 = (a1,0, . . . , a1,k−1) is a sequence of length k and
A2 = (a2,1, . . . , a2,n−k) is a sequence of length n − k, defined by
a1,j =(1
0
if j ∈ x,
if j /∈ x,
a2,i =(1
0
if n − i ∈ x,
if n − i /∈ x.
It is clear that the map ϕ1,n is one-to-one and onto. We will show that ϕ1,n is a
bilipschitz isomorphism of (P1,n, ρ) and (Wn, d∞).
For any A, B ∈ Tn, we denote by lgca(A, B) the length of the greatest common
ancestor of A and B in Tn. In this notation,
dT (A, B) = (A − lgca(A, B)) + (B − lgca(A, B)),
and for (A1, A2), (B1, B2) ∈ Wn,
d1((A1, A2), (B1, B2)) = dT (A1, B1) + dT (A2, B2)
= 2(n − (lgca(A1, B1) + lgca(A2, B2))).
In particular, for all (A1, A2), (B1, B2) ∈ Wn we have
d1((A1, A2), (B1, B2)) ≤ 2n.
(2.3)
(2.4)
To continue we need to estimate the distance ρ((x, k), (y, l)), where (x, k), (y, l) ∈
Gn and ρ is the distance in the Cayley graph of Gn with the generating set {t, ta}.
We use the following observation: to get from (x, k) to (y, l) we need
• to traverse at least one of the two paths from k to l on the n-cycle (the graph
of Zn with respect to the generating set {±1});
• to visit all positions j ∈ Zn which are not on the selected path, but which
belong to x△y, and to change the state of all lamps at these positions.
Denote by p1 and p2 the lengths of the two distinct paths from k to l on Zn,
and by g1 and g2 – the sizes of the largest "gaps" in these paths, that is, the largest
distances between distinct vertices for which there is no element of x△y in between
(observe that g1 and g2 are at least 1 each). With this notation it is easy to see the
validity of the leftmost inequality in
min{p1 + 2(p2 − g2), p2 + 2(p1 − g1)} ≤ ρ((x, k), (y, l))
≤ min{p1 + 2(p2 − g2), p2 + 2(p1 − g1)} + 2.
(2.5)
The rightmost inequality in (2.5) holds because discrepancies with the equality
in (2.5) can occur only at one of the endpoints of the 'interval' on Zn consisting
7
of all vertices that are visited by an optimal tour from k to l that establishes the
distance between (x, k) and (y, l). One of the cases when the distance exceeds the
minimum by 2 is the following: the distance from k to l in the positive direction is
significantly smaller than in the negative direction and x△y = {k, l}. In this case,
if we start at k we need to do one step in the negative direction in order to change
the status of the lamp at k, and then head back in the positive direction. Note that
the position l is reached by the step ta in order to change the status of the lamp
there, so no additional steps are needed at this endpoint.
For the rest of the proof we fix (x, k), (y, l) ∈ P1,n, and (A1, A2) = ϕ1,n(x, k),
(B1, B2) = ϕ1,n(y, l) in Wn.
Observation 2.1. The sum lgca(A1, B1) + lgca(A2, B2) is equal to the number of
vertices of the set E constructed as a union of two, possibly empty, 'intervals' in Zn.
One of the intervals starts at 0 ∈ Zn, goes in the 'positive' direction and ends at
the first vertex which belongs to x△y ∪ {k, l}, excluding this vertex, in particular if
0 ∈ x△y∪{k, l}, then the interval is empty. The other interval starts at (n−1) ∈ Zn,
goes in the 'negative' direction and ends at the first vertex which is in x△y ∪ {k, l},
this interval includes its end if and only if it does not belong to x△y. Since the
number of vertices in the interval E is equal to either its length E, if E is empty,
or to E + 1, otherwise, we obtain
E ≤ lgca(A1, B1) + lgca(A2, B2) ≤ E + 1.
(2.6)
Note that (2.3) and (2.6) immediately imply that
d1(ϕ1,n(x, k), ϕ1,n(y, l)) ≥ 2(n − g),
where g is the number of vertices in the largest 'interval' I in Zn which is disjoint
with {k, l} ∪ x△y. Since it is clear that there exists a 'tour' of the lamplighter
that travels from k to l, passes through every vertex of the difference x△y, visits
each vertex at most twice, and stays outside I except possibly one vertex (see the
paragraph after (2.5)), for all pairs (x, k), (y, l) ∈ P1,n we get the inequality
ρ((x, k), (y, l)) ≤ 2 + 2(n − g) ≤ 2 + d1(ϕ1,n(x, k), ϕ1,n(y, l))
≤ 2 + 2d∞(ϕ1,n(x, k), ϕ1,n(y, l)).
To get the inequality in the other direction we observe that ρ((x, k), (y, l)) is at
least τ − 1, where τ is the number of vertices in the smallest 'interval' J on Zn that
contains all elements of {k, l} ∪ x△y. Note that J is nonempty, and thus τ = J + 1.
If the interval J does not contain 0, then J and the interval E constructed in
Observation 2.1 have at most one point in common, and the union J ∪ E covers Zn.
8
Thus the length of J is at least n − E − 1. In this case, by (2.3) and (2.6), we get
ρ((x, k), (y, l)) ≥ τ − 1 ≥ n − E − 1
≥
≥
1
2
1
2
(d1(ϕ1,n(x, k), ϕ1,n(y, l))) − 1
(d∞(ϕ1,n(x, k), ϕ1,n(y, l))) − 1.
If the interval J contains 0, then it contains at least 1
3 n − 1 vertices, because by
the definition of P1,n we have k, l ∈ P1 =(cid:2) 1
ρ((x, k), (y, l)) ≥
n − 2
6 n, 5
6 n(cid:3). So in this case, by (2.4), we get
1
3
1
6
1
6
≥
≥
max{d1((A1, A2), (B1, B2)) : (A1, A2), (B1, B2) ∈ Wn} − 2
(d∞(ϕ1,n(x, k), ϕ1,n(y, l))) − 2.
This ends the proof that the bijections ϕ1,n : (P1,n, d∞) → (Wn, d∞) have uni-
formly bounded distortions.
Since, by (2.2), the spaces P2,n and P3,n are isometric to P1,n, the maps ϕ2,n
def=
ϕ1,n ◦ θn, and ϕ3,n
are bijections from P2,n and P3,n, respectively, onto
(Wn, d∞), and they have the same Lipschitz and co-Lipschitz constants as the maps
ϕ1,n.
def= ϕ1,n ◦ θn
2
2.3 Step 2
We will apply the well-known theory of Lipschitz retracts.
For basic theory of Lipschitz retracts in metric spaces see e.g.
[3, Chapter 1]
and [10, Propositions 2.1, 2.2, and the comment at the top of page 303]. For the
convenience of the reader, we briefly recall the definitions and results that we use.
A metric space M is called injective if for every metric space B, every A ⊆ B,
and every Lipschitz function f : A → M, there exists a Lipschitz extension of f ,
that is, a function ¯f : B → M, so that ¯f A = f and Lip( ¯f ) = Lip(f ).
A metric space M is called a λ-absolute Lipschitz retract (where 1 ≤ λ < ∞)
provided that whenever X is isometrically contained in a metric space Y , there
exists a retraction, r, from Y onto X, with Lip(r) ≤ λ.
Theorem 2.2 ([10, Proposition 2.2]). A metric space M is injective if and only if
it is an absolute 1-Lipschitz retract.
A metric space M is said to have the binary intersection property if every collec-
tion of mutually intersecting closed balls in M has a common point.
A metric space M is said to be metrically convex if for every x0, x1 ∈ M and
for every 0 < t < 1 there is a point xt ∈ M such that d(x0, xt) = td(x0, x1) and
d(x1, xt) = (1 − t)d(x0, x1).
9
Proposition 2.3 ([3, Proposition 1.4]). A metric space M is an absolute 1-Lipschitz
retract if and only it is metrically convex and has the binary intersection property.
A metric space M is called a metric tree if it is complete, metrically convex, and
for any pair of vertices there is a unique continuous curve joining them in M. Given
a tree T in Graph Theory sense, one can construct the corresponding metric tree
by attaching between any two adjacent vertices mutually disjoint curves isometric
to the interval (0, 1).
Proposition 2.4 (Corollary of [7, Lemma 2.1]). A metric binary tree of any height
is a 1-absolute Lipschitz retract.
Observation 2.5. The binary intersection property and metric convexity are pre-
served under ℓ∞-sums.
Proof sketch. Binary intersection property: If balls {B((xi, yi), ri)}i in M1 ⊕∞ M2
are mutually intersecting, then the same happens for projected balls {B((xi, ri)}i
in M1 and {B((yi, ri)}i in M2. Thus each of the projected collections of balls has
a nonempty intersection. Let x and y be some points in the intersections. Using
the definition of the ℓ∞-sum we get that {B((xi, yi), ri)}i are Cartesian products of
{B((xi, ri)}i and {B((yi, ri)}i. Thus (x, y) is in the intersection of {B((xi, yi), ri)}i.
Metric convexity: Let (x0, y0) and (x1, y1) be two points in the ℓ∞-sum. Let {xt}
be a suitable family for M1 and {yt} be a suitable family for M2. Then {(xt, yt)} is
a suitable family for the ℓ∞-sum.
Corollary 2.6. Let Πn be the ℓ∞-sum of two metric binary trees of depth n each.
Then Πn is a 1-absolute Lipschitz retract.
As an immediate consequence we obtain the objective of Step 2.
Corollary 2.7. For i = 1, 2, 3, let ϕi,n : Pi,n → Wn be the maps defined in Step 1.
Since Wn ⊆ Tn ⊕∞ Tn ⊆ Πn, there exist Lipschitz extensions of ϕi,n to maps ¯ϕi,n :
Gn → Πn which have uniformly bounded Lipschitz constants.
2.4 The Final Step of the proof of Theorem 1.3
By [4] binary trees {Tn}∞
n=1 admit bilipschitz embeddings into any nonsuperreflexive
space X with uniformly bounded distortions (see also [25] and [23]), and, without loss
of generality, we may assume that these embeddings do not decrease any distances.
Using Mazur's method of constructing basic sequences (see [12, pp. 4-5]), as in
[22, Section 3], one can construct embeddings jn, which do not decrease distances
and have uniformly bounded distortions, of an ℓ∞-sum of six copies of Tn into any
non-superreflexive space X
jn : Tn ⊕∞ Tn ⊕∞ Tn ⊕∞ Tn ⊕∞ Tn ⊕∞ Tn → X.
10
Using [21, Lemma 3.3] or a direct argument, one can extend embeddings jn to
bilipschitz embeddings of Πn⊕∞Πn⊕∞Πn into X with uniformly bounded distortions
(recall that Πn denotes the ℓ∞-sum of two metric binary trees of depth n each).
However, for our purpose, it suffices to extend embeddings jn to Lipschitz maps
from Πn ⊕∞ Πn ⊕∞ Πn into X with uniformly bounded Lipschitz constants. This
can be done directly: we extend the maps from a binary tree to the metric edges
using 'linear interpolation': a point ut (0 < t < 1) on the edge joining vertices u0
and u1 in the metric tree corresponding to Tn with d(u0, ut) = t and d(ut, u1) = 1 − t
is mapped onto the corresponding convex combination of the images of u0 and u1.
We denote the obtained Lipschitz maps by
En : Πn ⊕∞ Πn ⊕∞ Πn → X.
By construction, the maps {En}n are non-contractive on Tn ⊕∞ Tn ⊕∞ Tn ⊕∞ Tn ⊕∞
Tn ⊕∞ Tn, and their Lipschitz constants are bounded by a universal constant C.
We are now ready to define the embeddings Fn : Gn → X. We put
Fn(x, k) def= En( ¯ϕ1,n(x, k), ¯ϕ2,n(x, k), ¯ϕ3,n(x, k)).
It follows from Step 1 and our construction that the Lipschitz constants of the
maps Fn are uniformly bounded . To prove that maps Fn are uniformly co-Lipschitz,
we observe that for any (x, k), (y, l) ∈ Gn, there exists i ∈ {1, 2, 3} such that both
(x, k) and (y, l) are in Pi,n. By Step 1 and since the restriction of En to Tn ⊕∞ Tn ⊕∞
Tn ⊕∞ Tn ⊕∞ Tn ⊕∞ Tn is noncontractive, we get
kFn(x, k) − Fn(y, l)kX ≥ d∞(ϕi,n(x, k), ϕi,n(y, l)) ≥
1
4
ρ((x, k), (y, l)),
which completes the proof of Theorem 1.3.
3 Proof of Corollary 1.7
For finite groups. As we mentioned in the Introduction, it is enough to prove that
for all n, the Cayley graph of Gn contains a subset that is bilipschitz equivalent with
an absolute constant to the binary tree of depth n/2. As mentioned earlier, this fact
was proved in [13] for trees of depth n/c, for some c > 0, but below we obtain it as
a direct consequence of our construction in Section 2.2.
Indeed, let Wn be the set defined in Section 2.2, and let U be the subset of Wn
consisting of all pairs (A, B) ∈ Wn so that A ≤ n
2 , and B is a sequence consisting
of n − A zeroes. The distance in (Wn, d∞) between any two elements of the set U is
bounded below by the tree distance of the corresponding As, and is bounded above
by twice the the tree distance of the corresponding As. Thus U is 2-equivalent to
the binary tree of depth n/2. Since, by Step 1, P1,n ⊆ Gn is bilipschitz equivalent
with Wn, and by [4], the proof for finite groups is complete.
11
For the infinite group, by Remark 1.6, we only need to show that the bilipschitz
embeddability of Z2 ≀ Z into a Banach space X implies that X is nonsuperreflexive.
This follows similarly as in the case of finite groups. By [28], the Cayley graph of
Z2 ≀ Z with the generating set S = {t, ta} coincides with the horocyclic product of
two infinite binary trees, i.e. infinite trees whose every vertex has degree 3, and the
identification is obtained by mapping every element (x, k) ∈ Z2 ≀ Z to an element
(A1, A2, k) of two infinite sequences of 0s and 1s with an (arbitrary) finite number
of nonzero terms, cf. also [27]. By Theorem 1.5, the metric on Z2 ≀ Z is bilipschitz
equivalent with the metric inherited from the ℓ1-sum of two tree metrics. Thus,
as in the finite case, we see that Z2 ≀ Z contains a bilipschitz copy of an infinite
rooted binary tree by taking the set of all elements of the form (A1, A2, k), where
k = 0, 1, 2, . . . , A1 is any sequence of 0s and 1s so that all terms with indices larger
than k are equal to 0, and all terms in the sequence A2 are equal to 0.
The fact that bilipschitz embeddability of a binary tree into X implies nonsuper-
reflexivity of X follows from [4], see also [2].
Acknowledgements: We would like to thank Florent Baudier for suggesting
the problem on lamplighter groups to us. The first named author was supported by
the National Science Foundation under Grant Number DMS–1700176. Both authors
thank the Fields Institute (Toronto) for partial funding to attend the Workshop on
Large Scale Geometry and Applications, where we started our work on lamplighter
groups.
References
[1] L. Bartholdi, W. Woess, Spectral computations on lamplighter groups and Diestel-
Leader graphs. J. Fourier Anal. Appl. 11 (2005), no. 2, 175–202.
[2] F. Baudier, Metrical characterization of super-reflexivity and linear type of Banach
spaces, Archiv Math., 89 (2007), no. 5, 419–429.
[3] Y. Benyamini, J. Lindenstrauss, Geometric nonlinear functional analysis. Vol. 1.
American Mathematical Society Colloquium Publications, 48. American Mathemat-
ical Society, Providence, RI, 2000.
[4] J. Bourgain, The metrical interpretation of superreflexivity in Banach spaces, Israel
J. Math., 56 (1986), no. 2, 222–230.
[5] C. Drutu, M. Kapovich, Geometric group theory. With an appendix by Bogdan Nica.
American Mathematical Society Colloquium Publications, 63. American Mathemat-
ical Society, Providence, RI, 2018.
[6] P. de la Harpe, Topics in geometric group theory. Chicago Lectures in Mathematics.
University of Chicago Press, Chicago, IL, 2000.
12
[7] W. B. Johnson, J. Lindenstrauss, D. Preiss, G. Schechtman, Lipschitz quotients from
metric trees and from Banach spaces containing ℓ1. J. Funct. Anal. 194 (2002), no.
2, 332–346.
[8] W. B. Johnson, G. Schechtman, Diamond graphs and super-reflexivity, J. Topol.
Anal., 1 (2009), no. 2, 177–189.
[9] B. Kloeckner, Yet another short proof of the Bourgain's distortion estimate for em-
bedding of trees into uniformly convex Banach spaces, Israel J. Math., 200 (2014),
no. 1, 419–422.
[10] U. Lang, Injective hulls of certain discrete metric spaces and groups. J. Topol. Anal.
5 (2013), no. 3, 297–331.
[11] S. L. Leung, S. Nelson, S. Ostrovska, M. I. Ostrovskii, Distortion of embeddings of
binary trees into diamond graphs. Proc. Amer. Math. Soc. 146 (2018), no. 2, 695–
704.
[12] J. Lindenstrauss, L. Tzafriri, Classical Banach spaces. I. Sequence spaces. Ergebnisse
der Mathematik und ihrer Grenzgebiete, Vol. 92. Springer-Verlag, Berlin-New York,
1977.
[13] R. Lyons, R. Pemantle, Y. Peres, Random walks on the lamplighter group. Ann.
Probab. 24 (1996), no. 4, 1993–2006.
[14] J. Matousek, On embedding trees into uniformly convex Banach spaces, Israel J.
Math., 114 (1999), 221–237.
[15] M. Mendel, A. Naor, Markov convexity and local rigidity of distorted metrics, J. Eur.
Math. Soc. (JEMS), 15 (2013), no. 1, 287–337; Conference version: Computational
geometry (SCG'08), 49–58, ACM, New York, 2008.
[16] A. Naor, An introduction to the Ribe program, Jpn. J. Math., 7 (2012), no. 2, 167–
233.
[17] A. Naor, Y. Peres, Embeddings of discrete groups and the speed of random walks.
Int. Math. Res. Not. IMRN 2008, Art. ID rnn 076, 34 pp.
[18] S. Ostrovska, M. I. Ostrovskii, Nonexistence of embeddings with uniformly bounded
distortions of Laakso graphs into diamond graphs, Discrete Math., 340 (2017), no. 2,
9–17.
[19] M. I. Ostrovskii, Embeddability of locally finite metric spaces into Banach spaces is
finitely determined, Proc. Amer. Math. Soc., 140 (2012), 2721–2730.
[20] M. I. Ostrovskii, Metric Embeddings: Bilipschitz and Coarse Embeddings into Banach
Spaces, de Gruyter Studies in Mathematics, 49. Walter de Gruyter & Co., Berlin,
2013.
13
[21] M. I. Ostrovskii, Different forms of metric characterizations of classes of Banach
spaces, Houston. J. Math., 39 (2013), no. 3, 889–906.
[22] M. I. Ostrovskii, Metric characterizations of superreflexivity in terms of word hyper-
bolic groups and finite graphs. Anal. Geom. Metr. Spaces 2 (2014), 154–168.
[23] M. I. Ostrovskii, Metric characterizations of some classes of Banach spaces, in: Har-
monic Analysis, Partial Differential Equations, Complex Analysis, Banach Spaces,
and Operator Theory, Celebrating Cora Sadosky's life, M. C. Pereyra, S. Marcan-
tognini, A. M. Stokolos, W. U. Romero (Eds.), Association for Women in Mathematics
Series, Vol. 4, pp. 307–347, Springer-Verlag, Berlin, 2016.
[24] M. I. Ostrovskii, B. Randrianantoanina, A new approach to low-distortion embeddings
of finite metric spaces into non-superreflexive Banach spaces. J. Funct. Anal. 273
(2017), no. 2, 598–651.
[25] G. Pisier, Martingales in Banach spaces, Cambridge Studies in Advanced Mathemat-
ics 155. Cambridge, Press Cambridge University Press, 2016.
[26] M. Stein, J. Taback, Metric properties of Diestel-Leader groups. Michigan Math. J.
62 (2013), no. 2, 365–386.
[27] J. Taback Lamplighter groups. in: M. Clay, D. Margalit (Editors), Office hours with
a geometric group theorist, 310–330, Princeton Univ. Press, Princeton, NJ, 2017.
[28] W. Woess, Lamplighters, Diestel-Leader graphs, random walks, and harmonic func-
tions. Combin. Probab. Comput. 14 (2005), no. 3, 415–433.
[29] W. Woess, What is a horocyclic product, and how is it related to lamplighters?
Internat. Math. Nachrichten of the Austrian Math. Soc. 224 (2013), 1–27; see also a
corrected version on the author's website.
Department of Mathematics and Computer Science, St. John's Uni-
versity, 8000 Utopia Parkway, Queens, NY 11439, USA
E-mail address: [email protected]
Department of Mathematics, Miami University, Oxford, OH 45056,
USA
E-mail address: [email protected]
14
|
1712.04338 | 1 | 1712 | 2017-12-09T15:32:15 | Liftings for ultra-modulation spaces, and one-parameter groups of Gevrey type pseudo-differential operators | [
"math.FA"
] | We deduce one-parameter group properties for pseudo-differential operators $\operatorname{Op} (a)$, where $a$ belongs to the class $\Gamma ^{(\omega _0)}_*$ of certain Gevrey symbols. We use this to show that there are pseudo-differential operators $\operatorname{Op} (a)$ and $\operatorname{Op} (b)$ which are inverses to each others, where $a\in \Gamma ^{(\omega _0)}_*$ and $b\in \Gamma ^{(1/\omega _0)}_*$.
We apply these results to deduce lifting property for modulation spaces and construct explicit isomorpisms between them. For each weight functions $\omega ,\omega _0$ moderated by GRS submultiplicative weights, we prove that the Toeplitz operator (or localization operator) $\operatorname{Tp} (\omega _0)$ is an isomorphism from $M^{p,q}_{(\omega )}$ onto $M^{p,q}_{(\omega /\omega _0)}$ for every $p,q \in (0,\infty ]$. | math.FA | math |
LIFTINGS FOR ULTRA-MODULATION SPACES, AND
ONE-PARAMETER GROUPS OF GEVREY TYPE
PSEUDO-DIFFERENTIAL OPERATORS
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Abstract. We deduce one-parameter group properties for pseudo-differential
operators Op(a), where a belongs to the class Γ(ω0)
of certain Gevrey symbols.
We use this to show that there are pseudo-differential operators Op(a) and
Op(b) which are inverses to each others, where a ∈ Γ(ω0)
and b ∈ Γ(1/ω0)
.
∗
∗
∗
We apply these results to deduce lifting property for modulation spaces
and construct explicit isomorpisms between them. For each weight functions
ω, ω0 moderated by GRS submultiplicative weights, we prove that the Toeplitz
operator (or localization operator) Tp(ω0) is an isomorphism from M p,q
(ω) onto
M p,q
(ω/ω0) for every p, q ∈ (0, ∞].
0. Introduction
The topological vector spaces V1 and V2 is said to possess lifting property if there
exists a "convenient" homeomorphisms (a lifting) between them. For example, for
any weight ω on Rd, p ∈ (0, ∞] and s ∈ R the convenient mappings f 7→ ω · f
and f 7→ (1 − ∆)s/2f are homeomorphic from the (weighted) Lebesgue space Lp
(ω)
and the Sobolev space H p
0 , with inverses f 7→ ω−1 · f
and f 7→ (1 − ∆)−s/2f , respectively. (Cf. [30] and Section 1 for notations.) Hence,
these spaces possess lifting properties.
s , respectively, into Lp = H p
It is often uncomplicated to deduce lifting properties between (quasi-)Banach
spaces of functions and distributions, if the definition of their norms only differs by
a multiplicative weight on the involved distributions, or on their Fourier transforms,
which is the case in the previous homeomorphisms. Here recall that multiplications
on the Fourier transform side are linked to questions on differentiation of the in-
volved elements. A more complicated situation appear when there are some kind
of interactions between multiplication and differentiation in the definition of the
involved vector spaces.
An example where such interactions occur concerns the extended family of
Sobolev spaces, introduced by Bony and Chemin in [3] (see also [34]). More pre-
cisely, let ω, ω0 be suitable weight functions and g a suitable Riemannian met-
ric, which are defined on the phase space W ≃ T ∗Rd ≃ R2d. Then Bony and
Chemin introduced in [3] the generalised Sobolev space H(ω, g) which fits the
Hormander-Weyl calculus well in the sense that H(1, g) = L2, and if a belongs
to the Hormander class S(ω0, g), then Weyl operator Opw(a) with symbol a is con-
tinuous from H(ω0ω, g) to H(ω, g). Moreover, they deduced group algebras, from
which it follows that to each such weight ω0, there exist symbols a and b such that
Opw(a) ◦ Opw(b) = Opw(b) ◦ Opw(a) = I,
a ∈ S(ω0, g), b ∈ S(1/ω0, g).
(0.1)
2010 Mathematics Subject Classification. Primary: 35S05, 47D06, 46B03, 42B35, 47B35,
46F05; Secondary: 46G15, 47L80.
Key words and phrases. Gelfand-Shilov spaces, Banach Function spaces, ultra-distributions,
confinement of symbols, comparable symbols.
1
2
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Here I is the identity operator on S ′. In particular, by the continuity properties
of Opw(a) it follows that H(ω0ω, g) and H(ω, g) possess lifting properties with the
homeomorphism Opw(a), and with Opw(b) as its inverse.
The existence of a and b in (0.1) is a consequence of solution properties of the
evolution equation
(∂ta)(t, · ) = (b + log ϑ)#a(t, · ),
a(0, · ) = a0 ∈ S(ω, g), ϑ ∈ S(ϑ, g),
(0.2)
which involve the Weyl product # and a fixed element b ∈ S(1, g). It is proved that
(0.2) has a unique solution a(t, · ) which belongs to S(ωϑt, g) (cf. [3, Theorem 6.4]
or [34, Theorem 2.6.15]). The existence of a and b in (0.1) will follow by choosing
ω = a0 = 1, t = 1 and ϑ = ω0.
If g is the constant euclidean metric on the phase space R2d, then S(ω0, g)
equals S(ω0)(R2d), the set of all smooth symbols a which satisfies ∂αa . ω0. We
notice that also for such simple choices of g, (0.1) above leads to lifting properties
that are not trivial.
In fact, let ω and ω0 be polynomially moderate weight on
the phase space, and let B be a suitable translation invariant BF-space. Then it
is observed in [26] that the continuity results for pseudo-differential operators on
modulation spaces in [48,50] imply that Opw(a) in (0.1) is continuous and bijective
from M (ω0ω, B) to M (ω, B) with continuous inverse Opw(b).
In particular, by
choosing B to be the mixed norm space Lp,q(R2d) of Lebesgue type, then M (ω, B)
is equal to the classical modulation space M p,q
(ω), introduced by Feichtinger in [12].
Consequently, Opw(a) above lifts M p,q
(ω0ω) into M p,q
(ω).
An important class of operators in quantum mechanics and time-frequency anal-
ysis concerns Toeplitz, or localisation operators. The main issue in [26, 27] is to
show that the Toeplitz operator Tp(ω0) lifts M p,q
(ω) for suitable ω0. The
assumptions on ω0 in [26] is that it should be polynomially moderate and satisfies
ω0 ∈ S(ω0). In [27], it is only assumed that ω0 is moderated by a GRS weight, but
instead it is here required that ω0 is radial in each phase shift, i. e. ω0 should satisfy
(ω0ω) into M p,q
ω0(x1, . . . , xd, ξ1, . . . , ξd) = ϑ(r1, . . . , rd),
rj = (xj , ξj),
for some weight ϑ.
The approaches in [26, 27] are also different. In [27], the lifting properties for
Tp(ω0) is reached by using the links between modulation spaces and Bargmann-
Foch spaces in combination of suitable estimates for a sort of generalised gamma-
functions. The approach in [26] relies on corresponding lifting properties for pseudo-
differential operators, as follows:
(1) Tp(ω0) = Opw(c) for some c ∈ S(ω0);
(2) by the definitions, it follows by straightforward computations that if ϑ =
ω
1
2
0 , then Tp(ω0) is a homeomorphism from M p,q
(ϑ) to M p,q
(ϑ);
(3) combining (0.1) with Wiener's lemma for (S(1), #) ensures that the inverse
of Tp(ω0) in (2) is a pseudo-differential operator Opw(b) with the symbol
b in S(1/ω0);
(4) by (1), (3) and duality,
T1 ≡ Opw(b) ◦ Tp(ω0) and T2 ≡ Tp(ω0) ◦ Opw(b)
are both the identity operator on S ′(Rd), since T1 is the identity operator
on M p,q
(ϑ), T2 is the identity operator on M p,q
(1/ϑ), and S ⊆ M p,q
(5) by (4), T1 = T2 = Opw(1) is the identity operator on each M p,q
(ϑ) ∩ M p,q
(ω). Since
(1/ϑ).
Tp(ω0) = Opw(c) : M p,q
(ω0ω) → M p,q
(ω)
and Opw(b) : M p,q
(ω) → M p,q
(ω0ω)
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO
3
are continuous (cf. [48, 50]) and inverses to each other, it follows that they
are homeomorphisms.
In the first part of the paper we deduce an analog of (0.1) for the Gevrey type
symbol classes Γ(ω0)
s
and Γ(ω0)
0,s of orders s ≥ 1, the set of all a ∈ C∞ such that
∂αa(X) . hαα!sω(X)
for some h > 0, respectively for every h > 0, considered in [4]. That is, in Section
3 we show that there exist symbols a and b such that
Opw(a) ◦ Opw(b) = Opw(b) ◦ Opw(a) = I,
a ∈ Γ(ω0
, b ∈ Γ(1/ω0
s
,
(0.3)
s
are replaced by Γ(ω0)
s
and Γ(1/ω0)
and similarly when Γ(ω0)
For general ω0 it is clear that Γ(ω0)
0,s , Γ(ω2)
weights ω1, ω2 and ω3 in Γ(ω1)
to PE,s(R2d), P 0
, respectively.
s ⊆ S(ω0). On the other hand, for the
and S(ω3) we always assume that they belong
E,s(R2d) and P(R2d), respectively. That is, they should satisfy
0,s ⊆ Γ(ω0)
0,s and Γ(1/ω0)
0,s
s
s
ω1(X + Y ) . ω1(X)er1Y
1
s , ω2(X + Y ) . ω2(X)er2Y
1
s ,
and ω3(X + Y ) . ω3(X)(1 + Y )N ,
for some r1 > 0 and N > 0, and every r2 > 0. Since it is clear that P ⊆
P 0
E,s ⊆ PE,s, it follows by straightforward computations that there are admissible
a1 ∈ Γ(ω1)
and S(ω0), and admissible
a2 ∈ Γ(ω2)
0,s which are not contained in any admissible Γ(ω0)
which are not contained in any S(ω0).
s
s
As in [3], (0.3) is obtained by proving that the evolution equation
(∂ta)(t, · ) = (b + log ϑ)#a(t, · ),
, ϑ ∈ Γ(ϑ)
,
analogous to (0.2), has a unique solution a(t, · ) which belongs to Γ(ωϑt)
ilarly when the Γ(ω)
Section 3.
-spaces are replaced by corresponding Γ(ω)
a(0, · ) = a0 ∈ Γ(ω)
s
s
s
s
(and sim-
0,s -spaces), given in
(0.4)
In Sections 4 and 5 we use the framework in [26] in combination with (0.3) to
extend the lifting properties in [26] in such ways that the involved weights are
allowed to belong to P 0
E,s or in PE,s instead of the smaller set P which is the
assumption in [26].
Our main result, which is similar to [26, Theorem 0.1], can be stated as follows.
Theorem 0.1. Let s ≥ 1, ω, ω0 ∈ P 0
Then the Toeplitz operator Tpφ(ω0) is an isomorphism from M p,q
E,s(R2d), p, q ∈ (0, ∞] and let φ ∈ Ss(Rd).
(ω)(R2d) onto M p,q
(ω/ω0)(R2d).
We note that, in contrast to [26, 27], our lifting properties also hold for modula-
tion spaces which may fail to be Banach spaces, since p and q in Theorem 0.1 are
allowed to be smaller than 1.
We will establish several related result. Firstly, the window function may be
chosen in certain modulation spaces that are much larger than the Gelfand-Shilov
space Ss. Secondly, the theorem holds for a more general family of modulation
spaces that includes the classical modulation spaces. Finally, we also establish
isomorphisms given by pseudo-differential operators rather than Toeplitz operators.
In contrast to [27], we do not impose in Theorem 0.1 and in its related results in
Section 5 that ω0 should be radial in each phase shift (cf. e. g. [27, Theorem 4.3]).
Summing up, our lifting results in Section 5 extend the lifting results in [26, 27].
The paper is organised as follows. In Section 1 we introduce some notation, and
discuss modulation spaces and Gelfand-Shilov spaces of functions and distributions,
4
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
and pseudo-differential calculus. In Section 2 we introduce and discuss basic prop-
erties for confinements of symbols in Γ(ω0)
0,s . These considerations are
related to the discussions in [3,34], but here adapted to symbols that possess Gevrey
regularity, e. g. when the symbols belong to Γ(ω0)
and in Γ(ω0)
s
s
or Γ(ω0)
0,s .
In contrast to the classical Hormander symbol classes Sr
1,0 and the SG-classes
SGm,µ
1,1 , techniques on asymptotic expansions are absent for symbols in the classes
Γ(ω0)
0,s , and might be absent for symbols in the general Hormander class
s
S(m, g). The approach with confinements is, roughly speaking, a sort of stand-in
of these absent asymptotic expansion techniques.
or in Γ(ω0)
In Section 3 we show that the (0.4) has a unique solution with the requested
properties, which leads to (0.3). In Sections 4 and 5 we use the results from Sec-
tion 3 to deduce lifting properties for modulation spaces under pseudo-differential
operators and Toeplitz operators with symbols in Γ(ω0)
or in Γ(ω0)
0,s
s
1. Preliminaries
In this section we recall some basic facts on modulation spaces, Gelfand-Shilov
spaces of functions and distributions and pseudo-differential calculus (cf. [11 -- 16,18,
21, 25, 30, 31, 34, 38, 42, 45, 47 -- 51]).
1.1. Weight functions. A weight on Rd is a positive function ω ∈ L∞
that 1/ω ∈ L∞
v-moderate, if
loc(Rd) such
loc(Rd). If ω and v are weights on Rd, then ω is called moderate or
ω(x + y) ≤ Cω(x)v(y),
(1.1)
for some constant C. The set of all moderate weights on Rd is denoted by PE(Rd).
The weight v on Rd is called submultiplicative, if it is even and (1.1) holds for ω = v
and C = 1. From now on, v always denote a submultiplicative weight if nothing
else is stated. In particular, if (1.1) holds and v is submultiplicative, then it follows
by straightforward computations that
x, y ∈ Rd,
C−1 ω(x)
v(y)
≤ ω(x + y) ≤ Cω(x)v(y),
(1.2)
v(x + y) ≤ v(x)v(y)
and v(x) = v(−x) ≥ 1,
x, y ∈ Rd.
If ω is a moderate weight on Rd, then there is a submultiplicative weight v on
Rd such that (1.1) and (1.2) hold (cf. [47,48,52]). Moreover if v is submultiplicative
on Rd, then
(1.3)
for some constant r > 0 (cf. [24]). Here and in what follows we write A(θ) . B(θ),
θ ∈ Ω, if there is a constant c > 0 such that A(θ) ≤ cB(θ) for all θ ∈ Ω.
In
particular, if ω is moderate, then
1 . v(x) . erx
ω(x + y) . ω(x)ery
and e−rx ≤ ω(x) . erx,
x, y ∈ Rd
(1.4)
for some r > 0.
Next we introduce suitable subclasses of PE.
Definition 1.1. Let s > 0. The set PE,s(Rd) (P 0
PE(Rd) such that
E,s(Rd)) consists of all ω ∈
ω(x + y) . ω(x)ery
1
s ,
x, y ∈ Rd;
(1.5)
holds for some (every) r > 0.
By (1.4) it follows that P 0
E,s1 = PE,s2 = PE when s1 < 1 and s2 ≤ 1. For
convenience we set P 0
E(Rd) = P 0
E,1(Rd).
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO
5
1.2. Gelfand-Shilov spaces. We let F be the Fourier transform given by
(F f )(ξ) = bf (ξ) ≡ (2π)−d/2ZRd
f (x)e−ihx,ξi dx
when f ∈ L1(Rd). Here h · , · i denotes the usual scalar product on Rd.
Definition 1.2. The Gelfand-Shilov space Sσ
Beurling type), σ > 0, s > 0, consists of all f ∈ S (Rd) such that
s (Rd) of Roumieu type (Σσ
s (Rd) of
f (x) . e−rx
1
s
and
x, ξ ∈ Rd
(1.6)
1
(F f )(ξ) . e−rξ
σ ,
s and Σs = Σs
s.
for some r > 0 (for all r > 0). We set Ss = Ss
The classes Sσ
s (Rd) and related generalizations were widely studied, and used in
the applications to partial differential equations, see for example [2, 6, 9, 28, 35, 38].
We recall the following characterisations of Sσ
s (Rd).
Proposition 1.3. Let s, σ > 0, p ∈ [1, ∞] and let f ∈ S (Rd). Then the following
conditions are equivalent:
(1) f ∈ Sσ
(2) for some (every) h > 0 it holds
s (Rd) (f ∈ Σσ
s (Rd));
kxαf kLp . hαα!s
and
(3) for some (every) h > 0 it holds
kξβbf kLp . hββ!σ, α, β ∈ Nd;
kxαf kLp . hαα!s
and
k∂βf kLp . hββ!σ, α, β ∈ Nd;
(4) for some (every) h > 0 it holds
kxα∂βf (x)kLp . hα+βα!s β!σ, α, β ∈ Nd;
(5) for some (every) h, r > 0 it holds
ker ·
1
s ∂αf kLp . hα(α!)σ α ∈ Nd.
Remark 1.4. Let h, s, s0, σ, σ0 ∈ R+ be such that s + σ > 1 and s0 + σ0 ≥ 1, and
let Sσ
s,h(Rd) be the set of all f ∈ C∞(Rd) such that
kf kS σ
s,h
≡ sup
xβ∂αf (x)
hα+βα!σβ!s
is finite. Here the supremum is taken over all α, β ∈ Nd and x ∈ Rd.
Obviously Sσ
s,h(Rd) is a Banach space which increases as h, s and σ increase,
and is contained in S (Rd), the set of Schwartz functions on Rd. Furthermore
Sσ
s,h(Rd) and [h>0
Sσ0
s0,h(Rd)
are dense in S (Rd). Hence, the dual (Sσ
which contains S ′(Rd).
s,h)′(Rd) of Sσ
s,h(Rd) is a Banach space
The spaces Sσ
s (Rd) and Σσ
s (Rd) are the inductive and projective limits, respec-
tively, of Sσ
s,h(Rd) with respect to h. This implies that
s,h(Rd) and Σσ
Sσ
s,h(Rd),
Sσ
(1.7)
Sσ
s (Rd) = [h>0
s (Rd) = \h>0
and that the topology for Sσ
sion map from Sσ
(2) -- (5) in Proposition 1.3 induce the same topology for Sσ
s (Rd) is the strongest possible one such that each inclu-
s (Rd) is continuous. Moreover, any of the conditions
s,h(Rd) to Sσ
s (Rd) and Σσ
s (Rd).
6
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
The Gelfand-Shilov distribution spaces (Sσ
s )′(Rd) and (Σσ
s )′(Rd) are the projec-
tive and inductive limit respectively of (Sσ
s,h)′(Rd). Hence
(Sσ
s )′(Rd) = \h>0
s )′(Rd) = [h>0
(Sσ
s,h)′(Rd) and (Σσ
(Sσ
s,h)′(Rd).
(1.7)′
We have that (Sσ
(see [37]).
s )′ and (Σσ
s )′ are the topological duals of Sσ
s and Σσ
s , respectively
Remark 1.5. Let s, σ > 0. Then Σσ
h > 0. Moreover, Sσ
only if s + σ ≥ 1 and (s, σ) 6= ( 1
2 , 1
s (Rd) 6= {0} if and only if s + σ ≥ 1, and Σσ
s (Rd) is a Fr´echet space with seminorms k · kS σ
,
s (Rd) 6= {0} if and
s,h
Σσ
s (Rd) ⊆ Sσ
If in addition (s, σ) 6= ( 1
s+ε (Rd) ⊆ S (Rd) ⊆ S ′(Rd) ⊆ (Σσ+ε
s (Rd) ⊆ Σσ+ε
2 , 1
2 ), then
2 ). Moreover, if ε > 0 and s + σ ≥ 1, then
s+ε )′(Rd) ⊆ (Sσ
s )′(Rd).
(Sσ
s )′(Rd) ⊆ (Σσ
s )′(Rd).
The Gelfand-Shilov spaces are invariant or possess convenient mapping prop-
erties under several basic transformations. For example they are invariant under
translations, dilations, and under (partial) Fourier transformations.
The Fourier transform F on S (Rd) extends uniquely to homeomorphisms on
s(Rd), and restricts to homeomorphisms on Ss(Rd) and
S ′(Rd), S′
Σs(Rd), and to a unitary operator on L2(Rd).
s(Rd) and Σ′
We also recall some mapping properties of Gelfand-Shilov spaces under short-
time Fourier transforms. Let φ ∈ S (Rd) be fixed. For every f ∈ S ′(Rd), the
short-time Fourier transform Vφf is the distribution on R2d defined by the formula
(Vφf )(x, ξ) = F (f φ( · − x))(ξ) = (f, φ( · − x)eih · ,ξi).
(1.8)
We recall that if T (f, φ) ≡ Vφf when f, φ ∈ S1/2(Rd), then T is uniquely extendable
to sequentially continuous mappings
T : S′
s(Rd) × Ss(Rd) → S′
T : S′
s(Rd) × S′
s(Rd) → S′
s(R2d),
s(R2d)\ C∞(R2d),
and similarly when Ss and S′
S ′, respectively (cf. [7, 52]). We also note that Vφf takes the form
s are replaced by Σs and Σ′
s, respectively, or by S and
Vφf (x, ξ) = (2π)−d/2ZRd
f (y)φ(y − x)e−ihy,ξi dy
(1.8)′
(ω)(Rd) for some ω ∈ PE(Rd), φ ∈ Σ1(Rd) and p ≥ 1. Here Lp
when f ∈ Lp
when p ∈ (0, ∞] and ω ∈ PE(Rd), is the set of all f ∈ Lp
Lp(Rd).
(ω)(Rd),
loc(Rd) such that f · ω ∈
1.3. Suitable function classes with Gelfand-Shilov regularity. The next re-
sult shows that for any ω ∈ PE(Rd) one can find an equivalent weight ω0 which
satisfies similar regularity estimates as functions in Gelfand-Shilov spaces.
Proposition 1.6. Let ω ∈ PE(Rd) and s > 0. Then there is an ω0 ∈ PE(Rd) ∩
C∞(Rd) such that the following is true:
(1) ω0 ≍ ω;
(2) ∂αω0(x) . ω0(x)hαα!s ≍ ω(x)hαα!s for every h > 0.
where the last inequality follows from (1.4) and the fact that φ is bounded by a
super exponential function. This gives the first part of (2).
The equivalences in (1) follows in the same way as in [52]. More precisely, by
(1.4) we have
In the same way, (1.4) gives
ω0(x) =Z ω(y)φ(x − y) dy =Z ω(x + (y − x))φ(x − y) dy
. ω(x)Z ecx−yφ(x − y) dy ≍ ω(x).
ω0(x) =Z ω(y)φ(x − y) dy =Z ω(x + (y − x))φ(x − y) dy
& ω(x)Z e−cx−yφ(x − y) dy ≍ ω(x),
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO
7
Proof. We may assume that s < 1. It suffices to prove that (2) should hold for
some h > 0. Let φ0 ∈ Σs
1−s(Rd), giving
that
1−s(Rd) \ 0, and let φ = φ02. Then φ ∈ Σs
∂αφ(x) . hαe−cx
for every h > 0 and c > 0. Now let ω0 = ω ∗ φ.
1
1−s α!s,
We have
∂αω0(x) =(cid:12)(cid:12)(cid:12)(cid:12)Z ω(y)(∂αφ)(x − y) dy(cid:12)(cid:12)(cid:12)(cid:12)
. hαα!sZ ω(y)e−cx−y
. hαα!sZ ω(x + (y − x))e−cx−y
. hαα!sω(x)Z e− c
1−s dy
1
1
1
1−s dy
2 x−y
1−s dy ≍ hαα!sω(x),
and (1) as well as the second part of (2) follow.
(cid:3)
A weight ω0 which satisfies Proposition 1.6 (2) is called elliptic or s-elliptic.
Definition 1.7. Let s ≥ 0 and ω ∈ PE(Rd). The class Γ(ω)
consists of all f ∈ C∞(Rd) such that
s
(Rd) (Γ(ω)
0,s (Rd))
Dαf (x) . ω(x) hαα!s,
x ∈ Rd,
for some h > 0 (for every h > 0).
Evidently, by Proposition 1.6 it follows that the family of symbol classes in
Definition 1.7 is not reduced when the assumption ω ∈ PE(R2d) is replaced by the
more restrictive assumption ω ∈ PE,s(R2d) or by ω ∈ P 0
E,s(R2d).
By similar arguments as in the proof of Proposition 1.6 we get the following
analog of Proposition 2.3.16 in [33]. The details are left for the reader.
Proposition 1.8. Let s > 1/2, ω ∈ PE(R2d), and φ ∈ Σs(R2d). Then ω ∗ φ
belongs to Γ(ω)
0,s .
The following definition is motivated by Lemma 2.6.13 in [33].
Definition 1.9. Let s ≥ 1, ω ∈ PE(Rd) and ϑ0 = 1 + log ω. Then c is called
comparable to ω with respect to s ≥ 1 if
8
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
(1) kc − log ωkL∞ < ∞
(2) c ∈ Γ(ϑ0)
(Rd) and ∂αc ∈ Γ(1)
s
s (Rd), when α = 1.
Proposition 1.10. Let ω, v ∈ PE(Rd) be such that v is submultiplicative and ω
is v-moderate. Also let
v1(x) ≡ 1 + log v(x)
and ω1(x) ≡ 1 + log ω(x).
Then v1 is submultiplicative and ω1 is v1-moderate, satisfying (1.1) with 1+log C ≥
1 in place of C.
Proof. By (1.2) we get
v1(x + y) = 1 + log v(x + y) ≤ 1 + log v(x) + log v(y)
≤ (1 + log v(x)) (1 + log v(y)) = v1(x) v1(y),
and
ω1(x + y) = 1 + log ω(x + y)
≤ 1 + log C + log ω(x) + log v(y)
≤ (1 + log C)(1 + log ω(x)) (1 + log v(y))
≤ (1 + log C) ω1(x) v1(y),
as claimed.
(cid:3)
Lemma 1.11. Let s ≥ 1, ω ∈ PE(Rd) and ϑ0 = 1 + log ω. Then the following
is true:
(1) there exists an elliptic weight ω0 ∈ PE(Rd) ∩ Γ(ω)
s
(Rd) such that ω ≍ ω0
and 1 + log ω0 ∈ PE(Rd) ∩ Γ(ϑ0)
s
(Rd);
(2) there exists an element c which is comparable with ω0.
Proof. If ω1(x) is equal to ω(x) when ω(x) ≥ e or ω(x) ≤ e−1, and 3 otherwise,
then ω1 is equivalent to ω. The result now follows from Proposition 1.6 and its
proof, with ω1 in place of ω.
(cid:3)
1.4. Modulation spaces. Let φ ∈ Σ1(Rd) \ 0, p, q ∈ (0, ∞] and ω ∈ PE(R2d) be
fixed. Then the modulation space M p,q
(ω)(Rd) consists of all f ∈ Σ′
1(Rd) such that
kf kM p,q
(ω)
≡(cid:16)Z (cid:16)Z Vφf (x, ξ)ω(x, ξ)p dx(cid:17)q/p
dξ(cid:17)1/q
< ∞
(1.9)
(with the obvious modifications when p = ∞ and/or q = ∞). We set M p
and if ω = 1, then we set M p,q = M p,q
(ω) and M p = M p
(ω).
(ω) = M p,p
(ω),
The following proposition is a consequence of well-known facts in [12, 20, 21, 54].
Here and in what follows, we let p′ denotes the conjugate exponent of p, i. e.
∞
p
p − 1
when p ∈ (0, 1]
when p ∈ (1, ∞)
1
when p = ∞ .
p′ =
Proposition 1.12. Let p, q, pj, qj, r ∈ (0, ∞] be such that r ≤ min(1, p, q), j = 1, 2,
let ω, ω1, ω2, v ∈ PE(R2d) be such that ω is v-moderate, φ ∈ M r
(v)(Rd) \ 0, and let
f ∈ Σ′
1(Rd). Then the following is true:
(1) f ∈ M p,q
(ω)(Rd) if and only if (1.9) holds, i. e. M p,q
of the choice of φ. Moreover, M p,q
(1.9), and different choices of φ give rise to equivalent norms;
(ω)(Rd) is independent
(ω) is a Banach space under the norm in
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO
9
(2) if p1 ≤ p2, q1 ≤ q2 and ω2 ≤ Cω1 for some constant C, then
Σ1(Rd) ⊆M p1,q1
(ω1) (Rd) ⊆M p2,q2
(ω2) (Rd) ⊆ Σ′
1(Rd).
Proposition 1.12 (1) allows us to be rather vague about to the choice of φ ∈
(v) \ 0 in (1.9). For example, if C > 0 is a constant and Ω is a subset of Σ′
M r
1, then
≤ C for every a ∈ Ω, means that the inequality holds for some choice of
kakM p,q
φ ∈ M r
(v) \ 0, a
similar inequality is true although C may have to be replaced by a larger constant,
if necessary.
(v) \ 0 and every a ∈ Ω. Evidently, for any other choice of φ ∈ M r
(ω)
Let s, t ∈ R. Then the weights
(x, ξ) 7→ hξishxit
and (x, ξ) 7→ h(x, ξ)is,
(1.10)
are common in the applications. It follows that they belong to P(R2d) for every
If ω ∈ P(R2d), then ω is moderated by any of the weights in (1.10)
s, t ∈ R.
provided s and t are chosen large enough.
Remark 1.13. For modulation spaces of the form M p,q
(ω) with fixed p, q ∈ [1, ∞]
the norm equivalence in Proposition 1.12(1) can be extended to a larger class of
windows. In fact, assume that ω, v ∈ PE(R2d) with ω being v-moderate and
x, ξ ∈ Rd,
1 ≤ r ≤ min(p, p′, q, q′) .
Let φ ∈ M r
to M p,q
φ ∈ M r
in [51].)
(ω)(Rd), if and only if Vφf ∈ Lp,q
(v)(Rd) \ {0} in kVφf kLp,q
(ω)
(v)(Rd) \ {0}. Then a Gelfand-Shilov distribution f ∈ Σ′
1(Rd) belongs
(ω)(R2d). Furthermore, different choices of
give rise to equivalent norms. (Cf. Theorem 3.1
1.5. A broader family of modulation spaces. As announced in the introduc-
tion we consider in Section 2 mapping properties for pseudo-differential operators
when acting on a broader class of modulation spaces, which are defined by impos-
ing certain types of translation invariant solid BF-space norms on the short-time
Fourier transforms. (Cf. [11 -- 16, 18].)
Definition 1.14. Let B ⊆ Lr
loc(Rd) be a quasi-Banach of order r ∈ (0, 1], and
let v ∈ PE(Rd). Then B is called a translation invariant Quasi-Banach Function
space on Rd, or invariant QBF space on Rd, if there is a constant C such that the
following conditions are fulfilled:
(1) if x ∈ Rd and f ∈ B, then f ( · − x) ∈ B, and
kf ( · − x)kB ≤ Cv(x)kf kB;
(1.11)
(2) if f, g ∈ Lr
loc(Rd) satisfy g ∈ B and f ≤ g, then f ∈ B and
kf kB ≤ CkgkB.
If v belongs to PE,s(Rd) (P 0
E,s(Rd)) , then B in Definition 1.14 is called an
invariant BF-space of Roumieu type (Beurling type) of order s.
We notice that the quasi-norm k · kB in Definition 1.14 should satisfy
kf + gkB ≤ 2
1
r −1(kf kB + kgkB)
f, g ∈ B.
(1.12)
By Aoki and Rolewi´c in [1, 39] it follows that there is an equivalent quasi-norm to
the previous one which additionally satisfies
kf + gkr
B ≤ kf kr
B + kgkr
B
f, g ∈ B.
(1.13)
From now on we suppose that the quasi-norm of B has been chosen such that both
(1.12) and (1.13) hold true.
10
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
It follows from (2) in Definition 1.14 that if f ∈ B and h ∈ L∞, then f · h ∈ B,
and
kf · hkB ≤ Ckf kBkhkL∞.
(1.14)
If r = 1, then B in Definition 1.14 is a Banach space, and the condition (2) means
that a translation invariant QBF-space is a solid BF-space in the sense of (A.3)
in [15]. The space B in Definition 1.14 is called an invariant BF-space (with
respect to v) if r = 1, and Minkowski's inequality holds true, i. e.
kf ∗ ϕkB ≤ Ckf kBkϕkL1
(v)
,
f ∈ B, ϕ ∈ Σ1(Rd)
(1.15)
for some constant C which is independent of f ∈ B and ϕ ∈ Σ1(Rd).
Example 1.15. Assume that p, q ∈ [1, ∞], and let Lp,q
f ∈ L1
loc(R2d) such that
1 (R2d) be the set of all
if finite. Also let Lp,q
2 (R2d) be the set of all f ∈ L1
loc(R2d) such that
kf kLp,q
1
kf kLp,q
2
≡(cid:16)Z (cid:16)Z f (x, ξ)p dx(cid:17)q/p
≡(cid:16)Z (cid:16)Z f (x, ξ)q dξ(cid:17)p/q
dξ(cid:17)1/q
dx(cid:17)1/p
is finite. Then it follows that Lp,q
respect to v = 1.
1 and Lp,q
2 are translation invariant BF-spaces with
For translation invariant BF-spaces we make the following observation.
Proposition 1.16. Assume that v ∈ PE(Rd), and that B is an invariant BF-
space with respect to v such that (1.15) holds true. Then the convolution mapping
0 (Rd) × B to B extends uniquely to a continuous mapping
(ϕ, f ) 7→ ϕ ∗ f from C∞
(v)(Rd).
from L1
(v)(Rd) × B to B, and (1.15) holds true for any f ∈ B and ϕ ∈ L1
The result is a straightforward consequence of the fact that C∞
(v).
0
The quasi-Banach space B above is usually a mixed quasi-normed Lebesgue
space, given as follows. Let E be a non-degenerate parallelepiped in Rd which is
spanned by the ordered basis κ(E) = {e1, . . . , ed}. That is,
is dense in L1
E = { x1e1 + · · · + xded ; (x1, . . . , xd) ∈ Rd, 0 ≤ xk ≤ 1, k = 1, . . . , d }.
The corresponding lattice, dual parallelepiped and dual lattice are given by
ΛE = { j1e1 + · · · + jded ; (j1, . . . , jd) ∈ Zd },
E′ = { ξ1e′
1 + · · · + ξde′
d ; (ξ1, . . . , ξd) ∈ Rd, 0 ≤ ξk ≤ 1, k = 1, . . . , d },
and
Λ′
E = ΛE ′ = { ι1e′
1 + · · · + ιde′
d ; (ι1, . . . , ιd) ∈ Zd },
1, . . . , e′
respectively, where the ordered basis κ(E′) = {e′
d} of E′ satisfies
hej, e′
ki = 2πδjk
for every j, k = 1, . . . , d.
Note here that the Fourier analysis with respect to general biorthogonal bases has
recently been developed in [41].
The basis e′
1, . . . , e′
d is called the dual basis of e1, . . . , ed. We observe that there
d are the images of the standard
is a matrix TE such that e1, . . . , ed and e′
basis under TE and TE ′ = 2π(T −1
1, . . . , e′
E )t, respectively.
In the following we let
max q = max(q1, . . . , qd)
and min q = min(q1, . . . , qd)
when q = (q1, . . . , qd) ∈ (0, ∞]d, and χΩ be the characteristic function of Ω.
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 11
Definition 1.17. Let E be a non-degenerate parallelepiped in Rd spanned by the
ordered set κ(E) ≡ {e1, . . . , ed} in Rd, p = (p1, . . . , pd) ∈ (0, ∞]d and r = min(1, p).
If f ∈ Lr
loc(Rd), then
kf kLp
κ(E)
≡ kgd−1kLpd (R)
where gk(zk), zk ∈ Rd−k, k = 0, . . . , d − 1, are inductively defined as
g0(x1, . . . , xd) ≡ f (x1e1 + · · · + xded),
and
gk(zk) ≡ kgk−1( · , zk)kLpk (R),
k = 1, . . . , d − 1.
loc(Rd) such that kf kLp
κ(E)(Rd) consists of all f ∈ Lr
The space Lp
is called E-split Lebesgue space (with respect to p and κ(E)).
Definition 1.18. Let E0 ⊆ Rd be a non-degenerate parallelepiped with dual
parallelepiped E′
0, and E ⊆ R2d be a parallelepiped spanned by the ordered set
κ(E) ≡ {e1, . . . , e2d}. Then E is called a phase-shift split parallelepiped (with re-
spect to E0) if E is non-degenerate and d of the vectors in {e1, . . . , e2d} span E0
and the other d vectors is the corresponding dual basis which span E′
0.
is finite, and
κ(E)
Next we consider the extended class of modulation spaces which we are interested
in.
Definition 1.19. Assume that B is a translation invariant QBF-space on R2d,
ω ∈ PE(R2d), and that φ ∈ Σ1(Rd) \ 0. Then the set M(ω) = M (ω, B) consists of
all f ∈ Σ′
1(Rd) such that
kf kM(ω) = kf kM(ω,B) ≡ kVφf ωkB
is finite.
Obviously, we have M p,q
(ω)(Rd) = M (ω, B) when B = Lp,q
1 (R2d) (cf. Example
1.15). It follows that many properties which are valid for the classical modulation
spaces also hold for the spaces of the form M (ω, B). For example we have the
following proposition, which shows that the definition of M (ω, B) is independent
of the choice of φ when B is a Banach space. We omit the proof since it follows
by similar arguments as in the proof of Proposition 11.3.2 in [21]. (See also [36] for
topological aspects of M (ω, B).)
Proposition 1.20. Let B be an invariant BF-space with respect to v0 ∈ PE(R2d)
for j = 1, 2. Also let ω, v ∈ PE(R2d) be such that ω is v-moderate, M (ω, B) is
the same as in Definition 1.19, and let φ ∈ M 1
1(Rd). Then
M (ω, B) is a Banach space, and f ∈ M (ω, B) if and only if Vφf ω ∈ B, and
different choices of φ gives rise to equivalent norms in M (ω, B).
(v0v)(Rd) \ 0 and f ∈ Σ′
We refer to [11 -- 16, 18, 20, 21, 40, 54] for more facts about modulation spaces.
1.6. Pseudo-differential operators. We use the notation M(d, Ω) for the set of
d×d-matrices with entries in the set Ω. Let s ≥ 1/2, a ∈ Ss(R2d), and A ∈ M(d, R)
be fixed. Then, the pseudo-differential operator OpA(a) is the linear and continuous
operator on Ss(Rd) given by
(OpA(a)u)(x) = (2π)−dZZ a(x − A(x − y), ξ) f (y) eihx−y,ξi dydξ
(1.16)
when f ∈ Ss(Rd). For general a ∈ S′
OpA(a) is defined as the continuous operator from Ss(Rd) to S′
tribution kernel given by
s(R2d), the pseudo-differential operator
s(Rd) with dis-
Ka,A(x, y) = (2π)−d/2(F −1
2 a)(x − A(x − y), x − y).
(1.17)
12
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Here F2F is the partial Fourier transform of F (x, y) ∈ S′
y variable. This definition makes sense, since the mappings
s(R2d) with respect to the
F2
and F (x, y) 7→ F (x − A(x − y), y − x)
(1.18)
s(R2d). In particular, the map a 7→ Ka,A is a homeo-
are homeomorphisms on S′
morphism on S′
s(R2d).
Remark 1.21. For any K ∈ S′
ping from Ss(Rd1) to S′
s(Rd2), defined by the formula
s(Rd1+d2), let TK be the linear and continuous map-
(TKf, g)L2(Rd2 ) = (K, g ⊗ f )L2(Rd1+d2 ).
(1.19)
It is well-known that if t ∈ R, then it follows from e. g. [5, 32] that the Schwartz
kernel theorem also holds in the context of Gelfans-Shilov spaces. That is, the
mappings K 7→ TK and a 7→ Opt(a) are bijective from S′
s(R2d) to the set of linear
s(Rd). Similar facts hold true if Ss and
and continuous mappings from Ss(Rd) to S′
s are replaced by Σs and Σ′
S′
As a consequence of Remark 1.21 it follows that for each a1 ∈ S′
s, respectively (or by S and S ′, respectively).
s(R2d) and
s(R2d) such that OpA1(a1) = OpA2 (a2).
A1, A2 ∈ M(d, R), there is a unique a2 ∈ S′
The relation between a1 and a2 is given by
OpA1(a1) = OpA2 (a2)
⇔
a2(x, ξ) = eih(A1−A2)Dξ,Dxia1(x, ξ).
(1.20)
(Cf. [30].) Note here that the right-hand side makes sense, since it is equivalent to
S′
s when A ∈ M(d, R).
ba2(ξ, x) = ei(A1−A2)hx,ξiba1(ξ, x), and that the map a 7→ eihAx,ξia is continuous on
s(R2d) be fixed. Then a is called a rank-one element
with respect to A, if the corresponding pseudo-differential operator is of rank-one,
i. e.
Let A ∈ M(d, R) and a ∈ S′
for some f1, f2 ∈ S′
fulfilled, if and only if a = (2π)d/2W A
defined by the formula
f1,f2
OpA(a)f = (f, f2)f1,
(1.21)
s(Rd). By straightforward computations it follows that (1.21) is
it the A-Wigner distribution
, where W A
f ∈ Ss(Rd),
f1,f2
W A
f1,f2(x, ξ) ≡ F (f1(x + A · )f2(x − (I − A) · ))(ξ),
(1.22)
which takes the form
W A
f1,f2(x, ξ) = (2π)−d/2Z f1(x + Ay)f2(x − (I − A)y)e−ihy,ξi dy,
when f1, f2 ∈ Ss(Rd). Here I ∈ M(d, R) is the identity matrix. By combining
these facts with (1.20) it follows that
W A2
f1,f2
= eih(A1−A2)Dξ,DxiW A1
f1,f2
,
(1.23)
for each f1, f2 ∈ S′
important, we set W A
)Wigner distribution of f1 and f2.
s(Rd) and A1, A2 ∈ M(d, R). Since the Weyl case is particularly
2 I, i. e. Wf1,f2 is the usual (cross-
f1,f2 = Wf1,f2 when A = 1
For future references we note the link
(OpA(a)f, g)L2(Rd) = (2π)−d/2(a, W A
g,f )L2(R2d),
a ∈ S′
s(R2d) and f, g ∈ Ss(Rd)
(1.24)
between pseudo-differential operators and Wigner distributions, which follows by
straightforward computations (see also e. g. [55]).
Next we discuss the Weyl product, the twisted convolution and related objects.
Let s ≥ 1/2 and let a, b ∈ S′
s(R2d). Then the Weyl product a#b between a and b is
the function or distribution which fulfills Opw(a#b) = Opw(a) ◦ Opw(b), provided
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 13
the right-hand side makes sense as a continuous operator from Ss(Rd) to S′
More generally, if A ∈ M(d, R), then the product #A is defined by the formula
s(Rd).
OpA(a#Ab) = OpA(a) ◦ OpA(b),
(1.25)
provided the right-hand side makes sense as a continuous operator from Ss(Rd) to
S′
s(Rd), in which case a and b are called suitable or admissible. We also use the
notation # instead of #A when A = 1
2 I (i. e. in the Weyl case).
The Weyl product can also, in a convenient way, be expressed in terms of the
symplectic Fourier transform and the twisted convolution. More precisely, let s ≥
1/2. Then the symplectic Fourier transform for a ∈ Ss(R2d) is defined by the
formula
(Fσa)(X) = π−dZ a(Y )e2iσ(X,Y ) dY,
where σ is the symplectic form given by
σ(X, Y ) = hy, ξi − hx, ηi,
X = (x, ξ) ∈ R2d, Y = (y, η) ∈ R2d.
We note that Fσ = T ◦(F ⊗(F −1)), when (T a)(x, ξ) = a(ξ, x). In particular, Fσ
is continuous on Ss(R2d), and extends uniquely to a homeomorphism on S′
s(R2d),
and to a unitary map on L2(R2d), since similar facts hold for F . Furthermore, F 2
σ
is the identity operator.
Let s ≥ 1/2 and a, b ∈ Ss(R2d). Then the twisted convolution of a and b is
defined by the formula
(a ∗σ b)(X) =(cid:18) 2
π(cid:19) d
2Z a(X − Y )b(Y )e2iσ(X,Y ) dY.
The definition of ∗σ extends in different ways. For example, it extends to a con-
tinuous multiplication on Lp(R2d) when p ∈ [1, 2], and to a continuous map from
s(R2d), then a#b makes sense if and only
S′
s(R2d) × Ss(R2d) to S′
s(R2d). If a, b ∈ S′
(1.26)
(1.27)
(1.28)
if a ∗σbb makes sense, and then
a#b = (2π)− d
2 a ∗σ (Fσb).
We also remark that for the twisted convolution we have
Fσ(a ∗σ b) = (Fσa) ∗σ b = a ∗σ (Fσb),
where a(X) = a(−X) (cf. [46, 51, 53]). A combination of (1.27) and (1.28) gives
(1.29)
We now define the subspace of symbols in S ′(R2d) which give rise to L2(Rd)-
2 (Fσa) ∗σ (Fσb).
Fσ(a#b) = (2π)− d
bounded pseudo-differential operators, which will be useful in the sequel.
Definition 1.22. The set sw
is linear and continuous on L2(Rd), and we set
∞(R2d) consists of all a ∈ S ′(R2d) such that Opw(a)
kaksw
∞ ≡ k Opw(a)kL2(Rd)→L2(Rd).
Remark 1.23. By Remark 1.21 it follows that the map a 7→ Opw(a) is an isometric
bijection from sw
∞(R2d) to the set of linear continuous operators on L2(Rd).
s and s > 1
2 are replaced by Σs, Σ′
Remark 1.24. We remark that the relations in this subsection hold true after Ss,
s and s ≥ 1
S′
Next we recall some algebraic properties and characterisations of Γ(ω)
Γ(ω)
0,s (R2d), and begin with the following. We refer to [4] for its proof.
Proposition 1.25. Let s ≥ 1, ωj ∈ P 0
let ω0,r(X, Y ) = ω0(X)e−rY
E,s(R2d), Aj ∈ M(d, R) for j = 1, 2, and
s when r > 0. Then the following is true:
2 respectively, in each place.
(R2d) and
s
1
14
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
(1) If a1, a2 ∈ Σ′
s(R2d) satisfy OpA1(a1) = OpA2 (a2), then a1 ∈ Γ(ω0)
s
(R2d) if
(R2d).
(2) Γ(ω1)
(3) Γ(ω0)
and only if a2 ∈ Γ(ω0)
s ⊆ Γ(ω1ω2)
s #Γ(ω2)
s
.
s = [r>0
s
M ∞,1
(1/ω0,r) = [r≥0
M∞,1
(1/ω0,r).
Proposition 1.26. Let s ≥ 1, ωj ∈ PE,s(R2d), Aj ∈ M(d, R) for j = 1, 2, and
let ω0,r(X, Y ) = ω0(X)e−rY
s when r > 0. Then the following is true:
1
(1) If a1, a2 ∈ Σ′
s(R2d) satisfy OpA1(a1) = OpA2 (a2), then a1 ∈ Γ(ω0)
0,s (R2d) if
(2) Γ(ω1)
(3) Γ(ω0)
and only if a2 ∈ Γ(ω0)
0,s ⊆ Γ(ω1ω2)
0,s #Γ(ω2)
.
0,s (R2d).
0,s = \r>0
0,s
M ∞,1
(1/ω0,r) = \r≥0
M∞,1
(1/ω0,r).
In time-frequency analysis one also considers mapping properties for pseudo-
differential operators between modulation spaces or with symbols in modulation
spaces. Especially we need the following two results, where the first one is a gen-
eralisation of [44, Theorem 2.1] by Tachizawa, and the second one is a weighted
version of [21, Theorem 14.5.2]. We refer to [56] for the proof of the first two
propositions and to [56] for the proof of the third one.
E,s(R2d),
Proposition 1.27. Assume that A ∈ M(d, R), s ≥ 1, ω, ω0 ∈ P 0
a ∈ Γ(ω)
(R2d), and that B is an invariant BF-space on R2d of Beurling type.
Then OpA(a) is continuous from M (ω0ω, B) to M (ω0, B), and also continuous on
Ss(Rd) and on S′
s
s(Rd).
Proposition 1.28. Assume that A ∈ M(d, R), s ≥ 1, ω, ω0 ∈ PE,s(R2d),
a ∈ Γ(ω)
0,s (R2d), and that B is an invariant BF-space on R2d of Roumieu type.
Then OpA(a) is continuous from M (ω0ω, B) to M (ω0, B), and also continuous on
Σs(Rd) and on Σ′
s(Rd).
Proposition 1.29. Assume that p, q ∈ (0, ∞], r ≤ min(p, q, 1), ω ∈ PE(R2d ⊕
R2d) and ω1, ω2 ∈ PE(R2d) satisfy
ω2(X − Y )
ω1(X + Y )
≤ Cω(X, Y ),
X, Y ∈ R2d,
(1.30)
for some constant C. If a ∈ M∞,r
tinuous map from M p,q
(ω1)(Rd) to M p,q
(ω2)(Rd).
(ω) (R2d), then Opw(a) extends uniquely to a con-
Finally we need the following result concerning mapping properties of modulation
spaces under the Weyl product. The result is a special case of Theorem 0.3 in [10].
Proposition 1.30. Assume that ωj ∈ PE(R2d ⊕ R2d) for j = 0, 1, 2 satisfy
ω0(X, Y ) ≤ Cω1(X − Y + Z, Z)ω2(X + Z, Y − Z),
(1.31)
for some constant C > 0 independent of X, Y, Z ∈ R2d. Then the map (a, b) 7→
a#b from Σ1(R2d) × Σ1(R2d) to Σ1(R2d) extends uniquely to a mapping from
M∞,1
(ω1)(R2d) × M∞,1
In the proof of our main theorem we will need the following consequence of
(ω2)(R2d) to M∞,1
(ω0)(R2d).
Proposition 1.30.
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 15
Proposition 1.31. Let s ≥ 1, ω0, v0, v1 ∈ P 0
v0-moderate. Set ϑ = ω1/2
, and
0
E,s(R2d ⊕ R2d) be such that ω0 is
ω1(X, Y ) =
v0(2Y )1/2v1(2Y )
ϑ(X + Y )ϑ(X − Y )
,
ω2(X, Y ) = ϑ(X − Y )ϑ(X + Y )v1(2Y ) ,
v2(X, Y ) = v1(2Y ) .
Then
Γ(1/ϑ)
s
#M∞,1
(ω1)#Γ(1/ϑ)
s
Γ(1/ϑ)
s
#M∞,1
(v2)#Γ(1/ϑ)
s
⊆ M∞,1
(v2) ,
⊆ M∞,1
(ω2) .
The same holds with PE,s and Γ(1/ϑ)
0,s
at each occurrence.
in place of P 0
E,s and Γ(1/ϑ)
s
(1.32)
(1.33)
(1.34)
respectively,
Proof. Since Γ(1/ϑ)
it suffices to argue with the symbol class M∞,1
of Γ(1/ϑ)
=Sr≥0
M∞,1
.
s
s
(ϑr) for some sufficiently large r instead
(ϑr) with ϑr(X, Y ) = ϑ(X)erY
1
s (Proposition 1.25(3)),
Introducing the intermediate weight
ω3(X, Y ) =
v1(2Y )ϑ(X + Y )
ω0(X − Y )
.
we will show that for suitable r
ω3(X, Y ) ≤ Cω1(X − Y + Z, Z)ϑr(X + Z, Y − Z)
v1(2Y ) ≤ Cϑr(X − Y + Z, Z)ω3(X + Z, Y − Z) .
(1.35)
(1.36)
Proposition 1.30 applied to (1.35) shows that M∞,1
implies that Γ(1/ϑ)
Since ϑ is v1/2
-moderate and v0 ∈ P 0
(v2), and (1.33) follows.
(ω3) ⊆ M∞,1
#M∞,1
E,s, we have
0
s
(ω1)#Γ(1/ϑ)
s
⊆ M∞,1
(ω3), and (1.36)
ϑ(X − Y )−1 ≤ v0(2Z)1/2ϑ(X − Y + 2Z)−1
and ϑ(X + Y ) ≤ ϑ(X + Z)erY −Z
1
s
for suitable r > 0. Using these inequalities repeatedly in the following, a straight-
forward computation yields
ω3(X, Y ) =
v1(2Y )ϑ(X + Y )
ϑ(X − Y )2
≤ C1
v0(2Z)1/2v1(2Z)ϑ(X + Z)erY −Z
1
s
ϑ(X − Y + 2Z)ϑ(X − Y )
= C1ω1(X − Y + Z, Z)ϑr(X + Z, Y − Z),
for some C1 > 0 and r > 0.
16
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Likewise we obtain
v1(2Y ) =
ϑ(X − Y )v1(2Y )ϑ(X − Y )
ϑ(X − Y )2
≤ C1
≤ C2
ϑ(X − Y )v0(2Y )1/2v1(2Y )ϑ(X + Y )
ϑ(X − Y )2
ϑ(X − Y + Z)erZ
1
s v0(2(Y − Z))1/2v1(2(Y − Z))ϑ(X + Y )
ϑ(X − Y + 2Z)2
= C2ϑr(X − Y + Z, Z)ω3(X + Z, Y − Z) .
The twisted convolution relation (1.34) is proved similarly. Let
be the intermediate weight. Then the inequality
ω4(X, Y ) = ϑ(X − Y )v1(2Y )
ω4(X, Y ) = ϑ(X − Y )v1(2Y ) ≤ Cϑ(X − Y + Z)erZ
1
s v1(2(Y − Z))
= Cϑr(X − Y + Z, Z)v2(X + Z, Y − Z)
implies that Γ(1/ϑ)
s
#M∞,1
(v2) ⊆ M∞,1
(ω4).
Similarly we obtain
ω2(X, Y ) ≤ Cϑ(X − Y )v1(2Z)ϑ(X + Z)erZ−Y
1
s
= Cω4(X − Y + Z, Z)ϑ(X + Z)erZ−Y
1
s
= Cω4(X − Y + Z, Z)ϑr(X + Z, Y − Z),
and thus M∞,1
(ω4)#Γ(1/ϑ)
s
⊆ M∞,1
(ω2).
The case PE,s and Γ(1/ϑ)
0,s
in place of P 0
E,s and Γ(1/ϑ)
s
occurrence, is treated in similar ways and is left for the reader.
respectively, at each
(cid:3)
1.7. The Wiener Algebra Property. As a further crucial tool in our study of
the isomorphism property of Toeplitz operators we need to combine these continu-
ity results with convenient invertibility properties. The so-called Wiener algebra
property of certain symbol classes asserts that the inversion of a pseudo-differential
operator preserves the symbol class and is often referred to as the spectral invariance
of a symbol class.
Proposition 1.32. Let A ∈ M(d, R). Then the following are true:
(1) If s > 1, a ∈ Γ(1)
0,s(R2d) and OpA(a) is invertible on L2(Rd), then OpA(a)−1 =
OpA(b) for some b ∈ Γ(1)
0,s(R2d).
(2) If s ≥ 1, a ∈ Γ(1)
s (R2d) and OpA(a) is invertible on L2(Rd), then OpA(a)−1 =
OpA(b) for some b ∈ Γ(1)
s (R2d).
(3) If s ≥ 1, v0 ∈ P 0
E,s(R2d) is submultiplicative, v(X, Y ) ≡ v0(Y ), X, Y ∈
(v) (R2d) and OpA(a) is invertible on L2(Rd), then OpA(a)−1 =
R2d, a ∈ M ∞,1
OpA(b), for some b ∈ M ∞,1
(v) (R2d).
(4) If s > 1, v0 ∈ PE,s(R2d) is submultiplicative, v(X, Y ) ≡ v0(Y ), X, Y ∈
(v) (R2d) and OpA(a) is invertible on L2(Rd), then OpA(a)−1 =
R2d, a ∈ M ∞,1
OpA(b), for some b ∈ M ∞,1
(v) (R2d).
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 17
Proof. The results follows essentially from [22, Corollary 5.5] or [23]. More precisely,
Suppose s > 1, a ∈ Γ(1)
s (R2d), OpA(a) is invertible on L2(Rd), and let vr(X, Y ) =
erY f rac1s
when r ≥ 0. Then a ∈ M ∞,1
(vr) (R2d) for some r > 0. By [22, Corollary
5.5], Op(M ∞,1
(vr) (R2d)) is a Wiener algebra, giving that Op(a)−1 = Op(b) for some
b ∈ M ∞,1
s (R2d). This gives (2) in the case s > 1.
(vr) (R2d) ⊆ Γ(1)
If instead s = 1, then by [17, Theorem 4.4] there is an r0 > 0 such that Op(a)−1 =
Op(b) for some b ∈ M ∞,1
(vr0 )(R2d) ⊆ Γ(1)
1 (R2d), and (2) follows for general s ≥ 1.
By similar arguments, (1), (3) and (4) follow. The details are left for the reader.
(cid:3)
1.8. Toeplitz Operators. Fix a symbol a ∈ S (R2d) and a window ϕ ∈ S (Rd).
Then the Toeplitz operator Tpϕ(a) is defined by the formula
(Tpϕ(a)f1, f2)L2(Rd) = (a Vϕf1, Vϕf2)L2(R2d) ,
(1.37)
when f1, f2 ∈ S (Rd). Obviously, Tpϕ(a) is well-defined and extends uniquely to
a continuous operator from S ′(Rd) to S (Rd).
The definition of Toeplitz operators can be extended to more general classes of
windows and symbols by using appropriate estimates for the short-time Fourier
transforms in (1.37).
We state two possible ways of extending (1.37). The first result follows from [8,
Corollary 4.2] and its proof, and the second result is a special case of [57, Theorem
3.1]. We use the notation L(V1, V2) for the set of linear and continuous mappings
from the topological vector space V1 into the topological vector space V2. We also
set
ω0,t(X, Y ) = v1(2Y )t−1ω0(X)
for X, Y ∈ R2d .
(1.38)
Proposition 1.33. Let 0 ≤ t ≤ 1, p, q ∈ [1, ∞], and ω, ω0, v1, v0 ∈ P(R2d) be
such that v0 and v1 are submultiplicative, ω0 is v0-moderate and ω is v1-moderate.
Set
v = vt
1v0
and ϑ = ω1/2
,
0
and let ω0,t be as in (1.38). Then the following are true:
(1) The definition of (a, ϕ) 7→ Tpϕ(a) from S (R2d)×S (Rd) to L(S (Rd), S ′(Rd))
extends uniquely to a continuous map from M∞
L(S (Rd), S ′(Rd)).
(1/ω0,t)(R2d) × M 1
(v)(Rd) to
(2) If ϕ ∈ M 1
(v)(Rd) and a ∈ M∞
a continuous map from M p,q
(ϑω)(Rd) to M p,q
(ω/ϑ)(Rd).
(1/ω0,t)(R2d), then Tpϕ(a) extends uniquely to
Proposition 1.34. Let ω, ω1, ω2, v ∈ P(R2d) be such that ω1 is v-moderate, ω2 is
v-moderate and ω = ω1/ω2. Then the following are true:
(1) The mapping (a, ϕ) 7→ Tpϕ(a) extends uniquely to a continuous map from
(ω)(R2d) × M 2
L∞
(v)(Rd) to L(S (Rd), S ′(Rd)).
(2) If a ∈ L∞
(1/ω)(R2d) and ϕ ∈ M 2
(v)(Rd), then Tpϕ(a) extends uniquely to a
continuous operator from M 2
(ω1)(Rd) to M 2
(ω2)(Rd).
1.9. Weyl formulation of Toeplitz operators. We finish this section by recall-
ing some important relations between Weyl operators, Wigner distributions, and
Toeplitz operators. For instance, the Weyl symbol of a Toeplitz operator is the con-
volution between the Toeplitz symbol and a Wigner distribution. More precisely,
if a ∈ S (R2d) and ϕ ∈ S (Rd), then
Tpϕ(a) = (2π)−d/2 Opw(a ∗ Wϕ,ϕ) .
(1.39)
18
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Our analysis of Toeplitz operators is based on the pseudo-differential operator
representation given by (1.39). Furthermore, any extension of the definition of
Toeplitz operators to cases which are not covered by Propositions 1.33 and 1.34
is based on this representation. Here we remark that this leads to situations were
certain mapping properties for the pseudo-differential operator representation make
sense, whereas similar interpretations are difficult or impossible to make in the
framework of (1.37) (see Remark 5.7 in Section 2). We refer to [50] or Section 2 for
extensions of Toeplitz operators in context of pseudo-differential operators.
2. Confinement of the symbol classes Γ(ω)
s
(Rd) and Γ(ω)
0,s (Rd)
In this section we introduce and discuss basic properties for confinements for
symbols in Γ(ω0)
0,s . These considerations follow lines similar to the
discussions in [3,34], but are here adapted to symbols that possess Gevrey regularity.
In particular, this requires the deduction of various types of delicate estimates,
concerning the compositions of symbols that are confined in different ways.
and in Γ(ω0)
s
We recall that if σ is the (standard) symplectic form on R2d, namely,
σ(X, Y ) = hy, ξi − hx, ηi,
X = (x, ξ) ∈ R2d, Y = (y, η) ∈ R2d,
then the symplectic Fourier transform Fσ is defined by
and the twisted convolution ∗σ is given by
(Fσa)(X) =ba(X) ≡ π−dZ a(Y )e2iσ(X,Y ) dY,
π(cid:19) d
(a ∗σ b)(X) ≡(cid:18) 2
2Z a(X − Y )b(Y )e2iσ(X,Y ) dY,
1/2(R2d). The twisted convolution is linked to the Weyl product
(a#b)(X) = π−dZ a(X − Y )bb(Y )e2iσ(X,Y ) dY.
σ is the identity map.
We also note that F 2
In what follows we let aY = a( · − Y ) when a ∈ S′
analogous ways, bY , φY , ϕY , ψY etc. are defined when b, φ, ϕ, ψ ∈ S′
admissible a and b we have
1/2(R2d) and Y ∈ R2d, and in
1/2(R2d). For
(a#b)Y = aY #bY .
(2.1)
We also recall that if ϕ ∈ Ss(R2d), then there are functions φ, ψ ∈ Ss(R2d) such
that ϕ = φ#ψ (cf. [5, 58]). The same is true if Ss is replaced by Σs or by S . In
particular, by choosing ϕ such thatR ϕ(X) dX = 1, (2.1) gives the following.
2 . Then there are φ, ψ ∈ Ss(R2d) such that
Proposition 2.1. Let s ≥ 1
ψY #φY dY = 1.
(2.2)
ZR2d
The same holds true with Σs or S in place of Ss, provided s > 1
2 .
For independent translations in Weyl products we have the following.
Proposition 2.2. Let s ≥ 1
2 and let φ, ψ ∈ Ss(R2d). Then
(φY #ψZ)(X) = Ψ(X − Y, X − Z)
for some Ψ ∈ Ss(R2d × R2d). The same holds true with Σs or S in place of Ss.
for suitable a, b ∈ S′
by the formula
hence,
a#b = (2π)− d
2 a ∗σ (Fσb),
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 19
Proof. We only prove the result when φ, ψ ∈ Ss(R2d). The other cases follow by
similar arguments and are left for the reader.
We have
(φY #ψZ )(X) = π−dZ φ(X − Y − Y1)bψ(Y1)e2iσ(Y1,Z)e2iσ(X,Y1) dY1
Ψ(X, Z) = π−dZ φ(X − Y1)bψ(Y1)e2iσ(Z,Y1) dY1.
= π−dZ φ((X − Y ) − Y1)bψ(Y1)e2iσ(X−Z,Y1) dY1 = Ψ(X − Y, X − Z),
We note that
where
Ψ = (Fσ,2 ◦ T )(φ ⊗ bψ),
where (T Φ)(X, Z) = Φ(X − Z, Z) when Φ ∈ Ss(R2d × R2d), and Fσ,2Φ is the
partial symplectic Fourier transform of Φ(X, Z) with respect to the Z variable.
Since (φ, ψ) 7→ φ⊗bψ is continuous from Ss(R2d)×Ss(R2d) to Ss(R2d ×R2d), and T
and Fσ,2Φ are continuous on Ss(R2d × R2d), it follows that Ψ ∈ Ss(R2d × R2d). (cid:3)
Since Ψ in Proposition 2.2 belongs to similar types of spaces as φ and ψ, it
follows that estimates of the form
DαΨ(X, Y ) . hαα!se−(X
1
s +Y
1
s )/h
hold true. In particular, the following is an immediate consequence of Proposition
2.2 and some standard manipulations in Gelfand-Shilov theory.
Corollary 2.3. Let s ≥ 1
2 . If φ, ψ ∈ Ss(R2d) (φ, ψ ∈ Σs(R2d)), then
Dα
X Dβ
Y Dγ
Z(φY #ψZ )(X) . hα+β+γ(α!β!γ!)se−(X−Y
1
s +X−Z
1
s )/h
(2.3)
for some h > 0 (for every h > 0).
Proof. Obviously, as stated in the previous Proposition 2.2, we can write φY #ψZ
as Ψ(X − Y, X − Z) for some Ψ ∈ Ss(R2d × R2d). Thus
Indeed, using the fact that Pδ≤α(cid:0)α
(n + k)! ≤ 2n+kn!k!, then
δ(cid:1) = 2α and that n! ≤ 2n(n − k)!k!, which implies
(α + β + γ)! =
(αj + βj + γj)! ≤
4αj+βj +γj αj!βj!γj! = 4α+β+γα!β!γ!.
dYj=1
dYj=1
Thus, inequality (2.3) holds with 2 · 4sh in place of h.
(cid:3)
The next fundamental result is a consequence of Theorem 4.12 in [4].
Dα
X Dβ
Y Dγ
2
1 Dγ
1 Dγ+α−δ
ZΨ(X − Y, X − Z) =(cid:12)(cid:12)(cid:12)Dα
X(cid:16)Dβ
≤Xδ≤α(cid:18)α
δ(cid:19)(cid:12)(cid:12)(cid:12)(cid:16)Dβ+δ
Ψ(cid:17) (X − Y, X − Z)(cid:12)(cid:12)(cid:12)
≤ hα+β+γXδ≤α(cid:18)α
Xδ≤α(cid:18)α
2 Ψ(cid:17) (X − Y, X − Z)(cid:12)(cid:12)(cid:12)
δ(cid:19) ((β + δ)!(γ + α − δ)!)s e−r(cid:16)X−Y
δ(cid:19) ((β + δ)!(γ + α − δ)!)s ≤ 2α4sα+β+γ (α!β!γ!)s .
1
s +X−Z
1
s (cid:17).
Moreover, we have
20
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Proposition 2.4. Let s ≥ 1
continuous from Σs(R2d) × Γ(ϑ)
2 and ϑ ∈ PE(R2d).Then, the map (φ, a) 7→ φ#a is
s
(R2d) to Ss(R2d).
The next lemma concerns uniform estimates of the Weyl product between ele-
ments in sets Ωj,Yj , Yj ∈ R2d, j = 1, 2, such that related sets Ω∪
j , given by
Ω∪
j ≡ [Y ∈R2d
{ a( · + Y ) ; a ∈ Ωj,Y },
(2.4)
are bounded in Ss(R2d) or in Σs(R2d), j = 1, 2.
j be as in (2.4),
j and choose Yj ∈ R2d such that aj ∈ Ωj,Yj , j = 1, 2. Then the following is
2 and let Ωj,Y ⊆ S (R2d), Y ∈ R2d, Ω∪
Lemma 2.5. Let s ≥ 1
aj ∈ Ω∪
true:
(1) if Ω∪
j are bounded in Ss(R2d), then there are constants C > 0 and h > 0
which are independent of Yj ∈ R2d and aj, j = 1, 2, and such that
(a(α1)
1 #a(α2)
2
)(X) ≤ Chα1+α2(α1!α2!)se−(X−Y1
1
s +X−Y2
1
s +Y1−Y2
1
s )/h
and
Dα(a1#a2)(X) ≤ Chαα!se−(X−Y1
1
s +X−Y2
1
s +Y1−Y2
1
s )/h
(2.5)
(2.6)
(2) if Ω∪
j are bounded in Σs(R2d), then for every h > 0, there is a constant
C > 0 which is independent of Yj ∈ R2d and aj, j = 1, 2, and such that
(2.5) and (2.6) hold.
Proof. We only prove (2). The assertion (1) follows by similar arguments and is
left for the reader.
We have DX(a1#a2) = (DX a1)#a2+a1#(DX a2), which implies that the Leibnitz
rule is valid for the Weyl product. Hence, (2.6) follows for every h > 0 if we prove
that (2.5) holds for every h > 0.
Let Y = Y1, Z = Y2, a = a1, b = a2, aY = a( · + Y ) and bZ = b( · + Z). Then
a1#a2(X) = π−dZ aY ((X − Y ) − Y1)Fσ(bZ( · − Z))(Y1)e2iσ(X,Y1) dY1
= π−dZ aY ((X − Y ) − Y1)Fσ(bZ )(Y1)e2iσ(X−Z,Y1) dY1
= Fσ(aY (X1 − · )(FσbZ))(X2),
where X1 = X − Y and X2 = X − Z. That is a1#a2(X) = G(X1, X2), where
G = GY,Z = (Fσ,2 ◦ T )(aY ⊗ bZ).
Here Fσ,2F is the partial symplectic Fourier transform of F ( · , X2) with respect to
the variable X2 ∈ R2d, and
(T F )(X, Y ) = F (X − Y, Y ).
Since both Fσ,2 and T are continuous on Σs(R2d × R2d), it follows from the bound-
edness of the sets Ω∪
j that for every h > 0, there is a constant Ch which is indepen-
dent of aj ∈ Ω∪
j such that
Dα
X1 Dβ
X2
G(X1, X2) ≤ Chhα+β(αβ)se−(X1
1
s +X2
1
s )/h.
By straightforward computations it follows that
Dα
X1 Dβ
X2
G(X1, X2) = (a(α)#b(β)),
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 21
since
Y1 − Y2
1
s ≤ c(X − Y1
1
s + X − Y2
1
s )
for some constant c which only depends on s. The estimate (2.5) follows from these
relations.
(cid:3)
Lemma 2.6. Let s ≥ 1
ψZ = ψ( · − Z). Then the following is true:
2 , φ, ψ ∈ Σs(R2d), ω, ϑ ∈ PE(R2d), φY = φ( · − Y ), and
(1) if a ∈ Γ(ω)
s
(R2d) (a ∈ Γ(ω)
0,s (R2d)), then
Dα
X Dβ
Y (φY a)(X) . hα
1 hβ
2 (α!β!)se−X−Y
1
s /h1 min(ω(X), ω(Y ))
(2.7)
and
Dα
X Dβ
Y (φY #a)(X) . hα
1 hβ
2 (α!β!)se−X−Y
1
s /h1 min(ω(X), ω(Y )),
(2.8)
for some h1 > 0 (for every h1 > 0) and every h2 > 0;
(2) if a1 ∈ Γ(ω)
s
(R2d) and a2 ∈ Γ(ϑ)
s
then
(R2d) (a1 ∈ Γ(ω)
0,s (R2d) and a2 ∈ Γ(ϑ)
0,s (R2d)),
Dα
X Dβ
Y Dγ
Z((φY a1)#(ψZ a2))(X)
. hα+β
1
hγ
2 (α!β!γ!)se−(X−Y
1
s +X−Z
1
s +Y −Z
1
s )/h1
for some h1 > 0 (for every h1 > 0) and every h2 > 0.
min
X1,X2∈{X,Y,Z}(cid:0)ω(X1)ϑ(X2)(cid:1),
Proof. We only consider the case when a ∈ Γ(ω)
other cases follow by similar arguments and are left for the reader.
0,s (R2d) and b ∈ Γ(ϑ)
0,s (R2d). The
Let
By Leibniz rule we get
Ψ(X, Y ) = φ(X − Y )a(X).
Dα
X Dβ
. 2α sup
Y Ψ(X, Y ) ≤Xγ≤α(cid:18)α
γ(cid:19)φ(α+β−γ)(X − Y )a(γ)(X)
γ≤α(cid:18)hα+β((α + β − γ)!γ!)se−X−Y
1
s /hω(X)(cid:19)
≤ (21+sh)α+β(α!β!)se−X−Y
1
s /hω(X) . (21+sh)α+β(α!β!)se−X−Y
1
s /(2h)ω(Y ),
for every h > 0 which is chosen small enough. Here we have used the fact that for
some c > 0
ω(X) . ω(Y )ecX−Y . ω(Y )eX−Y
1
s /(2h),
since ω is a moderate function. This gives (2.7).
Next we prove (2). Let
Ω1,Y =( X 7→
Ω2,Z =(cid:26) X 7→
Dβ
Y (φY a)
hββ!sω(Y )
Dγ
Z(ψZ b)
hγγ!sϑ(Z)
; β ∈ Nd)
; γ ∈ Nd(cid:27) ,
22
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
and let Ω∪
in Σs(R2d). Hence, Lemma 2.5 shows that
2 be as in (2.4). By (2.7) it follows that Ω∪
1 and Ω∪
1 and Ω∪
2 are bounded
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Dα
X Dβ
hββ!sω(Y )! #(cid:18) Dγ
Y (φY a)
Z(ψZ b)
hγγ!sϑ(Z)(cid:19)! (X)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
. hαα!se−(X−Y
1
s +X−Z
1
s +Y −Z
1
s )/h
for every h > 0, or equivalently,
Dα
X Dβ
Y Dγ
Z((φY a)#(ψZ b))(X)
. hα+β+γ(α!β!γ!)se−(X−Y
1
s +X−Z
1
s +Y −Z
1
s )/hω(Y )ϑ(Z).
The assertion now follows from the latter estimate and the fact that ω and ϑ are
moderate weights, giving that
ω(Y ) . ω(X)eX−Y
1
s /(2h) . ω(Z)e(X−Y
1
s +X−Z
1
s )/(2h),
and similarly for ϑ.
Finally, the estimate in (2) also holds for (φY b)#(ψZa) in place of (φY a)#(ψZ b),
and (2.8) follows from this estimate by letting γ = 0, b = 1, ϑ = 1, and then
integrate with respect to Z. The proof is complete.
(cid:3)
Lemmas 2.5 and 2.6 imply the following characterisation of Γ(ω)
s
(R2d).
1(R2d), φ ∈ Σs(R2d) have
Proposition 2.7. Let s > 1/2, ω ∈ PE(R2d), a ∈ Σ′
non-vanishing integrals, and let φY = φ( · − Y ). Then the following conditions are
equivalent:
s
;
(1) a ∈ Γ(ω)
(2) φY a is smooth and satisfies (2.7) for some h1 > 0 and every h2 > 0;
(3) φY #a is smooth and satisfies (2.8) for some h1 > 0 and every h2 > 0;
(4)
Dα
X(φY a)(X) . hα
1 α!se−X−Y
1
s /h1 min(ω(X), ω(Y )),
(2.9)
for some h1 > 0;
(5)
Dα
X (φY #a)(X) . hα
1 α!se−X−Y
1
s /h1 min(ω(X), ω(Y )),
(2.10)
for some h1 > 0.
Lemma 2.8. Let s ≥ 1
are bounded families in Ss(R2d). Then, for any X, Y, Z ∈ R2d, α ∈ N2d,
2 . Assume that {a1( · +Y, Y )}Y ∈R2d and {a2( · +Z, Z)}Z∈R2d
Dα
X (a1( · + Y, Y )#a2( · + Z, Z))(X) . hαα!se−r(X−Y
1
s +X−Z
1
s ),
(2.11)
for some h > 0.
Proof. The result is a straightforward consequence of Corollary 2.3 and its proof,
since the involved constants on the right hand side of (2.3) depend continuously on
φ and ψ in Ss(R2d).
(cid:3)
We notice that, by straightforward computation, for some other h, r > 0, (2.11)
gives, for any X, Y, Z ∈ R2d, α ∈ N2d,
Dα
X (a1( · +Y, Y )#a2( · +Z, Z))(X) . hαα!se−r(X−Y
1
s +X−Z
1
s +Y −Z
1
s ). (2.12)
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 23
and Γ(1)
0,s. Let IR = [−R, R] and E0 = E0
∞(R2d)), with the symbol subspace sw
h,s}∞
2.1. A family related to Γ(1)
s
L∞(IR × R2d; sw
We shall consider suitable decreasing family {En
aim, let
Gn = {(Y, T1, . . . , Tn) ∈ R2d(n+1) : Y, Tj ∈ R2d with Tj ≤ 1, j = 1, . . . , n}, n ∈ Z+.
We define En
h,s =
∞(R2d) from Definition 1.22.
n=0 of Banach spaces. To this
h,s, n ≥ 1, as the set of all a ∈ E0 such that
kak(n) = sup
1≤k≤n
sup
t∈IR
sup
(Y,T1,...,Tk)∈Gk
khT1, DXi · · · hTk, DX ia(t, Y, · )ksw
∞
hk(k!)s
< ∞,
with the norm
kakEn
h,s
≡ kakE0 + kak(n).
We also let E∞
h,s be the set of all
a ∈ \n≥0
En
h,s
kakE∞
h,s
≡ sup
n≥0
kakEn
h,s
such that
is finite.
Lemma 2.9. Let n ≥ 0. Then En
h,s is a Banach space.
Proof. Let {aj}j≥0 be a Cauchy sequence in En
h,s, n ≥ 1. By definition, this
sequence clearly has a limit a ∈ E0, and for some X 7→ bk(t, Y, T1, . . . , Tk, X) ∈
sw
∞(R2d) we have
lim
j→∞
sup
khT1, DX i · · · hTk, DX iaj(t, Y, · ) − bk(t, Y, T1, . . . , Tk, · )ksw
∞
hk(k!)s
= 0,
where the supremum is taken over all
k ∈ {1, . . . , n},
We need to prove that a ∈ En
h,s, and aj → a in En
h,s.
t ∈ IR and (Y, T1, . . . , Tk) ∈ Gk.
The conditions here above are equivalent to
lim
j→∞
sup
t∈IR
sup
Y ∈R2d
kaj(t, Y, · ) − a(t, Y, · )ksw
∞ = 0
(2.13)
and
lim
j→∞
sup
k(−1)khT1, Di · · · hTk, Diaj(t, Y, · ) − bk(t, Y, T1, . . . , Tk, · )ksw
∞
hk(k!)s
= 0,
(2.14)
where the latter supremum should be taken over all
k ∈ {1, . . . , n},
t ∈ IR and (Y, T1, . . . , Tk) ∈ Gk.
Since sw
(2.14) that
∞(R2d) is continuously embedded in S ′(R2d), it follows from (2.13) and
has the limit
X 7→ (−1)khT1, DXi · · · hTk, DXiaj (t, Y, X)
X 7→ (−1)khT1, DX i · · · hTk, DX ia(t, Y, X)
in S ′(R2d), and the limit
X 7→ bk(t, Y, T1, . . . , Tk, X)
in sw
∞(R2d), and thereby in S ′(R2d), as j tends to ∞. Hence
bk(t, Y, T1, . . . , Tk, X) = (−1)khT1, DX i · · · hTk, DX ia(t, Y, X)
24
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
and it follows that En
n ≥ 0.
h,s is a Banach space for every h > 0, s > 0 and integer
(cid:3)
The spaces E∞
h,s can be related to Γ(1)
s
and Γ(1)
0,s, as the following lemma shows.
The details are left for the reader.
Lemma 2.10. Let a ∈ L∞(IR × R2d; sw
uniformly bounded family in Γ(1)
s (R2d) ( Γ(1)
∞(R2d)). Then {a(t, Y, · )}t∈IR,Y ∈R2d is a
0,s(R2d)), if and only if
for some h > 0 (for every h > 0).
kakE∞
h,s
< ∞
Later on we also need the following result of differential equations with functions
depending on a real variable with values in E∞
h,s. The proof is omitted since the
result can be considered as a part of the standard theory of ordinary differential
equations of first order in Banach spaces.
Lemma 2.11. Suppose s ≥ 0 and T > 0, and let K be an operator from E∞
E∞
h,s for every h > 0 such that
h,s to
kKakE∞
h,s
≤ CkakE∞
h,s
,
a ∈ E∞
h,s,
(2.15)
for some constant C which only depend on h > 0. Then
dc(t)
dt
= K(c(t)),
c(0) ∈ E∞
h,s,
has a unique solution t 7→ c(t) from [−T, T ] to E∞
h,s which satisfies
where C is the same as in (2.15).
kc(t)kE∞
h,s
≤ kc(0)kE∞
h,s
eCT ,
3. One-parameter group of elliptic symbols in the classes Γ(ω)
s
(Rd)
s
In this section we show that for suitable s and ω0, there are elements a ∈ Γ(ω0)
and b ∈ Γ(1/ω0)
such that a#b = b#a = 1. This is essentially a consequence of The-
orem 3.8, where it is proved that the evolution equation (0.4) has a unique solution
a(t, · ) which belongs to Γ(ωϑt)
, thereby deducing needed semigroup properties for
scales of pseudo-differential operators. Similar facts hold for corresponding Beurling
type spaces (cf. Theorem 3.9).
s
s
First we have the following result on certain logarithms of weight functions.
Theorem 3.1. Let ω ∈ PE(R2d) ∩ Γ(ω)
that ω is v-moderate, ϑ(X) = 1 + log v(X) and let
s0 (R2d), s0 ∈ (0, 1], v ∈ PE(R2d), be such
c(X, Y ) = log
ω(X + Y )
ω(Y )
.
Then,
• {c( · , Y )}Y ∈R2d is a uniformly bounded family in Γ(ϑ)
• for α 6= 0, {(∂α
(R2d), s ≥ 1;
X c)( · , Y )}Y ∈R2d is a uniformly bounded family in Γ(1)
s
s (R2d),
s ≥ 1.
For the proof of Theorem 3.1 we will need the following multidimensional version
of the well-known Fa`a di Bruno formula for the derivatives of composed functions.
It can be found, e.g., setting q = p = 1, n = 2d, in equations (2.3) and (2.4) in [29].
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 25
Lemma 3.2. Let f : R → R, g : R2d → R. Then, for any α ∈ N2d, α 6= 0,
∂αf (g(x))
α!
= X1≤k≤α
f (k)(g(x))
k!
Xβ1+···+βk=α
βj6=0, j=1,...,k Y1≤j≤k
(∂βj g)(x)
βj!
.
(3.1)
We will also need the next factorial estimate, for expressions involving decompo-
sitions of α ∈ N2d, α 6= 0, into the sum of k nontrivial multiindeces βj, j = 1, . . . , k,
as in (3.1), and corresponding products of (powers of) factorials.
Lemma 3.3. Let s0 ∈ (0, 1], α ∈ N2d, α 6= 0. Then, for a suitable C0 > 0,
depending only in d,
X1≤k≤α
1
k Xβ1+···+βk=α
βj6=0, j=1,...,k Y1≤j≤k
(βj!)s0−1 . Cα
0 .
(3.2)
The proof of Lemma 3.3 can be found in the Appendix.
Proof of Theorem 3.1. We have to show that c( · , Y ) satisfies Γ(ϑ)
estimates, uni-
formly with respect to Y ∈ R2d. If c(X, Y ) ≥ 0, then it follows by submultiplica-
tivity of ω, that
s
c(X, Y ) = log ω(Y + X) − log ω(Y ) . log ω(Y ) + log v(X) − log ω(Y )
≤ ϑ(X),
for any Y ∈ R2d. Again by moderateness, when c(X, Y ) ≤ 0, recall that ω(X+Y ) ≥
ω(Y )
v(X) , so that
c(X, Y ) & log
ω(Y )
v(X)
− log ω(Y ) ≥ − log v(X) ≥ −ϑ(X),
and we can conclude c(X, Y ) . ϑ(X), X ∈ R2d. Now, for α ∈ N2d, α 6= 0, (3.1)
gives
∂α
X c(X, Y ) = α! X1≤k≤α
(−1)k+1
k [ω(X + Y )]k Xβ1+···+βk=α
βj 6=0, j=1,...,k Y1≤j≤k
(∂βj ω)(X + Y )
βj!
We can then estimate, in view of (3.2),
∂α
1
X c(X, Y ) . α! X1≤k≤α
= hα α! X1≤k≤α
k [ω(X + Y )]k Xβ1+···+βk=α
k Xβ1+···+βk=α
βj 6=0, j=1,...,k Y1≤j≤k
βj 6=0, j=1,...,k Y1≤j≤k
1
ω(X + Y )hβj(βj!)s0
βj!
(βj !)s0−1 . (C0h)α(α!)s,
which concludes the proof.
(cid:3)
Proposition 3.4. Assume s > 1
for some r > 0. Let
{a( · , Y )}Y ∈R2d be a uniformly bounded family in Σs(R2d) and {c( · , Z)}Z∈R2d be
a bounded family in Γ(ω)
2 and ω(X) . erX
(R2d). Then,
s
1
s
{a( · , Y )#c( · , Z)}Y,Z∈R2d and {c( · , Z)#a( · , Y )}Y,Z∈R2d
are bounded families in Ss(R2d).
26
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
Proof. An immediate consequence of (2.8) in Lemma 2.6 is the following
Dα
X (φ#a)(X) ≤ Chαα!se−rX
1
s ,
(3.3)
which implies that, for φ ∈ Σs and a ∈ Γ(ω)
, (3.3) holds if and only if φ#a belongs
to Ss. Then by the proof of (3.3) it follows that the constant C, h and r can be
chosen depending continuously on φ ∈ Σs(R2d) and a ∈ Γ(ω)
(R2d). Hence if Ω1
is a bounded family in Σs(R2d) and Ω2 is a bounded family in Γ(ω)
(R2d), then
{φ#a}φ∈Ω1,a∈Ω2 is a bounded family in Ss(R2d).
(cid:3)
s
s
s
The following result can be found, e.g., in [46].
Lemma 3.5. Let a ∈ S ′(R2d). Then,
kaksw
∞ ≤ C Xα≤d+1
kakL∞ ≤ C Xα≤2d+1
k∂αakL∞
k∂αaksw
∞ ,
(3.4)
(3.5)
and
for some constant C > 0 depending only on the dimension d.
Proposition 3.6. Let a ∈ S ′(R2d), s ≥ 1
α, β ∈ N2d. Then the following conditions are equivalent:
2 and set bαβ(X) = ∂α(X βa(X)) when
(1) a ∈ Ss(R2d);
(2) for some h > 0 it holds
kbαβkL∞ . hα+β(α!β!)s,
α, β ∈ N2d;
(3) for some h > 0 it holds
kbαβksw
∞ . hα+β(α!β!)s,
α, β ∈ N2d.
Proof. The equivalence between (1) and (2) is well-known. The proof of the equiv-
alence of (2) and (3) follows by a straightforward application of Lemma 3.5.
In
fact, assume that (2) holds true. Then (3.4) gives
kbαβksw
∞ ≤ C Xγ≤d+1
= hα+β(α!β!)s Xγ≤d+1
k∂γbαβkL∞(R2d) . Xγ≤d+1
hγ(γ!)s(cid:18) (α + γ)!
α! γ! (cid:19)s
. (2sh)α+β(α!β!)s,
hα+β+γ((α + γ)!β!)s
with a constant depending only on d and h, since
Xγ≤d+1
hγ(γ!)s(cid:18) (α + γ)!
α! γ! (cid:19)s
≤ C1 · 2s(α+d+1) ≤ C22sα+β,
where the constants C1 and C2 only depend on d and h. Then, (3) holds true,
as claimed. The proof of the converse follows by similar argument, employing
(3.5).
(cid:3)
We have now the following.
Theorem 3.7. Let a ∈ S ′(R2d) and s > 0. Then the following conditions are
equivalent:
(1) a ∈ Γ(1)
s (R2d);
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 27
(2) there exists h > 0 such that
k∂αakL∞(R2d) . hα(α!)s,
for all α ∈ Z2d
+ ;
(3) there exists h > 0 such that
k∂αaksw
∞ . hα(α!)s,
for all α ∈ Z2d
+ ;
(4) there exists h > 0 such that
khT1, DX i · · · hTm, DX iaksw
∞ . hm(m!)s,
for any T1, . . . , Tm ∈ R2d such that Tj ≤ 1, j = 1, . . . , m, m ≥ 1.
Proof. The equivalence between (1) and (2) is well known. The equivalence of (2)
and (3) is proved by an argument completely similar to the one employed in the
proof of Proposition 3.6, using Lemma 3.5. It remains to prove only the equivalence
with (4). Assume that (3) holds true, and let
for the standard basis (el)l=1,...,d of Rd and the dual basis (εl)l=1,...,d. Recall that
so that the symbol hT1, DX i · · · hTm, DX ia is in the span of symbols of the form
hTk, DX ia =
tkl
∂a
∂xl
τkl
∂a
∂ξl
,
Tj =
tjlel +
τjlεl,
dXl=1
dXl=1
mXj=1
dXl=1
+
dXl=1
mXj=1
mYj=1
j τ γj
tβj
x ∂γ
ξ a,
j ∂β
βj = β,
γj = γ, β + γ = m,
where the summation contains at most (2d)m terms. Since Tj ≤ 1, j = 1, . . . , m,
by (3.4) we find
khT1, DX i · · · hTm, DX iaksw
∞ ≤ (2d)m sup
α=m
k∂αaksw
∞
hα+γ ((α + γ)!)s
. sup
α=m Xγ≤d+1
= sup
α=m
hα(α!)s Xγ≤d+1
. (2s+1h)m(m!)s,
hγ(γ!)s(cid:18) (α + γ)!
α! γ! (cid:19)s
which proves that (4) holds true. As in the proof of Proposition 3.6, the converse
implication is obtained by a completely similar argument, employing (3.5).
(cid:3)
We are now ready to state and prove the first main result of this section, the
following Theorem 3.8.
It deals with the existence of one-parameter groups of
pseudo-differential operators, obtained as solutions to suitable evolution equations.
Theorem 3.8. Let s ≥ 1, ω, ϑ ∈ P 0
ϑ ∈ Γ(ϑ)
smooth map (t, X) 7→ a(t, X) ∈ C such that a(t, · ) ∈ Γ(ω ϑt)
and
(R2d) and
s (R2d). Then, there exists a unique
(R2d) for all t ∈ R,
E,s(R2d) be such that ω ∈ Γ(ω)
(R2d), and let a0 ∈ Γ(ω)
(R2d), b ∈ Γ(1)
s
s
s
s
((∂ta)(t, · ) = (b + log ϑ)#a(t, · )
a(0, · ) = a0.
(3.6)
28
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
If ω ≡ a0 ≡ 1, then a(t, X) also satisfies
((∂ta)(t, · ) = a(t, · )#(b + log ϑ)
a(0, · ) = a0,
(3.7)
and
a(t1, · )#a(t2, · ) = a(t1 + t2, · ),
a(t, · ) ∈ Γ(ϑt)
s
(R2d),
t, t1, t2 ∈ R.
(3.8)
Proof. First suppose that a solution a(t, X) of (3.6) exists. Then,
with
a(t, X) = a0(X) +Z t
0
c(u, X) du
c(t, · ) = (b + log ϑ)#a(t, · ) ∈ Γ(ωhlog ϑiϑt)
s
(R2d),
in view of Theorem 3.1 and the properties of the Weyl product in the Γ(ω)
(R2d)
classes, see [4]. This implies that the map t 7→ a(t, · ) is C1 from [−R, R] into the
symbol space
s
Γ(ωhlog ϑi(ϑ+ϑ−1)R)
s
(R2d).
Choose s0 < s, and φ, ψ ∈ Ss0 (R2d) such that (2.2) holds true. Let
c1(t, Y, · ) = ω(Y )−1ϑ(Y )−t φY #a(t, · ).
(3.9)
By Lemma 2.6 (1) we have that, for any Y ∈ R2d, t 7→ c1(t, Y, · ) is a C1 map from
[−R, R] into Ss(R2d). Moreover,
∂tc1(t, Y, · ) = ω(Y )−1ϑ(Y )−t φY #(b + log ϑ)#a(t, · ) − c1(t, Y, · ) log ϑ(Y ).
Let
Then,
with
and
f (Y, X) = b(X) + log
ϑ(X)
ϑ(Y )
.
(∂tc1)(t, Y, · ) = ω(Y )−1 ϑ(Y )tZ φY #f (Y, · )#ψZ #φZ #a(t, · ) dZ
=Z KY,Z(t, · )#c1(t, Z, · ) dZ
KY,Z(t, · ) =
ω(Z) ϑ(Z)t
ω(Y ) ϑ(Y )t φY #f (Y, · )#ψZ,
(3.10)
c1(0, Y, · ) = ω(Y )−1φY #a0.
We also need to consider the similar equation where f (Y, · ) is replaced by f (Z, · ),
that is
∂tc2(t, Y, · ) =Z eKY,Z(t, · )#c2(t, Z, · ) dZ,
eKY,Z(t, · ) =
ω(Z) ϑ(Z)t
ω(Y ) ϑ(Y )t φY #f (Z, · )#ψZ ,
c2(0, Y, · ) = c1(0, Y, · ) = ω(Y )φY #a0.
(3.11)
(3.12)
where
and
We consider the operators K and eK on E0, defined by
(Ka)(t, Y, X) =Z (KY,Z(t, · )#a(t, Z, · ))(X) dZ,
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 29
and
and show that
(eKa)(t, Y, X) =Z (eKY,Z(t, · )#a(t, Z, · ))(X) dZ,
kKakEn
h,s
≤ C(n + 1)kakEn
h,s
,
kKakE∞
h,s
≤ CkakE∞
h,s
,
≤ C(n + 1)kakEn
h,s
and
h,s
≤ CkakE∞
h,s
(3.13)
for some constant C, which is independent of h, n and s.
In order to prove this, it is convenient to let Pk be the family of all subsets of
{1, . . . , k}, k ≥ 1. For each P ∈ Pk, a ∈ sw
∞(R2d), we set
h,s
keKakE∞
when P = ∅,
keKakEn
H(a, P ) =(a
hTj1, DX i · · · hTjl, DX ia when P = {j1 < · · · < jl}, l ≤ k.
We now estimate
k(hT1, DX i · · · hTk, DX iKa)(t, Y, · )ksw
∞(R2d)
hk(k!)s
.
when a ∈ En
h,s. Since
(hT1, DX i · · · hTk, DX iKa)(t, Y, X)
= hT1, DX i · · · hTk, DX iZ (KY,Z(t, · )#a(t, Z, · ))(X) dZ
= XP ∈PkZ (H(KY,Z(t, · ), P )#H(a(t, Z, · ), P c))(X) dZ,
we find
k(hT1, DXi · · · hTk, DX iKa)(t, Y, · )ksw
∞
hk(k!)s
·
kH(a(t, Z, · ), P c)ksw
∞
hk−l((k − l)!)s
dZ
≤
kXl=0 XP =l(cid:18)k
l(cid:19)−sZ kH(KY,Z(t, · ), P )ksw
kXl=0 XP =l
kakEk−l
hl l!s
≤
∞
h,s(cid:18)k
kXl=0 XP =l(cid:18)k
l(cid:19)−sZ kH(KY,Z(t, · ), P )ksw
hl l!s
∞
dZ
l(cid:19)−1Z kH(KY,Z(t, · ), P )ksw
hl l!s
∞
. kakEn
h,s
dZ,
(3.14)
where the last inequality follows from the fact that s ≥ 1 and kakEn
with n.
h,s
increases
We have now to estimate kH(KY,Z(t, · ), P )ksw
ties on the right-hand side of (3.10). Since ω and ϑ belong to P 0
for every r > 0,
∞ , and study the different quanti-
E,s, it follows that
ω(Z) ϑ(Z)t
ω(Y ) ϑ(Y )t =
ω(Z)
ω(Y )(cid:18) ϑ(Z)
ϑ(Y )(cid:19)t
. erY −Z
1
s (cid:18)erY −Z
1
s(cid:19)t
. er(1+t)Y −Z
Y, Z ∈ R2d.
(3.15)
1
s ,
30
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
For the Weyl product in (3.10) we have
φY #f (Y, · ) = φ( · − Y )#(cid:16)b( · ) + log
=(cid:16)φ#b( · + Y )(cid:17)Y
=(cid:16)φ#b( · + Y )(cid:17)Y
ϑ( · )
ϑ( · )
ϑ(Y )(cid:17)
+ φ( · − Y )#(cid:16) log
+(cid:16)φ# log
and nφ# log
ϑ(Y )(cid:17)
ϑ(Y ) (cid:17)Y
ϑ(Y ) oY ∈R2d
ϑ( · + Y )
ϑ( · + Y )
.
nφ#b( · + Y )oY ∈R2d
By Theorem 3.1 and Proposition 3.4,
are uniformly bounded families in Ss(R2d). Note that
(3.16)
a2(Z, X) = ψZ (X) ⇒ {a2(Z, · + Z)}Z∈R2d = {ψ}Z∈R2d,
which is evidently a uniformly bounded family in Ss(R2d). Combining this last
observation with the computations on φY #f (Y, · ) above, using Lemma 2.8 and
(2.12), we finally obtain, for some h, r0 > 0,
Dα
X (φY #f (Y, · )#ψZ)(X) . hαα!se−r0(X−Y
1
s +X−Z
1
s +Y −Z
1
s ),
X, Y, Z ∈ R2d, α ∈ N2d.
(3.17)
By Theorem 3.7, (3.15) and (3.17) we get for all P ∈ Pk Y, Z ∈ R2d and some
r0, h > 0 that
kH(KY,Z(t, · ), P )ksw
∞ ≤ Chll!se−r0Y −Z
1
s ,
l = P ,
where C is independent of k. Hence, (3.14) gives
kKa(t, Y, · )ksw
∞ . kakEn
h,s
,
and
khT1, DX i · · · hTk, DXiKa(t, Y, · )ksw
∞
hk(k!)s
≤ C1kakEn
h,s
1
s dZ
kXl=0 XP =l(cid:18)k
l(cid:19)−1Z e−r0Y −Z
kXl=0 XP =l(cid:18)k
l(cid:19)−1
= C2kakEn
h,s
= C2(k + 1)kakEn
,
1 ≤ k ≤ n,
h,s
as claimed, where C1 and C2 are independent of k, n and h.
In fact, by similar arguments that lead to (3.16) it follows that
By a completely similar argument, an analogous result can be obtained for eK.
{b( · + Z)#ψ}Z∈R2d
and (cid:26)log
ϑ( · + Z)
ϑ(Z)
#ψ(cid:27)Z∈R2d
are bounded in Ss(R2d), given that (3.17) holds with f (Z, · ) in place of f (Y, · ).
This gives the first and third inequalities in (3.13). From these estimates we get
kKakEn+1
h,s
≤ CkakEn
h,s
≤ CkakEn
h,s
,
and keKakEn+1
h,s
which give the other inequalities in (3.17).
We have proven that for any T > 0, then
kKtkEn
h,s→En
h,s
≤ C(n + 1),
t ≤ T,
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 31
where C is independent of n. As a consequence, since ω(Y )−1φY #a0 belongs to
E∞
h,s, the equation
dc1
dt
= Kc1,
c1(0) = ω(Y )−1φY #a0
(3.18)
has a unique solution on R belonging to E∞
sition 2.7 it follows that c1(t, Y, · ) ∈ Γ(1)
t.
h,s, in view of Lemma 2.11. By Propo-
s (R2d), uniformly in Y and for bounded
In order to prove the uniqueness of the solution a of (3.6), first we assume the
existence and by what we have proven above i.e. that c1(t, Y, · ) in (3.9) satisfies
(3.18) which implies the uniqueness of the solution of (3.6), since
ω(Y )ϑ(Y )tψY #c1(t, Y, · ) dY.
(3.19)
a(t, · ) =ZR2d
ψY #φY #a(t, · ) dY =ZR2d
a(t, · ) =ZR2d
To prove the existence of a solution of (3.6), we consider the solution c2(t, Y, · ) of
(3.11) with the initial data (3.12), and we let
ω(Y )ϑ(Y )tψY #c2(t, Y, · ) dY.
(3.20)
By Propositions 2.7 and 1.8, the family {ψY #c2(t, Y, · )}Y ∈R2d belongs to Ss and
a(t, · ) belongs to Γ(wϑt)
. Moreover,
s
da(t, · )
dt
ω(Y )ϑ(Y )t log ϑ(Y )ψY #c2(t, Y, · ) dY
= (b + log ϑ)#a(t, · ),
with the initial data
a(0, · ) =ZR2d
ω(Y )ψY #(ω(Y )−1φY #a0) dY = a0,
which provide a solution of (3.6).
In order to prove the last part we consider the unique solution a(t, · ) of (3.6)
with the initial data a(0, · ) ≡ 1. If ω ≡ 1, then for u ∈ R the mappings
t 7→ a(t + u, · ) and t 7→ a(t, · )#a(u, · )
are both solutions of (3.6) with value a(u, · ) at t = 0, and
a(t + u, · ) = a(t, · )#a(u, · ),
(3.21)
by the uniqueness property for the solution of (3.6).
Using (3.21) we have for all t ∈ R, a(t, · )#a(−t, · ) = 1. Taking the derivative
we get
d
dt
0 =
(a(t, · )#a(−t, · )) = (b+log ϑ)#a(t, · )#a(−t, · )−a(t, · )#(b+log ϑ)#a(−t, · ).
That is (b + log ϑ) = a(t, · )#(b + log ϑ)#a(−t, · ), implying the commutation for
the sharp product of a(t, · ) with (b + log ϑ), and the result follows.
(cid:3)
By similar argument as for the previous result we get the following.
=ZR2d
=ZR2d
=ZR2d
+ZR2dZR2d
+ZR2dZR2d
ω(Y )ϑ(Y )tψY #eKY,Z(t, · )#c2(t, Z, · ) dY dZ
ω(Z)ϑ(Z)t log ϑ(Z)ψZ #c2(t, Z, · ) dZ
ω(Z) ϑ(Z)tψY #φY #f (Z, · )#ψZ #c2(t, Z, · ) dY dZ
ω(Z)ϑ(Z)t(b + log ϑ)ψZ #c2(t, Z, · ) dZ
32
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
(R2d), and let a0 ∈ Γ(ω)
Theorem 3.9. Let s ≥ 1, ω, ϑ ∈ PE,s(R2d) be such that ω ∈ Γ(ω)
ϑ ∈ Γ(ϑ)
smooth map (t, X) 7→ a(t, X) ∈ C such that a(t, · ) ∈ Γ(ω ϑt)
and a(t, · ) satisfies (3.6).
(R2d) and
0,s(R2d). Then, there exists a unique
(R2d) for all t ∈ R,
0,s (R2d), b ∈ Γ(1)
s
s
s
Moreover, if ω ≡ a0 ≡ 1, then a(t, X) also satisfies (3.7) and
a(t1, · )#a(t2, · ) = a(t1 + t2, · ),
a(t, · ) ∈ Γ(ϑt)
0,s (R2d),
t, t1, t2 ∈ R.
4. Lifting of pseudo-differential operators and Toeplitz operators
on modulation spaces
In this section we apply the group properties in Theorems 3.8 and 3.9 to deduce
lifting properties of pseudo-differential operators on modulation spaces. Thereafter
we combine these results by the Wiener property of certain pseudo-differential op-
erators with symbols in suitable modulation spaces to get lifting properties for
Toeplitz operators with weights as their symbols.
We begin to apply Theorems 3.8 and 3.9 to get the following.
E,s(R2d), and let
Theorem 4.1. Let s ≥ 1, p ∈ (0, ∞]2d, A ∈ M(d, R), ω ∈ P 0
B be an invariant BF-space on R2d, or B = Lp,E(R2d) for some phase-shift split
parallelepiped E in R2d. Then the following are true:
(1) There exist a ∈ Γ(ω)
s
(R2d) and b ∈ Γ(1/ω)
s
(R2d) such that
OpA(a) ◦ OpA(b) = OpA(b) ◦ OpA(a) = IdS ′
s(Rd) .
(4.1)
Furthermore, OpA(a) is an isomorphism from M (ω0, B) onto M (ω0/ω, B),
for every ω0 ∈ P 0
E,s(R2d).
(2) Let a0 ∈ Γ(ω)
s
(ω1/ω)(Rd) for some ω1 ∈ P 0
(R2d) be such that OpA(a0) is an isomorphism from M 2
(ω1)(Rd)
E,s(R2d). Then OpA(a0) is an iso-
to M 2
E,s(R2d).
morphism from M (ω2, B) to M (ω2/ω, B), for every ω2 ∈ P 0
Furthermore, the inverse of OpA(a0) is equal to OpA(b0) for some b0 ∈
Γ(1/ω)
s
(R2d).
Theorem 4.2. Let s > 1, p ∈ (0, ∞]2d, A ∈ M(d, R), ω ∈ PE,s(R2d), and let
B be an invariant BF-space on R2d, or B = Lp,E(R2d) for some phase-shift split
parallelepiped E in R2d. Then the following are true:
(1) There exist a ∈ Γ(ω)
0,s (R2d) and b ∈ Γ(1/ω)
0,s
(R2d) such that
OpA(a) ◦ OpA(b) = OpA(b) ◦ OpA(a) = IdΣ′
s(Rd) .
(4.2)
Furthermore, OpA(a) is an isomorphism from M (ω0, B) onto M (ω0/ω, B),
for every ω0 ∈ PE,s(R2d).
(2) Let a0 ∈ Γ(ω)
0,s (R2d) be such that OpA(a0) is an isomorphism from M 2
(ω1)(Rd)
to M 2
(ω1/ω)(Rd) for some ω1 ∈ PE,s(R2d). Then OpA(a0) is an iso-
morphism from M (ω2, B) to M (ω2/ω, B), for every ω2 ∈ PE,s(R2d).
Furthermore, the inverse of OpA(a0) is equal to OpA(b0) for some b0 ∈
Γ(1/ω)
0,s
(R2d).
We only prove Theorem 4.2. Theorem 4.1 follows by similar arguments and is
left for the reader.
Proof of Theorem 4.2. The existence of a ∈ Γ(ω)
(R2d) such
that (4.2) holds is guaranteed by Theorem 3.9. By [56, Theorems 2.3 and 2.6] it
0,s (R2d) and b ∈ Γ(1/ω)
0,s
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 33
follows that
and
OpA(a) : M (ω0, B) →M (ω0/ω, B)
OpA(b) : M (ω0/ω, B) →M (ω0, B)
(4.3)
(4.4)
are continuous. By (4.2) and the fact that M (ω0, B) and M (ω0/ω, B) are contained
in Σ′
s(R2d), it follows that (4.3) and (4.4) are homeomorphic, and (1) follows.
(2) It suffices to prove the result in the Weyl case, A = 1
2 I, in view of Proposition
1.25. By (1), we may find
a1 ∈ Γ(ω1)
0,s ,
b1 ∈ Γ(1/ω1)
0,s
,
a2 ∈ Γ(ω1/ω)
0,s
,
b2 ∈ Γ(ω/ω1)
0,s
satisfying the following properties:
• Opw(aj) and Opw(bj) are inverses to each others on Σ′
• For arbitrary ω2 ∈ PE,s(R2d), the mappings
s(Rd) for j = 1, 2;
Opw(a1) : M 2
Opw(b1) : M 2
Opw(a2) : M 2
Opw(b2) : M 2
(ω2/ω1),
(ω2ω1),
(ω2ω/ω1),
(ω2) → M 2
(ω2) → M 2
(ω2) → M 2
(ω2) → M 2
(ω2ω1/ω)
(4.5)
are isomorphisms.
In particular, Opw(a1) is an isomorphism from M 2
(ω1) to L2, and Opw(b1) is an
isomorphism from L2 to M 2
(ω1).
Now set c = a2#a#b1. Then by [4, Theorem 4.14], the symbol c satisfies
c = a2#a#b1 ∈ Γ(ω1/ω)
s
#Γ(ω)
s #Γ(1/ω1)
s
⊆ Γ(1)
s .
Furthermore, Opw(c) is a composition of three isomorphisms and consequently
Opw(c) is boundedly invertible on L2.
By Proposition 1.32 (2), Opw(c)−1 = Opw(c1) for some c1 ∈ Γ(1)
0,s. Hence, by
(1) it follows that Opw(c) and Opw(c1) are isomorphisms on M (ω2, B), for each
ω2 ∈ PE,s(R2d). Since Opw(c) and Opw(c1) are bounded on every M (ω, B), the
factorization of the identity Opw(c) Opw(c1) = Id is well-defined on every M (ω, B).
Consequently, Opw(c) is an isomorphism on M (ω, B).
Using the inverses of a2 and b1, we now find that
Opw(a) = Opw(b2) ◦ Opw(c) ◦ Opw(a1)
is a composition of isomorphisms from the domain space M (ω2, B) onto the im-
age space M (ω2/ω, B) (factoring through some intermediate spaces) for every
ω2 ∈ PE,s(R2d) and every invariant BF-space B. This proves the isomorphism
assertions for Opw(a).
Finally, the inverse of Opw(a) is given by
Opw(b1) ◦ Opw(c1) ◦ Opw(a2).
which is a Weyl operator with symbol in Γ(1/ω)
0,s
, and the result follows.
(cid:3)
34
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
5. Mapping properties for Toeplitz operators
In this section we study the isomorphism properties of Toeplitz operators be-
tween modulation spaces. We first state results for Toeplitz operators that are
well-defined in the sense of (1.37) and Propositions 1.33 and 1.34. Then we state
and prove more general results for Toeplitz operators that are defined only in the
framework of pseudo-differential calculus.
We start with the following result about Toeplitz operators with smooth symbols.
(R2d)
Theorem 5.1. Let s ≥ 1 ω, ω0, v ∈ P 0
and that ω0 is v-moderate, and let B be an invariant BF-space on R2d or B =
Lp,E(R2d) for some phase-shift split parallelepiped E in R2d. If φ ∈ M 1
(v)(Rd),
then Tpφ(ω0) is an isomorphism from M (ω, B) to M (ω/ω0, B).
E,s(R2d) be such that ω0 ∈ Γ(ω0)
s
In the next result we relax our restrictions on the weights but impose more
restrictions on B.
Theorem 5.2. Let s > 1, 0 ≤ t ≤ 1, p, q ∈ [1, ∞], and ω, ω0, v0, v1 ∈ PE,s(R2d)
1v0, ϑ = ω1/2
be such that ω0 is v0-moderate and ω is v1-moderate. Set v = vt
and
let ω0,t be the same as in (1.38). If φ ∈ M 1
(v)(Rd) and ω0 ∈ M∞
(1/ω0,t)(R2d), then
Tpφ(ω0) is an isomorphism from M p,q
(ϑω)(Rd) to M p,q
(ω/ϑ)(Rd).
0
Before the proofs of Theorems 5.1 and 5.2 we state the following consequence of
Theorem 5.2 which was the original goal of our investigations.
Corollary 5.3. Let s ≥ 1, ω, ω0, v1, v0 ∈ PE,s(R2d) and that ω0 is v0-moderate
and ω is v1-moderate. Set v = v1v0 and ϑ = ω1/2
(v)(Rd), then Tpφ(ω0) is
an isomorphism from M p,q
(ω/ϑ)(Rd) simultaneously for all p, q ∈ [1, ∞].
(ϑω)(Rd) to M p,q
. If φ ∈ M 1
0
Proof. Let ω1 ∈ PE,s(R2d) ∩ Γ(ω1)
0,s (R2d) be such that C−1 ≤ ω1/ω0 ≤ C, for some
constant C. Hence, ω/ω0 ∈ L∞ ⊆ M ∞. By Theorem 2.2 in [50], it follows that
(ω2)(R2d), when ω2(x, ξ, η, y) = 1/ω0(x, ξ). The result
ω = ω1 · (ω/ω1) belongs to M ∞
now follows by setting t = 1 and q0 = 1 in Theorem 5.2.
(cid:3)
In the proofs of Theorems 5.1 and 5.2 we consider Toeplitz operators as defined by
an extension of the form (1.37). Later on we present extensions of these theorems
(cf. Theorems 5.1′ and 5.2′ below) for those readers who accept to use pseudo-
differential calculus to extend the definition of Toeplitz operators. Except for the
interpretation of Tpφ(ω) the proofs of Theorems 5.1 and 5.2 are identical to those
of Theorems 5.1′ and 5.2′.
We need some preparations and start with the following lemma.
Lemma 5.4. Let s ≥ 1, ω, v ∈ PE,s(R2d) be such that ϑ = ω1/2 is v-moderate.
Assume that φ ∈ M 2
(ϑ)(Rd) onto
M 2
(v). Then Tpφ(ω) is an isomorphism from M 2
(1/ϑ)(Rd).
Proof. Recall from Remark 1.13 that for φ ∈ M 2
ϑkL2 defines an equivalent norm on M 2
to φ are well defined.
(v)(Rd) \ {0} the expression kVφf ·
(ϑ). Thus the occurring STFTs with respect
Since Tpφ(ω) is bounded from M 2
(ϑ) to M 2
k Tpφ(ω)f kM 2
(1/ϑ)
holds for all f ∈ M 2
(ϑ).
Next we observe that
(1/ϑ) by Proposition 1.34, the estimate
. kf kM 2
(5.1)
(ϑ)
(Tpφ(ω)f, g)L2(Rd) = (ωVφf, Vφg)L2(R2d) = (f, g)M 2,φ
(ϑ)
,
(5.2)
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 35
for f, g ∈ M 2
sition 1.12(3)) now yields the following identity:
(ϑ)(Rd) and φ ∈ M 2
(v)(Rd). The duality of modulation spaces (Propo-
kf kM 2
(ϑ)
≍
≍
sup
kgkM 2
(ϑ)
sup
kgkM 2
(ϑ)
=1
=1
(f, g)M 2
(ϑ)
(Tpφ(ω)f, g)L2 ≍ k Tpφ(ω)f kM 2
(1/ϑ)
.
(5.3)
A combination of (5.1) and (5.3) shows that kf kM 2
equivalent norms on M 2
(ϑ)
(ϑ).
and k Tpφ(ω)f kM 2
(1/ϑ)
are
In particular, Tpφ(ω) is one-to-one from M 2
(1/ϑ) with closed range. Since
Tpφ(ω) is self-adjoint with respect to L2, it follows by duality that Tpφ(ω) has
dense range in M 2
(1/ϑ). By Banach's theorem
Tpφ(ω) is an isomorphism from M 2
(cid:3)
(1/ϑ). Consequently, Tpφ(ω) is onto M 2
(ϑ) to M 2
(ϑ) to M 2
(1/ϑ).
We need a further generalization of Proposition 1.33 to more general classes of
symbols and windows. Set
ω1(X, Y ) =
v0(2Y )1/2v1(2Y )
ω0(X + Y )1/2ω0(X − Y )1/2 .
(5.4)
Proposition 1.33′. Let s ≥ 1, 0 ≤ t ≤ 1, p, q, q0 ∈ [1, ∞], and ω, ω0, v0, v1 ∈
PE,s(R2d) be such that v0 and v1 are submultiplicative, ω0 is v0-moderate and ω
is v1-moderate. Set
r0 = 2q0/(2q0 − 1),
v = vt
1v0
and ϑ = ω1/2
0
,
and let ω0,t and ω1 be as in (1.38) and (5.4). Then the following are true:
(1) The definition of (a, φ) 7→ Tpφ(a) from Σs(R2d)×Σs(Rd) to L(Σs(Rd), Σ′
s(Rd))
extends uniquely to a continuous map from M∞,q0
L(Σs(Rd), Σ′
s(Rd)).
(1/ω0,t)(R2d) × M r0
(v)(Rd) to
(2) If φ ∈ M r0
a0 ∈ M∞,1
M p,q
(ϑω)(Rd) to M p,q
(ω/ϑ)(Rd).
(v)(Rd) and a ∈ M∞,q0
(1/ω0,t)(R2d), then Tpφ(a) = Opw(a0) for some
(ω1)(R2d), and Tpφ(a) extends uniquely to a continuous map from
For the proof we need the following result, which follows from [49, Proposition
2.1] and its proof.
Lemma 5.5. Assume that s ≥ 1, q0, r0 ∈ [1, ∞] satisfy r0 = 2q0/(2q0 − 1). Also
assume that v ∈ PE,s(R2d) is submultiplicative, and that κ, κ0 ∈ PE,s(R2d ⊕ R2d)
satisfy
κ0(X1 + X2, Y ) ≤ Cκ(X1, Y ) v(Y + X2)v(Y − X2) X1, X2, Y ∈ R2d,
(5.5)
for some constant C > 0. Then the map (a, φ) 7→ Tpφ(a) from Σs(R2d) ×
s(Rd)) extends uniquely to a continuous mapping from
Σs(Rd) to L(Σs(Rd), Σ′
M∞,q0
(v)(Rd)
and a ∈ M∞,q0
(κ) (R2d), then Tpφ(a) = Opw(b) for some b ∈ M∞,1
(κ0).
s(Rd)). Furthermore, if φ ∈ M r0
(v)(Rd) to L(Σs(Rd), Σ′
(ω) (R2d) × M r0
Proof of Proposition 1.33 ′. We show that the conditions on the involved parame-
ters and weight functions satisfy the conditions of Lemma 5.5.
First we observe that
vj(2Y ) ≤ Cvj (Y + X2)vj (Y − X2),
j = 0, 1
36
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
for some constant C which is independent of X2, Y ∈ R2d, because v0 and v1 are
submultiplicative. Refering back to (5.4) this gives
ω1(X1 + X2, Y ) =
v0(2Y )1/2v1(2Y )
ω0(X1 + X2 + Y )1/2ω0(X1 + X2 − Y )1/2
≤ C1
v0(2Y )1/2v1(2Y )v0(X2 + Y )1/2v0(X2 − Y )1/2
ω0(X1)
= C1v1(2Y )1−t v0(2Y )1/2v1(2Y )tv0(X2 + Y )1/2v0(X2 − Y )1/2
ω0(X1)
Hence
≤ C2v1(2Y )1−t v1(X2 + Y )tv1(X2 − Y )tv0(X2 + Y )v0(X2 − Y )
ω0(X1)
.
ω1(X1 + X2, Y ) ≤ C
v1(2Y )1−tv(X2 + Y )v(X2 − Y )
ω0(X1)
.
(5.6)
By letting κ0 = ω1 and κ = 1/ω0,t, it follows that (5.6) agrees with (5.5). The
result now follows from Lemma 5.5.
(cid:3)
In the remaining part of the paper we interpret Tpφ(a) as the extension of a
Toeplitz operator provided by Proposition 1.33′. (See also Remark 5.7 below for
more comments.)
Proposition 1.33′ can be applied to Toeplitz operators with smooth weights as
symbols.
Proposition 5.6. Assume that s ≥ 1, ω0 ∈ P 0
that v ∈ P 0
M 2
(v)(Rd), then Tpφ(ω0) = Opw(b) for some b ∈ Γ(ω0)
E,s(R2d) is submultiplicative, and that ω1/2
0
(R2d).
s
is v-moderate.
If φ ∈
E,s(R2d) be such that ω0 ∈ Γ(ω0)
s
(R2d),
Proof. By Propositions 1.25 and 1.26 with t = 0 we have ω0 ∈ M∞,1
(1/ω0,r0 )(R2d)
s . Furthermore, by letting
1
for some r0 ≥ 0, where ω0,r0(X, Y ) = ω0(X)e−r0Y
v1(Y ) = er0Y
s , and ω1 in (5.4) we have
1
ω1(X, Y ) &
er02Y
1
s v(2Y )1/2
ω0(X + Y )1/2ω0(X − Y )1/2
1
s
er0Y
ω0(X)
.
&
Proposition 1.33′ implies that existence of some b ∈ M∞,1
(1/ω0,r0 )(R2d) ⊆ Γ(ω0)
s
(R2d).
(cid:3)
The following generalization of Theorem 0.1 is an immediate consequence of
Theorem 4.1, Lemma 5.4 and Proposition 5.6, since it follows by straightforward
computations that Ss ⊆ M 2
(v) when v satisfies the hypothesis in Proposition 5.6.
Theorem 0.1′. Let s ≥ 1, ω, ω0 ∈ P 0
E,s(R2d), p ∈ (0, ∞]2d, B be an invariant
BF-space on R2d or B = Lp,E(R2d) for some phase-shift split parallelepiped E in
R2d, and let φ ∈ Ss(Rd). Then the Toeplitz operator Tpφ(ω0) is an isomorphism
from M (ω, B) onto M (ω/ω0, B).
Remark 5.7. As remarked and stated before, there are different ways to extend
the definition of a Toeplitz operator Tpφ(a) (from φ ∈ Σ1(Rd) and a ∈ Σ1(R2d))
to more general classes of symbols and windows. For example, Propositions 1.33
and 1.34 are based on the "classical" definition (1.37) of such operators and a
straightforward extension of the L2-form on Σ1. Proposition 5.6 interprets Tpφ(ω)
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 37
as a pseudo-differential operator. Let us emphasize that in this context the bilinear
E,s(R2d)
form (1.37) may not be well defined, even when φ ∈ M 2
satisfies ω ∈ Γ(ω)
(R2d). (See also [26, Remark 3.8].)
(v)(Rd) and ω ∈ P 0
s
In Theorems 5.1′ and 5.2′ below, we extend the definition of Toeplitz opera-
tors within the framework of pseudo-differential calculus and we interpret Toeplitz
operators as pseudo-differential operators. With this understanding, the stated
mapping properties are well-defined.
The reader who is not interested in full generality or does not accept Toeplitz
operators that are not defined directly by an extension of (1.37) may only consider
the case when the windows belong to M 1
(v) and stay with Theorems 5.1 and 5.2.
For the more general window classes in Theorems 5.1′ and 5.2′ below, however, one
should then interpret the involved operators as "pseudo-differential operators that
extend Toeplitz operators".
The following generalization of Theorem 5.1 is an immediate consequence of
Theorem 4.1, Lemma 5.4 and Proposition 5.6.
(R2d)
Theorem 5.1′. Let s ≥ 1, ω, v, v0 ∈ P 0
and that ω0 is v-moderate, and let B be an invariant BF-space on R2d or B =
(v)(Rd),
Lp,E(R2d) for some phase-shift split parallelepiped E in R2d. If φ ∈ M 2
then Tpφ(ω0) is an isomorphism from M (ω, B) to M (ω/ω0, B).
E,s(R2d) be such that ω0 ∈ Γ(ω0)
s
Theorem 5.1′ holds only for smooth weight functions.
In order to relax the
conditions on the weight function ω0, we use the Wiener algebra property of M∞,1
(v)
instead of Γ(1)
s . On the other hand, we have to restrict our results to modulation
spaces of the form M p,q
(ω) instead of M (ω, B).
Theorem 5.2′. Let s > 1, 0 ≤ t ≤ 1, p, q, q0 ∈ [1, ∞] and ω, ω0, v0, v1 ∈ PE,s(R2d)
be such that ω0 is v0-moderate and ω is v1-moderate. Set r0 = 2q0/(2q0 − 1),
(v)(Rd) and
v = vt
ω0 ∈ M∞,q0
(ω/ϑ)(Rd).
1v0, ϑ = ω1/2
(1/ω0,t), then Tpφ(ω0) is an isomorphism from M p,q
and let ω0,t be the same as in (1.38). If φ ∈ M r0
(ϑω)(Rd) to M p,q
0
Proof. First we note that the Toeplitz operator Tpφ(ω0) is an isomorphism from
(1/ϑ) in view of Lemma 5.4. With ω1 defined in (5.4), Proposition 1.33′
M 2
implies that there exist b ∈ M∞,1
(ϑ) to M 2
s(R2d) such that
(ω1) and c ∈ S′
Tpφ(ω0) = Opw(b) and Tpφ(ω0)−1 = Opw(c) .
Let ω2 be the "dual" weight defined as
ω2(X, Y ) = ϑ(X − Y )ϑ(X + Y )v1(2Y ).
(5.7)
We shall prove that c ∈ M∞,1
proved the existence of such a symbol c. Then we may proceed as follows.
(ω2)(R2d). Let us assume for now that we have already
After checking (1.30), we can apply Proposition 1.29 and find that each of the
mappings
Opw(b) : M p,q
(ωϑ) → M p,q
(ω/ϑ)
and Opw(c) : M p,q
(ω/ϑ) → M p,q
(ωϑ)
(5.8)
is well-defined and continuous.
In order to apply Proposition 1.30, we next check condition (1.31) for the weights
ω1, ω2, and
ω3(X, Y ) =
ϑ(X + Y )
ϑ(X − Y )
.
38
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
In fact we have
ω1(X − Y + Z, Z)ω2(X + Z, Y − Z)
= (cid:16)
v0(2Z)1/2v1(2Z)
ϑ(X − Y + 2Z)ϑ(X − Y )(cid:17) ·(cid:0)ϑ(X − Y + 2Z)ϑ(X + Y )v1(2(Y − Z))(cid:1)
v0(2Z)1/2v1(2Z)v1(2(Y − Z)) ϑ(X + Y )
=
&
ϑ(X − Y )
ϑ(X + Y )
ϑ(X − Y )
= ω3(X, Y ) .
(ϑ) to M 2
(ω3)(R2d), or equivalently, b#c ∈ M∞,1
Therefore Proposition 1.30 shows that the Weyl symbol of Opw(b) ◦ Opw(c) belongs
to M∞,1
(ω3). Since Opw(b) is an isomorphism from
(1/ϑ) with inverse Opw(c), it follows that b#c = 1 and that the constant
M 2
symbol 1 belongs to M∞,1
(ω3). By similar arguments it follows that c#b = 1. Therefore
the identity operator Id = Opw(b) ◦ Opw(c) on M p,q
(ω/ϑ), and
thus Opw(b) = Tpφ(ω0) is an isomorphism from M p,q
(ω/ϑ) with inverse
Opw(c). This proves the assertion.
(ωϑ) factors through M p,q
(ωϑ) onto M p,q
It remains to prove that c ∈ M∞,1
Section 3 (cf. Theorem 3.9), we choose a ∈ Γ(1/ϑ)
that the map
0,s
(ω2)(R2d). Using once again the basic result in
0,s (R2d) such
(R2d) and a0 ∈ Γ(ϑ)
Opw(a) : L2(Rd) → M 2
(ϑ)(Rd)
is an isomorphism with inverse Opw(a0). By Propositions 1.25 and 1.26, Opw(a) is
also bijective from M 2
(1/ϑ)(Rd) to L2(Rd). Furthermore, by Theorem 4.2 it follows
that a ∈ M∞,1
(ϑr) when r ≥ 0, where
ϑr(X, Y ) = ϑ(X)erY
1
s .
Let b0 = a#b#a. From Proposition 1.31 we know that
b0 ∈ M∞,1
(v2)(R2d), where
v2(X, Y ) = v1(2Y )
(5.9)
is submultiplicative and depends on Y only. Since Opw(b) is bijective from M 2
(ϑ)
to M 2
(1/ϑ) by Lemma 5.4 (2), Opw(b0) is bijective and continuous on L2.
Since v2 is submultiplicative and in PE,s(R2d), M∞,1
(v2) is a Wiener algebra by
Proposition 1.32. Therefore, the Weyl symbol c0 of the inverse to the bijective
operator Opw(b0) on L2 belongs to M∞,1
(v2)(R2d).
Opw(c0) = Opw(b0)−1 = Opw(a)−1 Opw(b)−1 Opw(a)−1,
Since
we find
Opw(a0) = Opw(b)−1 = Opw(a) Opw(c0) Opw(a),
or equivalently,
a0 = a#c0#a, where a ∈ Γ(1/ϑ)
0,s
and c0 ∈ M∞,1
(v2) .
(5.10)
The definitions of the weights are chosen such that Proposition 1.31 implies that
a0 ∈ M∞,1
(cid:3)
(ω2), and the result follows.
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 39
6. Examples on bijective pseudo-differential operators on
modulation spaces
In this section we construct explicit isomorphisms between modulation spaces
with different weights. Applying the results of the previous sections, these may be
in the form of pseudo-differential operators or of Toeplitz operators.
E,s(R2d), and let B be an invariant BF-
Proposition 6.1. Let s ≥ 1, ω0, ω ∈ P 0
space on R2d or B = Lp,E(R2d) for some phase-shift split parallelepiped E in R2d.
For λ = (λ1, λ2) ∈ R2
+ let Φλ be the Gaussian
Φλ(x, ξ) = Ce−(λ1x2+λ2ξ2) .
(1) ω0 ∗ Φλ belongs to P 0
E,s(R2d) ∩ Γ(ω0)
0,1
for all λ ∈ R2
+ and
ω0 ∗ Φλ ≍ ω0.
(2) If λ1 · λ2 < 1, then there exists ν ∈ R2
+ and a Gauss function φ on Rd such
that Opw(ω0∗Φλ) = Tpφ(ω0∗Φν) is bijective from M (ω, B) to M (ω/ω0, B)
for all ω ∈ PE,s(R2d).
(3) If λ1·λ2 ≤ 1 and in addition ω0 ∈ Γ(ω0)
s
(R2d), then Opw(ω0∗Φλ) = Tpφ(ω0)
is bijective from M (ω, B) to M (ω/ω0, B) for all ω ∈ PE,s(R2d).
Proof. The assertion (1) follows easily from the definitions.
(2) Choose µj > λj such that µ1 · µ2 = 1. Then the Gaussian Φµ is a multiple
of a Wigner distribution, precisely Φµ = cW (φ, φ) with φ(x) = e−µ1x2/2. By the
semigroup property of Gaussian functions (cf. e.g., [19, 21]) there exists another
Gaussian, namely Φν, such that Φλ = Φµ ∗ Φν . Using (1.39), this factorization
implies that the Weyl operator with symbol ω0 ∗ Φλ is in fact a Toeplitz operator,
namely
Opw(ω0 ∗ Φλ) = Opw(ω0 ∗ Φν ∗ Φµ)
= Opw(ω0 ∗ Φν ∗ cW (φ, φ))
= c(2π)d/2Tpφ(ω0 ∗ Φν ).
By (1) ω0 ∗Φν ∈ P 0
0,1 (R2d) is equivalent to ω0. Hence Theorem 5.1′
shows that Opw(ω0 ∗ Φλ) is bijective from M (ω, B) to M (ω/ω0, B). This proves
(2).
E,s(R2d)∩Γ(ω0)
(3) follows from (2) in the case λ1 · λ2 < 1.
Φλ = cW (φ, φ) for φ(x) = e−λ1x2/2 and thus
If λ1 · λ2 = 1, then as above
Opw(ω0 ∗ Φλ) = Tpw
φ (ω0)
is bijective from M (ω, B) to M (ω/ω0, B), since ω0 ∈ P 0
E,s(R2d) ∩ Γ(ω0)
s
(R2d). (cid:3)
Appendix A. Proof of Lemma 3.3
Lemma A.1. Let α = (α1, . . . , αd) ∈ Nd. Then the number of elements in the set
Ωk,α ≡ { (β1, . . . , βk) ∈ Nkd ; β1 + · · · + βk = α }
(A.1)
is equal to
dYj=1(cid:18)αj + k
k (cid:19).
40
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
For the proof we recall the formula
kXj=0(cid:18)n + j
j (cid:19) =(cid:18)n + k + 1
(cid:19),
k
which follows by a standard induction argument.
(A.2)
Proof. Let N be the number of elements in the set (A.1), which is the searched
number, and let Nj be the number of elements of the set
{ (β0
1, . . . , β0
k) ∈ Nk ; β0
1 + · · · + β0
k = αj },
j = 1, . . . , d
By straightforward computations it follows that N = N1 · · · Nd, and it suffices to
prove the result in the case d = 1, and then α = α1.
In order to prove the result for d = 1, let γ ∈ N,
S1(γ) =
γXβ=0
1 = γ + 1,
and define inductively
By straightforward computations it follows that N = N1 = Sk(α). We claim
Sj(β),
j = 1, 2, . . . .
Sj+1(γ) =
γXβ=0
Sj(γ) =(cid:18)γ + j
j (cid:19),
j = 1, 2, . . . .
(A.3)
In fact, (A.3) is obviously true for j = 1. Assume that (A.3) holds for j = n,
and consider Sn+1(γ). Then (A.2) gives
Sn+1(γ) =
Sn(β) =
γXβ=0
γXβ=0(cid:18)β + n
n (cid:19)
γXβ=0(cid:18)β + n
=
β (cid:19) =(cid:18)γ + n + 1
γ
(cid:19) =(cid:18)γ + n + 1
n + 1 (cid:19),
which gives (A.3) when j = n + 1. This proves (A.3), and the result follows.
(cid:3)
Lemma A.2. Let α ≥ 1 be an integer, s0 ∈ (0, 1], and let Ωk,α be the same as in
(A.1). Then there is a constant C which is independent of α such that
Proof. By Lemma A.1 and the fact that s0 − 1 < 0 we get
(β!)s0−1 ≤ 16α.
αXk=1
αXk=1
1
k Xβ∈Ωk,α
Xβ∈Ωk,α
αXk=1
1
k Xβ∈Ωk,α
β!s0−1 ≤
1 =
αXk=1(cid:18)α + k
k (cid:19) ≤ (1 + α)22α ≤ 16α. (cid:3)
References
[1] T. Aoki Locally bounded topological spaces, Proc. Japan Acad. 18 (1942), 588 -- 594.
[2] H.A. Biagioni,T. Gramchev Fractional derivative estimates in Gevrey classes, global regu-
larity and decay for solutions to semilinear equations in Rd. J. Differential Equations 194
(2003), 140 -- 165.
[3] J. M. Bony, J. Y. Chemin Espaces Functionnels Associ´es au Calcul de Weyl-Hormander,
Bull. Soc. math. France 122 (1994), 77 -- 118.
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 41
[4] M. Cappiello, J. Toft Pseudo-differential operators in a Gelfand-Shilov setting, Math. Nachr.
290 (2017), 738 -- 755.
[5] Y. Chen, M. Signahl, J. Toft Factorizations and singular value estimates of operators with
Gelfand-Shilov and Pilipovi´c kernels, J. Fourier Anal. Appl. (appeared online 2017).
[6] J. Chung, S.Y. Chung, D. Kim Characterization of the Gelfand-Shilov spaces via Fourier
transforms. Proc. Amer. Math. Soc. 124 (1996), 2101 -- 2108.
[7] 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.
[8] E. Cordero, K. Grochenig Time-Frequency Analysis of Localization Operators, J. Funct.
Anal. (1) 205 (2003), 107 -- 131.
[9] E. Cordero, S. Pilipovi´c, L. Rodino, N. Teofanov Localization operators and exponential
weights for modulation spaces. Mediterr. J. Math. 2 (2005), 381 -- 394.
[10] E. Cordero, J. Toft, P. Wahlberg Sharp results for the Weyl product on modulation spaces,
J. Funct. Anal. 267 (2014), 3016 -- 3057.
[11] H. G. Feichtinger Banach convolution algebras of Wiener's type, in: Proc. Functions, Se-
ries, Operators in Budapest, Colloquia Math. Soc. J. Bolyai, North Holland Publ. Co.,
Amsterdam Oxford NewYork, 1980.
[12] H. G. Feichtinger Modulation spaces on locally compact abelian groups. In Proceedings of
"International Conference on Wavelets and Applications" 2002, pages 99 -- 140, Chennai,
India, 2003. Updated version of a technical report, University of Vienna, 1983.
[13] H. G. Feichtinger Atomic characterizations of modulation spaces through Gabor-type repre-
sentations, in: Proc. Conf. on Constructive Function Theory, Rocky Mountain J. Math. 19
(1989), 113 -- 126.
[14] H. G. Feichtinger Wiener amalgams over Euclidean spaces and some of their applications,
in: Function spaces (Edwardsville, IL, 1990), Lect. Notes in pure and appl. math., 136,
Marcel Dekker, New York, 1992, pp. 123 -- 137.
[15] H. G. Feichtinger Modulation spaces: Looking back and ahead, Sampl. Theory Signal Image
Process. 5 (2006), 109 -- 140.
[16] H. G. Feichtinger and K. H. Grochenig Banach spaces related to integrable group represen-
tations and their atomic decompositions, I, J. Funct. Anal. 86 (1989), 307 -- 340.
[17] C. Fernandez, A. Galbis, J. Toft Spectral invariance for matrix algebras, J. Fourier Anal.
Appl. 20 (2014), 362 -- 383.
[18] H. G. Feichtinger and K. H. Grochenig Banach spaces related to integrable group represen-
tations and their atomic decompositions, II, Monatsh. Math. 108 (1989), 129 -- 148.
[19] G. B. Folland Harmonic analysis in phase space, Princeton Univ. Press, Princeton, 1989.
[20] Y. V. Galperin, S. Samarah Time-frequency analysis on modulation spaces M p,q
m , 0 < p, q ≤
∞, Appl. Comput. Harmon. Anal. 16 (2004), 1 -- 18.
[21] K. H. Grochenig Foundations of Time-Frequency Analysis, Birkhauser, Boston, 2001.
[22] K. H. Grochenig Composition and spectral invariance of pseudodifferential operators on
modulation spaces, J. Anal. Math. 98 (2006), 65 -- 82.
[23] K. H. Grochenig Time-frequency analysis of Sjostrand's class, Rev. Mat. Iberoam. 22 (2006),
703 -- 724.
[24] 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.
[25] K. Grochenig and T. Strohmer. Pseudodifferential operators on locally compact abelian
groups and Sjostrand's symbol class, J. Reine Angew. Math., 613:121 -- 146, 2007.
[26] K. Grochenig, J. Toft Isomorphism properties of Toeplitz operators and pseudo-differential
operators between modulation spaces, J. Anal. Math. 114 (2011), 255 -- 283.
[27] K. Grochenig, J. Toft The range of localization operators and lifting theorems for modulation
and Bargmann-Fock spaces, Trans. Amer. Math. Soc. 365 (2013), 4475 -- 4496.
[28] K.H. Grochenig, G. Zimmermann Spaces of test functions via the STFT. J. Funct. Spaces.
Appl. 2 (2004), 25 -- 53.
[29] H. Gzyl Multidimensional Extension of Fa`a di Bruno's formula, J. Math. Anal. Appl. 116
(1986), 450 -- 455.
[30] L. Hormander The Weyl calculus of pseudo-differential operators, Comm. Pure Appl. Math.
32 (1979), 359 -- 443.
[31] L. Hormander The Analysis of Linear Partial Differential Operators, vol I, III, Springer-
Verlag, Berlin Heidelberg NewYork Tokyo, 1983, 1985.
[32] Z. Lozanov-Crvenkovi´c, D. Perisi´c, M. Taskovi´c Gelfand-Shilov spaces structural and kernel
theorems, (preprint), arXiv:0706.2268v2.
[33] N. Lerner The Wick calculus of pseudo-differential operators and some of its applications,
CUBO, 5 (2003), 213 -- 236.
42
AHMED ABDELJAWAD, SANDRO CORIASCO, AND JOACHIM TOFT
[34] N. Lerner Metrics on the Phase Space and Non-Selfadjoint Pseudo-Differential Operators,
Birkhauser Springer, (2010).
[35] B.S. Mitjagin Nuclearity and other properties of spaces of type S , Amer. Math. Soc. Transl.,
Ser. 93 (1970), 49 -- 59.
[36] C. Pfeuffer, J. Toft Compactness properties for modulation spaces (under preparation).
[37] S. Pilipovi´c Tempered ultradistributions, Boll. Un. Mat. Ital., VII Ser., 2 (1988), 235 -- 251.
[38] S. Pilipovi´c, N. Teofanov Pseudodifferential operators on ultramodulation spaces, J. Funct.
Anal. 208 (2004), 194 -- 228.
[39] S. Rolewicz On a certain class of linear metric spaces, Bull. Acad. Polon. Sci. S´er. Schi.
Math. Astronom. Phys. 5 (1957), 471 -- 473.
[40] M. Ruzhansky, M. Sugimoto, N. Tomita, J. Toft Changes of variables in modulation and
Wiener amalgam spaces, Math. Nachr. 284 (2011), 2078 -- 2092.
[41] M. Ruzhansky, N. Tokmagambetov Nonharmonic analysis of boundary value problems, Int.
Math. Res. Notices 12 (2016), 3548 -- 3615.
[42] J. Sjostrand An algebra of pseudodifferential operators, Math. Res. L. 1 (1994), 185 -- 192.
[43] J. Sjostrand. Pseudodifferential operators and weighted normed symbol spaces, Serdica Math.
J. 34 (2008), no. 1, 1 -- 38.
[44] K. Tachizawa The boundedness of pseudo-differential operators on modulation spaces, Math.
Nachr. 168 (1994), 263 -- 277.
[45] N. Teofanov Modulation spaces, Gelfand-Shilov spaces and pseudodifferential operators,
Sampl. Theory Signal Image Process, 5 (2006), 225 -- 242.
[46] J. Toft Continuity Properties for non-Commutative Convolution Algebras with Applications
in Pseudo-Differential Calculus, Bull. Sci. Math. (2) 126 (2002), 115 -- 142.
[47] J. Toft Continuity properties for modulation spaces with applications to pseudo-differential
calculus, II, Ann. Global Anal. Geom., 26 (2004), 73 -- 106.
[48] J. Toft Continuity and Schatten properties for pseudo-differential operators on modulation
spaces in: J. Toft, M. W. Wong, H. Zhu (Eds) Modern Trends in Pseudo-Differential Op-
erators, Operator Theory: Advances and Applications 172, Birkhauser Verlag, Basel, 2007,
pp. 173 -- 206.
[49] J. Toft Continuity and Schatten properties for Toeplitz operators on modulation spaces in: J.
Toft, M. W. Wong, H. Zhu (Eds) Modern Trends in Pseudo-Differential Operators, Operator
Theory: Advances and Applications 172, Birkhauser Verlag, Basel, 2007, pp. 313 -- 328.
[50] J. Toft Pseudo-differential operators with smooth symbols on modulation spaces, Cubo. 11
(2009), 87 -- 107.
[51] J. Toft Multiplication properties in pseudo-differential calculus with small regularity on the
symbols, J. Pseudo-Differ. Oper. Appl. 1 (2010), 101 -- 138.
[52] J. Toft The Bargmann transform on modulation and Gelfand-Shilov spaces, with applica-
tions to Toeplitz and pseudo-differential operators, J. Pseudo-Differ. Oper. Appl. 3 (2012),
145 -- 227.
[53] J. Toft Multiplication properties in Gelfand-Shilov pseudo-differential calculus in: S. Mola-
hajlo, S. Pilipovi´c, J. Toft, M. W. Wong (eds) Pseudo-Differential Operators, Generalized
Functions and Asymptotics, Operator Theory: Advances and Applications 231, Birkhauser,
Basel Heidelberg NewYork Dordrecht London, 2013, pp. 117 -- 172.
[54] J. Toft Gabor analysis for a broad class of quasi-Banach modulation spaces, in: S. Pilipovi´c,
J. Toft (eds) Pseudo-Differential Operators and Generalized Functions, Operator Theory:
Advances and Applications Vol 245, Birkhauser, Basel Heidelberg NewYork Dordrecht Lon-
don, 2015, pp. 249 -- 278.
[55] J. Toft Matrix parameterized pseudo-differential calculi on modulation spaces in: M. Ober-
guggenberger, J. Toft, J. Vindas, P. Wahlberg (eds), Generalized functions and Fourier
analysis, Operator Theory: Advances and Applications Vol 260, Birkhauser/Springer, Basel,
2017, pp. 215 -- 235.
[56] J. Toft Continuity of Gevrey-Hormander pseudo-differential operators on modulation
spaces, (preprint) arXiv:1710.11366 (2017).
[57] J. Toft, P. Boggiatto, Schatten classes for Toeplitz operators with Hilbert space windows on
modulation spaces, Adv. Math. 217 (2008), no. 1, 305 -- 333.
[58] J. Toft, A. Khrennikov, B. Nilsson, S. Nordebo Decompositions of Gelfand-Shilov kernels
into kernels of similar class, J. Math. Anal. Appl. 396 (2012), 315 -- 322.
LIFTINGS FOR MODULATION SPACES, AND ONE-PARAMETER GROUPS OF ΨDO 43
Dipartimento di Matematica "G. Peano", Universit´a degli Studi di Torino
E-mail address: [email protected]
Dipartimento di Matematica "G. Peano", Universit´a degli Studi di Torino
E-mail address: [email protected]
Department of Mathematics, Linnaeus University, Sweden
E-mail address: [email protected]
|
1309.3146 | 1 | 1309 | 2013-09-12T13:09:06 | Further results on regular Fredholm pairs and chains | [
"math.FA"
] | The main objective of the present article is to characterize regular Fredholm pairs and chains in terms of Fredholm operators. | math.FA | math |
FURTHER RESULTS ON REGULAR FREDHOLM
PAIRS AND CHAINS
ENRICO BOASSO
The main objective of the present article is to characterize regular Fredholm pairs and
chains in terms of Fredholm operators.
AMS 2000 Subject Classification: Primary 47A13; Secondary 47A53.
Key words: Fredholm pairs and chains, regular operators.
1. INTRODUCTION
In multiparameter operator theory there exist objects that in the context
of Hilbert spaces can be characterized using linear and bounded maps, to
mention only some of the works related to this subject, see for example [9, 10,
11, 12]. Since the adjoint plays a key role in this relationship between one and
several variable operator theory, it is not possible to directly translate these
results to the frame of Banach spaces. However in [7] R. Harte and W. Y.
Lee showed that generalized inverses could be used to reformulate the Hilbert
space situation for regular Fredholm Banach space chains and complexes.
On the other hand, Fredholm pairs and chains have been recently intro-
duced and their main properties have been studied, see [1, 2, 4, 5, 8] and
the monograph [3]. These objects consist in generalizations of the notions of
Fredholm operators and Fredholm Banach space complexes respectively. In
particular, in [4] Fredholm pairs and chains in Hilbert spaces were charac-
terized using the ideas mentioned in the previous paragraph. Furthermore,
in [5] regular Fredholm pairs, that is Fredholm pairs whose operators admit
generalized inverses, were characterized and classified.
The main objective of the present article is to characterize regular Fredholm
pairs and chains in Banach spaces using Fredholm operator extending to these
objects the characterizations of [4] using the approach of R. Harte and W. Y.
Lee in [7]. Naturally, as in [7], since the adjoint can not be considered any
more, suitable generalized inverses must be defined to prove the main results
of this work.
The article is organized as follows. In the next section the definitions of
the objects under consideration will be recalled. In section 3 two character-
izations of regular Fredholm pairs will be given. Finally, in section 4 the
results of section 3 will be applied to regular Fredholm chains to obtain two
characterizations of them.
1
2
ENRICO BOASSO
2. PRELIMINARY DEFINITIONS
From now on X and Y will denote two Banach spaces and L(X, Y ) will
stand for the algebra of all linear and continuous operators defined on X with
values in Y . As usual, when X = Y , L(X, X) is denoted by L(X). For every
S ∈ L(X, Y ), the null space of S is denoted by N (S) = {x ∈ X : S(x) = 0},
and the range of S by R(S) = {y ∈ Y : ∃ x ∈ X such that y = S(x)}.
Next follow the definitions of Fredholm pairs and chains, see for instance
[1, 3, 8].
Definition 2.1. Let X and Y be two Banach spaces and let S ∈ L(X, Y )
and T ∈ L(Y, X) be such that the following dimensions are finite:
(i) a : = dim N (S)/(N (S) ∩ R(T )), b : = dim R(T )/(N (S) ∩ R(T )),
(ii) c : = dim N (T )/(N (T ) ∩ R(S)), d : = dim R(S)/(N (T ) ∩ R(S)).
A pair (S, T ) with the above properties is said to be a Fredholm pair.
Let P (X, Y ) denote the set of all Fredholm pairs. If (S, T ) ∈ P (X, Y ), then
the index of (S, T ) is defined by the equality
ind(S, T ) : = a − b − c + d.
In particular, if (S, T ) ∈ P (X, Y ) is such that ST = 0 and T S = 0, that
is if b = d = 0, then (S, T ) and (T, S) are Fredholm chains in the sense of [6,
section 10.6] and [7].
Definition 2.2. A Fredholm chain (Xp, δp)p∈Z is a sequence of Banach spaces
Xp and bounded operators δp ∈ L(Xp, Xp−1) such that there is a natural
number n with the property Xp = 0, p < 0 and p ≥ n + 1, δp = 0, p ≤ 0 and
p > n, and
N (δp)/(N (δp) ∩ R(δp+1)) and R(δp+1)/(N (δp) ∩ R(δp+1))
are finite dimensional subspaces of Xp, p ∈ Z.
Given a Fredholm chain, it is possible to associate an index to it. In fact,
if (Xp, δp)p∈Z is such an object, then define
ind(Xp, δp)p∈Z =
X
(−1)pdp,
n
p=0
where dp = dim N (δp)/(N (δp) ∩ R(δp+1)) − dim R(δp+1)/(N (δp) ∩ R(δp+1)),
see [8].
Recall that in [8] the more general concept of semi-Fredholm chain was in-
troduced. However, since the main concern of this article consists in Fredholm
objects, only Fredholm chains will be considered.
Remark 2.3. There is a natural relationship between Fredholm pairs and
In fact, given as in Definition 2.2 a sequence of spaces and maps
chains.
(Xp, δp)p∈Z , consider the Banach spaces
X = M
Xp,
Y = M
Xp,
p=2k
p=2k+1
FREDHOLM PAIRS AND CHAINS
3
and the Banach space operators S ∈ L(X, Y ) and T ∈ L(Y, X) defined by
S = M
δp,
p=2k
T = M
δp,
p=2k+1
where Xp = 0, p < 0 and p ≥ n + 1, δp = 0, p ≤ 0 and p > n, and n is a
natural number.
Now well, it is not difficult to prove that dim R(ST ) and dim R(T S) are
finite dimensional if and only if dim R(δp+1)/(N (δp) ∩ R(δp+1)) are finite
dimensional, p = 0, . . ., n. Furthermore, a straightforward calculation shows
that necessary and sufficient for N (S)/(N (S)∩R(T )) and N (T )/(N (T )∩R(S))
to be finite dimensional is the fact that N (δp)/(N (δp) ∩ R(δp+1)) are finite
dimensional, p = 0, . . ., n. Consequently, (Xp, δp)p∈Z is a Fredholm chain if
and only if (S, T ) is a Fredholm pair. Finally, in this case,
ind(Xp, δp)p∈Z = ind(S, T ),
see [4, Remark 2.4].
On the other hand, recall that an operator T ∈ L(X, Y ) is called regular,
if there is S ∈ L(Y, X) such that T = T ST . The map S is said to be a
generalized inverse of T . In addition, if T is also a generalized inverse of S,
that is if S = ST S, then S is said to be a normalized generalized inverse of T .
Note that if T is regular, then T always has a normalized generalized inverse.
In fact, if S is a generalized inverse of T , then S ′ = ST S is a normalized
generalized inverse of T .
In the following definition, the notions of regular
Fredholm pairs and chains will be recalled, see [5].
Definition 2.4. Let X and Y be two Banach spaces and consider S ∈
L(X, Y ) and T ∈ L(Y, X) such that (S, T ) ∈ P (X, Y ). The pair (S, T ) will be
said to be a regular Fredholm pair, if S and T are regular operators. Similarly,
given a Fredholm chain (Xp, δp)p∈Z , then (Xp, δp)p∈Z will be said to be a
regular Fredholm chain, if the Fredholm pair defined by (Xp, δp)p∈Z is regular.
Note that given (S, T ) ∈ P (X, Y ), several statements equivalent to the fact
that (S, T ) is a regular Fredholm pair were considered in [5, Proposition 2.4].
Concerning regular Fredholm chains, see [5, Remark 2.7].
3. CHARACTERIZATIONS OF REGULAR FREDHOLM PAIRS
First of all, a preliminary remark is presented.
Remark 3.1. Let X and Y be two Banach spaces, and let S ∈ L(X, Y )
and T ∈ L(Y, X) be two operators such that R(ST ) and R(T S) are finite
dimensional subspaces of Y and X respectively. Then, define the Banach
spaces X = X/R(T S) and Y = Y /R(ST ), and the linear and bounded maps
S ∈ L(X , Y) and T ∈ L(Y, X ) as the factorization of S and T through the
respective invariant subspaces. Clearly, S T = 0 and T S = 0, that is ( S, T )
and ( T , S) are chains in the sense of [7]. Moreover, the regularity of S and T
is equivalent to the one of S and T .
4
ENRICO BOASSO
THEOREM 3.2. In the same conditions of Remark 3.1, the following state-
ments are equivalent.
(i) S and T are regular operators.
(ii) S and T are regular operators.
Proof. Since S ∈ L(X, Y ) is a regular operator, according to [6, Theorem
3.8.2], there is M a linear subspace of Y such that R(S) ⊕ M = Y . Conse-
quently, R( S) + πY (M ) = Y, where πY : Y → Y is the quotient map.
Next suppose that there is x ∈ X such that S(x) = m ∈ πY (M ).
In
particular, there exist x ∈ X, m ∈ M , and y ∈ Y such that πX(x) = x,
πY (m) = m, and S(x) − m = ST (y), where πX : X → X is the quotient map.
However, in this case S(x − T (y)) = m. Then, m ∈ M ∩ R(S) = 0, and
S(x) = m = 0. Therefore, R( S) ⊕ πY (M ) = Y.
On the other hand, according to [6, Theorem 3.8.2], there exists N, a vector
subspace of X such that N (S) ⊕ N = X. However, since
(N (S) + R(T ))/N (S) ∼= R(T ))/(N (S) ∩ R(T )) ∼= R(ST ),
there is a finite dimensional subspace X1 such that N (S)⊕X1 = N (S)+R(T ).
Then, a straightforward calculation proves that there exists a closed vector
subspace R of X such that (N (S) + R(T )) ⊕ R = X.
Now well, according to [1, Remark 2.1], N ( S) = πX(N (S) + R(T )).
particular, N ( S) + πX (R) = X .
In
Next suppose that there are n ∈ N (S), y ∈ Y , and r ∈ R such that
πX(n + T (y)) = πX(r). Then, there is x0 ∈ X such that n + T (y) − r =
T S(x0). In particular, n + T (y − S(x0)) = r. Thus r = 0, which implies that
πX(r) = πX(n + T (x)) = 0. Therefore, N ( S) ⊕ πX(R) = X . As a result,
according to what has been proved and to [6, Theorem 3.8.2], the operator S
is regular.
Interchanging S with T and S with T , it can be proved that T is regular.
In order to prove the converse implication, suppose that S and T are regular
operators.
Consider V, a closed vector subspace of Y, such that R( S) ⊕ V = Y. Let
V1 = π−1
Y (V)∩R(ST ). Since V1 has finite dimension, there is a vector subspace
W1 such V1 ⊕ W1 = π−1
Y (V). Moreover, since V1 ⊆ R(ST ) and πY is surjective,
V = πY (W1). Consequently, πY (R(S) + W1) = R( S) + V = Y, which implies
that R(S)+W1 +R(ST ) = Y . However, since R(ST ) ⊆ R(S), R(S)+W1 = Y .
Next define L = R(S) ∩ W1. Then, πY (L) ⊆ R( S) ∩ V = 0. Thus, L ⊆
R(ST ). However, since W1 ∩ R(ST ) = 0, L = 0 and R(S) ⊕ W1 = X.
Similarly, suppose that U is a closed subspace of X such that N ( S) ⊕ U =
X (U ) ∩ R(T S). Since U1 is a finite dimensional subspace,
X (U ). Furthermore, since
X . Let U1 = π−1
there exists a subspace Z1 such that U1 ⊕ Z1 = π−1
πX : X → X is a surjecive map, πX(Z1) = U .
FREDHOLM PAIRS AND CHAINS
5
Now well, according to [1, Remark 2.1], πX(N (S)+R(T )+Z1) = πX (N (S)+
R(T )) + πX (Z1) = N ( S) + U = X . Therefore, since R(T S) ⊆ R(T ), N (S) +
R(T ) + Z1 = X.
Let P = (N (S) + R(T )) ∩ Z1. According to [1, Remark 2.1], πX(P ) ⊆
N ( S)∩U = 0. Thus, P ⊆ R(T S). Consequently, since Z1 ∩R(T S) = 0, P = 0,
and (N (S) + R(T )) ⊕ Z1 = X. However, since N (S) ⊕ X1 = N (S) + R(T ),
where X1 is a finite dimensional vector subspace of X, N (S) ⊕ (X1 + Z1) = X,
which, according to what has been proved and to [6, Theorem 3.8.2], implies
that S is a regular operator.
operator.
Interchanging S with T and S with T , it can be proved that T is a regular
(cid:3)
In order to prove the first characterization concerning regular Fredholm
pairs, some preparation is needed.
Remark 3.3. Let X, Y , S, T , X , Y, S, and T be as in Remark 3.1. Recall
that S T = 0 and T S = 0. Furthermore, according to Theorem 3.2, necessary
and sufficient for S and T to be regular bounded and linear maps is the fact
that ( S, T , S) is a regular chain in the sense of [7, p. 283-284]. In addition,
according again to [7, p. 284], there exist S ′ ∈ L(Y, X ) and T ′ ∈ L(X , Y) such
that ( T ′, S ′, T ′) is a regular chain in the sense of [7] with the property that S ′
(respectively T ′) is a normalized generalized inverse of S (respectively T ).
Now well, since R(ST ) and R(T S) are finite dimensional subspaces of Y
and X respectively, X and Y are isomorphic to finite codimensional closed
subspaces of X and Y respectively. Consequently, using Banach space iso-
morphisms, S ′ and T ′ can be thought of operators defined on and to finite
codimensional closed subspaces of X and Y . Donote by S ′ ∈ L(Y, X) and
T ′ ∈ L(X, Y ) the extension of S ′ and T ′ respectively, such that S ′ = 0 on
R(T S) and T ′ = 0 on R(ST ). Note that S ′ and T ′ are regular maps.
Next follows the first characterization of regular Fredholm pairs.
THEOREM 3.4. In the same conditions of Remark 3.3, the following state-
ments are equivalent.
(i) (S, T ) is a regular Fredholm pair.
(ii) S + T ′ ∈ L(X, Y ) is a Fredholm operator.
(iii) T + S ′ ∈ L(Y, X) is a Fredholm operator.
Futhermore, in this case
ind(S, T ) = ind(S + T ′) = − ind(T + S ′).
Proof. According to [1, Remark 2.1] and to Theorem 3.2, (S, T ) is a regular
Fredholm pair if and only if ( S, T , S) is a Fredholm regular chain in the sense
of [7]. Then, according to [7, Theorem 5], the first statement of the Theorem
is equivalent to the fact that S + T ′ ∈ L(X , Y) is a Fredholm operator.
Now well, since as in Remark 3.3 X and Y can be thought of finite codi-
mensional closed subspaces of X and Y respectively, it is possible to define
S1 ∈ L(Y, X) as the extension of S with the property S1 = 0 on R(T S). Con-
sequently, since R(ST ) and R(T S) are finite dimensional vector subspaces,
6
ENRICO BOASSO
according to what has been proved, the first statement of the Theorem is
equivalent to the fact that S1 + T ′ ∈ L(Y, X) is a Fredholm operator. How-
ever, since S − S1 is a compact operator, the first and the second statements
of the Theorem are equivalent.
As regard the index, according to [1, Remark 2.1] and to [7, Theorem 5]
ind(S, T ) − dim R(T S) + dim R(ST ) = ind( S, T ) = ind( S + T ′).
ind(S, T ) = ind(S + T ′).
On the other hand, note that
ind(S + T ′) = ind(S1 + T ′) = ind( S + T ′) + dim R(T S) − dim R(ST ).
Therefore,
A similar argument proves that the first and the third statements are equiv-
(cid:3)
The second characterization of regular Fredholm pairs is more general in
the sense that the generalized inverses can be chosen more freely, however, the
formula regarding the index can not be considered. On the other hand, to
state this characterization, a Banach space operator is introduced.
alent as well as the relationship between the indexes.
Remark 3.5 Let X, Y , S, T , X , Y, S, and T be as in Remark 3.3. Since
S, and T are regular operators, there exist S ′ ∈ L(Y, X ) and T ′ ∈ L(X , Y)
generalized inverses of S and T respectively.
In addition, proceeding as in
Remark 3.3, denote by S ′ ∈ L(Y, X) and T ′ ∈ L(X, Y ) any extensions of S ′
and T ′ respectively.
On the other hand, define the map V ∈ L(X⊕Y ) as follows: V X ∈ L(X, Y ),
V X = S + T ′, and V Y ∈ L(Y, X), V Y = T + S ′.
THEOREM 3.6. In the same conditions of Remark 3.5, the following state-
ments are equivalent.
(i) (S, T ) is a regular Fredholm pair.
(ii) S ′S + T T ′ and T ′T + SS ′ are Fredholm operators.
(iii) V is a Fredholm operator.
Proof. As in Theorem 3.4, according to [1, Remark 2.1] and to Theorem 3.2,
the first statement is satisfied if and only if ( S, T , S) is a Fredholm regular
chain in the sense of [7], which, according to [7, Theorem 3], is equivalent to
the fact that S ′ S + T T ′ ∈ L(X ) and T ′ T + S S ′ ∈ L(Y) are Fredholm operators.
Now well, since R(T S) and R(ST ) are finite dimensional subspaces of X and Y
respectively, it is not difficult to prove that what has been proved is equivalent
to the second statement of the Theorem.
On the other hand, V is a Fredholm operator if and only if V 2 also is
Fredholm. Now well, a straightforward calculation proves that there exists a
finite range operator F ∈ L(X ⊕ Y ) such that V 2 − F is a diagonal operator
with entries S ′S + T T ′ ∈ L(X) and SS ′ + T T ′ ∈ L(Y ), which clearly implies
that the second and the third statements are equivalent.
(cid:3)
FREDHOLM PAIRS AND CHAINS
7
4. CHARACTERIZATIONS OF REGULAR FREDHOLM CHAINS
In order to prove the main results of this section, some preliminary facts
will be considered.
Remark 4.1 Let (Xp, δp)p∈Z be a sequence of spaces and maps such that
Xp = 0, p < 0 and p ≥ n + 1, and δp = 0, p ≤ 0 and p > n, where n is a
natural number. In addition, suppose that R(δpδp−1) is finite dimensional and
δp ∈ L(Xp, Xp−1) is a regular operator, p ∈ Z. Then, if X, Y , S ∈ L(X, Y ),
and T ∈ L(Y, X) are as in Remark 2.3, the properties of (Xp, δp)p∈Z are
equivalent to the fact that R(ST ) and R(T S) are finite dimensional and S
and T are regular operators.
Now well, if X , Y, S and T are as in Remark 3.3, note that
X = M
Xp,
p=2k
S = M
δp,
p=2k
Y = M
Xp,
p=2k+1
T = M
δp,
p=2k+1
where Xp = Xp/R(δp+1δp+2) and δp : Xp → Xp−1 is the quotient map induced
by δp : Xp → Xp−1, p ∈ Z. Since S T = 0 and T S = 0, the sequence of spaces
and maps (Xp, δp)p∈Z is a finite complex of Banach spaces. In addition, ac-
cording to Theorem 3.2, S and T are regular operators, which is equivalent
to the fact that δp : Xp → Xp−1 is regular, p ∈ Z. A straightforward calcula-
tion proves that there exists a sequence of linear and bounded maps (δ′
p)p∈Z,
p is a normalized generalized inverse of δp and
p : Xp−1 → Xp such that δ′
δ′
(Xp, δ′
p)p∈Z is a finite complex of Banach spaces. Consequently, if
S ′ = M
δ′
p,
p=2k
T ′ = M
δ′
p,
p=2k+1
then ( T ′, S ′, T ′) is regular chain in the sense of [7] such that S ′ and T ′ are
normalized generalized inverses of S and T . Finally, as in Remark 3.3, consider
S ′ ∈ L(Y, X) and T ′ ∈ L(X, Y ), the extensions of S ′ and T ′ such that S ′ = 0
on R(T S) and T ′ = 0 on R(ST ), and denote by δ′
p : Xp−1 → Xp the extension
of δ′
p such that δ′
THEOREM 4.2. In the same conditions of Remark 4.1, the following state-
p = 0 on R(δpδp−1), p ∈ Z.
ments are equivalent.
(i) (Xp, δp)p∈Z is a regular Fredholm chain.
(ii) ⊕p=2k(δp + δ′
p+1) is a Fredholm operator.
(iii) ⊕p=2k+1(δp + δ′
Furthermore, in this case,
p+1) is a Fredholm operator.
ind(Xp, δp)p∈Z = ind ⊕p=2k (δp + δ′
p+1) = − ind ⊕p=2k+1 (δp + δ′
p+1).
Proof. Apply Remark 2.3, Theorem 3.4, and Remark 4.1.
(cid:3)
8
ENRICO BOASSO
Remark 4.3 In the same conditions of Remark 4.1, consider the same spaces
and maps but, as in Remark 3.5, S ′ and T ′ denote any generalized inverses of
S and T . In addition, S ′ and T ′ can be any extentions of S ′ and T ′. Similarly,
p is any extension of δ′
δ′
THEOREM 4.4. In the conditions of Remark 4.3, the following statements
p, p ∈ Z.
are equivalent.
(i) (Xp, δp)p∈Z is a regular Fredholm chain.
(ii) δp+1δ′
pδp ∈ L(Xp) is a Fredholm operator, p ∈ Z.
p+1 + δ′
Proof. Apply Remark 2.3, Theorem 3.6, and Remarks 4.1 and 4.3.
(cid:3)
References
[1] C.-G. Ambrozie, On Fredholm index in Banach spaces, Integral Equations Operator
Theory 25 (1996), 1-34.
[2] C.-G. Ambrozie, The Euler characteristic is stable under compact perturbations, Proc.
Amer. Math. Soc. 124 (7) (1996), 2041-2050.
[3] C.-G. Ambrozie and F.-H. Vasilescu, Banach Space Complexes, Kluwer Academic Plub-
lishers, Dordrecht - Boston - London, 1995.
[4] E. Boasso, Characterizations of Fredholm pairs and chains in Hilbert spaces, Rev.
Roumaine Math. Pures Appl. 51 (2) (2006), 151-165.
[5] E. Boasso, Regular Fredholm pairs, J. Operator Theory 55 (2) (2006), 311-337.
[6] R. Harte, Invertibility and Singularity for Bounded Linear Operators, Marcel Dekker,
Inc., New York and Basel 1988.
[7] R. Harte and W. Y. Lee, An index formula for chains, Studia Math. 116 (3) (1995),
283-294.
[8] V. Muller, Stability of index for semi-Fredholm chains, J. Operator Theory 37 (1997),
247-261.
[9] F.-H. Vasilescu A characterization of the joint spectrum in Hilbert spaces, Rev.
Roumaine Math. Pures Appl. 22 (1977), 1003-1009.
[10] F.-H. Vasilescu On pairs of commuting operators, Studia Math. 62 (1978), 203-207.
[11] F.-H. Vasilescu Analytic perturbations of the δ-operator and integral representation
fromulas in Hilbert spaces, J. Operator Theory 1 (1979), 187-205.
[12] F.-H. Vasilescu The stability of the Euler characteristic for Hilbert complexes, Math.
Ann. 248 (1980), 109-116.
Enrico Boasso
E-mail address: enrico [email protected]
|
1710.05286 | 2 | 1710 | 2017-10-27T07:43:29 | Generalization of The Results on Fixed Point For Coupling on Metric Spaces | [
"math.FA"
] | The purpose of this paper is to introduce the concept of self-cyclic maps, g-coupling and Banach type g-coupling which is the generalization of couplings introduced by Choudhury et al. In our main result we prove the existence theorem of coupled coincidence point for Banach type g-couplings which extends the results by Choudhury et al. We give an example in support of our result. | math.FA | math |
Generalization of The Results on Fixed Point
For Couplings on Metric Spaces
Tawseef Rashid 1 and Q. H. Khan 2
The purpose of this paper is to introduce the concept of self-cyclic maps,
g-coupling, Banach type g-coupling which is the generalization of couplings in-
troduced by Choudhury et al.[5].
In our main result we prove the existence
theorem of coupled coincidence point for Banach type g-couplings which extends
the results by choudhury et al. [5]. We give an example in support of our result.
Keywords : g-coupling, coupled coincidence point, strong coupled coinci-
dence point, self-cyclic maps, Banach type g-coupling.
2010MSC: 47H10, 54H25.
1. Introduction and Preliminaries
The concept of coupled fixed point was introduced in the work of Guo et al.
[10]. Then after the coupled contraction mapping theorem by Bhaskar and Lak-
shmikantham [9], and the introduction of coupled coincidence point by Laksh-
mikantham.V and ´Ciri´c.L [13], the coupled fixed and coupled coincidence point
results reported in papers [2, 3, 4, 6, 7, 11, 13, 14, 15, 16, 17, 18, 19, 20].
Kirk et al.
[12] gave the concept of cyclic mappping. Recently Choudhury
et al.[5] introduced the concept of couplings which are actually coupled cyclic
mappings with respect to two given subsets of a metric space. Choudhury et
al. [5] proved the existence of strong coupled fixed point for a coupling ( w.r.t.
subsets of a complete metric space). H. Aydi et al. [1] proved the theorem for
existence and uniqueness of strong coupled fixed point in partial metric spaces.
In this paper we extend the concept of Banach type coupling to Banach type
g-coupling by extending the concept of coupling to g-coupling. We have intro-
duced self-cyclic maps and generalize the results by Choudhury et al [5]. Now
we recall some definitions.
Definition 1.1 (Coupled fixed point) [9]. An element (x, y) ∈ X ×X, where X is
any non-empty set, is called a coupled fixed point of the mapping F : X×X → X
if F (x, y) = x and F (y, x) = y.
Definition 1.2 (Strong coupled fixed point) [8]. An element (x, y) ∈ X × X,
where X is any non-empty set, is called a strong coupled fixed point of the
mapping F : X × X → X if (x, y) is coupled fixed point and x = y; that is if
F (x, x) = x.
1Department of Mathematics, Aligarh Muslim University, Aligarh-202002, India.
Email : [email protected]
2Department of Mathematics, Aligarh Muslim University, Aligarh-202002, India.
Email : [email protected]
1
Definition 1.3 (Coupled Banach Contraction Mapping) [9]. Let (X, d) be a
metric space. A mapping F : X × X → X is called coupled Banach contraction
if there exists k ∈ (0, 1) s.t ∀ (x, y), (u, v) ∈ X × X, the following inequality is
satisfied:
d(F (x, y), F (u, v)) ≤
k
2
[d(x, u) + d(y, v)].
Definition 1.4 (Cyclic mapping) [12]. Let A and B be two non-empty subsets
of a given set X. Any function f : X → X is said to be cyclic (with respect to
A and B) if
f (A) ⊂ B and f (B) ⊂ A.
Definition 1.5 (Coupling) [8]. Let (X, d) be a metric space and A and B be
two non-empty subsets of X. Then a function F : X × X → X is said to be a
coupling with respect to A and B if
F (x, y) ∈ B and F (y, x) ∈ A whenever x ∈ A and y ∈ B.
Definition 1.6 (Banach type coupling) [5]. Let A and B be two non-empty
subsets of a complete metric space (X, d). A coupling F : X × X → X is called
a Banach type coupling with respect to A and B if it satisfies the following
inequality:
d(F (x, y), F (u, v)) ≤
k
2
[d(x, u) + d(y, v)].
where x, v ∈ A, y, u ∈ B and k ∈ (0, 1).
Definition 1.7 (Coupled coincidence point of F and g) [13]. An element (x, y) ∈
X × X is called a coupled coincidence point of the mappings F : X × X → X
and g : X → X if F (x, y) = g(x) and F (y, x) = g(y).
Definition 1.8 (Commutative mappings) [13]. For any set X, we say mappings
F : X × X → X and g : X → X are commutative if
g(F (x, y)) = F (g(x), g(y)), ∀ x, y ∈ X.
2. Main Result
Here we give some definitions in support of our main result.
Definition 2.1 (Self-cyclic mapping). Let A and B be two non-empty subsets
of any set X. Then a mapping g : X → X is said to be self-cyclic (with respect
to A and B) if
g(A) ⊆ A and g(B) ⊆ B.
Definition 2.2 (Strong coupled coincidence points of F and g). A coupled
coincidence point (x, y) ∈ X × X of F : X × X → X and g : X → X is said to
be strong coupled coincidence point of F and g, if x = y i.e. F (x, x) = g(x).
2
Definition 2.3 (g-coupling). Let (X, d) be a metric space and A and B be two
non-empty subsets of X. Let functions F and g are such that F : X × X →
X and g : X → X. Then F is said to be g-coupling (with respect to A and B)
if
F (x, y) ∈ g(A) ∩ B and F (y, x) ∈ g(B) ∩ A, whenever x ∈ A and y ∈ B.
Remark 2.4: It should be noted that every g-coupling is a coupling but converse
is not true in general.
Proof : Let F : X × X → X be g-coupling, where g : X → X. Then by
Definition (2.3), we have F (x, y) ∈ (g(A) ∩ B) ⊂ B and F (y, x) ∈ (g(B) ∩ A) ⊂
A.
this shows that F is a coupling (w.r.t. A and B).
Clearly converse is not true in general. If F is a coupling (w.r.t A and B), it
doesn't imply that F is a g-coupling for any g : X → X (w.r.t. A and B) as
F (x, y) ∈ B doesn't imply that F (x, y) ∈ g(A), and similarly for F (y, x).
Definition 2.5 (Banach type g-coupling). Let A and B be two non-empty
subsets of a complete metric space (X, d). Then a g-coupling F : X × X → X
is said to be Banach type g-coupling (with respect to A and B) if the following
inequality holds:
d(F (x, y), F (u, v)) ≤
k
2
[d(gx, gu) + d(gy, gv)].
y, u ∈ B and k ∈ (0, 1). where g :X → X is a self-cyclic
whenever x, v ∈ A,
mapping (with respect to A and B).
Note : If g = I (the identity mapping) which is also self-cyclic, then Banach
type g-coupling becomes Banach type coupling.
Theorem 2.6: Let A and B be any two subsets of a complete metric space
(X, d). If there exists a Banach type g-coupling F : X × X → X (with respect
to A andB), where g : X → X is self-cyclic (with respect to A and B).
If
g(A) and g(B) are closed subsets of (X, d). Then
(i) g(A) ∩ g(B) 6= ∅,
(ii) F and g have a coupled coincidence point in A × B, i.e. there exists (a, b)
∈ A × B, s.t F (a, b) = g(a) and F (b, a) = g(b).
Proof : Here F : X × X → X is given to be Banach type g-coupling, i.e.
(1)
d(F (x, y), F (u, v)) ≤
k
2
where x, v ∈ A, y, u ∈ B and k ∈ (0, 1).
Also as g : X → X is self-cyclic (w.r.t. A and B), so
[d(gx, gu) + d(gy, gv)].
gx, gv ∈ g(A) ⊆ A and gy, gu ∈ g(B) ⊆ B
Let x0 ∈ A and y0 ∈ B, then by defnition of g-coupling, we have
F (x0, y0) ∈ g(A) ∩ B and F (y0, x0) ∈ g(B) ∩ A
in particular F (x0, y0) ∈ g(A) and F (y0, x0) ∈ g(B).
If
F (x0, y0) = g(x0) and F (y0, x0) = g(y0),
3
then (x0, y0) is the coupled coincidence point of F and g, so we are done in this
case. Otherwise ∃ x1 ∈ A and y1 ∈ B , s.t.
F (x0, y0) = g(x1) and F (y0, x0) = g(y1).
Now if,
F (x1, y1) = g(x1) and F (y1, x1) = g(y1),
then (x1, y1) is a coupled coincidence point of F and g, and we are through
otherwise ∃ x2 ∈ A and y2 ∈ B s.t
F (x1, y1) = g(x2) and F (y1, x1) = g(y2).
continuing in this way, we get sequences {gxn} and {gyn} in g(A) and g(B)
respectively, such that
g(xn+1) = F (xn, yn) and g(yn+1) = F (yn, xn).
(2)
Now using (1) and (2), we get
and
d(gx1, gy2) = d(F (x0, y0), F (y1, x1))
≤
k
2
[d(gx0, gy1) + d(gy0, gx1)]
d(gy1, gx2) = d(F (y0, x0), F (x1, y1))
≤
k
2
[d(gy0, gx1) + d(gx0, gy1)].
from above two inequalities, we have
d(gx1, gy2) + d(gy1, gx2) ≤
k
2
[d(gx0, gy1) + d(gy0, gx1) + d(gy0, gx1) + d(gx0, gy1)]
= k[d(gx0, gy1) + d(gy0, gx1)].
or,
d(gx1, gy2) + d(gy1, gx2)
2
≤
k
2
[d(gx0, gy1) + d(gy0, gx1)].
(3)
using (1), (2) and (3), we have
d(gx2, gy3) = d(F (x1, y1), F (y2, x2))
≤
≤
k
2
k2
2
[d(gx1, gy2) + d(gy1, gx2)]
[d(gx0, gy1) + d(gy0, gx1)],
4
and
d(gy2, gx3) = d(F (y1, x1), F (x2, y2))
≤
≤
k
2
k2
2
[d(gy1, gx2) + d(gx1, gy2)]
[d(gx0, gy1) + d(gy0, gx1)].
Let for some integer n,
d(gxn, gyn+1) ≤
kn
2
[d(gx0, gy1) + d(gy0, gx1)].
kn
2
Now using (1), (4) and (5), we have
d(gyn, gxn+1) ≤
[d(gx0, gy1) + d(gy0, gx1)].
d(gxn+1, gyn+2) = d(F (xn, yn), F (yn+1, xn+1))
[d(gxn, gyn+1) + d(gyn, gxn+1)]
kn[d(gx0, gy1) + d(gy0, gx1)]
≤
≤
=
k
2
k
2
kn+1
2
[d(gx0, gy1) + d(gy0, gx1)].
similarly as above we can show that
d(gyn+1, gxn+2) ≤
kn+1
2
[d(gx0, gy1) + d(gy0, gx1)].
Thus (4) and (5) remains also true for n + 1,
Hence by principle of mathematical induction, we have ∀ n ≥ 1,
d(gxn, gyn+1) ≤
d(gyn, gxn+1) ≤
kn
2
kn
2
[d(gx0, gy1) + d(gy0, gx1)].
[d(gx0, gy1) + d(gy0, gx1)].
Again by (1) and (2), we have
d(gx1, gy1) = d(F (x0, y0), F (y0, x0))
≤
k
2
[d(gx0, gy0) + d(gy0, gx0)]
= kd(gx0, gy0).
(4)
(5)
(6)
(7)
that is,
d(gx1, gy1) ≤ kd(gx0, gy0).
(8)
5
Then from (1), (2) and (8), we have
d(gx2, gy2) = d(F (x1, y1), F (y1, x1))
≤
k
2
[d(gx1, gy1) + d(gy1, gx1)]
= kd(gx1, gy1)
≤ k2d(gx0, gy0).
Let for some integer n, we have
d(gxn, gyn) ≤ knd(gx0, gy0).
(9)
Then from (1), (2) and (9), we get
d(gxn+1, gyn+1) = d(F (xn, yn), F (yn, xn))
≤
k
2
[d(gxn, gyn) + d(gyn, gxn)]
= kd(gxn, gyn)
≤ kn+1d(gx0, gy0).
This shows that (9) remains also true for n + 1, thus by principle of mathemtical
induction, we say that
d(gxn, gyn) ≤ knd(gx0, gy0), ∀ n ≥ 1.
(10)
Now by (6), (7) and (10) and by triangular inequality, we have ∀ n ≥ 1
d(gxn, gxn+1) + d(gyn, gyn+1) ≤ d(gxn, gyn) + d(gyn, gxn+1) + d(gyn, gxn) + d(gxn, gyn+1)
= 2d(gxn, gyn) + [d(gyn, gxn+1) + d(gxn, gyn+1)]
≤ 2knd(gx0, gy0) + kn[d(gx0, gy1) + d(gy0, gx1)].
since k ∈ (0, 1), it follows that P d(gxn, gxn+1) + P d(gyn, gyn+1) < ∞.
Thus sequences {gxn} and {gyn} are Cauchy sequences in g(A) and g(B).
As g(A) and g(B) are closed subsets of complete metric space (X, d), so {g(xn)} and {g(yn)}
are convergent in g(A) and g(B) respectively.
Therefore ∃ u ∈ g(A) and v ∈ g(B), s.t.
gxn → u and gyn → v as n → ∞.
(11)
using (10), as k ∈ (0, 1), we have
d(gxn, gyn) → 0
as n → ∞.
therefore from (11), we have
u = v.
(12)
As u ∈ g(A) and v ∈ g(B) ⇒ u ∈ g(A) ∩ g(B)
This proves part (i) that g(A) ∩ g(B) 6= ∅.
6
Now, since u ∈ g(A) and v ∈ g(B), therefore ∃ a ∈ A and b ∈ B, s.t
u = g(a) and v = g(b).
then from (11) and (12), we have
and
gxn → g(a) and gyn → g(b).
g(a) = g(b).
(13)
(14)
Now by (1), (2), (13), (14) and triangular inequality, we have
d(g(a), F (a, b)) ≤ d(g(a), gyn+1) + d(gyn+1, F (a, b))
= d(g(a), gyn+1) + d(F (yn, xn), F (a, b))
≤ d(g(a), gyn+1) +
[d(gyn, g(a)) + d(gxn, g(b))]
k
2
n → ∞.
k
2
n → ∞.
→ 0
as
thus, we have
F (a, b) = g(a).
(15)
Again from(1), (2), (13), (14) and traingular inequality, we have
d(g(b), F (b, a)) ≤ d(g(b), gxn+1) + d(gxn+1, F (b, a))
= d(g(b), gxn+1) + d(F (xn, yn), F (b, a))
≤ d(g(b), gxn+1) +
[d(gxn, g(b)) + d(gyn, g(a))]
→ 0
as
thus, we get
F (b, a) = g(b).
(16)
Hence from (15) and (16) we get, F (a, b) = g(a) and F (b, a) = g(b),
where a ∈ A and b ∈ B.
Thus (a, b) ∈ A × B is the coupled coincidence point of F and g.
Remark 2.7. It is worth noting that the above theorem also gives the condition
for the existence of symmetric point of F in A × B, i.e. ∃ (a, b) ∈ A × B s.t
F (a, b) = F (b, a).
As from (14) g(a) = g(b) so, F (a, b) = F (b, a).
Existence and Uniqueness of Strong coupled coincidence point of F
and g.
Theorem 2.8: If in addition to above condition in Theorem 2.6, g is one-one,
then
(i) A ∩ B 6= ∅, and
(ii) F and g have unique strong coupled coincidence point in A ∩ B.
Proof: since g is given to be one-one
then from (14) of Theorem 2.6, we get
a = b.
7
since, a ∈ A and b ∈ B
⇒ a = b ∈ A ∩ B. Hence A ∩ B 6= ∅, this proves our part (i).
Also from (15) of Theorem 2.6, we have
F (a, a) = g(a).
which shows that F and g have strong coupled coincidence point in A ∩ B.
Uniqueness: Let us suppose if possible there exists two strong coupled coinci-
dence points l, m ∈ A ∩ B of F and g
then by definition we have,
F (l, l) = g(l) and F (m, m) = g(m).
(17)
then from (1) of Theorem 2.6, we have
d(g(l), g(m)) = d(F (l, l), F (m, m))
≤
k
2
[d(g(l), g(m)) + d(g(l), g(m))]
= k[d(g(l), g(m))].
Which is a contradiction as k ∈ (0, 1) and is only possible if d(g(l), g(m)) = 0,
i.e. g(l) = g(m) or l = m because g is one-one.
Hence F and g have a unique strong coupled coincidence point in A ∩ B.
Corrollary 2.9 : If F : X × X → X and g : X → X are commutative and g is
one-one.
if (gx, gy) is a coupled coincidence point of F and g, where x, y ∈ X, then
(x, y) is also a coupled coincidence point of F and g.
Proof : As (gx, gy) is a coupled coincidence point of F and g, we have
F (gx, gy) = gx,
F (gy, gx) = gy.
(18)
(19)
Now by (18) and commutavity of F and g, we have
g(F (x, y)) = F (gx, gy)
= gx.
as g is one-one, we have from above
Now again by (19) and commutavity of F and g, we have
F (x, y) = x.
(20)
g(F (y, x)) = F (gy, gx)
= gy.
as g is one-one, we have from above
F (y, x) = y.
(21)
8
Thus from (20) and (21), we get
F (x, y) = x and F (y, x) = y.
10 , where x, y ∈ X.
i.e. (x, y) is the coupled coincidence point of F and g.
The following example illustrates our results.
Example 2.10. Let X = R with the metric defined as d(x, y) = x − y , where
x, y ∈ X.
Let A = [0, 2] and B = [0, 3].
Let F be defined as F (x, y) = x+y
and let g : X → X is defined by g(x) = x
2 .
Then g(A) = [0, 1] and g(B) = [0, 3
X.
Also g(A) ⊆ A and g(B) ⊆ B, so g is self-cyclic.
Now we show that F is g-coupling.
As g(A) ∩ B = [0, 1] and g(B) ∩ A = [0, 1], so ∀ x ∈ A and y ∈ B, we have
0 ≤ F (x, y) ≤ 1
i.e. F (x, y) ∈ g(A) ∩ B and F (y, x) ∈ g(B) ∩ A, which shows that F is
g-coupling with respect to A and B.
Again for x, v ∈ A and y, u ∈ B, we have
2 ], so g(A) and g(B) are closed subsets of
2 and 0 ≤ F (y, x) ≤ 1
2 .
d(F (x, y), F (u, v)) =
=
≤
=
=
1
5
1
5
1
5
k
2
[
[
x + y
u + v
−
10
x − u
10
y − v
+
2
x − u
2
−
u
2
x
2
2
y − v
+
+
]
v
2
]
2
−
y
2
[d(gx, gu) + d(gy, gv)], where k =
2
5
∈ (0, 1).
which shows that F is a Banach type g-coupling (w.r.t. A and B).
Thus all the conditions of Theorem (2.6) are satisfied, therefore ∃ (a, b) ∈ A×B
s.t.
F (a, b) = g(a) and F (b, a) = g(b).
i.e.
a + b
=
and
a
2
b + a
10
=
b
2
(22)
or
10
a + b
5
= a = b.
which is only possible with a = 0 and b = 0.
thus (0, 0) is the unique strong g-coupled coincidence point of F and g.
Note : The uniqueness of the strong g-coupled coincidence point in the above
example is because of g is one-one.
9
Corrollary 2.11: Let g : X → X is a Banach contraction, Then every Banach
type g-coupling (w.r.t. A and B) is a Banach type coupling (w.r.t. A and B),
where A and B are subsets of X.
Proof : since g is a contraction, therefore ∃ α ∈ (0, 1), s.t
d(gx, gy) ≤ αd(x, y) ∀ x, y ∈ X.
(23)
Let F : X × X → X be a Banach type g-coupling (w.r.t. A and B), i.e.
k
2
d(F (x, y), F (u, v)) ≤
[d(gx, gu)+d(gy, gu)], where x, v ∈ A, y, u ∈ B and k ∈ (0, 1).
(24)
Also F is a g-coupling (w.r.t. A and B) and hence by remark (2.4), F is a
coupling (w.r.t. A and B).
Now using (23) in (24), we get
d(F (x, y), F (u, v)) ≤
≤
k
2
αk
2
[αd(x, u) + αd(y, v)]
[d(x, u) + d(y, v)].
as k ∈ (0, 1) and α ∈ (0, 1), theref ore kα = k1 ∈ (0, 1), so
d(F (x, y), F (u, v)) ≤
k1
2
[d(x, u) + d(y, v)].
where x, v ∈ A, y, u ∈ B and k1 ∈ (0, 1).
Hence F is a Banach type coupling (w.r.t. A and B).
3. Conclusion
In this paper we introduce some new results that extend the concept of coupling
and Banach type coupling introduced by Choudhury et al. [5] to g-coupling and
Banach type g-coupling resp. and proved the existence and uniqueness theorem
for coupled coincidence point and strong coupled coincidence point.
References
[1] H. Aydi, M. Barakat, A. Felhi, H. Isik, On φ-contraction type couplings in
partial metric spaces, Journal of Mathematical Analysis, 8(2017), 78-89.
[2] H. Aydi, Coupled fixed point results in ordered partial metric spaces, Sel¸cuk
J. Appl. Math. 13(2012), 23-33.
[3] V. Berinde, Generalized coupled fixed point theorems in partially ordered
metric spaces and applications, Nonlin. Anal. 74(2011), 7347-7355.
[4] N. Bilgili, I. M. Erham, E. Karapinar, D. Turkoglu, A note on coupled fixed
point theorems for mixed g-monotone mappings in partially ordered metric
spaces, Fixed Point Theory Appl. 2014(2014):120, 6 pages.
10
[5] B.S. Choudhury, P. Maity, P. Konar, Fixed point results for couplings on
metric spaces, U. P. B. Sci. Bull., series A, 79(1)(2017). 1-12.
[6] B.S. Choudhury, A. Kundu, A coupled coincidence result in partially or-
dered metric spaces for compatible mappings, Nonlinear Anal. 73(2010),
2524-2531.
[7] B.S. Choudhury, A. Kundu, Two coupled weak contraction theorems in
partially ordered metric spaces,Revista de la Real Academia de ciencias
Exactas, Fisicas y Naturales. Series A.Mathematicas.108(2014), 335-351.
[8] B.S. Choudhury, P. Maity, Cyclic coupled fixed point result using Kannan
type contractions, Journal Of Operators. 2014(2014), Article ID 876749, 5
pages.
[9] T. Gnana Bhaskar, V. Lakshmikantham, Fixed point theorems in partially
ordered metric spaces and applications, Nonlin. Anal. 65(2006), 1379-1393.
[10] D. Guo, V. Lakshmikantham, Coupled fixed points of non-linear operators
with applications, Nonlin. Anal. 11(1987), 623-632.
[11] Harjani. J, Lop´ez. B, Sadarangani. K, Fixed point theorems for mixed
monotone operators and applications to integral equations, Nonlinear Anal.
74(2011), 1749-1760.
[12] W.A. Kirk, P.S. Srinivasan and P. Veeramani, Fixed points for mappings
satisfying cyclical contractive conditions, Fixed Point Theory. 4(2003), 78-
89.
[13] Lakshmikantham. V, L. ´Ciri´c, Coupled fixed point theorems for nonlinear
contractions in partially ordered metric spaces, Nonlinear Anal. 70(2009),
4341-4349.
[14] Luong. NV, Thuan. NX, Coupled fixed point theorems in partially ordered
metric spacesdepended on another function, Bull. Math. Anal. 3(2011), 129-
140.
[15] N. V. Luong and N. X. Thuan, Coupled points in ordered generalized metric
spaces and application to integro-dierential equations, Analele Universitatii
"Ovidius" Constanta Seria Matematica, vol. 21, no. 3, pp. 155-180, 2013.
[16] Mujahid Abbas, Bashir Ali, Yusuf I.Suleiman, Generalized coupled common
fixed point results in partially ordered A-metric spaces, Fixed Point Theory
and Application, (2015) 2015:64, 24 pages.
[17] Rosouli. SH, Bahrampour. M, A remark on the coupled fixed point theorems
for mixed monotone operators in partially ordered metric spaces, J. Math.
Comput. Sci. 3(2011), 246-261.
11
[18] W. Shatanawi, H.K. Nashine, N. Tahat, Generalization of some coupled
fixed point results on partial metric spaces, International Journal of Math-
ematics and Mathematical Sciences, 2012(2012), Article ID 686801, 10
pages.
[19] W. Shatanawi, B. Samet, M. Abbas, Coupled fixed point theorems for mixed
monotone mappings in ordered partial metric spaces, Mathematical and
Computer Modelling, 55(2012), 680-687.
[20] Wasfi Shatanawi, Mujahid Abbas, Hassen Aydi, Nedal Tahat, Common
coupled coincidence and coupled fixed points in G-metric spaces, Nonlinear
Anal. and Application, 2012(2012), Article ID jnaa-00162, 16 pages.
12
|
1505.01945 | 1 | 1505 | 2015-05-08T07:28:30 | Composition operators on vector-valued analytic function spaces: a survey | [
"math.FA"
] | We survey recent results about composition operators induced by analytic self-maps of the unit disk in the complex plane on various Banach spaces of analytic functions taking values in infinite-dimensional Banach spaces. We mostly concentrate on the research line into qualitative properties such as weak compactness, initiated by Liu, Saksman and Tylli (1998), and continued in several other papers. We discuss composition operators on strong, respectively weak, spaces of vector-valued analytic functions, as well as between weak and strong spaces. As concrete examples, we review more carefully and present some new observations in the cases of vector-valued Hardy and BMOA spaces, though the study of composition operators has been extended to a wide range of spaces of vector-valued analytic functions, including spaces defined on other domains. Several open problems are stated. | math.FA | math |
COMPOSITION OPERATORS ON VECTOR-VALUED
ANALYTIC FUNCTION SPACES: A SURVEY
JUSSI LAITILA AND HANS-OLAV TYLLI
Abstract. We survey recent results about composition operators in-
duced by analytic self-maps of the unit disk in the complex plane on
various Banach spaces of analytic functions taking values in infinite-
dimensional Banach spaces. We mostly concentrate on the research line
into qualitative properties such as weak compactness, initiated by Liu,
Saksman and Tylli (1998), and continued in several other papers. We
discuss composition operators on strong, respectively weak, spaces of
vector-valued analytic functions, as well as between weak and strong
spaces. As concrete examples, we review more carefully and present
some new observations in the cases of vector-valued Hardy and BM OA
spaces, though the study of composition operators has been extended
to a wide range of spaces of vector-valued analytic functions, including
spaces defined on other domains. Several open problems are stated.
1. Introduction
Let D = {z ∈ C : z < 1} be the open unit disk in the complex plane
C, and let ϕ : D → D be a fixed analytic map. The classical study of the
analytic composition operators Cϕ, where
f 7→ Cϕ(f ) = f ◦ ϕ,
originates from the work of Ryff (1966) and Nordgren (1968). For instance,
they observed that any Cϕ defines a bounded operator H p → H p as a conse-
quence of the Littlewood subordination principle. Recall that for 1 ≤ p < ∞,
the analytic function f : D → C belongs to the Hardy space H p if
kf kp
H p = sup
0≤r<1ZT
f (rξ)pdm(ξ) < ∞
where T = ∂D = [0, 2π] and dm(eit) = dt/2π. The space H ∞ consists of
the bounded analytic functions. Subsequently an extensive literature has
emerged, where a very wide variety of properties of analytic composition op-
erators has been addressed on a large number of spaces of analytic functions.
Date: 20.3.2014.
2010 Mathematics Subject Classification. 47B33, 46E15, 46E40, 47B07.
Key words and phrases. Analytic functions, Banach space, compactness, composition
operator, weak compactness, vector-valued function.
1
2
JUSSI LAITILA AND HANS-OLAV TYLLI
We refer to [47] and [11] for comprehensive accounts of the theory until ca.
1995.
This survey reviews more recent results about composition operators on
various Banach spaces of vector-valued analytic functions including the vec-
tor-valued Hardy space H p(X), where X is a complex Banach space. Let
f : D → X be a vector-valued analytic function and let 1 ≤ p < ∞. Then
f ∈ H p(X), if
kf kp
H p(X) = sup
0≤r<1ZT
kf (rξ)kp
X dm(ξ) < ∞.
Moreover, f ∈ H ∞(X), if kf kH∞ = supz∈D kf (z)kX < ∞. In this notation
H p = H p(C). Above the analyticity of f : D → X means that the scalar-
valued function x∗◦f is analytic D → C for any functional x∗ ∈ X ∗ (that is, f
is weakly analytic). This is equivalent to the requirement that the X-valued
derivative f ′(z) exists for all points z ∈ D (that is, f is strongly analytic).
For the basics of vector-valued analytic functions, see for example [20].
Qualitative properties of the vector-valued composition operators f 7→ f ◦ϕ
on H p(X) and certain other spaces were first systematically studied by Liu,
Saksman and Tylli [35]. Independently Hornor and Jamison [21] considered
the operators f 7→ f ◦ ϕ on H p(X) with different aims, and Sharma and
Bhanu [49] looked at some of their basic operator properties on H 2(X),
where X is a Hilbert space.
We mostly concentrate on qualitative properties, such as weak compact-
ness, of composition operators on several Banach spaces of vector-valued
analytic functions of both strong and weak type defined on D. Weak type
spaces were introduced into this context by Bonet, Domanski and Lindström
[4], and in this case the techniques differ from those of the strong type spaces.
In section 2 we introduce a general framework for vector-valued composition
operators in order to provide a convenient general perspective into the study,
and we review results that illustrate both similarities and differences com-
pared to the scalar-valued case X = C. We also highlight new phenomena
that do not have any counterparts for scalar composition operators. For in-
stance, composition operators can be studied between a weak and a strong
space. In the final section we briefly discuss attempts to generalize the larger
class of weighted composition operators to the vector-valued setting. Some
vector-valued arguments are sketched, but we mostly assume that the basic
scalar theory is known from [47] and [11].
Composition operators of different nature occur in various other settings.
For instance, there is a well-developed theory of the composition operators
S 7→ A ◦ S ◦ B, where A and B are fixed bounded operators, on spaces of
linear operators, see e.g. the survey [44]. Properties of such composition
operators will actually be required in section 5 below.
SURVEY OF COMPOSITION OPERATORS
3
2. a general framework
We first introduce a flexible general framework for the study of qualita-
tive properties of vector-valued composition operators, which will facilitate
a discussion of some common features.
Suppose that A is a Banach space of analytic functions D → C and let
A(X) be an associated vector-valued Banach space of analytic functions D →
X, where X is a complex Banach space. Assume that the following properties
hold for the pair (A, A(X)) for all Banach spaces X:
(AF1) The constant maps f (z) ≡ c belong to A for all c ∈ C.
(AF2) f 7→ f ⊗ x defines a bounded linear operator Jx : A → A(X) for any
x ∈ X, where (f ⊗ x)(z) = f (z)x for z ∈ D.
(AF3) g 7→ x∗ ◦ g defines a bounded linear operator Qx∗ : A(X) → A for any
x∗ ∈ X ∗.
(AF4) The point evaluations δz, where δz(f ) = f (z) for f ∈ A(X), are
bounded A(X) → X for all z ∈ D.
It follows from (AF1) and (AF2) that the vector-valued constant maps
z 7→ fx(z) ≡ x, that is fx = 1 ⊗ x, belong to A(X) for all x ∈ X.
It is
easy to check that (AF1) -- (AF4) are satisfied for the pair (H p, H p(X)) for
any X. Note that for Banach spaces of analytic functions defined on other
domains, such as a half-plane or the plane C, condition (AF1) may not be
relevant and the above framework cannot be applied in this form.
Suppose that A and B are Banach spaces of analytic functions D → C
so that (A, A(X)) and (B, B(X)) satisfy (AF1) -- (AF4), where A(X) and
B(X) are X-valued Banach spaces of analytic functions on D associated with
A, respectively B. Let ϕ : D → D be a given analytic self-map, and suppose
In order to distinguish between composition
operators acting on different spaces we will in the sequel use Cϕ : A → B for
the composition operator f 7→ f ◦ ϕ in the scalar-valued setting, that is, in
that the vector-valued composition operator fCϕ is bounded A(X) → B(X),
where f 7→ fCϕ(f ) = f ◦ ϕ.
the case X = C, and fCϕ for its vector-valued version A(X) → B(X).
The following general formulation is partly motivated by [4, Prop. 1].
Proposition 1. The following factorizations hold.
(F1) Let x ∈ X, x∗ ∈ X ∗ be norm-1 vectors so that hx∗, xi = 1. Then
A(X)
Jx
A
fCϕ
Cϕ
/ B(X)
Qx∗
/ B
commutes.
/
O
O
/
4
JUSSI LAITILA AND HANS-OLAV TYLLI
(F2) Let j(x) = fx for x ∈ X, where fx(z) ≡ x for all z ∈ D. Then
A(X)
j
X
fCϕ
IX
/ B(X)
δ0
/ X
commutes, where IX is the identity operator on X.
The above factorizations place some inherent restrictions on possible qual-
Proof. Note towards (F1) that x∗(fCϕ(f ⊗ x)) = x∗((f ◦ ϕ) ⊗ x) = Cϕ(f ) for
f ∈ A, while δ0(fCϕ(fx)) = δ0(fx) = x for x ∈ X.
itative properties of the vector-valued operators fCϕ. Roughly speaking, part
(3) below states that fCϕ : A(X) → B(X) cannot have any qualitative prop-
properties of X also influence (qualitative) properties of fCϕ.
erties inherited under composition of linear operators that are not shared
by Cϕ : A → B and the identity operator IX : X → X. Thus Banach space
Corollary 2. Let X be a complex Banach space.
(cid:3)
X is finite-dimensional. In particular, if X is infinite-dimensional,
(1) If fCϕ is bounded A(X) → B(X), then Cϕ is bounded A → B.
(2) If fCϕ : A(X) → B(X) is compact, then Cϕ is compact A → B and
(3) Let I be an operator ideal in the sense of Pietsch [43]. If fCϕ : A(X) →
In fact, by (F1) and (F2) the compactness of fCϕ : A(X) → B(X) implies
then fCϕ is never compact A(X) → B(X).
that both Cϕ : A → B and IX are compact, that is, X is finite-dimensional.
Part (3) is verified in a similar fashion. For a converse of (2), see Proposition
7.
B(X) belongs to I, then IX as well as Cϕ : A → B belong to I.
3. Weak compactness on H 1(X) and other vector-valued spaces
Let ϕ : D → D be any analytic map. It was observed independently in [35]
and [21] that fCϕ is bounded on H p(X), while [49] contains the case H 2(X),
where X is a Hilbert space. Boundedness can be verified in the following
manner by a small modification of an argument for scalar H p spaces. Note
first that z 7→ kf (z)kX is a subharmonic map on D for any analytic function
f : D → X, since
kf (z)kX = sup
kx∗k≤1
hx∗, f (z)i,
z ∈ D.
Consequently, if ϕ(0) = 0, then the Littlewood inequality [11, Thm. 2.22]
yields that kfCϕ(f )kH p(X) ≤ kf kp
H p(X) for f ∈ H p(X).
/
O
O
/
SURVEY OF COMPOSITION OPERATORS
5
For the general case let σa : D → D be the Möbius transformation defined
by σa(z) = a−z
1−az for a ∈ D. If ϕ(0) 6= 0 , let ψ = σϕ(0) ◦ ϕ, so that ψ(0) = 0
and Cψ is a contraction H p(X) → H p(X). Since σ−1
ϕ(0) = σϕ(0), we get that
Cϕ = Cψ ◦ Cσϕ(0) is bounded on H p(X) once we have checked that Cφ is
bounded on H p(X) for any Möbius transformation φ. This can be verified
by the change of variables w = φ(z) inside the integral
ZT
kf (rφ(ξ))kp
X dm(ξ)
defining kCφ(f (r·))kH p(X) for 0 < r < 1 and letting r → 1.
The minor point of difference between the above scalar- and vector-valued
arguments for the Hardy spaces relates to the potential absence of radial
In fact, it is well known that any f ∈ H p has a.e. radial limits
limits.
f (ξ) = limr→1− f (rξ) on T, but this is not always true for functions in
H p(X): the bounded analytic function f : D → c0, where
f (z) = (zn),
z ∈ D,
does not have radial limits anywhere on T.
In fact, the existence of a.e.
radial limits for any f ∈ H p(X), and any fixed 1 ≤ p ≤ ∞, characterizes the
analytic Radon-Nikodým property (ARNP) of the complex Banach space X.
The ARNP and the above result by Bukhvalov and Danilevich (1982) is not
needed here, but the reader may keep in mind that e.g. every reflexive Banach
space has the ARNP. See, e.g., [28, p. 723] for references and a discussion of
the ARNP.
Let X be an infinite-dimensional Banach space. According to Corollary
2.(2) there are no compact compositions fCϕ : H p(X) → H p(X). This raises
the general question of which are the relevant qualitative properties for com-
position operators on vector-valued spaces such as H p(X). In [35] the au-
thors considered weak compactness, and related properties, for which there
are satisfactory results.
Let X and Y be Banach spaces. Recall that the bounded linear operator
U : X → Y is weakly compact if there is a weakly convergent subsequence
(U xnk ) for any bounded sequence (xn) ⊂ X. If X and Y are non-reflexive
spaces, then weakly compact U : X → Y are relatively small operators.
A fundamental result of Shapiro [46] (see also [47] and [11]) says that for
1 ≤ p < ∞ the composition operator Cϕ is compact H p → H p if and only if
lim
w→1
N (ϕ, w)
log(1/w)
= 0.
(1)
Above N (ϕ, w) =Pz∈ϕ−1(w) log(1/z), where w ∈ D \ {ϕ(0)}, is the Nevan-
linna counting function of ϕ. Several other equivalent criteria for the com-
pactness of Cϕ : H p → H p are known in the literature, but (1) suffices for
our purposes. Littlewood's inequality implies that N (ϕ, w) ≤ C · log(1/w)
6
JUSSI LAITILA AND HANS-OLAV TYLLI
as w → 1 for some constant C = C(ϕ) for any analytic map ϕ : D → D,
see [47, 10.4]. Shapiro's condition (1) is interpreted as a little-oh condition
describing the rate of decrease of the affinity of ϕ for the values w as w → 1.
H 1(X) and the compactness of Cϕ on H 1. Note that Corollary 2.(3) implies
There is a precise connection between the weak compactness of fCϕ on
that X is reflexive, that is, IX is weakly compact, whenever fCϕ is weakly
compact on H p(X). Hence only p = 1 or p = ∞ are interesting for weak
compactness, since H p(X) is itself reflexive if 1 < p < ∞ and X is reflexive,
because H p(X) is then a closed subspace of the reflexive space Lp(T, X).
The vector-valued part of the following result comes from [35].
Theorem 3. Let X be a complex reflexive Banach space, and ϕ : D → D be
an analytic map. Then the following conditions are equivalent.
(1) fCϕ : H 1(X) → H 1(X) is weakly compact
(2) Cϕ : H 1 → H 1 is weakly compact
(3) Cϕ : H 1 → H 1 is compact
(4) Shapiro's condition (1) holds.
Proof. The implication (1) ⇒ (2) follows from Corollary 2.(3). Sarason [45]
proved that the weak compactness of Cϕ : H 1 → H 1 actually yields the com-
pactness of Cϕ : H 1 → H 1, in other words that (2) ⇒ (3). The equivalence
of (3) and (4) is contained in Shapiro's theorem. There remains to show
condition.
We outline the proof of the implication (4) ⇒ (1). The argument is based
that fCϕ is weakly compact H 1(X) → H 1(X) whenever ϕ satisfies Shapiro's
on a Littlewood-Paley type formula for kfCϕ(f )kH 1(X) derived from a for-
mula of Stanton for continuous subharmonic maps. His formula [51, Thm.
2] implies that
kf ◦ ϕkH 1(X) = kf (0)kX +
1
2π Z N (ϕ, w)d[∆(kf kX )](w)
(2)
for f ∈ H 1(X), where d[∆(kf kX )] denotes the distributional Laplacian as-
sociated to the subharmonic map z 7→ kf (z)kX on D.
Define de la Vallee-Poussin operators Vn for any n ∈ N by
Vnf (z) =
nXk=0
fkzk +
2n−1Xk=n+1
2n − k
n
fkzk
k=0
P∞
for analytic functions f : D → X having the Fourier expansion f (z) =
fkzk. Then (Vn) is a uniformly bounded sequence of operators on
H 1(X) and Vn : H 1(X) → H 1(X) are weakly compact for any n if X is a re-
flexive Banach space (in fact, Vn factors through a finite direct sum of copies
of X). Moreover, given ε > 0 and 0 < r < 1 there is n0 = n0(ε, r) ∈ N so
SURVEY OF COMPOSITION OPERATORS
that
kf (z) − Vn0f (z)kX ≤ ε · kf kH 1(X)
holds for all z ≤ r and f ∈ H 1(X).
7
(3)
If Shapiro's condition (1) holds and ε > 0 is arbitrary, then there is r ∈
(0, 1) such that N (ϕ, w) ≤ ε · log(1/w) for all w > r. Fix n0 as in (3)
corresponding to ε and r. By applying (2) to f − Vn0f we get for f ∈ H 1(X)
that
1
2π Z{r<z<1}
N (ϕ, w)d[∆(kf − Vn0f kX)](w)
kfCϕ(f ) − fCϕ(Vn0f )kH 1(X) =
1
2π Z{z≤r}
+
N (ϕ, w)d[∆(kf − Vn0f kX)](w) ≡ I1 + I2.
The choice of r ∈ (0, 1) and (2) applied to ψ(z) = z give
I1 ≤
ε
2π ZD
log(1/w)d[∆(kf − Vn0f kX)](w) = εkf − Vn0f kH 1(X)
≤ C · εkf kH 1(X)
for a uniform constant C. Moreover, it can be shown that I2 ≤ 4εkf kH 1(X)
by using the estimates N (ϕ, w) ≤ log(1/w) and (3). For this it is convenient
to introduce a cut-off function ψ ∈ C ∞
0 (D) satisfying 0 ≤ ψ ≤ 1, ψ = 1 on
{z : z ≤ r} and ψ = 0 on {z ∈ D : z ≥ (1 + r)/2}. We refer to [35, Prop. 2
and Thm. 3] for the complete technical details.
Thus kfCϕ −fCϕVn0k ≤ C ′ · ε, where C ′ does not depend on ε, so that fCϕ is
well approximated by the weakly compact operators fCϕVn0 for suitable n0.
This means that fCϕ is weakly compact H 1(X) → H 1(X).
Theorem 3 corresponds to the following template for many results about
weak compactness, as well as other qualitative properties, of analytic com-
position operators on vector-valued spaces.
(cid:3)
Proposition 4. Let A be a Banach space of analytic functions on D and
A(X) a vector-valued version of A, such that (A, A(X)) satisfies (AF1) --
(AF4). Suppose that X is a complex reflexive Banach space and ϕ : D → D
that the following conditions hold:
(C1) if Cϕ is weakly compact A → A, then Cϕ is compact A → A, and
is an analytic map, so that fCϕ is bounded A(X) → A(X). Assume moreover
(C2) if Cϕ is compact A → A, then the vector-valued composition fCϕ is
(C) fCϕ is weakly compact A(X) → A(X) ⇔ Cϕ is compact A → A.
Then one has the characterization
weakly compact A(X) → A(X).
8
JUSSI LAITILA AND HANS-OLAV TYLLI
We stress that the above general scheme is only a guiding principle and
in practice the techniques for establishing (C2) depend on A and its vector-
valued extension A(X). Moreover, the criteria for the compactness of the
operator Cϕ : A → A usually depend on A. It is straightforward to modify the
scheme of Proposition 4 to apply to vector-valued compositions fCϕ : A(X) →
B(X) between different spaces, where (A, A(X)) and (B, B(X)) satisfy the
properties (AF1) -- (AF4).
Condition (C1) is a problem of independent interest for composition op-
erators A → A. Recently Lefevre, Li, Queffelec and Rodriguez-Piazza [31]
constructed the first example of a Banach space A of complex-valued analytic
functions on D, where (C1) fails for some symbol ϕ, see Example 10 below.
We next look at cases where Proposition 4 apply. Let vα(z) = (1 − z2)α
for z ∈ D and α > −1. The analytic function f : D → X belongs to the
weighted Bergman space Aα
p (X) if
kf kp
Aα
p (X) =ZD
kf (z)kp
X vα(z)dA(z) < ∞,
where dA is the area Lebesgue measure normalized by A(D) = 1 and 1 ≤ p <
∞. The classical Bergman space Ap(X) is obtained for α = 0. The following
result was established in [4], but the special case A1(X) was already contained
in [35].
Theorem 5. Let X be a complex reflexive Banach space, ϕ : D → D an
analytic map and α > −1. Then the following conditions are equivalent.
(1) fCϕ : Aα
1 (X) → Aα
1 → Aα
(2) Cϕ : Aα
(3) ϕ satisfies the condition
1 is compact
1 (X) is weakly compact
lim sup
w→1
Nα+2(ϕ, w)
(log(1/w))α+2 = 0.
Above Nβ(ϕ, ·) is the generalized Nevanlinna counting function defined for
β > 0 by
Nβ(ϕ, w) = Xz∈ϕ−1(w)
(log(1/z)β , w ∈ D \ {ϕ(0)},
so that N (ϕ, ·) = N1(ϕ, ·). Actually, [4, Thm. 8] contains the estimate
dist(fCϕ, W (Aα
1 (X))) ≤ C · lim sup
w→1
Nα+2(ϕ, w)
(log(1/w))α+2 ,
where W (Aα
pact operators Aα
1 (X)) denotes the linear subspace consisting of the weakly com-
1 (X) → Aα
1 (X) and C is an absolute constant.
We list some further Banach spaces A(X) for which the characterization
(C) for weak compactness of composition operators are known to hold. We
SURVEY OF COMPOSITION OPERATORS
9
emphasize that the arguments establishing (C1) and (C2) usually are specific
for A(X), and the relevant compactness conditions for Cϕ depend on A. One
verifies by inspection that these pairs (A, A(X)) satisfy (AF1) -- (AF4).
• Let v : D → (0, ∞) be a bounded continuous weight function, and
f : D → X be an analytic function. Recall that f ∈ H ∞
v (X) if
kf kH∞
v (X) = sup
z<1
v(z)kf (z)kX < ∞.
v = H ∞
v (C). The case v ≡ 1 gives the classical spaces H ∞(X)
Let H ∞
and H ∞ of bounded analytic functions. It was shown in [35] that
(C) holds on H ∞(X) and this was extended to the case H ∞
v (X) by
different means in [4].
• The vector-valued Bloch space B(X) [35]. Recall that f ∈ B(X) if
(1 − z2)kf ′(z)kX < ∞.
sup
z∈D
See [4] for an alternative approach via weak spaces (section 5 below).
• The space CT (X) of vector-valued Cauchy transforms [28]. The ar-
gument proceeds via composition operators on the vector-valued har-
monic Hardy space h1(X). The compactness criterion for CT is due
to Bourdon, Cima and Matheson.
• Vector-valued BM OA(X)-spaces, see section 4.
• Weak vector-valued versions of the above spaces, see section 5.
A modified general scheme as in Proposition 4 also applies to other op-
erator ideal properties, namely, just replace weak compactness by the rele-
vant ideal property in (C1) and (C2). We state two results of this kind for
H 1(X), respectively A1
α(X), from [35, Thm. 7] and [4, Cor. 9]. The opera-
tor U : X → Y is called weakly conditionally compact if (U xn) has a weak
Cauchy subsequence (U xnk ) for any bounded sequence (xn) ⊂ X. Recall that
by Rosenthal's ℓ1-theorem, see [33, 2.e.5], IX is weakly conditionally compact
if and only if X that does not contain any subspaces linearly isomorphic to
ℓ1. By Proposition 1 this is the relevant class of spaces here.
Theorem 6. Suppose that the Banach space X does not contain any sub-
spaces linearly isomorphic to ℓ1, and ϕ : D → D is an analytic map. Let
A = H 1 or A = A1
A(X) → A(X) if and only if Cϕ is compact A → A.
α for α > −1. Then fCϕ is weakly conditionally compact
We mention for completeness that the cases dim(X) < ∞ are similar to
the scalar case.
Proposition 7. Suppose that dim(X) < ∞, (A, A(X)) satisfies (AF1) --
(AF4), and ϕ : D → D is an analytic map. Then fCϕ is compact A(X) →
A(X) if and only if Cϕ is compact A → A.
10
JUSSI LAITILA AND HANS-OLAV TYLLI
Proof. Let n = dim(X) and fix a biorthogonal system {(xr, x∗
n} for X, so that x = Pn
be written as f (z) =Pn
Consequently, if Cϕ is compact A → A, then f 7→ fCϕ(f ) = Pn
s) : 1 ≤ r, s ≤
k(x)xk for x ∈ X. Hence any f ∈ A(X) can
k ◦ f ∈ A for k = 1, . . . , n.
k=1 Cϕ(fk)xk
is compact A(X) → A(X). For the converse note that Section 2 applies to
this setting.
(cid:3)
k=1 fk(z)xk, where fk = x∗
k=1 x∗
Other results. Hornor and Jamison [21] characterized the isometrically
equivalent compositions fCϕ and fCψ on H p(X), respectively Sp(X), for p 6= 2
and X a Hilbert space. Here f ∈ Sp(X) if the derivative f ′ ∈ H p(X).
Sharma and Bhanu [49] studied e.g. normal and unitary compositions on
H 2(X), where X is a Hilbert space. Bonet and Friz [5] characterized the
weakly compact compositions on weighted vector-valued locally convex spaces
of analytic functions on D. Composition operators on the Hilbert space-
valued Fock space of entire functions on C were considered by Ueki [53]. See
also [54] for results on the vector-valued Nevanlinna class.
4. Vector-valued BM OA-spaces
In this section we discuss in more detail the case of composition operators
on vector-valued BM OA spaces, since there are several natural vector-valued
versions of BM OA, and condition (C1) is a problem of independent interest.
Recall that the analytic function f : D → C belongs to BM OA, the space
of analytic functions of bounded mean oscillation, if
f ∗ = sup
a∈D
kf ◦ σa − f (a)kH 2 < ∞,
where σa(z) = (a − z)/(1 − az) for z ∈ D. The Banach space BM OA is
equipped with the norm kf kBM OA = f (0)+f ∗. BM OA is often considered
as a Möbius-invariant version of H 2, but its Banach space structure is comp-
licated, see e.g. [41]. Recall also that (H 1)∗ ≈ BM OA
There are by now several equivalent characterizations of compact com-
positions Cϕ : BM OA → BM OA, see [27] for a list. The following double
criterion due to W. Smith [50] is the most relevant one for our purposes:
Let ϕ : D → D be an analytic map. Then Cϕ is compact BM OA → BM OA
if and only if
lim
ϕ(a)→1
0<w<1
sup
w2N (σϕ(a) ◦ ϕ ◦ σa, w) = 0,
lim
t→1
sup
{a : ϕ(a)≤R}
m(cid:0){ζ ∈ ∂D : (ϕ ◦ σa)(ζ) > t}(cid:1) = 0.
Subsequently it was observed in [26] that (S1) can be restated as
lim
ϕ(a)→1
kσϕ(a) ◦ ϕ ◦ σakH 2 = 0.
(S1)
(S2)
(L)
SURVEY OF COMPOSITION OPERATORS
11
Let f : D → X be an analytic function. We say that f ∈ BM OA(X) if
f ∗,X = sup
a∈D
kf ◦ σa − f (a)kH 2(X) < ∞,
and let kf kBM OA(X) = kf (0)kX + f ∗,X. There are also other natural possi-
bilities. By departing from a well-known characterization of BM OA in terms
of Carleson measures, see [18, Thm. VI.3.4], let f ∈ BM OAC(X) if
f C,X = sup
kf ′(z)k2
X (1 − σa(z)2)dA(z) < ∞.
a∈DZD
The norm in BM OAC(X) is kf kBM OAC(X) = kf (0)kX + f C,X. Blasco
[3] showed that BM OA(X) = BM OAC(X), with equivalent norms, if and
only if X is linearly isomorphic to a Hilbert space. Thus BM OA(X) and
BM OAC(X) are different vector-valued versions of BM OA. (In section 5
we will meet yet another vector-valued version of BM OA.)
Laitila [24], [25] initiated the study of composition operators on vector-
BM OA(X) and BM OAC(X) → BM OAC(X) for any self-map ϕ : D → D.
Moreover, if X is a reflexive Banach space and ϕ satisfies conditions (S1)
valued BM OA-spaces. He observed that fCϕ is bounded BM OA(X) →
and (S2), then fCϕ is weakly compact both BM OA(X) → BM OA(X) and
BM OAC(X) → BM OAC(X). In order to obtain a complete characterization
following Proposition 4 one has to verify condition (C1) for BM OA. This
was actually a problem stated by Tjani in her Ph.D. thesis [52] and Bourdon,
Cima and Matheson [7], which was eventually solved in [27] as follows.
Theorem 8. The following conditions are equivalent for ϕ : D → D:
(1) Cϕ : BM OA → BM OA is compact
(2) Cϕ : BM OA → BM OA is weakly compact
(3) (S1) holds (alternatively, (L) holds)
It is part of the solution that condition (S2) is redundant in Smith's char-
acterization above. The combination of Theorem 8 with [24], [25] completes
the following result for these vector-valued BM OA-spaces.
Theorem 9. Let X be a reflexive Banach space, and ϕ : D → D an analytic
function. Then the following conditions are equivalent.
(1) fCϕ is weakly compact BM OA(X) → BM OA(X)
(2) fCϕ is weakly compact BM OAC(X) → BM OAC(X)
(3) Cϕ : BM OA → BM OA is compact, that is, condition (S1) holds.
The argument for Theorem 8 is quite intricate. It applies measure density
ideas for the radial limits of ϕ combined with a criterion due to Leibov (1986),
respectively Müller and Schechtman (1989), which allows to extract copies of
the unit vector basis in c0 from bounded sequences in the subspace V M OA
of BM OA. We refer to [27] for the full technical details.
12
JUSSI LAITILA AND HANS-OLAV TYLLI
The results of sections 3 and 4 might suggest that the weak compactness
of Cϕ : A → A always implies its compactness A → A for any Banach space
A of analytic functions on D. However, this is not the case [31, Thm. 4.1]:
Example 10. Let ϕ be the lens map
ϕ(z) =
(1 + z)1/2 − (1 − z)1/2
(1 + z)1/2 + (1 − z)1/2
,
z ∈ D.
Then there is an Orlicz function ψ so that Cϕ is weakly compact H ψ →
H ψ, but not compact, where H ψ is the non-reflexive Hardy-Orlicz space of
analytic functions of D defined by ψ.
Question 11. Characterize the weakly compact compositions on the space
H ψ(X) above. The operator Cϕ in Example 10 factors through H 4 by con-
struction, and fCϕ through the reflexive space H 4(X) for reflexive spaces X.
Thus fCϕ is weakly compact H ψ(X) → H ψ(X), so that (C) cannot hold for
H ψ(X).
To the best of our knowledge the first example of a weakly compact ana-
lytic composition operator which is not compact was obtained in the context
of uniform algebras defined on infinite-dimensional domains. Let UE be the
open unit ball of the Tsirelson space E and ϕ : UE → UE the map x 7→ x/2.
It was shown in [1, Example 3] that the composition operator f 7→ f ◦ ϕ is
weakly compact, but non-compact, H ∞(UE) → H ∞(UE). Here H ∞(UE) is
the uniform algebra of bounded scalar-valued analytic functions UE → C.
5. Weak vector-valued spaces
Bonet, Domanski and Lindström [4] introduced the class of weak spaces of
vector-valued analytic functions into the study of vector-valued composition
operators. One of their aims was to provide an alternative approach to [35],
but the weak spaces are in general different from the spaces considered in
sections 3 and 4. On the other hand, for the class of weak spaces there are
some general results concerning vector-valued compositions.
Suppose that E is a Banach space of analytic functions f : D → C satisfy-
ing the following conditions:
(W1) E contains the constant functions,
(W2) the closed unit ball BE is compact in the compact open topology τco
of D.
Recall that the vector-valued function f : D → X is analytic if and only
if x∗ ◦ f is analytic D → C for all x∗ ∈ X ∗. This fact suggests to define
f ∈ wE(X) if
kf kwE(X) = sup
kx∗kX ∗ ≤1
kx∗ ◦ f kE < ∞.
SURVEY OF COMPOSITION OPERATORS
13
By the closed graph theorem kf kwE(X) is finite if and only if x∗ ◦f ∈ E for all
x∗ ∈ X ∗. We will say that wE(X) is the weak space of vector-valued analytic
functions D → X modelled on E. The spaces appearing in sections 3 and 4,
whose norms involve pointwise norm quantities such as kf (z)kX , will in the
sequel be called strong spaces. Such a distinction between strong and weak
spaces is not precise, since e.g. wH ∞(X) = H ∞(X). If E is a Banach space
of harmonic functions on D which satisfies (W1) and (W2), then one may
similarly define the weak space wE(X) of vector-valued harmonic functions
D → X, see [28]. Weak type spaces first appeared in the theory of vector
measures, see e.g. [13, chap. 13]. The weak Hardy spaces wH p(X), and in
particular their harmonic versions whp(X), have been studied by Blasco [2],
as well as in [15, 16].
Weak spaces wE(X) have a dual nature, since they also admit a canonical
isometrically isomorphic representation as certain spaces of bounded opera-
tors. This general fact was observed in [4]. Note first that if E satisfies (W1)
and (W2), then
(W3) the evaluation maps δz ∈ E∗ for z ∈ D, where δz(f ) = f (z) for f ∈ E.
The Dixmier-Ng theorem [42] implies that E = V ∗, where
V = {u∗ ∈ E∗ : u∗ is τco-continuous on BE}.
The identification of f ∈ E with u∗ 7→ u∗(f ) gives the isometric isomorphism
E → V ∗. In addition, V = [δz ∈ E∗ : z ∈ D] by Hahn-Banach, where [B]
denotes the closed linear span of the subset B ⊂ E∗. We next formulate
the general linearization result from [4], which also implies that wE(X) is a
Banach space. An analogue holds for weak harmonic spaces, see [28].
Theorem 12. Suppose that E satisfies (W1) and (W2), let V = [δz ∈
E∗ : z ∈ D] and X be a complex Banach space. Then there is an isometric
isomorphism χ : L(V, X) → wE(X), so that
(χ(T ))(z) = T (δz),
(χ−1(f ))(δz) = f (z),
hold for T ∈ L(V, X), f ∈ wE(X) and z ∈ D.
Special cases and variants of this linearization result were known earlier.
The closest precursor is the general results of Mujica [40] that apply to the
case E = H ∞. An explicit operator representation was obtained by Blasco
[2] for the weak harmonic spaces whp(X), where 1 ≤ p ≤ ∞.
The study of composition operators between weak spaces of analytic func-
tions was initiated by Bonet, Domański and Lindström [4], and this was
extended in [28] to weak spaces of vector-valued harmonic functions. Propo-
sition 11 of [4] contains the result on weak compactness stated below in The-
orem 14 for the weak spaces, but [4] only explicitly discusses the weighted
wH ∞
v (X)-spaces and the weak Bloch space wB(X) as examples. However,
this approach applies to a large class of weak spaces of analytic functions, such
14
JUSSI LAITILA AND HANS-OLAV TYLLI
as the weak Hardy and weak Bergman spaces, as well as to wBM OA(X),
see the discussion below as well as in [28, 24, 25].
Let ϕ : D → D be an analytic map. The vector-valued composition fCϕ is
bounded wE(X) → wE(X) if and only if Cϕ is bounded E → E. In fact, if
x∗ ∈ X ∗ then
kx∗ ◦ (fCϕf )kE = kCϕ(x∗ ◦ f )kE ≤ kCϕk · kx∗ ◦ f kE,
so that kfCϕk ≤ kCϕk. For the converse it is worthwhile to point out that
the framework from section 2 applies to the weak spaces.
Lemma 13. If E satisfies (W1) and (W2), then the pair (E, wE(X)) satis-
fies (AF1) -- (AF4) for any Banach space X
Proof. Conditions (AF1) -- (AF3) are obvious. Towards (AF4) note that
x∗(eδz(f )) = δz(x∗ ◦ f ),
f ∈ wE(X), x∗ ∈ X ∗,
where we momentarily use eδz for the vector-valued evaluations f 7→ f (z)
taking wE(X) to X.
(cid:3)
We stress that the following basic weak compactness result from [4] for
vector-valued compositions holds on all weak spaces wE(X). The proof
uses different tools compared to the analytic arguments in Sections 3 and 4.
Recall again from Corollary 2 that fCϕ is never compact wE(X) → wE(X)
whenever X is infinite dimensional.
Theorem 14. Suppose that E is a Banach space of analytic functions on
D that satisfy (W1) and (W2). Let ϕ : D → D be an analytic map and X a
reflexive Banach space. If Cϕ : E → E is compact, then fCϕ is weakly compact
wE(X) → wE(X).
Proof. Assume that Cϕ : E → E is compact. Its adjoint (Cϕ)∗ : E∗ → E∗
satisfies
(Cϕ)∗(δz) = δϕ(z),
z ∈ D,
so that (Cϕ)∗(V ) ⊂ V . We obtain the factorization fCϕ = χ ◦ Uϕ ◦ χ−1, where
Uϕ is the operator composition map
T 7→ IX ◦ T ◦ (Cϕ)∗V ; L(V, X) → L(V, X),
and χ is the isometric isomorphism L(V, X) → wE(X) from Theorem 12.
Since (Cϕ)∗V is a compact operator V → V by duality, and IX is weakly
compact, it follows from a general result of Saksman and Tylli, see [44,
Prop. 2.3], that the operator composition Uϕ is weakly compact L(V, X) →
(cid:3)
L(V, X). Consequently fCϕ is weakly compact wE(X) → wE(X).
Theorem 14 verifies condition (C2) from Proposition 4 for the weak spaces
wE(X). The following observation includes many examples.
SURVEY OF COMPOSITION OPERATORS
15
Proposition 15. Suppose that E is a Banach space of analytic functions on
D that satisfy (W1) and (W2), let ϕ : D → D be an analytic map so that Cϕ
is bounded E → E and X a reflexive Banach space. Suppose moreover:
(C1) if Cϕ is weakly compact E → E, then Cϕ is compact E → E.
Then one has the characterization
(C) fCϕ is weakly compact wE(X) → wE(X) ⇔ Cϕ is compact E → E.
Moreover, (C1) holds e.g. if E is one of the following spaces: H 1, A1
α > −1, BM OA, H ∞
The preceding examples cover results for wH ∞
α for
v , where v is a bounded continuous weight on D, or B.
v (X) and wB(X) from [4],
wH 1(X) [28], and wBM OA(X) (combine Theorem 8 with [24, 25]).
The results of sections 3 -- 5 raise the question of what is the precise
connection between these strong and weak spaces of vector-valued analytic
functions. Clearly wH 2(ℓ2) ≈ L(ℓ2) by Theorem 12, whereas H 2(ℓ2) is a
separable Hilbert space, so the difference can be huge. On the other hand,
[4] observed that wH ∞
v (X) (equal norms) and wB(X) ≈ B(X)
(equivalent norms). It is evident that e.g. H p(X) ⊂ wH p(X), and
v (X) = H ∞
kf kwH p(X) ≤ kf kH p(X),
f ∈ H p(X),
where 1 ≤ p < ∞ and X is any Banach space. Blasco [2] observed that
h1(C(T)) wh1(C(T)) and hp(Lp′
) for 1 < p < ∞. Subse-
quently Freniche, García-Vázquez and Rodríguez-Piazza [15, 16] exhibited
functions f ∈ whp(X) \ hp(X) and g ∈ wH p(X) \ H p(X) for 1 ≤ p < ∞ and
any X. Fairly concrete functions of this kind were provided in [24, 28], and
[30] contains the analogous results for the weak vs. strong Bergman norms.
In fact, the norms
) whp(Lp′
k · kwH p(X) ≁ k · kH p(X)
are non-equivalent on H p(X). Strict inclusions BM OA(X) wBM OA(X)
and BM OAC(X) wBM OA(X) for any infinite-dimensional X were ob-
tained in [24, 25]. A common feature of these examples for arbitrary X is
the use of Dvoretzky's ℓn
2 -theorem to transfer from the Hilbert space setting
to X.
The linearization from Theorem 14 can also be used for other purposes.
The following result from [4] concern weak conditional compactness on the
spaces wE(X).
Theorem 16. Suppose that E is a Banach space of analytic functions on
D that satisfy (W1) and (W2). Let ϕ : D → D be an analytic map and X
a Banach space that does not contain any subspaces linearly isomorphic to
ℓ1.
wE(X) → wE(X).
If Cϕ : E → E is compact, then fCϕ is weakly conditionally compact
The proof is analogous to that of Theorem 14, but instead apply [34] to
deduce the weak conditional compactness of the operator composition Uϕ.
16
JUSSI LAITILA AND HANS-OLAV TYLLI
Since wH 2(ℓ2) ≈ L(ℓ2) is non-reflexive one may also look for a characteri-
ϕ satisfies Shapiro's condition (1).
observation is not included in Proposition 15 since H 2 is reflexive.
zation of weakly compact fCϕ : wH 2(ℓ2) → wH 2(ℓ2). Note that the following
Proposition 17. fCϕ is weakly compact wH 2(ℓ2) → wH 2(ℓ2) if and only if
ness of fCϕ : wH 2(ℓ2) → wH 2(ℓ2) implies condition (1). As in the proof of
Proof. In view of Theorem 14 there remains to show that the weak compact-
Theorem 14 let Uϕ be the operator composition map
S 7→ S ◦ (Cϕ)∗V , L(V, ℓ2) → L(V, ℓ2),
where V = [δz : z ∈ D] = H 2. We get that Uϕ = χ−1 ◦ fCϕ ◦ χ is weakly
compact L(H 2, ℓ2) → L(H 2, ℓ2). It is known, see e.g. [44, Example 2.6], that
for such operator compositions this yields the compactness of (Cϕ)∗V on V .
Hence Cϕ is compact H 2 → H 2, so that (1) holds.
(cid:3)
The corresponding picture for the general class wE(X) is quite compli-
cated for reflexive E, and remains open, since the spaces wE(X) can also
be reflexive. For instance, the weak Hardy spaces wH 2(ℓp) ≈ L(H 2, ℓp) =
K(H 2, ℓp) are reflexive for 1 < p < 2 by Pitt's theorem [33, Prop. 2.c.3] and
[23, Sect. 2, Cor. 2]. Here K(X, Y ) denotes the space of compact operators
X → Y .
6. Compositions from weak to strong spaces
A different line of study concerns the mapping properties of composition
operators from weak to strong spaces of analytic functions on D, such as
wH p(X) → H p(X). This line was initiated by Laitila, Tylli and Wang in [30]
for the Hardy and Bergman spaces, and subsequently the approach has been
extended to weighted Bergman and Dirichlet spaces by Wang [55, 56, 57].
The question which motivated [30] came from S. Kaijser for X = ℓ2. Recall
from section 5 that e.g. wH 2(ℓ2) ≈ L(ℓ2) while H 2(ℓ2) is a separable Hilbert
space, so that boundedness of a composition operator wH 2(ℓ2) → H 2(ℓ2)
entails strong compression.
Somewhat surprisingly, the boundedness of Cϕ : wH p(X) → H p(X) for
2 ≤ p < ∞ is related to composition operators in the Hilbert-Schmidt class
on H 2. Recall from [48, 11] that Cϕ is a Hilbert-Schmidt operator on H 2
precisely when
kCϕk2
HS =ZT
1
1 − ϕ(ζ)2 dm(ζ) < ∞.
The following result is taken from [30], which also contains a formally similar
result for the vector-valued Bergman spaces. Note that results of this type
have no counterparts in the scalar-valued theory.
SURVEY OF COMPOSITION OPERATORS
17
Theorem 18. Let X be any infinite-dimensional complex Banach space.
(1) If kCϕkHS < ∞, then Cϕ is bounded wH p(X) → H p(X) for any p
satisfying 1 ≤ p < ∞.
(2) The norm kCϕ : wH p(X) → H p(X)k is equivalent to kCϕk2/p
HS for
2 < p < ∞.
(3) kCϕ : wH 2(X) → H 2(X)k = kCϕkHS.
Parts (2) and (3) are obtained by explicit computations for X = ℓ2. The
2 -theorem
extension to arbitrary Banach spaces X is based on Dvoretzky's ℓn
and coefficient multiplier results [14] corresponding to bounded operators
Xk
akzk 7→Xk
λkakzk, H 2 → H p,
where (λk) is a fixed sequence. By contrast to Theorem 18, fCϕ is bounded
wBM OA(ℓ2) → BM OA(ℓ2) if and only if the scalar-valued operator Cϕ is
bounded B → BM OA, see [30, Example 4.1], where B is the Bloch space.
Question 19. (a) Does part (2) of Theorem 18 extend to 1 ≤ p < 2? The
corresponding coefficient multiplier theorems H 2 → H p for 1 ≤ p < 2 are
not readily useful.
(b) Characterize the weakly compact compositions fCϕ from wH 1(X) to
H 1(X) if X is reflexive. It is possible to show that if
0<r<1Zϕ(rζ)>s
sup
lim
s→1
1
1 − ϕ(ζ)2 dm(ζ) = 0,
then fCϕ is weakly compact from wH 1(X) to H 1(X) (details omitted). Note
that fCϕ is never compact wH 1(X) → H 1(X) for infinite-dimensional X by
section 2.
One may also consider the composition operators f :
7→ f ◦ ϕ from strong
to weak spaces, e.g. as acting H p(X) → wH p(X), but this case does not
produce new qualitative phenomena. This follows from the factorization
H p(X)
cCϕ
wH p(X)
fCϕ
%❏❏❏❏❏❏❏❏❏
J
9ssssssssss
H p(X)
H p(X) → wH p(X), and for p = 1 and reflexive spaces X one obtains that
where cCϕ denotes the composition operator acting H p(X) → wH p(X) and
J : H p(X) → wH p(X) is the continuous inclusion. Hence, any cCϕ is bounded
cCϕ : H 1(X) → wH 1(X) is weakly compact if and only if ϕ satisfies (1), that
is, Cϕ : H 1 → H 1 is compact. (For the "only if" -part note that section 2
applies here.)
/
/
%
9
18
JUSSI LAITILA AND HANS-OLAV TYLLI
7. Operator-weighted composition operators
In the final section we briefly discuss extensions of weighted composition
operators to the vector-valued setting. Let ψ : D → C and ϕ : D → D be
given analytic maps. The weighted composition operator
Wψ,ϕ : f 7→ ψ · (f ◦ ϕ)
defines a linear map H(D) → H(D), where H(D) denotes the linear space
of analytic functions D → C. Clearly Wψ,ϕ = Mψ ◦ Cϕ, where Mψ is the
pointwise multiplier defined by Mψf = ψ · f on H(D). Thus W1,ϕ = Cϕ and
Wψ,id = Mψ.
Weighted composition operators Wψ,ϕ have been extensively studied on
a range of complex-valued analytic function spaces, and characterizations of
e.g. boundedness and compactness are known for many classical spaces. The
case of the weighted spaces H ∞
v was resolved by Contreras and Hernández-
Díaz in [9] and Montes-Rodríguez in [39]. For 1 < p < ∞ and H p there
is a Carleson measure characterization in [10], and the analogous results for
the Bergman space Ap are found in [12]. The case of BM OA can be found
in [26]. We refer to e.g. [17] and [19] for other types of function-theoretic
conditions for the boundedness of Wψ,ϕ on H p. Moreover, by [10] all weakly
compact weighted compositions Wψ,ϕ : H 1 → H 1 are compact.
Independently Manhas [37] and the authors [29] proposed the following
natural analogue of weighted composition operators in the vector-valued set-
ting. Let ϕ : D → D be an analytic self-map and ψ : D → L(X, Y ) an ana-
lytic operator-valued map, where X and Y are complex Banach spaces. Here
L(X, Y ) denotes the space of bounded linear operators X → Y . Define the
operator-weighted composition operator Wψ,ϕ by f 7→ ψ(f ◦ ϕ), that is,
(Wψ,ϕ(f ))(z) = ψ(z)(f (ϕ(z)),
z ∈ D,
for analytic functions f : D → X. Note that z 7→ ψ(z)(f (ϕ(z)) is an analytic
map D → Y , so that Wψ,ϕ is a linear map H(D, X) → H(D, Y ). Again
Wψ,ϕ = Mψ ◦ Cϕ, where Mψ denotes the operator-valued pointwise multi-
plier defined by (Mψf )(z) = ψ(z)(f (z)) from H(D, X) to H(D, X). Thus
the operator-weighted composition operators form a much larger class com-
pared to its scalar-valued relative. Operator-weighted compositions appear
naturally:
for a large class of Banach spaces X all linear onto isometries
H ∞(X) → H ∞(X) have the form Wψ,ϕ, where ψ(z) ≡ U is a fixed onto
isometry of X and ϕ is an automorphism of D, see [32] and [8]. The above
definition of Wψ,ϕ is analogous to that of weighted compositions on spaces
C(K, X) of continuous functions, see e.g. [22], where K is a compact Haus-
dorff space.
The present knowledge of operator-weighted compositions is fairly rudi-
v (X) spaces and their
mentary and most of the results deal with weighted H ∞
SURVEY OF COMPOSITION OPERATORS
19
locally convex variants. Characterizations of boundedness and (weak) com-
pactness of Wψ,ϕ between H ∞
v (X) spaces were obtained in [29], and results
for certain locally convex spaces of analytic vector-valued functions are found
in [37] and [6]. Boundedness of the operator multipliers Mψ in related set-
tings has been considered e.g. in [36, 37, 38, 57].
We next state the main results from [29], which are vector-valued exten-
sions of scalar results from [9] and [39]. For a bounded continuous weight
function v : D → (0, ∞), put
v(z) = (sup{f (z) : kf kH∞
v ≤ 1})−1.
If ψ : D → L(X, Y ) is a given analytic operator-valued map define the auxil-
iary linear map Tψ by x 7→ ψ(·)x. It follows that Tψ is bounded X → H ∞
w (Y )
whenever Wψ,ϕ is bounded H ∞
v (X) → H ∞
w (Y ).
Theorem 20. (1) Let v and w be weight functions. Then
kWψ,ϕ : H ∞
v (X) → H ∞
w (Y )k = sup
z∈D
w(z)
v(ϕ(z))
kψ(z)kL(X,Y ).
(2) Assume that v and w are radial weight functions. Then Wψ,ϕ is com-
pact (respectively, weakly compact) H ∞
v (X) → H ∞
w (Y ) if and only if both
lim sup
ϕ(z)→1
w(z)
v(ϕ(z))
kψ(z)kL(X,Y ) = 0,
(4)
and Tψ is compact (respectively, weakly compact) X → H ∞
w (Y ).
Parts of Theorem 20 were independently obtained in [37] together with
other results. Clearly the case X = Y and ψ(z) ≡ IX yields the bound-
edness and weak compactness results from [35, 4] for fCϕ, since Tψ is then
(essentially) IX.
Theorem 20 points to some striking differences between scalar- and vector-
valued weighted compositions as well as between operator-weighted and stan-
dard composition operators. For instance, the auxiliary operators Tψ play no
role for scalar weighted compositions. Moreover, Wψ,ϕ can easily be compact
H ∞
w (Y ) for infinite-dimensional spaces X and Y . For instance,
let kϕk∞ < 1 and ψ(z) ≡ U ∈ K(X, Y ), whence (4) holds and Tψ is compact
X → H ∞
w (X)
do not always factor through IX as in Proposition 1, but (F2) is replaced by
the following factorization for any z ∈ D:
w (Y ). Operator-weighted compositions Wψ,ϕ : H ∞
v (X) → H ∞
v (X) → H ∞
H ∞
v (X)
Wψ,ϕ
/ H ∞
w (Y )
j
X
ψ(z)
δz
/ Y
/
O
O
/
20
JUSSI LAITILA AND HANS-OLAV TYLLI
There are also examples where Wψ,ϕ is weakly compact H ∞
w (X)
but not compact. For this one may use the fact [29, Thm. 4.4] that if w is
a radial weight and ψ ∈ H 0
w(L(X, Y ))), then Tψ is compact (respectively,
weakly compact) if and only if
v (X) → H ∞
ψ(D) ⊂ K(X, Y )
(respectively, ψ(D) ⊂ W (X, Y )).
Here H 0
nomials in H ∞
Y . There are further differences between H ∞
ting as studied in [6].
w(L(X, Y ))) denotes the closure of the analytic L(X, Y )-valued poly-
w (L(X, Y ))), and W (X, Y ) the weakly compact operators X →
v (X) and the locally convex set-
The following problem stated in [29] appears quite challenging.
Question 21. Characterize boundedness and (weak) compactness of operator-
weighted compositions Wψ,ϕ : H p(X) → H p(Y ) as well as Ap(X) → Ap(Y )
for 1 ≤ p < ∞. The argument from [10] based on Carleson measure tech-
niques does not readily extend to the vector-valued setting.
References
[1] R. Aron, P. Galindo and M. Lindström, Compact homomorphisms between alge-
bras of analytic functions, Studia Math. 123 (1997), 235 -- 247.
[2] O. Blasco, Boundary values of vector-valued harmonic functions considered as
operators, Studia Math. LXXXVI (1987), 19 -- 33.
[3] O. Blasco, Remarks on vector-valued BMOA and vector-valued multipliers, Posi-
tivity 4 (2000), 339 -- 356.
[4] J. Bonet, P. Domański and M. Lindström, Weakly compact composition operators
on analytic vector-valued function spaces, Ann. Acad. Sci. Fenn. Math. 26 (2001),
233 -- 248.
[5] J. Bonet and M. Friz, Weakly compact composition operators on locally convex
spaces, Math. Nachr. 245 (2002), 26 -- 44.
[6] J. Bonet, M.C. Gómez-Collado, D. Jornet and E. Wolf, Operator-weighted compo-
sition operators between weighted spaces of vector-valued analytic functions, Ann.
Acad. Sci. Fenn. Math. 37 (2012), 319 -- 338.
[7] P.S. Bourdon, J.A. Cima and A.L. Matheson, Compact composition operators on
BMOA, Trans. Amer. Math. Soc. 351 (1999), 2183 -- 2196.
[8] M. Cambern and K. Jarosz, Multipliers and isometries in H ∞
E , Bull. London
Math. Soc. 22 (1990), 463 -- 466.
[9] M. D. Contreras and A.G. Hernández-Díaz, Weighted composition operators in
weighted Banach spaces of analytic functions, J. Austral. Math. Soc. 69 (2000),
41 -- 60.
[10] M.D. Contreras and A.G. Hernández-Díaz, Weighted composition operators on
Hardy spaces, J. Math. Anal. Appl. 263 (2001), 224 -- 233.
[11] C.C. Cowen and B.D. MacCluer, Composition Operators on Spaces of Analytic
Functions, CRC Press, Boca Raton 1995.
[12] Z. Cucković and R. Zhao, Weighted composition operators on the Bergman space,
J. London Math. Soc. 70 (2004), 499 -- 511.
[13] N. Dinculeanu, Vector Measures, Pergamon Press, 1967.
[14] P.L. Duren, On the multipliers of H p spaces, Proc. Amer. Math. Soc. 22 (1969),
24 -- 27.
SURVEY OF COMPOSITION OPERATORS
21
[15] F.J. Freniche, J.C. García-Vázquez and L. Rodríguez-Piazza, The failure of Fa-
tou's theorem on Poisson integrals of Pettis integrable functions, J. Funct. Anal.
160 (1998), 28-41.
[16] F.J. Freniche, J.C. García-Vázquez and L. Rodríguez-Piazza, Operators into
Hardy spaces and analytic Pettis integrable functions,
in: Recent Progress in
Functional Analysis, North-Holland Mathematical Studies vol. 189 (2001), North-
Holland, pp. 349 -- 362.
[17] E.A. Gallardo-Gutiérrez and J.R. Partington, The role of BMOA in the bounded-
ness of weighted composition operators, J. Funct. Anal. 258 (2010), 3593 -- 3603.
[18] J. Garnett, Bounded Analytic Functions, Academic Press, 1981.
[19] G. Gunatillake, Compact weighted composition operators on the Hardy space,
Proc. Amer. Math. Soc. 136 (2008), 2895 -- 2899.
[20] E. Hille and R.S. Phillips, Functional analysis and semi-groups, Amer. Math. Soc.
Colloq. Publ. 31, Amer. Math. Soc., Providence, 1975.
[21] W.E. Hornor and J.E. Jamison, Isometrically equivalent composition operators on
spaces of analytic vector-valued functions, Glasgow Math. J. 41 (1999), 441 -- 451.
[22] J.E. Jamison and M. Rajagopalan, Weighted composition operators on C(K, E),
J. Operator Theory 19 (1988), 307 -- 317.
[23] N.J. Kalton, Spaces of compact operators, Math. Ann. 208 (1974), 267 -- 278.
[24] J. Laitila, Weakly compact composition operators on vector-valued BMOA, J.
Math. Anal. Appl. 308 (2005), 730 -- 745.
[25] J. Laitila, Composition operators and vector-valued BMOA, Integral Equ. Opera-
tor Theory 58 (2007), 487 -- 502.
[26] J. Laitila, Weighted composition operators on BMOA, Comput. Methods Funct.
Theory 9 (2009), 27 -- 46.
[27] J. Laitila, P. Nieminen, E. Saksman and H.-O. Tylli, Compact and weakly com-
pact composition operators on BMOA, Complex Anal. Oper. Theory 7 (2013),
163 -- 181.
[28] J. Laitila and H.-O. Tylli, Composition operators on vector-valued harmonic func-
tions and Cauchy transforms, Indiana Univ. Math. J. 55 (2006), 719 -- 746.
[29] J. Laitila and H.-O. Tylli, Operator-weighted composition operators on vector-
valued analytic function spaces, Illinois J. Math. 53 (2009), 1019 -- 1032.
[30] J. Laitila, H.-O. Tylli and M. Wang, Composition operators from weak to strong
spaces of vector-valued analytic functions, J. Operator Theory 62 (2009), 281 -- 295.
[31] P. Lefevre, D. Li, H. Queffelec and L. Rodriguez-Piazza, Some new properties of
composition operators associated with the lens map, Israel J. Math. 195 (2013),
801 -- 824.
[32] P.-K. Lin, The isometries of H ∞(E), Pacific J. Math. 143 (1990), 69 -- 77.
[33] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces I, Springer-Verlag, 1977.
[34] M. Lindström and G. Schlüchtermann, Composition of operator ideals, Math.
Scand. 84 (1999), 284 -- 296.
[35] P. Liu, E. Saksman and H.-O. Tylli, Small composition operators on analytic
vector-valued function spaces, Pacific J. Math. 184 (1998), 295 -- 309.
[36] J. S. Manhas, Multiplication operators on weighted locally convex spaces of vector-
valued analytic functions, Southeast Asian Bull. Math. 27 (2003), 649 -- 660.
[37] J.S. Manhas, Weighted composition operators on weighted spaces of vector-valued
analytic functions, J. Korean Math. Soc. 45 (2008), 1203 -- 1220.
[38] J.S. Manhas, Weighted composition operators between weighted spaces of vector-
valued holomorphic functions on Banach spaces, Appl. Math. Comput. 218
(2011), 929 -- 934.
22
JUSSI LAITILA AND HANS-OLAV TYLLI
[39] A. Montes-Rodríguez, Weighted composition operators on weighted Banach spaces
of analytic functions, J. London Math. Soc. 61 (2000), 872 -- 884.
[40] J. Mujica, Linearization of bounded holomorphic mappings on Banach spaces,
Trans. Amer. Math. Soc. 324 (1991), 295 -- 309.
[41] P.F.X. Müller, Isomorphisms between H 1-spaces, Monografie Matematyczne vol.
66, Birkhäuser, 2005.
[42] K. Ng, On a theorem of Dixmier, Math. Scand. 29 (1971), 279 -- 280.
[43] A. Pietsch, Operator Ideals, North-Holland, 1980.
[44] E. Saksman and H.-O. Tylli, Multiplications and elementary operators in the Ba-
nach space setting,
in: Methods in Banach space theory, London Math. Soc.
Lecture Note Ser. vol. 337 (2006), Cambridge University Press, Cambridge, pp.
253-292.
[45] D. Sarason, Weak compactness of holomorphic composition operators on H 1, in:
Functional Analysis and Operator Theory, Lecture Notes in Mathematics 1511
(1992), Springer-Verlag, pp. 75 -- 79.
[46] J.H. Shapiro, The essential norm of a composition operator, Ann. Math. 125
(1987), 375 -- 404.
[47] J.H. Shapiro, Composition Operators and Classical Function Theory, Springer-
Verlag, New York, 1993.
[48] J.H. Shapiro and P.D. Taylor, Compact, nuclear and Hilbert-Schmidt composition
operators on H 2, Indiana Univ. Math. J. 23 (1973/74), 471 -- 496.
[49] S.D. Sharma and U. Bhanu, Composition operators on vector-valued Hardy spaces,
Extracta Math. 14 (1999), 31 -- 39.
[50] W. Smith, Compactness of composition operators on BMOA, Proc. Amer. Math.
Soc. 127 (1999), 2715 -- 2725.
[51] C. Stanton, Counting functions and majorization theorems for Jensen measures,
Pacific J. Math. 125 (1986), 459 -- 468.
[52] M. Tjani, Compact composition operators on some Möbius invariant Banach
spaces, Ph.D. thesis, Michigan State University, 1996.
[53] S.-I. Ueki, Composition operators on the Fock space of vector-valued analytic func-
tions, Ars Combin. 100 (2011), 161 -- 167.
[54] M. Wang, Composition operators on analytic vector-valued Nevanlinna classes,
Acta Math. Sci. 25 (2005), 771 -- 780.
[55] M. Wang, Composition operators between vector-valued weighted Bergman spaces,
Far East J. Math. Sci. 26 (2007), 419 -- 431.
[56] M. Wang, Composition operators between different Dirichlet spaces, Complex Var.
Elliptic Equ. 53 (2008), 1093 -- 1102.
[57] M. Wang, Weighted composition operators between Dirichlet spaces, Acta Math.
Sci. Ser. B Engl. Ed. 31 (2011), 641 -- 651.
Jussi Laitila, Department of Biosciences, P.O. Box 65, FI-00014 University
of Helsinki, Helsinki, Finland
E-mail address: [email protected]
Hans-Olav Tylli, Department of Mathematics and Statistics, P.O. Box 68,
FI-00014 University of Helsinki, Helsinki, Finland
E-mail address: [email protected]
|
1301.4819 | 1 | 1301 | 2013-01-21T11:08:30 | Smoothing properties of the discrete fractional maximal operator on Besov and Triebel--Lizorkin spaces | [
"math.FA"
] | Motivated by the results of Korry and Kinnunen and Saksman, we study the behaviour of the discrete fractional maximal operator on fractional Hajlasz spaces, Hajlasz-Besov and Hajlasz-Triebel-Lizorkin spaces on metric measure spaces. We show that the discrete fractional maximal operator maps these spaces to the spaces of the same type with higher smoothness. Our results extend and unify aforementioned results. We present our results in general setting, but they are new already in the Euclidean case. | math.FA | math |
SMOOTHING PROPERTIES OF THE DISCRETE
FRACTIONAL MAXIMAL OPERATOR ON BESOV AND
TRIEBEL -- LIZORKIN SPACES
TONI HEIKKINEN AND HELI TUOMINEN
Abstract. Motivated by the results of Korry and Kinnunen and Saks-
man, we study the behaviour of the discrete fractional maximal operator
on fractional Haj lasz spaces, Haj lasz -- Besov and Haj lasz -- Triebel -- Lizorkin
spaces on metric measure spaces. We show that the discrete fractional
maximal operator maps these spaces to the spaces of the same type with
higher smoothness. Our results extend and unify aforementioned results.
We present our results in general setting, but they are new already in the
Euclidean case.
1. Introduction
Maximal functions are standard tools in harmonic analysis. They are usually
used to estimate absolute size, but recently there has been interest in studying
their regularity properties, see [1], [2], [3], [11], [12], [13], [15], [17], [18], [19],
[21], [22], [23], [25], [26], [28]. A starting point was [17], where Kinnunen
observed that the Hardy-Littlewood maximal operator is bounded on W 1,p(Rn)
for 1 < p ≤ ∞. In [22] and [23] Korry extended this result by showing that the
maximal operator preserves also fractional Sobolev spaces as well as Besov and
Triebel -- Lizorkin spaces. Another kind of extension was given in [20], where
Kinnunen and Saksman showed that the fractional maximal operator Mα,
defined by
rα
B(x, r) ZB(x,r)
Mα u(x) = sup
r>0
u(y) dy,
is bounded from W 1,p(Rn) to W 1,p∗(Rn), where p∗ = np/(n − αp), and from
Lp(Rn) to W 1,q(Rn), where q = np/(n−(α −1)p) and W 1,q(Rn) is the homoge-
nous Sobolev space. These results indicate that Mα has similar smoothing
properties as the Riesz potential.
It is natural to ask whether these results can be seen as special cases of the
behaviour of the fractional maximal operator on Besov and Triebel -- Lizorkin
spaces. In this paper we show that this is indeed the case, and that all these
results can be obtained by the same rather simple method.
Instead of the
Date: May 31, 2018.
2010 Mathematics Subject Classification. 42B25, 46E35.
Key words and phrases. Besov space, fractional maximal function, fractional Sobolev
space, Triebel -- Lizorkin space, metric measure space.
The research was supported by the Academy of Finland, grant no. 135561.
1
2
TONI HEIKKINEN AND HELI TUOMINEN
standard fractional maximal operator, we consider its variant, the so-called
discrete fractional maximal operator M∗
α. This allows us to present our results
in a setting of doubling metric measure spaces. In this generality, the standard
fractional maximal operator behaves quite badly. Indeed, one can construct
spaces, where the fractional maximal function of a Lipschitz function fails to be
continuous, see [3] and [15]. Since M∗
α and Mα are comparable, for practical
purposes it does not matter which one we choose. The discrete fractional
maximal operator was introduced in [18] and further studied in [21], [1] and
[13].
Among the many possible definitions of Besov and Triebel -- Lizorkin spaces,
the most suitable for our purposes is the one based on Haj lasz type pointwise
inequalities. This approach, introduced by Koskela, Yang and Zhou in [24],
provides a new point of view to the classical Besov and Triebel -- Lizorkin spaces.
On the other hand, it allows these spaces to be defined in the setting of metric
measure spaces.
By employing this definition, we can prove very general results using only
simple "telescoping" arguments and Poincar´e type inequalities. As special
cases, we obtain versions of the results of Kinnunen and Saksman as well as
those of Korry, see Remark 3.4 and Theorems 4.5 and 4.6. We prove our results
in doubling metric measure spaces but they are new even in Euclidean spaces.
Our main results (Theorems 4.3 and 4.4) imply that if α ≥ 0 and 0 < s+α < 1,
then M∗
p,q (Rn) for n/(n + s) < p, q < ∞ and
from Bs
p,q (Rn) for n/(n + s) < p < ∞, 0 < q < ∞, see Section 4
for the definition of Triebel -- Lizorkin and Besov spaces.
α is bounded from F s
p,q(Rn) to Bs+α
p,q(Rn) to F s+α
2. Preliminaries and notation
We assume that X = (X, d, µ) is a metric measure space equipped with a
metric d and a Borel regular outer measure µ, which satisfies 0 < µ(U) < ∞
whenever U is nonempty, open and bounded. We assume that the measure
is doubling, that is, there exists a fixed constant cd > 0, called the doubling
constant, such that
(2.1)
µ(B(x, 2r)) ≤ cdµ(B(x, r))
for every ball B(x, r) = {y ∈ X : d(y, x) < r}.
The doubling condition implies that
(2.2)
µ(B(y, r))
µ(B(x, R))
≥ C(cid:16) r
R(cid:17)Q
for every 0 < r ≤ R and y ∈ B(x, R) for some C and Q > 1 that only depend
on cD. In fact, we may take Q = log2 cd.
For the boundedness of the fractional maximal operator in Lp, we have to
assume, in Theorems 2.1 and 3.3.(b), that the measure µ satisfies the lower
bound condition
(2.3)
µ(B(x, r)) ≥ clrQ
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
3
with some constant cl > 0 for all x ∈ X and r > 0.
Throughout the paper, C will denote a positive constant whose value is not
necessarily the same at each occurrence.
The fractional maximal function. Let α ≥ 0. The fractional maximal
function of a locally integrable function u is
(2.4)
Mα u(x) = sup
r>0
rαZB(x,r)
u dµ,
where uB = RB u dµ = 1
α = 0, we have the usual Hardy-Littlewood maximal function
µ(B) RB u dµ is the integral average of u over B. For
M u(x) = sup
r>0ZB(x,r)
u dµ.
The following Sobolev type inequality for the fractional maximal operator fol-
lows easily from the boundedness of the Hardy-Littlewood maximal operator
in Lp, for the proof, see [4], [6] or [15].
Theorem 2.1. Assume that the measure lower bound condition holds. If p > 1
and 0 < α < Q/p, then there is a constant C > 0, depending only on the
doubling constant, constant in the measure lower bound, p and α, such that
for every u ∈ Lp(X) with p∗ = Qp/(Q − αp).
k Mα ukLp∗ (X) ≤ CkukLp(X),
Remark 2.2. If u is only locally integrable, then Mα u may well be identically
infinite. However, if Mα u(x0) < ∞ for some x0 ∈ X, then Mα u(x) < ∞ for
almost every x. This follows from the estimate
rαZB(x,r)
u dµ ≤
µ(B(x0, r + d(x, x0)))
µ(B(x, r))
Mα u(x0)
combined with the doubling condition and the fact that
lim
r→0
rαZB(x,r)
u dµ < ∞,
whenever x is a Lebesgue point of u.
The discrete fractional maximal function. We begin the construction of
the discrete maximal function with a covering of the space. Let r > 0. Since
the measure is doubling, there are balls B(xi, r), i = 1, 2, . . . , such that
X =
∞
[i=1
B(xi, r) and
∞
Xi=1
χB(xi,6r) ≤ N < ∞,
where χB(xi,6r) is the characteristic function of the ball B(xi, 6r). This means
that the dilated balls B(xi, 6r), i = 1, 2, . . . , are of bounded overlap. The
constant N depends only on the doubling constant and, in particular, it is
independent of r.
4
TONI HEIKKINEN AND HELI TUOMINEN
Then we construct a partition of unity subordinate to the covering B(xi, r),
i = 1, 2, . . . , of X. Indeed, there is a family of functions ϕi, i = 1, 2, . . . , such
that 0 ≤ ϕi ≤ 1, ϕi = 0 in X \ B(xi, 6r), ϕi ≥ ν in B(xi, 3r), ϕi is Lipschitz
with constant L/r with ν and L depending only on the doubling constant, and
∞
Xi=1
ϕi(x) = 1
for every x ∈ X.
The discrete convolution of a locally integrable function u at the scale 3r is
∞
for every x ∈ X, and we write uα
ur(x) =
ϕi(x)uB(xi,3r)
Xi=1
r = rαur.
Let rj, j = 1, 2, . . . be an enumeration of the positive rationals and let balls
B(xi, rj), i = 1, 2, . . . be a covering of X as above. The discrete fractional
maximal function of u in X is
M∗
uα
α u(x) = sup
rj(x)
j
for every x ∈ X. For α = 0, we obtain the Hardy-Littlewood type discrete
maximal function M∗ studied in [18], [21] and [1]. The discrete fractional
maximal function is easily seen to be comparable to the standard fractional
maximal function, see [13].
3. Fractional Haj lasz spaces
Let u be a measurable function and let s ≥ 0. A nonnegative measurable
function g is an s-Haj lasz gradient of u if there exists E ⊂ X with µ(E) = 0
such that for all x, y ∈ X \ E,
u(x) − u(y) ≤ d(x, y)s(g(x) + g(y)).
(3.1)
The collection of all s-Haj lasz gradients of u is denoted by Ds(u). A homoge-
neous Haj lasz space M s,p(X) consists of measurable functions u such that
kuk M s,p(X) = inf
g∈Ds(u)
kgkLp(X)
is finite. The Haj lasz space M s,p(X) is
norm
M s,p(X) ∩ Lp(X) equipped with the
kukM s,p(X) = kukLp(X) + kuk M s,p(X).
The space M 1,p(X), a counterpart of a Sobolev space in metric measure space,
was introduced in [9], see also [10]. The fractional spaces M s,p(X) were
introduced in [30] and studied for example in [16] and [14]. Notice that
M 0,p(X) = Lp(X).
The pointwise definition of the Haj lasz spaces implies the validity of Sobolev-
Poincar´e type inequalities without the assumption that the space admits any
weak Poincar´e inequality.
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
5
Lemma 3.1 ([7]). Let s ∈ [0, ∞) and let p ∈ (0, Q/s). There exists a constant
C such that for all measurable functions u with g ∈ Ds(u), all x ∈ X and
r > 0,
(3.2)
inf
c∈R(cid:18)ZB(x,r)
u(y) − cp∗(s) dµ(y)(cid:19)1/p∗(s)
≤ Crs(cid:18)ZB(x,2r)
gp dµ(cid:19)1/p
,
where p∗(s) = Qp/(Q − sp).
Moreover, if p ≥ Q/(Q + s) and g ∈ Ds(u) ∩ Lp(X), then (3.2) implies that
u is locally integrable and that
(3.3)
ZB(x,r)
u − uB(x,r) dµ ≤ Crs(cid:18)ZB(x,2r)
gp dµ(cid:19)1/p
.
For the case s = 1, see [9] and [10].
In the next theorem, we use the following simple result. If ui, i ∈ N, are
measurable functions with a common s-Haj lasz gradient g and u = supi ui is
finite almost everywhere, then g is an s-Haj lasz gradient of u.
Theorem 3.2. Assume that M∗
s-Haj lasz gradient of u.
α u 6≡ ∞. Let t ≥ Q/(Q + s) and let g be an
a) If 0 < s + α ≤ 1, then there exists a constant C > 0 such that
is an (s + α)-Haj lasz gradient of M∗
g = C(cid:0)M gt(cid:1)1/t
α u.
b) If s + α > 1, then there exists a constant C > 0 such that
is a 1-Haj lasz gradient of M∗
α u.
g = C(cid:0)Mt(s+α−1) gt(cid:1)1/t
Proof. We begin by proving the claims for uα
let x, y ∈ X.
r . Let r > 0, let g ∈ Ds(u) and
Assume first that r ≥ d(x, y). Let Ixy be a set of indices i for which x
or y belongs to B(xi, 6r). Then, for each i ∈ Ixy, B(xi, 3r) ⊂ B(x, 10r) ⊂
B(xi, 17r). This together with the doubling condition, the properties of the
functions ϕi, the fact that there are bounded number of indices in Ixy and
Poincar´e inequality (3.3) implies that
uα
r (x) − uα
r (y) ≤ rα Xi∈Ixy
ϕi(x) − ϕi(y)uB(xi,3r) − uB(x,10r)
(3.4)
u − uB(xi,3r) dµ
ZB(x,10r)
≤ Crα−1d(x, y) Xi∈Ixy
≤ Crα−1d(x, y)ZB(x,10r)
≤ Crs+α−1d(x, y)(cid:18)ZB(x,20r)
u − uB(x,10r) dµ
gt dµ(cid:19)1/t
.
6
TONI HEIKKINEN AND HELI TUOMINEN
If 0 < s + α ≤ 1, then by (3.4) and the assumption r ≥ d(x, y), we have that
uα
r (x) − uα
r (y) ≤ Cd(x, y)s+α(cid:0) M gt(x)(cid:1)1/t.
If s + α > 1, then by (3.4),
uα
r (x) − uα
r (y) ≤ Cd(x, y)(cid:0) Mt(s+α−1) gt(x)(cid:1)1/t.
This shows that Haj lasz gradient inequality (3.1) with desired exponent holds
when r ≥ d(x, y).
Assume then that r < d(x, y). Let R = d(x, y). Then B(y, r) ⊂ B(x, 2R)
and
(3.5)
uα
r (x) − uα
r (y) ≤ rα(cid:16)Xi∈Ix
+Xi∈Iy
ϕi(x)uB(xi,3r) − uB(x,9R)
ϕi(y)uB(xi,3r) − uB(x,9R)(cid:17),
where Ix is a set of indices i for which x belongs to B(xi, 6r) and Iy the
corresponding set for y. Let k ∈ N be the smallest integer such that 2kr ≥ R.
Assume first that 0 < s + α ≤ 1. If i ∈ Ix, then
(3.6)
uB(xi,3r) − uB(x,9R) ≤ uB(xi,3r) − uB(x,9r) +
+ uB(x,2k9r) − uB(x,9R).
k
Xi=1
uB(x,2i9r) − uB(x,2i−19r)
By the doubling condition and Poincar´e inequality (3.3), we have
rαuB(xi,3r) − uB(x,9r) ≤ CrαZB(x,9r)
u − uB(x,9r) dµ
gt dµ(cid:19)1/t
≤ Crs+α(cid:18)ZB(x,18r)
≤ CRs+α(cid:0) M gt(x)(cid:1)1/t
,
and, by the doubling condition, Poincar´e inequality (3.3), the fact that r ≤ 2i9r
for all i, and the selection of k,
uB(x,2i9r) − uB(x,2i−19r) ≤ Crα
u − uB(x,2i9r) dµ
(3.7)
rα
k
Xi=1
(3.8)
k
Xi=1ZB(x,2i9r)
(2i9r)s+α(cid:18)ZB(x,2i+19r)
Xi=1
k
≤ C
≤ CRs+α(cid:0) M gt(x)(cid:1)1/t
.
gt dµ(cid:19)1/t
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
7
Similarly we obtain that
(3.9)
rαuB(x,2k9r) − uB(x,9R) ≤ CRs+α(cid:18)ZB(x,36R)
≤ CRs+α(cid:0) M gt(x)(cid:1)1/t
.
gt dµ(cid:19)1/t
If i ∈ Iy, we use balls B(y, 2i9r) instead of balls B(x, 2i9r) in (3.6). Estimates
corresponding (3.7) and (3.8) are as above (x replaced by y) and, corresponding
to (3.9),
rαuB(y,2k9r) − uB(x,9R) ≤ CRs+α(cid:18)ZB(x,38R)
gt dµ(cid:19)1/t
(3.10)
Now, by (3.5)-(3.10) and the fact R = d(x, y), we have
≤ CRs+α(cid:0) M gt(x)(cid:1)1/t.
rαuB(xi,3r) − uB(x,R) ≤ Cd(x, y)s+α(cid:16)(cid:0) M gt(x)(cid:1)1/t + M gt(y)(cid:1)1/t(cid:17).
If s + α > 1, then similar estimates as above show that if i ∈ Ix ∪ Iy, then
(3.11)
rαuB(xi,3r) −uB(x,R) ≤ Cd(x, y)(cid:16)(cid:0) Mt(s+α−1) gt(x)(cid:1)1/t
+(cid:0) Mt(s+α−1) gt(y)(cid:1)1/t(cid:17).
These estimates together with (3.5) and the fact that there are bounded num-
ber of indices in Ix and Iy imply that Haj lasz gradient inequality (3.1) with
desired exponent holds when r < d(x, y).
The claim for uα
r follows from the estimates above and for M∗
discussion before the theorem.
Theorem 3.3. Let Q/(Q + s) < p < ∞.
a) If 0 < s + α ≤ 1, there exists a constant C > 0, such that
k M∗
α uk M s+α,p(X) ≤ Ckuk M s,p(X)
α u from the
(cid:3)
for all u ∈ M s,p(X) with M∗
α u 6≡ ∞.
b) If 1 < s + α ≤ 1 + Q/p and the measure lower bound condition holds,
there exists a constant C > 0 such that
k M∗
α uk M 1,q(X) ≤ Ckuk M s,p(X),
where q = Qp/(Q − (s + α − 1)p), for all u ∈ M s,p(X) with M∗
α u 6≡ ∞.
Proof. a) Let Q/(Q + s) ≤ t < p. By Theorem 3.2, the function C(M gt)1/t
is an (s + α)-gradient of M∗
α u. Since g ∈ Lp(X), the claim follows from the
boundedness of the Hardy -- Littlewood maximal operator in Lq(X) for q > 1.
b) Let Q/(Q + s) ≤ t < p. By Theorem 3.2, the function (Mt(s+α−1) gt)1/t
α u. Since g ∈ Lp(X), the claim follows from Theorem
(cid:3)
is a 1-gradient of M∗
2.1.
8
TONI HEIKKINEN AND HELI TUOMINEN
Remark 3.4. In the cases s = 0 and s = 1 of Theorem 3.3.b), we obtain
counterparts of the results of Kinnunen and Saksman.
Remark 3.5. As a special case of Theorems 3.2 and 3.3 we obtain boundedness
results for the discrete maximal operator M∗ in M s,p(X). If 0 < s ≤ 1, then
g = C(M gt)1/t is an s-Haj lasz gradient of M∗ u for all t ≥ Q/(Q + s) and
k M∗ uk M s,p(X) ≤ Ckuk M s,p(X)
for all u ∈ M s,p(X), p > Q/(Q + s).
If 1 < s ≤ 1 + Q/p, then g = C(Mt(s−1) gt)1/t is a 1-Haj lasz gradient of
M∗ u for all t ≥ Q/(Q + s) and
k M∗ uk M 1,q(X) ≤ Ckuk M s,p(X),
where q = Qp/(Q − (s − 1)p), for all u ∈ M s,p(X).
Moreover, when s = 1, we obtain boundedness results for the discrete max-
imal operator M∗ in (homogeneous) Haj lasz spaces M 1,p(X), proved earlier
for M 1,p(X) in [18] and [21].
4. Haj lasz -- Besov and Haj lasz -- Triebel -- Lizorkin spaces
Let u be a measurable function and let s ∈ (0, ∞). Following [24], we say
that a sequence of nonnegative measurable functions (gk)k∈Z is a fractional
s-Haj lasz gradient of u if there exists E ⊂ X with µ(E) = 0 such that
u(x) − u(y) ≤ d(x, y)s(gk(x) + gk(y))
for all k ∈ Z and all x, y ∈ X \ E satisfying 2−k−1 ≤ d(x, y) < 2−k. The
collection of all fractional s-Haj lasz gradients of u is denoted by Ds(u).
For p ∈ (0, ∞), q ∈ (0, ∞] and a sequence (fk)k∈Z of measurable functions,
we write
and
k(fk)k∈ZkLp(X, lq) = (cid:13)(cid:13)k (fk)k∈Z klq(cid:13)(cid:13)Lp(X)
k(fk)k∈Zklq(Lp(X)) = (cid:13)(cid:13)(cid:0)kfkkLp(X)(cid:1)k∈Z(cid:13)(cid:13)lq ,
where k(fk)klq = (Pk∈Z fkq)1/q if 0 < q < ∞ and k(fk)kl∞ = supk∈Z fk.
The homogeneous Haj lasz -- Triebel -- Lizorkin space M s
p,q(X) consists of mea-
surable functions u such that
kuk M s
p,q(X) = inf
(gk)∈Ds(u)
k(gk)kLp(X, lq)
is finite. The Haj lasz -- Triebel -- Lizorkin space M s
equipped with the norm
p,q(X) is
M s
p,q(X) ∩ Lp(X)
kukM s
p,q(X) = kukLp(X) + kuk M s
p,q(X).
The homogeneous Haj lasz -- Besov space N s
p,q(X) consists of measurable func-
tions u such that
kuk N s
p,q(X) = inf
(gk)∈Ds(u)
k(gk)klq(Lp(X))
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
9
is finite and the Haj lasz -- Besov space N s
with the norm
p,q(X) is
N s
p,q(X) ∩ Lp(X) equipped
kukN s
p,q(X) = kukLp(X) + kuk N s
p,q(X).
M s
Notice that
p,∞(X) is the homogeneous fractional Haj lasz space M s,p(X),
for the simple proof, see [24, Prop. 2.1]. The homogeneous Haj lasz -- Triebel --
M s
p,q(Rn) coincides with the classical homogeneous Triebel-
Lizorkin space
Lizorkin space F s
p,q(Rn) for s ∈ (0, 1), p ∈ (n/(n + s), ∞) and q ∈ (n/(n +
p,q(Rn) coincides with the classical homogeneous Besov
s), ∞]. Similarly,
space Bs
p,q(Rn) for s ∈ (0, 1), p ∈ (n/(n + s), ∞) and q ∈ (0, ∞] by [24, Thm
1.2]. For the definitions of F s
p,q(Rn), see [29].
p,q(Rn) and Bs
N s
If X supports a (weak) (1, p)-Poincar´e inequality with p ∈ (1, ∞), then for
p,q(X) are trivial, that is, they contain
all q ∈ (0, ∞), the spaces M 1
p,q(X) and N 1
only constant functions, see [7, Thm 4.1].
Lemma 4.1 ([7]). Let s ∈ (0, ∞) and p ∈ (0, Q/s). Then for every ε, ε′ ∈
(0, s) with ε < ε′ there exists a constant C > 0 such that for all measurable
functions u with (gj) ∈ Ds(u), x ∈ X and k ∈ Z,
(4.1)
inf
c∈R(cid:18)ZB(x,2−k)
≤ C2−kε′ Xj≥k−2
u(y) − cp∗(ε) dµ(y)(cid:19)1/p∗(ε)
2−j(s−ε′)(cid:18)ZB(x,2−k+1)
gp
j dµ(cid:19)1/p
,
where p∗(ε) = Qp/(Q − εp).
If p ≥ Q/(Q + ε), then (4.1) implies that
(4.2)
ZB(x,2−k)
u − uB(x,2−k) dµ ≤ C2−kε′ Xj≥k−2
2−j(s−ε′)(cid:18)ZB(x,2−k+1)
gp
j dµ(cid:19)1/p
.
We are now ready to state and prove our main results. Theorem 4.2 below
gives a formula for an (s + α)-Haj lasz gradient of M∗
α in terms of an s-Haj lasz
gradient of u. This easily implies the desired boundedness results for M∗
α in
homogeneous Haj lasz -- Besov and Haj lasz -- Triebel -- Lizorkin spaces. For related
results concerning Riesz potentials in the metric setting, see [31]
Theorem 4.2. Assume that M∗
α u 6≡ ∞ and that (gk) ∈ Ds(u). Let 0 <
s + α < 1, 0 < δ < 1 − s − α, 0 < ε < ε′ < s and t ≥ Q/(Q + ε). Then there
is a constant C > 0, indepent of u and (gk), such that (C gk), where
(4.3)
gk =
k
Xj=−∞
∞
2(j−k)δ(cid:0)M gt
j(cid:1)1/t +
Xj=k−7
2(k−j)(s−ε′)(cid:0)M gt
j(cid:1)1/t ,
is a fractional (s + α)-Haj lasz gradient of M∗
α u.
10
TONI HEIKKINEN AND HELI TUOMINEN
Proof. Let k ∈ Z and let x, y ∈ X such that 2−k−1 ≤ d(x, y) < 2−k. We will
show that
uα
r (x) − uα
r (y) ≤ Cd(x, y)s+α(gk(x) + gk(y)),
where C is independent of r and k.
Assume first that d(x, y) > r. Then
ur(x) − ur(y) ≤ ur(x) − uB(x,2−k+4) + ur(y) − uB(x,2−k+4)
≤ Xi∈Ix
+Xi∈Iy
ϕi(x)uB(xi,3r) − uB(x,2−k+4)
ϕi(y)uB(xi,3r) − uB(x,2−k+4),
where Ix is a set of indices i for which x belongs to B(xi, 6r) and Iy the
corresponding set for y. Let m ∈ Z be such that 2−m−1 < 9r ≤ 2−m. Since
r < d(x, y) < 2−k, it follows that m ≥ k − 4. If i ∈ Ix, we obtain
uB(xi,3r) − uB(x,2−k+4) ≤ uB(xi,3r) − uB(x,2−m) +
m−1
Xl=k−4
uB(x,2−l) − uB(x,2−l−1)
≤ C
m
Xl=k−4
ZB(x,2−l)
u − uB(x,2−l) dµ
and hence Poincar´e inequality (4.2) implies that
uB(xi,3r) − uB(x,2−k+4) ≤ C
2−lε′
∞
Xl=k−4
∞
∞
Xj=l−2
j+2
2−lε′
= C
Xj=k−6
j(x)(cid:1)1/t
2−j(s−ε′)(cid:0)M gt
j(x)(cid:1)1/t
Xl=k−4
j(x)(cid:1)1/t
2−j(s−ε′)(cid:0)M gt
j(x)(cid:1)1/t
2(k−j)(s−ε′)(cid:0)M gt
2−j(s−ε′)(cid:0)M gt
Xj=k−6
Xj=k−6
≤ C2−ksgk(x).
≤ C2−kε′
= C2−ks
∞
∞
Similarly, if i ∈ Iy, then
uB(xi,3r) − uB(x,2−k+4) ≤ uB(xi,3r) − uB(y,2−m) +
m−1
Xl=k−4
uB(y,2−l) − uB(y,2−l−1)
+ uB(y,2−k+5) − uB(x,2−k+4)
≤ C
m
Xl=k−5
ZB(y,2−l)
u − uB(y,2−l) dµ,
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
11
which implies that
It follows that
uB(xi,3r) − uB(x,2−k+4) ≤ C2−ksgk(y).
uα
r (x) − uα
r (y) ≤ Crα2−ks(gk(x) + gk(y))
≤ Cd(x, y)s+α(gk(x) + gk(y)).
Suppose then that d(x, y) ≤ r. Let Ixy be a set of indices i for which x or y
belongs to B(xi, 6r). Let l be such that 2−l−1 < 10r ≤ 2−l. Using the doubling
condition, the properties of the functions ϕi, the fact that there are bounded
number of indices in Ixy and Poincar´e inequality (4.2), we have that
(4.4)
uα
r (x) − uα
r (y) ≤ rα
∞
Xi=1
ϕi(x) − ϕi(y))uB(xi,3r) − uB(x,2−l)
≤ Cd(x, y)rα−12−lε′
∞
Xj=l−2
2−j(s−ε′)(cid:0)M gt
j(x)(cid:1)1/t .
Using the assumptions 0 < δ < 1 − α − s, r ≥ d(x, y) and d(x, y) < 2−k, we
have that
d(x, y)rα−12−lε′
≤ Cd(x, y)rs+α+δ−1 2l(s−ε′+δ) ≤ Cd(x, y)s+α+δ 2l(s−ε′+δ)
≤ Cd(x, y)α+s2(l−k)δ+l(s−ε′).
This together with (4.4) implies that
uα
r (x) − uα
r (y) ≤ Cd(x, y)s+α
∞
Xj=l−2
2(l−k)δ+(l−j)(s−ε′)(cid:0)M gt
j(x)(cid:1)1/t .
By splitting the sum in two parts and using the estimates l ≤ j + 2 and l ≤ k,
we obtain
∞
=
k−1
Xj=l−2
Xj=l−2
≤ C(cid:16)
2(l−k)δ+(l−j)(s−ε′)(cid:0)M gt
2(l−k)δ+(l−j)(s−ε′)(cid:0)M gt
Xj=−∞
2(j−k)δ(cid:0)M gt
j(x)(cid:1)1/t
j(x)(cid:1)1/t +
Xj=k
j(x)(cid:1)1/t +
k−1
∞
∞
Xj=k
2(l−k)δ+(l−j)(s−ε′)(cid:0)M gt
j(x)(cid:1)1/t(cid:17),
2(k−j)(s−ε′)(cid:0)M gt
j(x)(cid:1)1/t
which implies the claim for uα
proof of Theorem 3.2.
r . The claim for M∗
α u follows similarly as in the
(cid:3)
Theorem 4.3. Let 0 < s + α < 1 and Q/(Q + s) < p, q < ∞. Then there
exists a constant C > 0 such that
k M∗
α uk M s+α
p,q (X) ≤ Ckuk M s
p,q(X)
for all u ∈ M s
p,q(X) with M∗
α u 6≡ ∞.
12
TONI HEIKKINEN AND HELI TUOMINEN
2(1 − (s + α)), ε = 1
Proof. Let δ = 1
2(ε + s), where
r = min{p, q}, and let t = Q/(Q + ε). Then 0 < ε < ε′ < s and Q/(Q + s) <
t < min{p, q}. By Theorem 4.2, (C gk) defined by (4.3) is a fractional (s + α)-
Haj lasz gradient of M∗
2 max{s, s + Q−Qr
r }, ε′ = 1
It suffices to show that (gk) ∈ Lp(X, lq). We estimate the Lp(X, lq) norm of
α u.
k
Xj=−∞
(cid:16)
2(j−k)δ(cid:0)M gt
j(cid:1)1/t(cid:17)k∈Z
,
the other part can be estimated similarly. If q ≥ 1, we have, by the Holder
inequality, that
Xk∈Z(cid:16)
k
Xj=−∞
2(j−k)δ(cid:0) M gt
j(cid:1)1/t(cid:17)q
k
j(cid:1)q/t
∞
Xj=−∞
≤ CXk∈Z
≤ CXj∈Z (cid:0)M gt
≤ CXj∈Z (cid:0)M gt
2(j−k)δ(cid:0)M gt
j(cid:1)q/t
j(cid:1)q/t
Xk=j
.
2(j−k)δ
If q < 1, we obtain the same estimate by using the elementary inequality
(Pj aj)q ≤ Pj aq
j for aj ≥ 0.
By the Fefferman -- Stein vector valued maximal function theorem from [5]
(for a metric space version, see for example [27] or [8]), we obtain now the
desired estimate
k
Xj=−∞
(cid:13)(cid:13)(cid:13)(cid:16)
2(j−k)δ(cid:0)M gt
j(cid:1)1/t(cid:17)k∈Z(cid:13)(cid:13)(cid:13)Lp(X, lq)
1/t
Lp/t(X, lq/t)
≤ C(cid:13)(cid:13)(cid:0)M gt
k(cid:1)k∈Z(cid:13)(cid:13)
≤ Ck(gt
k)k∈Zk1/t
Lp/t(X, lq/t)
= Ck(gk)k∈ZkLp(X, lq).
(cid:3)
Theorem 4.4. Let 0 < s + α < 1, Q/(Q + s) < p < ∞ and 0 < q < ∞. Then
there exists a constant C > 0 such that
k M∗
α uk N s+α
p,q (X) ≤ Ckuk N s
p,q(X)
for all u ∈ N s
p,q(X) with M∗
α u 6≡ ∞.
2(1 − (s + α)), ε = 1
Proof. Let δ = 1
2(ε + s), and let
t = Q/(Q + ε). Then 0 < ε < ε′ < s and Q/(Q + s) < t < p. Then (C gk)
defined by (4.3) is a fractional (s + α)-Haj lasz gradient of M∗
α u by Theorem
4.2.
2 max{s, s + Q−Qp
p }, ε′ = 1
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
13
It suffices to show that k(gk)klq(Lp(X)) ≤ Ck(gk)klq(Lp(X)). By the Hardy --
Littlewood maximal theorem,
k
Xj=−∞
(cid:13)(cid:13)(cid:13)
2(j−k)δ(cid:0)M gt
j(cid:1)1/t(cid:13)(cid:13)(cid:13)Lp(X)
≤
≤
k
Xj=−∞
Xj=−∞
k
2(j−k)δ(cid:13)(cid:13)(cid:0)M gt
j(cid:1)1/t(cid:13)(cid:13)Lp(X)
2(j−k)δkgjkLp(X).
If q ≥ 1, we have by the Holder inequality,
k
Xj=−∞
kgjkq
k
Xj=−∞
Xk∈Z(cid:16)
2(j−k)δkgjkLp(X)(cid:17)q
≤ CXk∈Z
≤ CXj∈Z
≤ CXj∈Z
If q < 1, we use the inequality (Pj aj)q ≤ Pj aq
ity. The second part of (gk) can be estimated similarly.
2(j−k)δkgjkq
Lp(X)
Lp(X)
∞
Xk=j
2(j−k)δ
kgjkq
Lp(X).
j instead of the Holder inequal-
(cid:3)
Theorems 4.3, 4.4 and the Hardy -- Littlewood maximal theorem imply the
following results for the discrete maximal operator.
Theorem 4.5. Let 0 < s < 1.
a) If Q/(Q + s) < p, q < ∞, then there exist a constant C > 0 such that
k M∗ uk M s
p,q(X) ≤ Ckuk M s
p,q(X),
whenever u ∈ M s
p,q(X) and M∗ u 6≡ ∞.
b) If 1 < p, q < ∞, then there exist a constant C > 0 such that
k M∗ ukM s
p,q(X) ≤ CkukM s
p,q(X),
for all u ∈ M s
p,q(X).
Theorem 4.6. Let 0 < s < 1.
a) If Q/(Q + s) < p < ∞ and 0 < q < ∞, there exist a constant C > 0
such that
k M∗ uk N s
p,q(X) ≤ Ckuk N s
p,q(X)
for all u ∈ N s
p,q(X) with M∗ u 6≡ ∞.
b) If 1 < p < ∞ and 0 < q < ∞, there exist a constant C > 0 such that
k M∗ ukN s
p,q(X) ≤ CkukN s
p,q(X)
for all u ∈ N s
p,q(X).
14
TONI HEIKKINEN AND HELI TUOMINEN
References
[1] D. Aalto and J. Kinnunen, The discrete maximal operator in metric spaces, J. Anal.
Math. 111 (2010), 369 -- 390.
[2] J. M. Aldaz and J. P´erez L´azaro, Functions of bounded variation, the derivative of
the one dimensional maximal function, and applications to inequalities, Trans. Amer.
Math. Soc. 359 (2007), 2443-2461.
[3] S.M. Buckley, Is the maximal function of a Lipschitz function continuous?, Ann. Acad.
Sci. Fenn. Math. 24 (1999), 519 -- 528.
[4] D. Edmunds, V. Kokilashvili, A. Meskhi, Bounded and Compact Integral Operators,
Mathematics and its Applications, vol. 543, Kluwer Academic Publishers, Dordrecht,
Boston, London, 2002.
[5] C. Fefferman and E. M. Stein, Some maximal inequalities, Amer. J. Math. 93 (1971),
107 -- 115.
[6] I. Genebashvili, A. Gogatishvili, V. Kokilashvili and M. Krbec, Weight Theory for Inte-
gral Transforms on Spaces of Homogeneous Type, Addison Wesley Longman Limited,
1998.
[7] A. Gogatishvili, P. Koskela and Y. Zhou, Characterizations of Besov and Triebel --
Lizorkin Spaces on Metric Measure Spaces, to appear in Forum math.
[8] L. Grafakos, L. Liu and D. Yang, Vector-valued singular integrals and maximal func-
tions on spaces of homogeneous type, Math. Scand. 104 (2009), 296 -- 310.
[9] P. Haj lasz, Sobolev spaces on an arbitrary metric space, Potential Anal. 5 (1996),
403 -- 415.
[10] P. Hajlasz, Sobolev spaces on metric-measure spaces, (Heat kernels and analysis on
manifolds, graphs, and metric spaces (Paris, 2002)), 173 -- 218, Contemp. Math., 338,
Amer. Math. Soc., Providence, RI, 2003.
[11] P. Haj lasz and J. Maly, On approximate differentiability of the maximal function,
Proc. of AMS., 138 (2010), no. 1, 165 -- 174.
[12] P. Haj lasz and J. Onninen, On boundedness of maximal functions in Sobolev spaces,
Ann. Acad. Sci. Fenn. Math. 29 (2004), 167 -- 176.
[13] T. Heikkinen, J. Kinnunen, J. Nuutinen and H. Tuominen, Mapping properties of the
discrete fractional maximal operator in metric measure spaces, to appear in Kyoto J.
Math.
[14] T. Heikkinen, P. Koskela and H. Tuominen, Sobolev-type spaces from generalized
Poincar´e inequalities, Studia Math. 181 (2007), 1 -- 16.
[15] T. Heikkinen, J. Lehrback, J. Nuutinen and H. Tuominen, Regularity of the fractional
maximal function in metric measure spaces, preprint.
[16] J. Hu, A note on Haj lasz-Sobolev spaces on fractals, J. Math. Anal. Appl. 280 (2003),
91-101.
[17] J. Kinnunen, The Hardy-Littlewood maximal function of a Sobolev-function, Israel
J.Math. 100 (1997), 117-124.
[18] J. Kinnunen and V. Latvala, Lebesgue points for Sobolev functions on metric spaces,
Rev. Mat. Iberoamericana 18 (2002), no. 3, 685 -- 700.
[19] J. Kinnunen and P. Lindqvist, The derivative of the maximal function, J. Reine Angew.
Math. 503 (1998), 161-167.
[20] J. Kinnunen and E. Saksman, Regularity of the fractional maximal function, Bull.
London Math. Soc. 35 (2003), no. 4, 529 -- 535.
[21] J. Kinnunen and H. Tuominen, Pointwise behaviour of M 1,1 Sobolev functions, Math.
Z. 257 (2007), no. 3, 613 -- 630.
[22] S. Korry, A class of bounded operators on Sobolev spaces, Arch. Math. (Basel) 82
(2004), no. 1, 40 -- 50.
THE FRACTIONAL MAXIMAL OPERATOR ON BESOV AND T -- L SPACES
15
[23] S. Korry, Boundedness of Hardy-Littlewood maximal operator in the framework of
Lizorkin-Triebel spaces, Rev. Mat. Complut. 15 (2002), no. 2, 401 -- 416.
[24] P. Koskela, D. Yang and Y. Zhou, Pointwise Characterizations of Besov and Triebel-
Lizorkin Spaces and Quasiconformal Mappings, Adv. Math. 226 (2011), no. 4, 3579 --
3621.
[25] H. Luiro, Continuity of the Hardy-Littlewood maximal operator in Sobolev spaces,
Proc. of AMS., 135 (2007), no. 1, 243 -- 251.
[26] H. Luiro, On the regularity of the Hardy-Littlewood maximal operator on subdomains
of Rn, Proc. Edinb. Math. Soc. (2) 53 (2010), no. 1, 211 -- 237.
[27] Y. Sawano, Sharp estimates of the modified Hardy-Littlewood maximal operator on
the nonhomogeneous space via covering lemmas, Hokkaido Math. J. 34 (2005), no. 2,
435 -- 458.
[28] H. Tanaka, A remark on the derivative of the one-dimensional Hardy-Littlewood max-
imal function, Bull. Austral. Math. Soc. 65 (2002), 253 -- 258.
[29] H. Triebel, Theory of Function Spaces, Birkhauser Verlag, Basel, 1983.
[30] D. Yang, New characterizations of Haj lasz-Sobolev spaces on metric spaces, Sci. China
Ser. A 46 (2003), 675-689.
[31] D. Yang, Riesz Potentials in Besov and Triebel -- Lizorkin Spaces over Spaces of Homo-
geneous Type, Potential Anal. 19 (2003), 193 -- 210.
T.H., Department of Mathematics, P.O. Box 11100, FI-00076 Aalto Uni-
versity, Finland
[email protected]
H.T., Department of Mathematics and Statistics, P.O. Box 35, FI-40014
University of Jyvaskyla, Finland
[email protected]
|
1709.04968 | 1 | 1709 | 2017-09-14T20:26:59 | Toeplitz Quantization and Convexity | [
"math.FA",
"math.SG"
] | Let $T^m_f $ be the Toeplitz quantization of a real $ C^{\infty}$ function defined on the sphere $ \mathbb{CP}(1)$. $T^m_f $ is therefore a Hermitian matrix with spectrum $\lambda^m= (\lambda_0^m,\ldots,\lambda_m^m)$. Schur's theorem says that the diagonal of a Hermitian matrix $A$ that has the same spectrum of $ T^m_f $ lies inside a finite dimensional convex set whose extreme points are $\{( \lambda_{\sigma(0)}^m,\ldots,\lambda_{\sigma(m)}^m)\}$, where $\sigma$ is any permutation of $(m+1)$ elements. In this paper, we prove that these convex sets "converge" to a huge convex set in $L^2([0,1])$ whose extreme points are $ f^*\circ \phi$, where $ f^*$ is the decreasing rearrangement of $ f$ and $ \phi $ ranges over the set of measure preserving transformations of the unit interval $ [0,1]$. | math.FA | math |
Toeplitz Quantization and Convexity
O. El Hadrami, Mohamed Lemine,
Department of Mathematics
King Khalid Universtiy, Abha, September 14, 2017
1
Abstract
Let T m
0 , . . . , λm
f be the Toeplitz quantization of a real C∞ function defined
on the sphere CP(1). T m
is therefore a Hermitian matrix with spec-
f
trum λm = (λm
m). Schur's theorem says that the diagonal of a
Hermitian matrix A that has the same spectrum of T m
lies inside a fi-
f
nite dimensional convex set whose extreme points are {(λm
σ(m))},
where σ is any permutation of (m + 1) elements. In this paper, we prove
that these convex sets "converge" to a huge convex set in L2([0, 1]) whose
extreme points are f ∗ ◦ φ, where f ∗ is the decreasing rearrangement of
f and φ ranges over the set of measure preserving transformations of the
unit interval [0, 1].
σ(0), . . . , λm
1 Introduction and background
In their papers [3, 4], the authors have described similarities between the infinite
and finite dimensional Lie groups. They have strengthen the idea that the set
SDif f (CP(1)) of area preserving diffeomorphisms of the Riemann sphere is an
infinite dimensional analog of SU (n) by proving an infinite version of the SU (n)
Schur-Horn convexity theorem .
In the present paper, we want to show that the convex sets in the two versions
(finite and infinite) of Schur-Horn convexity theorem are related. In order to
do that, we will show first that each permutation of (m + 1) letters determine a
measure preserving transformation of the interval [0, 1]. Secondly we use the fact
that the eigenvalues of the Toeplitz quantization of f determine the decreasing
rearrangement of f .
But since the two convex sets are defined by inequalities arising from the theory
of majorization in Rm+1 developed in [11, 12] and its generalization to L1([0, 1])
by J.Ryff [13, 14, 15], we start by summarizing briefly here the main points.
Majorization is a partial ordering in Rm+1 defined as it follows:
12010 Mathematics Subject Classification. 70G65, 37A15.
Key words and phrases. Toeplitz quantization, measure preserving transformations.
1
For x ∈ Rm+1, let x∗ denote the vector obtained by rearranging the components
of x in non-increasing order. We say that x majorizes y, written y ≺ x, if
y∗
0 + y∗
1 + · · · + y∗
k ≤ x∗
0 + x∗
1 + · · · + x∗
k,
0 ≤ k ≤ m − 1
k=mXk=0
y∗
k =
k=mXk=0
x∗
k
Now we can state Schur' theorem for hermitian matrices.
Theorem 1.1 Let λm = (λ0, · · · , λm) ∈ Rm+1 be the eigenvalues of a hermi-
tian matrix A . Let diag(A) = (a00, · · · , amm) be the diagonal of A then λm
majorizes diag(A).
i.e
diag(A) ≺ λm
(1.1)
Before going on, let us first fix some notations:
1. Let x = (x0, · · · , xm) ∈ Rm+1.
Pm x is the orbit of x under the symmetric group of (m + 1) letters, i.e
the collection of points (xσ(0), · · · , xσ(m)), where σ ranges over all (m + 1)!
permutations.
2. For C ⊂ E where E is a vector space over R, co(C) denote the convex hull
of C.
Majorization and convexity are closely related as it is shown by the following
theorem.
Theorem 1.2 Let x ∈ Rm+1
• ( Rado's theorem)
• y ≺ x if and only if Pi=m
whose domain contains all the numbers xi, yi, 0 ≤ i ≤ m.
{y ∈ Rm+1, y ≺ x} = co(Xm
i=0 f (yi) ≤ Pi=m
x)
i=0 f (xi) for any convex function
Schur-Horn' theorem can be therefore restated in the following terms :
Theorem 1.3 (Schur-Horn's theorem) Let λm = (λ0, · · · , λm) ∈ Rm+1.
Let Hλm be the set of all hermitian matrices whose spectrum is λm.
Let pm : Hλm −→ Rm+1 be the map that picks the diagonal of a matrix.
Then the image of the map pm is the convex set co(Pm λm).
The convex set co(Pm λm) plays a very important role in symplectic geometry:
It is the image of a moment map [1, 8].
The concepts of majorization is also extended to integrable functions in the
2
following sense.
Let (X, µ) be a finite measurable space. For f measurable function on X, the
distribution function of f is the function Ff defined by
Ff (t) = µ({ω, f (ω) < t}).
Let df (t) = µ(X) − Ff (t).
Definition 1.1 The decreasing rearrangement of f is the function f ∗ defined
by:
f ∗(s) = inf{t > 0, df (t) < s}.
For f, g ∈ L1((X, µ)), let f ∗ and g∗ be their decreasing rearrangement
We say that f majorizes g ( written g ≺ f ) if
0
Z s
Z 1
0
g∗(z) dz ≤Z s
g∗(z) dz =Z 1
0
0
f ∗(z) dz,
0 ≤ s < 1
f ∗(z) dz
To stay conform with the notation of [4], let X = CP(1) be the Riemann sphere
, µ the measure defined by the the Fubini-Study symplectic form which in the
local coordinate [1, w] is given by
Ω =
i
(1 + ww)2
dw ∧ dw.
Set w = r exp(iθ) and introduce the real variable z ∈ [0, 1[ by z = r2\(1 + r2).
The symplectic form Ω becomes Ω = 2dz ∧ dθ. The infinite version of Schur
theorem is
Theorem 1.4 (Schur-type Theorems,[3]) Let L2(CP(1)) be the set of square
integrable functions on the sphere. Let P : L2(CP(1)) −→ L2[0, 1] be the projec-
tion P (f ) =
f (z, θ)dθ.
1
2π Z 2π
0
Then f ∗ majorizes P (f ).
i.e
P (f ) ≺ f ∗
We have also the equivalent of Rado's theorem in ∈ L2([0, 1]).
Theorem 1.5 Let f ∈ L2([0, 1]). The set Ω(f ) = {g ∈ L2([0, 1]), g ≺ f } is
weakly compact and convex. Its set of extreme points is
{f ∗ ◦ φ φ is a measure preserving transformation of
[0, 1]}.
The paper is organized in three sections: In §2, we have reviewed the topology
of the set of measure preserving transformations of the unit interval[0, 1] and
showed that the groups,Pm of permutations of (m+1) letters, can be identified
3
with a dense subset of the set of all invertible measure preserving transforma-
tions of [0, 1] endowed with strong operator topology.
In §3 we use Toeplitz quantization to show that the weak closure of the topo-
logical lim sup of the sets co (Pm ·λm) is the set
Ω(f ∗) = co ({f ∗ ◦ φ, φ measure preserving transformation of [0, 1]}).
2 measure preserving transformations of [0,1]
The Lebesgue measure on the unit interval I = [0, 1] will be always denoted by
· .
A map φ from [0, 1] to itself is a measure preserving transformation if
φ−1(A) = A,
for Borel set A
The set of all (non necessary invertible ) measure preserving transformation of
the unit interval will be denoted Smeas(I). The invertible ones will be denoted
by Imeas(I).
Each S ∈ Smeas(I) determine a bounded linear operator PS on L2([0, 1]) by
PS(f ) = f ◦ S. In this way, Smeas(I) can be identified to a subset of the set of
bounded linear operators of L2[0, 1]) and the strong operator topology induces
a topology on Smeas(I).
Evidently, a sequence Sn converges to S in the strong operator topology if for
every function f, f ◦ Sn converges to f ◦ S in L2([0, 1]).
σ(k) by ordinary translation.
k to the
To state our first main result, we need to define dyadic permutations.
Let I m
k = [k\(m + 1), (k + 1)\(m + 1)) , k = 0, 1, · · · , m; m = 0, 1, · · · .
invertible measure preserving transformation that sends the interval I m
interval I m
Let Pm be the group of permutations of (m + 1) letters. For σ ∈Pm, bσ is the
We call bσ a permutation of rank m. In this way, we can identified the group of
permutations Pm with a subgroup of Imeas(I).
Halmos ([10]) shows that the set of all permutations of different rank is dense
in Imeas(I) for the strong operator topology.
Also Brown, in [6] has proved that Smeas(I) is the closure of Imeas(I) for the
strong operator topology.
We can summarize these results in the following theorem.
Theorem 2.1 Let Pm be the symmetric group of (m + 1) letters. There exists
a one to one group homomorphism Ψm : Pm −→ Imeas(I) that sends σ to bσ
and if we identify Pm with Ψm(Pm), then SmPm is a dense set in Smeas(I)
for the strong operator topology.
4
3 Toeplitz Quantization and Convexity
3.1 Toeplitz Quantization
Consider the Riemann sphere CP(1) with the Fubini-Study symplectic form in
the local coordinate [1, w]
Ω =
i
(1 + ww)2 dw ∧ dw
and the standard hyperplane bundle L. The tensor power L⊗m has (m + 1) lin-
early independent sections which in the local coordinate w are just 1, w, ..., wm.
The bundle L⊗m comes equipped with the Hermitian metric
hs1, s2i(w) =
1
(1 + w2)m
s1(w)s2(w).
2 be the space of square-integrable sections of L⊗m and Γm
Now let Γm
hol the space
of holomorphic sections ( the span of 1, w, ..., wm).The orthogonal projection
Γm
2 −→ Γm
:
Γm
hol → Γm
hol is denoted by P m.The Toeplitz quantization of f is the map T m
hol defined by
f
T m
f = P m ◦ Mf ◦ P m
where Mf is multiplication by f . We refer the interested reader to [5] for a
detailed exposition on Toeplitz quantization.
The crucial result is the following theorem
Theorem 3.1 (Distribution of the Eigenvalues of Toeplitz Quantization )
Let λm = (λm
order and let Λm(s) be the real step function defined on the interval [0, 1[ by
f arranged in non-increasing
m) be the eigenvalues of T m
1 , · · · , λm
0 , λm
Λm(cid:18)(cid:20)
k
m + 1
,
k + 1
m + 1(cid:20)(cid:19) = λm
k ,
0 ≤ k ≤ m.
(3.1)
Then the sequence (Λm(s))m converges point-wise almost everywhere to the de-
creasing rearrangement f ∗(s) of the function f .
The proof of this Theorem is based on the the following theorem
Theorem 3.2 (Szego-type Theorem,[9] p: 248) Given a smooth real val-
f be Toeplitz quantization of f and let µm be its
ued function f on CP(1), let T m
spectral measure . Then
tends weakly to a limiting measure as m tends
µm
m + 1
to infinity, this limiting measure being
µ(φ) =
1
2πZCP(1)
φ(f (x))dΩ,
for φ ∈ C(R).
0 , λm
1 , · · · , λm
m) are the eigenvalues of T m
f ordered in non-increasing order,
i.e if (λm
then
lim
m−→+∞
k=mXk=0
φ(λk)
m + 1
=
1
2πZCP(1)
φ(f (x))dΩ.
(3.2)
5
If we use the step function Λm defined be (3.1) then (3.2) can be written
m−→+∞Z 1
lim
0
φ(Λm) =
1
2πZCP(1)
φ(f (x))dΩ.
(3.3)
But since f and f ∗ are equi-measurable, we have
1
2π ZCP(1)
(φ ◦ f )dΩ =Z 1
0
φ ◦ f ∗(t)dt
and (3.3) becomes
m7→+∞Z 1
lim
0
φ(Λm)(t) dt =Z 1
0
φ(f ∗)(t) dt.
(3.4)
where Λm is defined by (3.1).
Relation ( 3.4 ) is equivalent to: the sequence of step functions Λm converges
in distribution to the real function f ∗. (One can see [7] for more details on
convergence in distribution.)
In general convergence in distribution does not imply convergence point-wise.
Nevertheless, there exists another sequence gn with the same distribution as fn
that converges point-wise to a function g, that has the same distribution of f .
That is the content of Skorokhod's Theorem.
6
Theorem 3.3 (Skorokhod) Let (X, Σ, µ) be a finite measure space,and f, fn :
X → R be a sequence of measurable functions such that fn converge in distribu-
tion to f . Then on the Lebesgue measure space (I, B, ·), where I = (0, µ(X), B
is the Borel σ-algebra of I, and · is the Lebesgue measure, there exists measur-
able functions gn, g : I → R such that gn(t) → g(t) a.e[·], and
µ({ω : fn(ω) < x}) = {t : gn(t) < x}
µ({ω : f (ω) < x}) = {t : g(t) < x}
x ∈ R, n ≥ 1.
Let Fn and F be the distribution functions of fn and f . We can take gn and g
to be just the generalized inverse of Fn and F :
gn(t) = inf{x : Fn(x) > t},
g(t) = inf{x : F (x) > t}, 0 < t < 1.
It easily seen then that
gn(1 − t) = g∗
n(t) = f ∗
n(t),
g(1 − t) = g∗(t) = f ∗(t).
( [11] [p141-144] .)
Applying Skorokhod's Theorem to the the sequence Λm, we deduce that the
generalized inverses of the functions Λm converges point-wise almost everywhere
to the generalized inverse of the function f .
Consequently , the sequence of the decreasing rearrangements of Λm converges
to the decreasing rearrangement f ∗ of f .
lim
m7−→+∞
(Λm)∗ (s) = f ∗(s).
But since Λm(s) is a decreasing function, (Λm)∗ (s) = Λm(s), and therefore we
have
and that ends the proof of the theorem.
lim
m7−→+∞
Λm(s) = f ∗(s)
3.2 Convexity
Let pm : Rm+1 → L2([0, 1]) be the map that associates to the point (a0, a1, · · · , am)
k
the step function pm(a0, a1, · · · , am) = Pk=m
val (cid:20)
pm(co(Pm ·λm)).
m + 1(cid:20) and χIm
m + 1
k + 1
,
Now we are ready to state the main result about convexity.
k=0 akχIm
k , where I m
k
is the inter-
is the characteristic function of I m
k . Set Em =
k
Theorem 3.4 Let f ∈ C∞(CP(1)). Let f ∗ be the decreasing rearrangement of
f . Let T m
m) be the
f . Let Ω(f ∗) = {g ∈ L2([0, 1], g ≺ f ∗}. Then the closed convex
eigenvalues of T m
hull of the set Smeas(I) · f ∗ = {f ∗ ◦ φ φ ∈ Smeas(I)} is the weak closure of
f be the Toeplitz quantization of f and let λm = (λm
0 , · · · , λm
the topological lim sup of the convex sets Em = pm(co(Pm ·λm)).
7
We recall the definition of the closed limit.
Definition 3.1 ( [2], p. 114) Let {Em} be a sequence of subsets of a topolog-
ical space X. Then, a point x in X belongs to the topological lim sup of Em,
denoted LsEm, if for every neighborhood V of x there are infinitely many m with
V ∩ Em 6= ∅.
Clearly, LsEm is a closed sets and moreover
LsEm =
m=∞\m=1
m=∞[k=m
Ek.
In our case, X = L2([0, 1]) and Em = pm(co(Pm ·λm)) .
Since L2([0, 1], k·k2) is a normed vector space, every point of L2([0, 1] has a
countable basis of neighborhood and LsEm can be characterized in terms of
sequences:
g ∈ LsEm if and only if there exists a subsequence (gmk )k such that gmk ∈ Emk
and gmk converges to g ∈ L2([0, 1]).
Now we claim
(A) Smeas(I) · f ∗ ⊂ LsEm.
(B) Ω(f ∗) ⊂ LsEm
weak
⊂ Ω(f ∗).
Proof of( A):
Let f ∗ ◦ φ ∈ Smeas(I) · f ∗.
The set of dyadic permutations is a countable set and from theorem (2.1) is
dense in Smeas(I) for the strong operator topology; therefore there exists a
sequence (σn)n of permutations that converges to φ , i.e
∀g ∈ L2([0, 1]), ∀ǫ > 0, ∃N0, ∀n ≥ N0, kg ◦ σn − g ◦ φk2 <
ǫ
2
.
In particular for g = Λm, we have,
∀m, ∀ǫ > 0, ∃N0, ∀n ≥ N0, kΛm ◦ σn − Λm ◦ φk2 <
ǫ
2
.
(3.5)
Now since the sequence Λm converges almost everywhere to f ∗ and ∀m, ∀x ∈
[0, 1], Λm(x) ≤ kf ∗k∞, , it converges also to f ∗ in L2([0, 1]), ie
∀ǫ > 0, ∃M0, ∀m ≥ M0, kΛm ◦ φ − f ∗ ◦ φk2 <
ǫ
2
.
(3.6)
We conclude then from (3.5) and (3.6) that
∀ǫ > 0, ∃M0, ∃N0, ∀n ≥ N0, ∀m ≥ M0
kΛm ◦ σn − f ∗ ◦ φk2 ≤ kΛm ◦ σn − Λm ◦ φk2 + kΛm ◦ φ − f ∗ ◦ φk2 ≤ ǫ.
(3.7)
Now if σn is of order kn, and if we let m = kn in (3.7) we get
8
∀ǫ > 0, ∃M0, ∃N0 for every n ≥ N0 such that kn ≥ M0, we have
kΛkn ◦ σn − f ∗ ◦ φk2 ≤ ǫ.
Since Λkn ◦ σn ∈ Ekn , we conclude therefore that f ∗ ◦ φ ∈ LsEm .
Proof of (B):
It is shown in [ [3],p 523] that Ω(f ∗) = {g ∈ L2([0, 1]), g ≺ f ∗} is weakly compact
and convex. Its extreme points are precisely the elements f ∗ ◦ φ, φ ∈ Smeas(I).
It is also indicated in [ [15], p1030] that the set of extreme points is dense in
Ω(f ∗) for the weak topology. But we have just seen in part (A) of our claim
that Smeas(I) · f ∗ ⊂ LsEm . Taking the closure in the weak topology we get
Ω(f ∗) = Smeas(I) · f ∗
weak
⊂ LsEm
weak
.
weak
It remains to show that LsEm
Let g ∈ LsEm. Then by definition, there exists a subsequence (gmk )k such that
gmk ∈ Emk , and (gmk )k converges in L2([0, 1]) to g.
Now by Rado's theorem (1.2) , we have
⊂ Ω(f ∗).
gmk ∈ Emk ⇐⇒ gmk ≺ Λmk
But the sequence (gmk )k being convergent in L2([0, 1])), we can extract a sub-
l(cid:17)l
sequent (cid:16)gmk
that converges pointwise to g a.e. We have then
gmk
l
≺ Λmk
l
.
And taking limit (simple convergence) of both sides, we get g ≺ f ∗ i.e LsEm ⊂
Ω(f ∗). Taking the weak closure of both sets we conclude therefore that
Ω(f ∗) = LsEm
weak
.
Our next goal is to compare the co-adjoint orbits of SU (m + 1) and the co-
adjoint orbits of SDiff(CP(1), the group of area preserving diffeomorphisms of
the sphere.
Acknowledgment
I would like thank Professor H. Flaschka for valuable discussions.
References
[1] M.F.Atiyah, Convexity and commuting Hamiltonians, Bull. London Math.
Soc 14, 1-15, (1982)
[2] C. D. Aliprantis, K. C. Border, Infinite Dimensional Analysis, a Hitch-
hiker's Guide, 3rd edition, Springer (2005).
9
[3] A.Bloch, H.Flaschka, T.Ratiu, A Shur-Horn-Kostant Convexity Theorem
for the Diffeomorphism Group of the Annulus, Inv.Math. 113, 511-529,
(1993).
[4] A. Bloch, M. El Hadrami, H. Flaschka, T. Ratiu, Maximal Tori Of Some
Symplectomorphism Groups And Applications To Convexity, Proceedings
of Ascona Meeting, June 1996. (D. Sternheimer, J. Rawnsley, S. Gutt, eds.),
Mathematical Physics Studies 20, Kluwer Academic Publishers, 201-222,
(1997).
[5] M. Bordemann, E. Meinrenken and M. Schlichenmaier, Toeplitz Quantiza-
tion Of Kahler Manifolds And gl(N ), N → +∞ Limits., Commun. Math.
Phys. 165 , 281-296,(1994).
[6] J. R. Brown, Approximation
theorems
for Markov Operators,
Pac.J.Maths.16, 13-23, (1966).
[7] Richard Durrett Probability:Theory And Examples, Wadsworth &
brooks/cole Advanced Books & Software, 1991.
[8] V.Guillemin and S. Sternberg, Convexity properties of the momentum map-
ping I, Invent. Math, 67, 491-513,(1982).
[9] Victor Guillemin, Some Classical Theorems In Spectral Theory Revisited,
Seminar on Singulariries of Solutions of Linear Partial Differential Equa-
tions (L. H ormander ed) Annals of Mathematics Studies, 91, Princeton
University Press and University of Tokyo Press, [1979], 219- 259.
[10] P. Halmos Approximation Theorems for Measure Preserving Transforma-
tions, Trans.AMS 55, 1-18, (1944).
[11] M. M. Rao Measure Theory And Integration, 2nd Edition, Marcel Dekker
Inc, (2004).
[12] Jean-Michel Rakotoson R´earrangement Relatif,
Instrument
d'Estimation dans les Probl`emes aux Limites, Math´ematiques & Ap-
plications, 64, Springer, (2008).
Un
[13] J. V. Ryff, On the Representation of Doubly Stochastic Operators,
Pac.J.Maths.13, 1379-1386, (1963).
[14] J. V. Ryff, Orbits of L1 Functions Under Doubly Stochastic Transforma-
tions, Trans.AMS 117, 92-100, (1965).
[15] J. V. Ryff, Extreme points of some convex subsets of L1(0, 1), Proc. AMS
18 , 1026-1034, (1967).
King Khaled University, Abha, Saudi Arabia.
E-mail address: [email protected]
10
|
1704.01360 | 2 | 1704 | 2018-04-28T14:14:59 | Two modified proximal point algorithms in geodesic spaces with curvature bounded above | [
"math.FA",
"math.OC"
] | We obtain existence and convergence theorems on two variants of the proximal point algorithm for proper lower semicontinuous convex functions in complete geodesic spaces with curvature bounded above. | math.FA | math |
TWO MODIFIED PROXIMAL POINT ALGORITHMS IN
GEODESIC SPACES WITH CURVATURE BOUNDED ABOVE
YASUNORI KIMURA AND FUMIAKI KOHSAKA
Abstract. We obtain existence and convergence theorems for two variants
of the proximal point algorithm involving proper lower semicontinuous convex
functions in complete geodesic spaces with curvature bounded above.
1. Introduction
The aim of this paper is to study the asymptotic behavior of sequences generated
by two variants of the proximal point algorithm for proper lower semicontinuous
convex functions in admissible complete CAT(1) spaces. We focus not only on the
convergence of the sequences to minimizers of functions but also on the equivalence
between their boundedness and the existence of minimizers. Applications to convex
minimization problems in complete CAT(κ) spaces with a positive real number κ
are also included.
The proximal point algorithm introduced by Martinet [31] and studied more gen-
erally by Rockafellar [36] is an iterative method for finding zero points of maximal
monotone operators in Hilbert spaces. Bruck and Reich [12] also obtained some con-
vergence theorems for m-accretive operators in Banach spaces. It is known that this
algorithm has a wide range of applications including convex minimization problems,
variational inequality problems, minimax problems, and equilibrium problems.
For a proper lower semicontinuous convex function f of a Hilbert space H into
]−∞,∞], the proximal point algorithm generates a sequence {xn} by x1 ∈ H and
(1.1)
where {λn} is a sequence of positive real numbers and Jλnf is the resolvent of λnf
defined by
xn+1 = Jλn f xn
(n = 1, 2, . . . ),
Jλnf x = argmin
y∈H (cid:26)f (y) +
1
2λn ky − xk2(cid:27)
for all n ∈ N and x ∈ H. See also [8, 39] for more details on convex analysis in
Hilbert spaces.
The celebrated theorem by Rockafellar [36, Theorem 1] implies the following
existence and weak convergence theorems on the sequence {xn} defined by (1.1).
If inf n λn > 0, then {xn} is bounded if and only if the set argminH f of all min-
imizers of f is nonempty. Further, in this case, {xn} is weakly convergent to an
element of argminH f . Br´ezis and Lions [10, Th´eor`eme 9] also showed that {xn}
is weakly convergent to an element of argminH f if P∞n=1 λn = ∞ and argminH f
2010 Mathematics Subject Classification. Primary: 52A41, 90C25; Secondary: 47H10, 47J05.
Key words and phrases. CAT(1) space, convex function, fixed point, geodesic space, minimizer,
proximal point algorithm, resolvent.
1
2
Y. KIMURA AND F. KOHSAKA
is nonempty. Later, Guler [18, Corollary 5.1] and Bauschke, Matouskov´a, and Re-
ich [9, Corollary 7.1] found the counterexamples to the strong convergence of {xn}.
By assuming the so-called convergence condition, Nevanlinna and Reich [33, Theo-
rem 2] obtained a strong convergence theorem for m-accretive operators in Banach
spaces. In 2000, Solodov and Svaiter [38] and Kamimura and Takahashi [20] pro-
posed two different types of strongly convergent proximal-type algorithms in Hilbert
spaces.
On the other hand, a CAT(κ) space is a geodesic metric space such that every
geodesic triangle in it satisfies the CAT(κ) inequality, where κ is a real number. A
complete CAT(0) space is particularly called an Hadamard space. Since the concept
of CAT(κ) spaces includes several fundamental spaces, the fixed point theory and
the convex optimization theory in such spaces have been increasingly important.
See, for instance, [5, 11, 35] for more details in this direction.
In the 1990s, Jost [19] and Mayer [32] generalized the concept of resolvents of
convex functions to Hadamard spaces. According to [5, Section 2.2], [19, Lemma 2],
and [32, Section 1.3], if f is a proper lower semicontinuous convex function of an
Hadamard space X into ]−∞,∞], then the resolvent Jf of f given by
(1.2)
Jf x = argmin
y∈X (cid:26)f (y) +
1
2
d(y, x)2(cid:27)
for all x ∈ X is a single-valued nonexpansive mapping of X into itself. In this case,
the set F (Jf ) of all fixed points of Jf coincides with argminX f . See also [5] on
convex analysis in Hadamard spaces.
In 2013, Bac´ak [4] generalized the classical theorem by Br´ezis and Lions [10,
Th´eor`eme 9] to Hadamard spaces. Some related asymptotic results were found by
Ariza-Ruiz, Leu¸stean, and L´opez-Acedo [3, Corollary 6.6] and Bac´ak and Reich [6,
Proposition 1.5].
Theorem 1.1 ([4, Theorem 1.4]). Let X be an Hadamard space, f a proper lower
semicontinuous convex function of X into ]−∞,∞], Jηf the resolvent of ηf for all
η > 0, and {xn} a sequence defined by x1 ∈ X and (1.1), where {λn} is a sequence
of positive real numbers such that P∞n=1 λn = ∞. If argminX f is nonempty, then
{xn} is ∆-convergent to an element of argminX f .
Motivated by [2, 4, 20], the authors [23] recently obtained the following exis-
tence and convergence theorems for two variants of the proximal point algorithm
in Hadamard spaces. The algorithms (1.3) and (1.4) were originally introduced
by Eckstein and Bertsekas [14] and Kamimura and Takahashi [20] for maximal
monotone operators in Hilbert spaces, respectively.
Theorem 1.2 ([23, Theorem 4.2]). Let X, f , {Jηf}η>0 be the same as in Theo-
rem 1.1 and {xn} a sequence defined by x1 ∈ X and
(1.3)
(n = 1, 2, . . . ),
xn+1 = αnxn ⊕ (1 − αn)Jλn f xn
where {αn} is a sequence in [0, 1[ and {λn} is a sequence of positive real numbers
such that P∞n=1(1 − αn)λn = ∞. Then the following hold.
are ∆-convergent to an element x∞ of argminX f .
(i) The set argminX f is nonempty if and only if {Jλnf xn} is bounded;
(ii) if argminX f is nonempty and supn αn < 1, then both {xn} and {Jλnf xn}
TWO MODIFIED PROXIMAL POINT ALGORITHMS
3
Theorem 1.3 ([23, Theorem 5.1]). Let X, f , and {Jηf}η>0 be the same as in
Theorem 1.1, v an element of X, and {yn} a sequence defined by y1 ∈ X and
(1.4)
(n = 1, 2, . . . ),
yn+1 = αnv ⊕ (1 − αn)Jλnf yn
where {αn} is a sequence in [0, 1] and {λn} is a sequence of positive real numbers
such that limn λn = ∞. Then the following hold.
(i) The set argminX f is nonempty if and only if {Jλnf yn} is bounded;
(ii) if argminX f is nonempty, limn αn = 0, and P∞n=1 αn = ∞, then both {yn}
and {Jλnf yn} are convergent to P v, where P denotes the metric projection
of X onto argminX f .
Theorem 1.4 ([23, Theorem 5.4]). Let X, f , and {Jηf}η>0 be the same as in
Theorem 1.1, v an element of X, and {yn} a sequence defined by y1 ∈ X and (1.4),
where {αn} is a sequence in ]0, 1] and {λn} is a sequence of positive real numbers
such that
αn = 0,
lim
n→∞
∞
Xn=1
αn = ∞,
and
inf
n
λn > 0.
If argminX f is nonempty, then both {yn} and {Jλnf yn} are convergent to P v,
where P denotes the metric projection of X onto argminX f .
In 2015, Ohta and P´alfia [34, Definition 4.1 and Lemma 4.2] showed that the
resolvent Jf given by (1.2) is still well defined in a complete CAT(1) space such
that diam(X) < π/2, where diam(X) denotes the diameter of X. Using this result,
they [34, Theorem 5.1] obtained a ∆-convergence theorem on the proximal point
algorithm in such spaces. It should be noted that the condition that diam(X) < π/2
for a complete CAT(1) space X corresponds to the boundedness condition for an
Hadamard space. In fact, every sequence in a complete CAT(1) space X such that
diam(X) < π/2 has a ∆-convergent subsequence.
In 2016, the authors [22] introduced another type of resolvents of convex func-
tions in CAT(1) spaces. For a given proper lower semicontinuous convex function f
of an admissible complete CAT(1) space X into ]−∞,∞], they [22, Definition 4.3]
defined the resolvent Rf of f by
(1.5)
Rf x = argmin
y∈X (cid:8)f (y) + tan d(y, x) sin d(y, x)(cid:9)
for all x ∈ X. Following [24], we say that a CAT(1) space X is admissible
(1.6)
d(w, w′) <
π
2
for all w, w′ ∈ X.
Recently, the authors [24] obtained the following result on the proximal point al-
gorithm in CAT(1) spaces. We note that the ∆-convergence of {xn} in Theorem 1.5
was also found independently by Esp´ınola and Nicolae [16, Theorem 3.2].
Theorem 1.5 ([24, Theorems 1.1 and 1.2]). Let X be an admissible complete
CAT(1) space, f a proper lower semicontinuous convex function of X into ]−∞,∞],
Rηf the resolvent of ηf for all η > 0, and {xn} a sequence defined by x1 ∈ X and
xn+1 = Rλnf xn
(n = 1, 2, . . . ),
4
Y. KIMURA AND F. KOHSAKA
where {λn} is a sequence of positive real numbers such that P∞n=1 λn = ∞. Then
argminX f is nonempty if and only if
π
2
d(y, xn) <
inf
y∈X
lim sup
n→∞
and
sup
d(xn+1, xn) <
n
π
2
.
Further, in this case, {xn} is ∆-convergent to an element of argminX f .
Motivated by the papers mentioned above, we study the asymptotic behavior of
sequences generated by x1, y1, v ∈ X,
(1.7)
xn+1 = αnxn ⊕ (1 − αn)Rλnf xn
(n = 1, 2, . . . ),
and
yn+1 = αnv ⊕ (1 − αn)Rλnf yn
(1.8)
where {αn} is a sequence in [0, 1], {λn} is a sequence of positive real numbers,
X is an admissible complete CAT(1) space, f is a proper lower semicontinuous
convex function of X into ]−∞,∞], and Rλnf is the resolvent of λnf for all n ∈ N
given by (1.5). These algorithms correspond to (1.3) and (1.4) in Hadamard spaces,
respectively.
(n = 1, 2, . . . ),
This paper is organized as follows. In Section 2, we recall some definitions and
results needed in this paper. In Section 3, we obtain some fundamental proper-
ties of resolvents of convex functions in complete CAT(1) spaces.
In Sections 4
and 5, we study the asymptotic behavior of sequences generated by (1.7) and (1.8),
respectively. The three main results in this paper, Theorems 4.1, 5.1, and 5.2 in
admissible complete CAT(1) spaces, correspond to Theorems 1.2, 1.3, and 1.4 in
Hadamard spaces, respectively.
In Section 6, we deduce three corollaries of our
results in complete CAT(κ) spaces with a positive real number κ.
2. Preliminaries
Throughout this paper, we denote by N the set of all positive integers, R the set of
all real numbers, ]−∞,∞] the set R∪{∞}, R2 the two dimensional Euclidean space
with Euclidean metric ρR2 , S2 the unit sphere of the three dimensional Euclidean
space R3 with the spherical metric ρS2, H a real Hilbert space with inner product
h· , ·i and induced norm k·k, X a metric space with metric d, F (T ) the set of all
fixed points of a mapping T of X into itself, and argminX f or argminy∈X f (y) the
set of all minimizers of a function f of X into ]−∞,∞]. In the case where argminX f
is a singleton {p}, we sometimes identify argminX f with the single point p.
We need the following lemma.
Lemma 2.1 ([40, Lemma 2.5]; see also [1, Lemma 2.3]). Let {sn} be a sequence of
nonnegative real numbers, {αn} a sequence in [0, 1] such that P∞n=1 αn = ∞, and
{tn} a sequence of real numbers such that lim supn tn ≤ 0. Suppose that
(2.1)
for all n ∈ N. Then limn sn = 0.
sn+1 ≤ (1 − αn)sn + αntn
Saejung and Yotkaew [37] found the following variant of Lemma 2.1. Later,
Kimura and Saejung [25] filled in a slight gap in the original proof of this result.
Although it was assumed in [25, 37] that αn < 1 for all n ∈ N, the proof in [25,
Lemma 2.8] is also valid in the case below.
TWO MODIFIED PROXIMAL POINT ALGORITHMS
5
Lemma 2.2 ([25, Lemma 2.8] and [37, Lemma 2.6]). Let {sn} be a sequence of
nonnegative real numbers, {αn} a sequence in ]0, 1] such that P∞n=1 αn = ∞, and
{tn} a sequence of real numbers. Suppose that (2.1) holds for all n ∈ N and that
lim supi tni ≤ 0 whenever {ni} is an increasing sequence in N satisfying
Then limn sn = 0.
lim sup
i→∞ (cid:0)sni − sni+1(cid:1) ≤ 0.
Let κ be a nonnegative real number and Dκ the extended real number defined
by Dκ = ∞ if κ = 0 and π/√κ if κ > 0. A metric space X is said to be Dκ-geodesic
if for each x, y ∈ X with d(x, y) < Dκ, there exists a mapping c of [0, l] into X such
that c(0) = x, c(l) = y, and
d(cid:0)c(t1), c(t2)(cid:1) = t1 − t2
for all t1, t2 ∈ [0, l], where l = d(x, y). The mapping c is called a geodesic path
from x to y. The image of c is denoted by [x, y]c and is called a geodesic segment
between x and y. We denote by αx ⊕c (1 − α)y the point given by
αx ⊕c (1 − α)y = c(cid:0)(1 − α)l(cid:1)
for all α ∈ [0, 1]. A Dκ-geodesic metric space is also called a Dκ-geodesic space. An
∞-geodesic metric space is also called a geodesic metric space or a geodesic space.
A subset F of a Dκ-geodesic space X such that d(w, w′) < Dκ for all w, w′ ∈ F is
said to be convex if [x, y]c ⊂ F whenever x, y ∈ F and c is a geodesic path from
x to y. Although [x, y]c and αx ⊕c (1 − α)y depend on the choice of a geodesic
path c from x to y, we sometimes denote them simply by [x, y] and αx ⊕ (1 − α)y,
respectively. They are determined uniquely if the space X is uniquely Dκ-geodesic,
that is, for each x, y ∈ X with d(x, y) < Dκ, there exists a unique geodesic path
from x to y.
If H is a Hilbert space, then the unit sphere SH of H is a uniquely π-geodesic
complete metric space with the spherical metric ρSH defined by
for all x, y ∈ SH . For all distinct x, y ∈ SH with ρSH (x, y) < π, the unique geodesic
path c from x to y is given by
ρSH (x, y) = arccoshx, yi
c(t) = (cos t)x + (sin t) ·
y − hx, yi x
ky − hx, yi xk
for all t ∈ [0, ρSH (x, y)]. The space (SH , ρSH ) is called a Hilbert sphere. See [7, 11,
17] for more details on Hilbert spheres.
Let (Mκ, dκ) be the uniquely Dκ-geodesic space given by
(Mκ, dκ) =((cid:0)R2, ρR2(cid:1)
(cid:16)S2, 1
√κ ρS2(cid:17) (κ > 0).
(κ = 0);
If κ is a nonnegative real number, X is a Dκ-geodesic space, and x1, x2, x3 are
points of X satisfying
(2.2)
then there exist ¯x1, ¯x2, ¯x3 ∈ Mκ such that
d(x1, x2) + d(x2, x3) + d(x3, x1) < 2Dκ,
d(xi, xj) = dκ(¯xi, ¯xj )
6
Y. KIMURA AND F. KOHSAKA
for all i, j ∈ {1, 2, 3}; see [11, Lemma 2.14 in Chapter I.2]. The two sets ∆ and ¯∆
given by
∆ = [x1, x2] ∪ [x2, x3] ∪ [x3, x1] and
¯∆ = [¯x1, ¯x2] ∪ [¯x2, ¯x3] ∪ [¯x3, ¯x1]
are called a geodesic triangle with vertices x1, x2, x3 in X and a comparison triangle
for ∆, respectively. A point ¯p ∈ ¯∆ is called a comparison point for p ∈ ∆ if
p ∈ [xi, xj],
¯p ∈ [¯xi, ¯xj],
and d(xi, p) = dκ(¯xi, ¯p)
for some distinct i, j ∈ {1, 2, 3}. A metric space X is said to be a CAT(κ) space if
it is Dκ-geodesic and the CAT(κ) inequality
d(p, q) ≤ dκ(¯p, ¯q)
holds whenever ∆ is a geodesic triangle with vertices x1, x2, x3 ∈ X satisfying (2.2),
¯∆ is a comparison triangle for ∆, and ¯p, ¯q ∈ ¯∆ are comparison points for p, q ∈ ∆,
respectively. In this case, the space X is also uniquely Dκ-geodesic. Every CAT(κ)
space is a CAT(κ′) space for all κ′ ∈ ]κ,∞[. A complete CAT(0) space is particularly
called an Hadamard space. The class of Hadamard spaces includes nonempty closed
convex subsets of Hilbert spaces, open unit balls of Hilbert spaces with hyperbolic
metric, Hadamard manifolds, and complete R-trees. The class of complete CAT(1)
spaces includes Hadamard spaces and Hilbert spheres with spherical metric. We
say that a CAT(1) space X is admissible if (1.6) holds for all w, w′ ∈ X. If κ > 0,
then (X, d) is a complete CAT(κ) space such that d(w, w′) < Dκ/2 for all w, w′ ∈ X
if and only if (X,√κd) is an admissible complete CAT(1) space. See [5, 11, 13] for
more details on CAT(κ) spaces.
We know that if X is a CAT(1) space, x1, x2, x3 ∈ X satisfy (2.2) for κ = 1, and
cos d(cid:0)αx1 ⊕ (1 − α)x2, x3(cid:1) ≥ α cos d(x1, x3) + (1 − α) cos d(x2, x3).
We also know the following fundamental inequalities.
Lemma 2.3 ([27, Corollary 2.2]). If X is a CAT(1) space, x1, x2, x3 ∈ X sat-
isfy (2.2) for κ = 1, and α ∈ [0, 1], then
(2.4)
cos d(cid:0)αx1 ⊕ (1 − α)x2, x3(cid:1) sin d(x1, x2)
≥ cos d(x1, x3) sin(cid:0)αd(x1, x2)(cid:1) + cos d(x2, x3) sin(cid:0)(1 − α)d(x1, x2)(cid:1).
Lemma 2.4 ([28, Lemma 3.1]). If X is an admissible CAT(1) space, x1, x2, x3 ∈ X,
and α ∈ [0, 1], then
α ∈ [0, 1], then
(2.3)
cos d(cid:0)αx1 ⊕ (1 − α)x2, x3(cid:1)
≥ (1 − β) cos d(x2, x3) + β ·
(2.5)
where
sin d(x1, x2) tan(cid:0) α
cos d(x1, x3)
,
2 d(x1, x2)(cid:1) + cos d(x1, x2)
(x1 6= x2);
(x1 = x2).
sin(cid:0)(1 − α)d(x1, x2)(cid:1)
sin d(x1, x2)
1 −
α
β =
The concept of ∆-convergence was originally introduced by Lim [30] in metric
spaces. Later, Kirk and Panyanak [29] applied it to the study of geodesic spaces.
TWO MODIFIED PROXIMAL POINT ALGORITHMS
7
Let X be a metric space and {xn} a sequence in X. The asymptotic center A(cid:0){xn}(cid:1)
of {xn} is defined by
A(cid:0){xn}(cid:1) =(cid:26)z ∈ X : lim sup
n→∞
d(z, xn) = inf
y∈X
lim sup
n→∞
d(y, xn)(cid:27) .
The sequence {xn} is said to be ∆-convergent to p ∈ X if
A(cid:0){xni}(cid:1) = {p}
holds for each subsequence {xni} of {xn}. In this case, {xn} is bounded and its each
subsequence is also ∆-convergent to p. If X is a nonempty closed convex subset of
a Hilbert space, then the ∆-convergence coincides with the weak convergence. We
denote by ω∆(cid:0){xn}(cid:1) the set of all points q ∈ X such that there exists a subsequence
of {xn} which is ∆-convergent to q. Following [24], we say that a sequence {xn} in
a CAT(1) space X is spherically bounded if
inf
y∈X
We know the following lemmas.
lim sup
n→∞
d(y, xn) <
π
2
.
Lemma 2.5 ([15, Proposition 4.1 and Corollary 4.4]). Let X be a complete CAT(1)
space and {xn} a spherically bounded sequence in X. Then A(cid:0){xn}(cid:1) is a singleton
and {xn} has a ∆-convergent subsequence.
Lemma 2.6 ([26, Proposition 3.1]). Let X be a complete CAT(1) space and {xn}
a spherically bounded sequence in X. If {d(z, xn)} is convergent for each element
z of ω∆(cid:0){xn}(cid:1), then {xn} is ∆-convergent to an element of X.
Let X be an admissible CAT(1) space. A function f of X into ]−∞,∞] is said
to be proper if f (a) ∈ R for some a ∈ X. It is also said to be convex if
f(cid:0)αx ⊕ (1 − α)y(cid:1) ≤ αf (x) + (1 − α)f (y)
whenever x, y ∈ X and α ∈ ]0, 1[. We denote by Γ0(X) the set of all proper
lower semicontinuous convex functions of X into ]−∞,∞]. The set argminX f
It follows from (2.3) that
is obviously closed and convex for each f ∈ Γ0(X).
− cos d(·, z) belongs to Γ0(X) for all z ∈ X. For a nonempty closed convex subset
C of X, the indicator function iC for C, which is defined by iC(x) = 0 if x ∈ C
and ∞ otherwise, belongs to Γ0(X). See [21, 41] on convex functions in CAT(1)
spaces. A function f of X into ]−∞,∞] is said to be ∆-lower semicontinuous if
f (p) ≤ lim inf n f (xn) whenever {xn} is a sequence in X which is ∆-convergent to
p ∈ X. A function g of X into [−∞,∞[ is said to be concave if −g is convex.
Let X be an admissible complete CAT(1) space and f an element of Γ0(X). It
is known [22, Theorem 4.2] that for each x ∈ X, there exists a unique x ∈ X such
that
f (x) + tan d(x, x) sin d(x, x) = inf
y∈X {f (y) + tan d(y, x) sin d(y, x)} .
Following [22, Definition 4.3], we define the resolvent Rf of f by
Rf x = x
for all x ∈ X. In other words, Rf can be defined by (1.5) for all x ∈ X. If f is
the indicator function iC for a nonempty closed convex subset C of X, then the
8
Y. KIMURA AND F. KOHSAKA
resolvent Rf coincides with the metric projection PC of X onto C, that is,
Rf x = argmin
tan d(y, x) sin d(y, x) = argmin
d(y, x) = PC x
y∈C
y∈C
for all x ∈ X.
mapping of X into itself,
It is known [22, Theorems 4.2 and 4.6] that Rf is a well-defined and single-valued
(2.6)
and
(2.7)
F (Rf ) = argmin
X
f,
x(1 + C 2
(cid:0)C 2
≥ C 2
x(1 + C 2
y )Cy + C 2
y (1 + C 2
y ) cos d(Rf x, y) + C 2
x)Cx(cid:1) cos d(Rf x, Rf y)
y (1 + C 2
x) cos d(Rf y, x)
for all x, y ∈ X, where Cz the real number given by Cz = cos d(Rf z, z) for all
z ∈ X.
The following lemma was recently obtained in [24, Lemma 3.1 and Corollary 3.2].
The inequality (2.9) is a generalization of (2.7) and corresponds to [2, Lemma 3.1]
in Banach spaces.
Lemma 2.7 ([24, Lemma 3.1 and Corollary 3.2]). Let X be an admissible complete
CAT(1) space, f an element of Γ0(X), Rηf the resolvent of ηf for all η > 0, and
Cη,z the real number given by
Cη,z = cos d(Rηf z, z)
(2.8)
for all η > 0 and z ∈ X. Then
(2.9)
λ,x(1 + C 2
(cid:0)λC 2
≥ λC 2
µ,y)Cµ,y + µC 2
µ,y(1 + C 2
λ,x(1 + C 2
µ,y) cos d(Rλf x, y) + µC 2
λ,x)Cλ,x(cid:1) cos d(Rλf x, Rµf y)
µ,y(1 + C 2
λ,x) cos d(Rµf y, x)
holds for all x, y ∈ X and λ, µ > 0. Further,
π
2 1
C 2
λ,x
+ 1!(cid:0)Cλ,x cos d(u, Rλf x) − cos d(u, x)(cid:1) ≥ λ(cid:0)f (Rλf x) − f (u)(cid:1)
(2.10)
and
(2.11)
hold for all x ∈ X, u ∈ argminX f , and λ > 0.
cos d(Rλf x, x) cos d(u, Rλf x) ≥ cos d(u, x),
We also know the following results.
Lemma 2.8 ([22, Lemma 3.1]). If X is an admissible complete CAT(1) space, then
every f ∈ Γ0(X) is ∆-lower semicontinuous.
Theorem 2.9 ([24, Theorem 4.1]). Let X be an admissible complete CAT(1) space,
{zn} a spherically bounded sequence in X, {βn} a sequence of positive real numbers
such that P∞n=1 βn = ∞, and g the real function on X defined by
n
βk cos d(y, zk)
(2.12)
g(y) = lim inf
n→∞
1
l=1 βl
Pn
Xk=1
for all y ∈ X. Then g is a 1-Lipschitz continuous and concave function of X into
[0, 1] which has a unique maximizer.
TWO MODIFIED PROXIMAL POINT ALGORITHMS
9
It is obvious that if A is a nonempty bounded subset of R, I is a closed subset of
R containing A, and f is a continuous and nondecreasing real function on I, then
f (sup A) = sup f (A) and f (inf A) = inf f (A). Thus we obtain the following.
Lemma 2.10. Let I be a nonempty closed subset of R, {tn} a bounded sequence
in I, and f a continuous real function on I. Then the following hold.
(i) If f is nondecreasing, then f (lim supn tn) = lim supn f (tn);
(ii) if f is nonincreasing, then f (lim supn tn) = lim inf n f (tn).
3. Resolvents of convex functions in CAT(1) spaces
In this section, we obtain three fundamental lemmas on the resolvents of convex
functions in CAT(1) spaces.
The following lemma corresponds to [2, Lemmas 3.5 and 3.6] in Banach spaces.
Lemma 3.1. Let X be an admissible complete CAT(1) space, f an element of
Γ0(X), Rηf the resolvent of ηf for all η > 0, {λn} a sequence of positive real
numbers, p an element of X, and {xn} a sequence in X. Then the following hold.
(i) If inf n λn > 0, A(cid:0){xn}(cid:1) = {p}, and limn d(Rλnf xn, xn) = 0, then p is an
(ii) if limn λn = ∞, A(cid:0){Rλnf xn}(cid:1) = {p}, and supn d(Rλnf xn, xn) < π/2, then
Proof. Let Cη,z be the real number in ]0, 1] given by (2.8) for all η > 0 and z ∈ X.
It follows from (2.9) that
p is an element of argminX f .
element of argminX f ;
(cid:0)λnC 2
≥ λnC 2
λn,xn (1 + C 2
1,p) + C 2
1,p(1 + C 2
λn,xn(1 + C 2
1,p) cos d(Rλn f xn, p) + C 2
λn,xn )(cid:1) cos d(Rλnf xn, Rf p)
1,p(1 + C 2
λn,xn) cos d(Rf p, xn)
and hence
(3.1)
cos d(Rλn f xn, Rf p)
≥ cos d(Rλnf xn, p)
+
C 2
1,p
1 + C 2
1,p ·
1 + C 2
λnC 2
λn,xn
λn ,xn (cid:0)cos d(Rf p, xn) − cos d(Rf p, Rλnf xn)(cid:1)
for all n ∈ N.
and inf n λn > 0, the sequence
We first show (i). Suppose that the assumptions hold. Since limn Cλn ,xn = 1
( 1 + C 2
λnC 2
λn,xn
λn,xn )
is bounded. Since t 7→ cos t is 1-Lipschitz continuous and limn d(Rλn f xn, xn) = 0,
we have
cos d(Rf p, xn) − cos d(Rf p, Rλnf xn) ≤ d(Rf p, xn) − d(Rf p, Rλnf xn)
≤ d(xn, Rλn f xn) → 0
as n → ∞. Taking the lower limit in (3.1), we have
cos d(Rλnf xn, Rf p) ≥ lim inf
n→∞
lim inf
n→∞
cos d(Rλnf xn, p).
10
Y. KIMURA AND F. KOHSAKA
It then follows from Lemma 2.10 that
cos(cid:18)lim sup
n→∞
d(Rλn f xn, Rf p)(cid:19) ≥ cos(cid:18)lim sup
n→∞
d(Rλn f xn, p)(cid:19)
lim sup
n→∞
d(Rλnf xn, Rf p) ≤ lim sup
n→∞
d(Rλnf xn, p).
and hence
(3.2)
On the other hand, since limn d(Rλn f xn, xn) = 0, we have
(3.3)
lim sup
n→∞
d(Rλnf xn, y) = lim sup
n→∞
d(xn, y)
for all y ∈ X. By (3.2) and (3.3), we have
lim sup
n→∞
d(xn, Rf p) ≤ lim sup
n→∞
d(xn, p).
Since A(cid:0){xn}(cid:1) = {p}, we obtain Rf p = p. Consequently, it follows from (2.6) that
p ∈ argminX f .
We next show (ii). Suppose that the assumptions hold. Since
d(Rλn f xn, xn) <
sup
n
π
2
,
we know that
0 < cos(cid:18)sup
n
d(Rλn f xn, xn)(cid:19) = inf
n
Thus it follows from limn λn = ∞ that
≤
1 + C 2
λnC 2
0 <
λn ,xn
λn,xn
cos d(Rλn f xn, xn) = inf
n
Cλn,xn .
2
λn (inf m Cλm,xm)2 → 0
as n → ∞. Taking the lower limit in (3.1), we know that (3.2) holds. Then, since
A(cid:0){Rλnf xn}(cid:1) = {p}, we obtain Rf p = p and hence p ∈ argminX f .
We also need the following two lemmas.
(cid:3)
Lemma 3.2. Let X be an admissible complete CAT(1) space, f an element of
Γ0(X), Rηf the resolvent of ηf for all η > 0, and {λn} a sequence of positive real
numbers. If {xn} is a sequence in X such that
n→∞(cid:0)cos d(u, Rλnf xn) − cos d(u, xn)(cid:1) = 0
d(u, xn) <
sup
and
lim
π
2
n
for some u ∈ argminX f , then limn d(Rλn f xn, xn) = 0.
Proof. It follows from (2.11) that
(3.4)
cos d(u, Rλnf xn) ≥ cos d(Rλn f xn, xn) cos d(u, Rλnf xn) ≥ cos d(u, xn)
and hence
(3.5)
and
(3.6)
inf
n
cos d(u, Rλnf xn) ≥ inf
n
cos d(u, xn) = cos(cid:18)sup
n
d(u, xn)(cid:19) > 0
0 ≤ cos d(u, Rλnf xn)(cid:18)1 −
= cos d(u, Rλnf xn) − cos d(u, xn) → 0
cos d(u, Rλnf xn)(cid:19)
cos d(u, xn)
TWO MODIFIED PROXIMAL POINT ALGORITHMS
11
as n → ∞. Thus it follows from (3.5) and (3.6) that
(3.7)
cos d(u, xn)
= 1.
lim
n→∞
cos d(u, Rλnf xn)
By (3.4) and (3.7), we obtain
1 ≥ cos d(Rλn f xn, xn) ≥
cos d(u, xn)
cos d(u, Rλnf xn) → 1
(cid:3)
as n → ∞ and hence limn d(Rλnf xn, xn) = 0.
Lemma 3.3. Let X be an admissible complete CAT(1) space, F a nonempty closed
convex subset of X, P the metric projection of X onto F , and {xn} a spherically
bounded sequence in X. If ω∆(cid:0){xn}(cid:1) is a subset of F , then
for all v ∈ X.
Proof. Let v ∈ X be given. Since {xn} is spherically bounded, Lemma 2.5 ensures
that there exists a subsequence {xni} of {xn} which is ∆-convergent to some q ∈ X
and
cos d(P v, v) ≥ lim sup
n→∞
cos d(xn, v)
(3.8)
lim
i→∞
cos d(xni , v) = lim sup
n→∞
cos d(xn, v).
(3.9)
Since ω∆(cid:0){xn}(cid:1) ⊂ F , we have q ∈ F and hence the definition of P gives us that
Since − cos d(·, v) belongs to Γ0(X), Lemma 2.8 implies that it is ∆-lower semicon-
tinuous. Thus we have
cos d(P v, v) ≥ cos d(q, v).
(3.10)
By (3.8), (3.9), and (3.10), we obtain the conclusion.
− cos d(q, v) ≤ lim inf
i→∞ (cid:0)− cos d(xni , v)(cid:1) = − lim
i→∞
cos d(xni , v).
(cid:3)
4. ∆-convergent proximal-type algorithm
In this section, using some techniques from [24], we obtain the following theorem,
which is a generalization of Theorem 1.5. This is the first one of our three main
results in this paper.
Theorem 4.1. Let X be an admissible complete CAT(1) space, f an element of
Γ0(X), Rηf the resolvent of ηf for all η > 0, and {xn} a sequence defined by x1 ∈ X
and (1.7), where {αn} is a sequence in [0, 1[ and {λn} is a sequence of positive real
numbers such that P∞n=1(1 − αn)λn = ∞. Then the following hold.
(i) The set argminX f is nonempty if and only if {Rλnf xn} is spherically
(ii) if argminX f is nonempty and supn αn < 1, then both {xn} and {Rλnf xn}
bounded and supn d(Rλnf xn, xn) < π/2;
are ∆-convergent to an element x∞ of argminX f .
Proof. Let Cη,z be the real number given by (2.8) for all η > 0 and z ∈ X and let
{zn} be the sequence in X given by zn = Rλnf xn for all n ∈ N.
12
Y. KIMURA AND F. KOHSAKA
We first show the if part of (i). Suppose that {zn} is spherically bounded and
supn d(zn, xn) < π/2 and let {βn} and {σn} be the real sequences given by
βn =
(1 − αn)λnC 2
λn,xn
1 + C 2
λn ,xn
and σn =
βl
n
Xl=1
for all n ∈ N. Since αn < 1 and λn > 0 for all n ∈ N and X is admissible, we know
that {βn} is a sequence of positive real numbers. Since supn d(zn, xn) < π/2, we
also know that
0 < cos(cid:18)sup
n
d(zn, xn)(cid:19) = inf
n
cos d(zn, xn) = inf
n
Cλn,xn .
Thus, noting that
(1 − αn)λn (inf m Cλm,xm)2
2
βn ≥
and P∞n=1(1 − αn)λn = ∞, we obtain P∞n=1 βn = ∞. According to Theorem 2.9,
the real function g on X defined by (2.12) for all y ∈ X has a unique maximizer
p ∈ X. It then follows from (2.9) that
λkC 2
1 + C 2
λk ,xk
λk ,xk
λkC 2
1 + C 2
λk ,xk
λk ,xk
≥
cos d(Rλk f xk, Rf p)
cos d(Rλk f xk, p) +
and hence
(4.1)
βk cos d(zk, Rf p)
≥ βk cos d(zk, p) +
C 2
1,p
1 + C 2
1,p(cid:0)cos d(Rf p, xk) − cos d(Rf p, Rλkf xk)(cid:1)
(1 − αk)C 2
1,p
1 + C 2
1,p
(cid:0)cos d(Rf p, xk) − cos d(Rf p, zk)(cid:1)
for all k ∈ N. On the other hand, it follows from (2.3) and the definition of {xn}
that
(4.2)
cos d(Rf p, xk+1) ≥ αk cos d(Rf p, xk) + (1 − αk) cos d(Rf p, zk)
and hence, by (4.1) and (4.2), we have
βk cos d(zk, Rf p)
(4.3)
C 2
≥ βk cos d(zk, p) +
1,p(cid:0)cos d(Rf p, xk) − cos d(Rf p, xk+1)(cid:1)
for all k ∈ N. Summing up (4.3) with respect to k ∈ {1, 2, . . . , n}, we obtain
1,p
1 + C 2
βk cos d(zk, Rf p)
1
σn
(4.4)
n
Xk=1
Xk=1
1
σn
n
C 2
1
1,p
1 + C 2
≥
βk cos d(zk, p) +
σn(cid:0)cos d(Rf p, x1) − cos d(Rf p, xn+1)(cid:1)
for all n ∈ N. Since limn σn = ∞, it follows from (4.4) that g(Rf p) ≥ g(p). Since p
is the unique maximizer of g, we obtain Rf p = p. Consequently, it follows from (2.6)
that p ∈ argminX f . Therefore, the set argminX f is nonempty.
1,p ·
TWO MODIFIED PROXIMAL POINT ALGORITHMS
13
We next show the only if part of (i). Suppose that argminX f is nonempty and
let u ∈ argminX f be given. Then it follows from (2.11) that
(4.5)
min(cid:8)cos d(u, zn), cos d(zn, xn)(cid:9) ≥ cos d(u, zn) cos d(zn, xn)
and hence
≥ cos d(u, xn)
max(cid:8)d(u, zn), d(zn, xn)(cid:9) ≤ d(u, xn)
(4.6)
for all n ∈ N. By (2.3) and (4.5), we have
(4.7)
cos d(u, xn+1) ≥ αn cos d(u, xn) + (1 − αn) cos d(u, zn) ≥ cos d(u, xn).
It then follows from (4.7) and the admissibility of X that
(4.8)
d(u, xn+1) ≤ d(u, xn) ≤ d(u, x1) <
π
2
for all n ∈ N. Thus it follows from (4.6) and (4.8) that
lim sup
n→∞
d(u, zn) ≤ lim
n→∞
d(u, xn) ≤ d(u, x1) <
π
2
.
This implies the spherical boundedness of {xn} and {zn}. It also follows from (4.6)
and (4.8) that supn d(zn, xn) < π/2.
We finally show (ii). Suppose that argminX f is nonempty and supn αn < 1.
Then we know that (4.5), (4.6), (4.7), and (4.8) hold and that both {xn} and {zn}
are spherically bounded. Let u ∈ argminX f be given. It follows from (4.8) that
{d(u, xn)} tends to some β ∈ [0, π/2[. By (2.3) and (4.5), we have
cos d(u, xn+1)
≥ αn cos d(u, xn) + (1 − αn) cos d(u, zn)
≥ αn cos d(u, xn) + (1 − αn) ·
= cos d(u, xn) + (1 − αn) cos d(u, xn)(cid:18)
cos d(u, xn)
cos d(zn, xn)
1
cos d(zn, xn) − 1(cid:19)
and hence
(4.9)
0 ≤ (1 − αn)(cid:18)
1
cos d(zn, xn) − 1(cid:19) ≤
cos d(u, xn+1)
cos d(u, xn) − 1 →
cos β
cos β − 1 = 0
as n → ∞. Since supn αn < 1, it follows from (4.9) that
(4.10)
d(zn, xn) = 0.
lim
n→∞
On the other hand, it follows from (2.10) and (4.10) that there exists a positive
real number K such that
Kπ
(4.11)
for all n ∈ N. It then follows from (4.7) and (4.11) that
λn(cid:0)f (zn) − f (u)(cid:1) ≤
(1 − αn)λn(cid:0)f (zn) − f (u)(cid:1) ≤
Xn=1
(1 − αn)λn(cid:0)f (zn) − f (u)(cid:1) ≤
2 (cid:0)cos d(u, zn) − cos d(u, xn)(cid:1)
2 (cid:0)cos d(u, xn+1) − cos d(u, xn)(cid:1)
2 (cid:0)cos β − cos d(u, x1)(cid:1) < ∞.
Kπ
Kπ
∞
and hence
(4.12)
14
Y. KIMURA AND F. KOHSAKA
(4.13)
Since P∞n=1(1 − αn)λn = ∞, it follows from (4.12) that
n→∞ (cid:0)f (zn) − f (u)(cid:1) = 0.
By the definitions of {xn} and {zn} and the convexity of f , we also have
lim inf
−∞ < inf f (X) ≤ f (zn) ≤ f (zn) +
1
λn
tan d(zn, xn) sin d(zn, xn) ≤ f (xn)
and
−∞ < inf f (X) ≤ f (xn+1) ≤ αnf (xn) + (1 − αn)f (zn) ≤ f (xn)
for all n ∈ N. Thus {f (xn)} tends to some γ ∈ R and {f (zn)} is bounded. Let
{ni} be any increasing sequence in N. Since supn αn < 1, we have a subsequence
{nij} of {ni} such that {αnij } tends to some δ ∈ [0, 1[. Then, letting j → ∞ in
1
1 − αnij (cid:16)f(cid:0)xnij
+1(cid:1) − αnij
f(cid:0)xnij(cid:1)(cid:17) ≤ f(cid:0)znij(cid:1) ≤ f(cid:0)xnij(cid:1),
we obtain f(cid:0)znij(cid:1) → γ. Thus {f (zn)} also tends to γ. Consequently, it follows
from (4.13) that
(4.14)
f (xn) = γ = f (u) = inf f (X).
lim
n→∞
Let z be any element of ω∆(cid:0){xn}(cid:1). Then we have a subsequence {xmi} of {xn}
which is ∆-convergent to z. Since f is ∆-lower semicontinuous by Lemma 2.8, it
follows from (4.14) that
f (z) ≤ lim inf
i→∞
f (xmi ) = lim
n→∞
f (xn) = inf f (X)
and hence z ∈ argminX f . Thus ω∆(cid:0){xn}(cid:1) is a subset of argminX f . It then follows
from (4.8) that {d(z, xn)} is convergent for each z ∈ ω∆(cid:0){xn}(cid:1). Thus, Lemma 2.6
ensures that {xn} is ∆-convergent to some x∞ ∈ X. Since
f,
{x∞} = ω∆(cid:0){xn}(cid:1) ⊂ argmin
X
we know that x∞ belongs to argminX f . It then follows from (4.10) that
A(cid:0){zli}(cid:1) = A(cid:0){xli}(cid:1) = {x∞}
for each increasing sequence {li} in N. Consequently, we conclude that both {xn}
and {zn} are ∆-convergent to an element x∞ of argminX f .
(cid:3)
As a direct consequence of Theorem 4.1, we obtain the following corollary.
Corollary 4.2. Let (SH , ρSH ) be a Hilbert sphere, X an admissible closed convex
subset of SH , f an element of Γ0(X), and {xn} a sequence defined by x1 ∈ X
and (1.7), where {αn} is a sequence in [0, 1[ and {λn} is a sequence of positive
real numbers such that P∞n=1(1 − αn)λn = ∞. Then argminX f is nonempty if
and only if {Rλnf xn} is spherically bounded and supn ρSH (Rλn f xn, xn) < π/2.
Further, if argminX f is nonempty and supn αn < 1, then both {xn} and {Rλnf xn}
are ∆-convergent to an element x∞ of argminX f .
TWO MODIFIED PROXIMAL POINT ALGORITHMS
15
5. Convergent proximal-type algorithm
In this section, using some techniques from [28, Theorem 3.2], we first obtain the
following theorem. This is the second one of our three main results in this paper.
Theorem 5.1. Let X be an admissible complete CAT(1) space, f an element of
Γ0(X), Rηf the resolvent of ηf for all η > 0, v an element of X, and {yn} a
sequence defined by y1 ∈ X and (1.8), where {αn} is a sequence in [0, 1] and {λn}
is a sequence of positive real numbers such that limn λn = ∞. Then the following
hold.
bounded and supn d(Rλnf yn, yn) < π/2;
(i) The set argminX f is nonempty if and only if {Rλnf yn} is spherically
(ii) if argminX f is nonempty, limn αn = 0, and P∞n=1 α2
n = ∞, then both {yn}
and {Rλnf yn} are convergent to P v, where P denotes the metric projection
of X onto argminX f .
Proof. Let {zn} be the sequence in X given by zn = Rλn f yn for all n ∈ N.
We first show the if part of (i). Suppose that {zn} is spherically bounded and
supn d(zn, yn) < π/2. Then Lemma 2.5 implies that there exists p ∈ X such that
A(cid:0){zn}(cid:1) = {p}. Since
lim
n→∞
λn = ∞ and
sup
n
d(zn, yn) <
π
2
,
it follows from (ii) of Lemma 3.1 that p ∈ argminX f . Thus argminX f is nonempty.
We next show the only if part of (i). Suppose that argminX f is nonempty and
let P be the metric projection of X onto argminX f . It follows from (2.11) that
(5.1)
for all n ∈ N. It also follows from (2.3) and (2.11) that
max(cid:8)d(P v, zn), d(zn, yn)(cid:9) ≤ d(P v, yn)
cos d(P v, yn+1) ≥ αn cos d(P v, v) + (1 − αn) cos d(P v, yn).
This implies that
for all n ∈ N and hence
(5.2)
cos d(P v, yn) ≥ min(cid:8)cos d(P v, v), cos d(P v, y1)(cid:9)
d(P v, yn) ≤ max(cid:8)d(P v, v), d(P v, y1)(cid:9) <
π
2
for all n ∈ N, where the last inequality follows from the admissibility of X. Then,
by (5.1) and (5.2), we see that both {yn} and {zn} are spherically bounded and
(5.3)
d(zn, yn) <
sup
.
n
π
2
P∞n=1 α2
and {zn} are spherically bounded. By (2.5) and (5.1), we have
We finally show (ii). Suppose that argminX f is nonempty, limn αn = 0, and
n = ∞. Then we know that (5.1), (5.2), and (5.3) hold and that both {yn}
cos d(cid:0)P v, yn+1(cid:1)
≥ (1 − βn) cos d(P v, yn) + βn ·
cos d(P v, v)
(5.4)
,
sin d(zn, v) tan(cid:0) αn
2 d(zn, v)(cid:1) + cos d(zn, v)
16
where
(5.5)
Y. KIMURA AND F. KOHSAKA
sin(cid:0)(1 − αn)d(zn, v)(cid:1)
sin d(zn, v)
1 −
αn
(zn 6= v);
(zn = v)
βn =
for all n ∈ N. Note that if zn 6= v, then it follows from
sin(cid:0)(1 − αn)d(zn, v)(cid:1) ≥ αn sin 0 + (1 − αn) sin d(zn, v) = (1 − αn) sin d(zn, v)
that αn ≥ βn. Hence we have
(5.6)
for all n ∈ N. Letting
(5.7)
αn ≥ βn
sn = 1 − cos d(P v, yn)
and
(5.8)
tn = 1 −
we have from (5.4) that
cos d(P v, v)
sin d(zn, v) tan(cid:0) αn
sn+1 ≤ (1 − βn)sn + βntn
2 d(zn, v)(cid:1) + cos d(zn, v)
,
(5.9)
for all n ∈ N.
Let us show that P∞n=1 βn = ∞. Suppose that zn 6= v.
ϕ(t) = sin(γt)/ sin t for all t ∈ ]0, π/2], then we have
(cid:0)γ tan t − tan(γt)(cid:1)
(cid:16)γ tan t −(cid:0)(1 − γ) tan 0 + γ tan t(cid:1)(cid:17) = 0
cos(γt) cos t
cos(γt) cos t
ϕ′(t) =
sin2 t
sin2 t
≥
for all t ∈ ]0, π/2[. Thus the function
t 7→
sin(cid:0)(1 − αn)t(cid:1)
sin t
If γ ∈ [0, 1] and
is nondecreasing on ]0, π/2]. Using this property and the inequality 1− t2/4 ≥ cos t
on [0, π/2], we have
(1 − αn)π
βn ≥ 1 − sin
= 1 − cos
2
If zn = v, then βn = αn ≥ α2
n ≥ 16−1π2α2
n. Hence the inequality
π2
βn ≥
(5.10)
16
α2
n
=
αnπ
2 ≥
1
4(cid:16) αnπ
2 (cid:17)2
α2
nπ2
16
.
holds for all n ∈ N. It then follows from P∞n=1 α2
We next show that lim supn tn ≤ 0. Since limn αn = 0, we have
n = ∞ that P∞n=1 βn = ∞.
If lim supn cos d(zn, v) = 0, then we have
sin d(zn, v) tan
lim
n→∞
= 0.
αnd(zn, v)
2
cos d(P v, v)
lim sup
n→∞
tn = 1 − lim inf
n→∞
sin d(zn, v) tan(cid:0) αn
2 d(zn, v)(cid:1) + cos d(zn, v)
= −∞.
TWO MODIFIED PROXIMAL POINT ALGORITHMS
17
If lim supn cos d(zn, v) > 0, then it follows from Lemma 2.10 that
tn
lim sup
n→∞
= 1 − lim inf
n→∞
(5.11)
= 1 −
= 1 −
cos d(P v, v)
sin d(zn, v) tan(cid:0) αn
2 d(zn, v)(cid:1) + cos d(zn, v)
cos d(P v, v)
lim supn→∞(cid:16)sin d(zn, v) tan(cid:0) αn
lim supn→∞
cos d(P v, v)
cos d(zn, v)
.
2 d(zn, v)(cid:1) + cos d(zn, v)(cid:17)
On the other hand, if q is any element of ω∆(cid:0){zn}(cid:1), then there exists a subsequence
{zni} of {zn} which is ∆-convergent to some q ∈ X. Since
lim
i→∞
λni = ∞, A(cid:0){Rλni f yni}(cid:1) = {q},
and
sup
i
d(Rλni f yni , yni) <
π
2
,
it follows from (ii) of Lemma 3.1 that q ∈ argminX f . Thus ω∆(cid:0){zn}(cid:1) is a subset
of argminX f . It then follows from Lemma 3.3 that
(5.12)
cos d(zn, v).
cos d(P v, v) ≥ lim sup
n→∞
By (5.11) and (5.12), we know that lim supn tn ≤ 0.
Therefore, Lemma 2.1 yields that limn sn = 0 and hence
(5.13)
d(P v, yn) = 0.
lim
n→∞
Consequently, by (5.1) and (5.13), we conclude that both {yn} and {zn} are con-
vergent to P v.
(cid:3)
Xn=1
Using Lemma 2.2, we next obtain the last one of our three main results in this
paper.
Theorem 5.2. Let X, f , {Rηf}η>0, and v be the same as in Theorem 5.1, and
{yn} a sequence defined by y1 ∈ X and (1.8), where {αn} is a sequence in ]0, 1]
and {λn} is a sequence of positive real numbers such that
inf
n
αn = 0,
λn > 0.
and
∞
α2
n = ∞,
lim
n→∞
If argminX f is nonempty, then both {yn} and {Rλnf yn} are convergent to P v,
where P denotes the metric projection of X onto argminX f .
Proof. Let {zn} be the sequence in X given by zn = Rλn f yn for all n ∈ N.
As in the proof of Theorem 5.1, we can see that (5.1), (5.2), and (5.3) hold and
that both {yn} and {zn} are spherically bounded. Let {βn}, {sn}, and {tn} be
the real sequences defined by (5.5), (5.7), and (5.8), respectively. Then we can see
that (5.6), (5.9), and (5.10) hold for all n ∈ N. Since {αn} is a sequence in ]0, 1],
so is {βn}. Since P∞n=1 α2
Let {ni} be any increasing sequence in N such that
i→∞ (cid:0)sni − sni+1(cid:1) ≤ 0.
n = ∞, it follows from (5.10) that P∞n=1 βn = ∞.
lim sup
(5.14)
Then we show that
(5.15)
lim sup
i→∞
tni ≤ 0.
18
Y. KIMURA AND F. KOHSAKA
If lim supi cos d(zni , v) = 0, then we have
If lim supi cos d(zni , v) > 0, then it follows from Lemma 2.10 that
lim sup
i→∞
tni = 1 − lim inf
i→∞
(5.16)
lim sup
i→∞
It follows from (2.3) that
cos d(P v, v)
2 d(zni , v)(cid:1) + cos d(zni , v)
= −∞.
sin d(zni, v) tan(cid:0) αni
tni = 1 −
cos d(P v, v)
lim supi→∞
cos d(zni, v)
.
sni − sni+1 = cos d(P v, yni+1) − cos d(P v, yni )
≥ αni cos d(P v, v) + (1 − αni ) cos d(P v, zni) − cos d(P v, yni ).
Hence (2.11) yields that
(5.17)
sni − sni+1 + αni(cid:0)cos d(P v, zni ) − cos d(P v, v)(cid:1)
≥ cos d(P v, zni ) − cos d(P v, yni)
≥ 0.
Since limi αni = 0, it follows from (5.14) and (5.17) that
On the other hand, it follows from (5.2) that
lim
i→∞(cid:0)cos d(P v, zni) − cos d(P v, yni )(cid:1) = 0.
sup
i
d(P v, yni ) ≤ sup
n
d(P v, yn) <
π
2
.
Thus it follows from Lemma 3.2 that limi d(zni, yni) = 0. Let q be any element of
ω∆(cid:0){zni}(cid:1). Then there exists a subsequence {znij } of {zni} which is ∆-convergent
to some q ∈ X. Since
inf
j
λnij
> 0, A(cid:0){ynij}(cid:1) = {q},
and
lim
j→∞
d(cid:0)Rλnij
f ynij
, ynij(cid:1) = 0,
it follows from (i) of Lemma 3.1 that q ∈ argminX f . Thus ω∆(cid:0){zni}(cid:1) is a subset
of argminX f . Then, by Lemma 3.3, we know that
(5.18)
cos d(zni, v).
cos d(P v, v) ≥ lim sup
i→∞
By (5.16) and (5.18), we know that (5.15) holds.
Therefore, Lemma 2.2 yields that limn sn = 0 and hence
(5.19)
d(P v, yn) = 0.
lim
n→∞
Consequently, by (5.1) and (5.19), we conclude that both {yn} and {zn} are con-
vergent to P v.
(cid:3)
As direct consequences of Theorems 5.1 and 5.2, we obtain the following two
corollaries, respectively.
Corollary 5.3. Let (SH , ρSH ) be a Hilbert sphere, X an admissible closed convex
subset of SH , f an element of Γ0(X), v an element of X, and {yn} a sequence de-
fined by y1 ∈ X and (1.8), where {αn} is a sequence in [0, 1] and {λn} is a sequence
of positive real numbers such that limn λn = ∞. Then argminX f is nonempty if
and only if {Rλnf yn} is spherically bounded and supn ρSH (Rλnf yn, yn) < π/2. Fur-
ther, if argminX f is nonempty, limn αn = 0, and P∞n=1 α2
n = ∞, then both {yn}
TWO MODIFIED PROXIMAL POINT ALGORITHMS
19
and {Rλnf yn} are convergent to P v, where P denotes the metric projection of X
onto argminX f .
Corollary 5.4. Let (SH , ρSH ), X, f , and v be the same as in Corollary 5.3 and
{yn} a sequence defined by y1 ∈ X and (1.8), where {αn} is a sequence in ]0, 1] and
{λn} is a sequence of positive real numbers such that limn αn = 0, P∞n=1 α2
n = ∞,
and inf n λn > 0. If argminX f is nonempty, then both {yn} and {Rλnf yn} are
convergent to P v, where P denotes the metric projection of X onto argminX f .
6. Results in CAT(κ) spaces with a positive κ
In this final section, using Theorems 4.1, 5.1, and 5.2, we deduce three corollaries
in CAT(κ) spaces with a positive real number κ.
Throughout this section, we suppose the following.
• κ is a positive real number and Dκ = π/√κ;
• X is a complete CAT(κ) space such that d(w, w′) < Dκ/2 for all w, w′ ∈ X;
• f is a proper lower semicontinuous convex function of X into ]−∞,∞];
• Rηf is the mapping of X into itself defined by
Rηf x = argmin
y∈X (cid:26)f (y) +
1
η
tan(cid:0)√κd(y, x)(cid:1) sin(cid:0)√κd(y, x)(cid:1)(cid:27)
for all η > 0 and x ∈ X.
Since the space (X,√κd) is an admissible complete CAT(1) space, the mapping
Rηf is well defined and Theorems 4.1, 5.1, and 5.2 immediately imply the following
three corollaries, respectively.
Corollary 6.1. Let {xn} be a sequence defined by x1 ∈ X and
xn+1 = αnxn ⊕ (1 − αn) Rλnf xn
(n = 1, 2, . . . ),
where {αn} is a sequence in [0, 1[ and {λn} is a sequence of positive real numbers
such that P∞n=1(1 − αn)λn = ∞. Then argminX f is nonempty if and only if
d(y, Rλnf xn) <
Dκ
2
and
sup
n
d( Rλnf xn, xn) <
Dκ
2
.
inf
y∈X
lim sup
n→∞
Further, if argminX f is nonempty and supn αn < 1, then both {xn} and { Rλnf xn}
are ∆-convergent to an element x∞ of argminX f .
Corollary 6.2. Let v be an element of X and {yn} a sequence defined by y1 ∈ X
and
(6.1)
yn+1 = αnv ⊕ (1 − αn) Rλn f yn
(n = 1, 2, . . . ),
where {αn} is a sequence in [0, 1] and {λn} is a sequence of positive real numbers
such that limn λn = ∞. Then argminX f is nonempty if and only if
d( Rλn f yn, yn) <
d(y, Rλnf yn) <
sup
and
.
Dκ
2
n
Dκ
2
inf
y∈X
lim sup
n→∞
Further, if argminX f is nonempty, limn αn = 0, and P∞n=1 α2
n = ∞, then both
{yn} and { Rλnf yn} are convergent to P v, where P denotes the metric projection
of X onto argminX f .
20
Y. KIMURA AND F. KOHSAKA
Corollary 6.3. Let v be an element of X, and {yn} a sequence defined by y1 ∈ X
and (6.1), where {αn} is a sequence in ]0, 1] and {λn} is a sequence of positive real
numbers such that
αn = 0,
lim
n→∞
α2
n = ∞,
and
inf
n
λn > 0.
∞
Xn=1
If argminX f is nonempty, then both {yn} and { Rλnf yn} are convergent to P v,
where P denotes the metric projection of X onto argminX f .
Concluding remarks
As we stated in Section 1, it is known [34, Definition 4.1 and Lemma 4.2] that
the classical resolvent given by (1.2) is still well defined for any proper lower semi-
continuous convex function in a complete CAT(1) space whose diameter is strictly
less than π/2. However, it is also known [22, Corollary 3.3] that this diameter
condition on the space implies that such a function always has a minimizer.
On the other hand, according to [22, Theorem 4.2], we can define another type
of resolvent by (1.5) with the perturbation function tan d sin d in an admissible
complete CAT(1) space. This makes it possible for us to study the existence of
minimizers as well as the convergence to minimizers through the two proximal-type
algorithms defined by (1.7) and (1.8).
Finally, we point out that it is not clear whether there is any relationship between
the two types of resolvents and hence we cannot deduce any result for the classical
resolvents from the results obtained in this paper so far.
Acknowledgment
The authors would like to thank the anonymous referees for their helpful com-
ments on the original version of this paper. This work was supported by JSPS
KAKENHI Grant Numbers 15K05007 and 17K05372.
References
[1] K. Aoyama, Y. Kimura, W. Takahashi, and M. Toyoda, Approximation of common fixed
points of a countable family of nonexpansive mappings in a Banach space, Nonlinear Anal.
67 (2007), 2350–2360.
[2] K. Aoyama, F. Kohsaka, and W. Takahashi, Proximal point methods for monotone operators
in Banach spaces, Taiwanese J. Math. 15 (2011), 259–281.
[3] D. Ariza-Ruiz, L. Leu¸stean, and G. L´opez-Acedo, Firmly nonexpansive mappings in classes
of geodesic spaces, Trans. Amer. Math. Soc. 366 (2014), 4299–4322.
[4] M. Bac´ak, The proximal point algorithm in metric spaces, Israel J. Math. 194 (2013), 689–
701.
[5]
[6] M. Bac´ak and S. Reich, The asymptotic behavior of a class of nonlinear semigroups in
, Convex analysis and optimization in Hadamard spaces, De Gruyter, Berlin, 2014.
Hadamard spaces, J. Fixed Point Theory Appl. 16 (2014), 189–202.
[7] C. Bargetz, M. Dymond, and S. Reich, Porosity results for sets of strict contractions on
geodesic metric spaces, Topol. Methods Nonlinear Anal. 50 (2017), 89–124.
[8] H. H. Bauschke and P. L. Combettes, Convex analysis and monotone operator theory in
Hilbert spaces, Springer, New York, 2011.
[9] H. H. Bauschke, E. Matouskov´a, and S. Reich, Projection and proximal point methods: con-
vergence results and counterexamples, Nonlinear Anal. 56 (2004), 715–738.
[10] H. Br´ezis and P.-L. Lions, Produits infinis de r´esolvantes, Israel J. Math. 29 (1978), 329–345.
[11] M. R. Bridson and A. Haefliger, Metric spaces of non-positive curvature, Springer-Verlag,
Berlin, 1999.
TWO MODIFIED PROXIMAL POINT ALGORITHMS
21
[12] R. E. Bruck and S. Reich, Nonexpansive projections and resolvents of accretive operators in
Banach spaces, Houston J. Math. 3 (1977), 459–470.
[13] D. Burago, Y. Burago, and S. Ivanov, A course in metric geometry, American Mathematical
Society, Providence, RI, 2001.
[14] J. Eckstein and D. P. Bertsekas, On the Douglas-Rachford splitting method and the proximal
point algorithm for maximal monotone operators, Math. Programming 55 (1992), 293–318.
[15] R. Esp´ınola and A. Fern´andez-Le´on, CAT(k)-spaces, weak convergence and fixed points, J.
Math. Anal. Appl. 353 (2009), 410–427.
[16] R. Esp´ınola and A. Nicolae, Proximal minimization in CAT(κ) spaces, J. Nonlinear Convex
Anal. 17 (2016), 2329–2338.
[17] K. Goebel and S. Reich, Uniform convexity, hyperbolic geometry, and nonexpansive map-
pings, Marcel Dekker, Inc., New York, 1984.
[18] O. Guler, On the convergence of the proximal point algorithm for convex minimization, SIAM
J. Control Optim. 29 (1991), 403–419.
[19] J. Jost, Convex functionals and generalized harmonic maps into spaces of nonpositive cur-
vature, Comment. Math. Helv. 70 (1995), 659–673.
[20] S. Kamimura and W. Takahashi, Approximating solutions of maximal monotone operators
in Hilbert spaces, J. Approx. Theory 106 (2000), 226–240.
[21] W. S. Kendall, Convexity and the hemisphere, J. London Math. Soc. (2) 43 (1991), 567–576.
[22] Y. Kimura and F. Kohsaka, Spherical nonspreadingness of resolvents of convex functions in
geodesic spaces, J. Fixed Point Theory Appl. 18 (2016), 93–115.
[23]
[24]
, Two modified proximal point algorithms for convex functions in Hadamard spaces,
Linear Nonlinear Anal. 2 (2016), 69–86.
, The proximal point algorithm in geodesic spaces with curvature bounded above, Linear
Nonlinear Anal. 3 (2017), 133–148.
[25] Y. Kimura and S. Saejung, Strong convergence for a common fixed point of two different
generalizations of cutter operators, Linear Nonlinear Anal. 1 (2015), 53–65.
[26] Y. Kimura, S. Saejung, and P. Yotkaew, The Mann algorithm in a complete geodesic space
with curvature bounded above, Fixed Point Theory Appl. (2013), 2013:336, 1–13.
[27] Y. Kimura and K. Sato, Convergence of subsets of a complete geodesic space with curvature
bounded above, Nonlinear Anal. 75 (2012), 5079–5085.
[28]
, Halpern iteration for strongly quasinonexpansive mappings on a geodesic space with
curvature bounded above by one, Fixed Point Theory Appl. (2013), 2013:7, 1–14.
[29] W. A. Kirk and B. Panyanak, A concept of convergence in geodesic spaces, Nonlinear Anal.
68 (2008), 3689–3696.
[30] T. C. Lim, Remarks on some fixed point theorems, Proc. Amer. Math. Soc. 60 (1976), 179–
182.
[31] B. Martinet, R´egularisation d'in´equations variationnelles par approximations successives,
Rev. Fran¸caise Informat. Recherche Op´erationnelle 4 (1970), 154–158.
[32] U. F. Mayer, Gradient flows on nonpositively curved metric spaces and harmonic maps,
Comm. Anal. Geom. 6 (1998), 199–253.
[33] O. Nevanlinna and S. Reich, Strong convergence of contraction semigroups and of iterative
methods for accretive operators in Banach spaces, Israel J. Math. 32 (1979), 44–58.
[34] S. Ohta and M. P´alfia, Discrete-time gradient flows and law of large numbers in Alexandrov
spaces, Calc. Var. Partial Differential Equations 54 (2015), 1591–1610.
[35] S. Reich and I. Shafrir, Nonexpansive iterations in hyperbolic spaces, Nonlinear Anal. 15
(1990), 537–558.
[36] R. T. Rockafellar, Monotone operators and the proximal point algorithm, SIAM J. Control
Optim. 14 (1976), 877–898.
[37] S. Saejung and P. Yotkaew, Approximation of zeros of inverse strongly monotone operators
in Banach spaces, Nonlinear Anal. 75 (2012), 742–750.
[38] M. V. Solodov and B. F. Svaiter, Forcing strong convergence of proximal point iterations in
a Hilbert space, Math. Program. 87 (2000), 189–202.
[39] W. Takahashi, Introduction to nonlinear and convex analysis, Yokohama Publishers, Yoko-
hama, 2009.
[40] H.-K. Xu, Iterative algorithms for nonlinear operators, J. London Math. Soc. (2) 66 (2002),
240–256.
22
Y. KIMURA AND F. KOHSAKA
[41] T. Yokota, Convex functions and barycenter on CAT(1)-spaces of small radii, J. Math. Soc.
Japan 68 (2016), 1297–1323.
(Y. Kimura) Department of Information Science, Toho University, Miyama, Funabashi,
Chiba 274-8510, Japan
E-mail address: [email protected]
(F. Kohsaka) Department of Mathematical Sciences, Tokai University, Kitakaname,
Hiratsuka, Kanagawa 259-1292, Japan
E-mail address: [email protected]
|
1312.1059 | 1 | 1312 | 2013-12-04T08:57:22 | Fourier--Jacobi expansions in Morrey spaces | [
"math.FA",
"math.CA"
] | In this paper we obtain a characterization of the convergence of the partial sum operator related to Fourier--Jacobi expansions in Morrey spaces. | math.FA | math |
FOURIER -- JACOBI EXPANSIONS IN MORREY SPACES
ALBERTO ARENAS AND ´OSCAR CIAURRI
Abstract. In this paper we obtain a characterization of the convergence of the
partial sum operator related to Fourier -- Jacobi expansions in Morrey spaces.
1. Introduction and main results
For α, β > −1, we consider the Jacobi functions
p(α,β)
n
(x)(1 − x)α/2(1 + x)β/2,
(x) = d(α,β)
P (α,β)
n
n
x ∈ (−1, 1), n = 0, 1, 2, . . . ,
where P (α,β)
n
denotes the Jacobi polynomial of order (α, β) and degree n, and
1
d(α,β)
n
=(cid:18)Z 1
−1(cid:16)P (α,β)
n
(x)(cid:17)2
(1 − x)α(1 + x)β dx(cid:19)
1/2
.
The system of functions {p(α,β)
}n≥0 is orthonormal and complete in L2(−1, 1) with
the Lebesgue measure. Given an appropriate function f , its Fourier expansion
respect to the Jacobi functions, which we call Fourier -- Jacobi expansion, is given
by
n
f ∼
a(α,β)
k
(f )p(α,β)
k
,
a(α,β)
k
p(α,β)
k
(t)f (t) dt.
(f ) =Z 1
−1
∞
Xk=0
The convergence of the partial sum operator for the Fourier -- Jacobi expansions,
given by
n
Snf =
a(α,β)
k
(f )p(α,β)
k
,
Xk=0
which is equivalent to the uniform boundedness of Sn, has been widely analyzed
in different kinds of spaces. In the case of the Fourier series related to the Jacobi
polynomials in Lp spaces the first known results are due to Pollard [12] and Wing
[18] who treated the case α, β ≥ −1/2. In [10], Muckenhoupt extended the analysis
to the whole range α, β > −1 and included some weights. With the sufficient
conditions on the weights given in Muckenhoupt's paper and by using the results
about the necessary conditions in, for instance, [7] (about this question see [2, 8, 5]
also), the result for Fourier -- Jacobi functions establishes that
kSnf kLp(−1,1) ≤ Ckf kLp(−1,1) ⇐⇒
4
3
< p < 4.
A complete study of the boundedness properties of the partial sum operator in
Lp,∞ spaces can be seen in [4].
Date: February 5, 2018.
2010 Mathematics Subject Classification. Primary 42C10, 43A50.
Key words and phrases. Fourier -- Jacobi expansions, Morrey spaces, weighted inequalities.
Research of the author supported by grant MTM2012-36732-C03-02 of the DGI.
1
2
A. ARENAS AND ´O. CIAURRI
Our target in this paper is the study of the convergence of the Fourier -- Jacobi
expansions in Morrey spaces. To this end, for 1 ≤ p < ∞ and 0 ≤ λ < 1 we define
the Morrey space Lp,λ(−1, 1) as the set of functions f on (−1, 1) such that
kf kp,λ :=
x∈(−1,1),r>0 1
sup
rλ ZB(x,r)
f (t)p dt!1/p
< ∞,
where B(x, r) = {t ∈ (−1, 1) : t − x ≤ r}. It is clear that Lp,λ(−1, 1) are Banach
spaces. Morrey spaces can be defined in a more general way but this is enough for
our purposes. The Lp(−1, 1) spaces with the Lebesgue measure correspond with
the case λ = 0.
Morrey spaces were introduced by Morrey, see [9], in the setting of partial differ-
ential equations. In the last years, some classical operators from harmonic analysis
have been analyzed in the setting of Morrey spaces, see, for instance, [1], [13], [15],
[16], and the references therein.
Specially relevant for our purposes will be the boundedness with weights of the
Hilbert transform given in [15, Theorem 4.7 and Corollary 4.8]. In particular, we
will use the following version of that result: if
and
Hf (x) =Z 1
−1
f (t)
x − t
dt
k
wk(x) =
x − xjγj ,
Yj=1
with −1 ≤ x1 < x2 < · · · < xk−1 < xk ≤ 1, then for 1 < p < ∞ and 0 ≤ λ < 1 it
holds that
(1)
kwkHf kp,λ ≤ Ckwkf kp,λ ⇐⇒
λ − 1
p
< γj < 1 +
λ − 1
p
,
j = 1, . . . , k.
Our result about convergence of Fourier-Jacobi expansions on Morrey spaces is
the following.
Theorem 1. Let 0 ≤ λ < 3/4, 1 < p < ∞, and α, β ≥ 0. Then
if and only if
Snf −→ f,
in Lp,λ(−1, 1),
4
3
(1 − λ) < p < 4(1 − λ).
As a first step to prove Theorem 1 we start by establishing the equivalence
between the convergence of the partial sums and the uniform boundedness of the
operator Sn in Morrey spaces.
Theorem 2. Let 0 ≤ λ < 1, 1 ≤ p < ∞ and α, β > −1. Then
Snf −→ f,
in Lp,λ(−1, 1),
if and only if
where C is a constant independent of n and f .
kSnf kp,λ ≤ Ckf kp,λ,
FOURIER -- JACOBI EXPANSIONS IN MORREY SPACES
3
So, Theorem 1 will follow from the uniform boundedness of the partial sum
operator in Morrey spaces. The next theorem contains a characterization of the
interval of values of p for which this estimate holds.
Theorem 3. Let 0 ≤ λ < 3/4, 1 < p < ∞, and α, β ≥ 0. Then
(2)
kSnf kp,λ ≤ Ckf kp,λ,
where C is a constant independent of n and f , if and only if
(3)
4
3
(1 − λ) < p < 4(1 − λ).
The region where the partial sum Sn converges in Morrey spaces is the shadowed
one shown in Figure 1. Outside and in the border of that region the convergence
is not possible. As it occurs in Lp spaces, we think that in the border, with
dashed lines in the figure and corresponding with the values p = 4(1 − λ) and
p = max{4(1−λ)/3, 1}, some kind of weak boundedness of the partial sum operator
should hold.
p
4
4
3
1
p
=
4
(
1
−
λ
)
p
= 4
3(1
− λ)
0
1
4
3
4
λ
1
Figure 1: The region where the partial sum operator converges.
The proofs of Theorem 2 and Theorem 3 will be contained in Sections 2 and 3,
respectively. In the last section, we include the proof of some auxiliary results.
2. Proof of Theorem 2
Theorem 2 is a standard consequence of Banach-Steinhaus Theorem and the
following proposition.
Proposition 4. Let 1 ≤ p < ∞, 0 ≤ λ < 1, and α, β > −1. Then, span{p(α,β)
dense in Lp,λ(−1, 1).
n
} is
Proof. The case λ = 0 is well known and our proof for 0 < λ < 1 relies on it.
Let us suppose that span{p(α,β)
} is not dense in Lp,λ(−1, 1), for 0 < λ < 1.
Then, by a standard consequence of Hahn-Banach Theorem [14, Theorem 5.19],
n
4
A. ARENAS AND ´O. CIAURRI
there exists a non-zero functional T on Lp,λ(−1, 1) such that T p(α,β)
n ≥ 0.
n
= 0, for all
It is easy to check that for p ≤ r and λ ≤ µ, we have
Lr,µ(−1, 1) ⊂ Lp,λ(−1, 1) ⊂ Lp,0(−1, 1) = Lp(−1, 1).
So, by Hahn-Banach Theorem [14, Theorem 5.16], we can extend the functional T to
Lp(−1, 1). Then, there exists a unique function g ∈ Lq(−1, 1), with p−1 + q−1 = 1,
such that
T f =Z 1
−1
f (x)g(x) dx,
for each f ∈ Lp(−1, 1).
But the condition 0 = T p(α,β)
this is a contradiction because T was non-zero.
implies a(α,β)
n
n
(g) = 0 and g = 0. Then T = 0 and
(cid:3)
Proof of Theorem 2. By the Banach-Steinhaus Theorem [14, Theorem 5.8], the con-
vergence implies the uniform boundedness.
On the other hand, for each f ∈ Lp,λ(−1, 1) and given ε > 0, by Proposition
} such that kf − gkp,λ < ε. Moreover there exists
4, there exists g ∈ span{p(α,β)
N > 0 such that Sn(g) = g, for each n ≥ N . Then
n
kSnf − f kp,λ ≤ kSnf − Sngkp,λ + kg − f kp,λ
≤ (C + 1)kg − f kp,λ ≤ (C + 1)ε,
for n ≥ N and Snf −→ f in Lp,λ(−1, 1).
(cid:3)
3. Proof of Theorem 3
Before starting with the proof we are going to collect some facts that we will
use.
Given a function g defined in (−1, 1), for each 1 ≤ p < ∞ and 0 < λ < 1, we
define
kgk∗
q,λ := inf
x∈(−1,1)Z ∞
0
rλ/p−1kχ(B(x,r))cgkLq(−1,1) dr,
where (B(x, r))c = (−1, 1) \ B(x, r) and p−1 + q−1 = 1.
With the previous definition we can give an appropriate version of Holder in-
equality for Morrey spaces (see [3, Lemma 4.1]). In our case it reads so.
Lemma 1. Let 1 ≤ p < ∞, p−1 + q−1 = 1, and 0 < λ < 1. Then the inequality
(4)
Z 1
−1
f (x)g(x) dx ≤ Ckf kp,λkgk∗
q,λ,
holds with a constant C independent of f and g.
It is interesting to know some facts about the norm of certain functions in Morrey
spaces. From [16, Remark 4.4] we see that, for a ∈ (−1, 1) and νp > −1, the
function x − aνχB(x0,r0) ∈ Lp,λ(−1, 1) if and only if νp ≥ λ − 1, with λ > 0, and
(5)
kx − aνχB(x0,r0)kp,λ ≃ B(x0, r0)(1+νp−λ)/p.
With respect to the functions p(α,β)
n
we prove the following result.
FOURIER -- JACOBI EXPANSIONS IN MORREY SPACES
5
Lemma 2. Let 1 ≤ p < ∞ and α, β ≥ 0. Then
(6)
kp(α,β)
n
for 0 < λ ≤ 3/4, and
(7)
kp,λ ≃(1,
kp,0 ≃
n
kp(α,β)
n1/2−2(1−λ)/p,
1,
(log n)1/4,
n1/2−2/p,
if p ≤ 4(1 − λ),
if p > 4(1 − λ),
if p < 4,
if p = 4,
if p > 4.
The estimates in the previous lemma will be deduced by using a very sharp bound
for the Jacobi polynomials arising from a Hilb type formula for Jacobi polynomials
(see [17, Theorem 8.21.12]). In fact, we have
n
n1/2P (α,β)
(x) ≤ C(1 − x + n−2)−α/2−1/4(1 + x + n−2)−β/2−1/4,
(8)
for α, β > −1. Then, for α, β ≥ 0, we deduce in an easy way (note that d(α,β)
n1/2) the bounds
n
∼
(9)
where h±,n(x) = (1 ± x + n−2)−1/4, and
n
p(α,β)
(x) ≤ C(h+,n(x) + h−,n(x)),
(10)
for α, β ≥ 0.
p(α,β)
n
(x) ≤ C(1 − x2)−1/4,
To complete the boundedness of Sn, we will use Lemma 2 and to that end we
q,λ for the functions h±,n. That is the content of the next
have to estimate kgk∗
lemma.
Lemma 3. Let 1 ≤ p < ∞ and α, β > 0. Then kh±,nk∗
q,λ ≤ C, for p ≥ 4(1 − λ)/3.
Moreover, to analyze the norm of some functions in Morrey spaces, we have to
recall the following theorem from [7].
Theorem 5. Let dµ = w(x) dx be a measure on (−1, 1). If {pn}n≥0 is the sequence
of ortonormal polynomials on L2((−1, 1), dµ), then
(cid:16)Z 1
−1(cid:12)(cid:12)(cid:12)
g(x)
w(x)1/2(1 − x2)1/4(cid:12)(cid:12)(cid:12)
p
w(x) dx(cid:17)1/p
for 0 < p ≤ ∞ and for each measurable function g on (−1, 1).
≤ C lim inf
n→∞ (cid:16)Z 1
−1
pn(x)g(x)pw(x) dx(cid:17)1/p
,
As an immediate consequence of the previous result, we have
(11)
k(1 − x2)−1/4g(x)kp,λ ≤ C lim inf
n→∞
kp(α,β)
n
(x)g(x)kp,λ.
The last tool that we need to complete the proof of the theorem is an easy
observation related to the boundedness of the Hilbert transform with weights that
will be used to complete the necessity of the conditions (3).
Lemma 4. Let H be the Hilbert transform on (−1, 1) and let us suppose that
kuHgkp,λ ≤ Ckvgkp,λ,
vg ∈ Lp,λ(−1, 1),
for two nonnegative weights u and v. Then for r, s ∈ (−1, 1) such that r ≤ s we
have the inequality
(12)
(cid:18)Z r
−1
f (t)
s − t
dt(cid:19) kuχ(r,s)kp,λ ≤ Ckvf χ(−1,r)kp,λ,
vf ∈ Lp,λ(−1, 1),
6
A. ARENAS AND ´O. CIAURRI
with the same constant C.
The proofs of Lemma 2, Lemma 3, and Lemma 4 will be done in the last section.
Proof of Theorem 3. Let us start by assuming the conditions (3) to prove the uni-
form boundedness of the partial sum operators Sn. They can be written as
−1
Snf (x) =Z 1
Xk=0
n
Kn(x, t) =
Kn(x, t)f (t) dt,
p(α,β)
k
(x)p(α,β)
k
(t).
Now, we consider Pollard decomposition of the kernel, see [11],
Kn(x, t) = anT1(n, x, t) + anT2(n, x, t) + bnT3(n, x, t),
where
T1(n, x, t) = p(α,β)
n+1 (x)
T2(n, x, t) = T1(n, t, x) and T3(n, x, t) = p(α,β)
an and bn are bounded. This leads to
(1 − t2)1/2p(α+1,β+1)
n
(t)
,
x − t
n+1 (x)p(α,β)
n+1 (t). Moreover the sequences
where
with
and
Snf (x) = anW1,nf (x) − anW2,nf (x) + bnW3,nf (x),
W1,nf (x) = p(α,β)
W2,nf (x) = (1 − x2)1/2p(α+1,β+1)
n+1 (x)H((1 − (·)2)1/2p(α+1,β+1)
(x)H(p(α,β)
n
n+1
f )(x),
n
f )(x),
W3,nf (x) = p(α,β)
p(α,β)
n+1 (t)f (t) dt.
n+1 (x)Z 1
−1
Now, taking into account the estimate (10), the boundedness in Lp,λ(−1, 1) of
W1,n and W2,n will follow from the inequalities
k(1 − x2)±1/4Hg(x))kp,λ ≤ Ck(1 − x2)±1/4g(x))kp,λ.
Then, by using (1), it is enough that
λ − 1
p
<
−1
4
< 1 +
λ − 1
p
and
λ − 1
p
<
1
4
< 1 +
λ − 1
p
,
and this is implied by (3).
To treat W3,n we start by using (4), then
kW3,nf kp,λ ≤ kp(α,β)
p(α,β)
n+1 (t)f (t) dt ≤ kp(α,β)
n+1 kp,λkp(α,β)
n+1 k∗
q,λkf kp,λ.
n+1 kp,λZ 1
−1
Now, from (8), we have kp(α,β)
the estimate it is enough to apply Lemma 2 and Lemma 3. Indeed,
q,λ ≤ C(kh−,n+1k∗
q,λ+kh+,n+1k∗
n+1 k∗
q,λ). So, to conclude
kW3,nf kp,λ ≤ Ckp(α,β)
n+1 kp,λ(kh−,n+1k∗
q,λ + kh+,n+1k∗
q,λ)kf kp,λ ≤ Ckf kp,λ.
Let us show that conditions (3) are necessary for the uniform boundedness. We
consider the operators Tnf (x) = Snf (x) − Sn−1f (x), then it is clear that
Tnf (x) = a(α,β)
n
(f )p(α,β)
n
(x).
FOURIER -- JACOBI EXPANSIONS IN MORREY SPACES
7
By using the uniform boundedness for Sn, we have
(13)
kTnf kp,λ = a(α,β)
n
(f )kp(α,β)
n
kp,λ ≤ Ckf kp,λ,
with a constant C independent of n and f .
Taking f = sgn(p(α,β)
n
), (13) becomes
kp(α,β)
n
k1,0kp(α,β)
n
kp,λ ≤ C.
So, by using (7), the inequality kp(α,β)
implies, by (6), p ≤ 4(1 − λ).
n
kp,λ ≤ C has to be verified and this fact
Now, by considering f = sgn(p(α,β)
n
)p(α,β)
n
q−1, with p−1 + q−1 = 1, in (13), we
have
kp(α,β)
n
kq
q,0kp(α,β)
n
kp,λkp(α,β)
n
k−q/p
q,λ ≤ C.
Analyzing the different cases of the previous inequality for p ≤ 4(1 − λ) with the
estimates in Lemma 2, we deduce the restriction p ≥ 4(1 − λ)/3.
To conclude the necessity of (3), we have to check that for the cases p = 4(1−λ)/3
and p = 4(1 − λ) the inequality (2) is not possible.
For p = 4(1 − λ) the operators W2,n and W3,n are uniformly bounded. Let us
see that W1,n is unbounded. We proceed by contradiction. If W1,n is bounded, we
have the equivalent estimate
kp(α,β)
n+1 Hgkp,λ ≤ Ck(1 − (·)2)−1/2(p(α+1,β+1)
Now, from (12) with s = 1 and f (t) = χ(0,r)(t)p(α+1,β+1)
we have
n
n
)−1gkp,λ.
(t)(1 − t2)1/4, for r > 0,
p(α+1,β+1)
n
(t)(1 − t2)1/4
1 − t
dtkp(α,β)
n
χ(r,1)kp,λ ≤ Ck(1 − t2)−1/4χ(0,r)kp,λ.
Z r
0
By using (11), we obtain the inequality
Z r
0
dt
1 − t
k(1 − t)−1/4χ(r,1)kp,λ ≤ Ck(1 − t)−1/4χ(0,r)kp,λ,
which, by (5), is equivalent to − log(1 − r) ≤ C and this is impossible.
In the case p = 4(1 − λ)/3, W1,n and W3,n are bounded. Let us suppose that
W2,n is also bounded. Thus, we have the equivalent inequality
k(1 − (·))1/2p(α+1,β+1)
n+1
Hgkp,λ ≤ k(p(α,β)
n
)−1gkp,λ.
Then we can consider (12) with s = 1, r = 1/2 and f (t) = (p(α,β)
obtain
n
(t))4(1 − t), to
(14)
Z 1/2
−1
(p(α,β)
n
(t))4 dtk(1 − (·)2)1/2p(α+1,β+1)
n
χ(1/2,1)kp,λ
≤ Ck(1 − t)(p(α,β)
n
)3χ(−1,1/2)kp,λ.
It is clear that
k(1 − t)(p(α,β)
n
)3χ(−1,1/2)kp,λ ≤ Ck(1 + t)1/4χ(−1,1/2)kp,λ ≤ C,
where in the last step we have used (5). Following the estimate in [6] to evaluate
the Lp-norm of Jacobi polynomials, we have
Z 1/2
−1
(p(α,β)
n
(t))4 dt ∼ log n.
8
A. ARENAS AND ´O. CIAURRI
Finally, we observe that, for any ε > 0 and n big enough,
k(1 − (·)2)1/2p(α+1,β+1)
χ(1/2,1)kp,λ ≥ lim inf
k(1 − (·)2)1/2p(α+1,β+1)
n
n
n→
χ(1/2,1)kp,λ − ε
by using again (5). In this way, (14) implies log n ≤ C and this is not possible. (cid:3)
≥ k(1 − (·)2)1/4χ(1/2,1)kp,λ − ε ≥ C,
4. Proof of Lemmas
Proof of Lemma 2. The case λ = 0 is well known and it can be done, with the
proper modifications, following the ideas in [6, Proposition 1], so we omit it. From
(9), it will be enough to study the norms kh±,nkp,λ. We will focus on h−,n, the
other can be analyzed in the same way.
We have
kh−,nkp,λ =
sup
x∈(−1,1),r>0(cid:16)r−λnp/2−2I(cid:17)1/p
,
with I =R t
s (z + 1)−p/4 dz, s = (1 − x − r)n2 and t = (1 − x + r)n2. Note that in our
situation we can consider r bounded by a positive value. We will analyze different
cases.
When r ≤ n−2, with the estimate I ≤ C(t − s) we have
r−λnp/2−2I ≤ Cr1−λnp/2 ≤ np/2−2(1−λ).
Now, when r > n−2, we consider the cases s ≤ 1 < t and 1 ≤ s < t (when
s < t ≤ 1 it is verified that r ≤ n−2 and this case has been already treated).
If s ≤ 1 < t and r > n−2, for p ≤ 4(1 − λ), we have I ≤ Ct1−p/4. So, using that
in this case 1 − x ≤ Cr, we obtain
r−λnp/2−2I ≤ Cr−λ(1 − x + r)1−p/4 ≤ Cr1−λ−p/4 ≤ C.
For p > 4(1 − λ), it is clear that I ≤ CtλR t
s (1 + z)−p/4−λ dz ≤ Ct−λ and
r−λnp/2−2I ≤ Cr−λ(1 − x + r)λnp/2−2(1−λ) ≤ Cnp/2−2(1−λ).
Finally, for 1 ≤ s < t and r < n−2, when p ≤ 4(1 − λ) we can check easily that
I ≤ C(t1−p/4 − s1−p/4) ≃ t−p/4(t − s). Then
r−λnp/2−2I ≤ Cr1−λ(1 − x + r)−p/4 ≤ Cr1−λ−p/4 ≤ C.
For p > 4(1 − λ), we use that I ≤ CtλR t
tλ−1s1−p/4−λ(t − s). In this way
s z−p/4−λ ≃ tλ(s1−p/4−λ − t1−p/4−λ) ≃
r−λnp/2−2I ≤ C(1 − x − r)1−p/4−λ ≤ Cnp/2−2(1−λ),
and the proof of the upper bound is completed.
To obtain the lower estimate in the case p ≤ 4(1 − λ) we use that, for ε > 0 and
n big enough, by (11),
kp(α,β)
n
kp,λ ≥ Ck(1 − t2)−1/4kp,λ − ε.
Then, the required estimate follows by (5). In the case p > 4(1 − λ), we will use
that P (α,β)
n2 < x < 1. This is a well known consequence of the
Hilb type formula for the Jacobi polynomials. So,
(x) ≥ nα, for 1 − 1
n
kp(α,β)
n
kp,λ ≥ C(cid:16)n2λ+p(α+1/2)Z 1
1−1/n2
(1 − x)pα/2 dx(cid:17)1/p
≥ Cn1/2−2(1−λ)/p,
FOURIER -- JACOBI EXPANSIONS IN MORREY SPACES
and the proof is finished.
Proof of Lemma 3. We analyze h−,n, the other case is similar. It is clear that
9
(cid:3)
kh−,nk∗
q,λ ≤
=
x∈(1−1/n2,1)Z ∞
inf
0
inf
(J1 + J2),
x∈(1−1/n2,1)
rλ/p−1kχ(B(x,r))ch−,nkLq(−1,1) dr
where
and
0
J1 =Z 1−x
J2 =Z 1+x
1−x
rλ/p−1kχ(−1,x−r)∪(x+r,1)h−,nkLq(−1,1) dr
rλ/p−1kχ(−1,x−r)h−,nkLq(−1,1) dr.
It is easy to check that kh−,nkLq(−1,1) can be controlled by the right hand side in
(7) and then
J1 ≤ Ckh−,nkLq(−1,1)Z 1/n2
0
rλ/p−1 dr ≤ Cn−2λ/pkh−,nkLq(−1,1) ≤ C,
where in the last step we have used the condition p ≥ 4(1 − λ)/3.
For J2, when q < 4 we have J2 ≤ C because
J2 ≤ Ckh−,nkLq(−1,1)Z 1+x
1−x
In the case q > 4, we have
rλ/p−1 dr ≤ C.
kχ(−1,x−r)h−,nkLq(−1,1) ≤ C((1 + n−2)1/q−1/4 + (1 − x + r + n−2)1/q−1/4)
and
rλ/p−1((1 + n−2)1/q−1/4 + (1 − x + r + n−2)1/q−1/4) dr
1−x
J2 ≤ CZ 1+x
≤ C(cid:18)1 +Z 1+x
≤ C(cid:18)1 +Z 1+x
1−x
1−x
rλ/p−1(1 − x + r + n−2)1/q−1/4) dr(cid:19)
rλ/p+1/q−5/4 dr(cid:19)
≤ C(1 + (1 + x)λ/p+1/q−1/4 + (1 − x)λ/p+1/q−1/4) ≤ C,
where in the last step we have used the condition p ≥ 4(1 − λ)/3. When q = 4,
J2 ≤ CZ 1+x
1−x
r3λ/4−1 log(cid:18)
2n2 + 1
1 + n2(1 − x + r)(cid:19) dr.
Then, after applying integration by parts, we deduce that
J2 ≤ C(cid:18)1 + n2(1 − x)3λ/4 +Z 1+x
≤ C(cid:18)1 + n2(1 − x)3λ/4 +Z 1+x
1−x
1−x
≤ C(1 + n2(1 − x)3λ/4 + (1 − x)3λ/4)
r3λ/4−1 dr(cid:19)
n2r3λ/4
1 + n2(1 − x + r)
dr(cid:19)
and inf x∈(1−1/n2,1)(J1 + J2) ≤ C.
(cid:3)
10
A. ARENAS AND ´O. CIAURRI
Proof of Lemma 4. If g(t) = χ(−1,r)(t)f (t), for a function f such that vf ∈
Lp,λ(−1, 1), for r ≤ r ≤ s, we have
−1
f (t)
x − t
Hg(x) =Z r
dt ≥Z r
Hg(x) ≥ χ(s,r)(x)Z r
−1
f (t)
s − t
dt.
−1
f (t)
s − t
dt
and then
In this way, the boundedness of the Hilbert transform implies (12).
(cid:3)
References
[1] D. R. Adams and J. Xiao, Morrey spaces in harmonic analysis, Ark. Math. 50 (2012), 201 --
230.
[2] S. Chanillo, On the weak behaviour of partial sums of Legendre series, Trans. Amer. Math.
Soc. 268 (1981), 367-376.
[3] A. Gogatishvili and R. Ch. Mustafayev, New pre-dual space of Morrey space, J. Math. Anal.
Appl. 397 (2013), 678-692.
[4] J. J. Guadalupe, M. P´erez, and J. L. Varona, Weighted weak behaviour of Fourier -- Jacobi
series, Math. Nachr. 158 (1992), 161 -- 174.
[5] J. J. Guadalupe, M. P´erez, F. J. Ruiz, and J. L. Varona, Two notes on convergence and
divergence a.e. of Fourier series with respect to some orthogonal systems, Proc. Amer. Math.
Soc. 116 (1992), 457-464.
[6] J. Levesley and A. K. Kushpel, On the norm of the Fourier-Jacobi projection, Numer. Funct.
Anal. Optim. 22 (2001), 941-952.
[7] A. M´at´e, G. P. Nevai, and V. Totik, Neccesary conditions for weighted mean convergence of
Fourier series in orthogonal polynomials, J. Approx. Theory 46 (1986), 314 -- 322.
[8] C. Meaney, Divergent Jacobi polynomial series, Proc. Amer. Math. Soc. 87 (1983), 459-462.
[9] C. B. Morrey, On the solutions of quasi-linear elliptic partial differential equations, Trans.
Amer. Math. Soc. 43 (1938), 126 -- 166.
[10] B. Muckenhoupt, Mean convergence of Jacobi series, Proc. Amer. Math. Soc. 23 (1969),
306 -- 310.
[11] H. Pollard, The mean convergence of orthogonal series II, Trans. Amer. Math. Soc. 63 (1948),
355-367.
[12] H. Pollard, The mean convergence of orthogonal series III, Duke Math. J. 16 (1949), 189-191.
[13] M. Rosenthal and H. Triebel, Calder´on-Zygmund operators in Morrey spaces, Rev. Mat.
Complut., to appear. DOI 10.1007/s13163-013-0125-3
[14] W. Rudin, "Real and complex analysis", 3rd edition, McGraw-Hill International Editions,
Singapore, 1987.
[15] N. Samko, Weighted Hardy and singular operators in Morrey spaces, J. Math. Anal. Appl.
350 (2009), 56-72.
[16] N. Samko, On a Muckenhoupt-Type Condition for Morrey Spaces, Mediterr. J. Math. 10
(2013), 941-951.
[17] G. Szego, "Orthogonal polynomials", 3rd edition, Amer. Math. Soc. Colloq. Publ., vol. 23.
Amer. Math. Soc., Providence, R.I., 1967.
[18] G. M. Wing, The mean convergence of orthogonal series, Amer. J. Math. 72 (1950), 792-808.
Departamento de Matem´aticas y Computaci´on, Universidad de La Rioja, Calle Luis
de Ulloa s/n, 26004 Logrono, Spain
E-mail address: [email protected]
Departamento de Matem´aticas y Computaci´on, Universidad de La Rioja, Calle Luis
de Ulloa s/n, 26004 Logrono, Spain
E-mail address: [email protected]
|
1004.1527 | 1 | 1004 | 2010-04-09T11:01:06 | Slowly Changing Vectors and the Asymptotic Finite-Dimensionality of an Operator Semigroup | [
"math.FA"
] | Let $T:X\to X$ be a linear power bounded operator on Banach space. Let $X_0$ is a subspace of vectors tending to zero under iterating of $T$. We prove that if $X_0$ is not equal to $X$ then there exists $\lambda$ in Sp(T) such that, for every $\epsilon>0$, there is $x$ such that $|Tx-\lambda x|<\epsilon $ but $|T^nx|>1-\epsilon$ for all $n$. The technique we develop enables us to establish that if $X$ is reflexive and there exists a compactum $K$ in $X$ such that for every norm-one $x\in X$ $\rho\{T^nx, K\}<\alpha (T)<1$ for some $n=n_1, n_2,...$ then $codim(X_0)<\infty$. The results hold also for a one-parameter semigroup. | math.FA | math | SLOWLY CHANGING VECTORS AND THE ASYMPTOTIC
FINITE-DIMENSIONALITY OF AN OPERATOR SEMIGROUP
K. V. STOROZHUK
Abstract. Let X be a Banach space and let T : X → X be a linear power
bounded operator. Put X0 = {x ∈ X T nx → 0}. We prove that if X0 6= X
then there exists λ ∈ Sp(T ) such that, for every ε > 0, there is x such that
kT x − λxk < ε but kT nxk > 1 − ε for all n. The technique we develop enables
us to establish that if X is reflexive and there exists a compactum K ⊂ X
such that lim inf n→∞ ρ{T nx, K} < α(T ) < 1 for every norm-one x ∈ X then
codim X0 < ∞. The results hold also for a one-parameter semigroup.
0
1
0
2
r
p
A
9
]
.
A
F
h
t
a
m
[
1
v
7
2
5
1
.
4
0
0
1
:
v
i
X
r
a
1. Introduction
In this article, X is a complex Banach space, T : X → X is a linear operator
whose all powers are bounded by a constant C < ∞. We use the notations: BX is
the unit ball in X, Γ is the unit circle in C, X0 = {x ∈ X T nx →n→∞ 0}.
A vector x is called an ε-almost eigenvector (or simply an ε-eigenvector) if there
exists λ ∈ C such that kT x − λxk < ε. These vectors exist for each λ ∈ Sp(T ) ∩ Γ.
In Section 1, we study ε-eigenvectors that do not shorten much under the iterations
T n.
Definition 1. Suppose that ε > 0. Call a vector x ∈ X ε-slow if
∃λ ∈ Γ kT x − λxk < ε
and kT nxk > 1 − ε ∀n = 0, 1, 2 . . . .
For example, the eigenvectors x, T x = λx, λ ∈ Γ, are slow.
Example 1. T : l2 → l2 is the right shift, T (x1, x2, . . . ) = (0, x1, x2, . . . ). This
is an isometry; hence, every norm-one ε-eigenvector is ε-slow.
Example 1∗. T : l2 → l2 is the left shift, T (x1, x2, . . . ) = (x2, x3, . . . ). The
spectrum includes Γ but T nx → 0 for every x and there are no slow vectors.
If a vector x is ε-slow then T nx are Cε-slow for each n, since T (T nx) − λT nx =
T n(T x − λx). Thus, the angle between the (complex) lines T nx and T n+1x is slow
not only for n = 0 but for all n ∈ N (i.e., the vector changes slowly under the
iterations).
Remark. Our terminology is by no means connected with the terms "slow
vector" and "slow variable" of the classical theory of dynamical systems. In the
title, we call slow vectors slowly changing.
Definition 2. An operator T has slow vectors if, for every ε > 0, there exist
ε-slow vectors. An operator T has many slow vectors if dim X = ∞ and, for every
ε > 0 and n < ∞, there exist n-dimensional subspaces in X whose unit spheres
consist of ε-slow vectors.
If the powers kT nk are bounded below, i.e., there exists a number c such that
ckxk ≤ kT nxk for every x, then there are many slow vectors. Indeed, if kxk = 1
and x is a cε-eigenvector then x
c is ε-slow.
1
2
K.V. Storozhuk
If X0 = X then it is clear that there are no slow vectors.
It turns out that
the condition X0 = X is the only obstacle to the existence of slow vectors; if
codim X0 = ∞ then there are many slow vectors (Theorem 1.1).
In Section 2, slow vectors are used in the study of the asymptotic properties of
T n.
It is known that if there exists an attracting compact set K, i.e., such that
∀x ∈ BX lim
n→∞
ρ(T nx, K) = 0
then X = X0 ⊕ L, where L is a finite-dimensional invariant subspace in X. This
was proved in [1] for Markov semigroups in L1. For an arbitrary Banach space, this
was established in [2, 3]. In [4] it was proved that for the splitting X = X0 ⊕ L,
dim L < ∞, it suffices that a compact set K attract only sometimes, i.e.,
∀x ∈ BX lim inf
n→∞
ρ(T nx, K) = 0.
A semigroup T n : X → X or a one-parameter semigroup Tt : X → X is called
asymptotically finite-dimensional [5] if codim X0 < ∞. In [6, 1.3.33] the question is
posed whether a semigroup is asymptotically finite-dimensional if
∀x ∈ BX lim inf
n→∞
ρ(T nx, K) < α(T ) < 1.
(∗)
Clearly, the condition (∗) follows from the above-listed analogous conditions.
In Section 2 of the article, we prove that if X is reflexive then an operator
satisfying (∗) has few slow vectors and codim X0 < ∞ (Theorem 2.3). Thus, we give
a partial positive answer to the question of
[6]. Theorem 2.3 is easily generalized
to the case of a one-parameter semigroup of {Tt : X → X, t ≥ 0}.
Note that it is in the reflexive case that the condition codim X0 < ∞ for a
bounded semigroup implies the splitting X0 ⊕ L [7].
For nonreflexive X, the author does not know the answer to the question of [6,
1.3.33].
2. Slow Vectors
Theorem 1.1. Suppose that X is a Banach space, T : X → X, kT nk < C. If
X0 6= X then T has slow vectors. If codim X0 = ∞ then T has many slow vectors.
Proof. Without loss of generality, by passing to the equivalent norm kxk :=
supn{kT nxk}, we may assume that kT k ≤ 1.
The scheme of the proof is as follows: (a) X0 = 0 ⇒ slow vectors exist; (b)
X0 = 0 ⇒ there are many slow vectors; (c) codim X0 = ∞ ⇒ there are many slow
vectors.
(a) Introduce the norm kxkp := lim
n→∞
kT nxk on X. In this norm, T is an isometry.
The norms k kp and k k need not be equivalent but
kT kxk ∼k→∞ kT kxkp
(1.1)
for all x.
If the powers of T are not bounded below then the space (X, kkp) is
incomplete. Let bX be the completion of X in the norm k kp. Extend the isometry
T of (X, k kp) to bX and denote the extension by bT . Take λ ∈ Sp(bT ) ∩ Γ. Suppose
that ε > 0 and bx ∈ bX is a k kp-one ε-eigenvector corresponding to λ. Involving the
fact that bT : bX → bX is an isometry, we get
kbTbx − λbxkp < ε and ∀n kbT nbxkp = kbxkp = 1 > 1 − ε.
Slowly Changing Vectors
3
The set X is dense in bX.
all strict inequalities of the last formula remain valid. Thus, there exists a vector
x ∈ X such that
If a vector x ∈ X is sufficiently k kp-close to bx then
kT x − λxkp < ε and ∀n kT nxkp > 1 − ε.
(1.2)
Thus, x is an ε-slow vector of T in the norm k kp. Of course, x need not be slow in
the initial norm, since kT x−λxk may be large. However, applying the k kp-isometry
T k to (1.2), we infer
∀k kT (T kx) − λT kxkp < ε
and ∀n kT n(T kx)kp > 1 − ε.
(1.3)
By (1.1), starting from some k = k0, inequalities (1.3) also hold for the norm
k k, i.e., starting from some k, the vector T kx is as well slow in the initial norm
of X.
(b) Take a number λ ∈ Γ in the spectrum of the isometry bT of (X, k kp); λ has
many ε-eigenvectors (there exist even infinite-dimensional spheres of ε-eigenvectors
(see [8, Chapter IV, Theorems 5.33, 5.9]).
Let l < ∞ and let W be an l-dimensional subspace in (X, k kp) whose k kp-unit
sphere S consists of ε-eigenvectors. Perturbing W slightly, we may assume that
W ⊂ X. All vectors in S satisfy (1.2) and (1.3). By (1.1), for each x ∈ S, all
the vectors T k0x are slow vectors for the operator T : X → X starting from some
k0. The ellipsoid S is compact; therefore, k0 may chosen common for all x ∈ S.
Thus, X includes (l − 1)-dimensional ellipsoids of the form T k0(S) consisting of
small vectors.
(c) Consider the quotient space X/X0.
Its elements are [x] := x + X0. The
norm k[x]k is as follows: k[x]k = ρ(x, X0) = inf{kx − x0k, x0 ∈ X0}. We have
T (X0) ⊂ X0; therefore, the operator [T ] : X/X0 → X/X0 is defined. Clearly,
[T ]n = [T n]. It is easy to see that if [x] 6= [0] then [T ]n[x] 6→ 0. By (b), [T ] has
many slow vectors, i.e., for each l, in X/X0, there are l-dimensional ellipsoids of
slow vectors [x] for [T ] (moreover, we may assume that these ellipsoids have the
form [S], where S is an ellipsoid in X):
∀n ≥ 0 k[T ]n[x]k > 1 − ε
and k[T ][x] − λ[x]k < ε ∀[x] ∈ [S].
Since T kx0 → 0 for all x0 ∈ X0, we have
∀n ≥ 0 kT n(T kx)k > 1 − ε and kT (T kx) − λT kxk <k→∞ ε
(1.4)
for every x ∈ [x] = x + X0 ∈ [S].
The compactness of S enables us to assert now that, starting from some k, the
(cid:3)
ellipsoids T kS ⊂ X consist of slow vectors.
In what follows, we will need some properties of slow vectors.
Denote by Sm,λ the operator
Lemma 1.2. Suppose that δ > 0, m ∈ N. If T has slow vectors then there exists
λm(cid:1), the Cesaro mean of T /λ.
m+1(cid:0)Pm
1
T m
i=0
λ ∈ Γ such that
∃x ∈ BX kSm,λx − xk < δ
and ∀n kT n(Sm,λx)k > 1 − δ.
(1.5)
If T has many slow vectors then there exist subspaces W ⊂ X of an arbitrarily
large dimension whose unit spheres S consist of vectors x satisfying the inequality
of (1.5).
Proof. Involving a compactness of Γ, consider λ ∈ Γ to which there correspond
slow vectors. If ε := ε(δ, m) is very small and x is an ε-slow vector corresponding
4
K.V. Storozhuk
to λ then Sm,λ(x) ≈ x and Sm,λ(x) satisfies (1.5). The remaining part of the proof
is obvious.
(cid:3)
Remark. Geometrically, Lemma 1.2 means that, for each m, there exist spheres
of an arbitrarily large dimension that almost do not flatten under the mappings
T n(Sm,λ) for any n.
3. Asymptotic Finite-Dimensionality in the Reflexive Case
Throughout the section, we suppose that T : X → X satisfies (∗). We may
assume that K is a balanced set.
Lemma 2.1. Suppose that x ∈ BX . For each k, there exist vectors a1, . . . , ak ∈
K, numbers m1 > m2 > · · · > mk, and t1, . . . , tk, ti ≤ αi−1, such that
kT m1x − [t1T m2a1 + t2T m3a2 + · · · + tk−1T mk ak−1 + tkak]k ≤ αk.
(2.1)
Proof. We write down some inequalities for k = 1, 2, 3. The first is condition (∗),
and the validity of each subsequent inequality is guaranteed by an application (∗)
to the preceding inequality multiplied by α:
∃n1 kT n1x − t1a1k ≤ α, t1 ≤ 1,
∃n2 kT n2(T n1x − t1a1) − t2a2k ≤ α2, t2 ≤ α,
∃n3 kT n3(T n2(T n1x − t1a1) − t2a2) − t3a3k ≤ α3, t3 ≤ α2, . . . .
To finish, it remains to remove parenthesis and put mj = nj + · · · + nk.
The sum of the numbers ti in (2.1) does not exceed h := Pk
Hence, the convex hull bK of S∞
∀x ∈ BX ∀ε > 0 ∃n ∈ N ∃a ∈ bK kT nx − ak < ε.
i=0 T i(hK) attracts BX , i.e.,
We now show that T cannot act by multiplication by a scalar on the subspaces
(cid:3)
i=1 αi = 1
1−α .
(2.2)
X whose dimension is rather high.
Theorem 2.2. dim ker(T − λI) < ∞ for all λ ∈ Γ.
Proof. Choose a finite (1 − α)-net of k vectors for K and consider the subspace
Y spanned by the net. By the Kreın -- Krasnosel′skiı -- Milman Theorem [9], in every
subspace Z ⊂ X such that dim Z > dim Y , there exists a norm-one vector z such
that ρ(z, Y ) = 1. By (∗), ρ(T nz, Y ) < α + (1 − α) = 1 for some n. Therefore, T z
cannot have the form λ · z.
(cid:3)
Theorem 2.3. Suppose that X is reflexive. They T cannot have many slow
vectors and so codim X0 < ∞.
Proof. It suffices to prove that to no λ ∈ Γ there correspond many slow vectors.
We may assume that λ = 1.
By the Statistical Ergodic Theorem (see, for example, [10, § 2]), the operator
means Sm,1 = Sm = 1
m+1(cid:18) mP0
T k(cid:19) converge to the projection P of X onto ker(I −
T ). By Theorem 2.2, dim ker(I − T ) < ∞.
On a compact set K, the convergence of Sm − P to zero is uniform (for ex-
ample, by Arzel`a's Theorem). The operators Sm commute with T ; therefore, the
convergence (Sm − P ) → 0 is also uniform on bK. Hence, starting from some m,
kSm(a) − P (a)k is sufficiently small for all a ∈ bK, for example, less than 1
(2.2), for every x ∈ B and large n, the vector T nx is close to bK. Therefore,
∃m∀x ∈ BX k(Sm − P )T nxk = kT n(Smx) − P xk ≤n→∞ 1/3.
3 . But by
Slowly Changing Vectors
5
This implies, for example, that, under T n ◦ Sm, every k-dimensional sphere such
that k > dim ker(I − T ), "flattens" three times for large n along some radius x (x
must be chosen so that P x = 0).
This, by Lemma 1.2 and the remark thereto, means that the number λ = 1 does
not have many slow vectors.
The inequality codim X0 < ∞ follows now from Theorem 1.1.
(cid:3)
References
[1] Lasota A., Li T. Y., and Yorke J. A. Asymptotic periodicity of the iterates of Markov oper-
ators.// Trans. Amer. Math. Soc. 1984, v. 286, 2, 751 -- 764.
[2] Phong Vu Quoc. Asymptotic almost-periodicity and compactifying representations of semi-
groups.// Ukrainian Math. J. 1986, v. 38, 6, 576 -- 579.
[3] Sine R. Constricted systems.// Rocky Mountain J. Math. 1991, v. 21, 4, 1373 -- 1383.
[4] Storozhuk K. V. An extension of the Vu-Sine theorem and compact-supercyclicity.// J. Math.
Anal. Appl., 2007, v. 332, 2, 1365 -- 1370.
[5] Emel'yanov E. Yu. and Wolff Manfred P. H. Quasi-constricted linear operators on Banach
spaces.// Studia Math. 2001, v. 144, 2, 169 -- 179.
[6] Emel'yanov E. Yu., Non-Spectral Asymptotic Analysis of One-Parameter Operator Semi-
groups, Birkhauser, Basel (2007) (Oper. Theory, Advances Appl.; 173).
[7] Emel'yanov `E. Yu. Some conditions for a C0-semigroup to be asymptotically finite-
dimensional.// Siberian Math. J. 2003, v. 44, 5, 793 -- 796.
[8] Kato T. Perturbation Theory for Linear Operators. 1995 Springer-Verlag, Berlin etc.
[9] Kreın M. G., Krasnosel ′skiı M. A., and Milman V. D. On the defect numbers of linear
operators in Banach space and on some geometric properties.// Sb. Tr. Inst. Mat. Akad.
Nauk USSR 1948, v. 11, 97 -- 112.
[10] Lyubich Yu. I. Introduction to the Theory of Banach Group Representations. 1985 Vishcha
Shkola (Izdat. pri Khar′kov. Gos. Univ.), Khar′kov (in Russian)
Sobolev Institute of mathematics SB RAS
|
1812.06704 | 1 | 1812 | 2018-12-17T11:24:43 | Exhaustive families of representations of $C^*$-algebras associated to $N$-body Hamiltonians with asymptotically homogeneous interactions | [
"math.FA"
] | We continue the analysis of algebras introduced by Georgescu, Nistor and their coauthors, in order to study $N$-body type Hamiltonians with interactions. More precisely, let $Y$ be a linear subspace of a finite dimensional Euclidean space $X$, and $v_Y$ be a continuous function on $X/Y$ that has uniform homogeneous radial limits at infinity. We consider, in this paper, Hamiltonians of the form $H = - \Delta + \sum_{Y \in S} v_Y$, where the subspaces $Y$ belong to some given family S of subspaces. We prove results on the spectral theory of the Hamiltonian when $S$ is any family of subspaces and extend those results to other operators affiliated to a larger algebra of pseudo-differential operators associated to the action of $X$ introduced by Connes. In addition, we exhibit Fredholm conditions for such elliptic operators. We also note that the algebras we consider answer a question of Melrose and Singer. | math.FA | math |
EXHAUSTIVE FAMILIES OF REPRESENTATIONS OF
C∗-ALGEBRAS ASSOCIATED TO N -BODY HAMILTONIANS
WITH ASYMPTOTICALLY HOMOGENEOUS INTERACTIONS
J´ER´EMY MOUGEL AND NICOLAS PRUDHON
Abstract. We continue the analysis of algebras introduced by Georgescu,
Nistor and their coauthors, in order to study N -body type Hamiltonians with
interactions. More precisely, let Y ⊂ X be a linear subspace of a finite di-
mensional Euclidean space X, and vY be a continuous function on X/Y that
has uniform homogeneous radial limits at infinity. We consider, in this pa-
per, Hamiltonians of the form H = −∆ + PY ∈S vY , where the subspaces
Y ⊂ X belong to some given family S of subspaces. Georgescu and Nistor
have considered the case when S consists of all subspaces Y ⊂ X, and Nis-
tor and the authors considered the case when S is a finite semi lattice and
Georgescu generalized these results to any families. In this paper, we develop
new techniques to prove their results on the spectral theory of the Hamiltonian
to the case where S is any family of subspaces also, and extend those results
to other operators affiliated to a larger algebra of pseudo-differential operators
associated to the action of X introduced by Connes. In addition, we exhibit
Fredholm conditions for such elliptic operators. We also note that the algebras
we consider answer a question of Melrose and Singer.
An new approach in the study of Hamiltonians of N -body type with interactions
that are asymptotically homogeneous at infinity on a finite dimensional Euclidean
space X was initiated by Georgescu and Nistor [3, 5, 6].
For any finite real vector space Z, we let Z denote its spherical compactification. A
function in C(Z) is thus a continuous function on Z that has uniform radial limits
at infinity. Let SZ be the set of half-lines in Z, that is SZ := { a, a ∈ Z, a 6= 0 }
where a := {ra, r > 0}. We identify SZ = Z r Z .
For any subspace Y ⊂ X, πY : X → X/Y denotes the canonical projection. Let
(1)
H = −∆ + XY ∈S
vY ,
where vY ∈ C(X/Y ) is seen as a bounded continuous function on X via the projec-
tion πY : X → X/Y . The sum is over all subspaces Y ⊂ X, Y ∈ S and is assumed
to be uniformly convergent. One of the main results of [5, 9] describe the essential
spectrum of H extending the celebrated HVZ theorem [13]. The goal of this paper
is to explain how these results can be extended to any family of subspaces that
contains {0} and to more general operators using C∗-algebras techniques.
Let S be a family of subspaces of X with 0 ∈ S. We define the commutative sub-
C∗-algebra ES(X) of the commutative C∗-algebra C u
b (X) of bounded uniformly
1
2
J. MOUGEL AND N. PRUDHON
continous functions on X by
(2)
ES(X) = hC(X/Y ) ,
Y ∈ S i ⊂ C u
b (X).
The algebras ES(X) give an answer to a question of Melrose and Singer [8].
Theorem 1. Let n be an integer. Let S n be the semi-lattice of subspaces of X n
generated by S n
i ∪ S n
ij where
S n
i = {(x1, . . . , xn) ∈ X n ; xi = 0}
S n
ij = {(x1, . . . , xn) ∈ X n ; xi = xj}
Then the spectrum ΩS n of ES n (X n) is a compactification of X n satisfying the fol-
lowing properties :
(1) ΩS 1 is the spherical compactification X,
(2) The action of the symmetric group Sn on X n extends continuously to ΩSn ,
(3) The projections pn,k
(x1, . . . , xn) = (xi1 , . . . , xik ) extend
I
continuously to pn,k
: X n → X k, pn,k
: ΩSn → ΩSk ,
I
I
(4) The difference maps δij (x1, . . . , xn) = xi − xj from X n to X extend con-
tinuously to the compactifications.
Actually, the spectrum ΩS n have very strong connection with the space built by
Vasy in [15] and generalized by Kottke in the last section of [7].
The additive group X acts by translation on C u
invariant. So a crossed product C∗-algebra is obtained
b (X) and the subalgebra ES(X) is
ES(X) ⋊ X ,
(3)
which can be regarded as an algebra of operators on L2(X). Thanks to the as-
sumption 0 ∈ S, the algebra C0(X) belongs ES(X). Hence C0(X) ⋊ X is contained
in ES(X) ⋊ X.
It follows from the definition of crossed products algebras that
the C∗-algebra ES(X) ⋊ X is generated by two kinds of operators : multiplication
operators mf associated to functions f ∈ ES(X), and convolution operators
Cφu(x) := ZX
φ(y)u(x − y)dy
with φ ∈ Cc(X), a continuous compactly supported function. An immediate com-
putation shows that mf cφ (resp. cφmf ) is a kernel operator with kernel
(4)
K(x, y) = f (x)φ(y − x) ,
(resp. K(x, y) = f (y)φ(y − x)).
Proposition 2. (i) The subalgebra C0(X) ⋊ X is the algebra K(X) of compact
operators on L2(X).
(ii) For f ∈ C(X) and φ ∈ Cc(X) the commutator [mf , cφ] is compact.
The point (i) is a consequence of equation (4) because the kernel K has compact
support when f does and the result follows by density. Again, thanks to equation
(4), one sees that the commutator is a kernel operator with kernel
K(x, y) = φ(y − x)(f (x) − f (y)) .
Hence, in view of φ ∈ Cc(X), the support of K is contained in a band around
the diagonal. The distance between the border of the band and the diagonal is
bounded. Moreover, K goes to 0 at infinity because f has radial limits. So the
commutator is a limit of Hilbert-Schmidt operators, and hence is compact.
ESSENTIAL SPECTRUM
3
Recall that a self-adjoint operator P on L2(X) is said to be affiliated to a C∗-
algrebra A of bounded operators on L2(X) if for some (and hence any) function
h ∈ C0(R) then h(P ) belongs to A. For example,, it follows from the identity
(H + i)−1 = (−∆ + i)−1(cid:0)1 + V (−∆ + i)−1(cid:1)−1
,
that H is affiliated to ES(X)⋊X. More generally, for any C∗-algebra A, a morphism
h : C0(R) → A is called an operator affiliated to A. Following Connes [2] and Baaj
[1] we introduce the C∗-algebra of non positive order pseudo-differential operators
ΨDO(ES(X), X) together with the symbol map exact sequence
0 → ES(X) ⋊ X → ΨDO(ES (X), X)
σ0−→ C(SX , ES(X)) → 0 .
Positive order pseudo-differential operators are examples of operators affiliated to
the algebra of non positive order pseudo-differential operators ΨDO(ES(X), X).
Let α ∈ SX . For each x ∈ X, we let (Txf )(y) = f (y − x) denote the translation on
L2(X). For any operator P on L2(X), we let
(5)
τα(P ) = lim
r→+∞
T ∗
raP Tra ,
if α = a ∈ SX ,
whenever the strong limit exists.
Lemma 3. For f ∈ C(X/Y ) one has
τα(f )(x) = (cid:26) f (x)
f (πY (α))
if Y ⊃ α ,
else.
We define Sα = {Y ∈ S ; α ⊂ Y }. It follows from the previous lemma that on
ES(X), τα is the projection on the subalgebra ESα(X),
τα : ES(X) → ESα(X) .
Theorem 4.
(1) Let P be a self-adjoint operator affiliated to ΨDO(ES(X), X)
and α = a ∈ SX . Then the limit τα(P ) := limr→+∞ T ∗
raP Tra exists and
Specess(P ) = ∪α∈SX Spec(τα(P )) .
(2) Let P ∈ ΨDO(ES (X), X). Then P is a Fredholm operator if and only if P
is elliptic (i.e. σ0(P ) is invertible) and for all α ∈ SX , τα(P ) is invertible.
This extends theorems of [5, 9] in the following sense : only operators affiliated to
ES(X) ⋊ X are considered there, and the relation is
(6)
Specess(H) = ∪α∈SX Spec(τα(H))
in [5]. In [9] only finite semi-lattice S are considered. The equation (6) means that
the family (τα) is a faithful family of morphism of ES(X) ⋊ X. The stronger result
of [9] is obtained by showing that the family (τα ⋊ X)α∈SX is actually an exhaustive
family of representations of ES(X) ⋊ X, when S is a finite semi-lattice.
In the
framework of admissible locally compact group, decomposition of essential spectrum
involving exhaustive families can be found in [10] [11]. In fact, by [12, Proposition
3.12], exhaustive families are also strictly spectral families in the following sense.
Definition 5. [12, 14]
(1) A family (φi)i∈I of morphisms of a C∗-algebra A is said to be exhaustive
if any primitive ideal contains at least ker φi for some i ∈ I.
4
J. MOUGEL AND N. PRUDHON
(2) A family (φi)i∈I of morphisms of a unital C∗-algebra A is said to be strictly
spectral if
(∀a ∈ A)
Spec(a) = ∪i∈I Spec(φi(a))
Theorem 6. Let S be a family of subspaces of X with 0 ∈ S. Then the family
(τα ⋊ X)α∈SX is an exhaustive family of ES(X) ⋊ X/K(X).
Let us prove this result. Let π be an irreducible representation of ES(X) ⋊X/K(X).
It extends to an irreducible representation of ES(X) ⋊ X as well as to their mul-
tipliers algebras M(ES(X) ⋊ X/K(X)) and M(ES(X) ⋊ X). By proposition 2(i),
one obtains the following commutative diagram:
C(X)
֒→ ES(X)
→ M(ES(X) ⋊ X)
(7)
p
↓
C(SX )
φ
↓
→ M(ES(X) ⋊ X/K(X))
π
→ B(Hπ)
Lemma 7. The image φ(C(SX )) is central in M(ES(X) ⋊ X/K(X)).
In fact it is enough to show that any f ∈ C(X) commutes with any element of
ES(X) ⋊ X modulo a compact operator. But the result is true on the generators
by Proposition 2(ii), so the lemma follows by density.
By the Schur Lemma, we deduce that π ◦ φ is a character of C(SX ). Hence there
exists some α ∈ SX such that πC(X) = χαI, where χα is the character of C(X)
given by the evaluation at α ∈ SX .
Proposition 8. One has ker τα = (ker χα)ES(X).
Proof. We need to show that ES(X)/ ker τα = ESα(X) and ES(X)/(ker χα)ES (X)
have the same characters. By definition, for any character χ of ESα(X), there exists
a unique character χ′ of ES(X) such that χ′ = χ ◦ τα. In view of lemma 3, this is
equivalent to the following :
(8)
(∀Y ∈ S, α 6⊂ Y, ∀u ∈ C(X/Y )) χ(u) = u(πY (α)).
In particular, for Y = 0, we see that χC(X) = χα. Reciprocally it follows from [5,
Lemma 6.7] that if χC(X) = χα then relation (8) is true. On the other hand, the
characters of ES(X)/(ker χα)ES(X) are precisely the characters χ of ES(X) such
that χC(X) = χα. So ker τα = (ker χα)ES(X) as claimed.
(cid:3)
Now if πC(X) = χα, one has ker π ⊃ (ker χα)ES(X) = ker τα. Finally,
ker(τα ⋊ X) = (ker τα) ⋊ X ⊂ ker π .
It follows that (τα ⋊ X)α∈SX is an exhaustive family of morphisms.
Remark 9. The results presented here can easily be extended to pseudo-differential
operators with matrix coefficients. For example, Dirac operators DV = D + V , with
potentials V as in (1) may be considered and satisfy the condition of Theorem 4.
See also [4, Example 6.35] for others physical interesting operators.
ESSENTIAL SPECTRUM
5
Acknowledgments. The authors thank Victor Nistor for useful discussions.
References
[1] S. Baaj. Calcul pseudo-diff´erentiel et produits crois´es de C ∗-alg`ebres. I. C. R. Acad. Sci.
Paris S´er. I Math., 307(11):581 -- 586, 1988.
[2] A. Connes. C ∗ alg`ebres et g´eom´etrie diff´erentielle. C. R. Acad. Sci. Paris S´er. A-B,
290(13):A599 -- A604, 1980.
[3] V. Georgescu. On the structure of the essential spectrum of elliptic operators on metric spaces.
J. Funct. Anal., 260(6):1734 -- 1765, 2011.
[4] V. Georgescu and A. Iftimovici. Localizations at infinity and essential spectrum of quantum
Hamiltonians. I. General theory. Rev. Math. Phys., 18(4):417 -- 483, 2006.
[5] V. Georgescu and V. Nistor. On the essential spectrum of N -body Hamiltonians with asymp-
totically homogeneous interactions. J. Operator Theory, 77(2):333 -- 376, 2017.
[6] Vladimir Georgescu. On the essential spectrum of elliptic differential operators. J. Math.
Anal. Appl., 468(2):839 -- 864, 2018.
[7] C. Kottke. Functorial compactification of linear spaces. 2017. arXiv :1712.03902.
[8] R. Melrose and M. Singer. Scattering configuration spaces. 2008. arXiv :0808.2022.
[9] J. Mougel, V. Nistor, and N. Prudhon. A refined HVZ-theorem for asymptotically homo-
geneous interactions and finitely many collision planes. Rev. Roumaine Math. Pures Appl.,
62(1):287 -- 308, 2017.
[10] M. Mantoiu. C ∗-algebras, dynamical systems at infinity and the essential spectrum of gener-
alized Schrodinger operators. J. Reine Angew. Math., 550:211 -- 229, 2002.
[11] M. Mantoiu. Essential spectrum and Fredholm properties for operators on locally compact
groups. J. Operator Theory, 77(2):481 -- 501, 2017.
[12] V. Nistor and N. Prudhon. Exhaustive families of representations and spectra of pseudodif-
ferential operators. J. Operator Theory, 78(2):247 -- 279, 2017.
[13] M. Reed and B. Simon. Methods of modern mathematical physics. IV. Analysis of operators.
Academic Press [Harcourt Brace Jovanovich Publishers], New York, 1978.
[14] S. Roch. Algebras of approximation sequences: structure of fractal algebras. In Singular
integral operators, factorization and applications, volume 142 of Oper. Theory Adv. Appl.,
pages 287 -- 310. Birkhauser, Basel, 2003.
[15] A. Vasy. Propagation of singularities in many-body scattering. Ann. Sci. ´Ecole Norm. Sup.
(4), 34(3):313 -- 402, 2001.
Universit´e de Lorraine, UFR MIM, 3 rue Augustin Fresnel 57045 METZ, France
E-mail address: [email protected]
Universit´e de Lorraine, UFR MIM, 3 rue Augustin Fresnel 57045 METZ, France
E-mail address: [email protected]
|
1803.05052 | 2 | 1803 | 2018-06-17T16:36:02 | The weighted Property (A) and the greedy algorithm | [
"math.FA"
] | We investigate various aspects of the "weighted" greedy algorithm with respect to a Schauder basis. For a weight w, we describe w-greedy, w-almost-greedy and w-partially-greedy bases, and show some properties of w-semi-greedy bases. To achieve these goals, we introduce and study the w-Property (A). | math.FA | math |
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
Abstract. We investigate various aspects of the "weighted" greedy algorithm with respect
to a Schauder basis. For a weight w, we describe w-greedy, w-almost-greedy, and w-partially-
greedy bases, and examine some properties of w-semi-greedy bases. To achieve these goals, we
introduce and study the w-Property (A).
1. Introduction
i (ej) = 1 if i = j and e∗
n=1, with biorthogonal functionals (e∗
n)∞
n=1; that is,
nk} ≤ supn max{kenk,ke∗
In this paper, we investigate the operation of the "weighted" greedy algorithm, and its
efficiency. Throughout, (X,k · k) is a real Banach space with a semi-normalized Schauder basis
B = (en)∞
A1) 0 < c1 := inf n min{kenk,ke∗
A2) e∗
A3) X = span[ei : i ∈ N],
i=1 aiei) =
i=1 aiei. We denote by Kb the least value of K for which the preceding inequality holds,
A4) kSmk ≤ K for every m, where (Sm)m are partial sum operators -- that is, Sm(P∞
Pm
x =Pj e∗
We will refer to B as a basis. Of course, for every x ∈ X, there exists a unique expansion
i (x) 6= 0}, A denotes the cardinality of a set
j (x)ej. As usual, supp (x) = {i ∈ N : e∗
i (ej) = 0 for i 6= j,
and call it the basis constant.
nk} =: c2 < ∞,
A and
Nm = {A ⊂ N : A = m}, N<∞ =
Nm.
∞[m=0
Further notations will be often used:
if a and b are functions of some variable, a . b means
that there exists a constant c > 0 such that a ≤ c· b; if A and B are subsets of N, A < B means
that maxj∈A j < minj∈B j, PA is the projection operator, i.e, if A is a finite set, PA(Pj ajej) =
Pj∈A ajej and P c
A = I − PA is the complementary projection, 1εA =Pn∈A εnen for εn ∈ {±1}
and if εn ≡ 1, we write 1A.
Algorithm (TGA): for in x ∈ X we produce the sequence of greedy approximands
In 1999, S. V. Konyagin and V. N. Temlyakov introduced in [16] the Thresholding Greedy
Gm(x) =
e∗
π(n)(x)eπ(n),
mXn=1
2000 Mathematics Subject Classification. 46B15, 41A65.
Key words and phrases: thresholding greedy algorithm, unconditional basis, Property (A), w-greedy bases.
The first author was supported by a PhD fellowship FPI-UAM and the grants MTM-2016-76566-P (MINECO,
Spain) and 19368/PI/14 (Fundaci´on S´eneca, Regi´on de Murcia, Spain). The second author was supported
by the National Science Foundation under Grant Number DMS -- 1361461. The second and third authors were
supported by the Workshop in Analysis and Probability at Texas A&M University in 2017.
1
2
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
π(i)(x) ≥ e∗
where π is a greedy ordering, that is, π : {1, 2, ...,supp x} −→ supp x is a bijection such
that e∗
k(x)ek,
where Am(x) = {π(n) : n ≤ m} is a greedy set of x: inf k∈Am(x) e∗
k(x).
positive constant C such that
π(j)(x) for i ≤ j. Alternatively we can write Gm(x) = Pk∈Am(x) e∗
Also, they defined in [16] the quasi-greedy bases as those bases such that there exists a
k(x) ≥ supk /∈Am(x) e∗
kGm(x)k ≤ Ckxk, ∀x ∈ X,∀m ∈ N.
(1)
P. Wojtaszczyk proved in [18] that a basis is quasi-greedy if and only if the (TGA) converges
-- that is,
Of course, (1) is equivalent to the existence of a constant C ′ such that
m→∞kx − Gm(x)k = 0, ∀x ∈ X.
lim
kx − Gm(x)k ≤ C ′kxk, ∀x ∈ X,∀m ∈ N.
(2)
We denoted by Cq the least constant that satisfies (2), it is called the quasi-greedy constant
and we say that B is Cq-quasi-greedy.
On the other hand, the (TGA) is a good candidate to obtain the best m-term approximation
with regard to B. In this sense, S. V. Konyagin and V. N. Temlyakov defined in [16] the greedy
bases as those bases such that there exists a constant C ≥ 1 such that
anenk : A ⊂ N,A = m, an ∈ R}.
(3)
kx − Gm(x)k ≤ C inf{kx −Xn∈A
Furthermore, they showed that B is greedy if and only if B is democratic (that is, k1Ak . k1Bk,
for all A ≤ B) and unconditional.
Some years later, G. Kerkyacharian, D. Picard and V. N. Temlyakov [15] introduced the
following extension of the greedy bases: we consider a weight w = (wi)∞
i=1 ∈ (0,∞)N. If A ⊂ N,
w(A) =Pi∈A wi denote the w-measure of A. We define the error σw
δ (x) as
anenk : A ∈ N<∞, w(A) ≤ δ, an ∈ R}.
δ (x,B)X = σw
σw
δ (x) := inf{kx −Xn∈A
kx − Gm(x)k ≤ Cσw
Definition 1.1. We say that B is w-greedy if there exists a constant C ≥ 1 such that
w(Am(x))(x), ∀x ∈ X,∀m ∈ N.
(4)
We denote by Cg the least constant that satisfies (4) and we say that B is Cg-w-greedy.
Roughly, the greedy bases are those where the greedy approximation is "as effective as m-
term approximation can possibly be".
This generalization was motivated by the work of A. Cohen, R. A. DeVore and R. Hochmuth
in [8]. In their recent paper [6], the first author and ´O. Blasco characterize w-greedy bases
using the best m-term error in the approximation "with polynomials of constant coefficients".
Moreover, [17] characterizes w-greedy bases in terms of their w-democracy and unconditionality.
Definition 1.2. We say that B is w-democratic if there exists a constant C ≥ 1 such that
(5)
for any pair of sets A, B ∈ N<∞ with w(A) ≤ w(B). We denote by Cd the least constant that
satisfies (5) and we say that B is Cd-w-democratic.
k1Ak ≤ Ck1Bk,
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
3
Recall that a basis B in X is unconditional if any rearrangement ofPn e∗
n(x)en converges
in norm to x for any x ∈ X. This is equivalent to the uniform boundedness of basis projections:
(6)
kx − PA(x)k ≤ Kkxk, ∀x ∈ X,∀A ⊂ N.
We denote by Ku the least constant that satisfies (6), it is called the (suppression) unconditional
constant, and we say that B is Ku-(suppression) unconditional.
Other important w-type greedy basis in this context is the w-almost-greedy basis.
Definition 1.3. We say that B is w-almost-greedy if there exists a constant C ≥ 1 such that
(7)
kx − Gm(x)k ≤ C σw
w(Am(x)), ∀x ∈ X,∀m ∈ N,
where
δ (x,B)X = σw
σw
δ (x) := inf{kx − PA(x)k : A ∈ N<∞, w(A) ≤ δ}.
We denote by Cal the least constant that satisfies (7) and we say that B is Cal-w-almost-greedy.
Remark 1.4. If w ≡ 1, that is, w(A) = A, we recover the classical definition of almost-
greediness (resp. greediness, democracy and Property (A)-see the definition below), and we will
say that B is almost-greedy (resp. greedy, democratic, has the Property (A)).
In the classical sense, that is, when w ≡ 1, S. J. Dilworth, N. J. Kalton, D. Kutzarova and
V. N. Temlyakov gave in [12] a characterization of almost-greedy bases in terms of the quasi-
greediness and democracy. Recently, S. J. Dilworth, D. Kutzarova, V. N. Temalykov and B.
Wallis, in [13], gave a characterization of w-almost-greedy bases in terms of quasi-greedy and
w-democratic bases.
It is well known that, even for w ≡ 1, the w-democracy and unconditionality (resp. quasi-
greediness), cannot be used to determine whether a given basis is w-greedy (resp. w-almost-
greedy) with constant 1. For the weight w ≡ 1, F. Albiac and P. Wojtaszczyk introduced in
[4] the so called Property (A) (defined below) in order to obtain finer estimate for the greedy
constant Cg (and, in particular, to characterize bases with Cg = 1). The results of [4] were
further generalized in [11]; in [2], the Property (A) was used to estimate the almost-greedy
constant Cal.
Throughout the paper, we will be using a weighted version of Property (A):
Definition 1.5. We say that B satisfies the w-Property (A) if there exists a constant C ≥ 1
such that
kx + t1εAk ≤ Ckx + t1ηBk,
(8)
for any x ∈ X, for any A, B ∈ N<∞such that w(A) ≤ w(B), A∩ B = ∅, supp (x)∩ (A∪ B) = ∅,
for any ε, η ∈ {±1} and t ≥ supj e∗
j (x). We denote by Ca the least constant that satisfies (8)
and we say that B has the Ca-w-Property (A).
Remark 1.6. The definition of w-Property (A) was motivated by the "classical" Property (A)
(introduced in [4]), which states that
kx + t1εAk ≤ Ckx + t1ηBk,
whenever A = B < ∞, A ∩ B = ∅, supp (x) ∩ (A ∪ B) = ∅, ε, η ∈ {±1} and t ≥ supj e∗
j (x).
Proposition 3.2 shows that the classical Property (A) is equivalent to the w-Property (A) if
0 < inf n wn ≤ supn wn < ∞.
4
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
Another way of estimating the efficiency of greedy approximation is to compare the rate
of convergence with straightforward Schauder approximation. To this end we consider w-
partially-greedy bases.
In [12], the authors defined the partially-greedy bases as those
satisfying
kx − Gm(x)k ≤ Ckx − Sm(x)k, ∀x ∈ X,∀m ∈ N,
for some positive and absolute constant C. Moreover, they proved that B is partially-greedy if
and only if B is quasi-greedy and conservative (that is, k1Ak . k1Bk for all pair of finite sets
A, B such that A < B and A ≤ B). Here, we present the notion of w-partially-greedy bases
and we characterize these bases using w-conservative bases.
The paper is structured as follows. In Section 2 we describe the w-greedy and w-almost-
greedy bases in terms of their other properties (such as w-Property (A), unconditionality, or
being quasi-greedy). The main results are Theorems 2.1 and 2.2.
In Section 3, we collect basic facts about the w-Property (A). In addition, we consider the w-
semi-greedy bases -- that is, the bases where the Chebyshev greedy approximands are optimal.
It turns out (Theorem 3.7) such bases necessarily possess the w-Property (A).
Section 4 is devoted to properties (C) and (D), which arise naturally in the study of quasi-
greedy bases. In particular, it is shown that w-superdemocracy and Property (C) imply w-
Property (A) (Proposition 4.2). However, superdemocracy does not imply Property (C) (Ex-
ample 4.8). Further, we show that any w-semi-greedy basis has Property (C) if the weight w
is equivalent to a constant (Proposition 4.10).
In Section 5, we compare the efficiency of greedy approximation with that of the canonical
basis projections. This gives rise to the notion of an w-partially-greedy basis; such bases are
characterized in Theorem 5.7.
Finally, in Section 6 and Section 7 we state some open questions related to our results, and
prove some basic lemmas used throughout the paper.
We freely use the standard "greedy" terminology. The reader can consult e.g. [17] for more
information.
2. Characterization of w-greedy and w-almost-greedy bases
In this section we describe the w-(almost)-greediness of a basis in terms of its w-Property
(A) and unconditionality (resp. quasi-greediness). The corresponding results for the constant
weight w ≡ 1 can be found, for instance, in [17].
Theorem 2.1. Let B be a basis of a Banach space X.
a) If B is Cg-w-greedy, then the basis is Ku-unconditional and has the Ca-w-Property (A)
b) If B is Ku-unconditional and has the Ca-w-Property (A), then the basis is Cg-w-greedy
with constants Ku ≤ Cg and Ca ≤ Cg.
with Cg ≤ KuCa.
Theorem 2.2. Let B be a basis of a Banach space X.
a) If B is Cal-w-almost-greedy, then the basis is Cq-quasi-greedy and has the Ca-w-Property
b) If B is Cq-quasi-greedy and has the Ca-w-Property (A), then the basis is Cal-w-almost-
(A) with constants Cq ≤ Cal and Ca ≤ Cal.
greedy with Cal ≤ CqCa.
Later (Proposition 5.11) we will see examples of bases with w-Property (A) for a certain
weight w, but failing the "classical" Property (A).
For further use, we need the following reformulation of the w-Property (A) (inspired by [2]).
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
5
Proposition 2.3. A basis B has the Ca-w-Property (A) if and only if
kxk ≤ Cakx − PA(x) + 1ηBk,
(9)
j (x) ≤ 1, A, B ∈ N<∞, w(A) ≤ w(B), B ∩ supp (x) = ∅ and
for any x ∈ X with supj e∗
η ∈ {±1}.
The proof requires a technical result.
j (y).
j (x) = maxj e∗
Lemma 2.4. Suppose D is a finite subset of N, and x ∈ X\{0} satisfies supp (x) ∩ D = ∅.
Then for any ε > 0 there exists a finitely supported y ∈ X, so that kx−yk < ε, supp (y)∩D = ∅,
and maxj e∗
Proof. It suffices to consider ε < 1/(2c2). By scaling, we can assume that maxj e∗
(then kxk ≥ 1/c2). Clearly PD(x) = 0, and P c
D(x) = x. Now set δ = ε/(3c2
span [ej : j ∈ N] is dense in X, there exists a finitely supported z ∈ X so that kx−zk < δ/kP c
Let u = P c
C = maxj e∗
j (x) = 1
2kxk). As
Dk.
j (x − u) < c2δ, hence
j (y) = 1, and
D(z), then kx − uk = kP c
D(x − z)k < δ. For every j, e∗
j (x) ∈ (1 − c2δ, 1 + c2δ). Now let y = u/C. Then maxj e∗
kx − yk ≤ kx − uk + 1 − C −1kuk < δ +
(kxk + δ) < ε. (cid:3)
c2δ
1 − c2δ
Proof of Proposition 2.3. By Lemma 2.4, it suffices to restrict our attention to finitely supported
vectors x ∈ X only. So, throughout this proof, we assume supp (x) < ∞.
proposition with supj e∗
B, we obtain
Suppose that B has the Ca-w-Property (A), and x, A, B, ε, η are as in the statement of the
j (x) ≤ 1. Applying the definition of w-Property (A) to PAcx, A, and
kPAc(x) + 1εAk ≤ CakPAc(x) + 1ηBk = Cakx − PA(x) + 1ηBk.
To finish the proof, observe that x belongs to the convex hull of the set(cid:8)PAc(x) + 1εA}ε∈{±1}.
Now, suppose (9), and prove that the basis B has the w-Property (A) with the same constant.
j (x) ≤ 1, A, B ∈ N<∞ such that w(A) ≤ w(B), A ∩ B = ∅, supp (x) ∩
Take x ∈ X and supj e∗
(A ∪ B) = ∅ and ε, η ∈ {±1}. Define x′ = x + 1εA. Using (9),
kx + 1εAk = kx′k ≤ Cakx′ − PA(x′) + 1ηBk = Cakx + 1ηBk. (cid:3)
Proof of Theorem 2.1: Assume that B is Cg-w-greedy.
where
Unconditionality: Let x ∈ X and A ⊂ supp (x). Define y := PAc(x) +Pn∈A(α + e∗
n(x))en,
α > sup
j∈A e∗
j (x) + sup
j∈Ac e∗
j (x).
As A is a greedy set of y,
kx − PA(x)k = ky − PA(y)k ≤ Cgσw
w(A)(y) ≤ Cgky − α1Ak = Cgkxk.
w-Property (A): Fix x ∈ X, take t ≥ supn e∗
Thus, the basis is unconditional with constant Ku ≤ Cg.
n(x). Consider ε, η ∈ {±1} and finite sets A, B
such that A∩ B = ∅, w(A) ≤ w(B), and (A∪ B)∩ supp (x) = ∅. Set y := x + t1εA + (t + δ)1ηB
with δ > 0. Hence,
kx + t1εAk = ky − GB(y)k ≤ Cgσw
w(B)(y) ≤ Cgky − t1εAk = Cgkx + (t + δ)1ηBk.
Taking δ → 0, we obtain that the basis satisfies the w-Property (A) with constant Ca ≤ Cg.
6
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
w-greedy.
Next we prove that if B is Ku-unconditional and has the Ca-w-Property (A), then it is
Take x ∈ X and suppose that A is a greedy set of cardinality m for x ∈ X -- that is,
w(A)(x) + ε, with supp (y) = B
j (x)}, using the the
PA(x) = Gm(x). For ε > 0 find y ∈ X such that kx − yk < σw
and w(B) ≤ w(A). Then, taking t := min{e∗
reformulation of the w-Property (A) and Lemma 7.1, we obtain that
j (x) : j ∈ A} and η ≡ sgn{e∗
kx − Gm(x)k ≤ Cakx − PA(x) − PB\A(x) + t1η(A\B)k = CakP(A∪B)c(x − y) + t1η(A\B)k
= CakTt(I − PB)(x)k = CakTt(I − PB)(x − y)k ≤ KuCakx − yk.
Consequently, for any greedy set A we have kx − PAxk ≤ KuCaσw
Proof of Theorem 2.2. Assume that B is Cal-w-almost-greedy.
Quasi-greedy: Since
w(A)(x).
(cid:3)
kx − Gm(x)k ≤ Cal inf{kx −Xn∈B
n(x)enk : w(B) ≤ w(Am(x)), B ∈ N<∞},
e∗
we can select B = ∅. Then, we obtain that kx − Gm(x)k ≤ Calkxk, hence the basis is quasi-
greedy with constant Cq ≤ Cal.
w-Property (A): We can use the same argument as in Theorem 2.1.
Now, we will prove that if B is Cq-quasi-greedy and has the Ca-w-Property (A), then it is
For x ∈ X, let A be a greedy set of cardinality m. For ε > 0, find B such that kx− PB(x)k <
σw
w(A)(x) + ε, with w(B) ≤ w(A). Then, taking t := min{e∗
j (x)},
using the reformulation of the w-Property (A) and Lemma 7.1,
j (x) : j ∈ A} and η ≡ sgn{e∗
w-almost-greedy.
kx − Gm(x)k ≤ CakP(A∪B)c(x − y) + t1η(A\B)k
= CakTt(I − PB)(x)k ≤ CqCakx − PB(x)k.
This gives that, for any greedy set A, kx − PA(x)k ≤ CqCaσw(A)(x) as desired.
(cid:3)
Remark 2.5. In this paper, we focus on the situation when B is a Schauder basis. However, the
w-Property (A) can be defined for any complete biorthogonal system satisfying the conditions
(A1)-(A3); the proof of Proposition 2.3 goes through as well.
Moreover, in the definition of the w-Property (A), it suffices to show that (8) holds for with
maxj e∗
j (x) = t. More specifically, the following four statements are equivalent:
B ∩ supp (x) = ∅, w(A) ≤ w(B) and t ≥ supj e∗
w(B), A ∩ B = ∅, supp (x) ∩ (A ∪ B) = ∅, ε, η ∈ {±1} and s = supj e∗
B ∩ supp (x) = ∅, w(A) ≤ w(B) and s = supj e∗
(a) B satisfies the w-Property (A) (see Definition 1.5).
(b) There exists a constant C so that kxk ≤ Ckx− PA(x) + t1ηBk for any η ∈ {±1}, x ∈ X,
j (x).
(c) There exists a constant C ′ so that kx + s1εAk ≤ Ckx + s1ηBk for any x ∈ X, w(A) ≤
(d) There exists a constant C ′ so that kxk ≤ C ′kx−PA(x)+s1ηBk for any x ∈ X, η ∈ {±1},
Indeed, the implications (a) ⇒ (c) and (b) ⇒ (d) (with C ′ = C) are immediate. The equiv-
alence (a) ⇔ (b) (with the same constant C) has been established in Proposition 2.3. Minor
adjustments to that argument give us (c) ⇔ (d).
j (x). As before, we can
assume that x is finitely supported. Find k so that e∗
j (x). By replacing x by
To establish (d) ⇒ (b), take x, A, B, η as in (b) and t ≥ supj e∗
k(x) = supj e∗
j (x).
j (x).
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
7
−x if necessary, we can assume s = e∗
k(x) ≥ 0. Let c = t − s, and consider
x′ = x + cek = Xj∈supp (x)\{k}
e∗
j (x)ej + tek.
Note that kx − x′k ≤ cc2 ≤ tc2. Furthermore, x′ − PA(x′) equals either x − PA(x) (if k ∈ A),
or x − PA(x) + cek (if k /∈ A). In either case,
kx − PA(x) + t1ηBk ≥ kx′ − PA(x′) + t1ηBk − tc2.
By (d), we have kx′k ≤ C ′kx′ − PA(x′) + t1ηBk. By the above,
As kx − PA(x) + t1ηBk ≥ tc−1
kxk − tc2 ≤ C ′(kx − PA(x) + t1ηBk + tc2)
2 , we conclude that kxk ≤ (C ′ + 2c2
2)kx − PA(x) + t1ηBk.
3. Some remarks on the w-Property (A)
Definition 3.1. Let v = (vn)∞
to w, written v ≈ w, whenever there exist positive real constants 0 < a ≤ b < ∞ satisfying
n=1 be weights. We say that v is equivalent
n=1 and w = (wn)∞
avn ≤ wn ≤ bvn
for all n ∈ N.
Proposition 3.2. Let v, w weights and suppose that v ≈ w. Then every basis with the w-
Property (A) also has the v-Property (A).
Proof. Let x ∈ X with supp (x) < ∞ and supj e∗
v(B), A ∩ B = ∅, supp (x) ∩ (A ∪ B) = ∅ and ε, η ∈ {±1}. We set
j (x) ≤ 1, A and B finite satisfying v(A) ≤
Observe that
which gives us
Γ = {n ∈ A : wn ≥ w(B)} .
w(A) ≤ b · v(A) ≤ b · v(B) ≤
b
a · w(B),
w(A) ≥ w(Γ) ≥ Γ · w(B) ≥ Γ ·
a
b · w(A),
and hence Γ ≤ b/a. Next, we give the following partition of A \ Γ: A1 < . . . < Am, so that
for each i = 1, . . . , m, the set Ai is a maximal such that w(Ai) ≤ w(B). Due to maximality,
w(B) < w(Ai) + w(Ai+1) for all i = 1, . . . , m − 1.
Thus,
(m − 1) · w(B) <
m−1Xi=1
This gives us
[w(Ai) + w(Ai+1)] < 2 · w(A \ Γ) ≤ 2 · w(A) ≤
2b
a · w(B).
m ≤
2b
a
+ 1.
8
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
Hence, using the bounds of Γ, m and the condition of the w-Property (A),
kx + 1εAk ≤ k1Γk + kx +
mXi=1
1εAik ≤Xn∈Γ
kenk +
x
m
k
+ 1εAik
mXi=1
c2
2b
a kx + 1ηBk + Cakx + m1ηBk
x
m
+ 1ηBk ≤
≤ c2
2Γkx + 1ηBk + Camk
c2
2b
≤
a kx + 1ηBk + Camkx + 1ηBk + Ca(m − 1)kxk
c2
2b
a kx + 1ηBk + Camkx + 1ηBk + C 2
≤
2b + 2bC 2
a
≤ (cid:18)c2
a
(cid:19)kx + 1ηBk.
a(m − 1)kx + 1ηBk
(cid:3)
Remark 3.3. In a similar fashion, one can show that, if the weights w and v are equivalent,
then any w-democratic (w-superdemocratic, w-conservative -- for the definitions, see below)
basis is also v-democratic (resp. v-superdemocratic or v-conservative).
Remark 3.4. The converse to Proposition 3.2 does not hold in general. For example, suppose
the weights w, v belong to ℓ1. By [13], the family of w-democratic (or v-democratic) bases
consists precisely of those bases which are equivalent to the canonical basis of c0. However, w
and v need not be equivalent.
The rest of this section is motivated by the recent definition of w-semi-greedy bases introduced
in [13]. To give this notion, we need the Chebyshev Greedy Algorithm: for x ∈ X and m ∈ N, if
Am(x) is the greedy set of x with cardinality m, we define the Chebyshev Greedy Approximand
of order m as any Gm(x) ∈ span{ei : i ∈ Am(x)} such that
kx − Gm(x)k = min{kx − Xn∈Am(x)
kx − Gm(x)k ≤ Cσw
bnenk : bn ∈ R}.
Definition 3.5. We say that B is w-semi-greedy if there exists a constant C ≥ 1 such that
(10)
We denote by Csg the least constant that satisfies (10) and we say that B is Csg-w-semi-greedy.
w(Am(x))(x), ∀x ∈ X,∀m ∈ N.
By [13], any w-semi-greedy basis is w-superdemocratic.
Definition 3.6. We say that B is w-superdemocratic if there exists a constant C ≥ 1 such
that
k1εAk ≤ Ck1ηBk,
(11)
for any A, B ∈ N<∞ with w(A) ≤ w(B) and ε, η ∈ {±1}.
We denote by Cs the least constant that satisfies (11) and we say that B is Cs-superdemocratic.
Here, we show that
w − semi-greedy ⇒ w − Property (A) ⇒ w − superdemocracy.
Theorem 3.7. If a basis B is w-semi-greedy, then B has the w-Property (A).
(12)
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
9
We take ε, η, A, B and x in the conditions of the definition of the w-Property (A). In all of
Proof. Assume that kx − Gm(x)k ≤ Csgσw
w(Am(x)) for any x ∈ N and m ∈ N.
the following cases we consider x ∈ X such that supp (x) < ∞ and supn e∗
n(x) ≤ 1.
Case 1:P∞
Case 1.1: w(B) > lim supn→∞ wn. Since Pn wn = ∞, we can choose E and n0 ∈ N with
min E > max(A ∪ B ∪ supp (x)) and n0 > max E such that
n=1 wn = ∞ and supn wn < ∞.
Set F := E ∪ {n0}. Then, w(E) ≤ w(B) < w(F ) < 2w(B).
w(E) ≤ w(B) < w(E) + wn0 < 2w(B).
We define the element z := x + 1εA + (1 + δ)1F . For any scalar sequence (fn)n∈F , we have
kx + 1εAk ≤ Kbkx + 1εA +Pn∈F fnenk. As the basis B is w-semi-greedy with constant Csg,
and w(A) ≤ w(B) < w(F ), we conclude that
inf
fn kx + 1εA +Xn∈F
fnenk ≤ Csgσw
w(F )(z) ≤ Csgkx + (1 + δ)1Fk.
Consequently, kx + 1εAk ≤ KbCsgkx + (1 + δ)1Fk. Taking δ → 0,
kx + 1εAk ≤ KbCsgkx + 1Fk ≤ KbCsgkx + 1Ek + KbCsgken0k
(13)
≤ KbCsgkx + 1Ek + KbCsgc2 ≤ KbCsg(kx + 1ηBk + k1ηBk + k1Ek) + KbCsgc2.
Now, we set y := 1ηB + (1 + δ)1F . Reasoning as before, we obtain
k1ηBk ≤ Kb inf
Sending δ → 0, we obtain
cn k1ηB +Xn∈F
cnenk ≤ KbCsgσw
w(F )(y) ≤ KbCsgk(1 + δ)1Fk.
k1ηBk ≤ KbCsgk1Fk ≤ KbCsgk1Ek + KbCsgc2.
On the other hand, taking s := x + (1 + δ)1ηB + 1E,
bnen + 1Ek ≤ Csg(Kb + 1)σw
k1Ek ≤ (Kb + 1)kx +Xn∈B
Then, taking δ → 0,
k1Ek ≤ Csg(Kb + 1)kx + 1ηBk.
w(B)(s) ≤ Csg(Kb + 1)kx + (1 + δ)1ηBk.
Finally, using (13), (14) and (15), the basis satisfies the w-Property (A) with constant
K = O(C 3
sgK 3
b c2).
Case 1.2: w(A) ≤ w(B) ≤ lim supn→∞ wn. Using Proposition 3.5 of [13],
max{k1εAk,k1ηBk} ≤ 2KbCsgc2.
Since 1 = e∗
j (x + 1ηB) ≤ ke∗
jkkx + 1ηBk ≤ c2kx + 1ηBk for j ∈ B, then
kx + 1εAk ≤ kx + 1ηBk + k1ηBk + k1εAk
≤ kx + 1ηBk + 4KbCsgc2 ≤ (4KbCsgc2
2 + 1)kx + 1ηBk.
Case 2: IfPn wn < ∞ or supn wn = ∞, using the Proposition 3.5 of [13], B is equivalent to
the canonical basis of c0 and the result is trivial.
(14)
(15)
(16)
(cid:3)
10
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
Proposition 3.8. If B has the Ca-w-Property (A), then B is 2Ca-w-superdemocratic.
Proof. Take A, B ∈ N<∞ with w(A) ≤ w(B), and show that, for any choice of signs, k1ηAk ≤
2Cak1εBk. As in [7, Subsection 4.4], it is enough to prove our inequality for ε ≡ 1 (otherwise,
replace B = {en : n ∈ N} by {εnen : n ∈ N}). Since 1ηA ∈ 2S, where S = {PA′⊂A θA′1A′ :
PA′⊂A θA′ ≤ 1} (see [10, Lemma 6.4]), it suffices to show that
k1A′k ≤ Cak1Bk, ∀A′ ⊂ A.
Take A′ ⊂ A. Obviously, 1A′ = 1A′\B + 1B∩A′. Then, using the w-Property (A),
k1A′k = k1A′\B + 1B∩A′k ≤ Cak1B\A′ + 1B∩A′k = Cak1Bk.
We can apply the w-Property (A) because
w(A′) = w(A′\B)+w(A′∩B) ≤ w(A) ≤ w(B) = w(B\A′)+w(B∩A′) ⇒ w(A′\B) ≤ w(B\A′).
This completes the proof.
(cid:3)
With these results, we have proved the implications (12).
Remark 3.9. If w ≡ 1, we recover the classical definition of semi-greediness (resp. superdemoc-
racy), and we will say that B is semi-greedy (resp. superdemocratic).
Improving [13, Proposition 4.5], we prove that, in certain cases, any w-superdemocratic basis
has to contain a subsequence equivalent to the canonical basis of c0 (c0-basis henceforth), or
even to be equivalent to such basis.
Proposition 3.10. Suppose a basis B = (en)∞
n=1 is Cs-w-superdemocratic.
i) If A ∈ N<∞ and w(A) ≤ lim supn→∞ wn, then maxε∈{±1} k1εAk ≤ c2Cs.
ii) If supn wn = ∞, then B is equivalent to the c0-basis.
iii) If inf n wn = 0, then there exist i1 < i2 < . . . so that the sequence (eik )k∈N is equivalent
to the c0-basis. Moreover, if limn wn = 0, then for any infinite set A ⊂ N we can select
i1, i2, . . . ∈ A with the properties described above.
Proof. i) Find n ∈ N\A so that wn > w(A), then k1εAk ≤ Csk1{n}k ≤ c2Cs.
ii) By (i), k1εAk ≤ c2Cs for all choices of signs, which yields the desired equivalence.
iv) IfPn wn < ∞, then B is equivalent to the c0-basis.
iii) Suppose inf n wn = 0, and find i1 < i2 < . . . so that Pk wik < ∞. By convexity, it
for any finite set E ⊂ {i1, i2, . . .}. To this end, find N ∈ N so that PN
j=1 wj ≥Pk wik < ∞.
suffices the existence of a constant K with the property that the inequality k1εEk ≤ K holds
Let B = {1, . . . , N}\{i1, i2, . . .}, D = {1, . . . , N} ∩ {i1, i2, . . .}, and A = E\D. Note that
B,D ≤ N, hence, for every ε ∈ {−1, 1}N, k1εBk,k1εDk ≤ c2N. Then w(A) ≤ w(B) and
hence k1εAk ≤ Csk1εBk ≤ c2NCs. By the triangle inequality,
k1εEk ≤ k1εAk + k1εDk ≤ c2N(Cs + 1).
If limn wn = 0, then every infinite A ⊂ N contains i1 < i2 < . . . withPk wik < ∞. It remains
to invoke the preceding result.
iv) The proof proceeds as in (iii).
(cid:3)
From this we immediately obtain:
Corollary 3.11. If the weight w is unbounded, then a basis has the w-Property (A) if and only
if it is equivalent to the canonical basis of c0.
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
11
4. Properties (C) and (D)
Properties (C) and (D) (discussed below) naturally arise in the study of quasi-greedy bases.
Definition 4.1. We say that B satisfies the Property (C) if for any x ∈ X, there exists a
positive constant C such that
j∈Λ e∗
min
j (x)k1εΛk ≤ Ckxk,
(17)
for any greedy set Λ of x and ε ∈ {±1}. We denote by Cu the least constant that satisfies (17)
and we say that B has the Property (C) with constant Cu.
It is well known any quasy-greedy basis has Property (C) (see [7, Lemma 2.3]). Generalizing
[7, Lemma 2.2], we prove that any w-superdemocratic basis with the Property (C) has the
w-Property (A).
Proposition 4.2. If B is Cs-w-superdemocratic and satisfies the Property (C) with constant
Cu, then B has the Ca-w-Property (A) with Ca ≤ 3CuCs.
Proof. Take x, A, B, ε, η as in the definition of the w-Property (A) and assume that supj e∗
1. Then,
j (x) ≤
kx + 1εAk ≤ kx + 1ηBk + k1ηBk + k1εAk.
(18)
Using the w-superdemocracy and w(A) ≤ w(B), we obtain that k1εAk ≤ Csk1ηBk. Now, we
only have to estimate k1ηBk. For that, we consider the element y := x + 1ηB. It's clear that
1ηB is a greedy sum for y, so
j∈B e∗
min
j (y)k1ηBk = k1ηBk ≤ Cukyk = Cukx + 1ηBk.
(19)
Then, using (18) and (19),
kx + 1εAk ≤ kx + 1ηBk + 2CsCukx + 1ηBk ≤ 3CsCukx + 1ηBk.
Hence, the basis has the w-Property (A) with constant Ca ≤ 3CuCs.
Example 4.3. We next revisit a "pathological" basis constructed in Section 5.5 of [7] (using
some ideas from [9, Example 4.8]): a basis which has the Property (A), but fails to be quasi-
greedy. The initial proof of the Property (A) was unwieldy. Here we present a streamlined
proof that the basis has the Property (C), and then invoke Proposition 4.2.
(cid:3)
First recall the construction: Dk denote the set of all dyadic intervals I ⊂ [0, 1] with length
1 of all real sequences
I = 2−k, and consider D = ∪k≥0Dk. Now, we consider the space fq
a = (aI )I∈D such that
aIχ(1)
I
< ∞,
q!1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L1
kakf
q
1
XI
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
where χ(1)
I = I−1χI. By [14], the canonical basis {eI}I∈D is unconditional and democratic.
For every N ≥ 1, we shall pick a subset {k1, ..., kN} ⊂ N0 and look at the finite dimensional
space FN consisting of sequences supported in ∪N
j=1Dkj . We order the canonical basis by
∪N
j=1{eI}I∈Dkj
. We also consider in
, so we may as well write their elements as a = (aj)dim FN
j=1
12
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
FN the James norm
k(aj)kJq =
Now, set in FN a new norm
m0=0<m1<...Xk≥0
sup
kak = max{kakf
q
1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xmk<j≤mk+1
,kakJq}.
1/q
q
.
aj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Finally, we consider the Banach space X = ⊕ℓ1FN with B the consecutive union of the natural
bases in FN .
. To show that B
1 is unconditional: take a ∈ X
It's possible to show that B is superdemocratic and k1εAk ≈ A ≈ k1εAkfq
satisfies the Property (C), we use that the canonical basis in fq
and Λ a greedy set of a, then
1
n∈Λ ank1εΛk . min
min
n∈Λ ank1εΛkfq
1
. kakfq
1 ≤ kak.
Hence, the basis satisfies the Property (C). Also, since the basis is superdemocratic, using the
Proposition 4.2, the basis satisfies the Property (A).
Definition 4.4. We say that B is bidemocratic if there exists a constant C ≥ 1 such that
k1εAkk1∗
ηAk∗ ≤ CA, ∀ finite A,∀ε, η ∈ {±1}.
Here, k · k∗ is the norm of X∗, and 1∗
Lemma 4.5. If B is bidemocratic, then B satisfies the Property (C).
Proof. Here, we prove a stronger condition than Property (C). Take x ∈ X and A ⊂ supp (x).
Then, taking η = 1/sgn{e∗
i .
ηA =Pi∈A ηie∗
i (x)},
j (x)k1εAk . min
j∈A e∗
j (x)
j∈A e∗
min
ηAk∗ ≤ Pj∈A e∗
k1∗
A
k1∗
ηAk∗
j (x)
≤
1∗
ηA(x)
ηAk∗ ≤ kxk.
k1∗
(cid:3)
Corollary 4.6. All bidemocratic bases satisfy the "classical" Property (A).
Proof. If a basis B is bidemocratic, then it is superdemocratic. Now combine the preceding
lemma with Proposition 4.2.
(cid:3)
Relaxing the assumptions of Definition 4.1, we consider:
Definition 4.7. We say that B satisfies the Property (D) if there exists a positive constant
C such that
min
n∈A ank1Ak ≤ CkXn∈A
anenk,
for any finite set A and scalars (an)n∈A.
It's clear that if B satisfies the Property (C), then B satisfies the Property (D) as well.
Example 4.8. [Example of superdemocratic basis in a Banach space without the Property
(D)] Let X = ℓ1 ⊕ c0 and k(x, y)k = kxkℓ1 + kyk∞. Let (en)n be the canonical basis in ℓ1 and
(fm)m the canonical basis in c0. We define
E2n−1 =(cid:18)1
2
en, −1
2
fn(cid:19) , E2n =(cid:18) 1
4
en,
3
4
fn(cid:19) , n = 1, 2, ...,
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
13
and consider B = {En}n = {E2n−1, E2n}n. This basis is normalized.
To prove that this basis is superdemocratic, we show the following proposition:
Proposition 4.9. D(m) ≈ d(m) ≈ m, where D(m) := sup{k1εAk : A ≤ m, ε ∈ {±1}} and
d(m) := inf{k1εAk : A ≥ m, ε ∈ {±1}}.
Proof. Of course, d(m) ≤ D(m) ≤ m. We prove that d(m) ≥ 1
finite, we write
8m. To this end, given A ⊂ N
A1 = {k ∈ N : 2k ∈ A and 2k − 1 ∈ A},
A2 = {k ∈ N : 2k ∈ A and 2k − 1 6∈ A},
A3 = {k ∈ N : 2k 6∈ A and 2k − 1 ∈ A}.
Observe that the sets A1, A2, A3 are mutually disjoint, and 2A1 + A2 + A3 = A. For any
choice of signs,
k1εAk = kXk∈A1
= kXk∈A1(cid:18)[
+ Xk∈A2
≥ Xk∈A1
ε2k(
1
4
ε2kE2k + ε2k−1E2k−1 +Xk∈A2
1
4
1
2
ε2k +
ε2k−1]ek, [
ε2kE2k +Xk∈A3
ε2k−1]fk(cid:19)
1
2
ε2k−1E2k−1k
ε2k−1(
3
4
1
4
ε2k −
1
2
ek,−
1
2
+ Xk∈A3
fk)k
1
2
.
1
4
ek,
3
4
fk) +Xk∈A3
ε2k−1 +Xk∈A2
1
2
ε2k +
Therefore,
This finishes the proof.
k1εAk ≥
1
4A1 +
1
4A2 +
1
2A3 ≥
1
8A.
(cid:3)
Back to Example 4.8: to see that the basis does not have the Property (D), take z =
n=1 E2n−1. Then,
PN
n=1 2E2n −PN
NXn=1
Write z =Pi∈A aiEi. Then, mini∈A ai = 1 and
kzk = k
(0, 2fn)k = 2.
kXi∈A
Eik = k
E2n +
E2n−1k = k
NXn=1
NXn=1
(
3
4
NXn=1
en,
1
4
fn)k =
3
4
N +
1
4
.
This shows that the Property (D) fails.
Lemma 4.13 of [13] establishes that, if w is equivalent to the constant, and B is w-semi-greedy,
then B satisfies the Property (D). Here, we improve this result showing that the condition of
being w-semi-greedy implies the Property (C).
Proposition 4.10. Assume that w is equivalent to the constant, and the basis B is w-semi-
greedy. Then B satisfies the Property (C).
Proof. By Theorem 3.7, B has the w-Property (A). By Proposition 3.2, B also has the "classical"
Property (A). This, in turn, implies the Property (C).
(cid:3)
14
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
5. w-Partially-greedy bases
Partially-greedy and conservative bases were introduced in [12], in order to compare the
errors of greedy approximation with those of the canonical approximation relative to Schauder
basis (the "tails" of the basis expansion). In this section we define w-partially-greedy and w-
conservative bases and extend the characterization of partially-greedy bases proved in [12] to
this more general setting.
Definition 5.1. We say that B is w-partially-greedy if for all m and r such that w({1, ..., m}) ≤
w(Ar(x)), there exists a positive constant such that
kx − Gr(x)k ≤ Ck
e∗
i (x)eik.
(20)
∞Xn=m+1
We denote by Cp the least constant that satisfies (20) and we say that B is Cp-w-partially-greedy.
Definition 5.2. We say that B is w-conservative if there exists a positive constant C such
that
k1Ak ≤ Ck1Bk,
(21)
for all pair of A, B ∈ N<∞ such that A < B and w(A) ≤ w(B). We denote by Cc the least
constant that satisfies (21) and we say that B is Cc-w-conservative.
Remark 5.3. If w ≡ 1, we recover the classical definition of partially-greediness (resp. conser-
vativeness), and we will say that B is partially-greedy (resp. conservative).
Remark 5.4. Note that for some choices of weight w, the property of w-conservativeness can
be in some sense trivial. For instance, if w = (2−n)∞
n=1 then every seminormalized basis is
w-conservative. This is because there are no nonempty A, B ∈ N<∞ satisfying both A < B and
w(A) ≤ w(B).
Let us give a simple characterization of weights for which this occurs.
Proposition 5.5. Let w be a weight and set
sw := sup(cid:8)n ∈ N0 : there exist A ∈ Nn and B ∈ N<∞ such that A < B and w(A) ≤ w(B)(cid:9).
Then sw < ∞ if and only if every seminormalized basis is w-conservative.
Proof. (=⇒): Suppose sw < ∞. Let (en)∞
n=1 be a seminormalized basis for a Banach space X,
and select A, B ∈ N<∞ such that A < B and w(A) ≤ w(B). Observe that k1Ak ≤ c2·A ≤ c2sw.
It follows immediately that k1Ak ≤ c2sw ≤ c2
2sw)-w-conservative.
n=1 ⊂ N<∞ and
(Bn)∞
(⇐=): Suppose sw = ∞. Let's inductively construct sequences (An)∞
2swk1Bk. Hence, (en)∞
n=1 is (c2
n=1 ⊂ N<∞ satisfying
A1 < B1 < A2 < B2 < A3 < B3 < . . . ,
and also satisfying An ≥ n and w(An) ≤ w(Bn) for all n ∈ N. Let us begin by selecting
A1 ∈ N<∞ and B1 ∈ N<∞ with A1 = 1, A1 < B1, and w(A1) ≤ w(B1), which is possible
as sw ≥ 1. This is the base case; from now on, we proceed inductively. Since sw = ∞, we
may select bAn+1 ∈ N<∞ and Bn+1 ∈ N<∞ with bAn+1 > n + max Bn, bAn+1 < Bn+1, and
w(bAn+1) < w(Bn+1). Now set An+1 = bAn+1 \ {1, . . . , max Bn} so that we have An+1 > n,
An+1 < Bn+1, and w(An+1) < w(Bn+1). This completes the inductive step, and gives us our
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
15
intertwining sequences with the desired properties. We may now define a norm on c00 via the
rule
k(an)∞
n=1kX = k(an)∞
n=1k∞ ∨ sup
an ∀ (an)∞
n=1 ∈ c00,
k∈N Xn∈Ak
and denote by X the completion of c00 under this norm. It is clear that the standard canon-
ical basis for this space form a normalized 1-unconditional basis. However, it fails to be w-
conservative as k1AkkX = Ak ≥ k whereas k1BkkX = 1 for all k ∈ N.
(cid:3)
Proposition 5.6. Let w be a nonincreasing weight, i.e., wn+1 ≤ wn for all n ∈ N. Then every
conservative basis in a Banach space is w-conservative with the same constant.
Proof. Let (en)∞
satisfying both A < B and w(A) ≤ w(B). Now,
n=1 be a conservative basis in a Banach space X, and select any A, B ∈ N<∞
A · wmax A ≤ w(A) ≤ w(B) ≤ B · wmin B ≤ B · wmax A,
so that A ≤ B.
Theorem 5.7. A basis B is w-partially-greedy if and only if B is quasi-greedy and w-conservative.
Proof. Assume that B is Cp-w-partially-greedy.
(cid:3)
(1) w-conservative: take A and B such that A < B and w(A) ≤ w(B). Let m = max A and
define the set D = [1, .., m] \ A. Of course,
w({1, ..., m}) = w(A ∪ D) ≤ w(B ∪ D).
Define now x := 1A + (1 + δ)1B∪D. Then,
k1Ak = kx − GB∪D(x)k ≤ Cpk(1 + δ)1Bk.
Taking δ → 0, the basis is w-conservative.
(2) Quasi-greedy: here, we consider two cases.
1(x))e1, with t = maxe∗
e∗
a) Assume that the index 1 6∈ Ar(x). Define then x = te1 +P∞
hence x − Gr(x) =P∞
i (x) + δ with δ > 0. Then
Gr(x) = te1 + Gr−1(x),
i (x)ei − Gr−1(x). Thus, using the triangle inequality and
the fact that w({1}) ≤ w(Ar(x)),
i (x)ei = x + (t −
i=2 e∗
i=2 e∗
kGr−1(x)k ≤ kx − Gr(x)k + k
e∗
i (x)eik ≤ Cpk
∞Xi=2
∞Xi=2
e∗
i (x)eik + k
∞Xi=2
e∗
i (x)eik
≤ (Cp + 1)(1 + Kb)kxk.
That's implies that kGr(x)k ≤ ((Cp + 1)(Kb + 1) + c2
b) Assume now that 1 ∈ Ar(x). Taking the same x that in the above case,
2)kxk.
Gr(x) = Gr−1(x − e∗
1(x)e1) + te1,
16
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
i=2 e∗
i (x)ei − Gr−1(x − e∗
1(x)e1). Hence, using the same argument
than before,
so x − Gr(x) =P∞
kGr−1(x − e∗
1(x)e1)k ≤ kx − Gr(x)k + k
e∗
i (x)eik
∞Xi=2
∞Xi=2
e∗
≤ Cpk
i (x)eik + k
≤ (Cp + 1)(1 + Kb)kxk.
1(x)e1) + e∗
∞Xi=2
e∗
i (x)eik
Now, kGr−1(x − e∗
1(x)e1k = kGr(x) −
e∗
1(x)e1k. So, using the triangle inequality, we obtain that kGr(x)k ≤ ((Cp +1)(Kb +
1) + c2
1(x)e1)k = kGr−1(x − e∗
1(x)e1 − e∗
2)kxk.
Now, assume that B is Cc-w-conservative and Cq-quasi-greedy, and show that B is w-partially-
greedy. Take x ∈ X, m, and r as in the definition of w-partially-greedy, and consider the sets
D := {ρ(j) : j ≤ r, ρ(j) ≤ m}, B := {ρ(j) : j ≤ r, ρ(j) > m}, A := [1, ..., m] \ D,
where ρ is the greedy ordering. Then Ar(x) = B ∪ D, and w(A) = w({1, ..., m}) − w(D) ≤
w(Ar(x)) − w(D) = w(B).
x − Gr(x) =
e∗
i (x)ei − PB(x) + PA(x).
∞Xi=m+1
i (x)eik. On the other hand, using Lemmas 7.1 and
7.2 with η ≡ sgn (e∗
On the one hand, kPB(x)k ≤ 2CqkP∞
A e∗
kPA(x)k ≤ 4CqCc max
j (x)),
i=m+1 e∗
i (x)k1ηBk ≤ 4CqCc min
B e∗
i (x)k1ηBk
q CckPB(x)k ≤ 8C 3
q Cck
∞Xi=m+1
e∗
i (x)eik.
≤ 8C 2
q CckP∞
i=m+1 e∗
i (x)eik.
Then, kx − Gr(x)k . C 3
Remark 5.8. Note that if the inequality kx − Gr(x)k ≤ Ckx − Sm(x)k is satisfied for m and
r, then it is automatically satisfied -- with a different constant -- for any n < m and the same
r (since Ckx − Sm(x)k ≤ (1 + Kb)Ckx − Sn(x)k where Kb is the basis constant). So we only
need to check the condition in the definition of w-partially-greedy for the largest m satisfying
w([1, . . . , m]) ≤ w(Ar(x)).
Using the constant weight w ≡ 1, we recover the usual definition of a partially-greedy basis.
Indeed, for w ≡ 1, the largest m satisfying the definition is m = r, which recaptures the original
definition of partially-greedy given in [12].
(cid:3)
5.1. Example of conservative and not democratic basis. Define the set
√min A}.
Observe that S has the spreading property, i.e, if m ∈ N, (fi)m
fi ≤ gi for all i = 1, . . . , n, then (gi)m
B ∈ S.
S = {A ∈ N<∞ : A ≤
i=1 ∈ Nm with
i=1 ∈ S. It also hereditary, i.e., if A ∈ S and B ⊂ A then
i=1 ∈ S and (gi)m
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
17
Now, let X be the Banach space that we define like the completion of c00 under the norm
k(an)nk = sup
A∈SXn∈A
an.
Observe that this is a very slight modification of the Schreier space.
Of course, the canonical basis (en)n is a normalized 1-unconditional basis. Note that the
hereditary property guarantees that
k1Ak = sup
F ∈S,F ⊆AF.
Now, if A < B and A ≤ B, then there is F ∈ S with F ⊆ A such that k1Ak = F. By
the spreading property, we can "push out" F to obtain a set G ⊆ B such that G ∈ S and
G = F. Hence,
k1Ak = F = G ≤ k1Bk.
Thus, the basis is conservative with constant 1.
To prove that the basis is not democratic, we can select the sets A = {N 2 + 1, ..., N 2 + N}
and B = {1, ..., N} . Then, since A ∈ S, k1Ak = N. However, k1Bk ≤ √N, hence the basis is
not democratic: to prove this upper estimate, take a set A1 ∈ S such that k1Bk = A1. Then,
min A1 ≤ N, so A1 ≤ √N . Hence, k1Bk ≤ √N .
Remark 5.9. Of course, since the canonical basis is unconditional (hence, quasi-greedy) and
conservative, is partially-greedy, but not almost-greedy because is not democratic.
5.2. Example of a w-greedy basis which is not conservative (hence not greedy).
Definition 5.10. Fix 1 ≤ p < q ≤ ∞, and consider (en)∞
bases of ℓp and ℓq (or c0 if q = ∞). Let w = (wn)∞
space Xq,p,w as the closed subspace [fn ⊕ wnen]∞
basis of c0, we can define Xq,s,w similarly as the subspace [fn ⊕ wnsn]∞
n=1 the respective canonical
n=1 ∈ (0,∞)N. We define the Rosenthal-Woo
n=1 of ℓq ⊕∞ ℓp. For s = (sn)∞
n=1, the summing
n=1 and (fn)∞
n=1 of ℓq ⊕∞ c0.
It was mentioned in [13] that if w ∈ (0,∞)N satisfies w ∈ c0 \ ℓ1 then the basis formed by
completing c00 under the norm
k(an)∞
n=1k∞ ∨ ∞Xn=1
an2wn!1/2
,
(an)∞
n=1 ∈ c00,
forms a normalized 1-w-greedy basis which is not greedy. In fact, this is just the canonical basis
of the Rosenthal-Woo space X∞,2,w1/2. More generally, we have the following.
Proposition 5.11. Fix 1 ≤ p < ∞ and w ∈ (0,∞)N ∩ (c0 \ ℓ1). Then the canonical basis of
X∞,p,w1/p is 1-w-greedy, but it is not conservative.
Proof. Clearly it is unconditional with constant 1. To prove that the canonical basis is w-
greedy with constant 1, we need to show that it satisfies the w-Property (A) with constant 1
(Theorem 2.1). For that, take x ∈ X
j (x) ≤ 1, and consider A, B ⊂ N such
that A ∩ B = ∅, supp (x) ∩ (A ∪ B) = ∅, w(A) ≤ w(B). Then, if ε and η are arbitrary choice
∞,p,w1/p with supj e∗
18
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
of signs,
kx + 1εAk = 1 ∨ Xn∈supp (x)
≤ 1 ∨ Xn∈supp (x)
e∗
e∗
n(x)pwn + w(A)
n(x)pwn + w(B)
1/p
1/p
= kx + 1ηBk.
Then, the basis satisfies the w-Property (A) with constant 1, hence, using that the basis is
unconditional with constant 1, the basis is w-greedy with constant 1.
To see that it fails to be conservative, fix m ∈ N and set Am = {1, . . . , m} and Bm,k =
{k + 1, . . . , k + m} for each k ∈ N. Now observe that w ∈ c0 ensures that w(Bm,k) → 0 when
k → ∞ and hence k1Bm,kk∞,p,w1/p = 1 for sufficiently large k. Hence, we may select km ∈ N
so that Bm,km > Am and k1Bm,kmk∞,p,w1/p = 1. On the other hand, w /∈ ℓ1 guarantees that
w(Am) → ∞.
(cid:3)
Remark 5.12. The above proof works for X∞,s,w as well, except that the basis is no longer
unconditional. It yields an example of a subspace of c0 with a basis which is w-democratic but
not conservative, so long as w ∈ (0,∞)N ∩ (c0 \ ℓ1).
However, the situation is different for the canonical basis of Xq,p,w1/p when q 6= ∞. These
spaces fail to contain any copies of c0, and hence do not admit v-greedy bases for v non-
seminormalized. Even when v is seminormalized the canonical basis may not be v-democratic
(nor v-greedy), as we will see momentarily.
Proposition 5.13. Fix 1 ≤ p < q < ∞, and let w ∈ (0,∞)N be decreasing. Then the canonical
basis of Xq,p,w1/p is unconditional and w-conservative with constants 1.
Proof. Select A, B ∈ N<∞ with w(A) ≤ w(B) and A < B. Since w is decreasing, we must have
A ≤ B. Thus,
k1Akq,p,w1/p = (A)1/q ∨ w(A)1/p ≤ (B)1/q ∨ w(B)1/p = k1Bkq,p,w1/p.
(cid:3)
Proposition 5.14. Fix 1 ≤ p < q < ∞ and 0 < θ < 1 − p
n=1 ∈ (0,∞)N be
defined by wn = n−θ for n ∈ N. Then the canonical basis for Xq,p,w1/p is not conservative, and
not w-democratic.
q . Let w = (wn)∞
Proof. First establish that our basis is not conservative. As in the proof of Proposition 5.11,
for k, m ∈ N we set Am = {1, . . . , m} and Bm,k = {k + 1, . . . , k + m}, and for each m ∈ N we
find km ∈ N large enough that Am < Bm,km and k1Bm,kmkq,p,w1/p = m1/q. Meanwhile,
so that, due to 1 − θ − p/q > 0,
k1Amkq,p,w1/p ≥ w(Am)1/p = mXn=1
1 − θ (cid:19)1/p
k1Bm,kmkq,p,w1/p ≥(cid:18)m1−θ − 1
k1Amkq,p,w1/p
≥(cid:18)Z m
t−θ dt(cid:19)1/p
n−θ!1/p
m−1/q =(cid:18) m1−θ−p/q − m−p/q
1 − θ
=(cid:18)m1−θ − 1
1 − θ (cid:19)1/p
(cid:19)1/p
→ ∞.
1
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
19
We next sketch the proof of the lack of w-democracy. To this end, consider the sets An and
Bm,k as defined in the preceding paragraph. There exist universal constants c and C so that
w(An) ≥ cn1−θ, and k1Ank ≤ Cn1−θ. Then w(Bm,k) < mk−θ, while k1Bm,kk ≥ m1/q. For large
values of k, select m ∈ [cn1−θkθ/2, cn1−θkθ]. Then w(Bm,k) ≤ w(An), yet the ratio
k1Bm,kk
k1Ank ≈
m1/q
n1−θ ≈
n(1−θ)/qkθ/q
n1−θ
= kθ/qn−(1−θ)(1−1/q)
can be arbitrarily large (for large k), ruling out the possibility of w-democracy.
(cid:3)
Corollary 5.15. Fix 1 ≤ p < q < ∞ and 0 < θ < 1 − p
n=1 ∈ (0,∞)N be
defined by wn = n−θ for n ∈ N. Then the canonical basis for Xq,p,w1/p is not v-democratic for
any weight v ∈ (0,∞)N.
Proof. Suppose, for the sake of contradiction, that the canonical basis for Xq,p,w1/p is v-democratic.
Note that this basis contains no subsequences equivalent to the c0-basis. Proposition 3.10 shows
that 0 < inf vn ≤ sup vn < ∞ -- that is, the weight v is equivalent to a constant. Then, by
Remark 3.3, the canonical basis for Xq,p,w1/p has to be democratic, hence conservative. This,
however, contradicts Proposition 5.14.
(cid:3)
q . Let w = (wn)∞
Remark 5.16. Consider again the weight w from Proposition 5.14.
• It follows from Propositions 5.13 and 5.14 that the canonical basis of Xq,p,w1/p is w-
partially-greedy basis, but not w-almost-greedy. However, the space Xq,p,w1/p does have
Indeed, we recall from [3, Theorem 10.7.1] that if X has a
an almost-greedy basis.
complemented subspace with a symmetric basis and finite cotype then X admits an
almost-greedy basis. If w ∈ (0,∞)N ∩ (c0 \ ℓ(pq)/(q−p)) then the Woo-Rosenthal spaces
Xq,p,w contain complemented copies of ℓp and ℓq (or c0 if q = ∞; see [20, Corollary 3.2]),
and hence satisfy this condition.
• Just as in Proposition 5.14, one can show that the canonical basis of Xq,s,w is not
conservative when 0 < θ < 1− 1
q . However, it is not quasi-greedy, either. For θ = 1− 1
q ,
this basis becomes quasi-greedy and democratic (for q = 2, this was observed in [3,
Example 10.2.9], the argument is valid for all 1 < q < ∞).
6. Questions
• Does Property (D) imply Property (C)?
• Does Property (A) imply Property (D)?
• If a basis has Properties (C) and (A), is it necessarily semi-greedy?
• Is it possible to formulate a new property so that every conservative basis with this
• If w = (1, 1, ...) in Theorem 5.7, do we get the same constant as in the classic case ([12,
property is necessarily democratic?
Theorem 3.4])?
The purpose of this appendix is to show two basic lemmas. The first one resembles a result
7. Appendix
from [7]. For each λ > 0, we define the λ-truncation of z ∈ C by
Tλ(z) =(cid:26) λsgn (z)
z
if z ≥ λ
if z ≤ λ
20
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
We extend Tλ to an operator on X by
Tλ(x) =Xj
Tλ(e∗
j (x))ej = Xj∈Λλ
λsgn (e∗
j (x))en +Xj∈Λc
λ
e∗
j (x)ej,
where Λλ = {j : λ < e∗
Lemma 7.1. For all λ > 0 and x ∈ X, if B is Cq-quasi-greedy, we have
j (x)}.
kTλ(x)k ≤ Cqkxk, k(I − Tλ)(x)k ≤ (Cq + 1)kxk, αk1εΛk ≤ 2Cqkxk,
where α = minj∈Λ e∗
j (x)).
Moreover, if B is Ku-unconditional, for every set A with A < ∞,
j (x), Λ is a greedy set of x and ε ≡ sgn (e∗
kTλ(I − PA)(x)k ≤ Kukxk.
Proof.
Tλ(x) =Z 1
0 "Xj
(x) ](s)e∗
λ
e∗
j
j (x)ej# ds =Z 1
0
χ[0,
(I − PΛλ,s)x ds,
where Λλ,s = {j : λ
Minkowski's integral inequality,
s < e∗
j (x)} is a greedy set of x of finite cardinality. Then, using the
kTλ(x)k ≤Z 1
0 k(I − PΛλ,s)xkds ≤ Cqkxk.
0 PΛλ,s(x)ds, hence
Also, since (I − Tλ)x =R 1
Now, since α1εΛ = Tα(x) − PΛc(x) =R 1
k(I − Tλ)(x)k ≤ (Cq + 1)kxk.
0 (PΛ(x) − PΛα,s(x))ds,
αk1εΛk ≤ 2Cqkxk.
On the other hand, if A is a general set with A < ∞,
Tλ(I − PA)x =Z 1
0
(I − PΛλ,s)(I − PA)x ds =Z 1
0
(I − PA∪Λλ,s)x ds,
thus
kTλ(I − PA)(x)k ≤ Kukxk.
(cid:3)
The second lemma involves the concept of w-partially-greedy bases.
Lemma 7.2. If B is Cw-w-conservative and Cq-quasi-greedy, then
j∈A ajk1ηBk,
ajejk ≤ 4CqCw max
kXj∈A
for any sign η and A, B ∈ N<∞ such that w(A) ≤ w(B), A < B and any collection of scalars
(aj)j∈A.
THE WEIGHTED PROPERTY (A) AND THE GREEDY ALGORITHM
21
Proof. We prove that k1εAk ≤ 4CqCwk1ηBk for any signs ε and η. First, we can decompose
1εA = 1A+ − 1A−, where A± = {j ∈ A : εj = ±1}. Then,
Now, using the condition to be quasi-greedy, it is clear that k1Bk ≤ 2Cqk1ηBk, then
k1εAk ≤ k1A+k + k1A−k ≤ 2Cwk1Bk.
k1εAk ≤ 4CqCwk1ηBk.
Now, using convexity, we are done.
(cid:3)
Acknowledgments: The first author thanks the University of Murcia for partially support-
ing of his research stay in the University of Illinois at Urbana-Champaign in September 2017,
where this paper began.
References
[1] F. Albiac, J. L. Ansorena, Characterization of 1-quasi greedy bases, J. Approx. Theory, 201, 7-12 (2016).
[2] F. Albiac, J. L. Ansorena, Characterization of 1-almost greedy bases, Rev. Mat. Complut., 30(1), 13-24
(2017).
[3] F. Albiac and N. Kalton, Topics in Banach Space Theory, second edition, ISBN 978-3-319-31555-3
(2016).
[4] F. Albiac, P. Wojtaszczyk, Characterization of 1-greedy bases, J. Approx. Theory, 138 (2006), 65-86.
[5] P. M. Bern´a, ´O. Blasco, Characterization of greedy bases in Banach spaces, J. Approx. Theory, 205
(2017), 28-39.
[6] P. M. Bern´a, ´O. Blasco, The best m-term approximation with respect to polynomials with constant
coefficients, Anal. Math., 43 (2) (2017), 119-132.
[7] P. M. Bern´a, ´O. Blasco, G. Garrig´os, Lebesgue inequalities for the greedy algorithm in general bases,
Rev. Mat. Complut. 30 (2017), 369-392.
[8] A. Cohen, R. A. DeVore, R. Hochmuth, Restricted nonlinear approximation, Constr. Approx., 16
(2000), 85-113.
[9] S. J. Dilworth, N. J. Kalton, D. Kutzarova, On the existence of almost greedy bases in Banach
spaces, Studia Math. 159 (2003), 67 -- 101.
[10] S. J. Dilworth, D. Kutzarova, T. Oikhberg, Lebesgue constants for the weak greedy algorithm, Rev.
Mat. Complut. 28(2), 393 -- 409 (2015).
[11] S. J. Dilworth, D. Kutzarova, E. Odell, T. Schlumprecht, A. Zs´ak, Renorming spaces with
greedy bases, J. Approx. Theory 188 (2014), 39-56.
[12] S. J. Dilworth, N. J. Kalton, D. Kutzarova, V. N. Temlyakov, The thresholding greedy algorithm,
greedy bases, and duality, Constr. Approx. 19 (2003), no.4, 575-597.
[13] S. J. Dilworth, D. Kutzarova, V. N. Temlyakov, B. Wallis, Weight-Almost greedy bases. (Preprint)
http://arxiv.org/abs/1803.02932v1.
[14] G. Garrig´os, E. Hern´andez, T. Oikhberg, Lebesgue-type inequalities for quasi-greedy bases, Constr.
Approx. 38 (2013), 447-470.
[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 (2006), 103-118.
[16] S. V. Konyagin, V. N. Temlyakov, A remark on greedy approximation in Banach spaces, East J.
Approx. 5 (1999), 365-379.
[17] V. N. Temlyakov, Greedy approximation, Cambridge Monographs on Applied and Computational Math-
ematics, vol.20, Cambridge University Press, Cambridge, 2011.
[18] P. Wojtaszczyk, Greedy algorithm for general biorthogonal systems, J. Approx.Theory 107 (2) (2000),
293-314.
[19] P. Wojtaszczyk, Greedy type bases in Banach spaces, Constructive theory of functions, 136-155, DARBA,
Sofia, 2003.
[20] J. Y. T. Woo, On a class of universal modular sequence spaces, Israel Journal of Mathematics 20 (1975),
193-215.
22
P. M. BERN ´A, S. J. DILWORTH, D. KUTZAROVA, T. OIKHBERG, AND B. WALLIS
Pablo M. Bern´a, Departmento de Matem´aticas, Universidad Aut´onoma de Madrid, 28049
Madrid, Spain
E-mail address: [email protected]
Stephen J. Dilworth, Department of Mathematics, University of South Carolina, Columbia
SC 29208, USA
E-mail address: [email protected]
Denka Kutzarova, Department of Mathematics, University of Illinois Urbana-Champaign,
Urbana, IL 61801, USA; and Institute of Mathematics and Informatics, Bulgarian Academy of
Sciences
E-mail address: [email protected]
Timur Oikhberg, Department of Mathematics, University of Illinois Urbana-Champaign, Ur-
bana, IL 61801, USA
E-mail address: [email protected]
Ben Wallis, Department of Mathematical Sciences, Northern Illinois University, DeKalb,
IL 60115-2888, USA
E-mail address: [email protected]
|
1303.4040 | 1 | 1303 | 2013-03-17T07:48:36 | Criteria for Hankel operators to be sign-definite | [
"math.FA",
"math.SP"
] | We show that total multiplicities of negative and positive spectra of a self-adjoint Hankel operator $H$ with kernel $h(t)$ and of an operator of multiplication by some real function $s(x)$ coincide. In particular, $\pm H\geq 0$ if and only if $\pm s(x)\geq 0$. The kernel $h(t)$ and its "sign-function" $s(x)$ are related by an explicit formula. An expression of $h(t)$ in terms of $s(x)$ leads to an exponential representation of $h(t)$.
Our approach directly applies to various classes of Hankel operators. In particular, for Hankel operators of finite rank, we find an explicit formula for the total multiplicity of their negative and positive spectra. | math.FA | math |
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
D. R. YAFAEV
Abstract. We show that total multiplicities of negative and positive spectra of a
self-adjoint Hankel operator H with kernel h(t) and of an operator of multiplication
by some real function s(x) coincide. In particular, ±H ≥ 0 if and only if ±s(x) ≥ 0.
The kernel h(t) and its "sign-function" s(x) are related by an explicit formula.
An expression of h(t) in terms of s(x) leads to an exponential representation of
h(t). Our approach directly applies to various classes of Hankel operators.
In
particular, for Hankel operators of finite rank, we find an explicit formula for the
total multiplicity of their negative and positive spectra.
1.1. Hankel operators can be defined as integral operators
1. Introduction
(Hf )(t) =Z ∞
0
h(t + s)f (s)ds
(1.1)
in the space L2(R+) with kernels h that depend on the sum of variables only. Of
course H is symmetric if h(t) = h(t). There are very few cases where Hankel
operators can be explicitly diagonalized. We mention classical results by F. Mehler
[8], T. Carleman [3], W. Magnus [7] and M. Rosenblum [12]. They are treated in a
unified way in [16] where some new examples are also considered.
Our goal here is to find necessary and sufficient conditions for the positivity1 of
Hankel operators. This question seems to be of particular importance because of its
intimate relation to a representation of the function h(t) as the Laplace transform
of a (positive) measure. Such representations are continuous analogues of moment
problems. According to the Hamburger theorem (see, e.g., the book [1]) the posi-
tivity of a discrete Hankel operator is equivalent to the existence of a solution of the
corresponding moment problem.
2000 Mathematics Subject Classification. 47A40, 47B25.
Key words and phrases. Hankel operators, convolutions, necessary and sufficient conditions
for the positivity, the sign-function, operators of finite rank, the Carleman operator and its
perturbations.
1We always use the term "positive" instead of a more precise but lengthy term "non-negative."
1
b(ξ − η)g(η)dη,
−∞
(Bg)(ξ) =Z ∞
2πR ∞
R ∞
b(ξ) =
1
0 h(t)t−iξdt
0 e−tt−iξdt
.
(1.2)
(1.3)
2
D. R. YAFAEV
1.2. Our condition of the positivity2 of Hankel operators is quite explicit. Let B,
be the operator in the space L2(R) of the convolution with the function
Of course b(−ξ) = b(ξ) if h(t) = h(t) so that the operator B is symmetric. Our
main result is that the Hankel operator H ≥ 0 if and only if B ≥ 0. We call b(ξ)
the b-function of the Hankel operator H (or of the kernel h(t)), and we use the term
the sign-function or s-function for its Fourier transform s(x). So our result, roughly
speaking, means that a Hankel operator H is positive if and only if its sign-function
s(x) is positive. Note that in specific examples we consider, functions s(x) may be of
a quite different nature. For instance, s(x) may be a polynomial or, on the contrary,
it may be a distribution; for example, it may be a combination of delta functions
and their derivatives.
Our proofs rely on the identity
(Hf, f ) = (Bg, g)
(1.4)
where
g(ξ) = Γ(1/2 + iξ) f (ξ) =: (Ξf )(ξ),
(1.5)
f (ξ) is the Mellin transform of f (t) and Γ(·) is the gamma function. Actually, the
identity (1.4) or, equivalently,
(1.6)
allows us to find the numbers of positive and negative eigenvalues of a Hankel op-
erator H. For a self-adjoint operator A, we denote by N+(A) (by N−(A)) the total
mutiplicity of its strictly positive (negative) spectrum. Then the identity (1.6) shows
that
H = Ξ∗BΞ
N±(H) = N±(B).
(1.7)
This result can be compared with Sylvester's inertia theorem which states the
same for Hermitian matrices H and B related by equation (1.6) provided the matrix
Ξ is invertible. In contrast to the linear algebra, in our case the operators H and B
are of a completely different nature and B (but not H) admits an explicit spectral
analysis.
Of course the results above can be reformulated in terms of the sign-function
s(x) = √2π(Φ∗b)(x) where Φ is the Fourier transform. Let us introduce the operator
S = Φ∗BΦ of multiplication by s(x). Then (1.6) means that H = bΞ∗SbΞ where
bΞ = Φ∗Ξ, and (1.7) means that N±(H) = N±(S).
2We usually discuss conditions for H ≥ 0, but of course replacing H by −H we obtain conditions
for H ≤ 0.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
3
1.3. The precise meaning of formula (1.3) requires some discussion. Observe that
the denominator in (1.3) coincides with the numerator for the special case h(t) = e−t.
It equals Γ(1 − iξ) and hence exponentially tends to zero as ξ → ∞. Therefore
b(ξ) is a "nice" function of ξ only under very restrictive assumptions on the kernel
h(t). Thus to cover natural examples, we have, on the contrary, to extend a class
of kernels and to work with distributions h(t). The choice of appropriate spaces
of distributions is also very important.
In order to be able to divide in (1.3) by
an exponentially decaying function, we assume that the numerator belongs to the
0 (R)′. It means that the Fourier transform of the function
class of distributions C ∞
0 (R)′. The corresponding class of functions h(t)
θ(x) = exh(ex) should belong to C ∞
0 (R)′. The
will be denoted Z ′
Schwartz space S(R)′ is too restrictive for our purposes which is seen already on the
example of finite rank Hankel operators.
0 (R)′ for the validity of the identity (1.4) is very general. It
is satisfied for all bounded, but also for a wide class of unbounded, Hankel operators
H. More than that, it is not even required that H be defined by formula (1.1) on
some dense set. Therefore we work with quadratic forms (Hf, f ) which is more
convenient and yields more general results. In this context it is natural to consider
distributions h(t) which makes the theory self-consistent.
+. Under the assumption h ∈ Z ′
The condition Φθ ∈ C ∞
+ we have b ∈ C ∞
If however h ∈ L1
loc(R+), then θ ∈ S(R)′ and hence Φθ ∈ C ∞
0 (R)′ if
Z ∞
0
h(t)(1 + ln t)−κdt < ∞
(1.8)
for some κ. Condition (1.8) is quite general;
corresponding Hankel operator be bounded. For example, it admits kernels
it also does not require that the
h(t) = P (ln t)t−1
(1.9)
where P (x) is an arbitrary polynomial. Note that Hankel operators with such kernels
are bounded for P (x) = const only.
As far as test functions f (t) are concerned, we require that their Mellin transforms
f ∈ C ∞
0 (R). Then both sides of (1.4) are well defined and the identity holds. We
note that the distribution b(ξ) is "worse" than the kernel h(t). On the contrary, due
to the factor Γ(1/2 + iξ) in (1.5), the test function g(ξ) is "better" than f (t). In
the case of bounded operators H, this permits us to extend the main identity (1.4)
to all elements f ∈ L2(R+).
1.4. It turns out that the knowledge of the sign-function s(x) allows one to recover
the kernel h(t) by the formula
where
e−tλh♮(λ)λdλ
h(t) =Z ∞
h♮(λ) = λ−1s(− ln λ).
0
(1.10)
(1.11)
4
D. R. YAFAEV
Thus h(t) is the Laplace transform of the distribution λh♮(λ). It is noteworthy that
+ and the correspondence h 7→ h♮ is a continuous one-to-one mapping of Z ′
h♮ ∈ Z ′
onto itself.
Representation (1.10) does not require the positivity of H. If however H ≥ 0,
then combining our results with the Bochner-Schwartz theorem, we obtain that
λh♮(λ)dλ = dm(λ) where dm(λ) is a positive measure. In this case
+
h(t) =Z ∞
0
e−tλdm(λ).
(1.12)
This representation implies that h(t) is necessarily a completely monotonic function
and, in particular, h ∈ C ∞(R+). Note that the converse statement is also true: a
completely monotonic function admits representation (1.12). This is one of famous
Bernstein's theorems (see his paper [2], or the book by N. Akhiezer [1] or the book by
D. V. Widder [13]). In contrast to the Bernstein theorem, we deduce representation
(1.12) from the positivity of the Hankel operator with kernel h(t) and show that the
measure dm(λ) satisfies for some κ the condition
Z ∞
0
(1 + ln λ)−κλ−1dm(λ) < ∞.
(1.13)
Hankel operators can also be realized as operators in the space of sequences l2
+.
The relation between this discrete representation and the continuous representation
we consider is given by the unitary transformation of l2
+ onto L2(R+) constructed in
terms of the Laguerre polynomials. Thus all our results can in principle be translated
into the discrete representation. This hopefully will be discussed in another article
on this subject.
1.5. A large part of the paper is devoted to applying the general theory to various
classes of Hankel operators although we do not try to cover all possible cases. In
some examples the sign definiteness of H can also be verified or refuted with the
help of Bernstein's theorems. Note however that our approach yields additionally
an explicit formula for the total numbers of negative and positive eigenvalues of H.
In Section 5, we present such a formula for Hankel operators of finite rank. Then
we consider two specific examples. The first one is given by the formula
h(t) = tke−αt, α > 0, k ≥ −1.
(1.14)
Note that the Hankel operator H with such kernel has finite rank for k ∈ Z+ only.
We show that H is positive if and only if k ≤ 0. The second class of kernels is
defined by the formula
(1.15)
It turns out that the corresponding Hankel operator is positive if and only if r ≤ 1.
Section 6 is devoted to a study of Hankel operators H with non-smooth ker-
nels. In this case both numbers N±(H) are infinite, and we find the asymptotics of
eigenvalues of H.
h(t) = e−tr
r > 0.
,
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
5
Finally, in Section 7 we consider perturbations of the Carleman operator, that
is, of the Hankel operator with kernel h0(t) = t−1 by various classes of compact
Hankel operators. The Carleman operator can be explicitly diagonalized by the
Mellin transform. We recall that it has the absolutely continuous spectrum [0, π] of
multiplicity 2. The Carleman operator plays the distinguished role in the theory of
Hankel operators. In particular, it is important for us that its sign-function s(x) = 1.
As was pointed out by J. S. Howland in [6], Hankel operators are to a certain extent
similar to differential operators. In terms of this analogy, the Carleman operator
H0 plays the role of the "free" Schrodinger operator D2, D = −id/dx, in the space
L2(R). Furthermore, Hankel operators H with "perturbed" kernels h(t) = t−1 + v(t)
can be compared to Schrodinger operators D2 + V(x). The assumption that v(t)
decays sufficiently rapidly as t → ∞ and is not too singular as t → 0 corresponds
to a sufficiently rapid decay of the potential V(x) as x → ∞.
As shown in [17], the results on the discrete spectrum of the operator H lying above
its essential spectrum [0, π] are close in spirit to the results on the discrete spectrum
of the Schrodinger operator D2 + V(x). On the contrary, the results on the negative
spectrum of the Hankel operator H are drastically different. In particular, contrary
to the case of differential operators with decaying coefficients, the finiteness of the
negative spectrum of the Hankel operator H is not determined by the behaviour
of v(t) at singular points t = 0 and t = ∞. As an example, consider the Hankel
operator with kernel
h(t) = t−1 − γe−tr
,
r ∈ (0, 1).
Now the kernel of the perturbation is the function which decays faster than any
power of t−1 as t → ∞ and it has the finite limit as t → 0. Nevertheless we show
that the negative spectrum of H is infinite if γ > γ0 (here γ0 = γ0(r) is an explicit
constant) while H is positive if γ ≤ γ0. Such a phenomenon has no analogy for
Schrodinger operators with decaying potentials. However it occurs for three-particle
Schrodinger operators and is known as the Efimov effect.
We also study perturbations of the Carleman operator H0 by Hankel operators
V of finite rank. Here we obtain a striking result: the total numbers of negative
eigenvalues of the operators H = H0 + V and V coincide.
As examples, we consider only bounded Hankel operators in this paper. However,
our general results directly apply to a wide class of unbounded operators, such
as Hankel operators with kernels (1.9). Moreover, with slight modifications our
method works also for kernels (1.14) where α ≥ 0 and k is an arbitrary negative
number. In this case condition (1.8) is not satisfied. Hankel operators with kernels
(1.9) and (1.14) generalize the Carleman operator, and we call them quasi-Carleman
operators. They will be studied elsewhere.
1.6. Let us briefly describe the structure of the paper. We obtain the main iden-
tity (1.4) and the reconstruction formula (1.10) in Section 2. Necessary information
on bounded Hankel operators (including a continuous version of the Nehari theorem)
6
D. R. YAFAEV
is collected in Section 3. In Sections 2 and 3 we do not assume that the function
h is real, i.e., the corresponding Hankel operator H is not necessarily symmetric.
Spectral consequences of the formula (1.4) and, in particular, criteria of the sign-
definiteness of Hankel operators are formulated in Section 4. In Sections 5, 6 and 7
we apply the general theory to particular classes of Hankel operators.
Let us introduce some standard notation. We denote by Φ,
(Φu)(ξ) = (2π)−1/2Z ∞
−∞
u(x)e−ixξdx,
the Fourier transform. The space Z = Z(R) of test functions is defined as the subset
of the Schwartz space S = S(R) which consists of functions ϕ admitting the analytic
continuation to entire functions in the complex plane C and satisfying bounds
ϕ(z) ≤ Cn(1 + z)−ner Im z,
∀z ∈ C,
for some r = r(ϕ) > 0 and all n. We recall that the Fourier transform Φ : C ∞
0 (R) →
Z and Φ∗ : Z → C ∞
0 (R). The dual classes of distributions (continuous antilinear
functionals) are denoted S ′, C ∞
0 (R)′ and Z ′, respectively. In general, for a linear
topological space L, we use the notation L′ for its dual space. The Dirac function
is standardly denoted δ(·).
We use the notation hhh·,·iii and h·,·i for scalar products and duality symbols in
L2(R+) and L2(R), respectively. They are always linear in the first argument and
antilinear in the second argument. The letter C (sometimes with indices) denotes
various positive constants whose precise values are inessential.
2. The main identity
2.1. Let us consider a Hankel operator H defined by equality (1.1) in the space
L2(R+). Actually, it is more convenient to work with sesquilinear forms (Hf1, f2)
instead of operators.
Before giving precise definitions, let us explain our construction at a formal level.
It follows from (1.1) that
(Hf1, f2) =Z ∞
0 Z ∞
=Z ∞
h(t)F (t)dt =: hhhh, Fiii,
0
0
h(t + s)f1(s)f2(t)dtds
where
F (t) =Z t
0
f1(s)f2(t − s)ds =: ( ¯f1 ⋆ f2)(t)
(2.1)
(2.2)
is the Laplace convolution of the functions ¯f1 and f2. Formula (2.1) allows us to
consider h as a distribution with the test function F defined by (2.2). Thus the
Hankel form will be defined by the relation
h[f1, f2] = hhhh, ¯f1 ⋆ f2iii.
(2.3)
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
7
Let us introduce the test function
and the distribution
Ω(x) = F (ex) =: (RF )(x)
θ(x) = exh(ex)
defined for x ∈ R. Setting in (2.1) t = ex, we see that
hhhh, Fiii =Z ∞
−∞
θ(x)Ω(x)dx =: hθ, Ωi.
(2.4)
(2.5)
(2.6)
We are going to consider the form (2.6) on pairs F, h such that the corresponding
test function Ω defined by (2.4) is an element of the space Z of analytic functions
and the corresponding distribution θ defined by (2.5) is an element of the dual space
Z ′. The set of all such F and h will be denoted Z+ and Z ′
+, respectively, that is,
F ∈ Z+ ⇐⇒ Ω ∈ Z and h ∈ Z ′
Of course, the topology in Z+ is induced by that in Z and Z ′
that h ∈ Z ′
the corresponding function (2.5) satisfies the condition
(2.7)
+ is dual to Z+. Note
loc(R+) and integral (1.8) is convergent for some κ. In this case
+ ⇐⇒ θ ∈ Z ′.
+ if h ∈ L1
Z ∞
−∞ θ(x)(1 + x)−κdx < ∞,
and hence θ ∈ S ′ ⊂ Z ′.
Define the unitary operator U : L2(R+) → L2(R) by the equality
(Uf )(x) = ex/2f (ex).
(2.8)
Let the set D consist of functions f (t) such that Uf ∈ Z. Since
f (t) = t−1/2(Uf )(ln t)
and Z ⊂ S, we see that functions f ∈ D and their derivatives satisfy the estimates
f (m)(t) = Cn,mt−1/2−m(1 + ln t)−n
for all n and m. Obviously, f ∈ D if and only if ϕ(t) = t1/2f (t) belongs to the class
Z+.
Let us show that form (2.3) is correctly defined on functions f1, f2 ∈ D. To that
end, we have to verify that function (2.2) belongs to the space Z+ or, equivalently,
function (2.4) belongs to the space Z. This requires some preliminary study which
will also allow us to derive a convenient representation for form (2.3).
Recall that the Mellin transform M : L2(R+) → L2(R) is defined by the formula
(2.9)
f (t)t−1/2−iξdt.
(Mf )(ξ) = (2π)−1/2Z ∞
0
8
D. R. YAFAEV
Of course, M = ΦU where Φ is the Fourier transform and U is operator (2.8). Since
both Φ and U are unitary, the operator M is also unitary. The inversion of the
formula (2.9) is given by the relation
f (t) = (2π)−1/2Z ∞
−∞
f (ξ)t−1/2+iξdξ,
f = Mf.
(2.10)
Let Γ(z) be the gamma function. Recall that Γ(z) is a holomorphic function in
the right half-plane and Γ(z) 6= 0 for all z ∈ C. According to the Stirling formula
the function Γ(z) tends to zero exponentially as z → ∞ parallel with the imaginary
axis. To be more precise, we have
Γ(α + iλ) = eπi(2α−1)/4(2π/e)1/2λα−1/2eiλ(ln λ−1)e−πλ/2(cid:0)1 + O(λ−1)(cid:1)
(2.11)
for a fixed α > 0 and λ → +∞. Since Γ(α − iλ) = Γ(α + iλ), this yields also the
asymptotics of Γ(α + iλ) as λ → −∞.
If fj ∈ D, j = 1, 2, then fj = Mfj = ΦUfj ∈ C ∞
0 (R) and hence the functions
gj(ξ) = Γ(1/2 + iξ) fj(ξ),
j = 1, 2,
(2.12)
also belong to the class C ∞
g1 and g2,
0 (R). Let us introduce the convolution of the functions
(g1 ∗ g2)(ξ) =Z ∞
−∞
g1(ξ − η)g2(η)dη,
and set
(J g)(ξ) = g(−ξ).
We have the following result.
Lemma 2.1. Suppose that fj ∈ D, j = 1, 2, and define functions gj(ξ) by equality
(2.12). Let the function Ω(x) be defined by formulas (2.2) and (2.4). Then
(ΦΩ)(ξ) = (2π)−1/2Γ(1 + iξ)−1((J ¯g1) ∗ g2)(ξ).
(2.13)
Proof. Substituting (2.10) into definition (2.2), we see that
F (t) = (2π)−1Z t
0
dsZ ∞
−∞
f1(τ )(t − s)−1/2−iτ dτZ ∞
−∞
f2(σ)s−1/2+iσdσ.
Observe that
(t − s)−1/2−iτ s−1/2+iσds = ti(σ−τ ) Γ(1/2 − iτ )Γ(1/2 + iσ)
.
Γ(1 + i(σ − τ ))
Then using definition (2.12) we obtain the representation
ti(σ−τ )Γ(1 + i(σ − τ ))−1g1(τ )g2(σ)dτ dσ
−∞Z ∞
tiξΓ(1 + iξ)−1((J ¯g1) ∗ g2)(ξ)dξ,
−∞
0
Z t
F (t) =(2π)−1Z ∞
=(2π)−1Z ∞
−∞
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
9
whence
Ω(x) = (2π)−1Z ∞
−∞
eiξxΓ(1 + iξ)−1((J ¯g1) ∗ g2)(ξ)dξ.
This is equivalent to formula (2.13).
(cid:3)
Observe that the function Γ(1 + iξ)−1 in the right-hand side of (2.13) tends to
0 (R) because (J ¯g1) ∗ g2 ∈
infinity exponentially as ξ → ∞. Nevertheless ΦΩ ∈ C ∞
C ∞
0 (R) for g1, g2 ∈ C ∞
Corollary 2.2. Let fj ∈ D, j = 1, 2, and let the function Ω(x) be defined by
formulas (2.2) and (2.4). Then Ω ∈ Z or, equivalently, F ∈ Z+.
0 (R). Thus we have
Now we are in a position to give the precise definition.
Definition 2.3. Let h ∈ Z ′
form is defined by the relation (2.3).
+ and fj ∈ D, j = 1, 2. Then the Hankel sesquilinear
We shall see in subs. 2.4 that h ∈ Z ′
Of course definition (2.3) can be rewritten as
hhhh, ¯f1 ⋆ f2iii on f1, f2 ∈ D, that is, h = 0 if hhhh, ¯f1 ⋆ f2iii = 0 for all f1, f2 ∈ D.
+ is determined uniquely by the values
where
and θ is distribution (2.5).
h[f1, f2] = hθ, Ωi
Ω(x) = ( ¯f1 ⋆ f2)(ex)
(2.14)
We sometimes write h[f1, f2] as integral (2.1) keeping in mind that its precise
meaning is given by Definition 2.3.
2.2. Our next goal is to show that (2.14) is the sesquilinear form of the convolution
operator B, that is, it equals the right-hand side of (1.4). Here the representation
of Lemma 2.1 for the function
G(ξ) = √2πΓ(1 + iξ)(ΦΩ)(ξ)
(2.15)
plays the crucial role.
Since θ ∈ Z ′, its Fourier transform a = Φθ is correctly defined as an element of
0 (R)′. Formally,
C ∞
a(ξ) = (Φθ)(ξ) = (2π)−1/2Z ∞
0
h(t)t−iξdt,
(2.16)
that is, a(ξ) is the Mellin transform of the function h(t)t1/2. Let Ω ∈ Z. Passing to
the Fourier transforms and using notation (2.15), we see that
where G ∈ C ∞
0 (R) and the distribution b ∈ C ∞
0 (R)′ is given by the relation
hθ, Ωi = ha, ΦΩi = hb, Gi
b(ξ) = (2π)−1/2a(ξ)Γ(1 − iξ)−1
which is of course the same as (1.3). Thus we are led to the following
(2.17)
(2.18)
10
D. R. YAFAEV
+. The distribution b ∈ C ∞
0 (R)′ defined by formulas
Definition 2.4. Let h ∈ Z ′
(2.5), (2.16) and (2.18) is called the b-function of the kernel h(t) (or of the Hankel
operator H). Its Fourier transform s = √2πΦ∗b ∈ Z ′ is called the s-function or the
Recall that the distribution h♮ ∈ Z ′
+ was defined by relation (1.11). The following
assertion is an immediate consequence of formulas (2.5), (2.16) and (2.18).
sign-function.
Proposition 2.5. The mappings
h 7→ θ 7→ a 7→ b 7→ s 7→ h♮
yield one-to-one correspondences (bijections)
0 (R) → C ∞
Z ′
+ → Z ′ → C ∞
0 (R) → Z ′ → Z ′
+.
All of them, as well as their inverse mappings, are continuous.
Putting together equalities (2.6) and (2.17), we see that
hhhh, Fiii = hb, Gi.
(2.19)
Combining this relation with Lemma 2.1 and Definitions 2.3, 2.4 and using notation
(1.5), we obtain the main identity (1.4). To be more precise, we have the following
result.
0 (R)′ be the corresponding b-
Theorem 2.6. Suppose that h ∈ Z ′
function. Let fj ∈ D, j = 1, 2, and let the functions gj = Ξfj be defined by formula
(2.12). Then gj ∈ C ∞
+, and let b ∈ C ∞
0 (R) and the representation
hhhh, ¯f1 ⋆ f2iii = hb, (J ¯g1) ∗ g2i =: b[g1, g2]
(2.20)
holds.
Passing in the right-hand side of (2.20) to the Fourier transforms and using that
Φ∗((J ¯g1) ∗ g2) = (2π)1/2Φ∗g1Φ∗g2,
we obtain
Corollary 2.7. Let s ∈ Z ′ be the sign-function of h, and let uj = Φ∗gj = Φ∗Ξfj ∈
Z. Then
(2.21)
hhhh, ¯f1 ⋆ f2iii = hs, ¯u1u2i =: s[u1, u2].
Loosely speaking, equalities (2.20) and (2.21) mean that
hhhh, ¯f1 ⋆ f2iii =Z ∞
−∞Z ∞
=Z ∞
−∞
−∞
b(ξ − η)g1(η)g2(ξ)dξdη
(2.22)
s(x)u1(x)u2(x)dx.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
11
In the particular case h(t) = t−1, we have θ(x) = 1,
a(ξ) = (2π)1/2δ(ξ),
b(ξ) = δ(ξ),
s(x) = 1,
(2.23)
and hence (2.22) yields
hhhh, ¯f1 ⋆ f2iii =Z ∞
−∞
where
g1(ξ)g2(ξ)dξ =Z ∞
−∞ Γ(1/2 + iξ)2 f1(ξ) f2(ξ)dξ,
Γ(1/2 + iξ)2 =
π
cosh(πξ)
.
(2.24)
This leads to the familiar diagonalization of the Hankel operator H with kernel
h(t) = t−1. This operator is known as the Carleman operator and will be denoted
by C.
Let ΓΓΓα : C ∞
2.3. According to Proposition 2.5 the distribution h♮ determines uniquely the
distribution h. Let us now obtain an explicit formula for the mapping h♮ 7→ h. This
requires some auxiliary information.
0 (R) → C ∞
0 (R), α > 0, be the operator of multiplication by the
function Γ(α + iξ). Making the change of variables t = e−x in the definition of the
gamma function, we see that
Γ(α + iλ) =Z ∞
e−ttα+iλ−1dt =Z ∞
e−αxe−ixλdx, α > 0,
e−e−x
−∞
0
and hence
It follows that
−∞
(2π)−1Z ∞
(Φ∗ΓΓΓαΦΩ)(x) =Z ∞
−∞
Let us also introduce the operator Lα:
(LαF )(λ) = λαZ ∞
0
eixλΓ(α + iλ)dλ = e−e−x
e−αx.
eα(y−x)e−ey−x
Ω(y)dy.
(2.25)
(2.26)
e−tλtα−1F (t)dt,
λ > 0, α > 0.
Obviously, LαF ∈ C ∞(R+) for all bounded functions F (t) and, in particular, for
F ∈ Z+. Note that Lα is the Laplace operator L,
(LF )(λ) =Z ∞
0
e−tλF (t)dt,
(2.27)
sandwiched by the weights λα and tα−1.
Recall that the operator R defined by (2.4) is a one-to-one mapping of Z+ onto
We need the following result.
Z.
12
D. R. YAFAEV
Lemma 2.8. For all α > 0, the identity
Lα = R−1J Φ∗ΓΓΓαΦR
(2.28)
holds. In particular, Lα as well as its inverse are the one-to-one continuous mappings
of Z+ onto itself.
Proof. Putting Ω(y) = (RF )(y) = F (ey) in (2.26) and making the change of vari-
ables t = ey, we find that
(Φ∗ΓΓΓαΦRF )(x) = e−αxZ ∞
0
e−e−xttα−1F (t)dt.
Now making the change of variables λ = e−x, we arrive at the identity (2.28).
Consider the right-hand side of (2.28). All mappings R : Z+ → Z, Φ : Z →
C ∞
0 (R), ΓΓΓα : C ∞
0 (R) → Z, J : Z → Z are bijections. All of
them as well as their inverses are continuous. Therefore the identity (2.28) ensures
the same result for the operator Lα : Z+ → Z+.
(cid:3)
0 (R) → C ∞
0 (R), Φ∗ : C ∞
To recover h(t), we proceed from formula (2.19). Passing to the Fourier trans-
forms, we can write it as
where G is defined by formulas (2.4), (2.15), that is, G = (2π)1/2ΓΓΓ1ΦRF. Therefore
using the identity (2.28) for α = 1, we see that
hhhh, Fiii = (2π)−1/2hs, Φ∗Gi
hhhh, Fiii = hs,J RL1Fi =Z ∞
−∞
s(x)(L1F )(e−x)dx.
Making the change of variables λ = e−x in the right-hand side, we obtain the identity
Passing here to adjoint operators and using that F ∈ Z+ is arbitrary, we find that
(2.29)
h = L∗
1h♮,
hhhh, Fiii = hhhh♮, L1Fiii.
which gives the precise sense to formula (1.10). Of course formula (1.10) can also
be rewritten as
h(t) =Z ∞
−∞
e−te−x
e−xs(x)dx.
Let us state the result obtained.
+, and let s ∈ Z ′ be the corresponding sign-functions (see
Theorem 2.9. Let h ∈ Z ′
Definition 2.4). Define the distribution h♮ by formula (1.11). Then h♮(λ) belongs to
the class Z ′
+ and h can be recovered from h♮ by formula (2.29).
We emphasize that in the roundabout h 7→ h♮ 7→ h the mappings h 7→ h♮ as well
+ onto itself.
Let us also give a direct expression of u(x) = (Φ∗g)(x) in terms of f (t).
as its inverse h♮ 7→ h are one-to-one continuous mappings of the set Z ′
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
13
Lemma 2.10. Suppose that f ∈ D and put ϕ(t) = t1/2f (t). Let g(ξ) be defined by
formula (1.5) and u(x) = (Φ∗g)(x). Then
u(x) = (L1/2ϕ)(e−x).
(2.30)
Proof. Since (Rϕ)(x) = (Uf )(x), it follows from formula (2.28) for α = 1/2 that
(R−1J Φ∗ΓΓΓ1/2ΦUf )(λ) = (L1/2ϕ)(λ).
The left-hand side here equals (R−1J u)(λ) which after the change of variables λ =
e−x yields (2.30).
(cid:3)
Now we can rewrite identity (2.21) in a slightly different way.
Corollary 2.11. Let h ∈ Z ′
(1.11). Then for arbitrary fj ∈ D, j = 1, 2, and ϕj(t) = t1/2fj(t), we have
+, and let the distribution h♮ ∈ Z ′
+ be defined by formula
(2.31)
Proof. It suffices to make the change of variables x = − ln λ in the right-hand side
of (2.22) and to take equality (2.30) into account.
hhhh, ¯f1 ⋆ f2iii = hhhh♮, L1/2ϕ1L1/2ϕ2iii.
(cid:3)
We emphasize that according to Lemma 2.8, L1/2ϕj ∈ Z+ and hence
L1/2ϕ1L1/2ϕ2 ∈ Z+. Thus the right-hand side of (2.31) is correctly defined.
+ is determined uniquely by the
values hhhh, ¯f1 ⋆ f2iii on f1, f2 ∈ D. First we consider convolution operators. Let us
introduce the shift in the space L2(R):
2.4. Finally, we check that a distribution h ∈ Z ′
Since
(T (τ )g)(ξ) = g(ξ − τ ),
τ ∈ R.
(g1 ∗ g2)(ξ) =Z ∞
−∞
(T (τ )g1)(ξ)g2(τ )dτ,
∀g1, g2 ∈ C ∞
0 (R),
we have the formula
hb, (J ¯g1) ∗ g2i =Z ∞
−∞hb, T (τ )J ¯g1ig2(τ )dτ
(2.32)
(2.33)
where for b ∈ C ∞
0 (R)′ the function hb, T (τ )J ¯g1i is infinitely differentiable in τ ∈ R.
The following assertion is quite standard.
0 (R)′. Suppose that hb, (J ¯g1) ∗ g2i = 0 for all g1, g2 ∈
Lemma 2.12. Let b ∈ C ∞
C ∞
0 (R). Then b = 0.
Proof. If hb, (J ¯g1) ∗ g2i = 0 for all g2 ∈ C ∞
0 (R), then hb, T (τ )J ¯g1i = 0 for all τ ∈ R
according to formula (2.33). In particular, for τ = 0 we have hb,J ¯g1i = 0 whence
b = 0 because g1 ∈ C ∞
0 (R) is arbitrary.
(cid:3)
Next we pass to Hankel operators.
14
D. R. YAFAEV
+. Suppose that hhhh, ¯f1 ⋆ f2iii = 0 for all f1, f2 ∈ D.
Proposition 2.13. Let h ∈ Z ′
Then h = 0.
Proof. Let b ∈ C ∞
0 (R)′ be the b-function of h (see Definition 2.4). For arbitrary
0 (R), we can construct f1, f2 ∈ D by formula (2.12). Since hhhh, ¯f1 ⋆f2iii = 0,
g1, g2 ∈ C ∞
it follows from the identity (2.20) that hb, (J ¯g1)∗ g2i = 0. Therefore b = 0 according
to Lemma 2.12. Now Proposition 2.5 implies that h = 0.
(cid:3)
3. Bounded Hankel operators
Our main goal here is to show that the condition h ∈ Z ′
bounded Hankel operators H.
+ is satisfied for all
3.1. In this section we a priori only assume that h ∈ C ∞
Hankel form (2.3) on functions f1, f2 ∈ C ∞
restriction of the shift (2.32) on its invariant subspace L2(R+). Since
0 (R+)′ and consider the
0 (R+). Let T+(τ ) where τ ≥ 0 be the
( ¯f1 ⋆ f2)(t) =Z ∞
0
(T+(τ ) ¯f1)(t)f2(τ )dτ,
∀f1, f2 ∈ C ∞
0 (R+),
for all h ∈ C ∞
0 (R+)′ we have the formula
hhhh, ¯f1 ⋆ f2iii =Z ∞
0
hhhh, T+(τ ) ¯f1iiif2(τ )dτ.
(3.1)
Here the function hhhh, T+(τ ) ¯f1iii is infinitely differentiable in τ ∈ R+, and this function,
as well as all its derivatives, have finite limits as τ → 0. In the theory of Hankel
operators, formula (3.1) plays the role of formula (2.33) for convolution operators.
The proof of the following assertion is almost the same as that of Lemma 2.12.
0 (R+)′. Suppose that hhhh, ¯f1 ⋆ f2iii = 0 for all f1, f2 ∈
Proposition 3.1. Let h ∈ C ∞
C ∞
0 (R+). Then h = 0.
0 (R+), then hhhh, T+(τ ) ¯f1iii = 0 for all τ ∈ [0,∞)
Proof. If hhhh, ¯f1 ⋆f2iii = 0 for all f2 ∈ C ∞
according to formula (3.1). In particular, for τ = 0 we have hhhh, ¯f1iii = 0 which implies
that h = 0 because f1 ∈ C ∞
0 (R+) is arbitrary.
(cid:3)
Of course Propositions 2.13 and 3.1 differ only by the set of functions on which
the Hankel form is considered.
Assume now that
Then there exists a bounded operator H such that
hhhh, ¯f ⋆ fiii ≤ Ckfk2,
∀f ∈ C ∞
0 (R+).
(Hf1, f2) = hhhh, ¯f1 ⋆ f2iii,
∀f1, f2 ∈ C ∞
0 (R+).
We call H the Hankel operator associated to the Hankel form hhhh, ¯f1 ⋆ f2iii.
(3.2)
(3.3)
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
15
3.2. It is possible to characterize Hankel operators by some commutation rela-
tions. Let us define a bounded operator Σ in the space L2(R+) by the equality
Note that
esf (s)ds.
(Σf )(t) = −2e−tZ t
Σ = −2Z ∞
0
T+(τ )e−τ dτ.
0
(3.4)
Lemma 3.2. Let assumption (3.2) hold. Then the corresponding Hankel operator
H satisfies the commutation relations:
HT+(τ ) = T+(τ )∗H,
∀τ ≥ 0,
HΣ = Σ∗H.
(3.5)
(3.6)
and
Proof. Since
(T+(τ ) ¯f1) ⋆ f2 = ¯f1 ⋆ (T+(τ )f2),
∀τ ≥ 0,
relation (3.5) directly follows from definition (3.3). By virtue of formula (3.4),
relation (3.6) is a consequence of (3.5).
(cid:3)
Below we need the Nehari theorem; see the original paper [9], or the books [10],
+(R)
Chapter 1, §1 or [11], Chapter 1, §2. We formulate it in the Hardy space H2
of functions analytic in the upper half-plane. We denote by bΣ the operator of
multiplication by the function (µ − i)/(µ + i) in this space.
Theorem 3.3 (Nehari). Let ω ∈ L∞(R), and let a bounded operator bH in the space
H2
+(R) be defined by the relation
(bH f1, f2) =Z ∞
−∞
ω(µ) f1(−µ) f2(µ)dµ,
∀ f1, f2 ∈ H2
Then bHbΣ = bΣ∗bH. Conversely, if bH is a bounded operator in H2
then there exists a function ω ∈ L∞(R) such that representation (3.7) holds.
+(R) and bHbΣ = bΣ∗bH,
The following assertion can be regarded as a translation of this theorem into the
+(R) → L2(R+) is the unitary
space L2(R+). Recall that the Fourier transform Φ : H2
operator. Since
+(R).
(3.7)
Z ∞
−∞
µ − i
µ + i
e−iµtdµ = −4πe−t
for t > 0 and this integral is zero for t < 0, we have the relation
bΣ = Φ∗ΣΦ.
(3.8)
16
D. R. YAFAEV
Theorem 3.4. If representation (3.3) holds with h = Φω where ω ∈ L∞(R) (in this
case h ∈ S ′ ⊂ C ∞
0 (R+)′), then estimate (3.2) is true and the corresponding Hankel
operator satisfies commutation relation (3.5). Conversely, if a bounded operator H
satisfies (3.5), then representation (3.3) holds with h = Φω for some ω ∈ L∞(R).
Proof. Since
(Φ∗(f1 ⋆ ¯f2))(µ) = (J f1)(µ) f2(µ),
∀f1, f2 ∈ C ∞
0 (R+),
where f1 = Φ∗f1, f2 = Φ∗f2 and (J f1)(µ) = f1(−µ), we have
∀h ∈ S ′.
hhhh, ¯f1 ⋆ f2iii = hhhΦ∗h, (J f1) f2iii,
(3.9)
Therefore estimate (3.2) is satisfied if Φ∗h ∈ L∞(R). Relation (3.5) for the corre-
sponding Hankel operator H follows from Lemma 3.2.
Conversely, if a bounded operator H satisfies relation (3.5), then by virtue of (3.4)
it also satisfies relation (3.6). Hence it follows from (3.8) that bHbΣ = bΣ∗bH where
bH = Φ∗HΦ is a bounded operator in the space H2
exists a function ω ∈ L∞(R) such that representation (3.7) holds. It means that
ω(µ) f1(−µ) f2(µ)dµ,
(Hf1, f2) =Z ∞
+(R). Thus, by Theorem 3.3, there
∀f1, f2 ∈ L2(R+).
(3.10)
−∞
If h = Φω, then the right-hand sides in (3.9) and (3.10) coincide. This yields
representation (3.3).
(cid:3)
Corollary 3.5. For a bounded operator H, commutation relations (3.5) and (3.6)
are equivalent.
Proof. As was already noted, (3.6) follows from (3.5) according to formula (3.4).
Conversely, if H satisfies (3.5), then representation (3.3) holds according to Theo-
rem 3.4. Thus it remains to use Lemma 3.2.
(cid:3)
A function ω ∈ L∞(R) such that Φω = h is called the symbol of a bounded Hankel
operator H with kernel h(t). Of course if ω ∈ H∞
− (R), that is, ω admits an analytic
continuation to a bounded function in the lower half-plane, then the corresponding
Hankel operator is zero. Therefore the symbol is defined up to a function in the
class H∞
− (R).
3.3. Now we are in a position to check that the condition h ∈ Z ′
+ is satisfied for
all bounded Hankel operators. By definition (2.7) it means that distribution (2.5)
belongs to the class Z ′. We shall verify the stronger inclusion θ ∈ S ′.
To that end, it suffices to check that, for some N ∈ Z+ and some κ ∈ R,
hθ, Ωi ≤ C
NXn=0
max
x∈R (cid:0)(1 + x)κΩ(n)(x)(cid:1),
∀Ω ∈ C ∞
0 (R).
(3.11)
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
17
Putting F (t) = Ω(ln t), we see that (3.11) is equivalent to the estimate
hh, Fi ≤ C
NXn=0
max
t∈R+(cid:0)(1 + ln t)κtnF (n)(t)(cid:1), F ∈ C ∞
0 (R+).
(3.12)
If h ∈ L1
loc(R), then estimate (3.12) for N = 0 is equivalent to the convergence of
integral (1.8). If H is Hilbert-Schmidt, that is
Z ∞
0
h(t)2tdt < ∞,
then integral (1.8) converges for any κ > 1/2. Similarly, if h(t) ≤ Ct−1, then
integral (1.8) converges for any κ > 1.
For the proof of (3.12) in the general case, we use the following elementary result.
Its proof is given in Appendix A.
Lemma 3.6. If F ∈ C ∞
0 (R+), then for an arbitrary κ > 2 the estimate
kΦ∗FkL1(R) ≤ C
2Xn=0
max
t∈R+(cid:0)(1 + ln t)κtnF (n)(t)(cid:1)
holds.
Corollary 3.7. If h = Φω where ω ∈ L∞(R), then estimate (3.12) holds for N = 2
and an arbitrary κ > 2.
Since, by Theorem 3.4, for a bounded Hankel operator H, its kernel h = Φω for
some ω ∈ L∞(R), we arrive at the following result.
0 (R+)′ and that condition (3.2) is satisfied.
Theorem 3.8. Suppose that h ∈ C ∞
Then estimate (3.12) holds for N = 2 and an arbitrary κ > 2; in particular, h ∈ Z ′
+.
The following simple example shows that for N = 0 estimate (3.12) is not in
general true (for all κ).
Example 3.9. Let h(t) = e−it2. Then the corresponding Hankel operator H is
bounded because according to the formula e−i(t+s)2 = e−it2e−i2tse−is2, it is a product
of three bounded operators. However integral (1.8) diverges at infinity for all κ. In
this example condition (3.12) is satisfied for N = 1 and κ = 0. Indeed, integrating
by parts, we see that
Z ∞
h(t)F (t)dt = −Z ∞
where the function h1(t) = R t
0 e−is2ds is bounded and h1(t) = O(t) as t → 0.
Therefore integral (3.13) is bounded by maxt∈R+(cid:0)(1 + ln t)κtF ′(t)(cid:1) for κ > 1.
h1(t)F ′(t)dt
(3.13)
0
0
18
D. R. YAFAEV
Note that the symbol of H equals ω(µ) = (2π)−1e−πi/4eiµ2/4. More generally, one
can consider the class of symbols ω(µ) such that ω ∈ C ∞(R), ω(µ) = eiω0µα, ω0 > 0,
for large positive µ and ω(µ) = 0 for large negative µ. Of course Hankel operators
with such symbols are bounded. Using the stationary phase method, we find that
for α > 1 the corresponding kernel h(t) has the asymptotics
h(t) ∼ h0tβeiσtγ
,
(3.14)
where β = (1 − α/2)(α − 1)−1, γ = α(α − 1)−1 and h0, σ = ¯σ are some numbers.
Moreover, h(t) is a bounded function on all finite intervals. Similarly to Example 3.9,
it can be checked that for such kernels condition (3.12) is satisfied for N = 1 but
not for N = 0. If α ∈ (0, 1), then h(t) has asymptotics (3.14) for t → 0.
t → ∞,
3.4. Here we shall show that, for bounded Hankel operators H, the represen-
tations (2.20) and (2.21) extend to all f1, f2 ∈ L2(R+). By Theorem 3.8, we have
h ∈ Z ′
+. Let b and s be the corresponding b- and s-functions (see Definition 2.4).
Recall that the operator Ξ : L2(R+) → L2(R) was defined by formula (1.5). We de-
note by K the operator of multiplication by the functionpcosh(πξ)/π in the space
L2(R). It follows from identity (2.24) and the unitarity of the Mellin transform (2.9)
that
and hence the operator KΞ : L2(R+) → L2(R) is unitary. Therefore in view of the
identities (2.20) and (2.21) we have the following result.
kKΞfk = kfk,
Lemma 3.10. The inequalities (3.2),
and
hb, (J ¯g) ∗ gi ≤ CkKgk2,
∀g ∈ C ∞
0 (R),
(3.15)
(3.16)
are equivalent. The Hankel operator corresponding to form (2.3) is bounded if and
only if one of equivalent estimates (3.2), (3.15) or (3.16) is satisfied.
hs,u2i ≤ CkKΦuk2,
∀u ∈ Z,
These estimates can be formulated in a slightly different way. Let us introduce
the space E ⊂ L2(R) of exponentially decaying functions with the norm kgkE =
kKgk. Then the space W = Φ∗E consists of functions u(x) admitting the analytic
continuation u(z) in the strip Im z ∈ (−π/2, π/2); moreover, functions u(x + iy)
have limits in L2(R) as y → ±π/2. The identity
defines the Hilbert norm on W. We call W the exponential Sobolev space because it
is contained in standard Sobolev spaces Hl(R) for all l. The operators Ξ : L2(R+) →
E and bΞ := Φ∗Ξ : L2(R+) → W are of course unitary. Obviously, kKgk and kKΦuk
in the right-hand sides of (3.15) and (3.16) can be replaced by kgkE and kukW,
E = (2π)−1Z ∞
kΦuk2
−∞(cid:0)u(x + iπ/2)2 + u(x − iπ/2)2(cid:1)dx =: kuk2
W
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
19
are equivalent.
respectively. Note that the inclusions f ∈ L2(R+), g = Ξf ∈ E and u = bΞf ∈ W
If one of the equivalent estimates (3.2), (3.15) or (3.16) is satisfied, then all oper-
ators H : L2(R) → L2(R), B : E → E ′ and S : W → W ′ are bounded. Using that
relations fn → f in L2(R+), gn = Ξfn → g = Ξf in E and un = Φ∗gn → u = Φ∗g
in W are equivalent, we extend (2.20) and (2.21) to all f ∈ L2(R+). Thus we have
obtained the following result.
Proposition 3.11. If one of equivalent estimates (3.2), (3.15) or (3.16) is satisfied,
then the identities
(Hf1, f2) = (Bg1, g2) = (Su1, u2),
gj = Ξfj,
uj = Φ∗gj,
are true for all f1, f2 ∈ L2(R+).
Let Kl be the operator of multiplication by the function (1+ξ2)l/2. Then estimates
(3.15) or (3.16) are satisfied provided
hb, (J ¯g) ∗ gi ≤ ClkKlgk2
or
hs,u2i ≤ Clkuk2
Hl(R),
(3.17)
for some l; in this case
C = Clπ max
ξ∈R (cid:0)(1 + ξ2)l(cosh(πξ))−1(cid:1).
We note the following assertion.
Proposition 3.12. A Hankel operator H is bounded if its sign-function satisfies the
condition
s ∈ L1(R) + L∞(R).
(3.18)
If s ∈ L∞(R) and s(x) → 0 as x → ∞, then H is compact.
Proof. The first statement is obvious because under assumption (3.18) the second
estimate (3.17) is satisfied for l > 1/2. To prove the second statement, we observe
that the operator SΦ∗K −1 is compact because both S and K −1 are operators of
multiplication by bounded functions which tends to zero at infinity. It follows that
the operator
H = Ξ∗Φ(SΦ∗K −1)(KΞ)
is also compact.
(cid:3)
We emphasize, however, that as show already examples of Hankel operators H of
finite rank (see formula (5.9) for h(t) = e−αt), condition (3.18) is not necessary for
the boundedness of H.
20
D. R. YAFAEV
4. Criteria of the sign-definiteness
In this section we suppose that h(t) = h(t) so that the operator H is symmetric.
The results of Section 2 allow us to give simple necessary and sufficient conditions
for a Hankel operator H to be positive or negative. Moreover, they provide also
convenient tools for a evaluation of the total multiplicity of the negative and positive
spectra of H. We often formulate results only for the negative spectrum. The
corresponding results for the positive spectrum are obtained if H is replaced by
−H.
+ and introduce the b-function b ∈ C ∞
4.1. Actually, we consider the problem in terms of Hankel quadratic forms rather
than Hankel operators. This is both more general and more convenient. As usual,
0 (R)′
we suppose that a distribution h ∈ Z ′
and the s-function s ∈ Z ′ by Definition 2.4.
Below we use the following natural notation. Let h[ϕ, ϕ] be a real quadratic form
defined on a linear set D. We denote by N±(h) the maximal dimension of linear sets
M± ⊂ D such that ±h[ϕ, ϕ] > 0 for all ϕ ∈ M±, ϕ 6= 0. We apply this definition to
the forms h[f, f ] = hhhh, ¯f ⋆ fiii defined on D, to b[g, g] = hb,J ¯g ∗ gi defined on C ∞
0 (R)
and to s[u, u] = hs,u2i defined on Z. Of course, if D is dense in a Hilbert space H
and h[ϕ, ϕ] is closed on D, then for the self-adjoint operator H corresponding to h,
we have N±(H) = N±(h) .
Observe that formula (1.5) establishes one-to-one correspondence between the sets
0 (R). Moreover, the Fourier transform establishes one-to-one correspon-
0 (R) and Z. Therefore the following assertion is a direct
D and C ∞
dence between the sets C ∞
consequence of Theorem 2.6.
Theorem 4.1. Let h ∈ Z ′
+. Then
N±(h) = N±(b) = N±(s).
In particular, we have
Theorem 4.2. Let h ∈ Z ′
only if the form ±hb,J ¯g ∗ gi ≥ 0 for all g ∈ C ∞
all u ∈ Z.
+. Then the form ±hhhh, ¯f ⋆ fiii ≥ 0 for all f ∈ D if and
0 (R), or the form ±hs,u2i ≥ 0 for
4.2.
In many cases the following consequence of Theorem 4.1 is convenient.
According to Proposition 3.12, under the assumptions of Theorem 4.3, H is defined
as the bounded self-adjoint operator corresponding to the form hhhh, ¯f ⋆ fiii. Therefore
N±(h) = N±(H) is the total multiplicity of the (strictly) positive spectrum for the
sign "+" and of the (strictly) negative spectrum for the sign "−" of the operator
H. For definiteness, we consider the negative spectrum.
Theorem 4.3. Let h ∈ Z ′
+, and let the corresponding sign-function satisfy condition
(3.18). If s(x) ≥ 0, then the operator H is positive. If s(x) ≤ −s0 < 0 for almost
all x in some interval ∆ ⊂ R, then the operator H has infinite negative spectrum.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
21
Choose a function ϕ ∈ C ∞
Proof. If s(x) ≥ 0, then H ≥ 0 according to the second relation (2.22).
Let s(x) ≤ −s0 < 0 for x ∈ ∆. For an arbitrary N, we shall construct a linear set
L ⊂ Z of dimension N such that s[u, u] < 0 for all u ∈ L, u 6= 0. Then the second
statement will follow from Theorem 4.1.
0 (R) such that ϕ(x) = 1 for x ∈ [−δ, δ] and ϕ(x) = 0
for x 6∈ [−2δ, 2δ] where δ = δN is a sufficiently small number. Let points αj ∈ ∆,
j = 1, . . . , N, be such that αj+1 − αj = αj − αj−1 for j = 2, . . . , N − 1. Set
∆j = (αj − δ, αj + δ), e∆j = (αj − 2δ, αj + 2δ). For a sufficiently small δ, we
may suppose that e∆j ⊂ ∆ for all j = 1, . . . , N and that e∆j+1 ∩ e∆j = ∅ for
j = 1, . . . , N − 1. We set ϕj(x) = ϕ(x − αj). Since s(x) ≤ −s0 < 0 for x ∈ ∆, we
have
hs,ϕj2i =Z ∞
−∞
s(x)ϕj(x)2dx ≤ −2δs0 < 0.
(4.1)
The problem is that ϕj ∈ C ∞
The functions ϕ1, . . . , ϕN have disjoint supports and hence hs,u2i < 0 for an arbi-
trary non-trivial linear combination u of the functions ϕj.
0 (R) but ϕj 6∈ Z (and even ϕj 6∈ W). Thus we have
j ∈ Z. For a
to approximate these functions in the topology of S by functions ϕ(ε)
given κ > 1/2 and an arbitrary ε > 0, we can find a function ϕ(ε) ∈ Z such that
max
x∈R (cid:0)(1 + x)κϕ(x) − ϕ(ε)(x)(cid:1) < ε.
(4.2)
Now we set ϕ(ε)
j (x) = ϕ(ε)(x − αj).
independent. Assume that
Let us check that for ε small enough, the functions ϕ(ε)
1 , . . . , ϕ(ε)
N are linearly
NXj=1
λjϕ(ε)
j (x) = 0.
(4.3)
If x ∈ ∆k, then ϕ(ε)
the estimate
k (x) − 1 < ε and ϕ(ε)
j (x) < ε for j 6= k. Therefore (4.3) yields
(1 − ε)λk ≤ εXj6=k
λj.
Summing these estimates over k = 1, . . . , N, we see that
NXj=1
Hence λj = 0 for all j = 1, . . . , N if ε(N − 1) < 1 − ε.
It follows from (4.1), (4.2) that, for all j = 1, . . . , N,
λj ≤ ε(N − 1)
(1 − ε)
NXj=1
λj.
Z ∞
−∞
s(x)ϕ(ε)
j (x)2dx ≤ −2δs0 + Cε
22
and that
D. R. YAFAEV
Thus we have the estimate
j 6= k.
−∞
(cid:12)(cid:12)λj ¯λkZ ∞
Z ∞
−∞
s(x)ϕ(ε)
j (x)ϕ(ε)
k (x)dx(cid:12)(cid:12) ≤ Cε(λj2 + λk2),
j (x)2dx ≤ −
NXj=1
NXj=1
λjϕ(ε)
s(x)
λj2(2δs0 − Cε).
The right-hand side here is negative if ε is small enough and PN
Let L be spanned by the functions ϕ(ε)
N for sufficiently small ε. Then
dimL = N and s[u, u] < 0 for all u ∈ L, u 6= 0 according to (4.4). It follows from
Theorem 4.1 that N−(H) ≥ N, and hence N−(H) = ∞.
j=1 λj2 6= 0.
1 , . . . , ϕ(ε)
(4.4)
(cid:3)
Theorem 4.3 can be reformulated, although in a weaker form, in terms of the
functions b(ξ) and even h(t). Suppose, for example, that
(4.5)
Then its Fourier transform s(x) is a continuous function which tends to 0 as x → ∞.
The operator B defined by formula (1.2) is bounded, self-adjoint and
b ∈ L1(R).
spec(B) = [min
x∈R
s(x), max
x∈R
s(x)].
The result below follows directly from Theorem 4.3. Note that by Proposition 3.12
under assumption (4.5) the operator H is compact.
Proposition 4.4. Under assumption (4.5) the Hankel operator H is positive if and
only if s(x) ≥ 0. If minx∈R s(x) < 0, then necessarily H has an infinite negative
spectrum.
In particular, condition (4.5) is satisfied if
In this case a = Φθ ∈ C ∞
θ(ln t)
h(t) =
where
0 (R) and hence b ∈ C ∞
t
0 (R).
θ ∈ Z.
4.3. For a proof that a Hankel operator is not sign-definite it is sometimes even
not necessary to calculate the sign-function s(x) (the Fourier transform of b(ξ)). It
turns out that if b(ξ) grows as ξ → ∞, then the form b[g, g] = hb,J ¯g∗ gi cannot be
sign-definite. More precisely, we have the following statement about convolutions
with growing kernels b(−ξ) = b(ξ).
Theorem 4.5. Let b = b0+b∞ where b0 ∈ C p(R)′ for some p ∈ Z+ and b∞ ∈ L∞
loc(R).
Suppose that there exists a sequence of intervals ∆n = (rn − σn, rn + σn) where
rn → ∞ (or equivalently rn → −∞) and the sequence σn is bounded such that
lim
n→∞
σl
n min
ξ∈∆n
Re b∞(ξ) = ∞ or
lim
n→∞
σl
n max
ξ∈∆n
Re b∞(ξ) = −∞,
(4.6)
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
23
where l = 2 if p = 0 or p = 1 and l = p + 1 if p ≥ 2. Then for both signs N±(b) ≥ 1.
Proof. Since b can be replaced by −b, we can assume that, for example, the first
condition (4.6) is satisfied. Pick a real even function ϕ ∈ C ∞
0 (R) such that ϕ(ξ) ≥ 0,
ϕ(ξ) = 1 for ξ ≤ 1/4, ϕ(ξ) = 0 for ξ ≥ 1/2 and set
gn(ξ) = ϕ((ξ − rn/2)/σn) ± ϕ((ξ + rn/2)/σn).
(4.7)
An easy calculation shows that
((J gn) ∗ gn)(ξ) = 2σnψ(ξ/σn) ± σnψ((ξ − rn)/σn) ± σnψ((ξ + rn)/σn)
(4.8)
0 (R). The function ψ(ξ) is also even, ψ(ξ) ≥ 0, ψ(ξ) ≥ 1/8
where ψ = (J ϕ)∗ ϕ ∈ C ∞
for ξ ≤ 1/8 and ψ(ξ) = 0 for ξ ≥ 1.
Since hb0, gi ≤ CkgkC p, it follows from (4.8) that
hb0, (J gn) ∗ gni ≤ Cσ1−p
nZ ∞
hb∞, (J gn) ∗ gni = 2σ2
b∞(σnη)ψ(η)dη ± 2σ2
Moreover, again according to (4.8) we have
nZ ∞
n
.
−∞
−∞
(4.9)
Re b∞(σnη + rn)ψ(η)dη.
The first term in the right-hand side is O(σ2
estimate
(4.10)
n). For the second one, we use the
32Z ∞
−∞
Re b∞(σnη + rn)ψ(η)dη ≥ min
ξ−rn≤σn
Re b∞(ξ).
(4.11)
Let us first choose the sign " + " in (4.7). Then using representation (4.10) and
putting together estimates (4.9) and (4.11), we obtain the lower bound
hb, (J gn) ∗ gni ≥ −c(σ1−p
n + σ2
n) + σ2
n/16 min
Re b∞(ξ).
ξ−rn≤σn
If p = 0 or p = 1, then under the first condition (4.6) the right-hand side here tends
to +∞ as n → ∞. If p ≥ 2, it is bounded from below by
Re b∞(ξ)(cid:1)
n (cid:0) − c + σl
n/16 min
ξ−rn≤σn
σ1−p
where the expression in the brackets tends again to +∞ . Therefore hb, (J gn)∗gni >
0 for sufficiently large n. Similarly choosing the sign " − " in (4.7), we see that
hb, (J gn) ∗ gni < 0 for sufficiently large n.
Corollary 4.6. Instead of condition (4.6) assume that
(cid:3)
lim
ξ→∞
Re b∞(ξ) = ∞ or
lim
ξ→∞
Re b∞(ξ) = −∞.
Then for both signs N±(b) ≥ 1.
24
D. R. YAFAEV
In contrast to Theorem 4.5 there are no restrictions in Corollary 4.6 on the pa-
rameter p in the assumption b0 ∈ C p(R)′.
On the other hand, condition (4.6) permits Re b(ξ) to tend to ±∞ only on some
system of intervals. Moreover, the lengths of these intervals may tend to zero. In
this case, however, the growth of Re b(ξ) and the decay of these lengths should be
correlated and there are restrictions on admissible values of the parameters p and l.
Unlike Theorem 4.3, Theorem 4.5 does not guarantee that N = ∞; see subs. 5.4,
for a discussion of various possible cases.
4.4. Theorem 4.2 can be combined with the Bochner-Schwartz theorem (see,
e.g., Theorem 3 in §3 of Chapter II of the book [5]). It states that a distribution
b ∈ C ∞
0 (R) (such b are
sometimes called distributions of positive type) is the Fourier transform
0 (R)′ satisfying the condition hb,J ¯g ∗ gi ≥ 0 for all g ∈ C ∞
b(ξ) = (2π)−1Z ∞
−∞
e−ixξdM(x)
of a positive measure dM(x) such that
Z ∞
−∞
(1 + x)−κdM(x) < ∞
(4.12)
for some κ (that is, of at most polynomial growth at infinity). In particular, this
ensures that b ∈ S ′.
distribution b is of positive type.
determined by the measure dM(x):
Theorem 4.2 implies that if hhhh, ¯f ⋆ fiii ≥ 0 for all f ∈ D, then the corresponding
It means that the sign-function s(x) of h(t) is
hs, ϕi =Z ∞
−∞
ϕ(x)dM(x), ϕ ∈ S,
that is, s(x)dx = dM(x). Let us define the measure
dm(λ) = λdM(− ln λ),
λ ∈ R+.
(4.13)
It is easy to see that condition (4.12) is equivalent to condition (1.13) on measure
In terms of distribution (1.11), we have λh♮(λ)dλ = dm(λ). Therefore
(4.13).
Theorem 2.9 leads to the following result.
+ and hhhh, ¯f ⋆ fiii ≥ 0 for all f ∈ D. Then h(t) admits the
Theorem 4.7. Let h ∈ Z ′
representation (1.12) with a positive measure dm(λ), λ ∈ R+, satisfying for some κ
condition (1.13).
The representation (1.12) is of course a particular case of (1.10). It is much more
precise than (1.10) but requires the positivity of hhhh, ¯f ⋆ fiii. Theorem 4.7 shows that
the positivity of hhhh, ¯f ⋆ fiii imposes very strong conditions on h(t). Actually, we have
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
25
Corollary 4.8. Let h ∈ Z ′
and
+ and hhhh, ¯f ⋆ fiii ≥ 0 for all f ∈ D. Then h ∈ C ∞(R+)
(4.14)
(−1)nh(n)(t) ≥ 0
for all t > 0 and all n = 0, 1, 2, . . . (such functions are called completely monotonic).
The function h(t) admits an analytic continuation in the right-half plane Re t > 0
and it is uniformly bounded in every strip Re t ∈ (t1, t2) where 0 < t1 < t2 < ∞.
Moreover, for some κ ∈ R and C > 0 the estimate holds:
t > 0.
(4.15)
h(t) ≤ Ct−1(1 + ln t)κ,
All these assertions are direct consequences of the representation (1.12). In par-
ticular, under condition (1.13) we have
h(t) ≤ C max
λ≥0 (cid:0)e−tλλ(1 + ln λ)κ(cid:1)
which yields (4.15).
Note that according to the Bernstein theorem (see, e.g., Theorems 5.5.1 and 5.5.2
in [1]) condition (4.14) implies that the function h(t) admits the representation
(1.12) with some measure dm(λ). Of course, condition (4.14) does not impose any
restrictions on the measure dm(λ) (except that the integral (1.12) is convergent for
all t > 0). In contrast to the Bernstein theorem we deduce the representation (1.12)
from the positivity of the Hankel form. In this context condition (1.13) is due to
the assumption h ∈ Z ′
+.
We mention also a related result of H. Widom. He considered in [14] Hankel
operators H with kernels h(t) admitting the representation (1.12) and showed that
H is bounded if and only if m([0, λ)) = O(λ) as λ → 0 and as λ → ∞. In this case
h(t) ≤ Ct−1 for some C > 0. Thus Theorem 4.7 and Corollary 4.8 can be regarded
as an extension of Widoms's results to unbounded operators.
Under the positivity assumption the identity (2.21) takes a more precise form.
Proposition 4.9. Let h ∈ Z ′
positive measure dM(x) satisfying condition (4.12) for some κ such that
+ and hhhh, ¯f ⋆ fiii ≥ 0 for all f ∈ D. Then there exists a
hhhh, ¯f1 ⋆ f2iii =Z ∞
−∞
u1(x)u2(x)dM(x)
for all fj ∈ D, j = 1, 2, and uj = Φ∗Ξfj ∈ Z where the mapping Ξ is defined by
(1.5).
5. Applications and examples
********************************************************
5.1. Consider first self-adjoint Hankel operators H of finite rank. Recall that
integral kernels of Hankel operators of finite rank are given (this is the Kronecker
26
D. R. YAFAEV
theorem -- see, e.g., Sections 1.3 and 1.8 of the book [10]) by the formula
h(t) =
MXm=1
Pm(t)e−αmt
(5.1)
where Re αm > 0 and Pm(t) are polynomials of degree Km. If H is self-adjoint, that
is, h(t) = h(t), then the set {α1, . . . , αM} consists of points lying on the real axis
and pairs of points symmetric with respect to it. Let Im αm = 0 for m = 1, . . . , M0
and Im αm > 0, αM1+m = ¯αm for m = M0 + 1, . . . , M0 + M1. Thus M = M0 + 2M1;
of course the cases M0 = 0 or M1 = 0 are not excluded. The condition h(t) = h(t)
also requires that Pm(t) = Pm(t) for m = 1, . . . , M0 and PM1+m(t) = Pm(t) for
m = M0 + 1, . . . , M0 + M1. As is well known and as we shall see below,
rank H =
MXm=1
Km + M =: r.
For m = 1, . . . , M0, we denote by pm = ¯pm the coefficient at tKm in the polynomial
Pm(t).
The following assertion yields an explicit formula for the numbers N±(H). Its
proof will be given in [18].
Theorem 5.1. For m = 1, . . . , M0, set
− = (Km + 1)/2
if Km is odd
N (m)
+ = N (m)
+ − 1 = N (m)
N (m)
N (m)
+ = N (m)
− = Km/2
− − 1 = Km/2
if Km is even and pm > 0
(5.2)
if Km is even and pm < 0.
Then the total numbers N±(H) of (strictly) positive and negative eigenvalues of the
operator H are given by the formula
N±(H) =
M0Xm=1
N (m)
± +
M0+M1Xm=M0+1
Km + M1.
(5.3)
Formula (5.2) shows that every pair
Pm(t)e−αmt + Pm+M1(t)e−αm+M1 t, m = M0 + 1, . . . , M0 + M1,
(5.4)
of complex conjugate terms in (5.1) yields Km + 1 positive and Km + 1 negative
eigenvalues. The contribution of every real term Pm(t)e−αmt where m = 1, . . . , M0
also consists of the equal numbers (Km + 1)/2 of positive and negative eigenvalues
if the degree Km of the polynomial Pm(t) is odd.
If Km is even, then there is
one more positive (negative) eigenvalue if pm > 0 (pm < 0). In particular, in the
question considered, there is no "interference" between different terms Pm(t)e−αmt,
m = 1, . . . , M0, and pairs (5.4) in representation (5.1) of the kernel h(t).
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
27
According to (5.3) the operator H cannot be sign-definite if M1 > 0. Moreover,
± = 0 for m = 1, . . . , M0 if and only if Km = 0 and ∓pm > 0.
according to (5.2), N (m)
Therefore we have the following result.
Corollary 5.2. A Hankel operator H of finite rank is positive (negative) if and only
if its kernel is given by the formula
h(t) =
M0Xm=1
pme−αmt
where αm > 0 and pm > 0 (pm < 0).
Corollary 5.2 admits different proofs which avoid formula (5.3). For example,
one can use that although the functions Pm(t)e−αmt are analytic in the right-half
plane Re t > 0, they are bounded for t = τ + iσ as σ → ∞ for a constant Pm(t)
only. Therefore according to Corollary 4.8 such Hankel operators cannot be posi-
tive. Alternatively, using formula (5.16) below, one can deduce Corollary 5.2 from
Theorem 4.5.
5.2. Consider now Hankel operators H with kernels (1.14). Since the case k =
0, 1, . . . (finite rank Hankel operators) has been discussed in the previous subsection,
here we suppose that k 6= 0, 1, . . .. If k > −1, condition (1.8) is satisfied for all κ,
and the operators H are compact (actually, they belong to much better classes of
operators). If k = −1, then condition (1.8) is satisfied for κ > 1, and the operators
H are bounded but not compact.
Let us calculate the corresponding b- and s-functions. If k > −1, then function
(2.16) equals
a(ξ) = (2π)−1/2Z ∞
0
and hence function (2.18) equals
tke−αtt−iξdt = (2π)−1/2α−1−k+iξΓ(1 + k − iξ),
(5.5)
b(ξ) = α−1−k+iξ Γ(1 + k − iξ)
2πΓ(1 − iξ)
.
(5.6)
If k = −1, then in accordance with formulas (5.5) and (5.6), we have
a(ξ) = (2π)−1/2αiξ lim
ε→+0
Γ(ε − iξ),
b(ξ) = (2π)−1αiξi(ξ + i0)−1.
It yields the expression
s(x) = 0,
x > β,
(5.7)
for the function s = √2πΦ∗b. Formula (5.7) remains true for the Carleman operator
C (the Hankel operator with kernel h(t) = t−1) when α = 0. Indeed, in this case
according to (2.23) the sign-function s(x) = 1.
x < β, where β = − ln α,
s(x) = 1,
28
D. R. YAFAEV
Next, we calculate the Fourier transform of function (5.6). Assume first that
k ∈ (−1, 0). Then (see, e.g., formula (1.5.12) in the book [4])
t−k−1(t + 1)−1+iξdt =
Γ(−k)Γ(1 + k − iξ)
.
Γ(1 − iξ)
Making here the change of variables t + 1 = α−1e−x, we find that
0
Z ∞
Γ(−k)Z ∞
1
−∞
(e−x − α)−k−1
+
e−ixξdx = α−1−k−iξ Γ(1 + k − iξ)
Γ(1 − iξ)
.
Passing now to the inverse Fourier transform, we see that for k ∈ (−1, 0) the sign-
function s(x) = sk(x) of kernel (1.14) equals
s(x) =
1
Γ(−k)
(e−x − α)−k−1
+
.
(5.8)
Let us verify that this formula remains true for all non-integer k. To that end,
we assume that (5.8) holds for some non-integer k > −1 and check it for k1 = k + 1.
Since
we have
Γ(1 + k1 − iξ) = (k1 − iξ)Γ(1 + k − iξ),
sk1(x) = α−1(k1 − ∂)sk(x).
Substituting here formula (5.8) for sk(x) and differentiating this expression, we
obtain formula (5.8) for sk1(x). This concludes the proof of relation (5.8) for all
k ≥ −1.
Lemma 5.3. Let h(t) be given by formula (1.14) where k 6∈ Z+. Then the sign-
function is determined by relation (5.8).
Actually, relation (5.8) remains true for k ∈ Z+ if one takes into account that the
has poles at integer points. For example, for k = 0 we
distribution (e−x − α)−k−1
have
+
s(x) = α−1δ(x − β).
(5.9)
Obviously, s(x) = 0 for x > β = − ln α. If k = −1, then s(x) = 1 for x < β. If
k ∈ (−1, 0), then s(x) ≥ 0 and s ∈ L1(R). Therefore it follows from Theorem 4.3
that H ≥ 0.
If k > 0, then distribution (5.8) does not have a definite sign. Therefore it can
be deduced from Theorem 4.2 that the corresponding Hankel operator also is not
sign-definite.
Alternatively, for the proof of this result we can use Corollary 4.6. Formula (2.11)
implies that function (5.6) has the asymptotics
b(ξ) = (2π)−1α−1−k−iξ(−iξ)k(1 + O(ξ−1)),
ξ → ∞.
(5.10)
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
29
Making the dilation transformation in (1.14), we can suppose that α = 1. Then we
have
Re b(ξ) = (2π)−1 cos(πk/2)ξk + O(ξk−1),
(5.11)
Since cos(πk/2) 6= 0 unless k is an integer odd number, this expression tends to ±∞
if ± cos(πk/2) > 0. Thus Corollary 4.6 for the case b0 = 0 ensures that the Hankel
operator H is not sign-definite.
ξ → +∞,
Let us summarize the results obtained.
Proposition 5.4. The Hankel operator with kernel (1.14) is positive for k ∈ [−1, 0],
and it is not sign-definite for k > 0.
Explicit formulas for the sign-functions can also be used to treat more complicated
Hankel operators. For example, in view of (5.7) the following assertion directly
follows from Theorem 4.3.
Example 5.5. The Hankel operator with kernel
h(t) = t−1(e−α1t − γe−α2t),
is positive if and only if α2 ≥ α1 ≥ 0 and γ ≤ 1.
γ ≥ 0,
5.3. In this subsection, we consider the Hankel operator H with kernel (1.15).
Condition (1.8) is now fulfilled for all κ, and the operator H belongs of course to
the Hilbert-Schmidt class (actually, to much better classes). Observe that
a(ξ) = (2π)−1/2Z ∞
0
e−tr
t−iξdt = (2π)−1/2r−1Γ((1 − iξ)/r)
and define, as usual, the function b(ξ) by formula (2.18) so that
b(ξ) = (2πr)−1 Γ((1 − iξ)/r)
Γ(1 − iξ)
.
(5.12)
Consider first the case r > 1. It follows from the Stirling formula (2.11) that for
all r > 1 the modulus of function (5.12) exponentially grows and the periods of its
oscillations tend to zero only logarithmically as ξ → ∞. Therefore Theorem 4.5
implies that the Hankel operator with kernel (1.15) is not sign-definite.
The Hankel operator H with kernel h(t) = e−t2 can also be treated (see Appen-
dix B) in a completely different way which is perhaps also of some interest. This
method shows that both positive and negative spectra of the operator H are infinite.
If r = 1, then h(t) = e−t yields a positive Hankel operator of rank 1.
Let us now consider the case r < 1. Then, again according to the Stirling formula
(2.11), function (5.12) belongs to L1(R) so that its Fourier transform
s(x) = (2πr)−1Z ∞
−∞
Γ((1 − iξ)/r)
Γ(1 − iξ)
eixξdξ =: Ir(x)
(5.13)
30
D. R. YAFAEV
is a continuous function which tends to 0 as x → ∞. Therefore by Proposi-
tion 4.4the corresponding Hankel operator H ≥ 0 if and only if Ir(x) ≥ 0 for all
x ∈ R.
It turns out that Ir(x) ≥ 0. Surprisingly, we have not found a proof of this fact
in the literature, but it follows from our results. Only for r = 1/2, integral (5.13)
can be explicitly calculated. Indeed, according to formula (1.2.15) of [4]
Γ(2(1 − iξ))
Γ(1 − iξ)
= 21−2iξπ−1/2Γ(3/2 − iξ).
Therefore it follows from formula (2.25) that
I1/2(x) = 2−1π−1/2e3x/2e−ex/4
(5.14)
which is of course positive.
For an arbitrary r ∈ (0, 1), one can proceed from the Bernstein theorem on
completely monotonic functions (see subs. 4.4). Observe that if
ψ(t) = t−pe−tr
,
p ≥ 0,
(5.15)
then
ψ′(t) = −pt−p−1e−tr − rt−p+r−1e−tr
.
Further differentiations of ψ(t) change the sign and yield sums of terms having the
form (5.15). Thus the function h(t) = e−tr satisfies for all n condition (4.14) and
hence admits the representation (1.12) with some positive measure dm(λ). It follows
from (1.12) that
(Hf, f ) =Z ∞
0
(Lf )(λ)2dm(λ) ≥ 0,
∀f ∈ C ∞
0 (R+),
where L is the Laplace transform (2.27). Since the operator H is bounded, this
implies that H ≥ 0.
Thus we have obtained the following result.
Proposition 5.6. The Hankel operator with kernel (1.15) is positive for r ∈ (0, 1],
and it is not sign-definite for r > 1.
Putting together this result with Theorem 4.3, we see that integral (5.13) is pos-
itive for all r ∈ (0, 1). Our indirect proof of this fact looks curiously enough.
5.4. Let us now discuss convolution operators with growing kernels b(ξ). We
emphasize that condition (4.6) does not guarantee that the numbers N±(b) are
Indeed, consider the kernel h(t) = tke−αt where k is a positive integer.
infinite.
Formula (5.6) shows that for Im α = 0 the corresponding function b(ξ)
(5.16)
has a power asymptotics as ξ → ∞. According to Theorem 5.1 the positive and
negative spectra of the Hankel operator H with the kernel h(t) are finite; for example,
b(ξ) = (2π)−1α−1−k+iξ(1 − iξ)· · · (k − iξ)
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
31
H has exactly (k + 1)/2 positive and negative eigenvalues if k is odd. Moreover, if
Im α 6= 0, then in view of (5.16) the function b(ξ) exponentially grows as ξ → +∞
or ξ → −∞. Nevertheless the Hankel operator H with kernel h(t) = tk(e−αt + e− ¯αt)
has exactly k + 1 positive and negative eigenvalues.
On the other hand, for kernel (5.12) where r = 2 we have N±(b) = ∞. This
follows from Theorem 4.1 because, by Proposition B.1, the Hankel operator with
kernel h(t) = e−t2 has infinite number of positive and negative eigenvalues.
A similar phenomenon occurs for Hankel operators with non-smooth kernels. This
is discussed in the next section.
6. Hankel operators with non-smooth kernels
6.1. Let the symbol of the Hankel operator H be defined by the formula ω(µ) =
if t0 ≤ 0, and hence H = 0 in this case. If t0 > 0, then the
eit0µ. Evidently, ω ∈ H∞
integral kernel of H equals δ(t − t0) so that
−
(Hf )(t) = f (t0 − t).
Condition (3.12) is now satisfied for N = 0 and κ = 0.
The operator H admits an explicit spectral analysis. Indeed, observe first that
(Hf )(t) = 0 for t > t0 and hence L2(t0,∞) ⊂ Ker H. Since H 2f = f for f ∈
L2(0, t0), the restriction of H on its invariant subspace L2(0, t0) may have only ±1
as eigenvalues. Obviously, the eigenspace H± of H corresponding to the eigenvalue
±1 consists of all functions f (t) such that f (t) = ±f (t0 − t). Since
H+ ⊕ H− ⊕ L2(t0,∞) = L2(R+),
the spectrum of H consists of the eigenvalues 0, 1,−1 of infinite multiplicity each.
In this example, the b-function equals
b(ξ) =
t−iξ
0
2πΓ(1 − iξ)
.
(6.1)
Note that all functions (5.12) where r > 1, (6.1) as well as (6.2) below exponentially
grow and oscillate at infinity.
In these cases the corresponding Hankel operators
have infinite positive and negative spectra.
6.2. It follows from Corollary 4.8 that a Hankel operator H can be sign-definite
only for kernels h ∈ C ∞(R+). Actually, if h(t) or one of its derivatives h(l)(t) has a
jump discontinuity, then H has infinite number of both positive λ(+)
and negative
−λ(−)
n , n = 1, 2, . . . , eigenvalues, and we can even calculate their asymptotics as
n → ∞. Positive (negative) eigenvalues are of course enumerated in decreasing
(increasing) order with multiplicities taken into account.
n
Let us start with an explicit
32
D. R. YAFAEV
Example 6.1. Let h(t) = (t0 − t)l for some l = 0, 1, . . . if t ≤ t0 and h(t) = 0 if
t > t0. Then
(Hf )(t) =Z t0−t
0
(t0 − t − s)lf (s)ds,
t ∈ (0, t0),
and (Hf )(t) = 0 for t ≥ t0. For such h(t), the symbol equals
(iµt0)k(cid:1)
ω(µ) = l!(iµ)−l−1(cid:0)eiµt0 −
lXk=0
1
k!
and the b-function equals
b(ξ) =
l!tl+1−iξ
0
2πΓ(l + 2 − iξ)
.
Let us consider the spectral problem Hf = λf , that is,
Z t0−t
(t0 − t − s)lf (s)ds = λf (t),
Differentiating this equation k times, we find that
0
t ∈ (0, t0).
(−1)kl(l − 1)· · · (l − k + 1)Z t0−t
0
(t0 − t − s)l−kf (s)ds = λf (k)(t)
(6.2)
(6.3)
(6.4)
(6.5)
(6.6)
for k = 1, . . . , l. Differentiating the last equation where k = l once more, we see
that
Setting t = t0 in (6.3) and (6.4), we obtain the boundary conditions
l!f (t) = λ(−1)l+1f (l+1)(t0 − t),
t ∈ (0, t0).
f (t0) = f ′(t0) = · · · = f (l)(t0) = 0.
Conversely, if a function f (t) satisfies equation (6.5) and boundary conditions (6.6),
it satisfies also equation (6.3). This leads to the following intermediary result.
Lemma 6.2. Let the operator A be defined on the Sobolev class Hl+1(0, t0) by the
equation
(6.7)
Considered with boundary conditions (6.6), it is self-adjoint in the space L2(0, t0),
and its eigenvalues αn are linked to eigenvalues λn of the operator H by the equation
αn = l!λ−1
n .
(Af )(t) = (−1)l+1f (l+1)(t0 − t).
6.3. Clearly, A2 is a differential operator and the asymptotics of its eigenvalues
is described by the Weyl formula. However, to find the asymptotics of eigenvalues
of the operator A, we have to distinguish its positive and negative eigenvalues. For
this reason, it is convenient to introduce an auxiliary operator eA with symmetric
(with respect to the point 0) spectrum having the same asymptotics of eigenvalues
as A.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
33
We define the operator eA by the same formula (6.7) as A but consider it on
functions in Hl+1(0, t0/2) ⊕ Hl+1(t0/2, t0) satisfying the boundary conditions
f (k)(0) = f (k)(t0/2 − 0),
f (k)(t0/2 + 0) = f (k)(t0),
where k = 0, . . . , l for l even and
f (k)(0) = f (k)(t0/2 − 0) = f (k)(t0/2 + 0) = f (k)(t0) = 0,
where k = 0, . . . , (l − 1)/2 for l odd. The operator eA is self-adjoint in the space
and it is determined by the matrix
(6.8)
(6.9)
(6.10)
L2(0, t0) = L2(0, t0/2) ⊕ L2(t0/2, t0),
0 (cid:19) , A1,2 = A∗
eA =(cid:18) 0 A1,2
A2,1
2,1,
where A2,1 : L2(0, t0/2) → L2(t0/2, t0). The operator A2,1 is again given by relation
(6.7) on functions in Hl+1(0, t0/2) satisfying conditions (6.8) or (6.9) at the points
0 and t0/2 − 0. It follows from formula (6.10) that the spectrum of the operator eA
is symmetric with respect to the point 0 and consists of eigenvalues ±an where a2
are eigenvalues of the operator A∗
An easy calculation shows that A is the differential operator A = (−1)l+1∂2l+2
in the space L2(0, t0/2) defined on functions in the class H2l+2(0, t0/2) satisfying the
boundary conditions f (k)(0) = f (k)(t0/2) where k = 0, . . . , 2l + 1 for l even and the
boundary conditions
2,1A2,1 =: A.
n
f (k)(0) = f (k)(t0/2) = f (l+1+k)(0) = f (l+1+k)(t0/2) = 0,
where k = 0, . . . , (l− 1)/2 for l odd. The asymptotics of eigenvalues a2
by the Weyl formula, that is,
n of A is given
an = (2πt−1
0 n)l+1(1 + O(n−1)).
symmetric operator A0 with finite deficiency indices. For example, A0 can be defined
by formula (6.7) on C ∞ functions vanishing in some neighbourhoods of the points 0,
Let us now observe that the operators A and eA are self-adjoint extensions of a
t0/2 and t0. Therefore the operators A and eA have the same asymptotics of spectra.
Taking Lemma 6.2 into account, we obtain the following result.
Lemma 6.3. Eigenvalues of the Hankel operator H defined in Example 6.1 have the
asymptotics
λ(±)
n = l!(2π)−l−1tl+1
0 n−l−1(1 + O(n−1)).
Remark 6.4. In the case l = 0 we have the explicit formulas
n = (2π)−1t0(n − 3/4)−1,
λ(+)
n = (2π)−1t0(n − 1/4)−1, n = 1, 2, . . . .
λ(−)
34
D. R. YAFAEV
6.4. Now we are in a position to obtain the asymptotics of the spectrum of Hankel
operators whose kernels (or their derivatives) have jumps of continuity. We combine
Lemma 6.3 with the result by V. V. Peller (see Theorem 7.4 in Chapter 6 of his book
[10]) which implies that singular numbers sn(V ) of a Hankel operator V satisfy the
bound
sn(V ) = o(n−l−1)
if V has a symbol belonging to the Besov class Bl+1
(l+1)−1(R). Applying the Weyl
theorem on the preservation of the power asymptotics for the sum of operators, we
can state the following result.
Theorem 6.5. Let l ∈ Z+, and let v(t) be the Fourier transform of a function in
Bl+1
(l+1)−1(R). Set
for t ≤ t0 and h(t) = v(t) for t > t0. Then eigenvalues of the Hankel operator H
have the asymptotics
h(t) = h0(t0 − t)l + v(t)
n = h0l!(2π)−l−1tl+1
λ(±)
0 n−l−1(1 + o(1))
as n → ∞.
We emphasize that under the assumptions of this theorem the leading terms of
the asymptotics of positive and negative eigenvalues are the same. Of course if h(t)
becomes smoother (l increases), then eigenvalues of the Hankel operator H decrease
faster as n → ∞.
7. Perturbations of the Carleman operator
In this section we consider operators H = H0 + V where H0 is the Carleman
operator C (or a more general operator) and the perturbation V belongs to one of
the classes introduced in Section 5. Different objects related to the operator H0
will be endowed with the index "0", and objects related to the operator V will be
endowed with the index "v".
7.1. For perturbations V of finite rank, we have the following result.
Theorem 7.1. Let the sign-function s0(x) of a Hankel operator H0 be bounded and
positive. If V is a Hankel operator of finite rank and H = H0 + V , then
N−(H) = N−(V ).
In particular, H ≥ 0 if and only if V ≥ 0.
Applying Theorem 5.1 to the operator V , we get an explicit formula for the total
number of negative eigenvalues of the operator H.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
35
Corollary 7.2. Let the kernel v(t) of V be given by the formula
v(t) =
MXm=1
Pm(t)e−αmt
where Pm(t) is a polynomial of degree Km. Define the numbers N (m)
(5.2). Then N−(H) is given by formula (5.3).
±
by formula
Since for the Carleman operator C the sign-function equals 1, Theorem 7.1 applies
to H0 = C.
The inequality N−(H) ≤ N−(V ) is of course obvious because H0 ≥ 0. On the
contrary, the opposite inequality N−(H) ≥ N−(V ) looks surprising because the
operator H0 which may have the continuous spectrum is much "stronger" than the
operator V of finite rank. At a heuristic level the equality N−(H) = N−(V ) can
be explained by the fact that the supports of the sign-functions s0(x) and sv(x) are
essentially disjoint. Very loosely speaking, it means that the operators H0 and V
"live in orthogonal subspaces", and hence the positive operator H0 does not affect
the negative spectrum of H. The detailed proof of Theorem 7.1 as well as that of
Theorem 5.1 will be given in [18].
7.2. Let C be the Carleman operator, and let V be the Hankel operator with
kernel
v(t) = tke−αt, α > 0,
k > −1.
The operator V is compact, and hence the essential spectrum specess(Hγ) of the
operator
(7.1)
(7.2)
Hγ = C − γV,
γ ∈ R,
coincides with the interval [0, π]. Since the sign-function of the operator C equals
1, the sign-function sγ of the operator Hγ equals
sγ(x) = 1 − γsv(x)
where the function sv(x) is given by formula (5.8).
Let first k ∈ (−1, 0). Observe that sv(x) is continuous for x < β = − ln α and
sv(x) → +∞ as x → β − 0 but sv ∈ L1(R) Thus the function sγ(x) → −∞ as
x → β − 0 for all γ > 0, and hence it follows from Theorem 4.3 that the operator
Hγ has an infinite negative spectrum for all γ > 0.
In the case k > 0 we use the formula
(7.3)
and apply Corollary 4.6 with b0(ξ) = δ(ξ) and b∞(ξ) = bv(ξ). Since b0 ∈ C(R)′ and
b∞ has asymptotics (5.10), the operator Hγ has a negative spectrum for all γ 6= 0.
b(ξ) = δ(ξ) + bv(ξ)
Let us summarize the results obtained.
36
D. R. YAFAEV
Proposition 7.3. Let Hγ = C − γV where V is the Hankel operator with kernel
(7.1). Then:
10 The operator Hγ has an infinite negative spectrum for all γ > 0 if k ∈ (−1, 0).
20 The operator Hγ has negative eigenvalues for all γ 6= 0 if k > 0.
7.3. The result below directly follows from Theorem 4.3.
Proposition 7.4. Suppose that the sign-function sv(x) of a Hankel operator V is
continuous and sv(x) → 0 as x → ∞. Then the operator Hγ defined by formula
(7.2) is positive if and only if
γsv(x) ≤ 1,
∀x ∈ R.
If this condition is not satisfied, then Hγ has an infinite negative spectrum.
We note that, by Proposition 3.12, under the assumption of Proposition 7.4 on the
sign-function sv the operator V is compact. Of course this assumption is satisfied if
bv ∈ L1(R).
Example 7.5. Let v(t) = e−tr where r < 1. We have seen in subs. 5.3 that its
sign-function sv(x) = Ir(x) where Ir(x) is integral (5.13). Recall that Ir(x) is a
nonnegative continuous function of x ∈ R and Ir(x) → 0 as x → ∞. Set
νr = max
x∈R
Ir(x).
r
r
Then Hγ ≥ 0 if γ ≤ ν−1
γ > ν−1
and the operator Hγ has infinite negative spectrum for all
. Using explicit formula (5.14) it is easy to calculate ν1/2 = 3p6/πe−3/2.
In the case r > 1 we use formula (7.3). As shown in subs. 5.3, the modulus of the
function bv(ξ) exponentially grows and the periods of its oscillations tend to zero
only logarithmically as ξ → ∞. Therefore Theorem 4.5 yields the following result.
Proposition 7.6. Let v(t) = e−tr where r > 1. Then the operator Hγ has at least
one negative eigenvalue for all γ 6= 0.
Appendix A. Proof of Lemma 3.6
Set
t∈R+(cid:0)hln tiκtnF (n)(t)(cid:1)
where for shortness we use the notation hxi = (1 + x).
κ = max
Let us first consider (Φ∗F )(λ) for λ ∈ (−1, 1) =: I. We have
F (n)
√2π(Φ∗F )(λ) =Z k
0
F (t)eiλtdt +Z ∞
k
F (t)eiλtdt,
k = λ−1/2.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
37
The first integral in the right-hand side is bounded by F (0)
L1(I). In the second integral we integrate by parts
0
λ−1/2 which belongs to
Z ∞
k
F (t)eiλtdt = iλ−1F (k)eiλk + iλ−1Z ∞
k
F ′(t)eiλtdt.
(A.1)
The first term here is bounded by Cλ−1hln λi−κF (0)
The second term is bounded by
κ which belongs to L1(I) if κ > 1.
λ−1Z ∞
k
t−1hln ti−κdtF (1)
κ ≤ Cλ−1hln λi−κ+1F (1)
κ .
It belongs to L1(I) if κ > 2.
Next, we consider (Φ∗F )(λ) for λ ≥ 1. Integrating by parts, we see that
√2π(Φ∗F )(λ) = iλ−1Z k
0
F ′(t)eiλtdt + iλ−1Z ∞
k
F ′(t)eiλtdt.
(A.2)
The first term here is bounded by
λ−1Z k
0
t−1hln ti−κdtF (κ)
1 ≤ Cλ−1hln λi−κ+1F (κ)
1
.
It belongs to L1(R \ I) if κ > 2.
In the second integral in (A.2) we once more
integrate by parts, that is, we use formula (A.1) with F (t) replaced by F ′(t). The
function λ−2F ′(k) is bounded by λ−2k−1F (1)
. For the second term, we
use the estimate
0 = λ−3/2F (1)
0
(cid:12)(cid:12)λ−2Z ∞
k
F ′′(t)eiλtdt(cid:12)(cid:12) ≤ λ−2Z ∞
k
t−2dtF (2)
0 = λ−2k−1F (2)
0 = λ−3/2F (2)
0
.
Therefore the second term in (A.2) also belongs to L1(R \ I).
Appendix B. The Gaussian kernel
Here we return to the Hankel operator H with kernel h(t) = e−t2. Now we proceed
from the identity
(Hf, f ) = (Qψ, ψ),
(B.1)
where ψ(t) = e−t2f (t) and Q is the integral operator with real kernel e−2ts in the
space L2(R+). We shall use (B.1) essentially in the same way as the main identity
(1.4). Observe that operators Q with kernels q(ts) which depend only on the product
of variables can be explicitly diagonalized by the Mellin transform (see [15]). Under
fairly general assumptions on q the spectrum of Q consists of the interval [−γ, γ]
where
γ = √2π max
ξ∈R (Mq)(ξ)
38
D. R. YAFAEV
and M is the Mellin transform (2.9). In particular, for q(t) = e−2t the spectrum of
Q is absolutely continuous, simple and coincides with the interval [−pπ/2,pπ/2].
This allows us to check the following assertion.
Proposition B.1. The Hankel operator H with kernel h(t) = e−t2
number of positive and negative eigenvalues.
has infinite
1
j
, . . . , ∆(−)
Choose functions ϕ(±)
Proof. Choose some µ ∈ (0,pπ/2). For an arbitrary N,
(µ,pπ/2) and ∆(−)
and kϕ(±)
Let ϕ(±) = PN
± (Qϕ(±), ϕ(±)) = ±
N ⊂
N ⊂ (−pπ/2,−µ) be closed mutually disjoint intervals.
j k = 1, j = 1, . . . N.
, . . . , ϕ(±)
N .
be a linear combination of the functions ϕ(±)
j k2 = µkϕ(±)k2. (B.2)
αj2(Qϕ(±)
j = EQ(∆(±)
such that ϕ(±)
, . . . , ∆(+)
j=1 αjϕ(±)
j
let ∆(+)
)ϕ(±)
j
Then
, ϕ(±)
j
j
1
1
j
NXj=1
) ≥ µ
αj2kϕ(±)
NXj=1
0 (R+) such that kψ(±)
j ∈ C ∞
For an arbitrary ε > 0, we can choose ψ(±)
j −ϕ(±)
for all j = 1, . . . N. Since the functions ϕ(±)
are orthogonal, the functions ψ(±)
j
linearly independent if ε is small enough. Moreover, it follows from (B.2) that
j k < ε
are
j
±(Qψ(±), ψ(±)) ≥ 2−1µkψ(±)k2
j
j=1 αjψ(±)
and ε is small.
if ψ(±) =PN
Set now f (±)(t) = et2ψ(±)(t). Then f (±) ∈ L2(R+) and according to relation (B.1)
we have the inequality ±(Hf±, f±) > 0 on linear subspace of dimension N (except
f± = 0). Hence the Hankel operator H with kernel (1.15) has at least N positive
and N negative eigenvalues. Since N is arbitrary, this concludes the proof.
(cid:3)
We emphasize that the operator H is compact while the operator Q has the
continuous spectrum. Nevertheless the multiplicities of their positive and negative
spectra are the same (infinite).
References
[1] N. Akhiezer, The classical moment problem and some related questions in analysis, Olyver
and Boyd, Edinburgh and London, 1965.
[2] S. N. Bernstein, Sur les fonctions absolument monotones, Acta Math. 52 (1929), 1-66.
[3] T. Carleman, Sur les ´equations int´egrales singuli`eres `a noyau r´eel et sym´etrique, Almqvist
and Wiksell, 1923.
[4] A. Erd´elyi, W. Magnus, F. Oberhettinger, F. G. Tricomi, Higher transcendental functions.
Vol. 1, 2, McGraw-Hill, New York-Toronto-London, 1953.
[5] I. M. Gel'fand and N. Ya. Vilenkin, Generalized functions. Vol. 4, Academic Press, New York
and London, 1964.
[6] J. S. Howland, Spectral theory of operators of Hankel type. I, II Indiana Univ. Math. J. 41
(1992), no. 2, 409 -- 426, 427 -- 434.
[7] W. Magnus, On the spectrum of Hilbert's matrix, Amer. J. Math. 72 (1950), 405-412.
CRITERIA FOR HANKEL OPERATORS TO BE SIGN-DEFINITE
39
[8] F. G. Mehler, Uber eine mit den Kugel- und Cylinderfunctionen verwandte Function und ihre
Anwendung in der Theorie der Elektricitatsvertheilung, Math. Ann. 18 (1881), 161-194.
[9] Z. Nehari, On bounded bilinear forms, Ann. Math. 65 (1957), 153-162.
[10] V. V. Peller, Hankel operators and their applications, Springer Verlag, 2002.
[11] S. R. Power, Hankel operators on Hilbert space, Pitnam, Boston, 1982.
[12] M. Rosenblum, On the Hilbert matrix, I, II, Proc. Amer. Math. Soc. 9 (1958), 137-140, 581-585.
[13] D. V. Widder, The Laplace transform, Princeton Univ. Press, New Jersey, 1941.
[14] H. Widom, Hankel matrices, Trans. Amer. Math. Soc. 121 (1966), 1-35.
[15] D. R. Yafaev, The discrete spectrum in the singular Friedrichs model, in: Advances in Math.
Sciences, AMS, v. 189, 255-274, 1999.
[16] D. R. Yafaev, A commutator method for the diagonalization of Hankel operators, Funct. Anal.
and Appl. 44 (2010), 295-306.
[17] D. R. Yafaev, Spectral and scattering theory for perturbations of the Carleman operator,
ArXiv: 1210.5709, St.-Petersburg Math. J. 25, N 2.
[18] D. R. Yafaev, On finite rank Hankel operators, ArXiv, 2013.
IRMAR, Universit´e de Rennes I, Campus de Beaulieu, 35042 Rennes Cedex,
FRANCE
E-mail address: [email protected]
|
1804.01730 | 1 | 1804 | 2018-04-05T08:28:43 | Hypercyclic algebras | [
"math.FA"
] | We prove the existence of algebras of hypercyclic vectors in three cases: convolution operators, composition operators, and backward shift operators. | math.FA | math |
HYPERCYCLIC ALGEBRAS
FR ´ED´ERIC BAYART
Abstract. We prove the existence of algebras of hypercyclic vectors in three cases:
convolution operators, composition operators, and backward shift operators.
1. Introduction
When we work with a linear dynamical system (X, T ), it is natural to study how its linear
properties influence its dynamical properties. Here, X denotes a topological vector space
and T is a continuous linear operator on X. We are mainly interested in hypercyclic
operators: there exists x ∈ X, called a hypercyclic vector for T , such that {T nx; n ≥ 0}
is dense in X (when X is a second-countable Baire space, this amounts to saying that T
is topologically transitive). It is well known that the linear properties of (X, T ) reflects on
HC(T ), the set of hypercyclic vectors for T : HC(T )∪{0} always contains a dense subspace
([9]) and there are nice criteria to determine if it contains a closed infinite-dimensional
subspace (see [13], [14]).
In this paper we assume that X is also a topological algebra and we ask whether HC(T ) ∪
{0} contains a nontrivial algebra; we will call this a hypercyclic algebra for T . We explore
three relevant situations.
1.1. Convolution operators. Following the pioneering work of Birkhoff and MacLane,
Godefroy and Shapiro have shown in [12] that a nonconstant operator which commutes
with all translations τa is hypercyclic on H(C). Such an operator may be written φ(D),
where φ is an entire function of exponential type and D is the derivation operator. Bayart
and Matheron in [2] and independently Shkarin in [15] have shown that D admits a
hypercyclic algebra. The argument of [2], which is based on the Baire category theorem and
the fact thatSn ker(Dn) is dense in H(C), was refined by B`es, Conejero and Papathanasiou
in [5] to prove that P (D) supports a hypercyclic algebra for all nonzero polynomials P
with P (0) = 0 (see also the recent paper [11] for the existence of hypercyclic algebras for
weighted backward shifts in various Fr´echet algebras).
More recently, in the nice paper [6], the same authors provide further examples of entire
functions φ such that φ(D) admits such an algebra. For instance, this holds true for
φ(z) = cos(z) which does not satisfy φ(0) = 0 and which is not a polynomial.
In stark contrast with this, it was observed in [1] that the orbit of f 2 under λτa (which
corresponds to the case φ(z) = λeaz) can only contain functions for which the multiplicities
of their zeros is even.
Date: April 6, 2018.
The author was partially supported by the grant ANR-17-CE40-0021 of the French National Research
Agency ANR (project Front).
1
2
FR´ED´ERIC BAYART
Our first main theorem characterizes the existence of a hypercyclic algebra for φ(D) when
φ(0) < 1, or when φ(0) = 1 and φ has moderate growth.
Theorem 1.1. Let φ be a nonconstant entire function with exponential type.
(1) Assume that φ(0) < 1. Then the following assertions are equivalent:
(i) φ(D) supports a hypercyclic algebra.
(ii) φ is not a multiple of an exponential function.
(2) Assume that φ(0) = 1 and φ has subexponential growth. If either φ′(0) 6= 0 or φ
has order less than 1/2, then φ(D) supports a hypercyclic algebra.
In view of the previous result, it is tempting to conjecture that the assumption φ(0) ≤ 1
is a necessary condition for φ(D) to admit a hypercyclic algebra. Surprizingly, this is not
the case.
Theorem 1.2. Let φ(z) = 2 exp(−z) + sin(z). Then φ(D) supports a hypercyclic algebra.
Another natural conjecture is that φ(D) always suppport a hypercyclic algebra as soon
as φ is not a multiple of an exponential function. We do not know if this conjecture is
true. Nevertheless, we still get an interesting result if we weaken the conclusion. Recall
that a vector x ∈ X is a supercyclic vector for T ∈ L(X) provided {λT nx; λ ∈ C, n ≥ 0}
is dense in X. If X is a topological algebra, then any subalgebra of X consisting entirely
(but zero) of supercyclic vectors for T is said to be a supercyclic algebra.
Theorem 1.3. Let φ be a nonconstant entire function with exponential type. The following
assumptions are equivalent:
(i) φ(D) supports a supercyclic algebra.
(ii) There exists f ∈ H(C) such that, for all m ≥ 1, f m is a hypercyclic vector for φ(D).
(iii) φ is not a multiple of an exponential function.
To our knowledge, the existence of f ∈ H(C) such that f m ∈ HC(cid:0)φ(D)(cid:1) for all m ≥ 1
was only known when φ has subexponential type [3].
1.2. Composition operators. Let Ω ⊂ C be a simply connected domain and let ϕ be a
holomorphic self-map of Ω. The composition operator Cϕ(f ) = f ◦ϕ is a bounded operator
on H(Ω) which is hypercyclic if and only if ϕ is univalent and has no fixed point in Ω.
Moreover, in that case, if P is a nonconstant polynomial, then P (Cϕ) is hypercyclic (see
[4]). It was observed in [6] that Cϕ never supports a hypercyclic algebra and it was asked
whether P (Cϕ) can support such an algebra. We provide an affirmative answer.
Theorem 1.4. Let Ω ⊂ C be a simply connected domain and let ϕ be a holomorphic
self-map of Ω which is univalent and has no fixed point in Ω. Let also P be a nonconstant
polynomial which is not a multiple of z and which satisfies P (1) < 1. Then P (Cϕ)
supports a hypercyclic algebra.
1.3. Backward shift operators. So far, our examples live only in F -algebras (namely
in metrizable and complete topological algebras without assuming that the distance is
induced by a norm) and the proofs of Theorems 1.1 and 1.4 depend heavily on the
non Banach structure of the ambient space. We provide now an example in the Ba-
nach algebra ℓ1(N) = {(un)n≥0; Pn un < +∞} endowed with the convolution product
HYPERCYCLIC ALGEBRAS
3
j=0 ujvk−j. It was already sketched in [2] that, denoting by B the back-
ward shift operator, HC(2B) ∪ {0} contains a nontrivial algebra. Again, the density of
(u ⋆ v)(k) = Pk
Sn ker(Bn) was the key for the proof.
We go much further (we denote by D the open unit disk and by T its boundary the unit
circle).
Theorem 1.5. Let P ∈ C[X] be nonconstant and let B be the backward shift operator on
ℓ1(N). Then the following assertions are equivalent.
(i) P (B) is hypercyclic.
(ii) P (B) admits a hypercyclic algebra.
(iii) P (D) ∩ T 6= ∅.
1.4. Organization of the paper and strategy for the proofs. The proofs of our
results develop a method initiated in [6] and use eigenvalues and eigenvectors of our op-
erators. In Section 2, we use eigenvalues of modulus slightly bigger than 1 to prove half
of Theorem 1.1 and Theorem 1.4. In Section 3, we use eigenvalues with large modulus
to prove the remaining part of Theorem 1.1 whereas in Section 4, we use unimodular
eigenvalues to prove Theorem 1.5. There is also a significant difference between the first
two cases and the last one: the product of two eigenvectors can or cannot be still an
eigenvector. The latter situation is of course more difficult! In Section 5, we come back to
convolution operators and study the case φ(0) > 1. As a consequence we prove Theorems
1.2 and 1.3.
Of course, if we know that an operator admits a hypercyclic algebra, then it is natural to
ask how big it can be. Can it be dense? Can it be infinitely generated? These questions
were investigated very recently in [7] where it is shown that for most of the examples
exhibited in [6], one can improve the construction to get a dense and infinitely generated
hypercyclic algebra. In Section 6, we show how to modify our proofs to obtain a similar
improvement. We choose to postpone this in a separate section because the arguments
become more technical and we think that the ideas appear more clearly by handling
separately the case of singly generated algebras.
We end up this introduction with the following lemma taken from [2, Remark 8.28], which
gives a criterion for the existence of a hypercyclic algebra. It can be seen as a strong form
of the property of topological transitivity.
Lemma 1.6. Let T be a continuous operator on some separable F -algebra X. Assume
that, for any pair (U, V ) of nonempty open sets in X, for any open neighbourhood W of
zero, and for any positive integer m, one can find u ∈ U and an integer N such that
T N (un) ∈ W for all n < m and T N (um) ∈ V . Then T admits a hypercyclic algebra.
2. Small eigenvalues
2.1. A general result for operators with small eigenvalues. We shall deduce part
of Theorem 1.1 and Theorem 1.4 from the following general result.
Theorem 2.1. Let X be an F-algebra and let T ∈ L(X). Assume that there exist a func-
tion E : C → X and an entire function φ : C → C satisfying the following assumptions:
4
FR´ED´ERIC BAYART
(1) for all λ ∈ C, T E(λ) = φ(λ)E(λ);
(2) for all λ, µ ∈ C, E(λ)E(µ) = E(λ + µ);
(3) for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in X;
(4) φ is not a multiple of an exponential function;
(5) for all m ∈ N, there exist a, b ∈ C such that φ(mb) > 1 and, for all n ∈ {1, . . . , m},
all d ∈ {0, . . . , n}, with (n, d) 6= (m, m), φ(db + (n − d)a) < 1.
Then T supports a hypercyclic algebra.
We start with a lemma which explains why we have to exclude multiples of exponential
functions.
Lemma 2.2. Let φ be an entire function which is not a multiple of an exponential function.
Then, for any w0 ∈ C with φ(w0) 6= 0 and any δ > 0, there exist w1, w2 ∈ B(w0, δ),
w1 6= w2, such that the map [0, 1] → R, t 7→ log φ(tw1 + (1 − t)w2) is stricly convex.
Proof. Since φ(w0) 6= 0, there exist some neighbourhood V of w0 and a holomorphic
of an exponential function, we know that h is not an affine map. Thus there exists
w1 ∈ B(w0, δ) ∩ V such that h′′(w1) 6= 0. Without loss of generality, we assume that
function h : V → C such that φ(z) = exp(cid:0)h(z)(cid:1) for all z ∈ V . Since φ is not a multiple
w1 = 0 and we write h(z) =P+∞
log φ(z) = ℜe(cid:0)h(z)(cid:1)
= ℜe(a0) + (ℜe(a1)x − ℑm(a1)y)
k=0 akzk. Then
if z = x + iy. Since a2 6= 0, one may find (x0, y0) ∈ R2 with
+(cid:0)ℜe(a2)(cid:0)x2 − y2(cid:1) − 2ℑm(a2)xy(cid:1) + o(cid:0)x2 + y2(cid:1)
ℜe(a2)(cid:0)x2
0(cid:1) − 2ℑm(a2)x0y0 > 0.
0 − y2
Then g(t) = ℜe (h (t (x0 + iy0))) = b0 + b1t + b2t2 + o(t2) with b2 > 0 is strictly convex
around 0.
(cid:3)
Proof of Theorem 2.1. Let U, V, W be nonempty open subsets of X with 0 ∈ W and let
m ≥ 1. By Lemma 1.6, it suffices to find u ∈ U and N ∈ N so that
(1)
(2)
T N (un) ∈ W, n = 1, . . . , m − 1
T N (um) ∈ V.
The assumptions give us for this value of m two complex numbers a and b. We set
w0 = mb. We then consider δ > 0 very small and w1, w2 ∈ B(w0, δ) such that
• t ∈ [0, 1] 7→ log φ (tw1 + (1 − t)w2) is strictly convex;
• φ > 1 on [w1, w2];
• for all n ∈ {1, . . . , m}, for all d ∈ {0, . . . , n} with (n, d) 6= (m, m), for all λ1, . . . , λd ∈
[w1, w2] and all γ1, . . . , γn−d ∈ B(a, δ),
(3)
(cid:12)(cid:12)(cid:12)(cid:12)
φ(cid:18) λ1 + · · · + λd
m
+ γ1 + · · · + γn−d(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
< 1.
HYPERCYCLIC ALGEBRAS
5
We may ensure this last property because
λ1 + · · · + λd
m
+ γ1 + · · · + γn−d = db + (n − d)a + z
where the norm of z is controlled by δ.
Since B(a, δ) and [w1, w2] have accumulation points, we may find p, q ∈ N, complex
numbers a1, . . . , ap, b1, . . . , bq, complex numbers γ1, . . . , γp ∈ B(a, δ) and complex numbers
λ1, . . . , λq in [w1, w2] such that
p
q
alE (γl) ∈ U and
bjE (λj) ∈ V.
Xl=1
Xj=1
For N ≥ 1 and j ∈ {1, . . . , q}, let cj := cj(N ) be any complex number satisfying cm
j =
bj/(cid:0)φ(λj)(cid:1)N and define
u := uN =
p
q
alE (γl) +
cjE (λj/m) .
Xl=1
Xj=1
We claim that, for N large enough, uN belongs to U and satisfies (1) and (2). That uN
belongs to U is clear, since cj(N ) tends to zero as N goes to infinity. In order to prove
the other points, we need to compute un for 1 ≤ n ≤ m. To simplify the notations, let
Ip = {1, . . . , p} and for a multi-index l ∈ I d
p , al will stand for al1 · · · ald with the convention
that an empty product is equal to 1. Then we may write
un =
n
Xd=0 Xl∈I n−d
p
j∈I d
q
α(l, j, d, n)alcjE(cid:18)γl1 + · · · + γln−d +
λj1 + · · · + λjd
m
(cid:19)
for some coefficients α(l, j, d, n) that we do not try to compute, but which does not depend
on N . To prove that T N (un) belongs to W for N large enough and 1 ≤ n < m, we only
have to prove that, for any d ∈ {0, . . . , n} and any l ∈ I n−d
, any j ∈ I d
q ,
p
cj(N )(cid:18)φ(cid:18)γl1 + · · · + γln−d +
λj1 + · · · + λjd
m
(cid:19)(cid:19)N
−→N→+∞ 0.
This follows from (3) and the fact that cj(N ) tends to zero.
The case n = m is slightly different. We denote by Dq the diagonal of I m
m-uples (j, . . . , j), 1 ≤ j ≤ q. Then we decompose um into
q , namely the
um =
m−1
Xd=0 Xl∈I m−d
p
j∈I d
q
α(l, j, d, m)alcjE(cid:18)γl1 + · · · + γln−d +
λj1 + · · · + λjd
m
(cid:19)
α(j, m)cjE(cid:18) λj1 + · · · + λjm
m
(cid:19)
q \Dq
+ Xj∈I m
Xj=1
+
q
cm
j E (λj)
=: v1 + v2 + v3.
6
FR´ED´ERIC BAYART
The same considerations as above show that T N v1 tends to zero as N goes to infinity.
That T N v2 tends also to zero follows from a convexity argument. Indeed, for j ∈ I m
q \Dq,
the strict convexity of the map t 7→ log φ (tw1 + (1 − t)w2) implies that
< φ (λj1)1/m · · · φ (λjm)1/m .
Moreover,
m
φ(cid:18) λj1 + · · · + λjm
(cid:12)(cid:12)(cid:12)(cid:12)
cj(N ) ×(cid:12)(cid:12)(cid:12)(cid:12)
φ(cid:18) λj1 + · · · + λjm
m
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
N
Since the left hand side of this inequality goes to zero, we get that T N v2 tends to zero.
We conclude the proof by observing that
N
.
m
φ(cid:16) λj1 +···+λjm
(cid:17)
φ (λj1)1/m · · · φ (λjm)1/m(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ bj1/m(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
q
T N v3 =
bjE (λj) .
(cid:3)
For the applications, we emphasize two corollaries of Theorem 2.1.
Corollary 2.3. Let X be an F-algebra and let T ∈ L(X). Assume that there exist a func-
tion E : C → X and an entire function φ : C → C satisfying the following assumptions:
(1) for all λ ∈ C, T E(λ) = φ(λ)E(λ);
(2) for all λ, µ ∈ C, E(λ)E(µ) = E(λ + µ);
(3) for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in X;
(4) φ is not a multiple of an exponential function;
(5) for all ρ ∈ (0, 1), there exists w0 ∈ C with φ (w0) > 1 and, for all r ∈ (0, ρ],
φ (rw0) < 1.
Then T supports a hypercyclic algebra.
Proof. We show that Assumption (5) of Theorem 2.1 is satisfied. Let m ≥ 1 and ε ∈
(0, 1/m). Set ρ = m−1
m + mε. We get the existence of w0. We set b = w0/m and
a = εw0/m. Then, for any d ≤ m − 1 and any n ≤ m,
db + (n − d)a ≤(cid:18)m − 1
m
+ ε(cid:19) w0 ≤ ρw0
showing that φ(db + (n − d)a) < 1.
(cid:3)
Corollary 2.4. Let X be an F-algebra and let T ∈ L(X). Assume that there exist a func-
tion E : C → X and an entire function φ : C → C satisfying the following assumptions:
(1) for all λ ∈ C, T E(λ) = φ(λ)E(λ);
(2) for all λ, µ ∈ C, E(λ)E(µ) = E(λ + µ);
(3) for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in X;
(4) φ is not a multiple of an exponential function;
(5) φ(0) < 1.
Then T supports a hypercyclic algebra.
HYPERCYCLIC ALGEBRAS
7
Proof. We prove that Assumption (5) of Corollary 2.3 is satisfied. Denote by M (r) =
sup{φ(z); z = r} which is a continuous and increasing function of r satisfying M (0) < 1.
Let r0 > 0 with M (r0) = 1 and let r1 > r0 with ρr1 < r0. Any w0 ∈ C such that w0 = r1
and φ(w0) = M (r1) > 1 does the job.
(cid:3)
Remark 2.5. An operator on a Banach space cannot satisfy the assumptions of Theorem
2.1. Indeed, they imply that its spectrum is unbounded. We shall see later how it remains
possible to get a hypercyclic algebra in a Banach algebra context.
2.2. Applications to convolution operators. We now show how to deduce the first
half of Theorem 1.1 from Corollary 2.4. Thus, let φ be an entire function of exponential
type which is not a multiple of an exponential function. Then we let X = H(C) and
T = φ(D). The map E is defined by E(λ)(z) = exp(λz); it satisfies (1), (2) and (3) of
Corollary 2.4.
Corollary 2.3 may also be applied to functions satisfying φ(0) = 1. We give here two
examples which are not covered by Theorem 1.1.
Example 2.6. ([6]) Let φ(z) = cos(z). Then φ(D) supports a hypercyclic algebra.
Proof. Recall that if z = x + iy, then cos(z)2 = cos2 x + sinh2 y. Let us set ψ(t) =
cos2(2t) + sinh2(t). Using standard calculus one may prove that there exists t0 > 0 such
that ψ is decreasing on (0, t0) and increasing on (t0, +∞). Let t1 > 0 be such that
ψ(t1) = 1. It then suffices to consider w0 = (1 + η)(2 + i)t1 for some sufficiently small
η > 0.
(cid:3)
Example 2.7. Let φ(z) = ez − 2. Then φ(D) supports a hypercyclic algebra.
Proof. Let t1 > 0 be such that φ(t1) = 1. It suffices to consider w0 = (1 + η)t1 for some
sufficiently small η > 0.
(cid:3)
We now give a surprizing example of an entire function φ with φ(0) > 1 and φ(D)
supports a hypercyclic algebra.
Example 2.8. Let φ(z) = 2e−z + sin z. Then φ(D) supports a hypercyclic algebra.
Proof. We just need to show that φ satisfies Assumption (5) of Theorem 2.1. We let
a = kπ for some sufficiently large k and b = kπ + π
2m . Then
φ (db + (n − d)a) =(cid:12)(cid:12)(cid:12)(cid:12)
sin(cid:18)nkπ +
dπ
2m(cid:19) + 2 exp(cid:18)−nkπ −
.
dπ
2m(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
Provided k is large enough, this is less than 1 as soon as d < m, whereas φ(mb) > 1. (cid:3)
2.3. Applications to composition operators. In this subsection, we prove Theorem
1.4. Recall that given a simply connected domain Ω ⊂ C and ϕ a holomorphic self-map
of Ω, Cϕ is hypercyclic if and only if ϕ is univalent and without fixed points. We first
prove Theorem 1.4 when Ω = C. In that case ϕ is also entire hence ϕ is a translation
ϕ(z) = z + a, a 6= 0. Thus P (Cϕ) = φ(D), where φ(z) = P ◦ exp(az) and the result is a
particular case of Theorem 1.1.
Otherwise, by the Riemann mapping theorem, we may assume that Ω = D. We simplify
the proof by using the linear fractional model (see for instance [10]): since ϕ has no fixed
8
FR´ED´ERIC BAYART
points in D, there exists a univalent map σ : D → C and a linear fractional map ψ such
that σ ◦ ϕ = ψ ◦ σ. Moreover
• either ψ can be taken to be a dilation ψ(z) = rz for some 0 < r < 1;
• or ψ can be taken to be a translation ψ(z) = z + a for some a ∈ C\{0} and
σ(D) ⊂ {z; ℜe(z) > 0}.
The functional equation guarantees that σ(D) =: U is preserved by ψ and that P (Cϕ)
acting on H(D) and P (Cψ) acting on H(U ) are quasi-conjugate by Cσ. Since Cσ is a
multiplicative map, it is sufficient to prove that P (Cψ) admits a hypercyclic algebra (see
[6, Remark 6]).
◮ The translation case. We denote T = P (Cψ) and E(λ)(z) = exp(λz) so that T E(λ) =
φ(λ)E(λ) with φ(λ) = P (exp(aλ)). Then the assumptions of Corollary 2.4 are satis-
fied provided we are able to prove that, for any Λ ⊂ C with an accumulation point,
span (E(λ); λ ∈ Λ) is dense in H(U ). The proof is exactly similar to that for H(C), since
the polynomials are dense in H(U ) (recall that U is simply connected) - see for instance
[12].
◮ The dilation case. We still denote T = P (Cψ) but now we set E(λ) = zλ. This defines a
holomorphic function on U since U ⊂ {z; ℜe(z) > 0}. Moreover, T E(λ) = φ(λ)E(λ) with
φ(λ) = P (exp(λ log r)). Again the assumptions of Corollary 2.4 are satisfied provided that,
for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in H(U ). Let L be
a linear form on H(U ) which vanishes on span (E(λ); λ ∈ Λ). By the Riesz representation
theorem, there exists K a compact subset of U and µ a complex measure supported in
K such that, for all f ∈ H(U ), L(f ) = RK f dµ. The map λ 7→ L(zλ) is holomorphic and
has an accumulation point of zeros. Hence it is identically zero. Therefore, L vanishes
on each monomial zn, hence on H(U ) since U is simply connected. This shows that
span (E(λ); λ ∈ Λ) is dense in H(U ).
3. Large eigenvalues
As in Section 2, we shall deduce Part (2) of Theorem 1.1 from a more general statement.
Theorem 3.1. Let X be an F-algebra and let T ∈ L(X). Assume that there exist a
function E : C → X and a nonconstant entire function φ : C → C satisfying the following
assumptions:
(1) for all λ ∈ C, T E(λ) = φ(λ)E(λ);
(2) for all λ, µ ∈ C, E(λ)E(µ) = E(λ + µ);
(3) for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in X;
(4) φ(0) = 1, φ has subexponential growth and either φ′(0) 6= 0 or φ has order less
than 1/2.
Then T supports a hypercyclic algebra.
The proof of Theorem 3.1 shares many similarities with that of Theorem 2.1. Nevertheless,
we will now choose the complex numbers λj with φ (λj) very large (instead of being
slightly bigger than 1). In this way, because φ has subexponential growth, we will ensure
that φ(λj) is bigger than φ(2λj )1/2. Thus, when we will take the powers of u and apply
T N , the main term will change. We will also need a more careful interaction between the
HYPERCYCLIC ALGEBRAS
9
λj' and the γk'. This is the content of the following key lemma, which uses the fact that
we control the growth of φ.
Lemma 3.2. Let φ be a nonconstant entire function with subexponential growth and
φ(0) = 1. Assume that either φ′(0) 6= 0 or φ has order less than 1/2. Then for all
m ≥ 2, there exist z0 ∈ C\{0} and w0 = ρz0 for some ρ > 0 such that
• φ < 1 on (0, z0];
• φ(w0) > 1;
• φ(w0) > φ(dw0)1/d for all d = 2, . . . , m.
• t 7→ φ(w0 + tz0) is increasing on some interval [0, η), η > 0.
Proof. We first show the existence of z0, z1 ∈ C with z0 ∈ (0, z1), φ < 1 on (0, z0] and
φ(z1) > 1. The proof differs here following the assumptions made on φ. Assume first
that φ has order less than 1/2. Write φ(z) = eiθ0 + ρpeiθpzp + o(zp) with ρp > 0. Then,
φ(cid:0)te−i(θp−θ0+π)/p(cid:1) = eiθ0 − ρptpeiθ0 + o(tp) has modulus less than 1 provided t is small
enough. We then set z0 = te−i(θp−θ0+π)/p for some small t. We then find z1 since any
nonconstant entire function of order less than 1/2 cannot be bounded on a half-line (see
[8, Theorem 3.1.5]).
On the other hand, suppose now that φ′(0) 6= 0. Without loss of generality we may
assume that φ(z) = 1 − az + o(z) for some a > 0. Now, a nonconstant entire function
with subexponential growth cannot be bounded on a half-plane (see [8, Theorem 1.4.3]).
Thus, there exists z1 = r1eiα1 with r1 > 0 and α1 ∈ (−π/2, π/2) such that φ(z1) > 1. It
is easy to check that, for t > 0 small enough, φ(tz1) < 1 and we set z0 = tz1 for such a
small t > 0.
We now proceed with the construction of w0. Without loss of generality we may assume
z0 = 1. We set ψ(t) = φ(t)2. We proceed by contradiction and we assume that there
does not exist w0 > 0 such that the last three points of the lemma are satisfied. We fix a
n=1(1 − εn) > 0. We shall construct two sequences
sequence (εn) in (0, 1) such that κ :=Q+∞
(tn)n≥0 and (rn)n≥1 of positive real numbers such that, for all n ≥ 0,
ψ(tn) > 1, ψ′(tn) > 0
tn = rntn−1
rn ∈ (1, m]
ψ(tn) ≥ max(cid:0)ψ(tn−1)(1−εn)rn, ψ(tn−1)3/2(cid:1) .
First, the existence of z1 leads to some positive real number t0 > 1 such that ψ(t0) > 1
and ψ′(t0) > 0. Next, assume that the construction has been done until step n and let
us proceed with step n + 1. Since we assumed that the conclusion of Lemma 3.2 is false,
there exists kn+1 ∈ {2, . . . , m} such that
ψ(kn+1tn) ≥ ψ(tn)kn+1 > max(cid:16)ψ(tn)(1−εn+1)kn+1, ψ(tn)3/2(cid:17) .
If ψ′(kn+1tn) > 0, then we are done by choosing tn+1 = kn+1tn. Otherwise, let
τ := sup{t ∈ [tn, kn+1tn]; ψ′(t) > 0}.
Then
ψ(τ ) ≥ ψ(kn+1tn).
10
FR´ED´ERIC BAYART
Thus, there exists tn+1 ∈ [tn, τ ] such that ψ′(tn+1) > 0 and
Now, tn+1 = rn+1tn for some rn+1 ∈ (1, kn+1) so that
ψ(tn+1) > max(cid:16)ψ(tn)(1−εn+1)kn+1, ψ(tn)3/2(cid:17) .
ψ(tn+1) > max(cid:16)ψ(tn)(1−εn+1)rn+1, ψ(tn)3/2(cid:17)
as required to prove step n + 1. The sequence (tn) we have just built satisfies, for all
n ≤ N ,
ψ(tn) ≥ (ψ(t0))(3/2)n
In particular, (ψ (tn)), hence (tn), go to infinity. Now
.
ψ(tn) ≥ ψ(t0)Qn
k=1(1−εk)rk ≥ ψ(t0)κtn/t0.
This is a contradiction since ψ has subexponential growth.
(cid:3)
Proof of Theorem 3.1. Let w0, z0 6= 0 be given by Lemma 3.2. Then we may find γ1 ∈
(0, z0/m) which is sufficiently close to 0 so that
• φ (w0 + (m − 1)γ1) > 1;
• φ (w0 + (m − 1)γ1) > φ (dw0 + sγ1)1/d for all d ∈ {2, . . . , m} and for all s ∈
{0, . . . , m − d};
• φ (w0 + (m − 1)γ1) > φ (w0 + sγ1) for all s ∈ {0, . . . , m − 2}.
We then fix δ > 0 sufficiently small so that
• for all λ ∈ B(w0, δ), φ (λ + (m − 1)γ1) > 1;
• for all d ∈ {1, . . . , m}, for all s ∈ {0, . . . , m − d} with (d, s) 6= (1, m − 1), for all
λ, λ1, . . . λd ∈ B(w0, δ), for all z ∈ B(0, δ),
φ (λ + (m − 1)γ1) > φ (λ1 + · · · + λd + sγ1 + z)1/d .
Let p, q be integers, let a1, . . . , ap, b1, . . . , bq be complex numbers, let γ2, . . . , γp ∈ (0, z0/m)∩
B(0, δ/m) and let λ1, . . . , λq ∈ B(w0, δ) be such that
alE (γl) ∈ U
p
Xl=1
q
bjE (λj + (m − 1)γ1) ∈ V.
Xj=1
Without loss of generality we may assume a1 6= 0. For N ≥ 1 and j ∈ {1, . . . , q}, let
cj := cj(N ) be defined by
and let us set
u := uN =
p
cj =
mam−1
1
bj
(cid:0)φ (λj + (m − 1)γ1)(cid:1)N
Xl=1
alE (γl) +
Xj=1
q
cjE (λj)
HYPERCYCLIC ALGEBRAS
11
(observe that now we do not divide λj by m). As before, for N large enough, u belongs
to U . Moreover, the formula for un is similar:
n
un =
Xd=0 Xl∈I n−d
p
j∈I d
q
α(l, j, d, n)alcjE(cid:0)γl1 + · · · + γln−d + λj1 + · · · + λjd(cid:1) .
Assume first that n < m and let us show that, for all d ∈ {0, . . . , n}, for all l ∈ I n−d
all j ∈ I d
q ,
p
and
(4)
Assume first d ≥ 1 and let s = card {i; li = 1}. Then γl1 + · · · + γln−d + λj1 + · · · + λjd =
λj1 + · · · + λjd + sγ1 + z with z < δ and s ≤ m − 2. So, writing
N
cj(N ) ×(cid:12)(cid:12)φ(cid:0)γl1 + · · · + γln−d + λj1 + · · · + λjd(cid:1)(cid:12)(cid:12)
cj(N ) ×(cid:12)(cid:12)φ(cid:0)γl1 + · · · + γln−d + λj1 + · · · + λjd(cid:1)(cid:12)(cid:12)
N −→N→+∞ 0.
=
bj
mda1(m−1)d d
Yi=1
φ (λj1 + · · · + λjd + sγ1 + z)1/d
φ (λji + (m − 1)γ1)
!N
we observe that (4) is true. If d = 0 (in that case, cj(N ) = 1), (4) remains also true since
γl1 + · · · + γln ∈ (0, z0] so that φ (γl1 + · · · + γln) < 1. This yields that T N (un) tends to
zero. The case n = m requires small modifications. We now decompose um into
um = mam−1
cjE (λj + (m − 1)γ1)
q
p
1
m
+
Xj=1
Xd=0
d6=1 Xl∈I m−d
+ Xl∈I m−1
j∈I d
q
p
j∈I 1
q
α(l, j, d, m)alcjE(cid:0)γl1 + · · · + γln−d + λj1 + · · · + λjd(cid:1)
α(l, j, d, m)alcjE(cid:0)γl1 + · · · + γln−d + λj1 + · · · + λjd(cid:1)
l6=(1,...,1)
=: v1 + v2 + v3.
With exactly the same argument as above, one shows that T N (v2 + v3) tends to zero.
Furthermore,
T N (v1) =
which closes the argument.
bjE (λj + (m − 1)γ1) ∈ V,
m
Xj=1
(cid:3)
4. Unimodular eigenvalues
4.1. Proof of Theorem 1.5. In this section, we provide a proof for Theorem 1.5. As
before, P (B) admits a natural family of eigenvectors: for any λ ∈ D, (λk) is an eigenvector
of P (B) associated to P (λ). Nevertheless, we do not have a so simple formula for the
product of two eigenvectors.
12
FR´ED´ERIC BAYART
Lemma 4.1. Let Θ = (θ1, . . . , θn) ∈ Cn and let µ1, . . . , µr be pairwise distinct complex
numbers such that {θ1, . . . , θn} = {µ1, . . . , µr}. For j = 1, . . . , r, let κj = card{k; θk =
µj}. Then there exist polynomials PΘ,j, 1 ≤ j ≤ r, with degree less than or equal to κj − 1
such that
1(cid:17) ⋆ · · · ⋆(cid:16)θk
(cid:16)θk
n(cid:17) =
When all the θi are equal to the same λ, then
r
j(cid:17) .
Xj=1(cid:16)PΘ,j(k)µk
where deg(Pn) = n − 1.
(cid:16)λk(cid:17) ⋆ · · · ⋆(cid:16)λk(cid:17) =(cid:16)Pn(k)λk(cid:17)
The statement of this lemma motivates the study of the effect of P (B)N on the vectors
(kdλk).
Lemma 4.2. Let P ∈ C[X] and let d ≥ 0. Let also λ ∈ D with λP (λ)P ′(λ) 6= 0. There
exist complex numbers (Ad,N,s)N ≥0, 0≤s≤d, such that, for all N ≥ 0,
(5)
(cid:0)P (B)(cid:1)N(cid:16)kdλk(cid:17) =
d
Xs=0
P (λ)N +s−dAd,N,s(cid:16)ksλk(cid:17)
with Ad,N,s ∼N→+∞ ωd,sN d−s for some nonzero ωd,s.
We point out that in the statement of the previous lemma, the complex number ωd,s may
depend on λ; later we will sometimes denote them ωd,s(λ).
We will also need a density lemma.
Lemma 4.3. Let Λ ⊂ D with an accumulation point inside D. Then(cid:8)(cid:0)λk(cid:1) ; λ ∈ Λ(cid:9) spans
a dense subspace of ℓ1(N).
We postpone the proof of these lemmas to give that of Theorem 1.5.
Proof of Theorem 1.5. Since the spectrum of a hypercyclic operator has to intersect the
unit circle, if P (D) is hypercyclic, then P (D) ∩ T 6= ∅. Hence the only difficult implication
is (iii) =⇒ (ii). Thus we start with a nonconstant polynomial satisfying P (D) ∩ T 6= ∅.
Let Λ1 ⊂ D with an accumulation point in D such that P (λ) = 1 and λP ′(λ) 6= 0 for
all λ ∈ Λ1. Let also Λ2 ⊂ D with an accumulation point in D such that P (λ) < 1 and
λP ′(λ) 6= 0 for all λ ∈ Λ2.
Let U, V, W be nonempty open subsets of ℓ1 with 0 ∈ W . Let m ≥ 1. We may find
p, q ∈ N, complex numbers γ1, . . . , γp ∈ Λ2, λ1, . . . , λq ∈ Λ1, a1, . . . , ap, b1, . . . , bq such that
p
Xl=1
al(cid:16)γk
l(cid:17) ∈ U and
q
Xj=1
bj(cid:16)λk
j(cid:17) ∈ V.
We then set, for j = 1, . . . , q, N ≥ 0,
cj := cj(N ) =(cid:18)
ωm−1,0(λj)N m−1P (λj)N −m+1(cid:19)1/m
bj
HYPERCYCLIC ALGEBRAS
13
(we take any m-th root) and
u := uN =
p
Xl=1
al(cid:16)γk
l(cid:17) +
q
Xj=1
j(cid:17)
cj(cid:16)λk
so that, if N is large enough, u belongs to U . As usual, for n ∈ N,
un =
n
Xd=0 Xl∈I n−d
p
j∈I d
q
α(l, j, d, n)alcj(cid:16)γk
l1(cid:17) ⋆ · · · ⋆(cid:16)γk
ln−d(cid:17) ⋆(cid:16)λk
jd(cid:17) .
j1(cid:17) ⋆ · · · ⋆(cid:16)λk
Let us fix for a while n ≤ m, d ∈ {0, . . . , n}, l ∈ I n−d
p
and j ∈ I d
q . Applying Lemma
4.1, we observe that (cid:0)γk
µ ∈ {γl1, . . . , λjd} and s ≤ card {l; γl = µ} + card {j; λj = µ} − 1. Moreover, by Lemma
4.2,
jd(cid:17) writes as a linear combination of (cid:0)ksµk(cid:1) for some
l1(cid:1) ⋆ · · · ⋆(cid:16)λk
P (B)N(cid:16)ksµk(cid:17)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
cj(N ) ×(cid:13)(cid:13)(cid:13)
P (B)N(cid:16)ksµk(cid:17)(cid:13)(cid:13)(cid:13)
≤ CN sP (µ)N .
≤ C
N s
.
N d× m−1
m
In particular, if µ ∈ (cid:8)γl1, . . . , γln−d(cid:9), so that P (µ) < 1, then cj(N ) ×(cid:13)(cid:13)P (B)N(cid:0)ksµk(cid:1)(cid:13)(cid:13)
tends to zero. If µ ∈ {λj1, . . . , λjd}, then
(6)
Assume first that n < m. Then
s
d
≤
d − 1
d
≤
n − 1
n
<
m − 1
m
.
Hence, the right hand side of (6) goes to zero as N tends to +∞. This implies in particular
that P (B)N (un) goes to zero as N tends to +∞. Assume now that n = m. The same
we find that
j1(cid:17) ⋆ · · · ⋆(cid:16)λk
argument shows that cj(N ) ×(cid:13)(cid:13)P (B)N(cid:0)ksµk(cid:1)(cid:13)(cid:13) tends to zero, except if s = d − 1 = m − 1.
jd(cid:17)(cid:17) tends to zero except for
ln−d(cid:17) ⋆(cid:16)λk
Namely, cj(N )P (B)N(cid:16)(cid:0)γk
the terms cj(N )P (B)N(cid:16)(cid:16)λk
j(cid:17)(cid:17). Applying again Lemma 4.1 and Lemma 4.2,
j(cid:17) ⋆ · · · ⋆(cid:16)λk
cjP (B)N(cid:16)(cid:16)λk
l1(cid:1) ⋆ · · · ⋆(cid:16)γk
j(cid:17) ⋆ · · · ⋆(cid:16)λk
j(cid:17)(cid:17) = P (λ)N −m+1ωm−1,0(λj)N m−1(λk
j(cid:17) ⋆ · · · ⋆(cid:16)λk
P (B)N(cid:16)(cid:16)λk
j(cid:17)(cid:17) = bj(λk
Hence, by the definition of cj,
This achieves the proof that P (B)N (um) belongs to V provided N is large enough.
j ) + o(N m−1).
j ) + o(1).
(cid:3)
4.2. Proof of Lemma 4.1. The proof of Lemma 4.1 relies on the following facts and an
easy induction.
Fact 1. For any d ≥ 0, for any λ ∈ D, there exists a polynomial Qd with deg(Qd) = d + 1
such that
(cid:16)kdλk(cid:17) ⋆(cid:16)λk(cid:17) =(cid:16)Qd(k)λk(cid:17) .
For the induction step, we write using Fact 1
k
uk =
λ
µ
λjµk−j =
µ − λ(cid:16)µk(cid:17) .
λ − µ(cid:16)λk(cid:17) +
Xj=0
(cid:16)kd+1λk(cid:17) = a(cid:16)kdλk(cid:17) ⋆(cid:16)λk(cid:17) +(cid:16)P (k)λk(cid:17)
14
FR´ED´ERIC BAYART
Fact 2. For any d ≥ 0, for any λ, µ ∈ D with λ 6= µ, there exist a polynomial Qd,λ,µ with
deg(Qd,λ,µ) ≤ d and a complex number Bd,λ,µ such that
(cid:16)kdλk(cid:17) ⋆(cid:16)µk(cid:17) =(cid:16)Qd,λ,µ(k)λk(cid:17) + Bd,λ,µ(cid:16)µk(cid:17) .
The proof of Fact 1 is easy. Denoting (cid:0)kdλk(cid:1) ⋆(cid:0)λk(cid:1) by (uk), we have
k
uk =
jdλk
Xj=0
so that the result is proved with Qd the polynomial of degree d + 1 such that Qd(k) =
j=0 jd for all k ∈ N.
Pk
The proof of Fact 2 can be done by induction on d. For d = 0, we simply write(cid:0)λk(cid:1) ⋆(cid:0)µk(cid:1)
as (uk) with
with deg(P ) ≤ d (to simplify the notations, we do not write the subscripts on the complex
numbers and on the polynomials involved, but they clearly depend on d, λ, and later on
µ). Thus,
We then apply the induction hypothesis to both (cid:0)kdλk(cid:1) ⋆(cid:0)µk(cid:1) and (cid:0)P (k)λk(cid:1) ⋆(cid:0)µk(cid:1) to get
(cid:16)kd+1λk(cid:17) ⋆(cid:16)µk(cid:17) = a(cid:16)λk(cid:17) ⋆(cid:16)(cid:16)kdλk(cid:17) ⋆(cid:16)µk(cid:17)(cid:17) +(cid:16)P (k)λk(cid:17) ⋆(cid:16)µk(cid:17) .
(cid:16)kd+1λk(cid:17) ⋆(cid:16)µk(cid:17) = a(cid:16)λk(cid:17) ⋆(cid:16)(cid:16)Q(k)λk(cid:17) + b(cid:16)µk(cid:17)(cid:17) +(cid:16)R(k)λk(cid:17) + c(cid:16)µk(cid:17)
with deg(Q), deg(R) ≤ d and b, c ∈ C. We conclude by using again either Fact 1 or the
case d = 0.
4.3. Proof of Lemma 4.2. We first isolate the case N = 1.
Lemma 4.4. Let P ∈ C[X], d ≥ 0. There exist polynomials (Qd,s), 0 ≤ s ≤ d, such that,
for all λ ∈ D,
P (B)(cid:16)kdλk(cid:17) =
Qd,s(λ)(cid:16)ksλk(cid:17)
d
Xs=0
where Qd,d(λ) = P (λ) and Qd,d−1(λ) = dλP ′(λ).
Proof. We start from
Bn(cid:16)kdλk(cid:17) =(cid:16)(k + n)dλk+n(cid:17) = λn(cid:16)Pn,d(k)λk(cid:17)
where Pn,d is a monic polynomial of degree d. More precisely, Pn,d(k) = kd + dnkd−1 + · · · .
The result follows now from a linear combination of these equalities. Observe that if
(cid:3)
j=1 αjdjλj = dλP ′(λ).
P (X) =Pn
j=0 αjX j, then Qd,d−1(λ) =Pn
HYPERCYCLIC ALGEBRAS
15
Proof of Lemma 4.2. We first prove by induction on N that the relation (5), which is
clearly true for N = 0, holds for all N and we get an induction formula for the complex
numbers Ad,N,s. Indeed, assuming (5) is true for N and using Lemma 4.4, we have
(P (B))N +1 (kdλk) =
=
r
P (λ)N +r−dAd,N,r
Xs=0
Qr,s(λ)(cid:16)ksλk(cid:17)
P (λ)N +r−dAd,N,rQr,s(λ)(cid:16)ksλk(cid:17) .
Xr=s
d
Thus, (5) is true for N + 1 with the induction formula
(7)
Ad,N +1,s =
P (λ)r−s−1Ad,N,rQr,s(λ).
d
d
Xr=0
Xs=0
d
Xr=s
When s = d, using Qd,d(λ) = P (λ), this formula simply writes Ad,N +1,d = Ad,N,d so that
Ad,N,d = 1 for all N . When s = d−1, using Qd−1,d−1(λ) = P (λ) and Qd,d−1(λ) = dλP ′(λ),
we have
Ad,N +1,d−1 = dλP ′(λ)Ad,N,d + Ad,N,d−1
so that Ad,N,d−1 = N dλP ′(λ). Assume now that we have shown that Ad,N,r ∼N→+∞
ωd,rN d−r for r = s + 1, . . . , d and let us prove it for s. Rewriting (7) we get
d
Ad,N +1,s = Ad,N,s + (s + 1)λP ′(λ)Ad,N,s+1 +
P (λ)r−s−1Ad,N,rQr,s(λ).
Xr=s+2
We sum these equalities and use that Ad,N,r ∼N→+∞ ωd,rN d−r for r ≥ s + 1 to get
Ad,N,s = (s + 1)λP ′(λ)
The result follows now easily.
N
Xn=0
Ad,n,s+1 + O N
Xn=0
nd−(s+2)! .
(cid:3)
4.4. Proof of Lemma 4.3. Let u ∈ ℓ∞(N) which is orthogonal to all (cid:0)λk(cid:1) for λ ∈ Λ and
let F (λ) = hu,(cid:0)λk(cid:1)i. Then F is a holomorphic function in D with an accumulation point
of zeros inside D. Therefore, F and u are zero, which means that (cid:8)(cid:0)λk(cid:1) ; λ ∈ Λ(cid:9) spans a
dense subspace in ℓ1(N).
5. Convolution operators, the case φ(0) > 1
Example 2.8 relies clearly on the periodicity of the zeros of the sine function. We do not
know what happens for other examples of entire functions with φ(0) > 1, for instance for
φ(z) = ez − λ with λ − 1 > 1. Nevertheless, we have a general result for the existence of
powers of hypercyclic vectors.
Theorem 5.1. Let X be an F-algebra and let T ∈ L(X). Assume that there exist a func-
tion E : C → X and an entire function φ : C → C satisfying the following assumptions:
(1) for all λ ∈ C, T E(λ) = φ(λ)E(λ);
(2) for all λ, µ ∈ C, E(λ)E(µ) = E(λ + µ);
(3) for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in X;
16
FR´ED´ERIC BAYART
(4) φ is not a multiple of an exponential function.
Then there exists a residual set of vectors u ∈ X such that, for all m ≥ 1, um ∈ HC(T ).
Proof. Let us set E = {u ∈ X; ∀m ≥ 1, um ∈ HC(T )}. Fixing (Vj)j≥1 a basis of open
subsets of X, we set
Oj,m =(cid:8)u ∈ X; ∃N ≥ 1, T N (um) ∈ Vj(cid:9) .
Then E =Tj,m≥1 Oj,m so that, since each Oj,m is clearly open, one just has to prove that
these sets are dense. Thus, let U, V be nonempty open subsets of X and let m ≥ 1. We
are looking for a vector u ∈ U and for an integer N ≥ 1 such that T N (um) ∈ V . Let
a ∈ C be such that φ(a) < 1. Arguing as in the proof of Corollary 2.4, we may find
w0 ∈ C with φ(w0) > 1 and, for any z ∈ C with z − a ≤ m−1
m w0 − a, then φ(z) < 1.
Then we fix δ > 0 sufficiently small so that φ(z) < 1 if z − a ≤ m−1
m w0 − a + δ and
φ(w) > 1 if w ∈ B(w0, δ). Let finally, as usual(!), w1, w2 ∈ B(w0, δ) such that t ∈ [0, 1] 7→
log φ (tw1 + (1 − t)w2) is strictly convex. One may find p, q ∈ N, complex numbers
a1, . . . , ap, b1, . . . , bq, complex numbers γ1, . . . , γp ∈ B(a, δ) and λ1, . . . , λq ∈ [w1, w2] such
that
p
q
Xl=1
alE (γl/m) ∈ U and
bjE (λj) ∈ V.
Xj=1
For N ≥ 1 and j ∈ {1, . . . , q}, let cj := cj(N ) be any complex number satisfying cm
j =
bj/(cid:0)φ(λj)(cid:1)N and define
u := uN =
alE (γl/m) +
p
Xl=1
cjE (λj/m)
q
Xj=1
so that
m
um =
Xd=0 Xl∈I m−d
p
j∈I d
q
α(l, j, d, m)alcjE(cid:18) γl1 + · · · + γlm−d + λj1 + · · · + λjd
m
(cid:19) .
We write γl = a + zl with zl < δ and λj = w0 + z′
j with z′
j < δ. Then
γl1 + · · · + γlm−d + λj1 + · · · + λjd
m
=
(m − d)a + dw0
m
+ Z
with Z < δ. Moreover, provided m < d,
=
m
m
d
m
w0 − a ≤
(m − d)a + dw0
− a(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
φ(cid:16) γl1 +···+γlm−d +λj1 +···+λjd
(cid:17)(cid:12)(cid:12)(cid:12)
cj(N )T N E(cid:18) γl1 + · · · + γlm−d + λj1 + · · · + λjd
< 1 and
m
Therefore, (cid:12)(cid:12)(cid:12)
m − 1
m
w0 − a .
(cid:19) −→N→+∞ 0.
When d = m, we conclude exactly as we have done before.
(cid:3)
HYPERCYCLIC ALGEBRAS
17
Proof of Theorem 1.3. That (i) or (ii) implies (iii) is already contained in [1]. The proof
of (iii) implies (i) is easy if we observe that, for any λ ∈ C\{0}, a hypercyclic algebra
for λφ(D) is a supercyclic algebra for φ(D). Finally, the implication (iii) implies (ii) is a
consequence of Theorem 5.1 for T = φ(D).
(cid:3)
6. Infinitely generated and dense hypercyclic algebras
6.1. Notations. We use several specific notations for this section. We denote by N the
set of nonnegative integers and by N(∞) the set of sequences (α1, α2, . . . ) with αi ∈ N for
i=1 αi. If A is a
all i and αi = 0 for all large i. For α ∈ N(∞), α stands for the sum P+∞
finite subset of N(∞), A 6= ∅, then L(A) denotes sup{α; α ∈ A}.
For f ∈ X N and α ∈ N(∞), with αi = 0 for i > d, the notation f α simply means the
product f α1
d . If A is a finite subset of N(∞)\{(0, . . . )}, we will often consider it as
1
a subset of some Nd, since we may choose d ≥ 1 such that αi = 0 for all i ≥ d + 1 and all
α ∈ A.
· · · f αd
6.2. A criterion `a la Birkhoff for the existence of a dense and infinitely gener-
ated hypercyclic algebra. To prove the existence of a dense and infinitely generated
algebra of hypercyclic vectors, we need a reinforcement of Lemma 1.6. This is achieved
by the following natural proposition, which simplifies a statement of [7] since it does not
use the notion of pivot.
Proposition 6.1. Let T be a continuous operator on the separable F -algebra X. Let
≺ be a total order on N(∞). Assume that, for any d ≥ 1, for any finite and nonempty
subset A ⊂ N(∞)\{(0, . . . )}, for any nonempty open subsets U1, . . . , Ud, V of X, for any
neighbourhood W of 0, there exist u = (u1, . . . , ud) ∈ U1 × · · · × Ud and N ≥ 1 such that,
setting β = max(α; α ∈ A),
T N (uβ) ∈ V
T N (uα) ∈ W for all α ∈ A, α 6= β.
Then T admits a dense and not finitely generated hypercyclic algebra.
Proof. Let (Vk) be a basis of open neighbourhoods of X. For A ⊂ N(∞)\{(0, . . . )}, A 6= ∅,
A finite, for s, k ≥ 1, define
P (α)zα; P (βA) = 1 and P (α) ≤ s)
E(A, s) =(P (z) = Xα∈A
A(A, s, k) =nf ∈ X N; ∀P ∈ E(A, s), ∃N ≥ 1, T N (P (f )) ∈ Vko
where βA = max(α; α ∈ A). The set A being fixed, A may be considered as a subset
of Nd and E(A, s) as a subset of C[X1, . . . , Xd]. Moreover this set E(A, s) is compact.
By continuity of the maps (f, P ) 7→ T N (P (f )), this implies that each set A(A, s, k) is
open. Moreover, the assumptions of the theorem clearly imply that each such set is dense.
Observe also that the set of f in X N that induce a dense algebra in X is residual in X N
Hence, G :=TA,s,k A(A, s, k) is a residual subset of X N.
(see [7]). Hence we may pick f ∈ X N belonging to TA,s,k A(A, s, k) and inducing a dense
algebra in X.
18
FR´ED´ERIC BAYART
We show that for all nonzero polynomials P , P (f ) belongs to HC(T ). Let A be the
spectrum of P , let β = max(α; α ∈ A), let Q = 1
P and let s ≥ 1 be such that
P (β)
Q ∈ E(A, s). Since f belongs to Tk A(A, s, k), we conclude that Q(f ), hence P (f ), is a
hypercyclic vector for T .
It remains to show that the algebra generated by f is not finitely generated. Assume
on the contrary that it is generated by a finite number of f α(1), . . . , f α(p). In particular,
it is generated by a finite number of f1, . . . , fq. Then there exists a polynomial Q ∈
C[z1, . . . , zq] such that fq+1 = Q(f1, . . . , fq). Define P (z) = zq+1 − Q(z). Then P is a
nonzero polynomial. Nevertheless, P (f ) = 0, which contradicts the fact that P (f ) is a
hypercyclic vector for T .
(cid:3)
let
Remark 6.2. The last part of the proof may be formulated in the following way:
f ∈ X N be such that P (f ) is never zero for any nonzero polynomial P . Then the algebra
generated by f is not finitely generated. We could avoid this by using the following lemma,
proved in [7]: the set of sequences f ∈ X N whose induced algebra is not finitely generated
is residual in X. Nevertheless, the proof of this last statement seems more complicated.
6.3. On the existence of infinitely generated hypercyclic algebras. We now show
how to adapt our proofs to the existence of a dense and infinitely generated algebra. We
have to pay the price of additional technical difficulties and we restrict ourselves to an
analogue of Corollary 2.3.
Theorem 6.3. Let X be an F-algebra and let T ∈ L(X). Assume that there exist a func-
tion E : C → X and an entire function φ : C → C satisfying the following assumptions:
(1) for all λ ∈ C, T E(λ) = φ(λ)E(λ);
(2) for all λ, µ ∈ C, E(λ)E(µ) = E(λ + µ);
(3) for all Λ ⊂ C with an accumulation point, span (E(λ); λ ∈ Λ) is dense in X;
(4) φ is not a multiple of an exponential function;
(5) for all ρ ∈ (0, 1), there exists w0 ∈ C with φ (w0) > 1 and, for all r ∈ (0, ρ],
φ (rw0) < 1.
Then T supports a hypercyclic algebra which is dense and is not finitely generated.
Proof. We intend to apply Proposition 6.1. Thus, let d ≥ 1, let A be a finite and nonempty
subset of N(∞)\{(0, . . . )}. Enlarging d if necessary, we may and shall assume that A ⊂ Nd.
We choose for total order on Nd the lexicographical order and we denote by β the maximal
element of A for this order. Without loss of generality, we assume that β1 6= 0. Let also
U1, . . . , Ud, V be nonempty open subsets of X and let W be an open neighbourhood of 0.
We set
Iβ = {i ∈ {2, . . . , d}; βi 6= 0}
ΩA = {α ∈ A; α1 = β1} \{β}
∪nα ∈ Nd; ∀i ∈ Iβ, αi ≤ βi and ∃i ∈ Iβ, αi < βio .
We first consider the case Iβ 6= ∅ (the other case, which is easier, will be discussed at the
end of the proof). Observe that, for any α ∈ ΩA, there exists i0 ∈ Iβ such that, for all
HYPERCYCLIC ALGEBRAS
19
i ≤ i0, αi = βi and αi0 < βi0. Therefore it is easy to construct a sequence (ρi)i∈Iβ ⊂ (0, 1)
satisfying
Let η > 0 be such that
Xi∈Iβ
∀α ∈ ΩA, Xi∈Iβ
∀α ∈ ΩA, Xi∈Iβ
ρi = 1
ρi ×
αi
βi
< 1.
ρi ×
αi
βi
≤ 1 − η.
We finally choose ε > 0 and ρ ∈ (0, 1) satisfying
ρ > (1 − ε)
β1 − 1
β1
+ L(A)ε
ρ > (1 − ε) + (1 − η)ε = 1 − ηε.
This is possible for instance by setting ρ = 1 − ηε/2 for a sufficiently small ε > 0. For this
value of ρ, we get w0 ∈ C with φ(w0) > 1 and φ(rw0) < 1 if r ∈ (0, ρ]. We set κ = εw0
and z0 = (1 − ε)w0 = w0 − κ and we summarize some properties of w0, z0 and κ below:
(8)
(9)
(10)
φ(w0) > 1
φ(z0 + rκ) < 1 if r ∈ [0, 1 − η]
φ(tz0 + sκ) < 1 if t ≤
β1 − 1
β1
and s ≤ L(A).
By continuity, there exists δ > 0 such that these properties remain valid respectively in
B(w0, δ), B(z0 +rκ, δ), B(tz0 +sκ, δ). Let (as usual!) w1 6= w2 in B(w0, δ/2) such that t 7→
log φ(tw1+(1−t)w2) is stricly convex on [0, 1]. We then choose integers p, q, complex num-
bers a1,i, . . . , api,i, b1, . . . , bq, complex numbers γ1,i, . . . , γpi,i ∈ B(0, δ/2L(A)) ∩ (0, ρw0),
complex numbers λ1, . . . , λq in [w1, w2] such that
∀i ∈ {1, . . . , d},
pi
Xl=1
al,iE(γl,i) ∈ Ui and
bjE(λj) ∈ V.
q
Xj=1
We define for j = 1, . . . , q, zj = λj − κ. We then set
u1(N ) =
al,1E(γl,1) +
cj(N )E(zj /β1),
q
Xj=1
ui =
al,iE(γl,i) + ωE(ρiκ/βi),
p1
Xl=1
Xl=1
pi
ui =
pi
Xl=1
al,iE(γl,i)
provided i ∈ Iβ,
20
FR´ED´ERIC BAYART
otherwise, namely if i /∈ Iβ and i > 1. Above, ω is any positive real number small enough
so that all ui belong to Ui for i ≥ 2, and cj(N ) is any complex number such that
cj(N )β1 =
bj
.
Pi∈Iβ
βi
(cid:0)φ(λj)(cid:1)N ω
Since cj(N ) goes to zero, u1(N ) belongs to U1 for N large enough. It remains to show that
T N (uβ) ∈ V and T N (uα) ∈ W for α ∈ A\{β} and N large enough. Let us first compute
uβ. Let us examine uβ1
1 . As in the proof of Theorem 2.1, we can distinguish three different
kinds of terms:
• the terms cj(N )β1E(zj ), for j = 1, . . . , q;
• the terms cj(N )E(cid:16) zj1 +···+zjβ1
• a finite number of terms a(j, γ, N )E(cid:16) zj1 +···+zjl
(cid:17) for j ∈ I β1
γ < β1δ/2L(A).
β1
β1
q nondiagonal;
+ γ(cid:17) with j ∈ I l
q, l < β1 and
q for l > 0) but that they are uniformly bounded in N .
Observe that the terms a(j, γ, N ) may depend on N by involving cj(N ) (this happens if
j ∈ I l
Let us now inspect uβ2
Pi∈Iβ
d . For this product, we distinguish two kinds of terms:
• the term ω
• a finite number of terms b(r, γ′)E(rκ + γ′) with 0 ≤ r ≤ 1 − η and γ′ < (β2 +
αiρi/βi for some α ∈ ΩA, which
2 · · · uβd
βiE(κ);
· · · + βd)δ/2L(A) ≤ δ/2. Here, r is equal to Pi∈Iβ
explains why r ≤ 1 − η.
Thus, taking the product uβ, we get six kinds of terms:
• the terms cj(N )β1ω
Pi∈Iβ
βiE(λj) for j = 1, . . . , q. But the choice of cj(N ) ensures
that
cj(N )β1ω
Pi∈Iβ
q
T N
Xj=1
βiE(λj )
=
bjE(λj).
q
Xj=1
• the terms cj(N )β1b(r, γ′)E(zj + rκ + γ′), with 1 ≤ j ≤ q, 0 ≤ r ≤ 1 − η and
γ′ < δ/2. Since zj + rκ + γ′ − (z0 + rκ) < δ, we deduce from (9) that
T N(cid:16)cj(N )β1b(r, γ′)E(zj + rκ + γ′)(cid:17) →N→+∞ 0.
• the terms cj(N )ω
Pi∈Iβ
βiE(cid:18) λj1 +···+λjβ1
β1
(cid:19) for j ∈ I β1
proof of Theorem 2.1, the choice of [w1, w2] ensures that for all j ∈ I β1
q nondiagonal,
q nondiagonal. But as in the
T N(cid:18)cj(N )ω
Pi∈Iβ
βiE(cid:18) λj1 + · · · + λjβ1
β1
(cid:19)(cid:19) →N→+∞ 0.
• the terms cj(N )b(r, γ′)E(cid:16) zj1 +···+zjβ1
β1
+ rκ + γ′(cid:17). As before, (9) ensures that
T N(cid:18)cj(N )b(r, γ′)E(cid:18) zj1 + · · · + zjβ1
β1
+ rκ + γ′(cid:19)(cid:19) →N→+∞ 0.
HYPERCYCLIC ALGEBRAS
21
• the terms a(j, γ, N )ω
But writing
Pi∈Iβ
βiE(cid:16) zj1 +···+zjl
β1
+ κ + γ(cid:17) with l < β1, γ < β1δ/2L(A).
zj1 + · · · + zjl
β1
+ κ =
l
β1
z0 + κ + γ′
with γ′ < δ, we deduce from (10) that
T N(cid:18)a(j, γ, N )ω
Pi∈Iβ
βiE(cid:18) zj1 + · · · + zjl
β1
+ κ + γ(cid:19)(cid:19) →N→+∞ 0.
and r ∈ [0, 1 − η]. A similar proof shows that
• the terms a(j, γ, N )b(r, γ′)E(cid:16) zj1 +···+zjl
T N(cid:18)a(j, γ, N )b(r, γ′)E(cid:18) zj1 + · · · + zjl
+ γ + rκ + γ′(cid:17) with l < β1, γ+γ′ < δ/2
+ γ + rκ + γ′(cid:19)(cid:19) →N→+∞ 0.
β1
β1
Hence, as expected, provided N is large enough, T N (u(N )β) ∈ V . Let us now consider
α ∈ A with α ≺ β and let us show that T N (u(N )α) goes to zero as N goes to +∞. The
analysis is similar but simpler. Either α1 = β1 and in that case in uα2
d appear only
terms like b(r, γ)E(rκ + γ) with 0 ≤ r ≤ 1 − η and γ < (α2 + · · · + αd)δ/2L(A). Here,
i ≤ αi so that r ≤ 1 − η since α ∈ ΩA. Then applying T N to
α′
iρi/βi for some α′
2 · · · uαd
r =Pi∈Iβ
each term of uα will lead to a sequence going to zero.
Or α1 < β1, and now in uα1
1 appear only terms like
a(j, l, N )E(cid:18) zj1 + · · · + zjt
with t ≤ α1, namely terms like a(j, γ, N )E(cid:16) t
product uα2
+ γl1,1 + · · · + γlα1−t,1(cid:19)
z0 + γ(cid:17) with γ < δ/2. Expanding the
leads to terms like b(r, γ′)E(rκ + γ′) with γ′ < δ/2 and
β1
β1
2 · · · uαd
d
ρi
αi
βi
r ≤ Xi∈Iβ
≤ L(A).
A last application of (10) shows that
T N(cid:18)a(l, γ, N )b(r, γ′)E(cid:18) t
β1
z0 + rκ + γ + γ′(cid:19)(cid:19) →N→+∞ 0.
We need finally to consider the case Iβ = ∅. In that case, the proof of Theorem 2.1 works
almost mutatis mutandis. Indeed, the situation is simplified because now uβ = uβ1
1 . We
then set
u1(N ) =
ui =
p1
pi
Xl=1
Xl=1
al,1E(γl,1) +
q
Xj=1
β1(cid:19)
cj(N )E(cid:18) λj
al,iE(γl,i), i = 2, . . . , d
with cj(N )β1 = bj/ (φ(λj))N and we follow a completely similar proof. Observe in partic-
ular that if α ∈ A satisfies α ≺ β, then α1 < β1.
(cid:3)
22
FR´ED´ERIC BAYART
References
[1] R. M. Aron, J. A. Conejero, A. Peris, and J. B. Seoane-Sep´ulveda. Powers of hypercyclic functions
for some classical hypercyclic operators. Integral Equations Operator Theory, 58:591–596, 2007.
[2] F. Bayart and ´E. Matheron. Dynamics of linear operators, volume 179 of Cambridge Tracts in Math.
Cambridge University Press, 2009.
[3] L. Bernal-Gonz´alez. On universal functions with zero-free derivatives. Arch. Math., 68:145–150, 1997.
[4] J. B`es. Dynamics of composition operators with holomorphic symbol. Rev. Real Acad. Ciencias Exactas
Fisicas y Naturales Serie A (RACSAM), 107:437–449, 2013.
[5] J. B`es, A. Conejero, and D. Papathanasiou. Convolution operators supporting hypercyclic algebras.
J. Math. Anal. Appl., 445:1232–1238, 2017.
[6] J. B`es, A. Conejero, and D. Papathanasiou. Hypercyclic algebras for convolution and composition
operators. ArXiv e-prints, June 2017.
[7] J. B`es and D. Papathanasiou. Algebrable sets of hypercyclic vectors for convolution operators. ArXiv
e-prints, June 2017.
[8] R. P. Boas, Jr. Entire functions. Academic Press Inc., New York, 1954.
[9] P. S. Bourdon. Invariant manifolds of hypercyclic vectors. Proc. Amer. Math. Soc., 118:845–847, 1993.
[10] P. S. Bourdon and J. H. Shapiro. Cyclic phenomena for composition operators. Mem. Amer. Math.
Soc., 125(596), 1997.
[11] J. Falc´o and K-G. Grosse-Erdmann. Algebrability of the set of hypercyclic vectors for backward shift
operators. Preprint, 2018.
[12] G. Godefroy and J. H. Shapiro. Operators with dense, invariant, cyclic vector manifolds. J. Funct.
Anal., 98:229–269, 1991.
[13] M. Gonz´alez, F. Le´on-Saavedra, and A. Montes-Rodr´ıguez. Semi-Fredholm theory: hypercyclic and
supercyclic subspaces. Proc. London Math. Soc. (3), 81:169–189, 2000.
[14] Q. Menet. Hypercyclic subspaces and weighted shifts. Adv. in Math., 255:305–337, 2014.
[15] S. Shkarin. On the set of hypercyclic vectors for the differentiation operator. Israel J. Math., 180:271–
283, 2010.
Universit´e Clermont Auvergne, CNRS, LMBP, F-63000 Clermont–Ferrand, France.
E-mail address: [email protected]
|
1508.02036 | 2 | 1508 | 2018-07-18T00:06:41 | Functional calculus for $C_{0}$-groups using (co)type | [
"math.FA",
"math.NA"
] | We study the functional calculus properties of generators of $C_{0}$-groups under type and cotype assumptions on the underlying Banach space. In particular, we show the following.
Let $-iA$ generate a $C_{0}$-group on a Banach space $X$ with type $p\in[1,2]$ and cotype $q\in[2,\infty)$. Then $A$ has a bounded $\mathcal{H}^{\infty}$-calculus from $\mathrm{D}_{A}(\tfrac{1}{p}-\tfrac{1}{q},1)$ to $X$, i.e. $f(A):\mathrm{D}_{A}(\tfrac{1}{p}-\tfrac{1}{q},1)\to X$ is bounded for each bounded holomorphic function $f$ on a sufficiently large strip.
As a corollary of our main theorem, for sectorial operators we quantify the gap between bounded imaginary powers and a bounded $\mathcal{H}^{\infty}$-calculus in terms of the type and cotype of the underlying Banach space. For cosine functions we obtain similar results as for $C_{0}$-groups. We extend our results to $R$-bounded operator-valued calculi, and we give an application to the theory of rational approximation of $C_{0}$-groups. | math.FA | math |
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND
COTYPE
JAN ROZENDAAL
Abstract. We study the functional calculus properties of generators of C0-
groups under type and cotype assumptions on the underlying Banach space.
In particular, we show the following. Let −iA generate a C0-group on a
Banach space X with type p ∈ [1, 2] and cotype q ∈ [2, ∞). Then f (A) :
→ X is bounded for each bounded holomorphic function f
(X, D(A)) 1
p
q ,1
− 1
on a sufficiently large strip. As a corollary of this result, for sectorial opera-
tors we quantify the gap between bounded imaginary powers and a bounded
H∞-calculus in terms of the type and cotype of the underlying Banach space.
For cosine functions we obtain similar results as for C0-groups. We extend our
theorems to R-bounded operator-valued calculi, and we give an application to
the theory of rational approximation of C0-groups.
1. Introduction
Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach space X, and set
θ(U ) := infnθ ≥ 0(cid:12)(cid:12)(cid:12)∃M ≥ 1 : kU (s)k ≤ M eθs for all s ∈ Ro .
(1.1)
For each ω > 0 let
Stω := {z ∈ CIm(z) > ω} ,
(1.2)
and let H∞(Stω) be the space of bounded holomorphic functions on Stω with the
supremum norm. There is a natural definition of f (A) as an unbounded operator
on X for each ω > θ(U ) and each f ∈ H∞(Stω) (see Section 2.1). It was shown by
Boyadzhiev and deLaubenfels in [9] that, if X is a Hilbert space, then there exists
a constant C ≥ 0 such that f (A) ∈ L(X) with
(1.3)
for each ω > θ(U ) and f ∈ H∞(Stω). One says that A has a bounded H∞-calculus.
It is useful to know that an operator has a bounded H∞-calculus. The theory of
H∞-calculus has applications to questions of maximal regularity (see [2, 35, 38])
and played a crucial role in the solution to the Kato square root problem in [4, 6].
Also, functional calculus bounds can be used to determine convergence rates in the
numerical approximation theory for solutions to evolution equations (see [10,17,21]).
One can obtain nontrivial functional calculus results for C0-groups even when X
is not a Hilbert space. For example, it was shown in [25] that if −iA generates a
kf (A)kL(X) ≤ C kfkH∞(Stω)
2010 Mathematics Subject Classification. Primary 47A60; Secondary 47D03, 46B20, 42A45.
Key words and phrases. Functional calculus, Operator group, Type and cotype, Transference,
R-boundedness.
1
2
JAN ROZENDAAL
C0-group (U (s))s∈R on a UMD space X, then A has a bounded H∞1 (Stω)-calculus
for all ω > θ(U ). Here H∞1 (Stω) consists of all f ∈ H∞(Stω) such that
(1.4)
kfkH∞
1 (Stω ) := sup
z∈Stω f (z) + (1 + z)f′(z) < ∞.
In [29] a similar statement was obtained on general Banach spaces, where one
restricts the calculus to real interpolation spaces between X and the domain of A.
However, there are several drawbacks to functional calculus theory for function
spaces which are strictly contained in the class of H∞-functions. For example, in
applications one might be interested in functions f which are not contained in a
smaller function space such as H∞1 (Stω). But even when dealing with a function
f which is known to yield a bounded operator f (A), it can be of interest to know
that (1.3) holds, instead of an estimate for kf (A)kL(X) with respect to a larger
function norm. This is the case for numerical approximation schemes for solutions
to evolution equations, where one can often improve convergence rates for operators
with a bounded H∞-calculus (see [17, 20]).
Moreover, a bounded H∞-calculus yields square function estimates that arise
frequently in harmonic analysis (see [22, 36, 40]) and that are of use in the theory
of stochastic evolution equations [50, 51]. Also, a bounded H∞-calculus can be
bootstrapped to yield large classes of R-bounded operators, and the notion of R-
boundedness has various applications in the theory of evolution equations (see [38]).
It seems that such connections are not available in full generality for subspaces of the
class of H∞-functions (if A is self-adjoint then more can be said; see for example [5]).
Finally, it seems somewhat unsatisfactory that there is such a rough division
between functional calculus properties on Hilbert spaces and on UMD spaces, in the
sense that one goes from a bounded H∞-calculus on Hilbert spaces to a bounded
H∞1 -calculus on UMD spaces. One might expect a finer division of functional
calculus theorems within the class of UMD spaces, and on Lp-spaces one might
hope that functional calculus properties improve as p tends to 2. The latter is
indeed the case for symmetric contraction semigroups on Lp-spaces, cf. [12, 14].
In the present article we aim to address the issues above. We study H∞-calculus
for generators of C0-groups in terms of the type and cotype of the underlying Banach
space (see Definition 2.3). Our main result, proved as Theorem 4.1, is as follows.
(For the real interpolation space DA( 1
Theorem 1.1. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach space
X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ). Then there exists a
constant C ≥ 0 such that DA( 1
q , 1) see (1.7)).
p − 1
p − 1
q , 1) ⊆ D(f (A)) and
kf (A)xkX ≤ C kfkH∞(Stω ) kxkDA( 1
p− 1
q ,1)
for all f ∈ H∞(Stω) and x ∈ DA( 1
p − 1
q , 1).
Under additional geometric assumptions we improve Theorem 1.1. Indeed, in
Theorem 4.3 we show that, if X is isomorphic to a complemented subspace of a
p-convex and q-concave Banach lattice for p ∈ [1, 2] and q ∈ [2,∞), then for each
λ > ω > θ(U ) there exists a constant C ≥ 0 such that D((λ+ iA)− 1
q ) ⊆ D(f (A))
p + 1
and
(1.5)
kf (A)xkX ≤ C kfkH∞(Stω) k(λ + iA)− 1
p + 1
q xkX
for all f ∈ H∞(Stω) and x ∈ D((λ + iA)− 1
p + 1
q ).
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
3
One might say that Theorem 1.1 shows that each generator −iA of a C0-group
on a Banach space X with type p and cotype q has a bounded H∞-calculus from
q , 1) to X, and A has a bounded H∞-calculus from D((λ + iA)− 1
p − 1
DA( 1
q ) to X
under additional assumptions on X. Since 1
q = 0 if and only if X is isomorphic
to an L2-space, (1.5) recovers as a special case (1.3), the main result of [9].
p − 1
p + 1
The interpolation spaces DA(θ, 1) and the fractional domains D((λ + iA)−θ)
increase as θ tends to zero. Therefore the statements in this paper show that
many functional calculus properties depend in a quantitative manner on how close
the geometry of the underlying space is to that of a Hilbert space. Since each
UMD space has non-trivial type and finite cotype, Theorem 1.1 provides a scale of
functional calculus results on UMD spaces.
p − 1
q ≥ 1
q is attained.
For (Ω, µ) a measure space and p ∈ [1,∞), the space Lp(Ω, µ) is a p-convex
and p-concave Banach lattice with type min(p, 2) and cotype max(p, 2). Hence
our results show that the functional calculus properties of group generators on Lp-
spaces improve as p tends to 2. It should be noted that (1.5) holds without any
assumptions on the Banach space if 1
2 (see Remark 4.8). Therefore most
of the results in Sections 4 and 5 are only of interest on Lp-spaces when p ∈ (1,∞).
In Proposition 4.6 we deduce from Theorem 1.1 that each f ∈ H∞(Stω) with
polynomial decay of order α > 1
p − 1
q at infinity satisfies f (A) ∈ L(X). Under the
assumptions of (1.5) the case α = 1
p − 1
It was shown by McIntosh in [43] that, on Hilbert spaces, sectorial operators with
bounded imaginary powers have a bounded sectorial H∞-calculus. It is also known
that on general Banach spaces (even on Lp-spaces) an operator with bounded imag-
inary powers need not have a bounded H∞-calculus. In Theorem 5.1 we quantify
the gap between bounded imaginary powers and a bounded H∞-calculus, by show-
ing that each sectorial operator A with bounded imaginary powers has a bounded
sectorial H∞-calculus from Dlog(A)( 1
q , 1) to X if the underlying Banach space
X has type p ∈ [1, 2] and cotype q ∈ [2,∞). As in the classical case of a bounded
H∞-calculus on X, one obtains from this an unconditionality result and square
function estimates.
For generators of cosine functions we derive similar results as for C0-groups. In
particular, in Theorem 5.4 we show that each generator −A of a cosine function
on a Banach space X with type p and cotype q has a bounded parabola-type H∞-
calculus from DA( 1
We extend Theorem 1.1 and (1.5) to operator-valued functional calculi. Then in
Theorem 6.2 we show that A in fact has an R-bounded H∞-calculus from DA( 1
p −
1
q , 1) to X if X additionally has property (α). In Theorem 6.3 we obtain an R-
bounded version of (1.5). These results are sharp, as Example 6.5 shows.
q ), 1) to X.
p − 1
2 ( 1
p − 1
To indicate the use of Theorem 1.1 we give an application to the study of numer-
ical approximation methods for the solutions to evolution equations. In Proposition
7.3 we consider a method of rational approximation proposed in [34, 44], and for
exponentially stable groups on Lp-spaces we improve the rates of convergence which
were obtained in [17]. In Corollary 7.4 we show that many rational approximation
methods of exponentially stable C0-groups converge strongly on DA( 1
q , 1) if the
underlying space X has type p ∈ [1, 2] and cotype q ∈ [2,∞).
To prove Theorem 1.1 we use transference principles going back to [8,11,13]. The
transference technique has been applied to functional calculus theory in [14], [30]
p − 1
4
JAN ROZENDAAL
and [25, 26, 28, 29]. In [29] an interpolation version of the transference principle for
unbounded groups from [25] was established. This transference principle was then
combined with a theorem about Fourier multipliers on vector-valued Besov spaces.
In the present paper we also obtain a transference principle involving vector-valued
Besov spaces, but we combine it with a theorem about Fourier multipliers between
distinct Besov spaces. For (1.5) we use statements about Fourier multipliers from
Lp to Lq which were obtained recently in [47, 48]. We suspect that it is possible to
apply these Fourier multiplier theorems to the transference principle from [28] for
C0-semigroups, but this matter will not be explored in the present article.
This paper is organized as follows. In Section 2 we discuss some of the basics of
functional calculus theory, vector-valued Besov spaces and the notions of type and
cotype. We then state two Fourier multiplier results which are vital for that which
follows. In Section 3 we establish two new transference principles for C0-groups.
These are used in Section 4 to prove Theorem 1.1 and (1.5), as well as several
corollaries for C0-groups. In Section 5 we obtain results for sectorial operators and
generators of cosine functions, and in Section 6 we extend the results from previous
sections to R-bounded operator-valued calculi. Finally, in Section 7 we give an
application to the theory of rational approximation schemes.
1.1. Notation and terminology. We write N := {1, 2, 3, . . .} for the natural
numbers and N0 := N ∪ {0}. We write C+ for the open right half-plane of complex
numbers z ∈ C with Re(z) > 0, and C− := C \ C+.
We denote nonzero Banach spaces over the complex numbers by X and Y . The
space of bounded linear operators from X to Y is L(X, Y ), and L(X) := L(X, X).
The domain of a closed operator A on X is D(A), a Banach space with the norm
kxkD(A) := kxkX + kAxkX
(x ∈ D(A)).
The spectrum of A is σ(A) and the resolvent set ρ(A) := C \ σ(A). We write
R(λ, A) := (λ − A)−1 for the resolvent operator of A at λ ∈ ρ(A).
For p ∈ [1,∞] and (Ω, µ) a measure space, Lp(Ω; X) denotes the Bochner space of
equivalence classes of p-integrable X-valued functions on Ω. We often write k·kp =
k·kLp(R;C). For n ∈ N0 we let W n,p(R; X) be the Sobolev space of n times weakly
differentiable f ∈ Lp(R; X) such that f (n) ∈ Lp(R; X). The Holder conjugate of p
is denoted by p′ and is defined by 1 = 1
We let M(R) denote the space of complex Borel measures on R with the total
variation norm. For ω ≥ 0 we let Mω(R) be the convolution algebra of µ ∈ M(R) of
the form µ(ds) = e−ωsν(ds) for some ν ∈ M(R), with kµkMω(R) := keω·µkM(R).
For Ω 6= ∅ open in C, we denote by H∞(Ω) the space of bounded holomorphic
functions f : Ω → C, a Banach algebra with the norm
p + 1
p′ .
kfkH∞(Ω) := sup
z∈Ωf (z)
(f ∈ H∞(Ω)).
The class of X-valued rapidly decreasing smooth functions on R is S(R; X), and
the space of X-valued tempered distributions is S′(R; X). The Fourier transform
of Φ ∈ S′(R; X) is F Φ. If µ ∈ Mω(R) for ω > 0 then F µ ∈ H∞(Stω) is given by
F µ(z) :=ZR
e−izsµ(ds)
(z ∈ Stω).
An interpolation couple is a pair (X, Y ) of Banach spaces which are embedded
continuously in a Hausdorff topological vector space Z. The real interpolation
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
5
space of (X, Y ) with parameters θ ∈ [0, 1] and q ∈ [1,∞] is denoted by (X, Y )θ,q.
If T : X + Y → X + Y restricts to a bounded operator on X and Y then
(1.6)
kTkL((X,Y )θ,q ) ≤ kTk1−θ
L(X) kTkθ
L(Y )
for all θ ∈ (0, 1) and q ∈ [1,∞]. We mainly consider real interpolation spaces for
the interpolation couple (X, D(A)), where A is a closed operator on X. We write
(1.7)
DA(θ, q) := (X, D(A))θ,q
and
kxkθ,q := kxkDA(θ,q)
(x ∈ DA(θ, q)).
For an operator B on X and a continuously embedded space Y ֒→ X, the part
of B in Y is the operator BY on Y that satisfies BY y = By for y ∈ D(BY ) :=
{z ∈ D(B) ∩ Y Bz ∈ Y }. We write Bθ,q := BDA(θ,q) for θ ∈ [0, 1] and q ∈ [1,∞].
2. Functional calculus and Fourier multipliers
In this section we present the background on functional calculus and Fourier
multipliers which will be needed for the rest of the article.
2.1. Functional calculus. We assume that the reader is familiar with the basics
of the theory of C0-groups from [19]. For more on the functional calculus for
generators of C0-groups see [23, Chapter 4].
An operator A on a Banach space X is a strip-type operator of height ω0 ≥ 0 if
σ(A) ⊆ Stω0 , where St0 := R, and supλ∈C\Stω kR(λ, A)k < ∞ for all ω > ω0. For
ω > 0 set
E(Stω) :=(cid:8)g ∈ H∞(Stω)(cid:12)(cid:12)g(z) ∈ O(z−α) for some α > 1 as Re(z) → ∞(cid:9) .
The strip-type functional calculus for a strip-type operator A of height ω0 is defined
as follows. First, operators f (A) ∈ L(X) are associated with f ∈ E(Stω) for ω > ω0:
(2.1)
f (z)R(z, A) dz.
f (A) :=
1
2πiZ∂Stω′
Here ∂Stω′ is the positively oriented boundary of Stω′ for ω′ ∈ (ω0, ω). This proce-
dure is independent of the choice of ω′ by Cauchy's theorem, and yields an algebra
homomorphism E(Stω) → L(X), f 7→ f (A). The definition of f (A) is extended to
a larger class of functions by regularization:
f (A) := e(A)−1(ef )(A)
(2.2)
if there exists an e ∈ E(Stω) with e(A) injective and ef ∈ E(Stω). Then f (A) is
a closed unbounded operator on X, and the definition of f (A) is independent of
the choice of the regularizer e. Each f ∈ H∞(Stω) is regularizable by the function
z 7→ (λ − z)−2 for Im(λ) > ω.
Let −iA generate a C0-group (U (s))s∈R ⊆ L(X). Then A is a strip-type operator
of height θ(U ), with θ(U ) as in (1.1). Let M ≥ 1 and ω ≥ 0 be such that kU (s)k ≤
M eωs for all s ∈ R, and for µ ∈ Mω(R) set
(2.3)
U (s)x dµ(s)
Uµx :=ZR
(x ∈ X).
The mapping µ 7→ Uµ is an algebra homomorphism Mω(R) → L(X) called the
Hille-Phillips calculus, and the following lemma from [24, Lemma 2.2] shows that
this calculus is consistent with the strip-type calculus for A.
6
JAN ROZENDAAL
Lemma 2.1. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach space
X, and let ω > α > θ(U ). Then each f ∈ E(Stω) satisfies f = F µ and f (A) = Uµ ∈
L(X) for some µ ∈ Mα(R). Conversely, if µ ∈ Mα(R) then f := F µ ∈ H∞(Stα)
is such that f (A) = Uµ ∈ L(X).
In fact, it is observed in Remark 4.8 that the first statement in Lemma 2.1 holds
for all f ∈ H∞(Stω) such that f (z) = O(z−1/2) as Re(z) → ∞.
The following is a version of this lemma adapted to our setting.
A fundamental result in functional calculus theory is the Convergence Lemma.
Lemma 2.2 (Convergence Lemma). Let A be a densely defined strip-type operator
of height ω0 ≥ 0 on a Banach space X. Let Y be a Banach space continuously
embedded in X such that D(A2) ⊆ Y is dense. Let ω > ω0 and let (fj)j∈J ⊆
H∞(Stω) be a net satisfying the following conditions:
• supj∈J kfjkH∞(Stω ) < ∞;
• f (z) := limj fj(z) exists for all z ∈ Stω;
• supj∈J kfj(A)kL(Y,X) < ∞.
Then f ∈ H∞(Stω), f (A) ∈ L(Y, X), fj(A)x → f (A)x for all x ∈ Y and
kf (A)kL(Y,X) ≤ lim sup
kfj(A)kL(Y,X) .
j∈J
Proof. The proof is similar to the proofs of [23, Proposition 5.1.7] and [7, Theorem
3.1]. Vitali's theorem for nets from [3, Theorem 2.1] implies that f ∈ H∞(Stω)
and that fj → f uniformly on compact subsets of Stω. Let λ > ω. Applying the
dominated convergence theorem to (2.1) yields
fj(A)x =(cid:18) fj(·)
(iλ − ·)2(cid:19) (A)(iλ − A)2x →(cid:18) f (·)
(iλ − ·)2(cid:19) (A)(iλ − A)2x = f (A)x
and
kf (A)xkX ≤ lim sup
j
kfj(A)kL(Y,X) kxkY
for all x ∈ D(A2). The required statements now follow since D(A2) ⊆ Y is dense
and f (A) is a closed operator on X.
(cid:3)
2.2. Fourier multipliers and Banach space geometry. In this section we treat
Fourier multiplier operators under geometric assumptions on the underlying space.
For more on the prerequisite notions from Banach space geometry, as well as for
proofs of some of the statements below, see e.g. [15, 31, 32, 41]. Recall that a stan-
dard complex Gaussian random variable is a random variable γ on a probability
space (Ω, P) such that γ = γr +iγi√2
for independent standard real Gaussian random
variables γr, γi on Ω. A Gaussian sequence is a sequence of independent standard
complex Gaussian random variables.
Definition 2.3. Let X be a Banach space, (γk)k∈N a Gaussian sequence on a
probability space (Ω, P) and p ∈ [1, 2], q ∈ [2,∞].
• X has (Gaussian) type p if there exists a constant C ≥ 0 such that for all
n ∈ N and x1, . . . , xn ∈ X,
(2.4)
(cid:16)E(cid:13)(cid:13)(cid:13)
nXk=1
γkxk(cid:13)(cid:13)(cid:13)
2(cid:17)1/2
≤ C(cid:16) nXk=1
kxkkp(cid:17)1/p
.
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
7
(2.5)
• X has (Gaussian) cotype q if there exists a constant C ≥ 0 such that for
all n ∈ N and x1, . . . , xn ∈ X,
kxkkq(cid:17)1/q
nXk=1
with the obvious modification for q = ∞.
≤ C(cid:16)E(cid:13)(cid:13)(cid:13)
(cid:16) nXk=1
2(cid:17)1/2
γkxk(cid:13)(cid:13)(cid:13)
,
The minimal constants C in (2.4) and (2.5) are called the Gaussian type p con-
stant and the Gaussian cotype q constant and will be denoted by τp,X and cq,X .
We say that X has nontrivial type if X has type p ∈ (1, 2] and that X has finite
cotype if X has cotype q ∈ [2,∞). By the Kahane-Khintchine inequalities, one may
replace the exponent 2 in (2.4) and (2.5) by any r ∈ [1,∞). This does not change
the properties of type and cotype, only the minimal constants in (2.4) and (2.5).
It is common to replace the Gaussian sequence in Definition 2.3 by a Rademacher
sequence, i.e. a sequence (rk)k∈N of independent complex random variables on a
probability space (Ω, P) that are uniformly distributed on {z ∈ C
z = 1}.
This does not change the class of spaces under consideration, only the minimal
constants in (2.4) and (2.5). We choose to work with Gaussian sequences because
the constants τp,X and cq,X occur in Proposition 2.4.
Every Banach space X has type p = 1 and cotype q = ∞, with τ1,X = c∞,X = 1.
If X has type p and cotype q then it has type r with τr,X ≤ τp,X for all r ∈ [1, p] and
cotype s with cs,X ≤ cq,X for all s ∈ [q,∞]. A Banach space X has type p = 2 and
cotype q = 2 if and only if X is isomorphic to a Hilbert space, by Kwapie´n's result
[39]. A Banach space X with nontrivial type has finite cotype. Each UMD space has
nontrivial type. Let X be a Banach space with type p ∈ [1, 2] and cotype q ∈ [2,∞),
let r ∈ [1,∞) and let Ω be a measure space. Then Lr(Ω; X) has type min(p, r) and
cotype max(q, r) with τmin(p,r),Lr(Ω;X) ≤ Cp,rτp,X and cmax(q,r),Lr(Ω;X) ≤ Cq,rcq,X
for constants Cp,r, Cq,r ≥ 0 coming from the Kahane-Khintchine inequalities.
Let ψ ∈ C∞(R) be such that ψ ≥ 0, supp(ψ) ⊆ [ 1
ψ(2−ks) = 1
for all s ∈ (0,∞). For k ∈ N and s ∈ R let ϕk(s) := ψ(2−ks) and ϕ0(s) :=
1 −P∞k=1 ϕk(s). Let X be a Banach space and let p, q ∈ [1,∞] and r ∈ R. The
p,q(R; X) is the space of all f ∈ S′(R; X) such that
(inhomogeneous) Besov space Br
2 , 2] and P∞k=−∞
kfkBr
p,q(R;X) :=(cid:13)(cid:13)(cid:13)(cid:16)2kr(cid:13)(cid:13)F−1ϕk ∗ f(cid:13)(cid:13)Lp(R;X)(cid:17)k∈N0(cid:13)(cid:13)(cid:13)ℓq
< ∞,
p,q(R;X). Then Br
endowed with the norm k·kBr
p,q(R; X) is a Banach space, S(R; X) ⊆
Br
p,q(R; X) is dense if p, q < ∞, and a different choice of ψ yields an equivalent norm
on Br
p,q(R; X). More details on vector-valued Besov spaces can be found in [1, 49].
For X a Banach space and m ∈ L∞(R;L(X)), the Fourier multiplier operator
Tm : S(R; X) → S′(R; X) with symbol m is given by
Tm(f ) := F−1 (m · F f )
(f ∈ S(R; X)).
For each µ ∈ M(R),
(2.6)
defines a Fourier multiplier operator with symbol F µ ∈ L∞(R). We say that X is
a UMD space if the function 1[0,∞) − 1(−∞,0) is the symbol of a bounded Fourier
multiplier on L2(R; X).
Lµ(f ) := µ ∗ f
(f ∈ S(R; X))
8
JAN ROZENDAAL
Let X and Y be Banach spaces and let (rk)k∈N be a Rademacher sequence on a
probability space (Ω, P). A collection T ⊆ L(X, Y ) is R-bounded if there exists a
constant C ≥ 0 such that, for all n ∈ N, x1, . . . , xn ∈ X and T1, . . . , Tn ∈ T ,
(2.7)
.
nXk=1
(cid:16)E(cid:13)(cid:13)(cid:13)
rkTkxk(cid:13)(cid:13)(cid:13)
2
Y(cid:17)1/2
≤ C(cid:16)E(cid:13)(cid:13)(cid:13)
nXk=1
rkxk(cid:13)(cid:13)(cid:13)
The minimal constant C in (2.7) is denoted by R(T ). If we want to specify the
underlying spaces X and Y then we write RX,Y (T ) = R(T ), and we write RX (T ) =
R(T ) if T ⊆ L(X). By the Kahane contraction principle, each uniformly bounded
collection T ⊆ C is R-bounded as a subset of L(X), with RX (T ) equal to the
uniform bound of T .
The following result was obtained in [47, Theorem 1.1].
2
X(cid:17)1/2
Proposition 2.4. Let X be a Banach space with type p ∈ [1, 2] and cotype q ∈
[2,∞]. Let r ∈ R, s ∈ [1,∞] and m ∈ L∞(R;L(X)) be such that {m(s) s ∈ R} ⊆
L(X) is R-bounded. Then Tm ∈ L(Br+1/p−1/q
p − 1
q,s(R; X)) and
(R; X), Br
p,s
1
q,s(R;X)) ≤ 4
q τp,X cq,X RX ({m(s) s ∈ R}).
kTmkL(Br+1/p−1/q
p,s
(R;X),Br
Corollary 2.5. Let X be a Banach space with type p ∈ [1, 2] and cotype q ∈ [2,∞],
and let µ ∈ M(R). Then Lµ ∈ L(B1/p−1/q
(R;X),Lq(R;X)) ≤ 4
q τp,X cq,X kF µkL∞(R) .
(R; X), Lq(R; X)) and
kLµkL(B1/p−1/q
p− 1
p,1
p,1
1
Proof. By the Kahane contraction principle, RX ({F µ(s) s ∈ R}) = kF µkL∞(R).
q,1(R; X) is contractively embedded in Lq(R; X), the result follows by ap-
Since B0
plying Proposition 2.4 to Lµ = TF µ.
(cid:3)
Remark 2.6. If in Corollary 2.5 one assumes additionally that X is a UMD space,
then Lµ ∈ L(B1/p−1/q
p,p
(2.8)
(R; X), Lq(R; X)) and
kLµkL(B1/p−1/q
p,p
(R;X),Lq(R;X)) ≤ Cτp,X cq,X kF µkL∞(R)
for each µ ∈ M(R) and some C ≥ 0 independent of µ. This follows from Proposition
2.4 and the embedding B0
q,p(R; X) ⊆ Lq(R; X) from [52, Proposition 3.1].
We assume that the reader is familiar with the basics of Banach lattices from [41].
Definition 2.7. Let X be a Banach lattice and p, q ∈ [1,∞].
• X is p-convex if there exists a constant C ≥ 0 such that for all n ∈ N and
x1, . . . , xn ∈ X,
(cid:13)(cid:13)(cid:13)(cid:16) nXk=1
xkp(cid:17)1/p(cid:13)(cid:13)(cid:13) ≤ C(cid:16) nXk=1
kxkkp(cid:17)1/p
,
• X is q-concave if there exists a constant C ≥ 0 such that for all n ∈ N and
with the obvious modification for p = ∞.
x1, . . . , xn ∈ X,
(cid:16) nXk=1
kxkkq(cid:17)1/q
≤ C(cid:13)(cid:13)(cid:13)(cid:16) nXk=1
xkq(cid:17)1/q(cid:13)(cid:13)(cid:13),
with the obvious modification for q = ∞.
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
9
Every Banach lattice X is 1-convex and ∞-concave. If X is p-convex and q-
concave then it is r-convex and s-concave for all r ∈ [1, p] and s ∈ [q,∞]. By [41,
Proposition 1.f.3], if X is q-concave then it has cotype max(q, 2), and if X is p-convex
and q-concave for some q < ∞ then X has type min(p, 2). For (Ω, µ) a measure
space and r ∈ [1,∞], Lr(Ω, µ) is an r-convex and r-concave Banach lattice.
A subspace X0 ⊆ Y of a Banach space Y is said to be complemented if there
exists a projection P ∈ L(Y ) with P (Y ) = X0.
Proposition 2.8. Let X be isomorphic to a complemented subspace of a p-convex
and q-concave Banach lattice, for p ∈ [1, 2] and q ∈ [2,∞). Then there exists a
constant C ≥ 0 such that Tm ∈ L(Lp(R; X), Lq(R; X)) with
(2.9)
p − 1
1
q m(s) s ∈ R})
kTmkL(Lp(R;X),Lq(R;X)) ≤ CRX ({s
p − 1
1
for each m ∈ L∞(R;L(X)) such that RX ({s
Proof. Let X0 be a subspace of a p-convex and q-concave Banach lattice Y , S :
X → X0 an isomorphism and P ∈ L(Y ) a projection with P (Y ) = X0. Let
m ∈ L∞(R;L(X)) be such that RX ({s
q m(s) s ∈ R}) < ∞, and let m0 ∈
L∞(R;L(Y )) be given by m0(s) := Sm(s)S−1P ∈ L(Y ) for s ∈ R. Then
q m(s) s ∈ R}) < ∞.
p − 1
1
RY ({s
1
p − 1
q m0(s) s ∈ R}) ≤ kSkkS−1kkPkRX({s
1
p− 1
q m(s) s ∈ R}).
It follows from [48, Theorem 3.21] that there exists a constant C ≥ 0 independent
of m such that Tm0 ∈ L(Lp(R; Y ), Lq(R; Y )) with
kTm0kL(Lp(R;Y ),Lq(R;Y )) ≤ CRY ({s
This concludes the proof since S−1Tm0S = Tm.
q m0(s) s ∈ R}).
p− 1
(cid:3)
1
3. Transference principles
In this section we establish two new transference principles for C0-groups, both
based on the transference principles for unbounded groups from [25] and [29].
For ω ≥ 0 and µ ∈ Mω(R) let µω ∈ M(R) be given by
(3.1)
and note that
(3.2)
µω(ds) := cosh(ωs)µ(ds),
F µω(s) = F µ(s + iω) + F µ(s − iω)
2
for all s ∈ R.
Proposition 3.1. Let ω > ω0 ≥ 0, p ∈ [1, 2] and q ∈ [2,∞). Then there exists
a constant C ≥ 0 such that the following holds. Let X be a Banach space with
type p and cotype q, and let −iA generate a C0-group (U (s))s∈R ⊆ L(X) such that
kU (s)kL(X) ≤ M cosh(ω0s) for all s ∈ R and some M ≥ 1. Then
ZR
(cid:13)(cid:13)(cid:13)(cid:13)
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)X ≤ Cτp,X cq,X M 2kF µωkL∞(R)kxk1/p−1/q,1
for all µ ∈ Mω(R) and all x ∈ DA( 1
p − 1
q , 1).
10
JAN ROZENDAAL
p − 1
q ∈ (0, 1) (for 1
Proof. Since DA(0, 1) = {0} we may assume that 1
stronger result is obtained in Proposition 3.3 below). Let
√8ω
π
cosh(ωs)
cosh(2ωs)
cosh(2ωs)
ψ(s) :=
ϕ(s) :=
(3.3)
and
1
p − 1
q = 0 a
for s ∈ R. Define ι : X → Lp(R; X) by
ιx(s) := ψ(−s)U (−s)x
(3.4)
and P : Lq(R; X) → X by
P f :=ZR
(3.5)
ϕ(s)U (s)f (s) ds
Then ι is bounded and
(x ∈ X, s ∈ R)
(f ∈ Lq(R; X)).
(3.6)
kιkL(X,Lp(R;X)) ≤ M kψ(·) cosh(ω0·)kp .
By Holder's inequality, P is bounded and
(3.7)
kPkL(Lq(R;X),X) ≤ M kϕ(·) cosh(ω0·)kq′ .
Let x ∈ D(A). Then ιx ∈ C1(R; X) and
(ιx)′(s) = −ψ′(−s)U (−s)x + iψ(−s)U (−s)Ax
= −2ω
tanh(2ωs)
cosh(2ωs)
U (−s)x + i
1
cosh(2ωs)
U (−s)Ax
for all s ∈ R. Hence (ιx)′ ∈ Lp(R; X) and
cosh(ω0·)
k(ιx)′kp ≤ 2ωM ktanhkL∞(R)(cid:13)(cid:13)(cid:13)(cid:13)
cosh(2ω·)(cid:13)(cid:13)(cid:13)(cid:13)pkxkX + M(cid:13)(cid:13)(cid:13)(cid:13)
Now (3.6) implies that ιx ∈ W 1,p(R; X), with
cosh(ω0·)
cosh(2ω·)(cid:13)(cid:13)(cid:13)(cid:13)pkAxkX .
cosh(ω0·)
Hence ι : D(A) → W 1,p(R; X) is bounded and
(3.8)
kιxk1,p ≤ M (2ω ktanhkL∞(R) + 1)(cid:13)(cid:13)(cid:13)(cid:13)
kιkL(D(A),W 1,p(R;X)) ≤ M (2ω ktanhkL∞(R) + 1)(cid:13)(cid:13)(cid:13)(cid:13)
(R; X) =(cid:0)Lp(R; X), W 1,p(R; X)(cid:1) 1
cosh(2ω·)(cid:13)(cid:13)(cid:13)(cid:13)pkxkD(A) .
cosh(2ω·)(cid:13)(cid:13)(cid:13)(cid:13)p
By equation (5.9) in [1],
cosh(ω0·)
B1/p−1/q
p − 1
q ,1
p,1
.
with equivalent norms. Moreover, it follows from a direct sum argument that the
constant in the norm equivalence does not depend on X. Hence (1.6), (3.6) and
(3.8) imply that ι : DA( 1
(R; X) is bounded with
p − 1
kιkL(DA( 1
q , 1) → B1/p−1/q
q ,1),B1/p−1/q
p− 1
p,1
p,1
(R;X)) ≤ C1M,
(3.9)
for some constant C1 ≥ 0 independent of A and X.
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
11
It is shown in [25, Theorem 3.2] that ϕ ∗ ψ(s) =
cosh(ωs) for all s ∈ R. Let
Uµ ∈ L(X) be as in (2.3). Then the abstract transference principle from [26, Section
2] yields the commutative diagram
1
B1/p−1/q
p,1
(R; X)
ιx
DA( 1
p − 1
q , 1)
Lµω−−−−→ Lq(R; X)
yP
Uµ−−−−→
X
of bounded maps. Finally, estimate the norms of P and ι using (3.7) and (3.9) and
apply Corollary 2.5 to Lµω .
(cid:3)
Remark 3.2. If X is a UMD space then one can replace the space DA( 1
by the larger DA( 1
except that one uses (2.8) instead of Corollary 2.5 at the very end of the proof.
q , 1)
q , p). This follows in the exact same way as Proposition 3.1,
p − 1
p − 1
For Banach lattices we establish another transference principle. Recall that each
Banach space X with type p = 2 and cotype q = 2 is isomorphic to an L2-space,
by [39]. Hence the following proposition deals with the case p = q = 2 in Proposition
3.1.
Proposition 3.3. Let ω > ω0 ≥ 0, p ∈ [1, 2] and q ∈ [2,∞). Let X be isomorphic
to a complemented subspace of a p-convex and q-concave Banach lattice. Then there
exists a constant C ≥ 0 such that the following holds. Let −iA generate a C0-group
(U (s))s∈R ⊆ L(X) such that kU (s)kL(X) ≤ M cosh(ω0s) for all s ∈ R and some
M ≥ 1. Then
(cid:13)(cid:13)(cid:13)(cid:13)
ZR
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)X ≤ CM 2 sup
s∈R (cid:16)s
1
p− 1
q F µω(s)(cid:17)kxkX
for all µ ∈ Mω(R) and all x ∈ X.
Proof. For µ ∈ Mω(R) let Uµ ∈ L(X) be as in (2.3). Let ι and P be as in (3.4)
and (3.5), with ψ and ϕ as in (3.3). As in the proof of Proposition 3.1 one can
factorize Uµ as Uµ = P ◦ Lµω ◦ ι. Hence (3.6), (3.7) and Proposition 2.8 conclude
the proof.
(cid:3)
4. Results for C0-groups
In this section we obtain functional calculus results for generators of C0-groups.
4.1. The main result for C0-groups. We now prove our main functional calculus
result for C0-groups, already stated in the Introduction as Theorem 1.1.
Theorem 4.1. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach space
X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ). Then there exists a
constant C ≥ 0 such that DA( 1
(4.1)
q , 1) ⊆ D(f (A)) and
p − 1
q ,1),X) ≤ C kfkH∞(Stω )
p − 1
kf (A)kL(DA( 1
for all f ∈ H∞(Stω).
12
JAN ROZENDAAL
Proof. First consider f ∈ E(Stω) and let α ∈ (θ(U ), ω). By Lemma 2.1 there exists
a µ ∈ Mα(R) with f = F µ and f (A) = Uµ. By Proposition 3.1,
kf (A)xkX ≤ C kF µαkL∞(R) kxk 1
p − 1
q ,1
for all x ∈ DA(1/p− 1/q, 1) and some constant C ≥ 0 independent of f and x. This
implies (4.1) since kF µαkL∞(R) ≤ kfkH∞(Stω), as follows from (3.2).
For general f ∈ H∞(Stω), define τk(z) := −k2(ik − z)−2 for k ∈ N with k > ω
and z ∈ Stω. Then f τk ∈ E(Stω) for all k,
k kf τkkH∞(Stω) ≤ kfkH∞(Stω ) sup
sup
k kτkkH∞(Stω ) < ∞,
and (f τk)(z) → f (z) as k → ∞, for all z ∈ Stω. By what we have already shown,
kf τk(A)xkX ≤ C kf τkkH∞(Stω ) kxk 1
:= C supk kτkkH∞(Stω ). Since D(A2) ⊆ DA( 1
p − 1
for C′
lim supk kτkkH∞(Stω ) = 1, Lemma 2.2 yields f (A) ∈ L(DA( 1
q ,1),X) ≤ C kfkH∞(Stω) ,
kf (A)kL(DA( 1
p− 1
p − 1
q ,1 ≤ C′ kfkH∞(Stω) kxk 1
p− 1
q ,1
q , 1) is dense and since
p − 1
q , 1), X) with
which concludes the proof.
(cid:3)
Remark 4.2. If in Theorem 4.1 one assumes in addition that X is a UMD space,
then DA( 1
q , 1) may be replaced by DA( 1
q , p). This follows by using Remark
3.2 instead of Proposition 3.1 in the proof.
p − 1
p − 1
Under additional assumptions one can obtain stronger results. Note that (4.2)
improves (4.1), since DA(θ, 1) ⊆ D((λ + iA)θ) for each group generator A, λ ∈ R
sufficiently large and θ ∈ [0, 1], by [42, Proposition 4.1.7].
Theorem 4.3. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach space
X. Suppose that X is isomorphic to a complemented subspace of a p-convex and
q-concave Banach lattice, for p ∈ [1, 2] and q ∈ [2,∞). Let λ > ω > θ(U ). Then
there exists a constant C ≥ 0 such that D((λ + iA)
(4.2)
q ) ⊆ D(f (A)) and
p− 1
− 1
1
kf (A)kL(D((λ+iA)
1
p
q ),X) ≤ C kfkH∞(Stω )
for all f ∈ H∞(Stω).
p − 1
Proof. Write θ := 1
q . For f ∈ E(Stω) and α ∈ (θ(U ), ω) there exists a µ ∈
Mα(R) with f = F µα and f (A) = Uµ, by Lemma 2.1. Let x ∈ D((λ + iA)θ).
By [23, Corollary 3.3.6],
f (A)x = f (A)(λ + iA)−θ(λ + iA)θx = UµUν(λ + iA)θx = Uµ∗ν(λ + iA)θx,
where ν ∈ Mω(R) is given by ν(ds) := 1
ν)(s) = f (s)(λ + is)−θ for s ∈ R, (3.2) yields
Γ(θ) 1[0,∞)(s)sθ−1e−λsds. Since F (µ ∗
sup
s∈R sθF (µ ∗ ν)α(s) ≤ C1 kfkH∞(Stω)
for some constant C1 ≥ 0. Now Proposition 3.3 yields a constant C2 ≥ 0 such that
kf (A)xkX ≤ C2 sup
s∈R sθF (µ ∗ ν)α(s)k(λ + iA)θxkX
≤ C1C2 kfkH∞(Stω ) k(λ + iA)θxkX ,
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
13
which proves (4.2) for f ∈ E(Stω). Proceed as in the proof of Theorem 4.1 to obtain
(4.2) for general f ∈ H∞(Stω).
(cid:3)
Remark 4.4. It follows from the proofs of Theorems 4.1 and 4.3 that the constants
in (4.1) and (4.2) depend on A only through the norm kU (s)kL(X) of U (s) for all
s ∈ R. In (4.1) the constant C depends on the underlying space X only through the
Gaussian type p constant and the Gaussian cotype q constant of X. In particular,
the dependence of C on these constants is as in Proposition 2.4.
In (4.2) the
constant depends on the underlying space only through the constant in (2.9). Note
from the proof of Proposition 2.8 that, if (2.9) holds with constant C ≥ 0 on Y and
if X is complemented in Y by a projection P , then (2.9) holds on X with constant
C kPkL(Y ). This fact will be used in Theorem 6.3.
4.2. More results for C0-groups. Here we derive some additional results for
group generators from Theorems 4.1 and 4.3.
Proposition 4.5. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach
space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ) and α, λ ∈ C
with Re(α) > 1
q and Re(λ) > ω. Then there exists a constant C ≥ 0 such that
D((λ + iA)α) ⊆ D(f (A)) and
p − 1
(cid:13)(cid:13)f (A)(λ + iA)−α(cid:13)(cid:13)L(X) ≤ C kfkH∞(Stω )
for all f ∈ H∞(Stω).
p − 1
If 1
Proof. The case p = q = 2 follows from Theorem 4.3.
p − 1
by [42, Propositions 1.1.4 and 4.1.7], D((λ + iA)α) ⊆ DA( 1
Hence the proof is concluded by appealing to Theorem 4.1.
q ∈ (0, 1) then,
q , 1) continuously.
(cid:3)
One may equivalently formulate Proposition 4.5 as a statement about bounded-
ness on X of the calculus for functions with sufficient decay:
Proposition 4.6. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach
space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ) and α, λ ∈ C with
Re(α) > 1
q and Re(λ) > ω. Then there exists a constant C ≥ 0 such that the
following holds. Let f ∈ H∞(Stω) be such that f (z) ∈ O(z−α) as Re(z) → ∞.
Then f (A) ∈ L(X) with
(4.3)
p − 1
(cid:3)
kf (A)kL(X) ≤ C sup
z∈Stωλ + izαf (z).
Proof. Apply Proposition 4.5 to (λ + i·)αf (·) ∈ H∞(Stω).
p − 1
In Propositions 4.5 and 4.6 one may let α = 1
q if X is isomorphic to a
complemented subspace of a p-convex and q-concave Banach lattice, as follows
from Theorem 4.3.
Corollary 4.7. Let ω > 0 and let f ∈ H∞(Stω) be such that f (z) ∈ O(z−1/2) as
Re(z) → ∞. Then f = F µ, where µ ∈ Mω′(R) for all ω′ ∈ [0, ω). If in addition
f is bounded and holomorphic on {z ∈ C Im(z) > −w} then supp(µ) ⊆ [0,∞).
dt with maximal domain on X := L1(R, eω′tdt).
Proof. For ω′ ∈ [0, ω), let A := i d
Then −iA generates the left translation group (U (s))s∈R ⊆ L(X), and θ(U ) = ω′.
By Proposition 4.6, f (A) ∈ L(X). Now [24, Proposition 2.3] implies that f = F µ
for some µ ∈ Mω′(R). By uniqueness of the Fourier transform, µ is independent
14
JAN ROZENDAAL
of the choice of ω′ ∈ [0, ω). The final statement follows from an application of
Liouville's theorem.
(cid:3)
p − 1
q ≥ 1
Remark 4.8. It follows from Corollary 4.7 that the conclusion of Theorem 4.3
holds without any assumptions on the Banach space X if 1
2 . Indeed, let
−iA generate a C0-group (U (s))s∈R ⊆ L(X) on a general Banach space X, and
let λ > ω > θ(U ) and f ∈ H∞(Stω). By Corollary 4.7 and Lemma 2.1, f (A)(λ +
iA)−1/2 = (f (·)(λ + i·)−1/2)(A) ∈ L(X) and hence f (A) ∈ L(D((λ + iA)1/2), X).
Also note that Corollary 4.7 extends [28, Lemma 2.4] from α > 1/2 to α = 1/2.
Remark 4.9. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a UMD space X.
In [25] it is shown that A has a bounded H∞1 (Stω)-calculus for all ω > θ(U ), where
H∞1 (Stω) is defined by (1.4). Since X has type p ∈ (1, 2] and cotype q ∈ [2,∞), one
can compare our results with those in [25]. We note that our results do not imply
those in [25], nor does [25] imply the results in this article. Indeed, let ω > θ(U ).
Then f (z) := (λ + iz)−α defines an element of H∞1 (Stω) for all α > 0 and λ > ω,
but Proposition 4.6 does not apply to f if α ∈ (0, 1
q ). Also, the function
f ∈ H∞(Stω) given by f (z) := e−iz(λ + iz)−α is not an element of H∞1 (Stω) if
α ∈ ( 1
Although in both examples it is clear that f (A) ∈ L(X), the difference between
estimating kf (A)kL(X) by (1.4) or using (4.3) is relevant for numerical approxima-
tion methods, as is shown in Section 7.
p − 1
q at infinity.
q , 1) but decays with order α > 1
p − 1
p − 1
We now obtain a version of Theorem 4.1 for other interpolation spaces.
Proposition 4.10. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach
p + 1
space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ), r ∈ (0, 1 − 1
q )
and u ∈ [1,∞]. Then there exists a constant C ≥ 0 such that
q ,u
kf (A)xkr,u ≤ C kfkH∞(Stω) kxkr+ 1
p − 1
q , u).
p − 1
for all f ∈ H∞(Stω) and x ∈ DA(r + 1
Proof. Let f ∈ H∞(Stω) and first consider the case where u = 1. Then the part
−iAr,1 of −iA in DA(r, 1) generates the C0-group (U (s) ↾DA(r,1))s∈R ⊆ L(DA(r, 1))
which satisfies θ(U ↾DA(r,1)) ≤ θ(U ), by [29, Lemma 2.2]. Hence Theorem 4.1 yields
(4.4)
kf (Ar,1)xkr,1 ≤ C kfkH∞(Stω ) kxk(DA(r,1),D(Ar,1)) 1
− 1
q
,1
p
for all x ∈ (DA(r, 1), D(Ar,1)) 1
q ,1. Moreover, it follows from [42, Proposition
p− 1
3.1.5] that DA(r, 1) = (X, D(A2)) r
2 ,1 and D(Ar,1) = (X, D(A2)) r+1
2 ,1 with equiva-
lent norms. Hence, by the Reiteration Theorem (see [42, Theorem 1.3.5]) and again
by [42, Proposition 3.1.5],
(DA(r, 1), D(Ar,1)) 1
p − 1
q ,1 = ((X, D(A2)) r
= (X, D(A2)) 1
2 ,1, (X, D(A2)) r+1
2 ,1) 1
q ),1 = DA(r + 1
2 (r+ 1
p − 1
q ,1
p− 1
p − 1
q , 1).
Combine this with (4.4), using that f (Ar,1)x = f (A)x for all x ∈ D(f (Ar,1))
by [29, Lemma 2.2], to conclude the proof in the case where u = 1.
For general u ∈ [1,∞], the Reiteration Theorem yields
DA(r, u) = (DA(θ1, 1), DA(θ2, 1))θ,u
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
15
and
DA(r + 1
p − 1
p − 1
q , 1))θ,u
q , u) = (DA(θ1 + 1
q , 1), DA(θ2 + 1
p − 1
p + 1
q ) and θ ∈ (0, 1). Now apply (1.6) to what we have
(cid:3)
for certain θ1, θ2 ∈ (0, 1 − 1
already shown to conclude the proof.
Remark 4.11. For u < ∞ Proposition 4.10 can be proved directly, without using
Theorem 4.1. To do so, adapt Proposition 3.1 to the setting of Proposition 4.10,
using instead of Corollary 2.5 the more general Proposition 2.4, and then proceed
as in the proof of Theorem 4.1.
Similarly, Theorem 4.3 extends to other fractional domains.
Proposition 4.12. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a comple-
mented subspace X of a p-convex and q-concave Banach lattice, where p ∈ [1, 2] and
q ∈ [2,∞). Let λ > ω > θ(U ). Then there exists a constant C ≥ 0 such that
k(λ + iA)αf (A)xkX ≤ C kfkH∞(Stω ) k(λ + iA)α+ 1
p− 1
q xkX
for all α ∈ C, f ∈ H∞(Stω) and x ∈ D((λ + iA)α+ 1
p− 1
q ).
Proof. Apply Theorem 4.3 to (λ + iA)αx ∈ D((λ + iA)
iA)α+ 1
p − 1
q ).
1
p − 1
q ) for each x ∈ D((λ +
(cid:3)
5. Results for sectorial operators and cosine functions
In this section we derive from Theorem 4.1 some results for sectorial operators
and generators of cosine functions. By Remark 4.2, the results in this section can
be improved on UMD spaces, and by Remark 4.8 the statements hold on general
Banach spaces for 1
p − 1
q ≥ 1
2 .
5.1. Sectorial operators. For ϕ ∈ (0, π) let Sϕ := {z ∈ Carg(z) < ϕ}. An
operator A on a Banach space X is said to be a sectorial operator of angle ϕ if
σ(A) ⊆ Sϕ and sup{kzR(z, A)k ψ ∈ C \ Sψ} < ∞ for all ψ ∈ (ϕ, π).
as for strip-type operators. Define f (A) ∈ L(X) via a Cauchy-type integral for
For sectorial operators one can construct a functional calculus in a similar manner
f ∈ H∞0 (Sψ) :=ng ∈ H∞(Sψ)(cid:12)(cid:12)(cid:12)∃C, δ > 0 : g(z) ≤ C zδ
1 + z2δo,
set 1(A) := IX and (1 + ·)−1(A) := (1 + A)−1, and then extend linearly and
regularize as in (2.2). For details see [23, Chapter 2]. If A is an injective sectorial
operator of angle ϕ ∈ (0, π), then log(A) is defined via the sectorial calculus for
A, as is f (A) for all ψ ∈ (ϕ, π) and f ∈ H∞(Sψ). A sectorial operator A of angle
ϕ ∈ (0, π) has bounded imaginary powers if A is injective and if −i log(A) generates
a C0-group (U (s))s∈R ⊆ L(X). Then U (s) = A−is for all s ∈ R, and A is sectorial
of angle θA := θ(U ) if θ(U ) ∈ [0, π), by [46, Theorem 2]. We write A ∈ BIP(X).
If A ∈ BIP(X) and if there exists an ω ∈ [0, π) such that {e−ωsA−is s ∈ R} ⊆
L(X) is R-bounded, then A has a bounded H∞(Sψ) calculus for all ψ ∈ (ω, π).
Moreover, if X has property (α) as below, then a bounded H∞(Sψ)-calculus for
A implies that {e−ωsA−is s ∈ R} ⊆ L(X) is R-bounded for ω > ψ. See [36,
Theorem 7.5]. Hence the following result is only of use if A does not have R-bounded
imaginary powers.
16
JAN ROZENDAAL
Theorem 5.1. Let A ∈ BIP(X) with θA < π, where X is a Banach space with
type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ψ ∈ (θA, π). Then there exists a constant
C ≥ 0 such that Dlog(A)( 1
q , 1) ⊆ D(f (A)) and
p − 1
kf (A)kL(Dlog(A)( 1
p− 1
q ,1),X) ≤ C kfkH∞(Sψ)
for all f ∈ H∞(Sψ).
Proof. By [23, Theorem 4.2.4], (g ◦ log)(A) = g(log(A)) for all g ∈ H∞(Stψ). Since
g 7→ g◦log is an isometric isomorphism H∞(Stψ) → H∞(Sψ), Theorem 4.1 concludes
the proof.
(cid:3)
Remark 5.2. It was shown by Dore in [16] that each sectorial operator A with
dense range on a general Banach space X has a bounded H∞-calculus on the real
interpolation spaces (X, D(A)∩ ran(A))θ,r for all θ ∈ (0, 1) and q ∈ [1,∞]. We note
that this statement implies neither Proposition 4.10 nor Theorem 5.1. Indeed, after
rotation Proposition 4.10 deals with functions on strips around the imaginary axis,
whereas for unbounded group generators [16] only applies to f ∈ H∞(Sψ) for ψ > π
2 .
Moreover, for all θ ∈ (0, 1) and q ∈ [1,∞] it holds that (X, D(A) ∩ ran(A))θ,r ⊆
D(log(A)). Since in general D(log(A)) ( Dlog(A)( 1
q , 1), Theorem 5.1 does not
follow from [16].
p − 1
It follows from Proposition 4.10 that f (A) : Dlog(A)(r + 1
q , u) → Dlog(A)(r, u)
is bounded for each r ∈ (0, 1 − 1
kf (A)kL(Dlog(A)(r+ 1
p + 1
p − 1
p − 1
q ) and each u ∈ [1,∞], and
q ,u),Dlog(A)(r,u)) ≤ C kfkH∞(Sψ) .
In the same manner one can deduce from Proposition 4.6 that each f ∈ H∞(Sψ)
such that f (z) ∈ O(2π−i log(z)−α) as z → 0 or z → ∞ for some α > 1
q satisfies
f (A) ∈ L(X). If X is isomorphic to a complemented subspace of a p-convex and
q-concave Banach lattice then one may let α = 1
From Theorem 5.1 one obtains unconditionality of the functional calculus and
p − 1
q .
p− 1
square function estimates in the same manner as in [38, Theorem 12.2].
Corollary 5.3. Let A ∈ BIP(X) with θA < π, where X is a Banach space with type
p ∈ [1, 2] and cotype q ∈ [2,∞). Let ψ ∈ (θA, π), f ∈ H∞0 (Sψ), and let (rk)k∈Z be a
Rademacher sequence on a probability space (Ω, P). Then the following assertions
hold:
q ,1),X) n ∈ N, t > 0,ǫk = 1} < ∞;
sup
k=−n ǫkf (2ktA)kL(Dlog(A)( 1
∞Xk=−∞
for all x ∈ Dlog(A)( 1
• sup{kPn
p − 1
• there exists a constant C ≥ 0 such that
t>0(cid:13)(cid:13)(cid:13)
t>0(cid:13)(cid:13)(cid:13)
∞Xk=−∞
for all x∗ ∈ X∗.
rkf (2ktA)x(cid:13)(cid:13)(cid:13)L2(Ω;X) ≤ C kxkDlog(A)( 1
rkf (2ktA)∗x∗(cid:13)(cid:13)(cid:13)L2(Ω;Dlog(A)( 1
q , 1), and
p − 1
p − 1
sup
p − 1
q ,1)
q ,1)∗) ≤ C kx∗kX ∗
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
17
5.2. Cosine functions. For ω ≥ 0 let Πω := {z2 z ∈ Stω}. An operator A on
a Banach space X is of parabola-type ω if σ(A) ⊆ Πω and if for all ω′ > ω there
exists a Mω′ ≥ 0 such that
kR(λ, A)k ≤
(λ ∈ C \ Πω′ ).
Mω′
pλ(cid:16)Im(√λ) − ω′(cid:17)
For operators of parabola-type ω ≥ 0 there is a natural functional calculus, con-
structed similarly as the strip-type and sectorial functional calculi, and f (A) is
defined as an unbounded operator for all f ∈ H∞(Πω′ ), ω′ > ω. For details see [27].
A cosine function Cos : R → L(X) on a Banach space X is a strongly continuous
mapping such that Cos(0) = I and
Cos(t + s) + Cos(t − s) = 2Cos(t)Cos(s)
(s, t ∈ R).
Then
θ(Cos) := inf{ω ≥ 0 ∃M ≥ 0 : kCos(t)k ≤ M eωt for all t ∈ R} < ∞.
The generator of a cosine function is the unique operator −A on X that satisfies
λR(λ2,−A) =Z ∞
0
e−λtCos(t) dt
(λ > θ(Cos)).
Then A is an operator of parabola-type θ(Cos).
We now prove a version of Theorem 4.1 for generators of cosine functions.
Theorem 5.4. Let −A generate a cosine function (Cos(s))s∈R ⊆ L(X) on a Ba-
nach space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(Cos). Then
there exists a constant C ≥ 0 such that DA( 1
q ), 1) ⊆ D(f (A)) and
p − 1
2 ( 1
kf (A)kL(DA( 1
2 ( 1
p − 1
q ),1),X) ≤ C kfkH∞(Πω)
for all f ∈ H∞(Πω).
Proof. The proof follows the same lines as that of [29, Proposition 5.5]. It suffices
to assume that θ := 1
p − 1
q ∈ (0, 1). By [37, Theorem 2] there is a unique subspace
V ⊆ X such that D(A) ⊆ V and such that −iA generates a C0-group (U (s))s∈R ⊆
L(V × X) on V × X, where
A := i(cid:20) 0
IV
−A 0 (cid:21)
with domain D(A) := D(A) × V . Moreover, by [24, Theorem 6.2], θ(Cos) = θ(U ).
Hence Theorem 4.1 yields a constant C ≥ 0 such that g(A) ∈ L(DA(θ, 1), V × X)
with
kg(A)kL(DA(θ,1),V ×X) ≤ C kgkH∞(Stω )
(5.1)
for all g ∈ H∞(Stω).
kfkH∞(Πω ). Moreover, it is straightforward to see that
(5.2)
Let f ∈ H∞(Πω). Then [z 7→ g(z) := f (z2)] ∈ H∞(Stω) and kgkH∞(Stω) =
Now,
f (AV ) ⊕ f (A) = g(A).
A2 :=(cid:20) AV
0 A (cid:21)
0
18
JAN ROZENDAAL
with D(A2) = D(AV ) × D(A). By [42, Proposition 3.1.4],
D(A) × V ∈ K1/2(V × X, D(AV ) × D(A)) ∩ J1/2(V × X, D(AV ) × D(A)) ,
where K1/2 and J1/2 are as in [42, Definition 1.3.1]. Hence
V ∈ K1/2(X, D(A)) ∩ J1/2(X, D(A)) .
Now [42, Theorem 1.3.5] yields
DA(θ, 1) = (V × X, D(A) × V )θ,1 = (V, D(A))θ,1 × (X, V )θ,1
2 , 1(cid:1) , X(cid:1) with
Combining this with (5.1) and (5.2) yields f (A) ∈ L(cid:0)DA(cid:0) θ
kf (A)kL(DA( θ
2 ,1),X) ≤ kg(A)kL(DA(θ,1),V ×X) ≤ C kgkH∞(Stω ) = C kfkH∞(Πω ) ,
2 , 1(cid:1) × DA(cid:0) θ
= DA(cid:0) 1+θ
2 , 1(cid:1) .
as required.
(cid:3)
From Proposition 4.10 one deduces in a similar manner that, under the assump-
q )) and u ∈ [1,∞], there exists
tions of Proposition 5.4 and for all r ∈ (0, 1− 1
a constant C ≥ 0 such that
p − 1
2 ( 1
kf (A)xkr,u ≤ C kfkH∞(Πω ) kxkr+ 1
2 ( 1
p − 1
q ),u
2 ( 1
p − 1
q ), u). We leave the formulation of the
for all f ∈ H∞(Πω) and x ∈ DA(r + 1
obvious analogue of Proposition 4.6 for cosine functions to the reader.
Remark 5.5. Let −A generate a cosine function (Cos(s))s∈R ⊆ L(X), where X is
isomorphic to a complemented subspace of a p-convex and q-concave UMD Banach
lattice, for p ∈ (1, 2] and q ∈ [2,∞). Let λ > ω > θ(Cos). Then there exists a
constant C ≥ 0 such that
kf (A)xkX ≤ C kfkH∞(Πω ) k(λ2 + A)
1
2 ( 1
p− 1
q )xkX
q )). This follows from Theorem
for all f ∈ H∞(Πω) and x ∈ D((λ2 + A)
4.3 as in the proof of Proposition 5.4, using the complex interpolation method
and [23, Theorem 6.6.9] and [25, Theorem 5.5].
1
2 ( 1
p− 1
It should be noted that generators of cosine functions on UMD spaces have a
bounded sectorial H∞-calculus, by [25, Theorem 5.5]. Hence on UMD spaces The-
orem 5.4 is only of use when it does not suffice to obtain an estimate for kf (A)xkX
with respect to the supremum norm of f on a sector. The latter is e.g. the case if
f (z) = g(z2) for a g ∈ H∞(Stω) which is unbounded on any double sector Sψ ∪−Sψ
for ψ ∈ (0, π
2 ), such as g(z) = e−iz.
6. Operator-valued functional calculus
In this section we extend our main functional calculus theorems to R-bounded
operator-valued calculi. Since many of the ideas and proofs are similar to those in
the sections before, we leave some details to the reader.
Let A be a strip-type operator of height ω0 ≥ 0 on a Banach space X. Let
A ⊆ L(X) be the algebra of bounded operators that commute with A, and let
ω > ω0. For a bounded holomorphic f : Stω → A such that kf (z)kL(X) ≤ Cz−α
for all z ∈ Stω and certain C ≥ 0 and α > 1, define f (A) as in (2.1). Regularization
as in (2.2) extends this functional calculus to all bounded holomorphic f : Stω → A.
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
19
For ω ≥ 0 let Mω(R;A) consist of the A-valued Borel measures µ on R such that
eωsµ(ds) has bounded variation. If −iA generates a C0-group (U (s))s∈R ⊆ L(X)
then one can define Uµ ∈ L(X) as in (2.3) for all ω > θ(U ) and µ ∈ Mω(R;A).
Versions of Lemmas 2.1 and 2.2 hold for this operator-valued calculus, with the
same proofs.
For ω > 0 and T ⊆ L(X) let RH∞(Stω;T ) be the collection of bounded holo-
morphic functions f : Stω → T such that {f (z) z ∈ Stω} ⊆ L(X) is R-bounded.
If T is an algebra then RH∞(Stω;T ) is a Banach algebra with the norm
(f ∈ RH∞(Stω;T )).
kfkRH∞(Stω;T ) := RX ({f (z) z ∈ Stω})
The following result generalizes Theorem 4.1, since each f ∈ H∞(Stω) defines an
element f ∈ RH∞(Stω;A) by ef (z) = f (z)IX for z ∈ Stω, and kefkRH∞(Stω ;A) =
kfkH∞(Stω).
Proposition 6.1. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach
space X with type p ∈ [1, 2] and cotype q ∈ [2,∞), and let ω > θ(U ). Then there
exists a constant C ≥ 0 such that DA( 1
p − 1
kf (A)kL(DA( 1
p − 1
q , 1) ⊆ D(f (A)) and
q ,1),X) ≤ C kfkRH∞(Stω ;A)
for all f ∈ RH∞(Stω;A).
Proof. First extend Proposition 3.1 to all µ ∈ Mω(R;A) for which RX ({F µω(s)
s ∈ R}) < ∞. To this end, note that the abstract transference principle in [26,
Section 2] extends to A-valued measures, and appeal to Proposition 2.4 instead of
Corollary 2.5. Then proceed as in the proof of Theorem 4.1.
(cid:3)
The other results in the previous sections can be extended to statements about
operator-valued calculi in a similar manner. In particular, if in Proposition 6.1 X is
isomorphic to a complemented subspace of a p-convex and q-concave Banach lattice,
then for each λ > ω there exists a constant C ≥ 0 such that D((λ + iA)
q ) ⊆
D(f (A)) and
p− 1
1
(6.1)
kf (A)kL(D((λ+iA)
1
p
− 1
q ),X) ≤ C kfkRH∞(Stω;A)
for all f ∈ RH∞(Stω;A). Remark 4.4 applies to Proposition 6.1 and (6.1).
Let (rk)k∈N and (r′j)j∈N be mutually independent Rademacher sequences on
a probability space (Ω, P). We say that a Banach space X has property (α) if
there exists a constant C ≥ 0 such that, for all m ∈ N, {xj,k}m
j,k=1 ⊆ X and
{αj,k}m
2(cid:17)1/2
j,k=1 ⊆ C with αj,k ≤ 1 for all j, k ∈ {1, . . . , m},
mXj,k=1
2(cid:17)1/2
mXj,k=1
(6.2)
.
≤ C(cid:16)EE(cid:13)(cid:13)(cid:13)
rkr′jxj,k(cid:13)(cid:13)(cid:13)
(cid:16)EE(cid:13)(cid:13)(cid:13)
rkr′jαj,kxj,k(cid:13)(cid:13)(cid:13)
Each Banach lattice with finite cotype has property (α), and if X has property (α)
then so do the closed subspaces of X and Lp(Ω; X) for each σ-finite measure space
(Ω, µ) and each p ∈ [1,∞).
the H∞(Stω)-calculus for A is R-bounded from DA( 1
kfkH∞(Stω) ≤ 1} ⊆ L(DA( 1
q , 1), X) is R-bounded.
For T equal to the unit ball of C ⊆ L(X), the following theorem implies that
q , 1) to X. That is, {f (A)
p − 1
p − 1
20
JAN ROZENDAAL
Theorem 6.2. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach space
X with property (α), type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ). Then
there exists a constant C ≥ 0 such that
RDA( 1
q ,1),X ({f (A) f ∈ RH∞(Stω;A ∩ T )}) ≤ CRX (T )
p − 1
p − 1
k=1 rkxk (xk)n
k=1 ⊆ D(A)o
for each R-bounded T ⊆ L(X).
Proof. The proof is similar to that of [38, Theorem 12.8]. Write θ := 1
q and
let (rk)k∈N be a Rademacher sequence on [0, 1]. Fix n ∈ N and let Radn(X) :=
{Pn
k=1 ⊆ X} ⊆ L2([0, 1]; X). Let eA be the operator on Radn(X)
with domain
D(eA) =n nXk=1
k=1 rkxk) :=Pn
eU (s)(cid:16) nXk=1
rkxk ∈ Radn(X)(cid:12)(cid:12)(cid:12)(xn)n
k=1 rkAxk forPn
such that eA(Pn
k=1 rkxk ∈ D(eA). Then −ieA gener-
ates the C0-group (eU (s))s∈R ⊆ L(Radn(X)) given by
nXk=1
for s ∈ R and Pn
k=1 rkxk ∈ Radn(X). Note that keU (s)kL(Radn(X)) = kU (s)kL(X)
for all s ∈ R. Moreover, Radn(X) ⊆ L2([0, 1]; X) has type p and cotype q with
τp,Radn(X) ≤ Cpτp,X and cq,Radn(X) ≤ Cqcq,X for constants Cp, Cq ≥ 0 depending
only on p and q that come from the Kahane-Khintchine inequalities. By Proposition
6.1 and Remark 4.4 there exists a constant C1 ≥ 0 independent of n such that
(6.3)
rkxk(cid:17) =
rkU (s)xk
kf (eA)kL(D eA(θ,1),Radn(X)) ≤ C1 kfkRH∞(Stω; eA)
for all f ∈ RH∞(Stω; eA), where eA ⊆ L(Radn(X)) is the algebra of operators
commuting with eA.
Let T ⊆ L(X) be R-bounded and let f1, . . . , fn ∈ RH∞(Stω;A ∩ T ). Define
is well-defined.
j=1 ⊆ Stω and (yj)m
for z ∈ Stω and Pn
k=1 rkxk ∈ Radn(X). We will now show that the range of f is
R-bounded in L(Radn(X)), from which it will follow in particular that f : Stω → eA
Let (r′j )j∈N be a Rademacher sequence on [0, 1], independent of (rk)k∈N. Let
j=1 ⊆ Radn(X). Write yj = Pn
m ∈ N, (zj)m
k=1 rkxjk for
j ∈ {1, . . . , m} and (xjk)n
k=1 ⊆ X. Then [38, Lemma 4.11 and Remark 4.10] yield
a constant C2 ≥ 0 depending only on X such that
=Z 1
0 Z 1
0 (cid:13)(cid:13)(cid:13)
r′j(v)rk(u)f (zj)xjk(cid:13)(cid:13)(cid:13)
mXj=1
mXj=1
nXk=1
2 RX (T )2Z 1
0 Z 1
0 (cid:13)(cid:13)(cid:13)
mXj=1
≤ C2
r′jyj(cid:13)(cid:13)(cid:13)
2 RX (T )2(cid:13)(cid:13)(cid:13)
mXj=1
r′j(v)rk(u)xjk(cid:13)(cid:13)(cid:13)
r′j f (zj)yj(cid:13)(cid:13)(cid:13)
L2([0,1];Radn(X))
L2([0,1];Radn(X))
nXk=1
= C2
(cid:13)(cid:13)(cid:13)
dudv
dudv
2
X
2
X
2
2
.
f (z)(cid:16) nXk=1
rkxk(cid:17) =
nXk=1
rkfk(z)xk
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
21
(6.5)
(6.4)
(6.3) yields
k=1 rkxkkRadn(D(A))
Hence f ∈ RH∞(Stω; eA) with kfkRH∞(Stω ; eA) ≤ C2RX (T ). Combining this with
kf (eA)kL(D eA(θ,1),Radm(X)) ≤ C1C2RX (T ).
Note that(cid:13)(cid:13)Pn
k=1 rkxkkRadn(D(A)) ≤(cid:13)(cid:13)Pn
k=1 rkxkkD( eA) ≤ 2(cid:13)(cid:13)Pn
for all Pn
k=1 rkxk ∈ D(eA), hence D eA(θ, 1) = Radn(DA(θ, 1)) with
(cid:13)(cid:13)(cid:13)
rkxk(cid:13)(cid:13)(cid:13)Radn(DA(θ,1)) ≤(cid:13)(cid:13)(cid:13)
rkxk(cid:13)(cid:13)(cid:13)D eA(θ,1) ≤ 2(cid:13)(cid:13)(cid:13)
rkxk(cid:13)(cid:13)(cid:13)Radn(DA(θ,1))
nXk=1
nXk=1
nXk=1
for allPn
k=1 rkxk ∈ D eA(θ, 1). Also, it is straightforward to check (by regularization)
k=1 rkfk(A)xk for all Pn
k=1 rkxk) = Pn
that f (eA)(Pn
k=1 rkxk ∈ D(f (eA)). Hence
(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)f (eA)(cid:16) nXk=1
rkfk(A)xk(cid:13)(cid:13)(cid:13)L2([0,1];X)
rkxk(cid:17)(cid:13)(cid:13)(cid:13)Radn(X)
nXk=1
≤ C1C2RX (T )(cid:13)(cid:13)(cid:13)
rkxk(cid:13)(cid:13)(cid:13)D eA(θ,1)
nXk=1
rkxk(cid:13)(cid:13)(cid:13)L2([0,1];DA(θ,1))
≤ CRX (T )(cid:13)(cid:13)(cid:13)
nXk=1
for all x1, . . . , xn ∈ DA(θ, 1), where C ≥ 0 is independent of T ⊆ L(X), n ∈
N, f1, . . . fn ∈ RH∞(Stω;A ∩ T ) and x1, . . . , xn ∈ DA(θ, 1). This concludes the
proof.
(cid:3)
(6.4) and (6.5) yield
In the same manner we deduce an R-bounded version of Theorem 4.3, under the
extra assumption of p-convexity for some p > 1. By [41, Corollary 1.f.9] the latter
is equivalent to the assumption of nontrivial type. Recall that any closed subspace
of a Banach lattice with finite cotype has property (α).
Theorem 6.3. Let X be isomorphic to a complemented subspace of a p-convex
and q-concave Banach lattice, for p ∈ (1, 2] and q ∈ [2,∞). Let −iA generate a
C0-group (U (s))s∈R ⊆ L(X), and let λ > ω > θ(U ). Then there exists a constant
C ≥ 0 such that
R
D((λ+iA)
1
p
− 1
q ),X
({f (A) f ∈ RH∞(Stω;A ∩ T )}) ≤ CRX (T )
for each R-bounded T ⊆ L(X).
Proof. We use notation as in the proof of Theorem 6.2. It suffices to show that
there exists a constant C ≥ 0 independent of n ∈ N such that
(6.6)
− 1
q ),Radn(X)) ≤ C kfkRH∞(Stω ; eA)
kf (eA)kL(D((λ+i eA)
1
p
for all f ∈ RH∞(Stω; eA). Indeed, once this has been established, the rest of the
proof is identical to that of Theorem 6.2. To obtain (6.6) apply (6.1) to eA on
Radn(X). To see that the constant C that one gets from this can be chosen to be
independent of n, it suffices by Remark 4.4 to show that Radn(X) is complemented
in L2([0, 1]; X) by a projection Pn ∈ L(L2([0, 1]; X)) with kPnkL(L2([0,1];X)) ≤ C′
for some C′ ≥ 0 independent of n. The latter in turn follows from the fact that
22
JAN ROZENDAAL
X has nontrivial type and from Pisier's characterization in [45] of the K-convex
Banach spaces as the spaces with nontrivial type.
(cid:3)
We do not know whether the assumption in Theorem 6.3 that X has nontrivial
type is necessary.
From Theorems 6.2 and 6.3 one can deduce R-bounded versions of Theorems 5.1
and 5.4 in the obvious manner. Also, as a corollary of our results, for C0-groups we
improve Theorem [33, Theorem 6.1].
Corollary 6.4. Let −iA generate a C0-group (U (s))s∈R ⊆ L(X) on a Banach
space X with property (α), type p ∈ [1, 2] and cotype q ∈ [2,∞). Let ω > θ(U ).
Then {e−ωsU (s) s ∈ R} ⊆ L(DA( 1
If X is isomorphic to a complemented subspace of a p-convex and q-concave
Banach lattice for p ∈ (1, 2] and q ∈ [2,∞), then {e−ωsU (s) s ∈ R} ⊆ L(D(λ +
iA)
Proof. Let ω′ ∈ (θ(U ), ω) and let T be the unit ball of C ⊆ L(X). Now apply
Theorems 6.2 and 6.3 to {e−(ωs+is·) s ∈ R} ⊆ RH∞(Stω′ ;A ∩ T ).
q , X) is R-bounded for each λ > ω.
q , 1), X) is R-bounded.
p − 1
(cid:3)
1
p− 1
As the following example shows, Corollary 6.4 and Theorem 6.3 are sharp.
Example 6.5. Let p ∈ [1,∞) and let (U (s))s∈R ⊆ L(X) be the left translation
group on X := Lp(R) with generator −iA, where Af := if′ for f ∈ D(A) =
W1,p(R). Then D((iA)α) = Hα,p(R), where Hα,p(R) is a Bessel-potential space. It
is shown in [33, Example 6.2] that {U (s) s ∈ [−1, 1]} ⊆ L(Hα,p(R), Lp(R)) is not
R-bounded for α ∈ [0, 1
q ). Hence {e−ωsU (s) s ∈ R} ⊆ L(D((iA)α), X) is not
R-bounded for w ∈ R and α ∈ [0, 1
p − 1
q ).
p − 1
7. Rational approximation
In this section we give an application of the results in previous sections to the
theory of rational approximation of C0-groups. Note that the results in Section 6
can be used to replace the uniform bounds in this section by R-bounds.
Recall that a C0-semigroup (T (t))t≥0 ⊆ L(X) on a Banach space X is expo-
nentially stable if there exist M ≥ 1 and ω > 0 such that kT (t)kL(X) ≤ M e−ωt
for all t ≥ 0. We note that, if −A generates an exponentially stable C0-semigroup
(T (t))t≥0 ⊆ L(X) such that T (t) ∈ L(X) is invertible for each t ≥ 0, then −A in
fact generates the C0-group (U (s))s∈R ⊆ L(X), where U (s) := T (s) for s ≥ 0 and
U (s) := T (−s)−1 for s < 0. Then f (A) is defined as an unbounded operator for
each f ∈ H∞(C+) by a shifted version of the strip-type calculus from Section 2.1.
Lemma 7.1. Let −A generate an exponentially stable C0-semigroup (T (t))t≥0 on
a Banach space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Suppose that T (t) is
invertible for all t ≥ 0. Then there exists a constant C ≥ 0 such that
(7.1)
kf (A)kL(DA( 1
for all f ∈ H∞(C+). For each β > 1
(7.2)
for all f ∈ H∞(C+).
(cid:13)(cid:13)f (A)A−β(cid:13)(cid:13)L(X) ≤ C′ kfkH∞(C+)
q ,1),X) ≤ C kfkH∞(C+)
q there exists a constant C′ ≥ 0 such that
p − 1
p − 1
For n ∈ N let pn and qn be the unique polynomials of degree n and n + 1,
respectively, such that pn(0) = qn(0) = 1 and such that
q ,1),X) < ∞.
sup
p− 1
n∈N(cid:13)(cid:13)(1 − A)n(1 + A)−n(cid:13)(cid:13)L(DA( 1
qn(z) − ez(cid:12)(cid:12)(cid:12)(cid:12) ≤ Cz2n+2
pn(z)
(cid:12)(cid:12)(cid:12)(cid:12)
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
23
Proof. Let f ∈ H∞(C+) and apply Theorem 4.1 and Proposition 4.5 to the strip-
type operator −i(A − ω) and the function f (i · +ω) ∈ H∞(Stω) for ω > 0 large
enough. Then use the composition rule
f (i · +ω)(−i(A − ω)) = f (A),
which is straightforward to prove in the same manner as [23, Theorem 4.2.4]. (cid:3)
Lemma 7.1 applies to the important question of the power-boundedness of the
Cayley transform (1− A)(1 + A)−1 of A, which in turn is equivalent to the stability
of the Crank-Nicholson approximation scheme associated with A.
Corollary 7.2. Let −A generate an exponentially stable C0-semigroup (T (t))t≥0 ⊆
L(X) on a Banach space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Suppose that
T (t) is invertible for all t ≥ 0. Then
for all z in a neighborhood of 0 ∈ C. Let rn := pn
qn
Pad´e approximation of the exponential function.
. Then rn is the n-th subdiagonal
For 1
p − 1
q ∈ [0, 1
p − 1
2 ) the following proposition improves convergence rates ob-
tained in [17, Theorem 4.1] for uniformly bounded C0-semigroups on general Ba-
nach spaces. Note that, on a Banach space X with type p ∈ [1, 2] and cotype
q ∈ [2,∞), for α > 1
q we obtain strong convergence of rn(−tA) to T (t) on
D(Aα) with rate ∩a<α O(n−a+ 1
Proposition 7.3. Let −A generate an exponentially stable C0-semigroup (T (t))t≥0
on a Banach space X with type p ∈ [1, 2] and cotype q ∈ [2,∞). Suppose that T (t)
is invertible for all t ≥ 0. Let α > 1
q ). Then there exists
a constant C ≥ 0 such that
q and a ∈ (0, α − 1
q ), locally uniformly in t.
p − 1
p + 1
p − 1
krn(−tA)x − T (t)xkX ≤ Cta(n + 1)−a kAαxkX
for all t ∈ (0,∞), all n ∈ N with n > α
(rn(−tA))n∈N converges strongly on DA( 1
Proof. Let t ∈ (0,∞) and n ∈ N with n ≥ α
rn(−tz) − e−tz
f (z) :=
2 − 1 and all x ∈ D(Aα). Moreover,
p − 1
q , 1) to T (t), locally uniformly in t.
2 − 1. Set
za
(z ∈ C+).
Then Lemma 7.1 yields a constant C′ ≥ 0 such that
(cid:13)(cid:13)(rn(−tA) − T (t))A−α(cid:13)(cid:13)L(X) =(cid:13)(cid:13)f (A)A−α+a(cid:13)(cid:13)L(X) ≤ C′ kfkH∞(C+) .
By [17, Lemmas 3.3 and 3.5],
kfkH∞(C+) = ta sup
z∈C+
rn(−tz) − e−tz
(tz)a
≤ 2ta(n + 1)−a.
24
JAN ROZENDAAL
Therefore, with C := 2C′,
krn(−tA)x − T (t)xkX ≤(cid:13)(cid:13)(rn(−tA) − T (t))A−α(cid:13)(cid:13)L(X) kAαxkX
≤ Cta(n + 1)−a kAαxkX
for all x ∈ D(Aα), which proves the first statement.
Since krnkH∞(C−) ≤ 1 for all n ∈ N by [18], Lemma 7.1 yields that
q , 1), X)
{rn(−tA) − T (t) n ∈ N, t ≥ 0} ⊆ L(DA( 1
p − 1
is uniformly bounded. The proof is now concluded by what we have already shown
and by the fact that D(A2) is dense in DA( 1
(cid:3)
p − 1
q , 1).
p − 1
The same method that was used in Proposition 7.3 to yield strong convergence
on DA( 1
q , 1) also works for other rational approximation methods. Recall that
a rational function r ∈ H∞(C−) is said to be A-stable if krkH∞(C−) ≤ 1, and r is a
rational approximation (of the exponential function) of order k ∈ N if there exists a
constant C ≥ 0 such that r(z)− ez ≤ Czk+1 for all z in a complex neighborhood
of 0.
Corollary 7.4. Let r be an A-stable rational approximation of order k ∈ N. Let
−A generate an exponentially stable C0-semigroup (T (t))t≥0 on a Banach space X
with type p ∈ [1, 2] and cotype q ∈ [2,∞). Suppose that T (t) is invertible for all
t ≥ 0. Then (r(− t
q , 1) to T (t), locally
uniformly in t ≥ 0.
Proof. Lemma 7.1 yields a constant C ≥ 0 such that
n A)n)n∈N converges strongly on DA( 1
p − 1
(cid:13)(cid:13)r(− t
n A)n − T (t)(cid:13)(cid:13)L(DA( 1
p − 1
q ,1),X) ≤ C(cid:13)(cid:13)r(− t
n·)n − e−t·(cid:13)(cid:13)H∞(C+)
≤ C(krnkH∞(C−) + 1) ≤ 2C
for all n ∈ N and t ≥ 0. Since, by [10, Theorem 3], r(− t
uniformly in t to T (t) on D(Ak+1), the uniform boundedness of
n A)n − T (t) t ≥ 0, n ∈ N(cid:9) ⊆ L(DA( 1
(cid:8)r(− t
p − 1
and the fact that D(Ak+1) is dense in DA( 1
q , 1), X)
n A)n converges locally
p − 1
q , 1) yield the desired statement. (cid:3)
Remark 7.5. If X is isomorphic to a complemented subspace of a p-convex and
q-concave Banach lattice, for p ∈ [1, 2] and q ∈ [2,∞), then the case β = 1
q is
attained in Lemma 7.1. Hence, in the setting of Corollary 7.2,
p − 1
sup
n∈Nk(1 − A)n(1 + A)−n− 1
p + 1
qkL(X) < ∞.
Moreover, one obtains rate O(n−α+ 1
on D(A
q ) in Corollary 7.4.
p − 1
1
p− 1
q ) in Proposition 7.3, and strong convergence
Funding. This work was supported by the Netherlands Organisation for Scientific
Research (NWO) [grant number 613.000.908 "Applications of Transference Princi-
ples"].
Acknowledgements. The author thanks Mark Veraar for numerous helpful sugges-
tions, and the referee for carefully checking the manuscript.
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
25
References
[1] H. Amann. Operator-valued Fourier multipliers, vector-valued Besov spaces, and applications.
Math. Nachr., 186:5 -- 56, 1997.
[2] W. Arendt. Semigroups and evolution equations: Functional calculus, regularity and kernel
estimates. In C.M.Dafermos E. Feireisl, editor, Handbook of Differential Equations, pages
1 -- 85. Elsevier/North Holland, Amsterdam, 2004.
[3] W. Arendt and N. Nikolski. Vector-valued holomorphic functions revisited. Math. Z.,
234(4):777 -- 805, 2000.
[4] P. Auscher, S. Hofmann, M. Lacey, A. McIntosh, and Ph. Tchamitchian. The solution of
the Kato square root problem for second order elliptic operators on Rn. Ann. of Math. (2),
156(2):633 -- 654, 2002.
[5] P. Auscher, A. McIntosh, and A. Morris. Calder´on reproducing formulas and applications to
Hardy spaces. Rev. Mat. Iberoam., 31(3):865 -- 900, 2015.
[6] A. Axelsson, S. Keith, and A. McIntosh. Quadratic estimates and functional calculi of per-
turbed Dirac operators. Invent. Math., 163(3):455 -- 497, 2006.
[7] C. Batty, M. Haase, and J. Mubeen. The holomorphic functional calculus approach to oper-
ator semigroups. Acta Sci. Math. (Szeged), 79(1-2):289 -- 323, 2013.
[8] E. Berkson, T. A. Gillespie, and P. S. Muhly. Generalized analyticity in UMD spaces. Ark.
Mat., 27(1):1 -- 14, 1989.
[9] K. Boyadzhiev and R. deLaubenfels. Spectral theorem for unbounded strongly continuous
groups on a Hilbert space. Proc. Amer. Math. Soc., 120(1):127 -- 136, 1994.
[10] P. Brenner and V. Thom´ee. On rational approximations of semigroups. SIAM J. Numer.
Anal., 16(4):683 -- 694, 1979.
[11] A. P. Calder´on. Ergodic theory and translation-invariant operators. Proc. Natl. Acad. Sci.
USA, 59:349 -- 353, 1968.
[12] A. Carbonaro and O. Dragicevi´c. Functional calculus for generators of symmetric contraction
semigroups. Duke Math. J., 166(5):937 -- 974, 2017.
[13] R. R. Coifman and G. Weiss. Transference methods in analysis. Expository lectures from
the CBMS Regional Conference held at the University of Nebraska, May 31-June 4, 1976.
Regional Conference Series in Mathematics. No.31. Providence, R.I.: American Mathematical
Society (AMS) 59, 1977.
[14] M. Cowling. Harmonic analysis on semigroups. Ann. of Math. (2), 117(2):267 -- 283, 1983.
[15] J. Diestel, H. Jarchow, and A. Tonge. Absolutely summing operators, volume 43 of Cambridge
Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995.
[16] G. Dore. H∞ functional calculus in real interpolation spaces. II. Studia Math., 145(1):75 -- 83,
2001.
[17] M. Egert and J. Rozendaal. Convergence of subdiagonal Pad´e approximations of C0-
semigroups. J. Evol. Equ., 13(4):875 -- 895, 2013.
[18] B. L. Ehle. A-stable methods and Pad´e approximations to the exponential. SIAM J. Math.
Anal., 4:671 -- 680, 1973.
[19] K. Engel and R. Nagel. One-parameter semigroups for linear evolution equations, volume
194 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000. With contributions
by S. Brendle, M. Campiti, T. Hahn, G. Metafune, G. Nickel, D. Pallara, C. Perazzoli, A.
Rhandi, S. Romanelli and R. Schnaubelt.
[20] S. Flory, F. Neubrander, and L. Weis. Consistency and stabilization of rational approximation
schemes for C0-semigroups. In Evolution equations: applications to physics, industry, life sci-
ences and economics. Proceedings of the 7th international conference on evolution equations
and their applications, EVEQ2000 conference, Levico Terme, Italy, October 30 -- November
4, 2000, pages 181 -- 193. Basel: Birkhauser, 2003.
[21] A. Gomilko and Y. Tomilov. On convergence rates in approximation theory for operator
semigroups. J. Funct. Anal., 266(5):3040 -- 3082, 2014.
[22] B. Haak and M. Haase. Square function estimates and functional calculus. Online at
https://arxiv.org/abs/1311.0453, 2013.
[23] M. Haase. The functional calculus for sectorial operators, volume 169 of Operator Theory:
Advances and Applications. Birkhauser Verlag, Basel, 2006.
[24] M. Haase. Functional calculus for groups and applications to evolution equations. J. Evol.
Equ., 7(3):529 -- 554, 2007.
26
JAN ROZENDAAL
[25] M. Haase. A transference principle for general groups and functional calculus on UMD spaces.
Math. Ann., 345(2):245 -- 265, 2009.
[26] M. Haase. Transference principles for semigroups and a theorem of Peller. J. Funct. Anal.,
261(10):2959 -- 2998, 2011.
[27] M. Haase. The functional calculus approach to cosine operator functions. In Recent Trends
in Analysis. Proceedings of the Conference in honor of N.K. Nikolski held in Bordeaux 2011,
pages 123 -- 147. Theta Foundation, 2013.
[28] M. Haase and J. Rozendaal. Functional calculus for semigroup generators via transference.
J. Funct. Anal., 265(12):3345 -- 3368, 2013.
[29] M. Haase and J. Rozendaal. Functional calculus on real interpolation spaces for generators
of C0-groups. Math. Nachr., 289(2-3):275 -- 289, 2016.
[30] M. Hieber and J. Pruss. Functional calculi for linear operators in vector-valued Lp-spaces via
the transference principle. Adv. Differential Equations, 3(6):847 -- 872, 1998.
[31] T. Hytonen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach Spaces. Volume
I: Martingales and Littlewood-Paley Theory, volume 63 of Ergebnisse der Mathematik und
ihrer Grenzgebiete (3). Springer, 2016.
[32] T. Hytonen, J. van Neerven, M. Veraar, and L. Weis. Analysis in Banach spaces. Volume II:
Probabilistic Methods and Operator Theory, volume 67 of Ergebnisse der Mathematik und
ihrer Grenzgebiete (3). Springer, 2017.
[33] T. Hytonen and M. Veraar. R-boundedness of smooth operator-valued functions. Integral
Equations Operator Theory, 63(3):373 -- 402, 2009.
[34] P. Jara, F. Neubrander, and K. Ozer. Rational inversion of the Laplace transform. J. Evol.
Equ., 12(2):435 -- 457, 2012.
[35] N. J. Kalton and L. Weis. The H∞-calculus and sums of closed operators. Math. Ann.,
321(2):319 -- 345, 2001.
[36] N.J. Kalton and L.W. Weis. The H∞-calculus and square function estimates. In Nigel J.
Kalton Selecta, Volume 1, pages 715 -- 764. Springer, 2016.
[37] J. Kisy´nski. On cosine operator functions and one-parameter groups of operators. Studia
Math., 44:93 -- 105, 1972. Collection of articles honoring the completion by Antoni Zygmund
of 50 years of scientific activity. I.
[38] P. C. Kunstmann and L. Weis. Maximal Lp-Regularity for Parabolic Equations, Fourier
Multiplier Theorems and H∞-functional Calculus. In Functional Analytic Methods for Evo-
lution Equations (Levico Terme 2001), volume 1855 of Lecture Notes in Math., pages 65 -- 312.
Springer, Berlin, 2004.
[39] S. Kwapie´n. Isomorphic characterizations of inner product spaces by orthogonal series with
vector valued coefficients. Studia Math., 44:583 -- 595, 1972. Collection of articles honoring the
completion by Antoni Zygmund of 50 years of scientific activity, VI.
[40] C. Le Merdy. On square functions associated to sectorial operators. Bull. Soc. Math. Fr.,
132(1):137 -- 156, 2004.
[41] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. II, volume 97 of Ergebnisse der
Mathematik und ihrer Grenzgebiete [Results in Mathematics and Related Areas]. Springer-
Verlag, Berlin-New York, 1979. Function spaces.
[42] A. Lunardi. Interpolation theory. Appunti. Scuola Normale Superiore di Pisa (Nuova Serie).
[Lecture Notes. Scuola Normale Superiore di Pisa (New Series)]. Edizioni della Normale, Pisa,
second edition, 2009.
[43] A. McIntosh. Operators which have an H∞ functional calculus. In Miniconference on operator
theory and partial differential equations (North Ryde, 1986), volume 14 of Proc. Centre Math.
Anal. Austral. Nat. Univ., pages 210 -- 231. Austral. Nat. Univ., Canberra, 1986.
[44] F. Neubrander, K. Ozer, and T. Sandmaier. Rational approximations of semigroups without
scaling and squaring. Discrete Contin. Dyn. Syst., 33(11-12):5305 -- 5317, 2013.
[45] G. Pisier. Holomorphic semigroups and the geometry of Banach spaces. Ann. of Math. (2),
115(2):375 -- 392, 1982.
[46] J. Pruss and H. Sohr. On operators with bounded imaginary powers in Banach spaces. Math.
Z., 203(3):429 -- 452, 1990.
[47] J. Rozendaal and M. Veraar. Fourier multiplier theorems on Besov spaces under type and
cotype conditions. Banach J. Math. Anal., 11(4):713 -- 743, 2017.
[48] J. Rozendaal and M. Veraar. Fourier Multiplier Theorems Involving Type and Cotype. J.
Fourier Anal. Appl., 24(2):583 -- 619, 2018.
FUNCTIONAL CALCULUS FOR C0-GROUPS USING TYPE AND COTYPE
27
[49] H.-J. Schmeisser and W. Sickel. Vector-valued Sobolev spaces and Gagliardo-Nirenberg in-
equalities. In Nonlinear elliptic and parabolic problems, volume 64 of Progr. Nonlinear Dif-
ferential Equations Appl., pages 463 -- 472. Birkhauser, Basel, 2005.
[50] J. van Neerven. γ-radonifying operators -- a survey. In The AMSI-ANU Workshop on Spectral
Theory and Harmonic Analysis, volume 44 of Proc. Centre Math. Appl. Austral. Nat. Univ.,
pages 1 -- 61. Austral. Nat. Univ., Canberra, 2010.
[51] J. van Neerven, M. Veraar, and L. Weis. Stochastic maximal Lp-regularity. Ann. Probab.,
40(2):788 -- 812, 2012.
[52] M. Veraar. Embedding results for γ-spaces. In Recent trends in analysis. Proceedings of the
conference in honor of Nikolai Nikolski on the occasion of his 70th birthday, Bordeaux,
France, August 31 -- September 2, 2011, pages 209 -- 219. Bucharest: The Theta Foundation,
2013.
Mathematical Sciences Institute, Australian National University, Acton ACT 2601,
Australia, and Institute of Mathematics, Polish Academy of Sciences, ul. ´Sniadeckich
8, 00-656 Warsaw, Poland
E-mail address: [email protected]
|
1711.09479 | 1 | 1711 | 2017-11-26T22:50:34 | On hypercyclic rank one perturbations of unitary operators | [
"math.FA"
] | Recently, S. Grivaux showed that there exists a rank one perturbation of a unitary operator in a Hilbert space which is hypercyclic. Another construction was suggested later by the first and the third authors. Here, using a functional model for rank one perturbations of singular unitary operators, we give yet another construction of hypercyclic rank one perturbation of a unitary operator. In particular, we show that any Carleson set on the circle can be the spectrum of a perturbed (hypercyclic) operator. | math.FA | math |
ON HYPERCYCLIC RANK ONE PERTURBATIONS OF UNITARY
OPERATORS
ANTON BARANOV, VLADIMIR KAPUSTIN, ANDREI LISHANSKII
Abstract. Recently, S. Grivaux showed that there exists a rank one perturbation
of a unitary operator in a Hilbert space which is hypercyclic. Another construction
was suggested later by the first and the third authors. Here, using a functional
model for rank one perturbations of singular unitary operators, we give yet an-
other construction of hypercyclic rank one perturbation of a unitary operator. In
particular, we show that any Carleson set on the circle can be the spectrum of a
perturbed (hypercyclic) operator.
1. introduction
A continuous linear operator T in a Banach (or Fr´echet) space F is said to be
hypercyclic if there exists a vector f ∈ F such that its orbit {T nf }∞
n=0 is dense in
F . In this case the vector f is said to be hypercyclic for T . Though the property
to have a vector with a dense orbit may look rather exotic and "pathological", it
turns out that many natural operators on spaces of analytic functions possess it (e.g.,
some classes of Toeplitz and composition operators, pseudodifferential operators on
the space of all entire functions, weighted shifts, etc.). The theory of hypercyclic
operators and, more generally, dynamics of linear operators, were an active fields of
research during the last three decades. For a comprehensive account of the theory
we refer to the recent monographs [3, 9].
One of the natural questions about hypercyclicity is its stability (or instabil-
ity) with respect to small (in some sense) perturbations. In 1991 K.C. Chan and
J.H. Shapiro [5] showed that there exist hypercyclic operators in a Hilbert space of
the form I +K, where I is the identity operator and the compact operator K belongs
to any Schatten class. It is clear that I + R can not be hypercyclic when R is a finite
rank operator (since in this case I + R∗ has an eigenvalue). But if we replace I
by a unitary operator, hypercyclicity is possible. In 2010 S. Shkarin [16] produced
an example of a unitary operator U such that U + R is hypercyclic for some rank
two operator R. Shkarin then asked whether R can be taken to be of rank one. A
positive answer was soon given by S. Grivaux [8]:
1991 Mathematics Subject Classification. 47A16, 30A76, 30H10.
Key words and phrases. hypercyclic operator, rank one perturbation, inner function, model
space, functional model.
The results of Section 2 were obtained with the support of the RFBR grant 16-01-00635. The
results of Section 3 were obtained with the support of the Russian Science Foundation grant 14-21-
00035.
1
Theorem 1.1 (S. Grivaux, [8]). There exists a unitary operator U in the space ℓ2
and a rank one operator R such that U + R is hypercyclic.
In 2015 A. Baranov and A. Lishanskii [1] gave another proof of this theorem using
a functional model of rank one perturbations of unitary operators due to V. Ka-
pustin [11] and A.D. Baranov and D.V. Yakubovich [2]. The constructions of [8] and
[1] were both based on the following interesting theorem (also due to S. Grivaux)
which gives a condition sufficient for hypercyclicity in the case when the operator
has sufficiently many eigenvectors corresponding to unimodular eigenvalues with a
certain "continuity property".
Theorem 1.2 (S. Grivaux, [7]). Let X be a complex separable infinite-dimensional
Banach space, and let T be a bounded operator on X. Suppose that there exists a
sequence {un}n≥1 of vectors in X having the following properties:
(i) un is an eigenvector of T associated to an eigenvalue λn of T , with λn = 1
and λn are all distinct;
(ii) span{un : n ≥ 1} is dense in X;
(iii) for any n ≥ 1 and any ε > 0, there exists m 6= n such that kun − umk < ε.
Then T is hypercyclic.
Both in [8] and [1] the main technical difficulty was to prove the continuity prop-
erty (iii).
In this note we give yet another construction of hypercyclic rank one
perturbation of a unitary operator. This construction is related to a class of func-
tional spaces studied in [11, 12, 13]. One of the advantages of this approach is that
the continuity property of eigenvectors becomes almost automatic. Another (and
apparently more important) novelty is that now we have a lot of freedom in choosing
the spectrum of the perturbed (hypercyclic) operator (see Theorem 2.1 below).
2. Carleson sets and spaces of analytic functions
Let D = {z ∈ C : z < 1} be the unit disk and T = {z ∈ C : z = 1} be the unit
circle. We denote by H 2 the standard Hardy space in D (identifying it with subspace
− = L2(T) ⊖ H 2 = zH 2. We denote
of L2(T) via its boundary values) and we put H 2
by S and S∗ the shift and backward shift operators (considered on the Hardy space
or on the related spaces introduced below),
Sf (z) = zf (z),
S∗f (z) =
f (z) − f (0)
z
.
Let E be a closed subset of T with zero Lebesgue measure and let {Ij} be at most
countable set of disjoint open arcs Ij ⊂ T such that T \ E =Sj Ij. The set E is said
to be a Carleson set (or a set of finite entropy) if
Xj
Ij log
1
Ij
< ∞
(by I we always denote the normalized Lebesgue measure of I ⊂ T). By the classical
results of L. Carleson [4], sets of finite entropy are precisely those sets that may serve
2
as zero sets of smooth up to the boundary) analytic functions in D. Namely, a nonzero
function f ∈ Hol (D) ∩ C 1(D) (or f ∈ Hol (D) ∩ C ∞(D)) vanishing on E ⊂ T exists
if and only if E is a Carleson set.
Recall that a set E is said to be perfect if it is closed and has no isolated points.
One of the main results of this note is the following theorem:
Theorem 2.1. For any perfect Carleson set E on the unit circle T there exists a
hypercyclic operator of the form U + R, where U is unitary and R of rank one, such
that
σ(U + R) = σp(U + R) = E.
Here σ(T ) and σp(T ) denote the spectrum and the point spectrum of the operator
T , respectively.
Appearance of Carleson sets in this context is not random.
It is interesting to
compare Theorem 2.1 with the following remarkable result of N.G. Makarov [13]: a
set F ⊂ T is a point spectrum of some operator of the form U +K where U is unitary
and K is of trace class if and only if F is a countable union of Carleson sets. We do
not know for which F it is possible to find a hypercyclic operator U + R as above
with point spectrum F (and, in particular, whether there exist countable unions of
Carleson sets for which this is not true).
Another interesting problem is to describe those unitary operators U for which
there exists a hypercyclic rank one perturbation U + R. This seems to be a difficult
problem. Our results imply that the essential spectrum of such operator U can be
any perfect Carleson set.
2.1. Construction of the space H∗(ϕ, ψ). The following class of spaces of an-
alytic functions was introduced in [11] in connection with a model for rank one
perturbations of unitary operators. Let ϕ and ψ be the functions in H 2(T) with the
properties
ϕ = ψ a.e. on T
and
ψ(0) 6= 0.
Let H∗ = H∗(ϕ, ψ) be the space of functions defined by
H∗ =
H 2
ϕ
∩
H 2
−
¯ψ
= {f : f ϕ ∈ H 2, f ¯ψ ∈ H 2
−}
with the norm
kf kH∗ := kf ϕkH 2 = kf ¯ψkH 2
−
.
It is clear that the space H∗ is invariant with respect to S∗. It is also shown in [11]
that the inclusion f ∈ H 2
−/ψ and the fact that ψ(0) 6= 0 imply that we can consider
the values of f ∈ H∗ outside the unit disk and that there exists a finite limit
(zf )∞ = lim
z→∞
zf (z).
Assume now that E ⊂ T is a perfect Carleson set. Let ϕ ∈ C 1(D) be such that
ϕ(0) 6= 0 and {z ∈ T : ϕ(z) = 0} = E. Then it is clear that any function of the
3
form (z − λ)−1, λ ∈ E, belongs to the space H∗ = H∗(ϕ, ϕ) associated with the pair
(ϕ, ϕ). Indeed,
ϕ
z − λ
∈ H 2,
ϕ
z − λ
=
1
λ(cid:18) zϕ
λ − z(cid:19) ∈ H 2
−.
We will use the fact that the functions in the space H∗ have analytic continuation
through T \ E.
Lemma 2.2. Let µ ∈ T \ E. Then any function in H∗ admits analytic continuation
into a neighborhood of µ and the mapping f 7→ f (µ) is a bounded functional on H∗.
Proof. By definition, f ∈ H∗ if and only if ϕf ∈ H 2 and ϕ¯z ¯f ∈ H 2. The last
condition can be rewritten as follows: there exists a function g meromorphic and of
Nevanlinna class in {z > 1} such that g = f a.e. on T and
If µ ∈ T \ E, then there exist an open neighborhood V of µ and δ > 0 such that
1
z
g(cid:16)1
¯z(cid:17)ϕ(z) ∈ H 2.
ϕ(z) ≥ δ, z ∈ D ∩ V . It follows that, for some ε > 0,
1<r<1+εZ{eiθ−µ<ε}
sup
g(reiθ)2dθ < ∞.
Hence, g is a usual analytic continuation of f through the arc {eiθ − µ < ε}.
Moreover,
f (µ)2 ≤
f (z)2dm2(z) ≤
1
πε2 Z{z−µ<ε}
1
πε2δ2 kϕf k2
H 2.
Here m2 is the planar Lebesgue measure.
(cid:3)
3. Construction of a hypercyclic rank one perturbation.
In this section we use the spaces of analytic functions from the previous section to
construct a rank one perturbatrion of a unitary operator which satisfies the conditions
of Grivaux' theorem (Theorem 1.2).
As above let E ⊂ T be a perfect Carleson set and let ϕ ∈ C 1(D) be such that
ϕ(0) 6= 0 and {z ∈ T : ϕ(z) = 0} = E. Consider the space H∗ associated with the
pair (ϕ, ϕ). Then any function of the form (z − λ)−1 ∈ H∗ for λ ∈ E. It follows from
the inclusion ϕ ∈ C 1(D) that
whence, by the Lebesgue dominated convergence theorem,
(1)
→ 0 as µ → λ, µ ∈ E.
ϕ
z − µ
(cid:13)(cid:13)(cid:13)
ϕ(z)
z − λ(cid:12)(cid:12)(cid:12) < ∞,
ϕ
sup
λ∈E,z∈T(cid:12)(cid:12)(cid:12)
z − λ(cid:13)(cid:13)(cid:13)H 2
H0 = SpanH∗(cid:26) 1
−
: j ∈ J(cid:27).
z − λj
4
Now, let {λj}j∈J be a countable subset of E without isolated points and put
Theorem 3.1. The operator S∗H0 is a hypercyclic operator which is a rank one
perturbation of some unitary operator. Moreover, σ(S∗H0) = σp(S∗H0) = E where
E = {¯λ : λ ∈ E}.
z−λj
z−λj
cj
z−λj
cj
that S∗ acts isometrically on H1. We have the following lemma:
To prove the converse inclusion, let us note first that finite sums of the form
∈ H1 (we consider only finite sums), then g(0) = −Pj
λj (z−λj ), so (zS∗g)∞ = 0. Since (zf )∞ is a continuous linear
functional on H∗, we conclude that (zf )∞ = 0 whenever f ∈ S∗H1. Thus, S∗H1 ⊂
are dense in eH1. Let f ∈ H0 and (zf )∞ = 0. For any ε > 0 we can choose
z−λj(cid:13)(cid:13)(cid:13) < ε. Since h 7→ (zh)∞ is
Put H1 = {f ∈ H0 : f (0) = 0} and eH1 = {f ∈ H0 : (zf )∞ = 0}. Then it is clear
Lemma 3.2. S∗H1 = eH1.
Proof. If g =Pj
In this case S∗g = Pj
eH1.
P cj
such that (cid:13)(cid:13)(cid:13)f −P cj
a finite linear combination P cj
(cid:12)(cid:12)(cid:12)X cj(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18)z(cid:16)X cj
f (z) =X cj
(a finite sum) and (zf )∞ = P cj = 0. Put g = P λj cj
z(cid:18)X λjcj
a bounded functional on H0, there exists a constant A > 0 (independent of f ) such
that
− f(cid:17)(cid:19)∞(cid:12)(cid:12)(cid:12)(cid:12) < Aε.
− Pj cj
∈ eH1
z − λ0
Now let f (z) = P cj
Then g(0) = 0 (i.e., g ∈ H1) and
z−λj
Let λ0 ∈ E be fixed which is independent on f . Then it is easy to see that
and kf − f k ≤ ε(1 + Ak(z − λ0)−1kH ∗).
+X cj(cid:19) = f.
z − λj
z − λj
z − λj
cj
λj
= 0.
.
z−λj
1
S∗g =
Thus, eH1 ⊂ S∗H1.
(cid:3)
3.1. Proof of Theorem 3.1. First we verify that the operator S∗H0 satisfies the
conditions of Theorem 1.2. It is clear that the functions fj = 1
are eigenvectors
z−λj
of the operator S∗H0 with eigenvalues 1
λj
. Indeed,
S∗fj(z) =
1
z(cid:16) 1
z − λj
+
1
λj(cid:17) =
1
λj(z − λj)
.
By the definition of the space H0 the set {fj}j∈J is complete in H0. Since the set
{λj}j∈J has no isolated points, the continuity property (iii) in Theorem 1.2 follows
from (1). Thus, S∗H0 is hypercyclic.
5
Next we need to show that S∗H0 is a rank one perturbation of some unitary
It is clear that H0 = H1 ⊕ H2, where
operator. Let H1 and eH1 be as above.
H2 = Lin{g} for some g ∈ H0 such that g(0) = 1. By Lemma 3.2 S∗H1 = eH1
and so the closed subspace S∗H1 has codimension 1 in H0. Then we can define a
unitary operator U as follows: let UH1 = S∗H1 and let UH2 be an arbitrary unitary
operator of one-dimensional subspace H2 onto one-dimensional subspace H0 ⊖ S∗H1.
Clearly, S∗H0 will be a rank one perturbation of U.
It remains to show that σ(S∗H0) = σp(S∗H0) = E. It is clear that E ⊂ σp(S∗H0).
We show that for λ /∈ E the equation S∗f − λf = g has a unique solution f ∈ H0
for any g ∈ H0. Indeed, this equation is equivalent to
f =
zg + f (0)
1 − λz
.
If µ = λ−1 ∈ D or µ ∈ T\E, then by lemma 2.2 any function g ∈ H0 is analytic in the
neighborhood of µ and so there exists a unique c ∈ C, namely c = µg(µ), such that
λ(z−µ) belongs to ϕ−1H 2. The inclusion ϕ¯z ¯f ∈ H 2 follows immediately from the
f = zg−c
inclusion ϕ¯z¯g ∈ H 2. If λ ∈ D, then ϕf ∈ H 2 for any f of the form f = zg−c
λ(z−µ) . Also,
ϕ¯z ¯f = ϕ¯z
¯z¯g − ¯c
1 − ¯λ¯z
=
ϕ¯z¯g − ¯cϕ
z − ¯λ
∈ H 2
if and only if ¯c = (ϕ¯z¯g)(¯λ)/ϕ(¯λ). Thus, we have seen that there is a unique solution
f ∈ H∗ for any g ∈ H∗. If g is a finite linear combination of (z − λj)−1, then it is
easy to see that f also is of this form. Hence, f ∈ H0 whenever g ∈ H0. This proves
Theorem 3.1.
(cid:3)
3.2. A link to nearly invariant subspaces and Clark measures. In this sub-
section we discuss the relation of our construction with Hitt–Sarason theory of nearly
invariant subspaces. A closed subspace X ⊂ H 2 is said to be nearly invariant, if
for any f ∈ X such that f (0) = 0 we have S∗f ∈ X. Recall that by the Beurling
theorem any S∗-invariant subspace is of the form KΘ = H 2 ⊖ ΘH 2 for some inner
function Θ (see, e.g., [14]). The structure of nearly invariant subspaces was described
by D. Hitt [10] and D. Sarason [15].
Theorem 3.3 (Hitt, Sarason, [10, 15]). If X is nearly invariant, then X = gKΘ,
where Θ is an inner function, g is outer, and the mapping Mg : f 7→ gf is an
isometry on KΘ. In particular, Mg maps KΘ isometrically onto X.
Let us show that
is a nearly-invariant subspace of H 2. Indeed, for finite sums ϕP cj
dense in ϕH0) we have
z−λj
(which are
ϕH0 = ϕSpanH∗(cid:26) 1
z − λj(cid:12)(cid:12)(cid:12)(cid:12)z=0
X cj
6
: λ ∈ E(cid:27).
z − λ
= −
cj
λj
= 0,
and so
1
z(cid:18)ϕX cj
z − λj(cid:19) =
1
z(cid:18)ϕX(cid:16) cj
z − λj
+
cj
λj(cid:17)(cid:19) = ϕ(cid:18)X cj
λj(z − λj)(cid:19) ∈ ϕH0.
Therefore, by the Hitt–Sarason theorem ϕH0 = gKΘ for some g and Θ.
If f ∈ H0, then ϕf
g ∈ Kθ and
kf kH0 = kϕf kH 2 =(cid:13)(cid:13)(cid:13)
ϕf
g (cid:13)(cid:13)(cid:13)H 2
and so the mapping Mϕ/g : f 7→ ϕf /g is a unitary operator from H0 onto KΘ. It is
clear that operators S∗H1 (where H1 = {f ∈ H0 : f (0) = 0}) and S∗{f ∈KΘ: f (0)=0}
are unitarily equivalent. Since S∗H1 coincides on H1 with the unitary operator U
constructed in Theorem 3.1, we conclude that U = M −1
ϕ/gUMϕ/g is a unitary operator
on KΘ which is a rank one perturbation of S∗KΘ.
Unitary rank one perturbations of S∗KΘ were described by D.N. Clark in a seminal
paper [6]. This description involves a construction of a special family of measures.
Let α ∈ T. Then the function α+θ
α−θ has positive real part in D, and so there exists
a finite positive measure µα on T (singular with respect to the Lebesgue measure)
such that
Re
α + θ(z)
α − θ(z)
=
1
π ZT
1 − z2
τ − z2 dµα(τ ),
z ∈ D.
Clark showed that the set of unitary rank one perturbations of S∗KΘ can be
parametrized by α ∈ T and their spectral measures are exactly the measures µα.
Corollary 3.4. Let E ⊂ T be a perfect Carleson set and let ϕ be the corresponding
smooth function with boundary zero set E. Then the space ϕH0 is a nearly invariant
subspace. Moreover, if Θ is the associated inner function from the Hitt–Sarason
theorem. then the spectral measure of the unitary operator U which has a hypercyclic
rank one perturbation is the Clark measure µα for Θ.
It would be interesting to use this relation to characterize (at least a subclass of)
unitary operators U which have a hypercyclic rank one perturbation. The difficulty
here is that the proof of the Hitt–Sarason theorem is nonconstructive and, even if we
have explicit knowledge of ϕ and H0, it is difficult to translate it into the information
about Θ. In particular, it would be interesting to know whether the corresponding
Clark measure µα can have a nontrivial singular continuous (i.e., without atoms)
part.
References
[1] A. Baranov, A. Lishanskii, On S. Grivaux' example of a hypercyclic rank one perturbation of
a unitary operator, Arch. Math. 104 (2015), 3, 223–235.
[2] A. Baranov, D. Yakubovich, Completeness and spectral synthesis of nonselfadjoint one-
dimensional perturbations of selfadjoint operators, Adv. Math. 302 (2016), 740–798.
[3] F. Bayart, E. Matheron, Dynamics of Linear Operators, Cambridge University Press, 2009.
[4] L. Carleson, Sets of uniqueness for functions regular in the unit circle, Acta Math. 87 (1952),
325–345.
7
[5] K.C. Chan, J.H. Shapiro, The cyclic behavior of translation operators on Hilbert spaces of
entire functions, Indiana Univ. Math. J. 40, 4 (1991), 1421–1449.
[6] D.N. Clark, One-dimensional perturbations of restricted shifts, J. Anal. Math. 25 (1972),
169–191.
[7] S. Grivaux, A new class of frequently hypercyclic operators, Indiana Univ. Math. J. 60 (2011),
1177–1201.
[8] S. Grivaux, A hypercyclic rank one perturbation of a unitary operator, Math. Nachr. 285
(2012), 5-6, 533–544.
[9] K.-G. Grosse-Erdmann, A. Peris Manguillot, Linear Chaos, Universitext, Springer, London,
2011.
[10] D. Hitt, Invariant subspaces of H 2 of an annulus, Pacific J. Math. 134 (1988), 101–120.
[11] V.V. Kapustin, One-dimensional perturbations of singular unitary operators, Zapiski Nauchn.
Sem. POMI 232 (1996), 118–122; English transl.: J. Math. Sci. (New York) 92 (1998), 1,
3619–3621.
[12] V.V. Kapustin, Spectral analysis of almost unitary operators, Algebra i Analiz 13 (2001), 5,
44–68; English transl.: St. Petersburg Math. J. 13 (2002), 5, 739–756.
[13] N.G. Makarov, Unitary point spectrum of almost unitary operators, Zap. Nauchn. Sem. LOMI
126 (1983), 143–149; English transl.: J. Soviet Math. 27 (1984), 1, 2517–2520.
[14] N.K. Nikolski, Operators, Functions, and Systems: an Easy Reading, Math. Surveys Monogr.,
Vol. 92, AMS, Providence, RI, 2002.
[15] D. Sarason, Nearly invariant subspaces of the backward shift, Oper. Theory: Adv. Appl. 35
(1988), 481–493.
[16] S. Shkarin, A hypercyclic finite rank perturbation of a unitary operator, Math. Ann. 348
(2010), 379–393.
Anton Baranov,
Department of Mathematics and Mechanics, St. Petersburg State University,
St. Petersburg, Russia,
[email protected]
Vladimir Kapustin,
St. Petersburg Department of V.A. Steklov Mathematical Institute
St. Petersburg, Russia,
[email protected]
Andrei Lishanskii,
Chebyshev Laboratory, St. Petersburg State University,
St. Petersburg, Russia,
[email protected]
8
|
1111.5109 | 1 | 1111 | 2011-11-22T06:10:27 | Boundedness of oscillatory integral operators and their commutators on weighted Morrey spaces | [
"math.FA"
] | It is proved that both oscillatory integral operators and fractional oscillatory integral operators are bounded on weighted Morrey spaces. The corresponding commutators generated by $BMO$ functions are also considered. | math.FA | math |
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND
THEIR COMMUTATORS ON WEIGHTED MORREY SPACES
ZUNWEI FU, SHAOGUANG SHI♯, AND SHANZHEN LU
Abstract. It is proved that both oscillatory integral operators and fractional oscil-
latory integral operators are bounded on weighted Morrey spaces. The corresponding
commutators generated by BM O functions are also considered.
1. Introduction
To investigate the local behavior of solutions to second order elliptic partial differential
equations, Morrey [14] first introduced the classical Morrey space Mp,q(Rn) with the norm
kf kMp,q(Rn) = sup
B⊂Rn 1
B1− p
q ZB
f (x)pdx! 1
p
,
where f ∈ Lp
Mp,q(Rn) was an expansion of Lp(Rn) in the sense that Mp,p(Rn) = Lp(Rn).
loc(Rn) and 1 ≤ p ≤ q < ∞. Here and after, B denotes any balls in Rn.
In [2], Chiarenza and Frasca obtained the boundedness of Hardy-Littlewood maximal
function
on Mp,q(Rn).
Mf (x) = sup
B∋x
1
BZB
f (y)dy,
(1.1)
The Calder´on-Zygmund singular integral operator is defined by
K(x − y)f (y)dy,
(1.2)
where K is a Calder´on-Zygmund kernel(CZK). We say a kernel K ∈ C 1(Rn/{0}) is a
CZK if it satisfies
eT f (x) = p.v.ZRn
K(x) ≤
C
xn ,
∇K(x) ≤
C
xn+1
(1.3)
(1.4)
2010 Mathematics Subject Classification. Primary 42B20; Secondary 42B25.
Key words and phrases. weighted Morrey space, oscillatory integral, commutator, Ap weights.
♯ Corresponding author: [email protected].
This work was partially supported by NSF of China (Grant Nos. 10901076, 10931001 and 11171345)
and the Key Laboratory of Mathematics and Complex System (Beijing Normal University), Ministry of
Education, China.
1
2
and
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
Za<x<b
K(x)dx = 0,
(1.5)
for all a, b with 0 < a < b. Chiarenza and Frasca [2] showed the boundedness of eT on
Mp,q(Rn). Here and subsequently, C will denote a positive constant which may vary from
line to line but will remain independent of the relevant quantities.
For 0 < α < n, the fractional integral operator Iα is defined by
Iαf (x) =ZRn
f (y)
x − yn−α dy.
The boundedness of Iα on Mp,q(Rn) was established by Adams in [1]. For some works on
the boundedness for the multilinear singular integral operators on Morrey type spaces,
we refer to [5], [21] and [23].
In [7], Komori and Shirai gave the definition of weighted Morrey space, which is a
natural generalization of weighted Lebesgue space. Let 1 ≤ p < ∞, 0 < k < 1 and w be
a function. Then the weighted Morrey space Mp,k(w) was defined by
Mp,k(w) = {f ∈ Lp
loc(w) : kf kMp,k(w) < ∞},
where
kf kMp,k(w) = sup
B (cid:18) 1
w(B)kZB
f (x)pw(x)dx(cid:19) 1
p
,
and the supremum is taken over all balls B ⊂ Rn. It is obviously that if w = 1, k = 1 − p
q ,
then Mp,k(w) = Mp,q(Rn). For w ∈ Ap(1 ≤ p < ∞), if k = 0, then Mp,0(w) = Lp(w) and if
k = 1, Mp,1(w) = L∞(w). Also, Komori and Shirai [7] obtained the boundedness of Hardy-
Littlewoood maximal operator M and Calder´on-Zygmund singular integral operator eT
on Mp,k(w) with 1 < p < ∞. When p = 1, the corresponding weighted weak (1, 1) type
boundedness was also true [7]. Here Ap and the following A(p,q) denote the Muckenhoupt
classes [15]
and
Ap : sup
w(x)1−p′
B (cid:18) 1
BZB
B (cid:18) 1
BZB
w(x)dx(cid:19)(cid:18) 1
BZB
w(x)qdx(cid:19) 1
q(cid:18) 1
BZB
A(p,q) : sup
w(x)−p′
dx(cid:19)p−1
dx(cid:19) 1
p′
≤ C, 1 < p < ∞
≤ C, 1 < p, q < ∞
respectively. Here 1/p + 1/p′ = 1.
In the fractional case, we need to consider a weighted Morrey space with two weights
which also introduced by Komori and Shirai [7]. Let 1 ≤ p < ∞, 0 < k < 1. For two
weights w1 and w2,
Mp,k(w1, w2) =(cid:8)f : kf kMp,k(w1,w2) < ∞(cid:9) ,
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
3
with the norm
kf kMp,k(w1,w2) = sup
B (cid:18) 1
w2(B)kZB
f (x)pw1(x)dx(cid:19) 1
p
,
and the supremum is taken over all balls B ⊂ Rn. If w1 = w2 = w, then we denote
Mp,k(w1, w1) = Mp,k(w2, w2) = Mp,k(w).
In [7], the authors considered the weighted
estimates of the fractional maximal operator Mα
Mαf (x) = sup
B∋x
1
Bn−αZB
f (y)dy
(1.6)
and the fractional integral operator Iα with 0 < α < n.
It is worth pointing out that the kernel in (1.2) is convolution kernel. However, there
were many kinds of operators with non-convolution kernels, such as Fourier transform
and Radon transform [13] which both are versions of oscillatory integrals. The object we
consider in this paper is a class of oscillatory integrals due to Ricci and Stein [17]
T f (x) = p.v.ZRn
eiP (x,y)K(x − y)f (y)dy,
(1.7)
where P (x, y) is a real valued polynomial defined on Rn × Rn, and K is a CZK.
It is well known that the oscillatory factor eip(x,y) makes it impossible to establish the
weighted norm inequalities of (1.7) by the method as in the case of Calder´on-Zygmund
operators or fractional integrals. In [18], Sato established the weighted weak (1, 1) type
estimate of T . The strong type weighted norm inequality of T was proved by Lu and
Zhang in [11] with a more general case. For the other classical works about oscillatory
integral, we refer to [9], [10] and [12]. Inspired by [7] and [18], we will study the weak
type estimates for T on weighted morrey space, that is
Theorem 1.1. If K is a CZK, w ∈ A1 and 0 < k < 1, then there exists a constant C
independent on the coefficients of P such that
λw ({x ∈ B : T f (x) > λ}) ≤ Ckf kM1,k(w)w(B)k.
sup
λ>0
A distribution kernel K is called a standard Calder´on-Zygmund kernel(SCZK) if it
satisfies the following hypotheses
K(x, y) ≤
C
x − yn , x 6= y
and
∇xK(x, y) + ∇yK(x, y) ≤
C
x − yn+1 , x 6= y.
(1.8)
(1.9)
The corresponding Calder´on-Zygmund integral operator eS and oscillatory integral oper-
ator S are defined by
K(x, y)f (y)dy
(1.10)
eSf (x) = p.v.ZRn
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
Sf (x) = p.v.ZRn
eiP (x,y)K(x, y)f (y)dy.
(1.11)
where P (x, y) is a real valued polynomial defined on Rn × Rn. In 1992, Lu and zhang
[11] proved that S was bounded on Lp(w) with 1 < p < ∞, w ∈ Ap by the methods
of interpolation of operators with change of measures [22]. In [17], Ricci and Stein also
introduced the standard fractional Calder´on-Zygmund kernel(SFCZK) Kα with 0 < α <
n, where the condition (1.8) and (1.9) were replaced by
4
and
and
Kα(x, y) ≤
C
x − yn−α , x 6= y
(1.12)
(1.13)
(1.14)
x − yn+1−α , x 6= y.
The corresponding fractional oscillatory integral operator is defined by
∇xKα(x, y) + ∇yKα(x, y) ≤
C
Sαf (x) =ZRn
eiP (x,y)Kα(x, y)f (y)dy.
where P (x, y) is also a real valued polynomial defined on Rn × Rn. Obviously, when
α = 0, S0 = S and K0 = K. Recently, the authors of this paper obtained the weighted
boundedness of Sα in [20]. Partly motivated by the idea from [7] and the results of [11]
as well as [20], we now give another two results of this paper
Theorem 1.2. Let 1 < p < ∞, 0 < k < 1 and w ∈ Ap. If the Calder´on-Zygmund
singular integral operator eS is of type (L2(Rn), L2(Rn)), then for any real polynomial
P (x, y), there exists constant C > 0 such that
kSf (x)kMp,k(w) ≤ Ckf kMp,k(w),
where its norm depends only on the total degree of P , but not on the coefficients of P .
Theorem 1.3. Let 1
q = 1
p − α
n , 1 < p < n
α , 0 < k < p
q , 0 < α < n and w ∈ A(p,q). For
any real polynomial P (x, y), there exists constant C > 0 such that
kSαf (x)kM
q,
qk
p
(wq) ≤ Ckf kMp,k(wp,wq),
where its norm depends only on the total degree of P , but not on the coefficients of P . If
p = 1 and w ∈ A(1,q) with q = n
n−α , then for all λ > 0, there exists constant C > 0 such
that
wq ({x ∈ B : Sαf (x) > λ}) ≤
C
λq kf kq
M1,k(w,wq)(wq(B))kq.
We prove Theorem 1.1-Theorem 1.3 in Section 2. In Section 3, we set up the weighted
norm inequalities for the corresponding commutators of S and Sα, respectively. Through-
out this paper all definitions and notations are standard. A weight w is a locally integrable
function on Rn which takes values in (0, ∞) almost everywhere. B = B(x0, r) denotes the
ball with center x0 and radius r. Given λ > 0, λB = B(x0, λr).
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
5
2. Weighted estimates for oscillatory integral operators
We begin this section with some properties of Ap weight classes which play important
role in the proofs of our main results.
Lemma 2.1. [4] Let 1 ≤ p < ∞, and w ∈ Ap. Then the following statements are true
(1) There exists a constant C such that
w(2B) ≤ Cw(B).
When w satisfies this condition, we say w satisfies doubling condition.
(2) There exists a constant C > 1 such that
w(2B) ≥ Cw(B).
(2.1)
(2.2)
When w satisfies this condition, we say w satisfies reverse doubling condition.
(3) There exist two constant C and r > 1 such that the following reverse Holder
inequality holds for every ball B ⊂ Rn
(cid:18) 1
BZB
(4) For all λ > 1, we have
r
w(x)rdx(cid:19) 1
≤ C(cid:18) 1
BZB
w(x)dx(cid:19) .
w(λB) ≤ Cλnpw(B).
(2.3)
(2.4)
(5) There exist two constant C and δ > 0 such that for any measurable set Q ⊂ B
If w satisfies (2.5), we say w ∈ A∞.
w(Q)
w(B)
B(cid:19)δ
≤ C(cid:18) Q
.
(6) For all 1
p + 1
p′ = 1 we have
A∞ = [1<p<∞
Ap, w1−p′
∈ Ap′
.
(2.5)
(2.6)
Our argument based heavily on the following well-known results about T , S and Sα.
Lemma 2.2. [18] If K is a CZK, w ∈ A1, then there exist constant C > 0 independent
on the coefficients of P such that
λw ({x ∈ Rn : T f (x) > λ}) ≤ Ckf kL1(w).
sup
λ>0
Lemma 2.3. [11] If K is a SCZK, w ∈ Ap(1 < p < ∞) and the Calder´on-Zygmund
singular integral operator eS is of type (L2(Rn), L2(Rn)), then for any real polynomial
P (x, y), there exists C > 0 such that
kSf kLp(w) ≤ Ckf kLp(w),
where its norm depends only on the total degree of P , but not on the coefficients of P .
6
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
Lemma 2.4. [20], [7] Let w ∈ A(p,q) , 0 < α < n, 1 < p < n
α and 1
p − 1
q = α
n . Then there
exists constant C > 0 independent of P such that such that
kSαf kLq(wq) ≤ Ckf kLp(wp),
and
kIαf (x)kM
q,
qk
p
(wq) ≤ Ckf kMp,k(wp,wq).
When p = 1 and w ∈ A(1,q) with q = n
such that
n−α , then for all λ > 0, there exists constant C > 0
wq ({x ∈ B : Iαf (x) > λ}) ≤
C
λq kf kq
M1,k(w,wq)(wq(B))kq.
We first give the proof of Theorem 1.1. Decompose f = f χ2B + fχ(2B)c := f1 + f2. For
any given λ > 0, we write
w ({x ∈ B : T f (x) > λ}) ≤ w(cid:18)(cid:26)x ∈ B : T f1(x) >
:= I + II.
An application of (2.1) and Lemma 2.2 yields that
λ
2(cid:27)(cid:19) + w(cid:18)(cid:26)x ∈ B : T f2(x) >
λ
2(cid:27)(cid:19)
Next we turn to deal with the term II. An elementary estimate shows
λ
I ≤(cid:18)w(cid:26)x ∈ Rn : T f1(x) >
w(cid:18)(cid:26)x ∈ B : T f2(x) >
f (x)w(x)dx ≤
C
λ Z2B
2(cid:27)(cid:19) ≤
2(cid:27)(cid:19) =Z{x∈B:T f2(x)> λ
λZ{x∈B:T f (x)> λ
2 }
≤
C
λ
2 }
w(x)dx
T f2(x)w(x)dx.
C
λ
kf kM1,k(w)w(B)k. (2.7)
We note that for x ∈ B and y ∈ (2B)c, x0 − y < Cx − y. Applying (1.3), we conclude
that
eiP (x,y)K(x − y)f2(y)dy
f2(y)
x − yn dy
T f2(x) = Z(2B)c
≤ CZRn
≤ CZx0−y>2r
∞Xj=1Z2j r<x0−y<2j+1r
2jBZ2j+1B
∞Xj=1
≤ C
≤ C
1
f2(y)
x0 − yn dy
f (y)
x0 − yn dy
f (y)dy.
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
7
Holder inequality and the Ap condition imply that
II ≤
C
λ
∞Xj=1
1
2jBZ2j+1B
f (y)dyZ{x∈B:T f (x)> λ
2 }
w(x)dx ≤
C
λ
Then, Theorem 1.1 is a by-product of (2.7)-(2.8).
kf kM1,k(w)w(B)k.
(2.8)
We now give the proof of Theorem 1.2. As in [7], our method is adapted from [3] in
the case of Lebesgue measure. Let 1 < p < ∞, 0 < k < 1, K be a SCZK. It suffices to
show that
1
w(B)kZB
Sf (x)pw(x)dx ≤ Ckf kp
Mp,k(w).
(2.9)
For a fixed ball B = B(x0, r), we decompose f = f χ2B + fχ(2B)c := f1 + f2 and consider
the corresponding splitting
Sf (x) =Z2B
eiP (x,y)K(x, y)f (y)dy +Z(2B)c
=: Sf1(x) + Sf2(x).
eiP (x,y)K(x, y)f (y)dy
Since S is a linear operator, so we get
1
w(B)kZB
Sf (x)pw(x)dx ≤
1
w(B)kZB
(Sf1(x)p + Sf2(x)p)w(x)dx := I + II. (2.10)
It follows from Lemma 2.3 and (2.1) that
I ≤
1
w(B)kZRn
Sf1(x)pw(x)dx ≤
C
w(B)kZ2B
f (x)pw(x)dx ≤ Ckf kp
Mp,k(w).
(2.11)
We are now in a position to estimate the term II. We note that for x ∈ B and
y ∈ (2B)c, x0 − y < Cx − y. Applying (1.8), we conclude that
eiP (x,y)K(x, y)f2(y)dy
f2(y)
x − yn dy
Sf2(x) = Z(2B)c
≤ CZRn
≤ CZx0−y>2r
∞Xj=1Z2j r<x0−y<2j+1r
2jBZ2j+1B
∞Xj=1
≤ C
≤ C
f2(y)
x0 − yn dy
f (y)
x0 − yn dy
1
f (y)dy.
8
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
Holder inequality and the Ap condition imply that
Z2j+1B
f (y)dy ≤(cid:18)Z2j+1B
f (y)pw(y)dy(cid:19) 1
p(cid:18)Z2j+1B
p(cid:18)Z2j+1B
k
w(y)− p′
p′
p dy(cid:19) 1
dy(cid:19)1− 1
p′
≤ Ckf kMp,k(w)w(2j+1B)
w(y)1−p′
≤ Ckf kMp,k(w)2j+1Bw(2j+1B)
1
p (k−1).
Then, using (2.2) we obtain
II ≤ Ckf kp
Mp,k(w) ∞Xj=1
w(B)
1−k
p
w(2j+1B)
p !p
1−k
≤ Ckf kp
Mp,k(w).
(2.12)
We get (2.9) by (2.10), (2.11) and (2.12). The proof of Theorem 1.2 is completed.
For the last part of this section, we show the proof of Theorem 1.3. Theorem 1.3 is a
byproduct of Lemma 2.4 and the following observation
In fact, if 1 < p < ∞,
1
f (y)
Sαf (x) ≤Z
p Z Sαf (x)qwq ≤
qk
w(B)
x − yn−α dy = Iα(f )(x).
1
w(B)
p Z Iαf (x)qwq
qk
In the same manner, we can obtain the result for p = 1.
≤ Ckf kMp,k(wp,wq).
3. Weighted estimates for the commutators
The aim of this section is to set up the weighted boundedness for the commutators
formed by S(Sα) and BMO(Rn) functions.
A locally integrable function b is said to be in BMO(Rn) if for any ball B ⊂ Rn
kbkBM O(Rn) = sup
B
1
BZB
b(x) − bBdx < ∞,
where bB = 1
BRB b(x)dx.
We next formulate some remarks about BMO(Rn).
Lemma 3.1. [8],[24] Let 1 ≤ p < ∞, b ∈ BMO(Rn). Then for any ball B ⊂ Rn, the
following statements are true
(1) There exist constants C1, C2 such that for all α > 0
{x ∈ B : b(x) − bB > α} ≤ C1Be−C2α/kbkBM O(Rn ).
(3.1)
The inequality (3.1) is also called John-Nirenberg inequality.
kbkBM O(Rn) ∼ sup
kbkBM O(Rn) ∼ sup
B
p
.
b(x) − bBpdx(cid:19) 1
B (cid:18) 1
BZB
BZB
b(x) − adx.
inf
a∈R
1
(2)
(3)
(3.2)
(3.3)
(3.4)
(3.5)
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
9
(2)
b2λB − bB ≤ 2nλkbkBM O(Rn).
Lemma 3.2. [16] Let w ∈ A∞. Then the following statements are equivalent
(1)
kbkBM O(w) = sup
B
b(x) − bB,ww(x)dx.
1
w(B)ZB
where BMO(w) = {b : kbkBM O(w) < ∞} and bB,w = 1
w(B)RB b(y)w(y)dy.
For a locally integrable function b, the commutator formed by S(Sα) and b are defined
by
and
Sb := [b, S]f (x) = b(x)Sf (x) − S(bf )(x)
Sα,b := [b, Sα]f (x) = b(x)Sαf (x) − Sα(bf )(x).
Our main results of this section are
Theorem 3.3. Let p, k, w, K and eS be the same as Theorem 1.2. If b ∈ BMO(Rn),
then for any real polynomial P (x, y), there exists constant C > 0 such that
kSbf (x)kMp,k(w) ≤ Ckf kMp,k(w),
where its norm depends only on the total degree of P , but not on the coefficients of P .
Theorem 3.4. Let p, q, k, Kα and w be the same as in Theorem 1.3. Then for b ∈
BMO(Rn), there exists constant C > 0 such that
kSα,bf (x)kM
q,
qk
p
(wq) ≤ Ckf kMp,k(wp,wq).
The following results will play an important role in our analysis.
Lemma 3.5. [19] Suppose K is a SCZK, w ∈ Ap(1 < p < ∞) and the operator eS is
of type (L2(Rn), L2(Rn)). Then for any b ∈ BMO(Rn), there exists constants C > 0
independent on the coefficients of P such that
kSbf kLp(w) ≤ Ckf kLp(w).
10
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
Lemma 3.6. [7] Let 1 < p < n
q , 0 < α < n, 1
for b ∈ BMO(Rn), there exists constant C > 0 such that
α, 0 < k < p
q = 1
p − α
n and w ∈ A(p,q). Then
kIα,bf (x)kM
q,
qk
p
(wq) ≤ Ckf kMp,k(wp,wq).
Lemma 3.6 is the weighted version of Theorem 1 in [6].
The following Proposition is essential to the proof of Theorem 3.3.
Proposition 3.7. Let B = B(x0, r), 0 < k < 1 and 1 < p < ∞. Then the inequality
(cid:18)Zx0−y>2r
f (y)
x0 − yn bB,w − b(y)dy(cid:19)p
w(B)1−k ≤ Ckf kp
Mp,k(w)kbkp
BM O(Rn).
(3.6)
holds for every y ∈ (2B)c, where (2B)c = Rn/2B.
Proof. Using Holder's inequality to the left-hand-side of (3.6), we have
w(B)1−k
f (y)
x0 − yn bB,w − b(y)dy(cid:19)p
w(B)1−k
1
(cid:18)Zx0−y>2r
≤ ∞Xj=1Z2jr<x0−y<2j+1r
≤ ∞Xj=1
2jBZ2j+1B
≤ C" ∞Xj=1
2jB(cid:18)Z2j+1B
Mp,k(w)" ∞Xj=1
≤ Ckf kp
1
w(B)1−k
f (y)
x0 − yn bB,w − b(y)dy!p
f (y)bB,w − b(y)dy!p
f (y)pw(y)dy(cid:19) 1
(cid:18)Z2j+1B
p(cid:18)Z2j+1B
2jB
w(2j+1B)
k
p
p′#p
dy(cid:19) 1
w(B)1−k
bB,w − b(y)p′
w(y)1−p′
w(y)1−p′
p′#p
dy(cid:19) 1
bB,w − b(y)p′
w(B)1−k.
For the simplicity of analysis, we denote A as
(cid:18)Z2j+1B
bB,w − b(y)p′
w(y)1−p′
p′
dy(cid:19) 1
.
By an elementary estimate, we have
(b2j+1B,w1−p′ − b(y) + b2j+1B,w1−p′ − bB,w)p′
w(y)1−p′
p′
dy(cid:19) 1
A ≤(cid:18)Z2j+1B
≤(cid:13)(cid:13)(cid:13)(cid:13)
≤(cid:18)Z2j+1B
=: J + JJ.
b2j+1B,w1−p′ − b(·) + b2j+1B,w1−p′ − bB,w
w(·)
b2j+1B,w1−p′ − b(y)w(y)1−p′
(cid:13)(cid:13)(cid:13)(cid:13)Lp′ (w)
p′
dy(cid:19) 1
+ b2j+1B,w1−p′ − bB,ww1−p′
(2j+1B)
1
p′
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
11
For the term J, Lemma 3.2 implies
J ≤ CkbkBM O(w1−p′ )w1−p′
(2j+1B)
1
p′ ≤ Cw1−p′
(2j+1B)
1
p′ .
(3.7)
To deal with JJ, by (3.2), we have
b2j+1B,w1−p′ − bB,w ≤ b2j+1B,w1−p′ − b2j+1B + b2j+1B − bB + bB − bB,w
b(y) − b2j+1Bw(y)1−p′
dy + 2n(j + 1)kbkBM O(Rn)
≤
1
w1−p′(2j+1B)Z2j+1B
w(B)ZB
+
1
:= JJ1 + JJ2 + JJ3.
b(y) − bBw(y)dy
Combining (2.5) with (3.1),
JJ3 =
w({x ∈ B : b(y) − bB > α})dα
1
w(B)Z ∞
≤ CZ ∞
0
0
≤ C.
e−C2αδ/kbkBM O(Rn )dα
In the same manner we can see that
JJ1 ≤ C.
It follows immediately that
JJ ≤ C(2n(j + 1) + 2)w1−p′
(2j+1B)
1
p′ .
(3.8)
As a by-product of (3.7) and (3.8), we have
A ≤ C(j + 1)w1−p′
(2j+1B)
1
p′ .
Then, applying (2.2), the proof of (3.6) based on the following observation
" ∞Xj=1
≤ C" ∞Xj=1
w(2j+1B)
k
p
2jB
(cid:18)Z2j+1B
p #p
1−k
1−k
p (j+1)
w(B)
w(2j+1B)
b(y) − bB,wp′
w(y)1−p′
p′#p
dy(cid:19) 1
w(B)1−k
= C.
(cid:3)
Proof of Theorem 3.3. The task is now to find a constant C such that for fixed ball
B = B(x0, r), we can obtain
1
w(B)kZB
Sbf (x)p w(x)dx ≤ Ckf kp
Mp,k(w).
(3.9)
12
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
We decompose f = f χ2B + fχ(2B)c := f1 + f2, and consider the corresponding splitting
ZB
Sbf (x)p w(x)dx ≤ C(cid:18)ZB
Sbf1(x)pw(x)dx +ZB
=: K + KK.
Sbf2(x)pw(x)dx(cid:19)
An application of Lemma 3.5 and w ∈ Ap yields
K ≤ CZ2B
f (x)pw(x)dx ≤ Ckf kp
Mp,k(w)w(B)k.
(3.10)
To estimate the other term KK, we note that for x ∈ B and y ∈ (2B)c, x0 − y <
Cx − y. Then a further use of (1.8) derives that
eip(x,y)K(x, y)f2(y)(b(x) − b(y))dy(cid:12)(cid:12)(cid:12)(cid:12)
f2(y)b(x) − b(y)
p
dy(cid:19)p
Sbf2(x)p =(cid:12)(cid:12)(cid:12)(cid:12)ZRn
≤ C(cid:18)ZRn
≤ C(cid:18)Zx0−y>2r
x − yn
f (y)
x0 − yn {b(x) − bB,w + bB,w − b(y)}dy(cid:19)p
.
where bB,w = 1
w(B)RB b(x)w(x)dx. Then, we have
x0 − yn dy(cid:19)pZB
KK ≤ C(cid:18)Zx0−y>2r
f (y)
b(x) − bB,wpw(x)dx
+ C(cid:18)Zx0−y>2r
f (y)
x0 − yn b(y) − bB,wdy(cid:19)p
w(B)
:= KK1 + KK2.
A further use of proposition 3.7, we get
KK2 ≤ Ckf kp
Mp,k(w)w(B)k.
To get the desired estimate, we are led to estimate the term KK1. This estimate will
be done via (2.1), (2.3) and Lemma 3.2.
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
13
KK1 = ∞Xj=1Z2jr<x0−y<2j+1r
2jBZ2j+1B
2jB(cid:18)
≤ ∞Xj=1
∞Xj=1
≤ C
1
1
1
b(x) − bB,wpw(x)dx
f (y)
x0 − yn dy!pZB
f (y)dy!pZB
w(2j+1B)kZ2j+1B
p(cid:18)Z2j+1B
2jB (cid:18) 1
2j+1B− 1
w(y)− 1
k
p
b(x) − bB,wpw(x)dx
p
f (y)pw(y)dy(cid:19) 1
p−1 dy(cid:19) p−1
p ZB
w(y)dy(cid:19)− 1
2j+1BZ2j+1B
p
× w(2j+1B)
≤ Ckf kMp,k(w) ∞Xj=1
×ZB
b(x) − bB,wpw(x)dx
w(2j+1B)
k
p!p
b(x) − bB,wpw(x)dx
1−k
p
∞Xj=1 w(B)
w(2j+1B)
p !p
1−k
w(B)k
≤ Ckf kp
Mp,k(w)kbkp
BM O(Rn)
≤ Ckf kp
Mp,k(w)w(B)k
Hence
KK ≤ Ckf kp
Lp,k(w)w(B)k.
(3.11)
Combing (3.10), (3.11), we obtain (3.9), which is the desired conclusion.
Proof of Theorem 3.4. As in the proof of Theorem 1.3. Theorem 3.4 can be deduced
via Lemma 3.6 and the following observation
Sbf (x) ≤Z
f (y)
x − yn−β dy = Iα(f )(x).
References
[1] Adams D.R., A note on Riesz potentials, Duke Math. J. 42(1975), 765 -- 778.
[2] Chiarenza F. and Frasca M., Morrey spaces and Hardy-Littlewood maximal function, Rend.
Math. Appl. 7(7)(1987), 273 -- 279.
[3] Fan D.S., Lu S.Z. and Yang D.C., Regularity in Morry spaces of strong solutions to nondiver-
gence elliptic equations with VMO coefficients, Georgian Math. J. 5(5)(1998), 425 -- 440.
[4] Grafakos L., Classical and Modern Fourier Analysis, Pearson Education, Inc. Upper Saddle River,
New Jersey, 2004.
[5] Grafakos L. and Torres R., Multilinear Calder´on-Zygmund theory, Adv. Math, 165(2002),
124 -- 164.
[6] Komori Y. and Mizuhara T., Notes on commutators and Morrey spaces. Hokkaido Math. J.
32(2003), 345 -- 353.
14
ZUNWEI FU, SHAOGUANG SHI, AND SHANZHEN LU
[7] Komori Y. and Shirai S., Weighted Morrey spaces and a singular integral operator, Math. Nachr.
282(2009), 219 -- 231.
[8] Lu S.Z., Four Lectures on Real H p spaces, World Scientific Publishing, Singapore, 1995.
[9] Lu S.Z., Multilinar oscillatory integrals with Calder´on-Zygmund kernel, Sci. China(Ser.A) 42(1999),
1039 -- 1046.
[10] Lu S.Z., A class of oscillatory integrals, Int. J. Appl. Math. Sci. 2(1)(2005), 42 -- 58.
[11] Lu S.Z. and Zhang Y., Weighted norm inequality of a class of oscillatory singular operators,
Chin. Sci. Bull. 37(1992), 9 -- 13.
[12] Lu S.Z.and Xia X., Boundedness of commutators of an oscillatory singular operators, Studia
Math. 186(2008), 15 -- 27.
[13] Phong D.H. and Stein E.M., Singular integrals related to the Radon transform and boundary
value problems, Proc. Nat. Acad. USA. 80(1983), 7697 -- 7701.
[14] Morrey C.B., On the solutions of quasi-linear elliptic partial differential equations, Trans. Amer.
Math. Soc. 43(1938), 126 -- 166.
[15] Muckenhoupt B., Weighted norm inqualities for the Hardy maximal function, Trans. Amer. Math.
Soc. 165(1972), 207 -- 226.
[16] Muckenhoupt B. and Wheeden R., Weighted bounded mean oscillation and the Hilbert trans-
form, Studia Math. 54(1976), 221 -- 237.
[17] Ricci F. and Stein E.M., Harmonic analysis on Nilpotant groups and singular integrals I: Oscil-
latory Integrals, J. Funct. Anal. 73(1987), 179 -- 194.
[18] Sato S., Weighted weak type (1,1) estimates for oscillatory integrals, Studia Math. 47(2001), 1 -- 17.
[19] Shi S.G., Weighted boundedness for the commutators of one class of oscillatory integral operators,
J. Beijing Normal Univ. (Nat. Sci.) 47(2011), 344 -- 346.
[20] Shi S.G., Fu Z.W. and Lu S.Z., Weighted norm inequalities for fractional oscillatory integrals
and their commutators, arXiv:1111.3418v1.
[21] Shi Y.L., Tao X.X., Multilinear Riesz potential operators on Herz-type spaces and generalized
Morrey spaces, Hokkaido Math. J. 38 (2009), 635 -- 662.
[22] Stein E.M. and Weiss G., Interpolation of operators with change of measures, Trans. Amer.
Math. Soc. 87(1958), 159 -- 172.
[23] Tao X.X., Yu X., Zhang H.H., Multilinear Calder´on-Zygmund operators on variable exponent
Morrey spaces over domains, Appl. Math. J. Chinese Univ. 26(2)(2011), 187 -- 197.
[24] Torchinsky A., Real variable Methods in Harmonic Analysis Academic Press, San DIego, 1986.
School of Sciences
Linyi University
Linyi 276005
P. R. China
E-mail address: [email protected]
School of Mathematical Sciences
Beijing Normal University
Beijing 100875
and
School of Sciences
Linyi University
BOUNDEDNESS OF OSCILLATORY INTEGRAL OPERATORS AND THEIR COMMUTATORS
15
Linyi 276005
P. R. China
E-mail address: [email protected]
School of Mathematical Sciences
Beijing Normal University
Beijing 100875
P. R. China
E-mail address: [email protected]
|
1905.05971 | 1 | 1905 | 2019-05-15T06:39:36 | Orlicz Modules over Coset Spaces of Compact Subgroups in Locally compact Groups | [
"math.FA"
] | Let $H$ be a compact subgroup of a locally compact group $G$ and let $m$ be the normalized $G$-invariant measure on homogeneous space $G/H$ associated with Weil's formula. Let $\varphi$ be a Young function satisfying $\Delta_2$-condition. We introduce the notion of left module action of $L^1(G/H, m)$ on the Orlicz spaces $L^\varphi(G/H, m).$ We also introduce a Banach left $L^1(G/H, m)$-submodule of $L^\varphi(G/H, m).$ | math.FA | math | ORLICZ MODULES OVER COSET SPACES OF COMPACT SUBGROUPS
IN LOCALLY COMPACT GROUPS
VISHVESH KUMAR
Abstract. Let H be a compact subgroup of a locally compact group G and let m be the
normalized G-invariant measure on homogeneous space G/H associated with Weil's formula.
Let ϕ be a Young function satisfying ∆2-condition. We introduce the notion of left module action
of L1(G/H, m) on the Orlicz spaces Lϕ(G/H, m). We also introduce a Banach left L1(G/H, m)-
submodule of Lϕ(G/H, m).
9
1
0
2
y
a
M
5
1
]
.
A
F
h
t
a
m
[
1
v
1
7
9
5
0
.
5
0
9
1
:
v
i
X
r
a
1. Introduction
The abstract theory of Banach modules or Banach algebras plays an important role in var-
ious branches of Mathematics, for instance, abstract harmonic analysis, representation theory,
operator theory; see [4, 10, 11, 8] and the references therein. In particular, convolution structure
on the Orlicz spaces to be an Banach algebra or a Banach module over a locally compact group
or hypergroup were studied by many researchers [1, 9, 21, 13, 14, 15].
In [13], the author defined and studied the notion of abstract Banach convolution algebra
on Orlicz spaces over homogeneous spaces of compact groups. Recently, Ghaani Farashahi [8]
introduced the notion of abstract Banach convolution function module on the Lp-space on coset
spaces of compact subgroups in locally compact groups. The purpose of this article is to define
and study a new class of abstract Banach module on Orlicz spaces over coset spaces of compact
subgroups in locally compact groups. Let us remark that Orlicz spaces are genuine generalization
of Lebesgue spaces. It is worth mentioning that an appropriate use of Jensen's inequality [19,
pg. 62] plays a key role in this article.
In the next section, we present some basics of Orlicz spaces and some classical harmonic
analysis on a homogeneous space (the space of left cosets) of a locally compact group. Section 3
is devoted to the study of abstract convolution module structure on the Orlicz space Lϕ(G/H, m),
where H is a compact subgroup of a locally compact group G, m is the normalized G-invariant
measure on the homogeneous space G/H which satisfies Weil's formula and ϕ is a Young function
2010 Mathematics Subject Classification. Primary 43A85, 46E30; Secondary 43A15, 43A20.
Key words and phrases. Homogeneous space, Convolution function modules, Orlicz spaces.
1
2
VISHVESH KUMAR
satisfying ∆2-condition. In this section, we prove that Lϕ(G/H, m) is a Banach left L1(G/H, m)-
module with respect to a generalized convolution. We also introduce a Banach left L1(G/H, m)-
submodule of Lϕ(G/H).
2. Preliminaries
A non-zero convex function ϕ : R → [0, ∞] is called a Young function if it is even, left
continuous with ϕ(0) = 0 and lim
x→∞
ϕ(x) = ∞. Here we note that every Young function is an
integral of a non-decreasing left continuous function [19, Theorem 1].
Let Ω be a locally compact Hausdorff space and m be a positive Radon measure on Ω. Denote
the space of all equivalence classes of m-measurable functions on Ω by L0(Ω). A Young function
ϕ satisfies ∆2-condition if there exist a constant C > 0 and x0 > 0 such that ϕ(2x) ≤ Cϕ(x)
for all x ≥ x0 if m(Ω) < ∞ and ϕ(2x) ≤ Cϕ(x) for all x ≥ 0 otherwise. Given a Young function
ϕ, the modular function ρϕ : L0(Ω) → R is defined by ρϕ(f ) :=RΩ ϕ(f ) dm. We always assume
that Young function ϕ satisfies ∆2-condition. For a given Young function ϕ, the Orlicz space
Lϕ(Ω, m), in short Lϕ(Ω), is defined by
Lϕ(Ω) :=(cid:8)f ∈ L0(Ω) : ρϕ(af ) < ∞ for some a > 0(cid:9) .
Then the Orlicz space is a Banach space with respect to the norm k · k0
ϕ on Lϕ(Ω) called
Luxemburg norm or gauge norm which is defined by
kf k0
ϕ := inf(cid:26)k > 0 :ZG
ϕ(cid:18) f
k (cid:19) dm ≤ 1(cid:27) .
If ϕ(x) = xp 1 ≤ p < ∞ then Lϕ(Ω) is usual Lp-spaces, 1 ≤ p < ∞. An example of Young
function which satisfies ∆2-condition and gives an Orlicz space other than Lp-spaces is given by
ϕ(x) = (e + x) log(e + x) − e.
We denote the space of all continuous functions on Ω with compact support by Cc(Ω). It is
well known that if ϕ satisfies ∆2-condition then Cc(Ω) is a dense subspace of Lϕ(Ω). If A is a
measurable subset of Ω such that 0 < m(A) < ∞, then we have the Jensen's Inequality
ϕ(cid:18)RA f dm
m(A) (cid:19) ≤ RA ϕ(f ) dm
m(A)
.
We make use of the above inequality several times in this article. In this paper, we also employ
the notations of the author in [3, 8].
For a locally compact group G with the Haar measure dx and f, g ∈ Lϕ(G), define convolution
∗G on Lϕ(G, m) by
f ∗G g(x) =ZG
f (y) g(y−1x) dx (x ∈ G).
ORLICZ MODULES OVER HOMOGENEOUS SPACES
3
It is well-known that (Lϕ(G), k · k0
ϕ) is a Banach algebra with respect to the convolution product
∗G (see [12]) that is,
kf ∗G gk0
Φ ≤ kf k0
ϕ kgk0
ϕ
for all f, g ∈ Lϕ(G). Also, if f ∈ L1(G) and f ∈ Lϕ(G) then the above convolution define a
module action of L1(G) on Lϕ(G) which makes Lϕ(K) a Banach left L1(K)-module, that is,
(1)
kf ∗G gk0
ϕ ≤ kf k1 kgk0
ϕ
for all f ∈ L1(G) and g ∈ Lϕ(G) (see [14, 18]).
Let H be a compact subgroup of a locally compact group G with the normalized Haar measure
dh. The left coset space G/H can be seen as a homogeneous space with respect to the action
of G on G/H given by left multiplication. The canonical surjection q : G → G/H is given by
q(x) = xH. Define TH(f )(xH) =RH f (xh) dh, then
Cc(G/H) = {TH (f ) : f ∈ Cc(G)}
The homogeneous space G/H has a unique normalized G-invariant positive Radon measure
m that satisfies the Weil's formula
(2)
ZG/H
TH (f )(xH) dm(xH) =ZG
f (x) dx,
and hence kTH(f )kL1(G/H,m) ≤ kf kL1(G) for all f ∈ L1(G). For more details on harmonic
analysis on homogeneous spaces of locally compact groups see [3, 4, 5, 6, 7, 22, 8].
3. Orlicz modules over Coset Spaces of compact subgroups of locally compact
groups
Throughout this section, we assume that the Young function ϕ satisfies ∆2- condition, G
is a locally compact group with a Haar measure dx and H is a compact subgroup of G with
the normalized Haar measure dh. It is also assumed that the homogeneous space G/H has the
normalized G-invariant measure m satisfying the Weil's formula. In this section, we show that
the space Lϕ(G/H) becomes a Banach left L1(G/H, m)-module with respect to the convolution
on Cc(G/K) defined in [8]. We also define a subspace of Lϕ(G/H) and show that this subspace
is a Banach left L1(G/H, m)-submodule of Lϕ(G/H, m).
We begin this section with following result.
Theorem 3.1. Let G be a locally compact group and let H be a compact subgroup of G. Let
m be the normalized G-invariant measure on the coset space G/H. Then the linear map TH :
4
VISHVESH KUMAR
Cc(G) → Cc(G/H) satisfies
(3)
kTH(f )k0
Lϕ(G/H,m) ≤ kf k0
Lϕ(G)
for all f ∈ Cc(G). Further, if the Young function ϕ satisfies ∆2-condition. TH can be uniquely
extended to a linear map from Lϕ(G) onto Lϕ(G/H, m).
Proof. For f ∈ Cc(G) and k > 0, by using Weil's formula and Jensen's inequality, we have
ρϕ(cid:18) TH (f )
k
k
k
f (xh)
f (xh)
ϕ(cid:18) TH (f )(xH)
(cid:19) dm(xH)
k (cid:19) = ZG/H
ϕ(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)
dh(cid:12)(cid:12)(cid:12)(cid:12)
(cid:19) dm(xH)
ZH
= ZG/H
ϕ(cid:18)ZH
dh(cid:19) dm(xH)
≤ ZG/H
ϕ(cid:18) f (xh)
k (cid:19) dh dm(xH)
≤ ZG/HZH
k (cid:19) (xh) dh(cid:19) dm(xH)
ϕ(cid:18) f
= ZG/H(cid:18)ZH
TH(cid:18)ϕ(cid:18) f
k (cid:19)(cid:19) (xH) dm(xH)
= ZG/H
ϕ(cid:18) f
k (cid:19) dx = ρϕ(cid:18) f
k(cid:19) .
= ZG
Now,
kf k0
Lϕ(G) = inf{k > 0 : ρϕ(cid:18) f
k(cid:19) ≤ 1}
k (cid:19) ≤ 1} = kTH (f )k0
≥ inf{k > 0 : ρϕ(cid:18) TH(f )
Lϕ(G/H,m).
Therefore, we get kTH (f )k0
Lϕ(G/H,m) ≤ kf k0
Lϕ(G) for all f ∈ Cc(G).
Since φ is ∆2-regular, Cc(G) and Cc(G/H) are dense in Lϕ(G) and Lϕ(G/H, m) respectively.
Therefore, we can extend TH to a bounded linear map from Lϕ(G) onto Lϕ(G/H, m). We denote
this extension of TH again by TH and it satisfies (3).
(cid:3)
Proposition 3.2. Let G be a locally compact group and let H be a compact subgroup of G.
Let m be the normalized G-invariant measure on the coset space G/H. Suppose that ϕ satisfies
∆2-condition. If f ∈ Lϕ(G/H, m) and fq := f ◦ q then we have fq ∈ Lϕ(G) with
(4)
kfqk0
Lϕ(G) = kf k0
Lϕ(G/H,m).
ORLICZ MODULES OVER HOMOGENEOUS SPACES
5
Proof. For f ∈ Lϕ(G/H, m) and k > 0, by Weil's formula and the fact that H is compact, we
have
ρϕ(cid:18) fq
ϕ(cid:18) fq(x)
k (cid:19) dx
TH(cid:18)ϕ(cid:18) fq
k (cid:19) = ZG
= ZG/H
= ZG/H(cid:18)ZH
= ZG/H(cid:18)ZH
= ZG/H(cid:18)ZH
= ZG/H
k (cid:19)(cid:19) (xH) dm(xH)
(cid:19) dh(cid:19) dm(xH)
(cid:19) dh(cid:19) dm(xH)
(cid:19) dh(cid:19) dm(xH)
(cid:19) dm(xH)!(cid:18)ZH
k(cid:17) and consequently, we get kfqk0
ϕ(cid:18) fq(xh)
ϕ(cid:18) f (xhH)
ϕ(cid:18) f (xH)
ϕ(cid:18) f (xH)
k
k
k
k
dh(cid:19) = ρϕ(cid:18) f
k(cid:19) .
Lϕ(G) = kf k0
Lϕ(G/H,m).
(cid:3)
Therefore ρϕ(cid:16) fq
k (cid:17) = ρϕ(cid:16) f
Remark 1. Note that TH (fq) = f and therefore, it is clear from above corollary that for fq ∈
Lϕ(G), the equality in Theorem 3.1 holds.
Here we set certain terminologies for further use. For any continuous function f define the
left translation by Lhf (·) = f (h−1(·)) and the right translation by Rhf = f ((·)h). Let G be a
locally compact group and let H be a compact subgroup of G. We set
Cc(G : H) = {f ∈ Cc(G) : Rhf = f ∀ h ∈ H},
A(G : H) = {f ∈ Cc(G) : Lhf = f ∀ h ∈ H}
and A(G/H) = {g ∈ Cc(G/H) : Lhg = g ∀ h ∈ H}.
For a Young function ϕ ∈ ∆2, define
Aϕ(G : H) = {f ∈ Lϕ(G) : Lhf = f ∀ h ∈ H},
and also
Aϕ(G/H, m) = {g ∈ Lϕ(G/H, m) : Lhg = g ∀ h ∈ H}.
Note that Aϕ(G/H, m) is the topological closure of A(G/H) in Lϕ(G/H, m) and therefore
it is a closed subspace of Lϕ(G/H, m). Similarly Aϕ(G : H) is closed subspace of Lϕ(G). It is
known that TH maps Cc(G : H) and A(G : H) onto Cc(G/H) and A(G/H) respectively (see [8,
Proposition 4.2]).
6
VISHVESH KUMAR
Since A(G : H) is dense in Lϕ(G : H) and A(G/H) is dense in Aϕ(G/H), the next lemma
follows from the continuity of TH.
Lemma 3.3. Let G be a locally compact group and let H be a compact subgroup of G. Let
m be the normalized G-invariant measure on the coset space G/H. Suppose that ϕ satisfies
∆2-condition. Then the mapping TH maps Lϕ(G : H) onto Aϕ(G/H, m).
For g ∈ Cc(G/H), note that the mapping xH 7→ RH g(hxH) dh is in Cc(G/H). Define J :
Cc(G/H) → Cc(G/H) as
It is clear that J is a linear operator. In addition, the following theorem says that the norm of
Jg(xH) =ZH
g(hxH) dh.
J is bounded operator with the norm bounded by one.
Theorem 3.4. For f ∈ Cc(G/H), we have
kJf kLϕ(G/H,m) ≤ kf kLϕ(G/H,m).
Proof. For f ∈ Cc(G/H) we have,
ρϕ(cid:18)Jf
1
k
ϕ(cid:18) Jf (xH)
k (cid:19) = ZG/H
(cid:19) dm(xH)
ϕ(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)
f (hxH) dh(cid:12)(cid:12)(cid:12)(cid:12)
= ZG/H
k ZH
ϕ(cid:18) f (hxH)
≤ ZG/HZH
= ZH ZG/H
ϕ(cid:18) f (hxH)
= ZH ZG/H
ϕ(cid:18) f (xH)
= ZH ZG/H
ϕ(cid:18) f
(cid:19) dm(xH)
(cid:19) dh dm(xH)
(cid:19) dm(xH)! dh
(cid:19) dm(xH)! dh
k
k
k
k (cid:19) (xH) dm(xH)! dh = ρϕ(cid:18) f
k(cid:19) .
Consequently, we get kJf kLϕ(G/H,m) ≤ kf kLϕ(G/H,m) for all f ∈ Cc(G/H).
(cid:3)
It is shown in [8, Theorem 4.5 (2)] that J maps Cc(G/H) onto A(G/H). Now, the following
corollary is immediate.
Corollary 3.5. Let G be a locally compact group and let H be a compact subgroup of G. Let
m be the normalized G-invariant measure on the coset space G/H. Suppose that ϕ satisfies
ORLICZ MODULES OVER HOMOGENEOUS SPACES
7
∆2-condition. The bounded linear map J : Cc(G/H) → A(G/H) can be uniquely extended to a
bounded linear map Jϕ : Lϕ(G/H, m) → Aϕ(G/H, m) which satisfies
kJϕf kLϕ(G/H,m) ≤ kf kLϕ(G/H,m).
Further, the linear operator Jϕ : Lϕ(G/H, m) → Aϕ(G/H, m) is an onto map.
Proof. The extension of the map J from Cc(G/H) to Lϕ(G/H) follows from Theorem 3.4 and
the density of Cc(G/H, m) and A(G/H) in Lϕ(G/H, m) and Aϕ(G/H, m), respectively. Further,
for any f ∈ Lϕ(G/H) and z ∈ H, we have
Lz(Jf )(xH) = Jf (z−1xH) =ZH
f (hz−1xH) dh =ZH
f (hxH) dh = Jf (xH).
This shows that Jf ∈ Aϕ(G/H, m). Now we prove second part of the corollary. For any f ∈
Aϕ(G/H), we get,
Jf (xH) =ZH
f (hxH) dh =ZH
f (xH) dh = f (xH),
for all x ∈ G. Therefore, Jf = f. Hence Jϕ : Lϕ(G/H, m) → Aϕ(G/H, m) is a onto map.
(cid:3)
Remark 2. It can seen in the proof of Corollary 3.5 above that JAϕ(G/H,m) = IAϕ(G/H,m).
Now, we are ready to define the convolution product '∗G/H ' on Cc(G/H) same as given in
[8] as follows: let G be a compact group, H a closed subgroup and let m be the normalized
G-invariant measure on G/H. For f, g ∈ Cc(G/H), the convolution f ∗G/H g : G/H → C is given
by
(5)
f ∗G/H g(xH) =ZG/H
f (yH)Jg(y−1xH) dm(yH),
for all xH ∈ G/H. The convolution product '∗G/H ' has the following properties similar to the
usual convolution in Cc(G) (see [8, Proposition 4.10]).
(i) For any f, g ∈ Cc(G/H), (f, g) 7→ f ∗G/H g is a bilinear map from Cc(G/H) × Cc(G/H)
to Cc(G/H) and (Cc(G/H), ∗G/H ) is an algebra.
(ii) f ∗G/H g = TH(fq ∗G gq) and (f ∗G/H g)q = fq ∗G gq, where '∗G' is the usual convolution
in C(G).
(iii) Lx(f ∗G/H g) = (Lxf ) ∗G/H g.
The following result says that Cc(G/H) is a normed algebra with respect to the norm k ·
k0
Lϕ(G/H,m).
Lemma 3.6. If f, g ∈ Cc(G/H), then
(6)
kf ∗G/H gk0
Lϕ(G/H,m) ≤ kf kL1(G/H,m)kgk0
Lϕ(G/H,m).
8
VISHVESH KUMAR
Proof. Let f, g ∈ Cc(G/H). Using Proposition 3.2 we get
kf ∗G/H gk0
Lϕ(G/H,m) = k(f ∗G/H g)qk0
Lϕ(G) = kfq ∗G gqk0
Lϕ(G)
Since Lϕ(G) is a L1(G)-module then by Proposition 3.2 we get
kf ∗G/H gk0
Lϕ(G/H,m) ≤ kfqkL1(G)kgqk0
Lϕ(G) = kf kL1(G/H,m)kgk0
Lϕ(G/H,m).
(cid:3)
Theorem 3.7. Let G be a locally compact group and let H be a compact subgroup of G. Let
m be the normalized G-invariant measure on the coset space G/H. Suppose that ϕ satisfies
∆2-condition. Then the convolution ∗G/H : Cc(G/H) × Cc(G/H) → Cc(G/H) given by (5)
can be extended to a convolution ∗ϕ
G/H : L1(G/H, m) × Lϕ(G/H, m) → Lϕ(G/H, m) such that
Lϕ(G/H, m) is a Banach left L1(G/H, m)-module with respect to this extended convolution.
Proof. Let f ∈ L1(G/H, m) and g ∈ Lϕ(G/H, m). Since Cc(G/H) is dense in L1(G/H, m) and
in Lϕ(G/H, m) as ϕ ∈ ∆2, there exist {fn} and {gn} in Cc(G/H) such that fn → f and gn → g
as n → ∞. Now, define
f ∗ϕ
G/H g = lim
n→∞
fn ∗G/H gn.
Note that ∗ϕ
G/H : L1(G/H, m) × Lϕ(G/H, m) → Lϕ(G/H, m) is well-defined. By Lemma 3.6 we
have
kf ∗ϕ
G/H gk0
Lϕ(G/H,m) ≤ kf kL1(G/H,m)kgk0
Lϕ(G/H,m).
Thus Lϕ(G/H, m) is a Banach left L1(G/H, m)-module with respect to the extended convolu-
tion.
(cid:3)
The above theorem claims the existence of convolution product ∗ϕ
G/H but it does not reveal
any explicit formula for the convolution product. The following corollary fulfils this objective
whose proof is a consequence of the fact that Cc(G/H) is dense Lϕ(G/H, m).
Corollary 3.8. Let G be a locally compact group and let H be a compact subgroup of G. Let
m be the normalized G-invariant measure on the coset space G/H. Suppose that ϕ satisfies
∆2-condition. If f ∈ L1(G/H, m) and g ∈ Lϕ(G/H, m), then the convolution ∗ϕ
G/H is given by
(7)
for all xH ∈ G/H.
(f ∗ϕ
G/H g)(xH) =ZG/H
f (yH)Jϕg(y−1xH) dm(yH)
In the next corollary we present a Banach left L1(G/H, m)-submodule of Lϕ(G/H, m) whose
proof is a routine check.
ORLICZ MODULES OVER HOMOGENEOUS SPACES
9
Corollary 3.9. Let G be a locally compact group and let H be a compact subgroup of G. Let
m be the normalized G-invariant measure on the coset space G/H. Suppose that ϕ satisfies ∆2-
condition. Then the space Aϕ(G/H, m) is a Banach left L1(G/H, m)-submodule of Lϕ(G/H, m).
Acknowledgements. The author thanks Prof. V. Muruganandam for his support and encour-
agement.
References
[1] H. P. Aghababa, I. Akbarbaglu and S. Maghsoudi, The space of multipliers and convolution
of Orlicz spaces on a locally compact group, Studia Math. 291(1) (2013) 19-34.
[2] I. Akabarbaglu and S. Maghsoudi, Banach-Orlicz algebras on a locally compact group,
Mediterr. J. Math., 10 (2013) 1937-1947.
[3] A. Ghaani Farashahi, Abstract convolution function algebra over homogeneous spaces of
compact groups, Illinois J. Math., 59(4) (2012) 1025-1042.
[4] A. Ghaani Farashahi, A class of Abstract linear representation for convolution function
algebra over homogeneous space of compact groups, Canad. J. Math., DOI : 10.4153/CJM-
2016-043-9.
[5] A. Ghaani Farashahi, Abstract Plancherel (trace) formulas over homogeneous spaces of
compact groups, Canad. Math. Bull., 60(2) (2017) 111-121.
[6] A. Ghaani Farashahi, Abstract operator-valued Fourier transforms over homogeneous
spaces of compact groups, Groups Geom. Dyn. 11(4) (2017) 1437-1467.
[7] A. Ghaani Farashahi, Abstract measure algebras over homogeneous spaces of compact
groups, International Journal of Mathematics, 29(1) (2018) 1850005, 34 pp.
[8] A. Ghani Farashahi, Abstract Banach convolution function modules over coset spaces on
compact subgroups in locally compact spaces, Bull. Braz. Math. Soc. (New Series) (2019).
DOI: https://doi.org/10.1007/s00574-018-00129-6
[9] H. G. Feichtinger, Some new subalgebras of L1(G), Indag. Math. 36 (1974) 44-47.
[10] H. G. Feichtinger, On a class of convolution algebra of functions, Ann. Inst. Fourier (Greno-
ble), 27(3) (1977) 135-162.
[11] H. G. Feichtinger, Banach convolution algebra of functions II, Monatch. Math. 87 (1979)
181-207.
[12] H. Hudzik, A. Kamiska and J. Musielak, On some Banach algebras given by a modular,
in: Alfred Haar Memorial Conference, Budapest, Colloquia Mathematica Societatis J´anos
Bolyai (North-Holland, Amsterdam) 49 (1987) 445-463.
[13] V. Kumar, R. Sarma and N. Shravan Kumar, Orlicz algebras on homogeneous spaces of
compact groups and their abstract linear representations, Mediterr. J. Math. 15(4) Art.
186 (2018) 13 pp.
10
VISHVESH KUMAR
[14] V. Kumar, R. Sarma and N. Shravan Kumar, Orlicz spaces on hypergroups, Publ. Math.
Debrecen, 94(1-2) (2019) 399-414.
[15] , V. Kumar and R. Sarma, The Hausdorff-Young inequality for Orlicz spaces on compact
hypergroups, (accepted in Colloq. Math.) (2019).
[16] M. A. Krasnosel'skii and Ja. B. Rutickii, Convex functions and Orlicz spaces, Noordhoff,
Groningen (1961).
[17] A. Osancliol and S. Oztop, Weighted Orlicz algebra on locally compact groups, J. Aust.
Math. Soc., 99 (2015) 399-414.
[18] M. M. Rao, Convolution of vector field-I, Math. Z., 74 (1980) 63-79.
[19] M.M. Rao and Z.D. Ren, Theory of Orlicz spaces, Dekker, New York (1991).
[20] M. M. Rao, Convolution of vector fields-II : random walk models, Nonlinear Analysis, 47
(2001) 3599-3615.
[21] M. M. Rao, Convolutions of vector fields-III: Amenability and spectral properties, Real and
stochastic analysis, Trends Math., Birkhuser Boston, Boston, MA, (2004) 375-401.
[22] H. Reiter and J. D. Stageman, Classical harmonic analysis and locally compact groups,
Clarendon Press, Oxford, (2000).
Vishvesh Kumar
School of Mathematical Sciences
National Institute of Science Education and Research Bhubaneshwar, HBNI
At/Po- Jatni, District- Khurda, Odisha- 7520250, India.
E-mail address: [email protected]
|
1608.02463 | 1 | 1608 | 2016-08-08T14:48:56 | Dirichlet forms for singular diffusion on graphs | [
"math.FA"
] | We describe operators driving the time evolution of singular diffusion on finite graphs whose vertices are allowed to carry masses. The operators are defined by the method of quadratic forms on suitable Hilbert spaces. The model also covers quantum graphs and discrete Laplace operators. | math.FA | math |
Dirichlet forms for singular diffusion on graphs
Christian Seifert and Jurgen Voigt
Abstract
We describe operators driving the time evolution of singular diffusion on
finite graphs whose vertices are allowed to carry masses. The operators
are defined by the method of quadratic forms on suitable Hilbert spaces.
The model also covers quantum graphs and discrete Laplace operators.
MSC 2010: 47D06, 60J60, 47E05, 35Q99, 05C99
Keywords: gap diffusion, quantum graph, Dirichlet form, C0-semigroup,
positive, submarkovian
Introduction
The present paper is a continuation and extension of [2]. We present suitable
boundary or glueing conditions on graphs (quantum graphs) with singular sec-
ond order differential operators on the edges. In particular, we describe those
boundary conditions leading to positive and submarkovian C0-semigroups.
The graph consists of finitely many bounded intervals, the edges, whose end
points are connected with the vertices of the graph. On each of the edges e a
finite Borel measure µe is given, determining where particles may be located.
The particles move according to "Brownian motion" but are slowed down or
accelerated by the "speed measure" µe. Further, each of the vertices v is provided
with a weight µv > 0, and particles may also be located at those vertices v
with µv > 0.
The motivations for the treatment in [2] were twofold. The first issue was
to treat singular diffusion, including gap diffusion, on the edges of the graph,
in the framework of Dirichlet forms. The second aim was to describe glueing
conditions on the vertices, in the spirit of [4], and investigate conditions under
which the associated self-adjoint operator gives rise to a positive or submarkovian
C0-semigroup.
In the present paper, the extension with respect to [2] consists in two issues.
On the one hand, the boundary conditions we describe are more general than
glueing conditions. By glueing conditions or "local boundary conditions", we
understand conditions where, for a given vertex, only the values of a function on
adjacent edges and on the vertex itself can interact. In our treatment in Sections 2
1
2
C. Seifert, J. Voigt
and 3, however, the graph structure does not intervene at all, and we only specify
later the case of local boundary conditions, in Section 5. On the other hand, we
include the general case of vertices with masses, whereas in [2; Section 4] only a
special case was treated. These results have been obtained in [7].
The ultimate objective of the treatment is to obtain a semi-bounded (below)
self-adjoint operator H on a Hilbert space HΓ over the graph Γ which can then
be used in the initial value problem for the diffusion equation or heat equation
u′ = −Hu,
(0.1)
thus governing the time evolution of a process, i.e., giving rise to a C0-semigroup
on HΓ. For this equation it is of interest to obtain H in such a way that the
associated C0-semigroup is positive or submarkovian. The self-adjointness of H
is also of interest for the initial value problem for the Schrodinger equation
u′ = −iHu.
The part of the operator H acting on an edge e is of the form (Hf )e = −∂µe ∂fe,
where ∂µe is the derivative with respect to µe; cf. Section 1. The domain of H
is restricted by conditions on the boundary values of the functions on the edges
and the values at the vertices.
The Hilbert space HΓ is given by
L2([ae, be], µe) ⊕ KV ,
HΓ =Me∈E
where E is the set of edges, the interval [ae, be] corresponds to the edge e, and V
is the set of vertices; cf. Section 2 for more details. The operator H is obtained by
the method of forms. Avoiding all technicalities (which will be given in Section 2),
the form τ giving rise to H is of the form
τ (f, g) =Xe∈EZ be
ae
f ′
e(x)g′
e(x) dx + (L tr f tr g),
with domain
D(τ ) =(cid:8)f ∈ . . . ; tr f ∈ X(cid:9).
Here, tr f denotes the boundary values of f on the edges and the values of f on
the vertices, X is a subspace of the set of possible boundary values and values
on the vertices, and L is a self-adjoint operator (matrix) on X. The boundary
conditions for functions in the domain of H are encoded in the space X as well as
in the operator L; cf. Theorem 3.1. Our treatment includes the case that some
of the edges or vertices may have weight zero.
For the discussion of positivity and the submarkovian property in connection
with equation (0.1) we use the Beurling-Deny criteria for τ . These yield the
Dirichlet forms on graphs
3
result that the subspace X should satisfy lattice properties and L should satisfy
positivity properties; cf. Theorem 4.1.
The investigations mentioned so far did not take into account the graph struc-
ture of Γ.
In the description of glueing conditions, allowing only interactions
between vertices and adjacent edges, the space X and the operator L decompose
into parts corresponding to single vertices; cf. Corollaries 5.1 and 5.2.
In Section 1 we recall some notation and facts from the one-dimensional case on
an interval. In Section 2 we define the form in the Hilbert space HΓ on the graph
which then defines the operator driving the evolution. We show that the defined
form τ constitutes a form that is bounded below and closed. Let us point out
that our definition of the form looks somewhat different from the one given in [2;
Section 3]. In fact, looking at the definition of τ in [2; Section 3], one realises that
there is some interpretation needed in order to understand D(τ ) as a subset of the
Hilbert space HΓ. This interpreation is made explicit in the present paper by the
use of the mapping ι introduced in Sections 1 and 2. In Section 3 we describe the
operator H associated with the form τ (Theorem 3.1). In Section 4 we indicate
conditions for the C0-semigroup (e−tH )t>0 to be positive and submarkovian. In
Section 5 we describe the case of local boundary conditions.
1 One-dimensional prerequisites
In order to define the classical Dirichlet form we have to recall some notation and
facts for a single interval [a, b] ⊆ R, where a, b ∈ R, a < b. Let µ be a finite Borel
measure on [a, b], a, b ∈ spt µ, µ({a, b}) = 0. Our function spaces will consist of
K-valued functions, where K ∈ {R, C}. We define
Cµ[a, b] :=(cid:8)f ∈ C([a, b]; f affine linear on the components of [a, b] \ spt µ(cid:9),
2 (a, b) ∩ Cµ[a, b].
2,µ(a, b) := W 1
W 1
For later use we recall the following inequalities. There exists a constant C > 0
such that
kf k∞ 6 C(cid:0)kf ′k2
L2(a,b) + kf k2
for all f ∈ W 1
2 (a, b) ∩ C[a, b], and for all r ∈ (0, b − a] one has
L2([a,b],µ)(cid:1)1/2
(1.1)
f (a) 6 r1/2kf ′kL2(a,a+r) + kf kL2([a,a+r],µ) µ([a, a + r])−1/2,
(1.2)
and correspondingly for b; cf. [2; Lemma 1.4 and Remark 3.2(b)].
Let κ : W 1
be shown that R(κ) = R(κ W 1
2 (a, b) ∩ C[a, b] → L2([a, b], µ) be defined by κf := f . Then it can
2,µ(a,b) is
[2; Lemma 1.2]), and that κ W 1
2,µ(a,b)) (cf.
injective (cf. [2; lower part of p. 639]). We define ι := (cid:0)κ W 1
2,µ(a,b)(cid:1)−1. Thus, ι is
4
C. Seifert, J. Voigt
an operator from L2([a, b], µ) to W 1
2,µ(a, b),
D(ι) =(cid:8)f ∈ L2([a, b], µ); there exists g ∈ W 1
2 (a, b) ∩ C[a, b]
such that g = f µ-a.e.(cid:9),
2,µ(a, b) such that g = f µ-a.e.
and ιf is the unique element g ∈ W 1
In order to describe the operator associated with the form defined in the fol-
lowing section we need some additional notions and facts concerning derivatives
with respect to µ.
If f ∈ L1,loc(a, b), g ∈ L1([a, b], µ) are such that f ′ = gµ (where f ′ = ∂f denotes
the distributional derivative of f ), then we call g distributional derivative of f
with respect to µ, and we write
∂µf := g.
Note that then necessarily f ′ = 0 on [a, b] \ spt µ, i.e., f is constant on each of
the components of [a, b] \ spt µ. It is easy to see that this definition is equivalent
to
f (x) = c +Z(a,x)
g(y) dµ(y) a.e.,
(1.3)
with some c ∈ K. Thus, the function f has representatives of bounded variation
and these have one-sided limits (not depending on the representative) at all points
of [a, b].
2 The form on the graph
Let Γ = (V, E, γ) be a finite directed graph. This means that V and E are finite
sets, V ∩ E = ∅, V is the set of vertices (or nodes) of Γ, E the set of edges, and
γ = (γ0, γ1) : E → V × V associates with each edge e a "starting vertex" γ0(e),
and an "end vertex" γ1(e).
We assume that each edge e ∈ E corresponds to an interval [ae, be] ⊆ R (where
ae, be ∈ R, ae < be), and we assume that µe is a finite Borel measure on [ae, be]
satisfying either µe = 0 or else ae, be ∈ spt µe, µe({ae, be}) = 0. We denote
E0 := {e ∈ E; µe = 0},
E1 := E \ E0.
We further assume that, for each v ∈ V , we are given a weight µv > 0, and we
define
V0 :=(cid:8)v ∈ V ; µv = 0(cid:9),
V1 := V \ V0.
2.1 Remark. The sets E1 and V1 encode the parts of the graph Γ, where a
particle driven by the diffusion can be localised. In the present section we describe
general glueing conditions which do not take into account the correspondence of
the edges to the vertices.
In the case E1 = E, V1 = ∅ and µe the Lebesgue
measure on [ae, be], the model will describe quantum graphs; cf. [3], [4], [5]. In
the case E1 = ∅ we obtain (weighted) discrete diffusion on the vertices; cf.[1].
Dirichlet forms on graphs
5
We are going to describe the self-adjoint operator driving the evolution in the
Hilbert space
where on
we use the scalar product
HΓ := HE ⊕ KV1 ,
HE := Me∈E1
L2([ae, be], µe)
((fe)e∈E1 (ge)e∈E1)HΓ := Xe∈E1
(fe ge)L2([ae,be],µe),
and on KV1 we use the scalar product
((fv)v∈V1 (gv)v∈V1)HΓ := Xv∈V1
fvgv µv
In the following, the mapping ι defined in Section 1 will be applied in the
situation of the edges e ∈ E1, and will then be denoted by ιe. We then define the
(for f =(cid:0)(fe)e∈E1, (fv)v∈V1(cid:1), g =(cid:0)(ge)e∈E1, (gv)v∈V1(cid:1) ∈ HΓ).
operator ι from HΓ to Qe∈E1
2,µe(ae, be) × KV1, by
W 1
D(ι) :=(cid:8)f ∈ HΓ; fe ∈ D(ιe) (e ∈ E1)(cid:9),
(e ∈ E1),
(ιf )e := ιefe
(ιf )v := fv
(v ∈ V1).
We define the trace mapping (or boundary value mapping) tr : Qe∈E1
1 := E1 × {0, 1}, by
1∪V1, where E′
KV1 → KE ′
C[ae, be]×
tr f (e, j) :=(cid:26) fe(ae)
fe(be)
if e ∈ E1, j = 0,
if e ∈ E1, j = 1,
tr f (v) := fv
(v ∈ V1).
The space KE ′
1∪V1 will be provided with the scalar product
(ξ η) := X(e,j)∈E ′
1
ξ(e, j)η(e, j) + Xv∈V1
ξ(v)η(v) µv.
For the definition of the form we assume that X is a subspace of KE ′
1∪V1 and
that L is a self-adjoint operator in X. Then we define the form τ by
D(τ ) :=(cid:8)f ∈ D(ι); tr(ιf ) ∈ X(cid:9),
τ (f, g) := Xe∈E1Z be
ae
(ιefe)′(x)(ιege)′(x) dx +(cid:0)L tr(ιf )(cid:12)(cid:12) tr(ιg)(cid:1).
6
C. Seifert, J. Voigt
2.2 Remark. The subspace X encodes boundary conditions for the elements of
D(τ ). One would expect boundary conditions to be in the form of some equation
for tr(ιf ). Of course, if P denotes the orthogonal projection from KE ′
1∪V1 onto
X ⊥, then D(τ ) =(cid:8)f ∈ D(ι); P tr(ιf ) = 0(cid:9).
Further boundary conditions for the elements of the associated operator H are
encoded in the operator L; we refer to the description of H in Theorem 3.1.
2.3 Lemma. The form τ defined above is symmetric. D(τ ) is dense if and only
if
prV1(X) = KV1 ,
(2.1)
where prV1 denotes the canonical projection prV1 : KE ′
Proof. The symmetry of τ is obvious.
1∪V1 → KV1.
Assume that D(τ ) is dense. The image of the dense set D(τ ) under the
orthogonal projection
pr2 : HΓ → KV1
is dense in KV1, and therefore is equal to KV1. From the definition of D(τ ) it
follows that pr2(D(τ )) is contained in prV1(X), and therefore prV1(X) = KV1.
Now assume that (2.1) holds. For v ∈ V1 let ξv ∈ X be such that ξv(v) = 1
and ξv(w) = 0 for all w ∈ V1 \ {v}. Let gv ∈ D(ι) be defined by tr(ιgv) = ξv,
and gv affine linear on the edges. The affine linear interpolation of the prescribed
boundary values evidently yields an element of gv ∈ D(τ ).
Let f ∈ HΓ, and define
f := f − Xv∈V1
fvgv.
Then fv = 0 for all v ∈ V1. Because C 1
the function f can be approximated by functions in
c (ae, be) is dense in L2([ae, be], µe) (e ∈ E1),
Therefore f can be approximated by functions in
c (ae, be) (e ∈ E1), fv = 0 (v ∈ V1)(cid:9).
Dc :=(cid:8)f ∈ D(τ ); fe ∈ C 1
Dc +Xv∈V1
fvgv ⊆ D(τ ).
2.4 Remarks. (a) For the special case that X = KE ′
1∪V1 and L = 0 we denote
the corresponding form by τN (the index N indicating Neumann boundary con-
ditions). The form τN decomposes as the sum of the Neumann forms on each of
the edges and the null form on KV1. Therefore the closedness of τN follows from
the closedness in the one-dimensional cases; cf. [2; Section 1 and Remark 3.2].
(b) Condition (2.1) did not occur in the previous treatment [2]. The reason
is that it is obviously satisfied if the vertices do not have masses, i.e. V1 = ∅.
Also, in the case of vertices with masses, but with local boundary conditions of
continuity (see Example 5.3), condition (2.1) is automatically satisfied.
Dirichlet forms on graphs
7
2.5 Theorem. The form τ defined above is bounded below and closed.
Proof. For f ∈ D(τ ) we obtain the estimate
τ (f ) = τN(f ) +(cid:0)L tr(ιf )(cid:12)(cid:12) tr(ιf )(cid:1) > τN(f ) − kLk tr(ιf )2
(with τN defined in Remark 2.4(a)). From inequality (1.2) we obtain that the
mapping f 7→ tr(ιf ) E ′
is infinitesimally form small with respect to τN. The
remaining part of the trace, f 7→ tr(ιf ) V1, is bounded. These observations
imply that τ is bounded below and the that the embedding DτN ∋ f 7→ ιf ∈
C[ae, be] × KV1 , k · k∞) is continuous. (Here, DτN denotes D(τN), provided
1
(Qe∈E1
with the form norm.)
In order to obtain that τ is closed it is sufficient to show that D(τ ) is a
closed subset of DτN. This, however, is immediate from the continuity of the
mapping DτN ∋ f 7→ tr(ιf ) ∈ KE ′
1∪V1 (and the fact that X is a closed subspace
of KE ′
1∪V1).
3 The operator H associated with the form τ
We assume that the notation and the hypotheses are as in the previous section,
and that (2.1) holds.
Besides the trace mapping defined in the previous section we also need the
signed trace (or signed boundary values)
BV (ae, be) → KE ′
1 ⊆ KE ′
1∪V1
str : Ye∈E1
(where BV (ae, be) denotes the set of functions of bounded variation, with equiv-
alence of functions coinciding a.e.), defined by
str f (e, j) :=(cid:26) fe(ae+)
−fe(be−)
if e ∈ E1, j = 0,
if e ∈ E1, j = 1.
The inclusion KE ′
to be able to use str f also as an element of KE ′
1 ⊆ KE ′
1∪V1.
1∪V1 is to be understood in the canonical sense; we want
For the description of the self-adjoint operator H associated with the form τ
we use a maximal operator H for the differential part of the form. With the
notation described in Section 1, we define
D( H) :=(cid:8)f ∈ Ye∈E1
Hf := (−∂µe(ιefe)′)e∈E1
D(ιe); (ιefe)′ ∈ L1(ae, be), ∂µe(ιefe)′ exists,
∂µe(ιefe)′ ∈ L2([ae, be], µe) (e ∈ E1)(cid:9),
(f ∈ D( H)).
Thus, for f ∈ D( H), the signed trace str((ιefe)′)e∈E1 exists, and it describes the
"ingoing derivatives" from the endpoints of the intervals. It is to be understood
8
C. Seifert, J. Voigt
that for (ιefe)′ we choose representatives of bounded variation (which exist by
the explanation given at the end of Section 1), in order to be able to apply the
signed trace mapping.
Let
X0 :=(cid:8)x ∈ X; prV1 x = 0(cid:9),
which could also be expressed as X0 := X ∩ KE ′
as a subspace of KE ′
onto X0. Also, for v ∈ V1, let ξv ∈ X be such that ξv
Lemma 2.3).
1∪V1), and let Q0 be the orthogonal projection from KE ′
1 (with our understanding of KE ′
1∪V1
1
V1 = 1{v} (see the proof of
In the following, for f ∈ D(ι) we will use the shorthand notation (ιf )′ :=
.
(cid:0)(ιefe)′(cid:1)e∈E1
3.1 Theorem. The operator H associated with the form τ is given by
D(H) =(cid:8)f ∈ HΓ; (fe)e∈E1 ∈ D( H), tr(ιf ) ∈ X,
((Hf )e)e∈E1 = H(fe)e∈E1,
(Hf )v =
1
µv(cid:0)L tr(ιf ) − str(ιf )′(cid:12)(cid:12) ξv(cid:1)
Q0 str(ιf )′ = Q0L tr(ιf )(cid:9),
(v ∈ V1).
Proof. (i) A preliminary step: Let f ∈ D( H), g ∈ D(τ ). For all e ∈ E1 one has
Z be
ae
(ιefe)′(x)(ιege)′(x) dx = −Z(ae,be)
∂µe(ιefe)′(x)ge(x) dµe(x)
+ (ιefe)′(be−)ιege(be) − (ιefe)′(ae+)ιege(ae);
cf. [2; equ. (1.2)]. Summing this equation over all the edges in E1 we obtain
Xe∈E1Z be
ae
(ιefe)′(x)(ιege)′(x) dx = ( Hf g)HE
−(cid:0)str((ιefe)′)e∈E1(cid:12)(cid:12) tr (ιg)(cid:1)KE ′
1 .
(ii) Let f ∈ D(H), g ∈ D(τ ). From D(H) ⊆ D(τ ) we conclude that tr(ιf ) ∈
X. As in [2; proof of Theorem 1.9] one obtains that (fe)e∈E1 ∈ D( H), H(fe)e∈E1 =
. Using part (i) above we obtain
(cid:0)(Hf )e(cid:1)e∈E1
(Hf g)HΓ
= −Xe∈E1Z(ae,be)
= Xe∈E1Z be
ae
∂µe(ιefe)′(x)ge(x) dµe(x) + Xv∈V1
(Hf )vgv µv
(ιefe)′(x)(ιege)′(x) dx +(cid:0)str(ιf )′(cid:12)(cid:12) tr (ιg)(cid:1)KE ′
1 + Xv∈V1
(Hf )vgv µv.
9
(3.1)
Dirichlet forms on graphs
Because of
(Hf g)HΓ = Xe∈E1Z be
ae
we therefore obtain
(ιefe)′(x)(ιege)′(x) dx +(cid:0)L tr(ιf )(cid:12)(cid:12) tr(ιg)(cid:1)
Xv∈V1
(Hf )vgv µv =(cid:0)L tr(ιf ) − str(ιf )′(cid:12)(cid:12) tr(ιg)(cid:1).
For ξ ∈ X0 choose g ∈ D(τ ) satisfying tr(ιg) = ξ, and g affine linear on the
edges e ∈ E1. Then equation (3.1) implies
This shows that Q0L tr(ιf ) = Q0 str(ιf )′.
0 =(cid:0)L tr(ιf ) − str(ιf )′(cid:12)(cid:12) ξ(cid:1).
Let v ∈ V1, and choose g ∈ D(τ ) satisfying tr(ιg) = ξv, and g affine linear on
the edges e ∈ E1. Then equation (3.1) yields
(Hf )v µv = Xw∈V1
(Hf )wξv(w) µw =(cid:0)L tr(ιf ) − str(ιf )′(cid:12)(cid:12) ξv(cid:1).
This shows the second part of the formula for Hf .
(iii) Now let H denote the operator indicated on the right hand side of the
assertion, and let f ∈ D( H). Then f ∈ D(τ ). Let g ∈ D(τ ). Then ξ :=
tr(ιg) −Pv∈V1
gvξv ∈ X0, and therefore
(cid:0)L tr(ιf ) − str(ιf )′(cid:12)(cid:12) ξ(cid:1) =(cid:0)Q0(L tr(ιf ) − str(ιf )′)(cid:12)(cid:12) ξ(cid:1) = 0.
Using part (i) as well as the previous equality we obtain
( Hf g)HΓ
=(cid:0) H(fe)e∈E1(cid:12)(cid:12) (ge)e∈E1(cid:1)HE
= Xe∈E1Z be
ae
1
µv
+Xv∈V1
(L tr(ιf ) − str(ιf )′ ξv)gv µv
(ιefe)′(x)(ιege)′(x) dx +(cid:0)str(ιf )′(cid:12)(cid:12) tr (ιg)(cid:1)
+ (L tr(ιf ) − str(ιf )′ tr (ιg))
= τ (f, g).
The definition of H then yields that f ∈ D(H) and Hf = Hf .
3.2 Remarks. (a) For f ∈ D(H) and v ∈ V1, the expression given for (Hf )v
given in Theorem 3.1 does not depend on the choice of ξv.
(b) The case of a weight µv > 0 at a vertex leads to a case of Wentzell
boundary condition at v. The expression of (Hf )v in Theorem 3.1 generalises
the expression obtained at a boundary point in the case of a single interval; cf.
[8; Poposition 4.3].
10
C. Seifert, J. Voigt
4 Positivity and contractivity
In this section we indicate conditions for the C0-semigroup (e−tH )t>0 to be positive
or submarkovian. We assume that the hypotheses are as in Section 2 and that
(2.1) holds.
In the following we need the notion of a (Stonean) sublattice of Kn. We
consider Kn as the function space C({1, . . . , n}), and accordingly use the notation
x = (x1, . . . , xn), for x ∈ Kn, and x ∧ y = (x1 ∧ y1, . . . , xn ∧ yn), for x, y ∈ Rn.
A sublattice X of Kn is a subspace for which x ∈ X implies that x ∈ X. A
sublattice X is called Stonean if additionally x ∧ 1 ∈ X for all real x ∈ X.
We refer to [2; Appendix] for the description of (Stonean) sublattices of Kn
and of generators for positive (submarkovian) C0-semigroups on these sublattices.
4.1 Theorem. (a) Assume that X is a sublattice of KE ′
1∪V1 and that the semi-
group (e−tL)t>0 is positivity preserving. Then (e−tH )t>0 is positivity preserving.
(b) Assume that X is a Stonean sublattice of KE ′
1∪V1 and that the semigroup
(e−tL)t>0 is a submarkovian semigroup. Then (e−tH)t>0 is submarkovian.
This result was proved for the case of local boundary conditions (cf. Section 5)
and no vertex masses in [2; Theorem 3.5], and for the case of vertices with masses
and local boundary conditions of continuity (cf. Example 5.3) in [2; Theorem 4.2].
Its proof is completely analogous to [2; proof of Theorem 3.5]; so we refrain from
giving a complete proof but rather only mention the main ingredients. The proof
consists in an application of the Beurling-Deny criteria (cf. [6; Corollary 2.18];
see also [2; Remarks 1.6]). So, in order to prove part (a), it is equivalent to
prove that the normal contraction f 7→ f acts on D(τ ), and that τ (f ) 6 τ (f )
for all f ∈ D(τ ). That the inequality works on the differential part is a one-
dimensional issue which is taken care of in [2; Theorem 1.7]. For the trace part,
the main observation is the equation tr ιf = tr ιf . This is less obvious than it
might appear at the first glance since, in general, one does not have ιf = ιf .
However, this equality holds on spt µe, and therefore at the end points of the
intervals [ae, be], for all e ∈ E1. The reasoning for part (b) is analogous.
5 Local boundary conditions
So far, the structure of the graph did not enter the considerations; in fact the
function γ linking the edges to the vertices was not used at all.
In order to
explain what we understand by local boundary conditions, we need the following
definitions.
For v ∈ V , the sets
describe the sets of all edges having mass and starting or ending at v, respectively,
and the set
(j = 0, 1)
E1,v,j :=(cid:8)e ∈ E1; γj(e) = v(cid:9)
E1,v :=(cid:0)E1,v,0 × {0}(cid:1) ∪(cid:0)E1,v,1 × {1}(cid:1)
Dirichlet forms on graphs
11
is the set of all edges having mass connected with v (and where loops starting
and ending at v yield two contributions). Note that then E′
Recall that the boundary conditions are specified by the choice of a subspace
1∪V1 and a self-adjoint operator L in X. The boundary conditions will
X ⊆ KE ′
be called local if for each v ∈ V there exists a subspace
1 =Sv∈V E1,v.
Xv ⊆ KE1,v
if v ∈ V0,
Xv ⊆ KE1,v ∪{v}
if v ∈ V1,
and a selfadjoint operator Lv in Xv, such that
Xv,
X =Mv∈V
Lv.
L =Mv∈V
For v ∈ V , we define the "local trace mapping"
by
trv : Ye∈E1
C[ae, be] × KV1 →(cid:26) KE1,v
KE1,v∪{v}
if v ∈ V0,
if v ∈ V1
trv f :=( tr f E1,v
tr f E1,v∪{v}
if v ∈ V0,
if v ∈ V1.
Then for the form τ we obtain
D(τ ) =(cid:8)f ∈ D(ι); trv(ιf ) ∈ Xv (v ∈ V )(cid:9),
τ (f, g) = Xe∈E1Z be
(ιefe)′(x)(ιege)′(x) dx +Xv∈V (cid:0)Lv trv(ιf )(cid:12)(cid:12) trv(ιg)(cid:1).
ae
Xv,0 :=(cid:26) Xv
if v ∈ V0,
{ξ ∈ Xv; ξ(v) = 0} if v ∈ V1,
With
the condition (2.1) for D(τ ) to be dense then decomposes into
Xv,0 6= Xv
(v ∈ V1),
or expressed differently, for all v ∈ V1 there exists ξv ∈ Xv such that ξv(v) = 1.
It is an easy task to translate the description of the associated operator H,
given in Theorem 3.1, to the present case of local boundary conditions, as follows.
5.1 Corollary. The operator H associated with τ is given by
D(H) =(cid:8)f ∈ HΓ; (fe)e∈E1 ∈ D( H), trv(ιf ) ∈ Xv,
Qv,0 strv(ιf )′ = Qv,0Lv trv(ιf ) (v ∈ V )(cid:9),
((Hf )e)e∈E1 = H(fe)e∈E1,
(Hf )v =
1
µv(cid:0)Lv trv(ιf ) − strv(ιf )′(cid:12)(cid:12) ξv(cid:1)
(v ∈ V1).
12
C. Seifert, J. Voigt
Here, for v ∈ V the mapping strv : Qe∈E1
BV (ae, be) → KE1,v is defined by
strv f := (str f ) E1,v, and Qv,0 is the orthogonal projection onto Xv,0 in KE1,v , for
v ∈ V0, or in KE1,v∪{v}, for v ∈ V1. We will not put down further details here.
Similarly, the conditions for (e−tH )t>0 to be positive and submarkovian, Theorem
4.1, can be spelled out in terms of the spaces Xv and the operators Lv. The
statements are then analogous to [2; Theorem 3.5], where the case that E = E1
and V = V0 is treated.
5.2 Corollary. (a) Assume that Xv is a sublattice of KE1,v (v ∈ V0) or KE1,v∪{v}
(v ∈ V1) and that (e−tLv)t>0 positivity preserving, for all v ∈ V . Then (e−tH)t>0
is a positivitiy preserving C0-semigroup on HΓ.
(b) Assume that Xv is a Stonean sublattice of KE1,v (v ∈ V0) or KE1,v∪{v}
(v ∈ V1) and that (e−tLv )t>0 is a submarkovian C0-semigroup on Xv, for all
v ∈ V . Then (e−tH )t>0 is a submarkovian C0-semigroup on HΓ.
5.3 Example (local boundary conditions of continuity). This special case of
local boundary conditions was studied in [2; Section 4]. In our framework, this
example reads as follows. Let Xv = lin{1}, Lv ∈ R, lv := (Lv1 1) for v ∈ V .
Then Xv,0 = {0} for v ∈ V1 (which makes it clear that condition (2.1) is satisfied)
and hence
Qv,0 =(cid:26) (· 1)1 if v ∈ V0,
if v ∈ V1.
0
Functions f ∈ D(τ ) are continuous on Γ,
i.e., for v ∈ V1 we have f (v) =
(trv f )(e, j) for all (e, j) ∈ E1,v, and for v ∈ V0 there exists av(f ) ∈ K such
that av(f ) = (trv f )(e, j) for all (e, j) ∈ E1,v (note that we cannot write f (v)
in this case since f is not defined on V0). The second part of the boundary
conditions for f ∈ D(H) translates to
Xe∈E1,v,0
f ′
e(ae+) − Xe∈E1,v,1
f ′
e(be−) = lvav(f )
(v ∈ V0);
see also [2; Theorem 4.3]. In the setup considered in [4], these boundary condi-
tions are called δ-type conditions; cf. [4; Section 3.2.1].
References
[1] F. Chung and L. Lu: Complex graphs and networks. American Mathematical
Society, 2006
[2] U. Kant, T. Klauss, J. Voigt and M. Weber: Dirichlet forms for singular one-
dimensional operators and on graphs. J. Evol. Equ. 9, 637 -- 659 (2009).
[3] V. Kostrykin, J. Potthoff and R. Schrader: Contraction semigroups on met-
ric graphs. In: Analysis on Graphs and its Applications, P. Exner (ed.) et
Dirichlet forms on graphs
13
al., Amer. Math. Soc., Providence, RI, Proc. Symp. Pure Math. 77, 423 -- 458
(2008).
[4] P. Kuchment: Quantum graphs: I. Some basic structures. Waves Random Me-
dia 14, 107 -- 128 (2007).
[5] P. Kuchment: Quantum graphs: an introduction and a brief survey. In: Anal-
ysis on Graphs and its Applications, P. Exner (ed.) et al., Amer. Math. Soc.,
Providence, RI, Proc. Symp. Pure Math. 77, 291 -- 314 (2008).
[6] E. M. Ouhabaz: Analysis of heat equations on domains. Princeton University
Press, Princeton, 2005.
[7] C. Seifert: Behandlung singularer Diffusion mit Hilfe von Dirichletformen.
Diploma Thesis, 2009.
http://www-user.tu-chemnitz.de/seifch/Forschung/Diplomarbeit.pdf
[8] H. Vogt and J. Voigt: Wentzell boundary conditions on the context of Dirich-
let forms. Adv. Differ. Equ. 8, No. 7, 821-842 (2003)
Christian Seifert
Fakultat Mathematik
Technische Universitat Chemnitz
09107 Chemnitz, Germany
[email protected]
Jurgen Voigt
Fachrichtung Mathematik
Technische Universitat Dresden
01062 Dresden, Germany
[email protected]
|
1307.6741 | 1 | 1307 | 2013-07-25T13:43:14 | On eigenfunction expansions of first-order symmetric systems and ordinary differential operators of an odd order | [
"math.FA"
] | We study general (not necessarily Hamiltonian) first-order symmetric systems $J y'-B(t)y=\D(t) f(t)$ on an interval $\cI=[a,b\rangle $ with the regular endpoint $a$. It is assumed that the deficiency indices $n_\pm(\Tmi)$ of the minimal relation $\Tmi$ satisfy $n_+(\Tmi)< n_-(\Tmi)$. We define
$\l$-depending boundary conditions which are analogs of separated self-adjoint boundary conditions for Hamiltonian systems. With a boundary value problem involving such conditions we associate an exit space self-adjoint extension $\wt T$ of $\Tmi$ and the $m$-function $m(\cd)$, which is an analog of the Titchmarsh-Weyl coefficient for the Hamiltonian system. By using $m$-function we obtain the eigenfunction expansion with the spectral function $\Si(\cd)$ of the minimally possible dimension and characterize the case when spectrum of $\wt T$ is defined by $\Si(\cd)$. Moreover, we parametrize all spectral functions in terms of a Nevanlinna type boundary parameter. Application of these results to ordinary differential operators of an odd order enables us to complete the results by Everitt and Krishna Kumar on the Titchmarsh-Weyl theory of such operators. | math.FA | math |
ON EIGENFUNCTION EXPANSIONS OF FIRST-ORDER SYMMETRIC
SYSTEMS AND ORDINARY DIFFERENTIAL OPERATORS OF AN
ODD ORDER
VADIM MOGILEVSKII
Abstract. We study general (not necessarily Hamiltonian) first-order symmetric sys-
tems J y′ − B(t)y = ∆(t)f (t) on an interval I = [a, bi with the regular endpoint a. It
is assumed that the deficiency indices n±(Tmin) of the minimal relation Tmin satisfy
n+(Tmin) < n−(Tmin). We define λ-depending boundary conditions which are analogs
of separated self-adjoint boundary conditions for Hamiltonian systems. With a boundary
value problem involving such conditions we associate an exit space self-adjoint extension
eT of Tmin and the m-function m(·), which is an analog of the Titchmarsh-Weyl coefficient
for the Hamiltonian system. By using m-function we obtain the eigenfunction expansion
with the spectral function Σ(·) of the minimally possible dimension and characterize the
case when spectrum of eT is defined by Σ(·). Moreover, we parametrize all spectral func-
tions in terms of a Nevanlinna type boundary parameter. Application of these results
to ordinary differential operators of an odd order enables us to complete the results by
Everitt and Krishna Kumar on the Titchmarsh-Weyl theory of such operators.
1. Introduction
Let H and bH be finite dimensional Hilbert spaces and let
(1.1)
H0 := H ⊕ bH,
H := H0 ⊕ H = H ⊕ bH ⊕ H.
In the paper we study first-order symmetric systems of differential equations defined on
an interval I = [a, bi, −∞ < a < b ≤ ∞, with the regular endpoint a and regular or singular
endpoint b. Such a system is of the form [3, 12]
(1.2)
J y′ − B(t)y = ∆(t)f (t),
t ∈ I,
where B(t) = B∗(t) and ∆(t) ≥ 0 are the [H]-valued functions on I and
System (1.2) is called a Hamiltonian system if bH = {0}.
any λ ∈ C each solution y(·) of the equation
Throughout the paper we assume that system (1.2) is definite. The latter means that for
(1.4)
J y′ − B(t)y = λ∆(t)y
satisfying ∆(t)y(t) = 0 (a.e. on I) is trivial, i.e., y(t) = 0, t ∈ I.
2010 Mathematics Subject Classification. 34B08, 34B20, 34B40,34L10,47A06,47B25.
Key words and phrases. First-order symmetric system, boundary parameter, m-function, spectral func-
tion of a boundary problem, Fourier transform.
1
(1.3)
J =
0
0
IH
0 −IH
iI bH
0
0
0 : H ⊕ bH ⊕ H → H ⊕ bH ⊕ H.
2
VADIM MOGILEVSKII
In what follows we denote by H := L2
∆(I) the Hilbert space of H-valued Borel functions
f (·) on I (in fact, equivalence classes) satisfying f 2
(∆(t)f (t), f (t))H dt < ∞.
∆ :=RI
Studying of symmetric systems is basically motivated by the fact that system (1.4) is a
more general objet than a formally self-adjoint differential equation of an arbitrary order
with matrix coefficients. In fact, such equation is reduced to the system (1.4) of a special
form with J given by (1.3); moreover, this system is Hamiltonian precisely in the case when
the differential equation is of an even order (we will concern this questions below).
As is known, the extension theory of symmetric linear relations gives a natural framework
for investigation of the boundary value problems for symmetric systems (see [4, 7, 20, 28, 36]
and references therein). According to [20, 28, 36] the system (1.2) generates the minimal
linear relation Tmin and the maximal linear relation Tmax in H.
It turns out that Tmin
is a closed symmetric relation with not necessarily equal deficiency indices n±(Tmin) and
min. Since system (1.2) is definite, n±(Tmin) can be defined as a number of L2
Tmax = T ∗
∆-
solutions of (1.4) for λ ∈ C±.
A description of various classes of extensions of Tmin (self-adjoint, m-dissipative, etc.) in
terms of boundary conditions is an important problem in the spectral theory of symmetric
systems. For Hamiltonian system (1.2) self-adjoint separated boundary conditions were
described in [16]. Moreover, the Titchmarsh -- Weyl coefficient MT W (λ) of the boundary value
problem for Hamiltonian system with self-adjoint separated boundary conditions was defined
by various methods in [16, 25, 23]. Using MT W (·) one obtains the Fourier transform with
the spectral function Σ(·) of the minimally possible dimension NΣ = dim H (see [7, 18, 20]).
At the same time according to [34] non-Hamiltonian system (1.2) does not admit self-adjoint
separated boundary conditions. Moreover, the inequality n+(Tmin) 6= n−(Tmin), and hence
absence of self-adjoint boundary conditions is a typical situation for such systems. Therefore
the following problems are of certain interest:
• To find (may be λ-depending) analogs of self-adjoint separated boundary conditions
for general (not necessarily Hamiltonian) systems (1.2) and describe such type conditions;
• To describe in terms of boundary conditions all spectral matrix functions that have
the minimally possible dimension and investigate the corresponding Fourier transforms.
In the paper [2] these problems were considered for symmetric systems (1.2) satisfying
n−(Tmin) ≤ n+(Tmin). In the present paper we solve these problems in the opposite case
n+(Tmin) < n−(Tmin). It turns out that this case requires a somewhat another approach in
comparison with [2], although the ideas of both the papers are similar. Moreover, we show
that in the case n+(Tmin) < n−(Tmin) there is a class of exit space self-adjoint extensions
of Tmin with special spectral properties. We apply also the obtained results to ordinary
differential operators of an odd order with matrix valued coefficients and arbitrary deficiency
indices.
Let νb+ and νb− be indices of inertia of the skew-Hermitian bilinear form [y, z]b defined
on dom Tmax by
[y, z]b = lim
t↑b
(J y(t), z(t)),
y, z ∈ dom Tmax,
It turns out that system (1.2) with n+(Tmin) < n−(Tmin) has different properties depending
on the sign of νb+ − νb−. Within this section we present the results of the paper assuming
a simpler case νb+ − νb− ≤ 0.
A crucial role in our considerations is played by a symmetric linear relation T (⊃ Tmin) in H
Let a function y ∈ dom Tmax be represented as y(t) = {y0(t),by(t), y1(t)}(∈ H ⊕ bH ⊕ H).
ON EIGENFUNCTION EXPANSIONS
3
given by means of boundary conditions as follows:
T = {{y, f } ∈ Tmax : y1(a) = 0, by(a) = 0, eΓ0by = Γ1by = 0}
and Γ1by are singular boundary values of a function y ∈ dom Tmax at the endpoint b.
Here eΓ0b : dom Tmax → eHb and Γ1b : dom Tmax → Hb are linear mappings with special
properties, eHb and Hb(⊂ eHb) are auxiliary finite dimensional Hilbert spaces. In fact , eΓ0by
Recall that a linear relation eT = eT ∗ in a wider Hilbert space eH ⊃ H satisfying T ⊂ eT is
called an exit space self-adjoint extension of T . Moreover, a generalized resolvent R(·) and
a spectral function F (·) of T are defined by
t ∈ R,
R(λ) = PH(eT − λ)−1 ↾ H, λ ∈ C \ R, and F (t) = PHE(t) ↾ H,
exit space self-adjoint extensions exist for symmetric linear relations with arbitrary (possibly
unequal) deficiency indices.
where E(·) is the orthogonal spectral function (resolution of identity) of eT . As is known [1]
eR−(bH ⊕ eHb, Hb) of holomorphic operator pairs τ = {D0(λ), D1(λ)}. Such a pair is formed
To describe the set of all generalized resolvents of T we use the Nevanlinna type class
D0(λ) = (bD0(λ), eD0b(λ)) : bH ⊕ eHb → (bH ⊕ eHb) and D1(λ)(∈ [Hb, bH ⊕ eHb])
with special properties (see [35]). We show that each generalized resolvent y = R(λ)f, f ∈
H, is given as the L2
∆-solution of the following boundary-value problem with λ-depending
boundary conditions:
by defined on C− holomorphic operator functions
(1.5)
(1.6)
(1.7)
y1(a) = 0,
λ ∈ C−.
J y′ − B(t)y = λ∆(t)y + ∆(t)f (t),
t ∈ I,
ibD0(λ)by(a) + eD0b(λ)eΓ0by + D1(λ)Γ1by = 0,
Here (bD0(λ), eD0b(λ)) =: D0(λ) and D1(λ) are components of a pair τ = {D0(λ), D1(λ)} ∈
eR−(bH ⊕ eHb, Hb) (see (1.5)), so that the second equality in (1.7) is a Nevanlinna type bound-
ary condition involving boundary values of a function y at both endpoints a and b. Thus, in-
vestigation of boundary value problems for the system (1.2) in the case n+(Tmin) < n−(Tmin)
require use of boundary conditions of another class in comparison with the case n−(Tmin) ≤
n+(Tmin)(cf.[2]). One may consider a pair τ as a boundary parameter, since R(λ) runs over
the set of all generalized resolvents of T when τ runs over the set of all holomorphic operator
and F (t) = Fτ (t) for the generalized resolvents and spectral functions of T respectively.
pairs of the class eR−(bH ⊕ eHb, Hb). To indicate this fact explicitly we write R(λ) = Rτ (λ)
Moreover, we denote by eT = eT τ the exit space self-adjoint extension of T generating Rτ (·)
Next assume that ϕ(·, λ) and ψ(·, λ) are [H0, H]-valued operator solutions of the equation
and Fτ (·).
(1.4) satisfying the initial conditions
ϕ(a, λ) =(cid:18)IH0
0 (cid:19) (∈ [H0, H0 ⊕ H]), ψ(a, λ) =(cid:18)− i
−PH (cid:19) (∈ [H0, H0 ⊕ H]).
2 P bH
(here PH and P bH are the orthoprojectors in H0 onto H and bH respectively). We show that,
for each Nevanlinna boundary parameter τ = {D0(λ), D1(λ)} of the form (1.5), there exists
a unique operator function mτ (·) : C \ R → [H0] such that the operator solution
vτ (t, λ) := ϕ(t, λ)mτ (λ) + ψ(t, λ)
bf (s) =ZI
ϕ∗(t, s) ∆(t) f (t) dt
satisfies
(1.8)
4
VADIM MOGILEVSKII
(1.4) has the following property:
for every h0 ∈ H0 the function y = vτ (t, λ)h0
of Eq.
belongs to L2
∆(I) and satisfies the boundary condition
We call mτ (·) the m-function corresponding to the boundary value problem (1.6), (1.7); re-
ally, mτ (·) is an analog of the Titchmarsh-Weyl coefficient MT W (λ) for Hamiltonian systems.
It turns out that mτ (·) is a Nevanlinna operator function satisfying the inequality
bD0(λ)(iby(a) − P bH h0) + eD0b(λ)eΓ0by + D1(λ)Γ1by = 0, λ ∈ C−.
(Im λ)−1 · Im mτ (λ) ≥ZI
v∗
τ (t, λ)∆(t)vτ (t, λ) dt,
λ ∈ C \ R.
Next we study eigenfunction expansions of the boundary value problems for symmetric
systems. Namely, let τ be a boundary parameter and let Fτ (·) be the spectral function of
T generated by the boundary value problem (1.6), (1.7). A distribution operator-valued
function Στ (·) : R → [H0] is called a spectral function of this problem if, for each function
f ∈ H with compact support, the Fourier transform
((Fτ (β) − Fτ (α))f, f )H =Z[α,β)
(dΣτ (s)bf (s),bf (s))
for any compact interval [α, β) ⊂ R. We show that for each boundary parameter τ there
exists a unique spectral function Στ (·) and it is recovered from the m-function mτ (·) by
means of the Stieltjes inversion formula
(1.9)
Στ (s) = − lim
δ→+0
lim
ε→+0
1
πZ s−δ
−δ
Im mτ (σ − iε) dσ.
Below (within this section) we assume for simplicity that T is a (not necessarily densely
defined) operator, i.e., mul T = {0}.
one may define the inverse Fourier transform in the explicit form (see (5.30)). Moreover, if
ous extension to a contractive linear mapping V : H → L2(Στ ; H0) (for the strict definition
of the Hilbert space L2(Στ ; H0) see e.g.
[8, Ch.13.5]. As in [2] one proves that V is an
It follows from (1.8) that the mapping V f = bf (the Fourier transform) admits a continu-
isometry (that is, the Parseval equality bfL2(Στ ;H0) = f H holds for every f ∈ H) if and
only if the exit space extension eT τ in eH ⊃ H is an operator, i.e., muleT τ = {0}. In this case
V is an isometry, then there exists a unitary extension eV ∈ [eH, L2(Στ ; H0)] of V such that
the operator eT τ and the multiplication operator Λ in L2(Στ ; H0) are unitarily equivalent by
means of eV . Hence, the operators eT τ and Λ have the same spectral properties; for instance,
multiplicity of the spectrum of eT τ does not exceed dim H0(= dim H + dim bH).
D1(λ) = (0, D1(λ))⊤(∈ [Hb, bH ⊕ eHb]).
(1.10) D0(λ) = diag (I bH , D0(λ))(∈ [bH ⊕ eHb]),
mτ (λ) =(cid:18)mτ,1(λ) mτ,2(λ)
We show that in this case the corresponding m-function mτ (·) is of the triangular form
Now assume that the boundary parameter τ = {D0(λ), D1(λ)} is (cf. (1.5))
2 I bH (cid:19) : H ⊕ bH → H ⊕ bH,
λ ∈ C−,
0
− i
ON EIGENFUNCTION EXPANSIONS
5
so that the spectral function Στ (·) has the block matrix representation
(1.11)
Στ (s) =(cid:18)Στ,1(s) Στ,2(s)
Στ,3(s)
1
2π sI bH(cid:19) : H ⊕ bH → H ⊕ bH,
s ∈ R.
Here Στ,1(·) is an [H]-valued distribution function, which can be defined by means of the
Stieltjes inversion formula for mτ,1(λ).
It follows from (1.11), that in the case when a
boundary parameter τ = {D0(λ), D1(λ)} is of the form (1.10) and the Fourier transform is
spectral properties (for more details see Theorem 5.14):
an isometry, the corresponding exit space self-adjoint extension eT τ of T has the following
(S1) σac(eT τ ) = R, where σac(eT τ ) is the absolutely continuous spectrum of eT τ .
(S2) σs(eT τ ) = Ss(Στ,1), where σs(eT τ ) is the singular spectrum of eT τ and Ss(Στ,1) is
multiplicity of the singular spectrum of eT τ does not exceed dim H, which yields the same
estimate for multiplicity of each eigenvalue λ0 of eT τ .
Next, we show that all spectral functions Στ (·) can be parametrized immediately in terms
of the boundary parameter τ . More precisely, we show that there exists an operator function
a closed support of the measure generated by the singular component of Στ,1. Hence the
(1.12)
X−(λ) =(cid:18)m0(λ) Φ−(λ)
Ψ−(λ)
M−(λ)(cid:19) : H0 ⊕ (bH ⊕ eHb) → H0 ⊕ Hb,
λ ∈ C−,
such that for each boundary parameter τ = {D0(λ), D1(λ)} the corresponding m-function
mτ (·) is given by
(1.13)
mτ (λ) = m0(λ) + Φ−(λ)(D0(λ) − D1(λ) M−(λ))−1D1(λ)Ψ−(λ),
λ ∈ C−.
Thus, formula (1.13) together with the Stieltjes inversion formula (1.9) defines a (unique)
spectral function Στ (·) of the boundary value problem (1.6), (1.7). Note that entries of the
matrix (1.12) are defined in terms of the boundary values of respective operator solutions
of Eq.(1.4). We also describe boundary parameters τ for which the Fourier transform V is
an isometry and characterize the case when V is an isometry for every boundary parameter
τ (see Theorem 5.15). Note that a description of spectral functions for various classes of
boundary problems in the form close to (1.13), (1.9) can be found in [11, 13, 15, 22, 21, 33].
Clearly, all the foregoing results can be reformulated (with obvious simplifications) for
Hamiltonian systems (1.2).
We suppose that the above results about exit space extensions eT τ of T may be useful in
the case when eT τ is a self-adjoint relation in eH = L2
on the whole line R. More precisely, we assume that spectral properties of such eT τ may
be characterized in terms of the objects (m-function, boundary parameter etc.) associated
with the restriction of this system onto the semi-axis I = [0, ∞). These problems will be
considered elsewhere.
∆(R) induced by the symmetric system
If n+(Tmin) takes on the minimally possible value n+(Tmin) = dim H, then n+(Tmin) ≤
n−(Tmin) and the above results can be rather simplified. Namely, in this case T is a maximal
symmetric relation and hence there exists a unique generalized resolvent of T . Therefore
there is a unique m-function m(·) which has the triangular form
(1.14)
m(λ) =(cid:18)M (λ) N−(λ)
− i
0
2 I bH(cid:19) : H ⊕ bH → H ⊕ bH,
λ ∈ C−.
6
VADIM MOGILEVSKII
(M ∗(λ), N ∗
−(λ))⊤, where M (λ) and N−(λ) are taken from (1.14).
Moreover, the spectral function Σ(·) of the corresponding boundary value problem is of the
form (1.11) and the Fourier transform V is an isometry. Therefore a (unique) exit space
self-adjoint extension T0 of T has the spectral properties (S1) and (S2).
der [17], where the concept of the "rectangular" Titchmarsh-Weyl coefficient MT W (λ)(∈
Note that systems (1.2) with bH 6= {0} and both minimal deficiency indices n+(Tmin) =
dim H and n−(Tmin) = dim H + dim bH were studied in the paper by Hinton and Schnei-
[H, H ⊕ bH]), λ ∈ C+, was introduced. One can easily show that in fact MT W (λ) =
In the final part of the paper we consider the operators generated by a differential ex-
pression l[y] of an odd order 2m + 1 with [H]-valued coefficients (H is a Hilbert space with
r := dim H < ∞) defined on an interval I = [a, bi (see (6.1)). In the particular case of scalar
coefficients such differential operators have been investigated in the papers by Everitt and
Krishna Kumar [9, 10, 26], where the limiting process from the compact intervals [a, β] ⊂ I
was used for construction of the Titchmarsh-Weyl matrix MT W (λ) = (mjk(λ))m+1
j,k=1. Note
that the results of these papers can not be considered to be completed; in particular, an
attempt to define self-adjoint boundary conditions in [10] gave rise to hardly verifiable as-
sumptions even in the case of minimally possible equal deficiency indices n±(Lmin) = m + 1
of the minimal operator Lmin.
Our approach is based on the known fact [24] that the equation l[y] = λy is equivalent
to a special symmetric non-Hamiltonian system (1.4). This enables us to extend the ob-
tained results concerning symmetric systems to the expression l[y] with arbitrary (possibly
unequal) deficiency indices of Lmin. In particular, we describe self-adjoint and λ-depending
boundary conditions for l[y], which are analogs of self-adjoint separated boundary conditions
for differential expressions of an even order. This makes it possible to construct eigenfunc-
tion expansion with spectral matrix function Σ(·) of the dimension (m + 1)r × (m + 1)r and
to describe all Σ(·) immediately in terms of boundary conditions (for operators of an even
order and separated boundary conditions such a description was obtained in [33]).
In conclusion note that the above results are obtained with the aid of the method of
boundary triplets and the corresponding Weyl functions in the extension theory of symmetric
linear relations (see [6, 14, 29, 31, 32] and references therein).
2. Preliminaries
2.1. Notations. The following notations will be used throughout the paper: H, H denote
Hilbert spaces; [H1, H2] is the set of all bounded linear operators defined on the Hilbert
space H1 with values in the Hilbert space H2; [H] := [H, H]; A ↾ L is the restriction of an
operator A onto the linear manifold L; PL is the orthogonal projector in H onto the subspace
L ⊂ H; C+ (C−) is the upper (lower) half-plane of the complex plane.
Recall that a closed linear relation from H0 to H1 is a closed linear subspace in H0 ⊕ H1.
The set of all closed linear relations from H0 to H1 (in H) will be denoted by eC(H0, H1)
(eC(H)). A closed linear operator T from H0 to H1 is identified with its graph gr T ∈
eC(H0, H1).
For a linear relation T ∈ eC(H0, H1) we denote by dom T, ran T, ker T and mul T the
and adjoint linear relations of T are the relations T −1 ∈ eC(H1, H0) and T ∗ ∈ eC(H1, H0)
domain, range, kernel and the multivalued part of T respectively. Recall also that the inverse
ON EIGENFUNCTION EXPANSIONS
7
defined by
(2.1)
T ∗ = {{k1, k0} ∈ H1 ⊕ H0 : (k0, h0) − (k1, h1) = 0, {h0, h1} ∈ T }.
T −1 = {{h1, h0} ∈ H1 ⊕ H0 : {h0, h1} ∈ T }
In the case T ∈ eC(H0, H1) we write 0 ∈ ρ(T ) if ker T = {0} and ran T = H1, or
equivalently if T −1 ∈ [H1, H0]; 0 ∈ bρ(T ) if ker T = {0} and ran T is a closed subspace in
H1. For a linear relation T ∈ eC(H) we denote by ρ(T ) := {λ ∈ C : 0 ∈ ρ(T − λ)} and
bρ(T ) = {λ ∈ C : 0 ∈ bρ(T − λ)} the resolvent set and the set of regular type points of T
Recall also the following definition.
respectively.
Definition 2.1. A holomorphic operator function Φ(·) : C \ R → [H] is called a Nevanlinna
function if Im λ · ImΦ(λ) ≥ 0 and Φ∗(λ) = Φ(λ), λ ∈ C \ R.
(2.2)
H = H0 ⊕ mul A,
the following decompositions hold
A = gr A′ ⊕ dmul A,
E′(t)PH0 , where E′(·) : R → [H0] is the orthogonal spectral function of A′
operator in H0 (the operator part of A); moreover, A = A∗ if and only if A′ = (A′)∗.
2.2. Symmetric and self-adjoint linear relations. A linear relation A ∈ eC(H) is called
symmetric (self-adjoint) if A ⊂ A∗ (resp. A = A∗). For each symmetric relations A ∈ eC(H)
where dmul A = {0} ⊕ mul A and A′ is a closed symmetric (not necessarily densely defined)
A spectral function E(·) : R → [H] of the relation A = A∗ ∈ eC(H) is defined as E(t) =
For an operator A = A∗ we denote by σ(A), σac(A) and σs(A) the spectrum, the abso-
lutely continuous spectrum and the singular spectrum of A respectively [19, Section 10.1].
Next assume that H is a finite dimensional Hilbert space. A non-decreasing operator
function Σ(·) : R → [H] is called a distribution function if it is left continuous and satisfies
Σ(0) = 0. For each distribution function Σ(·) there is a unique pair of distribution functions
Σac(·) and Σs(·) such that Σac(·) is absolutely continuous, Σs(·) is singular and Σ(t) =
Σac(t) + Σs(t) (the Lebesgue decomposition of Σ). For a distribution function Σ(·) we
denote by S(Σ) the set of all t ∈ R such that Σ(t − δ) 6= Σ(t + δ) for any δ > 0. Moreover,
we let Sac(Σ) = S(Σac) and Ss(Σ) = S(Σs).
With each distribution function Σ(·) one associates the Hilbert space L2(Σ; H) of all
[8,
(dΣ(t)f (t), f (t)) < ∞ (for more details see e.g.
functions f (·) : R → H such that RR
Section 13.5]). The following theorem is well known.
Theorem 2.2. Let Σ(·) be a [H]-valued distribution function. Then the relations
dom ΛΣ = {f ∈ L2(Σ; H) : tf (t) ∈ L2(Σ; H)},
(ΛΣf )(t) = tf (t), f ∈ dom ΛΣ
define a self-adjoint operator Λ = ΛΣ in L2(Σ; H) (the multiplication operator) and
σ(ΛΣ) = S(Σ),
σac(ΛΣ) = Sac(Σ),
σs(ΛΣ) = Ss(Σ).
2.3. The class eR−(H0, H1). Let H0 be a Hilbert space, let H1 be a subspace in H0 and
let τ = {τ+, τ−} be a collection of holomorphic functions τ±(·) : C± → eC(H0, H1). In the
paper we systematically deal with collections τ = {τ+, τ−} of the special class eR−(H0, H1)
introduced in [35]. In the case H0 = H1 =: H this class turns into the well known class
8
VADIM MOGILEVSKII
eR(H) of Nevanlinna functions τ (·) : C \ R → eC(H) (see for instance [5]). If dim H0 < ∞,
then according to [35] the collection τ = {τ+, τ−} ∈ eR−(H0, H1) admits the representation
τ+(λ) = {(C0(λ), C1(λ)); H1}, λ ∈ C+;
τ−(λ) = {(D0(λ), D1(λ)); H0}, λ ∈ C−
(2.3)
by means of holomorphic operator pairs
(C0(λ), C1(λ)) : H0 ⊕ H1 → H1, λ ∈ C+;
(D0(λ), D1(λ)) : H0 ⊕ H1 → H0, λ ∈ C−
(more precisely, by equivalence classes of such pairs). The equalities (2.3) mean that
(2.4)
τ+(λ) = {{h0, h1} ∈ H0 ⊕ H1 : C0(λ)h0 + C1(λ)h1 = 0}, λ ∈ C+
τ−(λ) = {{h0, h1} ∈ H0 ⊕ H1 : D0(λ)h0 + D1(λ)h1 = 0}, λ ∈ C−.
In [35] the class eR−(H0, H1) is characterized both in terms of eC(H0, H1)-valued functions
τ±(·) and in terms of operator functions Cj(·) and Dj(·), j ∈ {0, 1}, from (2.3).
2.4. Boundary triplets and Weyl functions. Here we recall some definitions and results
from our paper [35].
Let A be a closed symmetric linear relation in the Hilbert space H, let Nλ(A) = ker (A∗ −
that H0 = H1 ⊕ H2. Denote by Pj the orthoprojector in H0 onto Hj, j ∈ {1, 2}.
Next assume that H0 is a Hilbert space, H1 is a subspace in H0 and H2 := H0 ⊖ H1, so
n±(A) := dim Nλ(A) ≤ ∞, λ ∈ C±, be deficiency indices of A. Denote by ExtA the set of
Definition 2.3. A collection Π− = {H0 ⊕ H1, Γ0, Γ1}, where Γj : A∗ → Hj, j ∈ {0, 1}, are
λ) (λ ∈ bρ(A)) be a defect subspace of A, let bNλ(A) = {{f, λf } : f ∈ Nλ(A)} and let
all proper extensions of A, i.e., the set of all relations eA ∈ eC(H) such that A ⊂ eA ⊂ A∗.
linear mappings, is called a boundary triplet for A∗, if the mapping Γ : bf → {Γ0bf , Γ1bf},bf ∈
holds for all bf = {f, f ′}, bg = {g, g′} ∈ A∗.
(f ′, g) − (f, g′) = (Γ1bf , Γ0bg)H0 − (Γ0bf , Γ1bg)H0 − i(P2Γ0bf , P2Γ0bg)H2
Proposition 2.4. Let Π− = {H0 ⊕ H1, Γ0, Γ1} be a boundary triplet for A∗. Then:
A∗, from A∗ into H0 ⊕ H1 is surjective and the following Green's identity
(1) dim H1 = n+(A) ≤ n−(A) = dim H0.
(2) ker Γ0 ∩ ker Γ1 = A and Γj is a bounded operator from A∗ into Hj, j ∈ {0, 1}.
(3) The equality
(2.5)
defines the maximal symmetric extension A0 ∈ ExtA such that C− ⊂ ρ(A0).
A0 := ker Γ0 = {bf ∈ A∗ : Γ0bf = 0}
(2.6)
(2.8)
(2.9)
Proposition 2.5. Let Π− = {H0 ⊕ H1, Γ0, Γ1} be a boundary triplet for A∗ (so that in view
of Proposition 2.4, (1) n+(A) ≤ n−(A)). Denote also by π1 the orthoprojector in H ⊕ H
onto H ⊕ {0}. Then:
(1) The operators P1Γ0 ↾ bNλ(A), λ ∈ C+, and Γ0 ↾ bNz(A), z ∈ C−, isomorphically map
bNλ(A) onto H1 and bNz(A) onto H0 respectively. Therefore the equalities
(2.7)
γ+(λ) = π1(P1Γ0 ↾ bNλ(A))−1, λ ∈ C+,
γ−(z) = π1(Γ0 ↾ bNz(A))−1, z ∈ C−,
M+(λ)h1 = (Γ1 − iP2Γ0){γ+(λ)h1, λγ+(λ)h1},
h1 ∈ H1,
λ ∈ C+
M−(z)h0 = Γ1{γ−(z)h0, zγ−(z)h0},
h0 ∈ H0,
z ∈ C−
ON EIGENFUNCTION EXPANSIONS
9
correctly define the operator functions γ+(·) : C+ → [H1, H], γ−(·) : C− → [H0, H] and
M+(·) : C+ → [H1, H0], M−(·) : C− → [H0, H1], which are holomorphic on their domains.
Moreover, the equality M ∗
+(λ) = M−(λ), λ ∈ C−, is valid.
(2) Assume that
M+(λ) = (M (λ), N+(λ))⊤ : H1 → H1 ⊕ H2,
M−(z) = (M (z), N−(z)) : H1 ⊕ H2 → H1,
λ ∈ C+
z ∈ C−
are the block representations of M+(λ) and M−(z) respectively and let
(2.10)
(2.11)
0
i
M(λ) =(cid:18) M (λ)
M(λ) =(cid:18)M (λ) N−(λ)
2 IH2(cid:19) : H1 ⊕ H2 → H1 ⊕ H2,
2 IH2(cid:19) : H1 ⊕ H2 → H1 ⊕ H2,
N+(λ)
− i
0
λ ∈ C+
λ ∈ C−.
Then M(·) is a Nevanlinna operator function satisfying the identity
(2.12)
M(µ) − M∗(λ) = (µ − λ)γ∗
−(λ)γ−(µ),
µ, λ ∈ C−.
Definition 2.6.
the Weyl functions, respectively, corresponding to the boundary triplet Π−.
[35] The operator functions γ±(·) and M±(·) are called the γ-fields and
It follows from (2.7) that for each h1 ∈ H1 and h0 ∈ H0 the following equalities hold
(2.13)
P1Γ0{γ+(λ)h1, λγ+(λ)h1} = h1,
Γ0{γ−(z)h0, zγ−(z)h0} = h0.
Proposition 2.7. Let Π− = {H0 ⊕ H1, Γ0, Γ1} be a boundary triplet for A∗ and let γ±(·)
and M±(·) be the corresponding γ-fields and Weyl functions respectively. Moreover, let the
spaces H0 and H1 be decomposed as
(so that
H0 = H1 ⊕ H2) and let
be the block representations of the operators Γ0 and Γ1. Then:
(1) The equalities
H1 = bH ⊕ H1,
Γ0 = (bΓ0, Γ0)⊤ : A∗ → bH ⊕ H0,
eA = {bf ∈ A∗ :bΓ0bf = Γ0bf = Γ1bf = 0},
H0 = bH ⊕ H0
Γ1 = (bΓ1, Γ1)⊤ : A∗ → bH ⊕ H1
eA∗ = {bf ∈ A∗ :bΓ0bf = 0}
define a closed symmetric extension eA ∈ ExtA and its adjoint eA∗.
(2) The collection Π− = { H0 ⊕ H1, Γ0 ↾ eA∗, Γ1 ↾ eA∗} is a boundary triplet for eA∗.
(3) The γ-fields γ±(·) and the Weyl functions
M±(·) corresponding to Π− are given by
γ+(λ) = γ+(λ) ↾ H1,
γ−(λ) = γ−(λ) ↾ H0,
M+(λ) = P H0
M−(λ) = P H1
M+(λ) ↾ H1,
M−(λ) ↾ H0,
λ ∈ C+
λ ∈ C−.
The proof of Proposition 2.7 is similar to that of Proposition 4.1 in [5].
Remark 2.8. If H0 = H1 := H, then the boundary triplet in the sense of Definition 2.3
turns into the boundary triplet Π = {H, Γ0, Γ1} for A∗ in the sense of [14, 29]. In this case
n+(A) = n−(A)(= dim H) and the γ-fields γ±(·) and Weyl functions M±(·) turn into the
γ-field γ(·) and Weyl function M (·) respectively introduced in [6, 29]. Observe also that
along with Π− we define in [32, 35] a boundary triplet Π+ = {H0 ⊕ H1, Γ0, Γ1} for A∗. Such
a triplet is applicable to symmetric relations A with n−(A) ≤ n+(A).
10
VADIM MOGILEVSKII
2.5. Generalized resolvents and exit space extensions. The following definitions are
well known.
(2.14)
(2.15)
1, H2
2].
Definition 2.9. Let eH be a Hilbert space and let H be a subspace in eH. A relation eA =
eA∗ ∈ eC(eH) is called H-minimal if span{H, (eA − λ)−1H : λ ∈ C \ R} =eH.
Definition 2.10. The relations Tj ∈ eC(Hj), j ∈ {1, 2}, are said to be unitarily equivalent
(by means of a unitary operator U ∈ [H1, H2]) if T2 = eU T1 with eU = U ⊕ U ∈ [H2
spectral function of A respectively if there exist a Hilbert space eH ⊃ H and a self-adjoint
relation eA ∈ eC(eH) such that A ⊂ eA and the following equalities hold:
Definition 2.11. Let A be a symmetric relation in a Hilbert space H. The operator func-
tions R(·) : C \ R → [H] and F (·) : R → [H] are called the generalized resolvent and the
R(λ) = PH(eA − λ)−1 ↾ H,
(in formula (2.15) E(·) is the spectral function of eA).
The relation eA in (2.14) is called an exit space extension of A.
space extension eA of A. Moreover, if the H-minimal exit space extensions eA1 ∈ eC(eH1) and
eA2 ∈ eC(eH2) of A induce the same generalized resolvent R(λ), then there exists a unitary
operator V ′ ∈ [eH1 ⊖ H,eH2 ⊖ H] such that eA1 and eA2 are unitarily equivalent by means of
eV = IH ⊕ V ′. By using these facts we suppose in the following that the exit space extension
eA in (2.14) is H-minimal, so that eA is defined by R(·) uniquely up to the unitary equivalence.
Note that in this case the equality mul eA = mul A holds for any exit space extension eA = eA∗
of A if and only if mul A = mul A∗ or, equivalently, if and only if the operator A′ (the
operator part of A) is densely defined.
According to [27] each generalized resolvent of A is generated by some H-minimal exit
F (t) = PHE(t) ↾ H,
λ ∈ C \ R
t ∈ R
3. Boundary triplets for symmetric systems
3.1. Notations. Let I = [a, bi (−∞ < a < b ≤ ∞) be an interval of the real line (the
symbol i means that the endpoint b < ∞ might be either included to I or not). For a given
finite-dimensional Hilbert space H denote by AC(I; H) the set of functions f (·) : I → H
which are absolutely continuous on each segment [a, β] ⊂ I and let AC(I) := AC(I; C).
Next assume that ∆(·) is a [H]-valued Borel functions on I integrable on each com-
∆(I) the semi-Hilbert
(∆(t)f (t), f (t))H dt < ∞ (see
pact interval [a, β] ⊂ I and such that ∆(t) ≥ 0. Denote by L2
space of Borel functions f (·) : I → H satisfying f 2
∆ := RI
that the latter condition is equivalent toRI
1
∆
2 (t)F (t)2 dt < ∞.
e.g.
(f, g)∆ = RI
[8, Chapter 13.5]). The semi-definite inner product (·, ·)∆ in L2
(∆(t)f (t), g(t))H dt, f, g ∈ L2
∆(I). Moreover, let L2
∆(I) is defined by
∆(I) be the Hilbert space
of the equivalence classes in L2
quotient map from L2
∆(I) onto L2
∆(I).
∆(I) with respect to the semi-norm · ∆ and let π be the
For a given finite-dimensional Hilbert space K denote by L2
operator-functions F (·) : I → [K, H] such that F (t)h ∈ L2
∆[K, H] the set of all Borel
∆(I) for each h ∈ K. It is clear
ON EIGENFUNCTION EXPANSIONS
11
3.2. Symmetric systems. In this subsection we provide some known results on symmetric
systems of differential equations.
Let H and bH be finite-dimensional Hilbert spaces and let
(3.1)
In the following we put
(3.2)
ν+ := dim H,
H0 = H ⊕ bH, H = H0 ⊕ H = H ⊕ bH ⊕ H.
bν := dim bH,
ν− := dim H0 = ν+ +bν, n := dim H = ν+ + ν−.
Let as above I = [a, bi (−∞ < a < b ≤ ∞) be an interval in R . Moreover, let B(·) and
∆(·) be [H]-valued Borel functions on I integrable on each compact interval [a, β] ⊂ I and
satisfying B(t) = B∗(t) and ∆(t) ≥ 0 a.e. on I and let J ∈ [H] be operator (1.3).
A first-order symmetric system on an interval I (with the regular endpoint a) is a system
of differential equations of the form
(3.3)
J y′(t) − B(t)y(t) = ∆(t)f (t),
t ∈ I,
where f (·) ∈ L2
∆(I). Together with (3.3) we consider also the homogeneous system
(3.4)
J y′(t) − B(t)y(t) = λ∆(t)y(t),
t ∈ I,
λ ∈ C.
A function y ∈ AC(I; H) is a solution of (3.3) (resp. (3.4)) if equality (3.3) (resp. (3.4)
holds a.e. on I. Moreover, a function Y (·, λ) : I → [K, H] is an operator solution of equation
(3.4) if y(t) = Y (t, λ)h is a (vector) solution of this equation for every h ∈ K (here K is a
Hilbert space with dim K < ∞).
In what follows we always assume that system (3.3) is definite in the sense of the following
definition.
Definition 3.1.
each solution y of (3.4) the equality ∆(t)y(t) = 0 (a.e. on I) implies y(t) = 0, t ∈ I.
[12, 24] Symmetric system (3.3) is called definite if for each λ ∈ C and
As it is known [36, 20, 28] symmetric system (3.3) gives rise to the maximal linear relations
Tmax and Tmax in L2
∆(I) and L2
Tmax = {{y, f } ∈ (L2
∆(I), respectively. They are given by
∆(I))2 : y ∈ AC(I; H) and
(3.5)
J y′(t) − B(t)y(t) = ∆(t)f (t) a.e. on I}
and Tmax = {{πy, πf } : {y, f } ∈ Tmax}. Moreover the Lagrange's identity
(3.6)
(f, z)∆ − (y, g)∆ = [y, z]b − (J y(a), z(a)),
{y, f }, {z, g} ∈ Tmax.
holds with
(3.7)
[y, z]b := lim
t↑b
(J y(t), z(t)),
y, z ∈ dom Tmax.
Formula (3.7) defines the skew-Hermitian bilinear form [·, ·]b on dom Tmax, which plays a
crucial role in our considerations. By using this form we define the minimal relations Tmin
in L2
∆(I) and Tmin in L2
∆(I) via
Tmin = {{y, f } ∈ Tmax : y(a) = 0 and [y, z]b = 0 for each z ∈ dom Tmax}.
and Tmin = {{πy, πf } : {y, f } ∈ Tmin}. According to [36, 20, 28] Tmin is a closed symmetric
linear relation in L2
∆(I) and T ∗
min = Tmax.
12
VADIM MOGILEVSKII
[28]) that the maximal relation Tmax induced by the
Remark 3.2. It is known (see e.g.
definite symmetric system (3.3) possesses the following property:
there exists a unique function y ∈ AC(I; H) ∩ L2
for any f ∈ ef . Below we associate such a function y ∈ AC(I; H) ∩ L2
{ey,ef} ∈ Tmax.
∆(I). Definition (3.5) of Tmax implies
(3.4) belonging to L2
For any λ ∈ C denote by Nλ the linear space of solutions of the homogeneous system
for any {ey,ef } ∈ Tmax
∆(I) such that y ∈ ey and {y, f } ∈ Tmax
∆(I) with each pair
Nλ = ker (Tmax − λ) = {y ∈ L2
∆(I) : {y, λy} ∈ Tmax}, λ ∈ C,
and hence Nλ ⊂ dom Tmax. As usual, denote by n±(Tmin) := dim Nλ(Tmin),
λ ∈ C±,
the deficiency indices of Tmin. Since the system (3.3) is definite, πNλ = Nλ(Tmin) and
ker (π ↾ Nλ) = {0}, λ ∈ C. This implies that dim Nλ = n±(Tmin), λ ∈ C±.
The following lemma is obvious.
Lemma 3.3. If Y (·, λ) ∈ L2
(3.8)
∆[K, H] is an operator solution of Eq. (3.4), then the relation
K ∋ h → (Y (λ)h)(t) = Y (t, λ)h ∈ Nλ.
defines the linear mapping Y (λ) : K → Nλ and, conversely, for each such a mapping Y (λ)
there exists a unique operator solution Y (·, λ) ∈ L2
∆[K, H] of Eq. (3.4) such that (3.8) holds.
Next assume that
(3.9)
U =(cid:18)u1 u2 u3
u4 u5 u6(cid:19) : H ⊕ bH ⊕ H → bH ⊕ H
is the operator satisfying the relations
(3.10)
(3.11)
(3.12)
iu2u∗
2 − u1u∗
3 + u3u∗
iu5u∗
1 = 0
1 = iI bH ,
5 + u6u∗
3 + u6u∗
2 − u4u∗
6 = 0
iu5u∗
4 − u4u∗
ran U = bH ⊕ H
: H ⊕ bH ⊕ H → H ⊕ bH ⊕ H,
One can prove that the operator (3.9) admits an extension to the J-unitary operator
(3.13)
(3.14)
u7 u8 u9
u1 u2 u3
u4 u5 u6
In view of (3.1) each function y ∈ AC(I; H) admits the representation
Using (3.13) and the representation (3.14) of y we introduce the linear mappings Γja :
i.e. the operator satisfying eU ∗JeU = J.
eU =
y(t) = {y0(t), by(t), y1(t)}(∈ H ⊕ bH ⊕ H),
AC(I; H) → H, j ∈ {0, 1}, andbΓa : AC(I; H) → bH by setting
Γ0ay = u7y0(a) + u8by(a) + u9y1(a),
Clearly, the mappingsbΓa and Γ1a are determined by the operator U , while Γ0a is determined
by the extension eU . Moreover, the mapping
Γa :=(cid:16)Γ0a, bΓa, Γ1a(cid:17)⊤
bΓay = u1y0(a) + u2by(a) + u3y1(a), Γ1ay = u4y0(a) + u5by(a) + u6y1(a).
: AC(I; H) → H ⊕ bH ⊕ H
y ∈ AC(I; H)
(3.17)
(3.16)
(3.15)
t ∈ I.
ON EIGENFUNCTION EXPANSIONS
13
(3.18)
satisfies Γay = eU y(a), y ∈ AC(I; H). Hence Γa is surjective and
(J y(a), z(a)) = −(Γ1ay, Γ0az) + (Γ0ay, Γ1az) + i(bΓay,bΓaz),
ϕ(·, λ) = ϕU (·, λ)(∈ [H0, H]), λ ∈ C, of Eq. (3.4) with the initial data
In what follows we associate with each operator U (see (3.9)) the operator solution
y, z ∈ AC(I; H).
(3.19)
(3.20)
(3.21)
.
H
{z
sin B 0 − cos B
0
I bH
cos B 0
0
0
I bH
cos B 0
u∗
iu∗
6
3
−iu∗
u∗
5
2
−u∗
4 −iu∗
1
ϕU (a, λ) =
where B = B∗ ∈ [H]. For such U the solution ϕU (·, λ) is defined by the initial data
U =(cid:18) 0
ϕU (a, λ) =
: H ⊕ bH
→ H ⊕ bH ⊕ H
}
{z }H0
One can easily verify that for each J-unitary extension eU of U one has
eU ϕU (a, λ) =(cid:18)IH0
0 (cid:19) : H0 → H0 ⊕ H.
The particular case of the operator U and its J-unitary extension eU is (cf. [17])
eU =
sin B ,
sin B(cid:19) ,
− cos B 0 : H ⊕ bH → H ⊕ bH ⊕ H.
n−(Tmin) = ν− + νb− = ν+ +bν + νb−
Lemma 3.4. For any pair of finite-dimensional Hilbert spaces Hb and bHb with dim Hb =
min{νb+, νb−} and dim bHb = νb+ − νb− there exists a surjective linear mapping
[y, z]b = i · sign (νb+ − νb−)(bΓby,bΓbz) − (Γ1by, Γ0bz) + (Γ0by, Γ1bz).
3.3. Decomposing boundary triplets. According to [4, 34] the skew-Hermitian bilinear
form (3.7) has finite indices of inertia νb+ and νb− and
Γb = (Γ0b,bΓb, Γ1b)⊤ : dom Tmax → Hb ⊕ bHb ⊕ Hb
Note that Γby is in fact a singular boundary value of a function y ∈ dom Tmax in the sense
(for ν± see (3.2)). Moreover, the following lemma is immediate from [34, Lemma 5.1].
such that for all y, z ∈ dom Tmax the following equality is valid
(3.22)
n+(Tmin) = ν+ + νb+,
sin B
0
0
I bH
(3.23)
(3.24)
of [8, Chapter 13.2] (for more details see Remark 3.5 in [2]).
the following two alternative cases may hold:
Assume that n+(Tmin) < n−(Tmin). Then by (3.22) and (3.2) bν > νb+ − νb− and hence
Case 1 . bν > νb+ − νb− > 0.
Case 2 . bν ≥ 0 ≥ νb+ − νb− andbν 6= νb+ − νb−(6= 0).
In Case 1 one has dim bH(= bν) > νb+ − νb− > 0. Let bH1 be a subspace in bH with
dim bH1 = νb+ − νb− and let Hb be a Hilbert space with dim Hb = νb−. Then by Lemma 3.4
Below we construct a boundary triplet for Tmax separately in each of these cases.
there exists a surjective linear mapping
(3.25)
Γb = (Γ0b,bΓb, Γ1b)⊤ : dom Tmax → Hb ⊕ bH1 ⊕ Hb
14
VADIM MOGILEVSKII
(3.26)
such that (3.24) holds for all y, z ∈ dom Tmax. Let bH be decomposed as bH = bH1 ⊕ bH2 with
bH2 := bH ⊖ bH1 and let
0 := H ⊕ bH1, so
be the block representation of the mappingbΓa (see (3.16)). Moreover, let H ′
bΓa = (bΓa1, bΓa2)⊤ : AC(I; H) → bH1 ⊕ bH2
that in view of (3.1) H0 admits the representation
In Case 1 we let
(3.27)
(3.28)
(3.29)
(3.30)
0
H0 = H ⊕ bH1
0 ⊕ bH2
⊕bH2 = H ′
{z }H ′
0 ⊕ bH2 ⊕ Hb, H1 = H ′
2 (bΓa1 +bΓb)y, −Γ1by}(∈ H ′
Γ0{ey,ef} = {−Γ1ay + i(bΓa1 −bΓb)y, ibΓa2y, Γ0by}(∈ H ′
Γ1{ey,ef} = {Γ0ay + 1
H0 = H ′
0 ⊕ Hb,
0 ⊕ Hb),
0 ⊕ bH2 ⊕ Hb),
{ey,ef } ∈ Tmax.
Now assume that Case 2 holds. Let Hb and bHb be Hilbert spaces with dim Hb = νb+
and dim bHb = νb− − νb+. Then by Lemma 3.4 there is a surjective linear mapping (3.23)
satisfying (3.24). Let eHb := Hb ⊕ bHb (so that Hb ⊂ eHb) and let eΓ0b : dom Tmax → eHb be
the linear mapping given by
eΓ0b = Γ0b +bΓb.
(3.31)
In Case 2 we put
(3.32)
(3.33)
(3.34)
H1 = H ⊕ Hb
H0 = H ⊕ bH ⊕ eHb,
Γ0{ey,ef } = {−Γ1ay, ibΓay,eΓ0by}(∈ H ⊕ bH ⊕ eHb)
{ey,ef} ∈ Tmax.
Γ1{ey,ef} = {Γ0ay, −Γ1by}(∈ H ⊕ Hb),
Note that in both Cases 1 and 2 H1 is a subspace in H0 and Γj is an operator from Tmax
to Hj, j ∈ {0.1}. Moreover, H2(= H0 ⊖ H1) = bH2 in Case 1 and H2 = bH ⊕ bHb in Case 2.
Proposition 3.5. Assume that eU is J-unitary operator (3.13) and Γ0a, Γ1a and bΓa are
linear mappings (3.15), (3.16). Moreover, let Γb be surjective linear mapping given either
by(3.25) (in Case 1) or (3.23) (in Case 2) and satisfying (3.24). Then a collection Π− =
{H0 ⊕ H1, Γ0, Γ1} defined either by (3.28) -- (3.30) (in Case 1) or (3.32) -- (3.34) (in Case 2)
is a boundary triplet for Tmax.
Proof. The immediate calculation with taking (3.18) and (3.24) into account gives
(Γ1{ey,ef}, Γ0{ez,eg}) − (Γ0{ey,ef }, Γ1{ez,eg}) − i(P2Γ0{ey,ef }, P2Γ0{ez,eg}) =
= [y, z]b − (J y(a), z(a)),
This and the Lagrange's identity (3.6) yield identity (2.5) for Γ0 and Γ1. Moreover, the
mapping Γ = (Γ0, Γ1)⊤ is surjective, because so are Γa (see (3.17)) and Γb.
(cid:3)
{ey,ef}, {ez,eg} ∈ Tmax.
Definition 3.6. The boundary triplet Π− = {H0 ⊕ H1, Γ0, Γ1} constructed in Proposition
3.5 is called a decomposing boundary triplet for Tmax.
ON EIGENFUNCTION EXPANSIONS
15
Proposition 3.7. Let in Case 1 U be operator (3.9) and letbΓa and Γ1a be linear mappings
(3.16). Moreover, let bH1 be a subspace in bH, let bH2 = bH ⊖bH1, letbΓaj, j ∈ {1, 2}, be defined
by (3.26) and let Γb be surjective linear mapping (3.25) satisfying (3.24). Then:
(1) The equalities
(3.35)
(3.36)
T = {{ey,ef} ∈ Tmax : Γ1ay = 0, bΓa1y =bΓby, bΓa2y = 0, Γ0by = Γ1by = 0}
T ∗ = {{ey,ef} ∈ Tmax : Γ1ay = 0, bΓa1y =bΓby}
define a symmetric extension T of Tmin and its adjoint T ∗. Moreover, the deficiency indices
(3.38)
Γ1{ey,ef} = −Γ1by,
(3.37)
is a boundary triplet for T ∗ and the (maximal symmetric) relation A0(= ker Γ0) is
of T are n+(T ) = νb− and n−(T ) =bν + 2νb− − νb+.
(2) The collection Π− = { H0 ⊕ Hb, Γ0, Γ1} with H0 = bH2 ⊕ Hb and the operators
Γ0{ey,ef} = {ibΓa2y, Γ0by}(∈ bH2 ⊕ Hb),
{ey,ef } ∈ T ∗,
A0 = {{ey,ef } ∈ Tmax : Γ1ay = 0, bΓa1y =bΓby, bΓa2y = 0, Γ0by = 0}.
{H0 ⊕ H1, Γ0, Γ1} be the decomposing boundary triplet (3.28) -- (3.30) for Tmax. Applying to
this triplet Proposition 2.7 one obtains the desired statements.
(cid:3)
Proof. Let eU be J-unitary extension (3.13) of U , let Γ0a be operator (3.15) and let Π− =
Proposition 3.8. Let in Case 2 U be operator (3.9) and letbΓa and Γ1a be linear mappings
(3.16). Moreover, let Γb be surjective linear mapping (3.23) satisfying (3.24), let eHb =
Hb ⊕ bHb and let eΓ0b be given by (3.31). Then:
(1) The equalities
(3.39)
(3.40)
T = {{ey,ef } ∈ Tmax : Γ1ay = 0, bΓay = 0, eΓ0by = Γ1by = 0}
T ∗ = {{ey,ef} ∈ Tmax : Γ1ay = 0}
define a symmetric extension T of Tmin and its adjoint T ∗. Moreover, n+(T ) = νb+ and
n−(T ) =bν + νb−.
(2) The collection Π− = { H0 ⊕ Hb, Γ0, Γ1} with H0 = bH ⊕ eHb and the operators
{ey,ef} ∈ T ∗
(3.41)
is a boundary triplet for T ∗ and the (maximal symmetric) relation A0(= ker Γ0) is
Γ0{ey,ef} = {ibΓay,eΓ0by}(∈ bH ⊕ eHb),
Γ1{ey,ef} = −Γ1by,
A0 = {{ey,ef} ∈ Tmax : Γ1ay = 0, bΓay = 0, eΓ0by = 0}.
We omit the proof of this proposition, because it is similar to that of Proposition 3.7.
(3.42)
4. L2
∆-solutions of boundary value problems
4.1. Basic assumption. In what follows we suppose (unless otherwise stated) that system
(3.3) satisfies n−(Tmin) < n+(Tmin) and the following assumptions are fulfilled:
mappings (3.16).
(A1) U is the operator (3.9) satisfying (3.10) - (3.12) and bΓa and Γ1a are the linear
(A2) In Case 1 bH is decomposed as bH = bH1 ⊕ bH2,bΓaj : AC(I; H) → bHj, j ∈ {1, 2}, are
the operators given by (3.26), Hb is a finite-dimensional Hilbert space and Γb is a
surjective operator (3.25) satisfying (3.24).
16
VADIM MOGILEVSKII
(3.31).
(A3) In Case 2 Hb and bHb are finite-dimensional Hilbert spaces, Γb is a surjective operator
(3.23) satisfying (3.24), eHb = Hb ⊕ bHb and eΓ0b : dom Tmax → eHb is the operator
Definition 4.1. In Case 1 a boundary parameter is a collection τ = {τ+, τ−} ∈ eR−(bH2 ⊕
Hb, Hb). A truncated boundary parameter is a boundary parameter τ = {τ+, τ−} satisfying
τ−(λ) ⊂ ({0} ⊕ Hb) ⊕ Hb, λ ∈ C−.
4.2. Case 1.
According to Subsection 2.3 a boundary parameter τ = {τ+, τ−} admits the representation
(4.1)
τ+(λ) = {(C0(λ), C1(λ)); Hb}, λ ∈ C+;
{τ+, τ−} admits the representation (4.1) with
τ−(λ) = {(D0(λ), D1(λ)); bH2 ⊕ Hb}, λ ∈ C−
with holomorphic operator functions C0(λ)(∈ [bH2 ⊕ Hb, Hb]), C1(λ)(∈ [Hb]) and D0(λ)(∈
[bH2 ⊕ Hb]), D1(λ)(∈ [Hb, bH2 ⊕ Hb]). Moreover, a truncated boundary parameter τ =
D1(λ)(cid:19) : Hb → bH2 ⊕Hb.
(4.2) D0(λ) =(cid:18)I bH2
Let τ = {τ+, τ−} be a boundary parameter (4.1) and let
0
0 D0(λ)(cid:19) : bH2 ⊕Hb → bH2 ⊕Hb, D1(λ) =(cid:18) 0
C0(λ) = (bC02(λ), C0b(λ)) : bH2 ⊕ Hb → Hb,
D0(λ) = (bD02(λ), D0b(λ)) : bH2 ⊕ Hb → (bH2 ⊕ Hb)
be the block representations of C0(λ) and D0(λ). For a given function f ∈ L2
the following boundary value problem:
∆(I) consider
(4.3)
(4.4)
(4.5)
(4.6)
(4.7)
J y′ − B(t)y = λ∆(t)y + ∆(t)f (t),
t ∈ I,
Γ1ay = 0,
λ ∈ C \ R,
bΓa1y =bΓby,
ibC02(λ)bΓa2y + C0b(λ)Γ0by + C1(λ)Γ1by = 0, λ ∈ C+
ibD02(λ)bΓa2y + D0b(λ)Γ0by + D1(λ)Γ1by = 0, λ ∈ C−.
A function y(·, ·) : I × (C \ R) → H is called a solution of this problem if for each λ ∈ C \ R
the function y(·, λ) belongs to AC(I; H) ∩ L2
∆(I) and satisfies the equation (4.4) a.e. on I
(so that y ∈ dom Tmax) and the boundary conditions (4.5) -- (4.7).
Theorem 4.2. Let in Case 1 T be a symmetric relation in L2
τ = {τ+, τ−} is a boundary parameter (4.1), then for every f ∈ L2
(4.4) - (4.7) has a unique solution y(t, λ) = yf (t, λ) and the equality
∆(I) defined by (3.35). If
∆(I) the boundary problem
R(λ)ef = π(yf (·, λ)),
ef ∈ L2
∆(I),
f ∈ ef ,
defines a generalized resolvent R(λ) =: Rτ (λ) of T . Conversely, for each generalized resol-
vent R(λ) of T there exists a unique boundary parameter τ such that R(λ) = Rτ (λ).
λ ∈ C \ R,
Proof. Let Π− be the boundary triplet for T ∗ defined in Proposition 3.7. Applying to this
triplet [35, Theorem 3.11] we obtain the required statements.
(cid:3)
Remark 4.3. Let τ0 = {τ+, τ−} be a boundary parameter (4.1) with
(4.8)
C0(λ) ≡ PHb, C1(λ) ≡ 0, D0(λ) ≡ I bH2⊕Hb
, D1(λ) ≡ 0,
ON EIGENFUNCTION EXPANSIONS
17
and let A0 be a symmetric relation (3.38). Then
(4.9)
Rτ0(λ) = (A∗
0 − λ)−1, λ ∈ C+ and Rτ0 (λ) = (A0 − λ)−1, λ ∈ C−.
Proposition 4.4. Let in Case 1 PH , P bH1
and P bH2
and bH2 respectively (see (3.27)) and let eP bH2
bH2 (resp. Hb). Then:
(3.4) satisfying
be the orthoprojectors in H0 onto H, bH1
(PHb ) be the orthoprojector in bH2 ⊕ Hb onto
∆[H0, H] of Eq.
(1) For any λ ∈ C \ R there exists a unique operator solution v0(·, λ) ∈ L2
Γ1av0(λ) = −PH ,
, Γ0bv0(λ) = 0, λ ∈ C \ R
(2) For any λ ∈ C+ (λ ∈ C−) there exists a unique operator solution u+(·, λ) ∈ L2
∆[Hb, H]
(resp. u−(·, λ) ∈ L2
,
λ ∈ C−
i(bΓa1 −bΓb)v0(λ) = P bH1
ibΓa2v0(λ) = P bH2
∆[bH2 ⊕ Hb, H]) of Eq. (3.4) such that
i(bΓa1 −bΓb)u±(λ) = 0,
ibΓa2u−(λ) = eP bH2
Γ0bu+(λ) = IHb , λ ∈ C+
Γ0bu−(λ) = PHb ,
Γ1au±(λ) = 0,
,
λ ∈ C±
λ ∈ C−.
(4.10)
(4.11)
(4.12)
(4.13)
(4.14)
(4.16)
(4.18)
(4.19)
In formulas (4.10) -- (4.14) v0(λ) and u±(λ) are linear mappings from Lemma 3.3 corre-
sponding to the solutions v0(·, λ) and u±(·, λ) respectively.
Proof. Let eU be the J-unitary extension (3.13) of U , let Γ0a be the operator (3.15) and let
Π− = {H0⊕H1, Γ0, Γ1} be the decomposing boundary triplet (3.28)-(3.30) for Tmax. Assume
also that γ±(·) are the γ-fields of Π−. Since the quotient mapping π isomorphically maps
Nλ onto Nλ(Tmin), it follows that for every λ ∈ C+ (λ ∈ C−) there exists an isomorphism
Z+(λ) : H1 → Nλ (resp. Z−(λ) : H0 → Nλ) such that
(4.15)
γ+(λ) = πZ+(λ), λ ∈ C+;
γ−(λ) = πZ−(λ), λ ∈ C−.
Let Γ′
0 and Γ′
1 be the linear mappings given by
Γ′
Γ′
0 =
−Γ1a + i(bΓa1 −bΓb)
ibΓa2
1 =(cid:18)Γ0a + 1
2 (bΓa1 +bΓb)
−Γ1b
Γ0b
: dom Tmax → H ′
(cid:19) : dom Tmax → H ′
0 ⊕ bH2 ⊕ Hb,
0 ⊕ Hb.
Then by (3.29) and (3.30) one has Γj{πy, λπy} = Γ′
this equality with (4.15) and (2.13) gives
jy, y ∈ Nλ, j ∈ {0, 1}. Combining of
(4.17)
PH ′
0⊕HbΓ′
0Z+(λ) = IH1 ,
λ ∈ C+;
Γ′
0Z−(λ) = IH0 ,
λ ∈ C−,
which in view of (4.16) can be written as
Γ0b
(cid:18)−Γ1a + i(bΓa1 −bΓb)
−Γ1a + i(bΓa1 −bΓb)
(cid:19) Z+(λ) =(cid:18)IH ′
Z−(λ) =
ibΓa2
IH ′
0
0
Γ0b
0
0
0
0
IHb(cid:19) , λ ∈ C+
, λ ∈ C−
0
0
IHb
0
I bH2
0
18
VADIM MOGILEVSKII
It follows from (4.18) and (4.19) that
(4.20)
Γ1aZ+(λ) = (−PH , 0),
Γ1aZ−(λ) = (−PH , 0, 0),
bΓa2Z−(λ) = (0, −iI bH2
Next assume that
2 P bH1
, 0), Γ0bZ+(λ) = (0, IHb )
1
2 (bΓa1 −bΓb)Z+(λ) = (− i
1
2 P bH1
, 0), Γ0bZ−(λ) = (0, 0, IHb )
2 (bΓa1 −bΓb)Z−(λ) = (− i
, 0, 0)
are the block representations of Z±(λ) and let
Z+(λ) = (r(λ), u+(λ)) : H ′
0 ⊕ Hb → Nλ, λ ∈ C+
v0(λ) := (r(λ), 0) : H ′
v0(λ) := (r(λ), ω−(λ)) : H ′
Z−(λ) = (r(λ), ω−(λ), eu−(λ)) : H ′
0 ⊕ bH2 ⊕ Hb → Nλ, λ ∈ C−
0 ⊕ bH2 → Nλ, λ ∈ C+
0 ⊕ bH2 → Nλ, λ ∈ C−
u−(λ) = (ω−(λ), eu−(λ)) : bH2 ⊕ Hb → Nλ, λ ∈ C−.
∆[H0, H], u+(·, λ) ∈ L2
(4.21)
(4.22)
(4.23)
(4.24)
(4.25)
(4.26)
(4.27)
(4.28)
(4.29)
where
(4.30)
(4.31)
(4.32)
(4.33)
Then in view of (3.27) the equalities (4.25) and (4.26) define the operator v0(λ) ∈ [H0, Nλ], λ ∈
C \ R. Moreover, (4.23) induces the operator u+(λ) ∈ [Hb, Nλ]. Introduce also the operator
Now assume that v0(·, λ) ∈ L2
Hb, H] are the operator solutions of Eq. (3.4) corresponding to v0(λ), u+(λ) and u−(λ) in
accordance with Lemma 3.3. Then combining of (4.23) -- (4.27) with (4.20) -- (4.22) yields the
relations (4.10) -- (4.14) for v0(·, λ) and u±(·, λ). Finally, by using uniqueness of the solution
of the boundary value problem (4.4) -- (4.7) (with f = 0) one proves uniqueness of v0(·, λ)
and u±(·, λ) in the same way as in [2, Proposition 4.3].
(cid:3)
∆[Hb, H] and u−(·, λ) ∈ L2
∆[bH2 ⊕
the solutions v0(·, λ) and u±(·, λ) we define the operator functions
Let eU be a J-unitary extension (3.13) of U and let Γ0a be the mapping (3.15). By using
Ψ+(λ)
X+(λ) =(cid:18)m0(λ) Φ+(λ)
X−(λ) =(cid:18)m0(λ) Φ−(λ)
M+(λ)(cid:19) : H0 ⊕ Hb → H0 ⊕ (bH2 ⊕ Hb),
M−(λ)(cid:19) : H0 ⊕ (bH2 ⊕ Hb) → H0 ⊕ Hb,
Ψ−(λ)
λ ∈ C+
λ ∈ C−,
m0(λ) = (Γ0a +bΓa)v0(λ) + i
Φ±(λ) = (Γ0a +bΓa)u±(λ), λ ∈ C±
,
2 P bH , λ ∈ C \ R
Ψ+(λ) = (bΓa2 − Γ1b)v0(λ) + iP bH2
Ψ−(λ) = −Γ1bv0(λ),
M+(λ) = (bΓa2 − Γ1b)u+(λ), λ ∈ C+
M−(λ) = −Γ1bu−(λ), λ ∈ C−
In the following proposition we specify a connection between the operator functions X±(·)
and the Weyl functions M±(·) of the decomposing boundary triplet Π− for Tmax.
ON EIGENFUNCTION EXPANSIONS
19
Proposition 4.5. Let Π− = {H0 ⊕ H1, Γ0, Γ1} be the decomposing boundary triplet (3.28) --
(3.30) for Tmax and let
M1(λ) M2+(λ)
N1+(λ) N2+(λ)
M3+(λ) M4+(λ) : H ′
M+(λ) =
M−(λ) =(cid:18) M1(λ) N1−(λ) M2−(λ)
M3−(λ) N2−(λ) M4−(λ)(cid:19) : H ′
0 ⊕ Hb → H ′
0 ⊕ bH2 ⊕ Hb, λ ∈ C+
0 ⊕ Hb, λ ∈ C−
0 ⊕ bH2 ⊕ Hb → H ′
be the block representations of the corresponding Weyl functions M±(·). Then the entries
of the operator matrices (4.28) and (4.29) are holomorphic on their domains and have the
following block matrix representations:
(4.34)
(4.35)
(4.36)
(4.37)
(4.38)
(4.39)
(4.40)
i
0
2 I bH2(cid:19) : H ′
0 ⊕ bH2
{z
}H0
2 I bH2(cid:19) : H ′
0 ⊕ bH2
{z
}
H0
N1+(λ)
m0(λ) =(cid:18) M1(λ)
m0(λ) =(cid:18)M1(λ) N1−(λ)
Φ+(λ) =(cid:18)M2+(λ)
Ψ+(λ) =(cid:18)N1+(λ)
0(cid:19) ,
M3+(λ)
− i
iI
0
N2+(λ)(cid:19) , λ ∈ C+; Φ−(λ) =(cid:18)N1−(λ) M2−(λ)
0 (cid:19) ,
−iI
λ ∈ C+; Ψ−(λ) = (M3−(λ), N2−(λ)),
λ ∈ C−
λ ∈ C−
λ ∈ C+
λ ∈ C−
→ H ′
,
H0
0 ⊕ bH2
{z
}
0 ⊕ bH2
{z
}
H0
,
→ H ′
M+(λ) = (N2+(λ), M4+(λ))⊤, λ ∈ C+;
M−(λ) = (N2−(λ), M4−(λ)), λ ∈ C−.
Moreover, the following equalities hold
(4.41) m∗
0(λ) = m0(λ), Φ∗
+(λ) = Ψ−(λ), Ψ∗
+(λ) = Φ−(λ),
M ∗
+(λ) = M−(λ), λ ∈ C−.
Proof. Let Γ′
Then by (4.15) and (2.8), (2.9) one has
0 and Γ′
1 be given by (4.16) and let Z±(·) be the same as in Proposition 4.4.
(Γ′
1 − iP bH2
Γ′
0)Z+(λ) = M+(λ),
λ ∈ C+;
Γ′
1Z−(λ) = M−(λ),
λ ∈ C−,
which in view of (4.34) and (4.35) can be represented as
(cid:18)Γ0a + 1
Γ0a + 1
−Γ1b
2 (bΓa1 +bΓb)
bΓa2
2 (bΓa1 +bΓb)
−Γ1b
M1(λ) M2+(λ)
N1+(λ) N2+(λ)
Z+(λ) =
M3+(λ) M4+(λ) , λ ∈ C+
(cid:19) Z−(λ) =(cid:18) M1(λ) N1−(λ) M2−(λ)
M3−(λ) N2−(λ) M4−(λ)(cid:19) , λ ∈ C−.
2 (bΓa1 +bΓb)Z+(λ) = P bH1
1
bΓa2Z+(λ) = (N1+(λ), N2+(λ)), Γ1bZ+(λ) = (−M3+(λ), −M4+(λ))
Γ0aZ−(λ) = PH (M1(λ), N1−(λ), M2−(λ))
(M1(λ), N1−(λ), M2−(λ))
1
2 (bΓa1 +bΓb)Z−(λ) = P bH1
Γ1bZ−(λ) = (−M3−(λ), −N2−(λ), −M4−(λ)).
Γ0aZ+(λ) = PH (M1(λ), M2+(λ)),
(M1(λ), M2+(λ))
This implies that
(4.42)
(4.43)
(4.44)
(4.45)
(4.46)
20
VADIM MOGILEVSKII
Summing up the second equality in (4.20) with the equalities (4.42) and the first equality
in (4.43) one obtains
Similarly, summing up the second equality in (4.21), the first equality in (4.22) and the
equalities (4.44) and (4.45) one gets
, M2+(λ) + N2+(λ)) : H ′
0 ⊕ Hb → H0.
Combining (4.47) and (4.43) with the block representation (4.23) of Z+(λ) and taking defi-
nition (4.25) of v0(λ) into account we obtain
, N1−(λ) − iI bH2
, M2−(λ)).
)PH ′
0
,
λ ∈ C+
λ ∈ C+
,
bΓa2u+(λ) = N2+(λ),
, Γ1bu+(λ) = −M4+(λ),
λ ∈ C+
λ ∈ C+.
(4.49)
(4.48)
2 P bH1
2 P bH1
(4.47) (Γ0a +bΓa)Z+(λ) = (M1(λ) + N1+(λ) − i
(Γ0a +bΓa)Z−(λ) = (M1(λ) − i
(Γ0a +bΓa)v0(λ) = (M1(λ) + N1+(λ) − i
(Γ0a +bΓa)u+(λ) = M2+(λ) + N2+(λ),
bΓa2v0(λ) = N1+(λ)PH ′
(Γ0a +bΓa)v0(λ) = (M1(λ) − i
(Γ0a +bΓa)u−(λ) = (N1−(λ) − iI bH2
Γ1bv0(λ) = −M3+(λ)PH ′
2 P bH1
(4.50)
(4.51)
(4.52)
(4.53)
(4.54)
(4.55)
2 P bH1
0
0
Combining of (4.30) with (4.49) and (4.53) yields
Moreover, (4.48) and (4.46) together with (4.24), (4.26) and (4.27) give
Γ1bv0(λ) = −(M3−(λ), N2−(λ)), Γ1bu−(λ) = −(N2−(λ), M4−(λ)),
)PH ′
0 + (N1−(λ) − iI bH2
)P bH2
,
λ ∈ C−
, M2−(λ)) : bH2 ⊕ Hb → H0,
λ ∈ C−
λ ∈ C−.
m0(λ) = (M1(λ) + N1+(λ) − i
2 P bH1
)PH ′
0 + i
+ i
2 (P bH1
m0(λ) = (M1(λ) − i
− iP bH2
2 P bH1
− i
2 P bH1
+ i
) = (M1(λ) + N1+(λ))PH ′
+ i
0 + (N1−(λ) − iI bH2
+ P bH2
) = M1(λ)PH ′
2 P bH = (M1(λ) + N1+(λ))PH ′
0 + i
, λ ∈ C+
2 P bH = M1(λ)PH ′
2 P bH2
0 + N1−(λ)P bH2
2 P bH2
)P bH2
− i
+ P bH2
)PH ′
2 (P bH1
0 + N1−(λ)P bH2
, λ ∈ C−,
−
0 − i
2 P bH1
+
which is equivalent to the block representations (4.36) and (4.37) of m0(λ). Moreover,
(4.31) together with (4.50) and (4.54) gives the block representations (4.38) of Φ+(λ) and
Φ−(λ). Combining (4.32) and (4.33) with (4.51), (4.52) and (4.55) we arrive at the block
M±(λ). Finally, holomorphy of the operator
representations (4.39) of Ψ±(λ) and (4.40) of
functions m0(·), Φ±(·), Ψ±(·) and M±(·) and the equalities (4.41) are implied by (4.34) --
(4.40) and the equality M ∗
(cid:3)
+(λ) = M−(λ), λ ∈ C−.
Theorem 4.6. Let in Case 1 τ = {τ+, τ−} be a boundary parameter (4.1), (4.3). Then:
(1) For each λ ∈ C \ R there exists a unique operator solution vτ (·, λ) ∈ L2
∆[H0, H] of Eq.
(3.4) satisfying the boundary conditions
(4.56)
(4.57)
(4.58)
(4.59)
bC02(λ)(ibΓa2vτ (λ) − P bH2
bD02(λ)(ibΓa2vτ (λ) − P bH2
Γ1avτ (λ) = −PH ,
λ ∈ C \ R
i(bΓa1 −bΓb)vτ (λ) = P bH1
,
λ ∈ C \ R
) + C0b(λ)Γ0bvτ (λ) + C1(λ)Γ1bvτ (λ) = 0,
λ ∈ C+
) + D0b(λ)Γ0bvτ (λ) + D1(λ)Γ1bvτ (λ) = 0,
λ ∈ C−
(here vτ (λ) is the linear map from Lemma 3.3 corresponding to the solution vτ (·, λ)).
ON EIGENFUNCTION EXPANSIONS
21
(2) vτ (·, λ) is connected with the solutions v0(·, λ) and u±(·, λ) from Proposition 4.4 by
(4.60)
(4.61)
−(λ) + M+(λ))−1Ψ+(λ),
vτ (t, λ) = v0(t, λ) − u+(t, λ)(τ ∗
vτ (t, λ) = v0(t, λ) − u−(t, λ)(τ−(λ) + M−(λ))−1Ψ−(λ),
λ ∈ C+,
λ ∈ C−,
where Ψ±(λ) and M±(λ) are the operator functions defined in (4.32) and (4.33).
Proof. It follows from (4.40) and (4.35) that
Therefore by Proposition 2.7, (3) M−(λ) is the Weyl function of the boundary triplet Π− for
T ∗ and in view of [35, Theorem 3.11] one has 0 ∈ ρ(τ−(λ)+ M−(λ)), λ ∈ C−. Hence for each
λ ∈ C \ R the equalities (4.60) and (4.61) correctly define the solution vτ (·, λ) ∈ L2
∆[H0, H]
of Eq. (3.4) and to prove the theorem it is sufficient to show that such vτ (·, λ) is a unique
solution of (3.4) belonging to L2
M−(λ) = PHb M−(λ) ↾ bH2 ⊕ Hb, λ ∈ C−.
∆[H0, H] and satisfying (4.56) -- (4.59).
Combining (4.60) and (4.61) with (4.10) and (4.12) one gets the equalities (4.56) and
(4.57). To prove (4.58) and (4.59) we let
(4.62) T+(λ) = (τ ∗
−(λ) + M+(λ))−1, λ ∈ C+,
so that τ ∗
−(λ) and τ−(λ) can be written as
T−(λ) = (τ−(λ) + M−(λ))−1, λ ∈ C−,
−(λ)}.
M+(λ)T+(λ))h, (−PHb + PHb
M+(λ)T+(λ))h}},
This and (4.63) yield
(4.63)
(4.64)
τ ∗
τ+(λ) = {{−h1 + iP2h0, −P1h0} : {h1, h0} ∈ τ ∗
τ−(λ) = {{T−(λ)h, (I − M−(λ)T−(λ))h} : h ∈ Hb}.
−(λ) = {{T+(λ)h, (I − M+(λ)T+(λ))h} : h ∈ bH2 ⊕ Hb},
By using definition of the class eR−(H0, H1) in [35] one can easily show that
(4.65) τ+(λ) = {{(−T+(λ) + ieP bH2
where eP bH2
and PHb are the same as in Proposition 4.4 and h runs over bH2 ⊕ Hb.
bΓa2v0(λ) = eP bH2
bΓa2u+(λ) = eP bH2
(ibΓa2vτ (λ) − P bH2
Combining (4.60) with (4.66), (4.67), (4.13) and the last equality in (4.10) one gets
M+(λ)T+(λ))Ψ+(λ),
) + Γ0bvτ (λ) = (−T+(λ) + ieP bH2
Γ1bvτ (λ) = (−PHb + PHb
M+(λ), Γ1bu+(λ) = −PHb
M+(λ),
, Γ1bv0(λ) = −PHbΨ+(λ),
λ ∈ C+
M+(λ)T+(λ))Ψ+(λ),
It follows from (4.32) that
− ieP bH2
− ieP bH2
(4.66)
(4.67)
(4.68)
λ ∈ C+.
λ ∈ C+.
Ψ+(λ) − iP bH2
Moreover, (4.61) together with (4.11), (4.14), (4.33) and the last equality in (4.10) yields
(ibΓa2vτ (λ) − P bH2
(4.69)
) + Γ0bvτ (λ) = −eP bH2
Hence by (4.64) and (4.65) one has
T−(λ)Ψ−(λ) − PHb T−(λ)Ψ−(λ) = −T−(λ)Ψ−(λ),
Γ1bvτ (λ) = −(I − M−(λ)T−(λ))Ψ−(λ),
λ ∈ C−.
{(ibΓa2vτ (λ) − P bH2
)h0 + Γ0bvτ (λ)h0, Γ1bvτ (λ)h0} ∈ τ±(λ), h0 ∈ H0, λ ∈ C±,
which in view of (2.4) and the block representations (4.3) yields (4.58) and (4.59). Finally,
uniqueness of vτ (·, λ) is implied by uniqueness of the solution of the boundary value problem
(4.4) -- (4.7) (see Theorem 4.2).
(cid:3)
22
VADIM MOGILEVSKII
4.3. Case 2.
Definition 4.7. A boundary parameter in Case 2 is a collection τ = {τ+, τ−} ∈ eR−(bH ⊕
eHb, Hb). A truncated boundary parameter is a boundary parameter τ = {τ+, τ−} satisfying
τ−(λ) ⊂ ({0} ⊕ eHb) ⊕ Hb, λ ∈ C−.
According to Subsection 2.3 a boundary parameter τ = {τ+, τ−} admits the representation
(4.70)
admits the representation (4.70) with
τ+(λ) = {(C0(λ), C1(λ)); Hb}, λ ∈ C+;
τ−(λ) = {(D0(λ), D1(λ)); bH ⊕ eHb}, λ ∈ C−
with holomorphic operator functions C0(λ)(∈ [bH ⊕ eHb, Hb]), C1(λ)(∈ [Hb]) and D0(λ)(∈
[bH ⊕eHb]), D1(λ)(∈ [Hb, bH ⊕eHb]). Moreover, a truncated boundary parameter τ = {τ+, τ−}
D1(λ)(cid:19) : Hb → bH ⊕ eHb.
(4.71) D0(λ) =(cid:18)I bH
Let τ = {τ+, τ−} be a boundary parameter (4.70) and let
0 D0(λ)(cid:19) : bH ⊕ eHb → bH ⊕ eHb, D1(λ) =(cid:18) 0
C0(λ) = (bC0(λ), eC0b(λ)) : bH ⊕ eHb → Hb,
D0(λ) = (bD0(λ), eD0b(λ)) : bH ⊕ eHb → (bH ⊕ eHb)
be the block representations of C0(λ) and D0(λ). For a given function f ∈ L2
the following boundary value problem:
∆(I) consider
(4.72)
0
J y′ − B(t)y = λ∆(t)y + ∆(t)f (t),
t ∈ I,
Γ1ay = 0,
λ ∈ C \ R,
(4.73)
(4.74)
(4.75)
(4.76)
ibC0(λ)bΓay + eC0b(λ)eΓ0by + C1(λ)Γ1by = 0,
ibD0(λ)bΓay + eD0b(λ)eΓ0by + D1(λ)Γ1by = 0, λ ∈ C−.
λ ∈ C+,
One proves the following theorem in the same way as Theorem 4.2.
Theorem 4.8. Let in Case 2 T be a symmetric relation in L2
∆(I) defined by (3.39). Then
statements of Theorem 4.2 hold with a boundary parameter τ = {τ+, τ−} of the form (4.70)
and the boundary value problem (4.73) - (4.76) in place of (4.4) - (4.7).
Remark 4.9. Let τ0 = {τ+, τ−} be a boundary parameter (4.70) with
(4.77)
C0(λ) ≡ PHb (∈ [bH ⊕ eHb, Hb]), C1(λ) ≡ 0, D0(λ) ≡ I bH⊕ eHb
and let A0 be the symmetric relation (3.42). Then Rτ0(λ) is of the form (4.9).
, D1(λ) ≡ 0
Proposition 4.10. Let in Case 2 PH (resp. P bH ) be the orthoprojector in H0 onto H (resp.
bH) and let eP bH (resp. P eHb
(3.4) satisfying
(1) For any λ ∈ C \ R there exists a unique operator solution v0(·, λ) ∈ L2
) be the orthoprojector in bH ⊕ eHb onto bH (resp. eHb). Then:
∆[H0, H] of Eq.
(4.78)
(4.79)
Γ1av0(λ) = −PH, λ ∈ C \ R
Γ0bv0(λ) = 0, λ ∈ C+;
ibΓav0(λ) = P bH ,
eΓ0bv0(λ) = 0, λ ∈ C−
ON EIGENFUNCTION EXPANSIONS
23
(2) For any λ ∈ C+ (λ ∈ C−) there exists a unique operator solution u+(·, λ) ∈ L2
∆[Hb, H]
(resp. u−(·, λ) ∈ L2
∆[bH ⊕ eHb, H]) of Eq. (3.4) satisfying
ibΓau−(λ) = eP bH ,
Γ1au±(λ) = 0,
λ ∈ C±
Γ0bu+(λ) = IHb , λ ∈ C+;
Proof. let Π− be the decomposing boundary triplet for Tmax constructed with the aid of
some J-unitary extension eU of U (see Proposition 3.5). By using γ-fields of this triplet one
proves in the same way as in Proposition 4.4 the existence of isomorphisms Z+(λ) : H1 →
Nλ, λ ∈ C+, and Z−(λ) : H0 → Nλ, λ ∈ C−, such that
eΓ0bu−(λ) = P eHb
, λ ∈ C−.
−Γ1a
0
Γ0b (cid:19) Z+(λ) =(cid:18)IH
(cid:18)−Γ1a
Z−(λ) =
ibΓa
eΓ0b
IH
0
0
0
IHb(cid:19) ,
0
I bH
0
0
0
I eHb
λ ∈ C+
,
λ ∈ C−
Γ1aZ+(λ) = (−IH , 0),
Γ0bZ+(λ) = (0, IHb)
It follows from (4.82) and (4.83) that
(4.80)
(4.81)
(4.82)
(4.83)
(4.84)
(4.85)
(4.88)
(4.89)
(4.90)
Assume that
(4.86)
(4.87)
Γ1aZ−(λ) = (−IH , 0, 0),
ibΓaZ−(λ) = (0, I bH , 0),
eΓ0bZ−(λ) = (0, 0, I eHb
).
Z+(λ) = (r(λ), u+(λ)) : H ⊕ Hb → Nλ, λ ∈ C+
Z−(λ) = (r(λ), ω−(λ), eu−(λ)) : H ⊕ bH ⊕ eHb → Nλ, λ ∈ C−
are the block representations of Z±(λ) (see (3.32) ) and let v0(λ) : H0 → Nλ, λ ∈ C \ R, be
the operator given by
It is clear that (4.86) induces the operator u+(λ) ∈ [Hb, Nλ]. Introduce also the operator
v0(λ) := (r(λ), 0) : H ⊕ bH → Nλ, λ ∈ C+
v0(λ) := (r(λ), ω−(λ)) : H ⊕ bH → Nλ, λ ∈ C−.
u−(λ) = (ω−(λ), eu−(λ)) : bH ⊕ eHb → Nλ, λ ∈ C−.
∆[Hb, H] and u−(·, λ) ∈ L2
∆[H0, H], u+(·, λ) ∈ L2
Let v0(·, λ) ∈ L2
solutions of Eq. (3.4) corresponding to v0(λ), u+(λ) and u−(λ) respectively (see Lemma
3.3). Then combining of (4.86) -- (4.90) with (4.84) and (4.85) gives the relations (4.78) --
(4.81). Finally, uniqueness of v0(·, λ) and u±(·, λ) follows from uniqueness of the solution of
the boundary value problem (4.73) -- (4.76).
(cid:3)
∆[bH ⊕eHb, H] be the operator
the operator functions
Let eU be a J-unitary extension (3.13) of U and let Γ0a be the mapping (3.15). Introduce
Ψ+(λ)
X+(λ) =(cid:18)m0(λ) Φ+(λ)
X−(λ) =(cid:18)m0(λ) Φ−(λ)
M+(λ)(cid:19) : H0 ⊕ Hb → H0 ⊕ (bH ⊕ eHb),
M−(λ)(cid:19) : H0 ⊕ (bH ⊕ eHb) → H0 ⊕ Hb,
Ψ−(λ)
λ ∈ C+
λ ∈ C−,
(4.91)
(4.92)
24
VADIM MOGILEVSKII
with the entries defined in terms of the solutions v0(·, λ) and u±(·, λ) as follows: m0(λ), Φ±(λ),
Ψ−(λ) and M−(λ) are given by (4.30), (4.31) and (4.33), while
(4.93) Ψ+(λ) = (bΓa − Γ1b − ibΓb)v0(λ) + iP bH ,
M+(λ) = (bΓa − Γ1b − ibΓb)u+(λ), λ ∈ C+.
Proposition 4.11. Let Π− = {H0 ⊕H1, Γ0, Γ1} be the decomposing boundary triplet (3.32) --
(3.34) for Tmax and let
M1(λ) M2+(λ)
N1+(λ) N2+(λ)
M+(λ) =
M−(λ) =(cid:18) M1(λ) N1−(λ) M2−(λ)
M3+(λ) M4+(λ) : H ⊕ Hb → H ⊕ bH ⊕ eHb, λ ∈ C+
M3−(λ) N2−(λ) M4−(λ)(cid:19) : H ⊕ bH ⊕ eHb → H ⊕ Hb, λ ∈ C−
be the block matrix representations of the corresponding Weyl functions. Then the entries
of the operator matrices (4.91) and (4.92) are holomorphic on their domains and satisfy
(4.94)
(4.95)
(4.96)
(4.97)
N1+(λ)
m0(λ) =(cid:18) M1(λ)
m0(λ) =(cid:18)M1(λ) N1−(λ)
− i
0
i
0
2 I bH(cid:19) : H ⊕ bH
{z }H0
2 I bH (cid:19) : H ⊕ bH
{z }H0
,
→ H ⊕ bH
{z }H0
→ H ⊕ bH
{z }H0
λ ∈ C+
,
λ ∈ C−
and (4.38) -- (4.40). Moreover, the equalities (4.41) are valid.
Proof. Let Z±(·) be the same as in Proposition 4.10. By using the reasonings similar to
those in the proof of Proposition 4.5 one proves the equalities
(4.98)
(4.99)
Γ0a
Z+(λ) =
bΓa
−Γ1b − ibΓb
−Γ1b(cid:19) Z−(λ) =(cid:18) M1(λ) N1−(λ) M2−(λ)
M3−(λ) N2−(λ) M4−(λ)(cid:19) ,
(cid:18) Γ0a
M3+(λ) M4+(λ) , λ ∈ C+,
M1(λ) M2+(λ)
N1+(λ) N2+(λ)
λ ∈ C−.
It follows from (4.98) that
Moreover, (4.99) and the second equality in (4.85) yield
(Γ0a +bΓa)Z+(λ) = (M1(λ) + N1+(λ), M2+(λ) + N2+(λ)),
(bΓa − Γ1b − ibΓb)Z+(λ) = (N1+(λ) + M3+(λ), N2+(λ) + M4+(λ)),
(Γ0a +bΓa)Z−(λ) = (M1(λ), N1−(λ) − iI bH , M2−(λ)),
−Γ1bZ−(λ) = (M3−(λ), N2−(λ), M4−(λ)),
λ ∈ C−.
λ ∈ C+.
ON EIGENFUNCTION EXPANSIONS
25
Combining these equalities with the block representations (4.86) and (4.87) of Z±(λ) and
taking (4.88) -- (4.90) into account we obtain
(Γ0a +bΓa)v0(λ) = (M1(λ) + N1+(λ))PH ,
(Γ0a +bΓa)u+(λ) = M2+(λ) + N2+(λ), λ ∈ C+
(bΓa − Γ1b − ibΓb)v0(λ) = (N1+(λ) + M3+(λ))PH , λ ∈ C+
(bΓa − Γ1b − ibΓb)u+(λ) = N2+(λ) + M4+(λ), λ ∈ C+
(Γ0a +bΓa)v0(λ) = M1(λ)PH + N1−(λ)P bH − iP bH ,
(Γ0a +bΓa)u−(λ) = (N1−(λ) − iI bH , M2−(λ)) : bH ⊕ eHb → H0,
−Γ1bu−(λ) = (N2−(λ), M4−(λ)),
λ ∈ C−
−Γ1bv0(λ) = (M3−(λ), N2−(λ)),
This and definitions (4.30), (4.31), (4.33) and (4.93) of m0(·), Φ±(·), Ψ±(·) and M±(·) yield
the equalities (4.96), (4.97) and (4.38) -- (4.40). Finally one proves (4.41) in the same way as
in Proposition 4.5.
(cid:3)
λ ∈ C−
λ ∈ C−
Theorem 4.12. Let in Case 2 τ = {τ+, τ−} be a boundary parameter defined by (4.70) and
(4.72). Then:
(1) Statement (1) of Theorem 4.6 holds with the boundary condition (4.56) and the fol-
lowing boundary conditions in place of (4.57) -- (4.59):
(4.100)
(4.101)
bC0(λ)(ibΓavτ (λ) − P bH ) + eC0b(λ)eΓ0bvτ (λ) + C1(λ)Γ1bvτ (λ) = 0,
bD0(λ)(ibΓavτ (λ) − P bH ) + eD0b(λ)eΓ0bvτ (λ) + D1(λ)Γ1bvτ (λ) = 0,
(2) The solution vτ (·, λ) is of the form (4.60), (4.61), where v0(·, λ) and u±(·, λ) are
defined in Proposition 4.10 and Ψ±(λ) and M±(λ) are given by (4.33) and (4.93).
λ ∈ C+
λ ∈ C−
Proof. As in Theorem 4.6 one proves that (4.60) and (4.61) correctly define the solution
vτ (·, λ) ∈ L2
∆[H0, H] of Eq. (3.4). Therefore it remains to show that such vτ (·, λ) is a unique
solution of (3.4) belonging to L2
∆[H0, H] and satisfying (4.56), (4.100) and (4.101).
First observe that (4.78) and (4.80) yield (4.56). Next assume that T±(λ) are defined by
(4.62). Then in the same way as in Theorem 4.6 one gets (4.64) and the equality
(4.102)
τ+(λ) = {{(−T+(λ) + iP bH⊕ bHb
− iP bH⊕ bHb
M+(λ)T+(λ))h,
(−PHb + PHb
M+(λ)T+(λ))h} : h ∈ bH ⊕ eHb},
(4.103)
(4.104)
where P bH⊕ bHb
It follows from (4.93) that
is the orthoprojector in bH ⊕ eHb onto bH ⊕ bHb(= (bH ⊕ eHb) ⊖ Hb).
bΓav0(λ) = eP bH Ψ+(λ) − iP bH , Γ1bv0(λ) = −PHb Ψ+(λ), λ ∈ C+
bΓbu+(λ) = iP bHb
bΓbv0(λ) = iP bHb
Ψ+(λ), λ ∈ C+
M+(λ),
M+(λ), Γ1bu+(λ) = −PHb
(4.105)
bΓau+(λ) = eP bH
where eP bH , PHb and P bHb
Moreover, (4.104), the last equality in (4.105) and the first equalities in (4.79) and (4.81)
give
are the orthoprojectors in bH ⊕ eHb onto bH, Hb and bHb respectively.
M+(λ),
(4.106)
Now combining (4.60) with (4.103) -- (4.106) one gets (4.68) and the equality
eΓ0bv0(λ) = iP bHb
Ψ+(λ),
eΓ0bu+(λ) = IHb + iP bHb
M+(λ), λ ∈ C+.
− iP bH⊕ bHb
M+(λ)T+(λ))Ψ+(λ), λ ∈ C+.
(ibΓavτ (λ) − P bH ) +eΓ0bvτ (λ) = (−T+(λ) + iP bH⊕ bHb
26
VADIM MOGILEVSKII
Moreover, (4.61) together with (4.79), (4.81) and (4.33) leads to (4.69) and the equality
(ibΓavτ (λ)−P bH )+eΓ0bvτ (λ) = −eP bHT−(λ)Ψ−(λ)−P eHb
Therefore in view of (4.64) and (4.102) one has
T−(λ)Ψ−(λ) = −T−(λ)Ψ−(λ), λ ∈ C−.
{(ibΓavτ (λ) − P bH )h0 +eΓ0bvτ (λ)h0, Γ1bvτ (λ)h0} ∈ τ±(λ),
which implies (4.100) and (4.101). Finally, uniqueness of vτ (·, λ) follows from uniqueness of
the solution of the boundary value problem (4.73) -- (4.76).
(cid:3)
h0 ∈ H0, λ ∈ C±,
5. m-functions and eigenfunction expansions
U and Γ0a is the mapping (3.15).
5.1. m-functions. In this subsection we assume that eU is a J-unitary extension (3.13) of
∆[H0, H] be the operator solution of
Eq. (3.4) defined in Theorems 4.6 and 4.12. Similarly to the case n−(Tmin) ≤ n+(Tmin) (see
[2]) we introduce the following definition.
Let τ be a boundary parameter and let vτ (·, λ) ∈ L2
Definition 5.1. The operator function mτ (·) : C \ R → [H0] defined by
(5.1)
is called the m-function corresponding to the boundary parameter τ or, equivalently, to the
boundary value problem (4.4) -- (4.7) (in Case 1 ) or (4.73) -- (4.76) (in Case 2 ).
mτ (λ) = (Γ0a +bΓa)vτ (λ) + i
2 P bH ,
λ ∈ C \ R,
From (4.56) it follows from that mτ (·) satisfies the equality
(5.2) eU vτ (a, λ)(cid:18)=(cid:18)Γ0a +bΓa
Γ1a (cid:19) vτ (λ)(cid:19) =(cid:18)mτ (λ) − i
−PH
2 P bH
(cid:19) : H0 → H0 ⊕ H,
It is easily seen that Proposition 5.3 in [2] remains valid in the case n+(Tmin) < n−(Tmin).
This means that for a given operator U (see (3.9)) and a boundary parameter τ the m-
functions m(1)
connected by m(1)
τ (·) corresponding to J-unitary extensions eU1 and eU2 of U are
A definition of the m-function mτ in somewhat other terms ia given in the following
τ (λ) + B, λ ∈ C \ R, with some B = B∗ ∈ [H0].
τ (·) and m(2)
τ (λ) = m(2)
λ ∈ C \ R.
proposition, which directly follows from Theorems 4.6 and 4.12.
Proposition 5.2. Let τ = {τ+, τ−} be a boundary parameter (4.1), (4.3) in Case 1 (resp.
(4.70), (4.72) in Case 2), let ϕU (·, λ)(∈ [H0, H]), λ ∈ C, be the operator solution of Eq.
(3.4) defined by (3.19) and let ψ(·, λ)(∈ [H0, H]), λ ∈ C, be the operator solutions of Eq.
(3.4) with
eU ψ(a, λ) =(cid:18)− i
2 P bH
−PH (cid:19) : H0 → H0 ⊕ H.
Then there exists a unique operator function m(·) = mτ (·) : C \ R → [H0] such that for any
λ ∈ C \ R the operator solution v(·, λ) = vτ (·, λ) of Eq. (3.4) given by
(5.3)
v(t, λ) := ϕU (t, λ)m(λ) + ψ(t, λ)
belongs to L2
(4.100), (4.101) in Case 2.
∆[H0, H] and satisfies the boundary conditions (4.57) -- (4.59) in Case 1 and
In the following theorem we give a description of all m-functions immediately in terms of
the boundary parameter τ .
ON EIGENFUNCTION EXPANSIONS
27
Theorem 5.3. Assume the following hypothesis:
(i) τ0 is a boundary parameter from Remark 4.3 in Case 1 and Remark 4.9 in Case 2.
(ii) m0(·), Φ−(·), Ψ−(·) and M−(·) are the operator functions given by (4.30), (4.31) and
(4.33) (this functions form the operator matrix X−(·) given by (4.29) in Case 1 and (4.92)
in Case 2).
(iii) M1(·) and Mj−(·), j ∈ {2, 3, 4}, are the operator functions defined by the block repre-
sentations (4.35) (in Case 1) and (4.95)(in Case 2) of the Weyl function M−(·) (according
to Propositions 4.5 and 4.11 these functions can be also defined as the entries of the block
matrix representations (4.36) -- (4.40) of m0(·), Φ−(·), Ψ−(·) and M−(·)).
Then: (1) m0(λ) = mτ0(λ), λ ∈ C \ R, and for any boundary parameter τ = {τ+, τ−}
defined by (4.1) in Case 1 and (4.70) in Case 2 the corresponding m-function mτ (·) is
(5.4)
mτ (λ) = m0(λ) + Φ−(λ)(D0(λ) − D1(λ) M−(λ))−1D1(λ)Ψ−(λ),
λ ∈ C−.
(2) For each truncated boundary parameter τ = {τ+, τ−} defined by (4.1), (4.2) in Case 1
and (4.70), (4.71) in Case 2 the corresponding m-function mτ (·) has the triangular block
matrix representation
(5.5)
mτ (λ) =(cid:18)mτ,1(λ) mτ,2(λ)
2 I (cid:19) ,
− i
0
λ ∈ C−,
with respect to the decomposition H0 = H ′
Case 2. Moreover, the operator function mτ,1(·) in (5.5) is
(5.6) mτ,1(λ) = M1(λ) + M2−(λ)(D0(λ) − D1(λ)M4−(λ))−1D1(λ)M3−(λ),
0 ⊕ bH2 in Case 1 (see (3.27)) and H0 = H ⊕ bH in
λ ∈ C−.
Proof. We prove the theorem only for Case 1, because in Case 2 the proof is similar.
(1) It is easily seen that v0(t, λ) = vτ0 (t, λ), where v0(·, λ) is defined in Proposition 4.4.
Hence by (4.30) one has m0(λ) = mτ0(λ). Next, applying the operator Γ0a +bΓa to the
equalities (4.60) and (4.61) with taking (4.30) and (4.31) into account one gets
−(λ) + M+(λ))−1Ψ+(λ),
mτ (λ) = m0(λ) − Φ+(λ)(τ ∗
mτ (λ) = m0(λ) − Φ−(λ)(τ−(λ) + M−(λ))−1S−(λ),
λ ∈ C+,
λ ∈ C−.
(5.7)
(5.8)
Moreover, according to [30, Lemma 2.1] 0 ∈ ρ(D0(λ) − D1(λ) M−(λ)) and
−(τ−(λ) + M−(λ))−1 = (D0(λ) − D1(λ) M−(λ))−1D1(λ),
λ ∈ C−,
which together with (5.8) yields (5.4).
(2) It follows from (4.2) and the second equality in (4.40) that (D0(λ)−D1(λ) M−(λ))−1 =
−D1(λ)N2−(λ) D0(λ) − D1(λ)M4−(λ)(cid:19)−1
(cid:18)
0
I
=(cid:18)I
∗
(D0(λ) − D1(λ)M4−(λ))−1(cid:19)
0
(the entry ∗ does not matter). Combining this equality with (5.4) and taking (4.37) -- (4.39)
into account one gets the equalities (5.5) and (5.6).
(cid:3)
By using the reasonings similar to those in the proof of Proposition 5.7 in [2] one proves
the following proposition.
Proposition 5.4. The m-function mτ (·) is a Nevanlinna operator function satisfying
(Im λ)−1 · Im mτ (λ) ≥ZI
v∗
τ (t, λ)∆(t)vτ (t, λ) dt,
λ ∈ C−.
28
VADIM MOGILEVSKII
5.2. Green's function. In the sequel we put H := L2
∆(I) and denote by Hb the set of all
ef ∈ H with the following property: there exists β ef ∈ I such that for some (and hence for
all) function f ∈ ef the equality ∆(t)f (t) = 0 holds a.e. on (β ef , b).
Let ϕU (·, λ) be the operator-valued solution (3.19), let τ be a boundary parameter and
∆[H0, H] be the operator-valued solution of Eq. (3.4) defined in Theorems
let vτ (·, λ) ∈ L2
4.6 and 4.12.
Definition 5.5. The operator function Gτ (·, ·, λ) : I × I → [H] given by
Gτ (x, t, λ) =(vτ (x, λ) ϕ∗
ϕU (x, λ) v∗
U (t, λ), x > t
τ (t, λ), x < t
, λ ∈ C \ R
(5.9)
will be called the Green's function corresponding to the boundary parameter τ .
(5.10)
RI
Theorem 5.6. Let τ be a boundary parameter and let Rτ (·) be the corresponding generalized
resolvent of the relation T (see Theorems 4.2 and 4.8). Then
(5.11)
f ∈ ef .
∆(I) the inequality
ef ∈ H,
Gτ (x, t, λ)∆(t)f (t) dt < ∞ is valid. This implies that formula
Proof. As in [2, Theorem 6.2] one proves that for each f ∈ L2
Gτ (·, t, λ)∆(t)f (t) dt(cid:19) ,
Rτ (λ)ef = π(cid:18)ZI
yf = yf (x, λ) :=ZI
following statement: for each ef ∈ H
To prove (5.12) first assume that ef ∈ Hb. We show that the function yf (·, λ) given by (5.11)
is a solution of the boundary problem (4.4) -- (4.7) in Case 1 (resp. (4.73) -- (4.76) in Case 2
). As in [2, Theorem 6.2] one proves that yf (·, λ) belongs to AC(I; H) ∩ L2
∆(I) and satisfies
(4.4) a.e. on I. Now it remains to show that yf satisfies the boundary conditions (4.5) -- (4.7)
in Case 1 (resp. (4.74) -- (4.76) in Case 2 ).
∆(I) and Rτ (λ)ef = π(yf (·, λ)),
correctly defines the function yf (·, ·) : I × C \ R → H. Therefore (5.10) is equivalent to the
f ∈ ef , λ ∈ C \ R.
Gτ (x, t, λ)∆(t)f (t) dt,
yf (·, λ) ∈ L2
λ ∈ C \ R
(5.12)
It follows from (5.11) and (5.9) that
(5.13)
yf (a, λ) = ϕU (a, λ)ZI
v∗
τ (t, λ)∆(t)f (t) dt,
ϕ∗
(5.14)
x ∈ (β ef , b).
U (t, λ)∆(t)f (t) dt,
yf (x, λ) = vτ (x, λ)ZI
Let eU be a J-unitary extension (3.13) of U and let Γa be the operator (3.17). Then Γayf =
eU yf (a, λ) and in view of (5.13) one has
Γayf = eU ϕU (a, λ)ZI
bΓayf = P bHZI
v∗
τ (t, λ)∆(t)f (t) dt,
v∗
τ (t, λ)∆(t)f (t) dt.
This and (3.20) imply
Γ1ayf = 0.
(5.15)
ON EIGENFUNCTION EXPANSIONS
29
Next assume that
(5.16)
h ef :=ZI
ϕ∗
U (t, λ)∆(t)f (t) dt(∈ H0)
Then in view of (5.14) one has (see [2, Remark 3.5 (1)])
(5.17)
Let us also prove the equality
bΓbyf =bΓbvτ (λ)h ef , Γ0byf = Γ0bvτ (λ)h ef , Γ1byf = Γ1bvτ (λ)h ef .
(5.18)
It follows from (5.2) that
bΓayf = (bΓavτ (λ) + iP bH )h ef .
eU vτ (a, λ) ↾ bH =(cid:18)(mτ (λ) − i
0
2 IH0 ) ↾ bH
(cid:19) : bH → H0 ⊕ H,
λ ∈ C \ R.
Therefore by (3.20) one has vτ (t, λ) ↾ bH = ϕU (t, λ)(mτ (λ)− i
τ (λ) = mτ (λ). This implies that
to Proposition 5.4 m∗
2 IH0 ) ↾ bH. Moreover, according
P bH v∗
τ (t, λ) = P bH (mτ (λ) + i
2 IH0 )ϕ∗
U (t, λ) = (P bH mτ (λ) + i
2 P bH )ϕ∗
U (t, λ)
and (5.15), (5.16) yield
(5.19)
bΓayf = (P bH mτ (λ) + i
2 P bH )ZI
ϕ∗
U (t, λ)∆(t)f (t) dt = (P bH mτ (λ) + i
2 P bH )h ef .
Since in view of (5.1) P bH mτ (λ) =bΓavτ (λ) + i
Next assume Case 1 and the hypothesis (A2). Then by (3.26) and (5.18)
The second equality in (5.15) gives the first condition in (4.5) and the condition (4.74).
2 P bH , the equality (5.19) gives (5.18).
(5.20)
bΓa1yf = (bΓa1vτ (λ) + iP bH1
)h ef ,
bΓa2yf = (bΓa2vτ (λ) + iP bH2
)h ef
and combining of (5.20) and (5.17) with (4.57) -- (4.59) shows that the second condition in
(4.5) and the conditions (4.6) and (4.7) are fulfilled for yf .
fulfillment of the conditions (4.75) and (4.76) for yf .
Thus yf (·, λ) is a solution of the boundary value problem (4.4) -- (4.7) in Case 1 (resp.
Now assume Case 2 and the hypothesis (A3). Then by (3.31) and (5.17) one haseΓ0byf =
eΓ0bvτ (λ)h ef . Combining of this equality and (5.17), (5.18) with (4.100) and (4.101) implies
(4.73) -- (4.76) in Case 2 ) and by Theorems 4.2 and 4.8 relations (5.12) hold (for ef ∈ Hb).
Finally, one proves (5.12) for arbitrary ef ∈ H in the same way as in [2, Theorem 6.2].
Hamiltonian systems (i.e., for systems (3.3) with bH = {0}). Moreover, for non-Hamiltonian
systems (3.3) in the case of minimally possible deficiency indices n±(Tmin) = ν± formulas
(5.9) and (5.10) were proved in [17].
Remark 5.7. Theorem 5.6 is a generalization of similar results obtained in [7, 16, 25] for
(cid:3)
5.3. Spectral functions and the Fourier transform. Let T be a symmetric relation in
H defined by (3.35) in Case 1 and (3.39) in Case 2 and let τ = {τ+, τ−} be a boundary
parameter given by (4.1) in Case 1 and (4.70) in Case 2 . According to Theorems 4.2
and 4.8 the boundary value problems (4.4) -- (4.7) and (4.73) -- (4.76) establish a bijective
correspondence between boundary parameters τ and generalized resolvents R(·) = Rτ (·) of
30
VADIM MOGILEVSKII
bf (s) =ZI
f ∈ ef .
(5.21)
Rτ (λ) = PH(eT τ − λ)−1 ↾ H, λ ∈ C \ R.
T . Denote by eT τ the (H-minimal) exit space self-adjoint extension of T in the Hilbert space
eH ⊃ H generating Rτ (·):
Clearly, formula (5.21) gives a parametrization of all such extensions eT = eT τ by means of a
boundary parameter τ . Denote by Fτ (·) the spectral function of T corresponding to eT τ .
In the following we assume that a certain J-unitary extension eU of U is fixed and hence
a choice of eU does not matter in our further considerations.
For each ef ∈ Hb introduce the Fourier transform bf (·) : R → H0 by setting
the m-function mτ (·) is defined (see (5.1)). Note that in view of the assertion just after (5.2)
U (t, s)∆(t)f (t) dt,
(5.22)
ϕ∗
(5.23)
Definition 5.8. A distribution function Στ (·) : R → [H0] is called a spectral function of
the boundary value problem (4.4) -- (4.7) in Case 1 (resp. (4.73) -- (4.76) in Case 2 ) if, for
each ef ∈ Hb and for each finite interval [α, β) ⊂ R, the Fourier transform (5.22) satisfies the
equality
((Fτ (β) − Fτ (α))ef ,ef )H =Z[α,β)
(dΣτ (s)bf (s),bf (s)).
In the following "the boundary value problem" means either the boundary value problem
(4.4) -- (4.7) (in Case 1 ) or the boundary value problem (4.73) -- (4.76) (in Case 2 ).
Let Στ (·) be the spectral function of the boundary value problem. Then in view of (5.23)
bf ∈ L2(Στ ; H0) and the same reasonings as in [2] give the Bessel inequality bf L2(Στ ;H0) ≤
ef H (for the Hilbert space L2(Σ; H) see Subsection 2.2). Therefore for each ef ∈ H there
exists a function bf ∈ L2(Στ ; H0) (the Fourier transform of ef ) such that
f ∈ ef ,
and the equality V ef = bf , ef ∈ H, defines the contraction V : H → L2(Στ ; H0)
Theorem 5.9. For each boundary parameter τ there exists a unique spectral function Στ (·)
of the boundary value problem. This function is defined by the Stieltjes inversion formula
bf −Z β
U (t, ·)∆(t)f (t) dtL2(Στ ;H0) = 0,
lim
β↑b
ϕ∗
a
(5.24)
Στ (s) = − lim
δ→+0
lim
ε→+0
Im mτ (σ − iε) dσ.
If in addition τ is a truncated boundary parameter, then the corresponding spectral function
Στ (·) has the block matrix representation
1
πZ s−δ
−δ
(5.25)
2π sI (cid:19) ,
Στ (s) =(cid:18)Στ,1(s) Στ,2(s)
Στ,3(s)
1
s ∈ R
with respect to the decomposition H0 = H ′
In (5.25) Στ,1(·) is an [H ′
0]-valued (resp. [H]-valued) distribution function defined by
0 ⊕ bH2 in Case 1 (resp. H0 = H ⊕ bH in Case 2).
πZ s−δ
Im mτ,1(σ − iε) dσ
−δ
1
(5.26)
Στ,1(s) = − lim
δ→+0
with mτ,1(·) taken from (5.5)
lim
ε→+0
ON EIGENFUNCTION EXPANSIONS
31
Proof. In the case of an arbitrary boundary parameter τ one proves the required statement
by using Theorem 5.6 and the Stieltjes-Livsic inversion formula in the same way as in [2,
Theorem 6.5]. The statements of the theorem for a truncated boundary parameter τ are
implied by (5.24) and the block matrix representation (5.5) of mτ (·).
(cid:3)
Let H be decomposed as
(5.27)
H = H0 ⊕ mul T
(5.28)
and let T ′ be the operator part of T . Moreover, let τ be a boundary parameter, let eT τ =
(eT τ )∗ be an exit space extension of T in the Hilbert space eH ⊃ H, let eH be decomposed as
and let T τ be the operator part of eT τ (so that T τ is a self-adjoint operator in eH0). Assume
also that Στ (·) is a spectral function of the boundary value problem and V ∈ [H, L2(Στ ; H0)]
is the Fourier transform. As in [2] one proves that V ↾ mul T = 0. Therefore by (5.27) H0
is the maximally possible subspace of H on which V may by isometric.
eH =eH0 ⊕ muleT τ
Definition 5.10. A spectral function Στ (·) is referred to the class SF0 if the operator
V0 := V ↾ H0 is an isometry from H0 to L2(Στ ; H0).
Similarly to [2] the following equivalence is valid:
Στ (·) ∈ SF0 ⇐⇒ muleT τ = mul T.
Hence all spectral functions Στ (·) of the boundary value problem belong to SF0 if and only
if mul T = mul T ∗ or, equivalently, if and only if the operator T ′ is densely defined.
inverse Fourier transform is
As in [2] one proves the following statement: if Στ (·) ∈ SF0, then for each ef ∈ H0 the
(5.29)
(5.30)
ef = π(cid:18)ZR
ϕU (·, s) dΣτ (s)bf (s)(cid:19) ,
∆(I).
where the integral converges in the semi-norm of L2
The following theorem can be proved in the same way as Theorem 6.9 in [2].
Theorem 5.11. Let τ be a boundary parameter such that Στ (·) ∈ SF0 and let V be the
corresponding Fourier transform. Then H0 ⊂ eH0 and there exists a unitary extension
eV ∈ [eH0, L2(Στ ; H0)] of the isometry V0(= V ↾ H0) such that the operator T τ and the
multiplication operator Λ = ΛΣτ in L2(Στ ; H0) are unitarily equivalent by means of eV .
Moreover, if mul T = mul T ∗, then the statements of the theorem hold for arbitrary bound-
ary parameter τ (that is, for any spectral function Στ (·)).
Since in the case of a densely defined T one has mul T = mul T ∗ = {0}, the following
theorem is implied by Theorem 5.11.
Theorem 5.12. If T is a densely defined operator, then for each boundary parameter τ and
the corresponding spectral function Στ (·) the following hold: (i) eT τ is an operator, that is,
eT τ = T τ ; (ii) the Fourier transform V is an isometry; (iii) there exists a unitary extension
eV ∈ [eH, L2(Στ ; H0)] of V such that eT τ and Λ are unitarily equivalent by means of eV .
Since n+(Tmin) 6= n−(Tmin), the equality σ(T τ ) = R holds for any boundary parameter
τ . Moreover, Theorem 5.11 yields the following corollary.
32
VADIM MOGILEVSKII
Corollary 5.13. (1) If τ is a boundary parameter such that Στ (·) ∈ SF0, then the spectral
multiplicity of the operator T τ does not exceed ν−(= dim H0).
(2) If mul T = mul T ∗, then the above statement about the spectral multiplicity of T τ holds
for any boundary parameter τ .
In the case of the truncated boundary parameter τ there is a somewhat more information
about the spectrum of T τ . Namely, the following theorem is valid.
Theorem 5.14. (1) Let τ be a truncated boundary parameter such that Στ (·) ∈ SF0 and let
Στ,1(·) be a distribution function given by the block matrix representation (5.25) of Στ (·).
Moreover, let E(·) be the orthogonal spectral measure of the operator T τ and let Es(·) be the
singular part of E(·). Then
σac(T τ ) = R,
σs(T τ ) = Ss(Στ,1)
and the multiplicity of the orthogonal spectral measure Es(·) does not exceed ν+ +νb+ −νb−(=
dim H ′
0) in Case 1 and ν+(= dim H) in Case 2 (this implies that the same estimates are
valid for multiplicity of each eigenvalue λ0 of T τ ).
(2) If mul T = mul T ∗, then the above statements about the spectrum of T τ hold for each
truncated boundary parameter τ .
Proof. (1) let Στ,ac(·), Στ,s(·) and Στ 1,ac(·), Στ 1,s(·) be absolutely continuous and singular
parts of Στ (·) and Στ,1(·) respectively. Then in view of (5.25) the block matrix representa-
tions
2π sI (cid:19) ,
Στ,ac(s) =(cid:18)Στ 1,ac(s) Στ,2(s)
Στ,3(s)
1
Στ,s(s) =(cid:18)Στ 1,s(s) 0
0(cid:19)
0
hold with respect to the same decompositions of H0 as in Theorem 5.9. This and Theorems
5.11 and 2.2 give the required statements about the spectrum of T τ .
The statement (2) of the theorem follows from the fact that in the case mul T = mul T ∗
the inclusion Στ (·) ∈ SF0 holds for any boundary parameter τ and hence for any truncated
boundary parameter τ .
(cid:3)
In the next theorem we give a parametrization of all spectral functions Στ (·) immediately
in terms of the boundary parameter τ .
Theorem 5.15. Let X−(·) be the operator matrix (4.29) in Case 1 (resp. (4.92) in Case 2).
Then for each boundary parameter τ = {τ+, τ−} of the form (4.1) in Case 1 (resp. (4.70)
in Case 2) the equality
(5.31)
mτ (λ) = m0(λ) + Φ−(λ)(D0(λ) − D1(λ) M−(λ))−1D1(λ)Ψ−(λ),
λ ∈ C−
together with (5.24) defines a (unique) spectral function Στ (·) of the boundary value problem.
If in addition τ = {τ+, τ−} is a truncated boundary parameter defined by (4.1), (4.2) in Case
1 and (4.70), (4.71) in Case 2, then the distribution function Fτ (·) in (5.25) is defined by
(5.6) and (5.26). Moreover, the following statements are valid:
(1) Let in Case 1 bC02(λ), C0b(λ) and M4+(λ), N2+(λ) be defined by (4.3) and (4.40)
and let PHb be the orthoprojector in bH2 ⊕ Hb onto Hb. Then there exist the limits
Bτ := lim
y→+∞
1
1
iy PHb (D0(iy) − D1(iy) M−(iy))−1D1(iy)
iy (C0b(iy) − C1(iy)M4+(iy) + ibC02(iy)N2+(iy))−1C1(iy) =
iy M4+(iy)(C0b(iy) − C1(iy)M4+(iy) + ibC02(iy)N2+(iy))−1C0b(iy)
1
= lim
y→−∞
bBτ := lim
y→+∞
ON EIGENFUNCTION EXPANSIONS
33
and the following equivalence holds:
(5.32)
(2) Let in Case 2 C0(λ) and M+(λ) has the block representations (cf. (4.72) and (4.40))
Στ (·) ∈ SF0 ⇐⇒ Bτ = bBτ = 0
C0(λ) = (C02(λ), C0b(λ)) : (bH ⊕ bHb) ⊕ Hb → Hb,
λ ∈ C+
M+(λ) = (N2+(λ), N4+(λ), M4(λ))⊤ : Hb → bH ⊕ bHb ⊕ Hb,
and let N+(λ) = (N2+(λ), N4+(λ))⊤ ∈ [Hb, bH ⊕ bHb], λ ∈ C+. Then there exist limits
iy (C0b(iy) − C1(iy)M4(iy) + iC02(iy)N+(iy))−1C1(iy) =
iy PHb (D0(iy) − D1(iy) M−(iy))−1D1(iy)
Bτ := lim
λ ∈ C+
y→+∞
1
1
= lim
y→−∞
bBτ := lim
y→+∞
1
iy M4(iy)(C0b(iy) − C1(iy)M4(iy) + iC02(iy)N+(iy))−1C0b(iy)
and the equivalence (5.32) holds.
(3) Each spectral function Στ (·) belongs to SF0 if and only if
lim
y→+∞
1
iy M4+(iy) = 0 in
Case 1 (resp.
lim
y→+∞
1
iy M4(iy) = 0 in Case 2) and
lim
y→−∞
y(Im( M−(iy)h, h) − 1
2 P2h2) = +∞,
h 6= 0,
where h ∈ bH2 ⊕ Hb and P2 is the orthoprojector in bH2 ⊕ Hb onto bH2 in Case 1 (resp.
h ∈ bH2 ⊕ eHb and P2 is the orthoprojector in bH2 ⊕ eHb onto bH2 ⊕ bHb in Case 2).
Proof. The main statement of the theorem directly follows from Theorems 5.3 and 5.9.
Next, consider the boundary triplet Π− = { H0⊕Hb, Γ0, Γ1} for T ∗ defined in Propositions
3.7 and 3.8. Since the Weyl functions M±(·) of the decomposing boundary triplet Π− for
Tmax have the block representations (4.34), (4.35) in Case 1 and (4.94), (4.95) in Case 2
, it follows from (4.40) and Proposition 2.7, (3) that the Weyl functions of the triplet Π−
are M+(λ) and M−(λ). Now applying to the boundary triplet Π− Theorems 4.12 and 4.13
from [35] and taking equivalence (5.29) into account one obtains statements (1) -- (3) of the
theorem.
(cid:3)
5.4. The case of minimal deficiency indices. It follows from (3.22) that for a given
system (3.3) minimally possible deficiency indices of the linear relation Tmin are
(5.33)
n+(Tmin) = ν+,
n−(Tmin) = ν−
and the first (second) equality in (5.33) holds if and only if νb+ = 0 (resp. νb− = 0).
If n+(Tmin) = ν+, then by (3.22) one has n+(Tmin) ≤ n−(Tmin); moreover, n+(Tmin) <
In the case of the minimal deficiency index n+(Tmin) the above results can be somewhat
simplified. Namely, assume the first equality in (5.33) and let n+(Tmin) < n−(Tmin). Then
the equality νb+ = 0 implies Case 2. Moreover, Lemma 3.4 gives Hb = {0} and hence
n−(Tmin) if in additionbν > 0(⇔ bH 6= {0}).
Γ0b = Γ1b = 0. Therefore eHb = bHb and (3.31) yieldseΓ0b =bΓb.
If the assumption (A1) from Subsection 4.1 is satisfied, then by Proposition 3.8 the
equality
(5.34)
T = {{ey,ef} ∈ Tmax : Γ1ay = 0, bΓay = 0, bΓby = 0}.
(5.35)
(5.36)
34
VADIM MOGILEVSKII
defines a maximal symmetric relation T in H(= L2
∆(I)) with the deficiency indices n+(T ) = 0
and n−(T ) = bν +bνb−. Moreover, in this case T coincides with the relation A0 defined by
(3.42). Hence T has a unique generalized resolvent R(λ) of the form (4.9), which in view of
Theorem 4.8 and Remark 4.9 is given by the boundary value problem
J y′ − B(t)y = λ∆(t)y + ∆(t)f (t),
t ∈ I,
Γ1ay = 0, λ ∈ C+;
Γ1ay = 0, bΓay = 0, bΓby = 0, λ ∈ C−.
m(·) of the problem (5.35), (5.36) is given by the relations
If eU is a J-unitary extension (3.13) of U , then according to Proposition 5.2 the m-function
and ibΓav(λ) = P bH , bΓbv(λ) = 0, λ ∈ C−.
Since Hb = {0}, the decomposing boundary triplet (3.32) -- (3.34) for Tmax takes the form
v(t, λ) := ϕU (t, λ)m(λ) + ψ(t, λ) ∈ L2
Π− = {H0 ⊕ H, Γ0, Γ1}, where
∆[H0, H], λ ∈ C \ R
Assume that the Weyl function M−(·) of Π− has the block matrix representation (cf. (4.95))
H0 = H ⊕ bH ⊕ bHb = H0 ⊕ bHb,
(5.37)
Γ0{ey,ef } = {−Γ1ay, ibΓay,bΓby} (∈ H ⊕ bH ⊕ bHb), Γ1{ey,ef} = Γ0ay,
M−(λ) = (M (λ), N−(λ), M−(λ)) : H ⊕ bH ⊕ bHb → H,
m(λ) =(cid:18)M (λ) N−(λ)
According to Theorem 5.3 m(λ) = m0(λ) and by (4.97) one has
(5.38)
− i
0
,
λ ∈ C−.
{ey,ef } ∈ Tmax.
λ ∈ C−.
Let M(·) be the operator function (2.10), (2.11) corresponding to the decomposing boundary
triplet Π−. Then by (5.37) and (5.38)
(5.39)
,
λ ∈ C−,
2 I bH(cid:19) : H ⊕ bH
{z }H0
2 I bHb(cid:19) : H0 ⊕ bHb
{z
}H0
→ H ⊕ bH
{z }H0
→ H0 ⊕ bHb
{z
}H0
0
− i
M(λ) =(cid:18)m(λ) M1(λ)
m(µ) − m∗(λ) = (µ − λ)ZI
Σ(s) =(cid:18) F (s) Σ1(s)
Σ2(s)
where M1(λ) = (M−(λ), 0)⊤(∈ [bHb, H ⊕ bH]). Using (5.39) and (2.12) one can show that
v∗(t, λ)∆(t)v(t, µ) dt, µ, λ ∈ C−.
The spectral function Σ(·) of the problem (5.35), (5.36) has the block matrix representation
1
2π sI bH(cid:19) : H ⊕ bH → H ⊕ bH,
s ∈ R,
where F (s) is an [H]-valued distribution function defined by the Stieltjes formula (5.26)
with M (λ) in place of mτ,1(λ). Moreover, since T is maximal symmetric, it follows that
mul T = mul T ∗ and, consequently, Σ(·) ∈ SF0. Hence the operator part T ′
0 of the (unique)
exit space self-adjoint extension T0 of T satisfies statements of Theorem 5.14 with T ′
0 and
F (·) in place of T τ and Fτ (·) respectively.
If in addition to the above assumptions n−(Tmin) = ν− (i.e., both the deficiency in-
dices n±(Tmin) are minimal), then bHb = {0} and bΓb = 0. This implies the corresponding
modification of the results of this subsection. In particular, formula (5.39) takes the form
m(λ) = M(λ), λ ∈ C \ R.
ON EIGENFUNCTION EXPANSIONS
35
6. Differential operators of an odd order
In this section we apply the above results to ordinary differential operators of an odd
order on an interval I = [a, bi (−∞ < a < b ≤ ∞) with the regular endpoint a.
Assume that H is a finite dimensional Hilbert space and
(6.1)
l[y] =
(−1)k(cid:16) i
mXk=0
2 [(qn−ky(k))(k+1) + (qn−ky(k+1))(k)] + (pn−ky(k))(k)(cid:17)
is a differential expression of an odd order 2m + 1 with sufficiently smooth operator valued
coefficients pk(·), qk(·) : I → [H] such that pk(t) = p∗
k(t) and 0 ∈ ρ(q0(t)).
Denote by y[k](·), k ∈ {0, . . . , 2m + 1}, the quasi-derivatives of y ∈ AC(I; H) and let dom l
be the set of all functions y ∈ AC(I; H) for which l[y] := y[2m+1] makes sense [37, 24].
k(t), qk(t) = q∗
As is known H = H +
t ⊕ H −
t , where H +
t
(H −
t ) is an invariant subspace of the operator
q0(t), on which q0(t) is strictly positive (resp. negative). We put
ν0+ = dim H +
t ,
ν0− = dim H −
t
(this numbers does not depend on t); moreover, we assume for definiteness that ν0− ≤ ν0+.
By using formula (1.27) in [24] one can easily show that there exist finite dimensional
Hilbert spaces H ′ and bH and an absolutely continuous operator function
t ∈ I,
such that 0 ∈ ρ(Q(t)) and the following holds:
iq0(t) = −Q∗
Q(t) = (Q1(t), bQ(t), Q2(t))⊤ : H → H ′ ⊕ bH ⊕ H ′,
2(t)Q1(t) + ibQ∗(t)bQ(t),
dimbH = ν0+ − ν0−.
1(t)Q2(t) + Q∗
dim H ′ = ν0−,
H = H ⊕ · · · ⊕ H
⊕H ′(= H m ⊕ H ′)
Introduce the finite dimensional Hilbert spaces (cf. (3.1))
(6.2)
(6.3)
t ∈ I
m terms
{z
}
Clearly, the space H admits the representation
H0 = H ⊕ bH = H m ⊕ H ′ ⊕ bH, H = H0 ⊕ H = H ⊕ bH ⊕ H
(6.4)
H =
H ⊕ · · · ⊕ H
m terms
For each function y ∈ dom l we let
z
H
}
{z
{
⊕H ′ ⊕bH ⊕
z
}
H ⊕ · · · ⊕ H
H
}
{z
m terms
}
⊕H ′
{
(6.5)
(6.6)
(6.7)
y0(t) = {y(t), . . . , y[m−1](t), Q1(t)y(m)(t)}(∈ H)
y1(t) = {y[2m](t), . . . , y[m+1](t), Q2(t)y(m)(t)}(∈ H)
y(t) = {y0(t), bQ(t)y(m)(t), y1(t)}(∈ H ⊕ bH ⊕ H = H).
Let K be a finite dimensional Hilbert space. For an operator valued solution Y (·) : I →
[K, H] of the equation
(6.8)
l[y] = λy
(λ ∈ C)
we define the operator function Y(·) : I → [K, H] by the following relations: if h ∈ K and
Y (t)h = y(t), then Y(t)h = y(t).
36
VADIM MOGILEVSKII
on I satisfyingRI
Next assume that H′ := L2(I) is the Hilbert space of all Borel H-valued functions f (·)
f (t)2 dt < ∞. Denote also by L2[K, H] the set of all operator functions
Y (·) : I → [K, H] such that Y (t)h ∈ H′, h ∈ K. Moreover, by using (6.4) we associate with a
function f (·) ∈ H′ the H-valued functions f (·) on I given by f (t) = {f (t), 0, . . . , 0}, t ∈ I.
According to [37] expression (6.1) induces in H′ the maximal operator Lmax and the
minimal operator Lmin. Moreover, Lmin is a closed densely defined symmetric operator and
L∗
min = Lmax.
It turns out that the expression l[y] is equivalent in fact to a certain symmetric system.
More precisely, the following proposition is implied by the results of [24].
Proposition 6.1. Let l[y] be the expression (6.1) and let
J =
0
0
IH
0 −IH
iI bH
0
0
0 ∈ [H ⊕ bH ⊕ H],
∆(t) =
P 0 0
0
0 0
0
0 0 ∈ [H ⊕ bH ⊕ H],
where P is the orthoprojector in H onto the first term in the right hand side of (6.3). Then
there exists a continuous operator function B(t) = B∗(t)(∈ [H]), t ∈ I, (defined in terms of
pj and qj) such that the first-order symmetric system
(6.9)
Jy′(t) − B(t)y(t) = ∆(t)f (t),
t ∈ I
and the corresponding homogeneous system
(6.10)
Jy′(t) − B(t)y(t) = λ∆(t)y(t),
t ∈ I, λ ∈ C
possess the following properties:
(2) Let Tmax be the maximal linear relation in L2
of the system (6.10). Moreover, Y (·, λ) ∈ L2[K, H] if and only if Y(·, λ) ∈ L2
(1) There is a bijective correspondence Y (·, λ) ↔ eY (·, λ) = Y(·, λ) between all [K, H]-
valued operator solutions Y (·, λ) of Eq. (6.8) and all [K, H]-valued operator solutions eY (·, λ)
∆(I) induced by the system (6.9). Then
the equality (V1y)(t) = y(t), y(·) ∈ dom Lmax, defines a bijective linear mapping V1 from
dom Lmax onto dom Tmax satisfying (V1y, V1z)∆ = (y, z)H′, y, z ∈ dom Lmax.
(3) Let Tmin and Tmax be minimal and maximal relations in H := L2
∆(I) induced by
system (6.9). Then Tmin is a densely defined operator and the equality V2f = π( f (·)), f =
f (·) ∈ H′, defines a unitary operator V2 ∈ [H′, H] such that (V2 ⊕ V2) (gr Lmin) = gr Tmin and
(V2 ⊕ V2) (gr Lmax) = gr Tmax (i.e., the operators Lmin and Tmin as well as Lmax and Tmax
are unitarily equivalent by means of V2).
∆[K, H].
By using Proposition 6.1 one can easily translate all the results of [2] and the present
paper to the expression (6.1). Below we specify only the basic points in this direction.
respectively. Using (6.5) -- (6.7) introduce the linear mappings Γja : dom l → H, j ∈ {0, 1},
Let eU ∈ [H] be a J-unitary operator given by (3.13) with H and bH in place of H and bH
andbΓa : dom l → bH by setting
y ∈ dom l
(6.11)
(6.12)
Γ0ay = u7y0(a) + u8bQ(a)y(m)(a) + u9y1(a),
bΓay = u1y0(a) + u2bQ(a)y(m)(a) + u3y1(a),
Γ1ay = u4y0(a) + u5bQ(a)y(m)(a) + u6y1(a).
(Jy(t), z(t)),
y, z ∈ dom Lmax.
[y, z]b := lim
t↑b
Next, assume that νb+ and νb− are indices of inertia of the bilinear form
ON EIGENFUNCTION EXPANSIONS
37
Combining (3.22) with (6.3) and (6.2) and taking Proposition 6.1, (3) into account one gets
the following equality for deficiency indices d± = n±(Lmin) of the operator Lmin:
(6.13)
d+ = m · dim H + ν0− + νb+,
d− = m · dim H + ν0+ + νb−.
Therefore d+ < d− if and only if one of the following two alternative cases holds:
Case 1. ν0+ − ν0− > νb+ − νb− > 0.
Case 2. ν0+ − ν0− ≥ 0 ≥ νb+ − νb− and ν0+ − ν0− 6= νb+ − νb−(6= 0)
Proposition 6.1, (2) enables one to identify dom Lmax and dom Tmax. Therefore in the
case d− ≤ d+ we may assume that the linear mapping Γb in [2, Lemma 3.4] is of the form
(6.14)
Γb = (Γ0b,bΓb, Γ1b)⊤ : dom Lmax → H0b ⊕ bH ⊕ H1b,
where H0b is a finite dimensional Hilbert space and H1b is a subspace in H0b. Similarly
in the case d+ < d− we define on dom Lmax the linear mappings Γ0b, Γ1b,bΓb (Case 1 ) and
eΓ0b, Γ1b (Case 2 ) in the same way as in Subsection 3.3.
In the following we suppose that: (i) U is the operator (3.9) (with H and bH in place
of H and bH) satisfying (3.10) -- (3.12); (ii) bΓa and Γ1a are the mappings (6.12). Then the
same boundary conditions as in the right hand sides of (3.35), (3.39) and [2, (3.40)] give a
symmetric operator T (⊃ Lmin) in H′. Moreover, we define a boundary parameter τ in the
same way as for symmetric systems in [2, Definition 5.1] (the case d− ≤ d+) and Definitions
4.1 and 4.7 (the case d+ < d−). A boundary parameter τ gives a parametrization of all
generalized resolvents Rτ (·) and, consequently, all spectral functions Fτ (·) of T . Such a
parametrization is generated by means of a boundary value problem involving the equation
(6.15)
l[y] − λy = f (t),
t ∈ I
and the boundary conditions [2, (4.2) and (4.3)] (d− ≤ d+), (4.5) -- (4.7) (d+ < d−, Case 1 )
or (4.74) -- (4.76) (d+ < d−, Case 2). In the following "the boundary value problem for l[y]"
means one of the listed above boundary value problems for the expression l[y].
If d+ = d−, then in (6.14) H0b = H1b =: Hb and Rτ (·) is a canonical resolvent of T
if and only if τ = {(C0, C1); Hb} is a self-adjoint operator pair (this means that C0, C1 ∈
[Hb], ImC1C ∗
0 = 0 and 0 ∈ ρ(C0 ± iC1)). In this case the corresponding boundary problem
involves equation (6.15) and self-adjoint boundary conditions
(6.16)
Γ1ay = 0,
bΓay =bΓby, C0Γ0by + C1Γ1by = 0,
(6.17)
T τ = Lmax ↾ dom T τ is a self-adjoint extension of Lmin with the domain
where Γ0b, Γ1b and bΓb are taken from (6.14). Moreover, Rτ (λ) = (T τ − λ)−1, where
Next assume that eU ∈ [H ⊕ bH ⊕ H] is a J-unitary extension (3.13) of U and Γ0a is the
dom T τ = {y ∈ dom Lmax : Γ1ay = 0, bΓay =bΓby, C0Γ0by + C1Γ1by = 0}.
Definition 6.2. The m-function of the expression l[y] corresponding to the boundary pa-
rameter τ is the m-function mτ (·) of the equivalent system (6.9).
mapping (6.11).
Definition 6.2 means that mτ (·) : C \ R → [H0] is a unique operator function such that
for any λ ∈ C \ R the operator solution vτ (·, λ) of Eq. (6.8) given by
(6.18)
belongs to L2[H0, H] and satisfies the boundary conditions [2, (4.41) -- (4.43)] (d− ≤ d+),
In (6.18)
(4.57) -- (4.59) (d+ < d−, Case 1 ) or (4.100) and (4.101) (d+ < d−, Case 2 ).
vτ (t, λ) := ϕU (t, λ)mτ (λ) + ψ(t, λ)
38
VADIM MOGILEVSKII
bf (s) =ZI
ϕU (·, λ)(∈ [H0, H]) and ψ(·, λ)(∈ [H0, H]) are the operator solutions of Eq. (6.8) with the
initial data
0 (cid:19) : H0 → H0 ⊕ H,
eUϕϕϕU (a, λ) =(cid:18)IH0
(note that ϕU (·, λ) does not depend on a choice of the extension eU ⊃ U ).
eUψψψ(a, λ) =(cid:18)− i
b the Fourier transform bf (·) : R → H0 is
−PH(cid:19) : H0 → H0 ⊕ H
b be the set of all functions f (·) ∈ H′ such that f (t) = 0 on some interval (β, b)
(depending on f ). For a function f (·) ∈ H′
ϕ∗
U (t, s)f (t) dt.
Let H′
(6.19)
2 P bH
Definition 6.3. Let τ be a boundary parameter. A distribution function Στ (·) : R → [H0]
is called a spectral function of the boundary value problem for l[y] if for each f (·) ∈ H′
b the
Fourier transform (6.19) satisfies (5.23) (with H′ in place of H).
V : H′ → L2(Στ ; H0).
f H′ . Therefore for each f ∈ H′ there exists the fourier transform (6.19) (the integral
Since Lmin is densely defined, formula (5.23) yields the Parseval equality bf L2(Στ ;H0) =
converges in the norm of L2(Στ ; H0)) and the equality V f = bf , f ∈ H′, define an isometry
In view of Proposition 6.1 each spectral function Στ (·) of the expression l[y] is a spectral
function of the equivalent symmetric system (6.9) and vice versa. Hence by [2, Theorem
6.5] and Theorem 5.9 for each boundary parameter τ there exits a unique spectral function
Στ (·) of the boundary value problem for l[y] and this function is defined by the Stieltjes
inversion formula (5.24). Moreover, for each f ∈ H′ the inverse Fourier transform is
f (t) =ZR
ϕU (t, s) dΣτ (s)bf (s).
For a given boundary parameter τ denote by T τ the (H-minimal) exit space self-adjoint
extension of T generating Rτ (λ). Since T is densely defined, T τ is a self-adjoint operator in
The following theorem is implied by Theorem 5.12 and [2, Theorem 6.11].
Theorem 6.4. Let τ be a boundary parameter and let Στ (·) and V be the corresponding
spectral function and Fourier transform respectively. Then there exists a unitary extension
the Hilbert space eH ⊃ H′.
eV ∈ [eH, L2(Στ ; H0)] of V such that T τ and the multiplication operator Λ in L2(Στ ; H0) are
unitarily equivalent by means of eV . Moreover, the following statements are equivalent:
(1) d+ = d− and τ = {(C0, C1); Hb} is a self-adjoint operator pair, so that T τ is the
canonical self-adjoint extension (6.17) of T ;
(2) V H′ = L2(Στ ; H0), that is the fourier transform V is a unitary operator.
If the statement (1) (and hence (2)) is valid, then the operators T τ and Λ are unitarily
equivalent by means of V .
Theorem 6.4 immediately implies that the spectral multiplicity of T τ does not exceed
m · dim H + ν0+(= dim H0).
Similarly one can translate to the expression l[y] Theorem 5.14 and parametrization of all
spectral functions Στ (·) by means of (5.31), [35, (5.22)] and the Stieltjes inversion formula.
It follows from (6.13) that for a given expression (6.1) minimally possible deficiency indices
of the operator Lmin are
n+(Lmin) = m · dim H + ν0−,
n−(Lmin) = m · dim H + ν0+.
ON EIGENFUNCTION EXPANSIONS
39
The routine reformulation of the results of Subsection 5.4 to the case of the expression l[y]
with minimal deficiency index n+(Lmin) of the operator Lmin is left to the reader.
Remark 6.5. Let H = C, so that l[y] is a scalar expression with real valued coefficients pk(·)
and qk(·). Then q0(t) > 0, bQ(t) = q
H = Cm,
0 (t), bH = C and the equalities (6.3) -- (6.7) take the form
H = Cm+1 ⊕ Cm = C2m+1
H0 = Cm ⊕ C = Cm+1,
y0(t) = {y(t), . . . , y[m−1](t)}(∈ Cm), y1(t) = {y[2m](t), . . . , y[m+1](t)}(∈ Cm),
1
2
y(t) = {y0(t), q
0 (t)y(m)(t), y1(t)}(∈ C2m+1).
1
2
In this case ϕU (·, λ) is the (m + 1)-component operator solution
ϕU (t, λ) = (ϕ1(t, λ), ϕ2(t, λ), . . . , ϕm(t, λ), ϕm+1(t, λ)) : Cm+1 → C
of Eq. (6.8) with the initial data
eUϕϕϕU (a, λ) =(cid:18)Im+1
0 (cid:19) : Cm+1 → Cm+1 ⊕ Cm
and (6.19) defines the Fourier transform bf (·) : R → Cm+1. Moreover, in a fixed basis of Cm+1
the m-function mτ (·) and spectral function Στ (·) can be represented as (m + 1) × (m + 1)-
matrix valued functions mτ (λ) = (mij (λ))m+1
i,j=1 and Στ (λ) = (Σij(λ))m+1
i,j=1 respectively.
In conclusion note that for a scalar expression (6.1) one has ν0+ = 1 and ν0− = 0.
Therefore for a scalar expression l[y] with n+(Lmin) < n−(Lmin) Case 1 is impossible. In
other words, for such an expression either n−(Lmin) ≤ n+(Lmin) or n+(Lmin) < n−(Lmin)
and Case 2 holds.
References
[1] N.I. Akhiezer, I.M. Glazman, Theory of linear operators in Hilbert space, Dover Publications, New
York, 1993.
[2] S. Albeverio,M.M. Malamud, V.I.Mogilevskii,On Titchmarsh-Weyl functions and eigenfunction ex-
pansions of first-order symmetric systems, arXiv:1303.6153[math.FA] 25 Mar 2013.
[3] F.V. Atkinson, Discrete and continuous boundary problems, Academic Press, New York, 1963.
[4] J. Behrndt, S. Hassi, H. de Snoo, R. Wiestma, Square-integrable solutions and Weyl functions for
singular canonical systems, Math. Nachr. 284 (2011), no 11-12, 1334 -- 1383.
[5] V.A. Derkach, S. Hassi, M.M. Malamud, H.S.V. de Snoo, Generalized resolvents of symmetric oper-
ators and admissibility, Methods of Functional Analysis and Topology 6 (2000), no 3 , 24 -- 55.
[6] V.A. Derkach, M.M. Malamud, Generalized resolvents and the boundary value problems for Hermit-
ian operators with gaps, J. Funct. Anal. 95 (1991),1 -- 95.
[7] A. Dijksma, H. Langer, H.S.V. de Snoo, Eigenvalues and pole functions of Hamiltonian systems with
eigenvalue depending boundary conditions, Math. Nachr. 161 (1993), 107 -- 153.
[8] N. Dunford and J.T. Schwartz, Linear operators. Part2. Spectral theory, Interscience Publishers, New
York-London, 1963.
[9] W.N. Everitt and V. Krishna Kumar, On the Titchmarsh-Weyl theory of ordinary symmetric differ-
ential expressions 1 : The general theory , Nieuw Arch. voor Wiskunde (3) 24 (1976), 1 -- 48.
[10] W.N. Everitt and V. Krishna Kumar, On the Titchmarsh-Weyl theory of ordinary symmetric differ-
ential expressions 2 : The odd-order case, Nieuw Arch. voor Wiskunde (3) 24 (1976), 109 -- 145.
[11] Ch. T. Fulton, Parametrizations of Titchmarsh's m(λ)-functions in the limit circle case, Trans.
Amer. Math. Soc. 229 (1977), 51 -- 63.
[12] I. Gohberg, M.G. Krein, Theory and applications of Volterra operators in Hilbert space, Transl. Math.
Monographs, 24, Amer. Math. Soc., Providence, R.I., 1970
[13] M.L. Gorbachuk, On spectral functions of a differential equation of the second order with operator-
valued coefficients, Ukrain. Mat. Zh. 18 (1966), no 2 , 3 -- 21.
40
VADIM MOGILEVSKII
[14] V.I. Gorbachuk, M.L. Gorbachuk, Boundary problems for differential-operator equations, Kluver
Acad. Publ., Dordrecht-Boston-London, 1991. (Russian edition: Naukova Dumka, Kiev, 1984).
[15] D.B. Hinton, J.K. Shaw, Parameterization of the M (λ) function for a Hamiltonian system of limit
circle type, Proc. Roy. Soc. Edinburgh Sect. A 93 (1982/83), no 3-4, 349-360.
[16] D.B. Hinton, A. Schneider, On the Titchmarsh-Weyl coefficients for singular S-Hermitian systems
I, Math. Nachr. 163 (1993), 323 -- 342.
[17] D.B. Hinton, A. Schneider, Titchmarsh-Weyl coefficients for odd order linear Hamiltonian systems,
J. Spectral Mathematics 1(2006),1-36.
[18] D.B. Hinton, A. Schneider, On the spectral representation for singular selfadjoint boundary eigen-
function problems, Oper. Theory: Advances and Applications, 106, Birkhauser, Basel, 1998.
[19] T. Kato,Perturbation theory of linear operators,Springer-Verlag, Berlin,1966.
[20] I.S. Kats, Linear relations generated by the canonical differential equation of phase dimension 2, and
eigenfunction expansion, St. Petersburg Math. J. 14, 429 -- 452 (2003).
[21] I.S. Kac, M.G. Krein, On spectral functions of a string, Supplement to the Russian edition of F.V.
Atkinson , Discrete and continuous boundary problems, Mir, Moscow, 1968.
[22] A.M. Khol'kin, Description of selfadjoint extensions of differential operators of an arbitrary order
on the infinite interval in the absolutely indefinite case, Teor. Funkcii Funkcional. Anal. Prilozhen.
44 (1985), 112 -- 122.
[23] V.I. Khrabustovsky, On the characteristic operators and projections and on the solutions of Weyl
type of dissipative and accumulative operator systems. 3. Separated boundary conditions, J. Math.
Phis. Anal. Geom., 2 (2006), no 4, 449 -- 473.
[24] V.I. Kogan, F.S. Rofe-Beketov, On square-integrable solutions of symmetric systems of differential
equations of arbitrary order, Proc. Roy. Soc. Edinburgh Sect. A 74 (1974/75), 5 -- 40.
[25] A.M. Krall, M (λ)-theory for singular Hamiltonian systems with one singular endpoint, SIAM J.
Math. Anal. 20 (1989), 664 -- 700.
[26] V. Krishna Kumar, On the Titchmarsh-Weyl theory of ordinary symmetric odd-order differential
expressions and a direct convergence theorem, Quaestiones Math. 5 (1982), 165-185.
[27] H. Langer, 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.
[28] M. Lesch, M.M. Malamud, On the deficiency indices and self-adjointness of symmetric Hamiltonian
systems, J. Differential Equations 189 (2003), 556 -- 615.
[29] M. M. Malamud, On the formula of generalized resolvents of a nondensely defined Hermitian oper-
ator, Ukr. Math. Zh. 44(1992), no 12, 1658-1688.
[30] M. M. Malamud, V. I. Mogilevskii, Krein type formula for canonical resolvents of dual pairs of linear
relations, Methods of Funct. Anal. and Topology 8, No.4 (2002), 72-100.
[31] M. Malamud, H. Neidhardt, Sturm-Liouville boundary value problems with operator potentials and
unitary equivalence, J. Differential Equations, 252 (2012), 5875-5922.
[32] V.I.Mogilevskii, Boundary triplets and Krein type resolvent formula for symmetric operators with
unequal defect numbers, Methods Funct. Anal. Topology 12 (2006), no 3, 258 -- 280.
[33] V.I.Mogilevskii, Description of spectral functions of differential operators with arbitrary deficiency
indices, Math. Notes 81 (2007), no 4, 553 -- 559.
[34] V.I.Mogilevskii, Boundary pairs and boundary conditions for general (not necessarily definite) first-
order symmetric systems with arbitrary deficiency indices, Math. Nachr.285 (2012), no14 -- 15, 1895 --
1931
[35] V.I.Mogilevskii, On exit space extensions of symmetric operators with applications to first order
symmetric systems, Methods Funct. Anal. Topology 19 (2013), to appear
[36] B.C. Orcutt, Canonical differential equations, Dissertation, University of Virginia, 1969.
[37] J. Weidmann, Spectral theory of ordinary differential operators, Lecture notes in mathematics, 1258,
Springer-Verlag, Berlin, 1987.
Department of Mathematical Analysis, Lugans'k National University, Oboronna Str. 2, 91011
Lugans'k, Ukraine
E-mail address: [email protected]
|
1703.01891 | 1 | 1703 | 2017-03-06T14:29:06 | Coarse embeddings into $c_0(\Gamma)$ | [
"math.FA"
] | Let $\lambda$ be a large enough cardinal number (assuming GCH it suffices to let $\lambda=\aleph_\omega$). If $X$ is a Banach space with $\text{dens}(X)\ge\lambda$, which admits a coarse (or uniform) embedding into any $c_0(\Gamma)$, then $X$ fails to have nontrivial cotype, i.e. $X$ contains $\ell_\infty^n$ $C$-uniformly for every $C>1$. In the special case when $X$ has a symmetric basis, we may even conclude that it is linearly isomorphic with $c_0(\text{dens}X)$. | math.FA | math |
Coarse embeddings into c0(Γ)
Petr H´ajek∗1 and Th. Schlumprecht† 2
1Institute of Mathematics, Academy of Science of the Czech
Republic, Zitn´a 25 115 67 Prague 1, Czech Republic, and
Department of Mathematics, Faculty of Electrical Engineering,
Czech Technical University in Prague, Zikova 4, 160 00 Prague,
Czech Republic.
email: [email protected]
2Department of Mathematics, Texas A&M University, College
Station, TX 77843, USA and Department of Mathematics, Faculty
of Electrical Engineering, Czech Technical University in Prague,
Zikova 4, 160 00 Prague, Czech Republic.
email: [email protected]
July 19, 2018
Abstract
Let λ be a large enough cardinal number (assuming GCH it suffices
to let λ = ℵω). If X is a Banach space with dens(X) ≥ λ, which admits
a coarse (or uniform) embedding into any c0(Γ), then X fails to have
nontrivial cotype, i.e. X contains ℓn
∞ C-uniformly for every C > 1. In the
special case when X has a symmetric basis, we may even conclude that it
is linearly isomorphic with c0(densX).
1 Introduction
The classical result of Aharoni states that every separable metric space (in
particular every separable Banach space) can be bi-Lipschitz embedded (the
definition is given below) into c0.
The natural problem of embeddings of metric spaces into c0(Γ), for an arbi-
trary set Γ, has been treated by several authors, in particular Pelant and Swift.
∗The first named author was supported by GA CR 16-073785 and RVO: 67985840.
†The second named author was supported by the National Science Foundation under the
Grant Number DMS -- 1464713.
2000 Mathematics Subject Classification Primary 46B26; Secondary 54E15, 54D20
1
1 INTRODUCTION
2
The characterizations that they obtained, and which play a crucial role in our
argument, are described below.
Our main interest, motivated by some problems posed in [3], lies in the case
of embeddings of Banach spaces into c0(Γ).
We now state the main results of this paper. We first define the following
cardinal numbers inductively. We put λ0 = ω0, and, assuming for n ∈ N0, λn
has been defined, we put λn+1 = 2λn. Then we let
λ = lim
n→∞
λn.
(1)
It is clear that assuming the generalized continuum hypothesis (GCH) λ = ℵω.
Theorem A. If X is a Banach space with density dens(X) ≥ λ, which admits
a coarse (or uniform) embedding into any c0(Γ), then X fails to have nontrivial
cotype, i.e. X contains ℓn
∞ C-uniformly for some C > 1 (equivalently, every
C > 1).
Our method of proof gives a much stronger result for Banach spaces with
a symmetric basis. Namely, under the assumptions of Theorem A, such spaces
are linearly isomorphic with c0(Γ) (Theorem (4.2)).
Theorem A will follow from the following combinatorial result which is of
independent interest.
Theorem B. Assume that Λ is a set whose cardinality is at least λ, n ∈ N,
and σ : [Λ]n → C is a map into an arbitrary set C. Then (at least) one of the
following conditions holds:
1. There is a sequence (Fj )∞
j=1 of pairwise disjoint elements of [Λ]n, so that
σ(Fi) = σ(Fj ), for all i, j ∈ N.
2. There is an F ∈ [Λ]n−1 so that σ(cid:0)(cid:8)F ∪ {γ} : γ ∈ Λ(cid:9)(cid:1) is infinite.
The above Theorem B was previously deduced in [6] from a combinatorial
result of Baumgartner, provided Λ is a weakly compact cardinal number (whose
existence is not provable in ZFC, as it is inaccessible [5] p. 325, p. 52). The au-
thors in [6] pose a question whether assuming that Γ is uncountable is sufficient
in Theorem B.
Theorem B is used in order to obtain a scattered compact set K of height
ω0, such that C(K) does not uniformly embed into c0(Γ). It is easy to check
that our version of Theorem B implies a ZFC example of such a C(K) space.
It is further shown in [6] that the space C[0, ω1] does not uniformly embed into
any c0(Γ).
Let us point out that a special case of Theorem A was obtained by Pelant
and Rodl [8], namely it was shown there that ℓp(λ), 1 ≤ p < ∞, spaces (which
are well known to have nontrivial cotype) do not uniformly embed into any
c0(Γ).
The paper is organized as follows. In Section 2 we recall Pelant's [7, 6] amd
Swift's [10] conditions for Lipschitz, uniform, and coarse embeddability into
2 PELANT'S AND SWIFT'S CRITERIA
3
c0(Γ). In Section 3 we provide a proof for Theorem B. Finally, in Section 4, we
provide a proof of Theorem A as well as the symmetric version of the result.
All set theoretic concepts and results used in our note can be found in [5],
whereas for facts concerning nonseparable Banach spaces [4] can be consulted.
2 Pelant's and Swift's criteria for Lipschitz, uni-
form, and coarse embeddability into c0(Γ)
In this section we recall some of the notions and results by Pelant [7, 6] and
Swift [10] about embeddings into c0(Γ).
For a metric space (M, d) a cover is a set U of subsets of M such that
M = SU∈U U . A cover U of M is called uniform if there is an r > 0 so that for
all x ∈ M there is a U ∈ U, so that Br(x) = {x′ ∈ M : d(x′, x) < r} ⊂ U . It is
called uniformly bounded if the diameters of the U ∈ U are uniformly bounded,
and it is called point finite if every x ∈ M lies in only finitely many U ∈ U. A
cover V of M is a refinement of a cover U, if for every V ∈ V there is a U ∈ U,
for which V ⊂ U .
Definition 2.1. [6] A metric space (M, d) is said to have the Uniform Stone
Property if every uniform cover U of M has a point finite uniform refinement.
Definition 2.2. [10] A metric space (M, d) is said to have the Coarse Stone
Property if every bounded cover is the refinement of a point finite uniformly
bounded cover.
Definition 2.3. Let (M1, d1) and (M2, d2) be two metric spaces. For a map
f : M1 → M2 we define the modulus of uniform continutiy wf : [0, ∞) → [0, ∞],
and the modulus of expansion ρ : [0, ∞) → [0, ∞] as follows
wf (t) = sup(cid:8)d2(cid:0)f (x), f (y)(cid:1) : x, y ∈ M1, d1(x, y) ≤ t(cid:9) and
ρf (t) = inf (cid:8)d2(cid:0)f (x), f (y)(cid:1) : x, y ∈ M1, d1(x, y) ≥ t(cid:9).
The map f is called uniform continuous if limt→0 wf (t) = 0, and it is called a
uniform embedding if moreover ρf (t) > 0 for every t > 0. It is called coarse
if wf (t) < ∞, for all 0 < t < ∞ and it is called a coarse embedding,
if
limt→∞ ρf (t) = ∞. The map f is called Lischitz continuous if
Lip(f ) = sup
x6=y
d2(f (x), f (y))
d1(x, y)
< ∞,
and a bi-Lipschitz embedding, if f is injective and Lip(f −1) is also finite.
The following result recalls results from [6](for (i) ⇐⇒ (ii)) and [10] (for
(ii) ⇐⇒ (iii) ⇐⇒ (iv) ⇐⇒ (v)).
Theorem 2.4. For a Banach space X the following properties are equivalent.
(i) X has the uniform Stone Property.
3 A COMBINATORIAL ARGUMENT
4
(ii) X is uniformly embeddable into c0(Γ), for some set Γ.
(iii) X has the coarse Stone Property.
(iv) X is coarsely embeddable into c0(Γ), for some set Γ.
(v) X is bi-Lipschitzly embeddable into c0(Γ), for some set Γ.
It is easy to see, and was noted in [6, 10], that the uniform Stone property
and the coarse Stone property are inherited by subspaces. The equivalence
(i) ⇐⇒ (ii) was used in [6] to show that C[0, ω1] does not uniformly embed
in any c0(Γ).
It was also used to prove that certain other C(K)-spaces do
not uniformly embed into c0(Γ): Let Λ be any set and denote for n ∈ N by
[Λ]≤n and [Λ]n the subsets of Λ which have cardinality at most n and exactly
n, respectively. Endow [Λ]≤n with the restriction of the product topology on
{0, 1}Λ (by identifying each set with its characteristic function). Then define
KΛ to be the one-point Alexandroff compactification of the topological sum of
the spaces [Λ]≤n, n ∈ N.
It was shown in [6] that if Λ satisfies Theorem B
then C(KΛ) is not uniformly Stone and thus does not embed uniformly into
any c0(Γ).
3 A combinatorial argument
We start by introducing property P (α) for a cardinal α as follows.
For every n ∈ N and any map σ : [α]n → C, C being an arbitrary set, (P (α))
(at least) one of the following two conditions hold:
There is a sequence (Fj) of pairwise disjoint elements of [λ]n, with σ(Fi) = σ(Fj ),
(2)
for any i, j ∈ N.
There is an F ∈ [λ]n−1, so that σ(cid:0)(cid:8)F ∪ {γ} : γ ∈ λ \ F(cid:9)(cid:1) is infinite.
As remarked in Section 2, if κ is an uncountable weakly compact cardinal
number, then P (κ) holds. But the existence of weakly compact cardinal numbers
requires further set theoretic axioms, beyond ZFC [5]. In [6, Question 3] the
authors ask if P (ω1) is true.
(3)
Theorem 3.1. For λ defined by (1), P (λ) holds.
For our proof of Theorem 3.1 it will be more convenient to reformulate it
into a statement about n-tuples, instead of sets of cardinality n. We will first
introduce some notation.
Let n ∈ N and Γ1, Γ2, . . . Γn be sets of infinite cardinality, and put Γ =
Qn
i=1 Γi. For a ∈ Γ and 1 ≤ i ≤ n we denote the i-the coordinate of a by
a(i). We say that two points a and b in Γ are diagonal, if a(i) 6= b(i), for all
i ∈ {1, 2, . . . , n}.
3 A COMBINATORIAL ARGUMENT
5
Let a ∈ Γi for i ∈ N. For i ∈ {1, 2 . . . , n} we call the set
H(a, i) = (cid:8)(b1, b2, . . . , bi−1, a(i), bi+1, . . . , αn) : bj ∈ Γj, for j ∈ {1, 2 . . . n} \ {i}(cid:9),
the Hyperplane through the point a orthogonal to i. We call the set
L(a, i) = (cid:8)(a(1), . . . , a(i − 1), bi, a(i + 1), . . . a(n)) : bi ∈ Γi(cid:9),
the Line through the point a in direction of i.
For a cardinal number β, we define recursively the following sequence of car-
dinal numbers (cid:0) exp+(β, n) : n ∈ N0(cid:1): exp+(β, 0) = β, and, assuming exp+(β, n)
has been defined for some n ∈ N0, we put
exp+(β, n + 1) = (cid:0)2exp+(β,n)+(cid:1)+
.
Here γ+ denotes the successor cardinal, for a cardinal γ, i.e., the smallest car-
dinal number γ′ with γ′ > γ. Note that since exp+(γ, 1) ≤ 222γ
, it follows for
the above defined cardinal number λ, that
λ = lim
n→∞
exp+(ω0, n).
Secondly, successor cardinals are regular [5], and thus every set of cardinality
γ, with γ being a successor cardinal, can be partitioned for n ∈ N into n
disjoint sets Γ1, Γ2, . . . , Γn, all of them having also cardinality γ, and the map
Γ1 × Γ2 × . . . × Γn → (cid:2)Sn
, (a1, a2, . . . , an) 7→ {a1, a2, . . . an}, is injective.
We therefore deduce that the following statement implies Theorem 3.1.
i=1 Γi(cid:3)n
Theorem 3.2. Let n ∈ N and a assume that the sets Γ1, Γ2, . . . Γn have cardi-
nality at least exp+(ω1, n2). For any function
σ : Γ :=
n
Y
i=1
Γi → C,
where C is an arbitrary set, at least one of the following two conditions hold
There is a sequence (a(j))∞
σ(a(i)) = σ(a(j)), for any i, j ∈ N.
There is a line L ⊂ Γ, for which σ(L) is infinite.
j=1, of pairwise diagonal elements in Γ, so that (4)
(5)
We will make the following observation before proving Theorem 3.2.
Lemma 3.3. Let n ∈ N and Γ1, Γ2, . . . Γn be non empty sets. Let
σ : Γ :=
n
Y
i=1
Γi → C,
be a function that fails both conditions (4) and (5).
3 A COMBINATORIAL ARGUMENT
6
Then there is a set C and a function
σ : Γ :=
n
Y
i=1
Γi → C,
that fails both (4) and (5) and moreover has the property that
for every c ∈ C there is a hyperplane Hc ⊂ Γ so that {b ∈ Γ : σ(b) = c} ⊂ Hc.
(6)
Proof. We may assume without loss of generality that σ is surjective. Since (4)
is not satisfied for each c ∈ C there exists an m(c) ∈ N and a (finite) sequence
(a(c,j))m(c)
j=1 ⊂ σ−1({c}), which is pairwise diagonal, and maximal, with this
property. Hence
σ−1({c}) ⊂
m(c)
n
[
j=1
[
i=1
H(a(c,j), i).
Indeed, from the maximality of (a(c,j))m(c)
j=1 ⊂ σ−1({c}), it follows that each
b ∈ σ−1({b}) must have at least one coordinate in common with at least one
element of (a(c,j))m(c)
j=1 ⊂ σ−1({c}).
We define
C = [
c∈C
{1, 2, . . . , m(c)} × {1, 2 . . . n} × {c},
and
σ : Γ → C,
c = σ(b),
b 7→ (c, j, i), where
n
j = minnj′ : b ∈
[
i′=1
H(a(c,j′), i′)o, and i = min{i′ : b ∈ H(a(c,j), i)}.
It is clear that σ satisfies (6). Since for every c ∈ C,
{b ∈ Γ : σ(b) = c} =
m(c)
n
[
j=1
[
i=1
{b ∈ Γ : σ(b) = (c, j, i)},
σ does not satisfy (4). In order to verify that (5) is not satisfied, assume L ⊂ Γ
is a line, and let {c1, c2, . . . , cp} be the image of L under σ. By construction,
σ(L) ⊂ {(j, i, ck), k ≤ p, j ≤ m(ck), ı ≤ n},
which is also finite.
Proof of Theorem 3.2. We assume that σ : Γ = Γ1 × Γ2 × . . . × Γn → C is a
map which fails both (4) and (5). By Lemma 3.3 we may also assume that σ
satisfies (6). For each a ∈ Γ we fix an i(a) ∈ {1, 2, . . . , n} so that σ−1(cid:0){σ(a)}(cid:1) ⊂
H(a, i(a)). It is important to note that, since (5) is not satisfied, it follows that
3 A COMBINATORIAL ARGUMENT
7
each line L, whose direction is some j ∈ {1, 2, . . . , n}, can only have finitely
many elements b for which i(b) = j.
Indeed, if i(b) = j then b is uniquely
determined by the value σ(b). To continue with the the proof the following
Reduction Lemma will be essential.
Lemma 3.4. Let β be an uncountable regular cardinal. Assume that Γ1 ⊂ Γ1,
Γ2 ⊂ Γ2, . . . , Γn ⊂ Γn are such that Γi ≥ exp+(β, n), for all i ∈ {1, . . . , n}.
Then, for any i ∈ {1, 2 . . . , n} there are a number Ki ∈ N, and subsets
1 ⊂ Γ1, Γ′
Γ′
2 ⊂ Γ2, . . . , Γ′
n ⊂ Γn, with Γ′
j ≥ β, so that
∀(a1, a2, . . . , ai−1, ai+1, . . . an) ∈
n
Y
j=1,j6=i
Γ′
i
(7)
Proof. We assume without loss of generality that i = n. Abbreviate βj =
exp+(β, j), for j = 1, 2 . . . n. We choose subsets Γ(0)
βn+1−j.
i : i(a1, a2, . . . ai−1, a, ai+1, . . . an) = i}(cid:12)(cid:12) ≤ Ki.
Γ(0)
(cid:12)(cid:12) =
j
j ⊂ Γj, for which (cid:12)(cid:12)
(cid:12)(cid:12){a ∈ Γ′
Since the βj's are regular, it follows for each j = 1, 2 . . . , n − 1 that
Γ(0)
(cid:12)(cid:12)
j
(cid:12)(cid:12) = βn+1−j
> 2βn−j
= 2Γ(0)
= (cid:12)(cid:12){f : Γ(0)
j+1×Γ(0)
j+1×...×Γ(0)
n
j+1 × Γ(0)
j+1 × . . . × Γ(0)
n → N}(cid:12)(cid:12).
Abbreviate for i = 1, . . . , n.
Fj = {f : Γ(0)
j × Γ(0)
j+1 × . . . × Γ(0)
n → N}
and consider the function
φ1 :
n−1
Y
j=1
Γ(0)
j → N,
(a1, a2, . . . , an−1) 7→ (cid:12)(cid:12){a ∈ Γ(0)
n : i(a1, a2, . . . , an−1, a) = n}(cid:12)(cid:12).
For fixed a1 ∈ Γ(0)
1 , φ1(a1, ·) ∈ F2, and the cardinality of F2 is by the above
estimates smaller than the cardinality of Γ(0)
1 , which is regular. Therefore we
1 ⊂ Γ(0)
can find a function Φ2 ∈ F2 and a subset Γ′
of cardinality βn so that
φ1(a1, ) = φ2 for all a1 ∈ Γ′
j ⊂ Γ(0)
,
for j = 1, 2 . . . , n − 2 of cardinality βn+1−j and functions φj ∈ Fj, for j =
1, 2, . . . , n − 1, so that for all (a1, a2, . . . , an−2) ∈ Qn−2
n−1, we
1. We continue the process and find Γ′
j and an−1 ∈ Γ(0)
j=1 Γ′
j
i
have
φ1(a1, a2, . . . , an−1) = φ2(a2, . . . , an−1) = . . . = φn−1(an−1).
(8)
Then, since φn−1 is N valued, we can finally choose an Kn ∈ N and a subset
Γ′
n−1, of cardinality at least β, so that φn−1(a1) ≤ Kn, which finishes our
argument.
4 PROOF OF THEOREM A
8
Continuation of the proof of Theorem 3.2. We apply Lemma 3.4 successively
to all i ∈ {1, 2 . . . n}, and the cardinals β(i) = exp+(ω1, n(n − i)). We obtain
numbers K1, K2, . . . Kn in N and infinite sets Λj ⊂ Γi, for j ≤ n, so that for all
i = 1, 2 . . . n and all a ∈ Qn
j=1 Λj
(cid:12)(cid:12){a ∈ Λi : i(a1, a2, . . . , ai−1, a, ai+1, . . . an) = i}(cid:12)(cid:12) ≤ Ki.
In order to deduce a contradiction choose for each j = 1, . . . n a subset Aj of Λj
of cardinality lj = (n + 1)Kj. Then it follows that
Y
j=1
Aj(cid:12)(cid:12)(cid:12)
n
Y
j=1
lj = (cid:12)(cid:12)(cid:12)
X
=
n
X
i=1
a∈Qn
j=1,j6=i Aj
(cid:12)(cid:12){a ∈ Ai : i(a1, a2, . . . , ai−1, a, ai+1, . . . an) = i}(cid:12)(cid:12)
≤
n
X
i=1
Ki
n
Y
j=1,j6=i
lj ≤
n
n + 1
n
Y
j=1
lj
which is a contradiction and finishes the proof of the Theorem.
We can now state the ZFC version of Theorem 4.1.
in [6], in which it was
shown that for weakly compact cardinalities κ0 the space C(Kκ0 ), where Kκ0
was defined at the end of Section 2, cannot be uniformly (or coarsely) embedded
into any c0(Γ), where Γ has any cardinality. Since the only property of κ0, which
is needed in [6], is the fact that (P (κ0)) holds, we deduce
Corollary 3.5. C(Kλ) does not coarsely (or uniformly) embed into c0(Γ), for
any cardinality Γ.
4 Proof of Theorem A
In this section we use our combinatorial Theorem B from Section 3 to show
Theorem A.
γ=0 such that x = PΓ
Recall that a long Schauder basis of a Banach space X is a transfinite se-
quence {eγ}Γ
γ=0 such that for every x ∈ X there exists a unique transfinite se-
quence of scalars {aγ}Γ
γ=0 aγ eγ. Similarly, a long Schauder
basic sequence in a Banach space X is a transfinite sequence {eγ}Γ
γ=0 which is a
long Schauder basis of its closed linear span. Recall that the w∗−dens(X ∗) is the
smallest cardinal such that there exists a w∗-dense subset of X ∗. Analogously
to the classical Mazur construction of a Schauder basic sequence in a separable
Banach space we have the following result, proved e.g.
in [4, p.135] (the fact
that the basis is normalized, i.e. keγk = 1, is not a part of the statement in [4],
but it is easy to get it by normalizing the existing basis).
Theorem 4.1. Let X be a Banach space with Γ = w∗ − densX ∗ > ω0. Then
X contains a monotone normalized long Schauder basic sequence of length Γ.
4 PROOF OF THEOREM A
9
Proof of Theorem A. Using the Hahn-Banach theorem it is easy to see that
w∗ − densX ∗ ≤ densX. On the other hand, since every x ∈ X is uniquely
determined by its values on a w∗-dense subset of X ∗, it is clear that
densX ≤ cardX ≤ 2w∗−densX ∗
It follows that for λ defined in (1) we get that λ = w∗ − densX ∗ holds if
and only if λ = densX. In order to prove Theorem A we may assume without
loss of generality that X has a long normalized and monotone Schauder basis
(eµ)µ<λ, of length λ, i.e. Γ = λ.
Set
Dn = {F ⊂ λ : F = n}, n ∈ N
Suppose that F = {γ1, . . . , γn} where γ1 < · · · < γn are elements of [0, λ)
arranged in an increasing order. Consider the corresponding finite set MF =
{Pn
i=1 εieγi : εi ∈ {−1, 1}}, containing 2n distinct vectors of X, and put a linear
order ≺ on this set according to the arrangement of the signs εi, setting
n
X
i=1
εieγi ≺
n
X
i=1
εieγi
if and only if for the minimal i, such that εi 6= εi, it holds εi < εi. In order
to prove Theorem A it suffices to show that if M = ∪F ∈Dn ,n∈NMF ⊂ X has
the coarse Stone property then X fails to have nontrivial cotype. To this end,
starting with U = {B2(x) : x ∈ M } we find a uniform bounded cover V, which
is point finite and so that U refines V, i.e., for all x ∈ M there is a Vx ∈ V with
B2(x) ⊂ Vx. Let r > 0 be such that each V ∈ V is a subset of a ball of radius r.
Let C be the set consisting of all finite tuples (V 1, . . . , V m), where V j ∈ V.
We now define the function σ : M → C as follows. If F ∈ Dn, F = {γ1, . . . , γn}
where γ1 < · · · < γn, we let
σ(F ) = (Vy1 , . . . , Vy2n ),
(9)
where y1 ≺ · · · ≺ y2n are the elements of MF arranged in the increasing
order. Applying Theorem B to the function σ, for a fixed n ∈ N, yields one of
two possibilities. Either there is an F = {γ1, . . . , γn−1}, where γ1 < · · · < γn−1,
so that σ(cid:0)(cid:8)F ∪ {τ } : τ ∈ λ \ F(cid:9)(cid:1) is infinite.
In this case, pick an infinite
sequence of distinct {τj}∞
j=1 witnessing the desired property. By passing to a
subsequence, we may assume without loss of generality that either there exists
k, 1 ≤ k ≤ n − 1, so that for all j ∈ N, γk < τj < γk+1, or τj < γ1 for all j ∈ N,
or γn−1 < τj for all j ∈ N. For simplicity of notation, assume the last case, i.e.
γ1 < · · · < γn−1 < τj holds for all j ∈ N. Denoting F j = {γ1, . . . , γn−1, τj}, we
conclude that there exists a fixed selection of signs ε1, . . . , εn such that the set
B = nVy : y =
n−1
X
i=1
εieγi + εnτj , j ∈ No
4 PROOF OF THEOREM A
10
is infinite.
Indeed, otherwise the set of values {σ({γ1, . . . , γn−1, τj}), j ∈ N},
which are determined by the definition (9), would have only a finite set of
options for each coordinate, and would therefore have to be finite. This is a
contradiction with the point finiteness of the system V, because
n−1
X
i=1
εieγi ∈ Vy,
for all Vy ∈ B.
It remains to consider the other case when there is a sequence (Fj) of pairwise
disjoint elements of [λ]n, with σ(Fi) = σ(Fj ), for any i, j ∈ N. In fact, it suffices
to choose just a pair of such disjoint elements (written in an increasing order
of ordinals) F = {γ1, . . . , γn}, G = {β1, . . . , βn}, such that σ(F ) = σ(G). This
means, in particular, that for every fixed selection of signs ε1, . . . , εn,
By our assumption, the elements of V are contained in a ball of radius r, hence
VPn
i=1 εieγi
= VPn
i=1 εieβi
k
n
X
i=1
εieγi −
n
X
i=1
εieβik ≤ 2r
(10)
holds for any selection of signs ε1, . . . , εn. Let uj = eγj − eβj , j ∈ {1, . . . , n}.
Because {eγ} is a monotonne normalized long Schauder basis, we have the trivial
estimate 1 ≤ kujk ≤ 2. The equation (10) means that
1 ≤ k
n
X
i=1
εiuik ≤ 2r
(11)
holds for any selection of signs ε1, . . . , εn. Since norm functions are convex, this
means that for the unite vector ball BE of of E = span(ui : i ≤) it follows that
n
n
X
j=1
ajuj : aj ≤
1
2ro ⊂ BE ⊂ n
n
X
j=1
ajuj : aj ≤ 2o,
which means that (uj)n
j=1 is 4r-equivalent to the unit vector basis of ℓn
∞.
In fact, our proof gives a much stronger condition than just failing cotype,
because our copies of ℓk
∞ are formed by vectors of the type eα − eβ. This fact
can be used to obtain much stronger structural results for spaces with special
bases. Recall that a long Schauder basis {eγ}Λ
γ=1 is said to be symmetric if
k
n
X
i=1
aieγik = k
n
X
i=1
aieβik
for any selection of ai ∈ R, and any pair of sets {γi}n
[1, Λ). It is well-known (c.f.
automatically unconditional, i.e. there exists K > 0 such that
i=1 ⊂
[9, Prop. II.22.2]), that each symmetric basis is
i=1 ⊂ [1, Λ), {βi}n
1
K
k
n
X
i=1
aieγik ≤ k
n
X
i=1
aieγik ≤ Kk
n
X
i=1
aieγik.
5 FINAL COMMENTS AND OPEN PROBLEMS
11
In particular,
whenever A ⊂ B.
1
K
kX
i∈A
aieγik ≤ kX
i∈B
aieγik
Theorem 4.2. Let X be a Banach space of density Λ ≥ λ, with a symmetric
basis {eγ}Λ
γ=1, which coarsely (or uniformly) embeds into some c0(Γ).
Then X is linearly isomorphic with c0(Λ).
Proof. By the proof of the above results, if X embeds into c0(Γ), there exists
an C > 0, such that for each k ∈ N there are some vectors {vi}k
i=1 of the form
vi = eγi − eβi satisfying the conditions
1
2
max
j
ai ≤ (cid:13)(cid:13)(cid:13)
k
X
i=1
aivi(cid:13)(cid:13)(cid:13)
≤ C max
j
aj.
(12)
Using the fact that the basis {eγ} is unconditional, (and symmetric) we
obtain by an easy manipulation that there exist some constants A, B > 0 such
that
k
X
i=1
A(cid:13)(cid:13)(cid:13)
aieγi(cid:13)(cid:13)(cid:13)
≤ (cid:13)(cid:13)(cid:13)
k
X
i=1
aivi(cid:13)(cid:13)(cid:13)
≤ B(cid:13)(cid:13)(cid:13)
k
X
i=1
aieγi(cid:13)(cid:13)(cid:13)
(13)
Combining (12) and (13) we finally obtain that for some D ≥ 1, and any
k ∈ N,
1
D
max
j
ai ≤ k
k
X
i=1
aieγik ≤ D max
j
aj.
(14)
for all {γ1, . . . , γk} ⊂ [1, Λ), which proves our claim.
5 Final comments and open problems
Let us mention in this final section some problems of interest.
First of all, we do not know whether or not Theorem A is true if we replace
λ by smaller cardinal numbers.
Problem 1. Assume that X is a Banach space with dens(X) ≥ ω1, and assume
that X coarsely embeds into c0(Γ) for some cardinal number Γ. Does X have
trivial co-type? If moreover X has a symmetric basis, must it be isomorphic to
c0(ω1)?
Of course Problem 1 would have a positive answer if the following is true.
Problem 2. Is Theorem B true for ω1?
Connected to Problems 1 and 2 is the following
REFERENCES
12
Problem 3. Does ℓ∞ coarsely embed into c0(κ) for some uncountable cardinal
number κ.
Another line of interesting problems asks which isomorphic properties do
non separable Banach spaces have which coarsely embed into c0(Γ)
Problem 4. Does a non separable Banach space which coarsely embeds into
some c0(Γ), Γ being uncountable, contain copies c0, or even c0(ω1).
References
[1] J. E. Baumgartner, Canonical partition relations, J. Symbolic Logic
40 (1975), 541 -- 554.
[2] J. Diestel, H. Jarchow, and A. Tonge, Absolutely Summing Operators,
Cambridge University Press (1993).
[3] G. Godefroy, G. Lancien, and V. Zizler, The non-linear geometry of
Banach spaces after Nigel Kalton, Rocky Mtn. J. Math. 44 (2014),
1529 -- 1583.
[4] P. H´ajek, V. Montesinos, J. Vanderwerff, and V. Zizler, Biorthogonal
Systems in Banach Spaces, CMS Books in Mathematics, Springer 2008.
[5] Th. Jech, Set Theory, Springer Monographs in Mathematics, fourth
edition, Springer Verlag, Berlin, Heidelberg, New York, 2006.
[6] J. Pelant, P. Holick´y, and O. Kalenda, C(K) spaces which cannot uni-
formly embedded into c0(Λ). Fund.Math. 192 (2006), 245 -- 254.
[7] J. Pelant, Embeddings into c0, Topol. Appl. 57 (1994), 259 -- 269.
[8] J. Pelant, and V. Rodl, On coverings of infinite-dimensional metric
spaces. Discrete .Math. 108 (1992), 75 -- 81.
[9] I. Singer, Bases in Banach spaces I, Grundlehren Math. Wiss. 154,
Springer-Verlag, Berlin (1970).
[10] A. Swift, On coarse Lipschitz embeddability into c0(κ), preprint, arX-
ive:1611.04623v1.
|
1204.0934 | 2 | 1204 | 2012-05-07T11:52:54 | Coherent states quantization of generalized bergman spaces on the unit ball of cn with a new formula for their associated berezin transforms | [
"math.FA",
"math-ph",
"math-ph"
] | While dealing with a class of generalized Bergman spaces on the unit ball, we construct for each of these spaces a set of coherent states to apply a coherent states quantization method. This provides us with another way to recover the Berezin transforms attached to these spaces. Finally, a new formula representing these transforms a functions of the Laplace-Beltrami operator is established in terms ofWilson polynomials by using the Fourier-Helgason transform. | math.FA | math |
COHERENT STATES QUANTIZATION OF GENERALIZED BERGMAN SPACES ON THE
UNIT BALL OF Cn WITH A NEW FORMULA FOR THEIR ASSOCIATED BEREZIN
TRANSFORMS
ABDELHAMID BOUSSEJRA & ZOUHAIR MOUAYN
ABSTRACT. While dealing with a class of generalized Bergman spaces on the unit ball, we construct
for each of these spaces a set of coherent states to apply a coherent states quantization method. This
provides us with another way to recover the Berezin transforms attached to these spaces. Finally,
a new formula representing these transforms a functions of the Laplace-Beltrami operator is estab-
lished in terms of Wilson polynomials by using the Fourier-Helgason transform.
1
INTRODUCTION
The Berezin transform introduced in [3] for certain classical bounded symmetric domains in
Cn is a transform linking the Berezin symbols and symbols for Toeplitz operators. It is present in
the study of the correspondence principle. The formula representing the Berezin transform as a
function of the Laplace operators ∆1, ..., ∆r ( r being the rank of the domain) plays a key role in the
Berezin quantization [4].
In this paper, we deal with the rank one symmetric domains. Namely the unit ball Bn in
(Cn,(cid:104),(cid:105)) endowed with its Bergman metric. We are precisely concerned with the L2-eigenspaces
(1.1)
ϕ ∈ L2(Bn, (1 − ξ2)−n−1dµ), Hν ϕ = ν,n
m ϕ
A2,ν
m (Bn) =
(cid:110)
(cid:111)
associated to the discrete spectrum
m = 4ν(2m + n) − 4m(m + n), m = 0, 1, 2, ..., [ν − n/2]
ν,n
of the Schrödinger operator with uniform magnetic field on Bn given by
(cid:32) n∑
(cid:0)δij−zizj
(cid:1)
(1.2)
(cid:33)
Hν = −4(1 − z2)
(1.3)
provided that ν > n/2. Above [x] denotes the greatest integer not exceeding x. For m ∈ Z+,
the Berezin transform associated with the space in (1.1) was obtained in [13] via the well known
formalism of Toeplitz operators as
∂zi∂zj
+ ν
i,j=1
j=1
(zj
) + ν2
+ 4ν2
∂2
n∑
∂
∂zj
− zj
∂
∂zj
Bν,n
m [ϕ] (z) =
m! (2ν − 2m − n) Γ (2ν − m) Γ(n)
n!Γ (2ν − m − n + 1) Γ(n + m)
P(n−1,2(ν−m)−n)
1 − 2ξ2(cid:17)(cid:17)2
×(cid:16)
(cid:16)
m
(cid:33)2(ν−m)
(cid:90)
(cid:16)
B
(1 − z2 (1 − ξ2)
(cid:32)
1 − ξ2(cid:17)n+1 dµ (ξ)
1 − (cid:104)z, ξ(cid:105)2
ϕ (ξ)
(1.4)
m
where P(α,β)
(.) denotes the Jacobi polynomial [16]. Moreover this transform have been expressed
as a function f (∆Bn ) of the Laplace-Beltrami operator ∆Bn in terms of an 3(cid:122)2-sum, see (5.24) be-
low. Our aim here is to construct for each of the eigenspaces in (1.1) a set of coherent states by
following a generalized formalism [11] in order to apply a coherent states quantization method.
Date: May 4, 2021.
1
2
A. BOUSSEJRA, Z. MOUAYN
This provides us with another way to recover the Berezin transforms in (1.4) attached to the L2-
eigenspace spaces in (1.1). Finally, we add a new formula expressing the transform (1.4) as a
function of the Laplace-Beltrami operator. The idea is to make the integral (1.4) appear as "con-
volution product" of the function ϕ with a specific radial function given in terms of the square
of a Jacobi polynomial. Next, a straightforward computation of the spherical transform of this
radial function with the use of a Clebsh-Gordon type linearisation [8] for the square of a Jacobi
polynomial amounts to a finite sum containing some integrals whose general form was given by
Koornwinder [17] in terms of Wilson polynomials.
This paper is summarized as follows. In Section 2, we recall briefly the formalism of coherent
states quantization we will be using. Section 3 deals with some needed facts on the generalized
Bergman spaces. In Section 4, we construct for each of these spaces a set of coherent states and we
apply the corresponding quantization scheme in order to recover their associated Berezin trans-
forms. In Section 5, we present the formula expressing these Berezin transforms as functions of
the Laplace-Beltrami operator by a different way and in a new form.
2 COHERENT STATES QUANTIZATION
Coherent states are mathematical tools which provide a close connection between classical and
quantum formalism. In general, they are a specific overcomplete set of vectors in a Hilbert space
satisfying a certain resolution of the identity condition. Here, we review a coherent states for-
malism starting from a measure space "as a set of data" as presented in [11]. Let X = {x x ∈ X}
(cid:9)+∞
be a set equipped with a measure dµ and L2(X, dµ) the space of dµ−square integrable functions
j=0. Let H be another
j=0. Then consider the family of states
on X. Let A2 be a subspace of L2(X, dµ) with an orthonormal basis(cid:8)Φj
(functional) space with a given orthonormal basis (cid:8)φj
(cid:9)+∞
{ x >}x∈X in H, through the following linear superposition:
where
x >:= (N (x))
− 1
2
Φj (x) φj >,
+∞
∑
j=0
N (x) =
+∞
∑
j=0
Φj (x) Φj (x).
These coherent states obey the normalization condition
(cid:104)x x(cid:105)H = 1
and the following resolution of the identity of H
1H =
x >< x N (x) dµ (x)
(2.1)
(2.2)
(2.3)
(2.4)
(2.5)
(cid:90)
X
(cid:90)
which is expressed in terms of Dirac's bra-ket notation x >< x meaning the rank-one -operator
ϕ (cid:55)→ (cid:104)ϕ x(cid:105)H . x > . The choice of the Hilbert space H define in fact a quantization of the space
X by the coherent states in (2.1), via the inclusion map x (cid:55)→ x >∈ H and the property (2.4) is cru-
cial in setting the bridge between the classical and the quantum mechanics. The Klauder-Berezin
coherent states quantization consists in associating to a classical observable that is a function f (x)
on X having specific properties the operator-valued integral
A f :=
x >< x f (x) N (x) dµ (x)
The function f (x) ≡ (cid:98)A f (x) is called upper (or contravariant) symbol of the operator A f and is
nonunique in general. On the other hand, the expectation value(cid:10)x A f x(cid:11) of A f with respect
X
to the set of coherent states { x >}x∈X is called lower ( or covariant) symbol of A f . Finally, asso-
ciating to the classical observable f (x) the obtained mean value(cid:10)x A f x(cid:11) , we get the Berezin
3
transform of this observable. That is,
B [ f ] (x) :=(cid:10)x A f x(cid:11) , x ∈ X.
(2.6)
(cid:18)a
(cid:19)
For all aspect of the theory of coherent states and their genesis, we refer to the survey [9] by
Dodonov or to the book by Gazeau [11].
3 THE SPACES A2,ν
m (Bn)
In this section, we review some results on the L2-concrete spectral analysis of the Schrödinger
operator Hν in (1.3) and acting in the Hilbert space L2(Bn, dµn),see [7], for more details. Let
Bn = {z ∈ Cn; z < 1} be the unit ball in Cn with the Lebesgue measure dµ normalized so that
µ(Bn) and let ∂Bn = {ω ∈ Cn, ω = 1} be the unit sphere with dσ the normalized measure on
it. Let G = SU(n, 1) be the group of all C-linear transforms g on Cn+1 that preserve the indefinite
j=1 zj 2 − zn+1 2, with det g = 1. Then G acts transitively on the unit ball by
hermitian form ∑n
G (cid:51) g =
(3.1)
As a homogeneous space we have the identification Bn = G/K where K = S(U(n) × U(1)) is
i,j ∂j∂j(Log(1 −
the stabilizer of 0. It is endowed with its usual Khaler-Bergman metric ds2 = − ∑n
z2))dzi ⊗ dzj. The Bergman distance and the volume element on Bn are given respectively by
: z → g.z = (az + b)(cz + d)−1.
b
c d
cosh2 d (z, w) =
1 − (cid:104)z, w(cid:105)2
(1− z 2)(1− w 2)
(3.2)
and dµn (z) = (1− z 2)−(n+1)dµ (z).
The group G acts unitarily on the space L2(Bn, dµn), via U(g)F(z) = F(g−1.z). Let consider the
magnetic gauge vector potential given through the canonical 1-form on Bn: θ = −i(∂− ∂)Log(1−
z 2), to which the Schrodinger operator
Hν = − (d + iνext (θ))
∗
(3.3)
can be associated. Here ν ≥ 0 is a fixed number d denotes the usual exterior derivative on
differential forms on Bn and ext(θ) is the exterior multiplication by θ while the symbol ∗ stands
for the adjoint operation with respect to the Hermitian scalar product induced by the Bergman
metric ds2 on differential forms. Note that when ν = 0, the operator in (3.3) reduces to
(d + iext (θ)) + 4ν2
which is the Laplace-Beltrami operator of the Bergman ball Bn. For general ν ≥ 0, the Schrodinger
operator Hν in (3.3) can be expressed in the complex coordinates (z1, ..., zn) by the formula (1.3)
see [2],[6] and [12].
Now, for an arbitrary complex number λ, a fundamental family of eigenfunctions of Hν with
eigenvalue λ2 + 4ν2 + n2 is given by the Poisson kernels :
(cid:18)
2 (iλ+1)(cid:32)
(cid:19) 1
(cid:33)ν
z (cid:55)→ Pν
λ(z, θ) =
1− z 2
1− < z, θ >2
1 − < z, θ >
1− < z, θ >
, z ∈ Bn.
(3.5)
Moreover, a complete description of the expansion of an eigenfunction f of Hν with eigenvalue
λ2 + 4ν2 + n2, in terms of the appropriate Fourier series in Bn have been given in [7, Proposition
H0 ≡ ∆Bn = 4(1− z 2)
∂2
∂zi∂zj
(3.4)
(cid:0)δij − zizj
(cid:1)
n∑
i,j=1
A. BOUSSEJRA, Z. MOUAYN
4
2.2]. Precisely,
f (z) = (1 − ρ2)
iλ+n
2
(cid:18) iλ + n
2
+∞
∑
p,q=0
ρp+q ·2 (cid:122)1
+ ν + p,
iλ + n
2
− ν + q, p + q + n; ρ2
(cid:19)
aλ,ν
p,q.hp,q(θ),
(3.6)
in C∞([0, 1[×∂Bn), z = ρθ, ρ ∈ [0, 1[ and θ = 1. Above 2F1 denotes the Gauss hypergeometric
function [14] and aλ,ν
p,q,j) ∈ Cd(n,p,q) are complex numbers, where
p,q = (aλ,ν
d(n, p, q) :=
(p + q + n − 1)(p + n − 2)!(q + n − 2)!
p!q!(n − 1)!(n − 2)!
(3.7)
is the dimension of the space H(p, q) of restrictions to the unit sphere ∂Bn of harmonic polynomi-
als h(z) on Cn, which are homogeneous of degree p in z and degree q in z, see [10] or [19] for more
details. The notation "." in (3.6) means the following finite sum
aλ,ν
p,q.hp,q(θ) =
d(n,p,q)∑
j=1
p,q,jhj
aλ,ν
p,q(θ),
(3.8)
where {hj
p,q}1≤j≤d(n,p,q) is an orthonormal basis of H(p, q). The spectral analysis of Hν have been
studied by many authors, see [7] and references therein. Actually, Hν is an elliptic densely de-
fined operator on the Hilbert space L2(Bn, (1 − (cid:104)z, z(cid:105))−(n+1)dµ) admitting a unique self-adjoint
realization also denoted by Hν. Its spectrum consists of a continuous part given by(cid:2)n2, +∞(cid:2) (cor-
responding to scattering states) and a finite number of infinitely degenerate eigenvalues ν,n
m given
by (1.2) (characterizing bound states) provided that 2ν > n. More precisely, ν,n
m + 4ν2 + n2,
with λm = i(2m + n − 2ν), m = 0, 1, ..., [ν − n/2]. Here, we focus on the discrete part of the spec-
trum, which is labeled by the integer m and the corresponding eigenspace A2,ν
m (Bn) defined in
(1.1). Taking into account (3.6) and expressing the involved hypergeometric in terms of Jacobi
polynomial, an orthonormal basis of A2,ν
m (Bn) can be given explicitly by
m = λ2
(cid:16)
1 − z2(cid:17)ν−m
(cid:16)
(cid:18) nΓ(2ν − m − n − q + 1)Γ(p + n + m)
P(n+p+q−1,2(ν−m)−n)
m−q
1 − 2z2(cid:17)
(cid:19)− 1
(m − q)!(2(ν − m) − n)Γ(2ν − m + p)
.
2
hj
p,q (z, z)
(3.9)
(3.10)
with
Φν,m,j
p,q
(z) = κν,m,n
p,q
κν,m,n
p,q =
for varying p = 0, 1, 2, ..., q = 0, 1, ...., m and j = 1, ...., d(n; p, q). Furthermore, the space A2,ν
m (Bn)
is a reproducing kernel Hilbert space. That is, there exists a unique complex valued function Kν,m
on Bn × Bn such that, denoting Kν,m
m (Bn) for any z ∈ Bn
and
for all functions f in A2,ν
the Bergman geodesic distance as
m (Bn) and all z ∈ Bn.Its expression can be given explicitly as function of
(w) = Kν,m(w, z), Kν,m
belongs to A2,ν
f (z) =< f , Kν,m
z >,
z
z
(cid:18) (1 − < z, w >)
1− < z, w >
(cid:19)ν
(3.11)
(1 − 2 tanh2 d(z, w))
Kν,m(z, w) =
(2 (ν − m) − n) Γ (2ν − m)
n!Γ (2ν − m − n + 1)
Remark 3.1. For m = 0, the space A2,ν
Bergman space of holomorphic function ψ on Bn satisfying the growth condition
0 (Bn) reduces further to be isomorphic to the weighted
×(cosh d(z, w))−2(ν−m))P(n−1,2(ν−m)−n)
(cid:90)
ψ (z)2 ((1 − (cid:104)z, z(cid:105))2ν−n−1dµ (z) < +∞.
m
Bn
This fact justify why the eigenspace A2,ν
index m.
m (Bn) have been also called a generalized Bergman spaces of
4 COHERENT STATES QUANTIZATION
5
Now, to adapt the defintion (2.1) of coherent states for the context of the generalized Bergman
spaces in (1.1) we first list the following notations.
1 − z2(cid:17)−(n+1)
(cid:16)
(cid:111)
m (Bn) ⊂ L2(Bn,
is the orthonormal basis of A2,ν
(cid:19)
1 − z2(cid:17)−n−1
dµ).
dµ
Bn,
(cid:18)
• (X, dη) :=
• x ≡ z ∈ Bn.
• A2 := A2,ν
(cid:16)
• {Φk (x)} ≡(cid:110)Φν,m
• {ϕk} ≡(cid:8)ϕp,q,j
• N (x) ≡ N (z) is a normalization factor.
(cid:9) is an orthonormal basis of another (functional) Hilbert space H.
m (Bn) in (3.8)
p,q,j (z)
, dη ≡ dµn is the volume element on Bn.
Definition 4.1. For each fixed integer m = 0, 1, ..., [n − ν/2] , a class of generalized coherent states asso-
ciated with the space A2,ν
m (Bn) is defined according to (2.1) by the form
z ≡ z, ν, m >:= (N (z))
φν,m
∑
− 1
2
Φν,m
p,q,j (z) ϕp,q,j
0≤q≤m,0≤p<+∞
1≤j≤d(n,p,q)
where N (z) is a normalization factor.
Proposition 4.1. The factor in (4.1) is given by
N (z) =
for every z ∈ Bn.
(2(ν − m) − n) Γ (2ν − m)
n!Γ (2ν − m − n + 1)
Γ (m + n)
m!Γ (n)
Proof. To calculate this factor, we start by writing the condition
(cid:104)φν,m
z
, φν,m
z
(cid:105)H = 1.
Equation (4.3) is equivalent to
(N (z))
−1
+∞
∑
p=0
m∑
q=0
d(n,p,q)∑
j=1
Φν,m
p,q,j (z) Φν,m
p,q,j (z) = 1
Making use of (3.9) and (3.11) for the particular case z = w, we get that
P(n−1,2(ν−m)−n)
(2(ν − m) − n) Γ (2ν − m)
n!Γ (2ν − m − n + 1)
Next, by the following fact on Jacobi polynomial [14]:
N (z) =
m
(1)
(4.1)
(4.2)
(4.3)
(4.4)
(4.5)
P(α,β)
m
(1) =
Γ (m + α + 1)
m!Γ (α + 1)
(4.6)
z ≡ z, ν, m > satisfy the resolution
for α = n − 1 to arrive at the announced result.The states φν,m
of the identity
(cid:90)
1H =
z, ν, m >< z, ν, m N (z) dν
(4.7)
and with the help of them we can achieve the coherent states quantization scheme described in
Sec.2 to rederive the Berezin transform Bν,n
m in (1.4) which was defined by Toeplitz operators
Bn
6
A. BOUSSEJRA, Z. MOUAYN
formalism in [13]. For this let us associate to any arbitrary function ϕ ∈ L2(Bn, (1 − ξ2)−n−1dµ)
the operator-valued integral
Aϕ :=
z, ν, m >< z, ν, m ϕ (z) N (z) (1 − z2)−n−1dµ
(4.8)
(cid:90)
Bn
The function ϕ (z) is a upper symbol of the operator Aϕ. On the other hand, we need to calculate
the expectation value
E{z,ν,m>}(cid:0)Aϕ
(cid:1) :=< z, ν, m Aϕ z, ν, m >
(4.9)
of Aϕ with respect to the set of coherent states { z, ν, m >}z∈Bn defined in (4.1). This will consti-
tute a lower symbol of the operator Aϕ.
Proposition 4.2. Let ϕ ∈ L2(Bn, (1 − ξ2)−n−1dµ). Then, the expectation value in (4.9) has the follow-
ing expression
Γ (n) m! (2 (ν − m) − n) Γ (2ν − m)
n!Γ (n + m) Γ (2ν − m − n + 1)
(cid:16)
P(n−1,2(ν−m)−n)
1 − 2ξ2(cid:17)(cid:17)2
×(cid:16)
(cid:16)
m
(cid:33)2(ν−m)
(4.10)
(1 − z2 (1 − ξ2)
(cid:32)
(cid:90)
1 − ξ2(cid:17)n+1 dµ (ξ)
1 − (cid:104)z, ξ(cid:105)2
ϕ (ξ)
B
E{z,ν,m>}(cid:0)Aϕ
(cid:1) =
for every z ∈ Bn.
Proof. We first write the action of the operator Aϕ in (4.8) on an arbitrary coherent state z, ν, m >
in terms of Dirac's bra-ket notation as
Aϕ z, ν, m >=
(cid:90)
Bn
Therefore, the expectation value reads
< z, ν, m Aϕ z, ν, m >
=
< z, ν, m w, ν, m > < z, ν, m w, ν, m >
w, ν, m >< w, ν, m z, ν, m >
(cid:90)
(cid:90)
< z, ν, m w, ν, m >2
Bn
ϕ (w)
=
Bn
N (w)
(1 − w2)n+1
dµ (w)
(4.11)
N (w)
(1 − w2)n+1
dµ (w)
N (w)
(1 − w2)n+1
dµ (w) .
(4.12)
(4.13)
Now, we need to evaluate the quantity < z, ν, m w, ν, m >2 in (4.13). For this, we write the
scalar product as
< z, ν, m w, ν, m >=
+∞
∑
p=0
m∑
q=0
d(n;p,q)∑
m∑
+∞
∑
r=0
s=0
j=1
(cid:10)ϕp,q,j, ϕp,q,l
l=1
d(n;p,q)∑
(cid:11)
Recalling that
since(cid:8)ϕp,q,j
(cid:9) is an orthonormal basis of H, the above sum in (4.14) reduces to
H = δj,lδp,rδq,s
Φν,m
p,q,j (z) Φν,m
(cid:112)N (z) N (w)
p,q,j (w)
(cid:10)ϕp,q,j, ϕp,q,l
(cid:11)
H
(4.14)
(4.15)
< z, ν, m w, ν, m >= (N (z) N (w))
− 1
2
∑
0≤q≤m,0≤p<+∞
1≤j≤d(n,p,q)
p,q,j (z) Φν,m
Φν,m
p,q,j (w).
(4.16)
(cid:32)
− 1
2
1 − (cid:104)z, w(cid:105)
1 − (cid:104)z, w(cid:105)
(cid:33)ν
1 − 2 tanh2 (d (z, w))
7
(4.17)
(cid:17)
.
< z, ν, m w, ν, m > =
(2 (ν − m) − n) Γ (2ν − m)
Now, taking account (3.9) and (3.11), Equation ( 4.16) takes the form
(N (z) N (w))
(cid:16)
−2(ν−m) P(n−1,2(ν−m)−n)
So that the square modulus of the scalar product in (4.17) reads
n!Γ (2ν − m − n + 1)
× (cosh (d (z, w)))
m
< z, ν, m w, ν, m >2 =
n!Γ (2ν − m − n + 1)
(cid:18) (2 (ν − m) − n) Γ (2ν − m)
−4(ν−m)(cid:16)
(cid:18) (2 [ν − m] − n) Γ (2ν − m)
(cid:19)2
P(n−1,2(ν−m)−n)
(cid:19)2
× (cosh (d (z, w)))
m
(cid:16)
(N (z) N (w))
−1
(4.18)
(cid:17)(cid:17)2
.
(4.19)
1 − 2 tanh2 (d (z, w))
−1
(N (z) N (w))
(cid:16)
1 − 2 tanh2 (d (z, w))
(1 − w2)n+1
N (w)
(cid:17)(cid:17)2
dµ (w) ,
Returning back to (4.12) , we get
E{z,ν,m>}(cid:0)Aϕ
(cid:1) =
(cid:90)
Bn
ϕ (w)
n!Γ (2ν − m − n + 1)
−4(ν−m)(cid:16)
× (cosh (d (z, w)))
(cid:90)
(cid:18) (2 (ν − m) − n) Γ (2ν − m)
P(n−1,2(ν−m)−n)
(cid:19)2
m
which can be also written as
E{z,ν,m>}(cid:0)Aϕ
(cid:1) =
Bn
ϕ (w)
n!Γ (2ν − m − n + 1)
−4(ν−m)(cid:16)
−4(ν−m)(cid:16)
m
× (cosh (d (z, w)))
(2(ν − m) − n) Γ (2ν − m) m!Γ (n)
n!Γ (2ν − m − n + 1) Γ (m + n)
× (cosh (d (z, w)))
(cid:90)
Bn
=
−1
(N (z))
(cid:16)
(1 − w2)n+1
ϕ (w)
(1 − w2)n+1
(cid:16)
P(n−1,2(ν−m)−n)
1 − 2 tanh2 (d (z, w))
(4.20)
dµ (w)
(4.21)
(cid:17)(cid:17)2
(cid:17)(cid:17)2
1 − 2 tanh2 (d (z, w))
Finally, we summarize the above discussion by considering the following definition.
Definition 4.3. The Berezin transform of the classical observable ϕ ∈ L2(Bn, (1 − ξ2)−n−1dµ) con-
structed according to the quantization by the coherent states { z, ν, m >} in (4.1) is obtained by associat-
ing to ϕ the obtained mean value in (4.10) . That is,
P(n−1,2(ν−m)−n)
dµ (w) .
m
m [ϕ] (z) = E{z,ν,m>}(cid:0)Aϕ
(cid:1)
Bν,n
for every z ∈ Bn.
Remark 4.4. For m = 0, the transform (4.10) reduces to the well known Berezin transform attached
to the weighted Bergman space A2,ν
0 (Bn) of holomorphic function ψ on Bn satisfying the growth
condition (3.12) and given by
Bν,n
0
[ϕ] (z) =
(2ν − n) Γ (2ν)
n!Γ (2ν − n + 1)
(cosh d (z, ξ))
−4ν
The latter one have also been written as a function of the Bergman Laplacian ∆Bn as
ϕ (ξ)
1 − ξ2(cid:17)n+1 dµ (ξ)
(cid:16)
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:112)−∆Bn − n2
(4.22)
(4.23)
(4.24)
(cid:90)
Bn
(cid:18)
(cid:12)(cid:12)(cid:12)(cid:12)Γ
Bν,n
0 =
1
Γ (α + 1) Γ (α + n + 1)
α + 1 +
n
2
+
i
2
firstly by Berezin. The above form, involving gamma factors, was derived by Peetre in [18, p. 182],
so that α there occurring in the weight of the Bergman space, corresponds to 2ν − n − 1.
8
A. BOUSSEJRA, Z. MOUAYN
Then Berezin transform Bν,n
Bν,n
m [ϕ] (z) = cν,n
m
with
Bn
cν,n
m =
where Wk(.) are Wilson polynomials,
γν,n,m
k
=
and the coefficients Aν,n,m
k
2m!Γ (n) (2 (ν − m) − n) Γ (2ν − m) (−1)k
Γ (n + m) Γ (2ν − m − n + 1) k!Γ2 (2 (ν − m) + k)
are given by (5.10) below.
× Aν,n,m
k
,
Proof. Recall that
Bν,n
where
m [ϕ] (z) = cν,n
(cid:16)
By this way, we have to compute the spherical transform F [hν,n
1 − ξ2(cid:17)2(ν−m)(cid:16)
(cid:90)
m (ϕ ∗ hν,n
m )(g),
P(n−1,2(ν−m)−n)
hν,n
m (ξ) :=
(cid:16)
m
z = g.0,
1 − 2ξ2(cid:17)(cid:17)2
m ] of hν,n
m (z)φ−λ(z)dµn(z), λ ∈ R
hν,n
F [hν,n
m ](λ) :=
, ξ ∈ Bn,
m . Namely
where φλ is the spherical function associated to ∆Bn, given by
Bn
φλ(z) = (1− z )
iλ+n
2
2F1(
Using Pfaff's transformation [14]
iλ + n
,
iλ + n
2
2
(cid:18)
2(cid:122)1 (a, b, c; x) = (1 − x)
−b
2(cid:122)1
b, c − a, c;
we rewrite φ−λ as
φ−λ(z) = 2F1
(cid:18)−iλ + n
2
,
iλ + n
2
, n;
z 2
z 2 −1
, n; z 2).
(cid:19)
x
x − 1
(cid:19)
5 AN EXPRESSION OF Bν,n
m AS FUNCTION OF ∆Bn
m associated the generalized Bergman space A2,ν
P(n−1,2(ν−m)−n)
1 − 2 tanh2 d (z, ξ)
m
(cid:17)(cid:17)2
(cid:16)
(cid:16)
(cid:90)
(cosh d (z, ξ))4(ν−m)
m (Bn) is given by
(cid:16)
dµ (ξ)
1 − ξ2(cid:17)n+1 ,
ϕ (ξ)
Γ (n) m! (2 (ν − m) − n) Γ (2ν − m)
n!Γ (n + m) Γ (2ν − m − n + 1)
m (z, w) be the kernel function of the above integral operator and set hν,n
Let Bν,n
z = g.0. Then the integral operator (5.1) can be written as a convolution product over G:
m (g) = Bν,n
m (z, 0),
Bν,n
m [ϕ] (z) = cν,n
m (ϕ ∗ hν,n
m )(g),
z = g.0,
from which it follows easily that the Berezin operator is an L2-bounded operator.
Since Bν,n
U(g) ◦ Bν,n
function of the G-invariant Laplacian ∆Bn of the unit ball. Below we give it explicitly.
m (z, w) is a G bi-invariant function it follows that Bν,n
m ◦ U(g), for every g ∈ G. Therefore Bν,n
m is a G-invariant operator. That is
m is, in the spectral theoretic sense, a
m = Bν,n
Theorem 5.1.The Berezin transform Bν,n
∆Bn as
m can be expressed as a function of the Laplace-Beltrami operator
(cid:18)
(cid:12)(cid:12)(cid:12)(cid:12)Γ
2m∑
k=0
Bν,n
m =
+
i
2
2 (ν − m) − n
2
∆Bn − n2
4
k Wk(−1
γν,m,n
4
(cid:112)−∆Bn − n2
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2
; 2(ν − m) − n
2
, 1 +
n
2
,
n
2
,
n
2
)
(5.1)
(5.2)
(5.3)
(5.4)
(5.5)
(5.6)
(5.7)
So that returning back to (5.5) we get
F [hν,n
m ](λ) = 2n
(cid:16)
1(cid:90)
0
ρ2n−1
(1 − ρ2)n+1−2(ν−m)
n − iλ
(cid:18) n + iλ
,
2
2
×2(cid:122)1
P(n−1,2(ν−m)−n)
m
(cid:19)
, n;
ρ2
ρ2 − 1
dρ
(cid:0)1 − 2ρ2(cid:1)(cid:17)2
9
(5.8)
To calculate this last integral, we first use a linearisation of the square of Jacobi polynomial in (5.8)
by making appeal to the following Clebsh-Gordon type formula see [8, p. 611],
P(κ,)
s
(u) P(τ,η)
(5.9)
for the particular case of parameters s = l = m, κ = τ = α = n − 1, = η = 2 (ν − m) − n and
δ = 2 (ν − m) . In our setting, the linearisation coefficients As,l (k) are of the form
(u) =
(u)
k=0
k
l
As,l (k) P(α,δ)
s+l∑
Aν,n,m
k
=
2:1
×(cid:122)2:2
−2m,−2m − n + 1 : 1 − 2ν, 1 − 2ν
(2 (ν − m) + n)k (n)2m (2k + 2 (ν − m) + n) (−1)k (2m)! ((2 (ν − m))2m)2
(cid:19)
(cid:18) −2m + k,−2ν − k − n : −m,−n − m + 1;−m,−m − n + 1
(n)k (2 (ν − m) + n)2m+k+1 (m!)2 (2m − k)! ((2 (ν − m))m)2
(cid:2)cp(cid:48)(cid:3)
(cid:18) (cid:0)ap
(cid:1) :(cid:0)bp(cid:48)(cid:1) ,(cid:0)cp(cid:48)(cid:1)
(cid:1)
(cid:0)aj
(5.11)
(dl) : (κl(cid:48) ) , (l(cid:48) )
s in which (x)s = x (x + 1) ... (x + s − 1) is the Pochhammer symbol.
l:l(cid:48) (.) denotes the Kampé de Fériet double hypergeometric function defined by [20, p. 63]
[dl]q+s [κl(cid:48) ]q [l(cid:48) ]s
(cid:2)bp(cid:48)(cid:3)
+∞
∑
q,s=0
(cid:2)ap
(cid:122)p:p(cid:48)
l:l(cid:48)
x, y
1, 1
(5.10)
(cid:19)
xq
q!
ys
s!
(cid:3)
q+s
=
q
s
Here (cid:122)p:p(cid:48)
(cid:3)
where (cid:2)ap
Therefore, inserting(cid:16)
s = ∏p
j=1
(cid:0)1 − 2ρ2(cid:1)(cid:17)2
P(n−1,2(ν−m)−n)
m
=
2m∑
k=0
Aν,n,m
k
P(n−1,2(ν−m))
k
(cid:0)1 − 2ρ2(cid:1)
into (5.8) the Fourier-Helgason transform of hν,n
m takes the form
2m∑
Aν,n,m
m ](λ) =
Iν,n,m
k
k
(λ) ,
k=0
where the last term in this sum is defined by the integral
F [hν,n
1(cid:90)
Iν,m
k
(λ) =
2nρ2n−1
(1 − ρ2)n+1−2(ν−m) P(n−1,2(ν−m))
(cid:18)1
(n − iλ) , n;
×2F1
(n + iλ) ,
0
k
1
2
2
(cid:0)1 − 2ρ2(cid:1)
(cid:19)
dρ
ρ2
ρ2 − 1
(cid:17)
(5.12)
(5.13)
(5.14)
To calculate this last integral we make the change of variable ρ = tanh t. Therefore (5.14) reads
Iν,m
k
(λ) =
+∞(cid:90)
2n (sinh t)2n−1 P(n−1,2(ν−m))
k
0
× (cosh t)
−4(ν−m)+1 .2F1
(cid:16)
(cid:18) n + iλ
1 − 2 tanh2 t
n − iλ
,
2
2
, n;− sinh2 t
(5.15)
(cid:19)
dt
10
A. BOUSSEJRA, Z. MOUAYN
Now, we make use of the result established by Koornwinder [17]
0
=
k
2
2
dt
(cid:17)
× Wk
×2 F1
1 − 2 tanh2 t
, α + 1;− sinh2 t
(cid:16)
+∞(cid:90)
(cosh t)−α+β−δ−µ(cid:48)−1 (sinh t)2α+1 P(α,δ)
(cid:19)
(cid:18) α + β + 1 + iλ
, α + β + 1 − iλ
2 (δ + µ(cid:48) + 1 − iλ)(cid:1)
2 (δ + µ(cid:48) + 1 + iλ)(cid:1) Γ(cid:0) 1
Γ (α + 1) (−1)k Γ(cid:0) 1
2 (α − β + δ + µ(cid:48) + 2) + k(cid:1)
2 (α + β + δ + µ(cid:48) + 2) + k(cid:1) Γ(cid:0) 1
k!Γ(cid:0) 1
(cid:18)1
(cid:48) + 1(cid:1) ,
(cid:0)δ + µ
(cid:48) + 1(cid:1) ,
(cid:0)δ − µ
(cid:0)x2, a, b, c, d(cid:1) := (a + b)k (a + c)k (a + d)k
(cid:12)(cid:12)(cid:12)(cid:12)Γ
(cid:18) −k, k + a + b + c + d − 1, a + ix, a − ix
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2
2nΓ (n) (−1)k
a + b, a + c, a + d
(α + β + 1) ,
×4 F3
+ i λ
2
4 λ2;
Iν,m
k
(λ) =
(cid:18)
Wk
1
2
1
2
1
2
1
2
(cid:19)
1
where β, δ, λ ∈ R, α, δ > −1 , δ + (cid:60)(µ)(cid:48) > −1 and Wk (.) is the Wilson polynomial given in terms
of the 4F3-sum as ([1],p. 158):
(5.16)
(cid:19)
(α − β + 1)
(5.17)
(5.18)
(5.19)
for the parameters α = n − 1, δ = 2(ν − m) − n, β = 0 and µ(cid:48) = 2(ν − m) − n − 1.We find that
Summarizing the above calculations
F [hν,n
m ] (λ) =
k!Γ2 (2 (ν − m) + k)
×Wk
(cid:18)1
4 λ2; 2 (ν − m) − n
2
, 1 +
,
,
n
2
n
2
n
2
2 (ν − m) − n
(cid:19)
2
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)2
+ i λ
2
2 (ν − m) − n
2
(cid:18)
(cid:12)(cid:12)(cid:12)(cid:12)Γ
(cid:32)
(cid:33)
× 2m∑
k=0
γν,n,m
k Wk
2
λ
4
; 2 (ν − m) − n
2
, 1 +
n
2
,
n
2
,
n
2
with the constants
γν,n,m
k
where the constants Aν,n,m
announced result.
k
:=
2m!Γ (n) (2 (ν − m) − n) Γ (2ν − m) (−1)k Aν,n,m
Γ (n + m) Γ (2ν − m − n + 1) k!Γ2 (2 (ν − m) + k)
is given by (5.10). Finally, replacing λ by(cid:112)−∆Bn − n2, we arrive at the
(5.20)
,
k
Remark 5.1. Setting m = 0 in the formula (5.3) in Theorem 5.1, we recover the result of Peetre [18].
Remark 5.2. We should note that the transform Bν,n
the Laplace-Beltrami operator ∆Bn in terms of the 3(cid:122)2-sum as
m have been expressed in [13] as a function of
Γ(cid:16)
(cid:16)
n − i(cid:112)−∆Bn − n2(cid:17)(cid:17)
Γ(cid:16)
(cid:16)
n + i(cid:112)−∆Bn − n2(cid:17)(cid:17)
2 (ν − m) − 1
n + i(cid:112)−∆Bn − n2(cid:17)
(cid:16)
1
2 (ν − m) + j + 1
n + i(cid:112)−∆Bn − n2(cid:17)
(cid:16)
(ν − m) + j + 1
, n + j, 1
2
(cid:16)
2
2
2
n + i(cid:112)−∆Bn − n2(cid:17)
, n
2
Bν,n
m =
2m∑
j=0
Cν,n,m
j
×3 (cid:122)2
(5.21)
1
where
Cν,n,m
j
=
(2 (ν − m) − n) Γ (n + m) (−1)j Γ (n + j)
m!Γ (2ν − n − m + 1) Γ (2ν − n)
× min(m,j)∑
p=max(0,j−m)
(m!)2 Γ (2ν − m) Γ (2ν − m + j − p)
(j − p)! (m + p − j)!p! (m − p)!Γ (n + j − p) Γ (n + p)
REFERENCES
11
(5.22)
.
[1] G. E. Andrews, R. Askey and R. Roy, Special Functions, Encyclopedia of Mathematics and its Applications, Cam-
[2] Ayaz, K.and Intissar, A.: Selberg trace formula for the heat and wave kernels of Maass Lpalacians of the complex
bridge University Press, 1999
hyperbolic space Hn (C) , n ≥ 2. Diff. Geom. App. 15, 1-31 (2001)
[3] Berezin, F.A.: General concept of quantization. Commun. Math. Phys. 40, 153-174 (1975)
[4] Berezin, F.A.: Quantization. Math. USSR, Izv. 8, 1109-1165 (1974)
[5] Berezin, F.A.: Quantization in complex symmetric spaces. Izv. Akad. Naud.SSSR. Ser.Mat(2) 39, 1109-1165 (1974)
[6] Bonami, N. Bruna, J. and Grellier, S.: On Hardy, BMO and Lipschitz spaces of invariantly harmonic functions in
the unit ball. Proc. London Math. Soc. 3, 665-696 (1998)
[7] Boussejra, A. and Intissar, A. L2-concrete Spectral Analysis of the Invariant Laplacian ∆
α,β in the Unit complex ball
Bn, J. Funct. Anal 160, 115-140 (1998)
[8] Chaggara, H. and Koepf, W.: On linearization coefficients of Jacobi polynomials. App. Math. Lett. 23, 609-614 (2010)
[9] Dodonov, V. V. : 'Noncalssical' states in quantum optics: a 'squeezed review of the first 75 years, J.opt.B:Quantum
[10] Folland, G. B.: Spherical harmonic expansion of the Poisson-Szegö kernel for the ball. Proc.Am. Math. Soc. 47 (2) ,
Semiclass.opt.4, R1-R33 (2002)
401-407 (1975)
[11] Gazeau, J.P.: Coherent states in quantum physics. Wiley, Weinheim (2009)
[12] Geller, D.: Some results in Hp theory for the Heisenberg group, Duke Math. J. 47, 365-391 (1980)
[13] Ghanmi, A. and Mouayn, Z.: A formula representing magnetic Berezin transforms on the Bergman ball of Cn as
functions of the Laplace-Beltrami operator, arxiv: 1101.1248v1 (2011)
[14] Gradshteyn I S and Ryzhik I M, "Table of Integrals, Series and Products", Academic Press, INC, Seven Edition
[15] Helgason, S.: A duality for symmetric spaces with applications to group representations. Advances in math., 5,
2007
1-154 (1970)
[16] Mourad E.H.Ismail.: Classical and Quantum Orthogonal Polynomials in one variable, Encyclopedia of Mathemat-
ics and its applications, Cambridge university press (2005)
[17] Koornwinder T.H.: Special orthogonal polynomial systems mapped onto each other by the Fourier-Jacobi trans-
form. Orthogonal polynomials and applications (Bar-le-Duc, 1984), 174-183, Lecture Notes in Math. 1171, Springer,
Berlin, 1985
[18] Peetre, J.: The Berezin transform and Ha-plitz operator. J. Operator Theory. 24, 165-186 (1990)
[19] Rudin, W.: Function theory in the unit ball of Cn. Grundlehern der mathematischen Wissenschften, vol.241.
[20] Srivastava, H. M. and Manocha A.: A treatise on Generating functions, John Willey and Sons, New York, Chich-
[21] Unterberger, A., Upmeir, H.: The Berezin transform and invariant differential operators. Commun. Math. Phys. 164,
Springer, New York (1980)
ester, Brisbane, Toraonto (1984)
563-597 (1994)
[22] Zernik, F. and Brinkman, H.C.: Hyperspharische Functionen und die in spahrischen bereichen orthogonalen poly-
nome. Proc. Kon. Akad. v. wet. Amsterdam 38, 161-170 (1935)
(A. B.) DEPARTMENT OF MATHEMATICS, FACULTY OF SCIENCES,
IBN TOFAIL UNIVERSITY, KENITRA, MOROCCO
E-mail address: [email protected]
(Z. M.) DEPARTMENT OF MATHEMATICS, FACULTY OF SCIENCES AND TECHNICS,
SULTAN MOULAY SLIMANE UNIVERSITY,BP. 123, BENI MELLAL, MOROCCO
E-mail address: [email protected]
|
1702.03201 | 1 | 1702 | 2017-02-10T15:12:23 | Kernel theorems for modulation spaces | [
"math.FA",
"math.NA"
] | We deal with kernel theorems for modulation spaces. We completely characterize the continuity of a linear operator on the modulation spaces $M^p$ for every $1\leq p\leq\infty$, by the membership of its kernel to (mixed) modulation spaces. Whereas Feichtinger's kernel theorem (which we recapture as a special case) is the modulation space counterpart of Schwartz' kernel theorem for temperate distributions, our results do not have a couterpart in distribution theory. This reveals the superiority, in some respects, of the modulation space formalism upon distribution theory, as already emphasized in Feichtinger's manifesto for a post-modern harmonic analysis, tailored to the needs of mathematical signal processing. The proof uses in an essential way a discretization of the problem by means of Gabor frames. We also show the equivalence of the operator norm and the modulation space norm of the corresponding kernel. For operators acting on $M^{p,q}$ a similar characterization is not expected, but sufficient conditions for boundedness can be sated in the same spirit. | math.FA | math |
KERNEL THEOREMS FOR MODULATION SPACES
ELENA CORDERO AND FABIO NICOLA
Abstract. We deal with kernel theorems for modulation spaces. We completely
characterize the continuity of a linear operator on the modulation spaces M p for
every 1 ≤ p ≤ ∞, by the membership of its kernel to (mixed) modulation spaces.
Whereas Feichtinger's kernel theorem (which we recapture as a special case) is
the modulation space counterpart of Schwartz' kernel theorem for temperate
distributions, our results do not have a couterpart in distribution theory. This
reveals the superiority, in some respects, of the modulation space formalism
upon distribution theory, as already emphasized in Feichtinger's manifesto for
a post-modern harmonic analysis, tailored to the needs of mathematical signal
processing. The proof uses in an essential way a discretization of the problem by
means of Gabor frames. We also show the equivalence of the operator norm and
the modulation space norm of the corresponding kernel. For operators acting
on M p,q a similar characterization is not expected, but sufficient conditions for
boundedness can be sated in the same spirit.
1. Introduction
Schwartz' kernel theorem certainly represents one of the most important results
in modern Functional Analysis. It states, in the framework of temperate distribu-
tions, that every linear continuous operator A : S(Rd) → S ′(Rd) can be regarded
as an integral operator in a generalized sense, namely
hAf, gi = hK, g ⊗ fi
f, g ∈ S(Rd),
in the sense of distributions, for some kernel K ∈ S ′(R2d), and viceversa [20]. We
write, formally,
(1)
Af (x) =ZRd
K(x, y) f (y)dy
f ∈ S(Rd).
This result provides a general big framework including most operators of interest
in Harmonic Analysis. However, for most applications in Time-frequency Analysis
and Mathematical Signal Processing the spaces S(Rd) and S ′(Rd) can be fruit-
fully replaced by the modulation spaces M 1(Rd) and M ∞(Rd), respectively, intro-
duced by H. Feichtinger in [13] and nowadays widely used as a fundamental tool in
2010 Mathematics Subject Classification. 42B35, 42C15, 47G30, 81Q20.
Key words and phrases. Time-frequency analysis, Gabor frames, modulation spaces,
1
2
ELENA CORDERO AND FABIO NICOLA
Time-frequency Analysis (see [18] and also Feichtinger's manifesto for a postmod-
ern Harmonic Analysis [17]). One of the obvious advantages is that M 1(Rd) and
M ∞(Rd) are Banach spaces; secondly their norm gives a direct and transparent
information on the time-frequency concentration of a function or temperate distri-
bution. The same holds, more generally, for a whole scale of intermediate spaces
M p(Rd), 1 ≤ p ≤ ∞, as well as the more general mixed norm spaces M p,q(Rd),
1 ≤ p, q ≤ ∞.
In short, for (x, ξ) ∈ Rd × Rd, define the time-frequency shifts π(x, ξ) by
π(x, ξ)f (t) = e2πitξf (t − x),
f ∈ S ′(Rd).
Then M p,q(Rd) is the space of temperate distributions f in Rd such that the function
(x, ξ) 7→ hf, π(x, ξ)gi,
for some (and therefore any) window g ∈ S(Rd)\{0}, is in Lq(Rd
x)), endowed
with the obvious norm (see Section 2 below for details). We also set M p(Rd) =
M p,p(Rd). Weighted versions are also used in the literature but here we limit
ourselves, for simplicity, to the unweighted case.
ξ; Lp(Rd
Now, a kernel theorem in the framework of modulation spaces was announced in
[9] and proved in [14]; see also [18, Theorem 14.4.1]. It states that linear continuous
operators A : M 1(Rd) → M ∞(Rd) are characterized by the memenbership of thier
distribution kernel K to M ∞(R2d).
In this note we provide a similar characterization of linear continuous operators
acting on the following spaces:
• M 1(Rd) → M p(Rd), for a fixed 1 ≤ p ≤ ∞
• M p(Rd) → M ∞(Rd), for a fixed 1 ≤ p ≤ ∞
• M p(Rd) → M p(Rd), for every 1 ≤ p ≤ ∞.
In all cases the characterization is given in terms of the membership of their
distribution kernels to certain mixed modulation spaces, first introduced and studied
in [3]. The proof uses in an essential way a discretization of the problem by means
of Gabor frames. We also show the equivalence of the operator norm and the
modulation space norm of the corresponding kernel.
These results seem remarkable, because no such a characterization is known,
for example, for linear continuous operators from S(Rd) into itself and similarly
from S ′(Rd) into itself. This again shows the superiority, in some respects, of the
modulation space framework upon distribution theory.
For operators acting on mixed-norm modulation spaces M p,q(Rd) → M p,q(Rd)
such a characterization is not expected, as we will see in Section 3 below. We have
however sufficient conditions for boundedness in the same spirit.
KERNEL THEOREMS FOR MODULATION SPACES
3
Let us also observe that in the literature there are plenty of sufficient conditions
for operators in certain classes (Fourier multipliers, localization operators, pseudo-
differential operators, Fourier integral operators) to be bounded M p(Rd) → M p(Rd)
for every 1 ≤ p ≤ ∞, or M p,q(Rd) → M p,q(Rd) for every 1 ≤ p, q ≤ ∞. For ex-
ample it is known that pseudodifferential operators with Weyl symbol in M ∞,1 are
in fact bounded M p,q(Rd) → M p,q(Rd) for every 1 ≤ p, q ≤ ∞ [6, 18, 19, 26, 27]
(see also [24]). On the other hand the Fourier multipier with symbol eiξ2
is still
bounded on these spaces [2], although that symbol does not belong to M ∞,1(R2d).
Similary, it is easy to see that any linear continuous operator A : S(Rd) → S ′(Rd)
whose Gabor matrix enjoys decay estimates such as
hAπ(z)g, π(w)gi ≤ C(1 + w − χ(z))−s
z, w ∈ R2d
for some s > 2d, g ∈ S(Rd) \ {0}, and bi-Lipschtiz mapping χ : R2d → R2d, is
bounded M p(Rd) → M p(Rd) for every 1 ≤ p ≤ ∞ [4] (see also [4, 7, 8]). But once
again this condition is not necessary at all.
The very simple characterization provided in the present paper when p = q
therefore includes, at least implicitly, all these continuity results on M p.
2. Backgrounds on time-frequency analysis
2.1. Mixed norm spaces. We summarize the main properties of mixed norm
spaces, we refer to [1] for a full treatment of this subject.
Consider measure spaces (Xi, µi), indices pi ∈ [1,∞], with i = 1, . . . , d. Then
the mixed norm space Lp1,p2,...,pd(X1 × X2 × · · · × Xd; µ1 × µ2 × · · · × µd) consists
of the measurable functions F : X1 × X2 × · · · × Xd → C such that the following
norm is finite:
(2)
kFkLp1,p2,...,pd = ZXd · · ·(cid:18)ZX1 F (x1, . . . , xd)p1 dµ1(x1)(cid:19)
1
pd
p2
p1
· · · dµd(xd)!
with standard modification when pi = ∞ for some i.
If Xi = R and µi is the Lebesgue measure on R for every i = 1, . . . , d, we simply
write Lp1,...,pd. If each Xi is a countable set and the corresponding measure µi is
the counting measure, we use the notation ℓp1,...,pd(X1 × X2 × · · · × Xd).
Banach spaces, generalizations of the classical Lp and ℓp spaces, cf. [1].
2.2. Modulation spaces. For x, ξ ∈ Rd, define the translation operator Tx and
modulation operator Mξ by
The mixed norm spaces Lp1,p2,...,pd(X1 × X2 × · · · × Xd; µ1 × µ2 × · · · × µd) are
Txf (t) = f (t − x) Mξf (t) = e2πitxf (t).
For z = (x, ξ), the composition operator π(z) = MξTx is called a time-frequency
shift of the phase space R2d.
4
ELENA CORDERO AND FABIO NICOLA
For a fixed non-zero g ∈ S(Rd), the short-time Fourier transform (STFT) of
f ∈ S ′(Rd) with respect to the window g is given by
(3)
f (t) g(t − x) e−2πiξt dt = hf, π(z)gi
Vgf (x, ξ) =ZRd
z = (x, ξ),
where the integral is intended as the (anti-)duality between S ′ and S. The short-
time Fourier transform gives information about the time-frequency content of the
signal f : indeed, roughly speaking, it can be seen as a localized Fourier transform
of f in a neighbourhood of x.
The STFT is the basic tool in the definition of modulation spaces.
Indeed,
modulation space norms are a measure of the joint time-frequency distribution of
f ∈ S ′. For their basic properties we refer [10, 11, 12] (see also [3, 21, 22, 23]) and
the textbook [18].
Definition 2.1. Fix g ∈ S(Rd), p = (p1, . . . , pd), q = (q1, . . . , qd), pi, qi ∈ [1,∞],
i = 1, . . . , d. Then
M p,q(Rd) = {f ∈ S ′(Rd) : kfkM p,q(Rd) < ∞},
where
kfkM p,q(Rd) = kVgfkLp1,...,pd ,q1,...,qd .
q = q1, . . . , qd, we come back to the classical modulation
If p = p1, . . . , pd,
spaces of distributions f ∈ S ′(Rd) such that
kfkM p,q(Rd) = ZRd(cid:18)ZRd Vgf (x, ξ)p dx(cid:19)
q
p
1
q
dξ!
< ∞
[13]). The extension of the original modulation spaces in Definition 2.1 was
(cf.
widely studied and applied to the investigation of boundedness and sampling prop-
erties for pseudodifferential operators in [21, 22, 23].
To state our kernel theorems we need a further extension of Definition 2.1, that
was introduced by S. Bishop in [3], in her study of Schatten p-class properties
for integral operators. This is a natural generalization of the classical modulation
spaces, as will be clear in the sequel.
We shall mainly use the notation in [3]. From now on we assume that c is a
permutation of the set {1, . . . , 2d}. We identify c with the linear bijection
c : R2d → R2d,
c(x1, . . . , x2d) = (xc(1), . . . , xc(2d)).
Definition 2.2. Consider g ∈ S(Rd) and let c be a permutation corresponding
to the map c as above. Then M(c)p1,p2,...,p2d is the mixed modulation space of
tempered distributions f ∈ S ′(Rd) for which
(4)
kfkM (c)p1,p2 ,...,p2d = kVgf ◦ ckLp1,p2 ,...,p2d < ∞.
KERNEL THEOREMS FOR MODULATION SPACES
5
If p = p1 = p2 = · · · = pd,
q = pd+1 = pd+2 = · · · = p2d we simply write
M(c)p,q; if p = p1 = p2 = · · · = pd = pd+1 = pd+2 = · · · = p2d, we shorten to M(c)p.
Remark 2.3.
(i) If c is the identity permutation and the indices satisfy
p = p1 = p2 = · · · = pd,
q = pd+1 = pd+2 = · · · = p2d
then M(c)p1,p2,...,p2d = M p,q(Rd) is the original modulation space introduced
by Feichtinger in [13].
(ii) If c is the identity permutation and p = (p1, . . . , pd), q = (q1, . . . , qd), we
come back to the modulation spaces in Definition 2.1.
(iii) If the indices satisfy
p = p1 = p2 = · · · = pd = pd+1 = pd+2 = · · · = p2d,
given any permutation c, a simple change of variables yields
M(c)p = M p(Rd).
(iv) In general the mixed modulation spaces differ from the classical ones of
Definition 2.1. A simple example is provided by the permutation
(5)
c0(1, 2, . . . , d, d + 1, . . . , 2d) = (d + 1, d + 2, . . . , 2d, 1, 2, . . . , d).
If we choose
p = p1 = p2 = · · · = pd,
then it is easy to see that
q = pd+1 = pd+2 = · · · = p2d
M p,q(c0) = W (F Lp, Lq)(Rd),
that is the space of distributions f ∈ S ′(Rd) such that
kfkW (F Lp,Lq)(Rd) = ZRd(cid:18)ZRd Vgf (x, ξ)p dξ(cid:19)
dx!
q
p
1
q
< ∞
for some window function g ∈ S(Rd) \ {0}.
It can be shown (cf. [3]) that the definition of mixed modulation spaces is in-
dependent of the choices of the window g in S(Rd), with different windows giv-
ing equivalent norms. Furthermore, this fact also holds for g in the larger space
M 1(Rd), as follows by the subsequent Theorem 2.4, which gives the inversion for-
mula on M(c)p1,...,p2d.
Theorem 2.4. Suppose c is a permutation of {1, . . . , 2d}, and for a measurable
function ψ : Rd → C define the operator Υψ by
(6)
F (x)π(c(x))ψ(t) dx.
ΥψF (t) =ZR2d
6
ELENA CORDERO AND FABIO NICOLA
(i) Given ψ, g ∈ M 1(Rd), p1, . . . , p2d ∈ [1,∞], and f ∈ M(c)p1,...,p2d, we have
Υψ(Vgf ◦ c) = hψ, gif.
(ii) Different windows in M 1(Rd) define equivalent norms on M(c)p1,...,p2d.
Corollary 2.5. For any p1, . . . , p2d ∈ [1,∞], M(c)p1,...,p2d is a Banach space.
We also have the expected duality result.
Theorem 2.6. If p1, . . . , p2d ∈ [1,∞), M(c)p′
where p′
i ∈ (1,∞] satisfies 1/pi + 1/p′
i = 1.
1,...,p′
2d is the dual space of M(c)p1,...,p2d,
2.3. Gabor frames. Given a lattice Λ in R2d and a non-zero square integrable
function g (called window) on Rd the system
G(g, Λ) = {π(λ) : λ ∈ Λ}
is called a Gabor frame if it is a frame for L2(Rd), that is there exist constants
0 < A ≤ B such that
Akfk2
(7)
hf, π(λ)gi2 ≤ Bkfk2
2,
∀f ∈ L2(Rd).
2 ≤Xλ∈Λ
If (7) holds, then there exists a γ ∈ L2(Rd) (so-called dual window), such that
G(γ, Λ) is a frame for L2(Rd) and every f ∈ L2(Rd) can be expanded as
(8)
f =Xλ∈Λ
hf, π(λ)giπ(λ)γ =Xλ∈Λ
hf, π(λ)γiπ(λ)g,
with unconditional convergence in L2(Rd).
From now on we shall work with lattices Λ = αZ2d, for suitable α > 0.
If the window function belongs to the Feichtinger's algebra M 1(Rd), then a Gabor
frame for L2(Rd) is a frame also for mixed modulation spaces (cf. [3, Theorem 4.6]).
Theorem 2.7. Let g be a window function in M 1(Rd) such that G(g, Λ) is a frame
for L2(Rd) with a dual frame G(γ, Λ). Consider p1, . . . , p2d ∈ [1,∞]. Then
(i) There exist constants 0 < A ≤ B independent of p1, . . . , p2d such that
AkfkM (c)p1 ,...,p2d ≤ kVgf ◦ cΛkℓp1,...,p2d ≤ BkfkM (c)p1,...,p2d ,
(ii) We have
∀f ∈ M(c)p1,...,p2d.
(9)
f =Xλ∈Λ
hf, π(λ)giπ(λ)γ =Xλ∈Λ
hf, π(λ)γiπ(λ)g,
with unconditional convergence in M(c)p1,...,p2d if p1, . . . , p2d ∈ [1,∞), weak*
convergence in M ∞(Rd) if p1, . . . , p2d ∈ [1,∞].
KERNEL THEOREMS FOR MODULATION SPACES
7
This generalizes the basic fact, valid for the identity permutation, that the coef-
ficient operator Cg and the reconstruction operator Dγ, defined by
(10)
satisfy
(Cgf )λ := hf, π(λ)gi
Dγc :=Xλ∈Λ
cλπ(λ)γ,
Cg : M p,q(Rd) → ℓp,q(Λ), Dγ : ℓp,q(Λ) → M p,q(Rd),
(11)
with DγCg = I, the identity on M p,q(Rd), cf. (9), and similarly if we interchange
the windows g and γ.
3. Kernel Theorems
In this section we will present the characterization of continuity properties on
modulation spaces of integral operators (1).
We need the following result (see e.g. Tao's Lecture Notes [25, Proposition 5.2]).
Proposition 3.1. Assume that (X, νX), (Y, µY ) are measure space and p ∈ [1,∞].
Suppose that K : X ×Y → C, and kK(·, y)kLp(X) is uniformly bounded on Y . Then
the operator A with kernel K maps L1(X) into Lp(Y ) and
(12)
kAkL1(X)→Lp(Y ) = ess sup
y∈Y
kK(·, y)kLp(X).
Similarly, we have a dual statment (see e.g. [25, Proposition 5.4]).
Proposition 3.2. Assume that (X, νX), (Y, µY ) are measure space and p ∈ [1,∞].
Suppose that K : X × Y → C, and kK(x,·)kLp′ (Y ) is uniformly bounded on Y .
Then the operator A with kernel K maps Lp(X) into L∞(Y ) and
(13)
kAkLp(X)→L∞(Y ) = ess sup
x∈X kK(x,·)kLp′ (Y ).
Let us introduce the permutation c1 of the set {1, . . . , 4d} with related c1 given
Recall that we only consider lattices of the form Λ = αZ2d.
Theorem 3.3. Suppose p ∈ [1,∞]. Let g be a window function in S(Rd) such
that G(g, Λ) is a frame for L2(Rd) with a dual frame G(γ, Λ). A linear continuous
operator A : S(Rd) → S ′(Rd) is bounded M 1(Rd) → M p(Rd) if and only if its
distrinution kernel K satisfies
(15)
VGK ◦ c1 ∈ ℓp,∞(Λ × Λ),
by, for xi ∈ Rd, i = 1, . . . , 4,
c1(x1, x2, x3, x4) =
(14)
Id 0d 0d 0d
Id 0d
0d 0d
Id 0d 0d
0d
0d 0d 0d
Id
x1
x2
x3
x4
= (x1, x3, x2, x4).
8
ELENA CORDERO AND FABIO NICOLA
where G = g ⊗ ¯γ.
Moreover
kAkM 1→M p ≍ kVGK ◦ c1kℓp,∞(Λ×Λ).
Proof. Assume first (15). For λ = (λ1, λ2), µ = (µ1, µ2) ∈ Λ, observe that the
Gabor matrix of the operator A is given by
(16)
Kλ,µ := hAπ(µ)γ, π(λ)gi
= hK, π(λ1, λ2)g ⊗ π(µ1,−µ2)¯γi
= VGK ◦ c1(λ1, λ2, µ1,−µ2).
This matrix turns out to be the integral kernel of the operator A := CgADγ, acting
on sequences on Λ, where the coefficient operator Cg and synthesis operator Dγ
are defined in (10). By (15) and Proposition 3.1, A is a bounded operator ℓ1(Λ) →
ℓp(Λ). Since Cg : M p(Rd) → ℓp(Λ) and Dγ : ℓp(Λ) → M p(Rd) continuously, we
have A = Dγ ACγ : M 1(Rd) → M p(Rd), with
kAkM 1→M p ≤ C sup
µ∈Λ kK·,µkℓp(Λ).
Vice-versa, consider the finite dimensional space
EN = {(aλ)λ∈Λ : aλ = 0 for λ > N}
and the natural projection PN : ℓ∞(Λ) → EN . Consider the composition operator
(17)
PN CgADγEN : (EN ,k · kℓ1) → (EN ,k · kℓp).
This operator is represented by the (finite) matrix
{Kλ,µ}λ,µ∈Λ,λ≤N,µ≤N .
Using formula (12) in Proposition 3.1 (we work on finite sets, so that the assump-
tions are trivially satisfied) we have
sup
µ∈Λ,µ≤N kKλ,µkℓp(Λ,λ≤N ) = kPN CgADγENkℓ1→ℓp ≤ kCgkkDγkkAkM 1→M p.
Since the right-hand side is independent of N, taking the supremum with respect
to N, we obtain
This concludes the proof.
µ∈Λ kK·,µkℓp(Λ) ≤ CkAkM 1→M p.
sup
We observe that, given a frame G(g, Λ) for L2(Rd), Λ = αZ2d, with dual frame
G(γ, Λ), then G(¯γ, Λ) is still a frame for L2(Rd) and G(g ⊗ ¯γ, Λ × Λ) is a frame for
L2(R2d). Using Theorem 2.7 we can therefore reformulate the previous result in
terms of mixed modulation spaces.
KERNEL THEOREMS FOR MODULATION SPACES
9
Corollary 3.4. Suppose p ∈ [1,∞]. A linear continuous operator A : S(Rd) →
S ′(Rd) is bounded from M 1(Rd) into M p(Rd) if and only if its distribution kernel
K satisfies
(18)
where c1 is identified with (14).
Moreover
K ∈ M(c1)p,∞,
kAkM 1→M p ≍ kKkM (c1)p,∞.
Remark 3.5. When p = ∞, condition (18) becomes K ∈ M ∞(R2d) and we recap-
ture Feichtinger's kernel Theorem [14].
A similar characterization of bounded operators A : M p(Rd) → M ∞(Rd) can be
obtained arguing as above, using Proposition 3.2, (16) and Theorem 2.7. Precisely,
define the permutation c2 of the set {1, . . . , 4d} with related c2 by
(19)
= (x3, x1, x4, x2),
c2(x1, x2, x3, x4) =
0d 0d
Id 0d
Id 0d 0d 0d
Id
0d 0d 0d
0d
Id 0d 0d
x1
x2
x3
x4
for xi ∈ Rd, i = 1, . . . , 4.
Theorem 3.6. Suppose p ∈ [1,∞]. A linear continuous operator A : S(Rd) →
S ′(Rd) is bounded from M p(Rd) into M ∞(Rd) if and only if its distribution kernel
K satisfies
(20)
where 1/p + 1/p′ = 1.
Moreover
K ∈ M(c2)p′,∞,
kAkM p→M ∞ ≍ kKkM (c2)p′ ,∞.
As a consequence of Theorems 3.3 and 3.6 we have the following result.
Corollary 3.7. Let A : S(Rd) → S ′(Rd) be an operator with distribution kernel
K ∈ S ′(R2d). The following conditions are equivalent:
(i) A : M p(Rd) → M p(Rd), for every 1 ≤ p ≤ ∞.
(ii) A : M 1(Rd) → M 1(Rd) and A : M ∞(Rd) → M ∞(Rd).
(iii) K ∈ M(c1)1,∞ ∩ M(c2)1,∞.
Proof. All the equivalences follow trivially from Theorems 3.3 and 3.6 but (ii) ⇒
(i), which follows by complex interpolation for modulation spaces [13].
10
ELENA CORDERO AND FABIO NICOLA
4. Operators acting on M p,q
Let us now consider the problem of boundedness of an integral operator on
M p,q(Rd), 1 ≤ p, q ≤ ∞. We do no longer expect a characterization as in the case
p = q, studied in the previous section (see Remark 4.2 below). However sufficient
conditions in the same spirit can be stated.
Consider the permutations c3 and c4 of the set {1, . . . , 4d} with related c3, c4
defined by
(21)
and
(22)
c3(x1, x2, x3, x4) =
c4(x1, x2, x3, x4) =
Id 0d 0d
0d
0d 0d
Id 0d
Id 0d 0d 0d
0d 0d 0d
Id
Id 0d 0d 0d
0d 0d 0d
Id
Id 0d 0d
0d
0d 0d
Id 0d
x1
x2
x3
x4
x1
x2
x3
x4
= (x2, x3, x1, x4),
= (x1, x4, x2, x3),
for xi ∈ Rd, i = 1, . . . , 4.
contained in [5, Proposition 2.4]).
The following result was proved in [7, Proposition 5.1] (a continuous version is
Proposition 4.1. Consider an at most countable index set J and the operator
defined on sequences (aλ1,λ2) on the set Λ = J × J by
(Aa)λ1,λ2 = Xµ1,µ2∈J
Kλ1,λ2,µ1,µ2aµ1,µ2.
kAkℓ∞,1→ℓ∞,1 ≤ kK ◦ c3kℓ1,∞,1,∞.
(i) If K ◦ c3 ∈ ℓ1,∞,1,∞(J × J × J × J), then A is continuous on ℓ∞,1(Λ). Moreover,
(23)
(ii) If K ◦ c4 ∈ ℓ1,∞,1,∞(J × J × J × J), then A is continuous on ℓ1,∞(Λ), with
(24)
(iii) If K satisfies assumptions (i) and (ii) and moreover K ◦ c0, K ∈ ℓ1,∞(Λ × Λ),
where the permutation c0 is defined in (5), then the operator A is continuous on
ℓp,q(Λ), for every 1 ≤ p, q ≤ ∞.
Remark 4.2. Observe that the reverse inequalites in (23), (24) do not hold, even
if one allows a multiplicative constant. Consider for example (23), with J =
{0, 1, . . . , N − 1}, N ∈ N, where
kAkℓ1,∞→ℓ1,∞ ≤ kK ◦ c4kℓ1,∞,1,∞.
Kλ1,λ2,µ1,µ2 = Fλ2,µ1 =
e−2πiλ2µ1
1
√N
KERNEL THEOREMS FOR MODULATION SPACES
11
is the Fourier matrix. We have
so that
but
(K ◦ c3)µ1,λ1,λ2,µ2 = Kλ1,λ2,µ1,µ2
kK ◦ c3kℓ1,∞,1,∞ = kFkℓ1 = N 3/2
(25)
which blows-up, as N → +∞, at a lower rate.
kAkℓ∞,1→ℓ∞,1 ≤ N,
Let us verify (25). We have
λ1kℓ1
λ2
kk Xµ1,µ2∈J
Kλ1,λ2,µ1,µ2aµ1,µ2kℓ∞
= kXµ1∈J
= kXµ1∈J
with aµ1 :=Pµ2∈J aµ1,µ2. Now, using the embeddings
kbkℓ2(J) ≤
√Nkbkℓ2(J),
kbkℓ1(J) ≤
aµ1,µ2kℓ1
λ2
Fλ2,µ1 Xµ2∈J
Fλ2,µ1aµ1kℓ1
λ2
√Nkbkℓ∞(J)
and the fact that F is a unitary transformation of ℓ2(J) we get (25).
As a consequence, we obtain the following boundedness result.
Consider the permutations c5 and c6 of the set {1, . . . , 4d} with related c5, c6
defined by
(26)
and
(27)
c5 = c1c3 =
c6 = c1c4 =
0d
Id 0d 0d
Id 0d 0d 0d
0d 0d
Id 0d
Id
0d 0d 0d
Id 0d 0d 0d
Id 0d 0d
0d
Id
0d 0d 0d
0d 0d
Id 0d
.
Theorem 4.3. (i) A linear continuous operator A : S(Rd) → S ′(Rd) is bounded
on M ∞,1(Rd) if its distribution kernel K satisfies
K ∈ M(c5)1,∞,1,∞.
(28)
(ii) A linear continuous operator A : S(Rd) → S ′(Rd) is bounded on M 1,∞(Rd) if
its distribution kernel K satisfies
(29)
K ∈ M(c6)1,∞,1,∞.
12
ELENA CORDERO AND FABIO NICOLA
Proof. The result is a consequence of (16), Proposition 4.1 and Theorem 2.7, by
using the same arguments as in the first part of the proof of Theorem 3.3.
By interpolation we also obtain the following result.
Corollary 4.4. Let A : S(Rd) → S ′(Rd) be an operator with distribution kernel
K ∈ M(c5)1,∞,1,∞ ∩ M(c6)1,∞,1,∞.
Then A : M p′,p(Rd) → M p′,p(Rd) continuously, for every 1 ≤ p ≤ ∞, with 1/p +
1/p′ = 1.
By combining Corollaries 3.7 and 4.4, we finally obtain the following result.
Corollary 4.5. Let A : S(Rd) → S ′(Rd) be an operator with distribution kernel
K ∈ M(c1)1,∞ ∩ M(c2)1,∞ ∩ M(c5)1,∞,1,∞ ∩ M(c6)1,∞,1,∞.
Then A : M p,q(Rd) → M p,q(Rd) continuously, for every 1 ≤ p, q ≤ ∞.
Acknowledgments
This research was partially supported by the Gruppo Nazionale per l'Analisi
Matematica, la Probabilit`a e le loro Applicazioni (GNAMPA) of the Istituto Nazionale
di Alta Matematica (INdAM), Italy.
References
[1] A. Benedek and R. Panzone. The space Lp with mixed-norm. Duke Math. J., 28:301–324,
1961.
[2] A. B´enyi, K. Grochenig, K.A. Okoudjou and L.G. Rogers. Unimodular Fourier multipliers
for modulation spaces. J. Funct. Anal., 246(2):366-384, 2007.
[3] S. Bishop. Mixed modulation spaces and their applications to pseudodifferential operators.
J. Math. Anal. Appl., 363:255–264, 2010.
[4] E. Cordero, K. Grochenig, F. Nicola and L. Rodino. Wiener algebras of Fourier integral
operators. J. Math. Pures Appl., 99:219–233, 2013.
[5] E. Cordero and F. Nicola. Boundedness of Schrodinger Type Propagators on Modulation
Spaces. J. Fourier Anal. Appl., 16:311-339, 2010.
[6] E. Cordero, F. Nicola. Pseudodifferential operators on Lp, Wiener amalgam and modulation
spaces. Intern. Math. Res. Not., 10:1860–1893, 2010.
[7] E. Cordero, F. Nicola and L. Rodino. Time-frequency analysis of Fourier integral operators.
Commun. Pure Appl. Anal., 9(1):1–21, 2010.
[8] E. Cordero, F. Nicola and L. Rodino. Sparsity of Gabor representation of Schrodinger prop-
agators. Appl. Comput. Harmon. Anal., 26(3):357–370, 2009.
[9] H.G. Feichtinger. Un espace de Banach de distributions temp´er´ees sur les groupes localement
compacts ab´eliens. Compt. Rend. Acad. Sci. Paris, Ser. A, 290(17):791–794, 1980.
[10] H. G. Feichtinger. Banach spaces of distributions of Wiener's type and interpolation,
in
"Proc. Conf. Functional Analysis and Approximation," Oberwolfach August 1980, Internat.
Ser. Numer. Math., Birkhauser, Boston, 69:153–165, 1981.
KERNEL THEOREMS FOR MODULATION SPACES
13
[11] H. G. Feichtinger. Banach convolution algebras of Wiener type, in "Functions, series, opera-
tors," Vol. I, II, (Budapest, 1980), North-Holland, Amsterdam, 509–524, 1983.
[12] H. G. Feichtinger. Generalized amalgams, with applications to Fourier transform, Canad. J.
Math., 42:395–409, 1990.
[13] H.G. Feichtinger. Modulation spaces on locally compact abelian groups, Technical Report,
University Vienna, 1983, and also in Wavelets and Their Applications, M. Krishna, R.
Radha, S. Thangavelu, editors, Allied Publishers, 99–140, 2003.
[14] H.G. Feichtinger, K. Grochenig. Gabor frames and time-frequency analysis of distributions,
J. Funct. Anal., 146(2):464–495, 1997.
[15] H.G. Feichtinger and T. Strohmer (eds.). Gabor Analysis and Algorithms: Theory and Ap-
plications. Birkhuser Boston, Boston, 1998.
[16] H.G. Feichtinger and T. Strohmer (eds.). Advances in Gabor analysis. Applied and Numerical
Harmonic Analysis. Birkhauser Boston Inc., Boston, 2003.
[17] H.G. Feichtinger. Elements of Postmodern Harmonic Analysis, in Operator-related Function
Theory and Time-Frequency Analysis. The Abel Symposium 2012, Oslo, Norway, August
20–24, 2012, Springer, 77–105, 2015.
[18] K. Grochenig. Foundations of Time-Frequency Analysis. Applied and Numerical Harmonic
Analysis, Birkhauser, Boston, MA, 2001.
[19] K. Grochenig. Time-frequency analysis of Sjstrand's class. Revista Mat. Iberoam., 22(2):703-
724, 2006.
[20] L. Hormander. The Analysis of Linear Partial Differential Operators I, Distribution Theory
and Fourier Analysis, Springer-Verlag, Berlin, second edition 1990.
[21] S. Molahajloo, K.A. Okoudjou and G. Pfander. Boundedness of Multilinear Pseudo-
differential Operators on Modulation Spaces. J. Fourier Anal. Appl., 22(6):1381–1415, 2016.
[22] S. Molahajloo and G. Pfander. Boundedness of pseudo-differential operators on Lp, Sobolev
and modulation spaces. Math. Model. Nat. Phenom., 8(1):175–192, 2013.
[23] G. Pfander. Sampling of operators. J. Fourier Anal. Appl., 19(3):612–650, 2013.
[24] J. Sjostrand. An algebra of pseudodifferential operators. Math. Res. Lett., 1(2):185–192, 1994.
[25] T. Tao. Lecture Notes 2 for 247 A: Fourier Analysis.
[26] J. Toft. Continuity properties for modulation spaces, with applications to pseudo-differential
www.math.ucla.edu/∼tao/247a.1.06f/notes2.pdf
calculus. I. J. Funct. Anal., 207(2):399–429, 2004.
[27] J. Toft. Continuity properties for modulation spaces, with applications to pseudo-differential
calculus. II. Ann. Global Anal. Geom., 26(1):73–106, 2004.
Universit`a di Torino, Dipartimento di Matematica, via Carlo Alberto 10, 10123
Torino, Italy
E-mail address: [email protected]
Dipartimento di Scienze Matematiche, Politecnico di Torino, corso Duca degli
Abruzzi 24, 10129 Torino, Italy
E-mail address: [email protected]
|
1911.10255 | 2 | 1911 | 2019-12-10T14:41:23 | On $C$-compact orthogonally additive operators | [
"math.FA"
] | We consider $C$-compact orthogonally additive operators in vector lattices. After providing some examples of $C$-compact orthogonally additive operators on a vector lattice with values in a Banach space we show that the set of those operators is a projection band in the Dedekind complete vector lattice of all regular orthogonally additive operators. In second part of the article we introduce a new class of vector lattices, called $C$-complete, and show that any laterally-to-norm continuous $C$-compact orthogonally additive operator from a $C$-complete vector lattice to a Banach space is narrow, which generalizes a result of Pliev and Popov. | math.FA | math |
ON C-COMPACT ORTHOGONALLY ADDITIVE OPERATORS
MARAT PLIEV and MARTIN R. WEBER
Abstract. We consider C-compact orthogonally additive operators in vector
lattices. After providing some examples of C-compact orthogonally additive
operators on a vector lattice with values in a Banach space we show that the
set of those operators is a projection band in the Dedekind complete vector
lattice of all regular orthogonally additive operators.
In the second part of
the article we introduce a new class of vector lattices, called C-complete, and
show that any laterally-to-norm continuous C-compact orthogonally additive
operator from a C-complete vector lattice to a Banach space is narrow, which
generalizes a result of Pliev and Popov.
1. Introduction
Orthogonally additive operators in vector lattices first were investigated in
[12]. Later these results were extended in [1, 2, 6, 7, 8, 20, 22, 23]). Recently,
some connections with problems of the convex geometry were revealed [24, 25].
Orthogonally additive operators in lattice-normed spaces were studied in [3]. In
this paper we continue this line of research. We analyze the notion of C-compact
orthogonally additive operator.
In the first part of the article we show that
set of all C-compact orthogonally additive operators from a vector lattice E to
an order continuous Banach lattice F is a projection band in the vector lattice
of all regular orthogonally additive operators from E to F (Theorem 3.9).
In
the final part of the paper we introduce a new class of vector lattices which we
call C-complete (the precise definition is given in section 4) and prove that any
laterally-to-norm continuous C-compact orthogonally additive operator from an
atomless C-complete vector lattice E to a Banach space X is narrow (Theorem
4.7). This is a generalization of the result of the article [18, Theorem 3.2].
Note that linear narrow operators in function spaces first appeared in [17].
Nowadays the theory of narrow operators is a well-studied object of functional
2010 Mathematics Subject Classification. Primary 47H30; Secondary 47H99.
Key words and phrases. C-compact operator, orthogonally additive operator, narrow oper-
ator, Urysohn operator, C-complete vector lattice, Banach lattice.
1
2
MARAT PLIEV and MARTIN R. WEBER
analysis and is presented in many research articles ([9, 11, 14, 15, 18, 23]) and in
the monograph [21].
2. Preliminaries
All necessary information on vector lattices one can find in [4]. In this article
all vector lattices are assumed to be Archimedean.
Two elements x, y of a vector lattice E are called disjoint (written as x ⊥ y),
if x ∧ y = 0. An element a > 0 of E is an atom if 0 ≤ x ≤ a, 0 ≤ y ≤ a
and x⊥y imply that either x = 0 or y = 0. The equality x =
n
xi means that
n
Fi=1
x =
xi and xi⊥xj for all i 6= j. In the case of n = 2 we write x = x1 ⊔ x2. An
Pi=1
element y of a vector lattice E is called a fragment (or a component) of x ∈ E, if
y ⊥ (x − y). The notation y ⊑ x means that y is a fragment of x. Two fragments
x1 and x2 of an element x are said to be mutually complemented if x = x1 ⊔ x2.
The set of all fragments of an element x ∈ E is denoted by Fx.
Definition 2.1. Let be E a vector lattice and X a real vector space. An operator
T : E → X is said to be orthogonally additive if T (x + y) = T x + T y for any
disjoint elements x, y ∈ E.
It is not hard to check that T (0) = 0. The set of all orthogonally additive
operators from E to X is a real vector space with respect to the natural linear
operations.
Definition 2.2. Let E and F be vector lattices. An orthogonally additive (in
general, nonlinear) operator T : E → F is said to be:
• positive if T x ≥ 0 for all x ∈ E,
• regular, if T = S1 − S2 for two positive, orthogonally additive operators
S1 and S2 from E to F .
The sets of all positive, regular orthogonally additive operators from E to F are
denoted by OA+(E, F ), OAr(E, F ), respectively, where the order in OAr(E, F )
is introduced as follows: S ≤ T whenever (T − S) ≥ 0. Then OAr(E, F ) becomes
an ordered vector space.
For a Dedekind complete vector lattice F we have the following property of
OAr(E, F ).
Proposition 2.3 ([19, Theorem 3.6]). Let E and F be a vector lattices, and
assume that F is Dedekind complete.Then OAr(E, F ) is a Dedekind complete
ON C-COMPACT OPERATORS
3
vector lattice. Moreover, for every S, T ∈ OAr(E, F ) and every x ∈ E the
following formulas1 hold
(T ∨ S)(x) = sup{T y + Sz : x = y ⊔ z};
(T ∧ S)(x) = inf{T y + Sz : x = y ⊔ z};
(1)
(2)
(3) T +(x) = sup{T y : y ⊑ x};
(4) T −(x) = − inf{T y : y ⊑ x};
(5)
T x ≤ T (x).
3. The projection band of C-compact orthogonally additive
operators
In this section we show that the set of all C-compact regular orthogonally
additive operators from a vector lattice E to a Banach lattice F with order
continuous norm is a band in the vector lattice of all orthogonally additive regular
operators from E to F .
Consider some examples.
Example 3.1. Assume that (A, Ξ, µ) and (B, Σ, ν) are σ-finite measure spaces.
We say that a map K : A × B × R → R is a Carath´eodory function if there hold
the conditions:
(1) K(·, ·, r) is µ × ν-measurable for all r ∈ R;
(2) K(s, t, ·) is continuous on R for µ × ν-almost all (s, t) ∈ A × B.
If additionally there holds K(s, t, ·) = 0 for µ × ν-almost all (s, t) ∈ A × B then
with the function K a nonlinear integral operator: T : Dom(K) → L0(µ)
(T f )(s) := ZB
is associated, where
K(s, t, f (t)) dν(t)
for µ-almost all s ∈ A
Dom(K) := {f ∈ L0(ν) : ZB
K(s, t, f (t)) dν(t) ∈ L0(µ)}.
Assume that E is a Banach function space in L0(ν) and E ⊆ Dom(K). Then
T defines an orthogonally additive integral operator acting from E to L0(µ). The
operator T is called Urysohn (integral) operator and, the function K is called the
kernel of this operator. If L0(µ) is replaced by R we say that T is an Urysohn
integral functional.
1 In the literature these formulas are known as the Riesz-Kantorovich formulas.
4
MARAT PLIEV and MARTIN R. WEBER
Example 3.2. Assume that (Ω, Σ, µ) is a measure space. Then the functional
N : L1(µ) → R defined by
N (f ) = kf kL1(µ) := ZΩ
f (t) dµ,
f ∈ L1(µ)
is positive and orthogonally additive.
Let E be a vector lattice. A linear operator S : E → E is called band preserving
if Sx ∈ {x}⊥⊥ for all x ∈ E. Clearly, every band preserving operators preserves
disjointness.
Example 3.3. Let (Ω, Σ, µ) be a measure space and S : L0(µ) → L0(µ) be a band
preserving linear operator. Consider the new operator T : L0(µ) → L0(µ) defined
by
T f = f S(f ),
f ∈ L0(µ).
Observe that T can be treated as the multiplication operator by the "variable"
function.
It is not hard to verify that T is an orthogonally additive operator.
Indeed, due to Sf ∈ {f }⊥⊥, for every disjoint f1, f2 ∈ L0(µ) we have
T (f1 + f2) = (f1 + f2)S(f1 + f2) =
f1S(f1 + f2) + f2S(f1 + f2) =
f1Sf1 + f2Sf2 = T f1 + T f2.
Definition 3.4. Let (Ω, Σ, µ) be a σ-finite measure space. We say that a function
N : Ω × R → R belongs to the class S (or N is a S-function) if the following
conditions hold:
(1) N(t, 0) = 0 for µ-almost all t ∈ Ω;
(2) N(·, g(·)) is µ-measurable for any g ∈ L0(µ).
If a function N satisfies only the condition (2), then it is called a superpositionally
measurable function or sup-measurable function.
Example 3.5. Let N : Ω × R → R be a S-function. Then with N there is
associated an orthogonally additive operator TN : L0(µ) → L0(µ) defined by
TN (g)(t) = N(t, g(t)) for µ-almost all t ∈ Ω and g ∈ L0(µ).
This class of operators (sometimes called nonlinear superposition operators or
Nemytskii operators) is widely represented in the literature (see e.g. [5]).
Definition 3.6. Let be E be a vector lattice and Y a normed space. An orthog-
onally additive operator T : E → Y is called
ON C-COMPACT OPERATORS
5
• AM-compact provided T maps order bounded subsets of E into relatively
compact subsets in Y ,
• C-compact, if T (Fx) is a relatively compact in Y for any x ∈ E.
For a Banach lattice F by COAr(E, F ) is denoted the space of all C-compact
regular orthogonally additive operators from E to F .
Example 3.7. We note that OAr(R, R) is exactly the set of all real-valued func-
tions such that f (0) = 0. Define an orthogonally additive operator T : R → R
by
1
x2 ,
0,
if x 6= 0
if x = 0.
T x =
Since any element x ∈ R is an atom one has Fx = {0, x} for any x ∈ R. It follows
that T is a C-compact operator. On the other hand T ([0, 1]) is an unbounded set
in R and therefore T is not AM-compact.
Proposition 3.8. Let (B, Σ, ν) be σ-finite measure spaces E be a Banach func-
tion space in L0(ν) and T : E → R be the Urysohn integral functional defined
by
T f = ZB
K(t, f (t)) dν(t),
f ∈ E
with the kernel K. Then T is C-compact.
Proof. Take f ∈ E. We note that Ff coincides with the set {f 1D : D ∈ Σ}. Now
for every D ∈ Σ we write
T f 1D = ZB
K(t, f 1D(t)) dν(t) = ZD
K(t, f (t)) dν(t) ≤
ZB
K(t, f (t)) dν(t) = M.
Hence the set T (Ff ) is order bounded in R and therefore the operator T is
C-compact.
We mention that a C-compact order bounded orthogonally additive operator
T : E → F from a Banach lattices E to a σ-Dedekind complete Banach lattice
F is AM-compact if, in addition, T is uniformly continuous on order bounded
subsets of E [13, Theorem 3.4].
The norm in a normed vector lattice is order continuous if xα ↓ 0 implies
kxαk ↓ 0. We point out that a Banach lattice with order continuous norm is
Dedekind complete (see [4, Theorem 12.9]).
Now we are ready to present the first main result of the article.
6
MARAT PLIEV and MARTIN R. WEBER
Theorem 3.9. Let be E a vector lattice and F a Banach lattice with order contin-
uous norm. Then the set of all C-compact regular orthogonally additive operators
from E to F is a projection band in OAr(E, F ).
In order to prove Theorem 3.9 we need some auxiliary propositions.
Proposition 3.10. [16, Proposition 3.9] Let E be a vector lattice and x, y ∈ E.
Then x ⊑ y if and only if x+ ⊑ y+ and x− ⊑ y−.
Proposition 3.11. Let E be a vector lattice and
yk for some (xi)n
i=1
k=1 ⊂ E. Then there exist a family of pairwise disjoint elements (zik) ⊂
and (yk)m
E, where i ∈ {1, . . . , n} and k ∈ {1, . . . , m} such that
n
m
xi =
Fk=1
Fi=1
(i) xi =
(ii) yk =
zik for any i ∈ {1, . . . , n};
zik for any k ∈ {1, . . . , m};
(iii)
n
Fi=1
zik =
xi =
n
Fi=1
yk.
m
Fk=1
m
n
Fk=1
Fi=1
Fk=1
m
Proof. Using Proposition 3.10, and applying induction arguments we have
n
Gi=1
Gi=1
n
x+
i = (cid:16)
i = (cid:16)
x−
n
Gi=1
Gi=1
n
xi(cid:17)+
xi(cid:17)−
= (cid:16)
= (cid:16)
m
Gk=1
Gk=1
m
yk(cid:17)+
yk(cid:17)−
=
=
m
Gk=1
Gk=1
m
y+
k ;
y−
k .
Now, by the Riesz Decomposition property there exist
z+
ik, z−
ik ∈ E+,
i ∈ {1, . . . , n} and k ∈ {1, . . . , m}
such that
m
m
x+
i =
y+
k =
Gk=1
Gi=1
n
Gk=1
Gi=1
n
z+
ik,
x−
i =
z−
ik, i ∈ {1, . . . , n},
z+
ik,
y−
k =
z−
ik, k ∈ {1, . . . , m}.
Set zik = z+
ik − z−
ik,
i ∈ {1, . . . , n} and k ∈ {1, . . . , m}. Hence
n
m
Gi=1
Gk=1
zik =
n
Gi=1
xi =
m
Gk=1
yk
and the proof is completed.
ON C-COMPACT OPERATORS
7
Lemma 3.12. Let E and F be vector lattices with F Dedekind complete, S, T ∈
OAr(E, F ) and x ∈ E. Then the following equalities hold:
(1)
(2)
(3)
(T ∨ S)x = supn n
Pi=1
(T ∧ S)x = inf n n
Pi=1
T x = supn n
Pi=1
T xi ∨ Sxi : x =
T xi ∧ Sxi : x =
n
n
xi, n ∈ No;
Fi=1
xi, n ∈ No;
Fi=1
T xi : x =
n
Fi=1
xi, n ∈ No.
Proof. Since x ∧ y = −(−x) ∨ (−y) and x = x ∨ (−x) for proving the lemma it
is sufficient to establish only the equation (1). Put
A(x) := n
n
Xi=1
T xi ∨ Sxi : x =
n
Gi=1
xi, n ∈ No.
We show that A(x) is an upward directed set. Indeed, take two disjoint decompo-
sitions x =
n
Fi=1
decomposition x =
xi and x =
n
m
Fi=1
Fk=1
yk of x. By Proposition 3.11 there exists a disjoint
m
Fk=1
zik such that
m
xi =
yk =
Gk=1
Gi=1
n
zik, i ∈ {1, . . . , n},
zik, k ∈ {1, . . . , m}.
We observe that
Sxi ∨ T xi ≤
m
Xk=1
T zik ∨ Szik
for any i ∈ {1, . . . , n}. Indeed
m
m
m
Szik ≤
T zik ∨ Szik,
Sxi = S(cid:16)
T xi = T(cid:16)
zik(cid:17) =
Gk=1
zik(cid:17) =
Gk=1
Xk=1
Xk=1
m
m
Xk=1
Xk=1
m
T zik ≤
T zik ∨ Szik =⇒
Sxi ∨ T xi ≤
m
Xk=1
T zik ∨ Szik.
Similar arguments show that
Syk ∨ T yk ≤
n
Xi=1
T zik ∨ Szik
8
MARAT PLIEV and MARTIN R. WEBER
for any k ∈ {1, . . . , m}. Now we may write
n
Xi=1
T xi ∨ Sxi ≤
n
m
Xi=1
Xk=1
T zik ∨ Szik
and
m
Xk=1
T yk ∨ Syk ≤
n
m
Xi=1
Xk=1
T zik ∨ Szik,
and deduce that A(x) is an upward directed subset of F . By Proposition 2.3
OAr(E, F ) is a Dedekind complete vector lattice and
(T ∨ S)x = sup{T y + Sz : x = y ⊔ z}.
Clearly T x ∨ Sx ≤ (T ∨ S)x for any x ∈ E. Thus for any decomposition x =
n
xi, n ∈ N we have that
Fi=1
n
Xi=1
T xi ∨ Sxi ≤
n
Xi=1
(T ∨ S)xi = (T ∨ S)x
and therefore the set A(x) is upper order bounded. Put fx := sup A(x). Then
fx ≤ (T ∨ S)x. On the other hand for any disjoint decomposition x = y ⊔ z we
have
T y + Sz ≤ supnT y ∨ Sy + T z ∨ Sz : x = y ⊔ zo ≤
n
n
supn
T xi ∨ Sxi : x =
Xi=1
Gi=1
xi, n ∈ No = fx.
Passing to the supremum in the last inequality over all disjoint decompositions
x = y ⊔ z of x, due to formula (1) of Proposition 2.3, we deduce (T ∨ S)x ≤ fx.
It follows (T ∨ S)x = fx which completes the proof.
Below we shall use the following elementary observation. Since
n
Xi=1
n
Sxi ∨ T xi : x =
n
Gi=1
xi, n ∈ No
is an upward downward directed subset of F the set
n
Xi=1
n
Sxi ∧ T xi : x =
n
Gi=1
xi, n ∈ No
is downward directed.
Lemma 3.13. Let be E a vector lattice, F a Banach lattice with order continuous
norm and T ∈ COAr(E, F ). Then FT ⊂ COAr(E, F ).
Proof. We recall that by definition S ∈ FT if S ∧ T − S = 0. We show that
S(Fx) is a relatively compact set2 in F for any S ∈ FT and x ∈ E. Indeed, fix
2 Due to the norm-completeness of F this is equivalent to its totally boundedness.
ON C-COMPACT OPERATORS
9
x ∈ E and ε > 0. Since n n
Pi=1
Sxi ∧ T − Sxi : x =
n
Fi=1
xi, n ∈ No is a downward
n
directed subset of F by Lemma 3.12 and the order continuity of the norm in the
Banach lattice F there exists a disjoint decomposition x =
xi of x such that
Fi=1
n
Xi=1
(cid:13)(cid:13)(cid:13)
Sxi ∧ T − Sxi(cid:13)(cid:13)(cid:13)
<
ε
4
.
Since T (Fxi) is a relatively compact set
finite subset Di ⊂ Fxi such that for any w ∈ Fxi there is u ∈ Di satisfying
for any i ∈ {1, . . . , n} there exists a
kT w − T uk <
ε
2n
.
Moreover, by Proposition 3.11 for any y ∈ Fx there exists a disjoint decomposition
n
y =
yi, where yi ⊑ xi, i.e. yi ∈ Fxi. Hence, for some ui ∈ Di the following
Fi=1
inequality
kT yi − T uik <
ε
2n
holds. We remark that for any positive orthogonally additive operator G : E → F
one has3 Gw ≤ Gxi for all w ∈ Fxi. Now, by taking into account the inequality
x ≤ x + y + x ∧ y, which holds in any vector lattice4 we have
Syi − Sui ≤
Syi − Sui + (T − S)yi − (T − S)ui + Syi − Sui ∧ (T − S)yi − (T − S)ui ≤
T yi − T ui + (Syi + Sui) ∧ ((T − S)yi + (T − S)ui) ≤
T yi − T ui + (Syi + Sui) ∧ (T − Syi + T − Sui) ≤
T yi − T ui + (Sxi + Sxi) ∧ (T − Sxi + T − Sxi) =
T yi − T ui + 2(cid:16)Sxi ∧ T − Sxi(cid:17).
Then we may write
n
Xi=1
Syi − Sui ≤
n
Xi=1
T yi − T ui + 2
n
Xi=1 (cid:16)Sxi ∧ T − Sxi(cid:17).
3 This immediately follows from y ∧ x − y = 0, the positivity of G and G(x) = G(y) +
G(x − y) ≥ G(y).
4 Indeed, since (see [26], Theorems 5.1 and 5.5) x + y ≥ x − y ≥ (x − y)+ we have
x − x ∧ y = x + (−x) ∨ (−y) = 0 ∨ (x − y) = (x − y)+ ≤ x + y.
10
MARAT PLIEV and MARTIN R. WEBER
n
Put D := {u =
ui : ui ∈ Di}. Clearly D is a finite subset of Fx. Thus for
any y ∈ Fx and u ∈ D we have
Fi=1
n
n
n
kSy − Suk = (cid:13)(cid:13)(cid:13)(cid:12)(cid:12)S(cid:16)
Syi − Sui(cid:13)(cid:13)(cid:13)
Xi=1
≤ (cid:13)(cid:13)(cid:13)
yi(cid:17) − S(cid:16)
Gi=1
T yi − T ui(cid:13)(cid:13)(cid:13)
n
Xi=1
= (cid:13)(cid:13)(cid:13)(cid:12)(cid:12)
Syi − Sui(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)
ui(cid:17)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)
Gi=1
+ 2(cid:13)(cid:13)(cid:13)
Xi=1 (cid:16)Sxi ∧ T − Sxi(cid:17)(cid:13)(cid:13)(cid:13)
n
≤
≤
n
(cid:13)(cid:13)(cid:13)
Xi=1
Xi=1
n
kT yi − T uik +
≤ ε.
ε
2
Hence {Su : u ∈ D} is a finite ε-net for S(Fx), and therefore the proof is com-
pleted.
Let V be a vector lattice and v ∈ V+. The order ideal generated by v is denoted
by Iv, i.e. Iv = {x ∈ V : ∃α > 0 such that x ≤ αv}. A v-step function is
any vector s ∈ V for which there exist pairwise disjoint fragments v1, . . . , vn of
vn
i and real numbers λ1, . . . , λn satisfying s =
λivi. The next
v with v = Fi=1
n
Pi=1
proposition is known as the Freudenthal Spectral Theorem ([4], Theorem 2.8).
Proposition 3.14. Let V be a vector lattice with the principal projection property
and let v ∈ V+. Then for every u ∈ Iv there exists a sequence (sn)n∈N of v-step
functions satisfying 0 ≤ u − sn ≤ 1
n v for each n and sn ↑ u. Moreover, 0 ≤ sn if
0 ≤ u.
Now we ready to prove the first main result.
Proof of Theorem 3.9. We prove some properties of COAr(E, F ):
a) Clearly, COAr(E, F ) is a vector subspace of OAr(E, F ).
b) We show that COAr(E, F ) is even a vector sublattice of OAr(E, F ). Take
S, T ∈ COAr(E, F ). Then T − S ∈ COAr(E, F ). By Lemma 3.13 FT ⊂
COAr(E, F ) and therefore T + ∈ COAr(E, F ). Therefore due to the equalities
S + (T − S)+ = S + (T − S) ∨ 0 = T ∨ S, S ∧ T = −(−S) ∨ (−T )
which are valid in OAr(E, F ) we obtain that COAr(E, F ) is a sublattice of
OAr(E, F ).
c) Now we show that T ∈ COAr(E, F ) if 0 ≤ Tλ ↑ T in OAr(E, F ) and any
Tλ ∈ COAr(E, F ). Indeed, take x ∈ E and ε > 0. Since the Banach lattice F
is order continuous it follows from Tλx ↑ T x that kT x − Tλ0xk < ε
4 for some λ0.
ON C-COMPACT OPERATORS
11
Then actually kT y − Tλ0yk < ε
some z ∈ E. Then
4 for any y ∈ Fx. Indeed, consider x = y ⊔ z for
0 ≤ T x − Tλ0x = T (y ⊔ z) − Tλ0(y ⊔ z) = T y − Tλ0y + T z − Tλ0 z ≥ T y − Tλ0y
implies kT y − Tλ0yk ≤ kT x − Tλ0xk. Since Tλ0 ∈ COAr(E, F ) there exists a
finite subset D of Fx with the property that for any y ∈ Fx there exists u ∈ D
satisfying
So we obtain
kTλ0u − Tλ0yk <
ε
2
.
kT u − T yk ≤ kT u − Tλ0u + Tλ0u − T y + Tλ0y − Tλ0yk ≤
kT u − Tλ0 uk + kT y − Tλ0 yk + kTλ0u − Tλ0yk < ε,
what establishes the relative compactness of T (FT ) in F .
d) Finally we prove that COAr(E, F ) is an order ideal in OAr(E, F ). Let
0 ≤ R ≤ T , where R ∈ OAr(E, F ) and T ∈ COAr(E, F ). Then R ∈ IT
and by Proposition 3.14 there5 exists a sequence (Sn)n∈N in OAr(E, F ) of T -
step-functions with 0 ≤ Sn ↑ R. Taking into account that6 Sn ∈ COAr(E, F ),
n ∈ N and what has been established in c) we deduce that R ∈ COAr(E, F ). So,
COAr(E, F ) is a band in OAr(E, F ).
e) Due to the Dedekind completeness of OAr(E, F ) it is a projection band.
4. C-compact and narrow orthogonally additive operators
In this section we consider a new class of vector lattices, where the condition of
Dedekind completeness is replaced by a much weaker property. For laterally-to-
norm continuous, C-compact orthogonally additive operators from a C-complete
vector lattice E to a Banach space X we show their narrowness.
Definition 4.1. A vector lattice E is said to be C-complete, if for each x ∈ E+
any subset D ⊂ Fx has a supremum and an infimum.
Clearly, every Dedekind complete vector lattice E is C-complete. The reverse
statement, in general, is not true.
5 This proposition can be applied since by Proposition 2.3 the vector lattice OAr(E, F ) is
Dedekind complete. Any T -step-function S ∈ OAr(E, F ) has the form S =
m
λiTi, where Ti
m
Pi=1
are disjoint fragments of T such that T =
Ti.
Fi=1
6 This follows from the fact that together with T each fragment Ti of T belongs to
COAr(E, F ).
12
MARAT PLIEV and MARTIN R. WEBER
Proposition 4.2. The vector lattice E = C[0, 1] of all continuous functions on
the interval [0, 1] is C-complete.
∞
Si=1
Proof. Fix f ∈ E+. Since the set Sf := {t ∈ [0, 1] : f (t) > 0} is open, by [10,
Chap.2, Theorem 5] there is the decomposition Sf =
(ai, bi) where (ai, bi) ∩
(aj, bj) = ∅, i 6= j. Take g ∈ Ff . We claim that for every i ∈ N either g(t) = 0 or
g(t) = f (t) for any t ∈ (ai, bi). Indeed, denote by fi and gi the restrictions of f
and g on the closed interval [ai, bi], respectively. It is clear that gi ∈ Ffi. Assume
that there exists a nonzero fragment vi ∈ Ffi such that vi ⊥ gi and fi = gi + vi.
Put Gi := {t ∈ (ai, bi) : g(t) > 0} and Vi = {t ∈ (ai, bi) : v(t) > 0}. Since fi
is strictly positive on (ai, bi) we deduce that (ai, bi) = Gi ⊔ Vi and therefore Gi
and Vi are open-closed subsets of (ai, bi). But the interval [ai, bi] is a connected
set and we come to the contradiction. Thus fi has no fragment 0 < vi < fi and
therefore, either fi(t) = gi(t) or gi(t) = 0 for any t ∈ (ai, bi). With each g ∈ Ff
there is associated the sequence (gi)i∈N, where
1,
0,
gi =
if g(t) = f (t)
if g(t) = 0
t ∈ (ai, bi).
Clearly, in this way a one-to-one correspondence between Ff and the set of all
0 - 1 sequences is established. Let D be any fixed subset of Ff . Put
D+ := {i ∈ N : ∃ g ∈ D such that gi = 1};
D− := {i ∈ N : such that gi = 1 for ∀ g ∈ D}.
Consider the pair of sequences (ui)i∈N and (vi)i∈N, where
1,
0,
1,
0,
ui =
vi =
if i ∈ D+
otherwise,
if i ∈ D−
otherwise.
With (ui)i∈N and (vi)i∈N there is associated the pair {u, v} of fragments of f .
Clearly u = sup D and v = inf D.
Since the vector lattice C[0, 1] is Archimedean but not Dedekind complete the
previous proposition shows that the set of all C-complete vector lattices is a
new subclass of vector lattices which strictly contains the class of all Dedekind
complete vector lattices. We note that in general a C-complete vector lattice is
not Archimedian.
ON C-COMPACT OPERATORS
13
Example 4.3. Let E = R2 equipped with the lexicographic order. That is, we
consider E as a vector lattice with the following order7 (x1, x2) ≥ (y1, y2), when-
ever either x1 > y1 or else x1 = y1 and x2 ≥ y2. The vector lattice E is not
Archimedian. On the other hand it is not hard to verify that the Boolean algebra
of all fragments Fx of an arbitrary element x = (x1, x2) ∈ E+ contains only two
elements:
Hence E is C-complete.
Fx = {(x1, x2), (0, 0)}.
We say that a set D ⊂ E is laterally bounded, if there exits x ∈ E such that
D ⊂ Fx. We say that a laterally bounded set D has a lateral supremum (infimum)
if there exists u ∈ E (v ∈ E) such that u = sup D (v = inf D) with respect to
the partial order ⊑ in Fx. Taking into account Proposition 3.10 we deduce that
a vector lattice E is C-complete if and only if every laterally bounded subset of
E has the lateral supremum and infimum.
Definition 4.4. Let E be a vector lattice and X be a normed space. An or-
thogonally additive operator T : E → X is called narrow, if for any x ∈ E and
ε > 0 there exists a pair x1, x2 of mutually complemented fragments of v, such
that kT x1 − T x2k < ε. In particular, if X = R, we call T a narrow functional.
Observe that the image of an atom under a narrow operator T is zero. Indeed.
The only disjoint fragments of an atom a are 0 and a. So, due to the narrowness
of T , for any ε > 0 one has kT uk < ε, what means T u = 0.
This is the reason for supposing the vector lattice E to be atomless in the Theo-
rems 4.7 and in most of the propositions of the current section.
Example 4.5. Let (Ω, Σ, µ) be a σ-finite measure space. Consider a map N :
L1(µ) → R defined by
N (f ) = kf kL1(µ),
f ∈ L1(µ).
In [23] (Proposition 2.5) it was shown that N is a narrow orthogonally additive
functional on L1(µ).
A net (xα)α∈Λ in a vector lattice E laterally converges to x ∈ E if xα ⊑ xβ ⊑ x
for all α ≤ β and xα order converges to x. This is written as xα
lat−→ x.
7 The cone E+ in this vector lattice consists of the open right half-space {(x1, x2) : x1 > 0}
joint with the half-ray {(x1, x2) : x1 = 0, x2 ≥ 0}.
14
MARAT PLIEV and MARTIN R. WEBER
Definition 4.6. An orthogonally additive operator T from a vector lattice E
to a normed space X is called laterally-to-norm continuous whenever for each
lat−→ x the net (T xα)α∈Λ converges with
laterally convergent net (xα)α∈Λ with xα
respect to the norm to in X to T x.
The following theorem is the second main result of the article.
Theorem 4.7. Let E be an atomless C-complete vector lattice and X be a Banach
space. Then every orthogonally additive laterally-to-norm continuous C-compact
operator T : E → X is narrow.
The next auxiliary proposition is well known (see e.g. [21, Lemma 10.20]).
Proposition 4.8. Let (vi)n
normed space V and (λi)n
i. Then there exists a set (θi)n
and
i=1 be a finite subset of elements in a finite dimensional
i=1 be non-negative numbers such that 0 ≤ λi ≤ 1 for each
i=1 of numbers such that θi ∈ {0, 1}, i ∈ {1, . . . , n}
n
Xi=1
(cid:13)(cid:13)(cid:13)
(λi − θi) vi(cid:13)(cid:13)(cid:13)
≤
dim V
2
max
i∈{1,...,n}
kvik.
Proposition 4.9. Let be E an atomless vector lattice, x ∈ E, X a Banach space
and T : E → X an orthogonally additive laterally-to-norm continuous operator.
Then for any ε > 0 there exists a decomposition x = y ⊔ z, where y, z are nonzero
fragments of x such that kT zk < ε.
Proof. Since E is an atomless vector lattice the Boolean algebra A := Fx has
infinite cardinality. We note that the set of all fragments of x is the net (xα)α∈A
laterally converges to x. The laterally-to-norm continuity of T implies the exis-
tence of α0 ∈ A such that kT x − T xαk < ε for all α ≥ α0. Consider for α0 the
disjoint decomposition x = (x − xα0) ⊔ xα0. Then
T x = T(cid:0)(x − xα0) ⊔ xα0(cid:1) = T (x − xα0) + T xα
implies kT (x − xα0)k < ε.
Assign y := xα0 and z = x − xα0. Then kT zk < ε and x = y ⊔ z is the desirable
disjoint decomposition.
Proposition 4.10. Let be E an atomless vector lattice, x ∈ E, X a Banach space
and T : E → X an orthogonally additive laterally-to-norm continuous operator.
Assume that (yn)n∈N is a sequence of fragments of x such that y1 = x, yn ⊑ ym
for n ≥ m; m, n ∈ N and Tn∈N
Fyn = {0}. Then lim
n→∞
kT ynk = 0.
ON C-COMPACT OPERATORS
15
Proof. By definition ym = yn ⊔ (ym − yn) for m, n ∈ N with n ≥ m. For xn :=
x − yn, n ∈ N with yn = ym − (ym − yn) we can write
xn = x − ym + (ym − yn) = x − ym + ((x − yn) − (x − ym)) = xm + (xn − xm).
Clearly xm ⊥ (xn − xm) for n, m ∈ N and n ≥ m and therefore xn ⊥ (x − xn) for
Fyn = {0}. Hence
all n ∈ N. Thus x = xn ⊔ (x − xn), n ∈ N and Tn∈N
we deduce that the sequence (xn)n∈N laterally converges8 to x. Now the relation
xn ⊥ (x − xn) for all n ∈ N implies
Fx−xn = Tn∈N
T x = T(cid:0)xn ⊔ (x − xn)(cid:1) = T xn + T (x − xn) and T (x − xn) = T x − T xn.
Thus by laterally-to-norm continuity of T we have that T xn is norm convergent
to T x in X and
lim
n→∞
kT x − T xnk = lim
n→∞
kT ynk = 0.
Proposition 4.11. Let be E an atomless C-complete vector lattice, X a Banach
space, T : E → X an orthogonally additive laterally-to-norm continuous operator,
x ∈ E and ε > 0. Then for some n ∈ N there exists a decomposition x =
n
where xi are nonzero fragments of x such that kT xik ≤ ε for any i ∈ {1, . . . , n}.
xi,
Fi=1
Proof. By Proposition 4.9 the set
Dx,T,ε := {z ∈ Fx : z 6= 0, kT zk ≤ ε}
is not empty. We note that Dx,T,ε is a partially ordered set with respect to the
relation ⊑. Let (uλ)λ∈Λ ⊆ Dx,T,ε be a chain, where Λ is a some linearly ordered
index set. Clearly uµ ⊑ uλ for all µ, λ ∈ Λ, µ ⊑ λ. By the C-completeness of
(uλ), where u ∈ Fx. Due to kT uλk ≤ ε
vector lattice E there exists u = (o)-lim
λ∈Λ
and the laterally-to-norm continuity of T we have that then T u = lim
(T uλ)
λ∈Λ
and therefore u ∈ Dx,T,ε. By Zorn's Lemma, there is a maximal element9 z ∈
Dx,T,ε with kT zk ≤ ε. Put y = x − z. If kT yk ≤ ε then we got the required
decomposition of x. Otherwise we apply Proposition 4.9 to y and get y = y1 ⊔ y2,
where y1 is a maximal element in Dy,T,ε with kT y1k ≤ ε. In case of necessity,
i.e. if kT y2k > ε, by further continuing in the same way with the corresponding
fragments y2, y4, . . . , y2k, . . . of y we construct a sequence of decompositions y2k =
y2k+1 ⊔ y2k+2, where y2k+1 is a maximal element in Dy2k,T,ε and satisfying the
8 Indeed, if some vector 0 6= u is a fragment of each vector x − xn then u ⊑ (x − xn) and
x − xn ≤ un for some sequence un ↓ 0 is impossible.
9 The element z is maximal in Dx,T,ε, if ∄ u ∈ Dx,T,ε such that z ⊑ u and z 6= u.
16
MARAT PLIEV and MARTIN R. WEBER
conditions kT y2k+1k ≤ ε and kT y2k+2k > ε, k ∈ N. We claim that there exists
l ∈ N such that y2l = y2l+1 ⊔ y2l+2 and both kT y2l+1k, kT y2l+2k ≤ ε. Assume the
contrary. Then the sequence (y2k)k∈N of fragments of y is such that kT y2kk > ε
Fy2k = {0}. Indeed, assume that there
for all k ∈ N. Nevertheless we show Tk∈N
exists a nonzero element v ∈ Tk∈N
2 = y2 − v, y′
y′
2m for m, n ∈ N with n ≥ m and Tn∈N
4 = y4 − v, . . . , y′
we have y′
2n ⊑ y′
Fy2k. Then for the sequence
2k = y2k − v, . . .
Fy′
2n = {0}. According to
Proposition 4.10 there exists n0 ∈ N such that kT y′
y′
2n0 ∈ Dy2n0 ,T,ε and y2n0+1 is a maximal element of Dy2n0 ,T,ε. We have
2n0k < ε. Thus y2n0 = y′
2n0 ⊔ v,
y2n0 = y2n0+1 ⊔ y2n0+2 = y′
2n0 ⊔ v.
Consider now the two cases:
Fy2n0+1 ∩ Fv = {0} and Fy2n0+1 ∩ Fv 6= {0}.
2n0, y2n0+1 6= y′
In the first case, since no non-zero fragment of v is a fragment of y2n0+1, we get
y2n0+1 ⊑ y′
2n0. Observe that - due to the maximality of y2n0+1 in Dy2n0 ,T,ε - the
relation y2n0+1 ⊑ y′
2n0 = y2n0+1, i.e.
y2n0+2 = v and, therefore v is not a fragment of y2n0+4. In the second one there
is a nonzero fragment w ∈ Fy2n0+1
∩ Fv. Thus v = (v − w) ⊔ w and w ⊑ y2n0+1,
i.e. the non-zero fragment w of v belongs to Fy2n0+1. Therefore the element v can
not be a fragment of y2n0+2.
2n0 is impossible. Hence y′
Fy2k = {0} and so, again by applying Proposition 4.10, we get
Hence Tk∈N
kT y2kk = 0. This is a contradiction and therefore the desirable l ∈ N exists.
lim
k→∞
Consider the following elements
x1 = y1, x2 = y3, . . . , xl = y2l−1, xl+1 = y2l+1, xl+2 = y2l+2, xl+3 = z.
Then x =
n
Fi=1
xi with n = l + 3 is the desirable decomposition of x.
Proposition 4.12. Let E be an atomless C-complete vector lattice and V a finite
dimensional Banach space. Then every orthogonally additive laterally-to-norm
continuous operator G : E → V is narrow.
Proof. Fix any x ∈ E and ε > 0. According to Proposition 4.11 there is a disjoint
dim V for any i ∈ {1, . . . , n}. Then by
xi such that kGxik < ε
decomposition x =
n
Fi=1
ON C-COMPACT OPERATORS
17
using Proposition 4.8 for λi = 1
such that
2 there exist numbers θi ∈ {0, 1} for i ∈ {1, . . . , n}
n
2
Xi=1(cid:16)1
2(cid:13)(cid:13)(cid:13)
Observe that for I0 = (cid:8)i ∈ {1, . . . , n(cid:9) : θi = 0(cid:9) and I1 = (cid:8)i ∈ {1, . . . , n} : θi = 1(cid:9)
the vectors yk = Fi∈Ik
xi for k ∈ {0, 1} are mutually complemented fragments of x
− θi(cid:17) Gxi(cid:13)(cid:13)(cid:13)
(4.1)
≤ dim V max
kGxik < ε.
i∈{1,...,n}
and by (4.1),
kGy1 − Gy2k = (cid:13)(cid:13)(cid:13) Xi∈I0⊔I1
hence the operator G is narrow.
(1 − 2θi) Gxi(cid:13)(cid:13)(cid:13)
< ε,
Definition 4.13. Let E be a vector lattice and F a vector space. An orthogonally
additive operator T : E → F is called of finite rank if the set T (E) generates a
finite-dimensional subspace in F .
Now we are in the position to prove the second main result (Theorem 4.7).
Before, however, we notice that a Banach space X can be considered as a closed
subspace of the Banach space
W := l∞(BX ∗) = {q : BX ∗ → R, sup q(f ) < ∞, f ∈ BX ∗}
(equipped with the supremum-norm) of all real-valued bounded functions on the
closed unit ball BX ∗ of the dual space X ∗ according to
X ֒→ X ∗∗ ֒→ W,
where the notation ֒→ means the isometric embedding
X ∋ x 7→ Fx ∈ X ∗∗
given by Fx(f ) := f (x), f ∈ BX ∗ .
Observe that, if H is a relatively compact subset of the Banach space W and
ε > 0, then there exists a linear finite rank operator R ∈ L(W ) such10 that
kw − Rwk ≤ ε for every w ∈ H [21, Lemma 10.25].
Proof of Theorem 4.7. Fix an arbitrary x ∈ E and ε > 0. Due to the C-
compactness of T the set K = T (Fx) is relatively compact in X and therefore
in W . It follows that there exists a finite rank operator S ∈ L(W ) such that
ky − Syk ≤ ε
4 for every y ∈ K. Then G = S ◦ T is an orthogonally additive
laterally-to-norm continuous finite rank operator. By Lemma 4.12 G is narrow
10 By L(W ) there is denoted the space of all linear bounded operators on W .
18
MARAT PLIEV and MARTIN R. WEBER
and consequently, there exist mutually complemented fragments x1, x2 of x such
that kGx1 − Gx2k < ε
2. Therefore,
kT x1 − T x2k =
kT x1 − T x2 + S(T x1) − S(T x2) − S(T x1) + S(T x2)k =
kT x1 − T x2 + Gx1 − Gx2 − Gx1 + Gx2k ≤
kGx1 − Gx2k + kT x1 − Gx1k + kT x2 − Gx2k <
ε
2
+
ε
2
= ε
since kT xi − Gxik < ε
4 for i ∈ {1, 2}. This completes the proof.
We remark that Theorem 4.7 generalizes the result of the article [18, Theo-
rem 3.2].
Acknowledgments. Marat Pliev was supported by the Russian Foundation
for Basic Research (grant number 17-51-12064). Martin Weber was supported by
the Deutsche Forschungsgemeinschaft (grant number CH 1285/5-1, Order pre-
serving operators in problems of optimal control and in the theory of partial
differential equations).
References
[1] N. Abasov, M. Pliev, On extensions of some nonlinear maps in vector lattices, J. Math.
Anal. Appl. 455 (2017), 516 -- 527.
[2] N. Abasov, M. Pliev, Disjointness preserving orthogonally additive operators in vector
lattices, Banach of Math. Anal. 12 (2018), 3, 730 -- 750.
[3] N. Abasov, M. Pliev, Dominated orthogonally additive operators in lattice-normed spaces,
Advances in Operator Theory, 4 (2019), 1, 251 -- 264.
[4] C. D. Aliprantis, O. Burkinshaw, Positive Operators, Springer, Dordrecht. (2006).
[5] J. A. Appell, P. P. Zabrejko, Nonlinear superposition operators, Cambridge University Press,
2008.
[6] W. A. Feldman, Lattice preserving maps on lattices of continuous functions, J. Math. Anal.
Appl., 404 (2013), 310 -- 316.
[7] W. A. Feldman, A characterization of non-linear maps satisfying orthogonality properties,
Positivity 21 (2017), 1, 85 -- 97.
[8] W. A. Feldman, A factorization for orthogonally additive operators on Banach lattices, J.
Math. Anal. and Appl., 472, (2019), 1, 238 -- 245.
[9] O. Fotiy, A. Gumenchuk, I. Krasikova, M. Popov, On sums of narrow and compact operators,
Positivity, doi.org/10.1007/s11117-019-00666-4.
[10] A. N. Kolmogorov, S. V. Fomin, Introductory Real Analysis, Dover Publication. (1975).
[11] O. Maslyuchenko, V. Mykhaylyuk, M. Popov, A lattice approach to narrow operators,
Positivity, 13 (2009), 459 -- 495.
[12] J. M. Maz´on, S. Segura de Le´on, Order bounded ortogonally additive operators, Rev.
Roumane Math. Pures Appl., 35 (1990), 4, 329 -- 353.
ON C-COMPACT OPERATORS
19
[13] J. M. Maz´on, S. Segura de Le´on, Uryson operators, Rev. Roumane Math. Pures Appl. 35
(1990), 5, 431 -- 449.
[14] V. Mykhaylyuk, M. Popov, On sums of narrow operators on Kothe function space, J. Math.
Anal. Appl. 404 (2013), 554 -- 561.
[15] V. Mykhaylyuk, On the sum of narrow and a compact operators, J. Funct. Anal. 266
(2014), 5912 -- 5920.
[16] V. Orlov, M. Pliev, D. Rode, Domination problem for AM -compact abstract Uryson oper-
ators, Archiv der Mathematik, 107, 5, (2016), 543 -- 552.
[17] Plichko A., M.Popov M. Symmetric function spaces on atomless probability spaces, Disser-
tationes Math. (Rozprawy Mat.), 306 (1990), 1 -- 85.
[18] M. Pliev, M. Popov, Narrow orthogonally additive operators, Positivity, 18 (2014), 4, 641 --
667.
[19] M. Pliev, K. Ramdane Order unbounded orthogonally additive operators in vector lattices,
Mediterranean J. Math. 15 (2018), 2, 20 p.
[20] M. Pliev, X. Fang, Narrow orthogonally additive operators in lattice-normed spaces,
Siberian Math. J., 58 (2017), 1, 134 -- 141.
[21] M. Popov, B. Randrianantoanina, Narrow Operators on Function Spaces and Vector Lat-
tices, De Gruyter Studies in Mathematics 45, De Gruyter (2013).
[22] M. A. Pliev, M. R. Weber, Disjointness and order projections in the vector lattices of ab-
stract Uryson operators, Positivity, 20 (2016), 3, 695 -- 707.
[23] M. A. Pliev, F. Polat, M. R. Weber, Narrow and C-compact orthogonally additive operators
in lattice-normed spaces, Results in Math., 74 (2019), 4, 19 pp.
[24] Tradacete, P., and Villanueva, I. Continuity and representation of valuations on star bodies,
Adv. Math. 329 (2018), 361 -- 391.
[25] Tradacete, P., and Villanueva, I. Valuations on Banach lattices, Int. Math. Res. Not.,
(2019), to appear.
[26] A. C. Zaanen, Introduction to Operator Theory in Riesz Spaces, Springer, Berlin Heidelberg
New York. (1997).
Southern Mathematical Institute of the Russian Academy of Sciences, Vladikavkaz,
362027 Russia
E-mail address: [email protected]
Technical University Dresden, Department of Mathematics, Institute of Anal-
ysis, 01062 Dresden, Germany
E-mail address: [email protected]
|
1112.5262 | 2 | 1112 | 2012-07-18T12:27:28 | Nonstationary Gabor Frames - Existence and Construction | [
"math.FA"
] | Nonstationary Gabor frames were recently introduced in adaptive signal analysis. They represent a natural generalization of classical Gabor frames by allowing for adaptivity of windows and lattice in either time or frequency. In this paper we show a general existence result for this family of frames. We then give a perturbation result for nonstationary Gabor frames and construct nonstationary Gabor frames with non-compactly supported windows from a related painless nonorthogonal expansion. Finally, the theoretical results are illustrated by two examples of practical relevance. | math.FA | math |
Nonstationary Gabor Frames - Existence and Construction
Monika Dorfler, Ewa Matusiak∗
Department of Mathematics, NuHAG, University of Vienna, Austria
Abstract
Nonstationary Gabor frames were recently introduced in adaptive signal analysis. They
represent a natural generalization of classical Gabor frames by allowing for adaptivity of
windows and lattice in either time or frequency. In this paper we show a general existence
result for this family of frames. We then give a perturbation result for nonstationary Gabor
frames and construct nonstationary Gabor frames with non-compactly supported windows
from a related painless nonorthogonal expansion. Finally, the theoretical results are illus-
trated by two examples of practical relevance.
Keywords: adaptive representations, nonorthogonal expansions, irregular Gabor frames,
existence
1. Introduction
The principal idea of Gabor frames was introduced in [14] with the aim to represent signals
in a time-frequency localized manner. Since the work of Gabor himself, a lot of research has
been done on the topic of atomic time-frequency representation. While it turned out that
the original model proposed by Gabor does not yield stable representations in the sense of
frames [4, 7, 10], the existence of Gabor frames was first established in the so called painless
case, [7], which requires the use of compactly supported analysis windows. The existence of
Gabor frames in more general situations was proved later [19, 23] and the proof often uses
an argument invoking the invertibility of diagonally dominant matrices.
Various irregular and adaptive versions of Gabor frames have been introduced over the
years, cf. [1, 5, 12, 21].
In these approaches, the irregularity usually concerns either the
sampling set, which is allowed to deviate from a lattice, or the window, which is allowed
to be modified. In [1], varying windows as well as irregular sampling points are allowed,
however, existence of a local frame is assumed, from which a global frame is constructed.
Nonstationary Gabor frames give up the strict regularity of the classical Gabor setting, but,
as opposed to irregular frames, maintain enough structure to guarantee efficient implementa-
tion and, possibly approximate, efficient reconstruction. In analogy to the classical, regular
case [7], painless nonstationary Gabor frames were introduced in [2], where the principal
idea is described and illustrated in detail. The construction of painless nonstationary Gabor
∗Corresponding author
Email addresses: [email protected] (Monika Dorfler), [email protected]
(Ewa Matusiak)
Preprint submitted to Elsevier
April 30, 2014
frames is similar to, but more flexible than the construction of windowed modified cosine
transforms and other lapped transforms [17, 24] that allow for adaptivity of the window
length. An efficient and perfectly invertible constant-Q transform was recently introduced
using nonstationary Gabor transforms [22]. In this and similar situations, redundancy of
the transform is crucial, since non-redundant versions of the constant-Q transform lead to
dyadic wavelet transforms, which are often inappropriate for audio signal processing.
Redundancy allows for good localization of both the analysis and synthesis windows, and
their respective Fourier transforms and often promote sparse representations in adaptive
processing.
Painless non-orthogonal expansions can only be devised if the involved analysis windows
are either compactly supported or band limited. This requirement may sometimes be too re-
strictive. For instance, one may be interested in designing frequency-adaptive nonstationary
Gabor frames with windows that are compactly supported in time, i.e. can be implemented
as FIR filters, cf. [11] and lend themselves to real-time implementation, cp. [9].
The present contribution addresses the case of nonstationary Gabor frames with more
general windows than used in the painless case. In Theorem 3.4, we derive the existence
of nonstationary Gabor frames directly from a generalized Walnut representation: under
mild uniform decay conditions on all windows involved, we show an existence result of
nonstationary Gabor frames in parallel to the result given in [23] for regular Gabor frames,
also compare [15, Theorem 6.5.1]. Note that the existence of a different class of nonstationary
Gabor frames, the quilted Gabor frames, [8], was recently proved in the general context of
spline type spaces in the remarkable paper [18].
This paper is organized as follows. In the next section, we introduce notation and state
some auxiliary results. In Section 3, we first define nonstationary Gabor frames and recall
known results for the painless case. In Section 3.2 a Walnut representation and a correspond-
ing bound of the frame operator in the general setting is derived and Section 3.3 provides the
existence of nonstationary Gabor frames. In Section 3.4 we pursue two basic approaches for
the construction of nonstationary Gabor frames . Using tools from the theory of perturbation
of frames, we construct nonstationary frames from an existing frame in Proposition 3.7. In
Corollary 3.8 we design nonstationary Gabor frames by exploiting knowledge about a related
painless frame, to obtain ”almost painless nonstationary Gabor frames”. In Section 4 we
provide examples based on the two introduced construction principles.
2. Notation and Preliminaries
Given a non-zero function g ∈ L2(R), let gk,l(t) = MblTakg(t) := e2πibltg(t − ak). Mbl is a
modulation operator, or frequency shift, and Tak is a time shift.
The set G(g, a, b) = {gk,l : k, l ∈ Z} is called a Gabor system for any real, positive a, b.
G(g, a, b) is a Gabor frame for L2(R), if there exist frame bounds 0 < A ≤ B < ∞ such that
for every f ∈ L2(R) we have
To every Gabor system, we associate the analysis operator Cg given by (Cgf )k,l = (cid:104)f, gk,l(cid:105),
k,l∈Z ck,lγk,l for c ∈ (cid:96)2. The operator
and the synthesis operator Uγ = C∗
A(cid:107)f(cid:107)2
(cid:104)f, gk,l(cid:105)2 ≤ B(cid:107)f(cid:107)2
2 .
(1)
2 ≤ (cid:88)
γ , given by Uγc =(cid:80)
k,l∈Z
2
Sg,γ associated to G(g, a, b) and G(γ, a(cid:48), b(cid:48)), where Sg,γ = UγCg reads
(cid:88)
k,l∈Z
Sg,γf =
(cid:104)f, gk,l(cid:105)γk,l .
The inequality (1) is equivalent to the invertibility and boundedness of the frame operator
Sg,g of G(g, a, b).
window g ∈ L2(R), the STFT of f ∈ L2(R) is
The analysis operator is the sampled short-time Fourier transform (STFT). For a fixed
Vgf (x, ω) =
f (t)e−2πiωtg(t − x) dt = (cid:104)f, MωTxg(cid:105) .
Setting (x, ω) = (ak, bl), leads to Vgf (ak, bl) = (Cgf )k,l.
When working with irregular grids, we assume that the sampling points form a separated
set: a set of sampling points {ak : k ∈ Z} is called δ-separated, if ak − am > δ for ak, am,
whenever k (cid:54)= m. χI will denote the characteristic function of the interval I.
A convenient class of window functions for time-frequency analysis on L2(R) is the Wiener
space.
Definition 2.1. A function g ∈ L∞(R) belongs to the Wiener space W (L∞, (cid:96)1) if
(cid:107)g(cid:107)W (L∞,(cid:96)1) :=
ess supt∈Qg(t + k) < ∞ , Q = [0, 1] .
(cid:90)
R
(cid:88)
(cid:88)
k∈Z
k∈Z
For g ∈ W (L∞, (cid:96)1) and δ > 0 we have [15]
ess supt∈R
g(t − δk) ≤ (1 + δ−1)(cid:107)g(cid:107)W (L∞,(cid:96)1) .
(2)
In dealing with polynomially decaying windows, we will repeatedly use the following
lemma.
Lemma 2.2. For p > 1 the following estimates hold:
(a) Let δ > 0, then
(b) Let {ak : k ∈ Z} ⊂ R be a δ−separated set. Then
(1 + δk)−p ≤ (1 + δ)−p(δ−1 + p)(p − 1)−1 .
∞(cid:88)
(cid:88)
(1 + t − ak)−p ≤ 2(cid:0)1 + (1 + δ)−p(δ−1 + p)(p − 1)−1(cid:1) .
k=1
ess supt∈R
k∈Z
Proof. To show (a) we write
(1 + δk)−p = (1 + δ)−p +
∞(cid:88)
k=1
∞(cid:88)
k=2
(1 + δk)−p = (1 + δ)−p +
3
(cid:90)
∞(cid:88)
k=2
[0,1]+k
(1 + δk)−p dt .
for t ∈ [k, k + 1], we have δt ≤ δ(k + 1) which implies that 1 + δ(t − 1) ≤ 1 + δk. Therefore,
(cid:90)
∞(cid:88)
(cid:90) ∞
(1 + δk)−p dt ≤
(1 + δ(t − 1))−p dt =
(1 + δ(t − 1))−p dt
k=2
[0,1]+k
k=2
[0,1]+k
= (1 + δ)−p+1δ−1(p − 1)−1 ,
2
(cid:90)
∞(cid:88)
(cid:88)
k∈Z
and the estimate follows.
To prove (b), fix t ∈ R. Since ak − al > δ for k (cid:54)= l, each interval of length δ contains at
most one point t − ak, k ∈ Z. Therefore we may write t − ak = δnk + xk for unique nk ∈ Z
and xk ∈ [0, δ), and by the choice of δ, we have nk (cid:54)= nl for k (cid:54)= l. We assume, without loss
of generality, that nk = 0 for k = 0, and we find that
(1 + t − ak)−p =
(1 + δnk + xk)−p
(cid:88)
k∈Z
≤ 1 +
≤ 1 +
≤ 1 +
(cid:32)
k∈Z ; nk>0
(cid:88)
(cid:88)
∞(cid:88)
∞(cid:88)
k=1
k∈Z ; nk>0
(1 + δk)−p +
= 2
1 +
(1 + δk)−p
(cid:88)
(cid:88)
k∈Z ; nk>0
(1 + δnk + xk)−p +
(1 + δnk − xk)−p
(1 + δnk)−p +
(1 + δnk − δ)−p
k∈Z ; nk>0
(1 + δ(k − 1))−p
∞(cid:88)
(cid:33)
≤ 2(cid:0)1 + (1 + δ)−p(δ−1 + p)(p − 1)−1(cid:1) .
k=1
k=1
The last expression is independent of t, and the claim follows.
Remark 1. When the set A = {ak : k ∈ Z} ⊂ R is relatively δ−separated, meaning
then the estimate (b) in Lemma 2.2 becomes
rel(A) := max
t∈R #{A ∩ ([0, δ] + t)} < ∞ ,
(1 + t − ak)−p ≤ 2 rel(A)(cid:0)1 + (1 + δ)−p(δ−1 + p)(p − 1)−1(cid:1) .
(cid:88)
k∈Z
ess supt∈R
Notice, that for a separated set, rel(A) = 1.
3. Nonstationary Gabor frames
Nonstationary Gabor systems provide a generalization of the classical Gabor systems of
time-frequency-shifted versions of a single window function.
Definition 3.1. Let g = {gk ∈ L2(R) : k ∈ Z} be a set of window functions and let b =
{bk : k ∈ Z} be a corresponding sequence of frequency-shift parameters. Set gk,l = Mbklgk.
Then, the set
G(g, b) = {gk,l : k, l ∈ Z}
is called a nonstationary Gabor system.
4
Note that, conceptually, we assume that the windows gk are centered at points {ak :
k ∈ Z}, in direct generalization of the regular case, where gk(t) = g(t − ak) for some time-
shift parameter a.
In this sense, we have a two-fold generalization: the sampling points
can be irregular and the windows can change for every sampling point. We are interested
in conditions under which a nonstationary Gabor system forms a frame. We first recall
the case of nonstationary Gabor frames with compactly supported windows, see [2] and
http://www.univie.ac.at/nonstatgab/ for further information.
3.1. Compactly supported windows: the painless case
Based on the support length of the windows gk, we can easily determine frequency-shifts
parameters bk, for which we obtain a frame. The following result is the nonstationary version
of the result given in [7].
Proposition 3.2 ([2]). Let g = {gk ∈ L2(R) : k ∈ Z} be a collection of compactly supported
functions with supp gk ≤ 1/bk. Then G(g, b) is a frame for L2(R) if there exist constants
A > 0 and B < ∞ such that
A ≤ G0(t) =
k gk(t)2 ≤ B a.e. .
b−1
k∈Z
The dual atoms are then γk,l(t) = MlbkG−1
Remark 2. An analogous theorem holds for bandlimited functions gk.
0 (t)gk(t).
The above theorem follows from the fact that the frame operator associated to the col-
lection of atoms described in the theorem can be written as
Sg,gf (t) =
k gk(t)2f (t) a.e. .
b−1
The diagonality of the frame operator in the painless case is derived from a generalized
Walnut representation for the frame operator Sg,g of nonstationary Gabor frames. In the
next section we will see that this representation immediately implies diagonality of Sg,g under
the assumptions of Proposition 3.2.
3.2. A Walnut representation for nonstationary Gabor Frames
Let us now consider nonstationary Gabor systems G(g, b) and G(γ, b), with all windows
gk and γk in W (L∞, (cid:96)1). The operator associated to G(g, b) and G(γ, b) reads
Sg,γf =
(cid:104)f, Mlbkgk(cid:105)Mlbkγk .
Proposition 3.3. The operator Sg,γ in (3) admits a Walnut representation
(cid:88)
k,l∈Z
Sg,γf =
k,l · Tlb−1
Gg,γ
k
f , where Gg,γ
k,l (t) = b−1
k gk(t − lb−1
k )γk(t) ,
5
(cid:88)
(cid:88)
k∈Z
(cid:88)
k,l∈Z
(3)
(4)
for f ∈ L2(R). Moreover, its operator norm can be bounded by
(cid:104)Sg,γf, h(cid:105) ≤(cid:16)
·(cid:16)
(cid:17)1/2(cid:16)
k )(cid:107)γk(cid:107)W (L∞,(cid:96)1)
(cid:17)1/2(cid:16)
k )(cid:107)gk(cid:107)W (L∞,(cid:96)1)
(1 + b−1
sup
k∈Z
(1 + b−1
sup
k∈Z
ess supt∈R
ess supt∈R
k∈Z
gk(t)(cid:17)1/2
(cid:88)
γk(t)(cid:17)1/2(cid:107)f(cid:107)2(cid:107)h(cid:107)2
(cid:88)
k∈Z
(5)
for all f, h ∈ L2(R).
Proof. First assume that f, h ∈ L2(R) are compactly supported. Since (cid:104)f, Mlbkgk(cid:105) =
(cid:91)(f ¯gk)(lbk) we can write Sg,γ as
Sg,γf (t) =
(cid:91)(f ¯gk)(lbk)Mlbkγk(t) =
mk(t)γk(t) ,
(6)
(cid:88)
k,l∈Z
(cid:88)
k∈Z
where mk(t) =(cid:80)
l∈Z
and by the Poisson formula can be written as
(cid:91)(f ¯gk)(lbk)e2πilbkt, for every k ∈ Z. The functions mk are b−1
k periodic
mk(t) = b−1
k
(f ¯gk)(t − lb−1
k ) .
(7)
(cid:88)
l∈Z
Therefore, substituting (7) in (6) yields the Walnut representation.
We next prove the boundedness (5). In the following chain of inequalities, we will use
Cauchy-Schwartz inequality for sums and integrals and, since all summands have absolute
value, Fubini’s theorem to justify changing the order of summation and integral. We thus
find
(cid:104)Sg,γf, h(cid:105) =
k )γk(·)f (· − lb−1
k gk(· − lb−1
b−1
k,l∈Z
k
R
k,l∈Z
b−1
(cid:12)(cid:12)(cid:12)(cid:68)(cid:88)
(cid:90)
≤ (cid:88)
(cid:20)(cid:90)
≤ (cid:88)
(cid:34)(cid:88)
(cid:90)
(cid:34)(cid:90)
f (t)2(cid:88)
b−1
b−1
k,l∈Z
k,l∈Z
≤
=
R
R
k
k
R
k,l∈Z
gk(t − lb−1
k )γk(t)f (t − lb−1
gk(t − lb−1
k )γk(t)f (t − lb−1
gk(t)γk(t + lb−1
k )f (t)2 dt
k gk(t)γk(t − lb−1
b−1
k ) dt
k ), h
(cid:69)(cid:12)(cid:12)(cid:12)
k )h(t) dt
(cid:21)1/2(cid:20)(cid:90)
(cid:35)1/2(cid:34)(cid:88)
(cid:35)1/2(cid:34)(cid:90)
k )2 dt
k,l∈Z
k
gk(t − lb−1
(cid:21)1/2
k )γk(t)h(t)2 dt
(cid:35)1/2
k )h(t)2 dt
(cid:35)1/2
k ) dt
k γk(t)gk(t − lb−1
b−1
γk(t)gk(t − lb−1
R
b−1
(cid:90)
h(t)2(cid:88)
R
k,l∈Z
.
(8)
R
6
The first term in the last expression can be bounded as follows
(cid:90)
R
f (t)2(cid:88)
≤(cid:88)
k,l∈Z
(cid:88)
k
R
b−1
k ess supt∈R
k gk(t)γk(t − lb−1
k ) dt =
b−1
(cid:32)
(cid:88)
(cid:32)
(cid:88)
(cid:32)
(cid:88)
b−1
k ess supt∈R
b−1
k ess supt∈R
l∈Z
b−1
(cid:88)
k∈Z
γk(t − lb−1
k )
(cid:90)
(cid:33)(cid:90)
(cid:33)(cid:88)
k )
γk(t − lb−1
(cid:33)(cid:32)
k )
γk(t − lb−1
k∈Z
l∈Z
R
(cid:90)
R
k∈Z
≤ sup
k∈Z
≤ sup
k∈Z
γk(t − lb−1
k )f (t)2gk(t) dt
l∈Z
f (t)2gk(t) dt
f (t)2gk(t) dt
(cid:33)
(cid:88)
k∈Z
Using relation (2) for ess supt∈R(cid:80)
l∈Z
l∈Zγk(t − lb−1
k ) and substituting (9) into (8) yields (5).
The estimate for the second term in (8) is obtained analogously. By the density of compactly
supported functions in L2(R), the estimate holds for all of L2(R).
ess supt∈R
gk(t)
(cid:107)f(cid:107)2
2 .
(9)
Remark 3. From (8) in the proof of Proposition 3.3, it follows that the operator Sg,γ is also
bounded by
Remark 4. In the case of a frame operator Sg,g, the Walnut representation (4) becomes
and the above bounds reduce to
(cid:104)Sg,gf, h(cid:105) ≤(cid:16)
(cid:32)
(cid:104)Sg,gf, h(cid:105) ≤
(1 + b−1
sup
k∈Z
ess supt∈R
k )(cid:107)gk(cid:107)W (L∞,(cid:96)1)
(cid:88)
k gk(t − lb−1
b−1
k,l∈Z
(10)
k ) a.e. ,
k )gk(t)f (t − lb−1
(cid:17)(cid:16)
gk(t)(cid:17)(cid:107)f(cid:107)2(cid:107)h(cid:107)2
(cid:88)
(cid:33)
k )gk(t)
k∈Z
(cid:107)f(cid:107)2(cid:107)h(cid:107)2
ess supt∈R
multiplication operator Sg,gf =(cid:80)
k∈Z Gg,g
k,0 · f = G0 · f .
Remark 5. Note that in the painless case of Theorem 3.2, the frame operator reduces to the
Remark 6. In the standard setting of Gabor frames, i.e. gk(t) = g(t − ak) for fixed a > 0,
and bk = b for all k ∈ Z, the above bound reduces to the well know bound
(cid:104)Sg,gf, h(cid:105) ≤ (1 + a−1)(1 + b−1)(cid:107)g(cid:107)2
W (L∞,(cid:96)1)(cid:107)f(cid:107)2(cid:107)h(cid:107)2
7
(cid:33)1/2
k )γk(t)
(cid:33)1/2
k )gk(t)
(cid:104)Sg,γf, h(cid:105) ≤
ess supt∈R
k gk(t − lb−1
b−1
(cid:32)
(cid:32)
(cid:88)
(cid:88)
k,l∈Z
k,l∈Z
(cid:88)
k,l∈Z
Sg,gf (t) =
k gk(t − lb−1
b−1
·
ess supt∈R
k γk(t − lb−1
b−1
(cid:107)f(cid:107)2(cid:107)h(cid:107)2 .
3.3. Existence of nonstationary Gabor frames
For windows gk that are neither compactly supported nor bandlimited, we are interested
in the existence of frames of the form G(g, b) and in the construction of the involved param-
eters. The following theorem derives a sufficient condition for the existence of nonstationary
Gabor frames and shows that this condition can be satisfied.
In this and the subsequent sections, [bL, bU ], [pL, pU ], [CL, CU ] are compact intervals of
positive real numbers.
Theorem 3.4. Let g = {gk ∈ W (L∞, (cid:96)1) : k ∈ Z} be a set of windows such that
i) for some positive constants A0, B0
0 < A0 ≤(cid:88)
k∈Z
gk(t)2 ≤ B0 < ∞ a.e. ;
(11)
ii) for all k ∈ Z, the windows decay polynomially around a δ-separated set {ak : k ∈ Z}
of time-sampling points ak
gk(t) ≤ Ck(1 + t − ak)−pk ,
(12)
where pk ∈ [pL, pU ] ⊂ R, pL > 2 and Ck ∈ [CL, CU ].
k}k∈Z, such that for bk ≤ b0
(cid:90)
Then there exists a sequence {b0
system G(g, b) forms a frame for L2(R).
Proof. Let f ∈ L2(R). Applying (10), we write
(cid:104)Sg,gf, f(cid:105) =
(cid:88)
k gk(t)2f (t)2 dt +
b−1
(cid:88)
(cid:88)
(cid:90)
k∈Z
R
R
(cid:88)
l∈Z\{0}
(cid:88)
k∈Z
Using similar arguments as in the derivation of (8), we obtain
(cid:12)(cid:12)(cid:12) ≤
k )f (t) dt
(cid:12)(cid:12)(cid:12)(cid:90)
R
l∈Z\{0}
k∈Z
k, k ∈ Z, the nonstationary Gabor
k gk(t)gk(t − lb−1
b−1
k )f (t − lb−1
k )f (t) dt
≤
k∈Z
l∈Z\{0}
≤ max
k∈Z {b−1
k )f (t − lb−1
k gk(t)gk(t − lb−1
b−1
ess supt∈R
(cid:88)
(cid:88)
(cid:32)
k } (cid:88)
(cid:124)
(cid:32)
(cid:32)
2 ≥ min
ess inf t∈R
l∈Z\{0}
k }
k∈Z{b−1
k∈Z {b−1
k }
ess supt∈R
ess supt∈R
k gk(t)gk(t − lb−1
k )
b−1
(cid:107)f(cid:107)2
(cid:33)
k )
gk(t)gk(t − lb−1
(cid:125)
2
(cid:88)
(cid:123)(cid:122)
k∈Z
R
(cid:88)
k }
gk(t)2 − maxk∈Z{b−1
k } R
mink∈Z{b−1
(cid:88)
(cid:33)
gk(t)2 + R
k∈Z
,
k∈Z
(cid:104)Sg,gf, f(cid:105)(cid:107)f(cid:107)−2
(cid:104)Sg,gf, f(cid:105)(cid:107)f(cid:107)−2
2 ≤ max
(cid:107)f(cid:107)2
2 .
(cid:33)
(13)
Therefore, lower and upper frame bounds are obtained from
8
We need to construct a sequence of bk, k ∈ Z, such that for all f ∈ L2(R), (13) is bounded
away from zero.
)1/pk. Then mink∈Z{b−1
(CL−1)1/p2, maxk∈Z{b−1
Let < CL and consider the sequence of frequency shifts bk = (
Ck
k } ≥
k } ≤ (CU −1)1/p1 and
maxk∈Z{b−1
k }
k } ≤ C 1/pL
mink∈Z{b−1
U C
−1/pU
L
1/pU−1/pL .
(14)
Since (1 + x + y)−p ≤ (1 + x)p(1 + y)−p for x, y ∈ R and p ≥ 0, using (12), we have, for
some µ with pL − 2 > µ > 0:
gk(t)gk(t − lb−1
k
(cid:125)
k )−pk+(1+µ)
k )−pk+(1+µ)
(1 + t − ak)−(1+µ) l−pk+(1+µ) 1−(1+µ)/pk .
k ) ≤ C 2
≤ C 2
≤ C 2
= C 1+(1+µ)/pk
k(1 + t − ak)−pk(1 + t − ak − lb−1
k(1 + t − ak)(1+µ)(1 + lb−1
k(1 + t − ak)−(1+µ) l−pk+(1+µ) bpk−(1+µ)
k(cid:124)
(cid:123)(cid:122)
k ) ≤(cid:88)
k∈Z Ek l−pL+(1+µ) 1−(1+µ)/pL (cid:88)
k∈Z Ek l−pL+(1+µ) 1−(1+µ)/pL 2(cid:0)1 + (1 + δ)−(1+µ)(δ−1 + 1 + µ)µ−1(cid:1) ,
Ek(1 + t − ak)−(1+µ) l−pk+(1+µ) 1−(1+µ)/pk
(1 + t − ak)−(1+µ)
k∈Z
k∈Z
Ek
Hence,(cid:88)
k∈Z
gk(t)gk(t − lb−1
≤ max
≤ max
(15)
where the last estimate follows from Lemma 2.2 (b). Summing the expression (15) over
l ∈ Z \ {0} using Lemma 2.2 ¡ , we see that R, as a function of , behaves like 1−(1+µ)/pL,
i.e. R ≈ 1−(1+µ)/pL, and R tends to 0 for → 0. Moreover,
maxk∈Z{b−1
k }
k } R ≈ 1−(2+µ)/pL+1/pU
mink∈Z{b−1
can be made arbitrarily small by choosing small since 1− (2 + µ)/pL + 1/pU > 0. Therefore,
)1/pk, maxk∈Z{b−1
k }
k } R < A0, then {Mlbkgk}k,l∈Z is a frame for all
if 0 is such that for b0
mink∈Z{b−1
bk ≤ b0
k.
k := ( 0
Ck
For completeness, we state the equivalent result for analysis windows gk with polynomial
decay in the frequency domain.
Corollary 3.5. Let g = {gk ∈ L2(R) : gk ∈ W (L∞, (cid:96)1), k ∈ Z} be a set of windows such
that
i) for some positive constants A0, B0
gk(t)2 ≤ B0 < ∞ a.e. ;
(16)
0 < A0 ≤(cid:88)
k∈Z
9
ii) for all k ∈ Z, the windows decay polynomially around a δ-separated set {bk : k ∈ Z}
of frequency-sampling points bk:
gk(t) ≤ Ck(1 + t − bk)−pk ,
(17)
where pk and Ck are chosen as in Theorem 3.4.
Then there exists a sequence {a0
system {Tlakgk : k, l ∈ Z} forms a frame for L2(R).
k}k∈Z, such that for ak ≤ a0
k, k ∈ Z, the nonstationary Gabor
3.3.1. Nonstationary Gabor frames on modulation spaces
Modulation spaces, cf. [13, 15], are considered as the appropriate function spaces for
time-frequency analysis and in particular, for the study of Gabor frames. By their definition,
modulation spaces require decay in both time and frequency. Under additional assumptions
on the windows gk, the collection G(g, b) is a frame for all modulation spaces M p, 1 ≤ p ≤ ∞.
Proposition 3.6. Let G(g, b) be a frame for L2(R) satisfying the uniform estimate
Vφgk(x, ω) ≤ C(1 + x − ak)−r−2(1 + ω)−r−2 ,
r > 2
(18)
where φ is a Gaussian window. Then the frame operator S is invertible simultaneously on
all modulation spaces M p for 1 ≤ p ≤ ∞.
Proof. Notice, that
Vφgk,l(x, ω) ≤ C(1 + (x − ak, ω − lbk))−r−2 ,
since (1+x+ω)−r ≥ (1+x)−r(1+ω)−r and the weights (1+(x, ω))r and (1+x+ω)r
are equivalent.
then the frame operator Sf =(cid:80)
A result on Gabor molecules [16] states that, if an L2-frame {gz : z = (z1, z2) ∈ Z ⊆ R2},
where Z is separable, satisfies the uniform estimate Vφgz(x, ω) ≤ C(1+(x−z1, ω−z2))−r−2,
z∈Z(cid:104)f, gz(cid:105)gz is invertible simultaneously on all M p for each
1 ≤ p ≤ ∞. The result hence follows from condition (18).
3.4. Constructing nonstationary Gabor frames
Theorem 3.4 shows that for windows with sufficient uniform decay, nonstationary Gabor
frames can always be constructed by choosing sufficient density in the frequency samples.
In the present section we assume the existence of a certain nonstationary Gabor frame and
explicitly construct a new frame by exploiting the prior information about the original one.
This is a situation of practical relevance, since we may often be interested in using windows
that decay fast and are negligible outside a support of interest. In particular, we will use
the fact that painless nonstationary Gabor frames are easily constructed and deduce new
frames from painless frames. The new frames thus obtained will be called almost painless
nonstationary Gabor frames.
We will subsequently assume bk ∈ [bL, bU ] for frequency-shift parameters and we let Ck ∈
[CL, CU ] and pk ∈ [pL, pU ] with pL > 1 for the constants involved in the decay assumptions
10
for the analysis windows. We then work with the following constants that depend on the
separation of the sampling points, the decay of the windows and the frequency-sampling
parameters:
E1 = 1 +
δ−1 + pL
(1 + δ)pL(pL − 1)
and E2 = 1 +
(1 + b−1
bU + pU
U )pL(pL − 1)
(19)
We first consider nonstationary Gabor frames obtained by perturbation of a known frame.
This result is in the spirit of similar results for regular Gabor frames [3, 6].
Proposition 3.7. Let {ak : k ∈ Z} be a δ-separated set and G(h, b) a nonstationary Gabor
frame with frame bounds Ah, Bh and frame operator Sh,h. Let gk ∈ L2(R) be a set of windows
such that for all k ∈ Z and for almost all t ∈ R
gk(t) − hk(t) ≤ Ck(1 + t − ak)−pk.
(20)
√
(cid:113)
If CU <
Ahλ−1 for
(21)
h )2.
λ = 4b−1
U λB−1
C 2
L · E1 · E2,
Bh(1 +
Proof. By applying Cauchy-Schwartz inequality, it is easy to see that, for G(g, b) to be a
2 for some R < Ah,
then G(g, b) is a frame for L2(R) with frame bounds A = Ah(1 −(cid:113)
frame for L2(R), it suffices to show that(cid:80)
also cf. [6, Proposition 4.1.]. Then, frame bounds of G(g, b) can be taken as Ah(1−(cid:112)R/Ah)2
and Bh(1 +(cid:112)R/Bh)2.
in (9):(cid:88)
To obtain the required error bound, we let ψk(t) := gk − hk and use the estimate given
k,l∈Z(cid:104)f, gk,l − hk,l(cid:105)2 ≤ R(cid:107)f(cid:107)2
h )2 and B =
U λA−1
C 2
ψk(t)(cid:17)(cid:107)f(cid:107)2
(cid:88)
ψk(t − lb−1
k )(cid:17)(cid:16)
b−1
k ess supt∈R
ess supt∈R
2 .
(22)
(cid:16)
(cid:104)f, ψk,l(cid:105)2 ≤ sup
k∈Z
k,l∈Z
(cid:88)
l∈Z
k∈Z
The first term of the above in the above inequality is bounded by assumption (20) and
we can simplify
ess supt∈R
(cid:88)
l∈Z
k )−pk is b−1
l∈Z(1 + t − ak − lb−1
(cid:88)
ψk(t − lb−1
k ) ≤ Ckess supt∈[0,b−1
(1 + t − lb−1
k )−pk .
k ]
l∈Z
k −periodic, therefore
(cid:88)
Lemma 2.2 (b):(cid:88)
To bound the second term, note that(cid:80)
ψk(t) ≤ CU
k∈Z
k∈Z
(cid:88)
k∈Z
(1 + t − ak)−pk ≤ CU
(1 + t − ak)−p1 ≤ 2CU E1 .
11
Hence, by Lemma 2.2, we obtain for t ∈ [0, b−1
k ]:
(cid:88)
l∈Z
(1 + t − lb−1
(1 + (l − 1)b−1
∞(cid:88)
k )−pk ≤ 1 +
∞(cid:88)
k )−pk
l=1
= 2
∞(cid:88)
k )−pk +
(cid:16)
∞(cid:88)
l=1
1 +
(1 + t − lb−1
(1 + t + lb−1
k )−pk
(1 + lb−1
k )−pk = 2
(1 + lb−1
k )−pk
l=0
l=1
(cid:17) ≤ 2E2.
Gathering all the estimates, we obtain
(cid:104)f, ψk,l(cid:105)2 ≤ C 2
U 4b−1
1 E1 · E2(cid:107)f(cid:107)2
2 = C 2
U λ(cid:107)f(cid:107)2
2 .
(cid:88)
k,l∈Z
By assumption C 2
U λ < Ah, and the proof is complete.
Using Proposition 3.7 we next construct a special class of nonstationary Gabor frames by
relying on knowledge of a painless nonstationary Gabor frame. We construct new windows
which are no more compactly supported, but coincide with the known, compact windows
on their support. We call the resulting new systems almost painless nonstationary Gabor
frames.
Corollary 3.8. Let g = {gk ∈ W (L∞, (cid:96)1) : k ∈ Z} be a set of windows, and let Ik be
the intervals Ik = [ak − (2bk)−1, ak + (2bk)−1] where {ak : k ∈ Z} forms a δ-separated set.
Assume that G(h, b), where hk = gkχIk, is a Gabor frame with lower frame bound Ah, and
that for ψk = gk − hk, all k ∈ Z and almost all tinR
Ck(1 + t − ak − (2bk)−1)−pk ,
ψk(t) ≤
0 ,
Ck(1 − t + ak − (2bk)−1)−pk ,
L · δ−1 · E1 · E2, then G(g, b) forms a nonstationary Gabor frame
(23)
t > ak + (2bk)−1 ;
t ∈ Ik ;
t < ak − (2bk)−1 .
√
If CU <
for L2(R).
Ahλ−1 for λ = 4b−2
Proof. The proof follows the steps of the proof of Proposition 3.7 with small changes on how
to approximate the terms in (22). First observe, that for any t ∈ R and all k, we have
(cid:2)(1 + t − ak − (2bk)−1)−pk + (1 + t − ak + (2bk)−1)−pk(cid:3)
ψk(t) ≤ Ck
Since the frequency shifts bk are taken from a finite interval and the set {ak : k ∈ Z} is
δ−separated, the sets Γ+ = {ak + (2bk)−1 : k ∈ Z} and Γ− = {ak − (2bk)−1 : k ∈ Z}
are relatively δ−separated with rel(Γ) = rel(Γ+) = rel(Γ−) = (cid:98)(2bLδ)−1(cid:99). Therefore, by
Remark 1, it follows that
ψk(t) ≤ CU
(1 + t − ak − (2bk)−1)−pL + CU
(1 + t − ak + (2bk)−1)−pL
(cid:88)
k∈Z
(cid:88)
k∈Z
(cid:88)
k∈Z
≤ 4 CU rel(Γ) (1 + (1 + δ)−pL(δ−1 + pL)(pL − 1)−1) .
12
l∈Zψk(t − lb−1
k ) is b−1
k −periodic. Let t ∈ Ik, then, by (23)
(1 + ak − (2bk)−1 − t + lb−1
k )−pk
(cid:35)
(1 − ak − (2bk)−1 + t − lb−1
k )−pk
Now, the expression(cid:80)
(cid:34)(cid:88)
(cid:88)
k ) ≤ Ck
(cid:88)
ψk(t − lb−1
l∈Z
l>0
+
l<0
≤ Ck2
∞(cid:88)
(1 + lb−1
k )−pk ≤ 2Ck(1 + (1 + b−1
k )−pk(bk + pk)(pk − 1)−1) ,
where the last estimate follows from Lemma 2.2(a).
l=0
4. Examples
We illustrate our theory with two examples.
In both examples, we consider a basic
window and dilations by 2 and 1
2, respectively. Since the dilation parameters take only three
different values, there are three kinds of windows, with support size 1/2, 1 and 2, respectively.
Note that, while theoretically possible, sudden changes in the shape and width of adjacent
windows turn out to be undesirable for applications, hence we only allow for stepwise change
in dilation parameters.
In the first example we consider a frame that arises as perturbation of a painless non-
stationary Gabor frame. The perturbation consists in the application of a bandpass filter in
order to obtain windows with compact support in the frequency domain.
Example 4.1. Let h be a Hann or raised cosine window, i.e. h(t) = 0.5 + 0.5 cos(2πt) for
t ∈ [−1/2, 1/2], and zero otherwise. We construct a painless nonstationary Gabor frame
2, respectively: Let {sk}k∈Z be a sequence with values from the set
by dilating h by 2 or 1
{−1, 0, 1} with the restriction that sk − sk−1 ∈ {0, 1} to avoid sudden changes between
adjacent windows. We then define corresponding shift-parameters by setting a0 = 0 and
ak+1 = ak + 2−sk · 5
6
ak+1 = ak + 2−sk+1 · 1
3
ak+1 = ak + 2−sk+1 · 5
6
if
if
sk > sk+1 ,
sk = sk+1 ,
if
sk < sk+1 .
The points ak, k ∈ Z, form a separated set with minimum separation δ = 1/3. Setting
2skh(2skt), the system {Mlbkhk : k, l ∈ Z} forms a painless
bk = 2sk and hk(t) = Tak
nonstationary Gabor frame with lower frame bound Ah = 0.5.
√
Let Ω = 0.02 and φ be a bandlimited filter given by
(cid:98)φ(ω) = 0.5 + 0.5 cos(2πΩ−1ω) on its support [−Ω/2, Ω/2] .
We build new windows gk by convolving φ with hk. The windows gk := φ ∗ hk are no more
compactly supported. Since φ(t) ≤ Ω(1 + t)−3, we rely on [20, Theorem 9.9] to deduce
13
the following bound, with C(cid:48) = (cid:107)hk(cid:107)∞ Ω
(1 + (t − ak) − 2−s(cid:48)
2 , Ik = [ak − 2−s(cid:48)
k, ak + 2−s(cid:48)
k)−2 − (1 + (t − ak) + 2−s(cid:48)
k)−2 − (1 − (t − ak) + 2−s(cid:48)
k] and s(cid:48)
k)−2
k)−2 − (1 − (t − ak) + 2−s(cid:48)
k)−2
gk(t) − hk(t) ≤ C(cid:48)
2 − (1 + (t − ak) + 2−s(cid:48)
(1 − (t − ak) − 2−s(cid:48)
t > ak + 2−s(cid:48)
t ∈ Ik
t < ak − 2−s(cid:48)
We obtain the joint bound, gk(t)− hk(t) ≤ CU (1 +t− ak)−2 by setting Ck = C(cid:48)· (1 + 2−s(cid:48)
√
k)2
for all k ∈ Z and CU = maxk∈Z Ck = 0.0282 <
Ahλ−1 = 0.0768, with λ as defined in (21).
Thus, by Proposition 3.7, {Mlbkgk : k, l ∈ Z} is a Gabor frame with a lower frame bound
A = 0.2.
k = sk + 1:
k)−2
k
k
Remark 7. Note that the construction presented in the Example 4.1 is of particular interest
for constructing frequency-adaptive frames with windows that are compactly supported in
time. This is a situation of considerable interest in applications, since it allows for real-time
implementation with finite impulse response filters, cp. [11].
In the second example we construct a nonstationary Gabor frame by applying Corol-
lary 3.8. The windows of the new frame coincide with the windows of a painless frame on
their support. The windows in this example are constructed in analogy to the windows used
in scale frames, introduced in [2] to automatically improve the resolution of transients in
audio signals.
Example 4.2. As in the previous example, let sk ∈ {−1, 0, 1} with sk − sk−1 ∈ {0, 1} for all
k ∈ Z. We consider a sequence of windows gk that are translated and dilated versions of the
Gaussian window g(t) = e−π(2.5t)2: gk(t) = Tak
ak+1 = ak + 2−sk+1−1
· 2−sk+1
· 2−sk
2skg(2skt) with a0 = 0 and
ak+1 = ak +
sk = sk+1 ,
sk > sk+1 ,
ak+1 = ak +
sk < sk+1 .
√
if
if
if
1
3
1
3
Here, the {ak : k ∈ Z} are separated with minimum distance δ = 1/4. We arrange the
windows as follows: after each change of window size, no change is allowed in the next step;
in other words, each window has at least one neighbor of the same size. An example of the
arrangement is shown in Figure 1.
Let Ik = [ak − 2−sk−1, ak + 2−sk−1] and define a new set of windows by hk(t) = gk(t)χIk.
Then {Mlbkhk : k, l ∈ Z} with bk = 2sk is a painless nonstationary Gabor frame. By
numerical calculations, its lower frame bound is Ah = 0.1609 and ψk(t) = gk(t) − hk(t) can
be bounded by
√
√
0
ψk(t) ≤
2skg(1/2)(1 + t − ak − 2−sk−1)−19
2skg(1/2)(1 + ak − 2−sk−1 − t)−19
t > ak + 2−sk−1
t ∈ Ik
t < ak − 2−sk−1 .
From Proposition 3.8 it follows that 4(δb2
k, l ∈ Z} is a nonstationary Gabor frame with a lower frame bound A = 0.1538.
L)−1 · C 2
U · E1 · E2 = 0.0071 < Ah, and {MlbkTakgk :
14
Figure 1: An example for the arrangement of dilated windows in Example 2
5. Acknowledgement
This work was supported by the WWTF project Audiominer (MA09-24) and the Austrian
Science Fund (FWF):[T384-N13] Locatif.
References
[1] A. Aldroubi, C. A. Cabrelli, and U. Molter. Wavelets on Irregular Grids with Arbitrary
Dilation Matrices, and Frame Atoms for L2(Rd). Appl. Comput. Harmon. Anal., Special
Issue on Frames II.:119–140, 2004.
[2] P. Balazs, M. Dorfler, F. Jaillet, N. Holighaus, and G. A. Velasco. Theory, imple-
mentation and applications of nonstationary Gabor Frames. J. Comput. Appl. Math.,
236:1481–1496, 2011.
[3] I. Carrizo and S. Favier. Perturbation of wavelet and Gabor frames. Anal. Theory
Appl., 19(3):238–254, 2003.
15
[4] O. Christensen. An Introduction to Frames and Riesz Bases. Applied and Numerical
Harmonic Analysis. Birkhauser, 2003.
[5] O. Christensen, S. Favier, and F. Z´o.
Irregular wavelet frames and Gabor frames.
Approx. Theory Appl., 17(3):90–101, 2001.
[6] O. Christensen and C. Heil. Perturbations of Banach frames and atomic decompositions.
Math. Nachr., 185:33–47, December 1997.
[7] I. Daubechies, A. Grossmann, and Y. Meyer. Painless nonorthogonal expansions. J.
Math. Phys., 27(5):1271–1283, May 1986.
[8] M. Dorfler. Quilted Gabor frames - A new concept for adaptive time-frequency repre-
sentation. Advances in Applied Mathematics, 47(4):668 – 687, Oct. 2011.
[9] N. Holighaus, M. Dorfler, G. Velasco, and T. Grill. A framework for invertible, real-time
constant-Q transforms. preprint, submitted, http://www.univie.ac.at/nonstatgab/slicq,
2012.
[10] R. J. Duffin and A. C. Schaeffer. A class of nonharmonic Fourier series. Trans. Amer.
Math. Soc., 72:341–366, 1952.
[11] G. Evangelista, M. Dorfler, and E. Matusiak. Phase Vocoders With Arbitrary Frequency
Band Selection. Proceedings of the 9th Sound and Music Computing Conference, July
11-14th 2012 Kopenhagen, 2012.
[12] S. J. Favier and R. A. Zalik. On the stability of frames and Riesz bases. Appl. Comput.
Harmon. Anal., 2(2):160–173, 1995.
[13] H. G. Feichtinger. On a new Segal algebra. Monatsh. Math., 92:269–289, 1981.
[14] D. Gabor. Theory of communication. J. IEE, 93(26):429–457, 1946.
[15] K. Grochenig. Foundations of Time-Frequency Analysis. Appl. Numer. Harmon. Anal.
Birkhauser Boston, 2001.
[16] K. Grochenig. Localization of Frames, Banach Frames, and the invertibility of the frame
operator. Proceedings of SPIE, San Diego, 2004.
[17] H. Malvar. Signal Processing with Lapped Transforms. Boston, MA: Artech House. xvi,
1992.
[18] J. L. Romero. Surgery of spline-type and molecular frames. J. Fourier Anal. Appl.,
17:135 – 174, 2011.
[19] A. Ron and Z. Shen. Weyl-Heisenberg frames and Riesz bases in L2(Rd). Duke Math.
J., 89(2):237–282, 1997.
[20] W. Rudin. Real and Complex Analysis. McGraw-Hill Book Company, New York, 1966.
16
[21] W. Sun and X. Zhou. Irregular wavelet/Gabor frames. Appl. Comput. Harmon. Anal.,
13(1):63–76, 2002.
[22] G. A. Velasco, N. Holighaus, M. Dorfler, and T. Grill. Constructing an invertible
constant-Q transform with non-stationary Gabor frames. In Proceedings of DAFX11,
Paris, September 2011.
[23] D. F. Walnut. Continuity properties of the Gabor frame operator. J. Math. Anal. Appl.,
165(2):479–504, 1992.
[24] Z. Xiong and H. Malvar. A nonuniform modulated complex lapped transform. IEEE
Signal Processing Letters, 8(9):257–260, September 2001.
17
|
1812.01419 | 3 | 1812 | 2019-05-30T18:59:57 | On existence of shift-type invariant subspaces for polynomially bounded operator | [
"math.FA"
] | A particular case of results from [K2] is as follows. Let the unitary asymptote of a contraction $T$ contain the bilateral shift (of finite or infinite multiplicity). Then there exists an invariant subspace $\mathcal M$ of $T$ such that $T|_{\mathcal M}$ is similar to the unilateral shift of the same multiplicity. The proof is based on the Sz.-Nagy--Foias functional model for contractions. In the present paper this result is generalized to polynomially bounded operators, but in the simplest case. Namely, it is proved that if the unitary asymptote of a polynomially bounded operator $T$ contains the bilateral shift of multiplicity $1$, then there exists an invariant subspace $\mathcal M$ of $T$ such that $T|_{\mathcal M}$ is similar to the unilateral shift of multiplicity $1$. The proof is based on a result from [B]. | math.FA | math |
ON EXISTENCE OF SHIFT-TYPE INVARIANT
SUBSPACES FOR POLYNOMIALLY BOUNDED
OPERATORS
MARIA F. GAMAL'
Abstract. A particular case of results from [K2] is as follows. Let the
unitary asymptote of a contraction T contain the bilateral shift (of finite
or infinite multiplicity). Then there exists an invariant subspace M of
T such that T M is similar to the unilateral shift of the same multi-
plicity. The proof is based on the Sz.-Nagy -- Foias functional model for
contractions. In the present paper this result is generalized to polynomi-
ally bounded operators, but in the simplest case. Namely, it is proved
that if the unitary asymptote of a polynomially bounded operator T
contains the bilateral shift of multiplicity 1, then there exists an invari-
ant subspace M of T such that T M is similar to the unilateral shift of
multiplicity 1. The proof is based on a result from [B].
1. Introduction
Let H be a (complex, separable) Hilbert space, and let L(H) be the
algebra of all (bounded, linear) operators acting on H. A (closed) subspace
M of H is called invariant for an operator T ∈ L(H), if TM ⊂ M. The
complete lattice of all invariant subspaces of T is denoted by Lat T . For a
(closed) subspace M of a Hilbert space H, by PM and IM the orthogonal
projection from H onto M and the identity operator on M are denoted,
respectively.
For Hilbert spaces H and K, let L(H,K) denote the space of (bound-
ed, linear) transformations acting from H to K. Suppose that T ∈ L(H),
R ∈ L(K), X ∈ L(H,K), and X intertwines T and R, that is, XT = RX.
If X is unitary, then T and R are called unitarily equivalent, in notation:
T ∼= R. If X is invertible, that is, X−1 ∈ L(K,H), then T and R are called
similar, in notation: T ≈ R. If X is a quasiaffinity, that is, ker X = {0}
and clos XH = K, then T is called a quasiaffine transform of R, in notation:
T ≺ R. If clos XH = K, we write T
≺ R. If T ≺ R and R ≺ T , then T and
R are called quasisimilar, in notation: T ∼ R. Clearly, T ≺ R if and only if
R∗ ≺ T ∗.
An operator T is called power bounded, if supn≥0 kT nk < ∞. An operator
T is called polynomially bounded, if there exists a constant C such that
kp(T )k ≤ C max{p(z) : z ≤ 1} for every (analytic) polynomial p. The
smallest such constant is called the polynomial bound of T and is denoted
here by Cpol,T .
d
2010 Mathematics Subject Classification. Primary 47A15, 47A60.
Key words and phrases. Shift-type invariant subspace, polynomially bounded operator,
similarity, unilateral shift.
1
2
M.F. GAMAL'
Every polynomially bounded operator can be represented as a direct sum
of an operator similar to a singular unitary operator and of an absolutely
continuous (a.c.) polynomially bounded operator, that is, an operator which
admits an H ∞-functional calculus, see [M] or [K3, Theorems 13, 17 and 23].
In the present paper, absolutely continuous polynomially bounded operators
are considered. (Although many results on polynomially bounded operators
that will be used in the present paper were originally proved by Mlak [M],
we will refer to [K3] for the convenience of references.)
An operator T is called a contraction, if kTk ≤ 1. Every contraction is
polynomially bounded with the constant 1 by the von Neumann inequality
(see, for example, [SFBK, Proposition I.8.3]). Clearly, every polynomially
bounded operator is power bounded. (It is well known that the converse
is not true, see [F] for the first example of a power bounded but not poly-
nomially bounded operator, and [P] for the first example of a polynomially
bounded operator which is not similar to a contraction.)
Let T ∈ L(H) be a power bounded operator. It is easy to see that the
space
HT,0 = {x ∈ H : kT nxk → 0}
is invariant for T (sf. [SFBK, Theorem II.5.4]). Classes Cab, a, b = 0, 1,·, of
power bounded operators are defined as follows (see [SFBK, Sec. II.4] and
[K1]). If HT,0 = H, then T is of class C0·, while if HT,0 = {0}, then T is of
class C1·. Furthermore, T is of class C·a, if T ∗ is of class Ca·, and T is of
class Cab, if T is of classes Ca· and C·b, a, b = 0, 1.
Let T and R be power bounded operators, and let T ≺ R.
It easily
follows from the definition that if R is of class C1· or of class C·0, then T
is of class C1· or of class C·0, too. Clearly, any isometry is of class C1·, a
unitary operator is of class C11, and the unilateral shift is of class C10.
The notions of isometric and unitary asymptotes will be used. Although
there exist further studies of these notions ([K3], [K4]), we restrict ourselves
to the case of power bounded operators. For a power bounded operator
T ∈ L(H) the isometric asymptote (X+, T (a)
+ ) can be defined using a Banach
limit, see [K1]. Here T (a)
+ is an isometry (on a Hilbert space), and X+ is
the canonical intertwining mapping: X+T = T (a)
+ X+. Recall that ker X+ =
d
≺ T (a)
HT,0, and the range of X+ is dense. Thus, T
+ . In particular, if T is
cyclic, then T (a)
+ is cyclic, too. A power bounded operator T is of class C11
if and only if T is quasisimilar to a unitary operator; then T (a)
+ is unitary
and T ∼ T (a)
The unitary asymptote (X, T (a)) of a power bounded operator T ∈ L(H)
is a pair where T (a) ∈ L(H(a)) (here H(a) is a some Hilbert space) is the
minimal unitary extension of T (a)
+ , and X is an extension of X+. Therefore,
+ (see [SFBK, Propositions II.3.4 and II.5.3]).
(1.1)
_n≥0
(T (a))−nXH = H(a).
SHIFT-TYPE INVARIANT SUBSPACES
3
In particular, if T is cyclic, then T (a) is cyclic, too. If T (a)
of course, (X+, T (a)
unitarily equivalence.
+ ) = (X, T (a)). Note that T (a)
+ is unitary, then,
+ and T (a) are defined up to
Set T0· = THT ,0 and T1· = PHT ,0THT ,0. By [K1, Lemma 1], T0· and T1·
are of classes C0· and C1·, respectively. Thus, every power bounded operator
has the triangulations of the form
(cid:18)C0·
O C1·(cid:19) and (cid:18)C·1
O C·0(cid:19) .
∗
∗
By [K1, Theorem 3], T (a) ∼= T (a)
1·
.
d
T
A particular case of [K2] is the following (see also [SFBK, Sec.
other hand, it follows from (1.1) that T
realized by the canonical intertwining mapping in general).
Let T be a power bounded operator, let U be a unitary operator, and let
d
≺ U . By [K1, Theorem 2], T (a) contains U as an orthogonal summand,
that is, there exists a unitary operator V such that T (a) ∼= U ⊕ V . On the
≺ T (a) (but this relation is not
IX.3]
and references in [K2] and [SFBK] for the history of a question). Let T
be a contraction, and let T (a) contain the bilateral shift of finite or infinite
multiplicity as an orthogonal summand. Then there exists M ∈ Lat T
such that TM is similar to the unilateral shift of the same multiplicity.
The proof is based on the Sz.-Nagy -- Foias functional model for contractions,
see [SFBK]. In the present paper this result is generalized to polynomially
bounded operators, but in the simplest case. Namely, it is proved that if T
is a polynomially bounded operator and T (a) contains the bilateral shift of
multiplicity 1, then there exists M ∈ Lat T such that TM is similar to the
unilateral shift of multiplicity 1. The proof is based on a result from [B].
Before formulating the main result of the paper, we introduce some no-
tation. T and D denote the unit circle and the open unit disc, respectively.
S is the unilateral shift of multiplicity 1, that is, S is the operator of mul-
tiplication by the independent variable on the Hardy space H 2 on T. The
normalized Lebesgue measure on T is denoted by m. For a measurable set
σ ⊂ T denote by Uσ the operator of multiplication by the independent vari-
able on L2(σ) := L2(σ, m). Clearly, UT is the bilateral shift of multiplicity
1. It is well known and easy to see that S and Uσ are a.c. contractions.
Theorem 1.1. Suppose that T is a cyclic a.c. polynomially bounded oper-
ator of class C1·, and T (a) ∼= UT. Then there exists M ∈ Lat T such that
TM ≈ S.
Corollary 1.2. Suppose that T is a polynomially bounded operator such that
≺ UT. Then for every c > 0 there exist M ∈ Lat T and W ∈ L(H 2,M)
T
such that W is invertible, W S = TMW and
d
kWkkW −1k ≤ (1 + c)(cid:0)√2(K 2 + 2)KCpol,T + 1(cid:1)qK 2C 2
Remark 1.3. Examples of polynomially bounded operators satisfying The-
orem 1.1 and not similar to contractions can be found in [G]. Moreover,
it follows from Theorem 1.1 and Corollary 1.2 that if T is a polynomially
pol,T + 1KC 2
pol,T .
4
M.F. GAMAL'
bounded operator such that T ∼ UT and the product of intertwining quasi-
affinities is an analytic function of UT, or if T ≺ S, then T is similar to an
operator constructed in [G, Proposition 2.7 or Corollary 2.3], respectively.
Remark 1.4. Let T satisfy the assumption of Corollary 1.2. By [K1, Theo-
rem 4], T ⊂ σ(T ). By [R], T either has a nontrivial hyperinvariant subspace
or is reflexive. Therefore, the result on existence of nontrivial invariant sub-
spaces of T is not new. The reflexivity of contractions satisfying the assump-
tions of Corollary 1.2 is proved in [T1] and [T2], see also [SFBK, Theorem
IX.3.8]. This proof can be generalized to polynomially bounded operators.
We expect to give a detailed proof later elsewhere.
Remark 1.5. Suppose that N ∈ N and T is a polynomially bounded oper-
≺ ⊕N
n=1UT. It is possible to prove by induction that there
ator such that T
exists M ∈ Lat T such that TM ≈ ⊕N
n=1S. In the case of infinite sum the
question remains open.
d
The paper is organized as follows. In Sec. 2, Bourgain's result [B] is cited,
and its corollaries which hold true for arbitrary a.c. polynomially bounded
operators are given. In Sec. 3 and 4, auxiliary results for some functions
and for operators intertwined with unitaries, respectively, are given. The
main part of the paper is Sec. 5, where Theorem 1.1 is proved. In Sec. 6
Corollary 1.2 is proved.
2. Bourgain's result
A denotes the disc algebra. For a positive finite Borel measure µ on T
set P 2(µ) = closL2(µ) A, and denote by Sµ the operator of multiplication by
the independent variable on P 2(µ).
Theorem A ([B]). There exists a universal constant K with the following
property. Let H be a Hilbert space, and let W ∈ L(A,H) (that is, W is a
bounded linear transformation from A to H). Then there exists a positive
Borel measure µ on T such that µ(T) = 1 and
kWϕk ≤ KkWk(cid:16)ZT ϕ2dµ(cid:17)1/2
for every ϕ ∈ A.
The following corollary is [BP, Lemma 2.1] emphasized for a.c. polyno-
mially bounded operators.
Corollary 2.1. Suppose that H is a Hilbert space, T ∈ L(H) is an a.c.
polynomially bounded operator, Cpol,T is the polynomial bound of T , and
x ∈ H. Then there exist ψ ∈ L2(T, m) such that RT ψ2dm ≤ 1 and
for every ϕ ∈ H ∞
kϕ(T )xk ≤ KCpol,Tkxk(cid:16)ZT ϕ2ψ2dm(cid:17)1/2
and W ∈ L(P 2(ψ2m),H) such that
W Sψ2m = T W,
and W ϕ = ϕ(T )x for every ϕ ∈ H ∞.
kWk ≤ KCpol,Tkxk
SHIFT-TYPE INVARIANT SUBSPACES
5
Proof. Define W ∈ L(A,H) by the formula
Wϕ = ϕ(T )x, ϕ ∈ A.
Let µ be the measure from Theorem A. Then µ = ψ2m + µs, where ψ ∈
L2(T, m) and µs is a positive Borel measure on T singular with respect to
m. We have
P 2(µ) = P 2(ψ2m) ⊕ L2(µs)
([C, Proposition III.12.3] or [N, Corollary A.2.2.1]), Sµs is a singular uni-
tary operator, and W has a continuous extension on P 2(µ) denoted by W .
Clearly, W Sµ = T W . Therefore, WL2(µs)Sµs = T WL2(µs). Since T is a.c.,
WL2(µs) = O ( [K3, Proposition 15] or [M]). Thus, µ can be replaced by
ψ2m.
Let ϕ ∈ H ∞. Set ϕr(ζ) = ϕ(rζ), ζ ∈ clos D, 0 < r < 1. Then ϕr ∈ A,
and ϕr → ϕ when r → 1 in the weak-∗ topology in H ∞ and in the norm
in L2(ψ2m) simultaneously. The conclusion of the corollary follows from
these convergences.
(cid:3)
Remark 2.2. For contractions T , Cpol,T = 1, and Corollary 2.1 is proved in
[BT, Lemma 3] with K = 1. The proof is based on the existence of isometric
dilations for contractions, see [SFBK, Theorem I.4.1].
Then
Remark 2.3. Let ψ ∈ L2(T). If RT log ψdm = −∞, then P 2(ψ2m) =
L2(ψ2m). If RT log ψdm > −∞, we accept that ψ is an outer function.
P 2(ψ2m) =
= khkH 2 , h ∈ H 2,
H 2
ψ
h
=n h
ψ
, h ∈ H 2o, (cid:13)(cid:13)(cid:13)
ψ(cid:13)(cid:13)(cid:13)P 2(ψ2m)
see [C, Ch. III.12, VII.10] or [N, Ch. A.4.1].
Lemma 2.4. Suppose that H is a Hilbert space, T ∈ L(H) is an a.c. polyno-
mially bounded operator, Cpol,T is the polynomial bound of T . Given N ∈ N
and {ϕn}N
n=1 ⊂ H ∞, define
A ∈ L(H,⊕N
n=1H), Ax = ⊕N
n=1ϕn(T )x, x ∈ H.
Then
kAk ≤ KCpol,T ess sup
ζ∈T (cid:16) NXn=1
ϕn(ζ)2(cid:17)1/2
.
6
M.F. GAMAL'
Proof. For x ∈ H, let ψx ∈ L2(T) be the function from Corollary 2.1 (applied
to T ). We have
kϕn(T )xk2
kAxk2 =
≤
NXn=1
NXn=1
K 2C 2
pol,Tkxk2ZT ϕn2ψx2dm
pol,Tkxk2ZT
NXn=1
ϕn2ψx2dm
ϕn2ZT ψx2dm
NXn=1
pol,Tkxk2 ess sup
ϕn2(cid:17) · kxk2.
NXn=1
pol,T ess sup
T
T
= K 2C 2
≤ K 2C 2
≤(cid:16)K 2C 2
(cid:3)
Theorem 2.5. Suppose that H is a Hilbert space, T ∈ L(H) is an a.c.
polynomially bounded operator, Cpol,T is the polynomial bound of T . Then
(cid:13)(cid:13)(cid:13)Xn∈N
ϕn(T )xn(cid:13)(cid:13)(cid:13) ≤ KCpol,T ess sup
ζ∈T (cid:16)Xn∈N
ϕn(ζ)2(cid:17)1/2(cid:16)Xn∈N
kxnk2(cid:17)1/2
for every {ϕn}n∈N ⊂ H ∞ and every {xn}n∈N ⊂ H such that the right part
of the above inequality is finite.
Proof. Let {ϕn}n∈N ⊂ H ∞ be such that
ess sup
ζ∈T Xn∈N
ϕn(ζ)2 := a < ∞.
that
Let N ∈ N. Let A be the transformation from Lemma 2.4 applied to T ∗ and
n=1, where eϕ(ζ) = ϕ(ζ) (ζ ∈ D, ϕ ∈ H ∞). Taking into account that
{fϕn}N
kAk = kA∗k and ϕ(T )∗ = eϕ(T ∗) ([K3, Proposition 24] or [M]), we obtain
ϕn(T )xn(cid:13)(cid:13)(cid:13) = kA∗(cid:0)⊕N
(cid:13)(cid:13)(cid:13)
NXn=1
≤ KCpol,T ess sup
n=1xn(cid:1)k
ζ∈T (cid:16) NXn=1
≤ KCpol,T a1/2(cid:16) NXn=1
kxnk2(cid:17)1/2
ϕn(ζ)2(cid:17)1/2(cid:16) NXn=1
kxnk2(cid:17)1/2
.
theorem follows.
Therefore, if Pn∈N kxnk2 < ∞, then Pn∈N ϕn(T )xn converges, and the
Lemma 2.6. Suppose that H is a Hilbert space, T ∈ L(H) is an a.c. poly-
nomially bounded operator, x ∈ H, ψk ∈ L2(T, m), k = 1, 2, are such that
(cid:3)
SHIFT-TYPE INVARIANT SUBSPACES
7
RT log ψkdm > −∞, and
kϕ(T )xk ≤(cid:16)ZT ϕ2ψk2dm(cid:17)1/2
for every ϕ ∈ H ∞,
k = 1, 2.
(2(1 + ε2))1/2
Set ψ(ζ) = min(ψ1(ζ),ψ2(ζ)), ζ ∈ T. Then for every 0 < ε < 1 there
exists an inner function ω ∈ H ∞ such that
for every ϕ ∈ H ∞.
kω(T )ϕ(T )xk ≤
Proof. We have T = τ1 ∪ τ2, where τ1 ∩ τ2 = ∅ and ψ = ψk a.e. on τk,
k = 1, 2. There exist outer functions ηk ∈ H ∞, k = 1, 2, such that
on τ1,
on τ2.
(cid:16)ZT ϕ2ψ2dm(cid:17)1/2
η2 =(ε ψ1
η1 =(1
on τ1,
on τ2,
1 − ε
ε ψ2
ψ1
and
ψ2
1
Clearly,
1 − ε ≤ η1 + η2 ≤ 1 + ε
a.e. on T.
Set ωη = η1 + η2, where ω, η ∈ H ∞, ω is inner, and η is outer. Let ϕ ∈ H ∞.
We have
2
2
2
2
ϕ
kω(T )ϕ(T )xk2 =(cid:13)(cid:13)(cid:13)ω(T )η(T )(cid:16) ϕ
=(cid:13)(cid:13)(cid:13)(η1 + η2)(T )(cid:16) ϕ
η(cid:17)(T )x(cid:13)(cid:13)(cid:13)
η(cid:17)(T )x(cid:13)(cid:13)(cid:13)
≤ 2(cid:16)(cid:13)(cid:13)(cid:13)η1(T )(cid:16) ϕ
η(cid:17)(T )x(cid:13)(cid:13)(cid:13)
η(cid:17)(T )x(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)η2(T )(cid:16) ϕ
2(cid:17)
ψ12dm +ZT η22(cid:12)(cid:12)(cid:12)
≤ 2(cid:16)ZT η12(cid:12)(cid:12)(cid:12)
η(cid:12)(cid:12)(cid:12)
η(cid:12)(cid:12)(cid:12)
ψ22dm(cid:17)
(1 − ε)2(cid:16)ZT η12ϕ2ψ12dm +ZT η22ϕ2ψ22dm(cid:17)
(1 − ε)2(cid:16)Zτ1 ϕ2ψ2dm + ε2Zτ2 ϕ2ψ2dm
+ ε2Zτ1 ϕ2ψ2dm +Zτ2 ϕ2ψ2dm(cid:17)
(1 − ε)2 ZT ϕ2ψ2dm.
2(1 + ε2)
≤
=
=
ϕ
2
2
2
(cid:3)
Remark 2.7. Applying Theorem 2.5, it is possible to prove an analog
of Lemma 2.6 for a finite family {ψk}N
k=1 with a constant not depended
on N , and for a countable family {ψk}k∈N under additional assumption
RT log ψdm > −∞, where ψ(ζ) = inf k∈N ψk(ζ), ζ ∈ T. We do not prove
these statements, because we will not apply them.
3. Preliminaries: function theory
Lemma 3.1. Suppose that ψ ∈ H 2 is an outer function. For t > 0 let ψt
be an outer function such that ψt = max(ψ, t). Then
(cid:13)(cid:13)(cid:13)
ψ
ψt − 1(cid:13)(cid:13)(cid:13)H 2→ 0 when t → 0.
8
M.F. GAMAL'
ψ(ζ) < t}. Since ψ 6= 0 a.e. on
H 2 ≥ m(T \ σt) → 1 when t → 0.
Proof. For 0 < t < 1 set σt = {ζ ∈ T :
T, m(σt) → 0 when t → 0. Therefore,
(3.1)
ψ
2
1 ≥(cid:13)(cid:13)(cid:13)
ψt(cid:13)(cid:13)(cid:13)
Furthermore,
0 ≤ (− log t)m(σt) ≤ −Zσt
because m(σt) → 0 andRT log ψdm > −∞. Therefore,
log ψdm → 0 when t → 0,
log ψ
t
log ψ
t
dm → 0 when t → 0.
Zσt
(0) = exp(cid:16)Zσt
dm(cid:17) → 1 when t → 0.
=(cid:13)(cid:13)(cid:13)
H 2 −(cid:12)(cid:12)(cid:12)
(0)(cid:13)(cid:13)(cid:13)
(0)(cid:12)(cid:12)(cid:12)
ψt(cid:13)(cid:13)(cid:13)
(0) − 1(cid:12)(cid:12)(cid:12) → 0 when t → 0. (cid:3)
+(cid:12)(cid:12)(cid:12)
(0)(cid:13)(cid:13)(cid:13)H 2
→ 0 when t → 0.
ψ
ψt −
ψ
ψt
ψ
ψt
ψ
ψt
ψ
ψt
H 2
ψ
2
2
2
Consequently,
(3.2)
ψ
ψt
By (3.1) and (3.2),
(cid:13)(cid:13)(cid:13)
(3.3)
ψ
ψt −
By (3.2) and (3.3),
ψ
ψt − 1(cid:13)(cid:13)(cid:13)H 2≤(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
Pn∈N ξ2
Proposition 3.2. Suppose that σ ⊂ T is a measurable set, {τn}n∈N is a
family of measurable sets such that σ = ∪n∈Nτn and τn ∩ τk = ∅, if n 6= k.
Suppose that C > 0 and {ξn}n∈N is a family of positive numbers such that
n < ∞. Suppose that {ηn}n∈N, {ϕn}n∈N, and {ψn}n∈N are families
of functions from L2(σ) such that
ηn ≤(1
ξn
on τn,
on σ \ τn,
and ϕn ≤ Cψn a.e. on σ. For every t > 0 and n ∈ N let ψnt ∈ L2(σ) be
such that ψn ≤ ψnt a.e. on σ and
Set
(cid:13)(cid:13)(cid:13)
ψn
ψnt − 1(cid:13)(cid:13)(cid:13)L2(σ) → 0 when t → 0
αt =Xn∈N
and α =Xn∈N
ϕn
ψnt
η2
n
η2
n
ϕn
ψn
.
for every n ∈ N.
ψnt − 1 ≤ 2 a.e. on σ. Therefore,
Then supt>0 kαtkL∞(σ) < ∞ and kαt − αkL2(σ) → 0 when t → 0.
Proof. Clearly, ψn
ψn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ηn2(cid:12)(cid:12)(cid:12)
αt − α ≤Xn∈N
n < ε and m(cid:16) [n≥N +1
Xn≥N +1
ψnt − 1(cid:12)(cid:12)(cid:12) ≤ 2C(cid:16)1 +Xk6=n
τn(cid:17) < ε.
Take ε > 0. There exists N ∈ N such that
k(cid:17) on τn.
ψn
ϕn
ξ2
ξ2
SHIFT-TYPE INVARIANT SUBSPACES
9
We have
ξ2
ε.
ξ2
ψk
ZSn≥N +1 τn αt − α2dm ≤ 4C 2(cid:16)1 +Xn∈N
ψnt − 1(cid:12)(cid:12)(cid:12) + Xk6=n,k≤N
ψnt − 1(cid:12)(cid:12)(cid:12)
k(cid:1) NXk=1(cid:12)(cid:12)(cid:12)
n(cid:17)2
ψkt − 1(cid:12)(cid:12)(cid:12) + 2 Xn≥N +1
k(cid:12)(cid:12)(cid:12)
ψkt − 1(cid:12)(cid:12)(cid:12)
+ Xk6=n,k≤N
ψkt − 1(cid:12)(cid:12)(cid:12)
+ 8C 2ε2 on
k(cid:12)(cid:12)(cid:12)
ψn
ψk
ψk
ξ4
ξ4
2
Therefore,
Let 1 ≤ n ≤ N . We have
ψn
αt − α ≤ C(cid:16)(cid:12)(cid:12)(cid:12)
αt − α2 ≤ C 2(cid:16)2N(cid:16)(cid:12)(cid:12)(cid:12)
≤ 2N C 2 max(cid:0)1, sup
Zσ(cid:12)(cid:12)(cid:12)
ψnt − 1(cid:12)(cid:12)(cid:12)
ψn
k∈N
There exists tε > 0 such that
2
dm ≤
ε
2N 2C 2 max(1, supk∈N ξ4
k)
for every 0 < t < tε and n = 1, . . . , N. Therefore,
2
ξ2
n(cid:17) on τn.
2(cid:17) + 8ε2(cid:17)
N[n=1
τn.
ZSN
n=1 τn αt − α2dm ≤ ε + 8C 2ε2
for every 0 < t < tε. (cid:3)
4. Preliminaries: operators intertwined with unitaries
Lemma 4.1. Suppose that T ∈ L(H) is an a.c. polynomially bounded op-
erator, Cpol,T is the polynomial bound of T , x ∈ H, and ψ ∈ L2(T) is
from Corollary 2.1. Furthermore, suppose that σ ⊂ T is a measurable set,
X ∈ L(H, L2(σ)), XT = UσX, and f = Xx. Then
f ≤ kXkKCpol,Tkxkψ
a.e. on σ.
Proof. We have
Zσ ϕ2f2dm = kϕ(Uσ)Xxk2 = kXϕ(T )xk2 ≤ kXk2kϕ(T )xk2
≤ kXk2K 2C 2
pol,Tkxk2ZT ϕ2ψ2dm
for every ϕ ∈ H ∞. It remains to apply standard reasons based on the fact
that for every measurable set τ ⊂ T and every a,b > 0 there exists ϕ ∈ H ∞
such that ϕ = a a.e. on τ and ϕ = b a.e. on T \ τ .
Lemma 4.2. Suppose that Lim is a Banach limit, H is a Hilbert space,
T ∈ L(H) is a power bounded operator, X is the canonical intertwining
mapping constructed using Lim. If M ∈ Lat T is such that TM is not of
class C0·, then kXMk ≥ 1.
Proof. We do not recall the construction of the canonical intertwining map-
ping from [K1] here. We recall only that kXxk2 = Limn kT nxk2 for every
x ∈ H. By assumption, there exists x ∈ M such that inf n≥0 kT nxk > 0.
Therefore, kXxk > 0. Since (T (a))nX = XT n and T (a) is unitary, we have
(cid:3)
kXxk = k(T (a))nXxk = kXT nxk ≤ kXMkkT nxk for every n ≥ 0.
10
M.F. GAMAL'
Consequently,
kXxk2 ≤ kXMk2 Lim
n kT nxk2 = kXMk2kXxk2.
(cid:3)
Since kXxk > 0, the conclusion of the lemma follows.
Lemma 4.3. Suppose that T ∈ L(H) is an a.c. polynomially bounded oper-
ator, Cpol,T is the polynomial bound of T , X∗ is the canonical intertwining
mapping for T ∗ constructed using a Banach limit. Let X be any transfor-
mation acting from H to some Hilbert space. For x ∈ H let ψ ∈ L2(T) be a
function from Corollary 2.1. IfRT log ψdm = −∞, then
Proof. Denote by H∗ the space on which (T ∗)(a) acts. Let W be from
Corollary 2.1. Since Sψ2m is unitary (see Remark 2.3), there exists B∗ ∈
L(H∗, P 2(ψ2m)) such that W ∗ = B∗X∗ and kB∗k ≤ kW ∗k (see [K1, The-
orem 2]). Set u = B∗
kXxk ≤ KCpol,TkXX∗
∗kkxk.
kuk ≤ kB∗
∗ 1, where 1(ζ) = 1 (ζ ∈ T). We have
∗kk1kP 2(ψ2m) ≤ kWk(ZT ψ2dm)1/2 ≤ KCpol,Tkxk,
∗ B∗
∗ 1 = XX∗
∗ u,
Xx = XW 1 = XX∗
∗kKCpol,Tkxk.
becauseRT ψ2dm ≤ 1. Taking into account that
we conclude that kXxk ≤ kXX∗
∗kkuk ≤ kXX∗
Proposition 4.4. Suppose that T ∈ L(H) is an a.c. polynomially bounded
operator, σ ⊂ T is a measurable set, X ∈ L(H, L2(σ)), XT = UσX, x ∈ H
is such that ∨n≥0T nx = H, f = Xx, ψ0 ∈ L2(T), andRT ψ02dm ≤ 1.
Furthermore, suppose that 0 < ε, δ < 1, τ ⊂ σ is a measurable set,
0 6= u ∈ H is such that Xu ∈ L2(τ ) and kXuk ≥ δkuk.
0(cid:17)1/2
(we assume that ε is so small that γ0,δ0 > 0). Let 0 < γ < γ0. By assump-
tion on x, there exists ϕ0 ∈ H ∞ such that kϕ0(T )x − uk ≤ ε. Then
γ0 =(cid:16)1 −
(kuk − ε)2δ2
δkuk − εkXk
ε2kXk2
kuk + ε
δ0 =
and
m(τ ) > 0,
Set
(cid:3)
τ 6⊂ {ζ ∈ σ :
ϕ0(ζ)f (ζ) ≤ γδ0kϕ0(T )xkψ0(ζ)}.
Proof. We have ϕ0f = Xϕ0(T )x, and
(ϕ0f )σ\τ = PL2(σ\τ )Xϕ0(T )x = PL2(σ\τ )X(ϕ0(T )x − u).
Therefore,
Also,
Thus,
Zσ\τ ϕ0f2dm ≤ kXk2kϕ0(T )x − uk2 ≤ ε2kXk2.
kXϕ0(T )xk ≥ kXuk − εkXk ≥ δkuk − εkXk
and δ0kϕ0(T )xk ≤ (kuk + ε)δ0.
kXϕ0(T )xk ≥ δ0kϕ0(T )xk.
δ2
0kϕ0(T )xk2 ≤ kXϕ0(T )xk2 =Zσ ϕ0f2dm
=Zσ\τ ϕ0f2dm +Zτ ϕ0f2dm
≤ ε2kXk2 +Zτ
≤ ε2kXk2 + γ2δ2
(becauseRτ ψ02dm ≤RT ψ02dm ≤ 1). Thus,
ε2kXk2
0kϕ0(T )xk2 ≤
ε2kXk2
0(kuk − ε)2 ,
0kϕ0(T )xk2
1 − γ2 ≤
δ2
δ2
γ2δ2
0kϕ0(T )xk2ψ02dm
SHIFT-TYPE INVARIANT SUBSPACES
11
Let us assume that
τ ⊂ {ζ ∈ σ :
ϕ0(ζ)f (ζ) ≤ γδ0kϕ0(T )xkψ0(ζ)}.
Then
(cid:3)
a contradiction with assumption on γ.
Proposition 4.5. Suppose that T ∈ L(H) is an a.c. polynomially bounded
operator, σ ⊂ T is a measurable set, T (a) ∼= Uσ, and X is the canonical in-
tertwining mapping for T and Uσ constructed using a Banach limit. Suppose
that V ∈ L(K) is a unitary operator, Y ∈ L(K,H) is such that Y V = T Y
and clos XY K = L2(σ).
Furthermore, suppose that x ∈ H and {ϕn}n∈N ⊂ H ∞ are such that
{ϕn(T )x}n∈N is dense in H. Set f = Xx. Suppose that {ψn}n∈N ⊂ L2(T)
are such that RT ψn2dm ≤ 1 for all n ∈ N. For n ∈ N and 0 < c < 1 set
ϕn(ζ)f (ζ) ≤ ckϕn(T )xkψn(ζ)}.
σn,c = {ζ ∈ σ :
For 0 < ε0, δ0, γ0 < 1 set
Nε0,δ0 = {n ∈ N : kXϕn(T )xk ≥ δ0kϕn(T )xk
and 1 − ε0 ≤ kϕn(T )xk ≤ 1 + ε0}
and
τε0,δ0,γ0 = \n∈Nε0,δ0
σn,γ0δ0.
Finally, let 0 < ε00, δ0, γ0 < 1 be given. Then there exists 0 < ε0 ≤ ε00
such that
m(τε0,δ0,γ0) = 0.
PK⊖E V K⊖E ≺ Uσ.
Proof. Set E = ker XY . Then E ∈ Lat V and XY K⊖E realizes the relation
(4.1)
Therefore, V ∗K⊖E is of class C·1. Consequently, V ∗K⊖E is unitary. Thus,
K ⊖ E ∈ Lat V , and V K⊖E ∼= Uσ by (4.1) and [SFBK, Proposition II.3.4].
Without loss of generality, K ⊖ E = L2(σ) and V K⊖E = Uσ. We have
ker XY L2(σ) = {0}. Therefore, there exists g ∈ L∞(σ) such that
a.e. on σ and XY h = gh for every h ∈ L2(σ).
(4.2)
g 6= 0
12
M.F. GAMAL'
For 0 < ε0, δ < 1 set
δ1 =
δ − ε0kXk
1 + ε0
and γ1 =(cid:16)1 −
0kXk2
ε2
(1 − ε0)2δ2
1(cid:17)1/2
.
Then γ1 → 1 and δ1 → 1 when δ → 1 and ε0 → 0. Choose 0 < ε0, δ < 1
such that 0 < ε0 ≤ ε00, γ0 < γ1 and δ0 ≤ δ1. We claim that m(τε0,δ0,γ0) = 0.
Set τ = τε0,δ0,γ0. Let us assume that m(τ ) > 0. Set M = clos Y L2(τ ).
By (4.2), XM 6≡ O. By Lemma 4.2, kXMk ≥ 1. Therefore, there exists
u ∈ M such that kuk = 1 and kXuk ≥ δ. We have
Xu ∈ clos XM = clos XY L2(τ ) = L2(τ ).
Since {ϕn(T )x}n∈N is dense in H, there exists n ∈ N such that
kϕn(T )x − uk ≤ ε0.
Therefore, n ∈ Nε0,δ0. By Proposition 4.4 applied to x and ϕn with ε0 and
δ, taking into account that γ0 < γ1, we obtain τ 6⊂ σn,γ0δ1. By the definition
of τ , we have τ ⊂ σn,γ0δ0 ⊂ σn,γ0δ1, a contradiction.
(cid:3)
The following lemma is a version of [K1, Theorem 3].
Lemma 4.6. Suppose that Lim is a Banach limit, T is a power bounded
operator, M ∈ Lat T , X∗ and Z are canonical intertwining mappings for T ∗
and (TM)∗ from their unitary asymptotes, respectively, constructed using
Lim. Set
n kPM(T ∗)nxk2 ≤ Lim
n k(T ∗)nxk2 =kX∗xk2
Following the proof of [K1, Theorem 3(b)] (applied to T ∗) we obtain a
L = _n≥0(cid:0)(T ∗)(a)(cid:1)−nX∗M⊥.
Then there exists a transformation B such that
kBk ≤
Z = BPL⊥X∗M and
√2kX∗k.
Proof. By the constructions of X∗ and Z,
kZxk2 = Lim
for every x ∈ M. Therefore, kZk ≤ kX∗k.
transformation(cid:18)B1 B2
(cid:18)B1 B2
O B(cid:19)(cid:18)X∗M⊥
n k(cid:0)(TM)∗(cid:1)nxk2 = Lim
O B(cid:19) such that
PL⊥X∗M(cid:19) =(cid:18)X∗M⊥ PLX∗M
(cid:19)
and(cid:13)(cid:13)(cid:13)(cid:18)B1 B2
O B(cid:19)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:18)X∗M⊥ PLX∗M
(cid:19)(cid:13)(cid:13)(cid:13) ≤
kBk ≤(cid:13)(cid:13)(cid:13)(cid:18)B1 B2
O B(cid:19)(cid:13)(cid:13)(cid:13) ≤
√2kX∗k. (cid:3)
Proposition 4.7. Set χ(ζ) = ζ and 1(ζ) = 1 (ζ ∈ T). Set
UTH 2
Therefore, BPL⊥X∗M = Z and
H 2
− = L2(T) ⊖ H 2
and S∗ = PH 2
PLX∗M
Z
.
−
Z
O
O
O
−
Suppose that N and K are Hilbert spaces, Lim is a Banach limit,
T0 ∈ L(N ) is power bounded,
√2kX∗k.
SHIFT-TYPE INVARIANT SUBSPACES
13
V0 ∈ L(K), Z0 ∈ L(N ,K), (Z0, V0) is the unitary asymptote of T ∗
structed using Lim, Y0 ∈ L(H 2,N ), and Y0S = T0Y0. Set
Z0 O
Y ∗
O
0
O IH 2
S∗ (cid:19) and X∗ =
T =(cid:18)T0
.
(·, χ)Y01
O
−
Then
0 con-
(Y0 ⊕ IH 2
(4.3)
T) is the unitary asymptote of T ∗ con-
T is power bounded, and (X∗, V0 ⊕ U ∗
structed using Lim. Furthermore, if T0 is polynomially bounded, then T is
polynomially bounded, and if T0 is a.c., then T is a.c..
)UT = T (Y0 ⊕ IH 2
),
−
−
Moreover, let (X0, UT) be the unitary asymptote of T0 constructed using
some Banach limit. Since (X0Y0)S = UT(X0Y0), there exists g1 ∈ L∞(T)
such that X0Y0h = g1h for every h ∈ H 2. Let
be defined by the formula
X ∈ L(N ⊕ H 2
−, L2(T))
X(x ⊕ h) = X0x + g1h
(x ∈ N , h ∈ H 2
−).
Then (X, UT) is the unitary asymptote of T constructed using the same
Banach limit as X0.
Proof. The equality (4.3) easy follows from the definition of T . Let p be an
(analytic) polynomial. It easy follows from (4.3) that
= Y0PH 2p(UT)H 2
PN p(T )H 2
(4.4)
.
−
−
The conclusions on power boundedness and polynomial boundedness of T
follow from (4.4). Since the right part of (4.4) is defined for every p ∈ H ∞,
the conclusion on a.c. polynomial boundedness of T follows from (4.4) and
[K3, Theorem 23].
Let x1, x2 ∈ N , let h1, h2 ∈ H 2
((T ∗)n(x1 ⊕ h1), (T ∗)n(x2 ⊕ h2))
χnY ∗
= (h1, h2) + (PH 2
−
−, and let n ∈ N. It follows from (4.4) that
0 x1, PH 2
−
χnY ∗
0 x2) + ((T ∗
0 )nx1, (T ∗
0 )nx2).
We have
lim
n
(PH 2
−
χnY ∗
0 x1, PH 2
−
χnY ∗
0 x2) = (Y ∗
0 x1, Y ∗
0 x2).
By the construction of Z0,
(Z0x1, Z0x2) = Lim
n
and by the construction of X∗,
((T ∗
0 )nx1, (T ∗
0 )nx2),
(X∗(x1 ⊕ h1), X∗(x2 ⊕ h2)) = Lim
n
((T ∗)n(x1 ⊕ h1), (T ∗)n(x2 ⊕ h2)).
Therefore,
(X∗(x1 ⊕ h1), X∗(x2 ⊕ h2)) = (h1, h2) + (Y ∗
0 x1, Y ∗
0 x2) + (Z0x1, Z0x2).
From the latest equality and [K1, Theorem 3] we obtain the conclusion on
the unitary asymptote of T ∗.
Suppose that (X0, UT) is the unitary asymptote of T0 constructed using
some Banach limit, and X is defined as above. Clearly, XT = UTX. Denote
14
M.F. GAMAL'
by XT the canonical intertwining mapping for T and T (a) constructed using
the same Banach limit as X0. By [K1, Theorem 3], T (a) ∼= (T0)(a) and
X0 = XTN . Therefore, T (a) ∼= UT, and it is easy to see from the definitions
of T and X and the intertwining properties of X and XT that XT = X. (cid:3)
5. Proof of Theorem 1.1
For convenience of applying Proposition 4.7, we denote the operator from
Theorem 1.1 by T0.
Suppose that N is a Hilbert space, T0 ∈ L(N ) is a power bounded oper-
ator of class C1·, and (T0)(a) ∼= UT. Let
∗
∗
O T·0(cid:19)
O C·0(cid:19), see Introduction and [K1].
T0 =(cid:18)T11
be the triangulation of T0 of the form(cid:18)C·1
Since T0 is of class C1·, we have T11 is of class C11. By [K1, Theorem 3] and
0 )(a) ∼= ((T11)(a))∗,
[SFBK, Propositions II.3.4 and II.5.3], T11 ∼ (T11)(a), (T ∗
and there exists a measurable set τ ⊂ T such that (T11)(a) ∼= Uτ .
First case: m(τ ) = 1. Then T11 ∼ UT.
Suppose that T0 is polynomially bounded. We can consider T11 instead of
T0. The proof becomes simpler. Proposition 4.7 is not used, because we
need not to represent T11 as a restriction of some operator of class C·1 on
its invariant subspace N . Lemma 2.6 is not used, because we need not to
request that the needed shift-type invariant subspace is contained in the
given subspace N . The detailed proof can be deduced from the proof of
second case and is not given here.
Two cases are possible.
Second case: m(τ ) < 1.
Suppose that T0 is a cyclic a.c. polynomially bounded operator. By
Corollary 2.1, there exists a quasiaffinity Y0 ∈ L(H 2,N ) such that Y0S =
T0Y0. Denote by X0 the canonical intertwining mapping for T0 and UT
constructed using a Banach limit. We have
(5.1)
X0T0 = UTX0
and _n≥0
U −n
T X0N = L2(T).
−. Let T ∈ L(H), X ∈ L(H, L2(T)), g1 ∈ L∞(T) be from
Set H = N ⊕ H 2
Proposition 4.7 applied to T0 and Y0. Since Y0 is a quasiaffinity, it follows
from the second equality in (5.1) that g1 6= 0 a.e. on T, and it follows from
(4.3) that T is of class C·1.
Let X∗ ∈ L(H, L2(τ ) ⊕ L2(T)) be the canonical intertwining mapping for
T ∗ constructed in Proposition 4.7 with V0 = U ∗
X∗
∗ (L2(τ ) ⊕ H 2) ⊂ N .
(5.2)
Furthermore, there exists g2 ∈ L∞(τ ) such that
τ . Then
(XX∗
∗ )(h2 ⊕ h1) = g2h2 + g1h1
for every h2 ∈ L2(τ ) and h1 ∈ L2(T).
Set
θ1 =
g1
(g12 + g22)1/2
and θ2 =
g2
(g12 + g22)1/2
,
SHIFT-TYPE INVARIANT SUBSPACES
15
and set
Clearly, E0 = ker XX∗
E0 = {θ1h ⊕ (−θ2h) : h ∈ L2(τ )}.
∗ and E0 is a reducing subspace for Uτ ⊕ UT, that is,
E0 ∈ Lat(Uτ ⊕ UT) and (L2(τ ) ⊕ L2(T)) ⊖ E0 ∈ Lat(Uτ ⊕ UT).
Define
by the formula
Set
(5.3)
J ∈ L(cid:0)L2(T), (L2(τ ) ⊕ L2(T)) ⊖ E0(cid:1)
(h ∈ L2(T)).
Jh = θ2hτ ⊕ (θ1hτ + hT\τ )
g =((g12 + g22)1/2
g1
on τ,
on T \ τ.
∗ Jh = gh for every h ∈ L2(T) and L2(τ ) ⊕ L2(T) = JL2(T) ⊕ E0.
Then g 6= 0 a.e. on T, J is unitary, JUT = (Uτ ⊕ UT)(L2(τ )⊕L2(T))⊖E0 J,
(5.4)
XX∗
Take σ ⊂ T such that
(5.5) KCpol,T√2kX∗k ess sup
Let h0 ∈ H 2 be such that
T\σ g > 0 and m(τ ∪ σ) < 1.
g < 1, ess inf
σ
(5.6)
There exists θ0 ∈ L∞(τ ) such that θ0 = 1 a.e. on τ and
(5.7)
h0 ≍ 1 a.e. on T.
Re θ1θ2θ0 = 0 a.e. on τ.
Set
(5.8)
h01 = (θ2θ0 + θ1)h0χτ + h0χT\τ
and
h02 = (−θ2 + θ1θ0)h0τ .
h01 = h0 a.e. on T.
Fτ,σ = θ2h01τ ∩σ + θ1h02,
FT,σ = θ1h01τ ∩σ + h01σ\τ − θ2h02,
Fτ,T\σ = θ2h01τ \σ,
FT,T\σ = θ1h01τ \σ + h01(T\τ )\σ.
By (5.7),
(5.9)
Set
(5.10)
and
(5.13)
Put
(5.14)
Then
(5.11) Fτ,σ ⊕ FT,σ = Jh01σ + θ1h02 ⊕ (−θ2h02),
(5.12)
Fτ,T\σ ⊕ FT,T\σ = Jh01T\σ,
θ1h02 ⊕ (−θ2h02) ∈ E0,
Fτ,σ ⊕ FT,σ + Fτ,T\σ ⊕ FT,T\σ = θ0h0τ ⊕ h0 ∈ L2(τ ) ⊕ H 2.
x0 = X∗
∗ (Fτ,σ ⊕ FT,σ).
16
M.F. GAMAL'
It follows from (5.11) and (5.4) that
Xx0 = gh01σ.
M0 = _k≥0
T kx0
and K0 = _k≥0
(Uτ ⊕ UT)k(Fτ,σ ⊕ FT,σ).
By (5.10), Fτ,σ ⊕ FT,σ ∈ L2(τ ) ⊕ L2(τ ∪ σ). It follows from third inequality
in (5.5) that (Uτ ⊕ UT)K0 is unitary. It follows from (5.6), (5.9), (5.15), and
the equality XT = UTX that
By [K1, Theorem 3], (XM0 , Uσ) is the unitary asymptote of TM0 con-
structed using the same Banach limit as X.
clos XM0 = L2(σ).
(5.15)
Set
Set
(5.16)
L = _n≥0
(U ∗
τ ⊕ U ∗
T)−nX∗M⊥
0 .
Denote by Z the canonical intertwining mapping for (TM0)∗ constructed
using the same Banach limit as X∗. Let B be the transformation from
Lemma 4.6 applied to T and M0. We have
XM0 Z ∗ = XPM0X∗
Therefore,
kXM0 Z ∗k ≤
It follows from (5.16) that X∗
∗ PL⊥B∗.
√2kX∗kkXPM0 X∗
∗L⊥ ⊂ M0. Therefore,
and XX∗
∗L⊥k.
XPM0X∗
(5.17)
Let u ∈ L⊥. By (5.4), there exist h ∈ L2(T) and v ∈ E0 such that u = Jh+v.
By (5.4) and (5.17),
∗L⊥ ⊂ L2(σ).
∗L⊥ = XX∗
∗L⊥
XX∗
Therefore, hT\σ = 0 and
∗ uk ≤ ess sup
= ess sup
kXX∗
σ
∗ u = XX∗
∗ Jh = gh ∈ L2(σ).
gkhkL2(σ) ≤ ess sup
gkuk
σ
σ
g(khk2
L2(T) + kvk2)1/2
(because J is unitary). We obtain that
Consequently,
By (5.5),
Take
kXM0 Z ∗k ≤
σ
kXX∗
∗L⊥k ≤ ess sup
g.
√2kX∗k ess sup
KCpol,TkXM0 Z ∗k < 1.
σ
g.
KCpol,TkXM0 Z ∗k < δ0 < 1 and 0 < ε00, γ0 < 1.
Let {ϕn}n∈N ⊂ H ∞ be such that {ϕn(T )x0}n∈N is dense in M0. Let ψn be
constructed by Corollary 2.1 applied to ϕn(T )x0 for every n ∈ N. Set
(5.18)
f = Xx0.
SHIFT-TYPE INVARIANT SUBSPACES
17
It is easy to see that TM0, XM0, (Uτ ⊕ UT)K0, X∗
Proposition 4.5. Let 0 < ε0 ≤ ε00 be such that
m(τε0,δ0,γ0) = 0,
∗K0, x0, {ψn}n∈N satisfy
where τε0,δ0,γ0 is from Proposition 4.5.
For convenience, relabel {ϕn}n∈Nε0,δ0
as {ϕn}n∈N. Then
kXϕn(T )x0k ≥ δ0kϕn(T )x0k
for every n ∈ N.
By Lemma 4.3 applied to TM0 and the choice of δ0 we have
(5.19)
log ψndm > −∞ for every n ∈ N.
ZT
We accept that ψn are outer functions (and ψn ∈ H 2), see Remark 2.3. Set
σn = σn,γ0δ0, n ∈ N.
By (5.18) and Lemma 4.1 applied to ϕn(T )x0,
(5.20)
ϕnf ≤ kXkKCpol,Tkϕn(T )x0kψn
≤ (1 + ε0)kXkKCpol,Tψn a.e. on σ for every n ∈ N.
By Proposition 4.5,
ϕnf > γ0δ0kϕn(T )xkψn ≥ (1 − ε0)γ0δ0ψn a.e. on σ \ σn
for every n ∈ N and
m(\n∈N
σn) = 0.
Set
Then
and
(5.21)
τ1 = σ \ σ1, τn = (σ \ σn) \ ∪n−1
[n∈N
τn ∩ τk = ∅,
τn = σ,
k=1τk, n ≥ 2.
if n 6= k, n, k ≥ 1,
ϕnf > (1 − ε0)γ0δ0ψn a.e. on τn, n ∈ N.
By (5.14),
ϕ(T )ϕn(T )x0 = ϕ(T )ϕn(T )X∗
∗ (Fτ,σ ⊕ FT,σ) = X∗
∗ (ϕϕnFτ,σ ⊕ ϕϕnFT,σ).
Therefore,
kϕ(T )ϕn(T )x0k2 ≤ kX∗
= kX∗
= kX∗
≤ kX∗
Easy calculation show that
∗k2(kϕϕnFτ,σk2 + kϕϕnFT,σk2)
∗k2(cid:16)Zτ ϕϕnFτ,σ2dm +ZT ϕϕnFT,σ2dm(cid:17)
∗k2ZT ϕϕn2(χτFτ,σ2 + FT,σ2)dm
(χτFτ,σ2 + FT,σ2)ZT ϕϕn2dm.
∗k2 ess sup
(τ \σ)∪(σ\τ ) h02).
h02, ess sup
τ ∩σ
T
ess sup
T
(χτFτ,σ2 + FT,σ2) = max(2 ess sup
18
Set
Then
M.F. GAMAL'
C1 = kX∗
kϕ(T )ϕn(T )x0k ≤ KCpol,T (1 + ε0)(cid:16)ZT ϕ2C 2
∗k(cid:0)max(2 ess supτ ∩σ h02, ess sup(τ \σ)∪(σ\τ ) h02)(cid:1)1/2
1ϕn2dm(cid:17)1/2
KCpol,T (1 + ε0)
.
Take 0 < c1 < 1. By Lemma 2.6, there exist inner functions ωn ∈ H ∞,
for every ϕ ∈ H ∞ and every n ∈ N.
n ∈ N, such that
kϕ(T )ωn(T )ϕn(T )x0k
≤ KCpol,T (1 + ε0)
(2(1 + c2
1 − c1
for every ϕ ∈ H ∞ and every n ∈ N.
ψ1n =(ψn
(5.22)
There exist outer functions ψ1n ∈ H 2 such that
min(ψn, C1ϕn)
on σ,
on T \ σ.
1))1/2
(cid:16)ZT ϕ2 min(ψn, C1ϕn)2dm(cid:17)1/2
Since min(ψn, C1ϕn) ≤ ψ1n a.e. on T, we have
(5.23)
(2(1 + c2
1 − c1
kϕ(T )ωn(T )ϕn(T )x0k ≤KCpol,T (1 + ε0)
1))1/2
(cid:16)ZT ϕ2ψ1n2dm(cid:17)1/2
for every ϕ ∈ H ∞ and every n ∈ N.
It follows from (5.22) that
Let {ξn}n∈N be a sequence of positive numbers such that
(cid:12)(cid:12)(cid:12)
ϕn
1
C1
ψ1n(cid:12)(cid:12)(cid:12) ≥
ξ :=Xn∈N
on T \ σ for every n ∈ N.
ξ2
n < ∞, Xn≥2
ξ2
n < 1
(5.24)
(5.25)
and
(5.26)
(1 + ε0)kXkKCpol,T ξ < (1 − ε0)γ0δ0.
For n ∈ N, define outer functions ηn ∈ H ∞ as follows:
(5.27)
1/2
1
ξ1
on T \ σ,
on τ1,
on σ \ τ1,
ηn =
η1 =
ϕn(cid:12)(cid:12)(cid:12)
ϕ1(cid:12)(cid:12)(cid:12)
ξn(cid:12)(cid:12)(cid:12) ψ1n
(cid:12)(cid:12)(cid:12) ψ11
Xn∈Nηn(ζ)2 ≤((1 +Pn≥2 ξ2
n)C1
1
ξn
1 + ξ
By (5.24) and (5.25),
for a.e. ζ ∈ T \ σ,
for a.e. ζ ∈ σ.
1/2
on T \ σ,
on τn,
on σ \ τn,
n ≥ 2.
For every t > 0 and every n ∈ N let ψnt be the outer function such that
(5.28)
ψnt = max(ψ1n, t)
a.e. on T.
SHIFT-TYPE INVARIANT SUBSPACES
19
Put
(5.29)
κnt(ζ) =
ηn(ζ)
ψnt(ζ)
,
ζ ∈ T, n ∈ N, t > 0.
Since ψnt ≥ t a.e. on T, we have κnt ∈ H ∞ for every n ∈ N and t > 0. Set
xnt = (κntωnϕn)(T )x0 = (κntωnϕn)(T )X∗
∗ (κntωnϕnFτ,T\σ ⊕ κntωnϕnFT,T\σ)
and ynt = X∗
∗ (Fτ,σ ⊕ FT,σ)
= (κntωnϕn)(T )X∗
∗ (Fτ,T\σ ⊕ FT,T\σ)
for every n ∈ N and t > 0.
(5.30)
We have
xnt + ynt = (κntωnϕn)(T )X∗
∗ (Fτ,σ ⊕ FT,σ + Fτ,T\σ ⊕ FT,T\σ).
By (5.13) and (5.2),
(5.31)
By (5.23) and (5.28),
xnt + ynt ∈ N .
(5.32)Xn∈N
kϕ(T )xntk2 =Xn∈N
and by (5.29),
(5.33)
(cid:17)2Xn∈NZT ϕκnt2ψnt2dm,
1))1/2
k(ϕκntωnϕn)(T )x0k2
(2(1 + c2
1 − c1
≤(cid:16)KCpol,T (1 + ε0)
Xn∈NZT ϕκnt2ψnt2dm =Xn∈NZT ϕ2ηn2dm
=ZTXn∈N
T Xn∈N
ηn2ϕ2dm ≤ ess sup
ηn2ZT ϕ2dm.
By Theorem 2.5, (5.32) and (5.33),
(5.34)
kXn∈N
(ηnϕ)(T )xntk
≤ K 2C 2
pol,T (1 + ε0)
1))1/2
(2(1 + c2
1 − c1
ess sup
T Xn∈N
ηn2(cid:16)ZT ϕ2dm(cid:17)1/2
for every t > 0.
For t > 0 put
(5.35)
βt =Xn∈N
η2
n
ωnϕn
ψnt
on T \ σ.
It follows from (5.27) and (5.28) that the series (5.35) converges in k·kL∞(T\σ)
and
(5.36)
kβtkL∞(T\σ) ≤ 1 +Xn≥2
ξ2
n
for all t > 0.
20
M.F. GAMAL'
For t > 0 and ϕ ∈ H ∞ let
Wtϕ =Xn∈N
(5.37)
(ηnϕ)(T )xnt + X∗
∗ (βtϕFτ,T\σ ⊕ βtϕFT,T\σ).
It follows from the convergence in k · kL∞(T\σ) of the series (5.35) that
(5.38)
X∗
∗ (βtϕFτ,T\σ⊕βtϕFT,T\σ)
X∗
ωnϕn
ψnt
n
∗(cid:16)η2
=Xn∈N
ϕFτ,T\σ ⊕ η2
n
ωnϕn
ψnt
ϕFτ,T\σ ⊕ η2
ϕFT,T\σ(cid:17).
ϕFT,T\σ(cid:17) = (ϕηn)(T )ynt.
∗ and (5.30) that
It follows from the intertwining property of X∗
It follows from (5.37), (5.38), (5.39), and (5.31) that
X∗
n
ωnϕn
ψnt
ωnϕn
ψnt
∗(cid:16)η2
Wtϕ ∈ N for every ϕ ∈ H ∞ and t > 0.
n
By (5.36) and (5.10),
By (5.9) and (5.10),
T\σ (cid:0)χτFτ,T\σ2 + FT,T\σ2(cid:1)ZT\σ ϕ2dm.
∗ (βtϕFτ,T\σ ⊕ βtϕFT,T\σ)k2
n(cid:1)2 ess sup
≤ kX∗
ξ2
ess sup
∗k2(cid:0)1 +Xn≥2
T\σ (cid:0)χτFτ,T\σ2 + FT,T\σ2(cid:1) = ess sup
T Xn∈N
(2(1 + c2
1 − c1
pol,T (1 + ε0)
1))1/2
ess sup
T\σ
C2 =K 2C 2
h02.
ηn2
+ kX∗
∗k(cid:0)1 +Xn≥2
ξ2
n(cid:1) ess sup
T\σ
h0.
It follows from (5.34) and (5.40) that the mapping Wt defined on H ∞ by
(5.37) is extended onto H 2 for every t > 0. Denoting the extension by the
same letter, we obtain
Wt ∈ L(H 2,N ) such that WtS = TN Wt
and kWtk ≤ C2 for every t > 0.
(5.39)
(5.40)
kX∗
Set
(5.41)
Put
(5.42)
αt =Xn∈N
η2
n
ωnϕngh01
ψnt
ωnϕngh01
η2
n
and α =Xn∈N
for every t > 0 and h ∈ H 2.
ψ1n
on T.
By (5.37), (5.30), (5.29), (5.11), (5.12), (5.4), (5.35),
(5.43)
XWth = αth
By (5.27), (5.20), (5.22), (5.18), (5.15), (5.28) and Lemma 3.1, αt and α
satisfy Proposition 3.2. By Proposition 3.2,
(5.44)
t>0 kαtkL∞(σ) < ∞ and kαt − αkL2(σ) → 0 when t → 0.
sup
SHIFT-TYPE INVARIANT SUBSPACES
21
An analog of (5.44) for T \ σ follows from (5.28), (5.27), (5.5), (5.6), (5.9)
and Lemma 3.1. Therefore,
(5.45)
t>0 kαtkL∞(T) < ∞ and kαt − αkL2(T) → 0 when t → 0.
sup
It follows from (5.41) that there exist a sequence {tj}j and W ∈ L(H 2,N )
such that tj →j 0 and Wtj →j W in the weak operator topology, W S =
TN W and kWk ≤ C2. It follows from (5.43) and (5.45) that
(5.46)
XW h = αh
for every h ∈ H 2.
We will show that
(5.47)
We have σ = ∪n∈Nτn. Fix n ∈ N. By (5.20), (5.22), (5.25), (5.27), taking
into account that f = gh01 a.e. on σ by (5.18) and (5.15), we obtain
1/α ∈ L∞(T).
ηk2ϕkf
ψk ≤ (1 + ε0)kXkKCpol,T Xk∈N,k6=n
ξ2
k
≤ (1 + ε0)kXkKCpol,T ξ
a.e. on τn.
From the latest estimate, (5.27) and (5.21) we conclude that
η2
k
(cid:12)(cid:12)(cid:12) Xk∈N,k6=n
ωkϕkf
ψ1k
(cid:12)(cid:12)(cid:12) ≤ Xk∈N,k6=n
ψn −(cid:12)(cid:12)(cid:12) Xk∈N,k6=n
α ≥ ϕnf
η2
k
ϕkf
ψk (cid:12)(cid:12)(cid:12)
≥ (1 − ε0)γ0δ0 − (1 + ε0)kXkKCpol,T ξ a.e. on τn.
It follows from (5.26) that
(5.48)
It follows from (5.27) that
1/α ∈ L∞(σ).
(cid:12)(cid:12)(cid:12)Xn∈N
η2
n
ωnϕn
ψ1n (cid:12)(cid:12)(cid:12) ≥(cid:12)(cid:12)(cid:12)η2
1
ω1ϕ1
ψ11 (cid:12)(cid:12)(cid:12) −Xn≥2(cid:12)(cid:12)(cid:12)η2
n
ωnϕn
ψ1n (cid:12)(cid:12)(cid:12) = 1 −Xn≥2
ξ2
n
on T \ σ.
It follows from (5.25), (5.6), (5.9) and the second inequality in (5.5) that
(5.49)
1/α ∈ L∞(T \ σ).
Now (5.47) follows from (5.48) and (5.49).
Define A ∈ L(L2(T)) by the formula Ah = h/α, h ∈ L2(T). Set
Z = AX.
It follows from (5.46) that
ZW h = h
for every h ∈ H 2.
Set
M = W H 2.
Then M ⊂ N and TM ≈ S. The conclusion of Theorem 1.1 for T0 follows
from the relation T0 = TN .
22
M.F. GAMAL'
6. Proof of Corollary 1.2
Lemma 6.1. Suppose that T ∈ L(H) is an a.c. polynomially bounded op-
erator, X ∈ L(H, H 2), Y ∈ L(H 2,H), XT = SX, Y S = T Y ,
clos XH = H 2
and
clos Y H 2 = H.
Then ker X = {0} and ker Y = {0}.
Proof. Since XY S = SXY and clos XY H 2 = H 2, there exists an outer
function g ∈ H ∞ such that XY = g(S). Therefore, ker Y = {0}. Since
Y XY = Y g(S) = g(T )Y and clos Y H 2 = H, we obtain that Y X = g(T ).
By [M], ker g(T ) = {0}. Therefore, ker X = {0}.
Lemma 6.2. Suppose that T is a cyclic a.c. polynomially bounded operator
(cid:3)
such that T
d
≺ S. Then T ∼ S.
d
Proof. Since T is cyclic, it follows from Corollary 2.1 that S
6.1, T ∼ S.
Lemma 6.3. Suppose that T is a cyclic a.c. polynomially bounded operator
≺ T . By Lemma
(cid:3)
d
≺ UT. Then T (a) ∼= UT.
such that T
Proof. Since T is an a.c. polynomially bounded operator, T (a) is a.c. by
[K3, Proposition 15]. By [K1, Theorem 2], T (a) contains UT as an orthogonal
summand. Since T (a) is a.c. and cyclic, we conclude that T (a) ∼= UT.
(cid:3)
Lemma 6.4. Suppose that T ∈ L(H) is an a.c. polynomially bounded op-
erator,
T =(cid:18)T0
∗O T1(cid:19)
with respect to some decomposition of H, and there exists M1 ∈ Lat T1 such
that T1M1 ∼ S. Then there exists M ∈ Lat T such that TM ∼ S.
Proof. Let x be a cyclic vector for T1M1. Set M = ∨n≥0T nx. It is easy
to see that PM1M realizes the relation TM
≺ T1M1. By Lemma 6.2,
TM ∼ S.
Lemma 6.5. Suppose that T ∈ L(H) is an a.c. polynomially bounded op-
erator, σ ⊂ T is a measurable set, X ∈ L(H, L2(σ)), XT = UσX and
Wk≥0 U −k
σ XH = L2(σ). Then there exists x ∈ H such that Xx =: f 6= 0
a.e. on σ.
Proof. Let {xn}N
n=1 be such that xn ∈ H, kxnk = 1 for all n, and
(cid:3)
d
where 1 ≤ N ≤ ∞. Set fn = Xxn. Then
T kxn = H,
N_n=1_k≥0
{ζ ∈ σ : fn(ζ) = 0}(cid:17) = 0.
m(cid:16) N\n=1
Take {an}N
n=1 such that an > 0 for all n, andPN
σ1n = {ζ ∈ σ :
fn(ζ) > 1}.
n=1 an < ∞. Set
SHIFT-TYPE INVARIANT SUBSPACES
23
Let {ϕn}N
n=1 ⊂ H ∞ be such that
ϕn =(an
an
fn
on T \ σ1n,
on σ1n.
Set x1n = ϕn(T )xn and f1n = Xx1n. Then kx1nk ≤ Cpol,T an and f1n ≤ an
a.e. on σ. Set f0 = supn f1n. Then 0 < f0 < ∞ a.e. on σ. Take 0 < δ < 1.
Set
Set
Then
σ2n = {ζ ∈ σ : δf0(ζ) ≤ f1n(ζ)}.
τ1 = σ21, τn = σ2n \ ∪n−1
k=1τk, n ≥ 2.
τn = σ and τn ∩ τk = ∅,
N[n=1
n=1 such that εn > 0 for all n, andPN
Take {εn}N
H ∞ be such that
if n 6= k, n, k ≥ 1.
n=1 εn < δ. Let {ηn}N
n=1 ⊂
f ≥ f1n −Xk6=n
NXk=1
εk(cid:17)f0
a.e. on τn.
Thus, x satisfies the conclusion of the lemma.
(cid:3)
Theorem 6.6. Suppose that T1 is a polynomially bounded operator such
that T1 ∼ S. Then for every c > 0 there exist M ∈ Lat T1 and W ∈
L(H 2,M) such that W is invertible, W S = T1MW and kWkkW −1k ≤
(1 + c)(cid:0)√2(K 2 + 2)KCpol,T1 + 1(cid:1)qK 2C 2
+ 1KC 2
pol,T1
.
pol,T1
Proof. Applying [K1, Theorem 1], it is easy to see that (T1)(a)
+ ∼= S. Denote
by X+ the canonical intertwining mapping for T1 and S constructed using
a Banach limit. Denote by H1 the space in which T1 acts. Let
1+c ≤ c3 <
c2 < 1. By Lemma 4.2, there exists x1 ∈ H1 such that kx1k = 1 and
kX+x1k ≥ c2. Set f1 = X+x1. Denote by ψx1 the function from Corollary
2.1 applied to T1 and x1. By Lemma 4.1,
1
(6.1)
f1 ≤ kX+kKCpol,T1ψx1 a.e. on T.
Since f1 ∈ H 2, we haveRT log ψx1dm > −∞. Taking into account Remark
2.3, we can assume that ψx1 is outer. Put g1 = f1/ψx1. Then g1 ∈ H ∞.
Put
Then
We have
on τn,
on T \ τn.
ηn(T )x1n.
ηnf1n.
εn
x =
ηn =(1
NXn=1
NXn=1
εkf1k ≥(cid:16)δ −
f =
24
M.F. GAMAL'
Acting as in Proposition 4.4, we obtain that
(6.2)
Namely, if f1 ≤ c3ψx1 a.e. on T, then
kg1k∞ > c3.
c2
2 ≤ kX+x1k2 = kf1k2 =ZT f12dm ≤ZT
3ψx12dm ≤ c2
c2
3,
becauseRT ψx12dm ≤ 1, a contradiction with the choice of c2 and c3.
Set N =Wk≥0 T k
1 x1, T0 = T1N , and X0 = X+N . Taking into account
that Sclos X+N ∼= S and deleting on the inner factor of f1, we can assume
that g1 is outer and we accept that X0 realizes the relation T0 ≺ S. By
Corollary 2.1, Remark 2.3, and Lemma 6.1 there exists Y0 ∈ L(H 2,N ) such
that kY0k ≤ KCpol,T1, Y0 realizes the relation S ≺ T0, and X0Y0 = g1(S).
Clearly T0 ∈ L(N ) satisfies the assumption of Theorem 1.1. We prove
Theorem 1.1 for T0 and obtain estimates of kWk and kZk for W and Z
constructed in the proof of Theorem 1.1.
(S)X0 and
If g1 ≥ c3 a.e. on T, then set W = Y0. We have W −1 = 1
g1
. Therefore, we assume that
kWkkW −1k ≤ (1 + c)KC 2
(6.3)
pol,T1
m({ζ ∈ T :
g1(ζ) ≥ c3}) < 1.
Applying Proposition 4.7 to T0, Y0, X0, and g1, we obtain
T ∈ L(N ⊕ H 2
−) and X ∈ L(N ⊕ H 2
−, L2(T)).
Taking into account the estimates of kY0k, of kg1k∞ by (6.1), and evident
inequality Cpol,T0 ≤ Cpol,T1, we obtain that
(6.4)
+ 1.
and kXk ≤ Cpol,T1qK 2C 2
pol,T1
Cpol,T ≤pK 2 + 2Cpol,T1
Since T0 is of class C10, we have τ = ∅, X∗
∗ = g1(UT).
Therefore, g = g1 and h01 = h0, where g and h01 are defined in (5.3) and
(5.8), respectively (h0 will be chosen below; g and h01 will be used in (5.42)).
We replace (5.5) by
∗ = Y0⊕IH 2
, and XX∗
−
(6.5)
It allows to improve the estimate of kZk constructed in the end of Theorem
1.1. By (6.3) and (6.2), 0 < m(σ) < 1.
g1(ζ) < c3}.
σ = {ζ ∈ T :
Set h0(ζ) = 1 (ζ ∈ T). Define x0 as in (5.14). Namely, set x0 = X∗
Take 0 < ε00, γ0, δ0 < 1 and apply Proposition 4.5 as in the proof of
Theorem 1.1. The first inequality in (5.5) and the choice of δ0 in the proof
of Theorem 1.1 was used to apply Lemmas 4.6 and 4.3 and obtain that ψn
constructed by Corollary 2.1 applied to ϕn(T )x0 satisfy (5.19). In the case
T0 ∼ S considered now (5.20) implies
∗ χσ.
Take t0 > 0 and set
log ψndm > −∞ for every n ∈ N.
Zσ
ψ0n =(ψn
max(ψn, t0)
on σ,
on T \ σ.
SHIFT-TYPE INVARIANT SUBSPACES
25
Then ψ0n satisfy (5.19) for all n ∈ N. Repeat the remaining part of the
proof of Theorem 1.1 with ψ0n instead of ψn. We obtain the constants C1,
C2, the function α defined as in (5.42) which satisfies (5.47) and intertwining
transformations W and Z with the following estimates:
C1 =
kX∗
∗k
KCpol,T (1 + ε0) ≤
1
(1 + ε0)
,
because kX∗k ≤ KCpol,T1 ≤ KCpol,T ;
(2(1 + c2
1 − c1
ξ2
C2 ≤ K 2C 2
pol,T (1 + ε0)
+ KCpol,T1(cid:0)1 +Xn≥2
n(cid:1);
1))1/2
max(cid:0)(1 +Xn≥2
ξ2
n)C1, 1 + ξ(cid:1)
kWk ≤ C2;
kZk ≤(cid:13)(cid:13)(cid:13)
1
α(cid:13)(cid:13)(cid:13)∞kXk.
Choosing c1, ε00, {ξn}n∈N close to 0 and c3, γ0, δ0 close to 1, taking into
account that ε0 ≤ ε00, the definition (5.25) of ξ, the estimates before (5.48)
and (5.49), the choice of σ by (6.5), and (6.4), we obtain the conclusion of
(cid:3)
the theorem.
Remark 6.7. There exist operators T which satisfy the assumptions of
Theorem 1.1 and have non-trivial factorization of the form (cid:18)C·1
see [SFBK, Sec.
IX.2]. Therefore, to obtain estimates of the norms of
intertwining transformations in terms of Cpol,T , a separate consideration of
operators of class C10 is needed in the present construction.
O C·0(cid:19),
∗
Proof of Corollary 1.2. We have T = Ta ∔ Ts, where Ta is a.c. and Ts
is similar to a singular unitary operator, see [M] or [K3]. Since there is no
nonzero transformation intertwining a.c. and singular unitaries, we conclude
d
that Ta
≺ UT. Let x be from Lemma 6.5 applied to Ta (with σ = T). Set
N0 = _k≥0
T k
a x.
d
Then N0 ∈ Lat T , TN0 is a cyclic a.c. polynomially bounded operator and
TN0
≺ UT. By Lemma 6.3,(cid:0)TN0(cid:1)(a) ∼= UT. Let
∗O T1(cid:19) ,
TN0 =(cid:18)T0
1 ∼=(cid:0)TN0(cid:1)(a). Since TN0 is cyclic and a.c.,
where T0 and T1 are of classes C0· and C1·, respectively, see Introduction and
[K1]. By [K1, Theorem 3], T (a)
we have T1 is cyclic and a.c., too. By Theorem 1.1 applied to T1, there exists
M1 ∈ Lat T1 such that T1M1 ≈ S. By Lemma 6.4, there exists M ∈ Lat T
such that TM ∼ S. It remains to apply Theorem 6.6 to TM.
(cid:3)
26
M.F. GAMAL'
References
[BP] H. Bercovici and B. Prunaru, Quasiaffine transforms of polynomially bounded oper-
ators, Arch. Math. (Basel), 71 (1998), 384 -- 387.
[BT] H. Bercovici and K. Takahashi, On the reflexivity of contractions on Hilbert spaces,
J. London Math. Soc. (2), 32 (1985), 149 -- 156.
[B] J. Bourgain, New Banach space properties of the disc algebra and H ∞, Acta Math.,
152 (1984), 1 -- 48.
[C] J. B. Conway, The theory of subnormal operators, Math. Surveys and Monographs,
36, AMS, 1991.
[F] S. R. Foguel, A counterexample to a problem of Sz.-Nagy, Proc. Amer. Math. Soc.,
15 (1964), 788-790.
[G] M. F. Gamal', Examples of cyclic polynomially bounded operators that are not similar
to contractions, Acta Sci. Math. (Szeged), 82 (2016), 597 -- 628.
[K1] L. K´erchy, Isometric asymptotes of power bounded operators, Indiana Univ. Math.
J., 38 (1989), 173 -- 188.
[K2] L. K´erchy, Shift-type invariant subspaces of contractions, J. Funct. Anal., 246 (2007),
281 -- 301.
[K3] L. K´erchy, Quasianalytic polynomially bounded operators, in: Operator Theory: the
State of the Art, Theta, Bucharest, 2016, 75 -- 101.
[K4] L. K´erchy, Quasianalytic n-tuples of Hilbert space operators, J. Operator Theory, 81
(2019), 3 -- 20.
[M] W. Mlak, Algebraic polynomially bounded operators, Ann. Polon. Math., 29 (1974),
133 -- 139.
[N] N. K. Nikolski, Operators, functions, and systems: an easy reading. Volume I: Hardy,
Hankel, and Toeplitz, Math. Surveys and Monographs, 92, AMS, 2002.
[P] G. Pisier, A polynomially bounded operator on Hilbert space which is not similar to
a contraction, J. Amer. Math. Soc., 10 (1997), no. 2, 351 -- 369.
[R] O. R´ejasse, Factorization and reflexivity results on polynomially bounded operators,
J. Oper. Theory, 60(2008), 219 -- 238.
[SFBK] B. Sz.-Nagy, C. Foias, H. Bercovici and L. K´erchy, Harmonic analysis of operators
on Hilbert space, Springer, New York, 2010.
[T1] K. Takahashi, The reflexivity of contractions with non-reductive ∗-residual parts,
Michigan Math. J., 34(1987), 153 -- 159.
[T2] K. Takahashi, On the reflexivity of contractions with isometric parts, Acta Sci. Math.
(Szeged), 53(1989), 147 -- 152.
M.F. GAMAL', St. Petersburg Branch, V. A. Steklov Institute of Math-
ematics, Russian Academy of Sciences, Fontanka 27, St. Petersburg, 191023,
Russia
E-mail address: [email protected]
|
1102.3998 | 1 | 1102 | 2011-02-19T16:20:16 | Almost periodic generalized functions | [
"math.FA",
"math.CA"
] | The aim of this paper is to introduce and to study an algebra of almost periodic generalized functions containing the classical Bohr almost periodic functions as well as almost periodic Schwartz distributions | math.FA | math |
ALMOST PERIODIC GENERALIZED FUNCTIONS
CHIKH BOUZAR AND MOHAMMED TAHA KHALLADI
Abstract. The aim of this paper is to introduce and to study
an algebra of almost periodic generalized functions containing the
classical Bohr almost periodic functions as well as almost periodic
Schwartz distributions.
1. Introduction
The theory of uniformly almost periodic functions was introduced
and studied by H. Bohr, since then, many authors contributed to the
development of this theory. There exist three equivalent definitions
of uniformly almost periodic functions, the first definition of H. Bohr,
S. Bochner's definition and the definition based on the approximation
property, see [1]. S. Bochner's definition is more suitable for exten-
sion to Schwartz distributions. L. Schwartz in [6] introduced the basic
elements of almost periodic distributions.
The new generalized functions of [2], [3], give a solution to the prob-
lem of multiplication of distributions, these generalized functions are
currently the subject of many scientific works, see [4] and [5].
The aim of this work is to introduce and to study an algebra of
almost periodic generalized functions containing the classical Bohr al-
most periodic functions as well almost periodic Schwartz distributions.
2. Almost periodic functions and distributions
We consider functions and distributions defined on the whole one
dimensional space R. Recall Cb the space of bounded and continuous
complex valued functions on R endowed with the norm k k∞ of uniform
convergence on R, (Cb, k k∞) is a Banach algebra.
Definition 1. (S. Bochner) A complex valued function f defined
and continuous on R is called almost periodic, if for any sequence
of real numbers (hn)n one can extract a subsequence (hnk)k such that
1991 Mathematics Subject Classification. Primary 46F30; Secondary 42A75.
Key words and phrases. Almost periodic functions; Almost periodic distribu-
tions; Colombeau algebra; Almost periodic generalized functions.
1
2
CHIKH BOUZAR AND MOHAMMED TAHA KHALLADI
(f (. + hnk))k converges in (Cb, k k∞). Denoted by Cap the space of al-
most periodic functions.
To recall Schwartz almost periodic distributions, we need some func-
tion spaces, see [6]. Let p ∈ [1, +∞] , the space
endowed with the topology defined by the countable family of norms
DLp :=(cid:8)ϕ ∈ C∞ : ϕ(j) ∈ Lp, ∀j ∈ Z+(cid:9)
ϕk,p :=Xj≤k(cid:13)(cid:13)ϕ(j)(cid:13)(cid:13)Lp , k ∈ Z+,
is a differential Frechet subalgebra of C∞. The topological dual of DL1,
denoted by D′
L∞, is called the space of bounded distributions.
Let h ∈ R and T ∈ D′, the translate of T by h, denoted by τhT , is
defined as :
hτhT , ϕi = hT , τ−hϕi , ϕ ∈ D,
where τ−hϕ (x) = ϕ (x + h) .
The definition and characterizations of an almost periodic distribu-
tion are summarized in the following results.
Theorem 1. For any bounded distribution T ∈ D′
statements are equivalent :
L∞, the following
i) The set {τhT, h ∈ R} is relatively compact in D′
ii) T ∗ ϕ ∈ Cap, ∀ ϕ ∈ D.
L∞.
iii) ∃ (fj)j≤k ⊂ Cap, T = Pj≤k
T ∈ D′
f (j)
j
.
L∞ is said almost periodic if it satisfies any (hence every) of
the above conditions.
Definition 2. The space of almost periodic distributions is denoted by
B′
ap.
Let recall the space of regular almost periodic functions.
Definition 3. The space of almost periodic infinitely differentiable
functions on R is defined and denoted by
Bap =(cid:8)ϕ ∈ DL∞ : ϕ(j) ∈ Cap , ∀j ∈ Z+(cid:9) .
Some, easy to prove, properties of Bap are given in the following
assertions.
Proposition 1. We have
i) Bap is a closed differential subalgebra of DL∞.
ii) If T ∈ B′
iii) Bap ∗ L1 ⊂ Bap .
iv) Bap = DL∞ ∩ Cap .
ap and ϕ ∈ Bap , then ϕT ∈ B′
ap.
ALMOST PERIODIC GENERALIZED FUNCTIONS
3
As a consequence of (iv) , we have the following result.
Corollary 1. If v ∈ DL∞ and v ∗ ϕ ∈ Cap, ∀ϕ ∈ D, then v ∈ Bap.
Remark 1. It is important to mention that Bap$C∞ ∩ Cap.
3. Almost periodic generalized functions
Let I = ]0, 1] and
ML∞ =n(uε)ε ∈ (DL∞)I , ∀k ∈ Z+, ∃m ∈ Z+,
NL∞ =n(uε)ε ∈ (DL∞)I , ∀k ∈ Z+, ∀m ∈ Z+,
uεk,∞ = O(cid:0)ε−m(cid:1) , ε −→ 0o
uεk,∞ = O (εm) , ε −→ 0o
Definition 4. The algebra of bounded generalized functions, denoted
by GL∞, is defined by the quotient
GL∞ =
ML∞
NL∞
Define
(3.1)
Map =n(uε)ε ∈ (Bap)I , ∀k ∈ Z+, ∃m ∈ Z+, uεk,∞ = O (ε−m) , ε −→ 0o
Nap =n(uε)ε ∈ (Bap)I , ∀k ∈ Z+, ∀m ∈ Z+, uεk,∞ = O (εm) , ε −→ 0o
The properties of Map and Nap are summarized in the following
proposition.
Proposition 2. i) The space Map is a subalgebra of (Bap)I .
ii) the space Nap is an ideal of Map.
Proof. i) It follows from the fact that Bap is an differential algebra.
ii) Let (uε)ε ∈ Nap and (vε)ε ∈ Map, we have
∀k ∈ Z+, ∃m′ ∈ Z+, ∃c1 > 0, ∃ε0 ∈ I, ∀ε < ε0, vεk,∞ < c1ε−m′
.
Take m ∈ Z+, then for m′′ = m+m′, ∃c2 > 0 such that uεk,∞ < c2εm′′
.
Since the family of the norms uεk,∞ is compatible with the algebraic
structure of DL∞, then ∀k ∈ Z+, ∃ck > 0 such that
uεvεk,∞ ≤ ck uεk,∞ vεk,∞ ,
consequently
uεvεk,∞ < ckc2εm
′′
c1ε−m′
≤ Cεm, where C = c1c2ck.
Hence (uεvε)ε ∈ Nap.
(cid:3)
The following definition introduces the algebra of almost periodic
generalized functions.
4
CHIKH BOUZAR AND MOHAMMED TAHA KHALLADI
Definition 5. The algebra of almost periodic generalized functions is
the quotient algebra
Gap =
Map
Nap
We have a characterization of elements of Gap similar to the result
(ii) of theorem (1) for almost periodic distributions.
Theorem 2. Let u = [(uε)ε] ∈ GL∞, the following assertions are equiv-
alent :
i) u is almost periodic.
ii) uε ∗ ϕ ∈ Bap, ∀ε ∈ I, ∀ϕ ∈ D.
Proof. i) =⇒ ii) If u ∈ Gap, so for every ε ∈ I we have uε ∈ Bap, then
uε ∗ ϕ ∈ Bap, ∀ε ∈ I, ∀ϕ ∈ D.
ii) =⇒ i) Let (uε)ε ∈ ML∞ and uε ∗ ϕ ∈ Bap, ∀ε ∈ I, ∀ϕ ∈ D,
therefor uε ∈ Bap follows from theorem (1) (ii); it suffices to show that
∀k ∈ Z+, ∃m ∈ Z+, uεk,∞ = O(cid:0)ε−m(cid:1) , ε −→ 0,
which follows from the fact that u ∈ GL∞.
(cid:3)
Remark 2. The characterization (ii) does not depend on representa-
tives.
Definition 6. Denote by Σ the subset of functions ρ ∈ S satisfying
Z ρ (x) dx = 1 and Z xkρ (x) dx = 0, ∀k = 1, 2, ...
Set ρε (.) = 1
ε ρ(cid:0) .
ε(cid:1) , ε > 0.
Proposition 3. Let ρ ∈ Σ, the map
iap : B′
ap −→
u −→ (u ∗ ρε)ε + Nap,
Gap
is a linear embedding which commutes with derivatives.
Proof. Let u ∈ B′
we have u = Pβ≤m
ap, by characterization of almost periodic distributions
f (β)
β , where fβ ∈ Cap, so ∀α ∈ Z,
1
consequently, there exists c > 0 such that
(cid:12)(cid:12)(cid:0)u(α) ∗ ρε(cid:1) (x)(cid:12)(cid:12) ≤Xβ≤m
x∈R(cid:12)(cid:12)(cid:0)u(α) ∗ ρε(cid:1) (x)(cid:12)(cid:12) ≤Xβ≤m
εα+βZR (cid:12)(cid:12)fβ (x − εy) ρ(α+β) (y)(cid:12)(cid:12) dy,
εα+β kfβkL∞(R)ZR (cid:12)(cid:12)ρ(α+β) (y)(cid:12)(cid:12) dy ≤
sup
1
c
,
εα+m
ALMOST PERIODIC GENERALIZED FUNCTIONS
5
i.e.
u ∗ ρεm′,∞ = Xα≤m′
sup
x∈R(cid:12)(cid:12)(cid:0)u(α) ∗ ρε(cid:1) (x)(cid:12)(cid:12) ≤
c′
εm+m′ , c′ = Xα≤m′
c
εα
,
u∗ρε = u in D′
L∞, but lim
ε−→0
u∗ρε = 0
this shows that (u ∗ ρε)ε ∈ Map. Let (u ∗ ρε)ε ∈ Nap, then lim
ε−→0
in D′
L∞, this shows that iap is an embedding.
Finally we note that iap is linear, this results from the fact that the
ε =
(cid:3)
convolution is linear and that iap(cid:0)w(j)(cid:1) =(cid:0)w(j) ∗ ρε(cid:1)ε = (w ∗ ρε)(j)
(iap (w))(j) .
The space Bap is embedded into Gap canonically, i.e.
σap : Bap −→
Gap
f −→ [(f )ε] = (f )ε + Nap
There is two ways to embed f ∈ Bap into Gap. Actually we have the
same result.
Proposition 4. The following diagram
Bap −→ B′
ap
σap ց ↓ iap
Gap
is commutative.
Proof. Let f ∈ Bap, we prove that (f ∗ ρε − f )ε ∈ Nap. By Taylor's
formula and the fact that ρ ∈ Σ, we obtain
kf ∗ ρε − f kL∞ ≤ εmsup
(−y)m
m!
x∈RZR (cid:12)(cid:12)(cid:12)(cid:12)
f (m) (x − θ (x) εy) ρ (y) dy(cid:12)(cid:12)(cid:12)(cid:12) ,
then ∃Cm > 0, such that
kf ∗ ρε − f kL∞ ≤ εmCm(cid:13)(cid:13)f (m)(cid:13)(cid:13)L∞ kymρkL1 .
The same result can be obtained for all the derivatives of f. Hence
(f ∗ ρε − f )ε ∈ Nap.
(cid:3)
The Colombeau algebra of tempered generalized functions on C is
denoted GT (C) , for more details on GT (C) see [2] or [4].
Proposition 5. Let u ∈ Gap and F ∈ GT (C), then
F ◦ u = [(F ◦ uε)ε]
is a well defined element of Gap.
6
CHIKH BOUZAR AND MOHAMMED TAHA KHALLADI
Proof. It follows from the classical case of composition, in context of
Colombeau algebra, we have F ◦uε ∈ Bap in view of the classical results
of composition and convolution.
(cid:3)
We recall a characterization of integrable distributions.
Definition 7. A distribution v ∈ D′ is said an integrable distribution,
denoted v ∈ D′
, where fi ∈ L1.
f (i)
i
Proposition 6. If u = [(uε)ε] ∈ Gap and v ∈ D′
u ∗ v defined by
L1, if and only if v =Pi≤l
(u ∗ v) (x) =ZR
uε (x − y) v (y) dyε
is a well defined almost periodic generalized function.
+ N [C]
L1, then the convolution
Proof. Let (uε)ε ∈ Map be a representative of u, then
∀k ∈ Z+, ∃m ∈ Z+, ∃C > 0, ∃ε0 ∈ I, ∀ε ≤ ε0, uεk,∞ < Cε−m,
since v ∈ D′
, where fi ∈ L1. For each ε ∈ I, uε ∗v is an
almost periodic infinitely differentiable function. By Young inequality
there exists C > 0 such that
i
f (i)
L1 then v =Pi≤l
(cid:13)(cid:13)(cid:13)(uε ∗ v)(j)(cid:13)(cid:13)(cid:13)L∞
kfikL1(cid:13)(cid:13)u(i+j)
uε ∗ vk,∞ = O(cid:0)ε−m(cid:1) , ε −→ 0,
≤ CXi≤l
ε
(cid:13)(cid:13)L∞ ,
consequently
this shows that (uε ∗ v)ε ∈ Map. Suppose that (wε)ε ∈ Map is another
representative of u, then there exists C > 0 such that
k(uε ∗ v − wε ∗ v)kL∞ ≤ Xi≤l
≤ CXi≤l
as (uε − wε)ε ∈ Nap, so ∀m ∈ Z+,
sup
R ZR (cid:12)(cid:12)(cid:12)(uε − wε)(i) (x − y)(cid:12)(cid:12)(cid:12) fi (y) dy
kfikL1(cid:13)(cid:13)(cid:13)(uε − wε)(i)(cid:13)(cid:13)(cid:13)L∞
,
(uε ∗ v − wε ∗ v) (x) = O (εm) , ε −→ 0.
We obtain the same result for (uε ∗ v − wε ∗ v)(j)
Nap.
ε
. Hence (uε ∗ v − wε ∗ v)ε ∈
(cid:3)
If u = [(uε)ε] ∈ Gap, taking the integral of each element uε on a
compact, we obtain an element of CI .
ALMOST PERIODIC GENERALIZED FUNCTIONS
7
Definition 8. Let u = [(uε)ε] ∈ Gap and x0 ∈ R, define the primitive
of u by
U (x) =
xZx0
uε (t) dtε
+ N [C] .
We give a generalized version of the classical Bohl-Bohr theorem.
Proposition 7. The primitive of an almost periodic generalized func-
tion is almost periodic if and only if it is bounded generalized function.
Proof. (=⇒) : It follows from the fact that Gap ⊂ GL∞. (⇐=) : Let
u = [(uε)ε] ∈ Gap and let U = [(Uε)ε] be its primitive, since for each
uε (t) dt ∈ DL∞ then by the classical result
ε ∈ I, uε ∈ Bap and Uε =
xRx0
uε (t) dt ∈ Bap. We
xRx0
of Bohl-Bohr and Bap, for every ε ∈ I we have
shows that (Uε)ε ∈ Map, i.e.
∀k ∈ Z+, ∃m ∈ Z+, Uεk,∞ =Xj≤k
ε
sup
x∈R(cid:12)(cid:12)U (j)
(x)(cid:12)(cid:12) = O(cid:0)ε−m(cid:1) , ε −→ 0.
If j = 0, since (Uε)ε ∈ ML∞. By hypothesis, we have
sup
x∈R
Uε (x) = O(cid:0)ε−m(cid:1) , ε −→ 0,
(x)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)u(j−1)
which shows that (Uε)ε ∈ Map. If j ≥ 1, we have(cid:12)(cid:12)(cid:12)U (j)
which gives Xj≤k
x∈R(cid:12)(cid:12)U (j)
(x)(cid:12)(cid:12) ,
x∈R(cid:12)(cid:12)u(j−1)
Uεk,∞ = O(cid:0)ε−m(cid:1) , ε −→ 0,
(x)(cid:12)(cid:12) ≤Xj≤k
i.e. (Uε)ε ∈ Map.
consequently
sup
sup
ε
ε
ε
ε
(x)(cid:12)(cid:12)(cid:12) ,
(cid:3)
As in the classical theory, we introduce the notion of mean value
within the algebra Gap.
Definition 9. Let u ∈ Gap, the generalized mean value of u, denoted
by Mg (u) , is defined by
Mg (u) = lim
X−→+∞
1
X
XZ0
uε (x) dxε
+ N [C] ,
where (uε)ε is a representative of u.
8
CHIKH BOUZAR AND MOHAMMED TAHA KHALLADI
The definition of Mg (u) is correct and does not depend on represen-
tatives.
uε (x) = O (ε−m) , ε −→ 0, hence
≤ sup
x∈R
uε (x) ,
lim
X→+∞
1
X
XZ0
Proof. i) Let ε ∈ I, we have
as (uε)ε ∈ Map, so ∃m ∈ Z+, sup
x∈R
Proposition 8. i) If u = [(uε)ε] ∈ Gap, then Mg (u) ∈eC.
ii) If (uε)ε ∈ Nap, then Mg (u) = 0 in eC.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
uε (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
uε (x) dxε
lim
XZ0
uε (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
XZ0
then Mg (u) = 0 in eC.
ii) If (uε)ε ∈ Nap, i.e.
∈ EM [C] .
lim
X→+∞
1
X
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1
X
X→+∞
distribution as stated in the following.
= O (εm) , ε −→ 0, ∀m ∈ Z+,
We have compatibility of the generalized mean value with that of a
(cid:3)
Proposition 9. If T ∈ B′
ap, then Mg (iap (T )) = M (T ) in C.
Proof. We have
Mg (iap (T )) = lim
X−→+∞
1
X
where ρ ∈ Σ and ρε (.) = 1
iap (T ) (h) dh = lim
XZ0
ε(cid:1) . Let ϕ ∈ D andRR
ε ρ(cid:0) .
X−→+∞
1
X
XZ0
(T ∗ ρε) (h) dhε
,
ϕ (x) dx = 1, then
M (T ) = lim
X−→+∞
1
X
(T ∗ ϕ) (h) dh.
XZ0
We have
Mg (iap (T )) − M (T ) = M (T ∗ (ρε − ϕ)) , ∀ε ∈ I.
ALMOST PERIODIC GENERALIZED FUNCTIONS
9
In view of formula (V I.9.2) of [6], we obtain
Mg (iap (T )) − M (T ) = M (T )ZR
(ρε (x) − ϕ (x)) dx,
as ∀ε ∈ I,RR
C.
(ρε (x) − ϕ (x)) dx = 0, then Mg (iap (T )) = M (T ) in
(cid:3)
Remark 3. We can introduce a new association within Gap with the
aid of the generalized mean value Mg.
Definition 10. A generalized trigonometric polynomial is a generalized
function [(Pε)ε] , where
Pε (x) =
cε,neiλε,nx ,
lXn=1
(cε,n)ε ∈eC and (λε,n)ε ∈eR, n = 1, ..., l.
Proposition 10. Every generalized trigonometric polynomial is an al-
most periodic generalized function.
Proof. Let Pε (x) =
n = 1, ..., l, we have
lPn=1
cε,neiλε,nx where (cε,n)ε ∈ eC and (λε,n)ε ∈ eR,
∀ε ∈ I,
cε,neiλε,nx ∈ Bap,
lXn=1
λε,nj cε,n
0. Consequently ∀k ∈ Z+, ∃C > 0 such that
moreover ∃m ∈ Z+, ∃m′ ∈ Z+, λε,n = O (ε−m) and cε,n = O(cid:0)ε−m′(cid:1) , ε −→
lXn=1
Pεk,∞ ≤ Xj≤k
λε,1j + λε,2j + ... + λε,lj! c′ε−m′
≤ Xj≤k
≤ k(cid:0)cε−m + c2ε−2m + ... + ckε−km(cid:1) c′ε−m′
, where C = c′kl(cid:0)c + c2 + ... + ck(cid:1) , m′′ = km + m′,
Let u ∈ Gap and eλ = [(λε)ε] ∈ eR, then ue−ieλx =(cid:0)uεe−iλεx(cid:1)ε ∈ Gap,
so the generalized mean value Mg(cid:16)ue−ieλx(cid:17) is a well defined element of
this show that (Pε)ε ∈ Map. In a similar way we show that if (λε,n)ε ∈
N [R] and (cε,n)ε ∈ N [C] , then (Pε (x))ε ∈ Nap.
(cid:3)
≤ Cε−m′′
10
CHIKH BOUZAR AND MOHAMMED TAHA KHALLADI
eC. Define
and the generalized spectra of u,
aeλ (u) = Mg(cid:16)ue−ieλx(cid:17) ,
Λg (u) =neλ ∈eR : aeλ (u) 6= 0 in eCo .
Remark 4. If a function or a distribution u is almost periodic, then
its spectra is at most countable, see [1] and [6].
Proposition 11. Let P =(cid:20)(cid:18) lPn=1
metric polynomial, then
cε,neiλε,nx(cid:19)ε(cid:21) be a generalized trigono-
Proof. A direct computation gives the result.
Λg (P ) =(cid:8)(cid:2)(λε,n)ε(cid:3) : n = 1, ..., l(cid:9) .
(cid:3)
Remark 5. We are indebted to the referee for the following example of
an almost periodic generalized function having uncountable generalized
spectra. Let u = [(uε)ε] , uε (x) = 1 + eix, ∀ε ∈ I, by proposition (11) ,
Λg (u) = {[(Aε)ε] : Aε = {0, 1}} , as the set {(λε)ε : λε = {0, 1} , ε ∈ I} ⊂
Λg (u) is uncountable, then the generalized spectra of u is too.
References
[1] H. Bohr. Almost periodic functions. Chelsea Publishing Company, 1947.
[2] J. F. Colombeau. New generalized functions and multiplication of distributions.
North-Holland, 1984.
[3] J. F. Colombeau. Elementary Introduction to New Generalized Functions.
North-Holland, 1985.
[4] M. Grosser, M. Kuzinger, M. Oberguggenberger,R. Steinbauer. Geometric the-
ory of generalised functions. Kluwer, 2001.
[5] A. Kaminski, M. Oberguggenberger, S. Pilipovic (Eds). Linear and non-linear
theory of generalized functions and its applications. Banach Center Publica-
tions, Vol. 88, 2010.
[6] L. Schwartz. Th´eorie des distributions. Hermann, 1966.
Oran Essenia University, Algeria
E-mail address: [email protected]
University of Adrar, Algeria
E-mail address: [email protected]
|
1801.06884 | 3 | 1801 | 2018-03-22T14:39:19 | On the Existence of Normal Square and Nth Roots of Operators | [
"math.FA",
"math.OA"
] | The primary purpose of this paper is to show the existence of normal square and nth roots of some classes of bounded operators on Hilbert spaces. Two interesting simple results hold. Namely, under simple conditions we show that if any operator $T$ is such that $T^2=0$, then this implies that $T$ is normal and so $T=0$. Also, we will see when the square root of an arbitrary bounded operator is normal. | math.FA | math |
ON THE EXISTENCE OF NORMAL SQUARE AND NTH ROOTS
OF OPERATORS
MOHAMMED HICHEM MORTAD
Abstract. The primary purpose of this paper is to show the existence of
normal square and nth roots of some classes of bounded operators on Hilbert
spaces. Two interesting simple results hold. Namely, under simple conditions
we show that if any operator T is such that T 2 = 0, then this implies that T
is normal and so T = 0. Also, we will see when the square root of an arbitrary
bounded operator is normal.
1. Introduction
Let H be a complex Hilbert space, and let B(H) denote the algebra of all
bounded linear operators on H.
By a square root of A ∈ B(H), we mean a B ∈ B(H) such that B2 = A. An
A ∈ B(H) is called positive if
< Ax, x >≥ 0, ∀x ∈ H.
1
It is known that any positive operator A has a unique positive square root (which
is denoted by √A or A
2 and these notations are exclusively reserved to the unique
positive square root). However, there are non normal N such that N 2 is normal.
As an extreme example, just consider any non-normal N such that N 2 = 0. In
fact, the identity 2× 2 matrix I has an infinite number of self-adjoint square roots.
Indeed, the self-adjoint
x
Ax = (cid:18)
√1 − x2
−x
represents a square root of I for each x ∈ [−1, 1].
There are some quite known research papers on this topic. For instance, readers
may wish to consult: [2], [4], [7], [8], [9] and [12]. In particular, Putnam [7] gave a
condition guaranteeing that the square root of a normal operator be normal. The
problem considered in this paper is of this sort.
√1 − x2
(cid:19)
self-adjoint. As it is customary, we denote A by ReT and B by ImT .
Recall that any T ∈ B(H) is expressible as T = A + iB where A, B ∈ B(H) are
It is also known (cf. [6]) that T = 0 iff A = B = 0. Therefore, if A+iB = C +iD,
It is also easily
then A = C and B = D whenever A, B, C, D are self-adjoint.
verifiable that T is normal iff AB = BA.
We also recall some known results which will be called on below (these are
standard facts, see [6] for proofs).
2010 Mathematics Subject Classification. Primary 47A62, Secondary 47A05.
Key words and phrases. Square roots. Normal, self-adjoint and positive operators. Real and
imaginary parts of an operator. Spectrum. Cartesian decomposition.
1
2
M. H. MORTAD
Theorem 1.1. Let A, B ∈ B(H) be self-adjoint. Then:
√B
0 ≤ A ≤ B =⇒
√A ≤
Proposition 1.2. Let A, B ∈ B(H) be self-adjoint. Then
where A = √A∗A.
A ≤ B =⇒ −B ≤ A ≤ B,
In the end, we assume that readers are familiar with other notions and results
on B(H).
2. Existence of Normal Roots
Theorem 2.1. Let C ∈ B(H) be self-adjoint and let T = A + iB ∈ B(H) be such
that T 2 = C.
(1) If σ(A) ∩ σ(−A) = ∅, then T is self-adjoint and invertible.
(2) If σ(B) ∩ σ(−B) = ∅, then T is skew symmetric (that is, T ∗ = −T ) and
invertible.
The proof is based upon the following known result:
Lemma 2.2. ([10] or [1]) Let A, B ∈ B(H) be such that σ(A) ∩ σ(B) = ∅. Then
the equation AX − XB = S has a unique solution X (in B(H)) for each S ∈ B(H).
Now, we prove Theorem 2.1:
Proof. We have
T 2 = C ⇐⇒ T 2 = A2 − B2 + i(AB + BA) = C.
By the self-adjointness of A and B, we obtain the self-adjointness of A2 − B2 and
AB + BA as well. Hence
(cid:26) A2 − B2 = C,
AB + BA = 0.
(1) If σ(A) ∩ σ(−A) = ∅, then Lemma 2.2 says that the equation
AB − B(−A) = AB + BA = 0
has a unique solution which is necessarily B = 0. Hence T = A is self-
adjoint. If 0 ∈ σ(T ), then 0 ∈ σ(A) and so 0 ∈ σ(−A) too. This, however,
would violate the assumption σ(A)∩σ(−A) = ∅. Therefore, T is invertible.
(2) When σ(B) ∩ σ(−B) = ∅, a similar method yields A = 0, i.e. T = iB.
Hence T is invertible because σ(B) ∩ σ(−B) = ∅.
(cid:3)
A result by Embry [3] may be readjusted as follows:
Lemma 2.3. Let A, B ∈ B(H) be normal and such that AB = −BA. Designate
the numerical range of A by W (A). If σ(A) ∩ σ(−A) = ∅ or 0 6∈ W (A) (resp.
σ(B) ∩ σ(−B) = ∅ or 0 6∈ W (B)), then B = 0 (resp. A = 0).
Using the same method as above and the foregoing lemma, we may establish:
Proposition 2.4. Let C ∈ B(H) be self-adjoint and let T = A + iB ∈ B(H) be
such that T 2 = C. If 0 6∈ W (A) (or 0 6∈ W (B)), then T is self-adjoint. As above,
in the first case T = A and in the second T = iB.
ROOTS OF OPERATORS
3
We know that if T ∈ B(H), then T 2 = 0 does not, in general, imply that T = 0.
We also know that if T satisfies kT 2k = kTk2 (for example, if T is self-adjoint or
normal or normaloid in general i.e. kT nk = kTkn for all n), then T 2 = 0 does imply
that T = 0. The following result is therefore of interest.
Proposition 2.5. Let T ∈ B(H) be such that T 2 = 0. If ReT ≥ 0 (or ImT ≥ 0),
then T is normal and so T = 0.
Proof. Write T = A + iB where A, B ∈ B(H) are self-adjoint where A = ReT and
B = ImT . Then clearly
So, if T 2 = 0, then
T 2 = A2 − B2 + i(AB + BA).
A2 − B2 + i(AB + BA) = 0 =⇒ (cid:26) A2 = B2,
AB = −BA.
Hence, if A ≥ 0 (a similar argument works when B ≥ 0), then
AB = −BA =⇒ A2B = −ABA = BA2 =⇒ AB = BA.
Therefore, T is normal. Accordingly,
kTk2 = kT 2k = 0 =⇒ T = 0,
as suggested.
Corollary 2.6. Let T ∈ B(H) be such that T 2 = 0. If any of σ(ReT ) or σ(ImT )
is a subset of either R+ or R−, then T is normal and so T = 0.
Remark. The condition ReT ≥ 0 (or ImT ≥ 0) in Proposition 2.5 is, in general, not
sufficient to make T normaloid. Indeed, if V is the Volterra operator on L2[0, 1]
say, then (cf. [6]) V is not normaloid as
(cid:3)
r(V ) = 0 6=
2
π
= kV k
where r(V ) denotes the spectral radius of V . It can, however, easily be checked
that ReV ≥ 0.
We finish this section with a general criterion guaranteeing the normality of the
square root (this generalizes Proposition 2.5):
Theorem 2.7. Let S = C + iD ∈ B(H) be such that T 2 = S where T = A + iB ∈
B(H). Then (if [·,·] denotes the usual commutator)
[B, C] = [A, D]
and
[A, C] = [B, D].
(Consequently, BC = CB ⇔ AD = DA and AC = CA ⇔ BD = DB).
If A ≥ 0 (or A ≤ 0), then
In particular, if S is self-adjoint (i.e. D = 0) and A ≥ 0 or A ≤ 0 or B ≥ 0 or
B ≤ 0, then T is always normal.
If B ≥ 0 (or B ≤ 0), then
T is normal ⇐⇒ AD = DA.
T is normal ⇐⇒ BD = DB.
4
M. H. MORTAD
Proof. By assumption,
A2 − B2 + i(AB + BA) = C + iD =⇒ (cid:26) A2 − B2 = C,
AB + BA = D.
Hence
and so
AB + BA = D =⇒ A2B + ABA = AD and ABA + BA2 = DA
Since A2 = B2 + C, we get
Also,
A2B − BA2 = AD − DA.
BC − CB = AD − DA.
AB + BA = D =⇒ AB2 + BAB = DB and BAB + B2A = BD
and so as above
and by invoking B2 = A2 − C, we obtain
B2A − AB2 = BD − DB
To show the last two assertions, we have from above that
AC − CA = BD − DB.
A2B = BA2 ⇐⇒ AD = DA and B2A = AB2 ⇐⇒ BD = DB.
So if A ≥ 0, then clearly
A2B = BA2 ⇐⇒ AB = BA,
and the previous holds iff T = A + iB is normal. A similar reasoning applies when
A ≤ 0. Finally, argue similarly in the event B ≥ 0 or B ≤ 0 and this completes the
proof.
(cid:3)
3. Explicit Construction of Roots
It is known (cf.
[11]) that if N is normal with a spectral integral Rσ(N ) λdE
√λdE is square root of N where √λ is
where E is a spectral measure, then Rσ(N )
a complex square root of λ.
Under an extra condition on the normal operator, we can have an even more
explicit formula.
Theorem 3.1. Let N = C + iD ∈ B(H) be normal with either D ≥ 0 or D ≤ 0
(equivalently, σ(D) ⊂ R+ or σ(D) ⊂ R−). When D ≥ 0, then
1
2
1
2
1
2
1
2
T = (cid:18)N + C
2
(cid:19)
+ i(cid:18)N − C
2
(cid:19)
is a normal square root of N . If D ≤ 0, then
T = (cid:18)N + C
2
(cid:19)
is another normal square root of N .
− i(cid:18)N − C
2
(cid:19)
ROOTS OF OPERATORS
5
Proof. The hardest part of the proof is the meticulousness! First C and D are
self-adjoint. Besides, CD = DC as N is normal. Then
N2 = N ∗N = (C − iD)(C + iD) = C2 + D2 ≥ C2
as D2 ≥ 0 because D is self-adjoint. By Theorem 1.1, we get N ≥ C. Hence, by
Proposition 1.2
and so
−N ≤ C ≤ N
Therefore, it makes sense to define their positive square roots. Consider (the self-
adjoint!)
N − C ≥ 0 and N + C ≥ 0.
A = (cid:18)N + C
2
(cid:19)
1
2
and B = (cid:18)N − C
2
(cid:19)
1
2
.
Since N commutes with C, it follows that A commutes with B. Consequently, the
operator M := A + iB is normal. Finally,
(1) If D ≥ 0, then
M 2 = (A + iB)(A + iB)
= A2 − B2 + i(AB + BA)
= N + C
− N − C
2
2
+ 2i(cid:18)N2 − C2
4
1
2
(cid:19)
1
2
= C + i(D2)
= C + iD
= N,
that is M is a normal square root of N .
(2) A similar argument applies when D ≤ 0. In this case,
(M ∗)2 = N,
that is, M ∗ is a normal square root of N . This marks the end of the proof.
(cid:3)
This approach, besides its explicit construction, does apply for higher powers of
the type 2n. In other language, the algorithm prescribed in the previous proof may
be applied to deal with biquadratic equations or in general equations of the form
T 2n
= N where n ∈ N. We have:
Corollary 3.2. Let N = C + iD ∈ B(H) be normal with either D ≥ 0 (or D ≤ 0).
Let T ∈ B(H) be such that T 4 = N . Then a normal 4th root of T is given by
where
T = (cid:18)S + ReS
2
(cid:19)
S = (cid:18)N + C
2
(cid:19)
1
2
1
2
+ i(cid:18)S − ReS
2
(cid:19)
1
2
+ i(cid:18)N − C
2
(cid:19)
1
2
6
M. H. MORTAD
Proof. Put S = T 2. Then T 4 = N becomes S2 = N . So, if e.g. N is such that
σ(ImN ) ⊂ R+, then S is normal. Moreover, by Theorem 3.1, we know that
S = (cid:18)N + C
2
(cid:19)
and it is normal. Since clearly
1
2
+ i(cid:18)N − C
2
(cid:19)
1
2
ImS = (cid:18)N − C
2
(cid:19)
1
2
≥ 0,
it follows that T 2 = S has a normal solution T given by
T = (cid:18)S + ReS
2
(cid:19)
1
2
+ i(cid:18)S − ReS
2
(cid:19)
1
2
.
Remark. By induction, we know how to find a root of order 2n of normal operators
very explicitly.
We finish with the case of general nth roots. First, the following result should
(cid:3)
be readily verified.
Lemma 3.3. Let A ∈ B(H). If k ∈ Z, then
ei(A+2kπI) = eiA.
Theorem 3.4. Let N ∈ B(H) be normal. Let n ∈ N. Then N has always an nth
root which is also normal and given by
n ei( A+2kπI
for some self-adjoint A ∈ B(H) and where k ∈ Z.
Proof. Since N is normal, it may be expressed as (see e.g. [11])
n = N
N
)
n
1
1
where U is unitary and P = √N ∗N = N. Since U is unitary, U = eiA for some
self-adjoint A ∈ B(H) (cf. Proposition 18.20 in [5]). Hence
N = U P = P U,
N = eiAP = P eiA.
Set for k ∈ Z
M = P
1
n ei( A+2kπI
n
)
1
n is the unique positive nth root of P . Then M is normal (cf. [6]) because
where P
it is a product of two commuting normal operators (in fact, a product of a positive
operator and a unitary one). By the commutativity of the factors and simple results,
we obtain
M n = (cid:16)P
as needed.
1
n ei( A+2kπI
n
)(cid:17)n
= P ei(A+2kπI) = P eiA = N,
(cid:3)
Remark. As is the case of many results on normal operators, the previous two
theorems too are inspired by results about complex numbers. This corroborates
and strengthens the ressemblance which is already known to readers.
ROOTS OF OPERATORS
7
References
[1] R. Bhatia, P. Rosenthal, How and why to solve the operator equation AX − XB = Y . Bull.
London Math. Soc., 29/1 (1997) 1-21.
[2] J. B. Conway, B. B. Morrel, Roots and logarithms of bounded operators on Hilbert space, J.
Funct. Anal., 70/1 (1987) 171-193.
[3] M. R. Embry, Similarities Involving Normal Operators on Hilbert Space, Pacific J. Math., 35
(1970) 331-336.
[4] S. Kurepa, On n-th roots of normal operators, Math. Z., 78 (1962) 285-292.
[5] R. Meise, D. Vogt, Introduction to Functional Analysis, Oxford G.T.M. 2, Oxford University
Press 1997.
[6] M. H. Mortad, An Operator Theory Problem Book, World Scientific Publishing Co., (to
appear in 2018).
[7] C. R. Putnam, On Square Roots of Normal Operators, Proc. Amer. Math. Soc., 8 (1957)
768-769.
[8] C. R. Putnam, On Square Roots and Logarithms of Self-adjoint Operators, Proc. Glasgow
Math. Assoc., 4 (1958) 1-2.
[9] H. Radjavi, P. Rosenthal, On Roots of Normal Operators, J. Math. Anal. Appl., 34 (1971)
653-664.
[10] M. Rosenblum, On the Operator Equation BX −XA = Q, Duke Math. J., 23 (1956) 263-269.
[11] W. Rudin, Functional Analysis, McGraw-Hill Book Co., Second edition, International Series
in Pure and Applied Mathematics, McGraw-Hill, Inc., New York, 1991.
[12] J. G. Stampfli, Roots of Scalar Operators, Proc. Amer. Math. Soc., 13/5 (1962) 796-798.
Department of Mathematics, University of Oran 1, Ahmed Ben Bella, B.P. 1524,
El Menouar, Oran 31000, Algeria.
Mailing address:
Pr Mohammed Hichem Mortad
BP 7085 Seddikia Oran
31013
Algeria
E-mail address: [email protected], [email protected].
|
1307.6208 | 1 | 1307 | 2013-07-19T17:06:12 | Difference Sequence Spaces Derived by Using Generalized Means | [
"math.FA"
] | This paper deals with new sequence spaces $X(r, s, t ;\Delta) $ for $X\in \{l_\infty, c, c_0\}$ defined by using generalized means and difference operator. It is shown that these spaces are complete normed linear spaces and the spaces $X(r, s, t ;\Delta)$ for $X\in \{c, c_0\}$ have Schauder basis. Furthermore, the $\alpha$-, $\beta$-, $\gamma$- duals of these sequence spaces are computed and also established necessary and sufficient conditions for matrix transformations from $X(r, s, t ;\Delta)$ to $X$. | math.FA | math |
Difference Sequence Spaces Derived by Using
Generalized Means
Atanu Manna∗, Amit Maji†, P. D. Srivastava ‡
Department of Mathematics, Indian Institute of Technology, Kharagpur
Kharagpur 721 302, West Bengal, India
Abstract
This paper deals with new sequence spaces X(r, s, t; ∆) for X ∈ {l∞, c, c0} defined by using generalized
means and difference operator. It is shown that these spaces are complete normed linear spaces and
the spaces X(r, s, t; ∆) for X ∈ {c, c0} have Schauder basis. Furthermore, the α-, β-, γ- duals of these
sequence spaces are computed and also established necessary and sufficient conditions for matrix trans-
formations from X(r, s, t; ∆) to X.
2010 Mathematics Subject Classification: 46A45, 46A35.
Keywords: Sequence spaces; Difference operator; Generalized means; α-, β-, γ- duals; Matrix transfor-
mations.
1
Introduction
The study of sequence spaces has importance in the several branches of analysis, namely, the structural
theory of topological vector spaces, summability theory, Schauder basis theory etc. Besides this, the
theory of sequence spaces is a powerful tool for obtaining some topological and geometrical results using
Schauder basis.
Let w be the space of all real or complex sequences x = (xn), n ∈ N0. For an infinite matrix A and a
sequence space λ, the matrix domain of A, which is denoted by λA and defined as λA = {x ∈ w : Ax ∈ λ}
[17]. Basic methods, which are used to determine the topologies, matrix transformations and inclusion
relations on sequence spaces can also be applied to study the matrix domain λA.
In recent times,
there is an approach of forming new sequence spaces by using matrix domain of a suitable matrix and
characterize the matrix mappings between these sequence spaces.
Kizmaz first introduced and studied the difference sequence spaces in [8]. Later on, several authors
including Ahmad and Mursaleen [1], C¸ olak and Et [5], Ba¸sar and Altay [3], Orhan [13], Polat and
Altay [15], Aydin and Ba¸sar [4] etc. have introduced and studied new sequence spaces defined by using
difference operator.
∗Corresponding author e-mail: [email protected]
†Author's e-mail: [email protected]
‡Author's e-mail: [email protected]
1
On the other hand, sequence spaces are also defined by using generalized weighted means. Some
of them can be viewed in Malkowsky and Sava¸s [11], Altay and Ba¸sar [2]. Mursaleen and Noman [12]
introduced a sequence space of generalized means, which includes most of the earlier known sequence
spaces. But till 2011, there was no such literature available in which a sequence space is generated by
combining both the weighted means and the difference operator. This was firstly initiated by Polat et
al. [14]. The authors in [14] have introduced the sequence spaces λ(u, v; ∆) for λ ∈ {l∞, c, c0} defined as
where u, v ∈ w such that un, vn 6= 0 for all n and the matrices G(u, v) = (gnk), ∆ = (δnk) are defined by
λ(u, v; ∆) =(cid:8)x ∈ w : (G(u, v).∆)x ∈ λ(cid:9),
gnk =( unvk
0
if 0 ≤ k ≤ n,
if k > n
respectively.
0
(−1)n−k
0
if 0 ≤ k < n − 1
if n − 1 ≤ k ≤ n,
if k > n,
δnk =
The aim of this article is to introduce new sequence spaces defined by using both the generalized
means and the difference operator. We investigate some topological properties as well as α-, β-, γ- duals
and bases of the new sequence spaces are obtained. Further, we characterize some matrix transformations
between these new sequence spaces.
2 Preliminaries
Let l∞, c and c0 be the spaces of all bounded, convergent and null sequences x = (xn) respectively,
with norm kxk∞ = sup
xn. Let bs and cs be the sequence spaces of all bounded and convergent series
n
respectively. We denote by e = (1, 1, · · · ) and en for the sequence whose n-th term is 1 and others are
zero and N0 = N ∪ {0}, where N is the set of all natural numbers. A sequence (bn) in a normed linear
space (X, k.k) is called a Schauder basis for X if for every x ∈ X there is a unique sequence of scalars
(µn) such that
x −
(cid:13)(cid:13)(cid:13)
k
Xn=0
µnbn(cid:13)(cid:13)(cid:13)
→ 0 as k → ∞,
µnbn. ([17], [9]).
i.e., x =
∞
Xn=0
For any subsets U and V of w, the multiplier space M (U, V ) of U and V is defined as
M (U, V ) = {a = (an) ∈ w : au = (anun) ∈ V for all u ∈ U }.
In particular,
U α = M (U, l1), U β = M (U, cs) and U γ = M (U, bs)
are called the α-, β- and γ- duals of U respectively [10].
Let A = (ank)n,k be an infinite matrix with real or complex entries ank. We write An as the sequence
of the n-th row of A, i.e., An = (ank)k for every n. For x = (xn) ∈ w, the A-transform of x is defined
as the sequence Ax = ((Ax)n), where
2
An(x) = (Ax)n =
ankxk,
∞
Xk=0
provided the series on the right side converges for each n. For any two sequence spaces U and V , we
denote by (U, V ), the class of all infinite matrices A that map U into V . Therefore A ∈ (U, V ) if and
only if Ax = ((Ax)n) ∈ V for all x ∈ U . In other words, A ∈ (U, V ) if and only if An ∈ U β for all n [17].
An infinite matrix T = (tnk)n,k is said to be triangle if tnk = 0 for k > n and tnn 6= 0, n ∈ N0.
3 Sequence spaces X(r, s, t; ∆) for X ∈ {l∞, c, c0}
In this section, we first begin with the notion of generalized means given by Mursaleen et al. [12].
We denote the sets U and U0 as
U =nu = (un)∞
n=0 ∈ w : un 6= 0 for all no and U0 =nu = (un)∞
n=0 ∈ w : u0 6= 0o.
Let r, t ∈ U and s ∈ U0. The sequence y = (yn) of generalized means of a sequence x = (xn) is defined
by
n
yn =
1
rn
sn−ktkxk
(n ∈ N0).
Xk=0
The infinite matrix A(r, s, t) of generalized means is defined by
(A(r, s, t))nk =( sn−ktk
rn
0
0 ≤ k ≤ n,
k > n.
Since A(r, s, t) is a triangle, it has a unique inverse and the inverse is also a triangle [7]. Take
0 = 1
s0
and
D(s)
D(s)
n = 1
sn+1
0
s1
s2
...
sn−1
sn
(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)
s0
s1
...
sn−2
sn−1
0
s0
...
sn−3
sn−2
0
0
0 · · ·
0 · · ·
...
sn−4 · · ·
sn−3 · · ·
s0
s1
(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)
for n = 1, 2, 3, · · ·
Then the inverse of A(r, s, t) is the triangle B = (bnk)n,k which is defined as
bnk =( (−1)n−k D(s)
0
n−k
tn
rk
0 ≤ k ≤ n,
k > n.
We now introduce the sequence spaces X(r, s, t; ∆) for X ∈ {l∞, c, c0} as
X(r, s, t; ∆) =nx = (xn) ∈ w :(cid:16) 1
rn
n
Xk=0
sn−ktk∆xk(cid:17)n
∈ Xo,
which is a combination of the generalized means and the difference operator ∆ such that ∆xk = xk−xk−1,
x−1 = 0. By using matrix domain, we can write X(r, s, t; ∆) = XA(r,s,t;∆) = {x ∈ w : A(r, s, t; ∆)x ∈ X},
where A(r, s, t; ∆) = A(r, s, t).∆, product of two triangles A(r, s, t) and ∆.
3
These sequence spaces include many known sequence spaces studied by several authors. For examples,
I. if rn = 1
un
, tn = vn and sn = 1 ∀ n, then the sequence spaces X(r, s, t; ∆) reduce to X(u, v; ∆) for
X ∈ {l∞, c, c0} introduced and studied by Polat et al. [14]
II. if tn = 1, sn = 1 ∀ n and rn = n + 1, then the sequence space X(r, s, t; ∆) for X = l∞ reduces to
ces∞(∆) studied by Orhan [13].
III. if rn = 1
n! , tn = αn
X ∈ {l∞, c, c0} reduce to eα
n! , sn = (1−α)n
∞(∆), eα
n!
, where 0 < α < 1, then the sequence spaces X(r, s, t; ∆) for
c (∆) and eα
0 (∆) respectively [15].
IV. if rn = n + 1, tn = 1 + αn, where 0 < α < 1 and sn = 1 ∀n, then the sequence spaces X(r, s, t; ∆)
0 (∆) studied by Aydin and Ba¸sar
for X ∈ {c, c0} reduce to the spaces of sequences aα
[4]. For X = l∞, the sequence space X(r, s, t; ∆) reduces to aα
∞(∆) studied by Djolovi´c [6].
c (∆) and aα
4 Main results
In this section, we begin with some topological results of the newly defined sequence spaces.
Theorem 4.1. The sequence spaces X(r, s, t; ∆) for X ∈ {l∞, c, c0} are complete normed linear spaces
under the norm defined by
= sup
(A(r, s, t; ∆)x)n.
n
Proof. Let u, v ∈ X(r, s, t; ∆) and α, β be any two scalars. Then
1
rn
kxkX(r,s,t;∆) = sup
n (cid:12)(cid:12)(cid:12)
sn−ktk∆(αuk + βvk)(cid:12)(cid:12)(cid:12)(cid:12)
n
sn−ktk∆xk(cid:12)(cid:12)(cid:12)
Xk=0
n (cid:12)(cid:12)(cid:12)(cid:12)
Xk=0
1
rn
n
1
rn
n
Xk=0
sup
n (cid:12)(cid:12)(cid:12)(cid:12)
≤ α sup
+ β sup
sn−ktk∆uk(cid:12)(cid:12)(cid:12)(cid:12)
1
rn
n
Xk=0
n (cid:12)(cid:12)(cid:12)(cid:12)
sn−ktk∆vk(cid:12)(cid:12)(cid:12)(cid:12)
and hence αu + βv ∈ X(r, s, t; ∆). Therefore X(r, s, t; ∆) is a linear space. It is easy to show that the
functional k.kX(r,s,t;∆) defined above gives a norm on the linear space X(r, s, t; ∆).
To show completeness, let (xm) be a Cauchy sequence in X(r, s, t; ∆), where xm = (xm
∈ X(r, s, t; ∆), for each m ∈ N0. Then for every ǫ > 0 there exists n0 ∈ N such that
k ) = (xm
0 , xm
1 , xm
2 , . . .)
The above implies that for each k ∈ N0,
kxm − xlkX(r,s,t;∆) < ǫ
for m, l ≥ n0.
A(r, s, t; ∆)(xm
k − xl
k) < ǫ for all m, l ≥ n0.
(4.1)
Therefore ((A(r, s, t).∆)xm
converges for each k. We write
k ) is a Cauchy sequence of scalars for each k ∈ N0 and hence ((A(r, s, t).∆)xm
k )
lim
m→∞
((A(r, s, t).∆)x(m)
k
) = ((A(r, s, t).∆)xk),
k ∈ N0.
Letting l → ∞ in (4.1), we obtain
A(r, s, t; ∆)(xm
< ǫ for all m ≥ n0 and each k ∈ N0.
(4.2)
Hence by definition kxm − xkX(r,s,t;∆) < ǫ for all m ≥ n0. Next we show that x ∈ X(r, s, t; ∆). Consider
(cid:12)(cid:12)(cid:12)
k − xk)(cid:12)(cid:12)(cid:12)
4
kxkX(r,s,t;∆) ≤ kxmkX(r,s,t;∆) + kxm − xkX(r,s,t;∆),
which is finite for m ≥ n0 and hence x ∈ X(r, s, t; ∆). This completes the proof.
Theorem 4.2. The sequence spaces X(r, s, t; ∆), where X ∈ {l∞, c, c0} are linearly isomorphic to the
spaces X ∈ {l∞, c, c0} respectively i.e. l∞(r, s, t; ∆) ∼= l∞, c(r, s, t; ∆) ∼= c and c0(r, s, t; ∆) ∼= c0.
Proof. We prove the theorem only for the case X = l∞. To prove this, we need to show that there exists
a bijective linear map from l∞(r, s, t; ∆) to l∞.
We define a map T : l∞(r, s, t; ∆) → l∞ by x 7−→ T x = y = (yn), where
yn =
1
rn
n
Xk=0
sn−ktk∆xk.
Since ∆ is a linear operator, so the linearity of T is trivial. It is clear from the definition that T x = 0
implies x = 0. Thus T is injective. To prove T is surjective, let y = (yn) ∈ l∞. Since y = (A(r, s, t).∆)x,
i.e.,
we can get a sequence x = (xn) as
x =(cid:0)A(r, s, t).∆(cid:1)−1
y = ∆−1.A(r, s, t)−1y,
xn =
Then
(−1)k D(s)
tk+j
k
n−j
Xk=0
rjyj,
n ∈ N0.
(4.3)
kxkl∞(r,s,t;∆) = sup
= sup
yn = kyk∞ < ∞.
n
n
Xj=0
n (cid:12)(cid:12)(cid:12)(cid:12)
1
rn
n
Xk=0
sn−ktk∆xk(cid:12)(cid:12)(cid:12)(cid:12)
Thus x ∈ l∞(r, s, t; ∆) and this shows that T is surjective. Hence T is a linear bijection from l∞(r, s, t; ∆)
to l∞. Also T is norm preserving. So l∞(r, s, t; ∆) ∼= l∞.
In the same way, we can prove that c0(r, s, t; ∆) ∼= c0, c(r, s, t; ∆) ∼= c. This completes the proof.
Since X(r, s, t; ∆) ∼= X for X ∈ {c0, c}, the Schauder bases of the sequence spaces X(r, s, t; ∆) are
the inverse image of the bases of X for X ∈ {c0, c}. So, we have the following theorem without proof.
Theorem 4.3. Let µk = (A(r, s, t; ∆)x)k for all k ∈ N0. Define the sequences b(j) = (b(j)
b(−1)
n
as
n ), j ∈ N0 and
(−1)k D(s)
tk+j
k
n−j
Xk=0
0
rj
if 0 ≤ j ≤ n
if j > n.
and b(−1)
n =
n
Xj=0
n−j
Xk=0
(−1)k D(s)
tk+j
k
rj.
b(j)
n =
Then the followings are true:
(i) The sequence (b(j))∞
unique representation of the form
j=0 is a basis for the space c0(r, s, t; ∆) and any sequence x ∈ c0(r, s, t; ∆) has a
x =
∞
Xj=0
µjb(j).
5
(ii) The sequence (b(j))∞
representation of the form
j=−1 is a basis for the space c(r, s, t; ∆) and any x ∈ c(r, s, t; ∆) has a unique
x = ℓb(−1)
n +
(µj − ℓ)b(j),
∞
Xj=0
where ℓ = lim
n→∞
(A(r, s, t; ∆)x)n.
Remark 4.1. In particular, if we choose rn = 1
un
X(r, s, t; ∆) reduce to X(u, v; ∆) for X ∈ {l∞, c, c0} [14]. With this choice of (sn), we have D(s)
D(s)
, tn = vn, sn = 1, ∀ n then the sequence spaces
0 =
n = 0 for n ≥ 2. Thus the sequences b(j) = (b(j)
n ), j ∈ N0 and b(−1)
1 = 1 and D(s)
reduce to
n
b(j)
n =
The sequences (b(j))∞
[14].
1
uj(cid:16) 1
1
vj
unvn
0
− 1
vj+1(cid:17) if
if
if
0 ≤ j < n
j = n
j > n.
and b(−1)
n =
n−1
Xj=0
1
uj(cid:16) 1
vj
−
1
vj+1(cid:17) +
1
unvn
.
j=0 and (b(j))∞
j=−1 are the bases for the spaces c0(u, v; ∆) and c(u, v; ∆) respectively
Let F be the collection of all nonempty finite subsets of the set of all natural numbers and A =
(ank)n,k be an infinite matrix satisfying the conditions:
∞
< ∞
ank < ∞
sup
K∈F
sup
n
ank(cid:12)(cid:12)(cid:12)
∞
Xn=0(cid:12)(cid:12)(cid:12) Xk∈K
Xk=0
lim
n
lim
n
lim
n
lim
n
lim
n
lim
n
ank = 0
∞
Xk=0
ank = 0 for all k
ank = 0
∞
Xk=0
ank exists for all k
ank − lim
n
ank = 0
ank exists
∞
∞
Xk=0
Xk=0
(4.4)
(4.5)
(4.6)
(4.7)
(4.8)
(4.9)
(4.10)
(4.11)
We now state some results of Stieglitz and Tietz [16] which are required to obtain the duals and to
characterize some matrix transformations.
Theorem 4.4. [16] (a) A ∈ (c0, l1), A ∈ (c, l1), A ∈ (l∞, l1) if and only if (4.4) holds.
(b) A ∈ (c0, l∞), A ∈ (c, l∞), A ∈ (l∞, l∞) if and only if (4.5) holds.
(c) A ∈ (c0, c0) if and only if (4.5) and (4.7) hold.
6
(d) A ∈ (l∞, c0) if and only if (4.6) holds.
(e) A ∈ (c, c0) if and only if (4.5), (4.7) and (4.8) hold.
(f ) A ∈ (c0, c) if and only if (4.5) and (4.9) hold.
(g) A ∈ (l∞, c) if and only if (4.5), (4.9) and (4.10) hold.
(h) A ∈ (c, c) if and only if (4.5), (4.9) and (4.11) hold.
4.1 The α-dual, γ-dual of X(r, s, t; ∆) for X ∈ {l∞, c, c0}
Theorem 4.5. The α-dual of the space X(r, s, t; ∆) for X ∈ {l∞, c, c0} is the set
Λ =na = (an) ∈ w : sup
K∈FXn (cid:12)(cid:12)(cid:12)Xj∈K
(−1)k D(s)
tk+j
k
n−j
Xk=0
< ∞o.
rj an(cid:12)(cid:12)(cid:12)
Proof. Let a = (an) ∈ w, x ∈ X(r, s, t; ∆) and y ∈ X for X ∈ {l∞, c, c0}. Then for each n ∈ N0, we have
anxn =
n
Xj=0
n−j
Xk=0
(−1)k D(s)
tk+j
k
rjanyj = (Cy)n,
where the matrix C = (cnj) is defined as
(−1)k D(s)
tk+j
k
n−j
Xk=0
0
rjan
if 0 ≤ j ≤ n
if j > n
cnj =
and xn is given by (4.3). Thus for each x ∈ X(r, s, t; ∆), (anxn)n ∈ l1 if and only if (Cy)n ∈ l1 where
y ∈ X for X ∈ {l∞, c, c0}. Therefore a = (an) ∈ [X(r, s, t; ∆)]α if and only if C ∈ (X, l1). By using
Theorem 4.4(a) , we have
[X(r, s, t; ∆)]α = Λ.
Theorem 4.6. The γ-dual of X(r, s, t; ∆) for X ∈ {l∞, c, c0} is the set
Γ =na = (an) ∈ w :
sup
m
∞
Xn=0
emn < ∞o,
where the matrix E = (emn) is defined by
E = (emn) =
Note: We mean
+(cid:16) D(s)
0
tn
−
D(s)
1
tn+1(cid:17)
m
Xj=n+1
aj +
m
Xj=n+2
(−1)j−n
D(s)
j−n
tj (cid:16)
m
Xk=j
ak(cid:17)(cid:21)
0 ≤ n ≤ m,
n > m.
rn(cid:20) an
s0tn
0
m
= 0 if n > m.
Pj=n
Proof. Let a = (an) ∈ w, x ∈ X(r, s, t; ∆) and y ∈ X, where X ∈ {l∞, c, c0}. Then by using (4.3), we
7
have
m
Xn=0
anxn =
k rjyjan
tk+j
n
n
m
n
n−j
(−1)k D(s)
(−1)k D(s)
=
n−j
n−1
m−1
Xk=0
Xj=0
Xn=0
Xk=0
Xj=0
Xn=0
=(cid:20) m−1
(−1)k D(s)
Xk=0
Xn=0
k
tk
+(cid:20) m−1
Xn=0
Xk=0
a0 +(cid:16) D(s)
=(cid:20) D(s)
+(cid:20) D(s)
a1 +(cid:16) D(s)
+(cid:16) D(s)
rn(cid:20) an
Xn=0
s0tn
0
tn
0
t1
0
t1
0
t0
0
t0
−
=
(−1)k D(s)
tk+1
−
m
k
(−1)k D(s)
tk+1
k
k rj yjan
tk+j
m
an +
m
m−j
+
Xk=0
Xj=0
(−1)k D(s)
k
tk
D(s)
1
m
m−1
aj +
an +
Xk=0
Xk=0
t1 (cid:17)
Xj=1
t2 (cid:17)
Xj=2
tn+1(cid:17)
Xj=n+1
D(s)
1
D(s)
1
aj +
m
m
−
m
Xj=2
Xj=3
m
(−1)k D(s)
k rj yjam
tk+j
am(cid:21)r0y0
am(cid:21)r1y1 + · · · +
D(s)
m
(−1)j
D(s)
0
tm
rmymam
j
tj (cid:16)
m
D(s)
j−1
ak(cid:17)(cid:21)r0y0
Xk=j
ak(cid:17)(cid:21)r1y1 + · · · +
tj (cid:16)
Xk=j
(−1)j−n D(s)
ak(cid:17)(cid:21)yn
tj (cid:16)
Xk=j
j−n
m
(−1)j−1
aj +
m
Xj=n+2
rmam
tm
D(s)
0 ym
m
Xn=0
anxn(cid:17)m
∈
= (Ey)m,
where E = (emn) is the matrix defined above.
Thus a ∈(cid:2)X(r, s, t; ∆)(cid:3)γ
if and only if ax = (anxn) ∈ bs for x ∈ X(r, s, t; ∆) if and only if(cid:16)
l∞, i.e. (Ey)m ∈ l∞, for y ∈ X. Hence by Theorem 4.4(b), we have
(cid:2)X(r, s, t; ∆)(cid:3)γ
= Γ.
4.2
β-dual and Matrix transformations
We now first discuss about the β-dual and then characterize some matrix transformations. Let T be a
triangle and XT be the matrix domain of T .
Theorem 4.7. ([7], Theorem 2.6) Let X be a BK space with AK property and R = St, the transpose
of S, where S = (sjk) is the inverse of the matrix T . Then a ∈ (XT )β if and only if a ∈ (X β)R and
W ∈ (X, c0), where the triangle W is defined by wmk =
ajsjk. Moreover if a ∈ (XT )β, then
∞
Pj=m
akzk =
∞
Xk=0
∞
Xk=0
Rk(a)Tk(z)
∀ z ∈ XT .
Remark 4.2. ([7], Remark 2.7) The conclusion of the Theorem 4.7 is also true for X = l∞.
Remark 4.3. ([10], [7]) We have a ∈ (cT )β if and only if Ra ∈ l1 and W ∈ (c, c). Moreover, if
8
a ∈ (cT )β then we have for all z ∈ cT
∞
akzk =
Rk(a)Tk(z) − ηγ,
∞
Xk=0
Xk=0
Xk=0
m
where η = lim
k→∞
Tk(z) and γ = lim
m→∞
wmk.
To find β-duals of the sequence spaces X(r, s, t; ∆) for X ∈ {l∞, c, c0}, we list the following sets:
∞
∞
m
m
lim
m→∞
lim
m→∞
Xk=0
Rk(a) < ∞o
B1 =na ∈ w :
wmk = 0 for all ko
B2 =na ∈ w :
B3 =na ∈ w : sup
wmk < ∞o
Xk=0
wmk = 0o
B4 =na ∈ w :
Xk=0
exists for all ko
B5 =na ∈ w :
wmk exists o,
B6 =na ∈ w :
Xk=0
(−1)l D(s)
− D(s)
tk+1(cid:17)
Pl=2
Pj=k+1
Pj=k+l
aj(cid:21).
(−1)l D(s)
Pl=m−k+1
Pj=k+l
lim
m→∞
lim
m→∞
wmk
tl+k
m
∞
∞
∞
l
l
l
1
∞
∞
∞
0
tk
s0tk
aj +
+(cid:16) D(s)
Pj=m
where Rk(a) = rk(cid:20) ak
wmk = rk(cid:20) m−k
(−1)l D(s)
Pl=0
Theorem 4.8. We have [c0(r, s, t; ∆)]β = B1T B2T B3, [l∞(r, s, t; ∆)]β = B1T B4 and [c(r, s, t; ∆)]β =
B1T B3T B5T B6.
aj(cid:21), R(a) = (Rk(a))k and
Proof. Here the matrix T = A(r, s, t).∆ = (t′
aj +
tl+k
tl+k
nk), where
So, T −1 = (A(r, s, t).∆)−1 = ∆−1.A(r, s, t)−1. Let S = (sjk) be the inverse of T . Then we easily get
t′
1
rn
s0tn
rn
0
nk =
sjk =
[sn−ktk − sn−k+1tk+1]
if 0 ≤ k < n
if k = n
if k > n.
(−1)l D(s)
l
tl+k
j−k
Pl=0
0
rk
if
if
0 ≤ k ≤ j
k > j.
9
To compute β-duals, we first determine W = (wmk) and R(a) = (Rk(a)), where R = St.
∞
ajsjk
rkak +
Rk(a) =
=
=
Xj=k
D(s)
0
tk
D(s)
0
tk
(−1)l D(s)
tl+k
l
rkaj
∞
j−k
1
Xl=0
Xj=k+1
(−1)l D(s)
Xl=0
tl+k
+(cid:16) D(s)
tk+1(cid:17)
D(s)
1
0
tk
−
l
rkak +
rkak+1 +
rkak+2 + · · ·
2
l
(−1)l D(s)
Xl=0
tl+k
(−1)l D(s)
Xl=2
tl+k
∞
l
∞
Xj=k+1
aj +
∞
Xj=k+l
aj(cid:21)
and
= rk(cid:20) ak
s0tk
∞
wmk =
ajsjk
∞
=
j−k
Xj=m
Xl=0
Xj=m
= rk(cid:20) m−k
Xl=0
l
(−1)l D(s)
tl+k
(−1)l D(s)
tl+k
l
rkaj
∞
Xj=m
aj +
∞
Xl=m−k+1
(−1)l D(s)
tl+k
l
∞
Xj=k+l
aj(cid:21).
Using Theorem 4.7 and Remark 4.2 & 4.3, we have [c0(r, s, t; ∆)]β = B1T B2T B3, [l∞(r, s, t; ∆)]β =
B1T B4 and [c(r, s, t; ∆)]β = B1T B3T B5T B6.
Theorem 4.9. ([7], Theorem 2.13) Let X be a BK space with AK property, Y be an arbitrary subset of
w and R = St. Then A ∈ (XT , Y ) if and only if BA ∈ (X, Y ) and W An ∈ (X, c0) for all n = 0, 1, 2, · · · ,
where BA is the matrix with rows BA
n = R(An), An are the rows of A and the triangles W An are defined
by
∞
anjsjk
Pj=m
0
:
:
0 ≤ k ≤ m
k > m.
wAn
mk =
Theorem 4.10. ([7]) Let Y be any linear subspace of w. Then A ∈ (cT , Y ) if and only if Rk(An) ∈
m
(c0, Y ) and W An ∈ (c, c) for all n and Rk(An)e − (γn) ∈ Y , where γn = lim
m→∞
and e = (1, 1, 1, · · · ).
Moreover, if A ∈ (cT , Y ) then we have
wAn
mk for n = 0, 1, 2 · · ·
Xk=0
Az = Rk(An)(T (z)) − η(γn)
for all z ∈ cT , where η = lim
k→∞
Tk(z).
To characterize the matrix transformations A ∈ (X(r, s, t; ∆), Y ) for X, Y ∈ {l∞, c, c0}, we consider
the following conditions:
10
∞
Xk=0
sup
n
lim
n→∞
Rk(An) < ∞
Rk(An) = 0
for all k
wAn
mk < ∞ for all n
wAn
mk = 0 for all n, k
Rk(An) exists for all k
∞
Rk(An) = 0
wAn
mk = 0
for all n
sup
m
m
Xk=0
lim
m→∞
lim
n→∞
lim
n→∞
lim
m→∞
∞
m
Xk=0
Xk=0
Xk=0(cid:12)(cid:12)(cid:12)
Xk=0
wAn
= 0
Rk(An)(cid:12)(cid:12)(cid:12)
lim
n→∞
Rk(An) − lim
n→∞
lim
m→∞
lim
m→∞
mk exists for all k, n
m
wAn
mk exists for all n
Rk(An)e − (γn) ∈ c0
for all γn, n = 0, 1, 2, · · ·
Rk(An)e − (γn) ∈ l∞ for all γn, n = 0, 1, 2, · · ·
Rk(An)e − (γn) ∈ c
for all γn, n = 0, 1, 2, · · · ,
m
(4.12)
(4.13)
(4.14)
(4.15)
(4.16)
(4.17)
(4.18)
(4.19)
(4.20)
(4.21)
(4.22)
(4.23)
(4.24)
where γn = lim
m→∞
s0tk
Rk(An) = rkh ank
mk = rk(cid:20) m−k
Pl=0
wAn
0
tk
wAn
mk,
Xk=0
+(cid:16) D(s)
Pj=m
tl+k
∞
(−1)l D(s)
l
1
− D(s)
tk+1(cid:17)
anj +
anj +
∞
l
(−1)l D(s)
Pl=2
(−1)l D(s)
tl+k
∞
l
tl+k
∞
∞
Pj=k+1
Pl=m−k+1
Pj=k+l
anji and
∞
Pj=k+l
anj(cid:21).
Theorem 4.11. (a) A ∈ (c0(r, s, t; ∆), c0) if and only if the conditions (4.12), (4.13), (4.14) and (4.15)
hold.
(b) A ∈ (c0(r, s, t; ∆), c) if and only if the conditions (4.12), (4.14), (4.15) and (4.16) hold.
(c) A ∈ (c0(r, s, t; ∆), l∞) if and only if the conditions (4.12), (4.14) and (4.15) hold.
Proof. We prove only the part (a) of this theorem. The other parts follow in a similar way. For this,
we first compute the matrices BA = Rk(An) and W An = (wAn
mk) for n = 0, 1, 2, · · · of Theorem 4.9 to
determine the conditions BA ∈ (c0, c0) and W An ∈ (c0, c0). Using the same lines of proof as used in
11
Theorem 4.8, we have
∞
Rk(An) =
∞
=
j−k
sjkanj
Xj=k
(−1)l D(s)
Xj=k+1
Xl=0
tl+k
+(cid:16) D(s)
= rkh ank
s0tk
0
tk
−
l
rkanj +
D(s)
0
tk
rkank
D(s)
1
tk+1(cid:17)
∞
Xj=k+1
anj +
(−1)l D(s)
tl+k
l
∞
Xl=2
∞
Xj=k+l
anji
and
∞
wAn
mk =
anjsjk
Xj=m
= rk(cid:20) m−k
Xl=0
(−1)l D(s)
tl+k
l
∞
Xj=m
anj +
∞
Xl=m−k+1
(−1)l D(s)
tl+k
l
∞
Xj=k+l
anj(cid:21).
Using Theorem 4.9, we have A ∈ (c0(r, s, t; ∆), c0) if and only if the conditions (4.12), (4.13), (4.14) and
(4.15) hold.
We can also obtain the following results.
Corollary 4.1. (a) A ∈ (l∞(r, s, t; ∆), c0) if and only if the conditions (4.17) and (4.18) hold.
(b) A ∈ (l∞(r, s, t; ∆), c) if and only if the conditions (4.12), (4.16), (4.18) and (4.19) hold.
(c) A ∈ (l∞(r, s, t; ∆), l∞) if and only if the conditions (4.12) and (4.18) hold.
Corollary 4.2. (a) A ∈ (c(r, s, t; ∆), c0) if and only if the conditions (4.12), (4.13), (4.14), (4.20), (4.21)
and (4.22) hold.
(b) A ∈ (c(r, s, t; ∆), c) if and only if the conditions (4.12), (4.14), (4.16), (4.20), (4.21) and (4.24) hold.
(c) A ∈ (c(r, s, t; ∆), l∞) if and only if the conditions (4.12), (4.14), (4.20), (4.21) and (4.23) hold.
5 Some applications
In this section, we justify our results in some special cases. Also, we illustrate the results related to
matrix transformations given by Djolovic[6], Polat and Altay [15], Aydin and Ba¸sar [4] etc.
(I) In particular, if we choose rn = 1
spaces X(r, s, t; ∆) for X ∈ {l∞, c0, c} reduce to the Euler difference sequence spaces eα
0 = 1, D(s)
and eα
D(s)
difference sequence spaces is the set
, where 0 < α < 1 then the sequence
∞(∆), eα
0 (∆)
1 = (1 − α),
, k ∈ N0. Thus the α-dual of the Euler
c (∆) respectively [15]. By the above choice of r, s and t, we have D(s)
and so on. Therefore D(s)
n! and sn = (1−α)n
2 = (1−α)2
3 = (1−α)3
k = (1−α)k
n! , tn = αn
, D(s)
n!
k!
3!
2!
n
K∈FXn (cid:12)(cid:12)(cid:12)(cid:12)Xj∈K
na ∈ w : sup
Xk=j
K∈FXn (cid:12)(cid:12)(cid:12)(cid:12)Xj∈K
=na ∈ w : sup
(−1)k−j (1 − α)k−j
(k − j)!
n
Xk=j
(−1)k−j(cid:18)k
.
1
j!
α−kk!an(cid:12)(cid:12)(cid:12)(cid:12)
< ∞o
j(cid:19)(1 − α)k−j α−kan(cid:12)(cid:12)(cid:12)(cid:12)
< ∞o.
12
Here we illustrate that how the characterization of matrix transformation A ∈ (eα
tained with the help of Corollary 4.1(c)
∞(∆), l∞) can be ob-
∞
Rk(An) =
anjsjk
Xj=k
+(cid:16) D(s)
= rk(cid:20) ank
+(cid:16) 1
αk −
ank
αk
s0tk
0
tk
=
−
1
αk+1(cid:17)
∞
∞
Xj=k+1
anj +
D(s)
1
tk+1(cid:17)
Xj=k+1
∞
anj +
l
∞
(−1)l D(s)
tk+l
Xl=2
(−1)l(cid:18)l + k
Xj=k+l
l (cid:19) (1 − α)l
αk+l
anj(cid:21)
Xj=k+l
∞
∞
Xl=2
anj.
∞
wAn
mk =
l
anjsjk
(−1)l D(s)
tk+l
Xj=m
= rk(cid:20) m−k
Xl=0
k!(cid:20) m−k
Xl=0
=(cid:20) m−k
(−1)l(cid:18)l + k
Xl=0
=
1
(−1)l (1 − α)l(l + k)!
l!αl+k
anj +
l (cid:19) (1 − α)l
αl+k
anj +
(−1)l(cid:18)l + k
l (cid:19))
(1 − α)l
αl+k
(−1)l D(s)
tk+l
l
∞
anj(cid:21)
Xj=k+l
(−1)l (1 − α)l(l + k)!
l!αl+k
anj
∞
Xj=m
∞
∞
Xl=m−k+1
Xj=m
Xj=m
∞
∞
Xm−k+1
Xm−k+1
∞
∞
Xj=k+l
Xj=k+l
∞
anj(cid:21)
anj(cid:21).
So A ∈ (eα
∞(∆), l∞) if and only if
and
sup
n
∞
Xk=0
Rk(An) < ∞
lim
m→∞
m
Xk=0
wAn
mk = 0 for all n.
(II) We choose sn = 1 ∀ n, rn = (n + 1), tn = 1 + αn, where 0 < α < 1 then the sequence
0 (∆) respectively. With this choice
mk) become
spaces X(r, s, t; ∆) for X ∈ {l∞, c, c0} reduce to aα
D(s)
k = 0 for all k ≥ 2. Therefore the matrices Rk(An) and W An = (wAn
0 = 1 = D(s)
c (∆) and aα
and D(s)
∞(∆), aα
1
Rk(An) = (k + 1)(cid:20) ank
1 + αk +(cid:16)
1
1 + αk −
1
1 + αk+1(cid:17)
∞
Xj=k+1
anj(cid:21),
and
wAn
0
tk
mk = rk(cid:20)(cid:16) D(s)
= (k + 1)(cid:20)(cid:16)
D(s)
1
−
tk+1(cid:17)
1 + αk −
1
∞
anj −
Xj=m
1 + αk+1(cid:17)
1
D(s)
1
tm+1
∞
Xj=m
∞
Xj=m+1
anj(cid:21)
anj −
1
1 + αm+1
∞
Xj=m+1
anj(cid:21).
13
By evaluating, we have
wAn
mk =
m
Xk=0
m
(k + 1)(cid:16)
Xk=0
1
1 + αk −
Therefore by Corollary 4.1(a), we have A ∈ (aα
∞(∆), c0) if and only if
1
1 + αk+1(cid:17)(cid:12)(cid:12)(cid:12)
∞
Xj=m
+ anm
m + 1
1 + αm+1 .
anj(cid:12)(cid:12)(cid:12)
lim
n→∞
∞
Xk=0
Rk(An) = 0
lim
m→∞
m
Xk=0
wAn
mk = 0 for all n.
and
References
[1] Z. U. Ahmad, M. Mursaleen, Kothe-Toeplitz duals of some new sequence spaces and their matrix
maps, Publ. Inst. Math.(Beograd), 42(56)(1987), 57-61.
[2] B. Altay, F. Ba¸sar, Generalization of the sequence space ℓ(p) derived by weighted mean, J. Math.
Anal. Appl., 330(2007), 174-185.
[3] B. Altay , F. Ba¸sar, The fine spectrum and matrix domain of the difference operator ∆ on the
sequece space ℓp, (0 < p < 1), Commun. Math. Anal., 2 (2)(2007), 1-11.
[4] C. Aydin, F. Ba¸sar, Some new difference sequence spaces, Appl. Math. Comput., 157(3)(2004),
677-693.
[5] R. C¸ olak, M. Et, On some generalized difference sequence spaces and related matrix transformations,
Hokkaido Math. J., 26(3)(1997), 483-492.
[6] I. Djolovi´c , On the space of bounded Euler difference sequences and some classes of compact
operators, Appl. Math. Comput., 182(2) (2006),1803-1811.
[7] A. M. Jarrah, E. Malkowsky, Ordinary, absolute and strong summability and matrix transforma-
tions, Filomat, 17(2003), 59-78.
[8] H. Kizmaz, On certain sequence spaces, Canad. Math. Bull., 24(2)(1981), 169-176.
[9] I. J. Maddox, Elements of Functioal Analysis, 2nd Edition, The University Press, Cambridge, 1988.
[10] E. Malkowsky, V. Rakocevi´c, On matrix domains of triangles, Appl. Math. Comput., 189(2)(2007),
1146-1163.
[11] E. Malkowsky, E. Sava¸s, Matrix transformations between sequence spaces of generalized weighted
means, Appl. Math. Comput., 147(2004), 333-345.
[12] M. Mursaleen, A. K. Noman, On generalized means and some related sequence spaces, Comput.
Math. Appl., 61(4)(2011), 988-999.
[13] C. Orhan, Matrix transformations on Cesro difference sequence spaces, Comm. Fac. Sci. Univ.
Ankara Sr. A1 Math., 33(1)(1984), 1-8.
14
[14] H. Polat, V. Karakaya, N. S¸im¸sek, Difference sequence spaces derived by using a generalized
weighted mean, Appl. Math. Lett., 24(5)(2011), 608-614.
[15] H. Polat, B. Atay, On some new Euler difference sequence spaces, Southeast Asian Bulletin Math,
30(2006), 209-220.
[16] M. Stieglitz, H. Tietz, Matrix trasformationnen von Folenraumen Eine Erebisubersicht, Mathema-
tische Zeitschrift(Math. Z.), 154(1977), 1-16.
[17] A. Wilansky, Summability through Functional Analysis, North-Holland Math. Stud., vol. 85, Else-
vier Science Publishers, Amsterdam, New York, Oxford, 1984.
15
|
1912.06846 | 1 | 1912 | 2019-12-14T13:45:55 | A class of sectorial relations and the associated closed forms | [
"math.FA"
] | Let $T$ be a closed linear relation from a Hilbert space ${\mathfrak H}$ to a Hilbert space ${\mathfrak K}$ and let $B \in \mathbf{B}({\mathfrak K})$ be selfadjoint. It will be shown that the relation $T^{*}(I+iB)T$ is maximal sectorial via a matrix decomposition of $B$ with respect to the orthogonal decomposition ${\mathfrak H}={\rm d\overline{om}\,} T^* \oplus {\rm mul\,} T$. This leads to an explicit expression of the corresponding closed sectorial form. These results include the case where ${\rm mul\,} T$ is invariant under $B$. The more general description makes it possible to give an expression for the extremal maximal sectorial extensions of the sum of sectorial relations. In particular, one can characterize when the form sum extension is extremal. | math.FA | math |
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED
CLOSED FORMS
S. HASSI AND H.S.V. DE SNOO
Dedicated to V.E. Katsnelson on the occasion of his 75th birthday
Abstract. Let T be a closed linear relation from a Hilbert space H to a
Hilbert space K and let B ∈ B(K) be selfadjoint. It will be shown that the
relation T ∗(I + iB)T is maximal sectorial via a matrix decomposition of B
with respect to the orthogonal decomposition H = dom T ∗ ⊕ mul T . This leads
to an explicit expression of the corresponding closed sectorial form. These
results include the case where mul T is invariant under B. The more general
description makes it possible to give an expression for the extremal maximal
sectorial extensions of the sum of sectorial relations. In particular, one can
characterize when the form sum extension is extremal.
1. Introduction
A linear relation H in a Hilbert space H is said to be accretive if Re (h′, h) ≥ 0,
{h, h′} ∈ H. Note that the closure of an accretive relation is also accretive. An
accretive relation H in H is said to be maximal accretive if the existence of an
accretive relation H ′ in H with H ⊂ H ′ implies H ′ = H. A maximal accretive
relation is automatically closed. In a similar way, a linear relation H in a Hilbert
space H is said to be sectorial with vertex at the origin and semi-angle α, α ∈
[0, π/2), if
(1.1)
Im (h′, h) ≤ (tan α) Re (h′, h),
{h, h′} ∈ H.
The closure of a sectorial relation is also sectorial. A sectorial relation H in a Hilbert
space H is said to be maximal sectorial if the existence of a sectorial relation H ′
in H with H ⊂ H ′ implies H ′ = H. A maximal sectorial relation is automatically
closed. Note that a sectorial relation is maximal sectorial if and only if it is maximal
as an accretive relation; see [6].
A sesquilinear form t = t[·, ·] in a Hilbert space H is a mapping from dom t ⊂ H
to C which is linear in its first entry and antilinear in its second entry. The adjoint
t∗ is defined by t∗[h, k] = t[k, h], h, k ∈ dom t; for the diagonal of t the notation t[·]
will be used. A (sesquilinear) form is said to be sectorial with vertex at the origin
and semi-angle α, α ∈ [0, π/2), if
(1.2)
ti[h] ≤ (tan α) tr[h],
h ∈ dom t,
2010 Mathematics Subject Classification. Primary 47B44; Secondary 47A06, 47A07, 47B65.
Key words and phrases. Sectorial relation, Friedrichs extension, Kreın extension, extremal
extension, form sum.
The second author thanks the University of Vaasa for its hospitality during the preparation of
this work.
1
2
S. HASSI AND H.S.V. DE SNOO
where the real part tr and the imaginary part ti are defined by
(1.3)
tr =
t + t∗
2
,
ti =
t − t∗
2i
,
dom tr = dom ti = dom t.
A sesquilinear form will be called a form in the rest of this note. Observe that the
form tr is nonnegative and that the form ti is symmetric, while t = tr + i ti. A
sectorial form t is said to be closed if
hn → h,
t[hn − hm] → 0 ⇒ h ∈ dom t and t[hn − h] → 0.
A sectorial form t is closed if and only if its real part tr is closed; see [7].
The connection between maximal sectorial relations and closed sectorial forms
is given in the so-called first representation theorem; cf. [1], [4], [7], [8].
Theorem 1.1. Let t be a closed sectorial form in a Hilbert space H with vertex
at the origin and semi-angle α, α ∈ [0, π/2). Then there exists a unique maximal
sectorial relation H in H with vertex at the origin and semi-angle α in H such that
(1.4)
dom H ⊂ dom t,
and for every {h, h′} ∈ H and k ∈ dom t one has
t[h, k] = (h′, k).
(1.5)
Conversely, for every maximal sectorial relation H with vertex at the origin and
semi-angle α, α ∈ [0, π/2), there exists a unique closed sectorial form t such that
(1.4) and (1.5) are satisfied.
This result contains as a special case the connection between nonnegative selfad-
joint relations and closed nonnegative forms. The nonnegative selfadjoint relation
Hr corresponding to the real part tr of the form t is called the real part of H; this
notion should not to be confused with the real part introduced in [6].
In the theory of sectorial operators one encounters expressions T ∗(I +iB)T where
T is a linear operator from a Hilbert space H to a Hilbert space K and B ∈ B(K)
is a selfadjoint operator. In the context of sectorial relations the operator T may
be replaced by a linear relation T . A frequently used observation is that when T is
a closed linear relation and the multivalued part mul T is invariant under B, then
the product is a maximal sectorial relation; cf. [4]. However, in fact, the relation
(1.6)
T ∗(I + iB)T
is maximal sectorial for any closed linear relation T . This will be shown in this note
via a matrix decomposition of B with respect to the orthogonal decomposition H =
dom T ∗ ⊕mul T . In addition the closed sectorial form corresponding to T ∗(I +iB)T
will be determined. The main argument consists of a reduction to the case where
T is an operator. For the convenience of the reader the arguments in the operator
case are included. Note that if T is not closed, then T ∗(I + iB)T is a sectorial
relation which may have maximal sectorial extensions, such as T ∗(I + iB)T ∗∗ and
some of these extensions have been determined in [5]; cf. [10].
It is clear that the sum of two sectorial relations is a sectorial relation and
there will be maximal sectorial extensions. In [5] the Friedrichs extension has been
determined in general, while the Kreın extension has been determined only under
additional conditions. As an application of the above results for the relation in
(1.6) the Kreın extension and, in fact, all extremal maximal sectorial extensions
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
3
of the sum of two sectorial relations will be characterized in general. With this
characterization one can determine when the form sum extension is extremal.
2. A preliminary result
The first case to be considered is the linear relation T ∗(I + iB)T , where T a
closed linear operator, which is not necessarily densely defined, and B ∈ B(K) is
selfadjoint.
In this case one can write down a natural closed sectorial form and
verify that T ∗(I + iB)T is the maximal sectorial relation corresponding to the form
via Theorem 1.1.
Theorem 2.1. Let T be a closed linear operator from a Hilbert space H to a Hilbert
space K and let the operator B ∈ B(K) be selfadjoint. Then the form t in H defined
by
(2.1)
t[h, k] = ((I + iB)T h, T k),
h, k ∈ dom t = dom T,
is closed and sectorial with vertex at the origin and semi-angle α ≤ arctan kBk and
the maximal sectorial relation H corresponding to the form t is given by
(2.2)
H = T ∗(I + iB)T,
with mul H = mul T ∗ = (dom T )⊥. A subset of dom t = dom T is a core of the form
t if and only if it is a core of the operator T . Moreover, the nonnegative selfadjoint
relation Hr corresponding to the real part (tH )r of the form t is given by
(2.3)
Hr = T ∗T.
Proof. It is straightforward to check that t in (2.1) is a closed sectorial form as
indicated, since
tr[h, k] = (T h, T k),
ti[h, k] = (BT h, T k).
Therefore, ti[h] = (BT h, T h) ≤ kBkkT hk2 = kBktr[h], so that t is closed and
sectorial with vertex at the origin and semi-angle α ≤ arctan kBk. Moreover, since
T is closed, it is clear that tr and hence t is closed.
Now let {h, h′} ∈ T ∗(I + iB)T , then there exists ϕ ∈ K such that
{h, ϕ} ∈ T,
{(I + iB)ϕ, h′} ∈ T ∗,
from which it follows that
Consequently, one sees that
(h′, h) = (ϕ, ϕ) + i(Bϕ, ϕ).
Im (h′h) = (Bϕ, ϕ) ≤ kBk kϕk2 = kBk Re (h′, h),
which implies that T ∗(I + iB)T is a sectorial relation with vertex at the origin and
semi-angle α ≤ arctan kBk. Furthermore, observe that the above calculation also
shows that mul T ∗(I + iB)T = mul T ∗.
To see that T ∗(I +iB)T is closed, let {hn, h′
n} ∈ T ∗(I +iB)T converge to {h, h′}.
Then there exist ϕn ∈ K such that
{hn, ϕn} ∈ T,
n, hn) = kϕnk2 shows that (ϕn) is a Cauchy sequence in K,
{(I + iB)ϕn, h′
n} ∈ T ∗,
and the identity Re (h′
so that ϕn → ϕ with ϕ ∈ K. Thus
{hn, ϕn} → {h, ϕ},
{(I + iB)ϕn, h′
n} → {(I + iB)ϕ, h′}.
4
S. HASSI AND H.S.V. DE SNOO
Since T and T ∗ are closed, one concludes that {h, ϕ} ∈ T and {(I + iB)ϕ, h′} ∈ T ∗,
which implies that {h, h′} ∈ T ∗(I + iB)T . Hence T ∗(I + iB)T is closed.
Now let H be the maximal sectorial relation corresponding to t in (2.1). Assume
that {h, h′} ∈ H, then for all k ∈ dom t = dom T
t[h, k] = (h′, k) or
((I + iB)T h, T k) = (h′, k),
which implies that
{(I + iB)T h, h′} ∈ T ∗
or {h, h′} ∈ T ∗(I + iB)T.
Consequently, it follows that H ⊂ T ∗(I + iB)T . Since T ∗(I + iB)T is sectorial and
H is maximal sectorial, it follows that H = T ∗(I + iB)T . In particular, one sees
that the closed relation T ∗(I + iB)T is maximal sectorial.
(cid:3)
With the closed linear operator T from H to K and the selfadjoint operator
B ∈ B(K), consider the following matrix decomposition of B
(2.4)
B∗
B =(cid:18)Baa Bab
ba Bbb(cid:19) : (cid:18)ker T ∗
ran T (cid:19) →(cid:18)ker T ∗
ran T (cid:19) .
Then it is clear that
(2.5)
t[h, k] = ((I + iB)T h, T k) = ((I + iBbb)T h, T k), h, h ∈ dom t = dom T,
which shows that only the compression of B to ran T plays a role in (2.1).
In
applications involving Theorem 2.1, it is therefore useful to recall the following
corollary.
Corollary 2.2. Let T ′ be a closed linear operator from the Hilbert space H to a
Hilbert space K′ and let the operator B′ ∈ B(K′) be selfadjoint. Assume that the
form t in Theorem 2.1 is also given by
t[h, k] = ((I + iB′)T ′h, T ′k),
h, k ∈ dom t = dom T ′.
Then there is a unitary mapping U from ran T onto ran T ′, such that
T ′ = U T, B′
bb = U BbbU ∗,
where Bbb and B′
respectively.
bb stand for the compressions of B and B′ to ran T and ran T ′,
Proof. By assumption ((I + iB′)T ′h, T ′k) = ((I + iB)T h, T k) for all h, k ∈ dom t.
This leads to
(T ′h, T ′k) = (T h, T k) and (B′T ′h, T ′k) = (BT h, T k)
for all h, k ∈ dom t. Hence the mapping T h 7→ T ′h is unitary, and denote it by U .
Then T ′ = U T and it follows that (B′T ′h, T ′k) = (BU ∗T ′h, U ∗T ′k).
(cid:3)
3. A matrix decomposition for T ∗(I + iB)T
Let T be a linear relation from H to K which is closed; observe that then the
subspace mul T is closed. The adjoint T ∗ of T is the set of all {h, h′} ∈ K × H for
which
(h′, f ) = (h, f ′)
for all {f, f ′} ∈ T.
Hence, the definition of T ∗ depends on the Hilbert spaces H and K in which T is
assumed to act. Let K have the orthogonal decomposition
(3.1)
K = dom T ∗ ⊕ mul T,
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
5
and let P be the orthogonal projection onto dom T ∗. Observe that P T ⊂ T , since
{0} × mul T ⊂ T . Therefore T ∗ ⊂ (P T )∗ = T ∗P , where the last equality holds
since P ∈ B(K). Then one has
(3.2)
(P T )∗ = T ∗ b⊕ (mul T × {0}).
The orthogonal operator part Ts of T is defined as Ts = P T . Hence Ts is an
operator from the Hilbert space H to the Hilbert space K and Ts ⊂ T . Note that
ran Ts ⊂ dom T ∗ = K ⊖ mul T . Thus one may interpret Ts as an operator from the
Hilbert space H to the Hilbert space dom T ∗ and one may also consider the adjoint
(Ts)× of Ts with respect to these spaces. It is not difficult to see the connection
between these adjoints: if {h, h′} ∈ K × H, then
(3.3)
{h, h′} ∈ T ∗ ⇔ {h, h′} ∈ (Ts)×.
The identity (3.2) shows the difference between (Ts)∗ and (Ts)×.
Let T be a closed linear relation from a Hilbert space H to a Hilbert space K and
let B ∈ B(K) be selfadjoint. In order to study the linear relation
T ∗(I + iB)T,
decompose the Hilbert space K as in (3.1) and decompose the selfadjoint operator
B ∈ B(K) accordingly:
(3.4)
B =(cid:18)B11 B12
12 B22(cid:19) : (cid:18)dom T ∗
mul T (cid:19) →(cid:18)dom T ∗
mul T (cid:19) .
B∗
Here the operators B11 ∈ B(dom T ∗) and B22 ∈ B(mul T ) are selfadjoint, while
B12 ∈ B(mul T, dom T ∗) and B∗
12 ∈ B(dom T ∗, mul T ).
By means of the decomposition (3.4) the following auxiliary operators will be
introduced. First, define the operator C0 ∈ B(dom T ∗) by
(3.5)
C0 = I + B12(I + B2
22)−1B∗
12.
Observe that C0 ≥ I and that (C0)−1 belongs to B(dom T ∗) and is a nonnegative
operator. Next, define the operator C ∈ B(dom T ∗) by
(3.6)
− 1
C = C
0
2
22)− 1
2 B22(I + B2
22)− 1
2 B∗
hB11 − B12(I + B2
12i C
− 1
0
2
,
which is clearly selfadjoint.
Lemma 3.1. Let T be a closed linear relation from a Hilbert space H to a Hilbert
space K, let Ts be the orthogonal operator part of T , and let the selfadjoint operator
B ∈ B(K) be decomposed as in (3.4). Let the operators C0 and C be defined by
(3.5) and (3.6). Then
(3.7)
T ∗(I + iB)T = (Ts)×C
1/2
0
(I + iC)C
1/2
0 Ts,
and, consequently, T ∗(I + iB)T is maximal sectorial and
(3.8)
mul T ∗(I + iB)T = mul T ∗ = mul (Ts)×.
Proof. In order to prove the equality in (3.7), assume that {h, h′} ∈ T ∗(I + iB)T .
This means that
(3.9)
{h, ϕ} ∈ T and {(I + iB)ϕ, h′} ∈ T ∗
for some ϕ ∈ K. Decompose the element ϕ as
(3.10)
ϕ = ϕ1 + ϕ2, ϕ1 ∈ dom T ∗, ϕ2 ∈ mul T.
6
S. HASSI AND H.S.V. DE SNOO
Since {0, ϕ2} ∈ T , it is clear that
(3.11)
{h, ϕ} ∈ T ⇔ {h, ϕ1} ∈ Ts.
Using (3.10) and the above decomposition (3.4) of B, one observes that
{(I + iB)ϕ, h′} =(cid:26)(cid:18)(I + iB11)ϕ1 + iB12ϕ2
12ϕ1 + (I + iB22)ϕ2(cid:19) , h′(cid:27) ,
iB∗
which implies that the condition {(I + iB)ϕ, h′} ∈ T ∗ is equivalent to
(cid:26) {(I + iB11)ϕ1 + iB12ϕ2, h′} ∈ T ∗,
12ϕ1 + (I + iB22)ϕ2 = 0,
iB∗
or, what is the same thing,
(3.12)
(cid:26) {[I + iB11 + B12(I + iB22)−1B∗
ϕ2 = −i(I + iB22)−1B∗
12ϕ1.
12]ϕ1, h′} ∈ T ∗,
Due to the definitions (3.5) and (3.6) and the identity
(I + iB22)−1 = (I + B2
22)− 1
2 (I − iB22)(I + B2
22)− 1
2 ,
observe that
I + iB11 + B12(I + iB22)−1B∗
12
= C0 + i[B11 − B12(I + B2
22)− 1
2 B22(I + B2
22)− 1
2 B∗
12]
1/2
= C
0
1/2
(I + iC)C
0
.
Therefore, it follows from (3.12), via the equivalence in (3.3), that
(3.13)
{(I + iB)ϕ, h′} ∈ T ∗ ⇔ (cid:26) {C
1/2
0
(I + iC)C
1/2
0 ϕ1, h′} ∈ (Ts)×,
ϕ2 = −i(I + iB22)−1B∗
12ϕ1.
Combining (3.11) and (3.13), one sees that
{h, h′} ∈ (Ts)×C
1/2
0
(I + iC)C
1/2
0 Ts.
Conversely, if this inclusion holds, then there exists ϕ1 ∈ dom T ∗, such that
{h, ϕ1} ∈ Ts
1/2
and {C
0
(I + iC)C
1/2
0 ϕ1, h′} ∈ (Ts)×.
Then define ϕ2 = −i(I + iB22)−1B∗
ϕ = ϕ1 + ϕ2. Hence {h, ϕ} ∈ T , and it follows from (3.13) that
12ϕ1, so that ϕ2 ∈ mul T . Furthermore, define
Therefore one can rewrite T ∗(I + iB)T in the form (3.7).
1
{h, h′} ∈ T ∗(I + iB)T.
Observe that C
2
0 Ts is a closed linear operator from the Hilbert space H to the
Hilbert space dom T ∗ whose adjoint is given by
(3.14)
(C
1/2
0 Ts)× = (Ts)× C
1/2
0
.
Hence, by Theorem 2.1 (Ts)×C
H and by the identity (3.7) the same is true for T ∗(I + iB)T .
(I + iC)C
1/2
0
1/2
0 Ts is a maximal sectorial relation in
The statement in (3.8) follows by tracing the above equivalences for an element
(cid:3)
{0, h′}.
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
7
Remark 3.2. Let ϕ = ϕ1 + ϕ2 ∈ K be decomposed as in (3.10). Then one has the
following equivalence:
(I + iB)ϕ ∈ dom T ∗ ⇔ (I + iB)ϕ = C
1/2
0
(I + iC)C
1/2
0 ϕ1.
To see this, let η = (I + iB)ϕ. Then η ∈ dom T ∗ if and only if
(cid:18)I + iB11
iB∗
12
iB12
I + iB22(cid:19)(cid:18)ϕ1
ϕ2(cid:19) =(cid:18)η
0(cid:19) ,
where dom T ∗ is interpreted as the subspace dom T ∗ × {0} of K. Now apply (3.13).
4. A class of maximal sectorial relations and associated forms
The linear relation T ∗(I +iB)T is maximal sectorial for any selfadjoint B ∈ B(K)
and any closed linear relation T from H to K. Now the corresponding closed sectorial
form will be determined. This gives the appropriate version of Theorem 2.1 in terms
of relations. In fact, the general result is based on a reduction via Lemma 3.1 to
Theorem 2.1.
Theorem 4.1. Let T be a closed linear relation from a Hilbert space H to a Hilbert
space K and let the selfadjoint operator B ∈ B(K) be decomposed as in (3.4). Let
the operators C0 and C be defined by (3.5) and (3.6). Then the form t defined by
1
1
(4.1)
t[h, k] = ((I + iC) C
2
0 Ts h, C
2
0 Ts k),
h, k ∈ dom t = dom T,
is closed and sectorial with vertex at the origin and semi-angle γ ≤ arctan kCk.
Moreover, the maximal sectorial relation H corresponding to the form t is given by
(4.2)
H = (Ts)× C
1/2
0
(I + iC)C
1/2
0 Ts = T ∗(I + iB)T.
A subset of dom t = dom T is a core of the form t if and only if it is a core of the
operator Ts. Moreover, the nonnegative selfadjoint relation Hr corresponding to the
real part (tH)r of the form t is given by
Hr = (Ts)×C0Ts.
1
2
Proof. Since C
0 Ts is a closed linear operator from the Hilbert space H to the Hilbert
1/2
space dom T ∗, Theorem 2.1 (with K replaced by dom T ∗, B by C, and T by C
0 Ts)
shows that the form t in (4.1) is closed and sectorial with vertex at the origin and
semi-angle γ ≤ arctan kCk. Moreover, the maximal sectorial relation associated
with the form t is given by
(C
1/2
0 Ts)×(I + iC)C
1/2
0 Ts = (Ts)×C
1/2
0
(I + iC)C
1/2
0 Ts,
cf. (2.1), (2.2), and (3.14). The identities in (4.2) are clear from Lemma 3.1. The
assertion concerning the core holds, since the factor C0 is bounded with bounded
inverse. The formula (4.2) shows that
1
1
(tH )r[h, k] = (C
and hence Hr = (C
1/2
0 Ts)×C
2
2
0 Tsk),
0 Tsh, C
1/2
0 Ts = (Ts)×C0Ts (cf. the discussion above).
h, k ∈ dom t = dom T,
(cid:3)
Recall that if {h, h′} ∈ T ∗(I + iB)T , then {h, ϕ} ∈ T and {(I + iB)ϕ, h′} ∈ T ∗.
The last inclusion implies the condition (I + iB)ϕ ∈ dom T ∗ ⊂ dom T ∗, giving
rise to ϕ2 = −i(I + iB22)−1B∗
12ϕ1. Thus, for instance, when B = diag (B11, B22),
it follows that ϕ2 = 0, so that it is immediately clear that γ ≤ arctan kB11k,
independent of B22. Note that the following assertions are equivalent:
8
S. HASSI AND H.S.V. DE SNOO
(i) B = diag (B11, B22);
(ii) B12 = 0;
(iii) C0 = I;
(iv) mul T is invariant under B,
in which case C = B11. Hence, if mul T is invariant under B, i.e., if any of the
assertions (i) -- (iv) hold, then Theorem 4.1 gives the following corollary, which co-
incides with [4, Theorem 5.1]. In the case where mul T = {0} the corollary reduces
to Theorem 2.1.
Corollary 4.2. Let T be a closed linear relation from a Hilbert space H to a Hilbert
space K, let Ts be the orthogonal operator part of T , and let mul T be invariant under
the selfadjoint operator B ∈ B(K), so that B = diag (B11, B22). Then the form t
defined by
t[h, k] = ((I + iB11)Tsh, Tsk),
h, k ∈ dom t = dom T,
is closed and sectorial with vertex at the origin and semi-angle γ ≤ arctan kB11k.
Moreover, the maximal sectorial relation H corresponding to the form t is given by
H = (Ts)×(I + iB11)Ts = T ∗(I + iB)T.
In the case that mul T is not invariant under B, one has C0 6= I, and the formulas
are different: for instance, the real part (tH )r in Theorem 4.1 is of the form
1
1
(tH )r[h, k] = (C
2
0 Tsh, C
2
0 Tsk),
h, k ∈ dom t = dom Ts = dom T.
Example 4.3. Assume that B11 6= 0 and
B11 = B12(I + B2
22)− 1
2 B22(I + B2
22)− 1
2 B∗
12,
so that C = 0. In this case the maximal sectorial relation H = T ∗(I + iB)T in
Theorem 4.1 is selfadjoint, i.e., H = Hr and the associated form t is nonnegative.
On the other hand, with such a choice of B the operator part of T determines the
maximal sectorial relation (Ts)∗(I + iB)Ts with semi-angle arctan kB11k > 0, while
T ∗(I + iB)T has semi-angle γ = 0.
5. Maximal sectorial relations and their representations
Let H be a maximal sectorial relation in H and let the closed sectorial form tH
correspond to H; cf. Theorem 1.1. Since the closed form tH is sectorial, one has
the inequality
(5.1)
(tH )i[h] ≤ (tan α)(tH )r[h],
h ∈ dom t,
and in this situation the real part (tH )r is a closed nonnegative form. Hence by the
first representation theorem there exists a nonnegative selfadjoint relation Hr, the
so-called real part of H, such that dom Hr ⊂ dom (tH )r = dom tH and
(tH )r[h, k] = (h′, k),
{h, h′} ∈ Hr,
k ∈ dom (tH )r = dom tH.
This real part Hr, not to be confused with the real part introduced in [6], will play
an important role in formulating the second representation theorem below. First
the case where H is a maximal sectorial operator will be considered, in which case
H is automatically densely defined; see [7].
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
9
Lemma 5.1. Let H be an maximal sectorial operator in H, let the closed sectorial
form tH correspond to H via Theorem 1.1, and let Hr be the real part of H. Then
there exists a unique selfadjoint operator G ∈ B(H) with kGk = tan α, of the form
G =(cid:18)0
0 Gbb(cid:19) : (cid:18)ker Hr
ran Hr(cid:19) →(cid:18)ker Hr
ran Hr(cid:19) ,
0
(5.2)
such that
(5.3)
tH[h, k] = ((I + iG)(Hr)
1
2 h, (Hr)
1
2 k),
h, k ∈ dom tH = dom H
2
r .
1
Moreover, the corresponding maximal sectorial operator H is given by
H = (Hr)
1
2 (I + iG)(Hr)
1
2 ,
with mul H = mul Hr.
Proof. The inequality
(tH )i[h, k]2 ≤ Ctr[h]tr[k] = CkH
1
1
2
r hkkH
2
r kk,
h, k ∈ dom ,
shows the existence of a selfadjoint operator G in H ⊖ ker H such that
(5.4)
(tH )i[h, k] = (G(Hr)
1
2 h, (Hr)
1
2 k),
h, k ∈ dom (Hr)
1
2 .
Extend G to all of H in a trivial way, so that the same formula remains valid; see
Corollary 2.2. It follows from the decomposition t = tr + iti, cf. (1.3), and the
identities (5.5) and (5.4), that
so that
1
1
1
1
tH = (tH )r + i(tH )i,
tH = [h, k] = ((Hr)
2
s h, (Hr)
2
s k) + i(G(Hr)
2
s h, (Hr)
2
s k).
This last identity immediately gives (5.3). The rest follows from Corollary 4.2. (cid:3)
Now let H be a maximal sectorial relation, let Hr be its real part, and let (Hr)s
be its orthogonal operator part. Then one obtains the representation
(5.5)
(tH )r[h, k] = (((Hr))
2
s h, ((Hr)s)
1
1
2 k),
h, k ∈ dom (tH )r = dom ((Hr)s)
1
2 ,
cf. Theorem 2.1. Now apply Corollary 4.2 and therefore one may formulate the
second representation theorem as follows.
Theorem 5.2. Let H be a maximal sectorial relation in H, let the closed sectorial
form tH correspond to H via Theorem 1.1, and let Hr be the real part of H. Then
there exists a selfadjoint operator G ∈ B(H) with kGk = tan α, such that G is
trivial on ker Hr ⊕ mul Hr, and
(5.6)
tH [h, k] = ((I + iG)((Hr)s)
1
2 h, ((Hr)s)
1
2 k),
h, k ∈ dom tH = dom H
2
r .
1
Moreover, the maximal sectorial relation H is given by
(5.7)
H = (((Hr)s)
1
2 )×(I + iG)((Hr)s)
1
2 ,
with mul H = mul Hr.
10
S. HASSI AND H.S.V. DE SNOO
Next, it is assumed that H is a maximal sectorial relation of the form H =
T ∗(I + iB)T , where T is a closed linear relation from a Hilbert space H to a Hilbert
space K and the operator B ∈ B(K) is selfadjoint. Let the operators C0 and C be
defined by (3.5) and (3.6), then
H = (Ts)× C
1/2
0
(I + iC)C
1/2
0 Ts,
while the corresponding closed sectorial form is given
1
1
t[h, k] = ((I + iC) C
2
0 Ts h, C
2
0 Ts k),
h, k ∈ dom t = dom T.
To compare these expressions with (5.6) and (5.7), observe that
1
1
( C
2
0 Ts h, C
2
0 Ts k) = ((Hr)s)
1
2 h, ((Hr)s)
1
2 k)
and
1
1
(C C
2
0 Ts h, C
2
0 Ts k) = (G((Hr)s)
1
2 h, ((Hr)s)
1
2 k).
1/2
It is clear from (4.1) that only the (selfadjoint) compression of C to ran C
0 Ts
contributes to the form (4.2), so that it is straightforward to set up a unitary
mapping; cf. Corollary 2.2.
6. Extremal maximal sectorial extensions of sums of maximal
sectorial relations
Let H1 and H2 be maximal sectorial relations in a Hilbert space H. Then the
sum H1 + H2 is a sectorial relation in H with
dom (H1 + H2) = dom H1 ∩ dom H2,
so that the sum is not necessarily densely defined. In particular, H1 + H2 and its
closure need not be operators, since
(6.1)
mul (H1 + H2) = mul H1 + mul H2.
To describe the class of extremal maximal sectorial extensions of H1 + H2 some
basic notations are recalled from [5], together with the description of the Friedrichs
and Kreın extensions
(H1 + H2)F
and (H1 + H2)K
of H1 + H2, respectively. In order to describe the whole class of extremal extensions
of H1 + H2 and the corresponding closed forms a proper description of the closed
sectorial form tK is essential. The results in Sections 3 and 4 allow a general
treatment that will relax the additional conditions in [5].
6.1. Basic notions. Let H1 and H2 be maximal sectorial relations and decompose
them as follows
1
1
(6.2)
Hj = A
2
j (I + iBj)A
2
j ,
1 ≤ j ≤ 2,
where Aj (the real part of Hj), 1 ≤ j ≤ 2, are nonnegative selfadjoint relations in
H and Bj, 1 ≤ j ≤ 2, are bounded selfadjoint operators in H which are trivial on
ker Aj ⊕ mul Aj; cf. Theorem 5.2. Furthermore, if A1 and A2 are decomposed as
Aj = Ajs ⊕ Aj∞,
1 ≤ j ≤ 2,
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
11
where Aj∞ = {0} × mul Aj , 1 ≤ j ≤ 2, and Ajs, 1 ≤ j ≤ 2, are densely defined
nonnegative selfadjoint operators (defined as orthogonal complements in the graph
sense), then the uniquely determined square roots of Aj, 1 ≤ j ≤ 2 are given by
1
1
A
2
j = A
2
js ⊕ Aj∞,
1 ≤ j ≤ 2.
Associated with H1 and H2 is the relation Φ from H × H to H, defined by
(6.3)
Φ =n {{f1, f2}, f ′
1 + f ′
2} : {fj, f ′
j} ∈ A
2
1
j , 1 ≤ j ≤ 2o .
Clearly, Φ is a relation whose domain and multivalued part are given by
dom Φ = dom A
2
1 × dom A
2
2 , mul Φ = mul H1 + mul H2.
1
1
The relation Φ is not necessarily densely defined in H × H, so that in general Φ∗ is
a relation as mul Φ∗ = (dom Φ)⊥. Furthermore, the adjoint Φ∗ of Φ is the relation
from H to H × H, given by
Φ∗ =n{h, {h′
1, h′
2}} : {h, h′
j} ∈ A
2
1
j , 1 ≤ j ≤ 2o .
The identity (6.4) shows that the (orthogonal) operator part (Φ∗)s of Φ∗ is given
by:
(6.4)
(6.5)
(6.6)
(6.7)
(6.8)
(Φ∗)s = n{h, {h′
= nnh, {A
1, h′
2}} : {h, h′
j} ∈ A
1
2
1sh, A
2
1
2sh}o : h ∈ dom A
2
1
js, 1 ≤ j ≤ 2o
1
2
1 ∩ dom A
2
1
2o .
The identities (6.4) and (6.5) show that
dom Φ∗ = dom A
2
1 ∩ dom A
2
2 , mul Φ∗ = mul H1 × mul H2, ran (Φ∗)s = F0,
1
1
where the subspace F0 ⊂ H × H is defined by
F0 =nnA
1
2
1sh, A
2
1
2sho : h ∈ dom A
1
2
1 ∩ dom A
2
1
2o .
The closure of F0 in H × H will be denoted by F. Define the relation Ψ from H to
H × H by
Ψ =nnh,nA
1
2
1sh, A
2
1
2shoo : h ∈ dom H1 ∩ dom H2o ⊂ H × (H × H).
It follows from this definition that
dom Ψ = dom H1 ∩ dom H2, mul Ψ = {0},
ran Ψ = E0,
where the space E0 ⊂ H × H is defined by
E0 =nnA
1
2
1sf, A
2
1
2sfo : f ∈ dom H1 ∩ dom H2o .
Observe that E0 ⊂ F0. The closure of E0 in H × H will be denoted by E. Hence,
(6.9)
E ⊂ F.
Comparison of (6.5) and (6.7) shows
(6.10)
Ψ ⊂ (Φ∗)s,
and thus the operator Ψ is closable and Ψ∗∗ ⊂ (Φ∗)s. It follows from dom Ψ∗ =
(mul Ψ∗∗)⊥ and mul Ψ∗ = (dom Ψ)⊥, that
dom Ψ∗ = H, mul Ψ∗ = (dom H1 ∩ dom H2)⊥.
12
S. HASSI AND H.S.V. DE SNOO
Next, define the relation K from H × H to H by
(6.11)
K = (cid:8){{(I + iB1)A
1
2
1sf, (I + iB2)A
2
2sf }, f ′
1 + f ′
2} :
1
1
{(I + iB1)A
1
2
1sf, f ′
1} ∈ A
2
1 , {(I + iB2)A
2
2sf, f ′
2} ∈ A
1
2
1
2(cid:9)
⊂ (H × H) × H.
Clearly, the domain and multivalued part of K are given by
dom K = D0, mul K = mul (H1 + H2),
where
(6.12)
D0 =n {(I + iB1)A
1/2
1s f, (I + iB2)A
1/2
2s f } : f ∈ dom H1 ∩ dom H2o .
The closure of D0 in H × H will be denoted by D.
Lemma 6.1. The relations K, Φ, and Ψ satisfy the following inclusions:
(6.13)
K ⊂ Φ ⊂ Ψ∗, Ψ ⊂ Φ∗ ⊂ K ∗.
Proof. To see this note that K ⊂ Φ follows from (6.3) and (6.11), and that Ψ ⊂ Φ∗
follows from (6.4) and (6.7). Therefore, also Φ∗ ⊂ K ∗ and Φ ⊂ Φ∗∗ ⊂ Ψ∗.
(cid:3)
6.2. The Friedrichs and the Kreın extensions of H1 + H2. The descriptions
of the Friedrichs extension and the Kreın extension (H1 + H2)F and (H1 + H2)K
of H1 + H2 are now recalled from [5]. For this, define the orthogonal sum of the
operators B1 and B2 in H × H by
B⊕ := B1 ⊕ B2 =(cid:18)B1
0 B2(cid:19) .
0
The descriptions of (H1 + H2)F and (H1 + H2)K incorporate the initial data on
the factorizations (6.2) of H1 and H2 via the mappings Φ, Ψ, and K in Subsection
6.1. The construction of the Friedrichs extension was given in [5, Theorem 3.2],
where some further details and a proof of the following result can be found. The
new additions in the next theorem are the second representations for (H1 + H2)F
and tF that will be needed in the rest of this paper.
Theorem 6.2. Let H1 and H2 be maximal sectorial and let Ψ be defined by (6.7).
Then the Friedrichs extension of H1 + H2 has the expression
(6.14)
(H1 + H2)F = Ψ∗(I + iB⊕)Ψ∗∗ = Ψ∗C
1/2
0
(I + iC)C
1/2
0 PD(Ψ∗∗)s.
The closed sectorial form tF associated with (H1 + H2)F is given by
(6.15)
tF [f, g] = ((I + iB⊕)Ψ∗∗f, Ψ∗∗g) = (C
1/2
0
(I + iC)C
1/2
0 PD(Ψ∗∗)sf, PD(Ψ∗∗)sg),
for all f, g ∈ dom tF = dom Ψ∗∗.
Proof. As indicated the first expressions for (H1 + H2)F in (6.14) and tF in (6.15)
have been proved in [5, Theorem 3.2] and, hence, it suffices to derive the second
expressions in (6.14) and (6.15).
By definition, one has ran Ψ = E0 (see (6.7), (6.8)), and by Lemma 6.1 one has
Ψ ⊂ Ψ∗∗ ⊂ K ∗, which after projection onto D = dom K yields
PDΨ∗∗ ⊂ PDK ∗ = (K ∗)s.
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
13
Notice that D0 = dom K = (I + iB⊕)E0 (see (6.8), (6.12)). Since the operator
I + iB⊕ is bounded with bounded inverse, one has the equality
(6.16)
D = (I + iB⊕)E.
It follows that the range of (I + iB⊕)Ψ∗∗ belongs to D = dom K. Now by Remark
3.2 this implies that for all f ∈ dom Ψ∗∗ one has the equality
(6.17)
This leads to
(I + B⊕)(Ψ∗∗)sf = C
1/2
0
(I + iC)C
1/2
0 PD(Ψ∗∗)sf.
Ψ∗(I + iB⊕)Ψ∗∗ = Ψ∗C
1/2
0
(I + iC)C
1/2
0 PD(Ψ∗∗)s,
which proves (6.14). Similarly by substituting (6.17) into the first formula for tF
and noting that PDC
(cid:3)
1/2
0 = PD, one obtains the second formula in (6.15).
Also the construction of the Kreın extension for the sum H1 + H2 can be found
in [5, Theorem 3.2]. However, the corresponding form tK was described only under
additional conditions to prevent the difficulty that appears by the fact that the
multivalued part of (H1 + H2)K is in general not invariant under the mapping
B⊕. Theorem 4.1 allows a removal of these additional conditions and leads to a
description of the form tK in the general situation.
For this purpose, decompose the Hilbert space H × H as follows
(6.18)
H × H = dom K ⊕ mul K ∗,
and let P be the orthogonal projection onto dom K. Moreover, decompose the
selfadjoint operator B⊕ ∈ B(H × H) accordingly:
(6.19)
B⊕ =(cid:18)B11 B12
12 B22(cid:19) : (cid:18)dom K
mul K ∗(cid:19) →(cid:18) dom K
mul K ∗(cid:19) .
B∗
Next define the operator C0 ∈ B(dom K ∗) by
(6.20)
C0 = I + B12(I + B2
22)−1B∗
12,
and the operator C ∈ B(dom K ∗) by
(6.21)
− 1
C = C
0
2
which is clearly selfadjoint.
hB11 − B12(I + B2
22)− 1
2 B22(I + B2
22)− 1
2 B∗
12i C
− 1
0
2
,
Theorem 6.3. Let H1 and H2 be maximal sectorial relations in a Hilbert space
H, let K be defined by (6.11), and let C0 and C be given by (6.20) and (6.21),
respectively. Then the Kreın extension of H1 + H2 has the expression
(H1 + H2)K = K ∗∗(I + iB⊕)K ∗ = ((K ∗)s)× C
1/2
0
1/2
(I + iC)C
0
(K ∗)s.
The closed sectorial form tK associated with (H1 + H2)K is given by
tK[f, g] = ((I + iC)C
1/2
0
(K ∗)sf, C
1/2
0
(K ∗)sg),
f, g ∈ dom tK = dom K ∗.
Proof. The first equality in the first statement is proved in [5, Theorem 3.2].
The second equality is obtained by applying Theorem 4.1 to the sectorial relation
K ∗∗(I + iB⊕)K ∗.
The statement concerning the form tK is a consequence of this second represen-
(K ∗)s is a closed operator and hence one can apply
(cid:3)
tation of (H1 + H2)K , since C
Theorem 2.1 to get the desired expression for the corresponding form tK.
1/2
0
14
S. HASSI AND H.S.V. DE SNOO
The form tK described in Theorem 6.3 can be used to give a complete descrip-
tion of all extremal maximal sectorial extensions of the sum H1 + H2. Namely, a
maximal sectorial extension eH of a sectorial relation S is extremal precisely when
the corresponding closed sectorial form t eH is a restriction of the closed sectorial
form tK generated by the Kreın extension SK of S; see e.g.
[4, Definition 7.7,
Theorems 8.2, 8.4, 8.5]. Therefore, Theorem 6.3 implies the following description
of all extremal maximal sectorial extensions of H1 + H2.
Theorem 6.4. Let H1 and H2 be maximal sectorial relations in H, let Ψ and K
be defined by (6.7) and (6.11), respectively, and let PD be the orthogonal projection
from H × H onto D = dom K. Then the following statements are equivalent:
(i) eH is an extremal maximal sectorial extension of H1 + H2;
(ii) eH = R∗(I + iC)R, where R is a closed linear operator satisfying
1/2
0 PDΨ∗∗ ⊂ R ⊂ C
(K ∗)s.
1/2
0
C
Proof. For comparison with the abstract results this statement will be proved by
means of the constructions used in [4]. Let S = H1 + H2 then the sectorial relation
S gives rise to a Hilbert space HS and a selfadjoint operator BS ∈ B(HS) such that
the Friedrichs extension SF and the Kreın extension SK of S are given by
SF = Q∗(I + iBS)Q∗∗,
tF = J ∗∗(I + iBS)J ∗,
with corresponding forms
tF [f, g] = ((I + iBS)Q∗∗f, Q∗∗g),
f, g ∈ dom Q∗∗,
and
tK[f, g] = ((I + iBS)J ∗f, J ∗g),
f, g ∈ dom J ∗;
see [4, Theorem 8.3]. Here Q : H → HS is an operator and J : HS → H is a densely
defined linear relation such that
J ⊂ Q∗, Q ⊂ J ∗;
in particular, the adjoint J ∗ is an operator.
Recall from Theorem 6.3 that
tK[f, g] = ((I + iC)C
1/2
0
(K ∗)sf, C
1/2
0
(K ∗)sg),
while Theorem 6.2 gives
tF [f, g] = ((I + iB⊕)Ψ∗∗f, Ψ∗∗g) = (C
1/2
0
(I + iC)C
1/2
0 PD(Ψ∗∗)sf, PD(Ψ∗∗)sg).
Now apply [4, Theorem 8.4] and Corollary 2.2.
(cid:3)
6.3. The form sum construction. The maximal sectorial relations H1 and H2
generate the following closed sectorial form
1
1
1
1
1
1
(6.22) ((I + iB1)A
2
1sh, A
2
1sk) + ((I + iB2)A
2
2sh, A
2
2sk),
h, k ∈ dom A
2
1 ∩ dom A
2
2 .
Observe that the restriction of this form to dom Ψ∗∗ is equal to
(6.23)
(Ψ∗∗h, Ψ∗∗k) = ((I + iB1)A
1sk) + ((I + iB2)A
1sh, A
2sh, A
2
2
2sk),
2
2
1
1
1
1
h, k ∈ dom Ψ∗∗,
since Ψ∗∗ ⊂ (Φ∗)s, cf. (6.5). Thus, the form in (6.22) has a natural domain which
is in general larger than dom Ψ∗∗.
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
15
Theorem 6.5. Let H1 and H2 be maximal sectorial relations in H, let Φ be given
by (6.3), and let E = clos E0 and F = clos F0 be defined by (6.8) amd (6.6). Then
the maximal sectorial relation
Φ∗∗(I + iB⊕)Φ∗
is an extension of the relation H1 + H2, which corresponds to the closed sectorial
form in (6.22).
Moreover, the following statements are equivalent:
(i) Φ∗∗(I + iB⊕)Φ∗ is extremal;
(ii) E = F.
Proof. The first statement is proved in [5, Theorem 3.5]. For the proof of the
equivalence of (i) and (ii) appropriate modifications are needed in the arguments
used in the proof of [5, Theorem 3.5]. The special case treated there was based
on the additional assumption that D = E, where D = dom K; a condition which
implies the invariance of mul K ∗ under the operator B⊕. In the present general
case such an invariance property cannot be assumed. Now for simplicity denote the
the inclusion F ⊂ E. By Theorem 6.4 and mul Φ∗ = mul H1 × mul H2 one sees
form sum extension of H1 + H2 briefly by bH = Φ∗∗(I + iB⊕)Φ∗.
(i) ⇒ (ii) Assume that bH is extremal. Since E ⊂ F by (6.9), it is enough to prove
bH = ((Φ∗)s)∗(I + iB⊕)(Φ∗)s = R∗(I + iC)R,
for some closed operator R satisfying
(6.24)
(6.25)
1/2
0 PDΨ∗∗ ⊂ R ⊂ C
1/2
0
C
(K ∗)s,
where PD is the orthogonal projection of H × H onto D = dom K. Recall that
(Φ∗)s ⊂ Φ∗ ⊂ K ∗ and hence PD(Φ∗)s ⊂ PDK ∗ = (K ∗)s. Moreover, one has
dom PD(Φ∗)s = dom (Φ∗)s = dom R, since by assumption these two domains co-
R, one has
incide with the corresponding joint form domain. Denoting bR = C −1/2
dom bR = dom PD(Φ∗)s and (6.24) can be rewritten as
bR,
where bR satisfies PDΨ∗∗ ⊂ bR ⊂ (K ∗)s. One concludes that PD(Φ∗)s = bR, since
bH = ((Φ∗)s)∗(I + iB⊕)(Φ∗)s = bR∗C
both operators are restrictions of (K ∗)s, and thus
1/2
(I + iC)C
0
(6.26)
1/2
0
0
(6.27)
Now one obtains from (6.26) the equalities
((Φ∗)s)∗PD = bR∗.
((Φ∗)s)∗(I + iB⊕)(Φ∗)s = bR∗C
= ((Φ∗)s)∗PDC
1/2
= ((Φ∗)s)∗C
0
Hence, for every f ∈ dom bH one has
(I + iB⊕)(Φ∗)sf − C
1/2
0
1/2
(I + iC)C
0
1/2
0
1/2
(I + iC)C
0
bR
bR
1/2
0
1/2
(I + iC)C
0
1/2
(I + iC)C
0
bR.
bRf ∈ ker ((Φ∗)s)∗.
1/2
(I + iC)C
0
1/2
Here C
0
1/2
Therefore, there exists ϕ ∈ E such that C
0
bRf ∈ D = dom K and D = dom K = (I + iB⊕)E; see (6.16).
bRf = (I + iB⊕)ϕ. On
1/2
(I + iC)C
0
16
S. HASSI AND H.S.V. DE SNOO
the other hand, (Φ∗)sf ∈ F = ran (Φ∗)s = (ker ((Φ∗)s)∗)⊥, see (6.5), (6.6). Since
ϕ ∈ E ⊂ F, this yields
((I + iB⊕)((Φ∗)sf − ϕ), (Φ∗)sf − ϕ) = 0,
and thus (Φ∗)sf − ϕ = 0. Consequently, for all f ∈ dom bH one has
Since dom bH is a core for the corresponding closed form, or equivalently, the closure
of (Φ∗)s↾ dom bH is equal to (Φ∗)s, the claim follows: F = ran (Φ∗)s ⊂ E.
(ii) ⇒ (i) Assume that E = F. Then F0 = ran (Φ∗)s ⊂ E and hence for all
f ∈ dom (Φ∗)s one has (I + B⊕)(Φ∗)sf ∈ dom K. By Remark 3.2 this implies that
(Φ∗)sf ∈ E.
(6.28)
(I + B⊕)(Φ∗)sf = C
1/2
0
(I + iC)C
1/2
0 PD(Φ∗)sf.
On the other hand, as shown above PD(Φ∗)s ⊂ PDK ∗ = (K ∗)s. Let bR be the closure
of (K ∗)s↾ dom (Φ∗)s. Then bR∗ satisfies the identity (6.27). Since Ψ∗∗ ⊂ (Φ∗)s (see
(6.10)) one obtains PDΨ∗∗ ⊂ bR. The identities (6.27) and (6.28) imply that for all
f ∈ dom bH the equalities
((Φ∗)s)∗(I + B⊕)(Φ∗)sf = ((Φ∗)s)∗PDC
1/2
0 PD(Φ∗)sf
(I + iC)C
1/2
0
1/2
0
1/2
(I + iC)C
0
= bR∗C
bRf
1/2
hold. Then the closed operator R = C
0
bR satisfies the inclusions (6.25) as well as
the desired identity ((Φ∗)s)∗(I + B⊕)(Φ∗)s = R∗(I + iC)R, and thus bH is extremal,
cf. Theorem 6.4.
(cid:3)
Theorem 6.5 is a generalization of [5, Theorem 3.5], where an additional invari-
ance of mul K ∗ under the operator B⊕ was used. Moreover, Theorem 6.5 generalizes
a corresponding result for the form sum of two closed nonnegative forms established
earlier in [3, Theorem 4.1].
The present result relies on Theorem 4.1, where the description of the closed
sectorial form generated by a general maximal sectorial relation of the form H =
T ∗(I + iB)T where T is a closed relation. This generality implies that with special
choices of B the relation H can be taken to be nonnegative and selfadjoint, i.e., the
corresponding closed form t becomes nonnegative; see Example 4.3.
References
[1] Yu.M. Arlinskiı, "Maximal sectorial extensions and closed forms associated with them",
Ukrainian Math. J., 48 (1996), 723 -- 739.
[2] M. Haase, The functional calculus for sectorial operators, Oper. Theory Adv. Appl., 169,
Birkhauser Verlag, 2006.
[3] S. Hassi, A. Sandovici, H.S.V. de Snoo, and H. Winkler, "Extremal extensions for the sum
of nonnegative selfadjoint relations", Proc. Amer. Math. Soc., 135 (2007), 3193 -- 3204.
[4] S. Hassi, A. Sandovici, H.S.V. de Snoo, and H. Winkler, "Extremal maximal sectorial exten-
sions of sectorial relations", Indag. Math., 28 (2017), 1019 -- 1055.
[5] S. Hassi, A. Sandovici, and H.S.V. de Snoo, "Factorized sectorial relations, their maximal
sectorial extensions, and form sums", Banach J. Math. Anal., 13 (2019), 538 -- 564.
[6] S. Hassi, H.S.V. de Snoo, and F.H. Szafraniec, "Componentwise and Cartesian decomposi-
tions of linear relations", Dissertationes Mathematicae, 465 (2009), 59 pages.
[7] T. Kato, Perturbation theory for linear operators, Springer-Verlag, Berlin, 1980.
[8] F.S. Rofe-Beketov, "Numerical range of a linear relation and maximal relations", Teor. Funkt-
sii Funktsional. Anal. i Prilozhen, 44 (1985), 103 -- 112 (Russian) [English translation: J. Math.
Sciences, 48 (1990), 329 -- 336].
A CLASS OF SECTORIAL RELATIONS AND THE ASSOCIATED CLOSED FORMS
17
[9] A. Sandovici, Contributions to the extension theory in Hilbert spaces, Doctoral dissertation,
University of Groningen, 2006.
[10] Z. Sebesty´en and Zs. Tarcsay, "T ∗T always has a positive selfadjoint extension", Acta Math.
Hungar., 135 (2012), 116 -- 129.
Department of Mathematics and Statistics, University of Vaasa, P.O. Box 700, 65101
Vaasa, Finland
E-mail address: [email protected]
Bernoulli Institute for Mathematics, Computer Science and Artificial Intelligence,
University of Groningen, P.O. Box 407, 9700 AK Groningen, Nederland
E-mail address: [email protected]
|
Subsets and Splits