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
|
---|---|---|---|---|---|---|---|---|---|
1806.06307 | 1 | 1806 | 2018-06-16T22:23:18 | The inner kernel theorem for a certain Segal algebra | [
"math.FA"
] | The Segal algebra ${\textbf{S}}_{0}(G)$ is well defined for arbitrary locally compact Abelian Hausdorff (LCA) groups $G$. Despite the fact that it is a Banach space it is possible to derive a kernel theorem similar to the Schwartz kernel theorem, of course without making use of the Schwartz kernel theorem. First we characterize the bounded linear operators from ${\textbf{S}}_{0}(G_1)$ to ${\textbf{S}}_{0}'(G_2)$ by distributions in ${\textbf{S}}_{0}'(G_1 \times G_2)$. We call this the "outer kernel theorem". The "inner kernel theorem" is concerned with the characterization of those linear operators which have kernels in the subspace ${\textbf{S}}_{0}(G_1 \times G_2)$, the main subject of this manuscript. We provide a description of such operators as regularizing operators in our context, mapping ${\textbf{S}}_{0}'(G_1)$ into test functions in ${\textbf{S}}_{0}(G_2)$, in a $w^{*}$-to norm continuous manner. The presentation provides a detailed functional analytic treatment of the situation and applies to the case of general LCA groups, without recurrence to the use of so-called Wilson bases, which have been used for the case of elementary LCA groups. The approach is then used in order to describe natural laws of composition which imitate the composition of linear mappings via matrix multiplications, now in a continuous setting. We use here that in a suitable (weak) form these operators approximate general operators. We also provide an explanation and mathematical justification used by engineers explaining in which sense pure frequencies "integrate" to a Dirac delta distribution. | math.FA | math | The inner kernel theorem for a certain Segal algebra
Hans G. Feichtinger and Mads S. Jakobsen
June 19, 2018
Abstract
0(G2) by distributions in S′
The Segal algebra S0(G) is well defined for arbitrary locally compact Abelian Hausdorff (LCA)
groups G. Despite the fact that it is a Banach space it is possible to derive a kernel theorem similar to
the Schwartz kernel theorem, of course without making use of the Schwartz kernel theorem. First we
characterize the bounded linear operators from S0(G1) to S′
0(G1 × G2). We
call this the "outer kernel theorem". The "inner kernel theorem" is concerned with the characterization
of those linear operators which have kernels in the subspace S0(G1 × G2), the main subject of this
manuscript. We provide a description of such operators as regularizing operators in our context,
mapping S′
0(G1) into test functions in S0(G2), in a w∗-to norm continuous manner. The presentation
provides a detailed functional analytic treatment of the situation and applies to the case of general LCA
groups, without recurrence to the use of so-called Wilson bases, which have been used for the case of
elementary LCA groups. The approach is then used in order to describe natural laws of composition
which imitate the composition of linear mappings via matrix multiplications, now in a continuous
setting. We use here that in a suitable (weak) form these operators approximate general operators.
We also provide an explanation and mathematical justification used by engineers explaining in which
sense pure frequencies "integrate" to a Dirac delta distribution.
8
1
0
2
n
u
J
6
1
]
.
A
F
h
t
a
m
[
1 Introduction
1
v
7
0
3
6
0
.
6
0
8
1
:
v
i
X
r
a
The focus of this paper is on the kernel theorem associated with the Segal algebra S0(G) introduced by the
first named author in [10]. Given a locally compact Abelian Hausdorff (LCA) group G we write bG for its
dual group, and for each ω ∈ bG we denote by Eωf (t) = ω(t)f (t), t ∈ G the modulation (frequency-shift)
operator. We define the set of test functions using convolution "∗" and the usual norm in L1:
S0(G) =nf ∈ L1(G) : Z bG
kEωf ∗ f k1 dω < ∞ o.
Any non-zero function g ∈ S0(G) (also called window or Gabor atom) defines a norm on S0(G) via
kf kS0,g = kf kS0(G),g :=Z bG
kEωf ∗ gk1 dω,
that turns S0 into a Banach space. These norms are pairwise equivalent and we therefore allow ourselves
to simply write k · kS0 without specifying the function g. The space S0(G) is a Fourier invariant Banach
algebra under convolution and pointwise multiplication. We call continuous linear functionals on this
space distributions. They form altogether the dual space S′
0(G), which is a Banach space itself. The action
σ ∈ S′
0(G) on a (test) function f ∈ S0(G) is described by the bilinear form
( · , · )S0,S′
0(G) : S0(G) × S′
0(G) → C,
(f, σ)S0,S′
0(G) = σ(f ).
(3)
Throughout the paper Bil(X × Y, Z) is the space of bilinear and norm continuous operators from the
normed space X × Y into the normed space Z and, similarly, Lin(X, Y ) is the space of linear and norm
continuous operators from X into Y , each of them endowed with their natural norm.
1
(1)
(2)
Using these spaces we can formulate the following result.
Theorem 1.1 (Outer kernel theorem for S0). For LCA groups G1 and G2 the four Banach spaces
0(G1 × G2), Bil(S0(G1) × S0(G2), C), Lin(S0(G1), S′
S′
0(G2)) and Lin(S0(G2), S′
0(G1))
are naturally isomorphic. In particular, given any σ ∈ S′
0(G1 × G2),
A ∈ Bil(S0(G1) × S0(G2), C), T ∈ Lin(S0(G1), S′
0(G2)) or S ∈ Lin(S0(G2), S′
0(G1))
the others are uniquely determined by the following identity, valid for f (1) ∈ S0(G1), f (2) ∈ S0(G2):
(f (1) ⊗ f (2), σ)S0,S′
0(G1×G2) = A(f (1), f (2)) = (f (2), T f (1))S0,S′
0(G2) = (f (1), Sf (2))S0,S′
0(G1).
The unique distribution σ ∈ S′
0(G1 × G2) associated with A, T or S is called the kernel of A, T or S,
respectively and we write κ(A) = κ(T ) = κ(S) = σ. The outer kernel theorem for S0 was first announced
in [9]. Its proof can be found in [15, 16, 21], for example.
This paper will consider the following questions:
Is there an analogue of Theorem 1.1 concerning operators that can be naturally identified with the
functions in S0(G1 × G2) (rather than its dual space S′
0(G1 × G2))?
This question has been considered and answered before in [6] and [16], however not in this generality (cf.
the comment following Theorem 1.3 below). As is well known (and as we will explain in detail in Section
2) there is a natural isomorphic copy of the Banach space of functions S0(G) inside its dual space S′
0(G).
We are therefore also interested in the following question:
Can we characterize those operators in Lin(S0(G1), S′
kernel σ ∈ S′
0(G1 × G2) which is induced by a function in S0(G1 × G2)?
0(G2)) ∼= Bil(S0(G1) × S0(G2), C) that have a
The main result of this paper, Theorem 1.3 and Theorem 3.2, answer these two questions. For their
formulation we need two auxiliary spaces:
Definition 1.2. For LCA groups G1 and G2 we define the following two sets of operators:
A(G1, G2) = {A ∈ Bil(S′
B(G1, G2) = {T ∈ Lin(S′
0(G2), C) : A is weak∗ continuous in each coordinate },
0(G1) × S′
0(G1), S0(G2)) : T maps bounded weak∗convergent nets in S′
0(G1)
into norm convergent nets in S0(G2) }.
In Section 4 we prove that the spaces A(G1, G2) and B(G1, G2) are complete with respect to their
natural subspace topologies. Furthermore, we shall show that all elements in B(G1, G2) are nuclear (and
thus, in particular, also compact) operators from S′
0(G1) into S0(G2) and that they are consequently trace
class operators for the case G1 = G2 (see Section 3.4).
We are now ready to formulate our first main result:
Theorem 1.3 (Inner kernel theorem for S0). For LCA groups G1 and G2 the four Banach spaces
S0(G1 × G2), A(G1, G2), B(G1, G2) and B(G2, G1)
are naturally isomorphic. In particular, if any
K ∈ S0(G1 × G2), A ∈ A(G1, G2), T ∈ B(G1, G2) or S ∈ B(G2, G1)
is given, then the others are uniquely determined such that, for all σ(i) ∈ S′
0(Gi), i = 1, 2,
(K, σ(1) ⊗ σ(2))S0,S′
0(G1×G2) = A(σ(1), σ(2)) = (T σ(1), σ(2))S0,S′
0(G2) = (Sσ(2), σ(1))S0,S′
0(G1).
(4)
2
If the groups G1 and G2 are elementary, i.e., isomorphic to Rn × Zm × Tl × F , where F is some finite
Abelian group and l, n, m ∈ N0, then a proof of Theorem 1.3 can be found in [16]. However, the methods
used there do not extend to general locally compact Abelian groups. The lack of a proof of the inner
kernel theorem for S0(G) on general locally compact Abelian groups also serves as a motivation for this
paper. We devote the entirety of Section 4 to the proof of Theorem 1.3.
Similar to the outer kernel theorem, given any A ∈ A(G1, G2), T ∈ B(G1, G2) or S ∈ B(G2, G1), the
function K ∈ S0(G1 × G2) satisfying (4) is called the kernel of A, T or S and we denote this function by
κ(A), κ(T ) or κ(S).
A combination of the inner and outer kernel theorem together with the continuous embedding of S0
0 (see Lemma 2.4) allows us to make the following diagram for any two LCA groups G1 and G2.
into S′
Here HS(G1, G2) are the Hilbert-Schmidt operators from L2(G1) into L2(G2).
Inner Kernel Theorem
A(G1, G2) ∼=
S0(G1 × G2)
∼=
B(G1, G2) ⊆ Lin(S′
0(G1), S0(G2)) ⊆ Lin(L2(G1), S0(G2)) ⊆ Lin(S0(G1), S0(G2))
⊆
⊆
⊆
Lin(S′
0(G1), L2(G2)) ⊆ Lin(L2(G1), L2(G2)) ⊆ Lin(S0(G1), L2(G2))
⊆
⊆ L2(G1 × G2) ∼= HS(G1, G2) ⊆
⊆
Hilbert-Schmidt Operators
Lin(S′
0(G1), S′
0(G2)) ⊆ Lin(L2(G1), S′
0(G2)) ⊆ Lin(S0(G1), S′
0(G2))
Bil(S0(G1) × S0(G2), C) ∼=
S′
0(G1 × G2)
Outer Kernel Theorem
∼=
Furthermore, we have the following (strict) inclusions for Banach spaces of operators:
B(G1, G2) ⊆ HS(G1, G2) ⊆ Lin(S0(G1), S′
0(G2)).
(5)
They even form a Banach Gelfand triple and have been investigated in [2],[6] and [16].
Both the inner and outer kernel theorem for S0 are analogous to the situation for nuclear spaces, cf.
Chapter 50 and 51 in Trèves book [28]. Further references to the theory of nuclear spaces and their kernel
theorems are Delcroix [8] and Hörmander [19]. Speaking about nuclear spaces, let us remark here that S0
contains the Schwartz(-Bruhat) space as a dense subspace ([10, Theorem 9]) and that S′
0 is a subspace of
the tempered distributions. For more on the Schwartz-Bruhat functions we refer to the original literature
[4, 24].
The paper is structured as follows. Section 2 recollects necessary facts about the function space S0(G)
and its continuous dual space S′
0(G). Section 3 is comprised of several smaller pieces. The first of which,
Section 3.1, states when the continuity of the operators in the spaces A and B can be described with the
notion of sequences rather than that of nets. Section 3.2 contains the second main result of this paper,
Theorem 3.2. This result gives a more quantitative description of the operators in Lin(S0(G1), S′
0(G2))
that have a kernel in S0 and establishes a more natural norm on those operators (rather than the subspace
topologies as mentioned following Definition 1.2). Section 3.3 shows similarities between the matrix rep-
resentation of operators between finite dimensional spaces and the space B(G1, G2). Examples of operator
with kernel in S0 and results concerning series representations, nuclearity and trace-class properties of the
operators in B are shown in Section 3.4.
In Section 3.5 we define and show examples of what we call
regularizing approximations of the identity. Finally, Section 3.6 contains some comments on extensions of
the theory and references to related work. As mentioned earlier, Section 4 is solely concerned with the
proof of the Theorem 1.3.
3
2 Preliminaries
2.1 Harmonic analysis on LCA groups
Throughout the paper we will be working with locally compact Abelian Hausdorff groups G. As any locally
compact group, they carry an (up to scaling) unique translation invariant measure, the Haar measure. The
dual group bG of an LCA group G is the multiplicative group of all continuous group homomorphisms from
G into the torus {z ∈ C : z = 1}. Under the topology of uniform convergence on compact sets the dual
group becomes a LCA group itself. As such it also carries a Haar measure. Without loss of generality we
always assume that these measures are normalized such that
f (x) =Z bG
f (ω) ω(x) dµ bG(ω) for almost every x ∈ G
for all f ∈ L1(G) with f ∈ L1(bG), where f is the Fourier transform of f , f (ω) = RG f (x) ω(x) dµG(x),
ω ∈ bG. Typically we will perform integration in the time-frequency plane (phase space) G × bG so that
we encounter integrals of the formRG× bG f (ν) dµG× bG(ν) for suitable complex valued functions f on G × bG.
From now on we shall simplify the notation and write RG . . . dx, R bG . . . dω, and RG× bG . . . dν, rather than,
e.g.,RG× bG . . . dµG× bG(ν). For more on integration on locally compact groups and abstract harmonic analysis
we refer to, e.g., [17, 23] and [25].
2.2 The space S0
In this section we summarize results on the space S0 and its dual space S′
0. Since we often will deal with
0(G1) and S′
functions in the spaces S0(G1) and S0(G2) and also with distributions in S′
0(G2) for typically
different locally compact Abelian groups Gi, i = 1, 2, we define once and for all that f (i) and σ(i) denote
a function and a distribution in S0(Gi) and S′
0(Gi), respectively. Different functions in S0(Gi) will be
denoted either by different letters, e.g., f (i), g(i) and h(i), or with an index, f (i)
j
.
For functions in S0(G1) and S0(G2) the tensor product
(cid:0)f (1) ⊗ f (2)(cid:1)(x(1), x(2)) = f (1)(x(1)) · f (2)(x(2)),
is a bilinear and bounded operator into S0(G1 × G2). In fact,
(x(1), x(2)) ∈ G1 × G2,
kf (1) ⊗ f (2)kS0(G1×G2),g(1)⊗g(2) = kf (1)kS0(G1),g(1) · kf (2)kS0(G2),g(2).
Any f ∈ S0(G1 × G2) can be written as a sum of tensor products of appropriately chosen functions.
Lemma 2.1. Given LCA groups G1 and G2 one has S0(G1 × G2) = S0(G1) ⊗S0(G2).
That is, any f ∈ S0(G1 × G2) can be written (in a non-unique way) as
f =Xj∈N
f (1)
j ⊗ f (2)
j
such that Xj∈N
kf (1)
j kS0 kf (2)
j kS0 < ∞,
(6)
where the sum is absolutely norm convergent in S0(G1 ×G2). Moreover, the S0(G1×G2)-norm is equivalent
to the projective tensor product norm
kf k = inf(cid:8)Xj∈N
kf (1)
j kS0 kf (2)
j kS0(cid:9),
(7)
where the infimum is taken over all possible representations of f as in (6). These statements were originally
proven in [10, Theorem 7] and can also be found in [21, Theorem 7.4].
Recall that the translation operator Tx and the modulation operator Eω, given by
Txf (t) = f (t − x),
Eωf (t) = ω(t)f (t),
4
t, x ∈ G, ω ∈ bG
act as linear and isometric operators on S0(G), and so do time-frequency shifts, given by
Besides the definition of S0 in the introduction, there is also an atomic characterization:
π(ν) = π(x, ω) = EωTx
for ν = (x, ω) ∈ G × bG.
Lemma 2.2. Fix a non-zero function g ∈ S0(G). For any f ∈ S0(G) there exists a sequence c ∈ ℓ1(N)
and elements νj ∈ G × bG, j ∈ N such that f = Pj∈N cj π(νj)g. Furthermore kf k = inf kck1, where the
infimum is taken over all possible representations of f as above, defines an equivalent norm on S0(G).
This result goes back to [11] and can also be found in [21, Theorem 7.2].
For each non-zero g ∈ S0(G) the dual space S′
0(G) is a Banach space with respect to the usual operator
topology induced by the family of equivalent norms
kσkS′
0(G),g =
sup
f ∈S0(G)\{0}
(f, σ)S0,S′
0(G)
kf kS0(G),g
, σ ∈ S′
0(G).
(8)
Lemma 2.3 (see [21, Proposition 6.11]). For any g ∈ S0(G)\{0}
k · kM∞
g
: S′
0(G) → R+
0 , kσkM∞
g = sup
ν∈G× bG
(π(ν)g, σ)S0,S′
0(G)
is a norm on S′
0(G) which is equivalent to the norm in (8).
In many situations the norm convergence in S′
0 is too strong and therefore we also have to make use
of the weak∗-topology. Recall that σ0 = w∗ − limα σα for a given net (σα) in S′
0(G) if
lim
α
(f, σα − σ0)S0,S′
0(G) = 0,
for anyf ∈ S0(G).
As for every Banach space (see [22, p. 98]), also for S0(G) the Hahn-Banach Theorem provides a
isometric embedding into its double dual S′′
0(G) via the canonical embedding
ι : S0(G) → S′′
0(G), ι(f ) = σ 7→ (f, σ)S0,S′
0(G), f ∈ S0(G), σ ∈ S′
0(G).
Moreover, ι(S0(G)) ⊆ S′′
That is, a linear and bounded functional ϕ : S′
into norm convergent nets in C if and only if ϕ is of the form ϕ(σ) = (f, σ)S0,S′
(see [22, Proposition 2.6.4]). Henceforth we view, if necessary, S0(G) as a closed subspace of S′′
fact is essential for our proof of Theorem 1.3 in Section 4.
0(G) is exactly the set of all bounded weak∗ continuous functionals on S′
0(G) → C sends bounded weak∗ convergent nets in S′
0(G).
0(G)
0(G) for some f ∈ S0(G)
0(G). This
Similar as for functions, we can define the tensor product σ(1) ⊗ σ(2) of two distributions σ(1) ∈ S′
0(G1)
and σ(2) ∈ S′
0(G2). It is the unique element in S′
0(G1 × G2) with the property that
(f (1) ⊗ f (2), σ(1) ⊗ σ(2))S0,S′
0(G1×G2) = (f (1), σ(1))S0,S′
0(G1) (f (2), σ(2))S0,S′
0(G2),
for all f (i) ∈ S0(Gi), i = 1, 2. One can show that
kσ(1) ⊗ σ(2)kM∞
g(1)⊗g(2)
= kσ(1)kM∞
g(1)
kσ(2)kM∞
g(2)
.
For a proof of this we refer to [21, Corollary 9.2].
(9)
(10)
As mentioned in the introduction, the space S0(G) is embedded into its dual space S′
0(G) in a very
natural, but non-isometric way. In order to properly formulate this result we define the modulation space
(for the parameter 1) as the subspace of S′
0(G) given by
M1(G) =nσ ∈ S′
0(G) : ZG× bG
(π(ν)g, σ)S0,S′
0(G) dν < ∞o,
(11)
5
where g is some non-zero function in S0(G). In Section 3.6 we give references to literature on the modulation
spaces. The norm
k · kM1,g : M1(G) → R+
0 ,
kσkM1,g =ZG× bG
(π(ν)g, σ)S0,S′
0(G) dν
(12)
turns M1(G) into a Banach space. Each function g ∈ S0(G)\{0} induces an equivalent norm on M1(G).
≤ c kσkM1 for all σ ∈ M1(G). That is,
One can show that there exists a constant c > 0 such that kσkS′
M1(G) is continuously embedded into S′
0
0(G).
Lemma 2.4. The Banach spaces S0(G) and M1(G) are naturally isomorphic. In particular we have:
(i) Via the Haar measure on G every h ∈ S0(G) induces a (unique) functional σ = σh ∈ S′
0(G):
(f, σh)S0,S′
0(G) =ZG
f (t) h(t) dt for all f ∈ S0(G).
(13)
This embedding of S0(G) into S′
0(G) is linear, continuous and injective.
(ii) If σ is a distribution in S′
0(G), then there exists a function h ∈ S0(G) such that (13) holds if and
only if σ ∈ M1(G). The function h ∈ S0(G) is characterized by the fact that for some g ∈ S0(G)\{0}
(and then for every such g) one has:
(h, σ)S0,S′
0(G) = kgk−2
2 ZG× bG(cid:0)π(ν)g, σ(cid:1)(cid:0)π(ν)g, σ(cid:1) dν for all σ ∈ S′
0(G).
(14)
One can verify that the embeddings in Lemma 2.4(i) and (ii) are inverses of one another (independently
of the choice of the function g in (ii)). The details can be found in [21, Theorem 6.12]. If h is any function
in Lp(G), p ∈ [1, ∞], then h also induces a functional in S′
0(G) as in (13).
By the natural isomorphism between S0(G) and M1(G) the function space S0(G) is continuously
0(G) we allow ourselves,
0(G), by which we mean the action that the function h has on f as
0(G) = (h, f )S0,S′
embedded into its dual space S′
for all f, h ∈ S0(G), to write (f, h)S0,S′
in Lemma 2.4(i). Note that (f, h)S0,S′
0(G). Due to this relation between S0(G) and S′
0(G).
The function space S0 is weak∗ dense in S′
0.
0(G) there exists a net (σα) in M1(G) ∼= S0(G)
Lemma 2.5 (see [21, Proposition 6.15]). For any σ ∈ S′
such that
lim
0 = 0 for all f ∈ S0(G) and kσαkS′
The translation and modulation operator can be uniquely extended from operators on S0(G) to weak∗-
0(G). We will denote these extensions by the same symbol. Specifically,
α (cid:12)(cid:12)(f, σ − σα)S0,S′
0(G) and ν = (x, ω) ∈ G × bG, they are characterized by the following identities:
weak∗ continuous operators on S′
for f ∈ S0(G), σ ∈ S′
0 ≤ kσkS′
0.
(f, Txσ)S0,S′
(f, Eωσ)S0,S′
(f, π(x, ω) σ)S0,S′
0(G) = (T−xf, σ)S0,S′
0(G) = (Eωf, σ)S0,S′
0(G) = ω(x) (π(−x, ω)f, σ)S0,S′
0(G),
0(G),
0(G).
In addition, for g, h ∈ S0(G), we define
(f, h · σ)S0,S′
(f, g ∗ σ)S0,S′
0(G) = (f · h, σ)S0,S′
0(G) = (f ∗ gX, σ)S0,S′
0(G),
0(G), gX(t) = g(−t), t ∈ G.
These formulas remain valid for h being a pointwise multiplier of S0(G) or g having a Fourier transform
with this property (defining a bounded convolution operator on S0(G)).
6
The complex conjugation of a distribution is defined by the relation
(f, σ)S0,S′
0(G) = (f , σ)S0,S′
0(G),
The reader may verify that these definitions are compatible with the embedding of S0(G) into S′
described in Lemma 2.4 and are in fact uniquely determined based on this consistency consideration.
0(G) as
The extension of the translation operator to S′
0(G) is not the same as its Banach space adjoint, which,
by definition, is the operator given by
(Tx)× : S′
0(G) → S′
0(G), (f, (Tx)×σ)S0,S′
0(G) = (Txf, σ)S0,S′
0(G).
However, it so happens that the Banach space adjoint of the modulation operator Eω : S0(G) → S0(G) is
the same as its unique extension to an operator on S′
Throughout the paper h·, ·i is the L2-inner product (with the anti-linearity in the second entry), which
is well-defined for functions in S0(G) as S0(G) ⊆ L2(G). In fact, S0(G) is continuously embedded into all
the Lp(G) spaces: for all p ∈ [1, ∞] and f ∈ S0(G),
0(G).
kf kp ≤ kgk−1
q kf kS0(G),g,
where p−1 + q−1 = 1 for p ∈ (1, ∞) and the usual convention if p = 1 or p = ∞ (this follows from [21,
Lemma 4.19]). Furthermore S0(G) is continuously embedded into C0(G) and hence S′
0(G) contains the
Dirac delta distribution δx : f 7→ f (x), x ∈ G, f ∈ S0(G).
We will make frequent use of the following equality.
Lemma 2.6 (see [21, Lemma 6.10(iv)]). If g ∈ S0(G)\{0}, then for any f ∈ S0(G) and σ ∈ S′
0(G)
(f, σ)S0,S′
0(G) = kgk−2
2 ZG× bG
hf, π(ν)gi (π(ν)g, σ)S0,S′
0(G) dν.
(15)
Lastly, we define the short-time Fourier transform with respect to a function g ∈ S0(G) to be the
operator
Vg : S′
0(G) → Cb(G × bG), Vgσ(ν) =(cid:0)π(ν)g, σ(cid:1)S0,S′
0(G)
for all σ ∈ S′
0(G), ν ∈ G × bG.
The operator maps L2(G) into L2(G × bG) and it maps S0(G) into S0(G × bG) (see [16, Section 6] or [21,
Theorem 5.3(ii)]. Note that if f ∈ L2(G), then Vgf (ν) = hf, π(ν)gi, ν ∈ G × bG. Using the short-time
Fourier transform we can reformulate (15) as kgk2
2 (f, σ)S0,S′
0(G) =RG× bG Vgf (ν)Vgσ(ν) dν.
3 Operators that have a kernel in S0
3.1 Nets versus sequences
The spaces of operators that are identified with S0(G1 × G2) by Theorem 1.3 are uniquely extended to S′
0
using weak∗ continuity in S′
0 is non-metrizable (unless S0 is finite dimensional,
[22, Proposition 2.6.12]) and it is therefore properly described using nets. However, in some cases, e.g., if
G = Rd, we may use the notion of sequences to describe the spaces A and B.
0. The weak∗ topology on S′
Lemma 3.1. If G1 and G2 are σ-compact and metrizable, then the Banach spaces A(G1, G2) and B(G1, G2)
can be described by the behavior of convergent sequences. Specifically,
A(G1, G2) = {A ∈ Bil(S′
B(G1, G2) = {T ∈ Lin(S′
convergent sequences in S0(G2)}.
0(G1) × S′
0(G1), S0(G2)) : T maps weak∗-convergent sequences in S′
0(G2), C) : A is sequentially weak∗ continuous in each coordinate}.
0(G1) into norm
7
σ-compact and metrizable [3, Section 3]. It is a fact that S0 can be described as a coorbit space associated
Proof. If a locally compact Abelian group G is σ-compact and metrizable then also its dual group bG is
to the Heisenberg representation of G × bG [15]. Coorbit theory [14, Theorem 6.1], together with the
fact that the time-frequency plane G × bG is σ-compact, implies the separability of S0(G). Thus, by the
assumption in the lemma, the spaces S0(Gi), i = 1, 2 are separable. The Banach-Alaoglu theorem implies
that the weak∗ topology on S′
0 on any bounded set is metrizable. Hence the notions of continuity by
bounded convergent nets and convergent sequences coincide.
Note that the commonly used locally compact Abelian groups R, Z, T, Z/N Z N = 1, 2, . . . and the
p-adic numbers are σ-compact and metrizable. The additive group R under the discrete topology is an
example of a non-σ-compact (albeit metrizable) locally compact Abelian group.
3.2 Identifying operators that have a kernel in S0
0(G1 × G2).
In this section we answer the second question posed in the introduction, which we expand on here. If
T is an operator in Lin(S0(G1), S′
0(G2)), then the outer kernel theorem implies that T has a kernel κ(T )
in S′
It may happen that this kernel is induced by a function in S0(G1 × G2). By the
inner kernel theorem we know that these operators are exactly the ones that belong to B(G1, G2) ⊆
Lin(S′
0(G1), S0(G2)). However, it is not immediately clear how (a) we verify that the domain of the
operator T can be extended from S0(G1) to S′
0(G1), (b) that its co-domain actually is S0(G2) rather
than S′
0(G2) and (c) how we can verify its continuity properties as described in Definition 1.2. Of course
we have similar issues for operators A ∈ Bil(S0(G1) × S0(G2), C) whose kernel might be induced by a
function in S0(G1 × G2). The following theorem characterizes in a quantitative way the operators in
Lin(S0(G1), S′
0(G2)) and Bil(S0(G1) × S0(G2), C) that have a kernel in S0(G1 × G2) and it describes how
their domain extends from S0 to S′
0.
Theorem 3.2. For i = 1, 2 fix a function g(i) ∈ S0(Gi)\{0} such that kg(i)k2 = 1.
(i) If A is an operator in Bil(S0(G1) × S0(G2), C), then its kernel κ(A) ∈ S′
0(G1 × G2) is induced by a
function in S0(G1 × G2), i.e. A ∈ A(G1, G2), if and only if
Z
(cid:12)(cid:12)A(cid:0)π(ν(1))g(1), π(ν(2))g(2)(cid:1)(cid:12)(cid:12) d(ν(1), ν(2)) < ∞.
G1× bG1×G2× bG2
In that case the operator A : S′
0(G1) × S′
0(G2) → C satisfies
A(σ(1), σ(2))
=
Z
G1× bG1×G2× bG2
Vg(1)σ(1)(ν(1)) · Vg(2)σ(2)(ν(2)) · A(cid:0)π(ν(1))g(1), π(ν(2))g(2)(cid:1) d(ν(1), ν(2)).
(16)
(17)
(ii) If T is an operator in Lin(cid:0)S0(G1), S′
0(G2)(cid:1), then its kernel κ(T ) ∈ S′
function in S0(G1 × G2), i.e. T ∈ B(G1, G2), if and only if
0(G1 × G2) is induced by a
Z
(cid:12)(cid:12)(cid:0)π(ν(2))g(2), T ◦ π(ν(1))g(1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12) d(ν(1), ν(2)) < ∞.
G1× bG1×G2× bG2
(18)
In that case the operators T : S′
0(G1) → S0(G2) satisfies
(T σ(1), σ(2))S0,S′
0(G2)
=
Z
G1× bG1×G2× bG2
Vg(1)σ(1)(ν(1)) · Vg(2)σ(2)(ν(2)) ·(cid:0)π(ν(2))g(2), T π(ν(1))g(1)(cid:1)S0,S′
0
d(ν(1), ν(2)).
(19)
8
Remark 1. The formula in (17) extends the domain of A from S0(G1) × S0(G2) to S′
(19) extends the domain of T from S0(G1) to S′
Remark 2. The condition in Theorem 3.2 that kg(i)k2 = 1 is only necessary to make the equalities in (17)
and (19) more pleasant. Otherwise the integrals need to be normalized by kg(1) ⊗ g(2)k−2
2 , see the details
in the proof.
0(G1) × S′
0(G2), and
0(G1).
Proof of Theorem 3.2. We will only prove (i) as the proof of (ii) is similar. By Theorem 1.1 and by
assumption we know that A has a kernel κ(A) ∈ S′
0(G1 × G2) so that
G1× bG1×G2× bG2
Z
Z
Z
(cid:12)(cid:12)A(cid:0)π(ν(1))g(1), π(ν(2))g(2)(cid:1)(cid:12)(cid:12) d(ν(1), ν(2))
(cid:12)(cid:12)(cid:0)π(ν(1))g(1) ⊗ π(ν(2))g(2), κ(A)(cid:1)S0,S′
(cid:12)(cid:12)(cid:0)Eω(1),ω(2)Tx(1),x(2)(g(1) ⊗ g(2)), κ(A)(cid:1)S0,S′
=
=
G1× bG1×G2× bG2
G1× bG1×G2× bG2
0(G1×G2)(cid:12)(cid:12) d(ν(1), ν(2)).
0(G1×G2)(cid:12)(cid:12) d(x(1), ω(1), x(2), ω(2)).
By Lemma 2.4 the last integral is finite if and only if the distribution κ(A) ∈ S′
0(G1 × G2) is induced by a
(unique) function in S0(G1 × G2), which we shall also call κ(A). By Theorem 1.3 this kernel is identifiable
with an operator A ∈ A ⊆ Bil(S′
0(G1) × S′
0(G2), C) which satisfies
A(σ(1), σ(2)) = (κ(A), σ(1) ⊗ σ(2))S0,S′
0(G1×G2).
By use of Lemma 2.6 (with g = g(1) ⊗ g(2), f = κ(A), σ = σ(1) ⊗ σ(2)) we can establish the desired equality.
(κ(A), σ(1) ⊗ σ(2))S0,S′
0(G1×G2)
Z
= kg(1) ⊗ g(2)k−2
2
(κ(A), π(ν(1))g(1) ⊗ π(ν(2))g(2))S0,S′
0(G1×G2)
G1× bG1×G2× bG2
· (π(ν(1))g(1) ⊗ π(ν(2))g(2), σ(1) ⊗ σ(2))S0,S′
0(G1×G2) d(ν(1), ν(2))
=
Z
G1× bG1×G2× bG2
A(cid:0)π(ν(1))g(1), π(ν(2))g(2)(cid:1)
· (π(ν(1))g(1), σ(1))S0,S′
0(G1) (π(ν(2))g(2), σ(2))S0,S′
0(G2) d(ν(1), ν(2)).
In the introduction we stated that the spaces A(G1, G2) and B(G1, G2) are Banach spaces with respect
to their subspace topologies which they naturally inherit from Bil(S′
respectively. It is clear that the induced norms fail to capture the continuity requirements for operators
in A(G1, G2) and B(G1, G2) as described in Definition 1.2. Hence the induced norm on A(G1, G2) can not
distinguish between operators in Bil(S′
0(G2)) that belong to A(G1, G2) and those that do not.
Similarly the norm on Lin(S′
0(G1), S0(G2)) can not detect if an operator actually belongs to B(G1, G2) or
not. The results from Theorem 3.2 show how we can define a norm on the spaces A(G1, G2) and B(G1, G2)
that exactly captures operators with a kernel in S0(G1 × G2).
Corollary 3.3. For i = 1, 2 fix a function g(i) ∈ S0(Gi)\{0}.
0(G2), C) and Lin(S′
0(G1) × S′
0(G1)×S′
0(G1), S0(G2)),
(i)
k · kA : A(G1, G2) → R+
0 ,
kAkA =
Z
(cid:12)(cid:12)A(cid:0)π(ν(1))g(1), π(ν(2))g(2)(cid:1)(cid:12)(cid:12) d(ν(1), ν(2)), A ∈ A(G1, G2),
G1× bG1×G2× bG2
defines a norm on A(G1, G2). This norm is equivalent to the subspace norm on A(G1, G2) induced
by the space Bil(S′
0(G1) × S′
0(G2), C).
9
(ii)
k · kB : B(G1, G2) → R+
0 ,
kT kB =
Z
(cid:12)(cid:12)(cid:0)π(ν(2))g(2), T π(ν(1))g(1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12) d(ν(1), ν(2)), T ∈ B(G1, G2),
G1× bG1×G2× bG2
defines a norm on B(G1, G2). This norm is equivalent to the subspace norm on B(G1, G2) induced
by the space Lin(S′
0(G1), S0(G2)).
Remark 3. Theorem 3.2 implies that the integrals used to define the norms in Corollary 3.3(i) and (ii) are
finite exactly when A and T belong to A(G1, G2) and B(G1, G2), respectively.
We can use (19) and Lemma 2.6 with respect to the time-frequency plane G2 × bG2 to show that an
operator in B(G1, G2) is uniquely determined by its action on all time-frequency shifts of a given function
in S0(G1).
Corollary 3.4. Fix g ∈ S0(G1)\{0}. An operator in B(G1, G2) is uniquely determined by its action on
the set {π(ν)g : ν ∈ G1 × bG1}. Specifically, for all T ∈ B(G1, G2) and σ(i) ∈ S′
Vgσ(1)(ν) · (cid:0)T π(ν)g, σ(2)(cid:1)S0,S′
2 ZG1× bG1
(T σ(1), σ(2)) = kgk−2
0(G2) dν.
0(Gi), i = 1, 2,
We refer to Corollary 3.4 by saying that the following identity holds true in the weak sense:
T σ = kgk−2
2 ZG1× bG1
Vgσ(ν) · T π(ν)g dν for all σ ∈ S′
0(G1).
(20)
3.2.1 A note on operator with kernel in S′
0
The results of Theorem 3.2 and Corollaries 3.3 and 3.4 are not restricted to the operators in A(G1, G2) ∼=
B(G1, G2), but can be formulated in a very similar form for the much larger spaces of operators that have
a kernel in S′
0(G1 × G2) (by use of the outer rather than the inner kernel theorem and Lemma 2.6). For
operators in Lin(S0(G1), S′
0(G2)) they take the following form.
Proposition 3.5. Given f (i), g(i) ∈ S0(Gi), g(i) 6= 0 for i = 1, 2 one has for any operator T ∈ Lin(S0(G1), S′
0(G2)):
kg(1) ⊗ g(2)k2
2 · (f (2), T f (1))S0,S′
0(G2)
(i)
(ii)
T f (1) = kg(1)k−2
=
Z
G1× bG1×G2× bG2
Vg(1)f (1)(ν(1)) · Vg(2)f (2)(ν(2)) ·(cid:16)π(ν(2))g(2), T ◦ π(ν(1))g(cid:17)S0,S′
2 ZG1× bG1
Vg(1)f (1)(ν(1)) · T(cid:0)π(ν(1))g(cid:1) dν(1),
d(ν(1), ν(2)),
0(G2)
(iii) and
k · kB′ : Lin(S0(G1), S′
0(G2)) → R+
0 ,
kT kB′ =
sup
ν(1)∈G1× bG1
ν(2)∈G2× bG2
(π(ν(2))g(2), T π(ν(1))g(1))S0,S′
0(G2)
defines a norm on Lin(S0(G1), S′
0(G2)) which is equivalent to the usual operator norm.
This result has the following consequence:
10
Corollary 3.6. For i = 1, 2 take g(i) ∈ S0(Gi)\{0}. Every continuous and bounded function F ∈ Cb(G1 ×
bG1 × G2 × bG2) defines a linear and bounded operator T : S0(G1) → S′
0(G2) via
(f (2), T f (1))S0,S′
0(G2) =
Vg(1)f (1)(ν(1)) · Vg(2)f (2)(ν(2)) · F (ν(1), ν(2)) d(ν(1), ν(2)).
(21)
Z
G1× bG1×G2× bG2
Conversely, for every linear and bounded operator T : S0(G1) → S′
0(G2) there exists a (non-unique)
function F ∈ Cb(G1 × bG1 × G2 × bG2) (which also depends on g(i)) such that (21) holds. Moreover, the
function F can be taken to be in L1(G1 × bG1 × G2 × bG2) if and only if T has a kernel in S0(G1 × G2).
0(G2)) where the operators
are represented by bounded and continuous functions (rather than abstract functionals as in the outer
kernel theorem).
This shows that it is possible to have a calculus for operators in Lin(S0(G1), S′
3.2.2 A note on Gabor frames
Recall that a function g ∈ S0(G) generates a Gabor frame for L2(G) with respect to a closed subgroup
Λ in G × bG (typically Λ is a discrete and co-compact subgroup, a lattice, in the time-frequency plane) if
there exist constants A, B > 0 such that
A kf k2
dλ ≤ B kf k2
2 ∀ f ∈ L2(G).
2 ≤ZΛ(cid:12)(cid:12)(cid:10)f, π(λ)g(cid:11)(cid:12)(cid:12)2
=ZΛ
(22)
(23)
In the positive case there exists a (not necessarily unique) function h ∈ S0(G) such that
(f, σ)S0,S′
0
hf, π(λ)gi (π(λ)h, σ) dλ ∀f ∈ S0(G), σ ∈ S′
0(G).
We will not go into details of how pairs of function g and h can be found or be characterized so that (23)
holds. For general Gabor and time-frequency analysis we refer to [5, 18] and [20].
We have already encountered a Gabor frame for L2(G) with respect to the subgroup G×bG: Lemma 2.6
shows that any non-zero function g ∈ S0(G) generates a Gabor frame for L2(G) with respect to Λ = G ×bG
(in (15) take σ to be induced by f ; since (15) holds for all f ∈ S0(G), which is dense in L2(G), it follows
that (22) is satisfied and that A = B = kgk2
2). In this case, if g is any other function in S0(G) such that
hg, gi 6= 0, then the pair (g, h), where h = (hg, gi)−1g satisfies (23).
Using (23) rather than (15) for the proofs of Section 3.2 leads to the following results for the operators
in B(G1, G2) (we leave the formulation of the corresponding results for A(G1, G2) to the reader).
Theorem 3.7. For i = 1, 2 let g(i) and h(i) be functions in S0(Gi) such that they generate Gabor frames
for L2(Gi) with respect to a closed subgroup Λi in Gi × bGi and such that (23) holds.
k · kB,g(1),g(2) : B(G1, G2) → R+
0 ,
(i)
kT k =ZΛ1×Λ2(cid:12)(cid:12)(cid:0)π(λ(2))g(2), T π(λ(1))g(1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12) d(λ(1), λ(2)), T ∈ B(G1, G2),
defines a norm on B(G1, G2). This norm is equivalent to the subspace norm on B(G1, G2) induced
by the space Lin(S′
only if T ∈ B(G1, G2).
0(G2)(cid:1) the norm is finite if and
0(G1), S0(G2)). For an operator T ∈ Lin(cid:0)S0(G1), S′
0(G2)(cid:1) with kernel κ(T ) ∈ S0(G1 × G2), then
(ii) Given T ∈ Lin(cid:0)S0(G1), S′
(T σ(1), σ(2))S0,S′
0(G2)
=ZΛ1×Λ2
Vg(1)σ(1)(λ(1)) · Vg(2)σ(2)(λ(2)) ·(cid:0)π(λ(2))h(2), T π(λ(1))h(1)(cid:1)S0,S′
0(G2) d(λ(1), λ(2)).
11
(iii) Any T ∈ B(G1, G2) satisfies
(T σ(1), σ(2)) =ZΛ1
Vg(1)σ(1)(λ(1)) · (T π(λ(1))h(1), σ(2))S0,S′
0(G2) dλ(1).
We can be much more concrete if G = Rn. In this case it is known that for any n ∈ N the Gaussian
function g(n)(x) = e−πx·x, x ∈ Rn generates a Gabor frame for L2(Rn) with respect to the lattice Λ =
aZ2n ⊂ R2n whenever 0 < a < 1. Hence in this case the integrals in Theorem 3.7 become a sum over
lattice points. In particular, any linear and bounded operator T from S0(Rn) into S′
0(Rm) has a kernel in
S0(Rn+m) if and only if
Xλ(n)∈aZ2n
λ(m)∈aZ2m
(cid:12)(cid:12)(cid:0)π(λ(m))g(m), T π(λ(n))g(n)(cid:1)S0,S′
0(Rm)(cid:12)(cid:12) < ∞.
3.3 Analogies with linear algebra
If A is an n2 × n1 matrix, then it defines an operator bA from Cn1 into Cn2,
bA : Cn1 → Cn2, bA(v(1)) = A · v(1), v(1) ∈ Cn1.
Conversely, if a linear operator bA from Cn1 into Cn2 is given and we use the standard basis for these spaces,
then the matrix representation of bA is
and then bA(v(1)) = A · v(1). If a matrix A is as above and if we let a matrix B ∈ Cn3×n2 define an operator
bB from Cn2 into Cn3, then their composition, bB ◦ bA, is represented by the product of the two matrices.
i = 1, . . . , n2, j = 1, . . . , n2
A(i, j) = (A(ej))⊤ · ei,
That is,
(24)
bB ◦ bA : Cn1 → Cn3, bB ◦ bA(v(1)) = B · A · v(1), v(1) ∈ Cn1,
(25)
according to usual matrix multiplication: (B · A)(i, j) =
B(i, k) · A(k, j), i = 1, . . . , n3, j = 1, . . . , n1.
n2Xk=1
The next two results show that the inner kernel theorem allows us to extend both (24) and (25) from
matrices to operators in B(G1, G2). In particular, the role of the unit vectors in Cn are replaced by the
Dirac delta distribution, δx : S0(G) → C, δx : f 7→ f (x), f ∈ S0(G), x ∈ G. If Gi = Z/niZ, i = 1, 2, then
the results reduce to the matrix case.
Lemma 3.8. Given T ∈ B(G1, G2) and A ∈ A(G1, G2) its kernel satisfies for x(i) ∈ Gi, i = 1, 2:
κ(T )(x(1), x(2)) = (T δx(1), δx(2))S0,S′
0(G2) and κ(A)(x(1), x(2)) = A(δx(1), δx(2)),
Proof. It is easy to verify the equality δx(1),x(2) = δx(1) ⊗ δx(2), since on has obviously
(cid:0)f (1) ⊗ f (2), δx(1) ⊗ δx(2)(cid:1)S0,S′
0(G1×G2) = f (1)(x(1)) · f (2)(x(2)) = (f (1) ⊗ f (2))(x(1), x(2)).
The desired result now follows from the inner kernel theorem:
κ(T )(x(1), x(2)) =(cid:0)κ(T ), δx(1),x(2)(cid:1)S0,S′
=(cid:0)κ(T ), δx(1) ⊗ δx(2)(cid:1)S0,S′
0(G1×G2) =(cid:0)T δx(1), δx(2)(cid:1)S0,S′
0(G1×G2)
0(G2).
The equality for the kernel of A follows in the same fashion.
The role of the "delta-basis" in Lemma 3.8 can also be taken by a continuous Gabor frame (cf. Section
3.2.2).
12
Corollary 3.9. For i = 1, 2 let g(i) ∈ S0(Gi)\{0} and x(i) ∈ Gi.
(i) For T ∈ B(G1, G2) one has
κ(T )(x(1), x(2)) = kg(1)k−2
2 ZG1× bG1(cid:0)π(ν(1))g(1)(cid:1)(x(1)) ·(cid:0)T ◦ π(ν(1))g(1)(cid:1)(x(2)) dν(1).
(ii) For A ∈ A(G1, G2) one has
κ(A)(x(1), x(2)) = kg(1) ⊗ g(2)k−2
2
Z
G1× bG1G2× bG2
(cid:0)π(ν(1))g(1)(cid:1)(x(1))(cid:0)π(ν(2))g(2)(cid:1)(x(2))
· A(cid:0)π(ν(1)g(1), π(ν(2)g(2)(cid:1) d(ν(1), ν(2)).
Proof. Combine equality (19) of Theorem 3.2, Corollary 3.4 with Lemma 3.8.
The composition rule of operators represented by matrices has the following analogous continuous
formulation for operators with kernel in S0.
Lemma 3.10. If T1 ∈ B(G1, G2) and T2 ∈ B(G2, G3), then T2 ◦ T1 ∈ B(G1, G3) and
κ(T2 ◦ T1)(x(1), x(3)) =ZG2
κ(T1)(x(1), x(2)) · κ(T2)(x(2), x(3)) dx(2), x(i) ∈ Gi, i = 1, 3.
Moreover, using the norm on B as defined in Corollary 3.3, there exists a constant c > 0 such that
kT2 ◦ T1kB ≤ c kT2kB kT1kB.
Corollary 3.11. The Banach space (B(G, G), k · kB) forms a Banach algebra under composition.
Proof of Lemma 3.10. Let us first show that the integral is well-defined. By Lemma 2.1 we can write
κ(T1) =Xj∈N
f (1)
j ⊗ f (2)
j
and κ(T2) =Xj∈N
h(2)
j ⊗ h(3)
j
for suitable f (i)
j ∈ S0(Gi), i = 1, 2 and h(i)
j ∈ S0(Gi), i = 2, 3 and where j ∈ N. Furthermore,
Xj∈N
kf (1)
j kS0 kf (2)
j kS0 < ∞ and Xj∈N
kh(1)
j kS0 kh(2)
j kS0 < ∞.
Because S0(G) is continuously embedded into L2(G) and into L∞(G) the following estimate applies: for
all x(i) ∈ Gi, i = 1, 2, 3
f (1)
j
k (x(2)) · h(3)
k (x(3)) dx(2)
(x(2)) · h(2)
j (x(1)) · f (2)
ZG2(cid:12)(cid:12)κ(T1)(x(1), x(2)) · κ(T2)(x(2), x(3))(cid:12)(cid:12) dx(2)
≤ZG2 Xj,k∈N
≤ Xj,k∈N
≤ Xj,k∈N
≤ c Xj,k∈N
k k∞ ZG2
k k∞ kf (2)
j k2 kh(2)
j k∞ kh(3)
j kS0 kh(3)
k kS0 kf (2)
j kS0 kh(2)
kf (1)
j k∞ kh(3)
(x(2)) · h(2)
kf (1)
kf (1)
f (2)
j
k k2
k kS0 < ∞,
k (x(2)) dx(2)
13
for some c > 0. This shows that the integral and thus the function κ(T2 ◦ T1) : G1 ×G3 → C is well-defined.
Note that
κ(T2 ◦ T1)(x(1), x(3)) =ZG2
κ(T1)(x(1), x(2)) κ(T2)(x(2), x(3)) dx(2)
=ZG2 Xj,k∈N
= Xj,k∈N
(f (2)
j
f (1)
j
(x(1)) f (2)
j (x(2)) h(2)
k (x(2)) h(3)
k (x(3)) dx(2)
, h(2)
k )S0,S′
0(G2)(cid:0)f (1)
j ⊗ h(3)
k (cid:1)(x(1), x(3)).
Hence κ(T1 ◦ T2) =Pj,k∈N(f (2)
j
, h(2)
k )S0,S′
0(G2) f (1)
j ⊗ h(3)
k . The above calculation shows that
Xj,k∈N
k(f (2)
j
, h(2)
k ) f (1)
j kS0 kh(3)
k kS0 ≤ ∞.
Hence κ(T2 ◦ T1) ∈ S0(G1) ⊗S0(G3). By Lemma 2.1 this implies that κ(T2 ◦ T1) ∈ S0(G1 × G3) as well as
the moreover-part of the lemma. Let us show that the function which we defined as κ(T2 ◦ T1) indeed is
the kernel of the operator T2 ◦ T1: if σ(i) ∈ S′
0(Gi), i = 1, 3, then
(T2 ◦ T1σ(1), σ(3))S0,S′
0(G3) = (κ(T2), T1σ(1) ⊗ σ(3))S0,S′
0(G2×G3)
(h(2)
k ⊗ h(3)
k , T1σ(1) ⊗ σ(3))S0,S′
0(G2×G3)
(T1σ(1), h(2)
k )S0,S′
0(G2) (h(3)
k , σ(3))S0,S′
0(G3)
(κ(T1), σ(1) ⊗ h(2)
k )S0,S′
0(G1×G2) (h(3)
k , σ(3))S0,S′
0(G3)
(f (1)
j ⊗ f (2)
j
, σ(1) ⊗ h(2)
k )S0,S′
0(G1×G2) (h(3)
k , σ(3))S0,S′
0(G3)
(f (2)
j
, h(2)
k )S0,S′
0(G2) f (1)
j ⊗ h(3)
k , σ(1) ⊗ σ(2)(cid:17)S0,S′
.
0(G1×G3)
=Xk∈N
=Xk∈N
=Xk∈N
= Xj,k∈N
=(cid:16) Xj,k∈N
3.4 Some examples, nuclearity and trace-class results
Example 3.12. The prototypical example of an element in S0(G1 × G2) is the tensor-product function
f = f (1) ⊗ f (2), f (i) ∈ S0(Gi), i = 1, 2. It is not difficult to show that the unique corresponding operators
A, T and S according to Theorem 1.3 are the following ones:
A : S′
T : S′
S : S′
0(G1) × S′
0(G1) → S0(G2), T (σ(1)) = (f (1), σ(1))S0,S′
0(G2) → S0(G1), S(σ(2)) = (f (2), σ(2))S0,S′
0(G2) → C, A(σ(1), σ(2)) = (f (1), σ(1))S0,S′
0(G1) · f (2),
0(G2) · f (1),
0(G1) (f (2), σ(2))S0,S′
0(G2),
where σ(i) ∈ S′
0(Gi), i = 1, 2.
Observe that the range of the operators T and S in Example 3.12 is one-dimensional. Naturally, not
all rank-one operators have a kernel in S0. For example, let f (1) be a function in S0(G1) and let f (2) be a
function in L2(G2) which is not also in S0(G2), then
T (σ(1)) = (f (1), σ(1))S0,S′
0(G1) · f (2)
is a bounded rank-one operator from S′
Indeed, if we restrict T to an operator from S0(G1) into L2(G2) ⊆ S′
0(G1) into L2(G2) which does not have a kernel in S0(G1 × G2).
0(G2), then (by the outer kernel
14
theorem) its kernel in S′
show that the operator
0(G1 × G2) is the functional induced by the function f 1 ⊗ f 2. Similarly, one can
S(h(2)) = hf (2), h(2)iL2(G2) · f (1), h(2) ∈ L2(G2),
is a linear and bounded rank-one operator from L2(G2) into S0(G1) with kernel f 2 ⊗ f 1.
Remark 4. By Lemma 2.1 every f ∈ S0(G1 × G2) has a representation
f =Xj∈N
f (1)
j ⊗ f (2)
j
such that Xj∈N
kf (1)
j kS0 kf (2)
j kS0 < ∞.
The inner kernel theorem implies that the corresponding operator T ∈ B(G1, G2) satisfies
T σ(1) =Xj∈N
(f (1)
j
, σ(1))S0,S′
0(G1) · f (2)
j
for all σ(1) ∈ S′
0(G1),
where the sum is absolutely convergent in the operator norm of Lin(S′
0(G1), S0(G2)).
This immediately leads to the following.
Corollary 3.13. Finite-rank operators in B(G1, G2) are norm dense in B(G1, G2), because any T ∈
B(G1, G2) can be written as absolutely convergent series of rank one operators (Tn)n∈N in B(G1, G2), i.e.
(cid:13)(cid:13)T −Xn∈N
Tn(cid:13)(cid:13)op,S′
0→S0
= 0 and Xn∈N
kTnkop,S′
0→S0 < ∞.
Because S0 is dense in L2 it follows that the finite-rank operators of B(G1, G2) are dense in the space
of Hilbert-Schmidt operators from L2(G1) into L2(G2).
Corollary 3.14. All operators in B(G1, G2) are nuclear operators from the Banach space S′
Banach space S0(G2).
0(G1) into the
Proof. By [27, Chapter III, §7] all nuclear operators from the Banach space S′
S0(G2) are of the form
0(G1) into the Banach space
T : S′
0(G1) → S0(G2), T σ =Xj∈N
ψ(1)
j (σ(1)) · f (2)
j
,
where (ψ(1)
j ) is a sequence in S′′
0(G1) and (f (2)
j
) is a sequence in S0(G2) such that
Xj∈N
kψ(1)
j kS′′
0 kf (2)
j kS0 < ∞.
Remark 4 combined with the fact that S0(G1) is continuously embedded into S′′
embedding implies that all operators in B(G1, G2) are nuclear.
0(G1) via the natural
By the embedding of S0 into S′
are also nuclear operators from S0(G1) into S0(G2); from S0(G1) into S′
0 as described in Lemma 2.4 it follows that all the operators in B(G1, G2)
0(G2).
0(G2); and from S′
0(G1) into S′
Corollary 3.15. An operator T in B(G, G) with kernel κ(T ) ∈ S0(G × G) is a trace-class operator on
both S0(G) and S′
0(G). Its trace satisfies tr(T ) =RG κ(T )(x, x) dx.
Proof. Following [26] we say that an operator T on a Banach space B is of trace class if it has the form
T : B → B, T x =Xj∈N
(x, σj)B,B′ · bj
for all x ∈ B,
for some suitable sequences (σj) in B′ and (bj) in B such that Pj kσjkB′ kbjkB < ∞. The trace of T is
tr(T ) = Pj∈N(bj, σj)B,B′. Remark 4 shows that for B = S0(G) or B = S′
have the desired form. Here we also use that S0 is continuously embedded into S′
0(G) the operators in B(G, G)
0 via Lemma 2.4
0 and S′′
15
and the canonical embedding ι : S0 → S′′
will first prove that the integral is well-defined. The inner kernel theorem states that κ(T ) is a function
in S0(G × G). Observe that the diagonal {(x, y) ∈ G × G : x = y} is a closed subgroup of G × G. It is a
fact (see [10, Theorem 7] or [21, Theorem 5.7]) that the restriction of an S0 function to a closed subgroup
again belongs to S0 of that subgroup. In our case this means that x 7→ κ(T )(x, x) is a function in S0(G)
0, respectively. Before we show that tr(T ) =RG κ(T )(x, x) dx we
and, in particular, that it is integrable. Now, since κ(T ) =Pj∈N f (1)
j ⊗ f (2)
j
,
j
tr(T ) =Xj∈N
=ZGXj∈N
0(G) =Xj∈NZG
(x) dx =ZG
(f (1)
, f (2)
j
)S0,S′
f (1)
j
(x) f (2)
j
(x) dx
f (1)
j
(x) f (2)
j
κ(T )(x, x) dx.
0, w∗)′ ∼= S0 it is reasonable to extend the
Remark 5. Since S0 is continuously embedded into S′
definition of trace-class operator from [26] used in the proof of Corollary 3.15 as follows: We say a (linear
and continuous) operator from S′
0(G) with the weak∗ topology into S0(G) with its norm topology is of
trace-class if
0 and (S′
T : S′
0(G) → S0(G), T σ =Xj∈N
(f (1)
j
, σ)S0,S′
0
· f (2)
j
for all σ ∈ S′
0(G),
for some suitable sequences (f (i)
operator is then tr(T ) = Pj∈N(f (1)
j
coincides exactly with the trace-class operators defined in this way.
j ) in S0(G) and such that Pj kf (1)
j k < ∞. The trace of such an
In that case, it is clear from Remark 4 that B(G, G)
j k kf (2)
, f (2)
.
)S0,S′
0
j
Let us consider another important example of elements in B(G, G).
Example 3.16 (Product-convolution operators). For any two functions h1 and h2 in S0(G) the product-
convolution operator
P Ch1,h2 : S′
0(G) → M1(G) ∼= S0(G), P Ch1,h2(σ) = (σ · h1) ∗ h2,
and the convolution-product operator
CPh1,h2 : S′
0(G) → M1(G) ∼= S0(G), CPh1,h2(σ) = (σ ∗ h1) · h2,
are linear and bounded operators, which send norm bounded weak∗ convergent nets in S′
0(G) into norm
convergent nets in M1(G) ∼= S0(G). That is, both operators belong to B(G, G). One can show that
κ(P Ch1,h2) = τ1(h1 ⊗ h2) and κ(CPh1,h2) = τ2(h1 ⊗ h2), where
τ1 : S0(G × G) → S0(G × G), τ1(f )(s, t) = f (s, t − s),
τ2 : S0(G × G) → S0(G × G), τ2(f )(s, t) = f (t − s, t).
Product-convolution operators can be used to prove Lemma 2.5 (see [21, Proposition 6.15] for the
details). The kernel theorems translate Lemma 2.5 into a statement for operators:
Lemma 3.17. For any operator T ∈ Lin(S0(G1), S′
bounded in Lin(S0(G1), S′
0(G2)) such that, for all f (i) ∈ S0(Gi), i = 1, 2,
0(G2)) there exists a net of operators (Tα) in B(G1, G2),
lim
α (cid:12)(cid:12)(cid:0)f (2), (T − Tα)f (1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12) = 0, kTαkop,S0→S′
0 ≤ kT kop,S0→S′
0.
Similar to Lemma 3.17, the inner kernel theorem can be used to translate Lemma 2.2 from a statement
of S0 to a statement of B(G1, G2).
16
Proposition 3.18. Let T0 be a non-trivial operator in B(G1, G2). The operators T in B(G1, G2) are
exactly those of the form
T =Xj∈N
cj π(ν(2)
j ) ◦ T0 ◦ π(ν(1)
j ),
where c ∈ ℓ1(N) and (ν(i)
furthermore kT k = inf kck1, where the infimum is taken over all possible representations of T as above,
defines an equivalent norm on B(G1, G2).
j ) are sequences in Gi × bGi, i = 1, 2. The sum converges in B(G1, G2) and
A similar statement is true for the space A(G1, G2). In that case, if A0 is a non-trivial element in
A(G1, G2), then all operators in A(G1, G2) are exactly those of the form
A(σ(1), σ(2)) =Xj∈N
with c and ν(i)
j
in Proposition 3.18.
cj A0(cid:0)π(ν(1)
j )σ(1), π(ν(2)
j )σ(2)(cid:1)
Proof of Proposition 3.18. By Lemma 2.2 we know that for any T ∈ B(G1, G2) there exists a sequence
c ∈ ℓ1(N) and a sequence (x(1)
j
, x(2)
j
, ω(1)
, ω(2)
j
j ) in G1 × G2 × bG1 × bG2 such that
cjEω(1)
j
,ω(2)
j
Tx(1)
j
,x(2)
j
κ(T0).
κ(T ) =Xj∈N
Hence
(T σ(1), σ(2))S0,S′
0(G2) = (κ(T ), σ(1) ⊗ σ(2))S0,S′
0(G1×G2)
=Xj∈N
=Xj∈N
=Xj∈N
=Xj∈N
j
j
j
j
,ω(2)
Tx(1)
cj(cid:0)Eω(1)
cj(cid:0)κ(T0), [T−x(1)
cj(cid:0)T0 ◦ T−x(1)
}
ω(1)
j (x(1)
∈ℓ1(N)
{z
j ) cj
j
,x(2)
j
κ(T0), σ(1) ⊗ σ(2)(cid:1)S0,S′
Eω(1)
j
σ(1)] ⊗ [T−x(2)
j
Eω(2)
j
0(G1×G2)
σ(2)](cid:1)S0,S′
0(G1,G2)
Eω(1)
j
σ(1), T−x(2)
j
Eω(2)
j
σ(2)(cid:1)S0,S′
0(G2)
j
(cid:0) Eω(2)
{z
=:π(ν(2)
)
j
}
Tx(2)
j
◦T0 ◦ Eω(1)
j
T−x(1)
j
=:π(ν(1)
)
{z
j
}
σ(1), σ(2)(cid:1)S0,S′
0(G2).
If we take T0 to be a rank-one operator in B(G1, G2) as in Example 3.12 with κ(T0) = f (1) ⊗ f (2), then
Proposition 3.18 states that any operator T ∈ B(G1, G2) has the form
T σ(1) =Xj∈N
cj(cid:0)π(ν(1)
j )f (1), σ(1)(cid:1)S0,S′
0(G1) π(ν(2)
j )f (2)
for all σ(1) ∈ S′
0(G1), for some suitable sequence c ∈ ℓ1(N) and sequences (νi
j) in Gi × bGi, i = 1, 2.
3.5 Regularizing approximations of the identity
Since S0 is weak∗ dense in S′
operator T in Lin(S0(G1), S′
in the weak∗sense towards κ(T ). The associated operators Tα ∈ B(G1, G2) satisfy
0 it is possible to approximate the kernel κ(T ) ∈ S′
0(G1 × G2) of a general
0(G2)) by a net (or sequence) of functions κα in S0(G1 × G2) that converges
lim
α (cid:12)(cid:12)(cid:0)f (2), (T − Tα)f (1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12) = 0 for all f (i) ∈ S0(Gi), i = 1, 2.
17
(26)
We saw this already in Lemma 3.17. In this section we propose a construction of a net of operators (Tα)
in B(G1, G2) such that (26) holds, that is not based on the modification of the kernel per se (which is the
idea behind Lemma 3.17), but rather by a composition of the given operator T with certain operators: we
introduce the idea of a regularizing approximations of the identity.
Definition 3.19. A regularizing approximation of the identity of S0(G) is a net of operators (Tα) in
B(G, G) resp. κ(Tα) ∈ S0(G × G) for each α and which satisfies the following conditions:
(i) limα kTαf − f kS0 = 0 for all f ∈ S0(G),
(ii) supα kTαkop,S0→S0 < ∞,
(iii) supα kTαkop,S′
0→S′
0 < ∞,
(iv) limα (f, Tασ − σ)S0,S′
0(G) = 0 for all f ∈ S0(G), σ ∈ S′
0(G).
Remark 6. Statement (i) and (ii) for the adjoint operators (T ×
α ) implies (iv) and (iii), respectively. Hence
we need only conditions (i) and (ii) for self-adjoint operators. Moreover, (i) implies (ii) for the case of a
sequence of operators, due to the Banach-Steinhaus principle.
We list three examples of such families of operators at the end of this section. It is straightforward to
show that the properties of a regularizing approximation of the identity implies convergence of L2.
Lemma 3.20. If (Tα) is a regularizing approximation of the identity for S0(G), then
sup
α
kTαkop,L2→L2 < ∞ and lim
α
kTαf − f k2 = 0 for all f ∈ L2(G).
Moreover, the net (κ(Tα)) in S0(G × G) ⊆ S′
in the weak∗sense.
0(G × G) converges towards the kernel of the identity operator
Proof. The first statement follows by interpolation theory for operators and assumptions (ii) and (iii) in
Definition 3.19. Now, since S0 is continuously embedded and dense in L2, Definition 3.19(i) implies that
kTαf − f k2 → 0 for all f ∈ L2(G).
lim
α
The moreover part follows from the fact that Definition 3.19(iv) implies (26)
Regularizing approximations of the identity allow us to construct a concrete family of operators that
0(G1×G2)
have kernels in S0(G1×G2), which approximate any given operator with an (abstract) kernel in S′
in the weak∗ sense.
Proposition 3.21. For i = 1, 2 let (T (i)
any operator T ∈ Lin(S0(G1), S′
α ) be a regularizing approximation of the identity for S0(Gi). For
0(G2)) the collection of operators (Tα), Tα := T (2)
α ◦ T ◦ T (1)
α
is such that
(i) Tα ∈ B(G1, G2) for each α, i.e., κ(Tα) ∈ S0(G1 × G2),
(ii) limα(cid:12)(cid:12)(cid:0)f (2), (T − Tα)f (1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12) = 0 for all f (i) ∈ S0(Gi), i = 1, 2,
(iii) κ(Tα) converges to κ(T ) in the weak∗ sense.
(iv) supα kTαkop,S0→S′
0
< ∞.
(v) For T ∈ Lin(L2(G1), L2(G2)) one has limα k(T − Tα)f kL2(G2) = 0 for all f ∈ L2(G1).
(vi) For T ∈ Lin(S0(G1), S0(G2)) one has limα k(T − Tα)f kS0(G2) = 0 for all f ∈ S0(G1).
18
Proof. (i). For any α and i = 1, 2 the operator T (i)
convergent nets in S′
that then Tα = T (2)
nets in S0(G2). Hence Tα ∈ B(G1, G2).
(ii). This is a simple estimate:
α ◦ T ◦ T (1)
0(Gi) into norm convergent nets in S0(Gi). Since T ∈ Lin(S0(G1), S′
α belongs to B(Gi, Gi) and thus maps bounded weak∗
0(G2)) it is clear
0(G1) into norm convergent
α also maps bounded weak∗ convergent nets in S′
lim
+ lim
≤ lim
≤ lim
α ◦ T ◦ T (1)
α ◦ T ◦ T (1)
α (cid:12)(cid:12)(cid:0)f (2), (T − T (2)
α (cid:12)(cid:12)(cid:0)f (2), (T − T (2)
α (cid:12)(cid:12)(cid:0)f (2), (T (2)
α (cid:12)(cid:12)(cid:0)f (2), (IdS′
≤ kf (2)kS0(cid:16) sup
0(G2)(cid:12)(cid:12)
α )f (1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12)
α ◦ T )f (1)(cid:1)S0,S′
α )f (1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12)
α )(T f (1))(cid:1)S0,S′
}
{z
0(cid:17) kT kop,S0→S′
α ◦ T − T (2)
− T (2)
kf (2)kS0kT (2)
0 kT kop,S0→S′
=0 (by Definition 3.19(iv))
α kop,S′
α kop,S′
+ lim
α
0 lim
α
kT (2)
0→S′
0→S′
α
0
0(G2)(cid:12)(cid:12)
0 k(IdS0 − T (1)
α )f (1)kS0
kTαf (1) − f (1)kS0 = 0.
(iii). This is implied by (ii).
(iv). By definition the operators T (1)
and S′
0. Thus
α and T (2)
α have uniformly bounded operator norms as operators on S0
sup
α
≤(cid:16) sup
α
kTαkop,S0→S′
0 = sup
α
kT (2)
α ◦ T ◦ T (1)
α kop,S0→S′
0
kT (2)
α kop,S′
0→S′
0(cid:17) kT kop,S0→S′
0(cid:16) sup
α
kT (1)
α kop,S0→S0(cid:17) < ∞
(vi). In case T ∈ Lin(S0(G1), S0(G2)) we make the following estimate: for all f ∈ S0(G1)
lim
α
≤ lim
α
≤(cid:0) sup
k(T − T (2)
α ◦ T ◦ T (1)
α )f kS0
k(T − T (2)
α ◦ T )f kS0
k(T (2)
α ◦ T − T (2)
α ◦ T ◦ T (1)
α )f kS0
+ lim
α
=0
{z
α kop,S0→S0(cid:1) kT kop,S0→S0 lim
}
α
kT (2)
kf − T (1)
α f kS0 = 0
α
The proof for (v) is similar.
Proposition 3.22. Consider an operator S ∈ Lin(L2(G1), L2(G2)) and T ∈ Lin(L2(G2), L2(G3)). Let
(Sα) and (Tα) be the nets of operators in B(G1, G2) and B(G2, G3) associated to S and T as in Proposition
3.21, respectively. In that case the kernel of the operator T ◦ S ∈ Lin(L2(G1), L2(G3)) is the weak∗ limit
of the kernels of the net of operators (Tα ◦ Sα), κ(Tα ◦ Sα) w∗
−→ κ(T ◦ S), i.e.,
lim
α (cid:12)(cid:12)(cid:0)f (3), (T ◦ S − Tα ◦ Sα)f (1)(cid:1)S0,S′
0(G3)(cid:12)(cid:12) = 0 for all f i ∈ S0(Gi), i = 1, 3.
Remark 7. The usefulness here is that the composition of the operators S and T can we approximated
in the weak∗ sense by a composition of operators Sα and Tα that have kernels in S0. Observe that the
composition Tα ◦ Sα is well understood, cf. the "continuous matrix-matrix product" in Section 3.3.
Proof of Proposition 3.22. By the estimates of Proposition 3.21 and the embedding of L2 into S′
0 one has:
lim
≤ lim
0(G3)(cid:12)(cid:12)
α (cid:12)(cid:12)(cid:0)f (3), (T ◦ S − Tα ◦ Sα)f (1)(cid:1)S0,S′
0(G3)(cid:12)(cid:12)
α (cid:12)(cid:12)(cid:0)f (3), (T ◦ S − Tα ◦ S)f (1)(cid:1)S0,S′
0(G3)(cid:12)(cid:12)
α (cid:12)(cid:12)(cid:0)f (3), (Tα ◦ S − Tα ◦ Sα)f (1)(cid:1)S0,S′
0(cid:1) lim
+ kf (3)kS0(cid:0) sup
≤ kf (3)kS0k(T − Tα)Sf (1)k2
kTαkop,S′
+ lim
0→S′
α
α
kSf (1) − Sαf (1)k2 = 0.
19
Let us apply Proposition 3.22 to a concrete example:
Example 3.23. The Fourier transform F is an operator from L2(G) onto L2(bG). Furthermore its inverse
F −1 is an operator from L2(bG) onto L2(G). It is clear that F −1 ◦ F = IdL2(G). It is not difficult to verify
0 (guaranteed by the outer kernel theorem) are as follows:
that their kernels in S′
• κ(F ) ∈ S′
so that (h, F f )S0,S′
• κ(F −1) ∈ S′
so that (f, F −1h)S0,S′
0( bG) =ZG× bG
0(G) =Z bG×G
0(G × bG) is induced by the function G × bG → C, (x, ω) 7→ ω(x),
0(bG × G) is induced by the function bG × G → C, (ω, x) 7→ ω(x),
0(G × G) is the functional defined by f1 ⊗ f2 7→ZG
f (x)h(ω) ω(x) d(x, ω) for all f ∈ S0(G), h ∈ S0(bG).
h(ω)f (x) ω(x) d(x, ω) for all f ∈ S0(G), h ∈ S0(bG).
f1(x)f2(x) dx,
• κ(IdL2(G)) ∈ S′
for all f1, f2 ∈ S0(G). This is typically expressed as κ(IdL2(G)) = δ(y − x).
Again the analogy to matrix analysis is helpful and gives these symbols a meaning.
• While we describe the distributional kernel for the identity operator as a distribution
of two variables, in the spirit of a Kronecker delta (describing the unit matrix), simply
given as δKron(F ) =ZG
F (x, x) dx, F ∈ S0(G × G) it has become a common understanding to
describe the kernel as a continuous collection of Dirac delta distributions δy,
or with the usual notation δ(y) this becomes just δ(y − x).
Let now (Fα), (F −1
Proposition 3.22. In that case
α ) be two nets of operators in B(G, bG) and B(bG, G) associated to F and F −1 as in
κ(Fα) w∗
κ(F −1
−→ κ(F ), κ(F −1
α ) w∗
−→ κ(F −1),
α ◦ Fα) w∗
−→ κ(F −1 ◦ F ) = κ(IdL2(G)).
At the same time Lemma 3.10 tells us that κ(F −1
α ◦ Fα) is the function in S0(G × G) given by
κ(F −1
α ◦ Fα)(x, y) =Z bG
κ(Fα)(x, ω) · κ(F −1
α )(ω, y) dω.
If "we take the limit" of the above integral, then we are lead to the following "identity", which is often
found in physics and engineering:
Z bG
⇔ Z bG
κ(F )(x, ω) · κ(F −1)(ω, y) dω = κ(IdL2(G))
ω(y − x) dω = δ(y − x).
Expressed in the familiar setting G = bG = R: ZR
We now consider examples of regularizing approximations of the identity.
e2πiω(y−x) dω = δ(y − x), x, y ∈ R.
Example 3.24. (Partial sums of Gabor frame operators) Let g ∈ S0(R) and a, b > 0 be such that
{π(λ)g}λ∈aZ×bZ is a Parseval Gabor frame for L2(R), i.e.,
kf k2
2 = Xλ∈aZ×bZ
hf, π(λ)gi2 for all f ∈ L2(R).
20
In that case the associated Gabor frame operator
Sg : S0(R) → S0(R), Sgf = Xλ∈aZ×bZ
hf, π(λ)giπ(λ)g, f ∈ S0(R)
is the identity on S0(R). Let (ΛN ), N ∈ N be a family of finite subsets of aZ × bZ so that for every point
λ ∈ aZ × bZ there exists an N0 ∈ N such that N > N0 implies that λ ∈ ΛN . For every N ∈ N we define
the operator
Sg,N : S0(R) → S0(R), Sg,N f = Xλ∈ΛN
hf, π(λ)giπ(λ)g.
It extends to an operator on S′
0(R) in the following way:
Sg,N : S′
0(R) → S′
0(R), (f, Sg,N σ)S0,S′
0(R) =(cid:0)f, Xλ∈λN
(π(α)g, σ)S0,S′
0(R) π(λ)g(cid:1)S0,S′
0(R).
The collection of operators (Sg,N )N ∈N is a regularizing approximation of the identity: It is straight forward
to write an explicit formula for the kernel of the operator Sg,N , namely
κ(Sg,N )(t1, t2) = Xλ∈ΛN
π(λ)g(t1) π(λ)g(t2),
t1, t2 ∈ R,
such that (f2, Sg,N f1)S0,S′
0(R) = (f1 ⊗ f2, κ(Sg,N ))S0,S′
(ii) and (iii) we need the following two inequalities:
constant c > 0 such that
0(R2). Hence κ(Sg,N ) ∈ S0(R2). Concerning condition
0(R) there exists a
for any f ∈ S0(R) and σ ∈ S′
Xλ∈aZ×bZ
hf, π(λ)gi ≤ c kf kS0 kgkS0 and
sup
λ∈aZ×bZ
(π(λ)g, σ) ≤ c kgkS0 kσkS′
0
.
(27)
We can then make the following estimates:
kSg,N kop,S0→S0 = sup
kSg,N f kS0
f ∈S0(R)
kf kS0 =1
≤ sup
f ∈S0(R)
kf kS0 =1(cid:13)(cid:13) Xλ∈ΛN
Xλ∈aZ×bZ
f ∈S0(R)
kf kS0 =1
≤ sup
hf, π(λ)giπ(λ)g(cid:13)(cid:13)S0
Xλ∈ΛN
≤ sup
f ∈S0(R)
kf kS0 =1
hf, π(λ)gi kgkS0
hf, π(λ)gi kgkS0
(27)
≤ c
sup
f ∈S0(R)
kf kS0 =1
kf kS0 kgk2
S0 = c kgk2
S0.
Hence supN kSg,N kop,S0→S0 < ∞. Similarly, also using (27), we can show that
Finally, because Sg is the identity on S0(R) we find that
kSg,N kop,S′
0→S′
0
≤ c kgk2
S0.
lim
N→∞
kSg,N f − f kS0 ≤ kgkS0
lim
N→∞ Xλ∈aZ×bZ\ΛN
hf, π(λ)gi = 0
where the last equality follows from the fact that for any two functions f, g ∈ S0(R) the sequence
{hf, π(λ)gi}λ∈aZ×bZ is absolutely summable. In a similar way one can show that (Sg,N ) satisfies condi-
tion (iv) in Definition 3.19.
Example 3.25. (Product-convolution operators) In sequel A(G) is the Fourier algebra A(G) = {f ∈
C0(G) : ∃h ∈ L1(bG) s.t. f = F bGh}, here F bG is the Fourier transform from L1(bG) into C0(G). The norm
in the Fourier algebra is defined by kf kA = khk1, where h is as before. We now construct regularizing
21
approximations of the identity with the help of product-convolution operators. As described in, e.g., [21,
Proposition 4.18], it is possible to find nets of functions (hα) ∈ S0(G) and (gα) ∈ S0(G) such that
lim
α
kf ∗ hα − f kS0 = 0 and lim
α
kf · gα − f k1 = 0 ∀ f ∈ S0(G),
where khαk1 ≤ 1 and kgαkA(G) ≤ 1 for all α. The net of operators
Tα : S′
0(G) → S0(G), Tασ = (σ · gα) ∗ hα, σ ∈ S′
0(G)
is a regularizing approximation of the identity.
Example 3.26. (Localization operators) Let (Hn) be a sequence of uniformly bounded functions in Cc(R2)
that converges uniformly over compact sets to the constant function 1 and take g to be a non-zero function
in S0(R) with kgk2 = 1. Then the operators
Tn : S′
0(R) → S0(R), Tnσ =ZR2
Hn(ν) (π(ν)g, σ)S0,S′
0(R) π(ν)g dν
form a regularizing approximation of the identity.
Similar statements can be obtained for Gabor multipliers with respect to tight Gabor families.
3.6 Kernel theorems for modulation spaces
The inner and outer kernel theorem characterize the operators that are linear and bounded from S′
0(G1)
into S0(G2) and from S0(G1) into S′
0(G2), respectively (with some added assumptions in the former case).
In between S0(G) and S′
0(G)
there is a well-studied family of spaces called the (unweighted) modulation spaces. We refer to [12, 13]
and the relevant chapters in [18] for more on those spaces. For our purpose here we only need to recall
the following.
0(G), or more precise, in between the embedding of S0(G) into S′
0(G) and S′
Definition 3.27. Given p ∈ [1, ∞] and 0 6= g ∈ S0(G). the modulation space Mp(G) can be defined by
Mp(G) =nσ ∈ S′
0(G) : (cid:18)ZG× bG(cid:12)(cid:12) (π(ν)g, σ)S0,S′
0(cid:12)(cid:12)p
dν(cid:19)1/p
< ∞o,
(28)
equipped with the norm kσkMp =(cid:0)RG× bG(cid:12)(cid:12) (π(ν)g, σ)S0,S′
in the obvious way.
0(cid:12)(cid:12)p dν(cid:1)1/p. In case p = ∞ the definition is modified
One can show that different g induce equivalent norms. As already mentioned in Section 2 we have
M1(G) ∼= S0(G) and M∞(G) = S′
0(G). For p ∈ (1, ∞), the modulation space Mp(G) is reflexive and
(Mp(G))′ ∼= Mp′(G), where 1/p + 1/p′ = 1. For any fixed function g ∈ S0(G)\{0}, the action of a
distribution σ ∈ Mp′(G) on a distribution f ∈ Mp(G) is given by
(f, σ)Mp,Mp′ (G) = kgk−2
2 ZG× bG(cid:0)π(ν)g, f(cid:1)S0,S′
0(G)(cid:0)π(ν)g, σ(cid:1)S0,S′
0(G) dν.
(29)
In light of the inner and outer kernel theorems we may therefore ask: can we characterize the bounded
It is straight forward to generalize
0(G2)) to be operators from
linear operators from Mp(G) into Mq(G) for some p, q ∈ [1, ∞].
Theorem 3.2 to the following sufficient condition for operators in Lin(S0(G1), S′
Mp′(G1) into Mq(G2).
Proposition 3.28. Fix any function g(i) ∈ S0(Gi)\{0}, i = 1, 2 and let p, q ∈ [1, ∞]. If an operator
T ∈ Lin(S0(G1), S′
0(G2)) satisfies the condition
ZG2× bG2(cid:16)ZG1× bG1(cid:12)(cid:12)(cid:0)π(ν(2))g(2), T π(ν(1))g(1)(cid:1)S0,S′
0(G2)(cid:12)(cid:12)p dν(1)(cid:17)q/p
dν(2) < ∞
22
then T is bounded from Mp′(G1) into Mq(G2). Hence, for σ(1) ∈ Mp′(G) and σ(2) ∈ Mq′(G),
kg(1) ⊗ g(2)k2
=
Z
G1× bG1×G2× bG2
2 (cid:0)σ(2), T σ(1)(cid:1)Mq′ ,Mq
Vg(1)σ(1)(ν(1)) · Vg(2)σ(2)(ν(2)) ·(cid:0)π(ν(2))g(2), T π(ν(1))g(1)(cid:1)S0,S′
0(G2) d(ν(1), ν(2)).
In general, the assumption in Proposition 3.28 is only sufficient for T to be a bounded operator from
Mp′(G1) to Mq(G2). For example, if p = q = 2, then the identity operator is bounded on L2(G), but its
kernel is not in L2(G × G).
In [1] and [7] it has been shown recently that for certain choices of p and q the condition in Proposition
3.28 is necessary for boundedness from Mp′(G1) into Mq(G2). Specifically in the cases where
(1) p = ∞ and q ∈ [1, ∞]
;
(2) p ∈ [1, ∞] and q = ∞.
Such results confirm the usefulness of coorbit spaces, here specifically of modulation spaces.
4 Proof of the inner kernel theorem
For the proof of Theorem 1.3 it is useful to introduce the space B(G1, G2):
Definition 4.1. Let G1 and G2 be locally compact abelian Hausdorff groups. We then define
B(G1, G2) = {T ∈ Lin(S′
into a norm convergent net in ι(S0(G2)) ⊆ S′′
0(G2)}.
0(G1), ι(S0(G2))) : T maps every bounded weak∗convergent net in S′
0(G1)
The identification of S0(G) with ι(S0(G)) implies that B(G1, G2) ∼= B(G1, G2).
Proof of Theorem 1.3. We will show that the three Banach spaces S0(G1 × G2), A and B(G1, G2) are
isomorphic. Since the roles of G1 and G2 can be interchanged (because S0(G1 × G2) ∼= S0(G2 × G1)), it
then automatically follows that also B(G2, G1) is isomorphic to these spaces. In order to prove the desired
identifications, we consider the following two operators.
c : S0(G1 × G2) → A, c(K) =h(σ(1), σ(2)) 7→ (K, σ(1) ⊗ σ(2))S0,S′
d : A → B(G1, G2), d(A) =hσ(1) 7→(cid:2)σ(2) 7→ A(σ(1), σ(2)) (cid:3) i,
0(G1×G2)i,
where K ∈ S0(G1 × G2), A ∈ A and σ(i) ∈ S′
that both these operators are well-defied, linear and bounded.
0(Gi), i = 1, 2. In Lemma 4.3 and Lemma 4.4 we will show
Furthermore, let S′
0(G1) ⊗ S′
0(G2) be the tensor product of S′
0(G1) and S′
0(G2), that is,
0(G1) ⊗ S′
S′
0(G2) = {σ ∈ S′
j=1 σ(1)
j ⊗ σ(2)
j
, N ∈ N }.
0(G1 × G2) : σ =PN
Then, for a given T ∈ B(G1, G2), we define the operator
e(T ) : S′
0(G1) ⊗ S′
0(G2) → C, e(T )(cid:16) NXj=1
σ(1)
j ⊗ σ(2)
j (cid:17) =
NXj=1
T (σ(1)
j )(σ(2)
j ).
So far it is not clear whether the value of e(T )(σ), σ ∈ S′
0(G1) ⊗ S′
0(G2) depends on the particular
representation PN
j=1 σ(1)
j ⊗ σ(2)
j
of σ. We will show in a moment that this is not the case.
In Lemma 4.5 we show that e(T ) is continuous with respect to the weak∗ topology induced by functions
0(G1 × G2) (the distributions induced by
0(G2)), there is a
0(G2) is weak∗ dense in S′
0(G1 × G2) and they are a subspace of S′
0(G1) ⊗ S′
in S0(G1 × G2). Because S′
S0(G1) ⊗ S0(G2) are weak∗ dense in S′
0(G1) ⊗ S′
23
unique weak∗ continuous extension of e(T ), which we also call e(T ), to a functional from S′
C. We can therefore define the operator
0(G1 × G2) to
e : B → ι(S0(G1 × G2)) ⊆ S′′
0(G1 × G2),
which, to every T ∈ B, assigns the operator e(T ) from above. Since ι(S0(G1 × G2)) ∼= S0(G1 × G2) we can
consider e as an operator from B into S0(G1 × G2).
Now, given K ∈ S0(G1 × G2), A ∈ A and T ∈ B(G1, G2) one can, simply by the definitions of the three
operators c, d and e, show that
e ◦ d ◦ c(K) = K, c ◦ e ◦ d(A) = A, d ◦ c ◦ e(T ) = T.
This implies that c, d and e are injective, surjective, and hence invertible. We conclude that e is the
(unique) inverse operator of d ◦ c, thus e(T )(σ) for σ ∈ S′
0(G2) can not depend on a particular
representation of σ as discussed earlier in the proof. Because S0(G1 × G2) is a Banach space, it follows
that also the normed vector spaces A and B(G1, G2) are Banach spaces. To complete the proof it remains
only to prove Lemma 4.3, 4.4, and 4.5.
0(G1) ⊗ S′
In order to verify weak∗ continuity of functionals the following result is essential to us.
Lemma 4.2 ([22, Corollary 2.7.9]). Let X be a Banach space and X ′ its continuous dual space. For a
functional ϕ : X ′ → C the following statements are equivalent:
(i) ϕ is weak∗ continuous, i.e., if (x′
α) is a weak∗ convergent net in X ′ with limit x′
0, then for all ǫ > 0
there exists a α0 such that for all α > α0 one has ϕ(x′
α − x′
0) < ǫ.
(ii) ϕ is continuous with respect to the bounded weak∗ topology, i.e., if (x′
α) is a (in X ′ norm) bounded
0, then for all ǫ > 0 there exists a α0 such that one has
weak∗ convergent net in X ′ with limit x′
ϕ(x′
0) < ǫ for all α > α0.
α − x′
Lemma 4.3. The follwoing operator is well-defined, linear and bounded:
c : S0(G1 × G2) → A, c(K) =(cid:2)(σ(1), σ(2)) 7→ (K, σ(1) ⊗ σ(2))S0,S′
0(G1×G2)(cid:3) .
Proof. Let a function K ∈ S0(G1 × G2) be given. Then for some constant a > 0
c(K)(σ(1), σ(2)) = (K, σ(1) ⊗ σ(2)) ≤ kKkS0kσ(1) ⊗ σ(2)kS′
0
(10)
≤ a kKkS0 kσ(1)kS′
0 kσ(2)kS′
0,
Hence c(K)(σ(1), σ(2)) is well-defined. The bilinearity of c(K) is clear. Also,
c(K)(σ(1), σ(2)) ≤ a kKkS0.
sup
(Gi)=1, i=1,2
kσ(i)kS′
0
(30)
(31)
This shows that c(K) is an element in Bil(S′
0(G2), C). Let us show that c(K) ∈ A, i.e., c(K) is
weak∗ continuous in each variable. In order to show this, let us first consider a function K ∈ S0(G1) ⊗
S0(G2) ⊆ S0(G1 × G2), that is, a function of the form
0(G1) × S′
K =
NXj=1
j ⊗ f (2)
f (1)
j
, (f (i)
j )N
j=1 in S0(Gi), i = 1, 2, N ∈ N.
If (σ(1)
α ) is a bounded weak∗ convergent net in S′
α − σ(1)
c(K)(σ(1)
0(G1) with limit σ(1)
(K, σ(1)
0 , σ(2)) = lim
α
0 , and σ(2) ∈ S′
α − σ(1)
0 , σ(2))
lim
α
0(G2), then
= lim
α
NXj=1
(f (1)
j
, σ(1)
α − σ(1)
0 ) (f (2)
j
, σ(2))
≤ a max
j
kf (2)
j kS0 kσ(2)kS′
0
, σ(1)
α − σ(1)
0 ) = 0.
(f (1)
j
lim
α
NXj=1
24
By Lemma 4.2 the operator c(K) is weak∗ continuous in the first coordinate. The continuity in the second
coordinate is proven in the same fashion. Let now K be any function in S0(G1 × G2). Then, given any
ǫ > 0, we can find function K ∈ S0(G1) ⊗ S0(G2) such that
kK − KkS0 <
ǫ
4
sup
α,{0}
kσ(1)
(·) kS′
0 kσ(2)kS′
0.
With this K fixed, there is, as we just showed, an index α0 such that for all α > α0
c( K)(σ(1)
α − σ(1)
0 , σ(2)) < ǫ/2.
Hence, for α > α0 we have that
c(K)(σ(1)
α − σ(1)
0 , σ(2))
α − σ(1)
= c(K − K + K)(σ(1)
α − σ(1)
0 , σ(2)) + c( K)(σ(1)
≤ c(K − K)(σ(1)
kσ(1)
< 2 kK − KkS0 sup
(·) kS′
0 + ǫ/2
0 kσ(2)kS′
0 , σ(2))
α,{0}
α − σ(1)
0 , σ(2))
< ǫ/2 + ǫ/2 = ǫ.
We have thus shown that c(K) is weak∗ continuous in the first coordinate for any K ∈ S0(G1 × G2).
The continuity in the second coordinate is proven in the same way. Consequently c is a mapping from
S0(G1 × G2) into A. The linearity of c is clear. Finally, the boundedness of c follows from the inequalities
concerning c(K) above, namely,
sup
K∈S0(G1×G2)
kKkS0 =1
kc(K)kBil(S′
0×S′
0,C) ≤ a,
where a is the same constant as in (30) and (31). Hence the operator c is well-defined, linear and bounded.
Lemma 4.4. The operator
d : A → B(G1, G2), d(A) =(cid:2)σ(1) 7→(cid:2)σ(2) 7→ A(σ(1), σ(2)) (cid:3) (cid:3) , σ(i) ∈ S′
0(Gi), i = 1, 2,
is well-defined, linear and bounded.
Proof. Let A be an operator in A. Let us show that d(A) is an operator in B(G1, G2). That is, we need to
show that d(A) ∈ Lin(S0(G1), ι(S′
0(G1)
into norm convergent nets in S′′
0(G2)
we have the estimate
0(G2))) and that d(A) maps bounded weak∗ convergent nets in S′
0(G1) and σ(2) ∈ S′
0(G2). Since A ∈ A it is clear that for all σ(1) ∈ S′
d(A)(σ(1))(σ(2)) = A(σ(1), σ(2)) ≤ kAkopkσ(1)kS′
0
kσ(2)kS′
0
< ∞.
(32)
Hence the functional
d(A)(σ1) : S′
0(G2) → C, d(A)(σ(1))(σ(2)) = A(σ(1), σ(2))
is well-defined. The bilineairty of A implies that d(A)(σ(1)) is linear. In order to show that the functional
is also bounded we use the estimate from (32). This yields
d(A)(σ(1))(σ(2))
(32)
≤
sup
σ(2)∈S′
0(G2)
kσ(2)k=1
sup
σ(2)∈S′
0(G2)
kσ(2)k=1
kAkop kσ(1)kS′
0
kσ(2)kS′
0
= kAkop kσ(1)kS′
0
< ∞.
(33)
25
Hence d(A)(σ(1)) is also bounded. The weak∗continuity of this functional is also easy to show: if (σ(2)
weak∗ convergent net in S′
coordinate,
α ) is a
0(G2), then, since A is weak∗ continuous in the second
0(G2) with limit σ(2)
0 ∈ S′
lim
α
d(A)(σ(1))(σ(2)
α − σ(2)
0 ) = lim
α
A(σ1, σ(2)
α − σ(2)
0 ) = 0.
Thus d(A)(σ(1)) ∈ ι(S0(G2)). Let us verify that d(A) is a bounded operator from S′
S′′
0(G2).
0(G1) into ι(S0(G2)) ⊆
kd(A)(σ(1))kS′′
0
(33)
≤
sup
kσ(1)kS′
0
(G1
≤1
sup
kσ(1)kS′
0
(G1)≤1
kAkop kσ(1)kS′
0
= kAkop.
(34)
We have thus shown that d(A) ∈ Lin(S′
convergent nets in S′
convergent net (σ(1)
α ) in S′
0(G1) with limit σ(1)
0 one has:
0(G1) into norm convergent nets in ι(S0(G2)) ⊆ S′′
0(G1), ι(S0(G2))). It is left to show that d(A) maps bounded weak∗
0(G2). Given a bounded weak∗
kd(A)(σ(1)
α − σ(1)
0 )kS′′
0
lim
α
= lim
α
= lim
α
sup
kσ(2)kS′
0
(G2)≤1
sup
kσ(2)kS′
0
(G2)≤1
d(A)(σ(1)
α − σ(1)
0 )(σ(2))
A(σ(1)
α − σ(1)
0 , σ(2)).
We need to show that the limit is equal to zero. Note that A is weak∗ continuous in the first and second
entry. By the Banach-Alaoglu Theorem ([22, Theorem 2.6.18]) the unit ball of S′
0(G2) is compact in the
weak∗ topology. Continuous mappings on compact sets are uniform continuous, therefore we conclude that
lim
α
kd(A)(σ(1)
α − σ(1)
0 )kS′′
0 = lim
α
sup
kσ(2)kS′
0(G2)≤1
A(σ(1)
α − σ(1)
0 , σ(2)) = 0.
Hence d(A)(σ(1)
well-defined operator, clearly linear. It is also bounded:
0(G2)-norm convergent net with limit d(A)(σ(1)
α ) is a S′′
0 ). Thus d(A) ∈ B and hence d is a
kdkop = sup
kAk=1
kd(A)kop,S′
0→S′′
0 = sup
kAk=1
k d(A)kop,S′
0→S′′
0
(34)
≤ 1.
Lemma 4.5. For every T ∈ B(G1, G2), the operator given by
e(T ) : S′
0(G1) ⊗ S′
0(G2) → C, e(T )(cid:16) NXj=1
σ(1)
j ⊗ σ(2)
j (cid:17) =
NXj=1
T (σ(1)
j )(σ(2)
j ).
is linear and continuous with respect to the weak∗ topology induced by the functions in S0(G1 × G2).
Proof. Let us first show that e(T ) is a well-defined and linear operator on S′
find that for all finite sequences (σ(i)
0(Gi), i = 1, 2, N ∈ N,
j=1 in S′
j )N
0(G1) ⊗ S′
0(G2). Indeed, we
NXj=1
T (σ(1)
j )(σ(2)
j )(cid:12)(cid:12) ≤ kT kop,S′
0→S′′
0
NXj=1
kσ(1)
j kS′
0
kσ(2)
j kS′
0
< ∞.
For e(T ) to be well-defined we should verify that the value of e(T )(σ), σ ∈ S′
j=1 σ(1)
j ⊗ σ(2)
j
0(G2) is independent
. This issue is resolved in the proof of Theorem 1.3. The
0(G1) ⊗ S′
NPj=1
j ⊗ σ(2)
σ(1)
(cid:12)(cid:12)e(T )(
j )(cid:12)(cid:12) =(cid:12)(cid:12)
of its particular representation PN
linearity of the operator e(T ) follows immediately from its definition.
Let us now show that e(T ) is weak∗ continuous. Let us start with bounded net of elementary tensors,
α ⊗ σ(2)
α ) is weak∗ convergent towards σ(1)
0 ⊗ σ(2)
0 . Then
(σ(1)
e(T )(σ(1)
α ⊗ σ(2)
α ) = e(T )(σ(1)
0 ⊗ σ(2)
0 ).
lim
α
(35)
26
2 kS′
0 ⊗ σ(2)
0(G2) as 0 = σ(1)
= 2. Assume now that (σ(1)
Since e(T ) is linear, it is enough to verify its weak∗ continuity at 0. We may write the zero element
0(G1) ⊗ S′
0(G2) with
0 ). Furthermore, we may assume without loss
0 = 1 for all α (in order to achieve this normalization
α = 0, then use that σ(1) ⊗ 0 = 0 ⊗ σ(2) = 0
in S′
kσ(0)
of generality that supα kσ(1)
use that σ(1) ⊗ σ(2) = ασ(1) ⊗ α−1σ(2) for all α ∈ C\{0}. If σ(2)
and then normalize appropriately).
is some non-zero element in S′
0(G1) < ∞ and kσ(2)
0 = 0 and σ(2)
−→ 0 = (0 ⊗ σ(2)
0 , where σ(1)
α ⊗ σ(2)
α ) w∗
α kS′
α kS′
0
0
Now, assume for a moment that (σ(1)
α ) w∗
X−→ 0 and that (σ(2)
α ) w∗
X−→ σ(2)
exist a function h(i) ∈ S0(Gi) and an ǫ(i) > 0 such that, for all index α(i)
(h(i), σ(i)
0(Gi) ≥ ǫ(i). This allows us, for sufficiently large α to achieve the inequality
α(i) − σ(i)
0 )S0,S′
0 . Then, for i = 1, 2 there
0 and
0 we have that α(i) > α(i)
ǫ(1)ǫ(2) ≤ (h(1), σ(1)
α )S0,S′
0(G1)(h(2), σ(2)
α − σ(2)
0 )S0,S′
0(G2).
(36)
On the other hand, because by assumption (σ(1)
high values of α,
α ⊗σ(2)
α ) w∗
−→ (0⊗σ(2)
0 ) = 0 we can ensure that, for sufficiently
(h(1), σ(1)
α )S0,S′
0(G1) (h(2), σ(2)
α − σ(2)
0 )S0,S′
0(G2) < ǫ(1)ǫ(2).
This is a contradiction to (36) and therefore the assumption that (σ(1)
wrong. We must therefore be in either of the following three situations:
α ) w∗
X−→ 0 and that (σ(2)
α ) w∗
X−→ σ(2)
0
is
(i) (σ(1)
α ) w∗
−→ 0 and (σ(2)
α ) w∗
X−→ σ(2)
0
(ii) (σ(1)
α ) w∗
−→ 0 and (σ(2)
α ) w∗
−→ σ(2)
0
(iii) (σ(1)
α ) w∗
X−→ 0 and (σ(2)
α ) w∗
−→ σ(2)
0
Assume for a moment that (σ(2)
α ) w∗
−→ σ(0)
2 . It follows from [22, Theorem 2.6.14] that this implies that
However, with our choice of normalization we find that
kσ(2)
0 kS′
0 ≤ lim inf
α
2 = kσ(2)
0 kS′
0 ≤ lim inf
α
kσ(2)
α kS′
0.
kσ(2)
α kS′
0 = 1,
which, clearly, can not be the case. We must therefore be in situation (i). We thus have that (σ(1)
σ(1)
0 = 0. Note that T maps bounded weak∗ convergent nets in S′
0(G2). Thus limα kT σ(1)
S′′
α ) w∗
−→
0(G1) into norm convergent nets in
0 = 0. We therefore find that
α kS′′
e(T )(σ(1)
α ⊗ σ(2)
α ) − e(T )(σ(1)
0 ⊗ σ(2)
0 )
lim
α
= lim
α
≤ lim
α
≤(cid:0) sup
α
T (σ(1)
α )(σ(2)
kT (σ(1)
α )kS′′
α ) − T (0)(σ(2)
0 kσ(2)
α kS′
0
0 ) = lim
α
T (σ(1)
α )(σ(2)
α )
kσ(2)
α kS′
kT (σ(1)
α )kS′′
0 = 0.
0(cid:1) lim
α
We have thus verified the continuity of e(T ) for elementary tensors with respect to the weak∗ topology
induced by functions in S0(G1 × G2). This continuity is preserved by finite linear combinations and as a
consequence e(T ) is continuous from S′
0(G2) into C.
0(G1) ⊗ S′
Acknowledgments
The work of M.S.J. was carried out during the tenure of the ERCIM 'Alain Bensoussan' Fellowship Pro-
gramme at NTNU.
27
References
[1] P. Balazs and K. Gröchenig. A guide to localized frames and applications to Galerkin-like repre-
sentations of operators. In Isaac Pesenson, Hrushikesh Mhaskar, Azita Mayeli, Quoc T. Le Gia, and
Ding-Xuan Zhou, editors, Novel methods in harmonic analysis with applications to numerical analysis
and data processing, Applied and Numerical Harmonic Analysis series (ANHA). Birkhäuser/Springer,
2017.
[2] S. Bannert. Banach-Gelfand Triples and Applications in Time-Frequency Analysis. Master's thesis,
University of Vienna, 2010.
[3] M. Bownik and K. Ross. The structure of translation-invariant spaces on locally compact abelian
groups. J. Fourier Anal. Appl., 21(4):849–884, 2015.
[4] F. Bruhat. Distributions sur un groupe localement compact et applications à l etude des représenta-
tions des groupes p-adiques. Bull. Soc. Math. France, 89:43–75, 1961.
[5] O. Christensen. An Introduction to Frames and Riesz Bases. Applied and Numerical Harmonic
Analysis. Birkhäuser Basel, Second edition, 2016.
[6] E. Cordero, H. G. Feichtinger, and F. Luef. Banach Gelfand triples for Gabor analysis. In Pseudo-
differential Operators, volume 1949 of Lecture Notes in Mathematics, pages 1–33. Springer, Berlin,
2008.
[7] E. Cordero and F. Nicola. Kernel theorems for modulation spaces. J. Fourier Anal. Appl., pages
1–14, 2017.
[8] A. Delcroix. Kernel theorems in spaces of generalized functions. In Linear and non-linear theory of
generalized functions and its applications, volume 88 of Banach Center Publ., pages 77–89. Polish
Acad. Sci. Inst. Math., Warsaw, 2010.
[9] H. G. Feichtinger. Un espace de Banach de distributions tempérées sur les groupes localement com-
pacts abéliens. C. R. Acad. Sci. Paris S'er. A-B, 290(17):791–794, 1980.
[10] H. G. Feichtinger. On a new Segal algebra. Monatsh. Math., 92:269–289, 1981.
[11] H. G. Feichtinger. Minimal Banach spaces and atomic representations. Publ. Math. Debrecen, 34(3-
4):231–240, 1987.
[12] H. G. Feichtinger. Modulation spaces of locally compact Abelian groups. In R. Radha, M. Krishna,
and S. Thangavelu, editors, Proc. Internat. Conf. on Wavelets and Applications, pages 1–56, Chennai,
January 2002, 2003. New Delhi Allied Publishers.
[13] H. G. Feichtinger. Modulation Spaces: Looking Back and Ahead. Sampl. Theory Signal Image
Process., 5(2):109–140, 2006.
[14] H. G. Feichtinger and K. Gröchenig. Banach spaces related to integrable group representations and
their atomic decompositions, I. J. Funct. Anal., 86(2):307–340, 1989.
[15] H. G. Feichtinger and K. Gröchenig. Gabor wavelets and the Heisenberg group: Gabor expansions
and short time Fourier transform from the group theoretical point of view. In C. K. Chui, editor,
Wavelets :a tutorial in theory and applications, volume 2 of Wavelet Anal. Appl., pages 359–397.
Academic Press, Boston, 1992.
[16] H. G. Feichtinger and W. Kozek. Quantization of TF lattice-invariant operators on elementary LCA
groups. In H. G. Feichtinger and T. Strohmer, editors, Gabor analysis and algorithms, Appl. Numer.
Harmon. Anal., pages 233–266. Birkhäuser Boston, Boston, MA, 1998.
28
[17] G. Folland. A Course in Abstract Harmonic Analysis. Textbooks in Mathematics. CRC Press, Boca
Raton, Second edition, 2016.
[18] K. Gröchenig. Foundations of Time-Frequency Analysis. Appl. Numer. Harmon. Anal. Birkhäuser,
Boston, MA, 2001.
[19] L. Hörmander. The Analysis of Linear Partial Differential Operators I: Distribution Theory and
Fourier Analysis. Classics in Mathematics. Springer, Reprint of the 2nd Edition 1990 edition, 2003.
[20] M. S. Jakobsen and J. Lemvig. Density and duality theorems for regular Gabor frames. J. Funct.
Anal., 270(1):229 – 263, 2016.
[21] M. S. Jakobsen. On a (no longer) New Segal Algebra: A Review of the Feichtinger Algebra. J. Fourier
Anal. Appl., preprint, 2018.
[22] R. Megginson. An Introduction to Banach Space Theory, volume 183 of Graduate Texts in Mathe-
matics. Springer-Verlag, New York, 1998.
[23] L. Nachbin. The Haar Integral. Princeton, N.J.-Toronto-New York-London: D. Van Nostrand Com-
pany, 1965.
[24] M. S. Osborne. On the Schwartz-Bruhat space and the Paley-Wiener theorem for locally compact
Abelian groups. J. Funct. Anal., 19:40–49, 1975.
[25] H. Reiter and J. D. Stegeman. Classical Harmonic Analysis and Locally Compact Groups. 2nd ed.
Clarendon Press, Oxford, 2000.
[26] A. Ruston. On the Fredholm theory of integral equations for operators belonging to the trace class
of a general Banach space. Proc. London Math. Soc. (2), 53:109–124, 1951.
[27] H. Schaefer and M. Wolff. Topological Vector Spaces, volume 3 of Graduate Texts in Mathematics.
Springer-Verlag, New York, Second edition, 1999.
[28] F. Treves. Topological Vector Spaces, Distributions and Kernels. Number 25 in Pure Appl. Math.
Academic Press, New York, 1967.
29
|
1502.07064 | 1 | 1502 | 2015-02-25T06:38:09 | The strong "zero-two" law for positive contractions of Banach-Kantorovich L_p-lattices | [
"math.FA"
] | In the present paper we study majorizable operators acting on Banach-Kantorovich $L_p$-lattices, constructed by a measure $m$ with values in the ring of all measurable functions. Then using methods of measurable bundles of Banach-Kantorovich lattices, we prove the strong "zero-two" law for positive contractions of the Banach-Kantorovich $L_p$-lattices. | math.FA | math | THE STRONG "ZERO-TWO" LAW FOR POSITIVE CONTRACTIONS OF
BANACH-KANTOROVICH Lp-LATTICES
INOMJON GANIEV, FARRUKH MUKHAMEDOV, AND DILMURAD BEKBAEV
Abstract. In the present paper we study majorizable operators acting on Banach-Kantorovich
Lp-lattices, constructed by a measure m with values in the ring of all measurable functions.
Then using methods of measurable bundles of Banach-Kantorovich lattices, we prove the strong
"zero-two" law for positive contractions of the Banach-Kantorovich Lp-lattices.
Mathematics Subject Classification: 37A30, 47A35, 46B42, 46E30, 46G10.
Key words and phrases: Banach-Kantorovich Lp-lattice, strong "zero-two" law, positive con-
traction.
5
1
0
2
b
e
F
5
2
]
.
A
F
h
t
a
m
[
1
v
4
6
0
7
0
.
2
0
5
1
:
v
i
X
r
a
1. Introduction
Starting from von Neumman's [20] pioneering work, the development of the theory of Banach
bundles had been stimulated by many works (see for example [13, 14]). There are many papers
were devoted to the applications of this theory to several branches of analysis [1, 16, 17, 23].
Moreover, this theory is well-connected with the theory of vector-valued Banach spaces [12, 13],
which has several applications (see for example, [18]). In the present paper, we concentrate
ourselves to the theory of Banach bundles of L0-valued Banach spaces (see for more details
[6, 13]). Note that such spaces are called Banach -- Kantorovich spaces. In [13, 14, 17]) the theory
of Banach -- Kantorovich spaces were developed. It is known [13] that the theory of measurable
bundles of Banach lattices is sufficiently well explored. Therefore, it is natural to employ
methods of measurable bundles of such spaces to investigate functional properties of Banach --
Kantorovich spaces. It is an effective tool which gives a good opportunity to obtain various
properties of these spaces [4, 5]. For example, in [7, 6] Banach-Kantorovich lattice Lp(∇, µ) is
represented as a measurable bundle of classical Lp -lattices. Naturally, these functional Banach --
Kantorovich spaces have many similar properties like the classical ones, constructed by the real
In [2, 10] this allowed to establish several weighted ergodic theorems for
valued measures.
positive contractions of Lp(∇, µ)-spaces.
In [5] the convergence theorems of martingales on
such lattices has been proved. Some other applications of the measurable bundles of Banach-
Kantorovich spaces can be found in [1, 11].
In [19] Ornstein and Sucheston proved that, for any positive contraction T on an L1-space,
one has either kT n − T n+1k1 = 2 for all n or lim
kT n − T n+1k1 = 0. An extension of this result
n→∞
to positive operators on L∞-spaces was given by Foguel [3]. In [24] Zahoropol generalized these
results, called "zero-two" laws, and his result can be formulated as follows:
Theorem 1.1. Let T be a positive contraction of Lp, p > 1, p 6= 2. If the following relation
holds (cid:13)(cid:13)T m+1 − T m(cid:13)(cid:13) < 2 for some m ∈ N ∪ {0}, then
kT n+1 − T nk = 0.
lim
n→∞
1
2
INOMJON GANIEV, FARRUKH MUKHAMEDOV, AND DILMURAD BEKBAEV
In [15] this theorem was established for Kothe spaces. In particularly, from that result it
follows the statement of the theorem for a case p = 2.
Furthermore, the strong "zero-two" law for positive contractions of Lp-spaces, 1 ≤ p < +∞
was proved in [22]. This result is formulated as follows:
Theorem 1.2. Let 1 ≤ p < +∞ and T be a positive contraction of Lp. If (cid:13)(cid:13)T m+1 − T m(cid:13)(cid:13) < 2
for some m ∈ N ∪ {0}, then
lim
n→∞(cid:13)(cid:13)T n+1 − T n(cid:13)(cid:13) = 0.
In [9] we have generalized Theorem 1.1 for the positive contractions of the Banach-Kantorovich
Lp-lattices. Namely, the following result was proved.
Theorem 1.3. Let T : Lp(∇, m) → Lp(∇, m), p > 1, p 6= 2 be a positive linear contraction
such that T 1 ≤ 1. If one had (cid:13)(cid:13)T m+1 − T m(cid:13)(cid:13) < 2 · 1 for some m ∈ N ∪ {0}. Then
kT n+1 − T nk = 0.
(o) − lim
n→∞
The main aim of this paper is to prove the strong "zero-two" law for the positive contrac-
tions of the Banach-Kantorovich lattices Lp(∇, m). To establish the main aim, we first study
majorizable operators acting on Banach-Kantorovich Lp-lattices (see Section 3). Then using
methods of measurable bundles of Banach-Kantorovich lattices, in section 4 we prove the main
result of the present paper.
2. Preliminaries
Let (Ω, Σ, µ) be a complete measure space with a finite measure µ. By L(Ω) (resp. L∞(Ω) )
we denote the set of all (resp. essentially bounded) measurable real functions defined on Ω a.e.
By the standard way, we introduce an equivalence relation on L(Ω) by putting f ∼ g whenever
f = g a.e. The set L0(Ω) of all cosets f ∼ = {g ∈ L(Ω) : f ∼ g}, endowed with the natural
algebraic operations, is an algebra with unit 1(ω) = 1 over the field of reals R. Moreover, with
respect to the partial order f ∼ ≤ g∼ ⇔ f ≤ g a.e., the algebra L0(Ω) is a Dedekind complete
Riesz space with weak unit 1, and the set B(Ω) := B(Ω, Σ, µ) of all idempotents in L0(Ω) is a
complete Boolean algebra. Furthermore, L∞(Ω) = {f ∼ : f ∈ L∞(Ω)} is an order ideal in L0(Ω)
generated by 1. In what follows, we will write f ∈ L0(Ω) instead of f ∼ ∈ L0(Ω) by assuming
that the coset of f is considered.
Let E be a linear space over the real field R. By k · k we denote a L0(Ω)-valued norm on E.
Then the pair (E, k · k) is called a lattice-normed space (LNS) over L0(Ω). An LNS E is said to
be d-decomposable if for every x ∈ E and the decomposition kxk = f + g with f and g disjoint
positive elements in L0(Ω) there exist y, z ∈ E such that x = y + z with kyk = f , kzk = g.
Suppose that (E, k · k) is an LNS over L0(Ω). A net {xα} of elements of E is said to be
(bo)-converging to x ∈ E (in this case we write x = (bo)-lim xα), if the net {kxα − xk} (o)-
converges to zero (here (o)-convergence means the order convergence) in L0(Ω) (written as
(o)-lim kxα − xk = 0). A net {xα}α∈A is called (bo)-fundamental if (xα − xβ)(α,β)∈A×A (bo)-
converges to zero.
An LNS in which every (bo)-fundamental net (bo)-converges is called (bo)-complete. A
Banach-Kantorovich space (BKS) over L0(Ω) is a (bo)-complete d-decomposable LNS over
L0(Ω). It is well known [16],[17] that every BKS E over L0(Ω) admits an L0(Ω)-module struc-
ture such that kf xk = f · kxk for every x ∈ E, f ∈ L0(Ω), where f is the modulus of a
THE STRONG "ZERO-TWO" LAW
3
function f ∈ L0(Ω). A BKS (U , k · k) is called a Banach-Kantorovich lattice if U is a vector
lattice and the norm k · k is monotone, i.e.
u1 ≤ u2 implies ku1k ≤ ku2k. It is known [16]
that the cone U+ of positive elements is (bo)-closed.
Let ∇ be an arbitrary complete Boolean algebra and let X(∇) be the Stone space of ∇.
Assume that L0(∇) := C∞(X(∇)) be the algebra of all continuous functions x : X(∇) →
[−∞, +∞] that take the values ±∞ only on nowhere dense subsets of X(∇). Finally, by
C(X(∇)) we denote the subalgebra of all continuous real functions on X(∇).
Given a complete Boolean algebra ∇, let us consider a mapping m : ∇ → L0(Ω). Such a
mapping is called a L0(Ω)-valued measure if one has
(i) m(e) ≥ 0 for all e ∈ ∇ and m(e) = 0 ⇔ e = 0;
(ii) m(e ∨ g) = m(e) + m(g) if e ∧ g = 0, e, g ∈ ∇;
(iii) m(eα) ↓ 0 for any net eα ↓ 0.
Following the well-known scheme of the construction of Lp-spaces, a space Lp(∇, m) can be
defined by
Lp(∇, m) =(cid:26)f ∈ L0(∇) : f p :=Z f pdm − exist (cid:27) ,
p ≥ 1
where m is a L0(Ω)-valued measure on ∇.
A L0(Ω)-valued measure m is said to be disjunctive decomposable (d-decomposable), if for
ai ∈ L0(B) there exit
a1 ∧ a2 = 0,
every e ∈ ∇ and the decomposition m(e) = a1 + a2,
e1, e2 ∈ ∇ such that e = e1 ∨ e2 and m(ei) = ai, i = 1, 2.
Theorem 2.1. [6] The following statements hold:
(i) The pair (Lp(∇, m), ·p) is (bo) -- complete lattice. Moreover, it is an ideal linear subspace
of L0(∇), i.e. from x ≤ y, y ∈ Lp(∇, m), x ∈ L0(∇) it follows that x ∈ Lp(∇, m)
and xp ≤ yp;
(ii) If 0 ≤ xα ∈ Lp(∇, m) and xα ↓ 0, then xαp ↓ 0;
(iii) If the measure m is d-decomposable, then α)xp = αxp for all α ∈ L0(Ω), x ∈
Lp(∇, m);
(iv) If the measure m is d-decomposable, then (Lp(∇, m), · p) is a Banach -- Kantorovich
space;
(v) One has L∞(∇, m) := C(X(∇)) ⊂ Lp(∇, m) ⊂ Lq(∇, m), 1 ≤ q ≤ p. Moreover,
L∞(∇, m) is (bo) -- dense in (L1(∇, m), k · k1).
Now we mention necessary facts from the theory of measurable bundles of Boolean algebras
and Banach spaces (see [13] for more details).
Let (Ω, Σ, µ) be the same as above and X be a mapping assigning an Lp-space constructed
by a real-valued measure mω, i.e. Lp(∇ω, mω) to each point ω ∈ Ω and let
L =(cid:26) nXi=1
αiei : αi ∈ R, ei(ω) ∈ ∇ω, i = 1, n, n ∈ N(cid:27)
4
INOMJON GANIEV, FARRUKH MUKHAMEDOV, AND DILMURAD BEKBAEV
be a set of sections. In [6] it has been established that the pair (X, L) is a measurable bundle
of Banach lattices and L0(Ω, X) is modulo ordered isomorphic to Lp(∇, µ).
Let ρ be a lifting in L∞(Ω) (see [13]). As before, let ∇ be an arbitrary complete Boolean
subalgebra of ∇(Ω) and m be an L0(Ω)-valued measure on ∇. By L∞(∇, m) we denote the set
of all essentially bounded functions w.r.t. m taken from L0(∇).
A mapping ℓ : L∞(∇, m)(⊂ L∞(Ω, X)) → L∞(Ω, X) is called a vector-valued lifting [13]
associated with the lifting ρ if it satisfies the following conditions:
(1) ℓ(u) ∈ u for all u such that dom(u) = Ω;
(2) kℓ(u)kLp(∇ω ,mω) = ρ(up)(ω);
(3) ℓ(u + v) = ℓ(u) + ℓ(v) for every u, v ∈ L∞(∇, m);
(4) ℓ(h · u) = ρ(h)ℓ(u) for every u ∈ L∞(∇, m), h ∈ L∞(Ω);
(5) ℓ(u) ≥ 0 whenever u ≥ 0;
(6) the set {ℓ(u)(ω) : u ∈ L∞(∇, m)} is dense in X(ω) for all ω ∈ Ω;
(7) ℓ(u ∨ v) = ℓ(u) ∨ ℓ(v) for every u, v ∈ L∞(∇, m).
In [6] the existence of the vector-valued lifting was proved.
Let Lp(∇, m) (p ≥ 1) be a Banach-Kantorovich lattice. A linear mapping T : Lp(∇, m) →
Lp(∇, m) is called positive if T f ≥ 0 whenever f ≥ 0. We say that T is a L0(Ω)-bounded
mapping if there exists a function k ∈ L0(Ω) such that T f p ≤ k f p for all f ∈ Lp(∇, µ). For
such a mapping we can define an element of L0(Ω) as follows
kT k = sup
f p≤1
T f p,
which is called an L0(Ω)-valued norm of T . A mapping T is said to be a contraction if one has
kT k ≤ 1. Some examples of contractions can be found in [10].
In the sequel we will need the following bundle representation of L0(Ω)-linear L0(Ω)-bounded
operators acting in Banach-Kantorovich lattices.
Theorem 2.2. [9] Let Lp(∇, m) (p ≥ 1) be a Banach-Kantorovich lattice, and Lp(∇ω, mω) be
the corresponding Lp-spaces constructed by real valued measures. Let T : Lp(∇, m) → Lp(∇, m)
be a positive linear contraction such that T 1 ≤ 1. Then for every ω ∈ Ω there exists a
positive contraction Tω : Lp(∇ω, µω) → Lp(∇ω, mω) such that Tωf (ω) = (T f )(ω) a.e. for every
f ∈ Lp(∇, m).
3. Majorizable operators in Banach-Kantorovich Lp-lattices
In this section, we are going to study majorizable operators in Banach-Kantorovich Lp-
lattices.
Theorem 3.1. Let T : L1(∇, m) → L1(∇, m) be an L0(Ω)- bounded linear operator in Banach-
Kantorovich lattice L1(∇, m). Then there exists a unique T - L0(Ω)- bounded linear operator
in L1(∇, m) such that
(a) kT k = kT k;
(b) one has T f ≤ T f, for all f ∈ L1(∇, m);
(c) for each f ∈ L1(∇, m) with f ≥ 0 one has T f = sup{T g : g ∈ L1(∇, m), g ≤ f };
(d) kT k∞ = kT k∞.
THE STRONG "ZERO-TWO" LAW
5
Proof. Let P denote the family of all finite measurable partitions π = {B1, B2, . . . , Bm} of Ω.
We partially order P in the usual way, i.e. for π = {B1, B2, . . . , Bm} and π′ = {B′
k}
we write π ≤ π
i}.
′ is a refinement of π, i.e. each set Bi is a union of sets {B′
′ if π
1, B′
2, . . . , B′
Given π ∈ P, and for every f ∈ L1(∇, m), f ≥ 0 we define
Tπ f =
mXi=1
T (χBi
f ).
Clearly π ≤ π′ implies Tπ f ≤ Tπ′ f . From f 1 =
χBi
f 1 we obtain Tπ f 1 ≤ kT k f1. Since
{Tπ f : π ∈ P} is increasing on P and is norm bounded, therefore one can define
mPi=1
We clearly have
(3.1)
T f := lim
π∈P
Tπ f ,
f ≥ 0.
T f1 ≤ kT k f 1, f ≥ 0
and T is linear on positive functions. Therefore T can be extended by the linearity to whole
L1(∇, m). This extension is again denoted by T .
For f ≥ 0 and g ≤ f we obtain T f ≥ T g by means of the approximation argument with
simple functions. This yields (b).
(c). From (b) we have T g ≥ T g, i.e. T has a positive majorant. Then by [21, Theorem
VIII 1.1] T is regular. Hence, using [21, formula (10),p.231] one finds T f = sup{T g : g ∈
L1(∇, m), g ≤ f }.
(a). Again from (b) we get kT k ≤ kT k and by (3.1) one finds kT k ≤ kT k. Hence,
kT k = kT k.
(d). Let f ∈ L∞( ∇, µ).
kT k∞k f k∞ which means kT k∞ ≤ kT k∞.
It is then clear that from T f ≤ T f one gets kT k∞k f k∞ ≤
Using (c) we obtain
T f = sup
g≤ f
T g ≤ sup
g≤ f
kT k∞kgk∞1 ≤ kT k∞k f k∞1.
Hence, kT k∞ ≤ kT k∞ and kT k∞ = kT k∞.
(cid:3)
Definition 3.2. A linear operator A : Lp(∇, m) → Lp(∇, m) is called majorizable if there
exists an L0(Ω)- bounded positive linear operator S : Lp(∇, m) → Lp(∇, m) such that
for all f ∈ Lp(∇, m). The operator S is called majorant.
A f ≤ S( f )
Theorem 3.3. Let T : Lp(∇, m) → Lp(∇, m) be a majorizable operator with a majorant S on
Banach-Kantorovich lattice Lp(∇, m). Then there exists a unique T - L0(Ω)- bounded linear
operator on Lp(∇, m) such that
(a) kT k ≤ kSk;
(b) one has T f ≤ T f, for all f ∈ Lp(∇, m);
(c) for each f ∈ Lp(∇, m), f ≥ 0 one has T f = sup{T g : g ∈ Lp(∇, m), g ≤ f };
6
INOMJON GANIEV, FARRUKH MUKHAMEDOV, AND DILMURAD BEKBAEV
Proof. The proof of the existence of T and (b), (c) are similar to the proof of Theorem 3.1.
Now we prove (a). From
T f = sup{T g : g ∈ Lp(∇, m), g ≤ f } ≤ sup{Sg : g ∈ Lp(∇, m), g ≤ f } = S f
we get
hence
T fp ≤ S f p ≤ kSkk f p
kT k ≤ kSk.
This completes the proof.
(cid:3)
Theorem 3.4. If A : Lp(∇, m) → Lp(∇, m) is a majorizable operator, and its majorant
S is a contraction with S1 ≤ 1, then for every ω ∈ Ω there exists a majorizable operator
Aω : Lp(∇ω, mω) → Lp(∇ω, mω) such that
for all f ∈ Lp(∇, m).
Aωf (ω) = (A f )(ω) a.e.
Proof. Since S is a contraction and S1 ≤ 1, we obtain that A(L∞(∇, m)) ⊂ L∞(∇, m).
Now we define a linear operator ϕω from {ℓ( f )(ω) : f ∈ L∞(∇, m)} into Lp(∇ω, mω) by
where ℓ is the vector lifting of L∞(∇, m) associated with the lifting ρ.
ϕω(ℓ( f )(ω)) = ℓ(A f )(ω)
From the majorizability of A one gets
ϕ(ω)(ℓ( f)(ω)) = ℓ(A f )(ω) = ℓ(A f )(ω) ≤ ℓ(S f )(ω) = S′
ω(ℓ( f )(ω)) = S′
ω(ℓ( f )(ω))
for any positive f ∈ L∞(∇, m), where S′
This means that ϕ(ω) is a majorizable operator on {ℓ( f )(ω) : f ∈ L∞(∇, m)}.
ω is a positive contraction on {ℓ( f )(ω) : f ∈ L∞(∇, m)}.
From S f p ≤ f p we obtain
kℓ(A f )(ω)kLp(∇ω ,mω) = ρ(A f p)(ω) ≤ ρ(S f p)(ω) ≤ ρ( f p)(ω) = kℓ( f )(ω)kLp(∇ω ,mω)
which implies that ϕω and S′
ω is positive (see
Theorem 2.2). Due to the density of {ℓ( f )(ω) : f ∈ L∞(∇, m)} in Lp(∇ω, mω), we can extend
ϕω and S′
ω, respectively, to Lp(∇ω, mω). We respectively denote the extensions by Aω and Sω.
One can see that Aω is bounded, and Sω is positive bounded.
ω are well defined and bounded. Moreover, S′
From
for any f ∈ L∞(∇, m) one finds
ϕ(ω)(ℓ( f)(ω)) ≤ S′
ω(ℓ( f )(ω))
Aω(f (ω)) ≤ Sω(f (ω))
i.e. Aω is majorizable.
Repeating the argument of the proof of [9, Theorem 2.1], we can prove that
Aωf (ω) = (A f )(ω)
for almost all ω ∈ Ω and for all f ∈ Lp(∇, m). This completes the proof.
(cid:3)
THE STRONG "ZERO-TWO" LAW
7
Theorem 3.5. If A : Lp(∇, m) → Lp(∇, m) is a majorizable operator, and its majorant S is a
contraction with S1 ≤ 1, then
kAωkp,ω = kAωkp,ω
for almost all ω ∈ Ω, where k · kp,ω is the norm of an operator from Lp(∇ω, mω) to Lp(∇ω, mω).
Proof. Due to −A ≤ A ≤ A we have −Aω ≤ Aω ≤ Aω which yields Aω ≤ Aω for almost
all ω ∈ Ω. Hence, kAωkp,ω ≥ kAωkp,ω for almost all ω ∈ Ω.
Let {πn} be an increasing sequence in P such that A f = (bo) − lim
n→∞
Aπn
f , for 0 ≤ f ∈
Lp(∇, m).
One can see that
(3.2)
(Aπn
f )(ω) =
for almost all ω ∈ Ω.
Now using
mXi=1
A(χBi
f )(ω) =
mXi=1
Aω(χBi(ω)f )(ω) = Aω,πnf (ω)
A f = (bo) − lim
n→∞
Aπn
f in Lp(∇, m),
with (3.2) we obtain Aπn
Hence,
f p
(o)
→ (cid:12)(cid:12)A f(cid:12)(cid:12)p or Aπn
f p(ω) → (cid:12)(cid:12)A f(cid:12)(cid:12)p(ω) for almost all ω ∈ Ω.
for almost all ω ∈ Ω.
On the other hand, one has
for almost all ω ∈ Ω. This means that
kAπn,ωf (ω)kLp(∇ω ,mω) →(cid:13)(cid:13)Aωf (ω)(cid:13)(cid:13)Lp(∇ω ,mω)
kAπn,ωf (ω)kLp(∇ω ,mω) ≤(cid:13)(cid:13)Aωf (ω)(cid:13)(cid:13)Lp(∇ω ,mω)
(cid:13)(cid:13)Aωf (ω)(cid:13)(cid:13)Lp(∇ω ,mω) ≤(cid:13)(cid:13)Aωf (ω)(cid:13)(cid:13)Lp(∇ω ,mω)
lim
n→∞
or
for almost all ω ∈ Ω. Hence
(cid:13)(cid:13)Aω(cid:13)(cid:13)p,ω ≤(cid:13)(cid:13)Aω(cid:13)(cid:13)p,ω
(cid:13)(cid:13)Aω(cid:13)(cid:13)p,ω =(cid:13)(cid:13)Aω(cid:13)(cid:13)p,ω
for almost all ω ∈ Ω. This completes the proof.
(cid:3)
4. The strong "zero-two" law
In this section we are going to prove an analog of the strong "zero-two" law for positive
contractions in the Banach-Kantorovich Lp-lattices. Before the formulation of the main result,
we need some auxiliary results.
Proposition 4.1. Let T, S : Lp(∇, m) → Lp(∇, m) be two positive linear contractions such
that T 1 ≤ 1, S1 ≤ 1. Then
here · means the modulus of an operator.
(cid:13)(cid:13)Tω − Sω(cid:13)(cid:13)p,ω ≥(cid:13)(cid:13)T − S(cid:13)(cid:13)(ω), a.e.
8
INOMJON GANIEV, FARRUKH MUKHAMEDOV, AND DILMURAD BEKBAEV
Proof. Due to (T − S)( f ) ≤ T ( f ) for any positive f ∈ Lp(∇, m) one gets
(T − S)( f ) ≤ T ( f )
for any f ∈ Lp(∇, m). Hence T − S is majorizable. Since T is a contraction and T 1 ≤ 1 by
Theorem 3.5, we obtain(cid:13)(cid:13)T − Sω(cid:13)(cid:13)p,ω =(cid:13)(cid:13)Tω − Sω(cid:13)(cid:13)p,ω for almost all ω ∈ Ω. By [8, Proposition
2] for any ε > 0 there exists f ∈ Lp(∇, m) with f p = 1 such that
Then
(cid:13)(cid:13)T − S(cid:13)(cid:13) − ε1 ≤(cid:12)(cid:12)T − S f(cid:12)(cid:12)p
.
(cid:13)(cid:13)T − S(cid:13)(cid:13)(ω) − ε1 ≤ (cid:12)(cid:12)T − S f(cid:12)(cid:12)p(ω) = k(T − S f )(ω)kLp(∇ω ,mω)
= kT − Sωf (ω)kLp(∇ω ,mω) ≤(cid:13)(cid:13)T − Sω(cid:13)(cid:13)p,ω
= (cid:13)(cid:13)Tω − Sω(cid:13)(cid:13)p,ω
for almost all ω ∈ Ω. The arbitrariness of ε > 0 implies the statement.
(cid:3)
Corollary 4.2. Let T, S : Lp(∇, m) → Lp(∇, m) be two positive linear contractions such that
T 1 ≤ 1, S1 ≤ 1. Then
(cid:13)(cid:13)Tω − Sω(cid:13)(cid:13)p,ω =(cid:13)(cid:13)T − S(cid:13)(cid:13)(ω), a.e.
The proof follows from [9, Proposition 3.2] and Proposition 4.1.
The next theorem is our main result of the present paper.
Theorem 4.3. Let T : Lp(∇, m) → Lp(∇, m) be a positive linear contraction such that T 1 ≤ 1.
If one has (cid:13)(cid:13)T m+1 − T m(cid:13)(cid:13) < 2 · 1 for some m ∈ N ∪ {0}. Then
n→∞(cid:13)(cid:13)T n+1 − T n(cid:13)(cid:13) = 0.
ω (cid:13)(cid:13)p,ω =(cid:13)(cid:13)T m+1 − T m(cid:13)(cid:13)(ω), a.e.
Proof. From Corollary 4.2 it follows that
(cid:13)(cid:13)T m+1
ω − T m
(o) − lim
on Ω. Therefore, due to(cid:13)(cid:13)T m+1−T m(cid:13)(cid:13) < 2·1 for some m ∈ N∪{0} we find(cid:13)(cid:13)T m+1
< 2
for almost all ω ∈ Ω. According to Theorem 2.2 we conclude that Tω is a positive contraction
on Lp(∇ω, mω). Hence, the contraction Tω satisfies the conditions of Theorem 1.2 for almost all
ω ∈ Ω, which yields that
ω (cid:13)(cid:13)p,ω
ω −T m
(cid:3)
for almost all ω ∈ Ω. Then again using Corollary 4.2 we obtain
lim
lim
ω − T n
ω (cid:13)(cid:13) = 0
n→∞(cid:13)(cid:13)T n+1
n→∞(cid:13)(cid:13)T n+1 − T n(cid:13)(cid:13)(ω) = 0
n→∞(cid:13)(cid:13)T n+1 − T n(cid:13)(cid:13) = 0.
for almost all ω ∈ Ω. Therefore,
This completes the proof.
(o) − lim
THE STRONG "ZERO-TWO" LAW
9
Acknowledgement
The first author acknowledges the MOE Grant FRGS13-071-0312. The second named author
thanks the MOE grant FRGS14-135-0376, and the Junior Associate scheme of the Abdus Salam
International Centre for Theoretical Physics, Trieste, Italy.
References
[1] Albeverio S., Ayupov Sh.A., Kudaybergenov K.K., Non commutative Arens algebras and their deriva-
tions, J. Funct. Anal.253 (2007), 287 -- 302.
[2] Chilin V.I., Ganiev I.G. An individual ergodic theorem for contractions in the Banach-Kantorovich lattice
Lp(∇, µ). Russian Math. (Iz. VUZ) 44 (2000), 77 -- 79.
[3] Foguel S.R. On the "zero-two" law. Israel J. Math. 10(1971), 275-280.
[4] Ganiev I.G. Measurable bundles of Banach lattices. Uzbek Math. Zh. (1998), N.5, 14 -- 21 (Russian).
[5] Ganiev I.G. The martingales convergence in the Banach-Kantorovich's lattices Lp(b∇,bµ), Uzb. Math.
[6] Ganiev I.G. Measurable bundles of lattices and their applications. In book : Studies on Functional
Jour. 2000, N.1, 18 -- 26.
Analysis and its Applications, pp. 9 -- 49. Nauka, Moscow (2006) (Russian).
[7] Ganiev, I. G., Chilin, V.I. Measurable bundles of noncommutative Lp-spaces associated with a center-
valued trace. Siberian Adv. Math. 12 (2002), no. 4, 19 -- 33.
[8] Ganiev I.G., Kudaybergenov K.K). The Banach-Steinhaus Uniform Boundedness Principle for Operators
in Banach-Kantorovich Spaces over L0, Siberian Adv. Math. 16(2006), 42 -- 53.
[9] Ganiev I.G., Mukhamedov F. On the "Zero-Two" law for positive contractions in the Banach-Kantorovich
lattice Lp(∇, µ), Comment. Math. Univ. Carolinae 47 (2006), 427 -- 436
[10] Ganiev I.G., Mukhamedov F. On weighted ergodic theorems for Banach-Kantorovich lattice Lp(∇, µ),
Lobachevskii Jour. Math. 34 (2013), 1 -- 10.
[11] Ganiev I., Mukhamedov F., Measurable bundles of C ∗-dynamical systems and its applications, Positivity
18(2014), 687 -- 702.
[12] Gierz, G. Bundles of Topological Vector Spaces and their Duality, Springer, Berlin, 1982.
[13] Gutman A.E. Banach bundles in the theory of lattice-normed spaces, III. Siberien Adv. Math. 3(1993),
n.4, 8 -- 40
[14] Gutman A.E. Banach fiberings in the theory of lattice-normed spaces. Order-compatible linear operators,
Trudy Inst. Mat. 29(1995), 63-211. (Russian)
[15] Katznelson Y., Tzafriri L. On power bounded operators, J. Funct. Anal. 68(1986), 313-328.
[16] Kusraev A.G. Vector duality and its applications, Novosibirsk, Nauka, 1985 (Russian).
[17] Kusraev A.G. Dominanted operators, Mathematics and its Applications, V. 519. Kluwer Academic Pub-
lishers, Dordrecht, 2000.
[18] Lee Y., Lin Y., Wahba G., Multicategory support vector machines: theory and application to the J.
Amer. Statist. Assoc. 99 (465) (2004) 67-81.
[19] Ornstein D., Sucheston L. An operator theorem on L1 convergence to zero with applications to Markov
kernels. Ann. Math. Statis. 41(1970), 1631-1639.
[20] von Neumann, J. On rings of operators III, Ann. Math. 41 (1940), 94 -- 161.
[21] Vulih B.Z. Introduction to theory of partially ordered spaces. Moscow, 1961 (Russian). English trans.
Noordhoff, Groningen, 1967.
[22] Wittman R. Ein starkes Null-Zwei Gesetz in Lp, Math. Z. 197(1988) 223 -- 229.
[23] Woyczynski W.A. Geometry and martingales in Banach spaces. Lect. Notes Math. 472(1975) 235 -- 283.
[24] Zaharopol R. The modulus of a regular liniear operators and the "zero-two" law in Lp-spaces (1 < p < ∞,
p 6= 2). J. Funct. Anal. 68(1986), 300-312.
10
INOMJON GANIEV, FARRUKH MUKHAMEDOV, AND DILMURAD BEKBAEV
Inomjon Ganiev, Department of Science in Engineering, Faculty of Engineering, Interna-
tional Islamic University Malaysia, P.O. Box 10, 50728, Kuala-Lumpur, Malaysia
E-mail address: [email protected], [email protected]
Farrukh Mukhamedov, Department of Computational & Theoretical Sciences, Faculty of
Science, International Islamic University Malaysia, P.O. Box, 141, 25710, Kuantan, Pahang,
Malaysia
E-mail address: farrukh [email protected], [email protected]
Dilmurad Bekbaev, Department of Computational & Theoretical Sciences, Faculty of Sci-
ence, International Islamic University Malaysia, P.O. Box, 141, 25710, Kuantan, Pahang,
Malaysia
|
1710.04893 | 1 | 1710 | 2017-10-13T12:48:50 | Some generalizations of the Aluthge transform of operators | [
"math.FA"
] | Let $A = U |A|$ be the polar decomposition of $A$. The Aluthge transform of the operator $A$, denoted by $\tilde{A}$, is defined as $\tilde{A} =|A|^{\frac{1}{2}} U |A|^{\frac{1}{2}}$. In this paper, first we generalize the definition of Aluthge transform for non-negative continuous functions $f, g$ such that $f(x)g(x)=x\,\,(x\geq0)$. Then, by using of this definition, we get some numerical radius inequalities. Among other inequalities, it is shown that if $A$ is bounded linear operator on a complex Hilbert space ${\mathscr H}$, then
\begin{equation*} h\left( w(A)\right) \leq \frac{1}{4}\left\Vert h\left( g^{2}\left( \left\vert A\right\vert \right) \right) +h\left( f^{2}\left( \left\vert A\right\vert \right) \right) \right\Vert +\frac{1}{2}h\left( w\left( \tilde{A}_{f,g}\right) \right) , \end{equation*} where $f, g$ are non-negative continuous functions such that $f(x)g(x)=x\,\,(x\geq 0)$, $h$ is a non-negative non-decreasing convex function on $[0,\infty )$ and $\tilde{A}_{f,g} =f(|A|) U g(|A|)$. | math.FA | math |
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF
OPERATORS AND THEIR CONSEQUENCES
MOJTABA BAKHERAD1 AND KHALID SHEBRAWI2
Abstract. Let A = U A be the polar decomposition of A. The Aluthge transform
of the operator A, denoted by A, is defined as A = A
generalize the definition of Aluthge transform for non-negative continuous functions
2 . In this paper, first we
2 U A
1
1
f, g such that f (x)g(x) = x (x ≥ 0). Then, by using of this definition, we get some
numerical radius inequalities. Among other inequalities, it is shown that if A is
bounded linear operator on a complex Hilbert space H , then
h (w(A)) ≤
1
4(cid:13)(cid:13)h(cid:0)g2 (A)(cid:1) + h(cid:0)f 2 (A)(cid:1)(cid:13)(cid:13) +
1
2
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17) ,
where f, g are non-negative continuous functions such that f (x)g(x) = x (x ≥ 0), h is
a non-negative non-decreasing convex function on [0, ∞) and Af,g = f (A)U g(A).
1. Introduction
Let B(H ) denote the C ∗-algebra of all bounded linear operators on a complex Hilbert
space H with an inner product h·, ·i and the corresponding norm k · k. In the case
when dimH = n, we identify B(H ) with the matrix algebra Mn of all n × n matrices
with entries in the complex field. For an operator A ∈ B(H ), let A = UA (U is a
partial isometry with kerU = rngA⊥) be the polar decomposition of A. The Aluthge
transform of the operator A, denoted by A, is defined as A = A
2 . In [12],
Okubo introduced a more general notion called t-Aluthge transform which has later
been studied also in detail. This is defined for any 0 < t ≤ 1 by At = AtUA1−t.
Clearly, for t = 1
2 we obtain the usual Aluthge transform. As for the case t = 1, the
operator A1 = AU is called the Duggal transform of A ∈ B(H ). For A ∈ B(H ), we
generalize the Aluthge transform of the operator A to the form
2 UA
1
1
Af,g = f (A)U g(A),
2010 Mathematics Subject Classification. Primary 47A12, Secondary 47A63, 47A30 .
Key words and phrases. Aluthge transform, Numerical
radius, Operator matrices, Polar
decomposition.
1
2
M. BAKHERAD AND K. SHEBRAWI
in which f, g are non-negative continuous functions such that f (x)g(x) = x (x ≥ 0).
The numerical radius of A ∈ B(H ) is defined by
w(A) := sup{hAx, xi : x ∈ H , kxk = 1}.
It is well known that w( · ) defines a norm on B(H ), which is equivalent to the usual
operator norm k · k. In fact, for any A ∈ B(H ), 1
2kAk ≤ w(A) ≤ kAk; see [6]. Let
r(·) denote to the spectral radius. It is well known that for every operator A ∈ B(H ),
we have r(A) ≤ w(A). An important inequality for ω(A) is the power inequality
stating that ω(An) ≤ ω(A)n (n = 1, 2, · · · ). The quantity w(A) is useful in studying
perturbation, convergence and approximation problems as well as integrative method,
etc. For more information see [3, 7, 8, 9] and references therein.
Let A, B, C, D ∈ B(H ). The operator matrices " A 0
the diagonal and off-diagonal parts of the operator matrix " A B
C 0 # are called
C D #, respectively.
0 D # and " 0 B
In [11], It has been shown that if A is an operator in B(H ), then
Several refinements and generalizations of inequality (1.1) have been given; see [1, 4,
14, 15]. Yamazaki [15] showed that for A ∈ B(H ) and t ∈ [0, 1] we have
w(A) ≤
1
2(cid:16)kAk + kA2k
1
2(cid:17) .
w(A) ≤
1
2(cid:16)kAk + w( At)(cid:17) .
(1.1)
(1.2)
Davidson and Power [5] proved that if A and B are positive operators in B(H ), then
kA + Bk ≤ max{kAk, kBk} + kABk
1
2 .
(1.3)
Inequality (1.3) has been generalized in [2, 13].
In [13], the author extended this
inequality to the form
kA + B∗k ≤ max{kAk, kBk} +
in which A, B ∈ B(H ) and t ∈ [0, 1].
1
2(cid:0)(cid:13)(cid:13)AtB∗1−t(cid:13)(cid:13) +(cid:13)(cid:13)A∗1−tBt(cid:13)(cid:13)(cid:1) ,
(1.4)
In this paper, by applying the generalized Aluthge transform of operators, we es-
tablish some inequalities involving the numerical radius.
In particular, we extend
inequality (1.2) and (1.4) for two non-negative continuous functions. We also show
some upper bounds for the numerical radius of 2 × 2 operators matrices.
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
3
To prove our numerical radius inequalities, we need several known lemmas.
2. main results
Lemma 2.1. [1, Theorem 2.2] Let X, Y, S, T ∈ B(H ). Then
r(XY + ST ) ≤
1
2
(w(Y X) + w(T S)) +
1
2q(w(Y X) − w(T S))2 + 4kY SkkT Xk.
Lemma 2.2. [15, 11] Let A ∈ B(H ). Then
(a) w(A) = max
θ∈R (cid:13)(cid:13)Re(cid:0)eiθA(cid:1)(cid:13)(cid:13) .
0 0 #! = 1
2 kAk.
(b) w " 0 A
Polarization identity: For all x, y ∈ H , we have
hx, yi =
1
4
3
Xk=0(cid:13)(cid:13)x + iky(cid:13)(cid:13)
2
ik.
Now, we are ready to present our first result. The following theorem shows a general-
ization of inequality (1.2).
Theorem 2.3. Let A ∈ B(H ) and f, g be two non-negative continuous functions on
[0, ∞) such that f (x)g(x) = x (x ≥ 0). Then, for all non-negative non-decreasing
convex function h on [0, ∞), we have
h (w(A)) ≤
1
4(cid:13)(cid:13)h(cid:0)g2 (A)(cid:1) + h(cid:0)f 2 (A)(cid:1)(cid:13)(cid:13) +
1
2
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17) .
4
M. BAKHERAD AND K. SHEBRAWI
Proof. Let x be any unit vector. Then
Re(cid:10)eiθAx, x(cid:11) = Re(cid:10)eiθU A x, x(cid:11)
1
1
4(cid:13)(cid:13)(cid:0)eiθf (A) − g (A) U ∗(cid:1) x(cid:13)(cid:13)
(by polarization identity)
2
2
1
=
2
2
≤
≤
1
1
−
= Re(cid:10)eiθU g (A) f (A) x, x(cid:11)
= Re(cid:10)eiθf (A) x, g (A) U ∗x(cid:11)
4(cid:13)(cid:13)(cid:0)eiθf (A) + g (A) U ∗(cid:1) x(cid:13)(cid:13)
4(cid:13)(cid:13)(cid:0)eiθf (A) + g (A) U ∗(cid:1) x(cid:13)(cid:13)
4(cid:13)(cid:13)(cid:0)eiθf (A) + g (A) U ∗(cid:1)(cid:13)(cid:13)
4(cid:13)(cid:13)(cid:0)eiθf (A) + g (A) U ∗(cid:1)(cid:0)e−iθf (A) + U g (A)(cid:1)(cid:13)(cid:13)
g2 (A) + f 2 (A) + eiθ Af,g + e−iθ(cid:16) Af,g(cid:17)∗(cid:13)(cid:13)(cid:13)
4(cid:13)(cid:13)(cid:13)
eiθ Af,g + e−iθ(cid:16) Af,g(cid:17)∗(cid:13)(cid:13)(cid:13)
4(cid:13)(cid:13)(cid:13)
4(cid:13)(cid:13)g2 (A) + f 2 (A)(cid:13)(cid:13) +
2(cid:13)(cid:13)(cid:13)
Re(cid:16)eiθ Af,g(cid:17)(cid:13)(cid:13)(cid:13)
4(cid:13)(cid:13)g2 (A) + f 2 (A)(cid:13)(cid:13) +
w(cid:16) Af,g(cid:17) .
4(cid:13)(cid:13)g2 (A) + f 2 (A)(cid:13)(cid:13) +
1
2
1
1
1
1
1
1
=
=
≤
=
≤
Now, taking the supremum over all unit vector x ∈ H and applying Lemma 2.2 in the
above inequality produces
w (A) ≤
1
4(cid:13)(cid:13)g2 (A) + f 2 (A)(cid:13)(cid:13) +
1
2
w(cid:16) Af,g(cid:17) .
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
5
Therefore,
+
2
2
1
2
1
2
g2 (A) + f 2 (A)
g2 (A) + f 2 (A)
h (w (A)) ≤ h(cid:18)1
4(cid:13)(cid:13)g2 (A) + f 2 (A)(cid:13)(cid:13) +
2(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)
= h(cid:18)1
(cid:13)(cid:13)(cid:13)(cid:13)
h(cid:18)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:19) +
h(cid:18) g2 (A) + f 2 (A)
2(cid:13)(cid:13)(cid:13)(cid:13)
(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
4(cid:13)(cid:13)h(cid:0)g2 (A)(cid:1) + h(cid:0)f 2 (A)(cid:1)(cid:13)(cid:13) +
w(cid:16) Af,g(cid:17)(cid:19)
w(cid:16) Af,g(cid:17)(cid:19)
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17)
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17)
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17)
1
2
1
2
1
1
2
1
2
1
≤
≤
=
+
2
(by the convexity of h)
(by the convexity of h).
(by the functional calculus)
Theorem 2.3 includes some special cases as follows.
Corollary 2.4. Let A ∈ B(H ). Then, for all non-negative non-decreasing convex
function h on [0, ∞) and all t ∈ [0, 1], we have
(cid:3)
h (w(A)) ≤
(2.1)
Corollary 2.5. Let A ∈ B(H ). Then, for all t ∈ [0, 1] and r ≥ 1, we have
wr(A) ≤
In particular,
1
4(cid:13)(cid:13)(cid:13)
h(cid:0)A2t(cid:1) + h(cid:16)A2(1−t)(cid:17)(cid:13)(cid:13)(cid:13)
k A2tr + A2(1−t)r k +
1
4
+
1
2
h(cid:16)w(cid:16) At(cid:17)(cid:17) .
wr(cid:16) At(cid:17) .
1
2
wr(A) ≤
1
2(cid:16)kAkr + wr(cid:16) A(cid:17)(cid:17) .
Proof. The first inequality follows from inequality (2.1) for the function h (x) = xr (r ≥
1). For the particular case, it is enough to put t = 1
2 .
(cid:3)
Theorem 2.3 gives the next result for the off-diagonal operator matrix " 0 A
B 0 #.
Theorem 2.6. Let A, B ∈ B(H ), f, g be two non-negative continuous functions on
[0, ∞) such that f (x)g(x) = x (x ≥ 0) and r ≥ 1. Then
wr " 0 A
B 0 #! ≤
1
4
+
1
max(cid:0)(cid:13)(cid:13)g2r (A) + f 2r (A)(cid:13)(cid:13) ,(cid:13)(cid:13)g2r (B) + f 2r (B)(cid:13)(cid:13)(cid:1)
4(cid:0)kf (B)g(A∗)kr + kf (A)g(B∗)kr(cid:1).
6
M. BAKHERAD AND K. SHEBRAWI
Proof. Let A = UA and B = V B be the polar decompositions of A and B, re-
spectively and let T = " 0 A
B 0 #.
A # that
T =" 0 U
0 #" B
V
0
0
It follows from the polar the composition of
Tf,g = f (T )" 0 U
V
0
= " f (B)
= "
0
f (A)V g(B)
0 # g(T )
f (A) #" 0 U
V
0
0
0 #" g(B)
# .
0
f (B)U g(A)
0
g(A) #
Using A∗2 = AA∗ = UA2U ∗ and B∗2 = BB∗ = V B2V ∗ we have g(A) =
U ∗g(A∗)U and g(B) = V ∗g(B∗)V for every non-negative continuous function g on
[0, ∞). Therefore,
w(cid:16) Tf,g(cid:17) = w "
0
f (A)V g(B)
0
0
≤ w " 0 f (B)U g(A)
= w " 0 f (B)U g(A)
= w " 0 f (B)U g(A)
0
0
0
0
f (B)U g(A)
0
0
0
#!
#! + w "
f (A)V g(B) 0 #!
# U!
#! + w U ∗" 0 f (A)V g(B)
#!
#! + w " 0 f (A)V g(B)
0
0
0
0
=
=
≤
1
2
1
2
1
2
kf (B)U g(A)k +
1
2
kf (A)V g(B)k
(by Lemma 2.1(b))
kf (B)U U ∗g(A∗)Uk +
1
2
kf (A)V V ∗g(B∗)V k
kf (B)g(A∗)k +
1
2
kf (A)g(B∗)k,
(2.2)
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
7
where U =" 0 I
I 0 # is unitary. Applying Theorem 2.3 and inequality (2.2), we have
wr (T ) ≤
≤
≤
1
1
1
1
4
2(cid:16)wr(cid:16) Tf,g(cid:17)(cid:17)
4(cid:13)(cid:13)g2r (T ) + f 2r (T )(cid:13)(cid:13) +
max(cid:0)(cid:13)(cid:13)g2r (A) + f 2r (A)(cid:13)(cid:13) ,(cid:13)(cid:13)g2r (B) + f 2r (B)(cid:13)(cid:13)(cid:1)
2(cid:20) 1
max(cid:0)(cid:13)(cid:13)g2r (A) + f 2r (A)(cid:13)(cid:13) ,(cid:13)(cid:13)g2r (B) + f 2r (B)(cid:13)(cid:13)(cid:1)
(kf (B)g(A∗)k + kf (A)g(B∗)k)(cid:21)r
kf (B)g(A∗)kr +
kf (A)g(B∗)kr
1
4
2
+
+
1
4
1
4
(by the convexity h(x) = xr).
Corollary 2.7. Let A, B ∈ B(H ). Then, for all t ∈ [0, 1] and r ≥ 1, we have
(cid:3)
(cid:3)
Proof. Applying the power inequality of the numerical radius, we have
w
r
2 (AB) ≤
1
4
+
1
A2tr + A2(1−t)r(cid:13)(cid:13)(cid:13)
max(cid:16)(cid:13)(cid:13)(cid:13)
4(cid:16)(cid:13)(cid:13)At B∗1−t(cid:13)(cid:13)
w
r
2 (AB) ≤ max(cid:0)w
r
r
2 (AB) , w
B2tr + B2(1−t)r(cid:13)(cid:13)(cid:13)(cid:17)
,(cid:13)(cid:13)(cid:13)
r
r(cid:17) .
+(cid:13)(cid:13)Bt A∗1−t(cid:13)(cid:13)
2 (BA)(cid:1)
r
r
= w
= w
0 BA #!
2 " AB 0
2
B 0 #2
" 0 A
B 0 #!
≤ wr " 0 A
A2tr + A2(1−t)r(cid:13)(cid:13)(cid:13)
max(cid:16)(cid:13)(cid:13)(cid:13)
4(cid:16)(cid:13)(cid:13)At B∗1−t(cid:13)(cid:13)
1
4
≤
+
1
r
,(cid:13)(cid:13)(cid:13)
B2tr + B2(1−t)r(cid:13)(cid:13)(cid:13)(cid:17)
r(cid:17)
+(cid:13)(cid:13)Bt A∗1−t(cid:13)(cid:13)
(by Theorem 2.6).
8
M. BAKHERAD AND K. SHEBRAWI
Corollary 2.8. Let A, B ∈ B(H ) be positive operators. Then, for all t ∈ [0, 1] and
r ≥ 1, we have
1
2 B
A
(cid:13)(cid:13)(cid:13)
r
1
2(cid:13)(cid:13)(cid:13)
≤
1
4
+
1
max(cid:0)(cid:13)(cid:13)Atr + A(1−t)r(cid:13)(cid:13) ,(cid:13)(cid:13)Btr + B2(1−t)r(cid:13)(cid:13)(cid:1)
4(cid:0)(cid:13)(cid:13)AtB1−t(cid:13)(cid:13)
r +(cid:13)(cid:13)BtA1−t(cid:13)(cid:13)
r(cid:1) .
Proof. Since the spectral radius of any operator is dominated by its numerical radius,
then r
1
2 (AB) ≤ w
1
2 (AB) . Applying a commutativity property of the spectral radius,
we get
r
r
2 (AB) = r
= r
= r
1
2 B
1
1
1
2(cid:17)
2(cid:17)
2(cid:17)∗(cid:17)
2(cid:17)∗(cid:13)(cid:13)(cid:13)
1
2
r
1
2 A
1
2 B
1
2 B
1
2 B
1
2 B
1
2 A
1
1
1
1
2 B
2 B
2 (cid:16)A
2 (cid:16)A
2(cid:13)(cid:13)(cid:13)
.
r
1
(2.3)
(cid:3)
r
r
2 (cid:16)A
2 (cid:16)A
2 (cid:16)A
1
r
A
2 B
1
2 B
A
= (cid:13)(cid:13)(cid:13)
= (cid:13)(cid:13)(cid:13)
Now, the result follows from Corollary 2.7.
An important special case of Theorem 2.6, which refines inequality (1.4) can be
stated as follows.
Corollary 2.9. Let A, B ∈ B(H ) and r ≥ 1. Then
kA + Bkr ≤
1
1
A2tr + A2(1−t)r(cid:13)(cid:13)(cid:13)
22−r max(cid:16)(cid:13)(cid:13)(cid:13)
22−r (cid:16)(cid:13)(cid:13)At B1−t(cid:13)(cid:13)
r(cid:17) .
+(cid:13)(cid:13)B∗t A∗1−t(cid:13)(cid:13)
,(cid:13)(cid:13)(cid:13)
B∗2tr + B∗2(1−t)r(cid:13)(cid:13)(cid:13)(cid:17)
r
+
In particular, if A and B are normal, then
kA + Bkr ≤
1
21−r max (kAkr , kBkr) +
1
21−r kABk
r
2 .
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
9
Proof. Applying Lemma 2.2 and Theorem 2.3, we have
kA + B∗kr = kT + T ∗kr
≤
2r
4
≤ 2rmax
= 2rwr (T )
r
θ∈R (cid:13)(cid:13)Re(cid:0)eiθT(cid:1)(cid:13)(cid:13)
max(cid:16)(cid:13)(cid:13)(cid:13)
4 (cid:16)(cid:13)(cid:13)At B∗1−t(cid:13)(cid:13)
2r
(by Theorem 2.6),
+
A2tr + A2(1−t)r(cid:13)(cid:13)(cid:13)
r
B2tr + B2(1−t)r(cid:13)(cid:13)(cid:13)(cid:17)
,(cid:13)(cid:13)(cid:13)
r(cid:17)
+(cid:13)(cid:13)Bt A∗1−t(cid:13)(cid:13)
where T = " 0 A
B 0 # . Now, the desired result follows by replacing B by B∗. For the
particular case, since A and B are normal, then B∗ = B and A∗ = A. Applying
equality (2.3) for the operators A
2 , we have
2 and B
1
1
1
2 B
A
(cid:13)(cid:13)(cid:13)
r
1
2(cid:13)(cid:13)(cid:13)
= r
r
2 (A B)
≤ kA Bk
r
2
= kU ∗AB∗V k
r
2
= kAB∗k
r
2 ,
where A = UA and B = V B are the polar decompositions of the operators A and
B. This completes the proof of the corollary.
(cid:3)
In the next result, we show another generalization of inequality (1.2).
Theorem 2.10. Let A ∈ B(H ) and f, g, h be non-negative non-decreasing continuous
functions on [0, ∞) such that f (x)g(x) = x (x ≥ 0). Then
h (w(A)) ≤
1
2(cid:16)h(cid:16)w(cid:16) Af,g(cid:17)(cid:17) + kh(A)k(cid:17).
Proof. Let A = UA be the polar decomposition of A. Then for every θ ∈ R, we have
kRe(cid:0)eiθA(cid:1) k = r(cid:0)Re(cid:0)eiθA(cid:1)(cid:1)
=
=
=
1
2
1
2
1
2
r(cid:0)eiθA + e−iθA∗(cid:1)
r(cid:0)eiθUA + e−iθAU ∗(cid:1)
r(cid:0)eiθU g(A)f (A) + e−iθf (A)g(A)U ∗(cid:1) .
(2.4)
10
M. BAKHERAD AND K. SHEBRAWI
Now, if we put X = eiθU g(A), Y = f (A), S = e−iθf (A) and T = g(A)U ∗ in
Lemma 2.1, then we get
r(cid:0)eiθU g(A)f (A) + e−iθf (A)g(A)U ∗(cid:1)
1
≤
1
2(cid:16)w(f (A)U g(A)) + w(g(A)U ∗f (A))(cid:17)
2p4ke−iθf (A)g(A)kkg(A)U ∗eiθU f (A)k
(by Lemma 2.1)
+
≤ w(f (A)U g(A)) +pkf (A)kkf (A)kkg(A)kkg(A)k
= w(f (A)U g(A)) +pf (kAk)g(kAk)g(kAk)f (kAk)
= w(f (A)U g(A)) +pkAkkAk
= w(cid:16) Af,g(cid:17) + kAk.
(by the functional calculus)
Using inequalities (2.4), (2.5) and Lemma 2.2 we get
ω(A) = max
θ∈R (cid:13)(cid:13)Re(cid:0)eiθA(cid:1)(cid:13)(cid:13) ≤
1
2(cid:16)w(cid:16) Af,g(cid:17) + kAk(cid:17).
Hence
h (w(A)) ≤ h(cid:18)1
≤
=
1
2
1
2
as required.
(by the monotonicity of h)
2hw(cid:16) Af,g(cid:17) + kAki(cid:19)
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17) +
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17) +
1
2
1
2
h (kAk)
kh(A)k ,
(by the convexity of h)
(2.5)
(cid:3)
Another proof for Theorem 2.3: We can obtain Theorem 2.3 from Theorem
2.10. To see this, first note that by the hypotheses of Theorem 2.3 we have
h(A) = h(g(A)f (A))
≤ h(cid:18) g2(A) + f 2(A)
2
(cid:19)
(by the arithmetic-geometric inequality)
(by the convexity of h).
(2.6)
1
≤
2(cid:0)h(cid:0)g2(A)(cid:1) + h(cid:0)f 2(A)(cid:1)(cid:1)
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
11
Hence, using Theorem 2.10 and inequality (2.6) we get
h (w(A)) ≤
≤
=
1
1
2hh(cid:16)w(cid:16) Af,g(cid:17)(cid:17) + kh(A)ki
2hh(cid:16)w(cid:16) Af,g(cid:17)(cid:17) +
h(cid:16)w(cid:16) Af,g(cid:17)(cid:17) +
1
2
1
1
2(cid:13)(cid:13)h(cid:0)g2(A)(cid:1) + h(cid:0)f 2(A)(cid:1)(cid:13)(cid:13)i
4(cid:13)(cid:13)h(cid:0)g2(A)(cid:1) + h(cid:0)f 2(A)(cid:1)(cid:13)(cid:13) .
Remark 2.11. For the special case f (x) = xt and g = x1−t (t ∈ [0, 1]), we obtain the
inequality (1.2)
w(A) ≤
1
2(cid:16)w(cid:16) At(cid:17) + kAk(cid:17) ,
where A ∈ B(H ) and t ∈ [0, 1].
Using Theorem 2.10, we get the following result.
Corollary 2.12. Let A, B ∈ B(H ) and f, g be two non-negative non-decreasing con-
tinuous functions such that f (x)g(x) = x (x ≥ 0). Then
2wr " 0 A
B 0 #! ≤ max{kAkr, kBkr} +
1
2(cid:0) kf (B)g(A∗)kr + kf (A)g(B∗)kr(cid:1),
where r ≥ 1.
Proof. Using Theorem 2.10 and inequality (2.2), we have
2wr " 0 A
r
B 0 #! ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
B 0 #(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
" 0 A
+ wr(cid:16) Tf,g(cid:17)
= max{kAkr, kBkr} +(cid:18)1
≤ max{kAkr, kBkr} +
and the proof is complete.
2(cid:2) kf (B)g(A∗)k + kf (A)g(B∗)k(cid:3)(cid:19)r
2(cid:0) kf (B)g(A∗)kr + kf (A)g(B∗)kr(cid:1)
1
(cid:3)
Corollary 2.13. Let A, B ∈ B(H ) and f, g be two non-negative non-decreasing con-
tinuous functions on [0, ∞) such that f (x)g(x) = x (x ≥ 0). Then
kA + Bk ≤ max{kAk, kBk} +
1
2(cid:16)(cid:13)(cid:13)f (B)g(A)(cid:13)(cid:13) +(cid:13)(cid:13)f (A∗)g(B∗)(cid:13)(cid:13)(cid:17).
12
M. BAKHERAD AND K. SHEBRAWI
Proof. Let T =" 0 A
B 0 #. Then
(cid:13)(cid:13)A + B
∗(cid:13)(cid:13) = (cid:13)(cid:13)T + T
≤ 2max
∗(cid:13)(cid:13)
θ∈R (cid:13)(cid:13)Re(cid:0)eiθT(cid:1)(cid:13)(cid:13)
= w(T )
≤ max{kAk, kBk} +
(by Lemma 2.2)
1
2(cid:16)(cid:13)(cid:13)f (B)g(A∗)(cid:13)(cid:13) +(cid:13)(cid:13)f (A)g(B∗)(cid:13)(cid:13)(cid:17)
(by Theorem 2.10).
If we replace B by B∗, then we get the desired result.
(cid:3)
In the last results, we present some upper bounds for operator matrices. For this
purpose, we need the following lemma.
Lemma 2.14. [10, Theorem 1] Let A ∈ B(H ) and x, y ∈ H be any vectors.
If
f , g are non-negative continuous functions on [0, ∞) which are satisfying the relation
f (x)g(x) = x (x ≥ 0), then
hAx, yi 2 ≤(cid:10)f 2(A)x, x(cid:11) (cid:10)g2(A∗)y, y(cid:11) .
Theorem 2.15. Let A, B, C, D ∈ B(H ) and fi, gi (1 ≤ i ≤ 4) be non-negative con-
tinues functions such that fi(x)gi(x) = x (1 ≤ i ≤ 4) for all x ∈ [0, ∞). Then
ω " A B
C D #! ≤ maxn(cid:13)(cid:13)f 2
1 (A) + g2
2(B∗) + f 2
+ maxnkg1(A∗)k ,(cid:13)(cid:13)f 2
2 (B) + g2
3(C ∗) + f 2
3 (C)(cid:13)(cid:13)
1
2 , kg4(D∗)ko
2o .
4 (D)(cid:13)(cid:13)
1
Proof. Let T = " A B
C D # and x = " x1
x2 # be a unit vector (i.e., kx1k2 + kx2k2 = 1).
Then
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
13
≤ hAx1, x1i + hBx2, x1i + hCx1, x2i + hDx2, x2i
1
1
1
= hAx1, x1i + hBx2, x1i + hCx1, x2i + hDx2, x2i
hT x, xi = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
x2 #+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
*" A B
C D #" x1
x2 # ," x1
x2 #+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
= (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Cx1 + Dx2 # ," x1
*" Ax1 + Bx2
(cid:10)f 2
2 +(cid:10)f 2
2 (cid:10)g2
1 (A)x1, x1(cid:11)
1(A∗)x1, x1(cid:11)
+(cid:10)f 2
2 +(cid:10)f 2
2 (cid:10)g2
3 (C)x1, x1(cid:11)
3(C ∗)x2, x2(cid:11)
2(B ∗)x1, x1(cid:11) +(cid:10)f 2
1 (A)x1, x1(cid:11) +(cid:10)g2
≤ (cid:0)(cid:10)f 2
2 (B)x2, x2(cid:11) +(cid:10)g2
1(A∗)x1, x1(cid:11) +(cid:10)f 2
+(cid:0)(cid:10)g2
3 (C)(cid:1) x1, x1(cid:11) +(cid:10)g2
= (cid:0)(cid:10)(cid:0)f 2
4 (D)(cid:1) x2, x2(cid:11) +(cid:10)g2
+(cid:0)(cid:10)(cid:0)f 2
3 (C)(cid:13)(cid:13) kx1k2 +(cid:13)(cid:13)g2
≤ (cid:0)(cid:13)(cid:13)f 2
+(cid:0)(cid:13)(cid:13)f 2
4 (D)(cid:13)(cid:13) kx2k2 +(cid:13)(cid:13)g2
2(B ∗) + f 2
2(B ∗) + f 2
3(C ∗) + f 2
3(C ∗) + f 2
2 (B) + g2
2 (B) + g2
1 (A) + g2
1 (A) + g2
1
(by the Cauchy-Schwarz inequality)
1
1
2
1
1
1
2
2
2 (cid:10)g2
2(B ∗)x1, x1(cid:11)
2 (B)x2, x2(cid:11)
2 (cid:10)g2
4 (D)x2x2(cid:11)
4(D∗)x2, x2(cid:11)
3 (C)x1, x1(cid:11) +(cid:10)g2
4(D∗)x2, x2(cid:11)(cid:1)
3(C ∗)x2, x2(cid:11) +(cid:10)f 2
4 (D)x1, x1(cid:11)(cid:1)
2(B ∗)x2, x2(cid:11)(cid:1)
1(A∗)x1, x1(cid:11)(cid:1)
4(D∗)(cid:13)(cid:13) kx2k2(cid:1)
1(A∗)(cid:13)(cid:13) kx1k2(cid:1)
2 .
2
1
2
1
1
2
1
1
2
Let
1 (A) + g2
2 (B) + g2
2(B∗) + f 2
3(C ∗) + f 2
α =(cid:13)(cid:13)f 2
µ =(cid:13)(cid:13)f 2
3 (C)(cid:13)(cid:13) ,
4 (D)(cid:13)(cid:13)
β =(cid:13)(cid:13)g2
and λ =(cid:13)(cid:13)g2
4(D∗)(cid:13)(cid:13) ,
1(A∗)(cid:13)(cid:13) .
It follows from
max
kx1k2+kx2k2=1(cid:0)αkx1k2 + βkx2k2(cid:1) = max
θ∈[0,2π]
(α sin2 θ + β cos2 θ) = max{α, β}
and
max
kx1k2+kx2k2=1(cid:0)λkx1k2 + µkx2k2(cid:1) = max
θ∈[0,2π]
(λ sin2 θ + µ cos2 θ) = max{λ, µ}
14
that
M. BAKHERAD AND K. SHEBRAWI
1
2
3(C ∗) + f 2
1 (A) + g2
2(B∗) + f 2
2(B∗) + f 2
1 (A) + g2
2 (B) + g2
hT x, xi ≤ (cid:0)(cid:13)(cid:13)f 2
+(cid:0)(cid:13)(cid:13)f 2
≤ maxn(cid:13)(cid:13)f 2
+ maxn(cid:13)(cid:13)g2
2 ,(cid:13)(cid:13)f 2
= maxn(cid:13)(cid:13)f 2
+ maxnkg1(A∗)k ,(cid:13)(cid:13)f 2
1(A∗)(cid:13)(cid:13)
1 (A) + g2
2 (B) + g2
3 (C)(cid:13)(cid:13) kx1k2 +(cid:13)(cid:13)g2
4(D∗)(cid:13)(cid:13) kx2k2(cid:1)
1(A∗)(cid:13)(cid:13) kx1k2(cid:1)
4 (D)(cid:13)(cid:13) kx2k2 +(cid:13)(cid:13)g2
2o
4(D∗)(cid:13)(cid:13)
2 ,(cid:13)(cid:13)g2
3 (C)(cid:13)(cid:13)
2o
4 (D)(cid:13)(cid:13)
2 , kg4(D∗)ko
3 (C)(cid:13)(cid:13)
2o .
4 (D)(cid:13)(cid:13)
3(C ∗) + f 2
3(C ∗) + f 2
2(B∗) + f 2
2 (B) + g2
1
1
1
Taking the supremum over all unit vectors x we get the desired result.
1
1
1
1
2
(cid:3)
Corollary 2.16. Let A, B, C, D ∈ B(H ). Then
ω " A B
1
C D #! ≤ maxn(cid:13)(cid:13)A2α + B∗2γ + C2µ(cid:13)(cid:13)
2 , kD∗ωko
+ max(cid:8)(cid:13)(cid:13)A∗β(cid:13)(cid:13) ,(cid:13)(cid:13)B2ζ + C ∗2ν + Dκ(cid:13)(cid:13)(cid:9) ,
0 #! ≤ max(cid:8)(cid:13)(cid:13)A∗β(cid:13)(cid:13) ,(cid:13)(cid:13)Bζ(cid:13)(cid:13)(cid:9) +(cid:13)(cid:13)A2α + B∗2γ(cid:13)(cid:13)
2 ,
1
ω " A B
0
where α + β = γ + ζ = µ + ν = ω + κ = 1. In particular,
in which α + β = γ + ζ = 1.
References
1. A. Abu-Omar and F. Kittaneh, A numerical radius inequality involving the generalized Aluthge
transform, Studia Math. 216 (2013), 69–75.
2. A. Abu-Omar and F. Kittaneh, Generalized spectral radius and norm inequalities for Hilbert space
operators, International Journal of Mathematics Vol. 26, No. 11 (2015) 1550097 (9 pages).
3. O. Axelsson, H. Lu and B. Polman, On the numerical radius of matrices and its application to
iterative solution methods, Linear Multilinear Algebra. 37 (1994), 225–238.
4. M. Bakherad and M.S. Moslehian, Complementary and refined inequalities of Callebaut inequality
for operators, Linear Multilinear Algebra 63 (2015), no. 8, 1678–1692.
5. K. Davidson and S.C. Power, Best approximation in C ∗-algebras, J. Reine Angew. Math. 368
(1986) 43-62.
6. K.E. Gustafson and D.K.M. Rao, Numerical Range, The Field of Values of Linear Operators and
Matrices, Springer, New York, 1997.
7. P.R. Halmos, A Hilbert Space Problem Book, 2nd ed., springer, New York, 1982.
SOME GENERALIZATIONS OF THE ALUTHGE TRANSFORM OF OPERATORS
15
8. O. Hirzallah, F. Kittaneh and K. Shebrawi, Numerical radius inequalities for commutators of
Hilbert space operators, Numer. Funct. Anal. Optim. 32 (2011) 739–749.
9. O. Hirzallah, F. Kittaneh and K. Shebrawi, Numerical radius inequalities for certain 2×2, operator
matrices, Integral Equations Operator Theory, 71 (2011) 129–147.
10. F. Kittaneh, Notes on some inequalitis for Hilbert space operators, Publ. Res. Inst. Math. Sci. 24
(2) (1988), 283–293.
11. F. Kittaneh, A numerical radius inequality and an estimate for the numerical radius of the Frobe-
nius companion matrix, Studia Math. 158 (2003), 11–17.
12. K. Okubo, On weakly unitarily invariant norm and the Aluthge transformation, Linear Algebra
Appl. 371 (2003) 369–375.
13. K. Shebrawi, Numerical radius inequalities for certain 2 × 2 operator matrices II, Linear Algebra
Appl. (2017), http://dx.doi.org/10.1016/j.laa.2017.02.019
14. K. Shebrawi and H. Albadawi, Numerical radius and operator norm inequalities, J. Math. Inequal.
(2009) Article ID 492154, 11 pages.
15. T. Yamazaki, On upper and lower bounds of the numerical radius and an equality condition, Studia
Math. 178 (2007), 83–89.
1Department of Mathematics, Faculty of Mathematics, University of Sistan and
Baluchestan, Zahedan, I.R.Iran.
E-mail address: [email protected]; [email protected]
2Department of Mathematics, Al-Balqa' Applied University, Salt, Jordan.
E-mail address: [email protected]; [email protected]
|
1407.4462 | 2 | 1407 | 2016-01-15T23:39:55 | Weighted discrete hypergroups | [
"math.FA"
] | Weighted group algebras have been studied extensively in Abstract Harmonic Analysis where complete characterizations have been found for some important properties of weighted group algebras, namely amenability and Arens regularity. One of the generalizations of weighted group algebras is weighted hypergroup algebras. Defining weighted hypergroups, analogous to weighted groups, we study Arens regularity and isomorphism to operator algebras for them. We also examine our results on three classes of discrete weighted hypergroups constructed by conjugacy classes of FC groups, the dual space of compact groups, and hypergroup structure defined by orthogonal polynomials. We observe some unexpected examples regarding Arens regularity and operator isomorphisms of weighted hypergroup algebras. | math.FA | math |
Weighted discrete hypergroups
Mahmood Alaghmandan and Ebrahim Samei
October 9, 2018
Abstract
Weighted group algebras have been studied extensively in Abstract Harmonic Analysis where
complete characterizations have been found for some important properties of weighted group
algebras, namely amenability and Arens regularity. One of the generalizations of weighted
group algebras is weighted hypergroup algebras. Defining weighted hypergroups, analogous to
weighted groups, we study Arens regularity and isomorphism to operator algebras for them.
We also examine our results on three classes of discrete weighted hypergroups constructed by
conjugacy classes of FC groups, the dual space of compact groups, and hypergroup structure
defined by orthogonal polynomials. We observe some unexpected examples regarding Arens
regularity and operator isomorphisms of weighted hypergroup algebras.
MSC 2010 classifications: 43A62, 43A77, 43A30, 20F24
Keywords:
groups, polynomial hypergroups, Arens regularity, injectivity, operator algebra isomorphism.
hypergroups, hypergroup algebras, weighted hypergroups, compact groups, FC
1 Introduction
Discrete hypergroups were defined as a generalization of (discrete) groups. Also, some objects
related to locally compact groups may be studied as discrete hypergroups. For instance, double
cosets of a locally compact group with respect to a compact open subgroup. In particular, this
class includes the hypergroup structures on conjugacy classes of an FC group (i.e. every conjugacy
class is finite). Also, for a compact group G, the set of equivalence classes of irreducible (unitary)
representations of G, denoted by (cid:98)G and called the dual of the group G, is a commutative discrete
hypergroup. On one hand these examples as well as hypergroups defined by orthogonal polynomials
connect the studies done on hypergroups to different topics in abstract harmonic analysis. On the
other hand, the similarities of hypergroups and groups suggest that one may be able to generalize
the studies on groups to hypergroups.
One of the topics related to hypergroups which has been initiated based on a similar study on
groups is weighted hypergroups and weighted hypergroup algebras, as they are defined in the following.
The weighted hypergroup algebra, as a Banach algebra can be the subject of study for different
properties of Banach algebras. The first studies over weighted hypergroup algebras may be tracked
back to [4, 11, 12].
1
In this manuscript, we study Arens regularity and isomorphism to operator algebras for weighted
hypergroup algebras. To recall, the second dual of a Banach algebra can be equipped with two
algebraic actions to form Banach algebras, we call a Banach algebra 'Arens regular' if these two
actions coincide. Also a Banach algebra A is called an operator algebra if there is a Hilbert space
H such that A is a closed subalgebra of B(H). The main result of [23] rules out Arens regularity
(and subsequently operator algebra isomorphism) of weighted hypergroup algebras for non-discrete
hypergroups. Consequently, this paper is only dedicated to discrete hypergroups, although many
results proved in Sections 3 hold for weights on non-discrete hypergroups as well. In this manuscript,
we particularly examine our results on various classes of weighted hypergroup algebras with respect
to these properties. One may note that, for the specific weight ω ≡ 1, the weighted case is reduced
back to regular hypergroups and their algebras.
The paper is organized as follows. We start this paper by Section 2 wherein we give the definition
of discrete hypergroups consistently and briefly go through three classes of hypergroup structures
we use in examples. Section 3 is devoted to weights on (discrete) hypergroups, their corresponding
algebras, and their examples. We continue this section by studying some examples. In particular,
in Subsection 3.3, we introduce and study some hypergroup weights on the dual of compact groups.
Arens regularity of weighted group algebras has been studied by Craw and Young in [8]. They
showed that a locally compact group G has a weight ω such that L1(G, ω) is Arens regular if and
only if G is discrete and countable. They also characterized the Arens regularity of weighted group
algebras with respect to one feature of the (group) weight, called 0-clusterness as described in [9]. In
Section 4, the Arens regularity of weighted hypergroup algebras for discrete hypergroups is studied
and it is shown (Theorem 4.4) that the strong 0-clusterness of the corresponding hypergroup weight
results in the Arens regularity of the weighted hypergroup algebra (strong 0-clusterness implies 0-
clusterness, [9]).
Injectivity and equivalently isomorphism of weighted group algebras to operator algebras have
been studied before, see [19, 24]. In Section 5, studying the hypergroup case, we demonstrate (The-
orem 5.2) that for hypergroup weights which are weakly additive and whose inverse is 2-summable
over the hypergroup, the weighted hypergroup algebra is injective and hence an isomorphism to
an operator algebra exists. To do so, we apply some results regarding Littlewood multipliers of
hypergroups. This machinery lets us to examine a class of hypergroup weights which are not weakly
additive, namely exponential weights, in Subsection 5.1.
In Sections 4 and 5 we present many examples to highlight some unexpected contrasts with
some results in the theory of weighted Fourier algebras on compact groups (Examples 4.6, 4.10 and
5.5).
Some results of this paper were first presented in the first author's Ph.D. thesis, [2], under the
supervision of Yemon Choi and Ebrahim Samei.
2 Discrete hypergroups and examples
In this paper, H is always a discrete hypergroup in the sense of [15] unless otherwise is stated.
For basic definitions and facts we refer the reader to the fundamental paper of Jewett, [15], or the
2
comprehensive book [6].
2.1 Definition
(cid:80)
Let H be a discrete set. Let (cid:96)1(H) denote the Banach space of all functions (bounded measures)
f : H → C which are absolutely summable with respect to the counting measure, i.e. (cid:107)f(cid:107)1 :=
x∈H f (x) < ∞. Let cc(H) and c0(H) denote respectively the space of all finitely supported and
vanishing at infinity elements of (cid:96)∞(H). We call H a discrete hypergroup if the following conditions
hold.
(H1) There exists an associative binary operation ∗ called convolution on (cid:96)1(H) under which (cid:96)1(H)
is a Banach algebra. Moreover, for every x, y in H, δx ∗ δy is a positive measure with a finite
support and (cid:107)δx ∗ δy(cid:107)(cid:96)1(H) = 1.
(H2) There exists an element (necessarily unique) e in H such that δe ∗ δx = δx ∗ δe = δx for all x
in H.
(H3) There exists a (necessarily unique) bijection x → x of H called involution satisfying (δx ∗ δy
δy ∗ δx for all x, y ∈ H.
) =
(H4) e belongs to supp(δx ∗ δy) if and only if y = x.
We call a hypergroup H commutative if (cid:96)1(H) forms a commutative algebra. The left translation
on (cid:96)∞(H) is defined by Lxf : H → C where Lxf (y) := f (δx ∗ δy) for each f in (cid:96)∞(H) and x, y ∈ H.
A non-zero, positive, left invariant linear functional h (possibly unbounded) on cc(H) is called a
Haar measure, [15]. For a discrete hypergroup the existence of a Haar measure is proved and it
is unique up to multiplication by a positive constant. Indeed, for a discrete hypergroup, a Haar
measure h : H → (0,∞) such that h(e) = 1 is defined by h(x) = (δx ∗ δx(e))−1 for all x ∈ H.
with respect to the Haar measure h equipped with the convolution f ∗h g :=(cid:80)
The hypergroup algebra, denoted by L1(H, h) is the Banach algebra of integrable functions on H
x∈H f (x)Lxgh(x).
It is easy to observe that f (cid:55)→ f h is an isometric algebra isomorphism from the Banach algebra
L1(H, h) onto the Banach algebra (cid:96)1(H). Due to this isomorphism, we focus our study on (cid:96)1(H)
without loss of generality.
2.2 The conjugacy classes of FC groups
Let G be a (discrete) group with the group algebra (cid:96)1(G) and Conj(G) is the set of all conjugacy
classes of G. We denote the centre of the group algebra by Z(cid:96)1(G). The group G is called an
FC or finite conjugacy group if for each C ∈ Conj(G), C < ∞. For such groups, Conj(G) forms a
commutative discrete hypergroup (which is the discrete case of the hypergroup structures defined
in [15, Subsection 8.3]). Let Ψ denote the linear mapping from Z(cid:96)1(G) to (cid:96)1(Conj(G)) defined by
Ψ(f )(C) = Cf (C) for C ∈ Conj(G) where Ψ(f )(C) := f (x) for (every) x ∈ C. Then one can
easily check that Ψ is an isometric Banach algebra isomorphism between (cid:96)1(Conj(G)) and Z(cid:96)1(G).
3
2.3 The dual of compact groups
i∈I Gi be the restricted direct product
of {Gi}i∈I. Then G is a discrete FC group and Conj(G) is the hypergroup generated by the
As an extension of finite products of hypergroups (or in particular groups), let {Hi}i∈I be a
i∈I Hi where for each x ∈ H, x = (xi)i∈I where xi is
the identity of the hypergroup Hi, eHi, for all i ∈ I except finitely many. H is called restricted direct
product of {Hi}i∈I which is a hypergroup (or a group if for every i, Hi is a group).
family of discrete hypergroups, then H :=(cid:76)
Example 2.1 For a family of FC groups {Gi}i∈I, let G :=(cid:76)
restricted direct product of {Conj(Gi)}i∈I, Conj(G) =(cid:76)
Let G be a compact group and (cid:98)G denotes the set of all irreducible unitary (necessary finite-
that the irreducible decomposition of the tensor products of elements of (cid:98)G leads to a discrete
π for dπ the dimension of π ∈ (cid:98)G. (See [6,
commutative hypergroup structure on (cid:98)G where h(π) = d2
(cid:99)SU(2) be the hypergroup of all irreducible representations on SU(2). It is known that (cid:99)SU(2) =
(π(cid:96))(cid:96)∈N0 where N0 := {0, 1, 2,···} and the dimension of π(cid:96) is (cid:96) + 1. Moreover, for all (cid:96), (cid:96)(cid:48), π(cid:96) = π(cid:96)
and π(cid:96) ⊗ π(cid:96)(cid:48) ∼= π(cid:96)−(cid:96)(cid:48) ⊕ π(cid:96)−(cid:96)(cid:48)+2 ⊕···⊕ π(cid:96)+(cid:96)(cid:48). This tensor decomposition is called "Clebsch-Gordan"
Example 1.1.14]).
Example 2.2 Let SU(2) be the compact Lie group of 2 × 2 special unitary matrices on C, and let
dimensional) representations of a compact group G, up to unitary equivalence relation. It is known
i∈I Conj(Gi).
decomposition formula. So using the Clebsch-Gordan formula, we have that
(cid:88)
(cid:96)+(cid:96)(cid:48)
2
r=(cid:96)−(cid:96)(cid:48)
δπ(cid:96) ∗ δπ(cid:96)(cid:48) =
(r + 1)
((cid:96) + 1)((cid:96)(cid:48) + 1)
δπr
where
(cid:88)
b
2
r=a
f (t) = f (a) + f (a + 2) + . . . + f (b − 2) + f (b).
(2.1)
Also π(cid:96) = π(cid:96) and h(π(cid:96)) = ((cid:96) + 1)2 for all (cid:96).
set I. Let G := (cid:81)
Example 2.3 Suppose that {Gi}i∈I is a non-empty family of compact groups for arbitrary indexing
the product topology. Then (cid:98)G is the restricted direct product of hypergroups {(cid:98)Gi}i∈I, (see [14,
i∈I Gi be the product of {Gi}i∈I i.e. G := {(xi)i∈I : xi ∈ Gi} equipped with
Theorem 27.43]).
2.4 Polynomial hypergroups
Let N0 = N ∪ {0}. Hypergroups related to systems of orthogonal polynomials in one variable have
been introduced and studied by Lasser [17] and Voit [25]. Such a hypergroup structure on N0 is
called a polynomial hypergroup which is also discrete and commutative ([6, Section 3.2]).
4
Example 2.4 Let N0 be equipped with the hypergroup convolution δn ∗ δm := (1/2)δn−m +
(1/2)δn+m. This hypergroup structure is called Chebyshev polynomial of the first type. One can
show that the hypergroup algebra of N0 is isomorphic to the subalgebra of symmetric functions on
the Fourier algebra of the torus, i.e. Z±1A(T) := {f + f : f ∈ A(T)} where f (x) = f (−x).
3 Weighted discrete hypergroups and examples
In this section we study weights on discrete hypergroups, their corresponding algebras, and their
examples. Specially we are interested to see concrete examples of weights defined on the classes of
commutative discrete hypergroups which were mentioned in Section 2.
3.1 General Theory
We believe all the definitions and observations in this subsection still hold for non-discrete hyper-
groups (see [11]), but here we are mainly interested in the discrete case.
Definition 3.1 Let H be a discrete hypergroup. We call a function ω : H → (0,∞) a weight if, for
every x, y ∈ H, ω(δx ∗ δy) ≤ ω(x)ω(y). Then we call (H, ω) a weighted hypergroup. Let (cid:96)1(H, ω) be
the set of all complex functions on H such that
(cid:107)f(cid:107)(cid:96)1(H,ω) :=
f (t)ω(t) < ∞.
(cid:88)
t∈H
Then one can easily observe that ((cid:96)1(H, ω),(cid:107) · (cid:107)(cid:96)1(H,ω)) equipped with the (extended) convolution
of (cid:96)1(H) forms a Banach algebra which is called a weighted hypergroup algebra.
• It is easy to see that if ω is a positive function on H such that ω(t) ≤ ω(x)ω(y) for all
t, x, y ∈ H where t ∈ supp(δx ∗ δy), then ω is a weight on H. We call such a weight a central
weight. We will show later that not all hypergroup weights are central. (See Examples 3.10
and 3.11)
• A hypergroup weight ω on H is called weakly additive, if for some C > 0, ω(δx ∗ δy) ≤
C(ω(x) + ω(y)) for all x, y ∈ H.
• Two weights ω1 and ω2 are called equivalent if there are constants C1, C2 such that C1ω1 ≤
ω2 ≤ C2ω2.
Example 3.2 Let {Hi}i∈I be a family of discrete hypergroups with corresponding weights {ωi}i∈I
i∈I ωi(xi) forms a
such that ωi(eHi) = 1 for all i ∈ I except finitely many. Then ω(xi)i∈I := (cid:81)
have H =(cid:83)
A discrete hypergroup H is called finitely generated if for a finite set F ⊆ H with F = F , we
hypergroup weight on the restricted direct product of hypergroups {Hi}i∈I.
n∈N F ∗n then F is called a finite symmetric generator of H. We define
τF : H → N ∪ {0}
5
(3.1)
by τF (x) := inf{n ∈ N : x ∈ F ∗n} for all x (cid:54)= e and τF (e) = 0. It is straightforward to verify that if
F (cid:48) is another finite symmetric generator of H, then for some constants C1, C2, C1τF (cid:48) ≤ τF ≤ C2τF (cid:48).
If there is no risk of confusion, we may just use τ instead of τF .
Definition 3.3 For a given β ≥ 0, ωβ(x) := (1 + τ (x))β is a central weight on H which is called a
Polynomial weight. Similarly, for given C > 0 and 0 ≤ α ≤ 1, σα,C(x) := eCτ (x)α is a central weight
on H which is called an Exponential weight.
Proposition 3.4 Let H, H(cid:48) be two discrete hypergroups and φ : H1 → H2 be a surjective hypergroup
homomorphism. If ω is a weight on H so that for every x ∈ H, ω(x) ≥ δ for some δ > 0. Then ω(cid:48)
defined by
ω(cid:48)(y) := inf{ω(x) : x ∈ H, φ(x) = y}
(y ∈ H(cid:48)),
is a weight on H(cid:48).
Proof. Proof is immediate if one note that φ : cc(H) → cc(H(cid:48)) satisfies (cid:107)φ(f )(cid:107)(cid:96)1(H(cid:48),ω(cid:48)) ≤ (cid:107)f(cid:107)(cid:96)1(H,ω)
and
ω(cid:48)(δφ(x) ∗ δφ(z)) = (cid:107)δφ(x) ∗ δφ(z)(cid:107)(cid:96)1(H(cid:48),ω(cid:48)) = (cid:107)φ(δx ∗ δz)(cid:107)(cid:96)1(H(cid:48),ω(cid:48)) ≤ (cid:107)δx(cid:107)(cid:96)1(H,ω)(cid:107)δz(cid:107)(cid:96)1(H,ω) = ω(x)ω(z)
for every pair x, z ∈ H.
(cid:3)
3.2 Weights on Conj(G)
Let (G, σ) be a weighted group i.e. σ(xy) ≤ σ(x)σ(y) for all x, y ∈ G. We use (cid:96)1(G, σ) to denote
the weighted group algebra constructed by σ. Let Z(cid:96)1(G, σ) denote the center of (cid:96)1(G, σ). It is not
hard to show that Z(cid:96)1(G, σ) is the set of all f ∈ (cid:96)1(G, σ) for them f (yxy−1) = f (x) for all x, y ∈ G.
The following proposition lets us apply group weights to generate hypergroup weights on
Conj(G). The proof is straightforward, so we omit it here.
by ωσ(C) := C−1(cid:80)
Proposition 3.5 Let G be an FC group possessing a weight σ. Then the mean function ωσ defined
t∈C σ(t) (C ∈ Conj(G)) is a weight on the hypergroup Conj(G). Further,
(cid:96)1(Conj(G), ωσ) is isometrically Banach algebra isomorphic to Z(cid:96)1(G, σ).
Remark 3.6 Let G be an FC group and let ω be a central weight on Conj(G). Then the mapping
σω, defined on G by σω(x) := ω(Cx), is a group weight on G. And (cid:96)1(Conj(G), ω) as a Banach
algebra is isometrically isomorphic to Z(cid:96)1(G, σω).
Example 3.7 Let G be a discrete FC group. The mapping ω(C) = C, for C ∈ Conj(G), is a
central weight on Conj(G).
6
Example 3.8 Let G = (cid:76)
i∈I Gi for a family of finite groups {Gi}i∈I. Given C = (Ci)i∈I ∈
Conj(G), define IC := {i ∈ I : Ci
(cid:54)= eGi}. For each α > 0, we define a mapping ωα(C) :=
(1 + Ci1 + ··· + Cin)α where ij ∈ IC. We show that ωα is a central weight on Conj(G). To do
so let E ⊆ CD for some E, C, D ∈ Conj(G). One can easily show that for each i ∈ I, Ei ⊆ CiDi;
IE ⊆ IC ∪ ID. Therefore,
ωα(C) = (1 +
Ei)α ≤ (1 +
CiDi)α
(by Example 3.7)
(cid:88)
1 +
(cid:88)
i∈IE
i∈IC
(cid:88)
α 1 +
(cid:88)
i∈IE
i∈ID
α
Di
≤
Ci
= ωα(C)ωα(D).
A group G is called a group with finite commutator group or FD if its derived subgroup is finite. It
is immediate that for a group G, for every C ∈ Conj(G), C ≤ G(cid:48) when G(cid:48) is the derived subgroup
of G. Therefore, the order of conjugacy classes of an FD group are uniformly bounded by G(cid:48).
The converse is also true, that is for an FC group G, if the order of conjugacy classes are uniformly
bounded, then G is an FD group, see [22, Theorem 14.5.11]. The following proposition implies that
every hypergroup weight on the conjugacy classes of an FD group which is constructed by a group
weight (as given in Proposition 3.5) is equivalent to a central weight. We omit the proof of the
following proposition as it is straightforward.
Proposition 3.9 Let (G, σ) be a weighted FD group. Then the hypergroup weight ωz(C) := G(cid:48)2ωσ(C),
for C ∈ Conj(G), forms a central weight. Here ωσ is defined as in Proposition 3.5.
In contrast to Proposition 3.9, we will see in the following examples that there exist weights on
FC groups (with infinite derived subgroup) which are not equivalent to any central weight.
Example 3.10 Let S3 be the symmetric group of order 6. Let ω be defined on Conj(S3) by ω(Ce) =
1, ω(C(12)) = 2, and ω(C(123)) = 5. One may verify that ω is a weight on Conj(S3). On the other
hand, since 5 = ω(C(123)) (cid:10) ω(C(12))2 = 4, ω is not a central weight.
Example 3.11 We generate the restricted direct product G =(cid:76)
ω(cid:48) :=(cid:81)
For each N ∈ N, define DN :=(cid:81)
n = Ce otherwise. One can verify that DN ∈ supp(δEN ∗ δEN ) for EN =(cid:81)
n∈N S3. Let us define the weight
n∈N ω on Conj(G) where ω is the hypergroup weight on Conj(S3) defined in Example 3.10.
n = C(123) for all n ∈ 1, . . . , N and
n ∈ Conj(G)
n ∈ Conj(G) where D(N )
n∈N D(N )
n∈N E(N )
n = C(12) for all n ∈ 1, . . . , N and E(N )
n = Ce otherwise. Therefore
D(N )
with E(N )
ω(cid:48)(DN )
ω(cid:48)(EN )2 =
N(cid:89)
n=1
ω(C(123))
ω(C(12))2 = (5/4)N → ∞
where N → ∞. Hence, ω(cid:48) is not equivalent to any central weight.
7
We close this subsection with the following corollary of Lemma 3.4.
Corollary 3.12 Let G be an FC group, N a normal subgroup of G, and ω a weight on Conj(G)
such that there is some δ > 0 such that ω(C) > δ, for any C ∈ Conj(G). Then the mapping
ω : Conj(G/N ) → R+ defined by ω(CxN ) := inf{ω(Cxy) : y ∈ N}, for CxN ∈ Conj(G/N ), forms a
weight on Conj(G/N ).
3.3 Weights on duals of compact groups
In this subsection, G is a compact group. We recall that for each π ∈ (cid:98)G and f ∈ L1(G),
(cid:98)f (π) :=
(cid:90)
G
f (x)π(x)dx
is the Fourier transform of f at π. Let V N (G) denote the group von Neumann algebra of G, i.e. the
von Neumann algebra generated by the left regular representation of G. It is well-known that the
predual of V N (G), denoted by A(G), is a Banach algebra of continuous functions on G; it is called
the Fourier algebra of G. Moreover, for every f ∈ A(G),
(cid:107)f(cid:107) :=
(cid:88)
π∈(cid:98)G
dπ(cid:107)(cid:98)f (π)(cid:107)1 < ∞,
where (cid:107) · (cid:107)1 denotes the trace-class operator norm (look at [14, Section 32]).
In an attempt to find the noncommutative analogue of weights on groups, Lee and Samei in [18]
defined a weight on A(G) to be a densely defined (not necessarily bounded) operator W affiliated
with V N (G) and satisfying certain properties mentioned in [18, Definition 2.4] (see also [20]).
Specially they assume that W has a bounded inverse, W −1, which belongs to V N (G). For a weight
W on A(G), the Beurling-Fourier algebra denoted by A(G, W ) is defined to be the set of all f ∈ A(G)
such that
(cid:107)f(cid:107)A(G,W ) :=
dπ(cid:107)(cid:98)f (π) ◦ W(cid:107)1 < ∞.
(cid:88)
π∈(cid:98)G
Indeed (A(G, W ),(cid:107) · (cid:107)A(G,W )) forms a Banach algebra with pointwise multiplication. For abelian
groups, the definition of Beurling-Fourier algebra corresponds the classical weighted group algebra
on the dual group. In [18], the authors also studied Arens regularity and isomorphism to operator
algebras for Beurling-Fourier algebras.
Definition 3.13 Let G be a compact group and W a weight on A(G). We define a function ωW :
(cid:98)G → (0,∞) by
(3.2)
where (cid:107)·(cid:107)1 denotes the trace norm and Iπ is the identity matrix corresponding to the Hilbert space
of π.
ωW (π) :=
dπ
(cid:107)Iπ ◦ W(cid:107)1
(π ∈ (cid:98)G),
8
dependently, in [20]), central weights on A(G) are defined.
As a specific class of weights on the Fourier algebra of a compact group G, in [18] (and in-
Indeed, [18, Theorem 2.12] implies
that each central weight W can be represent by a unique function ωW : (cid:98)G → (0,∞) such that
ωW (σ) ≤ ωW (π1)ωW (π2) for all π1, π2, σ ∈ (cid:98)G where σ ∈ supp(δπ1 ∗ δπ2). In this specific case of
hypergroup (cid:98)G. In the following we show that the same is true for a general weight on A(G) as well.
operator weights, ωW matches with our definition in Definition 3.13 for a central weight on the
Theorem 3.14 Let G be a compact group and W a weight on A(G). Then ωW is a weight on
Let us define ZA(G, W ) := {f ∈ A(G, W ) : f (yxy−1) = f (x) for all x ∈ G} which is a Banach
ωW (π) = 1 for every π ∈ (cid:98)G, ZA(G, W ) = ZA(G). For more on ZA(G), look at [3].
algebra with pointwise product and (cid:107) · (cid:107)A(G,W ). Note that for the operator weights W where
the hypergroup (cid:98)G and the weighted hypergroup algebra (cid:96)1((cid:98)G, ωW ) is isometrically isomorphic to
T : X (G) → cc((cid:98)G) by T (χπ) = dπδπ for each π ∈ (cid:98)G. Let f =(cid:80)n
i=1 αiχπi ∈ X (G) for πi ∈ (cid:98)G and
ZA(G, W ).
Proof. Let X (G) denote the linear span of all the characters of G. First define a linear mapping
αi ∈ C. In this case,
(cid:107)T (f )(cid:107)(cid:96)1((cid:98)G,ω) =
=
n(cid:88)
n(cid:88)
i=1
i=1
αidπiω(πi) =
n(cid:88)
i=1
dπi(cid:107) αi
dπi
Iπi ◦ W(cid:107)1 =
i=1
αidπi
n(cid:88)
dπi
(cid:107)Iπi ◦ W(cid:107)1
dπi(cid:107)αi(cid:98)χπi(πi) ◦ W(cid:107)1 = (cid:107)f(cid:107)A(G,W ).
Therefore, T forms a norm preserving linear mapping. To show that T is an algebra homomorphism
cc((cid:98)G) is dense in (cid:96)1((cid:98)G, ωW ). So T can be extended as an algebra isomorphism from ZA(G, W )
note that T (χπ1χπ2) = T (χπ1) ∗ T (χπ2). It is known that X (G) is dense in ZA(G, W ) and clearly
onto (cid:96)1((cid:98)G, ωW ) which preserves the norm. In particular, (cid:96)1((cid:98)G, ωW ) forms an algebra with respect
to its weighted norm and the convolution, and so ωW is actually a hypergroup weight on (cid:98)G.
Lemma 3.15 Let G be a compact group and (cid:98)G be the set of all irreducible representations of G as a
The proof of the following lemma is straightforward so we omit it here.
(cid:3)
discrete commutative hypergroup. Then ωβ(π) = dβ
π = h(dπ)β/2 is a central weight for each β ≥ 0.
In the following, recall
defined in (2.1).
Example 3.16 (Lifting weights from Z to (cid:99)SU(2)) Let σ be a weight on the group Z. We define
2
(cid:88)
(cid:88)
(cid:96)
(cid:96) + 1
2
r=−(cid:96)
9
ωσ(π(cid:96)) :=
1
σ(r)
((cid:96) ∈ N0).
(3.3)
Recall that elements of (cid:99)SU(2) can be regarded as π(cid:96) when (cid:96) ∈ N0. Suppose that m, n ∈ N0 and
without loss of generality n ≥ m. Then,
ωσ(πm)ωσ(πn) =
≥
=
=
=
σ(t)
n + 1
σ(s)
(cid:88)
m
1
m + 1
2
t=−m
1
(m + 1)(n + 1)
1
(m + 1)(n + 1)
1
m
(cid:88)
(cid:88)
n+m
2
t=−m
2
t=n−m
n
2
s=−n
n
(cid:88)
(cid:88)
(cid:88)
(cid:32)
2
s=−t
2
s=−n
t
1
σ(t + s)
σ(s)
(cid:88)
t
(†)
(‡)
(cid:33)
σ(s)
n+m
(cid:88)
(cid:88)
n+m
2
t=n−m
2
t=n−m
(t + 1)
(m + 1)(n + 1)
t + 1
2
s=−t
(t + 1)
(m + 1)(n + 1)
ωσ(πt)
= ωσ(δπm ∗ δπn).
To show that the summations (†) and (‡) are equal, let us arrange (†) as follows.
+σ(−m − n + 2) +··· +σ(−m + n − 2) +σ(−m + n)
σ(−m − n)
+σ(−m − n + 2) +σ(−m − n + 4) +··· +σ(−m + n)
...
+σ(m − 2)
+σ(m)
. . .
+··· +σ(m + n − 4)
+··· +σ(m + n − 2)
...
+σ(m)
+σ(m + 2)
...
+σ(−m + n + 2)
...
+σ(m + n − 2)
+σ(m + n) .
but the sum of all the entries in the first column and the last row is equal to
The next column and row give
2
s=−m−n
(cid:88)
m+n−2
(cid:88)
m+n
σ(s) .
σ(s) ,
2
s=−m−n+2
and so on. So by doing this finitely many times, we get (‡). Indeed, weight ωσ follows from the recipe
of Definition 3.13 using the non-central weight W on A(SU(2)) defined in (A.1) in Appendix A. So
instead of the above computations, one also could use Theorem 3.14 to prove that ωσ is actually a
weight on (cid:99)SU(2).
Fix β > 0. One may apply the construction in Example 3.16 for
σ((cid:96)) :=
(1 − (cid:96))β
0 ≤ (cid:96)
(cid:96) < 0
((cid:96) ∈ Z)
(3.4)
(cid:26) 1
10
to construct a hypergroup weight ωσ on (cid:99)SU(2). Observe that the weight ωσ is equivalent to the
weight ωβ defined in Lemma 3.15. We will see in Section 4, that this particular weight will give
interesting classes of examples. To construct weights from subgroups of compact groups, one can
look at [20, Proposition 4.11].
3.4 Weights on polynomial hypergroups
Recall that (cid:99)SU(2) is a particular example of polynomial hypergroup so-called Chebyshev polynomials.
Similar arguments can be applied to construct hypergroup weights on polynomial hypergroups
applying group weights of Z.
Example 3.17 Let f : N0 → R+ be an increasing function such that f (0) = 1. Then ωf (n) =
f (n) + 2 is a central weight on N0 when it is equipped with the Chebyshev polynomial hypergroup
structure of the first type. Applying the argument in Example 2.4, we can see that (cid:96)1(N0, ωf ) is
isomorphic to the symmetric subalgebra of A(T, σ), that is Z±1A(T, σ) := {f + f : f ∈ A(T, σ)},
for the group weight
(cid:26) 1
σf ((cid:96)) :=
4 Arens regularity
(cid:96) ≥ 0
(cid:96) < 0
f (−(cid:96))
((cid:96) ∈ Z)
In [16, Chaptetr 4], Kamyabi-Gol applied the topological center of hypergroup algebras to prove
some results about the hypergroup algebras and their second duals. For example, in [16, Corol-
lary 4.27], he showed that for a (not necessarily discrete and commutative) hypergroup H (which
possesses a Haar measure h), L1(H, h) is Arens regular if and only if H is finite.
Arens regularity of weighted group algebras has been studied by Craw and Young in [8]. They
showed that a locally compact group G has a weight ω such that L1(G, ω) is Arens regular if
and only if G is discrete and countable. The monograph [9] presents a thorough report on Arens
regularity of weighted group algebras. In the following we adapt the machinery developed in [9,
Section 8] for weighted hypergroups. In [9, Section 3], the authors study repeated limit conditions
and give a rich variety of results for them. Here, we will use some of them.
First let us recall the following definitions. Let A be a Banach algebra. For f, g ∈ A, φ ∈ A∗,
and F, G ∈ A∗∗, we define the following module actions.
(cid:104)f · φ, g(cid:105) := (cid:104)φ, gf(cid:105),
(cid:104)φ · F, f(cid:105) := (cid:104)F, f · φ(cid:105),
(cid:104)F♦G, φ(cid:105) := (cid:104)G, φ · F(cid:105),
(cid:104)φ · f, g(cid:105) := (cid:104)φ, f g(cid:105)
(cid:104)F · φ, f(cid:105) := (cid:104)F, φ · f(cid:105)
(cid:104)G(cid:3)F, φ(cid:105) := (cid:104)G, F · φ(cid:105).
Let F, G ∈ A∗∗, and let (fα)α and (gβ)β be nets in A such that fα → F and gβ → G in the weak∗
topology. One may show that for products (cid:3) and ♦ of A∗∗,
F(cid:3)G = w∗ − lim
α
w∗ − lim
β
fαgβ and F♦G = w∗ − lim
β
w∗ − lim
α
fαgβ.
11
The Banach space A∗∗ equipped with either of the multiplications (cid:3) or ♦ forms a Banach algebra.
The Banach algebra A is called Arens regular if two actions (cid:3) and ♦ coincide.
Let c0(H, ω−1) := {f : H → C : f ω−1 ∈ c0(H)}. Note that (cid:96)1(H, ω) is the dual of c0(H, ω−1).
Hence, (cid:96)1(H, ω)∗∗ can be decomposed as (cid:96)1(H, ω)(cid:76) c0(H, ω−1)⊥ when c0(H, ω−1)⊥ := {F ∈
F = (f, Φ) ∈ (cid:96)1(H, ω)(cid:76) c0(H, ω−1)⊥.
(cid:104)F, φ(cid:105) = 0 for all φ ∈ c0(H, ω−1)}. To see this decomposition, let F ∈ (cid:96)1(H, ω)∗∗,
(cid:96)1(H, ω)∗∗ :
it is clear that f := Fc0(H,ω−1) ∈ (cid:96)1(H, ω) and consequently Φ := F − f ∈ c0(H, ω−1)⊥. Therefore,
Proposition 4.1 Let (H, ω) be a weighted hypergroup. Then (cid:96)1(H, ω) is Arens regular if the multi-
plications (cid:3) and ♦ restricted to c0(H, ω−1)⊥ are constantly 0.
Proof. Now let F = (f, Φ) and G = (g, Ψ) belong to (cid:96)1(H, ω)∗∗. First, note that f(cid:3)Ψ = f♦Ψ and
(cid:3)
Φ(cid:3)g = Φ♦g. Thus F(cid:3)G = (f, Φ)(cid:3)(g, Ψ) = (f g, f(cid:3)Ψ + Φ(cid:3)g) = (f g, f♦Ψ + Φ♦g) = F♦G.
Let us define the bounded function Ωω : H × H → (0, 1] by
Ωω(x, y) :=
ω(δx ∗ δy)
ω(x)ω(y)
(x, y ∈ H).
(4.1)
If there is no risk of confusion, we may use Ω instead of Ωω.
For a weighted group (G, σ), the Arens regularity of weighted group algebras has been charac-
terized completely; [9, Theorem 8.11] proves that it is equivalent to the 0-clusterness of the function
Ωσ on G × G, that is
lim
n
lim
m
Ωσ(xm, yn) = lim
m
lim
n
Ωσ(xm, yn) = 0
whenever (xm) and (yn) are sequences in G, each consisting of distinct points, and both repeated
limits exist. A stronger version of 0-clusterness is called strong 0-clusterness (see [9, Section 3]). We
define strongly 0-cluster functions as presented in [9, Definition 3.6] for discrete topological spaces.
Definition 4.2 Let X and Y be two sets and f is a bounded function on X × Y into C. Then f
0-clusters strongly on X × Y if
lim
x→∞ lim sup
y→∞ f (x, y) = lim
y→∞ lim sup
x→∞ f (x, y) = 0.
Let us define Banach space isomorphism κ : (cid:96)1(H, ω) → (cid:96)1(H) where κ(f ) = f ω for each
f ∈ (cid:96)1(H, ω). Note that for κ∗∗ : (cid:96)1(H, ω)∗∗ → (cid:96)1(H)∗∗ and Φ ∈ c0(H, ω)⊥, one gets (cid:104)κ∗∗(Φ), φ(cid:105) =
(cid:104)Φ, κ∗(φ)(cid:105) which is 0 for all φ ∈ c0(H). Therefore κ∗∗(Φ) ∈ c0(H)⊥. The converse (which we do
not use here) is also true and straightforward to show.
The following theorem is a generalization of [9, Theorem 8.8]. In the proof we use some tech-
niques of the proof of [18, Theorem 3.16].
12
Theorem 4.3 Let (H, ω) be a weighted hypergroup and let Ω 0-cluster strongly on H × H. Then
Φ(cid:3)Ψ = 0 and Φ♦Ψ = 0 whenever Φ, Ψ ∈ c0(H, 1/ω)⊥.
Proof. Let us show the theorem for Φ(cid:3)Ψ, the proof for the other action is similar. Let Φ, Ψ ∈
c0(H, 1/ω)⊥. By Goldstine's theorem, there are nets (fα)α, (gβ)β ⊆ (cid:96)1(H) such that fα → κ∗∗(Φ)
and gβ → κ∗∗(Ψ) in the weak∗ topology of (cid:96)1(H)∗∗ while supα (cid:107)fα(cid:107)1 ≤ 1 and supβ (cid:107)gβ(cid:107)1 ≤ 1. So
for each ψ ∈ (cid:96)∞(H) and Φ, Ψ ∈ (cid:96)1(H, ω)∗∗,
(cid:104)ψω, κ∗∗(Φ(cid:3)Ψ)(cid:105) = (cid:104)κ∗(ψ), Φ(cid:3)Ψ(cid:105) = lim
(cid:104)ψω, κ−1(fα) ∗ κ−1(gβ)(cid:105) = lim
(cid:104)ψω, fα/ω ∗ gβ/ω(cid:105).
lim
β
α
lim
β
α
Thus
(cid:104)ψω, κ∗∗(Φ(cid:3)Ψ)(cid:105) = lim
α
lim
β
= lim
α
lim
β
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:88)
y∈H
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:104)ψω, fα/ω ∗ gβ/ω(cid:105)
(cid:88)
ψ(y)ω(y)
(cid:88)
x,z∈H
x,z∈H
fα(x)
ω(x)
≤ lim sup
α
lim sup
β
≤ (cid:107)ψ(cid:107)(cid:96)∞(H) lim sup
α
= (cid:107)ψ(cid:107)(cid:96)∞(H) lim sup
α
lim sup
β
lim sup
β
(cid:88)
(cid:88)
x,z∈H
x,z∈H
fα(x)
ω(x)
gβ(z)
ω(z)
gβ(z)
ω(z)
δx ∗ δz(y)
(cid:88)
fα(x)gβ(z)(cid:88)
y∈H
ψ(y)ω(y)δx ∗ δz(y)
ω(y)
ω(x)ω(z)
δx ∗ δz(y)
y∈H
fα(x)gβ(z)Ω(x, z).
For a given > 0, since by the hypothesis limx lim supz Ω(x, z) = 0, there is a finite set A ⊆ H
x :=
such that for each x ∈ Ac(= H \ A) there exists a finite set Bx ⊆ H such that for each z ∈ Bc
H \ B, Ω(x, z) ≤ . First note that
lim sup
lim sup
α
β
fα(x)gβ(z)Ω(x, z) ≤ lim sup
α
(cid:107)fα(cid:107)1(cid:107)gβ(cid:107)1 ≤ .
lim sup
β
(cid:88)
(cid:88)
x∈Ac
z∈Bc
x
Also according to our assumption about Φ and Ψ and since for each x ∈ H, δx ∈ c0(H, 1/ω),
limα fα(x) = 0 and limβ gβ(x) = 0. So for the given > 0, there is α0 such that for all α0 (cid:52) α,
fα(x) < /A for all x ∈ A. Moreover, for each x ∈ Ac there is some βx
0 such that for all β where
(cid:52) β, gβ(z) < /Bx for all z ∈ Bx (this is possible since A and Bx are finite). Therefore, since
βx
0
Ω(x, z) ≤ 1,
(cid:88)
(cid:88)
x∈A
z∈H
fα(x)gβ(z)Ω(x, z) ≤ lim sup
β
(cid:107)gβ(cid:107)1 =
lim sup
lim sup
α
and
lim sup
lim sup
α
β
β
(cid:88)
(cid:88)
x∈Ac
z∈Bx
(cid:88)
fα(x) lim sup
β
x∈Ac
(cid:107)fα(cid:107)1 = .
(cid:88)
z∈Bx
gβ(z)
fα(x)gβ(z)Ω(x, z) ≤ lim sup
α
≤ lim sup
α
13
But
(cid:88)
x,z∈H
fα(x)gβ(z)Ω(x, z) =
+
+
(cid:88)
(cid:88)
(cid:88)
x∈Ac,z∈Bc
x
x∈A,z∈H
x∈Ac,z∈Bx
fα(x)gβ(z)Ω(x, z)
fα(x)gβ(z)Ω(x, z)
fα(x)gβ(z)Ω(x, z),
and so, one gets that (cid:104)ψω, κ∗∗(Φ(cid:3)Ψ)(cid:105) ≤ 3(cid:107)ψ(cid:107)∞. Since > 0 was arbitrary, this proves the claim
(cid:3)
of the theorem.
Theorem 4.4 Let (H, ω) be a discrete weighted hypergroup and consider the following conditions:
(1) Ω 0-clusters strongly on H × H.
(2) Φ(cid:3)Ψ = Φ♦Ψ = 0 for all Φ, Ψ ∈ c0(H, 1/ω)⊥.
(3) (cid:96)1(H, ω) is Arens regular.
Then (1) ⇒ (2) ⇒ (3).
Proof. (1) ⇒ (2) by Theorem 4.3. (2) ⇒ (3) is implied from Proposition 4.1.
(cid:3)
Remark 4.5 Since in hypergroups, the cancellation does not necessarily exist, the argument of [8,
Theorem 1] cannot be applied to show (3) implies (1).
Example 4.6 Let N0 be equipped with Chebyshev polynomial hypergroup structure of the first
type and σf be the group weight defined in Example 3.17 for an increasing function f . One can
easily check that if limn,m f (n + m)/f (n)f (m) = 0, then Ωωf 0-clusters strongly on N0 × N0; hence,
(cid:96)1(N0, ωf ) is Arens regular. Indeed, Z±A(T, σf ) is Arens regular. But note that A(T, σf ) (which
is isomorphic to (cid:96)1(Z, σf ) through the Fourier transform) is not Arens regular, as Ωσf does not
0-cluster strongly on Z × Z (see [9, Theorem 8.11]).
Corollary 4.7 Let (H, ω) be a weighted discrete hypergroup such that ω is a weakly additive weight.
If 1/ω ∈ c0(H), then (cid:96)1(H, ω) is Arens regular.
Proof. We have
x→∞ lim sup
lim
y→∞
ω(δx ∗ δy)
ω(x)ω(y)
≤ lim sup
x→∞ lim sup
y→∞ C
ω(x) + ω(y)
ω(x)ω(y)
Therefore Ω 0-clusters strongly on H × H and hence (cid:96)1(H, ω) is Arens regular by Theorem 4.4. (cid:3)
= C lim sup
x→∞ lim sup
y→∞
1
ω(x)
+
1
ω(y)
= 0.
14
Corollary 4.8 Let H be a finitely generated hypergroup. Then for each polynomial weight ωβ (β > 0)
on H defined in Definition 3.3, (cid:96)1(H, ωβ) is Arens regular.
Proof. Here we only need to prove the case for an infinite hypergroup H. Let F be a finite generator
of the hypergroup H containing the identity of H rendering the central weight ωβ. Recall that ωβ
is weakly additive with constant C = min{1, 2β−1}. Moreover, for each N ∈ N, for x ∈ H \ F ∗N ,
τF (x) ≥ N ; hence, ωβ(x) = (1 + τF (x))β ≥ (1 + N )β. Therefore, 1/ωβ ∈ c0(H). Subsequently,
(cid:3)
(cid:96)1(H, ωβ) is Arens regular, by Corollary 4.7.
Remark 4.9 Every finitely generated hypergroup H admits a weight for which the corresponding
weighted algebra is Arens regular. On the other hand, an argument similar to [8, Corollary 1] may
apply to show that for every uncountable discrete hypergroup H, H does not have any weight ω
which 0-clusters strongly.
Example 4.10 Let ωβ be as defined in Lemma 3.15 for some β ≥ 0. Then Ωωβ also 0-clusters
strongly on (cid:99)SU(2) × (cid:99)SU(2). Therefore, (cid:96)1((cid:99)SU(2), ωβ), which is isometrically Banach algebra iso-
morphic to ZA(SU(2), ωβ), is Arens regular. On the other hand, A(SU(2), ωβ) is not Arens regular
if β > 0. To observe the later fact, first note that by applying [8], we obtain that (cid:96)1(Z, σ) is not
Arens regular for σ defined in (3.4). Therefore, A(T, σ) is not Arens regular. Note that, ωβ can
also be rendered using the weight σ through the argument of the last paragraph of Subsection 3.3.
For the dual spaces V N (T, σ) and V N (SU(2), ωβ), one may verify that V N (T, σ) embeds ∗-weakly
in V N (SU(2), ωβ) (the details of this embedding will appear in a manuscript by the second named
author and et al). Hence, A(T, σ) is a quotient of A(SU(2), ωβ) and consequently A(SU(2), ωβ) is
not Arens regular.
In the following, we generalize some results on SU(2) to all SU(n)'s, the group of all n×n special
unitary matrices on C, based on a recent study on the representation theory of SU(n), [7]. As an
example for Lemma 3.15, ((cid:99)SU(n), ωβ) is a discrete commutative hypergroup where ωβ(π) = dβ
correspondence between (cid:99)SU(n) and n-tuples (π1, . . . , πn) ∈ Nn
for some β ≥ 0. See [10] for the details of representation theory of SU(n). There is a one-to-one
0 such that π1 ≥ π2 ≥ ··· ≥ πn−1 ≥
πn = 0. This presentation of the representation theory of SU(n) is called dominant weight. Using
this presentation, we have the following formula which gives the dimension of each representation
by the formula
π
(cid:89)
dπ =
1≤i<j≤n
πi − πj + j − i
j − i
(4.2)
where π is the representation corresponding to (π1, . . . , πn). Suppose that π, ν, µ are representations
corresponding to (π1, . . . , πn), (ν1, . . . , νn), and (µ1, . . . , µn), respectively, such that π ∈ supp(δν ∗
δµ). Collins, Lee, and `Sniady showed in [7, Corollary 1.2] there exists some Cn > 0, for each n ∈ N,
such that
(cid:19)
dπ
dµdν
≤ Cn
+
1
1 + ν1
.
(4.3)
(cid:18) 1
1 + µ1
15
Applying (4.3), we prove that ωβ 0-clusters on (cid:99)SU(n).
Proposition 4.11 For every β > 0, (cid:96)1((cid:99)SU(n), ωβ) is Arens regular.
Proof. Let (µm)m∈N and (νk)k∈N be two arbitrary sequences of distinct elements of (cid:99)SU(n). Since,
the elements of (µm)m∈N are distinct, limm→∞ µ(m)
very same thing can be said for νk = (ν(k)
1 , . . . , ν(k)
π ∈ supp(δµm ∗ δνk ), we have
1 = ∞ where µm = (µ(m)
n ). The
n ). For each arbitrary pair (m, k) ∈ N × N, if
, . . . , µ(m)
1
dπ ≤ Cn(
1
1 + µ(m)
1
+
1
1 + ν(k)
1
)dµmdνk .
Hence
ωβ(π) ≤ Cβ
n (
1
1 + µ(m)
1
+
1
1 + ν(k)
1
)βωβ(µm)ωβ(νk).
Therefore
ωβ(δµm ∗ δνk ) =
(cid:88)
π∈(cid:99)SU(n)
Or equivalently
δµm ∗ δνk (π)ωβ(π) ≤ Cβ
n (
1
1 + µ(m)
1
+
1
1 + ν(k)
1
)βωβ(µm)ωβ(νk).
ωβ(δµm ∗ δνk )
ωβ(µm)ωβ(νk)
≤ Cβ
n (
1
+
1
)β.
Ωβ(µm, νk) :=
Hence, limm→∞ lim supk→∞ Ωβ(µm, νk) = limk→∞ lim supm→∞ Ωβ(µm, νk) = 0. Since (cid:99)SU(n) is
countable, this argument implies that Ωβ 0-clusters strongly on (cid:99)SU(n) × (cid:99)SU(n) and, by Theo-
rem 4.4, (cid:96)1((cid:99)SU(n), ωβ) is Arens regular.
(cid:3)
1
1 + µ(m)
1 + ν(k)
1
(cid:76)∞
Example 4.12 Let SL(2, 2n) denote the finite group of special linear matrices over the field F2n
with cardinal 2n, for given n ∈ N. As a direct result of the character table, [1], for each three
conjugacy classes say C1, C2, D ∈ Conj(SL(2, 2n)), D ≤ 2(C1 + C2) if D ⊆ C1C2 for all n.
Let us define the FC group G to be the restricted direct product of {SL(2, 2n)}n∈N i.e. G :=
n=1 SL(2, 2n). Therefore, one can easily show that the weight ωα, defined in Example 3.8, is a
weakly additive weight with the constant M = 2α min{1, 2α−1}. Moreover, since limC→∞ ωα(C) =
∞, (cid:96)1(Conj(G), ωα) is Arens regular, by Corollary 4.7.
Remark 4.13 Let ω be a central weight on Conj(G) for some FC group G. Then there is a group
weight σω, as defined in Remark 3.6, such that (cid:96)1(Conj(G), ω) is isometrically Banach algebra
isomorphic to Z(cid:96)1(G, σω). So one may also use the embedding (cid:96)1(Conj(G), ω) (cid:44)→ (cid:96)1(G, σω) to study
Example 4.12 by applying the theorems which are characterizing Arens regularity of weighted group
algebras.
16
Remark 4.14 Let G be an FC group and σ a group weight on G. We defined ωσ, the derived
weight on Conj(G) from σ in Proposition 3.5. Recall that in this case Z(cid:96)1(G, σ) is isomorphic to
the Banach algebra (cid:96)1(Conj(G), ωσ). If N is a normal subgroup of G, we defined a quotient mapping
Tωσ : (cid:96)1(Conj(G), ωσ) → (cid:96)1(Conj(G/N ), ωσ) in Corollary 3.12 where ωσ(CxN ) = inf{ωσ(Cxy) : y ∈
N} (CxN ∈ Conj(G/N )). Let us note that for an Arens regular Banach algebra A, every quotient
algebra A/I where I is a closed ideal of A is Arens regular as well (see [9, Corollary 3.15]).
Therefore, if (cid:96)1(Conj(G), ωσ) is Arens regular, for every normal subgroup N , (cid:96)1(Conj(G/N ), ωσ),
which is isomorphic to (cid:96)1(Conj(G), ωσ)/ Ker(Tωσ ), is Arens regular.
In the final result of this section, we apply some techniques of [8] to show that for restricted
direct product of hypergroups, product weights fail to admit Arens regular algebras.
Proposition 4.15 Let {Hi}i∈I be an infinite family of non-trivial discrete hypergroups and for each
i∈I Hi
i ∈ I, ωi is a weight on Hi such that ωi(eHi) = 1 for all except finitely many i ∈ I. Let H =(cid:76)
and ω =(cid:81)
i∈I ωi. Then (cid:96)1(H, ω) is not Arens regular.
Proof. Since I is infinite, suppose that N0 × N0 ⊆ I. Define vn = (xi)i∈I where xi = eHi for
all i ∈ I \ (n, 0) and x(n,0) be a non-identity element of H(n,0) for all n ∈ N. Similarly define
um = (xi)i∈I where xi = eHi for all i ∈ I \ (0, m) and x(0,m) be a non-identity element of H(0,m) for
all m ∈ N. Note that for each pair of elements (n, m) ∈ N× N, supp(δvn ∗ δum) forms a singleton in
H; moreover, ω(δvn ∗ δum) = ω(vn)ω(um). Hence, (δvn ∗ δum)(n,m)∈N×N forms a sequence of distinct
elements in (cid:96)1(H).
Let us define fn = δvn and gm = δum for all n, m ∈ N. Suppose that A := {(vn, um) : n > m}
and φ ∈ (cid:96)∞(H) is the characteristic function of the subset A. Clearly, κ−1(fn) = ω−1fn and
κ−1(gm) = ω−1gm belong to (cid:96)1(H, ω) for all n, m and κ∗(φ) = ωφ ∈ (cid:96)∞(H, ω−1), for the Banach
space isomorphism κ : (cid:96)1(H, ω) → (cid:96)1(H) where κ(f ) = f ω for each f ∈ (cid:96)1(H, ω). Note that
(cid:104)ω−1fn ∗ ω−1gm, κ∗(φ)(cid:105) = (cid:104)ω−1fn ∗ ω−1gm, ωφ(cid:105)
(cid:88)
(ω−1fn ∗ ω−1gm)(t)ω(t)φ(t)
=
=
t∈H
ω(vn ∗ um)
ω(vn)ω(um)
= φ(δvn ∗ δum) =
φ(δvn ∗ δum)
(cid:26) 1 if n > m
0 if n ≤ m
Let us recall that for each n and m, (cid:107)fn(cid:107)(cid:96)1(H,ω) = 1 and (cid:107)gm(cid:107)(cid:96)1(H,ω) = 1. So (fn)n∈N and (gm)m∈N,
as two nets in the unit ball of (cid:96)1(H, ω)∗∗, have two subnets (fα)α and (gβ)β such that fα and gβ
converge weakly∗ to some F and G in (cid:96)1(H, ω)∗∗, respectively. Note that for the specific element φ,
defined above, (cid:104)F(cid:3)G, φ(cid:105) = 0 while (cid:104)F♦G, φ(cid:105) = 1. Hence F(cid:3)G (cid:54)= F♦G and consequently (cid:96)1(H, ω)
(cid:3)
is not Arens regular.
17
5 Isomorphism to operator algebras
Let (H, ω) be a weighted discrete hypergroup. In this section, we study the existence of an algebra
isomorphism from (cid:96)1(H, ω) onto an operator algebra. A Banach algebra A is called an operator
algebra if there is a Hilbert space H such that A is a closed subalgebra of B(H). Let A be a Banach
algebra and m : A × A → A is the bilinear (multiplication) mapping m(f, g) = f g. Then A is
called injective, if m has a bounded extension from A ⊗ A into A, where ⊗ is the injective tensor
product. In this case, we denote the norm of m by (cid:107)m(cid:107).
[19, Corollary 2.2.] proves that if a
Banach algebra A is injective then it is isomorphic to an operator algebra. But the converse also
holds for weighted hypergroup algebras. The proof is similar to the group case in [19, Theorem
2.8] and it follows from the little Grothendieck inequality (see [21]). Note that a Banach algebra
which is isomorphic to an operator algebra is always Arens regular ([5, Corollary 2.5.4]).
Injectivity of weighted group algebras has been studied before. Initially Varopoulos, in [24],
studied the group Z equipped with the weight σα(n) = (1 + n)α for all α ≥ 0. This study looked
injectivity of (cid:96)1(Z, σα). He showed that (cid:96)1(Z, σα) is injective if and only if α > 1/2. The
at
manuscript [19], which studied the injectivity question for a wider family of weighted group alge-
bras, developed a machinery applying Littlewood multipliers. In particular, it partially extended
Varopoulos's result to finitely generated groups with polynomial growth. Following the structure
of [19], in this section, we study the injectivity or equivalently isomorphism to operator algebras
for weighted hypergroup algebras.
In this section, A ⊗γ B and A ⊗ B denote respectively the projective and injective tensor
products of Banach spaces A and B.
We know that (cid:96)1(H, ω) ⊗γ (cid:96)1(H, ω) is isometrically isomorphic to (cid:96)1(H × H, ω × ω). Moreover,
(cid:96)1(H × H, ω × ω)∗ is (cid:96)∞(H × H, ω−1 × ω−1). Since the injective tensor norm is minimal among all
cross-norm Banach space tensor norms, the identity map ι : (cid:96)1(H) × (cid:96)1(H) → (cid:96)1(H) × (cid:96)1(H) may
extend to a contractive mapping
ι : (cid:96)1(H) ⊗γ (cid:96)1(H) → (cid:96)1(H) ⊗ (cid:96)1(H).
Since, ι has a dense range,
ι∗ : ((cid:96)1(H) ⊗ (cid:96)1(H))∗ → ((cid:96)1(H) ⊗γ (cid:96)1(H))∗ = (cid:96)∞(H × H)
(5.1)
is an injective mapping. Therefore, applying ι∗, one may embed ((cid:96)1(H)⊗ (cid:96)1(H))∗ into (cid:96)∞(H × H),
as a linear subspace of (cid:96)∞(H × H).
Let H be a discrete hypergroup. We define Littlewood multipliers of H to be the set of all functions
f : H×H → C such that there exist functions f1, f2 : H×H → C where f (x, y) = f1(x, y)+f2(x, y)
for x, y ∈ G such that
(cid:88)
x∈H
sup
y∈H
f1(x, y)2 < ∞ and sup
x∈H
18
(cid:88)
y∈H
f2(x, y)2 < ∞.
We denote the set of all Littlewood multipliers by T 2(H) and define the norm (cid:107) · (cid:107)T 2(H) by
sup
y∈H
(cid:32)(cid:88)
x∈H
(cid:33)1/2
(cid:88)
y∈H
1/2
(cid:107)f(cid:107)T 2(H) := inf
f1(x, y)2
+ sup
x∈H
f2(x, y)2
where the infimum is taken over all possible decompositions f1, f2. Note that for a decomposition
f1, f2 of f ∈ T 2(H),
(cid:107)f(cid:107)(cid:96)∞(H×H) = sup
x,y∈H
f (x, y) ≤ sup
x,y∈H
f1(x, y) + sup
x,y∈H
f2(x, y)
≤ sup
y∈H
f1(x, y)2
+ sup
x∈H
(cid:33)1/2
1/2
< ∞,
f2(x, y)2
(cid:88)
y∈H
(cid:32)(cid:88)
x∈H
since for discrete space H, (cid:96)2(H) ⊆ (cid:96)∞(H) and (cid:107) · (cid:107)∞ ≤ (cid:107) · (cid:107)2. Since f1, f2, in the previous
equation are arbitrary, (cid:107)f(cid:107)(cid:96)∞(H×H) ≤ (cid:107)f(cid:107)T 2(H). Hence T 2(H) ⊆ (cid:96)∞(H × H). Furthermore, for
each φ ∈ (cid:96)∞(H × H) and f ∈ T 2(H), f φ ∈ T 2(H) and (cid:107)f φ(cid:107)T 2(H) ≤ (cid:107)f(cid:107)T 2(H)(cid:107)φ(cid:107)∞.
The following theorem is the hypergroup version of [19, Theorem 2.7]. Since the proof is very
similar to the group case, we omit it here (although with all the details it can be found in [2]).
Here we use KG to denote Grothendieck's constant. First in his celebrated "R´esum´e", Grothendieck
proved the existence of the constant KG in Grothendieck's inequality. For a detailed account of
Grothendieck's constant, its history, and approximations look at [21, Sections 3 and 4].
Theorem 5.1 Let I : T 2(H) → ((cid:96)1(H) ⊗γ (cid:96)1(H))∗ = (cid:96)∞(H × H) be the mapping which takes every
element of T 2(H) to itself as a bounded function on H×H. Then I(T 2(H)) ⊆ ι∗(((cid:96)1(H)⊗ (cid:96)1(H))∗)
for the mapping ι∗ defined in (5.1). Moreover, J := ι∗−1◦I : T 2(H) → ((cid:96)1(H)⊗ (cid:96)1(H))∗ is bounded
and (cid:107)J(cid:107) ≤ KG.
From now on, we identify ((cid:96)1(H) ⊗ (cid:96)1(H))∗) with its image through the mapping ι∗; hence,
J is the identity mapping which takes T 2(H) into ((cid:96)1(H) ⊗ (cid:96)1(H))∗. We present our first main
result of this section. This is a generalization of [19, Theorem 3.1].
Theorem 5.2 Let H be a discrete hypergroup and ω is a weight on H such that Ω, defined in
(4.1), belongs to T 2(H). Then (cid:96)1(H, ω) is injective and equivalently isomorphic to an operator
algebra. Moreover, for the multiplication map m on (cid:96)1(H, ω)⊗ (cid:96)1(H, ω), as defined before, (cid:107)m(cid:107) ≤
KG(cid:107)Ω(cid:107)T 2(H).
Proof. Let Γω : (cid:96)1(H × H, ω × ω) → (cid:96)1(H, ω) such that Γω(f ⊗ g) := f ∗ g for f, g ∈ (cid:96)1(H, ω). The
adjoint of Γω, Γ∗
ω, can be characterized as follows.
ω(φ)(x, y) = (cid:104)Γ∗
Γ∗
ω(φ), δx ⊗ δy(cid:105) = (cid:104)φ, Γω(δx ⊗ δy)(cid:105) = (cid:104)φ, δx ∗ δy(cid:105)
19
for all φ ∈ (cid:96)∞(H, ω−1) and x, y ∈ H. Now we define L from (cid:96)∞(H) to (cid:96)∞(H × H) such that the
following diagram commutes,
(cid:96)∞(H, ω−1)
Γ∗
ω /
/ (cid:96)∞(H × H, ω−1 × ω−1)
P
(cid:96)∞(H)
L
R
/ (cid:96)∞(H × H)
where P (ϕ)(x) = ϕ(x)ω(x) for ϕ ∈ (cid:96)∞(H) and R(φ)(x, y) = φ(x, y)ω−1(x)ω−1(y) for φ ∈ (cid:96)∞(H ×
H, ω−1 × ω−1) and x, y ∈ H. Hence, one gets
L(ϕ)(x, y) = R (Γ∗
ω ◦ P (ϕ)) (x, y) =
(Γ∗
ω ◦ P (ϕ)) (x, y)
ω(x)ω(y)
=
=
=
Γ∗
ω (ωϕ) (x, y)
ω(x)ω(y)
(cid:104)ϕω, δx ∗ δy(cid:105)
ω(x)ω(y)
δx ∗ δy(t)
(cid:88)
t∈H
ω(t)
ω(x)ω(y)
ϕ(t).
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:88)
t∈H
ϕ(t)
δx ∗ δy(t)
ω(t)
ω(x)ω(y)
ϕ(t) ≤ (cid:107)ϕ(cid:107)∞Ω(x, y)
for all ϕ ∈ (cid:96)∞(H). Hence,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:88)
t∈H
δx ∗ δy(t)
ω(t)
ω(x)ω(y)
So there is a function vϕ : H × H → C such that
(cid:104)δx ∗ δy, ωϕ(cid:105)
ω(x)ω(y)
= vϕ(x, y)(cid:107)ϕ(cid:107)∞Ω(x, y)
and (cid:107)vϕ(cid:107)∞ ≤ 1. Therefore L(ϕ) = Λ(ϕ)Ω where Λ(ϕ)(x, y) := vϕ(x, y)(cid:107)φ(cid:107)∞ for all ϕ ∈ (cid:96)∞(H).
Since Ω belongs to T 2(H) and T 2(H) is an (cid:96)∞(H × H)-module, L(ϕ) ∈ T 2(H) and (cid:107)L(ϕ)(cid:107)T 2(H) ≤
(cid:107)ϕ(cid:107)∞(cid:107)Ω(cid:107)T 2(H). Therefore L((cid:96)∞(H)) ⊆ T 2(H) ⊆ ((cid:96)1(H) ⊗ (cid:96)1(H))∗.
In this case, using the following diagram with A = R−1(((cid:96)1(H) ⊗ (cid:96)1(H))∗),
(cid:96)∞(H, ω−1)
P
(cid:96)∞(H)
Γ∗
ω
L
/ A
Rr
/ ((cid:96)1(H) ⊗ (cid:96)1(H))∗
ι
ι
(cid:96)∞(H × H, ω−1 × ω−1)
R
/ (cid:96)∞(H × H)
One can easily verify that A = ((cid:96)1(H, ω) ⊗ (cid:96)1(H, ω))∗. So, we have shown that Γ∗ is a map
ω is a map
projecting (cid:96)∞(H) into ((cid:96)1(H) ⊗ (cid:96)1(H))∗ as a subset of (cid:96)∞(H × H). we see that Γ∗
20
O
O
/
/
/
/
O
O
/
/
projecting (cid:96)∞(H, ω−1) into ((cid:96)1(H, ω)⊗ (cid:96)1(H, ω))∗. Hence, Γ∗
ω = m∗, where m is the multiplication
extended to (cid:96)1(H, ω) ⊗ (cid:96)1(H, ω). Therefore m is bounded and (cid:107)m(cid:107) = (cid:107)Γω(cid:107) = (cid:107)RΓωP(cid:107) = (cid:107)L(cid:107).
Moreover,
(cid:107)L(ϕ)(cid:107)((cid:96)1(H)⊗(cid:96)1(H))∗ ≤ (cid:107)J(cid:107) (cid:107)Γ∗(ϕ)(cid:107)T 2(H) ≤ KG (cid:107)Ω(cid:107)T 2(H) (cid:107)Λ(ϕ)(cid:107)(cid:96)∞(H×H)
≤ KG (cid:107)Ω(cid:107)T 2(H) (cid:107)ϕ(cid:107)(cid:96)∞(H)
for all ϕ ∈ (cid:96)∞(H). Consequently, (cid:107)m(cid:107) ≤ KG(cid:107)Ω(cid:107)T 2(H).
Example 5.3 Let ωβ be the dimension weight defined on (cid:99)SU(n) in Lemma 3.15. As we have shown
in the proof of Proposition 4.11, for the polynomial weight ωβ, β ≥ 0, and µ, ν ∈ (cid:99)SU(n),
(cid:3)
Ωβ(µ, ν) ≤ Cβ
n (
1
+
1
)β ≤ AβCβ
n
1 + µ1
1 + ν1
1
1
(1 + µ1)β +
(1 + ν1)β
where Aβ = min{1, 2β−1}. To study (cid:107) · (cid:107)T 2((cid:99)SU(2)) for Ωβ, let us note that for each k ∈ N ∪ {0},
there are less than (1 + k)n−2 many λ = (λ1, . . . , λn) ∈ (cid:99)SU(n) such that λ1 = k. Therefore
(cid:18)
(cid:19)
,
(cid:88)
λ∈(cid:99)SU(n)
1
(1 + λ1)2β ≤
(1 + k)n−2
(1 + k)2β
∞(cid:88)
k=0
where the right-hand side series converges if and only if 2β−n+2 > 1. Therefore, for β > (n−1)/2,
Ωβ ∈ T 2((cid:99)SU(n)) and by Theorem 5.2, (cid:96)1((cid:99)SU(2), ωβ) is injective and consequently isomorphic to an
operator algebra. Moreover, note that
(cid:107)Ωβ(cid:107)T 2((cid:99)SU(n)) ≤
≤
+
n
1 + µ1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(µ, ν) (cid:55)→ AβCβ
(cid:88)
(cid:88)
ν∈(cid:99)SU(n)
(cid:32) ∞(cid:88)
µ∈(cid:99)SU(n)
µ∈(cid:99)SU(n)
ν∈(cid:99)SU(n)
sup
sup
AβCβ
n
1 + ν1
+
n
1 + µ1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) AβCβ
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) AβCβ
n
1 + ν1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)T 2(H)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)21/2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)21/2
(cid:33)1/2
.
≤ AβCβ
n 2
1
(1 + k)2β−n+2
k=0
Hence, for Aβ = min{1, 2β−1},
(cid:107)m(cid:107) ≤ 2KGAβCβ
n
(cid:33)1/2
1
(1 + k)2β−n+2
(cid:32) ∞(cid:88)
k=0
21
Corollary 5.4 Let H be a discrete hypergroup and ω is a weakly additive weight on H with a corre-
sponding constant C > 0. Then (cid:96)1(H, ω) is injective if(cid:80)
(cid:32)(cid:88)
1
(cid:107)m(cid:107) ≤ 2CKG
x∈H ω(x)−2 < ∞. Moreover,
(cid:33)1/2
ω(x)2
x∈H
.
Proof. Suppose that(cid:80)
x∈H ω(x)−2 < ∞. Note that for each t ∈ supp(δx ∗ δy),
ω(t)
ω(x)ω(y)
≤ C
ω(x) + ω(y)
ω(x)ω(y)
=
C
ω(x)
+
C
ω(y)
.
Thus, for the functions f1(x, y) = ω(x)
−1 and f2(x, y) = ω(y)
−1,
(cid:107)Ω(cid:107)T 2(H) ≤
≤
(cid:13)(cid:13)(cid:13)(cid:13)(x, y) (cid:55)→ C
sup
(cid:32)(cid:88)
(cid:12)(cid:12)(cid:12)(cid:12) C
ω(x)
y∈H
x∈H
ω(x)
+
C
(cid:13)(cid:13)(cid:13)(cid:13)T 2(H)
(cid:12)(cid:12)(cid:12)(cid:12)2(cid:33)1/2
ω(y)
+ sup
x∈H
(cid:88)
y∈H
1/2 ≤ 2C
(cid:12)(cid:12)(cid:12)(cid:12) C
ω(y)
(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:32)(cid:88)
1
ω(x)2
x∈H
(cid:33) 1
2
.
Consequently, by Theorem 5.2, (cid:96)1(H, ω) is injective and (cid:107)m(cid:107) satisfies the mentioned inequality. (cid:3)
Example 5.5 Let ωf be the weight constructed by the group weight admitted by a positive increas-
ing function f (see Example 3.17). One can see that, if
(cid:88)
n∈N0
1
f (n)2 < ∞ and
sup
n,m∈N0
f (n + m)
f (n) + f (m)
< ∞,
then ωf satisfies the conditions of Corollary 5.4 and therefore, (cid:96)1(N0, ωf ) is isomorphic to an
operator algebra. On the other hand, (cid:96)1(N0, ωf ) can be embedded (isomorphically as a Banach
algebra) into A(T, σf ) which is not isomorphic to any operator algebra (as it is not even Arens
regular, see Example 4.6).
Remark 5.6 Note that the assumed condition for f in Example 5.5 implies the Arens regularity
condition required in Example 4.6. Compare it with this know fact that every Banach algebra
which is isomorphic to an operator algebra is Arens regular.
is a weakly additive weight. One can straightforwardly show that(cid:80)
Remark 5.7 Let (Conj(G), ωα) be the weighted hypergroup defined in Example 4.12. Note that ωα
C∈Conj(G) ω(C)−2 = ∞. Hence,
not all weakly additive weights are satisfying the other condition mentioned in Corollary 5.4.
22
For finitely generated hypergroups, we showed that that polynomial weights are weakly additive.
In the following, we study operator algebra isomorphism for weighted hypergroup algebras with
polynomial weights. Developing a machinery which relates exponential weights to polynomial ones,
we also study exponential weights in Subsection 5.1. For the case that H is a group, this has been
achieved in [19]
Corollary 5.8 Let H be a finitely generated hypergroup. If F is a generator of H such that F ∗n ≤
Dnd for some d, D > 0 and ωβ is the polynomial weight on H associated to F . Then (cid:96)1(H, ωβ) is
injective if 2β > d + 1. Moreover, for C = min{1, 2β−1},
(cid:32)
∞(cid:88)
Dnd
(1 + n)2β
n=1
(cid:33)1/2
.
(cid:107)m(cid:107) ≤ 2CKG
1 +
Proof. To prove this corollary, we mainly rely on Corollary 5.4. Recall that ωβ is weakly additive
whose constant is C = min{1, 2β−1}. To show the desired bound for (cid:107)m(cid:107), note that
1
ωβ(x)2 =
1
(1 + τ (x))2β =
1
(1 + n)2β
(cid:88)
x∈H
(cid:88)
x∈H
∞(cid:88)
≤ 1 +
∞(cid:88)
(cid:88)
n=0
{x∈F n\F n−1}
∞(cid:88)
Dnd
F n
(1 + n)2β ≤ 1 +
which is convergent if 2β > d + 1.
n=1
(1 + n)2β
n=1
(cid:3)
Example 5.9 For a polynomial hypergroup N0, as a finitely generated hypergroup with the gener-
ator F = {0, 1}, we have F ∗n = n + 1 ≤ 2n, as we have seen before. By Corollary 5.8, for the
polynomial weight ωβ with β > 1 associated to F , (cid:96)1(N0, ωβ) is injective. For C = min{1, 2β−1},
Corollary 5.4 implies that
(cid:32) ∞(cid:88)
(cid:33)1/2
.
1
n2β
(cid:107)m(cid:107) ≤ 2CKG
5.1 Hypergroups with exponential weights
n=1
The other class of weights introduced for finitely generated hypergroups is the class of exponential
weights. As we mentioned before, unlike polynomial weights, exponential weights are not necessarily
weakly additive. In this subsection, following [19], we study operator algebra isomorphism of these
weights by studying the cases for them Ω belongs to T 2(H). The following lemma is a hypergroup
adaptation of [19, Theorem 3.3]. Since the proof is similar to the one of [13, Lemma B.2], we omit
it here.
23
Lemma 5.10 Suppose that 0 < α < 1, C > 0, and β ≥ max
p : [0,∞) → R and q : (0,∞) → R by p(x) := Cxα − β ln(1 + x) and q(x) := p(x)
finitely generated hypergroup with a symmetric generator F and ω : H → (0,∞) such that
Cα(1−α)
1,
6
. Define the functions
x . Let H be a
(cid:110)
(cid:111)
ω(x) = ep(τF (x)) = eτF (x)q(τF (x))
for all x ∈ H.
Then ω(t) ≤ M ω(x)ω(y) for all t, x, y ∈ H such that t ∈ x ∗ y where
M = max{ep(z1)−p(z2)−p(z3) : z1, z2, z3 ∈ [0, 2K] ∩ N0}
and
K =
(cid:18)
β2
Cα(1 − α)
(cid:19)1/α
.
Theorem 5.11 Let H be a finitely generated hypergroup. If F is a symmetric generator of H such
that F ∗n ≤ Dnd for some d, D > 0 and σα,C is an exponential weight on H for some 0 < α < 1
and C > 0. Then (cid:96)1(H, σα,C) is injective and equivalently isomorphic to an operator algebra.
Proof. Let ωβ be the weight defined in Lemma 5.10. We define a function ω : H → (0,∞) by
ω(x) :=
σα,C(x)
ωβ(x)
= eCτF (x)α−β ln(1+τF (x)) (x ∈ H)
where ωβ is the polynomial weight defined on H associated to F and
β > max{1,
6
Cα(1 − α)
,
d + 1
2
}.
Therefore, by Lemma 5.10, ω(t) ≤ M ω(x)ω(y) for some M > 0 and all t, x, y ∈ H such that
t ∈ x ∗ y. Therefore
σα,C(t)
σα,C(x)σα,C(y)
≤ M
ωβ(t)
ωβ(x)ωβ(y)
.
Therefore,
σα,C(t)
σα,C(x)σα,C(y)
1
(1 + τ (x))β +
1
(1 + τ (y))β
≤ M(cid:48)(cid:18)
(cid:19)
for a modified constant M(cid:48) > 0. Therefore by the proof of Corollary 5.8, Ωσα,C ∈ T 2(H). Now
(cid:3)
Theorem 5.2 finishes the proof.
Example 5.12 As a result of Theorem 5.11, and to follow Example 5.9, if H is a polynomial
hypergroup on N0, for each exponential weight σα,C for 0 < α < 1 and C > 0, (cid:96)1(H, σα,C) is
injective. Note that this class of hypergroups includes (cid:99)SU(2).
24
Acknowledgements
For this research, the first author was supported by a Ph.D. Dean's Scholarship at University of
Saskatchewan and a Postdoctoral Fellowship form the Fields Institute For Research In Mathemat-
ical Sciences and University of Waterloo. The second name author was also supported by NSERC
Discovery grant no 409364 and a generous support from the Fields Institute. The first named
author also would like to express his deep gratitude to Yemon Choi and Nico Spronk for several
constructive discussions and suggestions which improved the paper significantly. The authors also
would like to thank the referee for his many productive comments.
A Appendix: Lifting weights from Z to weights on A(SU(2))
In this appendix, we briefly present a method to construct non-central weights on A(SU(2)) which
are related to Example 3.16. Here λG denotes the left regular representation of a compact group G
on L2(G) and V N (G) denotes the group von Neumann algebra generated by λG. We also identify
T with the (closed) subgroup of all matrices
(cid:21)
(cid:20) t 0
0 t
(t ∈ T)
,
in SU(2). It is an immediate consequence of Herz's restriction theorem that there is a canonical
embedding of V N (T) into V N (SU(2)). More precisely, the mapping Γ : V N (T) → V N (SU(2))
defined by
(cid:90)
T
Γ(λT(f )) =
f (t)λSU(2)(f )dt
is a weak∗-weak∗ isometric ∗-algebra homomorphism. We note that the integration in the definition
of Γ is the Bochnor integration in the weak operator topology of B(L2(SU(2))).
Now suppose that σ is a (group) weight on Z which is bounded below by some δ > 0, i.e. σ−1
belongs to (cid:96)∞(Z). Through the Fourier transform F on Z, F(σ−1) is an element in V N (T) defined
by
F(σ−1)χk = σ(k)−1χk,
where χk(t) = tk (k ∈ Z) are the characters of T.
We now consider the element Γ(F(σ−1)) in V N (SU(2)). Since SU(2) is compact, we can
write λSU(2) as the direct sum of the irreducible unitary representations of SU(2). Moreover, if
we take (cid:99)SU(2) = {π(cid:96) : (cid:96) ∈ N0}, where each π(cid:96) is a representation of dimension (cid:96) + 1, then, by
a straightforward computation based on [14, Theorem 29.18], we have that Γ(F(σ−1))(π(cid:96)) is the
25
(cid:77)
(cid:96)
−
W =
σ(−(cid:96))
0
...
0
0
0
σ(−(cid:96) + 2)
...
0
0
0
0
...
···
···
. . .
··· σ((cid:96) − 2)
···
0
0
0
...
0
σ((cid:96))
,
diagonal matrix diag(σ(−(cid:96))
−1, σ(−(cid:96) + 2)
−1, . . . , σ((cid:96) − 2)
−1, σ((cid:96))
−1). Therefore, if we define,
(A.1)
then W is a (possibly unbounded) operator L2(SU(2)) with W −1 = Γ(F(σ−1)) ∈ V N (SU(2)).
Moreover, it is straightforward to check that W satisfies the assumptions in [18, Definition 2.4]
so that, in particular, it is a Fourier algebra weight on A(SU(2)). We can apply the formula in
Definition 3.13 to W and define the (hypergroup) weight ωσ on SU(2) presented in Example 3.16.
References
[1] J. Adams. Character table of SL(2, F). http://www2.math.umd.edu/jda/characters/sl2/.
[2] Mahmood Alaghmandan. weighted hypergroups and some questions in abstract harmonic analysis. ProQuest
LLC, Ann Arbor, MI, 2013. Thesis (Ph.D.) -- The University of Saskatchewan (Canada).
[3] Mahmood Alaghmandan and Nico Spronk. Amenability properties of the algebra of central fourier algebra of a
compact group. (submited).
[4] H. N. Bhattarai and J. W. Fernandez. Joins of double coset spaces. Pacific J. Math., 98(2):271 -- 280, 1982.
[5] David P. Blecher and Christian Le Merdy. Operator algebras and their modules -- an operator space approach,
volume 30 of London Mathematical Society Monographs. New Series. The Clarendon Press, Oxford University
Press, Oxford, 2004. Oxford Science Publications.
[6] Walter R. Bloom and Herbert Heyer. Harmonic analysis of probability measures on hypergroups, volume 20 of
de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 1995.
[7] Benoıt Collins, Hun Hee Lee, and Piotr ´Sniady. Dimensions of components of tensor products of representations
of linear groups with applications to Beurling-Fourier algebras. Studia Math., 220(3):221 -- 241, 2014.
[8] I. G. Craw and N. J. Young. Regularity of multiplication in weighted group and semigroup algebras. Quart. J.
Math. Oxford Ser. (2), 25:351 -- 358, 1974.
[9] H. G. Dales and A. T.-M. Lau. The second duals of Beurling algebras. Mem. Amer. Math. Soc., 177(836):vi+191,
2005.
[10] William Fulton and Joe Harris. Representation theory, volume 129 of Graduate Texts in Mathematics. Springer-
Verlag, New York, 1991. A first course, Readings in Mathematics.
[11] F. Ghahramani and A. R. Medgalchi. Compact multipliers on weighted hypergroup algebras. Math. Proc.
Cambridge Philos. Soc., 98(3):493 -- 500, 1985.
[12] F. Ghahramani and A. R. Medgalchi. Compact multipliers on weighted hypergroup algebras. II. Math. Proc.
Cambridge Philos. Soc., 100(1):145 -- 149, 1986.
[13] Mahya Ghandehari, Hun Hee Lee, Ebrahim Samei, and Nico Spronk. Some Beurling-Fourier algebras on compact
groups are operator algebras. Trans. Amer. Math. Soc., 367(10):7029 -- 7059, 2015.
26
[14] Edwin Hewitt and Kenneth A. Ross. Abstract harmonic analysis. Vol. II: Structure and analysis for compact
groups. Analysis on locally compact Abelian groups. Die Grundlehren der mathematischen Wissenschaften, Band
152. Springer-Verlag, New York, 1970.
[15] Robert I. Jewett. Spaces with an abstract convolution of measures. Advances in Math., 18(1):1 -- 101, 1975.
[16] Rajab Ali Kamyabi-Gol. Topological center of dual branch algebras associated to hypergroups. ProQuest LLC,
Ann Arbor, MI, 1997. Thesis (Ph.D.) -- University of Alberta (Canada).
[17] Rupert Lasser. Orthogonal polynomials and hypergroups. Rend. Mat. (7), 3(2):185 -- 209, 1983.
[18] Hun Hee Lee and Ebrahim Samei. Beurling-Fourier algebras, operator amenability and Arens regularity. J.
Funct. Anal., 262(1):167 -- 209, 2012.
[19] Hun Hee Lee, Ebrahim Samei, and Nico Spronk. Some weighted group algebras are operator algebras. Proc.
Edinb. Math. Soc. (2), 58(2):499 -- 519, 2015.
[20] Jean Ludwig, Nico Spronk, and Lyudmila Turowska. Beurling-Fourier algebras on compact groups: spectral
theory. J. Funct. Anal., 262(2):463 -- 499, 2012.
[21] Gilles Pisier. Grothendieck's theorem, past and present. Bull. Amer. Math. Soc. (N.S.), 49(2):237 -- 323, 2012.
[22] Derek J. S. Robinson. A course in the theory of groups, volume 80 of Graduate Texts in Mathematics. Springer-
Verlag, New York, second edition, 1996.
[23] A. Ulger. Arens regularity of weakly sequentially complete Banach algebras. Proc. Amer. Math. Soc.,
127(11):3221 -- 3227, 1999.
[24] Nicholas Th. Varopoulos. Sur les quotients des alg`ebres uniformes. C. R. Acad. Sci. Paris S´er. A-B, 274:A1344 --
A1346, 1972.
[25] Michael Voit. Laws of large numbers for polynomial hypergroups and some applications. J. Theoret. Probab.,
3(2):245 -- 266, 1990.
Mahmood Alaghmandan
Department of Mathematical Sciences, Chalmers University of Technology and University of Gothenburg, Gothen-
burg SE-412 96, Sweden
[email protected]
Ebrahim Samei
Department of Mathematics and Statistics, University of Saskatchewan, 142 Wiggins road, Saskatoon, SK S7N
5E6, Canada
[email protected]
27
|
1203.3908 | 2 | 1203 | 2012-03-25T21:10:10 | Normal matrix compressions | [
"math.FA"
] | The recently developed theory of higher--rank numerical ranges originated in problems of error correction in quantum information theory but its mathematical implications now include a quite satisfactory understanding of \emph{scalar} compressions of complex matrices. Here our aim is to make some first steps in the more general program of understanding \emph{normal} compressions. We establish some general principles for the program and make a detailed study of rank--two normal compressions. | math.FA | math |
Normal matrix compressions
6 December 2011
John Holbrook, Nishan Mudalige, Rajesh Pereira
Abstract: The recently developed theory of higher -- rank numerical ranges
originated in problems of error correction in quantum information theory but
its mathematical implications now include a quite satisfactory understand-
ing of scalar compressions of complex matrices. Here our aim is to make
some first steps in the more general program of understanding normal com-
pressions. We establish some general principles for the program and make a
detailed study of rank -- two normal compressions.
AMS codes: MSC(2000) 47A12, 15A60, 15A90, 81P68
Key words and phrases: matrix compression, higher -- rank numerical ranges,
interlacing theorems, quantum information
1: Introduction
Given a linear operator T on a complex Hilbert space H, and any orthogonal
projection P , we say that P TP H is a compression of T . If H = CN and T
is represented by a matrix M ∈ MN (the N × N complex matrices), a second
matrix C represents a compression of T (or a compression of M) iff there is a
unitary matrix U such that C is a NW corner of UMU ∗. If C is k × k we say
it is a rank -- k compression of M. There is a rich history of results that allow
us to identify compressions by means of intrinsic criteria. A classic example
is the Cauchy interlacing theorem [Cau], along with its converse [FP], which
may be expressed as follows.
Theorem 1: If M ∈ MN is Hermitian, with eigenvalues
a1 ≤ a2 ≤ · · · ≤ aN ,
then C is a rank -- k compression of M iff C is Hermitian with eigenvalues bj
satisfying
a1 ≤ b1 ≤ aN −k+1, a2 ≤ b2 ≤ aN −k+2, . . . , ak ≤ bk ≤ aN .
1
In particular, C is a rank N − 1 compression iff
a1 ≤ b1 ≤ a2 ≤ b2 ≤ a3 ≤ ... ≤ aN −1 ≤ bN −1 ≤ aN ,
the classic "interlacing" of eigenvalues.
A much more recent example is provided by the theory of higher -- rank
numerical ranges. The striking development of this theory was motivated
originally by problems in quantum information theory. Since the introduc-
tion of this concept by Choi, Kribs, and Zyczkowski [CK Z1,CK Z2] only a few
years ago, it has indeed been effectively applied in the area of quantum in-
formation (see [CPMS Z,KPLRdS,LP,LPS1,MM Z], for example). It has also
inspired a remarkable development of its purely mathematical aspects (see,
for example, [CHK Z,CGHK,Wo,LS,LPS2,DGHP Z]). From this point of view
the theory of the higher -- rank numerical ranges may be described as a highly
successful analysis of scalar compressions of arbitrary matrices M ∈ MN .
This suggests a more general program: characterize the normal (diagonal)
compressions of M.
In what follows we begin to carry out this program,
although at present the program in its entirety seems out -- of -- reach.
The rank -- k numerical range of M, usually denoted in the literature by
Λk(M), was defined by Choi, Kribs, and Zyczkowski as the set of those
complex λ such that for some rank -- k orthogonal projection P we have
P MP = λP.
In terms of compressions, we see that λ ∈ Λk(M) iff λIk is a (matrix) com-
pression of M. Thus the following fundamental result of Li and Sze [LS] may
be placed in the same family as the Cauchy interlacing theorem (and, in fact,
the interlacing theorem plays a role in the argument of Li and Sze).
Theorem 2: Given M ∈ MN , let λj(θ) be an enumeration of the eigenvalues
of the (Hermitian)
Re(eiθM) = (eiθM + e−iθM ∗)/2
such that
λ1(θ) ≤ λ2(θ) ≤ · · · ≤ λN (θ).
For each real θ, let the half -- plane H(M, θ) be defined by
H(M, θ) = eiθ{z : Re(z) ≤ λN −k+1(−θ)}.
2
Then
Λk(M) = \{H(M, θ) : θ ∈ [0, 2π]}.
(1)
Our more general program seeks to describe all normal compressions of
M, ie to describe those complex a1, . . . , ak such that diag(a1, . . . , ak) is a
compression of M. Equivalently, we ask when there exist orthonormal
u1, u2, . . . , uk
such that (Mui, ui) = ai for each i and (Mui, uj) = 0 whenever i 6= j; in
particular, Λ1(M) is nothing but the classical numerical range
W (M) = {(Mu, u) : kuk = 1}
(hence the "higher -- rank numerical range" terminology).
In this work we
usually restrict our attention to the case where M itself is also normal, al-
though we occasionally comment on cases where either M or its compression
may not be normal.
Note that for normal M ∈ MN (C) Theorem 2 shows that Λk(M) can be
explicitly described in terms of the eigenvalues z1, . . . , zN of M:
Λk(M) = \#(J)=N −k+1
conv{zj : j ∈ J}.
(2)
We shall refer to this result, first proposed by Choi, Kribs, and Zyczkowski,
as the CK Z conjecture, although it is now a theorem. The CK Z conjecture
played an important role in the development of the theory of higher -- rank
numerical ranges. For example, while Li and Sze gave an effective description
of Λk(M) for non -- normal M (Theorem 2), their proof of the CK Z conjecture
was a key step towards the general result. Of course, the case k = 1 of (2) is
easy and well -- known: for normal M, W (M) = conv{z1, . . . , zN}.
The following observation is often useful.
Proposition 3: For every M ∈ MN , if k ≤ N, C is a rank-k compression of
M, and Q is a compression of rank N − k + 1, then
W (C) ∩ W (Q) 6= ∅.
3
Proof: Let S and T be the subspaces corresponding to compressions C
and Q. Since the dimensions add to more than N, S and T must intersect
non -- trivially; let u be a unit vector in S ∩ T . Then
(Mu, u) = (Mu, PSu) = (PSMu, u) = (Cu, u) ∈ W (C),
and similarly (Mu, u) ∈ W (Q). QED
Applying this observation to the normal case, we see that part of the CK Z
conjecture is straightforward.
Proposition 4: If M ∈ MN is normal with eigenvalues z1, . . . , zN , and the
rank -- k compression C is normal with eigenvalues c1, . . . , ck, then for every
index set J having #(J) = N − k + 1
In particular,
conv{c1, . . . , ck} ∩ conv{zj : j ∈ J} 6= ∅.
Λk(M) ⊆ \#(J)=N −k+1
conv{zj : j ∈ J}
(compare (2)).
Proof: We have noted that for normal (finite -- dimensional) operators the
numerical range is just the convex hull of the eigenvalues. Thus W (C) =
conv{c1, . . . , ck}. On the other hand, let Q be the compression to the span
of eigenvectors corresponding to {zj : j ∈ J}; then Q is normal and W (Q) =
conv{zj : j ∈ J}. Apply Proposition 3. In particular, for points λ ∈ Λk(M)
we may let c1 = c2 = · · · = ck = λ. QED
On the other hand, the fact that Λk(M) completely fills the RHS of (2) is
more subtle, in general, although for certain combinations of N and k it is
relatively easy to see. To illustrate this, and to introduce the preoccupations
of the present paper, consider the case N = 5, k = 2. In Figure 1 we see the
eigenvalues z1, . . . , z5 of a normal (in fact, unitary) M as the outer points
of the blue pentagram. It is easy to see that (2) implies that Λ2(M) is the
inner pentagon. As far as we know, there is no simple proof that Λk(M) fills
this pentagon, but three markedly disparate arguments may be found in the
literature:
(1) in [CHK Z] there is an argument based in part on topological concepts
4
such as simple connectivity and winding number;
(2) as it is easy to conclude (see section 2) that the vertices of the inner
pentagon are in Λ2(M), the fact that (whether or not M is normal) Λk(M)
is convex (see [CGHK] and [Wo])) -- a striking extension of the classical
Toeplitz -- Hausdorff Theorem for W (M) -- may be used;
(3) as we have noted, (2) is a direct consequence of the Li and Sze result
Theorem 2.
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
−1
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
Figure 1: Choosing a (red asterisk) at random in Λ2(M) (the inner pentagon),
we see that B(a) includes a "starfish" that covers Λ2(M) and more.
A fourth, and quite different yet again, approach can be obtained by consider-
ing those eigenvalue pairs a, b that can belong to rank -- 2 normal compressions
of M. Given a ∈ C we denote by B(a) the set of b that match a in this sense.
5
We shall prove in section 3 that for a in the inner pentagon B(a) includes a
"starfish" (outlined in green for the example of Figure 1) covering the (filled)
pentagon (our conjecture, in addition, is that the starfish is precisely B(a)).
Since a ∈ B(a) says that a ∈ Λ2(M), we conclude once again that Λ2(M)
fills the pentagon.
Plan of the paper: section 2 has some general results, section 3 treats the
case k = 2, section 4 examines continuity of B(·), and section 5 discusses
non -- normal compressions.
Acknowledgements: We have enjoyed many stimulating discussions of ma-
trix compression, particularly those with M. -- D. Choi, C. -- K. Li, Y. -- T. Poon,
N. -- S. Sze, and J. F. Queir´o. Versions of the material in this paper were
developed in [M]. The work of Holbrook and Pereira was supported in part
by Discovery Grants from NSERC of Canada.
2. Some general results (arbitrary k, N)
Note that if C is a rank -- k compression of M ∈ MN and C ′ is a rank -- k′
compression of C, then C ′ is a rank -- k′ compression of M. Thus Proposition
3 has the following consequence.
Proposition 5: If C is a compression of M ∈ MN then
W (C) ⊆ W (M).
Proof: Regard z ∈ W (C) as a rank -- 1 compression C ′ of C, hence of M and
apply Proposition 3 with k = 1, C replaced by C ′ and Q = M. QED
Whereas Proposition 4 supplies a necessary condition on the eigenvalues
c1, . . . , ck of a normal compression C of normal M, the following proposition
points out a sufficient condition that is sometimes useful. An interesting
analysis of such necessary vs sufficient conditions may be found in [QD].
If M ∈ MN is normal with eigenvalues z1, . . . , zN then
Proposition 6:
c1, . . . , ck ∈ C are eigenvalues of a normal compression C of M provided
that there exists a partition J1, . . . , Jk of {1, 2, . . . , N} such that for each
i = 1, . . . , k
ci ∈ conv{zj : j ∈ Ji}.
Proof: For each i let ci = Pj∈Ji tijzj represent ci as a convex combination.
6
Let u1, . . . , uN be an orthonormal basis of eigenvectors for M, with
For each i, let
Muj = zjuj.
wi = Xj∈Jiptijuj.
It is easy to check that w1, . . . , wk are orthonormal , that (Mwi, wi) = ci, and
that (Mwi, wh) = 0 if h 6= i. It follows that C = diag{c1, . . . , ck} represents
the compression of M to the subspace S = span{w1, . . . , wk}, ie
QED
C = PSMS.
In [CK Z1] Choi, Kribs, and Zyczkowski identified explicitly the higher -- rank
numerical ranges of Hermitian matrices, and their argument may be viewed,
along the lines of the proof of our next proposition, as an illustration of
the combined force of the necessary condition from Proposition 4 with the
sufficient condition from Proposition 6. Note that the result might also have
been obtained as a special case of the Fan -- Pall result, Theorem 1 (taking
b1 = b2 = · · · = bk).
Proposition 7: If M ∈ MN is Hermitian with (real) eigenvalues
a1 ≤ a2 ≤ · · · ≤ aN ,
then for each k ≤ N/2 we have
Λk(M) = [ak, aN −k+1].
If aN −k+1 < ak, then ΛK(M) = ∅.
Proof: If λ ∈ Λk(M) then taking c1 = ... = ck = λ in Proposition 4 we see
that
λ ∈ conv{ak, . . . , aN} = [ak, aN ].
Likewise, λ ∈ [a1, aN −k+1], so that Λk(M) ⊆ [ak, aN −k+1].
On the other hand, considering the partition of {1, . . . , N} into
J1 = {1, N}, J2 = {2, N − 1}, . . . , Jk = {k, N − k + 1}
7
we conclude from Proposition 6 that each λ ∈ [ak, aN −k+1] is in Λk(M). QED
As another example of such general arguments we treat the normal compres-
sion problem for the case k = N − 1. This result goes back to Fan-Pall [FP];
their proof is algebraic in character whereas ours is more geometric. We
restrict to the case where the matrix and its compression have no common
eigenvalues since this is where our general principles are most pertinent; Fan
and Pall also treat the general case by means of a direct sum construction.
Proposition 8: Let z1, . . . , zN and c1, . . . , cN −1 be two collections of complex
numbers having no elements in common. Then there is a normal M ∈ MN
with eigenvalues zj having a rank -- (N − 1) normal compression C with eigen-
values cj iff the zj are collinear and alternate with the cj (in some order)
along the common line.
Proof: Let us first show that if such M, C exist then the zj must be collinear.
Label the zj lying on the boundary of W (M) in counterclockwise order:
z1, . . . , zp. If the zj are not collinear there must be some zk−1, zk, zk+1 that
are not collinear, as in Figure 2. Proposition 4 requires that [zk−1, zk] meets
W (C) at some λ closest to zk; this λ is extreme in W (C) and so must be an
eigenvalue of C. Similarly we have an eigenvalue µ of C in [zk, zk+1], as in
Figure 2. Note that Proposition 4 also tells us that zk cannot be a repeated
eigenvalue of M, since it would then coincide with an eigenvalue of C.
Let u1, . . . , uN be an orthonormal set of eigenvectors of M, with Muj = zjuj,
and let orthonormal v, w be eigenvectors of C with Cv = λv and Cw = µw.
Expand v, w in terms of the uj:
then
v =
N
Xj=1
ajuj, w =
N
Xj=1
bjuj;
λ = (Cv, v) = (Mv, v) =
N
Xj=1
aj2zj,
so that aj = 0 unless zj lies on the line through zk−1, zk. Similarly bj = 0
unless zj lies on the line through zk, zk+1. Since zk is the only common point,
0 = (v, w) = akbk.
If ak = 0 we have λ = zk−1, which we have ruled out, while if bk = 0 we have
µ = zk+1, also ruled out.
8
zk
µ
zk+1
λ
zk−1
Figure 2: An example of the eigenvalue geometry ruled out in the proof of
Proposition 8.
Thus the eigenvalues all lie on a common line and by an affine map M →
αIN + βM this common line can be R, ie we are in the Hermitian case.
Proposition 1 then completes the argument, giving the interlacing property.
On the other hand, if the collinearity and interlacing conditions are met, the
same sort of affine map and Proposition 1 establish the existence of M and
C. QED
9
3: Results for k = 2 and small N
For 2×2 normal compressions diag(a, b), we can give a more detailed account
of the ab -- geometry, leading up to an understanding of the "starfish" seen in
Figure 1.
Recall that, given normal M ∈ MN and complex a, we denote by B(a) the
set of complex b such that diag(a, b) is a compression of M. Of course, in
order that B(a) should be nonempty we must have
a ∈ conv{z1, z2, . . . , zN},
where the zj are the eigenvalues of M. Note that Proposition 4 also requires
that for b ∈ B(a) we require that the line segment [a, b] intersect
for each i = 1, . . . , N.
conv{zj : j 6= i}
The simplest case to consider: N = 3 and the eigenvalues of M form a
nontrivial triangle.
Proposition 9: Suppose that the eigenvalues z1, z2, z3 of normal M ∈ M3
are not collinear. Then b ∈ B(a) iff either a is one of these eigenvalues, say
a = z1 and b ∈ [z2, z3] (the opposite side of the triangle formed by z1, z2, z3)
or a is in one of the sides, say [z2, z3], and b = z1.
Proof: Since [a, b] must meet each of the triangle's sides, the necessity of
the condition is clear. On the other hand, Proposition 6 shows that these
conditions suffice for a, b to be the eigenvalues of a normal compression. QED
Remark: Here we have a very simple case of the result of Fan and Pall [FP]
where they characterize in general the case k = N − 1.
When N = 4 we encounter more complex behaviour, such as that seen in
Figure 3, where B(a) is a curve interior to conv{z1, z2, z3, z4} (except for
endpoints).
10
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
−1
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
Figure 3: For a (red asterisk) strictly inside the upper quadrant (case (a)),
we see that B(a) is a curve in the opposite quadrant.
To analyse such behaviour, it will be convenient to assume in what follows
that the eigenvalues of M are generic in the sense that no three are collinear.
We may also assume that M = diag(z1, . . . , zN ), so that the eigenvectors of
M are the standard basis vectors ej.
Note that if b ∈ B(a) we have orthonormal u, w such that
(Mu, u) = a, (Mw, w) = b, and (Mu, w) = (Mw, u) = 0.
Thus a = PN
simplex, ie conv{e1, . . . , eN}; then u2 (where the operations are performed
componentwise) belongs to
1 uj2zj, a convex combination. Let ∆N denote the N -- dimensional
C(a) = {t ∈ ∆N : a =
N
X1
tjzj}.
11
By exchanging complex arguments between the components of u and w we
may assume that u ≥ 0; then the possible u lie in {√t : t ∈ C(a)}. The
conditions on w ∈ CN are then given by
kwk = 1, w ⊥ u, w ⊥ z ◦ u, and w ⊥ z ◦ u,
where ◦ indicates Schur (componentwise) multiplication, so that
z ◦ u = (z1u1, . . . , zN uN )′,
with ′ indicating transpose.
We may thus describe B(a) as follows.
Proposition 10: Given a ∈ W (M)(= conv{z1, . . . , zN}),
B(a) = [t∈C(a)
B(a, t),
where
B(a, t) = {
N
X1
wj2zj : kwk = 1, w ⊥ √t, z ◦ √t, z ◦ √t}.
Proof: To the discussion above we need only add the observation that
b = (Mw, w) =
N
X1
wj2zj.
QED
Clearly C(a) is a compact convex subset of ∆N . It is therefore the convex
hull of its extreme points, which are identified in the following result.
Proposition 11:The extreme points of C(a) are those t ∈ C(a) such that
at most three tk > 0.
Proof: Consider t ∈ C(a) such that tk > 0 for at least four values of k. We
show that t is not extreme. For convenience assume t1, t2, t3, t4 > 0. The
space
X = {x ∈ RN : xk = 0 for k > 4}
12
is 4 -- dimensional. Hence
Y = {x ∈ X :
4
X1
xk = 0,
4
X1
xkRe(zk) = 0,
4
X1
xkIm(zk) = 0} 6= {~0}.
Let ~0 6= y ∈ Y . Then for sufficiently small ǫ > 0 we have t ± ǫy ∈ ∆N and
Xk
tkzk = a,
(t ± ǫy)kzk = Xk
so that t ± ǫy ∈ C(a). Hence t is not extreme.
On the other hand, if at most three components, say t1, t2, t3 of t ∈ C(a)
are positive, and t is the average of t′, t′′ ∈ C(a), then t′
k = 0 for k > 3.
Because no three zj are collinear,
k, t′′
a = t1z1 + t2z2 + t3z3
is the unique representation of a as a convex combination of z1, z2, z3. Hence
t′ = t′′ = t. QED
For distinct indices i, j, l, let t(i, j, l) denote the element of C(a) (if it exists)
such that tk(i, j, l) = 0 whenever k 6= i, j, l. Note that such elements are
uniquely determined since
a = ti(i, j, l)zi + tj(i, j, l)zj + tl(i, j, l)zl
represents a uniquely as a point in the triangle conv{zi, zj, zl}; here again we
use the assumption that no three of the eigenvalues zj are collinear. Thus
C(a) = conv{t(i, j, l) : i, j, l are distinct and a ∈ conv{zi, zj, zl}}.
(3)
The complexity of B(a, t) increases with the number of nonzero tk. For
example, if only one tk > 0, then tk = 1 and a = zk. Here the simple
sufficient condition of Proposition 6 is also necessary:
B(a, t) = conv{zj : j 6= k}.
13
We see this as follows. Evidently, with u = √t = ek, u, w are orthonormal
exactly when w = Pj6=k αjej with Pj6=k αj2 = 1; then
b = (Nw, w) = Xj6=k
αj2zj ∈ conv{zj : j 6= k},
and any b ∈ conv{zj : j 6= k} can be obtained in this way.
The same sort of simplification occurs if only two or three tk > 0.
Proposition 12: (a) If t ∈ C(a) has exactly two positive components, say
t1, t2 > 0, then
(b) If t ∈ C(a) has exactly three positive components, say t1, t2, t3 > 0, then
B(a, t) = conv{zj : j > 2}.
B(a, t) = conv{zj : j > 3}.
Proof: (a) Since a ∈ conv{z1, z2}, Proposition 6 tells us that
B(a, t) ⊇ conv{zj : j > 2}.
On the other hand, with u = √t = (√t1,√t2, 0, . . . )′ we see that u, w are
orthonormal iff kwk = 1 and (w1, w2) ⊥ (√t1,√t2); similarly (Mu, w) = 0
only if (w1, w2) ⊥ (√t1z1,√t2z2). Since z1 6= z2, we have w1 = w2 = 0 so
that
b = (Mw, w) ∈ conv{zj : j > 2}.
(b) Since a ∈ conv{z1, z2, z3}, Proposition 6 tells us that
On the other hand, with u = √t we have u, w orthonormal iff kwk = 1 and
B(a, t) ⊇ conv{zj : j > 3}.
(w1, w2, w3) ⊥ (√t1,√t2,√t3)
and (Mu, w) = (Mw, u) = 0 only if
(w1, w2, w3) ⊥ (√t1Re(z1),√t2Re(z2),√t3Re(z3)), (√t1Im(z1),√t2Im(z2),√t3Im(z3)).
Since z1, z2, z3 are not collinear,
(1, 1, 1),
(Re(z1), Re(z2), Re(z3)),
(Im(z1), Im(z2), Im(z3))
14
are linearly independent. We must have w1 = w2 = w3 = 0 so that b =
(Mw, w) ∈ conv{zj : j > 3}. QED
We are now in a position to understand the features of Figure 3 and, indeed,
to analyse all the possibilities when N = 4. We treat in detail the case where
z1, z2, z3, z4 are all extreme in conv{z1, z2, z3, z4}; the case where one of the
eigenvalues lies in the interior of W (M) (eg z4 ∈ conv{z1, z2, z3}) can be
treated similarly.
Proposition 13: Let N = 4 and suppose that z1, z2, z3, z4 are all extreme
in W (M) and are numbered in counterclockwise order. The diagonals [z1, z3]
and [z2, z4] meet at q and divide W (M) into four quadrants. Consider a ∈
W (M); the possibilities for B(a) are as follows.
(a) See figure 3: a lies in the interior of one of the quadrants. For convenience,
assume that a ∈ conv{z1, z2, q}; let x = t(1, 2, 3), y = t(1, 2, 4). Then B(a)
is the curve traced out by the function b(r) defined for 0 < r < 1 by
b(r) =
4
Xk=1
(xk − yk)2
(1 − r)xk + ryk
zk.
4
Xk=1
(xk − yk)2
(1 − r)xk + ryk
.
Note that x4 = 0 and y3 = 0 so that
lim
r→0
b(r) = z4,
lim
r→1
b(r) = z3,
and we obtain a continuous curve parametrized on [0, 1] when we interpret
b(0) as z4 and b(1) as z3. Except for these endpoints, the curve lies in the
interior of the opposite quadrant conv{z3, z4, q}.
(b) If a lies in the interior of one of the sides of W (M) then B(a) is the
opposite side (eg if a is inside [z1, z2] then B(a) = [z3, z4]). If a = zk then
B(a) is the opposite triangle conv{zj : j 6= k}.
(c) See Figure 4: a lies interior to the diagonals but is not q; say a is interior
to [z1, q]. Then B(a) is the T -- shaped object [z2, z4] ∪ [q, z3].
(d) If a = q then B(a) is the union of the two diagonals.
15
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
−1
−1
z1
a
q
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
Figure 4: For a (red asterisk) strictly inside the segment [z1, q] (case (c)), we
see that B(a) is the T -- shaped object consisting of [z2, z4] ∪ [q, z3].
Proof: (a) Since a lies in the triangles conv{z1, z2, z3} and conv{z1, z2, z4}
but in no other triangle of eigenvalues, C(a) = [t(1, 2, 3), t(1, 2, 4)] = [x, y]
(recall the relation (3)). For 0 < r < 1 consider the t ∈ C(a) given by
t = (1 − r)x + ry. We shall see that B(a, t) consists of the single point b(r).
We take u = √t and note that the conditions on w are: w ⊥ u, w ⊥ u◦Re(z),
w ⊥ u◦Im(z), and kwk = 1. Thus w◦√t ⊥ ~14, Re(z), Im(z), where ~14 denotes
[1, 1, 1, 1]. Again we invoke linear independence of ~14, Re(z), Im(z): w ◦ √t
lies in the one -- dimensional space
dC 4 ⊖ span{~14, Re(z), Im(z)}.
There is a natural choice of (nonzero) vector in this space: x − y (because
(x,~14) = (y,~14) = 1, (x, Re(z)) = (y, Re(z)) = Re(a), and (x, Im(z)) =
(y, Im(z)) = Im(a)). Thus
w = α(x − y)/ ◦ √t,
16
where /◦ indicates entrywise division and α is some complex number. Re-
calling that kwk = 1, we derive our formula for (Mw, w) = b(r).
The necessary condition of Proposition 4 shows that the curve (ie B(a))
lies in both conv{z2, z3, z4} and conv{z1, z3, z4}, so that it must lie in the
(closed) opposite quadrant conv{z3, z4, q}. To see that the curve (except for
endpoints) lies in the interior of that quadrant, examine the arguments be-
low, showing that for b on the quadrant boundary (except for z3 and z4) a
matching a cannot be interior to the upper quadrant, and note that b ∈ B(a)
iff a ∈ B(b).
(b) If a is interior to one of the sides, say [z1, z2], then C(a) consists of
a single t with two positive components; apply Proposition 12(a) to see
that B(a) = B(a, t) = [z3, z4]. If a = z1, Propositions 4 and 6 imply that
B(a) = conv{z2, z3, z4}.
(c) Suppose a is interior to [z1, q]; then the relation (3) tells us that
C(a) = conv{t(1, 2, 3), t(1, 3, 4), t(1, 2, 4)}.
Let t(1, 2, 3) = x = [x1, 0, x3, 0]′; this is also t(1, 3, 4). Let t(1, 2, 4) = y =
[y1, y2, 0, y4]′, so that C(a) = {t(r) : 0 ≤ r ≤ 1}, where
t(r) = [(1 − r)x1 + ry1, ry2, (1 − r)x3, ry4]′.
For r = 0, Proposition 12(a) tells us that B(a, t(0)) = [z2, z4], while for
0 < r ≤ 1 we claim that B(a, t(r)) is a single point b(r) that moves along
[z3, q), covering it completely. Indeed, reasoning as in (a), we see that b(r) =
(Mw(r), w(r)) where w(r) is a normalized version of
Note that w2(r), w4(r) are proportional to −y2/√ry2,−y4/√ry4 respectively,
so that
(t(0) − t(1))/ ◦pt(r).
w2(r)2
w4(r)2 =
y2
y4
.
Since a = y1z1 + y2z2 + y4z4 lies on [z1, z3], we conclude that b(r) ∈ [z1, z3]
also. The necessary condition of Proposition 4 then tells us that b(r) ∈ [z3, q].
Since t2(r) and t4(r) tend to 0 as r → 0, limr→0 b(r) = q. Moreover, t(1) =
17
[y1, y2, 0, y4] so that Proposition 12(b) implies that b(1) = z3. Finally, since
b(r) is continuous over 0 < r ≤ 1, its values cover [z3, q).
(d) This case may be treated by an argument rather similar to that of (c).
QED
We now have the tools to continue the theme of Proposition 12, treating the
case when exactly four of the components of t ∈ C(a) are positive.
Proposition 14: Suppose that N > 4 and that t ∈ C(a) has exactly four
positive components; for convenience, assume that t1, t2, t3, t4 > 0 and that a
lies in the upper quadrant relative to Q = conv{z1, z2, z3, z4}, ie a is interior to
conv{z1, z2, q} (see Figure 3, with the understanding that it is now intended
to show only the relation of a to z1, z2, z3, z4, and Proposition 13). Let β be
the curve traced out by b(·) of Proposition 13(a) (and shown in Figure 3).
Then
Proof: With u = √t, we see that the conditions on w, namely
B(a, t) = conv{β, z5, z6, . . . , zN}.
w ⊥ u, u ◦ Re(z), u ◦ Im(z) and kwk = 1,
reduce to
w ⊥ u, u
where w = (w1, w2, w3, w4)′, u = (u1, u2, u3, u4)′ etc, and
◦Re(z), u
◦Im(z),
k wk2 +Xk>4
wk2 = 1.
Thus w/k wk is subject to the same conditions as w in the proof of Proposition
13(a). It follows that
(Mw, w) = k wk2b(r) +Xk>4
wk2zk
where b(r) can be any point on the curve β. QED
Proposition 14 allows us to understand, in large part, the phenomenon il-
lustrated in Figure 1. Let N = 5 and suppose that each eigenvalue zk is
an extreme point of W (M) = conv{z1, . . . , z5} (eg whenever M is unitary).
18
For convenience, label the zk in counterclockwise order. Suppose that a lies
strictly inside the central pentagon (which is known to be Λ2(M) in this
case). For each k let βk denote the curve obtained as in Proposition 14 by
regarding a as an element of the quadrilateral Qk = conv{zj : j 6= k}. Note
that βk connects zk+2 and zk+3 (numbering modulo 5) and lies in the quad-
rant of Qk opposite to the one containing a. We claim that (as illustrated in
Figure 1) B(a) includes the whole "starfish" region bounded by β1, β2, . . . , β5.
To see this note that the starfish is the union of the wedges Wk = conv{βk, zk},
so it suffices to show that each Wk ⊆ B(a). Since a ∈ Qk there is t ∈ C(a)
such that tk = 0. Then Proposition 14 tells us that B(a, c) = Wk.
Figure 1 was obtained by first computing C(a) via the relation (3) as
conv{t(k, k + 2, k + 3) : k = 1, 2, . . . , 5}
(note that for a in the inner pentagon, the only eigenvalue triangles containing
a correspond to the triples zk, zk+2, zk+3). To generate each of the thousands
of b's in B(a), plotted as green points in Figure 1, our MATLAB program
first chose a "random" point t ∈ C(a) (ie a random convex combination of
the five c(k, k + 2, k + 3)), put u = √t, then computed b = (Nw, w) where w
was chosen "randomly" in
C5 ⊖ span{u, u ◦ Re(z), u ◦ Im(z)}
(and normalized so that kwk = 1). The curves βk were added using the
formula of Proposition 13(a). Such simulations strongly suggest the following
"starfish conjecture", since no green dots fall outside the starfish: in such a
situation (and in particular when N = 5 and M is unitary), B(a) not only
contains the starfish but is equal to it.
We have seen in the discussion of Figure 1 that for N = 5 and a, b ∈ Λ2(M) we
always have a, b as eigenvalues of a normal compression of M. The following
proposition points out that this is true for any N -- and that N = 5 is, in
fact, the only subtle case.
Proposition 15: Let M be normal in MN and such that the eigenvalues
z1, . . . , zN are distinct and each is an extreme point of W (M) (eg M unitary).
Then a, b ∈ Λ2(M) implies that (cid:20)a 0
0 b(cid:21) is a compression of M.
19
Proof: For N ≤ 3, Λ2(M) = ∅. For even N ≥ 4, the relation (2) tells us that
Λ2(M) is the "inner N -- gon" cut off by the line segments [zj, zj+2] (indexing
modulo N). Thus for even N ≥ 4
Λ2(M) = conv{zj : j odd} ∩ conv{zj : j even},
and Proposition 6 suffices. For N = 5 the "starfish" discussion proves
our assertion. For odd N ≥ 7 we see that conv{zj
: j odd} ⊇ Λ2(M)
and conv{zj : j even} covers all of Λ2(M) except that part lying in Q =
conv{z1, z2, zN −1, zN}. Hence Proposition 6 suffices for a 6∈ Q, b ∈ Λ2(M).
The same argument applies for a 6∈ Q = conv{z2.z3, z4, z5} and because
N > 5 this covers any a ∈ Q. QED
4. Continuity of B(·)
A natural assertion of "continuity" for B(·) might be that dH(B(a′), B(a)) →
0 as a′ → a, where dH(X, Y ) is the Hausdorff distance between compact
nonempty sets X, Y ⊂ C. Recall that
where
dH(X, Y ) = max{ dH(X, Y ), dH(Y, X)},
dH(X, Y ) = max
x∈X
y∈Y x − y).
(min
However, we have seen simple examples where this fails: recall the analysis
of B(a) for various a ∈ conv{z1, z2, z3, z4} that was provided by Proposition
13. If a′ lies in the interior of [z1, z2] and a′ → a = z1, then B(a′) = [z3, z4]
"jumps" to B(a) = conv{z2, z3, z4}. A perhaps more surprising example: let
a be interior to [z1, q] as in Figure 4; for a′ approaching a from the interior
of conv{z1, z2, q} we see B(a′) as a curve joining z3 and z4 in conv{z3, z4, q},
whereas for a′ approaching a from the interior of conv{z1, z4, q} we see B(a′)
as a curve joining z2 and z3 in conv{z2, z3, q}.
In spite of such "failures" we'll show that B(·) is continuous with respect
to Hausdorff distance at most points of W (M) and enjoys a "one -- sided"
Hausdorff continuity in general.
Our standard set -- up for this discussion is as in section 3, ie we assume M
is normal in MN and is in diagonal form: M = diag(z), where no three
20
eigenvalues are collinear. Thus W (M) = conv{z1, . . . , zN} and B(a′) = ∅ if
a′ 6∈ W (M). Seeking continuity, we restrict attention to a′ → a with a′, a ∈
W (M). Note that if N = 3 and a′ is interior to W (M) = conv{z1, z2, z3},
we again have B(a′) = ∅, since b ∈ B(a′) and Proposition 4 would require
that [a′, b] meet each side of the triangle W (M). We therefore restrict also
to cases where N ≥ 4.
Proposition 16: If N ≥ 4, B(a) is a compact nonempty set for any a ∈
W (M).
Proof: Let t ∈ C(a). Since N ≥ 4,
CN ⊖ span{√t,√t ◦ Re(z),√t ◦ Im(z)}
is nontrivial (6= {~0}). Let w be a unit vector in this space; then b =
(Mw, w) ∈ B(a, t), so B(a) 6= ∅.
For compactness, consider bn ∈ B(a); there exist orthonormal pairs un, wn
such that
(Mun, un) = a,
(Mwn, wn) = bn,
(Mun, wn) = (Mwn, un) = 0.
Since the sequences un, wn are bounded, local compactness in C2N implies
that, for some subsequence nk,
Then u, w are orthonormal and
unk →k u, wnk →k w.
(Mu, u) = a,
(Mw, w) = lim
k
bnk = b,
(Mu, w) = (Mw, u) = 0.
The limit point b is in B(a). QED
A related argument shows that, in general, B(·) is continuous in a one -- sided
Hausdorff sense.
Proposition 17: If a, an ∈ W (M) and an → a, then
dH(B(an), B(a)) →n 0.
(4)
Proof: Recall that dH(X, Y ) = maxx∈X(miny∈Y x − y). Thus, if (4) were
to fail we'd have some ǫ > 0, subsequence nk, and bk ∈ B(ank) such that for
all b ∈ B(a)
bk − b ≥ ǫ.
21
By restricting to such a subsequence we may assume that bn ∈ B(an). Let
un, wn be orthonormal pairs such that
(Mun, un) = an,
(Mwn, wn) = bn,
(Mun, wn) = (Mwn, un) = 0.
There is a subsequence nk such that
Hence u, w are orthonormal and
unk →k u, wnk →k w.
(Mu, u) = lim
k
ank = a,
(Mw, w) = lim
k
bnk = b,
(Mu, w) = (Mw, u) = 0.
It follows that b = limk bnk ∈ B(a), contradicting bnk − b ≥ ǫ. QED
In terms of the obvious extension of Hausdorff distance to compact nonempty
subsets of ∆N , we note that C(·) is continuous and in fact satisfies a Lipschitz
condition for each fixed M.
Proposition 18: There is a constant K < ∞ depending only on M such
that for all a, a′ ∈ W (M)
dH(C(a), C(a′)) ≤ Ka − a′.
Proof: For each triple i, j, k of distinct indices, we have assumed that
zi, zj, zk are not collinear. Thus the matrix
T =
1
1
1
Re(zi) Re(zj) Re(zk)
Im(zi)
Im(zk)
Im(zj)
is nonsingular. Given a ∈ conv{zi, zj, zk}, consider tijk = t(i, j, k) as in (3).
Let tijk be the vector in R3 recording the i, j, k -- components of tijk, ie the
only components that may be positive. We have T tijk = (1, Re(a), Im(a))′ so
that tijk = T −1(1, Re(a), Im(a))′. In terms of the operator norm kT −1k we
have
ktijk − t′
ijkk ≤ kT −1ka − a′
for any other a′ ∈ conv{zi, zj, zk}. Let K be the maximum of kT −1k over all
such triangles conv{zi, zj, zk}.
The line segments [zi, zj] form a "grid" criss -- crossing W (M), dividing it into
22
regions. Suppose a, a′ lie in the same one of these regions (boundary points
allowed). Then the set Q of triples i, j, k such that a ∈ conv{zi, zj, zk} is
the same as that for a′. In view of (3), each t ∈ C(a) can be expressed as a
convex combination
t = Xijk∈Q
sijktijk.
Putting
t′ = Xijk∈Q
sijkt′
ijk,
we have t′ ∈ C(a′) and kt−t′k ≤ Ka−a′. The roles of a, a′ may be reversed,
so we see that if a, a′ are in the same region (boundary points allowed),
dH(C(a), C(a′)) ≤ Ka − a′.
Finally, for any a, a′ ∈ W (M), the line segment [a, a′] intersects the grid in
a sequence of points a0, a1, . . . , an ordered along [a, a′] with a0 = a, an = a′.
By the argument above,
dH(C(ak), C(ak+1)) ≤ Kak − ak+1,
so that (dH is a metric)
dH(C(a), C(a′)) ≤ K
n−1
Xk=0
ak − ak+1 = Ka − a′.
QED
Next we show that B(·) is dH -- continuous at any point that is "off the grid",
and that continuity is uniform if we stay bounded away from the grid.
Proposition 19: If a ∈ W (M) but a does not lie on any line segment [zi, zj],
then a′ → a implies that
dH(B(a′), B(a)) → 0.
In fact, on any subset S(d) ⊂ W (M) that is a positive distance d from the
grid
so that
G = [{[zi, zj] : i, j = 1, . . . , N},
S(d) = {a ∈ W (M) : min
g∈G a − g ≥ d},
23
the map a 7→ B(a) is uniformly continuous.
Proof: In this discussion i, j, k always denotes a triple of distinct indices.
Let
Q = [{C(a) : a ∈ S(d)};
we claim that
min
t∈Q
(max
i,j,k
titjtk)
is positive. Otherwise, by compactness, we'd have some a ∈ S(d) and
t ∈ C(a) such that maxi,j,k titjtk = 0. This can only happen if t has at
most two positive components, say ti, tj; then a ∈ [zi, zj], which we have
ruled out.
Given linearly independent q, r, s ∈ CN , let P (q, r, s) denote orthogonal pro-
jection onto
CN ⊖ span{q, r, s}.
The map (q, r, s) 7→ P (q, r, s) is uniformly continuous if we "stay away from
dependence"; to be precise, for any 0 < h < H < ∞ this map is uniformly
continuous on
Q(h, H) = {(q, r, s) : kqk,krk,ksk ≤ H, max
i,j,k det
qj
rj
qi
qk
ri
rk
si sj sk
≥ h}.
Now the values (√t,√t ◦ Re(z),√t ◦ Im(z)) where t ∈ Q lie in some fixed
Q(h, H) because each
det
1
1
1
Re(zi) Re(zj) Re(zk)
Im(zk)
Im(zi)
Im(zj)
is nonzero, so that
max
i,j,k ptitjtk det
1
1
1
Re(zi) Re(zj) Re(zk)
Im(zi)
Im(zk)
Im(zj)
≥ h
for some positive h. Thus the map t 7→ P (√t,√t ◦ Re(z),√t ◦ Im(z)) = P [t]
is uniformly continuous on Q: given ǫ1 > 0 there is δ1 > 0 such that t, t′ ∈ Q
24
and kt − t′k ≤ δ1 implies kP [t] − P [t′]k ≤ ǫ1.
In view of Proposition 18, there is δ > 0 such that a − a′ ≤ δ implies
dH(C(a), C(a′)) ≤ δ1. Consider b ∈ B(a); for some t ∈ C(a) we have b ∈
B(a, t) so that b = (Mw, w) for some unit w with P [t]w = w. Let t′ ∈ C(a′)
be such that kt − t′k ≤ δ1; then kw − P [t′]wk ≤ ǫ1. Note that
1 − ǫ1 ≤ kP [t′]wk ≤ 1,
and let w′ = P [t′]w/kP [t′]wk; then b′ = (Mw′, w′) ∈ B(a′) and
b − b′ = (Mw, w) − (Mw′, w′) ≤ 2kMkkw − w′k.
It is easy to see that kw− w′k ≤ 2ǫ1/(1− ǫ1), so that given any ǫ > 0 we have
b − b′ ≤ ǫ by an appropriate choice of ǫ1. We have shown that a − a′ ≤ δ
implies that dH(B(a), B(a′)) ≤ ǫ. Since the roles of a, a′ may be reversed,
we also have dH(B(a), B(a′)) ≤ ǫ. QED
Note that sometimes B(·) is continuous even at points that are on the grid.
For example, from Proposition 13(a) and 13(b) we can see that there is conti-
nuity everywhere on the boundary segments [zi, zi+1] except at the endpoints.
5. Related results
We offer some remarks on the apparently more difficult problem of charac-
terizing arbitrary compressions of a normal matrix M. Suppose again that
M is N × N, and is represented by the diagonal matrix diag(z) and that X
is a rank -- k compression of M, ie there is a k -- dimensional subspace S such
that X = PSMS. From Proposition 3 we obtain a necessary condition on
X: the (classical) numerical range W (X) of X must intersect the convex hull
of any subset of the eigenvalues zj having size N − k + 1.
When k = 2, ie X is represented by a 2 × 2 matrix, the numerical range
W (X) determines X uniquely as an operator. Indeed, W (X) is a (filled -- in)
ellipse in this case with the eigenvalues of X as foci and the length of the
minor axis is the modulus of the off -- diagonal entry of any upper -- triangular
matrix for X. Let's consider the problem of characterizing such compres-
sions X geometrically via the elliptical W (X) in the cases where N = 3 and
N = 4.
25
When N = 3, the necessary condition of above tells us that W (X) must be
tangent to each of the three sides of conv{z1, z2, z3} (recall that Proposition
5 tells us that in general we must have W (X) ⊆ W (M) = conv{zj : j =
1, . . . , n}). In fact, Williams showed long ago that the necessary condition is
also sufficient when N = 3 (see [Wi]).
When N = 4 we consider the case where the eigenvalues zj form a quadri-
lateral Q. The necessary condition above tells us that W (X) must intersect
each of the four triangles Ti = conv{zj : j 6= i}. Thus W (X) must intersect
each of the quadrants Ti ∩ Tk. This phenomenon is borne out by numerical
experiments such as Figure 5 illustrates, but it is not clear what additional
conditions must be satisfied by W (X), even in this N = 4 case. Of course,
if by chance W (X) is tangent to all three sides of some Ti, then Williams'
result tells us that X is indeed a 2 -- dimensional compression.
26
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
−1
−1
−0.8
−0.6
−0.4
−0.2
0
0.2
0.4
0.6
0.8
1
Figure 5: Shows the (elliptical) boundaries of the numerical ranges of several
(nonnormal) compressions of a 4 × 4 normal M, each compression having
a (red asterisk) as an eigenvalue (therefore seen as one of the foci of each
ellipse)
27
References:
[Cau] A. L. Cauchy, Sur l'´equation `a l'aide de laquelle on d´etermine les
in´egalit´es s´eculaires des mouvements des plan`etes, Oeuvres compl`etes, Sec-
ond Ser., IX, 174 -- 195
[CK Z1] M. -- D. Choi, D. W. Kribs, and K. Zyczkowski, Higher -- rank numer-
ical ranges and compression problems, Linear Algebra Appl. 418, 828 -- 839,
2006
[CK Z2] M. -- D. Choi, D. W. Kribs, and K. Zyczkowski, Quantum error cor-
recting codes from the compression formalism, Rep. Math. Phys. 58, 77 -- 91,
2006
[CHK Z] 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, 409 -- 426, 2007
[CGHK] M. -- D. Choi, M. Giesinger, J. A. Holbrook, and D. W. Kribs, Ge-
ometry of higher -- rank numerical ranges, Linear and Multilinear Algebra 56,
53 -- 64, 2008
[DGHP Z] C. F. Dunkl, P. Gawron, J. Holbrook, Z. Puchala, and K. Zyczkowski,
Numerical shadows: measures and densities on the numerical range, Linear
Algebra Appl. 434, 2042 -- 2080, 2011
[FP] K. Fan and G. Pall, Imbedding conditions for Hermitian and normal
matrices, Canadian J. Math. 9, 298 -- 304, 1957
[GPMS Z] P. Gawron, Z. Puchala, J. Miszczak, L. Skowronek, and K. Zyczkowski,
Restricted numerical range: a versatile tool in the theory of quantum infor-
mation, J. Math. Physics 51, 2010
[KPLRdS] D. W. Kribs, A. Pasieka, M. Laforest, C. Ryan, and M. P. da
Silva, Research problems on numerical ranges in quantum computing, Lin-
ear and Multilinear Algebra 57, 491-502, 2009
[LP] C. -- K. Li and Y. -- T. Poon, Generalized numerical ranges and quantum
28
error correction, J. Operator Theory 66, 335 -- 351, 2011
[LPS1] C. -- K. Li, Y. -- T. Poon, and N. -- S. Sze, Higher rank numerical ranges
and low rank perturbations of quantum channels, J. Math. Analysis Appl.
348, 843 -- 855, 2008
[LPS2] 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 57, 365 --
368, 2009
[LS] C. -- K. Li and N. -- S. Sze, Canonical forms, higher -- rank numerical ranges,
totally isotropic subspaces, and matrix equations, Proc. Amer. Math. Soc.
136, 3013 -- 3023, 2008
[MM Z] K. Majgier, H. Maassen, and K. Zyczkowski, Protected subspaces in
quantum information, Quantum Information Processing 9, 343 -- 367, 2010
[M] N. Mudalige, Higher Rank Numerical Ranges of Normal Operators, MSc
thesis, U of Guelph, 2010
[QD] J. F. Queir´o and A. L. Duarte, Imbedding conditions for normal ma-
trices, Linear Algebra Appl. 430, 1806 -- 1811, 2009
[Wi] J. P. Williams, On compressions of matrices, J. London Math. Soc. (2)
3, 526 -- 530, 1971
[Wo] H. Woerdeman, The higher -- rank numerical range is convex, Linear and
Multilinear Algebra 56, 65 -- 67, 2008
29
Author addresses:
John Holbrook
Dept of Mathematics and Statistics
University of Guelph
Guelph, Ontario, Canada N1G 2W1
[email protected]
Nishan Mudalige
Dept of Mathematics and Statistics
York University
Toronto, Ontario, Canada
[email protected]
Rajesh Pereira
Dept of Mathematics and Statistics
University of Guelph
Guelph, Ontario, Canada N1G 2W1
[email protected]
30
|
1810.05641 | 1 | 1810 | 2018-10-11T20:24:44 | On directional derivatives of trace functionals of the form $A\mapsto\Tr(Pf(A))$ | [
"math.FA"
] | Given a function $f:(0,\infty)\rightarrow\RR$ and a positive semidefinite $n\times n$ matrix $P$, one may define a trace functional on positive definite $n\times n$ matrices as $A\mapsto \Tr(Pf(A))$. For differentiable functions $f$, the function $A\mapsto \Tr(Pf(A))$ is differentiable at all positive definite matrices $A$. Under certain continuity conditions on~$f$, this function may be extended to certain non-positive-definite matrices $A$, and the \emph{directional} derivatives of $\Tr(Pf(A)$ may be computed there. This note presents conditions for these directional derivatives to exist and computes them. These conditions hold for the function $f(x)=\log(x)$ and for the functions $f_p(x)=x^p$ for all $p>-1$. The derivatives of the corresponding trace functionals are computed here. | math.FA | math |
On directional derivatives of trace functionals of the form
A 7→ Tr(P f (A))
Mark W. Girard1
1Institute for Quantum Computing at the University of Waterloo, Waterloo, ON, Canada
October 16, 2018
Abstract
Given a function f : (0, ∞) → R and a positive semidefinite n × n matrix P, one may define a trace func-
tional on positive definite n × n matrices as A 7→ Tr(P f (A)). For differentiable functions f , the function
A 7→ Tr(P f (A)) is differentiable at all positive definite matrices A. Under certain continuity conditions
on f , this function may be extended to certain non-positive-definite matrices A, and the directional deriva-
tives of Tr(P f (A) may be computed there. This note presents conditions for these directional derivatives
to exist and computes them. These conditions hold for the function f (x) = log(x) and for the functions
f p(x) = xp for all p > −1. The derivatives of the corresponding trace functionals are computed here.
1 Introduction
Let Hn denote the set of n × n Hermitian matrices over C, let Pn denote the subset of positive semidefinite
n × n matrices, and let P+
n denote the positive definite ones. Any function of real numbers f : (0, ∞) → R
can be extended to positive definite matrices by means of the spectral decomposition. Given a positive
matrix A ∈ P+
n with spectral decomposition
A =
n
∑
i=1
αi viv∗
i ,
where α1, . . . , αn ∈ (0, ∞) are the eigenvalues of A and v1, . . . , vn ∈ Cn are the corresponding normalized
eigenvectors, one defines f (A) as
f (A) =
n
∑
i=1
f (αi) viv∗
i ,
where v∗ denotes the conjugate transpose of a vector v ∈ Cn. For a positive semidefinite n × n matrix
P ∈ Pn and a function f : (0, ∞) → R, one may define a function fP : P+
n → R defined as
fP(A) = Tr(P f (A))
(1)
for all positive definite matrices A ∈ P+
n . Functions of this type arise frequently, for example, in the study
of quantum information theory [GGF14]. In this note, we investigate continuity and differentiability prop-
erties of functionals of the form in (1).
If the function f : (0, ∞) → R can be continuously extended to be defined at 0, the function fP can be
continuously extended to be defined at all non-positive-definite n × n matrices in the natural way. However,
if the limit limt→0+ f (t) does not exist, it is still possible to define fP(A) for certain non-positive-definite
n × n matrices A by restricting f to the subspace spanned by the eigenvectors corresponding to nonzero
1
eigenvalues of A. For example, the quantum relative entropy of two positive n × n matrices P, Q ∈ Pn is
defined by [Wat18]
S(PkQ) =(cid:26) Tr(P log P) − Tr(P log Q) whenever im(P) ⊆ im(Q)
otherwise,
+∞
where im(P) denotes the image of P and Tr(P log Q) has a natural interpretation whenever im(P) ⊆ im(Q).
Indeed, for any continuous function f : (0, ∞) → R and any matrix P ∈ Pn, it is natural to define fP(A) for
non-positive-definite matrices A ∈ Pn (with im(A) ⊆ im(P)) as
fP(A) =
r
∑
i=1
f (αi) Tr(Pviv∗
i ),
(2)
where we assume that A ∈ Pn has rank r with nonzero eigenvalues α1, . . . , αr > 0 and αr+1 = · · · = αn = 0
(see, e.g., equation (2.2) in [Ras11]).
1.1 Directional derivatives of matrix trace functions
Let g : A → R be a real-valued function on some subset A ⊂ Hn. For any A ∈ A and any matrix B ∈ Hn,
the (one-sided) directional derivative of g at A in the direction B is defined as
dg(A; B) = lim
t→0+
g(A + tB) − g(A)
t
.
Here we are interested in computing the directional derivatives of functions of the form fP(A) = Tr(P f (A)).
If f is differentiable and A is positive definite, then these directional derivatives certainly exist, since the
function f when extended to positive definite matrices is differentiable as a function of matrices. However,
if A is not necessarily positive definite, these directional derivatives may still be computed. Knowing the
derivatives is important for determining optimality conditions for certain types of optimization problems
that arise in quantum information [GGF14]. The directional derivatives are presented in Theorem 1 and
make use of the following notation.
Let f : (0, ∞) → R be a differentiable function. The first order divided differences of f defined as
f [1](x, y) =
f ′(x)
f (x) − f (y)
x − y
if x = y
if x 6= y
for all x, y ∈ (0, ∞). For any positive sedefinite n × n matrix A ∈ Pn, we may define a linear mapping
Φ f ,A : Hn → Hn of n × n matrices as follows. If A = diag(α1, . . . , αn) is diagonal, we can write A as
A =
n
∑
i=1
αi eie∗
i ,
where e1, . . . , en ∈ Cn are the standard orthonormal basis vectors of Cn such that the entries of any other
matrix B ∈ Hn are given by Bij = hei, Beji. The matrix of divided differences of A (restricted to the nonzero
eigenvalues of A) as the matrix D f ,A ∈ Hn whose entries are given by
(D f ,A)ij =(cid:26) f [1](αi, αj)
0
if αi > 0 and αj > 0
if αi = 0 or αj = 0,
and for all B ∈ Hn define Φ f ,A(B) as
Φ f ,A(B) = D f ,A ⊙ B
2
(3)
(4)
where X ⊙ Y denotes the entrywise product of matrices X, Y ∈ Hn with matrix elements (X ⊙ Y)ij = XijYij
for all i, j ∈ {1, . . . , n}. If A is not diagonal, there exists an n × n unitary matrix U such that U AU∗ is
diagonal, and one defines
(5)
for all B ∈ Hn, and this is independent of the choice of diagonalizing unitary U. We may now state the
main theorem of this work.
Φ f ,A(B) = U∗(cid:0)D f ,UAU∗ ⊙ (UBU∗)(cid:1)
Theorem 1. Let f : (0, ∞) → R be a differentiable function satisfying limt→0+ t f (t) = 0, let P ∈ Pn be a positive
semidefinite n × n matrix, and consider the function fP : P+
n → R as defined above. Let A ∈ Pn be a positive matrix
satisfying im(P) ⊆ im(A) such that we may define fP(A) as in (2), and let B ∈ Hn. Suppose there exists ε > 0
such that A + tB ∈ Pn holds for all t ∈ [0, ε). The directional derivative of f at A in the direction B exists and can be
computed by
d fP(A; B) = lim
t→0+
fP(A + tX) − fP(A)
t
= Tr(P Φ f ,A(B)),
(6)
where Φ f ,A : Hn → Hn is the linear mapping defined above in (5).
In the case when A is positive definite, we remark that the directional derivative in (6) coincides with
well known results in [Bha97, Theorem V.3.3] and [HP14, Theorem 3.25]. For non-positive-definite A ∈ Pn,
these derivatives were provided in [GGF14], but no proof of the existence of the directional derivatives
were provided there. We note that the function f in Theorem 1 must satisfy the condition that
t f (t) = 0
lim
t→0+
(7)
in order for the derivatives to be computed in this manner.
The remainder of this note is dedicated to the proof of Theorem 1 (which will be proved using matrix
perturbation methods) and to provide some examples.
In particular, we note that the condition in the
theorem is met for the function f (t) = log(t) and the functions f (t) = tp for all real values p > −1, as these
functions satisfy (7). The derivatives of the function A 7→ Tr(P log(A)) at non-positive-definite matrices
A were studied in [FG11]. For the function f (t) = t−1, the directional derivatives are no longer able to be
computed in this manner, as this function does not satisfy the condition in (7), however the expression in
(6) still provides a lower bound for the directional derivative.
The remainder of the note is organized as follows. Section 2 introduces the notation that will be used
in this note, recalls some basic notions of differentiation of matrix functions, and presents some facts from
perturbation theory for Hermitian matrices. The proof of Theorem 1 is presented in Section 3. In Section 4,
we consider the functions f (t) = log(t) and f (t) = tp for p ∈ (−1, 1) as examples, and provide alternate
proofs of the directional derivatives for these functions by the method of integral representations rather
than matrix perturbation methods. Finally, in Section 5, we show that this method finds a lower bound to
the directional derivatives for the choice of function f (t) = t−1.
2 Background
Notions of differentiability of matrix functions are recalled in Section 2.1. Some results on spectral pertur-
bation theory are reviewed in 2.2.
2.1 Derivatives of matrix functions
We refer to [Bha97] for more details. Let A ⊆ Hn be a subset of the n × n Hermitian matrices and let
f : A → Hn be a function of matrices. The function f is said to be (Fr´echet) differentiable at a matrix A ∈ A if
3
there exists a linear mapping Φ : Hn → Hn of matrices satisfying
lim
H→0
k f (A + tH) − f (A) − Φ(H)k
kHk
= 0,
where k·k denotes the spectral norm on the space of matrices n × n. If such a mapping exists, it is called the
(Fr´echet) derivative of f at A and is denoted by Φ = D f (A). In cases where the function is not differentiable
at a point, it may still possess directional derivatives.
Let f : (0, → R) be a differentiable functions. If A ∈ P+
tiable (as a function of matrices) at A with Fr´echet derivative
n is a positive definite matrix, then f is differen-
D f (A)(B) = Φ f ,A(B),
(8)
where Φ f ,A : Hn → Hn is the linear mapping defined in (5). Moreover, for any positive semidefinite matrix
P ∈ Pn, the directional derivatives of the function fP : P+
n → R (as defined in (1)) at any positive definite
matrix A ∈ P+
n are given by
d fP(A; B) = Tr(P D f (A)(B))
for all B ∈ Hn.
2.2 Spectral perturbation theory for Hermitian matrices
Consider now families of Hermitian matrices of the form A + tB for some choice of Hermitian matrices
A, B ∈ Hn and variable t ∈ R. It is a remarkable fact from perturbation theory of linear opeators (see, e.g.,
[Kat80, II.6.2]) that there exists a spectral decomposition of A + tB that behaves analytically in the variable t.
That is, there exist analytic functions λ1, . . . , λn : R → R for the eigenvalues of A + tB and analytic vector-
valued functions u1, . . . , un : R → Cn such that A + tB may be expressed as
A + tB =
n
∑
i=1
λi(t) ui(t)ui(t)∗,
(9)
for all t ∈ R. As this is a spectral decomposition of A + tB, one has that (A + tB)ui(t) = λi(t)ui(t) and that
hui(t), uj(t)i =(cid:26) 1
0
if i = j
if i 6= j
holds for all t ∈ R. Suppose that the eigenvalues α1, . . . , αn and the eigenvectors v1, . . . , vn of A are such
that αi = λi(0) and vi = ui(0) for each i ∈ {1, . . . , n}. The first-order derivatives λ′
i(0) can be
computed from B and the spectral decomposition of A, as the following propostion shows.
i(0) and u′
Proposition 2. Suppose A ∈ Hn and B ∈ Hn are Hermitian matrices and let λ1(t), . . . , λn(t) and u1(t), . . . , un(t)
denote the eigenvalues and coresponding eigenvectors (which are analytic as functions of t) of the matrix A + tB
comprising the spectral decomposition in (9). The following statements hold.
(i) For all i ∈ {1, . . . , n}, it holds that λ′
i(0) = hvi, Bvii.
(ii) For all i and j with i 6= j, it holds that(cid:0)αi − αj(cid:1)hvi, u′
(iii) For all i and j, it holds that hu′
j(0), vji + hvj, u′
j(0)i = hvi, Bvji.
j(0)i = 0.
Here, α1, . . . , αn are the eigenvalues and v1, . . . , vn are the eigenvectors of A such that αi = λi(0) and vi = ui(0) for
each i ∈ {1, . . . , n}
4
Proof. For each index j, note that the expression (A + tB)uj(t) − λj(t)uj(t) = 0 is constant with respect to t.
Differentiating this expression at t = 0 yields
= Bvj + Au′
j(0) − λ′
j(0)vj − αju′
j(0).
Taking the inner product of this expression with vi, one finds that
j(0)vj − αju′
j(0) − λ′
0 =
d
dt(cid:0)(A + tB)uj(t) − λj(t)uj(t)(cid:1)(cid:12)(cid:12)(cid:12)t=0
0 =(cid:10)vi, (cid:0)Bvj + Au′
= hvi, Bvji +(cid:0)αi − αj(cid:1)hvi, u′
= hvi, Bvji + αihvi, u′
j(0)(cid:1)(cid:11)
j(0)i − αihvi, u′
j(0)i − λ′
j(0)i − λ′
i(0)hvi, vji.
i(0)hvi, vji
Taking i = j yields property (i) while taking i 6= j yields property (ii). To prove (iii), note that hui(t), uj(t)i
is constant for all i and j. Taking the derivative yields
0 =
d
dt
hui(t), uj(t)i(cid:12)(cid:12)(cid:12)t=0
as desired.
3 Proof of Theorem 1
= hu′
i(0), uj(0)i + hui(0), u′
j(0)i = hu′
i(0), vji + hvi, u′
j(0)i,
Proof (of Theorem 1). Let λ1, . . . , λn : R → R and u1, . . . , un : R → Cn be the analytic functions denoting the
eigenvalues and corresponding orthonormal eigenvectors of A + tB, and let α1, . . . , αn be the eigenvalues
and v1, . . . , vn the corresponding orthonormal eigenvectors of A comprising the spectral decomposition
A =
n
∑
i=1
αi viv∗
i ,
such that λi(0) = αi and ui(0) = vi for all i ∈ {1, . . . , n}. We may assume that α1, . . . , αr > 0 are the nonzero
eigenvalues of A and that αr+1 = · · · = αn = 0. Define the value of fP(A) as
fP(A) =
r
∑
i=1
f (αi)hvi, Pvii,
where one sums only over the nonzero eigenvalues of A. Note from Proposition 2 that hvi, Bvii = λ′
i(0),
and moreover that λ′
i(0) ≥ 0 must hold by assumption for all i ∈ {r + 1, . . . , n} since A + tB is assumed
to be positive semidefinite for all t ∈ [0, ε). Furthermore, it may assumed without loss of generality that
hvi, Bvii > 0 for all i ∈ {r + 1, . . . , n}. Indeed, if it holds that hvi, Bvii = 0 for some i ∈ {r + 1, . . . , n} then
hvi, (A + tB)vii = 0 holds for all t, and one may restrict to the problem to the subspace perpendicular to vi.
One therefore has that
d fP(A; B) = lim
t→0+
fP(A + tB) − fP(A)
t
= lim
t→0+
= lim
t→0+
Tr(P f (A + tB)) − fP(A)
t
r
∑
i=1
f (λi(t))hui(t), Pui(t)i − f (αi)hvi, Pvii
t
+ lim
t→0+
n
∑
i=r+1
f (λi(t))hui(t), Pui(t)i
t
=
r
∑
i=1
f ′(αi)λ′
i(0)hui(0), Pui(0)i +
r
∑
i=1
f (αi)(cid:0)hu′
i(0), Pui(0)i + hui(0), Pu′
i(0)i(cid:1)
t
+
n
∑
i=r+1
lim
t→0+
f (λi(t))hi(t)
(10)
,
5
where we define the functions hi(t) = hui(t), Pui(t)i for each i ∈ {r + 1, . . . , n}. Note that each hi is analytic
with hi(0) = h′
i(0) = 0, since Pvi = 0 holds for all i ∈ {r + 1, . . . , n} by the assumption that im(P) ⊂ im(A).
The second sum in (10) reduces to
r
∑
i=1
f (αi)(cid:16)hu′
i(0), Pui(0)i + hui(0), Pu′
i(0)i(cid:17)
i(0), vjihvj, Pvii + hvi, Pvjihvj, u′
=
=
=
r
∑
i,j=1
r
∑
i,j=1
r
∑
i,j=1
αi=αj
f (αi)(cid:16)hu′
hvj, Pvii(cid:16) f (αi)hu′
f (αi)hvj, Pvii(cid:16)hu′
i(0)i(cid:17)
j(0)i(cid:17)
+
r
∑
i,j=1
αi6=αj
j(0)i(cid:17)
}
i(0), vji + f (αj)hvi, u′
i(0), vji + hvi, u′
=0
{z
hvj, Pvii(cid:16) f (αi)hu′
i(0), vji + f (αj)hvi, u′
j(0)i(cid:17)
where the term hu′
Thus the second sum in (10) further reduces to
i(0), vji + hvi, u′
j(0)i in the last line above vanishes by statement (iii) in Proposition 2.
hvj, Pvii f (αi)
hvi, Bvji
αj − αi
+ f (αj)
hvi, Bvji
αi − αj !
r
∑
i,j=1
αi6=αj
hvj, Pviihvi, Bvji
f (αi) − f (αj)
αi − αj
hvj, Pviihvi, Bvji f [1](αi, αj).
=
=
r
∑
i,j=1
αi6=αj
r
∑
i,j=1
αi6=αj
Noting from statement (ii) of Proposition 2 that hvi, Bvji = 0 for all pairs of indices i 6= j with αi = αj, the
first two sums in (10) reduce to
r
∑
i=1
f ′(αi)λ′
i(0)hui(0), Pui(0)i +
r
∑
i=1
f (αi)(cid:0)hu′
i(0), Pui(0)i + hui(0), Pu′
i(0)i(cid:1)
hvj, Pviihvi, Bvii f ′(αi) +
hvj, Pviihvi, Bvji f [1](αi, αj)
r
∑
i=1
=
=
n
∑
i,j=1
hvj, Pviihvi, Bvji f [1](αi, αj)
n
∑
i,j=1
αi6=αj
Finally, as λi(0) = αi = 0 and λ′
i(0) = hvi, Bvii > 0 for all i ∈ {r + 1, . . . , n}, it holds that
= Tr(P Φ f ,A(B)).
(11)
(12)
lim
t→0+
f (λi(t))hi(t)
t
= lim
t→0+
t f (λi(t))
t f (λi(t))
h′′
i (0)
2
= 0,
hi(t)
t2 = lim
t→0+
=0
{z
}
for all i ∈ {r + 1, . . . , n}, where the limit vanishes from the fact that λi(0) = 0 and λ′
i(0) > 0, and by
the assumption that limt→0+ t f (t) = 0. Plugging the results of (11) and (12) into (10) yields d fP(A; B) =
Tr(P Φ f ,A(B)), as desired.
6
4 Alternative proofs of differentiability via integral representations
In this section, alternative proofs for the computations of the directional derivatives of fP are provided in
the case when f (x) = log(x) or f (x) = x p for some value p ∈ (−1, 1) following the method in [VP98,
Thm. 3]. This method makes use of integral representations of these functions, which may be extended to
matrices in the usual way.
4.1 Directional derivatives of f p,P
Let p ∈ (−1, 0) ∪ (0, 1) and consider the function f p : (0, ∞) → R defined as f p(x) = x p for all x ∈ (0, ∞).
The divided differences of this function are given by
For a positive matrix P ∈ Pn, consider the function f p,P defined on positive matrices as
f
[1]
p (x, y) =
px p−1
x p − yp
x − y
x = y
x 6= y.
f p,P(A) =(cid:26) Tr(PAp)
+∞
im(P) ⊆ im(A) or p ∈ (0, 1)
else
for all A ∈ Pn. Note that f p,P is differentiable at all positive definite matrices A, as the function f p is
differentiable. As indicated by Theorem 1, the directional derivatives of f p,P at a positive semidefinite
matrix A can be computed as
d f p,P(A; B) = Tr(P Φ f p,A(B))
as long as im(P) ⊆ im(A), where B ∈ Hn is any Hermitian matrix such that A + tB is positive for all
t > 0 small enough. Here we show how to directly compute these directional derivatives using a method
of integral representations for f p.
The calculation is split into the cases p ∈ (−1, 0) and p ∈ (0, 1), which are considered in Sections 4.1.1
and 4.1.2 respectively. The following integral representations will be used.1 For all x ∈ (0, ∞) one has
x p =
and
x p =
− sin(pπ)
π
sin(pπ)
π
0
x + s
sp
Z ∞
sp(cid:18) 1
Z ∞
s
0
ds
for all p ∈ (−1, 0)
−
1
x + s(cid:19) ds
for all p ∈ (0, 1).
(13)
(14)
Furthermore, for all x, y ∈ (0, ∞) with x 6= y, and all p ∈ (−1, 1), one has
px p−1 =
sin(pπ)
π
x p − yp
x − y
=
sin(pπ)
π
and
0
Z ∞
Z ∞
0
sp
(x + s)2 ds
sp
(x + s)(y + s)
ds.
In particular, for all p ∈ (−1, 1), the divided differences of the function f p : (0, ∞) → R defined by f p = x p
can be given by
[1]
p (x, y) =
f
sin(pπ)
π
Z ∞
0
sp
(x + s)(y + s)
ds
(15)
for all x, y ∈ (0, ∞). For the function g : (0, ∞) → R defined as g(x) = x−1 for all x ∈ (0, ∞), note that the
divided differences can expressed compactly as
1c.f. [Car10, Lemma2.8]
g[1](x, y) = −
1
xy
7
(16)
for all x, y ∈ (0, ∞), since g′(x) = −1/x2 and g[1](x, y) = (1/x − 1/y)/(x − y) = −1/xy for all x, y ∈ (0, ∞)
with x 6= y. For any positive definite matrix A ∈ P+
n and any other matrix B ∈ Hn, one has that
(A + tB)−1 − A−1
t
lim
t→0
= Dg(A)(B),
where Dg(A) : Hn → Hn is the linear Fr´echet differential operator (as defined in (8)) for g(x) = x−1. In the
case when A = diag(α1, . . . αn) is diagonal and positive definite, the ij-entry of the matrix Dg(A)(B) are
computed as
(cid:0)Dg(A)(B)(cid:1)ij = g[1](αi, αj)Bij = −
Bij
αiαj
for any B ∈ Hn.
4.1.1 The case p ∈ (−1, 0)
First consider the case when p ∈ (−1, 0). Let A ∈ Pn be a positive matrix satisfying im(P) ⊆ im(A), which
we may suppose without loss of generality is diagonal with A = diag(α1, . . . , αn). We may assume that
α1, . . . , αr > 0 are the nonzero eigenvalues and that αr+1 = · · · = αn = 0. Let B ∈ Hn be an n × n Hermitian
matrix and suppose there exists a positive value ε > 0 such that A + tB ∈ Pn for all t ∈ [0, ε). One may
compute fα,P(A + tB) for any t ∈ [0, ε) using the integral representation in (13) as
fα,P(A + tB) = Tr(P(A + tB)p)
=
− sin(pπ)
π
Z ∞
0
Tr(cid:16)P(A + tB + s1)−1(cid:17)sp ds,
whre 1 denotes the n × n identity matrix. This holds even when t = 0. The directional derivative d f p,P(A; B)
can be computed as
d f p,P(A; B) = lim
t→0+
fα,P(A + tB) − fα,P(A)
t
− sin(pπ)
0
π
= lim
t→0+
=
=
− sin(pπ)
π
− sin(pπ)
π
t
Z ∞
Tr(cid:16)P(cid:16)(A + tB + s1)−1 − (A + s1)−1(cid:17)(cid:17)sp ds
!sp ds
Tr P lim
Z ∞
Z ∞
(A + tB + s1)−1 − (A + s1)−1
Tr (P Dg(A + s1)(B)) sp ds
t→0+
0
0
t
where g : (0, ∞) → R is the function (defined earlier) g(x) = x−1 for all x ∈ (0, ∞). Note that A + s1 is
positive definite and diagonal for all s ∈ (0, ∞) with eigenvalues αi + s. Extending g to all positive definite
matrices, one sees that g is Fr´echet differentiable at the positive definite matrix A + s1 for all s > 0 where
the matrix entries of the derivative Dg(A + s1)(B) are given by
(cid:0)Dg(A + s1)(B)(cid:1)i,j = g[1](αi + s, αj + s)hei, Beji = −
hei, Beji
(αi + s)(αj + s)
for all i, j ∈ {1, . . . , n}, and the divided differences are computed as in (16). As it has been assumed that
im(P) ⊆ im(A), it holds that Pij = Tr(P eie∗
j ) = 0 whenever i ∈ {r + 1, . . . , n} or j ∈ {r + 1, . . . , n}. It
follows that
Tr(P Dg(A + s1)(B)) = −
8
r
∑
i,j=1
hei, Bejihej, Peii
(αi + s)(αj + s)
for all s > 0, where one notes that the sum above is taken from 1 to r. Making use of the integral represen-
tation for the divided differences f
[1]
p (αi, αj) in (15), it follows that
d f p,P(A; B) =
r
∑
i,j=1
hei, Bejihej, Peii
sin(pπ)
π
Z ∞
0
sp
(αi + s)(αj + s)
ds
=
r
∑
i,j=1
e∗
a hei, Beji hej, Peii f
[1]
p (αi, αj)
= Tr(P Φ f p,A(B)),
where Φ f p,A is the linear mapping defined earlier.
4.1.2 The case p ∈ (0, 1)
Now let p ∈ (0, 1). One may compute f p,P(A + tB) using integral representation in (14) as
f p,P(A + tB) = Tr(P(A + tB)p)
for all t ∈ [0, ε). The directional derivative d f p,P(A, B) can be computed by
=
sin(pπ)
π
Z ∞
0
Tr(cid:16)P(cid:16)s−11 − (A + tB + s1)−1(cid:17)(cid:17)sp ds,
d f p,P(A; B) = lim
t→0+
f p,P(A + tB) − f p,P(A)
t
sin(pπ)
π
= lim
t→0+
0
t
Z ∞
Tr(cid:16)P(cid:16)(A + s1)−1 − (A + tB + s1)−1(cid:17)(cid:17)sp ds
!sp ds
Tr P lim
Z ∞
Z ∞
Tr(cid:16)P Dg(A + s1)(B)(cid:17)sp ds
(A + s1)−1 − (A + tB + s1)−1
t→0+
t
sin(pπ)
π
0
− sin(pπ)
π
0
=
=
=
r
∑
i,j=1
hei, Beji hej, Peii f
[1]
p (αi, αj)
= Tr(P Φ f p,A(B)),
using the same arguments as before.
4.2 Directional derivatives of Tr(P log(A))
The same methods can be used to compute the derivatives of Tr(P log(A)). One may use the integral
representation of logarithm function, which holds for all x ∈ (0, ∞):
0 (cid:18) 1
log(x) = Z ∞
s
−
1
x + s(cid:19) ds.
Let f : (0, ∞) → R be the function defined as f (x) = log x for all x ∈ (0, ∞). For a positive matrix P ∈ Pn,
define the function fP as
fP(A) = Tr(P log A) = Z ∞
0
Tr(cid:18)P(cid:18) 1
s
1 − (A + s1)−1(cid:19)(cid:19) ds
9
for all A ∈ Pn. Let A ∈ Pn be a matrix satisfying im(P) ⊆ im(A). One may suppose without loss of
generality that A = diag(α1, . . . , αn) is diagonal. Let B ∈ Hn and suppose there is a value ε > 0 such that
A + tB ∈ Pn holds for all t ∈ [0, ε). Then
d fP(A; B) = lim
t→0+
fP(A + tB) − fP(A)
t
Tr(P log(A + tB)) − Tr(P log(A))
Tr(cid:18)P lim
t→0+
t
(A + s1)−1 − (A + tB + s1)−1
t
(cid:19) ds
= lim
t→0+
0
= Z ∞
= −Z ∞
0
=
r
∑
i,j=1
Tr (P Dg(A + s1)(B)) ds
hei, Beji hej, Peii f [1](αi, αj)
= Tr(P Φ f ,A(B)),
where the steps are analogous to those in Section 4.1.2. This generalizes the method in [VP98, Theorem 3].
Note that the divided differences of the function f (t) = log(t) are given by
f [1](x, y) =
1
x
log(x) − log(y)
x − y
x = y
x 6= y
for all x, y ∈ (0, ∞).
5 Lower bound for derivative of A 7→ Tr(PA−1)
We now consider the function f : (0, ∞) → R defined by f (t) = t−1 and the corresponding trace functional
fP : P+
n → R defined as
fP(A) = Tr(PA−1)
for all positive definite matrices A. Let A ∈ Pn be a positive semidefinite matrix with spectral decomposi-
tion
A =
αi viv∗
i ,
n
∑
i=1
where α1, . . . , αr > 0 are the nonzero eigenvalues. Let B ∈ Hn and suppose there exists a positive value
ε > 0 such that A + tB ∈ Pn for all t ∈ [0, ε). Let λ1(t), . . . , λn(t) and u1(t), . . . , un(t) be the analytic
eigenvalues and eigenvectors of A + tB such that λi(0) = αi and ui(0) = vi for all i ∈ {1, . . . , n}. As in the
proof of Theorem 1, for each i ∈ {r, . . . , n} we define the function hi : R → R by hi(t) = hui(t), Pui(t)i such
that hi(0) = h′
i(0) = 0. Moreover, note that
i (0) = hu′
h′′
i(0), Pu′
i(0)i ≥ 0,
10
since P is positive semidefinite. Furthermore, we may assume (as in the proof of Theorem 1) that λ′
holds for all i ∈ {r + 1, . . . , n}. Then
i(0) > 0
d fP(A; B) = lim
t→0+
fP(A + tB) − fP(A)
t
= Tr(P Φ f ,A(B)) +
= Tr(P Φ f ,A(B)) +
n
∑
i=r+1
n
∑
i=r+1
lim
t→0+
lim
t→0+
hi(t)
tλj(t)
f (λi(t))hi(t)
t
,
(17)
where, for i ∈ {r + 1, . . . , n}, the limits in the final line reduce to
hi(t)
tλi(t)
=
h′′
i (0)
2
lim
t→0+
t
λi(t)
=
lim
t→0+
h′′
i (0)
2λ′
i(0)
≥ 0
since λ′
i(0) = 0 and h′′
i (0) ≥ 0. Thus Tr(P Φ f ,A(B)) provides the lower bound for the directional derivative,
d fP(A; B) ≥ Tr(P Φ f ,A(B)),
(18)
and this inequality is strict in general unless im(B) ⊆ im(A).
Indeed, to show that the inequality in (18) can be strict, consider the following example. Let A, P ∈ P2
and B ∈ H2 be the 2 × 2 matrices
A = P =(cid:18)1
0
0
0(cid:19)
and
B =(cid:18)0 1
1 1(cid:19) .
For the function f (x) = x−1, we may define fP(A) as fP(A) = 1, and the linear mapping Φ f ,A : H2 → H2
is given by
Φ f ,A(cid:18)(cid:18)a
∗ ∗(cid:19)(cid:19) =(cid:18)a
0 0(cid:19)
∗
0
(i.e., it simply picks out the entry in the upper-left corner and zeros out the other entries). It follows that
Tr(PΦ f ,A(B)) = 0 for these matrices, but that
(A + tB)−1 =(cid:18)1
t
t
t(cid:19)−1
=
1
1 − t(cid:18) 1 −1
1/t(cid:19)
−1
such that Tr(P(A + tB)−1) = 1/(1 − t) for all t > 0, and thus
d fP(A; B) = lim
t→0+
fP(A + tB) − fP(A)
t
= lim
t→0+
Tr(cid:0)P(A + tB)−1(cid:1) − 1
t
= lim
t→0+
1
1 − t
= 1.
Hence d fP(A; B) > Tr(PΦ f ,A(B)) for these matrices.
References
[Bha97] Rajendra Bhatia. Matrix Analysis, volume 169 of Graduate Texts in Mathematics. Springer, 1997.
[Car10] Eric A. Carlen. Trace inequalities and quantum entropy: an introductory course. In Roert Sims and
Daniel Ueltschi, editors, Entropy quantum Arizona Sch. Anal. with Appl., pages 73 -- 140. American
Mathematical Society, 2010.
11
[FG11]
Shmuel Friedland and Gilad Gour. An explicit expression for the relative entropy of entanglement
in all dimensions. J. Math. Phys., 52(5):052201, jul 2011.
[GGF14] Mark W. Girard, Gilad Gour, and Shmuel Friedland. On convex optimization problems in quan-
tum information theory. J. Phys. A Math. Theor., 47(50):505302, dec 2014.
[HP14]
Fumio Hiai and D´enes Petz. Introduction to Matrix Analysis and Applications. Universitext. Springer
International Publishing, 2014.
[Kat80] Tosio Kato. Perturbation Theory for Linear Operators. Springer, Berlin, 1980.
[Ras11] Alexey E. Rastegin. Upper continuity bounds on the relative q-entropy for q > 1. J. Math. Phys.,
52(6):1 -- 7, 2011.
[VP98] Vlatko Vedral and Martin B. Plenio. Entanglement measures and purification procedures. Phys.
Rev. A, 57(3):1619 -- 1633, mar 1998.
[Wat18]
John Watrous. Theory of Quantum Information. Cambridge University Press, 2018.
12
|
1811.01752 | 1 | 1811 | 2018-11-02T12:59:33 | Wave-front sets in non-quasianalytic setting for Fourier Lebesgue and modulation spaces | [
"math.FA"
] | We define and study wave-front sets for weighted Fourier-Lebesgue spaces when the weights are moderate with respect to the associated functions for general sequences $\{ M_p\} $ which satisfy Komatsu's conditions $(M.1) - (M.3)'$. In particular, when $\{ M_p\} $ is the Gevrey sequence ($M_p = p!^s$, $s>1$) we recover some previously observed results. Furthermore, we consider wave-front sets for modulation spaces in the same setting, and prove the invariance property related to the Fourier-Lebesgue type wave-front sets. | math.FA | math |
WAVE-FRONT SETS IN NON-QUASIANALYTIC
SETTING FOR FOURIER LEBESGUE AND
MODULATION SPACES
NENAD TEOFANOV
Abstract. We define and study wave-front sets for weighted Fourier-
Lebesgue spaces when the weights are moderate with respect to the
associated functions for general sequences {Mp} which satisfy Ko-
matsu's conditions (M.1) − (M.3)′. In particular, when {Mp} is
the Gevrey sequence (Mp = p!s, s > 1) we recover some previ-
ously observed results. Furthermore, we consider wave-front sets
for modulation spaces in the same setting, and prove the invariance
property related to the Fourier-Lebesgue type wave-front sets.
1. Introduction
Wave-front sets in the context of Fourier-Lebesgue spaces, together
with the study of corresponding pseudodifferential operatros, were first
considered in [40], see also [41, 42, 43]. They are recently used in [7]
for a mathematical explanation of phenomena related to the interfer-
ences in the Born-Jordan distribution. The conic neighborhoods in
the definition of such wave-front sets are replaced in [18] by a filter of
neighborhoods for the study of propagation of singularities of Fourier-
Lebesgue type for partial (pseudo)differential equations, whose symbol
satisfies generalized elliptic properties. An important extension of in-
vestigations from [41, 42] to general weighted Fourier Banach spaces is
given in [2, 3].
The above mentioned results are performed in the framework of
weights of polynomial growth and, consequently, within the realm of
tempered distributions. Spaces of ultradistributions in the context of
weighted Fourier-Lebesgue type spaces were first observed in [27], see
also [28]. The sequences of the form Mp = p!s, s > 1, are used there
to define the corresponding test function spaces. This in turn leads to
the analysis of weighted Fourier-Lebesgue spaces such that the growth
of the weight function at infinity is bounded by ek·1/s, for some k > 0.
2010 Mathematics Subject Classification. Primary 35A18,35S30,42B05,35H10.
Key words and phrases. Wave-front sets, weighted Fourier-Lebesgue spaces,
Gelfand-Shilov spaces, ultradistributions.
1
2
N. TEOFANOV
In this paper we extend the results from [27] to a more general con-
text when the spaces of test functions are given by the means of {Mp}
sequences which satisfy Komatsu's conditions (M.1) − (M.3)′, see Sec-
tion 2. Note that this allows "fine tuning" between the two Gevrey
type sequences, see Remark 2.1.
The paper is organized as follows. We end the introduction with
the basic notation, and a brief account on weight functions. Section2
contains a discussion on sequences and corresponding associate func-
tions, which are the basic notions in our analysis. We proceed with
an exposition of Gelfand-Shilov spaces and other test function spaces,
and their dual spaces of ultradistributions. Section 3 contains the def-
inition of wave-front sets for weighted Fourier-Lebesgue spaces when
the weights are submultiplicative with respect to the associated func-
tion of a given non-quasianalytic sequence {Mp}. We study its basic
properties, convolution relations, and discuss its relation to some other
types of wave-front sets.
In Section 4 we first study the short-time
Fourier transform in the context of test function spaces and their du-
als from Section 2, and then define modulation spaces and recall their
basic properties. Finally, in Section 5 we introduce wave-front sets for
modulation spaces and show that they coincide with appropriate wave-
front sets from Section 3. Since we consider general non-quasianalytic
sequences {Mp}, we recover the main results from [27, 28] where the
particular case Mp = p!s, s > 1, is observed.
1.1. Basic notation. We put N = {0, 1, 2, . . . }, hxi = (1 + x2)1/2,
x ∈ Rd, xy = x · y denotes the scalar product on Rd and
h(x, ω)is = hzis = (1 + x2 + ω2)s/2,
z = (x, ω) ∈ R2d,
s ∈ R.
i.e., p ∈ Nd
0 and pj ≥ 0, we write ∂p = ∂p1
1 · · · ∂pd
i=1 xpi
(x1, . . . , xd)(p1,...,pd) = Qd
The partial derivative of a vector x = (x1, . . . , xd) ∈ Rd with respect to
xj is denoted by ∂j = ∂
. Given a multi-index p = (p1, . . . , pd) ≥
∂xj
d and xp =
0,
i=1 hixi1/αi.
Moreover, for p ∈ Nd
+, we set (p!)α = (p1!)α1 . . . (pd!)αd. In
the sequel, a real number r ∈ R+ may play the role of the vector with
constant components rj = r, so for α ∈ Rd
+, by writing α > r we mean
αj > r for all j = 1, . . . , d. By X we denote an open set in Rd, and
K ⋐ X means that K is compact subset in X.
i . Similarly, h · x1/α = Pd
0 and α ∈ Rd
The Fourier transform is normalized to be
f (ω) = F f (ω) =Z f (t)e−2πitωdt.
3
We use the brackets hf, gi to denote the extension of the inner product
hf, gi = R f (t)g(t)dt on L2(Rd) to the dual pairing between a test
function space A and its dual A′: h·, ·i = A′h·, ·iA. We use the standard
notation for usual spaces of functions and distributions, e.g. Lp(Rd),
Lp
loc(Ω), 1 ≤ p ≤ ∞, denote Lebesgue spaces and their local versions
respectively, S(Rd) denotes the Schwartz space of rapidly decreasing
test functions, etc.
Translation and modulation operators, T and M respectively, when
acting on f ∈ L2(Rd) are defined by
Txf (·) = f (· − x)
and Mxf (·) = e2πix·f (·), x ∈ Rd.
(1.1)
Then for f, g ∈ L2(Rd) the following relations hold:
MyTx = e2πix·yTxMy, (Txf )= M−x
f , (Mxf )= Tx
f , x, y ∈ Rd.
These operators are extended to other spaces of functions and distri-
butions in a natural way.
Throughout the paper, A . B denotes A ≤ cB for a suitable con-
stant c > 0, whereas A ≍ B means that c−1A ≤ B ≤ cA for some c ≥ 1.
The symbol B1 ֒→ B2 denotes the continuous and dense embedding of
the topological vector space B1 into B2.
1.2. Weights. In general, a weight function is a non-negative function
in L∞
loc.
Definition 1.1. Let ω, v be non-negative functions. Then
(1) v is called submultiplicative if
v(x + y) ≤ v(x)v(y),
∀ x, y ∈ Rd;
(2) ω is called v-moderate if
ω(x + y) . v(x)ω(y),
∀ x, y ∈ Rd.
For a given submultiplicative weight v the set of all v-moderate weights
will be denoted by Mv.
If v is even and ω ∈ Mv, then 1/v . ω . v, ω 6= 0 everywhere and
1/ω ∈ Mv.
In the sequel we assume that v is an even submultiplicative function.
Submultiplicativity implies that v is dominated by an exponential func-
tion, i.e.
v ≤ Cek ·
for some C, k > 0.
For example, every weight of the form
v(z) = eskzkb
(1 + kzk)a logr(e + kzk)
for parameters a, r, s ≥ 0, 0 ≤ b ≤ 1 satisfies the above conditions.
4
N. TEOFANOV
Let s > 1. By M{s}(Rd) we denote the set of all weights which are
moderate with respect to a weight v which satisfies v ≤ Cek · 1/s for
some positive constants C and k. The weight v satisfy the Beurling-
Domar non-quasi-analyticity condition which takes the form
log v(nx)
n2
< ∞,
x ∈ Rd.
∞Xn=0
We refer to [21] for a detailed account on weights in time-frequency
analysis.
2. Spaces of test functions and their duals
Let (Mp)p∈N0 be a sequence of positive numbers monotonically in-
creasing to infinity which satisfies:
(M.1) M 2
p ∈ N;
(M.2) There exist positive constants A, H such that
p ≤ Mp−1Mp+1,
Mp ≤ AH p min 0≤q≤pMp−qMq, p, q ∈ N0,
or, equivalently, there exist positive constants A, H such that
Mp+q ≤ AH p+qMpMq, p, q ∈ N0;
p=1 Mp−1/Mp < ∞.
(M.3)′ P∞
positive constant.
We assume that M0 = 1, and that M 1/p
p
is bounded below by a
The condition (M.3)′ provides the existence of nontrivial compactly
supported smooth functions (and therefore partitions of unity) in the
corresponding spaces of test functions.
It is therefore known as the
non-quasianalyticity condition.
The Gevrey sequences Mp = p!s, p ∈ N, s > 1, are basic examples of
sequences which satisfy (M.1) − (M.3)′.
Let (Mp)p∈N0 and (Nq)q∈N0 be sequences which satisfy (M.1). We
write Mp ⊂ Nq ((Mp) ≺ (Nq), respectively) if there are constants
H, C > 0 (for any H > 0 there is a constant C > 0, respectively) such
that Mp ≤ CH pNp, p ∈ N0. Also, (Mp)p∈N0 and (Nq)q∈N0 are said to be
equivalent if Mp ⊂ Nq and Nq ⊂ Mp hold.
Remark 2.1. The conditions (M.1) and (M.2) can be described as fol-
lows. Let (sp)p∈N0 be a sequence of positive numbers monotonically
increasing to infinity (sp ր ∞) so that for every p, q ∈ N0 there exist
A, H > 0 such that
sp+j = sp+1 · · · sp+q ≤ AH ps1 · · · sq = AH p
qYj=1
sj.
(2.1)
qYj=1
Mp := p!
1
2
pYk=0
qYk=0
(M.1) and (M.2).
Then the sequence (Sp)p∈N0 given by Sp = Qp
Conversely, if (Sp)p∈N0 given by Sp =Qp
j=1 sj, sj > 0, j ∈ N, S0 = 1,
satisfies (M.1) then the sequence (sp)p∈N0 increases to infinity. If, in
addition, it satisfies (M.2) then (2.1) holds.
j=1 sj, S0 = 1, satisfies
Furthermore, if (Mp)p∈N0 and (Nq)q∈N0 are given by
5
lk = p!
1
2 Lp,
p ∈ N0, Nq := q!
1
2
rk = q!
1
2 Rq,
q ∈ N0
(2.2)
where (rp)p∈N0 and (lp)p∈N0 are sequences of positive numbers mono-
tonically increasing to infinity such that (2.1) holds with the letter s
replaced by r and l respectively, and which satisfy: For every α ∈ (0, 1]
and every k > 1 so that kp ∈ N, p ∈ N,
max{(
rkp
rp
)2, (
lkp
lp
)2} ≤ kα,
p ∈ N.
(2.3)
Then p! ≺ MpNp and the sequences (Rp)p∈N0 and (Lp)p∈N0 (Rp =
r1 · · · rp, Lp = l1 · · · lp, p ∈ N R0 = 1, and L0 = 1) satisfy (M.1) and
(M.2). Moreover,
max{Rp, Lp} ≤ p!α/2, p ∈ N,
for every α ∈ (0, 1]. (For p, q, k ∈ Nd
0 we have Lp = Qk≤p lk, and
Rq =Qk≤q rq.) Such sequences are used in the study of localization
operators in the context of quasianalytic spaces in [10].
The associated function for a given sequence (Mp) is defined by
M(ρ) = sup
p∈N
ln+
ρpM0
Mp
,
0 < ρ < ∞,
(2.4)
where ln+ t := max{ln t, 0}, t > 0. It is a non-negative monotonically
increasing function which vanishes for sufficiently small ρ, and tends to
infinity faster than ln ρp, as ρ → ∞. Moreover, if (Mp) satisfies (M.1)
and (M.3)′, then kpp!/Mp → 0 as p → ∞.
For example, the associated function for the Gevrey sequence Mp =
p!s, p ∈ N0, s > 1, behaves at infinity as · 1/s, cf.
In fact,
the interplay between the defining sequence and its associated function
plays an important role in the theory of ultradistributions.
[35].
The following result will be intensively used in this paper. We refer
to [1] for its proof.
6
N. TEOFANOV
Lemma 2.1. Let there be given sequence (Mp) which satisfies (M.1).
Then
M(
ρk) ≤
M(ρk), ρk > 0, k = 1, . . . , n.
(2.5)
nXk=1
nXk=1
If, in addition, (Mp) satisfies (M.2), then
2M(ρ) ≤ M(Hρ) + ln+(A), ρ > 0,
(2.6)
where A and H are the constants in (M.2)¿ Furthermore, if L ≥ 1,
then there is a constant C > 0 such that
M(Lρ) ≤
3
2
LM(ρ) + C, ρ > 0,
(2.7)
and there is a constant B > 0 and a constant KL > 0 which depends
on L, such that
LM(ρ) ≤ M(BL−1ρ) + KL, ρ > 0.
(2.8)
Remark 2.2. By Lemma 2.1,
it follows that estimates of the form
f (·) . eM (h·) for some/every h > 0 and f (·) . ekM (·) for some/every
k > 0 are equivalent. This observation will be often used in proofs.
2.1. Gelfand-Shilov spaces. We give here only the basic properties
and refer to [19, 34] for a more detailed discussion and applications in
partial differential equations.
Definition 2.1. Let there be given sequences of positive numbers
(Mp)p∈N0 and (Nq)q∈N0 which satisfy (M.1) and (M.2). Let S Nq,B
Mp,A(Rd)
be defined by
S Nq,B
Mp,A(Rd) = {f ∈ C ∞(Rd) kxα∂βf kL∞ ≤ CAαMαBβNβ, ∀α, β ∈ Nd
for some positive constant C, and A = (A1, . . . , Ad), B = (B1, . . . , Bd),
A, B > 0.
0},
Gelfand-Shilov spaces ΣNq
Mp(Rd) and S Nq
ductive limits of (Fr´echet) spaces S Nq,B
ΣNq
Mp(Rd) := proj
lim
A>0,B>0
S Nq ,B
Mp,A(Rd); S Nq
Mp(Rd) are projective and in-
Mp,A(Rd) with respect to A and B:
S Nq,B
Mp,A(Rd).
Mp(Rd) := ind lim
A>0,B>0
The corresponding dual spaces of ΣNq
Mp(Rd) are the
spaces of ultradistributions of Beurling and Roumieu type respectively:
Mp(Rd) and S Nq
(ΣNq
Mp)′(Rd) := ind lim
(S Nq,B
Mp,A)′(Rd);
A>0,B>0
(S Nq
Mp)′(Rd) := proj
lim
A>0,B>0
(S Nq ,B
Mp,A)′(Rd).
Gelfand-Shilov spaces are closed under translation, dilation, multi-
plication with x ∈ Rd, and differentiation. Moreover, they are closed
under the action of certain differential operators of infinite order (ul-
tradifferentiable operators in the terminology of Komatsu).
7
Whenever nontrivial, Gelfand-Shilov spaces contain "enough func-
tions" in the following sense. A test function space Φ is "rich enough"
if
Z f (x)ϕ(x)dx = 0, ∀ϕ ∈ Φ ⇒ f (x) ≡ 0 (a.e.).
The following theorem enlightens the fundamental properties of Gelfand-
Shilov spaces implicitly contained in their definition. Among other
things, it states that the decay and regularity estimates of f ∈ S Nq
Mp(Rd)
can be studied separately.
Theorem 2.1. Let there be given sequences of positive numbers (Mp)p∈N0
and (Nq)q∈N0 which satisfy (M.1), (M.2) and p! ⊂ MpNp (p! ≺ MpNp,
respectively). Moreover, let M(·) and N(·) denote the associated func-
tions for (Mp)p∈N0 and (Nq)q∈N0 respectively. Then the following con-
ditions are equivalent:
(1) f ∈ S Nq
(2) There exist constants A, B ∈ Rd, A, B > 0 (for every A, B ∈
Mp(Rd), respectively).
Mp(Rd) (f ∈ ΣNq
Rd, A, B > 0 respectively), and there exist C > 0 such that
keM (Ax)∂qf (x)kL∞ ≤ CBqNq,
∀p, q ∈ Nd
0.
(3) There exist constants A, B ∈ Rd, A, B > 0 (for every A, B ∈
Rd, A, B > 0, respectively), and there exist C > 0 such that
kxpf (x)kL∞ ≤ CApMp
∀p, q ∈ Nd
0.
(4) There exist constants A, B ∈ Rd, A, B > 0 (for every A, B ∈
and k∂qf (x)kL∞ ≤ CBqNq,
Rd, A, B > 0, respectively), and there exist C > 0 such that
kxpf (x)kL∞ ≤ CApMp
∀p, q ∈ Nd
0.
(5) There exist constants A, B ∈ Rd, A, B > 0 (for every A, B ∈
and kωq f (ω)kL∞ ≤ CBqNq,
Rd, A, B > 0, respectively), such that
kf (x)eM (Ax) kL∞ < ∞ and k f (ω) eN (Bω)kL∞ < ∞.
Theorem 2.1 is proved in [5] and reinvented many times afterwards,
see e.g. [9, 23, 30, 34, 38, 53].
By the above characterization F S Nq
Mp (Rd) = S {Mp}(Rd), and ΣMp
Nq we put SMp
over, the Fourier transform F extends to a homeomorphism on (S {Mp})′(Rd)
and on (S (Mp))′(Rd) in a usual way.
Mp(Rd) = SMp
Nq (Rd). When Mp =
Mp(Rd) = S (Mp)(Rd). More-
8
N. TEOFANOV
Next we discuss the important case when (Mp)p∈N0 and (Nq)q∈N0 are
hcosen to be the Gevrey sequences Mp = p!r, p ∈ N0 and Nq = q!s,
q ∈ N0, for some r, s ≥ 0, then we use the notation
S Nq
Mp(Rd) = S s
r (Rd) and ΣNq
Mp(Rd) = Σs
r(Rd).
If, in addition, s = r, then we put
S {s}(Rd) = S s
s (Rd) and Σ(s)(Rd) = Σs
s(Rd).
The choice of Gevrey sequences is the most often used choice in the
literature since it serves well in different contexts. For example, when
discussing nontriviality of Gelfand-Shilov spaces we have the following:
r (Rd) is nontrivial if and only if s+r > 1, or s+r = 1
(1) the space S s
and sr > 0,
(2) if s + r ≥ 1 and s < 1, then every f ∈ S s
r (Rd) can be extended
to the complex domain as an entire function,
(3) if s + r ≥ 1 and s = 1, then every f ∈ S s
r (Rd) can be extended
to the complex domain as a holomorphic function in a strip
{x + iy ∈ Cd : y < T } some T > 0
(4) the space Σs
r(Rd) is nontrivial if and only if s + r > 1, or, if
s + r = 1 and sr > 0 and (s, r) 6= (1/2, 1/2).
We refer to [19] or [34] for the proof in the case of S s
r (Rd), and to
[36] for the spaces Σs
r(Rd), see also [54].
The discussion here above shows that Gelfand-Shilov classes S s
r (Rd)
consist of quasi-analytic functions when s ∈ (0, 1). This is in a sharp
contrast with e.g. Gevrey classes Gs(Rd), s > 1, another family of
functions commonly used in regularity theory of partial differential
equations, whose elements are always non-quasi-analytic. Recall, for
1 < s < ∞ and an open set X ∈ Rd the Gevrey class Gs(X) is given
by
Gs(X) = {φ ∈ C ∞(X) (∀K ⋐ X)(∃C > 0)(∃h > 0)
We refer to [44] for microlocal analysis in Gervey classes and note that
∂αφ(x) ≤ Chαα!s}.
sup
x∈K
Gs
0(Rd) ֒→ S s
s (Rd) ֒→ Gs(Rd),
s > 1.
When the spaces are nontrivial we have the inclusions:
Σs
r(Rd) ֒→ S s
r (Rd) ֒→ S(Rd),
and S(Rd) can be revealed as the limiting case of spaces Ss
r (Rd), i.e.
S(Rd) = S∞
∞ (Rd) = lim
s,r→∞
S s
r (Rd),
when the passage to the limit when s and r tend to infinity is inter-
preted correctly, see [19, page 169].
9
Remark 2.3. Note that Σ1/2
s(Rd) is dense in the
Schwartz space whenever s > 1/2. One may consider a "fine tuning",
that is the spaces ΣNq
1/2(Rd) = {0} and Σs
Mp(Rd) such that
{0} = Σ1/2
1/2(Rd) ֒→ ΣNq
Mp(Rd) ֒→ S Nq
Mp(Rd) ֒→ Σs
s(Rd),
s > 1/2,
see also Remark 2.1.
We refer to [55] where it is shown how to overcome the minimality
condition (Σ1/2
1/2(Rd) = 0) by transferring the estimates for kxα∂βf kL∞
into the estimates of the form kH N f kL∞ . hN (N!)2s, for some (for
every ) h > 0, where H = x2 − ∆ is the harmonic oscillator.
We also mention that the Gelfand-Shilov space of analytic functions
S (1)(Rd) := Σ1
1(Rd) plays a prominent role in the theory since it is
isomorphic to the Sato test function space for the space of Fourier
hyperfunctions. More precisely, if f ∈ S (1)(Rd) then it can be extended
to a holomorphic function f (x+iy) in the strip {x+iy ∈ Cd : y < T }
for some T > 0. According to Theorem 2.1, we have
f ∈ S (1)(Rd) ⇐⇒ sup
x∈Rd
f (x)eh·x < ∞ and sup
ω∈Rd
f (ω)eh·ω < ∞,
for every h > 0. This representation is used to establish an isomorphism
between its dual space (S (1))′(Rd) and the space of Fourier hyperfunc-
tions, see [4] for details.
r (Rd) and S r
Already in [19] it is shown that the Fourier transform is a topological
isomorphism between S s
s ), which extends
to a continuous linear transform from (S s
s )′(Rd). In par-
1/2 (Rd)
ticular, if s = r and s ≥ 1/2 then F (S s
is the smallest non-empty Gelfand-Shilov space invariant under the
Fourier transform, cf.
[53, Remark 1.2]. Similar assertions hold for
Σs
r )′(Rd) onto (S r
s )(Rd) = S s
s (Rd), and S 1/2
s (Rd) (F (S s
r ) = S r
r(Rd).
2.2. Test function spaces on open sets. Since we are interested
in non-quasianalytic classes, we restrict our intention to the sequences
which satisfy (M.1)−(M.3)′, and refer to [31] for a more general setting.
Definition 2.2. Let there be given a sequence (Mp), p ∈ Nd, which
satisfies (M.1) − (M.3)′ and let X be an open set in Rd. For a given
compact set K ⊂ X and a constant A > 0 we denote by E Mp
A,K(X) the
10
N. TEOFANOV
space of all ϕ ∈ C ∞(X) such that the norm
kϕkMp,A,K = sup
p∈Nn
0
sup
x∈K
Ap
Mp
ϕ(p)(x) < ∞.
(2.9)
Note that k · kMp,A,K is a norm in E Mp
A,K(X).
The space of functions ϕ ∈ C ∞(X) such that (2.9) holds and supp ϕ ⊆
Let (Kn)n be a sequence of compact sets such that Kn ⊂⊂ Kn+1 and
K is denoted by DMp
A (K).
S Kn = X. Then
E (Mp)(X) = proj lim
n→∞
(proj lim
A→∞
E Mp
A,Kn)(X),
E {Mp}(X) = proj lim
n→∞
(ind lim
A→0
E Mp
A,Kn)(X),
D(Mp)(X) = ind lim
n→∞
(proj lim
A→∞
DMp
A (Kn))
= ind lim
n→∞
(D(Mp)
Kn ),
D{Mp}(X) = ind lim
n→∞
(ind lim
A→0
DMp
A (Kn))
= ind lim
n→∞
(D{Mp}
Kn
).
Obviously, D(Mp)(X) (D{Mp}(X) resp.) is the subspace of E (Mp)(X)
(of E {Mp}(X) resp.) whose elements are compactly supported.
Remark 2.4. Let ∗ denote (Mp) or {Mp}. Then D∗, S ∗ and E ∗ corre-
spond to C ∞
0 , S and C ∞, respectively, and
D∗ ⊆ C ∞
0 , S ∗ ⊆ S and E ∗ ⊆ C ∞.
The spaces of linear functionals over D(Mp)(X) and D{Mp}(X), de-
noted by (D(Mp))′(X) and (D{Mp})′(X) respectively, are called the spaces
of ultradistributions of Beurling and Roumieu type respectively, while
the spaces of linear functionals over E (Mp)(X) and E {Mp}(X), denoted
by (E (Mp))′(X) and (E {Mp})′(X), respectively are called the spaces of
ultradistributions of compact support of Beurling and Roumieu type
respectively. Clearly,
(E {Mp})′(X) ⊆ (E (Mp))′(X),
(E (Mp))′(X) ⊆ (E (Mp))′(Rd)
and
(E {Mp})′(X) ⊆ (E {Mp})′(Rd).
11
Moreover,
and
(E {Mp})′(Rd) ⊆ (S {Mp})′(Rd) ⊆ (D{Mp})′(Rd)
(E (Mp))′(Rd) ⊆ (S (Mp))′(Rd) ⊆ (D(Mp))′(Rd).
Any ultra-distribution with compact support can be viewed as an el-
ement of (S (1))′(Rd). More generally, by using similar reasoning as in
the case of distributions (see [24]), it follows that E ∗ are exactly those
elements in S ∗ or D∗ with compact support.
The following fact follows from the Paley-Wiener type theorems
which can be found e.g. in [31].
Theorem 2.2. Let there be given a sequence (Mp), p ∈ Nd, which
satisfies (M.1) − (M.3)′ and let K be a compact convex set in Rd.
Then ϕ ∈ D(Mp)
resp.) if and only if for every h > 0 there
is a constant C > 0 (there are constants h > 0 and C > 0 resp.) such
that
(ϕ ∈ D{Mp}
K
K
ϕ(ξ) ≤ Ce−hM (ξ),
ξ ∈ Rd.
3. Wave-front sets in weighted Fourier-Lebesgue spaces
Although in principle both Beurling and Roumieu cases could be
treated simultaneously (as we did in Section 2), in order to simplify
the exposition, from now on we will treat the Beurling case only. See
also [28] for a discussion related to a slight difference between the cases.
Throughout the section {Mp} will always denote a sequence satisfies
(M.1) − (M.3)′ and M(ρ) denotes its associated function. For the
notational convenience, the set of weights ω moderated with respect
to the weight eM (ρ) will be denoted by MM (ρ)(Rd) (instead of a more
cumbersome notation MeM (ρ)(Rd)).
Let q ∈ [1, ∞] and let ω ∈ MM (ρ)(Rd). The (weighted) Fourier
(ω)(Rd), i. e.
(ω)(Rd) is the inverse Fourier image of Lq
Lebesgue space F Lq
F Lq
(ω)(Rd) consists of all f ∈ (S (1))′(Rd) such that
kf kF Lq
(ω)
≡ kbf · ωkLq.
is finite. If ω = 1, then the notation F Lq is used instead of F Lq
note that if ω(ξ) = hξis, then F Lq
potential space H p
s .
(ω). We
(ω) is the Fourier image of the Bessel
Remark 3.1. We may permit an x dependency for the weight ω in
the definition of Fourier Lebesgue spaces. More precisely, for each
12
N. TEOFANOV
ω ∈ MM (ρ)(R2d) we let F Lq
such that
(ω) be the set of all ultradistributions f
kf kF Lq
(ω)
≡ kbf ω(x, · )kLq
is finite. Since ω is vk-moderate it follows that different choices of x
< ∞ is independent of x.
give rise to equivalent norms, hence kf kF Lq
Therefore, a F Lq
might
depend on x.
(ω)(Rd) is independent of x although k · kF Lq
(ω)
(ω)
Next we introduce local Fourier-Lebesgue spaces of ultradistributions
related to the given sequence {Mp}. Let X be an open set in Rd and
let ω ∈ MM (ρ)(Rd). The local Fourier Lebesgue space F Lq
(ω),loc(X)
consists of all f ∈ (S (1))′(Rd) such that ϕf ∈ F Lq
(ω)(Rd) for each
ϕ ∈ D(Mp)(X). It is a Fr´echet space under the topology given by the
, where ϕ ∈ D(Mp)(X), and the
family of seminorms f 7→ kϕf kF Lq
following simple properties hold.
(ω)
Lemma 3.1. Let there be given a sequence {Mp} with the associate
function M(ρ), ρ > 0. Let X be an open set in Rd and ω ∈ MM (ρ)(Rd).
Then
F Lq
(ω)(Rd) ⊆ F Lq
(ω),loc(Rd) ⊆ F Lq
(ω),loc(X).
Furthermore, let q1, q2 ∈ [1, ∞] and ω1, ω2 ∈ MM (ρ)(Rd). Then
(ω2),loc(X), when q1 ≤ q2 and ω2 . ω1.
(ω1),loc(X) ⊆ F Lq2
F Lq1
(3.1)
(3.2)
Proof. If f ∈ F Lq
gives
(ω)(Rd) and if ϕ ∈ D(Mp)(X), then Young's inequality
kϕf kF Lq
(ω)
= kF (ϕf ) ωkLq = (2π)−d/2k(bϕ ∗ bf ) ωkLq
,
(ω)
Remark 2.2 it follows that for every N > 0 we have
. kbϕ eM (·) ∗ bf ωkLq . kbf ωkLq = kf kF Lq
if kbϕ eM (·)kL1 is finite. Since ϕ ∈ D(Mp)(X), from Theorem 2.2 and
Therefore kbϕeM (·)kLp < ∞ for every p ∈ [1, ∞], and (3.1) is proved.
It remains to prove (3.2). The inclusion in (3.2) is clear when q1 = q2
and ω2 . ω1. It remains to show that F Lq
(ω),loc increases with respect
to q. Assume, without any loss of generality, that f ∈ (E (Mp))′(X),
and that ϕ ∈ D(Mp)(Rd) is such that ϕ ≡ 1 in the neighborhood of
bϕ(ξ)eM (ξ) . e−(N +1)M (ξ)eM (ξ) = e−N M (ξ).
(3.3)
supp f . Choose p ∈ [1, ∞] such that 1/q1 + 1/p = 1/q2 + 1. Then, for
a eM (·)-moderate weight ω, it follows from Young's inequality that
13
kf kF L
q2
(ω)
for some constant C, and the result follows.
. k(bϕ ∗ bf )ωkLq2 . kbϕeM (·)kLpkbf ωkLq1 = Ckf kF L
,
q1
(ω)
(cid:3)
Next we extend the definition of wave-front sets of Fourier-Lebesgue
type given in [27, 40, 41].
Let {Mp} satisfy (M.1) − (M.3)′ and let M(ρ) denote its associated
function. Furthermore, let q ∈ [1, ∞], and Γ ⊆ Rd \ 0 be an open cone.
If f ∈ (S (1))′(Rd) and ω ∈ MM (ρ)(R2d), then we define
f F Lq,Γ
(ω)
= f F Lq,Γ
(ω),x
(3.4)
≡(cid:16)ZΓ
bf (ξ)ω(x, ξ)q dξ(cid:17)1/q
(with obvious interpretation when q = ∞). We note that · F Lq,Γ
defines a semi-norm on (S (1))′(Rd) which might attain the value +∞.
Since ω is M(ρ)-moderate it follows that different x ∈ Rd gives rise
to equivalent semi-norms f F Lq,Γ
, see Remark 3.1. Furthermore, if
Γ = Rd \ 0, f ∈ F Lq
Fourier Lebesgue norm kf kF Lq
(ω)(Rd) and q < ∞, then f F Lq,Γ
agrees with the
(ω),x
(ω),x
(ω),x
of f .
(ω),x
For the sake of notational convenience we set
B = F Lq
(ω) = F Lq
(ω)(Rd),
and · B(Γ) = · F Lq,Γ
(ω),x
.
(3.5)
(ω)
(f ) be the set of all ξ ∈ Rd \ 0 such that
We let ΘB(f ) = ΘF Lq
f B(Γ) < ∞, for some open conical neighborhood Γ = Γξ of ξ. We also
let ΣB(f ) be the complement of ΘB(f ) in Rd\0. Then ΘB(f ) and ΣB(f )
are open respectively closed subsets in Rd \ 0, which are independent
of the choice of x ∈ Rd in (3.4).
Definition 3.1. Let there be given a sequence {Mp} which satisfies
(M.1) − (M.3)′ and let M(ρ) be its associated function. Furthermore,
let q ∈ [1, ∞], B be as in (3.5), and let X be an open subset of Rd. If
ω ∈ MM (ρ)(R2d), then the wave-front set of f ∈ (D∗)′(X), WFB(f ) ≡
(f ) with respect to B consists of all pairs (x0, ξ0) in X ×(Rd \0)
WFF Lq
such that ξ0 ∈ ΣB(ϕf ) holds for each ϕ ∈ D(Mp)(X) such that ϕ(x0) 6=
0.
(ω)
The set WFB(f ) is a closed set in Rd × (Rd \ 0), since it is obvious
that its complement is open. We also note that if x ∈ Rd is fixed and
ω0(ξ) = ω(x, ξ), then WFB(f ) = WFF Lq
(f ), since ΣB is independent
of x.
(ω0)
14
N. TEOFANOV
The following theorem shows that wave-front sets with respect to
F Lq
It also shows that
(ω) satisfy appropriate micro-local properties.
such wave-front sets are decreasing with respect to the parameter q,
and increasing with respect to the weight ω.
Theorem 3.1. Let there be given a sequence {Mp} which satisfy (M.1)−
(M.3)′ and let M(ρ) be its associated function. Furthermore, let q, r ∈
[1, ∞], X be an open set in Rd and ω, ϑ ∈ MM (ρ)(R2d) be such that
r ≤ q,
and ω(x, ξ) . ϑ(x, ξ).
Also let B be as in (3.5) and put B0 = F Lr
and ϕ ∈ D(Mp)(X) then
(ϑ)(Rd). If f ∈ (D(Mp))′(X)
WFB(ϕ f ) ⊆ WFB0(f ).
Proof. When Mp = p!s, s > 1, we recover [27, Theorem 2.1]. In fact,
the more general situation when {Mp} is an arbitrary sequence which
satisfies (M.1) − (M.3)′ can be proved by using the idea of the proof
of [27, Theorem 2.1] as follows.
By the definition it is sufficient to prove
ΣB(ϕf ) ⊆ ΣB0(f )
when ϕ ∈ D(Mp)(X), ϑ = ω and f ∈ (E (Mp))′(Rd), since the statement
only involves local assertions. For the same reasons we may assume
that ω(x, ξ) = ω(ξ) is independent of x. We prove the assertion for
r ∈ [1, ∞), and leave the case r = ∞ to the reader.
By using the idea of the proof of [44, Theorem 1.6.1] we conclude
that if f ∈ (E (Mp))′(Rd) then there exists N0 > 0 such that bf (ξ)ω(ξ) .
Choose open cones Γ1 and Γ2 in Rd such that Γ2 ⊆ Γ1. It is enough
eN0M (ξ).
to prove that for every N > 0, there exist CN > 0 such that
when Γ2 ⊆ Γ1.
ξ∈Rd(cid:0)bf (ξ)ω(ξ)e−N M (ξ)(cid:1)(cid:17)
ϕf B(Γ2) ≤ CN(cid:16)f B0(Γ1) + sup
Since ω ∈ MM (ρ)(Rd) by letting F (ξ) = bf (ξ)ω(ξ) and ψ(ξ) =
bϕ(ξ)eM (ξ) we have
ϕf B(Γ2) =(cid:16)ZΓ2
(3.6)
F (ϕf )(ξ)ω(ξ)q dξ(cid:17)1/q
.(cid:16)ZΓ2(cid:16)ZRd
ψ(ξ − η)F (η) dη(cid:17)q
dξ(cid:17)1/q
. J1 + J2,
where
J1 =(cid:16)ZΓ2(cid:16)ZΓ1
J2 =(cid:16)ZΓ2(cid:16)Z∁Γ1
,
ψ(ξ − η)F (η) dη(cid:17)q
ψ(ξ − η)F (η) dη(cid:17)q
dξ(cid:17)1/q
dξ(cid:17)1/q
15
.
Let q0 be chosen such that 1/r0 + 1/r = 1 + 1/q, and let χΓ1 be the
characteristic function of Γ1. Then Young's inequality gives
J1 ≤(cid:16)ZRd(cid:16)ZΓ1
ψ(ξ − η)F (η) dη(cid:17)q
dξ(cid:17)1/q
= kψ ∗ (χΓ1F )kLq ≤ kψkLr0 kχΓ1F kLr = Cψf B0(Γ1),
where Cψ = kψkLq0 < ∞.
To estimate J2, we note that since ϕ ∈ D(Mp)(X), then by Theorem
2.2 it follows that for every N > 0 there exist CN > 0 such that
ψ(ξ) = bϕ(ξ)eM (ξ) ≤ CN e−(N +1)M (ξ)eM (ξ) ≤ CN e−N M (ξ).
Furthermore, Γ2 ⊆ Γ1 implies that
(3.7)
ξ − η > 2c max(ξ, η)
≥ c(ξ + η),
ξ ∈ Γ2, η /∈ Γ1
(3.8)
holds for some constant c > 0, since this is true when 1 = ξ ≥ η.
Now, a combination of Lemma 2.1, (3.7) and (3.8) (together with the
monotone increasing property of M(ρ)) implies that for every N1 > 0
we have
ψ(ξ − η) . Ce−2N1(M (ξ)+M (η)),
which gives
J2 .(cid:16)ZΓ2(cid:16)Z∁Γ1
.(cid:16)ZΓ2(cid:16)Z∁Γ1
e−2N1(M (ξ)+M (η))F (η) dη(cid:17)r
e−2N1(M (ξ)+M (η))eN1M (η)(e−N1M (η)F (η)) dη(cid:17)r
dξ(cid:17)1/r
dξ(cid:17)1/r
. sup
η∈Rd
e−N1M (η)F (η)).
This implies (3.6) and the proof is finished.
(cid:3)
16
N. TEOFANOV
3.1. Comparisons to other types of wave-front sets. Let ω ∈
Mv(R2d) be moderated with respect to the weight v of a polyno-
mial growth at infinity, and let f ∈ D′(X). Then the wave frpont set
WFF Lq
(f ) in Definition 3.1 agrees with the wave-front set introduced
in [41, Definition 3.1]. Therefore, the information on regularity in the
background of wave-front sets of Fourier-Lebesgue type in Definition
3.1 might be compared to the information obtained from the classical
wave-front sets, cf. Example 4.9 in [41].
(ω)
Next we compare the wave-front sets introduced in Definition 3.1 to
the wave-front sets in spaces of ultradistributions given in [24, 37, 44].
Let s > 1 and let X be an open subset of Rd. The ultradistribu-
tion f ∈ (D(s))′(X) (f ∈ (D{s})′(X)) is (s)-micro-regular ({s}-micro-
regular) at (x0, ξ0) if there exists ϕ ∈ D(s)(X) (ϕ ∈ D{s}(X)) such that
ϕ(x) = 1 in a neighborhood of x0 and an open cone Γ which contains
ξ0 such that
F (ϕf )(ξ) . e−N ξ1/s
,
ξ ∈ Γ,
(3.9)
for each N > 0 (for some N > 0). The (s)-wave-front set ({s}-wave-
front set) of f , WF(s)(f ) (WF{s}(f )) is defined as the complement in
X × Rd \ 0 of the set of all (x0, ξ0) where f is (s)-micro-regular ({s}-
micro-regular), cf. [44, Definition 1.7.1].
The {s}-wave-front set WF{s}(f ) can be found in [37] and it coincides
with certain wave-front set WFL(f ) introduced in [24, Chapter 8.4].
Next we modify the definitions from [41, 27].
Let there be given a sequence {Mp} which satisfy (M.1)−(M.3)′ and
let M(ρ) be its associated function. Furthermore, let ωj ∈ MM (ρ)(R2d),
qj ∈ [1, ∞] when j belongs to some index set J, and let B be the array
of spaces, given by
(Bj) ≡ (Bj)j∈J, where Bj = F Lqj
j ∈ J. (3.10)
(ωj ) = F Lqj
(ωj)(Rd),
If f ∈ (D(Mp))′(Rd), and (Bj) is given by (3.10), then we let Θsup
(Bj )(f )
be the set of all ξ ∈ Rd \ 0 such that for some Γ = Γξ and each j ∈ J it
holds f Bj(Γ) < ∞. We also let Θinf
(Bj )(f ) be the set of all ξ ∈ Rd \ 0 such
that for some Γ = Γξ and some j ∈ J it holds f Bj(Γ) < ∞. Finally we
let Σsup
(Bj )(f ) and
Θinf
(Bj )(f ) be the complements in Rd \ 0 of Θsup
(Bj )(f ) and Σinf
(Bj )(f ) respectively.
Definition 3.2. Let there be given a sequence {Mp} which satisfy
(M.1) − (M.3)′ and let M(ρ) be its associated function. Furthermore,
let J be an index set, qj ∈ [1, ∞], ωj ∈ MM (ρ)(R2d) when j ∈ J, (Bj)
be as in (3.10), and let X be an open subset of Rd.
17
(1) The wave-front set of f ∈ (D(Mp))′(X), of sup-type with respect
(Bj )(f ), consists of all pairs (x0, ξ0) in X × (Rd \ 0)
(Bj )(ϕf ) holds for each ϕ ∈ D(Mp)(X) such that
to (Bj), WF sup
such that ξ0 ∈ Σsup
ϕ(x0) 6= 0;
(2) The wave-front set of f ∈ (D(Mp))′(X), of inf-type with respect
(Bj )(f ) consists of all pairs (x0, ξ0) in X × (Rd \ 0)
(Bj )(ϕf ) holds for each ϕ ∈ D(Mp)(X) such that
to (Bj), WF inf
such that ξ0 ∈ Σinf
ϕ(x0) 6= 0.
Now we are ready to rewrite the classical Gevrey wave-front sets
WF{s}(f ) and WF(s)(f ) in terms of wave-front sets introduced in Def-
inition 3.2.
Proposition 3.1. [27] Let s > 1, and let Bj be the same as in (3.10)
with qj ∈ [1, ∞] and ωj(ξ) ≡ ejξ1/s. Then the following is true:
(1) if f ∈ (D{s})′(Rd), then
WF inf
WFBj (f ) = WF{s}(f ) ⊆ WF(s)(f );
(2) if f ∈ (D(s))′(Rd), then
(Bj )(f ) =\j>0
WF(s)(f ) =[j>0
WFBj (f ) ⊆ WF sup
(Bj )(f ).
Remark 3.2. We recall that if f ∈ D′(Rd), and ωj(x, ξ) = hξij for
j ∈ J = N, then it follows that WF sup
(Bj )(f ) in Definition 3.2 is equal to
the standard wave front set WF(f ) in Chapter VIII in [24].
3.2. Convolution. We finish the section by recalling that the convo-
lution properties, valid for standard wave-front sets of Hormander type,
also hold for the wave-front sets of Fourier Lebesgue types, see [42, 43]
for related results in the framework of tempered distributions. More
generally, the following convolution result holds true.
Theorem 3.2. Let there be given a sequence {Mp} which satisfy (M.1)−
(M.3)′ and let M(ρ) be its associated function. Furthermore, let q, q1, q2 ∈
[1, ∞] and let ω, ω1, ω2 ∈ MM (ρ)(Rd) satisfy
1
q1
+
1
q2
=
1
q
and ω(ξ) . ω1(ξ)ω2(ξ).
(3.11)
Then the convolution map (f1, f2) 7→ f1 ∗ f2 from S (1)(Rd) × S (1)(Rd) to
S (1)(Rd) extends to a continuous mapping from F Lq1
(ω2)(Rd)
to F Lq
(ω)(Rd). This extension is unique if q1 < ∞ or q2 < ∞.
(ω1)(Rd)×F Lq2
18
N. TEOFANOV
If f1 ∈ F Lq1
supports, then
(ω1),loc(Rd), f2 ∈ (D(Mp))′(Rd) and f1 or f2 have compact
WFF Lq
(ω)
(f1∗f2) ⊆ { (x+y, ξ) ; x ∈ supp f1 and (y, ξ) ∈ WFF L
(f2) }.
q2
(ω2)
The proof is omitted, since the arguments for the first part of Theo-
rem are the same as in the proof of [42, Lemma 2.1], taking into account
that S (1) is dense in F Lq
(ω) when q < ∞. The second part of Theorem
3.2 can be proved in the same way as [28, Theorem 2.2].
4. Modulation Spaces
In this section we first recall the action of the short-time Fourier
transform on Gelfand-Shilov spaces and their dual spaces, and then
proceed with modulation spaces and their properties. Since the short-
time Fourier transform gives a phase-space description of a function or
distribution, we first extend Definition 2.1.
Definition 4.1. Let there be given sequences of positive numbers
(Mp)p∈N0, (Nq)q∈N0, ( Mp)p∈N0, ( Nq)q∈N0 which satisfy (M.1) and (M.2).
We define S Nq, Nq,B
(R2d) to be the set of smooth functions f ∈ C ∞(R2d)
Mp, Mp,A
such that
kxα1ωα2∂β1
x ∂β2
ω f kL∞ ≤ CAα1+α2Mα1 Mα2Bβ1+β2Nβ1 Nβ2,
∀α1, α2, β1, β2 ∈ Nd
0},
and for some A, B, C > 0. Gelfand-Shilov spaces are projective and
inductive limits of S Nq, Nq,B
Mp, Mp,A
(R2d):
ΣNq, Nq
Mp, Mp
S Nq, Nq
Mp, Mp
(R2d) := proj
lim
A>0,B>0
(R2d) := ind lim
A>0,B>0
S Nq, Nq,B
Mp, Mp,A
S Nq , Nq,B
Mp, Mp,A
(R2d);
(R2d).
Clearly, the corresponding dual spaces are given by
(ΣNq, Nq
Mp, Mp
(S Nq, Nq
Mp, Mp
)′(R2d) := ind lim
)′(R2d) := proj
(S Nq, Nq,B
Mp, Mp,A
(S Nq , Nq,B
Mp, Mp,A
A>0,B>0
lim
A>0,B>0
)′(R2d);
)′(R2d).
(R2d) to ΣMp, Mp
Nq, Nq
By Theorem 2.1, the Fourier transform is a homeomorphism from
ΣNq, Nq
(R2d) and, if F1f denotes the partial Fourier
Mp, Mp
transform of f (x, ω) with respect to the x variable, and if F2f denotes
the partial Fourier transform of f (x, ω) with respect to the ω variable,
then F1 and F2 are homeomorphisms from ΣNq, Nq
(R2d)
Mp, Mp
(R2d) to ΣNq, Mp
Mp, Nq
and ΣNq, Mp
Mp, Nq
replaced by S Nq, Nq
Mp, Mp
(R2d), respectively. Similar facts hold when ΣNq, Nq
Mp, Mp
(R2d), (ΣNq, Nq
Mp, Mp
)′(R2d) or (S Nq, Nq
Mp, Mp
)′(R2d).
When Mp = Mp and Nq = Nq we use usual abbreviated notation:
Mp(R2d) = S Nq, Nq
S Nq
Mp, Mp
(R2d) and similarly for other spaces.
19
(R2d) is
4.1. Short-time Fourier transform. Let (Mp)p∈N0 satisfy (M.1) and
(M.2). For any given f, g ∈ SMp
Mp(Rd), respectively) the
short-time Fourier transform (STFT) of f with respect to the window
g is given by
Mp (Rd) (f, g ∈ ΣMp
Vgf (x, ξ) = (2π)−d/2ZRd
f (y) g(y − x) e−ihξ,yi dy .
The following theorem (and its variations) is a folklore, in particu-
lar in the framework of the duality between S(R2d) and S ′(R2d). For
Gelfand-Shilov spaces we refer to e.g. [23, 50, 52, 54].
Theorem 4.1. Let there be given sequences (Mp)p∈N0 and (Nq)q∈N0
which satisfy (M.1), (M.2) and
{N.1} :
(∃H > 0)(∃A > 0) p!1/2 ≤ AH pMp, p ∈ N0.
If f, g ∈ S Nq
continuous map from (S Nq
Mp(Rd), then Vφf ∈ S Nq,Mp
Conversely, if Vφf ∈ S Nq,Mp
Next, assume that (Mp)p∈N0 and (Nq)q∈N0 satisfy (M.1), (M.2) and
Mp(Rd).
Mp)′(Rd) × (SMp
Mp,Nq (Rdd) then f, g ∈ S Nq
Nq )′(Rd) into (S Nq ,Mp
Mp,Nq )′(R2d).
Mp,Nq (Rdd) and extends uniquely to a
(N.1) :
(∀H > 0)(∃A > 0) p!1/2 ≤ AH pMp, p ∈ N0.
If f, g ∈ ΣNq
continuous map from (ΣNq
Mp(Rd), then Vφf ∈ ΣNq,Mp
Mp)′(Rd) × (ΣMp
Mp,Nq (Rdd) then f, g ∈ ΣNq
Conversely, if Vφf ∈ ΣNq,Mp
Mp,Nq(Rd) and extends uniquely to a
Nq )′(Rd) into (ΣNq,Mp
Mp,Nq)′(R2d).
Mp(Rd).
The conditions {N.1} and (N.1) are taken from [33] where they are
Mp(Rd)
called nontriviality conditions for the spaces SMp
respectively, see also [32].
Mp (Rd) and ΣMp
We will also need the following proposition when proving that the
wave-front sets of Fourier-Lebesgue and modulation space types are
the same. The first part is an extension of [9, Proposition 4.2].
Proposition 4.1. Let {Mp} satisfies (M.1) − (M.3)′ and let M(ρ)
denotes its associated function. Then the following is true:
20
N. TEOFANOV
(1) if f ∈ (E (Mp))′(Rd) and φ ∈ S (Mp)(Rd), then
Vφf (x, ξ) . e−hM (x)eεM (ξ),
(4.1)
for some ε > 0 and for every h > 0;
(2) if f ∈ (D(Mp))′(Rd) and φ ∈ D(Mp)(Rd)\0, then f ∈ (E (Mp))′(Rd),
if and only if supp Vφf ⊆ K × Rd for some compact set K, and
then
Vφf (x, ξ) . eεM (ξ),
(4.2)
for some ε > 0.
Proof. We only prove (1) and (3). The other statements follow by
similar arguments and are left for the reader. As before, we will use
Remark 2.2 in our calculations. Recall, f ∈ (E (Mp))′(Rd) implies that
for some ε > 0, cf. [44, Theorem 1.6.1].
For φ ∈ S (Mp)(Rd) and ψ ∈ D(Mp)(Rd) such that ψ = 1 in supp f by
bf (ξ) . eεM (ξ),
Theorem 4.1, Lemma 2.1 and Remark 2.2 it follows that
Vψφ(x, ξ) . e−hM (x)−kM (ξ),
for every h, k > 0. Now straight-forward calculations give
Vφf (x, ξ) = (Vφ(ψf ))(x, ξ) . (Vψφ(x, ·) ∗ bf)(ξ)
=Z Vψφ(x, ξ − η)bf (η) dη .Z e−hM (x)−2εM (ξ−η)eεM (η) dη
≤ e−hM (x)Z e−2εM (η)+2εM (ξ)+εM (η) dη . e−hM (x)+2εM (ξ),
and (1) follows.
Next we prove (3). First assume that φ ∈ D(Mp)(Rd) \ 0 and f ∈
(E (Mp))′(Rd). Since both φ and f have compact support, it follows that
supp(Vφf ) ⊆ K × Rd. Furthermore, by slightly modifying the proof of
[54, Theorem 2.5] we conclude thay
Vφf (x, ξ) . eε(M (x)+M (ξ)),
for some ε > 0, see also [27, Proposition 3.2]. Since Vφf (x, ξ) has
compact support in the x-variable, it follows that
Vφf (x, ξ) . eεM (ξ).
For the opposite direction, assume that supp Vφf ⊆ K ×Rd, for some
compact set K. Assume that supp φ ⊆ K and choose ϕ ∈ D(s)(Rd)
21
such that supp ϕ ∩ 2K = ∅. Then
(f, ϕ) = (kφkL2)−2(Vφf, Vφϕ) = 0,
which implies that f has compact support. Here the first equality is the
Moyal's identity (cf.
[20]). This implies that f has compact support
and the condition f ∈ (D(Mp))′(Rd) now gives f ∈ (E (Mp))′(Rd).
(cid:3)
4.2. Modulation spaces. The modulation space norms traditionally
measure the joint time-frequency distribution of f ∈ S ′, we refer, for
instance, to [11], [20, Ch. 11-13] and the original literature quoted there
for various properties and applications. It is usually sufficient to ob-
serve modulation spaces with weights which admit at most polynomial
growth at infinity. However the study of ultra-distributions requires a
more general approach that includes the weights of exponential or even
superexponential growth, cf. [9, 55]. Note that the general approach in-
troduced already in [11] includes the weights of sub-exponential growth.
We refer to [13, 14] for related but even more general constructions,
based on the general theory of coorbit spaces.
Depending on the growth of the weight function m, different Gelfand-
Shilov classes may be chosen as fitting test function spaces for modu-
lation spaces, see [9, 50, 55]. The widest class of weights allowing to
define modulation spaces is the weight class N . A weight function m
on Rd belongs to N if it is a continuous, positive function such that
m(z) = o(ecz2
),
for z → ∞,
∀c > 0,
(4.3)
with z ∈ Rd. For instance, every function m(z) = eszb, with s > 0
and 0 ≤ b < 2, is in N . Thus, the weight m may grow faster than
exponentially at infinity. For example, the choice m ∈ N \∪vMv, when
the weights v satisfy the Beurling-Domar condition from Introduction,
is related to the spaces of quasianalytic functions, [10]. We notice that
there is a limit in enlarging the weight class for modulation spaces,
imposed by Hardy's theorem: if m(z) ≥ Cecz2, for some c > π/2, then
the corresponding modulation spaces are trivial [22].
Definition 4.2. Let m ∈ N , and g a non-zero window function in
S 1/2
m (Rd) consists of
1/2 (Rd). For 1 ≤ p, q ≤ ∞ the modulation space M p,q
all f ∈ (S 1/2
m (R2d) (weighted mixed-norm
spaces). The norm on M p,q
1/2 )′(Rd) such that Vgf ∈ Lp,q
m is
kf kM p,q
m = kVgf kLp,q
m = ZRd(cid:18)ZRd
Vgf (x, ω)pm(x, ω)p dx(cid:19)q/p
dω!1/q
22
N. TEOFANOV
(with obvious changes if either p = ∞ or q = ∞). If p, q < ∞, the
1/2 in the M p,q
modulation space M p,q
m -
norm. If p = ∞ or q = ∞, then M p,q
1/2 in the
weak∗ topology.
m is the norm completion of S 1/2
m is the completion of S 1/2
When f, g ∈ S (1)(Rd), the above integral is convergent thanks to
Theorem 4.1. Namely, for a given m ∈ Mv there exist l > 0 such that
m(x, ω) ≤ Celk(x,ω)k and therefore
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ZRd(cid:18)ZRd
Vgf (x, ω)pm(x, ω)p dx(cid:19)q/p
≤ C(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ZRd(cid:18)ZRd
dω(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Vgf (x, ω)pelpk(x,ω)k dx(cid:19)q/p
< ∞
dω(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
since by Theorems 4.1 and Theorem 2.1 we have Vgf (x, ω) < Ce−sk(x,ω)k
for every s > 0. This implies S (1) ⊂ M p,q
m .
In particular, when m is a polynomial weight of the form m(x, ω) =
s,t (Rd) for the modulation spaces
hxithωis we will use the notation M p,q
which consists of all f ∈ S ′(Rd) such that
kf kM p,q
s,t
≡ ZRd(cid:18)ZRd
Vφf (x, ω)hxithωisp dx(cid:19)q/p
dω!1/q
< ∞
(with obvious interpretation of the integrals when p = ∞ or q = ∞).
m , and if m(z) ≡ 1 on R2d, then
m , and so on.
m instead of M p,p
m and M p,p
we write M p,q and M p for M p,q
If p = q, we write M p
In the next proposition we show that M p,q
m (Rd) are Banach spaces
whose definition is independent of the choice of the window g ∈ M 1
v \
{0}. In order to do so, we need the adjoint of the short-time Fourier
transform.
For given window g ∈ S (1) and a function F (x, ξ) ∈ Lp,q
m (R2d) we
(formally) define V ∗
g F by
hV ∗
g F, f i := hF, Vgf i.
Proposition 4.2. Let v be a submultiplicative weight. Fix m ∈ Mv
and g, ψ ∈ S (1), with hg, ψi 6= 0. Then
m (Rd), and
m ≤ CkVψgkL1
m (R2d) → M p,q
g F kM p,q
g : Lp,q
vkF kLp,q
m .
(1) V ∗
kV ∗
(4.4)
(2) The inversion formula holds: IM p,q
m = hg, ψi−1V ∗
g Vψ, where IM p,q
m
stands for the identity operator.
23
(3) M p,q
m (Rd) are Banach spaces whose definition is independent on
the choice of g ∈ S (1) \ {0}.
(4) The space of admissible windows can be extended from S (1) to
M 1
v .
Proof. We refer to [9] for the proof which is based on the proof of [20,
Proposition 11.3.2.]. Note that in (4) the density of S (1) in M p,q
m is es-
sential. This fact is not obvious, and we refer to [6] for the proof. Then
we may proceed by using the standard arguments, cf.
[20, Theorem
(cid:3)
11.3.7].
The following theorem lists some basic properties of modulation
spaces. We refer to [11, 20, 23, 38, 51, 54] for the proof.
Theorem 4.2. Let p, q, pj, qj ∈ [1, ∞] and s, t, sj, tj ∈ R, j = 1, 2.
Then:
(1) M p,q
s,t (Rd) are Banach spaces, independent of the choice of φ ∈
S(Rd) \ 0;
(2) if p1 ≤ p2, q1 ≤ q2, s2 ≤ s1 and t2 ≤ t1, then
s1,t1 (Rd) ⊆ M p2,q2
S(Rd) ⊆ M p1,q1
s,t (Rd) = S(Rd), ∪s,tM p,q
s2,t2 (Rd) ⊆ S ′(Rd);
s,t (Rd) = S ′(Rd);
(3) ∩s,tM p,q
(4) Let 1 ≤ p, q ≤ ∞, and let ws(x, ω) = esk(x,ω)k, x, ω ∈ Rd. Then
Σ1
(Rd),
M p,q
(Σ1
ws (Rd),
M p,q
1/ws
1)′(Rd) =[s≥0
1 )′(Rd) =\s>0
M p,q
ws (Rd),
(S 1
M p,q
1/ws
(Rd).
1(Rd) = S (1)(Rd) =\s≥0
1 (Rd) = S {1}(Rd) =[s>0
S 1
1
p + 1
p′ = 1
q + 1
q′ = 1.
(5) For p, q ∈ [1, ∞), the dual of M p,q
s,t (Rd) is M p′,q′
−s,−t(Rd), where
Remark 4.1. In the context of quasianalytic Gelfand-Shilov spaces, we
recall (a special case of) [54, Theorem 3.9]: Let s, t > 1/2 and set
wh(x, ω) ≡ eh(x1/t+ω1/s), h > 0, x, ω ∈ Rd.
Then
Σs
S s
t (Rd) = \h>0
t (Rd) = [h>0
M p,q
wh (Rd),
M p,q
wh (Rd),
(Σs
(S s
t )′(Rd) = [h>0
t )′(Rd) = \h>0
M p,q
1/wh
(Rd),
M p,q
1/wh
(Rd).
Modulation spaces include the following well-know function spaces:
(1) M 2(Rd) = L2(Rd), and M 2
t,0(Rd) = L2
t (Rd);
24
N. TEOFANOV
(2) The Feichtinger algebra: M 1(Rd) = S0(Rd);
(3) Sobolev spaces: M 2
(4) Shubin spaces: M 2
0,s(Rd) = H 2
s (Rd) = L2
s (Rd) = {f f (ω)hωis ∈ L2(Rd)};
s (Rd) = Qs(Rd), cf. [49].
s(Rd) ∩ H 2
5. The invariance property of Wave-front sets
Next we define wave-front sets with respect to modulation spaces
and show that they agree with corresponding wave-front sets of Fourier
Lebesgue types. More precisely, we prove that [41, Theorem 6.1] holds
if the weights of polynomial growth are replaced by more general sub-
multiplicative weights.
Let there be given a sequence {Mp} which satisfies (M.1)−(M.3)′ and
let M(ρ) denote its associated function. Furthermore, let p, q ∈ [1, ∞],
and Γ ⊆ Rd\0 be an open cone. If f ∈ (S (1))′(Rd) and ω ∈ MM (ρ)(R2d),
then we define
f B(Γ) = f B(φ,Γ) ≡(cid:16)ZΓ(cid:16)ZRd
Vφf (x, ξ)ω(x, ξ)p dx(cid:17)q/p
dξ(cid:17)1/q
when B = M p,q
(ω) = M p,q
(ω)(Rd).
(5.1)
when Γ = Rd \ 0 and φ ∈ S (s)(Rd), and
We note that f B(Γ) = kf kM p,q
that f B(φ,Γ) might attain +∞.
Furthermore, when B = M p,q
(ω)
(ω), the sets ΘB(f ), ΣB(f ) and WFB(f )
with respect to the modulation space B are defined in the same way as
in Section 3, after replacing the semi-norms of Fourier Lebesgue types
in (3.4) with the semi-norms in (5.1).
Proposition 5.1. Let there be given a sequence of positive numbers
(Mp)p∈N0 which satisfies (M.1) − (M.3)′, and let M(ρ), ρ > 0, be its
associated function. If f ∈ (D(Mp))′(Rd) then WFM p,q
(f ) is independent
of p and φ ∈ S (Mp)(Rd) \ 0 in (5.1) .
(ω)
Proof. We may assume that f ∈ (E (Mp))′(Rd) and that ω(x, ξ) = ω(ξ)
since the statements only concern local assertions.
(ω)
We follow the idea of the proof of [27, Theorem 3,1], and in order to
(f ) is independent of φ ∈ S (Mp)(Rd) \ 0, we assume
prove that WFM p,q
that φ, φ1 ∈ S (Mp)(Rd) \ 0 and let · C1(Γ) be the semi-norm in (5.1)
after φ has been replaced by φ1. Let Γ1 and Γ2 be open cones in Rd
such that Γ2 ⊆ Γ1. The asserted independency of φ follows if we prove
that
f C(Γ2) ≤ C(f C1(Γ1) + 1),
(5.2)
25
for some positive constant C. Let
Ω1 = { (x, ξ) ; ξ ∈ Γ1 } ⊆ R2d
and Ω2 = ∁Ω1 ⊆ R2d,
with characteristic functions χ1 and χ2 respectively, and set
Fk(x, ξ) = Vφ1f (x, ξ)ω(ξ)χk(x, ξ),
k = 1, 2,
and G = Vφφ1(x, ξ)eM (ξ). Since ω is v-moderate, it follows from [20,
Lemma 11.3.3] that
Vφf (x, ξ)ω(x, ξ) .(cid:0)(F1 + F2) ∗ G(cid:1)(x, ξ),
which implies that
f C(Γ2) . J1 + J2,
where
Jk =(cid:16)ZΓ2(cid:16)ZRd
(Fk ∗ G)(x, ξ)p dx(cid:17)q/p
dξ(cid:17)1/q
,
k = 1, 2.
By Young's inequality
J1 ≤ kF1 ∗ GkLp,q
1
≤ kGkL1kF1kLp,q
1
= Cf C1(Γ1),
where C = kGkL1 = kVφφ1(x, ξ)eM (ξ)kL1 < ∞, in view of Proposition
4.1.
Next we consider J2. For ξ ∈ Γ2 fixed and integrating over η ∈ ∁Γ1,
it follows from Propositon 4.1 and Lemma 2.1 that for some ε > 0 and
every N, h > 0 we have that (F2 ∗ G)(x, ξ) is bounded by
CZZR2d
e−N M (y)eεM (η)e−h(M (x−y)+M (ξ−η))eM (ξ−η) dydη,
for some constant C > 0. Therefore, there exist a constant c > 0 such
that
(F2 ∗ G)(x, ξ)
e−N M (y)eεM (η)e−hM (x−y)−hc(M (ξ)+M (η))e(M (ξ)+M (η)) dydη
.ZZR2d
. e(−N +h)M (x)e(1−hc)M (ξ)ZZR2d
e−hM (y)e(1+ε−hc)M (η) dydη,
. e(−N +h)M (x)e(1−hc)M (ξ) < ∞,
26
N. TEOFANOV
since N > 0 and h can be chosen arbitrarily. Therefore
J2 =(cid:16)ZΓ2(cid:16)ZRd
(F2 ∗ G)(x, ξ)p dx(cid:17)q/p
dξ(cid:17)1/q
.(cid:16)ZΓ2(cid:16)ZRd(cid:16)e(−N +h)M (x)e(1−hc)M (ξ)(cid:17)p
dx(cid:17)q/p
dξ(cid:17)1/q
< ∞.
This proves that (5.2), and hence WFC(f ) is independent of φ ∈
S (s)(Rd) \ 0.
(cid:3)
The main result of this section, Theorem 5.1, now follows from
Proposition 5.1 and calculations given in the proof of [27, Theorem
3.1]. For that reason we omit the proof.
Theorem 5.1. Let there be given a sequence of positive numbers (Mp)p∈N0
which satisfies (M.1) − (M.3)′, and let M(ρ), ρ > 0 be its associated
function. Let p, q ∈ [1, ∞] and ω ∈ MM (ρ)(R2d). If f ∈ (D(Mp))′(Rd)
then
WFF Lq
(ω)
(f ) = WFM p,q
(ω)
(f ).
(5.3)
Finally, note that for a given sequence of positive numbers (Mp)p∈N0
which satisfies (M.1)−(M.3)′, and its associated function M(ρ), ρ > 0,
when p, q ∈ [1, ∞], ω ∈ MM (ρ)(R2d) and f ∈ (E (Mp))′(Rd), then it
follows from the definition of wave-front sets that then
f ∈ B ⇐⇒ WFB(f ) = ∅,
when B is equal to F Lq
obtain
(ω) or M p,q
(ω). In particular, by Theorem 5.1 we
F Lq
(ω) ∩ (E (Mp))′(Rd) = M p,q
(ω) ∩ (E (Mp))′(Rd),
and we recover Corollary 6.2 in [41], Theorem 2.1 and Remark 4.6 in
[45].
Acknowledgement
This work is supported by MPNTR through Project 174024 and
DS028 -- Tifmofus.
References
1. R. D. Carmichael, A. Kami´nski, S. Pilipovi´c, Boundary Values and Convolution
in Ultradistribution Spaces World Scientific, 2007.
2. S. Coriasco, K. Johansson, J. Toft, Local wave-front sets of Banach and Fr´echet
types, and pseudo-differential operators, Monatsh. Math. 169 (3-4), 285 -- 316
(2013)
27
3. S. Coriasco, K. Johansson, J. Toft, Global wave-front sets of Banach, Fr´echet
and modulation space types, and pseudo-differential operators. J. Differential
Equations 254 (8), 3228 -- 3258 (2013)
4. Chung, J., Chung, S.-Y., Kim, D., A characterization for Fourier hyperfunc-
tions, Publ. Res. Inst. Math. Sci. 30 (2), 203 -- 208 (1994)
5. Chung, J., Chung, S.-Y., Kim, D., Characterization of the Gelfand -- Shilov
spaces via Fourier transforms, Proceedings of the American Mathematical So-
ciety 124(7), 2101 -- 2108 (1996)
6. E. Cordero, Gelfand -- Shilov window classes for weighted modulation spaces.
Integral Transforms Spec. Funct. 18 (11-12), 829 -- 837 (2007)
7. E. Cordero, M. de Gosson, F. Nicola, On the reduction of the interferences in
the Born-Jordan distribution. Appl. Comput. Harmon. Anal. 44 (2), 230 -- 245
(2018)
8. E. Cordero and K. Grochenig. Time-frequency analysis of localization operators.
J. Funct. Anal., 205(1):107 -- 131, 2003.
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, S. Pilipovi´c, L. Rodino, N. Teofanov, E. Cordero, S. Pilipovic,
L. Rodino, N. Teofanov - Quasianalytic Gelfand-Shilov spaces and localization
operators, Rocky Mountain Journal of Mathematics, vol. 40 (4) (2010), 1123-
1147.
11. H. G. Feichtinger Modulation spaces on locally compact abelian groups. Techni-
cal report, University of Vienna, Vienna, 1983; also in: M. Krishna, R. Radha,
S. Thangavelu (Eds) Wavelets and their applications, Allied Publishers Pri-
vate Limited, New Delhi, Mumbai, Kolkata, Chennai, Hagpur, Ahmadabad,
Bangalore, Hyderabad, Lucknow, 2003, pp. 99 -- 140.
12.
Modulation spaces: Looking back and ahead, Sampl. Theory Signal Im-
age Process., 5 (2006), 109 -- 140.
13. H. G. Feichtinger and K. H. Grochenig Banach spaces related to integrable group
representations and their atomic decompositions, I, J. Funct. Anal. 86(1989),
307 -- 340.
14. H. G. Feichtinger and K. H. Grochenig Banach spaces related to integrable
group representations and their atomic decompositions, II, Monatsh. Math. 108
(1989), 129 -- 148.
15. H. G. Feichtinger and K. Grochenig. Gabor frames and time-frequency analysis
of distributions, J. Funct. Anal. 146 (1997), 464 -- 495.
16. H. G. Feichtinger, T. Strohmer, editors, Gabor Analysis and Algorithms: The-
ory and Applications, Birkhauser, 1998.
17. H. G. Feichtinger, T. Strohmer, editors, Advances in Gabor Analysis,
Birkhauser, 2003.
18. G. Garello, A. Morando, Inhomogeneous microlocal propagation of singularities
in Fourier Lebesgue spaces. J. Pseudo-Differ. Oper. Appl. 9 (1), 47 -- 93 (2018)
19. I. M. Gelfand, G. E. Shilov, Generalized Functions II Academic Press, New
York, 1968.
20. K. Grochenig, Foundations of Time-Frequency Analysis, Birkhauser, Boston,
2001.
28
N. TEOFANOV
21. K. Grochenig, Weight functions in time-frequency analysis, In L. Rodino and
et al., editors, Pseudodifferential Operators: Partial Differential Equations and
Time-Frequency Analysis, volume 52, 343 - 366, 2007.
22. K. Grochenig and G. Zimmermann. Hardy's theorem and the short-time Fourier
transform of Schwartz functions. J. London Math. Soc., 63:205 -- 214, 2001.
23. K. Grochenig, G. Zimmermann, Spaces of test functions via the STFT, Journal
of Function Spaces and Applications, vol. 2 no. 1 (2004), 25 -- 53.
24. L. Hormander, The Analysis of Linear Partial Differential Operators, vol I,
Springer-Verlag, Berlin, 1983.
25. L. Hormander, Lectures on Nonlinear Hyperbolic Differential Equations,
Springer-Verlag, Berlin, 1997.
26. K. Johansson, S. Pilipovi´c, N. Teofanov, J. Toft, Gabor pairs, and a discrete
approach to wave-front sets, Monatsh. Math. 166 (2) (2012), 181199
27. K. Johansson, S. Pilipovi´c, N. Teofanov, J. Toft, Karoline Johansson, Stevan
Pilipovic, Nenad Teofanov, Joachim Toft Micro-local analysis in some spaces
of ultradistributions, Publications de lInstitut Mathematique, 92 (106) (2012),
124
28. K. Johansson, S. Pilipovi´c, N. Teofanov, J. Toft, A note on wave-front sets
of Roumieu type ultradistributions, chapter in Pseudo-Differential Opera-
tors, Generalized Functions and Asymptotics, Operator Theory: Advances
and Applications (S. Molahajloo, S. Pilipovic, J. Toft, M.W. Wong, editors),
Birkhauser, 231:239252 (2013)
29. Karoline Johansson, Stevan Pilipovic, Nenad Teofanov, Joachim Toft Resolu-
tion of wavefront set via discrete sets, PAMM Proc. Appl. Math. Mech. 13,
495-496 (2013)
30. Kami´nski, A., Perisi´c, D., Pilipovi´c, S., On Various Integral Transformations
of Tempered Ultradistributions, Demonstratio Math., 33(3), 641-655 (2000)
31. Komatsu, H., Ultradistributions, I, Structure theorems and a characterization,
J. Fac. Sci. Univ. Tokyo Sect. IA 20 (1973), 25 -- 105.
32. Langenbruch, M., Hermite functions and weighted spaces of generalized func-
tions, Manuscripta Math., 119, 269 -- 285 (2006)
33. Lozanov -- Crvenkovi´c, Z., Perisi´c, D., Hermite expansions of elements of
Gelfand-Shilov spaces in quasianalytic and non quasianalytic case, Novi Sad
J. Math., 37 (2), 129 -- 147 (2007)
34. Nicola, F., Rodino, L., Global Pseudo-differential calculus on Euclidean spaces,
Pseudo-Differential Operators. Theory and Applications 4, Birkhauser Verlag,
(2010)
35. Petche, H.-J., Generalized functions and the boundary values of holomorphic
functions, J. Fac. Sci. The University of Tokyo, Sec. IA, 31(2), 391 -- 431 (1984)
36. Pilipovi´c, S., Tempered ultradistributions, Bollettino della Unione Matematica
Italiana, 7(2-B), 235 -- 251 (1988)
37. S. Pilipovi´c, Microlocal analysis of ultradistributions, Proc. Amer. Math. Soc.,
126 (1998), 105-113.
38. Pilipovi´c, S., Teofanov, N., Wilson bases and ultra-modulation spaces, Math.
Nachr., 242, 179 -- 196 (2002)
39. S. Pilipovi´c and N. Teofanov, Pseudodifferential operators on ultra-modulation
spaces J. Funct. Anal. 208:194 -- 228, 2004.
29
40. S. Pilipovi´c, N. Teofanov, J. Toft, Wave-front sets in Fourier Lebesgue spaces,
Rendiconti del Seminario Matematico (Universita e Politecnico di Torino), 66
(4) (2008), 299 -- 319
41. S. Pilipovi´c, N. Teofanov, J. Toft, Micro-local analysis with Fourier Lebesgue
spaces. Part I, J. Fourier Anal. Appl., 17 (3) (2011), 374 -- 407.
42. S. Pilipovi´c, N. Teofanov, J. Toft, Micro-local analysis in Fourier Lebesgue and
modulation spaces. Part II, J. Pseudo-Differ. Oper. Appl. 1 (3) (2010), 341 --
376.
43. S. Pilipovi´c, N. Teofanov, J. Toft - Singular support and FLq continuity of
pseudo-differential operators, in Approximation and Computation, a volume
dedicated to 60th anniversary of G.V. Milovanovic (edited by W. Gautschi, G.
Mastroianni, and Th.M. Rassias), Springer, (2010) 357 - 376.
44. L. Rodino, Linear Partial Differential Operators in Gevrey Spaces, World Sci-
entific, 1993.
45. M. Ruzhansky, M. Sugimoto, N. Tomita, J. Toft Changes of variables in mod-
ulation and Wiener amalgam spaces, Math. Nachr., 284 (2011), 2078 - 2092.
46. M. Ruzhansky, V. Turunen, Pseudo-Differential Operators and Symmetries:
Background Analysis and Advanced Topics, Birkhauser, Boston, 2010.
47. M. Ruzhansky, V. Turunen, Quantization of pseudo-differential operators on
the torus, , J. Fourier Anal. Appl., published Online first, 2009.
48. L. Schwartz Th´eorie des Distributions, I -- II, Hermann & Cie, Paris, 1950 -- 51.
49. M. A. Shubin. Pseudodifferential Operators and Spectral Theory. Springer-
Verlag, Berlin, second edition, 2001.
50. Teofanov, N., Ultradistributions and time-frequency analysis,
in Pseudo-
differential Operators and Related Topics, Operator Theory: Advances and Ap-
plications, P. Boggiatto, L. Rodino, J. Toft, M.W. Wong, editors, Birkhauser,
164:173 -- 191, 2006.
51. Teofanov, N., Modulation spaces, Gelfand-Shilov spaces and pseudodifferential
operators, Sampl. Theory Signal Image Process, 5 (2), 225 -- 242 (2006)
52. Teofanov, N., Gelfand-Shilov spaces and localization operators, Funct. Anal.
Approx. Comput. 7 (2), 135-158 (2015)
53. Teofanov, N., The Grossmann-Royer transform, Gelfand-Shilov spaces, and
continuity properties of localization operators on modulation spaces, Mathe-
matical Analysis and ApplicationsPlenary Lectures, L. Rodino, J. Toft, editors,
Springer, 2018.
54. Toft, J., The Bargmann transform on modulation and Gelfand-Shilov spaces,
with applications to Toeplitz and pseudo-differential operators, J. Pseudo-
Differ. Oper. Appl., 3 (2), 145-227 (2012)
55. Toft, J., Images of function and distribution spaces under the Bargmann trans-
form, J. Pseudo-Differ. Oper. Appl. 8 (1), 83 -- 139 (2017)
56. Toft, J., Matrix Parameterized Pseudo-differential Calculi on Modulation
Spaces. In: Oberguggenberger M., Toft J., Vindas J., Wahlberg P. (eds) Gen-
eralized Functions and Fourier Analysis. Operator Theory: Advances and Ap-
plications, 260, Birkhauser (2017)
30
N. TEOFANOV
N. Teofanov, University of Novi Sad, Faculty of Sciences, Depart-
ment of Mathematics and Informatics, Trg Dositeja Obradovi´ca 4,
21000 Novi Sad, Serbia
E-mail address: [email protected]
|
1606.02937 | 2 | 1606 | 2016-09-02T07:41:04 | Uncertainty Relations in the Framework of Equalities | [
"math.FA",
"math-ph",
"math-ph",
"quant-ph"
] | We study the Schr\"odinger-Robertson uncertainty relations in an algebraic framework. Moreover, we show that some specific commutation relations imply new equalities, which are regarded as equality versions of well-known inequalities such as Hardy's inequality. | math.FA | math |
Uncertainty Relations in the Framework of Equalities
Tohru Ozawaa, Kazuya Yuasab
aDepartment of Applied Physics, Waseda University, Tokyo 169-8555, Japan
bDepartment of Physics, Waseda University, Tokyo 169-8555, Japan
Abstract
We study the Schrodinger-Robertson uncertainty relations in an algebraic framework. Moreover, we show
that some specific commutation relations imply new equalities, which are regarded as equality versions of
well-known inequalities such as Hardy's inequality.
Keywords: uncertainty relations
2010 MSC: 81S05, 26D, 46C
1. Introduction
In this paper, we study the Schrodinger-Robertson uncertainty relations as corollaries of equalities in
a scalar product space. Moreover, we give a number of characterizations in the case where the associated
inequalities are in fact equalities. Our presentation is based exclusively on an algebraic observation on the
standard Cauchy-Schwarz inequality and could presumably provide a clear and explicit understanding of
uncertainty relations from the point of view of orthogonality. As applications, we show that some specific
commutation relations, in the Hilbert space L2(Rn) of square integrable functions on the Euclidean space Rn
of dimensions n, imply new norm equalities in L2(Rn), which are regarded as equality versions of well-known
inequalities such as dilation and Hardy type inequalities. In particular, we give a method of recognizing
Hardy type inequalities in the framework of commutation relations of operators.
This paper is organized as follows. In Section 2, we characterize the Schrodinger-Robertson uncertainty
relations in the framework of equalities in a scalar product space. In Section 3, we give a number of examples
of uncertainty relations on the basis of equalities in L2(Rn). In the Appendix, we summarize basic theorems
on the Cauchy-Schwarz inequality in an algebraic setting.
Throughout the paper, H denotes a complex vector space endowed with scalar product ( · · ) : H × H ∋
(u, v) 7→ (uv) ∈ C, which is linear (resp. antilinear) in the first (resp. second) variable. The associated norm
is defined by kuk = (uu)1/2, u ∈ H.
There is a large literature on the uncertainty relations. We refer the readers to [8, 10, 6, 15, 26] and
references therein.
2. Uncertainty Relations
Let A and B be symmetric operators in H with domains D(A) and D(B), respectively. In this and next
sections, we use the terminology of operator theory (see [11, 23] for instance).
According to Mourre [19], we define the commutator [A, B] as a sesquilinear form on H × H by
([A, B]ϕψ) = (BϕAψ) − (AϕBψ),
ϕ, ψ ∈ M ≡ D(A) ∩ D(B).
(2.1)
It coincides with the usual definition AB − BA on D(AB) ∩ D(BA), which is smaller than M . In this paper,
we adopt the definition (2.1) to avoid the domain problem as much as possible (see [17]) and assume that
Preprint submitted to . . .
M 6= {0}
(2.2)
October 16, 2018
to avoid trivial cases.
Similarly, we define the anticommutator {A, B} by
({A, B}ϕψ) = (BϕAψ) + (AϕBψ),
ϕ, ψ ∈ M.
It is straightforward to verify that for all ϕ ∈ M ,
• ([A, B]ϕϕ) ∈ iR,
([A, B]ϕϕ) = −2i Im(AϕBϕ) = 2i Re i(AϕBϕ)
= 2i Im(BϕAϕ) = −2i Re i(BϕAϕ).
• ({A, B}ϕϕ) ∈ R,
({A, B}ϕϕ) = 2 Re(AϕBϕ) = 2 Im i(AϕBϕ)
= 2 Re(BϕAϕ) = 2 Im i(BϕAϕ).
• (AϕBϕ) =
• (BϕAϕ) =
1
2
1
2
({A, B}ϕϕ) −
({A, B}ϕϕ) +
1
2
1
2
([A, B]ϕϕ).
([A, B]ϕϕ).
(2.3)
(2.4)
(2.5)
(2.6)
(2.7)
(2.8)
(2.9)
We now summarize algebraic identities related to the Schrodinger-Robertson uncertainty relations. The
identities below are the direct consequences of the theorems in the Appendix.
Theorem 2.1. Let ϕ ∈ M satisfy Aϕ 6= 0, Bϕ 6= 0. Then,
Bϕ
Bϕ
(AϕBϕ) =
1
Aϕ
kAϕk
∓ i
Aϕ
kAϕk
∓
2! ,
±i([A, B]ϕϕ) = kAϕkkBϕk 2 −(cid:13)(cid:13)(cid:13)(cid:13)
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
2! ,
±({A, B}ϕϕ) = kAϕkkBϕk 2 −(cid:13)(cid:13)(cid:13)(cid:13)
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:16)([A, B]ϕϕ)2 + ({A, B}ϕϕ)2(cid:17)1/2
= kAϕkkBϕk
2!2
+ 1 −
1 −
2(cid:13)(cid:13)(cid:13)(cid:13)
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
+ eiθ Bϕ
2!
= kAϕkkBϕk 1 −
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
for any θ ∈ R, where sgn z = z/z, z ∈ C \ {0}, and sgn 0 = 1.
− [sgn(AϕBϕ)]
Aϕ
kAϕk
1
Aϕ
kAϕk
Bϕ
1
(2.10)±
(2.11)±
1/2
Aϕ
kAϕk
1
2(cid:13)(cid:13)(cid:13)(cid:13)
± ieiθ Bϕ
2!2
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
(2.12)
Remark 2.1. The standard uncertainty inequalities follow directly from equality (2.12). Indeed, equality
(2.12) implies the inequality
kAϕkkBϕk ≥ (AϕBϕ) =
,
(2.13)
1
2(cid:16)([A, B]ϕϕ)2 + ({A, B}ϕϕ)2(cid:17)1/2
which is called Schrodinger-Robertson uncertainty inequality [10, 13]. It is further bounded from below by
which is known as the Robertson uncertainty inequality [12].1
kAϕkkBϕk ≥
1
2
([A, B]ϕϕ),
(2.14)
1There are interesting developments on the Heisenberg uncertainty principle, beyond the Schrodinger-Robertson uncertainty
relations. See for instance [20, 21, 4, 28, 24, 14, 3, 16, 5] and references therein.
2
Moreover, characterizations of extremizers are given by:
Theorem 2.2. Let ϕ ∈ M . Then, the statements in each of the following Parts (1) -- (5) are equivalent:
(1)
(i) ({A, B}ϕϕ) = ±2kAϕkkBϕk.
(ii) kBϕkAϕ = ±kAϕkBϕ.
(iii) (AϕBϕ) = ±kAϕkkBϕk.
(2)
(i) i([A, B]ϕϕ) = ±2kAϕkkBϕk.
(ii) kBϕkAϕ = ±ikAϕkBϕ.
(iii) (AϕBϕ) = ±ikAϕkkBϕk.
(3)
(i) ({A, B}ϕϕ) = 2kAϕkkBϕk.
(ii) ([A, B]ϕϕ) = 0, (AϕBϕ) = kAϕkkBϕk.
(iii) 2kBϕk2Aϕ = ({A, B}ϕϕ)Bϕ.
(iv) 2kAϕk2Bϕ = ({A, B}ϕϕ)Aϕ.
(4)
(i) ([A, B]ϕϕ) = 2kAϕkkBϕk.
(ii) ({A, B}ϕϕ) = 0, (AϕBϕ) = kAϕkkBϕk.
(iii) 2kBϕk2Aϕ = −([A, B]ϕϕ)Bϕ.
(iv) 2kAϕk2Bϕ = ([A, B]ϕϕ)Aϕ.
(5)
(i) (AϕBϕ) = kAϕkkBϕk.
(ii) kBϕkAϕ = [sgn(AϕBϕ)]kAϕkBϕ.
(iii) kBϕk2Aϕ = (AϕBϕ)Bϕ.
(iv) kAϕk2Bϕ = (AϕBϕ)Aϕ.
3. Applications
In this section, we give a number of examples of commutation relations between operators in the Hilbert
space L2(Rn) of square integrable functions on Rn as well as related norm identities which are regarded as
equality versions of well-known inequalities. We follow the standard notation to denote a point in Rn by
n)1/2. The gradient
x = (x1, . . . , xn) ∈ Rn. The associated Euclidean length is defined as x = (x2
operator is defined as ∇ = (∂1, . . . , ∂n), where ∂j = ∂/∂xj is the partial differential operator in the jth
direction.
1 + · · · + x2
3.1. Momentum and Position Operators
Let A = −i∇ and B = x. More precisely,
Aϕ = (−i∂1ϕ, . . . , −i∂nϕ),
Bϕ = (x1ϕ, . . . , xnϕ)
(3.1)
(3.2)
for any ϕ ∈ C∞0 (Rn; C), compactly supported smooth functions on Rn. In fact, the natural domains of A
and B are given respectively by
D(A) = H 1(Rn) = {ϕ ∈ L2(Rn); ∂jϕ ∈ L2(Rn) for all j with 1 ≤ j ≤ n},
D(B) = {ϕ ∈ L2(Rn); xjϕ ∈ L2(Rn) for all j with 1 ≤ j ≤ n},
(3.3)
(3.4)
3
where the derivatives are understood to be distributional derivatives and H 1 denotes the standard Sobolev
space of order one. Since Aϕ and Bϕ are Cn-valued, the corresponding natural Hilbert space is given by
H = L2(Rn; Cn) with scalar product
((ϕ1, . . . , ϕn)(ψ1, . . . , ψn)) =
(ϕjψj) =
n
Xj=1
n
Xj=1ZRn
ϕjψj dx,
(3.5)
where ϕj, ψj ∈ L2(Rn; C), 1 ≤ j ≤ n. Since C∞0 (Rn) is dense in L2(Rn), all computations will be carried
out on C∞0 and then on M ≡ D(A) ∩ D(B) by density.
Theorem 3.1.
(1) Let ϕ ∈ M satisfy xϕ 6= 0, ∇ϕ 6= 0. Then, we have
= kxϕk2 + k∇ϕk2 − kxϕ + ∇ϕk2,
xϕ
kxϕk
+
2!
∇ϕ
k∇ϕk(cid:13)(cid:13)(cid:13)(cid:13)
nkϕk2 = −2 Re(xϕ∇ϕ)
= kxϕkk∇ϕk 2 −(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
1
xϕ
kxϕk
∇ϕ
2! .
k∇ϕk(cid:13)(cid:13)(cid:13)(cid:13)
and
(xϕ∇ϕ) = kxϕkk∇ϕk 1 −
− [sgn(xϕ∇ϕ)]
(2) Let ϕ ∈ M . Then, the following statements are equivalent:
(i) nkϕk2 = kxϕk2 + k∇ϕk2.
(ii) xϕ = −∇ϕ.
(iii) There exists θ ∈ R such that, for any x ∈ Rn, ϕ satisfies
ϕ(x) = eiθ 1
πn/4 kϕk exp(cid:18)−
x2
2 (cid:19) .
(3) Let ϕ ∈ M satisfy xϕ 6= 0. Then, the following statements are equivalent:
(i) nkϕk2 = 2kxϕkk∇ϕk.
(ii) k∇ϕkxϕ = −kxϕk∇ϕ.
(iii) There exists θ ∈ R such that, for any x ∈ Rn, ϕ satisfies
ϕ(x) = eiθ(cid:18) k∇ϕk
πkxϕk(cid:19)n/4
kϕk exp(cid:18)−
k∇ϕk
kxϕk
x2
2 (cid:19) .
(4) Let ϕ ∈ M satisfy xϕ 6= 0. Then, the following statements are equivalent:
(i) (xϕ∇ϕ) = kxϕkk∇ϕk.
(ii) k∇ϕkxϕ = [sgn(xϕ∇ϕ)]kxϕk∇ϕ.
(iii) There exists θ ∈ R such that, for any x ∈ Rn, ϕ satisfies
ϕ(x) = eiθ(cid:18)−[Re sgn(xϕ∇ϕ)]
k∇ϕk
πkxϕk(cid:19)n/4
kϕk exp(cid:18)[sgn (xϕ∇ϕ)]
k∇ϕk
kxϕk
x2
2 (cid:19) .
Note that for ∇ϕ 6= 0
Re(xϕ∇ϕ) = Re (xϕ∇ϕ) = −
n
2
kϕk2 < 0
and hence Re sgn(xϕ∇ϕ) = Re sgn (xϕ∇ϕ) < 0.
4
(3.6)
(3.7)
(3.8)
(3.9)
(3.10)
(3.11)
(3.12)
(3.13)
Remark 3.1. In general, one of the statements in Part (2) implies any of the statements in Part (3). The
converse implication holds if and only if kxϕk = k∇ϕk. Moreover, one of the statements in Part (3) implies
any of the statements in Part (4). The converse implication holds if and only if (xϕ∇ϕ) = −(xϕ∇ϕ).
Remark 3.2. For n = 1, the inequality
kxϕkk∇ϕk ≥
1
2
kϕk2
(3.14)
implied by equality (3.7) is known as the Kennard uncertainty inequality, which is a version of the Robertson
uncertainty inequality (2.14) specialized to A = −i∇ and B = x. The vector ϕ saturating the Kennard
uncertainty inequality (3.14) is given by (3.11), and is known to be a squeezed state in the field of quantum
optics [25, 27]. If furthermore the vector ϕ satisfies kxϕk = k∇ϕk = 1√2
kϕk, it is reduced to (3.10), and is
called coherent state [25, 27]. A tighter inequality than (3.14),
kxϕkk∇ϕk ≥ (xϕ∇ϕ) =
,
(3.15)
1
2(cid:16)kϕk2 + ((x · ∇ + ∇ · x)ϕϕ)2(cid:17)1/2
is available from equality (3.9) for n = 1, as a special version of the Schrodinger-Robertson uncertainty in-
equality (2.13). Inequality (3.15) is saturated by the vector ϕ given in (3.12), which is again a squeezed state.
The family of the extremizers (3.12) of the Schrodinger-Robertson uncertainty inequality (3.15) includes the
squeezed state (3.11) and the coherent state (3.10) as special cases.
Proof of Theorem 3.1.
(1) Let ϕ ∈ C∞0 . Then,
2 Re(xϕ∇ϕ) =Z x · ∇ϕ2 dx = −Z (div x)ϕ2 dx = −nkϕk2.
Moreover, we have
kxϕ + ∇ϕk2 = kxϕk2 + k∇ϕk2 + 2 Re(xϕ∇ϕ).
(3.16)
(3.17)
Then, equalities (3.6) -- (3.8) follow, by recalling (2.5) and (2.10)+. On the other hand, the identity
(3.9) follows from (2.12), by noting sgn z = 1 and (sgn z)−1 = sgn z = sgn z.
(2) The equivalence between (i) and (ii) follows from (3.8). If ϕ has the form (iii), then (ii) follows by a
direct calculation. Conversely, if (ii) holds, then
∇(cid:20)exp(cid:18) x2
2 (cid:19) ϕ(cid:21) = exp(cid:18) x2
2 (cid:19) (xϕ + ∇ϕ) = 0,
and therefore, for some c ∈ C, ϕ is represented as
ϕ(x) = c exp(cid:18)−
x2
2 (cid:19) .
(3.18)
(3.19)
Then, (iii) follows by evaluating kϕk.
(3) The equivalence between (i) and (ii) follows from (3.7). If ϕ has the form (iii), then (ii) follows by a
direct calculation. Conversely, if (ii) holds, then
∇(cid:20)exp(cid:18) k∇ϕk
kxϕk
x2
2 (cid:19) ϕ(cid:21) = exp(cid:18) k∇ϕk
kxϕk
x2
2 (cid:19)(cid:18) k∇ϕk
kxϕk
and therefore, for some c ∈ C, ϕ is represented as
ϕ(x) = c exp(cid:18)−
k∇ϕk
kxϕk
x2
2 (cid:19) .
Then, (iii) follows by evaluating kϕk.
5
xϕ + ∇ϕ(cid:19) = 0,
(3.20)
(3.21)
(4) The equivalence between (i) and (ii) follows from (3.9). If ϕ has the form (iii), then (ii) follows by a
direct calculation. Conversely, if (ii) holds, then
∇(cid:20)exp(cid:18)−[sgn (xϕ∇ϕ)]
k∇ϕk
kxϕk
x2
2 (cid:19) ϕ(cid:21)
= exp(cid:18)−[sgn (xϕ∇ϕ)]
k∇ϕk
kxϕk
x2
2 (cid:19)(cid:18)−[sgn (xϕ∇ϕ)]
k∇ϕk
kxϕk
xϕ + ∇ϕ(cid:19) = 0
by noting sgn z = (sgn z)−1, and therefore, for some c ∈ C, ϕ is represented as
ϕ(x) = c exp(cid:18)[sgn (xϕ∇ϕ)]
k∇ϕk
kxϕk
x2
2 (cid:19) .
Then, (iii) follows by evaluating kϕk.
We now rewrite (3.6) as
and regard (3.24) as an orthogonality relation. Then, as in [18] we notice that (3.24) yields a new equality:
Theorem 3.2. The following equality
holds for all ϕ ∈ D(A) with x · ∇ϕ ∈ L2(Rn). There does not exist ϕ ∈ D(A) \ {0} with x · ∇ϕ ∈ L2(Rn)
satisfying
kϕk2
(3.25)
kx · ∇ϕk2 =(cid:13)(cid:13)(cid:13)
ϕ(cid:17) = 0
n
2
Re(cid:16)x · ∇ϕ +
ϕ(cid:12)(cid:12)(cid:12)
ϕ(cid:13)(cid:13)(cid:13)
kx · ∇ϕk2 =(cid:16) n
2(cid:17)2
x · ∇ϕ +
n
2
n
2
kϕk ≤ kx · ∇ϕk
2
+(cid:16) n
2(cid:17)2
kϕk2.
Remark 3.3. The inequality
as well as the nonexistence of nontrivial extremizers has been proved in [22]. Theorem 3.2 is recognized as
an optimal description of (3.27) from the point of view of equalities.
Proof of Theorem 3.2. The equality (3.25) follows from (3.24), by noticing that
(3.22)
(3.23)
(3.24)
(3.26)
(3.27)
(3.28)
(3.29)
(3.30)
Moreover, since
(3.26) holds if and only if there exists a function f : Sn−1 → C satisfying
kx · ∇ϕk2 =(cid:13)(cid:13)(cid:13)(cid:16)x · ∇ϕ +
x · ∇ϕ +
2
n
2
n
2
2
ϕ(cid:13)(cid:13)(cid:13)
ϕ(cid:17) −
+(cid:16) n
2(cid:17)2
ϕ(cid:13)(cid:13)(cid:13)
n
2
kϕk2.
x · ∇ϕ +
ϕ = x−n/2x · ∇(xn/2ϕ),
ϕ(x) = x−n/2f(cid:16) x
x(cid:17)
=(cid:13)(cid:13)(cid:13)
n
2
ZSn−1
6
for all x ∈ Rn \ {0}, where Sn−1 is the unit sphere Sn−1 = {x ∈ Rn; x = 1}. Then, ϕ ∈ L2(Rn) if and only
if
f (ω)2 dσ(ω) = 0,
(3.31)
which in turn is equivalent to f = 0 and to ϕ = 0, where σ is the surface element, namely, the Lebesgue
measure on Sn−1. This proves the nonexistence of nontrivial extremizers of (3.27).
In [22], it has been proved that the inequality (3.27) is equivalent to the standard Hardy inequality for
n ≥ 3. The following theorem describes such relationship at the level of equalities:
Theorem 3.3. Let n ≥ 3. Then, the equality
for all ψ ∈ D(A) follows from (3.25). Conversely, (3.32) implies (3.25).
x
x
(cid:13)(cid:13)(cid:13)
2
· ∇ψ(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
x
x
· ∇ψ +
n − 2
2x
2 (cid:19)2
+(cid:18) n − 2
2
ψ(cid:13)(cid:13)(cid:13)
2
ψ
x(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
(3.32)
Proof. Since C∞0 (Rn \ {0}) is dense in H 1(Rn) = D(A) for n ≥ 3, we may assume that ϕ, ψ ∈ C∞0 (Rn \ {0}).
The following calculations are justified as long as ϕ, ψ ∈ C∞0 (Rn \ {0}) without restriction on the space
dimensions. First, suppose that (3.25) holds for all ϕ ∈ C∞0 (Rn \ {0}). Let ψ ∈ C∞0 (Rn \ {0}). Then, we
ψ(x), x ∈ Rn \ {0}. It follows that ϕ ∈ C∞0 (Rn \ {0}) and the left-hand side of (3.25)
define ϕ by ϕ(x) = 1
x
is rewritten as
x · ∇
x
x
x
x
x
x
x
x
x
x
kx · ∇ϕk2 =(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)
2
ψ
2
2
2
· ∇ψ −
x(cid:13)(cid:13)(cid:13)
· ∇ψ(cid:13)(cid:13)(cid:13)
· ∇ψ(cid:13)(cid:13)(cid:13)
· ∇ψ(cid:13)(cid:13)(cid:13)
· ∇ψ(cid:13)(cid:13)(cid:13)
ϕ(cid:13)(cid:13)(cid:13)
n
2
2
x · ∇ϕ +
(cid:16) n
2(cid:17)2
2
2
1
x
2
2
x
x
ψ
2
,
ψ
ψ
ψ
x
· ∇ψ
1
x
ψ(cid:13)(cid:13)(cid:13)
− 2 ReZ x
x(cid:13)(cid:13)(cid:13)
ψ dx +(cid:13)(cid:13)(cid:13)
−Z
x2 · ∇ψ2 dx +(cid:13)(cid:13)(cid:13)
x(cid:13)(cid:13)(cid:13)
+Z (cid:18)div
x2(cid:19) ψ2 dx +(cid:13)(cid:13)(cid:13)
x(cid:13)(cid:13)(cid:13)
+ (n − 1)(cid:13)(cid:13)(cid:13)
x(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
2 (cid:19)2
− (n − 1) =(cid:18) n − 2
x(cid:13)(cid:13)(cid:13)
ψ(cid:13)(cid:13)(cid:13)
n
2
n − 2
2x
· ∇ψ +
x
x
ψ
x
x · ∇
+
2
.
ψ
2
.
2
(3.33)
(3.34)
(3.35)
where we have used Gauss' divergence theorem, while the first term on the right-hand side of (3.25) is
rewritten as
Combining (3.33) and (3.34), we derive (3.32) from (3.25) with ϕ = 1
x
ψ, noticing that
Conversely, suppose that (3.32) holds for all ψ ∈ C∞0 (Rn \ {0}). Let ϕ ∈ C∞0 (Rn \ {0}). Then, we define
ψ = xϕ, x ∈ Rn \ {0}. It follows that ψ ∈ C∞0 (Rn \ {0}) and all the computations in (3.33) and (3.34) can
be traced backward to imply (3.25).
In [23], the standard Hardy type inequalities of the form
≤
(cid:13)(cid:13)(cid:13)
ψ
x(cid:13)(cid:13)(cid:13)
x
x
2
n − 2(cid:13)(cid:13)(cid:13)
· ∇ψ(cid:13)(cid:13)(cid:13)
≤
2
n − 2
k∇ψk
(3.36)
are referred to as the uncertainty principle lemma. Here we have derived (3.36) as a corollary to (3.24),
which is equivalent to (3.6), which in turn is regarded as an original form of the uncertainty relation between
the position and momentum operators.
7
3.2. Generator of Dilations and Free Hamiltonian
Let A = 1
2i (x · ∇ + ∇ · x) = −ix · ∇ − i n
2 and B = −∆ = −∇ · ∇ with natural domains
D(A) = {ϕ ∈ H 1(Rn); x · ∇ϕ ∈ L2(Rn)},
D(B) = H 2(Rn) = {ϕ ∈ L2(Rn); ∂j∂kϕ ∈ L2(Rn) for all j, k with 1 ≤ j, k ≤ n}.
(3.37)
(3.38)
The operator A is called the generator of dilations in the sense that one-parameter group of dilations
{T (θ); θ ∈ R} defined by
satisfies
T ′(0)ϕ =
d
dθ
(T (θ)ϕ)(x) = e
n
2 θϕ(eθx),
x ∈ Rn
= iAϕ.
T (θ)ϕ(cid:12)(cid:12)(cid:12)(cid:12)θ=0
[A, B]ϕ = 2iBϕ
The generator of dilations A and the free Hamiltonian B have a special commutation relation
for smooth functions ϕ ∈ C∞0 (Rn; C), where the commutator is understood to be AB − BA since C∞0 (Rn) ⊂
D(BA) ∩ D(AB) ⊂ D(A) ∩ D(B).
Theorem 3.4. Let ϕ ∈ M ≡ D(A) ∩ D(B) satisfy Aϕ 6= 0, Bϕ 6= 0. Then,
(3.39)
(3.40)
(3.41)
(3.42)
(3.43)
(3.44)
(3.45)
(3.46)
Remark 3.4. As a direct consequence, we have the inequality
2k∇ϕk2 = 2(Bϕϕ)
= −i([A, B]ϕϕ)
= −2 Im(AϕBϕ)
= kAϕkkBϕk 2 −(cid:13)(cid:13)(cid:13)(cid:13)
k∇ϕk2 ≤(cid:13)(cid:13)(cid:13)
x · ∇ϕ +
Aϕ
kAϕk
+ i
Bϕ
2! .
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
k∆ϕk,
n
2
ϕ(cid:13)(cid:13)(cid:13)
which might be new. The inequality (3.46) relates the information given by the momentum operator, the
generator of the dilations, and the free Hamiltonian.
Remark 3.5. By (3.25), we already know that
which implies (3.27) directly, as stated in Remark 3.3.
kAϕk2 = kx · ∇ϕk2 −(cid:16) n
2(cid:17)2
kϕk2,
(3.47)
Proof of Theorem 3.4. The equalities in (3.42) -- (3.45) follow from (3.41), (2.5), (2.10)−, and the equality
k∇ϕk2 = −(∆ϕϕ).
(3.48)
8
3.3. Radial Derivative and Coulomb Potential
2i ( x
x
Let n ≥ 3 as in Theorem 3.3. Let A = 1
domains
· ∇ + ∇ · x
x
) = −i x
x
· ∇ − i n−1
2x
and B = 1
x
with natural
D(A) = {ϕ ∈ L2(Rn); x
x
D(B) = {ϕ ∈ L2(Rn); 1
x
ϕ ∈ L2(Rn)},
· ∇ϕ, 1
x
ϕ ∈ L2(Rn)}.
(3.49)
(3.50)
The operator A is regarded as a symmetrized radial derivative defined by ∂r ≡ x
x
of the gradient has a pointwise decomposition
· ∇. The squared length
where Lj is the jth component of the spherical derivative defined by
∇ϕ2 = ∂rϕ2 +
Ljϕ2,
n
Xj=1
At the point x ∈ Rn, the unit outer vector is given by x
, 1 ≤ j ≤ n, where
x
ej = (0, . . . , 0, 1, 0, . . . , 0) is the standard unit vector in the jth direction. The corresponding one-parameter
family of operators acting on functions are given by
x
x
Lj ≡ ∂j −
∂r.
xj
x
and it is orthogonal to ej − xj
x
which satisfy
x
x(cid:17),
(T (θ)ϕ)(x) = ϕ(cid:16)x + θ
(Tj(θ)ϕ)(x) = ϕ(cid:16)x + θ(cid:16)ej −
T (θ)ϕ(cid:12)(cid:12)(cid:12)(cid:12)θ=0
Tj(θ)ϕ(cid:12)(cid:12)(cid:12)(cid:12)θ=0
d
dθ
d
dθ
T ′j(0)ϕ =
T ′(0)ϕ =
[A, B]ϕ = iB2ϕ
xj
x
x
x(cid:17)(cid:17),
= ∂rϕ,
= Ljϕ.
(3.51)
(3.52)
(3.53)
(3.54)
(3.55)
(3.56)
(3.57)
(3.58)
(3.59)
(3.60)
(3.61)
(3.62)
The (symmetrized) radial derivative and the Coulomb potential have a special commutation relation
for smooth functions ϕ ∈ C∞0 (Rn \ {0}; C), where the commutator is understood to be AB − BA since
C∞0 (Rn \ {0}) ⊂ D(BA) ∩ D(AB) ⊂ D(A) ∩ D(B).
Theorem 3.5. Let n ≥ 3 and let ϕ ∈ H 1(Rn) satisfy Aϕ 6= 0, Bϕ 6= 0. Then,
kBϕk2 = −i([A, B]ϕϕ)
= −2 Im(AϕBϕ)
= kAϕkkBϕk 2 −(cid:13)(cid:13)(cid:13)(cid:13)
4
Aϕ
kAϕk
+ i
Bϕ
2! ,
kBϕk(cid:13)(cid:13)(cid:13)(cid:13)
kBϕk2.
kAϕk2 = k∂rϕk2 −
(n − 1)(n − 3)
Remark 3.6. As a direct consequence of (3.60), we have the inequality
1
x
(cid:13)(cid:13)(cid:13)
ϕ(cid:13)(cid:13)(cid:13)
= kBϕk ≤ 2kAϕk = 2(cid:13)(cid:13)(cid:13)
∂rϕ +
n − 1
2x
,
ϕ(cid:13)(cid:13)(cid:13)
which might be new. The inequality (3.62) relates the information given by the radial derivative and
Coulomb potential. To be more specific, (3.62) shows that the Coulomb potential B is A (symmetrized
radial derivative)-bounded with relative bound 2.
9
Proof of Theorem 3.5. The equalities in (3.58) -- (3.60) follow from (3.57), (2.5), and (2.10)−. In the same
way as in the proof of Theorem 3.3 we calculate
· ∇ϕ +
n − 1
2x
kAϕk2 =(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)
to obtain (3.61).
x
x
x
x
x
x
x
x
2
2
2
· ∇ϕ(cid:13)(cid:13)(cid:13)
· ∇ϕ(cid:13)(cid:13)(cid:13)
· ∇ϕ(cid:13)(cid:13)(cid:13)
+ (n − 1) ReZ x
x
2
· ∇ϕ
1
x
ϕ(cid:13)(cid:13)(cid:13)
2 (cid:19)2
ϕ dx +(cid:18) n − 1
(cid:13)(cid:13)(cid:13)
2 (cid:19)2
x2 · ∇ϕ2 dx +(cid:18) n − 1
Z
(cid:13)(cid:13)(cid:13)
ϕ(cid:13)(cid:13)(cid:13)
2 (cid:19)2
x2(cid:19) ϕ2 dx +(cid:18) n − 1
Z (cid:18)div
(cid:13)(cid:13)(cid:13)
1
x
x
x
1
x
2
1
x
ϕ(cid:13)(cid:13)(cid:13)
ϕ(cid:13)(cid:13)(cid:13)
2
2
2
n − 1
n − 1
+
−
We now rewrite (3.57) or (3.59) as
Re(Bϕ − 2iAϕBϕ) = 0
2
(3.63)
(3.64)
and regard (3.64) as an orthogonality relation. Then, as in [18] we notice that (3.64) yields a new equality,
4kAϕk2 = k(2iAϕ − Bϕ) + Bϕk2
= k2iAϕ − Bϕ2 + kBϕk2,
where the right-hand side is exactly the same as
while the left-hand side is rewritten as the right-hand side of (3.61). Therefore, we have proved
4(cid:13)(cid:13)(cid:13)
4k∂rϕk2 = 4(cid:13)(cid:13)(cid:13)
x
x
· ∇ϕ +
n − 2
2x
x
x
· ∇ϕ +
n − 2
2x
2
ϕ(cid:13)(cid:13)(cid:13)
ϕ(cid:13)(cid:13)(cid:13)
2
,
1
x
ϕ(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)
+ (n − 2)2(cid:13)(cid:13)(cid:13)
2
1
x
2
ϕ(cid:13)(cid:13)(cid:13)
(3.65)
(3.66)
,
(3.67)
which is exactly the same as (3.32). We have thus derived the standard Hardy type inequality (the uncer-
tainty principle lemma [23]) from the orthogonality (3.64), which is equivalent to the commutation relation
between the radial derivative and Coulomb potential (3.57). By (3.51), the equality (3.32) or (3.67) is also
rewritten as
k∇ϕk2 −
n
Xj=1
kLjϕk2 = k∂rϕk2
2 (cid:19)2
=(cid:18) n − 2
x
x
1
x
(cid:13)(cid:13)(cid:13)
2
ϕ(cid:13)(cid:13)(cid:13)
+(cid:13)(cid:13)(cid:13)
· ∇ϕ +
n − 2
2x
.
2
ϕ(cid:13)(cid:13)(cid:13)
(3.68)
Appendix A. Basics of the Cauchy-Schwarz Inequality
In this appendix, we summarize algebraic identities related to the Cauchy-Schwarz inequality. For this
purpose, we introduce sign function sgn : C → R by
sgn z =
z/z,
z ∈ C \ {0},
1,
z = 0.
10
(A.1)
Theorem A.1. The following equalities hold for all u, v ∈ H \ {0}:
(uv) = kukkvk 1 −
± Re(uv) = kukkvk 1 −
± Im(uv) = kukkvk 1 −
− [sgn(uv)]
v
v
2! ,
kvk(cid:13)(cid:13)(cid:13)(cid:13)
2! ,
kvk(cid:13)(cid:13)(cid:13)(cid:13)
2! .
kvk(cid:13)(cid:13)(cid:13)(cid:13)
v
∓
1
1
1
u
u
u
kuk
kuk
2(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
kuk(cid:12)(cid:12)(cid:12)(cid:12)
− 2 Re(cid:18) u
kuk
∓ i
(A.2)
(A.3)±
(A.4)±
Proof. Let (uv) 6= 0. Then, we expand the square of the last norm on the right-hand side of (A.2) as
u
kuk
−
(uv)
(uv)
(cid:13)(cid:13)(cid:13)(cid:13)
2
v
kvk(cid:13)(cid:13)(cid:13)(cid:13)
=(cid:13)(cid:13)(cid:13)(cid:13)
u
kuk(cid:13)(cid:13)(cid:13)(cid:13)
= 2 − 2
2
Re[(uv)(uv)]
kukkvk(uv)
(uv)
kukkvk
,
= 2 − 2
(uv)
(uv)
(uv)
(uv)
v
kvk(cid:19) +(cid:13)(cid:13)(cid:13)(cid:13)
2
v
kvk(cid:13)(cid:13)(cid:13)(cid:13)
(A.5)
which yields (A.2). If (uv) = 0, then a similar calculation yields (A.2) in the trivial case. Equalities (A.3)±
can be proved in the same manner as (A.5). Or they follow from (A.2) by regarding H as a real vector space
with scalar product
Re( · · ) : H × H ∋ (u, v) 7→ Re(uv) ∈ R,
(A.6)
since the new scalar product Re(uv) satisfies sgn Re(uv) = 1 if and only if Re(uv) ≥ 0 while sgn Re(uv) =
−1 if and only if −Re(uv) = Re(uv) > 0. Substituting v by iv in (A.3)± implies (A.4)±.
Remark A.1. Equality (A.2) is regarded as an equality version of the Cauchy-Schwarz inequality
(uv) ≤ kukkvk.
(A.7)
Indeed, the former immediately implies the latter.
Remark A.2. Equalities (A.2), (A.3)+, and (A.4)+ have been noticed by Aldaz [1, 2] and are verified by
similar and simpler calculations as above, too. See also [7, 9] for related subjects.
Corollary A.1. The following equalities hold for all u, v ∈ H \ {0}:
v
1
1
u
±
kuk
+ 1 −
+ 1 −
Proof. The corollary follows from (A.3)±, (A.4)±, and the equality
(uv) = kukkvk
1 −
= kukkvk
1 −
2!2
kvk(cid:13)(cid:13)(cid:13)(cid:13)
2!2
kvk(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)
(uv) =(cid:16)[Re(uv)]2 + [Im(uv)]2(cid:17)1/2
kuk
±
u
1
1
v
u
kuk
u
kuk
1/2
1/2
.
(A.8)±
(A.9)±
± i
∓ i
v
2!2
kvk(cid:13)(cid:13)(cid:13)(cid:13)
2!2
kvk(cid:13)(cid:13)(cid:13)(cid:13)
v
.
(A.10)
Corollary A.1 is further generalized as:
11
Corollary A.2. The following equality holds for all u, v ∈ H \ {0} and θ ∈ R:
+ eiθ v
u
kuk
1
2(cid:13)(cid:13)(cid:13)(cid:13)
(uv) = kukkvk
1 −
2!2
kvk(cid:13)(cid:13)(cid:13)(cid:13)
Proof. Substituting v by eiθv in (A.8)+ [resp. (A.9)+] yields (A.11)+ [resp. (A.11)−].
Remark A.3. Equality (A.8)+ [resp. (A.9)+] follows from (A.11)+ [resp. (A.11)−] with θ ∈ 2πZ, while
equality (A.8)− [resp. (A.9)−] follows from (A.11)+ [resp. (A.11)−] with θ ∈ (2Z + 1)π.
Theorem A.2. Let u, v ∈ H. Then, the statements in each of the following Parts (1) -- (5) are equivalent:
2!2
kvk(cid:13)(cid:13)(cid:13)(cid:13)
+ 1 −
1
2(cid:13)(cid:13)(cid:13)(cid:13)
± ieiθ v
u
kuk
1/2
.
(A.11)±
(1)
(i) Re(uv) = ±kukkvk.
(ii) kvku = ±kukv.
(iii) (uv) = ±kukkvk.
(2)
(i) Im(uv) = ±kukkvk.
(ii) kvku = ±ikukv.
(iii) (uv) = ±ikukkvk.
(3)
(i) Re(uv) = kukkvk.
(ii) Im(uv) = 0, (uv) = kukkvk.
(iii) kvk2u = [Re(uv)]v.
(iv) kuk2v = [Re(uv)]u.
(4)
(i) Im(uv) = kukkvk.
(ii) Re(uv) = 0, (uv) = kukkvk.
(iii) kvk2u = [i Im(uv)]v.
(iv) kuk2v = −[i Im(uv)]u.
(5)
(i) (uv) = kukkvk.
(ii) kvku = [sgn(uv)]kukv.
(iii) kvk2u = (uv)v.
(iv) kuk2v = (uv)u.
Proof. If u = 0 or v = 0, then all of the equalities in the theorem trivially hold. Therefore, we assume that
u 6= 0 and v 6= 0.
(1) The equivalence between (i) and (ii) follows from (A.3)±. Given (ii), we calculate
kvk(uv) = (kvkuv) = (±kukvv) = ±kukkvk2,
(A.12)
which implies (iii) by dividing both sides by kvk > 0. Finally, (iii) implies (i) by taking the real part
of (uv).
(2) Part (2) follows from Part (1) by replacing v by iv.
12
(5) The equivalence between (i) and (ii) follows from (A.2). Given (i) and (ii), a direct calculation yields
(iii). Given (iii), we calculate
kvk2kuk2 = (kvk2uu) = ((uv)vu) = (uv)(vu) = (uv)2,
(A.13)
which implies (i) by taking its square root. This proves the equivalence among (i) -- (iii). A similar
argument shows the equivalence among (i), (ii) and (iv), or it follows by exchanging u and v in the
preceding argument.
(3) Given (i), we have the Cauchy-Schwarz inequality
kukkvk = Re(uv) ≤(cid:16)[Re(uv)]2 + [Im(uv)]2(cid:17)1/2
= (uv) ≤ kukkvk,
(A.14)
where all those inequalities turn out to be equalities, which in turn imply (ii). Given (ii), we have by
Part (5), which proved above
where the imaginary part of (uv) vanishes to imply (iii). Given (iii), we calculate
kvk2kuk2 = (kvk2uu) = ([Re(uv)]vu) = [Re(uv)](vu)
kvk2u = (uv)v,
(A.15)
(A.16)
and take its real part to obtain (i). This proves the equivalence among (i) -- (iii). A similar argument
shows the equivalence among (i), (ii), and (iv), or it follows by exchanging u and v in the preceding
argument.
(4) Part (4) follows from Part (3) by replacing v by iv.
Acknowledgments
This work is partially supported by the Top Global University Project from the Ministry of Education,
Culture, Sports, Science and Technology (MEXT), Japan. TO is supported by a Grant-in-Aid for Scientific
Research (A) (No. 26247014) from Japan Society for the Promotion of Science (JSPS). KY is supported by
a Grant-in-Aid for Scientific Research (C) (No. 26400406) from JSPS and by Waseda University Grants for
Special Research Projects (No. 2015K-202 and No. 2016K-215).
References
[1] J. M. Aldaz, A stability version of Holder's inequality, J. Math. Anal. Appl. 343, 842 (2008).
[2] J. M. Aldaz, Strengthened Cauchy-Schwarz and Holder inequalities, J. Inequal. Pure Appl. Math. 10, 116 (2009).
[3] C. Bastos, A. E. Bernardini, O. Bertolami, N. Costa Dias, and J. N. Prata, Robertson-Schrodinger-type formulation of
Ozawa's noise-disturbance uncertainty principle, Phys. Rev. A 89, 042112 (2014).
[4] C. Branciard, Error-tradeoff and error-disturbance relations for incompatible quantum measurements, Proc. Natl. Acad.
Sci. USA 110, 6742 (2013).
[5] P. Busch, P. Lahti, and R. F. Werner, Quantum root-mean-square error and measurement uncertainty relations, Rev.
Mod. Phys. 86, 1261 (2014).
[6] P. Dang, G.-T. Deng, and T. Qian, A sharper uncertainty principle, J. Funct. Anal. 265, 2239 (2013).
[7] J. J. Duistermaat and J. A. C. Kolk, Multidimensional Real Analysis II (Cambridge University Press, Cambridge, 2004).
[8] G. B. Folland and A. Sitaram, The uncertainty principle: A mathematical survey, J. Fourier Anal. Appl. 3, 207 (1997).
[9] K. Fujiwara and T. Ozawa, Stability of the Young and Holder inequalities, J. Inequal. Appl. 2014, 162 (2014).
[10] S. Furuichi and K. Yanagi, Schrodinger uncertainty relation, Wigner-Yanase-Dyson skew information and metric adjusted
correlation measure, J. Math. Anal. Appl. 388, 1147 (2012).
[11] S. J. Gustafson and I. M. Sigal, Mathematical Concepts of Quantum Mechanics, 2nd ed. (Springer, Berlin, 2011).
[12] M. Hayashi, S. Ishizaka, A. Kawachi, G. Kimura, and T. Ogawa, Introduction to Quantum Information Science (Springer,
Berlin, 2015).
[13] A. S. Holevo, Quantum Systems, Channels, Information: A Mathematical Introduction (Walter de Gruyter, Berlin, 2012).
13
[14] F. Kaneda, S.-Y. Baek, M. Ozawa, and K. Edamatsu, Experimental test of error-disturbance uncertainty relations by
weak measurement, Phys. Rev. Lett. 112, 020402 (2014).
[15] C. K. Ko and H. J. Yoo, Schrodinger uncertainty relation and convexity for the monotone pair skew information, Tohoku
Math. J. 66, 107 (2014).
[16] K. Korzekwa, D. Jennings, and T. Rudolph, Operational constraints on state-dependent formulations of quantum error-
disturbance trade-off relations, Phys. Rev. A 89, 052108 (2014).
[17] H. Kosaki, On intersections of domains of unbounded positive operators, Kyushu J. Math. 60, 3 (2006).
[18] S. Machihara, T. Ozawa, and H. Wadade, On the Hardy type inequalities, submitted.
[19] E. Mourre, Absence of singular continuous spectrum for certain self-adjoint operators, Commun. Math. Phys. 78, 391
(1981).
[20] M. Ozawa, Universally valid reformulation of the Heisenberg uncertainty principle on noise and disturbance in measure-
ment, Phys. Rev. A 67, 042105 (2003).
[21] M. Ozawa, Uncertainty relations for noise and disturbance in generalized quantum measurements, Ann. Phys. (N.Y.) 311,
350 (2004).
[22] T. Ozawa and H. Sasaki, Inequalities associated with dilations, Commun. Contemp. Math. 11, 265 (2009).
[23] M. Reed and B. Simon, Methods of Modern Mathematical Physics II: Fourier Analysis, Self-Adjointness (Academic Press,
San Diego, 1975).
[24] M. Ringbauer, D. N. Biggerstaff, M. A. Broome, A. Fedrizzi, C. Branciard, and A. G. White, Experimental joint quantum
measurements with minimum uncertainty, Phys. Rev. Lett. 112, 020401 (2014).
[25] M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University Press, Cambridge, 1997).
[26] A. Tawfik and A. Diab, Generalized uncertainty principle: Approaches and applications, Int. J. Mod. Phys. D 23, 1430025
(2014).
[27] D. F. Walls and G. J. Milburn, Quantum Optics, 2nd ed. (Springer, Berlin, 2008).
[28] Y. Watanabe, Formulation of Uncertainty Relation Between Error and Disturbance in Quantum Measurement by Using
Quantum Estimation Theory (Springer, Tokyo, 2014).
14
|
1207.0086 | 2 | 1207 | 2013-07-22T15:43:26 | Semispectral Measures and Feller markov Kernels | [
"math.FA"
] | We give a characterization of commutative semispectral measures by means of Feller and Strong Feller Markov kernels. In particular:
{itemize} we show that a semispectral measure $F$ is commutative if and only if there exist a self-adjoint operator $A$ and a Markov kernel $\mu_{(\cdot)}(\cdot):\Gamma\times\mathcal{B}(\mathbb{R})\to[0,1]$, $\Gamma\subset\sigma(A)$, $E(\Gamma)=\mathbf{1}$, such that $$F(\Delta)=\int_{\Gamma}\mu_{\Delta}(\lambda)\,dE_{\lambda},$$ \noindent and $\mu_{(\Delta)}$ is continuous for each $\Delta\in R$ where, $R\subset\mathcal{B}(\mathbb{R})$ is a ring which generates the Borel $\sigma$-algebra of the reals $\mathcal{B}(\mathbb{R})$. Moreover, $\mu_{(\cdot)}(\cdot)$ is a Feller Markov kernel and separates the points of $\Gamma$. we prove that $F$ admits a strong Feller Markov kernel $\mu_{(\cdot)}(\cdot)$, if and only if $F$ is uniformly continuous. Finally, we prove that if $F$ is absolutely continuous with respect to a regular finite measure $\nu$ then, it admits a strong Feller Markov kernel. {itemize} The mathematical and physical relevance of the results is discussed giving a particular emphasis to the connections between $\mu$ and the imprecision of the measurement apparatus. | math.FA | math |
Semispectral Measures and Feller Markov Kernels
Roberto Beneduci∗
Dipartimento di Matematica,
Universit`a della Calabria,
and
Istituto Nazionale di Fisica Nucleare, Gruppo c. Cosenza,
Abstract
We give a characterization of commutative semispectral measures by
means of Feller and Strong Feller Markov kernels. In particular:
• we show that a semispectral measure F is commutative if and only
if there exist a self-adjoint operator A and a Markov kernel µ(·)(·) :
Γ × B(R) → [0, 1], Γ ⊂ σ(A), E(Γ) = 1, such that
F (∆) =ZΓ
µ∆(λ) dEλ,
and µ(∆) is continuous for each ∆ ∈ R where, R ⊂ B(R) is a ring
which generates the Borel σ-algebra of the reals B(R). Moreover,
µ(·)(·) is a Feller Markov kernel and separates the points of Γ.
• we prove that F admits a strong Feller Markov kernel µ(·)(·), if
and only if F is uniformly continuous. Finally, we prove that if F
is absolutely continuous with respect to a regular finite measure ν
then, it admits a strong Feller Markov kernel.
The mathematical and physical relevance of the results is discussed giving
a particular emphasis to the connections between µ and the imprecision
of the measurement apparatus.
1
Introduction
A real semispectral measure (or Positive operator Valued measure) is a map
F : B(R) → L+
s (H) from the Borel σ-algebra of the reals to the space of posi-
tive self-adjoint operators on a Hilbert space H. If, F (∆) is a projection opera-
tor for each ∆ ∈ B(R), F is called spectral measure (or Projection Valued mea-
sure). Therefore, the set of spectral measures is a subset of the set of semispec-
tral measures. Moreover, spectral measures are in one-to-one correspondence
∗e-mail [email protected]
1
with self-adjoint operators (spectral theorem) [41] and are used in standard
quantum mechanics to represent quantum observables.
It was pointed out
[1, 20, 21, 31, 40, 43] that semispectral measures are more suitable than spec-
tral measures in representing quantum observables. The quantum observables
described by semispectral measures are called generalized observables or un-
sharp observables and play a key role in quantum information theory, quantum
optics, quantum estimation theory [20, 28, 31, 44] and in the phase space for-
mulation of quantum mechanics [44, 15, 16]. It is then natural to ask what are
the relationships between semispectral and spectral measures. A clear answer
can be given in the commutative case [1, 6, 7, 8, 9, 10, 11, 12, 13, 14, 30, 32].
Indeed [7, 32], a real positive semispectral measure F is commutative if and
only if there exist a bounded self-adjoint operator A and a Markov kernel
(transition probability) µ(·)(·) : σ(A) × B(R) → [0, 1] such that
F (∆) =Zσ(A)
µ∆(λ) dEλ
where, E is the spectral measure corresponding to A. In other words, F is a
smearing of the spectral measure E corresponding to A.
As an example we can consider the following unsharp position observable
hψ, Qf (∆)ψi :=Z[0,1]
µ∆(x) :=ZR
µ∆(x) dhψ, Qxψi, ∆ ∈ B(R), ψ ∈ L2([0, 1]),
(1)
χ∆(x − y) f (y) dy,
x ∈ [0, 1]
where, f is a positive, bounded, Borel function such that f (y) = 0, y /∈ [0, 1],
R[0,1] f (y)dy = 1, and Qx is the spectral measure corresponding to the position
operator
Q : L2([0, 1]) → L2([0, 1])
ψ(x) 7→ Qψ := xψ(x)
We recall that hψ, Q(∆)ψi is interpreted as the probability that a perfectly
accurate measurement (sharp measurement) of the position gives a result in ∆.
Then, a possible interpretation of equation (1) is that Qf is a randomization
of Q.
Indeed [40], the outcomes of the measurement of the position of a
particle depend on the measurement imprecision1 so that, if the sharp value
of the outcome of the measurement of Q is x then the apparatus produces
with probability µ∆(λ) a reading in ∆.
It is worth noticing that (see example 5.6 in section 5) the Markov kernel
µ∆(x) :=ZR
χ∆(x − y) f (y) dy,
x ∈ [0, 1]
1There are other possible interpretations of the randomization. For example, it could be
due to the existence of a no-detection probability depending on hidden variables [24].
2
in equation (1) above is such that the function x 7→ µ∆(x) is continuous for
each ∆ ∈ B(R). The continuity of µ∆ means that if two sharp values x and x′
are very close to each other then, the corresponding random diffusions are very
similar, i.e., the probability to get a result in ∆ if the sharp value is x is very
close to the probability to get a result in ∆ if the sharp value is x′. That is
quite common in important physical applications and seems to be reasonable
from the physical viewpoint. It is then natural to look for general conditions
which ensure the continuity of λ 7→ µ∆. That is one of the aims of the present
work. What we prove is that, in general, the continuity does not hold for all
the Borel sets ∆ but only for a ring of subsets which generates the Borel σ-
algebra of the reals. (Anyway, that is sufficient to prove the weak convergence
of µ(·)(x) to µ(·)(x′).) We also prove that the continuity for each Borel set is
equivalent to the uniform continuity of F which in its turn is equivalent to
require that the smearing in equation (1) can be realized by a strong Feller
Markov kernel.
It is our opinion that the continuity of µ∆ over a ring R which generates
the Borel σ-algebra of the reals could be helpful in dealing with problems
connected to the characterization of functions of the kind
Gf (x) =Z f (t) dµt(x).
A similar (but less general) problem arises in Ref. [12] where the relationships
between Naimark extension theorem and the characterization of commutative
semispectral measures as smearing of spectral measures are analyzed. That is
a second motivation for the analysis of the continuity properties of µ∆.
The results outlined above are contained in the two main theorems of the
present work.
The first is a stronger characterization of commutative semispectral measures.
In particular, we show (see theorems 4.3) that a semispectral measure is com-
mutative if and only if there exist a spectral measure E and a Markov kernel
µ(·)(·) : Γ × B(R) → [0, 1], Γ ⊂ σ(A), E(Γ) = 1, such that
F (∆) =ZΓ
µ∆(λ) dEλ
(2)
and µ∆(·) is continuous for each ∆ ∈ R where, R ⊂ B(R) is a ring which
generates the Borel σ-algebra of the reals B(R). It turns out that µ(·)(·) : Γ ×
B(R) → [0, 1] is a Feller Markov kernel [38, 42]. Therefore, F is commutative
if and only if there exists a Feller Markov kernel µ such that equation (2) is
satisfied.
We also prove that the family of functions {µ∆}∆∈B(R) separates the points
of σ(A) up to a null set (see theorems 3.1, and 4.3).
In other words, the
probability measures µ(·)(x) and µ(·)(x′) which represent the randomizations
corresponding to the sharp values x and x′ are different.
3
The second theorem is a characterization of the semispectral measures
which admit a strong Feller Markov kernel, i.e., a Markov kernel µ such that
the function λ 7→ µ∆(λ) is continuous for each ∆ ∈ B(R). In particular, we
prove (see theorem 5.5) that a semispectral measure F admits a strong Feller
Markov kernel if and only if it is uniformly continuous. As an example, we
develop the details for the unsharp position observable defined in equation (1)
above. Finally, we prove (see section 6) that a semispectral measure F which is
absolutely continuous with respect to a regular finite measure ν is uniformly
continuous (theorem 6.2). We give some examples of absolutely continuous
semispectral measures (see example 6.4) and analyze the unsharp position
observable which is obtained as the marginal of a phase space observable (see
section 6.1).
2 Some preliminaries about Semispectral measures
In what follows, we denote by B(R) and B([0, 1]) the Borel σ-algebra of R and
[0,1] respectively, by 0 and 1 the null and the identity operators, by Ls(H)
the space of all bounded self-adjoint linear operators acting in a Hilbert space
H with scalar product h·,·i, by F(H) = L+
s (H) the subspace of all positive,
bounded self-adjoint operators on H, by E(H) ⊂ F(H) the subspace of all
projection operators on H. We use the symbols POVM and PVM to denote
semispectral measures and spectral measures respectively.
Definition 2.1. A Semispectral measure or Positive Operator Valued measure
(for short, POVM) is a map F : B(R) → F(H) such that:
F(cid:0) ∞[n=1
∆n(cid:1) =
∞Xn=1
F (∆n).
where, {∆n} is a countable family of disjoint sets in B(R) and the series
converges in the weak operator topology. It is said to be normalized if
F (R) = 1
Definition 2.2. A POVM is said to be commutative if
(cid:2)F (∆1), F (∆2)(cid:3) = 0, ∀ ∆1 , ∆2 ∈ B(R).
Definition 2.3. A POVM is said to be orthogonal if
F (∆1)F (∆2) = 0 if ∆1 ∩ ∆2 = ∅.
(3)
(4)
Definition 2.4. A Spectral measure or Projection Valued measure (for short,
PVM) is an orthogonal, normalized POVM.
4
It is simple to see that for a PVM E, we have E(∆) = E(∆)2, for any ∆ ∈
B(R). Then, E(∆) is a projection operator for every ∆ ∈ B(R), and the PVM
is a map E : B(R) → E(H).
In quantum mechanics, non-orthogonal normalized POVM are also called gen-
eralised or unsharp observables and PVM standard or sharp observables.
In what follows, we shall always refer to real normalized POVM and we shall
use the term "measurable" for the Borel measurable functions. For any vector
x ∈ H the map
hF (·)x, xi : B(R) → R,
∆ 7→ hF (∆)x, xi,
is a Lebesgue-Stieltjes measure. There exists a one-to-one correspondence [5]
between POV measures F and POV functions Fλ := F ((−∞, λ]).
In the
following we will use the symbol dhFλx, xi to mean integration with respect to
the measure hF (·)x, xi. We shall say that a measurable function f : N ⊂ R →
f (N ) ⊂ R is almost everywhere (a.e.) one-to-one with respect to a POVM F
if it is one-to-one on a subset N ′ ⊂ N such that N − N ′ is a null set with
respect to F . We shall say that a function f : R → R is bounded with respect
to a POVM F , if it is equal to a bounded function g a.e. with respect to F ,
that is, if f = g a.e. with respect to the measure hF (·)x, xi, ∀x ∈ H. For
any real, bounded and measurable function f and for any POVM F , there is
a unique [18] bounded self-adjoint operator B ∈ Ls(H) such that
hBx, xi =Z f (λ)dhFλx, xi,
for each x ∈ H.
(5)
If equation (5) is satisfied, we write B = R f (λ)dFλ or B = R f (λ)F (dλ)
equivalently.
Definition 2.5. The spectrum σ(F ) of a POVM F is the closed set
(cid:8)λ ∈ R : F(cid:0)(λ − δ, λ + δ)(cid:1) 6= 0, ∀δ > 0, (cid:9) .
By the spectral theorem [23, 41], there is a one-to-one correspondence between
PV measures E and self-adjoint operators B, the correspondence being given
by
Notice that the spectrum of EB coincides with the spectrum of the corre-
sponding self-adjoint operator B. Moreover, in this case a functional calculus
can be developed. Indeed, if f : R → R is a measurable real-valued function,
we can define the self-adjoint operator [41]
B =Z λdEB
λ .
f (B) =Z f (λ)dEB
λ
5
where, EB is the PVM corresponding to B. If f is bounded, then f (B) is
bounded [41].
In the following we do not distinguish between PVM and the corresponding
self-adjoint operators.
Let Λ be a subset of R and B(Λ) the corresponding Borel σ-algebra.
Definition 2.6. A real Markov kernel is a map µ : Λ × B(R) → [0, 1] such
that,
1. µ∆(·) is a measurable function for each ∆ ∈ B(R),
2. µ(·)(λ) is a probability measure for each λ ∈ Λ.
Definition 2.7. Let ν be a measure on Λ. A map µ : Λ × B(R) → [0, 1] is a
weak Markov kernel with respect to ν if:
1. µ∆(·) is a measurable function for each ∆ ∈ B(R),
2. 0 ≤ µR(λ) ≤ 1,
3. µR(λ) = 1, µ∅(λ) = 0,
ν − a.e.,
ν − a.e.,
4. for any sequence {∆i}i∈N, ∆i ∩ ∆j = ∅,
Xi
µ(∆i)(λ) = µ(∪i∆i)(λ),
ν − a.e.
Definition 2.8. The map µ : Λ × B(R) → [0, 1] is a weak Markov kernel with
respect to a PVM E : B(Λ) → E(H) if it is a weak Markov kernel with respect
to each measure νx(·) := hE(·) x, xi, x ∈ H.
In the following, by a weak Markov kernel µ we mean a weak Markov kernel
with respect to a PVM E. Moreover the function λ 7→ µ∆(λ) will be denoted
indifferently by µ∆ or µ∆(·).
Definition 2.9. A POV measure F : B(R) → F(H) is said to be a smearing
of a POV measure E : B(Λ) → E(H) if there exists a weak Markov kernel
µ : Λ × B(R) → [0, 1] such that,
F (∆) =ZΛ
µ∆(λ)dEλ, ∆ ∈ B(R).
Example 2.10. In the standard formulation of quantum mechanics, the op-
erator
Q : L2(R) → L2(R)
ψ(x) ∈ L2(R) 7→ Qψ := xψ(x)
6
is used to represent the position observable. A more realistic description of the
position observable of a quantum particle is given by a smearing of Q as, for
example, the optimal position semispectral measure
F Q(∆) =
where,
1
l √2 πZ ∞
−∞(cid:16)Z∆
e− (x−y)2
2 l2 dy(cid:17) dEQ
x =Z ∞
−∞
µ∆(x) dEQ
x
µ∆(x) =
1
l √2 πZ∆
e− (x−y)2
2 l2 dy
defines a Markov kernel and EQ is the spectral measure corresponding to the
position operator Q.
In the following, the symbol µ is used to denote both Markov kernels and
weak Markov kernels. The symbols A and B are used to denote self-adjoint
operators.
Definition 2.11. Whenever F , A, and µ are such that F (∆) = µ∆(A), ∆ ∈
B(R), we say that (F, A, µ) is a von Neumann triplet.
The following theorem establishes a relationship between commutative semis-
pectral measures and spectral measures and gives a characterization of the
former. Other characterizations and an analysis of the relationships between
them can be found in Ref.s [1, 30, 4, 33].
Theorem 2.12 ([7, 32]). A semispectral measure F is commutative if and
only if there exist a bounded self-adjoint operator A and a Markov kernel (weak
Markov kernel) µ such that (F, A, µ) is a von Neumann triplet.
Corollary 2.13. A semispectral measure F is commutative if and only if it
is a smearing of a PV measure E with bounded spectrum.
Definition 2.14. The von Neumann algebra generated by the semispectral
measure F is the von Neumann algebra generated by the set {F (∆), ∆ ∈
B(R)}.
Definition 2.15. If A and F in theorem 2.12 generate the same von Neumann
algebra then A is named the sharp version of F .
Theorem 2.16. [7] The sharp version A is unique up to almost everywhere
bijections.
3 On the separation properties of µ
In the following, the symbol S denotes the family of open intervals in R with
rational end-points. The symbol R(S) denotes the ring generated by S. Notice
7
that S is countable. Then, by theorem c, page 24, in Ref.
countable too. Moreover, R(S) generates the Borel σ-algebra B(R).
A weak Markov kernel µ such that (F, A, µ) is a von Neumann triplet,
separates the point of Γ ⊂ σ(A) if the family of functions {µ∆}∆∈B(R) separates
the points of Γ or, in other words, if the set functions {µ(·)(λ)}λ∈Γ are distinct.
It is then natural to ask if in general µ has that property. The following
theorem answers in the positive.
[27], R(S) is
Theorem 3.1. Let (F, A, µ) be a von Neumann triplet and suppose that A is
a sharp version of F . Then, there exists a set Γ ⊆ σ(A), EA(Γ) = 1, such
that the family of functions {µ∆(·)}∆∈B(R) separates the points of Γ.
Proof. In the following, AW (F ) denotes the von Neumann algebra generated
by {F (∆)}∆∈B(R), O2 := {F (∆)}∆∈R(S) and AC(O2) is the C ∗-algebra gener-
ated by O2. The von Neumann algebra generated by AC(O2) coincides with
AW (F ) (see appendix A). Moreover, AW (F ) = AW (A) since A is the sharp
version of F and generates AW (F ). By the Gelfand-Naimark theorem [23, 39],
there is a * isomorphism φ between AC(O2) and the algebra of continuous
functions C(Λ2) where Λ2 is the spectrum of AC(O2). Moreover,
f ∈ C(Λ2) 7→ φ(f ) =ZΛ2
f (λ) deEλ
where, eE is the spectral measure from the Borel σ algebra B(Λ2) to E(H) whose
existence is assured by theorem 1, page 895, in Ref. [23]. The Gelfand-Naimark
isomorphism φ can be extended to a homomorphism between the algebra of
the Borel functions on Λ2 and the von Neumann algebra AW (F ) = AW (A)
generated by AC(O2) (see Ref. [22], page 360, section 3). Therefore, there is
a Borel function h such that
A =ZΛ2
h(λ) deEλ
(6)
Let {∆i}i∈N denote an enumeration of the set R(S). Since AC(O2) is the
smallest uniform closed algebra containing {F (∆i)}i∈N, C(Λ2) is the smallest
uniform closed algebra of functions containing {ν∆i := φ−1(F (∆i))}i∈N. In
other words {ν∆i}i∈N generates C(Λ2). The Stone-Weierstrass theorem [23]
assures that {ν∆i}i∈N separates the points in Λ2.
On the other hand, the fact that (F, A, µ) is a von Neumann triplet, implies
that, for each ∆i ∈ R(S), there is a Borel function µ∆i such that
ZΛ2
ν∆i(λ) deEλ = F (∆i) = µ∆i(A) =ZΛ2
µ∆i(h(λ)) deEλ.
Then, for each ∆i ∈ R(S), there is a set Mi ⊂ Λ2, eE(Mi) = 1, such that
µ∆i(h(λ)) = ν∆i(λ),
(7)
λ ∈ Mi.
8
Let M := ∩∞
i=1Mi. Then,
eE(M ) = lim
n→∞eE(∩n
i=1Mi) = lim
n→∞
and, for each i ∈ N,
nYi=1 eE(Mi) = 1
(µ∆i ◦ h)(λ) = ν∆i(λ),
λ ∈ M ⊆ Λ2.
(8)
Since {ν∆i}i∈N separates the points in Λ2, it separates the points in M . Then,
equation (8) implies that {µ∆i}i∈N separates the points in Γ := h(M ). More-
over2,
where, EA is the spectral measure defined by the relation
EA(Γ) = EA(h(M )) = eE[h−1(h(M ))] = 1
and such that,
EA(∆) = eE(h−1(∆))
A =Z x dEA
x
while, h−1(h(M )) is a Borel set containing M .
We have proved that the set of functions {µ∆i}i∈N separates the points of Γ
and that EA(Γ) = 1. In other words,
µ(·)(λ) 6= µ(·)(λ′),
λ 6= λ′, λ, λ′ ∈ Γ.
4 Characterization of Commutative Semi-spectral
Measures by means of Feller Markov kernels
As we have seen in the last section, theorem 2.12 asserts that a semispectral
measure F is commutative if and only if there exist a bounded self-adjoint
operator A and a weak Markov kernel (Markov kernel) µ such that F (∆) =
µ∆(A). In the present section we study the continuity of the functions µ∆.
In particular, we introduce the concept of strong Markov kernel, i.e., a weak
Markov kernel µ(·)(·) : Λ × B(R) → [0, 1] with respect to a PVM E : B(Λ) →
2 Notice that h(M ) is a Borel set. In order to prove that, we first recall that Λ2 is a
Polish space (that is, a complete, separable, space [35]). Indeed, by theorem 11, page 871, in
Ref. [23], it is homeomorphic to a closed subspace of the Cartesian product Q∞
σ(F (∆i)),
i=1
where σ(F (∆i)) is a complete separable metric space, and by theorem 2, page 406, and
theorem 6, page 156, in Ref. [36], it is complete and separable. Moreover, h is measurable
and injective on M . Therefore, Soulsin's theorem (see theorem 9 page 440 and Corollary 1
page 442 in Ref. [35]) assures that h(M ) is a Borel set.
9
E(H) such that µ(·)(λ) is a probability measure for each λ ∈ Γ ⊂ Λ, E(Γ) = 1.
Then, we prove (theorem 4.3) that in order to realize the smearing in corollary
2.13, one can use a strong Markov kernel µ such that µ∆ is continuous for each
∆ ∈ R, where R is a ring which generates the Borel σ-algebra of the reals. It
is worth remarking that µ(·)(·) : Γ × B(R) → [0, 1] is a Feller Markov kernel.
Therefore, F is commutative if and only if there exists a bounded self-adjoint
operator A and a Feller Markov kernel µ such that
F (∆) =ZΓ
µ∆(λ) dEλ.
Moreover, the family of functions {µ∆}∆∈R separates the points in Γ (see
theorems 3.1 and 4.3).
In order to prove the main theorem we need the following definitions.
Definition 4.1. Let E : B(Λ) → E(H) be a PVM. The map µ(·)(·) : Λ ×
B(R) → [0, 1] is a strong Markov kernel with respect to E if it is a weak
Markov kernel and there exists a set Γ ⊂ Λ, E(Γ) = 1, such that µ(·)(·) :
Γ × B(R) → [0, 1] is a Markov kernel with respect to E. A strong Markov
kernel is denoted by the symbol (µ, E, Γ ⊂ Λ).
Definition 4.2. A Feller Markov kernel is a Markov kernel µ(·)(·) : Λ ×
B(R) → [0, 1] such that the function
G(λ) =ZΛ
f (t) dµt(λ),
λ ∈ Λ
is continuous and bounded whenever f is continuous and bounded.
Theorem 4.3. A real POVM F : B(R) → F(H) is commutative if and only
if, there exists a bounded self-adjoint operator A = R λ dEλ with spectrum
σ(A) ⊂ [0, 1] and a strong Markov Kernel (µ, E, Γ ⊂ σ(A)) such that:
1) µ∆(·) : σ(A) → [0, 1] is continuous for each ∆ ∈ R(S),
2) F (∆) =RΓ µ∆(λ) dEλ, ∆ ∈ B(R).
3) µ separates the points in Γ.
Moreover, µ : Γ × B(R) → [0, 1] is a Feller Markov kernel.
Proof. Let AW (F ) be the von Neumann algebra generated by F . AW (F )
coincides with the von Neumann algebra generated by {F (∆)}∆∈R(S) where,
R(S) ⊂ B(R) is the ring generated by the family S of open intervals with
rational end-points (see appendix A for the proof). We recall that both S and
R(S) are countable (see theorem c, page 24, in Ref. [27]).
Now, we proceed to the proof of the existence of A. Let {∆i}i∈N be an
enumeration of the set R(S) and O2 := {F (∆)}∆∈R(S) . Let E(i) denote the
10
spectral measure corresponding to F (∆i) ∈ O2. We have F (∆i) =R x dE(i)
Therefore, for each i, k ∈ N there exists a division {∆(i,k)
x .
}j=1,...,mi,k of [0, 1]
j
such that
x(i,k)
j E(i)(∆(i,k)
j
mi,kXj=1
(cid:13)(cid:13)
) − F (∆i)(cid:13)(cid:13) ≤
1
k
.
(9)
By the spectral theorem [23] the von Neumann algebra AW (F ) contains all the
projection operators in the spectral resolution of F (∆), ∆ ∈ B(R). Therefore,
the von Neumann algebra AW (D) generated by the set D := {E(i)(∆i,k
j ), j ≤
mi,k, i, k ∈ N} is contained in AW (F ) and then
AW (D) ⊂ AW (F ) = AW (O2).
(10)
Moreover, the C ∗-algebra AC(D) generated by D contains the C ∗-algebra
AC(O2) generated by O2 (see equation (9)). Summing up the preceding ob-
servations, we have
By the double commutant theorem [34],
AC(O2) ⊂ AC(D) ⊂ AW (F ).
AW (F ) = [AC(O2)]′′ ⊂ [AC(D)]′′ = AW (D)
so that (see equation 10),
AW (D) = AW (F ).
i=1{0, 1}. Let π : Λ →Q∞
[23], the spectrum Λ of AC(D) is homeo-
i=1{0, 1} denote the
(11)
By theorem 11, page 871 in Ref.
homeomorphism between the two spaces.
morphic to a closed subset ofQ∞
Now, if we identify Λ with a closed subset of Q∞
π(λ) = ¯x := (x1, . . . , xn, . . . ) ∈Q∞
i=1{0, 1}. The function
i=1{0, 1}, we can prove the
existence of a continuous function distinguishing the points of Λ. Indeed, let
f (λ) =
xi
3i
∞Xi=1
is continuous and injective and then it distinguishes the points of Λ. Moreover,
since Λ and [0, 1] are Hausdorff, the map f : Λ → f (Λ) is a homeomorphism.
By theorem 1, page 895, in Ref.
B(Λ) → F(H) such that the map
[23], there exists a spectral measure eE :
T : C(Λ) → B(H)
g 7→ T (g) =ZΛ
g(λ)deEλ
defines an isometric ∗-isomorphism between AC(D) and C(Λ).
(12)
11
The fact that f distinguishes the points of Λ, implies that the self-adjoint
operator
A =ZΛ
f (λ) deEλ
is a generator of the von Neumann algebra AW (D) = AW (F ).
Indeed, by
the Stone-Weierstrass theorem, C(Λ) is singly generated, in particular f is a
generator. Then, the isomorphism between AC(D) and C(Λ) assures that
AC(D) is singly generated and that A is a generator. Hence, AW (F ) =
AW (D) = [AC(D)]′′ is singly generated. In particular, A generates AW (F ),
i.e., AW (F ) = AW (A).
Now, we proceed to the proof of the existence of the weak Markov kernel
By (12), for each ∆ ∈ R(S), there exists a continuous function γ∆ ∈ C(Λ)
such that
eν such that (F, A,eν) is a von Neumann triplet.
γ∆(λ) deEλ.
F (∆) =ZΛ
Now, we show that, for each ∆ ∈ R(S), there is a continuous function ν∆ :
σ(A) → [0, 1] from the spectrum of A to the interval [0, 1] such that ν∆(f (λ)) =
γ∆(λ), λ ∈ Λ, and F (∆) = ν∆(A).
To prove this, let us consider the function
ν∆(t) := (γ∆ ◦ f −1)(t), ∆ ∈ R(S).
It is continuous since it is the composition of continuous functions and,
ν∆(f (λ)) = γ∆(f −1(f (λ))) = γ∆(λ).
Moreover,
Indeed, by the change of measure principle (page 894, ref. [23]),
ν∆(A) = F (∆),
∀∆ ∈ R(S).
γ∆(λ) deEλ =ZΛ
F (∆) =ZΛ
γ∆(f −1(t)) dEt =Zσ(A)
=Zσ(A)
γ∆(f −1(f (λ))) deEλ
ν∆(t) dEt = ν∆(A)
where σ(A) = f (Λ) is the spectrum of A and E is the spectral measure corre-
sponding to A defined by the relation E(∆) = eE(f −1(∆)), ∆ ∈ B(σ(A)) (see
corollary 10, page 902, in Ref. [23]).
For each λ ∈ σ(A), the map ν(·)(λ) : R(S) → [0, 1] defines an additive set
function. Indeed, let ∆ ∈ R(S) be the disjoint union of the sets ∆1, ∆2 ∈
12
R(S). Then,
Z ν(∆1∪∆2)(λ) dEλ = F (∆1 ∪ ∆2) = F (∆1) + F (∆1)
=Z ν∆1(λ) dEλ +Z ν∆2(λ) dEλ
=Z (cid:2)ν∆1(λ) + ν∆2(λ)(cid:3) dEλ
so that, by the continuity of the functions ν(∆1)(λ) and ν(∆2)(λ), we get (see
theorem 1, page 895, in Ref. [23])
ν(∆1)(λ) + ν(∆2)(λ) = ν(∆1∪∆2)(λ),
∀λ ∈ σ(A).
Now, we extend ν to all the Borel σ-algebra of [0, 1].
Since A is the generator of AW (F ), for each ∆ ∈ B([0, 1]), there exists a Borel
function ω∆ such that.
Then, we can consider the mapeν : σ(A) × B(R) → [0, 1] defined as follows
(ω∆ ◦ f )(λ) deEλ
ω∆(t) dEt =ZΛ
F (∆) =Zσ(A)
eν∆(λ) =(ν∆(λ)
ω∆(λ)
if ∆ ∈ R(S)
if ∆ /∈ R(S).
which is the disjoint union of the sets {∆i}i∈N, ∆i ∈ B(R). Then,
Sinceeν coincides with ν on R(S) it is additive on R(S).
In order to prove thateν is a weak Markov kernel, let us consider a set ∆ ∈ B(R)
Z eν(∪∞
i=1∆i)(x) dEx =Z eν∆(x)dEx = F (∆) =
∞Xi=1
F (∆i)
=
∞Xi=1Z eν∆i(x) dEx =Z ∞Xi=1eν∆i(x) dEx
so that, by Corollary 9, page 900, in Ref. [23],
∞Xi=1eν∆i(x) =eν∆(x), E − a.e,
Now, we proceed to prove the existence of the Markov kernel µ : Γ×B(R) →
which implies that eν : σ(A) × B(R) → [0, 1] is a weak Markov kernel.
particular (F, A,eν) is a von Neumann triplet.
[7], starting from eν : σ(A) × R(S) → [0, 1] it is
possible to define a Markov kernel ω : σ(A)×B(R) → [0, 1] such that (F, A, ω)
[0, 1] such that items 1, 2, and 3 of the theorem are satisfied.
By corollary 1 in Ref.
In
13
is a von Neumann triplet. Since (F, A,eν) is a von Neumann triplet, for each
∆ ∈ B(R),
hence,
(13)
Now, let {∆i}i∈N be an enumeration of R(S). By equation (13), for each
i ∈ N, there is a set Ni ⊂ σ(A), E(Ni) = 0, such that
Z eν∆(λ) dEλ = F (∆) =Z ω∆(λ) dEλ
ω∆(λ) =eν∆(λ), E − a.e.
ω∆i(λ) =eν∆i(λ),
ω∆i(λ) =eν∆i(λ),
N := ∪∞
λ ∈ σ(A) − Ni.
λ ∈ σ(A) − N
i=1Ni, E(N ) = 0.
Then, for each i ∈ N,
where,
(14)
(15)
Therefore, for almost all λ ∈ σ(A), eν(·)(λ) is σ-additive on R(S).
Now, we can define the map
µ(·)(λ) =(eν(·)(λ) λ ∈ N
ω(·)(λ) λ ∈ σ(A) − N
If we put Γ = σ(A) − N , we have that µ(·)(·) : Γ × B(R) → [0, 1] is a Markov
kernel. Therefore, µ(·)(·) : σ(A) × B(R) → [0, 1] is a strong Markov kernel.
Notice that, for each ∆ ∈ R(S) and λ ∈ σ(A),
so that, µ∆ is continuous for each ∆ ∈ R(S) and additive on R(S). We also
have,
µ∆(λ) =eν∆(λ)
µ∆(A) = ω∆(A) = F (∆), ∆ ∈ R(S).
We have proved items 1, 2, and 3. Item 4 comes from theorem 3.1.
It remains to prove that µ is a Feller Markov kernel. By item 1, µ∆ is
continuous for each ∆ ∈ R(S). Notice that for each open set O ∈ B(R), there
is a countable family of sets ∆i ∈ R(S) such that O = ∪∞
i=1∆i. Therefore, by
theorem 2.2 in Ref. [19], µ(·)(λn) converges weakly to µ(·)(λ), i.e.,
n→∞Z f (t) µt(λn) =Z f (t) µt(λ),
lim
f ∈ Cb(R)
whenever limn→∞ λn = λ and Cb(R) is the space of bounded, continuous func-
tions.
Since F (∆) = µ∆(A) implies the commutativity of F , the theorem is
proved.
14
5 Characterization of Semi-spectral Measures which
admit strong Feller Markov Kernels
In the last section we proved that each commutative semispectral measure
admits a strong Markov kernel µ such that µ∆ is a continuous function for
each ∆ ∈ R(S) where, R(S) is a ring which generates the Borel σ-algebra
B(R).
In the present section we characterize the commutative semispectral measures
for which the Markov kernel µ, whose existence was proved in theorem 2.12, is
such that µ∆ is continuous for each ∆ ∈ B(R). Whenever such a Markov kernel
exists, we say that the semispectral measure admits a strong Feller Markov
kernel. In particular, we prove that a commutative semispectral measure F
admits a strong Feller Markov kernel if and only if F is uniformly continuous.
Definition 5.1. Let F : B(R) → F(H). Let ∆ = ∪∞
i=1∆i, ∆i ∩ ∆j = ∅. If
lim
n→∞
nXi=1
F (∆i) = F (∆)
in the uniform operator topology then we say that F is uniformly continuous.
Notice that the term uniformly continuous derives from the fact that the σ-
additivity of F in the uniform operator topology is equivalent to the continuity
in the uniform operator topology. Analogously, the σ-additivity of F in the
weak operator topology is equivalent to the continuity of F in the weak oper-
ator topology [18].
Definition 5.2. A Markov kernel µ(·)(·) : [0, 1] × B(R) → [0, 1] is said to be
strong Feller if µ∆ is a continuous function for each ∆ ∈ B(R).
Definition 5.3. We say that a commutative POVM admits a strong Feller
Markov kernel if there exists a strong Feller Markov kernel µ such that F (∆) =
R µ∆(λ) dEλ, where E is the sharp reconstruction of F .
In order to prove the main theorem of the section we need the following lemma.
Lemma 5.4. Let F be uniformly continuous. Let µ be a weak Markov kernel
and (F, A, µ) a von Neumann triplet. Suppose that µ∆ is continuous for each
∆ ∈ R(S). Then, for each λ ∈ σ(A), µ(·)(λ) is σ-additive on R(S).
Proof. Let ∆, ∆i ∈ R(S), ∆i ∩ ∆j = ∅, ∪∞
i=1∆i)(cid:1) = u − lim
n→∞Z (cid:0)µ∆(λ) −
n→∞(cid:0)F (∆) − F (∪n
nXi=1
µ∆i(λ)(cid:1) dEλ.
0 = u − lim
i=1∆i = ∆. Then,
15
By the uniform continuity of F and theorem 1, page 895, in Ref. [23], it follows
that, ∀ǫ > 0, there exists a number ¯n ∈ N, such that n > ¯n implies,
kµ∆(λ) −
nXi=1
By equation (16),
µ∆i(λ)k∞ = kZ (cid:0)µ∆(λ) −
nXi=1
µ∆i(λ)(cid:1) dEλk
= kF (∆) − F (∪n
i=1∆i)k ≤ ǫ.
(16)
µ∆(λ) −
nXi=1
µ∆i(λ) ≤ ǫ, ∀λ ∈ σ(A).
Theorem 5.5. A commutative POVM F : B(R) → F(H) admits a strong
Feller Markov kernel if and only if it is uniformly continuous.
steps.
Proof. Suppose F uniformly continuous. By theorem 4.3, there is a weak
Markov kernel µ : σ(A)×B(R) → [0, 1] such that µ∆(·) is continuous for every
∆ ∈ R(S) and a self-adjoint operator A such that (F, A, µ) is a von Neumann
triplet. By lemma 5.4, µ is σ-additive on R(S). Therefore (see proposition 2
in Ref. [7]), the map µ : σ(A) × R(S) → [0, 1] can be extended to a Markov
kernel eµ : σ(A) × B(R) → [0, 1] whose restriction to R(S) coincides with µ
and such that F (∆) =eµ∆(A).
Now we prove that eµ∆ is continuous for each ∆ ∈ B(R). We proceed by
1) eµ is continuous for each open interval. For each open interval, there
exists an increasing family of sets ∆i ∈ S such that ∆i ↑ ∆. Indeed, if ∆ =
(a, b), a, b ∈ R, the family of sets {(ai, bi) ∈ S}i∈N such that ai > ai+1 > a,
limi→∞ ai = a, bi < bi+1, limi→∞ bi = b, is increasing and ∪∞
i=1∆i = ∆. Then,
i→∞Z eµ∆i(λ) dEλ.
Z eµ∆(λ) dEλ = F (∆) = u − lim
keµ∆n(λ) −eµ∆m(λ)k∞ = kZ [eµ∆n(λ) −eµ∆m(λ)] dEλk
By the uniform continuity of F , it follows that, ∀ǫ > 0, there exists a number
¯n ∈ N, such that n, m > ¯n implies,
= kF (∆n) − F (∆m)k ≤ ǫ.
F (∆i) = u − lim
(17)
i→∞
By equation (17),
eµ∆n(λ) −eµ∆m(λ) ≤ ǫ, ∀λ ∈ σ(A).
16
(18)
Since eµ is a Markov kernel,
lim
i→∞eµ∆i(λ) =eµ∆(λ),
∀λ ∈ σ(A).
Moreover, by equation (18), the convergence is uniform and this proves the
continuous for each n ∈ N, and
continuity of eµ∆.
2) eµ∆ is continuous for each open set. Each open set ∆ is the disjoint
union of a countable family of open intervals, i.e., ∆ = ∪∞
i=1∆i. Therefore, e∆n ↑ ∆. Moreover, µ e∆n
Let us define the set e∆n := ∪n
F (e∆n) = F (∆).
Then, the same reasoning we used above allows us to conclude that the family
of continuous functions µ e∆n
i=1∆i, ∆i = (ai, bi).
is
converges uniformly to µ∆.
u − lim
i→∞
3) eµ∆ is continuous for each Borel set. Let ∆ ∈ B(R). Since F is regular,
there is a decreasing sequence of open sets Gi, ∆ ⊂ Gi, such that
s − lim
n→∞
F (∩n
i=1Gi) = s − lim
n→∞
F (Gi) = F (∆).
Moreover, by the uniform continuity of F ,
u − lim
n→∞
F (∩n
i=1Gi)) = F (∩∞
i=1Gi).
Therefore, F (∆) = F (∩∞
i=1Gi) and then,
u − lim
n→∞
F (∩n
i=1Gi) = u − lim
n→∞
F (Gi) = F (∆).
Then, the same reasoning we used in steps 1 and 2 allows us to conclude that
the family of continuous functions eµ(∩n
then the continuity of eµ∆.
In order to prove the second part of the theorem we show that the existence
of a strong Feller Markov kernel implies the uniform continuity of F . Suppose
that there exists a strong Feller Markov kernel µ such that F (∆) = µ∆(λ).
Since µ is a Markov kernel it is σ-additive. Then,
i=1∆i) converges uniformly to eµ∆ and
lim
n→∞(cid:0)µ∆(λ) −
where, ∆, ∆i ∈ B([0, 1]), ∪∞
By hypothesis,
i=1∆i = ∆.
nXi=1
µ∆i(λ)(cid:1) = 0,
λ ∈ σ(A).
µ∆(λ) −
nXi=1
µ∆i(λ) ∈ C(σ(A)),
∀n ∈ N.
17
Then, by theorem B1 in appendix B,
u − lim
n→∞(cid:0)µ∆(λ) −
nXi=1
µ∆i(λ)(cid:1) = 0.
nXi=1
By theorem 1, page 895, in Ref. [23], kF (∆)k = kµ∆k∞, hence
n→∞kF (∆) − F (∪n
lim
i=1∆i)k = lim
n→∞kµ∆ −
µ∆ik∞ = 0.
which proves that F is uniformly continuous.
Example 5.6. Let us consider the following unsharp position observable
Qf (∆) :=Z[0,1]
µ∆(x) :=ZR
µ∆(x) dQx, ∆ ∈ B(R),
(19)
χ∆(x − y) f (y) dy,
x ∈ [0, 1]
where, f is a bounded, continuous function such that f (y) = 0, y /∈ [0, 1] and
Z[0,1]
f (y) dy = 1,
and Qx is the spectral measure corresponding to the position operator
Q : L2([0, 1]) → L2([0, 1])
ψ(x) 7→ (Qψ)(x) := xψ(x)
Notice that, for each ∆ ∈ B(R), µ∆ : [0, 1] → [0, 1] is continuous. Indeed,
by the uniform continuity of f , for each ǫ > 0, there is a δ > 0 such that
x − x′ ≤ δ implies f (x − y) − f (x′ − y) ≤ ǫ, for each y. Therefore,
µ∆(x) − µ∆(x′) =(cid:12)(cid:12)(cid:12)ZR
=(cid:12)(cid:12)(cid:12)Z∆
χ∆(x − y) f (y) dy −ZR
χ∆(x′ − y) f (y) dy(cid:12)(cid:12)(cid:12)
[f (x − y) − f (x′ − y)] dy(cid:12)(cid:12)(cid:12) ≤ ǫZ∆∩[−1,1]
By theorem 5.5 and the continuity of µ∆, ∆ ∈ B(R), Qf is uniformly contin-
uous. That can be proved as follows. Suppose ∆i ↓ ∆ and f (y) ≤ M , y ∈ R.
Since, for each x ∈ [0, 1],
dy ≤ 2ǫ
µ∆i−∆(x) =Z∆i−∆
hψ, Qf (∆i − ∆)ψi =Z[0,1]
f (x − y) dy ≤ MZ(∆i−∆)∩[−1,1]
µ∆i−∆(x)ψ2(x) dx ≤ MZ(∆i−∆)∩[−1,1]
we have that, for each ψ ∈ H, ψ2 = 1,
dx
dx
which proves the uniform continuity of Qf .
18
In the case of uniformly continuous POV measures, we can prove a necessary
condition for the norm-1-property which has been recently used in Ref. [17] in
order to study the localization in phase space of massless relativistic particles.
Definition 5.7 ([29]). A semispectral measure F has the norm-1-property if
kF (∆)k = 1, for each ∆ ∈ B(R) such that F (∆) 6= 0.
Theorem 5.8. Let F be uniformly continuous. Then, F has the norm-1-
property only if kF ({λ})k 6= 0 for each λ ∈ σ(F ).
Proof. We proceed by contradiction. Suppose that F has the norm-1 property
and that there exists λ ∈ σ(F ), such that kF ({λ})k = 0. Let (ai, bi) ⊂ B([0, 1])
be a sequence of open intervals such that, ai < λ < bi, (ai+1, bi+1) ⊂ (ai, bi),
limi→∞ ai = λ, limi→∞ bi = λ. Then, (ai, bi) ↓ {λ}. Moreover, by the uniform
continuity of F and the norm-1 property,
1 = lim
i→∞kF ((ai, bi))k = lim
i→∞kF ((ai, bi)) − F ({λ})k = 0.
Example 5.9. Let Qf be as in example 5.6. Theorem 5.8 implies that Qf
cannot have the norm-1 property. Indeed, for each λ ∈ R,
Qf ({λ})ψ = lim
i→∞
Qf ([λ, λi))ψ = lim
i→∞
µ[λ,λi)(x)ψ(x) = 0,
∀ψ ∈ H
where, λ, λi ∈ R, λi → λ.
6 Absolutely continuous semispectral measures
In the present section, we prove that absolutely continuous commutative POV
measures admit a strong Feller Markov kernel. Then, we apply the result to
the case of the unsharp position observable.
Definition 6.1. [43, 44] A POV measure F : B(R) → F(H) is absolutely
continuous with respect to a measure ν : B(R) → [0, 1] if there exists a positive
number c such that kF (∆)k ≤ c ν(∆), for each ∆ ∈ B(R).
Theorem 6.2. Let F be absolutely continuous with respect to a finite measure
ν. Then, F is uniformly continuous.
Proof. Suppose ∆i ↑ ∆. We have
n→∞kF (∆) − F (∆i)k = lim
lim
≤ c lim
which proves that F is uniformly continuous.
n→∞
n→∞kF (∆ − ∆i)k
ν(∆ − ∆i) = 0.
19
Corollary 6.3. Let F be absolutely continuous with respect to a finite measure
ν. Then, F is commutative if and only if there exist a self-adjoint operator A
and a strong Feller Markov kernel µ : R × B(R) → [0, 1] such that:
F (∆) = µ∆(A), ∆ ∈ B(R)
(20)
Proof. By theorem 6.2, F is uniformly continuous. Then, theorem 5.5 implies
the thesis.
Example 6.4. Let us consider the unsharp position operator defined as fol-
lows.
Qf (∆) :=Z[0,1]
µ∆(x) :=ZR
µ∆(x) dQx, ∆ ∈ B(R),
(21)
χ∆(x − y) f (y) dy,
x ∈ [0, 1]
where, f is a positive, bounded, Borel function such that f (x) = 0, x /∈ [0, 1],
Z[0,1]
f (x)dx = 1,
and Qx is the spectral measure corresponding to the position operator
Q : L2([0, 1]) → L2([0, 1])
ψ(x) 7→ Qψ := xψ(x)
Qf is absolutely continuous with respect to the measure
ν(∆) = MZ∆∩[−1,1]
dx.
Indeed, for each ψ ∈ H, ψ2 = 1,
hψ, Qf (∆)ψi =Z[0,1]
µ∆(x) =Z∆
where, the inequality
µ∆(x) ψ2(x) dx ≤ MZ∆∩[−1,1]
f (x − y) dy ≤ MZ∆∩[−1,1]
dx
dx
has been used.
Therefore, by theorem 6.2, Qf (∆) is uniformly continuous.
20
6.1 Unsharp Position Observable
In the present subsection, we study an important kind of absolutely continuous
POV measures, the unsharp position observables obtained as the marginals of
a covariant phase space observable.
In the following H = L2(R), Q and P denote position and momentum observ-
ables respectively and ∗ denotes convolution, i.e. (f∗g)(x) =R f (y)g(x−y)dy.
Let us consider the joint position-momentum POV measure [1, 20, 21, 26, 31,
40, 44, 45]
F (∆ × ∆′) =Z∆×∆′
Uq,p γ U ∗
q,p dq dp
where, Uq,p = e−iqP eipQ and γ = fihf, f ∈ L2(R), kfk2 = 1. The marginal
Qf (∆) := F (∆ × R) =Z ∞
−∞
(1∆ ∗ f2)(x) dQx, ∆ ∈ B(R),
(22)
is an unsharp position observable. Notice that the map µ∆(x) := 1∆ ∗ f (x)2
defines a Markov kernel.
Moreover, Qf is absolutely continuous with respect to the Lebesgue measure.
Indeed,
Uq,p γ U ∗
Qf (∆) = F (∆ × R) =Z∆×R
dqZR
=Z∆
=Z∆ bQ(q) dq ≤Z∆
q,p dq dp
1 dq
Uq,p γ U ∗
q,p dp
bQ(q) =ZR
Uq,p γ U ∗
q,p dp.
where,
Although Qf is absolutely continuous with respect to the Lebesgue measure on
R, it is not uniformly continuous. That does not contradict theorem 6.2 since
the Lebesgue measure on R is not finite. Anyway, Qf is uniformly continuous
on each Borel set ∆ with finite Lebesgue measure.
Now, we show that Qf is not in general uniformly continuous. We give the
details of the following particular case.
Example 6.5 (Optimal Phase Space Representation). If we choose
f 2(x) =
1
l √2 π
e(− x2
2 l2 ),
l ∈ R − {0}.
in (22), we get an optimal phase space representation of quantum mechanics
[40]. In this case,
21
Qf (∆) =Z ∞
−∞(cid:16)Z∆ f (x − y)2) dy(cid:17) dQx
l √2 πZ ∞
−∞(cid:16)Z∆
e− (x−y)2
2 l2 dy(cid:17) dQx =Z ∞
1
−∞
=
where,
µ∆(x) =
defines a Markov kernel.
1
l √2 πZ∆
e− (x−y)2
2 l2 dy
µ∆(x) dQx
(23)
In order to prove that Qf is not uniformly continuous we consider the
family of sets ∆i = (−∞, ai), limi→∞ ai = −∞ such that ∆i ↓ ∅, and prove
that limi→∞ kQf (∆i)k = 1. For each i ∈ N,
2 l2 dy
lim
e− (x−y)2
µ∆i(x) = lim
1
x→−∞
x→−∞
l √2 πZ∆i
l √2 πZ(−∞, ai−x)
1
= lim
x→−∞
e− y2
2 l2 dy =
1
l √2 πZ ∞
−∞
e− y2
2 l2 dy = 1.
Now, we prove that kF (∆i)k = 1, i ∈ N. Indeed, if
ψn = χ[−n,−n+1](x),
n→∞hψn, Qf (∆i)ψni = lim
lim
µ∆i(x)ψn(x)2 dx
µ∆i(x) dx = 1.
(24)
(25)
n→∞Z ∞
n→∞Z[−n,−n+1]
−∞
= lim
Since, for each ∆ ∈ B(R), kQf (∆)k ≤ 1, equation (24) implies that kQf (∆i)k =
1, for each i ∈ N. Hence, limi→∞ kQf (∆i)k = 1 and Qf cannot be uniformly
continuous.
It is worth noticing that although Qf is not uniformly continuous, µ∆ is
continuous for each interval ∆ ∈ B(R). Indeed,
µ∆(x) − µ∆(x′) =
e− (x−y)2
1
l √2 π(cid:12)(cid:12)(cid:12)Z∆
l √2 π(cid:12)(cid:12)(cid:12)Z∆x
1
2 l2 dy −Z∆
2 l2 −Z∆x′
′
e− (x
−y)2
2 l2
dy(cid:12)(cid:12)(cid:12)
l √2 π(cid:12)(cid:12)(cid:12)Z∆
2 l2 dy(cid:12)(cid:12)(cid:12) ≤
1
e− (y)2
e− (y)2
e− (y)2
2 l2 dy(cid:12)(cid:12)(cid:12)
=
where,
∆x = {z ∈ R z = y − x, y ∈ ∆}, ∆x′ = {z ∈ R z = y − x′, y ∈ ∆}
and,
Therefore, x − x′ ≤ ǫ implies,
1
µ∆(x) − µ∆(x′) ≤
∆ = (∆x − ∆x′) ∪ (∆x′ − ∆x).
l √2 π(cid:12)(cid:12)(cid:12)Z∆
1
l √2 π Z∆
√2
l √π
ǫ.
dy =
e− (y)2
2 l2 dy(cid:12)(cid:12)(cid:12) ≤
22
Appendices
A AW (F ) coincides with the von Neumann algebra
generated by {F (∆)}∆∈R(S)
We recall that S ⊂ B(R) is the countable family of open intervals with rational
end-points and R(S) the ring generated by R. Theorem c, page 24, in Ref.
[37] ensures the countability of R(S).
Proof. Let R be the extended real line, M := {F (∆)}∆∈B(R), and AW (F ) =
AW (M ) the von Neumann algebra generated by F . Let G denote the family
of open subsets of R and O := {F (∆), ∆ ∈ G}. Since the POV measure F
is regular, for each Borel set ∆, there exists a decreasing family of open sets
Gi such that F (Gi) → F (∆) strongly. Then, O is dense in M and the von
Neumann algebra generated by M coincides with the von Neumann algebra
generated by O. Hence,
AW (F ) = AW (M ) = AW (O).
(26)
Now, let G1 denote the family of open intervals in R. Let us consider the set
O1 = {F (∆), ∆ ∈ G1}. Each open set ∆ is the disjoint union of a countable
family of open intervals ∆i, i.e. ∆ = ∪∞
i=1∆i. Therefore,
F (∆i)
∞Xi=1
i=1∆i) =
F (∆) = F (∪∞
nXi=1
= lim
n→∞
F (∆i) = lim
n→∞
F (∪n
i=1∆i).
Since the von Neumann algebra generated by O1 contains F (∪n
contain F (∆) = limn→∞ F (∪n
i=1∆i). Therefore,
i=1∆i), it must
AW (O1) = AW (O).
(27)
Now, we prove that the von Neumann algebra AW (O2) generated by O2 =
{F (∆)}∆∈R(S) coincides with AW (O1).
For each open interval (a, b), a, b ∈ R, there exists a disjoint family of sets
{∆i}i∈N ⊂ R(S), ∆i ⊂ (a, b), i ∈ N, such that (a, b) = ∪∞
i=1∆i. Then,
F (∆i)
∞Xi=1
i=1∆i) =
F (a, b) = F (∪∞
nXi=1
= lim
n→∞
F (∆i) = lim
n→∞
F (∪n
i=1∆i).
23
Since the von Neumann algebra generated by O2 contains F (∪n
n ∈ N, it must contain F (∆) = limn→∞ F (∪n
AW (O2) and, by equations (26) and (27),
i=1∆i) for each
i=1∆i). Therefore, AW (O1) =
AW (O2) = AW (O1) = AW (O) = AW (F )
(28)
which proves that AW (F ) coincides with the von Neumann algebra generated
by the set {F (∆)}∆∈R(S).
B Sequences of continuous functions
The following theorem is due to Dini. We give a proof based on the use of
sequences.
Theorem B1. Let {fn(λ)}n∈N be a non increasing sequence of continuous
functions defined on a compact set B ⊂ [0, 1] with values in [0, 1] and such
that fn(λ) → 0 point-wise. Then, fn(λ) → 0 uniformly.
Proof. Since fn+1(λ) ≤ fn(λ) for each λ ∈ B, we have kfn+1k∞ ≤ kfnk∞. If
kfnk∞ → 0 clearly fn(λ) → 0 uniformly.
Then, suppose kfnk∞ → a > 0. Since kfn+1k∞ ≤ kfnk∞, we have kfnk∞ ≥ a,
for each n ∈ N.
Let λn be such that fn(λn) = kfnk∞. Since {λn} is a bounded sequence of real
numbers, there exists a convergent subsequence {λnk}k∈N. Let β be its limit,
i.e., β := limk→∞ λnk . The compactness of B assures that β ∈ B. Moreover,
limk→∞ fnk (λnk ) = a.
Let us consider the sequence of numbers fnk(β). We prove that fnk (β) ≥ a
for each k ∈ N. We proceed by contradiction. Suppose that there exists ¯k ∈ N
such that fn¯k (β) < a. Then, there exists a neighborhood I(β) of β such that
fn¯k(λ) < a for each λ ∈ I(β). Moreover, since λnk → β, there exists l ∈ N
such that k > l implies λnk ∈ I(β). Take k > max{¯k, l}. Then, λnk ∈ I(β)
and fnk(λ) ≤ fn¯k (λ), for each λ ∈ B. Therefore,
fnk(λnk ) ≤ fn¯k (λnk ) < a
which contradicts the fact that fnk(λnk ) = kfnkk∞ ≥ a, for each k ∈ N.
We have proved that fnk (β) ≥ a, for each k ∈ N. This implies that limk→∞ fnk (β) ≥
a and contradicts one of the hypothesis of the lemma, i.e., limn→∞ fn(λ) = 0
for each λ ∈ B.
24
References
[1] S.T. Ali:
'A geometrical property of POV-measures and systems of co-
variance.' In: Doebner, H.-D., Andersson, S.I., Petry, H.R. (eds.) 'Dif-
ferential Geometric Methods in Mathematical Physics,' Lecture Notes in
Mathematics, vol. 905, pp. 207-228, Springer, Berlin (1982).
[2] S.T. Ali, G.G. Emch, 'Fuzzy observables in quantum mechanics,' J. Math.
Phys. 15 (1974) 176.
[3] S.T. Ali, E.D. Prugovecki, Physica A, 89 (1977) 501-521.
[4] S.T. Ali, C. Carmeli, T. Heinosaari, A. Toigo, Found. Phys. 39 (2009)
593-612 .
[5] R. Beals: Topics in Operator Theory, The University of Chicago Press,
Chicago, 1971.
[6] R. Beneduci, G. Nistic´o, J. Math. Phys. 44 (2003) 5461.
[7] R. Beneduci, J. Math. Phys. 47 (2006) 062104.
[8] R. Beneduci, Int. J. Geom. Meth. Mod. Phys. 3 (2006) 1559.
[9] R. Beneduci, J. Math. Phys. 48 (2007) 022102.
[10] R. Beneduci, Il Nuovo Cimento B, 123 (2008) 43-62.
[11] R. Beneduci, Int. J. Theor. Phys. 49 (2010) 3030-3038.
[12] R. Beneduci, Bull. Lond. Math. Soc. 42 (2010) 441-451.
[13] R. Beneduci, Linear Algebra and its Applications, 43 (2010) 1224-1239.
[14] R. Beneduci, International Journal of Theoretical Physics, Vol. 50, (2011)
3724-3736, doi: 10.1007/s10773-011-0907-7.
[15] R. Beneduci, J. Brooke, R. Curran, F. Schroeck Jr., International Journal
of Theoretical Physics, 50 (2011) 3682-3696, doi: 10.1007/s10773-011-
0797-8.
[16] R. Beneduci, J. Brooke, R. Curran, F. Schroeck Jr., International Journal
of Theoretical Physics, 50 (2011) 3697-3723, doi: 10.1007/s10773-011-
0869-9.
[17] Beneduci R.: F. Schroeck Jr., A note on the relationship between local-
ization and the norm-1 property, J. Phys. A: Math. Theor. vol. 46, 305303
(2013).
25
[18] S. K. Berberian, Notes on Spectral theory, Van Nostrand Mathematical
Studies, New York, 1966.
[19] P. Billingsley, Convergence of Probability Measures, John Wiley and Sons,
New York (1968).
[20] P. Busch, M. Grabowski, P. Lahti, 'Operational quantum physics,' Lec-
ture Notes in Physics, vol. 31, Springer-Verlag, Berlin, 1995.
[21] E.B.Davies, J.T. Lewis, Comm. Math. Phys. 17 (1970) 239.
[22] J. Dixmier, C ∗-Algebras, North-Holland, New York, 1977.
[23] N. Dunford, J. T. Schwartz, Linear Operators, part II, Interscience Pub-
lisher, New York, 1963.
[24] C. Garola, S. Sozzo, Int. J. Theor. Phys., 49, 31013117 (2009).
[25] M. C. Gemignani, Elementary topology, Dover, New York, (1972) pp.
223-227.
[26] W. Guz, Int. J. Theo. Phys. 23 (1984) 157-184.
[27] P. R. Halmos, Measure Theory, Springer-Verlag, New York (1974).
[28] C. W. Helstrom, Quantum Detection and Estimation Theory, Academic,
New York, 1976.
[29] T. Heinonen, P. Lahti, J. P. Pelloppaa, S. Pulmannova, K. Ylinen, 'The
norm-1 property of a quantum observable,' Journal of Mathematical
Physics 44 (2003) 1998-2008.
[30] A. S. Holevo, 'An analog of the theory of statistical decisions in non-
commutative probability theory,' Trans, Moscow Math. Soc. 26 (1972)
133.
[31] A. S. Holevo, Probabilistics and statistical aspects of quantum theory,
North Holland, Amsterdam, 1982.
[32] A. Jencov´a, S. Pulmannov´a, Rep. Math. Phys. 59 (2007) 257-266.
[33] A. Jencov`a, S. Pulmannov`a, 'Characterizations of Commutative POV
Measures,' Found. Phys. 39 (2009) 613-624.
[34] R. V. Kadison, J. R. Ringrose, Fundamentals of the theory of operator
algebras I and II, Academic Press, New York, 1986.
[35] K. Kuratowski, A. Mostowski, Set Theory with an introduction to de-
scriptive set theory, North-Holland, New York 1976.
26
[36] K. Kuratowski, Topology, Academic Press, New York, 1966.
[37] M. Lo`eve, Probability Theory I, 4th edition, Springer-Verlag, Berlin, 1977.
[38] B. Maslowski, J. Seidler, Probability Theory and Related Fields, 118
(2000) 187-210.
[39] M.A. Naimark, Normed Rings, Wolters-Noordhoff Publishing, Gronongen
(1972).
[40] E. Prugovecki, Stochastic Quantum Mechanics and Quantum Spacetime,
D. Reidel Publishing Company, Dordrecht, Holland, 1984.
[41] M.Reed, B.Simon, Methods of modern mathematical physics, Academic
Press, New York, 1980.
[42] D. Revuz, Markov Chains, North Holland, Amsterdam (1984).
[43] F. E. Schroeck, Jr., Int. J. Theo. Phys. 28 247 (1989).
[44] F. E. Schroeck, Jr., Quantum Mechanics on Phase Space, Kluwer Aca-
demic Publishers, Dordrecht, (1996).
[45] W. Stulpe, 'Classical Representations of Quantum Mechanics Related to
Statistically Complete Observables,' Wissenschaft und Technik Verlag,
Berlin 1997. Also available: quant-ph/0610122
27
|
1208.4013 | 2 | 1208 | 2012-08-21T11:42:30 | The simplified version of the Spielman and Srivastava algorithm for proving the Bourgain-Tzafriri restricted invertiblity theorem | [
"math.FA"
] | By giving up the best constants, we will see that the original argument of Spielman and Srivastava for proving the Bourgain-Tzafriri Restricted Invertibility Theorem \cite{SS} still works - and is much simplier than the final version. We do not intend on publishing this since it is their argument with just a trivial modification, but we want to make it available to the mathematics community since several people have requested it already. | math.FA | math |
THE SIMPLIFIED VERSION OF THE SPIELMAN AND
SRIVASTAVA ALGORITHM FOR PROVING THE
BOURGAIN-TZAFRIRI RESTRICTED INVERTIBLITY
THEOREM
PETER G. CASAZZA
Abstract. By giving up the best constants, we will see that the original
argument of Spielman and Srivastava for proving the Bourgain-Tzafriri Re-
stricted Invertibility Theorem [2] still works - and is much simplier than the
final version. We do not intend on publishing this since it is their argument
with just a trivial modification, but we want to make it available to the
mathematics community since several people have requested it already.
1. Introduction
Recently, Spielman and Sristave [2] made a stunning achievement by show-
ing that one of the deeper and most useful results in pure mathematics, the
Bourgain-Tzafriri Restricted Invertibility Theorem [1], can be proved directly
with an algorithm. The original proof had a technical error which they cor-
rected in a later version. But this correction doubled the degree of difficulty of
the proof. We will see that their original proof is still valid if we are willing to
give up the best constant in the theorem.
2. The Theorem and Their Original Proof Adjusted
Theorem 2.1 (Spielman and Srivastave). Let H be a Hilbert space with or-
thonormal basis {vi}n
i=1. Assume L : H → H is a linear operator with kLvik =
1 for all i = 1, 2, · · · , n and assume
A =
m
Xi=1
LviLvT
i ,
has m non-zero eigenvalues, all of which are greater than b, and b′ = b−δ > δ.
If
T r[LT (A − bI)−1L] ≤ −n −
then there exists a vector ω ∈ {Lvi}n
i=1 satisfying:
2kLk2
δ
,
The author was supported by NSF DMS 1008183; and NSF ATD 1042701; AFOSR
DGE51: FA9550-11-1-0245.
1
2
P.G. CASAZZA
1. ωT (A − b′I)−1ω < −1, and hence ω = Lvj for some m < j ≤ n.
2. T r[LT (A + ωωT − b′I)−1L] ≤ T r[LT (A − bI)−1L] ≤ −n − 2kLk2
δ
.
(Note that we added a 2 to the original constant in [2] (part (2) above)
and as a result we have to change their starting point barrier from (1 − ǫ) to
(1 − 2ǫ).)
Proof. Step I: We show:
(A − bI)−1 − (A − b′I)−1 ≥
δ
2
(A − b′I)−2,
Note: In the original paper the above inequality was stated to hold for δ
instead of δ/2. But this isn't true and is not even true for real numbers.
Our fix will change their perfect constant for the lower Riesz bound from their
(1 − ǫ)2 to (1 − 2ǫ)(1 − ǫ).
Proof: Note first that δ ≤ b′ implies 2b′ ≥ b′ + δ. Thus
and finally
Now,
Also,
and so
Hence,
= and thus
1
b′ + δ
≥
1
2b′ ,
1
b′(b′ + δ)
≥
1
2(b′)2 .
−1
b
−
−1
b′ =
b − b′
bb′ =
δ
b′(b′ + δ)
≥
δ
2(b′)2 .
(1)
λi − b ≤ λi − b′,
(λi − b)(λi − b′) ≤ (λi − b′)2.
1
(λi − b)(λi − b′)
≥
1
(λi − b′)2 ,
1
λi − b
−
1
λi − b′ =
b − b′
(λi − b)(λi − b′)
≥
δ
(λi − b′)2 .
SPIELMAN AND SRIVASTAVA
3
Step 2: We observe that
T r[LT (A − b′I)−1L] ≤ T r[LT (A − bI)−1L].
Proof: By Step I, we have
T r[LT (A − bI)−1L − LT (A − b′I)−1L] = T r[LT ((A − bI)−1 − (A − b′I)−1)L]
≥
δ
2
T r[LT (A − b′I)−2L] ≥ 0.
Step 3: We show
T r[LT (A − b′I)−1LLT (A − b′I)−1L]
≤ (T r[LT (A − bI)−1L] − T r[LT (A − b′I)−1L])(−n − T r[LT (A − b′I)−1L]).
Proof: Since
we have
and hence,
T r[LT (A − b′I)−1L] ≤ −n −
2kLk2
δ
,
kLk2 ≤
δ
2
(−n − T r[LT (A − b′I)−1L)]),
kLk2T r[LT (A − b′I)−2L] ≤
δ
2
T r[LT (A − b′I)−2L](−n − T r[LT (A − b′I)−1L]).
Applying the proof of Step I, and the facts: (A − b′I)−1LLT (A − b′I)−1 ≥ 0
and LLT ≤ kLk2I, we have
T r[LT (A − b′I)−1LLT (A − b′I)−1L]
(2)
≤ kLk2T r[LT (A − b′I)−2L]
≤
δ
2
T r[LT (A − b′I)−2L](−n − T r[LT (A − b′I)−1L])
≤ (T r[LT (A − bI)−1L] − T r[LT (A − b′I)−1L)(−n − T r[LT (A − b′I)−1L]).
Step 4: We pick a vector ω satisfying (1) and
T r[LT (A − b′I)−1L] −
ω(A − b′I)−1LLT (A − b′I)−1ω
1 + ω(A − b′I)−1ω
≤ T r[LT (A − bI)−1L].
Proof: Noting that ωT ω = 1, it follows from inequality 2 that there is a vector
ω ∈ {Lvi}n
i=1 so that
ωT (A − b′I)−1LLT (A − b′I)−1ω
(3)
≤ (T r[LT (A − bI)−1L] − T r[LT (A − b′I)−1L)](−1 − ωT (A − b′I)−1ω)
4
P.G. CASAZZA
Since the left-hand side of Equation 3 is non-negative, applying Step 1 we have
and hence
0 < −1 − ωT (A − b′I)−1ω,
ωT (A − b′I)−1ω < −1.
For 1 ≤ j ≤ m, if ω = Lvj, then
m
ωT (A − b′I)−1ω =
λi − b′ ω2
So ω = Lvj for m < j ≤ n. Now, Equation 3 implies
ωT (A − b′I)−1LLT (A − b′I)−1ω
Xi=1
1
i ≥ 0.
−1 − ωT (A − b′I)−1ω
and the result follows.
≤ T r[LT (A − bI)−1L] − T r[LT (A − b′I)−1L],
Step; 5: We check part (2) of the theorem.
Proof: We apply the Sherman-Morrison formula - which states, for a matrix
A,
(A + ωωT )−1 = A−1 −
A−1ωωT A−1
1 + ωT A−1ω
.
It follows that (Replacing A by A − b′I)
LT (A + ωωT − b′I)−1L = LT (A − b′I)−1L −
LT (A − b′I)−1ωωT (A − b′I)−1L
1 + ωT (A − b′I)−1ω
.
Thus,
T r[LT (A + ωωT − b′I)−1L] =
T r[LT (A − b′I)−1L] −
T r[LT (A − b′I)−1ωωT (A − b′I)−1L]
1 + ωT (A − b′I)−1ω
.
Using the fact that T r[AB] = T r[BA], we have that the above equals
T r[LT (A − b′I)−1L] −
T r[ωT (A − b′I)−1LLT (A − b′I)−1ω]
1 + ωT (A − b′I)−1ω
=
T r[LT (A − b′I)−1L] −
ωT (A − b′I)−1LLT (A − b′I)−1ω
1 + ωT (A − b′I)−1ω
.
We now have applying Step 4:
T r[LT (A+ωωT −b′I)−1L] = T r[(LT (A−b′I)−1L]−
ωT (A − b′I)−1LLT (A − b′I)−1ω
1 + ωT (A − b′I)−1ω
This completes the proof of the theorem.
(cid:3)
≤ T r[LT (A − bI)−1L]
SPIELMAN AND SRIVASTAVA
5
Corollary 2.2 (Bourgain-Tzafriri Restricted Invertibility Theorem). If we
iterate the algorithm k times, we get k vectors from {Lvi}m
i=1 with lower Riesz
bound for the operator A
1 − 2ǫ − (k − 1)δ = (1 − 2ǫ)(1 − (k − 1)
kLk2
ǫn
)
Hence,
1. If
then
k = ⌈
ǫ2n
kLk2 ⌉,
(1 − 2ǫ)(cid:20)1 − (k − 1)
kLk2
ǫn (cid:21) ≥ (1 − 2ǫ)(cid:20)1 −
ǫ2n
kLk2
kLk2
ǫn (cid:21)
= (1 − 2ǫ)(1 − ǫ).
which is BT.
2. If
then
k = ⌈
ǫn
kLk2 ⌉,
(1 − 2ǫ)(cid:20)1 − (k − 1)
kLk2
ǫn (cid:21) = (1 − 2ǫ)(cid:20)1 −
ǫn
kLk2
kLk2
ǫn (cid:21)
and the process stops.
= (1 − 2ǫ)0,
References
[1] J. Bourgain and L. Tzafriri, Invertibility of "large" submatrices and applications to
the geometry of Banach spaces and Harmonic Analysis, Israel J. Math. 57 (1987)
137-224.
[2] D.A. Spielman and N. Srivastava, , Israel Jour. Math. 19 No. 1 (2012) 83-91.
Department of Mathematics, University of Missouri, Columbia, MO 65211-
4100
E-mail address: [email protected]
|
1307.3523 | 2 | 1307 | 2013-07-16T08:39:47 | Virtual continuity of the measurable functions of several variables, and Sobolev embedding theorems | [
"math.FA"
] | Classical Luzin's theorem states that the measurable function of one variable is "almost" continuous. This is not so anymore for functions of several variables. The search of right analogue of the Luzin theorem leads to a notion of virtually continuous functions of several variables. This probably new notion appears implicitly in the statements like embeddings theorems and traces theorems for Sobolev spaces. In fact, it reveals their nature as theorems about virtual continuity. This notion is especially useful for the study and classification of measurable functions, aswell as in some questions on dynamical systems, polymorphisms and bistochastic measures. In this work we recall necessary definitions and properties of admissible metrics, define virtual continuity, describe some of applications. Detailed analysis is to be presented in another paper. | math.FA | math | VIRTUAL CONTINUITY OF
MEASURABLE FUNCTIONS OF SEVERAL
VARIABLES AND EMBEDDINGS
THEOREMS
A. M. Vershik, P. B. Zatitskiy, F. V. Petrov
.
A
F
h
t
a
m
[
2
v
3
2
5
3
.
7
0
3
1
:
v
i
X
r
a
Abstract
Classical Luzin's theorem states that the measurable function of
one variable is "almost" continuous. This is not so anymore for func-
tions of several variables. The search of right analogue of the Luzin
theorem leads to a notion of virtually continuous functions of sev-
eral variables. This probably new notion appears implicitly in the
statements like embeddings theorems and traces theorems for Sobolev
spaces. In fact, it reveals their nature as theorems about virtual conti-
nuity. This notion is especially useful for the study and classification
of measurable functions, as well as in some questions on dynamical
systems, polymorphisms and bistochastic measures. In this work we
recall necessary definitions and properties of admissible metrics, define
virtual continuity, describe some of applications. Detailed analysis is
to be presented in another paper.
St.
Petersburg Department of V.A.Steklov Institute of Mathematics RAS,
St. Petersburg State University. E-mail: [email protected], [email protected],
[email protected]. The work is supported by RFBR grant 11-01-00677-a, President
of Russia grant MK-6133.2013.1, by Chebyshev Laboratory in SPbSU, Russian Govern-
ment grant 11.G34.31.0026.
1
Introduction. Admissible metrics,
1
Luzin's theorem.
1.1 Admissible metrics
We consider the Lebesgue-Rokhlin standard continuous (atomless)
probabilistic measure space, isomorphic to the unit segment [0, 1] with
Lebesgue measure. The first author proposes [4, 8, 11] to consider on
the fixed standard space (X, A, µ) different (admissible) metrics, on the
contrast to usual approach, when metric space is fixed and different
Borel measures are considered. Such approach is useful and necessary
in ergodic theory and other situations. Agreement of the metric and
measure structures leads to the notion of admissible metric triple:
Definition 1. A metric or semimetric ρ on the space X is called
admissible if it is measurable, regarded as a function of two variables,
on the Lebesgue space (X × X, µ × µ) and there exists a subset X0 ⊂ X
of full measure such that the semimetric space (X0, ρ) is separable.
The standard probabilistic space (X, µ) with admissible (semi-
)metric ρ is called admissible metric triple or just admissible triple
(X, µ, ρ).
Properties of admissible metrics are studied in details by the au-
thors in [12], [17]. In particular, these works contain several equivalent
definitions of admissible triples.
Standartness of the space allows to get the following
Proposition 1. Let ρ be the admissible metric on (X, A, µ). Then
completed Borel sigma-algebra B = B(X, ρ) is a subalgebra of A. The
measure µ is inner regular with respect to metric ρ, i.e. for any A ∈ A
we have
µ(A) = sup{µ(K) : K ⊂ A, K is compact in metric ρ}.
So, for any admissible metric ρ the measure µ is Radon measure in
the metric space (X, ρ).
M. Gromov in the book [3] suggests to consider arbitrary metric
triples (X, µ, ρ), which he calls mm-spaces. Also, Gromov asks the
question about their classification, having in mind classical situations
(Riemannian manifolds and so on). It is natural to consider admissible
2
triples in this framework. Define equivalence of admissible triples up
to measure-preserving isometries: (X, µ, ρ) ∼ (X ′, µ′, ρ′), if
∃T : X → X ′; T µ = µ′;
ρ′(T x, T y) = ρ(x, y).
Here is the main result on this equivalence:
Theorem 1. (Gromov [3]; Vershik [4])
Consider the map Fρ : X∞ × X∞ → M∞(R) :
Fρ({xi, yj}(i,j)∈N×N) = {ρ(xi, yj)}(i,j)∈N×N,
and equip infinite product X∞ ×X∞ by the product-measure µ∞ ×µ∞.
Let Dρ denote the measure on the space of matrices (i.e.
random
matrix of distances), which is the Fρ-image of the measure µ∞ × µ∞.
Call it MATRIX DISTRIBUTION of the metric ρ. It is a complete
invariant of above equivalence of admissible metrics.
In other words,
(X, µ, ρ) ∼ (X ′, µ′, ρ′) ⇔ Dρ = Dρ′.
In [5] this result is generalized to the so called pure measurable
functions of several variables.
The following lemma is useful in the theory of admissible metrics:
Lemma 2. Let ρ1, ρ2 be admissible semimetrics on the standard space
(X, µ), and suppose that ρ1 is metric. Then for any ε > 0 there exists
measurable subset K ⊂ X such that µ(K) > 1 − ε and semimetric ρ2
(as a function of two variables) is continuous on K × K with respect
to metric ρ1.
We may choose K as a compact subset with respect to admissible
metric ρ = ρ1 + ρ2, if µ(K) > 1 − ε.
Lemma immediately implies the
Corollary 3. Let ρ1 and ρ2 be two admissible metrics on the standard
space (X, µ). Then for any ε > 0 there exists K ⊂ X such that µ(K) >
1 − ε and topologies defined by metrics ρ1 and ρ2 on K coincide.
1.2 Luzin's theorem on measurable functions
of one variable
Furthermore we consider (measurable) real-valued functions, though
most of our results remain true for maps into standard Borel space, in
3
particular into Polish spaces. Egorov's and Luzin's classical theorems
on measurable functions of one variable are well-known. The general-
ized Luzin's theorem for arbitrary admissible triple follows from above
results:
Corollary 4 (Luzin's theorem). Let ρ be an admissible metric on the
standard space (X, µ), let f be a measurable map from X into Polish
space (M, d). Then for any ε > 0 there exists a measurable subset
K ⊂ X such that µ(K) > 1 − ε and f is continuous on K with respect
to metric ρ.
Proof. Set ρ1(x, y) = ρ(x, y) + d(f (x), f (y)). Then ρ1 is a trivial
example of an admissible metric, with respect to which f is continuous.
By 3 there exist a subset K having measure µ(K) > 1 − ε, on which
this continuity implies continuity with respect to ρ.
But this fact does not hold true for functions of several variables.
2 Virtual continuity
2.1 Definitions and first examples
Let f (·, ·) be a measurable function of two variables. Then Luzin's
theorem analogue (continuity on the product X ′×Y ′ of sets of measure
> 1 − ε with respect to given metric ρ[(x1, y1), (x2, y2)] = ρX(x1, x2) +
ρY (y1, y2)) is not in general true. This leads to the following key
notion of this work. (Sum of metrics may be replaced to maximum
or other metric defining the topology of direct product. To stress this
we denote generic metric with such topology by ρX × ρY ).
Definition 2. Measurable function f (·, ·) on the product (X, µ)×(Y, ν)
of standard spaces is called virtually continuous, if for any ε > 0
there exist sets X ′ ⊂ X, Y ′ ⊂ Y , each of which having measure 1 − ε,
and admissible semimetrics ρX , ρY on X ′, Y ′ respectively such that
function f is continuous on (X ′ × Y ′, ρX × ρY ). virtual functions of
several variables are defined in the same way.
It is essential that admissible metric with respect to which function
becomes continuous is not arbitrary, but respects the structure of di-
rect product (in more general setting, it respects selected subalgebras,
see further). It is easy to verify that there does not exist universal
metric of such type (i.e. such a metric that virtual continuity implies
4
continuity in this metric).
defined notion.
It explains the non-trivial properties of
It is clear that any admissible metric (considered as a function
of two variables) is virtually continuous. So is any function, which is
continuous with the respect to product of admissible metrics. Degener-
i=1 ϕi(x)ψi(y),
where φi(·), ψi(·), i = 1, . . . n are arbitrary measurable functions, are
also virtually continuous. For the proof just use Luzin's theorem for
all functions ϕi(·), i = 1 . . . n, and ψi(·), i = 1 . . . n.
ated functions (or "finite rank functions") f (x, y) = Pn
less trivial examples of virtually continuous functions are given by
functions from some Sobolev spaces and kernels of trace class opera-
tors. For virtually continuous functions there exist well-define restric-
tions on some subsets of zero measure -- concretely, onto supporters
of (quasi)bistochastic measures, see next paragraph.
An easy example of not virtually continuous measurable function
on [0, 1]2 is provided by the characteristic function of the triangle
{x ≥ y}. In general, for functions on the square of a compact group
depending of the ratio of variables the criterion of virtual continuity
is simple:
Proposition 5. Let G be a metrizable compact group, f be a Haar
measurable function on G. Then the function F (x, y) := f (xy−1)
on G × G is virtually continuous if and only if f is equivalent to a
continuous function.
Stress once more that the definition of virtual continuity is not
topological, but measure-theoretical in nature. It applies to the choice
of various metrics on the measure space. So, the direct sense of the
proposition 5 is that group structure and measure-theoretical structure
allow to reconstruct topology.
2.2 Bistochastic measures and polymorphisms
From the measure-theoretical point of view a function of k variables on
the product of standard continuous spaces is nothing but the function
on the standard continuous space (due to isomorphism of all such
spaces). In order to deal with it as a function of k variables, we have
to introduce another category, then just measurable spaces.
namely, consider the following structure:
(X , A, m), with k selected sigma-subalgebras A1, . . . , Ak in A.
the measure space
It is
5
natural to suppose that those subalgebras generate the whole sigma-
algebra A.
The connection with general viewpoint is the following:
in the
space X = Qk
i=1(Xi, Ai, µi), m = Q µi, identify algebras Ai with
subalgebras of A = Q Ai by multiplying to trivial subalgebras on
other multiples.
measurable iff f depends only on i-th variable xi (i = 1, . . . , k).
In other words, function f (x1, . . . , xk) on X is Ai-
Definition 3. Measurable function on the standard space X with k
selected subalgebras, which generate the whole sigma-algebra, is called
general measurable function of k variables.
Consider some measure λ on the sigma-algebra A. It may be re-
stricted onto sigma-subalgebras Ai, i = 1, . . . , k. Let's consider such
measures λ for which those restrictions are absolutely continuous with
respect to the measure m (restricted onto Ai). If restrictions of λ onto
Ai coincide with m, i = 1 . . . k, such a measure λ is called multi-
stochastic with respect to given subalgebras (bistochastic for k = 2);
if restrictions are just equivalent to m for i = 1, . . . , k, we call λ almost
multistochastic. Finally, if λ(U ) ≤ m(U ) for any U ∈ Ai, i = 1, . . . , k,
we call λ submultistochastic.
Of course, bistochastic measure on X × Y may be singular with
respect to the product measure. For instance, in the case of direct
product of segments (X, µ) = (Y, ν) = [0, 1] there is a bistochastic
measure λ on diagonal {x = y} (with density dµ(x)).
Furthermore we suppose for simplicity that k = 2, i.e. consider
functions of two variables. But there is no serious difference for k > 2.
We consider not only independent variables, most of the notions may
be defined for general pair of sigma-algebras. But even the case of
independent variables is often useful to treat as a general case.
Bistochastic measure on the direct product of spaces define the
so called polymorphism of the space (X, µ) into (Y, ν) (see [9]), i.e.
"multivalued mapping" with invariant measure. The case of identified
variables (X, A, µ) = (Y, B, ν) is of special interest: polymorphism in
this case generalizes the concept of automorphism of measure space.
Almost bistochastic measures defines a polymorphism with quasi in-
variant measure. Bistochastic or almost bistochastic measure λ de-
fines also a bilinear (in general case k-linear) form (f (x), g(y)) →
R f (x)g(y)dλ(x, y), corresponding to the so called Markovian, resp.
quasi Markovian operator in corresponding functional spaces. Note
that this operator Uλ is a contraction, i.e. has norm at most 1, which
6
preserves the cone of non-negative functions. In the case of bistochas-
tic measure this operator (as well as adjoint operator) preserve con-
stants: Uλ1 = 1.
See [9, 10, 7] about many connections of polymorphisms (Marko-
vian operators, joinings, couplings, correspondences, Young measures,
bibundles etc). Bistochastic measures play a key role in the intensively
developing theory of continuous graphs [15].
2.3 Further properties of virtually continuous
functions
First of all, virtually continuous functions enjoy the properties a priori
stronger than required in the definition. On the one side, we may
require for sets X ′, Y ′ from the definition to have full measure:
Theorem 2. Let function f (·, ·) be virtually continuous. Then there
exist sets X ′ ⊂ X, Y ′ ⊂ Y of full measure and admissible semi-
metrics ρX, ρY on X ′, Y ′ respectively such that f is continuous on
(X ′ × Y ′, ρX × ρY ).
On the other hand, we may fix arbitrary admissible metrics:
Theorem 3. Let function f (·, ·) be virtually continuous. Then for
any admissible semimetrics ρX , ρY on X ′, Y ′ and for any ε > 0 there
exist sets X ′ ⊂ X, Y ′ ⊂ Y , each of which having measure 1 − ε, such
that function f is continuous on (X ′ × Y ′, ρX × ρY ).
A function of two variables on X × Y may be treated as a map
from X into the space of functions on Y (i.e. f (x, y) ≡ fx(y)). In [16]
such a viewpoint is used for classification problem. Virtual continuity
is described in those terms by the following equivalent definition:
Theorem 4. Virtual continuity of the function f (·, ·) is equivalent to
the following property of a function:
for any ε > 0 there exist sets
X ′ ⊂ X, Y ′ ⊂ Y having measures not less then 1 − ε, such that the
set of functions fx(·) on Y ′ (variable x runs over X ′) form a totally
bounded (precompact) family in L∞(Y ′).
This theorem-definition has an important corollary: continuity in
one variable implies virtual continuity.
Lemma 6. Let (X, µ), (Y, ν) be standard continuous probabilistic
spaces, ρY be an admissible metric on the set Y ′ of full measure. Let
7
measurable function f : X × Y → R be so that functions f (x, ·) are
continuous on (Y ′, ρY ) for µ-almost all x ∈ X. Then f is virtually
continuous.
Theorem 2 immediately implies the converse:
for any virtually
continuous function f such a metric ρY exists. So, the statement of
Lemma 6 ("continuity by appropriate metric on y for almost all fixed
x") is equivalent to virtual continuity.
It's remarkable that the spaces X and Y (i.e. arguments of the
function) play different roles in this definition. However, a posteriori
the property appears to be symmetric under the change of order of
variables. This is another demonstration of the non-triviality of the
virtual continuity concept.
As we have seen, measurable functions f (·, ·) are classified by ma-
trix distributions, i.e. by measures on the space of matrices (aij)∞
i,j=1,
induced by the map f → (ai,j = f (xi, yj)) (points xi in X and yi in Y ,
i = 1, 2, . . . are chosen independently). Virtual continuity also may
be characterized on this manner:
Theorem 5. Let x1, x2, . . . (resp. y1, y2, . . . ) be independent random
points in X (resp. in Y ). Virtual continuity of the measurable func-
tion f (x, y) is equivalent to the following condition: for any ε > 0 there
exist positive integer N such that the following event has probability
1:
there exist such a partitions of naturals {1, 2, . . . , } = ⊔N
i=0Ai =
⊔N
i=0Bi that upper density of the set A0 ∪ B0 is less than ε (i.e.
lim sup (A0 ∪ B0) ∩ [1, n]/n < ε) and f (xs, yt) − f (xr, yp) < ε for all
i, j > 0, s, r ∈ Ai, p, t ∈ Bj.
Aforementioned characteristics of the virtual continuity allow to
deduce that virtual continuous functions form a nowhere dense set
in the space of all measurable functions (with measure convergence
topology).
2.4 Thickness
Consider the space X × Y with product measure µ × ν. Choose two
subalgebras in its sigma-algebra, defined by projections onto X and
Y . For measurable set Z ⊂ X × Y define its thickness th(Z) as the
infimum of the value µ(X1) + ν(Y1) taken by all pairs of measurable
sets X1 ⊂ X, Y1 ⊂ Y such that
µ × ν(Z \ (X1 × Y ∪ X × Y1)) = 0.
(1)
8
Sets of the form X1 × Y , X × Y1 are exactly sets from our sigma-
subalgebras. It allows to define generalized thickness for other selected
subalgebras in the standard space.
The following properties of thickness are immediate:
• thickness of a set does not exceed 1 and equals 0 for and only
for sets of measure 0;
• thickness of a subset does not exceed a thickness of a set;
• thickness of a set does not exceed its measure;
• thickness of a finite or countable union of sets does not exceed
sum of thicknesses.
The following lemma is slightly less obvious.
Lemma 1. If th(Z) = 0, then X1, Y1 of zero measure may be chosen
in the definition 1.
By "arbitrarily thin set" we mean "the set of arbitrarily small
thickness". It allows to reformulate lemma 1 as follows: if the set may
covered may arbitrarily thin set, then its complement contains mod 0
the product of full measure sets.
the following equivalent definition of thickness is in some situations
more appropriate for using:
Lemma 2. For any set Z consider pairs of measurable functions f :
X → [0, 1], g : Y → [0, 1], such that f (x) + g(y) ≥ χZ (x, y) for µ × ν-
almost all pairs (x, y). Then thickness of Z is infimum of the sum of
integrals RX f dµ and RY gdν.
Applying this lemma and choosing weakly convergent subsequence
the lower semicontinuity of thickness may be proved:
Lemma 3. Let {Zn} -be increasing sequence of measurable sets, Z =
∪nZn. Then th(Z) = lim th(Zn).
Note that there is no upper semicontinuity: all sets {(x, y) : 0 <
x − y < 1/n} ⊂ [0, 1]2 have thickness 1, while their intersection is
empty.
Define the convergence of functions "in thickness" analogously to
convergence in measure. This is convergence in the metrizable topol-
ogy defined by the following distance:
9
Definition 4. Define the distance τ (f (x, y), g(x, y)) between arbitrary
measurable functions of two variables as infimum of such ε > 0 that
th{(x, y) : f (x, y) − g(x, y) > ε} ≤ ε.
Convergence in this τ -metric implies convergence in measure (but
not vice versa).
Let ξX : X = ⊔n
i=1Xi, ξY : Y = ⊔m
i=1Yi be finite partitions of the
spaces X, Y respectively onto measurable subsets of positive measure.
Functions which are constant mod 0 on each product Xi ×Yj are called
i=1 ai(x)bi(y) are called
functions of finite rank. The set of measurable functions is complete
in τ -metric.
step functions. Finite linear combinations PN
The following theorem connects finite rank functions and virtual
continuity.
Theorem 6. The τ -closure of the set of step functions (or the set of
finite rank functions) is exactly the set of virtually continuous func-
tions.
This definition of virtual continuity is even more explicitly
measure-theoretical,
it does not appeal to metrics at all. Also, it
has clear generalisation to arbitrary pair of sigma-subalgebras. See
[13] on close concepts.
2.5 Norm in the space of virtually continuous
functions
Defined convergence in τ -metric is a virtual continuity analogue of
the convergence in measure. Known Banach spaces of measurable
functions also have their analogues.
A measurable function h(·, ·) on the space (X × Y, µ × ν) is called
subbistochastic, if the measure with µ × ν-density h(·, ·) is subbis-
tochastic. Denote by S the set of subbistochastic functions.
Define a finite or infinite norm of a measurable function f (·, ·) as
kf k := inf{ZX
a(x)dµ(x)+ZY
b(y)dν(y) : f (x, y) ≤ a(x)+b(y) a.e.}
Next theorem is an analogue of known L. V. Kantorovich's duality
theorem [1] in the mass transportation problem (concretely, of duality
between measures space with Kantorovich distance and and the space
of Lipschitz functions, see also [18]).
10
Theorem 7.
kf k = sup{ZX×Y
f (x, y)h(x, y)dxdy : h ∈ S}.
Coincidence of infimum and supremum is a duality statement in
infinite-dimensional linear programming. But in our case the proof
requires more delicate arguments than the Monge-Kantorovich prob-
lem. The reason is that we consider L1-type space, in which the cone
of non-negative functions has empty interior (on the contrast to the
space of continuous functions, used in the transport problem). This
does not allow to apply directly standard separability theorems.
Theorem 8. The closure of step functions in above norm consists
exactly of all virtually continuous functions having finite norm (in
particular, each bounded virtually continuous function belongs to this
closure).
Denote this space by V C 1.
It is an analogue of the space L1
for virtually continuous functions and is a pre-dual for the space of
polymorphisms with bounded densities of projections.
Theorem 9. The space dual to V C 1 is a space of signed measures η
on X × Y with finite norm
∂P x
∗ η
∂µ (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)L∞(X,µ)
,(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
∂P y
∗ η
∂ν
},
(cid:12)(cid:12)(cid:12)L∞(Y,ν)
(cid:12)(cid:12)(cid:12)
kηkme = max{(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
where P x and P y are projections onto X and Y respectively and η is
a full variation of a signed measure η.
Corollary 1. Virtually continuous functions from V C 1 (in particular,
bounded virtually continuous functions) have a well defined integral not
only over sets of positive measure (as all summable functions do have),
but also also over bistochastic (singular) measures: for instance, over
Lebesgue measure on diagonal {x = y} ⊂ [0, 1]2, or over measure
concentrated on a graph of a map with quasiinvariant measure. To
summarize, virtually continuous functions have traces (restrictions)
on diagonal and other subsets in the sense of Sobolev trace theorems.
11
3 Applications:
theo-
rems, trace theorems and restrictions
of metrics
embeddings
Here we mention some applications of virtual continuity.
3.1 Sobolev spaces and trace theorems
Theorem 10. Let Ω1, Ω2 be domains of dimensions d1, d2 respectively,
suppose that pl > d2 or p = 1, l = d2. Then functions from the Sobolev
space W l
p(Ω1 × Ω2) (l-th generalized derivatives are summable with
power p) are virtually continuous as functions of two variables x ∈ Ω1,
p(Ω1 × Ω2) into V C 1(Ω1, K) is continuous for
y ∈ Ω2. Embedding W l
any compact subset K of the domain Ω2.
Proof. Using the theorem of embedding of Sobolev space into continu-
ous functions (see, for instance, [2, 6]), we have the following estimate
for functions h(y) ∈ W l
p(Ω2):
khkC(K) ≤ c(Ω2, K)khkW l
p(Ω2).
Let f (x, y) ∈ W l
p(Ω1 × Ω2) be a smooth function. Set
a(x) := kf (x, ·)kW l
p(Ω2).
Then by Fubini's theorem a ∈ L1(Ω1) and
Z a ≤ c(Ω1, Ω2)kf (x, y)kW l
p(Ω1×Ω2).
The following estimate holds on Ω1 × K:
f (x, y) ≤ kf (x, ·)kC(K) ≤ c(Ω2, K)a(x).
Summarizing this we have
kf kV C 1(Ω1,K) ≤ c(Ω1, Ω2, K)kf kW l
p(Ω1×Ω2).
(2)
Each function in the class W l
smooth functions, by (2) it is a limit in V C 1 as well.
p(Ω1 × Ω2) is a limit of a sequence of
So, under conditions of this theorem we may integrate functions
over quasibistochastic measures. It generalizes usual theorems about
traces on submanifolds.
12
3.2 Nuclear operators in Hilbert space
It is well known that the space of nuclear operators in the Hilbert
space L2 is a projective tensor product of Hilbert spaces. Their ker-
nels are measurable functions of two variables, which can hardly be
described directly. the following theorem claims that kernels of nuclear
operators are virtually continuous as functions of two variables. Note
that kernels of Hilbert-Schmidt operators are not in general virtually
continuous.
Theorem 11. Let (X, µ), (Y, ν) be standard spaces.
the space of
kernels of nuclear operators from L2(X) to L2(Y ) (with Schatten --
von Neumann norm) embeds continuously into V C 1.
It implies that such kernels may be integrated not only over diag-
onal when X = Y , which is well known, but by bistochastic measures.
But the space V C 1 is wider than kernels of nuclear operators.
If
we look at V C 1 as to the space of kernels of integral operators, it
is not unitray invariant, on the contrast to Schatten -- von Neumann
indeed, the definition of V C 1 essentially uses known sigma-
spaces.
subalgebras, which do not have necessary invariance. Close question
is considered in [14].
3.3 Restrictions of metrics
The following problem was one of origins of this paper. Let (X, µ)
be a standard space with continuous measure. Assume that ρ is an
admissible metric and ξ is a measurable partition of (X, µ) with parts
of null measure (say, ξ is a partition onto level sets of function which
is not constant on sets of positive measure). May we correctly restrict
our metric (a s a function of two variables) onto elements of this
partition?
It is not immediately clear, since the metric is a priori just a
measurable function. But admissible metric is virtually continu-
ous, and so for our goal
it suffices to define a bistochastic mea-
sure, onto which we have to restrict it. Suppose for simplicity that
X = [0, 1]2, µ is a Lebesgue measure, ξ is a partition onto ver-
tical lines. Then we say about restriction of virtually continuous
function defined on X 2 = [0, 1]4 onto three-dimensional submanifold
{(x1, x2, x3, x4) : x1 = x3}.
It is easy to see that such a submani-
fold equipped by a three-dimensional Lebesgue measure defines a bis-
tochastic measure on X × X.
13
4 Acknowledgements
We are grateful to L.Lovasz, who sent us his recent monograph [15],
in which close questions are discussed, and to A.Logunov for paying
our attention to the possibility of dual definition of the thickness.
14
References
[1] L. V. Kantorovich. On the Translocation of Masses, Dokl. Akad.
Nauk SSSR, 37, No. 7 -- 8, 227 -- 229 (1942). English translation:
Journal of Mathematical Sciences, 3-2006, Volume 133, Issue 4,
pp 1381-1382.
[2] V. G. Maz´ya. Sobolev spaces, Izdat. Leningrad. Univ., Leningrad
(1985). English translation: Springer-Verlag, Berlin, (1985).
[3] M. Gromov. Metric Structure for Riemannian and Non-
Riemannian Spaces. Birkhouser (1998).
[4] A. Vershik. The universal Uryson space, Gromov's metric triples,
and random metrics on the series of natural numbers. Uspekhi
Mat. Nauk 53, No.5, 57-64 (1998). English translation: Russian
Math. Surveys 53, No.5, 921-928 (1998).
[5] A. Vershik. Classification of measurable functions of several argu-
ments, and invariantly distributed random matrices (in Russian).
Funkts. Anal. Prilozh. 36, no.2, 12-28 (2002). English translation:
Funct. Anal. Appl. 36, no.2, 93-105 (2002).
[6] R. A. Adams, J. J. .F Fournier. Sobolev spaces. Academic press,
v. 140 (2003).
[7] A. Vershik. Three lectures on invariant measures and universal-
ity. In: Dynamics and Randomness II, A.Maass, S.Martinez, and
J.San Martin (eds.), Kluwer Academic Publ., Netherlands, 2004,
pp.199-228.
[8] A. Vershik. Random metric spaces and universality. Uspekhi Mat.
Nauk 59, No. 2(356), 65-104 (2004). English translation: Russian
Math. Surveys 59, No. 2, 259-295 (2004).
[9] A. Vershik. Polymorphisms, Markov processes, and quasi-
similarity. Discrete Contin. Dyn. Syst. 13, No. 5, 1305-1324
(2005).
[10] A. Vershik. Vershik A. What does a generic Markov operator
look like? Algebra i Analiz 17, No. 5, 91-104 (2005). English
translation: St. Petersburg Mathematical Journal. 17, No. 5. p.
763-772. (2006)
[11] A. Vershik. Dynamic of metrics in measure spaces and their
asymptotic invariants. Markov Processes and Related Fields, 16
No.1, 169-185 (2010).
15
[12] P. Zatitskiy, F. Petrov. Correction of metrics. Zapiski Nauch-
nykh Seminarov POMI, 390, 201-209 (2011). English translation:
arXiv:1112.2380.
[13] K-T. Sturm. The space of spaces:curvature bounds and gradient
flows on the space of metric measure space. (Preprint)
[14] M. Denker, M. Gordin. Limit theorem for von Mises statistics of
a vrasure preserving transformations. arXiv:1109.0635v2
[15] L. Lovasz. Large networks and graph limits. Colloquium Publi-
cations, Vol. 60. (2012)
[16] A. Vershik. On classification of measurable functions of several
variables, Zapiski Nauchn. Semin. POMI 403, 35-57 (2012). En-
glish version to appear in J. Math. Sci. 190, No. 3 (2013).
[17] F. Petrov, A. Vershik, P. Zatitskiy. Geometry and dynamics of
admissible metrics in measure spaces. Central Europ.J.Math. 11,
No.3, 379-400. (2013)
[18] A. Vershik. Long History of Monge-Kantorovich transportation
problem. Mathem.Intellegencer, 35, No.4. (2013).
16
|
1808.03866 | 1 | 1808 | 2018-08-11T21:11:43 | Log-majorization related to R\'enyi divergences | [
"math.FA"
] | For $\alpha,z>0$ with $\alpha\ne1$, motivated by comparison between different kinds of R\'enyi divergences in quantum information, we consider log-majorization between the matrix functions \begin{align*} P_\alpha(A,B)&:=B^{1/2}(B^{-1/2}AB^{-1/2})^\alpha B^{1/2}, \\ Q_{\alpha,z}(A,B)&:=(B^{1-\alpha\over2z}A^{\alpha\over z}B^{1-\alpha\over2z})^z \end{align*} of two positive (semi)definite matrices $A,B$. We precisely determine the parameter $\alpha,z$ for which $P_\alpha(A,B)\prec_{\log}Q_{\alpha,z}(A,B)$ and $Q_{\alpha,z}(A,B)\prec_{\log}P_\alpha(A,B)$ holds, respectively. | math.FA | math |
Log-majorization related to R´enyi divergences
Fumio Hiai1
1 Tohoku University (Emeritus),
Hakusan 3-8-16-303, Abiko 270-1154, Japan
Abstract
For α, z > 0 with α 6= 1, motivated by comparison between different kinds of
R´enyi divergences in quantum information, we consider log-majorization between
the matrix functions
Pα(A, B) := B1/2(B−1/2AB−1/2)αB1/2,
Qα,z(A, B) := (B
1−α
2z A
α
z B
1−α
2z )z
of two positive (semi)definite matrices A, B. We precisely determine the parame-
ter α, z for which Pα(A, B) ≺log Qα,z(A, B) and Qα,z(A, B) ≺log Pα(A, B) holds,
respectively.
2010 Mathematics Subject Classification: 15A42, 15A45, 47A64
Key words and phrases: Positive definite matrix, Log-majorization, R´enyi diver-
gence, Operator mean, Operator perspective, Weighted geometric mean, Unitar-
ily invariant norm
1
Introduction
For each n ∈ N we write Mn for the n × n complex matrices and M+
semidefinite n × n matrices. We write B > 0 if B ∈ Mn is positive definite.
n for the positive
Recall that for X, Y ∈ M+
n , the log-majorization X ≺log Y means that
λi(X) ≤
kYi=1
kYi=1
λi(Y ),
k = 1, . . . , n
with equality for k = n (i.e., det X = det Y ), where λ1(X) ≥ · · · ≥ λn(X) are the
eigenvalues of X in decreasing order counting multiplicities. As is well-known, X ≺log Y
i=1 λi(Y ) for k = 1, . . . , n.
The latter is equivalent to that kXk ≤ kY k holds for every unitarily invariant norm
k · k. See, e.g., [1, 7, 12, 18] for generalities on majorization theory for matrices.
implies the weak majorization X ≺w Y , i.e.,Pk
i=1 λi(X) ≤Pk
1E-mail address: [email protected]
1
Let α, z > 0 with α 6= 1, and let A, B ∈ M+
n with B > 0. In the present paper we
are concerned with the following two-variable matrix functions
Pα(A, B) := B1/2(B−1/2AB−1/2)αB1/2,
Qα,z(A, B) := (B
1−α
2z A
α
z B
1−α
2z )z.
Two special versions of Qα,z are
Qα(A, B) := Qα,1(A, B) = B
1−α
2 AαB
eQα(A, B) := Qα,α(A, B) = (B
1−α
2 ,
2α )α.
1−α
1−α
2α AB
Our motivation to consider these functions comes from different types of R´enyi diver-
gences that have recently been developed in quantum information. The conventional
(or standard) α-R´enyi divergence (due to Petz [21]) is
Dα(AkB) :=
1
α − 1
log
Tr Qα(A, B)
Tr A
,
the sandwiched α-R´enyi divergence [20, 24] is
eDα(AkB) :=
1
α − 1
log
TreQα(A, B)
Tr A
,
and the so-called α-z-R´enyi divergence [5] is
Dα,z(AkB) :=
1
α − 1
log
Tr Qα,z(A, B)
Tr A
.
In addition to Dα and eDα we define the maximal α-R´enyi divergence
Tr Pα(A, B)
1
log
α − 1
.
Tr A
bDα(AkB) :=
(For the term "maximal" here, see Remark 8.4.) See [13] and references therein for
more background information on quantum divergences.
We note that Pα(A, B) is a special case of operator perspective defined associated
with a function f on (0, ∞) by
Pf (A, B) := B1/2f (B−1/2AB−1/2)B1/2,
A, B ∈ M+
n , A, B > 0,
which was studied by Effros and Hansen [10] and others, with applications to quan-
tum information. Furthermore, note that when f is a non-negative operator monotone
function on (0, ∞) with f (1) = 1, Pf (A, B) is nothing but the operator mean B σf A
associated with f in the Kubo-Ando sense [17].
In particular, when 0 < α < 1,
Pα(A, B) = B #α A, the weighted geometric mean (first introduced by Pusz and
Woronowicz [22] in the case α = 1/2).
2
log-majorization [4] (see also [2]); indeed, Qα,z(A, B) is monotone decreasing in z > 0
The inequality eDα(AkB) ≤ Dα(AkB) is well-known as a consequence of Araki's
in the log-majorization order. However, the comparison between bDα and Dα,z (in
particular, Dα) has not fully been investigated so far, which motivate us to consider
the log-majorization between Pα and Qα,z.
In this paper we present the following
theorem which was announced without proofs in [13, Remark 4.6]:
Theorem 1.1. Let A, B ∈ M+
n with B > 0.
(1) For 0 < α < 1 and z > 0, Pα(A, B) ≺log Qα,z(A, B).
(2) For α > 1 and 0 < z ≤ min{α/2, α − 1}, Pα(A, B) ≺log Qα,z(A, B).
(3) For α > 1 and z ≥ max{α/2, α − 1}, Qα,z(A, B) ≺log Pα(A, B).
In particular, Pα(A, B) ≺log Qα(A, B) if 0 < α < 1 or α ≥ 2, and Qα(A, B) ≺log
Pα(A, B) if 1 < α ≤ 2.
The paper is organized as follows. In Sections 2 and 3 we prove Theorem 1.1. In
Section 4 we give an example showing that Theorem 1.1 is best possible with regard
to the assumptions on the parameters α, z, so that Theorem 1.1 is completed into
Theorem 4.1. In Section 5 we present the necessary and sufficient conditions on α, r, z
for which Pα,r(A, B) ≺log Qα,z(A, B) and Qα,z(A, B) ≺log Pα,r(A, B) hold, respectively,
where Pα,r(A, B) := Pα(A1/r, B1/r)r. Moreover, we give a log-majorization for Pα for
α ≥ 2, supplementing Ando-Hiai's log-majorization [2] for Pα for 0 < α < 1 and its
complementary version recently obtained by Kian and Seo [16] for Pα for 1 < α ≤ 2.
(Note that the negative power β ∈ [−1, 0) case in [16] can be rephrased into the case
of Pα for α = 1 − β ∈ (1, 2], see Section 5.) Applying our log-majorization results, in
Sections 6 and 7 we give norm inequalities for unitarily invariant norms and logarithmic
trace inequalities. The norm inequalities here improve those given in [16] and the
logarithmic trace inequalities here supplement those given in [2]. Finally in Section 8
we completely determine the parameters α, z for which bDα(AkB) ≤ Dα,z(AkB) and
Dα,z(AkB) ≤ bDα(AkB) hold, respectively.
2 Log-majorization (Part 1)
First, note that Araki's log-majorization [4] (see also [2]) implies that for every α > 0,
Qα,z ′(A, B) ≺log Qα,z(A, B)
if 0 < z ≤ z′.
(2.1)
The next proposition is an easy part of log-majorization results between Pα and Qα,z.
Proposition 2.1. Let A, B ∈ M+
n with B > 0.
3
(1) Assume that 0 < α < 1. Then for every z > 0,
Pα(A, B) ≺log Qα,z(A, B).
(2) Assume that 1 < α ≤ 2 and 0 < z ≤ α − 1. Then
Pα(A, B) ≺log Qα,z(A, B).
(3) Assume that α > 1 and z ≥ max{α/2, α − 1}. Then
Qα,z(A, B) ≺log Pα(A, B).
Proof. (1) Although this is an immediate consequence of well-known Araki's and Ando-
Hiai's log-majorization (see [4, 2]), we give a proof for the convenience of the reader.
By continuity we may assume that A > 0 as well as B > 0. From the Lie-Trotter
formula, letting z′ → ∞ in (2.1) gives
exp(α log A + (1 − α) log B) ≺log Qα,z(A, B),
z > 0.
(2.2)
On the other hand, when 0 < α < 1, the log-majorization in [2] says that
Pα(A, B) = B#αA ≺log (Bp#αAp)1/p,
0 < p < 1.
Letting p ց 0 and using [15, Lemma 3.3] we have
Pα(A, B) ≺log exp(α log A + (1 − α) log B).
(2.3)
Combining (2.2) and (2.3) implies the asserted log-majorization.
(2) By continuity we may assume that A > 0 as well as B > 0. The proof below is
an easy application of the standard anti-symmetric tensor power technique (see, e.g.,
[2]). To show that Pα(A, B) ≺log Qα,z(A, B), it suffices to prove that
kPα(A, B)k∞ ≤ kQα,z(A, B)k∞,
where k · k∞ denotes the operator norm. Due to the positive homogeneity in A, B of
both Pα and Qα,z (i.e., Pα(λA, λB) = λPα(A, B) for λ > 0 and similarly for Qα,z), it
also suffices to prove that
Qα,z(A, B) ≤ I =⇒ Pα(A, B) ≤ I.
(2.4)
Here recall the identity
(B−1/2AB−1/2)α = B−1/2A1/2(A1/2B−1A1/2)α−1A1/2B−1/2,
(2.5)
as seen from the well-known equality
Xf (X ∗X) = f (XX ∗)X
(2.6)
4
for every X ∈ Mn and every continuous function f on an interval containing the
eigenvalues of X ∗X (the proof is easy by approximating f by polynomials). Therefore,
Pα(A, B) = A1/2(A1/2B−1A1/2)α−1A1/2,
(2.7)
so that for (2.4) it suffices to prove that
1−α
2z A
α
z B
B
1−α
2z ≤ I =⇒ A1/2(A1/2B−1A1/2)α−1A1/2 ≤ I,
or equivalently,
α
z ≤ B
A
α−1
z
=⇒ (A1/2B−1A1/2)α−1 ≤ A−1.
(2.8)
Now, assume that 1 < α ≤ 2 and 0 < z ≤ α − 1, and that A
0 < z/(α − 1) ≤ 1,
α
z ≤ B
α−1
z . Since
B−1 = (B
and hence
α−1
z )− z
α−1 ≤ (A
α
z )− z
α−1 = A− α
α−1 ,
A1/2B−1A1/2 ≤ A1/2A− α
α−1 A1/2 = A− 1
α−1 .
Since 0 < α − 1 ≤ 1, we have
(A1/2B−1A1/2)α−1 ≤ (A− 1
α−1 )α−1 = A−1,
proving (2.8).
(3) As in the proof of (2) we may assume that both A, B are positive definite, and
prove the implication opposite to (2.4). In the present case, similarly to the above, it
suffices to prove that
(B−1/2AB−1/2)α ≤ B−1 =⇒ A
α
z ≤ B
α−1
z ,
or letting C := B−1/2AB−1/2 > 0, we may prove that
C α ≤ B−1 =⇒ (B1/2CB1/2)
α
z ≤ B
α−1
z .
(2.9)
Now, assume that α > 1 and z ≥ max{α/2, α − 1}. Note by (2.1) that if once
Qα,z(A, B) ≺log Pα(A, B) holds for z = z0 with some z0 > 0, then the same does for
all z ≥ z0. Hence we may further assume that z ≤ α. If C α ≤ B−1, then B ≤ C −α so
that C 1/2BC 1/2 ≤ C 1−α. Since 0 ≤ α
z − 1 ≤ 1, we have
(C 1/2BC 1/2)
α
z −1 ≤ (C 1−α)
Since by (2.5),
α
z −1 = C (1−α)( α
z −1).
(B1/2CB1/2)
α
z = B1/2C 1/2(C 1/2BC 1/2)
α
z −1C 1/2B1/2,
5
we have
(B1/2CB1/2)
α
z ≤ B1/2C 1+(1−α)( α
z −1)B1/2 = B1/2(C α)1− α−1
z B1/2.
Since the assumption on α, z implies that 0 ≤ 1 − α−1
z < 1, we have
(B1/2CB1/2)
proving (2.9).
α
z ≤ B1/2(B−1)1− α−1
z B1/2 = B
α−1
z
,
It should be noted that the above proofs of (2.8) and (2.9) are more or less similar
to that of [23, Theorem 3] for the Furuta inequality with negative powers.
Remark 2.2. When α = 2, since P2(A, B) = AB−1A is unitarily equivalent to
B−1/2A2B−1/2, P2(A, B) ≺log Q2,z(A, B) is equivalent to
B−1/2A2B−1/2 ≺log (B− 1
2z A
2
z B− 1
2z )z.
Assume that AB 6= BA. Then from Araki's log-majorization [4] and [11, Theorem 2.1]
we see that the above log-majorization holds true if and only if 0 < z ≤ 1. Similarly,
Q2,z(A, B) ≺log P2(A, B) holds if and only if z ≥ 1. These are of course consistent with
(2) and (3) of Proposition 2.1.
In particular, when A is a projection, we have:
Proposition 2.3. Let α > 1 and z > 0. Assume that E, B ∈ M+
a projection, B > 0 and EB 6= BE. Then:
n are such that E is
(a) Pα(E, B) ≺log Qα,z(E, B) if and only if z ≤ α − 1.
(b) Qα,z(E, B) ≺log Pα(E, B) if and only if z ≥ α − 1.
Proof. We write
Pα(E, B) = B1/2(B−1/2EB−1/2)αB1/2
= B1/2B−1/2E(EB−1E)α−1EB−1/2B1/2 = (EB−1E)α−1.
On the other hand, Qα,z(E, B) = (B
Hence Pα(E, B) ≺log Qα,z(E, B) is equivalent to
2z EB
1−α
1−α
2z )z is unitarily equivalent to (EB
1−α
2z E)z.
(EB−1E)α−1 ≺log (EB
1−α
2z E)z.
From [11, Theorem 2.1] we see that this holds if and only if α − 1 ≥ z. Similarly,
Qα,z(E, B) ≺log Pα(E, B) holds if and only if z ≥ α − 1.
6
3 Log-majorization (Part 2)
Our final goal is to completely determine the regions of {(α, z) : α, z > 0} for which
Pα ≺log Qα,z holds, or Qα,z ≺log Pα holds, or neither holds true, respectively. The next
step to the goal is to find a region in α ≥ 2 where Pα ≺log Qα,z holds true. Since
P2 ≺log Q2,z holds if and only if 0 < z ≤ 1 (see Remark 2.2), it would be reasonable
to conjecture that there is a region in α ≥ 2 touching {(2, z) : 0 < z ≤ 1} where
Pα ≺log Qα,z holds.
We show the next log-majorization result by elaborating the anti-symmetric tensor
power technique. The proof reveals essentially similar features to those of [2, Theorem
4.1] and [3, Theorem 2.1].
Proposition 3.1. Assume that α ≥ 2 and 0 < z ≤ α/2. Then for every A, B ∈ M+
n
with B > 0,
Pα(A, B) ≺log Qα,z(A, B).
Proof. By continuity we may assume that both A, B are positive definite. Assume that
α ≥ 2 and 0 < z ≤ α/2. Due to the anti-symmetric tensor power technique and the
positive homogeneity in A, B of Pα and Qα,z, it suffices to prove that
α
z ≤ B
A
α−1
z
=⇒ Pα(A, B) ≤ I.
So assume that A
2m ≤ α ≤ 2m + 1 for some m ∈ N, so write α = 2m + λ with 0 ≤ λ ≤ 1. Note that
z . We divide the proof into two cases. First, assume that
z ≤ B
α
α−1
Pα(A, B) = B1/2(B−1/2AB−1/2)m(B−1/2AB−1/2)λ(B−1/2AB−1/2)mB1/2
= (AB−1)m(B #λ A)(B−1A)m.
Since 0 < z/α ≤ 1/2, we have A ≤ B
α−1
α and hence
Therefore,
Since
B #λ A ≤ B #λ B
α−1
α = B1−λB
(α−1)λ
α = B
α−λ
α .
Pα(A, B) ≤ (AB−1)m−1AB− α+λ
α A(B−1A)m−1.
(α + λ)z
α(α − 1)
≤
α + λ
2(α − 1)
≤ 1
thanks to α = 2m + λ ≥ 2 + λ, we have
B− α+λ
α = (B
α−1
z )− (α+λ)z
α(α−1) ≤ (A
α
z )− (α+λ)z
α(α−1) = A− α+λ
α−1 ,
so that
Pα(A, B) ≤ (AB−1)m−1A
α−2−λ
α−1 (B−1A)m−1.
7
Since
we have
so that
(α − 2 − λ)z
α(α − 1)
≤
α − 2 − λ
2(α − 1)
≤ 1,
α−2−λ
α−1 = (A
α
z )
A
(α−2−λ)z
α(α−1) ≤ (B
α−1
z )
(α−2−λ)z
α(α−1) = B
α−2−λ
α
,
Pα(A, B) ≤ (AB−1)m−2AB− α+2+λ
α A(B−1A)m−2.
The above argument can be repeated to see that for k = 0, 1, . . . , m − 1,
Pα(A, B) ≤ (AB−1)m−1−kAB− α+2k+λ
α A(B−1A)m−1−k,
and hence
Pα(A, B) ≤ AB− α+2m−2+λ
α
A = AB− 2(α−1)
α A.
Finally, since 2z/α ≤ 1, we have
B− 2(α−1)
α ≤ (B
so that Pα(A, B) ≤ I.
α−1
z )− 2z
α ≤ (A
α
z )− 2z
α = A−2,
Secondly, assume that 2m+1 < α < 2m+2 for some m ∈ N, so write α = 2m+2−λ
with 0 < λ < 1. Note that
Pα(A, B) = B1/2(B−1/2AB−1/2)m+1(B−1/2AB−1/2)−λ(B−1/2AB−1/2)m+1B1/2
= (AB−1)mAB−1/2(B−1/2AB−1/2)−λB−1/2A(B−1A)m
= (AB−1)mA(B #λ A)−1A(B−1A)m.
Since 0 ≤ z/(α − 1) ≤ 1, we have B ≥ A
α
α−1 so that
Therefore,
Since
we have
so that
Since
(B #λ A)−1 ≤(cid:0)A
α
α−1 #λ A(cid:1)−1 =(cid:0)A
α(1−λ)
α−1 Aλ(cid:1)−1 = A− α−λ
α−1 .
Pα(A, B) ≤ (AB−1)mA
α−2+λ
α−1 (B−1A)m.
(α − 2 + λ)z
α(α − 1)
≤
α − 2 + λ
2(α − 1)
≤ 1,
α−2+λ
α−1 = (A
α
z )
A
(α−2+λ)z
α(α−1) ≤ (B
α−1
z )
(α−2+λ)z
α(α−1) = B
α−2+λ
α
,
Pα(A, B) ≤ (AB−1)m−1AB− α+2−λ
α A(B−1A)m−1.
(α + 2 − λ)z
α(α − 1)
≤
α + 2 − λ
2(α − 1)
≤ 1,
8
we have
so that
B− α+2−λ
α = (B
α−1
z )− (α+2−λ)z
α(α−1) ≤ (A
α
z )− (α+2−λ)z
α(α−1) = A− α+2−λ
α−1 ,
Pα(A, B) ≤ (AB−1)m−1A
α−4+λ
α−1 (B−1A)m−1.
Repeating the above argument we have
Pα(A, B) ≤ (AB−1)A
α−2m+λ
α−1
(B−1A) = AB−1A
2
α−1 B−1A.
2
α−1 = (A
α
z )
A
2z
α(α−1) ≤ (B
α−1
z )
2z
α(α−1) = B
2
α
Since
and
α = (B
we finally have Pα(A, B) ≤ AB− 2(α−1)
B− 2(α−1)
α A ≤ I.
α−1
z )− 2z
α ≤ (A
α
z )− 2z
α = A−2,
Now, Theorem 1.1 stated in the Introduction is proved from the log-majorization
results between Pα and Qα,z obtained so far in Propositions 2.1 and 3.1.
We note that some discussions involving Pα, Qα = Qα,1 and eQα = Qα,α were recently
given in [8, Sect. 5].
4 Main theorem
In this section we prove that Theorem 1.1 is best possible with regard to the assump-
tions on the parameters α, z.
Assume that α > 1. For each x, y > 0 and θ ∈ R define 2 × 2 positive definite
matrices
We write
sin θ
cos θ (cid:21)(cid:20)1 0
0 x(cid:21)(cid:20) cos θ
Aθ :=(cid:20)cos θ − sin θ
B−1/2AθB−1/2 =(cid:20)1 + (x − 1) sin2 θ
(1 − x)y−1/2 sin 2θ
2
− sin θ cos θ(cid:21) ,
sin θ
0 y(cid:21) .
B :=(cid:20)1 0
xy−1 + (1 − x)y−1 sin2 θ(cid:21)
(1 − x)y−1/2 sin 2θ
2
= G + θH + θ2K + o(θ2),
(4.1)
where
G :=(cid:20)1
0 xy−1(cid:21) , H :=(cid:20)
0
0
(1 − x)y−1/2
(1 − x)y−1/2
0
(cid:21) , K :=(cid:20)x − 1
0
0
(1 − x)y−1(cid:21) ,
and o(θ2) denotes a small value such that o(θ2)/θ2 → 0 as θ → 0. We apply the Taylor
formula with Fr´echet derivatives (see e.g., [11, Theorem 2.3.1]) to obtain
(B−1/2AθB−1/2)α = Gα + D(xα)(G)(θH + θ2K) +
1
2
D2(xα)(G)(θH, θH) + o(θ2),
9
where the second and the third terms in the right-hand side are the first and the second
Fr´echet derivatives of X 7→ X α (X ∈ M+
2 , X > 0) at G, respectively. By Daleckii and
Krein's derivative formula (see [7, Theorem V.3.3], [11, Theorem 2.3.1]) we have
D(xα)(G)(θH + θ2K)
(xα)[1](1, xy−1)
α
=(cid:20) (xα)[1](1, 1)
="
= θ"
1−(xy−1)α
1−xy−1
0
xα−yα
x−y (1 − x)y
(xα)[1](1, xy−1)
(xα)[1](xy−1, xy−1)(cid:21) ◦ (θH + θ2K)
1−(xy−1)α
α(xy−1)α−1# ◦ (θH + θ2K)
1−xy−1
xα−yα
x−y (1 − x)y
1
2 −α
0
1
2 −α
# + θ2(cid:20)α(x − 1)
0
α(1 − x)xα−1y−α(cid:21) ,
0
where (xα)[1] denotes the first divided difference of xα and ◦ means the Schur (or
Hadamard) product. For the second divided difference of xα we compute
(xα)[2](1, 1, xy−1) =
α − 1 − αxy−1 + xαy−α
(1 − xy−1)2
=
y{(α − 1)y − αx + xαy1−α}
(x − y)2
,
(xα)[2](1, xy−1, xy−1) =
=
1 − αxα−1y1−α + (α − 1)xαy−α
(1 − xy−1)2
y{y − αxα−1y2−α + (α − 1)xαy1−α}
(x − y)2
,
and hence we have
D2(xα)(G)(θH, θH) = θ2" (x−1)2{(α−1)y−αx+xαy1−α}
(x−y)2
0
1
2
0
(x−1)2{y−αxα−1y2−α+(α−1)xαy1−α}
(x−y)2
# .
(In the above computation we have assumed that x 6= y.) Therefore, it follows that
(B−1/2AθB−1/2)α ="1 + s(1)
s(3)
α θ
α θ2
s(3)
α θ
xαy−α + s(2)
α θ2# + o(θ2),
where
s(1)
α := α(x − 1) +
(x − 1)2{(α − 1)y − αx + xαy1−α}
(x − y)2
,
s(2)
α := αxα−1(1 − x)y−α +
(x − 1)2{y − αxα−1y2−α + (α − 1)xαy1−α}
(x − y)2
.
(The form of s(3)
below.) We hence arrive at
α is not written down here since it is unnecessary in the computation
Tr Pα(Aθ, B) = 1 + xαy1−α +(cid:0)s(1)
α + s(2)
α y(cid:1)θ2 + o(θ2).
10
(4.2)
Next, we write
1−α
2z A
B
α
z
θ B
1−α
(1 − x
z ="1 + (x
="1 + (x
(1 − x
α
z − 1) sin2 θ
1−α
2z sin 2θ
z )y
α
2
α
(1 − x
z )y
z + (1 − x
1−α
α
z y
x
α
z − 1)θ2
z )y
1−α
α
2z θ x
α
z y
α
1−α
(1 − x
z )y
z + (1 − x
2z θ
z )y
α
1−α
z )y
1−α
2z sin 2θ
2
1−α
α
z sin2 θ#
z θ2# + o(θ2).
1−α
Since
det(cid:16)tI − B
1−α
2z A
α
z
θ B
the eigenvalues of B
1−α
z (cid:17) = t2 −(cid:8)1 + x
α
1−α
1−α
2z A
z
θ B
z are
α
z y
1−α
z + (x
α
z − 1)(1 − y
1−α
z )θ2(cid:9)t + x
α
z y
1−α
z + o(θ2),
t±
α,z,θ =
1−α
z + (x
α
z − 1)(1 − y
1−α
z )θ2
α
z y
1−α
z )2 + 2(x
α
z − 1)(1 − y
1−α
z )(1 + x
α
z y
1−α
z )θ2(cid:21) + o(θ2).
α
1
z y
2(cid:20)1 + x
±q(1 − x
Assuming that 1 − x
x > 0), we have
α
z y
1−α
z > 0 (this is the case when we let y → ∞ for any fixed
α
1
z y
2(cid:20)1 + x
±(cid:26)1 − x
t±
α,z,θ =
so that
1−α
z + (x
α
z − 1)(1 − y
1−α
z )θ2
α
z y
1−α
z +
(x
α
z − 1)(1 − y
1−α
z )(1 + x
α
z y
1−α
z )
1 − x
1−α
z
α
z y
θ2(cid:27)(cid:21) + o(θ2),
t+
α,z,θ = 1 +
t−
α,z,θ = x
α
z y
Therefore, we have
(x
α
z − 1)(1 − y
1−α
z )
1 − x
α
z y
1−α
z
θ2 + o(θ2),
1−α
z (cid:26)1 −
(x
α
z − 1)(1 − y
1−α
z )
1 − x
1−α
z
α
z y
θ2(cid:27) + o(θ2).
Tr Qα,z(Aθ, B) = (t+
α,z,θ)z + (t−
α,z,θ)z
= 1 + z
(x
α
z − 1)(1 − y
1−α
z )
θ2
α
1 − x
z y
1−α
z
+ xαy1−α(cid:26)1 − z
= 1 + xαy1−α + z
(x
(x
α
z − 1)(1 − y
1−α
z )
α
1−α
z
1 − x
z y
z − 1)(1 − y
α
θ2(cid:27) + o(θ2)
1−α
z )(1 − xαy1−α)
z y
1−α
z
α
1 − x
11
θ2 + o(θ2).
(4.3)
Now, suppose that Qα,z(Aθ, B) ≺log Pα(Aθ, B) holds for all θ 6= 0. Then we
must have Tr Qα,z(Aθ, B) ≤ Tr Pα(Aθ, B).
(Since det Qα,z(Aθ, B) = det Pα(Aθ, B),
Qα,z(Aθ, B) ≺log Pα(Aθ, B) is indeed equivalent to Tr Qα,z(Aθ, B) ≤ Tr Pα(Aθ, B) in
the 2 × 2 case here.) So by (4.2) and (4.3) it follows that
z
(x
α
z − 1)(1 − y
1−α
z )(1 − xαy1−α)
z y
1−α
z
α
1 − x
≤ s(1)
α + s(2)
α y.
For any x > 0, let y → ∞; then the above left-hand side converges to z(x
α → α(x − 1) and s(2)
s(1)
x > 0,
z − 1), while
α → (x − 1)2 thanks to α > 1. Hence we must have for every
α
z(x
α
z − 1) ≤ α(x − 1) + (x − 1)2.
Letting x ց 0 gives −z ≤ −α + 1, i.e., z ≥ α − 1. Moreover, for any x > 1,
α
z − 1
x
x − 1
z
≤ x + α − 1,
which holds true only when α/z ≤ 2, i.e., z ≥ α/2. On the other hand, suppose that
Pα(Aθ, B) ≺log Qα,z(Aθ, B) holds for all θ 6= 0. Then, similarly to the above case,
z ≤ α − 1 and z ≤ α/2 must follow.
Thus, combining the above discussions with Theorem 1.1 proves our main theorem
as follows:
Theorem 4.1. Let α, z > 0 with α 6= 1.
(a) The following conditions are equivalent:
(i) Pα(A, B) ≺log Qα,z(A, B) for every A, B ∈ M+
(ii) Tr Pα(A, B) ≤ Tr Qα,z(A, B) for every A, B ∈ M+
(iii) Pα(A, B) ≺log Qα,z(A, B) for every A, B ∈ M+
n , n ∈ N, with B > 0;
n , n ∈ N, with B > 0;
2 with A, B > 0;
(iv) either 0 < α < 1 and z > 0 is arbitrary, or α > 1 and 0 < z ≤ min{α/2, α −
1}.
(b) The following conditions are equivalent:
(i)′ Qα,z(A, B) ≺log Pα(A, B) for every A, B ∈ M+
(ii)′ Tr Qα,z(A, B) ≤ Tr Pα(A, B) for every A, B ∈ M+
(iii)′ Qα,z(A, B) ≺log Pα(A, B) for every A, B ∈ M+
(iv)′ α > 1 and z ≥ max{α/2, α − 1}.
n , n ∈ N, with B > 0;
n , n ∈ N, with B > 0;
2 with A, B > 0;
The theorem says that neither Pα(A, B) ≺log Qα,z(A, B) nor Qα,z(A, B) ≺log Pα(A, B)
holds in general in the regions of 1 < α < 2 and α − 1 < z < α/2 and of α > 2 and
α/2 < z < α − 1.
12
5 Further extension
For A, B ∈ M+
account, we may define the two-parameter extension of Pα as
n with B > 0, taking the expression Qα,z(A, B) = Qα(A1/z, B1/z)z into
Pα,r(A, B) := Pα(A1/r, B1/r)r = {B1/2r(B−1/2rA1/rB−1/2r)αB1/2r}r,
α, r > 0.
The log-majorization in [2] says that when 0 < α < 1,
Pα,r(A, B) ≺log Pα,r′(A, B)
if 0 < r ≤ r′.
(5.1)
For every A, B > 0 and α > 0, note by (2.7) that Pα(A, B) = A1/2(A−1/2BA−1/2)βA1/2
where β := 1 − α (the right-hand side is often denoted by A♮βB when β 6∈ [0, 1] instead
of A#βB for β ∈ [0, 1]). Thus, the log-majorization recently obtained in [16, Theorem
3.1] is rephrased as follows: When 1 < α ≤ 2, for every A, B ∈ M+
n with B > 0,
Pα,r(A, B) ≺log Pα,r′(A, B)
if 0 < r′ ≤ r.
(5.2)
In particular, when α = 2, this reduces to Araki's log-majorization (see Remark 2.2).
For each α, r, z > 0 with α 6= 1, since it is easy to see that Pα,r(A, B) ≺log Qα,z(A, B)
(resp., Qα,z(A, B) ≺log Pα,r(A, B)) for every A, B ∈ M+
n with B > 0 if and only if
Pα(A, B) ≺log Qα,z/r(A, B) (resp., Qα,z/r(A, B) ≺log Pα(A, B)) for every A, B ∈ M+
n
with B > 0. Thus, we can extend Theorem 4.1 in the following way:
Proposition 5.1. Let α, r, z > 0 with α 6= 1
(a) The following conditions are equivalent:
(i) Pα,r(A, B) ≺log Qα,z(A, B) for every A, B ∈ M+
n , n ∈ N, with B > 0;
(ii) either 0 < α < 1 and r, z > 0 are arbitrary, or α > 1 and 0 < z/r ≤
min{α/2, α − 1}.
(b) The following conditions are equivalent:
(i)′ Qα,z(A, B) ≺log Pα,r(A, B) for every A, B ∈ M+
(ii)′ α > 1 and z/r ≥ max{α/2, α − 1}.
n , n ∈ N, with B > 0;
Although Proposition 5.1 is just a slight modification of Theorem 4.1, it can be used
to show the following log-majorization supplementary to (5.1) and (5.2):
Corollary 5.2. Assume that α ≥ 2. For every A, B ∈ M+
n with B > 0,
Pα,r(A, B) ≺log Pα,r′(A, B)
if 0 < r′ ≤ α
2(α−1) r.
Hence Pα,r(A, B) ≺log Pα,r′(A, B) for all α ≥ 2 if 0 < r′ ≤ r/2.
13
Proof. Let α ≥ 2. By Proposition 5.1 (a) we have
Pα,r(A, B) ≺log Qα,rα/2(A, B).
Since (rα/2)/r′ ≥ α − 1, Proposition 5.1 (b) implies that
Qα,rα/2(A, B) ≺log Pα,r′(A, B),
so that the asserted log-majorization follows.
Problem 5.3. Although the assumption β = 1 − α ∈ [−1, 0) (or 1 < α ≤ 2) seems
essential in the proof of (5.2) in [16], it is unknown whether (5.2) holds true even
for α > 2 (i.e., the bound α/2(α − 1) in the corollary can be removed) or not. For
example, when α = m + 1 ∈ N with m ∈ N, m ≥ 2, noting Pm+1(A, B) = (AB−1)mA
and replacing B−1 with B and 1/r with r, (5.2) is equivalent to the following extended
Araki's log-majorization for every A, B ∈ M+
n :
((AB)mA)r ≺log (ArBr)mAr
if r ≥ 1,
(5.3)
which seems difficult to hold in general, while no counter-example is at the moment
known to us. But Corollary 5.2 implies that ((AB)mA)r ≺log (ArBr)mAr for r ≥
2m/(m + 1). Here is a simple argument when m = 2. For m = 2, to prove (5.3),
it suffices to show that for 0 < p ≤ 1, ABABA ≤ I =⇒ ApBpApBpAp ≤ I.
Assume the left-hand inequality, i.e., (A1/2BA1/2)2 ≤ A−1; then B ≤ A−3/2 and so
Ap ≤ B−2p/3. Hence ApBpApBpAp ≤ ApB4p/3Ap. If p ≤ 3/4, then B4p/3 ≤ A−2p and
so ApBpApBpAp ≤ I. Therefore, (ABABA) ≺log ArBrArBrAr if r ≥ 4/3, which is
just the case α = 3 of the corollary. The same argument works well when α = m + 1
for any m ∈ N, m ≥ 2, proving directly the α = m + 1 case of the corollary.
6 Norm inequalities and their equality cases
A norm k·k on Mn is said to be unitarily invariant if kUXV k = kXk for all X ∈ Mn and
all unitaries U, V ∈ Mn. We say (see [11]) that a unitarily invariant norm k · k is strictly
increasing if for X, Y ∈ M+
n , X ≤ Y and kXk = kY k imply X = Y . For example, the
Schatten p-norm kXkp := (Tr Xp)1/p is strictly increasing when 1 ≤ p < ∞.
Theorem 1.1 implies the following:
Corollary 6.1. Let A, B ∈ M+
Mn.
n with B > 0 and k · k be a unitarily invariant norm on
(1) If 0 < α < 1, then kPα(A, B)k ≤ kQα,z(A, B)k for all z > 0.
(2) If α > 1 and 0 < z ≤ min{α/2, α − 1}, then kPα(A, B)k ≤ kQα,z(A, B)k.
14
(3) If α > 1 and z ≥ max{α/2, α − 1}, then kQα,z(A, B)k ≤ kPα(A, B)k.
Remark 6.2. The norm inequalities with negative power β in [16, Theorem 4.4] can
be rephrased as follows (by letting α = 1 − β): When A, B > 0, for every unitarily
invariant norm,
kQα,z(B, A)k ≤ kPα(B, A)k
kPα(B, A)k ≤ kQα,1/2(B, A)k ≤ kQα,z(B, A)k
if α ∈ (1, 2], z ≥ 2,
if α ∈ [3/2, 2], 0 < z ≤ 1/2.
These inequalities are indeed included in (2) and (3) of Corollary 6.1 (and (2.1)).
Lemma 6.3. Assume that α > 0 and α 6= 1. Let k · k be a strictly increasing unitarily
invariant norm on Mn. If kQα,z(A, B)k = kQα,z ′(A, B)k for some z, z′ > 0 with z 6= z′,
then AB = BA.
Proof. By [11, Theorem 2.1] the assumed norm equality implies that Aα and B1−α
commute and hence AB = BA.
Concerning the equality cases of the inequalities in Corollary 6.1 we have:
Proposition 6.4. Let k · k be a strictly increasing unitarily invariant norm on Mn.
Then we have AB = BA if kPα(A, B)k = kQα,z(A, B)k for some α, z satisfying one of
the following:
(1) 0 < α < 1 and z > 0,
(2) α > 1 and 0 < z < min{α/2, α − 1},
(3) α > 1 and z > max{α/2, α − 1}.
Proof. (1) Assume that kPα(A, B)k = kQα,z(A, B)k for some α, z in (1). Choose
z′ > z. By (2.1) and Corollary 6.1 (1) we have
kPα(A, B)k = kQα,z(A, B)k ≥ kQα,z ′(A, B)k ≥ kPα(A, B)k,
implying AB = BA by Lemma 6.3.
(2) Assume that kPα(A, B)k = kQα,z(A, B)k for some α, z in (2). Choose z′ with
z < z′ < min{α/2, α − 1}. By (2.1) and Corollary 6.1 (2),
kPα(A, B)k = kQα,z(A, B)k ≥ kQα,z ′(A, B)k ≥ kPα(A, B)k,
implying AB = BA by Lemma 6.3.
(3) Assume that kPα(A, B)k = kQα,z(A, B)k for some α, z in (3). Choose z′ with
z > z′ > max{α/2, α − 1}. Then AB = BA follows similarly to the proof for (2).
15
7 Logarithmic trace inequalities
For every p > 0 and every A, B ∈ M+
n with B > 0, the logarithmic trace inequalities
1
p
Tr A log(B−p/2ApB−p/2) ≤ Tr A(log A − log B)
≤
1
p
Tr A log(Ap/2B−pAp/2).
(7.1)
were shown in [15], and supplementary logarithmic trace inequalities were also in [2]. In
particular, the latter inequality for p = 1 was first proved in [14], giving the comparison
between the Umegaki relative entropy and the Belavkin-Staszewski relative entropy [6]
(see Remark 8.5 below). Recall that this can readily be verified by taking the derivatives
at α = 1 of Tr Pα(A, B) = Tr A(A1/2B−1A1/2)α−1 and Tr Qα(A, B) = Tr AαB1−α from
Corollary 6.1. By the derivatives at α = 2 we have more logarithmic trace inequalities
in the following:
Proposition 7.1. For every A, B ∈ M+
n with B > 0,
Tr AB−1A(log A − log B) ≤ Tr A1/2B−1A3/2 log(A1/2B−1A1/2)
= Tr B−1/2A2B−1/2 log(B−1/2AB−1/2)
= Tr A3/2B−1A1/2 log(A1/2B−1A1/2)
≤ Tr A2B−1(log A − log B).
(7.2)
Proof. To prove the inequalities and the equalities above, we may assume by continuity
that A > 0 as well as B > 0. The inequalities in the middle of (7.2) are easily verified
as
Tr A1/2B−1A3/2 log(A1/2B−1A1/2) = Tr A log(A1/2B−1A1/2) · A1/2B−1A1/2
= Tr A3/2B−1A1/2 log(A1/2B−1A1/2)
= Tr A3/2B−1/2 log(B−1/2AB−1/2) · B−1/2A1/2
= Tr B−1/2A2B−1/2 log(B−1/2AB−1/2),
where we have used (2.6) for the third equality. To prove the inequalities, we use
Corollary 6.1 (2) for z = 1 to have
Tr Pα(A, B) ≤ Tr Qα(A, B)
for α ≥ 2.
Since Tr P2(A, B) = Tr A2B−1 = Tr Q2(A, B), if follows that
d
dα
The left-hand side of (7.3) is
Tr Pα(A, B)(cid:12)(cid:12)(cid:12)(cid:12)α=2
≤
d
dα
Tr Qα(A, B)(cid:12)(cid:12)(cid:12)(cid:12)α=2
.
(7.3)
Tr B(B−1/2AB−1/2)2 log(B−1/2AB−1/2)
16
= Tr B1/2AB−1AB−/2 log(B−1/2AB−1/2)
= Tr B1/2AB−1A1/2 log(A1/2B−1A1/2) · A1/2B−1/2
= Tr A3/2B−1A1/2 log(A1/2B−1A1/2,
where we have used (2.6) again for the second equality. On the other hand, the right-
hand side of (7.3) is
Tr A2 log A · B−1 − Tr A2B−1 log B = Tr A2B−1(log A − log B).
Hence the latter inequality in (7.2) follows.
Next, set C := B−1/2AB−1/2 so that A = B1/2CB1/2. Then
Tr B−1/2A2B−1/2 log(B−1/2AB−1/2) = Tr CBC log C
and
Tr A2B−1(log A − log B)
= Tr B1/2CBCB−1/2(cid:0)log(B1/2CB1/2) − log B(cid:1)
= Tr C 1/2BCB−1/2 log(B1/2CB1/2) · B1/2C 1/2 − Tr CBC log B
= Tr C 1/2BC 3/2 log(C 1/2BC 1/2) − Tr CBC log B.
by (2.6) once again. Hence the latter inequality in (7.2) is rephrased as
Tr CBC(log C + log B) ≤ Tr C 1/2BC 3/2 log(C 1/2BC 1/2).
Replacing C, B with A, B−1, respectively, we have the first inequality in (7.2).
Remark 7.2. It is obvious that if A, B are commuting, then all the inequalities of
(7.1) and (7.2) become equality. In the converse direction, it is seen from [2, Theorem
5.1] and [11, Theorem 4.1] that the equality case of the second inequality of (7.1) (for
some p > 0) implies AB = BA. Here we note that if equality holds in both inequalities
of (7.2) then AB = BA. Indeed, the inequality between both ends of (7.2) means that
Tr AB−1A log B−1 ≤ Tr A2B−1 log B−1,
(7.4)
which is considered as a kind of so-called gathering inequalities (see, e.g., [9] and [3]).
To prove that the equality case of (7.4) implies AB = BA, we may assume that B is
diagonal, so B−1 = diag(λ1, . . . , λn). Then for A = [aij]n
i,j=1, equality in (7.4) means
that
nXi,j=1
nXi,j=1
nXi,j=1
17
which is rewritten as
aij2λi log λj =
aij2λi log λi,
aij2(λi − λj)(log λi − log λj) = 0.
Since (λi − λj)(log λi − log λj) > 0 when λi 6= λj, we must have aij = 0 for all i, j with
λi 6= λj, implying AB = BA.
We may naturally conjecture that if either inequality of (7.2) holds with equality
then AB = BA.
8 Applications to R´enyi divergences
In this section we apply our log-majorization results to the relations between R´enyi
type divergences Dα, eDα, bDα and Dα,z defined in the Introduction.
The equivalences (ii) ⇐⇒ (iv) and (ii)′ ⇐⇒ (iv)′ of Theorem 4.1 immediately yield
the following:
Corollary 8.1. Let α, z > 0 with α 6= 1.
(a) The following conditions are equivalent:
(i) bDα(AkB) ≤ Dα,z(AkB) for every A, B ∈ M+
B > 0;
(ii) α > 1 and z ≤ min{α/2, α − 1}.
n , n ∈ N, with A 6= 0 and
(b) The following conditions are equivalent:
(i)′ Dα,z(AkB) ≤ bDα(AkB) for every A, B ∈ M+
B > 0;
(ii)′ either 0 < α < 1 and z > 0 is arbitrary, or α > 1 and z ≥ max{α/2, α − 1}.
n , n ∈ N, with A 6= 0 and
Moreover, specializing Corollary 6.1 to z = 1, α and the trace-norm, we have:
Corollary 8.2. Let A, B ∈ M+
n with A 6= 0 and B > 0. If 0 < α ≤ 2 and α 6= 1 then
(8.1)
and if α ≥ 2 then
eDα(AkB) ≤ Dα(AkB) ≤ bDα(AkB),
eDα(AkB) ≤ bDα(AkB) ≤ Dα(AkB).
Corollary 8.3. Let A, B ∈ M+
n with A 6= 0 and B > 0. If some two of Dα(AkB),
Tr Pα. Also, note that if 1 < α < 2 then α > 1 > max{α/2, α − 1}, and if α > 2
then 1 < min{α/2, α − 1} and α > max{α/2, α − 1}. Hence by Proposition 6.4,
eDα(AkB), and bDα(AkB) are equal for some α ∈ (0, ∞) \ {1, 2}, then AB = BA.
Proof. Note that Dα = bDα means Tr Qα,1 = Tr Pα, and eDα = bDα means Tr Qα,α =
either equality of Dα(AkB) = bDα(AkB) or eDα(AkB) = bDα(AkB) implies AB = BA.
Furthermore, Dα(AkB) = eDα(AkB) implies AB = BA by Lemma 6.3.
18
Remark 8.4. In [13] we studied the standard f -divergence Sf (AkB) and the maximal
f -divergence bSf (AkB), which are defined as
Sf (AkB) := Tr B1/2f (LARB−1)(B1/2),
bSf (AkB) := Tr Pf (A, B)
n , A, B > 0 (and extended to general A, B ∈ M+
for A, B ∈ M+
n by convergences), where
LA is the left multiplication on Mn by A and RB−1 is the right multiplication by B−1.
It is known [13, Proposition 4.1] (see also [19]) that
Sf (AkB) ≤ bSf (AkB)
holds whenever f is an operator convex function on (0, ∞). When f (x) = −xα for
0 < α < 1 or f (x) = xα for 1 < α ≤ 2, this becomes the second inequality of (8.1).
Corollaries 8.2 and 8.3 say that this is no longer true if f is a general convex function
on (0, ∞). Furthermore, a special case of [13, Theorem 4.3] says that Dα(AkB) =
bDα(AkB) for some α ∈ (0, 2) \ {1} implies AB = BA, which is included in Corollary
8.3.
Remark 8.5. Let A, B ∈ M+
readily verified) that
n with A 6= 0 and B > 0 as above. It is well-known (and
lim
α→1
Dα(AkB) = D1(AkB) :=
D(AkB)
Tr A
,
where D(AkB) := Tr A(log A − log B), the Umegaki relative entropy. It is also known
[20] that
On the other hand, we note that
lim
α→1eDα(AkB) = D1(AkB).
lim
α→1bDα(AkB) =
1
Tr A
Tr B1/2AB−1/2 log(B−1/2AB−1/2) =
DBS(AkB)
Tr A
,
where DBS(AkB) := Tr A log(A1/2B−1A1/2), the Belavkin-Staszewski relative entropy
[6] (see also [13, Example 4.4]). By Corollary 8.2 we have D(AkB) ≤ DBS(AkB), which
was first obtained in [14].
Acknowledgments
This work was supported by JSPS KAKENHI Grant Number JP17K05266.
References
[1] T. Ando, Majorization, doubly stochastic matrices, and comparison of eigenvalues,
Linear Algebra Appl. 118 (1989), 163 -- 248.
19
[2] T. Ando and F. Hiai, Log majorization and complementary Golden-Thompson
type inequalities, Linear Algebra Appl. 197/198 (1994), 113 -- 131.
[3] T. Ando, F. Hiai and K. Okubo, Trace inequalities for multiple products of two
matrices, Math. Ineq. Appl. 3 (2000), 307 -- 318.
[4] H. Araki, On an inequality of Lieb and Thirring, Lett. Math. Phys. 19 (1990),
167 -- 170.
[5] K. M. R. Audenaert and N. Datta, α-z-relative entropies, J. Math. Phys. 56
(2015), 022202.
[6] V. P. Belavkin and P. Staszewski, C ∗-algebraic generalization of relative entropy
and entropy, Ann. Inst. H. Poincar´e Sect. A 37 (1982), 51 -- 58.
[7] R. Bhatia, Matrix Analysis, Springer-Verlag, New York, 1996.
[8] R. Bhatia, T. Jain and Y. Lim, Strong convexity of sandwiched entropy and related
optimization problems, Rev. Math. Phys., to appear.
[9] J.-C. Bourin, Some inequalities for norms on matrices and operator, Linear Algebra
Appl. 292 (1999), 139 -- 154.
[10] E. Effros and F. Hansen, Non-commutative perspectives, Ann. Funct. Anal. 5
(2014), 74 -- 79.
[11] F. Hiai, Equality cases in matrix norm inequalities of Golden-Thompson type,
Linear and Multilinear Algebra 36 (1994), 239 -- 249.
[12] F. Hiai, Matrix Analysis: Matrix Monotone Functions, Matrix Means, and Ma-
jorization, Interdisciplinary Information Sciences 16 (2010), 139 -- 248.
[13] F. Hiai and M. Mosonyi, Different quantum f -divergences and the reversibility of
quantum operations, Rev. Math. Phys. 29 (2017), 1750023, 80 pp.
[14] F. Hiai and D. Petz, The proper formula for relative entropy and its asymptotics
in quantum probability, Comm. Math. Phys. 143 (1991), 99 -- 114.
[15] F. Hiai and D. Petz, The Golden-Thompson trace inequality is complemented,
Linear Algebra Appl. 181 (1993), 153 -- 185.
[16] M. Kian and Y. Seo, Norm inequalities related to the matrix geometric mean of
negative power, Sci. Math. Japon., Online, 2018.
[17] F. Kubo and T. Ando, Means of positive linear operators, Math. Ann. 246 (1980),
205 -- 224.
20
[18] A. W. Marshall, I. Olkin and B. C. Arnold, Inequalities: Theory of Majorization
and Its Applications, Second ed., Springer-Verlag, New York, 2011.
[19] K. Matsumoto, A new quantum version of f -divergence, arXiv:1311.4722, 2014.
[20] M. Muller-Lennert, F. Dupuis, O. Szehr, S. Fehr and M. Tomamichel, On quantum
R´enyi entropies: A new generalization and some properties, J. Math. Phys. 54
(2013), 122203.
[21] D. Petz, Quasi-entropies for finite quantum systems, Rep. Math. Phys. 23 (1986),
57 -- 65.
[22] W. Pusz and S. L. Woronowicz, Functional calculus for sesquilinear forms and the
purification map, Rep. Math. Phys. 8 (1975), 159 -- 170.
[23] K. Tanahashi, The Furuta inequality with negative powers, Proc. Amer. Math.
Soc. 127 (1999), 1683 -- 1692.
[24] M. M. Wilde, A. Winter and D. Yang, Strong converse for the classical capacity of
entanglement-breaking and Hadamard channels via a sandwiched R´enyi relative
Entropy, Comm. Math. Phys. 331 (2014), 593 -- 622.
21
|
1512.05706 | 1 | 1512 | 2015-12-17T18:20:27 | Lower semicontinuity for an integral functional in BV | [
"math.FA"
] | We prove a lower semicontinuity result for a functional of linear growth initially defined by \[ \int_{\Omega}F\left(\frac{dDu}{d\mu}\right)\,d\mu \] for $u\in BV(\Omega;\mathbb{R}^N)$ with $Du\ll \mu$. The positive Radon measure $\mu$ is only assumed to satisfy $\mathcal L^n\ll \mu$. | math.FA | math |
Lower semicontinuity for an integral
functional in BV ∗
Jan Kristensen and Panu Lahti
Mathematical Institute, University of Oxford,
Andrew Wiles Building,
Radcliffe Observatory Quarter, Woodstock Road,
Oxford, OX2 6GG.
E-mail: [email protected],
[email protected]
July 17, 2018
Abstract
We prove a lower semicontinuity result for a functional of linear
growth initially defined by
ZΩ
F (cid:18) dDu
dµ (cid:19) dµ
for u ∈ BV(Ω; RN ) with Du ≪ µ. The positive Radon measure µ is
only assumed to satisfy Ln ≪ µ.
Acknowledgments: This research was done while P.L. was visiting the
Oxford Centre for Nonlinear Partial Differential Equations (OxPDE) from
September 2014 to July 2016. During this time, P.L. was supported by Aalto
University as well as the Finnish Cultural Foundation.
∗2010 Mathematics Subject Classification: 49J45, 26B30, 52A99.
1
1
Introduction
In this work we prove a lower semicontinuity result for a functional of linear
growth initially defined in an open set Ω ⊂ Rn by
ZΩ
F (cid:18) dDu
dµ (cid:19) dµ
(1.1)
for u ∈ BV(Ω; RN ) with Du ≪ µ. The measure µ is merely assumed to be a
positive finite Radon measure that satisfies Ln ≪ µ, where Ln is the Lebesgue
measure. For the integrand F we need somewhat stronger assumptions,
described in detail below. We refer the reader to Section 2 for precise notation
and terminology.
With the choice µ = Ln, this type of result was derived in [4] (see also
[5, Section 5.5]), and later in [10] for integrands depending also on x and
u. The problem was studied without a nonnegativity assumption on F in
[15]. These results relied mostly on blow-up techniques. The result in [15]
was generalized to x-dependent integrands in [14, Theorem 10], relying on
the theory of generalized Young measures, which were first introduced by
DiPerna and Majda in [8]. With a general measure µ, the problem was
studied in the case p > 1 in [3], and also in [13]. In a more general setting of
a metric measure space, the problem was studied in [11].
We first show that the functional (1.1), defined for general u ∈ BV(Ω; RN )
by relaxation, has an integral representation
ZΩ
F (cid:18)dDu
dµ (cid:19) dµ +ZΩ
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu,
where Ds,µu is the singular part of Du with respect to µ. Here we require
the integrand F : RN ×n → R be nonnegative and quasiconvex, with linear
growth mA ≤ F (A) ≤ M(1 + A) for some 0 < m ≤ M and all A ∈ RN ×n,
and with a continuous recession function F ∞.
Our proof will rely heavily on the theory of generalized Young measures,
particularly results derived in [14]. Once we have the above integral repre-
sentation, we can derive Jensen's inequalities for generalized Young measures
with respect to µ, as was done in [14, Theorem 9] with respect to the Lebesgue
measure. By using these inequalities, we can then prove the following lower
semicontinuity theorem (Theorem 4.4) which is the main result of this work:
Theorem 1.1. Let Ω ⊂ Rn be a bounded Lipschitz domain with inner bound-
ary normal νΩ, let µ be a positive finite Radon measure on Ω with Ln ≪ µ,
and let F : Ω× RN ×n → R be a µ×B(RN ×n)-measurable integrand with linear
2
growth 0 ≤ F (x, A) ≤ M(1 + A) for some M ≥ 0, a continuous recession
function F ∞, and such that A 7→ F (x, A) is quasiconvex for each fixed x ∈ Ω.
Then the functional
dDu
F (cid:18)x,
F (u) :=ZΩ
F ∞(cid:18)x,
+Z∂Ω
u
u
dµ (cid:19) dµ +ZΩ
F ∞(cid:18)x,
⊗ νΩ(cid:19) u dHn−1
dDs,µu
dDs,µu(cid:19) dDs,µu
is weakly* sequentially lower semicontinuous in BV(Ω; RN ).
We remark that an easier proof is possible when the Radon -- Nikodym
derivative of µ with respect to Lebesgue measure is bounded, and hence that
the main contribution is the proof covering the general case. This proof
seems to require the assumption about existence of a continuous recession
function for the integrand F .
2 Preliminaries
2.1 Notation
For N, n ∈ N, the matrix space RN ×n will always be equipped with the
Euclidean norm A := (cid:16)PN
, where i and j are the row and
column indices, respectively. We denote by B(x, r) the open ball in Rn with
center x and radius r. We denote by Bn the open unit ball in Rn and by ∂Bn
the unit sphere. For a ∈ RN and b ∈ Rn, we can define the tensor product
a ⊗ b = abT ∈ RN ×n.
i=1Pn
j=1 Ai
j(cid:17)1/2
We denote the n-dimensional Lebesgue measure by Ln and the s-dimen-
sional Hausdorff measure by Hs. Given any measure ν, the restriction of ν
to a set A is denoted by ν A, that is, ν A(B) = ν(A ∩ B). The Borel
σ-algebra on a set E ⊂ Rn is denoted by B(E). For open sets Ω, Ω′ ⊂ Rn,
by Ω ⋐ Ω′ we mean that Ω ⊂ Ω′ and that Ω is compact. We denote by 1E
the characteristic function of a set E.
If X is a locally compact separable metric space (usually an open or closed
subset of Rn), let Cc(X; Rl) be the space of continuous Rl-valued functions
with compact support in X and let C0(X; Rl) be its completion with respect
to the k · k∞-norm, l ∈ N. We denote by M(X; Rl) the Banach space
of vector-valued finite Radon measures, equipped with the total variation
norm µ(X) < ∞. By the Riesz representation theorem, M(X; Rl) can
be identified with the dual space of C0(X; Rl) through the duality pairing
3
ZΩ
f dµ :=
f dµ
1
µ(Ω)ZΩ
hφ, µi := RX φ · dµ := Pl
∗⇁ µ
in M(X) means hφ, µji → hφ, µi for all φ ∈ C0(X; Rl). We denote the
set of positive measures and probability measures by M+(X) and M1(X),
respectively.
i=1RX φi dµi. Thus weak* convergence µj
For a vector-valued Radon measure γ ∈ M(X; Rl) and a positive Radon
measure µ ∈ M+(X), we can write the Radon-Nikodym decomposition γ =
γa + γs = dγ
We write
dµ µ + γs of γ with respect to µ, where dγ
dµ ∈ L1(X, µ; Rl).
for integral averages (whenever they are defined).
For sets E ⊂ Rn, F ⊂ Rl open or closed, a parametrized measure
(νx)x∈E ⊂ M(F ) is a mapping from E to the set M(F ) of Radon measures
on F . It is said to be weakly* µ-measurable, for µ ∈ M+(E), if x 7→ νx(B) is
µ-measurable for all Borel sets B ∈ B(F ) (it suffices to check this for all rela-
tively open sets). Equivalently, (νx)x∈E is weakly* µ-measurable if the func-
tion x 7→RF f (x, y) dνx(y) is µ-measurable for every bounded Borel function
f : E × F → R (see [5, Proposition 2.26]). We denote by L∞
w∗(E, µ; M(F ))
the set of all weakly* µ-measurable parametrized measures (νx)x∈E ⊂ M(F )
with the property that ess supx∈E νx(F ) < ∞ (the essential supremum with
respect to µ). We omit µ in the notation if µ = Ln.
2.2 Functions of bounded variation
The theory of BV functions presented in this section can be found in e.g. the
monographs [5, 9, 19], and we will give specific references only for a few key
results. Let Ω ⊂ Rn be an open set. A function u ∈ L1(Ω; RN ) is a function of
bounded variation, denoted by u ∈ BV(Ω; RN ), if its distributional derivative
is a bounded RN ×n-valued Radon measure. This means that there exists
a (unique) measure Du ∈ M(Rn; RN ×n) such that for all ψ ∈ C 1
c (Ω), the
integration-by-parts formula
ZΩ
∂ψ
∂xj
ui dLn = −ZΩ
ψ dDui
j,
i = 1 . . . N,
j = 1, . . . , n
holds. We write the Radon-Nikodym decomposition of the variation measure
as Du = ∇u Ln Ω + Dsu.
The space BV(Ω; RN ) is a Banach space endowed with the norm
kukBV(Ω;RN ) := kukL1(Ω;RN ) + Du(Ω).
4
Furthermore, we say that a sequence (uj) ⊂ BV(Ω; RN ) converges weakly*
∗⇁ Du in
to u ∈ BV(Ω; RN ) if uj → u strongly in L1(Ω; RN ) and Duj
M(Ω, RN ×n). A norm-bounded sequence in BV(Ω; RN ), i.e.
(kukL1(Ω;RN ) + Duj(Ω)) < ∞,
sup
j∈N
always has a weakly* converging subsequence. Conversely, a weakly* con-
verging sequence is norm-bounded in BV(Ω, RN ), see [5, Proposition 3.13]. If
uj → u in L1(Ω; RN ) and Duj(Ω) → Du(Ω), we say that the uj converge
to u strictly. If even
hDuji(Ω) → hDui(Ω),
where for a measure ν ∈ M(Rn; RN ×n) with Radon-Nikodym decomposi-
tion ν = a Ln + µs, we define the measure (related to the minimal surface
functional)
hνi(A) :=ZAp1 + a2 dLn + µs(A),
A ∈ B(Rn),
then we speak of h·i-strict convergence. This notion is stronger than strict
convergence (this follows e.g. from Theorem 2.2 below), and one can show
that it implies that hDuji ∗⇁ hDui as measures.
For any bounded open set Ω ⊂ Rn and v ∈ BV(Ω; RN ), we can define the
Dirichlet class
BVv(Ω; RN ) :=(cid:8)u ∈ BV(Ω; RN ) : w ∈ BV(Rn; RN ) and Dw(∂Ω) = 0(cid:9) ,
where
w :=(u − v
0
in Ω,
in Rn \ Ω.
The following lemma is proved in e.g. [14, Lemma 1].
Lemma 2.1. Let Ω ⊂ Rn be a bounded open set, and let u ∈ BV(Ω; RN ).
Then there exists (vj) ⊂ BVu(Ω; RN ) ∩ C ∞(Ω; RN ) such that vj → u h·i-
strictly in Ω.
2.3 Generalized Young measures
Most of the theory of generalized Young measures presented in this section
is derived in [14].
The symbol Ω will always denote a bounded open set in Rn. We will
need the following linear transformations mapping C(Ω × Rl) to C(Ω × Bl)
5
and back, where Bl was the open unit ball in Rl:
g ∈ C(Ω × Bl), define
for f ∈ C(Ω × Rl) and
A
1 − A! , x ∈ Ω, A ∈ Bl,
and
A
1 + A(cid:19) , x ∈ Ω, A ∈ Rl.
(T f )(x, A) := (1 − A)f x,
(T −1g)(x, A) := (1 + A)g(cid:18)x,
It is an easy calculation to verify that T −1T f = f and T T −1g = g. We
consider the property
T f extends to a bounded continuous function on Ω × Bl.
(2.1)
In particular, this entails that f has linear growth at infinity, that is, there
exists a constant M ≥ 0 (in fact, M = kT f kL∞(Ω×Bl) will do) such that
f (x, A) ≤ M(1 + A)
for all x ∈ Ω, A ∈ Rl.
We collect all such integrands into the set
E(Ω; Rl) := {f ∈ C(Ω × Rl) : f satisfies (2.1)}.
For f ∈ E(Ω; Rl), the recession function f ∞ : Ω × Rl 7→ R is defined by
f ∞(x, A) := lim
x′→x
A′→A
t→∞
f (x′, tA′)
t
,
x ∈ Ω, A ∈ Rl.
(2.2)
The limit exists since it agrees with T f on Ω × ∂BN ×n, as can be seen by
substituting t = s/(1 − s), s ∈ (0, 1), and letting s → 1. The recession func-
tion is clearly positively 1-homogenous in A, that is, f ∞(x, sA) = sf ∞(x, A)
for all s ≥ 0, and thus takes finite values.
We also consider a second class of integrands that is larger than E(Ω; Rl)
and (partially) dispenses with continuity in the x-variable. A Carath´eodory
function is an Ln × B(Rl)-measurable function f : Ω × Rl → R such that
A 7→ f (x, A) is continuous for almost every x ∈ Ω. In fact, it can be shown
that it suffices to check measurability of x 7→ f (x, A) for all fixed A ∈ Rl
(see for example [5, Proposition 5.6]). With this notion, the representation
integrands are defined as follows:
R(Ω; Rl) := {f : Ω × Rl → R : f Carath´eodory with linear growth
at infinity and ∃f ∞ ∈ C(Ω × Rl)}.
(2.3)
6
A function f : RN ×n → R is said to be quasiconvex, which we denote by
f ∈ Q(RN ×n), if f is Borel measurable, locally bounded from below, and for
some bounded Lipschitz domain ω ⊂ Rn and every A ∈ RN ×n it holds that
ωf (A) ≤Zω
f (A + ∇ψ(x)) dLn(x)
for all ψ ∈ W 1,∞
0
(ω; RN ).
This definition does not depend on the particular choice of the Lipschitz do-
main ω (by an exhaustion argument) and it can be shown that quasiconvex
functions are rank one convex, meaning that they are convex along rank one
lines (see for example [5, Proposition 5.41]). See [7] for more on quasicon-
vexity.
A quasiconvex function does not necessarily have a recession function f ∞
in the sense of (2.2) (see [16, Theorem 2] for a counterexample), and the
notion can be relaxed in the following way: for f : RN ×n → R the generalized
recession function f # : RN ×n → R ∪ {±∞} is defined by
f #(A) := lim sup
A′→A
t→∞
f (tA′)
t
, A ∈ RN ×n.
Quasiconvex functions are globally Lipschitz continuous (see for example [6,
Lemma 2.2]) and hence for quasiconvex f
f #(A) = lim sup
t→∞
f (tA)
t
, A ∈ RN ×n.
(2.4)
By rank one convexity, the above holds as a limit for all matrices A of rank
one.
We have the following version of Reshetnyak's Continuity Theorem, see
the appendix of [15], as well as [17, Theorem 3] or [5, Theorem 2.39] for the
original result stated for 1-homogenous functions f .
Theorem 2.2. Let (γj) ⊂ M(Ω; Rl), γ ∈ M(Ω; Rl) with Radon-Nikodym
decompositions
γj = aj Ln Ω + γs
j ,
γ = a Ln Ω + γs.
If γj
∗⇁ γ in M(Ω; Rl) and hγji(Ω) → hγi(Ω), then for
F (γ) :=ZΩ
f (x, a(x)) dLn +ZΩ
f ∞(cid:18)x,
dγs
dγs
(x)(cid:19) dγs(x)
with f ∈ E(Ω, Rl), we have F (γj) → F (γ).
7
Let µ ∈ M+(Ω), and assume that µ(∂Ω) = 0.
The set of all generalized Young measures Y(Ω, µ; Rl) is defined to be the
set of all triples (νx, λν, ν∞
x ) such that
(νx)x ∈ L∞
(ν∞
x )x ∈ L∞
w∗(Ω, µ; M1(Rl)),
w∗(Ω, λν; M1(∂Bl)),
λν ∈ M+(Ω),
x 7→ h · , νxi ∈ L1(Ω, µ).
Under the duality pairing
hhf, νii :=ZΩ
=ZΩZRN ×n
hf ∞(x, ·), ν∞
hf (x, ·), νxi dµ(x) +ZΩ
f (x, A) dνx(A) dµ(x) +ZΩZ∂BN ×n
x i dλν(x)
f ∞(x, A) dν∞
x (A) dλν(x),
where f ∈ E(Ω; Rl) and ν ∈ Y(Ω, µ; Rl), the space of Young measures can
be considered a part of the dual space E(Ω; Rl)∗. We say that a sequence of
Young measures (νj) ⊂ Y(Ω, µ; Rl) converges weakly* to ν ∈ Y(Ω, µ; Rl) if
hhf, νjii → hhf, νii for every f ∈ E(Ω; Rl).
To every Radon measure γ ∈ M(Ω; Rl), with Radon-Nikodym decompo-
dµ µ+γs,µ, we associate an elementary
sition with respect to µ written as γ = dγ
Young measure εγ ∈ Y(Ω, µ; Rl) by
(εγ)x := δ dγ
dµ (x),
λεγ := γs,µ,
(εγ)∞
x := δp(x),
where p := dγs,µ
dγs,µ ∈ L1(Ω, γs,µ; ∂Bl).
Crucially, we have the following.
Theorem 2.3. Let µ ∈ M+(Ω) with µ(∂Ω) = 0, and let (γj) ⊂ M(Ω; Rl) be
a sequence of Radon measures that is bounded in the total variation norm,
that is, supj∈N γj(Ω) < ∞. Then there exists a subsequence (not relabeled)
and a generalized Young measure (νx, λν, ν∞
x ) with
(νx)x ∈ L∞
(ν∞
x )x ∈ L∞
w∗(Ω, µ; M1(Rl)),
w∗(Ω, λν; M1(∂Bl)),
λν ∈ M+(Ω),
x 7→ h · , νxi ∈ L1(Ω, µ),
such that hhf, εγj ii → hhf, νii, or equivalently
f(cid:18)x,
dγj
dµ
(x)(cid:19) µ + f ∞(cid:18)x,
j
dγs,µ
dγs,µ
j
(x)(cid:19) γs,µ
j
∗⇁ hf (x, ·), νxi µ + hf ∞(x, ·), ν∞
x i λν
in M(Ω)
(2.5)
(2.6)
for every f ∈ E(Ω; RN ×n).
8
Proof. This is proved in the case µ = Ln Ω in [14, Lemma 2, Corollary
2, Theorem 7], but the proofs run through also if we replace the Lebesgue
measure by a more general µ.
See also [2, Theorem 2.5] for a proof in the case γs
j ≡ 0.
Corollary 2.4. In the above theorem, (2.6) holds also for every µ × B(Rl)-
measurable f ∈ R(Ω; Rl).
In the case µ = Ln, (2.6) also holds for every Carath´eodory integrand
f : Ω × Rl → R possessing a recession function f ∞ : Ω × Rl → R in the
sense of (2.2) for (x, A) ∈ (Ω \ N) × Rl, and f ∞ is jointly continuous in
(Ω \ N) × Rl, where N ⊂ Ω is a Borel set with (Ln + λν)(N) = 0.
Note that a Carath´eodory function f : Ω × RN → R is by definition
Ln × B(RN )-measurable, but here we need the assumption of µ × B(RN )-
measurability.
Proof. Again, this is proved in the case µ = Ln Ω in [14, Proposition 2],
but the proof runs through also in the general case with the assumption of
µ × B(RN )-measurability.
In particular, given u ∈ BV(Ω; RN ), we can associate to its derivative
dµ µ + Ds,µu,
Du ∈ M(Ω; RN ×n) the Radon-Nikodym decomposition Du = dDu
and then the elementary Young measure εDu ∈ Y(Ω, µ; RN ×n) with
(εDu)x := δ dDu
dµ
,
λεDu := Ds,µu,
(εDu)∞
x := δp(x),
where p := Ds,µu
Ds,µu ∈ L1(Ω, Ds,µu; ∂BN ×n).
For a norm-bounded sequence (uj) ⊂ BV(Ω; RN ), we say that the deriva-
tives Duj generate the generalized Young measure
ν = (νx, λν, ν∞
x ) ∈ Y(Ω; RN ×n),
if for all f ∈ E(Ω, RN ×n) we have that hhf, εDuj ii → hhf, νii for all f ∈
E(Ω; RN ×n), or equivalently
f(cid:18)x,
dDuj
dµ (cid:19) µ + f ∞(cid:18)x,
dDs,µuj
dDs,µuj
(x)(cid:19) Ds,µuj
∗⇁ hf (x, ·), νxi µ + hf ∞(x, ·), ν∞
x i λν
in M(Ω).
(2.7)
We call such a generalized Young measure a gradient Young measure. Since
∗⇁ u for some u ∈ BV(Ω; RN ). The
(uj) is norm-bounded, we have uj
barycenter of a generalized Young measure ν ∈ Y(Ω; RN ×n) is defined as the
measure
hid, νxi µ + hid, ν∞
x iλν.
9
Note that by choosing f to be the identity on RN ×n in (2.7) (componentwise,
to be precise), we obtain that Du is the restriction of the barycenter to Ω.
In the case µ = Ln, we have the following Jensen's inequalities for gradient
Young measures, which are part of [14, Theorem 9].
Theorem 2.5. Let u ∈ BV(Ω; RN ) and let ν ∈ Y(Ω; RN ×n), ν = (νx, λν, ν∞
x )
be a gradient Young measure with barycenter Du and satisfying λ(∂Ω) = 0.
Then the following hold for any quasiconvex f : RN ×n → R with linear growth
(that is, F (A) ≤ M(A + 1) for all A ∈ RN ×n and some M ≥ 0):
f (∇u(x)) ≤ hf, νxi + hf #, ν∞
x i
dλν
dLn (x)
for Ln-almost every x ∈ Ω,
f #(cid:18) dDsu
dDsu(cid:19) Dsu ≤ hf #, ν∞
x iλs
ν
as measures.
3 The integral representation
Let F : RN ×n 7→ R be quasiconvex, with linear growth
mA ≤ F (A) ≤ M(1 + A)
for all A ∈ RN ×n,
for some 0 < m ≤ M, such that the recession function F ∞ exists in the
sense of (2.2). Let Ω ⊂ Rn be a bounded open set with Ln(∂Ω) = 0, and let
µ ∈ M+(Ω) with Ln ≪ µ. We define a Sobolev space with respect to µ by
W 1,1
µ (Ω; RN ) := {u ∈ BV(Ω; RN ) : Du ≪ µ}.
We consider the functional
j→∞ ZΩ
F∗(u, Ω) := infn lim inf
F (cid:18)dDuj
dµ (cid:19) dµ, uj ∈ W 1,1
µ (Ω; RN )
(3.1)
uj → u in L1(Ω; RN )o
for u ∈ BV(Ω; RN ). Note that the convergence above is in L1(Ω; RN ) with
respect to the Lebesgue measure Ln, not µ. We will prove an integral repre-
sentation for the above functional. The representation is
ZΩ
F (cid:18)dDu
dµ (cid:19) dµ +ZΩ
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu
(3.2)
for any u ∈ BV(Ω; RN ), where Ds,µu is the singular part of the variation
measure Du with respect to µ.
Initially we will work with a more restricted class of integrands, defined
as follows.
10
Definition 3.1. Define the class SQ(RN ×n) of special quasiconvex integrands
as quasiconvex functions F : RN ×n → R with linear growth F (A) ≤ M(1 +
A) for some M ≥ 0, such that for some parameters i, ri > 0, F (A) =
F ∞(A) − i for A ≥ ri, and F ∞(A) ≥ A/i for all A ∈ RN ×n.
Note that the existence of the recession function F ∞ in the sense of (2.2)
is part of the definition. (We could equally well require above that F (A) =
F #(A) − i for A ≥ ri, recall (2.4), as this would imply the existence of F ∞.)
Clearly SQ(RN ×n) ⊂ E(Ω; RN ×n) (constant in the x-variable).
Given F ∈ Q(RN ×n) with linear growth 0 ≤ F (A) ≤ M(1 + A) for some
M ≥ 0, we can define Gi(A) := max{F (A), F #(A) + A/i − i} for each
i ∈ N, and then it is shown in [12, Lemma 6.3] that Gi ∈ SQ(RN ×n) and
that Gi(A) ց F (A) and (Gi)∞(A) ց F ∞(A) for every A ∈ RN ×n. We will
use this fact on a number of occasions.
3.1 Estimate from below
In order to obtain the integral representation, we first prove the estimate
from below.
Proposition 3.2. Let Ω ⊂ Rn be a bounded open set with Ln(∂Ω) = 0, let
µ ∈ M+(Ω) with Ln ≪ µ, let F ∈ R(Ω; RN ×n) ∩ Q(RN ×n) with
mA ≤ F (A) ≤ M(1 + A)
for some 0 < m ≤ M, and let u ∈ BV(Ω; RN ). Then we have
F∗(u, Ω) ≥ZΩ
F (cid:18) dDu
dµ (cid:19) dµ +ZΩ
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu.
Write the Radon-Nikodym decomposition of µ as µ = a Ln + µs, with
a ∈ L1(Ω). We prove the theorem by considering separately the sets where
the absolutely continuous part and the singular part of µ are carried.
3.1.1 The absolutely continuous part
The following lemma gives, in essence, the estimate from below for the set
where the absolutely continuous part of µ is carried. At this point, we make
the extra assumption that F is a special quasiconvex integrand.
Lemma 3.3. Let Ω ⊂ Rn be a bounded open set with Ln(∂Ω) = 0, let
µ ∈ M+(Ω) with Ln ≪ µ, and let F ∈ SQ(RN ×n) with parameters i, ri > 0,
11
and with linear growth 0 ≤ F (A) ≤ M(1 + A). Then for any open U ⊂ Ω
and any sequence (uj) ⊂ W 1,1
µ (Ω; RN ) with uj → u in L1(Ω; RN ) and
lim inf
j→∞ ZΩ
F (cid:18) dDuj
dµ (cid:19) dµ < ∞,
(3.3)
we have
lim inf
j→∞ ZU
≥ZU
F (cid:18)dDuj
F (cid:18) ∇u
dµ (cid:19) dµ
a (cid:19) a dLn +ZU
F ∞(cid:18) dDsu
dDsu(cid:19) dDsu − (Mri + i)µs(U).
(3.4)
Proof. Since Ln ≪ µ, we can assume that a > 0 everywhere in Ω. Pick
a subsequence of (uj) (not relabeled) that gives the limit in (3.4). Since
F (A) = F ∞(A) − i ≥ A/i − i for all A ≥ ri, we have by (3.3) that (uj)
∗⇁ u in BV(Ω; RN ).
is a norm-bounded sequence in BV(Ω; RN ). Thus uj
By Theorem 2.3, the derivatives Duj generate a generalized Young measure
(νx, λν, ν∞
x ) with respect to the Lebesgue measure, with λν ∈ M+(Ω) and
(νx)x ∈ L∞
w∗(Ω; M1(RN ×n)),
(ν∞
x )x ∈ L∞
w∗(Ω, λν; M1(∂BN ×n)).
This means that for every representation integrand f ∈ R(Ω; RN ×n) and
every integrand satisfying the conditions of the latter part of Corollary 2.4,
we have
f (x, ∇uj(x)) Ln Ω + f ∞(cid:18)x,
dDsuj
dDsuj(cid:19) Dsuj
∗⇁ hf (x, ·), νxi Ln Ω + hf ∞(x, ·), ν∞
x i λν
(3.5)
in M(Ω).
First assume that Ln(∂U) = λν(∂U) = 0. Let us start computing
ZU
F (cid:18) dDuj
dµ (cid:19) dµ =ZU
=ZU
F (cid:18) dDuj
F (cid:18)∇uj
d(aLn)(cid:19) a dLn +ZU
a (cid:19) a dLn +ZU
F (cid:18)dDuj
dµs (cid:19) dµs
F (cid:18)dDsuj
dµs (cid:19) dµs
=: Ij + IIj.
(3.6)
We wish to analyze the term Ij by using the fact that Duj generates a
generalized Young measure. However, the function
(x, A) 7→ F (cid:18) A
a(x)(cid:19) a(x)1U (x)
12
does not necessarily satisfy the conditions of the latter part of Corollary
2.4: while it is a Carath´eodory function, its recession function need not be
continuous as required. To overcome this problem, we define the super-level
sets of a:
Em := {x ∈ Ω : a(x) > m},
m ∈ N.
Recall that a(x) > 0 for every x ∈ Ω. Denoting the minimum of a and m by
a ∧ m, by the fact that F (A) = F ∞(A) − i for all A ≥ ri we have for any
x ∈ U and A ∈ ∂BN ×n
F (cid:16) tA′
a(x′)∧m(cid:17)
t
a(x′) ∧ m = lim sup
A′→A
t→∞
F (tA′)
t
lim sup
x′→x
A′→A
t→∞
F ∞(tA′) − i
t
= lim sup
A′→A
t→∞
= lim sup
A′→A
t→∞
tF ∞(A′)
t
= lim sup
A′→A
F ∞(A′) = F ∞(A)
by the (Lipschitz) continuity of F ∞. Note that the first equality is not
necessarily true unless we take the minimum of a with m. Also, we now see
that all of the limit superiors above are in fact limits. We conclude that
(x, A) 7→ F (cid:18)
A
(a ∧ m)(x)(cid:19) (a ∧ m)(x)1U (x)
(3.7)
satisfies the conditions of the latter part of Corollary 2.4. Fix m ∈ N. By
the fact that F (A) = F ∞(A) − i for all A ≥ ri, we can write
m (cid:19) m dLn
F (cid:18)∇uj
a (cid:19) a − F (cid:18)∇uj
F ∞(cid:18)∇uj
a (cid:19) a − F (cid:18)∇uj
a (cid:19) a − F (cid:18)∇uj
m (cid:19) m dLn
a (cid:19) a − F ∞(cid:18)∇uj
m (cid:19) m dLn
m (cid:19) m − i(a − m) dLn
Ij −ZU
a ∧ m(cid:19) a ∧ m dLn =ZU ∩Em
F (cid:18) ∇uj
=ZU ∩Em∩{∇uj<ari}
F (cid:18)∇uj
+ZU ∩Em∩{∇uj≥ari}
=ZU ∩Em∩{∇uj<ari}
F (cid:18)∇uj
−ZU ∩Em∩{∇uj≥ari}
:= εm.
i(a − m) dLn
13
We have by the linear growth of F
(cid:12)(cid:12)(cid:12)(cid:12)
ZU ∩Em∩{∇uj<ari}
a (cid:19) a − F (cid:18)∇uj
m (cid:19) m dLn(cid:12)(cid:12)(cid:12)(cid:12)
a (cid:19) a + F (cid:18)∇uj
F (cid:18)∇uj
m (cid:19) m dLn
F (cid:18) ∇uj
≤ZU ∩Em∩{∇uj<ari}
≤ZU ∩Em
≤ZU ∩Em
2Ma(1 + ri) dµ.
M(1 + ri)a + M(1 + ari/m)m dµ
Clearly this last quantity converges to zero as m → ∞, as does the second
term of εm, so in total εm → 0 as m → ∞.
By the fact that the derivatives Duj generate a generalized Young mea-
sure (recall (3.5)) and the fact that the integrand (3.7) satisfies the conditions
of the latter part of Corollary 2.4 and has recession function F ∞ in U, we
have
F ∞(cid:18) dDsuj
Ij +ZU
F (cid:18) ∇uj
=ZU
→ZUZRN ×n
dDsuj(cid:19) dDsuj − εm
a ∧ m(cid:19) a ∧ m dLn +ZU
F (cid:18) A
F ∞(cid:18) dDsuj
dDsuj(cid:19) dDsuj
a ∧ m(cid:19) a ∧ m dνx(A) dLn +ZUZ∂BN ×n
F ∞(A) dν∞
x (A) dλν
(3.8)
as j → ∞. Recalling (3.6), let us then consider the term IIj. Since F (A) =
14
F ∞(A) − i for all A ≥ ri, we estimate
F ∞(cid:18) dDsuj
F (cid:18)dDsuj
IIj −ZU
=ZU
=ZU ∩{dDsuj /dµs<ri}
dDsuj(cid:19) dDsuj
dµs (cid:19) dµs −ZU
F (cid:18)dDsuj
−ZU ∩{dDsuj/dµs<ri}
dµs (cid:19) dµs
F ∞(cid:18)dDsuj
dµs (cid:19) dµs − iµs(U ∩ {dDsuj/dµs ≥ ri})
F ∞(cid:18) dDsuj
dµs (cid:19) dµs
≥ −iµs(U ∩ {dDsuj/dµs ≥ ri}) −ZU ∩{dDsuj /dµs<ri}
≥ −iµs(U) −ZU ∩{dDsuj/dµs<ri}
dDsuj
dµs
dµs
≥ −iµs(U) − Mriµs(U).
M(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
F ∞(cid:18)dDsuj
dµs (cid:19) dµs
(3.9)
Combining (3.6), (3.8), and (3.9), we get by Jensen's inequalities for gener-
alized Young measures given in Theorem 2.5,
x (A) dλν
lim inf
j→∞
(Ij + IIj)
F ∞(A) dν∞
+ εm − (Mri + i)µs(U)
j→∞ ZU
≥ZUZRN ×n
F (cid:18) ∇u
≥ZU
F (cid:18)∇u
≥ZU
F (cid:18)∇u
→ZU
dµ (cid:19) dµ = lim inf
F (cid:18)dDuj
F (cid:18) A
a ∧ m(cid:19) a ∧ m dνx(A) dLn +ZUZ∂BN ×n
F ∞(cid:18) dDsu
dDsu(cid:19) dDsu
a ∧ m(cid:19) a ∧ m dLn +ZU
F ∞(cid:18) dDsu
dDsu(cid:19) dDsu
a (cid:19) a ∧ m dLn +ZU
dDsu(cid:19) dDsu − (Mri + i)µs(U)
a (cid:19) a dLn +ZU
F ∞(cid:18) dDsu
+ εm − (Mri + i)µs(U)
+ εm − (Mri + i)µs(U)
as m → ∞, by the monotone convergence theorem. Finally, if U does not
satisfy λν(∂U) = 0 or Ln(∂U) = 0, we define
Uκ := {x ∈ U : dist(x, U c) > κ},
κ > 0,
15
and then λν(∂Uκ) = 0 and Ln(∂Uκ) = 0 for all but at most countably many
κ > 0 by the fact that these are finite measures on U. For such values of κ
we write
lim inf
j→∞ ZU
≥ lim inf
F (cid:18) dDuj
dµ (cid:19) dµ
F (cid:18) dDuj
dµ (cid:19) dµ
j→∞ ZUκ
F (cid:18)∇u
a (cid:19) a dLn +ZUκ
a (cid:19) a dLn +ZU
F (cid:18)∇u
≥ZUκ
→ZU
F ∞(cid:18) dDsu
F ∞(cid:18) dDsu
dDsu(cid:19) dDsu − (Mri + i)µs(Uκ)
dDsu(cid:19) dDsu − (Mri + i)µs(U)
as κ → 0, by the monotone convergence theorem.
3.1.2 The singular part
Let us then consider the set where µs is carried. We prove the following
lemma.
Lemma 3.4. Let Ω ⊂ Rn be a bounded open set, let µ ∈ M+(Ω), and let
F ∈ SQ(RN ×n) with F ≥ 0. Then for any sequence (uj) ⊂ W 1,1
µ (Ω; RN ) with
uj → u in L1(Ω; RN ) and
lim inf
j→∞ ZΩ
F (cid:18)dDuj
dµ (cid:19) dµ < ∞,
we have for any ball B(y, r) ⊂ Ω
ZB(y,r)
F (cid:18) dDsu
dµs (cid:19) dµs ≤ lim inf
j→∞ ZB(y,r)
F (cid:18)dDuj
dµ (cid:19) dµ.
(3.10)
Proof. Note again that it is enough to prove the result for a subsequence.
Let i, ri > 0 be the parameters of F , see Definition 3.1. Since F (A) =
F ∞(A) − i ≥ A/i − i for all A ≥ ri, the sequence dDuj
dµ is norm-bounded in
L1(Ω, µ; RN ×n), implying that (uj) is a norm-bounded sequence in BV(Ω; RN ).
By Theorem 2.3 we know that with respect to µ, a subsequence of Duj
(not relabeled) generates a generalized Young measure (νx, λν, ν∞
x ), with
λν ∈ M+(Ω) and
(νx)x ∈ L∞
w∗(Ω, µ; M1(RN ×n)),
16
(ν∞
x )x ∈ L∞
w∗(Ω, µ; M1(∂BN ×n)).
This means in particular that for every integrand f ∈ E(Ω; RN ×n),
f(cid:18)dDuj
dµ (cid:19)µ ∗⇁ hf, νxiµ + hf ∞, ν∞
=(cid:18)hf, νxia + hf ∞, ν∞
x i
x iλν
dλν
dLn(cid:19) Ln + hf, νxiµs + hf ∞, ν∞
x iλs
ν
(3.11)
in M(Ω). By Alberti's rank one theorem, see [1], we have µs-almost every-
where that
dDuj
dµs = ξj ⊗ η,
dDu
dµs = ξ ⊗ η,
(3.12)
where ξj, ξ ∈ RN and η ∈ ∂Bn. Note that η does not depend on j. We show
that for µs-almost every x ∈ Ω, the measure νx is carried on the hyperplane
RN ⊗ η(x). For this, fix ε > 0 and fix a point x0 ∈ Ω. Excluding a µs-
negligible set, we can assume by the Besicovitch differentiation theorem (see
e.g.
[5, Theorem 2.22]) that for some radius r > 0, we have B(x0, r) ⊂ Ω
and
ZB(x0,r)
η − η(x0) dµs < ε
and
ZB(x0,r)
Fix R ≥ 1 and define
a dLn < εµ(B(x0, r)).
(3.13)
f (A) := min{1, dist(A, (RN ⊗ η(x0)) ∪ B(0, R)c)};
note that there is no x-dependence, and f ∈ E(Ω; RN ×n). Since f (ξ ⊗ η) = 0
for ξ ≥ R and η = 1 and since f is 1-Lipschitz,
(cid:12)(cid:12)(cid:12)(cid:12)
ZB(x0,r)
f (ξj ⊗ η) dµs −ZB(x0,r)
≤ RZB(x0,r)
η − η(x0) dµs < Rε
f (ξj ⊗ η(x0)) dµs(cid:12)(cid:12)(cid:12)(cid:12)
(3.14)
by (3.13). Since hf, νxi ∈ L1(Ω, µ) by (2.5), excluding a further µs-negligible
set and possibly making r > 0 smaller, we can also assume that
ZB(x0,r)
hf, νxi − hf, νx0i dµ < ε.
(3.15)
Clearly f ∞ ≡ 0 and then by (3.11), we have
f(cid:18)dDuj
dµ (cid:19) µ ∗⇁ hf, νxia Ln + hf, νxi µs
in M(Ω).
17
Now by (3.15),
hf, νx0i ≤ZB(x0,r)
≤ lim inf
hf, νxi dµ + ε
f ≤1
≤ lim inf
(3.13)
≤ lim inf
j→∞ ZB(x0,r)
j→∞ ZB(x0,r)
j→∞ ZB(x0,r)
j→∞ ZB(x0,r)
j→∞ ZB(x0,r)
(3.12)
= lim inf
(3.14)
≤ lim inf
≤ 3Rε,
dµ (cid:19) dµ + ε
f(cid:18)dDuj
dµs (cid:19) dµs + RB(x0,r) a dLn
f(cid:18)dDuj
dµs (cid:19) dµs + 2ε
f(cid:18)dDuj
µ(B(x0, r))
f (ξj ⊗ η) dµs + 2ε
f (ξj ⊗ η(x0)) dµs + Rε + 2ε
+ ε
since f is zero on the hyperplane RN ⊗η(x0). Letting ε → 0, we get hf, νx0i =
0, implying that νx0 is carried on the set (RN ⊗ η(x0)) ∪ B(0, R)c. Letting
R → ∞, we obtain that νx0 is carried on the hyperplane RN ⊗ η(x0).
By choosing f to be the identity mapping on RN ×n in (3.11) (compo-
∗⇁ Du in M(Ω) (the fact
∗⇁ u in
nentwise, to be precise), and noting that Duj
that (uj) is a norm-bounded sequence in BV(Ω; RN ) implies that uj
BV(Ω; RN )), we get for the singular parts
Dsu = hid, νxi µs + hid, ν∞
x i λs
ν
(3.16)
in Ω. Using the fact that hid, νxi ∈ RN ⊗ η(x) for µs-almost every x ∈ Ω, we
get
hid, ν∞
x i
dλs
ν
dµs =
dDsu
dµs (x) − hid, νxi = ξ(x) ⊗ η(x) − hid, νxi ∈ RN ⊗ η(x)
for µs-almost every x ∈ Ω. Since F ∞ is quasiconvex and 1-homogenous,
we have F ∞(A) = (F ∞)c(A) for all rank one A ∈ RN ×n, where the convex
envelope is defined by
Gc(A) := sup {H(A) : H convex, H ≤ G} ,
see [12, Corollary 1.2]. According to [2, Lemma 5.5 (i)], for any convex
function g : RN ×n → R we have
g(A1 + A2) ≤ g(A1) + g∞(A2)
(3.17)
18
for all A1, A2 ∈ RN ×n. Note that in (3.16), all three terms belong to RN ⊗η(x)
for µs-almost every x ∈ Ω. Since ξ 7→ F (ξ ⊗ η(x)) is convex for a fixed x ∈ Ω
by the rank one convexity of F , we get by (3.17)
F (cid:18) dDsu
dµs (x)(cid:19) ≤ F (hid, νxi) + F ∞ (hid, ν∞
x i)
= F (hid, νxi) + (F ∞)c (hid, ν∞
x i)
Jensen
≤ hF, νxi + h(F ∞)c, ν∞
x i
dλs
ν
dµs
≤ hF, νxi + hF ∞, ν∞
x i
dλs
ν
dµs
dλs
ν
dµs
dλs
ν
dµs
for µs-almost every x ∈ Ω. Combining this with (3.11) -- note that F ∈
SQ(RN ×n) ⊂ E(Ω; RN ×n) -- we get for any ball B(y, r) ⊂ Ω
ZB(y,r)
F (cid:18)dDsu
dµs (cid:19) dµs ≤ lim inf
j→∞ ZB(y,r)
F (cid:18)dDuj
dµ (cid:19) dµ.
3.1.3 Combining the estimates
Now we combine the previous two lemmas to prove Proposition 3.2.
Proof of Proposition 3.2. We keep assuming that F ∈ SQ(RN ×n) with pa-
rameters i, ri ≥ 1 and linear growth 0 ≤ F (A) ≤ M(1 + A). Let (uj) ⊂
W 1,1
µ (Ω; RN ) be a sequence with uj → u in L1(Ω; RN ), and we can also
assume that (3.3) holds, so that the assumptions of both Lemma 3.3 and
Lemma 3.4 are satisfied.
Fix ε > 0. Let H ⊂ Ω be a Borel set with Ln(H) = 0 and µs(Ω \ H) = 0.
Also, let D ⊂ Ω be a Borel set with µ(D) = 0 and Ds,µu(Ω \ D) = 0. Take
an open set G ⊂ Ω with G ⊃ H \ D and
Ds,µu(G) +ZG
M(a + ∇u) dLn < ε.
(3.18)
Consider the fine cover {B(x, R)}x∈H\D of the set H \ D, with the balls
B(x, R) contained in G and satisfying µs(∂B(x, R)) = 0. By Vitali's covering
theorem, we can pick a countable, disjoint collection {Bi}i∈N := {B(xi, Ri)}i∈N
with
µs (H \ D) \
∞
[i=1
Bi! = 0 and thus µs Ω \
Bi! = 0.
∞
[i=1
(3.19)
19
Pick also m ∈ N such that
µ ∞
[i=m
Bi! + MZS∞
i=m Bi(cid:18)1 +
dDsu
dµs (cid:19) dµs < ε.
By (3.10) we have
lim inf
j→∞ ZSm
i=1 Bi
F (cid:18)dDuj
dµ (cid:19) dµ ≥ZSm
≥ZΩ
F (cid:18)dDsu
dµs (cid:19) dµs
dµs (cid:19) dµs − ε
F (cid:18)dDsu
i=1 Bi
by (3.20) and the linear growth of F .
By combining (3.19) and (3.20), we get
µs Ω \
m
[i=1
Bi! < ε.
(3.20)
(3.21)
(3.22)
Moreover, we can write (3.4) with the choice U = Ω \Sm
i=1 Bi:
lim inf
j→∞ ZΩ\Sm
≥ZΩ\Sm
i=1 Bi
dµ (cid:19) dµ
F (cid:18)dDuj
F (cid:18)∇u
a (cid:19) a dLn +ZΩ\Sm
i=1 Bi
F ∞(cid:18) dDsu
dDsu(cid:19) dDsu
i=1 Bi
Bi!
(3.22)
(3.18)
m
i=1 Bi
[i=1
− (Mri + i)ε
− (Mri + i)µs Ω \
F (cid:18)∇u
≥ ZΩ\Sm
F (cid:18)∇u
a (cid:19) a dLn +ZΩ
≥ ZΩ\Sm
≥ ZΩ
F (cid:18)∇u
− Mε − (Mri + i)ε
i=1 Bi
− ε − Mε − (Mri + i)ε.
(3.18)
i=1 Bi
a (cid:19) a dLn +ZΩ\Sm
a (cid:19) a dLn +ZΩ
dDs,µu(cid:19) dDs,µu
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu
F ∞(cid:18) dDs,µu
20
Combining this with (3.21), we get
lim inf
j→∞ ZΩ
dDs,µu(cid:19) dDs,µu
F ∞(cid:18) dDs,µu
dµ (cid:19) dµ
F (cid:18)dDuj
F (cid:18) ∇u
a (cid:19) a dLn +ZΩ
≥ZΩ
F (cid:18) dDsu
dµs (cid:19) dµs − 3(Mri + i)ε
+ZSm
F ∞(cid:18) dDs,µu
a (cid:19) a dLn +ZΩ
F (cid:18)∇u
≥ ZΩ
F (cid:18)dDsu
dµs (cid:19) dµs − 4(Mri + i)ε.
+ZΩ
i=1 Bi
(3.20)
dDs,µu(cid:19) dDs,µu
By letting ε → 0, we get the estimate from below.
Finally, we remove the assumption F ∈ SQ(RN ×n). By [12, Lemma 6.3],
we can find a sequence Fi ∈ SQ(RN ×n) with Fi(A) ց F (A) and F ∞
i (A) ց
F ∞(A) pointwise for every A ∈ RN ×n as i → ∞, and by making M slightly
larger, if necessary, we can also assume that mA ≤ Fi(A) ≤ M(1 + A) for
every i ∈ N.
As before, let (uj) ⊂ W 1,1
µ (Ω; RN ) with uj → u in L1(Ω; RN ). We can
again assume that (3.3) holds, and by the coercivity mA ≤ F (A), this
implies that (uj) is a norm-bounded sequence in BV(Ω; RN ). Thus by Theo-
rem 2.3, a subsequence of Duj (not relabeled) generates a generalized Young
measure (νx, λν, ν∞
x ) with respect to µ. Thus we have for any i ∈ N
dDs,µu(cid:19) dDs,µu
i (cid:18) dDs,µu
F ∞
dDs,µu(cid:19) dDs,µu
(3.23)
ZΩ
F ∞(cid:18) dDs,µu
F (cid:18) dDu
dµ (cid:19) dµ +ZΩ
Fi(cid:18) dDu
dµ (cid:19) dµ +ZΩ
≤ZΩ
dµ (cid:19) dµ
Fi(cid:18) dDuj
j→∞ ZΩ
hFi, νxi dµ +ZΩ
=ZΩ
≤ lim inf
hF ∞
i
, ν∞
x i dλν.
On the other hand, by Lebesgue's dominated convergence theorem, as well
as the fact that F ∈ Q(RN ×n) ∩ R(Ω; RN ×n) ⊂ E(Ω; RN ×n),
lim
i→∞(cid:18)ZΩ
hFi, νxi dµ +ZΩ
hF ∞
i
, ν∞
x i dλν(cid:19) =ZΩ
hF ∞, ν∞
x i dλν
hF, νxi dµ +ZΩ
F (cid:18) dDuj
j→∞ZΩ
dµ (cid:19) dµ.
= lim
(3.24)
21
By combining (3.23) and (3.24), we get the desired estimate from below.
3.2 Estimate from above
Recall from (3.1) the definition of the functional F∗ by relaxation. We prove
that the estimate from above holds for the integral representation of F∗. Here
our proof is not based on the theory of Young measures, so we can allow for
somewhat weaker assumptions on F .
Proposition 3.5. Let Ω ⊂ Rn be a bounded open set, let µ ∈ M+(Ω) with
Ln ≪ µ, let F ∈ Q(RN ×n) with
0 ≤ F (A) ≤ M(1 + A), A ∈ RN ×n
for some M ≥ 1, and let u ∈ BV(Ω; RN ). Then we have
F∗(u, Ω) ≤ZΩ
F (cid:18) dDu
dµ (cid:19) dµ +ZΩ
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu.
Proof. Again, by Definition 3.1 and [12, Lemma 6.3] we can find a sequence
Fi ∈ SQ(RN ×n) with parameters i ∈ N, ri > 0 such that Fi(A) ց F (A) and
F ∞
i (A) ց F ∞(A) pointwise for every A ∈ RN ×n as i → ∞. Moreover, by
making M slightly larger, if necessary, we have that 0 ≤ Fi(A) ≤ M(1 + A)
for all i ∈ N and A ∈ RN ×n. Fix i ∈ N.
The proof is based on mollifying the function u in a small set. Take a
Borel set D ⊂ Ω with Ds,µu(Ω \ D) = 0 and µ(D) = 0. Then take an open
set G ⊃ D with Ln(G) and µ(G) so small that
ZG
M(1 + ∇u) dLn +ZG
M(cid:12)(cid:12)(cid:12)(cid:12)
dDsu
dµ (cid:12)(cid:12)(cid:12)(cid:12)
dµ + M(ri + i)Ln(G) + M(1 + ri)µ(G)
(3.25)
is less than 1/i; this is possible by the absolute continuity of integrals. By
Lemma 2.1 we can pick a sequence (vj) ⊂ BVu(G; RN ) ∩ C ∞(G; RN ) (note
boundary values) that converges to u h·i-strictly in BV(G; RN ). Fix also
j ∈ N.
Using the linear growth of Fi, we estimate
ZG
F (cid:18) dDvj
dµ (cid:19) dµ ≤ZG
Fi(cid:18)dDvj
≤ZG∩{∇vj /a>ri}
dµ (cid:19) dµ =ZG
Fi(cid:18)∇vj
Fi(cid:18)∇vj
a (cid:19) a dLn
a (cid:19) a dLn + M(1 + ri)µ(G),
22
where by the fact that F (A) = F ∞(A) − i for A ≥ ri, the last integral
equals
ZG∩{∇vj/a>ri}(cid:18)F ∞
i (cid:18) ∇vj
F ∞
i (cid:18)∇vj
a (cid:19) a dLn
i
F ∞
(∇vj) dLn
i (∇vj) dLn
F ∞
a (cid:19) − i(cid:19) a dLn ≤ZG∩{∇vj /a>ri}
=ZG∩{∇vj/a>ri}
≤ZG
≤ZG∩{∇vj >ri}
=ZG∩{∇vj>ri}
≤ZG
F ∞
i
Fi (∇vj) dLn + M(ri + i)Ln(G).
(∇vj) dLn + MriLn(G)
(Fi (∇vj) + i) dLn + MriLn(G)
Now, since Fi ∈ SQ(RN ×n) ⊂ E(G; RN ×n) (constant in the x-variable) and
vj → u h·i-strictly in BV(G; RN ), we can apply Reshetnyak's continuity
theorem, Theorem 2.2, to obtain
lim inf
j→∞ ZG
F ∞
≤ lim inf
dµ (cid:19) dµ
Fi (∇vj) dLn + M(ri + i)Ln(G) + M(1 + ri)µ(G)
F (cid:18)dDvj
j→∞ ZG
Fi(∇u) dLn +ZG
=ZG
M(1 + ∇u) dLn +ZG
≤ZG
+ZG
≤ZG
i (cid:18) dDsu
dDsu(cid:19) dDsu
dµ (cid:12)(cid:12)(cid:12)(cid:12)
M(cid:12)(cid:12)(cid:12)(cid:12)
i (cid:18) dDs,µu
dDs µu(cid:19) dDs,µu + M(ri + i)Ln(G) + M(1 + ri)µ(G)
i (cid:18) dDs,µu
dDs µu(cid:19) dDs,µu + 1/i
+ M(ri + i)Ln(G) + M(1 + ri)µ(G)
dDsu
F ∞
F ∞
dµ
by (3.25). Then define for each j ∈ N
uj :=( vj
u
in G,
in Ω \ G.
The fact that vj ∈ BVu(G; RN ) implies by definition (given before Lemma
2.1) that Duj = Du Ω \ G + Dvj G. Thus it is clear that uj ∈ W 1,1
µ (Ω; RN ),
23
and also uj → u in L1(Ω; RN ), so that uj is an admissible sequence for
F∗(u, Ω). In total, we obtain
F∗(u, Ω) ≤ lim inf
F (cid:18)dDuj
dµ (cid:19) dµ
j→∞ ZΩ
F (cid:18) dDvj
dµ (cid:19) dµ +ZΩ\G
j→∞ ZG
i (cid:18) dDs,µu
dDs µu(cid:19) dDs,µu +ZΩ\G
i (cid:18) dDs,µu
dµ (cid:19) dµ +ZΩ
Fi(cid:18)dDu
F (cid:18)dDu
dµ (cid:19) dµ
dµ (cid:19) dµ + 1/i
F (cid:18)dDu
dDs µu(cid:19) dDs,µu + 1/i.
F ∞
= lim inf
F ∞
≤ZG
≤ZΩ
Letting i → ∞, by Lebesgue's monotone or dominated convergence we get
the desired estimate from above.
3.3 Some examples
Let us briefly consider why it is necessary to assume that Ln ≪ µ, at least
in order to obtain the integral representation (3.2). The reason is that the
estimate from above may be violated without this assumption. We note that
the integral representation (3.2) always takes a value at most
Mµ(Ω) + MDu(Ω),
which is finite for a BV function u ∈ BV(Ω; RN ). On the other hand, if
it is not true that Ln ≪ µ, then there can be a large set not "seen" by
the measure µ, and as a result it may simply be impossible to approximate
certain BV functions in the L1-sense by functions in the class W 1,1
µ (Ω; RN ).
Consider the following examples.
Example 3.6. Suppose that there is an open set B ⊂ Ω (which we can
assume to be a ball) with µ(B) = 0 but of course Ln(B) > 0. Take a
nonconstant u ∈ C 1
µ (Ω) satisfy
Duj(B) = 0 and are thus constant in the ball B. Thus there is no se-
quence of functions uj ∈ W 1,1
µ (Ω) with uj → u in L1(Ω), and consequently
F∗(u, Ω) = ∞.
c (B), and note that all functions uj ∈ W 1,1
Even if the support of µ is the whole of Ω, the estimate from above may
fail.
Example 3.7. Take Ω to be the open unit square on the plane, and let
A ⊂ Ω be a "fat" Sierpinski carpet, with L2(A) = 1/2. Then define the
24
weight w = 1Ω\A, and µ := w L2. Clearly the absolute continuity assumption
L2 ≪ µ is violated, but the support of µ is the whole of Ω. By using the
properties of BV functions restricted to lines, see e.g.
[5, Section 3.11], we
obtain that any function v ∈ W 1,1
µ (Ω) is constant almost everywhere in A. If
we define a BV function u ∈ BV(Ω) e.g. as u(x, y) := x, there is no sequence
uj ∈ W 1,1
µ (Ω) for which uj → u in L1(Ω), and consequently F∗(u, Ω) = ∞.
However, it is not clear to us whether the assumption Ln ≪ µ, or the
assumption on the integrand F ∈ R(Ω; RN ×n), are necessary in our main
result, Theorem 1.1.
4 The lower semicontinuity theorem
From the integral representation, we obtain the following lower semicontinu-
ity result.
Proposition 4.1. Let Ω ⊂ Rn be a bounded open set with Ln(∂Ω) = 0, let
µ ∈ M+(Ω) with Ln ≪ µ, and let F ∈ R(Ω; RN ×n) ∩ Q(RN ×n) with
mA ≤ F (A) ≤ M(1 + A)
for some 0 < m ≤ M. Then the functional
F (u) :=ZΩ
F (cid:18)dDu
dµ (cid:19) dµ +ZΩ
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) dDs,µu,
is lower semicontinuous with respect to convergence in L1(Ω; RN ).
u ∈ BV(Ω; RN ),
Proof. The relaxed functional F∗(u, Ω) given in (3.1) is obviously lower semi-
continuous with respect to convergence in L1(Ω; RN ), and by Proposition
3.2 and Proposition 3.5 it equals the functional F (u) given in this proposi-
tion.
We recall Jensen's inequalities for gradient Young measures with respect
to the Lebesgue measure Ln, given in Theorem 2.5. We can now partially
generalize these inequalities to the case of a general measure µ.
Theorem 4.2. Let Ω ⊂ Rn be a bounded open set with Ln(∂Ω) = 0, let
µ ∈ M+(Ω) with Ln ≪ µ, let F ∈ R(Ω; RN ×n) ∩ Q(RN ×n) with F ≥ 0,
let u ∈ BV(Ω; RN ), and let ν ∈ Y(Ω, µ; RN ×n) be a gradient Young measure
with λν(∂Ω) = 0 and with barycenter Du. Then the following hold:
F (cid:18)dDu
F ∞(cid:18) dDs,µu
dµ (cid:19) ≤ hF, νxi + hF ∞, ν∞
dDs,µu(cid:19) Ds,µu ≤ hF ∞, ν∞
x i
dλν
dµ
x iλs,µ
ν
(x)
for µ-almost every x ∈ Ω,
(4.1)
as measures.
(4.2)
25
Proof. Take a sequence (uj) ⊂ BV(Ω; RN ) that generates ν. We know
∗⇁ u in BV(Ω; RN ), see the discussion after (2.7). Note that F ∈
that uj
R(Ω; RN ×n) ∩ Q(RN ×n) ⊂ E(Ω; RN ×n) (constant in the x-variable), so that
F necessarily has linear growth F (A) ≤ M(1 + A) for some M ≥ 0. Let us
first also assume that F has the coercivity property mA ≤ F (A) for some
m > 0 and all A ∈ RN ×n. By combining our lower semicontinuity result,
Proposition 4.1, with the fact that F ∈ E(Ω; RN ×n), we obtain
ZΩ
F (cid:18)dDu
≤ lim inf
F ∞(cid:18) dDs,µu
dµ (cid:19) dµ +ZΩ
j→∞ (cid:18)ZΩ
F (cid:18) dDuj
hFi, νxi dµ +ZΩ
=ZΩ
dDs,µu(cid:19) dDs,µu
dµ (cid:19) dµ +ZΩ
, ν∞
x i dλν.
hF ∞
i
F ∞(cid:18) dDs,µuj
dDs,µuj(cid:19) dDs,µuj(cid:19)
We can equally well write the above inequality in any open U ⊂ Ω (in
particular, a ball) with λν(∂U) = 0. Thus we can differentiate the inequality
with respect to µ, and obtain (4.1) by the Besicovitch differentiation theorem
(see e.g. [5, Theorem 2.22]). By writing the above inequality for balls from
a suitable Vitali covering of Ω, we obtain (4.2).
The general case can be obtained by writing (4.1) and (4.2) for integrands
Fi(A) := max{F (A), A/i},
i ∈ N,
and letting i → ∞.
Corollary 4.3. With Ω, µ, u, and ν as in the previous theorem, there exist
sets E1, E2 ⊂ Ω with µ(E1) = 0 and Ds,µu(E2) = 0 such that for every
F ∈ R(Ω; RN ×n) ∩ Q(RN ×n) with F ≥ 0, we have
F (cid:18)dDu
dµ (cid:19) ≤ hF, νxi + hF ∞, ν∞
F ∞(cid:18) dDs,µu
dDs,µu(cid:19) ≤ hF ∞, ν∞
x i
x i
(x)
dλν
dµ
dλs,µ
ν
dDs,µu
for every x ∈ Ω \ E1,
(4.3)
for every x ∈ Ω \ E2.
(4.4)
The point is that we can find exceptional sets that do not depend on the
integrand F .
Proof. Again, we note that R(Ω; RN ×n) ∩ Q(RN ×n) ⊂ E(Ω; RN ×n) (constant
in the x-variable). Recalling the transformation T given in Section 2.3, we
have that
{T (F ) : F ∈ R(Ω; RN ×n) ∩ Q(RN ×n), F ≥ 0}
26
contains a countable dense subset {Gi}i∈N, since it is contained in the sep-
arable space C(BN ×m). Then (4.3) and (4.4) hold for some choice of sets
E1, E2 ⊂ Ω with µ(E1) = 0 and Ds,µu(E2) = 0, and with F = T −1Gi for
any i ∈ N. It is easy to see for any F ∈ R(Ω; RN ×n) ∩ Q(RN ×n), F ≥ 0 that
Fk(cid:18) dDu
dµ
(x)(cid:19) − F (cid:18)dDu
dµ
(x)(cid:19) → 0
for every x ∈ Ω \ E1, for a sequence (Fk) ⊂ {T −1Gi}i∈N with T (Fk) → T (F )
in C(BN ×m). The other terms are handled similarly, and so we get the desired
inequalities.
Now we can prove our semicontinuity result, where we also allow for
x-dependence of the integrand. The result could also be given without a
boundary term, but its inclusion simplifies our proof. In the case µ = Ln, an
analogous result was given in [14, Theorem 10].
Theorem 4.4. Let Ω ⊂ Rn be a bounded Lipschitz domain with inner bound-
ary normal νΩ, let µ ∈ M+(Ω) with Ln ≪ µ, and let F ∈ R(Ω; RN ×n) be
nonnegative and µ × B(RN ×n)-measurable such that A 7→ F (x, A) is quasi-
convex for each fixed x ∈ Ω. Then the functional
dDu
F (cid:18)x,
F (u) :=ZΩ
F ∞(cid:18)x,
+Z∂Ω
u
u
dµ (cid:19) dµ +ZΩ
F ∞(cid:18)x,
⊗ νΩ(cid:19) u dHn−1
dDs,µu
dDs,µu(cid:19) dDs,µu
is weakly* sequentially lower semicontinuous in BV(Ω; RN ).
Note that in the last term, u is a boundary trace, see e.g. [5, Section 3.7].
∗⇁ u in BV(Ω; RN ). Take a bounded Lipschitz domain Ω′ ⋑ Ω,
Proof. Let uj
j, ue the zero extensions of uj, u to Ω′ \ Ω. Since Ω is a
and denote by ue
bounded Lipschitz domain, we can use standard gluing theorems for BV
functions, see e.g.
[5, Proposition 3.21, Theorem 3.84, Theorem 3.86], to
obtain that ue
j ∈ BV(Ω′; RN ) with
Due
j = ∇uj Ln Ω + Dsuj + uj ⊗ νΩ Hn−1 ∂Ω
and kujkL1(∂Ω;RN ) ≤ CkujkBV(Ω;RN ) with C depending only on Ω; and simi-
larly for ue. By the weak* convergence, uj is a norm-bounded sequence in
j is a norm-bounded sequence in BV(Ω′; RN )
BV(Ω; RN ), so we have that ue
∗⇁ ue in BV(Ω′; RN ).
and that ue
j → ue in L1(Ω′; RN ). This implies that ue
j
27
Since F ∈ R(Ω, RN ×n), F ∞(x, A) is continuous on Ω × ∂BN ×n, which
is a compact set. By the Tietze extension theorem, we can extend F ∞
to Ω′ × ∂BN ×n as a continuous nonnegative function (F e)∞.
If we define
F e(x, tA) := t(F e)∞(x, A) for any t ≥ 0, A ∈ RN ×n, and x ∈ Ω′, we see
that our notation is consistent in that the recession function of F e is indeed
(F e)∞. We also extend µ by µe := µ Ω + LN (Rn \ Ω). Then we see that
F e ∈ R(Ω′; RN ×n) is nonnegative and µe × B(RN ×n)-measurable. We write
F e(ue
j) : =ZΩ′
=ZΩ
dDue
j
dDuj
F e(cid:18)x,
F (cid:18)x,
+Z∂Ω
dµe (cid:19) dµe +ZΩ′
dµ (cid:19) dµ +ZΩ
F ∞(cid:18)x,
uj
uj
(F e)∞(cid:18)x,
F ∞(cid:18)x,
⊗ νΩ(cid:19) uj dHn−1
= F (uj),
dDs,µeue
j
dDs,µeue
j(cid:19) dDs,µe
ue
j
dDs,µuj
dDs,µuj(cid:19) dDs,µuj
and similarly for ue. We conclude that we need to prove that F e(ue) ≤
lim inf j→∞ F e(ue
j). Pick first a subsequence (not relabeled) that gives this
limit, and then by Theorem 2.3 and Corollary 2.4 we can pick a further sub-
sequence (not relabeled) such that the sequence Due
j generates a generalized
Young measure ν = (νx, λν, ν∞
x ), with respect to µe. Clearly λν(∂Ω′) = 0,
and then the barycenter of ν is Due, see the discussion after (2.7). Note that
for any fixed x ∈ Ω, F e(x, ·) ∈ R(Ω, RN ×n) ∩ Q(RN ×n), so that we can apply
Corollary 4.3 to obtain
lim inf
j→∞
F e(ue
j) =ZΩ′
F e(cid:18)x,
hF e(x, ·), νxi dµe +ZΩ′
dµe (cid:19) dµe +ZΩ′
(F e)∞(cid:18)x,
dDue
≥ZΩ′
= F e(ue).
h(F e)∞(x, ·), ν∞
x i dλν(x)
dDs,µeue
dDs,µeue(cid:19) dDs,µe
ue
References
[1] G. Alberti, Rank one property for derivatives of functions with
bounded variation, Proc. Roy. Soc. Edinburgh Sect. A 123 (1993),
no. 2, 239 -- 274.
[2] J. J. Alibert and G. Bouchitt´e, Non-uniform integrability and gen-
eralized Young measures, J. Convex Anal. 4 (1997), no. 1, 129 -- 147.
28
[3] L. Ambrosio, G. Buttazzo, and I. Fonseca, Lower semicontinuity
problems in Sobolev spaces with respect to a measure, J. Math.
Pures Appl. (9) 75 (1996), no. 3, 211 -- 224.
[4] L. Ambrosio and G. Dal Maso, On the relaxation in BV(Ω; Rm) of
quasi-convex integrals, J. Funct. Anal. 109 (1992), no. 1, 76 -- 97.
[5] L. Ambrosio, N. Fusco, and D. Pallara, Functions of bounded vari-
ation and free discontinuity problems, Oxford Mathematical Mono-
graphs. The Clarendon Press, Oxford University Press, New York,
2000. xviii+434 pp.
[6] J. Ball, B. Kirchheim, and J. Kristensen, Regularity of quasiconvex
envelopes, Calc. Var. Partial Differential Equations 11 (2000), no.
4, 333 -- 359.
[7] B. Dacorogna, Direct methods in the calculus of variations, Second
edition. Applied Mathematical Sciences, 78. Springer, New York,
2008. xii+619 pp.
[8] R. J. DiPerna and A. J. Majda, Oscillations and concentrations in
weak solutions of the incompressible fluid equations, Comm. Math.
Phys. 108 (1987), no. 4, 667 -- 689.
[9] L. Evans and R. Gariepy, Measure theory and fine properties of
functions, Studies in Advanced Mathematics. CRC Press, Boca
Raton, FL, 1992. viii+268 pp.
[10] I. Fonseca and S. Muller, Relaxation of quasiconvex functionals in
BV(Ω, Rp) for integrands f (x, u, ∇u), Arch. Rational Mech. Anal.
123 (1993), no. 1, 1 -- 49.
[11] H. Hakkarainen, J. Kinnunen, P. Lahti, and P. Lehtela, Relaxation
and integral representation for functionals of linear growth on met-
ric measures spaces, preprint 2014.
[12] B. Kirchheim and J. Kristensen, On Rank-One Convex Functions
that are homogeneous of Degree One, preprint 2015.
[13] J. Kristensen, Lower semicontinuity in spaces of weakly differen-
tiable functions, Math. Ann. 313 (1999), no. 4, 653710.
[14] J. Kristensen and F. Rindler, Characterization of generalized gradi-
ent Young measures generated by sequences in W 1,1 and BV, Arch.
29
Ration. Mech. Anal. 197 (2010), no. 2, 539 -- 598. Erratum, Ibid.
203 (2012), 693 -- 700.
[15] J. Kristensen and F. Rindler, Relaxation of signed integral func-
tionals in BV, Calc. Var. Partial Differential Equations 37 (2010),
no. 1-2, 29 -- 62.
[16] S. Muller, On quasiconvex functions which are homogeneous of de-
gree 1, Indiana Univ. Math. J. 41 (1992), no. 1, 295 -- 301.
[17] Y. G. Reshetnyak, The weak convergence of completely additive
vector-valued set functions, Sibirsk. Mat. J. 9 1968, 1386 -- 1394.
[18] F. Rindler, Lower semicontinuity and Young measures in BV with-
out Alberti's rank-one theorem, Adv. Calc. Var. 5 2012, no. 2, 127 --
159.
[19] W.P. Ziemer, Weakly differentiable functions. Sobolev spaces and
functions of bounded variation Graduate Texts in Mathematics,
120. Springer-Verlag, New York, 1989.
30
|
1501.00785 | 1 | 1501 | 2015-01-05T08:38:43 | s-Numbers sequences for homogeneous polynomials | [
"math.FA"
] | We extend the well known theory of $s$-numbers of linear operators to homogeneous polynomials defined between Banach spaces.
Approximation, Kolmogorov and Gelfand numbers of polynomials are introduced and some well-known results of the linear and multilinear settings are obtained for homogeneous polynomials. | math.FA | math |
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
ERHAN C¸ ALIS¸KAN AND PILAR RUEDA
Abstract. We extend the well known theory of s-numbers of linear operators to ho-
mogeneous polynomials defined between Banach spaces. Approximation, Kolmogorov
and Gelfand numbers of polynomials are introduced and some well-known results of
the linear and multilinear settings are obtained for homogeneous polynomials.
1. Introduction
A. Pietsch [18] introduced s-numbers as a tool for the study of linear operators
between Banach spaces. The success of the linear operator theory gave rise to consider
a multilinear and polynomial analogue that was proposed firstly by A. Pietsch and
followed by many researchers in the last decades (see [8] and the references therein). The
theory of s-numbers of multilinear operators among Banach spaces has been recently
developed by D. L. Fernandez, M. Mastylo and E. B. da Silva [13]. While the properties
of s-numbers of linear operators are well-known, the analogous theory of homogeneous
polynomials has not been checked as far as it should have been. We mention only the
particular case by A. Brauns, H. Junek and E. Plewnia [6] and the unpublished [7].
The aim of this paper is to elaborate the corresponding theory of s-numbers in the
context of homogeneous polynomials. It is worth mentioning that in many situations
dealing with polynomials instead of multilinear mappings has proved to be a subtle
subject that has needed different approaches and that has yielded to different results.
For instance, when trying to generalize absolutely summing operators to a non linear
context, different approaches have been required, and whereas factorization theorems
have been stated in the multilinear case, the search for a factorization scheme for dom-
inated polynomials has turned out to be difficult and only partial results have been
obtained (see [2, 3, 4]. The main purpose of the present paper is to undertake a study
of the basics on s-number sequences for polynomials, that include: approximation
numbers an and their relation with the adjoint and the biadjoint of a homogeneous
Date: July 13, 2021.
2010 Mathematics Subject Classification. Primary: 46G20; Secondary 46B28, 46G25.
Key words and phrases. Banach spaces, homogeneous polynomials, s-numbers, approximation, Kol-
mogorov, Gelfand numbers, the measure of of non-compactness.
The second author was supported by MICINN MTM2011-22417.
1
2
E. C¸ ALIS¸KAN AND P. RUEDA
polynomial introduced by Aron and Schottenloher; Kuratowski and Hausdorff mea-
sures of non-compactness of homogeneous polynomials; Kolmogorov numbers dn and
their relation with the approximation numbers; Gelfand numbers cn, the equivalence
between P being compact and limn→∞ cn(P ) = 0 and their relation with Kolmogorov
numbers. Some of the proofs we present are inspired by the linear/multilinear ones,
whereas other use techniques from polynomial theory. Our aim is to present the theory
from the point of view of polynomials and relate it to the linear or multilinear case
with linearization techniques that will provide shorter proofs than coming from the
classical theory. Furthermore, the results obtained for homogeneous polynomials can
be considered extensions of the linear ones.
Section 2 is devoted to fix notation and state some basic definitions and preliminary
results. In a quite natural way, we introduce the notion of m−s-number sequence for m-
homogeneous polynomials and relate it with the classical s-number sequences of linear
operators. Section 3 contains the essentials of the n-th approximation number an(P )
of an homogeneous polynomial P and its coincidence with the n-th approximation
number of the adjoint P ∗ of P in the sense of [1]. Compactness of homogeneous
polynomials is treated in Section 4 by means of Kuratowski and Hausdorff measures
γ and γ respectively. We prove that γ(P ) ≤ γ(P ∗) and γ(P ∗) ≤ γ(P ) among other
inequalities. As an application, we recover the well-known result that a homogeneous
polynomial is compact if and only if its adjoint is compact. As an attempt to quantify
the non compactness character of a polynomial, we study the polynomial notion of
Kolmogorov numbers dn and the polynomial m-lifting property. In particular we prove
that dn(P ) = dn(PL), where PL is the linearization of an m-homogeneous polynomial
P defined on a Banach space X. Moreover, dn(P ) = an(P Q), where Q is the canonical
metric surjection from ℓ1(BX) onto X defined by Q({λx}) = Px∈BX
λxx, {λx} ∈
l1(BX). Finally, we deal with Gelfand's numbers cn adapted to the polynomial context.
We obtain characterizations of compactness of homogeneous polynomials, this time, in
terms of Gelfand numbers, and we prove that cn(P ∗) ≤ dn(P ), cn(P ) = dn(P ∗) and
cn(P ) ≤ 2√ncn(P ∗).
2. Notation and preliminaries
The symbol K represents the field of all real numbers or complex numbers, N repre-
sents the set of all positive integers.
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
3
The letters X, Y and Z will always represent (real or complex) Banach spaces. The
symbol BX represents the open unit ball of X and BX the closed unit ball. We denote
by X ∗ the dual Banach space of X, and by κX the canonical embedding of X into the
bidual X ∗∗ of X.
+
n=1
Given a subset C ⊂ X, let Γ(C) denote the closed balanced convex hull of C.
Let 1 ≤ p ≤ ∞, with the conjugate index p′ given by
1
p′ = 1 (where p′ = 1 if
p = ∞), let ℓp(X) (1 ≤ p < ∞) (resp., ℓ∞(X)) denote the set of all sequences (xn)∞
kxnkp < ∞ (resp., (xn)n is bounded), and let c0(X) denote the set
in X such that
n=1 in X such that xn −→ 0 in X.
of all sequences (xn)∞
∞X
1
p
n=1
m times
{z }
Y , defined by P (x) = A(x, . . . , x
Given a continuous m-linear mapping A : X × ··· × X → Y , the map P : X −→
) for every x ∈ X, is said to be a continuous m-
homogeneous polynomial. P(mX; Y ) will denote the vector space of all continuous m-
homogeneous polynomials from X into Y , which is a Banach space with norm kPk =
sup{kP (x)k : kxk ≤ 1}. When Y = K we will write P(mX) instead of P(mX; K) and
when m = 1, L(X; Y ) := P(1X; Y ) is the space of all continuous linear operators from
X to Y . Let P m := SX,Y P(mX; Y ), that is, P m is the class of all m-homogeneous
polynomials defined between Banach spaces. Denote by P := Sm P m the class of all
continuous homogeneous polynomials defined between Banach spaces.
Let P ∈ P(mX; Y ). We define the rank of P as the dimension of the linear span of
P (X) in Y : rank(P ) = dim([P (X)]).
In a natural way, we introduce the notion of an m − s-number sequence for m-
homogeneous continuous polynomials. Let m ∈ N and for each n ∈ N let sn : P m −→
[0,∞) be a mapping. The sequence s = (sn) is called an m − s-number sequence if the
following conditions are satisfied for any n, k ∈ N:
(S1) Monotonicity: For every P ∈ P(mX; Y ),
sn(SP T ) ≤ kSksn(P )kTkm.
(S4) Rank-property: Let P ∈ P(mX; Y ).
Furthermore, if m = 1 the following condition has to be added:
rank(P ) < n =⇒ sn(P ) = 0.
kPk = s1(P ) ≥ s2(P ) ≥ . . . ≥ 0.
(S2) Additivity: For every P, Q ∈ P(mX; Y ),
sk+n−1(P + Q) ≤ sk(P ) + sn(Q).
(S3) Ideal-property: For every P ∈ P(mX; Y ), S ∈ L(Y ; Z), T ∈ L(W ; X)
4
E. C¸ ALIS¸KAN AND P. RUEDA
(S5) Norming-property: sn(Id : ℓn
mapping on the n-dimensional Hilbert space ℓn
2 .
2 → ℓn
2 ) = 1, n ∈ N, where Id is the identity
If (sn) is an m− s-number sequence for each m ∈ N, then (sn) is called an s-number
sequence. Note that this notion coincides with the usual notion of s-number sequence
for linear operators whenever m = 1.
Given P ∈ P(mX; Y ), let PL denote the linearization of P ; that is the unique con-
tinuous linear operator PL : ⊗πs
m,sX → Y such that P (x) = PL(⊗mx). The correspon-
dence P ↔ PL determines an isometric isomorphism -- denoted by ImX,Y -- between
P(mX; Y ) and the space L(⊗πs
m,sX
to Y . Let Im : P m −→ L and I : P → L be the correspondences whose restrictions to
each component P(mX; Y ) is equal to ImX,Y .
m,sX → ⊗πs
m,sX; Y ) of all continuous linear operators from ⊗πs
If T ∈ L(X; Y ), let ⊗m,sT : ⊗πs
m,sY be the continuous linear map given by
⊗m,sT (⊗mx) = ⊗mT (x).
Our first interest is to relate the linear and the polynomial notions of s-number
sequences.
Proposition 2.1. If the mapping s = (sn) : L −→ [0,∞)N is an s-number sequence
(in the linear sense) then, s ◦ Im : P m −→ [0,∞)N is an m − s-number sequence.
Proof. We will pay attention just to the ideal property. This property follows from
the fact that (SP T )L = SPL ⊗m,s T and k ⊗m,s Tk = kTkm, for all P ∈ P(mX; Y ),
T ∈ L(W ; X) and S ∈ L(Y ; Z).
(cid:3)
Injectivity, surjectivity and multiplicativity have been proved useful tools for s-
number sequences. Let us extend these properties to the polynomial context.
(J) An m−s-number sequence s = (sn) is called injective if given any metric injection
j ∈ L(Y ; Z), i.e., kj(y)k = kyk for all y ∈ Y , sn(P ) = sn(jP ) for all P ∈ P(mX; Y )
and for all Banach spaces X.
(S) An m − s-number sequence s = (sn) is called surjective if given any metric
surjection q ∈ L(Z; X), i.e., q(BZ ) = BX , sn(P ) = sn(P q) for all P ∈ P(mX; Y ).
(M ) An s-number sequence s = (sn) is called multiplicative if, for u ∈ L(Y ; Z) and
P ∈ P(mX; Y ),
sk+n−1(u ◦ P ) ≤ sk(u)sn(P ), k, n ∈ N.
Proposition 2.2. Let s = (sn) : L −→ [0,∞)N be an s-number sequence for linear
operators.
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
5
(1) If s is injective then s ◦ Im is injective.
(2) If s is surjective then, s ◦ Im is surjective.
(3) If s is multiplicative then s ◦ Im is multiplicative.
Proof. (1) Let j ∈ L(Y ; Z) be a metric injection and P ∈ P(mX; Y ). Then,
sn ◦ Im(P ) = sn(PL) = sn(j ◦ PL) = sn((j ◦ P )L) = sn ◦ Im(j ◦ P ).
(2) Let q ∈ L(Z; X) be a metric surjection and P ∈ P(mX; Y ). If ⊗q : ⊗πs
⊗πs
m,sX denotes the linear map given by ⊗q(⊗x) := ⊗q(x), then
⊗ q(B⊗
m,sBX) = B⊗
Hence, for any P ∈ P(mX; Y ) and any metric surjection q ∈ L(Z; X) we have
m,sq(BZ )) = Γ(⊗πs
m,sZ ) = ⊗q(Γ(⊗πs
m,sBZ)) = Γ(⊗πs
πs
m,sZ −→
πs
m,sX.
sn ◦ Im(P ◦ q) = sn((P ◦ q)L) = sn(PL ◦ ⊗q) = sn(PL) = sn ◦ Im(P ).
(3) Let u ∈ L(Y ; Z) and P ∈ P(mX; Y ). Then,
sk+n−1 ◦ Im(u ◦ P ) = sk+n−1((u ◦ P )L) = sn+k−1(u ◦ PL)
≤ sk(u)sn(PL) = sk ◦ Im(u)sn ◦ Im(P ).
(cid:3)
The theory of ideals of homogeneous polynomials between Banach spaces has been
developed in the last decades by several authors, so the extension of the dual procedure
to polynomial ideals is a natural step.
In this paper we provide many results on
homogeneous polynomials in connection with their adjoint concerning measure of non-
compactness and s-numbers. First we need the definition of the adjoint of a continuous
homogeneous polynomial.
Definition 2.3. (Aron -- Schottenloher [1]) Given a continuous m-homogeneous poly-
nomial P ∈ P(mX; Y ) between the Banach spaces X and Y , the adjoint of P is the
following continuous linear operator:
P ∗ : Y ∗ −→ P(mX) , P ∗(ϕ)(x) = ϕ(P (x)).
It is clear that kP ∗k = kPk.
After this definition by R. Aron and M. Schottenloher, and after the works of R. Ryan
[21, 22], the adjoint of a polynomial became a standard tool in the study of spaces of
homogeneous polynomials and in infinite dimensional holomorphy (see, e.g. [9, 10, 17]
and references therein). We refer to [9] or [16] for the properties of polynomials in
infinite dimensional spaces, and to [15] for the theory of Banach spaces.
3. Approximation numbers of homogeneous polynomials
Similar to the linear case we define the n-th approximation number an(P ) of any
homogeneous polynomial P ∈ P(mX; Y ) by
6
E. C¸ ALIS¸KAN AND P. RUEDA
an(P ) := inf{kP − Qk : Q ∈ P(mX; Y ), rank(Q) < n}.
If we denote an(T ) := inf{kT − Lk : L ∈ L(X; Y ), rank(L) < n}, T ∈ L(X; Y ), then
an(P ) = an(PL).
If a = (an) is an s-number sequence on L, Proposition 2.1 gives that a = a ◦ Im is
an m − s-number sequence on P m. Therefore, a = a ◦ I is an s-number sequence.
Proposition 3.1. Let (sn) : P(mX; Y ) −→ [0,∞)N be an s-number sequence. Then
(i) For all P ∈ P(mX; Y ) we have sn(P ) ≤ an(P ), n ∈ N.
(ii) For all S ∈ L(Y ; Z), P ∈ P(mX; Y ) and all k, n ∈ N we have sk+n−1(SP ) ≤
s1(S)an(P ) and sk+n−1(SP ) ≤ ak(S)sn(P ).
Proof. (i) Let P ∈ P(mX; Y ). Then for any R ∈ P(mX; Y ) with rank(R) < n, we
have sn(P ) ≤ s1(P − R) + sn(R) = kP − Rk + sn(R) = kP − Rk. Hence, by definition
of an(P ) we have sn(P ) ≤ an(P ).
(ii) Let R ∈ P(mX; Y ) with rank(R) < n. Since rank(SR) < n, it follows that
sk+n−1(SP ) ≤ sk(S(P − R)) + sn(SR) = sk(S(P − R))
≤ kSksk(P − R)kIXkm ≤ kSks1(P − R) = kSkkP − Rk.
Hence by definition of (an) we get sk+n−1(SP ) ≤ s1(S)an(P ).
The proof of the second inequality can be obtained in a similar way.
(cid:3)
Remark 3.2. If P ∈ P(mX; Y ) has finite rank then PL has finite rank. Hence,
rank(P ) = rank(PL) = rank((PL)∗) = rank(ImX,K ◦ P ∗) = rank(P ∗).
It is worth mentioning that our use of polynomial techniques allows us to reduce to
the linear case many proofs instead of adapting all calculations to the new setting.
Proposition 3.3. Let m ≥ 2 and let X and Y be Banach spaces. For every polynomial
P ∈ P(mX; Y ) we have an(P ∗) ≤ an(P ), n ∈ N. Furthermore, if there exists a linear
projection π of norm 1 from Y ∗∗ onto κY (Y ) then, for every P ∈ P(mX; Y ) we have
that an(P ∗) = an(P ), n ∈ N.
Proof. Since ImX,K is an isometric isomorphism, it follows from, e.g., [19, p. 152, 11.7.3.
Proposition] that
an(P ∗) = an(ImX,K ◦ P ∗) = an((PL)∗) ≤ an(PL) = an(P ).
For the second assertion, we use the analogous property for linear operators [13,
Proposition 3.3] to get that
an(P ) = an(PL) = an(P ∗
L) = an(ImX,K ◦ P ∗) = an(P ∗).
(cid:3)
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
7
Remark 3.4. The technique we have used in this section makes use of the linearization
of continuous homogeneous polynomials. A similar technique works for continuous m-
linear mappings. For each integer m ∈ N, let L(X1, . . . , Xm; Y ) be the Banach space of
all continuous m-linear mappings A : X1× . . .× Xm 7−→ Y , endowed with the sup norm
kAk = sup{kA(x1, . . . , xm)k : kxik ≤ 1, i = 1, . . . , m}. If T ∈ L(X1, . . . , Xm; Y ) there
is a unique continuous linear operator TL ∈ L(X1 ⊗π ··· ⊗πXm; Y ) such that TL(x1 ⊗
. . . ⊗ xm) = T (x1, . . . , xm), and the correspondence T ↔ TL determines an isometric
isomorphism between L(X1, . . . , Xm; Y ) and L(X1 ⊗π ··· ⊗πXm; Y ). This could yield
to alternative proofs in [13] based in the well-known linear case.
4. Compactness of homogeneous polynomials
The results in this section shows that the natural extensions of Kuratowski and
Hausdoff measures to polynomials keeps the harmony between linear and non linear
theory.
Let X be a metric space. The Kuratowski measure α(A) of non-compactness of a
bounded set A ⊂ X is defined by
α(A) = inf{ε > 0 : A may be covered by finitely many sets of diameter ≤ ε}.
In case that we consider just finitely many balls of radius ≤ ε to cover A, the infimum
is called the Hausdorff ball measure β(A) of non-compactness of A, that is
β(A) = inf{ε > 0 : A may be covered by finitely many balls of radius ≤ ε}.
For every bounded set A we have that β(A) ≤ α(A) ≤ 2β(A).
Let X and Y be Banach spaces. Since continuous m-homogeneous polynomials are
bounded on bounded sets, we can extend the Kuratowski, and the Hausdorff measure
of non-compactness of linear operators to polynomials in a natural way: for any P ∈
P(mX; Y ) the Kuratowski and the Hausdorff measure, respectively, of non-compactness
of P is defined by
γ(P ) := α(P (BX )) and
eγ(P ) := β(P (B X))
Note that P is compact if and only if eγ(P ) = γ(P ) = 0.
Remark 4.1. (see [11, Theorem 2.9]) Suppose X and Y are Banach spaces and let
T ∈ L(X; Y ). Then γ(T ) ≤ eγ(T ∗) and γ(T ∗) ≤ eγ(T ).
Lemma 4.2. Let X be a Banach space. Then β(C) = β(Γ(C)) for any C ⊂ X. In
particular, β(Γ(P (BX ))) = γ(P ).
8
E. C¸ ALIS¸KAN AND P. RUEDA
Proof. Let ǫ, δ > 0. It suffices to be shown that if C can be covered by finitely many
balls of radius ǫ then, Γ(C) can be covered by finitely many balls of radius δ + ǫ.
Assume that there are x1, . . . , xN ∈ X such that C ⊂ ∪N
i=1xi + ǫBX . Take i, j ∈
α + β = 1} is compact, there are
i,j + δBX ). If xi ∈ xi + ǫBX, xj ∈ xj + ǫBX
{1, . . . , N}. Since the set Ci,j := {αxi + βxj :
i,j, . . . , zLi,j
z1
and k ∈ {1, . . . , Li,j} is such that αxi + βxj ∈ zk
i,j ∈ X such that Ci,j ⊂ ∪Li,j
k=1(zk
kαxi + β xj − zk
i,j + δBX then,
i,jk ≤ αkxi − xik + kαxi + βxj − zk
i,jk + βkxj − xjk
≤ αǫ + δ + βǫ = δ + ǫ.
i=1xi + ǫBX ) ⊂ ∪N
Hence,
Γ(∪N
Consider the m-homogeneous polynomial J : X −→ P(mX)∗ given by J(x)(B) :=
B(x), x ∈ X, B ∈ P(mX). Since P ∗∗◦J = κY ◦P , the following theorem can be proved
also with similar techniques to the ones given in [13, Theorem 2.1]. However, we will
i,j + (δ + ǫ)BX) ⊂ ∪N
i,j + (δ + ǫ)BX). (cid:3)
i,j=1(∪
i,j=1(∪
Li,j
k=1zk
Li,j
k=1zk
use polynomial techniques related to tensor products to show that the polynomial case
admits shorter proofs.
Theorem 4.3. Let m ≥ 2 and let X and Y be Banach spaces. Then
(1) γ(P ) ≤ eγ(P ∗) and γ(P ∗) ≤ eγ(P ),
(2) 1
2 γ(P ) ≤ γ(P ∗) ≤ 2γ(P ) and 1
2eγ(P ) ≤ eγ(P ∗) ≤ 2eγ(P ).
for every P ∈ P(mX; Y ).
Proof. (1) Since
PL(B⊗
πs
m,sX ) = PL(Γ(⊗m,sBX)) = Γ(PL(⊗m,sBX)) = Γ(P (BX )),
γ(PL) = α(PL(B⊗
πs
m,sX )) ≥ α(P (BX )) = γ(P ).
we obtain
Then,
γ(P ) ≤ γ(PL) ≤ γ(P ∗
L) = γ(ImX,K ◦ P ∗
L) = γ(P ∗),
where the first equality follows from being ImX,K an isometric isomorphism.
Now Lemma 4.2 gives that
γ(P ∗) = γ(ImX,K ◦ P ∗
(2) By using part (1),
L) = γ(P ∗
L) ≤ γ(PL) = β(PL(B⊗
πs
m,sX)) = β(Γ(P (BX ))) = γ(P ).
1
2 γ(P ) ≤ 1
2eγ(P ∗) = 1
2 β(P ∗(BY ∗)) ≤ 1
≤ eγ(P ) = β(P (BX )) ≤ α(P (BX )) ≤ 2α(P (B X)) = 2γ(P ),
2 α(P ∗(BY ∗)) ≤ α(P ∗(BY ∗)) = γ(P ∗)
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
and
1
2eγ(P ) = 1
2 β(P (B X)) ≤ 1
2 α(P (B X)) = 1
2 γ(P ) ≤ 1
2eγ(P ∗) ≤ eγ(P ∗)
= β(P ∗(BY ∗)) ≤ α(P ∗(BY ∗)) = γ(P ∗) ≤ eγ(P ) ≤ 2eγ(P ).
9
(cid:3)
As a consequence, we get R. Aron and M. Schottenloher result on compactness of
polynomials:
Corollary 4.4. [1, Proposition 3.6] Let m ≥ 2 and let X and Y be Banach spaces.
Then for every homogeneous polynomial P ∈ P(mX; Y ) we have that P is compact if
and only if its adjoint P ∗ is compact.
The next result generalizes [12, Proposition 2] (see [13, Theorem 3.1] for the multi-
linear case).
Proposition 4.5. Let m ≥ 2 and let X and Y be Banach spaces. Then
an(P ) ≤ an(P ∗∗) + 2eγ(P )
for all n ∈ N and all P ∈ P(mX; Y ).
Proof. an(P ) = an(PL) ≤ an(P ∗∗
I ∗
mX,K) + 2γ(P ) = an(P ∗∗) + 2γ(P ).
L ) + 2γ(PL) = an((ImX,K ◦ P ∗)∗) + 2γ(P ) = an(P ∗∗ ◦
(cid:3)
Corollary 4.6. Let m ≥ 2 and let X and Y be Banach spaces.
(1) If P ∈ P(mX; Y ) is a compact operator, then an(P ) = an(P ∗) for every n ∈ N.
(2) For every P ∈ P(mX; Y ) we have that an(P ) ≤ 5an(P ∗), n ∈ N.
Proof. (1) Since P ∗ is a linear continuous operator between Banach spaces, we have
an(P ∗∗) ≤ an(P ∗) (see, e.g., [19, p. 152, 11.7.3. Proposition]). If P is compact, then
eγ(P ) = 0 and hence, by Theorem 4.5 and Proposition 3.3, we get
an(P ) ≤ an(P ∗∗) + 2eγ(P ) = an(P ∗∗) ≤ an(P ∗) ≤ an(P ).
(2) an(P ) = an(PL) ≤ 5an(P ∗
L) = 5an(ImX,K ◦ P ∗) = 5an(P ∗).
(cid:3)
An alternative proof follows from the well known fact that P is compact if and only
if PL is compact (see [21]) and the corresponding property for linear operators, that is,
an(P ) = an(PL) = an(P ∗
L) = an(I ◦ P ∗) = an(P ∗), for all n ∈ N.
Let P ∈ P(mX; Y ). The quantity a(P ) := lim
an(P ) ≥ 0 does not help when trying
to measure the compactness of P . Even if P is approximable (and so compact) whenever
n→∞
a(P ) = 0 the converse is, in general, not true. If we consider the approximation property
(shortly, AP) on Y , then any compact m-homogeneous polynomial P ∈ P(mX; Y )
10
E. C¸ ALIS¸KAN AND P. RUEDA
can be approximated by finite-rank m-homogeneous polynomials (see [5, Proposition
2.5]). Hence, similarly to the (multi)linear case, if the space Y has the AP, then
P ∈ P(mX; Y ) is compact if and only if a(P ) = 0. However, by [1, Proposition 3.3]
(see also [17, Theorem 4.3]) we know that P(mX) has the AP if and only if, for every
Banach space Y , the space of all finite-rank polynomials Pf (mX; Y ) is norm-dense
in the space of all compact polynomials Pk(mX; Y ), or equivalently, any compact m-
homogeneous polynomial P ∈ P(mX; Y ) can be approximated arbitrarily and closely
by finite-rank m-homogeneous polynomials. Therefore if the space P(mX) has the AP,
then P ∈ P(mX; Y ) is compact if and only if a(P ) = 0. Let us remark that, there is a
reflexive separable Banach space X with basis such that P(2X) does not have the AP
(see [1]). Hence, for this space X, which has the AP, there is a Banach space Y such
that there is a compact polynomial P : X −→ Y which cannot be approximated by
finite-rank polynomials. Note that it turns out that this space Y also cannot have the
AP by [5, Proposition 2.5].
As in the (multi-)linear case, we use Kolmogorov numbers to measure how far a
polynomial is from being compact.
We define the n-th Kolmogorov number dn(P ) of a polynomial P ∈ P(mX; Y ) by
dn(P ) := inf{ε > 0 : P (BX) ⊂ Nε + εBY , Nε ⊂ Y, dim(Nε) < n}.
For P := T ∈ L(X; Y ) we write dn(T ) := edn(P ).
Kolmogorov numbers are related to approximation numbers via the equality dn(T ) =
an(T Q) , n ∈ N, T ∈ L(X; Y ). Recall that Q is the canonical metric surjection from
l1(BX) onto X, defined by
Q({λx}) = X
x∈BX
λxx,
{λx} ∈ l1(BX)
(see [19, p. 150-151], and for the multilinear case see [13, Theorem 4.1].) To get the
polynomial version of this result we will use the next proposition and the study of the
lifting property for polynomials by Gonz´alez and Guti´errez [14].
πs
Proposition 4.7. Given P ∈ P(mX; Y ), dn(P ) = dn(PL).
Proof. Clearly dn(P ) ≤ dn(PL). On the other hand, if P (BX) ⊂ Nǫ + ǫBY then
PL(B⊗
m,sX ) = Γ(P (BX )) ⊂ Γ(Nǫ + ǫBY ) = Nǫ + ǫBY ⊂ Nǫ + (ǫ + δ)B Y for all δ > 0.
Hence, dn(PL) ≤ ǫ + δ for all δ > 0 and so, dn(PL) ≤ dn(P ).
As in the linear case, P ∈ P(mX; Y ) is compact if and only if d(P ) := lim
dn(P ) = 0.
Also it is obvious that dn(P ) = 0 whenever rank(P ) < n. Propositions 2.1 and 4.7
n→∞
(cid:3)
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
11
imply that dn = dn ◦ I forms an s-number sequence. Proposition 3.1 implies that
dn(P ) ≤ an(P ), for every n ∈ N and Proposition 2.2 gives that ( dn) is a surjective
s-number sequence.
Let m ∈ N and let X be a Banach space. We say that X has the polynomial
m-lifting property if, for every continuous m-homogenous polynomial P from X to
any quotient space Y /N , there is eP ∈ P(mX; Y ) such that P = QY
N
denotes the canonical map of Y onto the quotient space Y /N . We say that X has
N eP , where QY
the polynomial metric m-lifting property if, for every ε > 0 and every continuous m-
homogenous polynomial P from X to any quotient space Y /N , there is eP ∈ P(mX; Y )
such that P = QY
N eP and kePk ≤ (1 + ε)kPk.
Proposition 4.8. Let m ∈ N. A Banach space X has the polynomial (metric) m-lifting
property if, and only if, ⊗πs
m,sX has the (respectively, metric) lifting property.
Proof. Assume first that X has the polynomial m-lifting property. Let T be a continu-
m,sX into some quotient space Y /N . Let P ∈ P(mX; Y /N )
N ◦ P = P .
N ◦ ( P )L ◦ δX = P , where δX is the m-homogeneous polynomial from X to
m,sX
ous linear operator from ⊗πs
be such that PL = T . By assumption, there is P ∈ P(mX; Y ) such that QY
Since QY
⊗πs
m,sX given by δX (x) = x ⊗ ··· ⊗ x (see [21]), then QY
has the lifting property.
N ◦ ( P )L = PL = T and ⊗πs
m,sX; Y /N ). By assumption, there is fPL ∈ L( ⊗πs
m,sX has the lifting property. Let P ∈ P(mX; Y /N ). Then
N ◦fPL =
m,sX; Y ) such that QY
We now assume that ⊗πs
PL ∈ L( ⊗πs
PL. Then eP := fPL ◦ δX satisfies P = QY
N ◦ eP .
The metric case follows from the fact that kPk = kPLk.
(cid:3)
As a consequence, if X has the polynomial metric m-lifting property, then for every
P ∈ P(mX; Y ) we have dn(P ) = dn(PL) = an(PL) = an(P ), for all n ∈ N.
Theorem 4.9. Let m ≥ 2 and let X and Y be Banach spaces. Let P ∈ P(mX; Y ),
and let Q be the canonical metric surjection from l1(BX) onto X. Then we have that
dn(P ) = an(P Q) , n ∈ N.
Proof. By [14, Theorem 1] l1(BX) has the polynomial metric lifting property and so,
from Proposition 4.8 ⊗πs
m,sl1(BX) has the metric lifting property. Then, dn(PL◦⊗mQ) =
an(PL ◦ ⊗mQ). Using that (dn) is surjective we get
dn(P ) = dn(PL) = dn(PL ◦ ⊗mQ) = an(PL ◦⊗mQ) = an((P ◦ Q)L) = an(P ◦ Q). (cid:3)
12
E. C¸ ALIS¸KAN AND P. RUEDA
Note that ( dn) is the largest surjective s-number sequence. Indeed, given any surjec-
tive s-number sequence (sn), then for the canonical metric surjection Q ∈ L(l1(BX); X)
we have, by Theorem 4.9, that for any P ∈ P(mX; Y )
sn(P ) = sn(P Q) ≤ an(P Q) = dn(P ).
By Propositions 2.2 and 4.7, ( dn) is multiplicative. As a consequence for any
surjective and multiplicative s-number sequence the following estimate holds for all
S ∈ L(Y ; Z) and all P ∈ P(mX; Y ):
sk+n−1(SP ) ≤ sk(S) dn(P ),
k, n ∈ N,
In fact, Theorem 4.9 yields
sk+n−1(SP ) = sk+n−1(SP Q) ≤ sk(S)sn(P Q) ≤ sk(S)an(P Q) = sk(S) dn(P ).
We end the paper with another example of s-number of homogeneous polynomi-
als, namely, Gelfand numbers, from which we will get alternative characterizations of
compactness of homogeneous polynomials. Motivated by [19, 11.5.1. Proposition] we
define the Gelfand numbers cn(P ) of an m-homogeneous polynomial P ∈ P(mX; Y ) by
cn(P ) := an(κY P ).
Clearly (cn) is an s-number sequence since (an) is an s-number sequence, and for
each n ∈ N we have that cn(P ) ≤ an(P ). We will just write cn(T ) := cn(P ) whenever
P = T ∈ L(X; Y ). Note that cn(P ) = cn(PL) for any P ∈ P(mX; Y ).
It follows
from Proposition 2.2 that (cn) is the largest injective s-number sequence, and satisfies
the multiplicavity property (M). We also have a polynomial version of Carl's mixing
multiplicavity of an injective s-number sequence (sn), that is, for all S ∈ L(Y ; Z) and
P ∈ P(mX; Y ), using Proposition 3.1 we get that sk+n−1(SP ) ≤ ck(S)sn(P ), k, n ∈ N.
cn(P ),
Considering the function c : P(mX; Y ) −→ [0,∞) given by c(P ) := lim
n−→∞
we have that c(P ) = c(PL). Now, compactness of homogeneous polynomials can be
quantified by means c and c as follows.
Proposition 4.10. Let m ≥ 2 and let X and Y be Banach spaces. The following
statements for a polynomial P ∈ P(mX; Y ) are equivalent.
(i) P is compact.
(ii) c(P ) = 0.
(iii) c(P ∗) = 0.
S-NUMBERS SEQUENCES FOR HOMOGENEOUS POLYNOMIALS
13
Proof. We know that P is compact if and only if PL is compact (see [21]), and P is
compact if and only if P ∗ is compact (see [1] or Corollary 4.4). Combining these facts
with [20, 2.4.11] we get the implications (i) ⇐⇒ (ii), and (i) ⇐⇒ (iii).
(cid:3)
Finally, following the lines of proof of [13, Theorem 5.1] we get the following result,
whose proof is omitted, which gives relation between Gelfand and Kolmogorov numbers
of polynomials.
Theorem 4.11. Let m ≥ 2 and let X and Y be Banach spaces. Then, for every
polynomial P ∈ P(mX; Y ) and n ∈ N we have that
(i) cn(P ∗) ≤ dn(P ),
(ii) cn(P ) = dn(P ∗),
(iii) cn(P ) ≤ 2√n cn(P ∗).
Acknowledgement: The authors are deeply indebted to R. Aron, who proposed the
research program on s-numbers for homogeneous polynomials and helped unselfishly
to improve the paper.
References
[1] R. M. Aron, M. Schottenloher: Compact holomorphic mappings on Banach spaces and the approx-
imation theory. J. Funct. Anal. 21 (1976), 7 -- 30.
[2] G. Botelho, D. Pellegrino, P. Rueda: Pietsch's factorization theorem for dominated polynomials. J.
Funct. Anal. 243 (2007), no. 1, 257269.
[3] G. Botelho, D. Pellegrino, P. Rueda: Preduals of spaces of homogeneous polynomials on Lp-spaces.
Linear Multilinear Algebra 60 (2012), no. 5, 565571.
[4] G. Botelho, D. Pellegrino, P. Rueda: On Pietsch measures for summing operators and dominated
polynomials. Linear Multilinear Algebra 62 (2014), no. 7, 860 -- 874.
[5] G. Botelho, L. Polac: A polynomial Hutton theorem with applications. - J. Math. Anal. Appl. 415
(2014), no. 1, 294 -- 301.
[6] H.-A. Braunss, H. Junek, E. Plewnia: Approximation numbers for polynomials. Finite or infinite
dimensional complex analysis (Fukuoka, 1999), 3546, Lecture Notes in Pure and Appl. Math., 214,
Dekker, New York, 2000.
[7] A. Brauns, H. Junek: Ideals of polynomials and multilinear mappings. - Unpublished Notes, 2003
[8] B. Carl: On s-numbers, quasi s-numbers, s-moduli and Weyl inequalities of operators in Banach
spaces. - Rev. Mat. Complut. 23:2, 2010, 467 -- 487.
[9] S. Dineen: Complex Analysis on Infinite Dimensional Spaces. Berlin. Springer Monographs in Math.
Springer, 1999.
[10] S. Dineen, J. Mujica: The approximation property for spaces of holomorphic functions on infinite
dimensional spaces II. - J. Funct. Anal. 259, 2010, 545 -- 560.
[11] D. E. Edmunds, W. D. Evans: Spectral theory and differential operators. - Oxford Math. Monogr.,
Oxford Science Publications, The Clarendon Press, Oxford Univ. Press, New York, 1987.
[12] D. E. Edmunds, H.-O. Tylli: On the entropy numbers of an operator and its adjoint. - Math.
Nachr. 126, 1986, 231 -- 239.
[13] D.L. Fernandez, M. Mastylo, E.B. da Silva: Quasi s-numbers and measures of non-compactness of
multilinear operators - Ann. Acad. Sci. Fenn. Math. 38, 2013, 805 -- 823.
[14] M. Gonz´alez, J. M. Guti´errez: Extension and lifting of polynomials. - Arch. Math. (Basel) 81,
2003, 431-438.
14
E. C¸ ALIS¸KAN AND P. RUEDA
[15] J. Lindenstrauss, L. Tzafriri: Classical Banach Spaces I. Sequence spaces. Berlin, Heidelberg, New
York. Springer-Verlag 1977.
[16] J. Mujica: Complex Analysis In Banach Spaces. Amsterdam. North-Holland Math. Stud. North-
Holland 1986.
[17] J. Mujica: Spaces of holomorphic functions and the approximation property. - IMI Graduate
Lecture Notes-1, Universidad Complutense de Madrid, 2009
[18] A. Pietsch: s-numbers of operators in Banach spaces. - Studia Math. 51, 1974, 201 -- 223.
[19] A. Pietsch: Operator ideals. - North-Holland, Amsterdam, 1980.
[20] A. Pietsch: Eigenvalues and s-numbers. - Cambridge Stud. Adv. Math. 13, Cambridge Univ. Press,
Cambridge, 1987.
[21] R. Ryan: Applications of topological tensor products to infinite dimensional holomorphy. Trinity
College, Ph.D. Thesis, Dublin, 1980.
[22] R. Ryan: Weakly compact holomorphic mappings on Banach spaces. - Pac. J. Math. 131, 1988,
179-190.
(E. C¸ alı¸skan) Yıldız Technical University, Faculty of Sciences and Arts, Department of
Mathematics, Davutpas¸a, 34210 Esenler, Istanbul, Turkey
E-mail address: [email protected]
(P. Rueda) Departamento de An´alisis Matem´atico, Universidad de Valencia, Doctor
Moliner 50, 46100 Burjasot (Valencia), Spain
E-mail address: [email protected]
|
1404.6479 | 2 | 1404 | 2015-12-16T20:58:59 | Fourier multipliers, symbols and nuclearity on compact manifolds | [
"math.FA",
"math.AP",
"math.SP"
] | The notion of invariant operators, or Fourier multipliers, is discussed for densely defined operators on Hilbert spaces, with respect to a fixed partition of the space into a direct sum of finite dimensional subspaces. As a consequence, given a compact manifold endowed with a positive measure, we introduce a notion of the operator's full symbol adapted to the Fourier analysis relative to a fixed elliptic operator. We give a description of Fourier multipliers, or of operators invariant relative to the elliptic operator. We apply these concepts to study Schatten classes of operators and to obtain a formula for the trace of trace class operators. We also apply it to provide conditions for operators between Lp-spaces to be r-nuclear in the sense of Grothendieck. | math.FA | math |
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON
COMPACT MANIFOLDS
JULIO DELGADO AND MICHAEL RUZHANSKY
Abstract. The notion of invariant operators, or Fourier multipliers, is discussed
for densely defined operators on Hilbert spaces, with respect to a fixed partition of
the space into a direct sum of finite dimensional subspaces. As a consequence, given
a compact manifold M endowed with a positive measure, we introduce a notion of
the operator's full symbol adapted to the Fourier analysis relative to a fixed elliptic
operator E. We give a description of Fourier multipliers, or of operators invariant
relative to E. We apply these concepts to study Schatten classes of operators on
L2(M ) and to obtain a formula for the trace of trace class operators. We also apply
it to provide conditions for operators between Lp-spaces to be r-nuclear in the sense
of Grothendieck.
1. Introduction
Let M be a closed manifold (i.e. a compact smooth manifold without bound-
ary) of dimension n endowed with a positive measure dx. Given an elliptic positive
pseudo-differential operator E of order ν on M, by considering an orthonormal basis
consisting of eigenfunctions of E we will associate a discrete Fourier analysis to the
operator E in the sense introduced by Seeley ([See65], [See69]). This analysis allows
us to introduce further a notion of invariant operators and of matrix-symbols corre-
sponding to those operators. The operators on M will be then analysed in terms of
the corresponding symbols relative to the operator E.
As a general framework, we first discuss invariant operators, or Fourier multipliers
in a general Hilbert space H. This notion is based on a partition of H into a direct
sum of finite dimensional subspaces, so that a densely defined operator on H can
be decomposed as acting in these subspaces. There are two main examples of this
construction discussed in the paper: operators on H = L2(M) for a compact manifold
M as well as operators on H = L2(G) for a compact Lie group G. The difference
in approaches to these settings is in the choice of partitions of H into direct sums
of subspaces:
in the former case they are chosen as eigenspaces of a fixed elliptic
pseudo-differential operator on M while in the latter case they are chosen as linear
spans of matrix coefficients of inequivalent irreducible unitary representations of G.
Date: October 13, 2018.
2010 Mathematics Subject Classification. Primary 35S05, 58J40; Secondary 22E30, 47B06,
47B10.
Key words and phrases. Compact manifolds, pseudo-differential operators, eigenvalues, Schatten
classes, nuclearity, trace formula.
The first author was supported by Marie Curie IIF 301599 and by the Leverhulme Grant RPG-
2014-02. The second author was supported by EPSRC grant EP/K039407/1. No new data was
collected or generated during the course of the research.
1
2
JULIO DELGADO AND MICHAEL RUZHANSKY
We note that for some results, the self-adjointness and ellipticity of E can be
dropped, see [RT15].
We give applications of these notions to the derivation of conditions characterising
those invariant operators on L2(M) that belong to Schatten classes. Furthermore,
we also give conditions for nuclearity on Lp-spaces and, more generally, for the r-
nuclearity of operators. While the theory of r-nuclear operators in general Banach
spaces has been developed by Grothendieck [Gro55] with numerous further advances
(e.g. in [HP10, Kon78, Olo72, Pie84, RL13]), in this paper we give conditions in terms
of symbols for operators to be r-nuclear from Lp1(M) to Lp2(M) for 1 ≤ p1, p2 < ∞
and 0 < r ≤ 1. Consequently, we determine relations between p1, p2, r and α ensuring
that the powers (I + E)−α are r-nuclear. Trace formulas are also obtained relating
operator traces to expressions involving their symbols.
In the recent work [DR14c] the authors found sufficient conditions for operators
to belong to Schatten classes Sp on compact manifolds in terms of their Schwartz
integral kernels. For p < 2, it is customary to impose regularity conditions on the
kernel because there are counterexamples to conditions formulated only in terms of
the integrability of kernels. Such examples go back to Carleman's work [Car16] and
their relevance to Schatten classes has been discussed in [DR14b]. A characteristic
feature of conditions of this paper is that no regularity is assumed neither on the
symbol nor on the kernel.
In the case of compact Lie groups, our results extend
results on Schatten classes and on r-nuclear operators on Lp spaces that have been
obtained in [DR13] and [DR14b]. We show this by relating the symbols introduced
in this paper to matrix-valued symbols on compact Lie groups developed in [RT13]
and in [RT10].
Schatten classes of pseudo-differential operators in the setting of the Weyl-Hor-
mander calculus have been considered in [Tof06], [Tof08], [BN04], [BN07], [BT10].
Conditions for symbols of lower regularity we given in [Sob14]. For the global analysis
of pseudo-differential operators on Rn see [BBR96], as well as [NR10, Chapter 4] also
for the basic general introduction to Schatten classes.
1
To formulate the notions more precisely, let H be a complex Hilbert space and let
T : H → H be a linear compact operator. If we denote by T ∗ : H → H the adjoint
of T , then the linear operator (T ∗T )
2 : H → H is positive and compact. Let (ψk)k
be an orthonormal basis for H consisting of eigenvectors of T = (T ∗T )
2 , and let
sk(T ) be the eigenvalue corresponding to the eigenvector ψk, k = 1, 2, . . . . The non-
negative numbers sk(T ), k = 1, 2, . . . , are called the singular values of T : H → H.
If 0 < p < ∞ and the sequence of singular values is p-summable, then T is said to
belong to the Schatten class Sp(H), and it is well known that each Sp(H) is an ideal
in L (H). If 1 ≤ p < ∞, a norm is associated to Sp(H) by
1
kT kSp = ∞Xk=1
p
(sk(T ))p! 1
.
If 1 ≤ p < ∞ the class Sp(H) becomes a Banach space endowed by the norm kT kSp.
If p = ∞ we define S∞(H) as the class of bounded linear operators on H, with
kT kS∞ := kT kop, the operator norm. For the Schatten class S2 we will sometimes
write kT kHS instead of kT kS2. In the case 0 < p < 1 the quantity kT kSp only defines
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
3
a quasi-norm, and Sp(H) is also complete. The space S1(H) is known as the trace
class and an element of S2(H) is usually said to be a Hilbert-Schmidt operator. For
the basic theory of Schatten classes we refer the reader to [GK69], [RS75], [Sim79],
[Sch70].
It is well known that the class S2(L2) is characterised by the square integrability
of the corresponding integral kernels, however, kernel estimates of this type are not
effective for classes Sp(L2) with p < 2. This is explained by a classical Carleman's
example [Car16] on the summability of Fourier coefficients of continuous functions
(see [DR14b] for a complete explanation of this fact). This obstruction explains the
relevance of symbolic Schatten criteria and here we will clarify the advantage of the
symbol approach with respect to this obstruction. With this approach, no regularity
of the kernel needs to be assumed.
In Section 6 we discuss the relation of our approach to that of the global analysis
on compact Lie groups. In particular, in the case of compact Lie groups the Fourier
coefficients can be arranged into a (square) matrix rather than in a column leading
to several simplifications. On general compact manifolds, this is not possible since
the multiplicities dj do not need to be all squares of integers.
We introduce ℓp-style norms on the space of symbols Σ, yielding discrete spaces
ℓp(Σ) for 0 < p ≤ ∞, normed for p ≥ 1. Denoting by σT the matrix symbol of an
invariant operator T provided by Theorem 4.1, Schatten classes of invariant operators
on L2(M) can be characterised concisely by conditions
and for 0 < p < ∞,
T ∈ L (L2(M)) ⇐⇒ σT ∈ ℓ∞(Σ),
T ∈ Sp(L2(M)) ⇐⇒ σT ∈ ℓp(Σ),
(1.1)
(1.2)
see (7.4) and (7.5). Here, the condition that T is invariant will mean that T is
strongly commuting with E (see Theorem 4.1). On the level of the Fourier transform
this means that
cT f (ℓ) = σ(ℓ)bf (ℓ)
for a family of matrices σ(ℓ), i.e. T assumes the familiar form of a Fourier multiplier.
In Section 2 in Theorem 2.1 we discuss the abstract notion of symbol for operators
densely defined in a general Hilbert space H, and give several alternative formulations
for invariant operators, or for Fourier multipliers, relative to a fixed partition of H
into a direct sum of finite dimensional subspaces,
H =Mj
Hj.
Consequently, in Theorem 2.3 we give the necessary and sufficient condition for the
bounded extendability of an invariant operator to L (H) in terms of its symbol,
and in Theorem 2.5 the necessary and sufficient condition for the operator to be in
Schatten classes Sr(H) for 0 < r < ∞, as well as the trace formula for operators in
the trace class S1(H) in terms of their symbols. As our subsequent analysis relies to
a large extent on properties of elliptic pseudo-differential operators on M, in Sections
3 and 4 we specify this abstract analysis to the setting of operators densely defined
on L2(M). The main difference is that we now adopt the Fourier analysis to a
4
JULIO DELGADO AND MICHAEL RUZHANSKY
fixed elliptic positive pseudo-differential operator E on M, contrary to the case of an
operator Eo ∈ L (H) in Theorem 2.2.
The notion of invariance depends on the choice of the spaces Hj. Thus, in the
analysis of operators on M we take Hj's to be the eigenspaces of E. However,
other choices are possible. For example, for H = L2(G) for a compact Lie group G,
choosing Hj's as linear spans of representation coefficients for inequivalent irreducible
unitary representations of G, we make a link to the quantization of pseudo-differential
operator on compact Lie groups as in [RT10]. These two partitions coincide when
inequivalent representations of G produce distinct eigenvalues of the Laplacian; for
example, this is the case for G = SO(3). However, the partitions are different when
inequivalent representations produce equal eigenvalues, which is the case, for example,
for G = SO(4). For the more explicit example on H = L2(Tn) on the torus see Remark
2.6. A similar choice could be made in other settings producing a discrete spectrum
and finite dimensional eigenspaces, for example for operators in Shubin classes on Rn,
see Chodosh [Cho11] for the case n = 1.
The analogous concept to Schatten classes in the setting of Banach spaces is the
It has applications to
notion of r-nuclearity introduced by Grothendieck [Gro55].
questions of the distribution of eigenvalues of operators in Banach spaces.
In the
setting of compact Lie groups these applications have been discussed in [DR14b] and
they include conclusions on the distribution or summability of eigenvalues of operators
acting on Lp-spaces. Another application is the Grothendieck-Lidskii formula which
is the formula for the trace of operators on Lp(M). Once we have r-nuclearity, most
of further arguments are then purely functional analytic, so they apply equally well
in the present setting of closed manifolds. Because of this we omit the repetition of
statements and refer the reader to [DR14b] for further such applications.
Some results of this paper have been announced in [DR14a], so here we provide
their proofs, including a correction to the formulation of [DR14a, Theorem 3.1, (iv)]
given by Theorem 4.1, (iv), of this paper.
The paper is organised as follows. In Section 2 we discuss Fourier multipliers and
their symbols in general Hilbert spaces. In Section 3 we associate a global Fourier
analysis to an elliptic positive pseudo-differential operator E on a closed manifold M.
In Section 4 we introduce the class of operators invariant relative to E as well as their
matrix-valued symbols, and apply this to characterise invariant operators in Schatten
classes in Section 5.
In Section 6 we relate the analysis developed so far to the
analysis on compact Lie groups from [RT13], [RT10], and establish formula relating
their matrix symbols in the case when M is a compact Lie group. In particular, we
will see that left-invariant operators on compact Lie groups are invariant in our sense.
In Section 7 we analyse the integral kernels of invariant operators on general closed
manifolds. Finally, in Section 8 we apply our analysis to study r-nuclear operators
on Lp-spaces.
Throughout the paper, we denote N0 = N ∪ {0}. Also δjℓ will denote the Kronecker
delta, i.e. δjℓ = 1 for j = ℓ, and δjℓ = 0 for j 6= ℓ.
The authors would like to thank V´eronique Fischer, Alexandre Kirilov, and Au-
gusto Almeida de Moraes Wagner for comments.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
5
2. Fourier multipliers in Hilbert spaces
In this section we present an abstract set up to describe what we will call invariant
operators, or Fourier multipliers, acting on a general Hilbert space H. We will give
several characterisations of such operators and their symbols. Consequently, we will
apply these notions to describe several properties of the operators, in particular, their
boundedness on H as well as the Schatten properties.
We note that direct integrals (sums in our case) of Hilbert spaces have been inves-
tigated in a much greater generality, see e.g. Bruhat [Bru68], Dixmier [Dix96, Ch 2.,
§2], [Dix77, Appendix]. The setting required for our analysis is much simpler, so we
prefer to adapt it specifically for consequent applications, also providing short proofs
for our statements.
The main application of the constructions below will be in the setting when M
is a compact manifold without boundary, H = L2(M) and H∞ = C ∞(M), which
will be described in detail in Section 3. However, several facts can be more clearly
interpreted in the setting of abstract Hilbert spaces, which will be our set up in this
section. With this particular example in mind, in the following theorem, we can think
of {ek
j } being an orthonormal basis given by eigenfunctions of an elliptic operator on
M, and dj the corresponding multiplicities. However, we allow flexibility in grouping
the eigenfunctions in order to be able to also cover the case of operators on compact
Lie groups.
Theorem 2.1. Let H be a complex Hilbert space and let H∞ ⊂ H be a dense linear
subspace of H. Let {dj}j∈N0 ⊂ N and let {ek
j }j∈N0,1≤k≤dj be an orthonormal basis of
H such that ek
k=1, and let Pj : H → Hj
be the orthogonal projection. For f ∈ H, we denote
j ∈ H∞ for all j and k. Let Hj := span{ek
j }dj
bf (j, k) := (f, ek
j )H
and let bf (j) ∈ Cdj denote the column of bf (j, k), 1 ≤ k ≤ dj. Let T : H∞ → H be a
linear operator. Then the following conditions are equivalent:
(A) For each j ∈ N0, we have T (Hj) ⊂ Hj.
(B) For each ℓ ∈ N0 there exists a matrix σ(ℓ) ∈ Cdℓ×dℓ such that for all ek
j
(C) If in addition, ek
j are in the domain of T ∗ for all j and k, then for each ℓ ∈ N0
there exists a matrix σ(ℓ) ∈ Cdℓ×dℓ such that
j (ℓ, m) = σ(ℓ)mkδjℓ.
dT ek
cT f (ℓ) = σ(ℓ)bf (ℓ)
for all f ∈ H∞.
The matrices σ(ℓ) in (B) and (C) coincide.
The equivalent properties (A) -- (C) follow from the condition
(D) For each j ∈ N0, we have T Pj = PjT on H∞.
If, in addition, T extends to a bounded operator T ∈ L (H) then (D) is equivalent to
(A) -- (C).
6
JULIO DELGADO AND MICHAEL RUZHANSKY
Under the assumptions of Theorem 2.1, we have the direct sum decomposition
H =
Hj, Hj = span{ek
j }dj
k=1,
(2.1)
∞Mj=0
and we have dj = dim Hj. The two applications that we will consider will be with
H = L2(M) for a compact manifold M with Hj being the eigenspaces of an elliptic
pseudo-differential operator E, or with H = L2(G) for a compact Lie group G with
Hj = span{ξkm}1≤k,m≤dξ
for a unitary irreducible representation ξ ∈ [ξj] ∈ bG. The difference is that in the
first case we will have that the eigenvalues of E corresponding to Hj's are all distinct,
while in the second case the eigenvalues of the Laplacian on G for which Hj's are the
eigenspaces, may coincide. In Remark 2.6 we give an example of this difference for
operators on the torus Tn.
In view of properties (A) and (C), respectively, an operator T satisfying any of the
equivalent properties (A) -- (C) in Theorem 2.1, will be called an invariant operator, or
a Fourier multiplier relative to the decomposition {Hj}j∈N0 in (2.1). If the collection
{Hj}j∈N0 is fixed once and for all, we can just say that T is invariant or a Fourier
multiplier.
The family of matrices σ will be called the matrix symbol of T relative to the
partition {Hj} and to the basis {ek
j }. It is an element of the space Σ defined by
Σ = {σ : N0 ∋ ℓ 7→ σ(ℓ) ∈ Cdℓ×dℓ}.
(2.2)
A criterion for the extendability of T to L (H) in terms of its symbol will be given
in Theorem 2.3.
For f ∈ H, in the notation of Theorem 2.1, by definition we have
f =
∞Xj=0
djXk=1 bf (j, k)ek
j
with the convergence of the series in H. Since {ek
system on H, for all f ∈ H we have the Plancherel formula
j≥0
j }1≤k≤dj
is a complete orthonormal
(2.3)
(2.4)
(2.5)
∞Xj=0
djXk=1
∞Xj=0
djXk=1
kf k2
H =
(f, ek
j )2 =
bf (j, k)2 = kbf k2
where we interpret bf ∈ Σ as an element of the space
∞Xj=0
ℓ2(N0,Σ) = {h : N0 →Yd
Cd : h(j) ∈ Cdj and
djXk=1
ℓ2(N0,Σ),
h(j, k)2 < ∞},
and where we have written h(j, k) = h(j)k. In other words, ℓ2(N0,Σ) is the space of
all h ∈ Σ such that
∞Xj=0
djXk=1
h(j, k)2 < ∞.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
7
We endow ℓ2(N0, Σ) with the norm
khkℓ2(N0,Σ) :=
∞Xj=0
djXk=1
h(j, k)2
1
2
.
(2.6)
We note that the matrix symbol σ(ℓ) depends not only on the partition (2.1) but
also on the choice of the orthonormal basis. Whenever necessary, we will indicate the
dependance of σ on the orthonormal basis by writing (σ, {ek
) and we also will
refer to (σ, {ek
) as the symbol of T . Throughout this section the orthonormal
basis will be fixed and unless there is some risk of confusion the symbols will be
denoted simply by σ. In the invariant language, as will be clear from the proof of
Theorem 2.1, we have that the transpose of the symbol, σ(j)⊤ = T Hj is just the
restriction of T to Hj, which is well defined in view of the property (A).
j }1≤k≤dj
j }1≤k≤dj
j≥0
j≥0
We will also sometimes write Tσ to indicate that Tσ is an operator corresponding
to the symbol σ. It is clear from the definition that invariant operators are uniquely
determined by their symbols. Indeed, if T = 0 we obtain σ = 0 for any choice of an
orthonormal basis. Moreover, we note that by taking j = ℓ in (B) of Theorem 2.1 we
obtain the formula for the symbol:
(2.7)
σ(j)mk =dT ek
j (j, m),
for all 1 ≤ k, m ≤ dj. The formula (2.7) furnishes an explicit formula for the symbol in
terms of the operator and the orthonormal basis. The definition of Fourier coefficients
tells us that for invariant operators we have
σ(j)mk = (T ek
(2.8)
j )L2(M ).
j , em
In particular, for the identity operator T = I we have σI(j) = Idj , where Idj ∈ Cdj ×dj
is the identity matrix.
Before proving Theorem 2.1, let us establish a formula relating symbols with respect
If {eα} and {fα} are orthonormal bases of H, we
to different orthonormal basis.
consider the unitary operator U determined by U(eα) = fα. Then we have
(T eα, eβ)H = (UT eα, Ueβ)H = (UT U ∗Ueα, Ueβ)H = (UT U ∗fα, fβ)H.
If (σT , {eα}) denotes the symbol of T with respect to the orthonormal basis {eα} and
(σU T U ∗, {fα}) denotes the symbol of UT U ∗ with respect to the orthonormal basis
{fα} we have obtained the relation
(σT , {eα}) = (σU T U ∗, {fα}).
(2.9)
Thus, the equivalence relation of basis {eα} ∼ {fα} given by a unitary operator U
induces the equivalence relation on the set Σ of symbols given by (2.9).
In view
of this, we can also think of the symbol as an element of the space Σ/ ∼ with the
equivalence relation given by (2.9).
We make another remark concerning part (C) of Theorem 2.1. We use the condition
that ek
j are in the domain Dom(T ∗) of T ∗ in showing the implication (B) =⇒ (C). Since
j 's give a basis in H, and are all contained in Dom(T ∗), it follows that Dom(T ∗)
ek
is dense in H.
In particular, by [RS80, Theorem VIII.1], T must be closable (in
8
JULIO DELGADO AND MICHAEL RUZHANSKY
part (C)). These conditions are not restrictive for the further analysis since they are
satisfied in the natural applications of this paper.
The principal application of the notions above will be as follows, except for in
the sequel we will need more general operators E unbounded on H.
In order to
distinguish from this general case, in the following theorem we use the notation Eo.
Theorem 2.2. Continuing with the notation of Theorem 2.1, let Eo ∈ L (H) be a
linear continuous operator such that Hj are its eigenspaces:
Eoek
j = λjek
j
for each j ∈ N0 and all 1 ≤ k ≤ dj. Then equivalent conditions (A) -- (C) imply the
property
(E) For each j ∈ N0 and 1 ≤ k ≤ j, we have T Eoek
j = EoT ek
j ,
and if λj 6= λℓ for j 6= ℓ, then (E) is equivalent to properties (A) -- (C).
Moreover, if T extends to a bounded operator T ∈ L (H) then equivalent properties
(A) -- (D) imply the condition
(F) T Eo = EoT on H,
and if also λj 6= λℓ for j 6= ℓ, then (F) is equivalent to (A) -- (E).
For an operator T = F (Eo), when it is well-defined by the spectral calculus, we
have
(2.10)
In fact, this is also well-defined then for a function F defined on λj, with finite values
which are e.g. j-uniformly bounded (also for non self-adjoint Eo). We first prove
Theorem 2.1.
σF (Eo)(j) = F (λj)Idj .
Proof of Theorem 2.1. (A) =⇒ (B). If T satisfies condition (A), we consider the ma-
trix of T Hj : Hj → Hj with respect to the orthonormal basis {ei
j : 1 ≤ i ≤ dj} of Hj
and denote it by β(j). Then
Consequenlty, we have
T ek
j =
djXi=1
β(j)kiei
j.
ℓ ) = β(j)kmδjℓ = β(ℓ)kmδjℓ.
We take then σ(ℓ) := β(ℓ)⊤; it belongs to Cdℓ×dℓ and satisfies (B).
j (ℓ, m) =(T ek
j , em
(B) =⇒ (A). Since ek
j ∈ H∞, writing the series (2.3) for T ek
j ∈ H, we have
dT ek
dℓXm=1dT ek
T ek
j =Xℓ
j (ℓ, m)em
σ(ℓ)mkδjℓem
ℓ =
ℓ =Xℓ
dℓXm=1
σ(j)mkem
j ∈ Hj.
(2.11)
dℓXm=1
Since {em
j
: 1 ≤ m ≤ dj} spans Hj, we obtain (A).
(B) =⇒ (C). We assume in addition that ek
j are in the domain of T ∗ for all j and
k. We also assume that for each ℓ ∈ N0 there exists a matrix σ(ℓ) ∈ Cdℓ×dℓ such that
j (ℓ, m) = σ(ℓ)mkδjℓ.
dT ek
(2.12)
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
9
Now, if f ∈ H∞, then T f ∈ H, and by the inversion formula (2.3) we have
f =
∞Xj=0
djXk=1 bf (j, k)ek
j .
Now, using this and the fact that all em
ℓ are in the domain of T ∗, we have
j , em
j , T ∗em
ℓ )
=(f, T ∗em
ℓ )
ℓ
ℓ (cid:1)
cT f (ℓ, m) =(T f, em
=
djXk=1 bf (j, k)ek
∞Xj=0
djXk=1 bf (j, k)(cid:0)T ek
∞Xj=0
djXk=1 bf (j, k)dT ek
∞Xj=0
djXk=1 bf (j, k)σ(ℓ)mkδjℓ
∞Xj=0
djXk=1
σ(ℓ)mkbf (ℓ, k),
where we also used (2.12). Hence cT f (ℓ) = σ(ℓ)bf (ℓ), yielding (C).
(C) =⇒ (B). If cT f (ℓ) = σ(ℓ)bf (ℓ), then
j (ℓ, m) = (cid:16)σ(ℓ)bek
dT ek
σ(ℓ)mibek
j (ℓ)(cid:17)m
which gives (B), even without any assumptions on T ∗.
djXi=1
djXi=1
=
=
=
=
j (ℓ, m)
=
j (ℓ, i) =
σ(ℓ)miδjℓδki = σ(ℓ)mkδjℓ,
(D) =⇒ (A). We take f ∈ Hj. Then Pjf ∈ Hj since Pjf = f , so that by assumption
(D) we have
implying (A).
T f = T Pjf = PjT f ∈ Hj,
(A) =⇒ (D). For this part we assume in addition that T extends to a bounded
operator T ∈ L (H). First, we show that this together with (A) implies that T (H ⊥
j )
is orthogonal to Hj. For g ∈ H ⊥
j , we can write
g =Xℓ6=j
dℓXk=1
(g, ek
ℓ )ek
ℓ
10
JULIO DELGADO AND MICHAEL RUZHANSKY
with the convergence in H, so that
T g =Xℓ6=j
dℓXk=1
(g, ek
ℓ )T ek
ℓ
with the convergence in H due to the boundedness of T on H. Since by (A) we have
T ek
ℓ ∈ Hℓ ⊂ H ⊥
j
Let now f ∈ H∞. Writing f = f1 + f2 with f1 := Pjf so that f1 ∈ Hj and f2 ∈ H ⊥
j
for ℓ 6= j we conclude that T g is orthogonal to Hj.
are both in H∞, we have
PjT f = PjT f1 + PjT f2 = T f1 = T Pjf,
since the proved claim Pjf2 = 0 implies that PjT f2 = 0.
(cid:3)
We now continue with the proof of Theorem 2.2 when the basis ek
j corresponds to
the eigenvectors of an operator Eo ∈ L(H).
Proof of Theorem 2.2. (A) =⇒ (E). Let us fix some ek
write
j . By condition (A) we can
for some constants αi. Then
T ek
j =
αiei
j
djXi=1
djXi=1
EoT ek
j = Eo
αiλjei
j = λj
αiei
j = λjT ek
j = T λjek
j = T Eoek
j ,
djXi=1
αiei
j =
djXi=1
which shows (E).
(E) =⇒ (A). We note that it is enough to prove that T ek
j ∈ Hj since {ek
dj} forms a basis of the finite dimensional space Hj. We can assume that T ek
since otherwise there is nothing to prove. We recall that Eoek
(E), we have
j = λjek
j : 1 ≤ k ≤
j 6= 0
j . Using property
λjT ek
j = T Eoek
j = EoT ek
j .
Hence T ek
j ∈ H is a non-zero eigenvector of Eo corresponding to the eigenvalue λj.
Consequently, since Hj are maximal eigenspaces corresponding to λj, we must have
T ek
j ∈ Hj.
(E) =⇒ (F). Since we have already shown that (A) -- (C) always imply (E), it is
enough to prove that (E) implies (F) under the additional assumption that T ∈
L (H).
Let us write S := Eo ◦ T, D := T ◦ Eo and let f ∈ H. Under the assumptions both
S and D are bounded on H, and hence the formula (2.3) implies
Sf = lim
N
(f, ek
j )Sek
j = lim
N
NXj=0
djXk=1
with the convergent series in H.
NXj=0
djXk=1
(f, ek
j )Dek
j = Df,
(F) =⇒ (A). We note that we require T ∈ L (H) in order for T Eo and EoT to make
sense on H. It is clear that (F) implies (E), and under the additional assumption
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
11
that λj 6= λℓ for j 6= ℓ we already know that (A) -- (C) and (E) are equivalent. If T is
bounded on H, then they are also equivalent to (D).
(cid:3)
We have the following criterion for the extendability of a densely defined invariant
operator T : H∞ → H to L (H), which was an additional hypothesis for properties
(D) and (F). In the statements below we fix a partition into Hj's as in (2.1) and the
invariance refers to it.
Theorem 2.3. An invariant linear operator T : H∞ → H extends to a bounded
operator from H to H if and only if its symbol σ satisfies sup
kσ(ℓ)kL (Hℓ) < ∞.
ℓ∈N0
Moreover, denoting this extension also by T , we have
kT kL (H) = sup
ℓ∈N0
kσ(ℓ)kL (Hℓ).
Proof. We will often abbreviate writing kσ(ℓ)kop := kσ(ℓ)kL (Hℓ). Let us first suppose
that kσ(ℓ)kop ≤ C for all ℓ ∈ N0. By the Plancherel formula (2.4) we have
kT f k2
ℓ2(N0,Σ)
H =kcT f k2
=Xℓ
=Xℓ
≤Xℓ
=(cid:18)sup
≤ sup
ℓ
ℓ
ℓ2(Cdℓ )
ℓ2(Cdℓ )
kσ(ℓ)k2
kcT f (ℓ)k2
kσ(ℓ)bf (ℓ)k2
opkbf (ℓ)k2
opXℓ
kbf (ℓ)k2
kσ(ℓ)kop(cid:19)2
kσ(ℓ)k2
kf k2
H.
ℓ2(Cdℓ )
ℓ2(Cdℓ )
Conversely, let us suppose that T is bounded on H. Then there exists a constant
C > 0 such that kT f kH ≤ C for all f such that kf kH = 1. We can take C := kT kL (H).
Hence
T Hj : Hj → Hj
is bounded and kT Hj kL (Hj ) ≤ C. On the other hand, let β(j) denote the matrix
of T Hj : Hj → Hj with respect to the orthonormal basis {ei
j : 1 ≤ i ≤ dj} of Hj
as in the proof of Part (A) implies (B) in Theorem 2.1. We consider an unitary
operator U : Hj → Cdj which defines coordinates in Cdj of vectors in Hj with respect
to the orthonormal basis {ek
j : 1 ≤ k ≤ dj} of Hj. We also consider the operator
A(j) : Cdj → Cdj induced by the matrix β(j). Then
T Hj = U ∗A(j)U,
and
kσ(j)kop = kβ(j)kop = kA(j)kop = kT Hj kL (Hj ) ≤ C,
completing the proof.
(cid:3)
We also record the formula for the symbol of the composition of two invariant
operators:
12
JULIO DELGADO AND MICHAEL RUZHANSKY
Proposition 2.4. If S, T : H∞ → H are invariant operators with respect to the
same orthonormal partition, and such that the domain of S ◦ T contains H∞, then
S ◦ T : H∞ → H is also invariant with respect to the same partition. Moreover, if
σS denotes the symbol of S and σT denotes the symbols of T with respect to the same
orthonormal basis then
i.e. σS◦T (j) = σS(j)σT (j) for all j ∈ N0.
σS◦T = σSσT ,
Proof. Recalling the definition of the composition of densely defined operators, the
domain of S ◦ T is the space of functions f in the domain of T such that T f is in the
domain of S, in which case we set (S ◦ T )f = S(T f ). The assumption says that we
are in the position to use Theorem 2.1. Applying the condition (C) of Theorem 2.1
repeatedly, we have
so that S ◦ T is invariant by Part (C) of Theorem 2.1.
\(S ◦ T )f (j) = \S(T f )(j) = σS(j)cT f (j) = σS(j)σT (j)bf (j),
(cid:3)
We now show another application of the above notions to give a characterisation
of Schatten classes of invariant operators in terms of their symbols.
Theorem 2.5. Let 0 < r < ∞. An invariant operator T ∈ L (H) with symbol σ is
in the Schatten class Sr(H) if and only if
In particular, if T is in the trace class S1(H), then we have the trace formula
kσ(ℓ)kr
.
(2.13)
Sr(Hℓ)!1/r
Tr(σ(ℓ)).
(2.14)
Moreover
kσ(ℓ)kr
Sr(Hℓ) < ∞.
∞Xℓ=0
kT kSr(H) = ∞Xℓ=0
∞Xℓ=0
Tr(T ) =
Proof. First, we claim that Schatten classes of invariant operators can be charac-
terised in terms of the projections to the eigenspaces Hℓ:
Let us prove (2.15). Since
kT kr
Sr(H) =
∞Xℓ=0
kT Hℓkr
Sr(Hℓ).
kT kSr = kT kSr
(2.15)
we can assume without loss of generality that T is positive definite. We first observe
that λ is an eigenvalue (singular value) of T if and only if λ is an eigenvalue (singular
value) of T Hℓ(λ) for some ℓ(λ). Indeed, if λ is an eigenvalue of T there exists ϕλ ∈
H\{0} such that T ϕλ = λϕλ. Using Part (D) of Theorem 2.1, we get that
T Pℓϕλ = λPℓϕλ
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
13
holds for every ℓ. Since ϕλ 6= 0, there exists ℓ(λ) such that Pℓ(λ)ϕλ 6= 0. Conse-
quently, λ is the eigenvalue of T Hℓ(λ) = T Pℓ(λ). Conversely, since T (Hℓ(λ)) ⊂ Hℓ(λ),
an eigenvalue of T Hℓ(λ) is also an eigenvalue of T . Therefore, we obtain (2.15).
Now, given (2.15), to prove (2.13), it is enough to check that
kσ(ℓ)kSr(Hℓ) = kT HℓkSr(Hℓ).
(2.16)
To prove (2.16) we consider an unitary operator U : Hℓ → Cdℓ which defines coordi-
nates in Cdℓ of functions in Hℓ with respect to the orthonormal basis {ek
ℓ : 1 ≤ k ≤ dℓ}
of Hℓ. We also consider the operator A(ℓ) : Cdℓ → Cdℓ
induced by the matrix
(σT (ℓ))⊤. Then
and basic properties of Schatten quasinorms imply that
T Hℓ = U ∗A(ℓ)U,
kT HℓkSr(Hℓ) = kA(ℓ)kSr(Cdℓ ) = kσ(ℓ)kSr,
completing the proof of (2.16) and of (2.13).
Finally, let us prove (2.14) for operators in the trace class S1(H). Since the trace
Tr(T ) does not depend on the choice of the orthonormal basis in H, using property
(C) and formula (2.11), we can write
Tr(T ) =Xℓ
dℓXk=1
(T ek
ℓ , ek
ℓ , ek
ℓ )
σ(ℓ)mk(em
dℓXm=1
dℓXk=1
ℓ ) =Xℓ
dℓXk=1
dℓXm=1
=Xℓ
σ(ℓ)mkδmk =Xℓ
completing the proof.
dℓXk=1
σ(ℓ)kk =Xℓ
Tr(σ(ℓ)),
(cid:3)
Remark 2.6. We note that the membership in L (H) and in the Schatten classes
Sr(H) does not depend on the decomposition of H into subspaces Hj as in (2.1).
However, the notion of invariance does depend on it. For example, let H = L2(Tn)
for the n-torus Tn = Rn/Zn. Choosing
Hj = span{e2πij·x},
j ∈ Zn,
we recover the construction of Section 6 on compact Lie groups and moreover, invari-
ant operators with respect to {Hj}j∈Zn are the translation invariant operators on the
torus Tn. However, to recover the construction of Section 4 on manifolds, we takefHℓ
to be the eigenspaces of the Laplacian E on Tn, so that
Hj = span{e2πij·x : j ∈ Zn and j2 = ℓ},
ℓ ∈ N0.
Then translation invariant operators on Tn, i.e. operators invariant relative to the
the Laplacian, in terminology of Section 4). If we have information on the eigenvalues
of E, like we do on the torus, we may sometimes also recover invariant operators rela-
partition {Hj}j∈Zn, are also invariant relative to the partition {fHℓ}ℓ∈N0 (or relative to
tive to the partition {fHℓ}ℓ∈N0 as linear combinations of translation invariant operators
composed with phase shifts and complex conjugation.
fHℓ = Mj2=ℓ
14
JULIO DELGADO AND MICHAEL RUZHANSKY
3. Fourier analysis associated to an elliptic operator
Our main application will be to study operators on compact manifolds, so we
start this section by describing the discrete Fourier series associated to an elliptic
positive pseudo-differential operator as an adaptation of the construction in Section
2. In order to fix the notation for the rest of the paper we may give some explicit
expressions for notions of Section 2 in the present setting.
Let M be a compact smooth manifold of dimension n without boundary, endowed
with a fixed volume dx. We denote by Ψν(M) the Hormander class of pseudo-
differential operators of order ν ∈ R, i.e. operators which, in every coordinate chart,
are operators in Hormander classes on Rn with symbols in Sν
[Shu01] or
[RT10]. In this paper we will be using the class Ψν
cl(M) of classical operators, i.e.
operators with symbols having (in all local coordinates) an asymptotic expansion of
the symbol in positively homogeneous components (see e.g.
[Dui11]). Furthermore,
we denote by Ψν
e (M)
the class of elliptic operators in Ψν
+(M) the class of positive definite operators in Ψν
cl(M), and by Ψν
1,0, see e.g.
cl(M). Finally,
Ψν
e (M)
+e(M) := Ψν
+(M) ∩ Ψν
will denote the class of classical positive elliptic pseudo-differential operators of order
ν. We note that complex powers of such operators are well-defined, see e.g. Seeley
[See67].
In fact, all pseudo-differential operators considered in this paper will be
classical, so we may omit explicitly mentioning it every time, but we note that we
could equally work with general operators in Ψν(M) since their powers have similar
properties, see e.g. [Str72].
We now associate a discrete Fourier analysis to the operator E ∈ Ψν
+e(M) inspired
by those constructions considered by Seeley ([See65], [See69]), see also Greenfield and
Wallach [GW73]. However, we adapt it to our purposes and in the sequel also prove
several auxiliary statements concerning the eigenvalues of E and their multiplicities,
useful to us in the subsequent analysis. In general, the construction below is exactly
the one appearing in Theorem 2.1.
The eigenvalues of E (counted without multiplicities) form a sequence {λj} which
we order so that
0 = λ0 < λ1 < λ2 < · · · .
(3.1)
For each eigenvalue λj, there is the corresponding finite dimensional eigenspace Hj
of functions on M, which are smooth due to the ellipticity of E. We set
dj := dim Hj, and H0 := ker E, λ0 := 0.
We also set d0 := dim H0. Since the operator E is elliptic, it is Fredholm, hence also
d0 < ∞ (we can refer to [Ati68], [Hor85a] for various properties of H0 and d0).
We fix an orthonormal basis of L2(M) consisting of eigenfunctions of E:
{ek
j }1≤k≤dj
j≥0
,
(3.2)
where {ek
: L2(M) → Hj be the
corresponding projection. We shall denote by (·, ·) the inner product of L2(M). We
is an orthonormal basis of Hj. Let Pj
j }1≤k≤dj
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
15
observe that we have
for f ∈ L2(M). The 'Fourier' series takes the form
Pjf =
(f, ek
j )ek
j ,
djXk=1
djXk=1
∞Xj=0
f =
(f, ek
j )ek
j ,
for each f ∈ L2(M). The Fourier coefficients of f ∈ L2(M) with respect to the
orthonormal basis {ek
j } will be denoted by
(F f )(j, k) := bf (j, k) := (f, ek
j ).
simply the Fourier coefficients of f .
We will call the collection of bf (j, k) the Fourier coefficients of f relative to E, or
forms a complete orthonormal system in L2(M), for all f ∈ L2(M)
Since {ek
j }1≤k≤dj
j≥0
we have the Plancherel formula (2.4), namely,
(3.3)
kf k2
L2(M ) =
(f, ek
j )2 =
ℓ2(N0,Σ),
(3.4)
∞Xj=0
djXk=1
∞Xj=0
djXk=1
bf (j, k)2 = kbf k2
where the space ℓ2(N0, Σ) and its norm are as in (2.5) and (2.6).
We can think of F = FM as of the Fourier transform being an isometry from
L2(M) into ℓ2(N0, Σ). The inverse of this Fourier transform can be then expressed
by
(F −1h)(x) =
j (x).
(3.5)
If f ∈ L2(M), we also write
...
h(j, k)ek
djXk=1
∞Xj=0
bf (j) = bf (j, 1)
∈ Cdj ,
bf (j, dj)
j (ℓ, m)(cid:17)dℓ
j (ℓ) =(cid:16)bek
bek
bek
j (ℓ, m) = δjℓδkm.
m=1
thus thinking of the Fourier transform always as a column vector. In particular, we
think of
as of a column, and we notice that
Smooth functions on M can be characterised by
f ∈ C ∞(M) ⇐⇒ ∀N ∃CN :
⇐⇒ ∀N ∃CN :
bf (j, k) ≤ CN (1 + λj)−N for all j, k
bf (j) ≤ CN (1 + λj)−N for all j,
(3.6)
(3.7)
16
JULIO DELGADO AND MICHAEL RUZHANSKY
where bf (j) is the norm of the vector bf (j) ∈ Cdj . The implication '⇐=' here is
immediate, while '=⇒' follows from the Plancherel formula (2.4) and the fact that
for f ∈ C ∞(M) we have (I + E)N f ∈ L2(M) for any N.
For u ∈ D′(M), we denote its Fourier coefficient
j ),
bu(j, k) := u(ek
and by duality, the space of distributions can be characterised by
f ∈ D′(M) ⇐⇒ ∃M ∃C : bu(j, k) ≤ C(1 + λj)M for all j, k.
We will denote by H s(M) the usual Sobolev space over L2 on M. This space can be
+e(M) is positive and elliptic
defined in local coordinates or, by the fact that E ∈ Ψν
with ν > 0, it can be characterised by
f ∈ H s(M) ⇐⇒ (I + E)s/νf ∈ L2(M) ⇐⇒ {(1 + λj)s/νbf (j)}j ∈ ℓ2(N0, Σ)
(1 + λj)2s/νbf (j, k)2 < ∞.
djXk=1
∞Xj=0
⇐⇒
the last equivalence following from the Plancherel formula (2.4). For the characteri-
sation of analytic functions (on compact manifolds M) we refer to Seeley [See69].
(3.8)
4. Invariant operators and symbols on compact manifolds
We now discuss an application of a notion of an invariant operator and of its symbol
from Theorem 2.1 in the case of H = L2(M) and H∞ = C ∞(M) and describe its
basic properties. We will consider operators T densely defined on L2(M), and we will
be making a natural assumption that their domain contains C ∞(M). We also note
that while in Theorem 2.2 it was assumed that the operator Eo is bounded on H, this
is no longer the case for the operator E here. Indeed, an elliptic pseudo-differential
operator E ∈ Ψν
+e(M) of order ν > 0 is not bounded on L2(M).
Moreover, we do not want to assume that T extends to a bounded operator on
L2(M) to obtain analogues of properties (D) and (F) in Section 2, because this is too
restrictive from the point of view of differential operators. Instead, we show that in
the present setting it is enough to assume that T extends to a continuous operator
on D′(M) to reach the same conclusions.
So, we combine the statement of Theorem 2.1 and the necessary modification of
Theorem 2.2 to the setting of Section 3 as follows.
We also remark that Part (iv) of the following theorem provides a correct formu-
lation for a missing assumption in [DR14a, Theorem 3.1, (iv)].
Theorem 4.1. Let M be a closed manifold and let T : C ∞(M) → L2(M) be a linear
operator. Then the following conditions are equivalent:
(i) For each j ∈ N0, we have T (Hj) ⊂ Hj.
(ii) For each j ∈ N0 and 1 ≤ k ≤ j, we have T Eek
(iii) For each ℓ ∈ N0 there exists a matrix σ(ℓ) ∈ Cdℓ×dℓ such that for all ek
j
j = ET ek
j .
j (ℓ, m) = σ(ℓ)mkδjℓ.
dT ek
(4.1)
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
17
(iv) If, in addition, the domain of T ∗ contains C ∞(M), then for each ℓ ∈ N0 there
exists a matrix σ(ℓ) ∈ Cdℓ×dℓ such that
for all f ∈ C ∞(M).
The matrices σ(ℓ) in (iii) and (iv) coincide.
cT f (ℓ) = σ(ℓ)bf (ℓ)
If T extends to a linear continuous operator T : D′(M) → D′(M) then the above
properties are also equivalent to the following ones:
(v) For each j ∈ N0, we have T Pj = PjT on C ∞(M).
(vi) T E = ET on L2(M).
If any of the equivalent conditions (i) -- (iv) of Theorem 4.1 are satisfied, we say that
the operator T : C ∞(M) → L2(M) is invariant (or is a Fourier multiplier) relative
to E. We can also say that T is E-invariant or is an E-multiplier. This recovers the
notion of invariant operators given by Theorem 2.1, with respect to the partitions
Hj's in (2.1) which are fixed being the eigenspaces of E. When there is no risk of
confusion we will just refer to such kind of operators as invariant operators or as
multipliers. It is clear from (i) that the operator E itself or functions of E defined
by the functional calculus are invariant relative to E.
We note that the boundedness of T on L2(M) needed for conditions (D) and (F) in
Theorem 2.1 and in Theorem 2.2 is now replaced by the condition that T is continuous
on D′(M) which explored the additional structure of L2(M) and allows application
to differential operators.
We call σ in (iii) and (iv) the matrix symbol of T or simply the symbol. It is an
element of the space Σ = ΣM defined by
ΣM := {σ : N0 ∋ ℓ 7→ σ(ℓ) ∈ Cdℓ×dℓ}.
(4.2)
j and not on the
Since the expression for the symbol depends only on the basis ek
operator E itself, this notion coincides with the symbol defined in Theorem 2.1.
Let us comment on several conditions in Theorem 4.1 in this setting. Assumptions
(v) and (vi) are stronger than those in (i) -- (iv). On one hand, clearly (vi) contains
(ii). On the other hand, as we will see in the proof, assumption (v) implies (i) without
the additional hypothesis that T is continuous on D′(M).
In analogy to the strong commutativity in (v), if T is continuous on D′(M), so that
all the assumptions (i) -- (vi) are equivalent, we may say that T is strongly invariant
relative to E in this case.
The expressions in (vi) make sense as both sides are defined (and even continuous)
on D′(M).
We also note that without additional assumptions, it is known from the general
theory of densily defined operators on Hilbert spaces that conditions (v) and (vi) are
generally not equivalent, see e.g. Reed and Simon [RS80, Section VIII.5]. If T is
a differential operator, the additional assumption of continuity on D′(M) for parts
(v) and (vi) is satisfied. In [GW73, Section 1, Definition 1] Greenfield and Wallach
called a differential operator D to be an E-invariant operator if ED = DE, which is
our condition (vi). However, Theorem 4.1 describes more general operators as well
as reformulates them in the form of Fourier multipliers that will be explored in the
sequel.
18
JULIO DELGADO AND MICHAEL RUZHANSKY
There will be several useful classes of symbols, in particular the moderate growth
class
S ′(Σ) := {σ ∈ Σ : ∃N, C such that kσ(ℓ)kop ≤ C(1 + λℓ)N ∀ℓ ∈ N0},
(4.3)
where
kσ(ℓ)kop = kσ(ℓ)kL (Hℓ)
denotes the matrix multiplication operator norm with respect to ℓ2(Cdℓ).
In the case when M is a compact Lie group and E is a Laplacian on G, left-
invariant operators on G, i.e. operators commuting with the left action of G, are also
invariant relative to E in the sense of Theorem 4.1; this will be shown in Proposition
6.1 after we investigate in Section 6 the relation between the symbol in Theorem
4.1 and matrix symbols of operators on compact Lie groups. However, we need an
adaptation of the above construction since the natural decomposition into Hj's in
(2.1) may in general violate the condition (3.1).
As in Section 2 since the notion of the symbol depends only on the basis, for the
identity operator T = I we have
σI (j) = Idj ,
where Idj ∈ CIdj ×Idj is the identity matrix, and for an operator T = F (E), when it
is well-defined by the spectral calculus, we have
σF (E)(j) = F (λj)Idj .
(4.4)
Proof of Theorem 4.1. Once the basis ek
follows from the equivalence of (A), (B) and (C) in Theorem 2.1.
j is fixed, the equivalence of (i), (ii) and (iv)
j is smooth, we have T ek
(ii) =⇒ (i). We first note that both ET and T E are well-defined on ek
j : for the
j ∈ L2(M) and hence in D′(M) where E is
former, since ek
well-defined as a pseudo-differential operator, while, for the latter, Eek
j ∈
Hj ⊂ C ∞(M) and hence it is in the domain of T . The rest of the proof is identical
to (E) =⇒ (A) in the proof of Theorem 2.2.
j = λjek
(i) =⇒ (ii). This is the same as (A) =⇒ (E) in the proof of Theorem 2.2.
(v) =⇒ (i). We take f ∈ Hj. Then Pjf = f ∈ C ∞(M) so that by assumption (v)
we have
implying (i).
T f = T Pjf = PjT f ∈ Hj,
(i) =⇒ (v). We now assume in addition that T is continuous on D′(M). First, we
j i = 0 in the
j ⊂ L2(M), we have hT g, ek
show that (i) implies that for any g ∈ H ⊥
sense of distributions. We can write
with the convergence in L2(M). Hence
g =Xℓ6=j
T g =Xℓ6=j
(g, ek
ℓ )ek
ℓ
(g, ek
ℓ )T ek
ℓ
dℓXk=1
dℓXk=1
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
19
with the convergence in D′(M). Since T ek
T g is orthogonal to Hj.
ℓ ∈ Hℓ ⊂ H ⊥
j
for ℓ 6= j we conclude that
Let now f ∈ C ∞(M). Writing f = f1 + f2 with f1 = Pjf so that f1 ∈ Hj and
f2 ∈ H ⊥
j are necessarily smooth, and Pjf2 = 0, we have
PjT f = PjT f1 + PjT f2 = T f1 = T Pjf,
since the above property implies that PjT f2 = 0.
(vi) =⇒ (ii). Trivial.
(ii) =⇒ (vi). For the following, we assume that T is continuous on D′(M). Let us
write S := E ◦ T, D := T ◦ E and let f ∈ L2(M). We can write
f =
∞Xj=0
djXk=1
(f, ek
j )ek
j
with the series convergent in L2(M). Since both S and D are continuous on D′(M),
we now have
Sf = lim
N
(f, ek
j )Sek
j = lim
N
(f, ek
j )Dek
j = Df.
NXj=0
djXk=1
NXj=0
djXk=1
The limit should be understood in D′(M). Indeed, if we write
fN =
NXj=0
djXk=1
(f, ek
j )ek
j ,
then fN → f in L2 and hence also in D′(M), which implies SfN → Sf and DfN →
Df in D′(M).
(cid:3)
We now discuss how invariant operators can be expressed in terms of their symbols.
Proposition 4.2. An invariant operator Tσ associated to the symbol σ can be written
in the following way:
Tσf (x) =
=
ℓ (x)
∞Xℓ=0
∞Xℓ=0
dℓXm=1
(σ(ℓ)bf (ℓ))mem
[σ(ℓ)bf (ℓ)]⊤eℓ(x),
(4.5)
where [σ(ℓ)bf (ℓ)] denotes the column-vector, and [σ(ℓ)bf (ℓ)]⊤eℓ(x) denotes the multi-
plication (the scalar product) of the column-vector [σ(ℓ)bf (ℓ)] with the column-vector
ℓ (x))⊤. In particular, we also have
ℓ (x), · · · , em
eℓ(x) = (e1
(Tσek
j )(x) =
σ(j)mkem
j (x).
(4.6)
djXm=1
If σ ∈ S ′(Σ) and f ∈ C ∞(M), the convergence in (4.5) is uniform.
20
JULIO DELGADO AND MICHAEL RUZHANSKY
Proof. Formula (4.5) follows from Part (iv) of Theorem 4.1, with uniform convergence
for f ∈ C ∞(M) in view of (4.3). Then, using (4.5) and (3.6) we can calculate
j (ℓ))mem
ℓ (x)
(σ(ℓ))mibek
j (ℓ, i)! em
ℓ (x)
(σ(ℓ))miδjℓδkiem
ℓ (x)
dℓXm=1
(σ(ℓ)bek
dℓXm=1 dℓXi=1
dℓXm=1
dℓXi=1
(Tσek
j )(x) =
=
=
=
∞Xℓ=0
∞Xℓ=0
∞Xℓ=0
djXm=1
(σ(j))mkem
j (x),
yielding (4.6).
(cid:3)
Theorem 2.3 characterising invariant operators bounded on L2(M) now becomes
Theorem 4.3. An invariant linear operator T : C ∞(M) → L2(M) extends to a
bounded operator from L2(M) to L2(M) if and only if its symbol σ satisfies
kσ(ℓ)kop < ∞,
sup
ℓ∈N0
where kσ(ℓ)kop = kσ(ℓ)kL (Hℓ) is the matrix multiplication operator norm with respect
to Hℓ ≃ ℓ2(Cdℓ). Moreover, we have
kT kL (L2(M )) = sup
ℓ∈N0
kσ(ℓ)kop.
This can be extended to Sobolev spaces. We will use the multiplication property
for Fourier multipliers which is a direct consequence of Proposition 2.4:
Proposition 4.4. If S, T : C ∞(M) → L2(M) are invariant operators with respect to
E such that the domain of S ◦ T contains C ∞(M), then S ◦ T : C ∞(M) → L2(M)
is also invariant with respect to E. Moreover, if σS denotes the symbol of S and σT
denotes the symbols of T with respect to the same orthonormal basis then
σS◦T = σSσT ,
i.e. σS◦T (j) = σS(j)σT (j) for all j ∈ N0.
Recalling Sobolev spaces H s(M) in (3.8) we have:
Corollary 4.5. Let an invariant linear operator T : C ∞(M) → C ∞(M) have symbol
σT for which there exists C > 0 and m ∈ R such that
kσT (ℓ)kop ≤ C(1 + λℓ)
m
ν
holds for all ℓ ∈ N0. Then T extends to a bounded operator from H s(M) to H s−m(M)
for every s ∈ R.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
21
Proof. We note that by (3.8) the condition that T : H s(M) → H s−m(M) is bounded
is equivalent to the condition that the operator
S := (I + E)
s−m
ν ◦ T ◦ (I + E)− s
ν
is bounded on L2(M). By Proposition 4.4 and the fact that the powers of E are
pseudo-differential operators with diagonal symbols, see (4.4), we have
σS(ℓ) = (1 + λℓ)− m
ν σT (ℓ).
But then kσS(ℓ)kop ≤ C for all ℓ in view of the assumption on σT , so that the
statement follows from Theorem 4.3.
(cid:3)
5. Schatten classes of operators on compact manifolds
In this section we give an application of the constructions in the previous section
to determine the membership of operators in Schatten classes and then apply it to a
particular family of operators on L2(M).
As a consequence of Theorem 2.5, we can now characterise invariant operators in
Schatten classes on compact manifolds. We note that this characterisation does not
assume any regularity of the kernel nor of the symbol. Once we observe that the
conditions for the membership in the Schatten classes depend only on the basis ek
j
and not on the operator E, we immediately obtain:
Theorem 5.1. Let 0 < r < ∞. An invariant operator T : L2(M) → L2(M) is in
Sr(L2(M)) if and only if
Sr < ∞. Moreover
kσT (ℓ)kr
∞Pℓ=0
kT kr
Sr(L2(M )) =
kσT (ℓ)kr
Sr.
∞Xℓ=0
∞Xℓ=0
Tr(T ) =
Tr(σT (ℓ)).
If an invariant operator T : L2(M) → L2(M) is in the trace class S1(L2(M)), then
Remark 5.2. In Section 6 we will establish a relation between the notion of symbol
introduced in Theorem 4.1 and the corresponding symbol in the setting of compact
Lie groups (cf. [RT10, RT13]). In particular the characterisation above extends the
one obtained in Theorem 3.7 of [DR13].
We now apply Theorem 5.1 to determining which powers of E belong to which
Schatten classes. But first we record a useful relation between the sequences λj and
dj of eigenvalues of E and their multiplicities.
Proposition 5.3. Let M be a closed manifold of dimension n, and let E ∈ Ψν
with ν > 0. Then there exists a constant C > 0 such that we have
+e(M),
for all j ≥ 1. Moreover, we also have
dj ≤ C(1 + λj)
n
ν
dj(1 + λj)−q < ∞ if and only if
q >
n
ν
.
∞Xj=1
(5.1)
(5.2)
as λ → ∞. This implies dj ≤ C(1 + λj)n/ν for sufficiently large λj, implying the
estimate (5.1).
To prove (5.2), let us denote T := (I + E)−q/2. Then the eigenvalues of T are
(1 + λj)−q/2 with multiplicities dj. This implies
dj(1 + λj)−q = kT k2
S2 ≍ kKk2
L2(M ×M ).
(5.3)
∞Xj=0
22
JULIO DELGADO AND MICHAEL RUZHANSKY
Proof. Since (1 + λj)1/ν are the eigenvalues of the first-order elliptic positive operator
(I +E)1/ν with multiplicities dj, the Weyl eigenvalue counting formula for the operator
(I + E)1/ν gives
Xj: (1+λj )1/ν ≤λ
dj = C0λn + O(λn−1)
By the functional calculus of pseudo-differential operators, we have T ∈ Ψ−νq/2(M),
and so its integral kernel K(x, y) is smooth for x 6= y, and near the diagonal x = y,
identifying points with their local coordinates, we have
K(x, y) ≤ Cαx − y−α,
for any α > n − νq/2, see e.g. [Dui11] or [RT10, Theorem 2.3.1]. Thus order is sharp
with respect to the order of the operator. Therefore, K ∈ L2(M × M) if and only if
there exists α such that n > 2α > 2n − νq. Together with (5.3) this implies (5.2). (cid:3)
Proposition 5.4. Let M be a closed manifold of dimension n, and let E ∈ Ψν
+e(M)
be a positive elliptic pseudo-differential operator of order ν > 0. Let 0 < p < ∞.
Then we have
(I + E)− α
ν ∈ Sp(L2(M)) if and only if α >
.
(5.4)
n
p
Proof. We note that the operator (I + E)− α
are (1 + λj)− α
ν with multiplicities dj. Therefore,
ν is positive definite, its singular values
k(I + E)− α
ν kp
Sp =
dj(1 + λj)− αp
ν ,
∞Xj=0
which is finite if and only if αp > n by (5.2), implying the statement.
(cid:3)
6. Relation to the setting of compact Lie groups
In the recent work [DR13] the authors studied Schatten classes of operators on
compact Lie groups. We now explore how the notion of the symbol from Theorem
4.1 corresponds to the matrix-valued symbols on compact Lie groups, and how the
results for Schatten classes correspond to each other when M = G is a compact Lie
group. In this and the following sections we assume that all operators are continuous
on D′(G) so that the integral kernels of such operators are distributions.
We will give two types of decompositions of L2(G) into Hj's as in (2.1). First,
we choose Hj's determined by unitary irreducible representations of G. However,
in this case the condition (3.1) may fail. Consequently, to view this analysis as a
special case of the construction on manifolds in Section 4 with condition (3.1), we
group representations corresponding to the same eigenvalue of the Laplacian together,
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
23
to form a coarser decomposition of L2(G) into a direct sum of finite dimensional
subspaces. The example of this types of partitions is given in Remark 2.6 in the case
of the torus Tn.
Now, we recall some basic definitions. Let G be a compact Lie group of dimension
classes of continuous irreducible unitary representations of G. Since G is compact,
n equipped with the normalised Haar measure. Let bG denote the set of equivalence
the set bG is discrete. For [ξ] ∈ bG, by choosing a basis in the representation space of ξ,
we can view ξ as a matrix-valued function ξ : G → Cdξ×dξ, where dξ is the dimension
of the representation space of ξ. By the Peter-Weyl theorem the collection
npdξ ξij : 1 ≤ i, j ≤ dξ, [ξ] ∈ bGo
FGf (ξ) ≡ bf (ξ) :=ZG
f (x)ξ(x)∗dx,
is the orthonormal basis of L2(G). If f ∈ L1(G) we define its group Fourier transform
at ξ by
where dx is the normalised Haar measure on G. If ξ is a matrix representation, we
have bf (ξ) ∈ Cdξ×dξ . We note that this Fourier transform is different from the one
we considered on manifolds in (3.3) which produced vector-valued Fourier coefficients
instead of the matrix-valued ones obtained in (6.1).
The Fourier inversion formula is a consequence of the Peter-Weyl theorem, so that
we have
(6.1)
(6.2)
f (x) = X[ξ]∈bG
dξ Tr(ξ(x)bf (ξ)).
For each [ξ] ∈ bG, the matrix elements of ξ are the eigenfunctions for the Lapla-
cian LG (or the Casimir element of the universal enveloping algebra), with the same
eigenvalues which we denote by −λ2
[ξ], so that we have
− LGξij(x) = λ2
[ξ]ξij(x)
for all 1 ≤ i, j ≤ dξ.
(6.3)
For a thorough discussion of Laplacians on compact Lie groups we refer to [Ste70].
The weight for measuring the decay or growth of Fourier coefficients in this setting
2 , the eigenvalues of the elliptic first-order pseudo-differential
1
is hξi := (1 + λ2
operator (I − LG)
[ξ])
2 . The Parseval identity takes the form
1
kf kL2(G) =X[ξ]∈bG
HS
dξkbf (ξ)k2
which defines the norm on ℓ2(bG).
valued symbol τA(x, ξ) ∈ Cdξ×dξ by
1
2
,
where kbf (ξ)k2
HS = Tr(bf (ξ)bf (ξ)∗),
For a linear continuous operator A from C ∞(G) to D′(G) we define its matrix-
τA(x, ξ) := ξ(x)∗(Aξ)(x) ∈ Cdξ×dξ.
Then one has ([RT10], [RT13]) the global quantization
Af (x) = X[ξ]∈bG
dξ Tr(ξ(x)τA(x, ξ)bf (ξ))
(6.4)
(6.5)
24
JULIO DELGADO AND MICHAEL RUZHANSKY
in the sense of distributions, and the sum is independent of the choice of a represen-
tation ξ from each equivalence class [ξ] ∈ bG. If A is a linear continuous operator from
C ∞(G) to C ∞(G), the series (6.5) is absolutely convergent and can be interpreted
in the pointwise sense. We will also write A = Op(τA) for the operator A given by
the formula (6.5). We refer to [RT10, RT13] for the consistent development of this
quantization and the corresponding symbolic calculus.
In the case of a left-invariant operator A, its symbol τA is independent of x, and
formula (6.4) reduces to
τA(ξ) = ξ(x)∗(Aξ)(x) = Aξ(e),
(6.6)
where e is the unit element of the group.
We can now establish a correspondence between the two frameworks, the one in this
paper and the one given in [DR13]. In the setting of compact Lie groups the unitary
dual being discrete, we can enumerate the representations as ξj for 0 ≤ j < ∞. The
indices (i, ℓ) of each matrix ξ(x) will be enumerated following the lexicographical order
((i, ℓ) ≤ (i′, ℓ′) if i < i′ or (i = i′ and ℓ ≤ ℓ′)). In this way, we fix the orthonormal
basis {ek
j } given by
where dj = d2
the lexicographical order described above. Then we have the subspaces
ξj and k represents an entry of the matrix of the representation following
Hj ≡ H[ξj] := span{(ξj)iℓ : 1 ≤ i, ℓ ≤ dξj }.
(6.8)
On a compact Lie group G we can consider E to be a bi-invariant Laplacian,
see Stein [Ste70] for a discussion of such operators. Then, in view of the Peter-
Weyl theorem, the functions {ek
j }1≤k≤dj are its eigenfunctions, with norm one in
L2(G) with respect to the normalised Haar measure, and corresponding to the same
eigenvalue λj. However, the condition (3.1) does not hold in general since non-
equivalent representations in bG may give the same eigenvalues of the Laplacian.
We now observe that there is also a correspondence between the vector-valued
Fourier transform introduced in (3.3) and the matrix-valued Fourier transform de-
fined in (6.1). Such correspondence can be established by applying once more the
lexicographical order to the matrix-valued Fourier transform (6.1).
In order to study such correspondence, for d ∈ N we will define a bijection from
the set of indices of the matrix-symbol {1, . . . , d}2 onto the set of indices {1, . . . , d2}
and calculate its inverse. If (j, k) ∈ {1, . . . , d}2 we define
Γd(j, k) := (j − 1)d + k.
The function Γd is surjective, indeed if t ∈ {1, . . . , d2}, j can be obtained from
{ek
j }1≤k≤dj =nqdξj (ξj)iℓo1≤i,ℓ≤dξj
,
(6.7)
where ⌊·⌋ denotes the function defined for x ≥ 0 by ⌊x⌋ = max{y ∈ N0 : y ≤ x}.
For the term k we observe that
j =(cid:22)t − 1
d (cid:23) + 1,
j − 1 =(cid:22)t − 1
d (cid:23) ,
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
25
hence
k = t −(cid:22)t − 1
d (cid:23) d.
Since we are dealing with finite sets with the same number of elements, the injectivity
of Γ follows.
We can now establish correspondences between the Fourier transforms on G = M,
for M viewed as both a compact manifold and a compact Lie group. Taking into
account (6.1) and (6.7) we obtain
(FM f )(i, t) = (f, et
i)L2 =pdξi((FGf )(ξi))
(t−(cid:22) t−1
ξi. In the another direction we have
for i ∈ N0, 1 ≤ t ≤ di = d2
dξi (cid:23)dξi ,(cid:22) t−1
dξi (cid:23)+1)
,
(6.9)
for 1 ≤ i, j ≤ dξℓ.
((FGf )(ξℓ))i,j =
1
pdξℓ
(FM f )(ℓ, Γdξℓ
(j, i)),
(6.10)
For the sake of simplicity, we introduce the following notation:
where t ∈ {1, . . . , d2}. With this notation formula (6.9) becomes
(6.11)
d (cid:23) + 1, φ(t, d) := t −(cid:22)t − 1
d (cid:23) d,
ψ(t, d) :=(cid:22)t − 1
(FM f )(ℓ, m) =pdξℓ((FGf )(ξℓ)(φ(m,dξℓ ),ψ(m,dξℓ )).
ek
j = (qdξj ξj)(ψ(k,dξj ),φ(k,dξj )).
(FG(ηrs)(η))ij =ZG
ηrs(x)ηji(x)dx =
1
dη
δ(i,j),(s,r),
We also have
In the calculations below we will use the following basic relations for the Fourier
transform on a compact Lie group G:
which means that FG(ηrs)(η) is the matrix of dimension dη × dη with the only entry
different from zero equal to 1
in the position (s, r). We will denote this matrix by
dη
1
(δ(i,j),(s,r))ij, and we have also δ(i,j),(s,r) = 1 if i = s and r = j, and δ(i,j),(s,r) = 0 if
dη
i 6= s or r 6= j.
Thus, for an invariant operator we obtain
(FG(T (ξrs)))(ξ) = τ (ξ)(FG(ξrs)(ξ)) = τ (ξ)
1
dξ
(δ(i,j),(s,r))ij.
(6.12)
In other words (FG(T (ξrs)))(ξ) is a matrix of dimension dξ ×dξ with all the columns
zero except for the r-column which is equal to the s-column of 1
dξ
τ (ξ).
We shall denote by σ the symbol corresponding to T and consider the orthonormal
basis {ek
j } defined in (6.7) in the sense of (4.1) on manifolds. The symbol introduced
in (6.4) in the sense of groups will be denoted by τ . We now can find formulae relating
26
JULIO DELGADO AND MICHAEL RUZHANSKY
the symbols τ and σ. We begin by finding a formula for σ in terms of τ . By (6.11),
(4.1) and (6.12) we obtain
σ(ℓ)mi =(FM (T ei
ℓ))(ℓ, m)
))
),ψ(m,dξℓ
ℓ))(ξℓ))(φ(m,dξℓ
=pdξℓ((FG(T ei
=pdξℓ((FG(T (pdξℓξℓ)ψ(i,dξℓ ),φ(i,dξℓ )))(ξℓ))(φ(m,dξℓ ),ψ(m,dξℓ ))
=dξℓ((FG(T (ξℓ)ψ(i,dξℓ ),φ(i,dξℓ )))(ξℓ))(φ(m,dξℓ ),ψ(m,dξℓ ))
=dξℓd−1
ξℓ
=τ (ξℓ)(φ(m,dξℓ ),φ(i,dξℓ ))δψ(i,dξℓ ),ψ(m,dξℓ ).
(τ (ξℓ)(δ((p,q),(φ(i,dξℓ ),ψ(i,dξℓ )))pq)(φ(m,dξℓ ),ψ(m,dξℓ ))
Therefore, we obtain
σ(ℓ)mi =(cid:26) τ (ξℓ)(φ(m,dξℓ ),φ(i,dξℓ )) ,
0
,
if ψ(m, dξℓ) = ψ(i, dξℓ),
otherwise.
(6.13)
We note that both functions φ and ψ are periodic with respect to the first parame-
ters i and m, implying that there is a periodic structure in the 'big' manifold-symbol
σ composed of some copies of the 'small' group-symbol τ .
We will now give a graphical description of the relations (6.13) between the two
symbols. The entries of τ (ξℓ) are distributed inside the matrix-symbol σ according
to (6.13): setting d := dξℓ it is
dξℓ
↓
τ (ξℓ)1d
τ (ξℓ)2d
...
τ (ξℓ)dd
dξℓ + 1
↓
0
0
...
0
i
0
0
...
0
0
0
...
0
...
...
0
0
...
0
τ (ξℓ)11
τ (ξℓ)21
...
τ (ξℓ)12
τ (ξℓ)22
...
τ (ξℓ)d1
...
...
τ (ξℓ)d2
...
...
0
0
...
0
0
0
...
0
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
0
0
...
0
τ (ξℓ)1d
τ (ξℓ)2d
...
τ (ξℓ)dd
...
...
0
0
...
0
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
↓
0
0
...
0
0
0
...
0
...
...
0
0
...
0
0
0
...
0
...
...
τ (ξℓ)11
τ (ξℓ)12
τ (ξℓ)21
...
τ (ξℓ)22
...
d2
ξℓ
↓
0
0
...
0
0
0
...
0
...
...
τ (ξℓ)1d
τ (ξℓ)2d
...
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
τ (ξℓ)d1
τ (ξℓ)d2
· · ·
τ (ξℓ)dd
τ (ξℓ)11
τ (ξℓ)21
...
τ (ξℓ)12
τ (ξℓ)22
...
τ (ξℓ)d1
τ (ξℓ)d2
0
0
...
0
...
...
0
0
...
0
0
0
...
0
...
...
0
0
...
0
On the other hand, given the symbol σ, an application of equations (6.13) for 1 ≤
m, i ≤ dξℓ gives
(6.14)
The proposition below shows that the Schatten quasi-norms k · kSr of the symbols
τ and σ are in agreement when M = G is a compact Lie group. Thus, our results
for 1 ≤ m, i ≤ dξℓ.
τ (ξℓ)mi = σ(ℓ)mi,
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
27
in Section 5 are an extension of those in [DR13] concerning Schatten classes.
In
particular, Theorem 5.1 extents Theorem 3.7 of [DR13] as announced in Remark 5.2.
We recall that on a compact Lie group G we take E to be a bi-invariant Laplacian.
Proposition 6.1. Let G be a compact Lie group. If a linear operator T : C ∞(G) →
L2(G), continuous on D′(G), is left-invariant then it is also invariant relative to the
family of Hj's as in (6.8) in the sense of Theorem 2.1 (in fact, it is also strongly
invariant).
Let T : C ∞(G) → L2(G) be a left-invariant operator, and let σ be its symbol in the
sense of Theorem 2.1 and τ its symbol in the sense of groups as in (6.6). Then these
symbols are related by formulae (6.13) -- (6.14).
Consequently, for a bounded left-invariant operator T : L2(G) → L2(G), for every
0 < r < ∞ we have
and, therefore,
kσ(ℓ)kr
Sr = dξℓkτ (ξℓ)kr
Sr ,
Xℓ
kσ(ℓ)kr
Sr =Xℓ
dξℓkτ (ξℓ)kr
Sr.
Proof. The invariance in the sense of groups as in (6.6) of the group-left-invariant ope-
rators follows from the relation (6.13) between symbols and from the characterisation
in Theorem 2.1.
For the following statements, since for Schatten quasi-norms we have
kBkSr = kBkSr,
we can assume that σ, τ are symmetric, and hence they can be also assumed diagonal.
On the other hand, using the relation between σ and τ in (6.13) and (6.14), and by
looking at the diagonal elements of σ in (6.13), we obtain
kσ(ℓ)kr
Sr =
d2
ξℓXm=1
σ(ℓ)mmr = dξℓ
dξℓXm=1
Sr and, therefore,Pℓ
τ (ξℓ)mmr = dξℓkτ (ξℓ)kr
Sr.
Sr =Pℓ
Thus kσ(ℓ)kr
Sr = dξℓkτ (ξℓ)kr
kσ(ℓ)kr
dξℓkτ (ξℓ)kr
Sr.
(cid:3)
We finish this section by describing an adaptation of the above construction to
put it in the framework of manifolds as described in Theorem 4.1. In the case of the
torus Tn this is indicated in Remark 2.6. Recalling the definition of H[ξ] in (6.8) for
each [ξ] ∈ bG, and the notation λ[ξ] for the eigenvalues as in (6.3), for the sequence
2 < . . . of eigenvalues of −LG counted without multiplicities we set
1 < λ2
0 < λ2
0 = λ2
span{ξik : 1 ≤ i, k ≤ dξ},
ℓ ∈ N0.
(6.15)
fHℓ := M[ξ]∈bG
λ[ξ]=λℓ
H[ξ] = M[ξ]∈bG
λ[ξ]=λℓ
The family offHℓ's is the collection of eigenspaces of the elliptic differential operator
LG for which the condition (3.1) is satisfied. The symbols σ and eσ of an invariant
28
JULIO DELGADO AND MICHAEL RUZHANSKY
operator T with respect to the partitions Hj's and fHℓ's, respectively, are related by
(6.16)
eσ(ℓ) = O[ξj]∈bG
λ[ξj ]=λℓ
σ(j),
witheσ(ℓ) ∈ Cedℓ×edℓ and
edℓ = X[ξj]∈bG
λ[ξj ]=λℓ
dj = X[ξj]∈bG
λ[ξj ]=λℓ
d2
ξj .
Recalling the relation (6.13) between the symbol σ in the sense of Theorem 2.1 and
the group symbol τ as in (6.6), given by
0
...
0
0
0
...
τ (ξj)
0
...
0
τ (ξj)
· · ·
· · ·
σ(j) ≡ σ(ξj) =
,
the formula (6.16) provides the further relation between the symbol eσ in the sense
λ[ξm] = λℓ for non-equivalent representations [ξ1], . . . , [ξm] ∈ bG, we have
of manifolds (in Theorem 4.1) and the group symbol τ . Therefore, if λ[ξ1] = . . . =
· · ·
· · · τ (ξj)
(6.17)
eσ(ℓ) =
· · ·
· · ·
σ(ξ1)
0
0
...
0
σ(ξ2)
...
0
· · ·
· · · σ(ξm)
0
0
...
.
(6.18)
In particular, we obtain
Corollary 6.2. Let G be a compact Lie group and let T : C ∞(G) → L2(G) be a
linear operator, continuous on D′(G). If T is left-invariant then it is also invariant
relative to the operator LG (in the sense of Theorem 4.1). The corresponding symbols
are related by formulae (6.16) -- (6.18).
7. Kernels of invariant operators on compact manifolds
In this section we describe invariant operators relative to E in terms of their kernels.
We first observe that if T = Tσ is invariant with symbol σ, expanding Proposition
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
29
4.2 we can write
Tσf (x) =
ℓ (x)
=
ℓ (x)
dℓXm=1
∞Xℓ=0
(σ(ℓ)bf (ℓ))mem
dℓXm=1
dℓXk=1
∞Xℓ=0
σ(ℓ)mkbf (ℓ)kem
ℓ (x)ZM
dℓXm=1
dℓXk=1
∞Xℓ=0
∞Xℓ=0
=ZM
dℓXm=1
dℓXk=1
σ(ℓ)mkem
σ(ℓ)mkem
=
f (y)ek
ℓ (y)dy
ℓ (x)ek
ℓ (y)! f (y)dy.
Hence, the integral kernel K(x, y) of Tσ is given by
K(x, y) =
∞Xℓ=0
dℓXm=1
dℓXk=1
On the other hand we note that
σ(ℓ)mkem
ℓ (x)ek
ℓ (y).
(7.1)
{em
ℓ ⊗ em′
ℓ′ }1≤m≤dℓ,1≤m′≤dℓ′
ℓ,ℓ′≥0
defines an orthonormal basis of L2(M × M).
If T is Hilbert-Schmidt on L2(M),
not necessarily invariant, then its kernel K is square integrable and we can write its
decomposition in this basis as
K(x, y) =
∞Xℓ=0
∞Xℓ′=0
dℓXm=1
((FM ⊗ FM )K)(ℓ, m, ℓ′, m′)em
ℓ (x)em′
ℓ′ (y),
dℓ′Xm′=1
(7.2)
where ((FM ⊗ FM )K)(ℓ, m, ℓ′, m′) denotes the Fourier coefficients of K with respect
to the basis {em
ℓ ⊗ em′
ℓ′ } given by
((FM ⊗ FM )K)(ℓ, m, ℓ′, m′) =(K, em
ℓ (x)em′
ℓ′ (y))L2(M ×M )
= ZM ×M
K(x, y)em
ℓ (x)em′
ℓ′ (y)dxdy.
We observe from (7.1) and (7.2) that T is invariant relative to (E, {em
only if
ℓ }1≤m≤dℓ
ℓ≥0
((FM ⊗ FM )K)(ℓ, m, ℓ′, m′) =(cid:26)
0,
σ(ℓ)mm′,
ℓ 6= ℓ′,
ℓ = ℓ′.
) if and
(7.3)
30
JULIO DELGADO AND MICHAEL RUZHANSKY
For example, from (7.1) we obtain
(K, em
ℓ (x)em′
σ(j)kiek
j (x)ei
ℓ (x)em′
ℓ′ (y)dxdy
ℓ′ (y))L2(M ×M ) = ZM ×M
∞Xj=0
djXi=1
djXk=1
∞Xj=0
=(cid:26)
0,
σ(ℓ)mm′,
=
djXk=1
djXi=1
σ(j)kiZM
ℓ 6= ℓ′,
ℓ = ℓ′.
j(y) em
ℓ (x)dxZM
ek
j (x)em
em′
ℓ′ (y)ei
j(y)dy
We now introduce some notation which will be useful in order to define a suitable
setting to study the above Fourier coefficients and the relation between operator's
kernel and symbol. Let
Σ(M × M) :=nσ = (σ(ℓ, m, ℓ′, m′))1≤m≤dℓ,1≤m′≤dℓ′
0≤ℓ,ℓ′<∞
: σ(ℓ, m, ℓ′, m′) = 0 if ℓ 6= ℓ′o ,
K := {K ∈ D′(M × M) : K defines an invariant operator relative to E}.
We now consider the mapping
K 7→ (FM ⊗ FM )K
from K into Σ(M × M). We can identify the family of symbols Σ(M × M) with the
matricesSℓ
Cdℓ×dℓ by letting
σ ≡ σ
such that σ(ℓ)mm′ = σ(ℓ, m, ℓ, m′). In this way we also get the identification
Σ(M × M) ≃ ΣM = Σ
with Σ from (4.2).
If 1 ≤ p < ∞ we define
ℓp(Σ) = {σ ∈ Σ :
On ℓp(Σ) we define the norm
kσ(ℓ)kp
Sp < ∞}.
∞Xℓ=0
kσkℓp(Σ) := ∞Xℓ=0
If p = ∞ we define
kσ(ℓ)kp
p
Sp! 1
, 1 ≤ p < ∞.
ℓ∞(Σ) = {σ ∈ Σ : sup
ℓ∈N0
kσ(ℓ)kop < ∞},
and we endow ℓ∞(Σ) with the norm
kσkℓ∞(Σ) := sup
ℓ∈N0
kσ(ℓ)kop.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
31
The integral operator with kernel K will be sometimes denoted by TK. We note that
in terms of the norms ℓp(Σ), for invariant operators Theorem 4.3 can be formulated
as
and Theorem 5.1 can be formulated as
T ∈ L (L2(M)) ⇐⇒ σT ∈ ℓ∞(Σ),
T ∈ Sp(L2(M)) ⇐⇒ σT ∈ ℓp(Σ)
for 0 < p < ∞.
(7.4)
(7.5)
For the formulation of the following theorem we will use the mixed-norm Lp spaces
x Lp2
y on the manifold M for 1 ≤ p1, p2 ≤ ∞. A measurable function K(x, y) is said
Lp1
to belong to Lp1
x Lp2
y (M × M) if
kkK(x, y)kL
y kL
p2
x < ∞.
p1
On Lp1
x Lp2
y (M × M) we consider the norm k · kL
p1
x L
p2
y
y kL
p2
x . We also define
p1
L(p1,p2)(M × M) := Lp1
x Lp2
y (M × M) ∩ Lp1
:= kk · kL
y Lp2
x (M × M),
endowed with norm
k · kL(p1,p2) := max{k · kL
p1
x L
y , k · kL
p2
p1
y L
x }.
p2
We note that in general L(p1,p2) 6= L(p2,p1). The basic properties of mixed-norm Lp
spaces for many variables were first studied by Benedek and Panzone in [BP61]. In
particular they proved a version of Stein's Interpolation of operators theorem and as
a consequence the Riesz-Thorin theorem in that setting. A slight modification allows
us to apply the Riesz-Thorin theorem when the operator T acts from a mixed-norm
Lp space to an ℓp(Σ)-space.
p′ = 1.
Theorem 7.1. If 1 ≤ p ≤ 2 and K ∈ K ∩ L(p′,p), then (FM ⊗ FM )K ∈ ℓp′(Σ), where
1
p + 1
Proof. If p = 2 we have p′ = 2. From
xL2
x = K ∩ L2
K ∈ K ∩ L2
x,y ⊂ L2
x,y
y ∩ L2
yL2
we get a Hilbert-Schmidt operator TK. On the other hand, by Theorem 5.1 with
S2 < ∞. Hence and by (7.3) we obtain
kσ(ℓ)k2
r = 2, if σ is the symbol of TK thenPℓ
(FM ⊗ FM )K ∈ ℓ2(Σ).
For p = 1 we have p′ = ∞. If
K ∈ K ∩ L∞
x L1
y ∩ L∞
y L1
x,
by Schur's Lemma we get TK ∈ L (Lr(M)) for all 1 ≤ r ≤ ∞. In particular TK ∈
L (L2(M)) and by Theorem 4.3 the symbol σ of TK satisfies
sup
kσ(ℓ)kop < ∞.
ℓ
By (7.3) we have
k(FM ⊗ FM )Kkℓ∞(Σ) = sup
ℓ
kσ(ℓ)kop.
Hence (FM ⊗ FM )K ∈ ℓ∞(Σ). We have shown that
(FM ⊗ FM ) : K ∩ L(2,2) −→ ℓ2(Σ)
32
and
JULIO DELGADO AND MICHAEL RUZHANSKY
(FM ⊗ FM ) : K ∩ L(∞,1) −→ ℓ∞(Σ).
By the Riesz-Thorin interpolation theorem between L(r,s) and ℓp(Σ) spaces (cf. [BP61,
Theorem 2]) we obtain
with 1
p 1
= 1−θ
2 + θ
∞ , 1
p 2
(FM ⊗ FM ) : K ∩ L(p1,p2) −→ ℓq(Σ),
= 1−θ
q = 1−θ
We observe that if p = 2
2 + θ
1 , 1
2
1 − θ
p1 =
, p2 =
2 + θ
∞ for 0 ≤ θ ≤ 1. Hence
2
2
, q =
.
1 + θ
p and 2
1 − θ
p−1 = p′. Thus
1+θ then θ = 2−p
1−θ = p
(FM ⊗ FM ) : K ∩ L(p′,p) −→ ℓp′
(Σ),
completing the proof.
(cid:3)
The following corollary is an immediate consequence of Theorems 7.1 and 5.1, it
furnishes a sufficient kernel condition for Schatten classes with index p′ ≥ 2.
Corollary 7.2. If 1 ≤ p ≤ 2 and K ∈ K ∩ L(p′,p)(M × M), then TK ∈ Sp′(L2(M)).
We recall that sufficient conditions of the type above in terms of kernels are not
allowed for 0 < p′ < 2 as a consequence of a Carleman's example. Corollary 7.2 is
known for general integral operators (cf. [Rus74, Theorem 3]). Here we have deduced
a particular version for invariant operators with a simple proof by applying the notion
of symbol.
We now describe another representation of the kernel as the 'generalised' Fourier
transform of the symbol. From formula (7.1) we have
K(x, y) =
=
=
=
∞Xℓ=0
∞Xℓ=0
∞Xℓ=0
∞Xℓ=0
where
dℓXm=1
dℓXk=1
σ(ℓ)mkem
ℓ (x)ek
ℓ (y)
Tr(eℓ(x)⊤σ(ℓ)eℓ(y))
Tr(σ(ℓ)eℓ(y)eℓ(x)⊤)
Tr(σ(ℓ)Qℓ(x, y)),
We notice that the matrix-valued function
Qℓ(x, y) = eℓ(y)eℓ(x)⊤ ∈ Cdℓ×dℓ.
(Qℓ(x, y))mk = em
ℓ (x)ek
ℓ (y)
is of rank one for every ℓ. Indeed, (Qℓ(x, y))mk is nothing else but the tensor product
of the vectors eℓ(x), eℓ(y) ∈ Cdℓ. Since on a normed space F we have ku ⊗ vkop =
kukF kvkF , we get
kQℓ(x, y)kop = keℓ(x)kℓ2(Cdℓ )keℓ(y)kℓ2(Cdℓ ).
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
33
From (7.2) we have
Hence
σ(ℓ) = ZM ×M
K(x, y)Qℓ(x, y)∗dxdy.
kσ(ℓ)kop ≤ kKkL1(M ×M ) sup
x,y
= kKkL1(M ×M ) sup
x,y
kQℓ(x, y)∗kop
keℓ(x)kℓ2(Cdℓ )keℓ(y)kℓ2(Cdℓ ).
Remark 7.3. We point out that the mere condition K ∈ L1(M × M) does not guar-
antee the L2 boundedness of the corresponding integral operator T . Indeed, consider
M = T1, g ∈ L1(T1)\L2(T1), h ≡ 1 ∈ L1(T1), and the kernel
K(θ, φ) := g(θ)h(φ) ∈ L1(T1 × T1).
It is easy to see that the kernel K(θ, φ) does not define an operator from L2(T1) into
L2(T1). For example, with f = 1 ∈ L2(T1) we have
(T 1)(θ) = g(θ)ZT1
h(φ)dφ = g(θ) /∈ L2(T1).
8. Applications to the nuclearity of operators in Lp(M)
We now turn to the study of nuclearity in Lp-spaces on closed manifolds. Sufficient
conditions for r-nuclearity on Lp on compact Lie groups have been established in
[DR14b]. The study of nuclearity on Lp in this section relies on the analysis of suitable
kernel decompositions and the relation between kernels and symbols described in
Section 7.
Let E and F be two Banach spaces and 0 < r ≤ 1, a linear operator T from E
into F is called r-nuclear if there exist sequences (x′
n) in E′ and (yn) in F so that
E ′kynkr
F < ∞.
(8.1)
kx′
nkr
T x =Xn
hx, x′
ni yn and Xn
When r = 1 they are known as nuclear operators, in that case this definition agrees
with the concept of trace class operator in the setting of Hilbert spaces (E = F = H).
More generally, Oloff proved in [Olo72] that the class of r-nuclear operators coincides
with the Schatten class Sr(H) when E = F = H and 0 < r ≤ 1.
The concept of r-nuclearity was introduced by Grothendieck [Gro55], and it has
application to questions of the distribution of eigenvalues of operators in Banach
spaces via e.g. the Grothendieck-Lidskii formula. We refer to [DR14b] for several
conclusions in the setting of compact Lie groups concerning summability and distri-
bution of eigenvalues of operators on Lp-spaces once we have information on their
r-nuclearity. Since these arguments are then purely functional analytic, they apply
equally well in the present setting of closed manifolds; we omit the repetition but
refer the reader to [DR14b] for several relevant applications.
The r-nuclear operators on Lebesgue spaces are characterised by the following the-
orem (cf. [Del10]). In the statement below we consider (Ω1, M1, µ1) and (Ω2, M2, µ2)
to be two σ-finite measure spaces.
34
JULIO DELGADO AND MICHAEL RUZHANSKY
Theorem 8.1. Let 1 ≤ p1, p2 < ∞, 0 < r ≤ 1 and let q1 be such that 1
= 1.
p1
An operator T : Lp1(µ1) → Lp2(µ2) is r-nuclear if and only if there exist sequences
(gn)n in Lp2(µ2), and (hn)n in Lq1(µ1) such that
Lq1 < ∞, and such
Lp2 khnkr
kgnkr
+ 1
q1
that for all f ∈ Lp1(µ1) we have
T f (x) =Z ∞Xn=1
gn(x)hn(y)! f (y)dµ1(y),
for a.e x.
∞Pn=1
In order to study nuclearity on Lp(M) spaces for a given compact manifold M of
dimension n, we introduce a function Λ(j, k; n, p) which controls the Lp-norms of the
family of eigenfunctions {ek
j } of the operator E, i.e. we will suppose that Λ(j, k; n, p)
is such that we have the estimates
kek
j kLp(M ) ≤ Λ(j, k; n, p).
(8.2)
In particular, if Λ is such a function we observe that
kek
j kLp(M ) ≤ vol(M)
1
p Λ(j, k; n, ∞).
When M = G is a compact Lie group efficient kek
j kLp(G) bounds can be obtained
[DR14b]). The estimation of Lp norms for eigenfunctions of differential elliptic
(cf.
operators on general closed manifolds has been largely studied, see for instance [SZ02].
Some examples will be given at the end of this section. An example can be also
obtained from the following simple lemma:
Lemma 8.2. Let f be such that kf kL2(M ) = 1, then
(i) kf kLp(M ) ≤ (vol(M))
2−p
2p
if 1 ≤ p ≤ 2.
p−2
(ii) kf kLp(M ) ≤ kf k
p
L∞(M ) if 2 ≤ p < ∞.
Proof. (i) By Holder inequality we have
ZM
f (x)pdx ≤(cid:18)ZM
f (x)p 2
p dx(cid:19) p
2(cid:18)ZM
1p 2
2
2−p dx(cid:19) 2−p
= (vol(M))
2−p
2 .
(ii) We also have
ZM
f (x)pdx =ZM
completing the proof.
f (x)p−2f (x)2dx ≤ kf kp−2
L∞(M ),
(cid:3)
We now formulate a sufficient condition for the r-nuclearity on Lp(M) spaces as
an application of the notion of the matrix-symbol on closed manifolds. Inspired by
Lemma 8.2, we will use the following function p for 1 ≤ p ≤ ∞:
0 ,
p−2
p ,
1,
if 1 ≤ p ≤ 2,
if 2 < p < ∞,
if p = ∞.
(8.3)
p :=
For p1, p2 we denote their dual indices by q1 := p′
1, q2 := p′
2.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
35
Theorem 8.3. Let 1 ≤ p1, p2 < ∞ and 0 < r ≤ 1. Let T : Lp1(M) → Lp2(M) be a
strongly invariant linear continuous operator. Assume that its matrix-valued symbol
σ(ℓ) satisfies
∞Xℓ=0
dℓXm,k=1
σ(ℓ)mkrΛ(ℓ, m; n, ∞) p2rΛ(ℓ, k; n, ∞) q1r < ∞.
Then the operator T : Lp1(M) → Lp2(M) is r-nuclear.
Proof. By (7.1) the kernel of T is given by
K(x, y) =
∞Xℓ=0
dℓXm=1
dℓXk=1
We set
σ(ℓ)mkem
ℓ (x)ek
ℓ (y).
gℓ,m,k(x) := σ(ℓ)mkem
ℓ (x), hℓ,k(y) := ek
ℓ (y).
Now, by Lemma 8.2 we have
where Cp = max{(vol(M))
2−p
2p , 1}. We now observe that
kem
ℓ kLp ≤ CpΛ(ℓ, m; n, ∞)p,
kgℓ,m,kkr
Lp2 khℓ,kkr
Lq1 =
Xℓ,m,k
=
∞Xℓ=0
∞Xℓ=0
dℓXm,k=1
dℓXm,k=1
≤(Cp2Cq1)r
kσ(ℓ)mkem
ℓ kr
Lp2 kek
ℓ kr
Lq1
σ(ℓ)mkrkem
ℓ kr
Lp2 kek
ℓ kr
Lq1
∞Xℓ=0
dℓXm,k=1
σ(ℓ)mkrΛ(ℓ, m; n, ∞) p2rΛ(ℓ, k; n, ∞) q1r,
finishing the proof in view of Theorem 8.1.
(cid:3)
In particular for formally self-adjoint invariant operators we can diagonalise each
matrix σ(ℓ), so that we have
Corollary 8.4. Let 1 ≤ p1, p2 < ∞ and 0 < r ≤ 1. Let T : Lp1(M) → Lp2(M)
be a strongly invariant formally self-adjoint continuous operator. Assume that its
matrix-valued symbol σ(ℓ) satisfies
∞Xℓ=0
dℓXm=1
σ(ℓ)mmrΛ(ℓ, m; n, ∞)( p2+ q1)r < ∞.
Then the operator T : Lp1(M) → Lp2(M) is r-nuclear.
In some cases it is possible to simplify the sufficient condition above when the
control function Λ(ℓ, m; n, ∞) is independent of m. For instance a classical result
(local Weyl law) due to Hormander ([Hor68, Theorem 5.1], [Hor85b, Chapter XXIX])
implies the following estimate:
36
JULIO DELGADO AND MICHAEL RUZHANSKY
Lemma 8.5. Let M be a closed manifold of dimension n. Let E ∈ Ψν
+e(M), then
kem
ℓ kL∞ ≤ Cλ
n−1
2ν
ℓ
.
(8.4)
Proof. In order to explain this estimate we first consider the family of eigenvalues
{λℓ} of E ordered in the increasing order
0 = λ0 ≤ λ1 ≤ · · · λℓ ≤ · · ·
and counted with multiplicity. For the projection Pℓ(f ) onto Hℓ, consider Eλf :=
Pℓ(f ) the associated partial sum operators. Its kernel is given by
Pλℓ≤λ
If p(x, ξ) is the principal symbol of E, by Theorem 5.1 of [Hor68] we have
em
ℓ (x)em
ℓ (y).
dℓXm=1
Eλ(x, y) = Xλℓ≤λ
dℓXm=1
em
Eλ(x, x) = Xλℓ≤λ
ℓ (x)2 = (2π)−n Zp(x,ξ)≤λ
with
R(x, λ) ≤ Cλ
n−1
ν , x ∈ M.
dξ + R(x, λ)
(8.5)
Since Eµ(x, x) is increasing right-continuous with respect to µ, the fact that the
dξ with respect to µ and by
spectrum of E is discrete, by the continuity of
taking left-hand limit in (8.5) we obtain
lim
µ→λ−
Eµ(x, x) = Xλℓ<λ
dℓXm=1
Hence
em
Rp(x,ξ)≤µ
ℓ (x)2 = (2π)−n Zp(x,ξ)≤λ
dℓXm=1
em
dξ + R(x, λ−).
Eλℓ(x, x) − Eλ−
ℓ
(x, x) =
ℓ (x)2 = R(x, λℓ) − R(x, λ−
ℓ ).
In particular, we have
em
ℓ (x) ≤ 2(pR(x, λℓ) +qR(x, λ−
ℓ ) ) ≤ 2Cλ
n−1
2ν
ℓ
,
(cid:3)
which proves Lemma 8.5.
Thus Λ(ℓ; n, ∞) = Cλ
2ν
furnishes an example of Λ independent of m. For controls
of type Λ(ℓ; n, ∞) we have a basis-independent condition:
n−1
ℓ
Corollary 8.6. Let 1 ≤ p1, p2 < ∞ and 0 < r ≤ 1. Let T : Lp1(M) → Lp2(M)
be a strongly invariant formally self-adjoint continuous operator. Assume that its
matrix-valued symbol σ(ℓ) satisfies
kσ(ℓ)kr
SrΛ(ℓ; n, ∞)( p2+ q1)r < ∞.
∞Xℓ=0
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
37
Then the operator T : Lp1(M) → Lp2(M) is r-nuclear. In particular, if its matrix-
valued symbol σ(ℓ) satisfies
kσ(ℓ)kr
Sr λ
(n−1)
2ν
ℓ
( p2+ q1)r
< ∞,
(8.6)
∞Xℓ=0
then the operator T : Lp1(M) → Lp2(M) is r-nuclear.
Proof. Since T is E-invariant and formally self-adjoint, each matrix σ(ℓ) can be as-
sumed diagonal, and the result follows from Corollary 8.4 since
σ(ℓ)mmr = Tr(σ(ℓ)r) = kσ(ℓ)kr
Sr,
dℓXm=1
completing the proof. The r-nuclearity under condition (8.6) follows by using Lemma
8.5 and taking Λ(ℓ; n, ∞) = Cλ
n−1
2ν
ℓ
.
(cid:3)
Remark 8.7. If M is a compact Lie group Corollary 8.6 absorbs Theorem 3.4 in
[DR14b] by taking E to be the Laplacian and the family of eigenfunctions {ek
ℓ } as in
1
(6.7). Indeed, since d
and taking into
(ξℓ)ij(x) ≤ d
2
ξℓ
account that, by Lemma 6.1, we have
one can choose Λ(ℓ; ∞) = d
1
2
ξℓ
1
2
ξℓ
kσ(ℓ)kr
Sr = dξℓkτ (ξℓ)kr
Sr ,
we obtain
Xℓ
kσ(ℓ)kr
SrΛ(ℓ; ∞)( p2+ q1)r =Xℓ
2 ( p2+ q1)r
1+ 1
ξℓ
d
kτ (ξℓ)kr
Sr ,
with a right-hand side equivalent to the term giving the sufficient condition in The-
orem 3.4 of [DR14b]. Indeed,
1
2
( p2 + q1) =
1
2(cid:18)1 −
2
max{2, p2}
1
−
2
max{2, q1}(cid:19)
+ 1 −
1
=1 −
=
max{2, q1}
1
−
max{2, p2}
1
,
min{2, p1}
max{2, p2}
which was the order obtained in [DR14b, Theorem 3.4] on compact Lie groups.
In order to give another example we recall Proposition 5.3 with useful relations
between the eigenvalues λj and their multiplicities dj. As a consequence of Corollary
8.6 and Proposition 5.3, for the negative powers of the operator E itself we obtain:
Corollary 8.8. Let 1 ≤ p1, p2 < ∞ and 0 < r ≤ 1. Let E ∈ Ψν
+e(M). If
α >
n
r
+ ( p2 + q1)
n − 1
2
then the operator (I + E)− α
ν : Lp1(M) → Lp2(M) is r-nuclear.
38
JULIO DELGADO AND MICHAEL RUZHANSKY
Note that if p1 = p2 = 2, we have p2 = q1 = 0, and since Schatten class Sr
and r-nuclear class coincide on L2(M), Proposition 5.4 shows that the statement of
Corollary 8.8 is sharp in this case of indices. However, it does depend on the bounds
for eigenvalues which can be improved in the presence of additional structures as
discussed in Remark 8.9.
Proof of Corollary 8.8. If we denote by λℓ the eigenvalues of E, for α > 0 we observe
that σ(I+E)− α
ν (ℓ) = (1 + λℓ)− α
ν Idℓ. Then
kσ(I+E)− α
ν (ℓ)kr
Sr = (1 + λℓ)− αr
ν dℓ.
Now by applying Corollary 8.6 we obtain
kσ(ℓ)kr
Srλ
(n−1)
2ν
ℓ
( p2+ q1)r
Xℓ
≤ CXℓ
dℓ(1 + λℓ)− αr
ν (1 + λℓ)( p2+ q1) (n−1)r
2ν
= CXℓ
dℓ(1 + λℓ)(−α+( p2+ q1) (n−1)
2
) r
ν < ∞,
if q = (α − ( p2 + q1) (n−1)
condition α > n
) r
ν > n
r + ( p2 + q1) n−1
2 .
2
ν by Proposition 5.3. But this is equivalent to the
(cid:3)
Remark 8.9. As we pointed out in Remark 8.7, on compact Lie groups we can always
choose E to be a Laplacian with an orthonormal basis given by rescaled matrix
elements of representations, for which we can take Λ(ℓ; ∞) = d
ℓ . At the same
time, if E is an operator of second order (so that ν = 2) the best we can hope
= d
1
2
ξℓ
1
4
n−1
4
ℓ
1
4
n
8
ℓ . λ
view of (5.1), we always have d
for on closed manifolds in general is Λ(ℓ; n, ∞) = Cλ
given by Lemma 8.5. In
ℓ , so that this choice on compact Lie groups is
better than the general bound Λ(ℓ; n, ∞) = Cλ
above. Partly, this is explained
by the presence of the additional (group) structure in this case. The other point is
that there is a difference in finding L∞-estimates for elements of any orthonormal
basis as opposed to estimates for a favourable one that may exist due to additional
assumptions or structures. However, the latter one seems to be the question much
less studied in the literature, see [SZ02] or [TZ02] for some partial discussions.
n−1
ℓ
4
We now give an example of the above remark in the case of the the sphere S3 ≃
SU(2). We consider the Laplacian (the Casimir element) E = −LS3. We will apply
ℓ . For the
2 , since the eigenvalues of I + E are of the form (1 + ℓ)ℓ we obtain
the condition given by Theorem 8.6 along with the control Λ(ℓ, ∞) = d
symbol of (I + E)− α
1
4
kσ(I+E)− α
2 (ℓ)kr
Sr = ((1 + ℓ)ℓ)− αr
2 dℓ ≈ ((1 + ℓ)ℓ)− αr
2 ℓ2 ≈ (1 + ℓ2)1− αr
2 .
Therefore, using dℓ ≈ ℓ2,
kσ(I+E)− α
2 (ℓ)kr
Xℓ
SrΛ(ℓ, ∞)( p2+ q1)r ≤Xℓ
≈Xℓ
(1 + ℓ2)1− αr
2 ℓ
1
2 ( p2+ q1)r
(1 + ℓ)2−αr+ 1
2 ( p2+ q1)r.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
39
The series on the right-hand side converges if and only if 2 − αr + 1
Thus, the condition
2( p2 + q1)r < −1.
α >
ensures the membership of (I + E)− α
we have proved the following:
+
1
2
( p2 + q1)
3
r
2 in the Schatten class of order r. Summarising,
Corollary 8.10. If α > 3
r-nuclear from Lp1(S3) into Lp2(S3).
r + 1
2( p2 + q1), 0 < r ≤ 1, the operator (I − LS3)− α
2
is
Corollary 8.10 gives a direct proof of Corollary 3.19 in [DR14b] which was proved
there in the group setting.
Remark 8.11. It is clear that the sharpness of the sufficient conditions obtained in this
section depends on how sharp is the Λ function we can choose. For instance the best
situation for Λ(ℓ, ∞) is when it can be chosen constant, i.e. when the eigenfunctions
are uniformly bounded. This is the case of the torus Tn and unfortunately may be
essentially the only one, see [TZ02].
References
[Ati68] M. F. Atiyah. Global aspects of the theory of elliptic differential operators. In Proc. In-
ternat. Congr. Math. (Moscow, 1966), pages 57 -- 64. Izdat. "Mir", Moscow, 1968.
[BBR96] P. Boggiatto, E. Buzano, and L. Rodino. Global hypoellipticity and spectral theory, vol-
ume 92 of Mathematical Research. Akademie Verlag, Berlin, 1996.
[BN04] E. Buzano and F. Nicola. Pseudo-differential operators and Schatten-von Neumann classes.
In Advances in pseudo-differential operators, volume 155 of Oper. Theory Adv. Appl., pages
117 -- 130. Birkhauser, Basel, 2004.
[BN07] E. Buzano and F. Nicola. Complex powers of hypoelliptic pseudodifferential operators. J.
Funct. Anal., 245(2):353 -- 378, 2007.
[BP61] A. Benedek and R. D. Panzone. The spaces Lp, with mixed norms. Duke. Math. J., 28:301 --
324, 1961.
[Bru68] F. Bruhat. Lectures on Lie groups and representations of locally compact groups. Tata
Institute of Fundamental Research, Bombay, 1968. Notes by S. Ramanan, Tata Institute
of Fundamental Research Lectures on Mathematics, No. 14.
[BT10] E. Buzano and J. Toft. Schatten-von Neumann properties in the Weyl calculus. J. Funct.
Anal., 259(12):3080 -- 3114, 2010.
[Car16] T. Carleman. Uber die Fourierkoeffizienten einer stetigen Funktion. Acta Math., 41(1):377 --
384, 1916. Aus einem Brief an Herrn A. Wiman.
[Cho11] O. Chodosh. Infinite matrix representations of isotropic pseudodifferential operators. Meth-
[Del10]
[Dix77]
[Dix96]
[DR13]
ods Appl. Anal., 18(4):351 -- 371, 2011.
J. Delgado. The trace of nuclear operators on Lp(µ) for σ-finite Borel measures on second
countable spaces. Integral Equations Operator Theory, 68(1):61 -- 74, 2010.
J. Dixmier. C ∗-algebras. North-Holland Publishing Co., Amsterdam-New York-Oxford,
1977. Translated from the French by Francis Jellett, North-Holland Mathematical Library,
Vol. 15.
J. Dixmier. Les alg`ebres d'op´erateurs dans l'espace hilbertien (alg`ebres de von Neu-
mann). Les Grands Classiques Gauthier-Villars. [Gauthier-Villars Great Classics]. ´Editions
Jacques Gabay, Paris, 1996. Reprint of the second (1969) edition.
J. Delgado and M. Ruzhansky. Schatten classes and traces on compact Lie groups.
arXiv:1303.3914v1, 2013.
[DR14a] J. Delgado and M. Ruzhansky. Kernel and symbol criteria for Schatten classes and r-
nuclearity on compact manifolds. C. R. Math. Acad. Sci. Paris, 352(10):779 -- 784, 2014.
40
JULIO DELGADO AND MICHAEL RUZHANSKY
[DR14b] J. Delgado and M. Ruzhansky. Lp-nuclearity, traces, and Grothendieck-Lidskii formula on
compact Lie groups. J. Math. Pures Appl. (9), 102(1):153 -- 172, 2014.
[DR14c] J. Delgado and M. Ruzhansky. Schatten classes on compact manifolds: kernel conditions.
[Dui11]
[GK69]
J. Duistermaat. Fourier
J. Funct. Anal., 267(3):772 -- 798, 2014.
J.
operators. Modern Birkhauser Classics.
Birkhauser/Springer, New York, 2011. Reprint of the 1996 edition [MR1362544],
based on the original lecture notes published in 1973 [MR0451313].
I. C. Gohberg and M. G. Kreın. Introduction to the theory of linear nonselfadjoint ope-
rators. Translated from the Russian by A. Feinstein. Translations of Mathematical Mono-
graphs, Vol. 18. American Mathematical Society, Providence, R.I., 1969.
integral
[Gro55] A. Grothendieck. Produits tensoriels topologiques et espaces nucl´eaires. Mem. Amer.
Math. Soc., 1955(16):140, 1955.
[GW73] S. J. Greenfield and N. R. Wallach. Remarks on global hypoellipticity. Trans. Amer. Math.
Soc., 183:153 -- 164, 1973.
[Hor68] L. Hormander. The spectral function of an elliptic operator. Acta Math., 121:193 -- 218,
1968.
[Hor85a] L. Hormander. The Analysis of linear partial differential operators, vol. III. Springer-
Verlag, 1985.
[Hor85b] L. Hormander. The Analysis of linear partial differential operators, vol. IV. Springer-
Verlag, 1985.
[HP10] A. Hinrichs and A. Pietsch. p-nuclear operators in the sense of Grothendieck. Math. Nachr.,
283(2):232 -- 261, 2010.
[Kon78] H. Konig. Eigenvalues of p-nuclear operators. In Proceedings of the International Confer-
ence on Operator Algebras, Ideals, and their Applications in Theoretical Physics (Leipzig,
1977), pages 106 -- 113, Leipzig, 1978. Teubner.
[NR10] F. Nicola and L. Rodino. Global pseudo-differential calculus on Euclidean spaces, volume 4
of Pseudo-Differential Operators. Theory and Applications. Birkhauser Verlag, Basel, 2010.
[Olo72] R. Oloff. p-normierte Operatorenideale. Beitrage Anal., (4):105 -- 108, 1972. Tagungsbericht
zur Ersten Tagung der WK Analysis (1970).
[Pie84] A. Pietsch. Grothendieck's concept of a p-nuclear operator. Integral Equations Operator
Theory, 7(2):282 -- 284, 1984.
[RL13] O. I. Reinov and Q. Laif. Grothendieck-Lidskii theorem for subspaces of Lp−spaces. Math.
Nachr., (2 -- 3):279 -- 282, 2013.
[RS75] M. Reed and B. Simon. Methods of modern mathematical physics. II. Fourier analysis,
self-adjointness. Academic Press [Harcourt Brace Jovanovich Publishers], New York, 1975.
[RS80] M. Reed and B. Simon. Methods of modern mathematical physics. I. Academic Press,
Inc. [Harcourt Brace Jovanovich, Publishers], New York, second edition, 1980. Functional
analysis.
[RT10] M. Ruzhansky and V. Turunen. Pseudo-differential operators and symmetries. Background
analysis and advanced topics, volume 2 of Pseudo-Differential Operators. Theory and Ap-
plications. Birkhauser Verlag, Basel, 2010.
[RT13] M. Ruzhansky and V. Turunen. Global quantization of pseudo-differential operators on
compact Lie groups, SU(2), 3-sphere, and homogeneous spaces. Int. Math. Res. Not.
IMRN, (11):2439 -- 2496, 2013.
[RT15] M. Ruzhansky and N. Tokmagambetov. Nonharmonic analysis of boundary value prob-
lems. Int. Math. Res. Notices, doi: 10.1093/imrn/rnv243, 2015.
[Rus74] B. Russo. The Norm of the Lp-Fourier Transform on Unimodular Groups. Trans. Amer.
Math. Soc., 192(2):293 -- 305, 1974.
[Sch70] R. Schatten. Norm ideals of completely continuous operators. Second printing. Ergebnisse
der Mathematik und ihrer Grenzgebiete, Band 27. Springer-Verlag, Berlin, 1970.
[See65] R. T. Seeley. Integro-differential operators on vector bundles. Trans. Amer. Math. Soc.,
117:167 -- 204, 1965.
FOURIER MULTIPLIERS, SYMBOLS AND NUCLEARITY ON COMPACT MANIFOLDS
41
[See67] R. T. Seeley. Complex powers of an elliptic operator. In Singular Integrals (Proc. Sympos.
Pure Math., Chicago, Ill., 1966), pages 288 -- 307. Amer. Math. Soc., Providence, R.I., 1967.
[See69] R. T. Seeley. Eigenfunction expansions of analytic functions. Proc. Amer. Math. Soc.,
21:734 -- 738, 1969.
[Shu01] M. A. Shubin. Pseudodifferential operators and spectral theory. Springer-Verlag, Berlin,
second edition, 2001. Translated from the 1978 Russian original by Stig I. Andersson.
[Sim79] B. Simon. Trace ideals and their applications, volume 35 of London Mathematical Society
Lecture Note Series. Cambridge University Press, Cambridge, 1979.
[Sob14] A. V. Sobolev. On the Schatten -- von Neumann properties of some pseudo-differential ope-
rators. J. Funct. Anal., 266(9):5886 -- 5911, 2014.
[Ste70] E. M. Stein. Topics in harmonic analysis related to the Littlewood-Paley theory. Annals of
Mathematics Studies, No. 63. Princeton University Press, Princeton, N.J., 1970.
[Str72] R. S. Strichartz. A functional calculus for elliptic pseudo-differential operators. Amer. J.
[SZ02]
[Tof06]
[Tof08]
[TZ02]
Math., 94:711 -- 722, 1972.
C. Sogge and S. Zelditch. Riemannian manifolds with maximal eigenfunction growth. Duke
Math. J., 114(3):387 -- 437, 2002.
J. Toft. Schatten-von Neumann properties in the Weyl calculus, and calculus of metrics
on symplectic vector spaces. Ann. Global Anal. Geom., 30(2):169 -- 209, 2006.
J. Toft. Schatten properties for pseudo-differential operators on modulation spaces. In
Pseudo-differential operators, volume 1949 of Lecture Notes in Math., pages 175 -- 202.
Springer, Berlin, 2008.
J. Toth and S. Zelditch. Riemannian manifolds with uniformly bounded eigenfunctions.
Duke Math. J., 111(1):97 -- 132, 2002.
Department of Mathematics, Imperial College London, 180 Queen's Gate, London
SW7 2AZ, United Kingdom
E-mail address: [email protected]
Department of Mathematics, Imperial College London, 180 Queen's Gate, London
SW7 2AZ, United Kingdom
E-mail address: [email protected]
|
1207.5749 | 1 | 1207 | 2012-07-24T17:09:12 | The Bochner-Riesz means for Fourier-Bessel expansions: norm inequalities for the maximal operator and almost everywhere convergence | [
"math.FA"
] | In this paper, we develop a thorough analysis of the boundedness properties of the maximal operator for the Bochner-Riesz means related to the Fourier-Bessel expansions. For this operator, we study weighted and unweighted inequalities in the spaces L^p((0,1),x^{2\nu+1}dx). Moreover, weak and restricted weak type inequalities are obtained for the critical values of p. As a consequence, we deduce the almost everywhere pointwise convergence of these means. | math.FA | math |
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL
EXPANSIONS: NORM INEQUALITIES FOR THE MAXIMAL
OPERATOR AND ALMOST EVERYWHERE CONVERGENCE
´OSCAR CIAURRI AND LUZ RONCAL
Abstract. In this paper, we develop a thorough analysis of the boundedness
properties of the maximal operator for the Bochner-Riesz means related to
the Fourier-Bessel expansions. For this operator, we study weighted and un-
weighted inequalities in the spaces Lp((0, 1), x2ν+1 dx). Moreover, weak and
restricted weak type inequalities are obtained for the critical values of p. As a
consequence, we deduce the almost everywhere pointwise convergence of these
means.
1. Introduction and main results
Let Jν be the Bessel function of order ν. For ν > −1 we have that
j, k = 1, 2, . . .
Jν(sjx)Jν (skx)x dx =
(Jν+1(sj))2δj,k,
1
2
Z 1
0
where {sj}j≥1 denotes the sequence of successive positive zeros of Jν. From the
previous identity we can check that the system of functions
√2
Jν+1(sj)
(1)
ψj(x) =
x−ν Jν(sjx),
j = 1, 2, . . .
is orthonormal and complete in L2((0, 1), dµν), with dµν (x) = x2ν+1 dx (for the
completeness, see [12]). Given a function f on (0, 1), its Fourier series associated
with this system, named as Fourier-Bessel series, is defined by
(2)
f ∼
∞
Xj=1
aj(f )ψj ,
with
aj(f ) =Z 1
0
f (y)ψj(y) dµν(y),
provided the integral exists. When ν = n/2− 1, for n ∈ N and n ≥ 2, the functions
ψj are the eigenfunctions of the radial Laplacian in the multidimensional ball Bn.
The eigenvalues are the elements of the sequence {s2
j}j≥1. The Fourier-Bessel series
corresponds with the radial case of the multidimensional Fourier-Bessel expansions
analyzed in [1].
For each δ > 0, we define the Bochner-Riesz means for Fourier-Bessel series as
R(f, x) =Xj≥1 1 −
Bδ
s2
j
R2!δ
+
aj(f )ψj(x),
Date: July 24, 2012.
2010 Mathematics Subject Classification. Primary: 42C10, Secondary: 42C20, 42A45.
Key words and phrases. Fourier-Bessel expansions, Bochner-Riesz means, almost everywhere
convergence, maximal operators, weighted inequalities.
Research supported by the grant MTM2009-12740-C03-03 from Spanish Government.
1
2
´O. CIAURRI AND L. RONCAL
where R > 0 and (1 − s2)+ = max{1 − s2, 0}. Bochner-Riesz means are a regular
summation method used oftenly in harmonic analysis. It is very common to analyze
regular summation methods for Fourier series when the convergence of the partial
sum fails. Ces`aro means are other of the most usual summation methods. B.
Muckenhoupt and D. W. Webb [14] give inequalities for Ces`aro means of Laguerre
polynomial series and for the supremum of these means with certain parameters
and 1 < p ≤ ∞. For p = 1, they prove a weak type result. They also obtain similar
estimates for Ces`aro means of Hermite polynomial series and for the supremum
of those means in [15]. An almost everywhere convergence result is obtained as a
corollary in [14] and [15]. The result about Laguerre polynomials is an extension of a
previous result in [18]. This kind of matters has been also studied by the first author
and J. L. Varona in [7] for the Ces`aro means of generalized Hermite expansions.
The Ces`aro means for Jacobi polynomials were analyzed by S. Chanillo and B.
Muckenhoupt in [3]. The Bochner-Riesz means themselves have been analyzed for
the Fourier transform and their boundedness properties in Lp(Rn) is an important
unsolved problem for n > 2 (the case n = 2 is well understood, see [2]).
The target of this paper is twofold. First we will analyze the almost everywhere
(a. e.) convergence, for functions in Lp((0, 1), dµν), of the Bochner-Riesz means for
Fourier-Bessel expansions. By the general theory [8, Ch. 2], to obtain this result
we need to estimate the maximal operator
R>0(cid:12)(cid:12)Bδ
Bδ(f, x) = sup
in the Lp((0, 1), dµν) spaces. A deep analysis of the boundedness properties of this
operator will be the second goal of our paper. This part of our work is strongly
inspired by the results given in [3] for the Fourier-Jacobi expansions.
R(f, x)(cid:12)(cid:12) ,
Before giving our results we introduce some notation. Being p0 = 4(ν+1)
2ν+3+2δ and
p1 = 4(ν+1)
2ν+1−2δ , we define
(3)
p0,
p0(δ) =(1,
p1(δ) =(∞,
p1,
δ > ν + 1/2 or − 1 < ν ≤ −1/2,
δ ≤ ν + 1/2 and ν > −1/2,
δ > ν + 1/2 or − 1 < ν ≤ −1/2,
δ ≤ ν + 1/2 and ν > −1/2.
Concerning to the a. e. convergence of the Bochner-Riesz means, our result
reads as follows
Theorem 1. Let ν > −1, δ > 0, and 1 ≤ p < ∞. Then,
Bδ
R(f, x) → f (x) a. e., for f ∈ Lp((0, 1), dµν)
if and only if p0(δ) ≤ p, where p0(δ) is as in (3).
Proof of Theorem 1 is contained in Section 2 and is based on the following
arguments. On one hand, to prove the necessity part, we will show the existence of
functions in Lp((0, 1), dµν) for p < p0(δ) such that Bδ
R diverges for them. In order
to do this, we will use a reasoning similar to the one given by C. Meaney in [13] that
we describe in Section 2. On the other hand, for the sufficiency, observe that the
convergence result follows from the study of the maximal operator Bδf . Indeed, it
is sufficient to get (p0(δ), p0(δ))-weak type estimates for this operator and this will
be the content of Theorem 3.
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
3
Regarding the boundedness properties of Bδf we have the following facts. First,
a result containing the (p, p)-strong type inequality.
Theorem 2. Let ν > −1, δ > 0, and 1 < p ≤ ∞. Then,
if and only if
(cid:13)(cid:13)Bδf(cid:13)(cid:13)Lp((0,1),dµν ) ≤ CkfkLp((0,1),dµν )
(1 < p ≤ ∞,
p0 < p < p1,
for −1 < ν ≤ −1/2 or δ > ν + 1/2,
for δ ≤ ν + 1/2 and ν > −1/2.
In the lower critical value of p0(δ) we can prove a (p0(δ), p0(δ))-weak type esti-
mate.
Theorem 3. Let ν > −1, δ > 0, and p0(δ) be the number in (3). Then,
with C independent of f .
(cid:13)(cid:13)Bδf(cid:13)(cid:13)Lp0(δ),∞((0,1),dµν ) ≤ CkfkLp0(δ)((0,1),dµν ),
Finally, for the upper critical value, when 0 < δ < ν + 1/2 and ν > −1/2, it is
possible to obtain a (p1, p1)-restricted weak type estimate.
Theorem 4. Let ν > −1/2 and 0 < δ < ν + 1/2. Then,
for all measurable subsets E of (0, 1) and C independent of E.
(cid:13)(cid:13)BδχE(cid:13)(cid:13)Lp1,∞((0,1),dµν ) ≤ CkχEkLp1 ((0,1),dµν ),
The previous results about norm inequalities are summarized in Figure 1 (case
−1 < ν ≤ −1/2) and Figure 2 (case ν > −1/2).
δ
δ
g
n
o
r
t
s
-
)
p
,
p
(
0
k
a
e
w
-
)
p
,
p
(
g
n
o
r
t
s
-
)
p
,
p
(
δ = ν + 1
2
(
p
,
p
)
-
r
e
s
t
r
i
c
.
1
p
1
0
w
e
a
k
2ν+1
4(ν+1)
k
a
e
w
-
)
p
,
p
(
( p, p )- w ea k
2ν+3
4(ν+1)
1
p
1
Figure 1: case −1 < ν ≤ − 1
2 .
Figure 2: case ν > − 1
2 .
At this point, a comment is in order. Note that J. E. Gilbert [9] also proves
weak type norm inequalities for maximal operators associated with orthogonal ex-
pansions. The method used cannot be applied in our case, and the reason is the
same as can be read in [3], at the end of Sections 15 and 16 therein. Following the
technique in [9] we have to analyze some weak type inequalities for Hardy opera-
tor and its adjoint with weights and these inequalities do not hold for p = p0 and
p = p1.
4
´O. CIAURRI AND L. RONCAL
The proof of the sufficiency in Theorem 2 will be deduced from a more general
result in which we analyze the boundedness of the operator Bδf with potential
weights. Before stating it, we need a previous definition. We say that the parame-
ters (b, B, ν, δ) satisfy the Cp conditions if
(4)
(5)
(6)
(7)
(8)
p
b > −2(ν + 1)
B < 2(ν + 1)(cid:18)1 −
b > 2(ν + 1)(cid:18) 1
2 −
B ≤ 2(ν + 1)(cid:18) 1
2 −
B ≤ b,
1
(≥ if p = ∞),
p(cid:19) (≤ if p = 1),
p(cid:19) − δ −
p(cid:19) + δ +
1
2
1
2
,
1
1
(≥ if p = ∞),
and in at least one of each of the following pairs the inequality is strict: (5) and (8),
(6) and (8), and (7) and (8) except for p = ∞. The result concerning inequalities
with potential weights is the following.
Theorem 5. Let ν > −1, δ > 0, and 1 < p ≤ ∞. If (b, B, ν, δ) satisfy the Cp
conditions, then
with C independent of f .
(cid:13)(cid:13)xbBδf(cid:13)(cid:13)Lp((0,1),dµν ) ≤ CkxBfkLp((0,1),dµν ),
A result similar to Theorem 5 for the partial sum operator was proved in [10,
Theorem 1]. It followed from a weighted version of a general Gilbert's maximal
transference theorem, see [9, Theorem 1]. The weighted extension of Gilbert's
result given in [10] depended heavily on the Ap theory and it can not be used
in our case because it did not capture all the information relative to the weights.
On the other hand, it is also remarkable the paper by K. Stempak [19] in which
maximal inequalities for the partial sum operator of Fourier-Bessel expansions and
divergence and convergence results are discussed.
The necessity in Theorem 2 will follow by showing that the operator Bδf is
neither (p1, p1)-weak nor (p0, p0)-strong for ν > −1/2 and 0 < δ ≤ ν + 1/2. This is
the content of the next theorems.
Theorem 6. Let ν > −1/2. Then
kf kLp1 ((0,1),dµν )=1kBδ
sup
RfkLp1,∞((0,1),dµν ) ≥ C(log R)1/p0 ,
if 0 < δ < ν + 1/2; and
kf kL∞((0,1),dµν )=1kBδ
sup
RfkL∞((0,1),dµν ) ≥ C log R,
if δ = ν + 1/2.
Theorem 7. Let ν > −1/2. Then
sup
E⊂(0,1)
kBδ
RχEkLp0 ((0,1),dµν )
kχEkLp0 ((0,1),dµν ) ≥ C(log R)1/p0 ,
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
5
if 0 < δ < ν + 1/2; and
sup
kf kL1((0,1),dµν )=1kBδ
RfkL1((0,1),dµν ) ≥ C log R,
if δ = ν + 1/2.
The paper is organized as follows.
In the next section, we give the proof of
Theorem 1. In Section 3 we first relate the Bochner-Riesz means Bδ
R to the Bochner-
Riesz means operator associated with the Fourier-Bessel system in the Lebesgue
measure setting. Then, we prove weighted inequalities for the supremum of this
new operator. With the connection between these means and the operator Bδ
R, we
obtain Theorem 5 and, as a consequence, the sufficiency of Theorem 2. Sections 4
and 5 will be devoted to the proofs of Theorems 3 and 4, respectively. The proofs
of Theorems 6 and 7 are contained in Section 6. One of the main ingredients in the
proofs of Theorems 6 and 7 will be Lemma 15, this lemma is rather technical and
it will be proved in the Section 7.
Throughout the paper, we will use the following notation: for each p ∈ [1,∞],
p′ = 1. We shall write X ≃ Y
we will denote by p′ the conjugate of p, that is, 1
when simultaneously X ≤ CY and Y ≤ CX.
p + 1
2. Proof of Theorem 1
The proof of the sufficiency follows from Theorem 3 and standard arguments.
In order to prove the necessity, let us see that, for 0 < δ < ν + 1/2 and ν > −1/2,
R(f, x) diverges.
there exists a function f ∈ Lp((0, 1), dµν ), p ∈ [1, p0), for which Bδ
We follow some ideas contained in [13] and [19].
First, we need a few more ingredients. Recall the well-known asymptotics for
the Bessel functions (see [20, Chapter 7])
Jν(z) =
zν
2νΓ(ν + 1)
+ O(zν+2),
z < 1,
arg(z) ≤ π,
(9)
and
(10)
π
z ≥ 1,
νπ
2 −
πz hcos(cid:16)z −
4(cid:17) + O(eIm(z)z−1)i ,
Jν(z) =r 2
where Dν = −(νπ/2 + π/4). It will also be useful the fact that (cf. [6, (2.6)])
(11)
For our purposes, we need estimates for the Lp norms of the functions ψj. These
estimates are contained in the following lemma, whose proof can be read in [5,
Lemma 2.1].
Lemma 1. Let 1 ≤ p ≤ ∞ and ν > −1. Then, for ν > −1/2,
arg(z) ≤ π − θ,
sj = O(j).
kψjkLp((0,1),dµν ) ≃
,
p
j(ν+1/2)− 2(ν+1)
(log j)1/p,
1,
if p > 2(ν+1)
ν+1/2 ,
if p = 2(ν+1)
ν+1/2 ,
if p < 2(ν+1)
ν+1/2 ,
and, for −1 < ν ≤ −1/2,
kψjkLp((0,1),dµν ) ≃(1,
jν+1/2,
if p < ∞,
if p = ∞.
6
´O. CIAURRI AND L. RONCAL
We will also use a slight modification of a result by G. H. Hardy and M. Riesz
for the Riesz means of order δ, that is contained in [11, Theorem 21]. We present
here this result, adapted to the Bochner-Riesz means. We denote by SR(f, x) the
partial sum associated to the Fourier-Bessel expansion, namely
SR(f, x) = X0<sj ≤R
The result reads as follows.
aj(f )ψj(x).
Lemma 2. Suppose that f can be expressed as a Fourier-Bessel expansion and
for some δ > 0 and x ∈ (0, 1) its Bochner-Riesz means Bδ
R(f, x) converges to c as
R → ∞. Then, for sn ≤ R < sn+1,
SR(f, x) − c ≤ Aδnδ
sup
0<t≤sn+1 Bδ
t (f, x).
By using this lemma, we can write
sup
0 > 2(ν+1)
ν+1/2 , and δ < ν + 1/2 − 2(ν+1)
0<t≤sj+1 Bδ
(12) aj(f )ψj(x) = (Ssj (f, x) − c) − (Ssj−1 (f, x) − c) ≤ Aδjδ
t (f, x).
Let us proceed with the proof of the necessity. Let 1 ≤ p < p0. Note that p′
0 = p1.
Therefore, p′ > p′
:= λ. By Lemma 1,
kψjkLp′ ((0,1),dµν ) ≥ Cjλ. Then, we have that the mapping f 7→ aj(f ), where
aj(f ) was given in (2), is a bounded linear functional on Lp((0, 1), dµν ) with norm
bounded below by a constant multiple of jλ. By uniform boundedness principle,
for p conjugate to p′ and each 0 ≤ ε < λ, there is a function f0 ∈ Lp((0, 1), dµν) so
that aj(f0)j−ε → ∞ as j → ∞. By taking ε = δ, we have that
(13)
p′
aj(f0)j−δ → ∞ as
j → ∞.
Suppose now that Bδ
R(f0, x) converges. Then, by Egoroff's theorem, it converges
on a subset E of positive measure in (0, 1) and, clearly, we can think that E ⊂ (η, 1)
for some fixed η > 0. For each x ∈ E, we can consider j such that sjx ≥ 1 and, by
(10),
aj(f0)ψj (x) =(cid:12)(cid:12)aj(f0)(cid:16)
√2
√2
x−ν Jν(sjx)
−
+ aj(f0)
Jν+1(sj)
πsjx(cid:17)1/2
x−ν(cid:16) 2
Jν+1(sj)
x−ν(cid:16) 2
√2
Jν+1(sj)
√2
Jν+1(sj )aj(f0)x−ν−1/2(cid:0)O((sj x)−1) + cos(sjx + Dν)(cid:1)
cos(sj x + Dν )(cid:12)(cid:12)
≃ aj(f0)x−ν−1/2(cos(sjx + Dν) + O((sj x)−1)).
cos(sjx + Dν)(cid:17)
πsjx(cid:17)1/2
= Cs−1/2
j
By (12) on this set E,
aj(f0)x−ν−1/2(cos(sjx + Dν) + O((j)−1)) ≤ Aδjδ
t (f0, x) ≤ KEjδ,
uniformly on x ∈ E. We also used (11) in the latter. The inequality above is
equivalent to
0<t≤sj+1 Bδ
sup
aj(f0)(cos(sj x + Dν) + O(j−1)) ≤ KExν+1/2jδ ≤ KEjδ.
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
7
Therefore,
(14)
Now, taking the functions
aj(f0)j−δ(cos(sjx + Dν) + O((j)−1)) ≤ KE.
Fj (x) = aj(f0)j−δ(cos(sj x + Dν) + O(j−1)),
x ∈ E,
and using an argument based on the Cantor-Lebesgue and Riemann-Lebesgue the-
orems, see [13, Section 1.5] and [21, Section IX.1], we obtain that
ZE Fj (x)2 dx ≥ Caj(f0)j−δ2E,
where, as usual, E denotes the Lebesgue measure of the set E. On the other hand,
by (14),
ZE Fj(x)2 dx ≤ K 2
EE.
Then, from the previous estimates, it follows that aj(f0)j−δ ≤ C, which contra-
dicts (13).
3. Bochner-Riesz means for Fourier-Bessel expansions in the
Lebesgue measure setting. Proof of Theorem 5
For our convenience, we are going to introduce a new orthonormal system. We
will take the functions
φj(x) =
√2xJν(sjx)
Jν+1(sj)
,
j = 1, 2, . . . .
These functions are a slight modification of the functions (1); in fact,
φj (x) = xν+1/2ψj (x).
In this case, the corresponding Fourier-Bessel expansion of a function f is
(15)
The system {φj(x)}j≥1 is a complete orthonormal basis of L2((0, 1), dx).
f (y)φj(y) dy(cid:19)
bj(f ) =(cid:18)Z 1
bj(f )φj (x),
f ∼
with
0
∞
Xj=1
provided the integral exists, and for δ > 0 the Bochner-Riesz means of this expan-
sion are
where R > 0 and (1 − s2)+ = max{1 − s2, 0}. It follows that
s2
j
R2!δ
+
bj(f )φj (x),
f (y)K δ
R(x, y) dy
s2
j
R2!δ
+
φj(x)φj (y).
Bδ
K δ
0
Bδ
R(f, x) =Xj≥1 1 −
R(f, x) =Z 1
R(x, y) =Xj≥1 1 −
Rf (x) =Z 1
Bδ
0
f (y)Kδ
R(x, y) dµν (y),
where
(16)
Our next target is the proof of Theorem 5. Taking into account that
8
where
´O. CIAURRI AND L. RONCAL
R(x, y) =Xj≥1 1 −
Kδ
s2
j
R2!δ
+
ψj(x)ψj (y),
it is clear, from (15), that Kδ
that the inequality
R(x, y) = (xy)−(ν+1/2)K δ
R(x, y). Then, it is verified
kxbBδ(f, x)kLp((0,1),dµν ) ≤ CkxBf (x)kLp((0,1),dµν )
is equivalent to
kxb+(ν+1/2)(2/p−1)Bδ(f, x)kLp((0,1),dx) ≤ CkxB+(ν+1/2)(2/p−1)f (x)kLp((0,1),dx),
that is, we can focus on the study of a weighted inequality for the operator Bδ
The first results about convergence of this operator can be found in [4].
R(f, x).
We are going to prove an inequality of the form
kxaBδ(f, x)kLp((0,1),dx) ≤ CkxAf (x)kLp((0,1),dx)
for δ > 0, 1 < p ≤ ∞, under certain conditions for a, A, ν and δ. Besides, a
weighted weak type result for supR>0 Bδ
R(f, x) will be proved for p = 1. The
abovementioned conditions are the following. Let ν > −1, δ > 0 and 1 ≤ p ≤ ∞;
parameters (a, A, ν, δ) will be said to satisfy the cp conditions provided
(17)
(18)
(19)
(20)
(21)
a > −1/p − (ν + 1/2) (≥ if p = ∞),
A < 1 − 1/p + (ν + 1/2) (≤ if p = 1),
a > −δ − 1/p (≥ if p = ∞),
A ≤ 1 + δ − 1/p,
A ≤ a
and in at least one of each of the following pairs the inequality is strict: (18) and
(21), (19) and (21), and (20) and (21) except for p = ∞.
The main results in this section are the following:
Theorem 8. Let ν > −1, δ > 0 and 1 < p ≤ ∞. If (a, A, ν, δ) satisfy the cp
conditions, then
kxaBδ(f, x)kLp((0,1),dx) ≤ CkxAf (x)kLp((0,1),dx),
with C independent of f .
Theorem 9. Let ν > −1 and δ > 0. If (a, A, ν, δ) satisfy the c1 conditions and
Eλ =(cid:26)x ∈ (0, 1) : xa sup
R>0(cid:0)Bδ
R(f, x)(cid:1) > λ(cid:27) ,
then
Eλ ≤ C kxAf (x)kL1((0,1),dx)
λ
,
with C independent of f and λ.
Note that, taking a = b + (ν + 1/2)(2/p − 1) and A = B + (ν + 1/2)(2/p − 1),
Theorem 5 follows from Theorem 8.
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
9
The proofs of Theorem 8 and Theorem 9 will be achieved by decomposing the
square (0, 1) × (0, 1) into five regions and obtaining the estimates therein. The
regions will be:
(22)
A1 = {(x, y) : 0 < x, y ≤ 4/R},
A2 = {(x, y) : 4/R < max{x, y} < 1, x − y ≤ 2/R},
A3 = {(x, y) : 4/R ≤ x < 1, 0 < y ≤ x/2},
A4 = {(x, y) : 0 < x ≤ y/2, 4/R ≤ y < 1},
A5 = {(x, y) : 4/R < x < 1, x/2 < y < x − 2/R}
∪ {(x, y) : y/2 < x ≤ y − 2/R, 4/R ≤ y < 1}.
Theorem 8 and Theorem 9 will follow by showing that, if 1 ≤ p ≤ ∞, then
sup
(23)
y−AxaK δ
holds for j = 1, 3, 4 and that
R>0Z 1
(cid:13)(cid:13)(cid:13)(cid:13)
R(x, y)f (y)χAj dy(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx) ≤ Ckf (x)kLp((0,1),dx)
0
(24)
y−AxaK δ
R(x, y)f (y)χAj dy ≤ CM (f, x),
Z 1
0
for j = 2, 5, where M is the Hardy-Littlewood maximal function of f , and C is
independent of R, x and f . These results and the fact that M is (1, 1)-weak and
(p, p)-strong if 1 < p ≤ ∞ complete the proofs.
K δ
To get (23) and (24) we will use a very precise pointwise estimate for the kernel
R(x, y), obtained in [4]; there, it was shown that
(25)
with
(26)
K δ
(xy)ν+1/2R2(ν+1),
R,
Φν (Rx)Φν (Ry)
R(x, y) ≤ C
Φν (t) =(tν+1/2,
Rδ x−yδ+1
1,
,
if 0 < t < 2,
if t ≥ 2.
(x, y) ∈ A1,
(x, y) ∈ A2
(x, y) ∈ A3 ∪ A4 ∪ A5,
The proof of (24) follows from the given estimate for the kernel K δ
In the case of A2, from K δ
R(x, y) and
y−Axa ≃ C in A2 ∪ A5 because A ≤ a.
R(x, y) ≤
CR we deduce easily the required inequality. For A5 the result is a consequence
of Φν(Rx)Φν (Ry) ≤ C and of a decomposition of the region in strips such that
Rx − y ≃ 2k, with k = 0, . . . , [log2 R] − 1; this can be seen in [4, p. 109]
In this manner, to complete the proofs of Theorem 8 and Theorem 9 we only
have to show (23) for j = 1, 3, 4 in the conditions cp for 1 ≤ p ≤ ∞, and this is
the content of Corollary 1 in Subsection 3.2. In its turn, Corollary 1 follows from
Lemmas 9 and 10 in the same subsection. Previously, Subsection 3.1 contains some
technical lemmas that will be used in the proofs of Lemmas 9 and 10.
3.1. Technical Lemmas. To prove (23) for j = 1, 3, 4 we will use an interpolation
argument based on six lemmas. These are stated below. They are small modifica-
tions of the six lemmas contained in Section 3 of [14] where a sketch of their proofs
can be found.
10
´O. CIAURRI AND L. RONCAL
Lemma 3. Let ξ0 > 0, if r < −1, r + t ≤ −1 and r + s + t ≤ −1, then for p = 1
xrχ[1,∞)(x)
sup
ξ0≤ξ≤x
ξsZ x
ξ
ytf (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),dx)
≤ Ckf (x)kLp((0,∞),dx)
with C independent of f . If r ≤ 0, r + t ≤ −1 and r + s + t ≤ −1 with equality
holding in at most one of the first two inequalities, then this holds for p = ∞.
Lemma 4. Let ξ0 > 0, if t ≤ 0, r + t ≤ −1 and r + s+ t ≤ −1, with strict inequality
in the last two in case of equality in the first, then for p = 1
xrχ[1,∞)(x)
sup
ξ0≤ξ≤x
ξsZ ∞
x
ytf (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),dx)
with C independent of f . If t < −1, r + t ≤ −1 and r + s + t ≤ −1, then this holds
for p = ∞.
Lemma 5. If s < 0, s + t ≤ 0 and r + s + t ≤ −1,with equality holding in at most
one of the last two inequalities, then for p = 1
≤ Ckf (x)kLp((0,∞),dx)
xrχ[1,∞)(x) sup
ξ≥x
ξsZ ξ
x
ytf (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),dx)
≤ Ckf (x)kLp((0,∞),dx)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
with C independent of f . If s < 0, s + t ≤ −1 and r + s + t ≤ −1 this holds for
p = ∞.
Lemma 6. If t ≤ 0, s + t ≤ 0 and r + s + t ≤ −1,with strict inequality holding in
the first two in case the third is an equality, then for p = 1
xrχ[1,∞)(x) sup
ξ≥x
ξsZ ∞
ξ
ytf (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),dx)
≤ Ckf (x)kLp((0,∞),dx)
with C independent of f . If t < −1, s + t ≤ −1 and r + s + t ≤ −1 then this holds
for p = ∞.
Lemma 7. If s < 0, r + s < −1 and r + s + t ≤ −1, then for p = 1
xrχ[1,∞)(x) sup
ξ≥x
ξsZ x
1
ytf (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),dx)
≤ Ckf (x)kLp((0,∞),dx)
with C independent of f . If s < 0, r + s ≤ 0 and r + s + t ≤ −1, with equality
holding in at most one of the last two inequalities, this holds for p = ∞.
Lemma 8. If r < −1, r + s < −1 and r + s + t ≤ −1, then for p = 1
xrχ[1,∞)(x) sup
1≤ξ≤x
ξsZ ξ
1
ytf (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),dx)
≤ Ckf (x)kLp((0,∞),dx)
with C independent of f . If r ≤ 0, r + s ≤ 0 and r + s + t ≤ −1, with equality in
at most one of the last two inequalities, this holds for p = ∞.
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
11
3.2. Proofs of Theorem 8 and Theorem 9 for regions A1, A3 and A4. This
section contains the proofs of the inequality (23) for regions A1, A3 and A4. The
results we will prove are included in the following
Lemma 9. If ν > −1, δ > 0, R > 0, j = 1, 3, 4 and (a, A, ν, δ) satisfy the c1
conditions, then (23) holds for p = 1 with C independent of f .
Lemma 10. If ν > −1, δ > 0, R > 0, j = 1, 3, 4 and (a, A, ν, δ) satisfy the c∞
conditions, then (23) holds for p = ∞ with C independent of f .
Corollary 1. If 1 ≤ p ≤ ∞, ν > −1, δ > 0, R > 0, (a, A, ν, δ) satisfy the cp
conditions and j = 1, 3, 4, then (23) holds with C independent of f .
Proof of Corollary 1. It is enough to observe that if 1 < p < ∞ and (a, A, ν, δ)
satisfy the cp conditions, then (a−1+1/p, A−1+1/p, ν, δ) satisfy the c1 conditions.
So, by Lemma 9
(cid:13)(cid:13)(cid:13)(cid:13)
Z 1
0
sup
R≥0Z 1
0
y−A+1−1/pxa−1+1/pK δ
R(x, y)χAj (x, y)f (y) dy(cid:13)(cid:13)(cid:13)(cid:13)L1((0,1),dx)
≤ Ckf (x)kL1((0,1),dx),
and this is equivalent to
xa+1/p(cid:18)sup
R≥0Z 1
0 K δ
R(x, y)χAj (x, y)f (y) dy(cid:19) dx
x ≤ CZ 1
0
xA+1/pf (x)
dx
x
,
where j = 1, 3, 4. Similarly, if (a, A, ν, δ) verify the cp conditions, then (a + 1/p, A +
1/p, ν, δ) satisfy the c∞ conditions. Hence, by Lemma 10
Now, we can use the Marcinkiewicz interpolation theorem to obtain the inequality
≤ CkxA+1/pf (x)kL∞((0,1),dx).
xa+1/p sup
R≥0Z 1
(cid:13)(cid:13)(cid:13)(cid:13)
Z 1
0 (cid:18)xa+1/p(cid:18)sup
0 K δ
R(x, y)χAj (x, y)f (y) dy(cid:13)(cid:13)(cid:13)(cid:13)L∞((0,1),dx)
R≥0Z 1
R(x, y)χAj (x, y)f (y) dy(cid:19)(cid:19)
≤ CZ 1
0 K δ
p
dx
x
0 (cid:16)xA+1/pf (x)(cid:17)p dx
x
,
for 1 < p < ∞ and the proof is finished.
Finally, we will prove Lemmas 9 and 10 for Aj , j = 1, 3 and 4, separately.
Proof of Lemma 9 and Lemma 10 for A1. First of all, we have to note that
R(f, x) = 0 when 0 < R < s1, being s1 the first positive zero of Jν. Using the
Bδ
estimate (25), the left side of (23) in this case is bounded by
Making the change of variables x = 4/u and y = 4/v, we have
xa+ν+1/2χ[0,1](x)
sup
s1<R≤4/x
R2(ν+1)Z 4/R
0
u−a−ν− 1
2 − 2
p χ[4,∞)(u)
sup
s1≤R≤u
R2(ν+1)Z ∞
R
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
C(cid:13)(cid:13)(cid:13)(cid:13)
.
y−A+ν+1/2f (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
p g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
2 )−2+ 2
vA−(ν+ 1
,
12
´O. CIAURRI AND L. RONCAL
where k · kLp((0,∞),du) denotes the Lp norm in the variable u, and
g(v) = v−2/pf (4v−1).
Note that function g(v) is supported in (1,∞) and kgkLp((0,∞),du) = kfkLp((0,1),dx).
The function g will be used through the subsection, but the value 4 may be changed
by another one, at some points, without comment. Now, splitting the inner integral
at u, we obtain the sum of
(27)
u−a−ν− 1
2 − 2
p χ[4,∞)(u)
sup
s1≤R≤u
R2(ν+1)Z u
R
vA−(ν+ 1
2 )−2+ 2
u−a−ν− 1
2 − 2
p χ[4,∞)(u)
sup
s1≤R≤u
R2(ν+1)Z ∞
u
vA−(ν+ 1
2 )−2+ 2
From Lemma 3 we get the required estimate for (27), using conditions (17) and
(21); Lemma 4 is applied to inequality (28), there we need conditions (18) and (21)
and the restriction on them. This completes the proof of Lemmas 9 and 10 for
j = 1.
Proof of Lemma 9 and Lemma 10 for A3. Clearly, the left side of (23) is
bounded by
p g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
p g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
.
Splitting the inner integral at 2/R, using the bound for the kernel given in (25) and
the definition of Φν, we have this expression majorized by the sum of
and
(28)
C(cid:13)(cid:13)(cid:13)(cid:13)
C(cid:13)(cid:13)(cid:13)(cid:13)
(29)
and
(30)
(cid:13)(cid:13)(cid:13)(cid:13)
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
xaχ[4/R,1](x) sup
4/x≤RZ x/2
0
y−AK δ
.
(Ry)ν+1/2y−A
R(x, y)f (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
Rδx − yδ+1 dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
Rδx − yδ+1 dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
f (y)y−A
.
xaχ[0,1](x) sup
4/x≤RZ 2/R
0
f (y)
xaχ[0,1](x) sup
4/x≤RZ x/2
2/R
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
For (29), taking into account that x − y ≃ x in A3, the changes of variables
x = 4/u, y = 2/v give us
u−a+(δ+1)− 2
p χ[4,∞)(u) sup
u≤R
R−δ+(ν+1/2)Z ∞
R
v−(ν+1/2)+A+ 2
p −2g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
.
Lemma 6 can be used here. The required conditions for p = 1 are (18), (20) and
(21) with the restriction in the pairs therein. For p = ∞ the same inequalities are
needed.
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
13
On the other hand, in (30), using again that x− y ≃ x, by changing of variables
x = 4/u and y = 2/v we have
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
u−a+(δ+1)− 2
p χ[4,∞)(u) sup
u≤R
R−δZ R
2u
u−a+(δ+1)− 2
p χ[4,∞)(u) sup
u≤R
≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
vA+ 2
p −2g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
R−δZ R
vA+ 2
u
p −2g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
.
Lemma 5 can then be applied. For p = 1, we need δ > 0, which is an hypothesis,
and (20) and (21) with its corresponding restriction. For p = ∞ the inequalities
are the same, with the requirement that (20) is strict. This completes the proof of
Lemmas 9 and 10 for j = 3.
Proof of Lemma 9 and Lemma 10 for A4. In this case, the left hand side
of (23) is estimated by
To majorize this, we decompose the R-range in two regions: 4 < R ≤ 2/x and
In this manner, with the bound for the kernel given in (25) and the
R ≥ 2/x.
definition of Φν, the previous norm is controlled by the sum of
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
xaχ[0,1/2](x) sup
R>4Z 1
max(4/R,2x)
xaχ[0,1/2](x)
4/R f (y)
sup
4<R≤2/xZ 1
R≥2/xZ 1
2x
xaχ[0,1/2](x) sup
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
.
y−AK δ
(Rx)ν+1/2y−A
f (y)y−A
R(x, y)f (y) dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
Rδx − yδ+1 dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
Rδx − yδ+1 dy(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,1),dx)
2 )Z R/4
p −2+(δ+1)g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
vA+ 2
vA+ 2
1
.
.
p −2+(δ+1)g(v) dv(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp((0,∞),du)
and
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
and
(32)
Next, using that x − y ≃ y in A4, with the changes of variables x = 2/u and
y = 1/v the previous norms are controlled by
(31)
u−a− 2
p −(ν+ 1
2 )χ[4,∞)(u) sup
4<R≤u
R−δ+(ν+ 1
u−a− 2
p χ[4,∞)(u) sup
R≥u
R−δZ u/4
1
C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
In (31), we use Lemma 8; for p = 1, conditions (17), (19) and (21) are needed; we
need the same for p = ∞. For (32), Lemma 7 requires the hypothesis δ > 0 and
conditions (19) and (21) for p = 1 and the same for p = ∞ with the restrictions in
the pairs therein. This proves Lemmas 9 and 10 for j = 4.
Now we shall prove Theorem 3. First note that, by (15), we can write
4. Proof of Theorem 3
R(f, x) =Z 1
Bδ
0
f (y)(cid:16) y
x(cid:17)ν+1/2
K δ
R(x, y) dy,
14
´O. CIAURRI AND L. RONCAL
where K δ
is enough to check that
R is the kernel in (16). By taking g(y) = f (y)yν+1/2, to prove the result it
ZE
dµν (x) ≤
0 g(x)px(ν+1/2)(2−p) dx,
C
λp Z 1
i=1 Ji, where
R(x, y) dy > λo and p = p0(δ).
where E =nx ∈ (0, 1) : supR>0 x−(ν+1/2)R 1
0 g(y)K δ
We decompose E into four regions, such that E =S4
x−(ν+1/2)Z 1
Ji =(cid:26)x ∈ (0, 1) : sup
R(x, y) dy > λ(cid:27)
0 g(y)χBi (x, y)K δ
for i = 1, . . . , 4, with B1 = A1, B2 = A2 ∪ A5, B3 = A3, and B4 = A4 where the
sets Ai were defined in (22). Note also thatRE dµν (x) ≤P4
i=1RJi
dµν(x), then we
need to prove that
0 g(x)px(ν+1/2)(2−p) dx,
dµν(x) ≤
(33)
R>0
C
λp Z 1
ZJi
for i = 1, . . . , 4 and p = p0(δ). At some points along the proof we will use the
notation
(34)
Ip :=Z 1
0 g(y)py(ν+1/2)(2−p) dy.
In J1, by applying (25) and Holder inequality with p = p0, we have
x−(ν+1/2)Z 1
0 g(y)χB1 (x, y)K δ
R(x, y) dy
0
≤ Cx−(ν+1/2)Z 4/R
≤ CR2(ν+1) Z 4/R
p0 Z 4/R
= CR
2(ν+1)
0
0
g(y)(xy)ν+1/2R2(ν+1) dy
g(y)p0 y(ν+1/2)(2−p0) dy!1/p0 Z 4/R
g(y)p0y(ν+1/2)(2−p0) dy!1/p0
≤ CR
0
y(2ν+1) dy!1/p′
0
2(ν+1)
p0
I 1/p0
p0
.
Therefore,
sup
R>0
x−(ν+1/2)Z 1
0 g(y)χB1 (x, y)K δ
R(x, y) dy ≤ C sup
R>0
χ[0,4/R](x)R
2(ν+1)
p0
I 1/p0
p0
≤ Cx− 2(ν+1)
p0
I 1/p0
p0
.
In the case p = 1, it is clear that
and
x−(ν+1/2)Z 1
x−(ν+1/2)Z 1
sup
R>0
Hence, for p = p0(δ),
0 g(y)χB1(x, y)K δ
R(x, y) dy ≤ CR2(ν+1)I1
0 g(y)χB1(x, y)K δ
R(x, y) dy ≤ Cx−2(ν+1)I1.
J1 ⊆ {x ∈ (0, 1) : Cx− 2(ν+1)
p
I 1/p
p > λ},
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
15
and this gives (33) for i = 1.
In J3, note first that
sup
R>0
= sup
R>0
x−(ν+1/2)Z 1
0 g(y)χB3 (x, y)K δ
x−(ν+1/2)χ[4/R,1](x) Z 2/R
0
R(x, y) dy
g(y)K δ
R(x, y) dy +Z x/2
2/R g(y)K δ
R(x, y) dy!
For R1, using (25), the inequality x/2 < x − y, which holds in B3, and Holder
inequality with p = p0,
:= R1 + R2.
R1 ≤ sup
R>0
x−(ν+3/2+δ)χ[4/R,1](x)Z 2/R
0
Rν+1/2−δyν+1/2g(y) dy
≤ sup
R>0
x−(ν+3/2+δ)χ[4/R,1](x)Rν+1/2−δR
− 2(ν+1)
p′
0
p0 ≤ Cx− 2(ν+1)
I 1/p0
p0
I 1/p0
p0
,
where Ip0 is the same as in (34). In the case p = 1, the estimate R1 ≤ Cx−2(ν+1)I1
can be obtained easily.
On the other hand, for R2, by using (25) and Holder inequality with p = p0
again,
R2 ≤ sup
R>0
≤ sup
R>0
x−(ν+3/2+δ)χ[4/R,1](x)I 1/p0
x−(ν+3/2+δ)χ[4/R,1](x)I 1/p0
2/R
p0 R−δ Z x/2
p0 R−δ Z x/2
2/R
y−(ν+1/2)
(2−p0 )p′
0
p0
0
dy!1/p′
y(ν+1/2) 2−p0
0
1−p0 dy!1/p′
.
Using that (ν + 1/2) 2−p0
1−p0
R−δ Z x/2
2/R
y(ν+1/2) 2−p0
< −1 and 4/R < x < 1, we have that
1−p0 dy!1/p′
−1(cid:17)1/p′
≤ C(cid:16)R−(ν+1/2) 2−p0
1−p0
0
0
R−δ = C
and the last inequality is true because the exponent of R is zero. Then
In the case p = 1 applying Holder inequality, then
R2 ≤ Cx
−2(ν+1)
p0
I 1/p0
p0
.
R2 ≤ sup
R>0
x−(ν+3/2+δ)χ[4/R,1](x)I1 R−δ
sup
y−(ν+1/2).
y∈[2/R,x/2]
Now, if ν + 1/2 > 0 and ν + 1/2 < δ,
χ[4/R,1](x)R−δ
sup
R>0
sup
y−(ν+1/2)
y∈[2/R,x/2]
= C sup
R>0
χ[4/R,1](x)Rν+1/2−δ ≤ Cx−ν−1/2+δ;
16
´O. CIAURRI AND L. RONCAL
and if ν + 1/2 ≤ 0,
sup
R>0
χ[4/R,1](x)R−δ
sup
y−(ν+1/2)
y∈[2/R,x/2]
= C sup
R>0
χ[4/R,1](x)R−δx−(ν+1/2) ≤ Cx−ν−1/2+δ.
In this manner
R2 ≤ Cx−2(ν+1)I1.
Therefore, collecting the estimates for R1 and R2 for p = p0 and p = 1, we have
shown that
J3 ⊆ {x ∈ (0, 1) : Cx
−2(ν+1)
p
(x)I 1/p > λ},
hence we can deduce (33) for i = 3.
For the region J4, we proceed as follows
sup
R>0
x−(ν+1/2)Z 1
≤ sup
0 g(y)χB4 (x, y)K δ
R(x, y) dy
x−(ν+1/2)χ[0,2/R](x)Z 1
R>0
4/R g(y)K δ
R(x, y) dy
+ sup
R>0
x−(ν+1/2)χ[2/R,1](x)Z 1
2x g(y)K δ
x−(ν+1/2)χ[0,2/R](x)(Rx)ν+1/2Z 1
4/R
R(x, y) dy
g(y)
Rδx − yδ+1 dy
≤ C sup
R>0
x−(ν+1/2)χ[2/R,1](x)Z 1
We first deal with S1, we use that y − x > y/2, then
+ C sup
R>0
2x
g(y)
Rδx − yδ+1 dy := S1 + S2.
S1 ≤C sup
R>0
χ[0,2/R](x)Rν+1/2−δZ 1
χ[0,2/R](x)Rν+1Z 1
4/R
4/R
g(y)
yδ+1 dy
g(y)√y
≤ C sup
R>0
dy ≤ Cx−(ν+1)Z 1
x
g(y)√y
dy.
p
Now for p = p0 or p = 1, we have that 2ν +1−p(ν +1) > −1 and Hardy's inequality
[17, Lemma 3.14, p. 196] is applied in the following estimate
0 (cid:18)Z 1
Z 1
0 S1(x)px2ν+1 dx ≤ CZ 1
g(y)√y
≤ CZ 1
y2ν+1−pν dy = CZ 1
0 (cid:12)(cid:12)(cid:12)(cid:12)
√y (cid:12)(cid:12)(cid:12)(cid:12)
S2 ≤ Cx−ν−1/2+δZ 1
Concerning S2, observe that supR>0 χ[2/R,1](x)R−δ ≤ Cxδ, thus
0 g(y)py(ν+1/2)(2−p) dy.
g(y)
yδ+1 dy.
x2ν+1−p(ν+1) dx
dy(cid:19)
g(y)
x
x
p
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
17
Since for p = p0 or p = 1 we have that 2ν + 1 − p(ν + 1/2 − δ) > −1, we can use
again Hardy's inequality to complete the required estimate. Indeed,
p
0 (cid:18)Z 1
Z 1
0 S2(x)px2ν+1 dx ≤ CZ 1
yδ+1 dy(cid:19)
g(y)
≤ CZ 1
0 (cid:12)(cid:12)(cid:12)(cid:12)
yδ+1(cid:12)(cid:12)(cid:12)(cid:12)
= CZ 1
g(y)
x
p
0 g(y)py(ν+1/2)(2−p) dy.
y2ν+1−p(ν+1/2−δ)+p dy
x2ν+1−p(ν+1/2−δ) dx
With the inequalities for S1 and S2, we can conclude (33) for i = 4.
To prove (33) for i = 2 we define, for k a nonnegative integer, the intervals
Ik = [2−k−1, 2−k],
Nk = [2−k−3, 2−k+2]
and the function gk(y) = g(y)χIk (y). By using (25) for x/2 < y < 2x, with
x ∈ (0, 1), we have the bound
K δ
R(x, y) ≤
Then
C
Rδ(x − y + 2/R)δ+1 .
Rδ(x − y + 2/R)δ+1 dy > Cλxν+1/2) .
gk(t)
∞
J2 ⊂(x ∈ (0, 1) : sup
Since at most three of these integrals are not zero for each x ∈ (0, 1)
R>0
x/2
Xk=0Z min {2x,1}
R>0Z min {2x,1}
x/2
J2 ⊂
⊂
∞
[k=0(x ∈ (0, 1) : 3 sup
[k=0nx ∈ Nk : M (gk, x) > Cλxν+1/2o
∞
where in the las step we have used that
gk(t)
Rδ(x − y + 2/R)δ+1 dy > Cλxν+1/2)
gk(t)
R>0Z min {2x,1}
x/2
sup
Rδ(x − y + 2/R)δ+1 dy ≤ CM (gk, x).
By using the estimate x ≃ 2−k for x ∈ Nk, we can check easily that
[k=1nx ∈ Nk : M (gk, x) > Cλ2−k(ν+1/2)o .
J2 ⊂
∞
Finally by using again that x ≃ 2−k for x ∈ Ik, Nk and the weak type norm
inequality for the Hardy-Littlewood maximal function we have
∞
ZJ2
x2ν+1 dx ≤ C
2−k(2ν+1)Z{x∈Nk:M(gk,x)>Cλ2−k(ν+1/2)}
dx
∞
2pk(ν+1/2)−k(2ν+1)
Xk=0
Xk=0
≤ C
λp Z 1
0 g(y)py(ν+1/2)(2−p) dy
λp
C
ZIk g(y)p dy
≤
and the proof is complete.
18
´O. CIAURRI AND L. RONCAL
5. Proof of Theorem 4
To conclude the result we have to prove (33) with g(x) = χE(x) and p = p1.
For J1 and J2 the result follows by using the steps given in the proof of Theorem
3 for the same intervals. To analyze J3 we proceed as we did for J4 in the proof of
Theorem 3. In this case we obtain that
sup
R>0
x−(ν+1/2)Z 1
0 g(y)χB3 (x, y)K δ
≤ C(cid:18)x−(ν+1)Z x
r (x, y)
g(y)√y
0 g(y)yδ dy(cid:19) .
Now taking into account that for p = p1 we have 2ν + 1 − p(ν + 1) < −1 and
2ν + 1 − p(ν + 3/2 + δ) < −1 we can apply Hardy's inequalities to obtain that
dy + x−(ν+3/2+δ)Z x
0
0 (cid:18)x−(ν+1)Z x
Z 1
0
g(y)√y
dy(cid:19)p
x2ν+1 dx ≤ CZ 1
0 g(y)py(ν+1/2)(2−p) dy
and
Z 1
0 (cid:18)x−(ν+3/2+δ)Z x
0 g(y)yδ dy(cid:19)p
x2ν+1 dx ≤ CZ 1
0 g(y)py(ν+1/2)(2−p) dy,
with these two inequalities we can deduce that (33) holds for J3 with p = p1 in this
case.
The main difference with the previous proof appears in the analysis of J4. To
deal with this case, we have to use the following lemma [3, Lemma 16.5]
Lemma 11. If 1 < p < ∞, a > −1, and E ⊂ [0,∞), then
(cid:18)ZE
xa dx(cid:19)p
≤ 2p(a + 1)1−pZE
x(a+1)p−1 dx.
In this case, it is enough to prove that
ZJ
dµν (x) ≤
where
C
λp Z 1
0
χE(y) dµν (y),
J =(cid:26)x ∈ (0, 1) : sup
R>0
x−(ν+1/2)Z 1
0
χE(y)χB4 (x, y)yν+1/2K δ
R(x, y) dy > λ(cid:27) ,
and this can be deduced immediately by using the inclusion
(35)
with
J ⊆ [0, min{1, H}]
H 2(ν+1) =
C
λp Z 1
0
χE(y) dµν(y).
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
19
Let's prove (35). By using (16) and the estimate y − x > y/2, we have
sup
R>0
x−(ν+1/2)Z 1
≤ C sup
R>0
0
χE(y)χB4 (x, y)yν+1/2K δ
R−δ+ν+1/2χ[0,2/R](x)Z 1
4/R
R(x, y) dy
χE(y)y−δ+ν−1/2 dy
+ C sup
R>0
R−δx−(ν+1/2)χ[2/R,1](x)Z 1
2x
χE(y)y−δ+ν−1/2 dy.
In the first summand we can use that R−δ+ν+1/2 ≤ Cxδ−ν−1/2 and in the second
one that R−δ ≤ xδ. Moreover observing that with p = p1 it holds −δ + ν + 1/2 =
2(ν + 1)/p we obtain that
sup
R>0
x−(ν+1/2)Z 1
0
χE(y)χB4 yν+1/2K δ
R(x, y) dy ≤ Cx−2(ν+1)/pZE
≤ Cx−2(ν+1)/pZE
y−1+2(ν+1)/p dy
dµν (y),
where in the last step we have used Lemma 11, and this is enough to deduce the
inclusion in (35).
6. Proofs of Theorem 6 and Theorem 7
This section will be devoted to the proofs of Theorem 6 and Theorem 7. To this
end we need a suitable identity for the kernel and in order to do that we have to
introduce some notation. H (1)
ν will denote the Hankel function of the first kind,
and it is defined as follows
where Yν denotes the Weber's function, given by
H (1)
ν (z) = Jν (z) + iYν(z),
Yν (z) =
Jν (z) cos νπ − J−ν(z)
sin νπ
, ν /∈ Z, and Yn(z) = lim
ν→n
Jν (z) cos νπ − J−ν (z)
sin νπ
.
From these definitions, we have
J−ν(z) − e−νπiJν (z)
H (1)
ν (z) =
i sin νπ
, ν /∈ Z, and H (1)
n (z) = lim
ν→n
J−ν(z) − e−νπiJν(z)
i sin νπ
.
For the function H (1)
ν , the asymptotic
(36) H (1)
ν (z) =r 2
πz
ei(z−νπ/2−π/4)[A + O(z−1)],
z > 1, −π < arg(z) < 2π,
holds for some constant A.
In [4, Lemma 1] the following lemma was proved
Lemma 12. For R > 0 the following holds:
K δ
R(x, y) = I δ
R,1(x, y) + I δ
R,2(x, y)
with
I δ
R,1(x, y) = (xy)1/2Z R
0
z(cid:18)1 −
z2
R2(cid:19)δ
Jν(zx)Jν (zy) dz
20
and
´O. CIAURRI AND L. RONCAL
I δ
R,2(x, y) = lim
ε→0
(xy)1/2
2
ZSε(cid:18)1 −
z2
R2(cid:19)δ zH (1)
Jν(z)
ν (z)Jν(zx)Jν (zy)
dz,
where, for each ε > 0, Sε is the path of integration given by the interval R + i[ε,∞)
in the direction of increasing imaginary part and the interval −R + i[ε,∞) in the
opposite direction.
Then, by Lemma 12 we have
Kδ
R(x, y) = I δ
R,1(x, y) + I δ
R,2(x, y)
where I δ
negative results will be the following lemma
R,j(x, y) = (xy)−(ν+1/2)I δ
R,j(x, y) for j = 1, 2. The main tool to deduce our
Lemma 13. For ν > −1/2, δ > 0, and R > 0 it is verified that
(yR)ν+δ+1 + I δ
R2(ν+1) Jν+δ+1(yR)
Kδ
R(0, y) =
2δ−νΓ(δ + 1)
Γ(ν + 1)
R,2(0, y),
where
(37)
Proof. From (9), it is clear that
R,2(0, y)(cid:12)(cid:12) ≤ C(R2ν−δ+1,
(cid:12)(cid:12)I δ
2νΓ(ν + 1)Z R
y−ν
0
I δ
R,1(0, y) =
Rν−δ+1/2y−(ν+1/2),
yR ≤ 1,
yR > 1.
zν+1(cid:18)1 −
z2
R2(cid:19)δ
Jν(zy) dz.
Now, by using Sonine's identity [20, Ch. 12, 12.11, p. 373]
Z 1
0
sν+1(cid:0)1 − s2(cid:1)δ
Jν (sy) ds = 2δΓ(δ + 1)
Jν+δ+1(y)
yδ+1
,
ν, δ > −1,
we deduce the leading term of the expression for Kδ
R(0, y).
To control the term
I δ
R,2(0, y) = lim
ε→0
y−(ν+1/2)
2
ZSε(cid:18)1 −
z2
R2(cid:19)δ zν+1/2H (1)
Jν(z)
ν (z)(zy)1/2Jν(zy)
dz,
we start by using the asymptotic expansions given in (36) and (10) for H (1)
ν (z)
and Jν(z). We see that on Sε, the path of integration described in Lemma 12, for
t = Im(z) the estimate
holds for t > 0. Now, from (9) and (10), it is clear that for z = ±R + it
≤ Ce−2t,
(cid:12)(cid:12)(cid:12)(cid:12)
Hν (z)
Jν (z)(cid:12)(cid:12)(cid:12)(cid:12)
√zyJν (zy) ≤ CeytΦν((R + t)y)
where Φν is the function in (26). Then
R,2(0, y) ≤ Cy−(ν+1/2)R−2δZ ∞
I δ
0
tδ(R + t)ν+δ+1/2Φν((R + t)y)e−(2−y)t dt.
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
21
If y > 1/R we have the inequality Φν((R + t)y) ≤ C, then
R,2(0, y) ≤ Cy−(ν+1/2)R−2δZ ∞
I δ
0
tδ(R + t)ν+δ+1/2e−(2−y)t dt
≤ Cy−(ν+1/2)R−δ(Rν+1/2 + R−δ) ≤ CRν−δ+1/2y−(ν+1/2)
and (37) follows in this case. If y ≤ 1/R we obtain the bound in (37) with the
estimate Φν((R + t)y) ≤ C(Φν (yR) + (yt)ν+1/2). Indeed,
R,2(0, y) ≤ Cy−(ν+1/2)R−2δΦν(yR)Z ∞
I δ
0
tδ(R + t)ν+δ+1/2e−(2−y)t dt
tν+δ+1/2(R + t)ν+δ+1/2e−(2−y)t dt
+ CR−2δZ ∞
≤ C(R2ν−δ+1 + Rν−2δ+1/2 + Rν−δ+1/2 + R−2δ) ≤ R2ν−δ+1.
0
Lemma 14. For ν > −1/2 and 0 < δ ≤ ν + 1/2, the estimate
kKδ
R(0, y)kLp0 ((0,1),dµν ) ≥ CRν−δ+1/2(log R)1/p0
holds.
Proof. We will use the decomposition in Lemma 13. By using (9) and (10) as was
done in [5, Lemma 2.1] we obtain that
(cid:3)
With the bound (37) it can be deduced that
(cid:13)(cid:13)(cid:13)(cid:13)
R2(ν+1) Jν+δ+1(yR)
(yR)ν+δ+1 (cid:13)(cid:13)(cid:13)(cid:13)Lp0 ((0,1),dµν ) ≥ CRν−δ+1/2(log R)1/p0 .
R,2(0, y)(cid:13)(cid:13)Lp0 ((0,1),dµν ) ≤ CRν−δ+1/2.
(cid:13)(cid:13)I δ
With the previous estimates the proof is completed.
(cid:3)
Finally, the last element that we need to prove Theorems 6 and 7 is the norm
inequality for finite linear combinations of the functions {ψj}j≥1 contained in the
next lemma. Its proof is long and technical and it will be done in the last section.
Lemma 15. For ν > −1/2, R > 0, 1 < p < ∞ and f a linear combination of
the functions {ψj}1≤j≤N (R) with N (R) a positive integer such that N (R) ≃ R, the
inequality
kfkL∞((0,1),dµν ) ≤ CR2(ν+1)/pkfkLp,∞((0,1),dµν )
holds.
Proof of Theorem 6. With the bound in Lemma 14 we have
(log R)1/p0 ≤ CR−2(ν+1)/p1(cid:13)(cid:13)Kδ
= CR−2(ν+1)/p1
= CR−2(ν+1)/p1
sup
R(0, y)(cid:13)(cid:13)Lp0 ((0,1),dµν )
Z 1
kf kLp1 ((0,1),dµν )=1(cid:12)(cid:12)(cid:12)(cid:12)
R(0, y)f (y) dµν(cid:12)(cid:12)(cid:12)(cid:12)
0 Kδ
kf kLp1 ((0,1),dµν )=1(cid:12)(cid:12)Bδ
Rf (0)(cid:12)(cid:12) .
sup
22
´O. CIAURRI AND L. RONCAL
From the previous estimate the result for δ = ν +1/2 follows. In the case δ < ν +1/2
it is obtained by using Lemma 15 because
R−2(ν+1)/p1
sup
kf kLp1 ((0,1),dµν )=1(cid:12)(cid:12)Bδ
Rf (0)(cid:12)(cid:12)≤ C
since Bδ
R.
Rf (x) is a linear combination of the functions {ψj}1≤j≤N (R) with N (R) ≃
(cid:3)
sup
kf kLp1 ((0,1),dµν )=1(cid:13)(cid:13)Bδ
Rf (x)(cid:13)(cid:13)Lp1,∞((0,1),dµν )
Proof of Theorem 7. In the case δ < ν + 1/2, the result follows from Theorem 6 by
using a duality argument. Indeed, it is clear that
sup
E⊂(0,1)
kBδ
RχEkLp0 ((0,1),dµν )
kχEkLp0 ((0,1),dµν )
= sup
E⊂(0,1)
(38)
=
sup
sup
kf kLp1 ((0,1),dµν )=1(cid:12)(cid:12)(cid:12)R 1
E⊂(0,1)(cid:12)(cid:12)(cid:12)R 1
0 f (y)Bδ
kχEkLp0 ((0,1),dµν )
0 χE(y)Bδ
kχEkLp0 ((0,1),dµν )
RχE(y) dµν(cid:12)(cid:12)(cid:12)
Rf (y) dµν(cid:12)(cid:12)(cid:12)
sup
.
kf kLp1 ((0,1),dµν )=1
By Theorem 6 it is possible to choose a function g such that kgkLp1((0,1),dµν ) = 1
and
kBδ
Then, with the notation
Rg(x)kLp1,∞((0,1),dµν ) ≥ C(log R)1/p0 .
we have
(39)
µν(E) =ZE
dµν ,
λp1 µν(A) ≥ C(log R)p1/p0 ,
for some positive λ and A = {x ∈ (0, 1) : Bδ
subsets of A
A1 = {x ∈ (0, 1) : Bδ
and we define D = A1 if µν (A1) ≥ µν (A)/2 and D = A2 otherwise. Then, by (39),
we deduce that
Rg(x) > λ}. Now, we consider the
A2 = {x ∈ (0, 1) : Bδ
Rg(x) < −λ}
Rg(x) > λ}
and
(40)
λ ≥ C
(log R)1/p0
µν (D)1/p1
.
Taking f = g and E = D in (38) and using (40), we see that
sup
E⊂(0,1)
kBδ
RχEkLp0 ((0,1),dµν )
kχEkLp0 ((0,1),dµν ) ≥ Cλ
µν(D)
kχDkLp0 ((0,1),dµν ) ≥ C(log R)1/p0
and the proof is complete in this case. For δ = ν + 1/2 the result follows from
Theorem 6 with a standard duality argument.
(cid:3)
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
23
7. Proof of Lemma 15
To proceed with the proof of Lemma 15 we need some auxiliary results that are
included in this section.
We start by defining a new operator. For each non-negative integer r, we consider
the vector of coefficients α = (α1, . . . , αr+1) and we define
r+1
Tr,R,αf (x) =
αℓBr
ℓRf (x).
Xℓ=1
This new operator is an analogous of the generalized delayed means considered in
[16]. In [16] the operator is defined in terms of the Ces`aro means instead of the
Bochner-Riesz means. The properties of Tr,R,α that we need are summarized in the
next lemma
Lemma 16. For each non-negative integer r and ν ≥ −1/2, the following state-
ments hold
a) Tr,R,αf is a linear combination of the functions {ψj}1≤j≤N ((r+1)R), where
N ((r + 1)R) is a non-negative integer such that N ((r + 1)R) ≃ (r + 1)R;
b) there exists a vector of coefficients α, verifying that αℓ ≤ A, for ℓ =
1, . . . , r + 1, with A independent of R and such that Tr,R,αf (x) = f (x) for
each linear combination of the functions {ψj}1≤j≤N (R) where N (R) is a
positive integer. Moreover, in this case, for r > ν + 1/2,
kT fr,R,αkL1((0,1),dµν ) ≤ CkfkL1((0,1),dµν )
and
kTr,R,αfkL∞((0,1),dµν ) ≤ CkfkL∞((0,1),dµν ),
with C independent of R and f .
Proof. Part a) is a consequence of the definition of Tr,R,α and the fact that the
m-th zero of the Bessel function Jν, with ν ≥ −1/2, is contained in the interval
(mπ + νπ/2 + π/2, mπ + νπ/2 + 3π/4).
To prove b) we consider f (x) =PN (R)
αℓ(cid:18)1 −
coefficients such that Tr,R,αf (x) = f (x) the equations
j=1 ajψj(x). In order to obtain the vector of
= 1,
s2
k
(ℓR)2(cid:19)r
r+1
for all k = 1, . . . , N (R), should be verified. After some elementary manipulations
each one of the previous equations can be written as
Xℓ=1
r
Xj=0
s2j
k (cid:18)r
j(cid:19) (−1)j
R2j
r+1
Xℓ=1
αℓ
ℓ2j = 1
and this can be considered as a polynomial in s2
we have the system of equations
k which must be equal 1, therefore
αℓ
ℓ2j = δj,0,
j = 0, . . . , r.
r+1
Xℓ=1
This system has an unique solution because the determinant of the matrix of coef-
ficients is a Vandermonde's one. Of course for each ℓ = 1, . . . , r + 1, it is verified
that αℓ ≤ A, with A a constant depending on r but not on N (R).
24
´O. CIAURRI AND L. RONCAL
The norm estimates are consequence of the uniform boundedness
kBδ
RfkLp((0,1),dµν ) ≤ CkfkLp((0,1),dµν ),
for p = 1 and p = ∞ when δ > ν + 1/2 (see [4]).
(cid:3)
In the next lemma we will control the L∞-norm of a finite linear combination of
the functions {ψj}j≥1 by its L1-norm.
Lemma 17. If ν > −1/2 and f (x) is a linear combination of the functions
{ψj}1≤j≤N (R) with N (R) a positive integer such that N (R) ≃ R, the inequality
kfkL∞((0,1),dµν ) ≤ CR2(ν+1)kfkL1((0,1),dµν )
holds.
Proof. It is clear that
N (R)
f (x) =
ψj(x)Z 1
0
f (y)ψj(y) dµν(y).
Xj=1
Now, using Holder inequality and Lemma 1 we have
kfkL∞((0,1),dµν ) ≤ C
N (R)
Xj=1
kψjk2
L∞((0,1),dµν )kfkL1((0,1),dµν )
≤ CkfkL1((0,1),dµν )
N (R)
Xj=1
j2ν+1 ≤ CR2(ν+1)kfkL1((0,1),dµν ).
(cid:3)
The following lemma is a version in the space ((0, 1), dµν) of Lemma 19.1 in [3].
The proof can be done in the same way, with the appropriate changes, so we omit
it.
Lemma 18. Let ν > −1, 1 < p < ∞ and T be a linear operator defined for
functions in L1((0, 1), dµν) and such that
kT fkL∞((0,1),dµν ) ≤ AkfkL1((0,1),dµν ) and kT fkL∞((0,1),dµν ) ≤ BkfkL∞((0,1),dµν ),
then
kT fkL∞((0,1),dµν ) ≤ CA1/pB1/p′
kfkLp,∞((0,1),dµν ).
Now, we are prepared to conclude the proof of Lemma 15.
Proof of Lemma 15. We consider the operator Tr,R,αf given in Lemma 16 b) with
r > ν + 1/2. By Lemma 16 and Lemma 17 we have
kTr,R,αfkL∞((0,1),dµν ) ≤ C((r + 1)R)2(ν+1)kTr,R,αfkL1((0,1),dµν )
≤ CR2(ν+1)kfkL1((0,1),dµν ).
From b) in Lemma 16 we obtain the estimate
So, by using Lemma 18, we obtain the inequality
kTr,R,αfkL∞((0,1),dµν ) ≤ CkfkL∞((0,1),dµν ).
kTr,R,αfkL∞((0,1),dµν ) ≤ CR2(ν+1)/pkfkLp,∞((0,1),dµν )
THE BOCHNER-RIESZ MEANS FOR FOURIER-BESSEL EXPANSIONS
25
for any f ∈ L1((0, 1), dµν). Now, since Tr,R,αf (x) = f (x) for a linear combination
of the functions {ψj}1≤j≤N (R), the proof is complete.
(cid:3)
References
[1] P. Balodis and A. C´ordoba, The convergence of multidimensional Fourier-Bessel series, J. Anal.
Math. 77 (1999), 269 -- 286.
[2] L. Carleson and P. Sjolin, Oscillatory integrals and a multiplier problem for the disc, Studia
Math. 44 (1972), 287 -- 299.
[3] S. Chanillo and B. Muckenhoupt, Weak type estimates for Ces`aro sums of Jacobi polynomial
series, Mem. Amer. Math. Soc. 102 (1993).
[4] ´O. Ciaurri and L. Roncal, The Bochner-Riesz means for Fourier-Bessel expansions, J. Funct.
Anal. 228 (2005), 89 -- 113.
[5] ´O. Ciaurri and L. Roncal, The wave equation for the Bessel Laplacian, preprint 2012.
[6] ´O. Ciaurri and K. Stempak, Conjugacy for Fourier-Bessel expansions, Studia Math. 176
(2006), 215 -- 247.
[7] ´O. Ciaurri and J. L. Varona, Two-weight norm inequalities for the Ces`aro means of generalized
Hermite expansions, J. Comput. Appl. Math. 178 (2005), 99 -- 110.
[8] J. Duoandikoetxea, "Fourier Analysis," Graduate Studies in Mathematics, 29, American
Mathematical Society, Providence, RI, 2001.
[9] J. E. Gilbert, Maximal theorems for some orthogonal series. I., Trans. Amer. Math. Soc. 145
(1969), 495 -- 515.
[10] 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.
[11] G. H. Hardy and M. Riesz, "A General Theory of Dirichlet Series," Cambridge Univ. Press,
Cambridge, 1915.
[12] H. Hochstadt, The mean convergence of Fourier-Bessel series, SIAM Rev. 9 (1967), 211 -- 218.
[13] C. Meaney, Divergent Ces`aro and Riesz means of Jacobi and Laguerre expansions, Proc.
Amer. Math. Soc. 131 (2003), 3123 -- 3128.
[14] B. Muckenhoupt and D. W. Webb, Two-weight norm inequalities for Ces`aro means of La-
guerre expansions, Trans. Amer. Math. Soc. 353 (2000), 1119 -- 1149.
[15] B. Muckenhoupt and D. W. Webb, Two-weight norm inequalities for the Ces`aro means of
Hermite expansions, Trans. Amer. Math. Soc. 354 (2002), 4525 -- 4537.
[16] E. Stein, Interpolation in polynomial classes and Markoff's inequality, Duke Math. J. 24
(1957), 467-476.
[17] E. Stein and G. Weiss, "Introduction to Fourier aAnalysis on Euclidean Spaces," Princeton
Univ. Press, Princeton, N. J. , 1970.
[18] K. Stempak, Almost everywhere summability of Laguerre series II, Stud. Math. 103 (1992),
317-327.
[19] K. Stempak, On convergence and divergence of Fourier-Bessel series, Elect. Trans. Num.
Anal. 14 (2002), 223-235.
[20] G. N. Watson, "A Treatise on the Theory of Bessel Functions," Cambridge Univ. Press,
1966.
[21] A. Zygmund, Trigonometric series. Vol. I, II, Reprinting of the 1968 version, Cambridge
Univ. Press, 1977.
Departamento de Matem´aticas y Computaci´on, Universidad de La Rioja, 26004 Logrono,
Spain
E-mail address: [email protected], [email protected]
|
1201.6584 | 1 | 1201 | 2012-01-31T15:56:10 | Polyhedron under Linear Transformations | [
"math.FA"
] | The image and the inverse image of a polyhedron under a linear transformation are polyhedrons. | math.FA | math |
Polyhedron under Linear Transformations
Zhang Zaikun†
May 6, 2008
Abstract
The image and the inverse image of a polyhedron under a linear transformation
are polyhedrons.
Keywords: polyhedron, linear transformation, Sard quotient theorem.
1
Introduction
All the linear spaces discussed here are real.
Definition 1.1.
i.) Suppose that X is a linear space, a subset P of X is said to be a polyhedron if it has
the form
P = {x ∈ X ; fk (x) ≤ λi},
k=1 ⊂ X ′ , and {λk }n
where n is a positive integer, {fk }n
k=1 ⊂ R.
If λk = 0 (k = 1, 2, 3, ..., n), then P is said to be a polyhedral cone.
ii.) Suppose that X is a TVS, a subset P of X is said to be a closed polyhedron if it has
the form
P = {x ∈ X ; fk (x) ≤ λi},
k=1 ⊂ X ∗ , and {λk }n
where n is a positive integer, {fk }n
k=1 ⊂ R.
If λk = 0 (k = 1, 2, 3, ..., n), then P is said to be a closed polyhedral cone.
It is obvious that both ∅ and X itself are (closed) polyhedral cones.
2 Main Results
Our main results are as follows.
Theorem 2.1. Suppose that X and Y are linear spaces, and T : X → Y is a linear
operator.
i.) If A ⊂ X is a polyhedron (polyhedral cone) and T is surjective, then T (A) is a
† Institute of Computational Mathematics and Scientific/Engineering Computing, Chinese Academy
of Sciences, Beijing 100190, CHINA.
1
polyhedron (polyhedral cone).
ii.) If B ⊂ Y is a polyhedron (polyhedral cone), then T −1 (B ) is a polyhedron (polyhedral
cone).
Theorem 2.2. Suppose that X and Y are Fr´echet spaces, and T : X → Y is a bounded
linear operator.
i.) If A ⊂ X is a closed polyhedron (closed polyhedral cone) and T is surjective, then
T (A) is a closed polyhedron (closed polyhedral cone).
ii.) If B ⊂ Y is a closed polyhedron (closed polyhedral cone), then T −1 (B ) is a polyhe-
dron (closed polyhedral cone).
The conclusions above will be verified in section 4.
3 A Lemma
The following conclusion is significant in our proof.
Lemma 3.1 (Sard Quotient Theorem).
i.) Suppose that X , Y and Z are linear spaces, and S : X → Y , T : X → Z are linear
operators with S surjective. If ker S ⊂ ker T , then there exists a uniquely specified linear
operator R : Y → Z , such that T = RS .
ii.) Suppose that X , Y and Z are TVS’, and S : X → Y , T : X → Z are bounded linear
operators with S surjective. If X and Y are F r ´echet spaces and ker S ⊂ ker T , then there
exists a uniquely specified bounded linear operator R : Y → Z , such that T = RS .
Proof. We will prove only ii.).
Define
and
S :X/ ker S → Y
[x] 7→ S x,
T :X/ ker S → Z
[x] 7→ T x.
Then both S and T are well defined (note that ker S ⊂ ker T ) and bounded. Besides, S is
bijective and S−1 is bounded, since X/ ker S and Y are both Fr´echet spaces. Now define
R = T S−1 ,
then it is easy to show that R satisfies the requirements.
The uniqueness of R is trivial.
■
2
4 Proofs of Main Results
We will prove only theorem 2.2, because the proof of theorem 2.1 is similar. Only the
polyhedron case will be discussed.
Proof of Theorem 2.2.
i.) Suppose that
A =
n
\
{x ∈ X ; fk (x) ≤ λk },
k=1
k=1 ⊂ X ∗ , and {λk }n
where n is a positive integer, {fk }n
k=1 ⊂ R. The proof will be
presented in four steps.
Step 1. We will prove that the conclusion holds if
ker T ⊂
n
\
k=1
ker fk .
..., n}, we can choose a functional gk ∈ Y ∗ such
In this case, for any k ∈ {1, 2, 3,
that fk = gk T (Sard quotient theorem). It can be shown without difficulty that
T (A) =
n
\
{y ∈ Y ; gk (y) ≤ λk }.
k=1
Step 2. We will prove that the conclusion holds if dim(ker T ) = 1. This is the most
critical part of the proof.
Suppose that ξ is a point in ker T \ {0}. Let
K+ = {k ; 1 ≤ k ≤ n and fk (ξ ) > 0},
K− = {k ; 1 ≤ k ≤ n and fk (ξ ) < 0},
K0 = {k ; 1 ≤ k ≤ n and fk (ξ ) = 0}.
For any i ∈ K+ and j ∈ K− , define
Then define
hij = fi −
fi (ξ )
fj (ξ )
fj .
{x ∈ X ; hij (x) ≤ λi −
fi(ξ )
fj (ξ )
λj },
{x ∈ X ; fk (x) ≤ λk }.
A1 = \
i∈K+ ,
j∈K−
A2 = \
k∈K0
If K+ = ∅ or K− = ∅, we take A1 as X . Similarly, if K0 = ∅, we take A2 as X . We will
prove that T (A) = T (A1 ∩ A2 ). It suffices to show that T (A1 ∩ A2 ) ⊂ T (A).
3
• If K+ = ∅ = K− , nothing needs considering.
• If K+ 6= ∅ = K− , fix a point x ∈ A1 ∩ A2 , define
s = min
i∈K+
λi − fj (x)
fi(ξ )
,
then it is easy to show that x + sξ ∈ A and T (x + sξ ) = T x. The case with K− 6=
∅ = K+ is similar.
• If K+ 6= ∅ 6= K− , fix a point x ∈ A1 ∩ A2 , define
t = max
j∈K−
λj − fj (x)
fj (ξ )
and consider x + tξ . It is obvious that
T (x + tξ ) = y
fj (x + tξ ) ≤ λj , ∀j ∈ K− ∪ K0 .
t =
λj0 − fj0 (x)
fj0 (ξ )
(j0 ∈ K− ),
and that
Suppose
then for any i ∈ K+ ,
fi (x + tξ )
=hij0 (x + tξ ) +
fj0 (x + tξ )
fi (ξ )
fj0 (ξ )
fi (ξ )
fj0 (ξ )
fi (ξ )
fj0 (ξ )
≤λi −
=λi .
Thus x + tξ ∈ A.
λj0 +
λj0
It has been shown that T (A1 ∩ A2 ) ⊂ T (A), and consequently T (A1 ∩ A2 ) = T (A). Ac-
cording to Step 1, the conclusion holds under the assumption dim(ker T ) = 1.
Step 3. We will prove by induction that the conclusion holds if dim(ker T ) is finite.
If dim(ker T ) = 0, then T is an isomorphism as well as a homeomorphism (inverse
mapping theorem), thus nothing needs proving. Now suppose that the conclusion holds
when dim(ker T ) ≤ n (n ≥ 0). To prove the case with dim(ker T ) = n + 1, choose a
4
point η in ker T \ {0}, find a functional F ∈ X ∗ such that F (η) = 1 (Hahn-Banach
theorem), and define
T :X → Y × R
x 7→ (T x, F (x)),
π :Y × R → Y
(y , λ) 7→ y .
Then we have
• T = π T ;
• dim(ker T ) = n;
• dim(ker π) = 1;
• both T and π are surjective bounded linear operators.
Thus by the induction hypothesis and the conclusion of Step 2, T (A) is a closed poly-
hedron.
Step 4. Now consider the general case.
Let
n
\
M = (
k=1
ker fk ) \(ker T ),
then M is a closed linear subspace of M , and therefore X/M is a Fr´echet space. Define
T :X/M → Y
[x] 7→ T x,
fk :X/M → R
[x] 7→ fk (x)
where k = 1, 2, 3, ..., n. Then T and fk are well defined, T is a bounded linear operator
from X/M onto Y , and { fk }n
k=1 ⊂ (X/M )∗ . Besides, we have
which implies that
Now let
n
\
(
k=1
ker fk ) \(ker T ) = {0},
dim(ker T ) ≤ n.
A =
n
{[x] ∈ X/M ; fk ([x]) ≤ λk },
\
k=1
5
then
T (A) = T ( A).
From what has been proved, it is easy to show that T (A) is a closed polyhedron.
Proof of part i.) has been completed.
ii.) This part is much easier. Suppose that
B =
m
\
{y ∈ Y ; gk (y) ≤ µk },
k=1
k=1 ⊂ Y ∗ , and {µk }m
where m is a positive integer, {gk }m
k=1 ⊂ R. One can show without
difficulty that
T −1 (B ) =
m
\
{x ∈ X ; gk (T x) ≤ µk },
k=1
which is a closed polyhedron in X .
■
5 Remarks
For part i) of theorem 2.2, the completeness conditions are essential. This can be seen
from the following examples.
Example 5.1. Suppose that (Y , k · kY ) is an infinite dimensional Banach space, and f is
an unbounded linear functional on it1 . Let X has the same elements and linear structure
as Y , but the norm on X is defined by
kxkX = kxkY + f (x).
It is clear that the identify mapping I : X → Y is linear, bounded and bijective. Now
consider ker f . It is a closed polyhedral cone in X , while its image under I is not closed
in Y .
Example 5.2. Suppose that X is ℓ1 . Let Y has the same elements and linear structure
as X , but the norm on Y is defined by
k(xk )k = sup
k≥1
xk .
Then f : (xk ) 7→ P xk is a bounded linear functional on X , while it is unbounded on
Y . Now consider the identify mapping again.
The preceding examples also imply that inverse mapping theorem and Sard quotient
theorem do not hold without completeness conditions.
1For a locally bounded TVS Y , there exist unbounded linear functionals on Y provided dim Y =
∞. One of them can be constructed as follows: Let U be a bounded neighborhood of 0, and {ek ; k ≥
1} ⊂ U be a sequence of linearly independent elements in Y . Let M = Span {ek ; k ≥ 1}, and define
g : M → R, P αk ek 7→ P kαk . Then extend g to Y .
6
|
1711.01659 | 1 | 1711 | 2017-11-05T20:46:32 | Besov classes on finite- and infinite-dimensional spaces and embedding theorems | [
"math.FA"
] | We give a new description of classical Besov spaces in terms of a new modulus of continuity. Then a similar approach is used to introduce Besov classes on an infinite-dimensional space endowed with a Gaussian measure. | math.FA | math |
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
AND EMBEDDING THEOREMS
EGOR D. KOSOV
Abstract. We give a new description of classical Besov spaces in terms of a new modulus of
continuity. Then a similar approach is used to introduce Besov classes on an infinite-dimensional
space endowed with a Gaussian measure.
Keywords: Besov class, fractional Sobolev class, Ornstein–Uhlenbeck semigroup, embedding
theorem
MSC: primary 46E35, secondary 28C20, 46G12
1. Introduction
In this work we continue the study of Nikolskii–Besov classes started in [7], where an equiva-
lent description of these classes was presented characterizing the inclusion of a function to the
Nikolskii–Besov class in terms of action on test functions in the spirit of the classical defini-
tions of Sobolev classes and the class of functions of bounded variation. Namely, a function
f ∈ Lp(Rn) belongs to the Nikolskii–Besov class Bα
p,∞(Rn) with 0 < α < 1 if and only if there
is a constant C such that
divΦ(x)f (x) dx ≤ CkΦkα
q kdivΦk1−α
q
(1.1)
for each vector field Φ of class C∞0 (Rn, Rn), where q = p/(p − 1). If we take α = 1 and p = 1,
we obtain the classical definition of a function of bounded variation. This new characterization
has already found some applications in the study of the distributions of polynomials on spaces
with Gaussian (and general log-concave) measures (see [11], [6], and also [5]).
In the present paper, we give a similar equivalent characterization for general Besov spaces
p,θ(Rn) with parameters α ∈ (0, 1), p ∈ [1,∞),
p,θ(Rn). We recall that the Besov space Bα
Bα
θ ∈ [1,∞] consists of all functions f ∈ Lp(Rn) such that the quantity
ZRn
is finite, where fh(x) := f (x− h) (see [2], [15], [16], and [17]). However, for further purposes, it
is more convenient to use another equivalent definition in terms of the Lp-modulus of continuity.
Recall that the Lp-modulus of continuity of a function f ∈ Lp(Rn) is defined by the equality
(cid:18)ZRn(cid:2)h−αkfh − fkp(cid:3)θh−ndh(cid:19)1/θ
ωp(f, ε) := sup
h≤εkfh − fkp.
Note that the function ωp(f,·) is nondecreasing and subadditive, which means that
A function f ∈ Lp(Rn) belongs to the class Bα
p,θ(Rn) if and only if the quantity
ωp(f, ε1 + ε2) ≤ ωp(f, ε1) + ωp(f, ε2),
ε1, ε2 > 0.
kfkα,p,θ :=(cid:18)Z +∞
0
(cid:2)s−αωp(f, s)(cid:3)θ
s−1ds(cid:19)1/θ
The author is a Young Russian Mathematics award winner and would like to thank its sponsors and jury.
This research was supported by the Russian Science Foundation Grant 17-11-01058 at Lomonosov Moscow State
University.
1
2
EGOR D. KOSOV
is finite. We define the Besov norm of a function f by the equality
kfkBα
p,θ(Rn) := kfkp + kfkα,p,θ.
Our equivalent characterization of Besov spaces is based on a new modulus of continuity
which is equivalent to ωp(f,·) and provides the known characterization (1.1) in the case of
θ = ∞. For a function f ∈ Lp(Rn) we introduce
σp(f, ε) := supnZRn
divΦ(x)f (x)dx, Φ ∈ C∞0 (Rn, Rn),kdivΦk p
p−1 ≤ 1,kΦk p
p−1 ≤ εo.
The first main result of the present paper asserts the equivalence of ωp(f,·) and σp(f,·): for
any function f ∈ Lp(Rn), one has
2−1ωp(f, 2ε) ≤ σp(f, ε) ≤ 6n ωp(f, ε).
Actually, the function σp(f,·) has appeared implicitly in the new definition of Nikolskii–Besov
spaces formulated above, since condition (1.1) can be reformulated in the following way:
s−ασp(f, s) < ∞.
sup
s≥0
So, this is the desired modulus of continuity. The above equivalence also shows that a function
f ∈ Lp(Rn) belongs to the Besov space Bα
p,θ(Rn) if and only if
(cid:16)Z ∞
0
(cid:2)s−ασp(f, s)(cid:3)θ
s−1ds(cid:17)1/θ
< ∞.
(1.2)
To illustrate how our approach to the fractional smoothness in terms of the modulus of
continuity σp(f,·) is related to the already known results, in Section 3 we propose the new
proof of the classical Ulyanov-type embedding theorems by means of the function σp(f,·).
We recall that in his seminal works [18], [19] P.L. Ulyanov obtained the following embedding
theorem.
Theorem. For a function f ∈ L1[0, 1] set
w1(f, ε) := sup
f (t + h) − f (t) dt
0≤h≤εZ 1−h
0
Then for any nondecreasing function U : [0,∞) → [0,∞) the following implications hold:
(i)
∞Xn=1
[U(n + 1) − U(n)]w1(f, 1/n) < ∞ ⇒Z 1
n−2U(36nw1(f, 1/n)) < ∞ ⇒Z 1
∞Xn=1
0
(ii)
0 f (t)U(f (t)) dt < ∞;
U(f (t)) dt < ∞.
Actually in the same works embedding theorems into Lr-spaces were obtained, but here we
discuss only the stated results as examples of embedding theorems. The multidimensional case
was considered in papers [8], [9] and [10], where the author obtained necessary and sufficient
conditions for such type of embeddings. The main method used by P.L. Ulyanov himself and
by other researchers in subsequent investigations of such embedding theorems is based on the
so-called equimeasurable rearrangements of functions (see [10] for a discussion of the method).
However, in Section 3 we employ another approach, based on the properties of the function
σp(f,·), and obtain similar simple sufficient conditions for embeddings into the classes LU(L)
and U(L). Actually, our conditions are a kind of integral form of Ulyanov's conditions stated
above and are similar to the multidimensional results [8, Theorem 1] and [10, Corollary 4.2],
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
3
which are slightly weaker than necessary and sufficient conditions [9, Theorem 5] and [10,
Theorem 4.4]. The main idea in Section 3 is to estimate the integral
ZA f (x)pdx
ZRn
divΦ(x)f (x) dx ≤ σp(f, r)
p−1 ≤ r and kdivΦk p
of a function f ∈ Lp(Rn) over a Borel set A in terms of Lebesgue measure of this set A.
Substituting A = {f ≥ s} we can estimate the behavior of the function f on sets of large
values, which is already sufficient to prove embedding theorems we are interested in. The
definition of the function σp(f,·) states that
for smooth vector fields Φ with kΦk p
p−1 ≤ 1. Taking Φ = ∇ϕ, solving the
Poisson equation div∇ϕ = ∆ϕ = u, estimating ∇ϕ in terms of u, and taking the supremum
over functions u with kukLq(A) = 1 we obtain the necessary bound.
Finally, in Section 4, we proceed to Besov classes on locally convex spaces endowed with
centered Gaussian measures. In paper [7], Nikolskii–Besov classes on a Gaussian space were
introduced by means of relation (1.1) as the definition, where in place of the divergence operator
on Rn the Gaussian divergence operator divγ was used. If we consider the standard Gaussian
measure γn on Rn, which is the measure with density (2π)−n/2 exp(−x2/2), then
divγnΦ =
(∂iΦi − xiΦi) = divΦ − hx, Φi.
nXi=1
In this paper we propose a similar approach (see Definitions 4.1 and 4.2) to general Besov classes
Bα
p,θ(γ) with respect to a Gaussian measure γ. The first main result of Section 4 (presented in
Theorem 4.7) provides an equivalent characterization of the introduced Besov classes in terms
of "shifts" on the Gaussian space, which is similar in a sense to the classical definition of Besov
spaces on Rn. Namely, the function f ∈ Lp(γ) with p > 1 belongs to the Besov class Bα
p,θ(γ) if
and only if the quantity
(cid:16)Z ∞
0 ht−α/2(cid:16)ZZ f (e−tx + √1 − e−2ty) − f (x)p γ(dx)γ(dy)(cid:17)1/piθ
t−1dt(cid:17)1/θ
is finite. This theorem can be also viewed as an analog of Theorem 3.2 from [1]. The second
main result of this section is the embedding theorem for Gaussian Besov classes. We recall
(see for example [4] and [13]) that for an arbitrary function f from the Gaussian Sobolev space
W 2,1(γ) the following logarithmic Sobolev inequality holds:
For the Sobolev class W 1,1(γ) there is also an embedding theorem of logarithmic type. Namely,
the space W 1,1(γ) is continuously embedded into the Orlicz space L log L1/2, which is defined
by the condition
Z f 2 ln(fkfk−1
2 )dγ ≤Z ∇f2dγ.
Z f[ln(1 + f)]1/2dγ < ∞
(see [12] and [3] for the case of functions of bounded variation). Both results mean that a
smoothness of a function provides some higher order of integrability. One may wonder whether
this effect remains in force for the Besov smoothness condition introduced in the present paper.
Theorem 4.8 is aimed to answer this question. It asserts that, for any α ∈ (0, 1), β ∈ (0, α),
p ∈ (1,∞), and θ ∈ [1,∞], there is a constant C = C(p, θ, α, β) such that for all functions
4
EGOR D. KOSOV
f ∈ Bα
p,θ(γ) one has
(cid:16)Z fp(cid:12)(cid:12)ln(fkfk−1
p )(cid:12)(cid:12)pβ/2
dγ(cid:17)1/p
≤ CkfkBα
p,θ(γ).
The main idea of the proof of this result is in spirit of the semigroup approach to the isoperi-
metric inequality on the Gaussian space proposed by M. Ledoux in [14] and [12]. Similarly
to the cited works, we use the short time behavior of the Ornstein-Uhlenbeck semigroup on
functions from the Besov class and the hypercontractivity property of the Ornstein–Uhlenbeck
semigroup. At the end of the paper we provide an estimate of the best approximation of a
function from L2(γ) by Hermite polynomials in terms of the introduced Gaussian modulus of
continuity.
Throughout the paper we assume that α is a fixed number from (0, 1]. Given p ∈ [1,∞],
we denote by q the dual number such that 1/p + 1/q = 1. The Lp-norm of a function f with
respect to a measure µ is defined as usual by
kfkp := kfkLp(µ) =(cid:16)Z fp dµ(cid:17)1/p
, p ∈ [1,∞),
and the limiting case of p = ∞ is treated also as usual. In Sections 2 and 3 the measure µ will
be the standard Lebesgue measure on Rn, but in Section 4 the measure µ will be a centered
Gaussian measure on a locally convex space. We denote the space of all infinitely differential
functions with compact support on Rn by C∞0 (Rn) and the space of all bounded infinitely
differential functions with bounded derivatives of every order is denoted by C∞b (Rn).
This section is devoted to obtaining a new characterization of Besov classes on Rn in terms
2. Besov classes on Rn
of the moduli of continuity σp(f,·) andeσp(f,·).
Let · denote the standard Euclidean norm on Rn generated by the standard Euclidean
inner product h·,·i. Let λn be the standard Lebesgue measure on Rn. We also need the heat
semigroup Pt on Rn, which is defined by the equality
Ptf (x) := (2πt)−n/2ZRn
f (y) exp(cid:18)−x − y2
2t (cid:19) dy,
f ∈ L1(Rn).
We start with the following key definitions (recall that q = p/(p − 1)).
Definition 2.1. Let f ∈ Lp(Rn). Set
Definition 2.2. Let f ∈ Lp(Rn). Set
σp(f, ε) := supnZRn
eσp(f, ε) := supnZRn
divΦ(x)f (x) dx : Φ ∈ C∞0 (Rn, Rn),kdivΦkq ≤ 1,kΦkq ≤ εo.
∂eϕ(x)f (x) dx : ϕ ∈ C∞0 (Rn),k∂eϕkq ≤ 1,kϕkq ≤ εo.
We now obtain several properties of the introduced functions.
Lemma 2.3. For any function f ∈ Lp(γ), the functions σp(f,·) andeσp(f,·) are nondecreasing,
subadditive, concave and continuous on (0, +∞).
Proof. We consider only the function σp(f,·), since for the second one the proof is essentially
the same. It is readily seen that this function is indeed nondecreasing and subadditive. We now
check that it is concave. Let a, b > 0, and t ∈ (0, 1). Then for an arbitrary pair of vector fields
Φ1, Φ2 ∈ C∞0 (Rn, Rn) with kdivΦ1kq ≤ 1,kΦ1kq ≤ a and kdivΦ2kq ≤ 1,kΦ2kq ≤ b we have
tZRn
divΦ1(x)f (x) dx + (1 − t)ZRn
divΦ2(x)f (x) dx =ZRn
div[tΦ1(x) + (1 − t)Φ2(x)]f (x) dx
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
5
and kdiv[tΦ1 + (1 − t)Φ2]kq ≤ 1, ktΦ1 + (1 − t)Φ2kq ≤ ta + (1 − t)b. Thus,
tσp(f, a) + (1 − t)σp(f, b) ≤ σp(f, ta + (1 − t)b).
The concavity implies the continuity.
(cid:3)
Let σ be a concave, nondecreasing and nonnegative function on (0, +∞). Let us introduce
the "adjoint" function
σ∗(s) := sσ(s−1).
In particular, we can consider
σ∗p(f, s) := sσp(f, s−1),
eσ∗p(f, s) := seσp(f, s−1).
Lemma 2.4. Let σ be a concave, nondecreasing and nonnegative function on (0, +∞). Then
the function σ∗ is also concave and nondecreasing on (0, +∞). If, in addition, we assume that
limt→0 t−1σ(t) = ∞, then the function σ∗ is strictly monotone.
Proof. Let s, t ∈ (0, +∞) and s > t. Then 1/s = (t/s)1/t + (1 − t/s)0. Due to the concavity
of the function σ we have
σ((t/s)1/t + (1 − t/s)ε) ≥ (t/s)σ(1/t) + (1 − t/s)σ(ε) ≥ (t/s)σ(1/t).
Due to the continuity of the function σ, taking the limit as ε → 0 in the above estimate, we
have σ(1/s) ≥ (t/s)σ(1/t) implying σ∗(s) ≥ σ∗(t).
Let again s > t > 0 and let κ ∈ (0, 1). We note that
1/s +
κs
=
1
κs + (1 − κ)t
Thus, by the concavity of the function σ one has
κs + (1 − κ)t
(1 − κ)t
κs + (1 − κ)t
1/t.
σ((κs + (1 − κ)t)−1) ≥
κs
κs + (1 − κ)t
σ(1/s) +
(1 − κ)t
κs + (1 − κ)t
σ(1/t)
and σ∗(κs + (1 − κ)t) ≥ κσ∗(s) + (1 − κ)σ∗(t), i.e. σ∗ is concave.
any point r > s one has
Assume that there are two points s > t such that σ∗(s) = σ∗(t). Then, by the concavity, for
σ∗(r) = σ∗(cid:18)r − t
s − t
s +(cid:16)1 −
r − t
s − t(cid:17)t(cid:19) ≤
Thus,
which contradicts the condition limt→0 t−1σ(t) = ∞.
lim
r→∞
rσ(1/r) = lim
r→∞
r − t
s − t
σ∗(s) +(cid:16)1 −
σ∗(r) ≤ σ∗(s)
r − t
s − t(cid:17)σ∗(t) = σ∗(s).
(cid:3)
decreasing on (0, +∞). If, in addition, we assume that limt→0 t−1σp(f, t) = ∞ (alternatively,
Corollary 2.5. Let f ∈ Lp(Rn). Then the functions σ∗p(f,·) andeσ∗p(f,·) are concave and non-
limt→0 t−1eσp(f, t) = ∞), then the function σ∗p(f,·) (eσ∗p(f,·), respectively) is strictly monotone.
We now proceed to the main result of this section showing the equivalence of σp(f,·),eσp(f,·),
2−1kf2h − fkp ≤eσp(f,h) ≤ σp(f,h) ≤ (2π)−n/2ZRn kfhz − fkp(1 + z)e− z2
and ωp(f,·). We start with the following technical lemma (the proof is similar to the proof of
Theorem 3.4 from [7]).
Lemma 2.6. For any function f ∈ Lp(Rn) one has
2 dz.
6
EGOR D. KOSOV
Proof. For every function ϕ ∈ C∞0 (Rn) and for every unit vector e ∈ Rn we can take Φ = eφ
and conclude that
Let now e = h−1h. For an arbitrary function ϕ ∈ C∞0 (Rn) with kϕkq ≤ 1 we can write
eσp(f, ε) ≤ σp(f, ε).
ZRn
ϕ(x)(fh(x) − f (x)) dx =ZRn
For the function
[ϕ(x + h) − ϕ(x)]f (x) dx
=ZRnZ h
0
∂eϕ(x + se) dsf (x) dx.
we have kψkq ≤ hkϕkq ≤ h and k∂eψkq ≤ 2kϕkq ≤ 2, since
ϕ(x + se) ds ∈ C∞0 (Rn)
0
ψ(x) =Z h
∂eψ(x) =(cid:12)(cid:12)(cid:12)(cid:12)Z h
ϕ(x)(fh(x) − f (x))dx =ZRn
0
ZRn
∂eϕ(x + se) ds(cid:12)(cid:12)(cid:12)(cid:12) = ϕ(x + h) − ϕ(x).
∂eψ(x)f (x) dx ≤ 2eσp(f,h/2).
Thus,
Taking the supremum over functions ϕ, we get the estimate kfh − fkp ≤ 2eσp(f,h/2).
Finally, for every smooth vector field Φ ∈ C∞0 (Rn, Rn) we can write
ZRn
divΦ(x)f (x) dx =ZRn
divΦ(x)(f (x) − Ptf (x)) dx +ZRn
kf − Ptfkp ≤ (2π)−n/2ZRn kf√tz − fkpe− z2
2 dz.
We note that
divΦ(x)Ptf (x) dx.
(2.1)
Thus, for the first term in equality (2.1) we have
Integrating by parts in the second term of equality (2.1), we have
divΦ(x)(f (x)−Ptf (x)) dx ≤ kdivΦkqkf−Ptfkp ≤ kdivΦkq(2π)−n/2ZRn kf√tz−fkpe− z2
ZRn
divΦ(x)Ptf (x) dx = −ZRnhΦ(x),∇Ptf (x)i dx
ZRn
2 dz.
= t−1/2ZRnZRnhΦ(x), (x − y)t−1/2if (y)(2πt)−n/2e− x−y2
= t−1/2ZRnZRnhΦ(x), zif (x −
2t dy dx
√tz)(2π)−n/2e− z2
2 dz dx.
Since
Z f (x)Z hΦ(x), zie− z2
2 dz dx = 0,
the above expression is equal to
t−1/2ZRn
(2π)−n/2e− z2
2 ZRnhΦ(x), zi(f (x −
√tz) − f (x)) dx dz
≤ t−1/2kΦkq(2π)−n/2ZRn ze− z2
2 kf√tz − fkp dz
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
7
Thus, we have
ZRn
divΦ(x)f (x) dx ≤ kdivΦkq(2π)−n/2ZRn kf√tz − fkpe− z2
2 dz
Hence
and taking √t = ε we conclude that
σp(f, ε) ≤ (2π)−n/2ZRn kf√tz − fkpe− z2
σp(f, ε) ≤ (2π)−n/2ZRn
The lemma is proved.
2 dz.
+ t−1/2kΦkq(2π)−n/2ZRn zkf√tz − fkpe− z2
2 dz + t−1/2ε(2π)−n/2ZRn zkf√tz − fkpe− z2
(1 + z)kfεz − fkpe− z2
2 dz.
2 dz
(cid:3)
subadditive, in particular,
As we have already mentioned in the introduction, the function ωp(f,·) is nondecreasing and
(2.2)
ωp(f, τ s) ≤ 2τ ωp(f, s)
for τ ≥ 1 and s > 0. Indeed, let k ∈ N be a number such that k ≤ τ < k + 1. Then
ωp(f, τ s) ≤ ωp(f, (k + 1)s) ≤ (k + 1)ωp(f, s) = k(1 + 1/k)ωp(f, s) ≤ 2τ ωp(f, s).
Now we are ready to prove the aforementioned equivalence.
Theorem 2.7. For any function f ∈ Lp(Rn), we have
2−1ωp(f, 2ε) ≤eσp(f, ε) ≤ σp(f, ε) ≤ 2(1 + √n + n)ωp(f, ε).
Proof. The first two inequalities are straightforward corollaries of Lemma 2.6. For the last one,
by the same lemma, we have
σp(f, ε) ≤ (2π)−n/2ZRn kfεz − fkp(1 + z)e− z2
= (2π)−n/2Zz≤1
ωp(f, εz)(1 + z)e− z2
ωp(f, ε)(2π)−n/2Zz≤1
2 dz ≤ (2π)−n/2ZRn
2 dz + (2π)−n/2Zz>1
(1 + z)e− z2
The first integral above is estimated by
2 dz ≤ 2ωp(f, ε)
ωp(f, εz)(1 + z)e− z2
ωp(f, εz)(1 + z)e− z2
2 dz.
2 dz
by the monotonicity of the function ωp(f,·). The second integral, by estimate (2.2), is not
greater than
ωp(f, ε)(2π)−n/2Zz>1
2z(1 + z)e− z2
2 dz ≤ ωp(f, ε)(2√n + 2n).
Combining these two estimates we get the announced bound.
(cid:3)
As a corollary of the above theorem we get an equivalent characterization of Besov classes
on Rn. Let us introduce the following notation.
Definition 2.8. Let f ∈ Lp(Rn), p ∈ [1,∞), θ ∈ [1,∞], and α ∈ (0, 1). Set
V p,θ,α(f ) =(cid:16)Z ∞
0
(cid:2)s−ασp(f, s)(cid:3)θ
s−1 ds(cid:17)1/θ
.
(i) f ∈ Bα
p,θ(Rn);
(ii) V p,θ,α(f ) < ∞;
Moreover,
(iii) eV p,θ,α(f ) < ∞.
2α−1kfkα,p,θ ≤ eV p,θ,α(f ) ≤ V p,θ,α(f ) ≤ 2(1 + √n + n)kfkα,p,θ.
3. Ulyanov embedding theorems
In this section we study embedding theorems by means of the obtained properties of the
Our first goal is to estimate the measure of the set {f ≥ t}. To provide such an estimate
function σp(f,·).
we need the following lemma.
Lemma 3.1. Let f ∈ Lp(Rn), p ∈ [1, n) or n = p = 1, and let u ∈ C∞0 (Rn) be a function such
that kukq ≤ 1 (recall that q = p/(p − 1)). Then
Z u(x)f (x)dx ≤ C(n, p)σp(cid:0)f,kukp/n
1 (cid:1),
n
where C(n, p) = 1 + ν−1/p
C(1, 1) = 1.
Proof. By approximation, for an arbitrary vector field Φ ∈ C∞(Rn) with kΦkq ≤ ε, kdivΦkq ≤ 1
one has
(n/p − 1)1/p−1, νn is the surface area of the unit sphere in Rn, and
Z divΦ(x)f (x)dx ≤ σp(f, ε).
Assume first that n > 2. Consider the function
ϕ(x) = −(n − 2)−1ν−1
n ZRn x − y−n+2u(y) dy.
It is known that div∇ϕ = ∆ϕ = u. Set
K1(x) = x−n+1Ind{x<R}(x), K2(x) = x−n+1Ind{x≥R}(x).
n ZRn x − y−n+1u(y) dy = ν−1
n (cid:0)K1 ∗ u(x) + K2 ∗ u(x)(cid:1).
8
EGOR D. KOSOV
Definition 2.9. Let f ∈ Lp(Rn), p ∈ [1,∞), θ ∈ [1,∞], and α ∈ (0, 1). Set
eV p,θ,α(f ) =(cid:16)Z ∞
0
(cid:2)s−αeσp(f, s)(cid:3)θs−1 ds(cid:17)1/θ
.
Corollary 2.10. For any function f ∈ Lp(Rn), the following conditions are equivalent:
Let us estimate ∇ϕ:
Thus,
k∇ϕkq ≤ ν−1
= R + ν−1/p
n
∇ϕ(x) ≤ ν−1
n (cid:0)kK1k1kukq + kK2kqkuk1(cid:1)
n Z{x<R}
≤ ν−1
(cid:0)(n − 1)(q − 1) − 1(cid:1)−1/q
x−n+1 dx + ν−1
n kuk1(cid:18)Z{x≥R}
R1−n/pkuk1 = R + ν−1/p
n
x−nq+q dx(cid:19)1/q
q−1/q(n/p − 1)1/p−1R1−n/pkuk1
(n/p − 1)1/p−1R1−n/pkuk1
n
Setting now R = kukp/n
1
, we obtain
k∇ϕkq ≤(cid:0)1 + ν−1/p
n
≤ R + ν−1/p
(n/p − 1)1/p−1(cid:1)kukp/n
1
.
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
9
Thus,
Z u(x)f (x)dx =Z div∇ϕ(x)f (x)dx ≤ σp(cid:0)f,(cid:0)1 + ν−1/p
n
(n/p − 1)1/p−1(cid:1)kukp/n
1 (cid:1)
where we have used the monotonicity of the function σ∗p(f,·) (see Corollary 2.5). We have
obtained the announced estimate in the case n > 2. For n = 2 we can take
(n/p − 1)1/p−1(cid:1)σp(cid:0)f,kukp/n
1 (cid:1),
n
≤(cid:0)1 + ν−1/p
ϕ(x) = −(2π)−1Z lnx − yu(y)dy
and argue as above.
Thus, only the case n = 1 remains. In that case we consider the function ϕ(x) =R x
For this function we can write
Z u(x)f (x)dx =Z ϕ′(x)f (x)dx ≤ σ1(f,kϕk∞) ≤ σ1(f,kuk1).
−∞
u(t)dt.
(cid:3)
The lemma is proved.
Corollary 3.2. Let f ∈ Lp(Rn), p ∈ [1, n) or n = p = 1, then
(cid:18)ZA f (x)p dx(cid:19)1/p
≤ C(n, p)σp(cid:0)f,(cid:0)λn(A)(cid:1)1/n(cid:1)
for an arbitrary Borel set A in Rn with C(n, p) = 1 + ν−1/p
surface area of the unit sphere in Rn, and C(1, 1) = 1.
Proof. Assume first that A is a bounded set, A ⊂ B(0, R), where B(0, R) is the ball of radius
R centered at the origin. Consider a function u ∈ Lq(A) with kukLq(A) ≤ 1. There is a
sequence of functions um ∈ C∞0 (Rn) such that supp(um) ⊂ B(0, 2R), kumkq ≤ 1, and um → u
can be constructed by means of convolutions with compactly supported smooth probability
densities. For each function um by the previous lemma we have
in Lq(cid:0)B(0, 2R)(cid:1) (λn-a.e. in case p = 1), where u(x) = 0 if x 6∈ A. For example, such a sequence
(n/p − 1)1/p−1, where νn is the
n
Z um(x)f (x) dx ≤ C(n, p)σp(cid:0)f,kumkp/n
1 (cid:1).
Since k · k1 is a continuous function on the space Lq(cid:0)B(0, 2R)(cid:1) for p > 1 (or by Lebesgue's
dominated convergence theorem in case p = 1), the above estimate is also valid for the function
u. Thus, for every function u ∈ Lq(A) with kukLq(A) ≤ 1 we have
ZA
u(x)f (x) dx ≤ C(n, p)σp(cid:0)f,kukp/n
1 (cid:1) ≤ C(n, p)σp(cid:0)f,(cid:0)λn(A)(cid:1)1/n(cid:1).
Taking the supremum over all functions u with kukLq(A) ≤ 1 we obtain the desired estimate for
bounded sets A. The case of an arbitrary set A can be obtained by passing to the limit.
(cid:3)
We now proceed to embedding theorems, which we formulate in terms of the function σp(f,·)
instead of ωp(f,·), since these functions are equivalent by Theorem 2.7.
Theorem 3.3. Let f ∈ Lp(Rn), where p ∈ [1, n) or n = p = 1. Let U : [0,∞) → [0,∞) be a
nondecreasing continuous function. Assume that there is a number N > 0 such that
where the integral is understood in the Lebesgue–Stieltjes sense. Then
N
Z +∞
(cid:2)σp(cid:0)f, t−p/n(cid:1)(cid:3)p
ZRn f (t)pU(f (t)kfk−1
dU(t) < ∞,
p ) dt < ∞.
since λn(As) ≤ s−p. Integrating both sides of the above estimate from N to +∞ with respect
to the locally bounded measure, generated by the monotone function U, we get
We now deal with the left-hand side of the above estimate, which is equal to
≤ C(n, p)p(cid:2)σp(cid:0)f, s−p/n(cid:1)(cid:3)p
,
(cid:2)σp(cid:0)f, s−p/n(cid:1)(cid:3)p dU(s) < ∞.
10
EGOR D. KOSOV
Proof. By Corollary 3.2 for the set As := {f ≥ kfkps} we have
N
ZAs f (x)p dx ≤ C(n, p)p(cid:2)σp(cid:0)f, (λn(As))1/n(cid:1)(cid:3)p
ZAs f (x)p dx dU(s) ≤ C(n, p)pZ +∞
Z +∞
p ) − U(N)(cid:3) dx
p ) dx −ZRn
I{f≥Nkfkp}f (x)p(cid:2)U(f (x)kfk−1
=ZRn f (x)pU(f (x)kfk−1
− U(N)ZRn
N
ZRn
Thus,
ZRn f (x)pU(f (x)kfk−1
The theorem is proved.
I{f<Nkfkp}f (x)pU(f (x)kfk−1
p ) dx
I{f≥Nkfkp}f (x)p dx ≥ZRn f (x)pU(f (x)kfk−1
p ) dx ≤ C(n, p)pZ +∞
N
(cid:2)σp(cid:0)f, s−p/n(cid:1)(cid:3)p dU(s) + U(N)kfkp
p.
p ) dx − U(N)kfkp
p.
(cid:3)
We now proceed to the second theorem. Let us introduce the following notation. Let
p ∈ [1, n) and let f ∈ Lp(Rn). Consider the function
vp(f, t) = σp(f, t
p
Since the function t → t
Moreover, limt→0 t−1vp(f, t) = lims→0 s−n/pσp(f, s). We note that, by concavity,
(3.1)
n is concave, the function vp(f,·) is also concave and nondecreasing.
n ).
p
σp(f, s) = σp(f, s · 1 + (1 − s) · 0) ≥ sσp(f, 1).
Thus, lims→0 s−n/pσ(f, s) = ∞ and the function v∗p(f, t) := tvp(f, t−1) is strictly increasing by
Lemma 2.4. To unify the notation, for the case n = p = 1 we also use the symbol vp(f,·),
which in this case coincides with σp(f,·).
Lemma 3.4. Let f ∈ Lp(Rn), where p ∈ [1, n) or n = p = 1. In the case n = p = 1 we also
assume that limt→0 t−1σ1(f, t) = ∞. Then
λn(f ≥ C(n, p)v∗p(f, t)) ≤ t−p,
v∗p(f, t) = tσp(f, t− p
n ),
where C(n, p) = 1 + ν−1/p
C(1, 1) = 1.
n
(n/p − 1)1/p−1, νn is the surface area of the unit sphere in Rn, and
Proof. As we have already mentioned, the function v∗ is strictly monotone. By Corollary 3.2,
for the set At := {f ≥ C(n, p)v∗p(f, t)} we have
C(n, p)v∗p(f, t)λn(At)1/p ≤(cid:18)ZAt f (x)p dx(cid:19)1/p
≤ C(n, p)σp(cid:0)f,(cid:0)λn(At)(cid:1)1/n(cid:1).
Thus,
By the strictly monotonicity of the function v∗p(f,·) we have the estimate
v∗p(f, t) ≤ v∗p(cid:0)f,(cid:0)λn(At)(cid:1)−1/p(cid:1).
λn(At) ≤ t−p
which completes the proof.
(cid:3)
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
11
We now proceed to the theorem itself.
Theorem 3.5. Let f ∈ Lp(Rn), where p ∈ [1, n) or n = p = 1. Let U : [0,∞) → [0,∞) be
a strictly increasing continuous function such that U(0) = 0, limt→∞ U(t) = ∞ and there are
positive constants a, r such that U(t) ≤ atp whenever 0 < t < r. Assume that there is a number
N > 0 such that
where C(n, p) = 1 + ν−1/p
C(1, 1) = 1. Then
n
(n/p − 1)1/p−1, νn is the surface area of the unit sphere in Rn, and
Z +∞
N
t−1−pU(cid:0)C(n, p)tσp(f, t−p/n)(cid:1) dt < ∞,
ZRn
U(f (x)) dx < ∞.
Proof. We first consider the case where either n > 1 or n = p = 1 and limt→0 t−1σp(f, t) = ∞.
Set ζ(s) = C(n, p)v∗p(f, s), where v∗ is defined by equality (3.1). Under our assumptions, the
function ζ(·) is continuous and strictly increasing. Set R = ζ(N).
um(x) = 0 at all other points x. For these functions we have
Consider the functions um such that um(x) = U(f (x)) if x ≤ m and U(f (x)) ≤ m and
ZRn
um(x) dx =Z m
=Z U (r)
λn(um ≥ t) dt ≤Z m
λn(U(f) ≥ t) dt +Z U (R)
U (r)
0
0
0
λn(U(f) ≥ t) dt
λn(U(f) ≥ t) dt +Z m
U (R)
λn(U(f) ≥ t) dt.
For the first term we have
Z U (r)
0
λn(U(f) ≥ t) dt =Z U (r)
0
λn(f ≥ U−1(t)) dt
≤Z U (r)
0
λn(f ≥ (a−1t)1/p) dt ≤Z ∞
0
λn(afp ≥ t)dt = aZRn f (x)p dx,
where in the second inequality we have used that a(U−1(t))p ≥ t if t ∈ (0, U(r)).
For the second term we have
U (r)
U (R)
For the third term, by Lemma 3.4, we have
Z U (R)
λn(U(f) ≥ t) dt ≤ (U(R) − U(r))λn(f ≥ r) ≤ (U(R) − U(r))r−pZRn f (x)p dx.
Z m
λn(U(f) ≥ t) dt =Z m
(t)(cid:3)−p
The last integral is the area under the graph of the strictly decreasing function(cid:2)(cid:0)U ◦ζ(cid:1)−1
[(U◦ζ)−1(m)]−p(cid:0)U ◦ ζ(cid:1)(t−1/p) dt − U(R)(cid:16)[ζ−1(R)]−p − [(U ◦ ζ)−1(m)]−p(cid:17)
λn(cid:16)f ≥ ζ(cid:0)ζ−1(U−1(t))(cid:1)(cid:17) dt ≤Z m
U (R)(cid:2)(cid:0)U ◦ ζ(cid:1)−1
Hence it is equal to
[ζ −1(R)]−p
Z
dt.
U (R)
.
(t)(cid:3)−p
The first integral in the above expression is equal to
Z (U◦ζ)−1(m)
N
s−1−pU(cid:0)ζ(s)(cid:1) ds ≤Z +∞
N
s−1−pU(cid:0)C(n, p)v∗p(f, s)(cid:1) ds < ∞,
+(cid:0)m − U(R)(cid:1)[(U ◦ ζ)−1(m)]−p.
12
EGOR D. KOSOV
since v∗p(f, s) = svp(f, 1/s) = sσp(f, s−p/n). We also note that
m[(U ◦ ζ)−1(m)]−p ≤ pZ +∞
(U◦ζ)−1(m)
t−1−pU(cid:0)ζ(t)(cid:1) dt.
Summing up these estimates and taking the limit as m tends to infinity, by Fatou's lemma, we
get
ZRn
U(f (x)) dx ≤ aZRn f (x)p dx + (U(C(n, p)v∗p(f, N)) − U(r))r−pZRn f (x)p dx
s−1−pU(cid:0)C(n, p)sσp(f, s−p/n)(cid:1) ds.
+Z +∞
N
which completes the proof in the case under consideration.
(3.2)
In the case where n = p = 1 and limt→0 t−1σp(f, t) = A for some constant A (the limit exists
by monotonicity) the function f has bounded variation and is bounded by the constant A.
Thus,
ZR
U(f (x)) dx ≤ aZ{f<r}
f (x) dx + U(A)λ(f ≥ r) ≤ (a + r−1U(A))ZR f (x) dx.
(cid:3)
The theorem is proved.
Remark 3.6. We note that the condition
N
is equivalent to the condition
Z +∞
Z +∞
N ′
t−1−pU(cid:0)C(n, p)tσp(f, t−p/n)(cid:1) dt < ∞,
s−1−nU(cid:0)C(n, p)sn/p−1σ∗p(f, s))(cid:1) ds < ∞
which can be verified by the change of variables t = sn/p.
Remark 3.7. Let us consider the case n = p = 1 in Theorems 3.3 and 3.5. In this case the
condition in the first theorem coincides with
and the condition in the second one coincides with
N
Z +∞
σp(cid:0)f, t−1(cid:1) dU(t) < ∞
Z +∞
t−2U(cid:0)tσp(f, t−1)(cid:1) dt < ∞.
N
Both conditions are integral forms of the classical Ulyanov conditions from [18] and [19], for-
mulated in the introduction.
4. Besov classes on spaces with Gaussian measures
We now proceed to the infinite-dimensional Gaussian case. Let X be a real Hausdorff locally
convex space with the topological dual space X∗.
We recall that a Borel measure γ on X is called Radon measure if for every Borel set B ⊂ X
and every ε > 0 there is a compact set K ⊂ B such that γ(B \ K) < ε. We also recall that a
Radon measure γ on X is a centered Gaussian measure if, for every continuous linear functional
l on X, the image measure γ ◦ l−1 is either Dirac's measure at zero or has a density of the form
(2πc2)−1/2 exp(−t2/2c2). From now on let γ be a Radon centered Gaussian measure on X.
For a function f ∈ Lp(γ) we set
kfkp := kfkLp(γ) :=(cid:18)ZX fp dγ(cid:19)1/p
.
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
13
Recall that the Cameron–Martin norm of a vector h ∈ X is defined by
hH = sup(cid:26)l(h) : ZX
l2 dγ ≤ 1, l ∈ X∗(cid:27).
Let H ⊂ X be the linear subspace of all vectors h ∈ X such that hH < ∞. This subspace H is
called the Cameron–Martin space of the measure γ. If γ is the standard Gaussian measure on
Rn, then its Cameron–Martin space is Rn itself and if γ is the countable power of the standard
Gaussian measure on the real line, then H is the classical Hilbert space l2. For a general Radon
centered Gaussian measure, the Cameron–Martin space is also a separable Hilbert space (see [4,
Theorem 3.2.7 and Proposition 2.4.6]) with the inner product h·,·iH generated by the Cameron–
Martin norm · H.
Let {li}∞i=1 ⊂ X∗ be an orthonormal basis in the closure X∗γ of the set X∗ in L2(γ). There is
an orthonormal basis {ei}∞i=1 in H such that li(ej) = δi,j (see [4]). We will use below that for
any orthonormal family l1, . . . , ln ∈ X∗γ the distribution of the vector (l1, . . . , ln), i.e., the image
of the measure γ, is the standard Gaussian measure γn on Rn, i.e., the measure with density
(2π)−n/2 exp(−x2/2) with respect to the standard Lebesgue measure on Rn.
Let FC∞(X) denote the set of all functions ϕ on X of the form ϕ(x) = ψ(l1(x), . . . , ln(x)),
where ψ ∈ C∞b (Rn), li ∈ X∗, and let FC∞0 (X) denote the set of all functions ϕ on X of the
form ϕ(x) = ψ(l1(x), . . . , ln(x)), where ψ ∈ C∞0 (Rn), li ∈ X∗. Let FC∞(X, H) be the set of all
vector fields Φ of the form
Φ(x) =
Ψi(g1(x), . . . , gn(x))hi,
nXi=1
where Ψi ∈ C∞b (Rn), gi ∈ X∗, hi ∈ H and let FC∞0 (X, H) be the subset of this class consisting
of mappings for which Ψi can be chosen with compact support. Note that here we can actually
take vectors hi orthogonal in H and functionals gi orthogonal in X∗γ such that gi(hj) = δij. We
will call such vectors and functionals biorthogonal.
For every ϕ ∈ FC∞(X) of the form ϕ(x) = ψ(l1(x), . . . , ln(x)) set
∂xj ψ(l1(x), . . . , ln(x))ej,
∇ϕ(x) =
nXj=1
where {li} and {ei} are biorthogonal. Let divγ be the "adjoint operator" to the gradient
operator ∇ with respect to γ, that is,
(divγΦ)ϕ dγ = −ZXhΦ,∇ϕiH dγ.
for arbitrary Φ ∈ FC∞(X, H) and φ ∈ FC∞(X). One can easily check that
divγΦ(x) =
∂xj Ψj(l1(x), . . . , ln(x)) − lj(x)Ψj(l1(x), . . . , ln(x))
for a vector field Φ ∈ FC∞(X, H) of the form
nXi=1
Φ(x) =
Ψi(l1(x), . . . , ln(x))ei
with biorthogonal {li} and {ei}. We note that for a vector field Φ from FC∞0 (X, H) its diver-
gence divγΦ is a bounded function.
We recall that the Ornstein–Uhlenbeck semigroup is defined by the equality
ZX
nXj=1
Ttf (x) :=ZX
f (e−tx + √1 − e−2ty) γ(dy)
for any function f ∈ L1(γ).
e−τ
√1 − e−2τ ≤ (2t)−1/2,
Let us define the Gaussian modulus of continuity σγ,p(f,·) which plays the same role as the
ct ≤ (2t)1/2, and lim
t→∞
ct = π/2.
function σp(f,·) introduced above in case of Rn.
Definition 4.1. Let f ∈ Lp(γ). Set
σγ,p(f, ε) := supnZ divγΦf dγ, Φ ∈ FC∞0 (X, H),kdivγΦkq ≤ 1,kΦkq ≤ εo.
We note that the function σγ,p(f,·) is continuous, concave, and nondecreasing on (0, +∞),
which can be proved similarly to Lemma 2.3. Thus, by approximation, in the definition of
the quantity σγ,p(f, ε) the supremum can be taken over all vector fields Φ ∈ FC∞(X, H) with
kdivγΦkq ≤ 1,kΦkq ≤ ε.
Using the previous definition we can now introduce Besov classes on a locally convex space
endowed with a Gaussian measure.
Definition 4.2. Let α ∈ (0, 1], p ∈ [1,∞), θ ∈ [1,∞]. We say that a function f ∈ Lp(γ)
belongs to the Gaussian Besov space Bα
p,θ(γ) if the quantity
14
EGOR D. KOSOV
We now fix an orthonormal basis {ln} ⊂ X∗ in X∗γ . For any function f ∈ L1(γ) let Enf be a
function on Rn such that
ZRn
ψEnf dγn =ZX
ψ(cid:0)l1(x), . . . , ln(x)(cid:1)f (x) γ(dx) ∀ ψ ∈ C∞b (Rn),
where γn is the standard Gaussian measure on Rn. This equality actually means that the
function Enf (l1, . . . , ln) is the conditional expectation of f with respect to the σ-field generated
by functions l1, . . . , ln. By the known property of conditional expectations, for any function
f ∈ Lp(γ), we have
We also introduce the following functions C(p) and ct to be used further:
kf − Enf (l1, . . . , ln)kp → 0, n → ∞.
C(p) :=(cid:18)(2π)−1/2ZR spe− s2
2 ds(cid:19)1/p
and ct :=Z t
0
e−τ
√1 − e−2τ
dτ.
We note that
V p,θ,α(f ) =(cid:16)Z ∞
0
(cid:2)s−ασγ,p(f, s)(cid:3)θs−1ds(cid:17)1/θ
is finite.
and
Gaussian Nikolskii–Besov class introduced in [7].
We note that in the case θ = ∞ the above definition coincides with the definition of the
We will give an equivalent description of these Gaussian Besov classes in terms of the following
two characteristics.
Definition 4.3. For a function f ∈ Lp(γ), p ∈ [1,∞), set
aγ,p(f, t) :=(cid:16)ZZ f (e−tx + √1 − e−2ty) − f (x)p γ(dx)γ(dy)(cid:17)1/p
Ap,θ,α
γ
(f ) :=(cid:16)Z ∞
0
(cid:2)t−α/2aγ,p(f, t)(cid:3)θ
t−1dt(cid:17)1/θ
.
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
15
We note that
kf − Ttfkp ≤ aγ,p(f, t).
In a sense, the function aγ,p(f,·) can be regarded as a Gaussian replacement for the finite-
dimensional modulus of continuity ωp(f,·), since we cannot directly use shifts fh of the function
f ∈ Lp(γ), since these shifts can fail to be in Lp(γ).
We need the following technical lemma.
Lemma 4.4. Let γn be the standard Gaussian measure on Rn. Then for any function f ∈ Lp(γn),
where p ∈ [1,∞), we have
aγn,p(f, t) ≤ 2σγn,p(f, 2−1C(p)ct).
Proof. For every function ϕ ∈ C∞0 (R2n) we can write
ZZ ϕ(x, y)[f (e−tx + √1 − e−2ty) − f (x)] γn(dx)γn(dy)
=Z f (u)Z [ϕ(e−tu −
√1 − e−2tv,√1 − e−2tu + e−tv) − ϕ(u, v)] γn(dv) γn(du)
=Z f (u)Z t
√1 − e−2sv,√1 − e−2su + e−sv) γn(dv).
∂
∂s
0
gs(u) ds γn(du),
We now note that for an arbitrary function ψ ∈ C∞0 (Rn) we have
where
gs(u) :=Z ϕ(e−su −
Z ψ(u)
∂
∂s
gs(u) γn(du)
=
∂
∂sZ ψ(u)Z ϕ(e−su −
√1 − e−2sv,√1 − e−2su + e−sv) γn(dv) γn(du)
=
∂
∂sZZ ϕ(x, y)ψ(e−sx + √1 − e−2sy) γn(dx) γn(dy)
=
where
=
e−s
√1 − e−2sZZ ϕ(x, y)h∇ψ(e−sx + √1 − e−2sy), e−sy −
√1 − e−2sZ D∇ψ(u),Z vϕ(e−su −
e−s
√1 − e−2sxi γn(dx) γn(dy)
√1 − e−2sv,√1 − e−2su + e−sv) γn(dv)E γn(du)
= −Z ψ(u)divγnGs(u) γn(du),
√1 − e−2sv,√1 − e−2su + e−sv) γn(dv) ∈ C∞b (Rn).
Gs(u) :=
e−s
√1 − e−2sZ vϕ(e−su −
Thus,
and
Z fZ t
0
∂
∂s
∂
∂s
gs(u) = divγ(cid:0)−Gs(u)(cid:1)
gs ds dγn =Z divγ(cid:16)−Z t
0
Gs ds(cid:17)f dγn.
16
EGOR D. KOSOV
We observe that
0
0
Moreover, we have
∂
∂s
gs(u) ds
Gs ds(cid:17) =Z t
divγn(cid:16)−Z t
0
and that
divγn(cid:0)−Gs(cid:1) ds =Z t
=Z [ϕ(e−tu −
(cid:13)(cid:13)(cid:13)divγn(cid:16)−Z t
(cid:13)(cid:13)(cid:13)Z t
√1 − e−2tv,√1 − e−2tu + e−tv) − ϕ(u, v)] γn(dv)
Gs, ds(cid:17)(cid:13)(cid:13)(cid:13)q ≤ 2kϕkLq(γn⊗γn).
Gs ds(cid:13)(cid:13)(cid:13)q ≤Z t
√1 − e−2sZZ hΨ(u), viϕ(e−su−
√1 − e−2skϕkLq(γn⊗γn)(cid:16)ZZ hΨ(u), vip γn(dv)γn(du)(cid:17)1/p
0 (cid:13)(cid:13)Gs(cid:13)(cid:13)q ds
and it remains to estimate (cid:13)(cid:13)Gs(cid:13)(cid:13)q. To do this, we note that for an arbitrary vector field
Z hΨ, Gsi dγn =
√1 − e−2sv,√1 − e−2su+e−sv) γn(dv)γn(du)
Ψ ∈ C∞0 (Rn)
e−s
e−s
≤
0
0
e−s
√1 − e−2s
C(p)kϕkLq(γn⊗γn)kΨkp.
Thus,
≤
e−s
and
(cid:13)(cid:13)Gs(cid:13)(cid:13)q ≤ C(p)
Gs ds(cid:13)(cid:13)(cid:13)q ≤Z t
√1 − e−2skϕkLq(γn⊗γn)
0 (cid:13)(cid:13)Gs(cid:13)(cid:13)q ds ≤ C(p)ctkϕkLq(γn⊗γn).
ZZ ϕ(x, y)[f (e−tx + √1 − e−2ty) − f (x)] γn(dx)γn(dy) =Z fZ t
Hence, for an arbitrary function ϕ ∈ C∞0 (R2n) with kϕkLq(γn⊗γn) = 1 we have
(cid:13)(cid:13)(cid:13)Z t
∂
∂s
0
0
gs ds dγn
Taking the supremum over functions ϕ ∈ C∞0 (R2n) with kϕkLq(γn⊗γn) = 1 we obtain the an-
nounced bound.
(cid:3)
=Z divγn(cid:16)−Z t
0
Gs ds(cid:17)f dγn ≤ 2σγn,p(f, 2−1C(p)ct).
The following theorem is a Gaussian analog of Theorem 2.7.
Theorem 4.5. For any function f ∈ Lp(γ), where p ∈ [1,∞), we have
If p > 1 we have the inverse bound:
aγ,p(f, t) ≤ 2σγ,p(f, 2−1C(p)ct).
σγ,p(f, ε) ≤(cid:0)1 + C(p/(p − 1))(cid:1)aγ,p(f, ε2).
Proof. To prove the first part of the lemma, we fix an orthonormal basis {ln} ⊂ X∗ in X∗γ . By
the previous lemma we have
aγn,p(Enf, t) ≤ 2σγn,p(Enf, 2−1C(p)ct) ≤ 2σγ,p(f, 2−1C(p)ct).
We observe that aγn,p(f, t) → aγ,p(f, t) as n tends to infinity, which completes the proof.
= C(q)t−1/2aγ,p(f, t)kΦkq.
for an arbitrary fixed point x. Thus, the last expression is equal to
−
e−t
√1 − e−2tZZ hΦ(x), yiH(cid:2)f (e−tx + √1 − e−2ty) − f (x)(cid:3) γ(dy)γ(dx)
≤ t−1/2aγ,p(f, t)(cid:16)ZZ hΦ(x), yiHq γ(dy)γ(dx)(cid:17)1/q
Z divγΦTtf dγ ≤ C(q)t−1/2aγ,p(f, t)kΦkq.
So, we have proved the estimate
Now we have
Z divγΦf dγ =Z divγΦ[f − Ttf ] dγ +Z divγΦTtf dγ.
The first term in the above expression is estimated by
aγ,p(f, t)kdivγΦkq
and the second term, as we have proved, is not greater than
C(q)t−1/2aγ,p(f, t)kΦkq.
Taking t = ε2 we obtain
σγ,p(f, ε) ≤ (1 + C(q))aγ,p(f, ε2),
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
17
write
Let now f ∈ Lp(γ) for some p > 1. For an arbitrary vector field Φ ∈ FC∞0 (X, H) we can
Z divγΦTtf dγ = e−tZ divγTtΦf dγ
√1 − e−2tZZ f (u)hΦ(e−tu −
√1 − e−2tv), e−tv + √1 − e−2tuiH γ(dv)γ(du)
= −
e−t
= −
e−t
√1 − e−2tZZ f (e−tx + √1 − e−2ty)hΦ(x), yiH γ(dy)γ(dx).
Z f (x)hΦ(x), yiH γ(dy) = 0
We observe that
which is the announced bound.
Corollary 4.6. For any function f ∈ Bα
u : R → R we have
p,θ(γ), where p ∈ [1,∞), and for any Lipschitz function
(cid:3)
where Lip(u) is the Lipschitz constant of the function u.
aγ,p(u(f ), t) ≤ 21−α(αθ)1/θLip(u)C(p)αcα
t V p,θ,α(f ),
Proof. By the Lipschitz continuity of the function u and by the previous lemma we can write
aγ,p(u(f ), t) =(cid:18)ZZ (cid:12)(cid:12)(cid:12)u(cid:0)f (e−tx +p1 − e−2ty)(cid:1) − u(cid:0)f (x)(cid:1)(cid:12)(cid:12)(cid:12)
≤ Lip(u)(cid:18)ZZ (cid:12)(cid:12)(cid:12)f (e−tx +p1 − e−2ty) − f (x)(cid:12)(cid:12)(cid:12)
p
γ(dy)γ(dx)(cid:19)1/p
γ(dy)γ(dx)(cid:19)1/p
p
We now note that
(cid:0)2−1C(p)ct(cid:1)−αθ(cid:2)σγ,p(f, 2−1C(p)ct)(cid:3)θ
≤ αθ
= Lip(u)aγ,p(f, t) ≤ 2Lip(u)σγ,p(f, 2−1C(p)ct).
r−αθ−1[σγ,p(f, r)]θdr ≤ αθ(cid:2)V p,θ,α(f )(cid:3)θ
∞Z2−1C(p)ct
.
18
Thus,
as announced.
EGOR D. KOSOV
aγ,p(u(f ), t) ≤ 21−α(αθ)1/θLip(u)C(p)αcα
t V p,θ,α(f )
(cid:3)
Also, as a corollary, we obtain that the conditions V p,θ,α
γ
(f ) < ∞ and Ap,θ,α
γ
(f ) < ∞ are
equivalent for p > 1.
Corollary 4.7. For any function f ∈ Bα
p,θ(γ), where p ∈ [1,∞), we have
Ap,θ,α
γ
(f ) ≤ 21−α+1/θC(p)αV p,θ,α
γ
(f ).
Moreover, for p ∈ (1,∞) we have the inverse statement, that is, if for a function f ∈ Lp(γ) the
quantity Ap,θ,α
p,θ(γ) and
γ
(f ) is finite, then f ∈ Bα
V p,θ,α
γ
Proof. For a function f ∈ Bα
(f ) ≤ 2−1/θ(cid:0)1 + C(p/(p − 1))(cid:1)Ap,θ,α
p,θ(γ), by Lemma 4.5, we have
γ
(f ).
aγ,p(f, t) ≤ 2σγ,p(f, 2−1C(p)ct) ≤ 2σγ,p(f, 2−1C(p)t1/2).
Thus,
γ
(cid:2)Ap,θ,α
(f )(cid:3)θ =Z ∞
0
(cid:2)t−α/2aγ,p(f, t)(cid:3)θt−1dt ≤ 2θZ ∞
= 21+θ−αθC(p)αθZ ∞
r−αθ(cid:2)σγ,p(f, r)(cid:3)θ
t−αθ/2(cid:2)σγ,p(f, 2−1C(p)t1/2)(cid:3)θt−1dt
r−1dr = 21+θ−αθC(p)αθ(cid:2)V p,θ,α
γ
0
0
,
(f )(cid:3)θ
which is the announced bound.
Conversely, for any function f ∈ Lp(γ) with p > 1 and finite Ap,θ,α
γ
(f ), Lemma 4.5 gives that
σγ,p(f, ε) ≤(cid:0)1 + C(q)(cid:1)aγ,p(f, ε2),
which yields
γ
=Z ∞
(cid:2)V p,θ,α
(f )(cid:3)θ
≤(cid:0)1 + C(q)(cid:1)θZ ∞
0
r−1dr
(cid:2)r−ασγ,p(f, r)(cid:3)θ
(cid:2)r−αaγ,p(f, r2)(cid:3)θr−1dt = 2−1(cid:0)1 + C(q)(cid:1)θZ ∞
0
This is the announced estimate.
0
(cid:2)t−α/2aγ,p(f, t)(cid:3)θt−1dt
(f )(cid:3)θ.
= 2−1(cid:0)1 + C(q)(cid:1)θ(cid:2)Ap,θ,α
γ
(cid:3)
We now proceed to a log-Sobolev-type embedding theorem for Besov classes with respect
to a Gaussian measure. As we have mentioned in the introduction, the main idea of the
proof is to use the short time behavior of the Ornstein-Uhlenbeck semigroup together with its
hypercontractivity property, similarly in a sense to the approach from [14].
Theorem 4.8. For any function f ∈ Bα
p,θ(γ), where p ∈ (1,∞), and for any number β ∈ (0, α)
the function f lnfβ/2 belongs to Lp(γ). Moreover, there is a constant C = C(p, θ, α, β),
depending only on parameters p, θ, α, and β, such that
(cid:16)Z fp(cid:12)(cid:12)ln(fkfk−1
p )(cid:12)(cid:12)pβ/2
dγ(cid:17)1/p
≤ C(cid:0)kfkp + V p,θ,α
γ
(f )(cid:1).
Proof. We recall the hypercontractivity property of the Ornstein–Uhlenbeck semigroup (see [4,
Theorem 5.5.3]): for any function f ∈ Lp(γ) one has
kTtfk1+(p−1)e2t ≤ kfkp.
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
19
For an arbitrary number s > 0, let As := {f ≥ s}. We note that the function τ 7→ max{τ, s}
is 1-Lipschitz. Thus, for an arbitrary function ϕ ∈ FC∞(X) and any number t > 0, by
Corollary 4.6 and by the hypercontractivity property, we have
Z ϕIAs(f − s) dγ =Z IAsϕ(max{f, s} − s) dγ
=Z IAsϕ(cid:2)max{f, s} − Tt(max{f, s})(cid:3) dγ +Z IAsϕTt(max{f, s} − s) dγ
≤ kϕkqk max{f, s} − Tt(max{f, s})kp + kIAsϕk 1+(p−1)e2t
(p−1)e2t kTt(max{f, s} − s)k1+(p−1)e2t
(p−1)e2t kIAs(f − s)kp.
(f )tα/2 + kIAsϕk 1+(p−1)e2t
≤ 21−α(αθ)1/θC(p)αkϕkqV p,θ,α
γ
We note that
1
1 + (p − 1)e2t
(p − 1)e2t = 1 +
(p − 1)e2t = q(1/q + 1/(pe2t)) ≤ q.
Thus, we can apply Holder's inequality to the expression kIAsϕk 1+(p−1)e2t
(1/p − 1/(pe2t))−1 and (1/q + 1/(pe2t))−1, which yields
(p−1)e2t ≤ [γ(As)]
kIAsϕk 1+(p−1)e2t
q+pe2t kϕkq.
(p−1)e2t
e2t −1
with the exponents
Taking the supremum over functions ϕ with kϕkq = 1 we obtain the estimate
kIAs(f − s)kp ≤ 21−α(αθ)1/θC(p)αV p,θ,α
γ
We now observe that
(f )tα/2 + [γ(As)]
e2t −1
q+pe2t kIAs(f − s)kp.
e2t − 1
q + pe2t = p−1
q−1(e2t − 1)
1 + q−1(e2t − 1) ≥ p−1
2q−1t
1 + 2q−1t ≥
t
pq
whenever t ≤ 1/2. Thus, whenever t ≤ 1/2, we have
kIAs(f − s)kp ≤ 21−α(αθ)1/θC(p)αV p,θ,α
γ
(f )tα/2 + [γ(As)](pq)−1tkIAs(f − s)kp.
For the sets As with γ(As) ≤ e−2pq we can take t = pq(− ln γ(As))−1 ≤ 1/2 and conclude that
kIAs(f − s)kp ≤ 21−α(αθ)1/θC(p)α(pq)α/2V p,θ,α
γ
(f )[− ln γ(As)]−α/2 + e−1kIAs(f − s)kp,
since
for such t. The obtained inequality can be rewritten in the form
[γ(As)](pq)−1t = e−(pq)−1[− ln γ(As)]t = e−1
kIAs(f − s)kp ≤ C(p, θ, α)V p,θ,α
γ
(f )[− ln γ(As)]−α/2,
where C(p, θ, α) = 21−α(αθ)1/θC(p)αe(e − 1)−1(pq)α/2. We now observe that γ(As) ≤ kfkp
and IAs(f − s) ≥ 2−1IA2sf. Thus, if t ≥ e2q, taking s = tkfkp, we have
ps−p
Z I{f≥2tkfkp}fp dγ ≤(cid:0)2p−α/2C(p, θ, α)(cid:1)p(cid:2)V p,θ,α
γ
(f )(cid:3)p[ln t]−pα/2.
Multiplying both sides of the inequality by t−1[ln t]−1+pβ/2 and integrating with respect to t
from e2q to +∞ we obtain
Z ∞
e2q
t−1[ln t]−1+pβ/2Z I{f≥2tkfkp}fp dγ dt
≤(cid:0)2p−α/2C(p, θ, α)(cid:1)p(cid:2)V p,θ,α
γ
(f )(cid:3)pZ ∞
=(cid:0)2p−α/2C(p, θ, α)(cid:1)p
e2q
[ln t]−1−p(α−β)/2t−1dt
2p−1(α − β)−1(2q)−p(α−β)/2(cid:2)V p,α
γ
.
(f )(cid:3)p
20
EGOR D. KOSOV
The left-hand side of the above estimate is equal to
Z fpI{f≥2e2qkfkp}Z fkfk−1
e2q
p 2−1
t−1[ln t]−1+pβ/2 dt dγ
= 2(pβ)−1Z fpI{f≥2e2qkfkp}(cid:0)[ln(fkfk−1
≥ 2(pβ)−1p−pβ/2Z fpI{f≥2e2qkfkp}[ln(fkfk−1
p 2−1)]pβ/2 − (2q)pβ/2(cid:1) dγ
p )]pβ/2 dγ − 2(pβ)−1(2q)pβ/2kfkp
p
= 2(pβ)−1p−pβ/2Z fp(cid:12)(cid:12)ln(fkfk−1
p )(cid:12)(cid:12)pβ/2
−2(pβ)−1p−pβ/2Z fpI{f<2e2qkfkp}(cid:12)(cid:12)ln(fkfk−1
p )(cid:12)(cid:12)pβ/2
− 2(pβ)−1(2q)pβ/2kfkp
p
dγ
Thus, since a ln aβ/2 ≤ 2e2q(2q + 1)β/2 if a ∈ [0, 2e2q], we have
Z fp(cid:12)(cid:12)ln(fkfk−1
p )(cid:12)(cid:12)pβ/2
C1(p, θ, α, β) =(cid:0)2p−α/2C(p, θ, α)(cid:1)p(α − β)−1(2q)−p(α−β)/2ppβ/2β
dγ ≤ C1(p, θ, α, β)(cid:2)V p,α
(f )(cid:3)p
γ
+ C2(p, α, θ, β)kfkp
p
with
and
C2(p, θ, α, β) = (2pe2qp(2q + 1)pβ/2 + (2q)pβ/2ppβ/2)
It is readily seen that the obtained bound is equivalent to the announced assertion. The theorem
is proved.
(cid:3)
Finally, let us discuss estimates for the best approximations by Hermite polynomials in L2(γ)
with respect to a Gaussian measure γ. Recall (see [4, Section 2.9]) that the space L2(γ) can be
decomposed into the direct sum of mutually orthogonal subspaces Hk consisting of the so-called
Hermite polynomials of a fixed degree k:
L2(γ) =
∞Mk=0
Hk.
This decomposition is also called the Wiener chaos decomposition. The space Hk is actually the
orthogonal complement of the space of all measurable polynomials of degree k − 1 in the space
of all measurable polynomials of degree k. Let Ik be the projection operator to the subspace
Hk. For any function f ∈ L2(γ) set
EN (f ) := inf{kf − fNk2; fN ∈
N−1Mk=0
Hk}.
The quantity EN (f ) is the value of the best approximation of the function f by linear combi-
nations of Hermite polynomials of the given degree. It is clear that
We now prove a Jackson–Stechkin-type inequality for the quantity EN (f ) involving the
EN (f ) = kf − I0(f ) − . . . − IN−1(f )k2.
Gaussian modulus of continuity σγ,2(f,·).
Theorem 4.9. For any function f ∈ L2(γ) we have
EN−1(f ) ≤ σγ,2(f,√2πN−1/2).
BESOV CLASSES ON FINITE- AND INFINITE-DIMENSIONAL SPACES
21
Proof. For an arbitrary function ϕ ∈ FC∞(X) with kϕk2 ≤ 1 we have
Z ϕ(f − I0(f ) − . . . − IN−1(f )) dγ =Z (ϕ − I0(ϕ) − . . . − IN−1(ϕ))f dγ
where we have used the equality
0 ∇Tt(ϕ − I0(ϕ) − . . . − IN−1(ϕ)) dt(cid:17)f dγ,
=Z divγ(cid:16)−Z ∞
LTtψ(x) dt = divγ(cid:16)−Z ∞
0 ∇Ttψ(x) dt(cid:17)
ψ(x) = −Z ∞
0
for an arbitrary function ψ ∈ FC∞(X) with R ψ dγ = 0, where L is the Ornstein–Uhlenbeck
operator (see [4, Section 1.4 and Remark 5.8.7]). Recall that
for any function ψ ∈ FC∞(X) and that
Ttg =
which yields the estimate
k∇Ttψk2 ≤
e−t
√1 − e−2tkψk2
∞Xk=0
e−ktIk(g),
kTt(g − I0(g) − . . . − IN−1(g))k2 ≤ e−N tkgk2,
for all g ∈ L2(γ). We now note that
(cid:13)(cid:13)(cid:13)Z ∞
0 ∇Tt(ϕ − I0(ϕ) − . . . − IN−1(ϕ)) dt(cid:13)(cid:13)(cid:13)2 ≤Z ∞
≤ kϕk2Z ∞
≤Z ∞
e−t/2
0
0
0
k∇Tt(ϕ − I0(ϕ) − . . . − IN−1(ϕ))k2 dt
√1 − e−tkTt/2(ϕ − I0(ϕ) − . . . − IN−1(ϕ))k2 dt
e−N t/2 dt = B((N + 1)/2, 1/2)kϕk2,
where B(x, y) is the standard beta function. It can be easily verified that
e−t/2
√1 − e−t
≤ √x,
Γ(x + 1/2)
Γ(x)
where Γ(·) is the standard gamma function. Indeed, introducing the probability density
ρx(t) := [Γ(x)]−1tx−1e−tI{t>0}
and applying Jensen's inequality we obtain
Γ(x + 1/2)
Γ(x)
=Z √tρx(t) dt ≤sZ tρx(t) dt =sΓ(x + 1)
Γ(x)
= √x.
Thus,
B((N + 1)/2, 1/2) =
Γ((N + 1)/2)Γ(1/2)
Γ(1 + N/2)
√πΓ(N/2 + 1/2)
N/2Γ(N/2) ≤
=
√2πN−1/2.
Therefore,
(cid:13)(cid:13)(cid:13)Z ∞
0 ∇Tt(ϕ − I0(ϕ) − . . . − IN−1(ϕ)) dt(cid:13)(cid:13)(cid:13)2 ≤
√2πN−1/2.
We also note that
(cid:13)(cid:13)(cid:13)divγ(cid:16)−Z ∞
0 ∇Tt(ϕ− I0(ϕ)− . . .− IN−1(ϕ)) dt(cid:17)(cid:13)(cid:13)(cid:13)2
= kϕ− I0(ϕ)− . . .− IN−1(ϕ)k2 ≤ kϕk2 ≤ 1.
22
So,
EGOR D. KOSOV
Z ϕ(f − I0(f ) − . . . − IN−1(f )) dγ ≤ σγ,2(f,√2πN−1/2),
which completes the proof.
(cid:3)
References
[1] Ambrosio L., Miranda Jr M. and Pallara D. "Some fine properties of BV functions on Wiener spaces",
Analysis and Geometry in Metric Spaces 3:1 (2015), 212–230.
[2] Besov O.V., Il'in V.P., Nikolskii S.M. Integral representations of functions and imbedding theorems, V. I, II.
Winston & Sons, Washington; Halsted Press, New York – Toronto – London, 1978, 1979.
[3] Fukushima M., Hino M. "On the space of BV functions and a related stochastic calculus in infinite
dimensions", J. Funct. Anal. 183:1 (2001), 245–268.
[4] Bogachev V.I. Gaussian measures, Amer. Math. Soc., Providence, Rhode Island, 1998.
[5] Bogachev V.I. "Distributions of polynomials on multidimensional and infinite-dimensional spaces with
measures", Uspehi Mat. Nauk 71:4 (2016), 107–154 (in Russian); English transl. in Russian Math. Surveys
71:4 (2016), 1–47.
[6] Bogachev V.I., Kosov E.D. and Zelenov G.I. "Fractional smoothness of distributions of polynomials and
a fractional analog of the Hardy–Landau–Littlewood inequality", to appear in Trans. Amer. Math. Soc.,
http://arxiv.org/abs/1602.05207.
[7] Bogachev V.I., Kosov E.D. and Popova S.N. "New approach to the Nikolskii–Besov classes", to appear,
https://arxiv.org/abs/1707.06477.
[8] Kolyada V.I. "On imbedding in classes φ(L)", Izv. AN. SSSR 39:2 (1975), 418–437 (in Russian); English
transl. in Math. USSR-Izv. 9:2 (1975), 395–413.
[9] Kolyada V.I. "Estimates of rearrangements and imbedding theorems", Mat. Sb. 136(178):1(5) (1988), 3–23
(in Russian); English transl. in Math. USSR-Sb. 64:1 (1989), 1–21.
[10] Kolyada V.I. "Rearrangements of functions and embedding theorems", Uspekhi Mat. Nauk 44:5(269) (1989),
61–95 (in Russian); English transl. in Russian Math. Surveys 44:5 (1989), 73–117.
[11] Kosov E.D. "Fractional smoothness of images of logarithmically concave measures under polynomials", to
appear, https://arxiv.org/abs/1605.00162.
[12] Ledoux M. "Isoperimetry and Gaussian analysis", Lecture Notes in Math. 1648 (1996), 165–294.
[13] Ledoux M. Concentration of measure and logarithmic Sobolev inequalities, Seminaire de probabilites
XXXIII, Springer Berlin Heidelberg, 1999.
[14] Ledoux M. "Semigroup proofs of the isoperimetric inequality in Euclidean and Gauss space", Bulletin des
sciences mathematiques 118:6 (1994), 485–510.
[15] Nikolskii S.M. Approximation of functions of several variables and imbedding theorems, Transl. from Rus-
sian. Springer-Verlag, New York – Heidelberg, 1975 (Russian ed.: Moscow, 1977).
[16] Triebel H. Theory of function spaces, V.II, Birkhauser Verlag, Basel, 1992.
[17] Stein E. Singular integrals and differentiability properties of functions, Princeton University Press, Prince-
ton, 1970.
[18] Ul'yanov P.L. "The imbedding of certain function classes H ω
p ", Izv. Akad. Nauk SSSR 32:3 (1968), 649–686
(in Russian); English transl. in Math. USSR-Izv. 2:3 (1968), 601–637.
[19] Ul'yanov P.L. "Imbedding theorems and relations between best approximations (moduli of continuity) in
different metrics", Mat. Sb. 81(123):1 (1970), 104–131 (in Russian); English transl. in Math. USSR-Sb. 10:1
(1970), 103–126.
|
1609.03514 | 1 | 1609 | 2016-09-12T18:11:03 | H\"older-Besov boundedness for periodic pseudo-differential operators | [
"math.FA"
] | In this work we give H\"older-Besov estimates for periodic Fourier multipliers. We present a class of bounded pseudo-differential operators on periodic Besov spaces with symbols of limited regularity. | math.FA | math |
H OLDER-BESOV BOUNDEDNESS FOR PERIODIC
PSEUDO-DIFFERENTIAL OPERATORS
DUV ´AN CARDONA 1
Abstract. In this work we give Holder-Besov estimates for periodic Fourier
multipliers. We present a class of bounded pseudo-differential operators on
periodic Besov spaces with symbols of limited regularity.
MSC 2010. Primary 43A22, 43A77; Secondary 43A15.
1. Introduction
In this paper we study the boundedness of periodic Fourier multipliers and
periodic pseudo-differential operators from Holder spaces into Besov spaces. Let
σ : Z → C be a symbol, the corresponding Fourier multiplier Op(σ) is the periodic
pseudo-differential operator formally defined by the formula
Op(σ(·))f = F −1(σ(ξ)F (f )),
(1.1)
where F is the Fourier transform on the torus T = [0, 2π) and F −1 is the in-
verse Fourier transform. In 1979, Agranovich [1] proposed a global quantization
of periodic pseudo-differential operators on the circle S1 ≡ T. Later, this theory
was widely developed by Ruzhansky and Turunen in [29], where the theory of
periodic pseudo-differential operators is considered in arbitrary dimensions. Pe-
riodic Besov spaces form a class of function spaces which are of special interest
in analysis and mathematical physics. They can be defined via dyadic decompo-
sition and form scales Br
p,q(T) carrying three indices: r ∈ R, 0 < p, q ≤ ∞. In
the special case p = q = ∞, Λr(T) = Br
∞,∞(T) is nothing else but the familiar
space of all Holder continuous functions of order 0 < r < 1. There are several
possibilities concerning the conditions to impose on a symbol σ in the attempt to
establish a periodic Fourier multiplier theorem of boundedness on Besov spaces
and Lebesgue spaces for its corresponding operator (1.1) (see [5, 6, 9, 10, 11]). In
this paper we investigate the action of periodic Fourier multipliers and periodic
pseudo-differential operators from Holder spaces into Besov spaces. Our work is
closely related with a classical result by Marcinkiewicz: if (σ(ξ))ξ∈Z is a sequence
satisfying the following condition, now known as variational Marcinkiewicz con-
dition:
kσkL∞(Z) + sup
j≥0 X2j≤ξ≤2j+1 σ(ξ + 1) − σ(ξ) < ∞,
(1.2)
Date: Received: xxxxxx; Revised: yyyyyy; Accepted: zzzzzz.
1 Universidad de los Andes, Mathematics Department, Bogot´a - Colombia.
Key words and phrases. Besov spaces, Fourier transform, Bernstein's theorem, Fourier
series, Toroidal pseudo-differential operators.
1
2
DUV ´AN CARDONA S ´ANCHEZ
p,q(T) into Br
then Op(σ) : Lp(T) → Lp(T) is a bounded operator for all 1 < p < ∞. Here one
may consider ∆σ(·) = σ(· + 1) − σ(·) as the first derivative of σ. As a particular
case of Theorem 4.2 in [4], every operator Op(σ) satisfying (1.2) is a bounded
operator from Br
p,q(T) for all 1 < p < ∞, r ∈ R and 1 ≤ q ≤ ∞. We
observe that, by Corollary 4.3 in [4], for every r ∈ (0, 1) there exists a Fourier
multiplier Op(σ) with σ satisfying (1.2), but with the property that Op(σ) is not
a Fourier multiplier from Br
∞,∞(T). In order to get, in particular,
boundedness of periodic Fourier multipliers on Holder spaces, we reformulate the
variational Marcinkiewicz condition by imposing the following inequality on the
symbol:
∞,∞(T) into Br
σ(ξ) ≤ Cξ−ρ,
(1.3)
uniformly on ξ 6= 0, for some 0 ≤ ρ ≤ 1. Later, by using estimates on Fourier
multipliers, we deduce the boundedness of operators with symbols σ(x, ξ) of finite
regularity on x. More precisely, symbols satisfying inequalities of the type
ξ ∂β
∆α
x σ(x, ξ) ≤ Cβξ−ρ−α,
(1.4)
for α ≤ l1,β ≤ l2, li < ∞. We note that, condition (1.4) is related with the
Hormander class of symbols on the torus proposed by Ruzhansky and Turunen
in [29]. In Section 3 we show that, under suitable conditions on the set of in-
dices p, q, r, s and ρ, the ρ-condition (1.4) implies the boundedness of Op(σ(·))
from Bs
p,q(T), then we extend these results to the case of pseudo-
differential operators on the torus. We end Section 3 with a discussion of our
main results and some applications.
∞,∞(T) into Br
Finally, let us give some references on the topic we use along this paper. The
boundedness of Fourier multipliers in Lp-spaces, Holder spaces and Besov spaces
has been considered by many authors for a long time.
In the general case of
Compact Lie groups we refer the reader to the works of Alexopoulos, Anker,
Coifman, Ruzhansky, Turunen and Wirth [2, 3, 18, 29, 30, 31, 32, 33]. The general
case of operator-valued Fourier multipliers on the torus has been investigated by
Arendt, Bu, Barraza, Denk, Hern´andez, and Nau in [4, 5, 6, 9, 10, 11]. Lp
and Holder estimates of periodic pseudo-differential operators can be found in
[12, 13, 14, 19] and [26]. The quantization process, L2-compactness, spectral
properties and Lp estimates of pseudo-differential operators on the circle S1 ≡ T
also can be found in the works of Delgado, Wong and Molahajloo [21, 24, 25, 26,
35]. Besov continuity of Fourier multipliers and pseudo-differential operators on
general compact Lie groups has been investigated by the author in [15].
2. preliminaries
We use the standard notation of pseudo-differential operators (see e.g.
[29]).
The Schwartz space S(Zn) denote the space of functions φ : Zn → C such that
(2.1)
∀M ∈ R,∃CM > 0, φ(ξ) ≤ CMhξiM ,
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
3
where hξi = (1 + ξ2)
C ∞(Tn) by
1
2 . The toroidal Fourier transform is defined for any f ∈
where dx is the Haar measure on the n-torus Tn = [0, 2π)n. The inversion formula
is given by
e−ihx,ξif (x)dx, ξ ∈ Zn,
(F f )(ξ) := bf (ξ) =ZTn
f (x) = Xξ∈Zn
eihx,ξibu(ξ), x ∈ Tn.
We now take up the Holder space Λs, 0 < s < 1. According to the usual
definition, a function f belongs to Λs if there exists a constant A so that f (x) ≤
A almost every where and
fΛs := sup
x,y
f (x − y) − f (x)
ys
≤ A.
(2.2)
We introduce the Besov spaces on the torus using the periodic Fourier transform
as follow. Let r ∈ R, 0 ≤ q < ∞ and 0 < p ≤ ∞. If f is a measurable function
on T, we say that f ∈ Br
p,q(T) if f satisfies
If q = ∞, Br
p,∞(T) consists of those functions f satisfying
kfkBr
p,q :=
∞Xm=0
kfkBr
p,∞ := sup
m∈N
2mrqk X2m≤ξ<2m+1
2mrk X2m≤ξ<2m+1
< ∞.
1
q
Lp(T)
eixξbf (ξ)kq
eixξbf (ξ)kLp(T) < ∞.
In the case of p = q = ∞ and 0 < r < 1 we obtain Br
Banach spaces together with the norm
∞,∞(T) = Λr(T), these are
kfkΛr = fΛr + sup
x∈T f (x).
Similarly to Besov spaces one defines the Triebel-Lizorkin spaces as follows. If
r ∈ R, 0 < p ≤ ∞ 0 < q < ∞, the Triebel-Lizorkin space F r
p,q(T) consists of
those functions satisfying
kfkF r
p,q(T) :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xs=0
2srq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
X2s≤ξ<2s+1
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
eixξbf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
with a similar modification as in Besov spaces in the case q = ∞. An interesting
property regarding Besov spaces and Triebel-Lizorkin spaces is that Br
p,p for
all 0 < p < ∞. Now, We introduce some classes of pseudo-differential operators.
The periodic Hormander class Sm
ρ,δ(Tn × Rn), 0 ≤ ρ, δ ≤ 1, consists of those
functions a(x, ξ) which are smooth in (x, ξ) ∈ Tn × Rn and which satisfy toroidal
symbols inequalities
p,p = F r
x ∂α
∂β
ξ a(x, ξ) ≤ Cα,βhξim−ρα+δβ.
(2.6)
(2.3)
(2.4)
(2.5)
4
DUV ´AN CARDONA S ´ANCHEZ
Symbols in Sm
which are 1-periodic in x. If a(x, ξ) ∈ Sm
differential operator is defined by
ρ,δ(Tn × Rn) are symbols in Sm
ρ,δ(Rn × Rn) (see [29]) of order m
ρ,δ(Tn × Rn), the corresponding pseudo-
a(X, Dx)u(x) =ZTnZRn
ei2πhx−y,ξia(x, ξ)u(y)dξdy.
(2.7)
(2.8)
(2.9)
The set Sm
smooth in x for all ξ ∈ Zn and which satisfy
ρ,δ(Tn × Zn), 0 ≤ ρ, δ ≤ 1, consists of those functions a(x, ξ) which are
∀α, β ∈ Nn,∃Cα,β > 0, ∆α
ξ ∂β
x a(x, ξ) ≤ Cα,βhξim−ρα+δβ.
The operator ∆ is the difference operator defined in [29]. The toroidal operator
with symbol a(x, ξ) is defined as
a(x, Dx)u(x) = Xξ∈Zn
ei2πhx,ξia(x, ξ)bu(ξ), u ∈ C ∞(Tn).
ρ,δ(Tn×Zn) (resp. Sm
ρ,δ(Tn × Zn), (resp. Ψm
The corresponding class of operators with symbols in Sm
ρ,δ(Tn×
ρ,δ(Tn × Rn)). There exists a
Rn)) will be denoted by Ψm
process to interpolate the second argument of symbols on Tn × Zn in a smooth
way to get a symbol defined on Tn × Rn.
Theorem 2.1. Let 0 ≤ δ ≤ 1, 0 < ρ ≤ 1. The symbol a ∈ Sm
only if there exists a Euclidean symbol a′ ∈ Sm
Moreover, we have
Ψm
ρ,δ(Tn × Zn) = Ψm
ρ,δ(Tn × Zn) if
ρ,δ(Tn × Rn) such that a = a′Tn×Zn.
ρ,δ(Tn × Rn).
Proof. The proof can be found in [29].
(cid:3)
The following results provide some properties about composition and invert-
ibility of periodic pseudo-differential operators. Proofs of these assertions can be
found in [29, 30].
Theorem 2.2. (Composition formula). Let 0 ≤ δ < ρ ≤ 1. The composition
τ (X, D)◦ σ(X, D) of two pseudo-differential operators with symbols τ ∈ Sl
ρ,δ(Tn×
Zn) and σ ∈ Sm
ρ,δ(Tn×Zn) is a pseudo-differential operator, and its toroidal symbol
ψ(x, ξ) has the following asymptotic expansion,
ψ(x, ξ) ≈Xγ≥0
1
γ!
∆γ
ξ τ (x, ξ) · D(γ)
x σ(x, ξ).
(2.10)
A pseudo-differential operator σ(x, ξ) ∈ Sm
ρ,δ is called elliptic, if for every M > 0,
there exists R > 0 such that σ(x, ξ) ≥ Rhξim if ξ ≥ M.
Theorem 2.3. (Parametrix existence). Let 0 ≤ δ < ρ ≤ 1. For every elliptic
pseudo-differential operators with symbol σ ∈ Sm
ρ,δ(Tn × Zn) there exists τ ∈
S−m
ρ,δ (Tn × Zn) such that
(2.11)
where, S, R are pseudo-differential operators with symbols in S−∞ = ∩mSm
ρ,δ.
σ(X, D) ◦ τ (X, D) = I + R, τ (x, D) ◦ σ(X, D) = I + S,
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
5
As a consequence of the Proposition 6 in [34] and Theorem 2.1, the continuity
property of pseudo-differential operators in the Holder spaces is contained in the
following theorems. First we consider the case of operators on Rn as follow.
Theorem 2.4. Suppose σ is a symbol in Sm
is a bounded mapping from Λs(Rn) into Λs−m(Rn) whenever m < s ≤ 1.
1,0(Rn×Rn). Then the operator σ(X, D)
Theorem 2.5. Suppose σ is a symbol in Sm
is a bounded mapping from Λs(Tn) into Λs−m(Tn) whenever m < s ≤ 1.
1,0(Tn × Zn). Then the operator Op(σ)
Our main results are analogues of the Theorem 2.5, but we consider symbols
with limited smoothness on the configuration variables (x, ξ).
3. Holder-Besov boundedness of periodic operators
3.1. Main results and proofs. In this section we present the proof of our
main results. Although all results in this paper are presented for the torus T1
only, extensions to the torus Tn are valid. First, we consider the Holder-Besov
boundedness of periodic Holder multipliers. Later we extend this result to the case
of pseudo-differential operators by considering the Sobolev embedding theorem.
This approach was used by Ruzhansky and Wirth [33], (see also [31] and [32]) in
order to get Lp multiplier theorems for non-invariant pseudo-differential operators
on compact Lie groups. We reserve the notaci´on A . B if there exists c > 0
independent of A and B such that A ≤ c · B.
Theorem 3.1. Let ρ ∈ [0, 1] and σ(ξ) be a symbol satisfying the ρ-condition.
Then, the corresponding Fourier multiplier Op(σ) : Bs
p,q(T) is a
bounded operator for all r + 1
2 − ρ < s ≤ 1, 0 < p ≤ ∞ and 0 < q < ∞. If we
assume r + 1
∞,∞ into
Br
∞,∞(T) → Br
2 − ρ ≤ s ≤ 1, we obtain the boundedness of Op(σ) from Bs
p,∞.
Proof. Let us consider f ∈ C ∞(T). In order to estimate the Besov norm of Op(σ)f
we use its dyadic decomposition. First we note that
Xξ∈Z−{0}
F (Op(σ)f )(ξ)2 =
∞Xm=0 X2m≤ξ<2m+1 F (Op(σ)f )(ξ)2.
(3.1)
6
DUV ´AN CARDONA S ´ANCHEZ
Now we estimate every dyadic decomposition as follow. If take h = 2π/3· 2m and
2m ≤ ξ ≤ 2m+1 we have e−iξh − 1 ≥ √3. Hence we get
X2m≤ξ<2m+1 F (Op(σ)f )(ξ)2 ≤ X2m≤ξ<2m+1 e−ihξ − 12F (Op(σ)f )(ξ)2
= X2m≤ξ<2m+1 e−ihξ − 12σ(ξ)F (f )(ξ)2
≤ X2m≤ξ<2m+1 e−ihξ − 12ξ−2ρF (f )(ξ)2
. X2m≤ξ<2m+1 e−ihξ − 122−2mρF (f )(ξ)2
≤ 2−2mρXξ∈Z e−ihξ − 12F (f )(ξ)2.
On the other hand, Fourier inversion formula guarantees that
f (t − h) − f (t) =Xξ∈Z
(e−iξh − 1)(F f )(ξ)eiξt.
(3.2)
By the Plancherel theorem we conclude that
Xξ∈Z e−ihξ − 12F (f )(ξ)2 = kf (· − h) − f (·)k2
L2(T) ≤ (
Hence
2π
3 · 2m )2skfk2
Λs(T).
X2m≤ξ<2m+1 F (Op(σ)f )(ξ)2 ≤ 2−2mρ(
2π
3 · 2m )2skfk2
Λs(T)
. 2−2m(ρ+s)kfk2
Λs.
(3.3)
(3.4)
By the Cauchy-Schwarz inequality, for all 0 < p ≤ ∞ we get
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X2m≤ξ<2m+1
F (Op(σ)f )(ξ)eixξ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
≤ 2π · X2m≤ξ<2m+1 F (Op(σ)f )(ξ)
≤ 2π · X2m≤ξ<2m+1 F (Op(σ)f )(ξ)2
. 2−m(ρ+s)+ 1
2 (m+1)kfkΛs.
1/2
1
2 (m+1)
2
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
7
Now, we consider the Besov-norm of Op(σ)f if 0 < p, q < ∞ : in fact, we have
kOp(σ)fkBr
p,q(T) :=
∞Xm=0
≤ ∞Xm=0
∞Xm=0
. (
2mrq(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X2m≤ξ<2m+1
q
eixξF ((Op(σ)f )(ξ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Λs! 1
2 q(m+1)kfkq
2 q(m+1))1/qkfkΛs.
q
2mrq2−mq(ρ+s)+ 1
2mrq2−mq(ρ+s)+ 1
1
q
Lp(T)
From the condition r + 1
2 − ρ < s ≤ 1 we obtain
∞Xm=0
2mrq2−mq(ρ+s)+ 1
2 q(m+1) = 2
1
2
2mq(r−ρ−s+ 1
2 ) < ∞.
(3.5)
∞Xm=0
Hence kOp(σ)fkBr
q < ∞. Now we consider the case q = ∞. In fact, if we assume r − ρ + 1
we have
p,q(T) . kfkΛs which shows the boundedness of Op(σ) when
2 ≤ s ≤ 1,
2mr(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X2m≤ξ<2m+1
2mr2−m(ρ+s)+ 1
eixξF ((Op(σ)f )(ξ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
2 (m+1)kfkΛs
kOp(σ)fkBr
p,∞(T) := sup
0≤m<∞
. sup
0≤m<∞
. kfkΛs.
With above inequality we end the proof.
Theorem 3.2. Let us consider 0 ≤ ρ ≤ 1 ≤ p < ∞, 0 < q < ∞ and r + 1
s ≤ 1. If σ(x, ξ) satisfies
(cid:3)
2 − ρ <
∂β
x σ(x, ξ) ≤ Cβξ−ρ, β ≤ [1/p] + 1, ξ 6= 0,
∞,∞(T) → Br
then the pseudo-differential operator Op(σ) : Bs
operator.
Proof. Let f ∈ C ∞(T). To prove this theorem we write
(3.6)
p,q(T) is a bounded
ei(x−y)ξσ(x, ξ)! f (y)dy
Op(σ)f (x) =Xξ∈Z
=ZT Xξ∈Z
eixξσ(x, ξ)bf (ξ) =ZT Xξ∈Z
eiyξσ(x, ξ)! f (x − y)dy.
κ(z, y) =Xξ∈Z
eiyξσ(z, ξ).
Hence, Op(σ)f (x) = (κ(x,·) ∗ f )(x), where
(3.7)
8
DUV ´AN CARDONA S ´ANCHEZ
Moreover, if we define Azf (x) = (κ(z,·) ∗ f )(x) for every z ∈ T, we have
For all 0 ≤ β ≤ [1/p] + 1 we have ∂β
we have
Axf (x) = Op(σ)f (x), x ∈ T.
z Azf (x) = Op(∂β
z σ(z,·))f (x). if 1 ≤ p < ∞
k X2m≤ξ<2m+1
By the Sobolev embedding theorem we have
p
dx
p
dx.
p
dx
≤ sup
eixξF (Op(σ)f )(ξ)kp
e−iyξOp(σ)f (y)dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Lp :=ZT(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eixξZT
X2m≤ξ<2m+1
e−iyξ(Ayf )(y)dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=ZT(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eixξZT
X2m≤ξ<2m+1
e−iyξ(Azf )(y)dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
z∈TZT(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eixξZT
X2m≤ξ<2m+1
e−iyξ(Azf )(y)dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
z∈TZT(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eixξZT
X2m≤ξ<2m+1
e−iyξ(Azf )(y)dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
. Xβ≤[1/p]+1ZTZT(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eixξZT
z X2m≤ξ<2m+1
z Azf ))(ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ Xβ≤[1/p]+1ZTZT(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
X2m≤ξ<2m+1
= Xβ≤[1/p]+1ZT k X2m≤ξ<2m+1
β≤[1/p]+1ZT k X2m≤ξ<2m+1
z Azf ))(ξ)kp
eixξF ((∂β
eixξF ((∂β
z Azf ))(ξ)kp
Lpdz
eixξF ((∂β
p
dx
. sup
dz dx
p
dxdz
Lpdz
∂β
p
sup
Hence,
k X2m≤ξ<2m+1
eixξF (Op(σ)f )(ξ)kLp
. sup
β≤[ 1
p ]+1ZT k X2m≤ξ<2m+1
eixξF ((Op(∂β
z σ(z,·))f ))(ξ)kp
Lpdz
1/p
.
Thus, considering 0 < q < ∞ we obtain
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
9
kOp(σ)fkBr
p,q(T)
=
∞Xm=0
.
∞Xm=0
.
∞Xm=0
.
∞Xm=0
2mrq
2mrq
2mrq(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2mrq
sup
sup
β≤[ 1
sup
β≤[ 1
eixξF ((Op(σ)f )(ξ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X2m≤ξ<2m+1
p ]+1ZT k X2m≤ξ<2m+1
p ]+1ZT
z∈T k X2m≤ξ<2m+1
p ]+1sup
z∈T k X2m≤ξ<2m+1
p,q(T) .
∞Xm=0
p ]+1,z∈TkOp(∂β
≤"
p ]+1,z∈TkOp(∂β
sup
β≤[ 1
2mrq
β≤[ 1
β≤[ 1
β≤[ 1
sup
sup
.
kOp(σ)fkBr
Hence, we can write (by using the Fatou's Lemma)
q
Lp(T)
1
q
eixξF ((Op(∂β
z σ(z,·))f ))(ξ)kp
eixξF ((Op(∂β
z σ(z,·))f ))(ξ)kp
eixξF ((Op(∂β
z σ(z,·))f ))(ξ)kp
1
q
1
q
q/p
Lpdz
q/p
Lpdz
q/p
Lp
1
q
.
sup
p ]+1,z∈Tk X2m≤ξ<2m+1
z σ(z,·))fkBr
p,q(T)
eixξF ((Op(∂β
z σ(z,·))f ))(ξ)kq
1
q
Lp
z σ(z,·))kB(Λs,Br
p,q)#kfkΛs.
With the last inequality we end the proof.
(cid:3)
Remark 3.3. In order to find connection of Holder-Besov estimates and Lp-
estimates, in the next theorem we endowed a Holder space of degree 0 < s < 1
with the norm
kfkBs
∞,∞,p := fΛs + kfkLp,
(3.8)
where 1 < p < ∞.
Theorem 3.4. Let 0 ≤ ρ ≤ 1, and σ(x, ξ) be a measurable function satisfying
ξ σ(x, ξ) ≤ Cβξ−ρ−α,
∞,∞,p(T) → Br
x ∆α
∂β
Then Op(σ) : Bs
s ≤ 1.
Proof. We use notation as in the proof of Theorem 3.2. If we consider the condi-
tion (3.9), in particular, we have
∞,∞,p(T) is a bounded operator for all r + 1
β ≤ [1/p] + 1,α ≤ 2. ξ 6= 0.
(3.9)
2 − ρ ≤
∂β
x σ(x, ξ) ≤ Cβξ−ρ,
β ≤ [1/p] + 1,α ≤ 2. ξ 6= 0.
(3.10)
10
DUV ´AN CARDONA S ´ANCHEZ
So, by Theorem 3.2, for every z ∈ T, the operator ∂β
Bs
∞,∞(T) → Br
Holder-norm of Op(σ) :
z σ(z,·)) :
∞,∞(T) extends to bounded operator. Next, we estimate the
z Az = Op(∂β
Op(σ)fΛr(T) = sup
= sup
x,h∈TOp(σ)f (x − h) − Op(σ)f (x)h−r
x,h∈TAx−hf (x − h) − Axf (x)h−r
x,h,z∈TAzf (x − h) − Azf (x)h−r
≤ sup
By using the Sobolev embedding Theorem we have that
x,h,z∈TAzf (x − h) − Azf (x)h−r
sup
≤ sup
z∈T ∂β
k sup
z (Azf (x − h) − Azf (x)) kLp(T)h−r
z∈T ∂β
sup
z (Azf (x − h) − Azf (x))h−r
x,h∈T∂β
sup
z (Azf (x − h) − Azf (x))h−r
sup
z∈T
≤ sup
x,h∈T Xβ≤[1/p]+1
x,h∈T Xβ≤[1/p]+1
≤ Xβ≤[1/p]+1
= Xβ≤[1/p]+1
≤ Xβ≤[1/p]+1
z∈T ∂β
sup
z (Azf )Λr
z∈T k∂β
sup
z (Az)kB(Λs,Λr)kfkΛs.
On the other hand, by Theorem 5.2 in [33] the operator Op(σ) is a Lp-bounded
operator for all 1 < p < ∞. Hence, kOp(σ)fkLp ≤ CkfkLp. With this in mind,
we conclude that
kOp(σ)fkBr
∞,∞,p(T) := Op(σ)fΛr(T) + kOp(σ)fkLp(T) . kfkBs
∞,∞,p.
(3.11)
Theorem 3.5. Let 0 < s < 1, 2 ≤ p < ∞ and 0 < q < ∞. If r + 1 − 2
and σ(ξ) satisfies the ρ-condition, then Op(σ) : Bs
linear operator. Moreover, if r +1− 2
∞,∞(T) → Br
p ≤ ρ ≤ 1, then Op(σ) : Bs
is a linear bounded operator.
Proof. First, we recall the Hardy-Littlewood inequality on the torus: If 2 ≤ p <
∞ then
(cid:3)
p < ρ ≤ 1
p,q(T) is a bounded
∞,∞(T) → Br
p,∞(T)
.
(3.12)
(1 + ξ)p−2bf (ξ)p!1/p
kfkLp(T) ≤ CpXξ∈Z
gm(x) = X2m≤ξ<2m+1
If we denote by gm(x) the function
eixξF (Op(σ)f )(ξ),
(3.13)
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
11
then gm = F −1[χ{2m≤ξ<2m+1} · F (Op(σ)f )(·)]. By (3.12) we have,
kgmkLp(T) ≤Cp · X2m≤ξ<2m+1
2mrqk X2m≤ξ<2m+1
therefore, for 0 < q < ∞ we obtain
∞Xm=0
(1 + ξ)p−2F (Op(σ)f )(ξ)p
1/p
,
q/p
.
Lp(T)
eixξF (Op(σ)f )(ξ)kq
∞Xm=0
∞Xm=0
∞Xm=0
2mrq X2m≤ξ<2m+1
2mrq X2m≤ξ<2m+1
2mrq(cid:2)2m(p−2)2−mρp(cid:3)q/p
.
.
(1 + ξ)p−2pF (Op(σ)f )(ξ)
2m(p−2)σ(ξ)F (f )(ξ)p
kbfkq
Lp(Z).
q/p
Considering that kbfkLp(Z) . kfkΛs for every 0 < s < 1 and 2 ≤ p < ∞ we get
∞Xm=0
p −mρqkfkq
Λs.
2mrqk X2m≤ξ<2m+1
eixξF (Op(σ)f )(ξ)kq
2mrq+m(p−2) q
∞Xm=0
Lp(T) .
(3.14)
Since r + 1 − 2
p < ρ we get
C =
∞Xm=0
2mrq+m(p−2) q
p −mρq < ∞.
So, kOp(σ)kBr
p,q . kfkΛs. Hence, we conclude the boundedness of
Op(σ) : Bs
∞,∞(T) → Br
p,q(T).
The proof of the boundedness of Op(σ) when q = ∞ is analogue.
(cid:3)
We extend Theorem 3.5 to case of non-invariant periodic operators as follows:
p < ρ ≤ 1.
Theorem 3.6. Let us consider 2 ≤ p < ∞, 0 < q < ∞ and r + 1 − 2
Let σ(x, ξ) be a symbol satisfying
∂β
x σ(x, ξ) ≤ Cβξ−ρ, β ≤ [1/p] + 1, ξ 6= 0.
Then Op(σ) is a bounded operator from Bs
∞,∞ into Br
p,q.
(3.15)
Proof. The proof of this theorem is similar to the proof of Theorem 3.2.
(cid:3)
Now, we prove results concerning Holder-Triebel boundedness of Fourier mul-
tipliers.
12
DUV ´AN CARDONA S ´ANCHEZ
Theorem 3.7. Let us consider 0 < q < ∞, 0 < p ≤ ∞, r < ρ ≤ 1 and 1
Then Op(σ) : Bs
satisfies the ρ-condition. If we assume r ≤ ρ ≤ 1, 1
Op(σ) : Bs
p,∞(T), is a bounded operator.
2 < s ≤ 1.
p,q(T) is a bounded operator if we consider that σ(ξ)
2 < s ≤ 1 and q = ∞ then
∞,∞(T) → F r
∞,∞(T) → F r
Proof. First we consider the case of 0 < q < ∞, 0 < p ≤ ∞ and r < ρ. By the
definition of Triebel-Lizorkin norm, we have
kOp(σ)fkF r
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
p,q(T) :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
eixξσ(ξ)bf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
∞Xm=0
X2m≤ξ<2m+1
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X2m≤ξ<2m+1 σ(ξ)bf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
∞Xm=0
X2m≤ξ<2m+1 σ(ξ)bf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
=
∞Xm=0
2mr−mρbf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ ∞Xm=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xξ∈Z
q!1/q
2mr−mρbf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤Xξ∈Z bf (ξ)"Xm=0
q!1/q
2qm(r−ρ)#1/q
1/q
.
.
∞Xm=0(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xξ∈Z
By using the Minkowski integral inequality (discrete version) we have
From the condition r < ρ and by using the Bernstein theorem (i.e kbfkL1(Z) .
2 < s ≤ 1) we have
kfkΛs, 1
kOp(σ)fkF r
p,q(T) . kfkΛs.
If q = ∞ and r ≤ ρ we observe that
kOp(σ)fkF r
p,q(T) := sup
m∈N
X2m≤ξ<2m+1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
eixξσ(ξ)bf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mr(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2mr X2m≤ξ<2m+1 m(ξ)bf(ξ)
2m(r−ρ)Xξ∈Z bf (ξ)
≤ sup
m∈N
m∈N
≤ sup
. kfkΛs.
(cid:3)
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
13
In order to get boundedness from Holder into Triebel-Lizorkin spaces, we
present the following lemma which is a generalization of the Bernstein Theo-
rem. (See [7, 8]). We recall the equivalence Λs(T) ≡ Bs
∞,∞(T) for the Holder
space of order s.
Lemma 3.8. Let 2/3 < p ≤ 2 and let sp = 1/p−1/2. Then, the Fourier transform
f 7→ F f from Λs(T) into Lp(T) is a bounded operator for all s, sp < s < 1.
Theorem 3.9. Let 0 ≤ ρ ≤ 1 < α ≤ 2, sα = 1
satisfies the ρ-condition, then Op(σ) : Bs
for all 0 < p ≤ ∞, 0 < q < ∞ and sα < s < 1. Moreover, if r + 1 − 1
operator Op(σ) : Bs
α < ρ ≤ 1. If σ(ξ)
p,q(T) is a bounded operator
α ≤ ρ, the
2, and r + 1− 1
∞,∞(T) → F r
α − 1
p,∞(T) is bounded.
∞,∞(T) → F r
Proof. From the proof of Theorem 3.7 we have
1/q
.
On the other hand, if 1 < α ≤ 2 and 1/α + 1/α′ = 1, by using the Holder
inequality we obtain
1/q
∞Xm=0
kOp(σ)fkF r
X2m≤ξ<2m+1 σ(ξ)bf(ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
p,q(T) ≤
X2m≤ξ<2m+1 σ(ξ)bf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
∞Xm=0
2mq(r−ρ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤
α X2m≤ξ<2m+1
Xξ∈Z bf (ξ)α! 1
∞Xm=0
≤ ∞Xm=0
q!1/q
2mq(r−ρ)(cid:12)(cid:12)(cid:12)(2
α′ )(cid:12)(cid:12)(cid:12)
kbfkLα(Z)
α′ )!1/q
. ∞Xm=0
kbfkLα(Z).
From Lemma 3.8 and the condition r + 1 − 1
α < ρ ≤ 1 we claim
1
2mq(r−ρ+ 1
α′(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
m+1
1
1/q
for all sα < s < 1. By a similar argument, we may prove
kOp(σ)fkF r
p,q(T) . kfkΛs
kOp(σ)fkF r
p,∞(T) . sup
m∈N
2m(r−ρ+1/α′)kfkΛs.
Hence,
for all sα < s < 1, 0 < p ≤ ∞ and r + 1 − 1
α ≤ ρ.
kOp(σ)fkF r
p,∞(T) . kfkΛs
(cid:3)
14
DUV ´AN CARDONA S ´ANCHEZ
Theorem 3.10. Let us consider the periodic pseudo-differential operator Op(σ)
with the symbol σ(x, ξ) satisfying
∂α
x σ(x, ξ) ≤ Cαξ−ρ, α ≤ [1/q] + 1, ξ 6= 0.
If 1 ≤ q < ∞, 0 < p ≤ ∞, r < ρ ≤ 1 and 1
Λs into F r
and r + 1 − 1
α < ρ ≤ 1, Op(σ) : Λs → F r
sα < s < 1.
p,q. Also, if we assume 0 ≤ ρ ≤ 1 < α ≤ 2, sα = 1
2 < s ≤ 1, then Op(σ) is bounded from
2 , 0 < p ≤ ∞
p,q is a bounded linear operator for all
α − 1
Proof. If 1 ≤ q < ∞, by the Sobolev embedding theorem we write,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
X2m≤ξ<2m+1
q
eixξF (Op(σ)f )(ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
eixξF (Op(σ(z,·))f )(ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eixξF (Op(∂α
≤ sup
z∈T(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
X2m≤ξ<2m+1
. Xα≤[1/q]+1ZT X2m≤ξ<2m+1
α≤[1/q]+1ZT X2m≤ξ<2m+1
. sup
z σ(z,·))f )(ξ)qdz
eixξF (Op(∂α
z σ(z,·))f )(ξ)qdz
From this inequality we deduce that
kOp(σ)fkF r
p,q(T)
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xm=0
.(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xm=0
.(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xm=0
. sup
α≤[1/q]+1
sup
eixξF (Op(∂α
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
X2m≤ξ<2m+1
2mrqZT
α≤[1/q]+1 X2m≤ξ<2m+1
z∈T X2m≤ξ<2m+1
2mrq X2m≤ξ<2m+1
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
eixξF −1[σ(x, ξ)bf ](ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
z σ(z,·))f )(ξ)qdz
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
z σ(z,·))f )(ξ)q
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
z σ(z,·))f )(ξ)q
z∈T(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
eixξF (Op(∂α
eixξF (Op(∂α
∞Xm=0
α≤[1/q]+1
2mrq
sup
sup
sup
.
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
15
Hence we get (by using the Fatou's Lemma)
kOp(σ)fkF r
p,q(T)
sup
α≤[1/q]+1,z∈T
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xm=0
α≤[1/q]+1,z∈TkOp(∂α
α≤[1/q]+1,z∈TkOp(∂α
sup
sup
.
.
≤
2mrq X2m≤ξ<2m+1
p,q(T)
z σ(z,·))fkF r
z σ(z,·))kB(Λs,F r
p,q)kfkΛs.
eixξF (Op(∂α
1/q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(T)
z σ(z,·))f )(ξ)q
So, by the last inequality, Theorem 3.7 and Theorem 3.9, we deduce the Bound-
edness of Op(σ(·,·)) from Λs into F r
p,q in the following cases:
• 1 ≤ q < ∞, 0 < p ≤ ∞, r < ρ, 1
• 0 ≤ ρ ≤ 1, 1 < α ≤ 2, 0 < p ≤ ∞, r + 1 − 1
2 < s ≤ 1.
α < ρ and sα < s < 1.
(cid:3)
Theorem 3.11. Let us consider Op(σ) be a Fourier multiplier with symbol sat-
isfying the ρ−condition. Then Op(σ) : Bs
p,q(T), 0 < p ≤ ∞,
1 < q < ∞ is a bounded operator if r + 1
2 − ρ ≤ s ≤ 1, then
Op(σ) : Bs
2 − ρ < s ≤ 1. If r + 1
∞,∞(T) → F r
p,∞(T) is continuous.
∞,∞(T) → F r
Proof. From the proof of Theorem 3.1 we have
Hence
2π
3 · 2m )2skfk2
Λs(T)
(3.16)
(3.17)
. 2−2m(ρ+s)kfk2
Λs.
X2m≤ξ<2m+1 F (Op(σ)f )(ξ)2 ≤ 2−2mρ(
X2m≤ξ<2m+1 σ(ξ)bf (ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
∞Xm=0
≤
X2m≤ξ<2m+1 σ(ξ)bf(ξ)2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2mrq(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
∞Xm=0
2mrq2−mq(ρ+s)2q(m+1)/2!1/q
. ∞Xm=0
1/q
q/2
kfkΛs.
1/q
2q(m+1)/2
From the condition r − ρ + 1
2 < s we deduce the boundedness of Op(σ), in fact
kOp(σ)fkF r
p,q(T) ≤ ∞Xm=0
2mrq2−mq(ρ+s)+q(m+1)/2!1/q
kfkΛs.
16
DUV ´AN CARDONA S ´ANCHEZ
A similar proof is valid for q = ∞ and r − ρ + 1
Remark 3.12. We observe that similar extensions that we give here of the Theorem
3.1 to the non-invariant case of pseudo-differential operators can be obtained if
in place of Br
p,q(T) we write F r
2 ≤ s.
p,q(T).
(cid:3)
We end this section with the following theorem on boundedness of periodic
pseudo-differential operators on Holder spaces.
Theorem 3.13. Let 2/3 < p ≤ 2, sp = 1/p − 1/2, 0 < r < 1 and sp < s < 1. If
r + 1
∞,∞(T)
is a bounded Fourier multiplier.
q ≤ ρ and σ(ξ) satisfies the ρ-condition, then Op(σ) : Bs
∞,∞(T) → Br
Proof. First we consider the case where σ depends only on the Fourier variable
ξ. So we get for s ≥ 0
kOp(σ)fkBr
eixξF (Op(σ)f )(ξ)kL∞(T)
∞,∞ = sup
s∈N
2srk X2s≤ξ<2s+1
2sr X2s≤ξ<2s+1 F (Op(σ)f )(ξ)
2sr X2s≤ξ<2s+1 bf (ξ)σ(ξ)
≤ sup
s∈N
= sup
s∈N
By Holder inequality we obtain
kOp(σ)fkBr
∞,∞ . sup
1/q
1/p X2s≤ξ<2s+1 σ(ξ)q2srq
1/q
s∈N X2s≤ξ<2s+1 f (ξ)p
s∈N kbfkLp(Z) X2s≤ξ<2s+1hξi−ρq2srq
2−sρq2srq
s∈N kbfkLp(Z) X2s≤ξ<2s+1
s∈N kbfkLp(Z)(cid:0)2sq(r−ρ)+s(cid:1)1/q
.
1/q
. sup
. sup
. sup
By Lemma 3.8 we have k f (ξ)kLp(Z) . kfkBs
we get,
∞,∞ for sp < s < 1. Since r + 1
q ≤ ρ
kOp(σ)fkBr
which proves the boundedness of Op(σ).
∞,∞ . kfkBs
∞,∞
(cid:3)
3.2. Remarks and examples. There exists a connection between the Lp bound-
edness of Fourier multipliers on compact Lie groups and its continuity on Besov
spaces. This fact was proved by the author in Theorem 1.2 of [15]. In fact, the
Lie group structure of the torus T implies that every periodic Fourier multiplier
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
17
p1,q into Br
bounded from Lp1 into Lp2 is bounded from Br
p2,q, r ∈ R and 0 < q ≤ ∞.
Since, in general, the boundedness of Fourier multipliers satisfying the ρ-condition
-- or pseudo-differential operators with symbols satisfying Hormander conditions
but with limited regularity -- fails for pi = ∞, we have concentrate our atten-
tion to this case in the preceding subsection, in order to give boundedness of
multipliers -- and of pseudo-differential operators -- in Holder spaces Λr ≡ Br
∞,∞.
Remark 3.14. With the discussion above in mind, periodic Fourier multipliers
with symbol σ(ξ) satisfying the variational Marcinkiewicz condition:
kσkL∞(Z) + sup
j≥0 X2j≤ξ≤2j+1 σ(ξ + 1) − σ(ξ) < ∞,
(3.18)
are bounded from Lp(T), 1 < p < ∞ and hence these operators are bounded on
every Besov space Br
p,q(T) but, its boundedness on Holder spaces fails (Corollary
4.3 of [4]).
It is important to mention that every Fourier multiplier satisfying
the ρ-condition (1.3) with 0 < ρ ≤ 1 also satisfies the variational Marcinkiewicz
condition and, as a consequence, these operators are bounded on Lp, 1 < p < ∞
and on every Besov space Br
p,q, r ∈ R, 1 < p < ∞ and 0 < q ≤ ∞.
∞,∞ into Besov spaces Br
Remark 3.15. Theorems 3.1, 3.5, 3.7, 3.9, 3.11 and 3.13 give boundedness of
Fourier multipliers from Holder spaces Bs
p,q or spaces of
Triebel Lizorkin F r
p,q. Theorem 3.1 shows a dependence of the parameters ρ, r and
s. Nevertheless, as a consequence of the Hardy-Littlewood inequality, Theorem 3.5
relaxes this type of conditions for 2 ≤ p < ∞ by imposing restrictions on ρ, r and
p. On the other hand, for 2
3 < p ≤ 2, Theorem 3.13 only consider a dependence on
the parameters r, ρ and q. These theorems have been proved by using non-trivial
modifications of the proof of the Bernstein Theorem [7]. Theorems 3.4, 3.2 and
3.10 have been proved using the Sobolev embedding theorem as a fundamental
tool.
Remark 3.16. Notice that the results of this section illustrate a very important
connection between Lp boundedness and Holder boundedness. Indeed, by observ-
ing the proof of Theorem 3.4, the condition on the symbol
∂β
x σ(x, ξ) ≤ Cβξ−ρ,
(3.19)
guarantees the boundedness of Op(σ), from Bs
2 ≤ ρ ≤ 1.
For the Lp boundedness, 1 < p < ∞ of Op(σ) (see Theorem 3.7 of [20]) it is
sufficient to consider the following condition
β ≤ [1/p] + 1. ξ 6= 0,
∞,∞ for r−s+ 1
∞,∞ into Br
for ρ = 2(1− ρ) 1
p − 1
ξ ∂β
∆α
x σ(x, ξ) ≤ Cβξ−ρ−ρα,
β ≤ [1/p] + 1,α ≤ 2. ξ 6= 0,
(3.20)
2, 0 ≤ ρ ≤ 1. Thus, if we consider the inequality (3.20), with
r − s +
1
p −
1
1
2 ≤ ρ = 2(1 − ρ)(cid:12)(cid:12)(cid:12)(cid:12)
2(cid:12)(cid:12)(cid:12)(cid:12) ≤ 1,
we obtain the boundedness of Op(σ) from Bs
∞,∞,p into Br
∞,∞,p.
We end this section with the following examples on operators satisfying the
ρ-condition and on elliptic regularity in Holder spaces.
18
DUV ´AN CARDONA S ´ANCHEZ
Example 3.17. Let X be a left-invariant real vector field on the torus T. By
Corollary 2.7 of [33], there exists an exceptional set C ⊂ iR, such that for all
c /∈ C , the operator X +c is invertible with inverse satisfying the ρ-condition with
ρ = 0. By Theorem 3.1, we have for r + 1
2 < s ≤ 1, 0 < p ≤ ∞ and 0 < q < ∞ :
(3.21)
kfkBr
p,q ≤ Ck(X + c)fkBs
∞,∞.
On the other hand, if we consider r + 1
2 ≤ s ≤ 1, we obtain the estimate,
kfkBr
p,∞ ≤ Ck(X + c)fkBs
∞,∞.
(3.22)
Analogous estimates may be obtained if we apply Theorem 3.5, Theorem 3.7 or
Theorem 3.13. Similar results also can be considered if we replace X by the
partial Riesz transform R = (−LT)− 1
2 ◦ X, of some negative power of the Laplace
operator LT.
Example 3.18. Let 0 < r < 1 and f ∈ Br
differential problem
∞,∞(T). Consider the toroidal pseudo-
Op(σ)u = f,
(3.23)
where Op(σ) is an elliptic operator with symbol σ(x, ξ) ∈ Sm
ρ,δ, m > 0, (in partic-
ular, Op(σ) can be an elliptic differential operator of the formP0≤i≤m ai(x)∂i
x,
m ≥ 1). By the existence of parametrices for elliptic operators, (see Theorems
2.2 and 2.3), there exists q ∈ S−m
ρ,δ and r ∈ S−∞ such that
(3.24)
Therefore, Op(q) ◦ Op(σ)u = u + Op(r)u = Op(q)f. By using Theorem 3.4, we
have for 0 < r ≤ s + m − 1
Op(q) ◦ Op(σ) = I + Op(r).
2, 0 < s < 1,
kOp(q)fkBr
∞,∞,p ≤ CkfkBs
∞,∞,p,
and considering that the operator Op(r) is a smoothing operator we get u ∈
Br
∞,∞,p. In conclusion, under the pseudo-differential problem considered, if f ∈
Bs
∞,∞,p(T) then u ∈ Br
∞,∞,p(T). A similar a priori estimate can be obtained if we
consider Theorem 3.10.
Acknowledgments. I would like to thank the anonymous referee for his remarks
which helped to improve the manuscript. The author is indebted with Alexander
Cardona for helpful comments on an earlier draft of this paper. This project was
partially supported by Universidad de los Andes, Mathematics Department.
References
1. Agranovich, M. S.:Spectral properties of elliptic pseudodifferential operators on a closed
curve Funct. Anal. Appl. Vol 13. pp 279-281.(1971)
2. Alexopoulos, G.: Spectral multipliers on Lie groups of polynomial growth. Proc. Amer.
Math. Soc. 120, 973 -- 979, (1994)
3. Anker, J.: Lp Fourier multipliers on Riemannian symmetric spaces of the noncompact type.
Ann. of Math., 132, 597 -- 628 (1990)
4. Arendt, W., Bu, S.: Operator-valued Fourier multipliers on periodic Besov spaces and
applications, Proceedings of the Edinburgh Mathematical Society, 47(1), 15 -- 33, (2004)
H OLDER-BESOV BOUNDEDNESS FOR PSEUDO-DIFFERENTIAL OPERATORS
19
5. Barraza, B., Gonz´alez, I., Hern´andez, J.: Operator-valued Fourier multipliers on periodic
Besov spaces. arXiv:1504.04408
6. Barraza Martnez, B., Denk, R., Hern´andez Monz´on, J., Nau, T.: Generation of Semigroups
for Vector-Valued Pseudodifferential Operators on the Torus. J. Fourier Anal. Appl. 22(4),
823 -- 853 (2016)
7. Bernstein, S.: Sur la convergence absolue des s´eries trigonom´etriques. Comptes Rendum
Hebdomadaires des S´eances de l'Academie des Sciences, Paris, 158, 1661 -- 1663, (1914)
8. Bloom, W. R.: Bernstein's inequality for locally compact Abelian groups. Journal the
Australian mathematical society 17, 88 -- 101 (1974)
9. Bu, S., and Kim, J.: Operator-valued Fourier multiplier theorems on Lpspaces Tn. Archiv.
der Math. 82 , 404 -- 414, (2004)
10. Bu, S., Kim, J.: Operator-valued Fourier Multipliers on Periodic Triebel Spaces, Acta
Mathematica Sinica, English Series, Vol. 21, No. 5, 1049 -- 1056 (2005)
11. Bu, S., Kim, J.: A note on operator-valued Fourier multipliers on Besov spaces, Math.
Nachr. 278, No. 14 1659 -- 1664 (2005)
12. Cardona, D.: Estimativos L2 para una clase de operadores pseudodiferenciales definidos en
el toro Rev. Integr. Temas Mat. 31(2), (2013) 147 -- 152.
13. Cardona, D.: Weak type (1, 1) bounds for a class of periodic pseudo-differential operators.
J. Pseudo-Differ. Oper. Appl., 5(4), (2014) 507-515.
14. Cardona, D.: Holder estimates for pseudo-differential operators on T 1. J. Pseudo-Differ.
Oper. Appl. 5 (4), 517 -- 525 (2014)
15. Cardona, D.: Besov continuity for Multipliers defined on compact Lie groups. Palest. J.
Math. Vol. 5(2) 35 -- 44 (2016)
16. Coifman, R., de Guzman, M.: Singular integrals and multipliers on homogeneous spaces,
Rev. Un. Mat. Argentina, 25 137 -- 143 (1970)
17. Coifman, R., Weiss, G.: Analyse Harmonique Non-Commutative sur Certains Espaces
Homogenes, Lecture Notes in Mathematics, Vol. 242, Springer-Verlag, Berlin Heidelberg
New York, 1971.
18. Coifman, R., Weiss, G.: Central multiplier theorems for compact Lie groups. Bull. Am.
Math. Soc., 80 124 -- 126, (1973)
19. Delgado, J.: Lp bounds for pseudo-differential operators on the torus Operators Theory,
advances and applications. 231, 103 -- 116 (2012)
20. Delgado, J. Ruzhansky, M.: Lp-bounds for pseudo-differential operators on compact Lie
groups. arXiv:1605.07027
21. Delgado, J., Wong, M.W.: Lp-nuclear pseudo-differential operators on Z and S1., Proc.
Amer. Math. Soc., 141 (2013) no. 11, 3935 -- 3942.
22. Fefferman, C.: Lp bounds for pseudo-differential operators, Israel J. Math. 14, 413 -- 417
(1973)
23. Hormander, L.: Estimates for translation invariant operators in Lp spaces. Acta Math.,
104, 93 -- 140 (1960)
24. Molahajloo, S.: A Characterization of Compact Pseudo-Differential Operators on S1, in
Pseudo-Differential Operators: Analysis, Applications and Computations 213, Birkha user,
Basel, 25 -- 29. 2011
25. Molahajloo, S., Wong, M. W.: Ellipticity, Fredholmness and Spectral Invariance of Pseudo-
Differential Operators on S1, J. Pseudo-Differ. Oper. Appl. 1 (2), 183 -- 205, (2010)
26. Molahajloo, S., Wong, M.W.: Pseudo-differential Operators on S1. New developments in
pseudo-differential operators, Eds. L. Rodino and M.W. Wong. 297 -- 306, (2008)
27. Ruzhansky, M., Turunen, V.: On the Fourier analysis of operators on the torus, Modern
trends in pseudo-differential operators, 87-105, Oper. Theory Adv. Appl., 172, Birkhauser,
Basel, 2007.
28. Ruzhansky M., Turunen V.: On the toroidal quantization of periodic pseudo-differential
operators, Numerical Functional Analysis and Optimization, 30, 1098 -- 1124 (2009)
20
DUV ´AN CARDONA S ´ANCHEZ
29. Ruzhansky, M., Turunen, V.:Pseudo-differential Operators and Symmetries: Background
Analysis and Advanced Topics Birkhauser-Verlag, Basel, (2010)
30. Ruzhansky, M., Turunen, V.: Quantization of Pseudo-Differential Operators on the Torus
J Fourier Annal Appl. Birkhauser Verlag, Basel, 16, 943 -- 982 (2010)
31. Ruzhansky, M., Turunen, V.: Global quantization of pseudo-differential operators on com-
pact Lie groups, SU(2), 3-sphere, and homogeneous spaces Int. Math. Res. Not., 11, 2439 --
2496 (2013) http://dx.doi.org/10.1093/imrn/rns122
32. Ruzhansky M., Turunen V., Wirth J.: Hormander class of pseudo-differential operators on
compact Lie groups and global hypoellipticity, J. Fourier Anal. Appl., 20, 476 -- 499 (2014)
33. Ruzhansky, M. Wirth, J.: Lp Fourier multipliers on compact Lie groups, Mathematische
Zeitschrift, 1432 -- 1823 (2015).
34. Stein, E.: Harmonic analysis: real-variable methods, orthogonality, and oscillatory inte-
grals. Princeton University Press, 1993.
35. Wong, M. W.: Discrete Fourier Analysis, Birkhauser, 2011.
1 Department of Mathematics, Universidad de los Andes, Colombia.
E-mail address: [email protected]; [email protected]
|
1303.0322 | 1 | 1303 | 2013-03-02T00:04:16 | Strong mixing measures for linear operators and frequent hypercyclicity | [
"math.FA",
"math.DS"
] | We construct strongly mixing invariant measures with full support for operators on F-spaces which satisfy the Frequent Hypercyclicity Criterion. For unilateral backward shifts on sequence spaces, a slight modification shows that one can even obtain exact invariant measures. | math.FA | math |
Strong mixing measures for linear operators and
frequent hypercyclicity
M. Murillo-Arcila and A. Peris∗
Abstract
We construct strongly mixing invariant measures with full support for operators
on F -spaces which satisfy the Frequent Hypercyclicity Criterion. For unilateral
backward shifts on sequence spaces, a slight modification shows that one can even
obtain exact invariant measures.1
1
Introduction
We recall that an operator T on a topological vector space X is called hypercyclic if
there is a vector x in X such that its orbit Orb(x, T ) = {x, T x, T 2x, . . . } is dense in
X. The recent books [5] and [15] contain the theory and most of the recent advances on
hypercyclicity and linear dynamics, especially in topological dynamics.
Here we are concerned with measure theoretic properties. Let (X, B, µ) be a probabil-
ity space, where X is a topological space and B denotes the σ-algebra of Borel subsets of
X. We will say that a Borel probability measure µ has full support if for each non-empty
open set U ⊂ X we have µ(U) > 0. A measurable map T : (X, B, µ) → (X, B, µ) is
called a measure-preserving transformation if µ(T −1(A)) = µ(A) for all A ∈ B). T is said
to be strongly mixing with respect to µ if
µ(A ∩ T −n(B)) = µ(A)µ(B)
(A, B ∈ B),
lim
n→∞
reader is referred to [20, 10] for a detailed account on the above properties.
n=0 T −nB then either µ(A) = 0 or µ(A) = 1. The interested
and it is exact if given A ∈T∞
Ergodic theory was first used for the dynamics of linear operators by Rudnicki [18]
and Flytzanis [11]. During the last few years it has been given special attention thanks
to the work of Bayart and Grivaux [2, 3]. The papers [1, 4, 6, 9, 13, 19], for instance,
contain recent advances on the subject.
The concept of frequent hypercyclicity was introduced by Bayart and Grivaux [3]
inspired by Birkhoff's ergodic theorem. They also gave the first version of a Frequent
Hypercyclicity Criterion, although we will consider the formulation of Bonilla and Grosse-
Erdmann [8] for operators on separable F -spaces. Another (probabilistic) version of it
was given by Grivaux [12].
∗IUMPA, Universitat Polit`ecnica de Val`encia, Departament de Matem`atica Aplicada, Edifici 7A, 46022
Val`encia, Spain. e-mail: [email protected]
1Keywords: hypercyclic operators, strongly mixing measures
2010 MSC: 37A25, 47A16.
1
We derive under the hypothesis of Bonilla and Grosse-Erdmann a stronger result
by showing that a T -invariant mixing measure can be obtained. Recently, Bayart and
Matheron gave very general conditions expressed on eigenvector fields associated with
unimodular eigenvalues under which an operator T admits a T -invariant mixing measure
[6]. Actually, on the one hand our results can be deduced from [6] in the context of
complex Fr´echet spaces, and on the other hand we only need rather elementary tools.
From now on, T will be an operator defined on a separable F -space X.
2
Invariant measures and the frequent hypercyclicity
criterion
set F ⊂ {N, N + 1, N + 2, . . . }.
We recall that a series Pn xn in X converges unconditionally if it converges and, for any
0-neighbourhood U in X, there exists some N ∈ N such thatPn∈F xn ∈ U for every finite
We are now ready to present our main result. The idea behind the proof is to con-
struct a "model" probability space (Z, µ) and a (Borel) measurable map Φ : Z → X,
where Z ⊂ NZ is such that σ(Z) = Z for the Bernoulli shift σ(. . . , n−1, n0, n1, . . . ) =
(. . . , n0, n1, n2, . . . ), µ is a σ−1-invariant strongly mixing measure, Y := Φ(Z) is a T -
invariant dense subset of X, Φσ−1 = T Φ, and then the Borel probability measure µ on
X defined by µ(A) = µ(Φ−1(A)), A ∈ B(X), is T -invariant and strongly mixing. We will
use the slight generalization of the Frequent Hypercyclicity Criterion for operators given
in [15, Remark 9.10].
Theorem 1. Let T be an operator on a separable F -space X. If there is a dense subset
X0 of X and a sequence of maps Sn : X0 → X such that, for each x ∈ X0,
(i) P∞
(ii) P∞
n=0 T nx converges unconditionally,
n=0 Snx converges unconditionally, and
(iii) T nSnx = x and T mSnx = Sn−mx if n > m,
then there is a T -invariant strongly mixing Borel probability measure µ on X with full
support.
Proof. We suppose that X0 = {xn ; n ∈ N} with x1 = 0 and Sn0 = 0 for all n ∈ N. Let
(Un)n be a basis of balanced open 0-neighbourhoods in X such that Un+1 + Un+1 ⊂ Un,
n ∈ N. By (i) and (ii), there exists an increasing sequence of positive integers (Nn)n with
Nn+2 − Nn+1 > Nn+1 − Nn for all n ∈ N such that
Skxmk ∈ Un+1, if mk ≤ 2l, for Nl < k ≤ Nl+1, l ≥ n. (1)
T kxmk ∈ Un+1 and Xk>Nn
Actually, this is a consequence of the completeness of X and the fact that, for each
Xk>Nn
0-neighbourhood U and for all l ∈ N, there is N ∈ N such that Pk∈F T kx ∈ U and
Pk∈F Skx ∈ U for any finite subset F ⊂]N, +∞[ and for each x ∈ {x1, . . . , x2l}.
We define K =Qk∈Z Fk where
Fk = {1, . . . m} if Nm < k ≤ Nm+1, m ∈ N, and Fk = {1}, if k ≤ N1.
1.-The model probability space (Z, µ).
2
Let K(s) := σs(K), s ∈ Z, where σ : NZ → NZ is the backward shift. K(s) is a compact
space when endowed with the product topology inherited from NZ, s ∈ Z.
We consider in NZ the product measure µ = Nk∈Z µk, where µk({n}) = pn for all
n ∈ N and µk(N) = P∞
n=1 pn = 1, k ∈ Z. The values of pn ∈]0, 1[ are selected such that,
if
∞
pi!Nj+1−Nj
, j ∈ N, then
βj > 0.
Yj=1
βj := j
Xi=1
Let Z =Ss∈Z K(s). We have
µ(Z) ≥ µ(K) = Yk≤N1
µk({1})
∞
Yl=1
YNl<k≤Nl+1
µk({1, . . . , l})
1
= p2N1+1
βl)2 > 0.
(
∞
Yl=1
It is well-known [20] that µ is a σ−1-invariant strongly mixing Borel probability measure.
Since σ(Z) = Z and it has positive measure, then µ(Z) = 1.
2.-The map Φ.
Given s ∈ Z we define the map Φ : K(s) → X by
Φ((nk)k∈Z) =Xk<0
S−kxnk + xn0 +Xk>0
T kxnk.
(2)
Φ is well-defined since, given (nk)k∈Z ∈ K(s) and for l ≥ s, we have nk ≤ 2l if Nl <
k ≤ Nl+1, which shows the convergence of the series in (2) by (1).
Φ is also continuous.
Indeed, let (α(j))j be a sequence of elements of K(s) that
converges to α ∈ K(s) and fix any n ∈ N with n > s. We will find n0 ∈ N such that
Φ(α(j)) − Φ(α) ∈ Un for n ≥ n0. To do this, by definition of the topology in K(s) there
exists n0 ∈ N such that
α(j)k = αk
if
k ≤ Nn+1 and j ≥ n0.
By (1) we have
Φ(α(j)) − Φ(α) = Xk<−Nn+1
S−k(xα(j)k − xαk ) + Xk>Nn+1
T k(xα(j)k − xαk ) ∈ Un
for all j ≥ n0. This shows the continuity of Φ : K(s) → X for every s ∈ Z.
The map Φ is then well-defined on Z, and Φ : Z → X is measurable (i.e., Φ−1(A) ∈
B(Z) for every A ∈ B(X)).
3.-The measure µ on X.
subset of X because Φσ−1 = T Φ.
L(s) := Φ(K(s)) is compact in X, s ∈ Z, and Y := Ss∈Z L(s) is a T -invariant Borel
We then define in X the measure µ(A) = µ(Φ−1(A)) for all A ∈ B(X). Obviously,
µ is well-defined and it is a T -invariant strongly mixing Borel probability measure. The
proof is completed by showing that µ has full support. Given a non-empty open set U in
X, we pick n ∈ N satisfying xn + Un ⊂ U. Thus
µ(U) ≥ µ({x = xn + Xk>Nn
T kxmk + Xk>Nn
Skxmk ; mk ≤ 2l for Nl < k ≤ Nl+1, l ≥ n})
3
≥ µ0({n}) Y0<k≤Nn
µk({1})
∞
Yl=n
YNl<k≤Nl+1
µk({1, . . . , 2l})
1
> pnp2Nn
βl)2 > 0.
(
∞
Yl=n
As we mentioned in the Introduction, Theorem 1 can be deduced from [6, Corollary 1.3]
when dealing with operators on separable complex Fr´echet spaces. Indeed, the argument
of ´E. Matheron is the following (we thank S. Grivaux for letting us know about it):
Let T : X → X be an operator on a separable complex Fr´echet space X satisfying
the hypothesis of the Frequent Hypercyclicity Criterion given in Theorem 1, and suppose
that X0 = {xn ; n ∈ N}. We define the following family of continuous T-eigenvector fields
for T
Em(λ) =Xn≥0
λ−nT nxm +Xn∈N
λnSnxm, λ ∈ T, m ∈ N.
They span X since, for any functional x∗ that vanishes on Em(λ) for each λ ∈ T and m ∈
N, the equality hx∗, Em(λ)i = 0 for fixed m and for all λ ∈ T implies that hx∗, T nxmi = 0
for every n ≥ 0. Thus, hx∗, xmi = 0 for each m ∈ N, and by density x∗ = 0.
The previous Theorem can be applied to different classes of operators. A distinguished
one is the class of weighted shifts on sequence F -spaces.
By a sequence space we mean a topological vector space X which is continuously
included in ω, the countable product of the scalar field K. A sequence F -space is a
sequence space that is also an F -space. Given a sequence w = (wn)n of non-zero weights,
the associated unilateral (respectively, bilateral ) weighted backward shift Bw : KN → KN
is defined by Bw(x1, x2, . . . ) = (w2x2, w3x3, . . . ) (respectively, Bw : KZ → KZ is defined
by Bw(. . . , x−1, x0, x1, . . . ) = (. . . , w0x0, w1x1, w2x2, . . . )). When a sequence F -space X
is invariant under certain weighted backward shift T , then T is also continuous on X by
the closed graph theorem.
We refer the reader to, e.g., Chapter 4 of [15] for more details about hypercyclic and
chaotic weighted shifts on Fr´echet sequence spaces. In particular, Theorems 4.6 and 4.12
in [15] (we refer the reader to [14] for the original results) remain valid for F -spaces, and
a bilateral (respectively, unilateral) weighted backward shift T : X → X on a sequence
F -space X in which the canonical unit vectors (en)n∈Z (respectively, (en)n∈N) form an
unconditional basis is chaotic if, and only if, Pn∈Z en (respectively, Pn∈N en) converges
unconditionally.
Corollary 2. Let T : X → X be a chaotic bilateral weighted backward shift on a sequence
F -space X in which (en)n∈Z is an unconditional basis. Then there exists a T -invariant
strongly mixing Borel probability measure on X with full support.
Remark 3. The preceding result can be improved if T is a unilateral backward shift
operator on a sequence F -space.
In that case, there exists a T -invariant exact Borel
probability measure on X with full support.
Proof. Let M = {zn ; n ∈ N} be a countable dense set in K with z1 = 0. Let (Un)n be a
basis of balanced open 0-neighbourhoods in X such that Un+1 + Un+1 ⊂ Un, n ∈ N. Since
n=1 en converges unconditionally, so there exists an increasing sequence of
positive integers (Nn)n with Nn+2 − Nn+1 > Nn+1 − Nn for all n ∈ N such that
αkek ∈ Un+1, if αk ∈ {z1, . . . , z2m}, for Nm < k ≤ Nm+1, m ≥ n.
(3)
T is chaotic, P∞
Xk>Nn
4
We define K =Qk∈N Fk where
Fk = {z1, . . . zm} if Nm < k ≤ Nm+1, m ∈ N, and Fk = {z1}, if k ≤ N1.
Let K(s) := σs(K), s ≥ 0. K(s) is a compact space when endowed with the product
topology inherited from M N, s ≥ 0. We consider in M N the product measure µ =
n=1 pn = 1, k ∈ N. As
Nk∈N µk, where µk({zn}) = pn for all n ∈ N and µk(M) = P∞
before, we select the sequence (pn)n of positive numbers such that, if
βj = j
Xi=1
pi!Nj+1−Nj
, then
βj > 0.
∞
Yj=1
It is known [20, §4.12] that µ is a σ-invariant exact Borel probability measure. By setting
Z =Ss≥0 K(s), we have µ(Z) = 1.
Now we define the map Φ : K(s) → X given by
Φ((αk)k∈N) =
αkek.
∞
Xk=1
Φ is (well-defined and) continuous, s ≥ 0. We have that Φ : Z → X is measurable.
Borel subset of X.
L(s) := Φ(K(s)) is compact in X, s ≥ 0, and Y := Ss≥0 L(s) = Φ(Z) is a T -invariant
We then define on X the measure µ(A) = µ(Φ−1(A)) for all A ∈ B(X). As in
Theorem 1, we conclude that µ is well-defined, and now it is a T -invariant exact Borel
probability measure with full support.
Devaney chaos is therefore a sufficient condition for the existence of strongly mixing
measures within the framework of weighted shift operators on sequence F -spaces. For
some natural spaces it is even a characterization of this fact. For instance, F. Bayart
and I. Z. Ruzsa [7] recently proved that weighted shift operators on ℓp, 1 ≤ p < ∞, are
frequently hypercyclic if, and only if, they are Devaney chaotic. It turns out that this
is equivalent to the existence of an invariant strongly mixing Borel probability measure
with full support on ℓp. Also, for the space ω, every weighted shift operator is chaotic
[14]. In particular, for the unilateral case we obtain exact measures.
Example 4. Every unilateral weighted backward shift operator on ω = KN admits an
invariant exact Borel probability measure with full support on ω.
We finish the paper by mentioning that a continuous-time version of Theorem 1 can
be given by using the Frequent Hypercyclicity Criterion for C0-semigroups introduced in
[16]. This is part of a forthcoming paper [17].
Acknowledgements
This work was supported in part by MEC and FEDER, Project MTM2010-14909, and
by GV, Project PROMETEO/2008/101. The first author was also supported by a grant
from the FPU Program of MEC. We thank the referee whose detailed report led to an
improvement in the presentation of this work.
5
References
[1] C. Badea and S. Grivaux, Unimodular eigenvalues, uniformly distributed sequences
and linear dynamics, Adv. Math. 211 (2007), 766 -- 793.
[2] F. Bayart and S. Grivaux, Hypercyclicity and unimodular point spectrum, J. Funct.
Anal. 226 (2005), 281 -- 300.
[3] F. Bayart and S. Grivaux, Frequently hypercyclic operators, Trans. Amer. Math.
Soc. 358 (2006), 5083 -- 5117.
[4] F. Bayart and S. Grivaux, Invariant Gaussian measures for operators on Banach
spaces and linear dynamics, Proc. Lond. Math. Soc. (3) 94 (2007), 181 -- 210.
[5] F. Bayart and ´E. Matheron, Dynamics of linear operators, Cambridge University
Press, Cambridge, 2009.
[6] F. Bayart and ´E. Matheron, Mixing operators and small subsets of the circle, preprint
(arXiv:1112.1289v1).
[7] F. Bayart and I. Z. Ruzsa, Frequently hypercyclic weighted shifts, preprint.
[8] A. Bonilla and K.-G. Grosse-Erdmann, Frequently hypercyclic operators and vec-
tors, Ergodic Theory Dynam. Systems 27 (2007), 383 -- 404. Erratum: Ergodic Theory
Dynam. Systems 29 (2009), 1993 -- 1994.
[9] M. De la Rosa, L. Frerick, S. Grivaux and A. Peris, Frequent hypercyclicity, chaos,
and unconditional Schauder decompositions, Israel Journal of Mathematics 190
(2012), 389 -- 399.
[10] M. Einsiedler and T. Ward. Ergodic theory with a view towards number theory, volume
259 of Graduate Texts in Mathematics. Springer-Verlag London Ltd., London, 2011.
[11] E. Flytzanis, Unimodular eigenvalues and linear chaos in Hilbert spaces, Geom.
Funct. Anal. 5 (1995), 1 -- 13.
[12] S. Grivaux, A probabilistic version of the frequent hypercyclicity criterion, Studia
Math. 176 (2006), 279 -- 290.
[13] S. Grivaux, A new class of frequently hypercyclic operators, Indiana Univ. Math. J.
60 (2011), 1177 -- 1202.
[14] K.-G. Grosse-Erdmann, Hypercyclic and chaotic weighted shifts, Studia Math. 139
(2000), 47 -- 68.
[15] K.-G. Grosse-Erdmann and A. Peris Manguillot. Linear chaos. Universitext,
Springer-Verlag London Ltd., London, 2011.
[16] E. Mangino and A. Peris, Frequently hypercyclic semigroups, Studia Math. 202
(2011), 227 -- 242.
[17] M. Murillo and A. Peris, Strong mixing measures for C0-semigroups, in preparation.
6
[18] R. Rudnicki, Gaussian measure-preserving linear transformations, Univ. Iagel. Acta
Math. 30 (1993), 105 -- 112.
[19] R. Rudnicki, Chaoticity and invariant measures for a cell population model, J. Math.
Anal. Appl., 393 (2012), 151 -- 165.
[20] P. Walters, An introduction to ergodic theory, Springer, New York-Berlin, 1982.
7
|
1305.2497 | 1 | 1305 | 2013-05-11T11:18:31 | An embedding theorem for weighted Sobolev classes on a John domain: case of weights that are functions of a distance to a certain h-set | [
"math.FA"
] | Let $\Omega$ be a John domain, and let $\Gamma\subset \partial \Omega$ be an $h$-set. For some functions $h$ and some weight functions depending on distance from $\Gamma$, embedding theorems for a weighted Sobolev class is obtained. | math.FA | math | An embedding theorem for weighted Sobolev
classes on a John domain: case of weights that are
functions of a distance to a certain h-set
A.A. Vasil'eva∗
1 Introduction
Let Ω ⊂ Rd be a bounded domain (an open connected set), and let g, v : Ω → R+
be measurable functions. For each measurable vector-valued function ϕ : Ω → Rm,
ϕ = (ϕk)16k6m, and for each p ∈ [1, ∞] we put
3
1
0
2
y
a
M
1
1
]
.
A
F
h
t
a
m
[
1
v
7
9
4
2
.
5
0
3
1
:
v
i
X
r
a
kϕkLp(Ω) =(cid:13)(cid:13)(cid:13)
max
16k6m
.
ϕk(cid:13)(cid:13)(cid:13)p
Let β = (β1, . . . , βd) ∈ Zd
f defined on Ω we write ∇rf = (cid:16)∂rf /∂xβ(cid:17)β=r
+ := (N ∪ {0})d, β = β1 + . . . + βd. For any distribution
(here partial derivatives are taken
in the sense of distributions), and denote by mr the number of components of the
vector-valued distribution ∇rf . Set
W r
p,g(Ω) =(cid:8)f : Ω → R(cid:12)(cid:12) ∃ϕ : Ω → Rmr : kϕkLp(Ω) 6 1, ∇rf = g · ϕ(cid:9)
∇rf
(cid:16)we denote the corresponding function ϕ by
g (cid:17),
kf kLq,v(Ω)=kf kq,v=kf vkLq(Ω),
Lq,v(Ω) = {f : Ω → R kf kq,v < ∞} .
We call the set W r
p,g(Ω) a weighted Sobolev class.
For properties of weighted Sobolev spaces and their generalizations, see the books
[15,36,55,57,58,60] and the survey paper [33]. Sufficient conditions for boundedness
and compactness of embeddings of weighted Sobolev spaces into weighted Lq-spaces
were obtained by Kudryavtsev [32], Necas [47], Kufner [34 -- 36], Yakovlev [63], Triebel
[55], Lizorkin and Otelbaev [43], Gurka and Opic [22 -- 24], Besov [5 -- 8], Antoci [4],
Gol'dshtein and Ukhlov [21], and other authors. Notice that in these papers weighted
∗E-mail address: [email protected]
1
Sobolev classes were defined as W r
for some different weight functions gi.
p,g(Ω)∩Lp,w(Ω) for some weight w, or as ∩r
l=0W l
p,gi(Ω)
For a Lipschitz domain Ω, a k-dimensional manifold Γ ⊂ Ω, and for weights
depending only on distance from x to Γ, the following results were obtained. The
case r = 1, p = q was considered in papers of Necas [47] (the case of power weights
and Γ = ∂Ω), Kufner [34] (weights are powers of distance from a fixed point),
Yakovlev [63] (weights depend on distance to k-dimensional manifold), Kadlec and
Kufner [30, 31] (here weights are powers with a logarithmic factor, Γ = ∂Ω), Kufner
[35] (here weights are arbitrary functions of distance from ∂Ω). For p = q, r ∈ N,
Γ = ∂Ω and for power type weights, the embedding theorem was obtained by El
Kolli [16]. By using Banach space interpolation, Triebel [54] extended this result
to the case p 6 q. For p = q, r = 1, a k-dimensional manifold Γ and general
weights Kufner and Opic [37] obtained some sufficient conditions for compactness of
embeddings. For p > q, r ∈ N, for an arbitrary k-dimensional manifold Γ and power
type weights the criterion of the embedding was obtained in [27 -- 29]. In addition,
in [29] for r = 1 the criterion was obtained for arbitrary functions depending on
distance from the manifold Γ.
Notice that for p > q in the proof of embedding theorems two-weighted Hardy-
type inequalities were applied.
∇f
(cid:13)(cid:13)(cid:13)
g (cid:13)(cid:13)(cid:13)Lp(Ω)
In [22] sufficient conditions for the embedding were obtained for r = 1 and
general weights. The norm in the weighted Sobolev space was defined by kf kg,w =
+ kwf kLp(Ω). The idea of the proof was the following. First the Besikovic
covering of Ω was constructed, then for each ball of this covering the Sobolev
embedding theorem was applied. After that the obtained estimates were summarized.
Here it was essential to use the second weight w, which satisfied rather tight restricti-
ons. If the boundary ∂Ω is Lipshitz and weight functions are powers of distance
from ∂Ω, then it is possible to take more weak restrictions on w. To this end, the
other method of proof is used (employing the Hardy inequality). In [23] embedding
theorems were obtained for a Holder domain Ω and power type weights depending
on distance from ∂Ω.
It is also worth noting the paper [38], where the result on embedding of W 1
p,g(Ω)
into Lp,v(Ω) was obtained for r = 1, p = q and weights that are powers of the
distance from the irregular boundary of ∂Ω.
In the present paper, we consider a John domain Ω, an h-set Γ ⊂ ∂Ω and weight
functions depending on distance from Γ (their form will be written below).
Let X, Y be sets, f1, f2 : X × Y → R+. We write f1(x, y) .
y
f2(x, y) (or
f2(x, y) &
y
f1(x, y)) if, for any y ∈ Y , there exists c(y) > 0 such that f1(x, y) 6
c(y)f2(x, y) for each x ∈ X; f1(x, y) ≍
y
f2(x, y) if f1(x, y) .
y
f2(x, y) and f2(x, y) .
y
f1(x, y).
For x ∈ Rd and a > 0 we shall denote by Ba(x) the closed Euclidean ball of
radius a in Rd centered at the point x.
2
Let · be an arbitrary norm on Rd, and let E, E′ ⊂ Rd, x ∈ Rd. We set
diam· E = sup{y − z : y, z ∈ E}, dist· (x, E) = inf{x − y : y ∈ E},
dist· (E′, E) = inf{x − y : x ∈ E, y ∈ E′}.
Definition 1. Let Ω ⊂ Rd be a bounded domain, and let a > 0. We say that
Ω ∈ FC(a) if there exists a point x∗ ∈ Ω such that, for any x ∈ Ω, there exists a
curve γx : [0, T (x)] → Ω with the following properties:
1. γx ∈ AC[0, T (x)], γx = 1 a.e.,
2. γx(0) = x, γx(T (x)) = x∗,
3. Bat(γx(t)) ⊂ Ω for any t ∈ [0, T (x)].
Definition 2. We say that Ω satisfies the John condition (and call Ω a John domain)
if Ω ∈ FC(a) for some a > 0.
For a bounded domain, the John condition is equivalent to the flexible cone
condition (see the definition in [9]). Reshetnyak in the papers [50, 51] constructed
the integral representation for functions defined on a John domain Ω in terms of
their derivatives of order r. This integral representation yields that in the case r
d −
p − 1
p (Ω) is compactly embedded in the space Lq(Ω) (i.e., the
conditions of the compact embedding are the same as for Ω = [0, 1]d).
> 0 the class W r
(cid:16) 1
q(cid:17)+
Remark 1. If Ω ∈ FC(a) and a point x∗ is such as in Definition 1, then
diam· Ω .
d,a,·
dist· (x∗, ∂Ω).
(1)
Denote by H the set of all non-decreasing positive functions defined on (0, 1].
We introduce the concept of h-set according to [10].
Definition 3. Let Γ ⊂ Rd be a compact set, and let h ∈ H. We say that Γ is an h-set
if there exists a finite measure µ on Rd such that supp µ = Γ and µ(Bt(x)) ≍ h(t)
for each x ∈ Γ, t ∈ (0, 1].
Notice that the measure µ is non-negative.
The concept of h-sets for functions h of a special type appeared earlier (see papers
of Edmunds, Triebel and Moura [13,14,45,56]). In these and some other papers (see,
for example, [11,12,48,49,59]) properties of the operator trΓ in Besov and Triebel --
Lizorkin spaces and its composition with the operator (∆)−1 were studied. Here trΓ
is the operator of restriction on the h-set Γ. In [25] Besov spaces with Muckenhoupt
weights were studied; weight functions depending on the distance from a certain
h-set were considered as examples.
3
In the sequel we suppose that
h(t) = tθΛ(t),
0 6 θ < d,
where Λ : (0, +∞) → (0, +∞) is an absolutely continuous function such that
tΛ′(t)
Λ(t)
→
t→+0
0.
(2)
(3)
Let Ω ∈ FC(a) be a bounded domain, and let Γ ⊂ ∂Ω be an h-set. In the
sequel for convenience we suppose that Ω ⊂ (cid:2)− 1
reduced to this case). Let 1 < p 6 ∞, 1 6 q < ∞, r ∈ N, δ := r + d
g(x) = ϕg(dist·(x, Γ)), v(x) = ϕv(dist·(x, Γ)),
2 , 1
2(cid:3)d (the general case can be
q − dp > 0,
ϕg(t) = t−βgΨg(t), ϕv(t) = t−βvΨv(t),
with absolutely continuous functions Ψg, Ψv such that
tΨ′
g(t)
Ψg(t)
→
t→+0
0,
tΨ′
v(t)
Ψv(t)
→
t→+0
0;
in addition, we suppose that
Also we assume that
−βvq + d − θ > 0.
a) βg + βv < δ − θ(cid:18)1
q
−
1
p(cid:19)+
or b) βg + βv = δ − θ(cid:18)1
q
−
1
p(cid:19)+
.
In the case b) we suppose that
(4)
(5)
(6)
(7)
Λ(t) = log tγτ ( log t), Ψg(t) = log t−αgρg( log t), Ψv(t) = log t−αvρv( log t),
(8)
functions ρg, ρv, τ are absolutely continuous,
lim
y→+∞
yτ ′(y)
τ (y)
= lim
y→+∞
yρ′
g(y)
ρg(y)
= lim
y→+∞
yρ′
v(y)
ρv(y)
= 0,
γ < 0 and α := αg + αv > (1 − γ)(cid:18)1
q
−
1
p(cid:19)+
.
(9)
(10)
It is easy to show that the functions Λ, Ψg and Ψv satisfy (3) and (5).
Remark 2. If functions Ψg and Ψv (respectively ρg and ρv) satisfy (5) (respectively
(9)), then their product and each degree of these functions satisfies the similar
condition.
4
Denote
β = βg + βv,
ρ(y) = ρg(y)ρv(y), Ψ(y) = Ψg(y)Ψv(y),
Z = (r, d, p, q, βg, βv, θ, Λ, Ψg, Ψv, a).
Let Pr−1(Rd) be the space of polynomials on Rd of degree not exceeding r − 1.
For a measurable set E ⊂ Rd, we put Pr−1(E) = {f E : f ∈ Pr−1(Rd)}.
Theorem 1. For any function f ∈ span W r
Pr−1(Ω) such that
p,g(Ω) there exists a polynomial P f ∈
kf − P f kLq,v(Ω) .
.
Z (cid:13)(cid:13)(cid:13)(cid:13)
∇rf
g (cid:13)(cid:13)(cid:13)(cid:13)Lp(Ω)
Here the mapping f 7→ P f can be extended to a linear continuous operator P :
Lq,v(Ω) → Pr−1(Ω).
Later we shall give a more general formulation of this theorem. It can be used in
p,g(Ω) by piecewise polynomial
problems on estimating of approximation of the class W r
functions in the space Lq,v(Ω) and in problems on estimating of n-widths.
We may assume that the norm on Rd is given by
(x1, . . . , xd) = max
16i6d
xi.
The paper is organized as follows. In Sections 2 and 3, we give necessary notations
and formulate the results which will be required in the sequel. In Section 4, we
describe the domain Ω in terms of a tree T (see [61]) and construct a special
partition of this tree. In Section 5, the discrete weighted Hardy-type inequality
on a combinatorial tree is obtained for p = q. If the tree is regular, i.e., the number
of vertices that follow the given vertex depends only on the distance between this
vertex and the root of the tree, then we employ some convexity arguments and
reduce the problem to the proof of a Hardy-type inequality for sequences. The tree
which was constructed in Section 4 is not regular in general; however, it satisfies
some more weak condition of regularity. For such trees it is possible to reduce the
problem to the case of regular trees. To this end, a discrete analogue for theorem of
Evans -- Harris -- Pick [20] is proved. At this step, some quantity BD emerges; it is
defined for subtrees D and can be calculated recursively. Under some conditions on
weights, we prove that BD can be estimated by some more simple quantity SD. Then
for any subtree D we construct a subtree D in some regular tree A, such that SD
can be estimated from above by S D. In Section 6, the discrete Hardy-type inequality
on a tree is proved for p 6= q. To this end, the problem is reduced to consider the
cases p = q and p = ∞; here the Holder inequality is applied. In Section 7, the
embedding theorem is proved. The problem is reduced to considering the case r = d
and employing the discrete Hardy-type inequality on a tree.
Embedding theorems and related results for function classes on metric and
combinatorial trees were studied by different authors. Naimark and Solomyak [46]
5
obtained Hardy-type inequalities on regular metric trees. For a weighted summation
operator (i.e., a Hardy-type operator) on a combinatorial tree acting from l2 into
l∞, Lifshits and Linde [40 -- 42] obtained estimates of entropy numbers. In [18, 19, 53]
Evans, Harris, Lang and Solomyak obtained estimates of approximation numbers
for weighted Hardy-type operators on metric trees. Also it is worth noting results
of Evans and Harris [17] on embeddings of Sobolev classes on ridged domains into
Lebesgue spaces; here the definition of a ridged domain was given in terms of metric
trees.
2 Notation
In what follows A (int A, mes A, card A, respectively) be, respectively, the closure
(interior, Lebesgue measure, cardinality) of A. If a set A is contained in some
subspace L ⊂ Rd of dimension (d − 1), then we denote by intd−1A the interior
of A with respect to the induced topology on the space L. We say that sets A,
B ⊂ Rd do not overlap if A ∩ B is a Lebesgue nullset. For a convex set A we denote
by dim A the dimension of the affine span of the set A.
A set A ⊂ Rd is said to be a parallelepiped if there are sj 6 tj, 1 6 j 6 d, such
that
d
d
Yj=1
(sj, tj) ⊂ A ⊂
[sj, tj].
Yj=1
If tj − sj = t1 − s1 for any j = 1, . . . , d, then a parallelepiped is referred to as a
cube.
Let K be a family of closed cubes in Rd with axes parallel to coordinate axes. For
a cube K ∈ K and s ∈ Z+ we denote by Ξs(K) the set of 2sd closed non-overlapping
Ξs(K).
cubes of the same size that form a partition of K, and write Ξ(K) :=Ss∈Z+
We generally consider that these cubes are close (except the proof of Lemma 4).
We recall some definitions from graph theory. Throughout, we assume that the
graphs have neither multiple edges nor loops.
Let Γ be a graph containing at most countable number of vertices. We shall
denote by V(Γ) and by E(Γ) the set of vertices and the set of edges of Γ, respectively.
Two vertices are called adjacent if there is an edge between them. We shall identify
pairs of adjacent vertices with edges that connect them. Let ωi ∈ V(Γ), 1 6 i 6 n.
The sequence (ω1, . . . , ωn) is called a path, if the vertices ωi and ωi+1 are adjacent
for any i = 1, . . . , n − 1. We say that a graph is connected if any two vertices are
connected by a finite path. A connected graph is a tree if it has no cycles.
Let (T , ω0) be a tree with a distinguished vertex (or a root) ω0. We introduce
a partial order on V(T ) as follows: we say that ω′ > ω if there exists a path
(ω0, ω1, . . . , ωn, ω′) such that ω = ωk for some k ∈ 0, n. In this case, we set
ρT (ω, ω′) = n + 1 − k and call this quantity the distance between ω and ω′. In
addition, we set ρT (ω, ω) = 0. If ω′ > ω or ω′ = ω, then we write ω′ > ω and denote
6
[ω, ω′] := {ω′′ ∈ V(T ) : ω 6 ω′′ 6 ω′} (this set of vertices is called a segment). This
partial order on T induces a partial order on its subtree.
For any j ∈ Z+ we set
Vj(ω) := VT
j (ω) := {ω′ > ω : ρT (ω, ω′) = j}.
Given ω ∈ V(T ), we denote by Tω = (Tω, ω) a subtree of T with the set of vertices
{ω′ ∈ V(T ) : ω′ > ω}.
(11)
Let G be a subgraph in T . Denote by Vmax(G) and by Vmin(G), respectively, the
sets of maximal and minimal vertices in G.
Let W ⊂ V(T ). We say that G ⊂ T is a maximal subgraph on the set of vertices
W if V(G) = W and if any two vertices ξ′, ξ′′ ∈ W that are adjacent in T are also
adjacent in G.
We need the concept of a metric tree. Let (T , ω∗) be a tree with a finite set of
vertices, and let ∆ : E(T ) → 2R be a mapping such that for any λ ∈ E(T ) the
set ∆(λ) = [aλ, bλ] is a non-trivial segment. Then the pair T = (T , ∆) is called a
metric tree. A point on the edge λ of the metric tree T is a pair (t, λ), t ∈ [aλ, bλ],
λ ∈ E(T ) (if ω′ ∈ V1(ω), ω′′ ∈ V1(ω′), λ = (ω, ω′), λ′ = (ω′, ω′′), then we set
(bλ, λ) = (aλ′, λ′)). The distance between two points of T is defined as follows:
if (ω0, ω1, . . . , ωn) is a path in the tree T, n > 2, λi = (ωi−1, ωi), x = (t1, λ1),
y = (tn, λn), the we set
n−1
y − xT = bλ1 − t1 +
bλi − aλi + tn − aλn;
Xi=2
if x = (t′, λ), y = (t′′, λ), then y − xT = t′ − t′′.
We say that (t′, λ′) 6 (t′′, λ′′) if λ′ 6 λ′′ and t′ 6 t′′ in the case λ′ = λ′′.
If (t′, λ′) 6 (t′′, λ′′) and (t′, λ′) 6= (t′′, λ′′), then we write (t′, λ′) < (t′′, λ′′). If a,
x ∈ T, a 6 x, then we set [a, x] = {y ∈ T : a 6 y 6 x}.
A subset A = {(t, λ) : λ ∈ E(T), t ∈ Aλ} is said to be measurable, if Aλ is
measurable for any λ ∈ E(T). The Lebesgue measure of A is defined by
A = Xλ∈E(T)
Aλ.
A function f : A → R is said to be integrable if fλ := f {(t, λ): t∈Aλ} is integrable for
fλ(t) dt is finite. In this case, we set
any λ ∈ E(T) and the sum Pλ∈E(T) RAλ
ZA
f (x) dx = Xλ∈E(T)ZAλ
fλ(t) dt.
Let D ⊂ T be a connected subset. Denote by TD the maximal subtree in T such
that for any λ ∈ E(TD) the set ∆(λ) ∩ D is a non-trivial segment. Set ∆D(λ) =
7
∆(λ) ∩ D, λ ∈ E(T ). Then (TD, ∆D) is a metric tree, which will be identified with
the set D and which will be called a metric subtree in T.
Let D be a metric subtree in T. A point t ∈ D is said to be maximal if x ∈ T\D
for any x > t.
3 Preliminary results
Let ∆ be a cube with a side of length 2−m, m ∈ Z. Set m(∆) = m. In particular, if
2 , 1
2, 1
We shall need Whitney's covering theorem (see, e.g., [39, p. 562]).
2(cid:3)d(cid:17), then ∆ ∈ Ξm(∆)(cid:16)(cid:2)− 1
∆ ∈ Ξ(cid:16)(cid:2)− 1
Theorem A. Let Ω ⊂(cid:2)− 1
pairwise non-overlapping cubes Θ(Ω) = {∆j}j∈N ⊂ Ξ(cid:16)(cid:2)− 1
2(cid:3)d be an open set. Then there exists a family of closed
2(cid:3)d(cid:17) with the following
2(cid:3)d(cid:17).
properties:
2 , 1
2 , 1
1. Ω = ∪j∈N∆j;
2. dist (∆j, ∂Ω) ≍
d
2−m(∆j );
3. for any j ∈ N
card {i ∈ N : dim(∆i ∩ ∆j) = d − 1} 6 12d;
4. if dim(∆i ∩ ∆j) = d − 1, then
m(∆j) − 2 6 m(∆i) 6 m(∆j) + 2.
(12)
(13)
Andersen and Heinig in [3, 26] proved discrete analogues of the two-weighted
Hardy-type inequality. We formulate a particular case of their result, which will be
used in the sequel.
Theorem B. Let 1 6 p 6 q < ∞, and let {un}n∈Z, {wn}n∈Z be nonnegative
sequences such that
Then, for any sequence {an}n∈Z,
C := sup
n!1/q m
Xn=−∞
wq
m∈Z ∞
Xn=m
q!1/q
ukak(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=−∞
wn
n
8
Xn=−∞(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
∞
up′
n!1/p′
< ∞.
.
p,q
C Xn∈Z
anp!1/p
.
(14)
Evans, Harris and Pick in [20] proved a criterion for boundedness of a two-
weighted Hardy-type operator on a metric tree.
Let T = (T , ∆) be a metric tree, x0 ∈ T, and let u, w : T → R+ be measurable
functions. We set Tx0 = {x ∈ T : x > x0},
Iu,w,x0f (x) = w(x)
x
Zx0
u(t)f (t) dt.
Denote by Jx0 = Jx0(T) a family of metric subtrees D ⊂ T with the following
properties:
1. x0 is a minimal vertex in D;
2. if x ∈ ∂D\{x0}, then x is a maximal point in D.
For D ∈ Jx0, we set
kf kLp(T) :
t
Zx0
αD = inf
f (x)u(x) dx = 1 for any t ∈ ∂D
.
Theorem C. Let 1 6 p 6 q 6 ∞. Then the operator Iu,w,x0 : Lp(Tx0) → Lq(Tx0) is
bounded if and only if
Cu,w := sup
D∈Jx0
kwχTx0 \DkLq(T)
αD
< ∞.
Moreover, Cu,w 6 kIu,wkLp(Tx0 )→Lq(Tx0 ) 6 4Cu,w.
The quantity αD is calculated recursively. The following theorem is also proved
in [20].
Theorem D. Let D ∈ Jx0, D = ∪m
j=0
m, Di ∩ Dj = {y0}, i 6= j. Then
Dj, D0 = [x0, y0], x0 < y0, Dj ∈ Jy0, 1 6 j 6
1
αD
=(cid:13)(cid:13)(cid:13)(cid:16)α−1
D0 , k(αDi)m
i=1k−1
lm
.
p (cid:17)(cid:13)(cid:13)(cid:13)l2
p′
Notice that if x0 = (t′, λ), λ ∈ E(T), t′ ∈ ∆(λ), and y0 is such as in Theorem D,
then y0 is a right end of ∆(λ).
The following theorem is proved in [1, 2, 52]; see also [44, p. 51] and [39, p. 566].
Theorem E. Let 1 < p < q < ∞, d ∈ N, r > 0, r
d + 1
q − 1
p = 0. Then the operator
T f (x) =ZRd
f (y)x − yr−d dy
is bounded from Lp(Rd) in Lq(Rd).
9
Reshetnyak [50, 51] constructed the integral representation for smooth functions
defined on a John domain Ω in terms of their derivatives of order r. We shall use
the following form of his result (see also [61]).
Theorem F. Let Ω ∈ FC(a), let the point x∗, the curves γx and the numbers T (x)
(x∗, ∂Ω), r ∈ N. Then there exist
be such as in Definition 1, and let R0 = distk·kld
measurable functions Hβ : Ω × Ω → R, β = (β1, . . . , βd) ∈ Zd
+, β = r, such that
the inclusion supp Hβ(x, ·) ⊂ ∪t∈[0, T (x)]Bat(γx(t)) and the inequality Hβ(x, y) .
a,d,r
2
x − yr−d hold for any x ∈ Ω. Moreover, the following representation holds:
f (x) = Xβ=rZΩ
Hβ(x, y)
∂rf (y)
∂βy
dy,
f ∈ C ∞(Ω),
f BR0/2(x∗) = 0.
The proof of the following lemma is straightforward and will be omitted.
Lemma 1. Let Φ : (0, +∞) → (0, +∞), ρ : (0, +∞) → (0, +∞) be absolutely
continuous functions and let lim
t→+0
yρ′(y)
ρ(y) = 0. Then for any ε > 0
tΦ′(t)
Φ(t) = 0,
lim
y→+∞
tε .
ε
Φ(t) .
ε
t−ε,
if
t ∈ (0, 1],
t−ε .
ε
ρ(t) .
ε
tε,
if
t ∈ [1, ∞).
Let σ ∈ R, µ < −1. Then for any sequence {kj}l
j=0 ⊂ Z+ such that k0 < k1 <
· · · < kl, the following estimates hold:
l
Xj=0
Xj=0
l
2σkj Φ(2−kj ) .
σ,Φ
2σk0Φ(2−k0), if σ < 0,
2σkj Φ(2−kj ) .
σ,Φ
2σklΦ(2−kl), if σ > 0,
l
Xj=0
kµ
j ρ(kj) .
µ,ρ
k1+µ
0
ρ(k0).
4 Construction of the partition of the tree
Let Θ ⊂ Ξ(cid:16)(cid:2)− 1
2 , 1
2(cid:3)d(cid:17) be a set of non-overlapping cubes.
Definition 4. Let G be a graph, and let F : V(G) → Θ be a one-to-one mapping. We
say that F is consistent with the structure of the graph G if the following condition
holds: for any adjacent vertices ξ′, ξ′′ ∈ V(G) the set Γξ′,ξ′′ := F (ξ′) ∩ F (ξ′′) has
dimension d − 1.
10
Let (T , ξ∗) be a tree, and let F : V(T ) → Θ be a one-to-one mapping consistent
with the structure of the tree T . For any adjacent vertices ξ′, ξ′′, we set Γξ′,ξ′′ =
int d−1Γξ′,ξ′′, and for each subtree T ′ of T , we put
ΩT ′,F =(cid:0)∪ξ∈V(T ′)int F (ξ)(cid:1) ∪(cid:16)∪(ξ′,ξ′′)∈E(T ′)Γξ′,ξ′′(cid:17) .
For ξ ∈ V(T ), ∆ = F (ξ), denote mξ = m(∆), Ω6∆ = Ω[ξ∗, ξ],F .
(15)
Let Θ(Ω) be a Whitney covering of Ω (see Theorem A). The following lemma is
proved in [61].
Lemma 2. Let Ω ⊂ (cid:2)− 1
, Ω ∈ FC(a). Then there exist a tree T and a one-
to-one mapping F : V(T ) → Θ(Ω) consistent with the structure of T and which
satisfies the following properties:
2, 1
2(cid:3)d
1. for any subtree T ′ of T ,
ΩT ′,F ∈ FC(b∗), where b∗ = b∗(a, d) > 0;
(16)
2. if x ∈ F (ξ), then a curve γx from Definition 1 can be chosen so that Bb∗t(γx(t)) ⊂
Ω6F (ξ) for any t ∈ [0, T (x)]; if ξT ′ is a minimal vertex of T ′, then the center
of the cube F (wT ′) can be taken as a point x∗ from Definition 1 with ΩT ′,F ∈
FC(b∗).
For ξ ∈ V(T ), we set Ωξ = ΩTξ,F ; the number kξ ∈ Z+ is chosen so that
2−kξ 6 dist·(F (ξ), Γ) < 2−kξ+1.
By Theorem A,
2−mξ ≍
d
dist·(F (ξ), ∂Ω) 6 dist·(F (ξ), Γ) ≍ 2−kξ;
hence, there exists ϑ(d) ∈ Z+ such that
kξ 6 mξ + ϑ(d).
(17)
(18)
(19)
Let zξ ∈ F (ξ) be such that dist·(zξ, Γ) = dist·(F (ξ), Γ), and let zξ be a center of
the cube F (ξ). Then the first relation in (18) together with
zξ − zξ 6 2−mξ
(20)
imply that for any x ∈ Ωξ
x − zξ 6 diam·Ωξ
(1)
.
a,d
dist·(zξ, ∂Ω) 6 dist·(F (ξ), ∂Ω) + zξ − zξ
(18),(20)
.
d
2−mξ.
11
Hence, there exists c(a, d) > 0 such that
x − zξ 6 c(a, d) · 2−mξ, x ∈ Ωξ.
Prove that
for any x ∈ F (ξ). Indeed,
dist·(x, Γ) ≍
d
2−kξ
2−kξ
(17)
6 dist·(F (ξ), Γ) 6 dist·(x, Γ) 6 dist·(F (ξ), Γ) + x − zξ
(17)
.
d
. 2−kξ+1 + 2−mξ
(19)
.
d
2−kξ.
Denote
W = {ξ ∈ V(T ) : mξ 6 kξ + 1 + log c(a, d)}.
From (19) it follows that for any ξ ∈ W
2−mξ ≍
a,d
2−kξ.
Let ξ /∈ W. We show that for any x ∈ Ωξ
dist·(x, Γ) ≍
a,d
2−kξ.
Indeed,
(21)
(22)
(23)
(24)
(25)
dist·(x, Γ) 6 dist·(zξ, Γ) + x − zξ
(17),(21)
.
a,d
2−kξ + 2−mξ
(19)
.
d
2−kξ,
dist·(x, Γ) > dist·(zξ, Γ) − x − zξ
(17),(21)
> 2−kξ − c(a, d) · 2−mξ
(23)
> 2−kξ−1.
Denote
Wν = {ξ ∈ W : kξ = ν}.
(26)
Then (21) and (24) imply that for any ξ ∈ Wν and for any tree T ′ ⊂ Tξ rooted at ξ
diam· ΩT ′, F ≍
a,d
2−ν.
(27)
Lemma 3. There exist a partition of the tree T into subtrees Tk,i with minimal
vertices ξk,i, k ∈ Z+, i ∈ Ik, Ik 6= ∅, and numbers νk ∈ N, satisfying the following
conditions:
12
1. ν0 < ν1 < · · · < νk < . . . ;
2. ξk,i ∈ Wνk;
3. dist·(x, Γ) ≍
a,d
2−νk for any x ∈ ΩTk,i,F ;
4. if ξk′,i′ < ξk,i, then k′ < k.
Proof. Let ν ∈ Z+, ξ ∈ Wν. Denote by T( ξ) a set of subtrees T ′ ⊂ T ξ with the
minimal vertex ξ such that
V(T ′)\ [l>ν+1
Wl! = ∅
(28)
(this set is nonempty, since { ξ} ∈ T( ξ)). Denote by S(T ξ) a subtree in T ξ such that
V(S(T ξ)) = ∪S∈T(ξ)V(S). Then S(T ξ) ∈ T( ξ).
Prove that there exists ν = ν(a, d) ∈ N such that for any x ∈ ΩS(T ξ),F
2−ν−ν 6 dist·(x, Γ) 6 2−ν+ν.
Indeed,
dist·(x, Γ) 6 x − z ξ + dist·(z ξ, Γ)
(17),(21),(26)
.
a,d
2−m ξ + 2−ν
(19)
.
d
2−ν.
Prove the estimate from below. Let x ∈ F (η), η ∈ V(S(T ξ)). Set
η = max{ W ∩ [ ξ, η]}.
Then η ∈ Wj for some j ∈ Z+; since S(T ξ) ∈ T( ξ), we have j 6 ν. If η = η, then
dist·(F (η), Γ)
(17)
> 2−j > 2−ν.
(29)
Let η > η, ζ ∈ [η, η] ∩ V1(η). Then ζ /∈ W, dim(F (η) ∩ F (ζ)) = d − 1, and for any
x ∈ F (η)
dist·(x, Γ)
(25)
≍
a,d
2−k ζ
(19)
&
d
2−mζ
(13)
> 2−mη−2 (24)
≍
a,d
(26),(29)
> 2−ν.
2−k η
Let ξ0 be a minimal vertex of the tree T . Prove that ξ0 ∈ W. Indeed, otherwise
2−kξ0 for any x ∈ ΩT ,F . Hence, Γ 6⊂ ∂Ω, which leads
(25) imply that dist·(x, Γ) ≍
a,d
to a contradiction.
The further arguments are the same as the arguments in Lemma 2 from [62].
13
Proposition 1. Let ξk,i < ξk′,i′, {ξ : ξk,i 6 ξ < ξk′,i′} ⊂ V(Tk,i). Then νk′ 6 νk + s,
with s = s(a, d).
Proof. Let ξ ∈ [ξk,i, ξk′,i′] be the direct predecessor of ξk′,i′. Then ξ ∈ V(Tk,i).
2−νk for any x ∈ F (ξ), as well as
By Assertion 3 of Lemma 3, dist·(x, Γ) ≍
a,d
2−νk′ for any x ∈ F (ξk′,i′). Since the mapping F is consistent with
2−νk′ . This
dist·(x, Γ) ≍
a,d
the structure of the tree T , then F (ξ) ∩ F (ξk′,i′) 6= ∅. Hence, 2−νk ≍
a,d
completes the proof.
Lemma 4. Let ξ ∈ Wν0. In addition, suppose that there exists c0 > 1 such that
∀j ∈ N,
t, s ∈ [2−j−1, 2−j+1]
c−1
0 6
h(t)
h(s)
6 c0.
Given ν > ν0, we denote
Then
Wν( ξ) = Wν ∩ V(T ξ).
card Wν( ξ) .
a,d,c0
h(2−ν0)
h(2−ν)
.
(30)
(31)
If, in addition, ξ is a root of the tree T , then there is k = k(a, d) ∈ N such that
card Wj( ξ) &
a,d,c0
h(2−ν0)
h(2−ν)
.
ν+k
Xj=ν
(32)
Proof. Throughout the proof of this lemma we suppose that a cube is a product of
j=1[aj, bj). Then any two non-overlapping cubes do not intersect.
If ξ ∈ Wν, then it follows from (24) that there is k∗(a, d) such that ν −k∗(a, d) 6
semi-intervals Qd
mξ 6 ν + k∗(a, d). Further,
diam Ωξ
(21),(24)
≍
a,d
2−ν0.
Therefore, if ξ ∈ Wν( ξ), then 2−ν ≍
a,d
2−mξ 6 diam Ωξ ≍
a,d
2−ν0. Hence,
2ν0,
if
Wν( ξ) 6= ∅.
2ν &
a,d
(33)
(34)
Let k ∈ Z, −k∗(a, d) 6 k 6 k∗(a, d). Denote by Wν,k( ξ) the set of ξ ∈ Wν( ξ) such
that mξ = ν + k.
It follows from (17) and (26) that
dist·(F (ξ), Γ) 6 2−ν+1,
ξ ∈ Wν( ξ).
14
This together with (33) implies that there exists a cube ∆0 and a number k0 =
k0(a, d) ∈ Z+ such that
F ( ξ) ∈ Ξ(∆0),
ν0 − k0(a, d) 6 m(∆0) 6 ν0 + k0(a, d),
Ωξ ⊂ ∆0,
∃x ∈ Γ ∩ ∆0 : dist·(x, ∂∆0) &
a,d
2−ν0,
∀ξ ∈ Wν( ξ) ∃x ∈ Γ ∩ ∆0 : dist·(x, F (ξ)) 6 2−ν+1.
(35)
(36)
(37)
Let j ∈ N, j > ν0 + k∗(a, d). In this case, if ∆ ∈ Ξj(cid:16)(cid:2)− 1
∆ ⊂ ∆0 or ∆ does not overlap with ∆0. It follows from the conditions F ( ξ) ∈ Ξ(∆0)
and j > ν0 + k∗(a, d) > m ξ. Denote by ∆0,j a cube that is obtained from ∆0 by a
dilatation in respect to its center, with a side length m(∆0) + 2 · 2−j. Set
2(cid:3)d(cid:17), then either
2 , 1
1
2
,
Θj(∆0) =(∆ ∈ Ξj (cid:20)−
Θj(∆0) =(∆ ∈ Ξj (cid:20)−
1
2
,
1
2(cid:21)d! : ∆ ⊂ ∆0, ∆ ∩ Γ 6= ∅) ,
2(cid:21)d! : ∆ ⊂ ∆0,j, ∆ ∩ Γ 6= ∅) .
1
Prove that
card Θj(∆0) ≍
a,d,c0
h(2−ν0)
h(2−j)
.
(38)
Let ∆ ∈ Θj(∆0). Since ∆ ∩ Γ 6= ∅, there is a cube K∆ centered at z∆, such that
∆ ∈ Ξ1(K∆),
dist·(z∆, Γ) 6 2−m(∆)−1.
(39)
Then there are
z∆ ∈ Γ,
t∆ &
d
2−j, t∆ .
d
2−j
such that Bt∆(z∆) ⊂ K∆ ⊂ Bt∆(z∆).
(40)
Let µ be a measure from the definition 3 (in particular, supp µ ⊂ Γ). Then (30) and
(40) imply that µ(K∆) ≍
d,c0
h(2−j). On the other hand, by (30), (35) and (36),
h(2−ν0) ≍
a,d,c0
h(2−ν0) ≍
a,d,c0
µ(∆0) = X∆∈Θj(∆0)
µ(∆0,j) = X∆∈ Θj(∆0)
µ(∆),
µ(∆).
15
µ(K∆) .
µ(∆). The first inequality holds since the measure µ is
Therefore, in order to prove (38) it is sufficient to check that P∆∈Θj (∆0)
P∆∈Θj(∆0)
d P∆∈ Θj(∆0)
nonnegative. Prove the second inequality. Since ∆ ∈ Ξ1(K∆), we have K∆ ⊂ ∆0,j.
Denote
µ(∆) 6
Θj,∆ = {∆′ ∈ Θj(∆0) : ∆′ ⊂ K∆} for ∆ ∈ Θj(∆0),
j,∆′ = {∆ ∈ Θj(∆0) : ∆′ ⊂ K∆} for ∆′ ∈ Θj(∆0).
Θ′
Since card Θ′
j,∆′ .
d
1 for any ∆′ ∈ Θj(∆0), we have
X∆∈Θj (∆0)
µ(K∆) = X∆∈Θj(∆0) X∆′∈Θj,∆
µ(∆′) 6 X∆′∈ Θj (∆0) X∆∈Θ′
j,∆′
µ(∆′) .
d X∆′∈ Θj (∆0)
µ(∆′).
This proves (38).
Show that if −k∗(a, d) 6 k 6 k∗(a, d), ν > ν0 + 2k∗(a, d), then
card Wν,k( ξ) .
card Θν+k(∆0)
a,d
(41)
(recall that Θj(∆0) was defined for j > ν0 + k∗(a, d)).
Set
A :=(∆′ ∈ Ξν+k (cid:20)−
1
2
,
1
2(cid:21)d! : ∃∆ ∈ Θν+k(∆0) : dist·(∆′, ∆) 6 2−ν+1) .
From (37) it follows that {F (ξ) : ξ ∈ Wν,k( ξ)} ⊂ A and card Wν,k( ξ) 6 card A .
a,d
card Θν+k(∆0).
If ν > ν0 + 2k∗(a, d), then (30), (38) and (41) imply (31). If Wν( ξ) 6= ∅ and
ν < ν0 + 2k∗(a, d), then by (34) we get 2−ν ≍
a,d
together with (24), (26), (35) and (36) yield (31).
2−ν0; hence, h(2−ν0 )
h(2−ν )
(30)
≍
a,d,c0
1. This
Let us prove (32). By (38), it is sufficient to check
ν+k
Xj=ν
card Wj( ξ) &
a,d,c0
card Θν(∆0).
Let ∆ ∈ Θl(∆0), and let K∆ be the cube defined above (see (39)),
x∆ ∈ K∆ ∩ Γ,
x∆ − z∆ 6 2−m(∆)−1.
(42)
(43)
Since Γ ⊂ ∂Ω, by Theorem A and Lemma 2 for any m ∈ N there is a vertex
ξ∆ ∈ V(T ) such that
dist (x∆, F (ξ∆)) < 2−m,
2−mξ∆ .
d
2−m.
(44)
16
If m is sufficiently large, then it follows from (43) and (44) that F (ξ∆) ⊂ K∆. Denote
η∆ = min{η ∈ [ ξ, ξ∆] : F (η) ⊂ K∆}. Show that
2−mη∆ ≍
a,d
2−l
and η∆ ∈ W for sufficiently large m.
(45)
Indeed, since ∆ ∈ Θl(∆0), we have 2−m(∆) = 2−l. The inclusion F (η∆) ⊂ K∆ and
2−m(∆).
the first relation in (39) imply that 2−mη∆ . 2−m(∆). Check that 2−mη∆ &
a,d
It follows from the definition of η∆ that ∂F (η∆) ∩ ∂K∆ 6= ∅. Let x ∈ ∂F (η∆) ∩ ∂K∆.
(39)
= 2−m(∆). By (44), there is a point y ∈ F (ξ∆) such that x∆ − y 6
Then x − z∆
2−m. Since F (ξ∆) ⊂ Ωη∆, we have
2−mη∆
(21)
&
a,d
diam·Ωη∆ > x − y >
> x − z∆ − z∆ − x∆ − x∆ − y
(43)
> 2−m(∆)−1 − 2−m > 2−m(∆)−2
for large m. The first relation in (45) is proved. Check the second relation. Let
η∆ /∈ W. Then, by (25), for any x ∈ Ωη∆ we have dist·(x, Γ) ≍
2−kη∆ . Taking
x ∈ F (ξ∆) ⊂ Ωη∆, we get
a,d
2−mη∆
(18)
.
a,d
2−kη∆ ≍
a,d
dist·(x, Γ) 6 x − x∆
(44)
.
d
2−m.
It follows from the proved first relation in (45) that 2−l .
a,d
2−m. It is impossible for
large m.
It follows from (24), (26) and (45) that η∆ ∈ Wj for some j ∈ Z+ such that
2−j ≍
a,d
2−l. Therefore,
l − l∗ 6 j 6 l + l∗, with l∗ = l∗(a, d) ∈ N.
(46)
Set k = 2l∗. In order to prove (32), we take j = ν + l∗ and apply (46). We get
card {η∆ : ∆ ∈ Θν+l∗(∆0)} 6
card Wl.
ν+k
Xl=ν
Hence, in order to prove (42) it is sufficient to check
card Θj(∆0) .
d
card {η∆ : ∆ ∈ Θj(∆0)}
(47)
and to apply (38) with (30). Let ∆, ∆′ ∈ Θj(∆0), η∆ = η∆′. Then K∆ ∩ K∆′ ⊃ ∆′′,
∆′′ ∈ Ξ1(K∆) and ∆′′ ∈ Ξ1(K∆′). Therefore, card {η∆′ : ∆′ ∈ Θj(∆0), η∆′ = η∆} .
d
1, which implies (47). This completes the proof.
17
Let m ∈ N. For 0 < t0 < t1 6 ∞ denote by Gt0, t1 the maximal subgraph in T
on the set of vertices
V(Gt0, t1) := [t06νk<t1 [i∈Ik
Tk,i
(the index set Ik was defined in Lemma 3); by {Dj,i}i∈ Ij we denote the set of all
connected components of the graph G1+mj, 1+m(j+1); by ξj,i = ξm
j,i denote the minimal
vertex of the tree Dj,i, j ∈ Z+. Then
1. ξj,i ∈ Wνk for some νk ∈ [1 + mj, 1 + m(j + 1)); in particular,
2. for any x ∈ ΩDj,i,F
diam ΩDj,i,F
(27)
≍
a,d,m
2−mj;
dist·(x, Γ) ≍
a,d,m
2−mj
(48)
(49)
(it follows from Assertion 3 of Lemma 3);
3. if ξj,i < ξj ′,i′, then j < j′ (indeed, ξj,i = ξk,t and ξj ′,i′ = ξk′,t′ for some k, t,
k′, t′; by Assertions 4 and 1 of Lemma 3, νk < νk′; it implies that j 6 j′; the
equality j = j′ is impossible; indeed, in this case the vertices ξj,i and ξj ′,i′ are
incomparable).
Let ξj,i < ξj ′,i′,
{ξ : ξj,i 6 ξ < ξj ′,i′} ⊂ Dj,i.
Then we say that the tree Dj ′,i′ follows the tree Dj,i.
Remark 3. Let s be such as in Proposition 1, let m > s, and let Dj ′,i′ follow the
tree Dj,i. Then j′ = j + 1.
Indeed, let ξt,s ∈ Dj,i, {ξ : ξt,s < ξ < ξj ′,i′} ⊂ V(Tt,s), ξj ′,i′ = ξt′,s′. By Proposition
1, 1 + mj′ 6 νt′ 6 νt + s < 1 + m(j + 1) + s. Hence, m(j′ − j − 1) < s. Since m > s,
the last inequality is possible only for j′ = j + 1.
Given j ∈ Z+, l ∈ N, t ∈ Ij, we denote
I l
j,t = I l,m
j,t = {i ∈ Ij+l : ξm
j+l,i > ξm
j,t}.
(50)
Lemma 5. Let m ∈ N be divisible by s. Suppose that (30) holds for some c0 > 1.
Then
card I l
j,t .
a,d,c0
h(2−mj)
h(2−m(j+l))
.
(51)
18
Proof. First consider the case m = s.
By the property 1 of the trees Dj,t and Dj+l,i,
ξj,t ∈
s(j+1)
[ν ′=1+sj
Wν ′,
ξj+l,i ∈
s(j+l+1)
[ν=1+s(j+l)
Wν( ξj,t)
(recall that Wν( ξj,t) = Wν ∩ V(T ξj,t
that
s(j+l+1)
)). Therefore, from Lemma 4 and (30) it follows
card I l
j,t 6
Xν=1+s(j+l)
card Wν( ξj,t) .
a,d,c0
h(2−sj)
h(2−s(j+l))
.
Consider the case m = m′s. Then ξm
j,t = ξs
j ′,t′ for some j′ > m′j, ξm
j+l,i = ξs
m′(j+l),i′
(by Remark 3). Hence,
card I l,m
j,t = card I m′l,s
m′j,t′ .
a,d,c0
h(2−sm′j)
h(2−sm′(j+l))
=
h(2−mj)
h(2−m(j+l))
.
This completes the proof.
5 The discrete Hardy-type inequality on a tree: case
p = q
5.1 The analogue of Evans -- Harris -- Pick theorem
Let (A, ξ0) be a tree with a finite vertex set, let 1 6 p 6 ∞, and let u, w : V(A) →
R+ be weight functions. Denote by SA,u,w the minimal constant C in the inequality
Xξ∈V(A)
wp(ξ) Xξ′6ξ
1/p
u(ξ′)f (ξ′)!p
6 C
Xξ∈V(A)
1/p
f p(ξ)
, f : V(A) → R+.
(52)
Remark 4. If D ⊂ A is a subtree, then SD,u,w 6 SA,u,w.
Let us obtain two-sided estimates for SA,u,w. We reduce this problem to estimating
the constant in the Hardy-type inequality on a metric tree and use the result from
the article [20].
Let ξ ∈ V(A), D ⊂ A ξ. We say that D ∈ J ′
ξ
if the following conditions hold:
1. ξ is the minimal vertex in D,
2. if ξ ∈ V(D) is not a maximal vertex D, then V1(ξ) ⊂ V(D).
19
Denote by D the subtree in D such that V( D) = V(D)\Vmax(D).
For any subgraph G ⊂ A and for any function f : V(G) → R, we denote
kf klp(G) =
Xω∈V(G)
f (ω)p
1/p
.
(53)
By lp(G) we denote the space of functions f : V(G) → R equipped with the norm
kf klp(G).
For D ∈ J ′
ξ
we set
βD = inf
kf klp(A) : Xξ6ξ′6ξ
Notice that if D = { ξ}, then
f (ξ′)u(ξ′) = 1, ∀ξ ∈ Vmax(D)
.
β{ ξ} = inf{f ( ξ) : f ( ξ)u( ξ) = 1} = u−1( ξ).
Lemma 6. Suppose that there exists C > 1 such that for any ξ ∈ V(A)
and let for any adjacent vertices ξ, ξ′ ∈ V(A)
card V1(ξ) 6 C,
Then
C −1 6
u(ξ)
u(ξ′)
6 C, C −1 6
w(ξ)
w(ξ′)
6 C.
SA ξ,u,w ≍
p, C
sup
D∈J ′
ξ
kwχA ξ\ Dklp(A ξ)
βD
.
(54)
(55)
(56)
(57)
Proof. If V(A) = {ξ0}, then the assertion is trivial.
Let V(A) 6= {ξ0}. Add to the set V(A) a vertex ξ∗ and join it with ξ0 by an edge.
Thus we obtain the tree ( A, ξ∗). Define the mapping ∆ by ∆(λ) = [0, 1], λ ∈ E( A).
Thus we get the metric tree A = ( A, ∆). For any function ψ : V(A) → R we define
ψ# : A → R as follows. Let e = (ξ′, ξ) ∈ E(A), ξ > ξ′. Then we set ψ#∆(e) = ψ(ξ).
Let λ ξ ∈ E(A) be an edge with the end ξ, x0 = (0, λ ξ) ∈ A. By Holder inequality,
kIu#,w#,x0kLp(Ax0 )→Lp(Ax0 ) ≍ SA ξ,u,w.
It follows from Theorem C that
SA ξ,u,w ≍ sup
D∈Jx0
kw#χAx0 \DkLp(Ax0 )
αD
,
20
with
kφkLp(Ax0 ) :
αD = inf
t
Zx0
φ(x)u#(x) dx = 1 ∀t ∈ ∂D
.
Applying the Holder inequality once again (see also [20]), we obtain that
αD = inf
here Ldiscr
the metric tree A.
p
kφkLp(Ax0 ) : φ ∈ Ldiscr
p
(Ax0),
t
Zx0
φ(x)u#(x) dx = 1 ∀t ∈ ∂D
;
(A) is the set of functions φ : A → R that are constants on each edge of
Let D = (D, ∆D) ∈ Jx0. Set D+ = (D, ∆), D− = ( D, ∆). Prove that D ∈ J ′
.
ξ
Indeed, let ξ ∈ V(D), and suppose that there exist vertices ξ′ ∈ V1(ξ)\V(D) and
ξ′′ ∈ V1(ξ) ∩ V(D). Let η be a vertex in A that is the direct predecessor of ξ. Then
the point (1, (η, ξ)) = (0, (ξ, ξ′)) = (0, (ξ, ξ′′)) belongs to the boundary of D, as
well as it is not maximal.
We have
kw#χAx0 \DkLp(A) 6 kw#χAx0 \D−kLp(A) = kwχA ξ\ Dklp(A),
kφkLp(Ax0 ) : φ ∈ Ldiscr
p
αD > inf
(Ax0),
t
Zx0
φ(x)u#(x) dx = 1 ∀t ∈ ∂D+
= βD.
This implies the upper estimate for SA ξ,u,w. Prove the lower estimate. Notice that
if D = D+, then αD = βD. If in addition Vmax(D) ∩ Vmax(A) = ∅, then
kw#χAx0 \DkLp(Ax0 ) = kwχA ξ\Dklp(A ξ)
(56),(57)
≍
p, C
kwχA ξ\ Dklp(A ξ).
Hence,
SA ξ,u,w &
p, C
sup( kwχA ξ\ Dklp(A ξ)
βD
: D ∈ J ′
ξ , Vmax(D) ∩ Vmax(A) = ∅) =: Σ.
Prove that
Σ ≍
p, C
sup(kwχA ξ\ Dklp(A ξ)
βD
: D ∈ J ′
ξ) .
To this end, it is sufficient to show that if V(D) 6= { ξ}, then
kwχA ξ\ Dklp(A ξ)
βD
21
Σ.
.
p, C
Indeed, set D1 = D. Then from (54), (56) and (57) it follows that kwχA ξ\ Dklp(A ξ) ≍
p, C
and Vmax(D1) ∩
kwχA ξ\ D1
Vmax(A) = ∅.
βD1. It remains to observe that D1 ∈ J ′
ξ
klp(A ξ) and βD ≍
p, C
Proposition 2. Let ξ∗ ∈ V(A), V1(ξ∗) = {ξ1, . . . , ξm}, Dj ∈ J ′
D = {ξ∗} ∪ D1 ∪ · · · ∪ Dm. Then
ξj , 1 6 j 6 m,
β−1
D =(cid:13)(cid:13)(cid:13)(cid:16)β−1
{ξ∗}, (cid:13)(cid:13)(βDj )m
j=1(cid:13)(cid:13)
−1
lm
p (cid:17)(cid:13)(cid:13)(cid:13)l2
p′
This assertion follows from Theorem D.
.
(58)
5.2 The reduction lemma
Let ψ : R+ → R+ be an increasing function, ψ(0) = 0, let (A, ξ0) be a tree with a
finite vertex set. In addition, suppose that there exists C∗ > 1 such that for any j0,
j ∈ Z+, j > j0, ξ ∈ Vj0(ξ0)
j−j0(ξ) 6 C∗2ψ(j)−ψ(j0).
Let u : V(A) → (0, +∞), u(ξ) = uj for ξ ∈ VA
card VA
(59)
j (ξ0). Suppose that there is
σ ∈ (0, 1) such that for any j ∈ N
uj2− ψ(j)
p
uj−12− ψ(j−1)
p
> σ− 1
p′ .
(60)
For each ξ∗ ∈ V(A) and for any subtree D ∈ J ′
ξ∗ we define the quantity βD by
(54). Then β{ξ}
We set
(55)
= u−1
j
for ξ ∈ Vj(ξ0), and if D 6= {ξ∗}, then (58) holds.
BD =
1
βD
,
SD =
− 1
p
.
Xξ∈Vmax(D)
u−p(ξ)
Let D ∈ J ′
ξ∗, ξ ∈ V(D), V1( ξ) = {ξ1, . . . , ξm1}. Then
D ξ = { ξ} ∪ Dξ1 ∪ · · · ∪ Dξm1
, Dξj ∈ J ′
ξj , 1 6 j 6 m1.
Let ε > 0, 1 6 i 6 m1. A vertex ξi is said to be (ε, D)-regular if
B−p
Dξi
> ε
B−p
Dξ1
+ · · · + B−p
Dξm1
m1
.
(61)
Notice that if ε < 1, then at least one of the vertices ξi is (ε, D)-regular. A path
(η0, . . . , ηl) in D is said to be (ε, D)-regular if η0 < η1 < · · · < ηl and for any
1 6 j 6 l the vertex ηj is (ε, D)-regular.
22
Lemma 7. There exists σ = σ(p, C∗) > 0 such that if (60) holds with σ ∈ (0, σ),
then for any ξ∗ ∈ V(A) and for any subtree D ∈ J ′
ξ∗
SD 6 BD 6 2SD.
(62)
Proof. Let
νD = max{j ∈ Z+ : Vj(ξ∗) 6= ∅}.
If νD = 0, then it follows from the definition that SD = BD. Let us prove the
assertion for νD > 0. In this case,
D = {ξ∗} ∪ D1 ∪ · · · ∪ Dm1, Dj ∈ J ′
ξj , ξj ∈ V1(ξ∗).
Notice that
Prove the first inequality, i.e.,
S−p
D =
S−p
Di .
m1
Xi=1
SD 6 BD.
(63)
(64)
(65)
Let ν ∈ Z+, and let the assertion be proved for any D such that νD 6 ν. Prove the
assertion for νD = ν + 1. From (58) and the induction assumption it follows that
Bp′
D = Bp′
{ξ∗} + m1
Xj=1
′
B−p
p
Dj!− p
′
p
S−p
Dj!− p
> m1
Xj=1
(64)
= Sp′
D .
Prove the second inequality. It is sufficient to check that
BD 6 ∞
Yj=1
(1 + σj/2)
2
p′! SD
(66)
holds for σ ∈ (0, σ(p, C∗)).
Let ε ∈ (0, 1) (it will be chosen later). Then the end of any (ε, D)-regular path
that has a maximal length and starts from ξ∗ is a maximal vertex in D (otherwise one
of its direct successors is (ε, D)-regular). Denote by lD the maximal length of (ε, D)-
regular paths that start in ξ∗. We show by induction on νD that for σ ∈ (0, σ(p, C∗))
BD 6 lD
Yj=1
(1 + σj/2)
2
p′! SD.
(67)
This implies (66).
If νD = 0, then D is a single vertex. Therefore, BD = SD and (67) is true.
23
Let νD > 0. Then (63) holds, and by (59)
m1 6 C∗2ψ(j0+1)−ψ(j0).
Let li = lDi. Denote by I1 the set of i ∈ {1, . . . , m1} such that ξi is (ε, D)-regular,
I2 = {1, . . . , m1}\I1. Set
l = max{li : i ∈ I1} + 1.
Then l = lD.
Prove that there exists σ∗ = σ∗(ε, C∗, p) > 0 such that for any σ ∈ (0, σ∗)
Suppose the converse, i.e.,
Let ξ∗ ∈ Vj0(ξ0). Then
β−1
{ξ∗} = uj0 = 2
ψ(j0)
p
β−p′
{ξ∗} 6 σ
β−p′
{ξ∗} > σ
l
2 m1
Xi=1
l
2 m1
Xi=1
′
B−p
p
Di!− p
′
B−p
p
Di!− p
.
.
uj0+j−12− ψ(j0+j−1)
p
uj0+j2− ψ(j0+j)
p
l
·
Yj=1
6 uj0+l2− ψ(j0+l)
· uj0+l2− ψ(j0+l)
p
(60)
6
p + ψ(j0)
p
l
p′ .
· σ
This together with (69) yields
m1
Xi=1
i.e.,
′
B−p
p
Di!− p
< σ
l
2 up′
j0+l
2− p
′
ψ(j0+l)
p
′
+ p
ψ(j0)
p
,
Di > σ− pl
B−p
2p′ 2ψ(j0+l)−ψ(j0)u−p
.
j0+l
m1
Xi=1
Let (ξ∗, η1, . . . , ηl) be an (ε, D)-regular path in D. Then
B−p
Dη1
>
ε
m1
m1
Xi=1
B−p
Di >
ε
m1
24
σ− pl
2p′ 2ψ(j0+l)−ψ(j0)u−p
.
j0+l
(68)
(69)
(70)
(71)
Let 2 6 j 6 l, Dηj−1 = {ηj−1} ∪ Dj,1 ∪ · · · ∪ Dj,mj . Then
Bp′
Dηj−1
(58)
= Bp′
{ηj−1} + mj
Xi=1
′
B−p
p
Dj,i!− p
> mj
Xi=1
′
B−p
p
Dj,i!− p
,
i.e., B−p
Dηj−1
6
mj
Pi=1
Therefore,
B−p
Dj,i. Since the vertex ηj is (ε, D)-regular, we have
B−p
Dηj
>
ε
mj
mj
Xi=1
B−p
Dj,i
>
ε
mj
B−p
Dηj−1
.
B−p
Dηl
>
ε
ml
B−p
Dηl−1
>
ε2
ml−1ml
B−p
Dηl−2
> . . . >
εl−1
m2 . . . ml
(71)
>
B−p
Dη1
>
εl
m1m2 . . . ml
2ψ(j0+l)−ψ(j0)u−p
j0+l
σ− pl
2p′ .
The vertex ηl is maximal in D. Hence, BDηl
= B{ηl} = uj0+l. In addition,
m1m2 . . . ml
(59)
6 C
l
∗
2ψ(j0+i+1)−ψ(j0+i) = C
l
∗2ψ(j0+l)−ψ(j0).
Thus,
u−p
j0+l
>
Cl
∗2ψ(j0+l)−ψ(j0)
2ψ(j0+l)−ψ(j0)u−p
j0+l
σ− pl
2p′ ,
l−1
Yi=0
εl
∗ σ− pl
i.e., 1 > εlC −l
This proves (68).
2p′ , or σ
p
2p′ > εC −1
∗ . For 0 < σ 6 (C−1
∗ ε)
2
2p′
p
we get the contradiction.
Now let us prove (67). We have
Bp′
D = β−p′
{ξ∗} + m1
Xi=1
′
B−p
Di!− p
p (68)
6 (cid:16)1 + σ
l/2(cid:17) m1
Xi=1
′
B−p
p
Di!− p
.
(72)
Show that there exists ε∗ = ε∗(p) ∈ (0, 1) such that for any ε ∈ (0, ε∗), 0 < σ <
min(cid:0) 1
(cid:16)B−p
2, σ∗(ε, C∗, p)(cid:1)
Dm1(cid:17)− 1
D1 + · · · + B−p
p
D1 + · · · + S−p
6(cid:16)S−p
p
Dm1(cid:17)− 1
l−1
(1 + σj/2)
Yj=1
2
p′
· (1 + σ
l/2)
1
p′ .
(73)
25
Then (72), (64) and (73) yield (67).
The relation (73) is equivalent to
B−p
Di −
m1
Xi=1
(1 + σl/2)
S−p
Di
m1
p
Pi=1
p′ Q
l−1
j=1(1 + σj/2)
> 0.
2p
p′
(74)
Consider separately sums in i ∈ I1 and in i ∈ I2. Let l = max16i6m1 li + 1. By the
induction hypotheses,
BDi 6 li
Yj=1
(1 + σj/2)
BDi 6 li
Yj=1
(1 + σj/2)
2
l−1
p′! SDi 6
Yj=1
p′! SDi 6 l−1
Yj=1
2
(1 + σj/2)
(1 + σj/2)
2
p′
SDi,
p′! SDi,
2
i ∈ I1,
(75)
i ∈ I2.
(76)
Hence,
>
= Xi∈I1
S−p
Di!
(1 + σl/2)
l−1
j=1(1 + σj/2)
(75)
>
2p
p′
S−p
Di
p
Pi∈I1
p′ Q
B−p
Di −
S−p
Di
Xi∈I1
Pi∈I1
−
(1 + σl/2)
Q
2p
p′
l−1
j=1(1 + σj/2)
(1 + σl/2)
l−1
j=1(1 + σj/2)
p′ − 1
p
p
(1 + σl/2)
2p
p′
p′ Q
S−p
Di
p
Pi∈I1
p′ Q
p Xi∈I1
&
l−1
j=1(1 + σj/2)
=
2p
p′
S−p
Di! σ
l/2
(65)
> Xi∈I1
B−p
Di! σ
l/2
2). Therefore, there exists C1(p) > 0
(the penultimate relation holds for 0 < σ < 1
such that
B−p
Di! σ
l/2.
(77)
If l = l, then the sum in i ∈ I2 is estimated similarly. In this case, (74) is proved.
B−p
Di −
Xi∈I1
(1 + σl/2)
S−p
Di
p
Pi∈I1
p′ Q
l−1
j=1(1 + σj/2)
Let l > l + 1. Then we have for 0 < σ < min(cid:0) 1
Pi∈I2
p′ Q
(1 + σl/2)
Xi∈I2
Di −
B−p
p
S−p
Di
2p
p′
> C1(p) Xi∈I1
2, σ∗(ε, C∗, p)(cid:1)
(76)
>
l−1
j=1(1 + σj/2)
2p
p′
26
S−p
Di
S−p
Di
2p
p′
>
−
(1 + σl/2)
j=1(1 + σj/2)
Pi∈I2
Ql−1
> Xi∈I2
> −C2(p) Xi∈I2
Pi∈I2
p′ Q
Di! 1 − (1 + σl/2)
p′ Ql−1
Ql−1
> −C3(p) Xi∈I2
Di! σ
j=1(1 + σj/2)
S−p
S−p
(76)
l/2
2p
p′
p
p
l−1
j=1(1 + σj/2)
j=l+1(1 + σj/2)
B−p
Di! σ
l/2,
>
2p
p′
2p
p′
>
where C2(p) > 0, C3(p) > 0. Thus,
B−p
Di −
Xi∈I2
(1 + σl/2)
l−1
j=1(1 + σj/2)
> −C3(p) Xi∈I2
2p
p′
B−p
Di! σ
l/2.
(78)
S−p
Di
p
Pi∈I1
p′ Q
From definitions of I1 and I2 we get
B−p
Di
Xi∈I2
m1
ε
m1
6Xi∈I2
Xj=1
Xi=1
Di −Xi∈I2
B−p
m1
B−p
Di =
Xi∈I1
B−p
Dj
6 ε
B−p
Dj ,
m1
Xj=1
B−p
Di
> (1 − ε)
B−p
Di .
m1
Xi=1
This together with (77) and (78) implies that
m1
Di −
B−p
Xi=1
> C1(p) Xi∈I1
B−p
Di! σ
(1 + σl/2)
l−1
j=1(1 + σj/2)
>
2p
p′
m1
p
S−p
Di
Pi=1
p′ Q
l/2 − C3(p) Xi∈I2
Xi=1
m1
B−p
Di! σ
l/2 >
B−p
Di > 0
> σl/2 ((1 − ε)C1(p) − εC3(p))
for sufficiently small ε. This completes the proof of (74).
Let w : V(A) → (0, ∞), w(ξ) = wj for ξ ∈ VA
j (ξ0). Suppose that there exists
σ ∈(cid:0)0, 1
2(cid:1) such that for any j ∈ N
wj · 2
ψ(j)
p
wj−1 · 2
ψ(j−1)
p
27
6 σ
1
p .
(79)
Given D ∈ J ′
ξ∗, we denote
RD =
Xξ∈Vmax(D)Xξ′>ξ
wp(ξ′)
1/p
, QD =
Xξ∈Vmax(D)
1/p
wp(ξ)
.
(80)
From (79) and (59) it follows that there exists σ∗ = σ∗(p, C∗) > 0 such that for any
0 < σ < σ∗
QD 6 RD 6 2QD.
(81)
j ∈ N.
(82)
(83)
(84)
Construct the function ψ∗ by induction as follows:
ψ∗(0) = 0,
Then 2ψ∗(j)−ψ∗(j−1) ∈ N and
2ψ∗(j)−ψ∗(j−1) =(cid:2)2ψ(j)−ψ∗(j−1)(cid:3) ,
2ψ∗(j) 6 2ψ(j) 6 2ψ∗(j)+1.
Let ξ∗ ∈ VA
j0(ξ0), and let ( A, ξ) be a tree such that
card V
A
1 (ξ) = 2ψ∗(j+1)−ψ∗(j),
ξ ∈ V
A
j−j0( ξ),
j > j0.
Lemma 8. Let D ⊂ A be a tree rooted at ξ∗ ∈ VA
j0(ξ0). Then there exists σ0 =
σ0(p, C∗) > 0 satisfying the following property: if (60) and (79) hold for some σ ∈
(0, σ0), then there exists a tree D ⊂ A rooted at ξ such that SD .
Q D.
p
S D and QD .
p,C∗
Proof. Set
{j1, . . . , js} = {j ∈ N : Vmax(D) ∩ Vj−j0(ξ∗) 6= ∅}, j1 < · · · < js.
For each 1 6 l 6 s, we denote Vl,D = Vmax(D) ∩ VA
jl−j0(ξ∗),
Us = VA
js−1−j0(ξ∗) ∩ V(D)\Vmax(D).
Then
Vs,D ⊂ ∪ξ∈Us
VA
js−js−1(ξ).
(85)
By (59) and (79), there exists σ1 = σ1(p, C∗) such that for any σ ∈ (0, σ1), 1 6 ν 6 s
wp
jν card VA
jν −jν−1(ξ) 6 wp
jν−1.
Show that for any σ ∈ (0, σ1)
wp
jtcard Vt,D 6 wp
jν · card VA
jν−j0(ξ∗).
s
Xt=ν
28
(86)
(87)
We use induction on s − ν. If s − ν = 0, then the inequality is trivial. Let s − ν > 1.
Denote by D the subtree in Aξ∗ with the set of maximal vertices (cid:0)∪s−1
and the root ξ∗. Then
t=1
Vt,D(cid:1) ∪ Us
wp
jtcard Vt,D
(85),(86)
6
s
Xt=ν
s−1
Xt=ν
s−1
wp
jtcard Vt,D + wp
js−1card Us =
wp
jtcard V
t, D 6 wp
jν card VA
jν −j0(ξ∗)
=
Xt=ν
(the last inequality holds by the induction assumption). This completes the proof of
(87).
l, D ⊂ V( A ξ) with the following
Applying induction on l, construct the set V
properties:
1. if 1 6 t < ν 6 l, then
V
ν, D ∩(cid:16)∪ξ∈V
t, D
V
A
jν −jt(ξ)(cid:17) = ∅;
2. if
∪l
t=1 ∪ξ∈V
t, D
V
A
jl−jt(ξ) = V
A
jl−j0( ξ),
then the tree D with the set of vertices
(88)
(89)
V( D) = ∪l
t=1 ∪ξ∈V
t, D [ ξ, ξ]
satisfies Vmax( D) = ∪16t6lV
t, D, SD .
p
S D and QD .
p,C∗
Q D;
3. if
∪l
t=1 ∪ξ∈V
t, D
V
A
jl−jt(ξ) 6= V
A
jl−j0( ξ),
(90)
then card V
t, D = card Vt,D for any 1 6 t 6 l.
If (89) holds for some l, then the construction is interrupted. In this case, D is the
desired tree. If (90) holds for any l 6 s, then we take as D the tree with the vertex
set ∪16t6s ∪ξ∈V
t, D [ ξ, ξ]. In this case, S D = SD and Q D = QD.
The base of induction. Let l = 1. If card V1,D < 1
1, D an arbitrary subset E1 ⊂ V
22ψ∗(j1)−ψ∗(j0), then we take
A
j1−j0( ξ) such that card E1 = card V1,D. By (84),
as V
we have (90).
Let card V1,D > 1
holds). Hence, V( D) = ∪j1−j0
j=0
2 2ψ∗(j1)−ψ∗(j0). Then we set V
A
j ( ξ), Vmax( D) = V
1, D = V
1, D and
V
A
j1−j0( ξ) (in this case, (89)
S−p
D
(84)
= 2ψ∗(j1)−ψ∗(j0)u−p
j1 , Qp
D
(84)
= 2ψ∗(j1)−ψ∗(j0)wp
j1.
29
Further,
S−p
D > card V1,D · u−p
j1
which implies SD .
p
S D. Prove that QD .
p, C∗
>
2ψ∗(j1)−ψ∗(j0)u−p
j1 ,
1
2
Q D. Indeed,
QD =
s
Xt=1
card Vt, D · wp
jt
(87)
6 wp
j1 · card VA
j1−j0(ξ∗)
(59),(83)
.
C∗
wp
j1 · 2ψ∗(j1)−ψ∗(j0) = Qp
D
.
The induction step. Let 1 6 l < s,
card Vt,D · 2ψ∗(jl)−ψ∗(jt) <
1
2
· 2ψ∗(jl)−ψ∗(j0).
(91)
l
Xt=1
Suppose that there are the sets V
t, D ⊂ V
A
jt−j0( ξ), 1 6 t 6 l, satisfying (88) and
card V
t, D = card Vt,D, 1 6 t 6 l.
(92)
Then
6
l
Xt=1
card V
A
jl−jt(ξ)
(84),(92)
6
l
Xt=1 Xξ∈V
t, D
card Vt,D · 2ψ∗(jl)−ψ∗(jt)
(91)
<
1
2
2ψ∗(jl)−ψ∗(j0)
(84)
< card V
A
jl−j0( ξ).
Therefore, properties 1 -- 3 of the sets V
t, D hold (property 2 is trivial, since (90) holds
instead of (89); property 3 follows from (92), property 1 holds since we supposed
that the sets satisfy (88)).
Construct the set V
Let
l+1, D ⊂ V
A
jl+1−j0( ξ)\ ∪l
t=1 ∪ξ∈V
t, D
V
A
jl+1−jt(ξ).
card Vl+1,D +
l
Xt=1
card Vt,D · 2ψ∗(jl+1)−ψ∗(jt) <
1
2
2ψ∗(jl+1)−ψ∗(j0).
(93)
In this case, we take an arbitrary subset
V
l+1, D ⊂ V
A
jl+1−j0( ξ)\ ∪l
t=1 ∪ξ∈V
t, D
V
A
jl+1−jt(ξ), card V
l+1, D = card Vl+1,D.
This set exists, since
card Vl+1,D+
l
Xt=1 Xξ∈V
t, D
card V
A
jl+1−jt(ξ)
(84)
= card Vl+1,D+
card V
t, D·2ψ∗(jl+1)−ψ∗(jt)
l
Xt=1
A
jl+1−j0( ξ).
(92),(93)
<
1
2
· 2ψ∗(jl+1)−ψ∗(j0)
(84)
< card V
30
Then we have (88), (91) and (92) with l + 1 instead of l. Hence, properties 1 -- 3 for
the sets {V
t, D}l+1
t=1 hold.
Let
card Vl+1,D +
Then we set
l
Xt=1
card Vt,D · 2ψ∗(jl+1)−ψ∗(jt) >
2ψ∗(jl+1)−ψ∗(j0).
(94)
1
2
V
l+1, D = V
A
jl+1−j0( ξ)\ ∪l
t=1 ∪ξ∈V
t, D
V
A
jl+1−jt(ξ).
(95)
By construction, we have property 1 of the sets {V
of l); i.e.,
t, D}l+1
t=1 and (89) (with l+1 instead
V
ν, D ∩(cid:16)∪ξ∈V
V
t, D
A
jν −jt(ξ)(cid:17) = ∅,
∪l+1
t=1 ∪ξ∈V
t, D
V
A
jl+1−jt(ξ) = V
A
jl+1−j0( ξ).
1 6 t < ν 6 l + 1,
(96)
Therefore, it is sufficient to check property 2. Define the tree D by
V( D) = ∪l+1
t=1 ∪ξ∈V
t, D [ ξ, ξ].
From (96) it follows that
Vmax( D) = ∪l+1
t=1
V
t, D.
We claim that SD .
p
S D and QD .
p,C∗
Q D. Indeed,
(92),(97)
=
S−p
D
l
Xt=1
jt card Vt,D + u−p
u−p
jl+1card V
l+1, D,
S−p
D >
l
Xt=1
jt card Vt,D + u−p
u−p
jl+1card Vl+1,D,
(97)
(98)
(99)
(92),(97)
=
Qp
D
l
Xt=1
jtcard Vt,D + wp
wp
jl+1card V
l+1, D,
(100)
Qp
D =
wp
jtcard Vt,D +
wp
jtcard Vt,D
(87)
6
s
Pt=l+1
l
Pt=1
l
Pt=1
.
6
jtcard Vt,D + wp
wp
jl+1card VA
jl+1−j0(ξ∗)
(59),(83)
.
C∗
(101)
wp
jtcard Vt,D + wp
jl+1
· 2ψ∗(jl+1)−ψ∗(j0).
l
Pt=1
31
In addition,
Case 1. Let
Then
Indeed,
card V
l+1, D 6 card V
A
jl+1−j0( ξ)
(84)
= 2ψ∗(jl+1)−ψ∗(j0).
(102)
card Vt,D · 2ψ∗(jl+1)−ψ∗(jt) <
1
4
2ψ∗(jl+1)−ψ∗(j0).
(103)
l
Xt=1
card Vl+1,D > 2ψ∗(jl+1)−ψ∗(j0)−2.
(104)
card Vl+1,D
(94),(103)
>
1
2
2ψ∗(jl+1)−ψ∗(j0) −
1
4
2ψ∗(jl+1)−ψ∗(j0) =
1
4
2ψ∗(jl+1)−ψ∗(j0).
From (98), (99), (102) and (104) it follows that SD .
p
S D.
Prove that QD .
p,C∗
Q D. By (100) and (101), it suffices to check that card V
l+1, D >
2ψ∗(jl+1)−ψ∗(j0)−1. We have
card V
l+1, D
(84),(92),(95)
>
card V
A
jl+1−j0( ξ) −
card Vt,D · 2ψ∗(jl+1)−ψ∗(jt)
(103)
>
l
Xt=1
= 2ψ∗(jl+1)−ψ∗(j0) −
1
4
· 2ψ∗(jl+1)−ψ∗(j0) > 2ψ∗(jl+1)−ψ∗(j0)−1.
card Vt,D · 2ψ∗(jl+1)−ψ∗(jt) >
1
4
2ψ∗(jl+1)−ψ∗(j0).
(105)
Then by (60), (79) and (83), there exists σ′
1 = σ′
1(p, C∗) such that for any σ ∈ (0, σ′
1)
Case 2. Let
l
Xt=1
l
l
Xt=1
Xt=1
l
u−p
jt card Vt,D >
wp
jtcard Vt,D >
Xt=1
Xt=1
l
u−p
jl+1
2ψ∗(jl+1)−ψ∗(jt)card Vt,D
wp
jl+12ψ∗(jl+1)−ψ∗(jt)card Vt,D
(105)
>
(105)
>
u−p
jl+1
4
wp
jl+1
4
2ψ∗(jl+1)−ψ∗(j0),
2ψ∗(jl+1)−ψ∗(j0).
This together with (98), (99), (100), (101) and (102) implies that
S−p
D >
u−p
jt card Vt,D ≍
p
S−p
D
,
l
Xt=1
Xt=1
l
Qp
D ≍
p,C∗
wp
jtcard Vt,D ≍
p
Qp
D
.
32
This completes the proof.
Let (56) and (57) hold, let σ be such as in Lemma 7, and let σ0 be such as in
Lemma 8. Take σ ∈ (0, min{σ, σ0}). By Lemma 6,
SAξ∗ ,u,w ≍
p, C
sup
D∈J ′
ξ∗
kwχAξ∗ \ Dklp(Aξ∗ )BD
(80)
= sup
D∈J ′
ξ∗
RDBD
(62),(81)
≍
p
sup
D∈J ′
ξ∗
QDSD. (106)
Lemma 9. Let ξ∗ ∈ VA
there exists σ2 = σ2(p, C∗) > 0 such that SAξ∗ ,u,w .
j0(ξ0), u(ξ) = uj, w(ξ) = wj for any ξ ∈ V
A
j−j0( ξ). Then
S A,u, w for any σ ∈ (0, σ2).
p, C,C∗
Proof. Suppose that the supremum of the right-hand side in (106) is attained at
ξ∗. Apply Lemma 8 and construct the tree D ⊂ A rooted at ξ such
the tree D ∈ J ′
Q D. Apply (106) to the trees D and D and notice that
that SD .
p
in respect to the tree D. We get
S D and QD .
p,C∗
D ∈ J ′
ξ
SAξ∗ ,u,w ≍
p, C
SDQD .
p,C∗
S DQ D
(62),(81)
6 B DR D .
p, C
S D,u, w 6 S A,u, w
(see Remark 4).
5.3 Estimates for the special class of weights
Let r = d, p = q and let the conditions (2), (3), (4), (5), (6), (7), (8), (9), (10) hold.
From (7) it follows that β 6 d.
Let T , F be the tree and the mapping such as in Lemma 2, and let s = s(a, d) ∈
N be such as in Proposition 1. Let m ∈ N be divisible in s. Consider the partition
{Dj,i}j∈Z+, i∈ Ij of the tree T defined at the page 18. Fix N ∈ N. Let A = A(m) be
the tree with the set of vertices {ηj,i}06j6N, i∈ Ij and with the set of edges defined by
VA
1 (ηj,i) = {ηj+1,s}s∈ I 1
.
j,i
j,i is defined in (50). By Remark 3, if Dj ′,i′ follows the tree Dj,i, then j′ = j + 1
j,i. Hence, card VA
l (ηj,i) = card I l
j,i for any l ∈ Z+.
By Lemma 5, for any j0, j ∈ {0, . . . , N}, j > j0, and for any ξ ∈ VA
j0(η0,1) we
Here I 1
and i′ ∈ I 1
have
with
card VA
j−j0(ξ) .
a,d,c0
h(2−mj0)
h(2−mj)
(2)
= 2ψ(j)−ψ(j0)
ψ(j) = mθj − log2 Λ(2−mj).
33
(107)
(108)
Denote ξ0 = η0,1. Set
uj := u(ξ) = ϕg(2−mj) · 2− mdj
p′
(4)
= 2mj(cid:16)βg− d
p′(cid:17)Ψg(2−mj),
wj := w(ξ) = ϕv(2−mj) · 2− mdj
p
(4)
= 2mj(βv− d
p )Ψv(2−mj), ξ ∈ VA
j (ξ0).
(109)
Lemma 10. There exists m∗ = m∗(Z) ∈ N such that for any m > m∗, ξ∗ ∈ VA
we have SAξ∗ ,u,w .
Z
we have SAξ∗ ,u,w .
Z
j0(ξ0)
2mj0(β−d)Ψ(2−mj0) in the case a) of (7); in the case b) for α > 0
j−α
0 ρ(j0); if α = 0 and ρ ≡ 1, then SAξ∗ ,u,w .
1.
Z
Proof. First suppose that
βg −
d
p′ −
θ
p
> 0.
(110)
We have
uj · 2− ψ(j)
p′ − θ
p = 2mj(cid:16)βg− d
p = 2mj(βv− d
p + θ
p(cid:17) · Ψg(2−mj)Λ
p ) · Ψv(2−mj)Λ− 1
1
ψ(j)
wj · 2
p (2−mj),
p (2−mj).
From (6) and (110) it follows that (60) and (79) hold with σ .
Z
λm
∗ , λ∗ = λ∗(Z) ∈
(0, 1). From (107) follows (59) with C∗ = C∗(a, d, c0). There exists m∗ such that
σ < σ2(p, C∗) for any m > m∗ (see Lemma 9). Let the tree ( A, ξ) satisfy (84) with
ψ∗ defined by (82), and let u(ξ) = uj, w(ξ) = wj for ξ ∈ Vj−j0( ξ0). By Lemma 9,
SAξ∗ ,u,w .
Z
S A,u, w.
The quantity S A,u,v equals to the minimal constant C in
wp(ξ)
Xξ∈V( A)
p
u(ξ′)f 1/p(ξ′)
Xξ6ξ′6ξ
6 C p Xξ∈V( A ξ)
f (ξ), f : V( A) → R+.
(111)
Denote by F (f ) the left-hand side of (111).
We claim that the function F is concave. Indeed, let λ ∈ [0, 1], f1, f2 : V( A ξ) →
R+. Applying the inverse Minkowski inequality and the homogeneity property, we
get
p
>
Xξ∈V( A)
wp(ξ)
Xξ6ξ′6ξ
> (1 − λ) Xξ∈V( A)
+λ Xξ∈V( A)
u(ξ′) ((1 − λ)f1(ξ′) + λf2(ξ′))1/p
wp(ξ)
wp(ξ)
Xξ6ξ′6ξ
Xξ6ξ′6ξ
(ξ′)
(ξ′)
u(ξ′)f 1/p
u(ξ′)f 1/p
+
.
p
p
2
1
34
Set nj = card V
A
1 (ξ), ξ ∈ V
A
j ( ξ), j ∈ Z+. It follows from (84) that this quantity
does not depend on ξ. Prove that
= sup{F (f ) : kf kl1( A) 6 1, ∀j ∈ Z+, ∀ξ′, ξ′′ ∈ Vj( ξ) f (ξ′) = f (ξ′′)}
sup{F (f ) : kf kl1( A) 6 1} =
(112)
(see the notation (53)).
Construct fk;i1, ..., ik, ik+1 by induction on k ∈ {0, 1, . . . , N − j0}. Set f0 = f ( ξ).
A
k ( ξ). Then we define
Let 0 6 k 6 N − j0 − 1, fk;i1, ..., ik = f (ξ) for some ξ ∈ V
fk+1;i1, ..., ik, ik+1 for 1 6 ik+1 6 nk so that
{fk+1;i1, ..., ik, ik+1}nk
ik+1=1 = {f (ξ′) : ξ′ ∈ V
A
1 (ξ)}.
Denote by Sj the set of permutations of j elements.
For 0 6 t 6 N − j0 − 1, σ ∈ Snt we set
(f t,σ)k;i1, ..., ik =(cid:26) fk;i1, ..., ik ,
for k 6 t,
fk;i1, ..., σ(it+1), ..., ik,
for k > t,
φ(t)(f ) =
f t,σ.
1
card Snt Xσ∈Snt
Since the function F is concave and F (f t,σ) = F (f ), we get
F (f ) 6 F (φ(0)(f )) 6 F (φ(1)φ(0)(f )) 6 . . . 6 F (φ(N −j0−1) . . . φ(0)(f )).
It remains to observe that (cid:0)φ(N −j0−1) . . . φ(0)(f )(cid:1) (ξ′) = (cid:0)φ(N −j0−1) . . . φ(0)(f )(cid:1) (ξ′′)
for any ξ′, ξ′′ ∈ Vk( ξ), 0 6 k 6 N − j0.
Thus, (112) holds. Hence, it suffices to find the minimal constant C in (111) for
the family of functions f such that fVk(ξ) = fk, 0 6 k 6 N − j0. Set mk = n0 . . . nk.
From (84) it follows that nk = 2ψ∗(j0+k+1)−ψ∗(j0+k) and
mk = 2ψ∗(j0+k+1)−ψ∗(j0) (83),(108)
≍
2θmk Λ(2−mj0)
Λ(2−m(j0+k+1))
.
(113)
Let xk = (mk−1fk)1/p, m−1 = 1. Then it follows from the definition of u and w that
(111) can be written as
N −j0
Xk=0
mk−1wp
k+j0 k
Xl=0
Applying Theorem B, we get
ul+j0m
− 1
p
l−1xl!p!1/p
6 C N −j0
Xk=0
xp
k!1/p
.
(114)
C ≍
p
sup
06k6N −j0 N −j0
Xl=k
ml−1wp
l+j0!
1
p k
Xl=0
35
up′
l+j0m
′
− p
p
l−1!
1
p′
.
Apply Lemma 1, taking into account Remark 2. From (6), (109) and (113) it follows
ml−1wp
l+j0 ≍
Z
mk−1wp
k+j0. The condition βg > d
p′ + θ
p yields the inequality
N −j0
Pl=k
up′
l+j0m
that
k
Pl=0
′
− p
p
l−1 ≍
Z
′
− p
k−1up′
p
m
k+j0. Therefore,
C ≍
Z
sup
06k6N −j0
uk+j0wk+j0
(109)
= sup
j06t6N
2mt(β−d)Ψ(2−mt) =: M.
In the case (7), a), we have M ≍
Z
2mj0(β−d)Ψ(2−mj0). In the case (7), b) for α > 0 we
get M ≍
Z,m
j−α
0 ρ(j0). If α = 0 and ρ ≡ 1, then M = 1.
Let, now, βg − d
p′ − θ
p 6 0. Since βv < d−θ
p , there exists βg > d
p′ + θ
p such that
βg + βv < d. Set u(ξ) = u(ξ) · 2(βg−βg)mj, ξ ∈ VA
j (ξ0). Then
SAξ∗ ,u,w 6 SAξ∗ ,u,w · 2mj0(βg−βg) .
Z
2mj0(β−d)Ψ(2−mj0).
This completes the proof.
6 The discrete Hardy-type inequality on the tree:
case p 6= q
Let the tree A be such as in the previous section, and let
u(ξ) = ϕg(2−mj) · 2− mdj
p′ = 2mj(cid:16)βg− d
p′(cid:17)Ψg(2−mj),
w(ξ) = ϕv(2−mj) · 2− mdj
q = 2mj(βv− d
q )Ψv(2−mj), ξ ∈ Vj(ξ0).
(115)
Let ξ∗ ∈ Vj0(ξ0). Denote by Sp,q
Aξ∗ ,u,v the minimal constant C in the inequality
Xξ∈V(Aξ∗ )
wq(ξ) Xξ∗6ξ′6ξ
u(ξ′)f (ξ′)!q
1
q
6 Ckf klp(Aξ∗ ).
(116)
Lemma 11. Let p > q. Then there exists m∗ = m∗(Z) ∈ N such that for any
m > m∗
Sp,q
Aξ∗ ,u,w .
Z
2mj0(β−d− d
q + d
p )Ψ(2−mj0)
for the case a) in (7); in the case b), for α >(cid:16) 1
q − 1
p(cid:17) (1 − γ)
Sp,q
Aξ∗ ,u,w .
Z
2−θ( 1
q − 1
p )mj0j
q − 1
p
−α+ 1
0
ρ(j0).
36
Proof. First consider the case p = ∞. Let βg > d. Then
1
q
6
u(ξ′)f (ξ′)!q
wq(ξ) Xξ∗6ξ′6ξ
Xξ∈V(Aξ∗ )
u(ξ′)!q
wq(ξ) Xξ∗6ξ′6ξ
Λ(2−mj) j
v(2−mj) · 2mθ(j−j0) Λ(2−mj0)
Xl=j0
1
q
6
Xξ∈V(Aξ∗ )
2mj(βvq−d)Ψq
kf kl∞(Aξ∗ )
(107),(108),(115)
.
a,d,c0,m
2ml(βg−d)Ψg(2−ml)!q!1/q
kf kl∞(Aξ∗ ).
. N
Xj=j0
This together with the condition βg > d, Lemma 1 and Remark 2 yield
S∞,q
Aξ∗ ,u,w .
Z N
Xj=j0
2mj(βq−d−dq+θ)Ψq(2−mj) · 2−θmj0 Λ(2−mj0)
Λ(2−mj)!1/q
.
In the case (7), a), the right-hand side can be estimated from above up to a
multiplicative constant by 2mj0(β−d− d
q )Ψ(2−mj0); in the case (7), b) it is estimated
by
(mj)−αqρq(mj) · 2−θmj0 jγ
0 τ (mj0)
jγτ (mj)!
1
q
−α+ 1
2− θmj0
q
q j
0
ρ(j0)
.
Z,m
N
Xj=j0
(here we use Lemma 1 and Remark 2 again). If βg 6 d, then we choose βg > d so
that βg + βv < d + d−θ
(it is possible by (6)), and we set u(ξ) = u(ξ) · 2mj(βg−βg),
q
ξ ∈ VA
j (ξ0). Then
S∞,q
Aξ∗ ,u,w .
Z
S∞,q
Aξ∗ ,u,w · 2mj0(βg−βg) .
Z
2mj0(β−d− d
q )Ψ(2−mj0).
Let, now, q < p < ∞. Let βg,1 + βg,2 = βg, βv,1 + βv,2 = βv,
u1(ξ) = 2mj(βg,1− d(p−q)
p
)Ψg(2−mj), u2(ξ) = 2mj(βg,2− d(q−1)
p
),
w1(ξ) = 2mj(βv,1− d(p−q)
pq )Ψv(2−mj), w2(ξ) = 2mj(βv,2− d
p ), ξ ∈ Vj(ξ0).
Then u1(ξ)u2(ξ) = u(ξ), w1(ξ)w2(ξ) = w(ξ). Applying the Holder inequality, we get
Xξ∈V(Aξ∗ )
wq
1(ξ)wq
2(ξ) Xξ∗6ξ′6ξ
1/q
u1(ξ′)u2(ξ′)f (ξ′)!q
6
37
6
Xξ∈V(Aξ∗ )
6
Xξ∈V(Aξ∗ )
=
Xξ∈V(Aξ∗ )
w
wq
wq
pq
p−q
1
1(ξ)wq
2(ξ) Xξ∗6ξ′6ξ
(ξ) Xξ∗6ξ′6ξ
1(ξ) Xξ∗6ξ′6ξ
p−q
1
u
p
p
u1(ξ′)!q
p
1
p
p
p
q
u
u
q − 1
p−q
1
2 (ξ′)f
(ξ′)!q(1− q
p ) Xξ∗6ξ′6ξ
(ξ′)!q
p
Xξ∈V(Aξ∗ )
p )
Xξ∈V(Aξ∗ )
q (ξ′)!
2(ξ) Xξ∗6ξ′6ξ
2(ξ) Xξ∗6ξ′6ξ
q (1− q
wp
wq
1
p
q
p
q
=
1
p
,
p
p
q
u
2 (ξ′)f
q (ξ′)!q
u2(ξ′)f2(ξ′)!q
q · q
p
p
1
1/q
6
q2
p
with u1(ξ) = u
(then f2 ∈ lq(Aξ∗)).
p−q
1
(ξ), w1(ξ) = w
p−q
1
(ξ), u2(ξ) = u
2 (ξ), w2(ξ) = w
2 (ξ), f2(ξ) = f
q (ξ)
Check that we can apply to each of multipliers the Hardy-type inequality with
(p1, q1) = (∞, q) and (p2, q2) = (q, q). Indeed, for ξ ∈ Vj(ξ0) we have
u1(ξ) = 2mj(βg,1
p
p−q −d)Ψ
p
p−q
g
(2−mj), u2(ξ) = 2mj(cid:16)βg,2
p
q − d
q′(cid:17),
w1(ξ) = 2mj(βv,1
p
p−q − d
q )Ψ
p
p−q
v
(2−mj), w2(ξ) = 2mj(βv,2
p
q − d
q ).
By Remark 2, since the functions Ψg and Ψv satisfy (5) (ρg and ρv satisfy (9),
respectively), we observe that their powers satisfy the similar conditions. First choose
βv,1 and βv,2 so that
βv,1 < (d − θ)(cid:18)1
q
−
1
p(cid:19) ,
βv,2 <
d − θ
p
,
βv,1 + βv,2 = βv
hold (it is possible, since βv < d−θ
require
q ). Then we choose βg,1, βg,2. In the case (7), a)
βg,1 + βv,1 <(cid:18)1 −
q
p(cid:19)(cid:18)d +
d
q
−
It is possible, since βg + βv < d + d
the case (7), b) we require
q − d
θ
q(cid:19) ,
p − θ(cid:16) 1
βg,1 + βv,1 =(cid:18)1 −
q
p(cid:19)(cid:18)d +
d
q
−
The condition α p
p−q > 1−γ
q holds by (10).
Also observe that kf2kq/p
Thus, in the case (7), a)
lq(Aξ∗ ) = kf klp(Aξ∗ ).
βg,2 + βv,2 <
qd
p
,
βg,1 + βg,2 = βg.
q − θ
p . In
q(cid:17) + qd
q − 1
p(cid:17)(cid:16)d + d
p(cid:17) =(cid:16)1 − q
q(cid:19) , βg,2 + βv,2 =
θ
qd
p
.
Sp,q
Aξ∗ ,u,w .
Z h2mj0((βg,1+βv,1) p
p−q −d− d
q )Ψ
p
p−q (2−mj0)i1− q
p ×
38
= 2mj0(β−d− d
q + d
p )Ψ(2−mj0),
×h2mj0((βg,2+βv,2) p
as well as in the case (7), b)
Sp,q
Aξ∗ ,u,w .
Z (cid:18)2− mθj0
− αp
q j
0
p−q + 1
q
q
p
q −d)i
p−q (j0)(cid:19)1− q
ρ
p
p
= 2−θ( 1
q − 1
p)mj0j
q − 1
p
−α+ 1
0
ρ(j0).
This completes the proof.
Lemma 12. Let p < q. Then there exists m∗ = m∗(Z) ∈ N such that for any
m > m∗
Sp,q
Aξ∗ ,u,w .
Z
2mj0(β−d− d
q + d
p )Ψ(2−mj0)
in the case a) of (7); in the case b), if α > 0, then
Sp,q
Aξ∗ ,u,w .
Z
j−α
0 ρ(j0).
Proof. Set λ = 1
Applying the Holder inequality, we get
q , and define the quantity p1 by 1
p − 1
p = 1−λ
p1
+ λ. Then 1
q = 1−λ
p1
.
S :=
wq(ξ) Xξ∗6ξ′6ξ
Xξ∈V(Aξ∗ )
wq(ξ) Xξ∗6ξ′6ξ
6
Xξ∈V(Aξ∗ )
6 max
ξ∈V(Aξ∗ )Xξ′6ξ
=
Xξ∈V(Aξ∗ )
wq(ξ) Xξ∗6ξ′6ξ
f p(ξ′)!λ
lp(Aξ∗ )
6 kf kλp
1
u
1−λ (ξ′)f
Xξ∈V(Aξ∗ )
Xξ∈V(Aξ∗ )
1
with w(ξ) = w
1−λ (ξ), u(ξ) = u
1−λ (ξ), f (ξ) = f
1
1/q
6
=
1/q
p
p
u(ξ′)f
q (ξ′)f 1− p
u(ξ′)f (ξ′)!q
q (ξ′)!q
q(1−λ) (ξ′)!(1−λ)q Xξ∗6ξ′6ξ
1−λ (ξ) Xξ∗6ξ′6ξ
wp1(ξ) Xξ∗6ξ′6ξ
u(ξ′) f (ξ′)!p1
1−λ (ξ′)f
w
u
p1
p
1
q(1−λ) (ξ). We have
1
q
6
1−λ
p1
6
f
1− p
q
λ (ξ′)!λq
q(1−λ) (ξ′)!p1
p
1−λ
p1
,
k fk1−λ
lp1 (Aξ∗ ) =
Xξ∈V(Aξ∗ )
1−λ
p1
f
pp1
q(1−λ)
1
q
=
Xξ∈V(Aξ∗ )
f p(ξ)
39
= kf k
p
q
lp(Aξ∗ ).
Hence,
S 6 kf k
1− p
lp(Aξ∗ ) · (Sp1,p1
q
Aξ∗ ,u, w)1−λkf k
If ξ ∈ Vj(ξ0), then
p
q
lp(Aξ∗ ) = (Sp1,p1
Aξ∗ ,u, w)1−λkf klp(Aξ∗ ).
w(ξ) = 2mj( βv
1−λ − d
(1−λ)q )Ψ
1
1−λ
v
(2−mj) = 2mj(cid:16) βv
1−λ − d
p1(cid:17)Ψ
1
1−λ
v
(2−mj),
u(ξ) = 2mj(cid:16) βg
1−λ − d
(1−λ)p′(cid:17)Ψ
1
1−λ
g
(2−mj).
Therefore,
u(ξ) = 2
mj(cid:18) βg− d
p′
with Ψg = Ψ
1
1−λ
g
, Ψv = Ψ
d − θ
(6)
> 0.
1
1−λ
v
1(cid:19) Ψg(2−mj),
, βg = βg
w(ξ) = 2mj(cid:16) βv− d
p1(cid:17) Ψv(2−mj),
1−λ − d
(1−λ)p′ + d
p′
1
and βv = βv
1−λ. Then − βvp1 +
In the case (7), a), we have β < d + d
< d(1 − λ), i.e., β − d + d
to βg + βv − d
that 1−λ
= 1
p1
p′ + d(1−λ)
q . Hence, by Lemma 10, (Sp1,p1
Consider the case (7), b). Then β = d + d
A,u, w)1−λ .
p′
1
Z
10 we get (Sp1,p1
A,u, w)1−λ .
Z
j−α
0 ρ(j0).
q − d
p . Check that βg + βv < d. It is equivalent
< 0. It remains to observe
p − d(1−λ)
p1
2mj0(β−d− d
q + d
p )Ψ(2−mj0).
q − d
p . Hence, βg + βv = d, and by Lemma
7 The proof of the embedding theorem
In this section we prove the main result of this article. In particular, we obtain
Theorem 1.
Let m∗ = m∗(Z) (see Lemmas 10, 11 and 12), and let {(Dj,i, ξj,i)}j∈Z+, i∈ Ij be
the partition of T for m = m∗ (see the definition on page 18).
Theorem 2. Let D ⊂ T ξj0,i0
any function f ∈ span W r
r − 1 such that
be a subtree with the minimal vertex ξj0,i0. Then for
p,g(Ω) there exists a polynomial P f of degree not exceeding
kf − P f kLq,v(ΩD,F ) .
Z
in the case (7), a),
kf − P f kLq,v(ΩD,F ) .
Z
2−m∗θ( 1
q − 1
∇rf
g (cid:13)(cid:13)(cid:13)(cid:13)Lp(ΩD,F )
2m∗j0(β−δ)Ψ(2−m∗j0)(cid:13)(cid:13)(cid:13)(cid:13)
g (cid:13)(cid:13)(cid:13)(cid:13)Lp(ΩD,F )
ρ(j0)(cid:13)(cid:13)(cid:13)(cid:13)
−α+( 1
j0j
0
∇rf
p )+
p )+
q − 1
(117)
(118)
in the case (7), b). Here the mapping f 7→ P f can be extended to a linear continuous
operator P : Lq,v(Ω) → Pr−1(Ω).
40
Proof. We shall denote Ω = ΩD,F .
Step 1. The set C ∞( Ω) ∩ W r
p,g( Ω) (it can be proved in the same
way as for a non-weighted case, see [44, p. 16]).1 Therefore, it is sufficient to check
(117) and (118) for f ∈ C ∞( Ω).
p,g( Ω) is dense W r
By Lemma 2, Ω ∈ FC(b∗), b∗ = b∗(a, d). Let x∗ ∈ Ω, γx(·), T (x) be such as in
(x∗, ∂ Ω). From assertion 2 of Lemma 2 it follows
Definition 1, and let R0 = distk·kld
that we can take the center of the cube F ( ξj0,i0) as the point x∗.
It is sufficient to show that if f ∈ C ∞(Ωw0), f BR0/2(x∗) = 0, then (117), (118)
hold with P f = 0 (the general case can be proved in the same way as in [62]; here
we can take as f 7→ P f the Sobolev's projection operators).
2
Let ϕ(x) = ∇rf (x)
g(x)
. By Theorem F, for any x ∈ Ω there exists a set Gx ⊂
∪t∈[0, T (x)]Bb∗t(γx(t)) such that
{(x, y) ∈ Ω × Ω : x ∈ Ω, y ∈ Gx}
is measurable,
f (x) .
r,d,aZGx
x − yr−dg(y)ϕ(y) dy.
By Assertion 2 of Lemma 2,
if x ∈ ∆, ∆ ∈ Θ(Ω), then Gx ⊂ Ω6∆.
(119)
Thus, it is sufficient to prove that
ZΩ
vq(x)
ZGx
q
g(y)ϕ(y)x − yr−d dy
dx
1/q
C(j0)kϕkLp( Ω),
.
Z
(120)
with C(j0) = 2m∗j0(β−δ)Ψ(2−m∗j0) in the case (7), a), and
C(j0) = 2−m∗θ( 1
q − 1
p )+
−α+( 1
j0j
0
q − 1
p )+
ρ(j0)
in the case (7), b).
Extending the function ϕ by zero to ΩT ξj0 ,i0
theorem, we may assume that V(D) = {ξ ∈ V(T ξj0,i0
N ∈ N.
,F \ΩD,F and applying the B. Levi's
) : ρT ( ξj0,i0, ξ) 6 N} for some
Step 2. Consider the case r = d. Let (A, ξ0) = (A(m∗), ξ0) be the tree defined
on the page 33. If ξ = ηj,i ∈ V(A), then we set D[ξ] = Dj,i, Ω[ξ] = ΩD[ξ],F ,
gξ = 2βgm∗jΨg(2−m∗j), vξ = 2βvm∗jΨv(2−m∗j).
(121)
1Here C∞(Ω) is the space of functions that are smooth on the open set Ω, but not necessarily
extendable to smooth functions on the whole space Rd.
41
By (48), the property 2 of the partition {Dj,i}j∈Z+, i∈ Ij and (4), we have
diam Ω[ξ] ≍
a,d
2−m∗j, g(x) ≍
Z
gξ, v(x) ≍
Z
vξ, x ∈ ΩD[ξ],F .
(122)
Set ξ∗ = ηj0,i0. Then
q
1/q
(119)
6
q
g(y)ϕ(y) dy
dx
g(y)ϕ(y) dy
dx
ϕ(y) dy
q
ZΩ
6
Xξ∈V(Aξ∗ )ZΩ[ξ]
≍
ξ ZΩ[ξ]
Xξ∈V(Aξ∗ )
.
Xξ∈V(Aξ∗ )
. C(j0)
vq(x)
ZGx
vq(x)
Xξ∗6ξ′6ξ ZΩ[ξ′]
gξ′ ZΩ[ξ′]
Xξ∗6ξ′6ξ
wq(ξ) Xξ∗6ξ′6ξ
Xξ∈V(Aξ∗ )
kϕkp
vq
u(ξ′)kϕkLp(Ω[ξ′])!q
Lp(Ω[ξ])
1/p
1/q
= C(j0)kϕkLp( Ω)
(122)
≍
Z
1/q
dx
(115),(121),(122)
.
Z
.
Z
1/q
(the penultimate relation follows from Lemmas 10, 11 and 12).
Step 3. Let r 6= d. Set
G1
x = {y ∈ Gx :
x − y > 2 dist·(x, Γ)}, G2
x = {y ∈ Gx : x − y < 2 dist·(x, Γ)}.
Then in order to prove (120) it suffices to check the inequalities
ZΩ
ZΩ
x
vq(x)
ZG1
vq(x)
ZG2
x
q
q
g(y)ϕ(y)x − yr−d dy
g(y)ϕ(y)x − yr−d dy
dx
dx
Prove (123). At first we check that for y ∈ G1
x
1/q
1/q
.
Z
.
Z
C(j0)kϕkLp( Ω),
(123)
C(j0)kϕkLp( Ω).
(124)
x − y ≍
a,d
dist·(y, Γ).
(125)
42
Indeed, let zx ∈ Γ, x − zx = dist·(x, Γ). Then
dist·(y, Γ) 6 y − zx 6 y − x + x − zx =
= x − y + dist·(x, Γ) 6 x − y +
x − y
2
=
3x − y
2
.
Prove the inverse inequality. Let y ∈ F (ω), ω ∈ V(T ). From (119) it follows that
2−mω . From assertion 2
x ∈ ΩTω,F . Since ΩTω,F ∈ FC(b∗), we have diam(ΩTω ,F ) .
a,d
2−mω . Hence,
of Theorem A it follows that dist·(y, ∂Ω) ≍
d
dist·(y, Γ) > dist·(y, ∂Ω) ≍
d
2−mω
(21)
&
a,d
diam ΩTω,F > x − y.
Thus, (125) is proved, and
ZΩ
x
vq(x)
ZG1
vq(x)
.
ZG1
ZΩ
g(y)ϕ(y)x − yr−d dy
dx
g(y)ϕ(y) dy
q
x
dx
1/q
q
1/q
(4)
.
Z
,
with g(y) = ϕg(dist·(y, Γ)),
ϕg(t) = ϕg(t) · tr−d = t−βgΨg(t), βg = βg + d − r.
Since β − δ = βg + βv − r − d
estimate which was obtained at the previous step.
p = βg + βv − d − d
q + d
q + d
p , it remains to apply the
Prove (124). If y ∈ G2
x, then
dist·(y, Γ) 6 dist·(x, Γ) + x − y 6 3 dist·(x, Γ).
(126)
Let x ∈ Ω[ηj,i], y ∈ Ω[ηj ′,i′]. From (119) and property 3 of the partition {Dj,i}j∈Z+,i∈ Ij
it follows that j′ 6 j. By (49), dist·(x, Γ) ≍
. This
together with (126) yield that there exists j∗ = j∗(a, d, m∗) such that j −j∗ 6 j′ 6 j.
Notice that
2−m∗j, dist·(y, Γ) ≍
2−m∗j ′
a,d,m∗
a,d,m∗
x − y .
2−m∗j.
a,d,m∗
(127)
Denote by Iηj,i,j∗ the maximal subgraph on the vertex set
V(Iηj,i,j∗) = [j ′>j−j∗ [ηj′ ,i′ 6ηj,i
V(Dηj′,i′ )
43
and set Ω[ηj,i] = ΩIηj,i,j∗ ,F . Then for any x ∈ Ω[ηj,i], the inclusion G2
In addition, from (121) and (122) it follows that
x ⊂ Ω[ηj,i] holds.
g(y) ≍
Z
gηj,i, y ∈ Ω[ηj,i].
(128)
By (107), for any ξ′ ∈ V(Aξ∗)
card {ξ ∈ V(Aξ∗) : ξ′ ∈ Iξ,j∗} .
1.
Z
Therefore,
We have
1/p
Lp( Ω[ξ])
kϕkLp( Ω).
.
Z
(129)
kϕkp
x
Xξ∈V(Aξ∗ )
vq(x)
ZG2
vq(x)
ZΩ[ξ]
ZΩ[ξ]
ξ ZΩ[ξ]
gq
ξ vq
ZΩ
6
Xξ∈V(Aξ∗ )ZΩ[ξ]
≍
Xξ∈V(Aξ∗ )
q
1/q
q
g(y)ϕ(y)x − yr−d dy
g(y)ϕ(y)x − yr−d dy
ϕ(y)x − yr−d dy
dx
dx
dx
q
6
1/q
1/q
(122),(128)
≍
Z
=: S.
Let ξ = ηj,i. By (122) and (127), Ω[ξ] and Ω[ξ] are contained in a ball of radius
Rξ ≍
a,d,m
2−m∗j.
(130)
Applying Theorem E and the Holder inequality, we get
S .
Z
Xξ∈V(Aξ∗ )
If p 6 q, then
gq
ξ vq
ξ Rδq
ξ kϕkq
1/q
Lp( Ω[ξ])
=: S1.
S1 6 max
ξ∈V(Aξ∗ )
gξvξRδ
ξ
Xξ∈V(Aξ∗ )
kϕkp
Lp( Ω[ξ])
44
1
p
(121),(129),(130)
.
Z
. 2m∗j0(β−δ)Ψ(2−m∗j0)kϕkLp( Ω) = C(j0)kϕkLp( Ω).
If p > q, then by the Holder inequality
S1 6
Xξ∈V(Aξ∗ )(cid:0)gξvξRδ
ξ(cid:1)
. ∞
Xj=j0
2−θm∗j0Λ(2−m∗j0)
2−θm∗jΛ(2−m∗j)
(see Lemma 1).
1
q − 1
pq
p−q
Lp( Ω[ξ])
1
q − 1
p
kϕkp
p
Xξ∈V(Aξ∗ )
p−q (2−m∗j)!
pq
2(β−δ) pq
p−q m∗jΨ
1
p
(121),(107),(129),(130)
.
Z
kϕkLp( Ω)
(7)
.
Z
C(j0)kϕkLp( Ω)
Notice that if the condition (7), a) is replaced by βg + βv > δ − θ(cid:16) 1
or if (7), b) holds and α < (1 − γ)(cid:16) 1
unbounded in Lq,v(Ω). Indeed, let ϕ ∈ C ∞
, then the set W r
0 (Ω) is
∇rϕ(x)p dx = 1, let
p,g(Ω) ∩ C ∞
p(cid:17)+
p(cid:17)+
q − 1
q − 1
,
k = k(a, d) ∈ N be such as in Lemma 4, and let ξ be the minimal vertex of the tree
T , ξ ∈ Wν0. Then for sufficiently large ν ∈ N
0 ([0, 1]d), ϕ > 0, R[0, 1]d
Set
ν+k
(32)
&
Z
h(2−ν0)
h(2−ν)
&
Z,ν0
card Wl
Xl=ν
{∆j}j∈Jν =nF (ξ) : ξ ∈ ∪ν+k
l=ν
(24),(26)
2νθ
Λ(2−ν)
Wlo .
.
(131)
Then ∆j = zj + tj[0, 1]d, tj
that
≍
Z
2−ν. In addition, from (4) and (25) it follows
g(x) ≍
Z
2νβgΨg(2−ν),
v(x) ≍
Z
2νβvΨv(2−ν),
x ∈ ∆j,
j ∈ Jν.
(132)
∇rϕν
Let p 6 q. Take j ∈ Jν and set ϕν(x) = cνϕ(cid:16) x−zj
g (cid:13)(cid:13)(cid:13)Lp(Ω)
2νβgΨg(2−ν)2ν( d
= 1. Then cν
(132)
(132)
p −r). Hence,
≍
Z
(cid:13)(cid:13)(cid:13)
kϕνkLq,v(Ω)
≍
Z
cν · 2νβvΨv(2−ν) · 2− νd
q ≍
Z
2ν(β−δ)Ψ(2−ν).
tj (cid:17), with cν > 0 such that
If β − δ > 0, then kϕνkLq,v(Ω) →
ν→∞
ν−αρ(ν) →
ν→∞
∞.
∞. If β = δ and α < 0, then kϕνkLq,v(Ω) ≍
Z
45
Let p > q. First consider the case β > δ−θ(cid:16) 1
where cν > 0 is such that (cid:13)(cid:13)(cid:13)
g (cid:13)(cid:13)(cid:13)Lp(Ω)
(card Jν)− 1
p ,
∇rϕν
q − 1
p(cid:17). Set ϕν(x) = cν Pj∈Jν
(132)
ϕ(cid:16) x−zj
tj (cid:17),
2νβgΨg(2−ν)2ν( d
p −r) ·
≍
Z
= 1. Then cν
kϕνkLq,v(Ω)
(132)
≍
Z
cν · 2νβvΨv(2−ν) · 2− νd
q (card Jν)
1
q ≍
Z
2ν(β−δ)Ψ(2−ν)(card Jν)
1
q − 1
p
(131)
&
Z
& 2ν(β−δ+θ( 1
q − 1
p ))Ψ(2−ν)(cid:0)Λ(2−ν)(cid:1)
1
p − 1
q .
Therefore, kϕνkLq,v(Ω) →
ν→∞
q − 1
Let β = δ − θ(cid:16) 1
∞.
p(cid:17), α < (1 − γ)(cid:16) 1
q − 1
p(cid:17). For s ∈ N denote
Ns = {s + l(k + 1) :
l ∈ Z+,
l(k + 1) 6 s}.
cνϕ(cid:16) x−zj
tj (cid:17), where cν > 0 is such that
= 1 and
p , j ∈ Jν, ν ∈ Ns. Then (cid:13)(cid:13)(cid:13)
p −r)(card Jν)− 1
p s− 1
p
∇rψs
g (cid:13)(cid:13)(cid:13)Lp(Ω)
(133)
Then card Ns ≍
a,d
s. Set ψs(x) = Pν∈Ns Pj∈Jν
p (card Ns)− 1
= (card Jν)− 1
cν
(132)
≍
Z
2νβgΨg(2−ν) · 2ν( d
(cid:13)(cid:13)(cid:13)
∇rψs
g (cid:13)(cid:13)(cid:13)Lp(∆j)
Hence,
kψskLq,v(Ω)
(132)
≍
Z Xν∈Ns
2νβvqΨq
v(2−ν)cq
ν · 2−νdcard Jν!
1
q (8),(131),(133)
&
Z
≍ Xν∈Ns
REFERENCES
2νβqΨq(2−ν) · 2−νδq · 2νθ(1− q
p )ν−γ(1− q
p )[τ (ν)]−1+ q
p s− q
p!
1
q
(8)
≍
Z
≍ s−αρ(s) · s( 1
q − 1
p)(1−γ)[τ (s)]− 1
q + 1
p →
s→∞
∞.
[1] D.R. Adams, "Traces of potentials. II", Indiana Univ. Math. J., 22 (1972/73),
907 -- 918.
[2] D.R. Adams, "A trace inequality for generalized potentials", Studia Math. 48
(1973), 99 -- 105.
[3] K.F. Andersen, H.P. Heinig, "Weighted norm inequalities for certain integral
operators", SIAM J. Math. Anal., 14 (1983), 834 -- 844.
46
[4] F. Antoci, "Some necessary and some sufficient conditions for the compactness
of the embedding of weighted Sobolev spaces", Ricerche Mat. 52:1 (2003), 55 --
71.
[5] O.V. Besov, "On the compactness of embeddings of weighted Sobolev spaces
on a domain with irregular boundary", Tr. Mat. Inst. im. V.A. Steklova, Ross.
Akad. Nauk, 232 (2001), 72 -- 93 [Proc. Steklov Inst. Math. 232 (2001), 66 -- 87].
"Sobolev's embedding theorem for a domain with irregular
[6] O.V. Besov,
boundary," Mat. Sb. 192:3 (2001), 3 -- 26 [Sb. Math. 192 (2001), 323 -- 346].
[7] O.V. Besov, "On the compactness of embeddings of weighted Sobolev spaces on
a domain with an irregular boundary," Dokl. Akad. Nauk 376:6 (2001), 727 -- 732
[Dokl. Math. 63:1 (2001), 95 -- 100].
[8] O.V. Besov,
"Integral estimates for differentiable functions on irregular
domains," Mat. Sb. 201:12 (2010), 69 -- 82 [Sb. Math. 201 (2010), 1777 -- 1790].
[9] O.V. Besov, V.P. Il'in, S.M. Nikol'skii, Integral representations of functions,
and imbedding theorems. "Nauka", Moscow, 1996. [Winston, Washington DC;
Wiley, New York, 1979].
[10] M. Bricchi,
"Existence and properties of h-sets", Georgian Mathematical
Journal, 9:1 (2002), 13(cid:22)32.
[11] M. Bricchi, "Compact embeddings between Besov spaces defined on h-sets",
Funct. Approx. Comment. Math., 30 (2002), 7 -- 36.
[12] A.M. Caetano, S. Lopes, "Spectral theory for the fractal Laplacian in the context
of h-sets", Math. Nachr., 284:1 (2011), 5 -- 38.
[13] D.E. Edmunds, H. Triebel, "Spectral theory for isotropic fractal drums", C. R.
Acad. Sci. Paris S´er. I Math., 326 (1998), 1269 -- 1274.
[14] D.E. Edmunds, H. Triebel,
"Eigenfrequencies of
isotropic fractal drums",
Operator Theory: Advances and Applications, 110 (1999), 81 -- 102.
[15] D.E. Edmunds, H. Triebel, Function spaces, entropy numbers, differential
in Mathematics, 120 (1996). Cambridge
operators. Cambridge Tracts
University Press.
[16] A. El Kolli, "n-i`eme ´epaisseur dans les espaces de Sobolev", J. Approx. Theory,
10 (1974), 268 -- 294.
[17] W.D. Evans, D.J. Harris, "Fractals, trees and the Neumann Laplacian", Math.
Ann., 296:3 (1993), 493 -- 527.
[18] W.D. Evans, D.J. Harris, J. Lang, "Two-sided estimates for the approximation
numbers of Hardy-type operators in L∞ and L1", Studia Math., 130:2 (1998),
171 -- 192.
[19] W.D. Evans, D.J. Harris, J. Lang, "The approximation numbers of Hardy-type
operators on trees", Proc. London Math. Soc. (3) 83:2 (2001), 390 -- 418.
[20] W.D. Evans, D.J. Harris, L. Pick, "Weighted Hardy and Poincar´e inequalities
on trees", J. London Math. Soc., 52:2 (1995), 121 -- 136.
47
[21] V. Gol'dshtein, A. Ukhlov, "Weighted Sobolev spaces and embedding theorems",
Trans. AMS, 361:7 (2009), 3829 -- 3850.
[22] P. Gurka, B. Opic, "Continuous and compact imbeddings of weighted Sobolev
spaces. I", Czech. Math. J. 38(113):4 (1988), 730 -- 744.
[23] P. Gurka, B. Opic, "Continuous and compact imbeddings of weighted Sobolev
spaces. II", Czech. Math. J. 39(114):1 (1989), 78 -- 94.
[24] P. Gurka, B. Opic, "Continuous and compact imbeddings of weighted Sobolev
spaces. III", Czech. Math. J. 41(116):2 (1991), 317 -- 341.
[25] D.D. Haroske, I. Piotrowska, "Atomic decompositions of function spaces with
Muckenhoupt weights and some relation to fractal analysis", Math. Nachr.,
281:10 (2008), 1476 -- 1494.
[26] H.P. Heinig, "Weighted norm inequalities for certain integral operators, II",
Proc. AMS, 95 (1985), 387 -- 395.
[27] Jain Pankaj, Bansal Bindu, Jain Pawan K.,
"Continuous and compact
imbeddings of weighted Sobolev spaces", Acta Sci. Math. (Szeged), 66:3 -- 4
(2000), 665 -- 677.
[28] Jain Pankaj, Bansal Bindu, Jain Pawan K., "Certain imbeddings of Sobolev
spaces with power type weights", Indian J. Math., bf 44:3 (2002), 303 -- 321.
[29] Jain Pankaj, Bansal Bindu, Jain Pawan K., "Certain imbeddings of weighted
Sobolev spaces", Math. Ineq. Appl., 6:1 (2003), 105 -- 120.
[30] J. Kadlec, A. Kufner, "Characterization of functions with zero traces by
integrals width weight functions, I", Casopis. pest. mat., 91 (1966), 463 -- 471.
[31] J. Kadlec, A. Kufner, "Characterization of functions with zero traces by
integrals width weight functions, II", Casopis. pest. mat., 92 (1967), 16 -- 28.
[32] L.D. Kudryavtsev, "Direct and inverse imbedding theorems. Applications to the
solution of elliptic equations by variational methods", Tr. Mat. Inst. Steklova,
55 (1959), 3 -- 182 [Russian].
[33] L.D. Kudryavtsev and S.M. Nikol'skii, "Spaces of differentiable functions of
several variables and imbedding theorems," in Analysis -- 3 (VINITI, Moscow,
1988), Itogi Nauki Tekh., Ser.: Sovrem. Probl. Mat., Fundam. Napravl. 26, pp.
5 -- 157; Engl. transl. in Analysis III (Springer, Berlin, 1991), Encycl. Math. Sci.
26, pp. 1 -- 140.
[34] A. Kufner,
Sobolevschen Raume mit
"Einige Eigenschaften
der
Belegungsfunktionen", Czech. Math. J., 15 (90) (1965), 597 -- 620.
[35] A. Kufner, "Imbedding theorems for general Sobolev weight spaces", Ann.
Scuola Sup. Pisa, 23 (1969), 373 -- 386.
[36] A. Kufner, Weighted Sobolev spaces. Teubner-Texte Math., 31. Leipzig: Teubner,
1980.
[37] A. Kufner, B. Opic, "Remark on compactness of imbeddings in weighted spaces",
Math. Nachr., 133 (1987), 63 -- 70.
48
[38] J. Lehrback,
"Weighted Hardy inequalities beyond Lipschitz domains",
arXiv:1209.0588v1.
[39] G. Leoni, A first Course in Sobolev Spaces. Graduate studies in Mathematics,
vol. 105. AMS, Providence, Rhode Island, 2009.
[40] M.A. Lifshits, "Bounds for entropy numbers for some critical operators", Trans.
Amer. Math. Soc., 364:4 (2012), 1797 -- 1813.
[41] M.A. Lifshits, W. Linde, "Compactness properties of weighted summation
operators on trees", Studia Math., 202:1 (2011), 17 -- 47.
[42] M.A. Lifshits, W. Linde, "Compactness properties of weighted summation
operators on trees (cid:22) the critical case", Studia Math., 206:1 (2011), 75 -- 96.
[43] P.I. Lizorkin and M. Otelbaev, "Imbedding and compactness theorems for spaces
of Sobolev type with weights. I, II", Mat. Sb. 108: 3 (1979), 358 -- 377 [Math.
USSR Sb. 36:3 (1980), 331 -- 349]; Mat. Sb. 112:1 (1980), 56 -- 85 [Math. USSR
Sb. 40:1 (1981), 51- -- 77].
[44] V.G. Maz'ja [Maz'ya], Sobolev spaces (Leningrad. Univ., Leningrad, 1985;
Springer, Berlin -- New York, 1985).
[45] S.D. Moura, Function spaces of generalized smoothness. Dissertationes Math.,
2001. 398:88 pp.
[46] K. Naimark, M. Solomyak, "Geometry of Sobolev spaces on regular trees and
the Hardy inequality", Russian J. Math. Phys., 8:3 (2001), 322 -- 335.
[47] J. Necas, "Sur une m´ethode pour r´esoudre les equations aux d´eriv´ees partielles
dy type elliptique, voisine de la varitionelle", Ann. Scuola Sup. Pisa, 16:4 (1962),
305 -- 326.
[48] I. Piotrowska, "Traces on fractals of function spaces with Muckenhoupt weights",
Funct. Approx. Comment. Math., 36 (2006), 95 -- 117.
[49] I. Piotrowska, "Entropy and approximation numbers of embeddings between
weighted Besov spaces", Function spaces VIII, 173 -- 185. Banach Center Publ.,
79, Polish Acad. Sci. Inst. Math., Warsaw, 2008.
[50] Yu.G. Reshetnyak,
"Integral representations of differentiable functions in
domains with a nonsmooth boundary", Sibirsk. Mat. Zh., 21:6 (1980), 108 -- 116
(in Russian).
[51] Yu.G. Reshetnyak, "A remark on integral representations of differentiable
functions of several variables", Sibirsk. Mat. Zh., 25:5 (1984), 198 -- 200 (in
Russian).
[52] S.L. Sobolev, "On a theorem of functional analysis", Mat. Sb., 4 (46):3 (1938),
471 -- 497 [Amer. Math. Soc. Transl., (2) 34 (1963), 39 -- 68.]
[53] M. Solomyak, "On approximation of functions from Sobolev spaces on metric
graphs", J. Approx. Theory, 121:2 (2003), 199 -- 219.
[54] H. Triebel, "Interpolation properties of ε-entropy and widths. Geometric
characteristics of function spaces of Sobolev -- Besov type", Mat. Sbornik, 98
(1975), 27 -- 41.
49
[55] H. Triebel, Interpolation theory. Function spaces. Differential operators (Dtsch.
Verl. Wiss., Berlin, 1978; Mir, Moscow, 1980).
[56] H. Triebel, Fractals and spectra. Birkhauser, Basel, 1997.
[57] H. Triebel, Theory of function spaces III. Birkhauser Verlag, Basel, 2006.
[58] D.E. Edmunds, W.D. Evans, Hardy Operators, Function Spaces and
Embeddings. Springer-Verlag, Berlin, 2004.
[59] H. Triebel, "Approximation numbers in function spaces and the distribution of
eigenvalues of some fractal elliptic operators", J. Approx. Theory, 129:1 (2004),
1 -- 27.
[60] B.O. Turesson, Nonlinear potential theory and weighted Sobolev spaces. Lecture
Notes in Mathematics, 1736. Springer, 2000.
[61] A.A. Vasil'eva, "Widths of weighted Sobolev classes on a John domain",
Proceedings of the Steklov Institute of Mathematics, 280 (2013), 91 -- 119.
[62] A.A. Vasil'eva, "Widths of weighted Sobolev classes on a John domain: strong
singularity at a point" (submitted to Revista Matematica Complutense).
[63] G.N. Yakovlev, "On a density of finite functions in weighted spaces", Dokl. Akad.
Nauk SSSR, 170:4 (1966), 797 -- 798 [in Russian].
50
|
1602.03982 | 1 | 1602 | 2016-02-12T08:39:41 | Controlled K-frames and their invariance under Compact Perturbation | [
"math.FA"
] | K-frames were recently introduced by L. G\v{a}vruta in Hilbert spaces to study atomic systems with respect to bounded linear operator. Also controlled frames have been recently introduced by Balazs, Antoine and Grybos in Hilbert spaces to improve the numerical efficiency of interactive algorithms for inverting the frame operator. In this manuscript, the concept of controlled K-frames will be studied and the stability of Controlled K-frames under compact perturbation will be discussed. | math.FA | math |
CONTROLLED K-FRAMES AND THEIR INVARIANCE
UNDER COMPACT PERTURBATION
A. RAHIMI1, SH. NAJAFZADEH2 AND M. NOURI3
Abstract. K-frames were recently introduced by L. Gavruta in Hilbert
spaces to study atomic systems with respect to bounded linear opera-
tor. Also controlled frames have been recently introduced by Balazs,
Antoine and Grybos in Hilbert spaces to improve the numerical effi-
ciency of interactive algorithms for inverting the frame operator. In this
manuscript, the concept of controlled K-frames will be studied and the
stability of Controlled K-frames under compact perturbation will be
discussed.
1. Introduction
Frames in Hilbert spaces were first proposed by Duffin and Schaeffer to
deal with nonharmonic Fourier series in 1952 [10] and widely studied from
1986 since the great work by Daubechies et al.[11]. Now frames play an im-
portant role not only in the theoretics but also in many kinds of applications
and have been widely applied in signal processing [14], sampling [12, 13], cod-
ing and communications [18], filter bank theory [3], system modeling [9] and
so on. For special applications many other types of frames were proposed,
such as the fusion frames [5, 6] to deal with hierarchical data processing, g-
frames [19] by Sun to deal with all existing frames as united object, oblique
dual frames [12] by Elder to deal with sampling reconstructions, and etc.
The notion of K-frames were recently introduced by L. Gavruta to study
the atomic systems with respect to a bounded linear operator K in Hilbert
spaces. K-frames are more general than ordinary frames in sense that the
lower frame bound only holds for the elements in the range of the K, where
K is a bounded linear operator in a separable Hilbert Space H.
One of the newest generalization of frames is controlled frames. Controlled
frames have been introduced recently to improve the numerical efficiency of
interactive algorithms for inverting the frame operator on abstract Hilbert
spaces [1], however they have been used earlier in [2] for spherical wavelets.
This concept generalized for fusion frames in [16] and for g-frames in [17].
In this paper, the concept of controlled K-frame will be defined and it will
be shown that any controlled K-frame is equivalent to a K-frame, finally we
will discuss the stability of compact perturbation for controlled K-frames.
2000 Mathematics Subject Classification. Primary 42C40; Secondary 41A58, 47A58,.
Key words and phrases. Bessel
sequence, Controlled frame, Frame, K-frame,
Perturbation.
1
2
A. RAHIMI1, SH. NAJAFZADEH2 AND M. NOURI3
Throughout this paper H is a separable Hilbert space, B(H) is the family
of all linear operators on H, GL(H) denotes the set of all bounded linear
operators which have bounded inverses and K ∈ B(H).
GL(H). Let GL+(H) be the set of all positive operators in GL(H).
It is easy to see that if S, T ∈ GL(H), then T ∗, T −1 and ST are also in
A bounded operator T ∈ B(H) is called positive (respectively, non-
negative), if hT f, fi > 0 for all f 6= 0 (respectively, hT f, fi ≥ 0 for all
f ). Every non-negative operator is clearly self-adjoint.
If A ∈ B(H) is
non-negative, then there exists a unique non-negative operator B such that
B2 = A. Furthermore B commutes with every operator that commutes
2 . Let B+(H) be the set of positive
with A. This will be denoted by B = A
operators on H. For self-adjoint operators T1 and T2, the notation T1 ≤ T2
or T2 − T1 ≥ 0 means
1
hT1f, fi ≤ hT2f, fi
,∀f ∈ H.
The following result is needed in the sequel, but straightforward to prove:
Proposition 1.1. [8] Let T : H → H be a linear operator. Then the
following condition are equivalent:
(1) There exist m > 0 and M < ∞, such that mI ≤ T ≤ M I;
(2) T is positive and there exist m > 0 and M < ∞, such that mkfk2 ≤
1
kT
2 fk2 ≤ Mkfk2 for all f ∈ H;
2 ∈ GL(H);
1
(3) T is positive and T
(4) There exists a self-adjoint operator A ∈ GL(H), such that
A2 = T ;
(5) T ∈ GL+(H);
(6) There exist constants m > 0 and M < ∞ and operator
(7) For every C ∈ GL+(H), there exist constants m > 0 and
C ∈ GL+(H), such that m′C ≤ T ≤ M′C;
M < ∞, such that m′C ≤ T ≤ M′C.
It is well-known that not all bounded operators U on a Hilbert space H
are invertible: an operator U needs to be injective and surjective in order
to be invertible. For doing this, one can use right-inverse operator. The
following lemma shows that if an operator U has closed range, there exists
a right-inverse operator U† in the following sense:
Lemma 1.2. [8] Let H1 and H2 be Hilbert spaces and suppose that U :
H2 → H1 is a bounded operator with closed range RU . Then there exists a
bounded operator U† : H1 → H2 for which
U U†x = x ,∀x ∈ RU .
The operator U† in the Lemma 1.2 is called the pseudo-inverse of U . In
the literature, one will often see the pseudo-inverse of an operator U with
closed range defined as the unique operator U† satisfying that
NU † = R⊥U , RU † = N⊥U , U U†x = x ,∀x ∈ RU .
CONTROLLED K-FRAMES AND THEIR INVARIANCE UNDER ...
3
A sequence {fi}i∈I in H is called a frame for H, if there exist constants
0 < A ≤ B < ∞ such that
Akfk2 ≤ X
i∈I
hf, fii2 ≤ Bkfk2 , ∀f ∈ H.
If A = B, then {fi}i∈I is called a tight frame and if A = B = 1, then it is
called a Parseval frame. A Bessel sequence {fi}i∈I is only required to fulfill
the upper frame bound estimate but not necessarily the lower estimate.
The frame operator Sf = Pi∈Ihf, fiifi associated with a frame {fi}i∈I
is a bounded, invertible and positive operator on H. This provides the
reconstruction formulas
f = S−1Sf = X
i∈I
hf, fiiS−1fi = X
i∈I
hf, S−1fiifi,∀f ∈ H.
Furthermore, AI ≤ S ≤ BI and B−1I ≤ S−1 ≤ A−1I.
Definition 1.3. Let C ∈ GL(H). A frame controlled by the operator C or
C-controlled frame is a family of vectors {fi}i∈I in H, such that there exist
constants 0 < mC ≤ MC < ∞, verifying
mCkfk2 ≤ X
i∈I
hf, fiihCfi, fi ≤ MCkfk2 , ∀f ∈ H.
The controlled frame operator S is defined by
Sf = X
i∈I
hf, fiiCfi,∀f ∈ H.
Because of the higher generality of K-frames, some properties of ordinary
frames can not hold for K-frames, such as the frame operator of a K-frame
is not an isomorphism. For more differences between K-frames and ordinary
frames, we refer to [20].
Definition 1.4. Let K ∈ B(H). A sequence {fn}∞n=1 ⊂ H is called a
K-frame for H, if there exist constants A, B > 0 such that
(1.1)
AkK∗fk2 ≤
∞
X
n=1
hf, fni2 ≤ Bkfk2, ∀f ∈ H.
we call A and B lower and upper frame bound for K-frame {fn}∞n=1 ⊂
H, respectively if only the right inequality of the above inequality holds,
{fn}∞n=1 ⊂ H is called a K-Bessel sequence.
Remark 1.5. If K = I, then K-frame are just the ordinary frame.
Remark 1.6. In the following, we will assume that R(K) is closed, since this
can assure that the pseudo-inverse K† of K exists.
Definition 1.7. [15] Let K ∈ B(H). A sequence {fn}∞n=1 ⊂ H is called an
atomic system for K, if the following conditions are satisfied:
(1) {fn}∞n=1 is a Bessel sequence.
4
A. RAHIMI1, SH. NAJAFZADEH2 AND M. NOURI3
(2) For any x ∈ H, there exists ax = {an} ∈ l2 such that
Kx =
anfn
∞
X
n=1
where kaxkl2 ≤ Ckxk, C is positive constant.
Suppose that {fn}∞n=1 is a K-frame for H. Obviously it is a Bessel se-
quence, so we can define the following operator
T : l2 → H, T a =
∞
X
n=1
anfn,
a = {an} ∈ l2,
it follows that
T ∗ : H → l2
T ∗f = {hf, fni}∞n=1,∀f ∈ H.
Let S = T T ∗, we obtain
Sf =
∞
X
n=1
hf, fnifn
,∀f ∈ H.
we call T, T ∗ and S the synthesis operator, analysis operator and frame
operator for K-frame {fn}∞n=1, respectively.
Theorem 1.8. Let {fn}∞n=1 be a Bessel sequence in H. Then {fn}∞n=1 is a
K-frame for H, if and only if there exists A > 0 such that
S ≥ AKK∗,
where S is the frame operator for {fn}∞n=1.
Proof. The sequence {fn}∞n=1 is a K-frame for H with frame bounds A, B
and frame operator S if and only if
AkK∗fk2 ≤
∞
X
K=1
(1.2)
that is
hf, fni2 = hSf, fi ≤ Bkfk2 , ∀f ∈ H,
hAKK∗f, fi ≤ hSf, fi ≤ hBf, fi , ∀f ∈ H.
so the conclusion holds.
(cid:3)
Remark 1.9. Frame operator of a K-frames is not invertible on H in gen-
eral, but we can show that it is invertible on the subspace R(K) ⊂ H.
In fact, since R(K) is closed, there exists a pseudo-inverse K† of K, such
that KK†f = f , ∀f ∈ R(K) , namely KK†R(K) = IR(K), so we have
I∗R(K) = (K†R(K))∗K∗. Hence for any f ∈ R(K), we obtain
kfk = k(K†R(K))∗K∗fk ≤ kK†k.kK∗fk,
that is, kK∗fk2 ≥ kK†k−2kfk2. Combined with (1.2) we have
hSf, fi ≥ AkK∗fk2 ≥ AkK†k−2kfk2 , ∀f ∈ R(K).
(1.3)
CONTROLLED K-FRAMES AND THEIR INVARIANCE UNDER ...
5
So, from the definition of K-frame we have
(1.4)
AkK†k−2kfk ≤ kSfk ≤ Bkfk , ∀f ∈ R(K),
which implies that S : R(K) → S(R(K)) is a homeomorphism, furthermore,
we have
B−1kfk ≤ kS−1fk ≤ A−1kK†k2kfk , ∀f ∈ S(R(K)).
2. Controlled K-frames
Controlled frames for spherical wavelets were introduced in [2] to get a
numerically more efficient approximation algorithm and the related theory.
For general frames, it was developed in [1]. For getting a numerical solu-
tion of a linear system of equations Ax = b, one can solve the system of
equations P Ax = P b, where P is a suitable preconditioning matrix. It was
the main motivation for introducing controlled frames in [2]. Controlled
frames extended to g-frames in [17] and for fusion frames in [16]. In this
section, the concept of controlled frames and controlled Bessel sequences
will be extended to K-frames and it will be shown that controlled K-frames
are equivalent K-frames.
Definition 2.1. Let C ∈ GL+(H) (C > 0) and let CK = KC. The family
{fn}∞n=1 is called C-controlled K-frame for H, if {fn}∞n=1 is a K-Bessel
sequence and there exist constants A > 0 and B < ∞ such that
hf, fnihf, Cfni ≤ Bkfk2 , ∀f ∈ H.
2 K∗fk2 ≤
∞
X
n=1
1
AkC
The constants A and B are called C-controlled K-frame bounds. If C = I,
the C-controlled K-frame {fn}∞n=1 is a K-frame for H with bounds A and
B.
If the second part of the above inequality holds, it called C-controlled
K-Bessel sequence with bound B.
The proof of the following lemmas is straightforward.
Lemma 2.2. Let C > 0 and C ∈ GL+(H). The K-Bessel sequence {fn}∞n=1
is C-controlled K-Bessel sequence if and only if there exists constant B < ∞
such that
∞
X
n=1
hf, fnihf, Cfni ≤ Bkfk2 , ∀f ∈ H.
Lemma 2.3. Let C ∈ GL+(H). A sequence {fn}∞n=1 ∈ H is a C-controlled
Bessel sequence for H if and only if the operator
LC : H → H , LC f =
∞
X
n=1
hf, fniCfn,
∀f ∈ H.
6
A. RAHIMI1, SH. NAJAFZADEH2 AND M. NOURI3
is well defined and there exists constant B < ∞ such that
hf, fnihf, Cfni ≤ Bkfk2 , ∀f ∈ H.
∞
X
n=1
Remark 2.4. The operator LC : H → H , LC f = P∞n=1hf, fniCfn, f ∈ H
is called the C-controlled Bessel sequence operator, also LC f = CSf .
The following lemma characterizes C-controlled K-frames in term of their
operators.
Lemma 2.5. Let {fn}∞n=1 be a C-controlled K-frame in H, for C ∈ GL+(H).
Then
AIkC
1
2 K†k2 ≤ LC ≤ BI.
Proof. Suppose that {fn}∞n=1 is a C-controlled K-frame with bounds A and
B. Then
AkC
1
2 K∗fk2 ≤
∞
X
n=1
hf, fnihf, Cfni ≤ Bkfk2 , ∀f ∈ H.
For f ∈ H
i.e.
AkC
1
2 K∗fk2 ≤ hf, LC fi ≤ Bkfk2
AkC
1
2 K∗k2I ≤ LC ≤ BI.
(cid:3)
The following proposition shows that for evaluation a family {fn}∞n=1 ⊂ H
to be a controlled K-frame it is suffices to check just a simple operator
inequality.
Proposition 2.6. Let {fn}∞n=1 be a Bessel sequence in H and C ∈ GL+(H).
Then {fn}∞n=1 is a C-controlled K-frame for H if and only if there exists
A > 0 such that CS ≥ CAKK∗.
Proof. The sequence {fn}∞n=1 is a controlled K-frame for H with frame
bounds A, B and frame operator S, if and only if
AkC
1
2 K∗fk2 ≤
∞
X
n=1
hf, fnihf, Cfni ≤ Bkfk2 , ∀f ∈ H.
That is,
hCAKK∗f, fi ≤ hCSf, fi ≤ hBf, fi, ∀f ∈ H.
(cid:3)
Proposition 2.7. Let {fn}∞n=1 be a C-controlled K-frame and C ∈ GL+(H).
Then {fn}∞n=1 is a K-frame for H.
CONTROLLED K-FRAMES AND THEIR INVARIANCE UNDER ...
7
Proof. Suppose that {fn}∞n=1 is a controlled K-frame with bounds A and B.
Then for any f ∈ H
2 K∗fk2
2 K∗fk2
hf, fnihf, C 0fni
1
1
2 C
≤ AkC
≤ kC
AkK∗fk2 = AkC− 1
2k2kC− 1
2k2 ∞
X
n=1
2k2 ∞
X
n=1
= kC
1
1
hf, fni2.
Hence for f ∈ H,
AkC
1
2k−2kK∗fk2 ≤
∞
X
n=1
hf, fni2
On the other hand for every f ∈ H,
∞
X
n=1
hf, fni2 = hf, Sfi
1
2 fi
1
1
= hf, C−1CSfi
= h(C−1CS)
= k(C−1CS)
≤ kC− 1
= kC− 1
≤ kC− 1
2 f, (C−1CS)
2 fk2
2k2k(CS)
2k2hf, CSfi
2k2Bkfk2.
2 fk2
1
2k2.
These inequalities yields that {fn}∞n=1 is a K-frame with bounds AkC
and BkC− 1
Proposition 2.8. Let C ∈ GL+(H) be a self adjoint and KC = CK, if
{fn}∞n=1 is K-frame for H, then {fn}∞n=1 is a C-controlled K-frame for H.
Proof. Suppose that {fn}∞n=1 be a K-frame with bounds A′ and B′. Then
for all f ∈ H
2k−2
(cid:3)
1
A′kK∗fk2 ≤
hf, fni2 ≤ B′kfk2.
∞
X
n=1
A′kC
1
2 K∗fk2 = A′kK∗C
1
2 fk2 ≤
∞
X
n=1
hC
1
2 f,
= hC
1
2 f, fnihC
∞
X
n=1
1
hfn, C
1
2 f, fni
1
2 fifni
1
2 f, C
= hC
2 Sfi = hf, CSfi.
8
A. RAHIMI1, SH. NAJAFZADEH2 AND M. NOURI3
1
2 K∗fk2 ≤ hf, CSfi for every f ∈ H. On the other hand for
Hence A′kC
every f ∈ H,
hf, CSfi2 = hC∗f, Sfi2 = hCf, Sfi2 ≤ kCfk2kSfk2 ≤ kCk2kfk2Bkfk2.
Hence
A′kC
1
2 K∗fk2 ≤ hf, CSfi ≤ B′kCkkfk2.
Therefore {fn}∞n=1 is a C-controlled K-frame with bounds A′ and B′kCk.
(cid:3)
3. Compact Perturbation for Controlled K-frames
One of the most important problems in the studying of frames and its
applications specially on wavelet and Gabor systems is the invariance of
these systems under perturbation. At the first, the problem of perturbation
studied by Paley and Wiener for bases and then extended to frames.There
are many versions of perturbation of frames in Hilbert spaces, Banach space,
Hilbert C∗-modules and etc. In the last decade, several authors have gener-
alized the Paley-Wiener perturbation theorem to the perturbation of frames
in Hilbert spaces. The most general result of these was the following ob-
tained by Casazza and Christensen [4].
Theorem 3.1. [4] Let {xj}j∈J be a frame for a Hilbert space H with frame
bounds C and D. Assume that {yj}j∈J is a sequence of H and that there
exist λ1, λ2, µ > 0 such that max{λ1 + µ√C
, λ2} < 1. Suppose one of the
following conditions holds for any finite scalar sequence {cj} and every x ∈
H. Then {yj}j∈J is also a frame for H.
1
1
(1) (Pj∈J hx, xj−yji2)
µkxk
(2) k Pn
i=1 cj2)
Moreover, if {xj}j∈J is a Riesz basis for H and {yj}j∈J satisfies (2), then
2 ≤ λ1(Pj∈J hx, xji2)
i=1 cj xjk+λ2k Pn
2 +λ2(Pj∈J hx, yji2)
i=1 cj yjk+µ(Pn
i=1 cj(xj−yj)k ≤ λ1k Pn
1
2 +
1
2
{yj}j∈J is also a Riesz basis for H.
Another type of the perturbation of frames is compact perturbation that
appeared in the paper [7] by Christensen and Heil:
Theorem 3.2. [7] Let {xj}j∈J be a frame for a Hilbert space H and {yj}j∈J
be a sequence in H. If the operator
K : ℓ2 → H, K{cj} = X cj(xj − yj)
is well-defined compact operator, then {yj}j∈J is a frame sequence.
The perturbation theorem investigated by X. Xiao, Y. Zhu, L. Gavruta
to K-frames [20].
CONTROLLED K-FRAMES AND THEIR INVARIANCE UNDER ...
9
Theorem 3.3. [20] Suppose that {fn}∞n=1 is a K-frame for H, and α, β ∈
[0,∞], such that max{α + γ√A−1kK +k, β} < 1.
If {gn}∞n=1 ⊂ H and satisfy
ck(fk − gk)k ≤ αk
ckfkk + βk
ckgkk + γ(
ck2)
k
2 ,
n
n
n
n
1
X
k=1
X
k=1
X
k=1
X
k=1
for any ci, i ∈ N, then {gn}∞n=1 is a PQ(R(K))K-frame for H, with frame
bounds
[√AkK +k−1(1 − α) − γ]2
,
(1 + β)2kKk2
[√B(1 + α) + γ]2
,
(1 − β)2
where PQ(R(K)) is a orthogonal projection operator for H to Q(R(K)),
Q = U T ∗, T, U are synthesis operator for {fn}∞n=1 and {gn}∞n=1 respectively.
Motivating the above theorems, we prove compact perturbation for con-
trolled K-frames.
Theorem 3.4. Let F = {fk}k∈I be a controlled K-frame for H, with op-
If G = {gk}k∈I is a sequence in H
erator S and frame bounds AF , BF .
and E = TF − TG be a compact operator, where TG{ck}k∈I = Pk∈I ckgk for
{ck}k∈I ∈ ℓ2, then G = {gk}k∈I is a controlled K-frame for H.
Proof. Let {fk}k∈I be a controlled K-frame with bounds AF , BF , then
kTFk2 ≤ BF . Let V = TF −E be an operator from l2(I) into H. Because TF
and E are bounded, then operator V is bounded. Therefore kV k = kV ∗k.
For any f ∈ H,
V ∗f = T ∗f − E∗f = {hf, fki}k∈I − {hf, fk − gki}k∈I
= {hf, fki}k∈I − {hf, fki − hf, gki}k∈I = {hf, gki}k∈I .
V ({ck}k∈I ) = X
k∈I
ckgk , SG = V V ∗.
hf, CSGfi = hf, CV V ∗fi = hC
2 V fk2 = kC
2k2kV fk2 = kC
1
1
= kC
1
2 V f, C
1
2 V fi
1
2k2k(TF − E)fk2
Therefore,
Therefore,
hf, CSGfi ≤ kTF − Ek2kfk2kC
1
2k2
≤ (kTFk2 + 2kTFkkEk + kEk2)kfk2kC
≤ (BF + 2pBFkEk + kEk2)kfk2kC
2k2
= BF (1 + kEk√BF
)2kfk2kC
2k2.
1
1
1
2k2
This inequality shows that {gk}k∈I is a K-Bessel sequence with bound
BF (1 + kEk√BF
)2kC
2k2.
1
10
A. RAHIMI1, SH. NAJAFZADEH2 AND M. NOURI3
In the next step, we prove that SG = V V ∗ is a surjective operator. We
have,
V V ∗ = (TF − E)(TF − E)∗ = (TF − E)(T ∗F − E∗)
= TF T ∗F − TF E∗ − ET ∗F + EE∗
= SF + EE∗ − TF E∗ − ET ∗F
s.t SF = TF T ∗F .
Since E, TF and SF are compact operators, then (EE∗ − TF E∗ − ET ∗F )S−1
is a compact operator. Therefore (EE∗ − TF E∗− ET ∗F )S−1
F + I is a bounded
operator with closed range. Thus, V V ∗ = EE∗ − TF E∗ − ET ∗F + SF is
a bounded operator with closed range. Therefore V V ∗ is an operator on
It is clear that V V ∗ is a injective. By lemma 1.2 it can
span{gk}k∈I .
be deduced that RV V ∗ = N†V V ∗ = span{gk}k∈I. Then SG is a surjective
operator. Therefore G = {gk}k∈I is a Controlled K-frame for span{gk}k∈I .
F
(cid:3)
References
[1] P. Balazs, J. P. Antoine, A. Grybos, Wighted and Controlled Frames, Int. J. Wavelets
Multiresolut. Inf. Process. 8(1) (2010) 109-132.
[2] I. Bogdanova, P. Vandergheynst, J .P . Antoine, L. Jacques, M. Morvidone, Stereo-
graphic wavelet frames on the sphere, Applied Comput. Harmon. Anal. (19) (2005)
223-252.
[3] H. Bolcskei, F. Hlawatsch , H. G. Feichtinger,Frame- theoretic analysis of over- sam-
pled filter banks, IEEE Trans. Signal process. 46 (1998) 3256-3268.
[4] P. Casazza, O. Christensen, Perturbation of operators and applications to frame the-
ory, J. Fourier Anal. Appl. 3 (1997) 543557.
[5] P. G. Casazza, G. Kutyniok, Frames of subspaces. Wavelets, frames and operator
theory, College Park, MD,Contemp. Math., vol.345. American Mathematical Society,
Providence,(2004) 87-113.
[6] P.G. Casazza, G. Li .S. Kutyniok, Fusion frames and distributed processing, Appl.
Comput. Harmon. Anal.25 (2008) 114-132.
[7] O. Christensen, C. Heil, Perturbation of Banach frames and atomic decompositions,
Math. Nach. 185 (1997) 33-47.
[8] O. Christensen, An introduction to frames and Riesz bases, Birkhauser, Boston 2003.
[9] N. E. Duday Ward, J. R. Partington, A construction of rational wavelets and frames
in Hardy-Sobolev space with applications to system modelling.SIAM.J.Control Op-
tim.36 (1998) 654-679.
[10] R.J. Duffin, A.C. Schaeffer, A class of nonharmonic Fourier series, Trans. Math.
Soc.72 (1952) 341-366.
[11] I. Daubechies, A. Grossmann, Y. Meyer, Painless non orthogonal expansions, J. Math.
Phys. 27(1986) 1271-1283.
[12] Y. C. Eldar, Sampling with arbitrary sampling and reconstruction spaces and oblique
dual frame vectors, J. Fourier. Anal. Appl. 9(1) (2003) 77-96.
[13] Y. C. Eldar, T. Werther, General framework for consistent sampling in Hilbert spaces,
Int. J. walvelets Multi. Inf. Process. 3(3) (2005) 347-359.
[14] P. J. S. G. Ferreira, Mathematics for multimedia signal processing II: Discrete fi-
nite frames and signal reconstruction, In: Byrnes, J.s. (ed.) signals processing for
multimedia, PP. 35-54.IOS press, Amsterdam (1999).
[15] L. Gavruta, Frames for operators, Appl. Comput. Harmon. Anal. 32 (2012) 139-144.
CONTROLLED K-FRAMES AND THEIR INVARIANCE UNDER ...
11
[16] A. Khosravi, K. Musazadeh, Controlled fusion frames, Meth. Func. Anal. Topol. Vol.
18 (2012), no. 3, 256265.
[17] A. Rahimi, A. Fereydooni, Controlled G-Frames and Their G-Multipliers in Hilbert
spaces , Analele Stiintifice ale Universitatii Ovidius Constanta, vol. 2(12), (2013),
223-236.
[18] T. Strohmer, R. Jr. Heath, Grass manian frames with applications to coding and
communications, Appl. Comput. Harmon. Anal. 14 (2003) 257- 275.
[19] W. Sun, G-frames and g-Riesz bases, J. Math. Anal. 322 (2006) 437-452.
[20] X. Xiao, Y. Zhu, L. Gavruta, Some Properties of K-Frames in Hilbert Spaces , Re-
sults. Math. 63 (2013), 1243-1255.
1Department of Mathematics, University of Maragheh, Maragheh, Iran.
E-mail address: [email protected]
2Department of Mathematics, Payame Noor University, Iran.
E-mail address: [email protected]
3Department of Mathematics, Payame Noor University, Iran.
E-mail address: [email protected]
|
1703.06441 | 1 | 1703 | 2017-03-19T14:17:14 | The Hautus test for non-autonomous linear evolution equation | [
"math.FA"
] | In this paper, we investigate the Hautus test for evolution equation with the operators depending on time. | math.FA | math |
The Hautus test for non-autonomous linear evolution
equation
Duc-Trung Hoang
Institute Mathematics of Bordeaux, France.
[email protected]
In this paper, we investigate the Hautus test for evolution equation with the
Abstract
operators depending on time.
1
Introduction
Controllability and observability are basis concepts in system theory and control
theory. They are important structural properties which have close relationships
with the stability of state feedback controllers abd state observers. In this paper,
we will study the controllability, the observabilty, the duality between these two
concepts for the non autonomous linear system. These properties were studied
well for the autonomous system. Let H be a Hilbert space. Considering U (t, s)
the evolution family of two variables generating by the family of operators A(t):
A(t) : D(A(t)) 7→ H. Let U be another Hilbert space and suppose C : H → U is a
linear operator. We consider the system :
x′(t) + A(t)x(t) = B(t)u(t)
x(0)
y(t)
= x0
= Cx(t)
(1.1)
For simplicity, we denote the above system as (A(t), B(t), C). We always assume
that the family of operator A(t) is bounded from H 7→ H. The solution is defined
as
x(t) = U (t, 0)x0 +Z t
0
U (t, s)B(s)u(s)ds,
t ≥ s
Definition 1.1.
A family {U (t, s)t,s} operators is called an evolution family if it satisfies the
following conditions :
(i) U (t, t)x = x for all t ≥ 0 and x ∈ H;
(ii) U (t, s) = U (t, r).U (r, s) for all t ≥ r ≥ s ≥ 0 Such an evolution family is called
continous if there exist M, ω > 0 such that
(iii) kU (t, s)k ≤ M eω(t−s)
(iv) U (t, s)x is jointly continuous with respect to t, s and x
1
Definition 1.2. The system (1.1) is said to be exactly controllable at time τ if for
every x0, x1 in H, there exist u ∈ L2(0, τ ; U ) such that the solution satisfy x1 = x(τ )
Definition 1.3. The system (1.1) is said to be exactly null controllable at time τ if
for every x0 in H, there exist u ∈ L2(0, τ ; U ) such that the solution satisfy x(τ ) = 0
Definition 1.4. The system (1.1) is said to be observable in [0, τ ] if the map O :
H → L2(0, τ ; U ) : x0 → y(.) is injective.
The definition express the fact that we can recover uniquely the initial state
from a knowledge of the output y(.) in the time interval [0, τ ]. When the system
is observable, we refer to (C, A(t)) as an observable pair. For one variable s fixed,
A(s) generate a strongly continuous semigroup Ts(t). We assume that the domaine
D(A(t)) is densed in H and independent of t. We consider the adjoint system
z′(t) − A(t)∗z(t) = 0
= zτ
z(τ )
= B(t)∗x(t)
h(t)
(1.2)
Russell and Weiss ([5]) showed that a necessary condition for exact observability
of exponentially systems is the following Hautus test : There exits a constant m > 0
such that for every s ∈ C− and every x ∈ D(A). The Hautus test can be use
for approximate observability of exponentially stable systems [3], for polynomially
stable system [4], for exact observability of strongly stable Riesz-spectral systems
with finite dimensional output spaces [5], and for exponentially stable C0−groups
[6]
k(sI − A)xk2 + ReskCxk2 ≥ Res2kxk2
where C− denotes the open left half plane.
2 Duality of controllability and observability
Let A(t) be such that the uncontrolled initial value problem
(cid:26) x′(t) + A(t)x(t) = 0
x(0)
= x0
(2.1)
admits a evolution (solution) family U (t, s). We observe that
U (t, s + h) − U (t, s) = U (t, s + h)[I − U (s + h, s)]
and so, dividing by h > 0 and letting h → 0+,
d
ds+ U (t, s) = −U (t, s)A(s)
Under mild extra conditions, this derivative will exist in both directions. Now we
take adjoints:
d
ds+hU (τ, s)∗zτ , xi = d
ds+hzτ , U (τ, s)xi = hzτ ,−U (τ, s)A(s)xi = h−A(s)∗U (τ, s)∗zτ , xi.
2
This holds for all x, so we may drop duality pairing and obtain that z(t) := U (τ, t)∗zτ
will solve the dual final time problem
(cid:26) z′(t)−A(t)∗z(t) = 0
z(τ )
= zτ
2.1 Duality
Now consider
(cid:26) x′(t) + A(t)x(t) = B(t)u(t)
= 0
x(0)
(2.2)
(2.3)
and assume, that the map Ψτ : L2(0, τ ; U ) → H defined by u 7→ x(τ ) is continuous
(i.e. that B(t)0≤t≤τ is an admissible family of control operators for the evolution
equation). Assume further exact controllability, i.e. that for any xτ ∈ X, we can
find some u ∈ L2(0, τ ; U ) such that the solution of the initial value problem (2.7)
satisfies x(τ ) = xτ . Then Ψτ is bounded and surjective.
Ψτ u = R τ
0 U (τ, s)B(s)u(s) ds and so (Ψ∗
τ x∗)(t) = B(t)∗U (τ, t)∗x∗.
According to the "standard lemma", Ψτ is surjective iff Ψ∗
τ allows lower esti-
mates, i.e. iff
δkx∗k ≤ kB(t)∗U (τ, t)∗x∗kL2(0,τ )
slow solution "a la main"
By the open mapping theorem we then have a constant C > 1 such that kukL2 ≤
Ckxτk. We can therefore simply let zτ = xτ in (2.2), and consider
d
dthx(t), z(t)i = hB(·)u(·) − A(·)x(·), z(t)i + hx(·), A(·)∗z(·)i = hB(·)u(·), z(t)i.
Integrating from 0 to τ (recall x(0) = 0), one obtains
0 hu(s), B(s)∗z(s)i ds
Now, by Cauchy-Schwarz and the hypothesis kukL2 ≤ Ckzτk
H = kzτk2
kxτk2
H = Z τ
kzτk2
H ≤ kukL2kB(·)∗z(·)kL2(0,τ ;H) ≤ CkzτkH(cid:16)Z τ
0 kB(s)∗z(s)k2
H ds(cid:17).
(2.4)
Dividing by kzτk, this gives the "observability estimate" of the adjoint problem
(2.2), that is, the estimate
kzτkH ≤ C(cid:16)Z τ
0 kB(s)∗z(s)k2
H ds(cid:17).
(2.5)
For the converse direction we assume (2.6), i.e. exact observability of the dual
system. We aim to obtain surjectivity of Ψτ .
Theorem 2.1. The system is exactly controllable on 0 ≤ τ < ∞ if and only if there
exists C > 0 such that for all x ∈ H, we have
kxτkH ≤ C(cid:16)Z τ
0 kB(s)∗U (τ, s)xk2
H ds(cid:17)
(2.6)
3
For the converse, we define the controllability Gramian as Wt = Ψτ Ψ∗
τ =
R t
0 U (t, s)B(s)B(s)∗U (t, s)∗ds be the operator depending on t. Now assuming that
kzτkH ≤ C(cid:16)R τ
H ds(cid:17). We have
0 kB(s)∗z(s)k2
kzτkH ≤ C(cid:16)Z τ
0 kB(s)∗U (τ, s)∗zτk2
H ds(cid:17)
Or
kzτk2
H ≤ C 2kΨ∗
τ zτk2 = C 2hΨ∗
τ zτ , Ψ∗
τ zτi = C 2hΨτ Ψ∗
τ zτ , zτi = hWτ zτ , zτi
Hence, we conclude that Wτ is self-adjoint, injective and coercive operator. Then
Wτ is boundedly invertible. Hence, Im(Wτ ) = D(Wτ )−1 = H). This implies Im(Ψτ )
= H because H = Im(Wτ ) ⊂ Im(Ψτ ). This indicates the controllability of the initial
system
2.2 Necessary condition
If we take G(t, s) = U (t, s)B(s)B(s)∗U (t, s)∗ be the function of two variables s, t
with 0 ≤ s ≤ t. Then
d
dt
G(t, s) = A(t)G(t) + G(t)A(t)∗
Now if we take the integral from 0 to t with respected to the variable s, we have:
The controllability W (t) is the unique solution of the equation
d
dt
W (t) = A(t)W (t) + W (t)A(t)∗ + B(t)B(t)∗
Noting that the operator Wt = ΨtΨ∗
t . We assume that Wτ is not invertible. Since,
Wτ ≥ 0, there exists the sequence zn ∈ H such that kznk = 1 and hzn, Wτ zni → 0.
It follows that
Z t
0
aukU (τ, s)∗B(s)∗znk2dt → 0
We also have a noting that the control function u(t) and the out put function
satisfy the following
Z τ
0 hu(t), y(t)i = 0
2.3 Null controllability
The system
(cid:26) x′(t) + A(t)x(t) = B(t)u(t)
= 0
x(0)
(2.7)
is exactly null controllable on [0, τ ] if for all x0, we can find u ∈ L2(0, τ ; U ) such
that
0 = U (τ, 0)x0 +Z τ
0
U (τ, s)B(s)u(s)ds
4
Ran(U (τ, s)) ⊂ Ran(u → Z τ
0
U (τ, s)B(s)u(s)ds)
We define the operator S : x0 → U (τ, s)x0 and T : u → R τ
0 U (τ, s)B(s)u(s)ds.
Lemma 2.2. ( see [6]) Suppose that Z1, Z2, Z3 are Hilbert spaces, the operators
F ∈ L(Z1, Z3) and G ∈ L(Z2, Z3). Then the following statements are equivalent:
(a) Ran(F ) ⊂ Ran(G)
(b) There exists a c > 0 such that kF ∗zkZ1 ≤ ckG∗zkZ2 for all z ∈ Z3
(c) There exist an operator U ∈ L(Z1, Z2) such that F = GU
We will assume that A(t)A(s)−1 are uniformly bounded for s, t ∈ [0,∞), A(∞)
and limt→∞k(A(t) − A(∞))A(0)−1k = 0. Then by [1] (Theorem 8.1, chapter 5.8)
there exists constant M ≥ 0 and v > 0 such that
kU (t, s)k ≤ M e−v(t−s)
Hence the operator S : x0 → U (τ, s)x0 is a bounded operator from H → H
Lemma 2.3. The operator T : u → R τ
from L2([0, τ ]; U ) → H
Proof. We have:
0 U (τ, s)B(s)u(s)ds is bounded linear map
0
0
U (τ, s)B(s)u(s)ds
0 kU (τ, s)B(s)u(s)kH ds
T (u) = Z τ
≤ Z τ
≤ Z τ
≤ M e−vτkB(t)kL(U,H)Z τ
≤ M e−vτkB(t)kL(U,H)(Z τ
≤ M e−vτkB(t)kL(U,H)
M e−v(τ −s)kB(s)kL(U,H)ku(s)kH ds
evsku(s)kH ds
e2vsdsZ τ
e2vτ − 1
v
0
1
2
(
0
H )1/2
0 ku(s)k2
)1/2ku(s)kL2([0,τ ];U )
Since S, T are bounded operator, using the lemma there exists a constant c > 0
such that
By computation: T ∗x = B∗(s)U (τ, s)∗x∗ and S∗x = U (τ, s)∗x∗. Then we obtain
kS∗xk ≤ ckT ∗xk
the inequality:
kB∗(s)U (τ, s)∗x∗k ≥ kU (τ, s)∗x∗k
5
2.4 Minimum cost controls
We have xτ = U (τ, 0)x0 +R τ
u = B(s)∗U (τ, s)∗W −1
τ
0 U (τ, s)B(s)u(s)ds. It is easy to check that the control
We will indicate that u takes the L2 minimum-norm. Suppoing that both u and
(xτ − U (τ, 0)x0) satisfies the equation.
u satisfy the equation. Then we have
Z τ
0
U (τ, s)B(s)(u(s) − u(s))ds = 0
For all η ∈ H, we have
hZ τ
0
If we choose η = W −1
τ
U (τ, s)B(s)(u(s) − u(s))ds, ηi = 0
(xτ − U (τ, 0)x0), then
This impiles kuk2
L2 ≥ kuk2
L2 . In fact,
hu(s) − u(s), u(s)i = 0
kuk2
0 hu, uids
0 hB(s)∗U (τ, s)∗W −1
L2 = Z τ
= Z τ
= hWτ (xτ − U (τ, 0)x0, (xτ − U (τ, 0)x0i
= kW ∗
τ (u)k
τ
(xτ − U (τ, 0)x0), B(s)∗U (τ, s)∗W −1
τ
(xτ − U (τ, 0)x0)ids
3 The Hautus test
Observe that
d/ds(cid:16)e−λsU (t, s)x(cid:17) = −λe−λsU (t, s)x − e−λsU (t, s)A(s)x
and so, integrating on [0, t],
−CU (t, s)x = e−λtCx +Z t
0
CU (t, s)(λ + A(s))xe−λs ds
If we have exact observability, i.e. δkxk ≤ kCU (t, 0)xkL2(0,τ ), this gives
δkxk ≤ kCxk/p2ℜ(λ) + kt 7→ Z t
0
CU (t, s)(λ + A(s))xe−λs dskL2
6
However, for g ∈ L2 of norm one, using admissibility (!) of C ∗ for U (t, s)∗,
hZ t
Z τ
0 Z t
0 kCU (t, s)(λ + A(s))xe−λskds.gi = (cid:12)(cid:12)(cid:12)(cid:12)
0 hCU (t, s)(λ + A(s))xe−λs, g(t)i ds dt(cid:12)(cid:12)(cid:12)(cid:12)
0 h(λ + A(s))x e−λs,Z τ
Z τ
= (cid:12)(cid:12)(cid:12)(cid:12)
U (t, s)∗C ∗g(t)i dt ds(cid:12)(cid:12)(cid:12)(cid:12)
≤ M Z τ
0 k(λ + A(s))x e−λskHkgkL2(s,τ ) ds
≤ M Z τ
0 k(λ + A(s))x e−λskH ds
s
we obtain the Hautus condition,
δkxk ≤ kCxk√2ℜ(λ)
+ M Z τ
0 k(λ + A(s))xe−λskH ds
Re(λ) > 0, x ∈ \s
D(A(s))
as a necessry condition for exact observability. Remark:
in case A(s) = A
this collapses down to the Hautus test of Russell-Weiss.
Remark 3.1. If A(s) ∈ B(H) for all s, we do not know whether we have "IFF" as
in the autonomous case.
3.0.1 The sufficient condition
We consider the case when C(s) = C. Supposing that C is admissible operator and
satisfy the inequality :
δkxk ≤ kCxk√ℜ(λ)
≤ kCxk√ℜ(λ)
+Z τ
0 k(λ + A(s))xe−λskH ds
Reλ Z τ
0 k(λ + A(s))Reλxe−λskH ds
+
1
Here we assume Reλ > η > 0, and uniformly stable. If Reλ < w, we use
k(λ + A(s))−1xk ≤
C
Reλkxk
even it is still true for C = 0. We have the following theorem
Theorem 3.2. (Alan's) Let D : Ω → L(X, Y ) be an operator-valued function ana-
lytic in an open set Ω ⊂ C. If D(λ) is left (resp.right) invertible for every λ ∈ Ωm
then there is an analytic operator function E : Ω → L(Y, X) such that
E(λ)D(λ) = IX
Proof. see (?)
Lemma 3.3. If we have R b
that λ(E) > 0 and kf (s)xk ≥ δ
b−akxk
a kf (s)xkds ≥ δkxk. There exist a subset E ⊂ [a, b] such
7
Proof. By using contradiction, it is easy to verify.
Due to the lemma, there exist a non-null set E ⊂ [0, τ ] such that for all λ ∈ C
and s ∈ E, we have
δ
2kxk ≤
Cx
√Reλ
+
1
Reλk(A(s) + λ)e−λsxk
We have the map x 7→ (Cx, (A(s) + λ)e−λsx) is left-invertible for 0 ≤ s ≤ ∞
and λ ∈ C. Hence there exists the analytic functions Us(λ) and Vs(λ) satisfying the
equation
Vs(λ)(A(s) + λ)e−λsx + Vs(λ)Cx = I
Vs(λ)e−λsx + Vs(λ)C(A(s) + λ)−1 = (A(s) + λ)−1
Intergating both sides we get
0 +
1
2πi Zλ=r
Xk
Vs(λ)C(A(s) + λ)−1dλ = I
Vk(s)CA(s)k = I
Then the map x 7→ (CA(s)kx)k≥0 is left-invertible. Now we suppose that the system
is not exactly observablem, then there exits a sequence znn≥1 such that kznk = 1
and hzn, Qzni → 0 .
Theorem 3.4. (Vitali's theorem)
Let fn(z) be a sequence of functions, each regular in a region D, let fn(z) ≤ M
for every n and z in D, and let fn(z) tend to a limit as n → ∞ at a set of points
having a limit point inside D. Then fn(z) tends uniformly to a limit in any region
bounded by a contour interior to D, the limit therefore being an analytic function
of z.
fn(λ) = C(t)U (λ, s)zn
kfn(.)kL∞ ≤ M on an open set D. We have fn(t) → 0 on the set with accu-
mulation points. By the Vitali's theoremm fn is uniformly convergent to f on a
compact subset of D.
Hypothese 3.5. The evolution family U (λ, s) is holomorphic. If A(t) is bounded
uniformly, could we infer that U (λ, s) is holomorphic.
Then there exists a subsequence of functions fnkl
such that fnkl → 0 uniformly
on a compact subset of D. The contour integral of fn at the point λ = ω is defined
as
1
2πi ZD
fn(λ)
(λ − ω)
= fn(λ)
8
Differentiating fn for n times at the point λ = ω gives
1
2πi ZD
fn(λ)
(λ − ω)n+1 = (
d
dλ
)nfn(λ)λ=ω = CA(t)nU (λ, s)xnλ=ω = CA(ω)nU (ω, s)xn
Since fn(λ) → 0 uniformly, CA(ω)nU (ω, s)xn → 0 uniformly, we have
WkCA(s)U (ω, s)xn = xn
Xk
Theorem 3.6. If C is admissible and A is boundedm then C is bounded.
Proof. First noting that, if f is C 1 and α-Holder function for α > 0 then we have
f (0) =
1
σ Z σ
0
f (s)ds −Z σ
0
In fact, Let σ > ǫ > 0. We have
1
t2 (Z t
0
f (t) − f (s)ds)dtf orallσ > 0
(3.1)
Z σ
ǫ
1
t2 (cid:18)Z t
0
ǫ
ǫ
ǫ
ǫ
0
0
t
t
1
f (t)
f (t)
f (t) − f (s) ds(cid:19) dt = Z σ
= Z σ
= Z σ
= Z σ
= Z σ
σ Z σ
σ Z σ
t2 (cid:18)tf (t) −Z t
dt −Z σ
ǫ Z t
dt −Z σ
0 Z σ
f (s)Z σ
dt −Z σ
dt −Z σ
f (s) ·(cid:18)
ǫ Z ǫ
f (s) ds −
f (s) ds −Z σ
ǫ Z ǫ
f (s) ds =
⇐⇒
f (t)
f (t)
ǫ
1
=
1
1
1
t
t
0
0
0
0
0
0
ǫ
f (s) ds(cid:19) dt
f (s)
t2 ds dt
f (s)
t2 dt ds
max{s,ǫ}
dt
t2 ds
max{s,ǫ}
1
max{s, ǫ} −
f (s) ds
1
σ(cid:19) ds
1
t2 (cid:18)Z t
0
f (t) − f (s) ds(cid:19) dt
Now let ǫ → 0, we get the result. Now if we take f (t) = CU (τ, t)x then
Cx =
0
1
1
σ Z σ
σ Z σ
CU (τ, t)xdt +Z σ
CU (τ, t)xdt +Z σ
(Z t
t2 (Z t
(Z r
t2 (Z t
1
1
0
0
s
0
0
0
0
=
CU (τ, r)A(r)xdr)ds)dt
CU (τ, r)A(r)xds)dr)dt
By triangle inequality,
kCxk ≤
1
√τ
(Z τ
0 kCU (τ, t)xk2)
1
2 +Z σ
0
1
t2 Z t
0
rkCU (τ, r)A(r)xkdrdt
9
By Cauchy-Swart inequality and use the fact that C be a admissible operator, we
have
rkCU (τ, r)A(r)xkdr ≤ (Z t
r2dr)
1
2 (Z t
0 kCU (τ, r)A(r)xdrk)
1
2
Z t
0
So, we have
0
1
√τ Z t
≤
0 kCU (τ, r)A(r)xdrk)
1
2 ≤
Mτ√τ kxk
kCxk ≤ (
Mτ√τ
+
2Mτ√τ
√3
kAkkxk)
Hence, C is a bounded operator.
4 Hautus test for the case of fix parameter
Lemma 4.1. If C is an admissible operator, i.e
Z τ
0 kCT (t)xk2dt ≤ Mkxk2
for all x ∈ H. Then we have
Z τ
0 kT (t)∗C ∗ykH dtk ≤ M√τkyk
for all y ∈ L2(0, τ ; H)
Proof. We have
(cid:12)(cid:12)Z τ
0 hCT (t)x, yiU dt(cid:12)(cid:12) = (cid:12)(cid:12)Z T
0 hx, T (t)∗C ∗yiU dt(cid:12)(cid:12) =(cid:12)(cid:12)hx,Z τ
≤(Z τ
0 kCT (t)xk2
U )
√Mkxk√τkykU
≤
T (t)∗C ∗yiH dt(cid:12)(cid:12)
2√τykU
0
1
Then
kZ τ
0
T (t)∗C ∗ydtkH = supxh
,Z τ
0
x
kxk
T (t)∗C ∗ydtiHdt ≤
√M√τkykU
Theorem 4.2. Suppose the operators A(t) is analytics in L(H). Suppose that for
all s ∈ [0, τ ] : (C, e−tA(s))t≥0 is exactly observable. Then we have (C, U (t, .)) is also
exactly observable.
10
Supposing that (C, (A(s) + λ)2e−tA(s))t≥0 is exactly observable. We denote
D(t) = (A(s) + λ)2e−tA(s). We obtain the inequality
m2kxk ≤ 2kCx.e−λtk2
L2 +Z τ
0 k(λ + (A(s) + λ)2e−A(s)s)e−(λ)sxkds
By triangle inequality
m2kxk ≤ 2kCx.e−λtk2
≤ 2 kCx.e−λtk2
≤ 2 kCx.e−λtk2
≤ 2 kCx.e−λtk2
0 k(λe−λs)xkL∞ +Z τ
L2 +Z τ
L2 + (1 − e−τ λ)kxk + k(A + λ)xkH Z τ
L2 + (1 − e−τ λ)kxk + k(A + λ)xkH (1 − e−τ (A+λ)
L2 + (1 − e−τ λ)kxk + k(A + λ)xkH (1 − e−τ (A+λ)
0 k(A + λ)2e−(A+λ)sxkH ds
0 k(A + λ)e−(A+λ)sxkH ds
Therefore, we can refer that:
(m2 + e−τ λ − 1)kxk ≤ 2kCx.e−λtk2
L2 + k(A + λ)xkH
The admissibility of observable operator C means that for some τ > 0, there exists
M ≥ 0 such that for any x ∈ D(A(s)),
Z τ
0 kCT (t)xk2dt ≤ M.kxk2
Since (C, e−tA(s))t≥0 is exactly observable,
0
m2kxk ≤ kCT (t)xk =kC(e−λtx +Z t
=2kCx.e−λtk2
=2kCx.e−λtk2
=2kCx.e−λtk2
≤2kCx.e−λtk2
≤2kCx.e−λtk2
T (t − s)(λ + A)e−λsds)k2
L2
0
L2)(H)
L2 + 2kt 7→ Z t
L2 + supkhkL2 ≤1Z τ
L2 + supkhkL2 ≤1Z τ
L2 +Z τ
L2 +Z τ
CT (t − s)(λ + A)xe−λsdsk2
0 Z t
0 h(λ + A)xe−λs,Z τ
0 k(λ + A(s)x)ke−Re(λ)s.kZ τ
0 k(λ + A(s)x)ke−Re(λ)s.Mkhk2
0 h(λ + A)xe−λs, T (t − s)∗C ∗h(t)dti
T (t − s)∗C ∗h(t)dti
T (t − s)∗C ∗h(t)kH
L
s
s
Finally, we obtain
m2
2 + M 2kxk2 ≤
Re(λ)kCxk2 +
For all λm there exists δλ positive such that
2
1
2
(Reλ)2k(λ + A)xk2
δλkxk2 ≤ kCxk2 + k(λ + A)xk2
Moreover, the functions Cx and (λ + A)x) are holormophic over the whole complex
plane. So that, the map x 7→ (Cx, (λ + A)x) is left-invertible and entire.
11
Theorem 4.3. (Alan's)
Let D : Ω → L(X, Y ) be an operator-valued function analytic in an open set
Ω ⊂ C. If D(λ) is left (resp.right) invertible for every λ ∈ Ωm then there is an
analytic operator function E : Ω → L(Y, X) such that
E(λ)D(λ) = IX
Hence there exists the analytic functions Us(λ) and Vs(λ) satisfying the equation
Us(λ)(A + λ)x + Vs(λ)Cx = x
Us(λ)x + Vs(λ)C(A + λ)−1x = (A + λ)−1x
We represent Vs(λ) = P+∞
1
0 +
Vs(λ)C(A + λ)−1dλ = I
k=0 λjVk(s). Intergating both sides we get
2πi Zλ=r
Xk
Vk(s)CAk = I
Then the map x 7→ (CAkx)k≥0 is left-invertible.
exist m > 0 such that
Now we suppose that the system is not exactly observable, i.e there does not
Z τ
0 kU (τ, t)∗Czk2dt ≥ mkzk2
for all z ∈ H, then there exists a sequence znn≥1 such that kznk = 1 and hzn, Qzni →
0 where Q = R +∞
0 C(s)U (τ, s)U ∗(τ, s)C ∗(s)ds
Theorem 4.4. (Vitali's theorem)
Let fn(z) be a sequence of functions, each regular in a region D, let fn(z) ≤ M
for every n and z in D, and let fn(z) tend to a limit as n → ∞ at a set of points
having a limit point inside D. Then fn(z) tends uniformly to a limit in any region
bounded by a contour interior to D, the limit therefore being an analytic function
of z.
fn(λ) = C(t)U (λ, s)zn
kfn(.)kL∞ ≤ M on an open set D. We have fn(t) → 0 on the set with accu-
mulation points. By the Vitali's theoremm fn is uniformly convergent to f on a
compact subset of D.
Hypothese 4.5. The evolution family U (λ, s) is holomorphic. If A(t) is bounded
uniformly, could we infer that U (λ, s) is holomorphic.
Then there exists a subsequence of functions fnk1
such that fnkl → 0 uniformly
on a compact subset of D. The contour integral of fn at the point λ = ω is defined
as
12
1
2πi ZD
fn(λ)
(λ − ω)
= fn(λ)
Differentiating fn for n times at the point λ = ω gives
1
2πi ZD
fn(λ)
(λ − ω)k+1 = (
d
dλ
)nfn(λ)λ=ω = CA(t)nU (λ, s)xnλ=ω = CA(ω)nU (ω, s)xn
fn(λ) → 0 uniformly on δD. So that
kCAkU (ω, s)xnk = k
1
2πi ZδD
fn(λ)
(λ − ω)k+1k ≤ kZδD
maxDfn(λ)
(λ − ω)k+1 dλk ≤ αn.
2πr
rk+1 = αn
2π
rn
where αn → 0 when n → +∞, and r > 1. Therefore, CAkU (ω, s)xn → 0 when
n → +∞. Using the estimation Pk Vk(s)CAk = I
VkCAkU (ω, s)xn = U (ω, s)xn
Xk
We finally obtain kU (ω, s)xnk → 0 when n → +∞. That is a contradiction because
we already assumed that kxnk = 1 for all n. As a result, the system (C, A) is exact
observable.
References
[1] A. Pazy, Semigroups of Linear Operators and Applications to Partial Differen-
tial Equations, ESAIM, 1983.
[2] B. Jacob, R. Schnaubelt, Observability of polynomially stable systems, Control
Lett., 56 (2007), pp. 277-284.
[3] B. Jacob, H. Zwart, observability of diagonal systems with a finite-dimensional
output operator, Control Lett., 43 (2001), pp. 101-109.
[4] B. Jacob, H. Zwart, On the Hautus test for exponentially stable C0-groups,
SIAM J.Control Optim, vol. 48, No.3, pp 1275-1288.
[5] D. L. Russell, G. Weiss, A general necessary condition for exact observability,
SIAM J. Control Optim, 32 (1), 123, 1994.
[6] M. Tucsnak, G. Weiss, Observation and Control for Operator Semigroups,
Birkhauser Verlag, Basel, 2009.
13
|
1801.04915 | 2 | 1801 | 2018-01-21T20:25:44 | Phillips symmetric operators and their extensions | [
"math.FA",
"math-ph",
"math-ph"
] | Let $S$ be a symmetric operator with equal defect numbers and let $\mathfrak{U}$ be a set of unitary operators in a Hilbert space $\mathfrak{H}$. The operator $S$ is called $\mathfrak{U}$-invariant if $US=SU$ for all $U\in\mathfrak{U}$. Phillips \cite{PH} constructed an example of $\mathfrak{U}$-invariant symmetric operator $S$ which has no $\mathfrak{U}$-invariant self-adjoint extensions. It was discovered that such symmetric operator has a constant characteristic function \cite{KO}. For this reason, each symmetric operator $S$ with constant characteristic function is called a \emph{Phillips symmetric operator}. | math.FA | math |
PHILLIPS SYMMETRIC OPERATORS AND THEIR
EXTENSIONS
SERGII KUZHEL,1 ∗ and LEONID NIZHNIK2
Abstract. Let S be a symmetric operator with equal defect numbers and
let U be a set of unitary operators in a Hilbert space H. The operator S
is called U-invariant if U S = SU for all U ∈ U. Phillips [21] constructed
an example of U-invariant symmetric operator S which has no U-invariant
self-adjoint extensions. It was discovered that such symmetric operator has a
constant characteristic function [13]. For this reason, each symmetric operator
S with constant characteristic function is called a Phillips symmetric operator.
The paper is devoted to the investigation of self-adjoint (and, more generally,
proper) extensions of a Phillips symmetric operator. Such extensions differ
from those that are commonly studied in the literature and they have a lot
of curious properties. In particular, proper extensions of a Phillips symmetric
operator that have real spectrum are similar to each other.
1. Introduction
Let S be a symmetric operator with equal defect numbers and let U be a family
of unitary operators in a Hilbert space H such that the inclusion U ∈ U implies
U∗ ∈ U. The operator S is called U-invariant if S commutes with all U ∈ U.
Does there exist at least one U-invariant self-adjoint extension of S? The answer
is definitely affirmative if S is assumed to be semibounded and the Friedrichs
extension of S gives the required example.
In general case of non-semibounded operators, R. Phillips constructed a sym-
metric operator S and a family U of unitary operators commuting with S such
that the U-invariant S has no U-invariant self-adjoint extensions [21, p. 382].
Precisely, the mentioned symmetric operator S acts in the Hilbert space l2(Z)
and it is defined as the Cayley transform S = i(V + I)(V − 1)−1 of the isometric
right shift operator
V {xn}n∈Z = {xn+1}n∈Z, D(V ) = {{xn}n∈Z ∈ l2(Z) : x0 = 0}.
The family U consists of unitary operators Uθ{xn}n∈Z = {yn}n∈Z (θ = 1), where
yn = θxn (n ∈ N) and yn = xn (n ∈ Z \ N). The operator S is U-invariant but
there are no U-invariant self-adjoint extensions of S.
It was discovered [13] that the characteristic function of the symmetric oper-
ator constructed in the Phillips work is a constant in the upper half-plane C+.
This fact can be used for the general definition of Phillips symmetric operators.
Namely, we will say that a symmetric operator S with equal defect numbers is
2010 Mathematics Subject Classification. Primary 47B25; Secondary 47A10.
Key words and phrases. symmetric operator, characteristic function, wandering subspace,
bilateral shift, Lebesgue spectrum.
1
2
S. KUZHEL and L. NIZHNIK
a Phillips symmetric operator (PSO) if its characteristic function is an operator-
constant on C+.
The concept of characteristic function of a symmetric operator was firstly in-
troduced by Shtraus [23] and, further, substantially developed by Kochubei [14]
on the base of boundary triplets technique [9]. Section 2 contains all necessary
results about characteristic functions which are used in the paper.
The present paper is devoted to the investigation of PSO as well as theirs self-
adjoint (and, more generally, proper1 extensions). Such self-adjoint extensions
differ from those that are commonly studied in the literature [1] and they have a
lot of curious properties.
Our original definition of PSO deals with the concept of characteristic function.
In many cases, an explicit calculation of a characteristic function is technically
complicated. For this reason, in Section 3, we establish equivalent descriptions of
PSO (Theorems 3.1, 3.4, Corollary 3.6) which can be employed as independent
definitions of PSO. These results lead to the conclusion that each simple2 PSO
coincides with orthogonal sum of simple maximal symmetric operators having
the same nonzero defect numbers in upper C+ and lower C
− half planes. Such
kind of decomposition means that every simple PSO S is unitary equivalent to
the momentum operator with one point interaction
S = i
d
dx
,
D(S) = {u ∈ W 1
2 (R, N) : u(0) = 0}
acting in the Hilbert space L2(R, N), where the dimension of the auxiliary Hilbert
space N coincides with the defect number of S.
Section 4 is devoted to proper extensions of PSO. The main result (Theorem
4.2) states that all proper extensions of a PSO S with real spectrum are similar
to each other. In fact we can say more: each proper extension with real spectrum
can be interpreted as self-adjoint extension of S for a special choice of inner
product equivalent to the initial one.
Some properties of PSO with defect numbers < 1, 1 > were established in [4].
In particular, analogues of Theorems 3.4, 4.2, and Corollary 4.3 were proved.
In Section 5, PSO are determined as the restrictions of a given self-adjoint op-
erator A. According to Theorem 5.2, Phillips symmetric operators which can be
obtained in this way are in one-to-one correspondence with the wandering sub-
spaces L of the Cayley transform U of A. This means that the set of restrictions
of A contains PSO only in the case where A has a reducing subspace H0 such
that A0 = A ↾D(A)∩H0 is a self-adjoint operator in H0 with Lebesgue spectrum on
R. The existence of a simple PSO is equivalent to the fact that A has Lebesgue
spectrum on R (Corollary 5.3).
In Section 6, examples of PSO are considered. We establish a useful (in our
opinion) characterization of wavelets as functions from the defect subspace N−i
of the specially chosen PSO (Proposition 6.1).
1an extension A of a symmetric operator S is called proper if S ⊂ A ⊂ S ∗
2a symmetric operator is called simple if its restriction to any nontrivial reducing subspace
is not a self-adjoint operator
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
3
Results of Sections 4-6 show that one-point interaction of the momentum op-
erator: i d
dx + αδ(x − y) leads to self-adjoint operators with Lebesgue spectrum
which are unitary equivalent to each other. This means that one should consider
more complicated perturbations of the momentum operator for the construction
of self-adjoint operators with non-trivial spectral properties.
In this way, self-
adjoint momentum operators acting in two intervals were studied in [12, 20].
The momentum operators defined on oriented metric graphs were investigated in
[8]. General nonlocal point interactions for first order differential operators were
introduced and studied in [3, 18].
In Section 7, we continue investigations of [3] and to focus on special classes of
perturbations which can be characterized as one point interaction defined by the
nonlocal potential γ ∈ L2(R).
2. Characteristic functions of symmetric operators
Let A be a linear operator acting in a Hilbert space H. Its domain is denoted
D(A), while A ↾D stands for the restriction of A onto a set D. Here and in the
following we denote by C+ (C
I. Let S be a closed symmetric densely defined operator with equal defect
numbers acting in a separable Hilbert space H with inner product (·,·) linear in
the first argument.
We denote by Nλ = ker(S∗ − λI) the defect subspaces of S and consider the
−) the open upper (resp. lower) half plane.
linear spaces
According to the von Neumann formulas (see, e.g., [15, 22]) each proper exten-
Mλ = Nλ +Nλ,
λ ∈ C \ R.
sion A of S is uniquely determined by the choice of a subspace M ⊂ Mλ:
Let us set M = Nλ in (2.1) and denote by
A = S∗ ↾D(A),
D(A) = D(S) +M,
(2.1)
(2.2)
Aλ = S∗ ↾D(Aλ), D(Aλ) = D(S) +Nλ,
λ ∈ C \ R
the corresponding proper extensions of S. The operators sign(Im λ)Aλ are max-
imal dissipative3 and A∗λ = Aλ. The resolvent set of every maximal dissipative
operator contains C
−. For this reason, the operator-function
Sh(λ) = (Aλ − iI)(Aλ + iI)−1 ↾Ni: Ni → N−i,
λ ∈ C+
(2.3)
is well-defined and it coincides with the characteristic function of symmetric op-
erator S defined by A. Shtraus [23].
Another (equivalent) definition of Sh(·) in [23] is based on the relation
D(Aλ) = D(S) +Nλ = D(S) +(I − Sh(λ))Ni,
λ ∈ C+,
(2.4)
which allows one to determine uniquely Sh(·).
The explicit construction of Sh(·) deals with the calculation of Nλ that, some-
times, is technically complicated. This inconvenience was overcame in [14] with
the use of boundary triplet technique. We recall [15, 19] that a triplet (H, Γ−, Γ+),
3An operator A is called dissipative if Im(Af, f ) ≥ 0 for all f ∈ D(A) and maximal dissipa-
tive if there are no dissipative extensions of A.
4
S. KUZHEL and L. NIZHNIK
where H is an auxiliary Hilbert space and Γ± are linear mappings of D(S∗) into
H, is called a boundary triplet of S∗ if
(S∗f, g) − (f, S∗g) = i[(Γ+f, Γ+g)H − (Γ−f, Γ−g)H],
f, g ∈ D(S∗)
(2.5)
holds and the map (Γ−, Γ+) : D(S∗) → H ⊕ H is surjective.
of operators Aλ in (2.2) admit the presentation
Let a boundary triplet (H, Γ−, Γ+) be given. Then the domains of definition
D(Aλ) =(cid:26)f ∈ D(S∗) :
Θ(λ)Γ+f = Γ−f, λ ∈ C+
Γ+f = Θ(λ)Γ−f, λ ∈ C
− (cid:27)
(2.6)
where Θ(·) is an operator in H.
The operator-valued function Θ(·) defined on C \ R is called the characteristic
function of S associated with boundary triplet (H, Γ−, Γ+). It follows from the
relation A∗λ = Aλ and (2.5) that Θ∗(λ) = Θ(λ).
The explicit form of characteristic function depends on the choice of a boundary
triplet. However, in any case, Θ(·) is a holomorphic operator-valued function
whose values are strong contractions in H (i.e., kΘ(λ)k < 1) [14].
The characteristic function determines a simple symmetric operator up to uni-
tary equivalence. Namely, the following result holds:
Theorem 2.1 ([14]). Simple symmetric operators S1 and S2 are unitary equiva-
lent if and only if some of their characteristic functions coincide.
The Shtraus characteristic function Sh(·) defined in (2.3) coincides (up to the
multiplication by unitary operator) with Θ(·) for special choice of boundary
triplet. Precisely, the simplest (inspired by the von Neumann formulas) boundary
triplets (Nµ, Γ−, Γ+) of S∗ can be constructed as follows
Γ−f =p2Im µV fµ, Γ+f =p2Im µfµ, f = u + fµ + fµ ∈ D(S∗),
(2.7)
where µ ∈ C+ and V : Nµ → Nµ is an arbitrary unitary mapping. Assume
that µ = i. Then, the characteristic function Θ(·) associated with the boundary
triplet (Ni, Γ−, Γ+) coincides with the function −V Sh(·) on C+.
Remark 2.2. There are various approaches to the definition of boundary triplets.
For instance, [9, 22], a triplet (H, Γ0, Γ1) where Γ0, Γ1 are linear mappings of
D(S∗) into H, is called a boundary triplet of S∗ if the Green's identity
f, g ∈ D(S∗)
(S∗f, g) − (f, S∗g) = (Γ1f, Γ0g)H − (Γ0f, Γ1g)H,
The operators Γ± in (2.5) and Γi are related as follows Γ± = 1√2
holds and the map (Γ0, Γ1) : D(S∗) → H ⊕ H is surjective.
(Γ1±iΓ0) and,
obviously, the definitions of boundary triplets (H, Γ−, Γ+) and (H, Γ0, Γ1) are
equivalent.
II. The characteristic function Θ(·) admits a natural interpretation in the Krein
space setting (see [2, 5] for the basic theory of Krein spaces and terminology). To
explain this point, we fix a boundary triplet (H, Γ−, Γ+) and rewrite (2.5) as:
(S∗f, g) − (f, S∗g) = i[Ψf, Ψg],
(2.8)
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
5
(2.9)
(2.10)
(2.11)
where
maps D(S∗) into the Krein space (H, [·,·]) with the indefinite inner product
Ψ =(cid:20)Γ+
Γ−(cid:21) : D(S∗) → H =(cid:20)H
H(cid:21) ,
x =(cid:20)x0
x1(cid:21) , y =(cid:20)y0
y1(cid:21) ∈ H.
[x, y] = (x0, y0) − (x1, y1),
It follows from the definition of boundary triplets that the mapping Ψ : D(S∗) →
In view of (2.10), the fundamental decomposition of the Krein space (H, [·,·])
H is surjective and ker Ψ = D(S).
coincides with
H = H+ ⊕ H−,
H+ =(cid:20)H0(cid:21) , H− =(cid:20) 0
H(cid:21) ,
where H+ = Ψ ker Γ− and H− = Ψ ker Γ+ are respectively, maximal uniformly
positive and negative subspaces with respect to the indefinite inner product [·,·].
By virtue of (2.1), each proper extension A of S is completely determined by
a subspace L = ΨD(A) = Ψ(D(S) +M) = ΨM of H. In other words, there is
a one-to-one correspondence between subspaces of H and proper extensions of
S. In particular, proper extensions Aλ in (2.2) are determined by the subspaces
Lλ = ΨD(Aλ), which are maximal uniformly positive (λ ∈ C+) and maximal
uniformly negative (λ ∈ C
Taking (2.6) and (2.9) into account, we arrive at the conclusion that the maxi-
mal uniformly positive subspace Lλ is decomposed with respect to the fundamen-
tal decomposition (2.11):
−) in (H, [·,·]), see [11].
Lλ = ΨD(Aλ) =(cid:26)(cid:20) Γ+f
ΘΓ+f(cid:21) : f ∈ D(Aλ)(cid:27) = {h+ + eΘ(λ)h+ : h+ ∈ H+},
λ ∈ C+.
(2.12)
where eΘ(·) : H+ → H− acts as follows:
eΘ(λ)h+ = eΘ(λ)(cid:20)h
0(cid:21) =(cid:20)
0
Θ(λ)h(cid:21) ,
This means that eΘ(λ) is the angular operator of the maximal uniformly positive
subspace Lλ with respect to the maximal uniformly positive subspace H+ of the
fundamental decomposition (2.11) (see [5] for the concept of angular operators).
Self-adjoint extensions A of S correspond to hypermaximal neutral subspaces
L = ΨD(A) of the Krein space (H, [·,·]). Each hypermaximal neutral subspace is
determined uniquely by a unitary mapping between subspaces H+ and H− of the
fundamental decomposition (2.11). This fact leads to the conclusion that each
self-adjoint extension A of S can be described as
A = S∗ ↾D(A),
D(A) = {f ∈ D(S∗) : TΓ+f = Γ−f},
where T is a unitary operator in H.
III. The explicit form of characteristic function depends on the choice of bound-
ary triplet. Let Θi(·) (i = 1, 2) be characteristic functions associated with bound-
ary triplets (Hi, Γi
+). Since the dimensions of the auxiliary Hilbert spaces Hi
−
, Γi
6
S. KUZHEL and L. NIZHNIK
coincide with the defect number of S, without loss of generality, we may assume
that H1 = H2 = H.
It is easy to see that the operator K : H → H defined by the formula
K(cid:20)Γ1
+f
Γ1
−
f(cid:21) =(cid:20)Γ2
+f
Γ2
−
f(cid:21) ,
f ∈ D(S∗).
is surjective in H and, moreover, K is a unitary operator in the Krein space
(H, [·,·]), i.e.
[Kx, Ky] = [x, y], x, y ∈ H (the latter relation follows from (2.8) -
(2.10)). Each unitary operator K in (H, [·,·]) determines the so-called interspher-
ical linear fractional transformation [5, Chapter III, section 3]
ΦK(Z) = (K21 + K22Z)(K11 + K12Z)−1,
K21 K22(cid:21) ,
K =(cid:20)K11 K12
where Kij are operator components of decomposition of K with respect to (2.11)
and a bounded linear operator Z maps H+ into H−. The interspherical trans-
formation ΦK(Z) is well defined for all Z : H+ → H− with kZk ≤ 1 (i.e.,
0 ∈ ρ(K11 + K12Z) ) and kΦK(Z)k ≤ 1.
It is known [14, 15] that
λ ∈ C+,
where eΘi(·) : H+ → H− are defined similarly to (2.12).
eΘ2(λ) = ΦK(eΘ1(λ)),
3. Phillips symmetric operator
(2.13)
We say that a symmetric operator with equal nonzero defect numbers is a
Phillips symmetric operator (PSO) if its characteristic function Θ(·) is an operator-
constant on C+.
By virtue of (2.13), this definition does not depend on the choice of boundary
triplet. However, in many cases, it is not easy to use it (because one should to
calculate the characteristic function). For this reason a series of statements which
can be used as (equivalent) definitions of PSO are presented below.
Theorem 3.1. A symmetric operator S with equal defect numbers is a Phillips
symmetric operator if and only if
Nλ ⊂ D(S) +Nµ,
for all λ, µ ∈ C+.
(3.1)
Proof. If S is a PSO, then its characteristic function Θ(·) associated with the
boundary triplet (Ni, Γ−, Γ+) determined by (2.7) has to be a constant. There-
fore, the Shtraus characteristic function Sh(λ) coincides with an operator U :
Ni → N−i for all λ ∈ C+. In particular, Sh(i) = U. By virtue of (2.4) with λ = i,
D(S) +Ni = D(S) +(I − U)Ni that is possible only for the case U = 0. Hence,
Sh(λ) ≡ 0 and (2.4) implies that
(3.2)
Let us assume that there exists fi ∈ Ni and µ ∈ C+ such that fi = v + fµ + fµ,
where v ∈ D(S) and fµ ∈ Nµ is non-zero. The last equality can be transformed
Nλ ⊂ D(S) +Ni,
∀λ ∈ C+.
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
7
to efi =ev + fµ with the use of (3.2). However, the obtained relation is impossible
because Im (S∗efi, efi) = kefik2 > 0 and, simultaneously,
Im (S∗efi, efi) = Im (S∗(ev + fµ),ev + fµ) = Im (S∗fµ, fµ) = −(Im µ)kfµk2 < 0.
Therefore, fi = v + fµ and Ni ⊂ D(S) +Nµ for all µ ∈ C. The obtained inequality
and (3.2) justify (3.1).
Conversely, if (3.1) holds, then, due to (2.4), D(S) +(I − Sh(λ))Ni ⊂ D(S) +Ni
that is possible only for Sh(λ) ≡ 0.
Remark 3.2. The inclusion (3.1) and its dual counterpart in C
(cid:3)
Nν ⊂ D(S) +Nξ,
for all
ν, ξ ∈ C
−.
(3.3)
are equivalent. Indeed, (3.1) means that the maximal dissipative operators Aλ
in (2.2) do not depend on the choice of λ ∈ C+, i.e., Aλ ≡ A+. Therefore, theirs
adjoint A∗λ = A∗µ = Aν = Aξ = A∗+ (ν = λ, ξ = µ) also do not depend on
ν, ξ ∈ C
Corollary 3.3. Simple Phillips symmetric operators with the same defect num-
bers are unitary equivalent.
−. This fact justifies (3.3).
−:
Proof. Let S be a PSO with defect numbers < m, m >. It follows from the proof
of Theorem 3.1 that the Shtraus characteristic function Sh(·) of S coincides with
the zero operator. Therefore, the characteristic function of S calculated in terms
of boundary triplet (Ni, Γ−, Γ+) (see (2.7)) is also zero operator acting in the
auxiliary space with the dimension m. Applying now theorem 2.1 for the case of
simple Phillips symmetric operators with the same defect numbers, we complete
the proof.
(cid:3)
−.
Theorem 3.4. A symmetric operator S with equal defect numbers is a Phillips
symmetric operator if and only if its defect subspaces Nλ and Nν are mutually
orthogonal for any λ ∈ C+ and ν ∈ C
Proof. Let S be a PSO and let λ, µ ∈ C+, λ 6= µ. By virtue of (3.1), fλ = u + fµ,
where fz ∈ Nz and u ∈ D(S). Therefore,
0 = (S∗ − λI)fλ = (S − λI)u + (µ − λ)fµ.
(3.4)
The obtained relation means that Nµ ⊂ R(S − λI) and, hence Nµ ⊥ Nλ. To
prove the orthogonality of Nµ and Nµ we use again (3.4) in order to rewrite
fλ = u + fµ as follows: fλ = (λ− µ)(S− λI)−1fµ + fµ. Let fµ be fixed and λ → µ.
Then fλ → fµ due to the last formula for fλ. Then,
(fµ, fµ) = lim
λ→µ
(fλ, fµ) = 0,
∀fµ ∈ Nµ, fµ ∈ Nµ.
Conversely, let Nλ ⊥ Nν for any λ ∈ C+, ν ∈ C
−. Assume now that the
relation (3.1) is not true for some µ ∈ C+. Then there exists fλ ∈ Nλ such that
fλ = u + fµ + fµ, where fµ 6= 0. Then
(λ − µ)fλ = (S∗ − µI)fλ = (S − µI)u + (µ − µ)fµ.
Due to our assumption Nλ ⊥ Nµ (ν = µ). Therefore, for any γµ ∈ Nµ,
0 = (λ − µ)(fλ, γµ) = (u, (S∗ − µI)γµ) + (µ − µ)(fµ, γµ) = (µ − µ)(fµ, γµ).
8
S. KUZHEL and L. NIZHNIK
That is possible only for fµ = 0. Therefore, the defect subspaces of S satisfy
(3.1) and S is a Phillips symmetric operator.
(cid:3)
Remark 3.5. Theorem 3.4 holds true if the weaker condition of orthogonality
Nλ ⊥ N−i will be used instead of Nλ ⊥ Nν. Indeed, since the inequalities (3.1)
and (3.2) are equivalent (see the proof of Theorem 3.1), it suffices to verify that
Nλ ⊥ N−i implies (3.2). The required implication is obtained by repeating the
end part of the proof of Theorem 3.4 with µ = i.
Corollary 3.6. A symmetric operator S in H with equal defect numbers <
m, m > is a Phillips symmetric operator if and only if the Hilbert space H can
be decomposed into the orthogonal sum H = H1 ⊕ H2 ⊕ H3 of Hilbert spaces Hj
leaving S invariant and such that
S = S1 ⊕ S2 ⊕ S3,
Sj = S ↾Hj ,
(3.5)
where S1 and S2 are simple maximal symmetric operators in H1 and H2 with defect
numbers < m, 0 > and < 0, m >, respectively and S3 is a self-adjoint operator in
H3.
Proof. If S has the decomposition (3.5), then its defect subspaces Nλ (λ ∈ C+)
coincide with the defect subspaces Nλ(S1) of the operator S1 (since Nλ(S2) = {0}
due to the defect numbers < 0, m > of S2) and therefore, Nλ ⊂ H1. Similarly,
the defect numbers < m, 0 > of S1 mean that Nν = Nν(S2) ⊂ H2 for all ν ∈ C
−.
Therefore, Nλ ⊥ Nν and S is a Phillips symmetric operator due to Theorem 3.4.
Conversely, each symmetric operator S with equal defect numbers is reduced
by the decomposition
H = Hα ⊕ H3,
R(S − µI),
(3.6)
H3 = \∀µ∈C−∪C+
where H3 is the maximal invariant subspace for S on which the operator S3 =
S ↾H3 is self-adjoint, while the subspace Hα coincides with the closed linear span
of all ker(S∗− µI) and the restriction Sα = S ↾Hα gives rise to a simple symmetric
operator in Hα with defect numbers < m, m > [10, p.9].
Assume now that S is a PSO, then its simple counterpart Sα is also PSO and
Nµ = ker(S∗ − µI) = Nµ(Sα) = ker(S∗α − µI) for all µ ∈ C
− ∪ C+. According to
Theorem 3.4, Nλ(Sα) ⊥ Nν(Sα) (λ ∈ C+, ν ∈ C
−). Therefore, we can decompose
Hα = H1 ⊕ H2, where H1 and H2 coincide with the closed linear spans of defect
subspaces {Nµ}µ∈C+ and defect subspaces {Nν}ν∈C−, respectively.
To complete the proof we should verify that Sα = S1 ⊕ S2, where Sj = Sα ↾Hj
(j = 1, 2) are maximal symmetric operators in Hj with defect numbers < m, 0 >
and < 0, m >, respectively. To that end we consider a simple symmetric operator
S = i
d
dx
,
D(S) = {u ∈ W 1
2 (R, N) : u(0) = 0}
(3.7)
acting in the Hilbert space L2(R, N), where N is an auxiliary Hilbert space with
the dimension m. It is easy to see that the defect subspaces Nµ, Nν (µ ∈ C+,
ν ∈ C
− ) are formed, respectively, by the functions
χR−(x)e−iµxn,
χR+(x)e−iνxn,
(3.8)
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
9
where n runs the Hilbert space N and χI(x) is the characteristic function of the
interval I. Therefore, the defect numbers of S is < m, m > and S is a Phillips
symmetric operator (since Nµ and Nν are mutually orthogonal).
By Corollary 3.3, the symmetric operator Sα in Hα is unitary equivalent to the
symmetric operator S acting in L2(R, N). For this reason, it sufficient to establish
the decomposition Sα = S1 ⊕ S2 for the case where Sα = S and Hα = L2(R, N).
Taking (3.8) into account, we decide that H1 = L2(R
−, N) and H2 = L2(R+, N).
Moreover, S = S1 ⊕ S2, where S1 = i d
−, N) : u(0) = 0}
is a maximal symmetric operator in L2(R
−, N) with defect numbers < m, 0 >
while, S2 = i d
2 (R+, N) : u(0) = 0} is maximal symmetric in
L2(R+, N) with defect numbers < 0, m >.
(cid:3)
dx , D(S1) = {u ∈ W 1
dx , D(S2) = {u ∈ W 1
2 (R
Remark 3.7. It follows from the proof of Corollary 3.6 that each simple PSO S
with defect numbers < m, m > is unitary equivalent to the symmetric operator
S defined by (3.7).
4. Proper extensions of a Phillips symmetric operator
It follows from (3.7) and Remark 3.7 that the spectrum of a simple PSO S is
continuous and it coincides with R. Furthermore, ker(S∗ − λI) = {0} for λ ∈ R.
Therefore, each proper extension A of a simple PSO has a continuous spectrum
on R without embedding eigenvalues. The lack of condition of being simple for
a PSO means that the spectra of the corresponding proper extensions coincides
with R but real point spectrum may appear due to a possible self-adjoint part S3
in (3.5).
Proposition 4.1. The spectrum σ(A) of a proper extension A of a Phillips sym-
metric operator S coincides with one of the following sets:
(i) σ(A) = R;
(ii) σ(A) = C
(iii) σ(A) = C.
− ∪ R or σ(A) = R ∪ C+;
Proof. Let us suppose that a proper extension A has a complex point λ0 ∈ ρ(A).
Without loss of generality we may assume that λ0 ∈ C
−. Then, the domain of A
admits the presentation
D(A) = {f = u + uλ0 + Φuλ0
: ∀u ∈ D(S), ∀uλ0 ∈ Nλ0},
where Φ : Nλ0 → Nλ0 is a bounded operator defined on Nλ0. The obtained
expression can be rewritten in terms of the boundary triplet (2.7) with µ = λ0:
where T = V Φ is a bounded operator in the auxiliary Hilbert space Nµ.
D(A) = {f ∈ D(S∗) : TΓ+f = Γ−f},
(4.1)
By virtue of [15, Theorem 4.2] (see also [14, Theorem 3]),
λ ∈ σ(A) ⇐⇒ 0 ∈ σ(Θ(λ) − T),
λ ∈ σ(A) ⇐⇒ 0 ∈ σ(I − Θ(λ)T),
λ ∈ C+,
−,
λ ∈ C
where Θ(·) is the characteristic function of S associated with the boundary triplet
(Nµ, Γ−, Γ+). Since S is PSO, the characteristic function Θ(·) is a constant on C+.
10
S. KUZHEL and L. NIZHNIK
Therefore, λ ∈ σ(A) ⇐⇒ 0 ∈ σ(Θ − T), where Θ(λ) = Θ for all λ ∈ C+. The
obtained relation means that either C+ belongs to σ(A) or C+ ⊂ ρ(A). Further,
due to the assumption above, there is a resolvent point λ0 ∈ C
− of A. Therefore
0 ∈ ρ(I − Θ∗T) and C
− ⊂ ρ(A). Summing up, the spectrum σ(A) corresponds
to the cases (i) or (ii) in dependence of either 0 ∈ ρ(Θ − T) or 0 ∈ σ(Θ − T).
The case λ0 ∈ C+ ∩ ρ(A) is considered in the same manner.
(cid:3)
Theorem 4.2. Proper extensions of a Phillips symmetric operator with real spec-
tra are similar to each other.
Proof. By virtue of Corollaries 3.3, 3.6, it is sufficient to consider proper exten-
sions of the simple PSO S determined by (3.7). In this case,
S∗f = i
df
dx
, D(S∗) = W 1
2 (R \ {0}, N)
and the defect subspaces Nµ, Nµ (µ ∈ C+) are formed, respectively, by the
functions χR−(x)e−iµxn and χR+(x)e−iµxn, where n runs the auxiliary Hilbert
space N.
Let us choose the unitary mapping V : Nµ → Nµ in the definition of boundary
triplet (2.7) as V χR+(x)e−iµxn = χR−(x)e−iµxn and consider the unitary mapping
W between Nµ and N as follows:
W χR−(x)e−iµxn =
n
.
p2(Im µ)
Then the modified boundary triplet (W Nµ, W Γ−, W Γ+) of the boundary triplet
(2.7) takes the form (N, Γ1
−
+), where
, Γ1
f = f (0+), Γ1
Γ1
−
+f = f (0−),
f ∈ D(S∗).
If a proper extension A of S has real spectrum, then its domain of definition
is determined by the formula (4.1), where T is a bounded operator in Nµ with
bounded inverse. This means that
A = S∗ ↾D(A), D(A) = {f ∈ D(S∗) : T f (0−) = f (0+)},
where T = W TW −1 is a bounded operator in N with bounded inverse.
(4.2)
Let F be a bounded operator with bounded inverse in N. Then, the operator
UF f =(cid:26) F f (x), x > 0
f (x), x < 0
,
f ∈ L2(R, N)
preserves these properties as an operator acting in L2(R, N) and U−1
F = UF −1.
Furthermore, UF : D(S∗) → D(S∗) and UF S∗ = S∗UF . These relations and (4.2)
lead to the conclusion that
UF AT f = UF S∗f = S∗UF f = AF T UF f,
where AT denotes the proper extension A in (4.2).
f ∈ D(AT ),
(4.3)
Let Aj be proper extensions of S with σ(Aj) = R. Then they are described in
(4.2) by bounded operators Tj with 0 ∈ ρ(Tj) (Aj ≡ ATj ). Due to (4.3),
AT1 = U−1
F AT2UF , with F = T2T −1
1
.
Therefore, Aj are similar to each other.
(4.4)
(cid:3)
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
11
Corollary 4.3. Self-adjoint extensions of a Phillips symmetric operator S are
unitary equivalent to each other. Precisely, there exists a collection of unitary
operators U = {Uξ}ξ∈I (I is the set of indices) with the properties
∀ξ ∈ I
Uξ ∈ T ⇐⇒ U∗ξ ∈ T,
UξS = SUξ,
and such that every pair of self-adjoint extensions A1, A2 of S satisfy the relation
UξA1 = A2Uξ
(4.5)
for some ξ ∈ I.
Proof. All self-adjoint extensions of the symmetric operator S are uniquely dis-
tinguished in (4.2) by the set of unitary operators T acting in N (see Section 2).
Therefore, the operators UT defined above are unitary operators in L2(R, N) and
(4.4) can be rewritten as (4.5) where Ai = ATi are self-adjoint extensions of S
and ξ = T2T −1
is unitary operator in L2(R, N).
1
It follows from the definition of UT that the set of U = {Uξ}ξ, where ξ runs
the set I of unitary operators in N satisfies the conditions of Corollary 4.3.
Therefore, the proof is complete for the simple PSO S defined by (3.7). This
result is extended to an arbitrary simple PSO S with the use of Corollary 3.3.
The required set U = {Uξ}ξ∈I for the general case of a Phillips symmetric
operator is obtained on the base of previously constructed (for simple operators)
set by the addition of the identity operator I3 acting in the subspace H3 (see
(3.5)) corresponding to the self-adjoint part of S.
(cid:3)
Remark 4.4. It follows from the construction of U = {Uξ}ξ∈I in Corollary 4.3
that the symmetric operator S is U-invariant. However, there are no U-invariant
self-adjoint extensions of S. Firstly, an example of such kind was constructed by
Phillips [21].
Corollary 4.5. Each self-adjoint extension of a simple PSO has Lebesgue spec-
trum on R.
Proof. In view of Corollary 3.3, a simple PSO is unitary equivalent to the sym-
metric operator S in (3.7). The momentum operator
A = i
d
dx
,
D(A) = W 1
2 (R, N)
(4.6)
is a self-adjoint extension of S in L2(R, N) and it has Lebesgue spectrum on R.
Applying now Theorem 4.2 we complete the proof.
(cid:3)
5. Phillips symmetric operators as the restriction of self-adjoint
ones
Lemma 5.1. Let A be a self-adjoint operator in a Hilbert space H. Each closed
densely defined symmetric operator that can be determined via certain restriction
of A is specified by the formula
SL = A ↾D(SL),
where L is a linear subspace of H such that L ∩ D(A) = {0}.
D(SL) = {u ∈ D(A) : ((A − iI)u, γ) = 0, ∀γ ∈ L},
(5.1)
12
S. KUZHEL and L. NIZHNIK
Proof. If S is a closed symmetric densely defined restriction of A, then S = SL,
where L = H ⊖ R(S − iI). Conversely, let SL be defined by (5.1). Obviously, SL
is a closed symmetric operator and, for any u ∈ D(SL) and p ∈ H,
(u, p) = ((SL − iI)u, (A + iI)−1p) = ((A − iI)u, (A + iI)−1p).
The obtained relation means that SL is nondensely defined ⇐⇒ (A + iI)−1p ∈
L ⇐⇒ L ∩ D(A) 6= {0}. Therefore, the condition L ∩ D(A) = {0} guarantees
that SL is densely defined.
(cid:3)
The symmetric operator SL in (5.1) turns out to be a PSO under certain choice
of L. To specify the required conditions, we consider the unitary operator
U = (A + iI)(A − iI)−1,
(5.2)
which is the Cayley transform of A and recall that a subspace L is called a
wandering subspace of U if U nL ⊥ L for any n ∈ N.
Theorem 5.2. The following statements are equivalent:
(A = i(U + I)(U − I)−1)
(i) the operator SL defined by (5.1) is a Phillips symmetric operator;
(ii) the subspace L is a wandering subspace of the unitary operator U.
Proof. (ii) ⇒ (i). Let L be wandering for U. First of all we should check that L∩
D(A) = {0}. Indeed, for all f ∈ L, (U nf, f ) = 0 and, hence R 2π
0 einλd(Eλf, f ) =
0, where Eλ is the spectral function of U. By the uniqueness theorem for the
Fourier-Stieltjes series, the last equality means that (Eλf, f ) = λ
2πkfk2.
It follows from (5.2) that
A = iZ 2π
eiλ + 1
eiλ − 1
with the domain D(A) = {f ∈ H : R 2π
f ∈ L,
0
0
cot(λ/2)dEλ
dEλ =Z 2π
0 cot2(λ/2)d(Eλf, f ) < ∞}. In the case of
2π Z 2π
Z 2π
cot2(λ/2)d(Eλf, f ) = kfk2
0
0
cot2(λ/2)dλ = ∞.
Therefore, L ∩ D(A) = {0} and the operator SL is densely defined.
order to describe other defect subspaces Nα of SL we consider the operator
It follows from (5.1) that the defect subspace N−i of SL coincides with L. In
Tα = (A + iI)(A − αI)−1,
− ∪ C+.
The formula Nα = TαL is verified directly with the use of (5.1).
the obtained expression for Tα can be rewritten as
Using (5.2) we get Tα = 2U[(1+iα)U +(1−iα)I]−1. In particular, if α = λ ∈ C+
α ∈ C
Tλ =
2U
1 − iλ
[I − tU]−1 =
2U
1 − iλ
∞Xn=0
tnU n,
t =
iλ + 1
iλ − 1
(5.3)
since ktUk = t < 1.
Since U nL ⊥ L for all n ∈ N, the relation (5.3) yields that TλL ⊥ L for λ ∈ C+.
Therefore, Nλ ⊥ N−i. Due to Remark 3.5 and Theorem 3.4, the operator SL is
PSO. The implication (ii) ⇒ (i) is proved.
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
13
(i) ⇒ (ii). If SL is a PSO, then the decomposition (3.5) and (5.1) imply that L
is a subspace of H1 ⊕ H2. Therefore, it is sufficient to assume that SL is a simple
symmetric operator.
Another important fact is that we can consider arbitrary self-adjoint extension
of SL in (5.1). Indeed, let A1 be a self-adjoint extension of SL such that A1 6= A.
Then, due to Corollary 4.3, there exists a unitary operator Uξ such that UξS =
SUξ and UξA = A1Uξ. Hence, the domain D(SL) can be described as
D(SL) = {v ∈ D(A1) : ((A1 − iI)v, g) = 0, ∀g ∈ L1 = UξL}.
Since the Cayley transformations U and U1 of the operators A and A1 are related
as UξU = U1Uξ, the existence of a wandering subspace L for U implies that
L1 = UξL will be a wandering subspace for U1. Therefore, it does not matter
which self-adjoint extension of SL we will consider in (5.1).
Simple Phillips symmetric operators with the same defect numbers are unitary
equivalent (Corollary 3.3). For this reason, we can consider a concrete Phillips
symmetric operator in (5.1). It is useful to work with PSO SL defined in H =
l2(Z, N) (N is an auxiliary Hilbert space) as follows:
SLu = i(. . . , x−3 + x−2, x−2 + x−1, x−1, x1, x1 + x2, . . .),
xj ∈ N,
(5.4)
where element at the zero position is underlined and
u ∈ D(S) ⇐⇒ u = (. . . , x−3 − x−2, x−2 − x−1, x−1,−x1, x1 − x2, . . .),
where Pi∈Z kxik2
N < ∞.
The operator SL defined by (5.4) is a simple PSO in l2(Z, N) and the operator
(5.5)
Au = i(. . . , x−3 + x−2, x−2 + x−1, x−1 + x0, x0 + x1, x1 + x2, . . .)
with the domain of definition u ∈ D(A) ⇐⇒
u = (. . . , x−3 − x−2, x−2 − x−1, x−1 − x0, x0 − x1, x1 − x2, . . .), Xi∈Z kxik2
N < ∞
is the self-adjoint extension of SL [15, 16].
It follows from (5.4) and (5.5) that D(SL) consists of those u ∈ D(A) for
which x0 = 0. Direct calculation with use of (5.5) shows that (A − iI)u =
2i(. . . , x−2, x−1, x0, x1, x2, . . .). Therefore, the formula (5.1) gives D(SL) if
L = {(. . . , 0, 0, x0, 0, 0, . . .) ∈ l2(Z, N) : ∀x0 ∈ N}.
It is easy to see that the Cayley transform U of A coincides with the bilateral
shift
U(. . . , x−2, x−1, x0, x1, x2, . . .) = (. . . , x−3, x−2, x−1, x0, x1, . . .)
in l2(Z, N). The subspace L is a wandering subspace for U. The proof is complete.
(cid:3)
Corollary 5.3. The set of symmetric restrictions of a given self-adjoint operator
A contains a Phillips symmetric operator if and only if there exists a reducing
subspace H0 of A such that the self-adjoint operator A0 = A ↾D(A)∩H0 has Lebesgue
spectrum on R.
14
S. KUZHEL and L. NIZHNIK
The existence of a simple PSO among symmetric restrictions of A is equivalent
to the fact that A has Lebesgue spectrum on R.
Proof. Due to Theorem 5.2, a PSO can appear only in the case where there
U0 = U ↾H0 is a bilateral shift in H0 and its Cayley transform A0 has Lebesgue
spectrum on R.
exists a wandering subspace L of U. Denote H0 = Pn∈Z ⊕U nL. The operator
According to Corollary 3.6, the existence of a simple PSO means that H3 = {0}
in the decomposition (3.5). Therefore, the corresponding wandering subspace L
(which determines a simple PSO with the help of (5.1)) has the property that
H =Pn∈Z ⊕U nL. This means that A has Lebesgue spectrum on R.
(cid:3)
Corollary 5.4. Let Wt = eiAt be the group of unitary operators generated by a
self-adjoint operator A. If there exists a nonzero h ∈ H such that
(Wth, h) = 0,
for all
t > c,
then the set of symmetric restrictions of A contains a PSO.
Proof. Denote by H0 the closure of linear span of {Wth} in H. The subspace H0
reduces Wt and, by virtue of [24, Lemma 1.2], the restriction Wt ↾H0 is a bilateral
shift in H0. Its generator A0 is a self-adjoint operator in H0 with the Lebesque
spectrum. Applying now Corollary 5.3 we complete the proof.
(cid:3)
The concept of Lebesgue spectrum on R for a self-adjoint operator A can be
defined in various (equivalent) ways which guarantee that the spectral type of
A is equivalent to the Lebesgue one and the multiplicity of the spectrum σ(A)
does not change for any real point. The last condition is obviously satisfied when
A is unitary equivalent to its shifts A − tI for any t ∈ R. Development of this
'translation-invariance' idea leads to the prominent Weyl commutation relation
which ensures the Lebesgue spectrum property of A. Namely, due to the von
Neumann theorem [17, p. 35], a self-adjoint operator A in H has the Lebesgue
spectrum on R if and only if there exists a strongly continuous group of unitary
operators Vt such that
VtAV−t = A − tI,
∀t ∈ R.
(5.6)
It should be mention that any simple PSO is also a solution of the Weyl com-
mutation relation (5.6). Indeed, any operator A which is the solution of (5.6) is
determined up to unitary equivalence. Therefore, it is sufficient to consider the
simple PSO S defined by (3.7) in L2(R, N) and to verify that A = S is a solution
of (5.6) with Vt acting as the multiplication on e−itx.
6. Examples of PSO
I. Let SL be a PSO that is determined by (5.1) as the restriction of a self-adjoint
operator A. Consider a unitary operator W that commutes with A. It is easy to
see that S′ = W SLW −1 is also PSO and S′ is determined by (5.1) with the new
wandering subspace L′ = WL, i.e., S′ = SWL. This simple observation gives rise
to infinitely many PSO which are symmetric restrictions of a given self-adjoint
operator A.
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
15
If A has Lebesgue spectrum on R and its multiplicity coincides with dimL,
then the obtained PSO is simple. Furthermore, the space H is presented as H =
Pn∈Z ⊕U nL. The last decomposition allows one to determine a unitary mapping
of H onto L2(R,L) in such a way that A corresponds to the multiplication by
independent variable: Af (δ) = δf (δ); the Cayley transform of A acts as the
multiplication operator: Uf (δ) = δ+iI
δ−iI f (δ); and the wandering subspace L (in H)
is mapped onto the wandering subspace 1
If W is a unitary operator in L2(R,L) that commutes with A, then W can be
realized as a multiplicative operator-valued function w(δ) on L into L which is
unitary for almost all δ (see, e.g., [17, Corollary 4.2, p.53]) :
δ+iL in L2(R,L) [17, Chapter 2].
W f = w(δ)f (δ),
f ∈ L2(R,L).
This means that the subspaces Lw ≡ W 1
δ+i L are wandering in L2(R,L)
and they determines infinitely many simple PSO Sw, which, due to (5.1), are
restrictions of the operator A of multiplication by δ onto linear manifolds
δ+iL = w(δ)
D(Sw) = {u ∈ D(A) :ZR
(u(δ), w(δ)v)Ldδ = 0, ∀v ∈ L}.
This result can be reformulated for the restrictions of self-adjoint momentum
operator A (see (4.6) where L = N) with the use of Fourier transformation
(F f )(x) = 1√2πRR e−iδxf (δ)dδ that relates Af (δ) = δf (δ) and Af = i df
dx . Taking
into account that AF = F A we decide that the subspaces FLw are wandering
for the Cayley transform of A in L2(R,L). Simple PSO Sw are the restrictions
of A onto those functions f ∈ W 1
2 (R,L) that satisfy the relation (c.f. (5.1)):
((A − iI)u, γ) = 0, ∀γ ∈ FLw. It is clear that Sw = F SwF −1.
Let us set w(δ) ≡ 1. Then the wandering subspace FLw coincides with the
subspace χR+(x)e−xL and the formula (5.1) leads, to the simple PSO S defined
by (3.7). The operator S is the result of one-point perturbation of the momentum
operator A supported at point x = 0. The symmetric operator
Sw = i
d
dx
,
D(Sw) = {u ∈ W 1
2 (R,L) : u(y) = 0}
(6.1)
corresponding to one-point interaction supported at real point x = y is deduced
from the formulas above with w(δ) = eiδy.
Assume now that w(δ) = δ+iI
formula (5.1) gives rise to the simple PSO
δ−iI . Then FLw = F (cid:0) 1
δ−iIL(cid:1) = χR−(x)exL and the
Sw = i
d
dx
,
D(Sw) = {u ∈ W 1
2 (R,L) : u(0) = 2Z 0
−∞
u(x)exdx},
(6.2)
which is an example of nonlocal point interaction of the momentum operator A.
II. Let Df = √2f (2x) and T f = f (x − 1) be the dilation and the translation
operators in L2(R). Denote A = i(D +I)(D−I)−1. The operator A is self-adjoint
in L2(R) and it has Lebesgue spectrum on R (since D is a bilateral shift).
Proposition 6.1. Let S be a simple PSO that is a restriction of the self-adjoint
operator A and let ψ ∈ N−i = ker(S∗ + iI) be a function such that {T kψ}k∈Z is
an orthonormal basis of N−i. Then ψ is a wavelet.
16
S. KUZHEL and L. NIZHNIK
Proof. If S is a restriction of A, then S is determined by (5.1), i.e., S = SL,
where L = N−i. Since S is assumed to be PSO, Theorem 5.2 implies that L is
a wandering subspace for the dilation operator D. Moreover, the simplicity of S
means that L2(R) = Pn∈Z ⊕DjL. Therefore, {DjT kψ}j,k∈Z is an orthonormal
basis of L2(R), i.e., ψ is a wavelet [7].
(cid:3)
7. Nonlocal point interactions
The above results show that one-point interaction of the momentum opera-
tor: A + αδ(x − y) leads to self-adjoint operators which are unitary equivalent
to each other and have Lebesgue spectrum on R. This means that non-trivial
spectral properties of self-adjoint operators associated with the momentum op-
erator should be obtained with the help of more complicated perturbations. In
the present section we consider special classes of general nonlocal one point in-
teractions [3] which can be characterized as one point interaction defined by the
nonlocal potential γ(x) ∈ L2(R).
{0}) by the differential expression
+ γ(x)fr
I. Let us consider the maximal operator Smax which is determined on W 1
Smaxf = i
2 (R \
df
dx
(x 6= 0),
fr =
(f (0+) + f (0−)),
1
2
where the non-local potential γ(x) belongs to L2(R). Direct calculation shows
that, for all f, g ∈ D(Smax) = W 1
2 (R \ {0}),
(Smaxf, g) − (f, Smaxg) = i[Γ+f Γ+g − Γ−f Γ−g],
where Γ± are determined by
Γ+f = f (0−) +
i
2
(f, γ),
Lemma 7.1. The operator
Γ−f = f (0+) −
i
2
(f, γ).
(7.1)
Smin = Smax ↾D(Smin),
D(Smin) = ker Γ− ∩ ker Γ+
is a closed densely defined symmetric operator in L2(R) and such that S∗min =
Smax. A triplet (C, Γ−, Γ+) , where the linear mappings Γ± : W 1
2 (R \ {0}) → C
are determined by (7.1) is a boundary triplet of Smax.
Proof. To complete the proof, by virtue of [6, Corollary 2.5] and Remark 2.2, it
is sufficient to verify that: (i) there is a unimodular c such that the operator A =
Smax ↾ker(cΓ+−Γ−) is self-adjoint in L2(R); (ii) the map (Γ−, Γ+) : D(Smax) → C2
is surjective.
dx with the
domain D(A) = {f∈W 1
2 (R \ {0}) : f (0−) = −f (0+)} is a self-adjoint operator.
2 (R \ {0})
such that f (0−) = h1 and f (0+) = h2. Let us fix u ∈ W 1
2 (R \ {0}) such that
u(0−) = u(0+) = 0 and (u, γ) 6= 0. Using now (7.1) we decide that the vector
ef = f − (f,γ)
(cid:3)
The condition (i) is satisfied if we choose c = −1. In this case A = i d
Let h = (h1, h2) be an arbitrary element of C2. There exists f ∈ W 1
(u,γ) u solves the equation (Γ−, Γ+)ef = (h1, h2), that justifies (ii).
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
17
The boundary triplet (C, Γ−, Γ+) constructed in Lemma 7.1 allows us to de-
termine self-adjoint operators
df
dx
Aθf = i
+ γ(x)fr,
2 (R \ {0}),
whose domains D(Aθ) consist of all functions f ∈ W 1
nonlocal boundary-value condition
f ∈ D(Aθ) ⊂ W 1
θ ∈ [0, 2π)
(7.2)
2 (R \ {0}) that satisfy the
eiθ[f (0−) +
i
2
(f, γ)] = f (0+) −
i
2
(f, γ).
These operators are mathematical models of one point interaction defined by the
nonlocal potential γ(x).
Each operator Aθ is a self-adjoint extension of the symmetric operator Smin =
S∗max = i d
dx with domain of definition
D(Smin) =(cid:26)f ∈ W 1
2 (R \ {0}) :
f (0−) + i
f (0+) − i
2(f, γ) = 0
2(f, γ) = 0 (cid:27) .
(7.3)
The symmetric operator Smin has defect numbers < 1, 1 > and its defect sub-
spaces Nλ, Nν (λ ∈ C+, ν ∈ C
fλ(x) = gλ(x) − 2[1 + gλ(0)]G+
respectively. Here
−) coincide with the linear span of the vectors
λ (x)
and fν(x) = gν(x) − 2[1 + gν(0)]G−ν (x),
gz = (A − zI)−1γ =(cid:26) ie−izxR ∞
−ie−izxR x
λ (x) =(cid:26) 0,
x > 0
e−iλx x < 0,
eizτ γ(τ )dτ,
x eizτ γ(τ )dτ,
z ∈ C+
z ∈ C
G−ν (x) =(cid:26) e−iνx, x > 0
−∞
x < 0,
−
0
G+
(7.4)
and
and
By virtue of (6.2) and (7.1), the characteristic function Θ(·) has the form
Θ(λ) =
Γ−fλ
Γ+fλ
=
fλ(0+) − i
fλ(0−) + i
2 (fλ, γ)
2 (fλ, γ)
= −I +
2
2 + gλ(0) − i
2(fλ, γ)
.
(7.5)
Let us consider a particular case assuming that γ = αχR+(x)e−x, α ∈ C. Then
gλ(x) =
iα
1 − iλ(cid:26) e−x,
e−iλx,
gλ(0) −
i
2
(fλ, γ) =
iα
1 − iλ
x > 0
x < 0
(1 − iα/4)
Therefore, the characteristic function (7.5) turns out to be a constant on C+
when β = 4i. In this case, the symmetric operator Smin in (7.3) is PSO and all
its self-adjoint extensions (7.2) have Lebesgue spectrum on R.
II. Let the maximal operator Smax be determined by the differential expression
Smaxf = i
df
dx
+ γ(x)fs
(x 6= 0),
fs = f (0+) − f (0−),
18
S. KUZHEL and L. NIZHNIK
where the non-local potential γ(x) belongs to L2(R). Similarly to the previous
case, the Green formula can be established
(Smaxf, g) − (f, Smaxg) = i[Γ+f Γ+g − Γ−f Γ−g],
2 (R \ {0})
where Γ+f = f (0−)− i(f, γ) and Γ−f = f (0+)− i(f, γ). The same arguments as
in the proof of Lemma 7.1 leads to the conclusion that (C, Γ−, Γ+) is a boundary
triplet of Smax and the corresponding symmetric operator Smin = Smax ↾D(Smin),
D(Smin) = ker Γ− ∩ ker Γ+ has the form
f, g ∈ D(Smax) = W 1
Smin = i
2 (R) : f (0) = i(f, γ)}.
Each self-adjoint extension Aα of Smin is determined by the formula
, D(Smin) = {f ∈ W 1
(7.6)
d
dx
Aαf = i
df
dx
+ γ(x)fs,
where D(Aθ) = {f ∈ D(Smax) : eiθ[f (0−) − i(f, γ)] = f (0+) − i(f, γ)}.
span of vectors
The defect subspaces Nλ, Nν (λ ∈ C+, ν ∈ C
−) of Smin coincide with the linear
fλ(x) = gλ(x) + G+
λ (x)
and fν(x) = gν(x) − G−ν (x),
respectively. Let us fix γ = αχR−(x)ex and specify for which α ∈ C the corre-
sponding symmetric operator Smin will be PSO. It follows from (7.4) that
iαχR−(x)
gλ(x) =
gν(x) = −
The obtained expressions allows one to calculate
(e−iλx − ex),
1 + iλ
iα
1 + iν (cid:26) e−iνx,
ex,
x > 0
x < 0.
(fλ, fν) =
α
2(1 − iλ)(1 − iν)
(2i − α),
λ ∈ C+,
ν ∈ C
−.
The obtained expression and Theorem 3.4 mean that the symmetric operator
Smin defined in (7.6) is PSO if and only if α = 2i. In this case, the PSO Smin
coincides with the operator Sw determined by (6.2).
References
1. S. Albeverio, F. Gesztesy, R. Høegh-Krohn, and H. Holden, Solvable Models in Quan-
tum Mechanics, Springer-Verlag, Berlin/New York, 1988; 2nd ed. (with an appendix by
P. Exner), AMS Chelsea Publishing, Providence, RI, 2005.
2. S. Albeverio and S. Kuzhel, PT -symmetric operators in quantum mechanics: Krein spaces
methods, Non-selfadjoint operators in quantum physics, 293 -- 343, Wiley, Hoboken, NJ, 2015.
3. S. Albeverio and L.P. Nizhnik, Schrodinger operators with nonlocal potentials Meth. Funct.
Anal. Topology, 19 (2013), 199 -- 210.
4. Yu. M. Arlinskii, V. A. Derkach and E. R. Tsekanovskii, On unitary equivalent quasi-
Hermitian extensions of Hermitian operators Mat. Fiz., 29 (1981), 72 -- 77 ( in Russian).
5. T. Ya. Azizov and I.S. Iokhvidov, Linear Operators in Spaces with an Indefinite Metric,
John Wiley & Sons, Chichester, 1989.
6. J. Behrndt and M. Langer, On the adjoint of a symmetric operator, J. Lond. Math. Soc.
82 (2010), 563-580.
7. O. Christensen, Functions, Spaces, and Expansions, Birkhauser, Basel, 2010.
PHILLIPS SYMMETRIC OPERATORS AND THEIR EXTENSIONS
19
8. P. Exner, Momentum operators on graphs, Spectral Analysis, Differential Equations and
Mathematical Physics: A Festschrift in Honor of Fritz Gesztesys 60th Birthday 87 (2012),
105 -- 118.
9. M. L. Gorbachuk and V. I. Gorbachuk, Boundary-Value Problems for Operator-Differential
Equations, Kluwer, Dordrecht, 1991.
10. M. L. Gorbachuk, V. I. Gorbachuk, M.G. Krein's Lectures on Entire Operators, Operator
Theory Advances and Applications 97, Birkhauser, Basel, 1997.
11. S. Hassi and S. Kuzhel, On J -self-adjoint operators with stable C-symmetries, Proc. Royal
12. P. Jorgensen, S. Pedersen, and F. Tian, Momentum operators in two intervals: Spectra and
Soc. Edinburgh 143A (2013), 141-167.
phase transition, Complex Anal. Operator Theory 7 (2013), 1735 -- 1773.
13. A. N. Kochubei, About symmetric operators commuting with a family of unitary operators,
Funk. Anal. Prilozh. 13(1979) 77 -- 78.
14. A. N. Kochubei, On extensions and characteristic functions of symmetric operators, Izv.
Akad. Nauk. Arm. SSR 15 (1980), 219 -- 232. (In Russian); English translation: Soviet J.
Contemporary Math. Anal. 15 (1980).
15. A. Kuzhel and S. Kuzhel, Regular Extensions of Hermitian Operators, VSP, Utrecht, 1998.
16. S. Kuzhel, O. Shapovalova, and L. Vavrykovych, On J -self-adjoint extensions of the Phillips
symmetric operator, Meth. Funct. Anal. Topology, 16 (2010), 333 -- 348.
17. P. D. Lax and R. F.g Phillips, Scattering Theory. Revised Edition, Academic Press, Inc.
San Diego, 1989.
18. L.P. Nizhnik, Inverse spectral nonlocal problem for the first order ordinary differential equa-
tion, Tamkang J. Math. 42 (2011) 385 -- 394.
19. C. R. de Oliveira, Intermediate Spectral Theory and Quantum Dynamics, Birkhuser, Basel,
2009.
20. S. Pedersen, J. D. Phillips, F. Tian, C. E. Watson, On the spectra of momentum operators,
Complex Anal. Oper. Theory 9 (2015), 1557 -- 1587.
21. R. S. Phillips, The extension of dual subspaces invariant under an algebra, in: Proceedings
of the International Symposium on Linear Spaces (Jerusalem, 1960), pp. 366-398, Jerusalem
Academic Press, 1961.
22. K. Schmudgen, Unbounded Self-adjoint Operators on Hilbert space, Springer, Berlin, 2012.
23. A. V. Shtraus, On extensions and characteristic functions of symmetric operators, Izv.
Akad. Nauk SSSR. Ser. Mat. 32 (1968), 186 -- 207. (Russian)
24. Ja. G. Sinai, Dynamical systems with countable Lebesgue spectrum. I., Izv. Akad. Nauk
SSSR Ser. Mat. 25 (1961) 899-924. (Russian)
1AGH University of Science and Technology, Krak´ow 30-059, Poland.
E-mail address: [email protected]
2Institute of Mathematics National Academy of Sciences of Ukraine, Kyiv
01-601, Ukraine
E-mail address: [email protected]
|
1503.06691 | 1 | 1503 | 2015-03-23T15:47:50 | Quasianalytic polynomially bounded operators | [
"math.FA"
] | Quasianalytic contractions form the crucial class in the quest for proper invariant and hyperinvariant subspaces for asymptotically non-vanishing Hilbert space contractions. The property of quasianalycity relies on the concepts of unitary asymptote and $H^\i$-functional calculus. These objects can be naturally defined in the setting of polynomially bounded operators too, which makes possible to extend the study of quasianalycity from contractions to this larger class. Carrying out this program we pose also several interesting questions. | math.FA | math |
Quasianalytic polynomially bounded operators
L´aszl´o K´erchy
University of Szeged
Abstract
Quasianalytic contractions form the crucial class in the quest for proper
invariant and hyperinvariant subspaces for asymptotically non-vanishing
Hilbert space contractions. The property of quasianalycity relies on the
concepts of unitary asymptote and H ∞-functional calculus. These objects
can be naturally defined in the setting of polynomially bounded operators
too, which makes possible to extend the study of quasianalycity from
contractions to this larger class. Carrying out this program we pose also
several interesting questions.
AMS Subject Classification (2010): 47A15, 47A45, 47A60.
Key words: unitary asymptote, polynomially bounded operator, Lebesgue
decomposition, intertwining relations, quasianalytic operator.
1 Introduction
Let H be a complex Hilbert space, and let L(H) stand for the C∗-algebra of
all bounded linear operators acting on H. Given any T ∈ L(H), the complete
lattice Lat T consists of those (closed) subspaces M of H, which are invariant
for T : T M ⊂ M. A subspace M is hyperinvariant for T , if it is invariant for
every operator C ∈ L(H), commuting with T : CT = T C. The complete lattice
of all hyperinvariant subspaces of T is denoted by Hlat T . The most challenging
open questions in operator theory are arguably the Invariant Subspace Problem
(ISP) and the Hyperinvariant Subspace Problem (HSP). The first question asks
whether Lat T is non-trivial (i.e., different from {{0}, H}), for every operator
T ∈ L(H), provided dim H ≥ 2; while the second question asks whether Hlat T
is non-trivial, whenever T is not a scalar multiple of the identity operator I.
For a thorough discussion of these problems see the monographs [RR] and [CP],
and the more recent papers [AM] and [GR].
Of course, dividing T by its norm, we may assume that T is a contraction:
kT k ≤ 1. An opposite, local estimate is made, when we assume that the contrac-
tion T is asymptotically non-vanishing, that is limn kT nhk > 0 holds for some
h ∈ H. Suprisingly enough, (ISP) and (HSP) are open under these assumptions
too. A useful extra tool available in this situation is the unitary asymptote
of T . Relying on this and on the H ∞-functional calculus, two spectral invari-
ants can be introduced for T . The first one is the residual set ω(T ), and the
second one is the quasianalytic spectral set π(T ) of T . The contraction T is
1
called quasianalytic, if these measurable subsets of the unit circle T coincide:
ω(T ) = π(T ). Quasianalytic contractions were investigated in the papers [K5],
[K8], [K9], [KT], [KSz2] and [KSz3]. A central theorem proved is that (HSP)
can be reduced to this class in the asymptotically non-vanishing case.
The aim of this paper is to extend these investigations from contractions to
polynomially bounded operators. It will turn out that this larger class is the
natural setting for the study of quasianalycity.
Our work is organized in the following way.
In Section 2 we introduce unitary asymptotes for an arbitrary operator T ∈
L(H) in a categorical sense, as it has been done for contractions in [BK] and
in Chapter IX of [NFBK]. The induced generalized Toeplitz operators, and the
connection of the associated symbolic calculus with the commutant mapping
are discussed. Existence of unitary asymptotes in the power bounded case is
shown by using Banach limits. This method originates in Sz.-Nagy's pioneering
work [N], and has been extended to many situations; see, e.g., [K1], [K2], [K3],
[K4], [K6], [KL], [KM] and [Pr]. A new characterization is given in terms of
norm-conditions.
It is well-known that every contraction T can be decomposed into the or-
thogonal sum T = T a ⊕ U s of an absolutely continuous contraction T a and a
singular unitary operator U s. Mlak showed that analogous Lebesgue-type de-
compositions can be given for polynomially bounded operators, or equivalently,
for bounded representations of the disk algebra A(T). Actually, Mlak considered
representations of more general function algebras; see [M1], [M2], [M3] and [M4].
Using Mlak's elementary measures, in Section 3 we give a detailed, streamlined
discussion of the Lebesgue decomposition in the particular case A(T), what is of
the main interest for us. Our purpose here (and partly in the next two sections)
is to make Mlak's important results more accessible for a wide range of readers.
In Section 4 we focus on intertwining relations. It is verified that absolute
continuity and singularity are preserved under quasisimilarity. A transparent
proof is given for the important known fact that every singular polynomially
bounded operator is similar to its unitary asymptote. Furthermore, the problem
of similarity to contractions is discussed.
Sz.-Nagy and Foias introduced the effective functional calculus, working with
functions in H ∞, for absolutely continuous contractions; see Chapter III in
[NFBK]. Section 5 is devoted to the study of properties and possible range
of this H ∞-functional calculus. A simplified proof is given for Mlak's result,
stating that exactly the absolutely continuous polynomially bounded operators
admit H ∞-functional calculae.
In Section 6 we turn to quasianalytic operators. The sets ω(T ) and π(T )
are introduced in a uniform way, relying on the local residual sets ω(T, x) (x ∈
H). The quasianalytic spectral set π(T ) is characterized also in terms of non-
monotone sequences. The central hyperinvariant subspace theorem is proved.
Furthermore, the effects of intertwining relations, the asymptotic behaviour,
orthogonal sums and restrictions are studied.
For the theory of Hardy spaces H p, see [Ho]. In connection with the theory of
contractions we refer to [NFBK]. Given operators T1 ∈ L(H1) and T2 ∈ L(H2),
2
the set of intertwining transformations is defined by
I(T1, T2) := {C ∈ L(H1, H2) : CT1 = T2C} .
The commutant of the operator T ∈ L(H) is {T }′ = I(T, T ), while the bicom-
mutant (or double commutant) {T }′′ of T consists of those operators B ∈ L(H)
which commute with every operator C in {T }′. Finally, N, Z, Z+, R, R+, C de-
note the set of positive integers, integers, non-negative integers, real numbers,
non-negative real numbers and complex numbers, respectively.
2 Unitary asymptotes
Let H be a complex Hilbert space, and let us consider an operator T ∈ L(H).
We say that (X, U ) is a unitary intertwining pair for T , if U is a unitary op-
erator acting on a (complex) Hilbert space K, and X ∈ I(T, U ). The pair is
n=1U −nXH = K. The unitary intertwining pair (X, U ) is called
minimal, if ∨∞
a unitary asymptote of T , if for any other unitary intertwining pair (X ′, U ′) of
T , there exists a unique transformation Z ∈ I(U, U ′) such that X ′ = ZX. The
uniqueness of Z implies that a unitary asymptote (X, U ) is necessarily minimal.
Let us assume that (X1, U1) and (X2, U2) are unitary asymptotes of T .
Then there exist Z1 ∈ I(U1, U2) and Z2 ∈ I(U2, U1) such that Z1X1 = X2 and
Z2X2 = X1. Therefore, I ·X1 = Z2X2 = (Z2Z1)X1, whence I = Z2Z1 follows by
the definition of the unitary asymptote. The equation Z1Z2 = I can be derived
similarly. Thus, the unitary intertwining pairs (X1, U1) and (X2, U2) are equiv-
alent, which means the existence of an invertible transformation Z ∈ I(U1, U2)
satisfying X2 = ZX1. Then the unitary operators U1 and U2 are necessarily
unitarily equivalent: U1 ≃ U2. Furthermore, if the unitary intertwining pair
(X2, U2) is equivalent to a unitary asymptote (X1, U1), then (X2, U2) is also a
unitary asymptote.
For any unitary intertwining pair (X, U ) of T we have kXhk = kU nXhk =
kXT nhk ≤ kXk · kT nhk (n ∈ N), and so
kXhk ≤ kXk · lim inf
n→∞
kT nhk
for all h ∈ H.
A lower estimate for kXhk yields universality.
Proposition 1 Let (X, U ) be a minimal unitary intertwining pair for T . If
there exists κ ∈ (0, ∞) such that
κ · lim inf
n→∞
kT nhk ≤ kXhk
holds for all h ∈ H, then (X, U ) is a unitary asymptote of T .
Proof. Let (X ′, U ′) be a unitary intertwining pair for T . For every h ∈ H we
have
kX ′hk ≤ kX ′k · lim inf
n→∞
kT nhk ≤ kX ′k · κ−1 · kXhk.
3
Hence there is a unique transformation Y+ ∈ L(K+, K′) such that Y+Xh =
X ′h (h ∈ H), where K+ = (XH)−. The relations
Y+U Xh = Y+XT h = X ′T h = U ′X ′h = U ′Y+Xh (h ∈ H)
show that Y+ ∈ I(U+, U ′), where U+ = U K+. It is immediate that the equa-
tions
Y U −nk := U ′−nY+k (k ∈ K+, n ∈ N)
define a norm-preserving extension Y ∈ I(U, U ′) of Y+. It is clear that Y X =
X ′, and that Y is uniquely determined by these properties.
(cid:3)
Question 2 Does κ lim inf n→∞ kT nhk ≤ kXhk (h ∈ H) hold with a κ ∈ (0, ∞),
whenever (X, U ) is a unitary asymptote of T ?
Let (X, U ) be any unitary intertwining pair for T . Then T ∗(X ∗X)T =
X ∗U ∗U X = X ∗X. Thus X ∗X belongs to the set
T (T ) := {B ∈ L(H) : T ∗BT = B}
of T -Toeplitz operators. It is clear that T (T ) is a selfadjoint linear manifold,
which is closed in the weak operator topology. Taking Ts(T ) = {B ∈ T (T ) :
B∗ = B}, we have T (T ) = Ts(T ) + iTs(T ). The set T+(T ) = {B ∈ Ts(T ) :
B ≥ 0} is of particular interest. We recall that A ∈ T+(T ) is universal (weakly
universal ) in T (T ), if for every B ∈ Ts(T ) (B ∈ T+(T ), respectively) there
exists β ∈ R+ such that
−βA ≤ B ≤ βA.
It is obvious that T (U ) = {U }′. Furthermore, for any F ∈ {U }′ we have
T ∗(X ∗F X)T = X ∗U ∗F U X = X ∗U ∗U F X = X ∗F X, that is X ∗F X ∈ T (T ).
The positive, bounded, ∗-linear mapping
ΨX : {U }′ → T (T ), F 7→ X ∗F X
is the symbolic calculus for T (T ), associated with (X, U ). It is easy to see that
ΨX is injective, when (X, U ) is minimal. We say that ΨX is universal, if its
range coincides with T (T ). The following theorem, proved in [K6] in a slightly
different form, establishes connections among these concepts.
Theorem 3 Let (X, U ) be a minimal unitary intertwining pair for T . Then
(i) ⇐⇒ (ii) ⇐ (iii) ⇐⇒ (iv), where
(i) (X, U ) is a unitary asymptote of T ;
(ii) A = X ∗X ∈ T+(T ) is weakly universal in T (T );
(iii) A = X ∗X ∈ T+(T ) is universal in T (T );
(iv) the symbolic calculus ΨX is universal.
4
We note that if A = X ∗X is universal in T (T ), then ΨX : {U }′ → T (T ) is
invertible, and so
−β0kBkA ≤ B ≤ β0kBkA
holds for every B ∈ Ts(T ), where β0 = kΨ−1
X k.
It may happen that there is not a universal operator in T (T ); we refer to
Example 5 in [K6]. Moreover, if T (T ) contains a non-zero operator B, then the
relations 0 < kBk = kT ∗nBT nk ≤ kBkkT nk2 (n ∈ N) show that the spectral
radius r(T ) ≥ 1.
Let us assume that (X, U ) is a unitary asymptote of T . Then, for every
C ∈ {T }′, XC ∈ I(T, U ) holds, and so there exists a unique D ∈ {U }′ such
that XC = DX. It is easy to verify that the commutant mapping
γX : {T }′ → {U }′, C 7→ D
is a (unital) algebra-homomorphism. Boundedness of γX follows from its rela-
tion to ΨX .
Proposition 4 Let (X, U ) be a unitary asymptote of T , and let us assume
that A = X ∗X is universal in T (T ). Then (ΨX ◦ γX )C = AC holds for every
C ∈ {T }′, and so kγX k ≤ kAk · kΨ−1
X k.
The simple proof is left to the reader.
Existence of unitary asymptotes can be verified in the class of power bounded
operators. We recall that T ∈ L(H) is a power bounded operator, if
MT := sup{kT nk : n ∈ Z+} < ∞.
In that case r(T ) ≤ 1 is evidently true. If r(T ) < 1, then limn→∞ kT nk = 0,
and so, for any unitary intertwining pair (X, U ) of T , the equations XT n =
U nX (n ∈ N) imply that X = 0. Thus U acts on the zero space, provided
(X, U ) is minimal. Therefore r(T ) = 1, whenever the power bounded operator
T has a non-trivial unitary asymptote.
Taking a Banach limit L : ℓ∞ → C, there exists a unique operator AL ∈ L(H)
such that
hALx, yi = L- lim
n→∞
hT ∗nT nx, yi
for all x, y ∈ H.
It is evident that AL ∈ T+(T ), and AL ≤ M 2
relations hBx, xi = hBT nx, T nxi and
T · I. Setting any B ∈ Ts(T ), the
−kBkhT nx, T nxi ≤ hBT nx, T nxi ≤ kBkhT nx, T nxi
(x ∈ H, n ∈ N)
show that
−kBkAL ≤ B ≤ kBkAL.
Therefore, AL is universal in T (T ).
Since T ∗ALT = AL implies kA1/2
L T hk = kA1/2
a unique isometry VL on the space K+,L = (A1/2
L hk (h ∈ H), there exists
L H)− satisying the condition
5
L = A1/2
VLA1/2
L T . Let UL ∈ L(KL) be the minimal unitary extension of VL, and
let XL ∈ L(H, KL) be defined by XLh := A1/2
L h (h ∈ H). Clearly, (XL, UL) is a
minimal unitary intertwining pair for T . Taking into account that X ∗
LXL = AL
is universal in T (T ), we infer by Theorem 3 that (XL, UL) is a unitary asymptote
of T .
We recall that the possible values of Banach limits on a real sequence ξ ∈ ℓ∞
comprise a closed interval determined by the numbers
B(ξ) = B- lim
n→∞
and
B(ξ) = B- lim
n→∞
ξ(n) := sup
ξ(n) := inf
For any h ∈ H, we have
lim inf
k→∞
1
r
lim sup
k→∞
ξ(nj + k) : n1, . . . , nr ∈ N, r ∈ N
rXj=1
ξ(nj + k) : n1, . . . , nr ∈ N, r ∈ N
rXj=1
.
1
r
kXLhk2 = hALh, hi = L- lim
n→∞
kT nhk2,
whence
B- lim
n→∞
kT nhk2 ≤ kXLhk2 ≤ B- lim
n→∞
kT nhk2.
These relations immediately yield that
lim inf
n→∞
kT nhk ≤ kXLhk ≤ lim sup
n→∞
kT nhk (h ∈ H).
We obtain the following characterization.
Theorem 5 Let (X, U ) be a minimal unitary intertwining pair for the power
bounded operator T ∈ L(H). Then (X, U ) is a unitary asymptote of T if and
only if there exists a κ ∈ (0, ∞) such that
κ · lim inf
n→∞
kT nhk ≤ kXhk
for all h ∈ H.
Proof. If (X, U ) is a unitary asymptote of T , then there exists an invertible
Z ∈ I(U, UL) such that XL = ZX, with a Banach limit L. Thus, for every
h ∈ H, we have
kXhk ≥ kZk−1kZXhk = kZk−1kXLhk ≥ kZk−1 lim inf
n→∞
kT nhk.
The reverse implication follows by Proposition 1.
As immediate consequences, we may derive the following statements.
(cid:3)
Corollary 6 Let (Xj, Uj) be a unitary asymptote of the power bounded operator
Tj, for j = 1, 2. Then (X1 ⊕ X2, U1 ⊕ U2) is a unitary asymptote of T1 ⊕ T2.
6
Corollary 7 If (X, U ) is a unitary asymptote of the power bounded operator
T ∈ L(H) and M ∈ Lat T , then (XM, U fM) is a unitary asymptote of the
restriction T M, where fM := ∨n∈NU −nXM.
n=1 is
convergent, for every h ∈ H. Suprisingly, these sequences are convergent also in
the case, when T is a ρ-contraction; see [E] and [CC].
If T ∈ L(H) is a contraction, then the decreasing sequence {kT nhk}∞
If T ∈ L(H) is an arbitrary power bounded operator, then for every h ∈ H
we have
lim sup
n→∞
kT nhk ≤ MT lim inf
n→∞
kT nhk.
In particular, inf{kT nhk : n ∈ N} = 0 implies limn→∞ kT nhk = 0. In view of
the previous observations we may infer the following statement.
Proposition 8 If (X, U ) is a unitary asymptote of the power bounded operator
T ∈ L(H), then the nullspace of X coincides with the hyperinvariant subspace
of those vectors, which are stable for T :
ker X = H0(T ) :=nh ∈ H :
lim
n→∞
kT nhk = 0o .
We conclude this section with norm-estimates concerning the symbolic cal-
culus and the commutant mapping.
Proposition 9 Setting a Banach limit L, let (XL, UL) be the corresponding
unitary asymptote of the power bounded operator T ∈ L(H). Then
(i) kF k ≤ kΨXL(F )k ≤ M 2
T kF k
for all F ∈ {UL}′;
(ii) M −2
T kALCk ≤ kγXL (C)k ≤ kALCk
for all C ∈ {T }′.
Proof. For (i) see the proof of Theorem 3 in [K6]. Statement (ii) is a conse-
quence of (i) and Proposition 4.
(cid:3)
3 Lebesgue decomposition
Let P(T) denote the algebra of analytic polynomials, restricted to the unit circle
T = {ζ ∈ C : ζ = 1}. C(T) is the abelian Banach algebra of all continuous
functions defined on T. The disk algebra A(T) is the norm-closure of P(T) in
C(T). We recall that the dual of C(T) can be identified with the Banach space
M (T) of all complex Borel measures on T (which are automatically regular).
Namely, by the Riesz Representation Theorem, the mapping
Ψ : M (T) → C(T)#, µ 7→ ψµ,
where ψµ(f ) =ZT
f dµ,
is a Banach space isomorophism.
7
We say that µ, ν ∈ M (T) are analytically equivalent measures, in notation:
µ a∼ ν, if
ZT
u dµ =ZT
u dν
for all u ∈ A(T).
It is clear that this is an equivalence relation. Furthermore, by a well-known
theorem of F. & M. Riesz, µ a∼ ν holds if and only if ν = µ + h dm with
a function h ∈ H 1
0 (see, e.g., [Ho]). Here and in the sequel m denotes the
normalized Lebesgue measure on T, and H 1
note that if the measures µ, ν are singular (with respect to m), then µ a∼ ν holds
exactly when µ = ν.
0 = {f ∈ H 1 : RT f dm = 0}. We
For any T ∈ L(H), the polynomial calculus
ΦT,0 : P(T) → L(H), p 7→ p(T )
is the uniquely determined (unital) algebra-homomorphism, which transforms
the identical function χ(ζ) = ζ (ζ ∈ T) into T . We say that T is a polyno-
mially bounded operator, if ΦT,0 is norm-continuous. In that case ΦT,0 can be
continuously extended to a bounded algebra-homomorphism
ΦT,1 : A(T) → L(H), u 7→ u(T ).
Our aim in this section is to decompose T into the direct sum of an abso-
lutely continuous component and a singular component. We shall follow Mlak's
method of using elementary measures.
More generally, we consider and arbitrary bounded, linear transformation
Φ : A(T) → L(H). For any x, y ∈ H, ϕx,y(C) = hCx, yi defines a bounded
linear functional on L(H). By the Hahn -- Banach Theorem ϕx,y ◦ Φ can be
extended to a bounded linear functional on C(T), even in a norm-preserving
way. The set of elementary measures of Φ at x, y is defined by
M (Φ, x, y) := {µ ∈ M (T) : ψµA(T) = ϕx,y ◦ Φ} .
Therefore, µ ∈ M (Φ, x, y) precisely when
hΦ(u)x, yi =ZT
u dµ
holds for all u ∈ A(T).
We say that Φ is absolutely continuous (a.c.), if for every x, y ∈ H the measures
in M (Φ, x, y) are a.c. (with respect to m). On the other hand, Φ is called
singular, if for every x, y ∈ H the set M (Φ, x, y) contains a measure, which is
singular (with respect to m).
It is well-known that every measure µ ∈ M (T) can be uniquely decomposed
into the sum µ = µa + µs, where µa ∈ M (T) is a.c. and µs ∈ M (T) is singular.
By the next proposition such Lebesgue-type decomposition is valid also for Φ.
Proposition 10 If Φ : A(T) → L(H) is a bounded linear transformation, then
8
(i) there exists a unique decomposition Φ = Φa + Φs, where the bounded
linear mapping Φa : A(T) → L(H) is a.c., and the bounded linear mapping
Φs : A(T) → L(H) is singular;
(ii) kΦak ≤ kΦk and kΦsk ≤ kΦk;
(iii) for every x, y ∈ H, we have M (Φa, x, y) = {µa : µ ∈ M (Φ, x, y)} and
M (Φs, x, y) = {µs + h dm : µ ∈ M (Φ, x, y), h ∈ H 1
0 }.
Proof. For every x, y ∈ H, the singular component of the measures in M (Φ, x, y)
is uniquely determined, denoted by µs
Φ,x,y. Taking a measure µ ∈ M (Φ, x, y)
with minimal norm, we obtain
kµs
Φ,x,yk ≤ kµk = kϕx,y ◦ Φk ≤ kΦk · kxk · kyk.
The mapping
ws
Φ : H × H → M (T), (x, y) 7→ µs
Φ,x,y
is sesquilinear (linear in x, and conjugate linear in y). For example, let us
check the linearity in x. Setting x1, x2, y ∈ H, c ∈ C, let us choose µj ∈
M (Φ, xj, yj) (j = 1, 2) and µ ∈ M (Φ, cx1 + x2, y). Then, for every u ∈ A(T),
we have
ZT
u dµ = hΦ(u)(cx1 + x2), yi = chΦ(u)x1, yi + hΦ(u)x2, yi
= cZT
u dµ1 +ZT
u dµ2 =ZT
u d(cµ1 + µ2),
whence µ a∼ cµ1 + µ2, and so µs = (cµ1 + µ2)s = cµs
2 follows.
For any g ∈ C(T), let us consider the bounded linear functional
1 + µs
Λg : M (T) → C, µ 7→ZT
g dµ,
with kΛgk = kgk. The bounded sesquilinear functional
ws
Φ,g := Λg ◦ ws
Φ : H × H → C
satisfies the condition kws
Φ,gk ≤ kgkkws
Φk ≤ kgkkΦk. There exists a unique
operator eΦs(g) ∈ L(H) such that
DeΦs(g)x, yE = ws
Φ,g(x, y) =ZT
holds for all x, y ∈ H; furthermore keΦs(g)k = kws
the integral in g yields that the mapping eΦs : C(T) → L(H) is linear; we obtain
also that keΦsk ≤ kΦk. Then
Φ,gk ≤ kΦk · kgk. Linearity of
g dµs
Φ,x,y
Φs :=eΦsA(T) : A(T) → L(H)
9
is also a bounded linear transformation, and kΦsk ≤ kΦk. Given any x, y ∈ H,
for every u ∈ A(T), we have
hΦs(u)x, yi =ZT
u dµs
Φ,x,y;
hence M (Φs, x, y) = {µs
Φ,x,y + h dm : h ∈ H 1
0 }, and so Φs is singular.
Let us consider the bounded linear transformation Φa := Φ − Φs. Given any
x, y ∈ H and setting µ ∈ M (Φ, x, y), for every u ∈ A(T), we have
hΦa(u)x, yi = hΦ(u)x, yi − hΦs(u)x, yi =ZT
u dµ −ZT
u dµs =ZT
u d(µ − µs).
Thus µa = µ − µs ∈ M (Φa, x, y), and so Φa is a.c.. Choosing µ so that kµk =
kϕx,y ◦ Φk, we obtain
hΦa(u)x, yi ≤ kuk · kµak ≤ kuk · kµk ≤ kukkΦkkxkkyk.
Therefore kΦa(u)k ≤ kΦk · kuk (u ∈ A(T)), and so kΦak ≤ kΦk.
Φa
1 + Φs
1 = Φa
2 + Φs
Finally, we turn to the uniqueness of the decomposition. Suppose that Φ =
2 − Φs
1
2 are Lebesgue decompositions. Then eΦ = Φa
is simultaneously a.c. and singular. Thus, the elementary measures of eΦ are
analytically equivalent to 0, and so eΦ = 0.
2 = Φs
1 − Φa
Remark 11 We note that these considerations can be carried out in the Banach
space setting also, with some restrictions. Let X be a complex Banach space,
and let Φ : A(T) → L(X ) be a bounded linear transformation. For any x ∈ X
and y ∈ X #, the measure µ ∈ M (T) belongs to M (Φ, x, y) if
(cid:3)
ZT
u dµ = [Φ(u)x, y]
for all u ∈ A(T).
For any g ∈ C(T), we can define the bounded bilinear functional
ws
φ,g : X × X # → C
as before. Then a necessary and sufficient condition for the existence of an
operator eΦs(g) ∈ L(X ) satisfying
heΦs(g)x, yi = ws
is that ws
Φ,g(x, y) be weak-∗ continuous in y.
Φ,g(x, y),
for all x ∈ X , y ∈ X #,
Now we turn to bounded representations of A(T). We recall that a map-
ping Φ : A(T) → L(H) is called a representation,
if it is a unital algebra-
homomorphism. Norm-continuous representations arise as functional calculae
for polynomially bounded operators: Φ = ΦT,1 where T = Φ(χ).
10
Uniqueness of the Lebesgue decomposition of measures implies transforma-
tion rules for elementary measures. Given any µ, ν ∈ M (T), if µ a∼ ν then
0 , and so µa a∼ νa and µs = νs. Furthermore,
ν = µ + h dm with some h ∈ H 1
for any v ∈ A(T), the equation v dµ = v dµa + v dµs yields that (v dµ)a = v dµa
and (v dµ)s = v dµs.
Lemma 12 Let Φ : A(T) → L(H) be a bounded representation. Setting x, y ∈
H and v ∈ A(T), let xv := Φ(v)x and y∗
v := Φ(v)∗y. If µx,y ∈ M (Φ, x, y), µxv ,y ∈
M (Φ, xv, y) and µx,y∗
v ∈ M (Φ, x, y∗
v), then
µa
xvy
a∼ v dµa
x,y, µs
xv ,y = v dµs
x,y, µa
x,y∗
v
a∼ v dµa
x,y, µs
x,y∗
v
= v dµs
x,y.
Proof. For every u ∈ A(T), we have
ZT
uv dµx,y = hΦ(uv)x, yi = hΦ(u)Φ(v)x, yi = hΦ(u)xv, yi =ZT
u dµxv,y,
a∼ v dµx,y, and so µa
x,y follow. The
whence µxv,y
other two relations can be derived similarly from the equation hΦ(uv)x, yi =
hΦ(u)x, y∗
xv,y = v dµs
x,y and µs
a∼ v dµa
xv,y
vi.
(cid:3)
We say that the polynomially bounded operator T ∈ L(H) is absolutely
continuous (singular) if its functional calculus ΦT,1 is an absolutely continuous
(singular, respectively) representation.
Theorem 13 Let T ∈ L(H) be a polynomially bounded operator, and let us
consider the Lebesgue decomposition Φ = Φa + Φs of the bounded representation
Φ = ΦT,1 : A(T) → L(H). Then
(i) Φa and Φs are also bounded representations;
(ii) P a = Φa(1) and P s = Φs(1) are complementary projections belonging to
the bicommutant {T }′′;
(iii) H = Ha + Hs, where Ha = P aH and Hs = P sH are hyperinvariant
subspaces of T ;
(iv) T a = T Ha is an a.c. and T s = T Hs is a singular polynomially bounded
operator; the Lebesgue decomposition T = T a + T s of T is unique; further-
more, Φa(u) = ΦT a,1(u) + 0 and Φs(u) = 0 + ΦT s,1(u) for all u ∈ A(T).
Proof. (i): Given x, y ∈ H, set µx,y ∈ M (Φ, x, y). Then µa
and v µa
we have
x,y ∈ M (Φa, x, y)
xv,y by Proposition 10 and Lemma 12. Hence, for u, v ∈ A(T),
a∼ µa
x,y
hΦa(uv)x, yi =ZT
uv dµa
x,y =ZT
u dµa
xv,y = hΦa(u)xv, yi = hΦa(u)Φ(v)x, yi .
11
Consequently,
Similarly,
Φa(uv) = Φa(u)Φ(v)
for all u, v ∈ A(T).
hΦa(uv)x, yi = ZT
v dµa
x,y∗
u
uv dµa
x,y =ZT
= hΦa(v)x, y∗
ui = hΦa(v)x, Φ(u)∗yi = hΦ(u)Φa(v)x, yi ,
and so
Then
Φa(uv) = Φ(u)Φa(v).
Φa(u)Φa(v) + Φa(u)Φs(v) = Φa(u)Φ(v) = Φa(uv) = Φ(u)Φa(v)
= Φa(u)Φa(v) + Φs(u)Φa(v),
whence
Φa(u)Φs(v) = Φs(u)Φa(v)
for all u, v ∈ A(T).
Fix v ∈ A(T), and set x, y ∈ H. Let x1 := Φs(v)x and x2 := Φa(v)x. Then, for
every u ∈ A(T), we have
ZT
Thus µa
x1,y
Therefore
u dµa
x1,y = hΦa(u)x1, yi = hΦa(u)Φs(v)x, yi
= hΦs(u)Φa(v)x, yi = hΦs(u)x2, yi =ZT
u dµs
x2,y.
a∼ µs
x2,y, and so µs
x2,y = 0, which implies that hΦs(u)Φa(v)x, yi = 0.
Φa(u)Φs(v) = Φs(u)Φa(v) = 0
for all u, v ∈ A(T).
Φa(uv) = Φa(u)Φ(v) = Φa(u)Φa(v) + Φa(u)Φs(v) = Φa(u)Φa(v),
Now
and so
Φs(uv) = Φ(uv) − Φa(uv) = (Φa(u) + Φs(u)) (Φa(v) + Φs(v)) − Φa(u)Φa(v)
= Φs(u)Φs(v).
Therefore, Φa and Φs are bounded representations.
(ii): It is clear that (P a)2 = (Φa(1))2 = Φa(12) = P a, and similarly (P s)2 =
P s. Moreover, P aP s = Φa(1)Φs(1) = 0 = Φs(1)Φa(1) = P sP a and P a + P s =
Φa(1) + Φs(1) = Φ(1) = I.
12
Let C ∈ {T }′ be arbitrary. Then CΦ(u) = Φ(u)C for any u ∈ A(T). Given
x, y ∈ H, set µCx,y ∈ M (Φ, Cx, y) and µx,C ∗y ∈ M (Φ, x, C∗y). Since, for every
u ∈ A(T), we have
ZT
u dµCx,y = hΦ(u)Cx, yi = hCΦ(u)x, yi = hΦ(u)x, C∗yi =ZT
u dµx,C ∗y,
it follows that µCx,y
Then, for every u ∈ A(T),
a∼ µx,C ∗y, and so µa
Cx,y
a∼ µa
x,C ∗y and µs
Cx,y = µs
x,C ∗y.
hΦa(u)Cx, yi = ZT
u dµa
x,C ∗y
u dµa
Cx,y =ZT
= hΦa(u)x, C∗yi = hCΦa(u)x, yi
for all x, y ∈ H.
Thus Φa(u)C = CΦa(u), whence Φs(u)C = CΦs(u) follows. Therefore
Φa(u), Φs(u) ∈ {T }′′
for all u ∈ A(T).
In particular, P a, P s ∈ {T }′′.
(iii): Since P a, P s ∈ {T }′′, it follows that the subspaces Ha = P aH and
Hs = P sH are hyperinvariant for T . Furthermore, P aP s = P sP a = 0 and
P a + P s = I imply that Ha + Hs = H.
(iv): Let us consider the decomposition T = T a + T s, where T a := T Ha and
T s := T Hs. For any u ∈ A(T), we have Φa(u)P a = P aΦa(u) = Φa(1)Φa(u) =
Φa(u), whence Φa(u) = (Φa(u)Ha) + 0 follows. The bounded representation
Φa
0 : A(T) → L(Ha), Φa
0(u) := Φa(u)Ha
0 = ΦT a,1. Thus T a is
satisfies the condition Φa
an a.c. polynomially bounded operator. Similarly, Φs(u) = 0 + (Φs(u)Hs) for
u ∈ A(T), and ΦT s,1 coincides with the singular representation
0(χ) = T a = ΦT a,1(χ), and so Φa
Φs
0 : A(T) → L(Hs), Φs
0(u) := Φs(u)Hs.
Thus T s is a singular polynomially bounded operator.
As for uniqueness, let us suppose that the decomposition H = H1 + H2 is
reducing for T , the restriction T1 = T H1 is a.c., and T2 = T H2 is singular.
For every u ∈ A(T), we have Φ(u) = ΦT,1(u) = ΦT1,1(u) + ΦT2,1(u). Let us
consider also the bounded representations Φ1, Φ2 : A(T) → L(H) defined by
Φ1(u) := ΦT1,1(u) + 0 and Φ2(u) := 0 + ΦT2,1(u). Since T1 is a.c., it follows
that Φ1 is a.c.; similarly, singularity of T2 implies that Φ2 is singular. Taking
into account that Φ = Φ1 + Φ2, the uniqueness part of Proposition 10 implies
that Φ1 = Φa and Φ2 = Φs. Thus H1 = Φ1(1)H = Φa(1)H = Ha and H2 =
Φ2(1)H = Φs(1)H = Hs.
(cid:3)
Concluding this section we examine the adjoint of a polynomially bounded
operator T ∈ L(H). For any function f : T → C, let ef (ζ) := f (ζ) (ζ ∈ T).
13
n=0 cnζn, and so
is also polynomially bounded, and kΦT ∗,0k = kΦT,0k. Taking uniform limits of
n=0 cnζn, we have ep(ζ) = PN
For a polynomial p(ζ) = PN
p(T ∗) = ep(T )∗. Since kp(T ∗)k = kep(T )k and kpk = kepk, it follows that T ∗
polynomials, we obtain that for any u ∈ A(T), eu ∈ A(T) and
u(T ∗) =eu(T )∗.
For any measure µ ∈ M (T), eµ ∈ M (T) is defined by eµ(ω) := µ(ω), upper
bar meaning complex conjugation. We shall use the notation M (T, x, y) =
M (ΦT,1, x, y).
Proposition 14 If T ∈ L(H) is polynomially bounded, then T ∗ is also polyno-
mially bounded, and
(i) M (T ∗, x, y) = {eµ : µ ∈ M (T, y, x)} for x, y ∈ H;
(ii) T is a.c. if and only if T ∗ is a.c.;
(iii) T is singular if and only if T ∗ is singular.
Proof. Given x, y ∈ H, set µ ∈ M (T, y, x) and ν ∈ M (T ∗, x, y). For every
u ∈ A(T), we have
ZT
= ZT
u dν = hu(T ∗)x, yi = heu(T )∗x, yi = heu(T )y, xi =ZTeu dµ
u(ζ) deµ(ζ) =ZT
whence ν a∼eµ follows. Sinceeh ∈ H 1
to see that eµ is a.c. (singular) if and only if µ is a.c. (singular, respectively),
which shows the validity of (ii) and (iii).
u deµ,
0 , for any h ∈ H 1
0 , we obtain (i). It is easy
(cid:3)
4 Intertwining relations
Let T1 ∈ L(H1) and T2 ∈ L(H2) be polynomially bounded operators, and let
us assume that Q ∈ I(T1, T2). Given x ∈ H1 and y ∈ H2, set µ ∈ M (T2, Qx, y)
and ν ∈ M (T1, x, Q∗y). For every u ∈ A(T), we have
ZT
u dµ = hu(T2)Qx, yi = hu(T1)x, Q∗yi =ZT
u dν,
hence µ a∼ ν, and so
M (T2, Qx, y) = M (T1, x, Q∗y).
14
Proposition 15 Let T1 ∈ L(H1) be an a.c. polynomially bounded operator, and
let T2 ∈ L(H2) be a singular polynomially bounded operator. Then I(T1, T2) =
{0} and I(T2, T1) = {0}.
Proof. Let Q ∈ I(T1, T2) be arbitrary. Given x ∈ H1 and y ∈ H2, and
setting µ ∈ M (T2, Qx, y) and ν ∈ M (T1, x, Q∗y), we obtain from the previous
discussion that µ a∼ ν. Since µ is singular and ν is a.c., it follows that µ = 0.
Then hQx, yi = RT 1 dµ = 0; and since x ∈ H1, y ∈ H2 were arbitrary, we
conclude that Q = 0. Turning to adjoints and applying Proposition 14, we
obtain that I(T2, T1) = {0} also holds.
We recall that T1 is a quasiaffine transform of T2, in notation: T1 ≺ T2,
if I(T1, T2) contains a quasiaffinity, i.e., an injective transformation with dense
range.
Proposition 16 Let T1 ∈ L(H1) and T2 ∈ L(H2) be polynomially bounded
operators, and let us assume that T1 ≺ T2. Then
(cid:3)
(i) T1 is a.c. if and only if T2 is a.c.;
(ii) T1 is singular if and only if T2 is singular.
2 + T s
2 ∈ L(H2) be the (oblique) projection onto Hs
Proof. Let us assume that T1 is a.c., and let us consider the Lebesgue de-
composition T2 = T a
2 of T2. There exists a quasiaffinity Q ∈ I(T1, T2).
Let P s
2. Since the transformation
Qs ∈ L(H1, Hs
2 , we
infer by Proposition 15 that Qs = 0. Thus QH1 ⊂ Ha
2, and since Q has dense
range, we obtain that Hs
2 = {0}. Therefore T2 is a.c.. Conversely, assuming
that T2 is a.c., the relation T ∗
1 is a.c., and then so is T1 too
by Proposition 14. Statement (ii) can be proved similarly.
2 Qx (x ∈ H1), intertwines T1 with T s
2), defined by Qsx := P s
2 ≺ T ∗
1 yields that T ∗
Let U ∈ L(K) be a unitary operator. Relying on the Gelfand transform
of the abelian C∗-algebra generated by U , it can be shown that there exists a
uniquely determined isometric ∗-representation Φ : C(σ(U )) → L(K) such that
Φ(χ) = U . Here σ(U ) ⊂ T is the spectrum of U . This Φ induces a contractive ∗-
(cid:3)
representation eΦ : C(T) → L(K), defined by f (U ) = eΦ(f ) := Φ(f σ(U )). Since
eΦP(T) = ΦU,0, it follows that U is polynomially bounded with kΦU,0k = 1;
furthermore eΦA(T) = ΦU,1.
It is known also that eΦ can be represented by
integration with respect to a uniquely determined spectral measure E : BT →
P(K); see, e.g., [C]. (Here BT denotes the σ-algebra of Borel sets on T, and
P(K) stands for the set of orthogonal projections on K.) Namely, for every
f ∈ C(T),
hf (U )x, yi =ZT
f dEx,y
(x, y ∈ K),
15
where Ex,y ∈ M (T) is the localization of E to x, y, defined by Ex,y(ω) :=
hE(ω)x, yi (ω ∈ BT). Therefore
M (U, x, y) = {Ex,y + h dm : h ∈ H 1
0 }.
We recall that the unitary operator U is called a.c. (singular, resp.), if Ex,y
is a.c. (singular, resp.) for every x, y ∈ K. The previous relation shows that
these properties coincide with the corresponding properties considering U as a
polynomially bounded operator.
Let us consider the Lebesgue decomposition U = U a + U s ∈ L(K = Ka + Ks).
Since Ka ∈ Hlat U and U ∗ = U −1 ∈ {U }′, it follows that Ka is reducing for
U , and so the orthogonal projection Qa ∈ L(K) onto Ka commutes with U .
Therefore QaKs ∈ I(U s, U a), whence QaKs = 0 follows by Proposition 15.
We conclude that Ka is orthogonal to Ks. (For a discussion of Lebesgue decom-
position of unitaries, based on spectral measures, see [Ha].)
Let us consider now a contraction T ∈ L(H) : kT k ≤ 1. It is known that T
can be decomposed into the orthogonal sum T = Tc ⊕ U ∈ L(H = Hc ⊕ Hu) of a
completely non-unitary contraction Tc and a unitary operator U . By Sz.-Nagy's
Dilation Theorem Tc has a minimal unitary dilation W , acting on a Hilbert space
K, containing Hc. We recall that W is an a.c. unitary operator, T n
c = PcW nHc
n=1W −nHc. (Here Pc ∈ L(K) is the orthogonal
for every n ∈ Z+, and K = ∨∞
projection onto Hc.) In connection with the theory of contractions we refer to
[NFBK]. For every p ∈ P(T), we have
kp(Tc)k = kPcp(W )Hck ≤ kp(W )k ≤ kpk,
what is called the von Neumann inequality, and so Tc is polynomially bounded.
Since, for any x, y ∈ Hc, hu(Tc)x, yi = hPcu(W )x, yi = hu(W )x, yi holds for
every u ∈ A(T), we infer that
M (Tc, x, y) = M (W, x, y).
Therefore Tc is a.c.. Considering the decomposition T = Tc ⊕ U a ⊕ U s, we can
see that the a.c. component of T is T a = Tc ⊕ U a, while the singular part is
T s = U s.
If T ∈ L(H) is a polynomially bounded operator, U ∈ L(K) is a singular
unitary operator, and T ≺ U , then T is singular by Proposition 16. The fol-
lowing theorem, which in different forms can be found in [AT] and [M4], states
that there must be a much stronger relation between T and U in that case.
Theorem 17 If T ∈ L(H) is a singular polynomially bounded operator, then
(i) the funcional calculus ΦT,1 : A(T) → L(H) can be extended to a bounded
representation eΦT,1 : C(T) → L(H);
(ii) the operator X ∈ I(T, U ) is invertible, whenever (X, U ) is a unitary
asymptote of T ; and so T is similar to the singular unitary operator U .
16
Proof. (i): We repeat the procedure carried out in the first part of the proof
of Proposition 10. We shall write Φ = ΦT,1 and eΦ = eΦT,1 for short. For any
x, y ∈ H, there exists a unique singular measure µx,y in M (T, x, y). Let us
consider the bounded sesquilinear mapping
wT : H × H → M (T), wT (x, y) = µx,y.
For any g ∈ C(T), Λg : M (T) → C is defined by Λg(µ) =RT g dµ. The bounded
sesquilinear functional
wT,g := Λg ◦ wT : H × H → C
uniquely determines an operator eΦ(g) ∈ L(H):
heΦ(g)x, yi = wT,g(x, y) =ZT
g dµx,y
(x, y ∈ H).
C(T),
f i =ZT
For v ∈ A(T), we have v dµx,y = µxv,y by Lemma 12. Thus, for any f ∈
Thus, the singular measures f dµx,y and µx,y∗
so they must coincide: f dµx,y = µx,y∗
f
f dµxv ,y = heΦ(f )xv, yi = heΦ(f )Φ(v)x, yi,
f =eΦ(f )∗y, for every v ∈ A(T),
We know that keΦ(g)k ≤ kΦkkgk. It is clear also that eΦ is linear. We have to
show yet that eΦ is multiplicative.
f v dµx,y =ZT
heΦ(f v)x, yi =ZT
whence eΦ(f v) =eΦ(f )Φ(v) follows. Setting y∗
ZT
vf dµx,y = heΦ(f v)x, yi = heΦ(f )Φ(v)x, yi = hΦ(v)x, y∗
heΦ(f g)x, yi =ZT
f i = heΦ(f )eΦ(g)x, yi
= heΦ(g)x, y∗
holds for every x, y ∈ H. Consequently, eΦ(f g) =eΦ(f )eΦ(g).
(ii): By the multiplicativity ofeΦ we infer that T is invertible andeΦ(χn) = T n
holds, for every n ∈ Z. Since kT nk ≤ keΦk = kΦk for all n ∈ Z, a well-
known theorem of Sz.-Nagy yields that T is similar to a unitary operator V ;
see [N]. Let Q ∈ I(T, V ) be invertible. The polynomially bounded operator T
is necessarily power bounded, and so it has a unitary asymptote (X, U ). There
exists a unique Y ∈ I(U, V ) such that Q = Y X. Since Q is invertible, it follows
that X is bounded from below. The isometry U X is singular by Proposition 16.
Considering its Wold decomposition, we can see that U X is unitary. Hence XH
is reducing for U , and so the minimality of (X, U ) yields that X is a surjection.
f g dµx,y =ZT
are analytically equivalent, and
v dµx,y∗
.
f
. Taking any g ∈ C(T),
f
g dµx,y∗
f
17
Therefore X is invertible, and then Y must be invertible too. Consequently, the
operators T, U and V are similar to each other.
(cid:3)
We conclude this section with a discussion of further intertwining relations.
Since Hilbert space contractions have a rich theory (see [NFBK]), it would be im-
portant to know how more general operators can be related to contractions. An-
swering a question posed by Sz.-Nagy, Foguel gave examples for power bounded
operators which are not similar to contractions; see [F]. Pisier answered nega-
tively also a more delicate question of Halmos, showing that not every polynomi-
ally bounded operator is similar to a contraction; see [Pi]. Muller and Tomilov
proved, giving negative answer for a question of the author posed in [K1], that
there are power bounded operators of class C11 which are not similar to con-
tractions; see [MT]. We recall that T ∈ C11 means H0(T ) = H0(T ∗) = {0},
and in that case T is quasisimilar to its unitary asymptote U : T ∼ U , that is
T ≺ U and U ≺ T hold simultaneously.
Question 18 If the polynomially bounded operator T is of class C11, is T sim-
ilar to a contraction?
For additional conditions under which a power bounded operator is similar
to a contraction see [G2]. We note yet that by a fundamental characterization
due to Paulsen, exactly those operators are similar to contractions which are
completely bounded; see [Pa].
Quasisimilarity is a much weaker relation than similarity, but it preserves
also numerous properties and plays important role in the classification of opera-
tors; see, e.g., [B] and [DH]. Muller and Tomilov showed that a power bounded
operator T is not necessarily quasisimilar to a contraction [MT]. On the other
hand, Bercovici and Prunaru proved that if T is a polynomially bounded oper-
ator, then there exist contractions T1 and T2 such that T1 ≺ T ≺ T2; see [BP].
As far as we know, the following questions are still open.
Question 19 Is every polynomially bounded operator quasisimilar to a contrac-
tion?
Question 20 If the power bounded operator T is quasisimilar to a singular
unitary operator V , does it follow that T is similar to V ?
The latter question is connected with Theorem 17, and was posed in [K1].
Partial answer for it can be found in [G1].
5 H ∞-functional calculus
The Hardy class H ∞ is the weak-∗ closed subalgebra of L∞(T), consisting of
those functions f whose Fourier coefficients satisfy the condition bf (−n) = 0
for n ∈ N. H ∞ can be identified as the dual of the Banach space L1(T)/H 1
0 .
18
For a detailed study of this class see [Ho]. Sz.-Nagy and Foias introduced H ∞-
functional calculus for a.c. contractions, and thoroughly exploited its properties
in their theory of contractions; see [NFBK] and [B]. In Chapter 2 of [CP] H ∞-
functional calculus is defined for a polynomially bounded operator T acting on
a complex Banach space X , which is stable, that is limn→∞ kT nxk = 0 holds
for avery x ∈ X . In [M1] and [M3] Mlak considered representations of general
function algebras. Here we follow Mlak's method of elementary measures, con-
centrating on H ∞, and providing detailed study of the calculus in this case.
We note that in [K6] our approach was based on the unitary asymptote in the
C1·-case, that is when H0(T ) = {0}.
We recall that L(H) is the dual of the Banach space C1(H) of trace class
operators. Namely, the mapping Λ : L(H) → C1(H)#, defined by [A, Λ(C)] =
tr(AC), is a Banach space isomorphism; see [Sch].
We say that the operator T ∈ L(H) admits an H ∞-functional calculus, if
there exists a weak-∗ continuous representation ΦT,2 : H ∞ → L(H) such that
ΦT,2(χ) = T . In that case we use the notation ΦT,2(f ) = f (T ) (f ∈ H ∞). In
the following proposition we collect some basic facts about this calculus.
Proposition 21 If T ∈ L(H) admits an H ∞-functional calculus, then
(i) ΦT,2 is uniquely determined;
(ii) ΦT,2 is norm-continuous;
(iii) T is polynomially bounded, ΦT,2A(T) = ΦT,1, and kΦT,2k = kΦT,1k;
(iv) H ∞(T ) := ran ΦT,2 ⊂ {T }′′;
(v) Lat T = Lat H ∞(T );
(vi) for every M ∈ Lat T, T M admits an H ∞-functional calculus, and
f (T M) = f (T )M (f ∈ H ∞).
Proof. (i): Let us assume that Φj : H ∞ → L(H) is a weak-∗ continuous repre-
sentation, for j = 1, 2, and that Φ1(χ) = Φ2(χ) holds. Then Φ1(p) = Φ2(p) is
true, for every polynomial p ∈ P(T). Taking an arbitrary f ∈ H ∞, the Cesaro
means
of the Fourier series of f converge to f in the weak-∗ topology, as n → ∞; see,
e.g., Chapter 2 in [Ho]. By continuity we obtain that Φ1(f ) = Φ2(f ).
(ii): For short we write Φ = ΦT,2. The unit ball (H ∞)1 := {f ∈ H ∞ :
kf k ≤ 1} is weak-∗ compact by the Banach -- Alaoglu Theorem. Since Φ is
weak-∗ continuous, it follows that Φ(H ∞)1 is weak-∗ compact in L(H). Thus
{[A, Φ(f )] : f ∈ (H ∞)1} is compact and so bounded in C, for every A ∈ C1(H).
Then the Uniform Boundedness Principle yields that Φ(H ∞)1 is bounded in
L(H), which means the norm-continuity of Φ.
σf,n =
nXk=0(cid:18)1 −
k
n + 1(cid:19)bf (k)χk
19
(iii): Since ΦP(T) = ΦT,0, we infer from (ii) that ΦT,0 is bounded: kΦT,0k ≤
kΦk. Hence T is polynomially bounded, and ΦA(T) = ΦT,1. Furthermore, for
any f ∈ H ∞, we have kΦ(σf,n)k ≤ kΦT,0k · kf k (n ∈ N). Taking into account
that w∗-limn Φ(σf,n) = Φ(f ), we conlude that kΦ(f )k ≤ kΦT,0k · kf k. Therefore
kΦk = kΦT,0k = kΦT,1k.
The statements (iv), (v), (vi) can be easily derived from the fact that H ∞
is the (sequentially) weak-∗ closure of P(T).
(cid:3)
The following lemma shows that it is enough to check seemingly weaker
conditions in order to prove that a mapping is an H ∞-functional calculus.
Lemma 22 If Φ : H ∞ → L(H) is a linear mapping, ΦA(T) is a representa-
tion, and limnhΦ(fn)x, yi = 0 for every x, y ∈ H whenever w∗-limn fn = 0, then
Φ is an H ∞-functional calculus.
Proof. Let us assume that the sequence {fn}∞
n=1 in H ∞ converges to zero in the
weak-∗ topology. By the assumption, the operators {Φ(fn)}∞
n=1 converge to zero
in the weak operator topology (wot). The Uniform Boundedness Principle yields
that M = sup{kΦ(fn)k : n ∈ N} < ∞. Taking an arbitrary operator A ∈ C1(H),
let us consider the Hilbert -- Schmidt decomposition A = Pk sk xk ⊗ yk. Here
{xk}k and {yk}k form orthonormal systems, sk ≥ 0 for all k, and Pk sk =
kAk1 < ∞. These conditions imply that
lim
n→∞
tr(Φ(fn)A) = lim
skhΦ(fn)xk, yki = 0.
n→∞Xk
Thus the functional ΛA ◦ Φ is sequentially weak-∗ continuous, where the weak-∗
continuous functional ΛA : L(H) → C is defined by ΛA(C) = tr(CA). By the
Krein -- Smulian Theorem ΛA ◦ Φ is weak-∗ continuous; see Corollary V.12.8 in
[C]. Therefore, Φ is also weak-∗ continuous; see Proposition 1.3.2 in [KR].
Setting any f, g ∈ H ∞, we know that Φ(σf,nσg,k) = Φ(σf,n)Φ(σg,k) holds
for every n, k ∈ N. Since w∗-limn σf,n = f , we infer by the sequentially weak-
∗ -- wot continuity of Φ that Φ(f σg,k) = Φ(f )Φ(σg,k). Tending now k to infinity,
we conclude that Φ(f g) = Φ(f )Φ(g).
(cid:3)
Theorem 23 The operator T ∈ L(H) admits an H ∞-functional calculus, if
and only if T is an absolutely continuous polynomially bounded operator.
Proof. Let us assume that T admits an H ∞-functional calculus. Then T is
necessarily polynomially bounded by Proposition 21. Suppose that T is not a.c.,
and consider the Lebesgue decomposition T = T a + T s ∈ L(H = Ha + Hs),
where Hs 6= {0}.
In view of Proposition 21, T s admits an H ∞-functional
calculus. We know by Theorem 17 that T s is similar to a singular unitary
operator U ∈ L(K), let Q ∈ I(T, U ) be invertible. It is clear that ΦU,2(f ) :=
QΦT,2(f )Q−1 (f ∈ H ∞) defines an H ∞-calculus for U . Let E denote the
spectral measure of U . Taking a non-zero vector x ∈ K, let us consider the
20
positive measure Ex,x. We know that Ex,x is singular and Ex,x(T) = kxk2 > 0.
In view of regularity we can find a compact set K ⊂ T such that m(K) = 0 and
Ex,x(K) > 0. By a result of Rudin there exists a function u ∈ A(T) such that
u(ζ) = 1 for all ζ ∈ K, and u(ζ) < 1 for all ζ ∈ T \ K; see page 81 in [Ho].
Since w∗-limn un = 0, regarding the continuity properties of ΦU,2 it follows that
limnhun(U )x, xi = 0. However,
lim
n→∞
hun(U )x, xi = lim
n→∞ZT
un dEx,x = Ex,x(K) > 0.
We arrived at a contradiction, and so the polynomially bounded operator T
must be absolutely continuous.
Let us assume now that T is an a.c. polynomially bounded operator. Let
f ∈ H ∞. Given any x, y ∈ H, let us choose an elementary measure µx,y ∈
M (T, x, y). Being abolutely continuous, µx,y = gx,y dm where gx,y ∈ L1(T);
M (T, x, y) is of the form ν = µx,y + h dm with h ∈ H 1
hence the integral RT f dµx,y = RT f gx,y dm can be formed. Any other ν ∈
0 , and so RT f dν =
RT f dµx,y +RT f h dm = RT f dµx,y, because f h ∈ H 1
0 . Let us consider the
well-defined mapping
wf : H × H → C, wf (x, y) =ZT
f dµx,y.
It is easy to check that wf is sesquilinear, and kwf k ≤ kf k · kΦT,1k. There exists
a unique operator Φ(f ) ∈ L(H) such that
hΦ(f )x, yi = wf (x, y) =ZT
f dµx,y
(x, y ∈ H).
It is clear that Φ : H ∞ → L(H) is linear, and ΦA(T) = ΦT,1 is a representation.
If {fn}∞
n=1 is a sequence in H ∞ converging to 0 in the weak-∗ topology, then
lim
n→∞
hΦ(fn)x, yi = lim
n→∞ZT
fngx,y dm = 0 (x, y ∈ H).
Applying Lemma 22 we obtain that Φ is an H ∞-functional calculus for T .
(cid:3)
Proposition 24 If T ∈ L(H) admits an H ∞-functional calculus, then so does
Proof. We have seen that T is an a.c. polynomially bounded operator if and
only if so is T ∗; see Proposition 14. Furthermore, for any polynomial p ∈ P(T)
its adjoint T ∗. Furthermore, f (T ∗) = ef (T )∗ holds for every f ∈ H ∞.
we have p(T ∗) = ep(T )∗. Let f ∈ H ∞ be arbitrary. Since w∗-limn σf,n = f
and w∗-limneσf,n = ef , it follows that {σf,n(T ∗)}n converges to f (T ∗), and
{eσf,n(T )}n converges to ef (T ) in the weak operator topology. The latter condi-
tion implies that {eσf,n(T )∗}n converges to ef (T )∗ in wot. Therefore, the equal-
ities σf,n(T ∗) =eσf,n(T )∗ (n ∈ N) yield that f (T )∗ = ef (T )∗.
(cid:3)
21
6 Quasianalytic operators
Let T ∈ L(H) be an a.c. polynomially bounded operator. Let (X, U ) be a
unitary asymptote of T , where U ∈ L(K). Taking the Lebesgue decomposition
U = U a ⊕ U s ∈ L(K = Ka ⊕ Ks), and considering the transformation Y ∈
I(T, U s), defined by Y h = P sXh, we infer by Proposition 15 that Y = 0.
Hence XH ⊂ Ka, and it follows by the minimality of (X, U ) that Ks = {0}.
Therefore, U is an a.c. unitary operator.
Let E : BT → P(K) denote the spectral measure of U . For any x, y ∈ H, the
localization of E at Xx, Xy is of the form EXx,Xy = wx,y dm, where wx,y ∈
L1(T) is the asymptotic density function of T at x, y. We note that
EXx,Xy ∈ M (T, x, Ay)
with A = X ∗X (see the beginning of Section 4), and EXx,Xx is the unique
positive measure in M (T, x, Ax).
The measurable set
ω(T, x) := {ζ ∈ T : wx,x(ζ) > 0}
is called the local residual set of T at x. Considering the functional model of
the unitary operator U , we may easily check that the hyperinvariant subspace
of U generated by Xx is just the spectral subspace corresponding to ω(T, x):
E(ω(T, x))K = ∨{F Xx : F ∈ {U }′} ∈ Hlat U.
More precisely, ω(T, x) is the smallest measurable set on T such that E(ω(T, x))K
contains Xx, that is Xx ∈ E(α)K implies m(ω(T, x) \ α) = 0; see, e.g., [C].
Proposition 25 The local residual set ω(T, x) is independent of the particular
choice of the unitary asymptote (X, U ).
Proof. Let (X ′, U ′) be another unitary asymptote of T ; E′ is the spectral
measure of U ′ ∈ L(K′), and ω′(T, x) is the local residual set of T at x, defined
via E′. We know that there is an invertible transformation Z ∈ I(U, U ′) such
that X ′ = ZX. Let us consider the polar decomposition Z = W Z, where
Z ≥ 0 is invertible and W is unitary. It is easy to verify that Z ∈ {U }′ and
W U = U ′W . Since the vector X ′x = ZXx belongs to the subspace
ZE(ω(T, x))K = W ZE(ω(T, x))K = W E(ω(T, x))ZK = E′(ω(T, x))K′,
it follows that m(ω′(T, x) \ ω(T, x)) = 0. Changing the roles of (X, U ) and
(X ′, U ′) we obtain that the symmetric difference ω(T, x) △ ω′(T, x) is of zero
Lebesgue measure, and so they can be considered identical. Note that the
Radon -- Nikodym derivative wx,x, and so ω(T, x) also are determined up to sets
of zero Lebesgue measure.
(cid:3)
22
Let us consider the functional calculus ΦT,2 : H ∞ → L(H), f 7→ f (T ) for T .
a
We write KT = kΦT,2k for short. We recall that for f, g ∈ H ∞, f
≺ g means
that f (z) ≤ g(z) for all z ∈ D := {ζ ∈ C : ζ < 1}. Then f = gh, where
h ∈ H ∞ and khk ≤ 1. Thus kf (T )xk = kh(T )g(T )xk ≤ KT kg(T )xk (x ∈ H),
a
≺ KT g(T ). (If T is a contraction, then KT = 1 and so ΦT,2 is
that is f (T )
a
monotone.) Let F = {fn}∞
≺ fn
for every n. The measurable limit function ϕF (ζ) = limn fn(ζ) is defined for
almost every ζ ∈ T. Set
n=1 be a decreasing sequence in H ∞, i.e. fn+1
NF := {ζ ∈ T : ϕF (ζ) > 0}.
We say that F is asymptotically non-vanishing on the measurable set α ⊂ T, if
m(α ∩ NF ) > 0.
Proposition 26 If inf{kfn(T )xk : n ∈ N} = 0, then limn→∞ kfn(T )xk = 0.
Furthermore,
H0(T, F ) := {x ∈ H :
lim
n→∞
kfn(T )xk = 0}
is a hyperinvariant subspace of T .
Proof. The first statement follows from the condition fn+1(T )
N), while the second one is a consequence of sup{kfn(T )k : n ∈ N} < ∞.
a
≺ KT fn(T ) (n ∈
(cid:3)
The following theorem shows that the local residual set is responsible for
local stability.
Theorem 27
(a) If x ∈ H0(T, F ), then m(Nf ∩ ω(T, x)) = 0.
(b) If m(NF ∩ ω(T, x)) = 0, then there exists a strictly increasing mapping
τ : N → N such that x ∈ H0(T, G), where G = {χτ (n)fn}∞
n=1.
Proof. For any x ∈ H, we have
kXfn(T )xk2 = kfn(U )Xxk2 =ZT
and the latter integral converges to RT ϕ2
F wx,x dm by Lebesgue's Dominated
Convergence Theorem, as n → ∞. Hence (a) follows, and we obtain (b) also by
applying Theorem 5. Note that ϕG = ϕF .
fn2wx,x dm (n ∈ N),
The previous theorem tells us that if the decreasing sequence F = {fn}∞
n=1
in H ∞ is asymptotically non-vanishing on ω(T, x), then x 6∈ H0(T, F ), that is
the vector-sequence {fn(T )x}∞
n=1 is asymptotically non-vanishing. On the other
hand, if F is asymptotically vanishing on ω(T, x), then a modified (strengthened)
sequence {T τ (n)fn(T )x}∞
n=1 is asymptotically vanishing (stable) with a suitable
mapping τ .
(cid:3)
23
It is known that the measurable sets on T form a complete lattice, if we
disregard of sets of measure zero. Hence, for any measurable sets α, β on T,
we shall write α ⊂ β, α = β or α 6= β, when m(α \ β) = 0, m(α △ β) = 0 or
m(α △ β) > 0, respectively. The following statement can be proved using the
Gelfand transform of the abelian Banach algebra L∞(T); see Section 11.13 in
[R]. Here we sketch an elementary proof.
Lemma 28 For any system of measurable sets {ωi}i∈I on T, there exist a
smallest measurable set ω = ∨i∈I ωi containing all ωi, and a largest measur-
able set ω = ∧i∈I ωi contained in all ωi. Furthermore, ω and ω are uniquely
determined.
Proof. Let Ω be the system of those measurable sets, which contain (in the
extended sense) every ωi (i ∈ I). It is clear that Ω is closed under countable
n=1 in
intersection. Set a = inf{m(ω) : ω ∈ Ω}, and select a sequence {bωn}∞
Ω so that limn m(bωn) = a. Then it is easy to verify that the measurable set
n=1bωn has the required properties. The statement about ω can be proved
ω = ∩∞
similarly, considering the system Ω of those measurable sets, which are contained
in every ωi (i ∈ I), and then taking b = sup{m(ω) : ω ∈ Ω}.
(cid:3)
The set
ω(T ) := ∨{ω(T, x) : x ∈ H}
is called the residual set of T . It is clear that ω(T ) is the measurable support of
the spectral measure E of U , that is E(α) = 0 exactly when m(α ∩ ω(T )) = 0.
Furthermore, ω(T ) = ∅ means that K = {0}, that is H0(T ) = H, in which case
T is called of class C0·.
The set
π(T ) := ∧{ω(T, x) : 0 6= x ∈ H}
In view of Theorem 27, this is
is called the quasianalytic spectral set of T .
the largest measurable set with the property that H0(T, F ) = {0}, whenever
n=1 is a decreasing sequence in H ∞, asymptotically non-vanishing on
F = {fn}∞
π(T ). We can characterize π(T ) also with more general sequences, as in the next
proposition. (For related results in connection with test sequences of stability,
see [KSz1].)
Proposition 29 If F = {fn}∞
lim supn kχπ(T )fnk2 > 0, then the hyperinvariant subspace
n=1 is a bounded sequence in H ∞ such that
H0(T, F ) := {x ∈ H :
lim
n→∞
kfn(T )xk = 0}
reduces to zero. Furthermore, π(T ) is the largest measurable set having this
property.
Proof. Let F = {fn}∞
lim supnRπ(T ) fn2 dm > 0. Set M = sup{kfnk∞ : n ∈ N} < ∞. Further-
more, given 0 6= x ∈ H, let α(x, N ) := {ζ ∈ π(T ) : wx,x(ζ) > N −1} for N ∈ N.
n=1 be a bounded sequence in H ∞ such that c =
24
We can select N so that M 2m(π(T ) \ α(x, N )) < c/2. For every n ∈ N, we have
ZT
Hence
fn2wx,x dm ≥ N −1Zα(x,N )
= N −1 Zπ(T )
≥ N −1(cid:0)kχπ(T )fnk2
n→∞ ZT
kXfn(T )xk2 = lim sup
lim sup
n→∞
fn2 dm
fn2 dm!
fn2 dm −Zπ(T )\α(x,N )
2 − M 2m(π(T ) \ α(x, N ))(cid:1)
fn2wx,x dm ≥ N −1c/2 > 0,
and so x is not stable for F . Therefore H0(T, F ) = {0}.
Let β be a measurable set on T, larger than π(T ) : β ⊃ π(T ) and β 6= π(T ).
By the definition of π(T ) we can find a non-zero x0 ∈ H so that β is not contained
in ω(T, x0), that is ∆ = β \ ω(T, x0) is of positive measure. Let f0 ∈ H ∞ be
an outer function satisfying the condition f0 = χ∆ + (1/2)χT\∆, and let us
consider the decreasing sequence F0 = {f n
In
view of Theorem 27, there exists a strictly increasing τ : N → N such that
x0 ∈ H0(T, F ) with F = {χτ (n)f n
n=1 in H ∞, with ϕF0 = χ∆.
n=1. It is clear also that
0 }∞
0 }∞
lim
n→∞
which completes the proof.
kχβχτ (n)f n
0 k2
2 = m(∆) > 0,
(cid:3)
The absolutely continuous polynomially bounded operator T ∈ L(H) is
quasianalytic, if π(T ) = ω(T ) 6= ∅. We note that π(T ) ⊂ ω(T ) is evident
from the definition. Theorem 3 in [KSz2] shows that our definition is consistent
with the one given in the contractive case.
The following theorem illuminates the importance of quasianalytic operators,
showing that the challenging Hyperinvariant Subspace Problem can be reduced
to this class in the asymptotically non-vanishing case.
Theorem 30 Let T ∈ L(H) be an absolutely continuous polynomially bounded
operator, which is asymptotically non-vanishing: H0(T ) 6= H. If T is not quasi-
analytic, then it has a non-trivial hyperinvariant subspace.
Proof. By our assumption we can find non-zero vectors x1, x2 ∈ H so that
ω(T, x2) is not contained in ω(T, x1), that is the set ∆ = ω(T, x2) \ ω(T, x1) is
of positive measure. There exists a decreasing sequence F = {fn}∞
n=1 in H ∞
such that x1 ∈ H0(T, F ) and NF = ∆; see the second part of the proof of
Proposition 29. The latter relation implies by Theorem 27 that x2 6∈ H0(T, F ).
Therefore, H0(T, F ) is a proper hyperinvariant subspace of T .
(cid:3)
The preceding proof immediately yields the following statement.
25
Proposition 31 Let T1 ∈ L(H1) and T2 ∈ L(H2) be a.c. polynomially bounded
operators. If ω(T2, x2) is not contained in ω(T1, x1), then there exists a decreas-
ing sequence F in H ∞ such that x1 ∈ H0(T1, F ) and x2 6∈ H0(T2, F ).
We proceed with some intertwining relations.
Proposition 32 Let T1 ∈ L(H1) and T2 ∈ L(H2) be a.c. polynomially bounded
operators.
(a) If Y ∈ I(T1, T2), then ω(T1, x) ⊃ ω(T2, Y x) holds for every x ∈ H1.
(b) If I(T1, T2) contains an injection Y , then π(T1) ⊃ π(T2).
(c) If both I(T1, T2) and I(T2, T1) contain injections, then π(T1) = π(T2).
(d) If I(T1, T2) contains a transformation Z with dense range, then ω(T1) ⊃
ω(T2).
(e) If T1 ∼ T2, then π(T1) = π(T2) and ω(T1) = ω(T2).
(f) If T1 ∼ T2 and T1 is quasianalytic, then T2 is also quasianalytic.
Proof. The equations Y fn(T1) = fn(T2)Y (n ∈ N) show that Y H0(T1, F ) ⊂
n=1 in H ∞. In view of
H0(T2, F ) holds, for any decreasing sequence F = {fn}∞
Proposition 31 we infer that ω(T1, x) ⊃ ω(T2, Y x) for every x ∈ H1. Assuming
that Y ∈ I(T1, T2) is injective, this relation yields
π(T2) = ∧{ω(T2, y) : 0 6= y ∈ H2} ⊂ ∧{ω(T1, x) : 0 6= x ∈ H1} = π(T1).
Let us assume now that Z ∈ I(T1, T2) has dense range. Let (Xj, Uj) be a
unitary asymptote of Tj (j = 1, 2). Since X2Z ∈ I(T1, U2), there exists a
1 X1H1)− = ∨n∈NU −n
2 X2H2 = K2.
unique eZ ∈ I(U1, U2) such that X2Z = eZX1. It follows that
(eZK1)− = (eZ ∨n∈N U −n
Considering the polar decomposition eZ = W eZ we obtain that W ∈ I(U1, U2)
is a coisometry, and so U2 is unitarily equivalent to the restriction of U1 to its
reducing subspace (ker W )⊥. Therefore, the measurable support ω(T1) of the
spectral measure of U1 contains the measurable support ω(T2) of the spectral
measure of U2.
2 eZX1H1 = ∨n∈NU −n
Quasianalicity determines the asymptotic behaviour of the operator.
Proposition 33 If the a.c. polynomially bounded operator T ∈ L(H) is quasi-
analytic, then T is of class C10 : H0(T ) = {0} and H0(T ∗) = H.
(cid:3)
26
Proof. Taking the decreasing sequence F0 = {χn}∞
n=1 in H ∞, we infer that
H0(T ) = H0(T, F0) = {0} since NF0 ∩ π(T ) = π(T ) 6= ∅; see Proposition 29.
Assuming that H0(T ∗) 6= H, let us consider a unitary asymptote (X ∗
∗ ) of
the adjoint T ∗. Then the a.c. unitary operator U∗ acts on a non-zero space K∗,
and X∗ ∈ I(U∗, T ) is a non-zero transformation. The equations X∗U∗ = T X∗
and X∗ = T X∗U ∗
∗ imply that the subspace ker X∗ is reducing for U∗, because of
ker T ⊂ H0(T ) = {0}. Let us consider the restriction U∗,0 of U∗ to the non-zero
reducing subspace K∗ ⊖ ker X∗. Since I(U∗,0, T ) contains an injection, we infer
by Proposition 32 that π(U∗,0) ⊃ π(T ). However, it is clear that π(U∗,0) = ∅
holds for the a.c. unitary operator U∗,0, and so π(T ) = ∅, what is a contradiction.
(cid:3)
∗ , U ∗
Now we examine the effect of taking orthogonal sums.
Proposition 34 Let Tj ∈ L(Hj ) be a.c. polynomially bounded operator acting
on non-zero space, for j = 1, 2, and let us form T = T1 ⊕ T2 ∈ L(H = H1 ⊕ H2).
(a) Then T is also an a.c. polynomially bounded operator with
π(T ) = π(T1) ∩ π(T2)
and
ω(T ) = ω(T1) ∪ ω(T2).
(b) T is quasianalytic if and only if T1 and T2 are quasianalytic and π(T1) =
π(T2).
Proof. Setting x = x1 ⊕ x2, y = y1 ⊕ y2 ∈ H, and µ1 ∈ M (T1, x1, y1), µ2 ∈
M (T2, x2, y2), we have
hu(T )x, yi = hu(T1)x1, y1i + hu(T2)x2, y2i =ZT
u d(µ1 + µ2)
for every u ∈ A(T); hence µ1 + µ2 ∈ M (T, x, y). Therefore T is an a.c. polyno-
mially bounded operator.
Let (Xj, Uj) be a unitary asymptote of Tj, for j = 1, 2. We know that (X, U )
will be a unitary asymptote of T , where X = X1 ⊕ X2 and U = U1 ⊕ U2; see
Corollary 6. Furthermore, if E is the spectral measure of U , then EKj will be
the spectral measure of Uj (j = 1, 2). Thus, for any x = x1 ⊕ x2 ∈ H we have
EXx,Xx = EX1x1,X1x1 + EX2x2,X2x2, whence ω(T, x) = ω(T1, x1) ∪ ω(T2, x2) can
be derived. From here the statement follows.
(cid:3)
We conclude with the properties of invariant subspaces.
Proposition 35 Let T ∈ L(H) be an a.c. polynomially bounded operator, and
let us assume that M ∈ Lat T is non-zero.
(a) Then T M is also an a.c polynomially bounded operator with
π(T ) ⊂ π(T M) ⊂ ω(T M) ⊂ ω(T ).
(b) If T is quasianalytic, then T M is also quasianalytic and π(T M) = π(T ).
27
Proof. It is clear that M (T M, x, y) = M (T, x, y) holds for every x, y ∈ M.
Hence, T M is an a.c. polynomially bounded operator. We know that if (X, U )
is a unitary asymptote of T , then (XM, U fM) is a unitary asymptote of T M;
see Corollary 7. Thus, ω(T M, x) = ω(T, x) holds, for every x ∈ M.
(cid:3)
We recall that if T ∈ L(H) is an a.c. contraction with ω(T ) = T, then
there exists M ∈ Lat T such that T M is similar to the simple unilateral shift
S ∈ L(H 2), Sf = χf ; even more, these shift-type invariant subspaces span the
whole space H (see Theorem IX.3.6 in [NFBK] and [K7]).
Question 36 Let T ∈ L(H) be an a.c. polynomially bounded operator with
ω(T ) = T (or π(T ) = T). Does there exist a non-zero M ∈ Lat T such that
T M is similar to a contraction?
If π(T ) = T, then such an M clearly contains a subspace M′ ∈ Lat T , where
T M′ is similar to S. It would be interesting to know whether Pisier's construc-
tion provides completely non-contractive (i.e., having no M ∈ Lat T \ {0}
such that T M is similar to a contraction) absolutely continuous polynomially
bounded operators in the case, when ω(T ) = T or π(T ) = T.
References
[AT]
[AM]
[B]
[BK]
[BP]
[CP]
[CC]
[C]
T. Ando and K. Takahashi, On operators with unitary ρ-dilations,
Ann. Polon. Math., 66 (1997), 11 -- 14.
S. A. Argyros and P. Motakis, A reflexive hereditarily indecom-
posable space with the hereditary invariant subspace property, Proc.
London Math. Soc. (3), 108 (2014), 1381 -- 1416.
H. Bercovici, Operator theory and arithmetic in H ∞, Mathematical
Surveys and Monographs 26, Amer. Math. Soc., Providence, Rhode
Island, 1988.
H. Bercovici and L. K´erchy, Spectral behaviour of C10-
contractions, Operator Theory Live, Theta, Bucharest, 2010, 17 -- 33.
H. Bercovici and B. Prunaru, Quasiaffine transforms of polynomi-
ally bounded operators, Arch. Math. (Basel ), 71 (1998), 384 -- 387.
I. Chalendar and J. R. Partington, Modern approaches to the
invariant-subspace problem, Cambridge Tracts in Mathematics 188,
Cambridge University Press, Cambridge, 2011.
B. Chevreau and A. Craciunescu, On a generalization of ρ-
contractions, Acta Sci. Math. (Szeged ), to appear.
J. B. Conway, A course in functional analysis, Springer-Verlag, New
York, 1990.
28
[DH]
[E]
[F]
[G1]
[G2]
[GR]
[Ha]
[Ho]
[KR]
[K1]
[K2]
[K3]
[K4]
[K5]
[K6]
[K7]
K. R. Davidson and D. A. Herrero, The Jordan form of a bitrian-
gular operator, J. Funct. Anal., 94 (1990), 27 -- 73.
G. Eckstein, Sur les op´erateurs de class Cρ, Acta Sci. Math. (Szeged ),
33 (1972), 349 -- 352.
S. Foguel, A counterexample to a problem of Sz.-Nagy, Proc. Amer.
Math. Soc., 15 (1964), 788 -- 790.
M. F. Gamal', On power bounded operators that are quasiaffine
transforms of singular unitaries, Acta Sci. Math. (Szeged ), 77 (2011),
589 -- 606.
M. F. Gamal', A class of power bounded operators which are similar
to contractions, Acta Sci. Math. (Szeged ), 80 (2014), 625 -- 637.
S. Grivaux and M. Roginskaya, A general approach to Read's type
constructions of operators without non-trivial invariant closed sub-
spaces, Proc. London Math. Soc. (3), 109 (2014), 596 -- 652.
P. R. Halmos, Introduction to Hilbert space and the theory of spectral
multiplicity, Chelsea Publishing Company, New York, 1951.
K. Hoffman, Banach spaces of analytic functions, Dover Publica-
tions, Inc., New York, 1988.
R. V. Kadison and J. R. Ringrose, Fundamentals of the theory of
operator algebras, Volume I: Elementary theory, Graduate Studies in
Mathematics 15, Amer. Math. Soc., Providence, Rhode Island, 1997.
L. K´erchy, Isometric asymptotes of power bounded operators, Indi-
ana Univ. Math. J., 38 (1989), 173 -- 188.
L. K´erchy, Operators with regular norm-sequences, Acta Sci. Math.
(Szeged ), 63 (1997), 571 -- 605.
L. K´erchy, Representations with regular norm-behaviour of discrete
abelian semigroups, Acta Sci. Math. (Szeged ), 65 (1999), 701 -- 726.
L. K´erchy, Hyperinvariant subspaces of operators with non-vanishing
orbits, Proc. Amer. Math. Soc., 127 (1999), 1363 -- 1370.
L. K´erchy, On the hyperinvariant subspace problem for asymptot-
ically nonvanishing contractions, Operator Theory Adv. Appl., 127
(2001), 399 -- 422.
L. K´erchy, Generalized Toeplitz operators, Acta Sci. Math. (Szeged ),
68 (2002), 373 -- 400.
L. K´erchy, Shift-type invariant subspaces of contractions, J. Funct.
Anal., 246 (2007), 281 -- 301.
29
[K8]
[K9]
[KL]
[KM]
[KT]
L. K´erchy, Quasianalytic contractions and function algebras, Indiana
Univ. Math. J., 60 (2011), 21 -- 40.
L. K´erchy, Unitary asymptotes and quasianalicity, Acta Sci. Math.
(Szeged ), 79 (2013), 253 -- 271.
L. K´erchy and Z. L´eka, Representations with regular norm-
behaviour of locally compact abelian semigroups, Studia Mathematica,
183 (2007), 143 -- 160.
L. K´erchy and V. Muller, Criteria of regularity for norm-sequences.
II, Acta Sci. Math. (Szeged ), 65 (1999), 131 -- 138.
L. K´erchy and V. Totik, Compression of quasianalytic spectral sets
of cyclic contractions, J. Funct. Anal., 263 (2012), 2754 -- 2769.
[KSz1] L. K´erchy and A. Szalai, Characterization of stability of contrac-
tions, Acta Sci. Math. (Szeged ), 79 (2013), 325 -- 332.
[KSz2] L. K´erchy and A. Szalai, Asymptotically cyclic quasianalytic con-
tractions, Studia Mathematica, 223 (2014), 53 -- 75.
[KSz3] L. K´erchy and A. Szalai, Spectral behaviour of quasianalytic con-
tractions, Proc. Amer. Math. Soc., to appear.
[M1] W. Mlak, Decompositions and extensions of operator valued repre-
sentations of function algebras, Acta Sci. Math. (Szeged ), 30 (1969),
181 -- 193.
[M2] W. Mlak, Decompositions of polynomially bounded operators, Bull.
Acad. Polon. Sci. S´er. Sci. Math. Astronom. Phys., 21 (1973), 317 --
322.
[M3] W. Mlak, Operator valued representations of function algebras, Lin-
ear operators and approximation, II (Proc. Conf., Oberwolfach Math.
Res. Inst., Oberwolfach, 1974), pp. 49 -- 79, Internat. Ser. Numer. Math.
25, Birkhauser, Basel, 1974.
[M4] W. Mlak, Algebraic polynomially bounded operators, Ann. Polon.
Math., 29 (1974), 133 -- 139.
[MT]
V. Muller and Y. Tomilov, Quasisimilarity of power bounded op-
erators and Blum -- Hanson property, J. Funct. Anal., 246 (2007), 385 --
399.
[N]
B. Sz.-Nagy, On uniformly bounded linear transformations in Hilbert
space, Acta Sci. Math. (Szeged ), 11 (1947), 152 -- 157.
[NFBK] B. Sz.-Nagy, C. Foias, H. Bercovici and L. K´erchy, Harmonic
analysis of operators on Hilbert space, Revised and enlarged edition,
Universitext, Springer, New York, 2010.
30
[Pa]
[Pi]
[Pr]
[RR]
[R]
[Sch]
V. Paulsen, Completely bounded maps and operator algebras, Cam-
bridge Studies in Advanced Mathematics 78, Cambridge University
Press, Cambridge, 2002.
G. Pisier, Similarity problems and completely bounded maps; Sec-
ond, expanded edition, Lecture Notes in Mathematics 1618, Springer-
Verlag, Berlin, 2001.
B. Prunaru, Toeplitz operators associated to commuting row con-
tractions, J. Funct. Anal., 254 (2008), 1626 -- 1641.
H. Radjavi and P. Rosenthal, Invariant subspaces, Second edition,
Dover Publications, INC., Mineola, New York, 2003.
W. Rudin, Functional analysis, McGraw-Hill Book Company, New
York, 1973.
R. Schatten, Norm ideals of completely continuous operators, Ergeb-
nisse der Mathematik und ihrer Grenzgebiete 27, Springer-Verlag,
Berlin, 1970.
L. K´erchy, Bolyai Institute, University of Szeged, Aradi v´ertan´uk tere 1, H-
6720 Szeged, Hungary; e-mail : [email protected]
31
|
1811.08206 | 1 | 1811 | 2018-11-20T12:27:51 | Continuous projections onto ideal convergent sequences | [
"math.FA",
"math.GN"
] | Let $\mathcal{I}\subseteq\mathcal{P}(\omega)$ be a meager ideal. Then there are no continuous projections from $\ell_\infty$ onto the set of bounded sequences which are $\mathcal{I}$-convergent to $0$. In particular, it follows that the set of bounded sequences statistically convergent to $0$ is not isomorphic to $\ell_\infty$. | math.FA | math |
CONTINUOUS PROJECTIONS ONTO IDEAL
CONVERGENT SEQUENCES
PAOLO LEONETTI
Abstract. Let I ⊆ P(ω) be a meager ideal. Then there are no continuous
projections from ℓ∞ onto the set of bounded sequences which are I-convergent
to 0. In particular, it follows that the set of bounded sequences statistically
convergent to 0 is not isomorphic to ℓ∞.
1. Introduction
A closed subspace X of a Banach space B is said to be complemented in B
if there exists a continuous projection from B onto X. It is known that c0, the
space of real sequences convergent to 0, is not complemented in ℓ∞, cf. [10, 12].
The aim of this note is to show the ideal analogue of this result.
Let I ⊆ P(ω) be an ideal, that is, a family closed under subsets and finite
unions. It is also assumed that Fin := [ω]<ω ⊆ I and ω /∈ I. Set I + := P(ω) \ I.
In particular, each I can be regarded as a subset of the Cantor space 2ω with the
product topology, so we can speak of Borel ideals, Fσ ideals, etc. An ideal I is
said to be a P-ideal if it is σ-directed modulo finite sets, i.e., for each sequence
(An) in I there exists A ∈ I such that An \ A is finite for all n ∈ ω. We refer to
[7] for a recent survey on ideals and filters.
A real sequence (xn) is said to be I-convergent to y if {n : xn /∈ U} ∈ I for all
neighborhoods U of y. We denote by c(I) [resp. c0(I)] the space of real sequences
which are I-convergent [resp. I-convergent to 0]. The set of bounded real I-
convergent sequences has been studied, e.g., in [2, 6, 8]. By an easy modification
of [8, Theorem 2.3], c0(I) ∩ ℓ∞ is a closed linear subspace of ℓ∞ (with the sup
norm).
The question addressed here, posed at the open problem session of the 45th
Winter School in Abstract Analysis (Czech Republic, 2017), follows:
Question 1. Is c0(I) ∩ ℓ∞ complemented in ℓ∞?
Before proving our main result, we recall the following:
2010 Mathematics Subject Classification. Primary: 40A35, 46B03.
Secondary: 54A20,
46B26.
Key words and phrases. Meager ideal, I-maximal almost disjoint family, complementability,
asymptotic density zero sets, I-convergent sequence.
2
Paolo Leonetti
Lemma 1.1. An infinite dimensional subspace X of ℓ∞ is complemented in ℓ∞
if and only if it is isomorphic to ℓ∞.
Proof. See [1, Proposition 2.5.2 and Theorem 5.6.5].
(cid:3)
Hence, Question 1 can be reformulated as:
Question 2. Is c0(I) ∩ ℓ∞ isomorphic to ℓ∞?
We will prove that the answer is negative for a large class of ideals. To state our
result, we recall that a family A ⊆ I + is said to be I-maximal-almost-disjoint
(in short, I-mad) if A is a maximal family (with respect to inclusion) such that
A ∩ B ∈ I for all distinct A, B ∈ A , so that for each X ∈ I + there exists
A ∈ A such that X ∩ A ∈ I +. (The minimal cardinality a(I) of an I-mad has
been studied in the literature: e.g., it is known that, if I is an analytic P-ideal,
a(I) > ω if and only if I is Fσ, cf. [4, 5].)
Our main result follows:
Theorem 1.2. Let I be an ideal for which there exists an uncountable I-mad
family. Then c0(I) ∩ ℓ∞ is not complemented in ℓ∞.
It can be shown that, if I is a meager ideal, there is an I-mad family of
cardinality c, see Lemma 2.3 below. In particular
Corollary 1.3. c0(I) ∩ ℓ∞ is not complemented in (and not isomorphic to) ℓ∞
whenever I is meager.
As an important example, the family of asymptotic density zero sets Z := {S ⊆
ω : S ∩ [1, n]/n → 0} is an analytic P-ideal, hence meager. Therefore:
Corollary 1.4. The set of bounded real sequences statistically convergent to 0
(i.e., c0(Z)) is not is isomorphic to ℓ∞.
Lastly, we obtain an analogue of the main result in [9] (for summability matri-
ces):
Corollary 1.5. c is complemented in c(I) ∩ ℓ∞ if and only if I = Fin.
It is worth noting that Theorem 1.2 cannot be extended to all ideals I. Indeed, if
I is maximal, then the set of bounded I-convergent sequences, which is isomorphic
to c0(I) ∩ ℓ∞, is exactly ℓ∞.
2. Preliminaries and Proofs
Thanks to Lemma 1.1, a negative question to Question 1 would follow if c0(I) ∩
ℓ∞ was separable (indeed ℓ∞ is nonseparable, hence they cannot be isomorphic).
However, this works only if I = Fin:
Lemma 2.1. c0(I) is separable if and only if I = Fin.
Projections onto ideal convergent sequences
3
Proof. The if part is known. Conversely, let us suppose that there exists A ∈
I ∩ [ω]ω. For each X ⊆ ω and ε > 0, let B(1X, ε) be the open ball with center 1X
and radious ε. The collection B := {B(1X, 1/2) : X ∈ [A]ω} is an uncountable
family of nonempty open sets which are pairwise disjoint, hence c0(I) is not
separable.
(cid:3)
At this point, recall the following characterization, see [11] and [3, Theorem
4.1.2]:
Lemma 2.2. I is a meager ideal if and only if there exists a finite-to-one function
f : ω → ω such that f −1(A) ∈ I if and only if A is finite.
In other words, the second condition is Fin ≤RB I, where ≤RB is the Rudin --
Blass ordering. This is sufficient to prove the existence of an uncountable I-mad
family:
Lemma 2.3. There exists an I-mad family of cardinality c, provided I is meager.
[12].
Proof. It is known that there is a Fin-mad family A of cardinality c, cf.
Then, thanks to Lemma 2.2, there exists a finite-to-one function f : ω → ω such
that f −1(A) ∈ I if and only if A is finite, hence {f −1(A) : A ∈ A } is the claimed
I-mad family.
(cid:3)
Let us prove our main result:
Proof of Theorem 1.2. Let us suppose for the sake of contradiction that c0(I)∩ℓ∞
is complemented in ℓ∞ and denote by
π : ℓ∞ → c0(I) ∩ ℓ∞
the canonical projection. Define T := I − π, hence T is bounded linear operator
such that T (x) = 0 for each x ∈ c0(I) ∩ ℓ∞. Note also that, if B /∈ I, then 1B
is a bounded sequence which is not I-convergent to 0, hence π(1B) 6= 1B and
T (1B) 6= 0.
At this point, let (Aj : j ∈ J) be an uncountable I-mad family, which exists by
hypothesis. We are going to show that there exists j ∈ J such that T (1Aj ) = 0,
which is impossible since Aj ∈ I +. Indeed, let us suppose that, for each j ∈ J,
there exists xj = (xj,n) ∈ ℓ∞ supported on Aj with T (xj) 6= 0 and, without
loss of generality, kxjk∞ = 1.
It follows that there exists m, k ∈ ω such that
J := {j ∈ J : xj,m ≥ 2−k} is uncountable. Also, by possibly replacing xj with
−xj, let us suppose without loss of generality that xj,m > 0 for all j ∈ J.
For each nonempty finite set F ⊆ J, define sF = (sF,n) := Pj∈F xj. In partic-
ular,
kT (sF )k∞ ≥ sF,m ≥ F 2−k.
(1)
Note also that I := S(Ai ∩ Aj), where the sum is extended over all distinct
i, j ∈ F , belongs to I. This implies that the sequence sF ↾ I is I-convergent to 0,
4
Paolo Leonetti
hence T (sF ) = T (sF ↾ I c). Therefore
kT (sF )k∞ = kT (sF ↾ I c)k∞ ≤ kT k · ksF ↾ I ck∞ ≤ kT k,
which, together with (1), implies F ≤ 2kkT k. This contradicts the fact the J is
infinite.
(cid:3)
Proof of Corollary 1.5. There is nothing to prove if I = Fin. Conversely, fix
I ∈ I \ Fin and define X := {x ∈ ℓ∞ : xi 6= 0 only if i ∈ I} and Y := X ∩ c0. It
is clear that
c ⊆ Y ⊆ X ⊆ c(I) ∩ ℓ∞
and that X and Y are isometric to ℓ∞ and c0, respectively. Hence, it is known
that c can be projected continuously onto Y , let us say through T , see [10].
To conclude the proof, let us suppose that there exists a continuous projection
H : c(I) ∩ ℓ∞ → c. Then the restriction T ◦ H ↾ X is a continuous projection
ℓ∞ → c0. This contradicts Theorem 1.2 (in the case I = Fin).
(cid:3)
2.1. Acknowledgments. The author is grateful to Tommaso Russo (Università
degli Studi di Milano, IT) for suggesting Question 1 and Lemma 1.1.
References
1. F. Albiac and N. J. Kalton, Topics in Banach space theory, Graduate Texts in Mathematics,
vol. 233, Springer, New York, 2006.
2. A. Bartoszewicz, S. Głab, and A. Wachowicz, Remarks on ideal boundedness, convergence
and variation of sequences, J. Math. Anal. Appl. 375 (2011), no. 2, 431 -- 435.
3. T. Bartoszyński and H. Judah, Set theory, A K Peters, Ltd., Wellesley, MA, 1995, On the
structure of the real line.
4. J. E. Baumgartner, Iterated forcing, Surveys in set theory, London Math. Soc. Lecture Note
Ser., vol. 87, Cambridge Univ. Press, Cambridge, 1983, pp. 1 -- 59.
5. B. Farkas and L. Soukup, More on cardinal invariants of analytic P -ideals, Comment. Math.
Univ. Carolin. 50 (2009), no. 2, 281 -- 295.
6. R. Filipów and J. Tryba, Ideal convergence versus matrix summability, Studia Math., to
appear.
7. M. Hrušák, Combinatorics of filters and ideals, Set theory and its applications, Contemp.
Math., vol. 533, Amer. Math. Soc., Providence, RI, 2011, pp. 29 -- 69.
8. P. Kostyrko, M. Mačaj, T. Šalát, and M. Sleziak, I -convergence and extremal I -limit
points, Math. Slovaca 55 (2005), no. 4, 443 -- 464.
9. J. Lindenstrauss, Mathematical Notes: A Remark Concerning Projections in Summability
Domains, Amer. Math. Monthly 70 (1963), no. 9, 977 -- 978.
10. A. Sobczyk, Projection of the space (m) on its subspace (c0), Bull. Amer. Math. Soc. 47
(1941), 938 -- 947.
11. M. Talagrand, Compacts de fonctions mesurables et filtres non mesurables, Studia Math. 67
(1980), no. 1, 13 -- 43.
12. R. Whitley, Mathematical Notes: Projecting m onto c0, Amer. Math. Monthly 73 (1966),
no. 3, 285 -- 286.
Projections onto ideal convergent sequences
5
Department of Statistics, Università "L. Bocconi", via Roentgen 1, 20136 Milan,
Italy
E-mail address: [email protected]
URL: https://sites.google.com/site/leonettipaolo/
|
1710.01464 | 1 | 1710 | 2017-10-04T05:25:34 | Positive solution to fractional thermostat model in Banach spaces via fixed point results | [
"math.FA"
] | The motive behind this manuscript is to set up the existence and uniqueness of a positive solution for a fractional thermostat model for certain values of the parameter $\lambda>0$. We accomplish sufficient conditions for the existence of a positive solution to the model, and afterwards formulate a non-trivial example to authenticate the grounds of our obtained results. Our findings are based on certain fixed point results of contractions depending on couple of altering distance functions $\phi$ and $\psi$ in the setting of Banach spaces that are discussed in this sequel. | math.FA | math | POSITIVE SOLUTION TO FRACTIONAL THERMOSTAT MODEL
IN BANACH SPACES VIA FIXED POINT RESULTS
HIRANMOY GARAI1, LAKSHMI KANTA DEY2, ANKUSH CHANDA3
Abstract. The motive behind this manuscript is to set up the existence and unique-
ness of a positive solution for a fractional thermostat model for certain values of the
parameter λ > 0. We accomplish sufficient conditions for the existence of a positive
solution to the model, and afterwards formulate a non-trivial example to authenticate
the grounds of our obtained results. Our findings are based on certain fixed point
results of contractions depending on couple of altering distance functions φ and ψ in
the setting of Banach spaces that are discussed in this sequel.
7
1
0
2
t
c
O
4
]
.
A
F
h
t
a
m
[
1
v
4
6
4
1
0
.
0
1
7
1
:
v
i
X
r
a
1. Introduction and Preliminaries
Metric fixed point theory is extensively employed in different mathematical branches
as well as in real world problems originating in applied sciences. The results on fixed
points of contractive maps considered on different underlying spaces are mostly applied
on the validation of the existence and uniqueness of solutions of functional, differential
or integral equations. The plurality of these types of problems elicits the probe for
more and better techniques, which is a salient feature of the recent research works in
this literature.
The dawning of fixed point theory on a complete metric space is integrated with the
Banach contraction principle due to S. Banach [6].
Theorem 1.1. Let (X, d) be a complete metric space and T be a self-mapping on X
satisfying
d(T x, T y) ≤ kd(x, y)
for all x, y ∈ X and k ∈ [0, 1). Then T has a unique fixed point z ∈ X, and for any
x ∈ X the sequence of iterates (T nx) converges to z.
Because of its inferences and huge usability in mathematical theory, Banach contrac-
tion principle has been improved and generalized in metric spaces, partially ordered
metric spaces, Banach spaces and many other spaces, see [1, 3, 4, 7, 10 -- 13, 16, 20].
In 1962, E. Rakotch [22] proved that the Theorem 1.1 still holds if the constant k
is replaced by a contraction monotone decreasing function. He proved the following
theorem as a corollary.
Theorem 1.2. Let (X, d) be a complete metric space and T : X → X be a mapping
such that
d(T x, T y) ≤ α(x, y)d(x, y)
1991 Mathematics Subject Classification. 47H10, 54H25.
Key words and phrases. Altering distance function, fixed point, Banach space, double sequence,
thermostat model.
1
2
H. GARAI, L.K. DEY, A. CHANDA
for all x, y ∈ X, where α is a function defined on [0,∞) satisfying the following condi-
tions:
i) α(x, y) = α(d(x, y)), i.e., α is dependent on the distance of x and y only;
ii) 0 ≤ α(τ ) < 1 for all τ > 0;
iii) α(τ ) is monotonically decreasing function of τ .
Then T has a unique fixed point.
In his research article, D.S. Jaggi [18] used the continuity and some different con-
tractive conditions on the mapping to attain the succeeding result.
Theorem 1.3. Let f be a continuous self-map defined on a complete metric space
(X, d). Further let, f satisfies the following condition:
d(f (x), f (y)) ≤
αd(x, f (x))d(y, f (y))
d(x, y)
+ βd(x, y)
for all x, y ∈ X, with x 6= y and for some α, β ∈ [0, 1) with α + β < 1. Then f has a
unique fixed point in X.
Thereafter, Khan et al. [19] extended and generalized the Banach principle using a
control function, known as altering distance function.
Definition 1.4. A function ϕ : [0,∞) → [0,∞) is called an altering distance function
if it satisfies the following conditions:
i) ϕ is monotone increasing and continuous;
ii) ϕ(t) = 0 if and only if t = 0.
In [19], the authors also proved the following fixed point theorem by means of the
newly originated concept of control functions.
Theorem 1.5. Let (X, d) be a complete metric space and ψ : [0,∞) → [0,∞). Also
suppose that f : X → X is a mapping satisfying
ψ(d(f x, f y)) ≤ aψ(d(x, y))
for all x, y ∈ X and for some 0 ≤ a < 1. Then f has a unique fixed point.
Alber and Guerre-Delabriere [2] introduced the notion of weak contractions in a
Hilbert space.
Definition 1.6.
weakly contractive if and only if
[2] Let (X, d) be a metric space. A mapping T : X → X is called
for all x, y ∈ X, where φ is an altering distance function.
d(T x, T y) ≤ d(x, y) − φ(d(x, y))
Afterwards, Rhoades [23] generalized the weak contraction condition in metric spaces
and proved the following fixed point result in complete metric spaces.
Theorem 1.7. Let (X, d) be a complete metric space.
contractive map, then T has a unique fixed point.
If T : X → X is a weakly
In their research paper, Dutta and Choudhury [15] generalized Theorem 1.5 and 1.7
to obtain the following theorem.
POSITIVE SOLUTION VIA FIXED POINT RESULTS
3
Theorem 1.8. Let (X, d) be a complete metric space and T : X → X be a mapping
satisfying
for all x, y ∈ X, where ψ and φ are two altering distance functions. Then T has a
unique fixed point.
ψ(d(T x, T y)) ≤ ψ(d(x, y)) − φ(d(x, y))
Fractional calculus has been explored for many decades mostly as a pure analytic
mathematical branch. Though in recent times, many authors are showing a lot of
interest in its applications for solving ordinary differential equations. Fractional dif-
ferential equations appear in different engineering and scientific branches as the math-
ematical modelling of systems and techniques in the domains of physics, chemistry,
aerodynamics, robotics and many more. For a few recent articles in this direction,
see [5, 8, 9, 14, 17, 21, 24] and the references in that respect.
Considering exclusively positive solutions are effective for several applications, in-
spired by the aforementioned works, in our draft, we set up an existence and uniqueness
theorem to find a positive solution for a fractional thermostat model with a positive
parameter. With a view to inspect the solution, we enquire into some new fixed point
results in a Banach space by considering a pair of altering distance functions in a
more adequate appearance. We also extend our results in a Banach space which is
equipped with an arbitrary binary relation and keeps the order preserving property of
the mappings. Finally, a suitable non-trivial example is furnished to substantiate the
effectiveness of our results.
2. Fixed Point Results
This section deals with the results on the existence and uniqueness of fixed points
of maps satisfying a contractive condition with a pair of control functions in a Ba-
nach space and also their proofs. Moreover, we formulate an example to elucidate our
attained results.
Theorem 2.1. Let (X,k.k) be a Banach space and C be a closed subset of X. Let
T : C → C be a mapping. Assume that, there exist two altering distance functions
φ, ψ : [0,∞) → [0,∞) such that
φ(kT x − T yk) ≤ φ(kT x − yk) − ψ(kx − yk)
(2.1)
for all x, y ∈ C. Then T has a unique fixed point in C.
Proof. Let x0 ∈ C be arbitrary but fixed. Consider, the iterated sequence {xn} where
xn = T nx0 for each natural number n.
Therefore, by given condition we have,
φ(kT xn−1 − T xm−1k) ≤ φ(kT xn−1 − xm−1k) − ψ(kxn−1 − xm−1k)
⇒ φ(kxn − xmk) ≤ φ(kxn − xm−1k) − ψ(kxn−1 − xm−1k)
which implies that,
φ(kxn − xmk) ≤ φ(kxn − xm−1k)
for all n, m ∈ N. Since φ is monotone increasing, we have
kxn − xmk ≤ kxn − xm−1k
for all n, m ∈ N.
Interchanging the role of xn and xm in above equation we get,
kxn − xmk ≤ kxn−1 − xmk
(2.2)
(2.3)
4
H. GARAI, L.K. DEY, A. CHANDA
Thus for each fixed n ∈ N, we can conclude that the sequence {s(n)
for all n, m ∈ N.
negative real numbers is monotone decreasing, where, s(n)
So, {s(n)
m }m∈N is convergent for each n ∈ N.
m }m∈N of non-
m = kxn−xmk for each m ∈ N.
Let,
for each n ∈ N.
Now from equation 2.2, we have,
lim
m→∞
s(n)
m = a(n)
φ(kxn − xmk) + ψ(kxn−1 − xm−1k) ≤ φ(kxn − xm−1k).
Keeping n fixed, taking limit as m → ∞ in both sides of above in-equation and using
the continuity of φ, ψ on [0,∞), we get
lim
m→∞
φ(kxn − xmk) + lim
m→∞
ψ(kxn−1 − xm−1k) ≤ lim
m→∞
φ(kxn − xm−1k)
⇒ φ( lim
m→∞kxn − xmk) + ψ( lim
m→∞kxn−1 − xm−1k) ≤ φ( lim
m→∞kxn − xm−1k)
⇒ φ(a(n)) + ψ(a(n−1)) ≤ φ(a(n))
⇒ ψ(a(n−1)) ≤ 0
⇒ ψ(a(n−1)) = 0
⇒ a(n−1) = 0 [since, ψ(t) = 0 if and only if t = 0].
m→∞kxn − xmk = 0 for all n ∈ N.
lim
Therefore a(n) = 0 for all n ∈ N, i.e.,
Now, we consider the sequence of functions {fm} defined on C by,
if x = xn for some n ∈ N;
otherwise.
fm(x) =(cid:26) kxn − xmk,
fm(x) = 0 for all x ∈ C. Thus the limit function f of the sequence
Therefore,
lim
m→∞
0,
of functions {fm} is given by
Now, let
Therefore,
f (x) = 0 for all x ∈ C.
Mm = sup
x∈C fm(x) − f (x).
Mm = sup
x∈C fm(x) [since, f (x) = 0 for all x ∈ C]
n fm(xn)
n kxn − xmk.
= sup
= sup
But, we know from 2.3 that,
kxn − xmk ≤ kxn−1 − xmk ≤ kxn−2 − xmk ≤ ... ≤ kx1 − xmk,
POSITIVE SOLUTION VIA FIXED POINT RESULTS
5
which implies that
sup
n kxn − xmk ≤ kx1 − xmk
m→∞
⇒ Mm ≤ kx1 − xmk
⇒ lim
Mm ≤ lim
Mm = 0.
⇒ lim
m→∞
lim
m→∞
m→∞kx1 − xmk = 0
Let ǫ > 0 be arbitrary. Since,
that
Mm = 0, so there exists a natural number N such
Mm < ǫ for all m ≥ N
x∈C fm(x) − f (x) < ǫ for all m ≥ N
⇒ sup
⇒ fm(x) − f (x) < ǫ for all m ≥ N and for all x ∈ C
⇒ fm(x) < ǫ for all m ≥ N and for all x ∈ C.
In particular, we have
Therefore, we can write,
fm(xn) < ǫ for all m ≥ N and for all n ∈ N.
(2.4)
for all n, m ≥ N .
Next, we consider the double sequence {snm}n,m∈N of real numbers, where
(cid:12)(cid:12)kxn − xmk − 0(cid:12)(cid:12) < ǫ
snm = kxn − xmk
for all n, m ∈ N. Here using 2.4, we have
This implies the double sequence {snm}n,m∈N converges to 0, i.e,
snm − 0 < ǫ for all n, m ≥ N.
n,m→∞kxn − xmk = 0.
lim
Thus, {xn} is a Cauchy sequence in C. C being complete, {xn} must converge to
Now from 2.1, we have,
some z ∈ C.
φ(kxn+1 − T zk) ≤ φ(kxn+1 − zk) − ψ(kxn − zk)
n→∞
φ(kxn+1 − T zk) ≤ lim
⇒ φ(kxn+1 − T zk) ≤ φ(kxn+1 − zk)
⇒ lim
⇒ φ( lim
⇒ φ( lim
⇒ φ( lim
⇒ lim
n→∞kxn+1 − T zk) ≤ φ( lim
n→∞kxn+1 − T zk) ≤ φ(0) = 0
n→∞kxn+1 − T zk) = 0
n→∞kxn+1 − T zk = 0.
n→∞
φ(kxn+1 − zk)
n→∞kxn+1 − zk)
The above equation shows that the sequence {xn} converges to T z. Thus, T z = z and
z is a fixed point of T .
Finally, we check the uniqueness of the fixed point z. To check this, let z1 be another
fixed point of T , i.e., T z1 = z1.
6
H. GARAI, L.K. DEY, A. CHANDA
From 2.1 and using Definition 1.4, we have
φ(kT z − T z1k) ≤ φ(kT z − z1k) − ψ(kz − z1k)
⇒ φ(kz − z1k) + ψ(kz − z1k) ≤ φ(kz − z1k)
⇒ ψ(kz − z1k) ≤ 0
⇒ ψ(kz − z1k) = 0
⇒ z = z1.
Therefore, z is the only fixed point of T .
(cid:3)
Now, we generalize Theorem 2.1 in a Banach space which is equipped with an arbi-
trary binary relation and state the subsequent theorem.
Theorem 2.2. Let (X,k.k) be a Banach space and R be an equivalence relation on X.
Assume that X has the property that if {xn} be any sequence in X converging to z ∈ X,
then xnRz for each natural number n. Let C be a closed subset of X and T : C → C
be a mapping such that T satisfies the following conditions:
i) T is order-preserving;
ii) φ(kT x − T yk) ≤ φ(kT x − yk) − ψ(kx − yk) for all x, y ∈ C such that xRy
where φ, ψ : [0,∞) → [0,∞) are two altering distance functions. Then T has a unique
fixed point in C if there exists x0 ∈ X such that x0RT x0.
Proof. The proof of this theorem is analogous to the previous one and so omitted. (cid:3)
In next portion of this section, we present a result which not only gives the guarantee
of existence of fixed point but also properly point out the fixed point.
Theorem 2.3. Let (X,k.k) be a Banach space and C be a closed subspace of X. Let
T : C → C be a mapping. Also assume that, there exist two altering distance functions
φ, ψ : [0,∞) → [0,∞) such that T satisfies the following conditions:
(i) φ(kT x − T yk) ≤ φ(kx − yk) − ψ(kx − yk);
(ii) φ(kT x − yk) ≤ φ(kx − yk) − ψ(kx − yk)
for all x, y ∈ C. Then the null vector of X is the only fixed point of T .
Proof. Let x0 ∈ C be arbitrary but fixed and consider the iterated sequence {xn} where
xn = T nx0 for all n ∈ N.
Let sn = kxn − xn+1k for all n ∈ N.
Now, by condition (i) we get
φ(kT xn − T xn+1k) ≤ φ(kxn − xn+1k) − ψ(kxn − xn+1k)
⇒ φ(kxn+1 − xn+2k) ≤ φ(kxn − xn+1k)
⇒ φ(sn+1) ≤ φ(sn)
⇒ sn+1 ≤ sn.
This is true for all natural number n, which implies that {sn} is a decreasing sequence
of non-negative reals and hence this sequence must converge. Let,
lim
n→∞
sn = a.
POSITIVE SOLUTION VIA FIXED POINT RESULTS
7
Again, from (i) we have,
φ(sn) − lim
n→∞
ψ(sn)
φ(sn+1) ≤ φ(sn) − ψ(sn)
n→∞
n→∞
⇒ lim
φ(sn+1) ≤ lim
⇒ φ(a) ≤ φ(a) − ψ(a)
⇒ ψ(a) ≤ 0
⇒ ψ(a) = 0
⇒ a = 0
⇒ lim
sn = 0.
n→∞
Therefore,
lim
n→∞kxn − xn+1 − θk = 0.
This shows that the sequence {un} in C converges strongly to θ, where θ is the null
vector in X and un = xn − xn+1 for all natural numbers n. Now,
φ(kT un − T θk) ≤ φ(kun − θk) − ψ(kun − θk)
φ(kun − θk) − lim
n→∞
n→∞
n→∞
ψ(kun − θk)
φ(kT un − T θk) ≤ lim
φ(kT un − T θk) ≤ 0
φ(kT un − T θk) = 0
n→∞kT un − T θk) = 0
n→∞
⇒ lim
⇒ lim
⇒ lim
⇒ φ( lim
⇒ lim
n→∞
n→∞kT un − T θk = 0.
Again, by condition (ii) we get,
φ(kT un − θk) ≤ φ(kun − θk) − ψ(kun − θk)
φ(kun − θk) − lim
n→∞
n→∞
φ(kT un − θk) ≤ lim
φ(kT un − θk) ≤ 0
φ(kT un − θk) = 0
n→∞kT un − θk) = 0
n→∞
n→∞
⇒ lim
⇒ lim
⇒ lim
⇒ φ( lim
⇒ lim
n→∞
n→∞kT un − θk = 0.
ψ(kun − θk)
Therefore by the uniqueness of limit, we obtain,
T θ = θ,
i.e., θ is a fixed point of T .
Finally, suppose z be another fixed point of T . Therefore,
φ(kT z − T θk) ≤ φ(kz − θk) − ψ(kz − θk)
⇒ φ(kz − θk) ≤ φ(kz − θk) − ψ(kz − θk)
⇒ ψ(kz − θk) ≤ 0
⇒ ψ(kz − θk) = 0
⇒ z = θ.
Therefore θ is the only fixed point of T in C.
(cid:3)
8
H. GARAI, L.K. DEY, A. CHANDA
Example 2.4. Consider the Banach space R endowed with the usual norm and define
a relation R on R by: for x, y ∈ R xRy if and only if either x, y ∈ [−(n + 1),−n] or
x, y ∈ [n, n + 1] for some n ∈ N or x = y = 0. Then clearly R is an equivalence relation
on R.
Now, let C = C1 ∪ C2 ∪ C3, where C1 = [−2,−1], C2 = [1, 2], C3 = {0}. Then C is
Define, a mapping T : C → C by
a closed subset of R.
T x =(cid:26) −x,
0,
if x ∈ C1;
if x ∈ C2 ∪ C3.
Therefore,
kT x − T yk =
x − y,
x,
0,
kT x − yk =(cid:26) x + y,
y,
if x, y ∈ C1;
if x ∈ C1 and y ∈ C2 ∪ C3;
if x, y ∈ C2 ∪ C3.
if x ∈ C1;
if x /∈ C1.
Consider the functions φ, ψ : [0,∞) → [0,∞) defined by
φ(t) = t2
ψ(t) =
t2
100000
for all t ∈ [0,∞).
such that xRy. Then the following cases arise.
Then, clearly φ, ψ are two altering distance functions. Let, x, y ∈ C be arbitrary
Case I: Let x, y ∈ C1. Then,
φ(kT x − T yk) + ψ(kx − yk) − φ(kT x − yk)
100000 − x + y2
= x − y2 + x − y2
(x − y)2
= −4xy +
100000
≤ 0
⇒ φ(kT x − T yk) ≤ φ(kT x − yk) − ψ(kx − yk).
Case II: Let x, y ∈ C2. Then,
φ(kT x − T yk) + ψ(kx − yk) − φ(kT x − yk)
(x − y)2
100000 − y2
= 0 +
≤ 0
⇒ φ(kT x − T yk) ≤ φ(kT x − yk) − ψ(kx − yk).
Case III: Let x, y ∈ C3. Then clearly the equality holds.
Thus,
φ(kT x − T yk) ≤ φ(kT x − yk) − ψ(kx − yk)
for all x, y ∈ C with xRy.
Also it is easily seen that T is order preserving and 0 is the only fixed point of T .
POSITIVE SOLUTION VIA FIXED POINT RESULTS
9
3. Application to Fractional Thermostat Model
The motivation of this section is to provide an application of the results discussed
in this manuscript. For this purpose, we consider the following fractional thermostat
model
subject to the boundary conditions:
CDαu(t) + λf (t, u(t)) = 0, t ∈ [0, 1],
(3.1)
(3.2)
where CDα stands for Caputo fractional derivative of order α, λ is a positive constant
and 1 < α ≤ 2, 0 ≤ η ≤ 1, β > 0 such that the following conditions hold:
u′(0) = 0, β CDα−1u(1) + u(η) = 0,
(1) βΓ(α) − (1 − η)(α−1) > 0;
(2) f : [0, 1] × R → R+ is a continuous function;
(3) u : [0, 1] → R is continuous.
Our aim is to derive some sufficient conditions under which the problem 3.1 with the
boundary conditions 3.2 possesses unique positive solution for certain values of the
parameter λ. To proceed further, we first recall the following lemmas.
Lemma 3.1.
boundary value problem
[21] Assume f ∈ C[0, 1]. A function u ∈ C[0, 1] is a solution of the
CDαu(t) + λf (t, u(t)) = 0, t ∈ [0, 1],
u′(0) = 0, β CDα−1u(1) + u(η) = 0,
if and only if it satisfies the integral equation
(3.3)
(3.4)
(3.5)
for s ≤ r and
where G(t, s) is the Green's function (depending on α) given by
0
G(t, s)f (s)ds
u(t) =Z 1
G(t, s) = β + Hη(s) − Ht(s)
and for r ∈ [0, 1], Hr(s) : [0, 1] → R is defined as Hr(s) = (r−s)α−1
Hr(s) = 0 for s > r, i.e.,
Γ(α)
G(t, s) =
Γ(α) + (η−s)α−1
Γ(α)
,
β − (t−s)α−1
β + (η−s)α−1
β − (t−s)α−1
Γ(α)
Γ(α)
,
,
β,
if 0 ≤ s ≤ η, s ≤ t;
if 0 ≤ s ≤ η, s ≥ t;
if η ≤ s ≤ 1, s ≤ t;
if η ≤ s ≤ 1, s ≥ t.
Lemma 3.2.
conditions:
[25] The function G(t, s) arising in Lemma 3.1 satisfies the following
i) G(t, s) is a continuous map defined on [0, 1] × [0, 1];
ii) for t, s ∈ (0, 1), we have G(t, s) > 0 .
Now we prove the following lemma.
Lemma 3.3. The Green's function G(t, s) derived in Lemma 3.1 satisfies
and
0
sup
t∈[0,1]Z 1
t∈[0,1]Z 1
inf
0
G(t, s)ds = β +
G(t, s)ds = β +
ηα
Γ(α + 1)
ηα − 1
Γ(α + 1)
.
10
H. GARAI, L.K. DEY, A. CHANDA
Proof. Let us consider the function ϕ defined on [0, 1] by
ϕ(t) =Z 1
0
G(t, s)ds
0
G(t, s)ds
for all t ∈ [0, 1].
Now, for t ∈ [0, 1] and t ≤ η, s ≥ η, we have t ≤ s and thus,
ϕ(t) = Z 1
= Z η
= Z t
= Z t
G(t, s)ds +Z 1
G(t, s)ds +Z η
(t − s)α−1
0 {β −
G(t, s)ds +Z 1
(η − s)α−1
}ds +Z η
t {β +
G(t, s)ds
G(t, s)ds
Γ(α)
Γ(α)
+
0
0
η
η
t
(η − s)α−1
Γ(α)
}ds +Z 1
η
βds
= β +
ηα − tα
Γ(α + 1)
.
0
G(t, s)ds
Again, for t ∈ [0, 1] and t ≥ η, s ≤ η, we have t ≥ s and so,
ϕ(t) = Z 1
= Z η
= Z η
= Z η
G(t, s)ds +Z 1
G(t, s)ds +Z t
(t − s)α−1
0 {β −
G(t, s)ds +Z 1
(η − s)α−1
}ds +Z t
η {β −
G(t, s)ds
G(t, s)ds
Γ(α)
Γ(α)
+
0
0
η
η
t
= β +
ηα − tα
Γ(α + 1)
.
(t − s)α−1
Γ(α)
}ds +Z 1
t
βds
Thus, from the above calculations we get,
for all t ∈ [0, 1].
Therefore,
ϕ(t) = β +
ηα − tα
Γ(α + 1)
ϕ′(t) = −αtα−1
Γ(α + 1)
< 0
for all t ∈ [0, 1].
This implies that the function ϕ is a decreasing function on [0, 1]. So,
sup
t∈[0,1]Z 1
0
G(t, s)ds = sup
t∈[0,1]
= φ(0)
ϕ(t)
= β +
ηα
Γ(α + 1)
,
POSITIVE SOLUTION VIA FIXED POINT RESULTS
11
and
inf
t∈[0,1]Z 1
0
G(t, s)ds = inf
t∈[0,1]
= φ(1)
ϕ(t)
= β +
ηα − 1
Γ(α + 1)
.
This completes the proof of the lemma.
(cid:3)
Lemma 3.4. For the Green's function G(t, s) derived in Lemma 3.1
G(t, s) ≤ β +
ηα−1
Γ(α)
for all t, s ∈ [0, 1] holds.
Proof. From the formulation of G(t, s) we get
∂G(t, s)
∂t
0,
Γ(α)
− (α−1)(t−s)α−2
− (α−1)(t−s)α−2
Γ(α)
0,
=
∂G(t, s)
∂t
,
,
if 0 ≤ s ≤ η, s ≤ t;
if 0 ≤ s ≤ η, s ≥ t;
if η ≤ s ≤ 1, s ≤ t;
if η ≤ s ≤ 1, s ≥ t.
≤ 0
Therefore, for any fixed s ∈ [0, 1], we have
for each t ∈ [0, 1] and thus G(t, s) is a decreasing function of t on [0, 1] for each fixed
s ∈ [0, 1].
Thus,
G(t, s) ≤ G(0, s) for all t, s ∈ [0, 1]
(3.6)
where
Therefore,
β,
Γ(α)
G(0, s) =( β + (η−s)α−1
=( −(α−1)(η−s)α−2
Γ(α)
∂s
0,
∂G(0, s)
,
if 0 ≤ s ≤ η;
if η ≤ s ≤ 1.
,
if 0 ≤ s ≤ η;
if η ≤ s ≤ 1.
This shows that ∂G(0,s)
on [0, 1]. Thus,
∂s ≤ 0 for all s ∈ [0, 1] and so G(0, s) is a decreasing function of s
G(0, s) ≤ G(0, 0) = β +
ηα−1
Γ(α)
for all s ∈ [0, 1].
From equations 3.6 and 3.7 we get
G(t, s) ≤ β +
ηα−1
Γ(α)
for all t, s ∈ [0, 1].
(3.7)
(cid:3)
Now we prove the following theorem concerning the existence and uniqueness of a
positive solution to the fractional thermostat model.
Theorem 3.5. Let us consider the fractional thermostat model with parameter λ > 0
given by equations 3.1 and 3.2. Assume that the following conditions hold:
12
H. GARAI, L.K. DEY, A. CHANDA
(i) βΓ(α + 1) + ηα > 1;
(ii) for all s ∈ [0, 1],
λf (s, u(s)) − f (s, v(s)) ≤ λf (s, u(s)) − λ sup
t∈[0,1]v(t) − ψ( sup
t∈[0,1]u(t) − v(t))
for some altering distance function ψ and for all real valued continuous func-
tions u(s), v(s) defined on [0, 1];
(iii) f is non-decreasing with respect to the second argument and there exists t0 ∈
(0, 1) such that f (t0, 0) > 0.
Then the fractional thermostat model with parameter λ given by equations 3.1 and 3.2
has a unique positive solution for λ ≥ 1
Proof. Consider the Banach space C[0, 1] of all real-valued continuous functions defined
on [0, 1] equipped with the sup norm.
k , where k = β + ηα−1
Γ(α+1) .
Define a mapping T : C[0, 1] → C[0, 1] by
T u(t) = λZ 1
0
G(t, s)f (s, u(s))ds
for all u ∈ C[0, 1], where G(t, s) is defined as in Lemma 3.1.
solution if and only if u(t) is a fixed point of T .
From Lemma 3.1, it is obvious that the thermostat model 3.1 and 3.2 has u(t) as a
Now, by condition (ii) we have,
λf (s, u(s)) − f (s, v(s)) ≤ λf (s, u(s)) − λ sup
t∈[0,1]v(t) − ψ( sup
t∈[0,1]u(t) − v(t))
= λf (s, u(s)) − λkvk − ψ(ku − vk).
Multiplying both sides by G(t, s), we get
⇒ λZ 1
−ψ(ku − vk)G(t, s)
λf (s, u(s)) − f (s, v(s))G(t, s) ≤ λf (s, u(s))G(t, s) − λkvkG(t, s)
0 f (s, u(s)) − f (s, v(s))G(t, s)ds ≤ λZ 1
0 f (s, u(s))G(t, s)ds
0 kvkG(t, s)ds −Z 1
−λZ 1
G(t, s)f (s, u(s))ds − λkvkZ 1
= λZ 1
0
0
0
ψ(ku − vk)G(t, s)ds
G(t, s)ds
−ψ(ku − vk)Z 1
0
G(t, s)ds
G(t, s)f (s, u(s))ds − λkvk inf
G(t, s)ds
t∈[0,1]Z 1
0
−ψ(ku − vk) inf
t∈[0,1]Z 1
0
G(t, s)ds
G(t, s)f (s, u(s))ds − λkkvk − kψ(ku − vk).
≤ λZ 1
0
≤ λZ 1
0
POSITIVE SOLUTION VIA FIXED POINT RESULTS
13
Now if λk ≥ 1, then from the above in-equation we obtain,
0 f (s, u(s)) − f (s, v(s))G(t, s)ds ≤ λZ 1
λZ 1
≤ λZ 1
≤ λZ 1
0
0
0
G(t, s)f (s, u(s))ds − kvk − kψ(ku − vk)
G(t, s)f (s, u(s))ds − v(t) − kψ(ku − vk)
G(t, s)f (s, u(s))ds − v(t)
(3.8)
Therefore using Equation 3.8 we get,
0
T u(t) − T v(t) = λZ 1
= λZ 1
≤ λZ 1
0
0
−kψ(ku − vk).
G(t, s)f (s, u(s))ds − λZ 1
G(t, s)(f (s, u(s)) − f (s, v(s)))ds
0
G(t, s)f (s, v(s))ds
G(t, s)f (s, u(s))ds − v(t) − kψ(ku − vk).
The above inequality holds for all t ∈ [0, 1] and so we have,
sup
t∈[0,1]T u(t) − T v(t) ≤ sup
t∈[0,1]
G(t, s)f (s, u(s))ds − v(t) − kψ(ku − vk)
λZ 1
0
(3.9)
⇒ kT u − T vk ≤ kT u − vk − kψ(ku − vk).
It is easily perceived by condition (i) that, k > 0.
Define two functions φ, ψ1 : [0,∞) → [0,∞) by
φ(t) = t and
for all t ∈ [0,∞). Then one can easily verify that φ, ψ1 are two altering distance
functions and also from equation 3.9 we get,
ψ1(t) = kψ(t)
φ(kT u − T vk) ≤ φ(kT u − vk) − ψ1(ku − vk).
(3.10)
The above inequality holds for all u, v ∈ C[0, 1] and so by Theorem 2.1, T has a unique
fixed point u(t), say, in C[0, 1].
Note that, Equation 3.10 holds if λk ≥ 1. So, T has u(t) as a fixed point if λk ≥ 1,
Now we have λ > 0, G(t, s) > 0 and f (s, u(s)) ≥ 0 for all t, s ∈ [0, 1]. Therefore it is
i.e., u(t) is a solution of the thermostat model 3.1 and 3.2 if λk ≥ 1, i.e., λ ≥ 1
k .
clear that
λZ 1
for all t ∈ [0, 1]. This means that T u(t) ≥ 0 for all t ∈ [0, 1] and which leads us to the
fact that u(t) ≥ 0 for all t ∈ [0, 1].
Finally, we show that the unique solution u(t) is always positive. To show this, first
we show that the zero function 0 is not a fixed point of T .
G(t, s)f (s, u(s))ds ≥ 0
0
Suppose in contrary, assume that the zero function 0 is a fixed point of T . Then, we
have
0 = λZ 1
0
G(t, s)f (s, 0)ds,
14
H. GARAI, L.K. DEY, A. CHANDA
for all t ∈ [0, 1]. Since G(t, s)f (s, 0) ≥ 0 for all t ∈ [0, 1] and for all s ∈ [0, 1], we have
G(t, s)f (s, 0) = 0,
for all t ∈ [0, 1] and for almost all s ∈ [0, 1]. This fact leads us to
f (s, 0) = 0 for almost all s ∈ [0, 1].
(3.11)
By condition (iii), there exists t0 ∈ (0, 1) such that f (t0, 0) > 0. Again, since f is
continuous at (t0, 0), there exists a subset A of [0, 1] of positive Lebesgue measure such
that f (s, 0) > 0 for all s ∈ A. This is a contradiction to 3.11. So the zero function 0 is
not a fixed point of T .
Now, let u(t1) = 0 for some t1 ∈ (0, 1). Therefore we have,
G(t1, s)f (s, u(s))ds = 0.
(3.12)
Z 1
0
But u(s) ≥ 0 for all s ∈ [0, 1] and f is non-decreasing with respect to the second
argument. Hence
0 ≥Z 1
0
G(t1, s)f (s, u(s))ds ≥ Z 1
0
Therefore from 3.12 and 3.13, we obtain
G(t1, s)f (s, 0)ds ≥ 0.
(3.13)
Z 1
0
G(t1, s)f (s, 0)ds = 0.
As G(t1, s)f (s, 0) ≥ 0, it follows that G(t1, s)f (s, 0) = 0 for almost all s ∈ [0, 1]. This
implies that f (s, 0) = 0 for almost all s ∈ [0, 1], which is a contradiction.
Hence it follows that, u(t) > 0 for all t ∈ (0, 1). Again, since u is continuous on
[0, 1], we have u(t) > 0 for all t ∈ [0, 1]. Thus the fractional thermostat model, given
by Equations 3.1 and 3.2, has a unique positive solution for λ ≥ 1
k , where k = β +
ηα−1
Γ(α+1) .
(cid:3)
Theorem 3.6. Let us consider the fractional thermostat model with parameter λ given
by equations 3.1 and 3.2. Assume that the following conditions hold:
(i) βΓ(α + 1) + ηα > 1;
(ii) for all s ∈ [0, 1],
λf (s, u(s)) − f (s, v(s)) ≤ λf (s, u(s)) − λ sup
t∈[0,1]v(t) − ψ( sup
t∈[0,1] u(t) − v(t)),
for some altering distance function ψ, for all u(s), v(s) in the set C = {u(s) ∈
C[0, 1] : 0 ≤ u(s) ≤ R, for all s ∈ [0, 1] and R is a positive constant},
(iii) Z 1
(iv) f is non-decreasing with respect to the second argument and there exists t0 ∈
, where k1 = β + ηα−1
Γ(α) ;
f (s, R)ds ≤
R
λk1
0
(0, 1) such that f (t0, 0) > 0.
Then the fractional thermostat model with parameter λ given by equations 3.1 and 3.2
has a unique positive solution in C for λ ≥ 1
Proof. Let us take the Banach space C[0, 1] endowed with the sup norm. Then it is
easily noticeable that C is a closed subset of C[0, 1].
k where k = β + ηα−1
Γ(α+1) .
Now, for any u(s) ∈ C[0, 1] we have
u(s) ≤ R.
POSITIVE SOLUTION VIA FIXED POINT RESULTS
15
The fact that f is non-decreasing with respect to the second argument gives us
Z 1
0
f (s, u(s))ds ≤ Z 1
0
R
λk1
≤
f (s, R))ds
.
Therefore,
λZ 1
0
i.e.,
G(t, s)f (s, u(s))ds ≤ λ(cid:0)β +
≤ λk1
R
λk1
,
ηα−1
Γ(α)(cid:1)Z 1
0
λZ 1
Next, we define a mapping T : C → C by
T u(t) = λZ 1
0
0
G(t, s)f (s, u(s)) ≤ R.
G(t, s)f (s, u(s))ds
f (s, u(s))ds
(3.14)
for all u ∈ C[0, 1], where G(t, s) is given by Lemma 3.1.
From Equation 3.14 one can easily check that T is well-defined on C.
We now define two functions φ, ψ1 : [0,∞) → [0,∞) by
φ(t) = t
ψ1(t) = kψ(t)
for all t ∈ [0,∞). Then it is clear that φ, ψ1 are altering distance functions.
Now proceeding as in Theorem 3.5 we get
φ(kT u − T vk) ≤ φ(kT u − vk) − ψ1(ku − vk)
for all u, v ∈ C if λ ≥ 1
k .
theorem, T has a unique fixed point in C, say, u(t).
Thus we see that all conditions of Theorem 2.1 are satisfied if λ ≥ 1
Thus, u(t) is the unique solution of the fractional thermostat model given by Equa-
tions 3.1 and 3.2, which follows by Lemma 3.1 and the definition of T . The fact that
u(t) is positive on [0, 1] follows by Theorem 3.5 using condition (iv). Hence, the frac-
tional thermostat model given by Equations 3.1 and 3.2 satisfying the hypotheses of
Theorem 3.6, has a unique positive solution for λ ≥ 1
k .
(cid:3)
k . So by the
Now, we demonstrate an example which validates the effectiveness of the aforemen-
tioned result.
Example 3.7. Let us consider the fractional thermostat model
C Dαu(t) + λf (t, u(t)) = 0, t ∈ (0, 1),
u′(0) = 0, βCDα−1u′(1) + u(η) = 0.
(3.15)
(3.16)
We choose,
3
2
α =
, β =
1
2
2 .√π − ( 1
5 . 1
Then, βΓ(α) − (1 − η)(α−1) = 4
2 )
, η =
4
5
1
2 > 0.
and f (t, u(t)) = ln(320 + t2) + t3 +
1
24 − u(t)
.
16
H. GARAI, L.K. DEY, A. CHANDA
with respect to the second argument and there exists 1
Clearly f : [0, 1] × R → R+ is a continuous function and also f is non-decreasing
2 , 0) > 0.
We take
2 ∈ (0, 1) such that f ( 1
i.e., here R = 20.
C = {u(s) ∈ C[0, 1] : 0 ≤ u(s) ≤ 20, for all s ∈ [0, 1]}
Now,
and
βΓ(α + 1) + ηα =
4
5
.
3
2
.
1
2
.√π + (
1
2
)
3
2 ≈ 1.4165 > 1,
k = β +
=
4
5
+
ηα − 1
Γ(α + 1)
( 1
2 )
2 − 1
2 .√π
2 . 1
3
3
≈ 0.3135
1
k ≈ 3.1897.
⇒
k1 = β +
=
4
5
+
ηα − 1
Γ(α)
( 1
2 )
2 .√π
1
1
2
We choose λ = 3.2 and clearly λ ≥ 1
k .
Now,
≈ 1.5981.
Z 1
0
f (s, R)ds ≤
1
320 +
1
4
+
1
4
1
1 + 320 −
R
λk1
≤
≈
≈ 3.9109.
Next, we define a mapping ψ : [0,∞) → [0,∞) by
20
3.2 × 1.5981
ψ(t) =(cid:26) t2,
Finally, for any u(s), v(s) ∈ C we have,
1,
if 0 ≤ t < 1;
if t ≥ 1.
Then it is an easy task to note that ψ is an altering distance function.
λf (s, u(s)) − f (s, v(s)) = 3.2(cid:12)(cid:12)
≤ 3.2(cid:0)
= 1.6.
1
24 − v(s)(cid:12)(cid:12)
1
24 − u(s) −
1
4
+
1
4(cid:1)
But,
λf (s, u(s)) − λ sup
t∈[0,1]u(t) − ψ( sup
t∈[0,1] u(t) − v(t)) ≥ 3.2 × 21 − 3.2 × 20 − ψ(40)
= 2.20.
POSITIVE SOLUTION VIA FIXED POINT RESULTS
17
Therefore,
λf (s, u(s)) − f (s, v(s)) ≤ λf (s, u(s)) − λ sup
t∈[0,1]u(t) − ψ( sup
t∈[0,1] u(t) − v(t))
for all u(s), v(s) ∈ C. So, by Theorem 3.6 the thermostat model, given by Equations
3.15 and 3.16 has a unique positive solution in C for λ = 3.2.
Acknowledgements:
The first named author would like to express his genuine appreciation to CSIR, New
Delhi, India for their financial supports. Also the third named author would like to
convey his cordial thanks to DST-INSPIRE, New Delhi, India for their financial aid
under INSPIRE fellowship scheme.
References
[1] R.P. Agarwal, M.A. El-Gebeily, and D. O'Regan. Generalized contractions in partially ordered
metric spaces. Appl. Anal., 87(1):109 -- 116, 2008.
[2] Ya.I. Alber and S. Guerre-Delabriere. Principle of weakly contractive maps in Hilbert spaces. Oper.
Theory Adv. Appl., 98:7 -- 22, 1997.
[3] I. Altun and H. Simsek. Some fixed point theorems on ordered metric spaces and application.
Fixed Point Theory Appl., 2010, 2010. Article ID 621469.
[4] A. Amini-Harandi and H. Emami. A fixed point theorem for contraction type maps in partially or-
dered metric spaces and application to ordinary differential equations. Nonlinear Anal., 72(5):2238 --
2242, 2010.
[5] Z. Bai and H. Lu. Positive solutions for boundary value problem of nonlinear fractional differential
equation. J. Math. Anal. Appl., 311(2):495 -- 505, 2005.
[6] S. Banach. Sur les op´erations dans les ensembles abstraits et leur application aux ´equations
int´egrales. Fund. Math., 3:133 -- 181, 1922.
[7] I. Beg and M. Abbas. Coincidence point and invariant approximation for mappings satisfying
generalized weak contractive condition. Fixed Point Theory Appl., 2006, 2006. Article ID 74503.
[8] A. Cabada and G. Wang. Positive solutions of nonlinear fractional differential equations with
integral boundary value conditions. J. Math. Anal. Appl., 389(1):403-411, 2012.
[9] J. Caballero, J. Harjani, and K. Sadarangani. Existence and uniqueness of positive solution for a
boundary value problem of fractional order. Abstr. Appl. Anal., 2011, 2011. Article ID 165641.
[10] J. Caristi. Fixed point theorems for mappings satisfying inwardness conditions. Trans. Amer.
Math. Soc., 215:241 -- 251, 1976.
[11] B.S. Choudhury. Unique fixed point theorem for weakly C-contractive mappings. Kathmandu Univ.
J. Sci. Engg. Tech., 5(1):6 -- 13, 2009.
[12] LB. ´Ciri´c. A generalization of Banach's contraction principle. Proc. Amer. Math. Soc., 45(2):267 --
273, 1974.
[13] LB. ´Ciri´c, B. Samet, and C. Vetro. Common fixed point theorems for families of occasionally
weakly compatible mappings. Math. Comput. Modelling, 53(5-6):631 -- 636, 2010.
[14] R. Dehghani and K. Ghanbari. Triple positive solutions for boundary value problem of a nonlinear
fractional differential equation. Bull. Iranian Math. Soc., 33(2):1 -- 14, 2007.
[15] P.N. Dutta and B.S. Choudhury. A generalisation of contraction principle in metric spaces. Fixed
Point Theory Appl., 2008, 2008. Article ID 406368.
[16] I. Ekeland. On the variational principle. J. Math. Anal. Appl., 47(2):324 -- 353, 1974.
[17] D. Gopal, M. Abbas, D.K. Patel, and C. Vetro. Fixed points of α-type F-contractive mappings with
an application to nonlinear fractional differential equation. Acta Math. Scientia., 36(3):957 -- 970,
2016.
[18] D.S. Jaggi. Some unique fixed point theorems. Indian J. Pure Appl. Math, 8(2):223 -- 230, 1977.
[19] M.S. Khan, M. Swalech, and S. Sessa. Fixed point theorems by altering distances between the
points. Bull. Austral. Math. Soc., 30(1):1 -- 9, 1984.
[20] A. Meir and E. Keeler. A theorem on contraction mappings. J. Math. Anal. Appl., 28:326 -- 329,
1969.
18
H. GARAI, L.K. DEY, A. CHANDA
[21] J.J. Nieto and J. Pimentel. Positive solutions of a fractional thermostat model. Bound. Value
Probl., 2013:5, 2013.
[22] E. Rakotch. A note on contractive mappings. Proc. Amer. Math. Soc., 13:459 -- 465, 1962.
[23] B.E. Rhoades. Some theorems on weakly contractive maps. Nonlinear Anal., 47(4):2683 -- 2693,
2001.
[24] T. Senapati and L.K. Dey. Relation-theoretic metrical fixed point results via w-distance with
applications. J. Fixed Point Theory Appl., 2017. DOI: 10.1007/s11784-017-0462-9.
[25] C. Shen, H. Zhou, and L. Yang. Existence and nonexistence of positive solutions of a fractional
thermostat model with a parameter. Math. Methods Appl. Sci., 39(15):4504 -- 4511, 2016.
1 Hiranmoy Garai, Department of Mathematics, National Institute of Technology
Durgapur, India.
E-mail address: [email protected]
2 Lakshmi Kanta Dey, Department of Mathematics, National Institute of Technology
Durgapur, India.
E-mail address: [email protected]
3 Ankush Chanda, Department of Mathematics, National Institute of Technology
Durgapur, India.
E-mail address: [email protected]
|
1409.6101 | 2 | 1409 | 2016-04-21T07:24:03 | Functional calculus on real interpolation spaces for generators of $C_{0}$-groups | [
"math.FA",
"math.OA"
] | We study functional calculus properties of $C_{0}$-groups on real interpolation spaces, using transference principles. We obtain interpolation versions of the classical transference principle for bounded groups and of a recent transference principle for unbounded groups. Then we show that each group generator on a Banach space has a bounded $H^{\infty}_{1}$-calculus on real interpolation spaces. Additional results are derived from this. | math.FA | math |
FUNCTIONAL CALCULUS ON REAL INTERPOLATION SPACES
FOR GENERATORS OF C0-GROUPS
MARKUS HAASE AND JAN ROZENDAAL
Abstract. We study functional calculus properties of C0-groups on real in-
terpolation spaces, using transference principles. We obtain interpolation ver-
sions of the classical transference principle for bounded groups and of a recent
transference principle for unbounded groups. Then we show that each group
generator on a Banach space has a bounded H∞
1 -calculus on real interpolation
spaces. Additional results are derived from this.
The classical transference principle by Berkson, Gillespie and Muhly from [5]
1. Introduction
yields an estimate
(1.1)
ZR
(cid:13)(cid:13)(cid:13)(cid:13)
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)X ≤ M 2 kLµkL(Lp(X)) kxkX
for all x ∈ X, where (U (s))s∈R ⊆ L(X) is a bounded C0-group of operators on a
Banach space X with uniform bound M , µ is a complex Borel measure on R and
Lµ is convolution with µ on Lp(X), the space of p-integrable X-valued functions,
for p ∈ [1,∞]. Under certain geometrical assumptions on X, the norm of Lµ can be
bounded in terms of a suitable norm of the Fourier transform F µ of µ. For instance,
if X is a Hilbert space then kLµkL(L2(X)) is equal to kF µk∞, by Plancherel's theo-
rem. If X is a UMD space and p ∈ (1,∞) then bounds for kLµkL(Lp(X)) follow from
the Mikhlin multiplier theorem. By combining this with (1.1), functional calculus
bounds for the generator A of (U (s))s∈R can be obtained, i.e., estimates of the form
kf (A)k ≤ C kfkF for all f in some function algebra F . Such bounds are important
for evolution equations, conform for instance [2, 17].
Useful as this procedure is, the assumptions on the space X restrict the generality
of the results. In particular, Hilbert and UMD spaces are reflexive. Therefore the
approach described above generally does not yield interesting results for groups of
operators on non-reflexive spaces, such as C(K)-spaces or L1-spaces. In this paper
we take a different approach and consider transference principles on interpolation
spaces. It is known that the functional calculus properties of operators improve
upon restriction to interpolation spaces, conform for instance the result of Dore [6]
that an invertible sectorial operator has a bounded sectorial H∞-calculus on real
interpolation spaces. However, we are interested in functional calculus on strips,
Date: September 4, 2018.
2010 Mathematics Subject Classification. 47A60, 47D03, 42B35, 42A45, 46B70.
Key words and phrases. Functional calculus, Transference, Operator semigroup, Fourier mul-
tiplier, Interpolation space.
The second-named author is supported by NWO-grant 613.000.908 "Applications of Transfer-
ence Principles".
1
2
MARKUS HAASE AND JAN ROZENDAAL
the more natural choice for group generators. We use that on Besov spaces Fourier
multiplier results hold that do not depend on the geometry of the underlying space
[9, 16]. Since Besov spaces are obtained from real interpolation between Lp and
Sobolev spaces, this fits into the setting of a transference principle on interpolation
In Proposition 3.2 we derive the following version of (1.1) on the real
spaces.
interpolation space (X, D(A))θ,q from (1.4). For the X-valued Besov space Bθ
p,q(X)
see Section 2.2.
Proposition 1.1. Let X be a Banach space and let θ ∈ (0, 1), p ∈ [1,∞) and
q ∈ [1,∞]. Then there exists a constant C ≥ 0 such that the following holds. If −iA
generate a C0-group (U (s))s∈R on a Banach space X with M := sups∈R kU (s)k <
∞, then
≤ CM 2 kLµkL(Bθ
p,q (X)) kxk(X,D(A))θ,q
(cid:13)(cid:13)(cid:13)(cid:13)
ZR
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)(X,D(A))θ,q
for all complex Borel meaures µ on R and x ∈ (X, D(A))θ,q.
Combining Proposition 1.1 with the aforementioned Fourier multiplier results on
Besov spaces yields the following, a consequence of Corollary 3.5.
Corollary 1.2. Let −iA generate a uniformly bounded C0-group (U (s))s∈R on a
Banach space X, and let θ ∈ (0, 1), q ∈ [1,∞]. Then there exists a constant C ≥ 0
such that
(cid:13)(cid:13)(cid:13)(cid:13)
ZR
s∈R F µ(s) + (1 + s)(F µ)′(s)kxk(X,D(A))θ,q
for all x ∈ (X, D(A))θ,q and for each µ ∈ M(R) such that F µ ∈ C1(R) with
sups∈R(1 + s)(F µ)′(s) < ∞.
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)(X,D(A))θ,q
≤ C sup
We also obtain an interpolation version of the transference principle for un-
bounded groups from [12], as Proposition 3.1. In terms of functional calculus for
the part of A in (X, D(A))θ,q, these transference principles yield a result for func-
tions in the analytic Mikhlin algebra
(1.2)
(1.3)
endowed with the norm
H∞
1 (Stω) :=(cid:26)f ∈ H∞(Stω)(cid:12)(cid:12)(cid:12)(cid:12)
sup
z∈Stω
kfkH∞
1 (Stω) := sup
z∈Stωf (z) + (1 + z)f ′(z)
(1 + z)f ′(z) < ∞(cid:27) ,
(f ∈ H∞
1 (Stω)).
Here Stω := {z ∈ C Im(z) < ω} for ω > 0. Note that definition (1.3) of the norm
in the analytic Mikhlin algebra is different from that in [12], where the quantity
kfk = supz∈Stωf (z)+zf ′(z) is considered. However, the two norms are equivalent
on domains containing zero, and (1.3) is more natural in the setting of transference
principles on (inhomogeneous) Besov spaces, since Fourier multiplier results on such
spaces require an inhomogeneous condition at zero. See also Remarks 3.7 and 4.2.
Our main functional calculus result is as follows. For the group type θ(U ) see
(2.1), and for a proof of this result see Theorem 4.1.
Theorem 1.3. Let −iA be the generator of a C0-group (U (s))s∈R on a Banach
space X, and let θ ∈ (0, 1), q ∈ [1,∞]. Then the part of A in (X, D(A))θ,q has a
bounded H∞
1 (Stω)-calculus for all ω > θ(U ). If (U (s))s∈R is uniformly bounded then
the constant bounding the H∞
1 (Stω)-calculus does not depend on ω > 0.
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
3
In [12], Theorem 1.3 is obtained for group generators on UMD spaces and func-
tional calculus for the operator A itself. Our result shows that on interpolation
spaces no assumptions on the geometry of the underlying space are required. This
means that even on spaces which are not UMD, such as C(K)-spaces and L1-spaces,
one can obtain functional calculus results if one is willing to restrict to interpolation
spaces. Moreover, our results reaffirm the philosophy that the functional calculus
properties of an operator improve when restricted to interpolation spaces, as was
already evidenced for functions on sectors by the result of Dore [6, Theorem 3.2].
From Theorem 1.3 we deduce other functional calculus statements, for sectorial
operators and generators of cosine functions.
Section 2 provides the necessary background on functional calculus and the the-
ory of Fourier multipliers on Besov spaces. In Section 3 we establish transference
principles on interpolation spaces, and in Section 4 we prove Theorem 1.3. Section
5 contains additional results that can be derived from this.
1.1. Notation and terminology. The natural numbers are N := {1, 2, . . .}, and
we write N0 := N ∪ {0}. The letters X and Y denote Banach spaces over the
complex number field, and L(X) is the Banach algebra of all bounded operators on
X. The domain D(A) ⊆ X of a closed unbounded operator A on X is a Banach
space when endowed with the norm
kxkD(A) := kxk + kAxk
(x ∈ D(A)).
The spectrum of A is σ(A) and the resolvent set ρ(A) := C \ σ(A). For z ∈ ρ(A)
the operator R(z, A) := (zI − A)−1 ∈ L(X) is the resolvent of A at z.
For p ∈ [1,∞], Lp(R; X) is the Bochner space of equivalence classes of X-valued
Lebesgue p-integrable functions on R. The Holder conjugate of p ∈ [1,∞] is p′ and
is defined by 1
p′ = 1. The norm on Lp(R; X) is usually denoted by k·kp. In the
case X = C we will simply write Lp(R) = Lp(R; C).
By M(R) we denote the space of complex-valued Borel measures on R with the
total variation norm. For ω ≥ 0 we let Mω(R) consist of those µ ∈ M(R) of the
form µ(ds) = e−ωsν(ds) for some ν ∈ M(R), with
p + 1
kµkMω(R) :=(cid:13)(cid:13)(cid:13)
eω·µ(cid:13)(cid:13)(cid:13)M(R)
.
Note that Mω(R) is a Banach algebra under convolution. A function g such that
[s 7→ g(s) eωs] ∈ L1(R) is usually identified with its associated measure µ ∈ Mω(R)
given by µ(ds) = g(s)ds.
For Ω 6= ∅ open in C we let H∞(Ω) be the unital Banach algebra of bounded
holomorphic functions on Ω with the supremum norm
kfkH∞(Ω) := sup
z∈Ωf (z)
(f ∈ H∞(Ω)).
We mainly consider the case where Ω is a strip of the form
Stω := {z ∈ C Im(z) < ω}
for ω > 0, with St0 := R.
The Schwartz class S(R; X) is the space of X-valued rapidly decreasing smooth
functions on R, and the space of X-valued tempered distributions is S′(R; X). The
Fourier transform of an X-valued tempered distribution Φ ∈ S′(R; X) is denoted
4
MARKUS HAASE AND JAN ROZENDAAL
by F Φ. For instance, if µ ∈ Mω(R) for ω > 0 then F µ ∈ H∞(Stω)∩ C(Stω) is given
by
F µ(z) :=ZR
e−iszµ(ds)
(z ∈ Stω).
If X and Y are Banach spaces that are embedded continuously into a Hausdorff
topological vector space Z, then we call (X, Y ) an interpolation couple. We let
K(t, z) := inf {kxkX + tkykY x ∈ X, y ∈ Y, x + y = z}
for t > 0 and z ∈ X + Y ⊆ Z. The real interpolation space of X and Y with
parameters θ ∈ [0, 1] and q ∈ [1,∞] is
(1.4)
(X, Y )θ,q :=(cid:8)z ∈ X + Y [t 7→ t−θK(t, z)] ∈ Lq((0,∞), dt/t)(cid:9) ,
(z ∈ (X, Y )θ,q).
kzk(X,Y )θ,q
a Banach space when equipped with the norm
If T : X +Y → X +Y restricts to a bounded operator on X and a bounded operator
on Y then
:=(cid:13)(cid:13)t 7→ t−θK(t, z)(cid:13)(cid:13)Lq((0,∞),dt/t)
L(X) kTkθ
kTkL((X,Y )θ,q ) ≤ kTk1−θ
(1.5)
for all θ ∈ (0, 1) and q ∈ [1,∞] [4, Theorem 3.1.2]. We mainly consider interpolation
spaces for the couple (X, D(A)), where A is a closed operator on X. We write
L(Y )
and
DA(θ, q) := (X, D(A))θ,q
kxkθ,q := kxkDA(θ,q)
(x ∈ DA(θ, q)).
For an operator B on X and a continuously embedded space Y ֒→ X, the operator
BY on Y that satisfies BY y = By for elements in its domain
D(BY ) := {y ∈ D(B) ∩ Y By ∈ Y }
is the part of B in Y . If Y = DA(θ, q) for θ ∈ (0, 1) and q ∈ [1,∞] then we write
Bθ,q := BDA(θ,q).
Throughout, an X-valued function space Φ(R; X) on the real line will be denoted
by Φ(X) whenever little confusion can arise.
2. Functional calculus and Fourier multipliers
2.1. Functional calculus. We assume that the reader is familiar with the basics
of the theory of C0-groups as developed in, for instance, [8], and merely recall some
of the notions and results in functional calculus theory that are used. Details on
functional calculus for group generators can be found in [10, Chapter 4].
Let −iA be the generator of a C0-group (U (s))s∈R on a Banach space X. Then
the group type of U ,
(2.1)
θ(U ) := infnω ≥ 0(cid:12)(cid:12)(cid:12)∃M ≥ 1 such that kU (s)k ≤ M eωs for all s ≥ 0o ,
is finite. Moreover, A is a strip-type operator of height ω0 := θ(U ), i.e., σ(A) ⊆ Stω0
and
λ∈C\Stω kR(λ, A)k < ∞ for all ω > ω0.
sup
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
5
The strip-type functional calculus for A is defined as follows. First, operators f (A) ∈
L(X) are associated with functions
f ∈ 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)
for ω > ω0, by a Cauchy-type integral
f (A) :=
f (z)R(z, A) dz.
1
2πiZδStω′
Here δStω′ 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). The definition of f (A) is extended to a larger
class of functions by regularization, i.e.
f (A) := e(A)−1(ef )(A)
if there exists e ∈ E(Stω) with e(A) injective and ef ∈ E(Stω). This yields a closed
unbounded operator f (A) on X, and the definition of f (A) is independent of the
choice of the regularizer e. The algebra of all meromorphic functions on Stω that
are regularizable for A is denoted by MA(Stω). Each f ∈ H∞(Stω) is regularizable
by the function z 7→ (λ − z)−2, for Im(λ) > ω.
Since −iA generates a C0-group, the Hille-Phillips functional calculus for A
yields certain functions f that give rise to bounded operators f (A). Fix M ≥ 1
and ω ≥ 0 such that kU (s)k ≤ M eωs for all s ∈ R. For µ ∈ Mω(R) define
(2.2)
U (s)x µ(ds)
(x ∈ X).
Uµx :=ZR
Then µ 7→ Uµ is an algebra homomorphism Mω(R) → L(X). The following lemma,
Lemma 2.2 in [11], shows that the Hille-Phillips calculus extends the strip-type
calculus for A.
Lemma 2.1. Let X, A and U be as above, and let ω′ > ω ≥ 0.
a) For each f ∈ E(Stω′ ) there exists µ ∈ Mω(R) such that f = F µ.
b) Let µ ∈ Mω(R) be such that F µ extends to an element of MA(Stω′). Then
f (A) = Uµ ∈ L(X) and
t∈R kf (t + A)k ≤ M kµkMω(R) .
sup
We now consider functional calculus for operators on interpolation spaces. The
following lemma shows that, in particular, the functional calculi for A and Aθ,q are
compatible.
Lemma 2.2. Let A be a strip-type operator of height ω0 on a Banach space X and
let θ ∈ (0, 1), q ∈ [1,∞] and m, n ∈ N0. Let Y := (D(Am), D(An))θ,q.
a) The part AY of A in Y is a strip-type operator of height ω0. Moreover,
f ∈ MAY (Stω) with f (AY ) = f (A)Y for all ω > ω0 and f ∈ MA(Stω).
erates the C0-group (U (s)Y )s∈R. In particular, D(AY ) is dense in Y .
b) If −iA generates a C0-group (U (s))s∈R on X and q < ∞, then −iAY gen-
Proof. a) First note that, for all k ∈ N0 and λ ∈ ρ(A), R(λ, A) leaves D(Ak)
invariant with kR(λ, A)kL(D(Ak)) ≤ kR(λ, A)kL(X). By (1.5), R(λ, A) leaves Y
invariant with
(2.3)
kR(λ, A)kL(Y ) ≤ kR(λ, A)kL(X) .
6
MARKUS HAASE AND JAN ROZENDAAL
By [10, Proposition A.2.8], σ(AY ) ⊆ σ(A) and R(λ, AY ) = R(λ, A)Y for all λ ∈
ρ(A). Hence (2.3) yields the first statement. Let ω > ω0 and f ∈ E(Stω) be given.
Then
f (AY )y =
f (z)R(z, AY )y dz =
f (z)R(z, A)y dz = f (A)y
1
2πiZΓ
1
2πiZΓ
for some contour Γ and all y ∈ Y . For a general f ∈ MA(Stω), note that e is a
regulariser for f in the functional calculus for AY if it is a regulariser for f in the
functional calculus for A, since then e(AY ) = e(A)Y is injective. The rest follows
by regularization.
b) By (1.5), kU (s)Y k ≤ kU (s)k for all s ∈ R. Hence (U (s)Y )s∈R is locally
bounded. Since it is strongly continuous on the dense subset D(Amax(n,m)) ⊆ Y
[19, Proposition 1.2.5], it is strongly continuous on Y . By [8, p. 60], −iAY is its
generator.
(cid:3)
Remark 2.3. Part b) of Lemma 2.2 ensures that the integral in (2.2) is well-defined
and converges in DA(θ, q), for x ∈ DA(θ, q) and q < ∞. Even though (U (s))s∈R
is not strongly continuous on DA(θ,∞) in general, the integral is well-defined and
converges in X. Since DA(θ, q) is continuously embedded in X for all θ ∈ (0, 1)
and q ∈ [1,∞], the value of the integral does not depend on the space in which
convergence takes place. Hence from now on we regularly will not specify in which
norm (2.2) converges.
Let A be a strip-type operator of height ω0 and ω > ω0. For a Banach algebra
F of functions that is continuously embedded in H∞(Stω), we say that A has a
bounded F -calculus if there exists a constant C ≥ 0 such that f (A) ∈ L(X) with
kf (A)kL(X) ≤ C kfkF
for all f ∈ F .
The next lemma from [10, Proposition 5.1.7] is fundamental.
Lemma 2.4 (Convergence Lemma). Let A be a densely defined strip-type operator
of height ω0 on a Banach space X. Let ω > ω0 and (fj)j∈J ⊆ H∞(Stω) be a net
satisfying the following conditions:
(1) supj∈J kfjkH∞(Stω) < ∞;
(2) f (z) := limj fj(z) exists for all z ∈ Stω;
(3) supj∈J kfj(A)kL(X) < ∞.
Then f ∈ H∞(Stω), f (A) ∈ L(X), fj(A) → f (A) strongly and
kf (A)k ≤ lim sup
j∈J
kfj(A)k .
2.2. Fourier multipliers on Besov spaces. Let us summarize some results about
Fourier multipliers on vector-valued Besov spaces which will be used later on. De-
tails can be found in [1] and [9].
Let ψ ∈ C∞(R) be a nonnegative function with support in [ 1
for all s ∈ (0,∞).
ψ(2−ks) = 1
∞
2 , 2] such that
For k ∈ N and s ∈ R let ϕk(s) := ψ(2−ks), and let ϕ0(s) := 1 −P∞
k=1 ϕk(s).
Let X be a Banach space and let p, q ∈ [1,∞] and r ∈ R be given. The (inhomo-
geneous) Besov space Br
p,q(R; X) consists of all X-valued tempered distributions
Xk=−∞
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
7
f ∈ S′(R; X) such that
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=0(cid:13)(cid:13)(cid:13)ℓq
p,q(R;X). Then Br
p,q(R; X) is a Banach space such that
p,q(R; X), and a different choice of ψ leads to an equivalent norm on
< ∞,
endowed with the norm kfkBr
S(R; X) ⊆ Br
Br
p,q(R; X).
For n ∈ N and p ∈ [1,∞] the Sobolev space
Wn,p(R; X) :=nf ∈ Lp(R; X) f (k) ∈ Lp(R; X) for all 1 ≤ k ≤ no ,
is a Banach space when endowed with the norm
(f ∈ Wn,p(R; X)).
kfkn,p := kfkWn,p(X) := kfkp + kf (n)kp
In the case X = C we simply write Wn,p(R) = W1,p(R; C).
The following lemma is equation (5.9) in [1]. The fact that the constant C does
not depend on the particular Banach space follows from a direct sum argument.
Lemma 2.5. Let θ ∈ (0, 1), p ∈ [1,∞), q ∈ [1,∞] and n ∈ N. Then there exists a
constant C > 0 such that, for any Banach space X, (Lp(X), Wn,p(X))θ,q = Bnθ
p,q(X)
with
1
C kfkBnθ
p,q (X) ≤ kfk(Lp(X),Wn,p(X))θ,q ≤ C kfkBnθ
p,q (X)
(f ∈ Bnθ
p,q(X)).
Let m ∈ L∞(R;L(X)), p, q ∈ [1,∞] and r ∈ R. We say that m is a bounded
p,q(X) →
p,q(X) such that
p,q(X) if there is a unique bounded operator Tm : Br
Fourier multiplier on Br
Br
(2.4)
Tm(f ) = F −1 (m · F f )
for all f ∈ S(X). Each µ ∈ M(R) induces a bounded Fourier multiplier F µ with
(2.5)
TF µ(f ) = Lµ(f ) := µ ∗ f
(f ∈ S(X)).
The main result about Fourier multipliers on Besov spaces that we use is the fol-
lowing, Corollary 4.15 from [9].
Proposition 2.6. There exists a constant C ≥ 0 such that the following holds.
Let X be a Banach space, p, q ∈ [1,∞] and r ∈ R. If m : R → C is such that
ϕkm ∈ B1/2
2,1 (R; C) for all k ∈ N0, and
M := sup
k∈N0
a>0k(ϕkm)(a·)kB1/2
inf
2,1 (R;C) < ∞,
then m is a bounded Fourier multiplier on Br
p,q(X) with kTmkL(Br
p,q(X)) ≤ CM .
Corollary 2.7. There exists a constant C ≥ 0 such that for all Banach spaces X,
p, q ∈ [1,∞], r ∈ R and all m ∈ C1(R; C) with
m is a bounded Fourier multiplier on Br
N := sup
s∈R m(s) + (1 + s)m′(s) < ∞,
p,q(X) with kTmkL(Br
p,q(X)) ≤ CN .
Proof. This follows as in [9, Corollary 4.11]. See also [9, Remark 4.16].
(cid:3)
8
MARKUS HAASE AND JAN ROZENDAAL
3. Transference principles
3.1. Unbounded groups. We first establish an interpolation version of the trans-
ference principle for unbounded groups from [12]. Note that, for each µ ∈ M(R) and
p ∈ [1,∞], the convolution operator Lµ from (2.5) extends to a bounded operator
on Lp(X), by Young's inequality. For ω ≥ 0 and µ ∈ Mω(R) let µω ∈ M(R) be
given by µω(ds) := cosh(ωs)µ(ds).
Proposition 3.1. Let 0 ≤ ω0 < ω, θ ∈ (0, 1), p ∈ [1,∞) and q ∈ [1,∞]. Then
there exists a constant C ≥ 0 such that the following holds. If −iA generates a
C0-group (U (s))s∈R on a Banach space 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)θ,q ≤ CM 2 kLµωkL(Bθ
p,q(X)) kxkθ,q
for all µ ∈ Mω(R) and x ∈ DA(θ, q).
Proof. Let µ ∈ Mω(R) be given and let Uµ be as in (2.2). By the proof of Theorem
3.2 in [12], we can factorize Uµ as Uµ = P ◦ Lµω ◦ ι, where
• ι : X → Lp(X) is given by
ιx(s) := ψ(−s)U (−s)x
(x ∈ X, s ∈ R),
with
for α > ω fixed.
ψ(s) :=
1
cosh(αs)
(s ∈ R)
• P : Lp(X) → X is given by
P f :=ZR
ϕ(s)U (s)f (s) ds
(f ∈ Lp(X)),
with
ϕ(s) :=
√8ω
π
cosh(ωs)
cosh(2ωs)
(s ∈ R).
Then, using Holder's inequality,
(3.1)
kιkL(X,Lp(X)) ≤ M kψ cosh(ω0·)kp ,
kPkL(Lp(X),X) ≤ M kϕ cosh(ω0·)kp′ .
(3.2)
We claim that ι : D(A) → W1,p(X) and P : W1,p(X) → D(A) are well-defined and
bounded. To prove this claim, first let x ∈ D(A). Then ιx ∈ C1(R) with
(ιx)′(s) = −ψ′(−s)U (−s)x + iψ(−s)U (−s)Ax
tanh(αs)
cosh(αs)
for all s ∈ R. Hence (ιx)′ ∈ Lp(R) with
= −α
U (−s)x + i
cosh(αs)
1
k(ιx)′kp ≤ αM ktanhkL∞(R)(cid:13)(cid:13)(cid:13)(cid:13)
cosh(ω0·)
cosh(α·) (cid:13)(cid:13)(cid:13)(cid:13)pkxkX + M(cid:13)(cid:13)(cid:13)(cid:13)
U (−s)Ax
cosh(ω0·)
cosh(α·) (cid:13)(cid:13)(cid:13)(cid:13)pkAxkX .
This shows that ι : D(A) → W1,p(X) is bounded. To prove the claim for P , fix
f ∈ S(X) and note that
kιxk1,p ≤ M (αktanhkL∞(R) + 1)(cid:13)(cid:13)(cid:13)(cid:13)
(U (h) − I)P f =ZR
U (s)
1
h
cosh(α·) (cid:13)(cid:13)(cid:13)(cid:13)pkxkD(A) .
ϕ(s − h)f (s − h) − ϕ(s)f (s)
ds
h
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
9
Combining this with (3.1) implies that ιx ∈ W1,p(R) with
cosh(ω0·)
(3.3)
for h > 0. The latter expression converges to −RR U (s)(ϕf )′(s) ds ∈ X as h → 0,
by the dominated convergence theorem. Hence P f ∈ D(A) with
AP f = lim
h→0
1
h
(U (h) − I)P f = −ZR
U (s)(ϕ′(s)f (s) + ϕ(s)f ′(s)) ds.
Another application of Holder's inequality yields
kAP fkX ≤ M kϕ′ cosh(ω0·)kp′ kfkp + M kϕ cosh(ω0·)kp′ kf ′kp .
Combining this with (3.2) implies
(3.4)
kP fkD(A) ≤ M(cid:16)kϕ cosh(ω0·)kp′ + kϕ′ cosh(ω0·)kp′(cid:17)kfk1,p .
As S(X) is dense in W1,p(X), P : W1,p(X) → D(A) is bounded.
Since Lµω ∈ L(W1,p(X)), we can factorize Uµ ∈ L(D(A)) as Uµ = P ◦ Lµω ◦ ι
via bounded maps through W1,p(X). Applying the real interpolation method with
parameters θ and q to the two factorizations of Uµ, through Lp(X) respectively
W1,p(X), yields the commutative diagram of bounded maps
Lµω−−−−→ (Lp(X), W1,p(X))θ,q
(Lp(X), W1,p(X))θ,q
ι
x
DA(θ, q)
Uµ−−−−→
P
y
DA(θ, q)
Finally, estimate the norms of ι and P in this diagram by applying (1.5) to (3.1)
and (3.3) respectively (3.2) and (3.4). This yields
kUµkL(DA(θ,q)) ≤ C′M 2 kLµkL((Lp(X),W1,p(X))θ,q )
(3.5)
for a constant C′ ≥ 0 independent of µ. Now Lemma 2.5 concludes the proof. (cid:3)
3.2. Bounded groups. In this section we establish a version of the classical trans-
ference principle from [5] on interpolation spaces, already stated in the Introduction
as Proposition 1.1. In the proof we use the convention 1/∞ := 0.
Proposition 3.2. Let θ ∈ (0, 1), p ∈ [1,∞) and q ∈ [1,∞]. Then there exists a
constant C ≥ 0 such that the following holds. If −iA generate a C0-group (U (s))s∈R
on a Banach space X with M := sups∈R kU (s)k < ∞, then
(3.6)
p,q (X)) kxkθ,q
ZR
(cid:13)(cid:13)(cid:13)(cid:13)
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)θ,q ≤ CM 2 kLµkL(Bθ
for all µ ∈ M(R) and x ∈ DA(θ, q).
10
MARKUS HAASE AND JAN ROZENDAAL
Proof. First note that it suffices to establish (3.6) for measures with compact sup-
port. Indeed, approximating by measures with compact support then extends (3.6)
to all µ ∈ M(R). So fix N > 0 and let µ ∈ M(R) be such that supp(µ) ⊆ [−N, N ].
We will factorize Uµ using the abstract transference principle from [13, Section 2].
To this end, let ρ ∈ C∞(R) be defined by
ρ(s) :=( c1 exp(cid:16) 1
0
s2−1(cid:17) s < 1
s ≥ 1
,
where c1 ≥ 0 is such that RR ρ(s) ds = 1. Fix α, β > 0 and define σ(s) := 1
α ρ(cid:0) s
α(cid:1)
for s ∈ R, and
σ ∗ 1[−(α+β),α+β].
ψ := σ ∗ 1[−(N +3α+β),N +3α+β]
2(α + β)
ϕ :=
and
1
Then ψ, ϕ ∈ C∞(R) are such that supp(ϕ) ⊆ [−(2α + β), 2α + β],
ψ ≡ 1 on [−(2α + N + β), 2α + N + β]
Z 2α+β
Hence ψ ∗ ϕ ≡ 1 on [−N, N ]. Let ι : X → Lp(X) be given by
(x ∈ X, s ∈ R),
ιx(s) := ψ(−s)U (−s)x
and
−(2α+β)
and P : Lp(X) → X by
P f :=ZR
ϕ(s)U (s)f (s) ds
(f ∈ Lp(X)).
ϕ(s) ds = 1.
Proposition 2.3 in [13] yields the factorization Uµ = P ◦ Lµ ◦ ι, where we use that
(ψ ∗ ϕ)µ = µ. By Holder's inequality,
kιkL(X,Lp(X)) ≤ M kψkp
(3.7)
Moreover, ι : D(A) → W1,p(X) and P : W1,p(X) → D(A) are bounded with
(3.8)
kPkL(Lp(X),X) ≤ M kϕkp′
and
and
kPkL(W1,p(X),D(A)) ≤ M kϕk1,p′ .
kιkL(D(A),W1,p(X)) ≤ M kψk1,p
This follows by arguments almost identical to those in the proof of Proposition 3.1.
Applying the real interpolation method with parameters θ and q to the two factor-
izations of Uµ, through Lp(X) and W1,p(X), produces the commutative diagram
of bounded maps
(Lp(X), W1,p(X))θ,q
Lµ−−−−→ (Lp(X), W1,p(X))θ,q
ι
x
DA(θ, q)
P
y
DA(θ, q)
Uµ−−−−→
Use (1.5) on (3.7) and (3.8) to estimate the norms of ι and P in this factorization
as kιk ≤ M kψk1,p and kPk ≤ M kϕk1,p′. This yields
(3.9)
To determine kψk1,p and kϕk1,p′ , note that
kUµkL(DA(θ,q)) ≤ M 2 kψk1,p kϕk1,p′ kLµkL((Lp(X),W1,p(X))θ,q) .
kψkp ≤ kσk1(cid:13)(cid:13)1[−(N +3α+β),N +3α+β](cid:13)(cid:13)p = (2(N + 3α + β))1/p,
2(α + β) kσk1(cid:13)(cid:13)1[−(α+β),α+β](cid:13)(cid:13)p′ = (2(α + β))−1/p,
kϕkp′ ≤
1
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
11
by Young's inequality. Since σ is an even function that is decreasing on [0, α] and
supported on [−α, α], its derivative satisfies
kσ′k1 = −2Z α
0
σ′(s) ds = 2(σ(0) − σ(α)) =
2ρ(0)
α
.
Let c2 := 2ρ(0). Another application of Young's inequality yields
1
kψ′kp ≤ kσ′k1(cid:13)(cid:13)1[−(N +3α+β),N +3α+β](cid:13)(cid:13)p =
2(α + β) kσ′k1(cid:13)(cid:13)1[−(α+β),α+β](cid:13)(cid:13)p =
kϕ′kp ≤
(cid:19)1/p
α(cid:17)2(cid:18) N + 3α + β
kUµkL(DA(θ,q)) ≤ M 2(cid:16)1 +
α + β
c2
Hence (3.9) becomes
Taking the infimum over α and β yields
c2
α
c2
α
(2(N + 3α + β))1/p,
(2(α + β))−1/p.
kLµkL((Lp(X),W1,p(X))θ,q) .
(3.10)
kUµkL(DA(θ,q)) ≤ M 2 kLµkL((Lp(X),W1,p(X))θ,q) .
Lemma 2.5 now establishes (3.6) and concludes the proof.
(cid:3)
Remark 3.3. Note that the constant C in Proposition 3.2 comes only from the
equivalence of the norms on (Lp(X), W1,p(X))θ,q and Bθ
p,q(X), whereas in Propo-
sition 3.1 a constant is present which is inherent to the transference method.
Remark 3.4. Let p ∈ [1,∞) and let (U (s))s∈R ⊆ L(Lp(C)) be the shift group
given by U (s)f (t) := f (t + s) for f ∈ Lp(C), s ∈ R and almost all t ∈ R. Then
(U (s)))s∈R is generated by −iA, where Af := if ′ for f ∈ D(A) = W1,p(C). Hence
DA(θ, q) = (Lp(C), W1,p(C))θ,q for θ ∈ (0, 1) and q ∈ [1,∞]. Moreover, for µ ∈
M(R) and f ∈ Lp(C),
ZR
U (s)f dµ(s) = µ ∗ f = Lµ(f ).
Hence, with Uµ as in (2.2),
kUµkL(DA(θ,q)) = kLµkL((Lp(C),W1,p(C))θ,q ) .
This shows that (3.10) is sharp in general, up to possibly a change of constant. By
Lemma 2.5, the same holds for (3.6).
Corollary 2.7 yields the following result, Corollary 1.2 from the Introduction.
Corollary 3.5. Let θ ∈ (0, 1) and q ∈ [1,∞]. Then there exists a constant C ≥ 0
such that the following holds. Let −iA generate a C0-group (U (s))s∈R on a Banach
space X with M := sups∈R kU (s)k < ∞, and let µ ∈ M(R) be such that F µ ∈ C1(R)
with sups∈R(1 + s)(F µ)′(s) < ∞. Then
(cid:13)(cid:13)(cid:13)(cid:13)
ZR
U (s)x µ(ds)(cid:13)(cid:13)(cid:13)(cid:13)θ,q ≤ CM 2 sup
for all x ∈ DA(θ, q).
s∈R F µ(s) + (1 + s)(F µ)′(s) kxkθ,q
12
MARKUS HAASE AND JAN ROZENDAAL
Remark 3.6. To obtain Corollary 3.5 we used Corollary 2.7, but there are other
ways to verify the conditions of Proposition 2.6, for instance Hormander type as-
sumptions, cf. [9, pp. 47-49]. More generally, one can define a norm on the space of
all bounded Fourier multipliers m on Br
p,q(X)) := kTmkL(Br
p,q(X)),
with Tm as in (2.4). Proposition 3.2 yields kUµkL(DA(θ,q)) ≤ C kF µkM(Br
p,q(X)),
which cannot be improved in general, cf. Remark 3.4.
p,q(X) by kmkM(Br
Remark 3.7. If X is a UMD space then (1.1) and the vector-valued Mikhlin
multiplier theorem [10, Theorem E.6.2 b] yield an estimate
(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 kxkX sup
s∈R F µ(s) + s(F µ)′(s)
for all x ∈ X. Corollary 3.5 then follows from (1.5), and moreover singularities of
(F µ)′ at zero are allowed. However, in our setting of general Banach spaces, the
inhomogeneity of the Besov space Br
p,q(X) implies that a condition at zero on the
multiplier is needed to deal with the term ϕ0m in Proposition 2.6.
Remark 3.8. Letting f := F µ, Corollary 3.5 yields an estimate
(3.11)
kf (Aθ,q)k ≤ C sup
s∈R f (s) + (1 + s)f ′(s).
This is a functional calculus statement for Aθ,q involving functions on the real line.
One may now ask to which functions f on the real line the definition of f (Aθ,q) can
be extended in a sensible manner such that (3.11) holds. We can take the closure of
the Fourier transforms of measures in the space consisting of all functions f ∈ C1(R)
for which sups∈R f (s) + (1 + s)f ′(s) is finite, or approximate by holomorphic
functions as in [18, Lemma 4.15], using Theorem 4.1. This will yield a definition
of f (Aθ,q) for a class of functions on the real line and a bound as in (3.11), but
the question then remains how this definition relates to other known extensions
of functional calculi. In the present article we restrict ourselves to results about
holomorphic functional calculi.
4. Functional calculus results
We now use the theory established in the previous sections to prove our main
functional calculus result, Theorem 1.3. Recall the definition of the analytic Mikhlin
algebra H∞
1 (Stω) from (1.2).
Theorem 4.1. Let −iA be the generator of a C0-group (U (s))s∈R on a Banach
space X and let θ ∈ (0, 1), q ∈ [1,∞] and ω > θ(U ) be given. Then there exists a
constant C ≥ 0 such that f (Aθ,q) ∈ L(DA(θ, q)) with
kf (Aθ,q)kL(DA(θ,q)) ≤ C kfkH∞
If (U (s))s∈R is uniformly bounded then C can be chosen
for all f ∈ H∞
1 (Stω).
independent of ω > 0.
Proof. First consider f ∈ H∞
1 (Stω) ∩ E(Stω) and fix α ∈ (θ(U ), ω) and p ∈ [1,∞).
By Lemma 2.1 there exists µ ∈ Mα(R) such that f = F µ. By Lemmas 2.1 and 2.2
and Proposition 3.1,
1 (Stω )
(4.1)
kf (Aθ,q)k = k(Uµ)θ,qk ≤ C1 kLµαkL(Bθ
p,q(X)) = C1 kTF µαkL(Bθ
p,q(X))
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
13
for some constant C1 ≥ 0, where TF µα is as in (2.4). Since
F µα(s) =
f (s + iα) + f (s − iα)
2
(s ∈ R),
Corollary 2.7 yields a constant C2 ≥ 0 such that
(4.2)
kf (Aθ,q)k ≤ C2 sup
s∈R F µα(s) + (1 + s)(F µα)′(s) ≤ C2 kfkH∞
1 (Stω ) .
For general f ∈ H∞
1 (Stω) first assume that q < ∞. By part b) of Lemma 2.2,
D(Aθ,q) is dense in DA(θ, q). Let τk(z) := −k2(ik − z)−2 for k ∈ N with k > ω and
z ∈ Stω. Then τk, f τk ∈ H∞
k kf τkkH∞
sup
1 (Stω) ∩ E(Stω),
1 (Stω ) ≤ kfkH∞
1 (Stω ) < ∞
k kτkkH∞
1 (Stω) sup
and f τk(z) → f (z) as k → ∞, for all z ∈ Stω. Now (4.2) yields
kf τk(Aθ,q)k ≤ C2 kf τkkH∞
1 (Stω) ≤ C kfkH∞
1 (Stω)
for some C ≥ 0. Hence the Convergence Lemma 2.4 implies f (A) ∈ L(X) and
(4.3)
Finally, for q = ∞ the Reiteration Theorem [4, Theorem 3.5.3] yields
kf (Aθ,q)k ≤ C kfkH∞
1 (Stω) .
DA(θ,∞) = (DA(θ1, 1), DA(θ2, 1))θ3,∞
with equivalence of norms, where θ1, θ2, θ3 ∈ (0, 1) are such that θ1 6= θ2 and
θ1(1 − θ3) + θ2θ3 = θ. Combining (4.3) and (1.5) concludes the proof of the first
statement.
In the case where (U (s))s∈R is uniformly bounded, use Proposition 3.2 instead
of 3.1 in (4.1) to obtain
kf (Aθ,q)k ≤ C1 kTF µkL(Bθ
p,q(X))
for all f ∈ H∞
of the proof is the same as before.
1 (Stω)∩E(Stω) and some constant C1 ≥ 0 independent of ω. The rest
(cid:3)
Remark 4.2. Compare Theorem 4.1 with Theorem 3.6 in [12]. There an estimate
(4.4)
kf (A)kL(X) ≤ C sup
z∈Stωf (z) + zf ′(z)
is obtained when the underlying space X is a UMD space, and the constant C
is independent of ω when the group in question is uniformly bounded. Theo-
rem 4.1 follows from (4.4) by interpolation, and this seems to yield a stronger
result since the term supz∈Stωf ′(z) does not appear in (4.4). In fact, the norms
supz∈Stωf (z) +zf ′(z) and kfkH∞
1 (Stω ) are equivalent, since 0 ∈ Stω for all ω > 0.
So for generators of unbounded groups (4.4) does not yield an essentially better
estimate than Theorem 4.1. This is different for generators of uniformly bounded
groups, since the norm equivalence of supz∈Stωf (z) +zf ′(z) and kfkH∞
1 (Stω ) fails
as ω ↓ 0. Hence for generators of uniformly bounded groups (4.4) yields a strictly
stronger result on DA(θ, q) than Theorem 4.1.
Remark 4.3. Let λ ∈ C with Re(λ) > ω. By [10, Corollary 6.6.3], D((λ− iA)α) ⊆
DA(α,∞) for each α ∈ (0, 1). Hence Theorem 4.1 yields f (A)(λ − iA)−α ∈ L(X)
for all ω > θ(U ), f ∈ H∞
1 (Stω) and α > 0. However, this already follows from [3,
Proposition 8.2.3] in a similar manner as in [15, Remark 5.2]. Moreover, using
arguments as in [15, Remark 3.9], [3, Proposition 8.2.3] already implies that f (A) :
14
MARKUS HAASE AND JAN ROZENDAAL
DA(θ, q) → DA(θ′, q′) is bounded for all θ′ < θ and q, q′ ∈ [1,∞]. The improvement
that Theorem 4.1 provides lies in going from θ′ < θ to θ′ = θ.
Remark 4.4. As already noted in Remark 3.6, we could have used Fourier mul-
tiplier results on Besov spaces other than Corollary 2.7. These lead to statements
about the boundedness of functional calculi for other function algebras.
For ϕ ∈ (0, π) define
(4.5)
Sϕ := {z ∈ C arg(z) < ϕ} ,
and for ψ ∈ (0, π/2) and ω > 0,
Σψ := Sψ ∪ −Sψ,
Vψ,ω := Stω ∪ Σψ.
Lemma 4.5. Let ω > ω′ > 0 and ψ ∈ (0, π/2). Then H∞(Vω,ψ) is continuously
embedded in H∞
1 (Stω′).
Proof. This follows in a straightforward manner from Lemma 4.5 in [12].
(cid:3)
Corollary 4.6. Let −iA be the generator of a C0-group (U (s))s∈R on a Banach
space X and let θ ∈ (0, 1) and q ∈ [1,∞]. Then Aθ,q has a bounded H∞(Vω,ψ)-
calculus for all ω > θ(U ) and ψ ∈ (0, π/2).
So far we have considered functional calculus on interpolation spaces for the
couple (X, D(A)). The next corollary extends our results to other interpolation
couples.
Corollary 4.7. Let −iA be the generator of a C0-group (U (s))s∈R on a Banach
space X and let θ ∈ (0, 1), q ∈ [1,∞] and m, n ∈ N0 with m 6= n. Then the part
of A in (D(Am), D(An))θ,q has a bounded H∞
1 (Stω)-calculus for all ω > θ(U ). If
(U (s))s∈R is uniformly bounded then the constant bounding the calculus is indepen-
dent of ω > 0.
Proof. First note that since
(D(Am), D(An))θ,q = (D(An), D(Am))1−θ,q
by [4, Theorem 3.4.1], we may assume that m < n. Using the similarity transform
R(λ, A)m : X → D(Am), it suffices to let m = 0. Suppose that nθ /∈ N. By Lemma
3.1.3 and Proposition 3.1.8 in [19],
(X, D(An))θ,q = (D(Ak), D(Ak+1))θ ′,q
for some k ∈ N0 and θ′ ∈ (0, 1). Another similarity transform shows that we can
let k = 0. Now Theorem 4.1 yields the statement.
If k := nθ ∈ N, the Reiteration Theorem [4, Theorem 3.5.3] yields
(X, D(An))θ,q =(cid:0)(D(Ak−1), D(Ak))1/2,q, (D(Ak), D(Ak+1))1/2,q(cid:1)1/2,q .
By what we have already shown and (1.5), this concludes the proof.
(cid:3)
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
15
5. Additional results
We now deduce several applications of Theorem 4.1. Corollary 4.7 can be applied
in this section to yield results for other interpolation couples.
We first state a proposition about the convergence of certain principal value
integrals, an interpolation version of [12, Theorem 4.4] on general Banach spaces.
If g ∈ L1[−1, 1] is even then by PV − g(s)/s we mean the distribution defined by
hPV − g(s)/s, ϕi := lim
ǫց0Zǫ≤s≤1
g(s)ϕ(s)
ds
s
for ϕ ∈ C∞(R) compactly supported. By BV[−1, 1] we denote the functions of
bounded variation on [−1, 1].
Proposition 5.1. Let −iA be the generator of a C0-group (U (s))s∈R on a Banach
space X. Let g ∈ BV[−1, 1] be even and set f := F (PV − g(s)/s). Then f (Aθ,q) ∈
L(DA(θ, q)) and
(5.1)
ǫց0Zǫ≤s≤1
for all θ ∈ (0, 1), q ∈ [1,∞) and x ∈ DA(θ, q).
Proof. By [12, Lemma 4.3], f ∈ H∞
1 (Stω) for all ω > 0. Theorem 4.1 now yields
the first statement. For (5.1) we may let q < ∞, since DA(θ,∞) ⊆ DA(θ′, 1) for
θ′ < θ [19, Proposition 1.1.4]. Now use the Convergence Lemma as in the proof
of [12, Theorem 4.4].
(cid:3)
f (A)x = lim
g(s)U (s)x
ds
s
Remark 5.2. Convergence in (5.1) takes place in DA(θ, q) for q < ∞. For q = ∞
the limit and the integral converge in X and in DA(θ′, q′) for θ′ < θ and q′ ∈ [1,∞).
Compare with Remark 2.3.
5.1. Results for sectorial operators and cosine functions. An operator A on
a Banach space X is sectorial of angle ϕ ∈ (0, π) if σ(A) ⊆ Sϕ, where Sϕ is as in
(4.5), and if sup{kzR(z, A)k ψ ∈ C \ Sψ} < ∞ for all ψ ∈ (ϕ, π). A functional
calculus for sectorial operators can be constructed by a method similar to the one
used for strip-type operators. For details see [10, Chapter 2].
If A is an injective sectorial operator of angle ϕ ∈ (0, π) then log(A) is defined,
as is f (A) for all f ∈ H∞(Sψ) and ψ ∈ (ϕ, π). A sectorial operator A has bounded
imaginary powers if A is injective and if −i log(A) is the generator of a C0-group
(U (s))s∈R on X. Then U (s) = A−is for all s ∈ R, and we write A ∈ BIP(X).
Moreover, A is sectorial of angle θA := θ(U ), by [10, Corollary 4.3.4].
log(Sψ) to be the unital Banach algebra of all f ∈ H∞(Sψ)
For ψ ∈ (0, π) define H∞
for which
kfkH∞
log(Sψ) := sup
z∈Sψf (z) + (1 + log(z))zf ′(z) < ∞,
log(Sψ).
endowed with the norm k·kH∞
Proposition 5.3. Let X be a Banach space and A ∈ BIP(X) such that θA < π.
Let θ ∈ (0, 1) and q ∈ [1,∞]. Set Y := (X, D(log(A)))θ,q. Then AY has a bounded
H∞
bounding the calculus is independent of ψ > 0.
log(Sψ)-calculus on Y for all ψ ∈ (θA, π). If sups∈R(cid:13)(cid:13)Ais(cid:13)(cid:13) < ∞ then the constant
16
MARKUS HAASE AND JAN ROZENDAAL
Proof. Let ψ ∈ (θA, π) be given and note that f 7→ f ◦ log is an isometric algebra
isomorphism H∞
log(Sψ). By Lemma 2.2 as well as Theorem 4.2.4 and
Proposition 6.1.2 from [10],
1 (Stψ) → H∞
f (log(A)Y ) = f (log(A))Y = (f ◦ log)(A)Y = (f ◦ log)(AY )
1 (Stψ). Now Theorem 4.1 concludes the proof.
for all f ∈ H∞
Remark 5.4. Let A be an injective sectoral operator of angle ϕ ∈ (0, π), and
let α > 0, θ ∈ (0, 1) and q ∈ [1,∞]. By [10, Corollary 6.6.3], a special case of
which was proved by Dore [7, Theorem 3.2], the part of A in (X, D(Aα)∩ R(Aα))θ,q
has a bounded H∞(Sψ)-calculus for all ψ ∈ (ϕ, π). Here R(A) is the range of A.
By [10, Corollary 6.6.3] and because log(A)Aαθ(1 + A)−2αθ ∈ L(X),
(cid:3)
(X, D(Aα) ∩ R(Aα))θ,q ⊆ (X, D(Aα))θ,q ⊆ D(Aαθ) ⊆ D(log(A)),
and in general D(log(A)) is strictly included in (X, D(log(A)))θ ′,q′ for all θ′ ∈ (0, 1)
and q′ ∈ [1,∞]. Hence the result of Dore does not imply Proposition 5.3.
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)
kR(λ, A)k ≤
Mω′
pλ(cid:16)Im(√λ) − ω′(cid:17)
1 (Πω) :=(cid:26)f ∈ H∞(Πω)(cid:12)(cid:12)(cid:12)(cid:12)
for all λ /∈ Πω′. For such operators there is a natural functional calculus, as before,
and a version of Lemma 2.2 holds. For details see [14]. For ω > 0 let
z∈Πωf (z) + (1 + z)f ′(z) < ∞(cid:27) ,
1 (Πω ) := sup
kfkH∞
H∞
a Banach algebra when endowed with the norm k·kH∞
Proposition 5.5. Let −A be the generator of a cosine function Cos on a Banach
space X and let θ ∈ (0, 1), q ∈ [1,∞]. Then the part Aθ,q of A in DA(θ, q) has
a bounded H∞
1 (Πω)-calculus for all ω > θ(Cos). If sups∈R kCos(s)k < ∞ then the
constant bounding the calculus is independent of ω > 0.
1 (Πω ).
Proof. We mainly follow [12, Theorem 5.3], providing extra details where necessary.
There is a unique subspace V ⊆ X, the Kisy´nski space, such that the operator −iA,
A := i(cid:20) 0
−A 0 (cid:21) ,
IV
for all s, t ∈ R. Then
θ(Cos) := infnω ≥ 0(cid:12)(cid:12)(cid:12)∃M ≥ 0 : kCos(t)k ≤ M eωt for all t ∈ Ro < ∞.
The generator of a cosine function is the unique operator −A on X that satisfies
λR(λ2,−A) =Z ∞
0
e−λtCos(t) dt
for λ > θ(Cos). Then A is an operator of parabola-type ω = θ(Cos). This means
that σ(A) ⊆ Πω, where Πω := (cid:8)z2 z ∈ Stω(cid:9), and that for all ω′ > ω there exists
Mω′ ≥ 0 such that
FUNCTIONAL CALCULUS ON INTERPOLATION SPACES
17
with domain D(A) := D(A)× V , generates a C0-group (U (s))s∈R on X × V . More-
over, θ(Cos) = θ(U ) [11, Theorem 6.2].
1 (Πω) if and only
1 (Πω ). Moreover,
Let ω > θ(Cos). Then f ∈ H∞(Πω) is an element of H∞
1 (Stω ) ≤ 4 kfkH∞
1 (Stω), with kgkH∞
if [z 7→ g(z) := f (z2)] ∈ H∞
f (A) ⊕ f (AV ) = g(A) and
for all f ∈ H∞
concludes the proof.
f (Aθ,q) ⊕ f (AV ) = (f (A) ⊕ f (AV ))Y = g(A)Y = g(AY )
1 (Πω), where Y := DA(θ, q) = DA(θ, q) × V . Hence Theorem 4.1
(cid:3)
Acknowledgements. We would like to thank our colleagues for interesting discus-
sions and helpful ideas.
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, C.J.K. Batty, M. Hieber, and F. Neubrander. Vector-valued Laplace transforms
and Cauchy problems, volume 96 of Monographs in Mathematics. Birkhauser/Springer Basel
AG, Basel, second edition, 2011.
[4] J. Bergh and J. Lofstrom. Interpolation spaces. An introduction. Springer-Verlag, Berlin,
1976. Grundlehren der Mathematischen Wissenschaften, No. 223.
[5] E. Berkson, T.A. Gillespie, and P.S. Muhly. Generalized analyticity in UMD spaces. Ark.
Mat., 27(1):1 -- 14, 1989.
[6] G. Dore. H∞ functional calculus in real interpolation spaces. Studia Math., 137(2):161 -- 167,
1999.
[7] G. Dore. H∞ functional calculus in real interpolation spaces. II. Studia Math., 145(1):75 -- 83,
2001.
[8] 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.
[9] M. Girardi and L. Weis. Operator-valued Fourier multiplier theorems on Besov spaces. Math.
Nachr., 251:34 -- 51, 2003.
[10] M. Haase. The functional calculus for sectorial operators, volume 169 of Operator Theory:
Advances and Applications. Birkhauser Verlag, Basel, 2006.
[11] M. Haase. Functional calculus for groups and applications to evolution equations. J. Evol.
Equ., 7(3):529 -- 554, 2007.
[12] M. Haase. A transference principle for general groups and functional calculus on UMD spaces.
Math. Ann., 345(2):245 -- 265, 2009.
[13] M. Haase. Transference principles for semigroups and a theorem of Peller. J. Fun. Anal.,
261(10):2959 -- 2998, 2011.
[14] 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.
[15] M. Haase and J. Rozendaal. Functional calculus for semigroup generators via transference.
J. Funct. Anal., 265(12):3345 -- 3368, 2013.
[16] T. Hytonen. Fourier embeddings and Mihlin-type multiplier theorems. Math. Nachr.,
274/275:74 -- 103, 2004.
[17] N.J. Kalton and L. Weis. The H∞-calculus and sums of closed operators. Math. Ann.,
321(2):319 -- 345, 2001.
[18] C. Kriegler. Spectral multipliers, R-bounded homomorphisms,
and analytic diffu-
sion semigroups. PhD thesis, Karlsruhe Institute of Technology, 2009. Online at
http://digbib.ubka.uni-karlsruhe.de/volltexte/1000015866 .
18
MARKUS HAASE AND JAN ROZENDAAL
[19] 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.
Delft Institute of Applied Mathematics, Mekelweg 4, 2628CD Delft, The Nether-
lands
E-mail address: [email protected]
Delft Institute of Applied Mathematics, Mekelweg 4, 2628CD Delft, The Nether-
lands
E-mail address, corresponding author: [email protected]
|
1906.05124 | 1 | 1906 | 2019-06-12T13:14:22 | Amenability and harmonic $L^p$-functions on hypergroups | [
"math.FA"
] | Let $K$ be a locally compact hypergroup with a left invariant Haar measure. We show that the Liouville property and amenability are equivalent for $K$ when it is second countable. Suppose that $\sigma$ is a non-degenerate probability measure on $K$, we show that there is no non-trivial $\sigma$-harmonic function which is continuous and vanishing at infinity. Using this, we prove that the space $H_\sigma^p(K)$ of all $\sigma$-harmonic $L^p$-functions, is trivial for all $1\leq p<\infty$. Further, it is shown that $H_\sigma^\infty(K)$ contains only constant functions if and only if it is a subalgebra of $L^\infty(K)$. In the case where $\sigma$ is adapted and $K$ is compact, we show that $H_\sigma^p(K)={\mathbb C}1$ for all $1\leq p\leq\infty$. | math.FA | math | AMENABILITY AND HARMONIC Lp-FUNCTIONS ON
HYPERGROUPS
MEHDI NEMATI1 AND JILA SOHAEI2
Abstract. Let K be a locally compact hypergroup with a left invariant Haar
measure. We show that the Liouville property and amenability are equivalent
for K when it is second countable. Suppose that σ is a non-degenerate proba-
bility measure on K, we show that there is no non-trivial σ-harmonic function
which is continuous and vanishing at infinity. Using this, we prove that the
σ(K) of all σ-harmonic Lp-functions, is trivial for all 1 ≤ p < ∞. Fur-
space H p
ther, it is shown that H ∞
σ (K) contains only constant functions if and only if it
is a subalgebra of L∞(K). In the case where σ is adapted and K is compact,
we show that H p
σ(K) = C1 for all 1 ≤ p ≤ ∞.
9
1
0
2
n
u
J
2
1
]
.
A
F
h
t
a
m
[
1
v
4
2
1
5
0
.
6
0
9
1
:
v
i
X
r
a
1. Introduction
Let σ be a complex Borel measure on a locally compact group G. A Borel func-
tion f on G is called σ-harmonic if it satisfies the convolution equation σ ∗ f = f .
It is a well-known result of [4] that if G is abelian, then the only bounded contin-
uous σ-harmonic function are constant functions when the support of σ generates
a dense subgroup of G. Bounded harmonic functions have been investigated by
several authors for various kinds of groups, e.g., nilpotent groups and compact
groups [7, 8, 10, 11, 12]. Moreover, it was shown in [5] that for 1 ≤ p < ∞,
any σ-harmonic Lp-function associated to an adapted probability measure σ on a
locally compact group G is trivial. Harmonic functions on groups play important
roles in analysis, geometry and probability theory [6].
Motivated by these observations, bounded continuous harmonic functions on
nilpotent, [IN] and central hypergroups have been studied in [1, 2].
In what follows, K denotes a locally compact hypergroup with a left-invariant
Haar measure. The purpose of this paper is to obtain some insight into the
harmonic functions problem for the Lp-spaces, 1 ≤ p ≤ ∞, of K.
In Section 3, for given a complex Borel measure σ on K with kσk = 1, we
first show that there is a contractive projection from Lp(K), 1 < p ≤ ∞, onto
H p
σ(K) = {f ∈ Lp(K) : σ ∗ f = f }. We also show that K is necessarily amenable
if it has the Liouville property; that is, there exists a probability measure σ on
K such that all σ-harmonic L∞-functions on K are constant. Further, we prove
that a second countable hypergroup possesses the Liouville property if and only
if it is amenable.
In Section 4, for the case that σ is a non-degenerate probability measure on
K, we show that the space of all σ-harmonic functions which are continuous and
2010 Mathematics Subject Classification. 43A62, 43A15, 43A07, 45E10.
Key words and phrases. Amenability, hypergroup, harmonic function, Liouville property.
1
2
M. NEMATI AND J. SOHAEI
vanishing at infinity are trivial. Using this we prove that for 1 ≤ p < ∞, any σ-
harmonic Lp-function is trivial. For such a measure σ, we also prove that H ∞
σ (K)
is a subalgebra of L∞(K) if and only if H ∞
σ (K) = C1. In the case where σ is
adapted and K is compact, we show that H p
σ(K) = C1 for all 1 ≤ p ≤ ∞. These
extend the results for the group case in [5].
2. Preliminaries
Let K be a locally compact Hausdorff space. The space K is a hypergroup
if there exists a bilinear, associative, weakly continuous convolution ∗ on the
Banach space M(K) of all bounded regular complex valued Borel measures on
K, such that (M(K), ∗) is an algebra and satisfies, for x, y ∈ K,
(i) δx ∗ δy is a probability measure on K with compact support,
(ii) the mapping K × K → C(K) , (x, y) 7→ supp(δx ∗ δy) is continuous with
respect to the Michael topology on the space C(K) of nonvoid compact sets in
K,
(ii) the mapping K × K → M(K), (x, y) 7→ δx ∗ δy is continuous,
(ii) there is an identity e ∈ K with δe ∗ δx = δx ∗ δe = δx,
(iv)there is a continuous involution on K such that (δx ∗ δy) = δy ∗ δx and
e ∈ supp(δx ∗ δy) if and only if x = y. The image measure of µ ∈ M(K) under
such involution is denoted by µ.
Given a (complex) Borel function f on K and x, y ∈ K the left translation xf
and the right translation fy are defined by
xf (y) = fy(x) = ZK
f (t)d(δx ∗ δy)(t) = f (x ∗ y),
if the integral exists, where f (x ∗ y) = RK f d(δx ∗ δy). For a Borel function f on K
the Borel function f is defined by f (x) = f (x) for all x ∈ K. Given µ, ν ∈ M(K),
their convolution is given by
hµ ∗ ν, f i = ZK
f d(µ ∗ ν) = ZK ZK
f (x ∗ y) dµ(x) dν(y)
(f ∈ C0(K))
and kµ ∗ νk ≤ kµkνk which shows that µ ∗ ν ∈ M(K). Also for a measure
µ ∈ M(K) and a Borel function f on K, we define the convolutions µ ∗ f and
f ∗ µ by
f (y ∗ x) dµ(y),
f ∗ µ(x) = ZK
µ ∗ f (x) = ZK
if the integrals exist. Note that in this case (µ ∗ f )= f ∗ µ. Moreover, if f is in
Cb(K), the Banach space of bounded complex continuous functions on K, then
µ ∗ f and f ∗ µ are in Cb(K) with kµ ∗ f k∞ ≤ kµkkf k∞ and hµ ∗ ν, f i = hν, µ ∗ f i.
We refer the reader to [3] for details of hypergroups.
f (x ∗ y) dµ(y)
(x ∈ K),
3. Amenability and Liouville property
Throughout of this paper, let K be a locally compact hypergroup with a left-
invariant Haar measure ω; that is, a non-zero positive Radon measure on K such
AMENABILITY AND HARMONIC Lp-FUNCTIONS ON HYPERGROUPS
3
that
δx ∗ ω = ω (x ∈ K).
Let C0(K) be the Banach space of complex continuous functions on K vanishing
at infinity. Then its dual identifies, via the Riesz representation theorem, with
the space M(K). Let Lp(K) be the complex Lebesgue spaces with respect to ω,
for 1 ≤ p ≤ ∞. Given a Borel measure σ on a hypergroup K, a Borel function f
on K satisfying the convolution equation
σ ∗ f = f
is called σ-harmonic. For 1 ≤ p ≤ ∞ define H p
Lp-functions; that is, H p
and g at least one of which is σ-finite, define the convolution f ∗ g on K by
σ(K) to be the set of all σ-harmonic
σ(K) = {f ∈ Lp(K) : σ ∗ f = f }. For Borel functions f
(f ∗ g)(x) = ZK
f (y)g(y ∗ x)dω(y).
We commence with the following lemma whose proof is similar to those given in
[5]. For completeness, we present the argument here.
Lemma 3.1. Let σ ∈ M(K) with kσk = 1 and let 1 < p ≤ ∞. Then there
is a contractive projection Pσ : Lp(K) → Lp(K) with Pσ(Lp(K)) = H p
σ(K).
Moreover, if 1 < p, q < ∞ with 1
q = 1, then Pσ is the dual map of the
projection Pσ : Lq(K) → Lq(K).
p + 1
Proof. Let U be a free ultra-filter on N, and define Pσ : Lp(K) → Lp(K) by the
weak∗ limit
Pσ(f ) = lim
U
1
n
n
Xk=1
σk ∗ f,
where σk is the k-times convolution of σ with itself. It is easy to see that Pσ(f ) =
σ(K). Moreover, if f ∈ Lp(K), then it is easily verified that
f for all f ∈ H p
σ ∗ Pσ(f ) = Pσ(f ) and so Pσ(f ) ∈ H p
σ = Pσ and
Pσ(Lp(K)) = H p
σ(K). These show that P 2
σ(K).
Suppose now that 1 < p < ∞. Then it is not hard to check that σ ∗ Pσ(f ) =
Pσ(σ ∗ f ) for all f ∈ Lp(K). Therefore, for each g ∈ Lq(K), we have
hσ ∗ P ∗
σ (f ), gi = hP ∗
σ (f ), σ ∗ gi = hf, Pσ(σ ∗ g)i
= hf, σ ∗ Pσ(g)i = hf, Pσ(g)i
= hP ∗
σ (f ), gi.
This shows that σ ∗ P ∗
that σ ∗ P ∗
σ (g) = P ∗
g ∈ Lq(K), we have
σ (f ) = P ∗
σ (f ) for all f ∈ Lp(K). Similarly, we can show
σ (g) for all g ∈ Lq(K). Consequently, for each f ∈ Lp(K) and
hPσP ∗
σ (f ), gi = hPσP ∗
σ (g), f i = hP ∗
σ (g), f i
= hPσ(f ), gi.
This shows that P ∗
σ (f ) = PσP ∗
σ (f ) = Pσ(f ) for all f ∈ Lp(K), as required.
(cid:3)
4
M. NEMATI AND J. SOHAEI
Remark 3.2. Let σ ∈ M(K) with kσk = 1 and let 1 < p, q < ∞ be such that 1
1
q = 1. Then we have linear isometric isomorphisms H p
H q
Lp(K)/H q
σ(K) ∼= Lp(K)/H q
σ(K) ∋ f 7→ f + H q
σ(K)∗, where the first isometry is given by H p
p +
σ(K)⊥ ∼=
σ(K)⊥ ∈
σ(K)⊥. Indeed, for each f ∈ H p
σ(K), we have
kf k ≥ inf{kf + gk : g ∈ H q
σ(K)⊥} ≥ inf{kPσ(f + g)k : g ∈ H q
≥ inf{kf k : g ∈ H q
σ(K)⊥}
σ(K)⊥} = kf k.
Recall that the hypergroup K is called amenable if there exists a topological
left invariant mean on L∞(K); that is, there exists m ∈ L∞(K)∗ such that
A topological right invariant mean on L∞(K) is a functional m ∈ L∞(K)∗ such
kmk = m(1) = 1 and m(g∗f ) = (RK g dω)m(f ) for all f ∈ L∞(K) and g ∈ L1(K).
that kmk = m(1) = 1 and m(f ∗ g) = (RK g dω)m(f ) for all f ∈ L∞(K) and
g ∈ L1(K). It is known that the involution on L1(K) can be canonically extended
to a linear involution ⋆ on L1(K)∗∗; see [9, Chapter 2]. Clearly, m ∈ L∞(K)∗ is
a topological left invariant mean if and only if m⋆ is a topological right invariant
mean. Therefore, the existence of a topological right invariant mean on L∞(K)
is equivalent to K being amenable.
Theorem 3.3. Let K be a hypergroup with the Liouville property; that is, there
exists a probability measure σ on K such that H ∞
σ (K) = C1. Then K is amenable.
Proof. Let Pσ : L∞(K) → H ∞
σ (K) be the contractive projection as defined in
Lemma 3.1. Then there is a unique functional m ∈ L∞(K)∗ such that Pσ(f ) =
m(f )1 for all f ∈ L∞(K). Since σ ∗ (f ∗ g) = (σ ∗ f ) ∗ g for all f ∈ L∞(K) and
g ∈ L1(K), it follows that Pσ(f ∗ g) = Pσ(f ) ∗ g. Moreover, since the projection
Pσ is positive and Pσ(1) = 1, we conclude that kmk = m(1) = 1. This shows
that m is a topological right invariant mean on L∞(K), which implies that K is
amenable.
(cid:3)
For a a locally compact hypergroup K consider the closed two sided ideal
L1
0(K) = (cid:26)f ∈ L1(K) : ZK
f dω = 0(cid:27)
in L1(K) and for each σ ∈ M(K) let Jσ be the norm closure of {f − σ ∗ f : f ∈
L1(K)} in L1(K). It is well known that L1
0(K) has codimension one in L1(K) and
if σ is a probability measure, then Jσ ⊆ L1
0(K). Moreover, it is easy to see that
σ = {f ∈ L∞(K) : σ ∗ f = f } = H ∞
J ⊥
We have the following lemma whose proof is similar to those given in [14,
Lemma 1.1 and Remark 3, p.210] for locally compact groups. Thus, we omit the
proof.
σ (K) = (L1(K)/Jσ)∗
σ (K) and hence H ∞
Lemma 3.4. Let K be a locally compact hypergroup and S be a norm closed,
convex subsemigroup of probability measuers on K. Let I be a separable, closed
subspace of L1(K) such that
(i) Jσ ⊆ I for every σ ∈ S; and
AMENABILITY AND HARMONIC Lp-FUNCTIONS ON HYPERGROUPS
5
(ii) for each ε > 0 and g ∈ I there is σ ∈ S such that
d(g, Jσ) = inf{kf − gk : f ∈ Jσ} < ε.
Then there is σ ∈ S such that I = Jσ.
Corollary 3.5. Let K be a second countable locally compact hypergroup. Then
the following conditions are equivalent.
(i) K is amenable.
(ii) There exists a probability measure σ on K such that L1
(iii) K has the Liouville property.
0(K) = Jσ.
Proof. (i)⇒(ii). Suppose that K is amenable. Then by [13, Corollary 4.2], there
is a net (fα) in P1(K) := {f ∈ L1(K) : kf k1 = RK f dω = 1} such that
kfα ∗ f − fαk1 → 0
for all f ∈ P1(K). In particular, for each g ∈ L1
0(K) we have kfα ∗ gk1 → 0.
Moreover, fα ∗ g − g ∈ Jσα for all α, where σα = fαω. This shows that the
condition (ii) of Lemma 3.4 is satisfied. Since L1(K) is separable, we give that
L1
0(K) = Jσ for some probability measure σ on K.
(iii)⇒(i). This follows from Theorem 3.3.
(ii)⇔(iii). This follows from the inclusion Jσ ⊆ L1
0(K)⊥ = C1.
L1
0(K) with the fact that
(cid:3)
Proposition 3.6. Let σ be a probability measure on K and let 1 < p ≤ ∞. Then
H p
σ(K) is generated by its non-negative elements.
Proof. Suppose that f ∈ H p
σ(K). Then σ ∗ f = (σ ∗ f ) = f . This shows that
H p
σ(K) is self-adjoint and consequently is generated by its real function parts.
Now let f ∈ H p
σ(K) be a real function and let f = f+ − f−, where f+, f−
are non-negative functions in Lp(K). Since σ is positive, the projection Pσ, as
defined in Theorem 3.1, is positive. It follows that Pσ(f+), Pσ(f−) ∈ H p
σ(K) are
non-negative. Moreover, f = Pσ(f ) = Pσ(f+) − Pσ(f−), and this completes the
proof.
(cid:3)
4. Harmonic Lp-functions
Let µ be a complex Borel measure on a locally compact hypergroup K. We
say that µ is non-degenerate if
∞
K =
(suppµ)n =
[n=1
∞
[n=1
suppµn,
where µn is the n-fold convolution of µ and suppµn equals the closure of
(suppµ)n. If µ satisfies the weaker condition that
K =
∞
[n=1(cid:16)suppµ ∪ (suppµ)(cid:17)n
then we say that µ is adapted.
6
M. NEMATI AND J. SOHAEI
Remark 4.1. Let µ ∈ M(K). Then it is not hard to check that non-degeneracy
n=1hµn, hi > 0 for every non-zero h ∈ Cc(K)+, or
of µ is equivalent to that P∞
equivalently there exists n ∈ N such that
hµn, hi > 0.
Therefore, if f ∈ Cb(K)+ is non-zero, then we may find h ∈ Cc(K)+ such that
khk∞ = 1 and f h 6= 0. It follows that f h ∈ Cc(K)+ and f h ≤ f . Therefore,
there exists n ∈ N such that
0 < hµn, f hi ≤ hµn, f i.
Theorem 4.2. Let σ be a probability measure on K. Then the following state-
ments are equivalent.
(i) H ∞
(ii) H ∞
(iii) H ∞
σ (K) is a subalgebra of L∞(K).
σ (K) is a von Neumann subalgebra of L∞(K)..
σ (K) = {f ∈ L∞(K) : ∀x ∈ K, f (y ∗ x) = f (x), for σ−a.e. y ∈ K}.
σ (K) is a weak∗ closed operator system, (i) implies that H ∞
Proof. Since H ∞
σ (K)
is a von Neumann subalgebra of L∞(K). The implication (iii)⇒(i) is trivial. We
need to prove (ii)⇒(iii). Let x ∈ K and f ∈ H ∞
σ (K). Without loss of generality
assume that f is real valued. Since (f − f (x))2 ∈ H ∞
σ (K), it follows that
ZK
(f (y ∗ x) − f (x))2dσ(y) = σ ∗ (f − f (x))2(x)
= (f (x) − f (x))2 = 0.
This implies that f (y ∗ x) = f (x) for σ−almost every y ∈ K.
(cid:3)
For a hypergroup K, we denote by LUC(K) to be the Banach space of all
bounded left uniformly continuous complex functions on K, consisting of bounded
continuous function f on K such that the map K ∋ x 7→ fx ∈ Cb(K) is continu-
ous.
Lemma 4.3. Let σ ∈ M(K). Then H ∞
σ (K)∩LUC(K) is weak∗ dense in H ∞
σ (K).
Let (φα) be a bounded approximate identity for L1(K) and f ∈ H ∞
σ (K). Then
σ (K) ∩ LUC(K) for all α, by [13, Lemma 2.2]. Moreover, for each
f ∗ φα ∈ H ∞
g ∈ L1(K), we have
lim
α
hf ∗ φα, gi = lim
α
hf, g ∗ φαi = hf, gi.
Thus, H ∞
σ (K) ∩ LUC(K) is weak∗ dense in H ∞
σ (K).
Corollary 4.4. Let σ be a non-degenerate probability measure on K. Then the
following conditions are equivalent.
(i) H ∞
(ii) H ∞
σ (K) is a subalgebra of L∞(K).
σ (K) = C1.
Proof. Suppose that (i) holds. Given f ∈ H ∞
σ (K) ∩ LUC(K), by Theorem 4.2
for each n ∈ N, we have f (y) = f (e) for all y ∈ (suppσ)n. It follows from non-
degeneracy of σ and continuity of f that f is constant. Since H ∞
σ (K) ∩ LUC(K)
is weak∗ dense in H ∞
(cid:3)
σ (K), we give that H ∞
σ (K) = C1.
AMENABILITY AND HARMONIC Lp-FUNCTIONS ON HYPERGROUPS
7
A subspace X of Lp(K), 1 ≤ p ≤ ∞, is called left (resp. right) translation
invariant if xf ∈ X (resp. fx ∈ X) for all f ∈ X and x ∈ K. The subspace X is
called translation invariant if it is left and right translation invariant. It is easy to
check that for each σ ∈ M(K) the space H ∞
σ (K) is a right translation invariant
subspace of L∞(K). Since RK(gx)(t)f (t)dω(t) = 1
f ∈ L∞(K), g ∈ L1(K) and x ∈ K, it follows that Jσ is also right translation
invariant in L1(K), where ∆ is the modular function on K. We recall that L∞(K)
is naturally a Banach L1(K)-bimodule by the following module actions
∆(x) RK g(t)(fx)(t)dω(t) for all
hg · f, hi = hf, h ∗ gi,
hf · g, hi = hf, g ∗ hi
(f ∈ L∞(K), g, h ∈ L1(K)).
It is easily verified that
(g · f )(x) = ZK
g(y)(xf )(y)dω(y),
(f · g)(x) = ZK
g(y)(fx)(y)dω(y)
for all x ∈ K.
Proposition 4.5. Let σ ∈ M(K). Then the following conditions are equivalent.
σ (K) is translation invariant.
(i) H ∞
(ii) Jσ is translation invariant.
(iii) Jσ is an ideal in L1(K).
(v) H ∞
(iv) RK f (y ∗ x)dσ(y) = RK f (x ∗ y)dσ(y) for all f ∈ H ∞
σ (K) is a sub-L1(K)-bimodule of L∞(K).
Proof. (i)⇔(ii). Since RK(xg)(t)f (t)dω(t) = 1
L∞(K), g ∈ L1(K) and x ∈ K, it follows that H ∞
if and only if Jσ is.
∆(x) RK g(t)(xf )(t)dω(t) for all f ∈
σ (K) is left translation invariant
σ (K) ∩ LUC(K).
(ii)⇒(iii). It suffices to show that Jσ is a left ideal in L1(K). To prove this,
given g ∈ L1(K), h ∈ Jσ and f ∈ H ∞
σ (K), we have
hf, g ∗ hi = ZK
= ZK
f (x)(ZK
g(y)(ZK
g(y)h(y ∗ x)dω(y))dω(x)
f (x)(yh)(x)dω(x))dω(y) = 0,
which implies that g ∗ h ∈ Jσ.
(iii)⇒(ii). Let (φα) be a bounded approximate identity for L1(K) and let
g ∈ Jσ. Since ((xφα) ∗ g) =x (φα ∗ g) and xφα ∈ L1(K) for all α and x ∈ K, the
proof follows from the fact that φα ∗ g → g.
(i)⇒(iv). Suppose that f ∈ H ∞
σ (K) ∩ LUC(K). Since H ∞
σ (K) is left transla-
tion invariant, we obtain that
ZK
f (y ∗ x)dσ(y) = x(σ ∗ f )(e)
= (xf )(e) = (σ ∗ (xf ))(e)
= ZK
f (x ∗ y)dσ(y).
8
M. NEMATI AND J. SOHAEI
(iv)⇒(i). Suppose that f ∈ H ∞
σ (K) ∩ LUC(K), by right translation invariance of H ∞
σ (K) ∩ LUC(K) and x ∈ K. Then fx ∈
σ (K) ∩ LUC(K). More-
H ∞
over, for each y ∈ K, we have
(xf )(y) =x (σ ∗ f )(y) = ZK
= ZK
= ZK
= ZK
f (t ∗ x ∗ y)dσ(t)
(fy)(t ∗ x)dσ(t)
(fy)(x ∗ t)dσ(t)
f (x ∗ t ∗ y)dσ(t)
= (σ ∗ (xf ))(y).
This shows that H ∞
by weak∗ density of H ∞
(i)⇒(v). Let f ∈ H ∞
σ (K)∩LUC(K) and hence H ∞
σ (K) is left translation invariant
σ (K) ∩ LUC(K) in H ∞
σ (K) and g ∈ L1(K). As g · f = f ∗ g , we give that
σ (K).
g · f ∈ H ∞
σ (K). Moreover, by assumption tf ∈ H ∞
σ (K) for all t ∈ K. Thus,
(σ ∗ (f · g))(x) = ZK
(f · g)(y ∗ x)dσ(y)
= ZK ZK
= ZK ZK
= ZK
g(t)(fy∗x)(t)dσ(y)dω(t)
g(t)(tf )(y ∗ x)dσ(y)dω(t)
g(t)(tf )(x)dω(t) = (g · f )(x).
This implies that f · g ∈ H ∞
σ (K).
(v)⇒(iii). It suffices to show that Jσ is a left ideal in L1(K). Indeed, given
f ∈ H ∞
σ (K), g ∈ L1(K) and h ∈ Jσ, we have
hf, g ∗ hi = hf · g, hi = 0,
which yields that g ∗ h ∈ Jσ, as required.
(cid:3)
Remark 4.6. It is obvious that under each of above equivalent conditions in Propo-
sition 4.5, the quotient space L1(K)/Jσ is a Banach algebra. This implies that
σ (K)∗ = (L1(K)/Jσ)∗∗ is a Banach algebra with respect to the two Arens
H ∞
products.
Theorem 4.7. Let σ be a non-degenerate probability measure on K. Then every
bounded continuous σ-harmonic function on K vanishing at infinity is constant.
Proof. Let f ∈ H ∞
σ (K)∩C0(K) be real-valued. Without loss of generality assume
that kf k∞ = 1. Therefor, we can find a probability measure µ on K such that
kf k∞ = hµ, f i. If f 6= 1, then the function 1−f is also non-negative and non-zero
σ-harmonic function in Cb(K). It is well known from [3, Proposition 1.2.16] that
AMENABILITY AND HARMONIC Lp-FUNCTIONS ON HYPERGROUPS
9
µ ∗ (1 − f ) is a non-negative and bounded continuous function. Moreover,
ZK
µ ∗ (1 − f )dω(x) = µ(K)ZK
(1 − f )dω(x) > 0,
which implies that µ ∗ (1 − f ) is non-zero. Since σ is non-degenerate, by Remark
4.1, there exists n ∈ N such that
hσn, µ ∗ (1 − f )i > 0.
On the other hand, since f is σ-harmonic, σn ∗ f = f . Therefore,
hσn, µ ∗ (1 − f )i = hσn, (1 − µ ∗ f )i = 1 − hσn, µ ∗ f i
= 1 − h σn, f ∗ µi = 1 − h σn ∗ µ, f i
= 1 − hσn ∗ f, µi = 1 − hf, µi = 0
which is a contradiction. Since H ∞
elements, the proof is complete.
σ (K) ∩ C0(K) is generated by its non-negative
(cid:3)
Corollary 4.8. Let K be non-compact and let σ be a non-degenerate probability
measure on K. Then the sequence (cid:0) 1
Proof. Let σ0 be a weak∗ cluster point of ( 1
k=1 σk) in M(K). Then σ ∗σ0 = σ0,
which implies that σ0 is an idempotent probability measure in M(K). Moreover,
for each f ∈ C0(G) we have
k=1 σk(cid:1) is weak∗-convergent to 0.
n Pn
n Pn
σ ∗ (σ0 ∗ f ) = (σ ∗ σ0) ∗ f = σ0 ∗ f.
This shows that σ0 ∗ f is σ-harmonic and so it is constant by Theorem 4.7.
Since K is non-compact, we must have σ0 ∗ f = 0 for all f ∈ C0(K). This
in M(K). By weak∗ compactness of the unit ball of M(K) we conclude that
1
(cid:3)
implies that σ0 = 0. Thus, 0 is the only weak∗ cluster point of (cid:0) 1
n Pn
Corollary 4.9. The hypergroup K is compact if and only if there is a non-
degenerate idempotent probability measure on K.
k=1 σk(cid:1)
k=1 σk w∗
n Pn
−→ 0.
Theorem 4.10. Let K be non-compact and let σ be a non-degenerate probability
measure on K. Then we have H p
Proof. Let 1 < p < ∞ and let f be a non-negative function in H p
kf kp = 1. Consider the probability measure σ0 on K defined by
σ(K) = {0} for all 1 < p < ∞.
σ(K) with
σ0 :=
∞
Xn=1
1
2n σn.
It is easy to see that σ0 ∗ f = f . Moreover, we may find g ∈ Lq(K) with kgkq ≤ 1
such that hf, gi = kf kp = 1, where 1 < q < ∞ and 1
q = 1. It follows from [3,
(1.4.11), (1.4.12)] that g ∗ f ∈ C0(K) and kg ∗ f k∞ ≤ kgkqkf kp ≤ 1. Therefore,
1 − g ∗ f ≥ 0 and
p + 1
hσ0, 1 − g ∗ f i = 1 − hσ0, g ∗ f i
= 1 − hσ0 ∗ f, gi
= 1 − hf, gi = 0.
10
M. NEMATI AND J. SOHAEI
It follows from non-degeneracy of σ that 1 = g ∗ f ∈ C0(K), contradicting K
being non-compact. Hence, f = 0. Thus, Proposition 3.6 implies that H p
σ(K) =
{0}.
(cid:3)
We use the following result which is proved in [2, Theorem 3.8] to show that
if σ is an adapted probability measure on compact hypergroup K, then each
σ-harmonic Lp-function is trivial for all 1 ≤ p ≤ ∞.
Theorem 4.11. If σ is an adapted probability measure on a compact hypergroup
K, then each σ-harmonic continuous function on K is constant.
Theorem 4.12. Let K be a compact hypergroup and let σ be an adapted proba-
bility measure on K. Then for 1 ≤ p ≤ ∞, we have H p
σ(K) = C1.
Proof. Let K be compact. Then we have L∞(K) ⊆ Lp(K) ⊆ L1(K) for all
1 ≤ p ≤ ∞. Thus, it suffices to prove the assertion for the case p = 1. Let
f ∈ H 1
σ(K) and let (φα) be a bounded approximate identity for L1(K) such
that φα is a bounded continuous function with compact support for all α; see [3,
Theorem 1.6.15]. Then [13, Lemma 2.2(i)] implies that f ∗ φα is continuous and
bounded for all α. Moreover, it is clear that f ∗ φα is σ-harmonic, and is therefore
constant by Theorem 4.11. This shows that f is also constant, as desired.
(cid:3)
Corollary 4.13. Let K be a compact hypergroup and let σ be a non-degenerate
probability measure on K. Then any weak∗ cluster point σ0 of the sequence
n Pn
(cid:0) 1
k=1 σk(cid:1) in M(K) is the normalized Haar measure on K.
Proof. Let σ0 be a weak∗ cluster point of ( 1
k=1 σk) in M(K). Then we have σ0
is an idempotent probability measure on K satisfying σ ∗ σ0 = σ0, and therefore
σ0 ∗ f ∈ H ∞
It follows from Theorem 4.11 and the
non-degeneracy of σ that for each f ∈ Cb(K) there exists λf ∈ C such that
σ0 ∗ f = λf 1. Let ω be the normalized Haar measure on K. Then for each
f ∈ Cb(K),
σ (K) for all f ∈ Cb(K).
n Pn
hω, f i = ZK
f (x)dω(x) = ZK
(σ0 ∗ f )(x)dω(x) = λf .
Moreover,
h σ0, f i = h σ0 ∗ σ, f i = hσ, σ0 ∗ f i = hσ, λf 1i = λf .
This shows that σ0 = ω. Since ω = ω, we conclude that σ0 = ω.
(cid:3)
Let σ ∈ M(K). We say that a measure µ ∈ M(K) is σ-harmonic if it satisfies
the convolution equation σ ∗ µ = µ. Define Hσ(K) to be the set of all σ-harmonic
measures.
Theorem 4.14. Let σ ∈ M(K) with kσk = 1. Then there is a contractive
projection Pσ : M(K) → M(K) with Pσ(M(K)) = Hσ(K).
Proof. Let U be a free ultra-filter on N, and define Pσ : M(K) → M(K) by the
weak∗ limit
Pσ(µ) = lim
U
1
n
n
σk ∗ µ.
Xk=1
AMENABILITY AND HARMONIC Lp-FUNCTIONS ON HYPERGROUPS
11
It is easy to see that Pσ(µ) = µ for all µ ∈ Hσ(K). Moreover, if µ ∈ M(K), then
it is easily verified that σ ∗ Pσ(µ) = Pσ(µ) and hence Pσ(µ) ∈ Hσ(K). These
show that P 2
(cid:3)
σ = Pσ and Pσ(M(K)) = Hσ(K).
Recall that a measure µ ∈ M(K) is non-negative if hµ, f i ≥ 0 for all f ∈
C0(K)+.
Proposition 4.15. Let σ be a probability measure on K. Then Hσ(K) is gener-
ated by its non-negative elements.
Proof. Suppose that µ ∈ Hσ(K). Then σ ∗ µ = (σ ∗ µ) = µ. This shows that
Hσ(K) is generated by its real measure parts. Now let µ ∈ Hσ(K) be a real
measure and let µ = µ+ − µ−, where µ+, µ− are non-negative measures in M(K).
Since σ is positive, the measures Pσ(µ+), Pσ(µ−) ∈ Hσ(K) are non-negative.
Moreover, µ = Pσ(µ) = Pσ(µ+) − Pσ(µ−), which completes the proof.
(cid:3)
Theorem 4.16. Let K be non-compact and let σ be a non-degenerate probability
measure on K. Then we have Hσ(K) = {0}.
Proof. Suppose that µ ∈ Hσ(K) is non-zero. By Proposition 4.15, we can as-
sume that µ is positive. Consider the probability measure σ0 defined by σ0 :=
It is clear that σ0 ∗ µ = µ. By [3, Theorem 1.6.9], we have that
δx ∗ µ = µ for all x ∈ suppσ0 = K. This shows that µ is the left Haar measure
that is finite on K, which is a contradiction with non-compactness of K.
P∞
n=1
1
2n σn.
(cid:3)
Corollary 4.17. Let K be non-compact and let σ be a non-degenerate probability
measure on K. Then we have H 1
σ(K) = {0}.
Proof. Suppose that f ∈ H 1
measure on K. Then we have
σ(K). Define µ := f ω, where ω is the left Haar
µ = (σ ∗ f )ω = σ ∗ (f ω) = σ ∗ µ.
Therefore, µ = 0 by Theorem 4.16 and hence f = 0.
(cid:3)
Theorem 4.18. let K be compact and let σ be an adabted probability measure on
K. Then we have Hσ(K) = Cω, where ω is the normalized Haar measure on K.
Proof. Because K is compact, the Haar measure ω is in M(K) and we have
σ ∗ ω = ω. Therefore, CωK ⊆ Hσ(K). To prove the converse, suppose that
µ ∈ Hσ(K). Then σ ∗ µ = µ. Let (φα) be a bounded approximate identity for
w∗
L1(K) such that φα ∈ C +
−→ δe; see [3, Theorem 1.6.15].
Then µ ∗ φα
σ(K) for all α. Thus, Theorem 4.12 implies
that µ ∗ φα is constant for all α. Hence, for every α there is λα ∈ C such that
µ ∗ φα = λα1. It follows that for each f ∈ Cb(K), we have
w∗
−→ µ and µ ∗ φα ∈ H 1
c (K) for every α and φα
hµ ∗ φα, f i = ZK
λαf dω = hλαω, f i.
Therefore, hλαωK, f i −→ hµ, f i for all f ∈ Cb(K). This shows that there exists
λ ∈ C such that µ = λω. It follows that Hσ(K) = Cω.
(cid:3)
12
M. NEMATI AND J. SOHAEI
References
1. M. Amini, Harmonic functions on [IN] and central hypergroups, Monatsh. Math. 169
(2013), 267 -- 284.
2. M. Amini and C.-H. Chu, Harmonic functions on hypergroups, J. Func. Anal. 261 (2011),
1835 -- 1864.
3. W. R. Bloom and H. Heyer, Harmonic Analysis of Probability Measures on Hypergroups,
de Gruyter Studies in Mathematics, Vol. 20. Walter de Gruyter, Berlin, 1995.
4. G. Choquet and J. Deny, Sur l'´equation de convolution µ ∗ σ = µ, C.R. Acad. Sci. Paris,
Ser. I Math. 250 (1960), 779 -- 801.
5. C.-H. Chu, Harmonic function spaces on groups, J. London Math. Soc. 70 (2004), 182 -- 198.
6. C.-H. Chu, Matrix convolution operators on groups, Lecture Notes in Math., vol. 1956.
Springer, Heidelberg, 2008.
7. C.-H. Chu and T. Hilberdink, The convolution equation of Choquet and Deny on nilpotent
groups, Integral Equ. Oper. Theory, 26 (1996), 1 -- 13.
8. C.-H. Chu and A. T. Lau, Harmonic functions on groups and Fourier algebras, Lecture
Notes in Math., vol. 1782, Springer-Verlag, Berlin, 2002.
9. H.G. Dales and A.T.-M. Lau, The second duals of Beurling algebras, Mem. Amer. Math.
Soc. 177 no. 836, 2005.
10. E.B. Dynkin and M.B. Malyutov, Random walks on groups with a finite number of gener-
ators, Soviet Math. Doklady, 2 (1961), 399 -- 402.
11. W. Jaworski, Ergodic and mixing probability measures on [SIN] groups, J. Theoret. Probab.
17 (2004), 741 -- 759.
12. B. E. Johnson, Harmonic functions on nilpotent groups, Integral Equ. Oper. Theory, 40
(2001), 454 -- 464.
13. M. Skantharajah, Amenable hypergroups, Illinois J. Math. 36 (1992), 15 -- 46.
14. G. A. Willis, Probability measures on groups and some related ideals in group algebras, J.
Funct. Anal. 92 (1990) 202 -- 263.
1Department of Mathematical Sciences, Isfahan Uinversity of Technology,
Isfahan 84156-83111, Iran;
School of Mathematics, Institute for Research in Fundamental Sciences (IPM),
P.O. Box: 19395 -- 5746, Tehran, Iran.
E-mail address: [email protected]
2 Department of Mathematical Sciences, Isfahan Uinversity of Technology,
Isfahan 84156-83111, Iran
E-mail address: [email protected]
|
1109.3795 | 1 | 1109 | 2011-09-17T15:44:09 | Test functions, Schur-Agler classes and transfer-function realizations: the matrix-valued setting | [
"math.FA",
"math.OA"
] | Given a collection of test functions, one defines the associated Schur-Agler class as the intersection of the contractive multipliers over the collection of all positive kernels for which each test function is a contractive multiplier. We indicate extensions of this framework to the case where the test functions, kernel functions, and Schur-Agler-class functions are allowed to be matrix- or operator-valued. We illustrate the general theory with two examples: (1) the matrix-valued Schur class over a finitely-connected planar domain and (2) the matrix-valued version of the constrained Hardy algebra (bounded analytic functions on the unit disk with derivative at the origin constrained to have zero value). Emphasis is on examples where the matrix-valued version is not obtained as a simple higher-multiplicity tensoring of the scalar-valued version. | math.FA | math |
TEST FUNCTIONS, SCHUR-AGLER CLASSES AND
TRANSFER-FUNCTION REALIZATIONS: THE
MATRIX-VALUED SETTING
JOSEPH A. BALL AND MOIS´ES D. GUERRA HUAM ´AN
Abstract. Given a collection of test functions, one defines the associated
Schur-Agler class as the intersection of the contractive multipliers over the
collection of all positive kernels for which each test function is a contractive
multiplier. We indicate extensions of this framework to the case where the
test functions, kernel functions, and Schur-Agler-class functions are allowed
to be matrix- or operator-valued. We illustrate the general theory with two
examples: (1) the matrix-valued Schur class over a finitely-connected planar
domain and (2) the matrix-valued version of the constrained Hardy algebra
(bounded analytic functions on the unit disk with derivative at the origin
constrained to have zero value). Emphasis is on examples where the matrix-
valued version is not obtained as a simple tensoring with CN of the scalar-
valued version.
1. Introduction
In honor of the work of Issai Schur (see [34]), it is common nowadays to refer to
the class of holomorphic functions s mapping the unit disk D into the closed unit
disk D as the Schur class S. We summarize some of the many characterizations of
the Schur class in the following theorem.
Theorem 1.1. For a given s : D → C, the following are equivalent:
(1) s ∈ S,
(2) the de Branges-Rovnyak kernel associated with s is a positive kernel on D:
Ks(z, w) :=
1 − s(z)s(w)
1 − zw
(cid:23) 0.
(1.1)
(3) s has a unitary transfer-function realization, i.e., there is a unitary colli-
gation matrix U = [ A B
C D ] : X ⊕ C → X ⊕ C so that
s(z) = D + zC(I − zA)−1B.
(1.2)
(4) s satisfies the von Neumann inequality: for any strict contraction operator
T on a Hilbert space K, ks(T )k ≤ 1.
A natural multivariable generalization of the Schur class from this point of view
is to consider functions s defined on the polydisk Dd (where d is a positive inte-
ger). It has been known for some time that the von Neumann inequality fails in
1991 Mathematics Subject Classification. 47A56; 47A48, 47A57, 47B32, 46E22.
Key words and phrases. Schur-Agler class, test functions, positive kernels, completely positive
kernels, reproducing kernel Hilbert spaces, transfer-function realization, internal tensor product
of correspondences, unitary colligation matrix, separation of convex sets, interior point of convex
hull, finitely connected planar domain, constrained H∞-algebra.
1
2
J.A. BALL AND M.D. GUERRA-HUAM ´AN
if d > 2 there is a holomorphic function s on Dd
more than two variables, i.e.:
(even a polynomial) with kskDd ≤ 1 and a commuting d-tuple T = (T1, . . . , Td) of
strict contraction operators on a Hilbert space K for which the multivariable von
Neumann inequality
(1.3)
fails. Nevertheless, the subclass of those Schur-class functions over Dd for which
(1.3) does hold, now called the Schur-Agler class, does have characterizations anal-
ogous to those given in Theorem 1.1 for the single-variable case (see [3, 5, 22]). Note
that the analogue of condition (4) in Theorem 1.1 is now used as the definition of
the Schur-Agler class. We then have the following analogue of Theorem 1.1
Theorem 1.2. Given s : Dd → C, the following are equivalent.
ks(T )k ≤ kskDd
(1) s ∈ SAd.
(2) There are positive kernels K1, . . . , Kd on Dd so that
dXk=1
1 − s(z)s(w) =
(1 − zkwk)Kk(z, w).
(1.4)
(3) There is a unitary colligation matrix U = [ A B
C D ] : X ⊕ C → X ⊕ C and a
collection {P1, . . . Pd} of orthogonal projections with PiPj = 0 for i 6= j and
withPd
j=1 Pj = IX so that
s(z) = D + C(I − Z(z)A)−1Z(z)B
where we have set Z(z) = z1P1 + ··· + zdPd.
(1.5)
In the test-function approach to defining generalized Schur-Agler classes, going
back to the unpublished preprint of Agler [2] and developed further in [6, 27, 29, 41],
one proceeds as follows. We here describe the scalar-valued function setting, al-
though the paper [27] deals with a more general semigroupoid setting. One replaces
the unit disk D (or unit polydisk Dd) with a completely general point set Ω and
supposes that one is given a collection of C-valued functions Ψ on Ω (the set of test
functions) subject to the condition that supψ∈Ψ ψ(z) < 1 for each z ∈ Ω. The
set Ψ carries with it a natural completely regular topology, namely, the weakest
topology with respect to which each of the functions
z ∈ Ω
E(z) : ψ → ψ(z),
is continuous. One then says that a positive kernel k is Ψ-admissible (written as
k ∈ KΨ) if multiplication by ψ is contractive as an operator on the reproducing
kernel Hilbert space H(k) associated with k, i.e., if the kernel Kψ,k(z, w) = (1 −
ψ(z)ψ(w)k(z, w) is positive for each ψ ∈ Ψ. We then say that the function s : Ω → C
is in the Ψ-Schur-Agler class SAΨ if multiplication by s is contractive on H(k) for
each k ∈ KΨ, i.e., if the kernel Ks,k(z, w) = (1 − s(z)s(w))k(z, w) is a positive
kernel for each k ∈ KΨ. We mention that the choice
(1.6)
leads to the classical Schur class while the choice
Ω = D, Ψ = {ψ0(z) = z}
Ω = Dd, Ψ = {ψk(z) = zk : k = 1, . . . , d}
(where z = (z1, . . . , zd) ∈ Dd) leads to the classical Schur-Agler class SAd.
with a general test-function collection Ψ.
The following is the main result concerning the Schur-Agler class SAΨ associated
(1.7)
(1.8)
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
3
Theorem 1.3. (See [27, 29] and [8] for an early version.) Given a function s : Ω →
C, the following are equivalent.
(1) s ∈ SAΨ.
(2) There is a measure ν on Ψβ (the Stone- Cech compactification of Ψ) and a
measurable family {Kψ : ψ ∈ Ψβ} of positive kernels on Ψβ so that
1 − s(z)s(w) =ZΨβ(cid:16)1 − ψ(z)ψ(w)(cid:17) Kψ(z, w) dν(ψ).
(1.9)
(3) There is a C(Ψβ)-unitary colligation, i.e., a bock unitary operator U =
C D ] : X ⊕ C → X ⊕ C together with a ∗-representation ρ of the C∗-algebra
[ A B
C(Ψβ) (continuous complex-valued functions on Ψβ) into L(X ) (bounded
linear operators on X ), so that
s(z) = D + C(I − ρ(E(z))A)−1ρ(E(z))B
(1.10)
(where E(z) is as in (1.6)).
Note that conditions (2) and (3) in Theorem 1.3 become conditions (2) and (3)
in Theorem 1.1 when Ω and Ψ are chosen as in (1.7), and conditions (2) and (3) in
Theorem 1.2 when Ω and Ψ are chosen as in (1.8).
A different type of extension of the classical Schur class over the unit disk is
the Schur-class SR over a bounded, finitely connected planar domain R. Here
R is a bounded domain in the complex plane with boundary consisting of m + 1
disjoint smooth Jordan curves ∂0, ∂1, . . . , ∂m, where ∂0 denotes the boundary of the
unbounded component of the complement of R, and we define SR as the class of all
holomorphic functions from R into the closed disk D−. Work in [27, 29] identifies
the Schur class SR over R as a test-function Schur-Agler class SAΨR for a certain
collection of test functions ΨR = {ψx : x ∈ TR} indexed by the so-called R-torus
TR defined as the Cartesian product of the connected components of ∂R:
x ∈ TR := ∂0 × ∂1 × ··· × ∂m.
(see Section 4.1 below for complete details). In particular, the decomposition (1.9)
in Theorem 1.3 for this case gives us the following: given s ∈ SR, there is a measure
ν on TR and a family of positive kernels {kx : x ∈ TR} so that
S(T ) = Xn∈Zd
+
Sn ⊗ T n if S(z) = Xn∈Zd
Snzn
1 − s(z)s(w) =ZTR(cid:16)1 − ψx(z)ψx(w)(cid:17) kx(z, w) dν(x).
(1.11)
We shall be interested in matrix- and operator-valued versions of these Schur and
Schur-Agler classes. The operator-valued version of the Schur class over R, which
we denote as SR(U,Y), consists of holomorphic functions S on R with values S(z)
equal to contraction operators between two Hilbert spaces U and Y. For the case
where R = D, we drop the subscript R and write simply S(U,Y); we also abbreviate
SR(U,U) to SR(U). There is also an operator-valued version of the Schur-Agler
class over Dd, namely: S : Dd → L(U,Y) is in the Schur-Agler class SAd(U,Y)
if S is a holomorphic map from Dd into L(U,Y) such that kS(T )k ≤ 1 for any
commutative tuple T = (T1, . . . , Td) of strictly contractive operators on a Hilbert
space K, where we use a tensor functional calculus to define S(T ):
4
J.A. BALL AND M.D. GUERRA-HUAM ´AN
where we use standard multivariable notation:
··· T nd
1 ··· znd
d ,
T n = T n1
1
zn = zn1
d
for n = (n1, . . . , nd) ∈ Zd
+.
Then Theorems 1.1 and 1.2 have seamless extensions to the matrix-/operator-valued
settings. Indeed, S ∈ S(U,Y) if and only if the de Branges-Rovnyak L(Y)-valued
kernel
KS(z, w) :=
IY − S(z)S(w)∗
1 − zw
that I − S(z)S(w)∗ =Pd
is a positive kernel on D if and only if there is a unitary colligation matrix U =
C D ] : X ⊕ U → X ⊕ Y so that S(z) = D + zC(I − zA)−1B. Similarly, S ∈
[ A B
SAd(U,Y) if and only if there are positive L(E)-valued kernels K1, . . . , Kd on Dd so
k=1(1− zkwk)Kk(z, w) if and only if S has a representation
as in (1.5) but with U acting from X ⊕ U to X ⊕ Y. We mention that this result
has inspired several variants where the polydisk Dd is replaced by a more general
domain DQ in Cd specified by a polynomial (or more generally analytic) matrix-
valued determining function Q: DQ = {z ∈ Cd : kQ(z)k < 1}; more generally
the technique of the proof going through the transfer-function realization naturally
leads to interpolation and commutant lifting versions of the result (see [22, 21, 53,
23, 10, 16, 9]). We mention that there is now also a noncommutative version of the
Schur-Agler class [19].
However, for the case SR(CN ), the expected matrix generalization of (1.11),
namely
I − S(z)S(w)∗ =ZTR(cid:16)1 − ψx(z)ψx(w)(cid:17) Kx(z, w) dν(x)
(1.12)
for a measurable family {Kx : x ∈ TR} of positive N × N matrix-valued kernels on
R, fails in general, at least in the case where R is a region with three holes having
some additional symmetry properties; indeed this phenomenon is a key ingredient
in the negative answer to the spectral set question for such regions R obtained by
Dritschel and McCullough in [28].
One of the main motivations for the present paper is to develop a framework
of test-function Schur-Agler class SAΨ for the case of matrix- or operator-valued
test functions Ψ and to recover a formula of the type (1.12) for the Schur class
SR(CN ) for an appropriately enlarged class ΨN
R of matrix-valued test functions.
We therefore develop a systematic extension of the work of [27, 29] to the matrix-
and operator-valued setting: this is the main content of Section 3 below. We also
emphasize the interpolation version of the main result, whereby one characterizes
which functions S0 defined on some subset Ω0 of Ω can be extended to a test-
function Schur-Agler-class function S defined on all of Ω. Most of the analysis
builds on the earlier work of [3, 5, 22, 10, 16, 8, 27, 29], but there are places where
new ideas and techniques were required.
In Section 4 we take two algebras which are intrinsically defined and identify
their unit balls as also arising as test-function Schur-Agler classes. The first has
already been mentioned: namely, the algebra of bounded holomorphic N×N matrix
functions over a multiply-connected planar domain R whose unit ball is the Schur
class SR(CN ). The second is the matrix-valued version of the constrained Hardy
algebra over the unit disk D (bounded holomorphic functions f on D subject to the
constraint that f ′(0) = 0). The first example has been an object of much study
over the years (see [1, 14, 18, 4, 28, 54]) while interest in the second is more recent
[26, 17, 50]. Motivation for study of the second algebra comes from the fact that
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
5
it is a model for the bounded analytic functions on the intersection of a variety V
embedded in C2 with the unit bidisk (see [7]). For these two examples we identify
an appropriate class of test functions ΨN so that the unit ball of the given algebra
is equal to the matrix-valued test-function Schur-Agler class SAΨN associated with
ΨN . It is always possible to choose ΨN simply as the unit ball of the given algebra;
the point is to find a valid class ΨN which is as small as possible. As has already
been mentioned for the first example, in both examples the test-function class Ψ1
identified in previous work ([29, 30]) for the scalar-valued version fails to work
for the matrix-valued case. For each of these two examples, we find a valid test-
function class ΨN as a linear-fractional transform of the set of extreme points of
a normalized matrix-valued Herglotz (positive real part) version of the algebra,
just as has been done for the scalar-valued case in [28, 28, 30]. Identification of
these extreme points for the matrix-valued case leads us to draw on results from
[20] concerning extreme points for a convex cone of matrix quantum probability
measures (positive matrix-valued measures with total mass equal to the identity
matrix). The resulting test-function classes are not as explicit as in the scalar-
valued settings; however, for the Schur class SR with R equal to an annulus, we are
able to use results of McCullough [38] to obtain a more explicit test-function class
and use the resulting matrix-valued continuous Agler decomposition (the matrix-
valued analogue of (1.9)) to obtain a variant of McCullough's positive solution of
the spectral set question for an annulus.
A criticism of the study of Schur-Agler classes in general is that their intrinsic
structure is a priori mysterious: after going through the several steps of the defini-
tion, one does not have any intrinsic characterization of the eventual result. Our
work in Section 4 (as well as the work in [29, 30]) counterbalances this concern by
starting with an intrinsically defined function algebra and identifying it as a Schur-
Agler class. There are now papers obtaining characterizations of which operator
algebras have unit balls equal to a Schur-Agler class (see [42, 36]). Other work [37]
characterizes families of kernels so that the associated contractive multipliers form
a test-function Schur-Agler class. It should be of interest to extend these results to
the matrix-valued setting in the spirit of the present paper.
The paper is organized as follows. Section 2 presents some preliminary mate-
rial on test functions, positive kernels, and structured unitary colligation matrices
needed in the sequel. Section 3 presents the main structure result (including the
interpolation version as well as a representation-theoretic version) for the general
matrix-valued test-function Schur-Agler class. Section 4 develops the two illustra-
tive examples of matrix-valued Schur classes which can be identified as test-function
Schur-Agler classes. Finally we mention that this paper together with [20] form an
enhanced version of the second author's dissertation [35].
2. Preliminaries
2.1. Test functions. We assume that we are given two coefficient Hilbert spaces
UT and YT and a collection Ψ of functions ψ on the abstract set of points Ω with
values in the space L(UT ,YT ) of bounded linear operators between UT and YT . We
say that Ψ is a collection of test functions if it happens that
sup{kψ(z)k : ψ ∈ Ψ} < 1 for each z ∈ Ω.
(2.1)
6
J.A. BALL AND M.D. GUERRA-HUAM ´AN
We view Ψ as a subset of B(Ω,BL(UT ,YT )) (the space of (bounded) maps from
Ω into the closed unit ball of bounded linear operators between UT and YT ). We
topologize B(Ω,BL(UT ,YT )) with the topology of pointwise weak-∗ convergence,
i.e., we view B(Ω,BL(UT ,YT )) as the Cartesian product ΠΩBL(UT ,YT ) with the
standard Cartesian product topology). As such B(Ω,BL(UT ,YT )) is compact by
Tychonoff's Theorem ([31, Theorem XI.1.4]), since each fiber BL(UT ,YT ) is com-
pact by the Banach-Alaoglu Theorem [51, Theorem 3.15]. As a subspace of the
completely regular space B(Ω,BL(UT ,YT )) (i.e., B(Ω,BL(UT ,YT )) is Hausdorff
and any closed set can be separated from a point disjoint from it by a contin-
uous function), Ψ is completely regular in the subspace topology inherited from
B(Ω,BL(UT ,YT )). The closure of Ψ in this topology is compact; however we shall
be more interested in the Stone- Cech compactification Ψβ of Ψ [31, Section XI.8].
Then the space Cb(Ψ,L(H,K)) of bounded continuous functions f from Ψ into
a space L(H,K) of bounded linear operators between two Hilbert spaces H and
K can be identified with the space C(Ψβ,L(H,K)) of continuous functions from
the Stone- Cech compactification Ψβ into L(H,K). An operator-valued version of
the Riesz representation theorem allows us to identify the dual of Cb(Ψ,L(H,K))
with regular, bounded, weakly countably additive C1(K,H)-valued measures on Ψβ,
where we use the notation C1(K,H) to denote the trace-class operators from K to H.
We note that there are continuous linear functionals L in C(Ψβ,L(H,K)) such that
allowing points of Ψβ \ Ψ to be part of the support of the corresponding measure
µL is essential (see [29, Section 5.2]).
For each ψ ∈ Ψ we define the map evψ : Cb(Ψ,L(H,K)) → L(K) by evψ : f →
f (ψ). A particular element of Cb(Ψ,L(UT ,YT )) which will often come up is the
function E(z) (for each z ∈ Ω) given by
evψ(E(z)) = E(z)(ψ) := ψ(z).
(2.2)
2.2. Positive operator-valued kernels and their multipliers. Let E be any
Hilbert space and suppose that K is a function on Ω × Ω with values in L(E). We
say that K is a positive kernel if the Aronszajn condition
NXi,j=1
hK(zi, zj)ej, eiiE ≥ 0 for all z1, . . . , zn ∈ Ω, e1, . . . , eN ∈ E, N = 1, 2, . . . .
(2.3)
The following equivalent versions of the positive-kernel condition are often used in
function-theoretic operator theory settings.
Theorem 2.1. (See e.g. [6].) Suppose that we are give a function K : Ω × Ω →
L(E). Then the following are equivalent:
(1) K is a positive kernel, i.e., condition (2.3) holds.
(2) There is a Hilbert space H(K) consisting of E-valued functions f such that
K(·, w)e ∈ H(K) for each w ∈ Ω and e ∈ E and has the reproducing
property:
hf, K(·, w)eiH(K) = hf (w), eiE for all f ∈ H(K).
(3) K has a Kolmogorov decomposition: there is an auxiliary Hilbert space X
and a function H : X → E so that
K(z, w) = H(z)H(w)∗.
(2.4)
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
7
In fact one can take X to be the reproducing kernel Hilbert space H(K)
described in (2) above with H(z) = evz : f 7→ f (z).
Rather than using a positive kernel to construct a reproducing kernel Hilbert
space as in condition (2) in Theorem 2.1, it is also possible to construct a repro-
ducing kernel Hilbert module as follows. By a Hilbert module over a C∗-algebra B
we mean a linear space E which is a right module over B which is also equipped
with an B-valued inner product and satisfies additional compatibility requirements
with respect to the algebra structure of B (see [49, Section 2.1]):
which satisfies the usual inner product axioms:
h·,·iE : E × E → B
(1) hλx + µy, zi = λhx, zi + µhy, zi,
(2) hx · b, yi = hx, yib,
(3) hx, yi∗ = hy, xi,
(4) hx, xi ≥ 0 (as an element of B),
(5) hx, xi = 0 implies that x = 0,
(6) E is complete in the norm given by kxk = khx, xik1/2
A
for all x, y, z ∈ E, b ∈ B and λ, µ ∈ C. (Here we follow the mathematicians'(rather
than the physicists') convention that inner products are linear in the left slot; this
departs from the standard usage in the operator-algebra literature.) By modifying
the construction of H(K) in Theorem 2.1, one can construct a C∗-module, denoted
as H(K), over the C∗-algebra L(E) characterized as follows.
Theorem 2.2. Suppose that K : Ω × Ω → L(E) is a positive kernel as in (2.3).
Then there is a uniquely determined C∗-module H(K) over B = L(E) with the
following properties:
(1) H(K) consists of L(E)-valued functions on Ω,
(2) for each w ∈ Ω, K(·, w) is in H(K) and the span of such elements is dense
(3) for each F ∈ H(K),
in H(K), and
hF, K(·, w)iH(K) = F (w) ∈ L(E).
Proof. Define an inner product on a pair of kernel elements K(·, w) and K(·, z) by
hK(·, w), K(·, z)iH(K) = K(z, w)
and extend by linearity to the space of kernel elements. Mod out by any linear
combinations having zero self inner product and take the completion to arrive at
the space H(K) having all the asserted properties. Note that there is a version
of the Cauchy-Schwarz inequality available (see [49, Lemma 2.5]) which guarantees
that the point evaluation map ev : f 7→ f (w) extends to elements of the completion,
and hence elements of the completion can also be identified as L(E)-valued functions
on Ω.
(cid:3)
It is natural now to take the next step and introduce the notion of C∗-corres-
pondence (see [43]). Given two C∗-algebras A and B, by an (A, B)-correspondence
we mean a Hilbert module E over B which also carries a left A-action x 7→ a · x
which is a ∗-representation of A with respect to the B-valued inner product on E:
ha · x, yiE = hx, a∗ · yiE.
8
J.A. BALL AND M.D. GUERRA-HUAM ´AN
Given three C∗-algebras A, B and C together with an (A, B)-correspondence E
and a (B, C)-correspondence F , the internal tensor product E ⊗ F of E and F is
defined to be the (A, C)-correspondence generated as the Hausdorff completion of
the span of pure tensors e ⊗ f (e ∈ E and f ∈ F ) in the C-valued inner product
given by
with left A-action given by
he ⊗ f, e′ ⊗ f ′iE⊗F = h(he, e′iE) · f, f ′iF
It is routine to verify that one then gets the balancing property
a · (e ⊗ f ) = (a · e) ⊗ f.
e ⊗ (b · f ) = (e · b) ⊗ f
(2.5)
(2.6)
(2.7)
for e ∈ E, f ∈ F and b ∈ B.
We shall need a couple of applications of this internal tensor-product construc-
tion. The first is as follows. For K an L(E)-valued positive kernel on Ω, we view the
C∗-module over B constructed in Theorem 2.2 as a (C,L(E))-correspondence. For
X another coefficient Hilbert space, let C2(X ,E) be the space of Hilbert-Schmidt
class operators from X into E. Then C2(X ,E) has a standard Hilbert-space inner
product
We also have a left action of the C∗-algebra L(E) on C2(X ,E) via left multiplication:
hT, T ′iC2(X ,E) = tr(T T ′∗).
X · T = XT for X ∈ L(E), T ∈ C2(X ,E)
and this action gives rise to a ∗-representation of L(E) on C2(X ,E):
hX · T, T ′iC2(X ,E) = hXT, T ′iC2(X ,E) = tr(XT T ′∗) = tr(T T ′∗X)
= tr(T (X ∗T ′)∗) = hT, X ∗ · T ′iC2(X ,E).
In this way we may view C2(X ,E) as an (L(E), C)-correspondence. We may then
form the internal C∗-correspondence tensor-product H(K) ⊗ C2(X ,E). Explicitly,
the inner product on pure tensors F ⊗ T (F ∈ H(K), T ∈ C2(X ,E) is given by
hF ⊗ T, F ′ ⊗ T ′iH(K)⊗C2(X ,E) = tr(cid:0)hF, F ′iH(K)T T ′∗(cid:1) .
When we evaluate the first factor F in a pure tensor F ⊗ T at a point w in Ω,
we get a tensor of the form
F (w) ⊗ T ∈ L(E) ⊗ C2(X ,E) ∼= C2(X ,E).
To interpret this tensor product as a C∗-correspondence internal tensor product, we
view L(E) as a (L(E),L(E))-correspondence with inner product hX, X ′i = X ′∗X ∈
L(E) and left action given by left multiplication: X ′ · X = X ′X. The balancing
property (2.7) then leads to the identification L(E) ⊗ C2(X ,E) ∼= C2(X ,E).
Using a linearity and approximation argument, one can show that in fact ele-
ments H of H(K)⊗C2(X ,E) can be viewed as C2(X ,E)-valued functions on Ω such
that K(·, w)U ∈ H(K)⊗C2(X ,E) for each w ∈ Ω and U ∈ C2(X ,E), and the kernel
element K(·, w)U has the reproducing property
hG, K(·, w)UiH(K)⊗C2(X ,E) = hG(w), UiC2 (X ,E) := tr (G(w)U ∗) .
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
9
Thus H(K)⊗C2(X ,U) is a reproducing kernel Hilbert space in the sense of Theorem
2.1 when we identify the range space L(E) of K as the subspace of L(C2(X ,E))
consisting of left multiplication operators by elements of L(E):
X ∈ L(E) 7→ LX ∈ L(C2(X ,E)) : LX : T 7→ XT
and we view C2(X ,E) as a Hilbert space in the inner product
hT, T ′iC2(X ,E) := tr (T T ′∗) .
In the sequel it will be convenient to use the shorthand notation
H(K)X := H(K) ⊗ C2(X ,E).
(2.8)
Note that in this notation, if H(K) is as in Theorem 2.1, then we have H(K) =
H(K)C.
Remark 2.3. The space H(K)X could just as well have been constructed as equal
to the space H(K) ⊗ C2(X , C) where the spaces H(K) (defined as in Theorem
2.1) and C2(X , C) (the dual version of the Hilbert space X ) are viewed as (C, C)-
correspondences (i.e., as ordinary Hilbert spaces), and the tensor product reduces
to the standard Hilbert-space tensor product.
Suppose that we are given two coefficient Hilbert spaces U and Y and an L(U,Y)-
valued function S on Ω. We define the right multiplication operator RS by
(RS(F )) (z) = F (z)S(z).
Thus RS maps C2(U,E)-valued functions on Ω to C2(U,E)-valued functions on Ω.
Given a positive L(E)-valued kernel K on Ω, it is of interest to determine exactly
when RS maps H(K)Y boundedly (or contractively) into H(K)U . The answer is
given by the following theorem.
Theorem 2.4. Let K be an L(E)-valued positive kernel on Ω and S an L(U,Y)-
valued function on Ω. Then the right multiplication operator RS is bounded as an
operator from H(K)Y to H(K)U with kRSk ≤ M if and only if the C-valued kernel
(2.9)
kX,S,K,M (z, w) := tr(cid:2)X(w)∗(M 2IU − S(w)∗S(z))X(z)K(z, w)(cid:3)
is a positive kernel on Ω for each choice of function X : Ω → C2(E,U).
Proof. By rescaling it suffices to consider the case M = 1 and kRSk ≤ 1.
The computation
hRSf, K(·, w)UiH(K)U = hf (w)S(w), UiC2(U ,E)
= tr (f (w)S(w)U ∗)
= tr (f (w)(U S(w)∗))
= hf, K(·, w)U S(w)∗iH(K)Y
shows that
(RS)∗ : K(·, w)U 7→ K(·, w)U S(w)∗
whenever RS is well defined as an element of L(H(K)Y ,H(K)U ). As elements of
j=1 K(·, zj)Uj are dense in H(K)U , we see that kRSk ≤ 1 holds if and
the formPN
10
only if
J.A. BALL AND M.D. GUERRA-HUAM ´AN
2
2
2
K(·, zj)Uj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
K(·, zj)Uj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
R∗
K(·, zj)Uj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
S
NXj=1
K(·, zj)UjS(zj)∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
−(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
NXj=1
2
0 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
NXj=1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
NXj=1
NXi,j=1
NXi,j=1
NXi,j=1
=
=
holds for all choices of z1, . . . , zN ∈ Ω and U1, . . . , UN ∈ C2(U,E) and N = 1, 2, . . . .
Expanding out self inner products and using the invariance of the trace under cyclic
permutations converts this condition to
0 ≤
tr (K(zi, zj)UjU ∗
i − K(zi, zj)UjS(zj)∗S(zi)U ∗
i )
tr (Uj(I − S(zj)∗S(zi))U ∗
i K(zi, zj))
tr (X(zj)∗(I − S(zj)∗S(zi))X(zi)K(zi, zj))
where we have set X(zi) = U ∗
i . This positivity condition holding for all choices
of z1, . . . , zN ∈ Ω and X(z1), . . . , X(zN ) ∈ C2(E,U) for all N = 1, 2, . . . in turn is
equivalent to the positivity of the kernel kX,S,K,1 on Ω for all choices of X : Ω →
C2(E,U).
(cid:3)
We shall also need a characterization of functional Hilbert spaces of the form
H(K)X .
Theorem 2.5. Suppose that H is a Hilbert space whose elements are C2(X ,E)-
valued functions on Ω. Then there is an L(E)-valued positive kernel K on Ω such
that H is isometrically equal to H(K)X if and only if
(1) the point evaluation map evw : f 7→ f (w) defines a bounded operator from
(2) H is a right module over L(X ) with the right action of L(X ) commuting
H into C2(X ,E) fo each w ∈ Ω, and
with each point evaluation map evw:
evw(f · X) = (evwf )X or (f · X)(w) = f (w)X for all w ∈ Ω.
(2.10)
Proof. By Theorem 2.1, from the fact that the point evaluations evw are bounded,
we get that H = H(K) for an L(C2(X ,E))-valued positive kernel K(z, w) = evz ·
(evw)∗. The additional condition (2.10) then implies that K(z, w) commutes with
the right multiplication operators RX : T 7→ T X on C2(X ,E) (X ∈ L(X )). This
is enough to force K(z, w) to be a left multiplication operator K(z, w) = LK(z,w)
for a K(z, w) ∈ L(E). One next verifies that K so constructed is an L(E)-valued
positive kernel and that we recover H as H = H(K)X .
(cid:3)
We shall also have use for a far-reaching generalization of the positive kernels
discussed so far introduced by Barreto, Bhat, Liebscher, and Skeide in [24]. Given
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
11
two C∗-algebras A and B, we say that a function Γ on Ω× Ω with values in L(A, B)
is a completely positive kernel if
NXi,j=1
b∗
i Γ(zi, zj)[a∗
i aj]bj ≥ 0 (in B)
(2.11)
for all choices of z1, . . . , zN ∈ Ω, a1, . . . , aN ∈ A, b1, . . . , bN ∈ B for all N = 1, 2, . . . .
The following characterization of completely positive kernels is the completely pos-
itive parallel to Theorems 2.1 and 2.2.
Theorem 2.6. (See [24, 15].) Given a function Γ on Ω× Ω with values in L(A, B),
the following are equivalent:
(1) Γ is a completely positive kernel, i.e., condition (2.11) holds.
(2) There is an (A, B)-correspondence H(Γ) whose elements consist of B-valued
functions f on Ω such that K(·, w)[a] ∈ H(Γ) for each w ∈ Ω and a ∈ A
and such that
hf, K(·, w)[a]iH(Γ) = (a∗ · f ) (w)
(3) K has a Kolmogorov decomposition of the following form:
for all f ∈ H(Γ), a ∈ A, and w ∈ Ω.
there is an
(A, B)-correspondence H and a function H on Ω with values in the space
L(H, B) of adjointable operators from H to B so that
K(z, w)[a] = H(z)π(a)H(w)∗.
Here a 7→ π(a) represents the left A-action on H: π(a)f = a· f for f ∈ H.
In case B = L(E) for a Hilbert space E, then we also have Hilbert space versions of
conditions (2) and (3):
(2′) There is an (A, C)-correspondence H(Γ) (i.e., a Hilbert space H(Γ) equipped
with a ∗-representation π : A → L(H(Γ)) of A) whose elements are E-valued
functions f on Ω such that K(·, w)[a]e ∈ H(Γ) for each w ∈ Ω, a ∈ A,
e ∈ E, and such that
hf, K(·, w)[a]eiH(Γ) = h(a∗ · f ) (w), eiE
(3′) There exists a Hilbert space H carrying a ∗-representation π of A and there
for all f ∈ H(Γ), a ∈ A, w ∈ Ω.
exists a function H : Ω → L(H,E) so that
K(z, w)[a] = H(z)π(a)H(w)∗.
Remark 2.7. The positivity condition in Theorem 2.4 can be equivalently formu-
lated as the condition that the kernel
kΓ,S,K(z, w) = [Γ(z, w)[I − S(w)∗S(z)], K(z, w)]C1(E)×L(E)
be a positive C-valued kernel on Ω for every choice of completely positive kernel
where the outside bracket
Γ : Ω × Ω → L(L(U),C1(E)),
is the duality pairing between the trace-class operators C1(E) and the bounded
linear operators L(E).
[·,·]C1(E)×L(E)
12
J.A. BALL AND M.D. GUERRA-HUAM ´AN
2.3. Ψ-unitary colligations. For the transfer-function realization
S(z) = D + zC(I − zA)−1B
in the operator-valued test-function setting to be developed in the sequel, we shall
need a more elaborate version of the unitary colligation matrix U = [ A B
C D ] which
we now describe. Given a collection of test functions Ψ as in Section 2.1, as de-
scribed there we view Ψ as a completely regular topological space. Then the space
Cb(Ψ,L(YT )) of bounded L(YT )-valued functions on Ψ is a C∗-algebra while the
space Cb(Ψ,L(YT ,UT )) of continuous L(YT ,UT )-valued functions is not (unless
UT = YT ). However we may view Cb(Ψ,L(YT ,UT )) as a (Cb(Ψ, ,UT ), Cb(Ψ,YT ))-
correspondence, with Cb(Ψ,L(YT ))-valued inner product given by
(cid:0)hF, F ′iCb(Ψ,L(YT ,UT ))(cid:1) (ψ) := F ′(ψ)∗F (ψ).
If X is a Hilbert space carrying a ∗-representation ρ of Cb(Ψ,L(YT )), then we
may view X as a (Cb(Ψ,L(YT )), C) correspondence (with the representation ρ pro-
viding the left Cb(Ψ,L(YT ))-action on X ) and form the internal tensor product
Cb(Ψ,L(YT ,UT )) ⊗ρ X . We shall say that a 2 × 2-block unitary matrix U = [ A B
C D ]
is a Ψ-unitary colligation if U has the form
U =(cid:20)A B
C D(cid:21) : (cid:20)X
U(cid:21) →(cid:20)Cb(Ψ,L(YT ,UT )) ⊗ρ X
Y
(cid:21)
for X equal to a Hilbert space carrying a ∗-representation ρ of Cb(Ψ,L(YT )).
in (2.2) (for a given z ∈ Ω). Hence the tensor multiplication operator
A particular element of Cb(Ψ,L(YT ,UT )) is the function E(z)∗, where E(z) is as
(2.12)
defines an operator from X to Cb(Ψ,L(YT ,UT ))⊗ρX ; one can verify that its adjoint
acting on pure tensors is given by
LE(z)∗ : x 7→ E(z)∗ ⊗ x
As a consequence we get the identity
L∗
E(z)∗ : g ⊗ x 7→ ρ(E(z)g)x.
L∗
E(z)∗ LE(w)∗x = L∗
E(z)∗(E(w)∗ ⊗ x) = ρ (E(z)E(w)∗) x.
(2.13)
In case YT = UT (the square case), then Cb(Ψ,L(YT ,UT )⊗ρX collapses down to X
(a consequence of the balancing property (2.7)), and then L∗
E(z)∗ can be identified
with L∗
E(z)∗ = ρ(E(z)). We conclude that the tensor-product construction is exactly
the technical tool needed to push the square case to the non-square case. This type
of colligation matrix appears in [8, 27, 29] for the square case and in [44] for the
nonsquare case.
3. The Schur-Agler class associated with a collection of test
functions
Suppose that we are given a collection Ψ of test functions ψ : Ω → L(UT ,YT )
satisfying the admissibility condition (2.1). For E any auxiliary HIlbert space and
K an L(E)-valued positive kernel on Ω, we say that K is Ψ-admissible, written as
K ∈ KΨ(E), if the operator Rψ : f (z) 7→ f (z)ψ(z) is contractive from H(K)YT to
H(K)UT for each ψ ∈ Ψ, or equivalently (by Theorem 2.4), if the C-valued kernel
(3.1)
kX,ψ,K(z, w) = tr (X(w)∗(I − ψ(w)∗ψ(z))X(z)K(z, w))
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
13
is a positive kernel for each choice of X : Ω → C2(E,UT ) and ψ ∈ Ψ. We then say
that the function S : Ω → L(U,Y) is in the Ψ-Schur-Agler class SAΨ(U,Y) if the
operator RS of right multiplication by S is contractive from H(Y)Y to H(Y)U for
each Ψ-admissible L(Y)-valued positive kernel K, or equivalently, if the kernel
(3.2)
is a positive C-valued kernel for each choice of Y : Ω → C2(Y,U) and K ∈ KΨ(Y).
kY,S,K(z, w) = tr(Y (w)∗(I − S(w)∗S(z))Y (z)K(z, w))
Our main result on the Schur-Agler class SAΨ(U,Y) is the following.
Theorem 3.1. Suppose that we are given a collection of test functions Ψ satisfying
condition (2.1) and S0 is a function on some subset Ω0 of Ω with values in L(U,Y).
Consider the following conditions:
(1) S0 can be extended to a function S defined on all of Ω such that S ∈
SAΨ(U,Y), i.e., the kernel (3.2) is a positive kernel for all choices of
L(Y,U)-valued functions Y on Ω0 and all choices of kernels K ∈ KΨ(Y).
(2) S0 has an Agler decomposition on Ω0, i.e., there is a completely positive
kernel Γ : Ω0 × Ω0 → L(Cb(Ψ,L(YT )),L(Y)) so that
I − S0(z)S0(w)∗ = Γ(z, w)[I − E(z)E(w)∗]
(3.3)
for all z, w ∈ Ω0 (where E(z) ∈ Cb(Ψ,L(UT ,YT )) is as in (2.2)).
(3) There is a Hilbert state space X which carries a ∗-representation of the
C∗-algebra Cb(Ψ,L(YT )) and a Ψ-unitary colligation U (see Section 2.3)
(3.4)
U =(cid:20)A B
Y
so that S0 has the transfer-function realization
E(z)∗A)−1L∗
C D(cid:21) : (cid:20)X
S0(z) = D + C(I − L∗
U(cid:21) →(cid:20)Cb(Ψ,L(YT ,UT )) ⊗ρ X
E(z)∗B
(3.5)
(cid:21)
for z ∈ Ω0.
Then (1) ⇒ (2) ⇔ (3); if dimYT < ∞, then also (2) ⇒ (1) and hence (1), (2),
(3) are all equivalent to each other.
We shall prove (1) ⇒ (2) ⇒ (3) ⇒ (2) and, if dim YT < ∞, then also (2) ⇒ (1).
Proof of (1) ⇒ (2): Step 1: Ω0 is a finite subset of Ω.
We define a cone C by
C ={Ξ : Ω0 × Ω0 → L(Y) : Ξ(z, w) = Γ(z, w)[I − E(z)E(w)∗] for some
completely positive kernel Γ : Ω0 × Ω0 → L(Cb(Ψ,L(YT )),L(Y))}.
Note that the elements of C can be viewed as matrices with rows and columns
indexed by the finite set Ω0 and matrix entries in L(Y). Thus we may view C as a
subset of the linear space V of all such matrices with topology of pointwise weak-∗
convergence. We shall need a few preliminary lemmas. It is easy to verify that C is
a cone in V.
Lemma 3.2. The cone C is closed in V.
Proof of Lemma. Suppose that {Ξα} is a net of elements of C such that {Ξα(z, w)}
converges weak-∗ to Ξ(z, w) for each z, w ∈ Ω0. Thus, for each index α there is a
choice of completely positive kernel Γα so that
Ξα(z, w) = Γα(z, w)[I − E(z)E(w)∗].
(3.6)
14
J.A. BALL AND M.D. GUERRA-HUAM ´AN
The computation
Γα(z, z)[I] = Γα(z, z)[(I − E(z)E(z)∗)1/2(I − E(z)E(w)∗)−1(I − E(z)E(z)∗)1/2]
1
1 − kE(z)k2(cid:19) (I − E(z)E(z)∗)1/2(cid:21)
≤ Γα(z, z)(cid:20)(I − E(z)E(z)∗)1/2(cid:18)
=(cid:18)
=(cid:18)
1 − kE(z)k2(cid:19) Γα(z, z)[I − E(z)E(z)∗]
1 − kE(z)k2(cid:19) Ξα(z, z)
1
1
shows that
kΓα(z, z)k ≤ MzkΞα(z, z)k where Mz =
1
1 − kE(z)k2 ,
(3.7)
where we used here the underlying assumption (2.1) for our set of test functions Ψ.
Since the block 2 × 2 matrix
(cid:20) Γα(z, z)[I] Γα(z, w)[I]
Γα(w, z)[I] Γα(w, w)[I](cid:21)
kΓα(z, w)k ≤ MzMwkΞα(z, w)k1/2kΞα(w, w)k1/2.
is positive semidefinite for each index α and each pair of points z, w ∈ Ω0, it follows
that
(3.8)
Since Ω0 is finite, we see that kΓα(z, w)k is in fact bounded uniformly with respect to
the indices α and the points z, w in Ω0. Since L(Cb(Ψ,L(YT )),L(Y)) is the Banach-
space dual of the projective tensor-product Banach space C1(Y) ⊗ Cb(Ψ,L(YT ))
(see e.g. [52, Theorem IV.2.3]), it follows from the Banach-Alaoglu theorem that
there is a subnet {Γβ} of {Γα} such that {Γβ(z, w)} converges weak-∗ to some
Γ∞(z, w) ∈ L(Cb(Ψ,L(YT )),L(Y)). It is straightforward to verify that the defining
property (2.11) for a completely positive kernel is preserved under such weak-∗
limits; hence Γ∞ is again a completely positive kernel. Moreover, from the fact
that {Ξα(z, w)} converges weak-∗ to Ξ(z, w), we get that the subnet {Ξβ(z, w)}
also converges weak-∗ to Ξ(z, w). Taking limits in the formula (3.6) leads us to the
representation
Ξ(z, w) = Γ∞(z, w)[I − E(z)E(w)∗]
for the limit kernel Ξ(z, w). We conclude that the limit kernel Ξ is again in C as
wanted.
Lemma 3.3. Suppose that Ξ(z, w) = H(z)H(w)∗ is a positive L(Y)-valued kernel
on Ω0. Then Ξ is in C.
Proof of Lemma. Let us say that Ξ(z, w) = H(z)H(w)∗ where H : Ω → L(X ,Y) for
some coefficient Hilbert space X . Let ψ0 be any fixed test function in Ψ. It suffices
so that
to find another coefficient Hilbert space eX and a function G : Ω0 → L(eX ⊗ YT ,Y)
Ξ(z, w) = G(z)(cid:0)I eX ⊗ (I − ψ0(z)ψ0(w)∗)(cid:1) G(w)∗,
for then we have the needed representation Ξ(z, w) = Γ0(z, w)[I − E(z)E(w)∗] with
Γ0 given by
(cid:3)
Γ0(z, w)[g] = G(z)(I eX ⊗ g(ψ0))G(w)∗.
Toward this end, choose a unit vector y0 in YT and note that
0(I − ψ0(z)ψ0(w)∗)y0 = 1 − y∗
y∗
0ψ0(z)ψ0(w)∗y0
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
15
is invertible (as an element of C) by our underlying assumption (2.1). Moreover we
have the geometric series representation for the inverse:
1
=
(y∗
0ψ0(z)ψ0(w)∗y0)n
0ψ0(z)ψ0(w)∗y0
1 − y∗
0ψ0(z)ψ0(w)∗y0)n is a positive kernel due to the Schur multiplier
(3.9)
∞Xn=0
where each term (y∗
theorem (see e.g. [48, Theorem 3.7]). Thus there exist functions gn : Ω0 → L(eGn, C)
0ψ0(z)ψ0(w)∗y0)n = gn(z)gn(w)∗.
so that
(y∗
Then we may rewrite (3.9) as
1
1 − y∗
0ψ0(z)ψ0(w)∗y0
=
∞Xn=0
gn(z)gn(w)∗.
(3.10)
We conclude that
Ξ(z, w) = H(z)H(w)∗
···(cid:3) ,
∞Mn=1eGn.
eX =
(cid:3)
Let us now note that the assertion of the condition (2) in the statement of the
Theorem is that the kernel ΞS0 (z, w) := I − S0(z)S0(w)∗ is in C. As V is a locally
convex linear topological vector space and C is closed in V, by a standard Hahn-
Banach separation principle (see [51, Theorem 3.49b)]), to show that ΞS ∈ C it
suffices to show: Re L(ΞS) ≥ 0 whenever L is a continuous linear functional on V
such that Re L(Ξ) ≥ 0 for each Ξ ∈ C.
With this strategy in mind let us suppose that L is a continuous linear functional
on V such that Re L(Ξ) ≥ 0 for each Ξ ∈ C. We then define L1 on V by
L1(Ξ) =
1
2(cid:16)L(Ξ) + L(Ξ∨)(cid:17)
where we set
Easy properties are that
Ξ∨(z, w) = Ξ(w, z)∗.
L1(Ξ) = Re L(Ξ) if Ξ∨ = Ξ.
(3.11)
0ψ0(z)ψ0(w)∗y0)IX(cid:19) H(w)∗
0ψ0(z)ψ0(w)∗y0)IX ) H(w)∗
0ψ0(z)ψ0(w)∗y0)I eGn(cid:17) gn(w)∗H(w)∗
0)(cid:16)I eGn ⊗ (I − ψ0(z)ψ0(w)∗)(cid:17) (gn(w)∗ ⊗ y0)H(w)∗
1
=
=
1 − y∗
0ψ0(z)ψ0(w)∗y0 · (1 − y∗
H(z) (gn(z)gn(w)∗(1 − y∗
H(z)gn(z)(cid:16)(1 − y∗
H(z)(gn(z) ⊗ y∗
= H(z)(cid:18)
∞Xn=0
∞Xn=0
∞Xn=0
= G(z)(cid:0)I eX ⊗ (I − ψ0(z)ψ0(w)∗)(cid:1) G(w)∗
G(z) =(cid:2)H(z)(g1(z) ⊗ y∗
=
0) H(z)(g2(z) ⊗ y∗
0)
where we set
16
J.A. BALL AND M.D. GUERRA-HUAM ´AN
For ǫ > 0 be an arbitrarily small but positive number, we use the functional L1
to define an inner product on the space HL1,ǫ of functions f : Ω0 → Y by
hf, giHL1,ǫ
where we have set
= L1(∆f,g) + ǫ2 Xw∈Ω0
tr (∆f,g(w, w))
∆f,g(z, w) = f (z)g(w)∗.
(3.12)
By Lemma 3.3 we know that ∆f,f ∈ C and hence Re L(∆f,f ) ≥ 0. Since ∆f,f =
∆∨
f,f , as a consequence of (3.11) we know that Re L(∆f,f ) = L1(∆f,f ). From
is a positive semidefinite inner product.
these observations it follows that h·,·iHL1,ǫ
Hence we can take the Hausdorff completion of HL1,ǫ to arrive at a Hilbert space,
still denoted as HL1,ǫ.
For X a coefficient Hilbert space, we shall be interested in the space HL1,ǫ ⊗
C2(X , C). The following lemma is crucial.
Lemma 3.4. The space HL1,ǫ ⊗ C2(X , C) can be identified with the space (HL1,ǫ)X
consisting of C2(X ,Y)-valued functions f on Ω with inner product given by
tr (∆f,g(w, w))
(3.13)
hf, giHL1,ǫ)X = L1(∆f,g) + ǫ2 Xw∈Ω0
where ∆f,g has the same form as in (3.12) (but where now the middle space is X
rather than C):
∆f,g(z, w) = f (z)g(w)∗.
Proof of lemma. For convenience of notation we drop the ǫ-term in the inner prod-
uct as the ǫ > 0 case proceeds in the same way but with more cumbersome notation.
For f ⊗ x∗ a pure tensor in HL1 ⊗ C2(X .C) (so f ∈ HL1 and x ∈ X ∼= L(C,X )) and
similarly for f ′ ⊗ x′∗, we have
hf ⊗ x∗, f ′ ⊗ x′∗iHL1 ⊗C2(X ,C) =(cid:10)hf, f ′iHL1
x∗, x′∗(cid:11)C2(X ,C)
= L1(∆f,f ′ )x∗x′∗ = L1(∆f,f ′ x∗x′)
where the last step follows since x∗x′ is just a complex number. Next observe that
∆f,f ′(z, w)x∗x = f (z)f ′(w)∗(x∗x′) = f (z)(x∗x′)f ′(w)∗
= (f (z)x∗) (f ′(w)x′∗)∗ = ∆f ·x∗,f ′·x′∗ (z, w).
By extending this calculation to linear combinations of pure tensors, the result
follows.
(cid:3)
With the formulation of the space (HL1,ǫ)X in hand, it makes sense to ask
whether the right multiplication operator Rψ : f (z) 7→ f (z)ψ(z) defines a con-
traction operator from (HL1,ǫ)YT
. The answer is given by the next
lemma.
to (HL1,ǫ)UT
Lemma 3.5. For each test function ψ ∈ Ψ, the right multiplication operator Rψ
defines a contraction operator form (HL1,ǫ)YT
Proof of Lemma. Rψ is contractive if and only if
(HL1,ǫ)YT − kRψfk2
(HL1,ǫ)UT ≥ 0
to (HL1,ǫ)UT
kfk2
.
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
17
for all f ∈ (HL1,ǫ)YT
. This translates to the condition that
for all such f . Observe that
L1(∆f,f − ∆f ψ,f ψ) + ǫ2 Xw∈Ω0
∆f,f (z, w) − ∆f ψ,f ψ(z, w) = f (z)(I − ψ(z)ψ(w)∗)f (w)∗
[∆f,f (w, w) − ∆f ψ,f ψ(w, w)] ≥ 0
from which we see that the kernel Ξ := ∆f,f − ∆f ψ,f ψ is in the cone C: note that
the kernel Γ(z, w)[g] = f (z)g(ψ)f (w)∗ is completely positive since its Kolmogorov
decomposition (condition (3′) in Theorem 2.6) is exhibited. Thus Re L(Ξ) ≥ 0,
and hence, since Ξ = Ξ∨, also L1(Ξ) ≥ 0. The ǫ-term is also nonnegative since
kψ(w)k < 1 for each w ∈ Ω0. It now follows that kRψk ≤ 1 as asserted.
(cid:3)
To make use of the hypothesis that S ∈ SAΨ(U,Y), we need to convert the space
HL1,ǫ to a reproducing kernel space. This is done as follows; it is at this point that
we make use of the ǫ-regularization of the HL1-inner product.
Lemma 3.6. The space (HL1,ǫ)Y is isometrically equal to a reproducing kernel
Hilbert spaces H(K)Y for a positive kernel K ∈ KΨ(Y).
Proof of lemma. We wish to apply Theorem 2.5 with E and X equal to Y and with
Ω0 equal to Ω. To this end, we note that elements of (HL1,ǫ)Y are C2(Y)-valued
functions, at least on the dense set before the Hausdorff-completion step is carried
out in the construction of the space. However, the presence of the term with
the ǫ2 factor in the definition of the (HL1,ǫ)Y -inner product guarantees that the
point-evaluation map evw : (HL1,ǫ)Y → C2(Y) is bounded with norm at most 1/ǫ.
Hence condition (1) in Theorem 2.5 is verified. Condition (2) is straightforward
since (HL1,ǫ)Y is itself a tensor-product space HL1,ǫ ⊗ C2(Y, C). We conclude that
(HL1,ǫ)Y is isometrically equal to a reproducing kernel Hilbert space H(K)Y for a
uniquely determined L(Y)-valued positive kernel K.
Finally we must verify that K is Ψ-admissible. But this is an immediate conse-
quence of Lemma 3.5.
(cid:3)
To conclude the proof of Step 1 (the case where Ω0 if finite), we proceed as
follows. Let K be the positive kernel identified in Lemma 3.6. Since K ∈ KΨ(Y),
we use the assumption that S is in the Schur-Agler class SAΨ(U,Y) to conclude
that the operator RS of right multiplication by S is contractive from H(K)Y to
H(K)U . As Lemma 3.6 also tells us that H(K)Y is isometrically equal to (HL1,ǫ)Y ,
trivially we can also say that RS is contractive from (cid:0)HL1,ǫ(cid:1)Y to (cid:0)HL1,ǫ(cid:1)U . The
criterion for this to be the case is that
(HL1,ǫ)U ≥ 0 for all f ∈ (HL1,ǫ)Y ,
or, equivalently
kfk2
(HL1,ǫ)Y − kRSfk2
L1 (∆f,f − ∆f S0,f S0) + ǫ2 Xw∈Ω0
tr (∆f,f (w, w) − ∆f S0,f S0(w, w)) ≥ 0 for all f,
where ∆f,f (z, w) − ∆f S0,f S0(z, w) = f (z)ΞS(z, w)f (w)∗.
In particular, taking
f (z) = Pn for all z ∈ Ω0 where {Pn} is an increasing sequence of finite-rank
orthogonal projections converging strongly to the identity operator IY gives us
L1(PnΞS0Pn) + ǫ2 Xz∈Ω0
tr (PnΞS0 (z, z)Pn) ≥ 0.
18
J.A. BALL AND M.D. GUERRA-HUAM ´AN
As this holds for all ǫ > 0, we may take the limit as ǫ → 0 (while holding n fixed)
to get
L1(PnΞS0 Pn) ≥ 0
for all n. By the weak-∗ continuity of L1 we have that
L1(PnΞS0Pn) = L1(ΞS0).
lim
n→∞
(3.14)
Taking limits in (3.14) then gives us L1(ΞS0 ) ≥ 0. As ΞS0 = Ξ∨
S0, this gives us
finally Re L(ΞS0 ) ≥ 0 as required, and we conclude that S0 ∈ C as wanted. This
concludes the proof of Step 1.
Step 2: Ω0 is not necessarily finite.
We now remove that assumption that Ω0 is finite. It is now understood how this
step is efficiently handled as an application of the Kurosh Theorem (see [27, 29]).
By Step 1, we know that for each finite subset ΩF of Ω, there is an associated
completely positive kernel ΓF (not necessarily uniquely determined) so that the
Agler decomposition
ΞS0 (z, w) := I − S0(z)S0(w)∗ = ΓΩF (z, w)[I − E(z)E(w)∗]
(3.15)
holds for all z, w ∈ ΩF . To set up the Kurosh Theorem, for each finite subset
ΩF ⊂ Ω, we let ΦΩF denote the collection
ΦΩF = {Ξ : Ξ completely positive kernel such that (3.15) holds for z, w ∈ ΩF}.
By applying the argument used in the proof of Lemma 3.2, one can see that ΦΩF is
compact in the pointwise weak-∗ convergence topology inherited from the space of
L(Cb(Ψ,L(YT )),L(Y))-valued functions on Ω × Ω. The Kurosh Theorem (see [11,
page 75]) tells us that, for each finite subset ΩF of Ω, there is a choice of completely
positive kernel ΓΩF for which (3.15) holds on ΩF such that, in addition, whenever
ΩF , ΩF ′ are two subsets of Ω with ΩF ⊂ ΩF ′, then ΓΩF ′ΩF ×ΩF = ΓΩF . We may
then define a completely positive kernel Γ on all of Ω × Ω by
Γ(z, w) = ΓΩF (z, w) where ΩF finite,
z, w ∈ ΩF .
The construction guarantees that Γ is well defined and the fact that each ΓΩF is
completely positive on ΩF guarantees that Γ is completely positive as a kernel on
all of Ω. We have now completed the proof of (1) ⇒ (2) in Theorem 3.1.
(cid:3)
Proof of (2) ⇒ (3). We are given a completely positive kernel Γ on Ω0 so that (3.3)
holds for z, w ∈ Ω0. By condition (3′) in Theorem 2.6, Γ has a decomposition of
the form
Γ(z, w)[g] = H(z)ρ(g)H(w)∗
where H : Ω0 → L(X ,Y) for an auxiliary Hilbert space X which also carries a
∗-representation ρ of the C∗-algebra Cb(Ψ,L(YT )). From (3.3) we then deduce
I − S0(z)S0(w)∗ = Γ(z, w)[I − E(z)E(w)∗]
= H(z)ρ(I − E(z)E(w)∗)H(w)∗
= H(z)H(w)∗ − H(z)L∗
where we use (2.13). This in turn can be rearranged as
E(z)∗LE(w)∗H(w)∗
H(z)L∗
E(z)∗LE(w)∗H(w)∗ + I = H(z)H(w)∗ + S0(z)S0(w)∗
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
19
which leads to the inner product identity
hLE(w)∗H(w)∗yw, LE(z)∗H(z)∗yziCb(Ψ,L(YT ,UT )⊗X + hyw, yziY
for arbitrary yw and yz in Y. It then follows that the mapping V given by
= hH(w)∗yw, H(z)∗yzxi + hS0(w)∗yw, S0(z)∗yziU
S0(w)∗yw(cid:21)
(cid:21) 7→(cid:20)H(w)∗yw
V : (cid:20)LE(w)∗H(w)∗yw
yw
Y
yw
extends by linearity and continuity to a well-defined isometry from the subspace
onto the subspace
R := span(cid:26)(cid:20)H(w)∗yw
D := span(cid:26)(cid:20)LE(w)∗H(w)∗yw
(cid:21) : yw ∈ Y, w ∈ Ω(cid:27) ⊂(cid:20)Cb(Ψ,L(YT ,UT )) ⊗ X )
S0(w)∗yw(cid:21) : yw ∈ Y, w ∈ Ω(cid:27) ⊂(cid:20)X
U(cid:21) .
By replacing X with X ′ = X ⊕ eX where eX is an infinite-dimensional Hilbert space
Y (cid:3)⊖D and(cid:2) X ′
if necessary, we can arrange that the defect spaces(cid:2) X ′
U (cid:3)⊖R have the
same dimension. We may also assume that eX is equipped with some representation
eρ of Cb(Ψ,L(YT )) and hence X ′ is equipped with the representation ρ′ = ρ⊕eρ. We
Y(cid:3) ⊖ D = dim [ X
loss of generality we have dim(cid:2) X
i⊖D onto
We now let V0 be any unitary transformation fromh Cb(Ψ,L(YT ,UT ))⊗X
i ∼= D ⊕(cid:16)h Cb(Ψ,L(YT ,UT ))⊗X
i ⊖ D(cid:17)
U∗ = V ⊕ V0 : h Cb(Ψ,L(YT ,UT ))⊗X
U ] ∼= R ⊕ ([ X
now assume that all this has been done and drop the prime notation; thus without
U ] ⊖ R and X is equipped with a
∗-representation ρ of Cb(Ψ,L(YT )).
U ] ⊖ R) .
We may then write out U∗ as a block 2 × 2-matrix
→ [ X
[ X
U ] ⊖ R and set
Y
Y
Y
Since U∗ is an extension of V given by (3.16), we have
U =(cid:20)A∗ C∗
B∗ D∗(cid:21) : (cid:20)Cb(Ψ,L(YT ,UT )) ⊗ X
(cid:20)A∗ C∗
B∗ D∗(cid:21)(cid:20)LE(w)∗H(w)∗yw
(cid:21) →(cid:20)X
U(cid:21) .
(cid:21) =(cid:20)H(w)∗yw
S0(w)∗yw(cid:21) .
yw
Y
(3.16)
(cid:21)
(3.17)
The first row of (3.17) gives
A∗LE(w)∗H(w)∗yw + C∗yw = H(w)∗yw.
Since supψ{kψ(w)k} < 1 by the assumption (2.1) and since kA∗k ≤ 1 as U is
unitary, we see that I − A∗LE(w)∗ is invertible and, by the arbitrariness of yw ∈ Y,
we can solve (3.17) to get
Plugging this into the second row of (3.17) then gives
H(w)∗ = (I − A∗LE(w)∗)−1C∗.
B∗LE(w)∗(I − A∗LE(w)∗)−1C∗ + D∗ = S0(w)∗.
Taking adjoints and replacing w by z ∈ Ω0 leads to the realization formula (3.5).
20
J.A. BALL AND M.D. GUERRA-HUAM ´AN
We actually get a little bit more. The right-hand side of (3.5) makes sense for
z equal to any point in Ω. Thus we have actually proved: (2) ⇒ (3′) where the
precise statement of (3′) is:
(3′) There is a Ψ-unitary colligation U as in (3.4) such that S0 has an extension
to an L(U,Y)-valued function S defined on all of Ω having the transfer-
function realization
S(z) = D + C(I − L∗
E(z)∗A)−1L∗
E(z)∗B
for z ∈ Ω.
(3.18)
(cid:3)
Proof of (3) ⇒ (2). We assume that we have a transfer-function realization (3.5)
and we must produce a completely positive kernel Γ so that (3.3) holds. There is a
natural candidate, namely:
Γ(z, w)[g] = C(I − L∗
E(z)∗A)−1ρ(g)(I − A∗LE(w)∗)−1C∗.
(3.19)
E(z)∗A)−1 and π = ρ). The verification of (3.3) amounts to the identity
The candidate is certainly a completely positive kernel since the formula (3.19)
exhibits its Kolmogorov decomposition (condition (3′) in Theorem 2.6 with H(z) =
C(I − L∗
I − S0(z)S0(w)∗ = C(I − L∗
Using the realization formula (3.5) for S0(z) and the relations
E(z)∗A)−1ρ(I − E(z)E(w)∗)(I − A∗LE(w)∗)−1C∗. (3.20)
AA∗ + BB∗ = I, AC∗ + BD∗ = 0, CC∗ + DD∗ = I
coming out of the coisometric property UU∗ = I of U then give us
I − S0(z)S0(w)∗
= I − [D + C(I − L∗
= I − DD∗ − C(I − L∗
E(z)∗A)−1L∗
− C(I − L∗
E(z)∗A)−1L∗
E(z)∗B][D∗ + B∗LE(w)∗(I − A∗LE(w)∗)−1C∗]
E(z)∗BD∗ − DB∗LE(w)∗(I − A∗LE(w)∗)−1C∗
E(z)∗A)−1L∗
E(z)∗BB∗LE(w)∗(I − A∗LE(w)∗)−1C∗
= CC∗ + C(I − L∗
E(z)∗A)−1L∗
E(z)∗AC∗ + CA∗LE(w)∗(I − A∗LE(w)∗)−1C∗
+ C(I − L∗
E(z)∗A)−1L∗
E(z)∗(AA∗ − I)LE(w)∗(I − A∗LE(w)∗)−1C∗
= C(I − L∗
where we have set X equal to
E(z)∗A)−1X(I − A∗LE(w)∗)−1C∗
(3.21)
(3.22)
X = (I − L∗
E(z)∗A)(I − A∗LE(w)∗) + L∗
E(z)∗A)A∗LE(w)∗ + L∗
E(z)∗A(I − A∗LE(w)∗)
E(z)∗AA∗LE(w)∗ − L∗
+ (I − L∗
E(z)∗LE(w)∗
= I − L∗
E(z)∗LE(w)∗
= ρ(I − E(z)E(w)∗)
where we used (2.13) for the last step. Combining (3.21) and (3.22) gives us (3.20)
as required.
(cid:3)
Proof of (2) ⇒ (1) if dim YT < ∞. We assume that we have an Agler decomposi-
tion (3.3) and must show that S0 can be extended to an S defined on all of Ω which
is in the Schur-Agler class SAΨ(U,Y). Toward this end, we note that the proof
of (2) ⇒ (3) really proved (3′), i.e., that S0 extends to an S defined on all of Ω
given by the realization formula (3.18). Therefore the argument behind (3) ⇒ (2)
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
21
actually gives us an Agler decomposition (3.3) valid for the extended S which holds
for z, w in all of Ω. In this way we may assume that S is given to us defined on all
of Ω and we are given the completely positive kernel Γ on all of Ω giving rise to the
Agler decomposition (3.3) for S.
To check that S is in the Schur-Agler class SAΨ(U,Y), we must verify that the
operator RS of right multiplication by S is contractive from H(K)Y to H(K)U
for any choice of admissible kernel K ∈ KΨ(Y). Toward this end, we reverse the
procedure used in the proof of (1) ⇒ (2) as follows.
Given an admissible kernel K ∈ KΨ and given any finite collection of points
z1, . . . , zN ∈ Ω, we must show that the kernel (3.2) is a positive kernel for all choices
of functions Y : {z1, . . . , zn} → C2(Y,U). It suffices to consider the restriction K0
of K to the finite set Ω0 = {z1, . . . , zN}. Since K ∈ KΨ(Y), we know that the
right multiplication operator Rψ is contractive from H(K0)YT to H(K0)UT for each
ψ ∈ Ψ. Consider the modified kernel
K0,ǫ(z, w) = K0(z, w) + ǫ2 Xz∈Ω0
δz,wIY
where δz,w is the Kronecker delta function equal to 1 for z = w and 0 otherwise.
Since the values of ψ are contractive, we see that Rψ is still contractive as an
operator from H(K0,ǫ)YT to H(K0,ǫ)UT for each ǫ > 0. Also, to show that RS is
contractive from H(K0)Y to H(K0)U , it is enough to show that RS is contractive
from H(K0,ǫ)Y to H(K0,ǫ)U for each ǫ > 0.
Our next goal is to construct a kernel Lǫ : Ω0 × Ω0 → L(Y) so that
tr (Lǫ(z, w)f (z)g(w)∗) .
(3.23)
hf, giH(K0,ǫ) = Xz,w∈Ω0
To do this, define L(z, w) ∈ L(Y) by
hLǫ(z, w)u, viY = hδzu, δwviH(K0,ǫ
where δz is the point-mass function
δz(z′) =(1
0
if z = z′,
otherwise.
In terms of the kernel function K0,ǫ, one can verify the block-matrix identity
[Lǫ(z, w)]z,w,∈Ω0 = ([K0,ǫ(z, w)]z,w∈Ω0 )−1 .
The fact that Rψ : H(K0,ǫ)YT → H(K0,ǫ)UT is contractive can be equivalently ex-
pressed as
tr (Lǫ(z, w)f (z)(I − ψ(z)ψ(w)∗)f (w)∗) ≥ 0 for all f : Ω → C2(YT ,Y).
(3.24)
To show that RS : H(K0,ǫ)Y → H(K0,ǫ)U is contractive can be expressed in a similar
way as
Xz,w,∈Ω0
Xz,w∈Ω0
tr (Lǫ(z, w)h(z)(I − S(z)S(w)∗)h(w)∗) ≥ 0 for all h : Ω0 → C2(Y).
(3.25)
By assumption we are given an Agler decomposition (3.3) for S. The completely
positive kernel Γ appearing in (3.3) in turn has a Kolmogorov decomposition as in
22
J.A. BALL AND M.D. GUERRA-HUAM ´AN
(3′) in Theorem 2.6:
Γ(z, w)[g] = H(z)ρ(g)H(w)∗
(3.26)
for a ∗-representation ρ : Cb(Ψ,L(YT )) → L(X ). We now use the assumption that
dimYT < ∞. This has the effect that Cb(Ψ,L(YT )) is a CCR C∗-algebra and that
any representation ρ of Cb(Ψ,L(YT )) is the direct integral of multiples of irreducible
representations, where an irreducible representation π0 : C(Ψβ,L(YT )) → L(YT )
has the point-evaluation form π0(g) = g(ψ0) for some ψ0 ∈ Ψβ; we refer to [13] and
[35, Section 2.3] for fuller discussion. Thus we may assume that there are mutually
singular measures µ∞, µ1, µ2, . . . defined on the Borel subsets of the Stone- Cech
compactification Ψβ of Ψ so that
ρ = ∞ · πµ∞ ⊕ 1 · πµ1 ⊕ 2 · πµ2 ⊕ ···
where
πµj (g) : f (ψ) 7→ g(ψ)f (ψ) on Hπj := L2
YT (µj ) = L2(µj ) ⊗ YT
and where in general n · π refers to the n-fold inflation of π:
nMj=1
Hπ.
(n · π)(g) =
π(g)
. . .
π(g)
on (Hπ)n :=
∞Mr=1
H(w)∗ =(cid:20) H∞(w)∗
r=1 Hr(w)∗(cid:21)
YT (µ∞)∞ ⊕
YT (µr)r.
L2
col∞
X = L2
Thus we may assume that the representation space X in (3.26) decomposes as
Therefore the operators H(w)∗ appearing in (3.26) decompose as
where each Hr(w)∗ is an operator from Y to L2
an operator-valued function Hr(w, ψ)∗ of ψ ∈ Ψβ according to
YT (µr)r. This enables us to define
Hr(w, ψ)∗y = ((Hr(w)∗y) (ψ).
Then the adjoint Hr(z) of Hr(z)∗ is given via an integral formula:
Hr(z) : G(ψ) 7→ZΨβ
Hr(z, ψ)G(ψ)dµr(ψ).
We conclude that the Agler decomposition (3.3) takes the more detailed form
I − S(z)S(w)∗ =ZΨβ
H∞(z, ψ) (Iℓ2 ⊗ (I − ψ(z)ψ(w)∗)) H∞(w, ψ)∗ dµ∞(ψ)
+
∞Xr=1ZΨβ
Hr(z, ψ) (ICr ⊗ (I − ψ(z)ψ(w)∗)) Hr(w, ψ)∗dµr(ψ).
(3.27)
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
23
ZΨβ
∞Xr=1ZΨβ
Plugging this into the left-hand side of the desired inequality in (3.25) and taking
the integral to the outside gives us the sum over z, w ∈ Ω0 of the following terms:
tr (Lǫ(z, w)h(z)H∞(z, ψ) (Iℓ2 ⊗ (I − ψ(z)ψ(w)∗)) H∞(w, ψ)∗h(w)∗) dµ∞(ψ)+
tr (Lǫ(z, w)h(z)Hr(z, ψ) (ICr ⊗ (I − ψ(z)ψ(w)∗)) Hr(w, ψ)∗h(w)∗) dµr(ψ).
From (3.24) we see that the sum over z, w ∈ Ω0 of the integrand in each of these
terms is nonnegative. Hence the sum over z, w of the integrals in nonnegative and
(3.25) follows as required.
(cid:3)
Remark 3.7. The interpolation problem for the class SAΨ(U,Y) can be formu-
lated as follows: Given a subset Ω0 of Ω and a function S0 : Ω0 → L(U,Y), give
necessary and sufficient conditions for the existence of an S ∈ SAΨ(U,Y) such that
SΩ0 = S0. Assuming that dimYT < ∞, one gets a solution criterion (arguably not
particularly practical at this level of generality) immediately from the equivalence
(1) ⇔ (2) in Theorem 3.1 (where we use (2) in the more concrete form (3.27)):
the SAΨ(U,Y)-interpolation problem has a solution if and only if there exists a
matrix-valued function (ψ, z) 7→ Hψ(z) on Ψβ × Ω0, bounded and measurable in ψ
for each z, together with a finite measure µ on Ψβ, so that
I − S0(z)S0(w)∗ =ZΨβ
Hψ(z)(cid:0)IXψ ⊗ (I − ψ(z)ψ(w)∗)(cid:1) Hψ(w)∗ dµ(ψ)
for each z, w ∈ Ω0. Not so apparent from the way Theorem 3.1 is formulated is
that condition (1) by itself is also a criterion for solving the interpolation problem.
Indeed, if we set ΨΩ0 equal to the collection of restricted functions
ΨΩ0 = {ψΩ0 : ψ ∈ Ψ},
(3.28)
we may view ΨΩ0 as itself a collection of test functions generating a Schur-Agler
class SAΨΩ0(U,Y) of L(U,Y)-valued functions defined only on Ω0. The only part
of the hypothesis that S0 extends to an S ∈ SAΨ used to prove (1) ⇒ (2) in
Theorem 3.1 is that then S0 ∈ SAΨΩ0. We conclude that we get another criterion
for solution of the interpolation problem: the SAΨ(U,Y)-interpolation problem has
a solution if and only if S0 ∈ SAΨΩ0. Let us say that the subset K0
Ψ(Y) of the
set of admissible kernels KΨ(Y) is a generating set for KΨ(Y) if, for each kernel
Ψ(Y) such that K is congruent to K 0 in the
K ∈ KΨ(Y), there is a kernel K 0 ∈ K0
sense that there is an operator function Y so that K(z, w) = Y (z)K 0(z, w)Y (w)∗.
It is easy to check that the kernels of the form (3.2) are positive on Ω0 for all Y and
admissible K if and only if all such kernels are positive when the admissible K is
restricted to those coming from the generating set K0
Ψ(Y). Hence we arrive at the
following dual criterion for solution of the SAΨ(U,Y)-interpolation problem: the
SAΨ(U,Y)-interpolation problem has a solution if and only if the kernel
k(z, w) = tr(cid:0)Y (w)∗(I − S0(w)∗S0(z))Y (z)K 0(z, w)(cid:1)
is a positive kernel on Ω0 for all Y : Ω0 → C2(Y,U) for all admissible kernels K from
the generating set K0
Ψ(Y). We illustrate these ideas on the examples discussed in
Section 4 below. This duality pairing between admissible kernels and test functions
is central to the operator-algebra point of view of Paulsen and Solazzo toward
interpolation theory (see [45, 46, 47]).
24
J.A. BALL AND M.D. GUERRA-HUAM ´AN
ΨΩ0
There is also an operator-algebra point of view toward the Schur-Agler class.
For convenience in the following discussion, we take all the coefficient spaces U, Y,
UT , and YT to be the same space U although this probably is not essential. We
abbreviate the notation SAΨ(U,U) to SAΨ(U). Let ΨΩ0 be as in (3.28) and let
H ∞
(U) denote the space of all L(U)-valued functions S0 on the subset Ω0 of Ω
such that there exists a positive M < ∞ so that the kernel kX,S0,K,M given by (2.9)
is a positive kernel on Ω0 for all choices of X : Ω0 → C2(E,U) and for all choices
of K for which the kernel kY,ψ,K,1 is positive for all choices of Y : Ω0 → C2(U) and
ψ ∈ Ψ, or, what is the same, such that the right multiplication operator RS has
norm at most M as an operator on H(K)U for all positive kernels K for which
Rψ has norm at most 1 on H(K)U for all ψ ∈ Ψ. We define the H ∞
-norm
(U) is an
kSkH∞
operator algebra with unit ball equal to the Schur-Agler class SAΨΩ0(U). The
following representation-theoretic characterization of the Schur-Agler class will be
convenient in Section 4.1 below.
Theorem 3.8. Suppose that Ψ, Ω0 ⊂ Ω, and S0 are as in Theorem 3.1 with
In addition to conditions (1), (2), (3) in Theorem 3.1,
U = Y = UT = YT .
consider:
as the infimum of all such positive numbers M . Then H ∞
ΨΩ0
ΨΩ0
ΨΩ0
(4) For any representation π : H ∞
ΨΩ0
ψ ∈ Ψ, it also holds that kπ(S0)k ≤ 1.
(U) → L(K) such that kπ(ψ)k ≤ 1 for all
ΨΩ0
(U) → L(H(K)U ) sending G ∈ H ∞
Then (4) ⇒ (1). If dim U < ∞, then also (2) ⇒ (4) and (1), (2), (3), and (4) are
all equivalent.
Proof. Assume (4) holds and suppose that K ∈ KΨΩ0 (U) is an admissible kernel.
We now view the map πK : H ∞
(U) to
the right multiplication operator RG on H(K)U as a representation (technically,
an anti-representation, but this does not affect the final results). By definition of
K ∈ KΨΩ0 (U), we have πK (ψ)k ≤ 1 for each ψ ∈ Ψ. Condition (4) then tells us
that π(S0)k ≤ 1, i.e., RS0 on H(K)U has norm at most 1. In this way we have
verified condition (1).
Conversely, we suppose dimYT = dim U < ∞ and that condition (2) holds. As in
the proof of (2) ⇒ (1) we see that (2) can be written in the more explicit form (3.27).
Given any L(U)-valued kernel K(z, w) with a factorization K(z, w) = F (z)G(w)∗
with F, G ∈ H ∞
(U), we use the hereditary functional calculus to extend a given
representation π of H ∞
ΨΩ0
ΨΩ0
ΨΩ0
(U) to such kernels according to the rule
π (F (z)G(w)∗) = π(F )π(G)∗.
Applying π to (3.27) (and using continuity to push π past the integral sign) gives
I − π(S0)π(S0)∗ =ZΨβ
+
∞Xr=1ZΨβ
π (H∞(·, ψ)) (Iℓ2 ⊗ (I − π(ψ)π(ψ)∗) π (H∞(·, ψ))∗
dµ∞(t)
π (Hr(·, ψ)) (ICr ⊗ (I − π(ψ)π(ψ)∗)) π (Hr(·, ψ))∗
dµr(ψ).
From the fact that kπ(ψ)k ≤ 1 for each ψ ∈ Ψ we read off from this last expression
that kπ(S0)k ≤ 1 as well, i.e., (4) is verified.
Remark 3.9. In the proof of Theorem 3.1 we drew on a lot of ideas which have
been used in previous versions of this type of result, starting with the seminal
(cid:3)
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
25
paper of Agler [3] and continuing with [5, 22, 10, 23, 32, 53, 16, 19, 8, 27, 29] as
well as commutant lifting versions [23, 21, 9, 41]. In particular, the cone separation
argument in the proof of (1) ⇒ (2) and the proof of (2) ⇒ (3) (the so-called
lurking-isometry argument) go back to [3]. However there are some new technical
difficulties in the test-function setting where some new ideas are required in order
to arrive at the final result; we now discuss some of these.
In the proof of (1) ⇒ (2), the use of the ǫ2-perturbation term in the definition of
the HL1,ǫ norm is the ploy needed to make the point-evaluations f 7→ f (w) bounded
and enables us to avoid the hypothesis that the set of test functions Ψ separates
the points of any finite subset ΩF of Ω, as used in [27, 29].
Our proof of (2) ⇒ (1) (with the hypothesis that dimYT < ∞) is close to the
proof of (3) ⇒ (1) in [29] (for the scalar-valued case) (which actually involves use of
the representation-theory formulation (4)). These authors make use of the spectral
theorem for a representation of Cb(Ψ, C), approximating a general representation ρ
by a "simple representation" (approximation of the general integral in (3.27) by a
simple-function integrand). Thus their proof also makes use of the CCR character
of Cb(Ψ, C), and hence does not appear to extend to the case dimYT = ∞.
4. Algebras arising from test functions
In this section, rather than starting with a set of test functions Ψ, we assume that
we are given a function algebra A and then seek to determine a set of test functions
ΨU ,Y so that the unit ball of the operator-valued version of A, say A ⊗ L(U,Y)
where U, Y are two coefficient Hilbert spaces, can be identified as the associated
Schur-Agler class SAΨU,Y (U,Y).
The classical example is the Hardy algebra over the unit disk A = H ∞(D). The
operator-valued version A ⊗ L(U,Y) has unit ball equal to the classical operator-
valued Schur class S(U,Y),
for which we have the now classical result: S ∈
S(U,Y) if and only if the associated de Branges-Rovnyak kernel KS(z, w) = [I −
S(z)S(w)∗]/(1− zw) is a positive kernel on D. If we let KS(z, w) = H(z)H(w)∗ be
the Kolmogorov decomposition of KS, then we arrive at
I − S(z)S(w)∗ = H(z) ((1 − zw)IX ) H(w)∗
which is exactly the Agler decomposition (3.3) corresponding to the singleton collec-
tion of test functions Ψ = {ψ0} with ψ0 equal to the coordinate function: ψ0(z) = z.
For this case, moving from the scalar-valued case to the matrix- or operator-valued
case necessitates no change in the choice of test-function set Ψ. A similar story holds
for the case of the Schur-Agler class over the polydisk [22], the Schur-multiplier class
over the Drury-Arveson space [23, 32], and the Schur-Agler class over more general
domains in Dd with matrix polynomial or analytic defining function [16, 9]. How-
ever the situation for the case where A is the algebra of bounded analytic functions
over a finitely connected planar domain R, or where A is the constrained Hardy
algebra over the unit disk (bounded holomorphic functions f on D with the extra
constraint that f ′(0) = 0) is quite different. We discuss each of these in turn.
4.1. The Schur class over a multiply connected planar domain. We let R
denote a bounded domain (connected, open set) in the complex plane C whose
boundary consists of m + 1 smooth Jordan curves ∂0, ∂1, . . . , ∂m with ∂0 denoting
the boundary of the unbounded component of the complement of R in C. We
let SR denote the space of holomorphic functions mapping R into the unit disk,
26
J.A. BALL AND M.D. GUERRA-HUAM ´AN
and SR(U,Y) the operator-valued version consisting of holomorphic functions on
R with values in the closed unit ball BL(U,Y) of bounded linear operators between
two coefficient Hilbert spaces U and Y. In [28] there was identified a collection of
inner functions {sx : x ∈ TR}, normalized to have value 1 at a fixed point ζ0 ∈ ∂0
and to satisfy s(t0) = 0 at a fixed point t0 ∈ R, having exactly m zeros in R (the
minimal number possible for a single-valued inner function on R), and indexed by
x belonging to the R-torus TR := ∂0 × ∂1 × ··· × ∂m, so that any scalar Schur
class function s ∈ SR has an Agler decomposition (3.3) with respect to the family
Ψ = {ψx : x ∈ TR} s in (1.11) (or (3.27) specialized to this case):
1 − s(z)s(w) =ZTR
hx(z)(cid:16)1 − sx(z)sx(w)(cid:17) hx(w) dν(x).
In more detail, the functions sx are constructed as follows. Let φ = {φ1, . . . , φm}
be real-valued continuous functions on ∂R such that
{φ1, . . . , φm} = basis for L2(ωt0) ⊖ [H 2(ωt0 ) + H 2(ωt0 )]
(4.2)
where ωt0 is the harmonic measure on ∂R for some fixed point t0 ∈ R (so h(t0) =
ciated Hardy space, and the overline indicates complex conjugation -- see e.g. [33].
0 , wx
Then given x = (x0, x1, . . . , xm) ∈ TR, there is a unique choice of weights wx
1 ,
. . . , wx
R∂R h(ζ) dωt0(ζ) for h harmonic on R and continuous on R−), H 2(ωt0) is the asso-
m, each positive with sum equal to 1, so that
(4.1)
wx
r φi(xr) = 0 for i = 1, . . . , m
(4.3)
mXr−0
(see [4, Theorem 3.1.17]). Given any x and the associated weights (wx,
we associate the probability measure on ∂R:
wx
µx :=
r δxr
0 wx
1 , . . . , wx
m)
where δxr is the unit point-mass measure at xr. The constraint (4.3) guarantees
that the harmonic function
mXr=0
hx(z) =Z∂R Pz(ζ) dµx(ζ)
(where Pz(ζ) is the poisson kernel normalized to have Pt0 (ζ) = 1) has single-valued
harmonic conjugate. We then define fx(z) to be the unique holomorphic function
on R with
Re fx(z) = hx(z) and fx(t0) = 1.
Finally we set
sx(z) =
fx(z) − 1
fx(z) + 1
.
(4.4)
Then sx are the inner functions appearing in (4.1), apart from the additional nor-
malization that sx(ζ0) = 1 at a fixed ζ0 ∈ ∂0. Then it is shown in [29] that SR =
SAΨR with the collection of test functions ΨR taken to be ΨR = {sx : x ∈ TR}.
There it is shown, at least for the annulus case (m = 1), that, with the addi-
tional normalization sx(ζ0) = 1 imposed, that ΨR is minimal in the sense that no
nonempty open subset of TR can be omitted and still have the decomposition (4.1)
hold for all s ∈ SR.
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
27
Before explaining the matrix generalization of (4.4), we first recall some ideas
from [20]. Suppose that we are given a collection
φ(1)
1
...
φ(1)
m
φ(n)
1
...
φ(n)
m
φ(1)
1 IN
...
φ(1)
m IN
φ(n)
1 IN
...
φ(n)
m IN
of n vectors in Rm. From φ we form the block column vectors
φ =
, . . . , φ(n) =
φ(1) =
φ ⊗ IN =
φ(1) ⊗ IN :=
, . . . , φ(n) ⊗ IN :=
in(cid:0)CN ×N(cid:1)m
the zero element 0 =" 0N ×N
0N ×N# of(cid:0)CN ×N(cid:1)m
there exist positive semidefinite N × N matrices W1, . . . , Wn withPn
nXr=1
. We say that 0 is in the interior of the C∗-
where we set φ(r) ⊗ Wr =
(m × 1-column vectors with entries of size N × N ). We then say that
is in the C∗-convex hull of φ ⊗ IN if
r=1 Wr = IN
convex hull of φ ⊗ IN if in addition the matrix weights {W1, . . . , Wn} have the
property that their range spaces {Ran W1, . . . , Ran Wn} are φ-constrained weakly
independent by which we mean: whenever T1, . . . , Tn are N × N complex Hermitian
matrices with Ran Tr ⊂ Ran Wr for each r = 1, . . . , n such that
1 Wr
φ(r)
...
φ(r)
m Wr
φ(r) ⊗ Wr
so that
(4.5)
0 =
...
Tr = 0 and
φi(xr)Tr = 0 for i = 1, . . . , n,
nXr=1
nXr=1
it follows that Tr = 0 for each r = 1, . . . , n. When all this happens, we refer to
{W1, . . . , Wn} as a choice of matrix barycentric coordinates of 0 with respect to φ.
By way of motivation for these notions, note that, in case N = 1 and all the
weights W1 = w1, . . . , Wn = wn (now complex numbers) are nonzero (which can
be arranged simply by discarding appropriate vectors φ(r) from the list of vectors
φ), then 0 = 0 ∈ Rm in the interior of the C∗-convex hull of φ ⊗ I1 = φ simply
means that the vector 0 ∈ Rm is in the interior of the simplex generated by the
vectors φ(1), . . . , φ(n) and that w1, . . . , wn are the classical barycentric coordinates
for 0 with respect to the simplex vertices φ(1), . . . , φ(m).
We are now ready to explain the matrix analogue of the R-torus TR used to
parametrize the set of scalar test functions (4.4). We define the matrix R-torus TN
R
to consist of all pairs (x, w) of the form (x, w) = (x1, . . . , xn; W1, . . . , Wn) where
x1, . . . , xn is a set of n distinct points in ∂R such that 0 is in the interior of the
C∗-convex hull of the set of vectors φ(x) ⊗ IN , where we set
φ(x) =
φ(x1) =
φ1(x1)
...
φm(x1)
, . . . , φ(xn) =
φ1(xn)
...
φm(xn)
,
(4.6)
28
J.A. BALL AND M.D. GUERRA-HUAM ´AN
with φ1, . . . , φm as in (4.2), and with {W1, . . . , Wn} is a choice of matrix barycentric
coordinates for 0 with respect to φ(x) ⊗ IN . In particular, the condition (4.5) in
the present context specializes to
φi(xr)Wr = 0 for i = 1, . . . , m.
(4.7)
nXr=1
For the case N = 1, necessarily n = m + 1, after a reindexing the collection of
points (x0, x1, . . . , xm) necessarily consists of exactly one point from each bound-
ary component ∂0, . . . , ∂m, and the associated scalar weights wx
m are
uniquely determined by x. For N > 1, the characterization of TN
R is not so explicit;
nevertheless it is nonempty and is a well-defined metrizable topological space which
is in one-to-one correspondence with a collection of quantum measures (positive ma-
trix measures with total mass equal to the identity matrix IN ) which we define next.
For additional information we refer to [20].
1 , . . . , wx
0 , wx
Given (x, w) ∈ TN
R, we associate a quantum measure µx,w by
µx,w =
nXr=1
Wrδxr if (x, w) = (x1, . . . , xn; W1, . . . , Wn) ∈ TN
R.
(4.8)
Then a consequence of (4.7) is that the matrix-valued harmonic function
Hx,w(z) =Z∂R Pz(ζ) dµx,w(ζ)
has a single-valued (matrix-valued) harmonic conjugate, and hence there is a unique-
ly determined holomorphic function Fx,w on R with
Re Fx,w(z) = Hx,w(z) and Fx,w(t0) = IN .
It can be shown that the collection of functions
{Fx,w : (x, w) ∈ TN
R}
(4.9)
is exactly the set of extreme points for the compact convex set HN (R)I of normal-
ized Herglotz functions over R given by
HN (R)I = {F : R 7→ CN ×N : F holomorphic, Re F (z) ≥ 0 for z ∈ R, F (t0) = IN}.
Finally, we set
Sx,w(z) = (Fx,w(z) + I)−1(Fx,w(z) − I).
(4.10)
Note that each Sx,w(z) is an N × N matrix inner function on R normalized to
satisfy S(t0) = 0. Then in [20] it is shown that any matrix-valued function S in
the Schur class SR(CN , CN ) has an Agler decomposition of the form
I − S(z)S(w)∗ =ZTN
Hx,w(z) (I − Sx,w(z)Sx,w(w)∗) Hx,w(w)∗ dν(x, w)
(4.11)
R
for appropriate matrix functions Hx,w(z) and probability measure ν on TN
R.
Following the arguments in [29] (adapted to the matrix-valued setting) leads to
the following identification of the matrix Schur class SR(CN ) with a matrix-valued
; the main ingredients of the proof also appear in
test-function Schur class SAΨN
the more involved proof of Theorem 4.4 below.
R
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
29
Theorem 4.1. Let ΨN
R be the collection of matrix inner functions
R = {Sx,w : (x, w) ∈ TN
ΨN
R}
(4.12)
with Sx,w as in (4.10), with the additional normalization Sx,w(ζ0) = IN at some
fixed point ζ0 ∈ ∂0. Then the matrix-valued Schur class SR(CN ) is identical to the
matrix-valued test-function Schur-Agler class SAΨN
associated with the collection
of test functions ΨN
R (as defined by (3.1) and (3.2)).
R
Combined with Theorem 3.1 and Remark 3.7, we arrive at the following dual
formulations of interpolation criteria for the Nevanlinna-Pick interpolation problem
for the matrix Schur class over R. Before stating the result we need a little more
background concerning function theory on R. There is a standard procedure (see
e.g.[1]) for introducing m disjoint simple curves γ1, . . . , γm so that R\ γ (where we
set γ equal to the union γ = γ1 ∪ ··· ∪ γm) is simply connected. For each cut γr
we assign some orientation, so that points z not on γr but in a sufficiently small
neighborhood of γr in R can be assigned a location of either "to the left" or "to the
right". For f a vector-valued function on R and z a point on some γr, we let f (z+)
denote the limit of f (ζ) as ζ approaches z from the right of γr in R, and similarly,
f (z−) the limit of f (ζ) as ζ approaches z from the left of γr in R, whenever these
limits exist. Given a U = (U1, . . . , Um) in U(N )m (m-tuples of unitary N × N
matrices), we define a Hardy space H 2(U) to consist of functions f : R → CN ,
holomorphic on R \ γ, subject to the jump conditions f (z−) = Urf (z+) for z ∈ γr
for each r = 1, . . . , m (so kf (z)k2 is continuous and single-valued on R), and so
that the well-defined integral
H2(U) =Z∂R kf (ζ)k2 dωt0
kfk2
is finite. Then the space H 2(U) is a reproducing kernel Hilbert space over R (with
some appropriate convention as to how elements are defined on γ); we denote
its CN ×N -valued reproducing kernel function by K U: H 2(U) = H(K U). These
kernels enter into the admissible-kernel formulation of the criterion for the SR(CN )-
interpolation problem to have a solution.
Theorem 4.2. Suppose that we are given an N × N matrix-valued function S0 on
the subset R0 of R. Then the following are equivalent:
(1) There is a function S in the Schur class SR(CN ) with SR0 = S0.
(2) There is a matrix-valued function ((x, w), z) 7→ Hx,w(z) on TN
R × R0,
bounded and measurable in (x, w) for each z ∈ R0, together with a finite
measure ν on TN
R so that
I−S0(z)S0(w)∗ =ZTN
R
for all z, w ∈ R0.
kernel
Hx,w(z)(cid:0)IXx,w ⊗ (I − Sx,w(z)Sx,w(w)∗(cid:1) Hx,w(w)∗ dν(x, w)
(3) For each U = (U1, . . . , Um) in U(N )m and for each Y : R0 → CN ×N , the
k(z, w) := tr(cid:0)Y (w)∗(I − S0(w)∗S0(z))Y (z))K U(z, w)(cid:1)
is a positive kernel on R0.
(4.13)
30
J.A. BALL AND M.D. GUERRA-HUAM ´AN
Proof. The equivalence of (1) and (2) is a consequence of Theorem 3.1, once the
result of Theorem 4.1 is plugged in.
The equivalence of (1) and (3) is a consequence of Remark 3.7, once it is verified
that the set
(ΨN
R)0 := {K U : U ∈ U(N )m}
(4.14)
(CN ). Rather than doing
is a generating set for the set of admissible kernels KΨN
this, we observe that a solution criterion for the SAR(CN )-interpolation problem
was obtained in [17, Theorem 1.5] (as a consequence of the lifting theorem from
[14]), but in a somewhat different, more convoluted form than the form (4.13). If
one works with right multiplication operators on the space H(K U )CN rather than
with left multiplication operators on a left-side tensoring of the reproducing kernel
Hilbert space consisting of row-vector functions as is done in [17], one arrives at
the solution criterion (4.13) as presented here.
(cid:3)
R
Remark 4.3. We note that the scalar-valued case N = 1 of criterion (3) in Theo-
rem 4.2 is due to Abrahamse [1] -- note that the extra parameter Y (z) washes out
in this case. It was later shown by Ball-Clancey [18] that no open subset of the
kernels K U (U ∈ U(1)m) can be omitted for the validity of this result. However, for
the case of the annulus, if one takes the set of interpolation nodes R0 to be finite
and prespecified, then two kernels suffice [54]. While the Abrahamse result extends
to the matrix-valued setting for the annulus case (using only scalar-valued kernels),
McCullough and Paulsen [39, 40], using the C∗-algebra approach to interpolation
theory, showed that the Fedorov-Vinnikov result fails for the matrix-valued case.
All this story is reviewed nicely in [26]. We do not address such minimality issues
here.
For the case of the annulus (m = 1), by using results of McCullough [38] it is
possible to obtain a more explicit test-function collection as follows. We take R to
have the concrete form R = Aq where
Aq = {z ∈ C : q < z < 1}
for a number q satisfying 0 < q < 1. It is established in [38] that there is a curve
t 7→ ϕt of inner functions on Aq (constructed from the Ahlfors function for Aq
based at the point √q ∈ Aq) with the following property: for a (U, t) ∈ U(N ) × Tn
(where U(n) denotes the set of N × N unitary matrices and Tn is the N -torus
{t = (t1, . . . , tn) : tj = 1 for 1 ≤ j ≤ N}), set
ΦU,t(z) = U
ϕt1 (z)
. . .
ϕtN (z)
and
then, for each (x, w) ∈ TN
(U, t) ∈ U(N ) × TN so that
RU,t(z) = (IN + ΦU,t(z))(I − ΦU,t(z))−1;
Aq there is a choice of invertible N × N matrix X and a
Fx,w(z) + Fx,w(z)∗ = X (RU,t(z) + RU,t(z)∗) X ∗ for all z ∈ Aq.
We are now ready to introduce a new test-function class for SN
Aq
, namely:
(4.15)
(4.16)
eΨN
Aq = {ΦU,t : (U, t) ∈ U(N ) × TN}.
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
31
Aq
Aq
Aq
where eΨN
Conversely suppose that S ∈ SAq (CN ). To show that S ∈ SA eΨN
We then have the following result.
Theorem 4.4. The matrix-valued Schur class over the annulus SAq (CN ) is identi-
cal to the matrix-valued test-function Schur-Agler class SA eΨN
Aq is given
by (4.16).
Proof. Suppose first that S ∈ SA eΨN
. Then the right multiplication operator RS
(CN ). Such kernels
is contractive on H(K)CN for each admissible kernel K in K eΨN
include the Fay kernel associated with the Hardy space H 2(ωt)⊗ CN over Aq. This
observation is enough to conclude that S ∈ SAq (CN ).
(Cn), by
(CN ) → L(K)
Theorem 3.8 it suffices to show:
such that kπ(ΦU,t)k ≤ 1 for all (U, t) ∈ U(N ) × TN , it follows that kπ(S)k ≤ 1.
By replacing π with r · π with r < 1 and then taking a limit as r tends to 1,
without loss of generality we may suppose that kπ(ΦU,t)k < 1 for each (U, t). Then
π(RU,t) = (I − π(ΦU,t))−1(I + π(ΦU,t)) is a well-defined bounded operator on K
such that
π(RU,t) + π(RU,t)∗ = 2 (I − π(ΦU,t))−1 (I − π(ΦU,t)π(ΦU,t)∗) (I − π(ΦU,t)∗)−1 > 0.
(4.17)
From (4.15), we see that, for each fixed (x, w) ∈ TN
Aq , π (Fx,w) is a well-defined
bounded operator on K satisfying
for any representation π : H ∞
eΨN
Aq
Aq
π (Fx,w) + π (Fx,w)∗ = (X ⊗ IK) (π(RU,t) + π(RU,t)∗) (X ∗ ⊗ IK) .
(4.18)
From (4.17) we read off that π (Fx,w) has positive real part. We next obtain π(Sx,w)
as a Cayley transform of π (Fx,w):
π(Sx,w) = (π(Fx,w) + I)−1 (π(Fx,w) − I) .
From the relation
I − π(Sx,w)π(Sx,w)∗ = 2 (π(Fx,w) + I)−1 (π(Fx,w) + π(Fx,w)∗) (π(Fx,w)∗ + I)−1
combined with (4.18), we see that kπ(Sx,w)k ≤ 1. Finally, since S ∈ SAq (CN ),
S has an Agler decomposition as in (4.11). Applying the hereditary functional
calculus with the representation π through this integral representation gives
π(Hx,w)(I − π(Sx,w)π(Sx,w)∗)π(Hx,w)∗ dν(x, w).
I − π(S)π(S)∗ =ZTN
R
Since kπ(Sx,w)k ≤ 1 for each (x, w) ∈ TN
that kπ(S)k ≤ 1
Aq , we read off from this last expression
(cid:3)
As a corollary of Theorem 4.4 combined with Theorem 3.1, we get the following
structure theorem for the Schur-Agler class over the annulus Aq. To this end we
introduce the space bTN
Aq = (U(N )/U(1)N ) × TN , where here U(1)N is identified
with unitary diagonal N × N matrices, and the action of U(1)N on U(N ) is given
by
u1
u : U 7→ U u for u =
. . .
uN
∈ U(1)N .
32
J.A. BALL AND M.D. GUERRA-HUAM ´AN
Φ[U],t = ΦU,tU ∗.
Aq , we abuse notation somewhat and set
Note that Φ[U],t is well-defined (independent of the choice of representative of the
coset [U ]). Note that each Φ[U],t is normalized to satisfy Φ[U],t(1) = IN as well
For ([U ], t) ∈eTN
as Φ[U],t(√q) = 0. Furthermore the expression I − ΦU,t(z)ΦU,t(w)∗ is independent
of choice of coset representative for [U ]. Also it is easily checked that the set of
admissible kernels KΨ associated with a given collection of test functions Ψ depends
on the functions ψ ∈ Ψ only through the expressions I − ψ(z)ψ(w)∗. Hence the
result of Theorem 4.4 can equally well be stated as:
(CN )
(4.19)
where we have set
Aq
Aq (Cn) = SA bΨN
SN
Aq = {Φ[U],t : ([U ], t) ∈bTN
bΨN
Aq}.
Then the following corollary is an immediate consequence of our man theorem on
the test-function Schur-Agler class, namely Theorem 3.1.
Corollary 4.5. Suppose that S ∈ SAq (CN ). Then the following hold:
(1) S has an Agler decomposition of the form
I − S(z)S(w)∗
=ZbTN
(2) There is a representation ρ of C(bTN
unitary colligation matrix
Aq
H[U],t(z)(cid:0)IX[U ],t ⊗ (I − Φ[U],t(z)Φ[U],t(w)∗(cid:1) H[U],t(w)∗ dν([U ], t). (4.20)
Aq ,L(CN )) on a Hilbert space X and a
U =(cid:20)A B
C D(cid:21) : (cid:20) XCN(cid:21) →(cid:20) XCN(cid:21)
so that S has the transfer-function realization
S(z) = D + C(I − ρ(E(z))A)−1ρ(E(z))B.
Remark 4.6. An appealing conjecture is that the Agler decomposition (4.20) is
minimal in the sense of [29, Section 5.1] and [30, Section 3.6].
4.2. The constrained Schur class over the unit disk. Following [26, 17], we
define the constrained Hardy space H ∞
1 over the unit disk D to consist of bounded
analytic functions s on D such that s′(0) = 0. One can check that this is still an
algebra. In this section we identify a class of test functions ΨN
1 for which the unit
ball B(H ∞)N ×N of the algebra of N × N matrices over H ∞
(with norm equal to
the multiplier norm as multiplication operators on (H 2)N ) can best identified as
the test-function Schur-Agler class SAΨN
The analysis parallels that of Section 4.1 for the Schur class over a finitely con-
nected planar domain. One first identifies the extreme points for the Herglotz class
HN
1 consisting of N × N matrix-valued functions F on D satisfying the normal-
ization F (0) = I together with the side constraint F ′(0) = 0. Such functions are
exactly the Cayley transforms
(CN ).
1
1
F (z) = (I − S(z))−1(I + S(z))
of functions S in the closed unit ball B(H ∞
Hardy algebra (H ∞
1 )N ×N of the matrix-valued constrained
1 )N ×N subject to the normalization S(0) = 0. As is the case
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
33
for any matrix-valued Herglotz function on D, there is a positive matrix-valued
measure µ on T so that F has the Herglotz representation
F (z) =ZT
ζ + z
ζ − z
dµ(ζ).
The constraint that F (0) = IN is equivalent to µ(T) = IN ; following the termi-
nology used in Section 4.1, we then say that µ is an N × N quantum probability
measure. The constraint that F ′(0) equals zero (i.e., that F ∈ HN
1 ) imposes the
constraints on the measure µ:
Taking real and imaginary part then gives us two real constraints
ζ dµ(ζ) = 0.
F ′(0) =ZT
ZT
ζ−1dµ(ζ) =ZT
ZT
Re ζ dµ(ζ) = 0,
Im ζ dµ(ζ) = 0.
(4.21)
We thus see that the convex set HN
over D) is affinely equivalent to the convex set of measures
I (the constrained matrix-valued Herglotz class
CN
1 = {µ : µ = N × N quantum probability measure such that (4.21) holds}.
This convex set of measures is compact in the weak-∗ topology (viewing complex
N ×N matrix-valued measures as the dual space of CN -valued continuous functions
on T) and hence, by the Kreın-Milman theorem, has extreme points. By the same
general results from [20] leading to the the identification of the set (4.9) of the
normalized Herglotz class H(R)I over the planar domain R, it follows that the
extreme points of CN
(t, w) where t = (t1, . . . , tn) is an n-tuple of points on the unit circle T (with
1 ≤ n ≤ 3N ) and w = (W1, . . . , Wn) is an n-tuple of N × N matrix weights such
that the following property holds: 0 = 0 ⊗ IN is in the interior of the C∗-convex
hull of φ(t) ⊗ IN , where we set
1 can be described as follows. We let bΘN consist of all pairs
φ(t) =(cid:26)(cid:20)Re t1
Im t1(cid:21) , . . . ,(cid:20)Re tn
Im tn(cid:21)(cid:27) ⊂ R2.
with a choice of matrix barycentric coordinates of 0 with respect to φ(t) ⊗ IN equal
to {W1, . . . , Wn} (refer back to Section 4.1 for the definition of terms). One conse-
quence of the definitions is that, for any such (t, w) = (t1, . . . , tn; W1, . . . , Wn) in
bΘN , it holds that
(Re tr)Wr = 0,
nXr=1
nXr=1
(Im tr)Wr = 0.
(4.22)
the unit disk given by
Associated with each (t, w) ∈ bΘN is a holomorphic N × N -matrix function on
Ft,w(z) =
Wr.
nXr=1
tr + z
tr − z
These functions are holomorphic on D with positive real part, and moreover, as a
consequence of (4.22), have the property that F ′
t,w(0) = 0. In fact, it can be shown
34
J.A. BALL AND M.D. GUERRA-HUAM ´AN
that the set of all such functions {Ft,w : (t, w) ∈ TN
points for the normalized constrained Herglotz class over D, i.e., the class
1 )IN :={F : D → CN ×N : F holomorphic, Re F (z) ≥ 0 for z ∈ D,
1 } is exactly the set of extreme
(HN
F (0) = IN , F ′(0) = 0}.
By using Choquet theory it then follows that a general element F of (HN
an integral representation of the form
1 )IN has
F (z) =Z bΘN
for some probability measure on bΘN .
We note that (HN
strained Schur class
Ft,w(z) dν(t, w)
1 )IN is exactly the Cayley transform of the normalized con-
(SN
1 )0 ={S : D → CN ×N : S holomorphic, kS(z)k ≤ 1 for z ∈ D,
S(0) = 0, S′(0) = 0},
i.e.,
S ∈ (SN
F ∈ (HN
1 )0 ⇔ F := (I − S)−1(I + S) ∈ (HN
1 )IN ⇔ S := (F + I)−1(F − I) ∈ (SN
1 )IN ,
1 )0.
In particular, for each (t, w) ∈ TN
turn leads us to the following collection of functions in (SN
1 we may define functions St,w ∈ (SN
1 )0:
1 )0 which in
ΨN
1 := {St,w(z) = (Ft,w(z) + I)−1(Ft,w(z) − I) : (t, w) ∈ bΘN}.
Following the proof of Theorem 5.4 in [20] (the parallel result for the matrix Schur
class over a planar domain R in place of SN
1 ) then leads to the integral Agler
decomposition for the normalized constrained Schur class: given S ∈ (SN
1 )0 there is
a function ((t, w), z) 7→ Ht,w(z) on bΘN×D, bounded and measurable in (t, w) ∈ TN
for each fixed z, together with a probability measure ν on bΘN , so that
I − S(z)S(w)∗ =Z bΘN
Ht,w(z) (I − St,w(z)St,w(w)∗) Ht,w(w)∗ dν(t, w).
(4.24)
(4.23)
1 )N ×N with kS(0)k < 1), then
If S is in the strict constrained Schur class (S ∈ B(H ∞
there is a choice of matrix Mobius transformation on the N × N -matrix ball TS(0)
so that TS(0)[S(z)] is in the normalized constrained Schur class (SN
1 ) (see e.g. [20,
Section 5]. Using this one can see that functions S in the strict but unnormalized
Schur class SN
1 )N ×N have the continuous Agler decomposition (4.24) as
well.
1 := B(H ∞
1
Once this Agler decomposition is in hand, by using the same techniques as used in
the proofs of Theorems 4.1 (adaptations to the matrix-valued setting of arguments
in [29] and [30]), one can arrive at the following result.
Theorem 4.7. With ΨN
1 ⊂ (SN
1 )0 given by (4.23), we have the identity
B(H ∞
1 )N ×N = SAΨN
.
1
There is also a dual pair of solution criteria for the interpolation problem for the
1 . We first need to introduce the generating set of admissible kernels for the
β ] with
(CN ) as follows. For each isometric 2N × 1 matrix, written as [ α
class SN
class KΨN
1
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
35
α and β equal to N × 1 column vectors satisfying α∗α + β∗β = 1, we introduce the
collection of N × N -matrix kernel functions
(ΨN
1 )0 ={K α,β(z, w) := (αz + zβ)(α∗ + wβ∗) +
α, β ∈ CN ×1, α∗α + β∗β = 1}.
z2w2
1 − zw
IN :
(4.25)
Then we have the following result.
Theorem 4.8. Suppose that we are given an N × N matrix-valued function S0 on
the subset D0 of the unit disk D. Then the following are equivalent:
(1) There is a function S in the restricted Schur class SN
bounded and measurable in (t, w) for each fixed z ∈ D0, together with a
(2) There is a matrix-valued function ((t, w), z) 7→ Ht,w(z) on bΘN × D0,
finite measure ν on bΘN , so that
I−S0(z)S0(w)∗ =Z bΘN
Ht,w(z)(cid:0)IXt,w ⊗ (I − St,w(z)St,w(w)∗)(cid:1) Ht,w(w)∗ dν(t, w).
(3) For each 2N × 1 isometric matrix [ α
β ] and for each Y : D0 → CN ×N , the
1 with SD0 = S0.
kernel
k(z, w) = tr(cid:0)Y (w)∗(I − S0(w)∗S0(z))Y (z))K α,β(z, w)(cid:1)
(where K α,β is given by (4.25)) is a positive kernel on D0.
(4.26)
Proof. The proof parallels that of Theorem 4.2. To verify the equivalence of con-
dition (2) with existence of a solution of the interpolation problem, use Theorem
4.7 in combination with Theorem 3.1. By Remark 3.7, the validity of condition (3)
follows if we can verify that the collection (ΨN
1 )0 given by (4.25) is a generating
(CN ). However, rather than doing
set for the collection of admissible kernels KΨN
this we use Theorem 1.3 from [17]. As was the case for the Schur class over a
domain R, the form presented there is somewhat different from the form (4.26) as
presented here. However, one can follow the argument in [17] and work with right
multiplication operators on H(K α,β)CN rather than left multiplication operators on
a left-sided tensor of the coefficient space with a reproducing kernel Hilbert space
of row-vector functions to arrive at the form (4.26) as the solution criterion.
(cid:3)
1
Remark 4.9. As was observed in connection with Corollary 4.5, the Schur-Agler
class SAΨ associated with a collection of test functions Ψ depends on the functions
ψ ∈ Ψ only through the kernels I − ψ(z)ψ(w)∗. Hence, for St,w in the test-function
class ΨN
1 we may define an equivalence relation St,w ∼ St′,w′ when there is a unitary
constant matrix U so that St′,w′(z) = St,w(z)U . To choose one representative out
of each equivalence class, we may normalize S ∈ ΨN
1 so that S(1) = IN . This
one of the points in the set of points t = (1, t2, . . . , tn) with associated weight W1
invertible; in this way we get a new smaller parameter space ΘN . Then we have
1 )N ×N = SA eΨN
B(H ∞
1 = {St,w : (t, w) ∈ ΘN} is this restricted class of
test functions.
For the case N = 1 (the scalar case), Theorem 4.7 is due to Dritschel-Pickering
has the effect of restricting the parameter (t, w) in bΘN to those such that 1 is
where eΨN
1
[30]. In this case the parameter space bΘ1 =: bΘ can be described in geometric terms
as consisting of (1) triples of points on the unit circle such that 0 is in the interior
of the associated triangle, with the weights then being the barycentric coordinates
36
J.A. BALL AND M.D. GUERRA-HUAM ´AN
2 , 1
of 0 with respect to this triangle, or (2) a pair of antipodal points on the unit circle
with weights then necessarily ( 1
2 ). When the reduction described in the previous
paragraph is carried out, one restricts to triples of points t = (1, t2, t3) which include
1 and there is only one antipodal pair of points (1,−1). These authors also show
that this space Θ with its natural topology is homeomorphic to the unit sphere.
1 is a minimal collection of test functions for
1 )N ×N in
1 is a minimal collection of test functions for B(H ∞
They also show that the collection eΨ1
1 . Whether eΨN
BH ∞
general we leave as an open question.
As we have seen, there is a dual issue of finding minimal generating sets for
admissible collections of kernels KΨ(CN ), as well as finding small generating sets for
R)0
such KΨ(CN ). In particular, it would be interesting to see a direct proof that (ΨN
1 )0 in (4.25) generates KΘN (CN ). We
in (4.14) generates KΨN
note that the proofs of the interpolation results from [1, 14, 26, 17] use the dual
factorization approach (see [25] for a unified setting); an independent proof of the
generating property for (ΨN
1 )0 would mean that Theorem 3.1 gives an
independent proof of these interpolation results.
and that the set (ΨN
R)0 and (ΨN
R
is C + z2H ∞. Many of the
Remark 4.10. An alternative description of H ∞
1
results concerning the space H ∞
1 have been generalized to more general algebras of
the form C + BH ∞ where B is a Blaschke product (see e.g. [50]). We believe that
the results from [20] are sufficiently flexible to lead to test-function Schur-Agler-
class characterizations of matrix-valued versions of these more general algebras as
well.
References
[1] M.B. Abrahamse, The Pick interpolation theorem for finite connected domains, Michigan
Math. J. 26 (1979), 195 -- 203.
[2] J. Agler, Interpolation, unpublished manuscript, circa 1988.
[3] J. Agler, On the representation of certain holomorphic functions defined on a polydisk, in:
Topics in Operator Theory: Ernst D. Hellinger Memorial Volume (ed. L. de Branges I. Go-
hberg, J. Rovnyak), OT 48, Birkhauser-Verlag, Basel, 1990.
[4] J. Agler, J. Harland, B.J. Raphael, Classical Function Theory, Operator Dilation Theory,
and Machine Computation on Multiply-Connected Domains, Memoirs of the American Math-
ematical Society, Number 92, January 2008.
[5] J. Agler and J.E. McCarthy, Nevanlinna-Pick interpolation on the bidisk, J. Reine
Angew. Math. 506 (1999), 191 -- 204.
[6] J. Agler and J.E. McCarthy, Pick Interpolation and Hilbert Function Spaces, Graduate Stud-
ies in Mathematics Vol. 44, American Mathematical Society, Providence, 2002.
[7] J. Agler and J.E. McCarthy, Distinguished varieties, Acta Math. 194 (2005), 133 -- 153.
[8] C.-G. Ambrozie, Remarks on the operator-valued interpolation for multivariable bounded
analytic functions, Indiana University Math. J. 53 (2004), 1551 -- 1576.
[9] C. Ambrozie and J. Eschmeier, A commutant lifting theorem on analytic polyhedra, in: Topo-
logical Algebras, Their Applications, and Related Topics, Banach Center Publications 67,
Institute of Mathematics, Polish Academy of Sciences, Warsaw, 2005.
[10] C.-G. Ambrozie and D. Timotin, A von Neumann type inequality for certain domains in Cn,
Proc. Amer. Math. Soc. 131 (2003), 859 -- 869.
[11] A.V. Arkhangel'skiı and V.V. Fedorchuk, I. The Basic Concepts and Constructions of Gen-
eral Topology, in General Topology I (ed. A.V. Arkhangel'skiı and L.S. Pontryagin), pp. 1 -- 90,
Encyclopaedia of Mathematical Sciences 17, Springer, New York, 1990.
[12] N. Aronszajn, Theory of reproducing kernels, Trans. Amer. Math. Soc. 68 (1950), 337 -- 404.
[13] W. Arveson, An Invitation to C ∗-Algebras, Graduate Texts in Mathematics 39, Springer-
Verlag, New York, 1976.
TEST FUNCTIONS AND TRANSFER-FUNCTION REALIZATION
37
[14] J.A. Ball, A lifting theorem for operator models of finite rank on multiply-connected domains,
J. Operator Theory 1 (1979), 3 -- 25.
[15] J.A. Ball, A. Biswas, Q. Fang, and S. ter Horst, Multivariable generalizations of the Schur
class: positive kernel characterization and transfer function realization, in: Recent Advances
in Operator Theory and Applications, pp. 17 -- 79, OT 187 Birkhauser-Verlag, Basel, 2008.
[16] J.A. Ball and V. Bolotnikov, Realization and interpolation for Schur-Agler-class functions on
domains with matrix polynomial defining function in Cd, J. Functional Analysis 213 (2004),
45-87.
[17] J.A. Ball, V. Bolotnikov, and S. ter Horst, A constrained Nevanlinna-Pick interpolation
problem for matrix-valued functions, Indiana University Math. J. 59 (2010) No. 1, 15 -- 51.
[18] J.A. Ball and K.F. Clancey, Reproducing kernels for Hardy spaces on multiply connected
domains, Integral Equations and Operator Theory 25 (1996), 35 -- 57.
[19] J.A. Ball, G. Groenewald, and T. Malakorn, Conservative structured noncommutative multi-
dimensional linear systems, in: The State Space Method: Generalizations and Applications
(ed. D. Alpay and I. Gohberg), pp. 179 -- 223, OT 161 Birkhauser-Verlag, Basel, 2005.
[20] J.A. Ball and M. Guerra Huaman, Convexity analysis and matrix-valued schur class over
finitely connected planar domains, preprint, 2011.
[21] J.A. Ball, W.S. Li, D. Timotin, and T.T. Trent, A commutant lifting theorem on the polydisc,
Indiana Univ. Math. J. 48 (1999), 653 -- 675.
[22] J.A. Ball and T.T. Trent, Unitary colligations, reproducing kernel Hilbert spaces and
Nevanlinna-Pick interpolation in several variables, J. Funct. Anal.157 (1998) no. 1, 1998.
[23] J.A. Ball, T.T. Trent, and V. Vinnikov, Interpolation and commutant lifting for multipliers
on reproducing kernel Hilbert spaces, in: The M.A. Kaashoek Anniversary Volume (Workshop
in Amsterdam, November 1997) (ed. H. Bart, I. Gohberg, A.C.M. Ran), pp. 89 -- 138, OT 122
Birkhauser-Verlag, Basel, 2001.
[24] S.D. Barreto, B.V.R. Bhat, V. Liebscher, and M. Skeide, Type I product systems of Hilbert
modules, J. Func. Anal. 212 (2004), 121 -- 181.
[25] K.R. Davidson and R. Hamilton, Nevanlinna-Pick interpolation and factorization of linear
functionals, Integral Equations and Operator Theory 70 (2011) no. 1, 125 -- 149.
[26] K.R. Davidson, V. Paulsen, M. Raghupathi, and D. Singh, A constrained Nevanlinna-Pick
interpolation problem, Indiana University Math. J. 58 (2009), 709 -- 732.
[27] M.A. Dritschel, S.A.M. Marcantognini and S. McCullough, Interpolation in semigroupoid
algebras, J. reine angew. Math. 606 (2007), 1 -- 40.
[28] M.A. Dritschel and S. McCullough, The failure of rational dilation on a triply connected
domain. J. Amer. Math Soc. 18 (2005), 873 -- 918.
[29] M.A. Dritschel and S. McCullough, Test functions, kernels, realizations and interpolation,
in: Operator Theory, Structured Matrices, and Dilations. Tiberiu Constantinescu Memorial
Volume (ed. M. Bakonyi, A. Gheondea, M. Putinar and J. Rovnyak), pp. 153 -- 179, Theta
Foundation, Bucharest, 2007.
[30] M.A. Dritschel
and J. Pickering, Test
functions
in constrained
interpolation,
Trans. Amer. Math. Soc., to appear.
[31] J. Dugundji, Topology, Allyn and Bacon, Boston, 1966.
[32] J. Eschmeier and M. Putinar, Spherical contractions and interpolation problems on the unit
ball, J. Reine Angew. Math. 542 (2002), 219 -- 236.
[33] S.D. Fisher, Function Theory on Planar Domains: A Second Course in Complex Analysis,
Wiley & Sons, New York, 1983; Second Edition: Dover Publications, New York, 2007.
[34] I. Gohberg (ed.), I. Schur Methods in Operator Theory and Signal Processing, OT 18,
Birkhauser, Basel, 1986.
[35] M.D. Guerra-Huaman, Schur class of finitely connected planar domains: the test-function
approach, Virginia Tech dissertation, 2011.
[36] M.T. Jury, Universal commutative operator algebras and transfer function realizations of
polynomials, arXiv:1009.6219v1 [math.FA], 30 Sep 2010.
[37] M.T. Jury, G. Knese and S. McCullough, Agler interpolation families of kernels, Oper. Ma-
trices 3(2009) no. 4, 571 -- 587.
[38] S. McCullough, Matrix functions of positive real part on an annulus, Houston J. Math. 21
(1995) no. 3, 489 -- 506.
[39] S. McCullough, Isometric representations of some quotients of H∞ of an annulus, Integral
Equations and Operator Theory 39 (2001), 335 -- 362.
38
J.A. BALL AND M.D. GUERRA-HUAM ´AN
[40] S. McCullough and V. Paulsen, C ∗-envelopes and interpolation theory,
Indiana
Univ. Math. J. 51 (2002), 479 -- 505.
[41] S. McCullough and S. Sultanic, Ersatz commutant lifting with test functions, Complex Anal-
ysis and Operator Theory 1 (2007), 581 -- 620.
[42] M. Mittal and V.I. Paulsen, Operator algebras of functions, J. Funct. Anal. 258 (2010) no. 9,
3195 -- 3225.
[43] P.S. Muhly and B. Solel, Tensor algebras over C ∗-correspondences: representations, dila-
tions, and C ∗-envelopes, J. Funct. Anal. 158 (1998), 389 -- 457.
[44] P.S. Muhly and B. Solel, Schur class operator functions and automorphisms of Hardy alge-
bras, Documenta Mathematica 13 (2008), 365 -- 411.
[45] V.I. Paulsen, Matrix-valued interpolation and hyperconvex sets, INtegral Equations and Op-
erator Theory 41 (2001), 38 -- 62.
[46] V.I. Paulsen, Operator algebras of idempotents, J. Func. Anal. 181 (2001), 209 -- 226.
[47] V.I. Paulsen and J.P. Solazzo, Interpolation and balls in C2, J. Operator Theory 60 (2008)
no. 2, 379 -- 398.
[48] V. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge Studies in Ad-
vanced Mathematics 78, Cambridge University Press, Cambridge, 2002.
[49] I. Raeburn and D.P. Williams, Morita Equivalence and Continuous-Trace C ∗-Algebras,
Mathematical Surveys and Monographs 60, American Mathematical Society, Providence,
1998.
[50] M. Raghupathi, Nevanlinna-Pick interpolation for C + BH∞, Integral Equations and Oper-
ator Theory 63 (2009), 103 -- 125.
[51] W. Rudin, Functional Analysis (Second Edition), McGraw-Hill, 1991.
[52] M. Takesaki, Theory of Operator Algebras I, Encyclopaedia of Mathematical Sciences 124,
Springer, New York, 1979.
[53] A.T. Tomerlin, Products of Nevanlinna-Pick kernel and operator colligations, Integral Equa-
tions and Operator Theory 38 (2000), 350 -- 356.
[54] V. Vinnikov and S.I. Fedorov, The Nevanlinna-Pick interpolation problem in multiply con-
nected domains, J. Math. Sciences 105 (2001) no. 4, 2109 -- 2126.
Department of Mathematics, Virginia Tech, Blacksburg, VA 24061-0123, USA
E-mail address: [email protected]
Department of Mathematics, Virginia Tech, Blacksburg, VA 24061-0123, USA
E-mail address: [email protected]
|
1710.10065 | 1 | 1710 | 2017-10-27T10:41:15 | Further results on the $(b, c)$-inverse, the outer inverse $A^{(2)}_{T, S}$ and the Moore-Penrose inverse in the Banach context | [
"math.FA"
] | In this article properties of the $(b, c)$-inverse, the inverse along an element, the outer inverse with prescribed range and null space $A^{(2)}_{T, S}$ and the Moore-Penrose inverse will be studied in the contexts of Banach spaces operators, Banach algebras and $C^*$-algebras. The main properties to be considered are the continuity, the differentiability and the openness of the sets of all invertible elements defined by all the aforementioned outer inverses but the Moore-Penrose inverse. The relationship between the $(b, c)$-inverse and the outer inverse $A^{(2)}_{T, S}$ will be also characterized. | math.FA | math |
Further results on the (b, c)-inverse, the outer inverse A(2)
T,S and
the Moore-Penrose inverse in the Banach context
Enrico Boasso
Abstract
In this article properties of the (b, c)-inverse, the inverse along an element, the outer inverse
with prescribed range and null space A(2)
S and the Moore-Penrose inverse will be studied
in the contexts of Banach spaces operators, Banach algebras and C ∗-algebras. The main
properties to be considered are the continuity, the differentiability and the openness of
the sets of all invertible elements defined by all the aforementioned outer inverses but the
Moore-Penrose inverse. The relationship between the (b, c)-inverse and the outer inverse
A(2)
S will be also characterized.
T
T
,
,
Keywords: (b, c)-inverse; outer inverse A(2)
C ∗-algebra; Banach space operator
T
,
S ; Moore-Penrose inverse; Banach algebra;
AMS Subjects Classifications: 46H05; 46L05; 47A05; 15A09
1
Introduction
Recently two outer inverses have been introduced: the (b, c)-inverse and the inverse along
an element, see [8] and [19], respectively. These two inverses are related; in fact, the latter
is a particular case of the former. It is worth noticing one of the main properties of these
inverses, namely, they encompass several well known outer inverses such as the Drazin inverse,
the group inverse and the Moore-Penrose inverse. Furthermore, several authors have studied
these notions, see for the (b, c)-inverse [4, 6, 8, 9, 14, 15, 25, 30] and for the invese along an
element [1, 2, 19 -- 22, 32, 33]. In particular, in [4] and [2] several properties of the (b, c)-inverse
and the inverse along an element were studied in the Banach context, respectively.
On the other hand, one of the most well known outer inverses is the outer inverse with
prescribed range and null space, i.e., the outer invers A(2)
T,S . This inverse has been studied in
the frames of matrices, Hilbert space operators and Banach space operators. To learn about
the A(2)
T,S outer inverse in Banach spaces, see for example [10, 18, 31].
The main objective of this article is to deepen the knowledge of the three aformentioned
outer inverses in the contexts of Banach algebras, C ∗-algebras, Banach space operators and
Hilbert space operators. However, as an application of the main results, properties of the
Moore-Penrose inverse will be also presented.
The article is organized as follows. In section 3, after having recalled some preliminary
definitions and facts in section 2, the relationship between the (b, c)-inverse (in particular
the inverse along an element) and the outer inverse A(2)
T,S will be studied. In section 4 both
the set of all (b, c)-invertible elements (in particular the set of all invertible elements along
a fixed element) and the set of all operator for which the outer inverse A(2)
T,S exists will be
1
proved to be open. The continuity of the (b, c)-inverse of Banach space operators and of
Banach algebra and C ∗-algebra elements will be characterized in section 5; two main notions
will be used to accomplish this aim: the gap between two subspaces and the Moore-Penrose
inverse. The diffrentiability of the (b, c)-inverse in Banach algebras and C ∗-algebras will be
studied in section 6; the Moore-Penrose inverse will be also applied in this section. Finally,
the continuity and differentiability of the outer inverse A(2)
T,S will be characterized in section
7 using again the gap between subspaces and the Moore-Penrose inverse.
In addition, in
section 5 and 6, as an application of the main results of these sections, the continuity and the
differentiability of the Moore-Penrose inverse for Banach algebra elements and Banach space
operators will be studied, respectively.
2 Preliminary definitions
From now on, A will denote a unitary Banach algebra with unit 1 while A−1 and A• will stand
for the set of invertible elements and the set of idempotents of A, respectively. A particular
case is L(X), the Banach algebra of all linear and bounded maps defined on and with values
in the Banach space X. However, in the present work it will be necessary to consider the
Banach space of all operators defined on the Banach space X with values in the Banach space
Y, which will be denoted by L(X, Y). Note that if T ∈ L(X, Y), then N(T ) ⊆ X and R(T ) ⊆ Y
will stand for the null space and the range of the operator T , respectively. For example, when
A is a unitary Banach algebra and x ∈ A, the operators Lx : A → A and Rx : A → A are the
maps defined as follows: given z ∈ A, Lx(z) = xz and Rx(z) = zx. Observe that since A is
unitary, then k La k=k a k=k Ra k. Moreover, the following notation will be used:
x−1(0) = N(Lx),
xA = R(Lx),
x−1(0) = N(Rx),
Ax = R(Rx).
Note that when no confusion is possible, the identity operator defined on the Banach space X
will be denoted by I ∈ L(X); otherwise it will be denoted by IX. In addition, given a Hilbert
M ∈ L(H)• wil stand for the orthogonal projector
space H and a closed subspace M ⊆ H, P ⊥
with range M.
An element a ∈ A will be said to be regular, if there exists x ∈ A such that a = axa. The
element x, which is not uniquely determined by a, will be said to be a generalized inverse or
an inner inverse of a. In addition, A will stand for the set of all regular elements of A and
given a ∈ A, a{1} will denote the set of all generalized inverses of a. On the other hand, if
y ∈ A satisfies yay = y, then y will be said to be an outer inverse of a. Moreover, an element
z will be said to be a normalized generalized inverse of a, if z is both an inner and an outer
inverse of a. Recall that if b is an inner inverse of a, then bab is a normalized generalized
inverse of a.
Now the definition of one of the key notion of this article will be recalled. Note that this
notion was originally introduced in the context of semigroup, however, since the frame of this
article are Banach algebras and Banach space operators, the notion under consideration, as
well as all the object considered in this work, will be introduced and studied in the Banach
context.
Definition 2.1 ([8, Definition 1.3]). Let A be a unitary Banach algebra and consider b, c ∈ A.
The element a ∈ A will be said to be (b, c)-invertible, if there exists y ∈ A such that the
following equations hold:
2
(i) y ∈ (bAy) ∩ (yAc),
(ii) b = yab, c = cay.
In the same conditions of Definition 2.1, if such an inverse exists, then it is unique ([8,
Theorem 2.1 (i)]). Thus in what follows, if the element y in Definition 2.1 exists, then it will
be denoted by a−(b, c). In addition, a−(b, c) is an outer inverse of a ([8, Theorem 2.1 (ii)]) and
b and c are regular ([4, Remark 2.2 (iii)] or [30, Proposition 3.3]).
A particular case of the (b, c)-inverse is the Bott-Duffin (p, q)-inverse.
Definition 2.2 ([8, Definition 3.2]). Let A be a unitary Banaxh algebra and consider p,
q ∈ A•. The element a ∈ A will be said to be Bott-Duffin (p, q)-invertible, if there exists
y ∈ A such that
(i) y = py = yq,
(ii) yap = p and qay = q.
Clearly, given p, q ∈ A•, the Bott-Duffin (p, q)-inverse is nothing but the (b, c)-inverse
when b and c are idempotents.
In addition, since there exists at most one (b, c)-inverse,
the Bott-Duffin (p, q)-inverse is unique, if it exists. According to what has been said, if
a ∈ A is Bott-Duffin (p, q)-invertible, then the element y in Definition 2.2 will be denoted
by a−(p, q). To learn more on the outer inverses recalled in Definition 2.1 and Definition 2.2,
see [4, 6, 8, 9, 14, 15, 25, 30].
Next the definition of the inverse along an element will be recalled. This is another
particular case of the (b, c)-inverse.
Definition 2.3 ([19, Definition 4]). Consider a, d ∈ A. The element a is invertible along d,
if there exists b ∈ A such that b is an outer inverse of a, bA = dA and Ab = Ad.
Recall that, in the same conditions of Definition 2.3, according to [19, Theorem 6], if
such b ∈ A exists, then it is unique. Therefore, the element b satisfying Definition 2.3 will
be said to be the inverse of a along d. In this case, the inverse under consideration will be
denoted by a−d. Moreover, according to [8, Proposition 6.1], the inverse along an element is
a particular case of the (b, c)-inverse. In fact, the element a is invertible along d if and only
if it is (d, d)-invertible. Furthermore, in this case, a−d = a−(d,d). To learn more on this outer
inverse, see [1, 2, 19 -- 22, 32, 33].
One of the most studied generalized inverses is the outer inverse A(2)
T,S , i.e., the outer
inverse of the operator A ∈ L(X, Y) with prescribed range T ⊆ X and null space S ⊆ Y ,
where X and Y are Banach spaces. This generalized inverse was studied for matrices and for
operators defined on Hilbert and on Banach spaces. Recall that this inverse is unique, when
it exists ([18, Lemma 1]).
Definition 2.4. Let X and Y be Banach spaces, A ∈ L(X, Y) and T and S two closed subspaces
of X and Y, respectively. If there exists a (necessarily unique) operator B ∈ L(Y, X) such that
B is an outer inverse of A, R(B) = T and N(B) = S, then B will be said to be the A(2)
T,S outer
inverse of A.
According to [18, Lemma 1], a necessary and sufficient condition for the existence of A(2)
T,S
: T → A(T)
is that T and S are complemented subspaces of X and Y, respectively, A A(T)
T
3
is invertible and A(T) ⊕ S = Y. In particular, using this latter decomposition, A(2)
following operator:
T,S is the
A(2)
T,S S= 0,
(A A(T)
T
)−1 = A(2)
T, S T
A(T) : A(T) → T.
To learn more properties of the A(2)
T,S outer inverse in Banach spaces, see [10, 18, 31].
To characterize the continuity of the outer inverses considered in this article, it is necessary
to recall the definition of the Moore-Penrose inverse. Let A be a C ∗-algebra. An element
a ∈ A will be said to be Moore-Penrose invertible, if there exists b ∈ A such that the following
equations hold:
a = aba,
b = bab,
(ab)∗ = ab,
(ba)∗ = ba.
To give the notion of a Moore-Penrose invertible Banach algebra element, the definition of
an hermitian element need to be recalled first. Given a Banach algebra A, an element z ∈ A is
said to be hermitian, if k exp(ita) k= 1, for all t ∈ R (see [29], [7, Chapter 4] and [5, Chapter
I, Section 10]). When A is a C ∗-algebra, a ∈ A is hermitian if and only if it is self-adjoint
([5, Proposition 20, Chapter I, Section 12]). Now Moore-Penrose Banach algebra elements
can be defined.
Given a Banach algebra A, an element x ∈ A will be said to be Moore-Penrose invertible, if
there exists b ∈ A such that b is a normalized generalized inverse of a and the elements ab and
ba are hermitian. Since the Moore-Penrose inverse is unique, when it exists (see [26, Lemma
2.1]), the element b will be denoted by a†. Moreover, A† will stand for the set of all Moore-
Penrose invertible elements of A. Note that A† ⊆ A. Furthermore, when A is a C ∗-algebra,
A† = A ([11, Theorem 6]); however, when the Banach algebra A is not a C ∗-algebra, in
general, this result does not hold ([3, Remark 4]). To learn the definition of the Moore-
Penrose inverse for matrices, see [24]; to learn more properties of this generalized inverse see,
for C ∗ algebras, [11, 12, 16], and for Banach algebras, [3, 26, 27].
To prove some of the main results of this article, the definition of the gap between two
subspaces need to be recalled. Let X be a Banach space and consider M and N two closed
subspaces in X. If M = 0, then set δ(M, N) = 0, otherwise set
δ(M, N) = sup{dist(x, N) : x ∈ M, k x k= 1},
where dist(x, N) = inf{k x − y k : y ∈ N}. The gap between the subspaces M and N is
bδ (M, N) = max{δ(M, N), δ(N, M)}.
To learn more on this notion, see [10, 13].
3 The relationship between the (b, c)-inverse and A(2)
T,S
The main objective of this section is to prove that given a Banach space X, the (B, C)-inverse
in L(X) consists in a reformulation of the outer inverse A(2)
T,S , in other words, the outer
inverse introduced in [8, Definition 1.3] is an extension to semigroups of the outer inverse
with prescribed range and null space.
4
First an equivalent formulation of Definition 2.1 will be given in the context of Banach
space operators. To this end, however, some preparation is needed. In the following remark
a particular operator will be introduced.
Remark 3.1. Recall that given a Banach space X, a necessary and sufficient condition for
F ∈ L(X) to be a regular operator is that N(F ) and R(F ) are closed and complemented
susbpsaces of X. Suppose that F ∈ L(X) satisfies this condition and let N and M be two
subspaces of X such that
N(F ) ⊕ N = X = R(F ) ⊕ M.
Consider the isomorphism F1 = F R(F )
follows:
N
: N → R(F ) and define the map SF ∈ L(X) as
SF M = 0,
SF R(F )= ιN,XF −1
1
,
where ιN,X : N → X is the inclusion map. The operator SF ∈ L(X) will be used in the proofs
of the next results.
Lemma 3.2. Let X be a Banach space and consider F ∈ L(X) a regular operator and SF ∈
L(X) the map defined in Remark 3.1. The following statements hold.
(i) If X ∈ L(X) is such that R(X) = R(F ), then X = F SF X.
(ii) If X ∈ L(X) is such that N(X) = N(F ), then X = XSF F .
Proof. Note that F SF R(F )= ιR(F ),X, where ιR(F ),X : R(F ) → X is the inclusion map.
R(X) = R(F ), then X = F SF X.
If
Similarly, SF F N= ιN,X, where N is the subspace of X considered in Remak 3.1.
If
N(X) = N(F ), then since N(F ) ⊕ N = X, X = XSF F .
The following proposition is a key step in the reformulation of Definition 2.1 for Banach
space operators.
Proposition 3.3. Let X be a Banach space and consider two regular operators F , G ∈ L(X).
(i) A necessary and sufficient condition for F L(X) = G L(X) is that R(F ) = R(G).
(ii) L(X)F = L(X)G if and only if N(F ) = N(G).
Proof. If there exist U , V ∈ L(X) such that F = GU and G = F V , then R(F ) = R(G).
On the other hand, if R(F ) = R(G), then according to Lemma 3.2 (i), G = F SF G and
F = GSGF , which implies that F L(X) = G L(X).
A similar argument, using in particular Lemma 3.2 (ii), proves the second statement.
Theorem 3.4. Let X be a Banach space and consider two regular operators B, C ∈ L(X).
The following statements are equivalent.
(i) The operator A ∈ L(X) is (B, C)-invertible.
(ii) There exists a bounded and linear map X ∈ L(X) such that
B = XAB, C = CAX, R(X) = B, N(X) = N(C).
5
Moreover, in this case X = A−(B,C).
Proof. According to Definition 2.1, if the (B, C)-inverse of A exists, then there are X, T1 and
T2 ∈ L(X) such that
B = XAB, C = CAX, X = BT1, X = T2C.
In particular, R(B) = R(X) and N(X) = N(C).
On the other hand, suppose that statement (ii) holds. Since B (respectively C) is regular
and R(X) = R(B) (respectively N(X) = N(C)), according to Lemma 3.2, X = BSBX
(respectively X = XSC C), where SB (respectively SC) is the operator considered in Remark
3.1. Therefore, X ∈ B L(X)X ∩ X L(X)C. Since the (B, C)-inverse is unique, when it exists,
X = A−(B,C).
Next the main result of this section will be proved
Theorem 3.5. Let X be a Banach space and consider two regular operators B, C ∈ L(X).
The following statements are equivalent.
(i) The operator A ∈ L(X) is (B, C)-invertible.
(ii) There exists X ∈ L(X) such that X = XAX, R(X) = R(B) and N(X) = N(C).
(iii) The outer inverse A(2)
R(B),N(C) exists.
(iv) A R(AB)
R(B)
: R(B) → R(AB) is invertible and R(AB) ⊕ N(C) = X.
Furthermore, in this case A−(B,C) = X = A(2)
R(B),N (C).
Proof. According to [8, Proposition 6.1], statement (i) is equivalent to the fact that there
exists an operator X ∈ L(X) such that X is an outer inverse of A, X L(X) = B L(X) and
L(X)X = L(X)C. However, according to the proof of Theorem 3.4, these conditions are
equivalent to statement (ii). In addition, since the (B, C)-inverse is unique, when it exists,
X = A−(B,C).
Statement (ii) and (iii) are equivalent; moreover X = A(2)
To prove the equivalence between statements (iii) and (iv), apply [18, Lemma 1] and recall
that, since B and C are regular, R(B) and N(C) are closed and complemented subspaces of
X.
R(B), N(C) ([18, Lemma 1]).
Note that Theorem 3.5 was proved for square matrices in [25, Theorem 1.5]. The following
results will be derived from what has been proved.
Corollary 3.6. Let X be a Banach space and consider A, B, C ∈ L(X) such that B and C
are regular. The following statements are equivalent.
(i) A−(B,C) exists.
(ii) A−(F,G) exists for any F , G ∈ L(X), F ,G regular, such that R(F ) = R(B) and N(G) =
N(C).
(iii) The Bott-Duffin inverse A−(P,Q) exists for any P , Q ∈ L(X)• such that R(P ) = R(B)
and N(Q) = N(C).
6
Furthermore, in this case, A−(B,C) = A−(F,G) = A−(P,Q).
Proof. Apply Theorem 3.5.
Note that Corollary 3.6 can be rephrased using the outer inverse A(2)
T,S.
Remark 3.7. Let X be a Banach space and consider S and T two closed and complemented
subspaces of X. Let A ∈ L(X). The following statement are equivalent.
(i) The outer inverse A(2)
T,S exists.
(ii) A−(B,C) exists for any B, C ∈ L(X), B, C regular, such that R(B) = T and N(C) = S.
(iii) The Bott-Duffin inverse A−(P,Q) exists for any P , Q ∈ L(X)• such that R(P ) = T and
N(Q) = S.
Moreover, in this case A(2)
The proof of the equivalence among statements (i)-(iii) and the last identity can be derived
from Theorem 3.5 and Corollary 3.6.
T,S = A−(B,C) = A−(P,Q).
It is worth noticing, as it has been mentioned in the first paragraph of this section, that
Theorem 3.5 and Remark 3.7 show that in L(X), X a Banach space, the (B, C)-inverse is a
reformulation of A(2)
T,S , so that [8, Definition 1.3] consists in an extension of the latter outer
inverse to semigroups. In fact, given A ∈ L(X), A(2)
T,S and the (B, C)-inverse of A ∈ L(X)
refer to the same object -under the conditions of Theorem 3.5 and Remark 3.7, but it could
be said that the former outer inverse is defined from a spatial point of view while the latter
from an algebraic point of view. Actually, in the first case subspaces of an underlying space
(a Banach space X or Cn, n ∈ N) are used in the definition, but in the second only elements
of a semigroup can be considered; precisely, in L(X) it is possible to associate two specific
subspaces to any element of this algebra -the range and the null space, which leads to the
aforementioned results.
Next the inverse along an operator will be considered.
Corollary 3.8. Let X be a Banach space and consider A, D ∈ L(X) such that D is regular.
The following statements are equivalent.
(i) A−D exists.
(ii) There exists X ∈ L(X) such that XAD = D = DAX, R(X) = R(D) and N(X) = N(D).
(iii) There exists Y ∈ L(X) such that Y is an outer inverse of A, R(Y ) = R(D) and N(Y ) =
N(D).
(iv) The outer inverse A(2)
R(D),N(D) exists.
(v) A R(AD)
R(D)
: R(D) → R(AD) is invertible and R(AD) ⊕ N(D) = X.
(vi) A−F exists for all F ∈ L(X) such that F is regular, R(F ) = R(D) and N(F ) = N(D).
Moreover, in this case, A−D = X = Y = A(2)
R(D),N(D) = A−F .
7
Proof. Recall that according to [8, Proposition 6.1], A−D = A−(D,D), when one of these outer
inverses exists. Apply now Theorem 3.4, Theorem 3.5 and Corollary 3.6.
In the following proposition the Banach algebra case will be considered.
Proposition 3.9. Let A be a unitary Banach algebra and consider a, b, c ∈ A such that b
and c are regular. The following statements hold.
(i) If a−(b,c) exists, then La−(b,c) = L−(Lb,Lc)
a
= (La)(2)
bA,c−1(0) ∈ L(A).
(ii) Suppose that L−(Lb,Lc)
a
= (La)(2)
bA,c−1(0) ∈ L(A) exists and there is z ∈ A such that
(La)(2)
bA,c−1(0) = Lz. Then, a−(b,c) exists and a−(b,c) = z.
(iii) If a−(b,c) exists, then Ra−(b,c) = R−(Rc,Rb)
a
= (Ra)(2)
Ac,b−1(0) ∈ L(A).
(iv) Suppose that R−(Rc,Rb)
a
= (Ra)(2)
Ac,b−1(0) ∈ L(A) exists and there is w ∈ A such that
(Ra)(2)
Ac,b−1(0) = Rw. Then, a−(b,c) exists and a−(b,c) = w.
Proof. Recall that according to [8, Proposition 6.1], necessary and sufficient for a−(b,c) to exist
is that a has an outer inverse, say y (y = a−(b,c)), such that yA = bA and Ay = Ac. Thus, if
a−(b,c) exists, then La−(b,c) ∈ L(A) is an outer inverse of La ∈ L(A) such that
R(La−(b,c)) = a−(b,c)A = bA = R(Lb),
N(La−(b,c) ) = (a−(b,c))−1(0).
However, note that (a−(b,c))−1(0) = c−1(0) = N(Lc). In fact, since Aa−(b,c) = Ac, there exist
u1, v1 ∈ A such that a−(b,c) = u1c and c = v1a−(b,c), which implies that (a−(b,c))−1(0) =
c−1(0). Therefore, according to [18, Lemma 1] and Theorem 3.5,
La−(b,c) = (La)(2)
bA,c−1(0) = L−(Lb,Lc)
a
.
Now suppose that statement (ii) holds. Note that according to Theorem 3.5, L−(Lb,Lc)
∈
bA,c−1(0) ∈ L(A) exists; moreover, in this case, both maps
a
L(A) exists if and only if (La)(2)
coincide.
Since (La)(2)
bA,c−1(0) ∈ L(A) is an outer inverse of La ∈ L(A), z is an outer inverse of a. In
addition,
zA = R(Lz) = R((La)(2)
z−1(0) = N(Lz) = N((La)(2)
bA,c−1(0)) = bA,
bA,c−1(0)) = c−1(0).
However, since c and z are regular, according to [1, Proposition 3.1 (b) (iii)], Az = Ac.
Consequently, according to [8, Proposition 6.1], a−(b,c) exists and a−(b,c) = z.
8
As in the proof of statement (i), if a−(b,c) exists, then Ra−(b,c)
is an outer inverse of
Ra ∈ L(A) such that
R(Ra−(b,c)) = Aa−(b,c) = Ac = R(Rc),
N(Ra−(b,c)) = (a−(b,c))−1(0).
However, since bA = a−(b,c)A, there exist u2, v2 ∈ A such that b = a−(b,c)u2 and a−(b,c) = bv2,
which implies that (a−(b,c))−1(0) = b−1(0) = N(Rb). Therefore, according to [18, Lemma 1]
and Theorem 3.5, statement (iii) holds.
If statement (iv) holds, then as in the proof of statement (ii), observe that according to
Ac,b−1(0) exists; moreover, in this case,
Ac,b−1(0) ∈ L(A) is an outer inverse of Ra ∈ L(A),
Theorem 3.5, R−(Rc,Rb)
both maps coincide. Furthermore, since (Ra)(2)
w is an outer inverse of a. In addition,
∈ L(A) exists if and only if (Ra)(2)
a
Aw = R(Rw) = R((Ra)(2)
w−1(0) = N(Rw) = N((Ra)(2)
Ac,b1−1(0)) = Ac,
Ac,b−1(0)) = b−1(0).
However, since b and w are regular, according to [1, Proposition 3.1 (b) (iv)], wA = bA.
Hence, according to [8, Proposition 6.1], a−(b,c) exists and a−(b,c) = w.
4 Openness
Given an unitary Banach algebra A, A−1 ⊂ A is an open set. In this section the set of all
(b, c)-invertible elements (b, c ∈ A) will be proved to be open. Similar results will be presented
for the inverse along an element and the outer inverse A(2)
T,S .
Let A be a unitary Banach algebra and consider b, c ∈ A. Let A−(b,c) be the set of all
(b, c)-invertible elements of A, i.e.,
A−(b,c) = {a ∈ A : a−(b,c) exists}.
Recall that if b or c is not regular, then A−(b,c) = ∅ ([4, Remark 2.2 (iii)]).
In addition,
note that if b = 0 = c, then A−(b,c) = A.
In fact, in this case, given a ∈ A, a−(b,c) = 0
satisfies Definition 2.1. Moreover, if b = 0 and c ∈ A is such that c 6= 0 (respectively if b ∈ A
is such that b 6= 0 and c = 0), according to Definition 2.1, then A−(b,c) = ∅. Actually, if
b = 0 (respectively c = 0) and a−(b,c) exists, then a−(b,c) = 0, which is impossible, for c 6= 0
(respectively b 6= 0). In the following theorem the set A−(b,c) will be proved to be open.
Theorem 4.1. Let A be a unitary Banach algebra and consider b, c ∈ A. Then A−(b,c) is an
open set. Furthermore, if b, c ∈ A \ {0}, a ∈ A−(b,c) and e ∈ A is such that k e k<
,
then a + e ∈ A−(b,c) and
ka−(b,c)k
1
(a + e)−(b,c) = (1 + a−(b,c)e)−1a−(b,c) = a−(b,c)(1 + ea−(b,c))−1.
9
Proof. According to what has been said in the first paragraph of this section, it is enough to
consider the case b, c ∈ A \ {0}. Recall, in addition, that under this hypothesis, a−(b,c) 6= 0.
In fact, according to Definition 2.1, a−(b,c) = 0 implies that b = 0 = c. Moreover, note also
that since k a−(b,c)e k< 1 (respectively k ea−(b,c) k< 1), 1 + a−(b,c)e ∈ A−1 (respectively
1 + ea−(b,c) ∈ A−1).
Consider the operators La, La−(b,c), Le ∈ L(A). According to Proposition 3.9 (i), La−(b,c) =
bA,c−1(0). Now, since A is a unitary algebra,
(La)(2)
k (La)(2)
bA,c−1(0) kk Le k=k a−(b,c) kk e k< 1.
Thus, according to [10, Lemma 3.4], (La+e)(2)
bA,c−1(0) exists and
(La+e)(2)
bA,c−1(0) = (I + La−(b,c) Le)−1La−(b,c) = La−(b,c)(I + LeLa−(b,c))−1.
However,
(I + La−(b,c) Le)−1 = L(1+a−(b,c)e)−1 ,
(I + LeLa−(b,c))−1 = L(1+ea−(b,c))−1 .
Consequently,
(1 + a−(b,c)e)−1a−(b,c) = a−(b,c)(1 + ea−(b,c))−1.
Set f = (1 + a−(b,c)e)−1a−(b,c) = a−(b,c)(1 + ea−(b,c))−1. Since (La+e)(2)
bA,c−1(0) = Lf ,
according to Proposition 3.9 (ii), (a + e)−(b,c) exists and (a + e)−(b,c) = f .
In the following theorem the case of the inverse along an element will be presented. To
this end, given a unitary Banach algebra A, let A−d be the set of all elements of A invertible
along d ∈ A, i.e.,
A−d = {a ∈ A : a−d exists}.
Theorem 4.2. Let A be a unitary Banach algebra and consider d ∈ A. Then A−d is an
open set. Furthermore, if d ∈ A \ {0}, a ∈ A−d and e ∈ A is such that k e k< 1
, then
a + e ∈ A−d and
ka−dk
(a + e)−d = (1 + a−de)−1a−d = a−d(1 + ea−d)−1.
Proof. Note that according to [8, Proposition 6.1], A−d = A−(d,d). Now apply Theorem
4.1.
For sake of completeness, the case of the outer inverse with prescribed range and null
space will be considered. Let T and S be two closed subspace of the Banach spaces X and Y,
T,S will stand for the set of all operators defined on X with values in Y
respectivley.
whose outer inverse with range T and null space S exists, i.e.,
L(X, Y)(2)
L(X, Y)(2)
T,S = {A ∈ L(X, Y) : A(2)
T,S exists}.
Theorem 4.3. Let X and Y be two Banach spaces and consider T and S two closed subspaces
of X and Y, respectively. Then, L(X, Y)(2)
T,S is an open set.
Proof. According to [18, Lemma 1], if T or S is not a complemented subspaces of X and Y
respectively, then L(X, Y)(2)
T,S = ∅. On the other hand, note that if A ∈ L(X, Y) is such that
T,S exists and A(2)
A(2)
0,Y = L(X, Y).
In fact, given A ∈ L(X, Y), A(2)
T,S = 0, then T = 0 and S = Y. In addition, in this case, L(X, Y)(2)
0,Y = 0. To end the proof, apply [10, Lemma 3.4].
10
5 Continuity of the (b, c)-inverse
Recall that the notion of the gap between subspaces was used to study the continuity of the
Moore-Penrose inverse ([27]) and the Drazin inverse ([17, 28]) in the Banach context. In this
section the aforementioned notion will be used to characterize the continuity of the (b, c)-
inverse for Banach space operators and Banach algebra elements. Another notion that will
be central to present more characterizations will be the Moore-Penrose inverse.
First a particular case will be presented.
n
n
n
n
n
n
n
n
) = R(A−(Bn,Cn)
n
= 0. In fact, if (A−(Bn,Cn)
)n∈N converges to 0, then (A−(Bn,Cn)
n
An) 6= 0, then (δ(R(A−(Bn,Cn)
= 0 for each n ≥ n0. The converse implication is evident.
Remark 5.1. (i) Let X be a Banach space and consider A, B, C ∈ L(X) such that B and C are
regular, A is (B, C)-invertible and A−(B,C) = 0. Let (An)n∈N ⊂ L(X) be such that (An)n∈N
converges to A ∈ L(X), and suppose that there exist two sequence of operators (Bn)n∈N,
(Cn)n∈N ⊂ L(X) such that for each n ∈ N, Bn and Cn are regular and An is (Bn, Cn)-invertible.
Then, (A−(Bn,Cn)
)n∈N converges to A−(B,C)(= 0) if and only if there exists n0 ∈ N such that for
each n ≥ n0, A−(Bn,Cn)
An)n∈N
converges to 0, and according to [17, Lemma 3.3], (δ(R(A−(Bn,Cn)
An), 0))n∈N converges to 0.
However, according to the definition of the gap between two subspaces (see [13, Chapter 2,
Section 2, Subsection 1]), if R(A−(Bn,Cn)
An), 0)) = 1. Therefore,
there exists n0 ∈ N such that for each n ≥ n0, R(A−(Bn,Cn)
An) = 0. As a
result, A−(Bn,Cn)
(ii) Let A be a Banach algebra and consider a ∈ A and b, c ∈ A such that a is (b, c)-
invertible and a−(b,c) = 0. Suppose that there exist three sequence (an)n∈N ⊂ A and (bn)n∈N,
(cn)n∈N ⊂ A such that for each n ∈ N, an is (bn, cn)-invertible and (an)n∈N converges to a.
Then, statement (i) can be extended to this case, i.e., necessary and sufficient for (a−(bn,cn)
)n∈N
to converge to a−(b,c)(= 0) is that there exists n0 ∈ N such that for all n ≥ n0, a−(bn,cn)
=
0. Actually, to prove this equivalent condition, it is enough to apply statement (i) to La,
L−(Lb,Lc)
∈ L(A) and (Lan)n∈N, (L
=
, see Proposition 3.9 (i)). Note that since A is a unitary Banach algebra, (an)n∈N
L
(respectively (a−(bn,cn)
)n∈N) converges to a (respectively to a−(b,c)) if and only if (Lan )n∈N
n
)n∈N converges to La (respectively to La−(b,c)).
(respectively (L
a−(bn ,cn)
n
(iii) In the same conditions of statement (ii), note that according to Definition 2.1, a−(b,c) = 0
implies that b = 0 = c . Similarly, for each n ≥ n0, bn = 0 = cn, i.e., the fact that
(a−(bn,cn)
)n∈N converges to 0 determines the elements bn and cn for which an is (bn, cn)-
invertible (n ≥ n0). Naturally, given a ∈ A−(0,0), a−(0,0) = 0.
(iv) It is worth noticing that if A is a unitary Banach algebra, a ∈ A and b, c ∈ A are such
that a is (b, c)-invertible with a−(b,c) 6= 0, then b 6= 0 and c 6= 0 (see Definition 2.1).
−(Lbn ,Lcn )
an
)n∈N ⊂ L(A), (L−(Lb,Lc)
a
a−(bn ,cn)
n
= La−(b,c), L
−(Lbn ,Lcn )
an
a
n
n
n
Now the continuity of the (b, c)-inverse will be studied using the gap between subspaces.
Next follows the characterization of the continuity of (B, C)-invertible Banach space opera-
tors.
Theorem 5.2. Let X be a Banach space and consider A, B, C ∈ L(X) such that B and C are
regular, A is (B, C)-invertible and A−(B,C) 6= 0. Suppose that there exist three sequences of
operators (An)n∈N, (Bn)n∈N, (Cn)n∈N ⊂ L(X) such that for each n ∈ N, Bn and Cn are regular
and An is (Bn, Cn)-invertible. If (An)n∈N converges to A, then the following statements are
equivalent.
11
(i) The sequence (A−(Bn,Cn)
n
)n∈N converges to A−(B,C).
(ii) The sequence (A−(Bn,Cn)
n
An)n∈N (respectively (AnA−(Bn,Cn)
n
(respectively to AA−(B,C)).
)n∈N) converges to A−(B,C)A
(iii) The sequence (A−(Bn,Cn)
n
(respectively to 0).
An)n∈N (respectively (δ(N(Cn), N(C)))n∈N) converges to A−(B,C)A
(iv) The sequence (AnA−(Bn,Cn)
n
)n∈N (respectively (δ(R(Bn), R(B)))n∈N) converges to AA−(B,C)
(respectively to 0).
(v) The sequences (δ(R(Bn), R(B)))n∈N and (δ(N(Cn), N(C)))n∈N converge to 0.
(vi) The sequences (δ(R(A−(Bn,Cn)
n
), R(A−(B,C))))n∈N and (δ(N(A−(Bn ,Cn)
n
), N(A−(B,C))))n∈N
converge to 0.
Proof. It is evident that statement (i) implies statement (ii).
Observe that, since A−(B,C) is an outer inverse of A and A−(Bn,Cn)
n
is an outer inverse of
An (n ∈ N), according to Theorem 3.4,
R(A−(B,C)A) = R(A−(B,C)) = R(B),
R(A−(Bn,Cn)
An) = R(A−(Bn,Cn)
n
N(AA−(B,C)) = N(A−(B,C)) = N(C),
n
) = R(Bn), N(AnA−(Bn,Cn)
n
) = N(A−(Bn,Cn)
n
) = N(Cn).
Consequently, according to [17, Lemma 3.3], statment (ii) implies statement (iii), which in
turn implies statement (v). In addition, applying again [17, Lemma 3.3], statement (ii) also
implies statement (iv), which in turn implies statement (v). Note also that statement (vi) is
an equivalent formulation of statement (v) (Theorem 3.5).
Suppose that statement (vi) holds. According to Theorem 3.5,
A−(Bn,Cn)
n
= An
(2)
R(Bn),N (Cn),
A−(B,C) = A(2)
R(B),N (C).
Let κ =k A kk A−(B,C) k and consider n0 ∈ N such that for all n ≥ n0,
un = δ(N(Cn), N(C)) <
1
3 + κ
,
vn = δ(R(Bn), R(B)) <
1
(1 + κ)2
,
zn =k A−(B,C) kk A − An k<
2κ
(1 + κ)(4 + κ)
.
Thus, according to [10, Theorem 3.5],
k A−(Bn,Cn)
n
− A−(B,C) k≤
(1 + κ)(vn + un) + (1 + un)zn
1 − (1 + κ)vn − κun − (1 + un)zn
k A−(B,C) k,
which implies statement (i).
Next the Banach algebra case will be considered.
Theorem 5.3. Let A be a unitary Banach algebra and consider a ∈ A and b, c ∈ A such that
a is (b, c)-invertible and a−(b,c) 6= 0. Suppose that there exist three sequences (an)n∈N ⊂ A and
(bn)n∈N, (cn)n∈N ⊂ A such that an is (bn, cn)-invertible, for each n ∈ N. If (an)n∈N converges
to a, then the following statements are equivalent.
12
(i) The sequence (a−(bn,cn)
n
)n∈N converges to a−(b,c).
(ii) The sequences (a−(bn,cn)
n
an)n∈N and (ana−(bn,cn)
n
)n∈N converge to a−(b,c)a and aa−(b,c),
respectively.
(iii) The sequence (a−(bn,cn)
n
an)n∈N (respectively (δ((cn)−1(0), c−1(0)))n∈N) converges to a−(b,c)a
(respectively to 0).
(iv) The sequence (ana−(bn,cn)
n
)n∈N (respectively (δ(bnA, bA))n∈N) converges to aa−(b,c) (re-
spectively to 0).
(v) The sequences (δ(bnA, bA))n∈N and (δ((cn)−1(0), c−1(0)))n∈N converge to 0.
(vi) The sequences (δ(a−(bn,cn)
n
A, a−(b,c)A))n∈N and (δ((a−(bn,cn)
n
)−1(0), (a−(b,c))−1(0)))n∈N con-
verge to 0.
(vii) The sequence (ana−(bn,cn)
n
(respectively to 0).
)n∈N (respectively (δ((bn)−1(0), b−1(0)))n∈N) converges to aa−(b,c)
(viii) The sequence (a−(bn,cn)
n
an)n∈N (respectively (δ(Acn, Ac))n∈N) converges to a−(b,c)a (re-
spectively to 0).
(ix) The sequences (δ(Acn, Ac))n∈N and (δ((bn)−1(0), b−1(0)))n∈N converge to 0.
(x) The sequences (δ(Aa−(bn,cn)
n
, Aa−(b,c)))n∈N and (δ((a−(bn,cn)
n
)−1(0), (a−(b,c))−1(0)))n∈N con-
verge to 0.
Proof. Recall that, according to Proposition 3.9 (i),
La−(b,c) = L−(Lb,Lc)
a
= (La)(2)
bA,c−1(0),
L
a−(bn ,cn)
n
= L
−(Lbn ,Lcn )
an
= (Lan)(2)
bnA,cn
−1(0),
for each n ∈ N. In addition, note that since A is a unitary Banach algebra, (an)n∈N converges
to a if and only if (Lan )n∈N converges to La. Similarly, a necessary and sufficient condition for
(a−(bn,cn)
)n∈N) to converge to a−(b,c)
(respectively to a−(b,c)a and aa−(b,c)) is that (L
Lan)n∈N
and (Lan L
)n∈N) converges to La−(b,c) (respectively to La−(b,c) La and LaLa−(b,c)).
)n∈N (respectively for (a−(bn,cn)
an)n∈N and (ana−(bn,cn)
)n∈N (respectively (L
a−(bn ,cn)
n
a−(bn ,cn)
n
n
n
n
a−(bn ,cn)
n
Now, to prove the equivalence between statements (i)-(vi), apply Theorem 5.2 to La and
(Lan)n∈N. Recall that according to the proof of Proposition 3.9, c−1(0) = (a−(b,c))−1(0) and
n (0) = (a−(bn,cn)
c−1
)−1(0).
n
A similar argument, using in particular Proposition 3.9 (iii) and Theorem 5.2, proves the
equivalence between statements (i), (ii) and (vii)-(x).
Next the continuity of the (b, c)-inverse will be characterized using the Moore-Penrose
inverse.
Theorem 5.4. Let A be a unitary Banach algebra and consider a ∈ A and b, c ∈ A† such that
a is (b, c)-invertible and a−(b,c) 6= 0. Suppose that there exist three sequences (an)n∈N ⊂ A
and (bn)n∈N, (cn)n∈N ⊂ A† such that an is (bn, cn)-invertible, for each n ∈ N. If (an)n∈N
converges to a, then the following statements are equivalent.
13
(i) The sequence (a−(bn,cn)
n
)n∈N converges to a−(b,c).
(ii) The sequence (a−(bn,cn)
n
an)n∈N converges to a−(b,c)a and the sequences
(c†c(1 − c†
ncn))n∈N,
(c†
ncn(1 − c†c))n∈N
converge to 0.
(iii) The sequence (ana−(bn,cn)
n
)n∈N converges to aa−(b,c) and the sequences
((1 − bb†)bnb†
n)n∈N,
((1 − bnb†
n)bb†)n∈N
converge to 0.
(iv) The sequences
((1 − bb†)bnb†
(c†c(1 − c†
n)n∈N,
ncn))n∈N,
((1 − bnb†
(c†
n)bb†)n∈N,
ncn(1 − c†c))n∈N
converge to 0.
(v) The sequence (ana−(bn,cn)
n
)n∈N converges to aa−(b,c) and the sequences
(bb†(1 − bnb†
n))n∈N,
(bnb†
n(1 − bb†))n∈N
converge to 0.
(vi) The sequence (a−(bn,cn)
n
an)n∈N converge to a−(b,c)a and the sequences
((1 − c†c)c†
ncn)n∈N,
((1 − c†
ncn)c†c)n∈N
converge to 0.
(vii) The sequences
(bb†(1 − bnb†
((1 − c†c)c†
n))n∈N,
ncn)n∈N,
(bnb†
((1 − c†
n(1 − bb†))n∈N,
ncn)c†c)n∈N
converge to 0.
(viii) The sequences (bnb†
n)n∈N and (c†
ncn)n∈N converge to bb† and c†c, respectively.
14
Proof. Note that
bA = bb†A,
Ac = Ac†c,
bnA = bnb†
Acn = Ac†
nA,
nc,
c−1(0) = (1 − c†c)A,
b−1(0) = A(1 − bb†),
c−1
n (0) = (1 − c†
ncn)A,
(bn)−1(0) = A(1 − bnb†
n).
Since A is a unitary Banach algebra, according to [27, Lemma 2.2],
n k, k (1 − bnb†
n)bb† k},
δ(bnA, bA) = max{k (1 − bb†)bnb†
δ((cn)−1(0), c−1(0)) = max{k c†c(1 − c†
δ((bn)−1(0), b−1(0)) = max{k bb†(1 − bnb†
δ(Acn, Ac) = max{k (1 − c†c)c†
ncn) k, k c†
ncn(1 − c†c k},
n) k, k bnb†
n(1 − bb†)} k},
ncn k, k (1 − c†
ncn)c†c k}.
Now, to prove the equivalence among statements (i)-(vii), apply Theorem 5.3 and use the
identities that have been proved.
Suppose that statement (i) holds. Observe that
bnb†
n = bb†bnb†
nbb† + bb†bnb†
n(1 − bb†) + (1 − bb†)bnb†
nbb† + (1 − bb†)bnb†
n(1 − bb†).
Thus, according to statements (iv) and (vii), if
fn = bb†bnb†
n(1 − bb†) + (1 − bb†)bnb†
nbb† + (1 − bb†)bnb†
n(1 − bb†),
then the sequence (fn)n∈N converges to 0.
sequence (bb†(1 − bnb†
to bb†. Therefore, (bnb†
converges to c†c.
n)bb†)n∈N converges to 0, which implies that (bb†bnb†
n)n∈N converges to bb†. A similar argument proves that (c†
In addition, according to statement (vii), the
nbb†)n∈N convergs
ncn)n∈N
It is evident that statement (viii) implies statement (vii).
In the following corollary, a particular case will be presented.
Corollary 5.5. Let A be a unitary Banach algebra and consider a ∈ A and b, c ∈ A† such
that a is (b, c)-invertible and a−(b,c) 6= 0. Suppose that there exist three sequences (an)n∈N ⊂ A
and (bn)n∈N, (cn)n∈N ⊂ A† such that an is (bn, cn)-invertible, for each n ∈ N. Suppose, in
addition, that the sequences (bn)n∈N, (cn)n∈N, (b†
n)n∈N converge to b, c, b† and c†,
respectively. Then, if (an)n∈N converges to a, the sequence (a−(bn,cn)
)n∈N converges to a−(b,c).
n)n∈N and (c†
n
Proof. Apply Theorem 5.4 (viii).
In the context of C ∗-algebras, Theorem 5.4 and Corollary 5.5 can be reformulated as
follows.
Theorem 5.6. Let A be a C ∗-algebra and consider a ∈ A and b, c ∈ A such that a is (b, c)-
invertible and a−(b,c) 6= 0. Suppose that there exist three sequences (an)n∈N ⊂ A and (bn)n∈N,
(cn)n∈N ⊂ A such that an is (bn, cn)-invertible, for each n ∈ N. If (an)n∈N converges to a,
then the following statements are equivalent.
15
(i) The sequence (a−(bn,cn)
n
)n∈N converges to a−(b,c).
(ii) The sequences (a−(bn,cn)
n
an)n∈N and (c†
ncn)n∈N converge to a−(b,c)a and c†c, respectively.
(iii) The sequences (ana−(bn,cn)
n
)n∈N and (bnb†
n)n∈N converge to aa−(b,c) and bb†, respectively.
(iv) The sequences (bnb†
In addition, if the sequence (bn)n∈N converges to b, then statement (iii) is equivalent to:
ncn)n∈N converge to bb† and c†c, respectively.
n)n∈N and (c†
(v) the sequences (ana−(bn,cn)
Moreover, if the sequence (cn)n∈N converge to c, then statement (ii) is equivalent to:
n)n∈N converge to aa−(b,c) and b†, respectively.
)n∈N and (b†
n
(vi) the sequences (a−(bn,cn)
n
an)n∈N and (c†
n)n∈N converge to a−(b,c)a and c†, respectively.
Furthermore, if the sequences (bn)n∈N and (cn)n∈N converge to b and c, respectively, then
statement (iv) is equivalent to:
(vii) the sequences (b†
n)n∈N and (c†
n)n∈N converge to b† and c†, respectively.
Proof. Note that since
((1 − bb†)bnb†
n)∗ = bnb†
n(1 − bb†),
((1 − bnb†
n)bb†)∗ = bb†(1 − bnb†
n),
according to the proof of statement (vii) implies statement (viii) in Theorem 5.4, statement
(iii) is equivalent to Theorem 5.4 (iii).
Now observe that,
(c†c(1 − c†
ncn))∗ = (1 − c†
ncn)c†c,
(c†
ncn(1 − c†c))∗ = (1 − c†c)c†
ncn.
Then, a similar argument proves that statement (ii) is equivalent to Theorem 5.4 (ii).
To prove statements (v)-(vii), apply [16, Theorem 1.6].
Since the inverse along an element is a particular case of the (b, c)-inverse, it is possible
to apply the results of this section to obtain characterizations of the continuity of the in-
verse along an element for Banach space operators and for Banach algebra and C ∗-algebra
elements. In fact, given a Banach algebra (or a C ∗-algebra) A, a ∈ A and d ∈ A, to state
the aforementioned characterizations, it is enough to apply the corresponding result to the
(d, d)-inverse (see section 2); in the Banach operator case it is possible to proceed in a similar
way.
In order not to extend unnecessarily this work, these results will not be stated; the
details are left to the interested reader.
To end this section, as an application, characterizations of the continuity of the Moore-
Penrose inverse in the contexts of Banch algebras and Banach space operators will be given;
in the former frame, compare with [27, Theorem 2.5].
Corollary 5.7. Let A be a unitary Banach algebra and consider a ∈ A† \ 0. Suppose that
there exists a sequence (an)n∈N ⊂ A† such that (an)n∈N converges to a. Then the following
statements are equivalent.
16
(i) The sequence (a†
n)n∈N converges to a†.
(ii) The sequence (a†
nan)n∈N converges to a†a and the sequences
(aa†(1 − ana†
n))n∈N,
(ana†
n(1 − aa†))n∈N
converge to 0.
(iii) The sequence (ana†
n)n∈N converges to aa† and the sequences
((1 − a†a)a†
nan)n∈N,
((1 − a†
nan)a†a)n∈N
converge to 0.
(iv) The sequences
((1 − a†a)a†
(aa†(1 − ana†
nan)n∈N,
n))n∈N,
((1 − a†
(ana†
nan)a†a)n∈N,
n(1 − aa†))n∈N
converge to 0.
(v) The sequence (ana†
n)n∈N converges to aa† and the sequences
(a†a(1 − a†
nan))n∈N,
(a†
nan(1 − a†a))n∈N
converge to 0.
(vi) The sequence (a†
nan)n∈N converge to a†a and the sequences
((1 − aa†)ana†
n)n∈N,
((1 − ana†
n)aa†)n∈N
converge to 0.
(vii) The sequences
(a†a(1 − a†
((1 − aa†)ana†
nan))n∈N,
n)n∈N,
(a†
((1 − ana†
nan(1 − a†a))n∈N,
n)aa†)n∈N
converge to 0.
(viii) The sequences (a†
nan)n∈N and (ana†
n)n∈N converge to a†a and aa†, respectively.
Proof. According to [8, Proposition 6.1], given x ∈ A†, x† = x−(x†,x†). To conclude the proof,
apply Theorem 5.4 to a, b = c = a†, an and bn = cn = a†
n (n ∈ N). Note that if x ∈ A†, then
x† ∈ A† and (x†)† = x.
17
Remark 5.8. In the same conditions of Corollary 5.7, note that according to [27, Lemma
2.2],
n
δ(R(La†
δ(N(L
δ(N(R
δ(R(R
a†
n
a†
n
), R(La† )) = δ(a†
), N(La† )) = δ((a†
), N(Ra† )) = δ((a†
), R(Ra† )) = δ(Aa†
nA, a†A) = max{k (1 − a†a)a†
n)−1(0), (a†)−1(0)) = max{k aa†(1 − ana†
n)−1(0), (a†)−1(0)) = max{k a†a(1 − a†
n, Aa†) = max{k (1 − aa†)ana†
a†
n
n k, k (1 − ana†
n)aa† k}.
nan k, k (1 − a†
nan)a†a} k},
n) k, k ana†
n(1 − aa†) k},
nan(1 − a†a) k},
nan) k, k a†
Compare Corollary 5.7 with [27, Theorem 2.5].
Corollary 5.9. Let X be a Banach space and consider a Moore-Penrose invertible operator
A ∈ L(X), A 6= 0. Suppose that there exists a sequence of operators (An)n∈N such that for
each n ∈ N, An is Moore-Penrose invertible. If (An)n∈N converges to A, then the following
statements are equivalent.
(i) The sequence (A†
n)n∈N converges to A†.
(ii) The sequence (A†
nAn)n∈N (respectively (AnA†
n)n∈N) converges to A†A (respectively to
AA†).
(iii) The sequence (A†
nAn)n∈N (respectively (δ(N(A†
n), N(A†)))n∈N) converges to A†A (respec-
tively to 0).
(iv) The sequence (AnA†
n)n∈N (respectively (δ(R(A†
n), R(A†)))n∈N) converges to AA† (respec-
tively to 0).
(v) The sequences (δ(R(A†
n), R(A†)))n∈N and (δ(N(A†
n), N(A†)))n∈N converge to 0.
Proof. Proceed as in the proof of Corollary 5.7 and apply Theorem 5.2.
6 Differentiation of the (b, c)-inverse
Next follows the first characterization of this section. Observe that if A is a Banach algebra,
J ⊆ R and there exist functions a : J → A and b, c : J → A such that for each t ∈ J,
a(t)−(b(t),c(t)) exists, then a−(b,c) : J → A is the following funciton: a−(b,c)(t) = a(t)−(b(t),c(t)).
Theorem 6.1. Let A be a Banach algebra and consider a function a : J → A, where J ⊆ R is
an open set. Let b, c : J → A be two functions such that for each t ∈ J , a(t)−(b(t),c(t)) exists.
Suppose that there exist functions g, h : J → A such that for each t ∈ J , g(t) ∈ b(t){1} and
h(t) ∈ c(t){1}. Then, given t0 ∈ J , if the functions a, bg, hc : J → A are differentiable at
t0, the following statements are equivalent.
(i) The funciton a−(b,c) : J → A is continuous at t0.
(ii) The funciton a−(b,c) : J → A is differentiable at t0.
18
Moreover, in this case,
(a−(b,c))′(t0) =a(t0)−(b(t0),c(t0))(hc)′(t0)a(t0)−(b(t0),c(t0))
− (1 − a(t0)−(b(t0),c(t0))a(t0))(bg)′(t0)a(t0)−(b(t0),c(t0))
− a(t0)−(b(t0),c(t0))a′(t0)a(t0)−(b(t0),c(t0)).
Proof. It is enough to prove that statement (i) implies statement (ii). This proof can be
deduced from [4, Corollary 7.12]. In fact, according to this result,
a(t)−(b(t),c(t)) − a(t0)−(b(t0),c(t0)) = a(t)−(b(t),c(t))(hc(t) − hc(t0))a(t0)−(b(t0),c(t0))
− (1 − a(t)−(b(t),c(t))a(t))(bg(t) − bg(t0))a(t0)−(b(t0),c(t0))
+ a(t)−(b(t),c(t))(a(t0) − a(t))a(t0)−(b(t0),c(t0)).
Now divide each term by t − t0 and note that the limt leads the formula of (a−(b,c))′(t0).
When the function b and c in Theorem 6.1 are such that b(J) ⊆ A† and c(J) ⊆ A†,
the aforementioned result can be reformulated as follows. Note first that if A is a Banach
algebra and x : J → A† is a function (J ⊆ R), then x† : J → A denotes the following function:
x†(t) = (x(t))†.
Corollary 6.2. Let A be a Banach algebra and consider a function a : J → A, where J ⊆ R
is an open set. Let b, c : J → A† be two functions such that for each t ∈ J , a(t)−(b(t),c(t))
exists. Then, given t0 ∈ J , if the functions a, bb†, c†c : J → A are differentiable at t0, the
following statements are equivalent.
(i) The funciton a−(b,c) : J → A is continuous at t0.
(ii) The funciton a−(b,c) : J → A is differentiable at t0.
Moreover, in this case,
(a−(b,c))′(t0) =a(t0)−(b(t0),c(t0))(c†c)′(t0)a(t0)−(b(t0),c(t0))
− (1 − a(t0)−(b(t0),c(t0))a(t0))(bb†)′(t0)a(t0)−(b(t0),c(t0))
− a(t0)−(b(t0),c(t0))a′(t0)a(t0)−(b(t0),c(t0)).
Proof. Apply Theorem 6.1 to the case under consideration.
As an application of Theorem 6.1, a characterization of the differentiability of the Moore-
Penrose inverse in the Banach context will be given.
Corollary 6.3. Let A be a Banach algebra and consider a function a : J → A†, where J ⊆ R is
an open set. Suppose that there exists t0 ∈ J such that the function a : J → A is differentiable
at t0. Then, the following statements are equivalent.
(i) The function a† : J → A is differentiable at t0.
19
(ii) The function a† : J → A is continuous at t0 and the functions aa†, a†a : J → A are
differentiable at t0.
Furthermore, in this case,
(a†)′(t0) =a†(t0)(aa†)′(t0)a†(t0) − (1 − a†a(t0))(a†a)′(t0)a†(t0) − a†(t0)a′(t0)a†(t0).
Proof. Recall that given x ∈ A†, x† = x−(x†,x†) ([8, Propostion 6.1])). To conclude the proof
apply Theorem 6.1 to the function a, b, c : J → A, where b = c = a†.
In the frame of C ∗-algebras, the hypotheses of Corollary 6.2 can be lightened.
Corollary 6.4. Let A be a C ∗-algebra and consider a function a : J → A, where J ⊆ R is
an open set. Let b, c : J → A be two functions such that for each t ∈ J , a(t)−(b(t),c(t)) exists.
Then, given t0 ∈ J , if the functions a, b, c : J → A are differentiable at t0, the functions
b†, c† : J → A are continuous at t0, and a(t0)−(b(t0),c(t0)) 6= 0, the following statements are
equivalent.
(i) The funciton a−(b,c) : J → A is continuous at t0.
(ii) The funciton a−(b,c) : J → A is differentiable at t0.
Moreover, in this case,
(a−(b,c))′(t0) =a(t0)−(b(t0),c(t0))((c†)′(t0)c(t0) + c†(t0)c′(t0))a(t0)−(b(t0),c(t0))
− (1 − a(t0)−(b(t0),c(t0))a(t0))(b′(t0)b†(t0) + b(t0)(b†)′(t0))a(t0)−(b(t0),c(t0))
− a(t0)−(b(t0),c(t0))a′(t0)a(t0)−(b(t0),c(t0)).
Proof. Note that the condition a(t0)−(b(t0),c(t0)) 6= 0 implies that b(t0) 6= 0 and c(t0) 6= 0
(Definition 2.1). Thus, according to [16, Theorem 2.1], the functions b†, c† : J → A are
differentiable at t0. To conclude the proof, apply Corollary 6.2.
As in section 5, the results concerning the differentiability of the inverse along an element
can be deduced from Theorem 6.1, Corollary 6.3 and Corollary 6.4. The details are left to
the interested reader.
7 The outer inverse A(2)
T,S
Although similar arguments to the ones used in sections 5 and 6 will be applied to study the
continuity and the differentiability of the outer inverse A(2)
T,S , the results of this section can not
be derived from the corresponding ones concerning the (b, c)-inverse. In fact, in what follows
operators between two different Banach spaces will be considered. First the gap between
subspaces will be used to characterize the continity of the A(2)
T,S .
Theorem 7.1. Let X and Y be Banach spaces and consider A ∈ L(X, Y) and two subspaces
T ⊆ X and S ⊆ Y such that A(2)
T,S exists. Let (An)n∈N ⊂ L(X, Y) and consider (Tn)n∈N and
(Sn)n∈N two sequences of subspaces of X and Y, respectively, such that (An)(2)
exists, for
each n ∈ N. Suppose that (An)n∈N converges to A. The following statements are equivalent.
Tn,Sn
20
(i) The sequence ((An)(2)
Tn,Sn
)n∈N converges to A(2)
T,S.
(ii) The sequences ((An)(2)
Tn,Sn
An)n∈N and (An(An)(2)
Tn,Sn
)n∈N converge to A(2)
T,SA and AA(2)
T,S,
respectively.
(iii) The sequence ((An)(2)
Tn,Sn
tively to 0).
An)n∈N (respectively (δ(Sn, S))n∈N) converges to A(2)
T,SA (respec-
(iv) The sequence (An(An)(2)
Tn,Sn
tively to 0).
)n∈N (respectively (δ(Tn, T))n∈N) converges to AA(2)
T,S (respec-
(v) The sequences (δ(Tn, T))n∈N and (δ(Sn, S))n∈N converge to 0.
Proof. First define
P = AA(2)
T,S,
Pn = An(An)(2)
Tn,Sn
,
Q = A(2)
T,SA,
Qn = (An)(2)
Tn,Sn
An
(P , Pn ∈ L(Y)• and Q, Qn ∈ L(X)•, n ∈ N). Note that
R(Q) = T,
R(Qn) = Tn,
N(P ) = S,
N(Pn) = Sn.
Suppose that statement (v) holds. If A(2)
It is evident that statement (i) implies statement (ii).
Now, according to [17, Lemma 3.3], if (Pn)n∈N converges to P (respectively (Qn)n∈N con-
verges to Q), then (δ(Sn, S))n∈N (respectively (δ(Tn, T))n∈N) converges to 0. Thus, statement
(ii) implies statement (iii) (respectively statement (iv)), which in turn implies statement (v).
T,S = 0, then T = 0 and S = Y. In particular,
(δ(Tn, 0))n∈N converges to 0. However, according to [13, Chapter 2, Section 2, Subsection
1], if Tn 6= 0, δ(Tn, 0) = 1. Thus, there exists n0 ∈ N such that for each n ≥ n0, Tn = 0,
which implies that (An)(2)
T,S 6= 0 and
appy [10, Theorem 3.5].
= 0 (n ≥ n0). To conclude the proof, assume that A(2)
Tn,Sn
In the follwing result the Moore-Penrose inverse will be used to characterize the continuity
of the outer inverse A(2)
T,S . Although it will not be used in this article, recall that given a
Banach space X and P and Q ∈ L(X)• such that P and Q are hermitian, if R(P ) = R(Q),
then P = Q ([23, Theorem 2.2]). In particular, given a subspace M ⊆ X, there exists at most
one hermitian idempotent R such that R(R) = M.
Theorem 7.2. Let X and Y be Banach spaces and consider A ∈ L(X, Y) and two subspaces
T ⊆ X and S ⊆ Y such that A(2)
T,S exists. Let (An)n∈N ⊂ L(X, Y) and consider (Tn)n∈N
and (Sn)n∈N two sequences of subspaces of X and Y, respectively, such that (An)(2)
exists,
for each n ∈ N. Suppose, in addition, that there exit hermitian idempotents U ∈ L(X)
(respectively V ∈ L(Y)) and Un ∈ L(X) (respectively Vn ∈ L(Y)) such that R(U ) = T and
R(Un) = Tn (respectively N(V ) = S and N(Vn) = Sn), n ∈ N. If (An)n∈N converges to A,
then the following statements are equivalent.
Tn,Sn
(i) The sequence ((An)(2)
Tn,Sn
)n∈N converges to A(2)
T,S.
21
(ii) The sequence ((An)(2)
An)n∈N converges to A(2)
Tn,Sn
T,SA and the sequences (V (I − Vn))n∈N
and (Vn(I − V ))n∈N converge to 0.
(iii) The sequence (An(An)(2)
)n∈N converges to AA(2)
Tn,Sn
T,S and the sequences ((I − U )Un)n∈N
and ((I − Un)U )n∈N converge to 0.
(iv) The sequences (V (I − Vn))n∈N, (Vn(I − V ))n∈N, ((I − U )Un)n∈N and ((I − Un)U )n∈N
converge to 0.
Proof. Note that δ(Tn, T) = δ(R(Un), R(U )). Thus, according to [27, Lemma 2.2], the
sequence (δ(Tn, T))n∈N converges to 0 if and only if the sequences ((I − U )Un)n∈N and
((I − Un)U )n∈N converge to 0. Similarly, since δ(Sn, S) = δ(R(I − Vn), R(I − V )), according
to [27, Lemma 2.2], the sequence (δ(Sn, S))n∈N converges to 0 if and only if the sequences
(V (I − Vn))n∈N and (Vn(I − V ))n∈N converge to 0. To conclude the proof, apply Theorem
7.1.
Next the continuity of the outer inverse A(2)
T,S will be studied in the context of Hilbert
spaces.
Theorem 7.3. Let H and K be Hilbert spaces and consider A ∈ L(H, K) and two subspaces
T ⊆ H and S ⊆ K such that A(2)
T,S exists. Let (An)n∈N ⊂ L(H, K) and consider (Tn)n∈N and
(Sn)n∈N two sequences of subspaces of H and K, respectively, such that (An)(2)
exists, for
each n ∈ N. If (An)n∈N converges to A, then the following statements are equivalent.
Tn,Sn
(i) The sequence ((An)(2)
Tn,Sn
)n∈N converges to A(2)
T,S.
(ii) The sequences ((An)(2)
Tn,Sn
An)n∈N and (An(An)(2)
Tn,Sn
)n∈N converge to A(2)
T,SA and AA(2)
T,S,
respectively.
(iii) The sequence ((An)(2)
Tn,Sn
to P ⊥
S⊥).
An)n∈N (respectively (P ⊥
S⊥
n
)n∈N) converges to A(2)
T,SA (respectively
(iv) The sequence (An(An)(2)
)n∈N (respectively (P ⊥
Tn
)n∈N) converges to AA(2)
T,S (respectively
Tn,Sn
to P ⊥
T ).
(v) The sequences (P ⊥
S⊥
n
)n∈N and (P ⊥
Tn
Proof. Note that
)n∈N converge to P ⊥
S⊥ and P ⊥
T , respectively.
U = P ⊥
T ,
Un = P ⊥
Tn ,
V = P ⊥
S⊥ ,
Vn = P ⊥
S⊥
n
,
satisfies the hypoteses of Theorem 7.2. First it will be proved that the sequences ((I −
U )Un)n∈N and ((I − Un)U )n∈N converge to 0 if and only if the sequence (P ⊥
)n∈N converges
Tn
to P ⊥
T .
Since U , Un ∈ L(H) are self-adjoint, the sequences ((I − U )Un)n∈N and ((I − Un)U )n∈N
converge to 0 if and only if the sequences (Un(I − U ))n∈N and (U (I − Un))n∈N converge to 0.
In particular, the sequences ((I − U )UnU )n∈N, ((I − U )Un(I − U ))n∈N and (U Un(I − U ))n∈N
22
converge to 0. In addition, since (U (I − Un))n∈N converges to 0, the sequence (U UnU )n∈N
converges to U . However, since
Un = U UnU + U Un(I − Un) + (I − U )UnU + (I − Un)Un(I − U ),
the sequence (P ⊥
Tn
)n∈N converges to P ⊥
T .
A similar argument proves that the sequences (V (I −Vn))n∈N and (Vn(I −V ))n∈N converge
to 0 if and only if the sequence (P ⊥
S⊥
n
)n∈N converges to P ⊥
S⊥.
To conlcude the proof, apply Theorem 7.1 and Theorem 7.2.
To study the differentiation of the outer inverse A(2)
T,S , it is necessary to present a prelimiery
result first.
Lemma 7.4. Let X and Y be two Banach spaces and consider A, B ∈ L(X, Y). Let T,
V ⊆ X and S, U ⊆ Y be two pairs of subspaces such that A(2)
V,U exist and consider
idempotents PT, PV ∈ L(X) and PS, PU ∈ L(Y) such that R(PT) = T, R(PV) = V, R(PS) = S
and R(PU) = U. Then
T,S and B(2)
B(2)
V,U − A(2)
T,S =B(2)
V,U(PS − PU)(IY − AA(2)
− B(2)
V,U(B − A)A(2)
T,S.
T,S) + (IX − B(2)
V,UB)(PV − PT)A(2)
T,S
Proof. Let PT, PV ∈ L(X)• be such that R(PT) = T and R(PV) = V. According to [18, Lemma
1],
(IX − B(2)
V,UB)PV = 0,
PTA(2)
T,S = A(2)
T,S.
Then,
B(2)
V,UBA(2)
T,S − A(2)
T,S = −(IX − B(2)
V,UB)PTA(2)
T,S = (IX − B(2)
V,UB)(PV − PT)A(2)
T,S.
Now consider PS, PU ∈ L(Y)• such that R(PS) = S and R(PU) = U. According to [18,
Lemma 1],
(IY − PS)(IY − AA(2)
T,S) = 0,
B(2)
V,U = B(2)
V,U(IY − PU).
Consequently,
B(2)
V,U − B(2)
V,UAA(2)
T,S = B(2)
= B(2)
V,U(IY − PU)(IY − AA(2)
V,U(PS − PU)(IY − AA(2)
T,S) = B(2)
T,S).
V,U((IY − PU) − (IY − PS))(IY − AA(2)
T,S)
23
Therefore,
B(2)
V,U − A(2)
T,S = B(2)
V,U(PS − PU)(IY − AA(2)
T,S) + B(2)
V,UAA(2)
T,S
+ (IX − B(2)
V,UB)(PV − PT)A(2)
T,S − B(2)
V,UBA(2)
T,S
= B(2)
V,U(PS − PU)(IY − AA(2)
T,S) + (IX − B(2)
V,UB)(PV − PT)A(2)
T,S
− B(2)
V,U(B − A)A(2)
T,S.
In what follows, the differentiation of the outer inverse A(2)
T,S will be studied.
Theorem 7.5. Let X and Y be Banach spaces and consider J ⊆ R an open set. Suppose that
there exist functions A : J → L(X, Y), P : J → L(X)• and Q : J → L(Y)• such that for each
t ∈ J , (A(t))(2)
R(P(t)),R(Q(t)) exists. If the functions A, P and Q are differentiable at t0, then
the following statements are equivalent.
(i) The function A
(2)
P,Q : J → L(Y, X) is continuous at t0,
(ii) the function A
(2)
P,Q : J → L(Y, X) is differentiable at t0,
(2)
P,Q(t) = (A(t))(2)
R(P(t)),R(Q(t)) .
where A
Furthermore,
(A
(2)
P,Q)′(t0) = − A
− A
(2)
P,Q(t0)Q′(t0)(IY − A(t0)A
(2)
P,Q(t0)A′(t0)A
(2)
P,Q(t0).
(2)
P,Q(t0)) + (IX − A
(2)
P,Q(t0)A(t0))P′(t0)A
(2)
P,Q(t0)
Proof. It is enough to prove that statement (i) implies statement (ii). This proof can be
derived from Lemma 7.4. In fact, according to this result,
A
(2)
P,Q(t) − A
(2)
P,Q(t0) = − A
(2)
P,Q(t)(Q(t) − Q(t0))(IY − A(t0)A
(2)
P,Q(t0))
+ (IX − A(2)
P,Q(t)A(t))(P(t) − P(t0))A(2)
P,Q(t0)
− A
(2)
P,Q(t)(A(t) − A(t0))A
(2)
P,Q(t0).
Now divide each term by t − t0 and note that the limit leads to (A
(2)
P,Q)′(t0).
In the Hilbert space operators context, Theorem 7.5 can be reformulated as follows.
Corollary 7.6. Let H and K be Hilbert spaces and consider J ⊆ R an open set. Suppose that
there exist functions A : J → L(X, Y), P⊥ : J → L(X)• and Q⊥ : J → L(Y)• such that for each
t ∈ J , P⊥(t) ∈ L(X) and Q⊥(t) ∈ L(Y) are orthogonal idempotents and (A(t))(2)
R(P⊥(t)),R(Q⊥(t))
exists. If the functions A, P⊥ and Q⊥ are differentiable at t0, then the following statements
are equivalent.
24
(i) The function A
(2)
P⊥,Q⊥ : J → L(Y, X) is continuous at t0,
(ii) the function A
(2)
P⊥,Q⊥ : J → L(Y, X) is differentiable at t0,
(2)
P⊥,Q⊥(t) = (A(t))(2)
R(P⊥(t)),R(Q⊥ (t))
.
where A
Furthermore,
(A
(2)
P⊥,Q⊥)′(t0) = − A
(2)
P⊥,Q⊥(t0)(Q⊥)′(t0)(IY − A(t0)A
P⊥,Q⊥(t0)A(t0))(P⊥)′(t0)A
(2)
(2)
P⊥,Q⊥(t0))
(2)
P⊥,Q⊥(t0)
+ (IX − A
− A
(2)
P⊥,Q⊥(t0)A′(t0)A
(2)
P⊥,Q⊥(t0).
Proof. Apply Theorem 7.5 to the case under consideration.
References
[1] J. Ben´ıtez, E. Boasso, The inverse along an element in rings, Electron. J. Linear Algebra
31 (2016), 572-592.
[2] J. Ben´ıtez, E. Boasso, The inverse along an element in rings with an involution, Banach
algebras and C ∗-algebras, Linear Multilinear Algebra 65 (2017), 284-299.
[3] E. Boasso, On the Moore-Penrose inverse, EP Banach space operators and EP Banach
algebra elements, J. Math. Anal. Appl. 339 (2008), 1003-1014.
[4] E. Boasso, G. Kant´un-Montiel, The (b, c)-inverse in rings and in the Banach context,
Mediterr. J. Math. 14 (2017), Art. 112, 21p.
[5] F. F. Bonsall, J. Duncan, Complete normed algebras, Springer Verlag, Berlin, Heidelberg,
New York 1973.
[6] N. Castro-Gonz´alez, J. Chen, L. Wang, Further results on generalized inverses in rings
with involution, Electron. J. Linear Algebra 30 (2015), 118-134.
[7] H. R. Dowson, Spectral Theory of Linear operators, Academic Press, Inc., London, New
York, San Francisco 1978.
[8] M. P. Drazin, A class of outer generalized inverses, Linear Algebra Appl. 436 (2012),
1909-1923.
[9] M. P. Drazin, Commuting properties of generalized inverses, Linear Multilinear Algebra
61 (2013), 1675-1681.
[10] F. Du, Y. Xue, Perturbation analysis of A(2)
T,S on Banach spaces, Electron. J. Linear
Algebra 23 (2012), 586-598.
[11] R. Harte, M. Mbekhta, On generalized inverses in C ∗-algebras, Studia Math. 103 (1992),
71-77.
25
[12] R. Harte, M. Mbekhta, On generalized inverses in C ∗-algebras II, Studia Math. 106
(1993), 129-138.
[13] T. Kato, Perturbation Theory for Linear Operators, Springer Verlag, Berlin, Heidelberg,
New York, 1980.
[14] Y. Ke, J. Chen, The Bott-Duffin (e, f )-inverses and their applications, Linear Algebra
Appl. 489 (2016), 61-74.
[15] Y. Ke, Z. Wang, J. Chen, The (b, c)-inverse for products and lower triangular matrices,
J. Algebra Appl. 16 (2017), doi: 10.1142/S021949881750222X.
[16] J. J. Koliha, Continuity and diffrentiability of the Moore-Penrose inverse in C ∗-algebras,
Math. Scand. 88 (2001), 154-160.
[17] J.J. Koliha, V. Rakocevi´c, Continuity of the Drazin inverse II, Studia Math. 131 (1998),
167-177.
[18] X. Liu, Y. Yu, J. Zhong, Y. Wei, Integral and limit representations of the outer inverse
in Banach space, Linear Multilinear Algebra 60 (2012), 333-347.
[19] X. Mary, On generalized inverse and Green's relations, Linear Algebra Appl. 434 (2011),
1836-1844.
[20] X. Mary, Natural generalized inverse and core of an element in semigroups, rings and
Banach and operator algebras, Eur. J. Pure Appl. Math. 5 (2012), 160-173.
[21] X. Mary, P. Patr´ıcio, The inverse along a lower triangular matrix, Appl. Math. Comput.
219 (2012), 886-891.
[22] X. Mary, P. Patr´ıcio, Generalized inverses modulo H in semigroups and rings, Linear
Multilinear Algebra 61 (2013), 1130-1135.
[23] T. Palmer, Unbounded normal operators on Banach spaces, Trans. Amer. Math. Soc.
133 (1968), 385-414.
[24] R. Penrose, A generalized inverse for matrices, Proc. Cambridge Philos. Soc. 51 (1955),
406-413.
[25] D. S. Raki´c, A note on Rao and Mitra's constrained inverse and Drazin's (b, c) inverse,
Linear Algebra Appl. 523 (2017), 102-108.
[26] V. Rakocevi´c, Moore-Penrose inverse in Banach algebras, Proc. Roy. Irish Acad. Sect. A
88 (1988), 57-60.
[27] V. Rakocevi´c, On the continuity of the Moore-Penrose inverse in Banach algebras, Facta
Univ. Ser. Math. Inform. 6 (1991), 133-138.
[28] V. Rakocevi´c, Continuity of the Drazin inverse, J. Operator Theory 41 (1999), 55-68.
[29] I. Vidav, Eine metrische Kennzeichnung der selbstadjungierten Operatoren, Math. Z. 66
(1956), 121-128.
26
[30] L. Wang, N. Castro-Gonz´alez, J. Chen, Characterizations of outer generalized inverses,
Canad. Math. Bull. 60 (2017), 861-871.
[31] Y. Yu, Y. Wei, The representation and computational procedures for the generalized
T, S of an operator A in Hilbert spaces, Numer. Funct. Anal. Optim. 30 (2009),
inverse A(2)
168-182.
[32] H. Zhu, P. Patr´ıcio, J. Chen, Y. Zhang, The inverse along a product and its applications,
Linear Multilinear Algebra 64 (2016), 834-841.
[33] H. Zou, J. Chen, T. Li, Y. Gao, Characterizations and representations of the inverse
along an element, Bull. Malays. Math. Sci. Soc. (2016), doi: 10.1007/s40840-016-0430-3.
Enrico Boasso
E-mail address: enrico [email protected]
27
|
1807.10809 | 2 | 1807 | 2019-12-22T13:58:37 | Almost-Orthogonality of Restricted Haar-Functions | [
"math.FA"
] | We consider the Haar functions $h_I$ on dyadic intervals. We show that if $p>\frac23$ and $E\subset[0,1]$ then the set of all functions $\|h_I1_E\|_2^{-1}h_I1_E$ with $|I\cap E|\geq p|I|$ is a Riesz sequence. For $p\leq\frac23$ we provide a counterexample. | math.FA | math |
Almost-Orthogonality of Restricted
Haar-Functions
Julian Weigt
Aalto University
December 24, 2019
Abstract
We consider the Haar functions hI on dyadic intervals. We show that
2 hI 1E with
3 we provide a counterexample.
if p > 2
I ∩E ≥ pI is a Riesz sequence. For p ≤ 2
3 and E ⊂ [0, 1] then the set of all functions khI 1Ek−1
1
Introduction
In this paper we prove a stability result for perturbed Haar functions. It grew
out of the author's Master's thesis [6], written in Bonn under the supervision of
Professor Christoph Thiele. It was motivated by an idea on how to to extend
the result in [3] to three general functions.
The Haar function of the interval I = [a, b) is given by
hI = −1I l + 1I r,
where I l = [a, a+b
consider the Haar functions of the dyadic intervals
2 ) and I r = [ a+b
2 , b) are the left and right halves of I. We
The main result of this paper is the following theorem.
D =(cid:8)(cid:2)k2n, (k + 1)2n(cid:1)(cid:12)(cid:12) n, k ∈ Z(cid:9).
Theorem 1.1. For each p > 2
3 there is a constant c > 0 such that for all
measurable sets E ⊂ [0, 1) and all sequences (aI )I∈D with aI = 0 if I ∩E < pI,
we have
kaI hI 1Ek2
2,
(1)
(cid:13)(cid:13)(cid:13)XI∈D
aI hI 1E(cid:13)(cid:13)(cid:13)
2
2
≥ cXI∈D
whenever the right-hand side is finite. For p ≤ 2
c > 0.
3 there is no such constant
Remark. The proof strategy of Theorem 1.1 resembles the well known Bellman
function technique as for example in [5]. A rephrasing of the proof which re-
sembles the Bellman function technique more closely can be found in Section
2.1.5 in [6].
The proof also yields an explicit value for c. We discuss its optimality in
Section 4. Furthermore, if the right-hand side of (1) converges, then the sum
1
on the left-hand-side converges in L2 because for any finite subset D0 ⊂ D we
have
2
2
(cid:13)(cid:13)(cid:13)XI∈D0
aI hI 1E(cid:13)(cid:13)(cid:13)
≤(cid:13)(cid:13)(cid:13)XI∈D0
aI hI(cid:13)(cid:13)(cid:13)
2
2
kaI hI k2
2 ≤
= XI∈D0
1
p XI∈D0
kaI hI 1Ek2
2.
(2)
This also implies (2) with D0 = D, which means that a reverse inequality of (1)
holds as for all p > 0.
In a more general setting, a sequence V in a Hilbert space is called a Bessel
sequence if
holds, and a Riesz sequence if in addition also
(cid:13)(cid:13)(cid:13)Xv∈V
(cid:13)(cid:13)(cid:13)Xv∈V
avv(cid:13)(cid:13)(cid:13)
avv(cid:13)(cid:13)(cid:13)
2
≤ C Xv∈V
av2
2
≥ cXv∈V
av2
holds, where the constants c > 0 and C < ∞ respectively are independent of
(av)v∈V ⊂ ℓ2(V ). Inserting
for aI in (2) shows that for all p > 0, the
sequence
khI 1E k2
aI
(3)
khI1Ek2 (cid:12)(cid:12)(cid:12)(cid:12)
V =(cid:26) hI 1E
I ∈ D, I ∩ E ≥ pI(cid:27)
p . Theorem 1.1 states that if p > 2
is a Bessel sequence with constant 1
3 then (3)
is also a Riesz sequence. A weaker result already follows from the well-known
Kadison-Singer Problem, which was resolved recently by Marcus, Spielman
and Srivastava in [4].
In doing so, they also solved the numerous equivalent
problems, one of which is the Feichtinger Conjecture, which states that every
Bessel sequence can be partitioned into finitely many Riesz sequences. This
means it can already be concluded from (2) that there is a finite partition of (3)
into Riesz sequences. Building upon [4], Bownik, Casazza, Marcus and Speegle
also proved a quantitative version of the Feichtinger Conjecture in [1]. For the
specific setting of restricted Haar functions, their Corollary 6.5 in [1] reads that
if p > 3
4 then (3) can be partitioned into two Riesz sequences. Theorem 1.1
improves on that because it already applies for p > 2
3 and states that (3) is
already a Riesz sequence prior to partitioning.
For more details on the relation of this paper to other work; see Section 4.
2 Proof of the Case p >
2
3
For n ∈ N0 denote Dn = {I ∈ D I ≥ 2−n}. The idea is to first prove a
weighted inequality,
2
(cid:13)(cid:13)(cid:13) XI∈Dn
aI hI 1E(cid:13)(cid:13)(cid:13)
L2(wn)
kaI hI 1Ek2
2.
≥ XI∈Dn
The weights are introduced in order to allow a proof by induction on n. They
will be uniformly bounded from above, so that the case p > 2
3 of Theorem 1.1
follows from the weighted inequality.
2
The weights wn look as follows: Fix 2
3 < p ≤ 1. Define g : [0, 1] → R by
g(q) =(1 +
g(p) q
p
p(2−p)
(3p−2)(3p−2q)
q ≥ p,
q ≤ p.
(4)
It is well defined because 2q ≤ 2 < 3p. Now on each interval I ∈ Dn+1 \ Dn
abbreviate qI = I∩E
if qI > 0. If qI = 0
then the value of wn does not matter. The properties of g that we need are
collected in the following proposition.
I and assign wn the constant value g(qI )
qI
Proposition 2.1. Let 2
properties:
3 < p ≤ 1. Then the function g has the following
1) For all q1, q2 ∈ [0, 1] and a ∈ R with q1+q2
2 ≥ p or a = 0 we have
(1 − a)2
2
g(q1) +
(1 + a)2
2
g(q2) − g(cid:16) q1 + q2
2
(cid:17) ≥ a2.
2) There is a constant C > 0 s.t. for all q ∈ [0, 1] we have
q ≤ g(q) ≤ Cq.
(5)
(6)
Proposition 2.1 is the crucial step. After that it requires not much more
3 of Theorem 1.1. In order to prove
than bookkeeping to conclude the case p > 2
Proposition 2.1, we first show the Lemmas 2.2 and 2.3.
Define g : [0, 1] → R by
g(q) = 1 +
p(2 − p)
(3p − 2)(3p − 2q)
,
so that g = g on q ≥ p.
Lemma 2.2. Let q1, q2 ∈ [0, 1], a ∈ R. Then
(1 − a)2
2
g(q1) +
(1 + a)2
2
Proof. For i = 1, 2 take xi s.t.
g(q2) − g(cid:16) q1 + q2
2
(cid:17) ≥ a2.
g(qi) = 1 +
1
xi
.
Then xi > 0 and
g(cid:16) q1 + q2
2
(cid:17) = 1 +
Hence it suffices to confirm the positivity of
2
x1 + x2
.
(1 − a)2
2
1
2
(cid:16)1 +
a2(cid:16) 1
x1x2(cid:20) 1
x1
1
2
+
=
=
(1 + a)2
1
2
x1(cid:17) +
x2(cid:17) + a(cid:16) 1
1
x2
−
1
(cid:16)1 +
x1(cid:17) +
1
2
1
+
x2(cid:17) −(cid:16)1 +
2(cid:16) 1
2(cid:16)x1 + x2 − 4
x1 + x2(cid:17) − a2
x1 + x2(cid:17)
x1 + x2(cid:17)(cid:21)
x1
1
1
x2
x1x2
− 4
1
a2(x1 + x2) + a(x1 − x2) +
3
which is a quadratic polynomial in a. Since x1 + x2 ≥ 0 and
positive leading coefficient. The discriminant is
1
x1x2
≥ 0 it has a
(x1 + x2)(cid:16)x1 + x2 − 4
x1x2
x1 + x2(cid:17) − (x1 − x2)2
= (x1 + x2)2 − 4x1x2 − (x1 − x2)2 = 0,
and so the minimum of the polynomial is zero.
Lemma 2.3.
Proof.
g(2p − 1) = g(2p − 1).
g(2p − 1) = g(p)
2p − 1
p
2p − 1
=(cid:20)1 +
p(2 − p)
p(3p − 2)(cid:21) 2p − 1
p
= 1 +
= g(2p − 1).
p
3p − 2
2p
=
3p − 2
p
Proof of Proposition 2.1. Lemma 2.2 with a = 0 implies that g is midpoint
convex and thus convex. By Lemma 2.3 and by the definition of g(p) we have
that g(q) = g(q) at the two values q = 2p − 1, p. This means that on [0, p] the
function g describes the line that passes through these two distinct points. On
[p, 1] recall that g = g. It follows from this that also g is convex. This means that
(5) holds for a = 0 and so it suffices to consider the case q1+q2
2 ∈ [p, 1]. There we
have q1 ≥ 2p − q2 ≥ 2p − 1 and similarly q2 ≥ 2p − 1. From the considerations
(cid:1) which
implies that for all a we have
(cid:1) = g(cid:0) q1+q2
above we then get g(q1) ≥ g(q1), g(q2) ≥ g(q2), g(cid:0) q1+q2
g(q2) − g(cid:16) q1 + q2
(cid:17)
g(q2) − g(cid:16) q1 + q2
(cid:17) ≥ a2,
(1 + a)2
(1 + a)2
(1 − a)2
(1 − a)2
g(q1) +
g(q1) +
≥
2
2
2
2
2
2
2
2
where the last inequality follows from Lemma 2.2. This finishes the proof of
(5).
The upper bound in (6) holds for C = g(1) because g is convex and non-
negative and g(0) = 0. For the lower bound, recall that for q ∈ [0, p] we have
g(q) = q
p g(p) for all q ∈ [0, 1]. It follows
from the definition of g that g(p) ≥ 1 and therefore g(q) ≥ q
p g(p), so that convexity implies g(q) ≥ q
p ≥ q.
The following lemma translates Proposition 2.1 into our setting of Haar
functions on weighted L2 spaces.
Lemma 2.4. For every n ∈ N we have
2
(cid:13)(cid:13)(cid:13) XI∈Dn+1
and
aI hI 1E(cid:13)(cid:13)(cid:13)
L2(wn+1)
2
−(cid:13)(cid:13)(cid:13) XI∈Dn
aI hI 1E(cid:13)(cid:13)(cid:13)
L2(wn)
≥ XI∈Dn+1\Dn
kaI hI 1Ek2
2 (7)
The constant C is the same as in Proposition 2.1.
1 ≤ wn ≤ C.
(8)
4
Proof. Recall that on I ∈ Dn+1 \ Dn we assign wn the constant value q(qI )
qI = I∩E
I
because the integrated function in (7) is zero a.e. on such I anyways.
if qI > 0. Where I ∩ E vanishes, the value of wn does not matter
with
qI
First note that (8) is equivalent to (6). We prove (7) using mostly (5).
Partition the domain of integration on the left-hand-side of (7) into Dn+1 \ Dn
so that the inequality becomes
XJ∈Dn+1\DnZE∩J(cid:20)(cid:16) XI∈Dn+1
≥ XJ∈Dn+1\DnZE
(aJ hJ )2.
aI hI(cid:17)2
wn+1 −(cid:16) XI∈Dn
aI hI(cid:17)2
wn(cid:21)
We prove this inequality for each summand J ∈ Dn+1 \ Dn individually. So fix
J ∈ Dn+1 \ Dn. Then for each I ∈ Dn the function hI is constant on J. That
means we may write
aI hI = bJ 1J
1J XI∈Dn
for some bJ ∈ R. For I ∈ Dn+1 \ Dn the function hI is nonzero if and only if
I = J, so that
aI hI = bJ 1J + aJ hJ .
1J XI∈Dn+1
So it suffices to show
ZE∩Jh(bJ 1J + aJ hJ )2wn+1 − (bJ 1J )2wni ≥ZE
in order to prove (7). Write
(aJ hJ )2
(9)
q1 =
J l ∩ E
J l
,
q2 =
J r ∩ E
J r
,
q1 + q2
2
=
J ∩ E
J
so that we have
1J l wn+1 =
g(q1)
q1
1J l ,
1J rwn+1 =
g(q2)
q2
1J r ,
1J wn =
2
g(cid:0) q1+q2
q1+q2
2
,
(cid:1)
1J .
if the respective denominators are positive. Evaluating the integrals, (9) then
reads
(bJ − aJ )2J lg(q1) + (bJ + aJ )2J rg(q2) − b2
J Jg(cid:16) q1 + q2
2
(cid:17) ≥ J
q1 + q2
2
a2
J ,
also if q1 or q2 are zero because g(0) = 0. Then divide both sides by J. For
bJ = 0 we obtain
g(q1) + g(q2)
q1 + q2
≥
2
a2
J .
a2
J
2
This inequality holds due to the lower bound in (6). In case bJ 6= 0 we addi-
tionally divide by b2
J and obtain
(cid:16)1 −
aJ
bJ(cid:17)2 1
2
g(q1) +(cid:16)1 +
aJ
bJ(cid:17)2 1
2
g(q2) − g(cid:16) q1 + q2
2
(cid:17) ≥
q1 + q2
2
bJ(cid:17)2
(cid:16) aJ
.
5
This inequality is a consequence of (5) because q1+q2
2 ≤ 1. Note that we can also
obtain the case bJ = 0 from (5) by sending aJ → ∞, instead of from (6). That
way we would even get the stronger inequality without the factor q1+q2
.
2
Proof of Theorem 1.1 in case p > 2
consists of only one interval we get
3 . We use Lemma 2.4. Because D0 = {[0, 1)}
2
L2(w0)
kaI hI 1Ek2
2
≥ XI∈D0
(10)
from the lower bound in (8). Summing up (10) and (7) for n = 1, . . . , k − 1 we
get
2
(cid:13)(cid:13)(cid:13)XI∈D0
(cid:13)(cid:13)(cid:13)XI∈Dk
C(cid:13)(cid:13)(cid:13)XI∈Dk
aI hI 1E(cid:13)(cid:13)(cid:13)
aI hI 1E(cid:13)(cid:13)(cid:13)
aI hI 1E(cid:13)(cid:13)(cid:13)
2
2
kaI hI 1Ek2
2,
L2(wk)
≥ XI∈Dk
aI hI 1E(cid:13)(cid:13)(cid:13)
≥(cid:13)(cid:13)(cid:13)XI∈Dk
2
.
L2(wk)
and the upper bound in (8) allows us to get rid of the weight
This proves the part p > 2
C if we only consider finite
sums. For infinite sums where the right hand side of (1) converges, we may also
pass to the limit n → ∞ with the help of (2).
3 of Theorem 1.1 with c = 1
3 Proof of the Case p ≤ 2
3
Fix E = [0, 2
intervals (I2n)n=0,1,..., defined inductively by
3 ]. We build the counterexample from the sequence of dyadic
I0 = [0, 1],
r,
I2n+1 = I2n
I2n+2 = I2n+1
l.
Lemma 3.1. For all n = 0, 1, . . . we have
In = 2−n,
I2n ∩ E =
2
3
I2n,
I2n+1 ∩ E =
1
3
I2n+1.
Proof. It is clear that we have In = 2−n. For the other statements we proceed
by induction on n. For n = 0 we have
(cid:12)(cid:12)(cid:12)(cid:12)
[0, 1] ∩h0,
(cid:12)(cid:12)(cid:12)(cid:12)h 1
, 1i ∩h0,
2
2
2
3i(cid:12)(cid:12)(cid:12)(cid:12)
3i(cid:12)(cid:12)(cid:12)(cid:12)
=
=
2
3
2
3
=
−
2
3
1
2
[0, 1],
=
1
6
=
.
2
1
3(cid:12)(cid:12)(cid:12)(cid:12)h 1
, 1i(cid:12)(cid:12)(cid:12)(cid:12)
Now assume the lemma holds for n. That means that the point 2
to the right from the left boundary of I2n+1, i.e. 2
left boundary of I2(n+1). Therefore we have I2(n+1) ∩ E = 2
in turn implies that the point 2
of I2(n+1), i.e. 1
Therefore we have I2(n+1)+1 ∩ E = 1
lemma for n + 1.
3 I2n+1
3 I2(n+1) to the right from the
3 I2(n+1). That
6 I2(n+1) to the right from the midpoint
3 I2(n+1)+1 to the right from the left boundary of I2(n+1)+1.
3 I2(n+1)+1, finishing the proof of the
3 lies 1
3 lies 1
6
Further set
a0 = 1,
a2n = 2n−1,
n ≥ 1.
The following proposition proves the case p ≤ 2
3 of Theorem 1.1.
Proposition 3.2. For each n we have
n
n
Xk=0
(cid:13)(cid:13)(cid:13)
Xk=0
ka2khI2k 1Ek2
2 =
2
2
=
a2khI2k 1E(cid:13)(cid:13)(cid:13)
+
n
6
,
.
2
3
2
3
(11)
(12)
Proof. By Lemma 3.1 we have I2n ∩ E = 2
3 2−2n. Thus
ka0hI0 1Ek2
2 =
ka2nhI2n 1Ek2
2 =
2
3
2
3
,
2−2n22(n−1) =
1
6
,
n ≥ 1,
which implies (11).
In order to prove (12), first note that the support of hI2(n+1) 1E is I2n
Therefore it follows by induction on n that
r ∩ E.
a2khI2k 1E = −1[0, 1
2 ) + 2n1I2n
r∩E,
n
Xk=0
since
2n1I2n
r∩E + 2nhI2(n+1) 1E = 2n+11I2(n+1)
r∩E.
By Lemma 3.1 we have I2n
r ∩ E = 1
3 I2n
r = 1
3 2−2n−1 so that we obtain
n
(cid:13)(cid:13)(cid:13)
Xk=0
2
2
a2khI2k 1E(cid:13)(cid:13)(cid:13)
=
1
2
· (−1)2 +
1
3
2−2n−122n =
2
3
.
4 Remarks
4.1 Optimality of the Constant c
From the proof we get an explicit expression for the constant in (1)
c =
1
C
=
1
g(1)
=
(3p − 2)2
(3p − 2)2 + p(2 − p)
=
and since
p(2 − p) = (p −
2
3
)(cid:20) 4
3
− (p −
2
3
)(cid:21) +
7
2
3(cid:20) 4
3
+ 1
3(cid:1)2
(cid:0)p − 2
3(cid:1)2
(cid:0)p − 2
)(cid:21) =
8
9
2
3
− (p −
9 p(2 − p)
+ O(cid:16)p −
2
3(cid:17)
we have
c =
3(cid:1)2
(cid:0)p − 2
81 + O(p − 2
3 )
8
=
81
8 (cid:16)p −
2
3(cid:17)2
1
1 + O(p − 2
3 )
=
81
8 (cid:16)p −
2
3(cid:17)2
+ O(cid:16)p −
.
2
3(cid:17)3
However this constant c is likely not maximal because g satisfies a stronger
bound than the required g(q) ≥ q, and because we dropped a factor q1+q2
in
the deduction of (5). We only did this because sending q1 → q2 in (5) leads
to an ODE with solution g, while with the factor q1+q2
in place we could not
solve the ODE. We did however minimize C in some respect: There are multiple
solutions to the ODE from (5) such that the corresponding g satisfies (6) and
(5) with some C. Among all those, g has the smallest C for p → 2
3 . For a proof
of this and for more details; see Section 2.1 in [6]. In Section 3 in [6] we also
provide a set E for which we prove an explicit sharp constant cs which satisfies
3 )3. We conjecture that (1) holds already with this
particular constant cs. In Section 4 of [6] we prove that this is indeed the case
at least for certain E.
cs = 27(cid:0)p − 2
3 )2 + O(p − 2
2
2
4.2 Further remarks on Theorem 1.1
For p > 2
i.e. allow a[0,2) 6= 0, even though usually [0, 2) ∩ E < p[0, 2).
3 inequality (1) still holds if we add the constant function to the sums,
Furthermore, Theorem 1.1 is not a consequence of the fact that (cid:8)hI 1E (cid:12)(cid:12)
I ∩ E ≥ pI(cid:9) is only a small perturbation of the orthogonal set(cid:8)hI (cid:12)(cid:12) I ∩ E ≥
pI(cid:9), in the sense that
khI − hI 1Ek2
2 ≤ (1 − p)khIk2
2 <
khI k2
2.
1
3
In order to see this, consider the following example. Assume that u1, . . . , un are
orthonormal. Abbreviate u = u1 + . . . + un and for i = 1, . . . , n set
u′
i = ui −
1
n
u.
kui − u′
ik2 =
1
n2 kuk2 =
1
n2
kuik2 =
1
n
.
n
Xi=1
Then
but
ku′
1 + . . . + u′
nk2 = ku − uk2 = 0 6& ku′
1k2 + . . . + ku′
nk2.
4.3 Related Topics
The starting point of this work was the following question, because its answer
could provide ideas on how to to extend the result in [3] to three general func-
tions.
Question 1. Let D be the set of dyadic intervals of [0, 1). Let [0, 1) = E0 ∪ E1
be a partition.
Is there a partition D = D0 ∪ D1 such that for i = 0, 1 the
equivalence
is true?
(cid:13)(cid:13)(cid:13)XI∈Di
aI hI 1Ei(cid:13)(cid:13)(cid:13)
2
2
kaI hI 1Eik2
2
∼ XI∈Di
(13)
8
We started investigating this question in Section 5 in [6]. An initial approach
to Question 1 could be to construct a partition by a majority decision: For
i = 0, 1 take Di s.t. for all I ∈ Di we have
I ∩ Ei ≥
1
2
I.
(14)
However by the counterexample of Theorem 1.1 for p = 2
2 , this strategy does
not result in the lower bound in (13). Although by (2) the majority decision
(14) at least leads to the upper bound in (13). Another idea was to use the
Feichtinger-Conjecture which was recently resolved by Marcus, Spielman and
Srivastava in [4]. Based on [4], Bownik, Casazza, Marcus and Speegle proved a
quantified version of the Feichtinger Conjecture in [1]. The following theorem
is a reformulation of Corollary 6.5 in [1].
3 ≥ 1
Theorem 4.1. Let C < 4
in a Hilbert space such that for all (av)v∈V ⊂ R
3 and c = C
2 −p2(C − 1)(2 − C). Let V be a sequence
≤ C Xv∈V
kavvk2.
2
Then there is a partition V = V0 ∪ V1 such that for i = 0, 1 we have
(cid:13)(cid:13)(cid:13)Xv∈V
(cid:13)(cid:13)(cid:13)Xv∈Vi
avv(cid:13)(cid:13)(cid:13)
avv(cid:13)(cid:13)(cid:13)
2
≥ c Xv∈Vi
kavvk2.
Unfortunately it is not clear if Theorem 4.1 can be used to answer Question
1. The closest consequence of Theorem 4.1 in that direction that we found is
the following corollary.
Corollary 4.2. Let p > 3
E = E0 ∪ E1 be a partition and for i = 0, 1 set
4 and c = 1
2p −q2( 1
p − 1)(2 − 1
p ). Let E ⊂ [0, 1) and
Then H0 ∪ H1 can be partitioned into G0 ∪ G1 where for i = 0, 1 we have
Hi =(cid:8)hI 1Ei (cid:12)(cid:12) I ∈ D, I ∩ Ei ≥ pI(cid:9).
(cid:13)(cid:13)(cid:13)Xv∈Gi
avv(cid:13)(cid:13)(cid:13)
2
≥ c Xv∈Gi
kavvk2.
(15)
Theorem 1.1 can be seen as an improvement of Corollary 4.2. That is because
by Theorem 1.1 the two sequences H0 and H1 already satisfy (15) with some
other constant c > 0, and since H0 and H1 are orthogonal to one another, also
their union satisfies (15), even without partitioning. And this already holds for
p > 2
3 .
References
[1] Marcin Bownik, Pete Casazza, Adam W Marcus, and Darrin Speegle. Im-
proved bounds in weaver and feichtinger conjectures. Journal fur die reine
und angewandte Mathematik (Crelles Journal), 2016.
9
[2] Ole Christensen. An introduction to frames and Riesz bases, volume 7.
Springer, 2003.
[3] Vjekoslav Kova, Christoph Thiele, and Pavel Zorin-Kranich. Dyadic trian-
gular hilbert transform of two general functions and one not too general
function. Forum of Mathematics, Sigma, 3:e25, 2015.
[4] Adam W Marcus, Daniel A Spielman, and Nikhil Srivastava.
Interlacing
families ii: Mixed characteristic polynomials and the kadison-singer problem.
Annals of Mathematics, pages 327 -- 350, 2015.
[5] Fedor Nazarov, Sergei Treil, and Alexander Volberg. The bellman functions
and two-weight inequalities for haar multipliers. Journal of the American
Mathematical Society, 12(4):909 -- 928, 1999.
[6] Julian Weigt.
restricted haar
functions.
Master's thesis, Rheinische Friedrich-Wilhelms-Universitat Bonn, 2018.
http://math.aalto.fi/~weigtj1/Thesis_updated.pdf.
Almost-orthogonality of
10
|
1506.05383 | 1 | 1506 | 2015-06-17T16:41:13 | Some trace monotonicity properties and applications | [
"math.FA",
"math-ph",
"math-ph"
] | We present some results on the monotonicity of some traces involving functions of self-adjoint operators with respect to the natural ordering of their associated quadratic forms. We also apply these results to complete a proof of the Wegner estimate for continuum models of random Schr\"odinger operators as given in \cite{co-hi1}. | math.FA | math |
SOME TRACE MONOTONICITY PROPERTIES AND
APPLICATIONS
JEAN-MICHEL COMBES AND PETER D. HISLOP
Abstract. We present some results on the monotonicity of some traces
involving functions of self-adjoint operators with respect to the natural or-
dering of their associated quadratic forms. We also apply these results to
complete a proof of the Wegner estimate for continuum models of random
Schrodinger operators as given in [4].
1. Statement of the Problem and Result
We consider two lower-semibounded self-adjoint operators A and B asso-
ciated with closed symmetric, lower-semibounded quadratic forms qA and qB
with form domains Q(A) and Q(B), respectively. We suppose that qA 6 qB.
This inequality means that Q(B) ⊂ Q(A) and that for all ϕ ∈ Q(B), we have
qA(ϕ) 6 qB(ϕ). Under these conditions, Kato proved [8, Theorem 2.21, chapter
VI] the following relationship between the resolvents of the two operators A and
B. For all a > − inf σ(A), we have
(B + a)−1 6 (A + a)−1.
(1)
This resolvent inequality may be used to derive several interesting relations
between the traces of functions of A and B under some additional assumptions.
We will prove that if f > 0 and g is a member of a class of functions L described
in Definition 2, then
Tr(f (B)g(B)) 6 Tr(f (B)g(A)),
see Theorem 5. We compare these inequalities with Lowner’s Theorem (see
section 3) on operator monotone functions. We also use these results to com-
plete a proof of Wegner’s estimate for random Schrodinger operators given in
[4]. These results rely on the following technical theorem.
Theorem 1. Suppose that A and B are two lower semibounded self-adjoint
operators with quadratic forms qA and qB and form domains Q(A) and Q(B).
Suppose that A and B are relatively form bounded in that
(1) the form domains satisfy Q(B) ⊂ Q(A), and
(2) for all ψ ∈ Q(B), we have qA(ψ) 6 qB(ψ).
Let PB project onto a B-invariant subspace so that for some m ∈ N, and for all
a > − inf σ(A), the operator PB(B + a)−m is in the trace class. Then we have
PDH was partially supported by NSF through grant DMS-1103104 and the Universit´e de
Toulon while some of this work was done.
2
J.-M. COMBES AND P. D. HISLOP
(1) For all n ∈ N large enough so that m < 2n and for all a > − inf σ(A),
Tr(PB(B + a)−2n
) 6 Tr(PB(A + a)−2n
);
(2) For any t > 0,
Tr(PBe−tB) 6 Tr(PBe−tA).
Proof. 1. The result of Kato [8, Theorem 2.21, chapter VI], stated above,
implies that (B + a)−1 6 (A + a)−1. We first suppose that PB is a non-zero
rank one projection PB = Pλ, so that BPλ = λPλ. From (1), it follows that for
a > − inf σ(A), we have
TrPλ = (λ + a) Tr(Pλ(B + a)−1)
6 (λ + a) Tr(Pλ(A + a)−1)
6 (λ + a)kPλk2kPλ(A + a)−1k2.
Since Pλ is a rank one projector, we have
kPλk2 = kPλk1 = TrPλ = 1,
and
kPλ(A + a)−1k2 = (TrPλ(A + a)−2)1/2.
Upon squaring inequality (2) and using the results (3)–(4), we obtain
TrPλ 6 (λ + a)2kPλ(A + a)−1k2
2
= (λ + a)2 Tr(Pλ(A + a)−2).
(2)
(3)
(4)
(5)
We continue by rewriting the trace on the right in (5) using the Hilbert-Schmidt
norm. We square the resulting inequality, use (3)–(4), and obtain
TrPλ 6 (λ + a)22
Tr(Pλ(A + a)−22
).
Continuing in this way, we obtain for any n ∈ N:
TrPλ 6 (λ + a)2n
Tr(Pλ(A + a)−2n
).
This may also be written as:
Tr(Pλ(B + a)−2n
) 6 Tr(Pλ(A + a)−2n
).
(6)
(7)
(8)
2. We now assume that PB is a projection operator diagonalizing B so that
PB =Pj Pλj with BPλj = λjPλj . If we take 2n > m, we can sum the inequal-
ities (8) over j to obtain
Tr(PB(B + a)−2n
) 6 Tr(PB(A + a)−2n
).
(9)
3. For the exponential bound, we first note that by assumption PB(B + a)−m
is trace class for some integer m > 0, so that PBe−tB is also trace class since
(B + a)me−tB is bounded for t > 0. For t > 0 and b ∈ R so that b > − inf σ(A),
MONOTONE TRACE PROPERTIES
3
we obtain
TrPBe−t(A+b) = lim
n→∞
> lim
n→∞
Tr"PB(cid:18)1 +
Tr"PB(cid:18)1 +
= TrPBe−t(B+b),
t(A + b)
2n (cid:19)−2n#
2n (cid:19)−2n#
t(B + b)
where we used (9) on the second line. It follows that
Tr(PB(I)e−t(B+b)) 6 Tr(PBe−t(A+b)).
This proves the second claim.
(10)
(11)
(cid:3)
2. An application to trace inequalities
Theorem 1 may be applied to a large class of functions of self-adjoint oper-
ators in order to obtain some inequalities relating traces of functions of self-
adjoint operators.
Definition 2. A real-valued function g is in the class L if there is a nonnegative
σ-finite Borel measure ρ supported on [0, ∞) so that for s > 0
g(s) =Z ∞
0
e−stdρ(t).
(12)
Theorem 3. Let self-adjoint operators A, B and projector PB be as in Theorem
1. Then for any g ∈ L such that PBg(B) is trace class, one has
TrPBg(B) 6 TrPBg(A).
(13)
Proof. By the representation of g in (12) and the inequality (10) with b = 0,
one has
TrPBg(B) = Z ∞
6 Z ∞
0
0
Tr(PBe−tB) dρ(t)
Tr(PBe−tA) dρ(t)
= Tr(PBg(A)).
(14)
(cid:3)
A particularly useful example of functions g are those related to powers of
the resolvent of a self-adjoint operator.
Corollary 4. Let self-adjoint operators A, B and projector PB be as in Theorem
1, and let a > − inf σ(A). For any β > m, where m is defined in Theorem 1,
we have
Tr(PB(B + a)−β) 6 Tr(PB(A + a)−β).
(15)
Proof. We use the Laplace transform formula valid for α > −1 and Re z > 0:
1
z1+α =
1
Γ(1 + α)Z ∞
0
e−zttα dt.
(16)
4
J.-M. COMBES AND P. D. HISLOP
This shows that the function g(s) = s−β is in the class L for any β > 0. The
result (15) follows from Theorem 3.
(cid:3)
We conclude this section with the following generalisation of Theorem 3. It
presents a trace comparison theorem for the class L of functions of self-adjoint
operators.
Theorem 5. Let A and B be two lower semibounded self-adjoint operators
satisfying the hypotheses of Theorem 1. Suppose g ∈ L and f > 0 so that
f (B)g(B) is trace class. We then have
Tr(f (B)g(B)) 6 Tr(f (B)g(A)).
(17)
Proof. Since f (B)g(B) is assumed to be trace class, the operator B must have
pure point spectrum {λj} on the support of the function f g. For any j, it
follows from Theorem 3 that
Tr(Pλj g(B)) 6 Tr(Pλj g(A)),
(18)
where, as above, BPλj = λjPλj . Multiplying both sides of (18) by f (λj) > 0,
and summing over j results in (17).
(cid:3)
We remark that if g(B) is in the trace class, we may take f = 1 and obtain
Tr(g(B)) 6 Tr(g(A)),
(19)
a result that also follows from the Min-Max Theorem since any function g ∈ L
is decreasing.
3. A relation with operator monotone functions
The following class of functions was introduced by Lowner [9] in 1934 and is
the object of his famous theorem that we now recall.
Definition 6. Let J ⊂ R be a finite interval or a half-line. A function g : J →
R is operator monotone increasing (respectively, decreasing) in J if for all pairs
of self-adjoint operators A, B with spectrum in J the operator inequality A 6 B
implies the operator inequality g(A) 6 g(B) (respectively g(B) 6 g(A).)
If g is operator monotone decreasing, then (17) holds for any f > 0 and for
all pairs of operators A 6 B such that f (B)g(B) is trace class. Because of this,
we study the relationship between operator monotone decreasing functions and
the class L of Definition 2. For simplicity, we assume that J = R+. We denote
by I (respectively, D) the class of operator monotone increasing (respectively,
decreasing) functions on R+. The map g ∈ D → g ∈ I defined by g(s) :=
g(1/s), for s > 0, is a bijective involution between D and I.
Lowner’s Theorem [9] (see also [1, 2, 3, 6, 7]) states that g is operator mono-
tone increasing if and only if g has an analytic extension to the upper-half
complex plane with a positive imaginary part. Such functions are known as
Herglotz or Pick functions and have integral representations. For example,
Hansen [7] proved the following representation.
MONOTONE TRACE PROPERTIES
5
Lemma 7. [7, Corollary 5.1] Let g be a positive operator monotone increasing
function on the half-line R+. Then there exists a bounded, positive measure µ
on R+ such that
g(s) =ZR+
s(1 + λ)
s + λ
dµ(λ), s > 0.
(20)
It follows easily from Kato’s result (1) that any function on R+ with a rep-
resentation as on the right of (20) is in the class I. The difficult part of the
proof of Lowner’s Theorem is the converse.
Using the bijection g → g between D and I described above, it follows that
if f ∈ D, then f has a representation of the form
f (s) =ZR+
1 + λ
1 + sλ
dµ(λ), s > 0,
(21)
for some bounded positive measure µ on R+. Using the Laplace transform
representation (14) with α = 0, we may write f as
where h is defined by
f (s) =ZR+
h(t) =ZR+(cid:18)1 +
e−sth(t) dt,
1
λ(cid:19) e− t
λ dµ(λ).
(22)
(23)
The function h ∈ L1
loc(R+) so by Definition 2, the function f ∈ L.
This shows that D ⊂ L. On the other hand, the functions on R+ such as
f (s) = e−as, with a > 0, or f (s) = (s + a)−ρ, with ρ > 1 and a > 0, belong to
the class L but not to the class D. As a consequence, we obtain the following
theorem.
Theorem 8. The class of operator monotone decreasing functions D is strictly
contained in the class of functions L.
4. An application to the proof of Wegner’s estimate
We complete the proof of the Wegner estimate given in [4]. Since this method
of proof seems to have been used in several subsequent papers, we wanted to
present the complete argument. The Wegner estimate is an upper bound on the
probability that a local Hamiltonian has eigenvalues in a given energy interval.
We considered a large cube Λ centered at the origin in Rd with odd integer
side length. We let Hω := −∆ + Vω be the random Schrodinger operator on
L2(Rd) with a standard Anderson-type random potential Vω > 0 (this condition
can be removed). We denote by HΛ the restriction of Hω to Λ with Dirichlet
boundary conditions. This operator has discrete spectrum. For any bounded
interval I = [I−, I+] ⊂ R, we let EΛ(I) be the spectral projection for HΛ and
interval I. The trace of this projection is finite and it is a random variable.
The Wegner estimate proved in [4, Proposition 4.5] is
P{TrEΛ(I) > 1} 6 CW IΛ,
(24)
where CW > 0 is a finite constant depending on I+.
6
J.-M. COMBES AND P. D. HISLOP
The proof of the Wegner estimate in [4] depends on a comparison of the
operator HΛ to a direct sum of operators defined on unit cubes in Λ. Let
Λ = Int{∪jΛ1(j)} be a decomposition of Λ into unit cubes centered at the
lattice points Λ of Λ. In the proof of Proposition 4.5 [4, section 4], we used the
operator inequality
HΛ > HN,Λ ≡ − ⊕j ∆N,j,
(25)
where −∆N,j is the Neumann Laplacian on a unit cube centered at j ∈ Λ∩ Zd, if
the boundary of the cube does not intersect the boundary of Λ, or the Laplacian
with mixed Neumann-Dirichlet boundary conditions if the cube’s boundary
intersects the boundary of Λ. This inequality is valid only in the operator form
sense. It cannot be used in conjunction with Jensen’s inequality as done after
equation (4.15) in [4] since the eigenfunctions φn of HΛ are not in the operator
domain of HN,Λ.
We apply Theorem 1 in order to complete the proof of Wegner’s estimate as
stated in [4, Proposition 4.5]. We divide the set of indices Λ of the unit cubes
in Λ into two sets: The set ∂ Λ associated with unit cubes whose boundary
intersects ∂Λ, and the set IntΛ of interior points. We take A = HN,Λ, as
defined in (25), and B = HΛ, the restriction of H to Λ with Dirichlet boundary
conditions.
We verify conditions (1) and (2) of Theorem 1. As quadratic forms, we have
0 (Λ), whereas Q(A) := Q(HN,Λ) = {⊕j∈IntΛH 1(Λ1(j))} ⊕
M (Λ1(j)) consists of functions in H 1(Λ1(j)) with
It follows that Q(HΛ) ⊂
Q(B) := Q(HΛ) = H 1
{⊕j∈∂ ΛH 1
Neumann boundary conditions along ∂Λ ∩ ∂Λ1(j).
Q(HN,Λ). The second condition of Theorem 1 holds identically.
M (Λ1(j))}, where H 1
We have inf σ(A) = 0 in this case. Then, with the notation of [4], the
projection PB is EΛ(Iη). From part 2 of Theorem 1, we have
TrEΛ(Iη) 6 eIη,+Tr(EΛ(Iη)e−HΛ)
6 eIη,+Tr(EΛ(Iη)e−HN,Λ)
= eIη,+
(26)
Xj∈Λ∩Zd
Tr(EΛ(Iη)e∆N,j χj)
,
where χj is the characteristic function for the unit cube Λ1(j) centered at j ∈ Zd.
In this way, we recover (4.16) of [4]. Following the remainder of the proof there,
since the operators −∆N,j do not depend on the random variables, we expand
the trace in the eigenfunctions of −∆N,j and apply the spectral averaging result
[4, Corollary 4.2]. In this manner, one obtains (24). We refer the reader to [5]
for a more general proof of the Wegner estimate.
References
[1] J. Bendat, S. Sherman, Monotone and convex operator functions, Trans. American
Math. Soc. 79 (1955) 58–71.
[2] R. Bhatia, K. Sinha, Variation of real powers of positive operators, Indiana U.
Math. J. 43 (1994) 913–925.
[3] P. Chansangiam, Operator monotone functions: Characterizations and integral
representations, arXiv:1305.2471v1.
MONOTONE TRACE PROPERTIES
7
[4] J.-M. Combes, P. D. Hislop, Localization for some continuous random Hamilto-
nians in d-dimensions, J. Funct. Anal. 124 (1994) 149–180.
[5] J.-M. Combes, P. D. Hislop, F. Klopp, An optimal Wegner estimate and its ap-
plication to the global continuity of the integrated density of states for random
Schrodinger operators, Duke Math. J. 140 (2007), 469–498.
[6] W. Donoghue, Monotone matrix functions and analytic continuation, Berlin:
Springer-Verlag, 1974.
[7] F. Hansen, The fast track to Lowner’s Theorem, Linear algebra and applications
438, No. 11 (2013) 4557–4571.
[8] T. Kato, Perturbation theory for linear operators, Berlin: Springer-Verlag, 1995.
[9] C. Lowner, Uber monotone matrix funktionen, Math. Z. 38 (1934) 177-216.
(J.-M. Combes) CPT CNRS, Luminy Case 907, F-13288 Marseille C´edex 9, France
E-mail address: [email protected]
(P. D. Hislop) Department of Mathematics, University of Kentucky, Lexing-
ton, Kentucky 40506-0027, USA
E-mail address: [email protected]
|
1310.2262 | 1 | 1310 | 2013-10-08T20:12:44 | A characterization of Hardy spaces associated with certain Schr\"odinger operators | [
"math.FA"
] | Let $\{K_t\}_{t>0}$ be the semigroup of linear operators generated by a Schr\"odinger operator $-L=\Delta - V(x)$ on $\mathbb R^d$, $d\geq 3$, where $V(x)\geq 0$ satisfies $\Delta^{-1} V\in L^\infty$. We say that an $L^1$-function $f$ belongs to the Hardy space $H^1_L$ if the maximal function $\mathcal M_L f(x) = \sup_{t>0} |K_tf(x)|$ belongs to $L^1(\mathbb R^d) $. We prove that the operator $(-\Delta)^{1\slash 2} L^{-1\slash 2}$ is an isomorphism of the space $H^1_L$ with the classical Hardy space $H^1(\mathbb R^d)$ whose inverse is $L^{1\slash 2} (-\Delta)^{-1\slash 2}$. As a corollary we obtain that the space $H^1_L$ is characterized by the Riesz transforms $R_j=\frac{\partial}{\partial x_j}L^{-1\slash 2}$. | math.FA | math |
A CHARACTERIZATION OF HARDY SPACES ASSOCIATED WITH
CERTAIN SCHR ODINGER OPERATORS
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
Abstract. Let {Kt}t>0 be the semigroup of linear operators generated by a Schro-
dinger operator −L = ∆ − V (x) on Rd, d ≥ 3, where V (x) ≥ 0 satisfies ∆−1V ∈ L∞.
We say that an L1-function f belongs to the Hardy space H 1
L if the maximal func-
tion MLf (x) = supt>0 Ktf (x) belongs to L1(Rd). We prove that the operator
(−∆)1/2L−1/2 is an isomorphism of the space H 1
L with the classical Hardy space
H 1(Rd) whose inverse is L1/2(−∆)−1/2. As a corollary we obtain that the space
L is characterized by the Riesz transforms Rj = ∂
H 1
∂xj
L−1/2.
1. Introduction and statement of the result
Let Kt(x, y) be the integral kernels of the semigroup {Kt}t>0 of linear operators on
Rd, d ≥ 3, generated by a Schrodinger operator −L = ∆ − V (x), where V (x) is a
non-negative locally integrable function which satisfies
(1.1)
∆−1V (x) = −cdZRd
1
x − yd−2 V (y) dy ∈ L∞(Rd).
Since V (x) is non-negative, the Fenman-Kac formula implies that
0 ≤ Kt(x, y) ≤ (4πt)−d/2e−x−y2/4t =: Pt(x − y).
(1.2)
It is known, see [14], that for V (x) ≥ 0 the condition (1.1) is equivalent to the lower
Gaussian bounds for Kt(x, y), that is, there are c, C > 0 such that
ct−d/2e−Cx−y2/t ≤ Kt(x, y).
(1.3)
We say that an L1-function f belongs to the Hardy space H 1
MLf (x) = supt>0 Ktf (x) belongs to L1(Rd). Then we set
kfkH 1
L
= kMLfkL1(Rd).
L if the maximal function
The Hardy spaces H 1
L associated with Schrodinger operators with nonnegative poten-
tials satisfying (1.1) were studied in [10]. It was proved that the map f (x) 7→ w(x)f (x)
is an isomorphism of H 1
L onto the classical Hardy space H 1(Rd), where
(1.4)
w(x) = lim
t→∞Z Kt(x, y) dy,
2000 Mathematics Subject Classification. 42B30, 35J10 (primary); 42B35 (secondary).
Key words and phrases. Hardy spaces, Schrodinger operators.
The research was supported by Polish funds for sciences, grants: DEC-2012/05/B/ST1/00672 and
DEC-2012/05/B/ST1/00692 from Narodowe Centrum Nauki.
1
2
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
which in particular means that
(1.5)
kf wkH 1(Rd) ∼ kfkH 1
L
,
see [10, Theorem 1.1]. The function w(x) is L-harmonic, that is, Ktw = w, and satisfies
0 < δ ≤ w(x) ≤ 1.
Let us remark that the classical real Hardy space H 1(Rd) can be thought as the space
H 1
L associated with the classical heat semigroup et∆, that is, L = −∆ + V with V ≡ 0 in
this case. Obviously, the constant functions are the only bounded harmonic functions
for ∆.
The present paper is a continuation of [10]. Our goal is to study the mappings
L1/2(−∆)−1/2
and (−∆)1/2L−1/2
which turn out to be bounded on L1(Rd) (see Lemma 2.6). Our main result is the
following theorem, which states another characterization of H 1
L.
Theorem 1.6. Assume that L = −∆ + V (x) is a Schrodinger operator on Rd, d ≥ 3,
with a locally integrable non-negative potential V (x) satisfying (1.1). Then the mapping
f 7→ (−∆)1/2L−1/2f is an isomorphism of H 1
L onto the classical Hardy space H 1(Rd),
that is, there is a constant C > 0 such that
(1.7)
(1.8)
k(−∆)1/2L−1/2fkH 1(Rd) ≤ CkfkH 1
L
,
kL1/2(−∆)−1/2fkH 1
L ≤ CkfkH 1(Rd).
As a corollary we immediately obtain the following Riesz transform characterization
of H 1
L.
Corollary 1.9. Under the assumptions of Theorem 1.6 an L1-function f belongs to the
space H 1
L−1/2f belong to L1(Rd) for j = 1, 2, ..., d. Moreover,
there is a constant C > 0 such that
L if and only if Rjf = ∂
∂xj
(1.10)
C −1kfkH 1
L ≤ kfkL1(Rd) +
dXj=1
kRjfkL1(Rd) ≤ CkfkH 1
L
.
Example 1. It is not hard to see that if for a function V (x) ≥ 0 defined on Rd, d ≥ 3,
there is ε > 0 such that V ∈ Ld/2−ε(Rd) ∩ Ld/2+ε(Rd), then V satisfies (1.1).
Example 2. Assume that (1.1) holds for a function V : Rd → [0,∞), d ≥ 3. Then
V (x1, x2) := V (x1) defined on Rd × Rn, n ≥ 1, fulfils (1.1).
The reader interested in other results concerning Hardy spaces associated with semi-
groups of linear operators, and in particular semigroups generated by Schrodinger op-
erators, is referred to [1], [2], [4], [5], [6], [7], [8], [12].
HARDY SPACES FOR SCHR ODINGER OPERATORS
3
2. Boundedness on L1
We define the operators:
0
(−∆)−1f (x) =Z ∞
L−1f (x) =Z ∞
(−∆)−1/2f = c1Z ∞
L−1/2f = c1Z ∞
0
0
Ptf (x) dt = cdZ
Ktf (x) dt =:Z Γ(x, y)f (y) dy,
f (y)
x − yd−2 dy =:Z Γ0(x − y)f (y) dy,
x − yd−1 f (y) dy =:Z eΓ0(x − y)f (y) dy,
1
dZ
= c′
dt
√t
Ptf
=:Z eΓ(x, y)f (y) dy,
0
Ktf
(2.1)
dt
√t
where c1 = Γ(1/2)−1. Clearly,
0 ≤ Γ(x, y) ≤ c′
dx − y−d+1,
The perturbation formula asserts that
Pt(x − y) = Kt(x, y) +Z t
= Kt(x, y) +Z t
(2.2)
0 < Γ(x, y) ≤ cdx − y−d+2.
0 Z Pt−s(x − z)V (z)Ks(z, y) dz ds
0 Z Kt−s(x, z)V (z)Ps(z − y) dz ds.
Multiplying the second inequality in (2.2) by w(x) and integrating with respect to dx
we get
(2.3)
Z Pt(x − y)w(x) dx = w(y) +ZRdZ t
0
w(z)V (z)Ps(z, y) ds dx,
since w is L-harmonic. The left-hand side of (2.3) tends to a harmonic function, which
is bounded from below by δ and above by 1, as t tends to infinity. Thus there is a
constant 0 < cw ≤ 1 such that
(2.4)
cw = w(y) +ZRd
w(z)V (z)Γ0(z − y) dz.
Similarly, integrating the first equation in (2.2) with respect to x and taking limit as t
tends to infinity, we get
(2.5)
1 = w(y) +ZRd
V (z)Γ(z, y) dz.
For a reasonable function f the following operators are well defined in the sense of
distributions:
(−∆)1/2f = c2Z ∞
L1/2 = c2Z ∞
0
0
(Ptf − f )
dt
t3/2 .
(Ktf − f )
dt
t3/2 , c2 = Γ(−1/2)−1,
4
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
Lemma 2.6. There is a constant C > 0 such that
(2.7)
(2.8)
k(−∆)1/2L−1/2fkL1 ≤ CkfkL1,
kL1/2(−∆)−1/2fkL1 ≤ CkfkL1.
Proof. From the perturbation formula (2.2) we get
(2.9)
0
(−∆)1/2L−1/2f (x) = c2Z ∞
= c2Z ∞
(Pt − Kt)L−1/2f (x)
(Kt − I)L−1/2f (x)
= c2Z ∞
0 Z t
0 ZZ Pt−s(x − z)V (z)Ks(z, y)L−1/2f (y) dy dz ds
(Pt − I)L−1/2f (x)
t3/2 + c2Z ∞
dt
t3/2
dt
0
0
dt
t3/2
dt
t3/2 + f (x).
Consider the integral kernel W (x, u) of the operator
f 7→Z ∞
0 Z t
0 ZZ Pt−s(x − z)V (z)Ks(z, y)L−1/2f (y) dy dz ds
dt
t3/2 ,
that is,
dt
t3/2 .
W (x, u) =Z ∞
0 Z t
0 ZZ Pt−s(x − z)V (z)Ks(z, y)Γ(y, u) dy dz ds
Clearly 0 ≤ W (x, u). Integration of W (x, u) with respect to dx leads to
Z W (x, u) dx =Z ∞
0 Z t
0 ZZ V (z)Ks(z, y)Γ(y, u) dy dz ds
dt
t3/2
= 2Z ∞
0 ZZ V (z)Ks(z, y)eΓ(y, u) dy dz
ds
√s
1 ZZ V (z)eΓ(z, y)eΓ(y, u) dy dz
≤ 2c−1
1 Z V (z)Γ(z, u)dz.
= 2c−1
(2.10)
Using (2.1) we see that R W (x, u) dx ≤ 2c−1
(2.7). The proof of (2.8) goes in the same way. We skip the details.
1 k∆−1V kL∞, which completes the proof of
(cid:3)
We finish this section by proving the following two lemmas, which will be used in the
sequel.
Lemma 2.11. Assume that f ∈ L1(Rd). Then
(2.12)
Z (−∆)1/2L−1/2f (x) dx =Z f (x)w(x) dx.
HARDY SPACES FOR SCHR ODINGER OPERATORS
5
Proof. From (2.9) and (2.10) we conclude that
Z (−∆)1/2L−1/2f (x) dx = c2Z Z W (x, u)f (u) dudx +Z f (x) dx
= 2c2c−1
1 Z V (z)Γ(z, u)f (u) dz du + Z f (x) dx
=Z (w(u) − 1)f (u) du +Z f (x) dx,
where in the last equality we have used (2.5).
(cid:3)
Lemma 2.13. Assume that f ∈ L1(Rd). Then
(2.14)
Z (L1/2(−∆)−1/2f )(x)w(x) dx = cwZ f (x) dx.
Proof. The proof is similar to that of Lemma 2.11. Indeed, by the perturbation formula
(2.2) we have
0
Z (L1/2(−∆)−1/2f )(x)w(x) dx
= c2Z Z ∞
(Kt − Pt)((−∆)−1/2)f )(x)
+ c2Z Z ∞
(Pt − I)((−∆)−1/2)f )(x)
= −c2Z Z ∞
0 Z t
0 ZZ w(x)Kt−s(x, z)V (z)
× Ps(z − y)((−∆)− 1
0
dt
t3/2 w(x) dx
dt
t3/2 w(x) dx
+Z w(x)f (x) dx
= −c2Z ∞
0 Z t
0 ZRdZRd
+Z w(x)f (x) dx,
2 f )(y) dydz ds
dt
t3/2 dx
w(z)V (z)Ps(z − y)((−∆)−1/2f )(y) dydz ds
dt
t3/2
6
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
where in the last equality we have used that w is L-harmonic. Integrating with respect
to dt and then with respect to ds yields
2c2
c1 Z Z w(z)V (z)eΓ0(z − y)((−∆)−1/2f )(y) dy dz +Z f (x)w(x) dx
Z (L1/2(−∆)−1/2f )(x)w(x) dx
= −
=Z w(z)V (z)Γ0(z − u)f (u) du dz +Z f (x)w(x) dx
=Z cwf (x) dx −Z w(y)f (y) dy +Z f (x)w(x) dx,
where in the last equality we have used (2.4).
(cid:3)
3. Atoms and molecules
Fix 1 < q ≤ ∞. We say that a function a is an (1, q, w)-atom if there is a ball
q −1, R a(x)w(x) dx = 0. The atomic
λjo,
B ⊂ Rd such that supp a ⊂ B, kakLq(Rd) ≤ B
norm kfkH 1at,q,w is defined by
= infn ∞Xj=1
kfkH 1
(3.1)
at,q,w
1
are (1, q, w)-atoms.
for the Hardy space H 1(Rd), which can be thought as H 1
where the infimum is taken over all representations f = P∞
j=1 λjaj, where λj ∈ C, aj
Clearly, if w0(x) ≡ 1, then the (1, q, w0)-atoms coincide with the classical (1, q)-atoms
As a direct consequence of Theorem 1.1 of [10] (see (1.5)) and the results about atomic
decompositions of the classical real Hardy spaces (see, e.g., [3], [13], [15]), we obtain
that the space H 1
L admits atomic decomposition into (1, q, w)-atoms, that is, there is a
constant Cq > 0 such that
−∆.
(3.2)
C −1
q kfkH 1
at,q,w ≤ kfkH 1
L ≤ CqkfkH 1
at,q,w
.
Let ε > 0, 1 < q < ∞. We say that a function b is a (1, q, ε, w)-molecule associated
with a ball B = B(x0, r) if
(cid:16)ZB b(x)q dx(cid:17) 1
q
(3.3)
and
(3.4)
≤ B
1
q
q −1, (cid:16)Z2kB\2k−1B b(x)q dx(cid:17) 1
Z b(x)w(x) dx = 0.
≤ 2kB
1
q −12−εk
Obviously every (1, q, w)-atom is a (1, q, ε, w)-molecule. It is also not hard to see that
for fixed q > 1 and ε > 0 there is a constant C > 0 such that every (1, q, ε, w) molecule
HARDY SPACES FOR SCHR ODINGER OPERATORS
7
b can be decomposed into a sum
∞Xn=1
where λn ∈ C, an are (1, q, w)-atoms.
The following lemma is easy to prove.
b(x) =
λnan,
∞Xn=1
λn ≤ C,
Lemma 3.5. Let 1 < q < ∞, δ, ε > 0 be such that δ > d(1 − 1
constant C > 0 such that if b(x) satisfies (3.4) and
q ) + ε. Then there is a
(3.6)
≤
then b is a (1, q, ε, w)-molecule associated with B(y0, r).
(cid:16)Z (cid:12)(cid:12)(cid:12)b(x)(cid:16)1 + x − y0
dx(cid:17)1/q
(cid:17)δ(cid:12)(cid:12)(cid:12)
r
q
r−d+d/q
C
,
In order to prove Theorem 1.6 we shall use general results about Hardy spaces as-
sociated with Schrodinger operators with non-negative potentials which were proved in
[9]. Let {Tt}t>0 be a semigroup of linear operators generated by a Schrodinger operator
−L = ∆ − V(x) on Rd, where V(x) is a non-negative locally integrable potential. The
Hardy space H 1
L is define by means of the maximal function, that is,
L = {f ∈ L1(Rd) : kfkH 1
H 1
t>0 Ttf (x)kL1(Rd) < ∞}.
:= k sup
L
We say that a function a is a generalized (1,∞,L)-atom for the Hardy space H 1
is a ball B = B(y0, r) and a function b such that
kbkL∞ ≤ B−1,
a = (I − Tr2)b.
supp b ⊂ B,
L if there
L ∼ kfkH 1
Then we say that a is associated with the ball B(y0, r). It was proved in Section 6 of [9]
that the space H 1
L admits atomic decomposition with the generalized (1,∞,L)-atoms,
is defined as in (3.1) with aj(x)
that is, kfkH 1
replaced by the general (1,∞,L)-atoms aj(x).
Lemma 3.7. There is a constant C > 0 such that for every a being a generalized
(1,∞,L) atom associated with B(y0, r) one has
, where the norm kfkH 1
at,∞,L
at,∞,L
L−1/2a(y) ≤ Cr1−d(cid:16)1 + y − y0
r
(cid:17)−d
.
Proof. The proof follows from functional calculi (see, e.g., [11]). Note that L−1/2a =
m(r)(L)b with m(r)(λ) = r(r2λ)−1/2(e−r2λ − 1) and b such that supp b ⊂ B(y0, r),
kbkL∞ ≤ B(y0, r)−1. From [11] we conclude that there is a constant C > 0 such that
for every r > 0 one has
m(r)(L)f (x) =ZRd
m(r)(x, y)f (y) dy,
with m(r)(x, y) satisfying
(3.8)
m(r)(x, y) ≤ Cr1−d(cid:16)1 + x − y
r
(cid:17)−d
.
8
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
Now the lemma can be easily deduced from (3.8) and the size and support property of
b.
(cid:3)
For real numbers n > 2, β > 0 let
4. Proof of Theorem 1.6
g(x) = (1 + x)−n−β,
gs(x) = s−n/2g(cid:0) x
√s(cid:1).
gs(x) ds ≤ Cx2−n(cid:16)1 + x√t(cid:17)−2−β
;
One can easily check that
0
(4.1)
(4.2)
r (cid:17)−n+2
Moreover, it is easily to verify that for 1 < q < ∞, d(1 − 1
(4.3)
Z t
Z ∞
gs(x) ds ≤ Cr2−n(cid:16)1 + x
(cid:13)(cid:13)(cid:13)xα−d(cid:16)1 + x√t(cid:17)−d−β(cid:13)(cid:13)(cid:13)Lq(Rd, dx)
r (cid:17)−d+γ
(cid:17)−β(cid:16)1 + y
Z z − y2−d(cid:16)1 + z − y
and
(4.4)
r2
r
= Cα,βt(α−d+d/q)/2
for 0 < γ < β < 2 .
for r > 0.
q ) < α ≤ d, β > 0 one has
dy ≤ Cr2(cid:16)1 + z
r (cid:17)−d+γ+2−β
Lemma 4.5. Assume that V (x) satisfies the assumptions of Theorem 1.6. Then for
0 < γ ≤ 2 and r > 0 one has
(4.6)
Proof. The left-hand side of (4.6) is bounded by
ZRd
Zz−y≤r
r
V (z)(cid:16)1 + z − y
z − y(cid:17)d−2
V (z)(cid:16)
r
d rd−2k∆−1V kL∞.
(cid:17)−d+γ
dz ≤ c−1
dz +Zz−y>r
≤ c−1
V (z)(cid:16)z − y
r
d rd−2k∆−1V kL∞.
(cid:17)−d+2
dz
(cid:3)
Proof of Theorem 1.6. We already have known that the operators (−∆)1/2L−1/2 and
L1/2(−∆)−1/2 are bounded on L1(Rd). It suffices to prove (1.7) and (1.8). Set γ = 1
and fix q > 1 and ε > 0 such that γ > d(1 − 1
q ) + ε. Set w0(x) ≡ 1. According to the
atomic and molecular decompositions (see Section 3) the proof of (1.7) will be done if we
verify that (−∆)1/2L−1/2a is a multiple of a (1, q, ε, w0)-molecule for every generalized
(1,∞, L)-atom a with a multiple constant independent of a. Identical arguments can
be then applied to show that L1/2(−∆)−1/2a is a (1, q, ε, w)-molecule for a being a
generalized atom for the classical Hardy space H 1(Rd) = H 1
−∆ with a multiple constant
independent of a.
10
HARDY SPACES FOR SCHR ODINGER OPERATORS
9
Let a = (I − Kr2)b be a generalized (1,∞, L)-atom for H 1
By Lemma 2.11, since R w(x)a(x) dx = 0, we have that
Z (−∆)1/2L−1/2a(x) dx = 0.
L associated with B(y0, r).
Set
(4.7)
J(x) =Z ∞
0 Z t
=Z r2
0 Z t
0 ZZ Pt−s(x − z)V (z)Ks(z, y)(L−1/2a)(y) dy dz ds
0 ZZ ... +Z ∞
ZZ +Z ∞
r2 Z t
t/2ZZ ...
r2 Z t/2
0
dt
t3/2
= J1(x) + J2(x) + J3(x).
Thanks to (2.9) and Lemma 3.5 it suffices to show that there is a constant Cq > 0,
independent of a(x) such that
(4.8)
(4.9)
Applying Lemma 3.7 and (4.1) with n = d + 1, we obtain
r
(cid:17)γ
J(x)(cid:13)(cid:13)(cid:13)Lq(Rd) ≤ Cqr−d+d/q.
(cid:13)(cid:13)(cid:13)(cid:16)1 + x − y0
J1(x) =(cid:12)(cid:12)(cid:12)Z r2
0 Z t
0 ZZ Pt−s(x − z)V (z)Ks(z, y)(L−1/2a)(y) dy dz ds
≤ CZ r2
0 Z t
0 Z Pt−s(x − z)V (z)r1−d(cid:16)1 + z − y0
≤ CZ r2
0 Z Ps(x − z)V (z)r1−d(cid:16)1 + z − y0
(cid:17)−d
≤ CNZ x − z1−d(cid:16)1 + x − z
V (z)r1−d(cid:16)1 + z − y0
(cid:17)−d
ds
√s
(cid:17)−N
dz ds
dz
r
r
r
r
dt
t3/2(cid:12)(cid:12)(cid:12)
dt
t3/2
(cid:17)−d
dz.
Consequently,
(4.10)
J1(x)(cid:16)1 + x − y0
≤ CN r1−dZ x − z−d+1(cid:16)1 + x − z
(cid:17)γ
r
r
(cid:17)−N +γ
V (z)(cid:16)1 + z − y0
r
(cid:17)−d+γ
dz.
Therefore, using the Minkowski integral inequality together with (4.3) and (4.6), we get
(4.11)
(cid:13)(cid:13)(cid:13)J1(x)(cid:16)1 + x − y0
r
(cid:17)γ(cid:13)(cid:13)(cid:13)Lq(dx) ≤ Cr−d+d/q.
In order to estimate J2(x) we use Lemma 3.7 and (4.1) with n = d to obtain
10
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
(4.12)
r
r
(cid:17)γ
J2(x)(cid:16)1 + x − y0
(cid:17)γZ t/2
ZZ t−d/2e−cx−z2/tV (z)
≤ CZ ∞
r2 (cid:16)1 + x − y0
× Ks(z, y)r1−d(cid:16)1 + y − y0
(cid:17)−d
≤ CZ ∞
× z − y2−d(cid:16)1 + z − y√t (cid:17)−N +γ
r2 ZZ t(2γ−d−3)/2e−cx−z2/tV (z)
r1−d−2γ(cid:16)1 + y − y0
dy dz ds
dt
t3/2
r
r
0
(cid:17)−d+γ
dy dz dt.
Setting N = β + γ with 0 < γ < β < 2 and applying the Minkowski integral inequality
together with (4.4) and (4.6) we conclude that
(4.13)
r
r1−d−2γ(cid:16)1 + y − y0
(cid:13)(cid:13)(cid:13)J2(x)(cid:16)1 + x − y0
(cid:17)γ(cid:13)(cid:13)(cid:13)Lq(dx)
≤ CZ ∞
r2 ZZ t−(d+3−2γ−d/q)/2V (z)
× z − y2−d(cid:16)1 + z − y√t (cid:17)−β
≤ CZ ∞
r2 ZZ t−(d+3−2γ−d/q)/2V (z)
(cid:17)−β(cid:16)√t
r (cid:17)β
r1−d−2γ(cid:16)1 + y − y0
× z − y2−d(cid:16)1 + z − y
≤ CZ r−2d+2+d/qV (z)(cid:16)1 + z − y0
(cid:17)−d+2+γ−β
≤ Cr−d+d/q.
(cid:17)−d+γ
dz
r
r
r
r
dy dz dt
(cid:17)−d+γ
dy dz dt
By Lemma 3.7 and (4.1) with n = d, we have
J3(x)
≤ CZ ∞
r2 Z t
t
2
ZZ Pt−s(x − z)V (z)t− d
×(cid:16)1 + y − y0
(cid:17)−d
r
2 e− cz−y2
t
r1−d dy dz ds
dt
3
2
t
V (z)
≤ CNZ ∞
× t− d
r2 ZZ x − z2−d(cid:16)1 + x − z√t (cid:17)−N
2 e−cz−y2/t(cid:16)1 + y − y0
(cid:17)−d
r
r1−d dy dz ds
dt
t3/2 .
HARDY SPACES FOR SCHR ODINGER OPERATORS
11
Hence,
By Minkowski's integral inequality combined with (4.3) we arrive to
tγV (z)
r1−d−2γ dy dz ds
dt
t3/2 .
(cid:17)−d+γ
r1−d−2γ dy dz dt.
r
(cid:17)γ
J3(x)(cid:16)1 + x − y0
≤ CZ ∞
× t− d
r2 ZZ x − z2−d(cid:16)1 + x − z√t (cid:17)−N +γ
2 e−c′z−y2/t(cid:16)1 + y − y0
(cid:17)γ(cid:13)(cid:13)(cid:13)Lq(dx)
(cid:13)(cid:13)(cid:13)J3(x)(cid:16)1 + x − y0
≤Z ∞
r2 ZZ t(−d+2+d/q)/2+γ−3/2V (z)
(cid:17)−d+γ
r
r
r
× t−d/2e−c′z−y2/t(cid:16)1 + y − y0
(cid:17)γ(cid:13)(cid:13)(cid:13)Lq(dx)
(cid:13)(cid:13)(cid:13)J3(x)(cid:16)1 + x − y0
≤ CZZ r2−3d+d/qV (z)(cid:16)1 + z − y
≤Z r2−2d+d/qV (z)(cid:16)1 + z − y0
≤ Cr−d+d/q.
r
r
r
Application of (4.2) with n = 2d + 1 − d
q − 2γ and then (4.6) yields
(cid:17)−2d+1+d/q+2γ(cid:16)1 + y − y0
(cid:17)−2d+1+d/q+3γ
dz
r
(cid:17)−d+γ
dy dz
The above inequality together with (4.11) and (4.13) gives desired (4.8) and, conse-
quently, the proof of (1.7) is complete.
Let us note that in the proof (1.7) we use only Lemmas 2.11, 3.7, and the upper
Gaussian bounds for the kernels. The proof of (1.8) goes identically to that of (1.7) by
replacing Lemma 2.11 by Lemma 2.13.
(cid:3)
5. Proof of the Riesz transform characterization of H 1
L
(5.1)
∂
∂xj
L−1/2f =
Proof of Corollary 1.9. Assume that f ∈ H 1
L. Then, thanks to Theorem 1.6, there is
g ∈ H 1(Rd) such that f = L1/2(−∆)−1/2g. By the characterization of the classical
Hardy space H 1(Rd) by the Riesz transforms we have
∂
∂xj
L−1/2f ∈ L1(Rd) for j = 1, 2, ..., d.
L−1/2L1/2(−∆)−1/2g =
Conversely, assume that for f ∈ L1(Rd) we have ∂
Set g = (−∆)1/2L−1/2f . Then by Lemma 2.6, g ∈ L1(Rd) and
∂
(5.2)
∂xj
(−∆)−1/2(−∆)1/2L−1/2f =
(−∆)−1/2g ∈ L1(Rd).
L−1/2f ∈ L1(Rd),
(−∆)−1/2g =
∂
∂xj
∂
∂xj
∂
∂xj
∂xj
12
JACEK DZIUBA ´NSKI AND JACEK ZIENKIEWICZ
which implies that g ∈ H 1(Rd). Consequently, by Theorem 1.6, f ∈ H 1
can be deduced from (5.1), (5.2), and Theorem 1.6.
L. Finally (1.10)
(cid:3)
References
[1] Auscher, P., Duong, X.T., McIntosh, A.: Boundedness of Banach space valued singular integral
operators and Hardy spaces, Unpublished preprint (2005)
[2] Bernicot, F., Zhao, J.: New abstract Hardy spaces, J. Funct. Anal., 255, 1761 -- 1796 (2008)
[3] Coifman, R.: A real variable characterization of H p, Studia Math., 51, 269 -- 274 (1974)
[4] Czaja, W., Zienkiewicz, J.: Atomic characterization of the Hardy space H 1
L(R) of one-dimensional
Schrodinger operators with nonnegative potentials, Proc. Amer. Math. Soc. 136, no. 1, 89 -- 94 (2008)
[5] Duong, X.T., Yan, L.X.: Duality of Hardy and BMO spaces associated with operators with heat
kernel bounds, J. Amer. Math. Soc., 18, 943 -- 973 (2005)
[6] Dziuba´nski, J., Garrig´os, G., Mart´ınez, T., Torrea, J.L., Zienkiewicz, J.: BMO spaces related to
Schrodinger operators with potentials satisfying a reverse Holder inequality, Math. Z., 249, 329 -- 356
(2005).
[7] Dziuba´nski, J., Zienkiewicz, J.: Hardy space H 1 associated to Schrodinger operator satisfying
reverse Holder inequality, Rev. Mat. Iberoamericana, 15, 279 -- 296 (1999)
[8] Dziuba´nski, J., Zienkiewicz, J.: Hardy spaces H 1 for Schrodinger operators with certain potentials,
Studia Math., 164, 39 -- 53 (2004)
[9] Dziuba´nski, J., Zienkiewicz, J.: On Hardy spaces associated with certain Schrodinger operators in
dimension 2, Rev. Mat. Iberoam., 28, no. 4, 1035 -- 1060 (2012)
[10] Dziuba´nski, J., Zienkiewicz, J.: On Isomorphisms of Hardy Spaces Associated with Schrodinger
Operators, J. Fourier Anal. Appl., 19, 447 -- 456 (2013)
[11] Hebisch, W.: A multiplier theorem for Schrodinger operators, Colloq. Math., 60/61, 659 -- 664
(1990)
[12] Hofmann, S., Lu, G.Z., Mitrea, D., Mitrea, M., Yan, L.X.: Hardy spaces associated with non-
negative self-adjoint operators satisfying Davies-Gafney estimates, Mem. Amer. Math. Soc., 214, no.
1007 (2011)
[13] Latter, R.H.: A decomposition of H p(Rn) in terms of atoms, Studia Math., 62 no. 1, 93 -- 101
(1978)
[14] Semenov, Yu.A.: Stability of Lp-spectrum of generalized Schrodinger operators and equivalence
of Green's functions, Internat. Math. Res. Notices, 12, 573 -- 593 (1997).
[15] Stein, E.: Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals,
Princeton University Press, Princeton, NJ, (1993)
HARDY SPACES FOR SCHR ODINGER OPERATORS
13
Instytut Matematyczny, Uniwersytet Wroc lawski, pl. Grunwaldzki 2/4, 50-384
Wroc law, Poland
E-mail address: [email protected], [email protected]
|
1501.05746 | 2 | 1501 | 2015-10-28T20:51:34 | The Riesz Capacity in Metric Spaces | [
"math.FA",
"math.CA"
] | We study a capacity theory based on a definition of a Riesz potential in metric spaces with a doubling measure. In this general setting, we study the basic properties of the Riesz capacity, including monotonicity, countable subadditivity and several convergence results. We define a modified version of the Hausdorff measure and provide lower bound and upper bound estimates for the capacity in terms of the modified Hausdorff content. | math.FA | math |
THE RIESZ CAPACITY IN METRIC SPACES
JUHO NUUTINEN AND PILAR SILVESTRE
Abstract. We study a capacity theory based on a definition of
a Riesz potential in metric spaces with a doubling measure.
In
this general setting, we study the basic properties of the Riesz ca-
pacity, including monotonicity, countable subadditivity and several
convergence results. We define a modified version of the Hausdorff
measure and provide lower bound and upper bound estimates for
the capacity in terms of the modified Hausdorff content.
1. Introduction
In this paper, we study a theory of capacity based on a metric version
of the Riesz potential in the setting of a general metric space (X, d)
equipped with a doubling measure µ. We define a related Hausdorff
measure and study the connections between the Riesz capacity and
the Hausdorff measure. With our definitions and results, we extend
the classical Riesz capacity theory from the Euclidean space, with the
Lebesgue measure, to the setting of a general metric measure space.
In Rn, the capacity theory for the Riesz potential can be found for
example in [2], [3], [29] and [30]. During the past twenty years, different
capacities have been studied in metric measure spaces for example in
[5], [15], [17], [19], [21], [22], [23], [24] and [25]. Also, a part of the
theory for Riesz capacity follows from general results in [12] and [32].
Here, we formulate the theory explicitly and state the results to keep
the paper self-contained.
We define a metric version of the Riesz potential of order γ, where
0 < γ < 1, as
Iγf (x) =ZX
f (y)
µ (B(x, d(x, y)))1−γ dµ(y).
One can find a similar definition for the Riesz potential in the works
of Kairema and Sjodin (see [20], [31] and [32]). In the definition, there
appears only the measure of balls in the Riesz kernel. Another defini-
tion for a metric version of the Riesz potential is such that it also has
the distance function as a part of the kernel. This version of the Riesz
potential can be found for example in [16], [18] and [26]. Also, other
Riesz potentials and fractional integral operators have been studied in
2010 Mathematics Subject Classification. 31E05, 31B15.
The research was supported by the Academy of Finland, grant no. 272886.
1
2
JUHO NUUTINEN AND PILAR SILVESTRE
the metric setting for example in [13], [14] and [27]. We emphasize
that, throughout the paper, we do not assume any type of (Ahlfors)
Q-regularity on the measure µ that would give uniform lower bounds
or upper bounds for the measure of balls in terms of the radii.
In
this generality, our definition of the Riesz potential, with no distance
function as a part of the kernel, works better.
In Section 3, we define a metric version of the Riesz capacity Cγ,p and
show that it satisfies the basic properties of capacity. These properties
include monotonicity, countable subadditivity and several convergence
results.
In particular, we show that the Riesz capacity is a Fatou
capacity. This lower semicontinuity property of capacity is an analogue
of Fatou’s lemma. We also study the capacitability of sets and show
that the Riesz capacity is a so called Choquet capacity. This means
that the capacity of a Borel set can be obtained by approximating with
compact sets from the inside and open sets from the outside. We finish
the section by briefly studying the dual Riesz capacity.
In the beginning of Section 4, we prove an upper bound estimate for
the capacity of balls in terms of the measure µ. This result leads us to
define a modified version of the standard Hausdorff content. The main
results of the section are lower bound and upper bound estimates for
the Riesz capacity in terms of this modified Hausdorff content. Similar
results have been studied by Sjodin in [31]. Here, we give direct proofs
to results that apply not only for compact sets. In particular, we do
not need to use Frostman’s lemma to obtain the results.
Acknowledgements. We would like to thank Professor Juha Kin-
nunen for proposing this project. We would also like to thank Juha
Lehrback and Heli Tuominen for useful discussions and comments on
the manuscript.
2. Notation and preliminaries
2.1. Riesz potential. We assume that X = (X, d, µ) is a locally com-
pact metric measure space equipped with a metric d and a Borel reg-
ular, doubling outer measure µ. The doubling property means that
there is a fixed constant cd ≥ 1, called the doubling constant of µ, such
that
(2.1)
µ(B(x, 2r)) ≤ cdµ(B(x, r))
for every ball B(x, r) = {y ∈ X : d(y, x) < r}. We also assume that the
measure of each open ball is positive and finite. The doubling condition
implies that
(2.2)
µ(B(y, r))
µ(B(x, R))
R(cid:17)Q
≥ C(cid:16) r
for every 0 < r ≤ R and y ∈ B(x, R) for some C > 0 and Q > 0
that only depend on cd. In fact, we may take Q = log2 cd and C = c−2
d
THE RIESZ CAPACITY IN METRIC SPACES
3
(see [5]). In addition, we assume that spheres are of measure zero, i.e.
(2.3)
µ ({y ∈ X : d(x, y) = r}) = 0,
for x ∈ X and B(x, r). This assumption is needed for the Riesz po-
tential, defined below, to satisfy lower semicontinuity properties (see
Remark 3.3) that are required for the capacity theory.
Definition 2.1. Let 0 < γ < 1. The Riesz potential of order γ of a
measurable function f is
(2.4)
Iγf (x) =ZX
f (y)
µ (B(x, d(x, y)))1−γ dµ(y).
Remark 2.2. (i) To be precise, we would need to define the kernel
separately for the cases x 6= y and x = y. However, we assume our
space X to be such that µ vanishes on sets which consist of a single
point. Then the domain of integration X \ {x} can be replaced by
X (see [20]). Since we have a doubling metric measure space, this is
equivalent to the condition that there are no isolated points in our
space X.
(ii) In the Euclidean space, with the n-dimensional Lebesgue measure,
we have, with the notation α = γn ∈ (0, n), the usual Riesz potential
Iαf (x) =ZRn
f (y)
x − yn−α
dy
of order α on Rn (up to a dimensional constant).
Another way to define a Riesz potential in a metric space, as in [16]
and [18], is
(2.5)
eIγf (x) =ZX
f (y)d(x, y)γ
µ (B(x, d(x, y)))
dµ(y).
If the measure µ is (Ahlfors) Q-regular, that is, there exists a constant
C > 1 such that
(2.6)
C −1rQ ≤ µ(B(x, r)) ≤ CrQ
for every x ∈ X and 0 < r < diam(X), then Iγf and eIγQf are compa-
rable in the sense that there exists a constant C ≥ 1 such that
C −1Iγf ≤ eIγQf ≤ CIγf.
In the next sections, we do not assume the (Ahlfors) Q-regularity
or any other estimates that would give uniform lower bounds or upper
bounds for the measure of balls in terms of the radii. We assume only
the doubling property (2.1) and develop the theory of Riesz capacity
based on the definition (2.4) of the Riesz potential. In particular, this
definition works better for our purposes in Section 4, where we define
a modified version of the standard Hausdorff measure and prove two
results that relate the Riesz capacity and the Hausdorff measure.
4
JUHO NUUTINEN AND PILAR SILVESTRE
2.2. Function spaces and capacities. We have by Cavalieri’s prin-
ciple that Lp(X) = Lp(X, µ) is the space of all µ-measurable functions
f in X such that
kf kLp(X) =(cid:18)Z ∞
0
ptp−1µ({z ∈ X : f (z) > t})dt(cid:19)1/p
< ∞,
which is a Banach space when 1 ≤ p < ∞. The weak Lp-space Lp,∞(X)
is defined by the condition
kf kLp,∞(X) := sup
t>0
tµ ({z ∈ X : f (z) > t})1/p < ∞.
We denote by Lp
+(X) the subset of Lp(X) of non-negative functions.
Definition 2.3. We define, on the family of µ-measurable subsets of X,
a capacity to be a non-negative set function C, which has the following
properties:
(a) C(∅) = 0,
(b) If A ⊂ B, then C(A) ≤ C(B),
(c) C (S∞
i=1 Ai) ≤P∞
i=1 C(Ai).
A1 ⊂ A2 ⊂ · · · are subsets of X and A =S∞
A capacity C is called a Fatou capacity if C(Ai) → C(A), whenever
i=1 Ai. We also say that a
property holds C-q.e. on X if it holds for all x ∈ X except those in a
set E with C(E) = 0.
The capacitary Lorentz spaces Lp,q(C), p, q > 0, are defined by the
condition
kf kLp,q(C) :=(cid:16)qZ ∞
0
tq−1C ({z ∈ X : f (z) > t})q/p dt(cid:17)1/q
< ∞,
when q < ∞, and in the case of q = ∞ by
kf kLp,∞(C) := sup
t>0
t C ({z ∈ X : f (z) > t})1/p < ∞.
The space Lp,∞(C) is called the weak capacitary Lp-space. For the
general facts and properties of the capacitary Lorentz spaces, we refer
to [6], [7] and [8]. Throughout the paper, we denote the characteristic
function of a set E ⊂ X by χE. In general, C will denote a positive
constant whose value is not necessarily the same at each occurrence.
3. Riesz Capacity
Definition 3.1. Let 1 < p < ∞ and 0 < γ < 1. The Riesz (γ, p)-
capacity of a set E ⊂ X is the number
Cγ,p(E) = inf
f ∈A(E)
f p
Lp(X),
where
A(E) = {f ∈ Lp
+(X) : Iγf ≥ 1 on E} .
THE RIESZ CAPACITY IN METRIC SPACES
5
If A(E) = ∅, we set Cγ,p(E) = ∞. Functions belonging to A(E) are
called admissible functions or test functions for E. From now on, we
always assume in this section that 1 < p < ∞ and 0 < γ < 1.
In the Euclidean space, with the Lebesgue measure, one can find
the basic properties of the Riesz capacity for example in [2], [3], [29]
and [30].
In the metric case, assuming only the doubling property
from the Borel regular measure µ, we begin by showing that the Riesz
(γ, p)-capacity is an outer measure. This means that the Riesz capacity
satisfies the properties of Definition 2.3.
Theorem 3.2. The Riesz (γ, p)-capacity is an outer measure.
Proof. Clearly Cγ,p(∅) = 0, since 0 is an admissible function. The
definition of the capacity also implies monotonicity, since if E1 ⊂ E2,
then A(E2) ⊂ A(E1).
To prove the countable subadditivity, let {Ai}∞
sets in X and let A = S∞
i=1 be a sequence of
i=1 Cγ,p(Ai) <
∞. Then, Cγ,p(Ai) < ∞ for all i ∈ N. Let ǫ > 0, and for each i ∈ N,
let fi ∈ A(Ai) be such that
fip
i=1 Ai. We may assume that P∞
We define f (x) := sup
i∈N
implies
f p
Lp(X) ≤
∞Xi=1
fip
Lp(X) < Cγ,p(Ai) + ǫ2−i.
fi(x). We have that f (x)p ≤P∞
∞Xi=1 (cid:0)Cγ,p(Ai) + ǫ2−i(cid:1) =
∞Xi=1
Lp(X) ≤
i=1 fi(x)p, which
Cγ,p(Ai) + ǫ.
Moreover, we have that Iγf (x) ≥ Iγfi(x), since f (x) ≥ fi(x) for all
x ∈ X and i ∈ N. Let x ∈ A. Then there exists j ∈ N such that
x ∈ Aj and hence Iγf (x) ≥ Iγfj(x) ≥ 1. Thus f is an admissible
function for A =S∞
Cγ,p(cid:16) ∞[i=1
i=1 Ai. Now
Ai(cid:17) ≤ f p
Lp(X) ≤
∞Xi=1
Cγ,p(Ai) + ǫ,
and the claim follows by letting ǫ → 0.
(cid:3)
Remark 3.3. The Riesz potential, as defined in (2.4), is lower semi-
continuous. For our purposes, it is enough to prove the lower semicon-
tinuity for functions f ∈ Lp
+(X). Let x0 ∈ X. Then
Iγf (x0) =ZX
f (y)
µ (B(x0, d(x0, y)))1−γ dµ(y).
We need to show that
Iγf (x0) ≤ lim inf
x→x0
Iγf (x),
when x → x0. Let x ∈ X. Since for any y ∈ X
B(x, d(x, y)) ⊂ B(x0, d(x, y) + d(x0, x)),
6
JUHO NUUTINEN AND PILAR SILVESTRE
we have by the monotonicity of µ that
µ (B(x, d(x, y))) ≤ µ (B(x0, d(x, y) + d(x0, x))) .
The above inequality and equality (2.3) imply that
lim sup
x→x0
µ (B(x, d(x, y))) ≤ lim
x→x0
µ(B(x0, d(x, y) + d(x0, x)))
= µ (B(x0, d(x0, y))) .
Now, for any y ∈ X,
f (y)
µ (B(x, d(x, y)))1−γ
x→x0
and, by using Fatou’s lemma, we get
f (y)
µ (B(x0, d(x0, y)))1−γ ≤ lim inf
Iγf (x0) =ZX
≤ZX
x→x0 ZX
lim inf
x→x0
≤ lim inf
f (y)
µ (B(x0, d(x0, y)))1−γ dµ(y)
f (y)
f (y)
µ (B(x, d(x, y)))1−γ dµ(y)
µ (B(x, d(x, y)))1−γ dµ(y)
= lim inf
x→x0
Iγf (x).
In addition, we get the following lower semicontinuity property of
the Riesz potential as an operator
Iγf ≤ lim inf
i→∞
Iγfi,
when fi → f weakly in Lp
+(X). The weak convergence implies that
fiµ → f µ converge weakly as measures with the vague topology of [12,
Section 1.1]. Also, because of equality (2.3) in the previous section,
we have that our Riesz kernel is continuous and hence lower semicon-
tinuous. The result then follows from [12, Lemma 2.2.1.b)] and [30,
Theorem 1.2. p.58].
Using the fact that the Riesz potential of a function f is lower semi-
continuous, we show that the Riesz capacity is an outer capacity. This
means that the capacity of a set E ⊂ X can be obtained by approxi-
mating with open sets from the outside.
Theorem 3.4. Cγ,p is an outer capacity, that is,
Cγ,p(E) = inf {Cγ,p(O) : O ⊃ E, O open} .
Proof. By the monotonicity, Cγ,p(E) ≤ inf {Cγ,p(O) : O ⊃ E, O open}.
To prove the inequality to the reverse direction, we may assume that
Cγ,p(E) < ∞. Let 0 < ǫ < 1 and let f ∈ Lp
+(X) be a function such
that Iγf ≥ 1 on E and
f p
Lp(X) < Cγ,p(E) + ǫ.
THE RIESZ CAPACITY IN METRIC SPACES
7
We define
and
fǫ :=
1
1 − ǫ
f
G := {x ∈ X : Iγfǫ(x) > 1} .
Since Iγfǫ is lower semicontinuous, G is an open set. We also have that
fǫ(x) > f (x) for all x ∈ X, since 0 < ǫ < 1. Now, if x ∈ E, then
Iγf (x) ≥ 1 and hence Iγfǫ(x) > 1. Thus we have that x ∈ G and
E ⊂ G. Moreover, fǫ is admissible for Cγ,p(G) and
Cγ,p(G) ≤ fǫp
Lp(X) =(cid:16) 1
1 − ǫ(cid:17)p
Lp(X)
≤ Cγ,p(E)(1 − ǫ)−p + ǫ(1 − ǫ)−p.
f p
Since we have that inf {Cγ,p(O) : O ⊃ E, O open} ≤ Cγ,p(G), letting
ǫ → 0 yields the inequality to the other direction.
(cid:3)
The next capacitary weak type lemma shows in particular that the
Riesz potential Iγf of a nonnegative Lp-function f belongs to the weak
capacitary Lp-space.
Lemma 3.5. If f ∈ Lp
+(X), then the capacitary weak type estimate
Cγ,p ({x ∈ X : Iγf (x) > a}) ≤ a−pf p
Lp(X)
holds for each 0 < a < ∞. Moreover, Iγ is bounded from Lp
Lp,∞(Cγ,p).
Proof. Let f ∈ Lp
+(X) and 0 < a < ∞. We define
+(X) to
and
fa :=
f
a
F := {x ∈ X : Iγf (x) > a} .
Since Iγfa = Iγ(cid:0) f
a(cid:1) ≥ 1 on F , fa is admissible for F and
Cγ,p(F ) ≤ fap
Lp(X) = a−pf p
Lp(X).
Moreover, the capacitary weak type estimate implies that
kIγf kLp,∞(Cγ,p) = sup
t>0
t Cγ,p ({x ∈ X : Iγf (x) > t})1/p ≤ f Lp(X),
and the second claim follows.
(cid:3)
We use the capacitary weak type estimate to prove the next theorem,
which in particular says that the Riesz potential of a function f ∈
Lp
+(X) is finite Cγ,p- q.e. It follows that Iγf , for f ∈ Lp(X), is well-
defined Cγ,p- q.e. and that the Riesz potential, as an operator, is linear
outside a set of capacity zero.
Theorem 3.6. Let E ⊂ X. Then Cγ,p(E) = 0 if and only if there
exists f ∈ Lp
+(X) such that Iγf (x) = ∞ for all x ∈ E.
8
JUHO NUUTINEN AND PILAR SILVESTRE
Proof. If Cγ,p(E) = 0, then for any integer j, we can find an admissible
function fj ∈ A(E) such that
fjp
Lp(X) =ZX
Then the function f :=P∞
P∞
Iγf (x) =ZX
fj(y)pdµ(y) < 2−j.
j=1 fj belongs to Lp(X) and
j=1 fj(y)
∞Xj=1
µ (B(x, d(x, y)))1−γ dµ(y) =
Iγfj(x) = ∞
for all x ∈ E, since Iγfj ≥ 1 on E for each j.
Conversely, if there exists a nonnegative function f ∈ Lp(X) such
that Iγf = ∞ on E, then by the capacitary weak type estimate
Cγ,p(E) ≤ Cγ,p ({x ∈ X : Iγf (x) > a}) ≤ a−pf p
Lp(X)
for every a > 0. By letting a → ∞, we see that Cγ,p(E) = 0.
(cid:3)
Corollary 3.7. Let f1, f2, f ∈ Lp(X). Then
Iγ(f1 + f2) = Iγ(f1) + Iγ(f2),
Cγ,p-q.e.
and
Iγ(af ) = aIγ(f ),
Cγ,p-q.e.,
where a is any finite constant.
Proof. If each term on the right side of the above equalities is finite at
a point x ∈ X, then the equalities hold at such point by the definition
of the Riesz potential. By Theorem 3.6, the sets where the equalities
can fail are of capacity zero.
(cid:3)
Next, we are going to prove several convergence results. We start by
defining the convergence of a sequence of functions in capacity.
Definition 3.8. We say that a sequence {fi} converges in capacity to
f , denoted fi → f in Cγ,p, if for every ǫ > 0
lim
i→∞
Cγ,p ({x ∈ X : fi(x) − f (x) > ǫ}) = 0.
We show that the Lp-convergence of functions implies that the corre-
sponding sequence of the Riesz potentials converges in capacity. Also,
for a subsequence, we have pointwise convergence except for a set of
capacity zero.
Theorem 3.9. Let {fi} ⊂ Lp
statements is a consequence of the previous one.
+(X). Each of following
+(X) and f ∈ Lp
(i) fi → f in Lp(X)
(ii) Iγfi → Iγf in Cγ,p
(iii) There exists a subsequence {fij } of {fi} such that
Iγfij → Iγf pointwise Cγ,p-q.e.
THE RIESZ CAPACITY IN METRIC SPACES
9
Proof. We show first that (i) implies (ii). Let ǫ > 0. By Theorem
3.6, the potentials Iγfi and Iγf are finite Cγ,p-q.e. Then, we have by
Corollary 3.7 and by Lemma 3.5 that
Cγ,p ({x ∈ X : Iγfi(x) − Iγf (x) > ǫ}) ≤ ǫ−pfi − f p
Lp(X),
which proves the claim.
Next, we assume that (ii) holds and show that it implies (iii). Let
ǫ = 2−j. Then, there exists a subsequence {fij } such that
Cγ,p(cid:0)(cid:8)x ∈ X : Iγ(fij )(x) − Iγf (x) > 2−j(cid:9)(cid:1) < 2−j.
We use the notation Aj = (cid:8)x ∈ X : Iγ(fij )(x) − Iγf (x) > 2−j(cid:9). The
upper limit set
A =
has zero capacity, since
∞\k=1
∞[j=k
Aj
Cγ,p(A) ≤
∞Xj=k
Cγ,p(Aj) ≤
∞Xj=k
2−j
for all k. Now, if x ∈ X \ A then there exists k = k(x) such that
x ∈ X \ Aj for j ≥ k, that is
Iγ(fij )(x) − Iγf (x) ≤ 2−j.
This implies that Iγ(fij )(x) → Iγf (x), as j → ∞, which proves the
claim.
(cid:3)
By using the above theorem, we can strengthen the lower semiconti-
nuity property of Iγ from Remark 3.3, at least outside a set of capacity
zero.
Theorem 3.10. Let {fi} ⊂ Lp(X) and f ∈ Lp(X).
(i) If fi → f weakly in Lp(X), then
Iγfi ≤ Iγf ≤ lim sup
lim inf
i→∞
i→∞
Iγfi,
Cγ,p-q.e.
(ii) If fi → f weakly in Lp
+(X), then
and
Iγf ≤ lim inf
i→∞
Iγfi everywhere
Iγf = lim inf
i→∞
Iγfi
Cγ,p-q.e.
Proof. We prove first the claim (i). By the Banach-Saks Theorem (see
[4]), there exists a subsequence {f ′
i} such that a sequence {gj}, where
gj = j−1
jXi=1
f ′
i ,
10
JUHO NUUTINEN AND PILAR SILVESTRE
converges to f in Lp(X). Then, by Theorem 3.9, there exists a subse-
quence {g′
j} such that
Iγf = lim
j→∞
Iγg′
j
Cγ,p-q.e.
Then, the left inequality in (i) follows due to the fact that
Iγf = lim
j→∞
Iγg′
j ≥ lim inf
i→∞
Iγf ′
i ≥ lim inf
i→∞
Iγfi
Cγ,p-q.e.
The right inequality in (i) follows by replacing fi and f by −fi and −f
in the previous argument.
If fi → f weakly in Lp
+(X), then by the lower semicontinuity of Iγ(·)
(see Remark 3.3), we have that
Iγf ≤ lim inf
i→∞
Iγfi everywhere
and it follows by (i) that
Iγf = lim inf
i→∞
Iγfi
Cγ,p-q.e.
We prove two more convergence results for the Riesz capacity. As a
corollary of Theorem 3.12, we get a lower semicontinuity property for
the capacity that is an analogue of Fatou’s lemma.
(cid:3)
Theorem 3.11. If X ⊃ K1 ⊃ K2 · · · are compact sets and K =
i=1 Kj, then
T∞
lim
i→∞
Cγ,p(Ki) = Cγ,p(K).
Proof. Clearly, by the monotonicity, limj→∞ Cγ,p(Kj) ≥ Cγ,p(K). On
the other hand, let O be an open set containing K. By the compact-
ness of K, we have that Kj ⊂ O for all sufficiently large j. Then
limj→∞ Cγ,p(Kj) ≤ Cγ,p(O). Finally, since Cγ,p is an outer capacity by
Theorem 3.4,
lim
j→∞
Cγ,p(Kj) ≤ inf {Cγ,p(O) : O ⊃ K, O open} = Cγ,p(K).
(cid:3)
Theorem 3.12. If A1 ⊂ A2 ⊂ · · · are subsets of X and A =S∞
then
i=1 Ai,
that is, the Riesz capacity Cγ,p is a Fatou capacity.
lim
i→∞
Cγ,p(Ai) = Cγ,p(A),
Proof. We may assume that limi→∞ Cγ,p(Ai) = l < ∞. Let fi ≥ 0 be a
test function for Cγ,p(Ai) such that
(3.1)
kfikp
Lp(X) ≤ Cγ,p(Ai) +
1
i
.
THE RIESZ CAPACITY IN METRIC SPACES
11
Then, the sequence {fi} is bounded in Lp(X) and there exists a sub-
sequence {fij } that converges weakly to a function f ∈ Lp
+(X). By
Theorem 3.10 (ii), we have that
and hence
Iγf ≥ 1 on Ai
Cγ,p-q.e.
Iγf ≥ 1 on A
Cγ,p-q.e.
Let E be the subset of A, where the previous inequality holds. Then,
by (3.1) and the weak convergence of the functions
Cγ,p(A) = Cγ,p(E) ≤ kf kp
Lp(X) ≤ lim inf
j→∞
fij p
Lp(X) ≤ l,
from which the result follows. Here, we also used the the lower semi-
continuity of · p
(cid:3)
p (see [3, Lemma 3.1.2. p.109]).
Corollary 3.13. If {Ai}∞
i=1 is a sequence of sets in X, then
Cγ,p(lim inf
i→∞
Ai) ≤ lim inf
i→∞
Cγ,p(Ai).
Proof. Let S := lim inf i→∞ Ai = SjTk≥j Ak and Si := Si
Then Si ⊂ Si+1 ⊂ · · · and S =S∞
i=1 Si, and by Theorem 3.12,
j=1Tk≥j Ak.
Cγ,p(S) = lim
i→∞
Cγ,p(Si) ≤ lim inf
i→∞
Cγ,p(Ai).
(cid:3)
The next definition extends the outer capacity property of Theorem
3.4 to the case where the Riesz capacity of a set E ⊂ X can also be
obtained by approximating with compact sets from the inside. By The-
orem 3.15, we have this inner capacity property for the Riesz capacity,
when considering analytic sets (for the definition of analytic sets, we
refer to e.g. [3], [9], [28]).
Definition 3.14. A set E ⊂ X is called Cγ,p-capacitable, if
Cγ,p(E) = sup{Cγ,p(K) : K ⊂ E, K compact}
= inf{Cγ,p(O) : O ⊃ E, O open}.
Capacitability has been studied in a very general context by Choquet
in [9]. Other references are [1], [2], [3], [11], [28] and [29], and the
references therein. Choquet’s capacitability theorem (see [3, pp.182–
184]) together with Theorems 3.11 and 3.12 give the next theorem,
which says that all analytic sets are Cγ,p-capacitable.
In particular,
we have that all Borel sets are Cγ,p-capacitable, which means that the
Riesz capacity is a so called Choquet capacity.
Theorem 3.15. All analytic sets, and hence all Borel sets, are Cγ,p-
capacitable.
12
JUHO NUUTINEN AND PILAR SILVESTRE
For the following variational problem
(3.2)
minnkf kp
Lp(X) : f ∈ Lp(X), Iγf ≥ 1 Cγ,p-q.e. on Ao,
we call a solution f a Cγ,p-capacitary distribution of A and Iγf a Cγ,p-
capacitary potential of A.
Theorem 3.16. If Cγ,p(A) < ∞, then A has a unique Cγ,p-capacitary
distribution f for which f ∈ Lp
Lp(X) = Cγ,p(A) and
+(X), kf kp
ZX
f (x)p−1g(x)dµ(x) ≥ 0
for all g ∈ Lp(X) such that
Iγg ≥ 0 Cγ,p-q.e. on A.
Proof. Using Clarkson’s inequality for Lp-norms, the theorem follows
as in [29, Theorem 9] due to previous results in the paper.
(cid:3)
Dual Riesz capacity. Let ν be a positive measure on X, 1 < p < ∞
and A ∈ F , where F is the σ-algebra of sets which are ν-measurable
for all positive measures ν with finite total variation in X. The total
variation of any such ν in X is
kνk = ν(X) = supnν(A) : A ⊂ X, A is measurableo.
In the case of a measure of this type, we define
Iγν(x) =ZX
dν(y)
µ(B(x, d(x, y)))1−γ
,
and following [29] we introduce a capacity, which uses measures as test
elements
cγ,p(A) = supnkνk : ν is a positive measure in X, ν(X \ A) = 0,
ν < ∞ and kIγνkLp′ (X) ≤ 1o,
where 1
p + 1
p′ = 1.
With the same techniques as in [29, Theorem 12, Theorem 14] or [3,
pp. 114–117], we see that cγ,p is an inner capacity on F that satisfies
(3.3)
Indeed, for the inequality cγ,p(A) ≤ Cγ,p(A)1/p, let f ∈ A(A) and let ν
be a test measure for cγ,p(A). Then, by Holder’s inequality,
cγ,p(A) = Cγ,p(A)1/p.
ν(A) ≤ZA
≤(cid:16)ZX(cid:16)ZA
Iγf (x)dν(x) =ZXZA
1
µ(B(x, d(x, y)))1−γ
≤ kf kLp(X)kIγνkLp′ (X) ≤ kf kLp(X)
1
dν(x)f (y)dµ(y)
µ(B(x, d(x, y)))1−γ
dν(x)(cid:17)p′
dµ(y)(cid:17) 1
p′(cid:16)ZX
f (y)pdµ(y)(cid:17) 1
p
THE RIESZ CAPACITY IN METRIC SPACES
13
and the inequality
cγ,p(A) ≤ Cγ,p(A)1/p
follows by taking the infimum over admissible functions f and the
supremum over the test measures ν. In order to obtain the equality
(3.3), one can show that (see [3] or [29])
+(X), kf kLp(X) ≤ 1(cid:9)
=(cid:16) inf
kνk=1
sup
f ∈Lp
+(X)
ZX
(Iγν)f dµ(cid:17)−1
Cγ,p(K)−1/p = sup
Iγf (x) : f ∈ Lp
x∈K
f (cid:8) inf
kIγνkLp′ (X)(cid:17)−1
and
cγ,p(K) =(cid:16) inf
kνk=1
for compact sets K, where ν is a positive measure supported on K
and kf kLp(X) ≤ 1. Thus, as in [29, Theorem 14] or [3, Theorem 3.6.1,
p.115], the equality (3.3) follows with the use of the Minimax Theorem
(see [10]) and for general sets A by a capacitability argument.
4. Capacity estimates and Hausdorff measure
In this section, we define a Hausdorff measure based on the upper
bound estimate for the capacity of balls. We prove an upper bound
estimate for the capacity of a set E ⊂ X in terms of this modified
Hausdorff content. We also show that the Hausdorff content, satisfying
a condition placed by γp, is zero if the capacity of the set is zero. For
the latter result, we assume that our space X satisfies inequality (4.1)
which holds in connected spaces. Similar results for compact sets can
be found in [31, Theorem 2.2], where the assumption of connectedness
is replaced by a density condition that gives an inequality equivalent
to (4.1). In this section, we give direct and short proofs to results that
apply not only for compact sets. In particular, we do not need to use
a version of Frostman’s lemma in our proofs. Also, unlike in [31], we
do not assume our space to be complete. We start by showing that the
Riesz capacity of a ball is bounded from above by a constant times the
measure of the ball to the power 1 − γp.
Lemma 4.1. Let 1 < p < ∞ and 0 < γ < 1 be such that γp < 1. Then
Proof. Choose
Cγ,p (B(x, r)) ≤ C µ (B(x, r))1−γp .
g = c2 3Q(1−γ)
χB(x,r)
µ (B(x, r))γ ,
where c > 0 and Q > 0 are some constants, for which the inequality
(2.2) holds. For each z ∈ B(x, r), we have
Iγg(z) =
c2 3Q(1−γ)
µ (B(x, r))γ ZB(x,r)
1
µ (B(z, d(z, y)))1−γ dµ(y).
14
JUHO NUUTINEN AND PILAR SILVESTRE
For each y ∈ B(x, r), we have that d(z, y) ≤ 2r, since z ∈ B(x, r). Now
µ (B(z, d(z, y))) ≤ µ (B(z, 2r)) ≤ µ (B(x, 3r)) ,
for each y ∈ B(x, r). Then
Iγg(z) ≥
c2 3Q(1−γ)
µ (B(x, r))γ ZB(x,r)
1
µ (B(x, 3r))1−γ dµ(y)
= c2 3Q(1−γ) ·
≥ c2 3Q(1−γ) ·
µ (B(x, r))1−γ
µ (B(x, 3r))1−γ
1
c2 ·(cid:16) r
3r(cid:17)Q(1−γ)
= 1 ,
where the last inequality follows by (2.2). Thus g is admissible and we
get the upper bound
Cγ,p (B(x, r)) ≤ gp
Lp(X)
= c2p 3Q(1−γ)p µ (B(x, r))
µ (B(x, r))γp
= C µ (B(x, r))1−γp .
(cid:3)
Lemma 4.1 leads us to define a modified version of the Hausdorff
measure that works in our generality. In this section, (and throughout
the paper) we do not assume the doubling measure µ to satisfy the
regularity (2.6) or any other estimates that would give uniform lower
bounds or upper bounds for the measure of balls in terms of the radii.
Recall that the usual definition for the λ-Hausdorff content of a set
E ⊂ X, for 0 < r ≤ ∞, is
Hλ
r (E) = inf(cid:8) ∞Xi=1
rλ
i : E ⊂
∞[i=1
B(xi, ri), xi ∈ E, ri ≤ r(cid:9),
and the λ-Hausdorff measure of E is Hλ(E) = limr→0 Hλ
Hausdorff dimension of E is the number
r (E). The
dim(E) = inf(cid:8)λ > 0 : Hλ(E) = 0(cid:9) .
Let 1 < p < ∞, 0 < γ < 1 and γp < 1. In our case, we define the
Hausdorff content of a set E ⊂ X, for 0 < r ≤ ∞, as
r (E) = inf(cid:8) ∞Xi=1
eHγ,p
µ(B(xi, ri))1−γp : E ⊂
∞[i=1
B(xi, ri), xi ∈ E, ri ≤ r(cid:9).
Then, the Hausdorff measure is
eHγ,p(E) = lim
r→0 eHγ,p
r (E).
THE RIESZ CAPACITY IN METRIC SPACES
15
Note that if the measure is Q-regular, then
r ≈ HQ(1−γp)
r
.
eHγ,p
In the next theorem, we show that the Riesz capacity of a set E ⊂ X
is bounded from above by a constant times the (modified) Hausdorff
content of the set E. In particular, this implies that compact sets with
positive capacity have positive Hausdorff measure (see [31, Theorem
2.2]).
Theorem 4.2. Let 1 < p < ∞ and 0 < γ < 1 be such that γp < 1.
∞ (E) for each E ⊂ X, where C is the same
constant as in Lemma 4.1.
Then Cγ,p(E) ≤ C eHγ,p
Proof. Suppose that eHγ,p
ǫ > 0, there is a countable covering {B(xi, ri)} of E such that
∞ (E) < ∞, otherwise the claim is obvious. For
∞Xi=1
µ(B(xi, ri))1−γp < eHγ,p
∞ (E) + ǫ.
Now, by the monotonicity and Theorem 4.1
Cγ,p(E) ≤
Cγ,p(B(xi, ri))
≤ C
∞Xi=1
∞Xi=1
< C(eHγ,p
µ(B(xi, ri))1−γp
∞ (E) + ǫ).
The claim follows by letting ǫ → 0.
(cid:3)
For the proof of the next theorem, we need an opposite inequality to
(2.2) which is true in connected spaces. Indeed, if X is connected then
by [5, Corollary 3.8] there exist constants C > 0 and s > 0 such that
for all balls B(y, R) in X, all z ∈ B(y, R) and all 0 < r ≤ R,
(4.1)
µ(B(z, r))
µ(B(y, R))
R(cid:17)s
≤ C(cid:16) r
.
Note that inequality (4.1) given by the connectedness (or uniform per-
fectness) of the space X is equivalent to the density condition assumed
in [31]. The proof of the next theorem is direct and we do not need
to use Frostman’s lemma to obtain the result. In the Euclidean space,
with the Lebesgue measure, we use the notation α = γn and consider
the usual Riesz potential of order α from Remark 2.2 (ii). Our result
implies the classical result that if the Riesz capacity of a set E is zero,
then E has Hausdorff dimension at most n − αp, where αp < n (see
e.g. [30, Section 5.2, Theorem 2.3]).
16
JUHO NUUTINEN AND PILAR SILVESTRE
Theorem 4.3. Assume that X satisfies (4.1). Let 1 < p < ∞, 1 <
p < ∞, 0 < γ < 1 and 0 < γ < 1 be such that γp < 1, γ p < 1 and
∞ (E) = 0.
γ p < γp. If Cγ,p(E) = 0, then eHγ,p
that E = S∞
Proof. In the following, we prove the result for bounded sets E ⊂ X.
If the set E is not bounded, then there exists bounded sets Ej such
j=1 Ej. We then use the countable subadditivity of the
Hausdorff content, the monotonicity of the Riez capacity and the result
for bounded sets to obtain the result for unbounded sets E.
Let E ⊂ X be a bounded set. For ǫ > 0, there is an admissible
function f ≥ 0 such that
f p
Lp(X) < ǫ.
For such a function f , at each point x ∈ E,
1 ≤ZX
f (y)
µ (B(x, d(x, y)))1−γ dµ(y).
Let x0 ∈ E. We choose R0 > diam(E) large enough such that E ⊂
B(x0, R0) and that the integral below is more than one half. Notice
that we can always find such a radius R0 but the selection depends on
the set E. Define R = 2R0 and ri = 2−iR, for i ∈ N. For each point
x ∈ E
1
2
≤ZB(x0,R0)
≤ZB(x,R)
f (y)
µ (B(x, d(x, y)))1−γ dµ(y)
f (y)
µ (B(x, d(x, y)))1−γ dµ(y)
and hence
1 ≤ 2ZB(x,R)
f (y)
µ (B(x, d(x, y)))1−γ dµ(y)
f (y)
= 2
≤ 2
≤ 2
∞Xi=0 ZB(x,ri)\B(x,ri+1)
∞Xi=0 ZB(x,ri)\B(x,ri+1)
µ (B(x, ri+1))1−γ ZB(x,ri)
∞Xi=0
1
µ (B(x, d(x, y)))1−γ dµ(y)
f (y)
µ (B(x, ri+1))1−γ dµ(y)
f (y) dµ(y).
Using Holder’s inequality and the doubling condition, we get
1 ≤ C
∞Xi=0
µ (B(x, ri))γ−1+ p−1
p (cid:16)ZB(x,ri)
f (y)p dµ(y)(cid:17)1/p
.
THE RIESZ CAPACITY IN METRIC SPACES
17
Next, we use the fact that for any δ > 0 there is a constant C > 0 such
that
Then, inequality (4.1) gives
2−iδ = C
1 = C
∞Xi=0
(cid:18) µ (B(x, ri))
µ(B(x0, R))(cid:19)δ/s
.
R(cid:17)δ
∞Xi=0 (cid:16) ri
R(cid:17)δ
∞Xi=0 (cid:16) ri
≤ C
C
∞Xi=0
= 1.
Now, by putting the measure of the ball B(x0, R) as part of the con-
stant, we have that
C
∞Xi=0
µ(B(x, ri))δ/s ≤
∞Xi=0
µ (B(x, ri))γ−1+ p−1
p (cid:16)ZB(x,ri)
f (y)p dµ(y)(cid:17)1/p
,
where the constant C depends on R. For δ > 0, there exists at least
one index ix ∈ N such that
µ (B(x, rix))γ−1+ p−1
p (cid:16)ZB(x,rix )
f (y)p dµ(y)(cid:17)1/p
and, by raising both sides to the power p, we get
≥ Cµ(B(x, rix))δ/s
ZB(x,rix )
f (y)p dµ(y) ≥ Cµ(B(x, rix))δp/s−(γ−1)p−p+1
= Cµ(B(x, rix))δp/s−γp+1.
We choose
δ =
γp − γ p
p
· s,
which is positive, as γp > γ p. We obtain for each x ∈ E a ball
B(x, rix) = Bx such that
(4.2)
µ(Bx)1−γ p ≤ CZBx
f (y)p dµ(y).
By using the basic 5r-covering theorem (see e.g. [18]), we obtain count-
ably many points xj ∈ E, such that the balls Bj = Bxj are pairwise
j=1 5Bj. Using the estimate (4.2), the doubling
property of the measure µ and the pairwise disjointness of the balls Bj,
we get
disjoint and E ⊂ S∞
∞Xj=1
∞Xj=1ZBj
eHγ,p
∞ (E) ≤
≤ C
µ(5Bj)1−γ p ≤ C
µ(Bj)1−γ p
∞Xj=1
f (y)p dµ(y) ≤ CZX
f (y)p dµ(y)
= C f p
Lp(X) < C ǫ.
Letting ǫ → 0 yields the claim.
(cid:3)
18
JUHO NUUTINEN AND PILAR SILVESTRE
We can see that in the metric space, with a doubling measure µ, the
Hausdorff content and the Hausdorff measure have the same null sets.
This relation has been studied in [22, Section 7] for slightly different
versions of the Hausdorff content and the Hausdorff measure. For our
definitions, the result of [22, Lemma 7.6] follows without any extra
assumptions. By the previous two theorems, we get as a corollary the
following result for arbitrary sets E.
Corollary 4.4. Let 1 < p < ∞, 1 < p < ∞, 0 < γ < 1 and 0 < γ < 1
be such that γp < 1, γ p < 1 and γ p < γp. Then
If we also assume our space X to be connected, then
Cγ,p(E) > 0 implies that eHγ,p(E) > 0.
eHγ,p(E) > 0 implies that Cγ,p(E) > 0.
References
[1] D.R. Adams, Choquet integrals in potential theory, Publ. Mat. 42 (1998),
3–66.
[2] D.R. Adams and L.I. Hedberg, Function Spaces and Potential Theory,
Springer-Verlag (1996).
[3] H. Aikawa, M.R. Ess´en, Potential theory: selected topics no.1633, Springer
(1996).
[4] S. Banach, S. Saks, Sur la convergence forte dans les champs Lp, Studia
Mathematica 2 (1930) 51–57.
[5] A. Bjorn and J. Bjorn, Nonlinear potential theory on metric spaces, EMS
Tracts in Mathematics 17, European Mathematical Society (EMS), Zurich
(2011).
[6] C. Bennett and B. Sharpley, Interpolation of Operators, Academic Press Inc.
(1988).
[7] J. Cerd`a, Lorentz capacitary spaces, AMS, Contemporary Mathematics 445
(2007), 49-55.
[8] J. Cerd`a, J. Mart´ın and P. Silvestre, Capacitary function spaces, Collectanea
Math. 62 (2011), 95–118.
[9] G. Choquet, Theory of capacities, Ann. Inst. Fourier (Grenoble) 5 (1953-54),
131–295.
[10] K. Fan, Minimax theorems, Proc. Nat. Acad. Sci. U.S.A. 39 (1953), 42–47.
[11] R.A. Fefferman, A theory of entropy in Fourier analysis, Adv. in Math. 30
(1978), 171–201.
[12] B. Fuglede, On the theory of potentials in locally compact spaces, Acta. Math.
103 (1960), 139–215.
[13] A.E. Gatto and S. V´agi, Fractional integrals on spaces of homogeneous type,
Analysis and Partial Differential Equations, C. Sadosky (ed.), Dekker (1990),
171–216.
[14] A.E. Gatto, C. Segovia and S. V´agi, On fractional differentiation and inte-
gration on spaces of homogeneous type, Rev. Mat. Iberoamericana 12 (1996),
111–145.
[15] V. Gol’dshtein and M. Troyanov, Capacities in metric spaces, Integral Equ.
Oper. Theory 44 (2002), 212–242.
[16] P. Haj lasz and P. Koskela, Sobolev met Poincar´e, Mem. Amer. Math. Soc.
145 (2000).
THE RIESZ CAPACITY IN METRIC SPACES
19
[17] H. Hakkarainen and J. Kinnunen, The BV-capacity in metric spaces,
Manuscripta Math. 132 (2010), 369–390.
[18] J. Heinonen, Lectures on analysis on metric spaces, Universitext, Springer-
Verlag New York (2001).
[19] J. Heinonen, P. Koskela, N. Shanmugalingam and J. Tyson, Sobolev spaces
on metric measure spaces: an approach based on upper gradients, New Math-
ematical Monographs. Cambridge University Press (2015).
[20] A. Kairema, Sharp weighted bounds for fractional integral operators in a space
of homogeneous type, Mathematica Scandinavica 114 (2014) no 2, 226–253.
[21] S. Kallunki and N. Shanmugalingam, Modulus and continuous capacity. Ann.
Acad. Sci. Fenn. Math. 26 (2001), 455–464.
[22] J. Kinnunen, R. Korte, N. Shanmugalingman and H. Tuominen, Lebesgue
points and capacities via the boxing inequality in metric spaces. Indiana Univ.
Math. J. 57 (2008), 401–430.
[23] J. Kinnunen and O. Martio, The Sobolev capacity on metric spaces, Ann.
Acad. Sci. Fenn. Math. 21 (1996), 367–382.
[24] J. Kinnunen and O. Martio, Choquet property for the Sobolev capacity in
metric spaces, In: Proceedings on Analysis and Geometry (Russian) (Novosi-
birsk Akademgorodok, 1999), 285–290. Izdat. Ross. Akad. Nauk Sib. Otd.
Inst. Math., Novosibirsk (2000)
[25] J. Lehrback, Neighbourhood capacities, Ann. Acad. Sci. Fenn. Math. 37
(2012), 35–51.
[26] J. Mal´y, Coarea integration in metric spaces, Nonlinear Analysis, Function
Spaces and Applications, Proceedings of the Spring School held in Prague,
July 17-22, 2002, Czech Academy of Sciences, Mathematical Institute, Praha,
7 (2003), 149–192.
[27] J. Mal´y and L. Pick, The sharp Riesz potential estimates in metric spaces,
Indiana Univ. Math. J., 51 (2002), 251–268.
[28] V.G. Maz’ya, Sobolev Spaces (translated from the Russian by T. O. Shaposh-
nikova), Springer-Verlag (1985).
[29] N.G. Meyers, A theory of capacities for potentials of functions in Lebesgue
classes, Math. Scand., 26 (1970), 255–292.
[30] Y. Mizuta, Potential Theory in Euclidean Spaces, Gakuto International Series
Mathematical Sciences and Applications Volume 6 (1996).
[31] T. Sjodin, A Note on Capacity and Hausdorff Measure in Homogeneous
Spaces, Potential Analysis 6 (1997), 87–97.
[32] T. Sjodin, Polar sets and capacitary potentials in homogeneous spaces, Ann.
Acad. Sci. Fenn. Math. 38 (2013), 771–783.
J.N., Department of Mathematics and Statistics, P.O. Box 35,
FI-40014 University of Jyvaskyla, Finland
[email protected]
P.S., Department of Mathematics, P.O. Box 11100, FI-00076 Aalto
University, Finland
[email protected]
|
1103.1032 | 1 | 1103 | 2011-03-05T09:38:17 | Subharmonicity of the modulus of quasiregular harmonic mappings | [
"math.FA"
] | In this note we determine all numbers $q\in \mathbf R$ such that $|u|^q$ is a subharmonic function, provided that $u$ is a $K-$quasiregular harmonic mappings in an open subset $\Omega$ of the Euclidean space $\mathbf R^n$. | math.FA | math |
SUBHARMONICITY OF THE MODULUS OF
QUASIREGULAR HARMONIC MAPPINGS
DAVID KALAJ AND VESNA MANOJLOVI ´C
Abstract. In this note we determine all numbers q ∈ R such that
uq is a subharmonic function, provided that u is a K−quasiregular
harmonic mappings in an open subset Ω of the Euclidean space Rn.
1. Introduction
By · we denote the Euclidean norm in Rn and let Ω be a region in Rn. In
this paper we consider K-quasiregular harmonic mappings, where K > 1.
We recall that a harmonic mapping u(x) = (u1(x), . . . , un(x)) : Ω → Rn
with formal differential matrix
is K-quasiregular if
Du(x) = {∂iuj(x)}n
i,j=1
(1.1)
K −1Du(x)n 6 Ju(x) 6 Kl(Du(x))n,
for all x ∈ Ω,
where Ju is the Jacobian of u at x,
and
Du := max{Du(x)h : h = 1},
l(Du) := min{Du(x)h : h = 1}.
See [7, p. 128] for the definition of quasiregular mappings in more general
setting. A quasiregular homeomorphism is called quasiconformal.
Let 0 < λ2
1 6 λ2
2 6 . . . 6 λ2
n be the eigenvalues of the matrix Du(x)Du(x)t.
Here Du(x)t is the transpose of the matrix Du(x). Then
(1.2)
(1.3)
and
(1.4)
n
λk,
Ju(x) =
Yk=1
Du = λn
l(Du) = λ1.
For the Hilbert-Schmidt norm of the matrix Du(x), defined by
kDu(x)k =pTrace(Du(x)Du(x)t)
1991 Mathematics Subject Classification. Primary 31A05; Secondary 31B10.
Key words and phrases. Quasiregular harmonic mappings, Subharmonic functions.
1
DAVID KALAJ AND VESNA MANOJLOVI ´C
2
we have
(1.5)
and
(1.6)
n
∂u
∂xk
∂u
∂xk •
kDu(x)k =vuut
Xk=1
kDu(x) =vuut
Xk=1
=vuut
2
n
Xk=1(cid:12)(cid:12)(cid:12)(cid:12)
∂u
∂xk(cid:12)(cid:12)(cid:12)(cid:12)
λ2
k.
n
Here • denotes the inner product between vectors. From (1.1), for a quasireg-
ular mapping we have
(1.7)
λn
λk
,
λk
λ1
6 K, k = 1, . . . , n.
It is well known that if u = (u1, . . . , un) is a harmonic mapping defined in
a region Ω of the Euclidean space Rn, then up is subharmonic for p > 1,
and that, in the general case, is not subharmonic for p < 1. Let us prove
this well-known fact. If u is harmonic, then by a result in [4, Lemma 1.4.]
(see also [3, Eq. (4.9)-(4.11)])
2
.
∆u = u(cid:13)(cid:13)(cid:13)(cid:13)
u(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
D(cid:18) u
If u(a) = 0, then
So ∆u > 0 for those points x, such that u(x) 6= 0.
we consider the harmonic mapping um(x) = u(x) + (1/m, 0, . . . , 0). Then
um(a) 6= 0, and ∆um(x) > 0 in some neighborhood of a. It follows from the
definition of subharmonic functions that the uniform limit of a convergent
sequence of subharmonic functions is still subharmonic. Since um(x) →
u(x), it follows that u is subharmonic in a. Since the function g(s) = sp,
is convex for p > 1, we obtain that up is subharmonic providing that u is
harmonic. (For the above facts we refer to [2, Chapter 2]).
Recently, several authors have proved the following two propositions,
which is the motivation for our study.
Proposition 1.1. [5] If f is a K-quasiregular harmonic map in a plane
domain, then fq is subharmonic for q > 1 − K −2.
Proposition 1.2. [1] If f is a K-quasiregular harmonic map in a space
domain, then fq is subharmonic for some q = q(K, n) ∈ (0, 1).
This paper is continuation of [1] in which Proposition 1.1 was extended to
the n-dimensional setting. In [1] the authors prove only the existence of an
exponent q ∈ (0, 1) without giving the minimal value of q. Here we improve
Proposition 1.2 by giving the optimal value of q. Our proof is completely
different from those given in [1] and [5]. Moreover for the first time we
consider the case q < 0.
Our proof is based on the following well-known explicit computation.
SUBHARMONICITY AND QUASIREGULAR HARMONIC MAPPINGS
3
Proposition 1.3. [6, Ch. VII 3, p.217]. Let u = (u1, . . . , un) : Ω → Rn, be
harmonic, let Ω0 = Ω \ u−1(0), let q ∈ R. Then for x ∈ Ω0
Xk=1(cid:18)u •
∆uq = q"uq−2
∇uk2 + (q − 2)uq−4
∂xk(cid:19)# .
Xk=1
1 +···+u2
Proof. Write v := uq = (u2
gives
n)p, for p := q/2 . A direct computation
∂u
n
n
vx1 = p(u2
= q(u2
1 + ··· + u2
1 + ··· + u2
n)p−1 · (2u1u1x1 + ··· + 2ununx1)
n)p−1 · (u1u1x1 + ··· + ununx1),
and further
vx1x1 = q{2(p − 1)(u2
1 + ··· + u2
+ (u2
n)p−2 · (u1u1x1 + ··· + unx1)2+
1 + ··· + u2
n)p−1 · [u1u1x1x1 + (u1x1)2 + ··· + ununx1x1 + (unx1)2]}.
Therefore
∆v = vx1x1 + ··· + vxnxn
n
2
n
2
n
+
2
}
2
n
n
n
u1
un
2
xk + . . .
xk )] + (q − 2)uq−4
Xk=1
= q{uq−2[(u1∆u1 + ··· + un∆un) + (
ujuj xk
Xk=1
Xj=1
xk ) + (q − 2)uq−4
Xk=1
Xk=1
= q{uq−2(
Xj=1 n
Xk=1
= quq−4{u2
Xj=1
∇uj2 + (q − 2)
Xk=1
xk! + (q − 2)
Xk=1
Xk=1(cid:18)u •
Xj=1
uj ·
∂xk(cid:19)2
= quq−4{u2
xk + ··· +
}.
∂u
un
u1
2
uj
n
n
n
n
n
2. Main result
n
n
Xk=1
Xj=1
∂xk
}
∂uj
2
uj ·
2
}
∂uj
∂xk
(cid:3)
Theorem 2.1. Let u be K−quasiregular harmonic in Ω ⊂ Rn. Then the
mapping g(x) = u(x)q is subharmonic in
(1) Ω for q > max{1 − n−1
(2) Ω \ u−1(0), for q 6 1 − (n − 1)K 2.
Moreover for 1− (n − 1)K 2 < q < 1− n−1
harmonic mapping such that uq is not subharmonic.
K 2 , 0};
K 2 , there exists a K-quasiconformal
4
DAVID KALAJ AND VESNA MANOJLOVI ´C
Remark 2.2. If n = 2 then 1 − n−1
K 2 = 1 − K −2. Thus Theorem 2.1 is an
extension of Proposition 1.1.
Remark 2.3. In the case 1 6 K 6 √n − 1 the function uq is subharmonic
for all q > 0.
Proof of Theorem 2.1. Let us fix such a map u : Ω → Rn and set Ω0 =
Ω \ u−1{0}. We have to find all positive real numbers q such that ∆uq > 0
on Ω0. Since u is quasiregular, the set Z = {x ∈ Ω0 : det Du(x) = 0} has
In particular,
measure zero (see [7]), it is also closed since u is smooth.
Ω1 = Ω0 \ Z is dense in Ω0 and thus it suffices to prove that ∆uq > 0 on
Ω1. From Proposition 1.3, we obtain
So we find all real q such that
After normalization, we see that it suffices to find all constants q < 2 such
Let 0 < λ2
1 6 λ2
2 6 . . . 6 λ2
n be the eigenvalues of the matrix Du(x)Du(x)t.
If q > 2, then ∆uq > 0. Assume that q > 0 and q < 2 such that
n
n
n
1
> 0.
2
6
x ∈ Ω1.
∆uq = q
2 − qu(x)2kDu(x)k2,
2
uj∇uj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
uq−2kDuk2 + (q − 2)uq−4(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
.
2
uj∇uj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
uq−2kDuk2 + (q − 2)uq−4(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
q
Xj=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
uj(x)∇uj(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
z∈Sn−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
zj∇uj(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
z∈Sn−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
z∈Sn−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2 − qkDu(x)k2,
x ∈ Ω1.
= λ2
n
n
Xj=1
= λ2
1
2
2
2
6
1
n
sup
n
sup
inf
(2.1)
that
(2.2)
Then
(2.3)
(2.4)
and
(2.5)
zj∇uj(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
zj∇uj(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=1
n
kDu(x)k2 =
λ2
k.
SUBHARMONICITY AND QUASIREGULAR HARMONIC MAPPINGS
5
Because u is K−quasiregular from (1.7) we have
(2.6)
6 K, k = 1, . . . , n − 1.
λn
λk
Thus (2.2) can be written as
(2.7)
λ2
n 6
1
2 − q
λ2
k.
n
Xk=1
By (2.5) and (2.6) we get that, the inequality (2.7) is satisfied whenever
(2.8)
i.e.
(2.9)
(2.10)
i.e.
If q < 0, then we should have
max(cid:26)0, 1 −
zj∇uj(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
n
inf
z∈Sn−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
1
1 + n−1
K 2
6
1
2 − q
K 2 (cid:27) 6 q < 2.
n − 1
>
1
2 − qkDu(x)k2,
x ∈ Ω1,
Xk=1
Because u is K−quasiregular from (1.7)
(2.11)
2 − q >
λk
λ1
6 K, k = 2, . . . , n.
n
λ2
k
λ2
1
.
Thus if
(2.12)
q 6 1 − (n − 1)K 2,
then (2.10) holds. To finish the proof we need the following lemma.
Lemma 2.4. For any 1 − (n − 1)K 2 < q < 1 − n−1
there is a (linear)
harmonic K-quasiconformal mapping u such that uq is not subharmonic.
Proof of Lemma 2.4. Assume first that q > 0. We will consider linear map-
ping u : Rn −→ Rn defined by
(2.13)
where K > 1. It is obviously harmonic and K−quasiconformal. If we put
this mapping in formula (2.1) we get
u(x1, . . . , xn−1, xn) = (x1, . . . , xn−1, Kxn),
K 2
[(n − 1) + K 2]u2 + (q − 2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
n−1
Xj=1
2
xjej + K 2enxn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≥ 0
6
DAVID KALAJ AND VESNA MANOJLOVI ´C
which is equivalent to
(n − 1 + K 2)"j−1
Xn=1
n−1
x2
j + K 2x2
n# + (q − 2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
(n − 1 + K 2)K 2 ≥ (2 − q)K 4
xjej + K 2enxn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
≥ 0.
By choosing x1 = ··· = xn−1 = 0 and xn = 1, we obtain
which is equivalent to
q ≥ 1 −
n − 1
K 2 .
For q < 0 we consider the linear mapping u : Rn −→ Rn defined by
(2.14)
u(x1, . . . , xn−1, xn) = (x1, . . . , xn−1, xn/K).
To finish the proof we only need to take u = uΩ, where u is defined in
(2.13) respectively in (2.14).
(cid:3)
(cid:3)
Acknowledgment. We thank the referee for providing constructive com-
ments and help in improving the contents of this paper.
References
[1] Arsenovi´c, M., Bozin, V. and Manojlovi´c, V. : Moduli of continuity of harmonic
quasiregular mappings in Bn: Potential Analysis DOI: 10.1007/s11118-010-9195-8.
[2] Hayman, W. K., Kennedy, P. B.: Subharmonic functions, I Academic Press,
London-New York, 1976, xvii+284 pp.
[3] Kalaj, D.: A priori estimate of gradient of a solution to certain differential inequality
and quasiconformal mappings. arXiv:0712.3580v3. 24 pp.
[4] Kalaj, D.: On the univalent solution of PDE ∆u = f between spherical annuli. J.
Math. Anal. Appl. 327 (2007), no. 1, 1 -- 11.
[5] Koji´c, V., Pavlovi´c, M.: Subharmonicity of f p for quasiregular harmonic functions
with applications. J. Math. Anal. Appl. 342 (2008), 742 -- 746 .
[6] Stein, E.M.: Singular integrals and differentiability properties of functions. Princeton
Mathematical Series, No. 30 Princeton University Press, Princeton, N.J. 1970 xiv+290
pp.
[7] Vuorinen, M.: Conformal geometry and quasiregular mappings. Lecture Notes in
Math., 1319. Springer, Berlin (1988).
University of Montenegro, Faculty of Natural Sciences and Mathematics,
Cetinjski put b.b. 8100 Podgorica, Montenegro
E-mail address: [email protected]
University of Belgrade, Faculty of Organizational Sciences, Jove Ili´ca
154, Belgrade, Serbia
E-mail address: [email protected]
|
1209.1822 | 1 | 1209 | 2012-09-09T17:32:17 | Analytic and Group-Theoretic Aspects of the Cosine Transform | [
"math.FA"
] | This is a brief survey of recent results by the authors devoted to one of the most important operators of integral geometry. Basic facts about the analytic family of cosine transforms on the unit sphere and the corresponding Funk transform are extended to the "higher-rank" case for functions on Stiefel and Grassmann manifolds. The main topics are the analytic continuation and the structure of polar sets, the connection with the Fourier transform on the space of rectangular matrices, inversion formulas and spectral analysis, and the group-theoretic realization as an intertwining operator between generalized principal series representations of SL(n, R). | math.FA | math |
Analytic and Group-Theoretic Aspects of the Cosine
Transform
G. ´Olafsson, A. Pasquale, and B. Rubin
Abstract. This is a brief survey of recent results by the authors devoted to
one of the most important operators of integral geometry. Basic facts about
the analytic family of cosine transforms on the unit sphere in Rn and the corre-
sponding Funk transform are extended to the "higher-rank" case for functions
on Stiefel and Grassmann manifolds. The main topics are the analytic continu-
ation and the structure of polar sets, the connection with the Fourier transform
on the space of rectangular matrices, inversion formulas and spectral analy-
sis, and the group-theoretic realization as an intertwining operator between
generalized principal series representations of SL(n, R).
1. Introduction
The cosine transform has a long and rich history, with connections to several
branches of mathematics. The name cosine transform was adopted by Lutwak
[50, p. 385] for the spherical convolution which is defined on the unit sphere Sn−1
in Rn by
f (v)u · v dv,
u ∈ Sn−1 .
(1.1)
(Cf )(u) =ZSn−1
The motivation for this name is that the inner product u · v is nothing but the
cosine of the angle between the unit vectors u and v.
The following list of references shows some branches of mathematics, where the
operator (1.1) and its generalizations arise in a natural way (sometimes implicitly,
without naming) and play an important role.
• Convex geometry: [1, 6, 23, 24, 32, 46, 50, 69, 71, 75].
• Pseudo-Differential Operators: [15, 61].
• Group representations: [2, 3, 11, 12, 57, 60].
• Harmonic Analysis and Singular Integrals: [4, 21, 22, 27, 48, 52, 58, 59, 63,
66, 73, 74, 79].
• Integral geometry: [5, 20, 26, 30, 62, 64, 65, 68, 70, 76, 86].
• Stochastic Geometry and Probability: [29, 49, 51, 77, 78].
The authors are thankful to Tufts University for the hospitality and support during the Joint
AMS meeting and the Workshop on Geometric Analysis on Euclidean and Homogeneous Spaces
in January, 2012. The research of G. ´Olafsson was supported by DMS-0801010 and DMS-1101337.
A. Pasquale gratefully acknowledges travel support from the Commission de Colloques et Congr`es
Internationaux (CCCI).
1
2
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
• Banach Space Theory: [38, 45, 47, 54, 72].
This list is far from being complete.
In most of the publications cosine-like
transforms serve as a tool for certain specific problems. At the same time, there are
many papers devoted to the cosine transforms themselves. The present article is
just of this kind. Our aim is to give a short overview of our recent work [57, 70] on
the cosine transform and explain some of the ideas and tools behind those results.
For a complex number λ, the λ-analogue of the operator (1.1) is the convolution
operator
(Cλf )(u) =ZSn−1
f (v)u · vλ dv,
u ∈ Sn−1,
(1.2)
where the integral is understood in the sense of analytic continuation, if necessary.
We adopt the name "the cosine transform" for (1.2) too. The same name will be
used for generalizations of these operators to be defined below.
In recent years more general, higher-rank cosine transforms attracted consid-
erable attention. This class of operators was inspired by Matheron's injectivity
conjecture [51], its disproval by Goodey and Howard [29], applications in group
representations [7, 12, 57, 60, 86] and in algebraic integral geometry [3, 5, 20].
To the best of our knowledge, the higher-rank cosine transform was explicitly pre-
sented (without naming) for the first time in [26, formula (3.5)].
As mentioned above, the present article gives a brief survey of recent results
by the authors [57, 70] in this area. The consideration grew up from specific
problems of harmonic analysis and group representations. However, we do not
focus on those problems, and mention them only for better explanation of the
corresponding properties of the cosine transforms and related operators of integral
geometry. Here we shall restrict ourselves to the case of real numbers, referring to
the above articles for the case of complex and quaternionic fields.
The paper is organized as follows. Section 2 contains basic facts about the
cosine transforms on the unit sphere. More general higher-rank transforms on
Stiefel or Grassmann manifolds are considered in Section 3, where the main tool
is the classical Fourier analysis. In Sections 4 and 5 we discuss the connections to
representation theory, and more precisely to the spherical representations and the
intertwining properties. Section 6 is devoted to explicit spectral formulas for the
cosine transforms.
2. Cosine transforms on the unit sphere
In this section we discuss briefly the cosine transform on the sphere Sn−1. We keep
the notation from the Introduction. For the analytic continuation of the cosine
transform it is convenient to normalize it by setting
(C λf )(u) = γn(λ)ZSn−1
f (v)u · vλ dv,
u ∈ Sn−1.
Here dv stands for the SO(n)-invariant probability measure on Sn−1 and the nor-
malizing coefficient γn(λ) is given by
γn(λ) =
π1/2 Γ(−λ/2)
Γ(n/2) Γ((1 + λ)/2)
,
Re λ > −1, λ 6= 0, 2, 4, . . . .
(2.1)
THE COSINE TRANSFORM
3
This normalization is chosen so that
C λ(1) =
Γ (−λ/2)
Γ((n + λ)/2)
.
Such a normalization is convenient in many occurrences, when harmonic analysis on
the sphere is performed in the multiplier language (in the same manner as analysis
of pseudo-differential operators is performed in the language of their symbols). We
shall see below that it also simplifies the spectrum of the cosine transform.
The limit case λ = −1 gives, up to a constant, the well-known Funk transform.
Specifically, if f ∈ C(Sn−1), then for every u ∈ Sn−1,
where
(C λf )(u) =
lim
λ→−1
π1/2
Γ((n − 1)/2)
(F f )(u),
(F f )(u) =
Z
f (v) duv .
{v∈Sn−1u·v=0}
(2.2)
(2.3)
In (2.3), duv stands for the rotational invariant probability measure on the (n − 2)-
dimensional sphere u · v = 0; see, e.g., [69, Lemma 3.1].
The operators Cλ and C λ were investigated by different approaches. A first one
employs the Fourier transform technique [45, 63, 76] and relies on the equality in
the sense of distributions
(cid:18) Eλ Cλf
Γ((1 + λ)/2)
, F ω(cid:19) = c1 (cid:18) E−λ−nf
Γ(−λ/2)
, ω(cid:19) ,
(2.4)
c1 = 2n+λ π(n−1)/2 Γ(n/2).
Here ω is a test function belonging to the Schwartz space S(Rn),
(F ω)(y) =ZRn
ω(x)eix·ydx,
and (Eλf )(x) = xλf (x/x) denotes the extension by homogeneity.
A second approach is based on the Funk-Hecke formula, so that for each spher-
ical harmonic Yj of degree j,
where
mj,λ =
C λYj = mj,λ Yj,
(−1)j/2
Γ(j/2 − λ/2)
Γ(j/2 + (n + λ)/2)
0
if j is even,
if j is odd;
(2.5)
(2.6)
see, e.g., [63]. The Fourier-Laplace multiplier {mj,λ} forms the spectrum of C λ.
Note that the normalizing coefficient in C λ was chosen so that only factors depend-
ing on j are involved in the spectral functions {mj,λ}. The spectrum of C λ encodes
important information about this operator. For instance, since mj,λmj,−λ−n = 1,
then for any f ∈ C∞(Sn−1) the following inversion formula holds:
C −λ−nC λf = f,
(2.7)
provided
λ ∈ C,
λ /∈ {−n, −n − 2, −n − 4, . . .} ∪ {0, 2, 4, . . .}.
provided
γ = Re λ +
n + 1
2
1
p
−
−(cid:12)(cid:12)(cid:12)
1
2(cid:12)(cid:12)(cid:12)(n − 1),
δ = Re λ +
n + 1
2
(2.9)
1
2(cid:12)(cid:12)(cid:12)(n − 1),
1
p
−
+(cid:12)(cid:12)(cid:12)
4
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
For the non-normalized transforms, (2.7) yields
C−λ−nCλf = ζ(λ) f,
ζ(λ) =
Γ2(n/2) Γ((1 + λ)/2) Γ((1 − λ − n)/2)
π Γ(−λ/2) Γ((n + λ)/2)
,
(2.8)
λ ∈ C,
λ /∈ {−1, −3, −5, . . .} ∪ {1 − n, 3 − n, 5 − n, . . .}.
Formula (2.6) reveals singularities, provides information about the kernel and
the image. Moreover, it plays a crucial role in the study of cosine transforms on
Lp functions. For instance, the following statement was proved in [64, p.11], using
the relevant results of Gadzhiev [21, 22] and Kryuchkov [48] for symbols of the
Calderon-Zygmund singular integrals operators.
Theorem 2.1. Let Lp
e(Sn−1) and Lγ
distributions), belonging to Lp(Sn−1) and the Sobolev space Lγ
Then
p,e(Sn−1) be the spaces of even functions (or
p(Sn−1), respectively.
p,e(Sn−1) ⊂ C λ(Lp
Lδ
e(Sn−1)) ⊂ Lγ
p,e(Sn−1)
λ /∈ {0, 2, 4, . . . } ∪ {−n − 1, −n − 3, −n − 5, . . . }.
The embeddings (2.9) are sharp.
Finally, one can use tools from the representation theory, as we will discuss in
more details in the second half of this article.
One can easily explain (2.5) -- but not (2.6) -- by the fact that the space of
harmonic polynomials of degree j is the underlying space of an irreducible rep-
resentation of K = SO(n). Then (2.5) follows from Schur's lemma and the fact
that C λ commutes with rotations. Note that the group K acts by the left regular
representation on L2(Sn−1) and, as a representation of K, we have the orthogonal
decomposition
L2(Sn−1) ≃K Mj∈N0
Y j,
(2.10)
where the set Y j of all spherical harmonics of degree j is an irreducible K-space.
As we shall see in Section 6, the spectral multiplier (2.6) can also be computed by
identifying C λ as a standard intertwining operator between certain principal series
representations of the larger group SL(n, R), see [57].
The fact that C λ is zero on the odd power harmonics follows from the obser-
vation that the kernel u · vλ is an even function of v. Hence C λ is actually an
integral transform on the projective space P(Rn), and here the analogue of (2.10)
is
L2(P(Rn)) ≃K Mj∈2N0
Y j .
3. Cosine transforms on Stiefel and Grassmann manifolds
In this section we introduce the higher-rank cosine transforms and collect some
basic facts about these transforms. The main results are presented in Theorems
3.2, 3.3, 3.6, 3.7, and 3.8.
THE COSINE TRANSFORM
5
3.1. Notation. We denote by Vn,m ∼ O(n)/O(n − m) the Stiefel manifold
of n × m real matrices, the columns of which are mutually orthogonal unit n-
vectors. For v ∈ Vn,m, dv stands for the invariant probability measure on Vn,m;
ξ = {v} denotes the linear subspace of Rn spanned by v. These subspaces form the
Grassmann manifold Gn,m ∼ O(n)/(O(n − m) × O(m)) endowed with the invariant
probability measure dξ. We write Mn,m ∼ Rnm for the space of real matrices
x = (xi,j ) having n rows and m columns and set
dx =
nYi=1
mYj=1
dxi,j ,
xm = det(xtx)1/2,
xt being the transpose of x. If n = m, then xm is just the absolute value of the
determinant of x; if m = 1, then x1 is the usual Euclidean norm of x ∈ Rn.
3.2. The Cos-function. We give two equivalent "higher-rank" substitutes for
u·v in (1.1). The first one is "more geometric", while the second is "more analytic".
For 1 ≤ m ≤ k ≤ n − 1, let η ∈ Gn,m and ξ ∈ Gn,k be linear subspaces of Rn of
dimension m and k, respectively. Following [2, 3, 57], we set
Cos(ξ, η) = volm(PrξE),
(3.1)
where volm(·) denotes the m-dimensional volume function, E is a convex subset of
η of volume one containing the origin, Prξ denotes the orthogonal projection onto
ξ. By affine invariance, this definition is independent of the choice of E.
The second definition [31] gives precise meaning to the projection operator
Prξ. Let u and v be arbitrary orthonormal bases of ξ and η, respectively. We
regard u and v as elements of the corresponding Stiefel manifolds Vn,k and Vn,m.
If k = m = 1, then u and v are unit vectors, as in (1.1). The orthogonal projection
Prξ is given by the k × k matrix uut, and we can define
Cos(ξ, η) ≡ Cos({u}, {v}) = (det(vtuutv))1/2 ≡ utvm.
(3.2)
This definition is independent of the choice of bases in ξ and η and yields u · v if
k = m = 1.
Remark 3.1. Note that vtuutv is a positive semi-definite matrix, and therefore,
det(vtuutv) ≡ det(utvvtu) ≥ 0. It means that Cos(ξ, η) = Cos(η, ξ) ≥ 0.
3.3. Non-normalized cosine transforms. According to (3.1) and (3.2), one
can use both Stiefel and Grassmannian language in the definition of the higher-rank
cosine transform, namely,
f (v) utvλ
m dv,
u ∈ Vn,k,
f (η) Cosλ(ξ, η) dη,
ξ ∈ Gn,k,
(Cλ
m,kf )(u) =ZVn,m
m,kf )(ξ) =ZGn,m
(Cλ
(3.3)
(3.4)
where dv and dη stand for the relevant invariant probability measures. The fact
that we have two ways of writing of the same operator, extends the arsenal of
techniques (some of them will be exhibited below). Both operators agree with Cλ
in (1.2), when k = m = 1. For brevity, we shall write Cλ
m = Cλ
m,m.
We remark that there are different shifts in the power λ in the literature, all for
different reasons. In particular, to make our statements in Sections 2-4 consistent
with those in [70], one should set λ = α − k. To adapt to the notation in [57] one
6
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
has to change λ to λ − n/2. For unifying the presentation of the results in [70] and
[57] we have preferred to adopt the unshifted notation as in (3.3) and (3.4).
Following [16, 28], the Siegel gamma function of the cone Ω of positive definite
m × m real symmetric matrices is defined by
Γm(α) =ZΩ
exp(−tr(r))rα−(m+1)/2
m
dr = πm(m−1)/4
m−1Yj=0
Γ(α−j/2)
(3.5)
and represents a meromorphic function with polar set
{(m − 1 − j)/2 j = 0, 1, 2, . . .}.
(3.6)
Theorem 3.2. Let 1 ≤ m ≤ k ≤ n − 1.
(i) If f ∈ L1(Vn,m) and Re λ > m − k − 1, then the integral (3.3) converges
for almost all u ∈ Vn,k.
(ii) If f ∈ C∞(Vn,m), then for every u ∈ Vn,k, the function λ 7→ (Cλ
m,kf )(u)
extends to the domain Re λ ≤ m − k − 1 as a meromorphic function with
the only poles m − k − 1, m − k − 2, . . . . These poles and their orders are
the same as of the gamma function Γm((λ + k)/2).
(iii) The normalized integral (Cλ
m,kf )(u)/Γm((λ + k)/2) is an entire function
of λ and belongs to C∞(Vn,k) in the u-variable.
A similar statement holds for (3.4). The proof of Theorem 3.2 can be found
in [70, Theorems 4.3, 7.1]. It relies on the fact that utvλ
m is a special case of the
composite power function (utv)λ with the vector-valued exponent λ ∈ Cm [16, 28].
The corresponding composite cosine transforms were studied in [58, 59, 70].
An important ingredient of the proof of Theorem 3.2 is the connection between
the cosine transform Cλ
m,kf on Vn,m and the Fourier transform
ϕ(y) = (F ϕ)(y) =ZMn,m
etr(iytx)ϕ(x) dx,
y ∈ Mn,m .
(3.7)
The corresponding Parseval equality has the form
( ϕ, ω) = (2π)nm (ϕ, ω),
ϕ(x)ω(x) dx.
(3.8)
(ϕ, ω) =ZMn,m
This equality with ω in the Schwartz class S(Mn,m) of smooth rapidly decreasing
functions is used to define the Fourier transform of the corresponding distributions.
We will need polar coordinates on Mn,m, so that for n ≥ m, every matrix
x ∈ Mn,m of rank m can be uniquely represented as x = vr1/2 with v ∈ Vn,m and
r = xtx ∈ Ω. Given a function f on Vn,m, we denote (Eλf )(x) = rλ/2
m f (v). The
following statement holds in the case k = m.
Theorem 3.3. Let f be an integrable right O(m)-invariant function on Vn,m,
ω ∈ S(Mn,m), 1 ≤ m ≤ n−1, Cλ
mf = Cλ
m,mf . Then for every λ ∈ C,
(cid:18)
EλCλ
mf
Γm((λ + m)/2)
, F ω(cid:19) = c (cid:18) E−λ−nf
Γm(−λ/2)
, ω(cid:19) ,
c =
2m(n+λ) πnm/2 Γm(n/2)
Γm(m/2)
,
where both sides are understood in the sense of analytic continuation.
(3.9)
THE COSINE TRANSFORM
7
The formula (3.9) agrees with (2.4). The more general statement for arbitrary
k ≥ m can be found in [70].
Remark 3.4. It is important to note that the domains, where the left-hand
side and the right-hand side of of (3.9) exist as absolutely convergent integrals,
have no points in common, when m > 1. This is the principal distinction from
the case m = 1, when there is a common strip of convergence −1 < Re λ < 0.
To perform analytic continuation, we have to switch from Cλ
m to the more general
m with λ ∈ Cm and then take the restriction to the
composite cosine transform Cλ
diagonal λ1 = · · · = λm = λ + m. This method of analytic continuation was first
used by Kh`ekalo (for another class of operators) in his papers [39, 41, 40] on Riesz
potentials on the space of rectangular matrices.
3.4. The Funk transform. The higher-rank version of the classical Funk
transform (2.3) sends a function f on Vn,m to a function Fm,kf on Vn,k by the
formula
(Fm,kf )(u) =Z{v∈Vn,m utv=0}
f (v) duv,
u ∈ Vn,k.
(3.10)
The condition utv = 0 means that subspaces {u} ∈ Gn,k and {v} ∈ Gn,m are
mutually orthogonal. Hence, necessarily, k + m ≤ n. The case k = m, when both f
and its Funk transform live on the same manifold, is of particular importance and
coincides with (2.3) when k = m = 1. We denote Fm = Fm,m.
If f is right O(m)-invariant, (Fm,kf )(u) can be identified with a function on
the Grassmannians Gn,m or Gn,n−m, and can be written "in the Grassmannian
language". For instance, setting ξ = {v} ∈ Gn,m, η = {u}⊥ ∈ Gn,n−k, and f (ξ) =
f (v), we obtain
f (ξ) dηξ = (Fm,kf )(u).
(3.11)
(Rm,n−k f )(η) ≡ Z
ξ⊂η
3.5. Normalized cosine transforms. Our next aim is to introduce a natural
m,kf of the normalized transform (2.1). "Natural" means that we
m,kf to obey the relevant higher-rank modifications of the properties (2.2)-
generalization C λ
expect C λ
(2.5).
Definition 3.5. Let 1 ≤ m ≤ k ≤ n − 1. For u ∈ Vn,k and v ∈ Vn,m, we define
(C λ
m,kf )(u) = γn,m,k(λ)ZVn,m
f (v) utvλ
m dv,
(3.12)
where
γn,m,k(λ) =
Γm(m/2)
Γm(n/2)
Γm(−λ/2)
Γm((λ + k)/2)
,
λ + m 6= 1, 2, . . . .
We denote C λ
m = C λ
m,m. The integral (3.12) is absolutely convergent if Re λ >
m − k − 1. The excluded values of λ belong to the polar set of Γm(−λ/2).
If
k = m = 1 this definition coincides with (2.1). Operators of this kind implicitly
arose in [26, pp. 367, 368].
Theorem 3.6. Let 1 ≤ m ≤ k ≤ n − 1, k + m ≤ n.
If f is a C∞ right
O(m)-invariant function on Vn,m, then for every u ∈ Vn,k,
a.c.
λ=−k
(C λ
m,kf )(u) =
Γm(m/2)
Γm((n − k)/2)
(Fm,kf )(u),
(3.13)
8
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
where "a.c." denotes analytic continuation and (Fm,kf )(u) is the Funk transform
(3.10).
This statement follows from [70, Theorems 7.1 (iv) and 6.1]. Note that if
m = k = 1, then (3.13) yields (2.2). However, unlike (2.2), the proof of which
is straightforward, (3.13) requires a certain indirect procedure, which invokes the
Fourier transform on the space of matrices and the relevant analogue of (3.9).
We point out that a pointwise inversion of the Funk transform can be obtained
by means of the dual cosine transform, which is defined by
∗
C λ
(
m,kϕ)(v) =ZVn,k
ϕ(u) utvλ
m du,
v ∈ Vn,m.
(3.14)
Indeed, the following result holds.
Theorem 3.7. (cf. [70, Theorems 7.4]) Let ϕ = Fm,kf , where f is a C∞ right
O(m)-invariant function on Vn,m, 1 ≤ m ≤ k ≤ n − m. Then, for every v ∈ Vn,m,
∗
C λ
(
m,kϕ)(v)
Γm((λ + k)/2)
a.c.
λ=m−n
= c f (v),
c =
Γm(n/2)
Γm(k/2) Γm(m/2)
.
(3.15)
Regarding other inversion methods of the higher-rank Funk transform (which
is also known as the Radon transform for a pair of Grassmannnians), see [31, 85]
and references therein.
In the case k = m the normalized cosine transform C λ
m,m has a number
of important features. If f ∈ C∞(Vn,m), then analytic continuation of (C λ
mf )(u)
is well-defined for all complex λ /∈ {1 − m, 2 − m, . . .} and belongs to C∞(Vn,m).
The following inversion formulas hold.
m = C λ
Theorem 3.8. (cf. [70, Theorems 7.7]) Let f ∈ C∞(Vn,m) be a right O(m)-
invariant function on Vn,m, 2m ≤ n. Then, for every u ∈ Vn,m,
(C −λ−n
m
C λ
mf )(u) = f (u),
λ, −λ − n /∈ {1 − m, 2 − m, . . .}.
(3.16)
In particular, for the non-normalized transforms,
(C−λ−n
m
Cλ
mf )(u) = ζ(λ) f (u),
λ + n, −λ /∈ {1, 2, 3, . . .},
(3.17)
where
ζ(λ) =
Γ2
m(n/2) Γm((m + λ)/2) Γm((m − λ − n)/2)
Γ2
m(m/2) Γm(−λ/2) Γm((n + λ)/2)
.
(3.18)
Both equalities (3.16) and (3.18) are understood in the sense of analytic continua-
tion.
In the case m = 1, the formulas (3.16) and (3.17) coincide with (2.7) and (2.8),
respectively, but the method of the proof is different.
4. Connection to Representation Theory
The cosine transform is closely related to the representation theory of semisimple
Lie groups.
In particular, as we shall now discuss, it has an important group-
theoretic interpretation as a standard intertwining operator between generalized
principal series representations of SL(n, R).
In the following we shall use the notation G = SL(n, R), K = SO(n), and
0 B(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
L = S(O(m)×O(n − m)) =(cid:26)(cid:18)A 0
A ∈ O(m)
B ∈ O(n − m)
, det(A)det(B) = 1(cid:27)
THE COSINE TRANSFORM
9
with m ≤ n − m. Then B ≡ K/L = Gn,m is the Grassmanian of m-dimensional
linear subspaces of Rn. We fix the base point
bo = {(x1, . . . , xm, 0, . . . , 0) x1, . . . , xm ∈ R} ∈ B,
so that B = K · b0 and every function on B can be regarded as a right L-invariant
function on K.
From now on, our main concern is the nonnormalized cosine transform (3.3)
with equal lower indices, that is, Cλ
m,m. We refer to [35, Chapter V] for
harmonic analysis on compact symmetric spaces and [42] for the representation
theory of semisimple Lie groups.
m ≡ Cλ
4.1. Analysis on B with respect to K. The first connection to representa-
tion theory is related to the left regular action of the group K on L2(B) by
(cid:0)ℓ(k)f(cid:1)(b) = f (k−1b),
k ∈ K , b ∈ B.
For an irreducible unitary representation (π, Vπ) of K, we consider the subspace
V L
π := {v ∈ Vπ π(k)v = v ∀k ∈ L},
L = S(O(m)×O(n − m)).
The representation (π, Vπ) is said to be L-spherical if V L
π 6= {0}. As B = K/L is a
symmetric space, the following result is a consequence of [35, Chapter IV, Lemma
3.6].
Proposition 4.1. If (π, Vπ) is L-spherical, then dim V L
π = 1.
Since V L
π 6= {0}, we can choose a unit vector eπ ∈ V L
π . Then we define a map
Φπ : Vπ → C∞(B) ⊂ L2(B) by the formula
(Φπv)(b) := d(π)−1/2hv, π(k)eπi ,
v ∈ Vπ,
b = k · bo ∈ B = K · bo,
(4.1)
where d(π) = dim Vπ. This definition is meaningful because k · bo = kk′ · bo for
every k′ ∈ L and eπ remains fixed under the action of π(k′). We also set
Φπ(v; b) := (Φπv)(b).
Recall, if (π, Vπ) and (σ, Vσ) are two representations of a Hausdorff topological group
H, then an intertwining operator between π and σ is a bounded linear operator
T : Vπ → Vσ such that T π(h) = σ(h)T for all h ∈ H. If π is irreducible and T
intertwines π with itself, then Schur's Lemma states that T = c id for some complex
number c, [17], p. 71. The map Φπ is a K-intertwining operator in the sense that
it intertwines the representation π on Vπ and the left regular representation ℓ on
L2(B), so that for b = h · bo and k ∈ K we have
Φπ(π(k)v; b) = hπ(k)v, π(h)eπi = hv, π(k−1h)eπi = ℓ(k)Φπ(v; b) .
Furthermore, the left regular representation ℓ on L2(B) is multiplicity free, see
e.g.
[84, Corollary 9.8.2]. Therefore, since (π, Vπ) is irreducible, any intertwining
operator Vπ → L2(B) is by Schur's Lemma of the form c Φπ for some c ∈ C.
We let L2
π(B) = Im Φπ. Denote by bKL the set of all equivalence classes of
irreducible L-spherical representations (π, Vπ) of K. Then, see [35, Chapter V,
Thm. 4.3], the decomposition of L2(B) as a K-representation is as follows.
Theorem 4.2. L2(B) ≃K Mπ∈ bKL
L2
π(B) .
10
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
The cosine transform is, as mentioned before, a K-intertwining operator, i.e.,
m(f ) for all k ∈ K and f ∈ L2(B). It follows by Schur's Lemma
m(ℓ(k)f ) = ℓ(k)Cλ
Cλ
that for each π ∈ bKL there exists a function ηπ on C such that
= ηπ(λ) id .
Cλ
mL2
π
(4.2)
Let f ∈ L2
meromorphic; cf. Theorem 3.2.
π(B) of norm one. Then ηπ(λ) = hCλ(f ), f i and it follows that ηπ(λ) is
4.2. Generalized spherical principal series representations of G. The
fact that Cλ
m is a K-intertwining operator does not indicate how to determine the
functions ηπ. In the case m = 1 and in some particular cases for the higher-rank
cosine transforms [58, 59] explicit expression for ηπ can be obtained using the
Funk-Hecke Theorem or the Fourier transform technique. It is a challenging open
problem to proceed the same way in the most general case, using, e.g., the relevant
results of Gelbart, Strichartz, and Ton-That, see, e.g., [25, 80, 82]. Below we
suggest an alternative way and proceed as follows.
To find ηπ explicitly, we observe that the cosine transform is an intertwining
operator between certain generalized principal series representations (πλ, L2(B)) of
G = SL(n, R) induced from a maximal parabolic subgroup of G. We can then use
the bigger group G, or better its Lie algebra, to move between K-types. We invoke
the spectrum generating technique introduced in [7] to build up a recursion relation
between the spectral functions ηπ. This finally allows us to determine all of them
by knowing ηtrivial.
The group G = SL(n, R) acts on B by
g · η := {gv v ∈ η} ,
where gv denotes the usual matrix multiplication. This action is transitive, as the
K-action is already transitive. The stabilizer of bo is the group
P = (cid:26)(cid:18)A X
0 B(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) X ∈ Mm,n−m ,
≃ S(GL(m) × GL(n − m)) ⋉ Mm,n−m ,
A ∈ GL(m, R)
B ∈ GL(n − m, R)
and det(A)det(B) = 1(cid:27)
where Mn,m is the space of n × m real matrices; see Section 3.1. We then have
B = G/P .
The K-invariant probability measure on B is not G-invariant. But there exists
a function j : G × B → R+ such that for all f ∈ L1(B) we have
ZB
f (b) db =ZB
f (g · b)j(g, b)n db ,
g ∈ G,
b ∈ B .
(4.3)
We include the power n to adapt our notation to [57]. By the associativity of the
action we have j(gg′, b) = j(g, g′ · b)j(g′, b) for all g ∈ G and b ∈ B. Hence, for each
λ ∈ C we can define a continuous representation πλ of G on L2(B) by
[πλ(g)f ](b) := j(g−1, b)λ+n/2f (g−1 · b) ,
g ∈ G,
f ∈ L2(B),
β ∈ B.
(4.4)
A simple change of variables shows that
hπλ(g)f, hiL2 = hf, π−λ(g−1)hiL2 ,
g ∈ G ,
f, h ∈ L2(B) .
In particular, πλ is unitary if and only if λ is purely imaginary. The representations
πλ are the so-called generalized (spherical) principal series representations (induced
THE COSINE TRANSFORM
11
from the maximal parabolic subgroup P ), in the compact picture. See e.g. [42], p.
169.
The representations πλ can also be realized on Stiefel manifolds as follows.
According to [70, Section 7.4.3], we introduce the radial and angular components
of a matrix x ∈ Mn,m of rank m by
rad(x) = (xtx)1/2 ∈ Ω,
ang(x) = x(xtx)−1/2 ∈ Vn,m,
so that x = ang(x) rad(x). Given λ ∈ C, we define a mapping which assigns to
every g ∈ GL(n, R) an operator πλ(g) acting on measurable functions f on Vn,m
by the rule
πλ(g)f (v) = rad(g−1v)−(λ+n/2) f (ang(g−1v)).
(4.5)
Clearly, πλ(In) is an identity operator. One can prove that if f is a measurable
right O(m)-invariant function on Vn,m, then
πλ(g1g2)f = πλ(g1) πλ(g2)f ,
g1, g2 ∈ GL(n, R).
(4.6)
The restriction of πλ to SL(n, R), acting on the space of square integrable right
O(m)-invariant functions on Vn,m, coincides with the representation defined by
(4.4).
4.3. The cosine transform as an intertwining operator. In this section
we follow the ideas in [57]. An alternative self-contained exposition (without using
the representation theory of semisimple Lie groups), can be found in [70].
The gain by using the representations πλ is that we now have a meromorphic
family of representations on L2(B) and that they are irreducible for almost all λ
and closely related to the cosine transform. For this, we recall some results from
[83].
Theorem 4.3 (Vogan-Wallach). There exists a countable collection {pn} of
non-zero holomorphic polynomials on C such that if pn(λ) 6= 0 for all n then πλ is
irreducible. In particular, πλ is irreducible for almost all λ ∈ C.
Proof. This is Lemma 5.3 in [83].
(cid:3)
Let θ : G → G be the involutive automorphism θ(g) = (g−1)t. We remark that
in [57] notation Cosλ = Cλ−n/2
m
was used.
Theorem 4.4. The cosine transform intertwines πλ and π−λ ◦ θ, namely,
m ◦ πλ+n/2 = (π−λ−n/2 ◦ θ) ◦ Cλ
Cλ
m,
whenever both sides of this equality are analytic functions of λ.
Proof. We refer to Theorem 2.3 and (4.10) in [57].
(4.7)
(cid:3)
In fact, it is shown in [57], Lemma 2.5 and Theorem 4.2, that Cλ−n/2
= J(λ),
where J(λ) is a standard intertwining operator, studied in detail among others by
Knapp and Stein in [43, 44] and Vogan and Wallach in [83]. These authors show,
in particular, that λ 7→ J(λ) has a meromorphic extension to all of C. Furthermore,
Vogan and Wallach show that if f ∈ C∞(B), then the map
m
{λ ∈ C Re (λ) > −1 + n/2} ∋ λ 7−→ J(λ)f ∈ C∞(B)
is holomorphic. As a consequence of Cλ−n/2
following theorem.
m
= J(λ) and [83, 1.6 Thm], we get the
12
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
Theorem 4.5. The map λ 7→ Cλ
m extends meromorphically to C. In particular,
for f ∈ C∞(B) the function λ 7→ Cλ
m(f )(b) extends to a meromorphic function on
C and the set of possibles poles can be chosen independent of f . In the complement
of the singular set we have Cλ
m(f ) ∈ C∞(B).
Notice that precise information about analiticity of more general cosine trans-
forms, including the structure of polar sets, is presented in Theorem 3.2 above.
The implication of (4.7) is that C−λ−n/2
intertwines πλ with itself
(in the sense of meromorphic family of operators). By Theorem 4.3 there exists a
meromorphic function η on C such that
◦ Cλ−n/2
m
m
C−λ−n/2
m
◦ Cλ−n/2
m
= η(λ) idC∞(B)
(4.8)
for all λ ∈ C for which the left-hand side is well defined. The shift by n/2 in the
definition is chosen so that the final formulas agree with those in [57] and make
some formulas more symmetric. The fact that η is meromorphic follows by noting
that η(λ) = hC−λ−n/2
◦ Cλ−n/2
(1), 1i.
m
m
Formula (4.8) is a symmetric version of (3.18) with λ replaced by λ − n/2.
The explicit value of η(λ) can be easily obtained from (3.17). An alternative,
representation-theoretic method to compute the function η(λ), is presented in Sec-
tion 6. The first step is the following lemma.
Lemma 4.6. Let c(λ) = Cλ−n/2
m
(1). Then η(λ) = c(λ)c(−λ).
Note that c(λ) is nothing but ηtrivial(λ) in Theorem 6.6.
Remark 4.7. There are several ways to prove the meromorphic extension of
the standard intertwining operators. The proof in [83] uses tensoring with finite
dimensional representations of G to deduce a relationship between Cλ
m and Cλ+2n
m .
In fact, there exists a family of (non-invariant) differential operators Dλ on B and
a polynomial b(λ), the Bernstein polynomial, such that
b(λ)Cλ
m(f ) = Cλ+2n
m (Dλ(f ))
(4.9)
[83, Thm. 1.4]. Another way to derive an equation of the form (4.9) is to convert the
m into an integral over the orbit of certain nilpotent group ¯N , as
integral defining Cλ
usually done in the study of standard intertwining operators, and then use the ideas
from [8, 55, 56]. In the case where G/P is a symmetric R-space (which contains
the case of Grassmann manifolds), the standard intertwining operators J(λ) have
been recently studied by Clerc in [9], using Loos' theory of positive Jordan triple
systems. In particular, Clerc explicitly computes the Bernstein polynomials b(λ)
in (4.9), and, hence, proves the meromorphic extension of J(λ) for this class of
symmetric spaces.
Finally, one can stick with the domain where λ 7→ Cλ
m is holomorphic and
determine the K-spectrum functions ηπ(λ) in (4.2). As rational functions of Γ-
factors, these functions have meromorphic extension to C. Hence, λ 7→ Cλ
m itself
has meromorphic extension by (4.2). We will comment more on that in Remark
6.8.
4.4. Historical remarks. We conclude this section with a few historical re-
marks. The standard intertwining operators J(λ), as a meromorphic family of
singular integral operators on K or ¯N , have been central objects in the study of
THE COSINE TRANSFORM
13
representation theory of semimisimple Lie groups since the fundamental works of
Knapp and Stein [43, 44], Harish-Chandra [33], and several others. In our case
¯N =(cid:26)(cid:18)Im
0
X In−m(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) X ∈ Mm,n−m(cid:27) .
Then, in the realization of the generalized principal series representations on L2(B),
the kernel of J(λ) is Cosλ−n/2(b, c). But in most cases there is neither an explicit
formula nor geometric interpretation of the kernel defining J(λ).
Apart of customary applications of the cosine transform in convex geometry,
probability, and the Banach space theory, similar integrals turned up independently
as standard intertwining operators between generalized principal series representa-
tions of SL(n, K), where K = R, C or H.
The real case was studied in [12], the complex case in [14], and the quaternionic
case in [60]. In these articles it was shown that integrals of the form
ZB
(x, y)λ−n/2f (x) dx,
with some modification for K = C or H, define intertwining operators between
generalized principal series representations induced from a maximal parabolic sub-
group in SL(n + 1, K). The K-spectrum was determined, yielding the cases of
irreducibility and, more generally, the composition series of those representations.
Among the applications, there were some embeddings of the complementary series
and the study of the so-called canonical representations on some Riemannian sym-
metric spaces of the noncompact type, [10, 11, 13]. However the connections of
these considerations to convex geometry, to the cosine transform and to the Funk
and Radon transforms was neither discussed nor mentioned. These connections
were first published in [57] in the context of the Grassmannians over R , C and
H. However, it was probably S. Alesker who was the first to remark in the unpub-
lished manuscript [2] that the cosine transform is a SL(n, R)-intertwining operator.
It was also shown in [86] that the Sinλ-transform (a transform related to the sine
transform) can be viewed as a Knapp-Stein intertwining operator. This was used to
construct complementary series representations for GL(2n, R). The Sinλ- transform
is then also naturally linked to reflection positivity, which relates complementary se-
ries representations of GL(2n, R) to the highest weight representations of SU(n, n),
[18, 19, 53, 37, 36]. Notice, however, that the definition of the Sinλ-transform in
[86] differs from the one in [66], [70]; see also [67] for the sine transform on the
hyperbolic space.
5. The spherical representations
The functions ηπ(λ) in (4.2) are parametrized by the L-spherical representations
of K. The main purpose of this section is to present this parametrization, which
is given by a semilattice in a finite dimensional Euclidean space associated with a
maximal flat submanifold of B. We will, therefore, have to study the structure of
the symmetric space B. We refer to [81] and the books by Helgason [34, 35] for
more detailed discussions and proofs. To bring the discussion closer to standard
references in Lie theory we also introduce some Lie theoretical notation which we
have avoided so far.
14
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
Let
g = {X ∈ Mn,n tr(X) = 0} ,
k = {X ∈ Mn,n X t = −X} ,
be the Lie algebras of G = SL(n, R) and K = SO(n), respectively. The derived
involution of θ on g, still denoted θ, is given by θ(X) = −X t. Hence, k = g(1, θ),
the eigenspace of θ on g with eigenvalue 1. We fix once and for all the G-invariant
m(n−m) tr(XY ) on g. Note that β is negative definite on k
bilinear form β(X, Y ) =
and hX, Y i = −β(X, θ(Y )) is an inner product on g such that ad(X)t = −ad(θ(X)),
where, as usual, ad(X)Y = [X, Y ] = XY − Y X. The normalization of β is chosen
so that it agrees with [57].
n
We recall that B is a symmetric space corresponding to the involution
τ (x) =(cid:18)Im
0 −In−m(cid:19) x(cid:18)Im
0 −In−m(cid:19) =(cid:18) A −B
−C D (cid:19)
0
0
for
x =(cid:18)A B
C D(cid:19) ,
where for r ∈ N we denote by Ir the r × r identity matrix. Note that τ in fact
defines an involution of G and that the derived involution on the Lie algebra g is
given by the same form.
We have k = l ⊕ q where l ≃ so(m) × so(n − m) is the Lie algebra of L and
q = k(−1, τ ) =(cid:26) Q(X) =(cid:18)0mm
−X t
X
0n−m,n−m(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) X ∈ Mm,n−m(cid:27) .
Let Eν,µ = (δiν δjµ)i,j denote the matrix in Mm,n−m with all entries equal to 0 but
the (ν, µ)-th which is equal to 1. For t = (t1, . . . , tm)t ∈ Rm we set
X(t) = −
mXj=1
tjEn−2m+j,j ∈ Mm,n−m ,
Y (t) = Q(X(t)) ∈ q .
Then b = {Y (t) t ∈ Rm} ≃ Rm is a maximal abelian subspace of q.
Pm
To describe the set bKL we note first that B is not simply connected. So we
cannot use the Cartan-Helgason theorem [35, p. 535] directly, but only a slight
modification is needed. Define ǫj(Y (t)) := itj. We will identify the element λ =
j=1 λj ǫj ∈ b∗
C with the corresponding vector λ = (λ1, . . . , λm).
If H ∈ b, then ad(H) is skew-symmetric on k with respect to the inner product
h · , · i. Hence ad(H) is diagonalizable over C with purely imaginary eigenvalues.
For α ∈ ib∗ let
kα
C := {X ∈ kC (∀H ∈ b) ad(H)X = α(H)X}
be the joint α-eigenspace. Let
∆k = {α ∈ ib∗ α 6= 0
and
kα
C 6= {0}} .
The dimension of kα
C is called the multiplicity of α (in kC).
Lemma 5.1. We have
∆k = {±ǫi ± ǫj (1 ≤ i 6= j ≤ m , ± independently), ±ǫi (1 ≤ i ≤ m) }
with multiplicities respectively 1 (and not there if m = 1), 2n − m (and not there if
m = n − m).
THE COSINE TRANSFORM
15
Proof. This follows from [34]: the table on page 518, the description of the
simple root systems on page 462 ff. and the Satake diagrams on pages 532 -- 533. (cid:3)
We let
∆+
k = {ǫi ± ǫj (1 ≤ i < j ≤ m ), ǫi (1 ≤ i ≤ m) } .
Lemma 5.2. Let ρk =
1
2 Xα∈∆+
k
dim(kα
C)α ∈ ib∗. Then ρk =
mXj=1(cid:16) n
2
− j(cid:17)ǫj.
Let now (π, Vπ) be a unitary irreducible representation of K. Then Vπ is
finite dimensional. Moreover, π(H) =
π(exp(tH)) is skew-symmetric, hence,
d
dt(cid:12)(cid:12)(cid:12)t=0
π
π ⊂ Vπ denote the joint eigenspace of eigenvalue µ. If X ∈ kα
diagonalizable, for all H ∈ b (in fact, π(H) is diagonalizable for all H ∈ k). Let
Γ(π) ⊂ ib∗ be the finite set of joint eigenvalues of π(H) with H ∈ b. For µ ∈ Γ(π),
C and v ∈ V µ
let V µ
π ,
then π(X)v ∈ V µ+α
π = {0}
for all α ∈ ∆+
k . This only uses that π is finite dimensional, but the irreducibility
implies that this µ is unique. It is called the highest weight of π. Finally we have
π ≃ σ if and only if µπ = µσ.
. Thus, there exists a µ = µπ ∈ Γ(π) such that π(kα
C)V µ
Let eK be the universal covering group of K. Then τ lifts to an involution eτ on
eK, eL := eK eτ is connected, and eB := eK/eL is the universal covering of B. Replacing
K by eK etc., we can talk about eL-spherical representations of eK and their highest
Theorem 5.3. The map π 7→ µπ sets up a bijection between the set of eL-
weights. The following theorem is a consequence of the Cartan-Helgason theorem
[35, p. 535].
spherical representations of eK and the semi-lattice
Λ+(eB) =(cid:26)µ ∈ ib∗ (cid:12)(cid:12)(cid:12)(cid:12) (∀α ∈ ∆+
Furthermore, if m = n/2, then
k )
hµ, αi
hα, αi
∈ Z+(cid:27) .
(5.1)
Otherwise,
Λ+(eB) = {(µ1, . . . , µm) ∈ Zm µ1 ≥ µ2 ≥ · · · ≥ µm−1 ≥ µm} .
Λ+(eB) = {(µ1, . . . , µm) ∈ Zm µ1 ≥ µ2 ≥ · · · ≥ µm−1 ≥ µm ≥ 0} .
sentation. Recall the notation Φπµ from (4.1). Let Λ+(B) denote the sublattice in
If µ ∈ Λ+(eB), then we write (πµ, Vµ) for the corresponding eL-spherical repre-
Λ+(eB) which corresponds to L-spherical representations of K. Then µ ∈ Λ+(B)
if and only if the functions Φπµ(v), which are originally defined on eB, factor to
µ and H ∈ b. We can normalize v and eπµ so
functions on B. For that, let v ∈ V µ
that
The same argument as for the sphere [81, Ch. III.12], proves the following theorem.
Φπµ(v; exp H) = eµ(H) .
Theorem 5.4. If m = n − m, then
Λ+(B) = {µ =
mXj=1
µjεj µj ∈ 2N0
and µ1 ≥ . . . ≥ µm−1 ≥ µm } .
16
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
In all other cases,
Λ+(B) = {µ =
mXj=1
µjεj µj ∈ 2N0
and µ1 ≥ . . . ≥ µm ≥ 0 } .
6. The generation of the K-spectrum
Recall from Section 5 the involution θ(X) = −X t on g. The Lie algebra g decom-
poses into eigenspaces of θ as g = k ⊕ s, where
s = g(−1, θ) = {X ∈ Mn,n θ(X) = −X and Tr(X) = 0} .
Then, except in the case n = 2, the complexification sC of s is an irreducible L-
spherical representation of K. For n = 2 this representation decomposes into two
one-dimensional representations.
Let
Ho =(cid:18) n−m
0
n Im
0
n In−m(cid:19) ∈ s .
− m
Then Ho is L-fixed and hH0, H0i = 1. Define a := RHo. The operator ad(H0) has
spectrum {0, 1, −1} and n = g(1, ad(H0)).
Let Ad(k) denote the conjugation by k. Define a map ω : sC → C∞(B) by
ω(Y )(k) := hY, Ad(k)Hoi = β(Y, Ad(k)Ho) =
n
m(n−m) Tr(Y kHok−1)
and note that
ω(Ad(h)Y )(k) = hAd(h)Y, Ad(k)Hoi = hY, Ad(h−1k)Hoi = ω(Y )(h−1k) .
Thus ω is a K-intertwining operator.
Fix an orthonormal basis X1, . . . , Xdim q of q such that X1, . . . , Xm, is an or-
j the corresponding positive definite
thonormal basis of b. Denote by Ω = −Pj X 2
Laplace operator on B. Then
where
ΩL2
µ(B) = ω(µ) id ,
ω(µ) = hµ + 2ρk, µi .
A simple calculation then gives:
Lemma 6.1. Let µ = (µ1, . . . , µm) ∈ Λ+(B). Then
ω(µ) =
m(n − m)
2n
mXj=1(cid:16)µ2
j + µj(n − 2j)(cid:17) .
For f ∈ C∞(B) denote by M (f ) : L2(B) → L2(B) the multiplication operator
g 7→ f g. Recall the notation π0 for the finite dimensional spherical representation
of highest weight 0 ∈ Λ+(B).
Theorem 6.2. Let Y ∈ s. Then [Ω, M (ω(Y ))] = 2π0(Y ).
Proof. This is Theorem 2.3 in [7].
(cid:3)
For µ ∈ Λ+(B) define Ψµ : L2
µ(B) ⊗ sC → L2(B) by
Ψµ(ϕ ⊗ Y ) := M (ω(Y ))ϕ .
Observe that for k ∈ K, Y ∈ sC, and ϕ ∈ L2
µ(B) we have
ℓ(k)(cid:0)M (ω(Y )ϕ(cid:1) =(cid:0)ℓ(k)ω(Y )(cid:1)(ℓ(k)ϕ) = M(cid:0)ω(Ad(k)Y )(cid:1)(ℓ(k)ϕ)
THE COSINE TRANSFORM
17
with Ad(k)Y ∈ sC and ℓ(k)ϕ ∈ L2
K-invariant. Define a finite subset S(µ) ⊂ Λ+(B) by
µ(B). Hence, Ψµ is K-equivariant and Im Ψµ is
Im Ψµ ≃K Mσ∈S(µ)
L2
σ(B) .
Lemma 6.3. Let µ ∈ Λ+(B). Then
S(µ) = {µ ± 2ǫj j = 1, . . . , m} ∩ Λ+(B) .
These representations occur with multiplicity one.
Denote by prσ the orthogonal projection L2(B) → L2
σ(B). The first spectrum
generating relation which follows from Theorem 6.2, see also [7, Cor. 2.6], states:
Lemma 6.4. Assume that µ ∈ Λ+(B). Let σ ∈ S(µ), Y ∈ sC, and λ ∈ C. Let
ωσµ(Y ) := prσ ◦ M (ω(Y ))L2
µ(B) : L2
µ(B) → L2
σ(B) .
Then
prσ ◦ πλ(Y )L2
µ(B) =
1
2
(ω(σ) − ω(µ) + 2 m(n−m)
n
λ)ωσµ(Y ) .
(6.1)
(6.2)
The spectrum generating relation that we are looking for can now easily be
deducted and we get:
Lemma 6.5. Let µ = (µ1, . . . , µm) ∈ Λ+(B) and λ ∈ C. Then
ηµ+2ǫj (λ)
ηµ(λ)
=
λ − µj + j − 1
λ + µj + n − j + 1
= −
−λ + µj − j + 1
λ + µj + n − j + 1
(6.3)
and η0(λ) = c(λ).
Proof. First we apply Cλ−n/2
m
mutes with prσ and that Cλ−n/2
We then get:
m
◦πλ(Y ) = π−λ ◦θ(Y )◦Cλ−n/2
m
to (6.2) from the left, using that Cλ−n/2
com-
= −π−λ(Y )◦Cλ−n/2
.
m
m
(cid:0)ω(σ) − ω(µ) + 2 m(n−m)
n
λ(cid:1)ησ(λ − n/2)ωσµ(Y ) =
−(cid:0)ω(σ) − ω(µ) − 2 m(n−m)
n
λ(cid:1)ηµ(λ − n/2)ωσµ(Y ) .
As ωσδ(Y ) is non-zero, for generic λ it can be canceled out.
Now insert the expression from Lemma 6.1 to get
ω(µ + 2ǫj) − ω(µ) = 2m(n−m)
n
(µj + n/2 − (j − 1))
and the claim follows. The last statement follows from the fact that πλ is irreducible
for generic λ, hence, iterated application of (6.1) will in the end reach all K-types
starting from the trivial K-type.
(cid:3)
Lemma 6.3 tells us that the evaluation of ηµ(λ) can be done in two steps. First
we determine the function η0(λ) and then use (6.3) as an inductive procedure to
determine the rest. The final result is given in the following theorem. It is presented
in terms of Γ-functions associated to the cone Ω of m× m positive definite matrices,
namely,
ΓΩ(λ) = πm(m−1)/4
mYj=1
Γ(λj − (j − 1)/2),
λ = (λ1, . . . , λm) ∈ Cm.
(6.4)
18
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
This integral is a generalization of Γm(λ) in (3.5); cf. [16, p. 123], [70, Sec. 2.2].
In the following the scalar parameters, which occur in the argument of ΓΩ, are
interpreted as vector valued, for instance, n ∼ (n, . . . , n), λ ∼ (λ, . . . , λ).
Theorem 6.6 ([57]). Let Λ+(B) be the sublattice in Theorem 5.4 parametrizing
the L-spherical representations of K, let µ = (µ1, . . . , µm) ∈ Λ+(B), and λ ∈ C.
Then the K-spectrum of the cosine transform Cλ
m is given by:
ηµ(λ) = (−1)µ/2 Γm (n/2)
Γm (m/2)
Γm ((λ + m)/2))
ΓΩ ((µ − λ)/2)
Γm (−λ/2)
ΓΩ ((λ + n + µ)/2)
.
(6.5)
Remark 6.7. Owing to (3.12), the spectrum of the normalized cosine transform
C λ
m has the simpler form
ηµ(λ) = (−1)µ/2
ΓΩ ((µ − λ)/2)
ΓΩ ((λ + n + µ)/2)
.
(6.6)
In the case m = 1 this formula coincides with (2.6).
Remark 6.8. In Section 4 we referred to the result of Vogan and Wallach on
the meromorphic continuation of the intertwining operator J(λ). This result is not
needed for the computation of ηµ(λ). Indeed, it is enough to know that J(λ) is
holomorphic on some open subset of C as that is all what is needed to determine
ηµ(λ) in Theorem 6.6. We can then extend Cλ
m meromorphically on each K-type.
Note, however, that this is weaker than the statement in [83] which extends Cλ
mf
for all smooth functions.
References
[1] A. D. Aleksandrov. On the theory of mixed volumes of convex bodies. II. New inequalities
between mixed volumes and their applications. Mat. Sbornik N.S., 2:1205 -- 1238, 1937.
[2] S. Alesker. The α-cosine transform and intertwining integrals on real grassmannians. 2003.
[3] S. Alesker and J. Bernstein. Range characterization of the cosine transform on higher Grass-
mannians. Adv. Math., 184(2):367 -- 379, 2004.
[4] R. Askey and S. Wainger. On the behavior of special classes of ultraspherical expansions. I,
II. J. Analyse Math., 15:193 -- 220, 1965.
[5] A. Bernig. Algebraic integral geometry. arXiv:1004.3145v3, 2011.
[6] W. Blaschke. Kreis und Kugel. Chelsea Publishing Co., New York, 1949.
[7] T. Branson, G. ´Olafsson, and B. Ørsted. Spectrum generating operators and intertwining
operators for representations induced from a maximal parabolic subgroup. J. Funct. Anal.,
135(1):163 -- 205, 1996.
[8] J.-L. Brylinski and P. Delorme. Vecteurs distributions H-invariants pour les s´eries princi-
pales g´en´eralis´ees d'espaces sym´etriques r´eductifs et prolongement m´eromorphe d'int´egrales
d'Eisenstein. Invent. Math., 109(3):619 -- 664, 1992.
[9] J.-L. Clerc. Intertwining operators for the generalized principal series on symmetric R-spaces.
arxiv:1209.0691v1, 2012.
[10] G. van Dijk and S. C. Hille. Canonical representations related to hyperbolic spaces. J. Funct.
Anal., 147(1):109 -- 139, 1997.
[11] G. van Dijk and S. C. Hille. Maximal degenerate representations, Berezin kernels and canoni-
cal representations. In Lie groups and Lie algebras, volume 433 of Math. Appl., pages 285 -- 298.
Kluwer Acad. Publ., Dordrecht, 1998.
[12] G. van Dijk and V. F. Molchanov. Tensor products of maximal degenerate series representa-
tions of the group SL(n, R). J. Math. Pures Appl. (9), 78(1):99 -- 119, 1999.
[13] G. van Dijk and A. Pasquale. Canonical representations of Sp(1, n) associated with represen-
tations of Sp(1). Comm. Math. Phys., 202(3):651 -- 667, 1999.
[14] A. H. Dooley and G. Zhang. Generalized principal series representations of SL(1 + n, C).
Proc. Amer. Math. Soc., 125(9):2779 -- 2787, 1997.
THE COSINE TRANSFORM
19
[15] G. I. Eskin. Boundary value problems for elliptic pseudodifferential equations, volume 52
of Translations of Mathematical Monographs. American Mathematical Society, Providence,
R.I., 1981.
[16] J. Faraut and A. Kor´anyi. Analysis on symmetric cones. Oxford Mathematical Monographs.
The Clarendon Press Oxford University Press, New York, 1994.
[17] G. B. Folland. A course in abstract harmonic analysis. Studies in Advanced Mathematics.
CRC Press, Boca Raton, FL, 1995.
[18] R. L. Frank and E. H. Lieb. Inversion positivity and the sharp Hardy-Littlewood-Sobolev
inequality. Calc. Var. Partial Differential Equations, 39(1-2):85 -- 99, 2010.
[19] R. L. Frank and E. H. Lieb. Spherical reflection positivity and the Hardy-Littlewood-Sobolev
inequality. concentration, functional inequalities and isoperimetry. Contemp. Math., 545:89 --
102, 2011.
[20] J. H. G. Fu. Algebraic integral geometry. arXiv: 1103.6256v2, 2012.
[21] A. D. Gadzhiev. Differential properties of the symbol of a singular operator in spaces of
Bessel potentials on a sphere. Izv. Akad. Nauk Azerbaıdzhan. SSR Ser. Fiz.-Tekhn. Mat.
Nauk, 3(1):134 -- 140, 1982. in Russian.
[22] A. D. Gadzhiev. Exact theorems on multipliers of spherical expansions and some of their
applications. In Special problems in function theory, No. IV (Russian), pages 73 -- 100. "`Elm",
Baku, 1989.
[23] R. J. Gardner. Geometric tomography, volume 58 of Encyclopedia of Mathematics and its
Applications. Cambridge University Press, Cambridge, 1995.
[24] R. J. Gardner and A. A. Giannopoulos. p-cross-section bodies. Indiana Univ. Math. J.,
48(2):593 -- 613, 1999.
[25] S. S. Gelbart. A theory of Stiefel harmonics. Trans. Amer. Math. Soc., 192:29 -- 50, 1974.
[26] I. M. Gel′fand, M. I. Graev, and R. Ro¸su. The problem of integral geometry and intertwining
operators for a pair of real Grassmannian manifolds. J. Operator Theory, 12(2):359 -- 383,
1984.
[27] I. M. Gel′fand and Z. Ya. Sapiro. Homogeneous functions and their extensions. Uspehi Mat.
Nauk (N.S.), 10(3(65)):3 -- 70, 1955. in Russian.
[28] S.G. Gindikin. Analysis on homogeneous domains. Russian Math. Surveys, 19(4):1 -- 89, 1964.
[29] P. Goodey and R. Howard. Processes of flats induced by higher-dimensional processes. Adv.
Math., 80(1):92 -- 109, 1990.
[30] P. Goodey, V. Yaskin, and M. Yaskina. Fourier transforms and the Funk-Hecke theorem in
convex geometry. J. Lond. Math. Soc. (2), 80(2):388 -- 404, 2009.
[31] E. L. Grinberg and B. Rubin. Radon inversion on Grassmannians via Garding-Gindikin frac-
tional integrals. Ann. of Math. (2), 159(2):783 -- 817, 2004.
[32] H. Groemer. Geometric applications of Fourier series and spherical harmonics, volume 61 of
Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge,
1996.
[33] Harish-Chandra. Harmonic analysis on real reductive groups iii. the Maass-Selberg relations
and the Plancherel formula. Ann. of Math. (2), 104(1):117 -- 201, 1976.
[34] S. Helgason. Differential geometry, Lie groups, and symmetric spaces, volume 80 of Pure and
Applied Mathematics. Academic Press Inc., New York, 1978.
[35] S. Helgason. Groups and geometric analysis. Integral geometry, invariant differential opera-
tors, and spherical functions, volume 83 of Mathematical Surveys and Monographs. American
Mathematical Society, Providence, RI, 2000.
[36] P. E. T. Jorgensen and G. ´Olafsson. Unitary representations of Lie groups with reflection
symmetry. J. Funct. Anal., 158(1):26 -- 88, 1998.
[37] P. E. T. Jorgensen and G. ´Olafsson. Unitary representations and Osterwalder-Schrader du-
ality. In The mathematical legacy of Harish-Chandra (Baltimore, MD, 1998), volume 68 of
Proc. Sympos. Pure Math., pages 333 -- 401. Amer. Math. Soc., Providence, RI, 2000.
[38] M. Kanter. The Lp norm of sums of translates of a function. Trans. Amer. Math. Soc.,
79:35 -- 47, 1973.
[39] S. P. Kh`ekalo. Riesz potentials in the space of rectangular matrices, and the iso-Hyugens
deformation of the Cayley-Laplace operator. Dokl. Math., 63(1):35 -- 37, 2001. Translated from
Dokl. Akad. Nauk, 376(2): 168 -- 170.
20
G. ´OLAFSSON, A. PASQUALE, AND B. RUBIN
[40] S. P. Kh`ekalo. The Cayley-Laplace differential operator on the space of rectangular matrices.
Izv. Math., 61(1):191 -- 219, 2005. Translated from Izv. Ross. Akad. Nauk Ser. Mat., 69:1,
195 -- 224, 2005.
[41] S. P. Kh`ekalo. The Igusa zeta function associated with a complex power function on the space
of rectangular matrices. Math. Notes, 78(5):719 -- 734, 2005. Translated from Mat. Zametki,
78:5, 773 -- 791, 2005.
[42] A. W. Knapp. Representation theory of semisimple groups, volume 36 of Princeton Math-
ematical Series. Princeton University Press, Princeton, NJ, 1986. An overview based on
examples.
[43] A. W. Knapp and E. M. Stein. Intertwining operators for semisimple groups. Ann. of Math.,
93:489 -- 578, 1971.
[44] A. W. Knapp and E. M. Stein. Intertwining operators for semisimple groups. II. Invent.
Math., 60(1):9 -- 84, 1980.
[45] A. Koldobsky. Inverse formula for the Blaschke-Levy representation. Houston J. Math.,
23(1):95 -- 108, 1997.
[46] A. Koldobsky. Fourier analysis in convex geometry, volume 116 of Mathematical Surveys and
Monographs. American Mathematical Society, Providence, RI, 2005.
[47] A. Koldobsky and H. Konig. Aspects of the isometric theory of Banach spaces. In Handbook
of the geometry of Banach spaces, Vol. I, pages 899 -- 939. North-Holland, Amsterdam, 2001.
[48] V. S. Kryuchkov. Differential properties of the symbol of the singular integral Calder´on-
Zygmund operator. Trudy Mat. Inst. Steklov., 170:148 -- 160, 276, 1984. Studies in the theory
of differentiable functions of several variables and its applications, X.
[49] P. Levy. Th´eorie de l'addition des variables al´eatoires. Gauthier-Villars, 1937.
[50] E. Lutwak. Centroid bodies and dual mixed volumes. Proc. London Math. Soc. (3), 60(2):365 --
391, 1990.
[51] G. Matheron. Un th´eor`eme d'unicit´e pour les hyperplans poissoniens. J. Appl. Probability,
11:184 -- 189, 1974.
[52] S. Meda and R. Pini. Spherical convolution with kernels having singularities on an equator.
Boll. Un. Mat. Ital. B (7), 5(2):275 -- 290, 1991.
[53] K.-H. Neeb and G. ´Olafsson. Reflection positivity and conformal symmetry. arXiv:1206.2039,
2012.
[54] A. Neyman. Representation of Lp-norms and isometric embedding in Lp-spaces. Israel J.
Math., 48(2-3):129 -- 138, 1984.
[55] G. ´Olafsson. Fourier and Poisson transformation associated to a semisimple symmetric space.
Invent. Math., 90(3):605 -- 629, 1987.
[56] G. ´Olafsson and A. Pasquale. On the meromorphic extension of the spherical functions on
noncompactly causal symmetric spaces. J. Funct. Anal., 181(2):346 -- 401, 2001.
[57] G. ´Olafsson and A. Pasquale. The Cosλ and Sinλ transforms as intertwining operators be-
tween generalized principal series representations of SL(n+1, K). Adv. Math., 229(1):267 -- 293,
2012.
[58] E. Ournycheva and B. Rubin. Composite cosine transforms. Mathematika, 52(1-2):53 -- 68,
2005.
[59] E. Ournycheva and B. Rubin. The composite cosine transform on the Stiefel manifold and
generalized zeta integrals. In Integral geometry and tomography, volume 405 of Contemp.
Math., pages 111 -- 133. Amer. Math. Soc., Providence, RI, 2006.
[60] A. Pasquale. Maximal degenerate representations of SL(n + 1, H). J. Lie Theory, 9(2):369 --
382, 1999.
[61] B. A. Plamenevskiı. Algebras of pseudodifferential operators, volume 43 of Mathematics and
its Applications (Soviet Series). Kluwer Academic Publishers Group, Dordrecht, 1989.
[62] B. Rubin. Fractional calculus and wavelet transforms in integral geometry. Fract. Calc. Appl.
Anal., 1(2):193 -- 219, 1998.
[63] B. Rubin. Inversion of fractional integrals related to the spherical Radon transform. J. Funct.
Anal., 157(2):470 -- 487, 1998.
[64] B. Rubin. Fractional integrals and wavelet transforms associated with Blaschke-Levy repre-
sentations on the sphere. Israel J. Math., 114:1 -- 27, 1999.
[65] B. Rubin. Inversion and characterization of the hemispherical transform. J. Anal. Math.,
77:105 -- 128, 1999.
THE COSINE TRANSFORM
21
[66] B. Rubin. Inversion formulas for the spherical Radon transform and the generalized cosine
transform. Adv. in Appl. Math., 29(3):471 -- 497, 2002.
[67] B. Rubin. Radon, cosine and sine transforms on real hyperbolic space. Adv. Math.,
170(2):206 -- 223, 2002.
[68] B. Rubin. Notes on Radon transforms in integral geometry. Fract. Calc. Appl. Anal., 6(1):25 --
72, 2003.
[69] B. Rubin. Intersection bodies and generalized cosine transforms. Adv. Math., 218(3):696 -- 727,
2008.
[70] B. Rubin. Funk, cosine, and sine transforms on stiefel and grassmann manifolds. To appear
in J. of Geom. Anal., 2012.
[71] B. Rubin and G. Zhang. Generalizations of the Busemann-Petty problem for sections of
convex bodies. J. Funct. Anal., 213(2):473 -- 501, 2004.
[72] W. Rudin. Lp-isometries and equimeasurability. Indiana Univ. Math. J., 25(3):215 -- 228, 1976.
[73] S. G. Samko. Generalized Riesz potentials and hypersingular integrals with homogeneous
characteristics; their symbols and inversion. Trudy Mat. Inst. Steklov., 156:157 -- 222, 263,
1980. Studies in the theory of differentiable functions of several variables and its applications,
VIII.
[74] S. G. Samko. Singular integrals over a sphere and the construction of the characteristic from
the symbol. Izv. Vyssh. Uchebn. Zaved. Mat., (4):28 -- 42, 1983.
[75] R. Schneider. Convex bodies:
the Brunn-Minkowski theory, volume 44 of Encyclopedia of
Mathematics and its Applications. Cambridge University Press, Cambridge, 1993.
[76] V. I. Semjanistyı. Some integral transformations and integral geometry in an elliptic space.
Trudy Sem. Vektor. Tenzor. Anal., 12:397 -- 441, 1963.
[77] E. Spodarev. On the rose of intersections of stationary flat processes. Adv. in Appl. Probab.,
33(3):584 -- 599, 2001.
[78] E. Spodarev. Cauchy-Kubota-type integral formulae for the generalized cosine transforms.
Izv. Nats. Akad. Nauk Armenii Mat., 37(1):52 -- 69 (2003), 2002.
[79] R. S. Strichartz. Convolutions with kernels having singularities on a sphere. Trans. Amer.
Math. Soc., 148:461 -- 471, 1970.
[80] R. S. Strichartz. The explicit Fourier decomposition of L2(SO(n)/SO(n − m)). Canad. J.
Math., 27:294 -- 310, 1975.
[81] M. Takeuchi. Modern spherical functions, volume 135 of Translations of Mathematical Mono-
graphs. American Mathematical Society, Providence, RI, 1994. Translated from the 1975
Japanese original.
[82] T. Ton-That. Lie group representations and harmonic polynomials of a matrix variable. Trans.
Amer. Math. Soc., 216:1 -- 46, 1976.
[83] D. A. Vogan, Jr. and N. R. Wallach. Intertwining operators for real reductive groups. Adv.
Math., 82(2):203 -- 243, 1990.
[84] J. A. Wolf. Harmonic analysis on commutative spaces, volume 142 of Mathematical Surveys
and Monographs. American Mathematical Society, Providence, RI, 2007.
[85] G. Zhang. Radon transform on real, complex, and quaternionic Grassmannians. Duke Math.
J., 138(1):137 -- 160, 2007.
[86] G. Zhang. Radon, cosine and sine transforms on Grassmannian manifolds. Int. Math. Res.
Not. IMRN, (10):1743 -- 1772, 2009.
Department of Mathematics, Louisiana State University, Baton Rouge, LA, 70803
USA
E-mail address: [email protected]
Laboratoire de Math´ematiques et Applications de Metz (UMR CNRS 7122), Univer-
sit´e de Lorraine, 57045 Metz cedex 1, France
E-mail address: [email protected]
Department of Mathematics, Louisiana State University, Baton Rouge, LA, 70803
USA
E-mail address: [email protected]
|
1211.6933 | 1 | 1211 | 2012-11-29T14:46:37 | Quasiadditivity of variational capacity | [
"math.FA"
] | We study the quasiadditivity property (a version of superadditivity with a multiplicative constant) of variational capacity in metric spaces with respect to Whitney type covers. We characterize this property in terms of a Mazya type capacity condition, and also explore the close relation between quasiadditivity and Hardy's inequality. | math.FA | math |
QUASIADDITIVITY OF VARIATIONAL CAPACITY
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
Abstract. We study the quasiadditivity property (a version of superadditivity with a
multiplicative constant) of variational capacity in metric spaces with respect to Whitney
type covers. We characterize this property in terms of a Mazya type capacity condition,
and also explore the close relation between quasiadditivity and Hardy's inequality.
1. Introduction
Given an open set Ω ⊂ Rn and a subset E ⊂ Ω, the relative p-capacity capp(E, Ω)
measures the minimal energy needed by a Sobolev function that vanishes on ∂Ω to take
on a value at least 1 on E. In potential theory this quantity can also be seen as a measure
of the amount of rectifiable curves connecting E to ∂Ω. Hence, greater the amount of ∂Ω
that is close to E the larger the relative p-capacity is.
It can be seen that E 7→ capp(E, Ω) is an outer measure on subsets of Ω. In particular,
capacity is countably sub-additivite: if Ek ⊂ Ω, k ∈ I ⊂ N, then
(1)
capp(cid:16)[k∈I
Ek, Ω(cid:17) ≤Xk∈I
capp(Ek, Ω).
Unlike for Borel regular measures, the equality in (1) does not (usually) hold even for nice,
well-separated sets. Indeed, the only sets that are measurable with respect to capp-outer
measure are sets of zero capacity and their complements, see for example [30, Theorem 4.8]
or [10, Theorem 2]. Nevertheless, in some cases a converse to (1), with a multiplicative
constant, can be shown to hold for certain of unions of sets; this is called the quasiadditivity
property of capacity. More precisely, we say that the p-capacity relative to an open set Ω
is quasiadditive with respect to a given cover (or a decomposition) W of Ω if there is a
constant A > 0 such that
for all E ⊂ Ω.
capp(E ∩ Q, Ω) ≤ A capp(E, Ω)
XQ∈W
The quasiadditivity property (for the linear case p = 2) was first considered by Landkof
[22, Lemma 5.5] (without the name) and Adams [1] for Riesz (and Bessel) capacities
with respect to annular decompositions of Rn \ {0}. Aikawa generalized these results
in [2], where he showed that if the complement Rn \ Ω has a sufficiently small dimension
(formulated in terms of a local version of packing condition), then the Riesz capacity of
Rn is quasiadditive with respect to Whitney decompositions of Ω. On the other hand, in [3]
Aikawa considered the Green capacity (obtained via the Green energy) and demonstrated
that if Rn \ Ω is uniformly regular (or, equivalently, uniformly 2-fat), then the Green
capacity is quasiadditive with respect to Whitney decompositions of Ω. Note that in this
case, conversely to the result of [2], the complement Rn \ Ω has a large dimension. A good
survey of these topics in the Euclidean setting can be found in [5, Section 7 of Part II].
Date: September 25, 2018.
2000 Mathematics Subject Classification. Primary 31E05, 31C45; Secondary 46E35, 26D15.
The first author acknowledges the support of the Academy of Finland, grant no. 252108. The second
author was partially supported by a grant from the Simons Foundation, Collaboration Grant # 200474
and NSF grant DMS-1200915.
1
2
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
See also [5, Section 16 of Part I] and [4] for related results for nonlinear (p 6= 2) setting for
which decompositions other than the Whitney decomposition are used.
The aim of this note is to study the quasiadditivity problem for the relative p-capacity
with respect to Whitney type covers in the setting of complete metric measure spaces
satisfying the 'standard' structural assumptions (see Section 2.1). Nevertheless, most of
our results are new for p 6= 2 (and for p = 2, obtained via new methods) even in Eu-
clidean spaces. Part of our motivation stems from a need to clarify the relation between
quasiadditivity and Hardy's inequality (a Sobolev-type inequality weighted with a power
of the distance-to-boundary function; see Section 3.2). Glimpses of a connection between
these concepts (and the related dimension bound of Aikawa from [2]) have appeared e.g.
in [3, 5, 21, 23, 25, 32], but now our main result -- Theorem 3.3, a characterization of
quasiadditivity in terms of a Mazya type capacity estimate -- reveals a simple equivalence
between quasiadditivity and Hardy's inequalities (Corollary 3.5). These results also link
quasiadditivity and the geometry of the boundary (or the complement) of the open set Ω.
The organization of this note is as follows. In Section 2 we recall some of the necessary
background material: The basic assumptions, the notions of (co)dimension for metric
spaces, Sobolev type spaces and the related capacities, and Whitney covers (substitutes
of the classical Whitney decompositions for our more general spaces). Since our proofs
are largely based on potential-theoretic (rather than PDE-based) tools, an overview of
these is given at the end of Section 2; of a particular importance for us is the weak
Harnack inequality for superminimizers. Section 3 contains our main characterization of
quasiadditivity and the aforementioned connection with Hardy's inequalities. A concrete
outcome of these considerations is that the uniform p-fatness of X \ Ω guarantees the
quasiadditivity for the relative p-capacity in Ω for 1 < p < ∞.
Aikawa's dimension bound dimA(Rn \ Ω) < n − p from [2] translates to more general
metric spaces as co-dimA(X \ Ω) > p. We show in Section 4 that this bound, together
with an additional discrete John type condition, is sufficient for the relative p-capacity to
be quasiadditive with respect to Whitney covers of Ω. Finally, in Section 5 we explain
how the results involving a large complement (uniform p-fatness) or a small complement
(Aikawa's condition) can be combined, allowing us to deal with more general open sets
whose complements consist of parts of different sizes.
For the notation we remark that C will denote positive constants whose value is not
If there exist constants c1, c2 > 0 such that
necessarily the same at each occurrence.
c1 F ≤ G ≤ c2F , we sometimes write F ≈ G and say that F and G are comparable.
2. Preliminaries
2.1. Doubling metric spaces. We assume throughout this note that X = (X, d, µ) is
a complete metric measure space, where µ is a Borel measure supported on X, with
0 < µ(B) < ∞ whenever B = B(x, r) is an open ball in X, and that µ is doubling, that is,
there is a constant C > 0 such that whenever x ∈ X and r > 0, we have
µ(B(x, 2r)) ≤ C µ(B(x, r)).
We make the tacit assumption that each ball B ⊂ X has a fixed center xB and radius
rad(B), and thus notation such as λB = B(xB, λ rad(B)) is well-defined for all λ > 0.
If µ is a doubling measure, then by iterating the doubling condition we find constants
Q > 0 and C > 0 such that
µ(B(y, r))
µ(B(x, R))
≥ C(cid:16) r
R(cid:17)Q
whenever 0 < r ≤ R < diam X and y ∈ B(x, R). Furthermore, if X is connected (this is
guaranteed in our setting by the below-mentioned Poincaré inequalities), then there exists
QUASIADDITIVITY OF VARIATIONAL CAPACITY
3
a constant Qu > 0 such that for all 0 < r < R < diam X and y ∈ B(x, R),
(2)
µ(B(y, r))
µ(B(x, R))
≤ C(cid:16) r
R(cid:17)Qu
.
In general, 1 ≤ Qu ≤ Q. However, if we have uniform upper and lower bounds for the
measures of the balls, i.e.
c−1rQ ≤ µ(B(x, r)) ≤ crQ
for every x ∈ X and all 0 < r < diam(X), we say that the measure µ is (Ahlfors) Q-regular.
When working with a (non-regular) doubling measure µ, it is often convenient to de-
scribe the sizes of sets in terms of codimensions (instead of dimensions). For instance, the
Hausdorff codimension of E ⊂ X (with respect to µ) is the number
where
co-dimH(E) = sup(cid:8)q ≥ 0 : Hµ,q
∞ (E) = inf(cid:26)Xk
∞ (E) = 0(cid:9),
rad(Bk)−qµ(Bk) : E ⊂[k
Hµ,q
Bk(cid:27)
is the q-codimensional Hausdorff content. If µ is Q-regular, then we have for all E ⊂ X
that Q − co-dimH(E) = dimH(E), the usual Hausdorff dimension.
Another notion of codimension that will be useful for us is the Aikawa codimension:
For E ⊂ X we define co-dimA(E) as the supremum of all q ≥ 0 for which there exists a
constant Cq such that
ZB(x,r)
dist(y, E)−q dµ(y) ≤ Cqr−qµ(B(x, r))
for every x ∈ E and all 0 < r < diam(E). Here we interpret the integral to be +∞ if q > 0
and E has positive measure. It is not hard to see that co-dimA(E) ≤ co-dimH(E) for all
E ⊂ X (cf. [25]). If µ is Ahlfors Q-regular, then we could define the Aikawa dimension
of a set E ⊂ X to be the number dimA(E) = Q − co-dimA(E). Nevertheless, it was
shown in [25] that for subsets of Ahlfors regular metric measure spaces this concept is
actually equal to the Assouad dimension of the subset; see [26] for the basic properties of
the Assouad dimension.
2.2. Sobolev-type function spaces in the metric setting. There are many analogs of
Sobolev-type function spaces in the metric setting. The one considered in this note is based
on the notion of upper gradients, generalizing the fundamental theorem of calculus. Given
a measurable function f : X → [−∞, ∞], we say that a Borel measurable non-negative
function g on X is an upper gradient of f if whenever γ is a compact rectifiable curve in
X, we have
f (y) − f (x) ≤ Zγ
g ds.
Here x, y denote the two endpoints of γ, and the above condition should be interpreted
as claiming that Rγ g ds = ∞ whenever at least one of f (x), f (y) is infinite. See [12]
and [6] for a good discussion on the notion of upper gradients. Using upper gradients as a
substitute for modulus of the weak derivative, we define the norm
kf kN 1,p(X) :=(cid:18)ZX
f p dµ + inf
g ZX
gp dµ(cid:19)1/p
,
where the infimum is taken over all upper gradients g of f . The Newtonian space N 1,p(X)
is the space
{f : X → [−∞, ∞] : kf kN 1,p(X) < ∞}/ ∼,
where the equivalence ∼ is given by u ∼ v if and only if ku − vkN 1,p(X) = 0 (see [28] or [6]
for more on this function space).
4
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
In addition to the doubling property, we will also assume throughout that the space X
supports a (1, p)-Poincaré inequality, that is, there exist constants C > 0 and λ ≥ 1 such
that whenever B = B(x, r) ⊂ X and g is an upper gradient of a measurable function f ,
we have
ZB
where
f − fB dµ ≤ C r (cid:18)ZλB
f dµ =:ZB
µ(B) ZB
fB :=
1
gp dµ(cid:19)1/p
f dµ.
Different notions of capacity are of fundamental importance in many questions related
to the behavior of the functions belonging to a certain class. Given a set E ⊂ X, the total
p-capacity of E, denoted Capp(E), is the infimum of kukp
N 1,p(X) over all functions u such
that u ≥ 1 on E. Just as sets of measure zero play the role of indeterminacy in the theory
of Lebesgue spaces Lp(X), sets of total p-capacity zero play the corresponding role in the
theory of Sobolev type spaces; see [6] or [28] for details. We say that a property holds
(p-)quasieverywhere (p-q.e.) if the exeptional set is of zero total capacity.
When the examinations are taking place in an open set Ω ⊂ X, then a more appropriate
version of capacity is the relative p-capacity. For a measurable set E ⊂ Ω this is defined as
the number
capp(E, Ω) := inf
u
inf
gu ZX
gp
u dµ,
where the infimum is taken over all u ∈ N 1,p(X) with u = 0 on X \ Ω, u = 1 on E,
0 ≤ u ≤ 1, and over all upper gradients gu of u. A function u satisfying the above
conditions is called a capacity test function for E. Should no such function u exist, we set
capp(E, Ω) = ∞.
When the variational capacity is taken with respect to Ω = X, it may be the case that
capp(E, X) = 0 for all bounded E ⊂ X; this is certainly true if X is bounded. If X is
unbounded and still capp(E, X) = 0 for all bounded E ⊂ X, then X is called p-parabolic,
but if capp(E, X) > 0 for some bounded E ⊂ X, then X is p-hyperbolic. These notions
will be relevant in the considerations of Section 4. See [14] and [15] for more on parabolic
and hyperbolic spaces. Notice in particular that if X is bounded or p-parabolic and Ω ⊂ X
is such that Capp(X \ Ω) = 0, then capp(E, Ω) = 0 for every E ⊂ Ω.
Besides measuring small (exeptional) sets, the relative capacity can also be used to give
conditions for the largeness of sets. For instance, a closed set E ⊂ X is said to be uniformly
p-fat if there exists a constant C > 0 such that
capp(E ∩ B, 2B) ≥ C capp(B, 2B)
for all balls B centered at E. Here capp(B, 2B) is actually comparable with rad(B)−pµ(B)
for all balls B of radius less than diam(X)/8. See [6, Chapter 6] for this and other basic
properties of the total and the variational capacity on metric spaces. We remark that the
uniform p-fatness can also be characterized using uniform density conditions for Hausdorff
contents; see e.g. [19].
Recall that we say the variational p-capacity capp(·, Ω) to be quasiadditive with respect
to a decomposition or a cover W of Ω if there exists a constant A > 0 such that
capp(E ∩ Q, Ω) ≤ A capp(E, Ω)
XQ∈W
for every E ⊂ Ω. In the next subsection we discuss the one particular family of covers we
are concerned with.
QUASIADDITIVITY OF VARIATIONAL CAPACITY
5
2.3. Whitney covers. Let Ω ⊂ X be an open set. We often write dΩ(x) = dist(x, X \ Ω)
for x ∈ Ω. When 0 < c ≤ 1/2, we fix a Whitney type cover Wc(Ω) = {Bi}i∈N, Ω ⊂Si∈N Bi,
controlled overlap: there exists 1 ≤ C < ∞ such that Pi χBi
consisting of balls Bi = B(xi, cdΩ(xi)), xi ∈ Ω, such that the balls Bi have uniformly
≤ C. Such a cover can
always be constructed by considering maximal packings (or, alternatively, '5r'-covers) of
the sets {x ∈ Ω : 2−k−1 ≤ dΩ(x) < 2−k}, k ∈ Z, with the balls of the above type. In
pathological situations we allow Bi = ∅ for some i, if necessary.
In our proofs, we need to be able to dilate the Whitney balls without having too much
overlap; the existence of such a cover is established in the next (elementary) lemma. For
a proof, see e.g. [7] (Theorem 3.1 together with Lemma 3.4) or [13, Chapter 3].
Lemma 2.1. Let Ω ⊂ X be an open set. Fix L ≥ 1 and let Wc(Ω) = {Bi}i∈N be a Whitney
cover of Ω with c ≤ (3L)−1. Then the balls LBi have a uniformly bounded overlap.
In the proof of our main characterization of the quasiadditivity, we will need for Whitney
balls B ∈ Wc(Ω) the estimate
(3)
c1 rad(B)−pµ(B) ≤ capp(B, Ω) ≤ c2 rad(B)−pµ(B),
where c1, c2 may depend on Wc(Ω) but not on the particular B ∈ Wc(Ω). The upper
bound in (3) is always true in our setting, and can be proved almost immediately by using
only the doubling condition and the test function u(x) = [1 − dΩ(x)/ rad(B)]+. The lower
bound is a bit more involved, and can in fact fail in some spaces satisfying our basic
assumptions. Thus, in the cases where we need the lower bound, we will need to have
some extra assumptions on Ω or X, e.g. those given in Lemma 2.2 below. However, we
stress that the lower bound in (3) is only needed in the proof of Lemma 2.3, which on
the other hand is only used to prove the Main Theorem 3.3, and then once again in the
considerations of Section 5, and thus these are the only instances where such assumptions
are needed.
Another important case when the lower bound is valid is when X \ Ω is uniformly p-
fat; the bound then follows easily (e.g.) from the p-Hardy inequality (see Section 3.2) for
capacity test functions of B ∈ Wc(Ω). In this case we only need the standing assumptions
that µ is doubling and X supports a (1, p)-Poincaré inequality.
If we want to weaken the assumption on X \ Ω, we need to assume more on µ and p. In
the following lemma we chose the condition that X is unbounded (in which case we have
µ(X) = ∞ by (2)). However, if X happens to be bounded, then we could impose a further
condition on Ω instead, such as diam(Ω) < 2 diam(X), in which case the constants depend
on γ = µ(Ω)/µ(X) < 1. The proof in this case is similar to the one below. Recall here
that Qu is the exponent from the 'upper mass bound' (2).
Lemma 2.2. Assume that 1 ≤ p < Qu and X is unbounded. Then
c rad(B)−pµ(B) ≤ capp(B, Ω)
for all Whitney balls B ∈ Wc(Ω).
Proof. Let u ∈ N 1,p(Ω) be a capacity test function for B = B(x, r) ∈ Wc(Ω), and let
g ∈ Lp(Ω) be an upper gradient of u. For positive integers j let Bj = 2jB and rj = 2jr.
As u ∈ Lp(X) and µ(X) = ∞ by the unboundedness of X (here we use (2)), there exists
K ∈ N (depending on u) such that uBK < 1/2. On the other hand, because u ≥ 1 on
B we know that uB = 1. Now a standard telescoping argument using the (1, p)-Poincaré
inequality yields
(4)
1
2 ≤ uB − uBK ≤ C
K
Xj=1
rj(cid:18)ZλBj
gp dµ(cid:19)1/p
.
6
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
It follows that there exists a constant C0 > 0 and some 1 ≤ j0 ≤ K such that
(r/rj0)(Qu−p)/p = (2−j0)(Qu−p)/p ≤ C0rj0µ(λBj0)−1/p(cid:18)ZλBj0
gp dµ(cid:19)1/p
(otherwise (4) would lead to a contradicition when compared to a geometric series). Thus
we obtain, using also (2) and the assumption 1 < p < Qu, that
ZΩ
as desired.
gp dµ ≥ C(r/rj0)Qu−pr−p
j0
µ(λBj0) ≥ Cr−pµ(B),
(cid:3)
Let us record the following easy consequence of estimate (3) for unions of Whitney balls.
Lemma 2.3. Let Ω ⊂ X be an open set and let Wc(Ω) = {Bi}i∈N be a Whitney cover of
Ω for which the lower bound in (3) holds. Furthermore, let U ⊂ Ω be a union of Whitney
balls, i.e., U =Si∈I Bi for some I ⊂ N. Then
dΩ(x)−p dµ ≈Xi∈I
ZU
capp(Bi, Ω).
Proof. This follows from the fact that dΩ(x) ≈ rad(Bi) for all x ∈ Bi, with constants only
depending on c, and that capp(Bi, Ω) ≈ rad(Bi)−pµ(Bi) by (3).
(cid:3)
2.4. Existence and properties of p-potentials. In computing the relative p-capacity
capp(E, Ω), should this capacity be finite, we can find a minimzing sequence of capacity
test functions uk ∈ N 1,p(X), i.e.,
lim
k→∞
gp
uk dµ = capp(E, Ω).
inf
gukZX
We will assume throughout that 1 < p < ∞; hence Lp(X) is reflexive, and so a standard
variational argument using Mazur's lemma on Lp(X) (as in Lemma 2.4 below) tells us that
if Ω ⊂ X is bounded and Capp(X \ Ω) > 0, then there is a p-potential u ∈ N 1,p(X) such
that 0 ≤ u ≤ 1 on X, u = 1 on E, u = 0 on X \ Ω, and
inf
gu ZΩ
gp
u dµ = inf
gu ZX
gp
u dµ = capp(E, Ω).
Such a p-potential is unique because X supports a (1, p)-Poincaré inequality; see [29] for
more details. Nevertheless, the following more general lemma tells us that a p-potential
u ∈ N 1,p
loc (X) exists in more general cases (e.g. if Capp(X \ Ω) = 0) as well; though, if X is
p-parabolic, we would have u be a constant. In addition, the below proof shows that the
reflexivity of N 1,p(X) is actually not needed for the existence of p-potentials.
Lemma 2.4. Let 1 < p < ∞ and let E ⊂ Ω be such that capp(E, Ω) < ∞. Then there is
a function u ∈ N 1,p
loc (X) such that u = 1 p-q.e. on E, u = 0 p-q.e. in X \ Ω, 0 ≤ u ≤ 1 on
X, and
inf
gu ZΩ
gp
u dµ = capp(E, Ω).
Proof. If Ω is bounded, then the following proof can be easily modified, or the results
of [29] can be used to obtain the desired conclusion. Hence here we will only give the proof
for the case that Ω, and hence X, is unbounded.
For each u ∈ N 1,p(X) there is a minimal (p-weak) upper gradient gu ∈ Lp(X); see for
example [6]. Hence, from now on, we let gu denote this minimal upper gradient of u.
QUASIADDITIVITY OF VARIATIONAL CAPACITY
7
Let {uk}k∈N be a sequence of functions in N 1,p(X) that satisfy 0 ≤ uk ≤ 1 on X, uk = 0
on X \ Ω p-q.e., uk = 1 p-q.e. on E, and
lim
k→∞ZΩ
gp
uk dµ = capp(E, Ω).
Fix x0 ∈ Ω and for each positive integer n let Bn = B(x0, n). Given that 0 ≤ uk ≤ 1,
the sequences {uk}k and {guk }k are bounded in Lp(Bn) for each positive integer n, and
hence because 1 < p < ∞, the uniform convexity of Lp(Bn) together with a Cantor
diagonalization argument tells us that {uk}k converges weakly to a function u in Lp
loc(X)
and that {guk }k converges weakly to g in Lp
Finally, an application of [16, Lemma 3.1] to {uk}k and {guk }k in Bn allows us to modify
u on a set of measure zero to obtain a function u ∈ Lp
loc(X) that has g as a p-weak upper
gradient, and furthermore, from [16, Lemma 3.1] and [28, Proof of Theorem 3.7], we can
conclude that u = 1 p-q.e. on E, u = 0 p-q.e. on X \ Ω, and that
loc(X).
ZΩ
gp
u dµ ≤ZΩ
k→∞ZΩ
gp dµ ≤ lim
gp
uk dµ = capp(E, Ω).
This, together with the definition of capp(E, Ω) now completes the proof.
(cid:3)
The results from [18] show that such u, if non-constant, satisfies u > 0 on Ω with u < 1
on X \ E. Of course, should capp(E, Ω) be infinite, then no such u exists.
It is clear that the p-potential u, corresponding to E ⊂ X, has the property that u is a
p-superminimizer in Ω and a p-subminimizer in X \ E; in particular, u is a p-minimizer
in Ω \ E. Here, we say that a function v ∈ N 1,p
loc (X) is a p-superminimizer in an open set
U ⊂ X if, whenever ϕ ∈ N 1,p(X) is a non-negative function such that ϕ = 0 on X \ U , we
have
inf
gv Zsupt(ϕ)
gp
v dµ ≤ inf
gv+ϕZsupt(ϕ)
gp
v+ϕ dµ,
and v is a p-subminimizer in U if −v is a p-superminimizer in U . We refer the interested
reader to [17] for information on minimizers; see also [6]. In particular, it is known that
if v is a p-superminimizer in U and w ∈ N 1,p
loc (X) is a p-minimizer in U such that w ≤ v
holds p-q.e. on X \ U , then w ≤ v on U as well. This is the so-called comparison principle.
Notice also that if Capp(X \ Ω) = 0 and v ∈ N 1,p
loc (X) is a minimizer in Ω, then v is a
minimizer in X; moreover, in this case, if u is a p-potential for capp(E, Ω) then it is a
p-potential for capp(E, X).
In the proofs of our results the following weak Harnack inequality for p-superminimizers
is of fundamental importance. See [18] for a proof of this lemma.
Lemma 2.5 (Weak Harnack). There exists constants A > 0, CA ≥ 1, and q > 0 such that
if u is a p-superminimizer in CAB ⊂ Ω, then
(cid:18)Z2B
uq dµ(cid:19)1/q
≤ A ess inf
B
u.
3. Characterizations of quasiadditivity
In this section we prove the main result of this note, Theorem 3.3, and provide a con-
nection between quasiadditivity and p-Hardy inequalities. Recall that we always assume
that 1 < p < ∞.
3.1. The main characterization. We begin by showing that quasiadditivity property
for unions of balls is in fact sufficient for the quasiadditivity for general sets. Below CA is
the constant from the weak Harnack inequality and λ is the dilatation constant from the
p-Poincaré inequality.
Xi∈I
8
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
Proposition 3.1. Let Ω ⊂ X be an open set with Ω 6= X and let Wc(Ω) = {Bi}i∈I be
a Whitney cover of Ω with c < min{(CA)−1, (30λ)−1}. Assume that the quasiadditivity
condition holds for unions of Whitney balls, i.e., if U = Si∈I Bi for some I ⊂ N, Bi ∈
Wc(Ω), then
capp(Bi, Ω) ≤ C1 capp(U, Ω).
Then the capacity capp(·, Ω) is quasiadditive with respect to Wc(Ω), i.e., there exists a
constant C > 0 such that
for every E ⊂ Ω.
capp(E ∩ Bi, Ω) ≤ C capp(E, Ω)
Xi∈N
Proof. The structure of the proof is based on the idea of Aikawa [2], but given the non-
linear nature of our setting, the tools we employ are completely different. Let E ⊂ Ω. If the
relative capacity capp(E, Ω) is infinite, then the desired inequality would follow. Therefore
we assume that capp(E, Ω) < ∞, and let u ∈ N 1,p
loc (Ω) be the p-potential corresponding
to this capacity. If capp(E, Ω) = 0, then by the monotonicity of capp(·, Ω), each term in
the sum on the left-hand side of the desired inequality is also zero, and the claim follows.
Therefore we will assume that capp(E, Ω) > 0, and so the p-potential u is non-constant.
Write Ei = E ∩ Bi. Choose C0 = (cid:0) 1
weak Harnack inequality (Lemma 2.5). We divide the union E =Si∈N Ei into the following
two parts: If u(x) ≥ C0 for q.e. x ∈ Bi, then i ∈ I1, and otherwise i ∈ I2. Note also that
the indices i for which capp(Ei, Ω) = 0 do not contribute to the sum on the left-hand side
of the desired quasiadditivity inequality. Hence in the following argument we only consider
the indices i for which capp(Ei, Ω) > 0.
A , where q and A are the constants from the
4(cid:1)1/q 1
It is immediate that u
C0
is an admissible test function for
Thus, using the assumption that quasiadditivity holds for unions of Whitney balls, we
obtain for all upper gradients gu of u that
capp(cid:16) [i∈I1
Bi, Ω(cid:17).
(5)
Xi∈I1
capp(Ei, Ω) ≤ Xi∈I1
On the other hand, if i ∈ I2, then by the weak Harnack inequality
Bi, Ω(cid:17) ≤ C1(cid:16) 1
C0(cid:17)pZΩ
gp
u dµ.
1
2 µ(2Bi).
Now write v = 1 − u, whence v ∈ N 1,p
Since 0 ≤ uq ≤ 1, it follows that for the set Ui = (cid:8)x ∈ 2Bi : uq ≥ 1
precisely the class of upper gradients of v as well. Also, Ui =(cid:8)x ∈ 2Bi : v ≤ 1 −(cid:0) 1
2(cid:9) we have µ(Ui) ≤
2(cid:1)1/q(cid:9),
loc (X) and the class of upper gradients of u is
and so
This gives a positive lower bound c1 for the mean-value of vp in 2Bi;
2 µ(2Bi) ≤ µ(2Bi \ Ui) =(cid:8)x ∈ 2Bi : v ≥ 1 −(cid:0) 1
2(cid:1)1/q(cid:9).
1
(vp)2Bi ≥ 1
2(cid:16)1 − 1
21/q(cid:17)p
=: c1.
capp(Bi, Ω) ≤ C1 capp(cid:16) [i∈I1
uq dµ(cid:19)1/q
≤ A ess inf
Bi
u < C0A,
uqdµ ≤ (C0A)qµ(2Bi) = 1
4 µ(2Bi).
(cid:18)Z2Bi
Z2Bi
and thus
(6)
QUASIADDITIVITY OF VARIATIONAL CAPACITY
9
We can now use the well-known Mazya's version of the (Sobolev -- )Poincaré inequality (see
e.g. [27, Chapter 10], and [8, Proposition 3.2] for the metric space version):
c1 <Z2Bi
vp dµ ≤
C
capp(Bi ∩ {v = 0}, 2Bi)Z10λBi
gp
v dµ,
where gv is an arbitrary upper gradient of v (and thus of u as well). Since Ei = Bi ∩{v = 0}
by the comparison principle, it follows from (7) that
capp(Ei, Ω) ≤ capp(Ei, 2Bi) = capp(cid:0)Bi ∩ {v = 0}, 2Bi(cid:1) ≤ C ′Z10λBi
gp
v dµ.
Using this and the fact that the balls 10λBi do not overlap too much, guaranteed by our
choice of the parameter c (with L ≥ 10λ) and Lemma 2.1, we conclude that
Xi∈I2
capp(Ei, Ω) ≤ C ′Xi∈I2Z10λBi
gp
v dµ ≤ CZΩ
gp
v dµ.
The claim now follows by taking the infima over all upper gradients of u in (5) and (8)
(cid:3)
and combining these two estimates.
The next lemma can be seen as a counterpart of Proposition 3.1 for a Mazya-type
condition (cf. [27, §2.3]):
Proposition 3.2. Let Ω ⊂ X be an open set and let Wc(Ω) = {Bi}i∈I be a Whitney cover
of Ω with c < min{(CA)−1, (30λ)−1}. Assume the existence of a constant C0 > 0 such that
(7)
(8)
(9)
dΩ(x)−p dµ(x) ≤ C0 capp(U, Ω)
dΩ(x)−p dµ(x) ≤ C capp(E, Ω)
ZU
ZE
whenever U ⊂ Ω is a union of Whitney balls. Then there exists a constant C > 0 such that
whenever E ⊂ Ω.
Proof. Let E ⊂ Ω. If capp(E, Ω) = ∞ the claim follows, and thus we may again assume
that capp(E, Ω) < ∞. Let u be the p-potential of E with respect to Ω, and let gu be an
upper gradient of u. We denote Ei = E ∩ Bi and split the union E =Si Ei into two parts:
We set i ∈ I1 if u2Bi < 1/2 and i ∈ I2 if u2Bi ≥ 1/2.
In the first case i ∈ I1 we have u − u2Bi ≥ 1/2 in Ei, and so, using the (p, p)-Poincaré
inequality (a consequence of the (1, p)-Poincaré inequality by [11, Theorem 5.1]) and the
bounded overlap of the balls 10λBi, we obtain
Xi∈I1ZEi
(10)
dΩ(x)−p dµ(x) ≤ CXi∈I1
r−p
i Z2Bi
≤ CXi∈I1Z10λBi
u − u2Bip dµ
gp
u dµ ≤ CZΩ
gp
u dµ.
In the second case i ∈ I2 it follows from u2Bi ≥ 1/2 and 0 ≤ u ≤ 1 that for Ui = {u ≥
1/4} ∩ 2Bi,
1
2 ≤
≤
from which we obtain
1
1
u dµ +Z2Bi\Ui
µ(2Bi)"ZUi
u dµ#
µ(2Bi)(cid:2)µ(Ui) + 1
4 µ(2Bi \ Ui)(cid:3) ≤
µ(cid:0){x ∈ 2Bi : u(x) ≥ 1/4}(cid:1) ≥ 1
µ(Ui)
µ(2Bi)
+ 1
4 ,
4 µ(2Bi).
10
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
Thus, by the weak Harnack inequality for superminimizers, we obtain that
inf
Bi
u ≥ A−1(cid:18)Z2Bi
uq dµ(cid:19)1/q
≥ C(cid:18)µ(2Bi)−1Z{u≥1/4}∩2Bi
(1/4)q dµ(cid:19)1/q
≥ C1
for each i ∈ I2. Hence the function u/C1 is an admissible test function for
Using the bounded overlap of the balls Bi and the assumption (9), we conclude that
Xi∈I2ZEi
(11)
Bi, Ω(cid:17).
capp(cid:16) [i∈I2
dΩ(x)−p dµ(x) ≤ Xi∈I2ZBi
≤ CZSi∈I2
≤ C capp(cid:16) [i∈I2
Bi
dΩ(x)−p dµ(x)
dΩ(x)−p dµ(x)
Bi, Ω(cid:17) ≤
C
C p
1 ZΩ
gp
u dµ.
The lemma follows from estimates (10) and (11) by taking the infimum over all upper
gradients of u.
(cid:3)
Combining the conditions from Propositions 3.1 and 3.2, we arrive at the main result of
this section:
Theorem 3.3. Let Ω ⊂ X be an open set, and let Wc(Ω) = {Bi}i∈N be a Whitney cover
of Ω with c < min{(CA)−1, (30λ)−1}. Then the following conditions are (quantitatively)
equivalent:
(a) There exist C > 0 such that
for all E ⊂ Ω.
dΩ(x)−p dµ ≤ C capp(E, Ω)
ZE
(b) There exist C > 0 such that
dΩ(x)−p dµ ≤ C capp(U, Ω)
whenever U =Si∈I Bi for Bi ∈ Wc(Ω) and I ⊂ N.
(c) There exist C > 0 such that
capp(E ∩ Bi, Ω) ≤ C capp(E, Ω)
for all E ⊂ Ω, and the capacity estimate (3) holds.
(d) There exist C > 0 such that
capp(Bi, Ω) ≤ C capp(U, Ω)
ZU
Xi∈N
Xi∈I
whenever U =Si∈I Bi for Bi ∈ Wc(Ω) and I ⊂ N, and the capacity estimate (3) holds.
Proof. The implications (a) =⇒ (b) and (c) =⇒ (d) are trivial and the implications
converse to these are Proposition 3.2 and Proposition 3.1, respectively. As a link between
these two equivalences we have (b) ⇐⇒ (d) from Lemma 2.3, and here the lower bound
of (3) is needed to pass from (d) to (b). Hence we assume the validity of (3) in parts (c)
and (d). Recall that the validity of (3) is guaranteed by Lemma 2.2 when the hypotheses
of this lemma are satisfied, or by the uniform p-fatness of X \ Ω.
(cid:3)
QUASIADDITIVITY OF VARIATIONAL CAPACITY
11
3.2. The Hardy connection. We say that an open set Ω ⊂ X admits a p-Hardy inequal-
ity if there exists a constant C > 0 such that the inequality
ZΩ
u(x)p dΩ(x)−p dµ(x) ≤ CZΩ
gp
u dµ,
holds for all u ∈ N 1,p(Ω) with u = 0 on X \ Ω and for all upper gradiets gu of u.
Let us record the following Mazya-type characterization for Hardy inequalities.
Lemma 3.4. An open set Ω ⊂ X admits a p-Hardy inequality if and only if
(12)
for all E ⊂ Ω.
dΩ(x)−p dµ(x) ≤ C capp(E, Ω)
ZE
Proof. For compact sets K ⊂ Ω, the above characterization is proven in the metric space
setting in [20, Theorem 4.1] (see also [27, §2.3] in the Euclidean setting). Thus it suffices
to show that if Ω admits a p-Hardy inequality and E ⊂ Ω is an arbitrary subset, then
estimate (12) holds. If capp(E, Ω) = ∞, then there is nothing to prove, and on the other
hand if capp(E, Ω) < ∞, then the p-Hardy inequality, used for capacity test functions uk
(cid:3)
with limk→∞RΩ guk dµ = capp(E, Ω), yields the desired estimate (12).
In other words, an open set Ω ⊂ X admits a p-Hardy inequality if and only if the
assertion (a) of Theorem 3.3 holds. This leads immediately to the following corollary.
Corollary 3.5. Let Ω ⊂ X be an open set and let Wc(Ω) be a Whitney cover of Ω with
a suitably small parameter 0 < c ≤ 1/2. Then Ω admits a p-Hardy inequality if and only
if the capacity capp(·, Ω) is quasiadditive with respect to Wc(Ω) and the capacity estimate
(3) holds for all balls B ⊂ Ω.
Since uniform p-fatness of the complement X \ Ω is a sufficient condition for p-Hardy
inequalities in our setting (see [9, Corollary 6.1] and [19]), we obtain a concrete sufficient
condition for the quasiadditivity of the p-capacity:
Corollary 3.6. Let Ω ⊂ X be an open set and let Wc(Ω) be a Whitney cover of Ω with
a suitably small parameter 0 < c ≤ 1/2. Assume further that the complement X \ Ω is
uniformly p-fat. Then the capacity capp(·, Ω) is quasiadditive with respect to Wc(Ω).
4. Quasiadditivity and the Aikawa dimension
In this section we focus on open sets Ω ⊂ X that satisfy co-dimA(X \ Ω) > p (recall
In this case we also have that co-dimH(X \ Ω) > p,
the definition from Section 2.1).
and hence it follows from [24, Proposition 4.1] that Capp(X \ Ω) = 0. Therefore, as was
remarked in Section 2.4, we know that capp(E, Ω) = capp(E, X) for every E ⊂ Ω. Recall
from Section 2.2 that if X is p-parabolic, then actually capp(E, X) = 0 for all bounded
E ⊂ X. Thus, if X is p-parabolic and Ω ⊂ X is such that Capp(X \ Ω) = 0, then Ω
satisfies the quasiadditivity property trivially; the same is also true if X is bounded and
Capp(X \ Ω) = 0. Hence in this section we assume that X is unbounded and p-hyperbolic.
We say that an open set Ω = X \ E and a related Whitney cover W = Wc(Ω) satisfy a
discrete John condition if there exist L > 1, a > 1, and C > 0 such that for each B ∈ W
we find a chain C(B) = {Bm}∞
m=0 of Whitney balls Bm ∈ W(Ω), with B0 = B, such
that Bm ∩ Bm+1 6= ∅, B ⊂ LBm, and rad(Bm) ≥ Cam rad(B) for each m ∈ N. This
condition is satisfied, for instance, if Ω is an unbounded John domain (see [31]); similar
chain conditions have been used e.g. in [11, 12]. Notice in particular that since our open
sets are unbounded, there can not exist a 'John center' as in the usual John condition for
bounded domains; essentially the 'point at infinity' acts as a John center. On the other
hand, the domain Ω = {(x, y) ∈ R2 : 0 < y < x + 1} satisfies the discrete John condition,
but is not an unbounded John domain (in the sense of [31]).
12
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
Proposition 4.1. Let Ω ⊂ X be an open set with co-dimA(X\Ω) > p. Assume furthermore
that Ω satisfies the above discrete John condition for a Whitney cover Wc(Ω) with c ≤
(6λ)−1. Then capp(·, Ω) is quasiadditive with respect to Wc(Ω) and Ω admits a p-Hardy
inequality.
Proof. By Theorem 3.3 and Corollary 3.5, it suffices to show that there is a constant
C > 0 such that if U =Si∈I Bi, Bi ∈ Wc(Ω), for some I ⊂ N, then
dΩ(x)−p dµ(x) ≤ C capp(U, Ω).
(13)
ZU
Fix such a set U , and write ri = rad(Bi), i ∈ I. We may clearly assume that capp(U, Ω) <
∞. Let u be a capacity test-function for U . Then uBi = 1 for each i ∈ I (and thus u2Bi ≥ α
for some α > 0 since 0 ≤ u ≤ 1). On the other hand, since u ∈ Lp(Ω), we find, using the
discrete John condition, for each i ∈ I a chain of Whitney balls Bi,m, m = 0, 1, . . . , Mi,
where Bi,0 = Bi, Bi,m−1 ∩ Bi,m 6= ∅, rad(Bi,m) ≥ Camri for all m = 1, . . . , Mi, and
u2Bi,Mi
≤ α/2. Indeed, since rad(Bi,m) ≥ Camri, we have by (2) that
µ(Bi)
µ(Bi,m)
and thus, by Hölder's inequality,
≤ C(cid:18)
ri
rad(Bi,m)(cid:19)Qu
≤ Ca−mQu,
u ≤ Cµ(Bi,m)−1/pkukLp(X)
m→∞−−−−→ 0,
Z2Bi,m
allowing us to choose the index Mi as above.
By a standard chaining argument using the (1, p)-Poincaré inequality (see e.g.
[11] or
[12]), we have that
Mi
(14)
α/2 ≤ u2Bi − u2Bi,Mi
≤ C
rad(Bi,m)(cid:18)Z2λBi,m
gp
u dµ(cid:19)1/p
.
Xm=0
Comparing the sum on the right-hand side of (14) with the convergent geometric series
m=0 a−mδ, we infer that if δ > 0, then there must exist a constant C1 > 0, independent
of u and Bi, and at least one index mi ∈ N such that
rad(Bi,mi)(cid:18)Z2λBi,mi
gp
u dµ(cid:19)1/p
≥ C1a−miδ ≥ C(cid:18)
ri
rad(Bi,mi)(cid:19)δ
.
Let us now fix q such that co-dimA(X \ Ω) > q > p and set δ = (q − p)/p > 0. We thus
obtain from (15) for each Bi a ball B∗
i = Bi,mi with radius r∗
i satisfying
P∞
(15)
(16)
Using estimate (16), and changing the summation to be over all Whitney balls, we calculate
rq−p
i ≤ C(r∗
i )qµ(cid:0)B∗
i(cid:1)−1Z2λB∗
i
gp
u dµ.
ZU
(17)
µ(Bi)r−q
i
i
i )(r∗
µ(Bi)r−p
µ(B∗
dΩ(x)−p dµ ≤ CXi
≤ CXi
≤ C XB∈W X{i:B=B∗
≤ C XB∈WZ2λB
gp
i
i
gp
u dµ
µ(Bi)r−q
i )−q Z2λB∗
µ(B) rad(B)−q Z2λB
u dµ X{i:B=B∗
i }
i }
gp
u dµ
µ(Bi)r−q
i
µ(B) rad(B)−q .
QUASIADDITIVITY OF VARIATIONAL CAPACITY
13
If B = B∗
i , that is, B ∈ W satisfies (16) for the ball Bi, then Bi ⊂ LB by the chain
i ≈ dΩ(x) for all x ∈ Bi, it follows from the bounded overlap of the
condition. Since r−q
Whitney balls Bi ⊂ LB and the assumption co-dimA(X \ Ω) > q, that
X{i:B=B∗
i }
(18)
µ(Bi)r−q
i
µ(B) rad(B)−q ≤ C
≤ C
≤ C
1
i }ZBi
dΩ(x)−q dµ
µ(B) rad(B)q X{i:B=B∗
µ(B) rad(B)q ZLB
µ(B) rad(B)q µ(LB) rad(LB)−q ≤ C.
dΩ(x)−q dµ
1
1
By the assumption c ≤ (6λ)−1 the overlap of the balls 2λB, where B ∈ Wc(Ω), is
uniformly bounded (Lemma 2.1), and so we conclude from (17) and (18) that
ZU
dΩ(x)−p dµ ≤ CZΩ
gp
u dµ.
The claim (13) follows by taking the infimum over all capacity test functions for U (and
their upper gradients).
(cid:3)
It has been shown in [25, Section 6] (following the considerations of [21]), that if Ω ⊂ X
admits a p-Hardy inequality, then either co-dimH(X \ Ω) < p − δ or co-dimA(X \ Ω) > p + δ
for some δ > 0 only depending on the data associated with the space X and the Hardy
inequality. Moreover, there is also a local version of such a dichotomy for the dimension
[25, Theorem 6.2]. These results, together with the above Proposition 4.1 (and see also the
following Section 5), show clearly that the condition co-dimA(X \ Ω) > p is very natural
in the context of Hardy inequalities and thus also for quasiadditivity. On the other hand,
the case co-dimH(X \ Ω) < p − δ includes open sets with uniformly p-fat complements; cf.
Corollary 3.6.
The main open question here is whether the discrete John condition is really necessary
in Proposition 4.1; we know of no counterexamples. In the Euclidean space Rn this extra
condition is certainly not needed. Indeed, as commented at the end of [21], the dimension
bound dimA(Rn \ Ω) < n − p implies by [2, Theorem 2] that
dΩ(x)−p dx ≤ CR1,p(E) ≤ C capp(E, Ω)
ZE
for all (measurable) E ⊂ Ω; here R1,p(E) is a Riesz capacity of E (cf. [2] or [5] for the
definition) and the second inequality is a well-known fact. Quasiadditivity for capp(·, Ω)
follows by Theorem 3.3.
Nevertheless, Proposition 4.1 still gives a partial answer to the question of Koskela and
Zhong [21, Remark 2.8], i.e., a q-Hardy inequality holds in their setting provided that Ω
satisfies the discrete John condition (and the Minkowski dimension in [21, Remark 2.8] is
replaced by the correct Aikawa (co)dimension).
5. Combining fat and small parts of the complement
The results studied in Section 3 gave us a criterion, uniform p-fatness of X \ Ω, under
which Ω supports quasiadditivity of p-capacity for the Whitney decompositions of Ω; this
condition requires X \Ω to be 'large'. Conversely, in Section 4 we gave a criterion, largeness
of the Aikawa co-dimension of X \ Ω (or, smallness of the Assouad dimension -- and
hence 'smallness' of X \ Ω), under which Ω supports quasiadditivity for the Whitney
decompositions of Ω. Nevertheless, requiring the whole complement to be either large
or small rules out many interesting cases. For instance, sometimes the complement of a
domain can be decomposed into two closed subsets such that one of them is 'large' and
one is 'small'; an easy example is the punctured ball B(0, 1) \ {0} ⊂ Rn. The aim of this
14
JUHA LEHRBÄCK AND NAGESWARI SHANMUGALINGAM
final section is to explain how the results of the previous Sections 3 and 4 can be combined
to address such more complicated sets.
In the Euclidean case, some results into this
direction can be found also in [23]. A full geometric characterization of domains supporting
quasiadditivity of p-capacity for Whitney decompositions still seems to be beyond our
reach. However, in the next lemma we demonstrate a technique which applies to a broad
class of sets.
Lemma 5.1. Assume that X is unbounded and that 1 < p < Qu. Suppose that Ω0 ⊂ X
is an open set such that X \ Ω0 is uniformly p-fat. Suppose also that F ⊂ Ω0 is a closed
set with co-dimA(F ) > p, and that X \ F satisfies the discrete John condition of Section 4.
Then Ω = Ω0 \ F satisfies a quasiadditivity property with respect to Whitney covers Wc(Ω)
with suitably small c > 0.
Proof. Let Wc(Ω) be a Whitney decomposition of Ω, where 0 < c < min{(CA)−1, (30λ)−1}.
Set W 1 to be the collection of all balls B(x, r) ∈ W satisfying dist(x, X \ Ω) = dist(x, F )
and, similarly, let W 2 be the collection of all balls B(x, r) ∈ W satisfying dist(x, X \ Ω) =
dist(x, X \ Ω0). It is clear that we can extend the collection W 1 to a Whitney cover W 1∗
of X \ F =: Ω1 and the collection W 2 to a Whitney cover W 2∗ of Ω0, both with the same
constant c but possibly with larger overlap constants.
As before, to prove the quasiadditivity property, it suffices to consider unions of Whitney
balls. Thus, let U = Si∈I Bi, where Bi ∈ Wc(Ω) and I ⊂ N; we may also assume that
capp(U, Ω) < ∞. Set U1 = SBi∈W 1 Bi, U2 = SBi∈W 2 Bi. Since co-dimA(X \ Ω1) =
co-dimA(F ) > p and the discrete John condition holds for Ω1, we know, by Proposition 4.1,
that
(19)
X{i∈I:Bi∈W 1}
capp(Bi, Ω1) ≤ C1 capp(U1, Ω1) ≤ C1 capp(U, Ω);
here we use the facts U1 ⊂ U and Ω ⊂ Ω1. On the other hand, an application of the results
of Section 3 yields
(20)
X{i∈I:Bi∈W 2}
capp(Bi, Ω0) ≤ C2 capp(U2, Ω0) ≤ C2 capp(U, Ω).
Here the last inequality follows since U2 ⊂ U and Ω0 \ Ω ⊂ F is of zero p-capacity. For
the same reason we have in (20) that capp(Bi, Ω0) = capp(Bi, Ω) for each Bi ∈ W 2. To
estimate the corresponding capacities on the left-hand side of (19), we use the capacity
bounds from (3) (with respect to Ω and Ω1; note that the assumptions of Lemma 2.2 are
valid in the latter case) to obtain for all Bi ∈ W 1 that
capp(Bi, Ω) ≤ C rad(Bi)−pµ(Bi) ≤ C3 capp(Bi, Ω1).
In conclusion,
Xi∈I
capp(Bi, Ω) ≤ C3(cid:18) X{i∈I:Bi∈W 1}
≤ C3(C1 + C2) capp(U, Ω),
capp(Bi, Ω1) + X{i∈I:Bi∈W 2}
capp(Bi, Ω0)(cid:19)
and the claim follows by Theorem 3.3.
(cid:3)
References
[1] D. Adams, 'Quasi-additivity and sets of finite Lp-capacity', Pacific J. Math. 79 (1978), no. 2, 283 -- 291.
[2] H. Aikawa, 'Quasiadditivity of Riesz capacity', Math. Scand. 69 (1991), no. 1, 15 -- 30.
[3] H. Aikawa, 'Quasiadditivity of capacity and minimal thinness', Ann. Acad. Sci. Fenn. Ser. A I Math.
18 (1993), no. 1, 65 -- 75.
[4] H. Aikawa and A. A. Borichev, 'Quasiadditivity and measure property of capacity and the tangen-
tial boundary behavior of harmonic functions', Trans. Amer. Math. Soc. 348 (1996), no. 3, 1013 -- 1030.
QUASIADDITIVITY OF VARIATIONAL CAPACITY
15
[5] H. Aikawa and M. Essén, 'Potential theory -- selected topics', Lecture Notes in Mathematics, 1633,
Springer-Verlag, Berlin, 1996.
[6] A. Björn, J. Björn, 'Nonlinear potential theory on metric spaces', EMS Tracts in Mathematics,
17, European Mathematical Society (EMS), Zürich, 2011. xii+403 pp.
[7] A. Björn, J. Björn, and N. Shanmugalingam, 'Sobolev extensions of Hölder continuous and
characteristic functions on metric spaces', Canadian J. Math. 59 (2007), No. 6, 1135 -- 1153.
[8] J. Björn, 'Boundary continuity for quasiminimizers on metric spaces', Illinois J. Math. 46 (2002),
no. 2, 383 -- 403
[9] J. Björn, P. MacManus and N. Shanmugalingam, 'Fat sets and pointwise boundary estimates
for p-harmonic functions in metric spaces', J. Anal. Math. 85 (2001), 339 -- 369.
[10] W.F. Donoghue Jr., 'Remarks on potential theory', Ark. Mat. 4 (1962) 461 -- 466.
[11] P. Hajłasz and P. Koskela, 'Sobolev met Poincaré', Mem. Amer. Math. Soc. 145 (2000), no. 688.
[12] J. Heinonen, 'Lectures on analysis in metric spaces', Universitext, Springer-Verlag, New York, 2001.
[13] J. Heinonen, P. Koskela, N. Shanmugalingam and J. Tyson, 'Sobolev spaces on metric measure
spaces: an approach based on upper gradients', in preparation.
[14] I. Holopainen, 'Nonlinear potential theory and quasiregular mappings on Riemannian manifolds',
Ann. Acad. Sci. Fenn. Ser. A I Math. Dissertationes 74 (1990), 45 pp.
[15] I. Holopainen and N. Shamugalingam, 'Singular functions on metric measure spaces', Collect.
Math. 53 (2002), no. 3, 313 -- 332.
[16] S. Kallunki and N. Shanmugalingam 'Modulus and continuous capacity', Ann. Acad. Sci. Fenn.
Math. 26 (2001) 455-464.
[17] J. Kinnunen and O. Martio, 'Nonlinear potential theory on metric spaces', Illinois J. Math. 46
(2002), no. 3, 857 -- 883.
[18] J. Kinnunen and N. Shanmugalingam,
Manuscripta Math. 105 (2001), no. 3, 401 -- 423.
'Regularity of quasi-minimizers on metric spaces',
[19] R. Korte, J. Lehrbäck and H. Tuominen, 'The equivalence between pointwise Hardy inequalities
and uniform fatness', Math. Ann. 351 (2011), no. 3, 711 -- 731.
[20] R. Korte and N. Shanmugalingam, 'Equivalence and self-improvement of p-fatness and Hardy's
inequality, and association with uniform perfectness', Math. Z. 264 (2010), no. 1, 99 -- 110.
[21] P. Koskela and X. Zhong, 'Hardy's inequality and the boundary size', Proc. Amer. Math. Soc.
131 (2003), no. 4, 1151 -- 1158.
[22] N. S. Landkof, 'Foundations of modern potential theory', Springer-Verlag, New York -- Heidelberg,
1972.
[23] J. Lehrbäck, 'Weighted Hardy inequalities and the size of the boundary', Manuscripta Math. 127
(2008), no. 2, 249 -- 273.
[24] J. Lehrbäck, 'Neighbourhood capacities', Ann. Acad. Sci. Fenn. Math. 37 (2012), no. 1, 35 -- 51.
[25] J. Lehrbäck and H. Tuominen, 'A note on the dimensions of Assouad and Aikawa', J. Math. Soc.
Japan, to appear.
[26] J. Luukkainen, 'Assouad dimension: antifractal metrization, porous sets, and homogeneous mea-
sures', J. Korean Math. Soc. 35 (1998), no. 1, 23 -- 76.
[27] V. G. Maz'ya, 'Sobolev Spaces', Springer-Verlag, Berlin, 1985.
[28] N. Shanmugalingam, 'Newtonian spaces: an extension of Sobolev spaces to metric measure spaces',
Rev. Mat. Iberoamericana 16 (2000), no. 2, 243 -- 279.
[29] N. Shanmugalingam, 'Harmonic functions on metric spaces', Illinois J. Math. 45 (2001), no. 3,
1021 -- 1050.
[30] M. Sion, 'On capacitability and measurability', Ann. Inst. Fourier (Grenoble) 13 (1963) 83 -- 98.
[31] J. Väisälä, 'Exhaustions of John domains', Ann. Acad. Sci. Fenn. Ser. A I Math. 19 (1994), no. 1,
47 -- 57.
[32] A. Wannebo, 'Hardy and Hardy PDO type inequalities in domains part I', arXiv:math/0401253v1
(http://arxiv.org/abs/math/0401253), (2004).
J.L.: Department of Mathematics and Statistics, P.O. Box 35 (MaD), FIN-40014 Univer-
sity of Jyväskylä, Finland
E-mail address: [email protected]
N.S.: Department of Mathematical Sciences, P.O. Box 210025, University of Cincinnati,
Cincinnati, OH 45221-0025, U.S.A.
E-mail address: [email protected]
|
0710.1753 | 5 | 0710 | 2015-06-28T18:26:26 | A generalisation of the Cauchy-Kovalevskaia theorem | [
"math.FA"
] | I prove, under mild assumptions, that solutions to linear evolution equations admit sectorial solutions. The size of the sector depends on the regularity of the initial data. If it is regular enough the solution is holomorphic and unique otherwise it is sectorial. I also prove that the result is optimal for many partial differential systems (which includes KdV and other examples). | math.FA | math |
A GENERALISATION OF THE
CAUCHY-KOVALEVSKAIA THEOREM
MAURICIO GARAY
Abstract. We prove that time evolution of a linear analytic ini-
tial value problem leads to sectorial holomorphic solutions in time.
1. Introduction
Among the class of systems of partial differential equations, evo-
lutionary ones form a minority. They are nevertheless of considerable
importance because they describe time evolution of physical data. This
can already be seen for ordinary differential equations where, among
differential equations, vector fields deserve a particular attention. The
aim of this paper is to prove that evolutionary linear partial differential
systems always admit sectorial analytic solutions, where the width of
the sector depends on the regularity of the initial condition.
Before stating our main theorem, let us recall the results obtained
by Kovalevskaıa in her thesis [24] (see also [1] for historical aspects).
We consider the vector space Cn with coordinates z1, . . . , zn and let
On be the algebra of germs of holomorphic functions at 0 ∈ Cn (series
in z1, . . . , zn which are analytic in some neighbourhood of the origin).
We put
∂I := ∂i1
z1∂i2
z2 . . . ∂in
zn, I = (i1, . . . , in)
and define the order σ(I) of the operator ∂I as the sum of the coordi-
nates of the vector I ∈ Nn:
σ(I) := i1 + · · · + in.
An evolution equation of order s is a partial differential equation of the
form
∂tx = F (x, ∂I1
z x), σ(Ij) ≤ s, x = (x1, . . . , xm), z = (z1, . . . , zn)
z x, . . . , ∂Ik
with some initial condition x(t = 0, ·) = x0, where F is a polynomial
mapping.
Date: Original version May 2012, (modified June 2015).
Key words and phrases. Cauchy-Kovalevskaıa theorems, initial value problem,
characteristic Cauchy data, Heat equation.
2000 Mathematics Subject Classification: 35A10.
1
2
MAURICIO GARAY
Kovalevskaıa proved that the formal solution to such an initial value
problem:
x(t, ·) :=Xk≥0
xktk, xk ∈ Om
n
exists and is unique. Then she proceeded to the analytic properties
of time evolution. For s = 1, she showed that the formal solution
is holomorphic, in any sufficient small neighbourhood of the origin in
Cn+1. A result now called the Cauchy-Kovalevskaıa theorem. For s = 2,
Kovalevskaıa considered the particular case of the one dimensional heat
equation and discovered that the formal solution might be divergent.
To state Kovalevskaıa's heat equation theorem, it is convenient to
n of class s Gevrey series in n variables [11, 12].
introduce the space Gs
These are formal power series:
such that
XI∈Nn
XI∈Nn
aIzI ∈ C[[z]]
aI
zI
(σ(I)!)s−1
is convergent in a sufficiently small neighbourhood of the origin. For
s = 1, Gevrey series are just analytic series, for s < 1 these are entire
functions and for s > 1 these are divergent power series.
Theorem 1.1 ([24]). The formal solution to the one dimensional heat
equation
∂tx = ∂zzx, x(t = 0, −) = x0, x0 ∈ Gs
1
1) is a Gevrey class 2 series;
2) has a unique holomorphic solution if x0 ∈ G1/2
3) is divergent if x0 /∈ G1/2
.
1
;
1
Time evolution for the heat equation theorem became a classical
subject and was treated in details in Hadamard's lectures on partial
differential equations [15].
In the eighties', Ouchi made an important step further, when he dis-
covered that the divergent series associated to time evolution of a single
linear partial differential equation are in fact asymptotic expansions of
sectorial solutions [22]. We will extend the results of Kovalevskaıa
thesis and Ouchi's theorem to arbitrary systems of linear partial dif-
ferential equations.
2. Statement of the theorem
According to the Kovalevskaıa heat equation theorem, formal solu-
tions will not be analytic in general. So we must look for sectorial
holomorphic solutions. Let us clarify this notion.
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM
3
By sector of width α ∈ [−∞, +∞], we mean a subset of the form:
Σ := {z ∈ Cn :
zi ≤ r; arg zi − θi ≤
α
2
}
for some direction θ = (θ1, . . . , θn) ∈ (S1)n and some radius r > 0. If
the width is negative or infinite the condition on the angle is empty
and the sector Σ is just a neighbourhood of the origin.
We denote by Dn(r) ⊂ Cn the polydisk:
Dn(r) := {z ∈ Cn : zi ≤ r, i = 1, . . . , n}
and by D∗
n(r) ⊂ C the "pointed polydisk":
Dn(r)∗ := {z ∈ (C∗)n : zi ≤ r, i = 1, . . . , n}
We denote by Gα
n(θ) the algebra of functions which are holomorphic
in an open subset containing some pointed polydisk of radius r and
which admit an asymptotic expansion of Gevrey class α inside any
sector Σ of width less than π/(α − 1), direction θ and radius r.
For instance the function e−1/t belongs to G2
1(0) which means that
for any sector Σ contained in the half plane
{t ∈ C, Re t > 0}
the asympotic expansion of e−1/t is of Gevrey class 2. This is indeed the
case since it is equal to zero. Note that there is a map which associates
to such a function its asymptotic expansion
and that it is not injective unless α ≤ 1.
Gα
n(θ) −→ Gα
n
Our theorem on time evolution requires some non-degeneracy con-
dition. We say that a linear partial differential operator of order s is
non-degenerate, if after a change of variables, it can be written in the
form:
∂s
zn − Xσ(I)<s
AI∂I
z , A ∈ M(m, Gα
n(θ)).
Here M(m, R) stands for m × m matrices with entries in the ring R for
some positive integer m.
When α = 1, for such a partial differential operator, the Cauchy-
Kovalevskaıa theorem shows that there is a unique holomorphic solu-
tion with initial data
x(−, zn = 0) = x0, . . . , ∂s
znx(−, zn = 0) = xs.
For instance, an initial value problem of the form
∂2
z x = ∂tx, x(z = 0, −) = x0, ∂zx(z = 0, −) = x1
4
MAURICIO GARAY
can be reduced to a system of order 1 :
(cid:26)∂zx = y
∂zy = ∂tx
with x(t = 0, −) = x0, y(t = 0, −) = x1. Therefore by Cauchy-
Kovalevskaıa theorem, it admits a unique holomorphic solution. This
is of minor interest for us, since we are interested in time evolution and
not in spatial evolution. The main result of this paper is the:
Theorem 2.1. LetPσ(I)≤s AI∂I
with AI ∈ M(m, Gα
z x be a non-degenerate linear operator
n(θ)) for some direction θ. The initial value problem
AI∂I
z x, x0 := x(t = 0, −) ∈ (Gα
n(θ))m
∂tx = Xσ(I)≤s
admits solutions x ∈ Gαs
1 (θ′) ⊗Gα
n(θ) for any θ′ ∈ S1.
In the statement of the theorem we used a topological tensor product
which can simply be understood as the space of functions in some
pointed polydisk Dn+1(r)∗ which have asymptotic expansions of Gevrey
class αs (resp. α) in the t variable (resp. z variable) inside the sector
Σ′ (resp. Σ). For more details on topological tensor products see [14].
Example 2.2. Consider the Kovalevskaıa example:
∂tx = ∂2
z x, x(t = 0, −) =
1
1 − z
.
Here α = 1, s = 2 so that Σ is of the form
Σ = Σ1 × D2(r)
where Σ1 is a sector of width π, that is, a half plane. There exists a
function
x(t, z) : {(t, z) ∈ D2(r) : Re t > 0} −→ C
which satisfies our initial value problem. We shall give examples of
such functions in the next section.
In real analysis, the solution of
the heat equation in the circle can be solved by Fourier series and the
flow is only defined for positive time. The situation is here completely
different, since it admits solutions both for positive and negative time.
Corollary 2.3. There is a unique holomorphic solution to an initial
value problem of the form
∂tx = Xσ(I)≤s
AI∂I
z x, x0 := x(t = 0, −) ∈(cid:0)G1/s
n (cid:1)m
, AI ∈ M(m, G1/s
n ).
For the heat equation (m = n = 1, s = 2), we recover Kovalevskaıa's
theorem (except for the statement on the divergence of the solution,
which is discussed in the appendix). The proof of Theorem 2.1 is based
on a generalisation of Cauchy's m´ethode des majorantes to general flows
in infinite dimensional spaces.
tn ⊗ an,
Xn≥0
that is the projective limit of the vector space k[[t]]/(tn) ⊗k E.
The map L induces a map id ⊗ L on E[[t]]:
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM
5
3. Formal evolution
We start with the definition of formal evolution. In the linear case,
this is quite obvious. Let E be a vector space over a field k and
L : E −→ E
a linear map. We denote by E[[t]] := k[[t]] ⊗E the vector space of
formal power series with coefficients in E:
(id ⊗ L) (Xn≥0
tn ⊗ an) =Xn≥0
tn ⊗ L(an).
We abuse notation and write L for id ⊗L. Similarly, we write ∂t instead
of ∂t ⊗ Id . We also identify E with the subspace 1 ⊗ E ⊂ E[[t]]. If the
field k is of characteristic zero then the exponential is well-defined and
et⊗Lx0 is the unique solution to the initial value problem
∂tx = Lx, x(t = 0, −) = x0.
The operator etL is called the time evolution of the operator L.
In order to extend this definition of evolution to non-linear operators,
we first construct the Lie derivative. So let E, F be locally convex
vector spaces and let U ⊂ E denote an open subset.
A mapping
f : E ⊃ U −→ F,
is called Gateaux differentiable at a point x ∈ U, if for any ξ ∈ E, the
following limits exists
Let
Df (x)ξ := lim
t7→0
f (x + tξ) − f (x)
t
.
f : E −→ F, X : E −→ E
be Gateaux differentiable mappings between locally convex spaces. The
Lie derivative of f along X is defined by
f 7→ Df (x)X(x).
We would like to find some vector space which is stable under the Lie
derivative. In finite dimensional differential geometry, one may choose
the space of C ∞ function but these are difficult to handle in the infinite
dimensional context, for general locally convex spaces. Therefore, we
now assume that k = C and consider the space of holomorphic maps
from E to F , denoted H(E, F ). These are defined as follows:
6
MAURICIO GARAY
Definition 3.1. A mapping f : E ⊃ U −→ F is called holomorphic
if it satisfies the following two conditions:
i) it is continuous,
ii) for any continuous linear mappings j : C −→ E, π : F −→ C the
map π ◦ f ◦ j is holomorphic.
For instance, a linear mapping is holomorphic if and only if it is
continuous. Like in the finite dimensional case, holomorphic functions
in infinitely many variables are convergent analytic power series (see
[7] for more details).
The elements of H(U, E) are called holomorphic vector fields in U.
By contracting the differential with a vector field X ∈ H(U, E), we
define the Lie derivative
LX : H(U, E) −→ H(U, F ), f 7→ [x 7→ Df (x)X(x)]
for general locally convex spaces. As the Lie derivative is linear map,
we constructed in this way the derivation associated to a vector field
in the infinite dimensional context.
The map
LX : eE −→ eF , eE = H(U, E), eF = H(U, F )
being linear, we can consider the time evolution of any function. Now
in the particular case E = F , we may consider the time evolution of
the identity mapping. This defines in turn time evolution for general
vector fields:
Definition 3.2. The formal flow of a holomorphic vector field X ∈
H(U, E) at x0 is the evaluation of the map etLX Id at x0, where LX is
the Lie derivative along X.
Note that by construction the flow is a solution of the differential
equation
dx
dt
= (LX Id )(x) = X(x).
Example 3.3. Consider the inviscid Burgers equation:
∂tx = x∂zx, x(t = 0, ·) = x0.
Denote by C{z} the vector space of convergent power series in one
variable z, it has a natural topology (see e.g.
[13]). The vector field
associated to our initial value problem is
C{z} −→ C{z}, x 7→ x∂zx.
The Lie derivative
L : H(C{z}) −→ H(C{z}), [x 7→ f (x)] 7→ [x 7→ Df (x)x∂zx]
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM
7
is linear and therefore admits a unique formal time evolution. Compu-
tation of time evolution up to order 2 gives:
(et⊗LId )x = x + tx∂zx +
t2
2 (cid:0)2x (∂zx)2 + x2∂2
z x(cid:1) + o(t2).
In simple words, the possibility to define differential calculus in the
space of holomorphic functions H(C{z}) allows us to define formal
flows like for the finite dimensional spaces. This explains the unicity
of the formal solution to initial value problems.
4. Generalisation of the heat equation theorem
We proceed to the Gevrey properties of formal solutions and first
recall Borel resummation procedure [4] (see also [2, 18]). Gevrey diver-
gent series can be considered as asymptotic expansions of exponential
integrals. For instance, the relation
shows that for any polynomial, we get:
More generally, if we consider a function
Z +∞
0
e− ξ
k!aktk.
nXk=0
t ξkdξ = k!tk.
1
0
e− ξ
tZ +∞
t nXk=0
ξ 7→ y(ξ) =Xk≥0
akξk! dξ =
akξk
holomorphic in a neighbourhood of the real line which has at most
exponential growth, then
is a holomorphic function whose asymptotic expansion at the origin is:
x(t) =
e− ξ
t y(ξ)dξ
0
1
tZ +∞
Xk≥0
k!aktk.
The real line can of course be replaced by any closed curve which starts
from 0 and does not come back to it, for instance a real segment [0, r].
But along the real positive half-line, if the integrand is well-defined
and decreases exponentially then the associated integral transformation
maps the ring Dt of partial differential operator on t to that on ξ in
8
MAURICIO GARAY
the following way:
Dt −→ Dξ
7→ ∂ξ
1
t
∂t 7→ ∂ξ + ξ∂2
ξ
This procedure can be generalised to Gevrey classes G1+s, s ≥ 1 using
the exponential integral:
x(t) :=
1
tsZ +∞
0
e− ξs
ts y(ξ)d(ξs)
Let us now apply this resummation procedure to the heat equation
with Kovalevskaıa's initial value:
∂tx = ∂zzx,
x0 : (C, 0) −→ (C, 0), z 7→
1
1 − z
.
The formal power series expansion of this Cauchy problem is of Gevrey
class 2
x(t, z) =
The substitution
leads to the analytic series
1
1 − zXj,k
(2k)!
k!
tk
(1 − z)2k .
k!tk 7→ ξk
y(ξ, z) =
1
1 − zXj,k
(2k)!
(k!)2
ξk
(1 − z)2k =
1
.
p(1 − z)2 − 4ξ
The initial condition x0 has a pole at z = 1 and the integrand has itself
a singularity at
ξ =
(1 − z)2
4
.
This is a particular case of the Lutz-Myiake-Schafke theorem which
states that the solution to the initial value problem "reproduces" the
singularities of the initial data [16]. Now let Γ be an arbitrary path
starting at 0, asymptotic to a line in the half plane
{ξ ∈ C : Re ξ ≥ 0}
and which avoids the singularity ξ = 1/4. For z sufficiently small, the
power series x is the asymptotic expansion at t = 0 of the function
xΓ(z, t) :=
1
tZΓ
e− ξ
t
1
p(1 − z)2 − 4ξ
dξ.
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM
9
As explained above, the ring of partial differential operators Dt,z is
mapped to Dξ,z. Thus our function
is a solution of the partial differential equation
y(ξ, z) =
1
p(1 − z)2 − 4ξ
(cid:0)∂ξ + ξ∂2
ξ(cid:1) y = ∂2
tZΓ(cid:0)∂ξ + ξ∂2
ξ − ∂2
z y.
1
(cid:0)∂t − ∂2
z(cid:1) xΓ =
z(cid:1) ydξ = 0.
But
This means that the functions xΓ are solutions to our initial value prob-
lem, for any such choice of the path Γ. Note that the singularity of
the integrand implies that the solutions obtained by choosing different
paths lead to different solutions and forces the divergence of the as-
ymptotic series. This can already be observed for the case of the Euler
equation:
In this case the Borel transform is
(−1)nn!tn+1.
t2 dx
dt
+ x = t, x(t) =Xn≥0
x(ξ) =Xn≥0
(−1)nξn+1 =
1
1 + ξ
.
The singularity at ξ = −1 is responsible for the non unicity of the
solution to our Cauchy problem (see for instance [6, 9]).
In any case, the first step to achieve Borel resummation of formal
solutions is to ensure that they lie in some Gevrey class. This is estab-
lished by the following
Theorem 4.1. The formal solution to an initial value problem
∂tx = Xσ(I)≤s
AI∂I
z x, x0 := x(t = 0, −) ∈ (Gα
n)m
of order s is of Gevrey class αs in the time variable, that is, time
evolution defines a map:
n)m ⊗Gαs, x0 7→ (etLId )x0
where L is the Lie derivative associated to the operator.
n)m −→ (Gα
etL : (Gα
For a single partial differential equation (m = 1), the theorem is
again due to Ouchi. Using techniques due to Boutet de Monvel and
Kree, Yonemura simplified the proof [5, 22, 25]. Gevrey properties for
some particular systems of partial differential equations (other than
the heat equation) is proved in [10]. We postpone the proof of our
theorem to the next section and first discuss the relation between formal
solutions and asymptotic ones.
10
MAURICIO GARAY
The Borel transform
Bs : Gs
1 −→ G1
1 = O1, Xk≥0
aktk 7→Xk≥0
ak
(k!)s ξk
associates a holomorphic function y(ξ) to a Gevrey series. In general
this function is defined only in a small neighbourhood of the origin,
therefore we cannot apply Borel resummation procedure to our formal
solution as we did for the Kovalevskaıa example. So we replace, our
integral formulas by
xr(t) :=
e− ξs
ts y(ξ)d(ξs)
1
tsZ r
0
where r is now some finite number. The resulting function is still as-
ymptotic to the formal solution x. But under this new integral transfor-
mation, the differential operator ∂t does not correspond to a differential
operator in the ξ variable anylonger. Therefore our function xr is, in
general, not a solution to the initial value problem but rather a solu-
tion up to a flat function. For instance, for r < 1/4, the asymptotic
expansion of the function
xr(z, t) :=
1
tZ r
0
e− ξ
t
1
dξ
p(1 − z)2 − 4ξ
is still a formal power series which satisfies the heat equation but the
function itself does not. In other words the function
(∂t − ∂zz)xr
is flat at t = 0 inside the half plane Re t > 0.
This might sound disappointing, and one might conclude that the
above theorem gives simply no information about local analytic solu-
tions, fortunately:
Lemma 4.2. The existence of formal solutions implies the existence
of asymptotic solutions, that is:
Theorem 4.1 =⇒ Theorem 2.1
Proof. Let
L : Gn(θ)m −→ Gn(θ)m
be a linear partial differential operator of order s ≥ 2. We consider an
initial value problem
∂tx = Lx, x(t = 0, −) = x0 ∈ Gn(θ)m.
Theorem 4.1 asserts that the formal solution is of Gevrey class s in
the time variable. It is therefore asymptotic expansion in any sector of
width ¡ π/(s − 1) of a holomorphic map u defined in the pointed disk.
The function u is a solution to our initial value problem up to a flat
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM 11
function. Substituting x by y + u in our system of partial differential
equation, we get a new system
(∗) ∂ty = Ly + (Lu − ∂tu).
As the initial system is non-degenerate, in appropriate coordinates L =
∂s
zn − L′ where L′ is of the form
L′ = Xσ(I)<s
AI∂I +
Bi∂s
zi.
n−1Xi=1
We re-write (∗) as
(∗∗) ∂s
zny = ∂ty + L′y + g, g := ∂tu − Lu.
For any initial data:
y0 = y(−, zn = 0), y1 = ∂zny(−, zn = 0), . . . , ys−1 = ∂s−1
zn y(−, zn = 0),
equation (∗∗) admits a unique holomorphic solution in the pointed
disk. This can be seen, for instance, by applying the abstract Cauchy-
Kovalevskaıa theorem to the Banach scale (see for instance [3, 19, 20,
21, 23]):
with
Er := C 0(Rr) ∩ O(Rr)
Rr := {z ∈ Cn : r ≤ zi ≤ 2r}.
We take the initial data y0 = · · · = ys = 0. The associated solution
admits an asymptotic expansion equal zero when t goes to zero inside
the given sector. Indeed, by adding ∂zny, . . . , ∂s−1
zn y as new variables:
y0 = y, y1 = ∂zny, . . . , ys−1 = ∂s−1
zn y.
We reduce the system of partial differential equation to a first order
system of the form
∂τ Y = T Y + h, Y = (y0, . . . , ys−1)
where τ = zn, T is linear and h is flat at t = 0 in Σ. The solutions are
of the form
Y = eτ T Y0 + α, α(τ = 0, −) = 0
where α satisfies the equation
∂τ α = T α + h.
As h is flat, α = 0 is a formal solution to this equation, but the for-
mal solution is unique thus α is flat at t = 0 in Σ. As the initial
condition is Y0 = 0, the solution to our system of partial differential
equations is constant Y = α and it is therefore flat. This shows that
the holomorphic mapping
provides a solution to our original initial value problem in Gn(θ)m. (cid:3)
x := u + y
12
MAURICIO GARAY
5. Generalisation of Cauchy's majorant method
We consider vector fields in the infinite dimensional Gevrey spaces
n)m and extend the classical Cauchy majorant method by compar-
(Gs
ing series in these different functional spaces. Via formal Borel trans-
form these topological vector spaces are isomorphic to spaces of con-
vergent power series. They are therefore endowed with a standard
topology (see [13] for the definition of the topology).
Recall that a formal power series
x := XI∈Nn
aIzI ∈ C[[z]], z = (z1, . . . , zp)
is majorated by another formal power series
y := XI∈Nn
bIzI ∈ R+[[z]]
if, for all I ∈ Nn, we have the estimates:
aI ≤ bI.
In such cases, we use the notation
x ≪ y.
In particular, x ≫ 0 means that x is a formal power series with real
non-negative coefficients.
Definition 5.1. Let X, Y be two vector fields in (Gs
X in (Gs
n)m majorates another one Y if
n)m. A vector field
x ≫ y =⇒ X(x) ≫ Y (y).
In particular X ≫ 0 means that:
x ≫ 0 =⇒ X(x) ≫ 0.
Example 5.2. Let X be a vector field associated to a linear partial
differential operator
X : (On)m −→ (On)m, x 7→ Xσ(I)≤s
AI∂I x.
Then X ≫ 0 provided that the entries of the matrices AI are analytic
series with real positive coefficients, i.e., AI ≫ 0.
The following proposition is a direct consequence of the exponential
formula for time evolution:
Proposition 5.3. Let X, Y be two vector fields defined in an open
subset of (Gs
n)m.
i) If X ≪ Y then the flow of X at x0 ≫ 0 is majorated by that of Y
at the same point,
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM 13
ii) If X ≫ 0 and y0 ≫ x0 then the flow of X at x0 is majorated by
that of X at y0.
We proceed to the proof of Theorem 4.1 and start with a
Proposition 5.4. The following assertions are equivalent
1) the flow of any linear initial value problem of order s in (Gα
Gevrey class αs in the time variable ;
2) the flow of any linear initial value problem of order s in (Gα
n)m is of
n)m is of
Gevrey class αs in the time variable ;
3) the flow of any linear initial value problem of order s at x0 =
Pn≥0 (n!)α−1 zn ∈ Gα
4) the flow of x 7→ x0∂s
class αs.
1 is of Gevrey class s ;
z x at the point x0 =Pn≥0 (n!)α−1 zn is of Gevrey
Proof. Let us first make a remark. Consider a vector field defined by a
linear differential operator:
n)m −→ (Gα
z x; I = (i1, . . . , in)
X : (Gα
AI∂I
n)m , x 7→ Xσ(I)≤s
at a point x0.
The map
Gα
n −→ Gα
n, α =XI
aIzI 7→ abs α :=XI
aIzI
induces a map on matrices with coefficients in Gα
the same way.
n that we denote in
Replace, in the initial value problem the AI's by abs AI's and x0 by
abs x0. By Proposition 5.3, if the solution of this new initial value prob-
lem is of Gevrey class k then X, x0 has the same property. Therefore
it is sufficient to consider the case X ≫ 0, x0 ∈ {x ≫ 0}.
2) =⇒ 1).
Let us consider the linear mapping
ψ : (Gα
n)m −→ Gα
n, (x1, . . . , xm) 7→
xk.
mXk=1
AI∂I
z x) =
we have
Write AI = (AI1, . . . , AIm) ≫ 0 and put fI =Pk AIk. For any x ≫ 0,
xk) = Xσ(I)≤s
ψ( Xσ(I)≤s
der ψ of the flow of X at x0 is majorated by the flow of PI fI∂I
The exponential formula for time evolution implies that the image un-
z at
z xk ≪ Xσ(I)≤s
mXk=1 Xσ(I)≤s
mXk=1
mXk=1
AIk)∂j
z (
AIk∂I
ψ(x0).
(
fI∂I
z ψ(x).
14
MAURICIO GARAY
3) =⇒ 2).
Consider the open subset U = {x ≫ 0} ⊂ Gα
n. The mapping
R : C −→ Cn, z 7→ (z, . . . , z)
induces a map
R∗ : Gα
n[[t]] ⊃ U[[t]] −→ Gα
1 [[t]],
and an element is of Gevrey class β in the t variable provided that it
is the case of its image under R∗.
The equalities
R∗∂zizk
j = kzk−1δij,
d
dz
R∗zk
j = kzk−1
give the estimate
R∗∂zi ≪
d
dz
R∗.
Consider a vector field in Gα
n for the form:
X : x 7→XI
fI∂I
z x, fI ≫ 0
As R∗∂zi ≪ ∂zR∗, the flow of the vector field
Gα
1 −→ Gα
1 , x 7→ Xσ(I)≤s
R∗fj
dσ(I)x
dzσ(I)
at R∗x0, x0 ≫ 0 majorates the image under R∗ of the flow of X.
Consider the Gevrey series
is analytic. Thus, by Hadamard's lemma, there exists A, r > 0 such
that
an
(n!)α−1 ≤ Arn.
This means that the series x0 is majorated by Af (rz). If X ≫ 0, the
formal flow passing through x0 is majorated by the formal flow passing
through Af (rz). Up to multiplication of x and z by constants, we may
assume that A = r = 1.
4) =⇒ 3).
and take
The series
(n!)α−1 zn ∈ Gα
1 .
f (z) :=Xn≥0
x0 =Xn≥0
Xn≥0
anzn ∈ Gα
1 .
an
(n!)α−1 zn
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM 15
Consider the flow of a vector field of the form:
at f defined above.
As
X : x 7→
aj
djx
dzj
sXj=0
ds
dzs ≫
dj
dzj
for any j < s, we get that the flow of X at x0 is majorated by that of
the vector field
x 7→ b(z)
dsx
dzs , b(z) :=
aj(z).
sXj=0
As before there exists constants A, r > 0 such that
b(z) ≪ Af (rz)
and without loss of generality we may assume, as above, that A = r =
1. This concludes the proof of the proposition.
(cid:3)
To conclude the proof of Theorem 4.1, it remains to prove that the
flow of the vector field
X = f (z)
with initial condition
is of Gevrey class αs.
ds
dzs , f (z) :=Xn≥0
(n!)α−1 zn
x0(z) = f (z)
Lemma 5.5. There exists a constant Cα > 0 such that for any real
positive increasing sequence (an) we have
(n!)α−1 anzn
(n!)α−1 anzn.
f g ≪ CαXn≥0
g :=Xn≥0
f g =Xn≥0
cnzn
with
Proof. Write
with
cn := Xi+j=n
(i!)α−1 (j!)α−1 aj ≤ Xi+j=n
(i!)α−1 (j!)α−1! an.
16
MAURICIO GARAY
One easily sees that
therefore
Xi+j=n−1
Xi+j=n
i!j! ≤ n!
i!j! ≤ 3n!
For α ≥ 1, as (an) is a positive increasing sequence, we get that:
If α ≤ 1 we have
Xn≥0 Xi+j=n
Consequently the sequence
cn ≤ 3α−1 (n!)α−1 an.
(i!)α−1 (j!)α−1! = (Xi≥0
Xi+j=n
(i!)α−1 (j!)α−1!
n∈N
(i!)α−1)2 < +∞.
tends to zero at infinity. This implies that cn = o(an) and concludes
the proof of the lemma.
(cid:3)
The above lemma gives a constant Cα such that:
We have
Therefore
(f ∂s
z)kf ≪ C k
(n!)α−1 zn.
(n!)α−1 zn.
∂s
n! (cid:19)α
zf =Xn≥0(cid:18)(n + s)!
αXn≥0(cid:18) (n + ks)!
α(cid:18)(n + ks)!
C k
n!
n!
(cid:19)α
(cid:19)α (n!)α−1
k!
x(t) ≪ Xn,k≥0
zntk.
Let us write an ≡ bn if the series an/bn is bounded by a geometric
series. By Stirling's formula, we have
(i + j)!
i!j!
≡
(i + j)i+j
iijj
= ei log(1+j/i)+j log(1+i/j) ≤ ei+j
thus
We get that:
(i + j)! ≡ i!j!
(cid:18) (n + ks)!
n!
(cid:19)α (n!)α−1
k!
≡ (k!)αs−1 (n!)α−1 .
Thus the flow is of Gevrey class αs in the time variable. This concludes
the proof of Theorem 4.1.
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM 17
Appendix A. On the divergence of formal solutions
In [17], Lysik proved a result similar to Kovalevskaıa divergence re-
sult for the Korteweg -- de Vries equation, namely that the solution to
the initial value problem:
∂tx = ∂ 3
z x + x∂zx, x(t = 0, ·) =
1
1 − z
is not holomorphic (see also [8]). More generally, one may wonder if
our Gevrey estimate for time evolution is optimal. This is indeed the
case under very general asumptions:
Theorem A.1. Consider an evolutionary initial value problem of or-
der s
∂tx = g(x, z)∂s
with x(t = 0, ·) = x0. Assume that g, G ≫ 0 and x0 ≫ 0.
convergence radius of the formal Borel transform
z1x + G(x, ∂I1x, ∂I2x, . . . , ∂Ikx,d∂s
z1x), σ(Iα) ≤ s,
If the
C −→ Cm, z1 7→ Bαx(z1, 0, . . . , 0)
is finite then the formal solution to this initial value problem is not of
Gevrey class (αs − ε), for any ε > 0.
Proof of Theorem A.1. The vector field associated to our initial
value problem majorates the vector field
Moreover, the flow of X at x0 obviously majorates that of
X : x 7→ g(x, z)∂s
z1
Y : x 7→ g(x0, z)∂j
z x
at the same point. Finally, let
αzI, α 6= 0, I ∈ Nn
be a monomial appearing with a non-zero coefficient in the Taylor
expansion of g. We have
Y ≫ αzI∂j
z1.
It remains to prove that the flow of
L = zI∂s
z1
at x0 is not of Gevrey class (αs − ε), for any ε > 0.
Given formal power series f, g, we write
f ≻ g
if there are infinitely many coefficients of f which are greater than that
of g. If f ≻ g and g is not of Gevrey class s then f cannot be of Gevrey
class s.
18
MAURICIO GARAY
Write
Define
x0 = (x0,1, . . . , x0,m).
f (z) :=Xn≥0
(n!)α−1 zn
1 .
The assumption on Bαs implies that for at least one of the components
of x0, say x0,j,there exists A, r > 0 such that :
x0,j ≻ Af (rz1).
Up to a multiplication of z1 and x0 by constants, we may assume that
A = r = 1.
Now, the majorant
Lk ≫ zkI∂ks
z1 , j ∈ N,
n!
Lkf ≫ zkIXn≥0(cid:18)(n + ks)!
(cid:19)α
etLk x0 ≻ Xk≥0,n≥0
Consequently
(n!)α−1 zn
1 ≫ zkIXn≥0
(k!)sα (n!)α−1 zn
1 .
zkI(k!)sα−1 (n!)α−1 zn
1 tk.
The right hand-side is not of Gevrey class sα − ε for any ε > 0. This
proves the theorem.
Acknowledgements. I thank Boris Dubrovin for discussions from which
this paper originated. Thanks also to Duco van Straten for encourage-
ments and suggestions.
References
1. M. Audin, Souvenirs sur Sofia Kovalevskaya, Calvage & Mounet, 2008.
2. W. Balser, From divergent power series to analytic functions, Lecture Notes in
Mathematics, vol. 1582, Springer Verlag, 1994.
3. M. S. Baouendi and C. Goulaouic, Remarks on the abstract form of nonlinear
Cauchy-Kovalevsky theorems, Comm. Partial Differential Equations 2 (1977),
no. 11, 1151 -- 1162.
4. ´E. Borel, Le¸cons sur les s´eries divergentes, vol. 2, Gauthier-Villars, 1901.
5. L. Boutet de Monvel and P. Kr´ee, Pseudo differential operators and Gevrey
classes, Annales de l'Institut Fourier 17 (1967), no. 1, 295 -- 323.
6. B. Candelpergher, J.C. Nosmas, and F. Pham, Approche de la r´esurgence, Her-
mann, 1993, 289 pp.
7. S. Dineen, Complex analysis on locally convex spaces, vol. 57, North Holland
Mathematical Studies, 1981, 492 pp.
8. A.V. Domrin and A.V. Domrina, On the divergence of the Kontsevich-Witten
series, Russ. Math. Surv. 109 (2008), 773 -- 775.
9. J. ´Ecalle, Les fonctions r´esurgentes, vol. 1, alg`ebres de fonctions r´esurgentes,
Pub. Math. Orsay (1981).
A GENERALISATION OF THE CAUCHY-KOVALEVSKAIA THEOREM 19
10. M.C. Fern´andez-Fern´andez and F.J. Castro-Jim´enez, Gevrey solutions of the ir-
regular hypergeometric system associated with an affine monomial curve, Trans.
Amer. Math. Soc. 363 (2011), 923 -- 948.
11. M. Gevrey, Comptes rendus `a l'acad´emie des sciences (1913), Note du 8
d´ecembre.
12.
, Sur la nature analytique des solutions des ´equations aux d´eriv´ees par-
tielles, Annales scientifiques de l'´Ecole Normale Sup´erieure 35 (1918), no. 3,
129 -- 190.
13. A. Grothendieck, Espaces vectoriels topologiques, Instituto de Matem`atica Pura
e Aplicada, Universidade de Sao Paulo, 1954, 240 pp., English Translation:
Topological vector spaces, Gordon and Breach, 1973.
14.
, Produits tensoriels topologiques et espaces nucl´eaires, Mem. of the Am.
Math. Soc. 16 (1955).
15. J. Hadamard, Le probl`eme de Cauchy et les ´equations aux d´eriv´ees partielles
hyperboliques, Paris Hermann et Cie, 1932, 542 pp.
16. D.A. Lutz, M. Miyake, and R. Schafke, On the Borel summability of divergent
solutions of the heat equation, Nagoya Math. Journal 154 (1999), 1 -- 29.
17. G. Lysik, Non-analyticity in time of solutions to the KdV equation, Z. Anal.
Anwendungen 23 (2004), no. 1, 67 -- 93.
18. B. Malgrange, Sommation des s´eries divergentes, Expositiones Mathematicae
13 (1995), no. 2/3, 163 -- 222.
19. M. Nagumo, Uber das Anfangswertproblem partieller Differentialgleichungen,
Jap. J. Math. 18 (1942), 41 -- 47.
20. L. Nirenberg, An abstract form of the nonlinear Cauchy-Kowalewski theorem,
J. Differential Geometry 6 (1972), 561 -- 576.
21. T. Nishida, A note on a theorem of Nirenberg, J. Differential Geom. 12 (1977),
no. 4, 629 -- 633.
22. S. Ouchi, Characteristic Cauchy problems and solutions of formal power series,
Ann. Inst. Fourier 33 (1983), 131 -- 176.
23. I.V. Ovsyannikov, A singular operator in a scale of Banach spaces, Soviet Math.
Dokl. 6 (1965), 1025 -- 1028.
24. S. von Kowalevsky, Zur Theorie der partiellen Differentialgleichungen, Journal
fur reine und angewandte Mathematik 80 (1875), 1 -- 32.
25. A. Yonemura, Newton polygons and formal Gevrey classes, Publ. RIMS 26
(1990), 197 -- 204.
Mauricio Garay, Lyc´ee Franco-Allemand, Rue Collin Mamet, 78530
Buc.
E-mail address: [email protected]
|
1701.08249 | 1 | 1701 | 2017-01-28T04:55:09 | A sharp Adams inequality in dimension four and its extremal functions | [
"math.FA",
"math.AP"
] | Let $\Omega$ be a smooth oriented bounded domain in $\mathbb R^4$, $H_0^2(\Omega)$ be the Sobolev space, and $\lambda_1(\Omega)= \inf \{\|\Delta u\|_2^2 : u\in H_0^2(\Omega), \|u\|_2 =1\}$ be the first eigenvalue of the bi-Laplacian operator $\Delta^2$ on $\Omega$. For $\alpha \in [0,\lambda_1(\Omega))$, we define $\|u\|_{2,\alpha}^2 = \|\Delta u\|_2^2 - \alpha \|u\|_2^2$, for $u \in H_0^2(\Omega)$. In this paper, we will prove the following inequality \[ \sup_{u\in H_0^2(\Omega),\, \|u\|_{2,\alpha} \leq 1} \int_{\Omega} e^{32 \pi^2 u(x)^2} dx < \infty. \] This strengthens a recent result of Lu and Yang \cite{LuYang}. We also show that there exists a function $u^*\in H_0^2(\Omega)\cap C^4(\overline{\Omega})$ such that $\|u^*\|_{2,\alpha} =1$ and the supremum above is attained by $u^*$. Our proofs are based on the blow-up analysis method. | math.FA | math |
A sharp Adams inequality in dimension four and its
extremal functions
Van Hoang Nguyen∗
September 19, 2018
Abstract
Let Ω be a smooth oriented bounded domain in R4, H 2
and λ1(Ω) = inf{k∆uk2
Laplacian operator ∆2 on Ω. For α ∈ [0, λ1(Ω)), we define kuk2
for u ∈ H 2
0 (Ω). In this paper, we will prove the following inequality
2 : u ∈ H 2
0 (Ω) be the Sobolev space,
0 (Ω),kuk2 = 1} be the first eigenvalue of the bi-
2 − αkuk2
2,
2,α = k∆uk2
sup
0 (Ω), kuk2,α≤1ZΩ
u∈H 2
e32π2u(x)2
dx < ∞.
This strengthens a recent result of Lu and Yang [30]. We also show that there exists
0 (Ω) ∩ C 4(Ω) such that ku∗k2,α = 1 and the supremum above is
a function u∗ ∈ H 2
attained by u∗. Our proofs are based on the blow-up analysis method.
1
Introduction
0
0
np
(Ω) ֒→ L
n−kp (Ω) holds if p < n/k, where W k,p
Let Ω be a smooth bounded domain in Rn. The Sobolev inequality says that the embed-
ding W k,p
(Ω) denotes the Sobolev space of
functions vanishing on boundary ∂Ω together their derivatives of order less than k − 1.
Such inequality plays an important role in many branch of mathematics such as analysis,
geometric, partial differential equations, calculus of variations, etc. However, when p = n/k
the embedding W k,n/k
(Ω) ֒→ L∞(Ω) does not holds. In this case, the Moser -- Trudinger and
0
∗Institut de Math´ematiques de Toulouse, Universit´e Paul Sabatier, 118 Route de Narbonne, 31062
Toulouse c´edex 09, France.
Email: [email protected]
2010 Mathematics Subject Classification: 46E35.
Key words and phrases: Adams inequality, blow-up analysis, sharp constant, extremal functions,
regularity theory.
1
Adams inequalities are perfect replacement. The Moser -- Trudinger inequality was estab-
lished independently by Yudovic [50], Pohozaev [35] and Trudinger [41]. This inequality
was sharpened by Moser [32] by finding its sharp constant. This sharp form asserts that
the existence of a constant C0 > 0 such that
1
n
n−1
(1.1)
exp(βf (x)
n−1 )dx ≤ C0,
ΩZΩ
for any β ≤ β0 = nω1/(n−1)
where ωn−1 denotes the surface area of the unit sphere of Rn, for
any bounded domain Ω and for any function f ∈ W 1,n
(Ω) with k∇fkn ≤ 1. If β > β0 then
the above inequality does not hold with uniform C0 independent of f . Moser -- Trudinger
is a crucial tool in studying the partial differential equation inequality with exponential
nonlinearity. Because of its importance, there are many generalization of Moser -- Trudinger
inequality, such as Moser -- Trudinger inequality on Heisenberg group, on complex sphere or
on compact Riemannian manifold [5, 6, 22]. It was also extended to entire Euclidean space
by Ruf [37] for dimension two and by Li and Ruf [25] for any dimension or entire Heisenberg
group by Lam and Lu [18], or on hyperbolic space by Wang and Ye [42]. In [39], Tian and
Zhu proved a Moser -- Trudinger type inequality for alomost plurisubharmonic functions on
any Kahler-Einstein manifolds with positive curvature.
0
The existence of the extremal function for Moser -- Trudinger inequality was first proved
by Carleson and Chang [4] for the unit ball in Rn. In [13], Flucher proved the existence
of extremal function for Moser -- Trudinger inequality for any smooth domain in R2. This
result was then extended to any dimension by Lin [28]. The existence of extremal function
for Moser -- Trudinger inequality on compact Riemannian manifold was studied by Li [23].
We refer the reader to [7, 8, 10, 11, 24, 25, 37, 42, 44, 45, 47 -- 49] for more existence results of
extremal functions for Moser -- Trudinger type inequalities.
Suggesting by the concentration -- compactness principle due to Lions [29], Adimurthi
and Druet established in [2] the following generalization of Moser -- Trudinger inequality on
any bounded domain Ω ⊂ R2
sup
0 (Ω), k∇uk2≤1ZΩ
u∈H 1
e4π(1+αkuk2
2)u2
dx < ∞,
(1.2)
for any 0 ≤ α < λ(Ω), where λ(Ω) = inf u∈H 1
2 is the first eigenvalue of
Laplace operator −∆. The existence of extremal function for (1.2) was proved by Yang
in [44]. This result was extended by Yang [45, 46] to the cases of high dimension and
compact Riemannian surfaces, by Lu and Yang [31] and Zhu [51] to the version of L−norm,
by Souza and do ´O [10, 11] to the whole Euclidean space, and by Tintarev [40] to the
following form
0 (Ω), kuk2≤1 k∇uk2
sup
0 (Ω), k∇uk2
u∈H 1
2−αkuk2≤1ZΩ
e4πu2
dx < ∞,
(1.3)
with 0 ≤ α < λ(Ω). Evidently, (1.3) implies (1.2). In [47], Yang generalized (1.3) to the
cases that large eigenvalues are involved, as well as to the manifold case. The existence
2
of extremal functions for (1.3) also obtained in [47]. In [48], Yang and Zhu studied the
singular version of (1.3). They proved the existence of extremal functions for the following
singular Moser -- Trudinger inequality
sup
0 (Ω), k∇uk2
u∈H 1
2−αkuk2≤1ZΩ
e4π(1−β)u2
x2β
dx < ∞, α < λ(Ω)
(1.4)
where Ω is a smooth bounded domain in R2 containing the origin in its interior and
0 ≤ β < 1. The same existence result for the singular Moser -- Trudinger inequality on
whole Euclidean space was recently proved by Yang and Zhu in [49].
Adams inequality is the version of higher order of derivatives of Moser -- Trudinger in-
equality. The study of this inequality was started by the work of Adams [1]. To state
Adams inequality, we use the symbol ∇mu with m is a positive integer, to denote the mth
order gradient for u ∈ C m, the class of mth order differentiable functions,
∇mu =(∆m/2u
if m even,
∇∆(m−1)/2u if m odd,
where ∇ and ∆ denotes the usual gradient operator and usual Laplacian respectively.
Adams proved in [1] that for any positive integer m less than n, there exists a constant
C0(n, m) such that for any bounded domain Ω ⊂ Rn, it holds
m, n
m
0
u∈W
sup
(Ω), k∇muk n
m ≤1
1
ΩZΩ
exp(βu
n
n−m )dx ≤ C0(n, m),
(1.5)
for any β ≤ β(n, m) with
β(n, m) =
if m odd,
if m even.
n
n
2 )
Γ( n−m+1
2 )
ωn−1 h πn/22mΓ( m+1
) i
ωn−1 h πn/22mΓ( m
) i
Γ( n−m
2
2
n
n−m
n
n−m
Furthermore, for β > β(n, m) the supremum above will be infinite. Notice that when
m = 1, (1.5) reduces to Moser -- Trudinger inequality (1.1).
n
Remark that the work of Moser and of Carleson and Chang was based on the rearrange-
ment argument to reduce problem to the one-dimensional problem. However, we can not
adapt this symmetrization technique in the case m ≥ 2 since we do not know whether the
m norm of the mth gradient of a function decreases under the rearrangement operator. In
L
order to establish (1.5), Adams use the representation of u in terms of its gradient function
∇mu using a convolution operator, and then apply O'Neil's idea [34] of rearrangement of
convolution of two functions together with the idea which originally goes back to Garcia.
Such an argument avoids in dealing with the issue of L
m norm preserving of the gradient
of the rearranged functions. This idea has also been developed to derive the sharp Adams
inequality on Riemannian manifolds without boundary by Fontana [14], on the measure
n
3
spaces by Fontana and Morpurgo [15]. The sharp Adams inequality was also generalized to
whole Euclidean space in the works of Fontana and Morgurgo [16], of Lam and Lu [19, 20]
and of Ruf and Sani [38]. The sharp Adams inequality was recently established on the
hyperbolic spaces by Karmakar and Sandeep [17].
It remains an open problem whether Adams inequality has an extremal function. Unlike
in Moser -- Trudinger inequality with first order derivative, we can not adapt Carleson --
Chang's idea [4] of symmetrization to establish the existence of extremal function for
inequalities of higher order derivatives. It is still a rather difficult problem to answer the
above question in the most generality. One interesting case of the above question when
n = 4 and m = 2 was addressed in [30]. Let Ω ⊂ R4 denote a smooth oriented bounded
domain, H 2
0 (Ω) denote the Sobolev space which is completion of the space of compactly
supported smooth functions in Ω under the Dirichlet norm kukH 2
0 (Ω) = k∆uk2. Then
Adams inequality in the case n = 4 and m = 2 states that
sup
0 (Ω), k∆uk2≤1ZΩ
u∈H 2
eγu2
dx < ∞,
(1.6)
for any γ ≤ 32π2. The existence of extremal function for inequality (1.6) was proved by Lu
and Yang [30]. Even, Lu and Yang established in [30] an improvement of (1.6) in spirit of
Adimurthi and Druet (for improvement of Moser -- Trudinger inequality (1.2)). Let λ1(Ω)
denote the first eigenvalue of the bi-Laplacian operator ∆2 on Ω, i.e.,
λ1(Ω) =
inf
0 (Ω), u6≡0
u∈H 2
2
k∆uk2
kuk2
2
.
An easy application of the variational method shows that λ1(Ω) > 0 and is attained. It
was proved by Lu and Yang that
sup
0 (Ω), k∆uk2≤1ZΩ
u∈H 2
e32π2q(kuk2
2)u2
dx < ∞,
(1.7)
where q(t) = 1 + a1t+· · ·+ aktk, k ≥ 1 is a polynomial of order k in R with 0 ≤ a1 < λ1(Ω),
0 ≤ a2 ≤ λ1(Ω)a1, . . . , 0 ≤ ak ≤ λ1(Ω)ak−1. Furthermore, if a1 ≥ λ1(Ω) then the
supremum above will be infinite.
The existence of extremal functions for inequality (1.7) was also studied in [30]. It was
proved that there exists a strictly positive constant ǫ0 < λ1(Ω) depending only on Ω such
that when 0 ≤ a1 < λ1(Ω), 0 ≤ a2 ≤ λ1(Ω)a1, . . . , 0 ≤ ak ≤ λ1(Ω)ak−1, we can find
u∗ ∈ H 2
0 (Ω) ∩ C 4(Ω) such that k∆u∗k2 = 1 and
e32π2q(ku∗k2
2)u∗ 2
dx =
ZΩ
sup
0 (Ω), k∆uk2≤1ZΩ
u∈H 2
e32π2q(kuk2
2)u2
dx.
Obviously, this implies the existence of extremal functions for Adams inequality (1.6).
4
The first aim of this paper is to strengthen Adams inequality (1.6) in the spirit of
Tintarev for the improvement of Moser -- Trudinger inequality (1.3).To do this, let us define
for any 0 ≤ α < λ1(Ω),
kuk2
2,α = k∆uk2
2 − αkuk2
2,
u ∈ H 2
0 (Ω).
Note that k · k2,α is a norm on H 2
will prove the following inequality.
0 (Ω) which is equivalent to k · kH 2
0 (Ω). In this paper, we
Theorem 1.1. Let Ω ⊂ R4 be a smooth, oriented bounded domain, λ1(Ω) be the first
eigenvalue of bi-Laplacian operator ∆2 on Ω. Then for any α with 0 ≤ α < λ1(Ω), we have
sup
0 (Ω), kuk2,α≤1ZΩ
u∈H 2
e32π2u2
dx < ∞.
(1.8)
Remark that when α = 0, (1.8) reduces to (1.6). Moreover, for any 0 ≤ α < λ1(Ω) and
0 (Ω) such that k∆uk2 ≤ 1 denote v = u/kuk2,α, then u2 ≤ v2, and kvk2,α = 1, thus
u ∈ H 2
(1.8) is indeed stronger than Adams inequality (1.6). The next result shows that (1.8) is
stronger than the inequality of Lu and Yang (1.7).
Proposition 1.2. Theorem 1.1 implies the inequality (1.7).
The second result of this paper is the existence of the extremal functions for the in-
equality (1.8). More precisely, we prove the following result.
Theorem 1.3. Let Ω ⊂ R4 be a smooth, oriented bounded domain, λ1(Ω) be the first
eigenvalue of bi-Laplacian operator ∆2 on Ω. Then for any α with 0 ≤ α < λ1(Ω), there
exists u∗ ∈ H 2
0 (Ω) ∩ C 4(Ω) such that ku∗k2,α = 1 and
e32π2u∗ 2
dx =
ZΩ
sup
0 (Ω), kuk2,α≤1ZΩ
u∈H 2
e32π2u2
dx.
Note that when α = 0 we obtain the existence of extremal function for Adams inequality
(1.6) which was already proved in [30]. Although, our inequality (1.8) is stronger than the
one of Lu and Yang (1.7), however the existence result in Theorem 1.3 does not imply the
existence result for the inequality (1.7). Also, contrary with the existence result of Lu and
Yang, our Theorem 1.3 gives the existence of extremal function for the inequality (1.8) for
any 0 ≤ α < λ1(Ω).
We conclude this introduction by mentioning about the method of proof of our main
Theorems. As usually, our method is based on the blow-up analysis method. We first
establish a concentration-compactness lemma of Lion's type and using it to prove the
existence of uǫ ∈ H 2
0 (Ω) ∩ C 4(Ω), ǫ ∈ (0, 32π2) such that kuǫk2,α = 1 and
ZΩ
0 (Ω), kuk2,α≤1ZΩ
e(32π2−ǫ)u2
e32π2uǫ
u∈H 2
sup
2
dx =
dx.
5
Thus, the Euler -- Lagrange equation of uǫ is given by
e(32π2−ǫ)u2
∆2uǫ = 1
ǫ uǫ + αuǫ
λǫ
kuǫk2,α = 1, uǫ = ∂uǫ
∂ν = 0
λǫ =RΩ e(32π2−ǫ)u2
ǫ u2
ǫ dx,
in Ω,
on ∂Ω,
sequence of test functions, we exclude the blow-up phenomena for the maximizing sequence
where ν denotes the outward unit normal vector to ∂Ω. Without loss of generality, let
cǫ = maxΩ uǫ = uǫ(xǫ). If cǫ is bounded, by the standard regularity theory we obtain
uǫ → u∗ in C 4(Ω) hence finishes our proof. If cǫ → ∞ (namely, the blow-up occurs) and
xǫ → p ∈ Ω, by using Pohozaev type identity and elliptic estimates, we exclude the case
p ∈ ∂Ω. We also show that cǫuǫ converges to some Green function weakly in H 2
0 (Ω) which
then immediately leads to Theorem 1.1. We also prove an upper bound for functional
RΩ e32π2u2dx when blow-up occurs by using some capacity estimates. By constructing a
of functionalRΩ e32π2u2dx. This leads to the existence result in Theorem 1.3. We emphasize
here that in our proof below, we do not require the sharp Adams inequality (i.e., γ = 32π2
in (1.6)), but only require the subcritical Adams inequality (i.e., γ < 32π2 in (1.6)). We also
would like to mention here that blow-up analysis technique have been already employed by
numerous authors in relevant but quite different setting in dealing with Sobolev inequalities
instead of Moser -- Trudinger inequality. We refer the interested reader to the works [3, 10 --
12, 21 -- 25, 30, 42, 44 -- 49], etc.
The rest of this paper is organized as follows. In section §2, we give the existence of
maximizers for subcritical functional. In section §3 we analyse the asymptotic behavior
of those maximizers functions. In section §4, we obtain an upper bound for the critical
functional under the assumption that blow-up occurs in the interior of Ω by using some
capacity estimates. We exclude the boundary bubble in section §5. The proof of Theorem
1.1 and Proposition 1.2 is given in section §6. In section §7, we construct a sequence of
test functions to conclude the existence of extremal function for the critical functional and
thus give the proof of Theorem 1.3.
2 Extremals for the subcritical Adams inequality
For any ǫ ∈ (0, 32π2), let us consider the subcritical problems
e(32π2−ǫ)u2
Cǫ =
dx.
sup
0 (Ω), kuk2,α≤1ZΩ
u∈H 2
(2.1)
In this section, we mainly prove that Cǫ < ∞ and the subcritical problem (2.1) is attained.
Noting that the existence of such extremals is nontrivial. In the proof, we need the following
Lion's type [29] concentration -- compactness principle.
6
Proposition 2.1. Let {uj}j ⊂ H 2
and uj ⇀ u0 weakly in H 2
0 (Ω) be a sequence of functions such that kujk2,α = 1
0 (Ω). Then for any p < (1 − ku0k2
2,α)−1,
lim sup
j→∞ ZΩ
e32π2pu2
j dx < ∞.
Proof. By Rellich -- Kondrachov theorem, we have kujk2 → ku0k2 as j → ∞. Denote
vj =
uj
k∆ujk2
=
uj
(1 + αkujk2
2)1/2 ,
then k∆vjk2 = 1 and vj ⇀ v0 = u0/(1 + αku0k2)1/2 weakly in H 2
0 (Ω). Applying the Lions
type concentration -- compactness principle of Lu and Yang (see Proposition 3.1 in [30]), we
have
lim sup
j→∞ ZΩ
e32π2qv2
j dx < ∞
(2.2)
for any q < 1/(1 − k∆v0k2
2). For any p < 1/(1 − ku0k2
2,α) we have
lim
j→∞
pk∆ujk2
2 = p(1 + αku0k2
1 + αku0k2
This implies the existence of j0 and q < 1/(1 − k∆v0k2
2) <
2
2 − k∆u0k2
2) such that
1 + αku0k2
2
=
pk∆ujk2
2 ≤ q <
1
1 − k∆v0k2
2
,
∀ j ≥ j0.
Thus by (2.2), we get
lim sup
j→∞ ZΩ
as our desire.
e32π2pu2
j dx = lim sup
j→∞ ZΩ
e32π2pk∆ujk2
2v2
j→∞ ZΩ
j dx ≤ lim sup
Our existence result is given in the following proposition.
1
1 − k∆v0k2
2
.
e32π2qv2
j dx < ∞,
Proposition 2.2. For any ǫ ∈ (0, 32π2), we have Cǫ < ∞ and there exists uǫ ∈ H 2
such that kuǫk2,α = 1 and
0 (Ω)
Note that 32π2 − ǫ can be replaced by any sequence {ρǫ}ǫ with ρǫ ↑ 32π2.
Proof. Let {uj}j ⊂ H 2
0 (Ω) be a sequence of functions with kujk2,α = 1 and
Cǫ =ZΩ
e(32π2−ǫ)u2
ǫ dx.
lim
j→∞ZΩ
e(32π2−ǫ)u2
j dx = Cǫ.
7
Since α ∈ [0, λ1(Ω)) then
1 = kujk2
2,α ≥(cid:18)1 −
α
λ1(Ω)(cid:19)k∆ujk2
2.
0 (Ω). Up to a subsequence, we can assume that uj ⇀ uǫ weakly
0 (Ω), uj → uǫ in Lp(Ω) for any 1 < p < ∞ and uj → uǫ a.e., in Ω. If uǫ = 0, then by
Thus {uj}j is bounded in H 2
in H 2
Rellich -- Kondrachov theorem, we have kujk2 → 0 as j → ∞. Define
vj =
uj
k∆ujk2
=
then k∆vjk2 = 1 and vj → 0 a.e., in Ω. Since
uj
(1 + αkujk2
2)2 ,
lim
j→∞
(32π2 − ǫ)(1 + αkujk2
2) = 32π2 − ǫ,
and by Adams inequality
then there exists p > 1 such that
sup
j≥1ZΩ
e32π2v2
j dx < ∞,
sup
j ZΩ
e(32π2−ǫ)pu2
j dx < ∞.
Thus, since vj → 0 a.e., in Ω, by letting j → ∞ we get
j→∞ZΩ
which is impossible. Hence uǫ 6≡ 0 and
lim
e(32π2−ǫ)u2
j dx = Ω,
0 < k∆uǫk2
It follows from Proposition 2.1 that
2 − αkuǫk2
2 ≤ lim inf
j→∞ kujk2
2,α ≤ 1,
sup
j≥1ZΩ
e32π2pu2
j dx < ∞
for any p < 1/(1 − kuǫk2
2,α). This together uj → uǫ a.e., in Ω implies
j→∞ZΩ
j dx =ZΩ
e(32π2−ǫ)u2
e(32π2−ǫ)u2
ǫ dx.
lim
This shows that Cǫ < ∞. Obviously, we must have kuǫk2,α = 1. Hence uǫ is a maximizer
for Cǫ.
8
An easy computation shows that the Euler -- Lagrange equation of uǫ is given by
eαǫu2
ǫ uǫ + αuǫ
uǫ = ∂uǫ
∆2uǫ = 1
λǫ
kuǫk2,α = 1,
λǫ =RΩ eαǫu2
αǫ = 32π2 − ǫ,
Lemma 2.3. It holds lim inf ǫ→0 λǫ > 0.
Proof. Using the inequality et ≤ 1 + tet for t ≥ 0, we get
∂ν = 0
ǫ u2
ǫ dx.
Cǫ =ZΩ
eαǫu2
ǫ dx ≤ Ω + αǫλǫ.
in Ω
on ∂Ω
(2.3)
(2.4)
It is evident that
lim sup
ǫ→0
Cǫ ≤
sup
0 (Ω), kuk2,α=1ZΩ
u∈H 2
e32π2u2
dx.
For any u ∈ H 2
0 (Ω) with kuk2,α = 1, by Fatou's lemma we have
dx ≤ lim inf
ǫ→0
ǫ→0 ZΩ
dx ≤ lim inf
e32π2u2
ZΩ
eαǫu2
Cǫ.
Taking the supremum over all such functions u, we have
sup
0 (Ω), kuk2,α=1ZΩ
u∈H 2
e32π2u2
dx ≤ lim inf
ǫ→0
Cǫ.
Thus we have shown that
Cǫ =
lim
ǫ→0
sup
0 (Ω), kuk2,α=1ZΩ
u∈H 2
e32π2u2
dx > Ω.
(2.5)
Combining (2.4) and (2.5) together we obtain the desired estimate.
3 Asymptotic behavior of extremals for subcritical
functionals
The crucial tool in studying the regularity of higher order equations is the Green's repre-
sentation formula. The Green function G(x, y) for ∆2 under the Dirichlet condition is the
solution of
∆2G(x, y) = δx(y)
in Ω,
G(x, y) =
∂G(x, y)
∂ν
= 0 on ∂Ω.
(3.1)
All functions u ∈ H 2
0 (Ω) ∩ C 4(Ω) satisfying ∆2u = f can be represented by
u(x) =ZΩ
G(x, y)f (y)dy.
9
We will need the following useful estimates [9] for G in the analysis below
G(x, y) ≤ C ln(2 + x − y−1),
∇iG(x, y) ≤ Cx − y−i,
i ≥ 1,
(3.2)
for some constant C > 0 and for all x, y ∈ Ω, x 6= y.
standard regularity to (2.3) we obtain uǫ → u∗ in C 4(Ω) for some u∗ ∈ H 2
with ku∗k2,α = 1. This then implies
Denote cǫ = maxx∈Ω uǫ(x) = uǫ(xǫ) for xǫ ∈ Ω. If cǫ is bounded, then applying the
0 (Ω) ∩ C 4(Ω)
e32π2u∗ 2
dx =
ZΩ
sup
0 (Ω), kuk2,α=1ZΩ
u∈H 2
e32π2u2
dx,
which leads to our desired results.
In the sequel, we assume that cǫ → ∞. Without loss of generality we assume that
cǫ = uǫ(xǫ) = max
x∈Ω uǫ(x) → ∞,
xǫ ∈ Ω,
xǫ → p ∈ Ω as
ǫ → 0.
(3.3)
As in [30], we call p the blow-up point. Here and in the sequel, we do not distinguish
sequence and subsequence, the reader can understand it from the context.
Since kuǫk2,α = 1 and α < λ1(Ω) then
k∆uǫk2
2 ≤ 1 +
α
λ1(Ω)
,
hence uǫ is bounded in H 2
in Ls(Ω) for any 1 < s < ∞ and uǫ → u0 a.e., in Ω.
concentration -- compactness principle (Proposition 2.1), there is p > 1 such that
0 (Ω), we can assume that uǫ ⇀ u0 weakly in H 2
0 (Ω), uǫ → u0
If u0 6≡ 0, then by Lions type
sup
ǫ>0ZΩ
e32π2pu2
ǫ dx < ∞.
Hence eαǫuǫ is bounded in Lr(Ω) for some r > 1 provided that ǫ is small enough. Ap-
plying the standard regularity theory to (2.3), we obtain the boundedness of cǫ which is
contradiction with (3.3). Hence, we have
weakly in H 2
in Lr(Ω) for any r > 1, and a.e., in Ω,
0 (Ω),
(3.4)
uǫ ⇀ 0
uǫ → 0
αǫ → 32π2.
In the rest of this section we focus on the case p ∈ Ω (the case p ∈ ∂Ω will be treated
below in §5). We claim that
∆uǫ2dx ⇀ δp
Indeed, if (3.5) does not hold. Since k∆uǫk2
and η > 0 such that Br(p) ⊂ Ω and
lim sup
in the sense of measure.
2 = 1 + αkuǫk2
ǫ→0 ZBr(p) ∆uǫ2dx ≤ 1 − η.
(3.5)
2 → 1 as ǫ → 0, we can find r > 0
10
From Sobolev embedding theorem and (3.4), we have ∇uǫ → 0 strongly in L2(Ω). Let
φ ∈ C∞0 (Br(p)) be a cut-off function with 0 ≤ φ ≤ 1 and φ = 1 on Br/2(p). We have
lim sup
ǫ→0 ZBr(p) ∆(φuǫ)2dx ≤ 1 − η.
ǫ
is bounded in L2/(2−η)(Ω) and hence eαǫu2
By Adams inequality, eαǫφ2u2
is bounded in
L2/(2−η)(Br/2(p)) provided that ǫ is small enough. Applying the standard regularity theory
to (2.3), we have that uǫ is bounded in C 1(Br/4(p)). This contradicts our assumption (3.3).
Hence, we obtain (3.5). In fact, we have shown that there is no other blow-up point if p
lies in Ω and kuǫk2,α = 1.
To proceed, we introduce the following quantities
ǫ
bǫ =
λǫ
ǫ dx
RΩ uǫeαǫu2
,
τ = lim
ǫ→0
cǫ
bǫ
,
σ = lim
ǫ→0 RΩ uǫeαǫu2
RΩ uǫeαǫu2
ǫ dx
ǫ dx
Note that τ ≥ 1 or τ = ∞, σ ≤ 1. We will show that σ = 1 at the end of this section.
Let
.
(3.6)
r4
ǫ =
λǫ
c2
ǫ
e−αǫc2
ǫ , Ωǫ = {x ∈ R4 : xǫ + rǫx ∈ Ω}.
We will show that rǫ converges to zero rapidly. Indeed, for any 0 < γ < 32π2, we have
r4
ǫ c2
ǫ eγc2
ǫ = e(γ−αǫ)c2
ǫZΩ
ǫ eαǫu2
u2
ǫ dx ≤ZΩ
ǫ eγu2
u2
ǫ dx → 0,
(3.7)
here we used Holder inequality, (3.4) and the fact 0 < γ < 32π2. In particular, rǫ → 0 and
Ωǫ → R4 as ǫ → 0. We next define two sequences of functions on Ωǫ by
ψǫ(x) =
uǫ(xǫ + rǫx)
cǫ
, ϕǫ(x) = bǫ(uǫ(xǫ + rǫx) − cǫ) = bǫcǫ(ψǫ(x) − 1).
Our next goal is to understand the asymptotic behavior of ψǫ and ϕǫ. Evidently, ψǫ ≤ 1
and
∆2ψǫ(x) = r4
ψǫ(x)eαǫuǫ(xǫ+rǫx)2
ǫ (cid:18) 1
λǫ
+ αψǫ(x)(cid:19) .
Thus, for any R > 0 and x ∈ BR(0) we have
∆2uǫ(x)2 ≤
1
c2
ǫ
+ αr4
ǫ → 0,
and
ZBR(0) ∆ψǫ2dx =
1
c2
ǫ ZBrǫR(xǫ) ∆uǫ2dx → 0.
These estimates and the standard regularity theory give ψǫ → ψ in C 4
loc(R4) with ∆ψ = 0
in R4. Note that ψǫ ≤ 1 and ψǫ(0) = 1, then ψ ≤ 1 and ψ(0) = 1. Using Liouville
theorem, we conclude that ψ ≡ 1 in R4. Thus, we have proved that
11
Lemma 3.1. It holds ψǫ → 1 in C 4
loc(R4).
We next investigate the convergence of ϕǫ.
Lemma 3.2. Let τ be defined in (3.6). Then ϕǫ → ϕ in C 4
loc(R4), where
ϕ(x) =( 1
16π2τ ln
0
1
1+ π
√6x2
if τ < ∞,
if τ = ∞,
(3.8)
for x ∈ R4.
Proof. Using Green representation formula, we have
for i = 1, 2. Thus, for any R > 0, x ∈ BR(0) and i = 1, 2, by using (3.2) we have
hence
eαǫuǫ(y)2
λǫ
uǫ(x) =ZΩ
∇iuǫ(x) =ZΩ ∇i
G(x, y)(cid:18) 1
xG(x, y)(cid:18) 1
λǫ
ri
∇iϕǫ(x) =(cid:12)(cid:12)(cid:12)(cid:12)
xG(xǫ + rǫx, y)(cid:18) 1
λǫ
uǫ(y)eαǫuǫ(y)2
xǫ + rǫx − yi +
ǫbǫZΩ ∇i
ǫZΩ 1
ǫ ZB2Rrǫ (xǫ)
λǫ
1
λǫ
≤ Cbǫri
≤ Cbǫri
uǫ(y) + αuǫ(y)(cid:19) dy
uǫ(y) + αuǫ(y)(cid:19) dy,
uǫ(y) + αuǫ(y)(cid:19) dy(cid:12)(cid:12)(cid:12)(cid:12)
eαǫuǫ(y)2
eαǫuǫ(y)2
xǫ + rǫx − yi! dy
αuǫ(y)
uǫ(y)eαǫuǫ(y)2
xǫ + rǫx − yi dy +ZΩ\B2Rrǫ (xǫ)
xǫ + rǫx − yi dy!
αuǫ(y)
1
ǫcǫZΩ
Ri + αbǫri
+ZΩ
x − zi +
dz
dy
xǫ + rǫx − yi!
1
λǫ
uǫ(y)eαǫuǫ(y)2
xǫ + rǫx − yi dy
(3.9)
cǫ ZB2R(0)
≤ C bǫ
≤ C(R),
here we use (3.7) and bǫ ≤ cǫ.
A straightforward computation shows that ϕǫ satisfies
∆2ϕǫ(x) =
bǫ
cǫ
ψǫ(x)eαǫ
cǫ
bǫ
(1+ψǫ(x))ϕǫ(x) + αbǫcǫr4
ǫ ψǫ(x).
(3.10)
Since bǫ ≤ cǫ, ψǫ → 1 in C 4
standard regularity theorey to (3.10) that ϕǫ → ϕ in C 4
loc(R4), (3.7), ϕǫ ≤ 0 and (3.9), we obtain by applying the
loc(R4) for some function ϕ. We
12
have two following cases.
• Case 1: τ < ∞. By letting ǫ → 0, then using Lemma 3.1, (3.7) and (3.10) we obtain
(3.11)
e64π2τ ϕ(x)dx < ∞.
Indeed, for any R > 0, we have ψǫ(x) = 1 + oǫ,R(1) where oǫ,R(1) means that
ϕ(x) ≤ ϕ(0) = 0,
e64π2τ ϕ(x),
∆2ϕ(x) =
ZR4
1
τ
oǫ,R(1) = 0 uniformly in BR(0).
lim
ǫ→0
ǫ (x) = 1 + oǫ,R(1) for x ∈ BR(0) or equivalently uǫ(xǫ + rǫx)2 = c2
ǫ (1 + oǫ,R(1)) for
Thus ψ2
x ∈ BR(0). Hence
λǫ =ZΩ
ǫ eαǫu2
u2
ǫ dx ≥ c2
ǫ (1 + oǫ,R(1))ZBRrǫ (xǫ)
eαǫu2
ǫ dx.
Applying Fatou's lemma, we have
ZBR(0)
e64π2τ ϕ2(x)dx ≤ lim inf
ǫ→0 ZBR(0)
ǫ→0 ZBR(0)
= lim inf
eαǫ
cǫ
bǫ
(1+ψǫ(x))ϕǫ(x)dx
eαǫ(uǫ(xǫ+rǫx)2−c2
ǫ )dx
= lim inf
ǫ→0
= lim inf
ǫ→0
≤ lim inf
ǫ→0
= 1,
ǫ )dx
r−4
eαǫ(uǫ(x)2−c2
ǫ ZBRrǫ (xǫ)
ǫRBRrǫ (xǫ) eαǫuǫ(x)2dx
λǫ
c2
c2
c2
ǫRBRrǫ (xǫ) eαǫuǫ(x)2dx
ǫ (1 + oǫ,R(1))RBRrǫ (xǫ) eαǫuǫ(x)2dx
for any R > 0. Letting R → ∞ we get RR4 e64π2τ ϕdx < ∞.
Moreover, we have
∆ϕǫ(x) = bǫr2
ǫ ZΩ
∆xG(xǫ + rǫx, y)(cid:18) 1
λǫ
eαǫuǫ(y)2
uǫ(y) + αuǫ(y)(cid:19) dy.
Hence, for any R > 0, by Fubini theorem we get
ZBR(0) ∆ϕǫ(x)dx ≤ Cbǫr2
ǫ ZΩ
1
λǫ
eαǫuǫ(y)2
uǫ(y)ZBR(0)
1
xǫ + rǫx − y2 dxdy
+ Cαbǫr2
ǫ ZΩ uǫ(y)ZBR(0)
1
xǫ + rǫx − y2 dxdy
≤ C′R2,
13
with C′ independent of R and ǫ. Letting ǫ → 0, we obtain
ZBR(0) ∆ϕ(x)dx ≤ C′R2,
for any R > 0 with C′ independent of R. This fact together (3.11) and the results in [27,43]
implies that
1
1
ϕ(x) =
x ∈ R4.
• Case 2: τ = ∞. From (3.9) we obtain by letting ǫ → 0 that
1 + π√6x2 ,
16π2τ
ln
∆ϕ(x) ≤
C
R2 ,
for any x ∈ BR(0) and for any R > 0 with C independent of R. Let R → ∞ we get
∆ϕ(x) = 0 for any x ∈ R4. Since ϕ(x) ≤ ϕ(0) = 0 for any x ∈ R4, then by Liouville
Theorem, we conclude that ϕ ≡ 0.
We next consider the asymptotic behavior of uǫ away from the blow-up point p. We
have the following result.
Lemma 3.3. bǫuǫ is bounded in H 2,r
constant C depending only on Ω, λ1(Ω) and α0 such that kbǫuǫkH 2,r
α ∈ [0, α0] with α0 < λ1(Ω).
Proof. Let vǫ be the solution of
0 (Ω) for any 1 < r < 2. In particular, there exists a
0 (Ω) ≤ C uniformly for
(∆2vǫ = 1
vǫ = ∂vǫ
λǫ
∂ν = 0
bǫuǫeαǫu2
ǫ
in Ω,
on ∂Ω.
By Green representation formula, we have
vǫ(x) =ZΩ
G(x, y)
1
λǫ
bǫuǫ(y)eαǫuǫ(y)2
dy
and hence for any i = 1, 2, it holds
∇ivǫ(x) ≤ C
bǫ
λǫ ZΩ x − y−iuǫ(y)eαǫuǫ(y)2
dy = CZΩ x − y−i
Applying Holder inequality, we obtain for any 1 < r < 2 that
uǫ(y)eαǫuǫ(y)2
RΩ uǫ(z)eαǫuǫ(z)2dz
dy.
uǫ(y)eαǫuǫ(y)2
RΩ uǫ(z)eαǫuǫ(z)2dz
Thus, by Fubini theorem, we have k∇ivǫkr ≤ C for i = 1, 2, whence
∇ivǫ(x)r ≤ C rZΩ x − y−ir
dy.
kvǫkH 2,r
0 ≤ C.
14
(3.12)
Let wǫ = bǫuǫ − vǫ, then wǫ satisfies
(∆2wǫ = αwǫ + αvǫ
wǫ = ∂wǫ
∂ν = 0,
in Ω,
on ∂Ω.
Using wǫ as testing function for this equation, we get
k∆wǫk2
2 = αkwǫk2
vǫwǫdx ≤
α
λ1(Ω)k∆wǫk2
2 +
Thus
2 + αZΩ
(cid:18)1 −
α
pλ1(Ω)kvǫk2k∆wǫk2.
α
λ1(Ω)(cid:19)k∆wǫk2 ≤
α
pλ1(Ω)kvǫk2,
which together (3.12) and Sobolev inequality yields k∆wǫk2 ≤ C with C depends on Ω,
λ1(Ω), and α0 < λ1(Ω) such that 0 ≤ α ≤ α0. Hence kwǫkH 2
0 (Ω) ≤ C which together (3.12)
implies that bǫuǫ is bounded in H 2,r
0 (Ω) for any 1 < r < 2.
We proceed by showing that bǫuǫ converges to some Green function.
Lemma 3.4. It holds bǫuǫ ⇀ Gα(·, p) in H 2,r
0 (Ω) for any 1 < r < 2 with
(∆2Gα(·, p) = σδp + αGα(·, p)
Gα(·, p) = ∂Gα(·,p)
∂ν = 0
in Ω,
on ∂Ω.
Furthermore, bǫuǫ → Gα(·, p) in C 4
loc(Ω \ {p}). Also, we have
Gα(x, p) = −
σ
8π2 ln x − p + Ap + ψ(x),
(3.13)
(3.14)
where Ap is constant depending on p and α, ψ ∈ C 3(Ω) and ψ(p) = 0.
Proof. By Lemma 3.3, there exists a function Gα(·, p) ∈ H 2,s
0 (Ω) such that bǫuǫ ⇀ Gα(·, p)
weakly in H 2,s
0 (Ω) for any 1 < s < 2. For any r > 0 such that Br(p) ⊂ Ω, by (3.5) we have
eαǫu2
ǫ is bounded in Ls(Ω\ Br(p)) for any s > 1 (this is based on Adams inequality and cut-
off function argument). Hence, by the standard regularity theory we obtain bǫuǫ → Gα(·, p)
in C 4
loc(Ω \ {p}). Notice that bǫuǫ satisfies
(∆2(bǫuǫ) = 1
λǫ
bǫuǫ = ∂(bǫuǫ)
in Ω,
on ∂Ω.
bǫuǫeαǫu2
ǫ + αbǫuǫ
∂ν = 0
(3.15)
For any φ ∈ C∞(Ω) we have
ZΩ
φ (cid:18) 1
λǫ
bǫuǫeαǫu2
ǫ + αbǫuǫ(cid:19) dx =ZΩ
(φ − φ(p))
1
+ φ(p)ZΩ
λǫ
bǫuǫeαǫu2
ǫ dx
1
λǫ
bǫuǫeαǫu2
ǫ dx + αZΩ
bǫuǫφdx.
(3.16)
15
Note that
and
We will show that
lim
ǫ→0ZΩ
bǫuǫφdx =ZΩ
Gα(x, p)φ(x)dx,
(3.17)
lim
ǫ→0ZΩ
1
λǫ
bǫuǫeαǫu2
ǫ dx = lim
lim
ǫ→0ZΩ
(φ − φ(p))
1
λǫ
ǫ→0 RΩ uǫ(x)eαǫuǫ(x)2dx
RΩ uǫ(x)eαǫuǫ(x)2dx
bǫuǫeαǫu2
ǫ dx = 0.
= σ.
(3.18)
(3.19)
Indeed, by Lebesgue dominated convergence theorem, we get that
lim
ǫ→0Z{uǫ≤1}
eαǫu2
ǫ dx = Ω.
This limit and (2.5) imply
lim inf
ǫ→0 ZΩ uǫeαǫu2
ǫ dx ≥ lim inf
= lim inf
ǫ dx
eαǫu2
ǫ→0 Z{uǫ≥1
ǫ→0 (cid:18)Cǫ −Z{uǫ≤1}
0 (Ω), kuk2,α=1ZΩ
sup
=
u∈H 2
eαǫu2
ǫ dx(cid:19)
dx − Ω
e32π2u2
> 0,
hence bǫ/λǫ is bounded. For any r > 0 with Br(p) ⊂ Ω, we know that eαǫu2
Ls(Ω \ Br(p)) for some s > 1 and uǫ → 0 in Lt(Ω) for any t > 1, hence
ǫ is bounded in
(φ − φ(p))
bǫ
λǫ
uǫeαǫu2
ǫ dx = 0.
(3.20)
In the other hand
(φ − φ(p))
bǫ
λǫ
(cid:12)(cid:12)(cid:12)(cid:12)
ZBr(p)
Thus
lim
ǫ→0ZΩ\Br(p)
ǫ dx(cid:12)(cid:12)(cid:12)(cid:12)
uǫeαǫu2
x∈Br(p)φ(x)−φ(p)
≤ sup
ǫ→0ZBr(p)
(φ − φ(p))
bǫ
λǫ
lim
lim
r→∞
bǫ
λǫZΩ uǫeαǫu2
ǫ dx = sup
x∈Br(p)φ(x)−φ(p).
uǫeαǫu2
ǫ dx = 0.
(3.21)
(3.19) follows from (3.20) and (3.21).
Plugging (3.17), (3.18) and (3.19) into (3.16) we obtain
lim
ǫ→0ZΩ
φ (cid:18) 1
λǫ
bǫuǫeαǫu2
ǫ + αbǫuǫ(cid:19) dx = σφ(p) + αZΩ
Gα(x, p)φ(x)dx,
16
for any φ ∈ C∞(Ω), hence
∆2Gα(·, p) = σδp + αGα(·, p)
in Ω.
The last conclusion was proved in the proof of Lemma 4.4 in [30]. Let is recall it here
for convenience of reader. Fix r > 0 such that B2r(p) ⊂ Ω and consider the cut-off function
φ ∈ C∞0 (B2r(p) such that φ ≡ 1 in Br(p). Let
g(x) = Gα(x, p) +
σ
8π2 η(x) lnx − p.
Then we have
with
(∆2g = f
g = ∂g
∂ν = 0 on ∂Ω,
in Ω,
f (x) = −
σ2
8π2 ∆2φ(x) lnx − p + 2∇∆φ(x) ∇ lnx − p + 2∆φ(x) ∆ lnx − p
+ 2∆(∇φ(x) ∇ lnx − p) + 2∇φ(x) ∇∆ lnx − p! + αGα(x, p).
Lemma 3.3 and Sobolev inequality implies that f ∈ Ls(Ω) for any s > 1. By the standard
regularity theory, we have g ∈ C 3(Ω). Let Ap = g(p) and
ψ(x) = g(x) − g(p) +
σ
8π2 (1 − φ(x)) lnx − p,
we obtain the desired result.
ǫ dx.
The following Pohozaev type identity is very useful in our analysis below.
We continue by using Pohozaev type identity to find an upper bound of RΩ eαǫu2
Lemma 3.5. Assume Ω′ ⊂ R4 is a smooth bounded domain. Let u ∈ C 4(Ω′) be a solution
of ∆2u = f (u) in Ω′. Then we have for any y ∈ R4
2Z∂Ω′
+Z∂Ω′(cid:18) ∂v
∂νhx − y,∇vi − h∇u,∇vihx − y, νi(cid:19) dω,
0 f (s)ds, v = −∆u and ν is the normal outward derivative of x on ∂Ω′.
F (u)dx =Z∂Ω′hx − y, νiF (u)dω +
∂νhx − y,∇ui +
4ZΩ′
v2hx − y, νidω + 2Z∂Ω′
∂u
∂ν
vdω
1
∂u
where F (u) =R u
17
The proof of this Pohozaev type identity can be found in [33, 36]. In the sequel, we will
ueαǫu2 + αu. Noting that v = −∆uǫ
apply it for Ω′ = Br(xǫ), y = xǫ, u = uǫ and f (u) = 1
λǫ
and F (u) = 1
2 u2. By Lemma 3.5, we have
eαǫu2 + α
2αǫλǫ
ZBr(xǫ)
eαǫu2
ǫ dx
= −
ααǫλǫ
b2
ǫ
αǫλǫ
4b2
ǫ
αǫλǫ
2b2
ǫ
+
−
r
(bǫuǫ)2dx +
ZBr(xǫ)
4Z∂Br(xǫ)
rZ∂Br(xǫ) ∆(bǫuǫ)2dω −
rZ∂Br(xǫ)(cid:18)2
∂∆(bǫuǫ)
∂ν
eαǫu2
ǫ dω +
ααǫλǫ
b2
ǫ
r
4Z∂Br(xǫ)
(bǫuǫ)2dω
∂(bǫuǫ)
∆(bǫuǫ)dω
αǫλǫ
b2
∂(bǫuǫ)
ǫ Z∂Br(xǫ)
∂ν − h∇∆(bǫuǫ),∇(bǫuǫ)i(cid:19) dω.
∂ν
(3.22)
Using the representation of Gα in Lemma 3.4 and xǫ → p, we have
ZBr(xǫ)
rZ∂Br(xǫ)
(bǫuǫ)2dx =ZBr(p)
Gα(x, p)2dx + oǫ,r(1) = or(1) + oǫ,r(1),
eαǫuǫdω = or(1) + oǫ,r(1),
rZ∂Br(xǫ)
(bǫuǫ)2dω = or(1) + oǫ,r(1),
rZ∂Br(xǫ) ∆(bǫuǫ)2dω = rZ∂Br(p) ∆Gα(x, p)2dω + oǫ,r(1) =
σ2
8π2 + or(1) + oǫ,r(1),
Z∂Br(xǫ)
rZ∂Br(xǫ)
∂(bǫuǫ)
∂ν
∆(bǫuǫ)dω =Z∂Br(p)
∂Gα(x, p)
∂ν
∆Gα(x, p)dω+oǫ,r(1) =
∂∆(bǫuǫ)
∂(bǫuǫ)
∂ν
∂ν
dω = rZ∂Br(p)
∂∆(Gα)
∂(Gα)
∂ν
∂ν
+ oǫ,r(1) = −
σ2
16π2 +or(1)+oǫ,r(1),
σ2
8π2 + or(1) + oǫ,r(1),
and
rZBr(xǫ)h∇∆(bǫuǫ),∇(bǫuǫ)idω = rZBr(p)h∇∆Gα,∇Gαidω+oǫ,r(1) = −
where oǫ,r(1) and or(1) mean that limǫ→0 oǫ,r(1) = 0 when r is fixed, and limr→0 or(1) = 0
respectively. Hence, we get
σ2
8π2 +or(1)+oǫ,r(1),
ZBr(xǫ)
eαǫu2
ǫ dx =
λǫ
b2
ǫ
(σ2 + or(1) + oǫ,r(1)) + or(1) + oǫ,r(1).
(3.23)
We claim that σ2 > 0. Indeed, if this is not true, then σ2 = 0, and we have
ZBr(xǫ)
eαǫu2
ǫ dx =
λǫ
b2
ǫ
(or(1) + oǫ,r(1)) + or(1) + oǫ,r(1).
18
Since ∆uǫ2dx ⇀ δp in the measure sense, then for any r > 0 with Br(p) ⊂ Ω, by using
Adams inequality and cut-off function argument, we have
lim
ǫ→0ZΩ\Br(p)
eαǫu2
ǫ dx = Ω − Br(p).
(3.24)
Fix a r0 > 0 such that B2r0(p) ⊂ Ω and or(1) ≤ 1/4 for any r ≤ 2r0. Choosing ǫ0 > 0
such that xǫ − p < r0 and oǫ,2r0(1) ≤ 1/4 for any ǫ ≤ ǫ0. Thus Br0(p) ⊂ B2r0(xǫ), and
hence
lim sup
ǫ→0 ZΩ\B2r0 (xǫ)
eαǫu2
ǫ dx ≤ Ω − Br0(p) ≤ Ω.
By Holder inequality, we have
λǫ
b2
ǫ
Thus
= (cid:16)RΩ uǫeαǫu2
ǫ eαǫu2
ǫ dx(cid:17)2
ǫ dx ≤ZΩ
RΩ u2
eαǫu2
ǫ dx ≤
1
2
λǫ
b2
ǫ
+
1
2
+ZΩ\B2r0 (xǫ)
eαǫu2
ǫ dx.
lim sup
ǫ→0
λǫ
ǫ ≤ 1 + 2Ω < ∞.
b2
This together the estimates above and σ2 = 0 implies
lim
r→0
lim
ǫ→0ZBr(xǫ)
eαǫu2
ǫ dx = 0.
For any r > 0 such that B2r(p) ⊂ Ω, we then have Br/2(p) ⊂ Br(xǫ) ⊂ B2p(p) for sufficiently
small ǫ > 0. Thus by (3.24), it holds
eαǫu2
ǫ dx = Ω.
(3.25)
lim
r→0
Finally, we get
lim
ǫ→0ZΩ\Br (xǫ)
ǫ→0ZΩ
eαǫu2
lim
ǫ dx ≤ Ω,
which is impossible. Then we must have σ2 > 0. This together (3.23) and (3.25) yields
lim
ǫ→0ZΩ
eαǫu2
ǫ dx = Ω + σ2 lim
ǫ→0
λǫ
b2
ǫ
.
(3.26)
We can further locate σ as follows.
Lemma 3.6. It holds σ = 1.
19
Proof. We know from Lemma 3.4 that bǫuǫ → Gα(·, p) in C 4
loc(Ω \ {p}) with
σ
8π2 lnx − p + Ap + ψ(x), ψ ∈ C 3(Ω), ψ(p) = 0.
Gα(x, p) = −
We also know that σ 6= 0. Suppose σ < 0, then Gα(·, p) ≤ −C in Br(p) for some r > 0
and C > 0. Hence uǫ < 0 in Br(p) \ {p} for ǫ small enough. In the other hand, by Holder
inequality, we have
thus
bǫ ≥ RΩ uǫeαǫu2
RΩ eαǫu2
bǫ ≥ 1 −
lim inf
ǫ→0
ǫ dx
ǫ dx ≥ 1 − R{uǫ≤1}
RΩ eαǫu2
eαǫu2
ǫ dx
,
ǫ dx
supu∈H 2
0 (Ω), kuk2,α=1RΩ e32π2u2dx
Ω
> 0.
Lemma 3.2 implies that uǫ > 0 on BRrǫ(xǫ) for any fixed R > 0 provided that ǫ > 0 is small
enough (since cǫ → ∞). However when ǫ is small enough, we then have BRrǫ(xǫ) ⊂ Br(p).
We thus get a contradiction on the sign of uǫ, hence σ > 0. Whence, Gα(·, p) ≥ C > 0 in
Br(p) \ {p} for some r > 0 and C > 0, hence uǫ > 0 in Br(p) \ {p} for sufficiently small
ǫ > 0, and then we have
ZBr(p) uǫeαǫu2
ǫ dx =ZBr(p)
uǫeαǫu2
ǫ dx.
Since ∆uǫ2dx ⇀ δp in the measure sense, and uǫ → 0 in Ls(Ω) for any s > 1, then by
using Adams inequality and cut-off function argument, we can show that
lim
ǫ→0ZΩ\Br(p) uǫeαǫu2
ǫ dx = 0.
Obviously,
ZΩ uǫeαǫu2
ǫ dx ≥Z{uǫ≥1}
eαǫu2
ǫ dx =ZΩ
eαǫu2
ǫ dx −Z{uǫ≤1}
eαǫu2
ǫ dx
hence by (2.5) and Lebesgue dominated convergence theorem we have
lim inf
ǫ dx ≥
ǫ→0 ZΩ uǫeαǫu2
σ − 1 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ǫ→0 RΩ uǫeαǫu2
RΩ uǫeαǫu2
sup
u∈H 2
0 (Ω), kuk2,α=1ZΩ
ǫ dx − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ 2 lim
ǫ dx
lim
Since
Hence σ = 1.
To summarize, we have the following result.
20
e32π2u2
dx − Ω > 0.
ǫ dx
ǫ→0RΩ\Br(p) uǫeαǫu2
RΩ uǫeαǫu2
ǫ dx
= 0.
Lemma 3.7. bǫuǫ ⇀ Gα(·, p) weakly in H 2,r
0 (Ω) for any 1 < r < 2 with
(∆2Gα(·, p) = δp + αGα(·, p)
Gα(·, p) = ∂Gα(·,p)
∂ν = 0
in Ω
on ∂Ω.
Furthermore, bǫuǫ → Gα(·, p) in C 4
loc(Ω \ {p}). Also we have
Gα(x, p) = −
1
8π2 ln x − p + Ap + ψ(x),
where Ap is constant depending on p and α, ψ ∈ C 3(Ω), with ψ(p) = 0.
4 Capacity estimates
We follow the argument in [30]. Notice that in this section, we still assume that uǫ blows
up and the blow-up point p ∈ Ω. We use capacity estimates to calculate the limit of
ǫ to estimate from above the supremum of the functional RΩ e32π2u2dx over functions
λǫ/b2
u ∈ H 2
0 (Ω) with kuk2,α = 1 under the assumption that uǫ blows up. The technique of using
capacity estimate applied to this kind of problems was discovery by Li [22] in dealing with
Moser -- Trudinger inequality of first order derivatives.
Let u∗ǫ be the function constructed by Lu and Yang in section §5 in [30]. The main
properties of this function are that u∗ǫ ∈ H 2(Bδ(xǫ) \ BRrǫ(xǫ)) and satisfies the boundary
conditions
and enery identity
ZBδ (xǫ)\BRrǫ (xǫ) ∆u∗ǫ2dx =ZBδ(xǫ)\BRrǫ (xǫ) ∆uǫ2dx +
o(1)
b2
ǫ
.
Now we start to derive the capacity estimates. Consider the variational problem
iδ,R,ǫ = infZBδ(xǫ)\BRrǫ (xǫ) ∆u2dx
where infimum takes all over functions belonging to H 2(Bδ(xǫ) \ BRrǫ(xǫ)) with the same
boundary conditions as u∗ǫ . It is well known (see [24, 26]) that this infimum is attained by
a bi-harmonic function T which is defined in the annular domain Bδ(xǫ) \ BRrǫ(xǫ) with
the same boundary condition as u∗ǫ . The explicit form of T is given by
T (x) = A lnx − xǫ + Bx − xǫ2 + Cx − xǫ−2 + D,
21
bǫ (cid:0) 1
u∗ǫ (x) = 1
u∗ǫ (x) = cǫ + 1
bǫ
∂u∗ǫ
∂ν = − 1
∂u∗ǫ
∂ν = 1
bǫrǫ
8π2δbǫ
∂ϕ
8π2 ln 1
on ∂BRrǫ(xǫ),
δ + Ap(cid:1) on ∂Bδ(xǫ),
rǫ (cid:17)
ϕ(cid:16) x−xǫ
rǫ (cid:17)
∂ν (cid:16) x−xǫ
on ∂Bδ(xǫ),
on ∂BRrǫ(xǫ),
(4.1)
(4.2)
with the explicit values of A,B was given in [30] (section §5) by solving a linear system.
Hence
iδ,R,ǫ = 8π2A2 ln
+ 32π2AB(δ2 − R2r2
ǫ ) + 32π2B2(δ4 − R4r4
ǫ ).
δ
Rrǫ
By the same proof of Lemma 5.1 in [30], we conclude that
1
c2
ǫ
lim
ǫ→0
ln
λǫ
c2
ǫ
= 0.
From the definition of rǫ, we have
ln
Rrǫ
δ
= ln
R
δ
+
ln λǫ
c2
ǫ − αǫc2
4
ǫ
.
(4.3)
(4.4)
(4.5)
According to the argument in [30] with the help of (4.4) and (4.5) and using the explicit
values of A and B, we obtain
αǫ 1 +
4π2 ln δ − 2Ap − 1
2ϕ(R) + Rϕ′(R) + 1
bǫcǫ
8π2A2 ln
δ
Rrǫ
32π2
=
8π2
+
ln λǫ
c2
ǫ
+ 8 + 4 ln R
r
αǫc2
ǫ
+ O(cid:18) 1
c4
ǫ
ln2 λǫ
c2
bǫcǫ(cid:19)! (4.6)
ǫ(cid:19) + o(cid:18) 1
ǫ ) = O(cid:18) 1
ǫ(cid:19) .
(4.7)
b2
and
32π2AB(δ2 − R2r2
ǫ ) = O(cid:18) 1
bǫcǫ(cid:19) ,
32π2B2(δ4 − R4r4
Remark that (4.6) is exactly the formula (5.12) in [30] with a mistake on the coefficient of
Rϕ′(R). We correct this mistake in (4.6). From (4.2) and definition of iδ,R,ǫ we have
iδ,R,ǫ ≤ZBδ(xǫ)\BRrǫ (xǫ) ∆uǫ2dx +
o(1)
b2
ǫ
= 1 + αkuǫk2
2 −ZΩ\Bδ (xǫ) ∆uǫ2dx −ZBRrǫ (xǫ) ∆uǫ2dx +
ǫ (cid:18)ZΩ\Bδ(p) ∆Gα2dx +ZBR(0) ∆ϕ2dx − αkGαk2
1
b2
= 1 −
o(1)
b2
ǫ
2(cid:19) +
o(1)
b2
ǫ
,
here we use Lemma 3.2 and Lemma 3.4. By integration by parts, we have
ZΩ\Bδ (p) ∆Gα2dx = −
1
16π2 −
1
8π2 ln δ + Ap + αkGαk2
2 + O(δ ln δ).
(4.8) together (4.9) gives
(4.8)
(4.9)
iδ,R,ǫ ≤ 1 −
1
b2
ǫ (cid:18)ZBr(0) ∆ϕ2dx −
1
16π2 −
1
8π2 ln δ + Ap + O(δ ln δ)(cid:19) +
o(1)
b2
ǫ
.
(4.10)
22
Plugging (4.3), (4.6) and (4.7) into (4.10) and using the fact 32π2/αǫ > 1, we obtain
32π2
8π2
ln λǫ
c2
ǫ
+ 8 + 4 ln R
r
αǫ 2ϕ(R) + Rϕ′(R) + 1
4π2 ln δ − 2Ap − 1
bǫcǫ
bǫcǫ(cid:19)
ǫ(cid:19) + O(cid:18) 1
ln2 λǫ
c2
ǫ (cid:18)ZBr(0) ∆ϕ2dx −
8π2 ln δ + Ap + O(δ ln δ)(cid:19) +
1
16π2 −
+ O(cid:18) 1
≤ −
αǫc2
ǫ
1
b2
!
c4
ǫ
+
1
o(1)
b2
ǫ
.
(4.11)
Multiplying both sides of (4.11) by αǫc2
calculation, we obtain
ǫ , using the fact bǫ ≤ cǫ and making a simple
(cid:20) 32π2
αǫ
+ O(cid:18) 1
c2
ǫ
ln
λǫ
c2
ǫ(cid:19)(cid:21) ln
λǫ
ǫ ≤ −
c2
αǫc2
ǫ
b2
ǫ (cid:18)ZBR(0) ∆ϕ2dx −
− 32π2 cǫ
bǫ (cid:18)2ϕ(R) + Rϕ′(R) +
1
8π2 ln δ(cid:19) −
32π2
αǫ
4 ln
R
δ
1
4π2 ln δ(cid:19) + O(cid:18) c2
ǫ(cid:19) .
ǫ
b2
Notice that
ln
λǫ
c2
ǫ
= ln
λǫ
b2
ǫ
+ ln
b2
ǫ
c2
ǫ
,
32π2
αǫ
= 1 + O(ǫ).
These equalities together (4.4) and the previous inequality imply
ln
λǫ
ǫ ≤ −(1 + o(1))
b2
− (1 + o(1))
1
αǫc2
ǫ
b2
αǫcǫ
ǫ (cid:18)ZBR(0) ∆ϕ2dx −
bǫ (cid:18)2ϕ(R) + Rϕ′(R) +
8π2 ln δ(cid:19) − (4 + o(1)) ln
4π2 ln δ(cid:19) − (1 + o(1)) ln
R
δ
1
c2
ǫ
b2
ǫ
+ O(cid:18) c2
ǫ(cid:19) .
ǫ
b2
(4.12)
Notice that by (2.5) and (3.26) we have limǫ→0 λǫ/b2
ǫ > 0, hence
ln
λǫ
ǫ ≥ −C0,
b2
for some C0 > 0. If τ = limǫ→o
cǫ
bǫ
ln
λǫ
ǫ ≤ (4 + o(1))
b2
c2
ǫ
b2
ǫ
= ∞ then ϕ ≡ 0 by Lemma 3.2 which shows that
+ O(cid:18)c2
ǫ(cid:19) .
ln δ − (4 + o(1)) ln
ln δ − (8 + o(1))
R
δ
cǫ
bǫ
ǫ
b2
Hence for a fixed R > 0, by choosing δ > 0 sufficiently small, we have
−C0 ≤ ln
λǫ
ǫ ≤ 2
b2
c2
ǫ
b2
ǫ
ln δ − (8 + o(1))
cǫ
bǫ
ln δ − (4 + o(1)) ln
R
δ
,
23
which is impossible since the right hand side tends to −∞ when ǫ → 0. This contradiction
proves that 1 ≤ τ < ∞. Whence ln λǫ
is also bounded from above by (4.12). Also by
(4.12) we have
b2
ǫ
λǫ
b2
ln
ǫ ≤(cid:2)4 (τ − 1)2 + o(1)(cid:3) ln δ − (4 + o(1)) ln R + (64π2 + o(1))τ (ϕ(R) + 2Rϕ′(R)) + O (1)
which then implies τ = 1 since otherwise by choosing ǫ > 0 and δ > 0 small enough we
would obtain a contradiction with ln λǫ
b2
ǫ ≥ −C0. With τ = 1, then
ϕ(x) =
1
16π2 ln
1
1 + π√6x2 .
In this situation, bǫ ∼ cǫ and the estimates in (4.7) are improved as (see formula (5.21)
in [30])
32π2AB(δ2 − R2r2
Consequently, (4.11) becomes
ǫ ) = o(cid:18) 1
ǫ(cid:19) ,
c2
32π2B2(δ4 − R4r4
ǫ ) = o(cid:18) 1
ǫ(cid:19) .
c2
2ϕ(R) + Rϕ′(R) + 1
≤ −
(1 + o(1))
c2
ǫ
+
ln λǫ
c2
ǫ
+ 8 + 4 ln R
δ
32π2c2
ǫ
! + O(cid:18) 1
ǫ(cid:19)
ln2 λǫ
c2
ǫ(cid:19) .
8π2 ln δ + Ap + O(δ ln δ)(cid:19) + o(cid:18) 1
c4
ǫ
c2
1
(4.13)
4π2 ln δ − 2Ap − 1
c2
ǫ
8π2
(cid:18)ZBr(0) ∆ϕ2dx −
1
16π2 −
Multiplying both sides of (4.13) by 32π2c2
ǫ we get
(1 + o(1)) ln
λǫ
c2
ǫ
1
1
4π2 ln δ − 2Ap −
1
≤ −32π2(cid:18)2ϕ(R) + Rϕ′(R) +
− (32π2 + o(1))(cid:18)ZBr(0) ∆ϕ2dx −
= −32π2(2ϕ(R) + Rϕ′(R)) − (32π2 + o(1))ZBr(0) ∆ϕ2dx + 32π2Ap
− 4 ln R − 2 + o(1)(1 − ln δ) + O(δ ln δ)
It was computed in [30] (see formula (5.22)) that
8π2(cid:19) − 8 − 4 ln
8π2 ln δ + Ap + O(δ ln δ)(cid:19) + o(1)
1
16π2 −
R
δ
(4.14)
Thus
ZBR(0) ∆ϕ2dx =
ZBR(0) ∆ϕ2dx =
1
16π2 ln(cid:18)1 +
π
√6
R2(cid:19) +
1
96π2 + O(R−2).
1
8π2 ln R +
1
16π2 ln
π
√6
+
1
96π2 + O(R−2).
24
It is easy to see that
and
ϕ(R) = −
1
8π2 ln R −
1
16π2 ln
π
√6
+ O(R−2),
Rϕ′(R) = −
Plugging these estimates into (4.13), we get
1
8π2 + O(R−2).
ln
lim
ǫ→0
λǫ
ǫ ≤
c2
5
3
+ 32π2Ap + ln
π2
6
.
Thus we have proved
sup
0 (Ω), kuk2,α=1ZΩ
u∈H 2
e32π2u2
dx ≤ Ω +
π2
6
5
3 +32π2Ap.
e
(4.15)
5 Nonexistence of boundary bubbles
0 (Ω), uǫ → 0 strongly in H 1
The main result of this section is that the boundary bubbles do not occur. Suppose without
loss of generality that cǫ = uǫ(xǫ) = maxx∈Ω uǫ → ∞ and xǫ → p ∈ ∂Ω. Note that we
have uǫ ⇀ 0 weakly in H 2
0 (Ω), strongly in Ls(Ω) for any s > 1
and a.e., in Ω.
Lemma 5.1. It holds ∆uǫ2dx ⇀ δp in the sense of measure.
Proof. Note that RΩ ∆uǫ2dx = 1 + αRΩ uǫ2dx → 1. If the conclusion of this lemma is
not true, then there is r > 0 small enough such that
lim
ǫ→0ZBr(p)∩Ω ∆uǫ2dx = η < 1.
Choosing φ be a cut-off function on C 4(Ω) such that 0 ≤ φ ≤ 1, φ = 1 on Ω ∩ Br/2(p),
φ = 0 on Ω \ Br(p), and ∇φ ≤ 4/r. Since uǫ ⇀ 0 weakly in H 2
0 (Ω) and uǫ → 0 strongly
in H 1
0 (Ω), hence
lim sup
ǫ→0 ZBp(r)∩Ω ∆(φuǫ)2dx ≤ η.
This together Adams inequality and (2.3) shows that φuǫ ∈ H 2
0 (Ω) is weak solution of
∆2(φuǫ) = fǫ with fǫ is bounded in Ls(Ω) for some s > 1. Applying the standard regularity
theory implies that φuǫ is bounded in C 3(Ω). In particular, cǫ is bounded which contradicts
with our assumption (3.3).
Lemma 5.1 proves that if there is a blow-up point on the boundary ∂Ω, then this is the
unique blow-up point in Ω. We next prove a convergence for bǫuǫ.
25
Lemma 5.2. It holds bǫuǫ ⇀ 0 in H 2,r
0 (Ω) for any 1 < r < 2.
Proof. By the same proof of Lemma 3.3, bǫuǫ is bounded in H 2,r
0 (Ω) for any 1 < r < 2.
Hence there is F ∈ H 2,r
0 (Ω). Using
the same method in the proof of Lemma 3.4, we get that F solves ∆2F = αF in Ω. Since
F ∈ H 2,r
0 (Ω) for any 1 < r < 2, by the standard regularity theory, we have F ∈ C 3(Ω).
However, α < λ1(Ω), we must have F ≡ 0.
0 (Ω) and bǫuǫ → H in H 1
0 (Ω) such that bǫuǫ ⇀ F in H 2,r
Applying Pohozaev type identity (Lemma 3.5) to equation (2.3) on the domain Ω ∩
Br(p), we obtain by the same way in the estimates for σ2 that
lim
ǫ→0ZΩ
eαǫu2
ǫ dx = Ω
which contradicts with (2.5). Therefore, the blow-up point p can not lie on ∂Ω.
6 Proof of Theorem 1.1
Let cǫ, xǫ, p and Ap as before. We have shown in section §3 that if blow-up occurs, i.e.,
cǫ → ∞ then the blow-up point p lies in the interior of Ω, and the supremum
sup
0 (Ω), kuk2,α=1ZΩ
u∈H 2
e32π2u2
dx ≤ Ω +
π2
6
5
3 +32π2Ap.
e
(6.1)
We are in position to prove Theorem
Proof of Theorem. If there exists a function u0 ∈ H 2
0 (Ω), kuk2,α=1ZΩ
e32π2u2
ZΩ
u∈H 2
0 (Ω) such that ku0k2,α = 1 and
0dx =
sup
e32π2u2
dx,
then our proof is finished. Otherwise, the blow-up case occurs, hence Theorem follows
from (6.1).
We finish this section by give a proof of Proposition 1.2 which shows that our inequality
(1.8) implies the one of Lu and Yang (1.7).
Proof of Proposition 1.2. Let a1, a2, . . . , ak, k ≥ 1 be the number such that 0 ≤ a1 < λ1(Ω),
0 ≤ a2 ≤ λ1(Ω)a1, . . . , ak ≤ λ1(Ω)ak−1. It is easy to see that
q(t) ≤ 1 + a1t + a1λ1(Ω)t2 + · · · + a1λ1(Ω)k−1tk.
Denote a = a1/λ1(Ω) < 1 and p(t) = 1 + at + · · · atk then
q(t) ≤ p(λ1(Ω)t).
26
We claim that there exist b ∈ (a, 1) such that
,
1
p(t) ≤
1 − bt
∀ t ∈ [0, 1].
(6.2)
Indeed, this claim is equivalent to
1 − bt
b
(1 + t + · · · + tk−1) ≤
1
a
,
∀ t ∈ [0, 1].
Note that
1 − bt
b
(1 + t + · · · + tk−1) =
1 − b
b
(1 + t + · · · + tk−1) + 1 − tk ≤ k
1 − b
b
+ 1.
Since a < 1, hence we can choose b ∈ (a, 1) such that (6.2) holds.
Denote α = bλ1(Ω) with b is given in (6.2). For any u ∈ H 2
then λ1(Ω)kuk2
2 ≤ 1. By our claim (6.2), we have
2) ≤ p(λ1(Ω)kuk2
q(kuk2
2) ≤
1
1 − αkuk2
2
.
0 (Ω) such that k∆uk2 ≤ 1,
Let
v =
u
(1 − αkuk2
2)1/2 ,
then kvk2,α ≤ 1 and v2 ≥ q(kuk2
2)u2. This together (1.8) implies (1.7).
7 Proof of Theorem 1.3
In this section, we construct functions φǫ ∈ H 2
0 (Ω) such that kφǫk2,α = 1 and
e32π2φ2
ǫ dx > Ω +
π2
6
5
3 +32π2Ap.
e
ZΩ
This fact together (4.15) shows that the blow-up case can not occur, and hence proves our
Theorem.
Denote r = x − p. Recall that
Gα(x, p) = −
1
8π2 ln r + Ap + ψ(x), ψ ∈ C 3(Ω), ψ(p) = 0.
Following the construction in [30] (section §7), let us define
a− 1
16π2 ln(cid:16)1+ π
√6
c
r2
ǫ2 (cid:17)
+ Ap+ψ
c + b
c r2,
if r ≤ Rǫ,
if r > Rǫ,
(7.1)
c +
1
c Gα
φǫ =
where a, b, c are constants determined later such that φǫ ∈ H 2
0 (Ω) and kφǫk2,α = 1.
27
We choose R = − ln ǫ. To ensure that φǫ ∈ H 2
0 (Ω), we choose a, b, c such that
lim
r↑Rǫ
φǫ = lim
r↓Rǫ
φǫ,
The simple computation shows that
lim
r↑Rǫ∇φǫ = lim
r↓Rǫ ∇φǫ.
16π2 ln(cid:16)1 + π√6
1
.
R2(cid:17) − ln(Rǫ)
8π2 − bR2ǫ2,
(7.2)
16π2R2ǫ2(cid:16)1+ π
√6
R2(cid:17)
a = −c2 + 1
b = −
16π2c2 (cid:18)ln
1
It was computed in [30] that
k∆φǫk2
2 =
π
√6ǫ2
+ 16π2Ap −
5
6(cid:19) +
α
c2kGαk2
2 + O(cid:18) 1
c2 ln2 ǫ(cid:19) .
We have
ZΩ\BRrǫ (p)
φ2
ǫ dx =
1
c2
ZΩ\BRrǫ (p)
G2
αdx =
1
c2kGαk2
2 −
1
c2 ZBRrǫ (p)
G2
αdx =
1
c2kGαk2
2 +
O(ǫ4 ln6(ǫ))
c2
.
On BRǫ(p) we have
φǫ(x) =
1
16π2c(cid:18)ln(cid:18)1 +
Ap + ψ
π
√6
−
r2
ǫ2(cid:19)(cid:19) −
R2(cid:19) − ln(cid:18)1 +
R2ǫ2(cid:18)1 −
π
√6
R2ǫ2(cid:19) ,
r2
b
c
ln(Rǫ)
8π2c
+
hence
c
Combining all these estimates together, we get
ZBRǫ(p)
φ2
ǫ dx =
1
c2 O(ǫ4 ln6 ǫ).
kφǫk2
2,α =
1
16π2c2(cid:18)ln
π
√6ǫ2
+ 16π2Ap −
5
6(cid:19) + O(cid:18) 1
c2 ln2 ǫ(cid:19) .
Thus we can choose c such that kφǫk2,α = 1 for ǫ small enough. Moreover, we have
We next compute RΩ e32π2φ2
ZΩ\BRǫ(p)
c2 =
1
5
π
√6ǫ2
+ 16π2Ap −
16π2(cid:18)ln
ǫ dx. On Ω \ BRǫ(p) we have
ǫ dx ≥ZΩ\BRǫ(p)(cid:18)1 +
e32π2φ2
32π2
c2 kGαk2
= Ω +
6(cid:19) + O(cid:18) 1
ln2 ǫ(cid:19) .
32π2
α(cid:19) dx
c2 G2
2 + O(cid:18) 1
ln2 ǫ(cid:19) .
28
(7.3)
(7.4)
On BRǫ(p), using (7.2) and (7.3) we have
φ2
ǫ ≥ c2 + 2(cid:18)a −
= −c2 + 2(a + c2) −
1
16π2 ln(cid:18)1 +
= −
+
1
π
√6ǫ2
16π2 ln
8π2 ln(cid:18)1 +
−
1
ǫ2(cid:19) + 2Ap + 2ψ + 2br2
r2
1
r2
ǫ2(cid:19) + Ap + ψ + br2(cid:19)
π
√6
π
√6
8π2 ln(cid:18)1 +
8π2 ln(cid:18)1 +
π
5
√6
96π2 +
ǫ2(cid:19) + Ap + O(cid:18) 1
ln2 ǫ(cid:19) ,
π
√6
R2(cid:19) −
r2
1
ln(Rǫ)
4π2
here we use the fact ψ(p) = 0, hence ψ = O(cid:0) 1
ln2 ǫ(cid:1) on BRǫ(p) since R = − ln ǫ and also
br2 = O(cid:0) 1
ln2 ǫ(cid:1) on BRǫ(p). Hence, on BRǫ(p), we have
R2(cid:19)4
ǫ ≥(cid:18) π2
ln2 ǫ(cid:19)(cid:19)
(Rǫ)−8(cid:18)1 +
π
√6
ǫ2(cid:19)−4(cid:18)1 + O(cid:18) 1
r2
6ǫ4(cid:19)−1
3 +32π2Apǫ−4(cid:18)1 +
3 +32π2Ap(cid:18)1 +
π
√6
ǫ2(cid:19)−4(cid:18)1 + O(cid:18) 1
π
√6
ln2 ǫ(cid:19)(cid:19) ,
e32π2φ2
π2
6
r2
=
e
e
5
5
since R = − ln ǫ. Integrating on BRǫ(p) and using a suitable change of variable, we get
ZBRǫ(p)
e32π2φ2
ǫ dx ≥(cid:18)1 + O(cid:18) 1
ln2 ǫ(cid:19)(cid:19) e
5
3 +32π2ApZBR(0)
(1 + x2)−4dx.
with R = π1/2R/61/4. Using polar coordinate we get
ZBR(0)
Finally, we have
(1 + r2)4 dr
2
0
r
r3
(1 + x2)−4dx = 2π2Z R
= π2Z R
= π2 1
6 −
6 (cid:18)1 + O(cid:18) 1
π2
=
1
0
2(1 + R
(1 + r)4 dr
ǫ dx ≥
Combining (7.4) together (7.5) we obtain
e32π2φ2
ZBRǫ(p)
π2
6
e
1
3(1 + R
)3!
2
+
2
)2
ln4 ǫ(cid:19)(cid:19) .
3 +32π2Ap + O(cid:18) 1
5
ln2 ǫ(cid:19) .
e32π2φ2
ǫ dx ≥ Ω +
π2
6
ZΩ
5
3 +32π2Ap +
e
29
32π2
c2 kGαk2
2 + O(cid:18) 1
ln2 ǫ(cid:19) .
(7.5)
This together (7.3) imply that for ǫ is sufficiently small
e32π2φ2
ǫ dx > Ω +
π2
6
5
3 +32π2Ap,
e
ZΩ
as our desire.
Acknowledgments
This work is supported by CIMI postdoctoral research fellowship.
References
[1] D. R. Adams, A sharp inequality of J. Moser for higher order derivatives, Ann. of
Math., 128 (2) (1988) 385-398.
[2] Adimurthi, and O. Druet, Blow-up analysis in dimension 2 and a sharp form of
Trudinger -- Moser inequality, Comm. Partial Differ. Equ., 29 (2004) 295 -- 322.
[3] T. Aubin, and Y. Y. Li, On the best Sobolev inequality, J. Math. Pures Appl., 78
(1999) 353 -- 387.
[4] L. Carleson, and S. Y. A. Chang, On the existence of an extremal function for an
inequality of J. Moser, Bull. Sci. Math., 110 (1986) 113-127.
[5] W. S. Cohn, and G. Lu, Best constants for Moser-Trudinger inequalities on the Heisen-
berg group, Indiana Univ. Math. J., 50 (2001) 1567-1591.
[6] W. S. Cohn, and G. Lu, Sharp constants for Moser-Trudinger inequalities on spheres
in complex space Cn, Comm. Pure Appl. Math., 57 (2004) 1458-1493.
[7] G. Csat´o, and P. Roy, Extremal functions for the singular Moser-Trudinger inequality
in dimension two, Calc. Var., 54 (2015) 2341 -- 2366.
[8] G. Csat´o, and P. Roye, Singular MoserTrudinger inequality on simply connected
domain, Commun. in PDE, (2016)
[9] A. Dall'Acqua, and G. Sweers, Estimates for Green function and Poisson kernels of
higher order Dirichlet boundary value problems, J. Differ. Equa., 205 (2004) 466-487.
[10] J. M. do ´O, and M. de Souza, A sharp Trudinger -- Moser type inequality in R2, Trans.
Amer. Math. Soc., 366 (2014) 4513 -- 4549.
[11] J. M. do ´O, and M. de Souza, A sharp inequality of Trudinger -- Moser type and extremal
functions in H 1,n(Rn), J. Differ. Equ., 258 (2015) 4062 -- 4101.
30
[12] O. Druet, and H. Emmanuel, F. Robert, Blow-up theory for elliptic PDEs in Rie-
mannian geometry, Math. Notes, vol. 45, Princeton University press, Princeton, NJ,
2004.
[13] M. Flucher, Extremal functions for the Trudinger-Moser inequality in 2 dimensions
Comment. Math. Helv., 67 (1992) 471 -- 497
[14] L. Fontana, Sharp borderline Sobolev inequalities on compact Riemannian manifolds,
Comment. Math. Helv., 68 (1993) 415 -- 454.
[15] L. Fontana, and C. Morpurgo, Adams inequalities on measure spaces, Adv. Maths.,
226 (2011) 5066 -- 5119.
[16] L. Fontana, and C. Morpurgo, Sharp Adams and Moser-Trudinger inequalities on Rn
and other spaces of infinite measure, preprint, arXiv:1504.04678v3.
[17] D. Karmakar, and K. Sandeep, Adams inequality on the hyperbolic space, J. Funct.
Anal., 270 (2016) 1792-1817.
[18] N. Lam, and G. Lu, Sharp Moser -- Trudinger inequality on the Heisenberg group at the
critical case and applications, Adv. Math., 231 (2012) 3259 -- 3287.
[19] N. Lam, and G. Lu, Sharp Adams type inequalities in Sobolev spaces W m, n
m (Rn) for
arbitrary integer m, J. Differential Equations, 253 (2012) 1143-1171.
[20] N. Lam, and G. Lu, A new approach to sharp Moser -- Trudinger and Adams type
inequalities: a rearrangement -- free argument, J. Differential Equations, 255 (213) 298-
325.
[21] Y. Y. Li, and M. Zhu, Sharp Sobolev trace inequalities on Riemannian manifolds with
boundary, Comm. Pure Appl. Math., 50 (1997) 449 -- 487.
[22] Y. Li, Moser -- Trudinger inequaity on compact Riemannian manifolds of dimension
two, J. Partial Differ. Equa., 14 (2001) 163-192.
[23] Y. Li, Extremal functions for the Moser-Trudinger inequalities on compact Rieman-
nian manifolds, Sci. China Ser. A, 48 (2005) 618648.
[24] Y. Li, and C. Ndiaye, Extremal functions for Moser -- Trudinger type inequality on
compact closed 4−manifolds, J. Geom. Anal., 17 (2007) 669-699.
[25] Y. Li, and B. Ruf, A sharp Trudinger-Moser type inequality for unbounded domains
in Rn, Indiana Univ. Math. J., 57 (2008) 451 -- 480.
[26] J. Li, Y. Li, and P. Liu, The Q−curvature on a 4−dimensional Riemannian manifold
(M, g) with RM QdVg = 8π2, Adv. Math., 231 (2012) 2194 -- 2223.
31
[27] C. Lin, A classification of solutions of conformally invariant fourth order equation in
R4, Comment. Math. Helv., 73 (1998) 203-231.
[28] K. Lin, Extremal functions for Moser's inequality, Trans. Amer. Math. Soc., 348
(1996) 2663 -- 2671.
[29] P. L. Lions, The concentration-compactness principle in the calculus of variations. The
limit case. II, Rev. Mat. Iberoamericana, 1 (1985) 45-121.
[30] G. Lu, and Y. Yang, Adams' inequalities for bi-Laplacian and extremal functions in
dimension four, Adv. Maths., 220 (2009) 1135 -- 1170.
[31] G. Lu, and Y. Yang, The sharp constant and extremal functions for Moser -- Trudinger
inequalities involving Lp norms, Discrete Contin. Dyn. Syst., 25 (2009) 963 -- 979.
[32] J. Moser, A sharp form of an inequality by N. Trudinger, Indiana Univ. Math. J., 20
(1970/71) 1077-1092.
[33] E. Mitidieri, A Rellich type identity and applications, Commun. Partial Differential
Equations, 18 (1993) 125-151.
[34] R. O'Neil, Convolution operators and L(p, q) spaces, Duke Math. J., 30 (1963) 129-
142.
[35] S. I. Pohozaev, On the eigenfunctions of the equation ∆u + λf (u) = 0, (Russian),
Dokl. Akad. Nauk. SSSR, 165 (1965) 36-39.
[36] F. Robert, and J. Wei, Asymptotic behavior of a forth order mean field equation with
Dirichlet boundary condition, Indiana Univ. Math. J., 57 (2008) 2039 -- 2060.
[37] B. Ruf, A sharp Trudinger-Moser type inequality for unbounded domains in R2, J.
Funct. Anal., 219 (2005) 340 -- 367.
[38] B. Ruf, and F. Sani, Sharp Adams-type inequalities in Rn, Trans. Amer. Math. Soc.,
365 (2013) 645 -- 670.
[39] G. Tian, and X. Zhu, A nonlinear inequality of Moser -- Trudinger type, Calc. Var.
Partial Differ. Equ., 10 (2000) 349-354.
[40] C. Tintarev, Trudinger -- Moser inequality with remainder terms, J. Funct. Anal., 266
(2014) 55 -- 66.
[41] N. S. Trudinger, On imbedding into Orlicz spaces and some applications, J. Math.
Mech., 17 (1967) 473-483.
[42] G. Wang, and D. Ye, A Hardy -- Moser -- Trudinger inequality, Adv. Math., 230 (212)
294 -- 320.
32
[43] J. Wei, and X. Xu, Classification of solutions of higher order conformally invariant
equations, Math. Ann., 313 (1999) 207-228.
[44] Y. Yang, Extremal functions for a sharp Moser -- Trudinger inequality, Internat. J.
Math., 17 (2006) 331-338.
[45] Y. Yang, A sharp form of Moser -- Trudinger inequality in high dimension, J. Funct.
Anal., 239 (2006) 100 -- 126.
[46] Y. Yang, A sharp form of Moser -- Trudinger inequality on a compact Riemannian
surfaces, Trans. Amer. Math. Soc., 359 (2007) 5761 -- 5776.
[47] Y. Yang, Extremal functions for Trudinger-Moser inequalities of Adimurthi-Druet type
in dimension two, J. Differ. Equ., 258 (2015) 3161 -- 3193.
[48] Y. Yang, and X. Zhu, Blow-up analysis concerning singular Trudinger -- Moser inequal-
ities in dimension two, J. Funct. Anal., in press.
[49] Y. Yang, and X. Zhu, Extremal functions for singular Trudinger -- Moser inequalities
in the entire Euclidean space, preprint, arXiv:1612.08247v1.
[50] V. I. Yudovic, Some estimates connected with integral operators and with solutions of
elliptic equations, (Russian), Dokl. Akad. Nauk. SSSR, 138 (1961) 805-808.
[51] J. Zhu, Improved Moser -- Trudinger inequality involving Lp norm in n dimensions,
Adv. Nonlinear Study, 14 (2014) 273 -- 293.
33
|
1706.08957 | 2 | 1706 | 2019-08-11T20:31:10 | Nonlinear Fokker-Planck equations with reaction as gradient flows of the free energy | [
"math.FA",
"math.AP"
] | We interpret a class of nonlinear Fokker-Planck equations with reaction as gradient flows over the space of Radon measures equipped with the recently introduced Hellinger-Kantorovich distance. The driving entropy of the gradient flow is not assumed to be geodesically convex or semi-convex. We prove new generalized dissipation inequalities, which allow us to control the relative entropy by its production. We establish the entropic exponential convergence of the trajectories of the flow to the equilibrium. Along with other applications, this result has an ecological interpretation as a trend to the ideal free distribution for a class of fitness-driven models of population dynamics. Our existence theorem for weak solutions under mild assumptions on the nonlinearity is new even in the absence of the reaction term. | math.FA | math |
NONLINEAR FOKKER-PLANCK EQUATIONS WITH REACTION AS GRADIENT
FLOWS OF THE FREE ENERGY
STANISLAV KONDRATYEV AND DMITRY VOROTNIKOV
Abstract. We interpret a class of nonlinear Fokker-Planck equations with reaction as gra-
dient flows over the space of Radon measures equipped with the recently introduced Hel-
linger-Kantorovich distance. The driving entropy of the gradient flow is not assumed to
be geodesically convex or semi-convex. We prove new generalized dissipation inequalities,
which allow us to control the relative entropy by its production. We establish the entropic
exponential convergence of the trajectories of the flow to the equilibrium. Along with other
applications, this result has an ecological interpretation as a trend to the ideal free distri-
bution for a class of fitness-driven models of population dynamics. Our existence theorem
for weak solutions under mild assumptions on the nonlinearity is new even in the absence
of the reaction term.
Keywords: functional inequalities, optimal transport, Hellinger-Kantorovich distance,
geodesic non-convexity
MSC [] D, Q, Q, B
. Introduction
.. Setting. Let Ω be an open connected bounded domain in Rd with sufficiently smooth
boundary and let ν be the outward unit normal along ∂Ω. We are interested in nonnega-
tive solutions of
∂tu = − div(u∇f ) + f u,
∂f
u
∂ν
= 0,
u = u0,
(x, t) ∈ Ω × (0,∞),
(x, t) ∈ ∂Ω × (0,∞),
(x, t) ∈ Ω × 0.
(.)
(.)
(.)
Here u is the unknown function, f = f (x, u(x, t)) is a known nonlinear function of x and u,
equation (.) is the no-flux boundary condition and the initial data u0 are nonnegative.
We refer to Section . for the motivation and background.
uf , ufx ∈ C(Ω × [0,∞))
fu < 0,
lim sup
u→∞
lim inf
u→+0
f (x, u) < 0 ∀x ∈ Ω,
f (x, u) > 0 ∀x ∈ Ω,
(ufx)(cid:12)(cid:12)(cid:12)u=0
f (x, u) + ufu(x, u) + ufxu(x, u) ≤ g(u)
= 0.
a. a. u > 0; g ∈ L1
loc[0,∞),
When needed, we also assume that
(.)
(.)
(.)
(.)
(.)
(.)
(.)
(.)
(.)
S. KONDRATYEV AND D. VOROTNIKOV
When considering problem (.) -- (.), we always make the following assumptions con-
cerning the function f : Ω × (0,∞) → R:
f ∈ C2(Ω × (0,∞))∩ L1
loc(Ω × [0,∞))
either fx = 0 for large u or
either fx = 0 for small u or
lim
u→∞
lim
u→+0
f (x, u) = −∞ ∀x ∈ Ω
f (x, u) = ∞ ∀x ∈ Ω
Remark .. We make comfortable assumptions about the smoothness of f . We do not
insist that f should be defined for u = 0 so as not to exclude the interesting cases such as
f = −(logu + V (x)) (which corresponds to the linear Fokker-Planck equation, cf. [, ])
and f = uα − 1, −1 < α < 0, (the fast diffusion, cf. []). However, we assume in (.) that
the functions uf and ufx admit continuous extensions to Ω × [0,∞). This ensures that
the terms in (.) make sense. Moreover, we assume (.) to avoid certain complications
with the entropy production to be defined below.
Remark .. Assumption (.) is essential, it ensures the parabolicity of (.). The equa-
tion may become degenerate or singular only if u = 0 or u is large. The latter does not
bother us as we only consider bounded solutions in what follows.
Remark .. Assumptions (.), (.) ensure the existence of a positive equilibrium, see
below.
Remark .. Estimate (.) ensures that the entropy and energy of the equation are well-
defined and well-behaved. Note that at least some restrictions on the growth of fu as u →
0 are inevitable, as the related very fast diffusion equation is known to behave abnormally
[].
Remark .. Conditions (.) and (.) are convenient technical assumptions needed for
L∞-bounds (hence for the existence theorem) and for controlling the energy for large u in
the proof of Theorem .. However, they are not necessary everywhere, so we explicitely
mention them when the need arises.
Remark .. The results of the paper remain valid if Ω is the periodic box Td .
FREE ENERGY FOKKER-PLANCK
It follows from (.) -- (.) that for any x ∈ Ω there exists a unique m(x) > 0 such that
f (x, m(x)) = 0.
Clearly, m ∈ C2(Ω). It is a stationary solution of (.), (.). As we will see, all non-zero
solutions of the problem converge to m.
.. Energy and entropy. Now we will introduce the energy and entropy functionals for
equation (.) as well as the notion of weak solution.
Put
Φ(x, u) = −Z u
0
ξfu(x, ξ) dξ, Ψ(x, u) =Z u
0
Φ(x, ξ) dξ.
It is easy to see that
Φ(x, 0) = Ψ(x, 0) = 0, Φu = −ufu, Φx = −Z u
0
ξfxu(x, ξ) dξ, Ψu = Φ.
Observe that both Φ and Ψ are nonnegative and strictly increase with respect to u.
Note that if u is a nonnegative function of x and possibly of t, an L∞-bound on u
is translated into an L∞-bound on Φ(·, u(·)), i. e., the superposition operator associated
with Φ is L∞-bounded. The same is true of Ψ.
Let u be a classical solution of (.) -- (.). Equation (.) can be cast in the equivalent
form
where we write Φ for Φ(x, u(x, t)), etc. Multiplying by Φ(x, u(x, t)) and integrating over Ω,
we obtain
∂tu = ∆Φ − div(Φx + ufx) + uf ,
(.)
(Φx + ufx)·∇Φ dx +ZΩ
uf Φ dx.
(.)
We call the functional
∂tZΩ
Ψ dx = −ZΩ ∇Φ2 dx +ZΩ
W (u) =ZΩ
Ψ(x, u(x)) dx
the energy of problem (.) -- (.) and equation (.), the energy identity. Thus, any classi-
cal solution of (.) -- (.) satisfies the energy identity (.).
For our purposes, the energy identity is useful because it allows us to control the inte-
gral!QT ∇Φ2 dx dt. In particular, we can define the weak solution of (.) -- (.) in a class
of functions u such that Φ(·, u(·)) ∈ L2(0, T ; H1(Ω)). It is easier to exploit this assumption
in the case of equation (.). Thus, we define the weak solution as follows:
Definition .. Let u0 ∈ L∞(Ω). A function u ∈ L∞(QT ) is called a weak solution of (.) --
(.) on [0, T ] if Φ(·, u(·)) ∈ L2(0, T ; H1(Ω)) and
0 ZΩ
Z T
(u∂tϕ + (−∇Φ + Φx + ufx)·∇ϕ + f uϕ) dx dt =ZΩ
u0(x)ϕ(x, 0) dx
(.)
S. KONDRATYEV AND D. VOROTNIKOV
for any function ϕ ∈ C1(Ω× [0, T ]) such that ϕ(x, T ) = 0. A function u ∈ L∞loc([0,∞); L∞(Ω))
is called a weak solution of (.) -- (.) on [0,∞) if for any T > 0 it is a weak solution
on [0, T ].
Now, let us address the entropy of the problem. Define
E(x, u) = −Z u
m(x)
f (x, ξ) dξ.
It follows from (.) that E is well-defined and continuous on Ω × [0,∞). As f decreases
with respect to u and f (x, m(x)) = 0, it is clear that E ≥ 0 and E(x, u) = 0 if and only if
u = m(x). The relative entropy of equation (.) is the functional
E(u) =ZΩ
Observe that it is well-defined at least for u ∈ L∞+ (Ω) as the superposition operator u 7→
E(·, u(·)) is bounded in the spaces L∞+ → L∞+ .
A straightforward computation shows that for a positive classical solution of (.) -- (.)
we have
∂tE(u) = −ZΩ
u(f 2 +∇f 2) dx.
E(x, u(x)) dx.
(.)
(.)
Equation (.) is called the entropy dissipation identity and the integral on the right-hand
side of (.) is called the entropy production. However, the termRΩ u∇f 2 dx may make no
sense for vanishing or non-smooth u. In order to generalise the definition of the entropy
production, we use the identity
u∇f 2 =
1
u−∇Φ + Φx + ufx2
(u > 0).
Given a function u ∈ L∞+ (Ω) such that Φ(·, u(·)) ∈ H1(Ω), the right-hand side of the last
identity is a nonnegative measurable function on [u > 0], so we can define the entropy
production for such functions by the formula
DE(u) =ZΩ
uf 2 dx +Z[u>0]
1
u −∇Φ + Φx + ufx2 dx,
where the second integral on the right-hand side may be infinite. Thus, we see that any
positive classical solution of (.) -- (.) satisfies the entropy dissipation identity
As usual, in the case of weak solutions we establish not the identities (.) and (.)
but rather corresponding inequalities, viz. the energy inequality
∂tE(u) = −DE(u).
∂tW (u) ≤ZΩ(cid:16)−∇Φ2 + (Φx + ufx)·∇Φ + uf Φ(cid:17) dx
and the entropy dissipation inequality
∂tE(u) ≤ −DE(u).
(.)
(.)
(.)
FREE ENERGY FOKKER-PLANCK
For functions u ∈ L∞+ (Ω) such that Φ ∈ L2(0, T ; H1(Ω)) we understand (.) and (.)
in the sense of measures, i. e., that for any smooth nonnegative compactly supported
function χ : (0, T ) → R we respectively have
−Z T
0
χ′(t)W (u) dt ≤"QT
χ(t)(cid:16)−∇Φ2 + (Φx + ufx)·∇Φ + uf Φ(cid:17) dx dt,
Z T
0
χ′(t)E(u) dt ≥Z T
0
χ(t)DE(u) dt.
If (.) holds in the sense of measures, the derivative ∂tE(u) is a nonpositive distribution
and hence a measure, while the entropy E(u) itself a. e. coincides with a non-increasing
function.
An important question is whether the entropy can be controlled by the entropy pro-
duction, since this would imply the exponential stability of the equilibrium. It turns out
that this is true provided that the L1-norm of u is bounded away from 0. Specifically, we
have
Theorem . (Entropy-entropy production inequality). Suppose that f satisfies (.) -- (.)
as well as (.). Let U ⊂ L∞+ (Ω) be a set of functions such that for any u ∈ U , we have
Φ(·, u(·)) ∈ H1(Ω) and
(.)
inf
u∈U kukL1(Ω) > 0.
Then there exists CU such that
Theorem . is a consequence of a fairly general functional inequality established in
E(u) ≤ CU DE(u)
(u ∈ U).
(.)
Section .
Theorem . (Existence of weak solutions). Suppose that f satisfies (.) -- (.) as well
as (.) and (.). Then for any u0 ∈ L∞+ (Ω) there exists a nonnegative weak solution u ∈
L∞(Ω × (0,∞)) of problem (.) -- (.) enjoying the following properties:
() (upper L∞-bound)
kukL∞(Ω×(0,∞)) ≤ inf(ξ ≥ 0 : sup
f (x, ξ) ≤ − ess sup
x∈Ω
x∈Ω
f −(x, u0(x))) ;
() u satisfies the energy inequality (.) in the sense of measures and
() u satisfies the entropy dissipation inequality (.) in the sense of measures and
ess lim sup
t→+0 W (u(t)) ≤ W (u0);
ess sup
t>0
E(u(t)) ≤ E(u0);
() (lower L1-bound)
ku(t)kL1(Ω) ≥ k min(u0, m)kL1(Ω)
a. a. t > 0.
(.)
(.)
(.)
(.)
S. KONDRATYEV AND D. VOROTNIKOV
Remark .. Theorem ., mutatis mutandis, is also valid in the case of the pure Fokker-
Planck equation (.). Even in this case, our conditions on the nonlinearity f are more
relaxed than the ones available in the literature, see, e.g., [, , , , , , , ] and the
references therein.
Remark .. In the general case, uniqueness of solutions cannot be expected due to the
non-Lipschitz reaction term. However, our weak solutions are unique provided the initial
data is bounded away from zero, see Theorem ..
Remark .. Under the hypotheses of Theorem ., the right-hand side of (.) is al-
ways finite (see Remark .). Moreover, if u0 satisfies an estimate ku0kL∞(Ω) ≤ a, inequal-
ity (.) provides an estimate kukL∞(Ω×(0,∞)) ≤ Ca.
The next theorem shows that the solutions that we have constructed exponentially con-
verge to m. Note that (.) is not needed for the long-time convergence.
Theorem . (Convergence to equilibrium). Assume (.) and suppose that a weak so-
lution u of (.) -- (.) with the initial data u0 . 0 satisfies the entropy dissipation inequal-
ity (.), inequality (.), and the lower L1-bound (.). Then u exponentially converges
to m in the sense of entropy:
where γ > 0 can be chosen uniformly over initial data satisfying
E(u(t)) ≤ E(u0)e−γt
a. a. t > 0,
with some c > 0.
k min(u0, m)kL1(Ω) ≥ c
Theorems ., ., and . are proved in Section ..
.. Motivation and background. The nonlinear Fokker-Planck equation
∂tu = − div(u∇(f (x, u)))
(.)
(.)
(.)
is intended to express the behaviour of stochastic systems coming from various branches
of physics, chemistry and biology, see [, , , ]. In order to take into account the
creation and annihilation of mass, the general drift-diffusion-reaction equation (.) was
suggested in []. In the considerations of [] (cf. also []), the crucial role is played
by the free energy functional that up to an additive constant coincides with our relative
entropy functional E from (.). We opt for this change of terminology (though for
thermodynamists the free energy involves the (physical) entropy, the internal energy, and
the temperature) because in mathematical analysis it is convenient to refer to the basic
Lyapunov functional of a system as the entropy, cf. [, p. ].
On the other hand, equation (.) is a general nonlinear model for the spatial dynamics
of a population that is tending to achieve the ideal free distribution [, ] (the distribu-
tion that happens if everybody is free to choose its location) in a heterogeneous environ-
ment. The dispersal strategy is determined by a local intrinsic characteristic of organisms
called fitness (see, e.g., [, ]). The fitness manifests itself as a growth rate, and simul-
taneously affects the dispersal as the species move along its gradient towards the most
FREE ENERGY FOKKER-PLANCK
favorable environment. In (.), u(x, t) is the density of organisms, and f (x, u) is the fit-
ness. The equilibrium u(x) ≡ m(x) when the fitness is constantly zero corresponds to the
ideal free distribution. The original model [, ] assumes a linear logistic fitness
f = m(x)− u
(.)
but in general it can be any nonlinear function of the spatial variable and the density, cf.
[]. The assumptions (.), (.), (.) are natural as they simply mean that the fitness
is decreasing with respect to the population density (as the resources are limited), being
positive for very small densities and negative for very large densities. Our Theorem .
indicates that the populations converge to the ideal free distribution with an exponential
rate.
The existence of weak solutions for the fitness-driven dispersal model (.) -- (.) with
the logistic fitness (.) was shown in [], and the entropic exponential convergence to
m was established in []. The same kind of results for cross-diffusion systems involving
several interacting populations (with logistic fitnesses) can be found in []. Related two-
species models were investigated in [, ], where one population uses the fitness-driven
dispersal strategy and the other diffuses freely or does not move at all. A system of two
interacting populations with a particular nonlinear fitness function has recently been con-
sidered in [], which is the only existing mathematical treatment of a non-logistic fitness
model that we are aware of.
But perhaps our main motivation to study (.) is that it is a gradient flow of the entropy
functional E with respect to the intriguing recently introduced distance on the space of
Radon measures, which is related to the unbalanced optimal transport (i.e., failing to pre-
serve the total transported mass), and that is referred to as the Hellinger-Kantorovich dis-
tance or the Wasserstein-Fisher-Rao distance [, , , , ]. This distance endows the
set of Radon measures with a formal (infinite dimensional) Riemannian metric h·,·i, and
provides first- and second-order differential calculus [] in the spirit of Otto [, , ].
In particular, one can compute the metric gradients of the functionals of the form
F (u) =ZΩ
F(x, u(x)) dx
by the formula
gradF (u) = − div(cid:18)u∇
δF
δu(cid:19) + u
δF
δu
,
(.)
where δF
δu = ∂uF(x, u) stands for the first variation with respect to u and ∇ = ∇x is the usual
gradient in space. We refer to [] for further details and explanations. Since f = −∂uE,
we can recast (.) as a gradient flow
The entropy dissipation identity (.), which by the way was already known to Frank
[], is then nothing but the archetypal property of gradient flows
∂tu = − gradE(u).
(.)
d
dtE(u) = −hgradE(u), gradE(u)iu.
S. KONDRATYEV AND D. VOROTNIKOV
In this connection, we recall that for the metric gradient flows like (.), the geodesic
convexity of the driving entropy functional (or at least semi-convexity, i.e., λ-convexity
with a negative constant λ) makes a difference [, , , , ]. The presence of convex-
ity allows one to apply minimizing movement schemes [, ] to construct solutions to
the gradient flow. Moreover, λ-convexity with λ strictly positive enables the Bakry-Emery
procedure that usually yields the exponential convergence of the relative entropy to zero.
Minimizing movement schemes for Hellinger-Kantorovich gradient flows of geodesically
convex functionals and for related reaction-diffusion equations were suggested in [, ].
Our entropy E is geodesically (−1/2)-convex with respect to the Hellinger-Kantorovich
structure if f = 1 − uα, α > 0, but fails to be semi-convex for f = uα − 1, α < 0, and
for f = − logu (the latter option corresponds to the interesting case of the Boltzmann
entropy). The spatial heterogeneity further complicates the situation. The quadratic (lo-
gistic) multicomponent entropy considered in [, ] is not even semi-convex. All this
can be observed by computing the Hessian of the entropy, cf. [, Section .]; the non-
convexity of the Boltzmann entropy with respect to the Hellinger-Kantorovich metric was
also mentioned in [, , , ]. We refer to [] for a more detailed discussion of ex-
amples of f and the corresponding geodesic non-convexity. However, Santambrogio []
emphasizes that the lack of geodesic convexity is not a universal obstacle for the study of
gradient flows; our results in the current paper and in [, , , , , ] illustrate
this idea.
. Generalized dissipation inequalities
.. Setting. Motivated by the expressions for the entropy and entropy production, we
forget for a while problem (.) -- (.) and consider the integrals
ZΩ
E(x, u(x)) dx,
(g(x, u(x)) + u∇xf (x, u(x))p) dx
ZΩ
(.)
(.)
on their own right. Here Ω a domain in Rd; p ≥ 1; the functions
E, g : Ω × (0,∞) → [0,∞),
f : Ω × (0,∞) → R
are fixed, and u varies over a set U of functions Ω → (0,∞). Observe that the nonnega-
tivity of E and g ensures the existence of the integrals (.) and (.), although they need
not be finite.
The functions f and E introduced in Section . are, of course, prototypes for the ones
appearing in (.) and (.), but we assume no formal relationship between them. In
particular, in this section we do not suppose that f satisfies (.) -- (.).
We would like to know whether (.) can be controlled by (.) uniformly with respect
to u ∈ U. In general, this is not the case, cf. a related discussion in []. However, we
FREE ENERGY FOKKER-PLANCK
show that under suitable assumptions on the functions E, f , and g, (.) does indeed
control (.) provided that the set U of admissible u is separated from 0 in some sense.
For simplicity, we concentrate on the regular case. Section . contains a discussion of
possible generalisations.
Theorem .. Let Ω be a bounded, connected, open domain in Rd admitting the relative
isoperimetric inequality. Let p ≥ 1. Suppose that functions E, g ∈ C(Ω × (0,∞)) and f ∈
C1(Ω × (0,∞)) satisfy
lim
ε→0
E(x, u) < ∞;
E ≥ 0, g ≥ 0;
sup
0<u≤ε
x∈Ω
g(x, u)
E(x, u)
> 0 ∀ε > 0,
inf
u>ε
x∈Ω
E(x,u),0
inf
0<u≤ε
x∈Ω
lim
ε→0
f (x, u) > lim
ε→0
f (x, u).
sup
u>0
E(x,u)<ε
(.)
(.)
(.)
(.)
Finally, suppose that a set U ⊂ C1(Ω) consisting of strictly positive functions contains no
sequence {un} such that {E(·, un(·))} is bounded in L1(Ω) and {un} converges to 0 in measure.
Then there exists a constant C = C(Ω, p, E, g, f , U) such that
ZΩ
E(x, u(x)) dx ≤ C ZΩ(cid:16)g(x, u(x)) + u(x)∇xf (x, u(x))p(cid:17) dx!
(u ∈ U).
(.)
Remark .. The isoperimetric inequality for Ω reads
P(A; Ω) ≥ cΩA
d−1
d , A ⊂ Ω, A ≤
1
2Ω,
(.)
where P(A; Ω) denotes the relative perimeter of a Lebesgue measurable set A of locally
finite perimeter with respect to Ω, cf. [, Remark .], []. We recall that the relative
perimeter is defined as
where µA := ∇1A is the Gauss-Green measure associated with A. The support of µA is
contained [] in the topological boundary of A.
P(A; Ω) = µA(Ω),
Remark .. If E ∈ C(Ω × R+), condition (.) is automatically true. If the set {(x, u) ∈ Ω ×
R+ : E(x, u) = 0} is compact, the right-hand side of (.) is simplified to maxE(x,u)=0 f (x, u)
and likewise, if f ∈ C(Ω × R+), the left-hand side of (.) can be written as minx f (x, 0).
As for (.), it is more tricky. In Section . we show that it always holds in a particular
setting relevant for gradient flows (Theorem .).
Remark .. The infimum in (.) depends on ε and may tend to zero as ε → 0, otherwise
the claim would be trivial.
S. KONDRATYEV AND D. VOROTNIKOV
.. Strategy of the proof of Theorem .. Before starting the proof of Theorem ., we
would like to informally outline the underlying ideas.
For simplicity, we will opt for an argument by contradiction. Of course, a direct proof
could be presented (as we have recently done in [] for a related inequality), and a quan-
titative constant could be derived from it. However, this would be much more cumber-
some, and the constant obtained in this way would anyway not be optimal. Any discussion
of quantitative constants lies beyond the scope of this article.
It easily follows from (.) that g controls E from above unless u is small. Moreover, we
infer (Lemma .) that if the constant in (.) blows up, the sets where either u or E are
small tend to grow and together occupy nearly all of Ω, while the 'transitional annulus' --
where neither is small -- collapses. At this point we must be prepared to face the situation
where the integral
(.)
E dx
g dx
Z[u≪1]
Z[u≪1]
Z[E≪1]
Z[u≫0,E≫0]
g dx
g dx
is controlled neither by
(because (.) is not applicable), nor by
(because g may be small), nor by
(because the 'annulus' is too small).
This is where the term with the gradient comes into play. The crucial observation is that
the total variation of f over the 'annulus' can be estimated from below. Actually, condition
(.) gives a universal lower bound on the variation of f between the 'inner boundary' of
the annulus (say, where u is small) and its 'outer boundary' (where E is small). All in
all, the integral (.) is controlled by the area of the set [u ≪ 1] (due to (.)), which is
controlled by the perimeter of this set (by the isoperimetric inequality), which is in turn
controlled by the total variation of f over the 'annulus'. This eventually leads to a con-
tradiction. Naturally, when this idea is implemented in Lemma . and the subsequent
reasoning, we must relate the total variation and the integralRΩ u∇f p dx. Then we use
the coarea formula and estimate the total variation of f by the perimeters of its superlevel
sets.
.. Proof of Theorem .. Here we prove Theorem .. We start with the following
observations.
FREE ENERGY FOKKER-PLANCK
Under the hypotheses of Theorem ., integral (.) is finite for u ∈ U whenever so is
ZΩ
g(x, u(x)) dx.
Indeed, according to (.) we can choose ε > 0 such that
E(x, u) < ∞.
A := sup
0<u≤ε
x∈Ω
By (.), we have
g(x, u)
E(x, u)
> 0
B := inf
u>ε
x∈Ω
E(x,u),0
(possibly B = ∞). Then E(x, u) ≤ g(x, u)/B whenever u > ε, so
E(x, u(x)) dx +Z[u>ε]
BZΩ
g(x, u(x)) dx < ∞,
E(x, u(x)) dx =Z[u≤ε]
≤ AΩ +
ZΩ
1
E(x, u(x)) dx
as claimed.
Assume that Theorem . is not true. Then there exists a sequence of functions {un} ⊂ U
En dx,
(.)
Take sequences {εn} and {ξn} such that εn > 0, εn → 0,
g(x, u)
E(x, u)
0 < ξn ≤ inf
u>εn
x∈Ω
(this is possible according to (.)), and ξn → 0.
such that
ZΩ
(gn + un∇fnp) dx ≤ εnξnZΩ
where
En(x) = E(x, un(x)),
fn(x) = f (x, un(x)),
gn(x) = g(x, un(x)).
Clearly, En, gn ∈ C(Ω) and fn ∈ C1(Ω). Moreover, it easily follows from (.) -- (.) that
En(x) ≥ 0, gn(x) ≥ 0;
En < ∞;
lim
n→∞
inf
[un≤εn]
and according to the choice of ξn, we have
fn > lim
ε→0
sup
[un≤εn]
lim
n→∞
sup
[En<ε]
fn,
gn ≥ ξnEn on [un > εn].
(.)
(.)
(.)
(.)
S. KONDRATYEV AND D. VOROTNIKOV
thus obtaining a contradiction.
We want to show that the sequence {En} is bounded in L1(Ω) and un → 0 in measure,
We use (.) to estimate
1
εnZΩ
En dx −
En dx −
un∇fnp dx ≤ ξnZΩ
≤ ξnZΩ
≤ ξnZΩ
= −ξn(ε−1
gn dx
1
1
εnZΩ
εnZ[un>εn]
εnZ[un>εn]
ξn
gn dx
En dx
En dx −
n − 1)Z[un>εn]
n − 1)Z[un>εn]
εnZ un∇fnp dx ≤ ξn sup
1
[un≤εn]
un∇fnp dx ≤ −ξn(ε−1
En dx + ξnZ[un≤εn]
En dx + ξnZ[un≤εn]
En[un ≤ εn].
Thus, we have
1
εnZΩ
For large n, the first term on the right-hand side is negative, so we conclude that
En dx.
En dx.
(.)
From (.) we get
Z[un>εn]
En dx ≤
1
ε−1
n − 1Z[un≤εn]
En dx ≤
sup[un≤εn] En[un ≤ εn]
ε−1
n − 1
and by (.), the last expression is bounded uniformly with respect to n. Hence the
sequence {En} is bounded in L1(Ω).
Lemma .. Given a > 0,
lim
n→∞[un > εn]∩ [En > a] = 0.
(.)
Proof. Using (.), we have:
[un > εn]∩ [En > a] ≤
En dx
1
1
aZ[un>εn]∩[En>a]
aZ[un>εn]
≤
≤ [un ≤ εn]
a(ε−1
n − 1)
sup
[un≤εn]
En dx
En → 0 (n → ∞),
where we have taken into account (.), so (.) is proved.
Lemma .. Given a > 0, for large n we have
[En > a] ≤ 2[un ≤ εn].
(.)
(.)
(cid:3)
(.)
Proof. Using the estimate
FREE ENERGY FOKKER-PLANCK
obtained in the proof of Lemma ., we get
[un > εn]∩ [En > a] ≤ [un ≤ εn]
n − 1)
[En > a] ≤ [un ≤ εn] +[un > εn]∩ [En > a] ≤ 1 +
a(ε−1
En
sup
[un≤εn]
sup[un≤εn] En
a(ε−1
n − 1) ![un ≤ εn],
and the lemma follows.
(cid:3)
It follows from (.) that we can choose a > 0, α, and β, all independent of n, such that
for large n we have
sup
[En≤a]
fn ≤ α < β ≤ inf
[un≤εn]
fn.
(.)
We can assume that the limit
exists. It follows from (.) that for large n the sets [un ≤ εn] and [En ≤ a] are disjoint, so
in view of Lemma . we have
lim
n→∞[un ≤ εn]
Thus, we actually face three logical possibilities:
[un ≤ εn] +[En ≤ a] → Ω
lim
n→∞[un ≤ εn] = Ω;
n→∞[un ≤ εn] = 0;
lim
lim
n→∞[un ≤ εn] = µ0 ∈ (0,Ω);
(.)
(.)
(.)
(.)
As εn → 0, (.) clearly implies un → 0 in measure, a contradiction.
In what follows we show that (.) and (.) are in fact impossible. The following
lemma is crucial.
Lemma .. We have
1
1
[En > a]∩ [un > εn]p−1 Z β
un∇fnpdx ≥
εnZΩ
un∇fnp dx ≥Z[En>a]∩[un>εn]∇fnp dx
α
1
εnZΩ
P([fn > t], Ω) dt!p
(.)
(.)
Proof. We have
[En > a]∩ [un > εn]p−1 Z[En>a]∩[un>εn]∇fn dx!p
1
≥
.
(.)
S. KONDRATYEV AND D. VOROTNIKOV
Using the coarea formula, we get:
Z[En>a]∩[un>εn]∇fn dx =Z ∞
≥Z β
−∞
α
P([fn > t]; [En > a]∩ [un > εn]) dx
P([fn > t]; [En > a]∩ [un > εn]) dx
Fix t ∈ (α, β). Evoking the definition of the relative perimeter, we have
P([fn > t]; [En > a]∩ [un > εn]) =(cid:12)(cid:12)(cid:12)µ[fn>t](cid:12)(cid:12)(cid:12) ([En > a]∩ [un > εn]),
where µ[fn>t] is the Gauss-Green measure. Obviously, we have
supp µ[fn>t] ∩ Ω ⊂ ∂Ω[fn > t] ⊂ [fn = t]
for any t ∈ (α, β). It follows from (.) that
[fn = t] ⊂ [En > a]∩ [un > εn],
so
(.)
(.)
and continuing (.), we obtain
supp µ[fn>t] ∩ Ω ⊂ [En > a]∩ [un > εn]
P([fn > t]; [En > a]∩ [un > εn]) =(cid:12)(cid:12)(cid:12)µ[fn>t](cid:12)(cid:12)(cid:12) ([En > a]∩ [un > εn])
=(cid:12)(cid:12)(cid:12)µ[fn>t](cid:12)(cid:12)(cid:12) (Ω)
= P([fn > t]; Ω).
Combining this with (.) and (.), we obtain (.).
(cid:3)
Let us show that (.) is impossible. Assume that it holds.
If at a point x we have fn(x) > t, t ∈ (α, β), (.) guarantees that En(x) > a. Hence, [fn >
t] ⊂ [En > a]. It follows from (.) and (.) that [En ≤ a] → Ω, and thus [En > a] → 0,
so we conclude that [fn > t] is uniformly in t small when n is large. For such large n we
can apply the isoperimetric inequality:
P([fn > t]; Ω) ≥ cΩ[fn > t]
d−1
d .
Now it follows from (.) that [un ≤ εn] ⊂ [fn > t], so we have
Plugging this estimate into (.), we obtain
P([fn > t]; Ω) ≥ cΩ[un ≤ εn]
d−1
d .
Estimating
1
εnZΩ
un∇fnp dx ≥
c
p
Ω(β − α)p[un ≤ εn]p(d−1)/d
[En > a]∩ [un > εn]p−1
.
[En > a]∩ [un > εn] ≤ [En > a] ≤ 2[un ≤ εn]
FREE ENERGY FOKKER-PLANCK
by virtue of (.), we obtain
1
εnZΩ
c
un∇fnp dx ≥
where C is independent of n.
p
Ω(β − α)p[un ≤ εn]p(d−1)/d
2p−1[un ≤ εn]p−1
= C[un ≤ ε]1−p/d ,
Combining obtained estimate with (.), we get:
C[un ≤ εn]1− p
d ≤ ξn sup
[un≤εn]
En[un ≤ εn],
whence
C ≤ ξn sup
[un≤εn]
En[un ≤ εn]
p
d → 0 (n → ∞),
as ξn → 0 and the suprema are bounded by (.). This contradicts the fact that the
left-hand side is a positive constant independent of n. Thus, (.) is impossible.
It remains to show that (.) is also impossible. Assume that it holds.
It is easy to check that in this case we have
P([fn > t]; Ω) ≥ p0
(α < t < β),
where p0 > 0 is independent of t and n. Indeed, we have the inclusions
(.)
and as in our case the measure of the first and third terms goes to µ0 as n → ∞, we also
have
[un ≤ εn] ⊂ [fn > t] ⊂ [En > a]
Now it suffices to apply the isoperimetric equality to [fn > t] if µ0 < 1/2 and to [fn ≤ t]
otherwise.
[fn > t] → µ0 uniformly in t ∈ (α, β).
Plugging (.) into (.), we get
un∇fnp dx ≥
Comparing this with (.), we obtain
1
εnZΩ
1
[un > εn]∩ [En > a]p−1 (β − α)pp
p
0.
1
[un > εn]∩ [En > a]p−1 (β − α)pp
p
0 ≤ ξn sup
[un≤εn]
En[un ≤ εn] → 0 (n → ∞).
As n → ∞, the left-hand side remains bounded away from , while the right-hand side
goes to , a contradiction.
.. Generalisations and specialisations. We start with the remark that Theorem .
can often be applied if U is a subset of a space X of functions defined on Ω provided that
C1(Ω) is dense in X and the integrals (.) and (.) are continuous with respect to the
topology of X. Indeed, if U1 = U ∩ C1(Ω) is dense in U, we apply the theorem to U1 and
proceed by density to make sure that the same constant works for U as well. On the other
hand, if U1 is not dense in U, we replace U with its small enlargement eU in the cone
S. KONDRATYEV AND D. VOROTNIKOV
density argument is used in the proof of Theorem . given in Section ..
of nonnegative functions in X and apply the same reasoning to eU. A more complicated
Another question is whether the constant C can be chosen uniformly with respect to
(E, g, f ) if the latter triple is allowed to vary over a set X . It turns out that Theorem . can
be easily extended to handle this case. Specifically, if the suprema and infima in (.) --
(.) are additionally taken over (E, g, f ) ∈ X , the constant C can be chosen independently
of (E, g, f ). The proof remains essentially the same. Assuming the converse, we have
violating sequences {(eEn, gn, fn)} ⊂ X and {un} ⊂ U such that (.) holds with
En(x) = En(x, un(x)),
fn(x) = fn(x, un(x)),
gn(x) = gn(x, un(x)).
Moreover, the functions En, gn, and fn satisfy (.) -- (.). The rest of the proof can be
reused verbatim.
It should also be noted that the bare u on the right-hand side of (.) can be replaced
by a nonnegative function v(x, u(x)). Of course, in this case it no longer makes sense to
require that U should consist exclusively of positive functions. The separation from 0
should be taken in the sense that no sequence {v(·, un(·))}, where un ∈ U and the sequence
{En(·, un(·))} is bounded in L1(Ω), converges to 0 in measure. However, if v is, for exam-
ple, an increasing function vanishing at 0, this new condition is clearly equivalent to the
original one.
Again, the proof remains essentially unchanged, the sets [un > εn] and [un ≤ εn] being
Summarising, we have the following strengthened version of Theorem .:
replaced by [vn > εn] and [vn ≤ εn], respectively (here vn(x) = v(x, un(x))).
Theorem .. Let Ω be a bounded, connected, open domain in Rd admitting the relative
isoperimetric inequality. Let p ≥ 1 and I be an interval (possibly unbounded). Let X =
{(E, g, f , v)} be a set of tuples such that E, g, v ∈ C(Ω × I), f ∈ C1(Ω × I), and
E ≥ 0, g ≥ 0, v ≥ 0 ∀(E, g, f , v) ∈ X ;
inf( g(x, u)
E(x, u)
lim
ε→0
sup{E(x, u) : (E, f , g, v) ∈ X , (x, u) ∈ Ω × I, v(x, u) ≤ ε} < ∞
: (E, f , g, v) ∈ X , (x, u) ∈ Ω × I, E(x, u) , 0, v(x, u) > ε) > 0 ∀ε > 0
(.)
(.)
(.)
(.)
lim
ε→0
> lim
ε→0
inf{f (x, u) : (E, f , g, v) ∈ X , (x, u) ∈ Ω × I, v(x, u) ≤ ε}
sup{f (x, u) : (E, f , g, v) ∈ X , (x, u) ∈ Ω × I, E(x, u) ≤ ε}
Finally, suppose that a set U ⊂ C1(Ω; I) satisfies the following requirement: for any sequences
{(En, gn, fn, vn)} ⊂ X} and {un} ⊂ U such that the sequence {En(·, un(·))} is bounded in L1(Ω),
the sequence {vn(·, un(·))} does not converge to 0 in measure. Then there exists a constant C
FREE ENERGY FOKKER-PLANCK
depending only on Ω, p, U and X such that
ZΩ
E(x, u(x)) dx
≤ C ZΩ(cid:16)g(x, u(x)) + v(x, u(x))∇xf (x, u(x))p(cid:17) dx!
((E, g, f , v) ∈ X , u ∈ U).
The proof is left to the reader.
Another option would be to allow for nonnegative instead of strictly positive u in The-
orem .. In this case one assumes that E ∈ C(Ω × [0,∞)) and that the supremum in (.)
is taken over 0 ≤ u ≤ ε and x ∈ Ω. The resulting inequality differs from (.) in that the
integral on the right-hand side is taken over [u > 0]. The only modification needed in the
proof is that whenever g or u∇f p are integrated over Ω, the domain of integration should
be changed to [u > 0]. Note that this does not fit into the previous theorem because f can
be undefined on [u = 0].
We conclude by showing that Theorem . is applicable in a situation relevant for gra-
dient flows. In the subsequent formulation, fu and Eu denote the derivatives of the func-
tions f and E, respectively, with respect to their second argument.
Theorem .. Suppose that functions E ∈ C(Ω × [0,∞)), f ∈ C1(Ω × (0,∞)), and m ∈ C(Ω)
satisfy
E(x, u) ≥ 0,
(x, u) ∈ Ω × [0,∞);
m(x) > 0,
E(x, m(x)) = 0,
x ∈ Ω;
x ∈ Ω;
Eu(x, u) = −f (x, u),
(x, u) ∈ Ω × (0,∞);
(.)
(.)
(.)
(.)
fu(x, u) < 0,
(.)
and let U ⊂ C1(Ω) be a set of strictly positive functions having the property that no sequence
{un} ⊂ U such that {E(·, un(·))} is bounded in L1(Ω), converges to 0 in measure. Finally, let
σ ∈ (0, minΩ m) and
(x, u) ∈ Ω × (0,∞)
ξ 2
vσ (ξ) =
max(ξ, σ)
.
Then we have
ZΩ
E(x, u(x)) dx ≤ CZΩ
where C > 0 depends on Ω, f , σ, and U .
vδ(u(x))(cid:16)(f (x, u(x)))2 +∇xf (x, u(x))2(cid:17) dx
(u ∈ U),
(.)
Remark .. Observe that under the hypotheses of Theorem ., the functions E and m
are uniquely determined by f . Indeed, if x ∈ Ω is fixed, E(x, u) as a function of u attains
its minimum at m(x) > 0, so Eu(x, m(x)) = 0, i. e., f (x, m(x)) = 0, according to (.). This
uniquely defines m(x), as it follows from (.) that f (x, u) strictly decreases with respect
to u. Now, E(x, u) is the antiderivative of −f (x, u) with respect to u vanishing at m(x).
S. KONDRATYEV AND D. VOROTNIKOV
Proof. We check the hypotheses of Theorem . with I = (0,∞), p = 2, g(x, u) = vσ (u)(f (x, u))2,
and the set X consisting of the single tuple (E, g, f , vσ ). Clearly, we have (.), while (.) --
(.) are equivalent to (.) -- (.).
Recalling Remark ., we see that (.) holds.
Let us check (.). Fix x ∈ Ω. The function E(x, u) is strictly convex in u and attains its
zero minimum only at u = m(x). As f (x, m(x)) = 0, we see that
On the other hand, as f decreases with respect to u, we have
lim
ε→0
sup
E<ε
f = max
E=0
f = 0.
lim
ε→0
inf
0<u≤ε
x∈Ω
so (.) indeed holds.
f (x, u) ≥ inf
x∈Ω
f (x, σ)
= inf
x∈ΩZ m(x)
σ
≥ min
σ≤u≤m(x)
x∈Ω
(−fu(x, u)) du
(−fu(x, u)) min
x∈Ω
(m(x)− σ) > 0,
It remains to check (.). Without loss of generality, assume that ε > 0 is such that
ε <
min
x∈Ω
1
2
ε < min
x∈Ω
(−2m(x)fu(x, m(x))),
(−fu(x, m(x))).
(.)
(.)
By Cauchy's mean value theorem, for any x ∈ Ω, u > σ, u , m(x), we have
g(x, u)
E(x, u)
=
g(x, u)− g(x, m(x))
E(x, u)− E(x, m(x))
=
gu(x, ξx,u )
Eu(x, ξx,u )
where ξx,u is some point between u and m(x).
= −f (x, ξx,u )− 2ξx,ufu(x, ξx,u ),
(.)
By uniform continuity, there exists δ ∈ (0, minΩ m− σ) such that
ξ − m(x) < δ
implies
− f (x, ξ)− 2ξfu(x, ξ) + 2m(x)fu(x, m(x)) < ε,
fu(x, ξ)− fu(x, m(x)) < ε.
Then from (.) and (.) we see that
Further, using (.) and (.), we have
ξ − m(x) < δ ⇒ −f (x, ξ)− 2ξfu(x, ξ) > ε.
−f (x, m(x) + δ) =Z m(x)+δ
(−fu(x, u) du) ≥ εδ,
m(x)
(.)
(.)
(.)
FREE ENERGY FOKKER-PLANCK
whence, recalling that fu is negative and f is decreasing, we conclude
ξ ≥ m(x) + δ ⇒ −f (x, ξ)− 2ξfu(x, ξ) > εδ.
(.)
Now, if u−m(x) < δ, the point ξx,u also satisfies ξ −m(x) < δ, so we use (.) to conclude
from (.) that
g(x, u)
E(x, u)
> ε.
(.)
If u ≥ m(x) + δ, then either m(x) < ξx,u < m(x) + δ and we again obtain (.), or ξx,u ≥
m(x) + δ and then we use (.) to get
g(x, u)
E(x, u)
> εδ.
g(x, u)
E(x, u) ≥ min
g(x, u)
E(x, u)
min
ε≤u≤m(x)−δ
x∈Ω
> 0,
, ε, εδ
Thus,
inf
u>ε
x∈Ω
E(x,u),0
since the function g/E is continuous and positive on the compact set
{(x, u) : x ∈ Ω, ε ≤ u ≤ m(x)− δ}.
We have showed that (.) holds.
Thus, the hypotheses of Theorem . are fulfilled and the inequality follows.
(cid:3)
.. Positive classical solutions. Let
. Technicalities
θ(s) =
1 if s > 0,
0 if s ≤ 0
be the Heaviside step function.
Lemma .. If nonnegative u, u ∈ C∞(Ω) satisfy the no-flux boundary condition (.), then
(.)
θ(u − u) div(u∇f − u∇ f ) dx ≥ 0,
ZΩ
where f and f stand for f (x, u(x)) and f (x, u(x)), respectively.
Proof. Without loss of generality, the functions u and u are defined and smooth on Rd.
Consider the set Υ := [u− u > 0]. First let us assume that 0 is a regular value of the function
u − u, then the boundary of Υ is smooth. Employing de Giorgi's Gauss-Green formula
S. KONDRATYEV AND D. VOROTNIKOV
[, Theorem .] and the formula for the Gauss-Green measure of an intersection [,
Theorem .], we compute
ZΩ
θ(u − u) div(u∇f − u∇ f ) dx =ZΥ∩Ω
=Z∂∗(Υ∩Ω)
div(u∇f − u∇ f ) dx
(u∇f − u∇ f )· νΥ∩Ω dHd−1 =Z∂Υ∩Ω
+ZΥ∩∂Ω
(u∇f − u∇ f )· νΩ dHd−1 +Z[νΥ=νΩ ]
(u∇f − u∇ f )· νΥ dHd−1
(u∇f − u∇ f )· νΩ dHd−1,
where νΥ∩Ω is the measure-theoretic outward unit normal vector along the reduced bound-
ary ∂∗(Υ∩ Ω) of the intersection []. Due to the no-flux boundary condition, the last two
integrals vanish. On ∂Υ ∩ Ω, we have u = u and consequently, f = f . Thus, we can write
ZΩ
θ(u − u) div(u∇f − u∇ f ) dx =Z∂Υ∩Ω
u∇(f − f )· νΥ dHd−1.
(.)
Due to the monotonicity of f , we have Υ = [f − f < 0]. We see then that whenever ∇(f − f ) ,
0 on ∂Υ, ∇(f − f ) is an outward normal vector along ∂Υ. Thus, ∇(f − f ) · νΥ ≥ 0 and
In the general case, take a decreasing sequence εn → 0 such that 0 is a regular value of
equality (.) gives (.).
u − u − εn. Set
By the above, we have
un = u + εn, fn = f (x, un(x)).
ZΩ
θ(u − un) div(u∇f − un∇ fn) dx ≥ 0.
(.)
(cid:3)
As θ is left-continuous, we have
θ(u − un) → θ(u − u) pointwise in Ω;
moreover, it is clear that
Passing to the limit in (.), we obtain (.).
fn → f in C2(Ω).
Lemma . (L1-contraction for positive classical solutions). Let u and u be classical solu-
tions of (.) -- (.) on [0, T ] with different initial data. Suppose that u and u satisfy
1
κ
with some κ > 0 and let Lκ > 0 be such that
κ ≤ u ≤
, κ ≤ u ≤
1
κ
in QT
u1f (x, u1)− u2f (x, u2) ≤ Lκu1 − u2 x ∈ Ω, ∀u1, u2 ∈(cid:18)κ,
Then for a. a. t > 0,
∂tZΩ
(u − u)+ dx ≤ LκZΩ
(u − u)+ dx.
1
κ(cid:19) .
(.)
(.)
FREE ENERGY FOKKER-PLANCK
Proof. We have:
∂tZΩ
(u − u)+ dx =ZΩ
= −ZΩ
+ZΩ
θ(u − u)(∂tu − ∂t u) dx
θ(u − u) div(u∇f − u∇ f ) dx
θ(u − u)(uf − u f ) dx =: −I1 + I2,
where f and f stand for f (x, u(x, t)) and f (x, u(x, t)), respectively. By Lemma ., we have
I1 ≥ 0. To estimate I2, we use (.) and the observation that the integrand vanishes where
u − u < 0, thus obtaining
I2 ≤ LκZΩ
(u − u)+ dx.
Inequality (.) follows.
For c ∈ R, define uc ∈ C2(Ω) by
f (x, uc(x)) = c.
(cid:3)
(.)
As f is monotonous in u, we see that the function uc is unique, but it does not need to
exist for a given c. Note that u0 = m.
Remark .. There is a simple formula for the L∞-norm of uc:
kuckL∞(Ω) = inf(ξ ≥ 0 : sup
f (x, ξ) ≤ c).
x∈Ω
It follows from the fact that due to the monotonicity of f , the inequality ξ ≥ kuckL∞(Ω)
or, equivalently, ξ ≥ uc(x) for all x ∈ Ω, holds if and only if f (x, ξ) ≤ f (x, uc(x)) ≡ c for all
x ∈ Ω, i. e., when
(.)
(.)
Remark .. If (.) holds, for any u ∈ L∞+ (Ω) the function uc with
f (x, ξ) ≤ c.
sup
x∈Ω
c = − ess sup
x∈Ω
f −(x, u(x))
is well-defined and u ≤ uc a. e. in Ω. Indeed, if the second alternative in (.) holds, for
any x ∈ Ω, the function f (x, ξ) assumes all the values in the interval (−∞, 0] as ξ varies
over [m(x),∞); in particular, f (x, ξ) attains the value c. If, on the other hand, the first
alternative in (.) holds, take ξ1 ≥ kukL∞ such that c1 := f (x, ξ1) is independent of x
and negative. Clearly, for any fixed x ∈ Ω, the function f (x, ξ) takes all the values in
the interval [c1, 0] as ξ varies over [m(x), ξ1]. Now it suffices to observe that due to the
monotonicity of f , we have c ∈ [c1, 0]. One can prove in the same way that if (.) holds,
for any function u essentially bounded away from 0 on Ω, there exists uc such that u ≥ uc a. e.
in Ω, and c ≥ 0.
S. KONDRATYEV AND D. VOROTNIKOV
Remark .. It follows from Remarks . and . that if (.) holds, the right-hand side
of (.) is finite for any u ∈ L∞+ (Ω).
Lemma . (Restricted L1-contraction). Let u be a classical solution of (.) -- (.) on [0,∞).
Then for c ≤ 0 we have
ZΩ
and likewise, for c ≥ 0 we have
ZΩ
(u − uc)+ dx ≤ZΩ
(u − uc)− dx ≤ZΩ
(u0 − uc)+ dx,
t > 0
(u0 − uc)− dx,
t > 0
(.)
(.)
provided that uc exists.
Proof. Let us prove (.) for c ≤ 0. Computing the derivative of the left-hand side, for
a. a. t > 0 we get
∂tZΩ
(u − uc)+ dx =ZΩ
= −ZΩ
+ZΩ
θ(u − uc)∂tu dx
θ(u − uc) div(u∇f ) dx
θ(u − uc)uf dx =: −I1 + I2.
As ∇f (x, uc(x)) ≡ 0, we can use Lemma . to get I1 ≥ 0. Now, the integrand of I2 can
only be non-zero where u > uc, in which case f ≤ c ≤ 0 due to the monotonicity of f ;
consequently, I2 ≤ 0. Thus, we have
∂tZΩ
(u − uc)+ dx ≤ 0
and (.) follows. Inequality (.) is proved in much the same way.
Lemma .. Suppose that f satisfies (.) and (.). Then for any smooth u0 : Ω → (0,∞)
satisfying the non-flux boundary condition, problem (.) -- (.) has a classical solution.
(cid:3)
Proof. Equation (.) can be cast in the form
∂tu = −ufu∆u −∇u · (fx + fu∇u)− u(fxx + 2fxu ·∇u + fuu∇u2 − f ).
If we show that a classical solution is a priori bounded and stays away from , we can
ignore the fact that the coefficient −ufu can be degenerate or singular at u = 0,∞ and infer
the existence of the solution from the classical theory of quasilinear parabolic equations.
Indeed, according to Remark ., we can find uc1 and uc2 such that c2 ≤ 0 ≤ c1 and
Then it follows from Lemma . that
uc1(x) ≤ u0(x) ≤ uc2(x)
(x ∈ Ω).
providing the required bounds.
uc1(x) ≤ u(x, t) ≤ uc2(x, t)
(x, t) ∈ Ω × (0,∞),
(cid:3)
FREE ENERGY FOKKER-PLANCK
.. Positive initial data. If the initial data (.) is bounded away from 0, we approxi-
mate it with smooth functions and prove the existence and uniqueness of weak solutions
to (.) -- (.) stated in Theorem . below.
Lemma .. Suppose that u ∈ L∞+ (QT ) satisfies the energy inequality (.) in the sense of
measures; then
kW (u)kL∞(0,T ) ≤ ess lim sup
k∇Φ(·, u(·))kL2(QT ) ≤ 2(ess lim sup
t→+0 W (u(t)) + CT ,
t→+0 W (u(t)) + CT ),
where C > 0 is determined by an upper bound on kukL∞(Ω).
Proof. The function
t 7→ W (u(t))−Z t
0 −ZΩ ∇Φ2 dx +ZΩ
(Φx + ufx)·∇Φ dx +ZΩ
(.)
(.)
uf Φ dx! dt
has a non-positive derivative in the sense of measures, so it a. e. coincides with a non-
increasing function. In other words, for a. a. t0, t1 ∈ (0, T ), t0 < t1, we have
t0 −ZΩ ∇Φ2 dx +ZΩ
uf Φ dx! dt ≤ 0.
(Φx + ufx)·∇Φ dx +ZΩ
W (u(t1))−W (u(t0))−Z t1
An upper bound on kukL∞(QT ) defines essential upper bounds on uf , Φ = Φ(x, u(x, t)),
Φx, and ufx, so for a. a. t ∈ (t0, t1) we can estimate
(Φx + ufx)·∇Φ dx +ZΩ
ZΩ
uf Φ dx
1
2ZΩ ∇Φ2 dx +
1
2ZΩ Φx + ufx2 dx +ZΩ
≤
uf Φ dx
1
≤
2Z t1
t0 ZΩ ∇Φ2 dx dt ≤ W (u(t0)) + C(t1 − t0).
W (u(t1)) +
2Z t1
0 ZΩ ∇Φ2 dx dt ≤ ess lim sup
1
t→+0 W (u(t)) + CT ,
W (u(t1)) +
1
2ZΩ ∇Φ2 dx + C,
whence
Passing to the essential upper limit as t0 → 0 and estimating t1 − t0 ≤ T , we obtain
whence (.) and (.) follow.
(cid:3)
Theorem . (Solvability for positive data). Suppose that f satisfies (.) -- (.) as well
as (.) and (.). Then for any u0 ∈ L∞+ such that
κ ≤ u0 ≤
1
κ
a. e. in Ω
S. KONDRATYEV AND D. VOROTNIKOV
with some constant κ > 0, there exists a unique weak solution
u ∈ L∞+ (Ω × [0,∞))∩ C([0,∞); L1(Ω))
satisfying the following properties: i) the upper bound (.) and lower bound (.); ii) the
energy and entropy dissipation inequalities as well as (.) and (.); iii) the restricted con-
traction
(u − uc)+ dx ≤ZΩ
ZΩ
ZΩ
(u − uc)− dx ≤ZΩ
(u0 − uc)+ dx
(u0 − uc)− dx
(c ≤ 0),
(c ≥ 0)
(.)
(.)
whenever uc is defined; iv) if u is another such solution with the initial data u0, the L1-
contraction holds:
k(u(t)− u(t))+kL1(Ω) ≤ eLκtk(u0 − u0)+kL1(Ω),
(.)
where Lκ is defined by (.).
Proof. Let {u0
tion such that
n} be a sequence of smooth functions satisfying the no-flux boundary condi-
and
κ ≤ u0
n(x) ≤
1
κ
in Ω
u0
n → u0
in L1(Ω) and a. e. in Ω.
(.)
(.)
Let un be the classical solution of (.) -- (.) on [0,∞) with the initial data u0
T > 0, it follows from Lemma . that
n. For any
kun − umkC([0,T ];L1(Ω)) ≤ eLκTku0
n − u0
mkL1,
so {un} is a Cauchy sequence in C([0, T ]; L1(Ω)). As T is arbitrary, we see that {un} con-
verges in C([0,∞); L1(Ω)) to some function u. We claim that it is the sought-for solution.
By Remark ., there exists uc (c ≤ 0) such that uc ≥ 1/κ; then uc dominates the initial
data u0
n and thus, the solutions un as well, which follows from Lemma .. Consequently,
the sequence {un} is bounded in L∞(Ω × (0,∞)), so it converges to u weakly* in this space,
whence u ∈ L∞(Ω × (0,∞)).
Put
fn = f (x, un(x, t)),
Φn = Φ(x, un(x, t)),
Ψn = Ψ(x, un(x, t)),
fxn = fx(x, un(x, t)),
Φxn = Φx(x, un(x, t)),
En = E(x, un(x, t)).
FREE ENERGY FOKKER-PLANCK
Fix T > 0. As the sequence {un} is bounded in L∞(QT ), so are the sequences {unfn}, {unfxn},
{Φn}, {Φxn}, {Ψn}, and {En}. Thus, there is no loss of generality in assuming
a. e. in QT ,
strongly in any Lp(QT ), 1 ≤ p < ∞,
weakly* in L∞(QT ),
and in the sense of distributions,
(.)
un → u
unfn → uf
unfxn → ufx
Φn → Φ
Φxn → Φx
Ψn → Ψ
En → E
where we write Φ for Φ(·, u(·)), etc. It follows from (.) that ∇Φn → ∇Φ in the sense of
distributions. The approximate solutions satisfy the energy inequality and (.) while
their initial energy is bounded, so we see from (.) that the sequence ∇Φn is bounded
in L2(QT ). Consequently, Φ ∈ L2(0, T ; H1(Ω)) and
∇Φn → ∇Φ weakly in L2(QT ).
(.)
Let us check that u is a weak solution of (.) -- (.) on [0, T ]. Take an admissible test
function ϕ. Writing the weak setting for the approximate solution, we have
0 ZΩ
Z T
(un∂tϕ + (−∇Φn + Φxn + unfxn)·∇ϕ + fnunϕ) dx dt =ZΩ
u0
n(x)ϕ(x, 0) dx.
(.)
It follows from (.), (.), and (.) that we can pass to the limit in (.) and ob-
tain (.) for u. Thus, u is indeed a weak solution.
Let us show that u satisfies the energy inequality on [0, T ] in the sense of measures.
Taking a smooth nonnegative test function ϕ ∈ C∞ vanishing outside of [0, T ], we write
the energy inequality in the sense of measures for the approximate solutions:
−"QT
Ψnϕ′(t) dx dt ≤ −"QT ∇Φn2ϕ(t) dx dt
+"QT
ϕ(t)(Φxn + unfxn)·∇Φn dx dt +"QT
unfnΦnϕ(t) dxcdt
Convergences (.) ensure that we can pass to the limit in all the terms but for the first
one on the right-hand side. As for the latter, it follows from (.) that √ϕ∇Φn → √ϕ∇Φ
weakly in L2(QT ), whence
"QT
ϕ∇Φ2 dx dt ≤ lim inf
n→∞ "QT
ϕ∇Φn2 dx dt,
and the energy inequality follows.
Let us check (.). The approximate solutions satisfy
ess lim sup
t→+0 W (un(t)) ≤ W (u0
n)
S. KONDRATYEV AND D. VOROTNIKOV
so by virtue of (.) we obtain
It follows from (.) and (.) that
ess sup
t∈(0,ε) W (un(t)) ≤ W (u0
n) + Cε.
W (un) → W (u) weakly* in L∞(0, ε),
W (u0
n) → W (u0),
so we get
ess sup
t∈(0,ε) W (u(t)) ≤ lim inf
n→∞
n→∞W (u0
≤ lim
= W (u0) + Cε.
ess sup
t∈(0,ε) W (un(t))
n) + Cε
Now sending ε → 0 we recover (.).
Let us show that u satisfies the entropy dissipation inequality on [0, T ] in the sense
of measures. Let ϕ ∈ C∞ be a smooth nonnegative test function vanishing outside of
[0, T ]. The approximate solutions satisfy the entropy dissipation inequality in the sense
of measures, so we have
−"QT
Enϕ′(t) dx dt ≤ −"QT
ϕ(t)unf 2
n dx dt −"un>0
ϕ(t)
un −∇Φn + Φxn + unfxn2 dx dt.
Consequently, for any δ > 0 we have
−"QT
Enϕ′(t) dx dt ≤ −"QT
ϕ(t)
max(un, δ)
(unfn)2 dx dt
−"QT
ϕ(t)
max(un, δ)−∇Φn + Φxn + unfxn2 dx dt.
(.)
Observe that
ϕ(t)
max(un, δ) →
ϕ(t)
max(u, δ)
a. e. in QT ,
strongly in any Lp, 1 ≤ p < ∞,
and weakly* in L∞(QT ),
vn := −∇Φn + Φxn + unfxn → −∇Φ + Φx + ufx weakly in L2(Ω)
(.)
(.)
We claim that
"QT
ϕ(t)
max(u, δ)−∇Φ + Φx + ufx2 dx dt
n→∞ "QT
≤ lim inf
ϕ(t)
max(un, δ)−∇Φn + Φxn + unfxn2 dx dt.
(.)
FREE ENERGY FOKKER-PLANCK
Then, taking into account (.), we can pass to the limit in (.) obtaining
−"QT
Eϕ′(t) dx dt ≤ −"QT
ϕ(t)
max(u, δ)
(uf )2 dx dt
max(u, δ)−∇Φ + Φx + ufx2 dx dt.
On the set {(x, t) ∈ QT : u(x, t) = 0} we have ufx = 0 (by virtue of (.)), Φx = 0 and Φ = 0,
whence also ∇Φ = 0 a. e. on this set. Thus, we can write
−"QT
Eϕ′(t) dx dt ≤ −"QT
(uf )2 dx dt
max(u, δ)
ϕ(t)
ϕ(t)
−"QT
−"u>0
ϕ(t)
max(u, δ)−∇Φ + Φx + ufx2 dx dt
To prove the technical claim (.), we use a variant of the Banach-Alaoglu theorem in
Letting δ → 0, by Beppo Levi's theorem we obtain the energy inequality.
varying L2(dµn) spaces:
Lemma .. Let O ⊂ RN be an open set, µn a sequence of finite non-negative Radon measures
narrowly converging to µ, and vn a sequence of vector fields on O. If
then there exists v ∈ L2(O, dµ) such that, up to extraction of some subsequence,
∀ ζ ∈ C∞c (O) :
and
kvnkL2(O,dµn) ≤ C,
lim
n→∞ZO
vn · ζ dµn =ZO
n→∞ kvnkL2(O,dµn).
kvkL2(O,dµ) ≤ lim inf
v · ζ dµ
(.)
(.)
ϕ(t)
The proof of this fact by optimal transport techniques can be found in []; this lemma
also follows from a variant of the Banach-Alaoglu theorem [, Proposition .]. We will
apply this lemma with O = QT , vn from (.), and the sequence of measures dµn(t, x) :=
max(un,δ) dx dt, which converges narrowly to dµ(t, x) = ϕ(t)
max(u,δ) dx dt due to the strong con-
vergence (.). Extracting a subsequence if needed, we see that there is a vector-field
v ∈ L2(O, dµ) verifying (.) and (.). On the other hand, by (.) and (.),
ϕ(t)
vn
max(un, δ) → (−∇Φ + Φx + ufx)
ϕ(t)
max(u, δ)
weakly in L1(QT ). Evoking (.), we find that
ZO
v · ζ dµ =ZO
(−∇Φ + Φx + ufx)· ζ dµ
for all test functions ζ. By density, we conclude that v = −∇Φ + Φx + ufx in L2(O, dµ), and
(.) follows from (.).
S. KONDRATYEV AND D. VOROTNIKOV
Inequality (.) is proved in the same way as (.) given that it holds for the approx-
imate solutions.
Inequalities (.) -- (.) follow from correspondent inequalities for approximate so-
lutions (Lemmas . and .), as we obviously have
(un(t)− un(t))+ → (u(t)− u(t))+) in L1(Ω), ∀t ≥ 0,
(un(t)− uc)± → (u(t)− uc)±
where the approximations un are constructed in the same way as un.
Contraction (.) implies the uniqueness of u.
To obtain the upper bound (.), we define c ≤ 0 by (.) and thus have u0 ≤ uc on Ω,
whence in view of contraction (.),
u(x, t) ≤ uc(x),
(x, t) ∈ Ω × (0,∞).
Recalling the formula (.) for the norm of uc, we obtain the upper bound.
To obtain the lower L1-bound (.), we take uc = m in (.), obtaining
(m− (u(t)− m)−) dx
(m− (u0 − m)−) dx = k min(u(t), m)kL1(Ω),
ku(t)kL1(Ω) ≥ k min(u(t), m)kL1(Ω) =ZΩ
≥ZΩ
as required.
(cid:3)
.. Nonnegative initial data. If initial data (.) is only nonnegative, we approximate
it with positive functions and reuse the proof of Theorem . to establish the existence of
solutions to (.) -- (.) as stated in Theorem . (but not uniqueness, owing to the loss of
contraction).
Proof of Theorem .. Take a decreasing sequence εn → 0 and set
u0
n = u0 + εn.
By Theorem ., there exists a weak solution un of (.) -- (.) with the initial data u0
n. Con-
traction (.) ensures the comparison principle for this sequence of solutions, whence
un+1 ≤ un a. e. in Ω × (0,∞). Consequently, there exists the monotone limit u ∈ L∞(Ω ×
(0,∞)) and moreover, we obviously have the convergences (.). From this moment on,
the proof copies that of Theorem . except that (.) and (.) hold almost everywhere
rather then everywhere.
(cid:3)
We conclude by proving Theorems . and ..
Proof of Theorem .. Let D = {(x, Φ(x, u)) : x ∈ Ω, u > 0} and consider the function Ξ : D →
[0,∞) implicitly defined by the equation
Φ(x, Ξ(x, φ)) = φ.
As Φ is monotonous with respect to its second argument, Ξ is uniquely defined. Clearly,
Ξ is C2.
FREE ENERGY FOKKER-PLANCK
Fix u ∈ U. We claim that there exists a sequence of functions Φn ∈ C(Ω)∩ C∞(Ω) such
that
(x, Φn(x)) ∈ D (x ∈ Ω),
in H1and a. e. in Ω
Φn → Φ(·, u(·))
Indeed, take a sequence {δn}, where δn > 0 and δn → 0, puteΦn(x) = Φ(x, u(x)) + δn, and let
eΦε
n be the mollification of eΦn. Observe that eΦn is strictly positive and so is eΦε
n. It suffices
to show that for any n sufficiently large there exists εn > 0 such that whenever ε < εn, we
have
n(x)) : x ∈ Ω} ⊂ D.
{(x,eΦε
ξfu(x, ξ) dξ ≥ Φ(x, 1)+Z 1
If the second alternative in (.) holds, for every x ∈ Ω we have
Φ(x, u) = Φ(x, 1)+Z 1
as u → +∞. This implies that D = Ω × (0,∞), so (.) obviously holds with any ε.
depend on x if ξ ≥ ξ0 and set
Assume the first alternative in (.). Take ξ0 ≥ kukL∞(Ω) such that f (x, ξ) does not
u
fu(x, ξ) dξ = Φ(x, 1)+f (x, 1)−f (x, u) → +∞
u
(.)
a = −Z ξ0+1
ξ0
ufu(x, u) dx > 0.
We have:
Φ(x, ξ0 + 1)−eΦn(x) = Φ(x, ξ0 + 1)− Φ(x, u(x))− δn
≥ Φ(x, ξ0 + 1)− Φ(x, ξ0)− δn = a− δn.
Thus, for large n we have
Upon mollification,
eΦn(x) ≤ Φ(x, ξ0 + 1)−
n(x) ≤ Φε(x, ξ0 + 1)−
eΦε
a
2
.
a
2
.
For a fixed n, the function Φ(·, ξ0 + 1) is continuous on Ω, so the mollifications Φε(·, ξ0 + 1)
converge to it uniformly on Ω as ε → 0. Consequently,
(x,eΦε
n(x)) ∈ {(x, φ) ∈ Ω × (0,∞) : φ ≤ Φ(x, ξ0 + 1)} ⊂ D
(.)
for all x ∈ Ω, proving (.).
Taking a sequence {Φn} as above, we can set un(x) = Ξ(x, Φn(x)), so that Φn(x) = Φ(x, un(x)).
Clearly, un ∈ C2(Ω) and un > 0. Further, the sequence {un} is bounded in L∞(Ω) because
so is {Φn}, and due to the continuity of Ξ we have
un → u a. e. in Ω.
S. KONDRATYEV AND D. VOROTNIKOV
As a consequence, for fn = f (x, un(x)) and En = E(x, un(x)) we have
un → u
unfn → uf
unfxn → ufx
Φxn → Φx
En → E
a. e. in Ω
and in any Lp(Ω), 1 ≤ p < ∞ ,
(.)
where we write f for f (·, u(·)), etc. In particular, there is no loss of generality in assuming
a lower bound
kunkL1(Ω) ≥ c :=
1
2
inf
u∈U kukL1(Ω) > 0
(positivity by virtue of (.)), where c is obviously independent not only of un but of u
as well.
Define
By Theorem ., there exist a function
eU = {w ∈ C1(Ω) : w > 0,kwkL1(Ω) ≥ c}.
v(ξ) =
ξ 2
max(ξ, σ)
,
where σ > 0, and a constant C > 0 such that
ZΩ
E(x, w(x)) dx ≤ CZΩ
In particular, as un ∈ eU, we see that
v(w(x))(f (x, w(x)) +∇f (x, w(x))2) dx
(w ∈ eU).
ZΩ
En dx ≤ CZΩ
vn(fn +∇fn2) dx,
(.)
where vn = v(un(x)).
Let us check that we can pass to the limit in (.). First, it follows from (.) that
ZΩ
En dx →ZΩ
E dx.
Next, note that we clearly have
1
max(un, σ) →
1
max(u, σ)
a. e. in Ω and weakly* in L∞(Ω)
and thus, again using (.), we obtain
ZΩ
vnf 2
n dx =ZΩ
1
max(un, σ)
(unfn)2 dx →ZΩ
1
max(u, σ)
(uf )2 dx.
FREE ENERGY FOKKER-PLANCK
Finally, as un is smooth and positive, we can write
ZΩ
vn∇fn2 dx =ZΩ
→ZΩ
1
max(un, σ)−∇Φn + Φxn + unfxn2 dx
max(u, σ)−∇Φ + Φx + ufx2 dx.
1
On the set [u = 0] we have ufx = 0 by (.), Φx = 0, and Φ = 0, the last equality implying
∇Φ = 0 a. e. on [u = 0]. Thus, we can write
vn∇fn2 dx →Z[u>0]
max(u, σ)−∇Φ + Φx + ufx2 dx.
ZΩ
1
To sum up, we have
ZΩ
E dx ≤ C ZΩ
u2
max(u, σ)
f 2 dx +Z[u>0]
1
max(u, σ)−∇Φ + Φx + ufx2 dx! ,
which is even stronger than (.).
(cid:3)
Proof of Theorem .. Let U ⊂ L∞+ be the set of functions such that for any v ∈ U, we
have Φ(·, v(·)) ∈ H1(Ω) and kvkL1(Ω) ≥ c. By Theorem . we have the entropy-entropy
production inequality (.) for U.
Let u be a weak solution of (.) -- (.) with the initial data satisfying (.). It follows
from the lower L1-bound (.) that u(t) ∈ U for a. a. t > 0. Combining the entropy
dissipation and entropy-entropy production inequalities, we obtain
∂tE(u(t)) ≤ −C−1
U E(u(t))
a. a. t > 0.
Letting e(t) = E(u(t))eC−1
U t, we see that ∂te(t) ≤ 0 in the sense of measures, whence e a. e.
coincides with a nonincreasing function. Moreover,
t→0
ess sup
t>0
t→0
E(u(t))eC−1
e(t) = ess lim sup
e(t) = ess lim sup
U t ≤ E(u0)
by virtue of (.), so e(t) ≤ E(u0) for a. a. t > 0, yielding (.) with γ = C−1
U .
Acknowledgment. The idea of this paper originated from conversations of the second
author with Goro Akagi and Yann Brenier during a stay at ESI in Vienna. He would like
to thank Goro Akagi and Yann Brenier for the inspiring discussions and correspondence,
Ulisse Stefanelli for the invitation to the thematic program Nonlinear Flows at ESI, and
ESI for hospitality. The research was partially supported by the Portuguese Government
through FCT/MCTES and by the ERDF through PT (projects UID/MAT//,
PTDC/MAT-PUR// and TUBITAK//).
(cid:3)
Conflict of interest: none.
S. KONDRATYEV AND D. VOROTNIKOV
References
[] H. W. Alt and S. Luckhaus. Quasilinear elliptic-parabolic differential equations. Math. Z., (): --
, .
[] L. Ambrosio, N. Gigli, and G. Savar´e. Gradient Flows: in Metric Spaces and in the Space of Probability
Measures. Basel: Birkhauser Basel, .
[] V. Barbu. Generalized solutions to nonlinear Fokker-Planck equations. J. Differential Equations,
(): -- , .
[] M. Bertsch and D. Hilhorst. A density dependent diffusion equation in population dynamics: stabiliza-
tion to equilibrium. SIAM J. Math. Anal., (): -- , .
[] T. Bodineau, J. Lebowitz, C. Mouhot, and C. Villani. Lyapunov functionals for boundary-driven nonlin-
ear drift-diffusion equations. Nonlinearity, (): -- , .
[] R. S. Cantrell, C. Cosner, Y. Lou, and C. Xie. Random dispersal versus fitness-dependent dispersal. J.
Differential Equations, (): -- , .
[] J. A. Carrillo, A. J ungel, P. A. Markowich, G. Toscani, and A. Unterreiter. Entropy dissipation methods
for degenerate parabolic problems and generalized Sobolev inequalities. Monatsh. Math., (): -- ,
.
[] L. Chizat, G. Peyr´e, B. Schmitzer, and F.-X. Vialard. An interpolating distance between optimal transport
and Fisher -- Rao metrics. Foundations of Computational Mathematics, (): -- , .
[] L. Chizat, G. Peyr´e, B. Schmitzer, and F.-X. Vialard. Unbalanced optimal transport: Dynamic and Kan-
torovich formulations. Journal of Functional Analysis, (): -- , .
[] C. Cosner. A dynamic model for the ideal-free distribution as a partial differential equation. Theoretical
Population Biology, (): -- , .
[] C. Cosner. Beyond diffusion: conditional dispersal in ecological models. In J. Mallet-Paret et al., editor,
Infinite Dimensional Dynamical Systems, pages -- . Springer, .
[] C. Cosner and M. Winkler. Well-posedness and qualitative properties of a dynamical model for the ideal
free distribution. Journal of mathematical biology, (-): -- , .
[] J. Filo and J. Kacur. Local existence of general nonlinear parabolic systems. Nonlinear Anal., (): --
, .
[] T. D. Frank. Asymptotic properties of nonlinear diffusion, nonlinear drift-diffusion, and nonlinear
reaction-diffusion equations. Ann. Phys., (-): -- , .
[] T. D. Frank. Nonlinear Fokker-Planck equations. Springer Series in Synergetics. Springer-Verlag, Berlin,
. Fundamentals and applications.
[] S. D. Fretwell. Populations in a seasonal environment. Princeton University Press, .
[] S. D. Fretwell and H. L. Lucas. On territorial behavior and other factors influencing habitat distribution
in birds I. Theoretical development. Acta Biotheoretica, (): -- , .
[] T. Gallouet, M. Laborde, and L. Monsaingeon. An unbalanced optimal transport splitting scheme for
general advection-reaction-diffusion problems. arXiv:., .
[] T. O. Gallouet and L. Monsaingeon. A JKO splitting scheme for Kantorovich-Fisher-Rao gradient flows.
SIAM J. Math. Anal., (): -- , .
[] R. Jordan, D. Kinderlehrer, and F. Otto. The variational formulation of the Fokker -- Planck equation.
SIAM journal on mathematical analysis, (): -- , .
[] A. J ungel. Entropy methods for diffusive partial differential equations. SpringerBriefs in Mathematics.
Springer, [Cham], .
[] J. Kacur. On a solution of degenerate elliptic-parabolic systems in Orlicz-Sobolev spaces. I. Math. Z.,
(): -- , .
[] J. Kacur. On a solution of degenerate elliptic-parabolic systems in Orlicz-Sobolev spaces. II. Math. Z.,
(): -- , .
[] S. Kondratyev, L. Monsaingeon, and D. Vorotnikov. A fitness-driven cross-diffusion system from popu-
lation dynamics as a gradient flow. J. Differential Equations, (): -- , .
FREE ENERGY FOKKER-PLANCK
[] S. Kondratyev, L. Monsaingeon, and D. Vorotnikov. A new optimal transport distance on the space of
finite Radon measures. Adv. Differential Equations, (-): -- , .
[] S. Kondratyev, L. Monsaingeon, and D. Vorotnikov. A new multicomponent Poincar´e-Beckner inequal-
ity. J. Funct. Anal., (): -- , .
[] S. Kondratyev and D. Vorotnikov. Convex Sobolev inequalities related to unbalanced optimal transport.
arXiv e-prints, page arXiv:., Apr .
[] S. Kondratyev and D. Vorotnikov. Spherical Hellinger-Kantorovich gradient flows. SIAM J. Math. Anal.,
(): -- , .
[] M. Liero, A. Mielke, and G. Savar´e. Optimal transport in competition with reaction: the Hellinger-
Kantorovich distance and geodesic curves. SIAM J. Math. Anal., (): -- , .
[] M. Liero, A. Mielke, and G. Savar´e. Optimal entropy-transport problems and a new Hellinger --
Kantorovich distance between positive measures. Inventiones mathematicae, (): -- , .
[] Y. Lou, Y. Tao, and M. Winkler. Approaching the ideal free distribution in two-species competition
models with fitness-dependent dispersal. SIAM J. Math. Anal., (): -- , .
[] A. D. MacCall. Dynamic geography of marine fish populations. Washington Sea Grant Program Seattle,
.
[] F. Maggi. Sets of Finite Perimeter and Geometric Variational Problems: An Introduction to Geometric Measure
Theory. Cambridge Studies in Advanced Mathematics. Cambridge University Press, .
[] V. G. Maz'ja. Sobolev spaces. Springer Series in Soviet Mathematics. Springer-Verlag, Berlin, . Trans-
lated from the Russian by T. O. Shaposhnikova.
[] F. Otto. The geometry of dissipative evolution equations: the porous medium equation. Comm. Partial
Differential Equations, (-): -- , .
[] F. Santambrogio. {Euclidean, metric, and Wasserstein} gradient flows: an overview. Bull. Math. Sci.,
[] W. Shi and D. Vorotnikov. The gradient flow of the potential energy on the space of arcs. Calculus of
(): -- , .
Variations and Partial Differential Equations, to appear, .
[] C. Tsallis. Introduction to nonextensive statistical mechanics. Springer, .
[] J. L. V´azquez. Failure of the strong maximum principle in nonlinear diffusion. Existence of needles.
Comm. Partial Differential Equations, (-): -- , .
[] J. L. V´azquez. The porous medium equation. Oxford Mathematical Monographs. The Clarendon Press,
Oxford University Press, Oxford, . Mathematical theory.
[] C. Villani. Topics in optimal transportation. American Mathematical Soc., .
[] C. Villani. Optimal transport: old and new. Springer Science & Business Media, .
[] Q. Xu, A. Belmonte, R. deForest, C. Liu, and Z. Tan. Strong solutions and instability for the fitness
gradient system in evolutionary games between two populations. J. Differential Equations, (): --
, .
(S. Kondratyev) CMUC, Department of Mathematics, University of Coimbra, - Coimbra, Por-
tugal
E-mail address: [email protected]
(D. Vorotnikov) CMUC, Department of Mathematics, University of Coimbra, - Coimbra, Por-
tugal
E-mail address: [email protected]
|
0811.4432 | 4 | 0811 | 2010-03-29T19:33:34 | Translation-finite sets, and weakly compact derivations from $\lp{1}(\Z_+)$ to its dual | [
"math.FA",
"math.CO"
] | We characterize those derivations from the convolution algebra $\ell^1({\mathbb Z}_+)$ to its dual which are weakly compact. In particular, we provide examples which are weakly compact but not compact. The characterization is combinatorial, in terms of "translation-finite" subsets of ${\mathbb Z}_+$, and we investigate how this notion relates to other notions of "smallness" for infinite subsets of ${\mathbb Z}_+$. In particular, we show that a set of strictly positive Banach density cannot be translation-finite; the proof has a Ramsey-theoretic flavour. | math.FA | math |
Translation-finite sets, and weakly compact derivations
from ℓ1(Z+) to its dual
Y. Choi, M. J. Heath∗
12th October 2009†
Abstract
We characterize those derivations from the convolution algebra ℓ1(Z+) to its
dual which are weakly compact, providing explicit examples which are not compact.
The characterization is combinatorial, in terms of "translation-finite" subsets of
Z+, and we investigate how this notion relates to other notions of "smallness" for
infinite subsets of Z+. In particular, we prove that a set of strictly positive Banach
density cannot be translation-finite; the proof has a Ramsey-theoretic flavour.
1
Introduction
The problem of determining the weakly compact or compact homomorphisms between
various Banach algebras has been much studied; the study of weakly compact or com-
pact derivations, less so. In certain cases, the geometrical properties of the underlying
Banach space play an important role. For instance, by a result of Morris [9], every
bounded derivation from the disc algebra A(D) to its dual is automatically weakly
compact. (It had already been shown in [2] that every bounded operator from A(D) to
A(D)∗ is automatically 2-summing, hence weakly compact; but the proof is significantly
harder than that of the weaker result in [9].)
In this article, we investigate the weak compactness or otherwise of derivations
from the convolution algebra ℓ1(Z+) to its dual. Unlike the case of A(D), the space of
derivations is easily parametrized: every bounded derivation from ℓ1(Z+) to its dual is
of the form
Dψ(δ0) = 0 and Dψ(δj)(δk) =
j
j + k
ψj+k
(j ∈ N, k ∈ Z+)
(1.1)
for some ψ ∈ ℓ∞(N). It was shown in the second author's thesis [7] that Dψ is compact
if and only if ψ ∈ c0, and that there exist ψ for which Dψ is not weakly compact.
The primary purpose of the present note is to characterize those ψ for which Dψ
is weakly compact (see Theorem 2.6 below). In particular, we show that there exist a
plethora of ψ for which Dψ is weakly compact but not compact. Our criterion is combi-
natorial and uses the notion, apparently due to Ruppert, of translation-finite subsets of
∗The 2nd author is supported by post-doctoral grant SFRH/BPD/40762/2007 from FCT (Portugal).
†MSC2000: Primary 43A20, 43A46, 47B07
1
a semigroup. A secondary purpose is to construct various examples of translation-finite
and non-translation-finite subsets of Z+, to clarify the connections or absence thereof
with other combinatorial notions of "smallness".
An example
We first resolve a question from [7], by giving a very simple example of a Dψ that is
non-compact but is weakly compact.
Proposition 1.1 Let ψ be the indicator function of {2n : n ∈ N} ⊂ N. Then Dψ is
non-compact, and the range of Dψ is contained in ℓ1(Z+).
In particular, since ℓ1(Z+) ⊂ ℓp(Z+) for every 1 < p < ∞, Dψ factors through a
reflexive Banach space and is therefore weakly compact.
Proof . Since ψ /∈ c0, we know by [7, Theorem 5.7.3] that Dψ is non-compact.
We have Dψ(δ0) = 0. For each j ∈ N, let Nj = min(n ∈ N : 2n ≥ j); then
kDψ(δj)k1 = Xk≥0
j
(j + k)
(cid:12)(cid:12)(cid:12)(cid:12)
= Xn≥Nj
j
2n =
j
2Nj −1 ≤ 2 .
ψj+k(cid:12)(cid:12)(cid:12)(cid:12)
By linearity and continuity we deduce that kDψ(a)k1 ≤ 2kak1 for all a ∈ ℓ1(Z+). The
last assertion now follows, by standard results on weak compactness of operators. (cid:3)
Remark 1.2 Since Dψ factors through the inclusion map ℓ1(Z+) → c0(Z+), which is
known to be 1-summing, it too is 1-summing.
This last remark raises the natural question:
is every weakly compact derivation
from ℓ1(Z+) to its dual automatically p-summing for some p < ∞? The answer, unsur-
prisingly, is negative: we have deferred the relevant counterexample to an appendix.
2 Characterizing weakly compact derivations
We need only the basic results on weak compactness in Banach spaces, as can be found
in standard references such as [8].
Recall that if X is a closed subspace of a Banach space Y , and K ⊆ X, then K is
weakly compact as a subset of X if and only if it is weakly compact as a subset of Y .
Since (by Equation (1.1)) our derivations Dψ take values in c0(Z+), we may therefore
work with the weak topology of c0(Z+) rather than that of ℓ∞(Z+).
Moreover, we can reduce the verification of weak compactness to that of sequential
pointwise compactness. This is done through some simple lemmas, which we give below.
Lemma 2.1 Let (yi) be a bounded net in c0(Z+), and let y ∈ c0(Z+). Then (yi)
converges weakly to y if and only if it converges pointwise to y.
The proof is straightforward and we omit the details.
2
Lemma 2.2 Let T : ℓ1(Z+) → c0(Z+) be a bounded linear map. Then the following
are equivalent:
(i) T is weakly compact;
(ii) every subsequence of (T (δn))n∈N has a further subsequence which converges point-
wise to some y ∈ c0(Z+).
Proof . Let B denote the closed unit ball of ℓ1(Z+), let E = {T (δn) : n ∈ N}, and let
τ denote the topology of pointwise convergence in c0. Note that the restriction of τ to
bounded subsets of c0 is a metrizable topology.
If (i) holds, then by (the trivial half of) Lemma 2.1, the bounded set T (B) is τ -
precompact, and hence sequentially τ -precompact. Thus (ii) holds.
Conversely, suppose that (ii) holds, i.e. that E is sequentially τ -precompact. Then
(again by metrizability) we know that E is τ -compact, and hence by Lemma 2.1 it is
weakly precompact. Therefore, by Krein's theorem (as it appears in Bourbaki, see
[1, §IV.5]), the closed balanced convex hull of E is weakly compact. Since this hull is
T (B), T is weakly compact.
(cid:3)
The following notation will be used frequently. Given a subset S ⊆ Z+ and n ∈ Z+,
we denote by S − n the set {t ∈ Z+ : t + n ∈ S}.
We need the following definition, due to Ruppert [11] in a more general setting.
Definition 2.3 ([11]) Let S ⊆ Z+. We say that S is translation-finite (TF for short)
if, for every sequence n1 < n2 < . . . in Z+, there exists k such that
k
(S − ni)
\i=1
is finite or empty.
(2.1)
(In the later paper [3], TF-sets are called "RW sets"; we believe that for our purposes
the older terminology of Ruppert is more suggestive and apposite.)
TF-sets were introduced by Ruppert in the investigation of weakly almost periodic
subsets of semigroups. Recall that a bounded function f on a semigroup S is said to
be weakly almost periodic if the set of translates {s · f : s ∈ S} ∪ {f · s : s ∈ S} is
weakly precompact in ℓ∞(S). Specializing to the case where the semigroup in question
is Z+, one of Ruppert's results can be stated as follows.
Theorem 2.4 ([11]) Let S ⊆ Z+. Then S is TF if and only if all bounded functions
S → C are WAP as elements of ℓ∞(Z+).
In particular, if S is a TF-set then the
indicator function of S belongs to WAP(Z+).
It is sometimes convenient to use an alternative phrasing of the original definition
(see [3] for instance).
Lemma 2.5 Let S ⊆ Z+. Then S is non-TF if and only if there are strictly increasing
sequences (an)n≥1, (bn)n≥1 ⊂ Z+ such that {a1, . . . , an} ⊆ S − bn for all n.
3
Proof . Suppose that there exist sequences (an), (bn) as described. Then for every
n ≥ m ∈ N we have {a1, . . . , am} ⊆ S − bn. Hence
{bn : n ≥ m} ⊆
m
(S − ak)
\k=1
where the set on the left hand side is infinite, for all m. Thus S is non-TF.
that Tk
Conversely, suppose S is non-TF: then there is a sequence a1 < a2 < . . . in Z+ such
j=1(S − aj) is infinite for all k ∈ N. Let b1 ∈ S − a1. We inductively construct
k=1(S − ak) is
a sequence (bn) as follows:
infinite it contains some bn+1 > bn. By construction the sequences (an) and (bn) are
strictly increasing, and {a1, . . . , an} ⊆ S − bn for all n.
(cid:3)
if we have already chosen bn, then since Tn+1
Our main result, which characterizes weak compactness of Dψ in terms of ψ, is as
follows.
Theorem 2.6 Let ψ ∈ ℓ∞(N). Then Dψ is weakly compact if and only if, for all ε > 0,
the set Sε := {n ∈ N : ψn > ε} is TF.
It is not clear to the authors if one can deduce Theorem 2.6 in a "soft" way from
Ruppert's characterization (Theorem 2.4). Instead, we give a direct argument. The
proof naturally breaks into two parts, both of which can be carried out in some gener-
ality.
Given ψ ∈ ℓ∞(Z+) and M ∈ ℓ∞(Z+ × Z+), define T M
ψ : ℓ1(Z+) → ℓ∞(Z+) by
T M
ψ (δj )(δk) = Mjkψj+k
(j, k ∈ Z+).
(2.2)
In particular, if we take ψ ∈ ℓ∞(N), identified with (0, ψ1, ψ2, . . . ) ∈ ℓ∞(Z+), and
take Mjk to be 0 for j = k = 0 and to be j/(j + k) otherwise, then T M
ψ ≡ Dψ.
Proposition 2.7 Let ψ ∈ ℓ∞(Z+) be such that Sε is TF for all ε > 0.
ℓ∞(Z+ × Z+) is such that limk→∞ Mjk = 0 for all j, then T M
ψ is weakly compact.
If M ∈
Proof . To ease notation we write Tψ for T M
condition on M implies that Tψ takes values in c0(Z+).
ψ throughout this proof. Note that the
Define ψε ∈ ℓ∞(Z+) as follows: set (ψε)n to be ψn if n ∈ Sε and 0 otherwise. Then
ψε is supported on Sε and Tψε → Tψ as ε → 0. It therefore suffices to prove that if ψ
has TF support, then Tψ is weakly compact.
Let ψ ∈ ℓ∞ be supported on a TF set S. Let (jn)n≥1 ⊂ Z+ be a strictly increasing
sequence and set (j0,n) = (jn), k0 = 0. For each i ≥ 1 we specify an integer ki ∈ Z+ and
a sequence (ji,n)n≥1 ⊂ Z+ recursively, as follows. If there exists k ∈ Z+ \ {k0, . . . , ki−1}
such that ji−1,n + k ∈ S for infinitely many n ∈ N, let ki be some such k. Otherwise,
let ki = ki−1. Let (ji,n)n≥1 be the enumeration of the set
{ji−1,n : n ∈ N, ji−1,n + ki ∈ S}
with ji,n < ji,n+1 for each n ∈ N. Then, by induction on i, (ji,n)n is a subsequence of
(jn)n and, for each l ∈ {1, . . . , i}, and each n ∈ N, we have ji,n + kl ∈ S.
4
In particular, for each i ∈ N, {ji,i + k1, . . . , ji,i + ki} ⊂ S. Hence, by our assumption
that S is TF, the set {ki : i ∈ Z+} is finite. Let i0 be the smallest i for which ki = ki+1:
then for each k ∈ Z+ \ {k0, . . . , ki0 }, there are only finitely many n ∈ N such that
ji0,n + k ∈ S.
Let K = {k0, . . . , ki0}. By the Heine-Borel theorem, there exists a subsequence
(jn(m))m of (ji0,n)n such that, for each k ∈ K, limm→∞ Tψ(δjn(m))(δk) exists. Moreover,
by the previous paragraph, for each k ∈ Z+ \ K there exist at most finitely many m
such that jn(m) + k ∈ S; hence there exists m(k) such that
Tψ(δjn(m))(δk) = Mjn(m),kψ(jn(m) + k) = 0
for all m ≥ m(k).
Thus Tψ(δjn(m)) converges pointwise to some function supported on K, and the result
follows by Lemma 2.2.
(cid:3)
Proposition 2.8 Let M ∈ ℓ∞(Z+ × Z+) satisfy
lim
k→∞
Mjk = 0 for all j, and inf
k
lim inf
j→∞
Mjk = η > 0 .
(2.3)
Let ψ ∈ ℓ∞, and suppose that T M
ψ is weakly compact. Then Sε is TF for all ε > 0.
Proof . We first note that (2.3) implies that T M
ψ has range contained in c0(Z+). Suppose
the result is false: then there exists ε > 0 such that Sε is non-TF. Hence, by Lemma 2.5,
there exist strictly increasing sequences (an), (bn) ⊂ Z+ such that
{a1 + bn, . . . , an + bn} ⊂ Sε
for all n.
Now
T M
ψ (δbn )(δam ) = Mbn,amψam+bn ≥ Mbn,amε
for all m ≤ n;
so, by the hypothesis (2.3), we have
inf
m
lim inf
n
T M
ψ (δbn)(δam ) ≥ ηε .
(2.4)
Since T M
ψ is weakly compact, by Lemma 2.2 the sequence (Tψ(δbn)) has a sub-
sequence that converges pointwise to some φ ∈ c0(Z+). But by (2.4) we must have
inf m φam ≥ ηε, so that φ /∈ c0(Z+). Hence we have a contradiction and the proof is
complete.
(cid:3)
Proof of Theorem 2.6. Sufficiency of the stated condition follows from Proposition 2.7;
necessity, from Proposition 2.8, once we observe that limk j/(j + k) = 0 for all j, and
limj j/(j + k) = 1 for all k.
(cid:3)
5
Remark 2.9 The set S = {2n : n ∈ N} is TF. In fact, it is not hard to show it has
the following stronger property:
for every n ∈ N, the set S ∩ (S − n) is finite or empty.
(†)
We therefore obtain another proof that the derivation constructed in Proposition 1.1
is weakly compact.
Subsets of Z+ satisfying the condition (†) seem not to have an agreed name. They
were called T -sets in work of Ramirez [10], and for ease of reference we shall use his
terminology.
3 Comparing the TF-property with other notions of size
Let S ⊂ Z+. For n ∈ N we define fS(n) to be the nth member of S.
Proposition 3.1 Let S ⊂ Z+. Then there exists a non-TF R ⊂ Z+ with fR(n) >
fS(n) for all n.
Proof . For n ∈ Z+, let tn = 1
2 n(n + 1) be the nth triangular number, so that tn =
tn−1 + n for all n ∈ N. Each n ∈ N has a unique representation as n = tk−1 + j
where k ≥ 1 and 1 ≤ j ≤ k. Enumerate the elements of S in increasing order as
s1 < s2 < s3 < . . ., and define a sequence (rn)n≥1 by
rtk−1+j = stk + j
(1 ≤ j ≤ k)
as indicated by the following diagram:
r1 = s1 + 1
r2 = s3 + 1 r3 = s3 + 2
r4 = s6 + 1 r5 = s6 + 2 r6 = s6 + 3
...
Since the sequence (sn) is strictly increasing,
stk ≥ stk−1 + (tk − tk−1) = stk−1 + k
(k ∈ N),
and so the sequence (rn) is strictly increasing. Put R = {rn : n ∈ N}: then clearly
: k ≥ m} for
(cid:3)
fR(n) = rn > sn = fS(n) for all n. Finally, since Tm
all m, R is not a TF-set.
j=1(R − j) ⊇ {st(k)
On the other hand, we can find T-sets S such that fS grows at a "nearly linear"
rate.
Proposition 3.2 Let g : N → Z+ be any function such that g(n)/n → ∞. Then there
is a strictly increasing sequence a1 < a2 < . . . in Z+, such that {an : n ∈ N} is a T-set,
while an < g(n) for all but finitely many n.
6
Proof . Let N ∈ N and set kN to be the smallest natural number such that g(n) > N n
for all n > kN . We now construct our sequence (an) recursively. Set a0 = 0 and
suppose that a0, . . . , an have been defined:
if an < k1 set an+1 = an + 1; otherwise,
if kN ≤ an < kN +1 for some N ∈ N, set an+1 = an + N . Thus, the elements of
[kN , kN +1] ∩ {an : n ∈ N} are in arithmetic progression with common difference N .
A simple induction gives that if kN ≤ n < kN +1, then an ≤ N n. Since for n ≥ kN
we also have that g(n) > N n, it follows that, for all n ≥ k1, an < g(n).
Finally, let i, j ∈ N. If ai − j ∈ {an : n ∈ N} it follows that ai < kj+1 + j. Thus,
(cid:3)
{an : n ∈ N} is a T-set.
Remark 3.3 Given that infinite arithmetic progressions are the most obvious exam-
ples of non-TF sets, it may be worth noting that if g(n)/n2 → 0, the T-set constructed
in the proof of Proposition 3.2 contains arbitrarily long arithmetic progressions.
The previous two results indicate that the growth of a subset in Z+ tells us nothing,
on its own, about whether or not it is TF. The main result of this section shows that,
nevertheless, certain kinds of density property are enough to force a set to be non-TF.
First we need some definitions.
Definition 3.4 Let S ⊂ Z+. The upper Banach density1 of S, denoted by Bd(S), is
Bd(S) := lim
d→∞
max
n
d−1S ∩ {n + 1, . . . , n + d}
(The limit always exists, by a subadditivity argument.)
For example, the set R constructed in the proof of Proposition 3.1 satisfies R ⊇
{st(m) + 1, . . . , st(m) + m} for all m, and so has a Banach density of 1.
Proposition 3.5 Let S ⊂ Z+ and suppose Bd(S) > 0. Then S is not TF.
The converse clearly fails: for example, the set S = {2i + j2 : i ∈ N, j ∈ {0, . . . , i}} is
not TF, but has Banach density zero.
The proof of Proposition 3.5 builds on some preliminary lemmas, which in turn
require some notation. Fix once and for all a set S ⊂ Z+ with strictly positive Banach
density, and choose ε ∈ (0, 1) such that Bd(S) > ε.
For shorthand, we say that a subset X ⊆ Z+ is recurrent in S if there are infinitely
many n ∈ N such that X + n ⊂ S. For each d ∈ N, let
Vd = {X ⊂ N : X is recurrent in S and d ≥ X ≥ dε}.
Lemma 3.6 For every d ∈ N, Vd is non-empty.
1What we call "Banach density" is also referred to as upper Banach density, and is in older sources
given a slightly different but equivalent definition. Some background and remarks on the literature can
be found in [6, §1], for example.
7
Proof . The key step is to prove that the set {i ∈ N :
S ∩ {i + 1, . . . , i + d} ≥ dε}
is infinite, which we do by contradiction. For, suppose it is finite, with cardinality N ,
say: then for any j, n ∈ N we have
(jd)−1S ∩ {n + 1, . . . , n + (jd)} = j−1
j−1
Xm=0
d−1S ∩ {n + md + 1, . . . , n + (m + 1)d}
≤ j−1(N + (j − N )ε) ,
so that
Bd(S) = lim
j
(jd)−1 sup
n
S ∩ {n + 1, . . . , n + (jd)} ≤ lim sup
j−1(N + (j − N )ε) = ε ,
j
which contradicts our original choice of ε.
It follows that there exists a strictly increasing sequence i1 < i2 < . . . in N, such
that S ∩ {in + 1, . . . , in + d} ≥ dε for all n. Since there are at most finitely many
subsets of {1, . . . , d}, by passing to a subsequence we may assume that the sequence of
sets ((S − in) ∩ {1, . . . , d})n≥1 is constant, with value X say. Clearly X ∈ Vd, which
completes the proof.
(cid:3)
Lemma 3.7 There exists a sequence 1 = d1 < d2 < . . . in N such that, for every j ∈ N
and any X ∈ Vdj+1, there exists Y ∈ Vdj such that Y ⊆ X and max(Y ) < max(X).
Proof . Put d1 = 1 and choose N1 ∈ N such that N1 > ε−1. We then inductively
construct our sequence (dn) as follows: if we have already defined dj for some j ∈ N,
let aj be the largest non-negative integer such that aj < djε. Then choose Nj ∈ N,
Nj ≥ 2, large enough that
1
Nj (cid:20)(Nj − 1)
aj
dj
+ 1(cid:21) < ε ,
(3.1)
and set dj+1 = Njdj.
To show that this sequence has the required properties, let j ∈ N. Given X ∈ Vdj+1,
put x0 = min(X), and for m = 0, 1, . . . , Nj − 1 put
Ym = X ∩ {x0 + mdj, . . . , x0 + (m + 1)dj − 1}.
Since X ≤ dj+1 and min(X) = x0, the sets Y0, . . . , YNj −1 form a partition of X. Since
X is recurrent in S, so is each of the subsets Ym, and by construction Ym ≤ dj for
all m.
We claim that there exist m(1) < m(2) ∈ {0, 1, . . . , Nj − 1} such that Ym(1) and
Ym(2) have cardinality ≥ djε. If this is the case then Ym(1) ∈ Vdj and max(Ym(1)) <
min(Ym(2)) ≤ max(X), so that we may take Y = Ym(1) in the statement of the lemma.
Suppose the claim is false. Then at least Nj − 1 of the sets Y0, . . . , YNj −1 have
cardinality < djε, and hence (by the definition of aj) at least Nj − 1 of these sets have
8
cardinality ≤ aj. Now since X ⊆ {x0, . . . , x0 + dj+1 − 1}, we have X = Y0 ⊔ . . . ⊔ YNj−1,
and so
Nj −1
Njdjε = dj+1ε ≤ X =
Xm=0
Ym ≤ dj + (Nj − 1)aj.
On dividing through by Njdj, we obtain a contradiction with (3.1), and our claim is
proved.
(cid:3)
The final ingredient in our proof of Proposition 3.5 is purely combinatorial: it is a
version of 'Konig's infinity lemma', which we isolate and state for convenience. We
shall paraphrase the formulation given in [5, Lemma 8.1.2], and refer the reader to that
text for the proof.
Lemma 3.8 Let G be a graph on a countably infinite vertex set V , and let V = `j≥1 Vj
be a partition of V into mutually disjoint, non-empty finite subsets. Suppose that for
each j ≥ 1, every v ∈ Vj+1 has a neighbour in Vj . Then there exists a sequence (vn)n≥1,
with vn ∈ Vn for each n, such that vn+1 is a neighbour of vn.
Proof of Proposition 3.5. Let (dj) be the sequence from Lemma 3.7. For each j, let
Vj be the set of all subsets of {1, . . . , dj } which are also members of Vdj . The proof of
Lemma 3.6 shows that Vj is non-empty, and clearly it is a finite set.
Regard `j≥1 Vj as the vertex set for a graph, whose edges are defined by the
following rule: for each j ∈ N and Y ∈ Vj, X ∈ Vj+1, join X to Y with an edge if and
only if there exists m ∈ Z+ with Y + m ⊆ X and max(Y ) + m < max(X). Then by
Lemma 3.7, every element of Vj+1 has a neighbour in Vj. Hence, by Lemma 3.8, there
exists a sequence (Yj)j≥1 of finite subsets of N, and a sequence (mj) ⊂ Z+, such that
(i) Yj ∈ Vj for all j;
(ii) Yj + mj ⊆ Yj+1 and max(Yj) + mj < max(Yj+1) for all j.
Now put X1 = Y1 and put Xj+1 = Yj+1 − (mj + · · · + m1) ⊂ N for j ≥ 1. An
easy induction using both parts of (ii) above shows that Xj ( Xj+1 for all j. Since
each Yi is recurrent in S, so is each Xi, and hence there exist infinitely many n such
that Xi + n ⊂ S. We may therefore inductively construct n1 < n2 < . . . such that
Xi + ni ⊂ S for all i.
Pick c1 ∈ X1 and for each i pick ci+1 ∈ Xi+1 \ Xi; then for all 1 ≤ i ≤ j we have
: i ∈ N} is infinite, by Lemma 2.5 S is
(cid:3)
ci + nj ∈ Xj + nj ⊂ S; and since the set {ci
not TF.
4 Closing thoughts
We finish with some remarks and questions that this work raises. Here and in the
appendix, it will be convenient to abuse notation as follows: given S ⊆ N, we write DS
for the derivation DχS , where χS is the indicator function of S. For example, with this
notation DN ≡ D1.
9
Combinatorics of TF subsets of Z+
We have been unable to find much in the literature on the combinatorial properties of
TF subsets of Z+. Here are some elementary facts.
• Let k ∈ N; then S is TF if and only if S + k is.
• Finite unions of T-sets are TF.
• The set of odd numbers is non-TF, as is the set of even numbers. In particular,
the complement of a non-TF set can be non-TF.
• Subsets of TF sets are TF. (Immediate from the definition.) In particular, the
intersection of two TF sets is TF.
• If S and T are TF then so is S ∪ T .
The last of these observations follows immediately if we grant ourselves Ruppert's
result (Theorem 2.4 above). It also follows from our Theorem 2.6: for if S and T are
TF subsets of Z+, then since S + 1, (S ∩ T ) + 1 and T + 1 are also TF, the derivations
DS+1, DT +1 and D(S∩T )+1 are all weakly compact; whence
D(S∪T )+1 = DS+1 − D(S∩T )+1 + DT +1
is also weakly compact, so that by the other direction of Theorem 2.6, (S ∪ T ) + 1 and
hence S ∪ T are TF. It also seems worth giving a direct, combinatorial proof, which to
our knowledge is not spelled out in the existing literature (cf. [11, Remark 18]).
Proof . Let A1, A2 be subsets of Z+ and suppose that A1 ∪ A2 is not TF. By Lemma 2.5
there exist a1 < a2 < . . . and b1 < b2 < . . . in Z+, such that {am + bn : 1 ≤ m ≤ n} ⊆
A1 ∪ A2. Let
E = {(m, n) ∈ N2 : m < n, am + bn ∈ A1} ,
F = {(m, n) ∈ N2 : m < n, am + bn ∈ A2 \ A1} .
The sets E and F can be regarded as a partition of the set of 2-element subsets of
N. Hence, by Ramsey's theorem [5, Theorem 9.1.2], there exists either an infinite
set S ⊂ N such that {(x, y) ∈ S2 : x < y} ⊆ E, or an infinite set T ⊂ N such that
{(x, y) ∈ T 2 : x < y} ⊆ F .
In the former case, enumerate S as s0 < s1 < s2 < . . ., and put cj = asj−1, dj = bsj
for j ∈ N. Then ci + dj ∈ A1 for all 1 ≤ i ≤ j; therefore, by Lemma 2.5, A1 is not TF.
In the latter case, a similar argument shows that A2 is not TF. We conclude that
(cid:3)
at least one of A1 and A2 is non-TF, which proves the desired result.
Generalizations to other (semigroup) algebras?
We have relied heavily on the convenient parametrization of Der(ℓ1(Z+), ℓ1(Z+)∗) by
elements of ℓ∞(N). There are analogous parametrizations for Zk
+, where k ≥ 2, but it
is not clear to the authors if they allow one to obtain reasonable higher-rank analogues
of Theorem 2.6.
We can at least make one general observation.
10
Definition 4.1 Let A be a Banach algebra and X a Banach A-bimodule. If x ∈ X,
we say that x is a weakly almost periodic element of X if both a 7→ ax and a 7→ xa are
weakly compact as maps from A to X. The set of all weakly almost periodic elements
of X will be denoted by WAP(X).
Combining Proposition 2.8 with Ruppert's result (Theorem 2.4), we see that if
Dψ is weakly compact then ψ ∈ WAP(ℓ∞(Z+)), where we identify ψ ∈ ℓ∞(N) with
(0, ψ1, ψ2, . . . ) ∈ ℓ∞(Z+). This is a special case of a more general result.
Proposition 4.2 Let A be a unital Banach algebra, let D : A → A∗ be a weakly
compact derivation, and let ψ ∈ A∗ be the functional D(·)(1). Then ψ ∈ WAP(A∗).
Proof . Let κ : A → A∗∗ be the canonical embedding. By Gantmacher's theorem,
D∗ : A∗∗ → A∗ is weakly compact, so D∗κ is also weakly compact. Note that D∗κ(a) =
D(·)(a) for all a ∈ A.
Let a ∈ A, and consider ψ · a ∈ A∗. For each b ∈ A we have
(ψ · a)(b) = ψ(ab) = D(ab)(1) = D(a)(b) + D(b)(a) ;
thus ψ · a = D(a) + D∗κ(a). Since D and D∗κ are weakly compact, this shows that the
map a 7→ ψ · a is weakly compact. A similar argument shows that the map a 7→ a · ψ
is weakly compact, and so ψ ∈ WAP(A∗) as claimed.
(cid:3)
When A = ℓ1(S) is the convolution algebra of a discrete monoid S, we may regard
A∗ = ℓ∞(S) as an algebra with respect to pointwise multiplication. The previous
proposition shows that the functional D(·)(δe) lies in WAP(A∗):
is it the case that
hD(·)(δe) lies in WAP(A∗) for every h ∈ ℓ∞(S)?
Acknowledgments
The authors thank the referee and N. J. Laustsen (as editorial adviser) for a close
reading of the text and for useful corrections.
A A weakly compact derivation which is not p-summing
Definition A.1 Let X and Y be Banach spaces and let p ∈ [1, ∞). We say that a
bounded linear map T : X → Y is p-summing if there exists C > 0 such that:
m
Xj=1
kT xjkp ≤ C p
sup
φ∈X ∗,kφk≤1
m
Xj=1
φ(xj)p,
for all m ∈ N and x1, . . . , xm ∈ X.
(A.1)
The least such C is denoted by πp(T ). If no such C exists (i.e. if T is not p-summing)
we write πp(T ) = ∞.
Recall that in Proposition 1.1, taking S to be the set of integer powers of 2 gives a
derivation DS that is 1-summing. In this appendix, we construct a T-set A such that
DA, while weakly compact, is not p-summing for any finite p. To do this, we shall need
11
some standard general results, which are collected in the following proposition for ease
of reference.
Proposition A.2 Let X and Y be Banach spaces and let T ∈ B(X, Y ).
(i) Let 1 ≤ p ≤ q < ∞. Then πp(T ) ≥ πq(T ).
(ii) If T is p-summing for some p ∈ [1, ∞), then it is weakly compact.
We refer to [4] for the proofs: part (i) may be found as [4, Theorem 2.8]; and
part (ii) follows from the Pietsch factorization theorem, see [4, Theorem 2.17].
We now specialize to operators of the form Dψ. The key observation is the following.
Lemma A.3 Let ψ ∈ ℓ∞(N) and p ∈ [1, ∞) and K < πp(Dψ). There exists N ∈ N
depending on ψ, on p and on K, such that for each ψ′ ∈ ℓ∞(N) satisfying ψ(k) = ψ′(k)
for all k < N , we have πp(Dψ′) > K.
Proof . There are x1, . . . , xm ∈ ℓ1(Z+) such that
m
Xj=1
kDψ(xj)kp > K p
sup
φ∈ℓ∞,kφk≤1
m
Xj=1
φ(xj )p.
(A.2)
Without any loss of generality we may take x1, . . . , xn ∈ c00; write xj = Pl(j)
i=0 αi,jδi.
For each j ∈ {1, . . . , m}, since Dψ(xj) ∈ c0, there exists n(j) ∈ N with Dψ(xj)(δn(j)) =
kDψ(xi)k.
Fix N > max{l(1) + n(1), . . . , l(m) + n(m)}, and let ψ′ ∈ ℓ∞(N) be such that
ψ(k) = ψ′(k) for all k < N . Observe now that for each j we have
Dψ(xj)(δn(j)) =
=
l(j)
Xi=1
l(j)
Xi=1
αi,j
αi,j
i
i + n(j)
i
i + n(j)
ψ(i + n(j))
ψ′(i + n(j)) = Dψ′(xj)(δn(j)) .
Therefore,
m
Xj=1
kDψ′ (xj)kp ≥
m
Xj=1
Dψ′ (xj)(δn(j))p =
m
Xj=1
Dψ(xj)(δn(j))p =
m
Xj=1
kDψ(xj)kp.
Combining this with Equation (A.2) yields πp(Dψ′ ) > K, and the result follows.
(cid:3)
We can now give the promised example.
Theorem A.4 There exists a T-set A such that DA is not p-summing for any p < ∞.
Proof . The set A will be the disjoint union of a sequence of finite arithmetic progressions
whose "skip size" tends to infinity. For each k ∈ Z+, we shall construct, recursively,
A(k) ⊂ N and ck ∈ Z+ such that
12
(a) ck > max A(k);
(b) πk(DB) > k for each B ⊂ N satisfying B ∩ {1, . . . , ck} = A(k) ∩ {1, . . . , ck};
(c) A(k) ⊃ A(k − 1) for all k ≥ 1.
Let A(0) = ∅ and c0 = 0. For each k ∈ N assume that we have already defined
A(k − 1) ⊂ N and ck−1 ∈ N satisfying conditions (a) and (b). Let S := A(k − 1) ∪
(ck−1 + kN). Since S contains an infinite arithmetic progression, it is non-TF. Hence
by Theorem 2.6 DS is not weakly compact, and so by part (i) of Proposition A.2 it is
not k-summing. In particular, πk(DS ) > k, so by applying Lemma A.3 with ψ = χS,
we see that there exists M such that
πk(DS∩{1,...,m}) > k for all m ≥ M .
(A.3)
Choose n such that ck−1 + kn ≥ M , and take
A(k) := S ∩ {1, . . . , ck−1 + kn} = A(k − 1) ∪ {ck−1 + k, ck−1 + 2k, . . . , ck−1 + nk} .
By construction this choice satisfies condition (c). Applying Lemma A.3 with ψ =
χA(k), we can choose ck satisfying conditions (a) and (b), and so our construction may
continue.
Set A := S∞
k=1 A(k). Then for each k ∈ N, A ∩ {1, . . . , ck} = A(k) ∩ {1, . . . , ck} and
so πk(DA) > k. Thus by Proposition A.2(i) πp(DA) = ∞ for all p ∈ [1, ∞). Finally,
if we enumerate the elements of A as an increasing sequence a1 < a2 < . . . , then
ai+1 − ai → ∞; it follows easily that A is a T-set.
(cid:3)
References
[1] N. Bourbaki, Topological vector spaces. Chapters 1 -- 5, Elements of Mathematics
(Berlin), Springer-Verlag, Berlin, 1987. Translated from the French by H. G.
Eggleston and S. Madan.
[2] J. Bourgain, 'New Banach space properties of the disc algebra and H ∞', Acta
Math. 152 (1984) 1 -- 48.
[3] C. Chou, 'Weakly almost periodic functions and thin sets in discrete groups',
Trans. Amer. Math. Soc. 321 (1990) 333 -- 346.
[4] J. Diestel, H. Jarchow, A. Tonge, Absolutely summing operators, vol. 43 of
Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cam-
bridge, 1995.
[5] R. Diestel, Graph theory, vol. 173 of Graduate Texts in Mathematics, Springer-
Verlag, Berlin, third ed., 2005.
[6] N. Hindman, 'On creating sets with large lower density', Discrete Math. 80 (1990)
153 -- 157.
13
[7] M. J. Heath, 'Bounded derivations from Banach algebras', PhD thesis, University
of Nottingham, 2008.
[8] R. E. Megginson, An introduction to Banach space theory, vol. 183 of Graduate
Texts in Mathematics, Springer-Verlag, New York, 1998.
[9] S. E. Morris, 'Bounded derivations from uniform algebras', PhD thesis, Univer-
sity of Cambridge, 1993.
[10] D. E. Ramirez, 'Uniform approximation by Fourier-Stieltjes coefficients', Proc.
Cambridge Philos. Soc. 64 (1968) 615 -- 623.
[11] W. A. F. Ruppert, 'On weakly almost periodic sets', Semigroup Forum 32 (1985)
267 -- 281.
D´epartement de math´ematiques
Departamento de Matem´atica,
et de statistique,
Pavillon Alexandre-Vachon
Universit´e Laval
Qu´ebec, QB
Canada, G1V 0A6
Instituto Superior T´ecnico,
Av. Rovisco Pais
1049-001 Lisboa
Portugal
Email: [email protected]
Email: [email protected]
14
|
1811.04103 | 1 | 1811 | 2018-11-09T19:17:50 | The algebra of bounded type holomorphic functions on the ball | [
"math.FA"
] | We study the spectrum $M_b(U)$ of the algebra of bounded type holomorphic functions on a complete Reinhardt domain in a symmetrically regular Banach space $E$ as an analytic manifold over the bidual of the space. In the case that $U$ is the unit ball of $\ell_p$, $1<p<\infty$, we prove that each connected component of $M_b(B_{\ell_p})$ naturally identifies with a ball of a certain radius. We also provide estimates for this radius and in many natural cases we have the precise value. As a consequence, we obtain that for connected components different from that of evaluations, these radii are strictly smaller than one, and can be arbitrarily small. We also show that for other Banach sequence spaces, connected components do not necessarily identify with balls. | math.FA | math | THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE
BALL
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
Abstract. We study the spectrum Mb(U ) of the algebra of bounded type holomorphic functions
on a complete Reinhardt domain in a symmetrically regular Banach space E as an analytic manifold
over the bidual of the space. In the case that U is the unit ball of ℓp, 1 < p < ∞, we prove that each
connected component of Mb(Bℓp ) naturally identifies with a ball of a certain radius. We also provide
estimates for this radius and in many natural cases we have the precise value. As a consequence,
we obtain that for connected components different from that of evaluations, these radii are strictly
smaller than one, and can be arbitrarily small. We also show that for other Banach sequence spaces,
connected components do not necessarily identify with balls.
1. Introduction
8
1
0
2
v
o
N
9
]
.
A
F
h
t
a
m
[
1
v
3
0
1
4
0
.
1
1
8
1
:
v
i
X
r
a
The study of the spectrum of the algebra of bounded type analytic functions on a Banach space
E was initiated by the seminal article of Aron, Cole and Gamelin [3]. Their main motivation was its
relation with the algebra H∞(BE) of bounded holomorphic functions on the unit ball. As in the one
or finite dimensional case, there is a natural projection defined on the spectrum M of H∞, which in
the infinite dimensional case, has range contained in the closed unit ball of the bidual BE ′′.
The results proved in [3] imply that the interior part of the spectrum M (i.e.
the subset of
homomorphisms which lie in the fibers of the interior points of the ball) naturally identifies with
the spectrum Mb(BE) of the algebra of bounded type holomorphic functions on the unit ball of the
Banach space E.
2000 Mathematics Subject Classification. Primary 46G20, 46E50, 46T25, 46E25. Secondary 58B12, 32D26, 32A38.
Key words and phrases. Holomorphic functions, spectrum of algebras, Riemann domains.
The first author was supported by conicet-pip 11220130100329CO, ANPCyT PICT 2015-2299 and UBACyT
20020130100474BA. The second author was supported by conicet-pip 11220130100329CO, ANPCyT PICT 2015-2224
and UBACyT 20020130300052BA. The third author was supported by FAPESP-Brazil, Proc. 2014/07373-0.
1
2
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
In [4], the authors continued the study of the spectrum of the algebra of bounded type analytic
functions. They showed that for symmetrically regular Banach spaces, the spectrum Mb(U) of the
algebra Hb(U) of bounded type holomorphic functions on an open set U ⊂ E may be endowed with
an analytic structure as an infinite dimensional Banach manifold modeled over the bidual E′′ of
E. This was applied, for example, to characterize the envelope of holomorphy of U in [7, 12]. The
analytic structure of Mb(X) for X a Riemann domain over a symetrically regular Banach space was
studied in [9].
In this article, we study the spectrum of the algebra of bounded type analytic functions on the unit
ball of E (or on a complete Reinhardt domain) from this point of view. More precisely, we aim to
give an accurate description of Mb(U) as analytic manifold. We show that whenever U is a complete
Reinhardt domain in a reflexive space with 1-unconditional basis, each connected component of
Mb(U) is (identified with) a complete Reinhardt set, which is not necessarily a multiple of U. We
also prove that, when U is the unit ball of ℓp, the connected components are identified with balls in
the following sense (see definitions below): they are all of the form
(1)
S = {ϕz
: kzk < r},
for some ϕ in the fiber of 0 and some 0 < r ≤ 1. Moreover, with the exception of the component
formed by evaluations, the radius r is strictly smaller that 1. Also, there are connected components
with arbitrary small radius. To show these facts, we give estimates of the radius of each connected
component and, for the components of most natural homomorphisms, we give their exact value.
This altogether provides a thorough description of Mb(Bℓp), which in turn gives information on the
spectrum of H∞(Bℓp) by [3].
The fact that connected components are identified with balls as in (1) is a particular (isometric)
property of ℓp: we exhibit an example of a Banach space E with 1-unconditional basis for which the
connected components of Mb(BE) are not balls. The example is actually a Banach space isomorphic
to ℓ2.
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
3
We refer to [8, 11] for general theory on complex analysis in Banach spaces, and to [5, 6, 8, 10, 13]
for background on the space of holomorphic functions of bounded type and its spectrum.
2. The spectrum of bounded type functions on complete Reinhardt domains
Let E be a complex Banach space. We denote by E′ its dual, and by BE its open unit ball.
Sometimes, when the underlying Banach space is clear, we use Br(x) to denote the open ball of
radius r centered at x and write Br when the ball is centered at the origin.
For an open subset U ⊂ E, a U-bounded set is a bounded set A ⊂ U whose distance to the
boundary of U, denoted by dU (A), is positive. A family (Un)n∈N of subsets of U is a fundamental
family of U-bounded sets if each Un is U-bounded, and if every U-bounded set is contained in some
Un. Every open set U admits a fundamental family of U-bounded sets, for instance
Un = {x ∈ U : kxk ≤ n, dU (x) ≥
1
n},
for every n ∈ N. A holomorphic function on U which is bounded on U-bounded sets is called of
bounded type on U. The algebra of all bounded type holomorphic functions on U is denoted by
Hb(U) and it is a Fr´echet algebra when it is endowed with the topology of uniform convergence
on U-bounded sets. The spectrum of Hb(U), i.e. the set of non-zero continuous complex valued
homomorphisms on Hb(U), is denoted by Mb(U). For each homomorphism ϕ ∈ Mb(U), there exists
a U-bounded subset A such that
(2)
ϕ(f ) ≤ kfkA,
for every f ∈ Hb(U),
where kfkA is the supremum of f over the set A. We will write ϕ ≺ A when (2) holds.
There is a natural projection π : Mb(U) → E′′, defined by π(ϕ) = ϕE ′ ∈ E′′, ϕ ∈ Mb(U). We thus
have the following commutative diagram:
4
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
U
δ
✲
Mb(U)
❍
❍
❍
❍
❍
❍
jE
❍
❍
❍
π
❍
❍❥
❄
E′′
where δ is the point evaluation mapping and jE : E → E′′ is the natural inclusion.
A Banach space E is symmetrically regular if every continuous symmetric linear mapping T : E →
E′ is weakly compact (an operator T : E → E′ is symmetric if T x1(x2) = T x2(x1) for all x1, x2 ∈ E).
Every reflexive Banach space is symmetrically regular. In [4], for E a symmetrically regular Banach
space and U ⊂ E an open subset, a topology is defined on Mb(U) so that the mapping π above is a
local homeomorphism that makes (Mb(U), π) a Riemann domain over E′′.
Let us briefly describe this topology (see [4] for details). Recall that any holomorphic function f
of bounded type on E may be extended to a function AB(f ) ∈ Hb(E′′) through the Aron-Berner
extension [1]. Given f ∈ Hb(U) and z ∈ E′′, the function
dnf (x)
x 7→ AB(cid:0)
n!
(cid:1)(z),
is a bounded type holomorphic function on U. For ϕ ∈ Mb(U), we denote by dU (ϕ) the supremum
of dU (A) over the U-bounded sets A satisfying ϕ ≺ A. If r < dU (ϕ), it is possible to define, for each
z ∈ E′′ with kzk < r, the homomorphism ϕz given by
ϕ(cid:16)AB(cid:0)
dnf (·)
n!
(cid:1)(z)(cid:17).
ϕz(f ) =
(3)
∞
X
n=0
When E is symmetrically regular, the sets {ϕz : kzk < r}, with ϕ ∈ Mb(U) and r < dU (ϕ), form a
basis of a Hausdorff topology for Mb(U), and each set {ϕz : kzk < r} is homeomorphic to the ball
π(ϕ) + rBE ′′ via the projection π. This endows Mb(U) with an analytic structure over E′′.
Definition 2.1. Let U be an open subset of a symmetrically regular Banach space. The connected
component of a homomorphism ϕ ∈ Mb(U) is called the sheet of ϕ in Mb(U) and is denoted by
SU (ϕ).
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
5
In the case of bounded type entire functions (i.e., U = E), the description of the connected
components of Mb(E) is simpler than for a general open set U, as pointed out in [4] and [8, Section
6.3]. Given z ∈ E′′, the function x 7→ τzf (x) := AB(f )(z +jEx) is an entire function of bounded type
on E. Thus, given ϕ ∈ Mb(E) and z ∈ E′′, the homomorphism ϕz can be equivalently constructed
as
The sheet of ϕ is exactly
ϕz(f ) := ϕ(τzf ).
SE(ϕ) := {ϕz : z ∈ E′′}.
Since π(ϕz) = π(ϕ) + z, π is a homeomorphism between SE(ϕ) and E′′.
Remark 2.2. If U ⊂ E is a balanced open set (or more generally, if U is such that entire functions
of bounded type are dense in Hb(U)), the spectrum Mb(U) is naturally embedded in Mb(E). Indeed,
given ϕ ∈ Mb(U) we can naturally associate a unique character on Hb(E) which is just the restriction
to the bounded type entire functions: ϕHb(E). When the context is clear we will denote this restriction
by ϕ. The natural projection defined on Mb(U) is just the restriction of the projection defined on
Mb(E), and we will denote both as π.
Suppose that U is balanced. The embedding of (Mb(U), π) into (Mb(E), π) is continuous (with
their topologies as Riemann domains), so each connected component of Mb(U) is embedded into a
connected component of Mb(E) (which is homeomorphic to E′′). Therefore, restricted to each sheet,
the projection πSU (ϕ) is a homeomorphism onto some open set of E′′. Our main goal is to describe
the connected components SU (ϕ), and a natural way to do this is to understand the image πSU (ϕ).
Under the same assumptions, given ϕ ∈ Mb(U) and ψ ∈ SU (ϕ) there exists z ∈ E′′ such that
ψ = (ϕ)z and then (ϕ)z belongs to Mb(U) (that is, it can be extended to Hb(U)). Thus, to describe
what the connected components of Mb(U) look like, it will be useful to determine for which z ∈ E′′
the homomorphism (ϕ)z belongs to Mb(U) (which means, again, that (ϕ)z can be extended to
Hb(U)).
The following lemma from [2] will be useful for our results, in particular for Lemma 2.4.
6
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
Lemma 2.3. [2, Lemma 1.7]. Let E be a Banach space with Schauder basis (ek)k∈N, and denote by
(e′k)k∈N its dual basic sequence. Let z ∈ E′′ and ϕ ∈ π−1(z). Then for f ∈ Hb(E) and N ∈ N:
ϕ(f ) = ϕ(cid:16)x 7→ f(cid:0)
N
X
k=1
z(e′k)ek +
∞
X
k=N +1
e′k(x)ek(cid:1)(cid:17).
Lemma 2.4. Let E be a Banach space with Schauder basis (ek)k∈N, and let ϕ ∈ Mb(E) ∩ π−1(0).
For each N ∈ N, the following assertions hold.
(1) For z ∈ E′′ and f ∈ Hb(E),
ϕz(f ) = ϕ(x 7→ AB(f )(z1, . . . , zN , xN +1 + zN +1, xN +2 + zN +2, . . . )).
(2) If ϕ ≺ A, then ϕ ≺ A(N ), where
A(N ) = {(0, . . . , 0, xN +1, xN +2, . . . ) : x = (xj) ∈ A}.
Proof: If ϕ ∈ π−1(0), then z = 0 in Lemma 2.3. Then
ϕ(f ) = ϕ(x 7→ f(cid:0)
∞
X
k=N +1
e′k(x)ek(cid:1)) = ϕ(x 7→ f (0, . . . , 0, xN +1, xN +2, . . . )).
(1) If f ∈ Hb(E), then ϕz(f ) = ϕ(x 7→ AB(f )(x + z)). If we denote g(x) = AB(f )(x + z), then
it follows from Lemma 2.3 that
ϕz(f ) = ϕ(x 7→ g(0, . . . , 0, xN +1, xN +2, . . . ) = AB(f )(z1, . . . , zN , xN +1 + zN +1, xN +2 + zN +2, . . . )).
(2) Since ϕ ≺ A, we have
ϕ(f ) = ϕ(x 7→ f (0, . . . , 0, xN +1, xN +2, . . . )) ≤ sup
x∈A f (0, . . . , 0, xN +1, xN +2, . . . ) = sup
A(N) f. ✷
We recall that a subset U of a Banach space with unconditional basis (ek)k∈N is complete Reinhardt
if P∞k=1 λkxkek ∈ U, whenever P∞k=1 xkek ∈ U and λk ≤ 1 for all k. Proposition 2.6 states that if
U is a complete Reinhardt domain in a Banach space with 1-unconditional basis, then each sheet
in the spectrum is also a complete Reinhardt domain. First we need the following lemma, which is
probably known.
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
7
Lemma 2.5. Let E be a Banach space with unconditional basis and let U ⊂ E be a complete
Reinhardt open set. Then U admits a fundamental system of U-bounded sets formed by complete
Reinhardt sets.
Proof: Any Banach space with unconditional basis can be renormed so that kλ · xk ≤ kxk
whenever kλk∞ ≤ 1. Assuming that E has such a norm, let us show that the sets Un = {x ∈ U :
kxk ≤ n, dU (x) ≥ 1
n} are complete Reinhardt. Note that it suffices to prove that if Bδ(x) ⊂ U and
kλk∞ ≤ 1, then Bδ(λ · x) ⊂ U.
Let y be a point in Bδ(λ · x) and define a vector z ∈ E by specifying its coordinates as follows:
zj =
yj,
xj
xj
max(xj,yj)
if xj = 0,
otherwise.
If the index j is such that xj < yj, then zj − xj = yj − xj ≤ yj − λjxj ≤ yj − λjxj by
the triangle inequality. And if j is such that xj ≥ yj, then zj − xj = 0 ≤ yj − λjxj. Thus
zj − xj ≤ yj − λjxj for every index j, so kz − xk ≤ ky − λ · xk < δ. In other words, z ∈ Bδ(x), so
z ∈ U. Since zj ≥ yj for every j, and U is a complete Reinhardt set, it follows that y ∈ U. But y
is an arbitrary point of Bδ(λ · x), so we conclude that Bδ(λ · x) ⊂ U.
✷
If we only look at the subset of homomorphisms that project to E, then the above topology
restricted to Mb(U) ∩ π−1(E) is well defined, even though E is not symmetrically regular. Thus, for
an arbitrary Banach space E, (Mb(U) ∩ π−1(E), ππ−1(E)) is a Riemann domain over E (see [7]).
Proposition 2.6. Let E be Banach space with 1-unconditional basis (ek)k∈N and let U ⊂ E be a
complete Reinhardt open subset. Then, in each sheet of Mb(U) there is a character ϕ ∈ Mb(U)∩π−1(0)
such that the set
is a complete Reinhardt subset of E.
{w ∈ E : (ϕ)w extends to Mb(U)}
Proof: Recall that since Hb(E) is dense in Hb(U), we have that Mb(U) is embedded in Mb(E).
Then, given ψ ∈ Mb(U) ∩ π−1(E) there exists ϕ ∈ Mb(E) ∩ π−1(0) and z ∈ E such that ψ = ϕz. We
8
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
must show that for every scalar sequence λ with kλk∞ ≤ 1, the vector w = λ · z satisfies that ϕw
extends to Mb(U) whenever ϕz extends to Mb(U). Note that since ϕw belongs to Mb(E), it suffices
to show that ϕw ≺ A for some U-bounded set A.
Let us start by assuming that z = PN
j=1 zjej. If f ∈ Hb(E), it follows by Lemma 2.4 that
ϕw(f ) = ϕ(x 7→ f (λ1z1, . . . , λN zN , xN +1, xN +2, . . . )).
Let us consider the entire function of bounded type,
fλ(x) = f (λ1x1, . . . , λN xN , xN +1, xN +2, . . . ),
then, applying again Lemma 2.4,
ϕz(fλ) = ϕ(x 7→ fλ(z1, . . . , zN , xN +1, xN +2, . . . )) = ϕ(x 7→ f (λ1z1, . . . , λN zN , xN +1, xN +2, . . . )) = ϕw(f ).
By the previous lemma we may take a complete Reinhardt U-bounded set, A, such that ϕz ≺ A.
Then,
ϕw(f ) = ϕz(fλ) ≤ sup
A fλ ≤ sup
A f.
Therefore ϕw ∈ Mb(U) and ϕw ≺ A.
3
3
Take now an arbitrary z ∈ E for which ϕz belongs to Mb(U) with ϕz ≺ A. Let us denote by πN
the projection onto the span of {e1, . . . , eN} and choose 0 < δ < dU (A)
. We can take N such that
kπN (z) − zk < δ < dU (A)
. Now, proceeding as in [4, page 550], we have ϕπN (z) ≺ Aδ := A + Bδ.
By the first part of the proof, for kλk∞ ≤ 1 we have ϕλ·πN (z) ≺ Aδ. Since dU (Aδ) > 2δ and
kλ · πN (z) − λ · zk < δ, we have ϕλ·z ≺ A2δ. Finally, since δ is arbitrary small, we conclude that
ϕλ·z ≺ A.
✷
If the Banach space E is reflexive (which obviously implies that E is symmetrically regular), the
above result tells us that the sheets of Mb(U) are complete Reinhardt domains.
Corollary 2.7. Let E be a reflexive Banach space with 1-unconditional basis and let U ⊂ E be
a complete Reinhardt open subset. Then for each sheet S of Mb(U) there exist a character ϕ ∈
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
9
Mb(U) ∩ π−1(0) and a complete Reinhardt domain V ⊂ E such that
S = {(ϕ)z ∈ Mb(U) : z ∈ V }.
3. The spectrum of bounded type functions on Bℓp
We now focus in the case where U is the unit ball of ℓp. The following theorem shows that each
sheet is also a ball centered at zero. We will see later in Theorem 3.3 that the radius of each sheet
other than the sheet of evaluations, is strictly smaller than 1.
Theorem 3.1. Let E = ℓp, 1 < p < ∞, and let U = Bℓp. Then all sheets are balls centered at 0,
that is, in each sheet there is some ϕ ∈ Mb(U) ∩ π−1(0), and
π(SU (ϕ)) = {w ∈ E : (ϕ)w ∈ Mb(U)} = rBℓp,
for some 0 < r ≤ 1.
Proof: By Corollary 2.7 we know that each sheet intersects π−1(0). So take ϕ ∈ Mb(U) ∩ π−1(0)
and suppose that (ϕ)z belongs to Mb(U) for some z ∈ E. The theorem will be proved if we show
that (ϕ)w ∈ Mb(U) whenever kwk < kzk.
If w = (wj)j∈N and z = (zj)j∈N are such that kwk < kzk, then there exists N1 ∈ N such that
kPN
j=1 wjejk < kPN
j=1 zjejk for every N ≥ N1. On the other hand, since (ϕ)z ∈ SU (ϕ), there exists
δ > 0 such that (ϕ)z+y ∈ SU (ϕ), for all kyk < δ. So let us take N ≥ N1 such that
∞
X
j=N +1
zjp < (cid:16) δ
3(cid:17)p
and
∞
X
j=N +1
wjp < (cid:16) δ
3(cid:17)p
.
Then, if v = (ΠN (z), (I − ΠN )(w)), where ΠN : ℓp −→ ℓp denotes the canonical projection, we
have that (ϕ)v also belongs to SU (ϕ). Note that kwk < kvk and that (I − ΠN )(w) = (I − ΠN )(v).
To show that (ϕ)w ∈ SU (ϕ), we will construct some auxiliary bounded linear transformations, as
follows. First, take γ : CN −→ C such that kγk = k(v1, . . . , vN )k−1 and γ(v1, . . . , vN ) = 1. Next, we
define SN : CN −→ CN by
SN (x) = γ(x)(w1, . . . , wN ),
10
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
which clearly satisfies kSNk ≤ 1 and SN (v1, . . . , vN ) = (w1, . . . , wN ). Finally, let TN : ℓp −→ ℓp be
given by TN (x) = (SN (ΠN (x)), (I − ΠN )(x)). In other words,
TN (x) = (SN (x1, . . . , xN ), xN +1, xN +2, . . . ),
for x ∈ ℓp.
Note that TN (v) = w and, since
kTN (x)kp = kSN (ΠN (x))kp + k(I − ΠN )(x)kp ≤ kSNkpkΠN (x)kp + k(I − ΠN )(x)kp
≤ kΠN (x)kp + k(I − ΠN )(x)kp = kxkp,
we also have kTNk ≤ 1.
If f ∈ Hb(E), then it follows from Lemma 2.4 that
(ϕ)v(f ) = ϕ(x 7→ f (ΠN (v), (I − ΠN )(x + v))
and that
(ϕ)w(f ) = (ϕ)TN (v)(f ) = ϕ(x 7→ f (ΠN (TN (v)), (I − ΠN )(x + TN (v))).
Since ΠN (TN (v)) = SN (ΠN (v)) and (I − ΠN )(TN (v)) = (I − ΠN )(v), we have
(4)
(ϕ)w(f ) = (ϕ)TN (v)(f ) = ϕ(x 7→ f (SN (ΠN (v)), (I − ΠN )(x + v)).
On the other hand, for f ∈ Hb(Bℓp), consider fN = f ◦ TNBℓp ∈ Hb(Bℓp). Then we have
fN (ΠN (v), (I − ΠN )(x + v)) = f ◦ TN(cid:0)ΠN (v), (I − ΠN )(x + v)(cid:1)
= f (TN (v1, . . . , vN , xN +1 + wN +1, xN +2 + wN +2, . . . ))
= f (SN (ΠN (v)), (I − ΠN )(x + v)).
Hence,
(ϕ)v(fN ) = ϕ(x 7→ fN (ΠN (v), (I − ΠN )(x + v)) = ϕ(x 7→ f (SN (ΠN (v)), (I − ΠN )(x + v))
= (ϕ)w(f ).
If A is a U-bounded ball such that ϕv ≺ A, then, using again that kTNk ≤ 1, we conclude that
(ϕ)w(f ) = (ϕ)v(fN ) ≤ sup
A f,
A fN = sup
TN (A)f ≤ sup
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
which shows that (ϕ)w ∈ SU (ϕ).
11
✷.
A natural question at this point is whether each sheet on Mb(BE) is necessarily a ball centered at
zero, for more general Banach spaces. The next example shows that this is not always true.
Example 3.2. Let E = he0i⊕∞ ℓ2. Take ϕ ∈ Mb(BE) to be any limit point of the sequence (δen/√2)n.
By Proposition 2.6 we know that the projection of the sheet of ϕ
π(SBE (ϕ)) = {x ∈ E : (ϕ)x ∈ Mb(BE)},
is a complete Reinhardt open set. Let us show that π(SBE (ϕ)) is not a ball centered at 0. For this
we will see that (ϕ)se0 ∈ Mb(BE) for every s < 1 but that (ϕ)te1 /∈ Mb(BE) for every t > 1/√2.
For the first assertion, just note that the set (se0 + en/√2)n is BE-bounded and clearly (ϕ)se0 ≺
(se0 + en/√2)n, thus (ϕ)se0 ∈ Mb(BE). For the second assertion, define the function f (x) = Pk≥1 x2
k.
Then f ∈ Hb(E) and for every m ∈ N, its mth-power satisfies kf mkBE = 1. On the other hand,
since (ϕ)te1 ∈ Mb(E), we know that for each m ∈ N, (ϕ)te1(f m) is a limit point of (δte1+en/√2f m)n.
Finally, since f (te1 + en/√2)m = (t2 + 1
2)m → ∞ as m → ∞, we conclude that (ϕ)te1 cannot be
extended to Hb(BE).
Now that we know that each sheet of Mb(Bℓp) is a ball centered at zero, we would like to estimate
its radius. Let us first recall some terminology from [3] that will be used in the next theorem. For
ϕ ∈ Mb(BE) and m ≥ 0 we associate ϕm ∈ P(mE)′, as ϕm := ϕP(mE). Recall also that R(ϕ), the
radius of ϕ, is defined as the infimum of all r > 0 such that ϕ ≺ rBE. In [3] it is shown that
R(ϕ) = lim sup
m∈N kϕmk
1
m = sup
m∈N kϕmk
1
m .
It should be mentioned that the definition of the radius and the above result were given for ϕ ∈
Mb(E), but it is easily checked that the same works for ϕ ∈ Mb(BE).
Theorem 3.3. Let E = ℓp, 1 < p < ∞, and let U = Bℓp. Given a sheet S, we take ϕ ∈ S ∩ π−1(0)
(which exists thanks to Theorem 3.1). Then,
(1 − R(ϕ)p)
p · Bℓp ⊂ π(S) ⊂ (cid:0)1 − sup
1
m≥p kϕmk(cid:1)1/⌈p⌉ · Bℓp,
12
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
where ⌈p⌉ denotes the smallest natural number which is ≥ p.
1
Proof: Let us first prove the lower inclusion. Take z ∈ (1 − R(ϕ)p)
p · Bℓp. Since Mb(Bℓp) embeds
in Mb(ℓp), we know that (ϕ)z ∈ Mb(ℓp). We must show that (ϕ)z belongs to Mb(Bℓp), that is, that
(ϕ)z is continuous with respect to the topology in Hb(Bℓp) of uniform convergence on Bℓp-bounded
sets. Recall that the seminorms qs(f ) = P∞n=0 sn(cid:13)(cid:13)
n! (cid:13)(cid:13), with 0 < s < 1, define the topology on
dnf (0)
Hb(Bℓp) (see [8]).
Let f ∈ Hb(ℓp) and let us denote by P∞n=0 Pn its Taylor series at the origin, then
(ϕ)z(f ) =
∞
X
n=0
ϕ(cid:0)x 7→ Pn(x + z)(cid:1).
Now, since kzkp + R(ϕ)p < 1, we can find N ∈ N and r < 1 such that for every y ∈ R(ϕ) · B(N )
have z + y ∈ rBℓp. Then, by the definition of R(ϕ) and Lemma 2.4, it follows that
ℓp
, we
Therefore,
ϕ(cid:0)x 7→ Pn(z + x)(cid:1) ≤
sup
y∈R(ϕ)·B(N)
ℓp
kPn(z + y)k ≤ rnkPnk.
(ϕ)z(f ) ≤
∞
X
n=0
ϕ(cid:0)x 7→ Pn(z + x)(cid:1) ≤
∞
X
n=0
rnkPnk = qr(f ).
This implies that (ϕ)z belongs to Mb(Bℓp).
Now we prove the upper inclusion. By Theorem 3.1 we already know that SBℓp (ϕ) is a ball centered
at zero. Let z = te1, with t⌈p⌉ + supm≥p kϕmk > 1 + δ, for some δ > 0. We will show that (ϕ)z is not
continuous on Hb(Bℓp). This will prove that the radius of the ball SBℓp (ϕ) is smaller than or equal
to (1 − supm≥p kϕmk)1/⌈p⌉.
Let 0 < r < 1 be such that ϕ ≺ rBℓp. Consider m0 ≥ p with t⌈p⌉ + kϕm0k > 1 + δ. For ε < δ, let
P0 ∈ P (m0E) be such that ϕ(P0) > kϕm0k − ε, and kP0k ≤ 1. Note that by Lemma 2.4, we have
that ϕ(P0) = ϕ(P0 ◦ (I − e′1 ⊗ e1)). Let Q0 = P0 ◦ (I − e′1 ⊗ e1). It follows from Lemma 2.4 that
(ϕ)te1(Q0) = ϕ(x 7→ Q0(x + te1)) = ϕ(x 7→ Q0(x)) = ϕ(Q0).
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
13
Consider the polynomial Q(x) = (e′1)⌈p⌉ + Q0(x). Since m0 ≥ p, we have supkxkp≤1 Q(x) ≤ 1.
Indeed, for kxkp ≤ 1,
Q(x) ≤ x1⌈p⌉ + P0 ◦ (I − e′1 ⊗ e1)(x) ≤ x1⌈p⌉ + k(I − e′1 ⊗ e1)(x)km0
p ≤ kxkp
p ≤ 1.
Moreover,
(ϕ)te1(Q) = (ϕ)te1(cid:16)(e′1)⌈p⌉ + Q0(cid:17) = t⌈p⌉ + ϕ(Q0), and then
(ϕ)te1(Q) =t⌈p⌉ + ϕ(Q0) > t⌈p⌉ + kϕm0k − ε > 1 + δ − ε > 1 + s,
for some s > 0.
Therefore it follows that (ϕ)te1(Qn) = (ϕ)te1(Q)n > (1 + s)n → ∞ when n → ∞, while kQnkBℓp ≤
1 for every n. Then (ϕ)te1 /∈ Mb(Bℓp).
✷
The only homomorphism ϕ such that ϕm = 0 for sufficiently large m is δ0, so the previous Theorem
allows us to conclude the following.
Corollary 3.4. Let 1 < p < ∞, and let S ⊂ Mb(Bℓp) be a sheet. Then π(S) = Bℓp if, and only if,
S is the sheet of evaluations.
Remark 3.5. The results of this section can be summarized in the following way. Given a connected
component S of Mb(Bℓp), there exists ϕ ∈ Mb(Bℓp) ∩ π−1(0) and 0 < r ≤ 1 such that
S = {ϕz
: kzk < r}.
Moreover, r and ϕ satisfy
(1 − R(ϕ)p)
1
p ≤ r ≤ (cid:0)1 − sup
m≥p kϕmk(cid:1)1/⌈p⌉.
Some comments deserve to be highlighted. If p is a natural number and ϕ is a homomorphism
such that R(ϕ) = supm∈N kϕmk
R(ϕ)p)
m is attained at m = p , then it follows that π(SBℓp (ϕ)) = B(0, (1 −
p ), and then we have an accurate description of the sheet of ϕ. It is interesting to mention that
1
1
this is not an artificial hypothesis, since the r-block homomorphisms considered in [7, Definition 5.3]
satisfy this condition. From this point of view, [7, Proposition 5.4] can be seen now as a consequence
of Theorem 3.3.
14
DANIEL CARANDO, SANTIAGO MURO, AND DANIELA M. VIEIRA
In [8, Section 6.3], the spectrum Mb(E) of a symmetrically regular Banach space was informally
referred to as the envelope of "bounded" holomorphy of E because each bounded type entire function
is proved to extend to a holomorphic function on Mb(E) which is of bounded type on each connected
component of Mb(E). However, as shown in [7, Proposition 5.1], the extension need not be of
bounded type on the whole Riemann domain, even for a homogeneous polynomial. In the case of
the unit ball, we do not know whether the extensions to the spectrum are of bounded type or not.
If for any ϕ ∈ Mb(Bℓp) the connected components would satisfy
π(SBℓp (ϕ)) = B(0, (1 − R(ϕ)p)
1
p )
(that is, if the left inclusion in Theorem 3.3 were always an equality), then it would be possible to
answer this question affirmatively.
Acknowledgments We would like to thank the anonymous reviewer for his/her comments, which
improved the presentation of the article. We are also indebted to Professor Harold P. Boas for
carefully reading the manuscript, for his valuable comments and, in particular, for providing us with
a simple proof for Lemma 2.5.
D. M. Vieira thanks the Departamento de Matem´atica of the Universidad de Buenos Aires and its
members for their kind hospitality.
References
[1] R. Aron, P. Berner, A Hahn-Banach extension theorem for analytic mappings, Bull. Soc. Math. France 106 (1978)
3-24.
[2] R. Aron, D. Carando, S. Lassalle, M. Maestre, Cluster values of holomorphic functions of bounded type, Trans.
Amer. Math. Soc. 368 (2016) 2355-2369.
[3] R. Aron, B. Cole, T. Gamelin, Spectra of algebras of analytic functions on a Banach space, J. Reine Angew. Math.
415 (1991) 51-93.
[4] R. Aron, P. Galindo, D. Garc´ıa, M. Maestre, Regularity and algebras of analytic functions in infinite dimensions,
Trans. Amer. Math. Soc. 348 (1996) 543-559.
[5] D. Carando, D. Garc´ıa, M. Maestre, Homomorphisms and composition operators on algebras of analytic functions
of bounde type, Adv. Math. 197 (2005) 607-629.
THE ALGEBRA OF BOUNDED TYPE HOLOMORPHIC FUNCTIONS ON THE BALL
15
[6] D. Carando, D. Garc´ıa, M. Maestre, P. Sevilla-Peris, On the spectra of algebras of analytic functions, Contempo-
rary Mathematics, 561 (2012) 165-198.
[7] D. Carando, S. Muro, Envelopes of holomorphy and extension of functions of bounded type, Adv. Math. 229 (2012)
2098-2121.
[8] S. Dineen, Complex Analysis on Infinite Dimensional Spaces, Springer-Verlag London, 1999.
[9] S. Dineen, M. Venkova. Extending bounded type holomorphic mappings on a Banach space, J. Math. Anal. Appl.
297 (2004) 645-658.
[10] T. Gamelin, Analytic functions on Banach spaces, Complex potential theory. Springer, Dordrecht, 1994. 187-233.
[11] J. Mujica, Complex Analysis in Banach Spaces, North-Holland Math. Stud. 120, Amsterdam, 1986.
[12] S. Muro, Funciones holomorfas de tipo acotado e ideales de polinomios homog´eneos en espacios de Banach, Tesis,
Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires, 2010.
[13] D. M. Vieira, Spectra of algebras of holomorphic functions of bounded type, Indag. Math. (N.S.) 18 (2007) 269-279.
Departamento de Matem´atica, Facultad de Cs. Exactas y Naturales, Universidad de Buenos Aires
and IMAS-UBA-CONICET, Argentina.
Departamento de Matem´atica Facultad de Cs. Exactas y Naturales, Universidad de Buenos Aires,
Argentina, and CIFASIS-CONICET
Departamento de Matem´atica, Instituto de Matem´atica e Estat´ıstica, Universidade de Sao Paulo,
Sao Paulo, Brasil
E-mail address: [email protected]
E-mail address: [email protected]
E-mail address: [email protected]
|
1506.09010 | 1 | 1506 | 2015-06-30T09:53:31 | Strong extensions for $q$-summing operators acting in $p$-convex Banach function spaces for $1 \le p \le q$ | [
"math.FA"
] | Let $1\le p\le q<\infty$ and let $X$ be a $p$-convex Banach function space over a $\sigma$-finite measure $\mu$. We combine the structure of the spaces $L^p(\mu)$ and $L^q(\xi)$ for constructing the new space $S_{X_p}^{\,q}(\xi)$, where $\xi$ is a probability Radon measure on a certain compact set associated to $X$. We show some of its properties, and the relevant fact that every $q$-summing operator $T$ defined on $X$ can be continuously (strongly) extended to $S_{X_p}^{\,q}(\xi)$. This result turns out to be a mixture of the Pietsch and Maurey-Rosenthal factorization theorems, which provide (strong) factorizations for $q$-summing operators through $L^q$-spaces when $1 \le q \le p$. Thus, our result completes the picture, showing what happens in the complementary case $1\le p\le q$, opening the door to the study of the multilinear versions of $q$-summing operators also in these cases. | math.FA | math |
STRONG EXTENSIONS FOR q-SUMMING
OPERATORS ACTING IN p-CONVEX BANACH
FUNCTION SPACES FOR 1 ≤ p ≤ q
O. DELGADO AND E. A. S ´ANCHEZ P´EREZ
Abstract. Let 1 ≤ p ≤ q < ∞ and let X be a p-convex Ba-
nach function space over a σ-finite measure µ. We combine the
structure of the spaces Lp(µ) and Lq(ξ) for constructing the new
space S q
(ξ), where ξ is a probability Radon measure on a certain
Xp
compact set associated to X. We show some of its properties, and
the relevant fact that every q-summing operator T defined on X
can be continuously (strongly) extended to S q
(ξ). This result
Xp
turns out to be a mixture of the Pietsch and Maurey-Rosenthal
factorization theorems, which provide (strong) factorizations for
q-summing operators through Lq-spaces when 1 ≤ q ≤ p. Thus,
our result completes the picture, showing what happens in the
complementary case 1 ≤ p ≤ q, opening the door to the study of
the multilinear versions of q-summing operators also in these cases.
1. Introduction
Fix 1 ≤ p ≤ q < ∞ and let T : X → E be a Banach space valued
linear operator defined on a saturated order semi-continuous Banach
function space X related to a σ-finite measure µ. In this paper we prove
an extension theorem for T in the case when T is q-summing and X is
p-convex. In order to do this, we first define and analyze a new class
Date: July 5, 2021.
2010 Mathematics Subject Classification. 46E30, 47B38.
Key words and phrases. Banach function spaces, extension of operators, order
continuity, p-convexity, q-summing operators.
The first author gratefully acknowledge the support of the Ministerio de
Econom´ıa y Competitividad (project #MTM2012-36732-C03-03) and the Junta
de Andaluc´ıa (projects FQM-262 and FQM-7276), Spain.
The second author acknowledges with thanks the support of the Ministerio de
Econom´ıa y Competitividad (project #MTM2012-36740-C02-02), Spain.
1
2
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
of Banach function spaces denoted by S q
Xp(ξ) which have some good
properties, mainly order continuity and p-convexity. The space S q
Xp(ξ)
is constructed by using the spaces Lp(µ) and Lq(ξ), where ξ is a finite
positive Radon measure on a certain compact set associated to X.
Corollary 5.2 states the desired extension for T . Namely, if T is
q-summing and X is p-convex then T can be strongly extended contin-
uously to a space of the type S q
Xp(ξ). Here we use the term "strongly"
for this extension to remark that the map carrying X into S q
Xp(ξ) is
actually injective; as the reader will notice (Proposition 3.1), this is
one of the goals of our result. In order to develop our arguments, we
introduce a new geometric tool which we call the family of p-strongly q-
concave operators. The inclusion of X into S q
Xp(ξ) turns out to belong
to this family, in particular, it is q-concave.
If T is q-summing then it is p-strongly q-concave (Proposition 5.1).
Actually, in Theorem 4.4 we show that in the case when X is p-convex,
T can be continuously extended to a space S q
Xp(ξ) if and only if T is
p-strongly q-concave. This result can be understood as an extension of
some well-known relevant factorizations of the operator theory:
(I) Maurey-Rosenthal factorization theorem: If T is q-concave and X
is q-convex and order continuous, then T can be extended to a
weighted Lq-space related to µ, see for instance [3, Corollary 5].
Several generalizations and applications of the ideas behind this
fundamental factorization theorem have been recently obtained,
see [1, 2, 4, 5, 9].
(II) Pietsch factorization theorem: If T is q-summing then it factors
through a closed subspace of Lq(ξ), where ξ is a probability Radon
measure on a certain compact set associated to X, see for instance
[6, Theorem 2.13].
In Theorem 4.4, the extreme case p = q gives a Maurey-Rosenthal
type factorization, while the other extreme case p = 1 gives a Pietsch
type factorization. We must say also that our generalization will allow
to face the problem of the factorization of several p-summing type
of multilinear operators from products of Banach function spaces -- a
topic of current interest -- , since it allows to understand factorization
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
3
of q-summing operators from p-convex function lattices from a unified
point of view not depending on the order relation between p and q.
As a consequence of Theorem 4.4, we also prove a kind of Kakutani
representation theorem (see for instance [7, Theorem 1.b.2]) through
the spaces S q
Xp(ξ) for p-convex Banach function spaces which are p-
strongly q-concave (Corollary 4.5).
2. Preliminaries
Let (Ω, Σ, µ) be a σ-finite measure space and denote by L0(µ) the
space of all measurable real functions on Ω, where functions which are
equal µ-a.e. are identified. By a Banach function space (briefly B.f.s.)
we mean a Banach space X ⊂ L0(µ) with norm k · kX, such that if
f ∈ L0(µ), g ∈ X and f ≤ g µ-a.e. then f ∈ X and kf kX ≤ kgkX.
In particular, X is a Banach lattice with the µ-a.e. pointwise order, in
which the convergence in norm of a sequence implies the convergence
µ-a.e. for some subsequence. A B.f.s. X is said to be saturated if there
exists no A ∈ Σ with µ(A) > 0 such that f χA = 0 µ-a.e. for all f ∈ X,
or equivalently, if X has a weak unit (i.e. g ∈ X such that g > 0 µ-a.e.).
Lemma 2.1. Let X be a saturated B.f.s. For every f ∈ L0(µ), there
exists (fn)n≥1 ⊂ X such that 0 ≤ fn ↑ f µ-a.e.
Proof. Consider a weak unit g ∈ X and take gn = ng/(1 + ng). Note
that 0 < gn < ng µ-a.e., so gn is a weak unit in X. Moreover, (gn)n≥1
increases µ-a.e. to the constant function equal to 1. Now, take fn =
gnf χ{ω∈Ω: f ≤n}. Since 0 ≤ fn ≤ ngn µ-a.e., we have that fn ∈ X, and
fn ↑ f µ-a.e.
(cid:3)
The Kothe dual of a B.f.s. X is the space X ′ given by the functions
h ∈ X ′. Here, as usual, BX denotes the closed unit ball of X. Each
h ∈ L0(µ) such that R hf dµ < ∞ for all f ∈ X. If X is saturated
then X ′ is a saturated B.f.s. with norm khkX ′ = supf ∈BXR hf dµ for
function h ∈ X ′ defines a functional ζ(h) on X by hζ(h), f i =R hf dµ
for all f ∈ X. In fact, X ′ is isometrically order isomorphic (via ζ) to a
closed subspace of the topological dual X ∗ of X.
From now and on, a B.f.s. X will be assumed to be saturated. If for
every f, fn ∈ X such that 0 ≤ fn ↑ f µ-a.e. it follows that kfnkX ↑
4
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
kf kX, then X is said to be order semi-continuous. This is equivalent to
ζ(X ′) being a norming subspace of X ∗, i.e. kf kX = suph∈BX′R f h dµ
for all f ∈ X. A B.f.s. X is order continuous if for every f, fn ∈ X
such that 0 ≤ fn ↑ f µ-a.e., it follows that fn → f in norm. In this
case, X ′ can be identified with X ∗.
For general issues related to B.f.s.' see [7], [8] and [10, Ch. 15] con-
sidering the function norm ρ defined as ρ(f ) = kf kX if f ∈ X and
ρ(f ) = ∞ in other case.
Let 1 ≤ p < ∞. A B.f.s. X is said to be p-convex if there exists a
constant C > 0 such that
(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
fip(cid:17)1/p(cid:13)(cid:13)(cid:13)X
≤ C(cid:16) nXi=1
kfikp
X(cid:17)1/p
for every finite subset (fi)n
i=1 ⊂ X. In this case, M p(X) will denote
the smallest constant C satisfying the above inequality. Note that
M p(X) ≥ 1. A relevant fact is that every p-convex B.f.s. X has an
equivalent norm for which X is p-convex with constant M p(X) = 1,
see [7, Proposition 1.d.8].
The p-th power of a B.f.s. X is the space defined as
Xp = {f ∈ L0(µ) : f 1/p ∈ X},
endowed with the quasi-norm kf kXp = k f 1/p kp
X, for f ∈ Xp. Note
that Xp is always complete, see the proof of [8, Proposition 2.22]. If X
is p-convex with constant M p(X) = 1, from [3, Lemma 3], k · kXp is a
norm and so Xp is a B.f.s. Note that Xp is saturated if and only if X
is so. The same holds for the properties of being order continuous and
order semi-continuous.
3. The space S q
Xp(ξ)
Let 1 ≤ p ≤ q < ∞ and let X be a saturated p-convex B.f.s. We can
assume without loss of generality that the p-convexity constant M p(X)
is equal to 1. Then, Xp and (Xp)′ are saturated B.f.s.'. Consider the
topology σ(cid:0)(Xp)′, Xp(cid:1) on (Xp)′ defined by the elements of Xp. Note
(Xp)′ of all positive elements of the closed unit ball of
that the subset B+
(Xp)′ is compact for this topology.
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
5
Let ξ be a finite positive Radon measure on B+
(Xp)′. For f ∈ L0(µ),
consider the map φf : B+
(Xp)′ → [0, ∞] defined by
φf (h) =(cid:16)ZΩ
f (ω)ph(ω) dµ(ω)(cid:17)q/p
for all h ∈ B+
(Xp)′. In the case when f ∈ X, since f p ∈ Xp, it follows
that φf is continuous and so measurable. For a general f ∈ L0(µ), by
Lemma 2.1 we can take a sequence (fn)n≥1 ⊂ X such that 0 ≤ fn ↑ f
µ-a.e. Applying monotone convergence theorem, we have that φfn ↑ φf
pointwise and so φf is measurable. Then, we can consider the integral
φf (h)dξ(h) ∈ [0, ∞] and define the following space:
(Xp)′
S q
Let us endow S q
(Xp)′(cid:16)ZΩ
RB+
Xp(ξ) =(f ∈ L0(µ) : ZB+
(ξ) = ZB+
= (cid:13)(cid:13)(cid:13) h →(cid:13)(cid:13)f h1/p(cid:13)(cid:13)Lp(µ)(cid:13)(cid:13)(cid:13)Lq(ξ)
(Xp)′(cid:16)ZΩ
Xp(ξ) with the seminorm
kf kS q
Xp
.
f (ω)ph(ω) dµ(ω)(cid:17)q/p
dξ(h) < ∞) .
f (ω)ph(ω) dµ(ω)(cid:17)q/p
dξ(h)!1/q
In general, k · kS q
(ξ) is not a norm. For instance, if ξ is the Dirac
Xp
measure at some h0 ∈ B+
(Xp)′ such that A = {ω ∈ Ω : h0(ω) = 0}
satisfies µ(A) > 0, taking f = gχA ∈ X with g being a weak unit of
X, we have that
kf kS q
Xp
and
(ξ) =(cid:16)ZA
g(ω)ph0(ω) dµ(ω)(cid:17)1/p
= 0
µ({ω ∈ Ω : f (ω) 6= 0}) = µ(A ∩ {ω ∈ Ω : g(ω) 6= 0}) = µ(A) > 0.
Proposition 3.1. If the Radon measure ξ satisfies
ZB+
(Xp)′(cid:16)ZA
h(ω) dµ(ω)(cid:17)q/p
dξ(h) = 0 ⇒ µ(A) = 0
(3.1)
Xp(ξ) is a saturated B.f.s. Moreover, S q
then, S q
p-convex (with constant 1) and X ⊂ S q
Xp(ξ) continuously.
Xp(ξ) is order continuous,
6
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
Xp(ξ) and kf kS q
Proof. It is clear that if f ∈ L0(µ), g ∈ S q
f ∈ S q
norm. Suppose that kf kS q
Xp
for every n ≥ 1. Since χAn ≤ nf and
(ξ) ≤ kgkS q
Xp
Xp(ξ) and f ≤ g µ-a.e. then
(ξ) is a
(ξ) = 0 and set An = {ω ∈ Ω : f (ω) > 1
n }
(ξ). Let us see that k · kS q
Xp
Xp
(Xp)′(cid:16)ZAn
ZB+
h(ω) dµ(ω)(cid:17)q/p
dξ(h) =(cid:13)(cid:13)χAn(cid:13)(cid:13)q
from (3.1) we have that µ(An) = 0 and so
(ξ) ≤ nqkf kq
S q
Xp
(ξ) = 0,
S q
Xp
µ({ω ∈ Ω : f (ω) 6= 0}) = lim
n→∞
µ(An) = 0.
Now we will see that S q
S q
Xp(ξ) whenever (fn)n≥1 ⊂ S q
Xp(ξ) is complete by showing thatPn≥1 fn ∈
Xp(ξ) with C =P kfnkS q
let us prove that Pn≥1 fn < ∞ µ-a.e. For every N, n ≥ 1, taking
NPn
n = {ω ∈ Ω : Pn
ZB+
(Xp)′(cid:16)ZAN
h(ω) dµ(ω)(cid:17)q/p
j=1 fj(ω) > N}, since χAN
j=1 fj, we have
dξ(h) = kχAN
(ξ) < ∞. First
AN
that
n ≤ 1
n kq
S q
Xp
(ξ)
Xp
n
≤
nXj=1
1
N q(cid:13)(cid:13)(cid:13)
q
fj(cid:13)(cid:13)(cid:13)
≤
(ξ)
C q
N q .
S q
Xp
Note that, for N fixed, (AN
applying twice the monotone convergence theorem, it follows that
n )n≥1 increases. Taking limit as n → ∞ and
ZB+
(Xp)′(cid:16)Z∪n≥1AN
n
h(ω) dµ(ω)(cid:17)q/p
dξ(h) ≤
C q
N q .
Then,
(Xp)′(cid:16)Z∩N ≥1∪n≥1AN
ZB+
n
h(ω) dµ(ω)(cid:17)q/p
dξ(h) ≤ lim
N→∞
C q
N q = 0,
and so, from (3.1),
µ(cid:16)nω ∈ Ω : Xn≥1
fn(ω) = ∞o(cid:17) = µ(cid:16) \N ≥1[n≥1
AN
n(cid:17) = 0.
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
7
Hence, Pn≥1 fn ∈ L0(µ). Again applying the monotone convergence
theorem, it follows that
(Xp)′(cid:16)ZΩ(cid:12)(cid:12)(cid:12)Xn≥1
ZB+
(Xp)′(cid:16)ZΩ(cid:0)Xn≥1
ZB+
n→∞ZB+
(Xp)′(cid:16)ZΩ(cid:0) nXj=1
lim
p
fn(ω)(cid:12)(cid:12)(cid:12)
h(ω) dµ(ω)(cid:17)q/p
fn(ω)(cid:1)ph(ω) dµ(ω)(cid:17)q/p
fj(ω)(cid:1)ph(ω) dµ(ω)(cid:17)q/p
fj(cid:13)(cid:13)(cid:13)
n→∞(cid:13)(cid:13)(cid:13)
nXj=1
lim
dξ(h) ≤
dξ(h) =
dξ(h) =
q
S q
Xp
≤ C q
(ξ)
and thusPn≥1 fn ∈ S q
Xp(ξ).
Note that if f ∈ X, for every h ∈ B+
(Xp)′ we have that
f (ω)ph(ω) dµ(ω) ≤ k f p kXpkhk(Xp)′ ≤ kf kp
X
and so
ZΩ
ZB+
(Xp)′(cid:16)ZΩ
Then, X ⊂ S q
particular, S q
S q
Xp(ξ).
Let us show that S q
f (ω)ph(ω) dµ(ω)(cid:17)q/p
(ξ) ≤ ξ(cid:0)B+
Xp
Xp(ξ) and kf kS q
Xp(ξ) is saturated, as a weak unit in X is a weak unit in
(Xp)′(cid:1)1/q kf kX for all f ∈ X. In
dξ(h) ≤ kf kq
X ξ(cid:0)B+
(Xp)′(cid:1).
Xp(ξ) is order continuous. Consider f, fn ∈ S q
Xp(ξ)
such that 0 ≤ fn ↑ f µ-a.e. Note that, since
(Xp )′(cid:16)ZΩ
ZB+
f (ω)ph(ω) dµ(ω)(cid:17)q/p
dξ(h) < ∞,
there exists a ξ-measurable set B with ξ(B+
f − fnph ↓ 0 µ-a.e. and f − fnph ≤ f ph µ-a.e. Then, applying
RΩ f (ω)ph(ω) dµ(ω) < ∞ for all h ∈ B. Fixed h ∈ B, we have that
the dominated convergence theorem,RΩ f (ω) − fn(ω)ph(ω) dµ(ω) ↓ 0.
(Xp)′\B) = 0 such that
8
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
Consider the measurable functions φ, φn : B+
(Xp)′ → [0, ∞] given by
φ(h) = (cid:16)ZΩ
φn(h) = (cid:16)ZΩ
f (ω)ph(ω) dµ(ω)(cid:17)q/p
f (ω) − fn(ω)ph(ω) dµ(ω)(cid:17)q/p
for all h ∈ B+
by the dominated convergence theorem, we obtain
(Xp)′. It follows that φn ↓ 0 ξ-a.e. and φn ≤ φ ξ-a.e. Again
kf − fnkq
S q
Xp
(ξ) =ZB+
(Xp)′
φn(h)dξ(h) ↓ 0.
Finally, let us see that S q
consider the measurable functions φi : B+
defined by
Xp(ξ) is p-convex. Fix (fi)n
Xp(ξ) and
(Xp)′ → [0, ∞] (for 1 ≤ i ≤ n)
i=1 ⊂ S q
fi(ω)ph(ω) dµ(ω).
for all h ∈ B+
(Xp)′. Then,
(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
fip(cid:17)1/p(cid:13)(cid:13)(cid:13)
q
S q
Xp
dξ(h)
(Xp)′(cid:16)ZΩ
fi(ω)ph(ω) dµ(ω)(cid:17)q/p
nXi=1
(Xp)′(cid:16) nXi=1
φi(h)(cid:17)q/p
kφikLq/p(ξ)(cid:17)q/p
dξ(h)
.
(ξ)
φi(h) =ZΩ
= ZB+
= ZB+
≤ (cid:16) nXi=1
fip(cid:17)1/p(cid:13)(cid:13)(cid:13)S q
Xp
Since kφikLq/p(ξ) = kfikp
S q
Xp
(ξ) for all 1 ≤ i ≤ n, we have that
(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
(ξ)
≤(cid:16) nXi=1
kfikp
S q
Xp
(ξ)(cid:17)1/p
.
(cid:3)
Example 3.2. Take a weak unit g ∈ (Xp)′ and consider the Radon
measure ξ as the Dirac measure at g. If A ∈ Σ is such that
0 =ZB+
(Xp )′(cid:16)ZA
h(ω) dµ(ω)(cid:17)q/p
dξ(h) =(cid:16)ZA
g(ω) dµ(ω)(cid:17)q/p
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
9
then, gχA = 0 µ-a.e. and so, since g > 0 µ-a.e., µ(A) = 0. That is, ξ
satisfies (3.1). In this case, S q
Xp(ξ) = Lp(gdµ) with equal norms, as
(Xp )′(cid:16)ZΩ
ZB+
f (ω)ph(ω) dµ(ω)(cid:17)q/p
dξ(h) =(cid:16)ZΩ
f (ω)pg(ω) dµ(ω)(cid:17)q/p
for all f ∈ L0(µ).
Example 3.3. Write Ω = ∪n≥1Ωn with (Ωn)n≥1 being a disjoint sequence
of measurable sets and take a sequence of strictly positive elements
on B+
(Xp)′, where δgχΩn is the Dirac measure at gχΩn with g ∈ (Xp)′
being a weak unit. Note that for every positive function φ ∈ L0(ξ), it
(αn)n≥1 ∈ ℓ1. Let us consider the Radon measure ξ = Pn≥1 αnδgχΩn
follows thatRB+
φ dξ =Pn≥1 αnφ(gχΩn). If A ∈ Σ is such that
(Xp )′(cid:16)ZA
0 =ZB+
g(ω) dµ(ω)(cid:17)q/p
h(ω) dµ(ω)(cid:17)q/p
dξ(h) =Xn≥1
then,RA∩Ωn
g(ω) dµ(ω) =Xn≥1ZA∩Ωn
ZA
αn(cid:16)ZA∩Ωn
g(ω) dµ(ω) = 0 for all n ≥ 1. Hence,
g(ω) dµ(ω) = 0
(Xp )′
and so gχA = 0 µ-a.e., from which µ(A) = 0. That is, ξ satisfies (3.1).
For every f ∈ L0(µ) we have that
ZB+
(Xp )′(cid:16)ZΩ
Xn≥1
f (ω)ph(ω) dµ(ω)(cid:17)q/p
αn(cid:16)ZΩn
dξ(h) =
Xp(ξ) can be described as the space of functions
.
f (ω)pg(ω) dµ(ω)(cid:17)q/p
n kf kLp(gχΩn dµ)(cid:1)n≥1 ∈ ℓq. Moreover,
for all f ∈ S q
Xp(ξ).
Then, the B.f.s. S q
f ∈ ∩n≥1Lp(gχΩndµ) such that(cid:0)α1/q
Lp(gχΩn dµ)(cid:17)1/q
(ξ) =(cid:16)Pn≥1 αn kf kq
kf kS q
Xp
4. p-strongly q-concave operators
Let 1 ≤ p ≤ q < ∞ and let T : X → E be a linear operator from a
saturated B.f.s. X into a Banach space E. Recall that T is said to be
10
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
q-concave if there exists a constant C > 0 such that
(cid:16) nXi=1
kT (fi)kq
E(cid:17)1/q
≤ C(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
fiq(cid:17)1/q(cid:13)(cid:13)(cid:13)X
for every finite subset (fi)n
i=1 ⊂ X. The smallest possible value of C will
be denoted by Mq(T ). For issues related to q-concavity see for instance
[7, Ch. 1.d]. We introduce a little stronger notion than q-concavity: T
will be called p-strongly q-concave if there exists C > 0 such that
(cid:16) nXi=1
kT (fi)kq
E(cid:17)1/q
≤ C
sup
(βi)i≥1∈Bℓr(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
βifip(cid:17)1/p(cid:13)(cid:13)(cid:13)X
p − 1
q .
for every finite subset (fi)n
i=1 ⊂ X, where 1 < r ≤ ∞ is such that
r = 1
1
In this case, Mp,q(T ) will denote the smallest constant
C satisfying the above inequality. Noting that r
p are conjugate
exponents, it is clear that every p-strongly q-concave operator is q-
concave and so continuous, and moreover kT k ≤ Mq(T ) ≤ Mp,q(T ).
As usual, we will say that X is p-strongly q-concave if the identity map
I : X → X is so, and in this case, we denote Mp,q(X) = Mp,q(I).
p and q
Our goal is to get a continuous extension of T to a space of the type
S q
Xp(ξ) in the case when T is p-strongly q-concave and X is p-convex.
To this end we will need to describe the supremum on the right-hand
side of the p-strongly q-concave inequality in terms of the Kothe dual
of Xp.
Lemma 4.1. If X is p-convex and order semi-continuous then
sup
(βi)i≥1∈Bℓr(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
βifip(cid:17)1/p(cid:13)(cid:13)(cid:13)X
= sup
h∈B+
(Xp)′(cid:16) nXi=1(cid:16)Z fiph dµ(cid:17)q/p(cid:17)1/q
for every finite subset (fi)n
1
r = 1
unit ball B(Xp)′ of (Xp)′.
q and B+
p − 1
i=1 ⊂ X, where 1 < r ≤ ∞ is such that
(Xp)′ is the subset of all positive elements of the closed
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
11
Proof. Given (fi)n
so, and (ℓq/p)∗ = ℓr/p, as r
i=1 ⊂ X, since Xp is order semi-continuous, as X is
p , we have that
p is the conjugate exponent of q
sup
(βi)∈Bℓr(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
p
X
βifip(cid:17)1/p(cid:13)(cid:13)(cid:13)
=
=
=
=
=
=
sup
(βi)∈Bℓr(cid:13)(cid:13)(cid:13)
sup
(βi)∈Bℓr
sup
sup
h∈B+
βifiph dµ
βifiph dµ
βifip(cid:13)(cid:13)(cid:13)Xp
nXi=1
h∈B(Xp)′Z
nXi=1
(Xp)′Z
nXi=1
βipZ fiph dµ
nXi=1
αiZ fiph dµ
nXi=1
(Xp)′(cid:16) nXi=1(cid:16)Z fiph dµ(cid:17)q/p(cid:17)p/q
sup
(αi)∈B+
(βi)∈Bℓr
sup
(Xp)′
ℓr/p
.
sup
(βi)∈Bℓr
sup
h∈B+
(Xp)′
sup
h∈B+
sup
h∈B+
(cid:3)
In the following remark, from Lemma 4.1, we obtain easily an exam-
ple of p-strongly q-concave operator.
Remark 4.2. Suppose that X is p-convex and order semi-continuous.
For every finite positive Radon measure ξ on B+
(Xp)′ satisfying (3.1), it
follows that the inclusion map i : X → S q
Xp(ξ) is p-strongly q-concave.
Indeed, for each (fi)n
i=1 ⊂ X, we have that
nXi=1
kfikq
S q
Xp
(ξ) =
(Xp)′(cid:16)ZΩ
nXi=1ZB+
(Xp)′(cid:1)
≤ ξ(cid:0)B+
sup
h∈B+
(Xp)′
fi(ω)ph(ω) dµ(ω)(cid:17)q/p
nXi=1(cid:16)ZΩ
dξ(h)
and so, Lemma 4.1 gives the conclusion for Mp,q(i) ≤ ξ(cid:0)B+
Now let us prove our main result.
fi(ω)ph(ω) dµ(ω)(cid:17)q/p
(Xp)′(cid:1)1/q
.
12
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
Theorem 4.3. If T is p-strongly q-concave and X is p-convex and
order semi-continuous, then there exists a probability Radon measure ξ
on B+
(Xp)′ satisfying (3.1) such that
kT (f )kE ≤ Mp,q(T )(cid:16)ZB+
(Xp)′(cid:16)ZΩ
f (ω)ph(ω) dµ(ω)(cid:17)q/p
dξ(h)(cid:17)1/q
(4.1)
for all f ∈ X.
Proof. Recall that the stated topology on (Xp)′ is σ((Xp)′, Xp), the
one which is defined by the elements of Xp. For each finite subset
(with possibly repeated elements) M = (fi)m
i=1 ⊂ X, consider the map
ψM : B+
for
h ∈ B+
(Xp)′. Note that ψM attains its supremum as it is continuous on a
compact set, so there exists hM ∈ B+
ψM (h) =
ψM (hM ). Then, the p-strongly q-concavity of T , together with Lemma
4.1, gives
(Xp)′ → [0, ∞) defined by ψM (h) = Pm
i=1(cid:0)RΩ fip h dµ(cid:1)q/p
(Xp)′ such that suph∈B+
(Xp)′
mXi=1
kT (fi)kq
E ≤ Mp,q(T )q
≤ Mp,q(T )q
mXi=1(cid:16)ZΩ
ψM (h)
fiph dµ(cid:17)q/p
sup
h∈B+
(Xp)′
sup
h∈B+
(Xp)′
= Mp,q(T )q ψM (hM ).
(4.2)
Consider now the continuous map φM : B+
(Xp)′ → R defined by
φM (h) = Mp,q(T )q ψM (h) −
kT (fi)kq
E
mXi=1
i )k
i=1, M ′ = (f ′
for h ∈ B+
every M = (fi)m
tφM + (1 − t)φM ′ = φM ′′ where M ′′ = (cid:0)t1/qfi(cid:1)m
(Xp)′. Take B = {φM : M is a finite subset of X}. Since for
i=1 ⊂ X and 0 < t < 1, it follows that
i=1,
we have that B is convex. Denote by C(B+
(Xp)′) the space of continuous
real functions on B+
(Xp)′, endowed with the supremum norm, and by A
the open convex subset {φ ∈ C(B+
(Xp)′}.
By (4.2) we have that A ∩ B = ∅. From the Hahn-Banach separation
theorem, there exist ξ ∈ C(B+
(Xp)′)∗ and α ∈ R such that hξ, φi < α ≤
hξ, φMi for all φ ∈ A and φM ∈ B. Since every negative constant
i=1 ∪(cid:0)(1 − t)1/qf ′
i(cid:1)k
(Xp)′) : φ(h) < 0 for all h ∈ B+
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
13
function is in A, it follows that 0 ≤ α. Even more, α = 0 as the
constant function equal to 0 is just φ{0} ∈ B.
It is routine to see
that hξ, φi ≥ 0 whenever φ ∈ C(B+
(Xp)′) is such that φ(h) ≥ 0 for all
h ∈ B+
(Xp)′) and
so it can be interpreted as a finite positive Radon measure on B+
(Xp)′.
Hence, we have that
(Xp)′. Then, ξ is a positive linear functional on C(B+
0 ≤ZB+
(Xp)′
φM dξ
for all finite subset M ⊂ X. Dividing by ξ(B+
(Xp)′), we can suppose
that ξ is a probability measure. Then, for M = {f } with f ∈ X, we
obtain that
kT (f )kq
E ≤ Mp,q(T )qZB+
(Xp)′(cid:16)ZΩ
f (ω)ph(ω) dµ(ω)(cid:17)q/p
dξ(h)
(cid:3)
and so (4.1) holds.
Actually, Theorem 4.3 says that we can find a probability Radon
measure ξ on B+
(Xp)′ such that T : X → E is continuous when X is
considered with the norm of the space S q
Xp(ξ). In the next result we
will see how to extend T continuously to S q
Xp(ξ). Even more, we will
show that this extension is possible if and only if T is p-strongly q-
concave.
Theorem 4.4. Suppose that X is p-convex and order semi-continuous.
The following statements are equivalent:
(a) T is p-strongly q-concave.
(b) There exists a probability Radon measure ξ on B+
(3.1) such that T can be extended continuously to S q
is a factorization for T as
(Xp)′ satisfying
Xp(ξ), i.e. there
X
i
T
S q
Xp(ξ)
E
eT
where eT is a continuous linear operator and i is the inclusion map.
If (a)-(b) holds, then Mp,q(T ) = keT k.
/
/
"
"
<
<
14
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
(Xp)′ satisfying (3.1) such that kT (f )kE ≤ Mp,q(T )kf kS q
Proof. (a) ⇒ (b) From Theorem 4.3, there is a probability Radon mea-
sure ξ on B+
for all f ∈ X. Given 0 ≤ f ∈ S q
(fn)n≥1 ⊂ X such that 0 ≤ fn ↑ f µ-a.e. Then, since S q
continuous, we have that fn → f in S q
(ξ)
Xp(ξ), from Lemma 2.1, we can take
Xp(ξ) is order
Xp(ξ) and so (cid:0)T (fn)(cid:1)n≥1 con-
verges to some element e of E. Define eT (f ) = e. Note that eT is well
defined, since if (gn)n≥1 ⊂ X is such that 0 ≤ gn ↑ f µ-a.e., then
Xp
kT (fn) − T (gn)kE ≤ Mp,q(T )kfn − gnkS q
Xp
(ξ) → 0.
Moreover,
keT (f )kE = lim
n→∞
kT (fn)kE
≤ Mp,q(T ) lim
n→∞
= Mp,q(T )kf kS q
Xp
(ξ).
kfnkS q
Xp
(ξ)
For a general f ∈ S q
Xp(ξ), writing f = f + − f − where f + and f − are
the positive and negative parts of f respectively, we define eT (f ) =
eT (f +) − eT (f −). Then, eT : S q
extending T . Moreover keT k ≤ Mp,q(T ). Indeed, let f ∈ S q
Xp(ξ) → E is a continuous linear operator
Xp(ξ) and
n ↑ f −
n )n≥1 ⊂ X such that 0 ≤ f +
n ↑ f + and 0 ≤ f −
n )n≥1, (f −
take (f +
µ-a.e. Then, f +
n → f in S q
n − f −
Xp(ξ) and
T (f +
n − f −
n ) = T (f +
n ) − T (f −
n ) → eT (f +) − eT (f −) = eT (f )
in E. Hence,
keT (f )kE = lim
n→∞
kT (f +
n − f −
kf +
n )kE
n − f −
≤ Mp,q(T ) lim
n→∞
= Mp,q(T )kf kS q
Xp
(ξ).
n kS q
Xp
(ξ)
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
15
(b) ⇒ (a) Given (fi)n
i=1 ⊂ X, we have that
kT (fi)kq
E =
nXi=1
nXi=1
= keT kq
≤ keT kq
keT (fi)kq
nXi=1ZB+
sup
h∈B+
(Xp)′
E ≤ keT kq
(Xp )′(cid:16)ZΩ
nXi=1(cid:16)ZΩ
(ξ)
kfikq
S q
Xp
nXi=1
fi(ω)ph(ω) dµ(ω)(cid:17)q/p
fi(ω)ph(ω) dµ(ω)(cid:17)q/p
.
dξ(h)
keT k.
That is, from Lemma 4.1, T is p-strongly q-concave with Mp,q(T ) ≤
(cid:3)
A first application of Theorem 4.4 is the following Kakutani type
representation theorem (see for instance [7, Theorem 1.b.2]) for B.f.s.'
being order semi-continuous, p-convex and p-strongly q-concave.
Corollary 4.5. Suppose that X is p-convex and order semi-continuous.
The following statements are equivalent:
(a) X is p-strongly q-concave.
(b) There exists a probability Radon measure ξ on B+
Xp(ξ) with equivalent norms.
(3.1), such that X = S q
(Xp)′ satisfying
Proof. (a) ⇒ (b) The identity map I : X → X is p-strongly q-concave
as X is so. Then, from Theorem 4.4, there exists a probability Radon
measure ξ on B+
(Xp)′ satisfying (3.1), such that I factors as
X
i
I
S q
Xp(ξ)
X
eI
where eI is a continuous linear operator with keIk = Mp,q(X) and i
is the inclusion map. Since ξ is a probability measure, we have that
(ξ) ≤ kf kX for all f ∈ X, see the proof of Proposition 3.1. Let
kf kS q
Xp
0 ≤ f ∈ S q
Xp(ξ). By Lemma 2.1, we can take (fn)n≥1 ⊂ X such that
0 ≤ fn ↑ f µ-a.e. Since S q
Xp(ξ) is order continuous, it follows that
fn → f in S q
Xp(ξ) and so fn = eI(fn) → eI(f ) in X. Then, there is a
/
/
"
"
<
<
16
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
X. For a general f ∈ S q
Xp(ξ), writing f = f + − f − where f + and f −
are the positive and negative parts of f respectively, we have that f =
subsequence of (fn)n≥1 converging µ-a.e. to eI(f ) and hence f =eI(f ) ∈
Xp(ξ) andeI is de identity
eI(f +) −eI(f −) = eI(f ) ∈ X. Therefore, X = S q
map. Moreover, kf kX = keI(f )kX ≤ keIk kf kS q
(ξ) = Mp,q(X)kf kS q
Xp
(b) ⇒ (a) From Remark 4.2 it follows that the identity map I : X →
(cid:3)
X is p-strongly q-concave.
for all f ∈ X.
(ξ)
Xp
Note that under conditions of Corollary 4.5, if X is p-strongly q-
Xp(ξ) with equal
concave with constant Mp,q(X) = 1, then X = S q
norms.
5. q-summing operators on a p-convex B.f.s.
Recall that a linear operator T : X → E between Banach spaces is
said to be q-summing (1 ≤ q < ∞) if there exists a constant C > 0
such that
(cid:16) nXi=1
kT xikq
E(cid:17)1/q
≤ C sup
x∗∈BX∗(cid:16) nXi=1
hx∗, xiiq(cid:17)1/q
for every finite subset (xi)n
i=1 ⊂ X. Denote by πq(T ) the smallest
possible value of C. Information about q-summing operators can be
found in [6].
One of the main relations between summability and concavity for
operators defined on a B.f.s. X, is that every q-summing operator is
q-concave. This is a consequence of a direct calculation which shows
that for every (fi)n
i=1 ⊂ X and x∗ ∈ X ∗ it follows that
(cid:16) nXi=1
hx∗, fiiq(cid:17)1/q
≤ kx∗kX ∗(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
fiq(cid:17)1/q(cid:13)(cid:13)(cid:13)X
see for instance [7, Proposition 1.d.9] and the comments below. How-
ever, this calculation can be slightly improved to obtain the following
result.
,
(5.1)
Proposition 5.1. Let 1 ≤ p ≤ q < ∞. Every q-summing linear
operator T : X → E from a B.f.s. X into a Banach space E, is p-
strongly q-concave with Mp,q(T ) ≤ πq(T ).
STRONG EXTENSIONS FOR q-SUMMING OPERATORS
17
Proof. Let 1 < r ≤ ∞ be such that 1
subset (fi)n
r = 1
i=1 ⊂ X. We only have to prove
p − 1
q and consider a finite
sup
x∗∈BX∗(cid:16) nXi=1
hx∗, fiiq(cid:17)1/q
Fix x∗ ∈ BX ∗. Noting that q
using the inequality (5.1), we have
(cid:16) nXi=1
hx∗, fiiq(cid:17)1/q
=
=
≤
p are conjugate exponents and
.
≤ sup
p and r
sup
(αi)i≥1∈B
(βi)i≥1∈Bℓr(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
ℓr/p(cid:16) nXi=1
(βi)i≥1∈Bℓr(cid:16) nXi=1
(βi)i≥1∈Bℓr(cid:13)(cid:13)(cid:13)(cid:16) nXi=1
βifip(cid:17)1/p(cid:13)(cid:13)(cid:13)X
αihx∗, fiip(cid:17)1/p
hx∗, βifiip(cid:17)1/p
βifip(cid:17)1/p(cid:13)(cid:13)(cid:13)X
sup
sup
.
Taking supremum in x∗ ∈ BX ∗ we get the conclusion.
(cid:3)
From Proposition 5.1, Theorem 4.4 and Remark 4.2, we obtain the
final result.
Corollary 5.2. Set 1 ≤ p ≤ q < ∞. Let X be a saturated order semi-
continuous p-convex B.f.s. and consider a q-summing linear operator
T : X → E with values in a Banach space E. Then, there exists a
probability Radon measure ξ on B+
(Xp)′ satisfying (3.1) such that T can
be factored as
X
i
T
S q
Xp(ξ)
E
eT
where eT is a continuous linear operator with keT k ≤ πq(T ) and i is
the inclusion map which turns out to be p-strongly q-concave, and so
q-concave.
Observe that what we obtain in Corollary 5.2 is a proper extension for
T , and not just a factorization as the obtained in the Pietsch theorem
for q-summing operators through a subspace of an Lq-space.
/
/
"
"
<
<
18
O. DELGADO AND E. A. S ´ANCHEZ P ´EREZ
References
[1] J. M. Calabuig, O. Delgado and E. A. S´anchez P´erez, Factorizing operators
on Banach function spaces through spaces of multiplication operators, J. Math.
Anal. Appl. 364 (2010), 88-103.
[2] J. M. Calabuig, J. Rodr´ıguez and E. A. S´anchez P´erez, Strongly embedded
subspaces of p-convex Banach function spaces, Positivity 17 (2013), 775-791.
[3] A. Defant, Variants of the Maurey-Rosenthal theorem for quasi Kothe function
spaces, Positivity 5 (2001), 153-175.
[4] A. Defant and E. A. S´anchez P´erez, Maurey-Rosenthal factorization of positive
operators and convexity, J. Math. Anal. Appl. 297 (2004), 771-790.
[5] O. Delgado and E. A. S´anchez P´erez, Summability properties for multiplication
operators on Banach function spaces, Integr. Equ. Oper. Theory 66 (2010),
197-214.
[6] J. Diestel, H. Jarchow and A. Tonge, Absolutely Summing Operators, Cam-
bridge University Press, Cambridge, 1995.
[7] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces II, Springer-Verlag,
Berlin, 1979.
[8] S. Okada, W. J. Ricker and E. A. S´anchez P´erez, Optimal Domain and Integral
Extension of Operators acting in Function Spaces, Operator Theory: Adv.
Appl., vol. 180, Birkhauser, Basel, 2008.
[9] E. A. S´anchez P´erez, Factorization theorems for multiplication operators on
Banach function spaces, Integr. Equ. Oper. Theory 80 (2014), 117-135.
[10] A. C. Zaanen, Integration, 2nd rev. ed., North-Holland, Amsterdam, 1967.
Departamento de Matem´atica Aplicada I, E. T. S. de Ingenier´ıa de
Edificaci´on, Universidad de Sevilla, Avenida de Reina Mercedes, 4 A,
Sevilla 41012, Spain
E-mail address: [email protected]
Instituto Universitario de Matem´atica Pura y Aplicada, Univer-
sitat Polit`ecnica de Val`encia, Camino de Vera s/n, 46022 Valencia,
Spain.
E-mail address: [email protected]
|
1608.03699 | 1 | 1608 | 2016-08-12T07:47:28 | Comparing the generalized roundness of metric spaces | [
"math.FA"
] | Motivated by the local theory of Banach spaces we introduce a notion of finite representability for metric spaces. This allows us to develop a new technique for comparing the generalized roundness of metric spaces. We illustrate this technique in two different ways by applying it to Banach spaces and metric trees. In the realm of Banach spaces we obtain results such as the following: (1) if $\mathcal{U}$ is any ultrafilter and $X$ is any Banach space, then the second dual $X^{\ast\ast}$ and the ultrapower $(X)_{\mathcal{U}}$ have the same generalized roundness as $X$, and (2) no Banach space of positive generalized roundness is uniformly homeomorphic to $c_{0}$ or $\ell_{p}$, $2 < p < \infty$. Our technique also leads to the identification of new classes of metric trees of generalized roundness one. In particular, we give the first examples of metric trees of generalized roundness one that have finite diameter. These results on metric trees provide a natural sequel to a paper of Caffarelli, Doust and Weston. In addition, we show that metric trees of generalized roundness one possess special Euclidean embedding properties that distinguish them from all other metric trees. | math.FA | math |
Comparing the generalized roundness of metric spaces
Lukiel Levy-Moore, Margaret Nichols, and Anthony Weston
Abstract. Motivated by the local theory of Banach spaces we introduce a notion of finite representability
for metric spaces. This allows us to develop a new technique for comparing the generalized roundness of
metric spaces. We illustrate this technique in two different ways by applying it to Banach spaces and metric
trees.
In the realm of Banach spaces we obtain results such as the following: (1) if U is any ultrafilter
and X is any Banach space, then the second dual X ∗∗ and the ultrapower (X)U have the same generalized
roundness as X, and (2) no Banach space of positive generalized roundness is uniformly homeomorphic to c0
or ℓp, 2 < p < ∞. Our technique also leads to the identification of new classes of metric trees of generalized
roundness one. In particular, we give the first examples of metric trees of generalized roundness one that
have finite diameter. These results on metric trees provide a natural sequel to a paper of Caffarelli et al. [6].
In addition, we show that metric trees of generalized roundness one possess special Euclidean embedding
properties that distinguish them from all other metric trees.
Direct calculation of the generalized roundness of an infinite metric space is, in general, a difficult task.
In this paper we develop a versatile technique for comparing the generalized roundness of metric spaces.
This leads to substantial new insights into the generalized roundness of Banach spaces and metric trees.
1. Introduction
Definition 1.1. The generalized roundness of a metric space (X, d), denoted by ℘(X,d) or simply ℘X ,
is the supremum of the set of all p ≥ 0 that satisfy the following condition: for all integers k ≥ 2 and all
choices of (not necessarily distinct) points a1, . . . , ak, b1, . . . , bk ∈ X, we have
X1≤i<j≤k
{d(ai, aj)p + d(bi, bj)p} ≤ X1≤i,j≤k
d(ai, bj)p.
(1.1)
The configuration of points Dk = [a1, . . . , ak; b1, . . . , bk] ⊆ X underlying (1.1) will be called a simplex in X.
We will say that p ≥ 0 is a generalized roundness exponent for (X, d) if (1.1) holds for every simplex in X.
The key idea of this paper is to take an indirect approach to the calculation of ℘(X,d) that is especially
well-suited to the analysis of infinite metric spaces.
The notion of generalized roundness was introduced by Enflo [8] to study universal uniform embedding
spaces. By showing that such spaces must have generalized roundness zero, Enflo was able to prove that
Hilbert spaces are not universal uniform embedding spaces. This resolved a prominent question of Smirnov.
Sometime later, Lennard et al. [25] exhibited an important connection between generalized roundness and
the classical isometric embedding notion of negative type. Lafont and Prassidis [23] used this connection
to show that if a finitely generated group Γ has a Cayley graph of positive generalized roundness, then Γ
must satisfy the coarse Baum-Connes conjecture, and hence the strong Novikov conjecture. The interplay
between these notions has a very interesting history. An overview is given by Prassidis and Weston [30].
The set of generalized roundness exponents of a given metric space (X, d) is always a closed interval of
the form [0, ℘] or [0,∞), including the possibility that ℘ = 0 in which case the interval degenerates to {0}.
This result is a direct consequence of Schoenberg [34, Theorem 2.7] and Lennard et al. [25, Theorem 2.4].
Faver et al. [10] have shown that the interval [0,∞) arises if and only if d is an ultrametric. For finite metric
spaces it is always the case that ℘ > 0. This is the main result in Weston [36].
2010 Mathematics Subject Classification. 46B07, 46B80, 05C05, 05C12.
Key words and phrases. Generalized roundness, negative type, uniform homeomorphism, scale isomorphism.
1
Enflo [8] constructed a separable metric space that is not uniformly embeddable in any metric space of
positive generalized roundness. Dranishnikov et al. [7] modified Enflo's example to construct a locally finite
metric space that is not coarsely embeddable in any Hilbert space, thereby settling a prominent question
of Gromov. Kelleher et al. [19] unified these examples to construct a locally finite metric space that is
not uniformly or coarsely embeddable in any metric space of positive generalized roundness. One may also
use generalized roundness as a highly effective isometric invariant by exploiting the connection between
generalized roundness and negative type due to Lennard et al. [25]. A general principle for using generalized
roundness as an isometric invariant was recently isolated by Kelleher et al. [20, Theorem 3.24]. It is therefore
a matter of great utility to be able to calculate the generalized roundness of certain metric spaces.
In recent work, S´anchez [33] has provided a method of calculating, at least numerically, the generalized
roundness of a given finite metric space (X, d). However, as the size of the space grows, S´anchez' method
rapidly becomes computationally intensive. Nevertheless, the method is an important tool for the analysis
of the generalized roundness of finite metric spaces. In [33], the method is used to calculate the generalized
roundness of certain finite graphs endowed with the usual combinatorial metric. The metric graphs that we
consider in this paper are countable metric trees and so we are unable to use S´anchez' method.
It is prudent at this point to pin down some basic definitions pertaining to metric graphs. A graph G
is connected if there is a (finite) path between any two vertices of G. A tree is an undirected, connected,
locally finite graph without cycles. These definitions imply that the vertex and edge sets of a tree are at
most countable. Assigning a positive length to each edge of a given tree T induces a shortest path metric d
on the vertices of the tree. The resulting metric space is denoted (T, d) and is called a metric tree.
Generalized roundness properties of metric trees have been studied by several authors. All additive
metric spaces, and hence all metric trees, have generalized roundness at least one. This fact is folklore and it
may be derived in several different ways. One such proof appears in Kelly [21, Theorem II]. Another proof
follows from Faver et al. [10, Proposition 4.1]. Examples of Caffarelli et al. [6] show that some countable
metric trees have generalized roundness exactly one. The situation is different for finite metric trees. Indeed,
Hjorth et al. [14] have shown that all finite metric trees have strict 1-negative type. This condition ensures
that all finite metric trees have generalized roundness greater than one. (One way to see this is to appeal to
Lennard et al. [25, Theorem 2.4] and Li and Weston [27, Corollary 4.2].) Hence metric trees of generalized
roundness one are necessarily countable. Simple examples show that the converse of this statement is not
true in general.
We conclude this introduction with some comments about the structure and main results of this paper.
In Section 2, motivated by the local theory of Banach spaces, we introduce a notion of finite representability
for metric spaces. Our purpose in introducing such a notion is to provide a new technique for comparing
the generalized roundness of metric spaces. The remainder of Section 2 is then devoted to a preliminary
investigation of this technique in the context of infinite-dimensional Banach spaces. We prove, for example,
that if U is any ultrafilter and X is any Banach space, then the second dual X∗∗ and the ultrapower (X)U
have the same generalized roundness as X. In other words, ℘X = ℘X ∗∗ = ℘(X)U . It is also noted that no
Banach space of positive generalized roundness is uniformly homeomorphic to c0 or ℓp, 2 < p < ∞.
Caffarelli et al. [6] identified several classes of metric trees of generalized roundness one. The types of
trees studied in [6] were spherically symmetric, infinitely bifurcating or comb-like trees endowed with the
usual combinatorial path metric. In other words, all edges in the trees were assumed to have length one and
all other distances were determined geodesically. In Sections 3 and 4 we relax this condition by considering
trees endowed with weighted path metrics. Section 3 focusses on trees that resemble jagged combs. Section 4
deals with spherically symmetric trees that have systematically weighted edges. We also make a distinction
between convergent and divergent spherically symmetric trees. In both cases we show that the generalized
roundness of such trees can easily be one. In particular, we identify a large class of metric trees of generalized
roundness one that have finite diameter.
In Section 5 we examine isometric embedding properties of metric trees of generalized roundness one.
We prove that all metric trees of generalized roundness one possess the stronger property of strict 1-negative
type. Due to the relationship between generalized roundness and negative type, it also follows that no metric
tree of generalized roundness one has p-negative type for any p > 1. Taken together, these facts imply the
following embedding phase transition: If (T, d) is a metric tree of generalized roundness one, then (1) the
metric transform (T,√d) is isometric to an affinely independent subset of ℓ2, and (2) the metric transform
2
(T,√dp) does not embed isometrically into ℓ2 for any p, 1 < p ≤ 2. Moreover, the only metric trees that
satisfy condition (2) are those of generalized roundness one.
2. Comparing the generalized roundness of metric and Banach spaces
In this section we develop a technique for comparing the generalized roundness of metric spaces. In order
to do this we introduce a metric space version of the Banach space notion of finite representability. This
important notion in the local theory of Banach spaces was introduced by James [16, 17]. Throughout this
section, all Banach spaces are assumed to be real and infinite-dimensional unless noted otherwise. The first
and second duals of a Banach space X are denoted by X∗ and X∗∗, respectively. All Lp-spaces are assumed
to be commutative unless noted otherwise.
Definition 2.1. Let X and X′ be Banach spaces.
(1) X is crudely represented in X′ if there exists an ε0 > 0 such that for each finite-dimensional
subspace E ⊂ X there exists a finite-dimensional subspace F ⊂ X′ (with dim E = dim F ) and a
one-to-one linear mapping T : E → F that satisfies kTkkT −1k ≤ 1 + ε0.
(2) X is finitely represented in X′ if for each ε > 0 and each finite-dimensional subspace E ⊂ X there
exists a finite-dimensional subspace F ⊂ X′ (with dim E = dim F ) and a one-to-one linear mapping
T : E → F that satisfies
(1 − ε)kxk ≤ kT xk ≤ (1 + ε)kxk
(2.1)
for all x ∈ E.
It is easy to see that an equivalent reformulation of the condition given in Definition 2.1 is the following:
for each ε > 0 and each finite-dimensional subspace E ⊂ X there exists a finite-dimensional subspace F ⊂ X′
(with dim E = dim F ) and a one-to-one linear mapping T : E → F that satisfies kTkkT −1k ≤ 1 + ε. While
this reformulation makes the the relationship between crude and finite representability plain, the metric
nature of Definition 2.1 (2) suits our purposes, not least because it motivates Definitions 2.2 and 2.3 below.
The notion of crude representability is particularly important in the uniform theory of Banach spaces.
Recall that two Banach spaces X and X′ are said to be uniformly homeomorphic if there exists a bijection
f : X → X′ such that f and f−1 are both uniformly continuous. A famous result of Ribe [31] asserts that
if a Banach space X is uniformly homeomorphic to a Banach space X′, then X is crudely represented in X′
and X′ is crudely represented in X. This result is sometimes known as Ribe's rigidity theorem.
A one-to-one linear mapping T : E → F that satisfies condition (2.1) is said to be a (1 + ε)-isomorphism.
A similar notion for metric spaces may be formulated as follows.
Definition 2.2. Let (X, d) and (X′, ρ) be metric spaces, and suppose that ε > 0. A one-to-one mapping
φ : X → X′ : x 7→ x′ is called a (1 + ε)-scale isomorphism if there exists a constant n = n(ε) > 0 such that
(1 − ε)nd(a, b) ≤ ρ(a′, b′) ≤ (1 + ε)nd(a, b)
for all a, b ∈ X.
It is worth noting that we will use the notation x′ to denote φ(x) throughout this section.
Definition 2.3. A metric space (X, d) is said to be locally represented in a metric space (X′, ρ) if for each
ε > 0 and each non-empty finite set X ♯ ⊆ X there exists a (1 + ε)-scale isomorphism φ : (X ♯, d) → (X′, ρ).
The following lemma notes that for Banach spaces, finite representation implies local representation.
Lemma 2.4. Let X and X′ be given Banach spaces. If X is finitely represented in X′, then X is locally
represented in X′.
Proof. Let X ♯ be a given non-empty finite subset of X and suppose that ε > 0. Let E denote the
linear span of X ♯ in X. Then E is a finite-dimensional subspace of X. As X is finitely represented in X′,
there exists a (1 + ε)-isomorphism T : E → X′. Setting φ to be the restriction of T to X ♯ we obtain a
(1 + ε)-scale isomorphism X ♯ → X′ (with constant n = 1). Hence then X is locally represented in X′. (cid:3)
We turn now to the main technical result of this section. It provides a new technique for comparing the
generalized roundness of metric spaces.
3
Theorem 2.5. If a metric space (X, d) is locally represented in a metric space (X′, ρ), then every
generalized roundness exponent of (X′, ρ) is a generalized roundness exponent of (X, d). Hence, ℘X ′ ≤ ℘X .
Proof. It suffices to prove that if p is not a generalized roundness exponent of (X, d) then p is not a
generalized roundness of (X′, ρ).
Suppose that p ≥ 0 is not a generalized roundness exponent of (X, d). We immediately have that p > 0
because 0 is a generalized roundness exponent of all metric spaces. From our definition there must be a
simplex [ai; bj] ⊆ X such that
Xi<j
(d(ai, aj)p + d(bi, bj)p) > Xi,j
d(ai, bj)p.
The limiting behavior y = xp in a neighborhood of x = 1 then ensures that we may choose an ε > 0 so that
(1 − ε)p ·Xi<j
(d(ai, aj)p + d(bi, bj)p) > (1 + ε)p ·Xi,j
d(ai, bj)p.
(2.2)
We now let X ♯ denote the finite subset of X that consists of the simplex points ai, bj. As (X, d) is
locally represented in (X′, ρ) and ε > 0, there must exist an injection φ : X ♯ → X′ : x 7→ x′ and a constant
n = n(ε) > 0 such that
(1 − ε)nd(a, b) ≤ ρ(a′, b′) ≤ (1 + ε)nd(a, b)
for all a, b ∈ X ♯. Furthermore, if we scale the metric on X′ by defining ω = ρ/n, we immediately obtain
(2.3)
for all a, b ∈ X ♯. It now follows from (2.2) and (2.3) that p is not a generalized roundness exponent for the
scaled metric space (X′, ω). Indeed,
(1 − ε)d(a, b) ≤ ω(a′, b′) ≤ (1 + ε)d(a, b)
Xi<j (cid:0)ω(a′i, a′j)p + ω(b′i, b′j)p(cid:1) ≥ (1 − ε)p ·Xi<j
> (1 + ε)p ·Xi,j
= Xi,j
≥ Xi,j
ω(a′i, b′j)p.
((1 + ε)d(ai, bj))p
(d(ai, aj)p + d(bi, bj)p)
d(ai, bj)p
This completes the proof because generalized roundness is preserved under any scaling of the metric ρ. (cid:3)
For the remainder of this section we will consider the application of Theorem 2.5 to Banach spaces. It is
germane to recall a few facts about the generalized roundness of Lp-spaces. If X is an Lp-space, then ℘X = p
if 1 ≤ p ≤ 2 and ℘X = 0 if p > 2. These results are due to Enflo [8] in the case 1 ≤ p ≤ 2 and Lennard
et al. [25] in the case p > 2. With the exception of the Schatten p-classes Cp, the generalized roundness of
non-commutative Lp-spaces has not been widely studied. In [25] the authors noted that ℘Cp = 0 if p > 2.
It is only relatively recently that Dahma and Lennard [3] have shown that ℘Cp = 0 if 0 < p < 2.
Corollary 2.6. If a Banach space X is finitely represented in a Banach space X′, then ℘X ′ ≤ ℘X . In
particular, if ℘X = 0, then ℘X ′ = 0.
Proof. Immediate from Lemma 2.4 and Theorem 2.5.
(cid:3)
Examples of Banach spaces that have generalized roundness zero include C[0, 1], ℓ∞, c0, ℓp if p > 2, and
the Schatten p-class Cp if p 6= 2. For each Banach space X and each integer n ≥ 2, Dineen [5] has shown
that ℓ∞ is finitely represented in the space P(nX) of bounded n-homogenous polynomials on X. Hence
P(nX) has generalized roundness zero by Corollary 2.6. For each p ∈ (1,∞), c0 is finitely represented in the
quasi-reflexive James space Jp. (This result is due to Giesy and James [11] in the case p = 2 and, for p 6= 2,
it is due to Bird et al. [1].) Hence, for each p ∈ (1,∞), Jp has generalized roundness zero by Corollary 2.6.
On the basis of existing theory and Corollary 2.6 we are able to isolate some situations where generalized
roundness functions as an invariant in the uniform theory of Banach spaces. For instance, as the next corollary
shows, no Banach space of positive generalized roundness is uniformly homeomorphic to ℓp for any p > 2.
4
Corollary 2.7. If a Banach space X is uniformly homeomorphic to ℓp (1 ≤ p < ∞), then ℘X ≤ ℘ℓp .
In particular, if p > 2, then ℘X = 0.
Proof. Ribe's rigidity theorem [31] implies that ℓp is crudely represented in X. However, if ℓp is
crudely represented in X, then ℓp is finitely represented in X. This follows from Krivine's theorem [22] (as
noted by Rosenthal [32] and Lemberg [24]) if 1 < p < ∞, and it is due to James [15] in the case p = 1.
Thus ℘X ≤ ℘ℓp by Corollary 2.6. Moreover, if p > 2, then ℘ℓp = 0. So for p > 2 we deduce that ℘X = 0. (cid:3)
The uniform structure of ℓp, 1 < p < ∞, is particularly well-understood. For instance, if a Banach space
X is uniformly homeomorphic to ℓp, 1 < p < ∞, then it is linearly isomorphic to ℓp. This deep theorem is
due to Enflo [9, Theorem 6.3.1] in the case p = 2 and Johnson et al. [18, Theorem 2.1] when p 6= 2. So if
1 < p < ∞, one may replace the phrase "uniformly homeomorphic" in the statement of Corollary 2.7 with
the phrase "linearly isomorphic" without losing any generality. The situation for c0 is somewhat similar.
Corollary 2.8. If a Banach space X is uniformly homeomorphic to c0, then ℘X = 0. In particular,
no Banach space of positive generalized roundness is uniformly homeomorphic to c0.
Proof. Ribe's rigidity theorem [31] implies that c0 is crudely represented in X. However, James [15]
has shown that if c0 is crudely represented in X, then c0 is finitely represented in X. Thus ℘X ≤ ℘c0 by
Corollary 2.6. Moreover, as ℘c0 = 0, we further deduce that ℘X = 0.
(cid:3)
The uniform structure of c0 is more beguiling and less well understood than that of ℓp, 1 < p < ∞.
Johnson et al. [18, Corollary 3.2] proved that if a complemented subspace of a C(K) space is uniformly
homeomorphic to c0, then it is linearly isomorphic to c0. Godefroy et al. [12, Theorem 5.6] have shown that
a Banach space which is uniformly homeomorphic to c0 is an isomorphic predual of ℓ1 with summable Szlenk
index. But it is not known whether a predual of ℓ1 with summable Szlenk index is linearly isomorphic to c0.
Thus, unlike ℓp (1 < p < ∞), it remains unclear whether c0 is determined by its uniform structure.
central to the local theory of Banach spaces and it is originally due to Lindenstrauss and Rosenthal [28].
In the proof of the following corollary we invoke the Principle of Local Reflexivity. This principle is
Corollary 2.9. If X is a Banach space, then ℘X = ℘X ∗∗ and ℘X ∗ = ℘X ∗∗∗.
Proof. As X embeds isometrically into X∗∗ we see that ℘X∗∗ ≤ ℘X . In addition, the Principle of Local
Reflexivity implies that X∗∗ is finitely represented in X. Hence ℘X ≤ ℘X ∗∗ by Corollary 2.6. By combining
these two inequalities we obtain ℘X = ℘X ∗∗. By replacing X with X∗ we also see that ℘X ∗ = ℘X ∗∗∗.
(cid:3)
Examples show that for a Banach space X we may have ℘X 6= ℘X ∗ . Indeed, if p ∈ [1, 2) and X = ℓp,
then ℘X = p and ℘X ∗ = 0. By way of comparison, if p 6= 2 and X = Cp, then ℘X = ℘X ∗ = 0. Thus, given
Banach space X, the entries of the sequence (℘X , ℘X ∗ , ℘X ∗∗, . . .) are restricted to take on at most two values
(in the interval [0, 2]) by Corollary 2.9.
Intimately related to the concept of finite representability is the notion of an ultrapower of a Banach
space. Given an ultrafilter U on a set I and a Banach space X there is a canonical procedure to construct a
large Banach space (X)U called the ultrapower of X. Importantly, (X)U contains a natural isometric copy of
X and it is finitely represented in X. For a detailed construction of (X)U , and a discussion of the interplay
between finite representability and ultrapowers, we refer the reader to H´ajek and Johanis [13].
Corollary 2.10. Let U be a given ultrafilter on a set I and let X be a Banach space. Then ℘X = ℘(X)U .
Proof. As X embeds isometrically into (X)U we see that ℘(X)U ≤ ℘X . In addition, (X)U is finitely
represented in X by Stern [35, Theorem 6.6]. Hence ℘X ≤ ℘(X)U by Corollary 2.6. By combining these two
inequalities we obtain ℘X = ℘(X)U .
(cid:3)
Lennard et al. [26, Theorem 2.3] noticed that if the infimal cotype of a Banach space X is greater than
two, then X must have generalized roundness zero. By utilizing deep theory and Corollary 2.6 we are able to
exhibit a more precise relationship between the supremal type and infimal cotype of a Banach space and its
generalized roundness. The notions of type and cotype have been paramount in the local theory of Banach
spaces for quite some time and are defined in the following manner.
5
Definition 2.11. A Banach space X is said to have type p if there exists a constant A ∈ (0,∞) such
that for all integers n > 0 and for all finite sequences (xj )n
j=1 in X, we have
Xǫ∈{−1,+1}n
n
Xj=1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ǫjxj
2n (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X
≤ A
n
Xj=1
X
kxjkp
1
p
.
(2.4)
Cotype p is defined similarly but with the inequality (2.4) reversed.
It is well-known that no Banach space can have type p > 2 or cotype q < 2. We let p(X) denote the
supremum of all p such that X has type p and q(X) denote the infimum of all q such that X has cotype q.
For an overview of theory of type and cotype we refer the reader to Diestel et al. [4].
A famous theorem of Maurey and Pisier [29] states that ℓp(X) and ℓq(X) are finitely represented in X.
This theorem and Corollary 2.6 provide an immediate link to generalized roundness.
Corollary 2.12. If X is a Banach space, then ℘X ≤ min{℘ℓp(X), ℘ℓq(X)}. In particular, if q(X) > 2,
then ℘X = 0.
Proof. By the Maurey-Pisier theorem, ℓp(X) and ℓq(X) are finitely represented in X. Hence ℘X ≤ ℘ℓp(X)
and ℘X ≤ ℘ℓq(X) by Corollary 2.6. In particular, if q(X) > 2, then ℘ℓq(X) = 0, and so ℘X = 0.
(cid:3)
There are some classical Banach spaces for which the inequality in Corollary 2.12 is an equality. For
example, if X is an Lp-space, 1 ≤ p < ∞, then ℘X = min{℘ℓp(X), ℘ℓq(X)}. In this case, ℘X = p if 1 ≤ p ≤ 2
and ℘X = 0 if p > 2. Moreover, it is well known that p(X) = min{p, 2} and q(X) = max{p, 2}. So, for
example, if p > 2, then ℘X = 0 and q(x) = p. Thus ℘ℓq(X) = 0. On the other hand, if X = Cp, 1 ≤ p < ∞,
then p(X) and q(X) have the same values as any Lp-space but, by inspection, ℘X = min{℘ℓp(X), ℘ℓq(X)} if
and only if p ≥ 2.
3. Comb-like graphs of generalized roundness one
In this section we apply Theorems 2.5 and 3.3 to analyze the generalized roundness of countable metric
trees that resemble combs. We first give sufficient conditions for the existence of a (1 + ε)-scale isomorphism
φ : (T, d) → (T, ρ), under the assumption that d and ρ are path weighted metrics on a given finite tree T .
In what follows, we let N denote the set of all non-negative integers. Moreover, given a positive integer m,
we let [m] denote the segment {0, 1, 2, . . . , m}.
Lemma 3.1. Let d and ρ be two path weighted metrics on a given finite tree T . Let
m = max(cid:26) ρ(a, b)
d(a, b)
: a, b ∈ T and a 6= b(cid:27) .
Then there must be an edge {x, y} in T such that m = ρ(x,y)
d(x,y) .
Proof. Suppose that a, c ∈ T are non-adjacent vertices such that m = ρ(a, c)/d(a, c). Then we may
choose a strictly intermediate vertex b ∈ T on the geodesic from a to c. Now let q = ρ(a, b)/d(a, b) and
r = ρ(b, c)/d(b, c). Without loss of generality we may assume that q ≥ r. Furthermore, as ρ and d are path
metrics on T , we have ρ(a, c) = ρ(a, b) + ρ(b, c) and d(a, c) = d(a, b) + d(b, c). In particular, it follows that
m =
=
=
ρ(a, c)
d(a, c)
ρ(a, b) + ρ(b, c)
d(a, b) + d(b, c)
qd(a, b) + rd(b, c)
d(a, b) + d(b, c)
qd(a, b) + qd(b, c)
d(a, b) + d(b, c)
≤
= q.
Therefore, by definition of m, it must be the case that m = q. This shows that we can always pass to a pair
of vertices connected by a geodesic with fewer edges and preserve the ratio m. Applying this logic finitely
many times gives the lemma.
(cid:3)
6
The following analogous lemma for minima may be proved in the same way.
Lemma 3.2. Let d and ρ be two path weighted metrics on a given finite tree T . Let
m∗ = min(cid:26) ρ(a, b)
d(a, b)
: a, b ∈ T and a 6= b(cid:27) .
Then there must be an edge {x, y} in T such that m∗ = ρ(x,y)
d(x,y).
Theorem 3.3. Let ε > 0 be given. Let d and ρ be two path weighted metrics on a given finite tree T . If
there exists a constant n = n(ε) > 0 such that
(1 − ε)n ≤
ρ(x, y)
d(x, y) ≤ n(1 + ε)
(3.1)
for each edge {x, y} in T , then the identity map φ : (T, d) → (T, ρ) : x 7→ x is a (1 + ε)-scale isomorphism.
Proof. Using the notation of Lemmas 3.1 and 3.2 it follows from (3.1) that
(1 − ε)n ≤ m∗ ≤ m ≤ n(1 + ε).
Thus, given any two distinct vertices a, b ∈ T , we deduce that (1 − ε)n ≤ ρ(a, b)/d(a, b) ≤ n(1 + ε) by
definition of m and m∗. Hence the identity map φ : (T, d) → (T, ρ) : x 7→ x is a (1+ε)-scale isomorphism. (cid:3)
We now apply Theorems 2.5 and 3.3 to analyze the generalized roundness of certain countable metric
trees that resemble combs.
Definition 3.4. The vertex set V of the infinite comb C consists of the points x0, xk and yk, where k
is any positive integer. The edge set E of C consists of the unordered pairs {xk−1, xk} and {xk, yk}, where
k is any positive integer.
For each positive integer m, the finite subtree of C that has vertex set {x1+k, y1+k k ∈ [m]} will be
called the m-comb. The m-comb will be denoted Cm.
We are interested in placing various path metrics on the infinite comb C and the m-comb Cm. One way
to do this is to adopt the following canonical procedure.
Definition 3.5. Let f : N → (0,∞) be a function. We define a path metric ρf on the infinite comb C
in the following manner:
(1) ρf (xk−1, xk) = f (k − 1), and
(2) ρf (xk, yk) = f (k) for each positive integer k.
All other distances in C are then determined geodesically. The resulting metric tree will be denoted C(f ).
If, moreover, we restrict ρf to the m-comb, the resulting metric tree will be denoted Cm(f ).
There are some special cases of Definition 3.5 worth highlighting. If f (k) = 1 for all k ≥ 0, the resulting
metric trees C(f ) and Cm(f ) will be denoted C(1) and Cm(1), respectively. In other words, C(1) and Cm(1)
are the combs C and Cm endowed with the usual combinatorial path metric δ. Caffarelli et al. [6] have shown
that ℘Cm(1) → 1 as m → ∞. Hence ℘C(1) = 1. We will see presently that by placing mild assumptions on
the function f it follows that ℘C(f ) = 1. Our arguments will be facilitated by the following lemma.
Lemma 3.6. If the m-comb Cm(1) is locally represented in a metric tree (T, d) for all integers m > 0,
then ℘(T,d) = 1.
Proof. All metric trees have generalized roundness at least one. By Theorem 2.5, ℘(T,d) ≤ ℘Cm(1) for
all m > 0. Moreover, Caffarelli et al. [6] have shown that ℘Cm(1) → 1 as m → ∞. Hence ℘(T,d) = 1.
(cid:3)
Definition 3.7. A function f : N → (0,∞) is said to be additively sub-exponential if
for each integer m > 0.
f (n+m)
f (n) = 1
lim
n→∞
The class of additively sub-exponential functions f : N → (0,∞) is very large. For instance, f could be
any rational function that takes positive values on N. Other interesting possibilities for f include inverse
tangent, logarithmic functions (translated suitably), and classically sub-exponential functions such as e√n.
Furthermore, if f : N → (0,∞) is an additively sub-exponential function, then so is 1/f .
Theorem 3.8. If f : N → (0,∞) is additively sub-exponential function, then ℘C(f ) = 1.
7
Proof. By Lemma 3.6, it suffices to prove that Cm(1) is locally represented in C(f ) for all m > 0.
Let m > 0 be a given integer. Let ε > 0 be given. As f is additively sub-exponential, we may choose an
integer n0 > 0 so that 1 − ε ≤ f (n + k)/f (n) ≤ 1 + ε for each k ∈ [m] and all n ≥ n0. In particular, we have
f (n0 + k)
1 − ε ≤
f (n0) ≤ 1 + ε
for each k ∈ [m]. Consider the subtree Y ′ of C(f ) that has vertices xn0+k and yn0+k for all k ∈ [m]. As
simple (unweighted) graphs, Y ′ and Cm(1) are one and the same graph; namely, Cm. Let φ : Cm(1) → Y ′
denote this natural identification. Let ρ denote the path metric that Y ′ inherits from C(f ). We may regard
the metrics on Y ′ and Cm(1) as path metrics on Cm. For each edge {s, t} in Cm we have, by choice of n0,
f (n0)(1 − ε) ≤ min
k∈[m]
f (n0 + k) ≤
ρ(φ(s), φ(t))
δ(s, t)
≤ max
k∈[m]
f (n0 + k) < f (n0)(1 + ε).
It follows from Theorem 3.3 that φ : Cm(1) → Y ′ is a (1 + ε)-scale isomorphism. As Cm(1) is a finite metric
space and as ε > 0 was arbitrary, we conclude that Cm(1) is locally represented in C(f ).
(cid:3)
It also follows from Theorem 3.8 that if f is an additively sub-exponential function, then ℘C(f ) = ℘C(1/f ).
4. Convergent and divergent spherically symmetric trees of generalized roundness one
Caffarelli et al. [6] considered the generalized roundness of spherically symmetric trees endowed with
the usual combinatorial path metric (wherein all edges in the tree are assumed to have unit length).
In
this section we consider a broader class of spherically symmetric trees by relaxing the requirement that all
edges have unit length. This allows one to make a distinction between convergent and divergent spherically
symmetric trees. In order to proceed we review the basic definitions and notations for spherically symmetric
trees. In addition, we introduce the notion of a downward length sequence for a spherically symmetric tree.
Given a vertex v0 in a tree T , we let r(T, v0) = sup{δ(v0, v) : v ∈ T}, where δ denotes the usual
combinatorial path metric on T . We call r(T, v0) the v0-depth of T . Naturally included here is the possibility
that the v0-depth of T may be infinite. A vertex v of T is a level k vertex of T if δ(v0, v) = k. The children
of a level k vertex v ∈ T consist of all level k + 1 vertices w ∈ T such that δ(v, w) = 1. We let dk(v) denote
the number of children of v.
Definition 4.1. A tree T is said to be spherically symmetric if we can choose a vertex v0 ∈ T so that
for any k, all level k vertices of T have the same number of children. Such a pair (T, v0) will be called a
spherically symmetric tree (SST).
Notice that if v is a level k vertex in a given SST (T, v0), then dk = dk(v) only depends upon k. Thus
dk is the number of children of any level k vertex in T . We call the (possibly finite) sequence (dk)0≤k<r(T,v0)
the downward degree sequence of (T, v0). We will say that (dk) is non-trivial provided dk > 1 for at least
one k such that 0 ≤ k < r(T, v0).
Now suppose that (T, v0) is a given SST with downward degree sequence (dk). If, for each k such that
0 ≤ k < r(T, v0), lk is a positive real number, we will call ℓ = (l0, l1, l2, . . .) a downward length sequence
for (T, v0). Given such a sequence ℓ, we may define a path metric ρℓ on T in the following manner. For
any k such that 0 ≤ k < r(T, v0), if w is a child of a level k vertex v ∈ T , we define ρℓ(v, w) = lk. All
other ρℓ-distances in T are then determined geodesically. The resulting metric tree (T, ρℓ) is said to be
convergent if P lk < ∞ and divergent if P lk = ∞. One significance of convergent SSTs is that they have
finite diameter.
We proceed to show that large classes of divergent and convergent SSTs have generalized roundness one.
The following lemma is a variation of [6, Theorem 2.1]. As the statement of the lemma is complicated, we
will comment on the intuition behind this result. Among all n-point metric trees endowed with the usual
combinatorial path metric, the complete bipartite graph K1,n−1 has the smallest generalized roundness.
Moreover, as n → ∞, the generalized roundness of K1,n−1 tends to one. The conditions placed on the SST
in the statement of the following lemma ensure that it contains a star-like structure that resembles K1,q,
q = d0d1 ··· dk, modulo scaling. Such an SST must have generalized roundness relatively close to one.
Lemma 4.2. Let (T, v0) be a finite SST with a non-trivial downward degree sequence (d0, d1, . . . , dn−1).
Suppose ℓ = (l0, l1, . . . , ln−1) is a downward length sequence for (T, v0) that satisfies 2l0 < l0 + l1 +··· + ln−1.
8
For each k, 1 ≤ k ≤ n, set Mk = Pk−1
non-negative integer k ≤ m such that d0d1 ··· dk > 1, we have
i=0 li and let m be the largest integer k such that Mk < 1
2 Mn. For each
℘(T,ρℓ) ≤
ln(cid:16)2 +
2
(d0d1···dk)−1(cid:17)
ln(cid:16)2 − 2Mk
Mn (cid:17)
.
(4.1)
If d0d1 ··· dm = 1, then we have the trivial bound ℘(T,ρℓ) ≤ 2.
Proof. Let (T, v0) be a finite SST that satisfies the hypotheses of the lemma. Because at least one
dj > 1 there must exist at least one vertex z ∈ T with at least two children x, y ∈ T . As children in T are
ρℓ-equidistant from their parent we see that
ρℓ(x, z) =
ρℓ(x, y)
2
= ρℓ(z, y).
The existence of such a metric midpoint in (T, ρℓ) ensures that ℘(T,ρℓ) ≤ 2.
Now assume that d0d1 ··· dm > 1 and consider any non-negative integer k ≤ m such that d0d1 ··· dk > 1.
Then there are d0d1 ··· dk−1 vertices at distance Mk from v0. For each of the dk children of such a vertex,
choose a leaf which is a descendent of that child (or the child itself if it is a leaf). This results in a total of
q = d0d1 ··· dk > 1 distinct leaves which we label a1, a2, . . . , aq. Set bj = v0 for all j such that 1 ≤ j ≤ q.
For all i and j we have ρℓ(ai, bj) = Mn. Moreover, for all i 6= j, we have ρℓ(ai, aj) ≥ 2(Mn − Mk) and
ρℓ(bi, bj) = 0. It follows that any generalized roundness exponent p of (T, ρℓ) must satisfy
1
2
q(q − 1) (2(Mn − Mk))p ≤ Xi<j
{ρℓ(ai, aj)p + ρℓ(bi, bj)p} ≤ Xi,j
ρℓ(ai, bj)p = q2M p
n.
(4.2)
By comparing the left and right sides of (4.2) it follows that p must satisfy:
p ≤
ln(cid:16)2 +
2
(d0d1···dk)−1(cid:17)
ln(cid:16)2 − 2Mk
Mn (cid:17)
.
As p was an arbitrary generalized roundness exponent of (T, ρℓ), we conclude that the lemma holds.
(cid:3)
Theorem 4.3. Let (T, v0) be a countable SST with downward degree sequence (d0, d1, d2, . . .) and down-
ward length sequence ℓ = (l0, l1, l2, . . .). If di > 1 for infinitely many i and P li = ∞, then ℘(T,ρℓ) = 1.
Proof. For each positive integer n let (Tn, v0) denote the finite SST with downward degree sequence
(d0, d1, . . . , dn−1) and downward length sequence ℓ = (l0, l1, . . . , ln−1). For each k, 1 ≤ k ≤ n, set Mk =
Pk−1
i=0 li. As P li = ∞ we will have 2l0 < l0 + l1 + ··· + ln−1 provided n is sufficiently large. Moreover, for
each such integer n, we may choose the largest integer k = k(n) such that Mk ≤ ln Mn. As n → ∞, the
quantities k, Mk, Mn and d0d1 ··· dk all tend to ∞. However, by construction, (2Mk)/Mn → 0 as n → ∞.
Thus, by Lemma 4.2, ℘(Tn,ρℓ) → 1 as n → ∞, and so we conclude that ℘(T,ρℓ) = 1.
(cid:3)
Theorem 4.4. Let f : N → (0,∞) be an additively sub-exponential function. Let (T, v0) be a countable
SST with downward degree sequence (dk). Let ℓ denote the downward length sequence (f (k)). If dk > 1 for
each integer k ≥ 0, then ℘(T,ρℓ) = 1.
Proof. The condition dk > 1 for each integer k ≥ 0 ensures that the infinite comb C(f ) is isometric to
a metric subspace of (T, ρℓ). Thus ℘(T,ρℓ) ≤ ℘C(f ). Moreover, ℘C(f ) = 1 by Theorem 3.8 and ℘(T,ρℓ) ≥ 1,
thereby forcing ℘(T,ρℓ) = 1.
(cid:3)
Theorem 4.4 provides examples of convergent SSTs with generalized roundness one. For instance, we
may simply set dk = 2 and f (k) = (k + 1)−2 for all k ≥ 0 to obtain a countable SST that is convergent and
has generalized roundness one. In particular, such SSTs have finite diameter.
9
5. Embedding properties of metric trees of generalized roundness one
We conclude this paper with some comments on the special Euclidean embedding properties of metric
trees of generalized roundness one that set them apart from all other metric trees. As noted at the outset
of this paper, the notions of generalized roundness and negative type are equivalent. In order to make this
statement more precise we recall the following definition, the roots of which can be traced back to an 1841
paper of Cayley [2].
Definition 5.1. Let p ≥ 0 and let (X, d) be a metric space. Then:
(1) (X, d) has p-negative type if and only if for all integers k ≥ 2, all finite subsets {x1, . . . , xk} ⊆ X,
and all choices of real numbers η1, . . . , ηk with η1 + ··· + ηk = 0, we have
X1≤i,j≤k
d(xi, xj)pηiηj ≤ 0.
(5.1)
(2) (X, d) has strict p-negative type if and only if it has p-negative type and the associated inequalities
(5.1) are all strict except in the trivial case (η1, . . . , ηk) = (0, . . . , 0).
Lennard et al. [25] proved that for all p ≥ 0, a metric space (X, d) has p-negative type if and only if p
is a generalized roundness exponent of (X, d). The significance of this result is that builds a bridge between
Enflo's [8] notion of generalized roundness and classical isometric embedding theory. These connections, in
conjunction with contemporary results on strict negative type, enable the following theorem.
Theorem 5.2. If (T, d) is a metric tree of generalized roundness one, then:
(1) the metric transform (T,√d) is isometric to an affinely independent subset of ℓ2, and
(2) the metric transform (T,√dp) does not embed isometrically into ℓ2 for any p, 1 < p ≤ 2.
Moreover, the only metric trees that satisfy condition (2) are those of generalized roundness one.
Proof. The vertex set of (T, d) is countable because it is a metric tree of generalized roundness one.
Our definitions imply that each finite subset of T is contained in a finite subtree of T . Hjorth et al. [14] have
shown that all finite metric trees have strict 1-negative type. Hence each finite metric subspace of (T, d) has
strict 1-negative type. This ensures that (T, d) has strict 1-negative type. Equivalently, the metric transform
(T,√d) has strict 2-negative type. Therefore (T,√d) is isometric to an affinely independent subset of ℓ2 by
Kelleher et al. [20, Theorem 5.6]. This establishes condition (1).
If the metric transform (T,√dp) were to embed isometrically into ℓ2 for some p ∈ (1, 2], this would imply
that (T, d) has p-negative type. But by Lennard et al. [25], this would mean that p is a generalized roundness
exponent of (T, d), thereby contradicting our assumption that ℘(T,d) = 1. This establishes condition (1).
On the other hand, if a metric tree (Z, d) is not of generalized roundness one, then it must be the case
that ℘(Z,d) > 1 (because 1 is a generalized roundness exponent of all metric trees). By Lennard et al. [25],
this implies that (Z, d) has p-negative type for some p ∈ (1, 2]. Consequently, the metric transform (T,√dp)
embeds isometrically into ℓ2 by Kelleher et al.
that satisfy condition (2) are those of generalized roundness one.
[20, Theorem 5.6]. We conclude that the only metric trees
(cid:3)
The proof of Theorem 5.2 shows that all metric trees have strict 1-negative type. Therefore every metric
tree satisfies Theorem 5.2 (1) by the result of Kelleher et al. [20].
Acknowledgements
The research in this paper was initiated at the 2011 Cornell University Summer Mathematics Institute
(NSF grant DMS-0739338) and completed at the University of South Africa (Unisa). The second named
author was partially supported by the National Science Foundation Graduate Research Fellowship Program
(NSF grant DGE-1144082). We are indebted to the Visiting Researcher Support Programme at Unisa, the
US National Science Foundation, the Department of Mathematics and the Center for Applied Mathematics
at Cornell University for their support.
10
References
[1] A. Bird, G. Jameson, N. J. Laustsen, The Giesy-James theorem for general index p, with an application to operator
ideals on the pth James space, J. Operator Theory 70 (2013), 291–307. 4
[2] A. Cayley, On a theorem in the geometry of position, Cambridge Mathematical Journal II (1841), 267–271. (Also in The
Collected Mathematical Papers of Arthur Cayley (Vol. I), Cambridge University Press, Cambridge (1889), pp. 1–4.) 10
[3] A. M. Dahma and C. J. Lennard, Generalized roundness of the Schatten class, Cp, J. Math. Anal. Appl. 412 (2014),
676–684. 4
[4] J. Diestel, H. Jarchow and A. M. Tonge, Absolutely Summing Operators, Cambridge Studies in Advanced Mathematics
43 (1995), Cambridge U. Press, 1–474. 6
[5] S. Dineen, A Dvoretzky theorem for polynomials, Proc. Amer. Math. Soc. 123 (1995), 2817–2821. 4
[6] I. Doust, E. Caffarelli and A. Weston, Metric trees of generalized roundness one, Aeq. Math. 83 (2012), 239–256. 1, 2, 7,
8
[7] A.N. Dranishnikov, G. Gong, V. Lafforgue and G. Yu, Uniform embeddings into Hilbert space and a question of Gromov,
Canad. Math. Bull. 45 (2002), 60–70. 2
[8] P. Enflo, On a problem of Smirnov, Ark. Mat. 8 (1969), 107–109. 1, 2, 4, 10
[9] P. Enflo, Uniform Structures and Square Roots in Topological Spaces II, Israel J. Math. 8 (1970), 253–272. 5
[10] T. Faver, K. Kochalski, M. Murugan, H. Verheggen, E. Wesson and A. Weston, Roundness properties of ultrametric
spaces, Glasgow Math. J. 56 (2014), 519–535. 1, 2
[11] D. P. Giesy and R. C. James, Uniformly non-ℓ1 and B-convex Banach spaces, Studia Math. 48 (1973), 61–69. 4
[12] G. Godefroy, N. J. Kalton and G. Lancien, Szlenk Indices and uniform homeomorphisms, Trans. Amer. Math. Soc. 353
(2001), 3895–3918. 5
[13] P. H´ajek and M. Johanis, Smooth Analysis in Banach Spaces, De Gruyter Series in Nonlinear Analysis and Applications
19 (2014), De Gruyter, 1–497. 5
[14] P. Hjorth, P. Lisonek, S. Markvorsen and C. Thomassen, Finite metric spaces of strictly negative type, Linear Algebra
Appl. 270 (1998), 255–273. 2, 10
[15] R. C. James, Uniformly nonsquare Banach spaces, Ann. of Math. 80 (1964), 542–550. 5
[16] R. C. James, Some self-dual properties of normed linear spaces, Sympos. Infinite Dimensional Topology, Ann. of Math.
Studies, no. 69, Princeton Univ. Press, Princeton, N. J., 1972. 3
[17] R. C. James, Super-reflexive Banach spaces, Canad. J. Math. 24 (1972), 896–904. 3
[18] W. B. Johnson, J. Lindenstrauss and G. Schechtman, Banach spaces determined by their linear structure, GAFA 6 (1996),
430–470. 5
[19] C. Kelleher, D. Miller, T. Osborn and A. Weston, Strongly non-embeddable metric spaces, Topology Appl. 159 (2011),
749–755. 2
[20] C. Kelleher, D. Miller, T. Osborn and A. Weston, Polygonal equalities and virtual degeneracy in Lp-spaces, J. Math.
Anal. Appl. 415 (2014), 247–268. 2, 10
[21] J. B. Kelly, Hypermetric spaces, in: Proc. Conf. Geometry of Metric and Linear Spaces, Lecture Notes in Math. 490
(Springer-Verlag, Berlin, 1975), 17–31. 2
[22] J. L. Krivine, Sous-espaces de dimension finie des espaces de Banach r´eticul´e, Ann. of Math. 104 (1976), 1–29. 5
[23] J-F. Lafont and S. Prassidis, Roundness properties of groups, Geom. Ded. 117 (2006), 137–160. 1
[24] H. Lemberg, Nouvelle d´emonstration d'un th´eor`eme de J. L. Krivine sur la finie repr´esentation de ℓp dans un espace de
Banach, Israel J. Math 39 (1981), 341–348. 5
[25] C. J. Lennard, A. M. Tonge and A. Weston, Generalized roundness and negative type, Michigan Math. J. 44 (1997),
37–45. 1, 2, 4, 10
[26] C. J. Lennard, A. M. Tonge and A. Weston, Roundness and metric type, J. Math. Anal. Appl. 252 (2000), 980–988. 5
[27] H. Li and A. Weston, Strict p-negative type of a metric space, Positivity 14 (2010), 529–545. 2
[28] J. Lindenstrauss and H. Rosenthal, The L
[29] B. Maurey and G. Pisier, S´eries de variables al´eatoires vectorielles ind´ependantes et propri´et´es g´eom´etriques des espaces
p spaces, Israel J. Math. 7 (1969), 325–349. 5
de Banach, Studia Math. 58 (1976), 45–90. 6
[30] E. Prassidis and A. Weston, Manifestations of non linear roundness in analysis, discrete geometry and topology. In:
Arzhantseva, G., Valette, A. (eds.) Limits of Graphs in Group Theory and Computer Science. Research Proceedings of
the ´Ecole Polytechnique F´ed´erale de Lausanne. CRC Press, Boca Raton (2009). 1
[31] M. Ribe, On uniformly homeomorphic normed spaces, Ark. Mat. 14 (1976), 237–244. 3, 5
[32] H. P. Rosenthal, On a theorem of J. L. Krivine concerning block finite representability of ℓp in general Banach spaces,
J. Funct. Anal. 28 (1978), 197–225. 5
[33] S. S´anchez, On the supremal p-negative type of a finite metric space, J. Math. Anal. Appl. 389 (2012), 98–107. 2
[34] I. J. Schoenberg, On certain metric spaces arising from euclidean spaces by a change of metric and their imbedding in
Hilbert space, Ann. of Math. 38 (1937), 787–793. 1
[35] J. Stern, Some applications of model theory in Banach space theory, Ann. Math. Logic 9 (1976), 49–122. 5
[36] A. Weston, On the generalized roundness of finite metric spaces, J. Math. Anal. Appl. 192 (1995), 323–334. 1
11
Department of Economics, New York University, New York, NY 10012, USA
E-mail address: [email protected]
Department of Mathematics, University of Chicago, Chicago, IL 60637, USA
E-mail address: [email protected]
Department of Mathematics and Statistics, Canisius College, Buffalo, NY 14208, USA
E-mail address: [email protected]
Department of Decision Sciences, University of South Africa, UNISA 0003, South Africa
E-mail address: [email protected]
12
|
1109.5736 | 1 | 1109 | 2011-09-26T23:02:27 | On The Uniqueness of The Strongly Irreducible Decompositions of Operators up to Similarity | [
"math.FA"
] | We give a generalization of the Jordan canonical form theorem for a class of bounded linear operators on complex separable Hilbert spaces in terms of direct integrals. Precisely, we study the uniqueness of strongly irreducible decompositions of the operators on the Hilbert spaces up to similarity. | math.FA | math |
J. OPERATOR THEORY
00:0(0000), 101 -- 110
© Copyright by THETA, 0000
ON THE UNIQUENESS OF THE STRONGLY IRREDUCIBLE
DECOMPOSITIONS OF OPERATORS UP TO SIMILARITY
RUI SHI
Communicated by Editor
ABSTRACT. We give a generalization of the Jordan canonical form theorem
for a class of bounded linear operators on complex separable Hilbert spaces in
terms of direct integrals. Precisely, we study the uniqueness of strongly irre-
ducible decompositions of the operators on the Hilbert spaces up to similarity.
KEYWORDS: Strongly irreducible operator, von Neumann algebra, K0 group, direct
integral.
MSC (2000): Primary 47A67; Secondary 47A15, 47C15.
1. INTRODUCTION
Throughout this article, all Hilbert spaces discussed are complex and sepa-
rable. Denote by L (H ) the set of bounded linear operators on a Hilbert space
H . An idempotent P is an operator in L (H ) satisfying P2 = P. A projection Q
is an idempotent such that kerQ = (ranQ)⊥ (See [5]). An operator A in L (H )
is said to be irreducible if its commutant {A}′ , {B ∈ L (H ) : AB = BA} con-
tains no projections other than 0 and the identity operator I on H , introduced by
P. Halmos in [11]. (The separability assumption is necessary because on a non-
separable Hilbert space every operator is reducible.) An operator A in L (H )
is said to be strongly irreducible if XAX−1 is irreducible for every invertible oper-
ator X in L (H ) [10]. This shows that the commutant of a strongly irreducible
operator contains no idempotents other than 0 and I. Strong irreducibility stays
invariant up to similar equivalence while irreducibility is only an invariant up to
unitary equivalence. An idempotent P in {A}′ is said to be minimal if every idem-
potent Q in {A}′ ∩ {P}′ satisfies QP = P or QP = 0. For a minimal idempotent
P in {A}′, it can be observed that the restriction AranP is strongly irreducible on
ranP. An operator A in L (H ) is said to have a finite strongly irreducible decompo-
i=1 in {A}′ such that
sition if there exist finitely many minimal idempotents {Pi}n
∑n
i=1 Pi = I and PiPj = PjPi = 0 for 1 ≤ i 6= j ≤ n. By the above observation, an
102
RUI SHI
operator A in L (H ) having a finite strongly irreducible decomposition can be
expressed as a direct sum of finitely many strongly irreducible operators.
On finite dimensional Hilbert spaces, every strongly irreducible operator is
similar to a Jordan block. In [12], D. A. Herrero and C. Jiang proved that for ev-
ery operator T in L (H ), there exists a sequence {Tn}∞
n=1 in L (H ) such that
limn→∞ kT − Tnk = 0, where every operator Tn is similar to a direct sum of
finitely many strongly irreducible operators. Y. Cao, J. Fang and C. Jiang [4] stud-
ied the uniqueness of finite strongly irreducible decompositions of operators in
L (H ) up to similar equivalence by the K0 groups of Banach algebras. For more
work around this subject, the reader is referred to [7, 8, 9, 13, 14, 15, 16, 17, 19].
Inspired by the ideas and results in [4], we study operators in L (H ) which may
have no finite strongly irreducible decompositions. In particular, there are many
operators in L (H ) whose commutants contain no minimal idempotents. To
represent these operators, direct sums of strongly irreducible operators need to
be generalized to direct integrals with some regular Borel measures. In [18], C.
Jiang and the author of the present paper proved that an operator A in L (H ) is
similar to a direct integral of strongly irreducible operators if and only if its com-
mutant {A}′ contains a bounded maximal abelian set of idempotents. A direct
integral of strongly irreducible operators means the integrand is strongly irre-
ducible almost everywhere on the domain of integration. For related concepts
and results about direct integrals and abelian von Neumann algebras, the reader
is referred to [3, 5, 6, 20, 21].
Following the notation of [18], we generalize a definition mentioned above.
An operator A in L (H ) is said to have a strongly irreducible decomposition if its
commutant {A}′ contains a bounded maximal abelian set of idempotents. Fur-
thermore, a strongly irreducible decomposition of the operator A is said to be
unique up to similarity if for bounded maximal abelian sets of idempotents P and
Q in {A}′, there is an invertible operator X in {A}′ such that XP X−1 = Q.
As a corollary of the main theorems, a normal operator in L (H ) has unique
strongly irreducible decomposition up to similarity if and only if the multiplicity
function mN for N is finite a. e. on σ(N) with respect to the scalar-valued spectral
measure µN . By this, the tensor product IH ⊗ N does not have unique strongly
irreducible decomposition up to similarity, if dimH = ∞.
To simplify the statements of the main theorems, we need to introduce
the upper triangular representation for operator-valued matrices. Assume A in
L (H ) is a direct integral of strongly irreducible operators in the form
A =
∞Mn=1
ZΛn
A(λ)dµ(λ),
(1)
with respect to a partitioned measure space {Λ, µ, {Λn}∞
n=1}, where µ is a regular
Borel measure on a compact set Λ and {Λn}∞
n=1 is a Borel partition of Λ, and
the equation µ(Λn) = 0 holds for all but finitely many n in N (0 /∈ N), and the
dimension of the fibre space H
λ ([1], §2) is n for almost every λ in Λn.
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
103
By ([2], Corollary 2), there is a unitary operator U such that
U AU∗ =
Z
∞Mn=1
Λn
Mφn Mφn
12
Mφn
13
0 Mφn Mφn
23
Mφn
0
...
...
0
0
0
...
0
· · · Mφn
1n
· · · Mφn
2n
· · · Mφn
3n
...
. . .
· · · Mφn
n×n
(λ)dµn(λ),
(2)
where µn = µΛn for 1 ≤ n < ∞ and φn, φn
cation operators. Denote by νn = µn ◦ φ−1
n
Mφn. Let the set {Γnm}m=∞
spect to the νn-measurable multiplicity function mφn
such that mφn
ij ∈ L∞(µn), and Mφn, Mφn
the scalar-valued spectral measure for
m=1 be the Borel partition of the spectrum σ(Mφn) with re-
for Mφn defined on σ(Mφn)
(λ) = m for almost every λ in Γnm. Write νnm for νnΓnm, 1 ≤ m ≤ ∞.
For a class of operators in L (H ) having unique strongly irreducible de-
compositions up to similarity, we give a necessary and sufficient condition by
K-theory for Banach algebras. Precisely, we prove the following theorems.
are multipli-
ij
THEOREM 1.1. Assume that an operator A in L (H ) is stated as in (1) and ex-
pressed as in (2) such that
(i) the νn-measurable multiplicity function mφn
spectrum σ(Mφn) for every n in N and
is simple and may take ∞ on the
(ii) every superdiagonal entry as in (2) is invertible for n in {n ∈ N : µ(Λn) > 0}.
Then the following statements are equivalent.
(a) The strongly irreducible decomposition of A is unique up to similarity.
(b) There exists a bounded N-valued simple function rA on σ(A) such that
V({A}′) ∼= { f (λ) ∈ N(rA (λ)) : f is Borel and bounded on σ(A)} and
K0({A}′) ∼= { f (λ) ∈ Z(rA
(λ)) : f is Borel and bounded on σ(A)}.
THEOREM 1.2. If an operator A in L (H ) is expressed as in (1) and (2) such
is simple and bounded on σ(Mφn) for
that the νn-measurable multiplicity function mφn
every n in N, then there exists a sequence of operators {Ak}∞
k=1 in L (H ) required as
in Theorem 1.1 and having unique strongly irreducible decompositions up to similarity
such that limk→∞ kAk − Ak = 0.
2. PROOFS
The following lemma describes an important property of the superdiagonal
entries in (2).
104
RUI SHI
LEMMA 2.1. An upper triangular matrix in Mn(C) of the form
α11
0
0
...
0
α12
α22
0
...
0
α13
α23
α33
...
0
· · ·
· · ·
· · ·
. . .
· · ·
α1n
α2n
α3n
...
αnn
is strongly irreducible if and only if the equation α11 = α22 = · · · = αnn and the inequal-
ity αi,i+1 6= 0 for 1 ≤ i ≤ n − 1 both hold.
Proof. If the matrix is strongly irreducible, then the equation α11 = · · · = αnn
holds. Write α for αii, 1 ≤ i ≤ n. Because every strongly irreducible matrix is
similar to a Jordan matrix, we know that there is an invertible matrix in Mn(C)
such that
α11
0
0
...
0
α12
α22
0
...
0
=
x11
x21
x31
...
xn1
α13
α23
α33
...
0
x12
x22
x32
...
xn2
· · ·
· · ·
· · ·
. . .
· · ·
x13
x23
x33
...
xn3
α1n
α2n
α3n
...
αnn
· · ·
· · ·
· · ·
. . .
· · ·
x1n
x2n
x3n
...
xnn
x11
x21
x31
...
xn1
x12
x22
x32
...
xn2
x13
x23
x33
...
xn3
· · ·
· · ·
· · ·
. . .
· · ·
x1n
x2n
x3n
...
xnn
α 1
0
0 α 1
0
...
0
· · ·
· · ·
0 α · · ·
...
. . .
· · ·
0
...
0
0
0
0
...
α
.
This equation yields that xij = 0 for i > j and xii = αi,i+1xi+1,i+1 for i = 1, 2, . . . , n −
1. Hence, we obtain xkk = ∏n−1
i=k αi,i+1 xnn. If αi,i+1 = 0 for some i in {1, 2, . . . , n − 1},
then the matrix (xij)1≤i,j≤n is not invertible. Therefore the inequality αi,i+1 6= 0
holds for 1 ≤ i ≤ n − 1.
On the other hand, if αi,i+1 6= 0 holds for i = 1, 2, . . . , n, then every matrix
in Mn(C) commuting with the matrix (αij)1≤i,j≤n can be expressed in the form
X =
x11
0
0
...
0
x12
x11
0
...
0
x13
x23
x11
...
0
· · ·
· · ·
· · ·
. . .
· · ·
x1n
x2n
x3n
...
x11
.
If X is an idempotent, then it must be I or 0. Thus the matrix (αij)1≤i,j≤n is strongly
irreducible.
Applying this lemma, we obtain the following corollary.
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
105
COROLLARY 2.2. In (2), the function φn
almost everywhere on Λn for i = 1, 2, . . . , n − 1.
i,i+1 in L∞(µn) satisfies φn
i,i+1(λ) 6= 0
In this corollary, the Multiplication operator Mφn
induced by the function
φn
i,i+1 is not invertible in general. But Mφn
can be approximated by a sequence
of invertible Multiplication operators in L (L2(µn)). Meanwhile, replacing the
superdiagonal entries with invertible ones enable us to simplify the problem.
That is why we add the hypothesis (ii) in Theorem 1.1. Precisely, we obtain the
following two lemmas.
i,i+1
i,i+1
LEMMA 2.3. If an operator An is a direct integral of strongly irreducible operators
stated as in (2) in the form
then for every positive integer k, there exists an operator Ank in the form
An =
Ank =
Mφn Mφn
12
Mφn
13
0 Mφn Mφn
23
Mφn
0
...
...
0
0
0
...
0
· · · Mφn
1n
· · · Mφn
2n
· · · Mφn
3n
...
. . .
· · · Mφn
Mφn Mφn
12,k
Mφn
13
Mφn Mφn
23,k
Mφn
...
0
0
...
0
· · · Mφn
1n
· · · Mφn
2n
· · · Mφn
3n
...
. . .
· · · Mφn
0
0
...
0
,
n×n
n×n
1
k
.
1
kn
;
with invertible Mφn
i,i+1,k
for 1 ≤ i ≤ n − 1 such that kAn − Ankk <
Proof. For λ in Λn, we construct φn
i,i+1,k in the form
φn
i,i+1,k(λ) =
Thus kMφn
i,i+1
− Mφn
i,i+1,k
k <
φn
i,i+1(λ),
φn
i,i+1(λ)
knφn
i,i+1(λ)
1
kn
,
,
if φn
i,i+1(λ) ≥
;
1
kn
i,i+1(λ) <
if 0 < φn
if φn
i,i+1(λ) = 0.
1
k(n − 1)
,
1 ≤ i ≤ n − 1. Therefore we obtain
kAn − Ankk ≤
n−1
∑
i=1
kMφn
i,i+1
− Mφn
i,i+1,k
k < (n − 1)
1
k(n − 1)
=
1
k
.
By the definition, the operator Mφn
i,i+1,k
is invertible for 1 ≤ i ≤ n − 1.
106
RUI SHI
LEMMA 2.4. If an operator An is a direct integral of strongly irreducible operators
stated as in (2) in the form
An =
Mφn Mφn
12
Mφn
13
0 Mφn Mφn
23
Mφn
0
...
...
0
0
0
...
0
· · · Mφn
1n
· · · Mφn
2n
· · · Mφn
3n
...
. . .
· · · Mφn
n×n
such that Mφn
invertible operator Xn in L ((L2(µn))(n)) such that Xn AnX−1
n
is invertible in L (L2(µn)) for i = 1, 2, . . . , n − 1, then there exists an
is in the form
i,i+1
Xn AnX−1
n =
Mφn
I
0
I
0 Mφn
0
...
0
0 Mφn
...
...
0
0
· · ·
0
· · ·
0
· · ·
0
...
. . .
· · · Mφn
.
(3)
Proof. We construct an invertible upper triangular operator-valued matrix
Xn in L ((L2(µn))(n)) as follows.
the equation
Choose an invertible operator M fnn in L (L2(µn)). Fix an operator M fii
by
Mφn
i,i+1
M fi+1,i+1
= M fii
for i = 1, . . . , n − 1. Let {M fii
Notice that every operator in the set {M fii
Choose an operator M f n
n−1,n
equation
}n
i=1 be the main diagonal (0-diagonal) entries of Xn.
}n
i=1 is invertible in L (L2(µn)).
in L (L2(µn)). Fix an operator M f n
by the
i,i+1
Mφn
i,i+1
M f n
i+1,i+2
+ Mφn
i,i+2
M f n
i+2,i+2
= M f n
i,i+1
for i = 1, . . . , n − 2. Let {M f n
i,i+1
Choose an operator M f n
n−l,n
}n−1
i=1 be the 1-diagonal entries of Xn.
in L (L2(µn)), where l is a positive integer such
that 1 ≤ l ≤ n − 1. Fix an operator M f n
by the equation
i,i+l
Mφn
i,i+1
M f n
i+1,i+l+1
+ Mφn
i,i+2
M f n
i+2,i+l+1
+ · · · + Mφn
i,i+l+1
M f n
i+l+1,i+l+1
= M f n
i,i+l
for i = 1, . . . , n − l − 1. Let {M f n
Choose an operator M f n
1n
}n−l
i=1 be the l-diagonal entries of Xn.
i,i+l
in L (L2(µn)) to be the n-diagonal entry of Xn.
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
107
Therefore we obtain an invertible operator-valued matrix Xn in the form
Xn =
12
M f11 M f n
M f n
13
0 M f22 M f n
0 M f33
0
...
...
...
0
0
0
23
· · · M f n
1n
· · · M f n
2n
· · · M f n
3n
...
. . .
· · · M fnn
such that the equation (3) holds.
By Lemma 2.4, we can reduce equation (2) to the form
A =
Z
∞Mn=1
Λn
Mφn
I
0
I
0 Mφn
0
...
0
0 Mφn
...
...
0
0
· · ·
0
· · ·
0
· · ·
0
...
. . .
· · · Mφn
n×n
(λ)dµn(λ),
(4)
in the sense of similar equivalence.
For a regular Borel measure ν on C with compact support K, define Nν on
L2(ν) by Nν f = z · f for each f in L2(ν).
LEMMA 2.5. Let an operator An be in the form
An =
N
(∞)
νn
0
0
...
0
N
I
(∞)
νn
0
...
0
0
I
(∞)
νn
N
...
0
0
0
0
...
· · ·
· · ·
· · ·
. . .
· · · N
(∞)
νn
n×n
,
where νn is a regular Borel measure and supported on some compact set Kn such that
0 < νn(Kn) < ∞. Then the strongly irreducible decomposition of An is not unique up
to similarity.
Proof. To prove this lemma, we need to construct two bounded maximal
abelian sets of idempotents in {An}′ which are not similar.
(∞)
νn
We can write N
in the form Nνn ⊗ Il2, where Il2 is the identity operator
on l2. Denote by P the set of all the spectral projections of Nνn . This set forms
a bounded maximal abelian set of idempotents in {Nνn }′. Let {ek}∞
k=1 be an or-
thonormal basis for l2. Denote by Ek the projection such that ranEk = {λek :
, {P ∈ L (l2) : P = P∗ = P2 ∈ {Ek : k ∈ N}′′}. De-
λ ∈ C}. Let Q1
note by χS the characteristic function for a Borel subset S in the interval [0, 1]
∈ L (L2[0, 1]) : S ⊂ [0, 1] is Borel.}. There is a unitary op-
and let
erator U : L2[0, 1] → l2 such that UPU∗ ∈ L (l2) for every P ∈ Q2. The sets
, U Q2U∗ and Q1 are two bounded maximal abelian sets of idempotents in
Q2
, {MχS
Q2
108
RUI SHI
L (l2) but they are not unitarily equivalent. By the fact that W∗(P) ⊗ W∗(Q1)
and W∗(P) ⊗ W∗(Q2) are both maximal abelian von Neumann algebras, we ob-
tain that
F1
, {P ∈ W∗(P) ⊗ W∗(Q1) : P = P∗ = P2}
and
F2
, {P ∈ W∗(P) ⊗ W∗(Q2) : P = P∗ = P2}
are both maximal abelian sets of idempotents in {Nνn ⊗ Il2}′ = L∞(νn) ⊗ L (l2).
is a bounded maximal abelian set of idempo-
We need to prove that F (n)
i
tents in {An}′ for i = 1, 2.
An operator X in {An}′ can be expressed in the form
X =
X11 X12 X13
X21 X22 X23
X31 X32 X33
...
...
Xn1 Xn2 Xn3
...
· · · X1n
· · · X2n
· · · X3n
...
. . .
· · · Xnn
n×n
.
(5)
We prove that Xij is in {Nνn ⊗ Il2}′. Note that P (∞) is the set of all the spectral
projections of Nνn ⊗ Il2. Fix an projection P in (P (∞))(n). The operator A can be
expressed in the form
An = An1 ⊕ An2,
where
Ani =
N
(∞)
νni
0
0
...
0
N
I
(∞)
νni
0
...
0
0
I
(∞)
νni
N
...
0
0
0
0
...
· · ·
· · ·
· · ·
. . .
· · · N
(∞)
νni
, i = 1, 2.
The measures νn1 and νn2 are mutually singular and their supports depend on
the characteristic functions corresponding to P and I − P. Hence X can also be
expressed in the form
X = (cid:18)Y11 Y12
Y21 Y22(cid:19) ranP
ran(I − P)
.
The equations An1Y12 = Y12 An2 and An2Y21 = Y21 An1 yield that Y12 = Y21 = 0.
Therefore P reduces X and Xijs are in {Nνn ⊗ Il2}′. A computation shows that the
equation Xij = 0 holds for i > j and Xii = X11 for i = 2, . . . , n in (5). Furthermore,
if X as in (5) is an idempotent, then so is every main diagonal entry Xii of X.
We assume that X is an idempotent in {An}′ and commutes with F (n)
.
Hence Xii commutes with F1. The fact that F1 is a maximal abelian set of
idempotents implies that Xii belongs to F1. Thus Xii commutes with Xij. For
the 1-diagonal entries, the equation 2XiiXi,i+1 − Xi,i+1 = 0 yields Xi,i+1 = 0,
1
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
109
for i = 1, . . . , n − 1. By this way, the k-diagonal entries of X are all zero, for
k = 2, . . . , n. Therefore X is in F (n)
are bounded maximal
abelian sets of idempotents in {An}′.
. Both F (n)
and F (n)
1
1
2
We prove that F (n)
and F (n)
2
{An}′ can be written in the form
1
are not similar in {An}′. Every operator X in
X = Zσ(Nνn )
X(λ)dνn(λ).
Suppose that there is an invertible operator X in {An}′ such that
XF (n)
2 X−1 = F (n)
1
.
(n)
For each P in F
2
λ in σ(Nνn). But there exists an projection Q in F (n)
for almost every λ in σ(Nνn). This is a contradiction. Therefore F (n)
are not similar in {An}′.
, the projection P(λ) is either of rank ∞ or 0, for almost every
such that Q(λ) is of rank n,
and F (n)
1
1
2
By ([22], Theorem 3.3), we have the following corollary.
COROLLARY 2.6. Let an operator An be in the form
An =
N
(m)
νn
0
0
...
0
N
I
(m)
νn
0
...
0
0
I
(m)
νn
N
...
0
0
0
0
...
· · ·
· · ·
· · ·
. . .
· · · N
(m)
νn
n×n
,
where m is a positive integer and νn is a regular Borel measure supported on some compact
set Kn such that 0 < νn(Kn) < ∞. Then the strongly irreducible decomposition of An is
unique up to similarity.
For a regular Borel measure ν with compact support, Denote by Jm(ν) an
operator in the form
Jm(ν) =
Nν
I
0 Nν
0
...
0
0
I
0 Nν
...
...
0
0
· · ·
0
· · ·
0
· · ·
0
...
. . .
· · · Nν
m×m
L2(ν)
L2(ν)
L2(ν)
...
L2(ν)
.
(6)
Mφ1
Mφ2
0 Mφ1
...
...
0
0
· · · Mφm
· · · Mφm−1
...
. . .
· · · Mφ1
,
Mφ11
0
...
0
Mφ12
Mφ22
...
0
· · · Mφ1m
· · · Mφ2,m−1
. . .
· · · Mφmm
...
.
110
RUI SHI
LEMMA 2.7. Every operator in {Jm(ν)}′ is in the form
where φi is in L∞(ν) for 1 ≤ i ≤ m.
erator in {Jm(ν)}′ is in the form
Proof. By a similar computation as in Lemma 2.5, we obtain that every op-
By the equation
Nν Mφi,j+1
+ Mφi+1,j+1
= Mφi,j
+ Mφi,j+1
Nν,
the k-diagonal entries are as required for 1 ≤ k ≤ n.
LEMMA 2.8. Let m1 and m2 be two positive integers such that m1 > m2. Then
the following equations hold:
(i) {B ∈ L ((L2(ν))(m2), (L2(ν))(m1)) : Jm1
= {(CT, 0)T : C ∈ {Jm2
(ν)}′}.
(ii) {B ∈ L ((L2(ν))(m1), (L2(ν))(m2)) : Jm2
= {(0, C) : C ∈ {Jm2
(ν)}′}.
(ν)B = BJm2
(ν)}
(ν)B = BJm1
(ν)}
Proof. We only need to prove the first equation. The second equation can be
obtained by the same method. Let B = (CT, DT)T such that C and D are in the
form
C =
B11
B12
B21
B22
...
...
Bm2 1 Bm2 2
B1m2
B2m2
...
· · ·
· · ·
. . .
· · · Bm2 m2
, D =
Bm2 +1,1
...
Bm1 1
· · · Bm2 +1,m2
...
· · ·
Bm1 m2
.
By a similar computation as in Lemma 2.5, we can obtain that P(m1
) B = BP(m2 )
for every spectral projection P of Nν. Thus, every Bij belongs to {Nν}′, for 1 ≤ i ≤
m1 and 1 ≤ j ≤ m2. For i = 1, . . . , m1 − 1, the equation NνBi1 + Bi+1,1 = Bi1Nν
yields Bi+1,1 = 0. For i = 2, . . . , m1 − 1, the equation NνBi2 + Bi+1,2 = Bi2Nν
yields Bi+1,2 = 0. By this way, we can obtain Bij = 0 for i > j. Hence D = 0 and
a further computation shows that C ∈ {Jm2
(ν)}′.
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
111
LEMMA 2.9. Let m1, m2, r1, r2 be positive integers. If an idempotent P in Mn(C)
is in the form
P =
Im1
0
0
0
0
0r1
0
0
R11 R12
R21 R22
Im2
0
0r2
0
,
where Im is the identity operator in Mm(C), then there exists an invertible operator X in
the form
X =
Im1
0
0
0
0
Ir1
0
0
0
−R21
Im2
0
such that
and X−1 =
Im1
0
0
0
0
Ir1
0
0
0 −R12
R21
Im2
0
0
0
Ir2
R12
0
0
Ir2
Im1
0
0
0
XPX−1 =
0
0r1
0
0
0
0
Im2
0
0
0
0
0r2
.
Note that the equation P2 = P implies that R11 = R22 = 0 and the construc-
tion of X depends on P. In the following example, we construct an operator A
and prove the strongly irreducible decomposition of A is unique up to similarity.
EXAMPLE 2.10. Let A = J
(2)
1 (ν). We prove that for every
two bounded maximal abelian sets of idempotents P and Q in {A}′, there is an
invertible operator X in {A}′ such that the equation P = XQX−1 holds and
(3)
2 (ν) ⊕ J
(2)
3 (ν) ⊕ J
V({A}′) = { f is bounded Borel : σ(Nν) → N ⊕ N ⊕ N},
K0({A}′) = { f is bounded Borel : σ(Nν) → Z ⊕ Z ⊕ Z}.
Denote by Pm1
the set of all the idempotents in {Jm1
(ν)}′. Note that Pm1
the set of all
. Denote by Em2
the set of all the projections in
(m1 )
equals the set of all the spectral projections of N
ν
(C) and by Fm1 ,m2
the diagonal projections in Mm2
{Pm1
}′′.
⊗ Em2
Let P = F3,2 ⊕ F2,3 ⊕ F1,2. We can verify that P is a bounded maximal
abelian set of idempotents in {A}′. Then we only need to prove that for every
bounded maximal abelian set of idempotents Q in {A}′, there is an invertible
operator X in {A}′ such that P = XQX−1.
We reduce the rest into two claims:
(i) For every idempotent P in {A}′, there is an invertible operator X in {A}′
such that XPX−1 belongs to P.
112
RUI SHI
(ii) There are seven idempotents {Qk}7
k=1 in Q such that for almost every
k=1 and Q(λ) generate the same bounded maximal
λ in σ(Nν), {Qk(λ)}7
abelian set of idempotents.
Every operator B in {A}′ can be expressed in the form
B =
B11 B12
B21 B22
...
...
B71 B72
· · · B17
· · · B27
...
. . .
· · · B77
,
where
b11
1
0
0
B11 =
B31 = (cid:18)0
B61 = (cid:0)0 0
0
b11
2
b11
1
0
b31
1
0
b31
b13
1
0
0
b11
3
b11
2
b11
1
, B13 =
B33 =(cid:18)b33
1 (cid:19) ,
B63 =(cid:0)0
1 (cid:1) ,
b31
2
b31
1
0
b16
1
0
0
b13
2
b13
1
0
, B16 =
,
0 (cid:19) ,
B36 = (cid:18)b36
1 (cid:19) ,
B66 = (cid:0)b66
1 (cid:1) ,
1 (cid:1) ,
b33
2
b33
1
b33
other Bijs are expressed as follows:
• For 1 ≤ i, j ≤ 2, Bijs are of the same form;
• For 3 ≤ i, j ≤ 5, Bijs are of the same form;
• For 6 ≤ i, j ≤ 7, Bijs are of the same form;
• For 1 ≤ i, j ≤ 2 and 3 ≤ j ≤ 5, Bijs are of the same form;
• For 1 ≤ i, j ≤ 2 and 6 ≤ j ≤ 7, Bijs are of the same form;
• For 3 ≤ i, j ≤ 5 and 1 ≤ j ≤ 2, Bijs are of the same form;
• For 3 ≤ i, j ≤ 5 and 6 ≤ j ≤ 7, Bijs are of the same form;
• For 6 ≤ i, j ≤ 7 and 1 ≤ j ≤ 2, Bijs are of the same form;
• For 6 ≤ i, j ≤ 7 and 3 ≤ j ≤ 5, Bijs are of the same form,
where bij
k s belong to {Nν}′ for 1 ≤ i, j ≤ 7 and 1 ≤ k ≤ 3.
For B expressed in the above form, there is a unitary operator U1 such that
U1BU∗
1 =
B11 B12 B13
B22 B23
0
B33
0
0
,
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
113
where Bijs are in the form
B11 =
b11
1
b21
1
· · ·
0
0
b12
1
b22
1
· · ·
0
0
0
· · ·
0
· · ·
0
0
0
0
...
...
...
...
...
...
...
b14
1
b24
1
· · ·
b34
1
b44
1
b54
1
· · ·
0
0
b15
1
b25
1
· · ·
b35
1
b45
1
b55
1
· · ·
0
0
...
...
...
...
...
...
...
b16
1
b26
1
· · ·
b36
1
b46
1
b56
1
· · ·
b66
1
b76
1
7×7
b17
1
b27
1
· · ·
b37
1
b47
1
b57
1
· · ·
b67
1
b77
1
,
(7)
b15
2
b25
2
· · ·
b35
2
b45
2
b55
2
· · ·
0
0
7×5
b35
1
b45
1
b55
1
b15
1
b25
1
· · ·
, B13 =
b11
3
b21
3
· · ·
0
0
0
· · ·
0
0
b12
3
b22
3
· · ·
0
0
0
· · ·
0
0
7×2
,
, B23 =
b11
2
b21
2
· · ·
0
0
0
b12
2
b22
2
· · ·
0
0
0
,
B33 = (cid:18)b11
1
b21
1
1 (cid:19) .
b12
1
b22
b13
1
b23
1
· · ·
b33
1
b43
1
b53
1
· · ·
0
0
b14
2
b24
2
· · ·
b34
2
b44
2
b54
2
· · ·
0
0
B12 =
b11
2
b21
2
· · ·
0
0
b12
2
b22
2
· · ·
0
0
0
· · ·
0
· · ·
0
0
...
...
...
...
...
...
...
b13
2
b23
2
· · ·
b33
2
b43
2
b53
2
· · ·
0
0
0
0
B22 =
b11
1
b21
1
· · ·
b12
1
b22
1
· · ·
0
0
0
0
0
0
...
...
...
...
...
b13
1
b23
1
· · ·
b33
1
b43
1
b53
1
b14
1
b24
1
· · ·
b34
1
b44
1
b54
1
114
RUI SHI
If P is an idempotent in {U1 AU∗
1 }′, then by the proof of ([22], Lemma 3.4),
we can construct an invertible operator X in {U1 AU∗
1 }′ of the form
where the main diagonal blocks as in (7) are diagonal projections. There is also a
unitary operator U2 in {U1 AU∗
1 }′ of the form
such that
such that the equation
X2
X1
X =
XPX−1 =
X3
⊕(cid:18)X1
P11 P12 P13
P22 P23
0
P33
0
0
X2(cid:19) ⊕ X1,
,
U2 =
U2
1
U2
2
U2
3
⊕(cid:18)U2
1
2(cid:19) ⊕ U2
1,
U2
=
U2
1
U2
2
U2
1
P11
U2
3
U2
2
∗
U2
3
(λ)
Is1
0
· · ·
0
0
· · ·
0
0
0
0t1
· · ·
0
0
· · ·
0
0
...
...
...
...
...
...
∗
∗
· · ·
Is2
0
· · ·
0
0
∗
∗
· · ·
0
0t2
· · ·
0
0
...
...
...
...
...
...
∗
∗
∗
· · ·
∗
∗
· · ·
Is3
0
∗
· · ·
∗
∗
· · ·
0
0t3
bP = U2XPX−1U∗
2 =
bP11
0
0
bP12
bP22
0
.
bP13
bP23
bP33
holds for almost every λ in σ(Nν), where si and ti are non-negative integers.
Write U2XPX−1U∗
2 in the form
By Lemma 2.9, we can construct an invertible operator Y1 in {U1 AU∗
1 }′ such that
thermore, we can construct an invertible operator Y2 in {U1 AU∗
bP11, bP22, and bP33 become diagonal projections after similar transformation. Fur-
diagonal blocks of Y1bPY−1
1 }′ such that the 1-
1 vanish after similar transformation. And then we can
1 }′ such that the 2-diagonal blocks
construct an invertible operator Y3 in {U1 AU∗
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
115
2 vanish after similar transformation. Thus we finish the proof of
1 Y−1
of Y2Y1bPY−1
claim (i).
To prove claim (ii), we need to define a ν-measurable function rQ with re-
spect to an idempotent Q in {A}′. Without loss of generality, we assume that
Q ∼ (P31 ⊕ P32) ⊕ (P21 ⊕ P22 ⊕ P23) ⊕ (P11 ⊕ P12) ∈ F3,2 ⊕ F2,3 ⊕ F1,2. Define
rQ (λ) , 1
3
+
+ Tr(P11(λ) + P12(λ)), λ ∈ σ(Nν),
Tr(P31(λ) + P32(λ))
1
2
Tr(P21(λ) + P22(λ) + P22(λ))
where Tr stands for the standard trace of a square matrix. Note that rQ stays
invariant up to similarity. By the proof of ([22], Lemma 3.5), we can obtain that
there are seven idempotents Qi in Q such that
• the equation rQi (λ) = 1 holds a. e. on σ(Nν) for i = 1, . . . , 7 and
• the equation QiQj = 0 holds for i 6= j.
The idempotent Qi may be in the form
Qi(λ) ∼
I3 ⊕ 0, λ ∈ Λ3;
I2 ⊕ 0, λ ∈ Λ2;
I1 ⊕ 0, λ ∈ Λ1,
where {Λi}3
(7)
projections of N
ν
belongs to a Pm for m = 1, 2, 3. We finish the proof of claim (ii).
i=1 is a Borel partition of σ(Nν). We can choose finitely many spectral
to cut Qis and to piece together new Qis such that every Qi
By the proof of ([22], Lemma 3.6) and the idempotents {Qi}7
i=1 constructed
above, we can obtain an invertible operator X in {A}′ such that XQX−1 = P.
Therefore the strongly irreducible decomposition of A is unique up to similarity.
Assume that Q1, Q2 and Q3 are in P1, P2 and P3 respectively. Then there
is a group isomorphism α such that α([Q1]) = (1, 0, 0), α([Q2]) = (0, 1, 0), and
α([Q3]) = (0, 0, 1), where [Qi] stands for the similar equivalence class of Qi in
n=1 Mn({A}′)/ ∼. Thus we obtain α([I]) = (2, 3, 2), where I is
the identity operator in {A}′. Furthermore, a routine computation yields that
V({A}′) and K0({A}′) are of the forms at the beginning of this example.
V({A}′) = S∞
Proof of Theorem 1.1. By the calculating in the above example, we can prove
(ν) is unique up to simi-
that the strongly irreducible decomposition ofLk
larity, where mi, ni and k are all positive integers.
There is a unitary operator V such that VAV∗ can be expressed in the form
as described at the beginning of Example 2.10. Then we can apply the above
lemmas to perform calculation as we need. Note that the equation
(ni)
mi
i=1 J
{(
k1Mi=1
J
(ni)
mi
(ν1)) ⊕ (
k2Mj=1
(nj)
mj
J
(ν2))}′ = {
k1Mi=1
J
(ni)
mi
(ν1)}′ ⊕ {
k2Mj=1
(nj)
mj
J
(ν2)}′
116
RUI SHI
holds for mutually singular Borel measures ν1 and ν2.
By Lemma 2.5, if the strongly irreducible decomposition of A is unique up
to similarity, then every multiplicity function mφn
is bounded. Then we can obtain
that V({A}′) and K0({A}′) are as described in the theorem. On the other hand,
if the strongly irreducible decomposition of A is not unique up to similarity, then
there is a number m in {n ∈ N : µ(Λn) > 0} such that the multiplicity function
takes ∞ in its codomain on a Borel subset Γm1 of measure nonzero in its
mφm
domain. Therefore in K0({A}′), every Borel function f vanishes on Γm1. This is a
contradiction.
The proof of Theorem 1.2 is an application of Lemma 2.3.
Acknowledgements. The author is grateful to Professor Chunlan Jiang and Professor
Guihua Gong for their advice and comments on writing this paper. Also the author was
supported in part by NSFC Grant (No. 10731020) and NSFC Grant (No.10901046).
REFERENCES
[1] M. B. ABRAHAMSE, Multiplication operators. Hilbert space operators (Proc. Conf.,
Calif. State Univ., Long Beach, Calif., 1977), 17 -- 36, Lecture Notes in Math., 693,
Springer, Berlin, 1978. MR0526530 (80b:47042)
[2] EDWARD A. AZOFF, Borel measurability in linear algebra. Proc. Amer. Math. Soc. 42
(1974), 346 -- 350. MR0327799 (48 #6141)
[3] E. AZOFF, C. FONG, F. GILFEATHER, A reduction theory for non-self-adjoint operator
algebras. Trans. Amer. Math. Soc. 224 (1976), 351 -- 366. MR0448109 (56 #6419)
[4] Y. CAO, J. FANG, C. JIANG, K-groups of Banach algebras and strongly irreducible
decompositions of operators. J. Operator Theory 48 (2) (2002), 235 -- 253. MR1938796
(2004f:47003)
[5] JOHN B. CONWAY, A course in functional analysis. Second edition. Graduate Texts in
Mathematics, 96. Springer-Verlag, New York, 1990. xvi+399 pp. ISBN: 0-387-97245-5.
MR1070713 (91e:46001)
[6] KENNETH R. DAVIDSON, C∗-algebras by example. Fields Institute Monographs, 6.
American Mathematical Society, Providence, RI, 1996. xiv+309 pp. ISBN: 0-8218-0599-
1. MR1402012 (97i:46095)
[7] M. ENOMOTO, Y. WATATANI, Relative position of four subspaces in a Hilbert space.
Adv. Math. 201 (2) (2006), 263 -- 317. MR2211531 (2009d:46044)
[8] M. ENOMOTO, Y. WATATANI, Exotic indecomposable systems of four subspaces in a
Hilbert space. Integral Equations and Operator Theory 59 (2) (2007), 149 -- 164. MR2345993
(2009g:47194)
[9] M. ENOMOTO, Y. WATATANI, Indecomposable representations of quivers on infinite-
dimensional Hilbert spaces. J. Funct. Anal. 256 (4) (2009), 959 -- 991. MR2488332
(2010c:47238)
ON THE STRONGLY IRREDUCIBLE DECOMPOSITIONS OF OPERATORS
117
[10] F. GILFEATHER, Strong reducibility of operators. Indiana Univ. Math. J. 22 (4) (1972),
393 -- 397. MR0303322 (46 #2460)
[11] P. HALMOS, Irreducible operators. Michigan Math. J. 15 (1968), 215 -- 223. MR0231233
(37 #6788)
[12] DOMINGO A. HERRERO, CHUNLAN JIANG, Limits of strongly irreducible operators,
and the Riesz decomposition theorem. Michigan Math. J. 37 (1990), no. 2, 283´lC291.
MR1058401 (91k:47035)
[13] C. JIANG, Similarity classification of Cowen-Douglas operators. Canad. J. Math. 56 (4)
(2004), 742 -- 775. MR2074045 (2006d:47037)
[14] C. JIANG, Z. WANG, The spectral picture and the closure of the similarity orbit of
strongly irreducible operators. Integral Equations Operator Theory 24 (1) (1996), 81 -- 105.
MR1366542 (97h:47011)
[15] C. JIANG, X. GUO, K. JI, K-group and similarity classification of operators. J. Funct.
Anal. 225 (1) (2005), 167 -- 192. MR2149922 (2006c:47023)
[16] C. JIANG, Z. WANG, Structure of Hilbert space operators. World Scientific Publish-
ing Co. Pte. Ltd., Hackensack, NJ, 2006. x+248 pp. ISBN: 981-256-616-3 MR2221863
(2008j:47001)
[17] C. JIANG, K. JI, Similarity classification of holomorphic curves. Adv. Math. 215 (2)
(2007), 446 -- 468. MR2355596 (2008g:46081)
[18] C. JIANG, R. SHI, Direct Integrals of Strongly Irreducible Operators. J. Ramanujan
Math. Soc. 26 (2) (2011), 165 -- 180.
[19] Z. JIANG, S. SUN, On completely irreducible operators. Front. Math. China 1 (4) (2006),
569 -- 581. MR2257195 (2007g:47003)
[20] H. RADJAVI, P. ROSENTHAL, Invariant subspaces. Ergebnisse der Mathematik und
ihrer Grenzgebiete, Band 77. Springer-Verlag, New York-Heidelberg, 1973. xi+219 pp.
MR0367682 (51 #3924)
[21] J. T. SCHWARTZ, W∗-algebras. Gordon and Breach, New York, 1967. vi+256 pp.
MR0232221 (38 #547)
[22] R. SHI, On a generalization of the Jordan canonical form theorem on separable Hilbert
spaces. arXiv:1109.4224v1 [math.FA].
RUI SHI, SCHOOL OF MATHEMATICAL SCIENCES, DALIAN UNIVERSITY OF TECH-
NOLOGY, DALIAN, 116024, CHINA
E-mail address: [email protected]
Received Month dd, yyyy; revised Month dd, yyyy.
|
1606.03122 | 2 | 1606 | 2016-09-20T23:17:47 | Asymptotically Hilbertian Modular Banach Spaces: Examples of Uncountable Categoricity | [
"math.FA",
"math.LO"
] | We give a criterion ensuring that the elementary class of a modular Banach space E (that is, the class of Banach spaces, some ultrapower of which is linearly isometric to an ultrapower of E) consists of all direct sums E\oplus_m H, where H is an arbitrary Hilbert space and \oplus_m denotes the modular direct sum. Also, we give several families of examples in the class of Nakano direct sums of finite dimensional normed spaces that satisfy this criterion. This yields many new examples of uncountably categorical Banach spaces, in the model theory of Banach space structures. | math.FA | math |
ASYMPTOTICALLY HILBERTIAN
MODULAR BANACH SPACES:
EXAMPLES OF UNCOUNTABLE CATEGORICITY
C. WARD HENSON* AND YVES RAYNAUD
Abstract. We give a criterion ensuring that the elementary class of a modular Banach
space E (that is, the class of Banach spaces, some ultrapower of which is linearly isometric
to an ultrapower of E) consists of all direct sums E ⊕m H, where H is an arbitrary Hilbert
space and ⊕m denotes the modular direct sum. Also, we give several families of examples
in the class of Nakano direct sums of finite dimensional normed spaces that satisfy this
criterion. This yields many new examples of uncountably categorical Banach spaces, in
the model theory of Banach space structures.
1. Introduction
The aim of this paper is to give some new examples of uncountably categorical Banach
space structures. The motivation is model-theoretic, but here we formulate our objectives
and methods of proof in the framework of ordinary Banach space theory, using the well
known ultrapower construction [4].
To begin, we give some terminology and explain briefly the model-theoretic background.
Two Banach spaces X, Y are elementarily equivalent if for some ultrafilters U and V, their
respective ultrapowers XU and YV are linearly isometric. This is an equivalence relation,
although its transitivity is not evident.
In fact this relation is identical to that of ap-
proximate elementary equivalence in the first author's logic for normed space structures
[5, Discussion p. 5 and Theorem 10.7] and also to that of elementary equivalence in con-
tinuous logic applied to the unit balls [2, Definition 4.3, Corollary 5.6 and Theorem 5.7].
These logical counterparts are defined to mean that the two Banach spaces satisfy the
same sentences having the appropriate syntactic form, and this makes it clear that the el-
ementary equivalence relation defined using ultrapowers is transitive. The class of Banach
spaces that are elementarily equivalent to a given Banach space is called the elementary
class of this Banach space; it consists of all ultraroots of ultrapowers of the given space.
(An ultraroot of a Banach space X is a Banach space Y , an ultrapower of which is linearly
isometric to X).
It follows from basic results of model theory, the Compactness Theorem as well as
the Lowenheim-Skolem Theorems (in either of the two logics mentioned above), that the
elementary class of any infinite dimensional Banach space contains spaces of all infinite
density characters.
If κ is an infinite cardinal number, an infinite dimensional Banach
space X is said to be κ-categorical if its elementary class contains exactly one member
2010 Mathematics Subject Classification. Primary: 46B04, 46B45; Secondary: 03C20, 03C35.
Key words and phrases. Banach spaces, ultraproducts, elementary equivalence, categoricity, isometric
embeddings, modular direct sums.
(*) Research of this author was supported by grant #202251 from the Simons Foundation.
Part of the work was carried out during the Universality and Homogeneity program at the Hausdorff
Institute of the University of Bonn (Fall 2013).
1
2
C. W. HENSON AND Y. RAYNAUD
of density character κ. It is called uncountably categorical if it is κ-categorical for some
uncountable cardinal κ; in that case it is κ-categorical for all uncountable cardinals κ, as
proved independently in [1] and [12].
Note that uncountable categoricity is a property of the elementary class of X rather than
of X by itself, so we should speak of an uncountably categorical class of Banach spaces. So
we say that a class C of Banach spaces is uncountably categorical if and only if C is closed
under ultrapowers and under ultraroots (hence it is closed under elementary equivalence),
and it contains only one member Xκ of density character κ for some uncountable cardinal
κ;
in that case all infinite dimensional members of C will necessarily be elementarily
equivalent to Xκ.
Indeed, up to linear isometry there
A trivial example is the class of Hilbert spaces.
is only one infinite dimensional Hilbert space of any given density character; moreover,
ultrapowers of any Hilbert space are Hilbert spaces, and so are closed subspaces (hence
also ultraroots).
Although the structure of uncountably categorical Banach space structures has begun
to be investigated [13], very few examples are known. In addition to the class of Hilbert
spaces, some less trivial examples consist of certain finite dimensional perturbations of
Hilbert spaces: namely given a finite dimensional normed space E, the class CE of all
2-direct sums E ⊕2 H, where H is any infinite dimensional Hilbert space, is κ-categorical
for all infinite cardinal κ. This fact was known to the authors for a few years; its proof is
included here (in Section 6) for the sake of completeness and because it is closely tied to
the other results that are expounded here. The main purpose of this article is, however, to
give some less trivial examples, which include many spaces that are not linear-topologically
isomorphic to a Hilbert space; one of them is even not linear-topologically embeddable
in any Banach lattice with nontrivial concavity. So from the point of view of Banach
space geometry, they are not close to the class of Hilbert spaces. On the other hand, in a
certain asymptotic sense, which will be explained in Section 3, they have a very hilbertian
character. (In a sense that is not yet well understood, the main result of [13] says that an
uncountably categorical Banach space must be very closely related to the class of Hilbert
spaces.)
Some other model-theoretic properties of our examples will be discussed in a future
paper. Here we concentrate on the methods needed to prove their categoricity, using
purely Banach space theoretic arguments which have their own interest, in particular in
modular sequence space theory.
This paper is organized as follows:
in Section 2 we describe a general pattern of the
examples that we present, in the framework of modular Banach spaces. We give a crite-
rion guaranteeing that the elementary class of a (separable, infinite dimensional) modular
space E, considered as a Banach space with no extra structure, consists exactly of all
the modular direct sums E ⊕m H, where H is a Hilbert space of arbitrary hilbertian di-
mension; this property clearly implies uncountable categoricity. All of our examples fit
into this framework. (The concepts of modular Banach space and modular direct sum are
explained in Section 2. We note that every Banach space can be considered as a modular
space by taking its modular function to be the square of the norm.) We then make a
brief presentation of the concrete examples whose properties are demonstrated later in the
paper. All these examples are given as Nakano direct sums of a sequence of finite dimen-
sional spaces. The Nakano (sequence) spaces appearing there are associated to a sequence
of exponents converging to 2. In Section 3, ultrapowers of these Nakano direct sums are
described; it is proved that when the geometry of the sequence of finite dimensional spaces
converges to the geometry of Hilbert space in a very weak sense, then the Nakano direct
sum is asymptotically hilbertian.
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
3
The main content of this paper is in Sections 4 and 5. There are two main families of
examples discussed there: in the first one, which is treated in Section 4, all exponents of the
underlying Nakano sequence space are distinct from 2; in the second one, treated in Section
5, all the exponents are equal to 2. In both cases, the sequence of finite dimensional spaces
used in the direct sum must converge in a suitable local sense to Hilbert space. There is
more flexibility in the case of exponents different from 2: here the factor spaces may be
even one-dimensional; in particular, the Nakano space itself is an example. This space is
not linear-topologically equivalent to Hilbert space if the convergence of the exponents to
2 is slow enough.
We finish with section 6, which treats the finite dimensional perturbations of the class
of Hilbert spaces already mentioned above, as an addendum to section 5.
Throughout this paper, the Banach spaces considered may be either real or complex,
and our proofs make no distinction between real and complex scalars. We denote the scalar
field by K. For general notions about ultrapowers the reader is referred to [4, Section I].
(Our notation differs slightly from that of this author; e.g., we use EU , [xi]U , in place of
(E)U , (xi)U , etc).
2. A general pattern
The Banach space examples E presented in this paper have the common feature that
every ultrapower of E is linearly isometric to a direct sum E ⊕ H, where H is a Hilbert
space.
(It can be shown that this condition implies that E is reflexive; we omit the
argument since we have easier direct proofs of reflexivity for the examples presented here.)
Indeed, if E is one of our examples, we show that for any ultrafilter U we have a direct
sum decomposition
EU = D(E) ⊕ H
(1)
where D : E → EU is the canonical embedding of E into its ultrapower EU (namely,
D assigns to each x ∈ E the element represented in EU := EI /U by the constant family
(x)i∈I ) and where H is a closed linear subspace of EU that is linearly isometric to a Hilbert
space.
xi
Then D(E) is complemented in EU by the weak limit projection P : [xi]U 7→ w- lim
i,U
(which exists since E is reflexive). If ker P is linearly isometric to a Hilbert space (i.e.,
equation (1) holds with H = ker P ) we say that E is asymptotically hilbertian (in the iso-
metric sense). Indeed this is an isometric version of the notion of asymptotically hilbertian
space considered, e.g., in [11, pp. 220-221], which corresponds to the case where ker P is
only linear-topologically isomorphic to a Hilbert space. The examples that we present here
are indeed asymptotically hilbertian (in the isometric sense), but the reader will observe
that this property is not formally required by the main theoretical tool of the present
section (Theorem 2.1).
To go further toward a similar description of all members of the elementary class of E,
we need to make hypotheses on the nature of the direct sum. For example, if the direct
sum in equation (1) is always a 2-sum, that is, if
kx + hk2 = kxk2
E + khk2
holds for all x ∈ E, h ∈ H, then for any Hilbert space K:
H
(E ⊕2 K)U = EU ⊕2 KU = (E ⊕2 H) ⊕2 KU = E ⊕2 (H ⊕2 KU ) = E ⊕2 K′
where K′ = H ⊕2 KU is still a Hilbert space. It follows that the class C of direct sums
E ⊕2 K, where K is a Hilbert space, is closed under ultrapowers. Moreover, if E is
separable, then for every uncountable cardinal κ, the only member of C with density
4
C. W. HENSON AND Y. RAYNAUD
character κ is E⊕2 Hκ, where Hκ is the Hilbert space of hilbertian dimension κ. Therefore,
if C turns out to be closed under ultraroots, it is necessarily uncountably categorical.
Furthermore, there are other simple kinds of direct sums, apart from the 2-sums, for
which the preceding reasoning is valid, namely the modular direct sums. A convex modular
on a linear space X is a convex function Θ : E → R+ that satisfies the following conditions:
Θ(0) = 0, Θ is symmetric (Θ(λx) = Θ(x) for any scalar λ with λ = 1), and Θ is faithful
(Θ(x) = 0 =⇒ x = 0). An associated norm on X is then defined by the Luxemburg
formula
(2)
kxk = inf{λ > 0 : Θ(x/λ) ≤ 1}.
Since we require that the convex modular has finite values, the Luxemburg norm is im-
plicitly defined by the equation
(3)
Θ(cid:18) x
kxk(cid:19) = 1.
The modular direct sum X1⊕m X2 of two modular spaces (X1, Θ1), (X2, Θ2) is their linear
topological direct sum equipped with the modular
Θ(x1 + x2) = Θ1(x1) + Θ2(x2) x1 ∈ X1, x2 ∈ X2 .
Xi
Note that if norms kxkXi on Xi are given (for i = 1, 2), then convex modulars can be
defined by Θi(x) = kxk2
; moreover, the Luxemburg norms associated to these modulars
coincide with the given norms and the m-direct sum coincides with the 2-direct sum. We
shall systematically equip Hilbert spaces with their trivial modular ΘH(x) = kxk2.
We say that a modular space (X, Θ) has a modular direct decomposition X = Y ⊕m Z if
Y , Z are linear subspaces of X and for every y ∈ Y, z ∈ Z we have Θ(y + z) = Θ(y)+ Θ(z).
Given two modular spaces (X1, Θ1) and (X2, Θ2), we say that a linear map T : X1 → X2
is modular preserving, or preserves modulars, if Θ2 ◦ T = Θ1. Note that such a T is
necessarily isometric (for Luxemburg norms). Two modular spaces (X1, Θ1) and (X2, Θ2)
are said to be linearly isomodular if there exists a modular preserving surjective linear
map from X1 onto X2. This implies that X1 and X2 are isometric, and the converse is
trivially true in the special case where Θi(x) = kxk2, i = 1, 2.
We say that a modular Θ satisfies the ∆2 condition with constant C if we have
Θ(2x) ≤ C · Θ(x)
for all x ∈ X.
In this case the modular is bounded by C n on the ball B(0, 2n), and thus, by the general
theory of convex functions, it is Lipschitz of constant at most C n on each ball B(0, 2n).
One may thus define unambiguously a function ΘU on each ultrapower XU by
(4)
ΘU ([xi]U ) = [Θ(xi)]U .
It is immediate that ΘU is convex and symmetric. It is faithful since ΘU ([xi]U ) = 0 means
Θ(xi) →i,U 0 which implies xi →i,U 0 by equivalence of modular and norm convergence
under the ∆2 condition. It is also easy to see using (3) that the Luxemburg norm associated
to ΘU coincides with the ultrapower norm on XU .
categoricity for the specific examples treated in the rest of the paper.
Our next result provides the main theoretical tool that we use in proving uncountable
2.1. Theorem. Let E be an infinite dimensional ∆2 modular Banach space such that:
i) Every ultrapower of E is linearly isomodular to E ⊕m H for some Hilbert space H.
ii) Every linear isometric operator E → E ⊕m H, where H is any Hilbert space, maps
E onto E.
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
5
then the elementary class of E consists exactly of all modular direct sums E ⊕m H, where
H is an arbitrary Hilbert space, and hence this class is uncountably categorical (and E is
separable).
Furthermore, if E0 is a finite dimensional ∆2 modular Banach space such that
ii') Any linear isometric operator E0 ⊕m ℓ2 → E0 ⊕m H, where H is any infinite
dimensional Hilbert space, maps E0 onto E0
then the elementary class of E = E0 ⊕m ℓ2 consists exactly of all modular direct sums
E0 ⊕m H, where H is an arbitrary infinite dimensional Hilbert space, and hence this class
is κ-categorical for every infinite cardinal κ.
2.2. Remark. Condition (i) implies that every ultrapower EU has a modular direct sum
decomposition EU = E1 ⊕m H, where E1 ⊂ EU is isomodular with E and H with an
Hilbert space. Condition (ii) applied to the diagonal embedding D : E → EU implies that
E1 = D(E), thus E verifies eq. (1). However H has no clear reason to be the kernel of
the weak limit projection, and so E may not be asymptotially hilbertian.
Proof of Theorem 2.1. Assume that E is infinite dimensional and let C be the class of all
Banach spaces that are linearly isometric to E ⊕m K, for some Hilbert space K. This
class is closed under ultrapowers since
(E ⊕m K)U = EU ⊕m KU is isomodular with (E ⊕m H) ⊕m KU = E ⊕m K′
where K′ = H ⊕m HU is also a Hilbert space. On the other hand, for every uncountable
cardinal κ strictly bigger than the density character of E, the only member (up to linear
isometry) of the class C having density character κ is E ⊕m Hκ, where Hκ is the Hilbert
space of hilbertian dimension κ. To complete the proof of the Theorem it remains only to
prove that the class C is also closed under ultraroots. Once this is proved, it follows that
E must be separable, since the elementary class of E contains a separable space.
Let X be a Banach space such that some ultrapower XU of X is linearly isometric
to a space E ⊕m H. Let J : E ⊕m H → XU be such an isometry, DX : X ֒→ XU be the
diagonal embedding and i = J−1DX be the resulting embedding of X into E⊕mH. Taking
ultrapowers, we get an embedding iU : XU ֒→ (E ⊕m H)U . By the preceding argument
we have a modular direct decomposition (E ⊕m H)U = E1 ⊕m K, where E1 is isometric
(in fact, isomodular) to E and K is an Hilbert space. Let D : E ⊕m H ֒→ (E ⊕m H)U
be the diagonal embedding. Using assumption (ii) we have that DE is an isometry from
E onto E1, in particular E1 = D(E). We set S = iU J; this is an isometric embedding
E⊕m H ֒→ (E⊕m H)U which need not coincide with the diagonal embedding D. However,
using assumption (ii) again, we do have that SE is an isometry from E onto D(E). To
summarize, we have the following commutative diagram of linear isometries:
XU
J
iU
◆◆◆◆◆◆◆◆◆◆◆◆
7♣♣♣♣♣♣♣♣♣♣♣
D
S
E ⊕m H
/ (E ⊕m H)U = D(E) ⊕m K
X
DX
;✈✈✈✈✈✈✈✈✈✈
❍❍❍❍❍❍❍❍❍
i
i
E ⊕m H
Note that iU (XU ) ∩ D(E) ⊂ Di(X). Indeed if ξ = [xi]U ∈ XU and z ∈ E are such that
z. Since i(X) is a closed subspace of E ⊕m H
iU ([xi]U ) = [i(xi)]U = D(z) then i(xi)−→U
this implies that z ∈ i(X).
'
'
;
/
/
#
#
O
O
/
7
6
C. W. HENSON AND Y. RAYNAUD
For every x ∈ E we have iU J(x) = S(x) ∈ D(E); hence by the preceding argument
there is x′ ∈ X such that S(x) = Di(x′). As SE : E → D(E) is surjective this shows that
D(E) ⊂ Di(X) and thus E ⊂ i(X).
Let π : E⊕m H → E be the first projection on the direct sum E⊕m H. By the preceding
argument, the range of π is contained in i(X). Let π′ be the restriction of π to i(X); since
E ⊂ i(X), the range of π′ is E. Its kernel H0 is contained in that of π, and thus H0 ⊂ H
is a Hilbert space. Finally i(X) = E ⊕ H0, with the norm induced by that of E ⊕m H.
That is, i(X) = E ⊕m H0.
Now suppose E0 is a finite dimensional ∆2 modular Banach space that satisfies (ii).
Let C be the class of all Banach spaces that are linearly isometric to E0 ⊕m K, for some
Hilbert space K. Since (E0)U = E0 for any ultrafilter U , the reasoning above shows that
C is closed under ultrapowers and ultraroots. However, the members of this class are
not all mutually elementarily equivalent, since any finite dimensional member X of C has
trivial ultrapowers (XU = X), and thus cannot be elementarily equivalent to an infinite
dimensional one. The subclass C∞ consisting of all the infinite dimensional members of C
is also closed under ultrapowers and ultraroots, and has a unique member of any density
character (up to linear isometry). Thus C∞ is the elementary class of its unique separable
member E = E0 ⊕m ℓ2.
It is routine to verify that the previous argument for proving closedness by ultraroots
also works for C∞ under hypothesis (ii'), which is formally weaker than (ii).
(cid:3)
We now introduce the examples that are treated in the next sections and summarize
their connections to the hypotheses of the preceding Theorem; proofs of what we state
here are given in Sections 3, 4, and 5.
Typical examples of modular spaces satisfying condition (i) of Theorem 2.1 are the
Nakano sequence spaces ℓ(pn) associated to a sequence of exponents (pn) converging to
2. The space N := ℓ(pn) is the linear space of sequences of (real or complex) scalars
x = (x(n)) such that
Θ(x) :=
∞
Xn=1
x(n)pn < ∞
and Θ is a natural convex modular on N which verifies condition ∆2 (since the sequence
(pn) is bounded). The Luxemburg norm on N is then defined by (2), and N is complete
for this norm. Since x ≤ y clearly implies Θ(x) ≤ Θ(y), it implies also kxk ≤ kyk;
i.e., N is a Banach lattice (for the Luxemburg norm). It is easy to see that if (xn) is a
decreasing sequence of elements of N which converges to zero coordinatewise then Θ(xn)
converges to zero and so does kxnk; hence N is order-continuous.
If pn 6= 2 for all n, then N also satisfies condition (ii) of Theorem 2.1 (if pn is allowed
to equal 2, this condition will only be true for the subspace N0 = span [en : pn 6= 2]).
The space N is linear-topologically isomorphic to ℓ2 if a certain summation condition due
to Nakano holds (see Fact 3.3 below). On the other hand, it is possible to choose the
exponents to yield Nakano spaces that satisfy both conditions of Theorem 2.1 but are not
linear-topologically isomorphic to a Hilbert space, and they provide new kinds of examples
of uncountable categoricity.
A bigger variety of examples appears when we consider vector-valued Nakano sequence
spaces, that is, direct sums of a sequence of finite dimensional Banach spaces. The elements
of this direct sum are sequences of vectors, the norms of which form a sequence belonging
to a given Nakano space. We denote by (⊕n
En)N the Nakano direct sum associated to the
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
7
family of spaces (En) and the Nakano space N . Thus
(cid:0)⊕n
En(cid:1)N = {(x(n)) ∈ Yn
En : (kx(n)kEn ) ∈ N}.
A convex modular and norm are defined on the vector-valued Nakano sequence space by
taking the Nakano modular of the sequence of norms:
Θ(x) :=
∞
Xn=1
kx(n)kpn
En
and then using the associated Luxemburg norm. We show that the Nakano direct sum
En)N satisfies condition (i) of Theorem 2.1 whenever the Jordan-von-Neumann con-
(⊕n
stants (defined below, at the beginning of Section 3) of the spaces En converge to 1 (here
pn may take the value 2). These constants measure the degree of approximation to which
the spaces satisfy the parallelogram inequality. It is equivalent to require that the Banach-
Mazur distances from 2-dimensional subspaces of En to the 2-dimensional Hilbert space
converge uniformly to 1. (Note that this is a far weaker condition than saying that the
Banach-Mazur distances from En to the Hilbert space of the same dimension converge to
1.)
On the other hand, provided pn 6= 2 for all n, the Nakano direct sum (⊕n
En)N satisfies
condition (ii) of Theorem 2.1 without any condition on the spaces En except that they
are finite dimensional. The reason is that isometries distinguish the spaces En (or H) by
the value of the corresponding exponent pn (resp. 2 for H).
In the opposite case where pn is constantly 2, in which case N = ℓ2 and the N -direct
sum is a 2-sum, the preceding argument does not work, and isometries recognize the En
spaces rather by their geometric properties. For this reason the conditions on the spaces
En that we assume in this case are far more restrictive than in the preceding case (the
examples are essentially the ℓp
n spaces or their non-commutative analogues, the Schatten
classes Sp
n).
3. Asymptotically hilbertian Nakano direct sums
For a normed space X, its Jordan-von Neumann constant a(X) is defined by
(5)
a(X) =
1
2
sup{kx + yk2 + kx − yk2 : x, y ∈ X,kxk2 + kyk2 = 1}.
By setting u = x + y, v = x − y it is immediate that we also have
(6)
It follows that a(X) ≥ 1, and that a(X) = 1 iff X is linearly isometric to a Hilbert space
[7]. Note that a(X) is the norm of the operator
a(X) = 2 sup{kxk2 + kyk2 : x, y ∈ X,kx + yk2 + kx − yk2 = 1}.
MX : ℓ2
2(X) → ℓ2
2(X) : (x, y) 7→
1
√2
(x + y, x − y) .
The conjugate operator is easily seen to be MX ∗, so that a(X∗) = a(X).
Let N be a Nakano sequence space with exponent sequence (pn) converging to 2.
Let (en) be the sequence of units of N (en is the sequence (δkn)k∈N). For every n ∈ N and
x ∈ N let Pn(x) = Pn
k=1 x(k)ek. Clearly Pn is a projection of norm one on N . On the
other hand Θ(x−Pn(x)) = P∞k=n+1 x(k)pk → 0 when n → ∞, which by the ∆2-condition
for Θ implies that kx− Pn(x))k → 0. It follows that (en) is a Schauder basis for N and the
generic element of N can be written x = P∞n=1 x(n)en. (This basis is clearly unconditional
with constant 1; in fact, it consists of mutually disjoint atoms of the Banach lattice N ).
8
C. W. HENSON AND Y. RAYNAUD
Let E = (cid:18)⊕n
be the N -direct sum of a family (En) of Banach spaces and Θ its
En(cid:19)N
modular, as defined in section 2. Note that since (pn) is bounded, the modular Θ satisfies
a ∆2 condition.
If we set ν(x) = (kx(n)kEn ) we have clearly ΘE(x) = ΘN (ν(x)) and
kxkE = kν(x)kN . The map ν : X → N is clearly 1-Lipschitz.
If (en) is the natural basis of the Nakano space (as above), it can be useful to denote
the generic element x = (xn) of E by Pn en ⊗ x(n). Setting Pn(x) = Pn
k=1 ek ⊗ x(k), the
map Pn is a linear projection on E of norm one and for every x ∈ E
kx − Pn(x)kE = kν(x) − Pn(ν(x))kN → 0
when n → ∞.
Since the Banach lattice N is order continuous, its dual space N∗ is also a Banach
sequence space. It is well known that the dual space N∗ to the N -direct sum E = (⊕n
En)N
E∗n)N ∗. If moreover (pn) is bounded away from 1, then N∗ is the Nakano
is then E∗ = (⊕n
sequence space with the conjugate exponent sequence (p∗n), with an equivalent norm, and
since the spaces En are finite dimensional we have that E∗∗ = (⊕n
En)N = E
(with the same norm). Thus in this case E is reflexive. This remains true if a finite
number of the pn equal 1 and the remainder of the exponents are bounded away from 1
(e.g., when the sequence (pn) converges to 2).
E∗∗n )N = (⊕n
3.1. Proposition. If the Nakano sequence space N has its exponent sequence converging to
2, and if the linear spaces En are finite dimensional, then every ultrapower of their Nakano
direct sum E = (cid:0)⊕n
En(cid:1)N has the modular decomposition EU = E ⊕m H, where E is the
diagonal copy of E in EU and H is the kernel of the weak limit projection P : EU → E.
Moreover
ΘU (x + h) = Θ(x) + khk2
for every x ∈ E, h ∈ H. If, moreover, the Jordan-von Neumann constants a(En) converge
En(cid:1)N is
to 1, then the space H is linearly isometric to a Hilbert space, and E = (cid:0)⊕n
asymptotically hilbertian.
Proof. 1) The modular decomposition of EU . As explained at the beginning of section 2,
the canonical image D(E) of E in EU is the range of the weak limit projection P . Let
H = ker P ; we prove first that the direct sum EU = E ⊕ H is modular.
Note that if ξ ∈ H, then for every bounded family (xi) in E representing ξ and every
n ∈ N we have
w − lim
i,U
xi(n) = 0 .
If x ∈ E has finite support relative to N , that is nx = sup{n : x(n) 6= 0} < ∞, and ξ ∈ H,
we can find a representing family (xi) for ξ with xi(n) = 0 for every n ≤ nx and i. Then
Θ(x + xi) = Θ(x) + Θ(xi), which shows that
ΘU (x + ξ) = Θ(x) + ΘU (ξ) .
This equality extends to every x ∈ E, by density of finitely supported elements in E and
continuity of ΘU .
2) ΘU (ξ) = kξk2 whenever ξ ∈ H. It suffices to prove that the modular restricted to H
is 2-homogeneous. Given ξ ∈ H and n ∈ N, we can choose a family (xi) in E, representing
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
9
ξ and such that xi(k) = 0 for k = 1 . . . n and all i ∈ I. Then for every λ ∈ K and i ∈ I
(cid:12)(cid:12)Θ(λxi) − λ2Θ(xi)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
kλxi(k)kpk − λ2 ∞
Xk=n
kxi(k)kpk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
∞
Xk=n
∞
k≥n (cid:12)(cid:12)λpk − λ2(cid:12)(cid:12) Θ(xi) .
Xk=n(cid:12)(cid:12)λpk − λ2(cid:12)(cid:12)kxi(k)kpk ≤ max
≤
It follows that
k≥n (cid:12)(cid:12)λpk − λ2(cid:12)(cid:12) ΘU (ξ) .
Then since pn → 2, by letting n → ∞ we obtain ΘU (λξ) = λ2ΘU (ξ).
(cid:12)(cid:12)ΘU (λξ) − λ2ΘU (ξ)(cid:12)(cid:12) ≤ max
3) H is linearly isometric to a Hilbert space. If x, y ∈ En and pn ≤ 2 we have
kx + ykpn + kx − ykpn
2
≤ a(En)pn/2(cid:0)kxk2 + kyk2(cid:1)pn/2
(cid:19)pn/2
≤ (cid:18)kx + yk2 + kx − yk2
≤ a(En)pn/2 (kxkpn + kykpn) .
2
If pn ≥ 2 the inequalities are reversed
kx + ykpn + kx − ykpn
2
≥ a(En)−pn/2(kxkpn + kykpn) .
Setting u = x + y and v = x − y we obtain in this case
u + v
pn
kukpn + kvkpn
2
≥ a(En)−pn/2(cid:18)(cid:13)(cid:13)(cid:13)(cid:13)
2 (cid:13)(cid:13)(cid:13)(cid:13)
pn(cid:19)
+(cid:13)(cid:13)(cid:13)(cid:13)
u − v
2 (cid:13)(cid:13)(cid:13)(cid:13)
and relabelling the variables and rearranging the preceding inequality
kx + ykpn + kx − ykpn
2
≤ 2pn−2a(En)pn/2 (kxkpn + kykpn) .
Set αn = a(En)pn/2 max(1, 2pn−2); we then have αn → 1 when n → ∞.
Now assume that x, y ∈ E with x(k) = 0 = y(k) whenever k < n. We then have
Θ(x + y) + Θ(x − y)
≤ βn(Θ(x) + Θ(y))
where βn = sup{αk : k ≥ n}. Observe that βn → 1 when n → ∞.
Passing to the ultrapower, consider ξ, η ∈ EU that are represented by families (xi) and
(yi) respectively, with xi(k) = 0 = yi(k) for all k < n and i ∈ I. By the previous argument
we have
2
≤ βn(ΘU (ξ) + ΘU (η)) .
When ξ, η ∈ H, the preceding inequality is valid for all n ∈ N, hence
2
ΘU (ξ + η) + ΘU (ξ − η)
ΘU (ξ + η) + ΘU (ξ − η)
2
≤ ΘU (ξ) + ΘU (η) .
Since ΘU (ξ) = kξk2 whenever ξ ∈ H, the Banach space H satisfies the parallelogram
inequality and thus is linearly isometric to a Hilbert space.
(cid:3)
3.2. Remark. Proposition 3.1 suggests that Nakano spaces like N (as in Corollary 4.4)
are "close to being hilbertian". However they need not be linearly isomorphic to a Hilbert
space. Indeed, from a result of Nakano himself [9] it is easy to deduce the following fact:
3.3. Fact. The Nakano space N = ℓ(pn) has an equivalent hilbertian norm iff for some
c > 0 the series P∞n=1 c
2pn
pn−2 is convergent.
10
C. W. HENSON AND Y. RAYNAUD
Indeed by Theorem 1 in [9] this condition is necessary and sufficient for the unit vector
basis of N to be equivalent to the ℓ2 basis.
If N has an equivalent hilbertian norm,
its unit vector basis is not necessarily an orthonormal basis for the hilbertian structure,
but it remains unconditional in the hilbertian norm (since unconditionality is preserved
by linear isomorphisms, although the unconditionality constant may, of course, change).
Since unconditional bases in a Hilbert space are all equivalent to the ℓ2 unit basis, Nakano's
condition must then hold true.
4. First example: Nakano direct sums
The main result of this section (Corollary 4.4) states that every Nakano direct sum of
finite dimensional normed spaces associated to a Nakano space N with exponent sequence
converging to 2, but different from 2, satisfies condition (ii) of Theorem 2.1. The isometries
of general Nakano spaces with exponent function strictly greater than 2 were studied in
the article [6]. Here we can avoid the latter restriction on (pn) by taking advantage of
the fact that the Banach lattice N is atomic. We state first a Proposition where the
condition pn 6= 2 is not required, with a view to getting some partial results also in this
case (Corollary 4.5 and Remark 4.7).
4.1. Proposition. Let H be any Hilbert space, and N be a Nakano sequence space with
exponent sequence pn → 2. Further, let (En) be a sequence of finite dimensional normed
spaces and E = (⊕n
En)N their N -direct sum. Consider also the partial direct sums Eh
and Enh corresponding respectively to the set of indices {n : pn = 2}, and {n : pn 6= 2}.
Then every linear isometric embedding from E into E ⊕m H is modular preserving and
maps Enh onto Enh and Eh into Eh ⊕m H.
We first give two lemmas in preparation for the proof of Proposition 4.1. Fix N ,
(En), and E = (⊕n
En)N as in the hypotheses of the Proposition. We regard K as the
1-dimensional modular space by taking its modular to be the square of the absolute value.
If a sequence (xn) in a Hilbert space H is weakly null, then for every x ∈ H we have
kx + xnk2 − (kxk2 + kxnk2) → 0 .
In particular if kxnk → a then kx + xnk2 → kxk2 + a2. An analogous property for E is
given by the next lemma.
4.2. Lemma. Let (xn) be a weakly null sequence in E with Θ(xn) → 1. Then for every
x ∈ E and t ∈ K, we have
Θ(x + txn) → Θ(x) + t2 and kx + txnkE → kx + tkE⊕mK .
Proof. Let U be a nonprincipal ultrafilter on N. Since xn → 0 weakly, the corresponding
element ξ = [xn]U of EU belongs to ker P = H. Thus
Θ(x + txn) = ΘU (x + tξ) = Θ(x) + t2ΘU (ξ) = Θ(x) + t2 .
lim
n,U
Since the U -limit does not depend on U , it is also the ordinary limit. Similarly
lim
n,U kx + txnkE = kx + tξkE⊕mH = inf{λ > 0 : Θ(cid:16) x
λ(cid:17) +
t2
λ2 = 1} = kx + tkE⊕mK .
The next lemma, valid for any modular space M , yields an estimation of kx + tkM⊕mK
for small x ∈ M . By homogeneity of the norm, it suffices to do this for t = 1.
(cid:3)
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
11
4.3. Lemma. Let M be a modular space with convex modular Θ satisfying a ∆2 condition.
Then for x → 0 in M we have
kx + 1kM⊕mK − 1 ∼
Θ(x) .
1
2
Proof. We have
hence
1 = Θ(cid:18) x + 1
kx + 1km(cid:19) = Θ(cid:18)
m − 1 = kx + 1k2
kx + 1k2
1
x
kx + 1km(cid:19) +
m Θ(cid:18)
kx + 1k2
kx + 1km(cid:19) ∼ Θ(x)
x
m
m − 1 ∼ 2(kx + 1km − 1).
and the result follows since kx + 1k2
(cid:3)
Proof of Proposition 4.1. Let T be an isometric embedding from E into E ⊕m H. We
denote by Θ the modular on E ⊕m H; that is, Θ(x + h) = Θ(x) + khk2 whenever x ∈
E, h ∈ H. Let (¯pk)k∈N be an enumeration of the distinct values of the exponent sequence
(pn), and ¯Ek = span {en ⊗ En : pn = ¯pk}. It will be sufficient to prove that T maps each
¯Ek into itself, and Eh into Eh ⊕ H:
indeed since T is injective and the spaces ¯Ek with
¯pk 6= 2 have finite dimension, so T will in fact map each of those ¯Ek onto itself, and thus
Enh onto Enh. Moreover for x ∈ ¯Ek, Θ(x) = kxk¯pk , and for x ∈ H, Θ(x) = kxk2, so that
the isometry T will be modular preserving.
Let j ∈ N be fixed, and for every n ∈ N choose xn ∈ En with kxnk = 1. Then xn → 0
weakly in E, and for every x ∈ E we have by Lemma 4.2
kx + xnkE → kx + 1kE⊕mK .
Then since T is an isometry
(7)
kT x + T xnkE⊕mH = kx + xnkE −→n→∞kx + 1kE⊕mK .
On the other hand, since xn → 0 weakly in E as n → ∞, we see T xn → 0 weakly in E⊕H.
If we let T xn = un + vn, un ∈ E, vn ∈ H be the decomposition of T en in the direct sum
E ⊕ H, then we have separate weak convergences un → 0 in E and vn → 0 in H. Consider
a subsequence (nk) such that Θ(unk ) and kvnkk2 converge to, say, a2 and b2 respectively.
Note that a2 + b2 = limk Θ(T xnk ) = 1 since kT xnk = kxnk = 1. Let T x = u + v, with
u ∈ E and v ∈ H. Then for every λ > 0
Θ((λT x + T xnk )) = Θ(λ(u + unk )) + kλ(v + vnk )k2
→ Θ(λu) + λ2a2 + kλvk2 + λ2b2
= Θ(λT x) + λ2 .
(by Lemma 4.2)
Since the limit is independent of the subsequence, Θ(λ(T x + T xn)) → Θ(λT x) + λ2 and
choosing λ = (kT x + 1k(E⊕mH)⊕mK)−1 it follows that
(8)
kT x + T xnkE⊕mH → kT x + 1k(E⊕mH)⊕mK .
Comparing (7) and (8), we see that
Assume now that x ∈ ¯Ej. By Lemma 4.3 we have for λ → 0
kx + 1kE⊕mK = kT x + 1k(E⊕mH)⊕mK .
kλx + 1km − 1 ∼
Θ(λx) =
1
2
1
2λ¯pjkxk¯pj
and
kλT x + 1km − 1 ∼
1
2
Θ(λT x)
12
hence
C. W. HENSON AND Y. RAYNAUD
Θ(λT x) ∼ λ¯pjkxk¯pj .
Assume that kxk = 1. We decompose T x = Pk uk + v, where uk ∈ ¯Ek and v ∈ H. Then
(9)
Θ(λT x) = Xk
Θ(λuk) + kλvk2
H = Xk
λ¯pk Θ(uk) + λ2kvk2
H
and hence
(10)
1 ∼ λ−¯pj Θ(λT x) = Xk
λ¯pk−¯pj Θ(uk) + λ2−¯pjkvk2
H .
This implies that Θ(uk) = 0 if ¯pk < ¯pj and v = 0 if 2 < ¯pj (otherwise the right side of
equation (10) goes to +∞ when λ → 0). On the other hand
λ¯pk−¯pj Θ(uk) → 0 and λ2−¯pjkvk2
H → 0 if 2 > ¯pj
Xk: ¯pk>pj
thus the right side of (10) is equivalent to Θ(uj). We have thus Θ(uj) = 1. Since
kT xk = kxk = 1 we have Θ(T x) = 1 and thus equality (9) with λ = 1 shows that
Θ(uk) = 0 for all k 6= j, as well as kvkH = 0 if 2 6= ¯pj. Finally, we conclude T x ∈ ¯Ej as
desired, except if ¯pj = 2, in which case T x ∈ ¯Ej ⊕ H.
(cid:3)
By Propositions 3.1 and 4.1 and Theorem 2.1 we conclude:
4.4. Corollary. Let N be a Nakano sequence space with exponent sequence (pn), where
pn 6= 2 for all n, and pn converges to 2. Let (En) be a sequence of finite dimensional
normed spaces whose Jordan-von Neumann constants converge to 1. Then the elementary
class of the Nakano direct sum E = (⊕En)N is equal to the class of all modular direct
sums E ⊕m H of E with arbitrary Hilbert spaces, and hence it is uncountably categorical.
In the scalar case we can drop the condition pn 6= 2. Denote by Nnh the Nakano space
associated to the subsequence (which may be finite or not) that consists of the exponents
that differ from 2 .
4.5. Corollary. Let N be a Nakano sequence space with exponent sequence (pn) converging
to 2. Then the elementary class of N consists of:
space H, if an infinity of exponents pn differ from 2.
-- the class of all modular direct sums Nnh ⊕m H of Nnh with an arbitrary Hilbert
-- the class of all modular direct sums Nnh ⊕m H of Nnh with an infinite dimensional
Hilbert space H, if not.
In both cases this class is uncountably categorical.
Proof. If Nnh is infinite dimensional, then by Corollary 4.4, its elementary class consists of
all modular direct sums Nnh⊕m H of Nnh with an arbitrary Hilbert space H. In particular
N is elementarily equivalent to Nnh, and thus its elementary class is the same.
If Nnh is finite dimensional, then it follows from Proposition 4.1 that every isometric
linear embedding from Nnh ⊕m ℓ2 into Nnh⊕m H sends Nnh onto itself. Then by Theorem
2.1, the elementary class of Nnh ⊕m ℓ2 consists of all modular direct sums Nnh ⊕m H of
Nnh with an arbitrary infinite dimensional Hilbert space H. This class contains N , the
elementary class of which is thus the same.
(cid:3)
4.6. Remark. We have a similar result to Corollary 4.5 for Nakano direct sums if we
require that En is 1-dimensional whenever pn = 2.
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
13
Fn)ℓ2,
4.7. Remark. In section 5 we shall give several examples of 2-direct sums F = (⊕n
where the normed spaces Fn are finite dimensional, such that F satisfies the hypotheses
En)N ,
of Theorem 2.1. Given such an F , we consider further a Nakano direct sum E = (⊕n
where N is a Nakano sequence space with exponents (pn) all distinct from 2 and such
that either (pn) is finite or it converges to 2. Then the modular sum E ⊕m F satisfies the
hypotheses of Theorem 2.1. Indeed, if H is any Hilbert space, it follows from Proposition
4.1 that every linear isometric map from E ⊕m F into E ⊕m F ⊕m H maps E onto E and
F into F ⊕m H, and thus F onto F .
In the spirit of the preceding Remark we now give a general lemma about isometries of
modular direct sums, which will have several applications in the next section.
4.8. Lemma. Let E1, F1, E2, F2 be modular spaces, and T : E1 ⊕m F1 → E2⊕m F2 a linear
isometric embedding. If T (E1) = E2 then T (F1) ⊂ F2.
Proof. Let f ∈ F1 with kfk = 1 and decompose T f = x + g with x ∈ E2 and g ∈ F2. We
have
(11)
Since T : E1 → E2 is onto we can find y ∈ E1 with x = T y, and thus g = T (f − y). Then
1 = Θ(T f ) = Θ(x) + Θ(g) .
kgk = kT (f − y)k = kf − yk ≥ kfk = 1
because in the modular sum, factor projections are contractive. Using (11) we find that
Θ(x) = 0 and thus x = 0.
(cid:3)
En)ℓ2, where the normed
5. Second example: 2-direct sums
In this section we consider direct sums of the form E = (⊕n
spaces En are finite dimensional. The norm of x = (xn) is given by kxk2 = P∞n=1 kxnk2
.
Since this is a special case of Nakano direct sum, with constant exponent function pn ≡ 2,
Proposition 3.1 of Section 3 applies: if an(En) → 1 then E is asymptotically hilbertian; in
fact EU = E ⊕2 H for some Hilbert space H depending on U . However in contrast to the
Nakano direct sums treated in Section 4, here we need some relatively strong hypotheses on
the En's for proving that the isometric embeddings from E into E ⊕2 H send each En into
itself. See Propositions 5.2 and 5.6 and Corollary 5.8 for the kinds of assumptions under
which we have been able to carry out the required arguments. Example 5.9 summarizes
the specific examples E = (⊕n
En)ℓ2 for which we prove uncountable categoricity in this
section.
5.1. Remark. If the Banach-Mazur distance d(En, ℓdn
2 ) is not bounded, where dn =
dim En, then the space E is not linear-topologically isomorphic to a Hilbert space. Indeed
if E is C-linearly isomorphic to a Hilbert space, so is every closed linear subspace of E.
En
5.2. Proposition. Let (En) be a sequence of finite dimensional Banach spaces. Assume
that for some sequence of exponents pn > 2, with pn → 2, the following conditions are
satisfied:
a) For every n, and every x, y ∈ En we have
kx + yk2
En + kx − yk2
En ≥ 2(kxkpn
En
+ kykpn
En
)2/pn ;
b) For every n there exists a basis Bn of En such that for every y ∈ Bn there is x ∈ Bn
such that (x, y) is an ℓpn-pair, that is kx + λykpn
= 1 + λpn for every λ ∈ R.
Then the hypotheses of Theorem 2.1 are satisfied, and the Banach space E = (⊕n
En)ℓ2
En
is uncountably categorical.
14
C. W. HENSON AND Y. RAYNAUD
Proof. 1) By the hypothesis (a) and Holder's inequality we have for every x, y ∈ En
= 1
En + kyk2
En)
2 − 1
pn
En + kx − yk2
En ≥ 2 × 2−2/rn(kxk2
kx + yk2
with 1
. Thus by (6), a(En) ≤ 22/rn → 1 as n → ∞, and it follows from
rn
Proposition 3.1 that every ultrapower of E is of the form E ⊕2 H, where H is a Hilbert
space.
2) Let S : En → E ⊕2 H be a linear isometric embedding, we show that its range
is included in E (this will require only condition (b)). This will easily imply that any
linear isometric embedding T : E → E ⊕2 H sends E into E. Let Bn be a basis of En
as in the condition (b) of the proposition, it is sufficient to prove that Sy ∈ E for every
y ∈ Bn. Let x ∈ Bn be choosen such that (x, y) is an ℓpn-pair. Let u = Sx, v = Sy and
u = Pk u(k) + uH , v = Pk v(k) + vH, with u(k), v(k) ∈ Ek and uH , vH ∈ H. Then
ku + λvk2 + ku − λvk2
=
ku(k) + λv(k)k2 + ku(k) − λv(k)k2
2
(12)
2
∞
Xk=1
+ kuH + λvHk2 + kuH − λvHk2
2
.
Since ku ± λvkpn = kx ± λykpn = 1 + λpn, the left side of equation (12) is equal to
(1 + λpn)2/pn. On the other hand, by convexity of k · k2 and the parallelogram identity
in H, the right side of (12) is bigger than
∞
Xk=1
ku(k)k2 + (kuHk2 + λ2kvHk2) = kuk2 + λ2kvHk2 = 1 + λ2kvHk2 .
Thus, since pn > 2, we have
kvHk2 ≤
(1 + λpn)2/pn − 1
λ2
≤
2
pnλpn−2 −→λ→0
0 ;
hence vH, the H-component of v, is 0.
3) Now assuming both conditions (a) and (b) we show that the range of any isometric
Em. We keep the notation of the
embedding S : En → E ⊕2 H is included in ⊕pm≥pn
preceding part. For every m ≥ 1, the right side of equation (12) is by condition (a) greater
than
ku(k)k2 + (ku(m)kpm + λpmkv(m)kpm )2/pm + kuHk2 .
Xk6=m
Hence
(1 + λpn)2/pn ≥ 1 + (ku(m)kpm + λpmkv(m)kpm )2/pm − ku(m)k2
for every λ ∈ R. If u(m) = 0 we get
(1 + λpn)2/pn ≥ 1 + λ2kv(m)k2
and deduce v(m) = 0 in the same way we did for vH. If u(m) 6= 0 we get
(1 + λpn)2/pn − 1 ≥ ku(m)k2(cid:0)(1 + λkv(m)k/ku(m)k)pm − 1(cid:1)2/pm
2
pmku(m)k2−pmkv(m)kpmλpm as λ → 0 ;
∼
thus if pm < pn we get
kv(m)kpm .
pm
pn ku(m)kpm−2λpn−pm −→λ→0
0 .
Hence v(m), the Em-component of v, must vanish. Finally Sy ∈ ⊕pm≥pn
Em as was claimed.
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
15
4) Now let T : E → E ⊕2 H be a linear isometric embedding. Let us denote by (¯pk)
an enumeration of the distinct values of the pn's (for fixing ideas we may assume the
sequence (¯pk) to be strictly decreasing). Note that since pn > 2 and pn → 2, each set
En. By part (3) above, we
Ak = {n : pn = ¯pk} is finite. For every k ≥ 1 set Gk = ⊕pn≥¯pk
have that T (Gk) ⊂ Gk. Since Gk is finite dimensional and T is isometric it follows that
T (Gk) = Gk. Hence the range of T contains Sk Gk, a dense subspace of E, and since this
range is closed it contains E.
(cid:3)
5.3. Remark. We have in fact the more precise result that T ( ¯Ek) = ¯Ek for every k ≥ 1,
En. For k = 1 we have ¯E1 = G1 and thus T ( ¯E1) = ¯E1. For k ≥ 2 it will
where ¯Ek = ⊕pn=¯pk
be sufficient to prove that T ( ¯Ek) ⊂ ¯Ek. This is done inductively using Gk = ¯Ek ⊕m Gk−1
and Lemma 4.8.
5.4. Example. E = (cid:0) ⊕ ℓdn
pn(cid:1)2 with pn > 2, pn → 2 and dn ≥ 2 satisfies the hypotheses
of Proposition 5.2. Condition (b) is clearly satisfied. As for condition (a), we have for
x, y ∈ ℓp, p ≥ 2:
kx + yk2
= kx + y2kp/2 + kx − y2kp/2
p
p + kx − yk2
2
(by convexity since p/2 ≥ 1)
2
x + y2 + x − y2
2
≥ (cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)p/2
= kx2 + y2kp/2 = kx2 + y21/2k2
≥ k(xp + yp)1/pk2
p
= (kxkp
p)2/p .
p + kykp
p
(since p ≥ 2)
Thus E is uncountably categorical. On the other hand if d(En, ℓdn
E is not linear-topologically isomorphic to a Hilbert space.
1
2− 1
2 ) = d
n
pn
→ ∞ then
Let Bn be the basis of Sdn
pn(cid:1)2 with pn > 2, pn → 2, and dn ≥ 2, where Sd
5.5. Example. E = (cid:0) ⊕ Sdn
p is the
Schatten class of exponent p and dimension d2 (consisting of d × d matrices with complex
coefficients).
pn consisting of the matrix units (ei,j)1≤i,j≤dn. For each matrix
unit ei,j consider another matrix unit ek,ℓ with i 6= k, j 6= ℓ. Then the pair (eij, ekℓ) is
a ℓpn-pair in Sdn
pn , and the condition (b) in Proposition 5.2 is satisfied. As for condition
(a) we reason by interpolation. Indeed condition (a) means exactly that for p = pn the
inverse of the operator M : (x, y) 7→ ( x+y√2
p ). Since
2(Sd
these spaces are complex interpolation spaces, more precisely ℓ2
∞))θ
and ℓ2
p , it is sufficient to verify contractivity in the
cases p = 2 and p = ∞. For p = 2, Sd
2 is a Hilbert space and condition (b) follows
from the parallelogram identity. For p = ∞, Sd
p = Md(C) with the matrix norm (which
we denote by k · k∞) and we have for x, y ∈ Md
) is contractive from ℓ2
2(Sd
∞))θ for θ = 1 − 2
p = Sd
2(Sd
p ) = (ℓ2
p ) = (ℓ2
p ) to ℓ2
2(Sd
p(Sd
2 ), ℓ2
∞(Sd
2(Sd
2 ), ℓ2
, x−y√2
p(Sd
kx + yk2
∞ + kx − yk2
∞
= k(x + y)∗(x + y)k∞ + k(x − y)∗(x − y)k∞
2
2
2
(x + y)∗(x + y) + (x − y)∗(x − y)
≥ (cid:13)(cid:13)(cid:13)(cid:13)
= kx∗x + y∗yk∞
≥ max(kx∗xk∞,ky∗yk∞) = (cid:0) max(kxk∞,kyk∞)(cid:1)2 .
(cid:13)(cid:13)(cid:13)(cid:13)∞
16
C. W. HENSON AND Y. RAYNAUD
2− 1
Thus here again E is uncountably categorical. Note that if d
→ ∞, it follows from
n
[10, Theorem 2.1] that E is not linear-topologically embeddable in any space with local
unconditional structure with nontrivial cotype (in particular, any Banach lattice with
nontrivial concavity).
pn
1
Next we present another criterion for uncountable categoricity, similar to Proposition
5.2, but with exponents strictly less than 2.
5.6. Proposition. Assume that for some sequence of exponents 1 ≤ pn < 2, with pn → 2,
the following conditions are satisfied:
a) For every n, and every x, y ∈ En we have
kx + yk2
En + kx − yk2
En ≤ 2(kxkpn
En
+ kykpn
En
)2/pn .
b) For every n there exists a basis Bn of En such that for every y ∈ Bn there is x ∈ Bn
such that (x, y) is an ℓpn-pair, that is kx + λykpn
= 1 + λpn for every λ ∈ R.
Then the hypotheses of Theorem 2.1 are satisfied, and the Banach space E = (⊕n
En)ℓ2
En
is uncountably categorical.
Proof. 1) Hypothesis (a) implies that for every x, y ∈ En
En + kyk2
En )
En + kx − yk2
kx + yk2
2 , and therefore an(En) ≤ 22/rn by (5).
En ≤ 2 × 22/rn (kxk2
= 1
where 1
rn
pn's rearranged now in increasing order).
En(cid:1)2, where ¯p1 < ¯p2 < . . .
pn − 1
2) We prove that if S : ¯E1 → E ⊕2 H is an isometric embedding then S ¯E1 = ¯E1. (As
before, ¯Ek = (cid:0) ⊕pn=¯pk
is the sequence of distinct values of the
Let (x, y) be an ℓ¯p1-pair in ¯E1; we claim that the H-component as well as the En-
components for pn > ¯p1 of v = Sy all vanish. Let u = Sx, v = Sy; then for every
λ ∈ R
(1 + λ¯p1)2/¯p1 = ku + λvk2 + ku − λvk2
2
(13)
∞
Xn=1(cid:0)ku(n)k¯pn + kλv(n)k¯pn(cid:1)2/¯pn + kuHk2 + kλvHk2
≤
≤ Xpn=¯p1(cid:0)ku(n)k¯p1 + kλv(n)k¯p1(cid:1)2/¯p1 + Xpn≥¯p2(cid:0)ku(n)k¯p2 + kλv(n)k¯p2(cid:1)2/¯p2
+(cid:0)kuHk¯p2 + kλvHk¯p2(cid:1)2/¯p2 .
v(n) be the components of u, v in ¯E1; then by the
u(n), v ¯E1 = Ppn=¯p1
Let u ¯E1 = Ppn=¯p1
reverse Minkowski inequality in ℓ¯p1/2 (note that ¯p1/2 ≤ 1) we have
Xpn=¯p1(cid:0)ku(n)k¯p1 + kλv(n)k¯p1(cid:1)2/¯p1 ≤ (cid:18)(cid:0) Xpn=¯p1
ku(n)k2(cid:1)¯p1/2 +(cid:0) Xpn=¯p1
kλv(n)k2(cid:1)¯p1/2(cid:19)2/¯p1
Set G1 = (cid:0) ⊕k≥2
¯Ek ⊕ H(cid:1)2, and let uG1, vG1 be the components of u, v in G1; treating
similarly the last two terms in (13), we obtain
= (cid:0)ku ¯E1k¯p1 + kλv ¯E1k¯p1(cid:1)2/¯p1 .
(14)
(1 + λ¯p1)2/p1 ≤ (cid:0)ku ¯E1k¯p1 + kλv ¯E1k¯p1(cid:1)2/¯p1 +(cid:0)kuG1k¯p2 + kλvG1k¯p2(cid:1)2/¯p2 .
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
17
The left side of inequality (14) is
1 +
2
p1λ¯p1 + o(λ¯p1)
while the right side of (14) is
2
¯p1u ¯E1k2−¯p1kv ¯E1k¯p1λp1 + o(λ¯p1) + kuG1k2 +
ku ¯E1k2 +
2
¯p1u ¯E1k2−¯p1kv ¯E1k¯p1λ¯p1 + o(λ¯p1) .
= 1 +
2
¯p2uG1k2−¯p2kvG1k¯p2λ¯p2 + o(λ¯p2)
Comparing the leading terms of both sides of inequality (14) we obtain 1 ≤ u ¯E1k2−¯p1kv ¯E1k¯p1;
since u ¯E1k ≤ kuk = 1, kv ¯E1k ≤ kvk = 1 this implies ku ¯E1k = kv ¯E1k = 1. Then
uG1k2 = 1 − u ¯E1k2 = 0 and similarly vG1k2 = 0, and u, v ∈ ¯E1 as was claimed. Thus
S( ¯E1) ⊂ ¯E1, and in fact S( ¯E1) = ¯E1 since the dimension is finite.
3) If now T : E → E ⊕ H is an isometric embedding, then by part (2) we have
T ( ¯E1) = ¯E1. It follows that also T (G1) ⊂ G1 by Lemma 4.8. Now starting with G1 in
place of E ⊕ H, the reasoning of part (2) shows that T ( ¯E2) = ¯E2, etc.
(cid:3)
5.7. Examples. E = (cid:0) ⊕ ℓdn
pn(cid:1)2 with 1 ≤ pn < 2, pn → 2, and dn ≥ 2
pn(cid:1)2 and E = (cid:0) ⊕ Sdn
satisfy the hypotheses of Proposition 5.6.
The proof that these examples satisfy condition (a) of Proposition 5.6 is by duality
(a(En) = a(E∗n)).
In certain cases we can mix the examples of Propositions 5.2 and 5.6.
En)2 and F = (⊕n
Fn)2 be two direct sums satisfying re-
5.8. Corollary. Let E = (⊕n
spectively the hypotheses of Propositions 5.6 and 5.2 with respective exponent sequences
1 ≤ pn < 2 and 2 < qn < ∞. Assume moreover that for some constant C we have
(15)
∀x, y ∈ Fn, kx + yk2 + kx − yk2 ≤ 2(kxk2 + kCyk2)
(that is, the spaces Fn are uniformly 2-uniformly smooth in the sense of [3]).
Then every linear isometric embedding T of E ⊕2 F into E ⊕2 F ⊕2 H, where H is any
Hilbert space, maps E onto E and F onto F . In particular the hypotheses of Theorem 2.1
are satisfied by E ⊕2 F , and hence that space is uncountably categorical.
Proof. The proof that T maps E onto itself is the same as in Proposition 5.6 except that
we have to replace the Hilbert space H by the direct sum G = H ⊕ F . It follows from
Lemma 4.8 that T maps G into G. Then apply Proposition 5.2 to the restriction of T to
G.
(cid:3)
satisfy the hypotheses of Corollary 5.8.
pn(cid:1)2 and (cid:0) ⊕ Sdn
5.9. Examples. (cid:0) ⊕ ℓdn
pn(cid:1)2 with 1 ≤ pn < ∞, pn 6= 2, pn → 2, and dn ≥ 2
Proof. We recall a proof of (15) in the case of Lp-spaces, p ≥ 2. For any scalars x, y, by an
inequality of Beckner (see [8, 1.e.14] for the real case; the complex case is a special case
of [8, 1.e.15]) we have:
x + yp + x − yp
2
≤ (cid:18)x + Cp y2 + x − Cp y2
2
(cid:19)p/2
18
C. W. HENSON AND Y. RAYNAUD
with Cp = √p − 1. We deduce when x, y ∈ ℓd
x + yp
p + x − yp
p
p
2
p/2
p/2
2
(cid:18)x + Cp y2 + x − Cp y2
(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)
≤ (cid:13)(cid:13)(cid:13)(cid:13)
≤ (cid:0)kx2kp/2 + (p − 1)ky2kp/2(cid:1)p/2
= (cid:0)kxk2
p(cid:1)p/2
.
p + (p − 1)kyk2
Then by concavity of the function f (t) = t2/p:
= (cid:13)(cid:13)x2 + (p − 1)y2(cid:13)(cid:13)
p/2
p/2
(triangular inequality in ℓp/2)
x + y2
p + x − y2
2
p
≤ (cid:18)x + yp
p + (p − 1)kyk2
p .
For the Schatten class case use [3, Th. 1] and the preceding concavity argument.
(cid:3)
5.10. Remark. In examples 5.9 we can drop the condition dn ≥ 2 and show that these
spaces are uncountably categorical by reasoning similarly as in the proof of Corollary 4.5.
≤ kxk2
2
p + x − yp
p
(cid:19)2/p
6. Addendum: finite dimensional perturbations of Hilbert spaces
As a footnote to Section 5, we prove here that for any finite dimensional normed space
E, the direct sum E ⊕2 ℓ2 is κ-categorical for every infinite cardinal number κ. This is
a relatively simple example of categoricity and the proof is reasonably short; we present
it here for completeness. For a similar but partial result for finite dimensional modular
spaces and the modular direct sum, see Corollary 4.5 and the remark following it.
6.1. Definition. A linear projection P in a Banach space X is called a 2-projection if
∀x ∈ X kxk2 = kP xk2 + k(I − P )xk2 .
A closed linear subspace E in X is called a 2-summand if it is the range of a 2-projection.
In other words E is a 2-summand iff X = E ⊕2 F for some closed linear subspace F of
X.
6.2. Theorem. Let E0 is a finite dimensional normed space. The following assertions are
equivalent:
i) E0 has no one dimensional 2-summand;
ii) Any linear isometric embedding T of E0 into the 2-direct sum E0 ⊕2 H of E0 with
some Hilbert space H maps E0 onto E0.
To prepare for the proof of Theorem 6.2 we prove the following lemma:
6.3. Lemma. Let E0 be a Banach space without one dimensional 2-summand, and T be
an isometric linear embedding of E0 into a 2-direct sum E0 ⊕2 H of E0 with a Hilbert
space K. Then T (E0) ∩ H = (0).
Proof. It suffices to prove that if Y is a linear subspace of X := E0 ⊕2 H then any vector
ξ ∈ Y ∩ H generates a 2-summand in Y . Note that since H is a Hilbert space, every closed
subspace is a 2-summand in H, in particular H = Kξ ⊕2 ξ⊥. Then X = Kξ ⊕2 (E0 ⊕2 ξ⊥),
and Kξ is a 2-summand in X. Let P : X → X be the corresponding 2-projection in X
with range Kξ, then its restriction PY is a 2-projection in Y with range Kξ.
(cid:3)
Proof of Theorem 6.2. ii) =⇒ i): It is clear that if E0 has a 2-direct decomposition E0 =
E1 ⊕2 Kξ0, with ξ0 6= 0, one can define an isometric embedding T : E0 → E0 ⊕2 H by
defining T as the identity on E1 and T ξ0 = ξ1 ∈ H, with kξ1k = kξ0k.
Let us prove now the implication i) =⇒ ii). Let T : E0 → X := E0 ⊕2 H be an
isometric embedding and P : X → E0 be the 2-projection on E0 with kernel H. Then
ASYMPTOTICALLY HILBERTIAN SPACES AND UNCOUNTABLE CATEGORICITY
19
P T : E0 → E0 is a linear isomorphism, since P T x = 0 is equivalent to T x ∈ H, and
T (E0) ∩ H = (0) by Lemma 6.3. Since E0 is finite dimensional, P T : E0 → E0 is onto.
Let Q = IdX − P : X → X, which is the 2-projection on H with kernel E0. We then have:
kxk2 = kP T xk2 + kQT xk2
= k(P T )2xk2 + kQT P T xk2 + kQT xk2
= . . .
= k(P T )nxk2 +
n−1
Xk=0
kQT (P T )kxk2 .
Equivalently, we have a sequence of isometric linear embeddings Tn of E0 into E0 ⊕2 Hn,
where Hn = ℓn
2 (H) is the 2-sum of n copies of H, defined by
Tnx = ((P T )nx, QT x, QT P T x, . . . , QT (P T )n−1x) .
All these Hilbert spaces Hn can be isometrically embedded in the infinite 2-sum H∞ =
ℓ2(H). For defining a limit embedding of E0 into E0 ⊕2 H, fix a free ultrafilter U on N
and set
where
∞x = (SU x, QT x, QT P T x, . . . , QT (P T )nx, . . . )
T U
SU x = lim
n,U
(P T )nx .
This ultrafilter limit is well defined because P T is a contraction and E0 is finite dimen-
sional. Then SU : E0 → E0 is a linear contraction. Note that although SU x depends a
priori on the ultrafilter U , the norm kSU xk does not. In fact the sequence (k(P T )nxk) is
non-increasing, and thus convergent, so that
kSU xk = lim
n,U k(P T )nxk = lim
n→∞k(P T )nxk .
If P∞ denotes the 2-projection E0 ⊕2 H∞ → E0 with kernel H∞, we have clearly SU =
P∞T U . Since T U is a linear isometry it results again that SU is a linear isomorphism of
E0 onto E0. This map is contractive; let us show that it is in fact an isometry. We have
(SU )2x = SU lim
(P T )m+nx
n,U
SU (P T )nx = lim
n,U
(P T )nx = lim
(P T )m(P T )nx = lim
lim
lim
m,U
n,U
n,U
m,U
k→∞k(P T )kxk = kSU xk .
m,U k(P T )m+nxk = lim
kSU SU xk = lim
n,U
hence
Since SU : E0 → E0 is surjective, it follows that kSU yk = kyk for every y ∈ E0. Then
lim
∀x ∈ E0,
kxk = kSU xk ≤ kP T xk ≤ kxk
and it follows that, for each x ∈ E0, kP T xk = kT xk and thus T x ∈ E0.
6.4. Corollary. If E is a finite dimensional Banach space, then the elementary class
of E ⊕2 ℓ2 consists exactly of all spaces E ⊕2 H, where H is any infinite dimensional
Hilbert space. Therefore, the elementary class of E ⊕2 ℓ2 is totally categorical (i.e., it is
κ-categorical for every infinite cardinal number κ).
(cid:3)
Proof. Let K be a hilbertian subspace of E of largest dimension such that K is a 2-
summand of E, and let E0 be a subspace of E for which E = E0 ⊕2 K. Evidently E0
has no one dimensional 2-summand. Regarding E0 as a modular space with the modular
Φ(x) := kxk2 and applying Theorem 6.2, we see that condition (ii) in Theorem 2.1 is
satisfied by E0 (and a fortiori condition (ii')). Hence the elementary class of E0 ⊕2 ℓ2
20
C. W. HENSON AND Y. RAYNAUD
consists exactly of all spaces E0 ⊕2 H, where H is any infinite dimensional Hilbert space.
The proof is completed by noting that for any infinite dimensional Hilbert space H, the
spaces E ⊕2 H and E0 ⊕2 H are linearly isometric.
(cid:3)
References
[1] I. Ben Yaacov, Uncountable dense categoricity in cats, J. Symb. Logic 70 (2005), no. 3, 829 -- 860.
[2] I. Ben Yaacov, A. Berenstein, C. W. Henson, and A. Usvyatsov, Model theory for metric structures.
Model Theory with Applications to Algebra and Analysis, Vol. 2 (Z. Chatzidakis, D. Macpherson, A.
Pilllay, and A. Wilkie ed.), Cambridge Univ. Press (2008), 315 -- 427.
[3] K. Ball, E. A. Carlen, and E. H. Lieb, Sharp uniform convexity and smoothness inequalities for trace
norms, Invent. Math. 115 (1994), no. 3, 463-482.
[4] S. Heinrich, Ultraproducts in Banach Spaces Theory. J. Reine Angew. Math. 313 (1980), 72-104.
[5] C. W. Henson, J. Iovino, Ultraproducts in Analysis. Analysis and Logic (C. Finet, C. Michaux ed.),
Cambridge Univ. Press (2003), 1 -- 113.
[6] J.E. Jamison, A. Kami´nska, and Pei-Kee Lin, Isometries of Musielak-Orlicz spaces II, Studia Math.
104 (1993), 75-89.
[7] P. Jordan, J. von Neumann, On Inner Products in Linear, Metric Spaces, Ann. of Math. (2) 36 (1935),
no. 3, 719-723.
[8] J. Lindenstrauss, L. Tzafriri, Classical Banach spaces II, Springer-Verlag, 1979.
[9] H. Nakano, Modulared sequence spaces, Proc. Japan Acad. 27 (1951), 508-512.
[10] G. Pisier, Some results on Banach spaces without local unconditional structure, Compositio Math. 37
(1978), no. 1, 3 -19.
[11] G. Pisier, The volume of convex bodies and Banach space geometry, Cambridge Univ. Press, 1989.
[12] S. Shelah, A. Usvyatsov, Model theoretic stability and categoricity for complete metric spaces. Israel
J. Math. 182 (2011), 157-198.
[13] S. Shelah, A. Usvyatsov, Minimal types in stable Banach spaces, preprint, (2014), arXiv:1402.6513.
C. Ward Henson, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA
Yves Raynaud, Institut de Math´ematiques de Jussieu-Paris Rive Gauche, CNRS/UPMC
(Univ.Paris 06)/Univ. Paris-Diderot, 4 place Jussieu, F-75252 Paris Cedex 05, France
|
1009.1751 | 3 | 1009 | 2012-01-03T12:33:57 | Average best $m$-term approximation | [
"math.FA",
"math.NA",
"math.ST",
"math.ST"
] | We introduce the concept of average best $m$-term approximation widths with respect to a probability measure on the unit ball of $\ell_p^n$. We estimate these quantities for the embedding $id:\ell_p^n\to\ell_q^n$ with $0<p\le q\le \infty$ for the normalized cone and surface measure. Furthermore, we consider certain tensor product weights and show that a typical vector with respect to such a measure exhibits a strong compressible (i.e. nearly sparse) structure. | math.FA | math |
Average best m-term approximation
Jan Vyb´ıral∗
October 23, 2018
Abstract
We introduce the concept of average best m-term approximation widths with
p . We
respect to a probability measure on the unit ball or the unit sphere of ℓn
estimate these quantities for the embedding id : ℓn
q with 0 < p ≤ q ≤ ∞
for the normalized cone and surface measure. Furthermore, we consider certain
tensor product weights and show that a typical vector with respect to such
a measure exhibits a strong compressible (i.e. nearly sparse) structure. This
measure may be therefore used as a random model for sparse signals.
p → ℓn
AMS subject classification (MSC 2010): Primary: 41A46, Secondary: 52A20,
60B11, 94A12.
Key words: nonlinear approximation, best m-term approximation, average widths,
random sparse vectors, cone measure, surface measure.
1 Introduction
1.1 Best m-term approximation
Let m ∈ N0 and let Σm be the set of all sequences x = {xj}∞
j=1 with
kxk0 := # supp x = #{n ∈ N : xn 6= 0} ≤ m.
Here stands #A for the number of elements of a set A. The elements of Σm are
said to be m-sparse. Observe, that Σm is a non-linear subset of every ℓq := {x =
{xj}∞
j=1 : kxkq < ∞}, where
kxkq :=
j=1 xjq(cid:17)1/q
(cid:16)P∞
supj∈N xj,
,
0 < q < ∞,
q = ∞.
For every x ∈ ℓq, we define its best m-term approximation error by
σm(x)q := inf
y∈Σm kx − ykq.
∗Johann Radon Institute for Computational and Applied Mathematics, Austrian Academy of
Sciences, Altenbergerstrasse 69, A-4040 Linz, Austria, email: [email protected], Tel: +43
732 2468 5262, Fax: +43 732 2468 5212.
1
Moreover for 0 < p ≤ q ≤ ∞, we introduce the best m-term approximation widths
σp,q
m := sup
x:kxkp≤1
σm(x)q.
The use of this concept goes back to Schmidt [44] and after the work of Oskolkov
[39], it was widely used in the approximation theory, cf.
[15, 18, 45]. In fact, it is
the main prototype of nonlinear approximation [17]. It is well known, that
m ≤ (m + 1)1/q−1/p, m = 0, 1, 2, . . . .
2−1/p(m + 1)1/q−1/p ≤ σp,q
(1)
The proof of (1) is based on the simple fact, that (roughly speaking) the best m-term
approximation error of x ∈ ℓp is realized by subtracting the m largest coefficients
taken in absolute value. Hence,
j=m+1(x∗
j )q(cid:19)1/q
,
m+1 = supj≥m+1 x∗
x∗
j ,
0 < q < ∞,
q = ∞,
(cid:18)P∞
σm(x)q =
where x∗ = (x∗
the vector (x1,x2,x3, . . . ).
from above then follows by
1, x∗
2, . . . ) denotes the so-called non-increasing rearrangement [6] of
Let us recall the proof of (1) in the simplest case, namely q = ∞. The estimate
σm(x)∞ = sup
j≥m+1
x∗
j = x∗
m+1 ≤(cid:18)(m + 1)−1
The lower estimate is supplied by taking
m+1
Xj=1
(x∗
j )p(cid:19)1/p
≤ (m + 1)−1/pkxkp.
(2)
x = (m + 1)−1/p
ej,
m+1
Xj=1
(3)
j=1 are the canonical unit vectors.
where {ej}∞
Holder's inequality
For general q, the estimate from above in (1) may be obtained from (2) and
kxkq ≤ kxkθ
p · kxk1−θ
∞ , where
1
q
=
θ
p
.
(4)
The estimate from below follows for all q's by simple modification of (3).
The discussion above exhibits two effects.
(i) Best m-term approximation works particularly well, when 1/p − 1/q is large,
i.e. if p < 1 and q = ∞.
(ii) The elements used in the estimate from below (and hence the elements, where
the best m-term approximation performs at worse) enjoy a very special struc-
ture.
Therefore, there is a reasonable hope, that the best m-term approximation could
behave better, when considered in a certain average case. But first we point out two
different interesting points of view on the subject.
2
1.2 Connection to compressed sensing
The interest in ℓp spaces (and especially in their finite-dimensional counterparts ℓn
p )
with 0 < p < 1 was recently stimulated by the impressive success of the novel and
vastly growing area of compressed sensing as introduced in [8, 10, 11, 19]. Without
going much into the details, we only note, that the techniques of compressed sensing
allow to reconstruct a vector from an incomplete set of measurements utilizing the
prior knowledge, that it is sparse, i.e. kxk0 is small. Furthermore, this approach
may be applied [14] also to vectors, which are compressible, i.e. kxkp is small for
(preferably small) 0 < p < 1. Indeed, (1) tells us, that such a vector x may be very
well approximated by sparse vectors. We point to [9, 24, 25, 42] for the current state
of the art of this field and for further references.
This leads in a very natural way to a question, which stands in the background
of this paper, namely:
How does a typical vector of the ℓn
p unit ball look like?
or, posed in an exact way:
Let µ be a probability measure on the unit ball of ℓn
p . What is the mean value of
σm(x)q with respect to this measure?
Of course, the choice of µ plays a crucial role. There are several standard proba-
p in a natural way, namely
bility measures, which are connected to the unit ball of ℓn
(cf. Definitions 2 and 9)
(i) the normalized Lebesgue measure,
(ii) the n − 1 dimensional Hausdorff measure restricted to the surface of the unit
ball of ℓn
p and correspondingly normalized,
(iii) the so-called normalized cone measure.
Unfortunately, it turns out, that all these three measures are "bad" -- a typical
vector with respect to any of them does not involve much structure and corresponds
rather to noise then signal (in the sense described below). Therefore, we are looking
for a new type of measures (cf. Definition 13), which would behave better from this
point of view.
1.3 Random models of noise and signals
Random vectors play an important role in the area of signal processing. For example,
if n ∈ N is a natural number, ω = (ω1, . . . , ωn) is a vector of independent Gaussian
variables and ε > 0 is a real number, then εω is a classical model of noise, namely
the white noise. This model is used in the theory but also in the real life applications
of signal processing.
The random generation of a structured signal seems to be a more complicated
task. Probably the most common random model to generate sparse vectors, cf.
[7, 13, 30, 40], is the so-called Bernoulli-Gaussian model. Let again n ∈ N be a
3
natural number and ε > 0 be a real number. Also ω = (ω1, . . . , ωn) stands for a
vector of independent Gaussian variables. Furthermore, let 0 < p < 1 be a real
number and let = (1, . . . , n) be a vector of independent Bernoulli variables
defined as
i =(1, with probability p,
0, with probability 1 − p.
The components of the random Bernoulli-Gaussian vector x = (x1, . . . , xn) are then
defined through
xi = εi · ωi,
i = 1, . . . , n.
(5)
Obviously, the average number of non-zero components of x is k := pn. Unfortu-
nately, if k is much smaller than n, then the concentration of the number of non-zero
components of x around k is not very strong. This becomes better, if k gets larger.
But in that case, the model (5) resembles more and more the model of white noise.
In some sense, (5) represents rather a randomly filtered white noise then a structured
signal. It is one of the main aims of this paper to find a new measure, such that a
random vector with respect to this measure would show a nearly sparse structure
without the need of random filtering.
1.4 Unit sphere
Let us describe the situation in the most prominent case, when p = 2, m = 0 and
µ = µ2 is the normalized surface measure on the unit sphere Sn−1 of ℓn
2 . Furthermore,
we denote by γn the standard Gaussian measure on Rn with the density
1
(2π)n/2 e−kxk2
2/2,
x ∈ Rn.
We use polar coordinates to calculate
2/2dx
ZRn
j=1,...,nxj dγn(x) =
max
=
=
=
1
Ωn
(2π)n/2 ZRn
(2π)n/2 Z ∞
(2π)n/2 Z ∞
(2π)n/2 Z ∞
Ωn
Ωn
0
0
0
max
j=1,...,nxj · e−kxk2
rn−1ZSn−1
rne−r2/2dr ·ZSn−1
rne−r2/2dr ·ZSn−1
max
j=1,...,nrxje−krxk2
2/2dµ2(x) dr
max
j=1,...,nxjdµ2(x)
(6)
σ0(x)∞dµ2(x),
where Ωn denotes the area of Sn−1. This formula connects the expected value of
σ0(x)∞ with the expected value of maximum of n independent Gaussian variables.
Using that this quantity is known to be equivalent to plog(n + 1), cf. [33, (3.14)],
rne−r2/2dr = 2(n−1)/2Γ((n + 1)/2)
and Ωn =
2πn/2
Γ(n/2)
,
Z ∞
0
one obtains
ZSn−1
σ0(x)∞dµ2(x) ≈r log(n + 1)
n
, n ∈ N.
(7)
Several comments on (6) and (7) are necessary.
4
(i) Quantities similar to the left-hand side of (7) have been used in the study of
geometry of Banach spaces and local theory of Banach spaces since many years
and are treated in detail in the work of Milman [23, 35, 36]. Especially, if k·kK
is a norm in Rn and K := {x ∈ Rn : kxkK ≤ 1} denotes the corresponding
unit ball, then the quantity
AK =ZSn−1 kxkK dµ2(x)
(and the closely connected median MK of kxkK over Sn−1) plays a crucial role
in the Dvoretzky theorem [20, 22, 35] and, in general, in the study of Euclidean
sections of K, cf.
[36, Section 5]. Furthermore, it is known that the case of
K = [−1, 1]n, when
AK =ZSn−1
max
j=1,...,nxjdµ2(x) =ZSn−1
σ0(x)∞dµ2(x),
is extremal, cf. [35].
(ii) The connection between the estimated value of a maximum of independent
Gaussian variables and the estimated value of the largest coordinate of a ran-
dom vector on Sn−1 is given just by integration in polar coordinates and is
one of the standard techniques in the local theory of Banach spaces. Due to
the result of [43], this holds true also for other values of p, even for p < 1,
with Gaussian variables replaced by variables with the density cpe−tp
. This
approach is nowadays classical in the study of the geometry and concentration
of measure phenomenon on the ℓn
p -balls, cf. [2, 3, 4, 5, 37, 38, 41].
(iii) For every x ∈ Sn−1 we obtain easily that max
j=1,...,nxj ≥(cid:16) 1
n
x2
j(cid:17)1/2
n
Xj=1
= 1/√n.
Estimate (7) shows that the average value of max
j=1,...,nxj over Sn−1 is asymp-
totically larger only by a logarithmic factor. The detailed study of the concen-
tration of max
j=1,...,nxj around its estimated value (or its mean value) is known
as concentration of measure phenomena [32, 33, 36] and gives more accurate
information then the one included in (7). As our main interest lies in esti-
mates of average best m-term widths, cf. Definition 1, we do not investigate
the concentration properties in this paper and leave this subject to further
research.
(iv) The calculation (6) is based on the use of polar coordinates. For p 6= 2, the
normalized cone measure is exactly that measure, for which a similar formula
holds, cf. (13). The estimates for n − 1 dimensional surface measure are later
obtained using its density with respect to the cone measure, cf. Lemma 10.
(v) As we want to keep the paper self-contained as much as possible and to make
it readable also for readers without (almost) any stochastic background, we
prefer to use simple and direct techniques. For example we use rather the
simple estimates in Lemma 5, than any of their sophisticated improvements
available in literature.
5
(vi) The connection to random Gaussian variables explains, why a random point
of Sn−1 is sometimes referred to as white (or Gaussian) noise. It is usually not
associated with any reasonable (i.e. structured) signal, rather it represents a
good model for random noise.
1.5 Basic Definitions and Main Results
1.5.1 Definition of average best m-term widths
After describing the context of our work we shall now present the definition of the
so-called average best m-term widths, which are the main subject of our study.
First, we observe, that
σm((x1, . . . , xn))q = σm((ε1x1, . . . , εnxn))q = σm((x1, . . . ,xn))q
holds for every x ∈ Rn and ε ∈ {−1, +1}n. Also all the measures, which we shall
consider, are invariant under any of the mappings
(x1, . . . , xn) → (ε1x1, . . . , εnxn),
and therefore we restrict our attention only to Rn
ε ∈ {−1, +1}n
+ in the following definition.
Definition 1. Let 0 < p ≤ q ≤ ∞ and let n ≥ 2 and 0 ≤ m ≤ n − 1 be natural
numbers.
(i) We set
∆n
p =({(t1, . . . , tn) ∈ Rn
{(t1, . . . , tn) ∈ Rn
j=1 tp
+ :Pn
+ : maxj=1,...,n tj = 1},
j = 1},
p < ∞,
p = ∞.
(ii) Let µ be a Borel probability measure on ∆n
p . Then
σp,q
m (µ) =Z∆n
p
σm(x)qdµ(x)
is called average surface best m-term width of id : ℓn
(iii) Let ν be a Borel probability measure on [0, 1] · ∆n
p → ℓn
p . Then
q with respect to µ.
σp,q
m (ν) =Z[0,1]·∆n
p
σm(x)qdν(x)
is called average volume best m-term width of id : ℓn
p → ℓn
q with respect to ν.
Let us observe, that the estimates
m (µ) ≤ σp,q
σp,q
m and σp,q
m (ν) ≤ σp,q
m
follow trivially by Definition 1. Furthermore, the mapping x → σm(x)q is continuous
and, therefore, measurable with respect to the Borel measure µ.
6
dp
dµp
where
(x) = c−1
p,n(cid:18) n
Xi=1
p(cid:18) n
Xi=1
x2p−2
i
cp,n =Z∆n
x2p−2
i
(cid:19)1/2
,
(cid:19)1/2
dµp(x)
1.5.2 Main results
After introducing new notion of average best m-term width in Definition 1, we
study its behavior for the measures on ∆n
p , which are widely used in literature. A
prominent role among them is played by the so-called normalized cone measure given
by
µp(A) =
λ([0, 1] · A)
λ([0, 1] · ∆n
p )
, A ⊂ ∆n
p .
In Theorem 7 and Proposition 8 we provide basic estimates of σp,q
m (µp) for q = ∞
and q < ∞, respectively. Surprisingly enough, it turns out that (7) has its direct
counterpart for all 0 < p < ∞. This means (as described above), that the coordinates
of a "typical" element of the surface of the ℓn
p unit ball are well concentrated around
the value n−1/p. So, roughly speaking, it is only ℓp-normalized noise.
Another well known probability measure on ∆n
p is the normalized surface measure
p, cf. Definition 9. We calculate in Lemma 10 the density of p with respect to µp
to be equal to
is the normalizing constant. This result (which is a generalization of the work of
Naor and Romik [38] to the non-convex case 0 < p < 1) might be of independent
interest for the study of the geometry of ℓn
p spheres. One observes immediately,
that if p < 1 and one or more coordinates of xi are going to zero, then this density
has a polynomial singularity and, therefore, gives more weight to areas closed to
coordinate hyperplanes.
We then obtain in Theorem 12 an estimate of σp,∞
(p) from above. Although
the measure p concentrates around coordinate hyperplanes, it turns out, that the
estimate from above of σp,∞
(µp) as obtained in Theorem 7 and the estimate of
Theorem 12 differ only in the constants involved.
0
0
The last part of this paper is devoted to the search of a new probability measure
on ∆n
p , which would "promote sparsity" in the sense, that the mean value of σm(x)q
decays rapidly with m. One possible candidate is presented in Definition 13 by
introducing a new class of measures θp,β, which are given by their density with
respect to the cone measure µp
dθp,β
dµp
(x) = c−1
p,β ·
n
Yi=1
xβ
i ,
x ∈ ∆n
p ,
where cp,β is a normalising constant. We refer also to Remark 4 for an equivalent
characterisation.
We show, that for an appropriate choice of β, namely β = p/n− 1, the estimated
p -unit sphere decays expo-
m−1(θp,p/n−1), which
value of the m-th largest coefficient of elements of the ℓn
nentially with m. Namely, Theorem 16 provides estimates of σp,∞
7
σp,∞
m−1(θp,p/n−1) ≤ lim sup
n→∞
σp,∞
m−1(θp,p/n−1) ≤
C 2
p
p + 1(cid:17)m
(cid:16) 1
(8)
at the end imply that
C 1
p
p + 1(cid:17)m ≤ lim inf
(cid:16) 1
n→∞
for two positive real numbers C 1
p and C 2
p , which depend only on p.
This result (which is also simulated numerically in the very last section of this
paper) is in a certain way independent of n. This gives a hope, that one could
apply this approach also to the infinite-dimensional spaces ℓp or, using a suitable
discretization technique (like wavelet decomposition), also to some function spaces.
This remains a subject of our further research.
Of course, the class θp,β provides only one example of measures with rapid decay
of their average best m-term widths. We leave also the detailed study of other
measures with such properties open to future work.
Note added in the proof: Let us comment on the relation of our work with
recent papers of Cevher [12] and Gribonval, Cevher, and Davis [29]. Cevher uses
in [12] the concept of Order Statistics [16] to identify the probability distributions,
whose independent and identically distributed (i.i.d.) realizations result typically in
p-compressible signals, i.e.
x∗
i ≤ C R · i−1/p.
Our approach here is a bit different and more connected to the geometry of ℓn
p spaces.
In accordance with [43], this leads to the study of ℓn
p -normalized vectors with i.i.d.
components. This again allows us to better distinguish between the norm of such a
vector (i.e. its size or energy) and its direction (i.e. its structure).
The approach of the recent preprint [29] (which was submitted during the review
process of this work) comes much closer to ours. Their Definition 1 of "Compressible
priors" introduces the quantity called relative best m-term approximation error as
¯σm(x)q =
σm(x)q
kxkq
,
x ∈ Rn
+.
mn
The asymptotic behavior of this quantity for x = (x1, . . . , xn) being a vector with
i.i.d. components and lim inf n→∞
n ≥ κ ∈ (0, 1) is then used to define q-compressible
probability distribution functions. In contrary to [29], we consider ℓq approximation
of ℓp normalized vectors and therefore our widths depend on two integrability pa-
rameters p and q. Furthermore, we do not pose any restrictions on the ratio m/n
to any specific regime and consider the average best m-term widths σp,q
m (µ) for all
0 ≤ m ≤ n − 1. In the only case, when we speak about asymptotics (i.e. (37) of
Theorem 16), we suppose m to be constant and n growing to infinity. Furthermore,
Theorem 1 of [29] shows that all distributions with bounded fourth moment do not
fit into their scheme and do not "promote sparsity". As we are interested in distri-
butions, which are connected to the geometry of ℓn
p -balls (i.e. generalized Gaussian
distribution and generalized Gamma distribution), it is exactly that reason why
we change the parameters of the distribution θp,β in dependence of n. Although
quite inconvenient from the mathematical point of view, it is not really clear if this
presents a serious obstacle for application of our approach. But the investigation of
this goes beyond the scope of this work.
8
1.5.3 Structure of the paper
The paper is structured as follows. The rest of Section 1 gives some notation used
throughout the paper. Sections 2 and 3 provide estimates of this quantity with
respect to the cone and surface measure, respectively.
In Section 4, we study a
new type of measures on the unit ball of ℓn
p . We show, that the typical element
with respect to those measures behaves in a completely different way compared
to the situations discussed before. Those results are illustrated by the numerical
experiments described in Section 5.
1.6 Notation
We denote by R the set of real numbers, by R+ := [0,∞) the set of nonnegative
real numbers and by Rn and Rn
+ their n-fold tensor products. The components of
x ∈ Rn are denoted by x1, . . . , xn. The symbol λ stands for the Lebesgue measure
on Rn and H for the n − 1 dimensional Hausdorff measure in Rn. If A ⊂ Rn and
I ⊂ R is an interval, we write I · A := {tx : t ∈ I, x ∈ A}.
We shall use very often the Gamma function, defined by
Γ(s) :=Z ∞
0
ts−1e−tdt,
s > 0.
(9)
In one case, we shall use also the Beta function
B(p, q) :=Z 1
0
tp−1(1 − t)q−1dt =
Γ(p)Γ(q)
Γ(p + q)
,
p, q > 0
(10)
and the digamma function
Ψ(s) :=
d
ds
log Γ(s) =
Γ′(s)
Γ(s)
,
s > 0.
We recommend [1, Chapter 6] as a standard reference for both basic and more
advanced properties of these functions. We shall need the Stirling's approximation
formula (which was implicitly used already in (7)) in its most simple form
Γ(x) =r 2π
x (cid:16) x
j=1 and b = {bj}∞
e(cid:17)x(cid:18)1 + O(cid:18) 1
x(cid:19)(cid:19) ,
x > 0.
(11)
If a = {aj}∞
j=1 are real sequences, then aj . bj means, that
there is an absolute constant C > 0, such that aj ≤ C bj for all j = 1, 2, . . . . Similar
convention is used for aj & bj and aj ≈ bj. The capital letter C with indices (i.e.
Cp) denotes a positive real number depending only on the highlighted parameters
and their meaning can change from one occurrence to another. If, for any reason,
we shall need to distinguish between several numbers of this type, we shall write for
example C 1
p as already done in (8).
p and C 2
9
2 Normalized cone measure
In this section, we study the average best m-term widths as introduced in Definition
1 for the most important measure (the so-called cone measure) on ∆n
p , which is
well studied in the literature within the geometry of ℓn
[38, 4, 37, 5].
Essentially, we recover in Theorem 7 an analogue of the estimate (7) for all 0 < p <
∞.
Definition 2. Let 0 < p ≤ ∞ and n ≥ 2. Then
λ([0, 1] · A)
λ([0, 1] · ∆n
p )
, A ⊂ ∆n
p spaces, cf.
µp(A) =
p
is the normalized cone measure on ∆n
p .
If νp denotes the p-normalized Lebesgue measure, i.e.
νp(A) =
λ(A)
λ([0, 1] · ∆n
p )
, A ⊂ Rn
+,
then the connection between νp and µp is given by
νp(A) = nZ ∞
0
rn−1µp(cid:18){x ∈ A : kxkp = r}
r
(cid:19)dr.
(12)
The proof of (12) follows directly for sets of the type [a, b]·A with 0 < a < b < ∞ and
A ⊂ ∆n
p and is then finished by standard approximation arguments. The formula
(12) may be generalized to the so-called polar decomposition identity, cf. [4],
f (x)dλ(x)
ZRn
λ([0, 1] · ∆n
p )
+
= nZ ∞
0
rn−1Z∆n
p
f (rx)dµp(x)dr,
(13)
which holds for every f ∈ L1(Rn
+).
The formula (13) allows to transfer immediately the results for the average sur-
face best m-term approximation with respect to µp to the average volume approxi-
mation with respect to νp.
Proposition 3. The identity
σp,q
m (νp) = σp,q
m (µp) ·
n
n + 1
holds for all 0 < p ≤ q ≤ ∞, all n ≥ 2 and all 0 ≤ m ≤ n − 1.
Proof. We plug the function
f (x) = σm(x)q · χ[0,1]·∆n
p (x)
10
into (13) and obtain
Z[0,1]·∆n
p
σm(x)qdλ(x)
λ([0, 1] · ∆n
p )
σm(x)qdνp(x)
p
=Z[0,1]·∆n
= nZ 1
0
rn−1Z∆n
p
σm(rx)qdµp(x)dr = nZ 1
0
rndr · σp,q
m (µp),
which gives the result.
Proposition 3 shows, that the ratio between approximation with respect to µp and
p . Furthermore,
for
νp is equal to 1 + 1/n. This justifies our interest in measures on ∆n
it shows that the quantities σp,q
n → ∞) very similarly.
random variables. Then
Let p = 2 and let ω1, . . . , ωn be independent normally distributed Gaussian
m (µp) behave asymptotically (i.e.
m (νp) and σp,q
2(A) = µ2(A) = P (ω1, . . . ,ωn)
j=1 ω2
j(cid:1)1/2 ∈ A!,
(cid:0)Pn
A ⊂ ∆n
2 .
As noted in [43], this relation may be generalized to all values of p with 0 < p < ∞.
Let ω1, . . . , ωn be independent random variables on R+ each with density
with respect to the Lebesgue measure, where cp = p
Γ(1/p) =
1
Γ(1/p+1) .
cpe−tp
,
t ≥ 0
Then, cf. [43, Lemma 1],
µp(A) = P (ω1, . . . , ωn)
j=1 ωp
j(cid:1)1/p ∈ A!,
(cid:0)Pn
A ⊂ ∆n
p .
(14)
We shall fix ω1, . . . , ωn to the end of this paper. Also the symbols E and P are always
taken with respect to these variables.
2.1 The case q = ∞
In this section we deal with uniform approximation, i.e. with the case q = ∞. To
be able to imitate the calculation (6), we shall need several tools, which are subject
of Lemmas 4, 5 and 6. Our main result of this section (Theorem 7) then provides
the estimate of σp,∞
m (µp) from above for all m with 0 ≤ m ≤ n − 1. Furthermore, it
is shown that in the range 0 ≤ m ≤ εpn this estimate is also optimal.
Lemma 4. Let 0 < p < ∞ and let n ≥ 2 and 1 ≤ m ≤ n be natural numbers. Then
x∗
mdµp(x) =
Z∆n
p
Γ(n/p)
Γ(n/p + 1/p) · E x∗
m.
p and C 2
Furthermore, there are two positive real numbers C 1
such that
p depending only on p,
C 1
p ·
E x∗
m
n1/p ≤Z∆n
p
E x∗
m
n1/p
.
mdµp(x) ≤ C 2
x∗
p ·
11
Proof. We put f (x) = x∗
me−xp
1−···−xp
n and use the polar decomposition identity (13)
ZRn
+
me−xp
x∗
1−···−xp
ndλ(x)
λ([0, 1] · ∆n
p )
0
= nZ ∞
= nZ ∞
0
(rx∗
rn−1Z∆n
rn−1 · re−rp
p
m) · e−(rx1)p−···−(rxn)p
drZ∆n
x∗
mdµp(x)
p
dµp(x)dr
or, equivalently,
Z∆n
p
The identity
mdµp(x) = ZRn
+
x∗
me−xp
x∗
1−···−xp
ndλ(x)
λ([0, 1] · ∆n
p ) · nR ∞
0 rne−rpdr
.
(15)
Z ∞
0
rne−rp
dr =
Γ(n/p + 1/p)
p
,
follows by a simple substitution. Furthermore, we shall need the classical formula
of Dirichlet for the volume of the unit ball Bℓn
p , cf. [21, p. 157],
p of ℓn
λ([0, 1] · ∆n
p ) =
λ(Bℓn
p )
2n =
Γ(1/p + 1)n
Γ(n/p + 1)
.
This allows us to reformulate (15) as
x∗
mdµp(x) =
Z∆n
p
Γ(n/p + 1) E x∗
m
cn
p · n/p · Γ(n/p + 1/p)Γ(1/p + 1)n =
Γ(n/p) E x∗
m
Γ(n/p + 1/p)
.
Finally, we use Stirling's formula (11) to estimate
n1/p · Γ(n/p)
Γ(n/p + 1/p) ≤ C 1
p
n1/p(n/p)n/p−1/2
(n/p + 1/p)n/p+1/p−1/2 ≤ C 2
n + 1(cid:19)n/p+1/p−1/2
p(cid:18) n
≤ C 3
p
and similarly for the estimate from below.
Lemma 5. Let α ∈ R and δ > 0. Then
δ
1
Z ∞
1,
1−α/δ ,
uαe−udu ≤ δαe−δ ·
δ(cid:1)α · α/δ
(cid:0) α
Proof. If α ≤ 0, we may estimate
uαe−udu ≤ δαZ ∞
Z ∞
1−δ/α ,
δ
δ
if α ≤ 0,
if α > 0
if α > 0
and
and
α
δ < 1,
α
δ > 1.
e−udu = δαe−δ.
If 0 < α ≤ 1, we use partial integration and obtain
Z ∞
δ
uαe−udu = δαe−δ + αZ ∞
δ
uα−1e−udu ≤ δαe−δ(1 + αδ−1).
12
This is smaller than
δαe−δ(1 +
α
δ
+
α2
δ2 + . . . ) = δαe−δ ·
1
1 − α/δ
if α/δ < 1 and smaller than
δαe−δ α
δ
(1 +
δ
α
+
δ2
α2 + . . . ) = δαe−δ α
δ ·
1
1 − δ/α
.
if α/δ > 1.
If k − 1 < α ≤ k for some k ∈ N, we iterate the partial integration and arrive at
Z ∞
uαe−udu ≤ δαe−δ(1 + αδ−1 + α(α − 1)δ−2 + ··· + α(α − 1) . . . (α − k + 1)δ−k)
δ
α
δ
+
≤ δαe−δ(1 +
≤ δαe−δ( 1
δ(cid:1)α+1
(cid:0) α
1−α/δ ,
α2
αk
δk )
δ2 + ··· +
if α/δ < 1,
1
1−δ/α ,
if α/δ > 1.
Lemma 6. Let 0 < p < ∞. Then there is a positive real number Cp, such that
E x∗
m ≤ Cp log1/p(cid:16) en
m(cid:17)
for all 1 ≤ m ≤ n.
Proof. We estimate
P(ω∗
m > t)dt
E x∗
0
P(ω∗
m =Z ∞
≤ δ +(cid:18) n
= δ +(cid:18) n
m > t)dt = δ +Z ∞
m(cid:19)Z ∞
m(cid:19)Z ∞
P(ω1 > t)mdt.
δ
δ
δ
P(ω1 > t, ω2 > t, . . . , ωm > t)dt
(16)
The parameter δ > max(1, 3(1/p − 1))1/p is to be chosen later on. We substitute
v = up and obtain
P(ω1 > t) = cpZ ∞
t
e−up
du =
cp
p Z ∞
tp
v1/p−1e−vdv.
Using the first two estimates of Lemma 5 (recall that tp ≥ δp > max(1, 3(1/p− 1))),
we arrive at
P(ω1 > t) ≤ Cpt1−pe−tp
,
where Cp depends only on p. We plug this estimate into (16) and obtain
E x∗
m ≤ δ +(cid:18) n
m(cid:19)(Cp)mZ ∞
δ
13
tm(1−p)e−mtp
dt.
(17)
If p ≥ 1, then
Z ∞
δ
tm(1−p)e−mtp
dt ≤ δm(1−p)Z ∞
δ
Altogether, we obtain
e−mtp
mδp
dt ≤ δm(1−p)Z ∞
m(cid:19)(Cp)me−mδp
.
E x∗
m ≤ δ +(cid:18) n
e−uu1/p−1du ≤ e−mδp
.
m )m and choosing δ = C ′
p ln( en
m )1/p finishes the proof.
If p < 1, we use again the second estimate of Lemma 5
m(cid:1) ≤ ( en
Using (cid:0) n
Z ∞
δ
tm(1−p)e−mtp
dt =
≤
m(cid:1) ≤ ( en
Using (17) and (cid:0) n
E x∗
1
mp · m(1/p−1)(m+1)Z ∞
mp · δ(1−p)(m+1)e−mδp
·
1
mδp
m )m again, we get
u(1/p−1)(m+1)e−udu
1
1 − 2(1/p−1)
δp
pδ(1−p)(m+1)e−mδp
≤ C ′
.
1 ≤ δ + exp(−mδp + m ln(en/m) + (1 − p)(m + 1) ln δ + m ln Cp + ln C ′
p)
≤ δ + exp[−m(δp + Cp ln(en/m) + 2(1 − p) ln δ)]
The choice δ = C ′
p ln( en
m )1/p with C ′
p large enough ensures, that
δp
2 ≥ Cp ln(en/m)
and
δp
2 ≥ 2(1 − p) ln δ
and finishes the proof.
The following theorem gives the basic estimates of σp,∞
m (µp).
Theorem 7. Let 0 < p ≤ ∞ and let n ≥ 2.
(i) Let 0 ≤ m ≤ n − 1. Then
m (µp) ≤ Cp" log(cid:16) en
m+1(cid:17)
n
σp,∞
#1/p
.
(18)
(ii) There is a number 0 < εp < 1, such that for 0 ≤ m ≤ εpn the following estimate
holds
σp,∞
m (µp) ≥ Cp(cid:20) log( en
m+1 )
n
(cid:21)1/p
.
(19)
Proof. Lemma 4 and Lemma 6 imply immediately the first part of the theorem if
p < ∞. If p = ∞, the proof is trivial.
The proof of the second part is divided into two steps.
Step 1. We start first with the case m = 0.
If p = ∞, then x∗
1 = 1 for all x ∈ ∆n
assume, that p < ∞. According to Lemma 4, we have to estimate E x∗
p and the proof is trivial. Let us therefore
1 from below.
14
This was done in [43, Lemma 2]. We include a slightly different proof for readers
convenience. For every t0 > 0, it holds
E x∗
1 ≥ t0 P(x∗
1 > t0) = t0 P( max
1≤j≤n
xj > t0) ≥ t0[nP(x1 > t0) −(cid:18)n
2(cid:19)P(x1 > t0)2].
We define t0 by P(x1 > t0) = 1
n and obtain E x∗
From the simple estimate
1 ≥ t0/2.
cp
p Z ∞
T p
u1/p−1e−udu ≥ Cpe−2T p
,
T > 1,
it follows, that there is a positive real number γp > 0, such that
P(x1 > γp(log(en))1/p) ≥ 1/n.
1 ≥ Cp(log(en))1/p.
This gives t0 ≥ γp(log(en))1/p and E x∗
Step 2. Let 0 ≤ m ≤ εpn, where εp > 0 will be chosen later on.
We shall use the inequality
1
m
m
Xj=1
log1/p(cid:16) en
j (cid:17) ≤ Cp log1/p(cid:16) en
m(cid:17),
1 ≤ m ≤ n,
(20)
which follows by direct calculation for p = 1, by Holder's inequality for 1 < p < ∞
and by replacing the sum by the corresponding integral and integration by parts if
0 < p < 1.
We denote
By Lemma 6 and (20),
kxk(m) =
1
m
m
Xj=1
x∗
j .
Ekxk(m) =
1
m
m
Xj=1
E x∗
j ≤
Cp
m
m
Xj=1
log1/p(cid:16) en
j (cid:17) ≤ C 1
p log1/p(cid:16) en
m(cid:17).
(21)
To estimate Ekxk(m) from below, we assume that 1 ≤ m ≤ n and that n/m is
an integer (otherwise one has to slightly modify the argument at the cost of the
constants involved). We partition the set {1, . . . , n} = A1 ∪ ··· ∪ Am, where each
one of the disjoint sets Aj has n/m elements. Then we have
kxk(m) ≥
1
m
m
Xj=1
max
l∈Aj
xl
and by the first step we obtain
Ekxk(m) ≥
1
m
m
Xj=1
E max
l∈Aj
xl ≥ C 2
p log1/p(cid:16) en
m(cid:17).
(22)
15
Let Np < 1/εp be a natural number to be chosen later on. Combining (21) with
(22) gives finally
1
E x∗
m ≥
Npm
Xk=m
1
Np
E x∗
Npm
C 1
p
Np
k ≥ Ekxk(Npm) −
Ekxk(m)
p log1/p(cid:16) en
log1/p(cid:16) en
Npm(cid:17) −
m(cid:17)
≥ C 2
m(cid:17)
p
= log1/p(cid:16) en
1 −
log(cid:16) en
m(cid:17)
log(Np)
C 1
p
Np
C 2
−
1/p
.
An appropriate choice of Np and εp (i.e. Np > 21/pC 1
with
p /C 2
p and εp < min(1/Np, e/N 2
p ))
C 2
p
1 −
log(Np)
log(cid:16) e
εp(cid:17)
gives the result.
1/p
C 1
p
Np
−
> 0
(i) Theorem 7 provides basic estimates of average best m-term widths
Remark 1.
σp,∞
In the case m = 0 a stronger result on concentration of µp was
m (µp).
obtained already in [43, Theorem 3 and Remark 2]. It would be certainly of
interest to obtain a similar statement also for other values of m > 0, but this
would go beyond the scope of this paper and we leave this direction open for
further study.
(ii) Theorem 7 may be interpreted in the sense of the discussion after formula
(7). Namely, the average coordinate of x ∈ ∆n
p is n−1/p. Theorem 7 shows,
that the average value of the largest coordinate is only slightly larger (namely
c[ln(en)]1/p times larger). In this sense, the average point of ∆n
p is only slightly
modified (and properly normalized) white noise.
(iii) Using the interpolation formula (4), one may immediately extend this result
to all 0 < p ≤ q < ∞. But we shall see later on, that in the case q < ∞, one
may prove slightly better estimates.
(iv) The behavior of σp,∞
m (µp) was studied in detail in [28, Example 10] for p = 2.
It was shown that if xi are independent N (0, 1) Gaussian random variables
and m ≤ n/2 + 1, then
crln
2n
m ≤ E x∗
m ≤ Crln
2n
m
,
where c and C are absolute positive constants. Furthermore, if m ≥ n/2 + 1,
then
r π
2
n − m + 1
n + 1 ≤ E x∗
m ≤
√2π
n − m + 1
n
.
(v) The method used in the proof of the second part of Theorem 7 may be found
for example in [27].
16
2.2 The case q < ∞
We discuss briefly also the case when q < ∞. It turns out, that in this case the
logarithmic term disappears. We do not go much into details and restrict ourselves
to the case m = 0.
Proposition 8. Let n ≥ 2 and 0 < p ≤ q < ∞. Then
(i) C 1
p,qn1/q ≤ Ekxkq ≤ C 2
p,qn1/q,
(ii)
C 1
p,q ·
Ekxkq
n1/p ≤ σp,q
0 (µp) =Z∆n
p kxkqdµp(x) ≤ C 2
p,q ·
Ekxkq
n1/p
and
p,qn1/q−1/p ≤ σp,q
(iii) C 1
0 (µp) ≤ C 2
p,qn1/q−1/p,
where in all these estimates C 1
p.
p and C 2
p are positive real numbers depending only on
Proof. (i) The following two inequalities may be easily proved by Holder's and
Minkowski inequality.
n
(cid:18) n
(Exj)q(cid:19)1/q
Xj=1
Xj=1
(cid:0)
This gives for q ≥ 1
Exq
Xj=1
≤ E(cid:0)
j(cid:1)1/q ≤ E(cid:0)
Xj=1
n
n
n
xq
xq
Exq
j(cid:1)1/q ≤(cid:0)
j(cid:1)1/q,
Xj=1
j(cid:1)1/q ≤(cid:18) n
(Exj)q(cid:19)1/q
Xj=1
q ≥ 1,
,
q ≤ 1.
and for q ≤ 1
Ekxkq ≤ n1/q(Exq
j)1/q
and Ekxkq ≥ n1/q Exj
Ekxkq ≤ n1/qExj
and Ekxkq ≥ n1/q(Exq
j )1/q.
Let us note, that the value of Exj and (Exq
and q.
j )1/q does not depend on n, only on p
(ii) The proof of the second part resembles very much the proof of Lemma 4 and
is left to the reader.
(iii) The last point follows immediately from (i) and (ii).
Remark 2. A similar statement to Proposition 8 is included in [43, Lemma 2, point
4].
17
3 Normalized surface measure
In this section we study the average best m-term widths for another classical measure
on ∆n
p , namely the normalized Hausdorff measure, cf. Definition 9. Intuitively, this
measure gives more weight to those areas, where one or more components of x ∈ ∆n
p
are close to zero. It turns out, that this is really the case - with the mathematical
formulation given in Lemma 10 below. This relation is then used together with
Lemma 11 in Theorem 12 to provide estimates of σp,∞
(p) from above.
0
Definition 9. Let n ≥ 2 be a natural number. We denote by
p(A) = H(A)
H(∆n
p )
,
A ⊂ ∆n
p
the normalized n − 1 dimensional Hausdorff measure on ∆n
p .
Let us mention, that for p ∈ {1, 2,∞} the measure p coincides with µp. The
following lemma provides a relationship between the normalized surface measure p
and the cone measure µp. For p ≥ 1, it was given by [38]. We follow closely their
approach and it turns out, that it may be generalized also to the non-convex case
of 0 < p < 1.
Lemma 10. Let 0 < p < ∞ and n ≥ 2. Then p is an absolutely continuous measure
with respect to µp and for µp almost every x ∈ ∆n
p it holds
dp
dµp
(x) =
nλ([0, 1] · ∆n
p )
H(∆n
p )
where
= c−1
p,n(cid:18) n
Xi=1
x2p−2
i
(cid:19)1/2
,
(cid:13)(cid:13)(cid:13)∇(k · kp)(x)(cid:13)(cid:13)(cid:13)2
p(cid:18) n
(cid:19)1/2
Xi=1
x2p−2
i
cp,n =Z∆n
dµp(x)
is the normalizing constant.
Proof. The proof imitates the proof of [38, Lemma 1 and Lemma 2], where the
statement was proven for 1 ≤ p < ∞. Hence, we may assume, that 0 < p < 1. First,
we introduce some notation.
p , such that
We fix x = (x1, . . . , xn) ∈ ∆n
• the mapping y → kykp is differentiable at x,
• x is a density point of H, i.e.
lim
ε→0+
H(B(x, ε) ∩ ∆n
p )
εn−1Vn−1
= 1,
(23)
where Vn−1 denotes the Lebesgue volume of the n − 1 dimensional Euclidean
unit ball.
• xi > 0 for all i = 1, . . . , n.
18
Obviously, p-almost every x ∈ ∆n
example to [34, Theorem 16.2] for the second one).
p satisfies all the three properties (we refer for
Furthermore, we put z := ∇(k · kp)(x). This means, that
where
kx + ykp = 1 + hz, yi + r(y),
: 0 < kyk2 ≤ δ(cid:27) ,
θ(δ) := sup(cid:26)r(y)
kyk2
(24)
δ > 0
tends to zero if δ tends to zero. Using (24) for y = δx, one observes, that hz, xi = 1.
We denote by H = x + z⊥ the tangent hyperplane to ∆n
p at x. Let us note, that
for 0 < p < 1 the set Rn
p is convex. Next, we show, that
hz, yi ≥ 1 for every y ∈ [1,∞) · ∆n
p = [1,∞) · ∆n
+ \ [0, 1) · ∆n
p . Indeed,
1 ≤ kx + λ(y − x)kp = 1 + hz, λ(y − x)i + r(λ(y − x))
= 1 − λ + λhz, yi + r(λ(y − x))
Dividing by λ > 0 and letting λ → 0 gives the statement.
The proof of the lemma is based on the following two inclusions, namely
p(cid:17)
[0, 1] ·(cid:16)B(x, ε(1 − θ(ε))) ∩ H(cid:17) ⊂ [0, 1] ·(cid:16)B(x, ε) ∩ ∆n
(25)
and
[0, 1] ·(cid:16)B(x, ε) ∩ ∆n
First, we prove (25). To given 0 ≤ s ≤ 1 and v ∈ B(x, ε(1 − θ(ε)) ∩ H we need
p(cid:17) ⊂ [0, 1 + εθ(ε)] ·(cid:16)B(x, ε(1 + θ(ε)kxk2)) ∩ H(cid:17),
which hold for all ε > 0 small enough.
(26)
p , such that sv = tw. To do this, we set
to find 0 ≤ t ≤ 1 and w ∈ B(x, ε) ∩ ∆n
kvkp ∈ ∆n
w :=
v
p
and t := skvkp.
We need to show, that t ≤ 1 and kx − wk2 ≤ ε.
We choose 0 < ε ≤ mini xi. Then
xi ≤ xi − vi + vi ≤ kx − vk2 + vi ≤ ε + vi
for every i = 1, . . . , n, which implies, that vi ≥ 0 and v ∈ Rn
v ∈ Rn
+ we deduce, that kvkp ≤ 1. Hence t = skvkp ≤ kvkp ≤ 1.
Next, we write
+. From v ∈ H and
kx − wk2 =(cid:13)(cid:13)(cid:13)
x −
v
kvkp(cid:13)(cid:13)(cid:13)2 ≤ kx − vk2 +(cid:13)(cid:13)(cid:13)
v −
v
kvkp(cid:13)(cid:13)(cid:13)2
1 − kvkp
kvkp ≤ ε(1 − θ(ε)) + 1 − kvkp
≤ ε(1 − θ(ε)) + kvk2 ·
= ε(1 − θ(ε)) + 1 − {1 + hv − x, zi + r(v − x)}
= ε(1 − θ(ε)) + r(v − x) ≤ ε.
19
Next, we prove (26). We need to find to given 0 ≤ t ≤ 1 and w ∈ B(x, ε) ∩ ∆n
some 0 ≤ s ≤ 1 + εθ(ε) and v ∈ B(x, ε(1 + θ(ε)kxk2)) ∩ H, such that tw = sv. We
put
p
s := thw, zi
and v :=
w
.
hw, zi
Let us recall, that we have shown above, that w ∈ ∆n
s ≤ 1 + εθ(ε) and kv − xk2 ≤ ε(1 + θ(ε)kxk2).
Of course, tw = sv and v ∈ H (as hv, zi = 1). Hence, it remains to show, that
The application of (24) gives
p implies that hw, zi ≥ 1.
1 = kwkp = kx + (w − x)kp = 1 + hw − x, zi + r(w − x),
which again forces hw, zi ≤ 1 + εθ(ε). Then s = thw, zi ≤ hw, zi ≤ 1 + εθ(ε).
Finally, we write
w
w
x
kv − xk2 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
hw, zi − x(cid:13)(cid:13)(cid:13)2 ≤(cid:13)(cid:13)(cid:13)
≤ kw − xk2
hw, zi
+(cid:13)(cid:13)(cid:13)
hw, zi −
+ kxk2hw, zi − 1
hw, zi ≤ ε + εθ(ε)kxk2.
hw, zi(cid:13)(cid:13)(cid:13)2
hw, zi − x(cid:13)(cid:13)(cid:13)2
x
Equipped with (25) and (26), we may finish the proof of the lemma. We write
lim
ε→0
p(B(x, ε) ∩ ∆n
p )
µp(B(x, ε) ∩ ∆n
p )
H(B(x, ε) ∩ ∆n
p )
= lim
ε→0
H(∆n
p )
λ([0, 1] · ∆n
p )
H(∆n
p )
· lim
ε→0
=
·
εn−1Vn−1
εn−1Vn−1 ·
εn−1Vn−1
λ([0, 1] · [B(x, ε) ∩ ∆n
p ])
λ([0, 1] · ∆n
p )
λ([0, 1] · [B(x, ε) ∩ ∆n
p ])
,
(27)
where we have used (23). As the perpendicular distance between zero and H is equal
to 1/kzk2, we observe, that
vol(B(x, a) ∩ H) =
an−1Vn−1
nkzk2
holds for every a > 0. Using this, we get from (25) and (26)
[ε(1 − θ(ε))]n−1Vn−1
λ(cid:16)[0, 1] ·(cid:16)B(x, ε(1 − θ(ε))) ∩ H(cid:17)(cid:17) =
p(cid:17)(cid:17)
≤ λ(cid:16)[0, 1] ·(cid:16)B(x, ε) ∩ ∆n
≤ λ(cid:16)[0, 1 + εθ(ε)] ·(cid:16)B(x, ε(1 + θ(ε)kxk2)) ∩ H(cid:17)(cid:17)
= [1 + εθ(ε)]n ·
[ε(1 + θ(ε)kxk2)]n−1Vn−1
nkzk2
.
nkzk2
Combining these estimates with (27) gives the result.
Following lemma is analogous to Lemma 4 and reduces the calculation of σp,∞
(p)
to inequalities for the estimated values of functions of the random variables x1, . . . , xn.
0
20
Lemma 11. Let 0 < p < ∞. There exists two positive real numbers C 1
such that
p and C 2
p ,
n
n
E x∗
C 1
p·
x2p−2
i
x2p−2
i
(cid:17)1/2
1(cid:16)
Xi=1
(cid:17)1/2
E(cid:16)
Xi=1
= Z∆n
1(cid:16)
Xi=1
Z∆n
p(cid:16)
Xi=1
for all n ≥ 2.
x2p−2
i
x∗
n
n
p
(cid:17)1/2
(cid:17)1/2
dµp(x)
· n−1/p ≤ σp,∞
0
(p) =Z∆n
p
x∗
1dp
(28)
x2p−2
i
dµp(x)
≤ C 2
p
n
x2p−2
i
x2p−2
i
E x∗
1(cid:16)
Xi=1
E(cid:16)
Xi=1
n
(cid:17)1/2
(cid:17)1/2
· n−1/p
Proof. Only the inequalities need a proof. It resembles the proof of Lemma 4 and
is again based on the polar decomposition formula (13).
We plug the functions
f1(x) = x∗
n
1(cid:16)
Xi=1
x2p−2
i
(cid:17)1/2
e−xp
1−···−xp
n
n
and f2(x) =(cid:16)
Xi=1
x2p−2
i
(cid:17)1/2
e−xp
1−···−xp
n
into (13) and obtain
σp,∞
0
+
+
0
f1(x)dx ·Z ∞
(p) = ZRn
f2(x)dx ·Z ∞
ZRn
(cid:17)1/2
1(cid:16)
Xi=1
(cid:17)1/2
E(cid:16)
Xi=1
x2p−2
i
x2p−2
i
E x∗
=
n
n
0
rn+p−2e−rp
rn+p−1e−rp
dr
dr
Γ(n/p + 1 − 1/p)
Γ(n/p + 1)
.
·
By Stirling's formula, the last expression is equivalent to n−1/p with constants of
equivalence depending only on p.
Theorem 12. Let 0 < p < ∞. Then there is a positive real number Cp, such that
σp,∞
0
(p) ≤ Cp(cid:20) log(n + 1)
n
(cid:21)1/p
(29)
for all n ≥ 2.
Proof. We define a probability measure αp,n on R+
n by the density
c−1
p,n · n
Xi=1
x2p−2
i !1/2
e−xp
1−···−xp
n,
cp,n :=ZRn
+ n
Xi=1
x2p−2
i !1/2
21
e−xp
1−···−xp
ndx
with respect to the Lebesgue measure. Let us note, that due to the inequality
x2p−2
i !1/2
≤
n
Xi=1
xp−1
i
n
Xi=1
the integral in the definition of cp,n really converges and αp,n is well defined.
According to Lemma 11, we need to estimate
x∗
1dαp,n(x).
ZRn
+
We calculate for δ > 1, which is to be chosen later on,
ZRn
+
x∗
1dαp,n(x) =Z ∞
0
αp,n(x∗
1 > t)dt ≤ δ +Z ∞
δ
αp,n(x∗
1 > t)dt
αp,n(x1 > t)dt.
≤ δ + nZ ∞
δ
+ . Then
αp,n(x1 > t) = c−1
We write x′ = (x2, . . . , xn) ∈ Rn−1
1ZRn−1
1ZRn−1
e−xp
e−xp
+
t
p,nZ ∞
p,nZ ∞
≤ c−1
p,nZ ∞
= c−1
t
t
e−xp
1 xp−1
1
+ c−1
p,nZ ∞
t
:= I1 + I2.
The inequality
e−xp
xp−1
x2p−2
i !1/2
+ n
Xi=1
1 + n
Xi=2
dx1 ·ZRn−1
1 dx1 ·ZRn−1
+ n
Xi=2
e−xp
+
e−xp
2−···−xp
ndx′dx1
x2p−2
i !1/2
2−···−xp
ndx′
e−xp
2−···−xp
ndx′dx1
x2p−2
i !1/2
e−xp
2−···−xp
ndx′
cn
p cp,n = cn
e−xp
x2p−2
p ZRn
p ZRn
≥ cn
p Z ∞
i !1/2
+ n
Xi=1
i !1/2
+ n
Xi=2
+ n
1 dx1ZRn−1
Xi=2
x2p−2
e−xp
= cn
e−xp
0
1−···−xp
ndx
1−···−xp
ndx
x2p−2
i !1/2
(30)
cp,n−1
e−xp
2−···−xp
ndx′ = cn−1
p
shows, that
I1 =
e−xp
1 dx1
cpR ∞
t xp−1
1
cn
p cp,n
e−xp
1 dx1
cpR ∞
t xp−1
1
cpcp,1
≤
= c−1
p,1 ·
e−tp
p
.
22
Using (30) again, we get also
I2 = c−1
p,n · cp,n−1Z ∞
t
t
e−xp
1 dx1 ≤ cpZ ∞
I1 + I2 ≤ Cpe−tp
1dαp,n(x) ≤ δ + CpnZ ∞
δ
x∗
If p ≥ 1, we get
and
ZRn
+
e−xp
1 dx1 =
cp
p ·Z ∞
tp
s1/p−1e−sds.
,
t > 1
(31)
e−tp
dt ≤ δ + C ′
pne−δp
.
By choosing δ = Cp log(n + 1)1/p, we get the result.
If p < 1, we use the second estimate of Lemma 5 and replace (31) with
I1 + I2 ≤ Cpt1−pe−tp
,
t > t0
for t0 > 1 large enough and the result again follows by the choice of δ.
Remark 3.
(i) Theorem 12 shows, that the average size of the largest coordinate
of x ∈ ∆n
p taken with respect to the normalized Hausdorff measure is again
only slightly larger than n−1/p. Hence, also in this case, the typical element of
∆n
p seems to be far from being sparse and resembles rather properly normalized
white noise in the sense described in Introduction.
(ii) Using interpolation inequality (4), one may again obtain a similar estimate
also for 0 < p ≤ q < ∞, namely
σp,q
0 (p) ≤ Cp,q(cid:20) log(n + 1)
n
(cid:21)1/p−1/q
.
It would be probably possible to avoid the logarithmic terms and provide
improved estimates also for m > 0, but we shall not go into this direction. Our
main aim of this section was to show, that normalized Hausdorff measure does
not prefer sparse (or nearly sparse) vectors, and this was clearly demonstrated
by Theorem 12.
4 Tensor product measures
As discussed already in the Introduction and proved in Theorem 7 and Theorem
12, the average vectors of ∆n
p with respect to the cone measure µp and with respect
to surface measure p behave "badly" meaning that (roughly speaking) many of
their coordinates are approximately of the same size. As promised before, we shall
now introduce a new class of measures, for which the random vector behaves in
a completely different way. These measures are defined through their density with
respect to the cone measure µp. This density has a strong singularity near the points
with vanishing coordinates.
23
Definition 13. Let 0 < p < ∞, β > −1 and n ≥ 2. Then we define the probability
measure θp,β on ∆n
p by
where
dθp,β
dµp
(x) = c−1
p,β ·
n
Yi=1
xβ
i ,
x ∈ ∆n
p ,
cp,β =Z∆n
p
n
Yi=1
xβ
i dµp(x).
(32)
(33)
Remark 4.
(i) If 0 > β > −1, then (32) defines the density of θp,β with respect
to µp only for points, where xi 6= 0 for all i = 1, . . . , n. That means, that this
density is defined µp-almost everywhere. The definition is then complemented
by the statement, that θp,β is absolutely continuous with respect to µp.
(ii) We shall see later on, that the condition β > −1 ensures, that (33) is finite.
(iii) It was observed already in [4], that the measures θp,β allow a formula similar
p into (13), where
i=1 xβ
i e−kxkp
n
p , and obtain
to (14). We plug the function f (x) = χ[0,∞)·AQn
A is any µp-measurable subset of ∆n
p )·n·Z ∞
Z[0,∞)·A
We use a similar formula also for A = ∆n
pdλ(x) = λ([0, 1]·∆n
i e−kxkp
xβ
Yi=1
0
p , which leads to
rn−1+nβe−rp
dr·ZA
n
Yi=1
xβ
i dµp(x).
ZA
1d θp,β = ZA
Z∆n
p
n
n
Yi=1
Yi=1
xβ
i dµp(x)
xβ
i dµp(x)
n
i e−kxkp
xβ
= Z[0,∞)·A
Yi=1
ZRn
i e−kxkp
xβ
Yi=1
n
+
pdx
pdx
.
1, . . . , ω′
Let ω′ = (ω′
n) be a vector with independent identically distributed com-
ponents with respect to the density cp,βtβe−tp
dt
is a normalizing constant. Up to a simple substitution, this is the well known
gamma distribution. We observe that the distribution of random points with
respect to θp,β equals to the distribution of ℓn
p normalized vectors ω′, i.e.
p,β =R ∞
, t > 0, where c−1
tβe−tp
0
θp,β(A) = P (ω′
(cid:0)Pn
1, . . . , ω′
n)
j=1 (ω′
j)p(cid:1)1/p ∈ A!,
A ⊂ ∆n
p .
(34)
(iv) Of course, the same procedure might be considered also for other distributions.
We leave this to future work. We also refer to the discussion on the recent
work of Gribonval, Cevher, and Davies [29] in the Introduction.
Lemma 14. Let 0 < p < ∞, β > −1 and n ≥ 2.
24
(i) Let 1 ≤ m ≤ n. Then
n
σp,∞
m−1(θp,β) =Z∆n
p
x∗
mdθp,β =
E
E x∗
m
xβ
i
n
Yi=1
Yi=1
xβ
i
Γ(n(β + 1)/p)
Γ(n(β + 1)/p + 1/p)
.
·
(ii)
E
n
Yi=1
xβ
i =(cid:20) cp
p · Γ((β + 1)/p)(cid:21)n
.
Proof. The proof of the first part follows again by (13), this time used for the
functions
f1(x) = x∗
n
m(cid:16)
Yi=1
1−···−xp
n
xβ
i(cid:17)e−xp
n
and f2(x) =(cid:16)
Yi=1
1−···−xp
n.
xβ
i(cid:17)e−xp
The proof of the second part is straightforward.
It follows directly from (9), that Γ(s) tends to infinity, when s tends to zero. The
following lemma quantifies this phenomenon. Although the statement seems to be
well known, we were not able to find a reference and we therefore provide at least a
sketch of the proof.
Lemma 15. Let C ≃ 0.577 . . . denote the Euler constant. Then
lim
n (cid:19)n
n→∞(cid:18) Γ(1/n)
= e−C .
Proof. It is enough to show, that
lim
n→∞
n · log(Γ(1 + 1/n)) = −C,
which (by using the l'Hospital rule) follows from
0 s1/ne−s log s ds
0 s1/ne−sds
lim
n→∞R ∞
R ∞
= −C.
But the numerator of this fraction is equal to Γ′(1 + 1/n) and its denominator to
Γ(1 + 1/n). The whole fraction is therefore equal to Ψ(1 + 1/n) and Ψ(1 + 1/n) →
Ψ(1) = −C as n tends to infinity, cf. [1, Section 6.3.2, p. 258].
Next theorem shows, that if β = p/n−1, then the measure θp,β promotes sparsity
and one may even consider limiting behavior of n growing to infinity.
Theorem 16. Let 0 < p < ∞ and let n ≥ 2 and 1 ≤ m ≤ n be integers. Then
σp,∞
m−1(θp,p/n−1) ≥ C 1
p ·
Γ(n + 1)
Γ(n − m + 1) ·
Γ(n/p + n − m + 1)
Γ(n/p + n + 1)
,
(35)
25
and
σp,∞
m−1(θp,p/n−1) ≤ C 2
p ·
Γ(n + 1)
Γ(n − m + 1)(cid:26) Γ(n/p + n − m + 1)
Γ(n/p + n + 1)
+
1
m! ·(cid:18) e−1
Γ(1/n)(cid:19)m(cid:27)(36)
where C 1
p and C 2
p are positive real numbers depending only on p.
Furthermore, for every fixed m ∈ N,
C 1
p
p + 1(cid:17)m ≤ lim inf
(cid:16) 1
n→∞
σp,∞
m−1(θp,p/n−1) ≤ lim sup
n→∞
σp,∞
m−1(θp,p/n−1) ≤
C 2
p
p + 1(cid:17)m ,
(cid:16) 1
(37)
where C 1
p and C 2
p are positive real numbers depending only on p.
Proof. First observe, that n(β + 1)/p = 1 for β = p/n − 1 and therefore
Γ(n(β + 1)/p)
Γ(n(β + 1)/p + 1/p)
=
1
Γ(1 + 1/p)
depends only on p. Due to Lemma 14, we have to estimate
E x∗
m(cid:18) n
Yi=1
xp/n−1
i
(cid:19) = cn
pZRd
+
x∗
m
n
Yi=1
xp/n−1
i
e−xp
1−···−xp
ndx.
(38)
Let t = x∗
m and let us assume, that there is only one coordinate j = 1, . . . , n, such
that xj = t. Obviously, this assumption holds almost everywhere. Of course, we
have n possibilities for j. Furthermore, m− 1 from the remaining n− 1 components
of x are bigger than t and the remaining n− m components are smaller. This allows
to rewrite (38) as
cn
p n(cid:18) n − 1
m − 1(cid:19)Z ∞
0
tp/ne−tp(cid:18)Z t
0
up/n−1e−up
du(cid:19)n−m
×
cn
p n
pn (cid:18) n − 1
m − 1(cid:19)Z ∞
0
×(cid:18)Z ∞
t
×(cid:18)Z ∞
ω
dt
up/n−1e−up
du(cid:19)m−1
ω1/p+1/n−1e−ω(cid:18)Z ω
s1/n−1e−sds(cid:19)m−1
0
dω.
s1/n−1e−sds(cid:19)n−m
×
=
Let us denote
γ = Γ(1/n) =Z ∞
0
s1/n−1e−sds
and y(ω) = γ−1 ·Z ω
0
s1/n−1e−sds.
(39)
Then y(ω) is a non-decreasing function of ω, y(0) = 0 and limω→∞ y(ω) = 1. We
denote by ω(y) its inverse function, i.e.
y = γ−1 ·Z ω(y)
0
s1/n−1e−sds,
0 ≤ y ≤ 1.
(40)
26
Using this notation, we obtain
E x∗
m(cid:18) n
Yi=1
xp/n−1
i
(cid:19) =
p γn
cn
pn n(cid:18) n − 1
m − 1(cid:19)Z 1
0
ω(y)1/pyn−m(1 − y)m−1dy
and
σp,∞
m−1(θp,p/n−1) =
Γ(n + 1)
Γ(m)Γ(n − m + 1)Z 1
0
ω(y)1/pyn−m(1 − y)m−1dy,
(41)
where ω(y) is given by (40).
Step 1. Estimate from below
The estimate
γy =Z ω(y)
0
s1/n−1e−sds ≤Z ω(y)
0
s1/n−1ds = nω(y)1/n
implies together with Lemma 15
with c independent of n. This gives finally
σp,∞
m−1(θp,p/n−1) ≥ c1/p ·
= c1/p ·
= c1/p ·
Γ(n + 1)
≥ cyn
ω(y) ≥(cid:16) γy
n (cid:17)n
Γ(m)Γ(n − m + 1) ·Z 1
Γ(m)Γ(n − m + 1) · B(n/p + n − m + 1, m)
Γ(n − m + 1) ·
Γ(n/p + n − m + 1)
Γ(n/p + n + 1)
Γ(n + 1)
Γ(n + 1)
yn/p+n−m(1 − y)m−1dy
0
,
where we used the Beta function (10) and the proof of (35) is complete.
Step 2. Estimate from above
Let us first take y, such that 1 − e−1/γ ≤ y ≤ 1. Then − ln(γ(1 − y)) ≥ 1 and
Z ∞
− ln(γ(1−y))
s1/n−1e−sds ≤Z ∞
− ln(γ(1−y))
e−sds = γ(1 − y).
Hence,
ω(y) ≤ − ln(γ(1 − y)),
1 − e−1/γ ≤ y ≤ 1.
(42)
Finally, we observe, that
f : y →Z ∞
Cyn
s1/n−1e−sds
is a convex function on R+, f (0) = γ and
f (1 − e−1/γ) =Z ∞
≤Z ∞
1
s1/n−1e−sds
C(1−e−1/γ)n
s1/n−1e−sds ≤ e−1,
27
if we choose C so large, that C(1 − e−1/γ)n ≥ 1 for all n ∈ N. This is indeed
possible, while a byproduct of Lemma 15 is also a relation limn→∞ γ/n = 1. Using
the convexity of f , we obtain
f (y) ≤ γ(1 − y),
0 ≤ y ≤ 1 − e−1/γ,
which further leads to
ω(y) ≤ Cyn,
0 ≤ y ≤ 1 − e−1/γ.
We insert (42) and (43) into (41) and obtain
σp,∞
m−1(θp,p/n−1) ≤
Γ(n + 1)
Γ(m)Γ(n − m + 1)nC 1/pI1 + I2o ,
(43)
(44)
where
and
0
yn/p+n−m(1 − y)m−1dy
I1 :=Z 1−e−1/γ
1−e−1/γ ln(γ(1 − y))1/pyn−m(1 − y)m−1dy.
I2 :=Z 1
The first integral may be estimated again using the Beta function, which gives
I1 ≤ B(n/p + n − m + 1, m).
(45)
We denote by k the uniquely defined integer, such that 1/p ≤ k < 1/p + 1 holds,
and estimate
I2 ≤Z 1
1−e−1/γ ln(γ(1 − y))1/p(1 − y)m−1dy ≤ Ik,m :=Z e−1/γ
ln(γy)kym−1dy.
0
Next, we use partial integration to estimate Ik,m. We obtain
Ik,m =
1
γ (cid:19)m
m(cid:18) e−1
+
k
m · Ik−1,m.
Together with I0,m = 1/m · (e−1/γ)m, this leads finally to
Ik,m ≤
(k + 1)!
γ (cid:19)m
m (cid:18) e−1
.
This, together with (44) and (45) finishes the proof of (36).
The proof of (37) then follows directly by Stirling's formula (11).
Remark 5.
(i) Let us take m = 0. Then the formula (37) describes an essen-
tially different behavior compared to the normalized cone and surface mea-
sure. Namely, the expected value of the largest coordinate of x ∈ ∆n
p with
respect to θp,p/n−1 does not decay to zero with n growing to infinity. We shall
demonstrate this effect also numerically in next section.
28
(ii) If m > 0, then (37) shows, that σp,∞
m (θp,p/n−1) decays exponentially fast with
m, as soon as n is large enough. That means, that for n large enough, the
average vector of ∆n
p exhibits a strong sparsity-like structure. Namely, its m-th
largest component decays exponentially with m.
(iii) We have chosen in (32) a different β for each n, namely βn = p/n − 1 > −1.
This was of course a crucial ingredient in the proof of Theorem 16. It is not
difficult to modify the analysis of the proof of Theorem 16 to the situation,
when β > −1 is fixed for all n ∈ N.
In this case we obtain again, that
(up to logarithmic factors) σp,∞
(θp,β) is equivalent to n−1/p with constants of
equivalence depending on p > 0 and β > −1.
0
(iv) Last, but not least, we observe, that one may choose p = 1 or even p = 2 in
Theorem 16 and still obtains the exponential decay of coordinates as described
by (37). It seems, that there is no significant connection between sparsity of
an average vector of x ∈ ∆n
p and the size of p > 0.
5 Numerical experiments
5.1 Cone measure
We would like to demonstrate the most significant effects of the theory also by
numerical experiments. We start with the case of the cone measure. The key role is
played by (14). It may be interpreted in the following way. To generate a random
point on ∆n
p with respect to the normalized cone measure, it is enough to generate
ω1, . . . , ωn with respect to the density cpe−tp
, t > 0 and then calculate
(ω1, . . . , ωn)
j=1 ωp
(cid:0)Pn
p .
j(cid:1)1/p ∈ ∆n
This method is very practical, as the running time of this algorithm depends only
linearly on n.
Let us note, that the values of ωi may be generated very easily. For example the
package GNU Scientific Library [26] implements a random number generator with
respect to the gamma distribution using the method described in the classical work of
Knuth [31]. Using this package, we generated 108 random points x ∈ ∆n
p for n = 100
x∗
and p ∈ {1/2, 1, 2} to approximate numerically the value of n1/p·R∆n
mdµp(x). The
result may be found in the Figure 1.
p
5.2 Tensor measures
As pointed out in Remark 4, point (iii), a random point on ∆n
1, . . . , ω′
may be generated in the following way. We generate ω′
density cp,βtβe−tp
consider the vector
p with respect to θp,β
n with respect to the
dt is a normalizing constant and we
, t > 0, where c−1
tβe−tp
0
(ω′
p,β =R ∞
(cid:0)Pn
1, . . . , ω′
n)
j=1(ω′
p .
j)p(cid:1)1/p ∈ ∆n
29
Also this may be easily done with the help of [26]. We generated again 108 random
points x ∈ ∆n
p with respect to θp,p/n−1 for n = 100 and p ∈ {1/2, 1, 2}. Then we used
those points to numerically approximate the expression log10(R∆n
x∗
mdθp,p/n−1).
p
14
12
10
8
6
4
2
0
0
20
40
60
80
100
0
−20
−40
−60
−80
−100
−120
−140
−160
−180
0
20
40
60
80
100
(a) n1/p
· R∆n
p
x∗
mdµp(x)
(b) log10(R∆n
p
x∗
mdθp,p/n−1)
x∗
Figure 1: Approximations of n1/p ·R∆n
mdθp,p/n−1)
(right) for n = 100, p = 1/2(◦), p = 1(•) and p = 2(×) based on sampling of 108
random points.
mdµp(x) (left) and log10(R∆n
x∗
p
p
Acknowledgments
I would like to thank to Stephan Dahlke, Massimo Fornasier, Aicke Hinrichs, Erich
Novak and Henryk Wo´zniakowski for their interest in this topic and the anonymous
referees for their valuable comments and remarks, which helped to greatly improve
the quality of the presented paper. In particular, the proof of the second part of
Theorem 7 was suggested by one of the referees. I acknowledge the financial support
provided by the START-award "Sparse Approximation and Optimization in High
Dimensions" of the Fonds zur Forderung der wissenschaftlichen Forschung (FWF,
Austrian Science Foundation).
References
[1] M. Abramowitz and I. A. Stegun, Handbook of mathematical functions with
formulas, graphs, and mathematical tables, U.S. Government Printing Office,
Washington, D.C. 1964.
[2] M. Anttila, K. Ball and I. Perissinaki, The central limit problem for convex
bodies, Trans. Amer. Math. Soc. 355 (2003), no. 12, 4723 -- 4735.
[3] K. Ball and I. Perissinaki, The subindependence of coordinate slabs in ℓn
p balls,
Israel J. Math. 107 (1998), 289 -- 299.
30
[4] F. Barthe, M. Csornyei and A. Naor, A note on simultaneous polar and Carte-
in: Geometric Aspects of Functional Analysis, Lecture
sian decomposition,
Notes in Mathematics, Springer, Berlin, 2003.
[5] F. Barthe, O. Gu´edon, S. Mendelson and A. Naor, A probabilistic approach to
the geometry of the ln
p -ball, Ann. Probab. 33 (2005), no. 2, 480 -- 513.
[6] C. Bennett and R. Sharpley, Interpolation of operators, Pure and Applied Math-
ematics, 129, Academic Press, Boston, 1988.
[7] J. Bobin, J.-L. Starck, J. M. Fadili, Y. Moudden and D. L. Donoho, Morpho-
logical Component Analysis: An Adaptive Thresholding Strategy, IEEE Trans.
Image Process. 16 (2007), no. 11, 2675 -- 2681.
[8] E. J. Cand´es, J. K. Romberg and T. Tao, Stable signal recovery from incomplete
and inaccurate measurements, Comm. Pure Appl. Math. 59 (2006), no. 8, 1207 --
1223.
[9] E. J. Cand´es, Compressive sampling, In Proceedings of the International
Congress of Mathematicians, Madrid, Spain, 2006.
[10] E. J. Cand´es, J. K. Romberg and T. Tao, Robust uncertainty principles: ex-
act signal reconstruction from highly incomplete frequency information, IEEE
Trans. Inform. Theory 52 (2006), no. 2, 489 -- 509.
[11] E. J. Cand´es and T. Tao, Decoding by linear programming, IEEE Trans. Inform.
Theory 51 (2005), no. 12, 4203 -- 4215.
[12] V. Cevher, Learning with compressible priors, In Neural Information Processing
Systems (NIPS), 2009.
[13] F. Champagnat, Y. Goussard and J. Idier, Unsupervised deconvolution of sparse
spike trains using stochastic approximation, IEEE Trans. Signal Process. 44
(1996), no. 12, 2988 -- 2998.
[14] A. Cohen, W. Dahmen, and R. DeVore, Compressed sensing and best k-term
approximation, J. Amer. Math. Soc. 22 (2009), no. 1, 211 -- 231.
[15] S. Dahlke, E. Novak and W. Sickel, Optimal approximation of elliptic problems
by linear and nonlinear mappings I, J. Complexity 22 (2006), no. 1, 29-49.
[16] H. A. David and H. N. Nagaraja, Order Statistics, Wiley-Interscience, 2004
[17] R. A. DeVore, Nonlinear approximation, Acta Num. 51 -- 150, (1998).
[18] R. A. DeVore, B. Jawerth and V. Popov, Compression of wavelet decomposi-
tions, Amer. J. Math. 114 (1992), no. 4, 737-785.
[19] D. L. Donoho, Compressed sensing, IEEE Trans. Inform. Theory 52 (2006), no.
4, 1289 -- 1306.
31
[20] A. Dvoretzky, Some results on convex bodies and Banach spaces, Proc. Internat.
Sympos. Linear Spaces - Jerusalem 1960, (1961), 123 -- 160.
[21] J. Edwards, A treatise on the integral calculus, Vol. II, Chelsea Publishing Com-
pany, New York, 1922.
[22] T. Figiel, A short proof of Dvoretzky's theorem on almost spherical sections of
convex bodies, Compositio Math. 33 (1976), no. 3, 297 -- 301.
[23] T. Figiel, J. Lindenstrauss and V. D. Milman, The dimension of almost spherical
sections of convex bodies, Acta Math. 139 (1977), no. 1-2, 53 -- 94.
[24] M. Fornasier, Numerical methods for sparse recovery, Theoretical Foundations
and Numerical Methods for Sparse Recovery, (Massimo Fornasier Ed.) Radon
Series on Computational and Applied Mathematics 9, 2010.
[25] S. Foucart and H. Rauhut, A mathematical introduction to compressive sensing,
Appl. Numer. Harmon. Anal., Birkhauser, Boston, in preparation.
[26] GNU Scientific Library, http://www.gnu.org/software/gsl/
[27] E. D. Gluskin, An octahedron is poorly approximated by random subspaces,
Funktsional. Anal. i Prilozen. 20 (1986), no. 1, 14 -- 20, 96.
[28] Y. Gordon, A. E. Litvak, C. Schutt, and E. Werner, On the minimum of several
random variables, Proc. Amer. Math. Soc. 134 (2006), no. 12, 3665 -- 3675.
[29] R. Gribonval, V. Cevher, and M. Davies, Compressible priors for high-
dimensional statistics, preprint, 2011.
[30] R. Gribonval and K. Schnass, Dictionary identification - sparse matrix fac-
torisation via ℓ1 minimisation, IEEE Trans. Infor. Theory 56 (2010), no. 7,
3523 -- 3539.
[31] D. E. Knuth, The Art of Computer Programming, Vol. 2: Seminumerical Al-
gorithms, 3rd ed., Addison-Wesley 1998.
[32] M. Ledoux, The concentration of measure phenomenon, AMS, 2001.
[33] M. Ledoux and M. Talagrand, Probability in Banach spaces. Springer-Verlag,
Berlin, 1991.
[34] P. Mattila, Geometry of sets and measures in Euclidean Spaces, Cambridge
University Press, 1995.
[35] V. D. Milman, A new proof of A. Dvoretzky's theorem on cross-sections of
convex bodies, Funkcional. Anal. i Prilozen. 5 (1971), no. 4, 28 -- 37.
[36] V. D. Milman and G. Schechtman, Asymptotic theory of finite-dimensional
normed spaces, Lecture Notes in Mathematics, 1200, Springer-Verlag, Berlin,
1986.
32
[37] A. Naor, The surface measure and cone measure on the sphere of ln
p . Trans.
Amer. Math. Soc. 359 (2007), no. 3, 1045 -- 1079.
[38] A. Naor and D. Romik, Projecting the surface measure of the sphere of ln
p , Ann.
Inst. H. Poincar´e Probab. Statist. 39 (2003), no. 2, 241 -- 261.
[39] K. Oskolkov, Polygonal approximation of functions of two variables, Math.
USSR Sbornik 35, 851 -- 861, (1979).
[40] J. C. Pesquet, H. Krim, D. Leporini and E. Hamman, Bayesian approach to
best basis selection, In Proc. IEEE Int. Conf. on Acoustics, Speech, and Signal
Proc., pages 2634 -- 2637, 1996.
[41] S. T. Rachev and L. Ruschendorf, Approximate independence of distributions on
spheres and their stability properties, Ann. Probab. 19 (1991), no. 3, 1311 -- 1337.
[42] H. Rauhut, Compressive sensing and structured random matrices, Theoretical
Foundations and Numerical Methods for Sparse Recovery, (Massimo Fornasier
Ed.) Radon Series on Computational and Applied Mathematics 9, 2010.
[43] G. Schechtman and J. Zinn, On the volume of the intersection of two Ln
p balls,
Proc. AMS 110 (1), 217 -- 224, (1990).
[44] E. Schmidt, Zur Theorie der linearen und nichtlinearen Integralgleichungen I,
Math. Anal. 63, 433 -- 476, (1907).
[45] V. N. Temlyakov, Nonlinear methods of approximation, Found. Comput. Math.
3 (2003), no. 1, 33-107.
33
|
1901.03777 | 2 | 1901 | 2019-09-18T09:17:54 | Multi-marginal maximal monotonicity and convex analysis | [
"math.FA",
"math.OC"
] | Monotonicity and convex analysis arise naturally in the framework of multi-marginal optimal transport theory. However, a comprehensive multi-marginal monotonicity and convex analysis theory is still missing. To this end we study extensions of classical monotone operator theory and convex analysis into the multi-marginal setting. We characterize multi-marginal c-monotonicity in terms of classical monotonicity and firmly nonexpansive mappings. We provide Minty type, continuity and conjugacy criteria for multi-marginal maximal monotonicity. We extend the partition of the identity into a sum of firmly nonexpansive mappings and Moreau's decomposition of the quadratic function into envelopes and proximal mappings into the multi-marginal settings. We illustrate our discussion with examples and provide applications for the determination of multi-marginal maximal monotonicity and multi-marginal conjugacy. We also point out several open questions. | math.FA | math |
Multi-marginal maximal monotonicity
and convex analysis
Sedi Bartz∗, Heinz H. Bauschke†, Hung M. Phan‡, and Xianfu Wang§
September 12, 2019
Abstract
Monotonicity and convex analysis arise naturally in the framework of multi-marginal op-
timal transport theory. However, a comprehensive multi-marginal monotonicity and convex
analysis theory is still missing. To this end we study extensions of classical monotone operator
theory and convex analysis into the multi-marginal setting. We characterize multi-marginal
c-monotonicity in terms of classical monotonicity and firmly nonexpansive mappings. We
provide Minty type, continuity and conjugacy criteria for multi-marginal maximal monotonic-
ity. We extend the partition of the identity into a sum of firmly nonexpansive mappings and
Moreau's decomposition of the quadratic function into envelopes and proximal mappings into
the multi-marginal settings. We illustrate our discussion with examples and provide applica-
tions for the determination of multi-marginal maximal monotonicity and multi-marginal con-
jugacy. We also point out several open questions.
2010 Mathematics Subject Classification: Primary 47H05, 26B25; Secondary 49N15, 49K30, 52A01, 91B68.
Keywords: c-convexity, c-monotonicity, c-splitting set, cyclic monotonicity, Kantorovich duality, maximal
monotonicity, Minty Theorem, Moreau envelope, multi-marginal, optimal transport.
1
Introduction
Our discussion stems from multi-marginal optimal transport theory: Let (X1, µ1), . . . , (XN, µN)
be Borel probability spaces. We set X = X1 × · · · × XN and we denote by Π(X) the set of all
Borel probability measures π on X such that the marginals of π are the µi's. Let c : X → R
∗Mathematics, University of Massachusetts Lowell, MA 01854, USA. E-mail: sedi [email protected].
†Mathematics, University of British Columbia, Kelowna, B.C. V1V 1V7, Canada. E-mail: [email protected].
‡Mathematics, University of Massachusetts Lowell, MA 01854, USA. E-mail: hung [email protected].
§Mathematics, University of British Columbia, Kelowna, B.C. V1V 1V7, Canada. E-mail: [email protected].
1
be a cost function. A cornerstone of multi-marginal optimal transport theory is Kellerer's [16]
generalization of the Kantorovich duality theorem to the multi-marginal case. Kellerer's duality
theorem asserts that, in a suitable framework,
min
π∈Π(X)ZX
c(x)dπ(x) =
max
ui ∈ L1(µi),
∑1≤i≤N ui ≤ c
∑
1≤i≤NZXi
ui(xi)dµi(xi).
(1)
It follows that if π is a solution of the left-hand side of (1) and (u1, . . . , uN) is a solution of the
right-hand side of (1), then π is concentrated on the subset Γ of X where the equality c = ∑1≤i≤N ui
holds. In recent publications (see, for example, [5, 15, 17]) such subsets Γ of X are referred to as
c-splitting sets: Let N ≥ 2 be a natural number and I = {1, . . . , N} an index set. Let X1, . . . , XN be
nonempty sets, X = X1 × · · · × XN and c : X → R a function.
Definition 1.1 (c-splitting set) Let Γ ⊆ X. We say that Γ is a c-splitting set if for each i ∈ I there exists
a function ui : Xi → ]−∞, +∞] such that
and
∀x = (x1, . . . , xN) ∈ X,
∀x = (x1, . . . , xN) ∈ Γ,
c(x1, . . . , xN) ≤ Mi∈I
c(x1, . . . , xN) = Mi∈I
ui! (x) := ∑
i∈I
ui(xi)
ui! (x) := ∑
i∈I
ui(xi).
(2)
(3)
In this case we say that (u1, . . . , uN) is a c-splitting tuple of Γ. Given functions ui : Xi → ]−∞, +∞] that
satisfy (2), we call the set of all points (x1, . . . , xN) ∈ X that satisfy (3) the c-splitting set generated by the
tuple (u1, . . . , uN).
In the case N = 2, splitting sets are natural in convex analysis as graphs of subdifferentials.
Indeed, by the Young-Fenchel inequality the graph of the subdifferential ∂ f is the c-splitting set
generated by the pair ( f , f ∗) where c = h·, ·i is the classical pairing between a linear space and its
dual. Similar to the two-marginal case, in the multi-marginal case monotonicity arises naturally
as well:
Definition 1.2 (c-cyclic monotonicity) The subset Γ of X is said to be c-cyclically monotone of order n,
n-c-monotone for short, if for all n tuples (x1
N) in Γ and every N permutations
σ1, . . . , σN in Sn,
N), . . . , (xn
1 , . . . , xn
1, . . . , x1
n
∑
j=1
c(xσ1(j)
1
, . . . , xσN (j)
N
) ≤
n
∑
j=1
c(xj
1, . . . , xj
N);
(4)
Γ is said to be c-cyclically monotone if it is n-c-monotone for every n ∈ {2, 3, . . . }; and Γ is said to be
c-monotone if it is 2-c-monotone. Finally, Γ is said to be maximally n-c-monotone if it has no proper
n-c-monotone extension.
2
Cyclic monotonicity was first introduced by Rockafellar [24] in the framework of classical con-
vex analysis. During the late 80s and early 90s (see [8, 23, 26]) the concept was generalized to
c-cyclic monotonicity in order to hold for more general cost functions c in the framework of two-
marginal optimal transport theory. Currently, it lays at the foundations of the theory (see for
example [11, 28, 30]) and plays a role also in recent refinements (see, for example, [2, 3]). Ex-
tending the role it plays in two-marginal optimal transport theory, in the past two and a half
decades multi-marginal c-monotonicity and aspects of c-convex analysis are becoming an integral
part of the fast evolving multi-marginal optimal transport theory as can be seen, for example, in
[1, 5, 7, 9, 10, 12, 13, 14, 15, 17, 18, 19, 20, 21, 22, 27]. An important instance of an extension from
the two-marginal case relating Definition 1.1 with Definition 1.2 is the known fact that c-splitting
sets are c-cyclically monotone (see, for example, [5, 15, 17, 18]).
Before attending our convex analytic discussion we remark that in order to make optimal trans-
port compatible with our discussion, one should exchange min for max in the left-hand side of (1),
exchange max for min in the right-hand side of (1) and, finally, exchange the constraint ∑i ui ≤ c
in the right-hand side of (1) with the constraint c ≤ ∑i ui as we did in Definition 1.1 and Defini-
tion 1.2.
In the framework of multi-marginal optimal transport, presumably the most traditional and
well studied cost functions are classical extensions of the pairing between a linear space and its
dual:
For the remainder of our discussion, for each 1 ≤ i ≤ N, we assume that Xi = H is a real
Hilbert space with inner product h·, ·i and induced norm k · k. We let c : X → R be the cost
function defined by
c(x1, . . . , xN) = ∑
1≤i<j≤Nhxi, xji.
It follows from straightforward computation (see for example [5]) that a set Γ ⊆ X is n-c-monotone
if and only if it is n-c-monotone with respect to each of the functions
(x1, . . . , xN) 7→ − ∑
1≤i<j≤N
1
2kxi − xjk2
and
(x1, . . . , xN) 7→ 1
N
∑
i=1
2(cid:13)(cid:13)(cid:13)(cid:13)
2
.
xi(cid:13)(cid:13)(cid:13)(cid:13)
Although classical convex analysis and monotonicity are instrumental in multi-marginal optimal
transport, and although several multi-marginal convex analytic results are already available (as
we recall in our more specific discussion further below), to the best of our knowledge, a com-
prehensive multi-marginal monotonicity and convex analysis theory is still lacking. To this end,
in the present paper we lay additional foundations and provide several extensions of classical
monotone operator theory and convex analysis into the multi-marginal settings.
The remainder of the paper is organized as follows. In Section 2 we provide a characterization of
multi-marginal c-monotonicity in terms of classical monotonicity. We employ this characterization
in order to provide several equivalent criteria, including a Minty-type criterion, a criterion based
3
on the partition of the identity into a sum of firmly nonexpansive mappings, and other criteria for
multi-marginal maximal c-monotonicity. In Section 3 we provide a continuity criterion for multi-
marginal maximal monotonicity. In Section 4 we focus on multi-marginal convex analysis. In
particular, we extend Moreau's decompositions and provide criteria for maximal c-monotonicity
of c-splitting sets, the multi-marginal extensions of subdifferentials. We show that the same cri-
teria also imply multi-marginal c-conjugacy of c-splitting functions. In the case N = 3 we also
provide a class of c-splitting triples for which c-conjugacy implies maximal c-monotonicity. Sec-
tion 5 contains examples and applications of our results to the problem of determining maximal
c-monotonicity of sets and c-conjugacy of c-splitting tuples, thus reducing the need of further
challenging computations of multi-marginal c-conjugate tuples. Additionally, we point out sev-
eral open problems.
In the remainder of this section we collect standard notations and preliminary facts from clas-
sical monotone operator theory and convex analysis which, largely, follow [6]. Let A : H ⇒ H be
a set-valued mapping. The domain of A is the set dom A = {x ∈ H Ax 6= ∅}. The range of A is
the set ran A = A(H) = Sx∈H Ax, the graph of A is the set gra A = {(x, u) ∈ H × H u ∈ Ax}
and the inverse mapping of A is the mapping A−1 satisfying x ∈ A−1u ⇔ u ∈ Ax. A is said to be
monotone if
(∀(x, u) ∈ gra A)(∀(y, v) ∈ gra A)
hx − y, u − vi ≥ 0.
A is said to be maximally monotone if there exists no monotone operator B such that gra A is a
proper subset of gra B. The resolvent of A is the mapping JA = (A + Id)−1 where Id is the identity
mapping. The mapping T : dom T ⊆ H → H is said to be firmly nonexpansive if
(∀x ∈ dom T)(∀y ∈ dom T) kTx − Tyk2 + k(Id −T)x − (Id −T)yk2 ≤ kx − yk2,
: H → ]−∞, +∞] is said to be proper if dom f
where dom T ⊆ H. The function f
H f (x) < ∞} 6= ∅. The Fenchel conjugate of the function f is the function f ∗ defined by
:= {x ∈
We set q(·) = 1
2k · k2. The Moreau envelope of f is the function defined by the infimal convolution
f ∗(u) = sup
x∈H(cid:0)hu, xi − f (x)(cid:1).
The subdifferential of the proper function f is the mapping ∂ f : H ⇒ H defined by
e f (s) = ( f (cid:3)q)(s) = inf
x∈H(cid:0) f (x) + q(s − x)(cid:1).
(5)
(6)
∂ f (x) = (cid:8)u ∈ H (cid:12)(cid:12) f (x) + hu, y − xi ≤ f (y), ∀y ∈ H(cid:9).
The indicator function of a subset C of H is the function ιC : H → ]−∞, +∞] which vanishes on C
and equals +∞ on H r C.
Fact 1.3 (Minty's Theorem [6, Theorem 21.1]) Let A : H ⇒ H be monotone. Then A is maximally
monotone if and only if ran(Id +A) = H.
Fact 1.4 ([6, Proposition 23.8]) Let A : H ⇒ H. Then
4
(i) JA is firmly nonexpansive if and only if A is monotone;
(ii) JA is firmly nonexpansive and dom JA = H if and only if A is maximally monotone.
Let f be a proper lower semicontinuous convex function. The proximity operator [6, Defini-
tion 12.23] of f is defined by
Prox f : H → H : x 7→ Prox f x = argmin
y∈H (cid:0) f (y) + q(y − x)(cid:1).
(7)
For all s ∈ H, [6, Proposition 12.15] implies that there is a unique minimizer of f (·) + q(s − ·)
over all x ∈ H; thus, the proximity operator of f is well defined. Furthermore, we also have
Prox f = J∂ f .
Additional properties of the Moreau envelope are:
Fact 1.5 (Moreau envelope) Let f be a proper lower semicontinuous convex function. The following
assertions hold:
(i) (Moreau decomposition) e f + e f ∗ = q.
(ii) x = Prox f s ⇔ e f (s) = f (x) + q(s − x).
(iii) ([6, Proposition 12.30]) e f is Fr´echet differentiable with ∇e f = Id − Prox f .
Finally, we set the marginal projections Pi : X → Xi : (x1, . . . , xN) 7→ xi for i in {1, . . . , N} and
the two-marginal projections Pi,j : X → Xi × Xj : (x1, . . . , xN) 7→ (xi, xj) for i < j in {1, . . . , N}.
Given a subset Γ of X, we set
Γi = Pi(Γ)
and
Γi,j = Pi,j(Γ)
(8)
We also define Ai,j : Xi ⇒ Xj via
gra Ai,j = Γi,j.
The notation Ai is reserved for a different purpose and introduced in Section 2.
2 A characterization of multi-marginal c-monotonicity and Minty type
criteria for c-monotonicity
Let S : H × H → H be the mapping defined by S(x, y) = x + y. For any mapping A : H ⇒ H, we
have the identity [25, Lemma 12.14]
(9)
JA−1 = Id−JA.
If, in addition, A is monotone, then by Fact 1.4, JA and JA−1 are single-valued, thus,
JA + JA−1 = IdS(gra A),
5
(10)
which is equivalent to gra A being parameterized by
gra A = (cid:8)(JAs, JA−1 s) s ∈ S(gra A)(cid:9).
(11)
Given a set Γ ⊆ X, we now associate with Γ monotone mappings as follows.
Definition 2.1 Let Γ ⊆ X be a set. For each index set ∅ 6= K ( I, we define the mapping AK : H ⇒ H
by
gra AK = (cid:26)(cid:16) ∑
i∈K
xi, ∑
i∈I\K
and for each i ∈ I we set Ai = A{i}.
(x1, . . . , xN) ∈ Γ(cid:27)
xi(cid:17)(cid:12)(cid:12)(cid:12)
(12)
Our first aim is to characterize the c-monotonicity of a set Γ in terms of the monotonicity of its
AK's, and furthermore, extend (10) and (11) to the multi-marginal settings. To this end we will
employ the sum mapping
and the following fact which follows by a straightforward computation (see, e.g., [5, Fact 3.3]).
S : X → H : (x1, . . . , xN) 7→ ∑
i∈I
xi,
(13)
Fact 2.2 Let x ∈ X. If the subset Γ of X is n-c-cyclically monotone, then so is Γ + x.
Lemma 2.3 Let Γ ⊆ X be a set. Then the following assertions are equivalent:
(i) Γ is c-monotone;
(ii) For each ∅ 6= K ( I, the mapping AK is monotone;
(iii) For each ∅ 6= K ( I, the mapping JAK : S(Γ) → H is firmly nonexpansive.
In this case,
equivalently, Γ can be parameterized by
JA1 + · · · + JAN = Id S(Γ),
and, furthermore, for each ∅ 6= K ( I,
Γ = (cid:8)(JA1s, . . . , JAN s) s ∈ S(Γ)(cid:9);
JAK = ∑
i∈K
JAi.
(14)
(15)
(16)
Proof. (i) ⇔ (ii): First we characterize the c-monotone relations of the set {z, 0} in X. We employ a
similar computation to the one in [5, Lemma 4.1]: For z = (z1, . . . , zN) ∈ X and ∅ 6= K ( I we set
zK = (zK
1 , . . . , zK
N) ∈ X by
zK
i = (zi,
0,
i ∈ K;
i ∈ I \ K.
6
From Definition 1.2 it follows that {z, 0} is c-monotone if and only if for each ∅ 6= K ( I
0 ≤ c(z) + c(0) − c(zK) − c(zI\K)
i,j∈I, i<jhzK
= ∑
i,j∈I, i<jhzi, zji + 0 − ∑
i,j∈I, i<jhzi, zji − ∑
= ∑
i,j∈K, i<jhzi, zji − ∑
i , zK
i
j
, zI\K
i,j∈I, i<jhzI\K
j i − ∑
i
i,j∈I\K, i<jhzi, zji = (cid:28) ∑
i∈K
zi, ∑
i∈I\K
zi(cid:29).
In general, from Definition 1.2 it follows that the set Γ ⊆ X is c-monotone if and only if for any
x ∈ Γ and y ∈ Γ, the set {x, y} is c-monotone, which, in turn, by invoking Fact 2.2, is equivalent to
the set {x − y, 0} being c-monotone. Summing up, we see that Γ is c-monotone if and only if for
any x = (x1, . . . , xN), y = (y1, . . . , yN) ∈ Γ and any ∅ 6= K ( I, by letting z = x − y,
0 ≤ (cid:28) ∑
i∈K
xi − ∑
i∈K
yi, ∑
i∈I\K
xi − ∑
i∈I\K
yi(cid:29),
i.e., AK is monotone.
(ii) ⇔ (iii): By the definition of AK, it follows that dom JAK = S(Γ). Thus, the equivalence (ii) ⇔
(iii) follows immediately from Fact 1.4(i).
Finally, (14), (15) and (16) follow from (iii) and the definition of AK.
(cid:4)
We now address maximal c-monotonicity. Equivalent statements of Minty's characterization
are: Let A : H ⇒ H be a monotone mapping. Then A is maximally monotone if and only if
equivalently,
S(cid:0) gra(A)(cid:1) = H,
gra(A) + gra(− Id) = H × H.
In order to extend our discussion of these formulas into the multi-marginal settings we will
employ the following definitions and notations. We denote by ∆ the subset of X = X1 × · · · × XN
defined by
∆ = (cid:8)(x, . . . , x)(cid:12)(cid:12) x ∈ H(cid:9).
Consequently, ∆⊥ = n(x1, . . . , xN) ∈ X (cid:12)(cid:12)(cid:12)
Corollary 2.4 Let Γ ⊆ X be a c-monotone set. Then for every u, v ∈ Γ,
u − v ∈ ∆⊥
⇔
u = v.
Proof. Let u = (u1, . . . , uN) and v = (v1, . . . , vN) belong to Γ and suppose that
u − v = d = (d1, . . . , dN) ∈ ∆⊥.
7
N
∑
i=1
xi = 0o.
(17)
(18)
We prove that di = 0 for each 1 ≤ i ≤ N. To this end, set 1 ≤ i0 ≤ N. By Lemma 2.3, Ai0 is
monotone. Consequently we see that
0 ≤ (cid:28)ui0 − vi0, ∑
i6=i0
ui − ∑
i6=i0
vi(cid:29) = (cid:28)di0, ∑
i6=i0
di(cid:29) = hdi0, −di0i = −kdi0k2 ≤ 0.
(cid:4)
Combining Lemma 2.3 and Corollary 2.4 with classical two-marginal monotone operator theory,
we arrive at the following result.
Theorem 2.5 (multi-marginal maximal c-monotonicity) Let Γ ⊆ X be a c-monotone set. Then the
following assertions are equivalent:
(i) For each ∅ 6= K ( I the mapping AK defined by (12) is maximally monotone;
(ii) There exists ∅ 6= K ( I such that the mapping AK is maximally monotone;
(iii) Γ + ∆⊥ = X;
(iv) JA1 + · · · + JAN = Id;
(v) For each ∅ 6= K ( I the firmly nonexpansive mapping JAK : H → H has full domain and JAK =
∑i∈K JAi;
(vi) S(Γ) = H.
In this case, Γ is maximally c-monotone.
Proof. (i) ⇒ (ii) is trivial.
(ii) ⇒ (iii): Suppose that AK is maximally monotone and let a = (a1, . . . , aN) ∈ X. We will prove
that there exist x = (x1, . . . , xN) ∈ Γ and d = (d1, . . . , dN) ∈ ∆⊥ such that x + d = a. Indeed, the
maximal monotonicity of AK implies that ran(AK + Id) = H. Consequently, by the definition of
AK, there exists x = (x1, . . . , xN) ∈ Γ such that ∑N
i=1 ai. For each 1 ≤ i ≤ N we let
di = ai − xi. Then ∑N
i=1 di = 0, that is d = (d1, . . . , dN) ∈ ∆⊥ and x + d = a.
i=1 xi = ∑N
(iii) ⇒ (iv): Fix 1 ≤ i0 ≤ N. We prove that Ai0 + Id is onto.
Indeed, let s ∈ H. We
Indeed, let
i=1 hi = s. Then (iii) implies the existence of x = (x1, . . . , xN) ∈ Γ
i=1 xi = s which implies that
i=1 xi(cid:17) ∈ gra(Ai0 + Id). Thus, since Ai0 is monotone, we conclude that its re-
is firmly nonexpansive and has full domain. This is true for each 1 ≤ i0 ≤ N and
i=1 xi = s, we conclude that
prove that there exists x = (x1, . . . , xN) ∈ Γ such that (xi0, s) ∈ gra(Ai0 + Id).
h = (h1, . . . , hN) ∈ X such that ∑N
and d = (d1, . . . , dN) ∈ ∆⊥ such that x + d = h. Consequently, ∑N
(xi0, s) = (cid:16)xi0, ∑N
solvent JAi0
since for any s ∈ H there exists x = (x1, . . . , xN) ∈ Γ such that ∑N
i=1 JAi (s) = ∑N
∑N
i=1 xi = s, that is, (iv) holds.
8
(iv) ⇒ (v): Since AK is monotone for every ∅ 6= K ( I, the resolvent JAK is firmly nonexpansive
and (iv) implies it has full domain. Furthermore, by employing our notations from the previous
step, we see that for every s ∈ H, ∑i∈K JAi (s) = ∑i∈K xi = JAK (s), that is, we have arrived at (v).
(v) ⇒ (i): Let ∅ 6= K ( I. Since the resolvent JAK is firmly nonexpansive and has full domain,
AK is maximally monotone.
Summing up, we have established (i) ⇒ (ii) ⇒ (iii) ⇒ (iv) ⇒ (v) ⇒ (i).
(iv) ⇒ (vi): Since for each 1 ≤ i ≤ N, dom(JAi ) = S(Γ), then (iv) ⇒ (vi).
(vi) ⇒ (iii): Suppose that S(Γ) = H and let y ∈ X. Then there exist x ∈ Γ such that S(y) = S(x).
Consequently, y − x ∈ ∆⊥, which implies that y = x + (y − x) ∈ Γ + ∆⊥.
Finally, we prove that (iii) implies the maximal c-monotonicity of Γ. Indeed, suppose that u is c-
monotonically related to Γ. We then write u = d + v where d ∈ ∆⊥ and v ∈ Γ. Since u, v ∈ Γ ∪ {u}
which is c-monotone and u − v ∈ ∆⊥, Corollary 2.4 implies that u = v ∈ Γ.
Remark 2.6 To the best of our knowledge, the question whether the multi-marginal generalization
of the other direction of Minty's characterization of maximal monotonicity holds, namely, whether
the maximal c-monotonicity of the set Γ implies that Γ + ∆⊥ = X, equivalently, that JA1 + · · · +
JAN = Id, is still open.
(cid:4)
Remark 2.7 In the partition of the identity in (14) and in Theorem 2.5(iv) we conclude from (16)
and Theorem 2.5(v) that any partial sum of the firmly nonexpansive mappings is also firmly non-
expansive. This is not the case for general partitions of the identity into sums of firmly nonexpan-
sive mappings; indeed, an example where partial sums of a partition of the identity into firmly
nonexpansive mappings fail to be firmly nonexpansive is provided in [4, Example 4.4]. We elabo-
rate further on this in Example 5.7 below.
3 Multi-marginal maximal c-monotonicity via continuity
In the classical two-marginal case an important class of maximally monotone operators is the one
of continuous monotone operators. A continuity criterion guarantees maximality in the multi-
marginal framework as well:
Theorem 3.1 Let Γ ⊆ X be a c-monotone set. Suppose that Γ is the graph of a continuous mapping
T = (T2, . . . , TN) : X1 → ΠN
i=2Xi, i.e.,
where for each 2 ≤ i ≤ N the mapping Ti : H → H is continuous. Then Γ is maximally c-monotone.
Γ = gra(T) = (cid:8)(x, T2x, . . . , TN x)(cid:12)(cid:12) x ∈ H(cid:9)
We provide two proofs for Theorem 3.1. We begin with a direct proof.
9
Proof. Let u = (u1, . . . , uN) be c-monotonically related to Γ. We prove that u ∈ Γ. Since A1, induced
from the c-monotone set Γ ∪ {u}, is monotone,
∀x ∈ H,
0 ≤ (cid:28)u1 − x,
N
∑
i=2
(ui − Tix)(cid:29).
For t > 0 we let xt = u1 + t ∑N
i=2(ui − Tiu1). Then xt −→ u1 as t → 0+ and
0 ≤ t(cid:28) N
(ui − Tixt)(cid:29).
(Tiu1 − ui),
N
∑
i=2
∑
i=2
Since each Ti is continuous, we deduce that
(Tiu1 − ui),
N
∑
i=2
∑
i=2
0 ≤ (cid:28) N
−−−→ (cid:28) N
t→0+
∑
i=2
(Tiu1 − ui),
(ui − Tixt)(cid:29)
(ui − Tiu1)(cid:29) = −(cid:13)(cid:13)(cid:13)(cid:13)
N
∑
i=2
2
,
N
∑
i=2
(ui − Tiu1)(cid:13)(cid:13)(cid:13)(cid:13)
N
∑
i=2
ui =
N
∑
i=2
Tiu1;
(u1, . . . , uN) − (u1, T2u1, . . . , TNu1) ∈ ∆⊥.
which implies
equivalently,
Thus, by Corollary 2.4, we have (u1, . . . , uN) = (u1, T2u1, . . . , TNu1) ∈ gra T.
(19)
(cid:4)
The second proof of Theorem 3.1 employs the classical two-marginal fact that a monotone and
continuous mapping is maximally monotone [6, Corollary 20.28], Lemma 2.3 and Theorem 2.5.
Proof. Since A1(x) = ∑N
i=2 Ti(x) for every x ∈ H, by employing Lemma 2.3 it follows that A1 is
a monotone and continuous mapping, hence, maximally monotone. Consequently, by employing
Theorem 2.5 we conclude that Γ is maximally monotone.
(cid:4)
4 Maximal c-monotonicity of c-splitting sets, c-conjugate tuples and
multi-marginal convex analysis
We begin our discussion of c-splitting tuples by a known observation regarding the subdifferen-
tials of the splitting functions: As in [12, 18, 27] we observe that if ( f1, . . . , fN) is a c-splitting tuple
10
of Γ ⊆ X, then given x = (x1, . . . , xN) ∈ Γ and for any x′1 ∈ X1,
N
∑
i=1
fi(xi) = c(x1, . . . , xN)
and
c(x′1, x2, . . . , xN) ≤ f1(x′1) +
N
∑
i=2
fi(xi).
Summing up these two inequalities followed by simplifying, we see that
f1(x1) + hx′1, x2 + · · · + xNi ≤ f1(x′1) + hx1, x2 + · · · + xNi,
that is,
N
∑
i=2
xi ∈ ∂ f1(x1).
Similarly, we conclude that for each 1 ≤ i0 ≤ N,
xi ∈ ∂ fi0 (xi0).
∑
i6=i0
Since gra Ai0 = n(cid:0)xi0, ∑i6=i0 xi(cid:1)(cid:12)(cid:12)(cid:12)
(x1, . . . , xN) ∈ Γo, this implies
gra(Ai0) ⊆ gra(∂ fi0 ).
(20)
(21)
Similar observations and c-monotonicity properties of Γ from Section 2 are also related to the
Wasserstein barycenter as can be seen, for example, in [1].
We continue our discussion by a characterization of c-splitting tuples and their generated c-
splitting sets in terms of the Moreau envelopes of the splitting functions.
i=1 fi if and only if
Theorem 4.1 For each 1 ≤ i ≤ N, let fi : Xi → ]−∞, +∞] be proper, lower semicontinuous, and convex.
Then c ≤ LN
Now assume this is the case, and let Γ ⊆ X be the c-splitting set generated by ( f1, . . . , fN). Then equality
in (22) holds if and only if s = x1 + · · · + xN where (x1 . . . , xN) ∈ Γ.
Proof. The inequality c ≤ LN
i=1 fi holds if and only if for all (x1, . . . , xN) ∈ X,
∀s ∈ H,
e f ∗1
(s) + · · · + e f ∗N
(s) ≤ q(s).
(22)
c(x1, . . . , xN) ≤
fi(xi)
N
∑
i=1
(23)
(24)
⇔ q(x1 + · · · + xN) = c(x1, . . . , xN) +
N
∑
i=1
q(xi) ≤
N
∑
i=1(cid:0) fi(xi) + q(xi)(cid:1).
We see that (23) holds with equality only when (x1, . . . , xN) ∈ Γ if and only if (24) holds with
equality only when (x1, . . . , xN) ∈ Γ. Let ϕ : X → R be defined by
ϕ(x1, . . . , xN) = q(x1 + · · · + xN).
11
Then, using [6, Corollary 15.28(i)], we have
∀(x1, . . . , xN) ∈ X,
ϕ∗(x1, . . . , xN) = q(x1) + ι∆(x1, . . . , xN).
Since for each 1 ≤ i ≤ N, ( fi + q)∗ = e f ∗i
(see, for example, [6, Proposition 14.1]), we arrive at
N
Mi=1
(cid:16)
( fi + q)(cid:17)∗ =
N
Mi=1
( fi + q)∗ =
N
Mi=1
e f ∗i
.
Consequently, (classical) Fenchel conjugation transforms (24) into (22) and vise versa.
We now address the case of equality in (22). Let (x1 . . . , xN) ∈ X and s = x1 + · · · + xn. Then
for each 1 ≤ i ≤ N, by the Fenchel-Young inequality,
hs, xii ≤ ( fi + q)∗(s) + ( fi + q)(xi) = e f ∗i
(s) + ( fi + q)(xi)
(25)
with equality if and only if xi ∈ ∂( fi + q)∗(s), i.e., since ( fi + q)∗ = e f ∗i
(see, e.g., [6, Proposition 12.30]), xi = ∇e f ∗i
(s) . By summing up (25) over i, we obtain
is Fr´echet differentiable
hs, si =
N
∑
i=1hs, xii ≤
N
∑
i=1(cid:0)e f ∗i
(s) + ( fi + q)(xi)(cid:1)
(26)
with equality if and only if xi = ∇e f ∗i
(s) for every 1 ≤ i ≤ N.
(⇐): Suppose that x = (x1, . . . , xN) is in the c-splitting set Γ generated by ( f1, . . . , fN) and set
s = S(x). We prove equality in (22). It follows from (20) that for each 1 ≤ i ≤ N,
xj = s − xi ∈ ∂ fi(xi) ⇔ s ∈ ∂( fi + q)(xi),
(27)
∑
j6=i
which, in turn, implies that xi ∈ ∂( fi + q)∗(s), that is, xi = ∇e f ∗i
equality in (26) and in (24), we obtain equality in (22).
(s). Since in this case there is
(⇒): Let s ∈ H be a point where equality in (22) holds. Since ∑N
i=1 e f ∗i
entiable and ∑N
i=1 e f ∗i ≤ q, then at the point of equality s we have
e f ∗i (cid:17)(s) = ∇q(s) = s.
∇(cid:16)
N
∑
i=1
and q are Fr´echet differ-
(s) (see, e.g., [6, eq (14.7)]). Then it follows that
For each 1 ≤ i ≤ N, set xi = Prox fi(s) = ∇e f ∗i
s = x1 + · · · + xN. Thus, in order to complete the proof it is enough to prove that (x1, . . . , xN) ∈
Γ or, equivalently, that there is equality in (24). Indeed, Moreau's decomposition (see, e.g., [6,
Remark 14.4]) implies that e fi + e f ∗i
= q for each 1 ≤ i ≤ N. Consequently,
N
∑
i=1
e f ∗i
(s) = q(s)
is equivalent to
N
∑
i=1
e fi(s) = (N − 1)q(s).
12
We also note that for each 1 ≤ i ≤ N, xi = Prox fi(s) implies that
e fi(s) = min
x∈H(cid:0) fi(x) + q(s − x)(cid:1) = fi(xi) + q(s − xi).
Thus, we arrive at
N
∑
N
∑
N
∑
i=1(cid:0)q(s − xi) + fi(xi)(cid:1) = (N − 1)q(s)
N
∑
i=1hs, xii +
−
i=1(cid:0) fi(xi) + q(xi)(cid:1) = q(s).
i=1(cid:0) fi(xi) + q(xi)(cid:1) = −q(s)
⇔
⇔
We now address c-conjugation.
Definition 4.2 (c-conjugate tuple) For each 1 ≤ i ≤ N, let fi : Xi → ]−∞, +∞] be a proper function.
We say that ( f1, . . . , fN) is a c-conjugate tuple if for each 1 ≤ i0 ≤ N,
fi0 (xi0) = (cid:16)Mi6=i0
fi(cid:17)c
(xi0) := sup
i6=i0, xi∈Xi
c(x1, . . . , xi0, . . . , xN) − ∑
i6=i0
fi(xi),
xi0 ∈ Xi0.
(cid:4)
It follows that if ( f1, . . . , fN) is a c-conjugate tuple, then fi is lower semicontinuous and convex
for each 1 ≤ i ≤ N. Furthermore, it is known (see [12] and [10]) that given a c-splitting tuple
(u1, . . . , uN) of a set Γ ⊆ X, it can be relaxed into a c-conjugate c-splitting tuple ( f1, . . . , fN) of Γ by
setting
inductively,
and finally
fi0 = (cid:16) M1≤i≤i0−1
ui(cid:17)c
f1 = (cid:16) M2≤i≤N
ui(cid:17)c
fi ⊕ Mi0+1≤i≤N
fN = (cid:16) M1≤i≤N−1
for 2 ≤ i0 ≤ N − 1,
.
fi(cid:17)c
In the case N = 2, let f1 : X1 → ]−∞, +∞] be proper, lower semicontinuous and convex, let
f2 = f ∗1 : X2 → ]−∞, +∞] be its conjugate and let Γ = gra(∂ f1) ⊆ X1 × X2. Then it is well known
that Γ is maximally monotone, see, e.g., [6, Theorem 20.25]. Since f1 = f ∗∗1 = f c
2 and also f2 = f c
1 ,
then we can restate as follows:
13
Let Γ ⊆ X1 × X2 be the c-splitting set generated by the c-conjugate pair ( f1, f2). Then
Γ is maximally c-monotone and determines its c-conjugate c-splitting tuple ( f1, f2)
uniquely up to an additive constant pair (ρ, −ρ) with ρ ∈ R.
A generalization to an arbitrary N ≥ 2 would be
Let Γ ⊆ X be the c-splitting set generated by the c-conjugate tuple ( f1, . . . , fN). Then Γ
is maximally c-monotone and determines its c-conjugate c-splitting tuple ( f1, . . . , fN)
uniquely up to an additive constant tuple (ρ1, . . . , ρN) such that ∑N
i=1 ρi = 0.
To the best of our knowledge, whether or not this latter assertion is true in general is still open. We
do, however, provide a positive answer in a more particular case in Theorem 4.6 and additional
insight in Theorem 4.3.
Furthermore, we note that in the case N = 2, given a conjugate pair ( f1, f2), Moreau's decom-
position can be restated as
e f ∗1
+ e f ∗2
= q
and
Prox f1 + Prox f2 = Id .
(28)
Combining our discussion with Theorems 4.1 and 2.3, we arrive at the following generalized
multi-marginal convex analytic assertions which, in particular, generalize the decomposition (28).
To this end, we again recall that for each 1 ≤ i0 ≤ N,
gra Ai0 = n(cid:0)xi0, ∑
xi(cid:1)(cid:12)(cid:12)(cid:12)
(x1, . . . , xN) ∈ Γo.
i6=i0
: Xi → ]−∞, +∞] be convex, lower semicontinuous, and
Theorem 4.3 For each 1 ≤ i ≤ N, let fi
proper. Suppose that Γ ⊆ X is the c-splitting set generated by ( f1, . . . , fN). Then the following assertions
are equivalent:
(i) There exist 1 ≤ i0 ≤ N such that Ai0 is maximally monotone;
(ii) There exist 1 ≤ i0 ≤ N such that Ai0 = ∂ fi0 ;
(iii) Ai = ∂ fi
for each 1 ≤ i ≤ N;
(iv) Prox f1 + · · · + Prox f N = Id;
(v) e f ∗1
+ · · · + e f ∗N
= q.
In this case
(A) Γ is maximally c-monotone (and, consequently, maximally c-cyclically monotone);
14
(B) ( f1, . . . , fN) is a c-conjugate c-splitting tuple of Γ. Moreover, Γ determines its c-conjugate c-splitting
tuple ( f1, . . . , fN) uniquely up to an additive constant tuple (ρ1, . . . , ρN) such that ∑N
i=1 ρi = 0.
Proof. (i) ⇒ (ii): ∂ fi0 is monotone and gra(Ai0) ⊆ gra(∂ fi0 ) (see (21)). Consequently, since Ai0 is
maximally monotone, it follows that Ai0 = ∂ fi0 .
(ii) ⇒ (iii): Ai0 = ∂ fi0 is maximally monotone as the subdifferential of a proper lower semicon-
tinuous convex function. Consequently, it follows from Theorem 2.5(i)&(ii) that Ai is maximally
monotone for each 1 ≤ i ≤ N. Now, ∂ fi is monotone and gra(Ai) ⊆ gra(∂ fi) (see (21)). Conse-
quently, since Ai is maximally monotone, it follows that Ai = ∂ fi.
(iii) ⇒ (iv): Follows from Theorem 2.5(i)&(iv) since Ai = ∂ fi is maximally monotone and
Prox fi = J∂ fi = JAi.
(iv) ⇒ (v): By integrating (iv) we obtain the equality in (v) up to an additive constant. Theo-
rem 4.1 implies that equality in (v) holds on S(Γ); thus, the additive constant vanishes.
(v) ⇒ (i): By Theorem 4.1 equality in (v) holds only on S(Γ). Consequently, (v) implies that
S(Γ) = H. By employing Theorem 2.5(vi)&(i), we obtain (i).
In this case Theorem 2.5 also implies Γ is maximally c-monotone. Thus, it remains to prove (B).
By our preliminary discussion there exists a c-conjugate c-splitting tuple (h1, . . . , hN) of Γ. From
(iii) and from (21) we conclude that gra(∂ fi) = gra(Ai) ⊆ gra(∂hi) which, by maximality, implies
that ∂ fi = ∂hi for each 1 ≤ i ≤ N. Here there exists a constant tuple (ρ1, . . . , ρN) ∈ RN such
that ( f1, . . . , fN) = (h1, . . . , hN) + (ρ1, . . . , ρN). For (x1, . . . , xN) ∈ Γ the equality ∑N
i=1 fi(xi) =
i=1 hi(xi) implies that ∑N
∑N
i=1 ρi = 0. Consequently, the fact that for each 1 ≤ i0 ≤ N
fi0 − ρi0 = (cid:16)
N
Mi6=i0
( fi − ρi)(cid:17)c
implies that ( f1, . . . , fN) is a c-conjugate tuple.
(cid:4)
We now provide a smoothness criteria in the 3-marginal case where Theorem 4.3(i) -- (v)&(B) are
equivalent and imply maximal c-monotonicity. To this end we will employ the following facts.
Fact 4.4 ([6, Theorem 14.19]) Let g : H → ]−∞, +∞] be proper, let h : H → ]−∞, +∞] be proper,
lower semicontinuous and convex. Set
Then
f : H → [−∞, +∞] : x 7→ (g(x) − h(x),
+∞,
x ∈ dom(g);
x /∈ dom(g).
f ∗(y) = sup
v∈dom(h∗)(cid:0)g∗(y + v) − h∗(v)(cid:1).
Fact 4.5 ([29, Corollary 2.3]) Let f : Rn → R be proper and lower semicontinuous. If f ∗ is essentially
smooth, then f is convex.
15
Theorem 4.6 Let n ∈ N, N = 3 and H = Rn. Let g : X2 → ]−∞, +∞] and h : X3 → ]−∞, +∞] be
proper, lower semicontinuous and convex functions. Suppose that f = (g ⊕ h)c (in particular if ( f , g, h)
is a c-conjugate triple) and that f is essentially smooth. Let Γ be the c-splitting set generated by ( f , g, h).
Then assertions (i) -- (v) of Theorem 4.3 hold and Γ is maximally c-monotone.
Proof. Since f = (g ⊕ h)c and dom(g + q)∗ = dom(eg∗ ) = Rn, then by employing Fact 4.4 in (29)
and then Moreau's decomposition in (30) we see that
( f + q)(x) = sup
= sup
y,z∈Rn(cid:0)c(x, y, z) − g(y) − h(z) + q(x)(cid:1)
y,z∈Rn(cid:0)hx, yi + hy, zi + hz, xi + q(x) − g(y) − h(z)(cid:1)
y∈Rn(cid:0)hx, yi + h∗(x + y) + q(x) − g(y)(cid:1)
y∈Rn(cid:0)h∗(x + y) + q(x + y) − (g(y) + q(y))(cid:1)
= (cid:0)(h∗ + q)∗ − (g + q)∗(cid:1)∗(x)
= (eh − eg∗ )∗(x) = (q − eg∗ − eh∗ )∗(x).
= sup
= sup
Since f + q is essentially smooth, Fact 4.5 implies that q − eg∗ − eh∗ is convex. Consequently,
e f ∗ = ( f + q)∗ = (q − eg∗ − eh∗ )∗∗ = q − eg∗ − eh∗,
that is, e f ∗ + eg∗ + eh∗ = q.
(29)
(30)
(cid:4)
Remark 4.7 In our discussion in the last paragraph of Section 2 we pointed out that in the par-
tition of the identity in Theorem 2.5(iv) any partial sum of the firmly nonexpansive mappings
is again firmly nonexpansive and, furthermore, that general partitions of the identity into firmly
nonexpansive mappings partial sums may fail to be firmly nonexpansive. Thus, in the context of
c-splitting sets a natural question is: Given a partition of the identity into proximal mappings, are
partial sums also proximal mappings? Unlike general firmly nonexpansive mappings, a positive
answer to this question is provided by [4, Theorem 4.2].
5 Examples, observations and remarks
We now apply our results in order to determine maximality of c-monotone sets. Given a multi-
marginal c-cyclically monotone set Γ ⊆ X, the problem of constructing a c-splitting tuple is, in gen-
eral, nontrivial. Nevertheless, constructions which are independent of maximality and uniqueness
considerations are available for some classes of c-cyclically monotone sets (for example, see [5] for
the case N ≥ 3). We also note that c-splitting tuples can be constructed via (21) if it is known, in
addition, that the antiderivatives fi are unique up to additive constants, as guaranteed by Theo-
rem 4.3. Now, suppose that a c-splitting tuple is already given. The computation and classification
16
of the c-splitting tuple as being a c-conjugate tuple were, thus far, nontrivial. We employ our re-
sults for such classifications in the following examples. For these cases, we are able to conclude
c-conjugacy without additional nontrivial computations of multi-marginal conjugates. In addi-
tion, we demonstrate finer aspects of multi-marginal maximal monotonicity.
Example 5.1 For each 1 ≤ i ≤ N, set Xi = Rd and let Qi ∈ Rd×d be symmetric, positive definite,
and pairwise commuting. Set
For each 1 ≤ i ≤ M, define Mi ∈ Rd×d by
Γ = (cid:8)(Q1v, . . . , QNv)(cid:12)(cid:12) v ∈ Rd(cid:9).
Mi = (cid:16) ∑
k6=i
Qk(cid:17)Q−1
i
.
In [5, Example 3.4], it was established that
∀(x1, . . . , xN) ∈ X,
c(x1, . . . , xN) = ∑
1≤i<j≤Nhxi, xji ≤ ∑
1≤i≤N
qMi(xi),
where qMi(x) = 1
2hx, Mixi, and equality holds if and only if (x1, . . . , xN) ∈ Γ.
Thus, we conclude that Γ is the c-splitting set generated by the tuple (qM1, . . . , qMN ), and that
Ai = Mi = ∇qMi for each 1 ≤ i ≤ N. Consequently, Theorem 4.3 implies that (qM1, . . . , qMN ) is a
c-conjugate c-splitting tuple of Γ, and that Γ is maximally c-monotone.
The maximal c-monotonicity of Γ is also implied by Theorem 3.1 via continuity of a parametriza-
tion, say,
Γ = (cid:8)(v, Q2Q−1
1 v . . . , QNQ−1
1 v)(cid:12)(cid:12) v ∈ Rd(cid:9).
As a simple application of Example 5.1, we now generalize the well-known classical fact that the
only conjugate pair of the form ( f , f ) is ( f , f ) = (q, q) and that in this case the generated splitting
set is the graph of the identity mapping.
Corollary 5.2 (self c-conjugate tuple) The only c-conjugate tuple of the form ( f , . . . , f ) is
In this case, the generated c-splitting set is Γ = ∆.
( f , . . . , f ) = (N − 1)(q, . . . , q).
In the settings of Example 5.1 we let Qi = Id for each 1 ≤ i ≤ N. Then Γ = ∆ and
Proof.
qMi = (N − 1)q for each 1 ≤ i ≤ N. We conclude that (N − 1)(q, . . . , q) is a c-conjugate c-splitting
tuple and generates the c-splitting set ∆. We now prove that it is the only c-conjugate tuple of this
form. Let ( f , . . . , f ) be a c-conjugate tuple. Then for 1 ≤ i0 ≤ N and for xi0 ∈ Xi0,
f (xi0 ) = sup
i6=i0, xi∈H(cid:16)c(x1, . . . , xi0, . . . , xN) − ∑
i6=i0
f (xi)(cid:17).
(31)
17
By letting xi = xi0 for every i in the supremum in (31) we see that
f (xi0 ) ≥ c(xi0, . . . , xi0) − (N − 1) f (xi0 ) ⇒ N f ≥ N(N − 1)q ⇒ f ≥ (N − 1)q.
Consequently,
f = (cid:16)Mi6=i0
f(cid:17)c
≤ (cid:16)Mi6=i0
q(cid:17)c
= (N − 1)q.
(cid:4)
A similar type of construction to the one of Example 5.1, however, a nonlinear one, is available
when the marginals are one-dimensional.
Example 5.3 For each 1 ≤ i ≤ N, let αi : R → R be a continuous, strictly increasing and surjective
function with αi(0) = 0. Let Γ be the curve in RN defined by
and for each 1 ≤ i ≤ N, let
In [5, Example 4.3], it was established that
Γ = n(cid:0)α1(t), . . . , αN(t)(cid:1)(cid:12)(cid:12)(cid:12)
fi(xi) = Z xi
αk(cid:0)α−1
0 (cid:18) ∑
k6=i
t ∈ Ro
(t)(cid:1)(cid:19)dt.
i
(32)
∑
N
∑
i=1Z xi
1≤i<j≤N
xixj ≤
0 (cid:18) ∑
(t)(cid:1)(cid:19)dt
and that equality in (33) holds if and only if xj = αj(cid:0)α−1
(xi)(cid:1) for every 1 ≤ i < j ≤ N, namely, if
(x1, . . . , xN) ∈ Γ. We now conclude that Γ is the c-splitting set generated by the tuple ( f1, . . . , fN)
and that for each 1 ≤ i ≤ N,
∀(x1, . . . , xN) ∈ RN
αk(cid:0)α−1
(33)
k6=i
i
i
Ai = ∇ fi = ∑
k6=i
αk ◦ α−1
i
.
Consequently, Theorem 4.3 implies that ( f1, . . . , fn) is a c-conjugate c-splitting tuple of the maxi-
mally c-monotone curve Γ. Similar to Example 5.1, the maximal c-monotonicity of Γ can also be
deduced via continuity.
A linear example of a different type, where none of the two marginal projections of Γ is mono-
tone, but where, however, Γ is c-cyclically monotone, is available for N = 3 and 2-dimensional
marginals.
Example 5.4 Suppose that N = 3 and that X1 = X2 = X3 = R2. We set
M1 = 2(cid:18)1 0
0 0(cid:19) , M2 = 2(cid:18)1 0
0 1(cid:19) , M3 =
1
7(cid:18)8 3
3 2(cid:19)
18
and
Set
∆2 = (cid:8)(a, a)(cid:12)(cid:12) a ∈ R(cid:9) ⊆ R2.
f2 = ι∆2 + qM2 = ι∆2 + 2q,
and
f3 = qM3.
f1 = ιR×{0} + qM1,
Furthermore, set v1 = (cid:0)(0, 0), (−1, −1), (1, −5)(cid:1), v2 = (cid:0)(1, 0), (2, 2), (0, 7)(cid:1) and
Γ = span{v1, v2} = n(cid:0)(s, 0), (2s − t, 2s − t), (t, 7s − 5t)(cid:1)(cid:12)(cid:12)(cid:12)
s, t ∈ Ro.
It was established in [5, Example 3.5] that
(x1, x2, x3) ∈ (cid:0)R2(cid:1)3
hx1, x2i + hx2, x3i + hx3, x1i ≤ f1(x1) + f2(x2) + f3(x3)
for all
with equality if and only if (x1, x2, x3) ∈ Γ, namely, Γ is the c-splitting set generated by the tuple
( f1, f2, f3) and that none of the two marginal projections Γ1,2, Γ1,3 and Γ2,3 of Γ, is monotone.
We observe that the matrix representation of the mapping
(t, 7s − 5t) 7→ (s, 0) + (2s − t, 2s − t)
s, t ∈ R
is M3. Consequently, we see that A3 = M3 = ∇ f3. Thus, by employing Theorem 4.3 we conclude
that ( f1, f2, f3) is a c-conjugate c-splitting tuple of the maximally c-monotone subspace Γ of(cid:0)R2(cid:1)3
.
In all of our examples thus far, the set Γ was a maximally c-monotone c-splitting set. We now
present maximally c-monotone sets which are not c-splitting sets. To this end, we note the fol-
lowing simple fact: Suppose that the set Γ ⊆ X is n-c-monotone, then for each 1 ≤ i0 ≤ N the
mapping Ai0 : H ⇒ H is n-monotone. Indeed, let Γ be n-c-monotone and assume, without the loss
1, . . . , x1
of generality, that i0 = 1. Let (x1
N) ∈ Γ and σ ∈ Sn. Then a straightforward
computation implies that the inequality
N), . . . , (xn
1 , . . . , xn
n
∑
j=1
leads to the inequality
c(xj
1, xσ(j)
2
, . . . , xσ(j)
N ) ≤
n
∑
j=1
c(xj
1, . . . , xj
N)
n
∑
j=1(cid:28)xj
1,
N
∑
i=2
xσ(j)
i (cid:29) ≤
n
∑
j=1(cid:28)xj
1,
N
∑
i=2
xj
i(cid:29).
Thus, we see that if Γ is n-c-monotone, then A1 is n-monotone. To sum up,
if for some 1 ≤ i0 ≤ N the mapping Ai0 is not cyclically monotone, then the set Γ is not
a c-splitting set.
19
Indeed, otherwise, Γ would have been c-cyclically monotone (as we recollected after Defini-
tion 1.2) and, by the above argument, for all 1 ≤ i0 ≤ N the mapping Aio would have been
cyclically monotone.
We now address a trivial embedding of all classical maximally monotone operators in the multi-
marginal framework. In particular, we obtain maximally c-monotone mappings which are not
c-cyclically monotone.
Example 5.5 Let A : H ⇒ H be a maximally monotone mapping. We set Γ ⊆ X by
Γ = (cid:8)(x1, x2, 0, . . . , 0) x2 ∈ Ax1(cid:9).
Then Γ is c-monotone and we see that A1 = A is maximally monotone. Consequently, by invoking
Theorem 2.5 (ii) we conclude that Γ is maximally c-monotone. In addition, we see that A is n-
monotone if and only if Γ is n-c-monotone. Therefore, if A is not n-monotone for some n ≥ 3, then
Γ is not n-c-monotone. Furthermore, since the n-c-monotonicity of a set is invariant under shifts,
the set Γ = (cid:8)(x1, x2, ρ3, . . . , ρN) x2 ∈ Ax1(cid:9) is also maximally monotone for any constant vectors
ρ3, . . . , ρN ∈ H.
Our next example of a maximally c-monotone set which is not a c-splitting set does not follow
from an embedding of the type in Example 5.5.
Example 5.6 Set N = 3 and for each 1 ≤ i ≤ 3 set Xi = R2. Let Rθ denote the counterclockwise
rotation by the angle θ in R2. Let the set Γ ⊆ X = (cid:0)R2(cid:1)3
be defined by
(34)
Γ = n(cid:16)x,
√3
2 R−π/2x,
√3
x ∈ R2o.
2 R−π/2x(cid:17)(cid:12)(cid:12)(cid:12)
x ∈ R2o =⇒ A1 = √3R−π/2.
It follows that
Since Γ = n(cid:16) 2√3
gra A1 = n(x, √3R−π/2x)(cid:12)(cid:12)(cid:12)
Rπ/2x, x, x(cid:17)(cid:12)(cid:12)(cid:12)
x ∈ R2o, we have
2
Rπ/2x(cid:17)(cid:12)(cid:12)(cid:12)
gra A2 = gra A3 = n(cid:16)x, x +
√3
x ∈ R2o =⇒ A2 = A3 = r 7
3
Rarctan(2/√3).
We see that A1, A2, and A3 are maximally monotone. Consequently, for each ∅ 6= K ( {1, 2, 3},
the mapping AK is maximally monotone and it now follows from Theorem 2.5 that Γ is maximally
c-monotone in X. Furthermore, since A1 is not 3-c-cyclically monotone, it is not c-cyclically mono-
tone and, consequently, Γ is not a c-splitting set. By a straightforward computation, it follows
that
JA1 =
1
2
Rπ/3,
JA2 = JA3 =
√3
4
R−π/6
and
JA1 + JA2 + JA3 = Id .
Finally, from (34) it is easy to see that Γi,j is monotone for all 1 ≤ i < j ≤ 3.
20
We see that in the case N = 3 the set Γ is c-monotone if and only if the mappings A1, A2 and A3
are monotone. In the following example we demonstrate that the monotonicity of all of the Ai's
no longer implies the c-monotonicity of Γ in the case when N ≥ 4.
Example 5.7 In [4, Lemma 4.2 and Example 4.3] it was established that: In X = R2, let n ∈
counterclockwise rotator by θ. Then the following hold:
{2, 3, . . .}, let θ ∈ (cid:3)arccos(1/√2), arccos(1/√2n)(cid:3), set α = 1/(2n cos(θ)), and denote by Rθ the
(i) αRθ and αR−θ are firmly nonexpansive.
(ii) nαRθ and nαR−θ are not firmly nonexpansive.
(iii) nαRθ + nαR−θ = Id.
We employ these facts to construct a set Γ as follows. We set N = 2n and
Ti = (αRθ,
1 ≤ i ≤ n;
αR−θ, n + 1 ≤ i ≤ 2n.
Define
It then follows that for each 1 ≤ i ≤ N, the mapping JAi = Ti is firmly nonexpansive with full
domain. We conclude that the set Γ possesses the following properties:
Γ = (cid:8)(T1x, . . . , T2nx) x ∈ R2(cid:9) ⊆ X = (R2)
2n
.
(iv) for each 1 ≤ i ≤ N, the mapping Ai is maximally monotone,
(v) JA1 + · · · + JAN = Id.
However, due to (ii), the mappings
JA{1,...,n}
=
n
∑
i=1
JAi =
n
∑
i=1
Ti = nαRθ,
and similarly JA{n+1,...,2n}
= nαR−θ
are not firmly nonexpansive, equivalently, A{1,...,n} = R−2θ and A{n+1,...,2n} = R2θ are not mono-
tone. Consequently, by employing Lemma 2.3 we conclude that despite the fact that Γ possesses
properties (iv) and (v), it is not a c-monotone set.
Remark 5.8 In [5] the two marginal projections Γi,j of a set Γ ⊆ X were employed, it was estab-
lished that if the Γi,j's are cyclically monotone, then Γ is c-cyclically monotone and an explicit
construction of a c-splitting tuple is provided. However, it was also established that this is a
sufficient condition for c-cyclic monotonicity of Γ but not a necessary one, in general, as can be
seen in Example 5.4 where we provide a maximally c-cyclically monotone set such that all of its
two-marginal projections are not monotone. In the one dimensional case (i.e., Xi = R for each
1 ≤ i ≤ N), it was established that Γ is c-monotone if and only if all of its two marginal projections
21
Γi,j are monotone. With the exception of Example 5.4, in all of our examples of c-monotone sets
in this section the set Γ had monotone two-marginal projections Γi,j. Thus, a natural question is:
How does the monotonicity and maximal monotonicity of the two-marginal projections Γi,j relate to the
c-monotonicity and maximal c-monotonicity of Γ?
Proposition 5.9 Let Xi = Rd for each 1 ≤ i ≤ N. Let Γ ⊆ X be a set. Suppose that for each 1 ≤ i < j ≤
N the set Γi,j is monotone. Then Γ is c-monotone.
Proof. The mapping AK is monotone if and only if for every (x1, . . . , xN), (y1, . . . , yN) ∈ Γ,
0 ≤ *∑
xi − yi, ∑
j6∈K
xj − yj+ .
Since the right-hand side is equal to ∑ i∈K
0 ≤ hxi − yi, xj − yji, we see that AK is monotone.
i∈K
j6∈Khxi − yi, xj − yji and since, by the monotonicity of Γi,j,
(cid:4)
To the best of our knowledge, the question whether the maximal monotonicity of the Γi,j's im-
plies the maximal c-monotonicity of Γ is still open.
Finally, we note that the maximal c-monotonicity of Γ does not imply the maximal monotonic-
ity of the Γi,j's even when the Γi,j's are monotone. Indeed, in Example 5.5, we see that although Γ
is maximally c-monotone, Γi,j is a singleton for all 3 ≤ i < j ≤ N, thus Γi,j is monotone but not
maximally monotone. Even in the case N = 3, Γ1,3 is a proper subset of the graph of the zero map-
ping whenever Γ is generated by a maximally monotone mapping A without a full domain. We
conclude in this case that Γ is maximally c-monotone, however, Γ1,3 is not maximally monotone.
Acknowledgments
We thank three anonymous referees for their kind and useful remarks. Sedi Bartz was partially
supported by a University of Massachusetts Lowell startup grant. Heinz Bauschke and Xianfu
Wang were partially supported by the Natural Sciences and Engineering Research Council of
Canada. Hung Phan was partially supported by Autodesk, Inc.
References
[1] M. Agueh and G. Carlier, Barycenters in the Wasserstein Space, SIAM Journal on Mathematical
Analysis 43 (2011), 904 -- 924.
[2] S. Bartz and S. Reich, Abstract convex optimal antiderivatives, Annales de l'Institut Henri
Poincare (C) Non Linear Analysis 29 (2012), 435 -- 454.
22
[3] S. Bartz and S. Reich, Optimal pricing for optimal transport, Set-Valued and Variational Analy-
sis 22 (2014), 467 -- 481.
[4] S. Bartz, H.H. Bauschke and X. Wang, The resolvent order: a unification of the orders by
Zarantonello, by Loewner, and by Moreau, SIAM Journal on Optimization 27 (2017), 466 -- 477.
[5] S. Bartz, H.H. Bauschke and X. Wang, A class of multi-marginal c-cyclically monotone sets
with explicit c-splitting potentials, Journal of Mathematical Analysis and Applications 461 (2018),
333 -- 348.
[6] H.H. Bauschke and P.L. Combettes, Convex Analysis and Monotone Operator Theory in Hilbert
Spaces, 2nd edition, Springer, 2017.
[7] M. Beiglb ock and C. Griessler, An optimality principle with applications in optimal transport,
arXiv preprint, arXiv:1404.7054 (2014).
[8] H. Brezis, Liquid crystals and energy estimates for S2-valued maps, Theory and Applications of
Liquid Crystals (Minneapolis, Minn., 1985), The IMA Volumes in Mathematics and its Applica-
tions Volume 5, Springer, (1987), 31 -- 52.
[9] G. Carlier, On a class of multidimensional optimal transportation problems, Journal of Convex
Analysis 10 (2003), 517 -- 529.
[10] G. Carlier and B. Nazaret, Optimal transportation for the determinant, ESAIM: Control, Opti-
misation and Calculus of Variations 14 (2008), 678 -- 698.
[11] W. Gangbo and R. McCann, The geometry of optimal transportation, Acta Mathematica 177
(1996), 113 -- 161.
[12] W. Gangbo and A. Swiech, Optimal maps for the multidimensional Monge-Kantorovich
problem, Communications on Pure and Applied Mathematics 51 (1998), 23 -- 45.
[13] N. Ghoussoub and B. Maurey, Remarks on multi-marginal symmetric Monge-Kantorovich
problems, Discrete and Continuous Dynamical Systems 34 (2013), 1465 -- 1480.
[14] N. Ghoussoub and A. Moameni, Symmetric Monge-Kantorovich problems and polar decom-
positions of vector fields, Geometric and Functional Analysis 24 (2014), 1129 -- 1166.
[15] C. Griessler, c-cyclical monotonicity as a sufficient criterion for optimality in the multi-
marginal Monge-Kantorovich problem, Proceedings of the American Mathematical Society 146
(2018), 4735 -- 4740.
[16] H.G. Kellerer, Duality theorems for marginal problems, Zeitschrift f ur Wahrscheinlichkeitstheo-
rie und Verwandte Gebiete 67 (1984), 399 -- 432.
[17] Y.-H. Kim and B. Pass, A general condition for Monge solutions in the multi-marginal optimal
transport problem, SIAM Journal on Mathematical Analysis 46 (2014), 1538 -- 1550.
[18] M. Knott and C.S. Smith, On a generalization of cyclic monotonicity and distances among
random vectors, Linear Algebra and its Applications 199 (1994), 363 -- 371.
23
[19] S. Di Marino, L. De Pascale and M. Colombo, Multimarginal optimal transport maps for 1-
dimensional repulsive costs, Canadian Journal of Mathematics 67 (2015), 350 -- 368.
[20] S. Di Marino, A. Gerolin and L. Nenna, Optimal transportation theory with repulsive costs,
Topological Optimization and Optimal Transport: In the Applied Sciences 9 (2017), 204 -- 256.
[21] B. Pass, On the local structure of optimal measures in the multi-marginal optimal transporta-
tion problem, Calculus of Variations and Partial Differential Equations 43 (2012), 529 -- 536.
[22] B. Pass, Multi-marginal optimal transport: theory and applications, ESAIM: Mathematical
Modelling and Numerical Analysis 49 (2015), 1771 -- 1790.
[23] J.-C. Rochet, A necessary and sufficient condition for rationalizability in a quasilinear context,
Journal of Mathematical Economics 16 (1987), 191 -- 200.
[24] R.T. Rockafellar, Characterization of the subdifferentials of convex functions, Pacific Journal of
Mathematics 17 (1966), 497 -- 510.
[25] R.T. Rockafellar and R.J-B. Wets, Variational Analysis, Springer, 1998.
[26] L. R uschendorf, On c-optimal random variables, Statistics and Probability Letters 27 (1996),
267 -- 270.
[27] L. R uschendorf and L. Uckelmann, On Optimal multivariate couplings, chapter in Distribu-
tions with given Marginals and Moment Problems, Springer (1997), 261 -- 273.
[28] F. Santambrogio, Optimal Transport for Applied Mathematicians, Birkhauser, 2015.
[29] V. Soloviov, Duality for nonconvex optimization and its applications, Analysis Mathematica 19
(1993), 297 -- 315.
[30] C. Villani, Optimal Transport: Old and New, Springer, 2009.
24
|
1608.07516 | 2 | 1608 | 2017-03-02T09:10:41 | Local characterizations for the matrix monotonicity and convexity of fixed order | [
"math.FA"
] | We establish local characterizations of matrix monotonicity and convexity of fixed order by giving integral representations connecting the Loewner and Kraus matrices, previously known to characterize these properties, to respective Hankel matrices. Our results are new already in the general case of matrix convexity and our approach significantly simplifies the corresponding work on matrix monotonicity. We also obtain an extension of the original characterization for matrix convexity by Kraus, and tighten the relationship between monotonicity and convexity. | math.FA | math |
LOCAL CHARACTERIZATIONS FOR THE MATRIX MONOTONICITY AND
CONVEXITY OF FIXED ORDER
OTTE HEIN AVAARA
Abstract. We establish local characterizations of matrix monotonicity and convexity of fixed
order by giving integral representations connecting the Loewner and Kraus matrices, previously
known to characterize these properties, to respective Hankel matrices. Our results are new
already in the general case of matrix convexity and our approach significantly simplifies the
corresponding work on matrix monotonicity. We also obtain an extension of the original char-
acterization for matrix convexity by Kraus, and tighten the relationship between monotonicity
and convexity.
1. Introduction
For an open interval (a, b), we say that f : (a, b) → R is matrix monotone (increasing) of
order n (or n-monotone) if for any n × n Hermitian matrices A, B with spectra in (a, b) and
A ≤ B we have f (A) ≤ f (B).1 Analogously, f : (a, b) → R is matrix convex of order n (or
n-convex) if for any n × n Hermitian matrices A, B with spectra in (a, b) and λ ∈ [0, 1] we have
f (λA + (1 − λ)B) ≤ λf (A) + (1 − λ)f (B).
Ever since Charles Loewner (then known as Karl Lowner) introduced matrix monotone functions
in 1934 [12], this class has been characterized in various ways. See for example [2, 8] for survey and
recent progress. The famous theorem established in the Loewner's paper states that a function
that is matrix monotone of all orders on an interval, extends to upper half-plane as a Pick-
Nevanlinna function: an analytic function with non-negative imaginary part. Loewner's proof of
this jewel is based on an important characterization in terms of divided differences here denoted by
[·, ·, . . . , ·]f . Recall that divided differences are defined recursively by [λ]f = f (λ) and for distinct
λ1, λ2, . . . , λn ∈ (a, b),
[λ1, λ2, . . . , λn]f =
[λ1, λ2, . . . , λn−1]f − [λ2, λ3, . . . , λn]f
λ1 − λn
.
If f ∈ C n−1(a, b), divided difference has continuous extension to all tuples of not necessarily
distinct n numbers on the interval [4].
Theorem 1 (Loewner). A function f : (a, b) → R is n-monotone (for n ≥ 2) if and only if
f ∈ C1(a, b) and the Loewner matrix
(2)
L = ([λi, λj]f )1≤i,j≤n
is positive2 for any tuple of numbers (λi)n
i=1 on the same interval.
2010 Mathematics Subject Classification. Primary 26A48; Secondary 26A51, 47A63.
Key words and phrases. Matrix monotone functions, Matrix convex functions.
1As usual, the space of Hermitian matrices is equipped with the Loewner order, i.e. the partial order induced
by the convex cone of positive semi-definite matrices.
2Here and in the following, positivity of matrix means that it is positive semi-definite.
1
2
OTTE HEIN AVAARA
Similarly Kraus, a student of Loewner introduced the matrix convexity in [11] and established
similar characterization:
Theorem 3 (Kraus). A function f : (a, b) → R is n-convex (for n ≥ 2) if and only if f ∈ C2(a, b)
and the Kraus matrix
(4)
Kr = ([λi, λj , λ0]f )1≤i,j≤n
is positive for any tuple of numbers (λi)n
i=1 ∈ (a, b)n and λ0 ∈ (λi)n
i=1.
A different, local characterization for monotonicity was given by another student of Loewner,
Dobsch in [5]:
Theorem 5 (Dobsch, Donoghue). A C2n−1 function f : (a, b) → R is n-monotone if and only
if the Hankel matrix
(6)
is positive for any t ∈ (a, b).
M (t) = (cid:18) f (i+j−1)(t)
(i + j − 1)!(cid:19)1≤i,j≤n
By employing standard regularization techniques, one could further extend this to merely C2n−3
functions with convex derivative of order (2n−3), a class of functions for which the property makes
sense for almost every t, to obtain the complete local characterization of the matrix monotonicity
of fixed order. The result has a striking consequence: n-monotonicity is a local property, meaning
that if function has it in two overlapping intervals, it has it for their union. This property is
actually used in the proof, and although it was noted by Loewner to be easy ([12, p. 212, Theorem
5.6]), no rigorous proof was given until 40 years later in the monograph of Donoghue [6], and the
proof is rather long when n > 2.
The main results of this paper establish novel integral representations connecting Hankel ma-
trices to the Loewner and Kraus matrices. These identities give rise to a new simple proof for
Theorem 5, and more importantly, settle the conjecture in [9] (see also [10]) by establishing similar
local characterization for the matrix convex functions.
Theorem 7. A C2n function f : (a, b) → R is n-convex if and only if the Hankel matrix
(8)
is positive for any t ∈ (a, b).
K(t) = (cid:18) f (i+j)(t)
(i + j)! (cid:19)1≤i,j≤n
Again, with regularizations we may extend this to give a complete local description of matrix
convexity of fixed order, which as an immediate corollary gives the expected local property theorem
for convexity.
Corollary 9. For any positive integer n, n-convexity is a local property.
As another byproduct, we obtain a slight improvement to Theorem 3, where λ0 may now vary
freely. This also implies through divided differences a rather direct connection between matrix
monotonicity and convexity.
LOCAL CHARACTERIZATIONS FOR THE MATRIX MONOTONICITY AND CONVEXITY
3
2. Matrix monotone functions
2.1. Integral representation. In this section we construct the integral representations for the
Loewner matrices alluded to in the introduction.
Let n ≥ 2, (a, b) be an interval, and Λ = (λi)n
i=1 ∈ (a, b)n be an arbitrary sequence of distinct
points in (a, b).
In the following the Loewner and respective Hankel matrices, introduced in the introduction
in (2) and (6), for sufficiently smooth f : (a, b) → R and λ0 ∈ (a, b) are denoted by L(Λ, f ) and
Mn(t, f ) respectively.
Recall that as one easily verifies with Cauchy's integral formula and induction, the divided
differences can be written as
(10)
[λ1, λ2, . . . , λn]f =
for analytic f and suitable closed curve γ.3
1
2πi Zγ
f (z)
(z − λ1) · · · (z − λn)
dz,
Divided differences also admit a natural generalization for the mean value theorem [4]. Namely,
for an open interval (a, b), f ∈ C n−1(a, b) and any tuple of (not necessarily distinct) real numbers
Λ = (λi)n
i=1 ∈ (a, b)n we have
(11)
for some ξ ∈ [min(Λ), max(Λ)].
[λ1, λ2, . . . , λn]f =
f (n−1)(ξ)
(n − 1)!
We shall also need the very basic properties of regularizations. Namely for even, non-negative
and smooth function φ supported on [−1, 1] and with integral 1, and integrable f : (a, b) → R,
regularization (or ε-regularization, to be precise) of f , denoted by fε : (a + ε, b − ε) → R is the
convolution
fε(x) = Z ∞
−∞
f (x − εy)φ(y)dy.
This is a smooth function, and for any continuity point x ∈ (a, b) of f we clearly have limε→0 fε(x) =
f (x). Note that regularizations of matrix monotone (convex) functions are obviously matrix mono-
tone (convex) functions on a slightly smaller interval.
Define the functions gj for 1 ≤ j ≤ n by
(12)
gj,Λ(t, y) = Yk6=j
(1 + y(t − λk)).
Define also the matrix C(t) := C(t, Λ) by setting Ci,j to be the coefficient of yi−1 in the polynomial
gj(t, y), i.e. we have
(13)
gj(t, y) = C1,j(t) + C2,j(t)y + . . . + Cn,j(t)yn−1.
Define polynomial pΛ with pΛ(t) := Qn
setting hz(x) = (z − x)−1.
i=1(t − λi). Also for any z ∈ C define function hz by
Lemma 14. For Λ = (λi)n
i=1 as before, t ∈ R, and z ∈ C distinct from t, we have
C T (t, Λ) Mn (t, hz) C (t, Λ) = L(Λ, hz)
pΛ(z)2
(z − t)2n .
3For our purposes, it is enough to consider f analytic in an open half-plane and γ a circle in this half-plane
enclosing the points λ1, λ2, . . . , λn.
4
OTTE HEIN AVAARA
Proof. Write D = C T (t, Λ)Mn(t, hz)C(t, Λ). Note that as we have h(k)
may write Mn(t, hz) = 1
(z−t)2 vvT with v = (1, 1
(z−t)2 , . . . ,
z−t ,
1
1
(z−t)n−1 )T . Thus
z (t)/k! = (z − t)−k−1, we
D =
1
(z − t)2 (C(t, Λ)T v)(C(t, Λ)T v)T .
One also easily sees that (C(t, Λ)T v)i = gi(t, 1
Di,j =
gi(t, 1
z−t )gj(t, 1
(z − t)2
z−t )
=
1
(z − t)2 Yk6=i
z−t ) so that finally
(cid:18)1 +
z − t (cid:19) Yk6=j
t − λk
(cid:18)1 +
t − λk
z − t (cid:19) = [λi, λj ]hz
pΛ(z)2
(z − t)2n .
(cid:3)
Consider now the function
S(z, t) := SΛ(z, t) := −
(z − t)2n−2
pΛ(z)2
.
As S(z, t) decays as z−2, with the residue theorem we see that for suitable closed curve γ we have
0 =
1
2πi Zγ
S(z, t)dz =
n
Xi=1
Res
z=λi
S(z, t).
Defining now the weight functions Ii := Ii,Λ for 1 ≤ i ≤ n by
and
Ii(t) = Res
z=λi
S(z, t),
I(t) := IΛ(t) := X1≤i≤n
λi<t
Ii(t),
we see by simple computation that Ii's are polynomials such that Ii(λi) = 0 and I is hence
piecewise polynomial, continuous function supported on [min(Λ), max(Λ)].
Note that with Cauchy's integral formula we can also write I in the form
I(t) =
whenever t /∈ Λ.
1
2πi Z t+i∞
t−i∞
S(z, t)dz,
Remark 15. The weight function I and the analogous weight J to be introduced in the convex
setting are examples of weights called Peano kernels or B-splines. The properties of these kernels
are discussed for example in [3]. To stay self-contained, we give proofs of the crucial properties
used in our discussion.
Lemma 16. For Λ = (λi)n
i=1 as before and z ∈ C outside the interval [min(Λ), max(Λ)], we have
(2n − 1)Z ∞
−∞
I(t)
(z − t)2n dt =
1
pΛ(z)2 .
LOCAL CHARACTERIZATIONS FOR THE MATRIX MONOTONICITY AND CONVEXITY
5
Proof. We simply compute that
(2n − 1)Z ∞
−∞
I(t)
(z − t)2n dt = (2n − 1)
= −(2n − 1)
(z − t)2n dt
n
λi
Ii(t)
Z ∞
w=λiZ ∞
Res
Xi=1
Xi=1
λi
n
(w − t)2n−2
pΛ(w)2(z − t)2n dt
=
= −
n
Xi=1
Xi=1
n
(1 − z−w
z−λi
)2n−1 − 1
(w − z)pΛ(w)2
Res
w=λi
Res
w=λi
1
(w − z)pΛ(w)2
= Res
w=z
1
(w − z)pΛ(w)2
−
1
2πi Zγ
dw
(w − z)pΛ(w)2
=
1
pΛ(z)2 ,
where we used the residue theorem for the function (w 7→ (w − z)−1pΛ(w)−2).
(cid:3)
We are then ready to formulate and prove the integral representation of the Loewner matrix.
Theorem 17. For f ∈ C2n−1(a, b) and Λ as before, we have
L(Λ, f ) = (2n − 1)Z ∞
−∞
C T (t, Λ)Mn(t, f )C(t, Λ)IΛ(t)dt.
Proof. For entire f , by Lemmas 14, 16, Fubini and (10) we have
(2n − 1)Z ∞
−∞
C T (t)Mn(t, f )C(t)I(t)dt =
=
=
1
1
−∞
2πi Zγ (cid:18)(2n − 1)Z ∞
2πi Zγ
2πi Zγ
L(Λ, hz)f (z)dz
1
L(Λ, hz)(cid:18)(2n − 1)Z ∞
−∞
C T (t)Mn(t, hz)C(t)I(t)dt(cid:19) f (z)dz
pΛ(z)2
(z − t)2n I(t)dt(cid:19) f (z)dz
= L(Λ, f ).
The general case now follows by uniformly approximating f and its derivatives up to order (2n −
1) by entire functions on [min(Λ), max(Λ)], say, by polynomials with a suitable application of
Weierstrass approximation theorem.
(cid:3)
2.2. Positivity of the weight. In this section we prove the non-negativity of the weight function
I introduced in the previous section. We begin with a simple lemma.
Lemma 18. Let n be a positive integer and numbers Z = (ζi)n
Qn
i=1(ζi − t)−1, then for any non-negative integer k and t < 0 we have
f (k)(t) ≥ 0.
i=1 non-negative. Now if f (t) =
Proof. The case of n = 1 is trivial; the general case follows now immediately from the product
rule.
(cid:3)
6
OTTE HEIN AVAARA
Lemma 19. For Λ as before, IΛ is non-negative.
Proof. We may clearly assume that Λ is strictly increasing. When checking the non-negativity at
a point t, we may without loss of generality assume that t = 0 ∈ [λ1, λn]. Also by continuity we
may further assume that all the λi's are non-zero. We are left to investigate
1
2πi Z i∞
−i∞
S(z, 0)dz = −
1
2πi Z i∞
−i∞
z2n−2dz
pΛ(z)2 .
Making the change of variable w = 1
z , we are to check that
1
2πi Z i∞
−i∞
dw
pZ(w)2
≥ 0,
Let k (< n) be the number of the negative ζi's and denote Z− = (ζi)k
i=1. Note that if we further
where Z = 1
Λ , that is ζi = 1
write f (t) = (cid:0)Qi>k(t − ζi)(cid:1)−2
λi
.
1
2πi Z i∞
−i∞
dw
pZ(w)2 =
, we have by suitable variant of (10)
1
2πi Z i∞
−i∞
f (w)dw
pZ−(w)2 = [ζ1, ζ1, ζ2, ζ2, . . . , ζk, ζk]f ,
which is positive in the view of (11) and Lemma 18.
(cid:3)
2.3. Characterizations for the matrix monotonicity.
Proof of Theorem 5. The necessity of the condition can be found in [5]. For sufficiency note that
by Theorem 17 we can write
L(Λ) = (2n − 1)Z ∞
−∞
C T (t)M (t)C(t)I(t)dt
Now if M (t) ≥ 0 for any t ∈ (a, b), also C T (t)M (t)C(t) ≥ 0 for any t ∈ (a, b). It follows from
Lemma 19 that the integrand is a positive matrix, so indeed, L is positive as an integral of positive
matrices. But now f is n-monotone by Theorem 1.
(cid:3)
Putting everything together we obtain complete characterizations of the class of n-monotone
functions.
Theorem 20 (Loewner, Dobsch, Donoghue). Let n ≥ 2, and (a, b) be an open interval. Now
for f : (a, b) → R the following are equivalent
(i) f is n-monotone.
(ii) f ∈ C1(a, b) and the Loewner matrix L(Λ, f ) is positive for any tuple Λ ∈ (a, b)n.
(iii) f ∈ C2n−3(a, b), f (2n−3) is convex, and the Hankel matrix Mn(t, f ), which makes sense
almost everywhere, is positive for almost every t ∈ (a, b).
Proof. As noted before, (i) ⇔ (ii) was proven in the original paper of Loewner [12]. For C2n−1
functions, (i) ⇔ (iii) is Theorem 5, and for merely C2n−3 functions the claim follows from standard
regularization procedure, details of which can be found in [6]. For an alternate approach to the
latter equivalence, see again [6].
(cid:3)
Corollary 21. For any positive integer n, n-monotonicity is a local property.
LOCAL CHARACTERIZATIONS FOR THE MATRIX MONOTONICITY AND CONVEXITY
7
3. Matrix convex functions
3.1. Integral representation. In this section we construct the integral representations for the
Kraus matrices alluded to in the introduction.
Again, let n ≥ 2, (a, b) be an interval, and Λ = (λi)n
i=1 ∈ (a, b)n be an arbitrary sequence of
distinct points in (a, b).
In the following the Kraus and the respective Hankel matrices, introduced in the introduction,
for sufficiently smooth f : (a, b) → R and λ0 ∈ (a, b) are denoted by Kr(λ0, Λ, f ) and Kn(t, f ),
respectively.
The integral representation for the Kraus matrix is similar to that of the Loewner matrix. Fix
again n ≥ 2, open interval (a, b) and Λ = (λi)n
i=1 ∈ (a, b)n, an arbitrary sequence of distinct
points on (a, b). For fixed λ0 ∈ (a, b) the weights Ji,λ0 := Ji,λ0,Λ, now for 0 ≤ i ≤ n, are defined
analogously as the residues at λi's of
and
Tλ0(z, t) := Tλ0,Λ(z, t) := −
(z − t)2n−1
(z − λ0)pΛ(z)2
Jλ0 (t) := Jλ0,Λ(t) := X0≤i≤n
λi<t
Ji,λ0 (t).
Lemma 22. For Λ = (λi)n
we have
i=1, as before, λ0 ∈ (a, b) and z ∈ C outside the interval [min(Λ), max(Λ)],
2nZ ∞
−∞
Jλ0 (t)
(z − t)2n+1 dt =
1
(z − λ0)pΛ(z)2 .
Proof. Proof is almost identical to that of Lemma 16; we just perform the residue trick with the
map (w 7→ (w − z)−1(w − λ0)−1pΛ(w)−2) instead.
(cid:3)
Theorem 23. For f ∈ C2n(a, b), Λ as before, and λ0 ∈ (a, b), we have
Kr(λ0, Λ, f ) = 2nZ ∞
C T (t, Λ)Kn(t, f )C(t, Λ)Jλ0 ,Λ(t)dt.
Proof. After noting that Kn(t, hz) = 1
of Theorem 17, using Lemma 22 instead of Lemma 16.
−∞
z−t Mn(t, hz), the calculation is carried out as in the proof
(cid:3)
3.2. Positivity of the weight.
Lemma 24. For Λ = (λi)n
i=1 as before and λ0 ∈ (a, b), Jλ0,Λ is non-negative.
Proof. As in the proof of Lemma 19, we can assume that t = 0 is our point of inspection and
that Λ is strictly increasing. We also make the same change of variables Z = 1
Λ . Note that we
may well assume that ζ0 > 0, since the other case would follow by reflecting the variables, that
is considering the sequence −Z and −λ0, instead. Now the inequality is reduced to an equivalent
form
1
2πi Z i∞
−i∞
dw
(ζ0 − w)pZ (w)2
≥ 0.
But as in the proof of Lemma 19, the left hand side can be again written as
[ζ1, ζ1, ζ2, ζ2, . . . , ζk, ζk]f
where f (t) = (ζ0 − t)−1(cid:0)Qi>k(t − ζi)(cid:1)−2
and k is the number of negative ζi's.
(cid:3)
8
OTTE HEIN AVAARA
3.3. Characterizations for the matrix convexity.
Proof of Theorem 7. The necessity of the condition was proven in [9]. For the other direction, by
Lemma 23 we can write
Kr(λ, Λ) = 2nZ ∞
−∞
C T (t)K(t)C(t)Jλ0 (t)dt.
But as in the proof of Theorem 5, we see now that the Kraus matrix is an integral of positive
matrices, hence positive, and Theorem 3 finishes the claim.
(cid:3)
The next theorem finally completes the characterization of n-convex functions. The original
characterization of Kraus is also improved.
Theorem 25. Let n ≥ 2, and (a, b) be an open interval. Now for f : (a, b) → R the following are
equivalent
(i) f is n-convex.
(ii) f ∈ C2(a, b) and the Kraus matrix Kr(λ0, Λ, f ) is positive for any tuple Λ ∈ (a, b)n and
λ0 ∈ Λ.
(iii) f ∈ C2(a, b) and the Kraus matrix Kr(λ0, Λ, f ) is positive for any tuple Λ ∈ (a, b)n and
λ0 ∈ (a, b).
(iv) f ∈ C2n−2(a, b), f (2n−2) is convex, and the Hankel matrix Kn(t, f ), which makes sense
almost everywhere, is positive for almost every t ∈ (a, b).
Proof. (i) ⇔ (ii) was proven in [11]. For C2n functions (i) ⇔ (iv) is Theorem 7; the proof of
Theorem 7 also gives (iv) ⇒ (iii) in this case. For merely C2n−2 functions these claims follow
from regularization techniques as in the monotone case. (iii) ⇒ (ii) is trivial.
(cid:3)
We also get an interesting corollary connecting the monotonicity to convexity, extending a result
in [1].
Corollary 26. Let n ≥ 2, and (a, b) be an open interval. If f : (a, b) → R is n-convex, then for
any λ0 ∈ (a, b) the function g = (x 7→ [x, λ0]f ) is n-monotone.
Proof. Simply note that L(Λ, g) = Kr(λ0, Λ, f ).
(cid:3)
Remark 27. The ideas introduced in the paper can be generalized to characterize more general
class of functions called matrix k-tone functions, introduced in [7]. A paper discussing related
questions in this more general setting is in preparation.
4. Acknowledgements
We thank the open-source mathematical software Sage [13] for invaluable support in discovering
the main identities of this paper. We are also truly grateful to O. Hirviniemi, J. Junnila and E.
Saksman, and anonymous reviewers for their helpful comments on the earlier versions of the
manuscript.
References
[1] J. Bendat and S. Sherman. Monotone and convex operator functions. Trans. Amer. Math. Soc., 79:58–71, 1955.
[2] P. Chansangiam. A survey on operator monotonicity, operator convexity, and operator means. Int. J. Anal.,
pages Art. ID 649839, 8, 2015.
[3] C. de Boor. A practical guide to splines, volume 27 of Applied Mathematical Sciences. Springer-Verlag, New
York-Berlin, 1978.
LOCAL CHARACTERIZATIONS FOR THE MATRIX MONOTONICITY AND CONVEXITY
9
[4] C. de Boor. Divided differences. Surv. Approx. Theory, 1:46–69, 2005.
[5] O. Dobsch. Matrixfunktionen beschrankter Schwankung. Math. Z., 43(1):353–388, 1938.
[6] W. F. Donoghue, Jr. Monotone matrix functions and analytic continuation. Springer-Verlag, New York-
Heidelberg, 1974. Die Grundlehren der mathematischen Wissenschaften, Band 207.
[7] U. Franz, F. Hiai, and E. Ricard. Higher order extension of Lowner's theory: operator k-tone functions. Trans.
Amer. Math. Soc., 366(6):3043–3074, 2014.
[8] F. Hansen. The fast track to Lowner's theorem. Linear Algebra Appl., 438(11):4557–4571, 2013.
[9] F. Hansen and J. Tomiyama. Differential analysis of matrix convex functions. Linear Algebra Appl., 420(1):102–
116, 2007.
[10] F. Hansen and J. Tomiyama. Differential analysis of matrix convex functions. II. JIPAM. J. Inequal. Pure
Appl. Math., 10(2):Article 32, 5, 2009.
[11] F. Kraus. Uber konvexe Matrixfunktionen. Math. Z., 41(1):18–42, 1936.
[12] K. Lowner. Uber monotone Matrixfunktionen. Math. Z., 38(1):177–216, 1934.
[13] The Sage Developers. SageMath,
the Sage Mathematics Software System (Version 7.1),
2016.
http://www.sagemath.org.
University of Helsinki, Department of Mathematics and Statistics, P.O. Box 68 (Gustaf Hallstromin
katu 2b), FI-00014 University of Helsinki
E-mail address: [email protected]
|
1607.01666 | 3 | 1607 | 2016-11-10T09:48:50 | $L^p$-$L^q$ off-diagonal estimates for the Ornstein--Uhlenbeck semigroup: some positive and negative results | [
"math.FA",
"math.CA"
] | We investigate $L^p(\gamma)$-$L^q(\gamma)$ off-diagonal estimates for the Ornstein-Uhlenbeck semigroup $(e^{tL})_{t > 0}$. For sufficiently large $t$ (quantified in terms of $p$ and $q$) these estimates hold in an unrestricted sense, while for sufficiently small $t$ they fail when restricted to maximal admissible balls and sufficiently small annuli. Our counterexample uses Mehler kernel estimates. | math.FA | math |
Submitted to the Bulletin of the Australian Mathematical Society
doi:10.1017/S . . .
Lp-Lq OFF-DIAGONAL ESTIMATES FOR THE
ORNSTEIN–UHLENBECK SEMIGROUP: SOME POSITIVE
AND NEGATIVE RESULTS
ALEX AMENTA ✉ and JONAS TEUWEN
Abstract
We investigate Lp(γ)–Lq(γ) off-diagonal estimates for the Ornstein–Uhlenbeck semigroup (etL)t>0. For
sufficiently large t (quantified in terms of p and q) these estimates hold in an unrestricted sense, while
for sufficiently small t they fail when restricted to maximal admissible balls and sufficiently small annuli.
Our counterexample uses Mehler kernel estimates.
2010 Mathematics subject classification: primary 47D06; secondary 43A99.
Keywords and phrases: Ornstein–Uhlenbeck semigroup, off-diagonal estimates, Mehler kernel.
Consider the Gaussian measure
1. Introduction
dγ(x) := π−n/2e−x2 dx
(1.1)
on the Euclidean space Rn, where n ≥ 1. Naturally associated with this measure space
is the Ornstein–Uhlenbeck operator
1
2
∆ − hx, ∇i = −
∇∗∇,
L :=
1
2
where ∇∗ is the adjoint of the gradient operator ∇ with respect to the Gaussian measure.
This operator generates a heat semigroup (etL)t>0 on L2(γ) = L2(Rn, γ), called the
Ornstein–Uhlenbeck semigroup, with an explicit kernel: for all u ∈ L2(γ) and all
x ∈ Rn we have
Mt(x, y)u(y) dγ(y),
etLu(x) = ZRn
(1 − e−2t)n/2 exp(cid:18)−e−t x − y2
1
1 − e−2t(cid:19) exp(cid:18)2e−t hx, yi
1 + e−t(cid:19)
where
Mt(x, y) =
(1.2)
The first author acknowledges financial support from the Australian Research Council Discovery Grant
DP120103692 and the ANR project "Harmonic analysis at its boundaries" ANR-12-BS01-0013. The
second author acknowledges partial financial support from the Netherlands Organisation for Scientific
Research (NWO) by the NWO-VICI grant 639.033.604.
2
A. Amenta and J. Teuwen
is the Mehler kernel. If we equip Rn with the Euclidean distance and the Gaussian
measure, and if we consider operators associated with the Ornstein–Uhlenbeck op-
erator, we find ourselves within the realm of Gaussian harmonic analysis: here, the
Ornstein–Uhlenbeck operator takes the place of the Laplace operator ∆.1 For a deeper
introduction to Gaussian harmonic analysis see the review of Sjögren [10] and the
introduction of [11].
In this article we investigate whether the Ornstein–Uhlenbeck semigroup satisfies
Lp(γ)–Lq(γ) off-diagonal estimates: that is, estimates of (or similar to) the form
(cid:18)ZF
1/q
etL1E f q dγ(cid:19)
. t−θ exp(cid:16)−c
dist(E, F)2
t
(cid:17)(cid:18)ZE
1/p
f p dγ(cid:19)
,
(1.3)
f ∈ Lp(γ), and
for some parameters c > 0 and θ ≥ 0, where 1 ≤ p < q ≤ ∞,
for some class of testing sets E, F ⊂ X. Often such estimates hold whenever E and
F are Borel, but in applications we generally only need E to be a ball and F to be
an annulus associated with E. Such estimates serve as a replacement for pointwise
kernel estimates in the harmonic analysis of operators whose heat semigroups have
rough kernels, or no kernels at all, most notably in the solution to the Kato square
root problem [2] (see also [4]). Even though the Ornstein–Uhlenbeck semigroup
has a smooth kernel, it would be useful to show that it satisfies some form of off-
diagonal estimates, as this would suggest potential generalisation to perturbations of
the Ornstein–Uhlenbeck operator, whose heat semigroups need not have nice kernels.
Various notions of off-diagonal estimates, including (1.3), have been considered
by Auscher and Martell [3]. However, they only consider doubling metric measure
spaces, ruling out the non-doubling Gaussian measure. Mauceri and Meda [7]
observed that γ is doubling when restricted to admissible balls, in the sense that
γ(B(x, 2r)) . γ(B(x, r)) when r ≤ min(1, x−1). Therefore it is reasonable to expect
that the Ornstein–Uhlenbeck semigroup may satisfy some form of Lp(γ)–Lq(γ) off-
diagonal estimates if we restrict the testing sets E, F to admissible balls and sufficiently
small annuli.
Here we demonstrate both the success and failure of off-diagonal estimates of the
form (1.3), as a first step in the search for the 'right' off-diagonal estimates. First
we give a simple positive result (Theorem 2.3): for p ∈ (1, 2), and for t sufficiently
large (depending on p), (1.3) is satisfied for all Borel E, F ⊂ Rn. This is proven
by interpolating between L2(γ)–L2(γ) Davies–Gaffney-type estimates and Nelson's
Lp(γ)–L2(γ) hypercontractivity. We follow with a negative result (Theorem 3.1): for
1 ≤ p < q < ∞ and for t sufficiently small (again depending on p and q), (1.3)
fails when E is a 'maximal' admissible ball B(cB, cB−1) and when F is a sufficiently
small annulus Ck(B), in the sense that the implicit constant in (1.3) must blow up
exponentially in cB. This is shown by direct estimates of the Mehler kernel.
1 The multiplicative factor 1/2, which is not present in the usual definition of the Laplacian, arises
naturally from the probabilistic interpretation of the Ornstein–Uhlenbeck operator.
Off-diagonal estimates for the Ornstein–Uhlenbeck semigroup
3
Notation Throughout the article we will work in finite dimension n ≥ 1. We will
write Lp(γ) = Lp(Rn, γ). Every ball B ⊂ Rn is of the form
B = B(cB, rB) = {x ∈ Rn : x − cB < rB}
for some unique centre cB ∈ Rn and radius rB > 0. For each ball B and each scalar
λ > 0 we define the expansion λB = λB(cB, rB) := B(cB, λrB), and we define annuli
(Ck(B))k∈N by
For two sets E, F ⊂ Rn we write
2B
k = 0,
2k+1B \ 2kB k ≥ 1.
Ck(B) :=
dist(E, F) := inf{x − y : x ∈ E, y ∈ F}.
For two non-negative numbers A and B, we write A .a1,a2,... B to mean that A ≤ CB,
where C is a positive constant depending on the quantities a1, a2, . . .. This constant
will generally change from line to line.
2. A positive result
The Ornstein–Uhlenbeck semigroup satisfies the following 'Davies–Gaffney-type'
L2(γ)–L2(γ) off-diagonal estimates. These appear in [13, Example 6.1], where they
are attributed to Alan McIntosh.
Theorem 2.1 (McIntosh). There exists a constant C > 0 such that for all Borel subsets
E, F of Rn and all u ∈ L2(γ),
k1FetL(1Eu)kL2(γ) ≤ C
t
dist(E, F)
dist(E, F)2
2t
exp(cid:16)−
(cid:17)k1EukL2(γ).
Furthermore, Nelson [8] established the following hypercontractive behaviour of
the semigroup.1
Theorem 2.2 (Nelson). Let t > 0 and p ∈ (1 + e−2t, 2]. Then etL is a contraction from
Lp(γ) to L2(γ).
Note that p > 1 + e−2t if and only if t > 1
p−1. Thus the hypercontractive
behaviour of the Ornstein–Uhlenbeck semigroup is much more delicate than that of
the usual heat semigroup et∆ on Rn, which is a contraction from Lp(Rn) into Lq(Rn) for
all 1 ≤ p ≤ q ≤ ∞ and all t > 0.
2 log 1
As indicated in the proof of [1, Proposition 3.2], one can interpolate between
Theorems 2.1 and 2.2 to deduce certain Lp(γ)-L2(γ) off-diagonal estimates for the
Ornstein–Uhlenbeck semigroup.
1 This is done only for n = 1 in this reference, and a full proof for general n is given in Nelson's seminal
1973 paper [9]. These papers won him the 1995 Steele prize.
4
A. Amenta and J. Teuwen
Theorem 2.3. Suppose that E, F are Borel subsets of Rn. Let t > 0 and p ∈ (1+e−2t, 2].
Then for all u ∈ Lp(γ),
k1FetL(1Eu)kL2(γ) ≤ (cid:18)
Ct
dist(E, F)
dist(E, F)2
2t
exp(cid:16)−
1−δ(p,t)
(cid:17)(cid:19)
k1EukLp(γ),
where C is the constant from Theorem 2.1 and where
Proof. Write
Theorem 2.1 says that
δ(p, t) :=
1
2 − 1
p
2 − 1
1
1+e−2t
∈ [0, 1).
CM :=
Ct
dist(E, F)
exp(cid:16)
dist(E, F)2
2t
(cid:17).
ketLkL2(γ,E)→L2(γ,F) ≤ CM.
For all p0 ∈ (1 + e−2t, p) we have
ketLkLp0 (γ,E)→L2(γ,F) ≤ ketLkLp0 (γ)→L2(γ) ≤ 1
by Theorem 2.2. Therefore by the Riesz–Thorin theorem we get
ketLkLp(γ,E)→Lp(γ,F) ≤ Cθ(p0)
M ,
where p−1 = (1 − θ(p0))/p0 + θ(p0)/2, or equivalently
θ(p0) =
1
p − 1
p0
1
2 − 1
p0
= 1 −
1
2 − 1
p
1
2 − 1
p0
.
Taking the limit as p0 → 1 + e−2t gives
ketLkLp(γ,E)→Lp(γ,F) ≤ C1−δ(p,t)
M
and completes the proof.
(cid:3)
Remark 2.4. For 1 < p < q < ∞, a Lp(γ)–Lq(γ) version of Theorem 2.3 could
be proven by first establishing Lq(γ)–Lq(γ) off-diagonal estimates-which may be
obtained by interpolating between boundedness on Lq(γ) and the Davies–Gaffney type
estimates-and then arguing by the Lp(γ)–Lq(γ) version of Nelson's theorem.
This positive result does not rule out the possibility of some restricted Lp(γ)–L2(γ)
off-diagonal estimates for p ≤ 1 + e−2t. In the next section we show one way in which
these can fail.
Off-diagonal estimates for the Ornstein–Uhlenbeck semigroup
5
3. Lower bounds and negative results
In this section we show that the Lp(γ)–Lq(γ) off-diagonal estimates of (1.3) are not
satisfied for admissible balls and small annuli when t is sufficiently small (depending
on p and q). More precisely, we show that (1.3) fails when E is a maximal admissible
ball B, i.e. a ball for which rB = min(1, cB−1), and F is an annulus Ck(B) with k
sufficiently small. These sets typically appear in applications of off-diagonal estimates.
Theorem 3.1. Suppose that 1 ≤ p < q < ∞, and that
1
2
or equivalently that
−
p
q(cid:19),
et + 1
> 1 −(cid:18) 1
p − 1
t < log 1 + ( 1
q)
q)!.
p − 1
1 − ( 1
(3.1)
Then the off-diagonal estimates (1.3) do not hold for the class of testing sets
{(E, F) : E = B(cB, cB−1), F = Ck(B), 2k ≤ cB}.
q ∈ (0, 1), so we always obtain some range of t for which the off-
Note that 1
p − 1
diagonal estimates (1.3) fail.
Let us compare Theorems 3.1 and 2.3. Having fixed p ∈ (1, 2), we get failure of
Lp(γ)–L2(γ) off-diagonal estimates for maximal admissible balls and small annuli for
p−1 the off-diagonal estimates hold for
2 log 1
etL when t < log(cid:18) 1+( 1
1−( 1
p − 1
2 )
p − 1
2 )(cid:19), and when t > 1
all Borel sets. We do not know what happens for the remaining values of t.
To prove Theorem 3.1 we rely on the following lower bound.
Lemma 3.2. Suppose k ≥ 1 is a natural number, 1 < q < ∞, and let B be a maximal
admissible ball with cB ≥ 2k. Then
1/q
(etL1B)(y)q dγ(y)(cid:19)
q ) exp(cid:18)cB2(cid:18) 2
&k,n,t cB−n(1+ 1
(cid:18)ZCk(B)
q(cid:19)(cid:19).
et + 1
− 1 −
1
Proof of Lemma 3.2. Suppose x ∈ B and y ∈ C j(B). We argue by computing a lower
bound for the Mehler kernel Mt(x, y) as given in (1.2).
First we focus on the factor involving the inner product hx, yi, where x =
(x1, x2, . . . , xn) and y = (y1, y2, . . . , yn). By symmetry we may assume that cB = cBe1.
Using rB = cB−1, we get that
x1y1 ≥ (cB − rB)(cB − 2k+1rB) ≥ cB2 + O(1),
where we use the big-O notation O(1) to mean that x1y1 − cB2 is bounded as cB → ∞.
If n ≥ 2, then by using xiyi = O(1) for i ≥ 2 we get that
evidently this estimate remains true when n = 1.
hx, yi ≥ cB2 + O(1);
6
A. Amenta and J. Teuwen
Using the Mehler kernel representation of etL, for all y ∈ Ck(B) we thus have
etL1B(y) &n,t ZB
exp(cid:18)−e−t x − y2
1 − e−2t(cid:19) exp(cid:18) 2cB2
et + 1(cid:19) dγ(x).
Since x − y < 2k+1rB ≤ 2, using rB = cB−1 ≤ 2−k, this gives
etL1B(y) &n,t exp(cid:18) 2cB2
et + 1(cid:19)γ(B)
&n cB−n exp(cid:18) 2cB2
≃ cB−n exp(cid:18)cB2(cid:18)
et + 1
2
et + 1
− (cB + cB−1)2(cid:19)
− 1(cid:19)(cid:19)
(3.2)
using a straightforward estimate on γ(B). Next, we estimate
γ(Ck(B)) &n Ck(B)e−(cB+2k+1rB)2
≃n 2knrn
B exp(cid:18)−(cB2 + 2k+2cBrB + 2k+2r2
B)(cid:19)
≃k,n cB−ne−cB2
.
Combining this with (3.2) gives
(cid:18)ZCk(B)
(etL1B(y))q dγ(y)(cid:19)
1/q
&n,t cB−n exp(cid:18)cB2(cid:18) 2
&k,n cB−n(1+ 1
et + 1
q ) exp(cid:18)cB2(cid:18) 2
et + 1
− 1(cid:19)(cid:19)γ(Ck(B))1/q
q(cid:19)(cid:19),
− 1 −
1
as claimed.
(cid:3)
Proof of Theorem 3.1. We argue by contradiction. Suppose that etL satisfies the
Lp(γ)–Lq(γ) off-diagonal estimates (1.3) for some θ ≥ 0, and for (E, F) as stated.
Fix a natural number k ≥ 1 and let B be a maximal admissible ball with cB > 2k.
Lemma 3.2 and the off-diagonal estimates for E = B, F = Ck(B), and f = 1B then
imply
cB−n(1+ 1
et + 1
q ) exp(cid:18)cB2(cid:18) 2
.k,n,t,θ exp(cid:18)−c
≃ γ(B)1/p
− 1 −
1
q(cid:19)(cid:19)
(2k+1 − 1)2r2
B
t
(cid:19)γ(B)1/p
for some c > 0. Since
γ(B)1/p
.n B1/pe− 1
p (cB−rB)2
≃n cB−n/p exp(cid:18)−
cB2
p (cid:19),
Off-diagonal estimates for the Ornstein–Uhlenbeck semigroup
7
this implies
exp(cid:18)cB2(cid:18)
2
et + 1
− 1 +
1
p
−
1
q(cid:19)(cid:19) .k,n,t,θ cB
n(1−( 1
p − 1
q ))
.
The left hand side grows exponentially in cB when (3.1) is satisfied. However, the
right hand side only grows polynomially in cB. Thus we have a contradiction.
(cid:3)
Remark 3.3. By the same argument we can prove failure of Lp(γ)–Lq(γ) off-diagonal
estimates for the derivatives (LmetL)m∈N of the Ornstein–Uhlenbeck semigroup, with
the same conditions on (p, q, t) and the same class of testing sets (E, F). This relies
on an identification of the kernel of LmetL, which has been done by the second author
[12].
In this article we only considered off-diagonal estimates with respect to the Gaus-
sian measure γ. In future work it would be very interesting to consider appropriate
weighted measures, following in particular [5] and [6], in which (among many other
things) it is shown that estimates of the form ketL f kL2(γ) . k f VtkL1(γ) hold, where Vt
is a certain weight depending on t. Thus the Ornstein–Uhlenbeck semigroup does
satisfy a form of 'ultracontractivity', but with the caveat that one must keep track of
t-dependent weights. It seems that this has not yet been explored in the context of
Gaussian harmonic analysis.
Acknowledgements
The authors thank Mikko Kemppainen, Jan van Neerven, and Pierre Portal for
valuable discussions and encouragement on this topic. We also thank an anonymous
referee for their suggested simplification of the proof of Lemma 3.2.
References
[1]
[2]
[3]
P. Auscher. 'On necessary and sufficient conditions for Lp-estimates of Riesz transforms associ-
ated to elliptic operators on Rn and related estimates'. Mem. Amer. Math. Soc. 186 (871) (2007),
xviii+75.
P. Auscher, S. Hofmann, M. Lacey, A. McIntosh, and P. Tchamitchian. 'The solution of the Kato
square root problem for second order elliptic operators on Rn'. Ann. Math. 156 (2002), 633–654.
P. Auscher and J. M. Martell.
'Weighted norm inequalities, off-diagonal estimates and elliptic
operators part ii: Off-diagonal estimates on spaces of homogeneous type'. J. Evol. Equ. 7 (2)
(2007), 265–316.
[4] A. Axelsson, S. Keith, and A. McIntosh. 'Quadratic estimates and functional calculi of perturbed
Dirac operators'. Invent. Math. 163 (2006), 455–497.
[5] D. Bakry, F. Bolley, and I. Gentil.
'Dimension dependent hypercontractivity for Gaussian
kernels'. Probab. Theory Related Fields 154 (3-4) (2012), 845–874.
[6] D. Bakry, F. Bolley, I. Gentil, and P. Maheux. 'Weighted Nash inequalities'. Rev. Mat. Iberoam.
28 (3) (2012), 879–906.
[7] G. Mauceri and S. Meda. 'BMO and H1 for the Ornstein-Uhlenbeck operator'. J. Funct. Anal.
252 (1) (2007), 278–313.
[8] E. Nelson. 'A quartic interaction in two dimensions'. In Mathematical Theory of Elementary
Particles (Proc. Conf., Dedham, Mass., 1965) (M.I.T. Press, Cambridge, Mass., 1966), 69–73.
8
A. Amenta and J. Teuwen
[9] E. Nelson.
'Construction of quantum fields from Markoff fields'. J. Funct. Anal. 12 (1973),
[10]
[11]
[12]
[13]
'Operators associated with the Hermite semigroup-a survey'.
97–112.
P. Sjögren.
In Proceedings of
the conference dedicated to Professor Miguel de Guzmán (El Escorial, 1996), Volume 3 (1997),
813–823.
J. Teuwen. 'A note on Gaussian maximal functions'. Indag. Math. 26 (2015), 106–112.
J. Teuwen. 'On the integral kernels of derivatives of the Ornstein-Uhlenbeck semigroup'. Infin.
Dimens. Anal. Quantum Probab. Relat. Top. 19 (2016), to appear.
J. van Neerven and P. Portal. 'Finite speed of propagation and off-diagonal bounds for Ornstein–
Uhlenbeck operators in infinite dimensions'. Ann. Mat. Pura Appl. 195 (6) (2016), 1889–1915.
Alex Amenta, Delft Institute of Applied Mathematics, Delft University of Technology,
P.O. Box 5031, 2628 CD Delft, The Netherlands
e-mail: [email protected]
Jonas Teuwen, Division of Radiation Oncology, Netherlands Cancer Institute/Antoni
van Leeuwenhoek, Plesmanlaan 121, 1066 CX Amsterdam, The Netherlands
Department of Imaging Physics, Optics Research Group, Delft University of Tech-
nology
e-mail: [email protected]
|
1804.06983 | 1 | 1804 | 2018-04-19T02:50:58 | Characterizations of Nonsmooth Robustly Quasiconvex Functions | [
"math.FA"
] | Two criteria for the robust quasiconvexity of lower semicontinuous functions are established in terms of Fr\'echet subdifferentials in Asplund spaces. | math.FA | math |
Characterizations of Nonsmooth Robustly Quasiconvex
Functions
Hoa T. Bui∗, Pham Duy Khanh†,Tran Thi Tu Trinh‡
April 20, 2018
Abstract Two criteria for the robust quasiconvexity of lower semicontinuous functions are es-
tablished in terms of Fr´echet subdifferentials in Asplund spaces.
Keywords Quasiconvexity, robust quasiconvexity, quasimonotone, Fr´echet subdifferential, ap-
proximate mean value theorem
Mathematics Subject Classification (2010) 26A48, 26A51, 49J52, 49J53
1
Introduction
The question of characterizing convexity and generalized convexity properties in terms of subdifferentials
receives tremendous attention in optimization theory and variational analysis. For decades, there have been
received many significant contributions devoted to this question such as [8, 9, 11, 14, 16] for convex functions,
[2, 3, 4, 6, 12, 13] for quasiconvex functions and [6] for robustly quasiconvex functions.
This paper follows this stream of research. Our aim is to establish the first-order characterizations for the
robust quasiconvexity of lower semicontinuous functions in Asplund spaces. First, some existing results re-
garding to the properties of subdifferential operators of convex, quasiconvex functions are recalled in Section
2, where the definitions and some basic results are given as well. Besides, necessary and sufficient first-order
conditions for a lower semicontinuous function to be quasiconvex are reconsidered. Those characterizations
moreover could be used to characterize the Asplund property of the given space. Second, two criteria for the
robust quasiconvexity of lower semicontinuous functions in Asplund spaces are obtained by using Fr´echet
subdifferentials in Section 3. Each criterion corresponds to each type of analogous conditions for quasicon-
vexity. The first one is based on the zero and first order condition for quasiconvexity (see Theorem 2.2(b)
in Section 2). It extends [6, Proposition 5.3] from finite dimensional spaces to Asplund spaces. Moreover,
its proof also overcomes a glitch in the proof of the sufficient condition of [6, Proposition 5.3]. The second
criterion is totally new.
It is settled from the equivalence of the quasiconvexity of lower semicontinuous
functions and the quasimonotonicity of their subdifferential operators (see Theorem 2.2(c) in Section 2).
∗Centre for Informatics and Applied Optimization, Faculty of Science and Technology, Federation University Australia, POB
663, Ballarat, Vic, 3350, Australia. E-mail: [email protected]
†Department of Mathematics, HCMC University of Pedagogy, 280 An Duong Vuong, Ho Chi Minh, Vietnam and Cen-
ter for Mathematical Modeling, Universidad de Chile, Beauchef 851, Edificio Norte, Piso 7, Santiago, Chile. E-mails: pd-
[email protected]; [email protected]
‡Department of Mathematics and Statistics, Oakland University, 318 Meadow Brook Rd, Rochester, MI 48309, USA. Email:
[email protected]
1
2 Preliminaries
Let X be a Banach space and X ∗ its dual space. X is called an Asplund space, or has the Asplund property,
if every separable subspace Y of X has separable continuous dual space Y ∗. The duality pairing on X × X ∗
is denoted by h., .i. In what follows, R :=] − ∞, ∞]; Br(x) is the open ball of radius r > 0 centered at x ∈ X
and B∗ ⊂ X ∗ is the closed ball of radius 1 centered at 0X ∗. The extended real-valued function ϕ : X → R
considered mostly is proper lower semicontinuous (l.s.c), i.e. ϕ is not identically +∞, and the lower level
sets ϕ≤
α := {x ∈ X : ϕ(x) ≤ α} are closed for all α ∈ R. As usual domϕ stands for the domain of ϕ, defined
as
For a set-valued mapping A : X ⇒ X ∗, the domain of A is written
domϕ := {x ∈ X : ϕ(x) < ∞}.
domA := {x ∈ X : A(x) 6= ∅}.
The graphs of ϕ and A are respectively defined as
graphϕ := {(x, α) ∈ X × R : ϕ(x) = α},
graphA := {(x, x∗) ∈ X × X ∗ : x∗ ∈ A(x)}.
A subset U of X is convex if it contains all closed segments connecting two points in U . The function ϕ
is said to be convex if the domain of ϕ is convex and for any α ∈ [0, 1], x, y ∈ domϕ we always have the
inequality ϕ(αx + (1 − α)y) ≤ αϕ(x) + (1 − α)ϕ(y).
As usual, the Fr´echet subdifferential of a proper lower semicontinuous function ϕ is the set-valued mapping
When ϕ is convex, the Fr´echet subdifferential reduces to the convex analysis subdifferential
for all x ∈ domϕ.
for all x ∈ domϕ.
b∂ϕ : X ⇒ X ∗ defined by
y→x
ky − xk
ϕ(y) − ϕ(x) − hx∗, y − xi
b∂ϕ(x) :=(cid:26)x∗ ∈ X ∗ : lim inf
b∂ϕ(x) = ∂ϕ(x) := {x∗ ∈ X ∗ : hx∗, y − xi ≤ ϕ(y) − ϕ(x)},
≥ 0(cid:27) ,
An operator A is monotone if for all x, y ∈ domA, one has hx∗ − y∗, x − yi ≥ 0 with x∗ ∈ A(x), y∗ ∈ A(y). It
is well-known that when ϕ is convex, the operator b∂ϕ is monotone [16]. The inverse implication also holds
in Asplund space [11, Theorem 3.56]; but it is not true in general Banach spaces. The reader is referred to
the proof of the reverse implication in [10, Theorem 2.4] for a counter-example.
Let us recall some notions of generalized convex functions.
Definition 2.1 A function ϕ : X → R is
1. quasiconvex if
∀x, y ∈ X, λ ∈]0, 1[,
f (λx + (1 − λ)y) ≤ max{f (x), f (y)}.
(1)
2. α-robustly quasiconvex with α > 0 if, for every v∗ ∈ αB∗, the function ϕv∗ : x 7→ ϕ(x) + hv∗, xi is
quasiconvex.
Clearly, ϕ is α-robustly quasiconvex iff the function ϕv∗ is quasiconvex for all v∗ ∈ X ∗ such that kv∗k < α.
Tracing back to the original definition of robustly quasiconvex functions, they were first defined in [15]
under the name "s-quasiconvex" or "stable quasiconvex", and then renamed "robustly quasiconvex" in [5].
2
This class of functions holds a notable role, as many important optimization properties of generalized convex
functions are stable when disturbed by a linear functional with a sufficiently small norm (for instance, all
lower level sets are convex, each minimum is global minimum, each stationary point is a global minimizer).
For interested readers, we refer to [15] again, and further related works [1, 5].
Definition 2.2 An operator A : X ⇒ X ∗ is quasimonotone if for all x, y ∈ X and x∗ ∈ A(x), y∗ ∈ A(y) we
have min{hx∗, y − xi, hy∗, x − yi} ≤ 0.
Significant contributions concerning dual criteria for quasiconvex functions are in [2, 4]. Those character-
izations are applicable for a wide range of subdifferentials, for instance Rockafellar-Clarke subdifferentials
in Banach spaces, and Fr´echet subdifferentials in reflexive spaces. These results are still unclear for Fr´echet
subdifferentials in Asplund spaces. Below, we give a short proof to clarify this. Our proof relies on the proof
scheme of [2] and the following approximate mean value theorem [11, Theorem 3.49].
Theorem 2.1 Let X be an Asplund space and ϕ : X → R be a proper lower semicontinuous function finite
at two given points a 6= b. Consider any point c ∈ [a, b) at which the function
ψ(x) := ϕ(x) −
ϕ(b) − ϕ(a)
ka − bk
kx − ak
attains its minimum on [a, b]; such a point always exists. Then, there are sequences xk
satisfying
Moreover, when c 6= a one has
lim inf
k→∞
hx∗
k, b − xki ≥
ϕ(b) − ϕ(a)
ka − bk
kb − ck,
lim inf
k→∞
hx∗
k, b − ai ≥ ϕ(b) − ϕ(a).
lim
k→∞
hx∗
k, b − ai = ϕ(b) − ϕ(a).
ϕ
→ c and x∗
k ∈ b∂ϕ(xk)
(2)
(3)
(4)
Theorem 2.1 allows us to deduce the following three-points lemma which is similar to [3, Lemma 3.1].
Lemma 2.1 Let ϕ : X → R be a proper, lower semicontinuous function on an Asplund space X. Let
u, v, w ∈ X such that v ∈ [u, w], ϕ(v) > ϕ(u) and λ > 0. Then, there are ¯x ∈ domϕ and ¯x∗ ∈ b∂ϕ(¯x) such
that
¯x ∈ Bλ([u, v]) and h¯x∗, w − ¯xi > 0,
where
Bλ([u, v]) := {x ∈ X : ∃y ∈ [u, v] such that kx − yk < λ}.
We are in position to establish characterizations of quasiconvexity in terms of Fr´echet subdifferentials in
Asplund spaces.
Theorem 2.2 Let ϕ : X → R be a proper lower semicontinuous function on an Asplund space X. The
following statements are equivalent
(a) ϕ is quasiconvex;
(b) If there are x, y ∈ X such that ϕ(y) ≤ ϕ(x), then hx∗, y − xi ≤ 0 for all x∗ ∈ b∂ϕ(x).
(c) b∂ϕ is quasimonotone.
3
Proof.
ϕ(x)}. Since ϕ is quasiconvex, then Sx is a convex set. Thus, we have the function f := δSx + ϕ(x) is convex,
where δSx is equal to 0 for u ∈ Sx and to ∞ otherwise. On the other hand, f (x) = ϕ(x) and f (u) ≥ ϕ(u)
(a)⇒(b) Assume that x, y ∈ X, ϕ(x) ≥ ϕ(y), and x∗ ∈ b∂ϕ(x). Consider Sx := {u ∈ X : ϕ(u) ≤
for all u ∈ X, thus b∂ϕ(x) ⊂ b∂f (x). By the definition of convex subdifferential, since x∗ ∈ b∂ϕ(x) ⊂ b∂f (x),
(b)⇒(c) Assume that there are x, y ∈ X and x∗ ∈ b∂ϕ(x), y∗ ∈ b∂ϕ(y) such that hx∗, x − yi < 0 and
we have hx∗, y − xi ≤ 0.
hy∗, x − yi > 0. Then, by (b), ϕ(x) < ϕ(y) and ϕ(y) < ϕ(x), which is a contradiction.
(c)⇒(a) By using Lemma 2.1, the proof of this assertion is similar to one in [2, Theorem 4.1].
Remark 2.1 Observe that the implications (a) ⇒ (b) and (b) ⇒ (c) hold in Banach spaces while (c) ⇒ (a)
only holds in Asplund spaces.
In fact, the equivalence of these statements actually can characterize the
Asplund property in the sense that if X is not an Asplund space, then there is a function ϕ whose Fr´echet
subdiferential satisfies (b) and (c) but is not quasiconvex. Such a function ϕ can be found in [10, Theorem 2.4].
3 Characterizations of Robustly Quasiconvex Functions
A zero and first order characterization of robust convexity was given in [6, Proposition 5.3] for finite di-
mensional spaces. We remark that there is an oversight in the proof given there; although the function f
is only assumed to be lower semicontinuous, the existence of z in the second paragraph actually requires
continuity. Here we show that this conclusion is still correct not only when f is assumed just to be lower
semicontinuous, but also when X is only assumed to be an Asplund space. To derive this generalization, we
need the following lemmas, revealing that quasiconvex functions have certain nice properties which resemble
those of convex functions.
Lemma 3.1 If ϕ : X → R is a quasiconvex and lower semicontinuous function, and u, v ∈ X are such that
ϕ(v) ≥ ϕ(u) then
ϕ(v + t(u − v)) = ϕ(v).
lim
t↓0
(5)
Suppose that u, v ∈ X and that ϕ(v) ≥ ϕ(u). Since ϕ is quasiconvex, for all t ∈]0, 1[, we have
Proof.
ϕ(v + t(u − v)) ≤ max{ϕ(v), ϕ(u)} = ϕ(v). It follows that lim supt↓0 ϕ(v + t(u − v)) ≤ ϕ(v). Combining the
latter with the lower semicontinuity of ϕ we get (5).
✷
Lemma 3.2 Let ϕ : X → R be a quasiconvex function and u, v, w ∈ X such that v ∈]u, w[, ϕ(u) ≤ ϕ(w).
Suppose that there exist v∗ ∈ X ∗ and z ∈]u, v[ such that ϕv∗ (z) > max{ϕv∗ (u), ϕv∗ (w)}. Then
ϕ(u) < ϕ(z) ≤ ϕ(v) ≤ ϕ(w).
(6)
Proof. Since z ∈]u, v[⊂]u, w[, ϕ(u) ≤ ϕ(w) and ϕ is quasiconvex we have ϕ(z) ≤ max{ϕ(u), ϕ(w)} = ϕ(w).
Hence, the latter and the inequality ϕv∗ (z) > ϕv∗ (w) implies that hv∗, zi > hv∗, wi. Again, z ∈]u, w[ implies
hv∗, zi < hv∗, ui. Therefore, the inequality ϕv∗ (u) < ϕv∗ (z) yields ϕ(u) < ϕ(z). Since z ∈]u, v[ and v ∈]z, w[,
we deduce ϕ(z) ≤ ϕ(v) ≤ ϕ(w) from the latter inequality and the quasiconvexity of ϕ. Hence, (6) holds. ✷
Lemma 3.3 Let ϕ : X → R be a quasiconvex, proper, and lower semicontinuous function, and v∗ ∈ X ∗. If
ϕv∗ is not quasiconvex then there exist u, v, w ∈ X such that v ∈]u, w[ and
ϕ(w) ≥ ϕ(v) > ϕ(u),
ϕv∗ (v) > max{ϕv∗ (u), ϕv∗ (w)},
∀γ > 0, ∃vγ ∈ Bγ(v)∩]v, w[ : ϕv∗ (v) > ϕv∗ (vγ).
4
(7)
(8)
(9)
Since ϕv∗ is not quasiconvex, there exist u, w ∈ X such that u 6= w, ϕ(u) ≤ ϕ(w) and v0 ∈]u, w[
Proof.
such that ϕv∗ (v0) > max{ϕv∗ (u), ϕv∗ (w)}. Applying Lemma 3.1, we get limt↓0 ϕ(w + t(u − w)) = ϕ(w), and
so limt↓0 ϕv∗ (w + t(u − w)) = ϕv∗ (w). Since ϕv∗ (w) < ϕv∗ (v0), there exists t0 ∈]0, 1[ such that
ϕv∗ (w + t(u − w)) < ϕv∗ (v0),
∀t ∈]0, t0[.
(10)
Consider the set
L := {z ∈]u, w[: ϕv∗ (z) ≥ ϕv∗ (v0)}.
Clearly, L 6= ∅ and for each z ∈ L we have kz − wk ≥ t0ku − wk by (10). It follows that
r := inf{kz − wk : z ∈ L } ∈ [t0ku − wk, ku − wk[ ⊂ ]0, ku − wk[,
v := w + r
u − w
ku − wk
∈ ]u, w[.
We will show that v ∈ L and so (8) holds. Suppose on the contrary that v /∈ L . Then v0 ∈]u, v[ and we
get ϕ(u) < ϕ(v0) ≤ ϕ(v) ≤ ϕ(w) by Lemma 3.2. Applying Lemma 3.1, we get limt↓0 ϕ(v + t(u − v)) = ϕ(v),
and so limt↓0 ϕv∗ (v + t(u − v)) = ϕv∗ (v). By the definition of r, there exists a sequence (zn) ⊂ L such that
kzn − wk → r and kzn − wk > r for all n ∈ N. Therefore,
ϕv∗ (v) = lim
t↓0
ϕv∗ (v + t(u − v))
kzn − wk − r
(u − v)(cid:19)
= lim ϕv∗(cid:18)v +
= lim ϕv∗(cid:18)v −
= lim ϕv∗(cid:18)v −
= lim ϕv∗(cid:18)w +
ku − vk
r
ku − vk
r
(u − v) +
(u − w) +
ku − wk
kzn − wk
ku − wk
(u − w)(cid:19)
kzn − wk
ku − vk
kzn − wk
ku − wk
(u − v)(cid:19)
(u − w)(cid:19)
= lim ϕv∗ (zn)
≥ ϕv∗ (v0),
which is a contradiction. Now we show that v satisfies (9). Let γ be any positive real number and
vγ := w +
r − rγ
ku − wk
(u − w) with rγ := min{r/2, γ/2} > 0.
Since 0 < r − rγ < r < ku − w, it implies that vγ ∈]v, w[ \ L . Therefore, ϕv∗ (vγ) < ϕv∗ (v0) ≤ ϕv∗ (v).
Furthermore,
kvγ − vk = (cid:13)(cid:13)(cid:13)(cid:13)w +
Hence, v satisfies (9).
r − rγ
ku − wk
(u − w) − w − r
u − w
ku − wk(cid:13)(cid:13)(cid:13)(cid:13) = rγ < γ.
✷
Theorem 3.1 Let ϕ : X → R be a proper lower semicontinuous function on a Banach space X, and α > 0.
Consider the following statements
(a) ϕ is α−robustly quasiconvex;
(b) For every x, y ∈ X
ϕ(y) ≤ ϕ(x) =⇒ hx∗, y − xi ≤ − min {αky − xk, ϕ(x) − ϕ(y)} , ∀x∗ ∈ b∂ϕ(x).
(11)
5
Then (a)⇒(b). Additionally, if X is an Asplund space, then (b)⇒(a).
Proof.
Suppose that ϕ is α−robustly quasiconvex, and x, y ∈ X satisfy ϕ(y) ≤ ϕ(x). Assume that
x∗ ∈ b∂ϕ(x). We will prove
hx∗, y − xi ≤ − min {αky − xk, ϕ(x) − ϕ(y)} .
If x = y, the above inequality is trivial. Otherwise, we consider two cases:
Case 1.
We then need to prove that
αky − xk ≤ ϕ(x) − ϕ(y)
hx∗, y − xi ≤ −αky − xk.
(12)
By the Hahn-Banach theorem, there exists v∗ ∈ X ∗, kv∗k = 1 such that hv∗, y − xi = ky − xk. Consider the
function f : X → R given by
f (z) = ϕ(z) + αhv∗, z − xi ∀z ∈ X.
Then f (x) = ϕ(x), and
f (y) = ϕ(y) + αhv∗, y − xi = ϕ(y) + αky − xk ≤ ϕ(x) = f (x),
i.e., max{f (x), f (y)} = f (x). Since ϕ is α−robustly quasiconvex, f is quasiconvex. Therefore for each
t ∈ [0, 1], we always have
ϕ(x) = f (x) = max{f (x), f (y)} ≥ f (x + t(y − x))
= ϕ(x + t(y − x)) + tαhv∗, y − xi
= ϕ(x + t(y − x)) + tαky − xk,
(13)
(14)
(15)
which implies that
ϕ(x) − tαky − xk ≥ ϕ(x + t(y − x)).
Since x∗ ∈ b∂ϕ(x), for any γ > 0, there exists a number r > 0 such that
ϕ(z) ≥ ϕ(x) + hx∗, z − xi − γkz − xk ∀z ∈ Br(x).
Let t ∈]0, 1[ such that x + t(y − x) ∈ Br(x). It follows from (13) and (14) that
ϕ(x) − tαky − xk ≥ ϕ(x) + thx∗, y − xi − tγky − xk,
and so
On taking limit on both sides of the above inequality as γ → 0+, we get (12).
hx∗, y − xi ≤ −αky − xk + γky − xk.
Case 2.
We have ¯αky − xk = ϕ(x) − ϕ(y), where
αky − xk > ϕ(x) − ϕ(y)
Since ϕ is ¯α−robustly quasiconvex, we derive from Case 1 that
¯α :=
ϕ(x) − ϕ(y)
ky − xk
∈]0, α[.
hx∗, y − xi ≤ − ¯αky − xk = ϕ(y) − ϕ(x)
= − min {αky − xk, ϕ(x) − ϕ(y)} .
6
Conversely, assume that X is Asplund, and (b) holds. It follows from Theorem 2.2 that ϕ is quasiconvex.
Suppose that ϕ is not α−robustly quasiconvex, i.e., there exists v∗ ∈ X ∗ \ {0}, kv∗k < α such that ϕv∗
is not quasiconvex. By Lemma 3.3, there are u, w ∈ X and v ∈]u, w[ satisfying (7),(8), and (9). Since
ϕv∗ (v) > ϕv∗ (u), there exists δ > 0 such that ¯v∗ := (1 + δ)v∗ satisfies k¯v∗k < α and ϕ¯v∗ (v) > ϕ¯v∗ (u). Thus,
we have ϕ(v) > ϕ(u), ϕv∗ (v) > ϕv∗ (u), ϕ¯v∗ (v) > ϕ¯v∗ (u) and the lower semicontinuity of ϕ, ϕv∗ , and ϕ¯v∗ .
This implies the existence of γ > 0 satisfying
ϕ(z) > ϕ(u), ϕv∗ (z) > ϕv∗ (u), ϕ¯v∗ (z) > ϕ¯v∗ (u) ∀z ∈ Bγ(v).
(16)
By the assertion (9), there is vγ ∈ Bγ(v)∩]v, w[ such that ϕv∗ (v) > ϕv∗ (vγ). Then, vγ can be written as
vγ := v + λ(w − v) with λ ∈(cid:21)0, min(cid:26)1,
γ
kw − vk(cid:27)(cid:21) .
Since ϕv∗ (v) > ϕv∗ (w) and ϕ(v) ≤ ϕ(w), we have hv∗, w − vi < 0 and so
ϕ¯v∗ (vγ) − ϕ¯v∗ (v) = ϕv∗ (vγ) − ϕv∗ (v) + δhv∗, vγ − vi
= ϕv∗ (vγ) − ϕv∗ (v) + δλhv∗, w − vi < 0.
Applying Lemma 2.1 for ϕ¯v∗ , v ∈ [vγ, u] with ϕ¯v∗ (v) > ϕ¯v∗ (vγ), there exist x ∈ domϕ¯v∗ and x∗ ∈ b∂ϕ¯v∗ (x)
such that
x ∈ [vγ, v] + (r − kvγ − vk)B and hx∗, u − xi > 0.
(17)
Then x ∈ Bγ(v) and so ϕ(x) > ϕ(u) by (16). By the assumption (b) and the second inequality of (17),
−h¯v∗, u − xi < hx∗ − ¯v∗, u − xi ≤ − min{αku − xk, ϕ(x) − ϕ(u)}.
Since h¯v∗, u − xi ≤ k¯v∗kku − xk < αku − xk, the above inequality implies that h¯v∗, u − xi > ϕ(x) − ϕ(u), i.e.,
ϕ¯v∗ (x) < ϕ¯v∗ (u) and this contradicts (16).
✷
We next construct a completely new characterization for the robust quasiconvexity.
It is based on
the equivalence of the quasiconvexity of a lower semicontinuous function and the quasimonotonicity of its
subdifferential operator.
Theorem 3.2 Let ϕ : X → R be proper, lower semicontinuous on an Asplund space X and α > 0. Then,
ϕ is α−robustly quasiconvex if and only if for any (x, x∗), (y, y∗) ∈ graph b∂ϕ, we have
min{hx∗, y − xi, hy∗, x − yi} > −αky − xk =⇒ hx∗ − y∗, x − yi ≥ 0.
(18)
Proof.
that
Suppose that ϕ is α−robustly quasiconvex and that there exist (x, x∗), (y, y∗) ∈ graph b∂ϕ such
min{hx∗, y − xi, hy∗, x − yi} > −αky − xk.
Since ϕ is quasiconvex, b∂ϕ is quasimonotone by Theorem 2.2. It follows that
min{hx∗, y − xi, hy∗, x − yi} ≤ 0.
Combining (19) and (20), we have
0 ≤ − min(cid:26)(cid:28)x∗,
y − x
ky − xk(cid:29) ,(cid:28)y∗,
x − y
kx − yk(cid:29)(cid:27) < α.
7
(19)
(20)
Without loss of generality, we may assume
(cid:28)x∗,
y − x
ky − xk(cid:29) = min(cid:26)(cid:28)x∗,
Let r > 0 be such that
−(cid:28)x∗,
y − x
ky − xk(cid:29) ,(cid:28)y∗,
ky − xk(cid:29) < r ≤ α.
y − x
x − y
kx − yk(cid:29)(cid:27) .
(21)
By the Hahn-Banach theorem, there exists v∗ ∈ X ∗ satisfying hv∗, y − xi = rky − xk and kv∗k = r ≤ α. It
follows that
hx∗, y − xi + hv∗, y − xi > −rky − xk + rky − xk = 0.
(22)
Consider ϕv∗ : X → R given by ϕv∗ (u) = ϕ(u) + hv∗, ui for any u ∈ X. Then, we have b∂ϕv∗ (u) = b∂ϕ(u) + v∗
for u ∈ domϕ. Hence, by the quasiconvexity of ϕv∗ and by Theorem 2.2, we have
min{hx∗, y − xi + hv∗, y − xi, hy∗, x − yi + hv∗, x − yi} ≤ 0.
Combining with (22), it implies
hy∗, x − yi + hv∗, x − yi ≤ 0, i.e., hy∗, x − yi ≤ hv∗, y − xi = rkx − yk.
ky−xkE, we obtain hy∗, x − yi ≤ hx∗, x − yi and thus (18) holds.
Letting r → −Dx∗, y−x
Conversely, assume that (18) holds for all x, y ∈ X and x∗ ∈ b∂ϕ(x), y∗ ∈ b∂ϕ(y). Taking any v∗ in αB∗, we
the quasimonotonicity of b∂ϕv∗ . Taking any x, y ∈ X and x∗ ∈ b∂ϕv∗ (x), y∗ ∈ b∂ϕv∗ (y), then x∗ − v∗ ∈ b∂ϕ(x),
y∗ − v∗ ∈ b∂ϕ(y). We then consider two cases.
next prove that ϕv∗ : X → R, defined by ϕv∗ (u) = ϕ(u) + hv∗, ui for any u ∈ X, is quasiconvex by showing
Case 1. min{hx∗ − v∗, y − xi, hy∗ − v∗, x − yi} ≤ −αky − xk
Without loss of generality, assume that
hx∗ − v∗, y − xi = min{hx∗ − v∗, y − xi, hy∗ − v∗, x − yi}.
Since kv∗k ≤ α, we have
min{hx∗, y − xi, hy∗, x − yi} ≤ hx∗, y − xi = hx∗ − v∗, y − xi + hv∗, y − xi
≤ −αky − xk + kv∗kky − xk ≤ 0.
Case 2. min{hx∗ − v∗, y − xi, hy∗ − v∗, x − yi} > −αky − xk
Since (18) is satisfied, we have
h(x∗ − v∗) − (y∗ − v∗), x − yi ≥ 0,
i.e., hx∗ − y∗, x − yi ≥ 0. It implies that
2 min{hx∗, y − xi, hy∗, x − yi} ≤ hx∗, y − xi + hy∗, x − yi ≤ 0.
Hence, b∂ϕv∗ is quasimonotone and thus ϕv∗ is quasiconvex for any v∗ ∈ αB∗ by Theorem 2.2. This yields
the α-robust quasiconvexity of ϕ.
✷
8
4 Conclusions
Using Fr´echet subdifferentials, we have obtained two first-order characterizations for the robust quasicon-
vexity of lower semicontinuous functions in Asplund spaces. The first one is a generalization of [6, Proposi-
tion 5.3] from finite dimensional spaces to Asplund spaces and its proof also overcomes a glitch in the proof
of the sufficient condition of [6, Proposition 5.3]. The second criterion is totally new and it is settled from
the equivalence of the quasiconvexity of lower semicontinuous functions and the quasimonotonicity of their
subdifferential operators. Further investigations are needed to apply those characterizations in partial differ-
ential equations with connections to differential geometry, mean curvature, tug-of-war games, and stochastic
optimal control [5, 6, 7].
Acknowlegement
This work was completed while the second author was visiting Vietnam Institute for Advanced Study in
Mathematics (VIASM). He would like to thank VIASM for the very kind support and hospitality.
References
1. An P.T.: Stability of generalized monotone maps with respect to their characterizations. Optimization
55, 289–299 (2006)
2. Aussel D., Corvellec J.-N., Lassonde M.: Subdifferential characterization of quasiconvexity and convexity.
J. Convex Anal. 1, 195-201 (1994)
3. Aussel D., Corvellec J.-N., Lassonde M.: Mean-value property and subdifferential criteria for lower
semicontinuous functions. Trans. Amer. Math. Soc. 347, 4147–4161 (1995)
4. Aussel D.: Subdifferential properties of quasiconvex and pseudoconvex functions: Unified approach.
Optimization 97, 29–45 (1998)
5. Barron E.N., Goebel R., Jensen R.R.: Function which are quasiconvex under linear perturbations. SIAM
J. Optim. 22, 1089–1108 (2012)
6. Barron E.N., Goebel R., Jensen R.R.: The quasiconvex envelope through first-order partial differential
equations which characterize quasiconvexity of nonsmooth functions. Discrete Contin. Dyn. Syst. Ser. B
17, 1693–1706 (2012)
7. Barron, E.N., Goebel, R., Jensen, R.R.: Quasiconvex functions and nonlinear PDEs. Trans. Amer. Math.
Soc. 365, 4229–4255 (2013)
8. Clarke F.H.: Optimization and Nonsmooth Analysis, Wiley-Interscience, New-York (1983)
9. Correa R., Jofr´e A., Thibault L.: Characterization of lower semicontinuous convex functions. Proc.
Amer. Math. Soc. 116, 67–72 (1992)
10. Trang N.T.Q.: A note on an approximate mean value theorem for Fr´echet subgradients. Nonlinear Anal.
75, 380–383 (2012)
11. Mordukhovich B.S.: Variational Analysis and Generalized Differentiation I: Basic Theory. Springer,
Berlin (2006)
9
12. Luc D.T.: Characterisations of quasiconvex functions. Bull. Austral. Math. Soc. 48, 393–406 (1993)
13. Penot J.-P., Quang P.H.: Generalized convexity of functions and generalized monotonicity of set-valued
maps. J. Optim. Theory Appl. 92, 343–356 (1997)
14. Poliquin R.A.: Subgradient monotonicity and convex functions. Nonlinear Anal. 14, 305–317 (1990)
15. Phu H.X., An P.T.: Stable generalization of convex functions. Optimization 38, 309–318 (1996)
16. Rockafellar R.T.: On the maximal monotonicity of subdifferential mappings, Pacific J. Math. 33. 209–
216 (1970)
10
|
1609.06214 | 1 | 1609 | 2016-09-20T15:13:16 | Anisotropic Shubin operators and eigenfunctions expansions in Gelfand-Shilov spaces | [
"math.FA"
] | We derive new results on the characterization of Gelfand--Shilov spaces $\mathcal{S}^\mu_\nu (\R^n)$, $\mu,\nu >0$, $\mu+\nu \geq 1$ by Gevrey estimates of the $L^2$ norms of iterates of $(m,k)$ anisotropic globally elliptic Shubin (or $\Gamma$) type operators, $(-\Delta)^{m/2} +| x |^k$ with $m,k\in 2\N$ being a model operator, and on the decay of the Fourier coefficients in the related eigenfunction expansions. Similar results are obtained for the spaces $\Sigma^\mu_\nu (\R^n)$, $\mu,\nu >0$, $\mu+\nu > 1$, cf. \eqref{GSdef}. In contrast to the symmetric case $\mu = \nu$ and $k=m$ (classical Shubin operators) we encounter resonance type phenomena involving the ratio $\kappa:=\mu/\nu$; namely we obtain a characterization of $\mathcal{S}^\mu_\nu(\R^n)$ and $\Sigma^\mu_\nu(\R^n)$ in the case $\mu=kt/(k+m), \nu= mt/(k+m), t \geq 1$, that is, when $\kappa=k/m \in \Q$. | math.FA | math |
ANISOTROPIC SHUBIN OPERATORS AND EIGENFUNCTION
EXPANSIONS IN GELFAND-SHILOV SPACES
MARCO CAPPIELLO, TODOR GRAMCHEV, STEVAN PILIPOVIC, AND LUIGI RODINO
Abstract. We derive new results on the characterization of Gelfand -- Shilov
spaces S µ
ν (Rn), µ, ν > 0, µ + ν ≥ 1 by Gevrey estimates of the L2 norms
of iterates of (m, k) anisotropic globally elliptic Shubin (or Γ) type operators,
(−∆)m/2 +xk with m, k ∈ 2N being a model operator, and on the decay of the
Fourier coefficients in the related eigenfunction expansions. Similar results are
obtained for the spaces Σµ
ν (Rn), µ, ν > 0, µ+ν > 1, cf. (1.2). In contrast to the
symmetric case µ = ν and k = m (classical Shubin operators) we encounter
resonance type phenomena involving the ratio κ := µ/ν; namely we obtain
a characterization of S µ
ν (Rn) in the case µ = kt/(k + m), ν =
mt/(k + m), t ≥ 1, that is, when κ = k/m ∈ Q.
ν (Rn) and Σµ
1. Introduction and statement of the results
The main goal of the paper is to prove results on the characterization of the
ν (Rn), µ, ν > 0, µ + ν ≥ 1 by
non-symmetric (µ 6= ν) Gelfand -- Shilov spaces Sµ
Gevrey estimates of the L2 norms of the iterates P ℓu, ℓ = 1, 2, . . . , u ∈ S (Rn),
of positive anisotropic globally elliptic Shubin differential operators P of the type
(m, k), m, k being even natural numbers, and on the decay of the Fourier coefficients
j=1 stands
for an orthonormal basis of eigenfunctions associated to the operator P . The (m, k)
Shubin elliptic differential operators are modelled by
uj, j ∈ N, in the eigenfunction expansions u = P∞
j=1 ujϕj, where {ϕj}∞
(1.1)
Hm,k
n
:= (−∆)m/2 + xk,
x = qx2
1 + . . . + x2
n, k, m ∈ 2N.
We recall that for µ > 0, ν > 0, the inductive (respectively, projective) Gelfand-
Shilov classes Sµ
ν (Rn), µ + ν > 1), are defined as
the set of all u ∈ S (Rn) for which there exist A > 0, C > 0 (respectively, for every
A > 0 there exists C > 0) such that
ν (Rn), µ + ν ≥ 1 (respectively, Σµ
(1.2)
xβ∂α
x u(x) ≤ CAα+β(α!)µ(β!)ν , α, β ∈ Nn,
see [2, 12, 14, 17, 25] and [27, Chapter 6]. These spaces have recently gained a wide
importance in view of the fact that they represent a suitable functional setting
both for microlocal analysis and PDE and for Fourier and time-frequency analysis
[1, 3, 6 -- 10, 13, 20, 35].
Concerning the investigation in the present paper, we can cite different sources of
motivations. First, we recall the fundamental work of Seeley [33] on eigenfunction
expansions of real analytic functions on compact manifolds (see also the recent
paper of Dasgupta and Ruzhansky [15], extending the result of [33] for all Gevrey
2010 Mathematics Subject Classification. Primary 46F05; Secondary 34L10, 47F05.
Key words and phrases. anisotropic Shubin-type operators, Gelfand-Shilov spaces, eigenfunc-
tion expansions.
1
2 MARCO CAPPIELLO, TODOR GRAMCHEV, STEVAN PILIPOVIC, AND LUIGI RODINO
spaces Gσ, σ > 1, on compact Lie groups). Secondly, we mention the work [19]
µ (Rn) by means of
on the characterization of symmetric Gelfand-Shilov spaces Sµ
estimates of iterates and the decay of the Fourier coefficients in the eigenfunction
expansions associated to globally elliptic (or Γ elliptic) differential operator. We also
refer to [37], where general Gevrey sequences Mp are used. Finally, we mention as
additional motivation the results on hypoellipticity in Sµ
ν (Rn) for elliptic operators
of the type Hm,k
for µ ≥ k/(m + k), ν ≥ m/(m + k), k, m being even natural
numbers, cf. [7] (see also the older work [6]).
n
Before stating our main results we need some preliminaries.
As counterpart of an elliptic operator in a compact manifold, we consider in
Rn the decay of the Fourier coefficients in the eigenfunction expansions associated
to Hm,k
n . In contrast to the symmetric case µ = ν and k = m (classical Shubin
operators) we encounter new resonance type phenomena involving κ := µ/ν, namely
we can characterize the spaces Sµ
ν (Rn), µ + ν > 1)
by iterates and eigenfunction expansions defined by Hm,k
iff κ is rational number,
κ = k/m.
ν (Rn), µ + ν ≥ 1 (respectively Σµ
n
Our basic example of operator will be the anisotropic quantum harmonic oscil-
lator appearing in Quantum Mechanics
(1.3)
H2,k
n = −△ + xk,
k ∈ 2N,
with recovering for k = 2 the standard harmonic oscillator whose eigenfunctions
are the Hermite functions
(1.4)
hα(x) = Hα(x)e−x2/2, α = (α1, ..., αn) ∈ Nn,
where Hα(x) is the α-th Hermite polynomial. See for example [24,29,31] for related
Hermite expansions as well as [18, 38] for connections with a degenerate harmonic
oscillator.
Here we shall consider a more general class of operators with polynomial coeffi-
cients in Rn, namely (m, k) anisotropic operators:
(1.5)
Set
(1.6)
P = Xα
m + β
k ≤1
cαβxβ Dα
x , Dα = (−i)α∂α
x .
Λm,k(x, ξ) = (1 + x2k + ξ2m)1/2,
(x, ξ) ∈ R2n, m, k ∈ 2N.
The global ellipticity for P in (1.5) is defined by imposing
(1.7)
Xα
m + β
k =1
cαβxβξα 6= 0 for (x, ξ) 6= (0, 0).
or equivalently, there exist C1 > 0, C2 > 0, R > 0 such that
(1.8)
C2 ≤
p(x, ξ)
Λm,k(x, ξ)
≤ C1,
(x, ξ) ≥ R.
Under the assumption (1.7) (or (1.8)), the following estimate holds for every u ∈
S (Rn):
(1.9)
cf. [4].
kxβDα
x ukL2 ≤ C(kP ukL2 + kukL2),
Xα
m + β
k ≤1
ANISOTROPIC SHUBIN OPERATORS AND EIGENFUNCTION EXPANSIONS IN GELFAND-SHILOV SPACES3
For these operators, the counterpart of the standard Sobolev spaces are the
spaces Qs
m,k(Rn), s ∈ R, defined, for example, by requiring that
(1.10)
where
(1.11)
kΛ(x, D)sukL2 < ∞,
Λ(x, ξ) = (1 + x2k + ξ2m)1/2 max{k,m},
k, m ∈ 2N.
Under the global ellipticity assumption (1.7),
P : Qs
m,k(Rn) → L2(Rn), s = max{k, m},
is a Fredholm operator. The finite-dimensional null-space Ker P is given by func-
tions in the Schwartz space S (Rn).
We assume, as in [19], that P is a positive anisotropic elliptic operator, which
implies that k and m are even numbers. This guarantees the existence of an or-
thonormal basis of eigenfunctions ϕj, j ∈ N, with eigenvalues λj,
λj = +∞
lim
j→∞
(see [34]). Moreover we have that
(1.12)
λj ∼ Cj
mk
n(m+k)
as
j → +∞.
for some C > 0, cf. [4,34]. Hence, given u ∈ L2(Rn), or u ∈ S ′(Rn), we can expand
(1.13)
u =
∞
Xj=1
ujϕj
where the Fourier coefficients uj ∈ C are defined by
(1.14)
uj = (u, uj)L2, j = 1, 2, . . .
with convergence in L2(Rn) or S ′(Rn) for (1.13).
By the hypoellipticity results of [7] the eigenfunctions ϕj belong to Sk/(m+k)
m/(m+k)(Rn).
We first state an assertion on the characterization of the anisotropic Sobolev
spaces Qs
m,k(Rn) and the Schwartz class S (Rn).
Theorem 1.1. Suppose that P is (m, k)-globally elliptic cf. (1.5), (1.7), and posi-
tive. Then:
(i) u ∈ Qs
m,k(Rn) ⇐⇒
uj2λs/ max{m,k}
j
< ∞, s ∈ N.
∞
Pj=1
(ii) u ∈ S (Rn) ⇐⇒ uj = O(λ−s
j ), j → ∞ ⇐⇒ uj = O(j−s), j → ∞ for all
s ∈ N.
ν (Rn) and Σµ
Let us now come to the characterization of the spaces Sµ
ν (Rn) in
the case κ := µ/ν ∈ Q. We may link µ, ν with an operator of the form (1.5) for a
suitable choice of k and m. In fact, observe first that we may write µ = tµo, ν = tνo
for some t > 0 with µo = κ/(1 + κ), ν0 = 1/(1 + κ) so that µo + νo = 1. If µ + ν ≥ 1
we have t ≥ 1, if µ + ν > 1 then t > 1. On the other hand, for any given µo ∈ Q
we may write µo = k/(k + m) for two positive integers k and m, and consequently
νo = 1 − µo = m/(k + m). Multiples of k and m work as well, in particular we may
assume k and m to be even natural numbers so that the symbol of Λm,k in (1.6)
is a smooth function which is necessary for the proof of the hypoellipticity result
of [7]. So we have
µ =
kt
k + m
,
ν =
mt
k + m
.
4 MARCO CAPPIELLO, TODOR GRAMCHEV, STEVAN PILIPOVIC, AND LUIGI RODINO
For given even integers k and m, an example of globally elliptic positive operator
is given by (1.1).
The first main result of the paper characterizes the Gelfand-Shilov spaces in
terms of estimates of the iterates of P and reads as follows.
Theorem 1.2. Let P be an operator of the form (1.5) for some integers k ≥
1, m ≥ 1, be globally elliptic, namely satisfy (1.7) and let u ∈ S (Rn). Then
(Rn), t > 1) if and only if there exist
(Rn), t ≥ 1 (respectively u ∈ Σ
k+m
u ∈ S
k+m
mt
kt
kt
mt
k+m
k+m
C > 0, R > 0 (respectively for every C > 0 there exists R > 0) such that:
(1.15)
kP M ukL2 ≤ RCM (M !)
kmt
k+m
for every integer M ≥ 1.
Remark 1.3. Theorem 1.2 suggests the possibility of considering new function spaces
defined by the estimates (1.15) also for 0 < t < 1 (respectively 0 < t ≤ 1).
Corresponding Gelfand-Shilov classes are empty in that case as well known from [17]
and the equivalence in Theorem 1.2 fails. Nevertheless such definition in terms of
(1.15) deserves interest, cf. also [11, 36].
Using Theorem 1.2 we can prove the following result.
Theorem 1.4. Let P be a positive operator of the form (1.5) for some integers
k ≥ 1, m ≥ 1, satisfying (1.7) and let u ∈ S (Rn). Let the eigenvalues λj and the
Fourier coefficients uj be defined as before. The following conditions are equivalent:
i) u ∈ S
kt
k+m
mt
k+m
(Rn), t ≥ 1 (respectively u ∈ Σ
kt
k+m
mt
k+m
(Rn), t > 1);
ii) there exists ε > 0 such that (respectively for every ε > 0) we have
(1.16)
uj2eǫλ
k+m
kmt
j
< ∞;
∞
Xj=1
iii) there exists ε > 0 such that (respectively for every ε > 0) we have
(1.17)
uj2eǫλ
k+m
kmt
j
< ∞.
sup
j∈N
iv) there exists ε > 0 such that (respectively for every ε > 0) we have for some
C > 0:
uj ≤ Ce−εj
tn ,
1
j ∈ N.
The somewhat surprising fact that in iv) the estimates do not depend on the
ν regularity
couple (m, k), that is on (µ, ν), may find intuitive explanation in the Sµ
of the eigenfunctions ϕj , cf. [7].
2. Proof of the main results
Proof of Theorem 1.1. The proof of Theorem 1.1 is easy, by using the r-th power
of P, r ∈ R, that we may define as
P ru =
∞
Xj=1
λr
j ujϕj ,
ANISOTROPIC SHUBIN OPERATORS AND EIGENFUNCTION EXPANSIONS IN GELFAND-SHILOV SPACES5
and by observing that the norms kP rukL2, r = s/ max{k, m} and kΛ(x, D)sukL2
are equivalent, see [4, 27, 34]. On the other hand, by Parseval identity
kP ruk2
Xj=1
and i) follows. Since S (Rn) = Ts∈N
L2 = k
∞
λr
j ujϕjk2
L2 =
∞
Xj=1
λ2r
j uj2
Qs
m,k(Rn) we also obtain ii).
(cid:3)
The proof of Theorem 1.2 needs some preparation. We first define, for fixed
r ≥ 0 and u ∈ L2(Rn):
(2.1)
ur = Xα
m + β
k =r
kxβDαukL2
First it is useful to characterize Gelfand-Shilov spaces in terms of the norms ur
as follows.
Proposition 2.1. Let u ∈ L2(Rn). Then u ∈ S
kt
k+m
mt
k+m
(Rn), t ≥ 1 (respectively u ∈
Σ
kt
k+m
mt
k+m
(Rn), t > 1) if and only if there exist C > 0, R > 0 (respectively for every
C > 0 there exists R > 0) such that
(2.2)
for every r > 0.
ur ≤ RCrr
kmrt
k+m
We have the following preliminary result.
Lemma 2.2. There exists a constant C > 0 such that, for any given p ∈ N, (α, β) ∈
N2n, with α/m + β/k = r, p < r < p + 1, and for every ε > 0, the following
estimate holds true:
(2.3)
for all u ∈ S (Rn).
ur ≤ εup+1 + Cε− r−p
p+1−r up + Cp(p + 1)!
km
k+m u0
The proof follows the same lines as the proof of Proposition 2.1 in [5], cf. also [23],
and it is omitted.
Next, fixed λ > 0, p ∈ N and u ∈ L2(Rn), we set:
σp(u, λ) = λ−p(p!)− kmt
k+m up.
(2.4)
Lemma 2.3. For every p ∈ N and for λ > 0 sufficiently large, we have:
(2.5)
σp+1(u, λ) ≤ (p + 1)− kmt
k+m σp(P u, λ) +
p
Xh=0
σh(u, λ)
for every u ∈ S (Rn).
Proof. For p = 0 the assertion is a direct consequence of (1.9) if λ is large enough.
Fix now p ∈ N, p ≥ 1 and let α, β ∈ Nn such that α/m + β/k = p + 1. It is easy
to verify that we can find γ, δ ∈ Nn, with γ ≤ α, δ ≤ β such that γ/m + δ/k = p
and α − γ/m + β − δ/k = 1. Then by (1.9) we can write
kxβDαukL2 ≤ kxβ−δDα−γ(xδDγu)kL2 + kxβ−δ[xδ, Dα−γ]DγukL2
≤ CkP (xδDγu)kL2 + kxβ−δ[xδ, Dα−γ]DγukL2
≤ I1 + I2 + I3,
6 MARCO CAPPIELLO, TODOR GRAMCHEV, STEVAN PILIPOVIC, AND LUIGI RODINO
where
I1 = CkxδDγ(P u)kL2,
Let now
Jh = Xα
m + β
k =p+1
Then, obviously we have
I2 = Ck[P, xδDγ]ukL2,
I3 = kxβ−δ[xδ, Dα−γ]DγukL2.
Ih,
Yh = λ−p−1(p + 1)!− kmt
k+m Jh,
h = 1, 2, 3.
up+1 ≤ J1 + J2 + J3,
σp+1(λ, u) ≤ Y1 + Y2 + Y3.
Now, since J1 ≤ C1P up for some C1 > 0, then we have Y1 ≤ (p+1)− kmt
if λ ≥ C−1
1 . To estimate J2 and Y2 we observe that
k+m σp(λ, P u),
[P, xδDγ]u = X α
m + β
k ≤1
c α β[x
βD α, xδDγ]u,
and that
[x
C αδτ xδ+ β−τ Dγ+ α−τ u − X06=τ ≤ β,τ ≤γ
βD α, xδDγ]u = X06=τ ≤ α,τ ≤δ
where the constants C αδτ and C βγτ can be estimated by C2 pτ for some positive
constant C2 independent of p. We observe now that in both the sums above we
have
C βγτ xδ+ β−τ Dγ+ α−τ u.
γ + α − τ
δ + β − τ
r =
+
= p +
+
−
τ ≤ p + 1 −
km
hence in particular we have 0 ≤ r < p + 1 since τ > 0. Moreover, we have
km
m
k
α
m
β
k
m + k
m + k
τ ,
In view of these considerations, we easily obtain
τ ≤
km
m + k
(p + 1 − r).
J2 ≤ C3(J ′
2 + p
km
k+m up + J ′′
2 ),
where
J ′
2 = Xp<r<p+1
2 = X0≤r<p
J ′′
km
k+m (p+1−r)ur,
p
km
k+m (p+1−r)ur.
p
Now, applying Lemma 2.2 to J ′
2 with
and using standard factorial inequalities we obtain
ε = (4C3)−1p− km
k+m (p+1−r),
J ′
2 ≤ (4C3)−1up+1 + C4p
Similarly, writing
km
k+m up + Cp+1
5
(p + 1)!
km
k+m u0.
J ′′
2 = p
km
k+m (p+1)u0 +
p−1
Xq=0 Xq<r<q+1
km
k+m (p+1−r)ur
p
and applying Lemma 2.2 to each term of the sum above with
ε = p− km
k+m (q+1−r),
ANISOTROPIC SHUBIN OPERATORS AND EIGENFUNCTION EXPANSIONS IN GELFAND-SHILOV SPACES7
we get
2 ≤ Cp+1
J ′′
6
(p + 1)!
km
k+m u0 + C7
≤ Cp+1
8
(p + 1)!
km
k+m u0 + C9
from which we get
p−1
Xq=0hp
Xq=1
p
p
km
k+m (p−q)uq+1 + p
km
k+m (p−q+1)uqi
km
k+m (p−q+1)uq,
J2 ≤
1
4
up+1 + Cp+1(p + 1)!
km
k+m u0 + C′
p
Xq=1
km
k+m (p−q+1)uq
p
for some positive constants C′, C independent of p. From the estimates above,
taking λ sufficiently large and using the fact that t ≥ 1, we obtain
Y2 = λ−p−1(p + 1)!− kmt
k+m J2 ≤
1
4
p+1
Xh=0
σh(λ, u).
Analogous estimates can be derived for Y3 and yield (2.5). We leave the details for
the reader.
(cid:3)
Starting from (2.5) and arguing by induction on p it is easy to prove the following
result. We omit the proof for the sake of brevity.
Lemma 2.4. For every p ∈ N, t ≥ 1 and λ > 0 sufficiently large we have
σp(u, λ) ≤ 2pσ0(u, λ) +
p
Xℓ=1
2p−ℓ(cid:18)p
ℓ(cid:19)(ℓ!)− kmt
k+m σ0(P ℓu, λ).
Proof of Theorem 1.2. The fact that the Gelfand-Shilov regularity of u implies
(1.15) is easy to prove and we omit the details.
In the opposite direction, by
Proposition 2.1 it is sufficient to prove that u satisfies (2.2) for every r > 0. From
the previous estimate, we have, for every p ∈ N:
σp(u, λ) ≤ C +
p
Xℓ=1
2p−ℓ(cid:18)p
ℓ(cid:19)Cℓ+1 ≤ C(2 + C)p+1.
Therefore
up ≤ Cp+1p!
kmt
k+m
for a new constant C > 0, which gives (2.2) in the case r ∈ N. If r > 0 is not
integer, then p < r < p + 1 for some p ∈ N and we can apply Lemma 2.2 which
yields
ur ≤ εup+1 + Cε− r−p
k+m u0
km
p+1−r up + Cp(p!)
k+m + Cp
1 ε− r−p
kmt
≤ εCp+1
1
(p + 1)!
p+1−r (p + 1)!
kmt
k+m + Cp
1 (p + 1)!
kmt
k+m ≤ Cr+1
2
kmrt
k+m .
r
Then, by Proposition 2.1 we conclude that u ∈ S
u ∈ Σ
kt
k+m
mt
k+m
(Rn).
kt
k+m
mt
k+m
(Rn). Similarly we argue for
(cid:3)
Proof of Theorem 1.4. The equivalence between ii) and iii) is obvious. Moreover
(Rn)
iii) is equivalent to iv) in view of (1.12). The arguments are similar for S
k+m
mt
kt
k+m
8 MARCO CAPPIELLO, TODOR GRAMCHEV, STEVAN PILIPOVIC, AND LUIGI RODINO
and Σ
kt
k+m
mt
k+m
(Rn) classes. To conclude the proof we will show the equivalence between
i) and iv). We first observe that
kP M uk2
L2 = k
∞
Xj=1
ujP M ϕj k2
L2 =
∞
Xj=1
λ2M
j
uj2,
in view of Parseval identity. By (1.12) it follows that
(2.6)
C1kP M uk2
L2 ≤
∞
Xj=1
j2Mkm/(n(k+m))uj2 ≤ C2kP M uk2
L2
for suitable positive constants C1, C2. Now if iv) holds, then we have
uj2 ≤ e−ǫj1/(nt)
for some new constant ǫ > 0. Then from the first estimate in (2.6) we have for
some C > 0
(2.7)
(2.8)
with
∞
kP M uk2
L2 ≤ C
Xj=1
≤ C sup
j∈N
j2Mkm/(n(m+k))e−ǫj1/(nt)
j2Mmk/(n(m+k))e−ǫj1/(nt)
C = C
∞
Xj=1
e−ǫj1/(nt)
.
Moreover, for any fixed ω > 0 we have
eωj1/(nt)
=
∞
XM=0
ωM jM/(nt)
M !
.
This implies that for every M ∈ N:
(2.9)
jM/(nt)e−ωj1/(nt)
≤ ω−M M !
Taking the 2kmt/(k + m)-th power of both sides of (2.9) and applying in the last
estimate in (2.8) with
ω = 2ǫkmt/(k + m),
we obtain
kP M uk2
L2 ≤ Cω− 2M kmt
k+m (M !)
2mkt
m+k ,
which gives i) in view of Theorem 1.2.
i) ⇒ ii) Viceversa assume that u ∈ S
kt
k+m
mt
k+m
(Rn). In view of iv) it is sufficient to show
that
(2.10)
uj2eǫj
1
nt < +∞.
sup
j∈N
Theorem 1.2 and the second inequality in (2.6) imply that
2M km
n(k+m)
j
CM (M !)
2kmt
k+m
uj2 ≤ C
ANISOTROPIC SHUBIN OPERATORS AND EIGENFUNCTION EXPANSIONS IN GELFAND-SHILOV SPACES9
for every j, M ∈ N and for some C independent of j and M . Taking the supremum
of the left-hand side over M we get (2.10) with ǫ = 2kmt
2kmt . This concludes
the proof.
(cid:3)
k+m C− k+m
3. Generalizations
We list some possible generalizations of the preceding results. First, one can
replace the hypothesis of positivity for the operator P by assuming that P is nor-
mal, i.e. P ∗P = P P ∗. This guarantees the existence of an orthonormal basis of
eigenfunctions ϕj, j ∈ N, with eigenvalues λj, lim
λj = +∞, see [34], and we may
j→∞
then proceed as before, cf. [33].
Another possible generalization consists in replacing L2 norms with Lp norms,
1 < p < ∞. Let us observe that the basic estimate (1.9) is valid also for Lp norms,
see [16, 26], and it seems easy to extend Theorem 1.2 in this direction.
A much more challenging problem is an analogous characterization of the classes
ν (Rn) when κ = µ/ν = k/m is irrational. First difficulty, in this case, is given by
Sµ
an appropriate choice of the operator P . In fact, the natural candidates
P = (−∆)m/2 + (1 + x2)k/2,
m ∈ 2N, k > 0, k /∈ 2N
can be easily treated in the setting of temperate distributions but results of Gelfand-
Shilov regularity, extending those in [7], are missing for them.
Note. With great sorrow, Marco Cappiello, Stevan Pilipovic and Luigi Rodino
inform that their friend Todor Gramchev passed away on October 17, 2015. He
inspired and collaborated to the initial version of the present paper and appears
here as co-author.
References
1. A.Ascanelli, M.Cappiello, Hölder continuity in time for SG hyperbolic systems, J. Differential
Equations 244 (2008), 2091 -- 2121.
2. A. Avantaggiati, S-spaces by means of the behaviour of Hermite-Fourier coefficients, Boll.
Un. Mat. Ital. 6 (1985), 487 -- 495.
3. H.A. Biagioni, T. Gramchev, Fractional derivative estimates in Gevrey spaces, global regu-
larity and decay for solutions to semilinear equations in Rn, J. Differential Equations 194
(2003), 140 -- 165.
4. P. Boggiatto, E. Buzano, L. Rodino, Global hypoellipticity and spectral theory. Math. Res. 92,
Akademie Verlag, Berlin, 1996.
5. D. Calvo, L. Rodino, Iterates of operators and Gelfand-Shilov functions, Int. Transf. Spec.
Funct. 22 (2011), 269 -- 276.
6. M. Cappiello, T. Gramchev, L. Rodino, Super-exponential decay and holomorphic extensions
for semilinear equations with polynomial coefficients. J. Funct. Anal. 237 (2006), 634 -- 654.
7. M. Cappiello, T. Gramchev, L. Rodino, Entire extensions and exponential decay for semilinear
elliptic equations, J. Anal. Math. 111 (2010), 339 -- 367.
8. M. Cappiello, T. Gramchev, L. Rodino, Sub-exponential decay and uniform holomorphic ex-
tensions for semilinear pseudodifferential equations, Comm. Partial Differential Equations 35
(2010), n. 5, 846-877.
9. M. Cappiello, L. Rodino, SG-pseudo-differential operators and Gelfand-Shilov spaces, Rocky
Mountain J. Math.36 (2006) n. 4, 1117 -- 1148.
10. M. Cappiello, J. Toft, Pseudo-differential operators in a Gelfand-Shilov setting, Math. Nachr.
(2016). To appear.
11. Y. Chen, M. Signahl, J. Toft, Factorizations and singular value estimates of operators with
Gelfand -- Shilov and Pilipović kernels, arXiv:1511.06257 (2016).
10 MARCO CAPPIELLO, TODOR GRAMCHEV, STEVAN PILIPOVIC, AND LUIGI RODINO
12. J. Chung, S. Y. Chung, D. Kim, Characterization of the Gelfand-Shilov spaces via Fourier
transforms, Proc. Am. Math. Soc. 124 (1996), 2101 -- 2108.
13. E. Cordero, F. Nicola, L. Rodino, Wave packet analysis of Schrödinger equations in analytic
function spaces, Adv. Math. 278 (2015), 182 -- 209.
14. E. Cordero, S. Pilipović, L. Rodino, N. Teofanov, Localization operators and exponential
weights for modulation spaces, Mediterranean J. Math. 2 (2005), 381 -- 394.
15. A. Dasgupta, M. Ruzhansky, Eigenfunction expansions of ultradifferentiable func-
to appear. Available at
tions and ultradistributions, Trans. Amer. Math. Soc.,
https://arxiv.org/abs/1410.2637.
16. G. Garello, A. Morando, Lp-bounded pseudo-differential operators and regularity for multi-
quasi-elliptic equations, Integral Equations Operator Theory 51 (2005), 501 -- 517.
17. I.M. Gelfand, G.E. Shilov, Generalized functions II. Academic Press, New York, 1968.
18. T. Gramchev, S. Pilipović, L. Rodino, Global Regularity and Stability in S-Spaces for Classes
of Degenerate Shubin Operators. Pseudo-Differential Operators: Complex Analysis and Partial
Differential Equations Operator Theory: Advances and Applications 205 (2010), 81-90.
19. T. Gramchev, S. Pilipović, L. Rodino, Eigenfunction expansions in Rn, Proc. Amer. Math.
Soc. 139 (2011), 4361 -- 4368.
20. K. Gröchenig, G. Zimmermann, Spaces of test functions via the STFT, J. Funct. Spaces Appl.
2 (2005), 1671 -- 1716.
21. B. Helffer, Théorie spectrale pour des opérateurs globalement elliptiques. Astérisque 112,
Société Mathématique de France, Paris, 1984.
22. H. Komatsu, A proof of Kotake and Narashiman's Theorem. Proc. Japan Acad. 38 (1962),
615-618.
23. T. Kotake, M.S. Narasimhan, Regularity theorems for fractional powers of a linear elliptic
operator. Bull. Soc. Math. France, 90 (1962), 449-471.
24. M. Langenbruch, Hermite functions and weighted spaces of generalized functions. Manuscripta
Math. 119 (2006), 269 -- 285.
25. B.S. Mitjagin, Nuclearity and other properties of spaces of type S, Amer. Math. Soc. Transl.,
Ser. 2, 93 (1970), 45 -- 59.
26. A. Morando, Lp-regularity for a class of pseudo-differential operators in Rn, J. Partial Dif-
ferential Equations 18 (2005), 241 -- 262.
27. F. Nicola, L. Rodino, Global pseudo-differential calculus on Euclidean spaces, Birkhauser,
Basel, 2010.
28. S. Pilipović, Generalization of Zemanian spaces of generalized functions which have orthonor-
mal series expansions, SIAM J. Math. Anal. 17 (1986), 477-484.
29. S. Pilipović, Tempered ultradistributions. Boll. Unione Mat. Ital. VII. Ser. B 2 (1988), 235 -- 251.
30. S. Pilipović, N. Teofanov, Pseudodifferential operators on ultramodulation spaces, J. Funct.
Anal. 208 (2004), 194 -- 228.
31. M. Reed, B. Simon, Methods of modern mathematical physics Vol 1. Academic Press, San
Diego Ca., 1975.
32. R.T. Seeley, Integro-differential operators on vector boundes, Trans. Am. Math. Soc. 117
(1965), 167-204.
33. R.T. Seeley, Eigenfunction expansions of analytic functions, Proc. Am. Math. Soc. 21 (1969),
734 -- 738.
34. M. Shubin, Pseudodifferential operators and spectral theory. Springer Series in Soviet Math-
ematics, Springer Verlag, Berlin, 1987.
35. J. Toft, Multiplication properties in Gelfand-Shilov pseudo-differential calculus. In: Molahajlo,
S., Pilipović, S., Toft, J., Wong, M.W. (eds.) Pseudo-Differential Operators, Generalized
Functions and Asymptotics, Operator Theory: Advances and Applications, Birkhüser, Basel,
231 (2013), 117-172.
36. J. Toft, Images of function and distribution spaces under the Bargmann transform, J. Pseudo-
Differ. Oper. Appl. DOI 10.1007/s11868-016-0165-9 (2016).
37. J. Vindas, Dj. Vuckovic, Eigenfunction expansions of ultradifferentiable functions and ultra-
distributions in Rn arXiv:1512.01684 (2016).
38. M.W. Wong, The heat equation for the Hermite operator on the Heisenberg group. Hokkaido
Math. J. 34 (2005), 393 -- 404.
ANISOTROPIC SHUBIN OPERATORS AND EIGENFUNCTION EXPANSIONS IN GELFAND-SHILOV SPACES11
Dipartimento di Matematica, Università di Torino, Via Carlo Alberto 10, 10123
Torino, Italy
E-mail address: [email protected]
Dipartimento di Matematica e Informatica, Università di Cagliari, Via Ospedale
72, 09124 Cagliari, Italy
Institute of Mathematics, University of Novi Sad, trg. D. Obradovica 4, 21000
Novi Sad, Serbia
E-mail address: [email protected]
Dipartimento di Matematica, Università di Torino, Via Carlo Alberto 10, 10123
Torino, Italy
E-mail address: [email protected]
|
1708.01231 | 2 | 1708 | 2018-05-19T09:02:22 | Optimal constants for a non-local approximation of Sobolev norms and total variation | [
"math.FA",
"math.OC"
] | We consider the family of non-local and non-convex functionals proposed and investigated by J. Bourgain, H. Brezis and H.-M. Nguyen in a series of papers of the last decade. It was known that this family of functionals Gamma-converges to a suitable multiple of the Sobolev norm or the total variation, depending on the summability exponent, but the exact constants and the structure of recovery families were still unknown, even in dimension one.
We prove a Gamma-convergence result with explicit values of the constants in any space dimension. We also show the existence of recovery families consisting of smooth functions with compact support.
The key point is reducing the problem first to dimension one, and then to a finite combinatorial rearrangement inequality. | math.FA | math | Optimal constants for a non-local approximation of
Sobolev norms and total variation
Clara Antonucci
Massimo Gobbino
Scuola Normale Superiore
Universit`a degli Studi di Pisa
PISA (Italy)
PISA (Italy)
e-mail: [email protected]
e-mail: [email protected]
Matteo Migliorini
Nicola Picenni
Scuola Normale Superiore
Scuola Normale Superiore
PISA (Italy)
PISA (Italy)
e-mail: [email protected]
e-mail: [email protected]
8
1
0
2
y
a
M
9
1
]
.
A
F
h
t
a
m
[
2
v
1
3
2
1
0
.
8
0
7
1
:
v
i
X
r
a
Abstract
We consider the family of non-local and non-convex functionals proposed and in-
vestigated by J. Bourgain, H. Brezis and H.-M. Nguyen in a series of papers of the last
decade. It was known that this family of functionals Gamma-converges to a suitable
multiple of the Sobolev norm or the total variation, depending on the summability expo-
nent, but the exact constants and the structure of recovery families were still unknown,
even in dimension one.
We prove a Gamma-convergence result with explicit values of the constants in any
space dimension. We also show the existence of recovery families consisting of smooth
functions with compact support.
The key point is reducing the problem first to dimension one, and then to a finite
combinatorial rearrangement inequality.
Mathematics Subject Classification 2010 (MSC2010): 26B30, 46E35.
Key words: Gamma-convergence, Sobolev spaces, bounded variation functions, mono-
tone rearrangement, non-local functional, non-convex functional.
1
Introduction
where
Let p ≥ 1 and δ > 0 be real numbers, let d be a positive integer, and let Ω ⊆ Rd be an
open set. For every measurable function u : Ω → R we set
(1.1)
δp
Λδ,p(u, Ω) :=ZZI(δ,u,Ω)
y − xd+p dx dy,
I(δ, u, Ω) :=(cid:8)(x, y) ∈ Ω2 : u(y) − u(x) > δ(cid:9) .
This family of non-convex and non-local functionals was introduced, motivated and
investigated in a series of papers by H.-M. Nguyen [13, 14, 15, 16, 17], J. Bourgain and
H.-M. Nguyen [4], H. Brezis and H.-M. Nguyen [8] (see also [6] and [7]).
We point out that the dependence on u is just on the integration set. The fixed
integrand is divergent on the diagonal y = x, and the integration set is closer to the
diagonal where the gradient of u is large. This suggests that Λδ,p(u, Ω) is proportional,
in the limit as δ → 0+, to some norm of the gradient of u, and more precisely to the
functional
Λ0,p(u, Ω) :=
ZΩ ∇u(x)p dx
total variation of u in Ω
+∞
if p > 1 and u ∈ W 1,p(Ω),
if p = 1 and u ∈ BV (Ω),
otherwise.
(1.2)
It is natural to compare the family (1.1) with the classical approximations of Sobolev
or BV norms, based on non-local convex functionals such as
Gε,p(u, Ω) :=ZZΩ
u(y) − u(x)p
y − xp
ρε(y − x) dx dy,
(1.3)
where gradients are replaced by finite differences weighted by a suitable family ρε of
mollifiers. The idea of approximating integrals of the gradient with double integrals of
difference quotients, where all pairs of distinct points interact, has been considered in-
dependently by many authors in different contexts. For example, E. De Giorgi proposed
an approximation of this kind to the Mumford-Shah functional in any space dimension,
in order to overcome the anisotropy of the discrete approximation [9]. The resulting
theory was put into paper in [11], and then extended in [12] to more general free dis-
continuity problems, and in particular to Sobolev and BV spaces. In the same years,
the case of Sobolev and BV norms was considered in details in [3] (see also [18]).
The result, as expected, is that the family Gε,p(u, Rd) converges as ε → 0+ to a
suitable multiple of Λ0,p(u, Rd), both in the sense of pointwise convergence, and in the
sense of De Giorgi's Gamma-convergence. This provides a characterization of Sobolev
functions (if p > 1), and of bounded variation functions (if p = 1), as those functions
for which the pointwise limit or the Gamma-limit is finite.
1
From the heuristic point of view, the non-convex approximating family (1.1) seems
to follow a different paradigm. Indeed, it has been observed by J.M. Morel (as quoted
at page 4 of the transparencies of the conference [5]) that this definition involves some
sort of "vertical slicing" that evokes the definition of integral `a la Lebesgue, in contrast
to the definition `a la Riemann that seems closer to the "horizontal slicing" of the finite
differences in (1.3).
From the mathematical point of view, the asymptotic behavior of (1.1) exhibits
some unexpected features. In order to state the precise results, let us introduce some
notation. Let Sd−1 := {σ ∈ Rd : σ = 1} denote the unit sphere in Rd. For every p ≥ 1
we consider the geometric constant
Gd,p :=ZSd−1 hv, σip dσ,
(1.4)
where v is any element of Sd−1 (of course the value of Gd,p does not depend on the
choice of v), and the integration is intended with respect to the (d − 1)-dimensional
Hausdorff measure.
The main convergence results obtained so far can be summed up as follows.
• Pointwise convergence for p > 1. For every p > 1 it turns out that
∀u ∈ Lp(Rd).
Gd,p Λ0,p(u, Rd)
1
p
lim
δ→0+
Λδ,p(u, Rd) =
(1.5)
• Pointwise convergence for p = 1.
In the case p = 1 equality (1.5) holds true
c (Rd), but there do exist functions u ∈ W 1,1(Rd) for which the
for every u ∈ C 1
left-hand side is infinite (while of course the right-hand side is finite). A precise
characterization of equality cases is still unknown.
• Gamma-convergence for every p ≥ 1. For every p ≥ 1 there exists a constant Cd,p
such that
Γ– lim
δ→0+
Λδ,p(u, Rd) =
1
p
Gd,pCd,p Λ0,p(u, Rd)
∀u ∈ Lp(Rd),
where the Gamma-limit is intended with respect to the usual metric of Lp(Rd)
(but the result would be the same with respect to the convergence in L1(Rd) or
in measure). Moreover, it was proved that Cd,p ∈ (0, 1), namely the Gamma-limit
is always nontrivial but different from the pointwise limit.
As a consequence, again one can characterize the Sobolev space W 1,p(Rd) as the set
of functions in Lp(Rd) for which the pointwise limit or the Gamma-limit are finite. As
for BV (Rd), in this setting it can be characterized only through the Gamma-limit.
Some problems remained open, and were stated explicitly in [16, 8].
• Question 1. What is the exact value of Cd,p, at least in the case d = 1?
2
• Question 2. Does Cd,p depend on d?
• Question 3. Do there exist recovery families made up of continuous functions, or
even of functions of class C ∞?
In this paper we answer these three questions. Concerning question 1 and 2, we
prove that Cd,p does not depend on d, and coincides with the value Cp conjectured
in [14] for the one-dimensional case, namely
1
p − 1(cid:18)1 −
1
2p−1(cid:19)
log 2
if p > 1,
if p = 1.
(1.6)
Cp :=
Concerning the third question, we prove that smooth recovery families do exist. Our
main result is the following.
Theorem 1.1 (Gamma-convergence). Let us consider the functionals Λδ,p and Λ0,p
defined in (1.1) and (1.2), respectively.
Then for every positive integer d and every real number p ≥ 1 it turns out that
Γ– lim
δ→0+
Λδ,p(u, Rd) =
1
p
Gd,pCp Λ0,p(u, Rd)
∀u ∈ Lp(Rd),
where Gd,p is the geometric constant defined in (1.4), and Cp is the constant defined in
(1.6). In particular, the following two statements hold true.
(1) (Liminf inequality) For every family {uδ}δ>0 ⊆ Lp(Rd), with uδ → u in Lp(Rd)
as δ → 0+, it turns out that
lim inf
δ→0+
Λδ,p(uδ, Rd) ≥
1
p
Gd,pCp Λ0,p(u, Rd).
(1.7)
(2) (Limsup inequality) For every u ∈ Lp(Rd) there exists a family {uδ}δ>0 ⊆ Lp(Rd),
with uδ → u in Lp(Rd) as δ → 0+, such that
1
p
Λδ,p(uδ, Rd) ≤
lim sup
δ→0+
Gd,pCp Λ0,p(u, Rd).
We can also assume that the family {uδ} consists of functions of class C ∞ with
compact support.
The proof of this result requires a different approach to the problem, which we
briefly sketch below. In previous literature the constant Cd,p was defined through some
sort of cell problem as
1
p
Gd,pCd,p := inf(cid:26)lim inf
δ→0+
Λδ,p(cid:0)uδ, (0, 1)d(cid:1) : uδ → u0 in Lp(cid:0)(0, 1)d(cid:1)(cid:27) ,
3
where u0(x) = (x1 + . . . + xd)/√d. Unfortunately, this definition is quite implicit and
provides no informations on the structure of the families that approach the optimal
value. This lack of structure complicates things, in such a way that just proving that
Cd,p > 0 requires extremely delicate estimates (this is the content of [4]). On the
Gamma-limsup side, since Λδ,p is quite sensitive to jumps, what is difficult is glueing
together the recovery families corresponding to different slopes, even in the case of
a piecewise affine function in dimension one. This requires a delicate surgery near
the junctions (see [16]). Finally, as for question 3, difficulties originate from the lack
of convexity or continuity of the functionals (1.1), which do not seem to behave well
under convolution or similar smoothing techniques.
The core of our approach consists in proving that Λδ,p in dimension one behaves well
under vertical δ-segmentation and monotone rearrangement. We refer to Section 3.1 for
the details, but roughly speaking this means that monotone step functions whose values
are consecutive integer multiples of δ are the most efficient way to fill the gap between
any two given levels. The argument is purely one-dimensional, and it is carried out
in Proposition 3.2. In turn, the proof relies on a discrete combinatorial rearrangement
inequality, which we investigate in Theorem 2.2 under more general assumptions.
We observe that this strategy, namely estimating the asymptotic cost of oscillations
by reducing ourselves to a discrete combinatorial minimum problem, is the same ex-
ploited in [11, 12], with the remarkable difference that now the reduction to the discrete
setting is achieved through vertical δ-segmentation, while in [11, 12] it was obtained
through a horizontal ε-segmentation (see Figure 1).
Figure 1: vertical δ-segmentation vs horizontal ε-segmentation (δ is the distance between the
parallel lines on the left, ε is the distance between the parallel lines on the right)
The asymptotic estimate on the cost of oscillations opens the door to the Gamma-
liminf inequality in dimension one, which at this point follows from well established
techniques. As for the Gamma-limsup inequality, in dimension one we just need to
exhibit a family that realizes the given explicit multiple of Λ0,p(u, R), and this can be
achieved through a vertical δ-segmentation `a la Lebesgue (see Proposition 3.7). This
produces a recovery family made up of step functions, and it is not difficult to modify
them in order to obtain functions of class C ∞ with asymptotically the same energy
4
(see Proposition 3.9). Finally, passing from dimension one to any dimension is just
an application of the one-dimensional result to all the one-dimensional sections of a
function of d variables.
At the end of the day, we have a completely self-contained proof of Theorem 1.1
above, and a clear indication that the true difficulty of the problem lies in dimension
one, and actually in the discretized combinatorial model. We hope that these ideas
could be extended to the more general functionals considered in [8]. Some steps in this
direction have already been done in [2] (see also the note [1]).
This paper is organized as follows. In Section 2 we develop a theory of monotone
rearrangements, first in a discrete, and then in a semi-discrete setting. In Section 3
we prove our Gamma-convergence result in dimension one. In Section 4 we prove the
Gamma-convergence result in any space dimension.
We would like to thank an anonymous referee for drawing our attention to the
rearrangement inequalities of [10], which could be used in place of our Theorem 2.4 in
the appropriate point of the proof of Proposition 3.2. It is not difficult to realize that
those rearrangement inequalities, and their discrete combinatorial counterpart known
as Taylor's Lemma (see [19]), are actually equivalent to ours. Keeping this equivalence
into account, we think that our Section 2 could be interesting not only because it makes
the paper as self-contained as possibile, but also because it provides new proofs of some
of the results of [19, 10], starting from a different perspective (`a la Lebesgue here, and `a
la Riemann in previous works). It was surprising to discover that similar combinatorial
problems originated in the 1970s in a completely different context.
2 An aggregation/segregation problem
In this section we study the minimum problem for two simplified versions of (1.1), which
we interpret as optimizing the disposition of some objects of different types (actually
dinosaurs of different species). The first problem is purely discrete, namely with a finite
number of dinosaurs of a finite number of species. The second one is semi-discrete,
namely with a continuum of dinosaurs belonging to a finite number of species.
2.1 Discrete setting
Let us consider
• a positive integer n,
• a function u : {1, . . . , n} → Z,
• a symmetric subset E ⊆ Z2 (namely any subset with the property that (i, j) ∈ E
if and only if (j, i) ∈ E),
• a nonincreasing function h : {0, 1, . . . , n − 1} → R.
5
Let us introduce the discrete interaction set
J(E, u) :=(cid:8)(x, y) ∈ {1, . . . , n}2 : x ≤ y, (u(x), u(y)) ∈ E(cid:9) ,
and let us finally define
H(h, E, u) := X(x,y)∈J(E,u)
h(y − x).
(2.1)
(2.2)
Just to help intuition, we think of u as an arrangement of n dinosaurs placed in the
points {1, . . . , n}. There are different species of dinosaurs, indexed by integer numbers,
so that u(x) denotes the species of the dinosaur in position x. The subset E ⊆ Z2 is the
list of all pairs of species that are hostile to each other. A pair of points (x, y) belongs
to J(E, u) if and only if x ≤ y and the two dinosaurs placed in x and y belong to hostile
species, and in this case the real number h(y − x) measures the "hostility" between the
two dinosaurs. As expected, the closer are the dinosaurs, the larger is their hostility.
Keeping this jurassic framework into account, sometimes in the sequel we call u a
"discrete arrangement of n dinosaurs", we call E an "enemy list", we call h a "discrete
hostility function", and H(h, E, u) the "total hostility of the arrangement". At this
level of generality, we admit the possibility that (i, i) ∈ E for some integer i, namely
that a dinosaur is hostile to dinosaurs of the same species, including itself. For this
reason, the hostility function h(x) is defined also for x = 0. This generality turns out
to be useful in the proof of the main result for discrete arrangements.
In the sequel we focus on the special case where E coincides with
Ek := {(i, j) ∈ Z2 : j − i ≥ k + 1}
(2.3)
for some positive integer k.
In this case it is quite intuitive that the arrangements
that minimize the total hostility are the "monotone" ones, namely those in which all
dinosaurs of the same species are close to each other, and the groups corresponding to
different species are sorted in ascending or descending order. To this end, we introduce
the following notion.
Definition 2.1 (Nondecreasing rearrangement – Discrete setting). Let n be a positive
integer, and let u : {1, . . . , n} → Z be a function. The nondecreasing rearrangement of
u is the function Mu : {1, . . . , n} → Z defined as
Mu(x) := min(cid:8)j ∈ Z : {y ∈ {1, . . . , n} : u(y) ≤ j} ≥ x(cid:9) ,
where A denotes the number of elements of the set A.
As the name suggests, Mu turns out to be a nondecreasing function, and it is
uniquely characterized by the fact that the two level sets
(cid:8)x ∈ {1, . . . , n} : u(x) = j(cid:9) ,
have the same number of elements for every j ∈ Z.
hostility with respect to the enemy list Ek.
(cid:8)x ∈ {1, . . . , n} : Mu(x) = j(cid:9)
As expected, the main result is that monotone arrangements minimize the total
6
Theorem 2.2 (Total hostility minimization – Discrete setting). Let n and k be two
positive integers, let Ek ⊆ Z2 be the subset defined by (2.3), and let h : {0, . . . , n− 1} →
R be a nonincreasing function. Let u : {1, . . . , n} → Z be any function, let Mu be the
nondecreasing rearrangement of u introduced in Definition 2.1, and let H(h, Ek, u) be
the total hostility defined in (2.2).
Then it turns out that
H(h, Ek, u) ≥ H(h, Ek, Mu).
(2.4)
2.2 Semi-discrete setting
Let us consider
• an interval (a, b) ⊆ R,
• a measurable function u : (a, b) → Z with finite image,
• a symmetric subset E ⊆ Z2,
• a nonincreasing function c : (0, b − a) → R (note that c(σ) might diverge as
σ → 0+).
Let us introduce the semi-discrete interaction set
I(E, u) :=(cid:8)(x, y) ∈ (a, b)2 : (u(x), u(y)) ∈ E(cid:9) ,
and let us finally define
F (c, E, u) :=ZZI(E,u)
c(y − x) dx dy.
(2.5)
(2.6)
In analogy with the discrete setting, we interpret u(x) as a continuous arrangement
of dinosaurs of a finite number of species, c(y−x) as the hostility between two dinosaurs
of hostile species placed in x and y, and we think of F (c, E, u) as the total hostility of
the arrangement u with respect to the enemy list E.
Once again, we suspect that monotone arrangements minimize the total hostility
with respect to the enemy list Ek. This leads to the following notion.
Definition 2.3 (Nondecreasing rearrangement – Semi-discrete setting). Let u : (a, b) →
Z be a measurable function with finite image. The nondecreasing rearrangement of u
is the function Mu : (a, b) → Z defined as
Mu(x) := min(cid:8)j ∈ Z : meas{y ∈ (a, b) : u(y) ≤ j} ≥ x − a(cid:9) ,
where meas(A) denotes the Lebesgue measure of a subset A ⊆ (a, b).
7
The function Mu is nondecreasing and satisfies
meas{x ∈ (a, b) : u(x) = j} = meas{x ∈ (a, b) : Mu(x) = j}
∀j ∈ Z.
The following result is the semi-discrete counterpart of Theorem 2.2.
Theorem 2.4 (Total hostility minimization – Semi-discrete setting). Let (a, b) ⊆ R be
an interval, let k be a positive integer, let Ek ⊆ Z2 be the subset defined by (2.3), and
let c : (0, b − a) → R be a nonincreasing function. Let u : (a, b) → Z be any measurable
function with finite image, let Mu be the nondecreasing rearrangement of u introduced
in Definition 2.3, and let F (c, Ek, u) be the total hostility defined in (2.6).
Then it turns out that
F (c, Ek, u) ≥ F (c, Ek, Mu).
(2.7)
Remark 2.5. Theorem 2.4 above is stated in the form that we need in the proof of
Proposition 3.2. With a further approximation step in the proof, one can show that
the same conclusion (2.7) holds true also without assuming that the image of u is finite
and contained in Z, and without assuming that k is a positive integer (but just a real
number greater than −1).
Theorem 2.2 is equivalent to Taylor's lemma (see [19] and [10, Theorem 1.2]).
This extended result is equivalent to [10, Theorem 1.1], in the same way as our
2.3 Proof of Theorem 2.2
Despite the quite intuitive statement, the proof requires some work. To begin with, we
introduce some notation, and we develop some preliminary results. Since the function
h is fixed once for all, in the sequel we simply write H(E, u) instead of H(h, E, u).
Left-right gap Let L and R be two finite sets of positive integers. We call left-right
gap the quantity
G(L, R) := h(0) +Xℓ∈L
h(ℓ) +Xr∈R
h(r) − X(ℓ,r)∈L×R(cid:0)h(ℓ + r − 1) − h(ℓ + r)(cid:1) .
We prove that
(2.8)
(2.9)
G(L, R) ≤
L+RXi=0
h(i).
To this end, we argue by induction on the number of elements of R. If R = ∅, from
(2.8) we deduce that
G(L, R) := h(0) +Xℓ∈L
h(ℓ) ≤
LXi=0
h(i) =
L+RXi=0
h(i),
8
where the inequality is true term-by-term because h is nonincreasing.
Let us assume now that the conclusion holds true whenever R has n elements, and
let us consider any pair (L, R) with R = n + 1. Let us set
a := max R,
b := min{n ∈ N \ {0} : n 6∈ L},
and let us consider the new pair (L1, R1) defined as
L1 := L ∪ {b},
R1 := R \ {a}.
In words, we have removed the largest element of R, and added the smallest possible
element to L. We observe that R1 = n and L1 + R1 = L + R. Therefore, if we
show that
(2.10)
G(L, R) ≤ G(L1, R1),
then (2.9) follows from the inductive assumption.
In order to prove (2.10), we expand the left-hand side and the right-hand side
according to (2.8). After canceling out the common terms, with some algebra we obtain
that inequality (2.10) holds true if and only if
h(a) +Xr∈R1(cid:0)h(b + r − 1) − h(b + r)(cid:1) ≤ h(b) +Xℓ∈L(cid:0)h(ℓ + a − 1) − h(ℓ + a)(cid:1) .
(2.11)
All terms in the sums are nonnegative because h is nonincreasing. Let us consider
the left-hand side. If a > 1 we know that R1 ⊆ {1, . . . , a − 1}, and hence
h(a) +Xr∈R1(cid:0)h(b + r − 1) − h(b + r)(cid:1) ≤ h(a) +
a−1Xr=1(cid:0)h(b + r − 1) − h(b + r)(cid:1)
= h(a) + h(b) − h(a + b − 1).
(2.12)
The same inequality is true for trivial reasons also if a = 1.
Let us consider now the right-hand side of (2.11).
If b > 1 we know that L ⊇
{1, . . . , b − 1}, and hence
h(b) +Xℓ∈L(cid:0)h(ℓ + a − 1) − h(ℓ + a)(cid:1) ≥ h(b) +
b−1Xℓ=1(cid:0)h(ℓ + a − 1) − h(ℓ + a)(cid:1)
= h(b) + h(a) − h(a + b − 1).
As before, the same inequality is true for trivial reasons also if b = 1.
Combining (2.13) and (2.12) we obtain (2.11), which in turn is equivalent to (2.10).
(2.13)
This completes the proof of (2.9).
9
Reduction of an arrangement and hostility gap For every function v : {1, . . . , n} → Z,
let us set
µ := max{v(i) : i ∈ {1, . . . , n}},
and let m ∈ {1, . . . , n} be the largest index such that v(m) = µ.
If n ≥ 2, we call reduction of v the function Rv : {1, . . . , n − 1} → Z defined by
[Rv](i) :=(cid:26) v(i)
v(i + 1)
if i < m,
if i ≥ m.
In terms of dinosaurs, Rv is the arrangement obtained from v by removing the
rightmost dinosaur of the species indexed by the highest integer, and by shifting one
position to the left all subsequent dinosaurs.
When passing from v to Rv, the total hostility changes by an amount that we call
hostility gap, defined as
∆(E, v) := H(E, v) − H(E, Rv).
We observe that interactions between any two dinosaurs placed on the same side of
the removed one are equal before and after the removal, and therefore they cancel out
when computing the gap. On the contrary, if two hostile dinosaurs are placed within
distance d on opposite sides of the removed one, their hostility changes from h(d) to
h(d − 1) after the removal. It follows that the hostility gap can be written as
∆(E, v) = Xi∈J1(E,u,m)
h(m − i) − X(i,j)∈J2(E,u,m)(cid:0)h(j − i − 1) − h(j − i)(cid:1) ,
(2.14)
where
and
J1(E, u, m) :=(cid:8)i ∈ {1, . . . , n} : (u(i), u(m)) ∈ E(cid:9)
J2(E, u, m) :=(cid:8)(i, j) ∈ {1, . . . , n}2 : i < m < j, (u(i), u(j)) ∈ E(cid:9) .
The first sum in (2.14) keeps into account the interactions of the removed dinosaur
with the rest of the world, the second sum represents the increment of the total hostility
due to the reduction of distances among the others.
Monotone rearrangement decreases the hostility gap We prove that
∆(Ek, v) ≥ ∆(Ek, Mv)
(2.15)
for every arrangement v of n dinosaurs.
In order to prove this inequality, we consider the new enemy list
Ehµi := Z2 \ {µ, µ − 1, . . . , µ − k}2,
10
and we claim that
∆(Ek, v) ≥ ∆(Ehµi, v) ≥ ∆(Ehµi, Mv) = ∆(Ek, Mv),
(2.16)
which of course implies (2.15).
The equality between the last two terms of (2.16) follows from formula (2.14). In-
deed, since Mv is nondecreasing, the removed dinosaur is the rightmost one, and there-
fore in both cases the second sum in (2.14) is void. Also the first sum in (2.14) is the
same in both cases, because a dinosaur of the highest species is hostile to another di-
nosaur with respect to the enemy list Ek if and only if it is hostile to the same dinosaur
with respect to the enemy list Ehµi.
The inequality between the first two terms of (2.16) follows again from formula
(2.14). Indeed, the first sum has the same terms both in the case of the enemy list Ek
and in the case of the enemy list Ehµi, as observed above. As for the second sum, the
interactions with respect to Ek are also interactions with respect to Ehµi, and therefore
when passing from Ek to Ehµi the second sum cannot decrease. Since the second sum
appears in (2.14) with negative sign, the hostility gap with respect to Ehµi is less than
or equal to the hostility gap with respect to Ek.
It remains to prove that
∆(Ehµi, v) ≥ ∆(Ehµi, Mv).
To this end, we introduce the complement enemy list
(2.17)
Ec
hµi := {µ, µ − 1, . . . , µ − k}2 = Z2 \ Ehµi.
Since Z2 is the disjoint union of Ehµi and Ec
hµi, and the total hostility is additive
with respect to the enemy list, we deduce that
H(Ehµi, w) = H(Z2, w) − H(Ec
hµi, w)
for every arrangement w, and for the same reason
∆(Ehµi, w) = ∆(Z2, w) − ∆(Ec
hµi, w).
Moreover, we observe that the total hostility with respect to Z2 depends only on
the number of dinosaurs, and in particular ∆(Z2, v) = ∆(Z2, Mv). As a consequence,
proving (2.17) is equivalent to showing that
∆(Ec
hµi, v) ≤ ∆(Ec
hµi, Mv).
(2.18)
The advantage of this "complement formulation" is that hostility gaps with respect
hµi depend only on the relative positions of the removed dinosaur with respect to
to Ec
the other dinosaurs of the species with indices between µ − k and µ.
11
To be more precise, let us compute the left-hand side of (2.18). Let m denote as
usual the position of the dinosaur that is removed from v to Rv, and let us set
R(v) := {r ≥ 1 : v(m + r) ∈ {µ, µ − 1, . . . , µ − k}} ,
L(v) := {ℓ ≥ 1 : v(m − ℓ) ∈ {µ, µ − 1, . . . , µ − k}} .
In other words, this means that
{m − ℓ : ℓ ∈ L(v)} ∪ {m} ∪ {m + r : r ∈ R(v)}
is the set of all integers i ∈ {1, . . . , n} such that v(i) ∈ {µ, µ − 1, . . . , µ − k}, namely
the set of positions where the dinosaurs of the last k + 1 species are placed.
With this notation, the first sum in (2.14) is
h(0) + Xℓ∈L(v)
h(ℓ) + Xr∈R(v)
h(r)
(we recall that in this "complement formulation" the dinosaur in m is also hostile to
itself), while the second sum in (2.14) is
X(ℓ,r)∈L(v)×R(v)(cid:0)h(ℓ + r − 1) − h(ℓ + r)(cid:1) .
h(ℓ) + Xr∈R(v)
h(r) − X(ℓ,r)∈L(v)×R(v)(cid:0)h(ℓ + r − 1) − h(ℓ + r)(cid:1)
Therefore, it turns out that
∆(Ec
hµi, v) = h(0) + Xℓ∈L(v)
= G(L(v), R(v)),
(2.19)
where G is the left-right gap defined in (2.8).
On the other hand, in the nondecreasing arrangement Mv the rightmost dinosaur
has L(v) + R(v) dinosaurs of the last k + 1 species exactly on its left, and therefore
∆(Ec
hµi, Mv) =
L(v)+R(v)Xi=0
h(i).
(2.20)
Keeping (2.19) and (2.20) into account, inequality (2.18) is exactly (2.9).
Conclusion We are now ready to prove (2.4). To this end, we argue by induction on
the number n of dinosaurs.
In the case n = 1 there is nothing to prove.
Let us assume now that the conclusion holds true for all arrangements of n dinosaurs.
Let u be any arrangement of n + 1 dinosaurs, and let Ru denote its reduction. Since
Ru is an arrangement of n dinosaurs, from the inductive assumption we know that
H(Ek, Ru) ≥ H(Ek, M(Ru)).
12
On the other hand, from (2.15) we know that
Since M(Ru) = R(Mu), we finally conclude that
∆(Ek, u) ≥ ∆(Ek, Mu).
H(Ek, u) = H(Ek, Ru) + ∆(Ek, u)
≥ H(Ek, M(Ru)) + ∆(Ek, Mu)
= H(Ek, R(Mu)) + ∆(Ek, Mu)
= H(Ek, Mu),
which proves the conclusion for arrangements of n + 1 dinosaurs. (cid:3)
2.4 Proof of Theorem 2.4
The proof relies on the following approximation result (we omit the proof, which is an
exercise in basic measure theory).
Lemma 2.6. Let m be a positive integer, and let D1, . . . , Dm be disjoint measurable
subsets of (0, 1) such that
Then for every ε > 0 there exist disjoint subsets D1,ε, . . . , Dm,ε of [0, 1] such that
Di = (0, 1).
Di,ε = (0, 1),
m[i=1
m[i=1
and such that for every i = 1, . . . , m it turns out that
• Di,ε is a finite union of intervals with rational endpoints,
• the Lebesgue measure of the symmetric difference between Di and Di,ε is less than
or equal to ε. (cid:3)
We are now ready to prove Theorem 2.4. First of all, we observe that (2.7) is invari-
ant by translations and homotheties. As a consequence, there is no loss of generality
in assuming that (a, b) = (0, 1) and c : (0, 1) → R. Then we proceed in three steps. To
begin with, we prove (2.7) in the special case where the hostility function c is bounded
and the arrangement u has a very rigid structure, then for general u but again bounded
hostility function, and finally in the general setting.
13
Step 1 We prove (2.7) under the additional assumption that the hostility function
c : (0, 1) → R is bounded, and that there exists a positive integer d such that u(x) is
constant in each interval of the form ((i − 1)/d, i/d) with i = 1, . . . , d.
Indeed, this is actually the discrete setting. To be more precise, we introduce the
discrete arrangement v : {1, . . . , d} → Z defined as
d (cid:19)
v(i) := u(cid:18)i − 1/2
∀i ∈ {1, . . . , d},
and the discrete hostility function h : {0, . . . , d − 1} → R defined as
h(i) :=Z 1/d
0
dxZ (i+1)/d
i/d
c(y − x) dy
∀i ∈ {0, . . . , d − 1},
which represents the contribution to the total hostility of two intervals of length 1/d
occupied by hostile dinosaurs, and placed at distance i/d from each other. Then for
every enemy list Ek it turns out that
F (c, Ek, u) = 2H(h, Ek, v),
where H(h, Ek, v) is the discrete total hostility defined in (2.2), and the factor 2 keeps
into account that both (x, y) and (y, x) are included in the semi-discrete interaction
set I(Ek, u), while only one of them is included in the discrete counterpart J(Ek, v)
(see (2.1) and (2.5)). Moreover, the monotone rearrangement Mv of v is related to the
monotone rearrangement Mu of u by the formula
Mv(i) = Mu(cid:18) i − 1/2
d (cid:19)
∀i ∈ {1, . . . , d},
and again it turns out that
F (c, Ek, Mu) = 2H(h, Ek, Mv)
for every enemy list Ek. At this point, (2.7) is equivalent to
which in turn is true because of Theorem 2.2.
H(h, Ek, v) ≥ H(h, Ek, Mv),
Step 2 We prove (2.7) for a general arrangement u : (0, 1) → Z, but again under the
additional assumption that the hostility function c : (0, 1) → R is bounded.
To this end, let z1 < z2 < . . . < zm denote the elements in the image of u, and let
Di := {x ∈ (0, 1) : u(x) = zi}
∀i ∈ {1, . . . , m}
14
denote the set of positions of dinosaurs of the species zi. For every ε > 0, let us consider
the sets D1,ε, . . . , Dm,ε given by Lemma 2.6, and the function uε : (0, 1) → Z defined
as
uε(x) = zi
∀x ∈ Di,ε.
Since the hostility function c is bounded, and the symmetric difference between Di
and Di,ε has measure less than or equal to ε, there exists a constant Γ (depending on
m and c, but independent of ε) such that
F (c, Ek, Mu) − F (c, Ek, Muε) ≤ Γε.
F (c, Ek, u) − F (c, Ek, uε) ≤ Γε
On the other hand, the function uε satisfies the assumptions of the previous step,
and
and therefore
From all these inequalities it follows that
F (c, Ek, uε) ≥ F (c, Ek, Muε).
F (c, Ek, u) ≥ F (c, Ek, Mu) − 2Γε.
Since ε > 0 is arbitrary, (2.7) is proved in this case.
Step 3 We prove (2.7) without assuming that the hostility function c(x) is bounded.
To this end, for every n ∈ N we consider the truncated hostility function
We observe that
cn(x) := min{c(x), n}
∀x ∈ (0, 1).
F (c, Ek, u) ≥ F (cn, Ek, u)
∀n ∈ N
because c(x) ≥ cn(x) for every x ∈ (0, 1), and
F (cn, Ek, u) ≥ F (cn, Ek, Mu)
∀n ∈ N
because of the result of the previous step applied to the bounded hostility function
cn(x). As a consequence, we obtain that
∀n ∈ N.
On the other hand, by monotone convergence we deduce that
F (c, Ek, u) ≥ F (cn, Ek, Mu)
(2.21)
F (c, Ek, Mu) = sup
n∈N F (cn, Ek, Mu),
and therefore (2.7) follows from (2.21). (cid:3)
15
3 Gamma-convergence in dimension one
In this section we prove Theorem 1.1 for d = 1, in which case
G1,p = 2
∀p ≥ 1.
(3.1)
To begin with, we introduce the notion of vertical δ-segmentation, which is going to
play a crucial role in many parts of the proof.
Definition 3.1 (Vertical δ-segmentation). Let X be any set, let w : X → R be any
function, and let δ > 0. The vertical δ-segmentation of w is the function Sδw : X → R
defined by
δ (cid:23)
Sδw(x) := δ(cid:22) w(x)
∀x ∈ X.
(3.2)
The function Sδw takes its values in δZ, and it is uniquely characterized by the fact
that Sδw(x) = kδ for some k ∈ Z if and only if kδ ≤ w(x) < (k + 1)δ.
3.1 Asymptotic cost of oscillations
Let us assume that a function uδ(x) oscillates between two values A and B in some
interval (a, b). Does this provide an estimate from below for Λδ,p(uδ, (a, b)), at least
when δ is small enough? The following Proposition and the subsequent corollaries give
a sharp quantitative answer to this question. They are the fundamental tool in the
proof of the liminf inequality.
Proposition 3.2 (Limit cost of vertical oscillations). Let p ≥ 1 be a real number, let
(a, b) ⊆ R be an interval, and let {uδ}δ>0 ⊆ Lp((a, b)) be a family of functions.
Let us assume that there exist two real numbers A ≤ B such that
lim inf
δ→0+
meas{x ∈ (a, b) : uδ(x) ≤ A + ε} > 0
∀ε > 0,
and
lim inf
δ→0+
meas{x ∈ (a, b) : uδ(x) ≥ B − ε} > 0
∀ε > 0.
Then it turns out that
lim inf
δ→0+
Λδ,p(uδ, (a, b)) ≥
2
p · Cp ·
(B − A)p
(b − a)p−1 ,
where Cp is the constant defined in (1.6).
(3.3)
(3.4)
(3.5)
16
Proof To begin with, we observe that (3.5) is trivial if A = B, or if the left-hand side
is infinite. Up to restricting ourselves to a sequence δk → 0+, we can also assume that
the liminf is actually a limit. Therefore, in the sequel we assume that the left-hand side
of (3.5) is uniformly bounded from above, and that A < B.
Let us fix ε > 0 such that 4ε < B − A. Due to assumptions (3.3) and (3.4), there
exist η > 0 and δ0 > 0 such that
meas{x ∈ (a, b) : uδ(x) ≤ A + ε} ≥ η
meas{x ∈ (a, b) : uδ(x) ≥ B − ε} ≥ η
∀δ ∈ (0, δ0),
∀δ ∈ (0, δ0).
(3.6)
(3.7)
Truncation, δ-segmentation and monotone rearrangement
In this section of the
proof, we replace {uδ} with a new family {buδ} of monotone piecewise constant functions
that still satisfies (3.3) and (3.4), without increasing the left-hand side of (3.5). To this
end, we perform three operations on uδ(x).
The first operation is a truncation between A and B. To be more precise, we define
TA,Buδ : (a, b) → R by setting
TA,Buδ(x) :=
We observe that the implication
A
uδ(x)
B
if uδ(x) < A,
if A ≤ uδ(x) ≤ B,
if uδ(x) > B.
TA,Buδ(y) − TA,Buδ(x) > δ =⇒ uδ(y) − uδ(x) > δ
holds true for every x and y in (a, b), and hence
Λδ,p(TA,Buδ, (a, b)) ≤ Λδ,p(uδ, (a, b))
∀δ > 0.
We also observe that (3.6) and (3.7) remain true if we replace uδ(x) by TA,Buδ(x).
The second operation is a vertical δ-segmentation, namely we replace TA,Buδ by the
function SδTA,Buδ defined according to (3.2).
Again we observe that the implications
SδTA,Buδ(y) − SδTA,Buδ(x) > δ =⇒ SδTA,Buδ(y) − SδTA,Buδ(x) ≥ 2δ
=⇒ TA,Buδ(y) − TA,Buδ(x) > δ
hold true for every x and y in (a, b), and hence
Λδ,p(SδTA,Buδ, (a, b)) ≤ Λδ,p(TA,Buδ, (a, b))
∀δ > 0.
As for (3.6) and (3.7), we set δ1 := min{ε, δ0}, and we observe that now
∀δ ∈ (0, δ1),
meas{x ∈ (a, b) : SδTA,Buδ(x) ≤ A + 2ε} ≥ η
(3.8)
17
meas{x ∈ (a, b) : SδTA,Buδ(x) ≥ B − 2ε} ≥ η
∀δ ∈ (0, δ1).
(3.9)
The third and last operation we perform is monotone rearrangement, namely we
replace SδTA,Buδ with the nondecreasing function MSδTA,Buδ in (a, b) whose level sets
have the same measure of the level sets of SδTA,Buδ (see Definition 2.3).
From (3.8) and (3.9) we deduce that now
MSδTA,Buδ(x) ≤ A + 2ε
MSδTA,Buδ(x) ≥ B − 2ε
Moreover, we claim that
∀x ∈ (a, a + η),
∀x ∈ (b − η, b),
∀δ ∈ (0, δ1),
∀δ ∈ (0, δ1).
Λδ,p(MSδTA,Buδ, (a, b)) ≤ Λδ,p(SδTA,Buδ, (a, b))
∀δ > 0.
(3.10)
(3.11)
(3.12)
This is a straightforward consequence of Theorem 2.4. To be more formal, let us
consider the semi-discrete arrangement vδ : (a, b) → Z defined by
vδ(x) :=
SδTA,Buδ(x)
∀x ∈ (a, b)
1
δ
(we recall that SδTA,Buδ takes its values in δZ, and hence vδ(x) is integer valued), and
the hostility function c : (0, b − a) → R defined as c(σ) := δpσ−1−p. We observe that
MSδTA,Buδ(x) = δMvδ(x)
∀x ∈ (a, b),
where Mvδ is the nondecreasing rearrangement of vδ according to Definition 2.3.
We observe also that for every pair of points x and y in (a, b) it turns out that
(x, y) ∈ I(δ, SδTA,Buδ, (a, b)) ⇐⇒ vδ(y) − vδ(x) ≥ 2 ⇐⇒ (x, y) ∈ I(E1, vδ),
where E1 is the enemy list defined in (2.3), and I(E1, vδ) is the semi-discrete interaction
set defined according to (2.5). It follows that
Λδ,p(SδTA,Buδ, (a, b)) = F (c, E1, vδ),
and therefore (3.12) is equivalent to (2.7).
Λδ,p(MSδTA,Buδ, (a, b)) = F (c, E1, Mvδ),
In conclusion, the three operations described so far delivered us a family
satisfies (3.10) and (3.11), and
of nondecreasing functions such that the image of buδ is contained in δZ. This family
(3.13)
∀δ > 0.
In the sequel we are going to show that any such family satisfies
buδ := MSδTA,Buδ
Λδ,p(uδ, (a, b)) ≥ Λδ,p(buδ, (a, b))
lim inf
δ→0+
Λδ,p(buδ, (a, b)) ≥
18
Due to (3.13) and the arbitrariness of ε > 0, this is enough to prove (3.5).
2
p · Cp ·
(B − A − 4ε)p
(b − a)p−1
.
(3.14)
Extension of the integrals to a vertical strip
In this section of the proof we modify
the domain of integration in order to simplify the computation of Λδ,p(buδ, (a, b)).
To begin with, we observe that
δp
δp
y − x1+p dx dy ≥ZZBδ
Λδ,p(buδ, (a, b)) =ZZAδ
y − x1+p dx dy,
Aδ := I(δ,buδ, (a, b)) =(cid:8)(x, y) ∈ (a, b)2 : buδ(y) −buδ(x) > δ(cid:9) ,
Bδ :=(cid:8)(x, y) ∈ (a + η, b − η) × (a, b) : buδ(y) −buδ(x) > δ(cid:9) .
y − x1+p dx dy =ZZBδ ∪Cδ
y − x1+p dx dy −ZZCδ
δp
δp
δp
y − x1+p dx dy,
Then we write the last integral in the form
where
ZZBδ
where
Cδ := (a + η, b − η) × (R \ (a, b)).
In other words, the set Bδ ∪ Cδ consists of the vertical strip (a + η, b− η)× R minus
Now we observe that
the set of points (x, y) ∈ (a + η, b − η) × (a, b) such that buδ(y) −buδ(x) ≤ δ.
ZZCδ
δp
y − x1+p dx dy = 2δpZ b−η
a+η
dxZ +∞
b
1
y − x1+p dy.
From the convergence of the last double integral it follows that
and therefore
lim inf
δ→0+
δ→0+ZZCδ
lim
δp
y − x1+p dx dy = 0,
δ→0+ ZZBδ
δ→0+ ZZBδ∪Cδ
δp
y − x1+p dx dy
δp
y − x1+p dx dy.
Λδ,p(buδ, (a, b)) ≥ lim inf
= lim inf
Computing the integrals
lim inf
δ→0+ ZZBδ∪Cδ
In this last part of the proof we show that
(B − A − 4ε)p
δp
.
2
p · Cp ·
y − x1+p dx dy ≥
(b − a)p−1
(3.15)
(3.16)
Recalling (3.15), this proves (3.14), and hence also (3.5).
function with finite image. Let us consider the partition
To this end, we need to introduce some notation. We know thatbuδ is a nondecreasing
a = x0 < x1 < . . . < xn = b
19
and different intervals correspond to different constants. Let us set
of (a, b) with the property thatbuδ(x) is constant in each interval of the form (xi−1, xi),
h := min{i ∈ {1, . . . , n} : xi ≥ a + η},
k := max{i ∈ {0, . . . , n − 1} : xi ≤ b − η}.
δp
Of course n, h, k, as well as the partition, do depend on δ. Now we claim that
2
p · Cp ·
δp(k − h − 1)p
y − x1+p dx dy ≥
ZZBδ∪Cδ
To begin with, we show that the values of buδ(x) in neighboring intervals are con-
secutive multiples of δ, namely if buδ(x) = mδ in (xi−1, xi) for some m ∈ Z, then
buδ(x) = (m + 1)δ in (xi, xi+1). Let us assume indeed thatbuδ(x) ≥ (m + 2)δ in (xi, xi+1).
In this case it turns out that
∀δ ∈ (0, δ1).
(b − a)p−1
(3.17)
δp
Λδ,p(buδ, (a, b)) ≥Z xi
xi−1
dxZ xi+1
xi
(y − x)1+p dy,
which is absurd because the left-hand side is uniformly bounded from above and the
integral in the right-hand side is divergent.
With these notations it turns out that
Now we distinguish two cases.
• If p = 1, computing the integrals we obtain that
ZZBδ∪Cδ
δ
(y − x)2 dx dy ≥ δ
k−1Xi=h+1
log(cid:18)xi+1 − xi−1
xi+1 − xi
xi+1 − xi−1
xi − xi−1 (cid:19) .
·
If ℓi := xi − xi−1 denotes the length of the i-th interval of the partition, and we
apply the inequality between arithmetic and geometric mean, we obtain that
ZZBδ∪Cδ
δ
(ℓi + ℓi+1)2
ℓi · ℓi+1
log
(y − x)2 dx dy ≥ δ
≥ δ
= 2 log 2 · δ(k − h − 1),
log 4
k−1Xi=h+1
k−1Xi=h+1
20
ZZBδ∪Cδ
δp
y − x1+p dx dy
k−1Xi=h+1(cid:18)Z xi
≥
k−1Xi=h+1(cid:18)Z xi
δp
p
xi−1
=
δp
xi+1
y − x1+p dy +Z xi+1
dxZ +∞
(xi+1 − x)p dx +Z xi+1
dxZ xi−1
(x − xi−1)p dx(cid:19) .
−∞
1
1
xi
xi
xi−1
δp
y − x1+p dy(cid:19)
which proves (3.17) in this case.
• If p > 1, computing the integrals we obtain that
ZZBδ ∪Cδ
δp
y − x1+p dx dy ≥
δp
p(p − 1)
k−1Xi=h+1 1
ℓp−1
i+1
+
1
i −
ℓp−1
(ℓi+1 + ℓi)p−1! ,
2
where we set ℓi := xi−xi−1 as before. Therefore, with two applications of Jensen's
inequality to the convex function t → t1−p, we obtain that
ZZBδ ∪Cδ
δp
y − x1+p dx dy ≥
≥
≥
=
δp
2p − 2
(ℓi+1 + ℓi)p−1
(k − h − 1)p
k−1Xi=h+1
p(p − 1)
δp(2p − 2)
p(p − 1) ·
(cid:16)Pk−1
i=h+1(ℓi+1 + ℓi)(cid:17)p−1
δp(2p − 2)
p(p − 1) ·
2
p · Cp ·
(k − h − 1)p
(2(b − a))p−1
δp(k − h − 1)p
,
(b − a)p−1
which proves (3.17) also in this case.
Now it remains to estimate δ(k−h−1). To this end, from (3.10) and the minimality
of h we deduce that
∀x ∈ (xh−1, xh).
Similarly, from (3.11) and the maximality of k we deduce that
∀x ∈ (xk, xk+1).
A + 2ε ≥buδ(x) =: mAδ
B − 2ε ≤buδ(x) =: mBδ
Since the values ofbuδ in consecutive intervals are consecutive multiples of δ, it turns
mB = mA + (k − h + 1),
out that
and therefore
(k − h − 1)δ = (k − h + 1)δ − 2δ = (mB − mA)δ − 2δ ≥ B − A − 4ε − 2δ.
Plugging this inequality into (3.17), and letting δ → 0+, we obtain (3.16), which
completes the proof. (cid:3)
The following result is a straightforward consequence of Proposition 3.2.
Corollary 3.3. Let us assume that uδ → u in Lp(R), and let (a, b) ⊆ R be an interval
whose endpoints a and b are Lebesgue points of u.
Then it turns out that
lim inf
δ→0+
Λδ,p(uδ, (a, b)) ≥
2
p · Cp · u(b) − u(a)p
(b − a)p−1
.
21
It is enough to apply Proposition 3.2 with A := min{u(a), u(b)} and B :=
Proof
max{u(a), u(b)}. Assumptions (3.3) and (3.4) are satisfied because a and b are Lebesgue
points of the limit of the sequence uδ. (cid:3)
We conclude with another variant of Proposition 3.2. We do not need this statement
in the sequel, but we think that it clarifies once more the relation between oscillations
of uδ and values of Λδ,p(uδ, (a, b)).
Corollary 3.4. Let (a, b) ⊆ R be an interval, let {uδ}δ>0 ⊆ Lp((a, b)) be a family of
functions, and let osc(uδ, (a, b)) denote the essential oscillation of uδ in (a, b).
Then it turns out that
lim inf
δ→0+
Λδ,p(uδ, (a, b)) ≥
2
p
Cp
1
(b − a)p−1(cid:18)lim inf
δ→0+
osc(uδ, (a, b))(cid:19)p
.
Proof Let iδ and sδ denote the essential infimum and the essential supremum of uδ(x)
in (a, b), respectively. Let us assume that iδ and sδ are real numbers (otherwise an
analogous argument works with standard minor changes). Let us set wδ(x) := uδ(x)−iδ,
and let us observe that
Λδ,p(uδ, (a, b)) = Λδ,p(wδ, (a, b))
∀δ > 0.
Now it is enough to apply Proposition 3.2 with A := 0 and
B := lim inf
δ→0+
(sδ − iδ) = lim inf
δ→0+
osc(uδ, (a, b)). (cid:3)
3.2 Gamma-liminf
Corollary 3.3 represents a "localized" version of the liminf inequality (1.7), which now
follows from well established techniques (see for example [11, 12]). To this end, we need
the following characterization of Λ0,p(u, R) (we omit the standard proof, based on the
convexity of the norm).
Lemma 3.5 (Piecewise affine horizontal segmentation). Let p ≥ 1 be a real number,
and let u ∈ Lp(R).
Then there exists c ∈ R such that c + q is a Lebesgue point of u for every q ∈ Q.
Moreover, if for every positive integer k we consider the piecewise affine function
vk : R → R such that
then it turns out that
k(cid:19) = u(cid:18)c +
i
k(cid:19)
∀i ∈ Z,
i
vk(cid:18)c +
k→+∞ZR v′
Λ0,p(u, R) = lim
k(x)p dx = sup
k(x)p dx. (cid:3)
k≥1ZR v′
22
We are now ready to prove (1.7) in the case d = 1. Let uδ → u be any family
converging in Lp(R), and let c and vk be as in Lemma 3.5. For every i ∈ Z, we set
ck,i := c + i/k, and we apply Corollary 3.3 in the interval (ck,i, ck,i+1). We obtain that
lim inf
δ→0+
Λδ,p (uδ, (ck,i, ck,i+1)) ≥
Cpu(ck,i+1) − u(ck,i)p
(1/k)p−1
2
p
=
2
p
CpZ ck,i+1
ck,i
v′
k(x)p dx.
Since
we deduce that
Λδ,p(uδ, R) ≥Xi∈Z
Λδ,p(uδ, (ck,i, ck,i+1))
∀δ > 0,
lim inf
δ→0+
Λδ,p(uδ, R) ≥ lim inf
Λδ,p(uδ, (ck,i, ck,i+1))
δ→0+ Xi∈Z
≥ Xi∈Z
≥
=
2
p
2
p
lim inf
δ→0+
Λδ,p(uδ, (ck,i, ck,i+1))
v′
k(x)p dx
ck,i
CpXi∈ZZ ck,i+1
CpZR v′
k(x)p dx.
Letting k → +∞, and recalling (3.1), we obtain exactly (1.7). (cid:3)
3.3 Gamma-limsup
This subsection is devoted to a proof of statement (2) of Theorem 1.1 in the case d = 1.
It is well-known that we can limit ourselves to showing the existence of recovery
families for every u belonging to a subset of Lp(R) that is dense in energy with respect
to Λ0,p(u, R). Classical examples of subsets that are dense in energy are the space
c (R) of functions of class C ∞ with compact support, or the space of piecewise affine
C ∞
functions with compact support. Here for the sake of generality we consider the space
P C 1
Definition 3.6. Let u : R → R be a function. We say that u ∈ P C 1
c (R) if u has
compact support, it is Lipschitz continuous, and there exists a finite subset S ⊆ R such
that u ∈ C 1(R \ S).
c (R) of piecewise C 1 functions with compact support, defined as follows.
We show that for every u ∈ P C 1
c (R) the family Sδu of vertical δ-segmentations of u
is a recovery family. This proves the Gamma-limsup inequality in dimension one.
Proposition 3.7 (Existence of recovery families). Let p ≥ 1 be a real number, and
c (R) be a piecewise C 1 function with compact support according to Defini-
let u ∈ P C 1
tion 3.6. For every δ > 0, let Sδu denote the vertical δ-segmentation of u according to
Definition 3.1.
23
Then it turns out that
lim sup
δ→0+
Λδ,p(Sδu, R) ≤
2
p
CpZR u′(x)p dx.
(3.18)
Proof To begin with, we introduce some notation. Let R0 ≥ 1 be any real number
such that the support of u is contained in [−R0 + 1, R0 − 1]. Let L be the Lipschitz
constant of u in R, and let S ⊆ R be a finite set such that u ∈ C 1(R \ S). For every
x ∈ R and every δ > 0 we set
J(δ, u, x) := {y ∈ R : Sδu(y) − Sδu(x) > δ},
(3.19)
δp
y − x1+p dy,
Hδ,p(x) :=ZJ(δ,u,x)
Λδ,p(Sδu, R) =ZR
Hδ,p(x) dx
∀δ > 0.
(3.20)
In the sequel we call Hδ,p(x) the "pointwise hostility function". It represents the
contribution of each point x to the double integral defining Λδ,p(Sδu, R).
Strategy of the proof The outline of the proof is the following. First of all, we show
that
Then we define an averaged pointwise hostility function bHδ,p(x) with the property
−∞
lim
δ→0+Z −R0
Z R0
−R0
Hδ,p(x) dx = lim
R0
δ→0+Z +∞
Hδ,p(x) dx =Z R0
−R0 bHδ,p(x) dx.
We also show that the averaged pointwise hostility function satisfies the uniform
Hδ,p(x) dx = 0.
(3.21)
and
so that
that
bound
(3.22)
(3.23)
(3.24)
and the asymptotic estimate
2
p
Lp
bHδ,p(x) ≤
δ→0+ bHδ,p(x) ≤
lim sup
∀x ∈ [−R0, R0],
∀δ > 0,
2
p
Cpu′(x)p
∀x ∈ [−R0, R0] \ S.
24
At this point, from Fatou's lemma we deduce that
lim sup
δ→0+ Z R0
−R0
Hδ,p(x) dx = lim sup
δ→0+ Z R0
−R0 bHδ,p(x) dx
δ→0+ bHδ,p(x) dx
CpZ R0
−R0 u′(x)p dx.
lim sup
−R0
≤ Z R0
2
p
≤
Keeping (3.20) and (3.21) into account, this estimate implies (3.18).
Reducing integration to a bounded interval We prove (3.21).
To this end, let us consider any x ≤ −R0. We observe that in this case the set
J(δ, u, x) defined in (3.19) is contained in the support of u, and hence
Z −R0
−∞
Hδ,p(x) dx ≤ δpZ −R0
−∞
dxZ R0−1
−R0+1
1
y − x1+p dy.
At this point the first limit in (3.21) follows from the convergence of the double
integral. The proof of the second limit is analogous.
Uniform bound on the pointwise hostility function We prove that
Hδ,p(x) ≤
2
p
Lp
∀x ∈ [−R0, R0], ∀δ > 0.
(3.25)
To this end, we observe that the implication
Sδu(y) − Sδu(x) > δ =⇒ u(y) − u(x) > δ
holds true for every (x, y) ∈ R2. Since u is Lipschitz continuous, we deduce that
Sδu(y) − Sδu(x) > δ =⇒ y − x ≥
δ
L
,
and hence
as required.
Hδ,p(x) ≤Zy−x≥δ/L
δp
y − x1+p dy = 2Z +∞
δ/L
δp
z1+p dz =
2
p
Lp,
25
The averaged pointwise hostility function bHδ,p : R → R is defined as
Hδ,p(s) ds
1
b − aZ b
a
bHδ,p(x) :=
Averaged pointwise hostility function In this part of the proof we introduce the
averaged pointwise hostility function. To this end, we consider the open set
A(u, δ) := {x ∈ (−R0, R0) : u(x) 6∈ δZ}.
A connected component (a, b) of A(u, δ) is called monotone if [a, b] ∩ S = ∅, and
u′(x) ≥ δ for every x ∈ [a, b]. In this case there exists k ∈ Z such that u(a) = kδ and
u(b) = kδ ± δ, where the sign depends on the sign of u′(x) in (a, b). From the Lipschitz
continuity of u we deduce that A(u, δ) has only a finite number of monotone connected
components.
otherwise.
At this point, inequality (3.23) follows from (3.25), while (3.22) is true because the
if x ∈ [a, b) for some monotone connected component of A(δ, u), and bHδ,p(x) := Hδ,p(x)
integrals of Hδ,p(x) and bHδ,p(x) are the same both in all monotone connected compo-
nents, and in the complement set.
Asymptotic estimate in stationary points We prove that (3.24) holds true for every
x ∈ (−R0, R0) \ S with u′(x) = 0.
To begin with, we observe that in this case x 6∈ [a, b) for every monotone connected
component (a, b) of A(δ, u) (because u′(x) is strictly positive in the closure of every
If J(δ, u, x) = ∅ for every δ > 0, then u is identically null, and the conclusion is
trivial. Otherwise J(δ, u, x) 6= ∅ when δ is small enough. In this case, let rδ be the
largest positive real number such that
monotone connected component), and therefore bHδ,p(x) = Hδ,p(x) for every δ > 0.
(x − rδ, x + rδ) ∩ J(δ, u, x) = ∅,
so that
Hδ,p(x) ≤Z x−rδ
−∞
δp
y − x1+p dy +Z +∞
x+rδ
δp
y − x1+p dy =
2
p(cid:18) δ
rδ(cid:19)p
.
Let δk → 0+ be any sequence such that
δ
rδ
lim sup
δ→0+
= lim
k→+∞
δk
rδk
.
(3.26)
Up to subsequences, we can also assume that rδk tends to some r0. If r0 > 0, then
the limit in the right-hand side of (3.26) is 0, which proves (3.24) in this case. If r0 = 0,
26
then from the maximality of rδk we deduce that u(x ± rδk) − u(x) = δk for a suitable
choice of the sign, which might depend on k. In any case, the limit in the right-hand
side of (3.26) turns out to be
lim
k→+∞
δk
rδk
= lim
k→+∞
u(x ± rδk ) − u(x)
rδk
= u′(x) = 0,
which proves (3.24) also in this case.
Asymptotic estimate in non-stationary points We prove that (3.24) holds true for
every x ∈ (−R0, R0) \ S with u′(x) > 0.
Let us assume, without loss of generality, that u′(x) > 0 (the other case is analogous).
Then for every δ > 0 small enough it turns out that x lies in the closure of a monotone
connected component of A(δ, u). More precisely, there exist four real numbers aδ, bδ,
cδ, dδ with
and kδ ∈ Z such that
u(aδ) = (kδ − 1)δ,
and
aδ < bδ ≤ x < cδ < dδ,
u(bδ) = kδδ,
u(cδ) = (kδ + 1)δ,
u(dδ) = (kδ + 2)δ,
u(y) ∈ ((kδ − 1)δ, kδδ)
u(y) ∈ (kδδ, (kδ + 1)δ)
∀y ∈ (aδ, bδ),
∀y ∈ (bδ, cδ),
u(y) ∈ ((kδ + 1)δ, (kδ + 2)δ)
∀y ∈ (cδ, dδ).
We observe that aδ, bδ, cδ, and dδ tend to x as δ → 0+, and hence
lim
δ→0+
δ
bδ − aδ
= lim
δ→0+
u(bδ) − u(aδ)
bδ − aδ
= u′(x).
Similarly it turns out that
lim
δ→0+
δ
cδ − bδ
= lim
δ→0+
δ
cδ − aδ
= lim
δ→0+
δ
dδ − cδ
δ
dδ − bδ
= u′(x),
=
u′(x)
2
.
and also
lim
δ→0+
From (3.27) through (3.29) we deduce that
(3.27)
(3.28)
(3.29)
(3.30)
(3.31)
(3.32)
J(δ, u, s) ⊆ (−∞, aδ] ∪ [dδ, +∞)
∀s ∈ (bδ, cδ).
It follows that
Hδ,p(s) ≤ZR\(aδ ,dδ)
δp
y − s1+p dy =
δp
p (cid:18)
1
(dδ − s)p +
1
(s − aδ)p(cid:19)
∀s ∈ [bδ, cδ),
27
and hence
bHδ,p(x) =
1
cδ − bδZ cδ
bδ
Hδ,p(s) ds ≤
δp
p
1
bδ (cid:18)
cδ − bδZ cδ
1
(dδ − s)p +
1
(s − aδ)p(cid:19) ds (3.33)
for every x ∈ [bδ, cδ). Now we distinguish two cases.
• If p = 1, computing the integrals in (3.33) we obtain that
cδ − aδ
δ
δ
bHδ,p(x) ≤
cδ − bδ
log(cid:18)dδ − bδ
δ
·
dδ − cδ ·
δ
δ
bδ − aδ(cid:19) ,
·
and therefore (3.24) follows from (3.30) through (3.32).
• If p > 1, computing the integrals in (3.33) we obtain that
bHδ,p(x) ≤
1
δ
p(p − 1)
cδ − bδ ·
(dδ − cδ)p−1 +
δp−1
·(cid:26)
δp−1
(bδ − aδ)p−1 −
δp−1
(dδ − bδ)p−1 −
δp−1
(cδ − aδ)p−1(cid:27) ,
and therefore also in this case (3.24) follows from (3.30) through (3.32).
This completes the proof. (cid:3)
3.4 Smooth recovery families
The aim of this subsection is refining the Gamma-limsup inequality by showing the
existence of recovery families consisting of C ∞ functions with compact support. To this
end, we introduce the following notion.
Definition 3.8 (δ-step functions). Let δ be a positive real number. A function u :
R → R is called a δ-step function if there exists a positive integer n, a (n + 1)-uple
x0 < x1 < . . . < xn of real numbers, and (k1, . . . , kn) ∈ Zn such that
• u(x) = 0 for every x ∈ (−∞, x0) ∪ (xn, +∞),
• u(x) = kiδ in (xi−1, xi) for every i = 1, . . . , n,
• k1 = kn = 1 and ki − ki−1 = 1 for every i = 2, . . . , n.
The values of u(x) for x ∈ {x0, x1, . . . , xn} are not relevant (just to fix ideas, we can
i and the limit of u(x)
define u(xi) as the maximum between the limit of u(x) as x → x+
as x → x−
i ).
28
Now we show that, for every fixed δ > 0, every δ-step function can be approximated
in energy by functions of class C ∞ with compact support. Roughly speaking, this is
possible because the rigid structure of δ-step functions allows to control the effect of
convolutions, which otherwise is unpredictable due to the sensitivity of the integration
region in (1.1) to small perturbations.
Proposition 3.9 (Smooth approximation of δ-step functions). Let δ > 0 and p ≥ 1 be
real numbers, and let u : R → R be a δ-step function.
Then there exists a family {uε}ε>0 ⊆ C ∞
c (R) such that
and
lim
ε→0+
uε = u
in Lp(R),
lim
ε→0+
Λδ,p(uε, R) = Λδ,p(u, R).
Proof Let n, xi and ki be as in the definition of δ-step functions, and let
τ := min{xi − xi−1 : i = 1, . . . , n}
be the length of the smallest interval of the partition. We observe that points in
neighboring intervals do not contribute to the computation of Λδ,p(u, R). In particular,
if we write as usual
Λδ,p(u, R) :=ZZI(δ,u,R)
δp
y − x1+p dx dy,
then it turns out that
y − x ≥ τ
∀(x, y) ∈ I(δ, u, R).
(3.34)
c (R) with
Let us fix a mollifier ρ ∈ C ∞
• ρ(x) ≥ 0 for every x ∈ R,
• ρ(x) = 0 for every x ∈ R with x ≥ 1,
• RR ρ(x) dx = 1,
and let us consider the usual regularization by convolution
uε(x) :=ZR
u(x + εy)ρ(y) dy.
It is well-known that uε ∈ C ∞
Let us assume that 2ε < τ , let us consider the two open sets
out that uε → u in Lp(R) as ε → 0+.
c (R) for every ε > 0, and that for every p ≥ 1 it turns
Aε :=
n[i=0
(xi − ε, xi + ε) ⊆ R,
Bε := (Aε × R) ∪ (R × Aε) ⊆ R2,
29
and let us write
Λδ,p(uε, R) =ZZI(δ,uε,R)∩Bε
δp
y − x1+p dx dy +ZZI(δ,uε,R)\Bε
δp
y − x1+p dx dy.
Since the support of ρ is contained in [−1, 1], it turns out that uε(x) = u(x) for
every x ∈ R \ Aε. It follows that
I(δ, uε, R) \ Bε = I(δ, u, R) \ Bε,
and therefore
ε→0+ZZI(δ,uε,R)\Bε
lim
δp
y − x1+p dx dy = lim
ε→0+ZZI(δ,u,R)\Bε
δp
y − x1+p dx dy = Λδ,p(u, R),
where the last equality follows from Lebesgue's dominated convergence theorem because
Bε shrinks to a set of null measure. So it remains to show that
ε→0+ZZI(δ,uε,R)∩Bε
lim
δp
y − x1+p dx dy = 0.
(3.35)
To this end, from (3.34) and the properties of the support of the mollifier, we deduce
that now
and therefore
ZZI(δ,uε,R)∩Bε
y − x ≥ τ − 2ε
∀(x, y) ∈ I(δ, uε, R),
δp
y − x1+p dx dy ≤ 2
≤ 2
4
p
=
δp
z1+p dz
xi−ε
nXi=0Z xi+ε
dxZz≥τ −2ε
nXi=0Z xi+ε
τ − 2εp dx
τ − 2εp · 2ε(n + 1),
2
p
xi−ε
δp
δp
which implies (3.35). (cid:3)
We are now ready to show the existence of smooth recovery families. As usual, it is
enough to show the existence of such a family for every u in a subset of Lp(R) which is
dense in energy for Λ0,p(u, R). In this case we consider the space PAc(R) of piecewise
affine functions with compact support.
Since piecewise affine functions are piecewise C 1, we know from Proposition 3.7
that the family Sδu of vertical δ-segmentations of u is a (non-smooth) recovery family
for u. The key point is that the vertical δ-segmentation of a piecewise affine function
with compact support is a δ-step function according to Definition 3.8. Thus from
Proposition 3.9 we deduce the existence of a function uδ ∈ C ∞
c (R) such that
kuδ − SδukLp(R) ≤ δ
and
Λδ,p(uδ, R) ≤ Λδ,p(Sδu, R) + δ
for every δ > 0. This implies that {uδ} is a smooth recovery family for u. (cid:3)
30
4 Gamma-convergence in any dimension
It remains to prove Theorem 1.1 in any space dimension. This follows from well es-
tablished sectioning techniques. For every σ ∈ Sd−1, let hσi⊥ denote the hyperplane
orthogonal to σ, namely
hσi⊥ := {z ∈ Rd : hz, σi = 0}.
Given any u : Rd → R, for every σ ∈ Sd−1 and every z ∈ hσi⊥, we consider the
one-dimensional section uσ,z : R → R defined as
uσ,z(x) := u(z + σx)
∀x ∈ R.
The main idea is that Sobolev norms, total variation, and functionals such as Λδ,p
computed in u are a sort of average of the same quantities computed on the one-
dimensional sections uσ,z. The result is the following.
Proposition 4.1 (Integral-geometric representation). Let u : Rd → R be any measur-
able function. Let Λδ,p and Λ0,p be the functionals defined in (1.1) and (1.2), respectively.
(1) For every p ≥ 1 it turns out that
Λ0,p(uσ,z, R) dz = Gd,p Λ0,p(u, Rd),
where Gd,p is the geometric constant defined in (1.4).
(2) For every δ > 0 and every p ≥ 1 it turns out that
Λδ,p(uσ,z, R) dz = 2Λδ,p(u, Rd). (cid:3)
ZSd−1
dσZhσi⊥
ZSd−1
dσZhσi⊥
We skip the details of the proof of Proposition 4.1, which is a simple application
of variable changes in multiple integrals. More generally, for every σ ∈ Sd−1 and every
g ∈ L1(Rd) it turns out that
ZRd
g(y) dy =Zhσi⊥
dzZR
g(z + σx) dx,
and this is the main ingredient in the proof of statement (1).
Similarly, for every g ∈ L1(Rd × Rd) it turns out that
ZZRd×Rd
g(u, v) du dv =
1
2ZSd−1
dσZhσi⊥
dzZZR×R
g(z + σx, z + σy) · y − xd−1 dx dy,
and this is the main ingredient in the proof of statement (2).
We are now ready to prove Theorem 1.1.
31
Gamma-liminf Let us assume that uδ → u in L1(Rd). Then for every σ ∈ Sd−1 it
turns out that
for almost every z ∈ hσi⊥. Therefore, from the integral-geometric representations of
Proposition 4.1, Fatou's lemma, and the one-dimensional result, we obtain that
(uδ)σ,z → uσ,z
in L1(R)
lim inf
δ→0+
Λδ,p(uδ, Rd) = lim inf
δ→0+
Λδ,p((uδ)σ,z, R) dz
dσZhσi⊥
1
2ZSd−1
dσZhσi⊥
dσZhσi⊥
1
2ZSd−1
2ZSd−1
1
1
p
Gd,pCp Λ0,p(u, Rd).
lim inf
δ→0+
Λδ,p((uδ)σ,z, R) dz
2
p
Cp Λ0,p(uσ,z, R) dz
≥
≥
=
≤
≤
=
Gamma-limsup Let u ∈ C ∞
c (Rd) be any function with compact support. For every
δ > 0 we consider the vertical δ-segmentation Sδu of u, and we observe that this
operation commutes with the one-dimensional sections, in the sense that
(Sδu)σ,z = Sδ(uσ,z)
∀σ ∈ Sd−1,
∀z ∈ hσi⊥.
Therefore, from the integral-geometric representations of Proposition 4.1, Fatou's
lemma, and the one-dimensional result, we obtain that
lim sup
δ→0+
Λδ,p(Sδu, Rd) = lim sup
δ→0+
Λδ,p((Sδu)σ,z, R) dz
dσZhσi⊥
1
2ZSd−1
dσZhσi⊥
dσZhσi⊥
1
2ZSd−1
2ZSd−1
1
1
p
Gd,pCp Λ0,p(u, Rd).
lim sup
Λδ,p((Sδu)σ,z, R) dz
δ→0+
2
p
Cp Λ0,p(uσ,z, R) dz
The δ-independent bounds on Λδ,p((Sδu)σ,z, R) needed in order to apply Fatou's
lemma follow from the Lipschitz continuity of u and the boundedness of its support.
It remains to show the existence of smooth recovery families.
Smooth recovery families
The strategy is analogous to the one-dimensional case, and therefore we limit ourselves
to outlining the argument, sparing the reader all technicalities.
32
To begin with, we observe that it is enough to construct smooth recovery families
for every u ∈ PAc(Rd).
In this case, a non-smooth recovery family is provided by
the vertical δ-segmentations Sδu of u. On the other hand, vertical δ-segmentations of
piecewise affine functions with compact support are δ-step functions, and these functions
can be approximated in energy by smooth functions. It follows that for every δ > 0
there exists uδ ∈ C ∞
c (Rd) such that
kuδ − SδukLp(Rd) ≤ δ
and
Λδ,p(uδ, Rd) ≤ Λδ,p(Sδu, Rd) + δ,
and therefore {uδ} is the required recovery family.
The last approximation step can be proved by convolution as we did in Proposi-
tion 3.9. To be more precise, a δ-step function in dimension d is a function v : Rd → R
with the property that there exist a finite set P1, . . . , Pm of disjoint open polytopes
(bounded intersections of half-spaces), and integer numbers k1, . . . , km such that
• v(x) = kiδ in Pi for every i = 1, . . . , m,
• v(x) = 0 in the open set P0 defined as the complement set of the closure of
P1 ∪ . . . ∪ Pm,
• ki − kj ≤ 1 whenever the closure of Pi intersects the closure of Pj,
• ki ≤ 1 whenever the closure of Pi intersects the closure of P0.
In words, the level sets of a δ-step function are finite unions of polytopes, and values
in adjacent regions differ by δ.
The key point is that for every δ-step function v there exists a positive real number
τ such that
(x, y) ∈ I(δ, v, Rd) =⇒ y − x ≥ τ.
As a consequence, when we define vε as the convolution of v with a mollifier whose
support is contained in the ball with center in the origin and radius ε, we obtain that
(x, y) ∈ I(δ, vε, Rd) =⇒ y − x ≥ τ − 2ε,
and at this point the conclusion follows exactly as in the proof of Proposition 3.9. (cid:3)
Acknowledgments
The second author has been introduced to this family of non-local functionals by the
inspiring talk [5] given by H. Brezis during the congress "A mathematical tribute to
Ennio De Giorgi", held in Pisa in September 2016 in the 20-th anniversary of his death.
The same author had been introduced to non-local approximations of free discontinuity
problems by E. De Giorgi himself in the last year of his life.
We are all deeply grateful to both of them.
33
References
[1] C. Antonucci, M. Gobbino, M. Migliorini, N. Picenni. On the gap be-
tween Gamma-limit and pointwise limit for a non-local approximation of the total
variation. arXiv:1712.04413.
[2] C. Antonucci, M. Gobbino, N. Picenni. On the gap between gamma-
limit and pointwise limit for a non-local approximation of the total variation.
arXiv:1708.01231.
[3] J. Bourgain, H. Brezis, P. Mironescu. Another look at Sobolev spaces. In
Optimal control and partial differential equations, pages 439–455. IOS, Amsterdam,
2001.
[4] J. Bourgain, H.-M. Nguyen. A new characterization of Sobolev spaces. C. R.
Math. Acad. Sci. Paris 343 (2006), no. 2, 75–80.
[5] H. Brezis. Another triumph for De Giorgi's Gamma convergence. URL
https://www.youtube.com/watch?v=1Y6fvZX1fx8. Conference held during the
congress "A mathematical tribute to Ennio De Giorgi" (Pisa, September 2016).
[6] H. Brezis. New approximations of the total variation and filters in imaging. Atti
Accad. Naz. Lincei Rend. Lincei Mat. Appl. 26 (2015), no. 2, 223–240.
[7] H. Brezis, H.-M. Nguyen. Non-convex, non-local functionals converging to the
total variation. C. R. Math. Acad. Sci. Paris 355 (2017), no. 1, 24–27.
[8] H. Brezis, H.-M. Nguyen. Non-local Functionals Related to the Total Variation
and Connections with Image Processing. Ann. PDE 4 (2018), no. 1, 4:9.
[9] A. Chambolle. Image segmentation by variational methods: Mumford and Shah
functional and the discrete approximations. SIAM J. Appl. Math. 55 (1995), no. 3,
827–863.
[10] A. M. Garsia, E. Rodemich. Monotonicity of certain functionals under rear-
rangement. Ann. Inst. Fourier (Grenoble) 24 (1974), no. 2, vi, 67–116.
[11] M. Gobbino. Finite difference approximation of the Mumford-Shah functional.
Comm. Pure Appl. Math. 51 (1998), no. 2, 197–228.
[12] M. Gobbino, M. G. Mora. Finite-difference approximation of free-discontinuity
problems. Proc. Roy. Soc. Edinburgh Sect. A 131 (2001), no. 3, 567–595.
[13] H.-M. Nguyen. Some new characterizations of Sobolev spaces. J. Funct. Anal.
237 (2006), no. 2, 689–720.
34
[14] H.-M. Nguyen. Γ-convergence and Sobolev norms. C. R. Math. Acad. Sci. Paris
345 (2007), no. 12, 679–684.
[15] H.-M. Nguyen. Further characterizations of Sobolev spaces. J. Eur. Math. Soc.
(JEMS) 10 (2008), no. 1, 191–229.
[16] H.-M. Nguyen. Γ-convergence, Sobolev norms, and BV functions. Duke Math.
J. 157 (2011), no. 3, 495–533.
[17] H.-M. Nguyen. Estimates for the topological degree and related topics. J. Fixed
Point Theory Appl. 15 (2014), no. 1, 185–215.
[18] A. C. Ponce. A new approach to Sobolev spaces and connections to Γ-
convergence. Calc. Var. Partial Differential Equations 19 (2004), no. 3, 229–255.
[19] H. Taylor. Rearrangements of incidence tables. J. Combinatorial Theory Ser. A
14 (1973), 30–36.
35
|
1209.3624 | 1 | 1209 | 2012-09-17T11:02:22 | The $\epsilon-\epsilon^\beta$ property, the boundedness of isoperimetric sets in $\R^N$ with density, and some applications | [
"math.FA"
] | We show that every isoperimetric set in R^N with density is bounded if the density is continuous and bounded by above and below. This improves the previously known boundedness results, which basically needed a Lipschitz assumption; on the other hand, the present assumption is sharp, as we show with an explicit example. To obtain our result, we observe that the main tool which is often used, namely a classical "\epsilon-\epsilon" property already discussed by Allard, Almgren and Bombieri, admits a weaker counterpart which is still sufficient for the boundedness, namely, an "\epsilon-\epsilon^\beta" version of the property. And in turn, while for the validity of the first property the Lipschitz assumption is essential, for the latter the sole continuity is enough. We conclude by deriving some consequences of our result about the existence and regularity of isoperimetric sets. | math.FA | math |
THE ε − εβ PROPERTY, THE BOUNDEDNESS OF ISOPERIMETRIC SETS
IN RN WITH DENSITY, AND SOME APPLICATIONS
E. CINTI AND A. PRATELLI
Abstract. We show that every isoperimetric set in RN with density is bounded if the density
is continuous and bounded by above and below. This improves the previously known bound-
edness results, which basically needed a Lipschitz assumption; on the other hand, the present
assumption is sharp, as we show with an explicit example. To obtain our result, we observe
that the main tool which is often used, namely a classical "ε − ε" property already discussed
by Allard, Almgren and Bombieri, admits a weaker counterpart which is still sufficient for the
boundedness, namely, an "ε − εβ" version of the property. And in turn, while for the validity
of the first property the Lipschitz assumption is essential, for the latter the sole continuity is
enough. We conclude by deriving some consequences of our result about the existence and
regularity of isoperimetric sets.
1. Introduction
This paper deals with the isoperimetric problem in RN with density. More precisely, we
consider a given l.s.c. function f : RN → R+ (the "density") and we define the weighted volume
and perimeter of a set E ⊆ RN as
(cid:90)
(cid:90)
E
Ef :=
f (y) dH N−1(y) ,
f (x) dH N (x) ,
∂∗E
Pf (E) :=
(1.1)
where for every set E of locally finite perimeter we denote as usual by ∂∗E its reduced boundary,
while Pf (E) = +∞ for every set which is not locally of finite perimeter (the basic properties of
sets of finite perimeter will be briefly recalled in Section 1.2). The isoperimetric problem, then,
consists in searching for sets of minimal (weighted) perimeter among those of fixed (weighted)
volume. The study of the isoperimetric problem in RN with density has been deeply studied in
last years, also because of its close connection with the isoperimetric problems on Riemannian
manifolds (a short and incomplete list is [9, 10, 13, 14, 17, 22, 24, 25]).
The three main questions which one wants to understand are usually the existence, the
boundedness, and the regularity of isoperimetric sets (only in very specific examples one can
try to determine explicitely the minimizers). Concerning the boundedness, it is to be pointed
out that it is not only interesting by itself, but it is also important when proving the existence
(roughly speaking, when trying to show the existence of an isoperimetric set for volume m, it is
useful to know that there do not exist unbounded isoperimetric sets for volumes less than m).
We will be able to give some new results concerning all the three questions; in particular,
we will find a sharp boundedness theorem (Theorem 1.1).
One important tool in many of the works on isoperimetric problems is a classical "ε − ε"
property already discussed by Allard, by Almgren, and by Bombieri since the 1970's (see for
1
2
E. CINTI AND A. PRATELLI
instance [1, 2, 3, 4, 8]); this property basically means that a certain set can be locally modified
in order to increase its volume by ε, while the perimeter increases at most by Cε (we leave the
formal definitions to Section 1.1). This property is trivial to establish if the density and the set
are supposed to be regular enough, but its validity is also known if the set is even just of locally
finite perimeter and the density is only Lipschitz continuous (see [21, 23]); on the other hand,
it is very easy to observe that the validity may fail as soon as the density is not Lipschitz. This
is more or less the reason why most of the different boundedness and regularity results in this
context use at least a Lipschitz assumption on the density.
N (this is the content of our Theorem A). Then, our main result implies that the ε−ε
In this paper, we start from the following observation. It is classical and very easy to prove
that, if an isoperimetric set fulfills the ε − ε property, then it must be bounded; but in fact, to
get its boundedness, a weaker property is really needed, namely, the "ε − εβ" property, which
states that it is possible to increase its volume by ε and its perimeter at most by Cεβ, for some
N−1
β ≥ N−1
N
property is always true for any set of locally finite perimeter whenever the density is continuous.
N < β ≤ 1 such that,
Moreover, for every 0 < α ≤ 1 there exists some β = β(α, N ), with N−1
if the density is C0,α, then the ε − εβ property holds (this will be proved in Theorem B).
Putting together the two results, the following consequence will be immediate (the meaning of
"essentially bounded" density is clarified in Definition 1.7, anyway any density which is bounded
from above and below is essentially bounded).
Theorem 1.1. Assume that f is continuous and essentially bounded. Then every isoperimetric
set is bounded.
We underline that this result is sharp.
In fact, many examples (see for instance those
in [24]) show the existence of unbounded isoperimetric sets for densities which are unbounded
from above or from below; and on the other hand, in Section 4 we are able to build the example
of an unbounded isoperimetric set for a density which is bounded both from above and from
below but not continuous. As we said above, Theorem 1.1 is stronger than the previously known
results, which all needed at least a Lipschitz assumption. More precisely, to give a comparison,
we just recall the following very recent boundedness result.
Theorem 1.2 ([24], Corollary 5.11). Let E be an isoperimetric set in RN with a C1 density f .
Then E is bounded if any of the following three hypotheses hold:
(1) N = 2 and f is increasing, or
(2) f is radial and increasing, or
(3) Df ≤ Cf .
Let us now briefly pass to describe our contributions to the questions of the existence and
regularity of isoperimetric sets, which come as applications of Theorems A and B. Concerning
the existence, we will only recall that some existence results available in the literature require,
as an a priori assumption, the boundedness of the isoperimetric sets. Since all these results
concern densities which are essentially bounded and continuous, the boundedness assumption
can be removed, because it now directly follows from Theorem 1.1.
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
3
Concerning the regularity, instead, we will recall some very classical results, and we check
their consequences in view of the ε − εβ property that we have established. The theorem that
we obtain, Theorem 5.7, says that if the density f is essentially C0,α then any isoperimetric set
is of class C0,
2N (1−α)+2α . Stronger regularity results are contained in the forthcoming paper [11].
Observe that, as usual, results on RN with density admit counterparts in the context of
Riemannian manifolds; however, we do not study the extension here (for an overview of the
known results in this direction, see for instance [21]).
α
The plan of the paper is the following. In Section 1.1 we give all the formal definitions and
the claims of our main results, while in Section 1.2 we recall some basic properties of the sets of
finite perimeter. Then, in Sections 2 and 3 we give the proof of Theorems A and B. Later on, in
Section 4 we give the example of a situation where an isoperimetric set is unbounded, while the
density is essentially bounded but not continuous. Finally, in Section 5 we discuss the questions
of the existence and regularity of isoperimetric sets.
1.1. Preliminary definitions and claims of the main theorems. This section is devoted
to present the relevant definitions that we will need during the paper, and to claim our main
results.
We consider a given l.s.c. function f : RN → R+ = [0, +∞], such that the points x for which
f (x) = 0 or f (x) = +∞ are locally finite, and we work with the weighted notion of volume and
f := f H k for any k ∈ {0, . . . , N}, so that
perimeter given by (1.1). For brevity we will denote H k
definition (1.1) can be rewritten as Ef = H N
(∂∗E); given an open
set A ⊆ RN , we will denote the relative perimeter of E in A as Pf (E, A) = H N−1
(A ∩ ∂∗E).
Sometimes we will need to consider the Euclidean volume, or perimeter, or relative perimeter,
of a set E, which will be denoted by Eeucl, Peucl(E) or Peucl(E, A) respectively (while we will
not use the notation P (E) or E to avoid ambiguity). We will always call BR the open ball
centered at the origin and with radius R, and BR(x) the ball centered at x and with radius R.
The next definition is sometimes useful.
Definition 1.3. For every set E ⊆ RN , we define the spherical rearrangement E∗ ⊆ RN as
f (E) and Pf (E) = H N−1
f
f
where g : R+ → R+ is the unique function such that for every R > 0 one has
E∗ :=
(cid:110)
x ∈ RN : x1 ≥ g(cid:0)x(cid:1)(cid:111)
(cid:1) = H N−1(cid:0)E ∩ ∂BR
,
(cid:1) .
H N−1(cid:0)E∗ ∩ ∂BR
The definition of spherical rearrangement is not so much useful for a generic density f , but
it becomes very important when f is radial, thanks to the following result, whose proof can be
found for instance in [24] or [16].
Theorem 1.4. Assume that f is radial and E ⊆ RN . Then, one has
(cid:12)(cid:12)E∗(cid:12)(cid:12)f =(cid:12)(cid:12)E(cid:12)(cid:12)f ,
Pf (E∗) ≤ Pf (E) .
4
E. CINTI AND A. PRATELLI
Let us give now the particular definitions that we will use in this paper.
Definition 1.5. Let E ⊆ RN be a set of locally finite perimeter, 0 ≤ β ≤ 1, and C > 0. We say
that E fulfills the ε − εβ property with constant C if there exist a ball B and a constant ¯ε > 0
such that, for every ε < ¯ε, there is a set F ⊆ RN such that
F(cid:52)E ⊂⊂ B ,
Ff − Ef = ε ,
Pf (F ) ≤ Pf (E) + Cεβ .
We give also the following simple definition, which will only be used within the subsequent
(cid:9)
Definition 1.7.
Definition 1.6. A family(cid:8)Uδ
has H N−1(cid:0)∂∗E ∩ Uδ
(cid:1) > 0 for some arbitrarily small δ.
(cid:1) = 0.
H N−1(cid:0)RN \(cid:83)
a well-decreasing family(cid:8)Uδ
bounded, and there exist a well-decreasing family (cid:8)Uδ
δ>0 Uδ
(cid:9)
In the paper, we will always assume one of the following hypotheses on f .
Definition 1.7. The l.s.c. function f : RN → R+ is said to be essentially bounded if there exist
M < f (x) < M for every
x ∈ Uδ. Analogously, f is said to be essentially α-Holder for some 0 ≤ α ≤ 1 if f is essentially
δ>0 and constants M = M (δ) such that
f (x)− f (y) ≤ Mx− yα for every x, y ∈ Uδ. Observe that f is essentially α-Holder with α = 0
if and only if it is essentially bounded.
δ>0 and constants M = M (δ) such that 1
(cid:9)
δ>0 of open subsets of RN is said well-decreasing if one has that
RN \ Uδ is bounded for every δ > 0, and for any measurable set E (cid:54)= ∅ of finite perimeter one
Notice that, for instance, the sets Uδ = RN \ Bδ form a well-decreasing family; more in
general, for a family of open sets with RN \ Uδ bounded, to be well-decreasing it is enough that
Notice that, clearly, if f is bounded by above and below then it is also essentially bounded,
and similarly if it is α-Holder continuous then it is also essentially α-Holder, hence we could
have simply restricted our attention to standard bounded or Holder densities. Nevertheless, we
prefer to use the above slightly more complicate definitions for two reasons. First of all, this
allows to consider also the typical class of examples where the density may vanish of explode at
the origin, such as
f (x) =
xp
1 + xp ,
or
f (x) = 1 +
1
xp
for p > 0. Second, as one can see later, this choice does not effect at all any of the proofs, so we
can obtain slightly stronger results for free.
Another comment deserves to be done concerning the choice of considering only densities
which are (at least in the essential sense) bounded both from above and from below. In fact, this
is unavoidable, since many examples (see for instance those in [24]) enlighten that both existence
and boundedness can easily fail otherwise, thus no general result can be obtained without this
assumption.
We are now in position to claim our two main theorems.
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
5
Theorem A (Boundedness of isoperimetric sets). Assume that f is essentially bounded and that
E is an isoperimetric set fulfilling the ε− εβ property, either with β > N−1
N and some C > 0, or
with β = N−1
N and every small C > 0. Then, E is bounded.
Notice that, in the above theorem, we do not need any continuity assumption on f : the sole
ε−εβ property together with the essential boundedness of f is enough to ensure the boundedness
of any isoperimetric set E.
Theorem B (The ε − εβ property). Assume that f is essentially α-Holder for some 0 ≤ α ≤ 1.
Then every set E of locally finite perimeter fulfills the ε − εβ property with some C, where β is
defined by
β = β(α, N ) :=
.
(1.2)
α + (N − 1)(1 − α)
α + N (1 − α)
Moreover, if f is essentially bounded and continuous, then every set E of locally finite perimeter
still fulfills the ε − εβ property with β = N−1
N and with any constant C > 0.
Notice that, once N is fixed, β is a continuous and strictly increasing function of α with
N for α = 0, and β = 1 for α = 1, and moreover β > α for every 0 ≤ α < 1. Notice also
β = N−1
that Theorem 1.1 is an immediate consequence of the two above theorems.
Remark 1.8. It is essential to underline a very delicate point in the claim of Theorem B,
namely, that in the case when f is just essentially bounded and continuous, the ε − εβ property
with β = N−1
N holds true for every constant C > 0 (of course, when C becomes smaller, so
in fact, as the proof
does the constant ¯ε in Definition 1.5). This is of primary importance:
of Theorem A will enlighten, when a set fulfills the ε − εβ property with some β > N−1
N , then
the constant C is unessential; on the other hand, if β = N−1
N , then it is fundamental that the
constant C can be chosen arbitrarily small. And indeed, if f is essentially bounded but not
N−1
continuous then the ε − ε
N property holds true, but not with any constant C, and hence one
cannot apply Theorem A and in fact an isoperimetric set can be unbounded, as we will show
with the example in Section 4.
1.2. Basic properties of sets of finite perimeter. Since a basic knowledge of the theory of
sets of finite perimeter is needed for the proof of our results, we recall here very briefly what we
are going to use. For a general tractation of this subject, and for the proof of all the claims of
this section, the interested reader should refer for instance to [6].
Let then E ⊆ RN be a set of locally finite volume. We say that E is of locally finite perimeter
if the characteristic function χE of E is a BVloc function. In other words, we require µE := DχE
to be a vector valued and locally finite Radon measure. If E is a set of locally finite perimeter,
one defines the reduced boundary ∂∗E of E as the set of all those points x ∈ RN such that there
exists a (necessarily unique) vector νE(x) ∈ SN−1 with the property that
(cid:12)(cid:12)E ∩ B(x, r)(cid:12)(cid:12)eucl
(cid:12)(cid:12)B(x, r)(cid:12)(cid:12)eucl
lim
r(cid:38)0
= lim
r(cid:38)0
(cid:12)(cid:12)(cid:12)E ∩ B(x, r) ∩(cid:8)y : (y − x) · νE(x) < 0(cid:9)(cid:12)(cid:12)(cid:12)eucl
(cid:12)(cid:12)B(x, r)(cid:12)(cid:12)eucl
=
1
2
.
The vector νE(x) is called the (measure theoretical) outer normal of E at x.
6
E. CINTI AND A. PRATELLI
One can prove that µE coincides with the vector valued measure νE(x)H N−1 ∂∗E; finally,
one says that the (Euclidean) perimeter of E is defined as
Peucl(E) = µE(RN ) = H N−1(cid:0)∂∗E(cid:1) =
(cid:90)
∂∗E
1 dH N−1(x) .
It is easy to show that, if the set E is regular enough, then this general notion of perimeter
coincides with the usual perimeter, the reduced boundary coincides with the usual topological
boundary, and the measure theoretical outer normal coincides with the usual outer normal.
However, there exist sets of finite perimeter for which the topological boundary and the reduced
boundary do not coincide; for instance, the set of the points in RN with rational coordinates has
an empty reduced boundary and so null perimeter, but its topological boundary is the whole
RN . We conclude by recalling three classical results, that we will use extensively in the sequel.
Theorem 1.9 (Blow-up Theorem). Let E ⊆ RN and x ∈ ∂∗E. For every ε > 0, define the
ε (E − x), and call µε := µEε and H = {x ∈ RN : x · ν(x) < 0}. Then, when
blow-up set Eε := 1
ε (cid:38) 0, one has that the sets Eε converge to H in the L1
loc sense, while the measures µε (resp.,
µε) weak* converge to the measure νE(x)H N−1 ∂H (resp., H N−1 ∂H).
To state the following result, the Vol'pert Theorem, we need a simple preliminary piece of
notation.
Definition 1.10. Let E ⊆ RN be a Borel set. We define the vertical section of E at any level
y ∈ RN−1, and the horizontal section of E at any level t ∈ R as
Ey := {t ∈ R : (y, t) ∈ E} ,
Et := {y ∈ RN−1 : (y, t) ∈ E} .
y ∈ RN−1 the vertical section Ey is a set of finite perimeter in R, and ∂∗(Ey) = (cid:0)∂∗E(cid:1)
Theorem 1.11 (Vol'pert). Let E be a set of locally finite perimeter. Then, for H N−1-a.e.
y.
Analogously, for H 1-a.e. t ∈ R the horizontal section Et is a set of finite perimeter in RN−1,
and ∂∗(Et) = (∂∗E)t up to an H N−2-negligible set.
The proof of this result can be found in [6, 28] for the case of the vertical sections, while the
analogous for the horizontal sections (or in general, for sections of any codimension) is proved
in [16, 15, 7].
x ∈ ∂∗E, the outer normal of E at x as νE(x) =(cid:0)ν(cid:48)(x), νN (x)(cid:1) ∈ RN−1 × R. For every Borel
Theorem 1.12 (Coarea formula). Let E be a set of locally finite perimeter and denote, for
(cid:90)
(cid:90)
function g : RN → R+ it is
(cid:112)
and analogously
g(x)
∂∗E
1 − νN (x)2 dH N−1(x) =
(cid:112)
1 − ν(cid:48)(x)2 dH N−1(x) =
(cid:90)
g(x)
∂∗E
(cid:90)
(cid:90) +∞
(cid:90)
−∞
RN−1
∂∗Ey
g(y, t) dH N−2(y) dt ,
∂∗Et
g(y, t) dH 0(t) dH N−1(y) .
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
7
2. Proof of Theorem A
This section is entirely devoted to give a proof of Theorem A.
Proof (of Theorem A). Let E be an isoperimetric set. Since f is essentially bounded, we can
M ≤ f ≤ M outside BR(cid:48). Let us now use the assumption
take some R(cid:48) > 0 and M > 1 such that 1
that E fulfills the ε − εβ property, finding constants C, R(cid:48)(cid:48), ¯ε > 0 with R(cid:48)(cid:48) ≥ R(cid:48) such that, for
every 0 < ε < ¯ε, there exists a set F with F = E outside the ball BR(cid:48)(cid:48), and
Pf (F ) ≤ Pf (E) + Cεβ ;
Ff = Ef + ε ,
(2.1)
in addition, we are also allowed to assume
C ≤ 2N ω1/N
N
M 2
Let us now define, for every R > R(cid:48)(cid:48),
(cid:12)(cid:12)f ,
ε(R) :=(cid:12)(cid:12)E \ BR
(cid:90)
(cid:90) ∞
ε(R) =
R
∂Br
g(R) := H N−1
f
χE(x)f (x) dH N−1(x) dr =
if β =
.
(2.2)
N − 1
N
(cid:90)
(cid:1) =
(cid:0)E ∩ ∂BR
E∩∂BR
(cid:0)E ∩ ∂BR
(cid:90) ∞
H N−1
f
R
f (x) dH N−1(x) ;
(cid:1) dr =
(cid:90) ∞
R
g(r)dr ,
the function R (cid:55)→ ε(R) is decreasing and goes to 0 as R goes to ∞. Moreover, observe that
hence ε ∈ W1,1
competitor (cid:101)E by cutting away the part of E which is outside the ball BR, and then using the
loc(R(cid:48)(cid:48), +∞) and ε(cid:48)(R) = −g(R). Pick now any R > R(cid:48)(cid:48): we can consider a
ε − εβ property to recover the correct volume without increasing too much the perimeter. More
(cid:1) + 2g(R) . (2.3)
(cid:0)E ∩ BR
(cid:1) = Pf (E)− Pf
precisely, first of all we notice that
(cid:1) ≥ 1
(cid:0)E \ BR
(cid:1) = Pf (E)− Pf
(cid:12)(cid:12) N−1
(cid:0)E ∩ ∂BR
(cid:12)(cid:12)E \ BR
Then, we apply the standard Euclidean isoperimetric inequality and the bounds on f to get
(cid:0)E \ BR
N ω1/N
N−1
N ,
N ω1/N
N ε(R)
1
2N−1
eucl ≥
N
Peucl
Pf
Pf
N
M
N
f
(cid:0)E \ BR
(cid:1) + 2H N−1
(cid:0)E \ BR
(cid:1) ≥ 1
(cid:1) ≤ Pf (E) −
M
M
(cid:0)E ∩ BR
Pf
1
2N−1
N
M
N ω1/N
N ε(R)
N−1
N + 2g(R) .
(2.4)
which inserted into (2.3) gives
We apply now (2.1) with ε = ε(R) to find a set F ⊆ RN with Ff = Ef + ε and Pf (F ) ≤
Pf (E) + Cεβ. Since F coincides with E outside BR(cid:48)(cid:48) and R > R(cid:48)(cid:48), we can finally define the
competitor (cid:101)E as F ∩ BR. By construction we find
(cid:12)(cid:12)(cid:101)E(cid:12)(cid:12)f =(cid:12)(cid:12)E ∩ BR
(cid:12)(cid:12)f + ε(R) = Ef ,
(cid:1) + Pf (F ) − Pf (E) ≤ Pf (E) − N ω1/N
N
2N−1
while by (2.4) it is also
(cid:0)(cid:101)E(cid:1) = Pf
(cid:0)E ∩ BR
Pf
N−1
N + 2g(R) + Cε(R)β .
ε(R)
M
N
8
E. CINTI AND A. PRATELLI
Since E is an isoperimetric set and (cid:101)E has the same volume as E, we derive that for every R > R(cid:48)(cid:48)
it must be Pf ((cid:101)E) ≥ Pf (E), which implies
g(R) ≥ N ω1/N
ε(R)β .
ε(R)
(2.5)
N
2N−1
N
2M
N−1
N − C
2
We claim then what follows: there exist two positive constants γ and ε ≤ ¯ε such that
g(R) ≥ γε(R)
N−1
N
∀ R > R(cid:48)(cid:48) : ε(R) ≤ ε .
(2.6)
It is immediate to prove the validity of this estimate by considering separately the case β = N−1
and the case β > N−1
N . Indeed, in the first case (2.6) is just an immediate consequence of (2.5)
N−1
N (cid:29) εβ when ε is small enough,
and of the choice (2.2). On the other hand, if β > N−1
then (2.6) readily follows by (2.5). Since R (cid:55)→ ε(R) is a continuous and decreasing function, we
can select R(cid:48)(cid:48)(cid:48) ≥ R(cid:48)(cid:48) such that ε(R) ≤ ε for every R ≥ R(cid:48)(cid:48)(cid:48).
N , then ε
Let us now directly show the boundedness of E:
if E ⊆ BR(cid:48)(cid:48)(cid:48), then there is nothing to
prove; otherwise, let j ∈ N be such that ε(R(cid:48)(cid:48)(cid:48)) ≥ 2−j, and for every i ≥ j let R(i) be such that
ε(Ri) = 2−i. Then, recalling (2.6) and the fact that ε(cid:48) = −g, for every i ≥ j we can evaluate
Ri+1 − Ri
2(i+1) N−1
2i+1 = ε(cid:0)Ri
2i − 1
(cid:1) − ε(cid:0)Ri+1
(cid:90) Ri+1
(cid:90) Ri+1
N−1
N dR ≥ γ
g(R) dR ≥
(cid:1) =
γε(R)
Ri
Ri
N
N
,
1
1
2i+1 =
from which we deduce that
and in turn this implies that ε(R) = 0 for every R ≥ R∞, that is, E ⊆ BR∞ is bounded.
(cid:3)
3. Proof of Theorem B
In this section we give the proof of our main result, Theorem B. Since the proof is quite
involved, we have divided it for simplicity in three parts and several steps.
Proof (of Theorem B). Let us consider a function f as in the claim, and an isoperimetric set E.
In the first part, we will show that the ε− εα property holds. Since α < β unless β = α = 1, the
property is in fact weaker than what we need; nevertheless, we prefer to start with this somehow
easier case, because the proof of the stronger ε − εβ property, which will be done in the second
part, will be a careful modification of the same argument. And in turn, also the case when f is
only essentially continuous will eventually be treated, in the third part, with the same strategy.
By Definition 1.7, there exists an open set U ⊆ RN with U ∩ ∂∗E (cid:54)= ∅ and such that, for a
suitable M > 1, one has
< f (x) < M ∀ x ∈ U ,
f (x) − f (y) ≤ Mx − yα ∀ x , y ∈ U .
1
M
Let ¯x be a point in U ∩ ∂∗E. We can assume for simplicity that ¯x coincides with the origin in
RN , and that the outer normal of E at ¯x is the vertical direction ν(¯x) = (0, 1) ∈ RN−1 × R.
This immediately implies that, for every (cid:96) ≥ j,
Ri+1 − Ri ≤ 1
γ
2− i+1
N .
R(cid:96) ≤ Rj +
1
γ
(cid:0)2− 1
N(cid:1)i+1 ≤ Rj +
1
1
γ
(cid:0)2− 1
N(cid:1)i+1 =: R∞ < +∞ ,
1
(cid:96)−1(cid:88)
i=j
∞(cid:88)
i=j
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
9
Part I. The ε − εα property.
We start considering the case when f is essentially α-Holder, and prove the ε− εα property; the
proof is divided in many steps.
Step (i). Choice of the cube QN .
In this first step, we will select a suitably small constant a, and from now on we will restrict
our attention to the cube QN = (−a/2, a/2)N , which is entirely contained inside U as soon as
a (cid:28) 1. Let us denote by Q = (−a/2, a/2)N−1 the horizontal cube, and by ϕ : RN → RN the
constant vector field ϕ ≡ (0, 1) ∈ RN−1 × R; let moreover ρ > 0 be a sufficiently small constant,
that will be precised later. We aim to choose a > 0 such that QN is contained in the open set
U defined above, and moreover all the following properties hold:
≤ 1 + ρ ,
aN−1
aN−1
1 − ρ ≤ H N−1(cid:0)∂∗E ∩ QN ∩ {−aρ < xN < aρ}(cid:1)
H N−1(cid:0)∂∗E ∩ QN \ {−aρ < xN < aρ}(cid:1)
H N(cid:0)E ∩ QN ∩ {xN < 0}(cid:1)
H N(cid:0)E ∩ QN ∩ {xN > 0}(cid:1)
(cid:90)
∂∗E ∩(cid:0)∂Q × (−a/2, a/2)(cid:1)(cid:17) ≤ 2N +1aN−2 .
ϕ · dµE ≥ (1 − ρ)aN−1 ,
H N−2(cid:16)
≥ 1 − ρ ,
≤ ρ ,
aN /2
aN /2
≤ ρ ,
QN
(3.1)
(3.2)
(3.3)
(3.4)
(3.5)
(3.6)
(cid:90) ¯a/2
We show now that such a choice of a is possible. Indeed, the first five conditions (3.1) -- (3.5) are
true for every a small enough, say a ≤ ¯a, as a direct consequence of the blow-up Theorem 1.9.
It remains then only to show that there exists some a ≤ ¯a satisfying also condition (3.6). To do
so observe that, also by (3.1) and (3.2),
∂∗E ∩(cid:0)∂(−t, t)N−1 × (−¯a/2, ¯a/2)(cid:1)(cid:17)
H N−2(cid:16)
∂∗E ∩(cid:0)∂Q × (−a/2, a/2)(cid:1)(cid:17) ≤ H N−2(cid:16)
dt ≤ H N−1(cid:16)
∂∗E ∩(cid:0) − ¯a/2, ¯a/2(cid:1)N(cid:17)
≤(cid:0)1 + 2ρ(cid:1)¯aN−1 .
∂∗E ∩(cid:0)∂(−a/2, a/2)N−1 × (−¯a/2, ¯a/2)(cid:1)(cid:17)
Therefore, there exists a ∈ (¯a/2, ¯a) for which
H N−2(cid:16)
¯a/4
(cid:0)1 + 2ρ(cid:1)¯aN−1
≤ 4
≤ 5¯aN−2 ≤ 2N +1aN−2 .
¯a
E ∩(cid:0)∂Q × (−a/2, a/2)(cid:1)(cid:17)
∂∗E ∩(cid:0)∂Q × (−a/2, a/2)(cid:1) = ∂∗(cid:16)
Notice that, thanks to Vol'pert Theorem, without loss of generality we can assume
H N−2 − a.e. .
(3.7)
We have then found some a ≤ ¯a for which also (3.6) holds true. This concludes the first step.
Step (ii). Definition of A, B, G and Γ.
In this step, we subdivide the horizontal cube Q into four sets A, B, G and Γ, depending on
the properties of ∂∗(E ∩ QN )x(cid:48). Since in the whole proof we are concentrated only on what
10
E. CINTI AND A. PRATELLI
happens inside QN , we will always consider the horizontal and vertical sections inside the cube,
even without specifying it; in other words, we will write Ey or Et (respectively ∂∗Ey or ∂∗Et)
instead of (E ∩ QN )y or (E ∩ QN )t (respectively ∂∗(E ∩ QN )y or ∂∗(E ∩ QN )t). This is a slight
abuse of notation, but it will simplify a lot the formulas in the rest of the proof. The sets are
the following
A := {x(cid:48) ∈ Q : ∂∗(Ex(cid:48)) (cid:54)= (∂∗E)x(cid:48)} ,
B := {x(cid:48) ∈ Q \ A : H 0(∂∗Ex(cid:48)) = 0} ,
G := {x(cid:48) ∈ Q \ A : H 0(∂∗Ex(cid:48)) = 1, ∂∗Ex(cid:48) ⊆ (−aρ, aρ), Ex(cid:48) ⊆ (−a/2, aρ)} ,
Γ := Q \ (A ∪ B ∪ G) .
Let us briefly discuss the meaning of these sets: thanks to Step (i), we can imagine E ∩ QN to
be close to the half-cube QN ∩ {xN < 0}, thus we expect the vertical sections Ex(cid:48) to be close to
(−a/2, 0). The "good" set G is precisely the set of those x(cid:48) ∈ Q for which this holds, namely,
Ex(cid:48) is a "lower" segment starting at −a/2 and ending between −aρ and aρ. All the other x(cid:48) ∈ Q
are then contained in the "bad" sets A, B and Γ. More precisely, A collects those x(cid:48) for which
Vol'pert Theorem does not hold true (keep in mind that we know by Theorem 1.11 that A is
H N−1 negligible, but this does not imply that the sections corresponding to A do not carry
perimeter!). Instead, B is the set of the sections which have no boundary, thus are either the
full segment (−a/2, a/2), or they are empty. Finally, Γ is the set of the sections having some
boundary, but not contained in G. Observe that this can happen for several different reasons: if
∂∗Ex(cid:48) contains exactly one point, then either this point is not between −aρ and aρ, or Ex(cid:48) is an
"upper" segment starting between −aρ and aρ, and ending at a/2. On the other hand, if ∂∗Ex(cid:48)
has more than one point, then the points can be finitely many or infinitely many. We further
subdivide Γ in four subsets according to the above possibilities, namely, we define
Γ0 :=(cid:8)x(cid:48) ∈ Γ : H 0(∂∗E)x(cid:48) = 1, ∂∗Ex(cid:48) /∈ (−aρ, aρ)(cid:9) ,
Γ1 :=(cid:8)x(cid:48) ∈ Γ \ Γ0 : H 0(∂∗E)x(cid:48) = 1(cid:9) ,
Γ2 :=(cid:8)x(cid:48) ∈ Γ \ (Γ0 ∪ Γ1) : ∂∗Ex(cid:48) contains a finite number of points(cid:9) ,
Γ3 := Γ \(cid:0)Γ0 ∪ Γ1 ∪ Γ2
(cid:1) .
The aim of this step is to show that G fills a big portion of Q, and that the perimeter of E in the
sections not belonging to G is extremely small. Let us start by observing that, thanks to (3.5),
one has
(1 − ρ)aN−1 ≤
ϕ · dµE =
ϕ · dµE +
ϕ · dµE
A×(−a/2,a/2)
B×(−a/2,a/2)
ϕ · dµE +
G×(−a/2,a/2)
Γ×(−a/2,a/2)
(3.8)
ϕ · dµE .
(cid:90)
QN
(cid:90)
(cid:90)
+
(cid:90)
(cid:90)
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
11
We have now to estimate each of the terms in the right-hand side of last inequality. First, since
by construction dµE = 0 on the set B × (−a/2, a/2), we have
(cid:90)
B×(−a/2,a/2)
ϕ · dµE = 0 .
(3.9)
We address now the integral on Γ0 × (−a/2, a/2). Recall that, as already observed, if x(cid:48) ∈ Γ0
and (x(cid:48), xN ) ∈ ∂∗E, then xN /∈ (−aρ, aρ). Therefore, using (3.2), we get
(cid:90)
(cid:90)
ϕ · dµE ≤
ϕ dµE = H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ0})
Γ0×(−a/2,a/2)
Γ0×(−a/2,a/2)
≤ H N−1(∂∗E ∩ QN \ {−aρ < xN < aρ}) ≤ ρaN−1 .
Concerning A, we just recall that by Vol'pert Theorem 1.11 it is
H N−1(A) = 0 .
Let us pass now to Γ3; as already observed, for every x(cid:48) ∈ Γ3 the set ∂∗(Ex(cid:48)) = (∂∗E)x(cid:48) contains
infinitely many points. Then, since for any K ≥ 1 it is clearly
H N−1(cid:16)(cid:8)x(cid:48) ∈ Q \ A : H 0(∂∗Ex(cid:48)) ≥ K(cid:9)(cid:17) ≤ 1
H N−1(∂∗E ∩ QN ) ,
K
(3.10)
(3.11)
(3.12)
(3.13)
by sending K → ∞ we derive
H N−1(Γ3) = 0 .
(cid:90)
Thanks to (3.11) and (3.12), the coarea formula (Theorem 1.12) directly gives
(Γ3∪A)×(−a/2,a/2)
QN∩∂∗E∩{x(cid:48)∈Γ3∪A}
(cid:90)
(cid:90)
(cid:90)
ϕ · dµE =
=
=
(0, 1) · νE dH N−1
(cid:113)
(cid:32)(cid:90)
1 − ν(cid:48)
E(x)2 dH N−1(x)
(cid:33)
1 dH 0(xN )
dH N−1(x(cid:48)) = 0 .
QN∩∂∗E∩{x(cid:48)∈Γ3∪A}
QN∩Γ3∪A
∂∗Ex(cid:48)
We address now Γ1. Recall that, by definition, if x(cid:48) ∈ Γ1 then ∂∗Ex(cid:48) = {p} with p = p(x) ∈
E(x(cid:48), p) < 1, that is, the levels
x(cid:48) such that the outer normal at (x(cid:48), p) is not horizontal. We remark the well known fact that
(−aρ, aρ). Call then (cid:101)Γ1 the set of those x(cid:48) ∈ Γ1 for which ν(cid:48)
H N−1(cid:0)Γ1 \(cid:101)Γ1
(cid:1) = 0. Using again the coarea formula, denoting by δp the Dirac mass at p ∈ R
we find
µE
(cid:0)(cid:101)Γ1 × (−a/2, a/2)(cid:1) =
δp ⊗ H N−1 (cid:101)Γ1
(cid:112)1 − ν(cid:48)
νE(x(cid:48), p) · (0, 1) δp ⊗ H N−1 (cid:101)Γ1 .
1
E(x(cid:48), p)2
1
=
E. CINTI AND A. PRATELLI
(cid:0)Γ1 × (−a/2, a/2)(cid:1)
(cid:0)(cid:101)Γ1 × (−a/2, a/2)(cid:1)
12
(cid:90)
Hence we have
Γ1×(−a/2,a/2)
ϕ · dµE =
=
=
for every x(cid:48) ∈(cid:101)Γ1.
Q
(0, 1) · νE(x) dµE
(cid:90)
(cid:90)
(0, 1) · νE(x) dµE
(cid:90)
(0, 1) · νE(x(cid:48), p)
(cid:101)Γ1
Q
Note that the "−" sign comes from that fact that νE(x(cid:48), p) has clearly a negative last component
The very same argument used for Γ1 can be repeated for G, recalling that for every x(cid:48) ∈ G
one has that ∂∗Ex(cid:48) = {q} with some q = q(x) ∈ (−aρ, aρ). Therefore, since νE(x(cid:48), q) has a
positive last component, in place of (3.14) we find now
(0, 1) · νE(x(cid:48), p) dH N−1(x(cid:48)) = −H N−1((cid:101)Γ1) = −H N−1(Γ1) .
(3.14)
(cid:90)
G×(−a/2,a/2)
(cid:90)
ϕ · dµE = H N−1(G) .
(3.15)
Finally, we address Γ2. First of all, recall that H 0(∂∗Ex(cid:48)) ≥ 2 for almost every x ∈ Γ2. Thus
H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2}) ≥
H 0(∂∗Ex(cid:48)) dH N−1(x(cid:48)) ≥ 2H N−1(Γ2) .
(3.16)
Γ2
Moreover, by definition, for every x(cid:48) ∈ Γ2 there exist k ∈ N and pi ∈ (−a/2, a/2), for i = 1, ..., k,
such that ∂∗Ex(cid:48) =(cid:83)k
i=1 pi. Let us call again(cid:101)Γ2 the set of those x(cid:48) for which all the corresponding
pi have a non-horizontal outer normal, where again H N−1(cid:0)Γ2 \(cid:101)Γ2
(cid:1) = 0. Using the coarea
µE ((cid:101)Γ2 × (−a/2, a/2)) = αx(cid:48) ⊗ H N−1 (cid:101)Γ2 ,
formula exactly as we did for Γ1 and G, we have that
where
αx(cid:48) =
1
νE(x(cid:48), pi) · (0, 1) δpi .
Therefore we have, analogously as in (3.14) or (3.15),
(cid:90)
i=1
k(cid:88)
(cid:90)
(cid:101)Γ2
k(cid:88)
i=1
ϕ · dµE =
Γ2×(−a/2,a/2)
(0, 1) · νE(x(cid:48), pi)
(0, 1) · νE(x(cid:48), pi) dH N−1(x(cid:48)) .
(3.17)
discuss carefully the signs. For x(cid:48) ∈ (cid:101)Γ2, we have that Ex(cid:48) = (cid:83)h
Recall that, in the last expression, the integer k and the points pi depend on x(cid:48). Observe now
that, in the right hand side of last equation, the integrand is always a finite sum of ±1; let us
j=1(bj, cj) is a finite union of
segments. Moreover, by construction {pi}i=1,...,k = {bj}j=1,...,h ∪{cj}j=1,...,h \{−a/2, a/2}, since
(0, 1) · νE(x(cid:48), bj)
(0, 1) · νE(x(cid:48), bj) = −1
we are working within the open cube QN . In addition, one has that
for every j = 1, ..., h such that bj (cid:54)= −a/2, since the normal vector at any point (x(cid:48), bj) has a
(0, 1) · νE(x(cid:48), cj)
(0, 1) · νE(x(cid:48), cj) = 1 for every j = 1, ..., h such
negative last component; similarly, we have
that cj (cid:54)= a/2. In conclusion, for every x(cid:48) ∈ Γ2 the value of the integrand in (3.17) is either
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
13
−1, or 0, or 1, depending whether or not b1 = −a/2 or ch = a/2. Therefore we have, also
recalling (3.16),
ϕ · dµE ≤ H N−1(Γ2) ≤ H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2})
2
.
(3.18)
Γ2×(−a/2,a/2)
(cid:90)
Combining (3.8), (3.9), (3.10), (3.13), (3.14), (3.15), and (3.18), we have
H N−1(cid:0)∂∗E ∩ {(x(cid:48), xN )x(cid:48) ∈ G}(cid:1) +
H N−1(cid:0)∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2}(cid:1)
1
2
H N−1(cid:0)∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2}(cid:1)
≥ H N−1(G) +
(cid:90)
=
G×(−a/2,a/2)
1
2
ϕ · dµE +
H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2}) ≥ (1 − 2ρ)aN−1 .
1
2
On the other hand, using (3.1) and (3.2) of we get the upper bound
H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ G}) + H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2})
≤ H N−1(∂∗E ∩ QN ) ≤ (1 + 2ρ)aN−1 .
Combining together (3.19) and (3.20), we deduce
H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ Γ2}) ≤ 8ρaN−1 .
This, together with (3.19) again, implies that
H N−1(∂∗E ∩ {(x(cid:48), xN ) : x(cid:48) ∈ G}) ≥ H N−1(G) ≥ (1 − 6ρ)aN−1 .
Finally, this last estimate implies on one hand, since A ∪ G ∪ B ∪ Γ = Q, that
(3.19)
(3.20)
(3.21)
(3.22)
(3.23)
(3.24)
and on the other hand, recalling (3.1) and (3.2), that
H N−1(A) + H N−1(B) + H N−1(Γ) ≤ 6ρaN−1 ,
H N−1(cid:16)
∂∗E ∩(cid:0)A ∪ B ∪ Γ(cid:1) × (−a/2, a/2)
(cid:17) ≤ 8ρaN−1 .
Step (iii). Definition of σ−, σ+ and F .
Definition of σ+. Let ¯δ (cid:28) ρa be a fixed number; we can take H :=
disjoint
horizontal strips Si := Q × (σi, σi + ¯δ) ⊆ QN with aρ < σi < a/2 − 2ρ for every 1 ≤ i ≤ H. By
assumptions (3.2) and (3.4), we have
¯δ
(cid:20) a(1/2 − 3ρ)
(cid:21)
H(cid:88)
i=1
aH N−1(cid:0)∂∗E ∩ Si
≤ aH N−1(cid:16)
(cid:1) + H N(cid:0)E ∩ Si
∂∗E ∩(cid:0)Q × [aρ, a/2)(cid:1)(cid:17)
(cid:1)
+ H N(cid:16)
E ∩(cid:0)Q × [aρ, a/2)(cid:1)(cid:17) ≤ 2ρaN .
Therefore there exists i ∈ {1, ..., H} such that
aH N−1(∂∗E ∩ Si) + H N (E ∩ Si) ≤ 2ρaN
H
≤ 6¯δρaN−1 ,
(3.25)
recalling that by definition H ≥ a
3¯δ
. We set σ+ := σi, for such an i.
To show that this is possible, we apply (3.2) and Vol'pert Theorem 1.11 to obtain that
(3.26)
14
E. CINTI AND A. PRATELLI
Definition of σ−. We now select a horizontal level σ− ∈ (−a/2,−aρ) such that ∂∗(Eσ−
(∂∗E)σ−
in the H N−2 sense, and
) =
H N−2(cid:0)∂∗Eσ−(cid:1)
ρ ≥ H N−1(cid:0)∂∗E ∩ QN ∩ {xN < −aρ}(cid:1)
(cid:18) 1
aN−1
H N−2(∂∗Et)
(cid:19)
(cid:90) −aρ
aN−2
=
− ρ
2
--
−a/2
aN−2
dt ≥ 1
3
≤ 3ρ .
(cid:90) −aρ
(cid:90) −aρ
−a/2
≥
--
−a/2
H N−2(∂∗Et)
aN−1
dt
H N−2(∂∗Et)
aN−2
dt ,
from which the existence of some σ− satisfying (3.26) immediately follows.
Definition of F . We want now to construct the competitor F . To do so, we take a constant
¯δ/(4M 2) ≤ δ ≤ ¯δ, and we define the set F = F (δ) as
x ∈ F ⇐⇒
x ∈ E \ QN ,
x ∈ E ∩ QN ∩(cid:0){xN ≤ σ−} ∪ {xN > σ+ + δ(cid:1) ,
(x(cid:48), σ−) ∈ E ∩ QN and σ− < xN ≤ σ− + δ ,
(x(cid:48), xN − δ) ∈ E ∩ QN and σ− + δ < xN ≤ σ+ + δ .
In words, we eliminate the intersection of E with the strip Q × (σ+, σ+ + δ) ⊆ S¯i, and we move
up of a distance δ all the part of E between the levels σ− and σ+. Notice that by definition
σ+ + δ < a/2, so nothing changes outside of QN .
Step (iv). Evaluation of the volume and perimeter of F .
We are now ready to evaluate the volume and the perimeter of the set F , in order to obtain the
ε − εα property.
Volume. By the definition of F , it is easy to expect that its volume should equal that of E
plus something similar to aN−1δ, since we are moving up of a distance δ a set within the cube
of (N − 1)-dimensional volume equal to aN−1. This is exactly what we are going to prove, but
some care is required since, in passing from E to F , we could also lose some volume, basically
for two reasons. First, because we are eliminating the strip Q× (σ+, σ+ + δ) ⊆ S¯i (with which E
has a small intersection, though). Second, because the density is not constant, and then there
is in principle the risk of moving the mass where the density is lower.
1 := E \ F ∩(cid:0)Q × (σ−, σ+)(cid:1), and E−
2 := E \ F ∩
2 ). By construction we have E+ ∩ E = ∅,
1 ∪ E−
(cid:0)Q × (σ+, σ+ + δ)(cid:1), so that F = E ∪ E+ \ (E−
Let us define the sets E+ := F \ E, E−
E−
1 ∩ E−
2 = ∅, and E−
1 ∪ E−
2 ⊆ E, thus
Ff − Ef = E+f − E−
1 f − E−
2 f .
Let us estimate the terms on the right-hand side of this equality, starting with E−
f ≤ M in QN and (3.25), and recalling that ¯δ/(4M 2) ≤ δ ≤ ¯δ, we have that
≤(cid:12)(cid:12)E ∩ S¯i
(cid:12)(cid:12)f ≤ M H N(cid:0)E ∩ S¯i
(cid:1) ≤ 6ρM aN−1¯δ
(cid:12)(cid:12)E−
2
(cid:12)(cid:12)f ≤(cid:12)(cid:12)(cid:12)E ∩(cid:0)Q × (σ+, σ+ + δ(cid:1)(cid:12)(cid:12)(cid:12)f
(3.27)
2 . Using that
(3.28)
≤ 24ρM 3aN−1δ .
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
15
We pass now to E+. Observe that E+ =(cid:83)
x(cid:48) ⊇(cid:83)
x(cid:48) and that, if x(cid:48) ∈ G, then E+
x(cid:48)
is a segment of length δ. Therefore, using (3.22) and recalling that f ≥ 1/M on QN , we deduce
x(cid:48)∈Q E+
x(cid:48)∈G E+
δaN−1(1 − 6ρ) .
E+f ≥ 1
M
δH N−1(G) ≥ 1
M
(3.29)
Since we need also an upper bound for E+f , we consider separately the sets G and Q \ G. In
G we have
(cid:12)(cid:12)E+ ∩ (G × (−a/2, a/2))(cid:12)(cid:12)f ≤ M δH N−1(G) ≤ M δaN−1 .
(3.30)
Recall now that Q\ G = A∪ B ∪ Γ. By definition of B, if x(cid:48) ∈ B then Ex(cid:48) is either empty or the
whole segment (−a/2, a/2): in both cases, E+
x(cid:48) = ∅. Therefore, recalling also (3.11) and (3.12),
we have
(cid:12)(cid:12)(cid:12)E+ ∩(cid:0)(A ∪ B ∪ Γ3) × (−a/2, a/2)(cid:1)(cid:12)(cid:12)(cid:12)f
Finally, observe that
(cid:12)(cid:12)(cid:12)(cid:0)E+ ∪ E−
1
Therefore, using (3.24), we deduce that
for every x(cid:48) ∈ Γ0 ∪ Γ1 ∪ Γ2 ,
1 )x(cid:48)
≤ M
(cid:90)
(cid:90)
(E+)x(cid:48) ∪ (E−
H 1(cid:16)
(cid:17)(cid:12)(cid:12)(cid:12)f
(cid:1) ∩(cid:16)(cid:0)Γ0 ∪ Γ1 ∪ Γ2
(cid:1) × (−a/2, a/2)
(cid:17)
H 1(cid:16)
≤ M δH N−1(cid:16)
(cid:1) × (−a/2, a/2)
∂∗E ∩(cid:0)Γ0 ∪ Γ1 ∪ Γ2
1 ∩(cid:0)(A ∪ B ∪ G ∪ Γ3) × (−a/2, a/2)(cid:1)(cid:12)(cid:12)(cid:12)f
(cid:12)(cid:12)(cid:12)E−
1 . Observe that if x(cid:48) ∈ B ∪ G then (E−
H 0(∂∗Ex(cid:48)) dH N−1(x(cid:48))
(E+)x(cid:48) ∪ (E−
dH N−1(x(cid:48))
Γ0∪Γ1∪Γ2
Γ0∪Γ1∪Γ2
≤ M δ
1 )x(cid:48)
= 0 .
(cid:17) ≤ δH 0(∂∗Ex(cid:48)) .
(cid:17) ≤ 8ρM δaN−1 .
(3.31)
(3.32)
It remains to estimate E−
as we did to get (3.31), using (3.11) and (3.12), we have
1 )x(cid:48) = ∅: then, arguing exactly
(3.33)
We have now all the ingredients to estimate Ff − Ef both from above and below, thanks
to (3.27). Indeed, on one hand, combining (3.29), (3.32), (3.33) and (3.28), and up to take ρ
sufficiently small, we get
= 0 .
Ff − Ef ≥ δaN−1
(1 − 6ρ) − 8ρM − 24M 3ρ
≥ δaN−1
2M
.
(3.34)
(cid:18) 1
M
(cid:19)
On the other hand, combining (3.30), (3.31), and (3.32), we also find
Ff − Ef ≤ E+f ≤ M δaN−1(1 + 8ρ) ≤ 2M δaN−1 .
(3.35)
Perimeter. We are then left to find an upper bound for Pf (F ) − Pf (E) in terms of δ. This
will be the only point in this Part where we are going to use the α-Holder assumption on f .
We start pointing out that the change in perimeter has four contributions. First, since we
move upwards the set E of a distance δ inside the cube QN , on the lateral boundary ∂Q ×
(−a/2, a/2) we are adding a surface T +
1 :=(cid:0)∂∗F \ ∂∗E(cid:1) ∩(cid:0)∂Q × (−a/2, a/2)(cid:1) .
1 of "height" δ, namely,
T +
2 as soon as ∂∗Eσ−
2 := ∂∗Eσ− × (σ−, σ− + δ) .
T +
16
E. CINTI AND A. PRATELLI
Second, since in the strip Q× (σ−, σ− + δ) the set F is defined as F = Eσ− × (σ−, σ− + δ), then
we are creating some new surface T +
is not empty. More precisely, we set
Third, since we are cutting away the set E ∩(cid:0)Q × (σ+, σ+ + δ)(cid:1), then we are removing some
3 in the strip, but at the same time we might also create some new surface T +
3 at the
surface T −
level σ+ + δ. Hence, we call
3 := ∂∗E ∩(cid:0)Q × (σ+, σ+ + δ)(cid:1) ,
T −
3 := π(cid:48)(cid:0)T −
3
(cid:1) ,
being π(cid:48) : QN → Q ×(cid:8)xN = σ+ + δ(cid:9) the projection on the last variable. The last contribution
comes from the fact that, since we are slightly moving ∂∗E ∩ QN between the levels σ− and σ+,
we have to take into account that the density is changing. We set then finally
T +
4 := ∂∗E ∩(cid:0)Q × (σ−, σ+)(cid:1) ,
T −
By construction, we can write
∂∗F ⊆(cid:16)
∂∗E \ (T −
3 ∪ T −
4 )
(cid:9) .
T +
4 :=(cid:8)(x(cid:48), xN + δ) : (x(cid:48), xN ) ∈ T −
(cid:17) ∪(cid:16)
1 ∪ T +
T +
2 ∪ T +
3 ∪ T +
(cid:17)
4
4
.
Thus, since the sets T +
can estimate
3 and T −
i are H N−1-essentially pairwise disjoint, and so are also T −
(cid:17)
(cid:17)
1 ) + H N−1
(T +
+
(cid:16)H N−1
(cid:16)H N−1
4 ) − H N−1
(T +
3 ) − H N−1
(T +
(T −
3 )
(T −
4 )
(T +
2 )
.
f
f
f
f
f
f
+
Pf (F ) − Pf (E) ≤ H N−1
4 , one
(3.36)
We now estimate the terms in the right hand side of the above inequality one by one: while the
first two terms (H N−1(T +
i ) for i = 1, 2) are small by the sole essential boundedness of f , to
show that the last two terms (H N−1(T +
i ) for i = 3, 4) are small one needs to use
the essential α-Holder assumption on f .
i )− H N−1(T −
Let us begin by considering T +
1 : by the definition, and also recalling (3.7), it is easy to show
the inclusion
of course to be intended in the H N−1-sense. Therefore, by using (3.6), we directly find
(cid:111)
,
1 ⊆(cid:110)
T +
(x(cid:48), xN ) ∈ ∂Q ×(cid:0) − a/2, a/2(cid:1) : ∃ (x(cid:48), t) ∈ ∂∗E, xN − δ ≤ t ≤ xN
∂∗E ∩(cid:0)∂Q × (−a/2, a/2)(cid:1)(cid:17) ≤ 2N +1M δaN−2 .
1 ) ≤ M δH N−2(cid:16)
(cid:1) ≤ M δH N−2(∂∗Eσ−) ≤ 3M ρaN−2δ .
(cid:0)T +
H N−1
2 , it is sufficient to recall (3.26) in order to obtain
(T +
f
2
H N−1
f
Concerning T +
(3.37)
(3.38)
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
17
We compare now T +
we have by (3.25) and recalling that ¯δ/(4M 2) ≤ δ ≤ ¯δ that
3 and T −
3 . Since the projection π(cid:48) is 1-Lipschitz and f is α-Holder on QN ,
H N−1
f
3 ) − H N−1
(T +
f
(T −
3 ) =
f (x) dH N−1(x) −
(cid:90)
(cid:90)
(cid:90)
(cid:17)
dH N−1(y) ≤ M δαH N−1(cid:0)T −
(cid:1) ≤ 24M 3δα+1ρaN−2 .
≤ M δαH N−1(cid:0)∂∗E ∩ S¯i
f (x) dH N−1(x)
f (π(cid:48)(y)) − f (y)
(cid:16)
T −
T −
≤
T +
3
3
3
3
(cid:90)
(cid:90)
T +
4
(cid:16)
(cid:90)
4 as follows
f (x) dH N−1(x)
4 and T −
(cid:17)
T −
4
f (x(cid:48), xN + δ)) − f (x(cid:48), xN )
dH N−1(x)
T −
=
≤ M δαH N−1(T −
4
4 ) ≤ M δαaN−1(1 + 2ρ) .
(cid:1)
(3.39)
(3.40)
Finally, using again the α-Holder property, we can compare T +
H N−1
f
4 ) − H N−1
(T +
f
(T −
4 ) =
f (x) dH N−1(x) −
Plugging (3.37), (3.38), (3.39) and (3.40) into (3.36), and recalling that ρ, a and δ are as small
as we desire, we conclude
Pf (F ) − Pf (E) ≤ 2M δαaN−1 .
(3.41)
Step (v). Conclusion of the ε − εα property.
We can now conclude very quickly the ε − εα property. Indeed, take a very small ε > 0, and let
¯δ = 2M ε/aN−1. Then, observe that the volume of the set F = F (δ) is a continuous function
of δ ∈(cid:0)¯δ/(4M 2), ¯δ(cid:1): thus, thanks to (3.34) and (3.35), there exists some admissible δ for which
Ff − Ef = ε. In particular, δ satisfies
ε
2M
≤ δaN−1 ≤ 2M ε .
(3.42)
Therefore, (3.41) immediately implies Pf (F ) − Pf (E) ≤ (2M )1+αa(N−1)(1−α)εα.
To finish the proof of the ε − εα property, we have to consider the case when ε < 0 and
ε (cid:28) 1. To do so, still assuming for simplicity that the origin of RN belongs to U ∩ ∂∗E and
the outer normal of E at the origin is the vertical direction, we define E(cid:48) = B(1) \ E. Of course
E(cid:48) is a set of finite perimeter, and ∂∗E = ∂∗E(cid:48) inside the unit ball. We can then apply all the
preceding construction to the set E(cid:48), finding a new set F (cid:48) such that
being
F (cid:48) \ QN = E(cid:48) \ QN ,
(cid:12)(cid:12)F (cid:48)(cid:12)(cid:12)f =(cid:12)(cid:12)E(cid:48)(cid:12)(cid:12)f + ε ,
Defining then F =(cid:0)E \ QN(cid:1) ∪(cid:0)QN \ F (cid:48)(cid:1), we clearly have
(cid:12)(cid:12)F(cid:12)(cid:12)f =(cid:12)(cid:12)E(cid:12)(cid:12)f + ε ,
F \ QN = E \ QN ,
C = (2M )1+αa(N−1)(1−α) .
Pf (F (cid:48)) ≤ Pf (E(cid:48)) + Cεα ,
Pf (F ) ≤ Pf (E) + Cεα ,
so the ε − εα property is finally established.
Part II. The ε − εβ property.
This second part of the proof is devoted to show the ε − εβ property for E, where β = β(α, N )
18
E. CINTI AND A. PRATELLI
is defined in (1.2). Notice that we can assume 0 ≤ α < 1, since otherwise β = α and then the
property has been already shown in Part I.
Our idea is to follow exactly the construction of Part I except for a single, yet fundamental,
detail. To explain this, recall that in Part I we have selected a cube QN , of small but fixed side a;
then, for any small constant ε > 0, we have defined F by moving up the set E in the whole cube
QN = Q× (−a/2, a/2) of a distance δ ≈ ε -- in the sense of (3.42). What we will do now, instead,
will be the following: for every small constant ε > 0, we will find a smaller (N − 1)-dimensional
horizontal cube Qε ⊆ Q of side aεγ, being γ a suitable constant to be specified later. Then, we
will define F by moving up the set E only inside Qε × (−a/2, a/2), and of a bigger distance
δ ≈ ε1−(N−1)γ. Of course, this can make sense for arbitrarily small ε only if 0 < γ < 1
N−1 .
Once had this idea, the proof is only a quite simple modification of the argument of Part I;
basically, one only has to select carefully the small (N − 1)-dimensional cube Qε, write down
the new form of all the estimates already found in Part I, and then select the right constant γ.
We will again split the proof in some steps. First of all, we fix the constant a > 0 and the cube
QN = (−a/2, a/2)N exactly as in Step (i) of Part I, and we also let 0 < γ < 1
N−1 be a constant,
which will be explicitely chosen later.
Step (i). Selection of H "candidate cubes" satisfying (3.45).
Let 0 < ε (cid:28) 1 be a suitably small positive number. We aim to select a (N − 1)-dimensional
cube Qε ⊆ Q of side aεγ; to do so, we will proceed in two different steps. In this first one, we
select a high number of cubes satisfying the new version of the boundary estimate (3.6), namely,
the estimate (3.45) below; then, in next step, we will choose one of those cubes, which will fulfill
also all the other conditions that we need.
We start setting
(cid:35)
(cid:34)
ε(1−N )γ
2N +1
,
(cid:9)
H :=
j=1, ..., 2H contained in Q and having side 2aεγ; this is of
course possible by definition of H as soon as ε is small enough. Let us now concentrate on a
the cube centered at ¯x(cid:48) and
and selecting 2H disjoint cubes (cid:8)(cid:101)Qj
single cube (cid:101)Qj, which is centered at some ¯x(cid:48) ∈ Q, and call (cid:101)Q1/2
having side aεγ. For every x(cid:48) ∈ (cid:101)Q1/2
(cid:19)
, we call
(cid:18)
j
j
N−1(cid:89)
2
i=1
ensures that
the cube of side aεγ centered at x(cid:48), which is of course contained in (cid:101)Qj. A simple rough estimate
H N−1(cid:16)
∂∗E ∩(cid:0)(cid:101)Qj × (−a/2, a/2)(cid:1)(cid:17)
(cid:90)
H N−2(cid:16)
(cid:101)Q1/2
H N−2(cid:16)
∂∗E ∩(cid:0)∂Q(x(cid:48)) × (−a/2, a/2)(cid:1)(cid:17)
∂∗E ∩(cid:0)∂Q(x(cid:48)) × (−a/2, a/2)(cid:1)(cid:17)
2(N − 1)(aεγ)N−2
dH N−1(x(cid:48)) .
dH N−1(x(cid:48))
aεγ
≥
=
1
j
Q(x(cid:48)) :=
i − aεγ
x(cid:48)
, x(cid:48)
i +
aεγ
2
(cid:90)
(cid:101)Q1/2
--
j
2(N − 1)
j ∈ (cid:101)Q1/2
j
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
aεγ
such that
H N−1(cid:16)
As a consequence, there exists some x(cid:48)
H N−2(cid:16)
∂∗E∩(cid:0)∂Q(x(cid:48)
Observe now that, since the cubes (cid:101)Qj are disjoint and contained in Q, it is
H N−1(cid:16)
2H(cid:88)
∂∗E ∩(cid:0)(cid:101)Qj × (−a/2, a/2)(cid:1)(cid:17) ≤
∂∗E∩(cid:0)(cid:101)Qj ×(−a/2, a/2)(cid:1)(cid:17)
j)×(−a/2, a/2)(cid:1)(cid:17) ≤ 2(N − 1)
∂∗E ∩(cid:0)(cid:101)Qj × (−a/2, a/2)(cid:1)(cid:17) ≤ H N−1(cid:0)∂∗E ∩ QN(cid:1) ≤(cid:0)1 + 2ρ(cid:1)aN−1 .
≤ 2N +1(cid:0)1 + 3ρ(cid:1)(cid:0)aεγ(cid:1)N−1 .
Among the 2H cubes (cid:101)Qj, there are then at least H cubes which satisfy
(cid:0)1 + 2ρ(cid:1)aN−1
H N−1(cid:16)
j=1
H
19
. (3.43)
(3.44)
joint (N −1)-dimensional cubes(cid:8)Qj
Up to renumbering, we can assume that those "good" cubes correspond to the indices j =
1, 2, . . . , H. Hence, for any such j we define Qj := Q(x(cid:48)
j). Summarizing, we have found H dis-
j=1, ..., H of side aεγ contained inside Q, and inserting (3.44)
(cid:9)
in (3.43) we find that each of these cubes satisfies the estimate
H N−2(cid:16)
∂∗E ∩(cid:0)∂Qj × (−a/2, a/2)(cid:1)(cid:17) ≤ 2N +2N(cid:0)aεγ(cid:1)N−2 ,
(3.45)
which can be seen as the new version of (3.6). Exactly as in (3.7), Vol'pert Theorem 1.11 allows
us to assume, without loss of generality, that for every 1 ≤ j ≤ H it is
∂∗E ∩(cid:0)∂Qj × (−a/2, a/2)(cid:1) = ∂∗(cid:16)
E ∩(cid:0)∂Qj × (−a/2, a/2)(cid:1)(cid:17)
H N−2 − a.e. .
(3.46)
Step (ii). Choice of the cube Qε.
In this step, we will select one of the H cubes found in Step (i), and we will call it Qε. We will
denote by Aε, Bε, Gε and Γε the intersections of A, B, G and Γ with Qε, where we consider the
decomposition Q = A ∪ B ∪ G ∪ Γ already presented in Step (ii) of Part I, and we will write for
ε = Qε × (−a/2, a/2). We aim to choose Qε in such a way that the following holds:
brevity QN
≤ 1 + 2N +4 · 8ρ ,
ε
(cid:1)
H N−1(cid:0)∂∗E ∩ QN
(cid:0)aεγ(cid:1)N−1
ε \ {−aρ < xN < aρ}(cid:1)
H N−1(cid:0)∂∗E ∩ QN
(cid:0)aεγ(cid:1)N−1
ε ∩ {xN > 0}(cid:1)
H N(cid:0)E ∩ QN
H N−1(cid:0)Gε
(cid:1)
(cid:0)aεγ(cid:1)N−1 ≥ 1 − 2N +4 · 6ρ .
aN εγ(N−1)/2
≤ 2N +4ρ ,
≤ 2N +4ρ ,
(3.47)
(3.48)
(3.49)
(3.50)
Let us show that this is possible. First of all, writing for brevity QN
every 1 ≤ j ≤ H, we can apply (3.2) to find
aN−1ρ ≥ H N−1(cid:0)∂∗E ∩ QN \ {−aρ < xN < aρ}(cid:1) ≥ H(cid:88)
H N−1(cid:0)∂∗E ∩ QN
j = Qj × (−a/2, a/2) for
j \ {−aρ < xN < aρ}(cid:1) .
j=1
20
E. CINTI AND A. PRATELLI
As a consequence, strictly more than 75% of the H cubes satisfy
aN−1ρ
j \ {−aρ < xN < aρ}(cid:17) ≤ 4
H N−1(cid:16)
∂∗E ∩ QN
≤ 2N +4ρ(cid:0)aεγ(cid:1)N−1 ,
H
that is, (3.48). In the very same way, applying (3.4) we observe that more than 75% of the cubes
satisfy (3.49), and applying (3.22) we observe than more than 75% of the cubes satisfy (3.50).
In fact, let π : QN → Q be the
Some additional care is required to obtain also (3.47).
projection on the horizontal variables, and define the measure µ ∈ M(Q) as
(cid:0)∂∗E ∩ QN(cid:1)(cid:17) − H N−1 G .
(cid:16)H N−1
∂∗E ∩(cid:8)(x(cid:48), xN ) ∈ QN : x(cid:48) ∈ V(cid:9)(cid:17) − H N−1(cid:0)V ∩ G(cid:1) .
µ := π#
µ(V ) := H N−1(cid:16)
In other words, for every (N − 1)-dimensional Borel set V ⊆ Q, we set
By construction and by definition of G, one clearly has that µ is a positive measure; hence,
by (3.1), (3.2) and (3.22) we deduce
(cid:107)µ(cid:107) = µ(Q) = H N−1(cid:0)∂∗E ∩ QN(cid:1) − H N−1(G) ≤ 8ρaN−1 .
Thus, arguing as before, we find that more than 75% of the cubes satisfy
µ(cid:0)Qj
(cid:1) ≤ 4
(cid:1) = H N−1(cid:0)Qj ∩ G(cid:1) + µ(cid:0)Qj
H N−1(cid:0)∂∗E ∩ QN
8ρaN−1
H
j
≤ 2N +4 · 8ρ(cid:0)aεγ(cid:1)N−1 .
(cid:1) ≤(cid:0)1 + 2N +4 · 8ρ(cid:1)(cid:0)aεγ(cid:1)N−1 ,
For each of those cubes, it is clearly
thus (3.47) holds.
As a consequence, there must be at least one of the cubes Qj which satisfies contemporar-
ily (3.47), (3.48), (3.49) and (3.50), hence we conclude this step by calling Qε one of those
"good" cubes. Recall that, since Qε is one of the H cubes found in Step (i), then also (3.45)
holds true.
Step (iii). Definition of F and evaluation of its volume and perimeter.
We can now easily give the definition of σ+, σ− and F , and evaluate the volume and perimeter
of F , performing the very same arguments done in Steps (iii) and (iv) of Part I. In fact, we will
see that (3.45) -- (3.50) are the analogous of everything that we really needed there.
We start again by fixing some ¯δ (cid:28) a: then, exactly as we proved (3.25), we can use (3.48)
and (3.49) to find aρ < σ+ < a/2 − 2ρ such that the horizontal strip S¯i = Qε × (σ+, σ+ + ¯δ)
satisfies
aH N−1(∂∗E ∩ Si) + H N (E ∩ Si) ≤ 2N +6¯δρ(aεγ)N−1 .
(3.51)
Moreover, exactly as we used (3.2) to prove (3.26), we can use (3.48) to get the existence of
some −a/2 < σ− < −aρ such that
H N−2(cid:0)∂∗(E ∩ QN
ε )σ−(cid:1) ≤ 3 · 2N +4ρ
(cid:0)aεγ(cid:1)N−1
a
≤ 3 · 2N +4ρ(cid:0)aεγ(cid:1)N−2 .
(3.52)
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
21
We can now generalize also the definition of the competitor set in the obvious way. More
precisely, for every ¯δ/(4M 2) ≤ δ ≤ ¯δ we set F = F (δ) as
x ∈ F ⇐⇒
ε ∩(cid:0){xN ≤ σ−} ∪ {xN > σ+ + δ(cid:9)) ,
ε and σ− < xN ≤ σ− + δ ,
x ∈ E \ QN
ε ,
x ∈ E ∩ QN
(x(cid:48), σ−) ∈ E ∩ QN
(x(cid:48), xN − δ) ∈ E ∩ QN
ε and σ− + δ < xN ≤ σ+ + δ .
2 f .
2 : recalling that f ≤ M in QN , and using (3.51),
1 f −E−
Let us evaluate now the volume and perimeter of F . Concerning the volume, similarly as in
Part I we define
2 := E \ F ∩(cid:0)Qε × (σ+, σ+ + δ)(cid:1) ,
Now, since Aε, Bε, Gε and Γ3,ε are contained by definition in A, B, G and Γ3, by (3.31)
and (3.33) we immediately deduce that
1 ∪ E−
2
(cid:12)(cid:12)E−
2
E+f ≥ 1
M
δH N−1(Gε) ≥ 1
M
E+ := F \ E , E−
so that F = E ∪ E+ \(cid:0)E−
We start with the estimate of the volume of E−
we get
1 := E \ F ∩(cid:0)Qε × (σ−, σ+)(cid:1) , E−
(cid:12)(cid:12)f ≤(cid:12)(cid:12)E ∩ S¯i
To estimate from below the volume E+, it is enough to recall that E+ ⊇ ∪x(cid:48)∈GεE+
x(cid:48) is a segment of length δ for every x(cid:48) ∈ Gε. Thus, by (3.50) we get
E+
(cid:1) and (3.27) holds, namely, Ff −Ef = E+f −E−
(cid:12)(cid:12)f ≤ 2N +6M ¯δρ(aεγ)N−1 ≤ 2N +8M 3δρ(aεγ)N−1 .
δ(cid:0)aεγ(cid:1)N−1(cid:0)1 − 2N +4 · 6ρ(cid:1) ,
and conversely (cid:12)(cid:12)E+ ∩ (Gε × (−a/2, a/2))(cid:12)(cid:12)f ≤ M δH N−1(Gε) ≤ M δ(cid:0)aεγ(cid:1)N−1 .
1 ∩(cid:0)Gε × (−a/2, a/2)(cid:1)(cid:12)(cid:12)(cid:12)f
(cid:17)(cid:12)(cid:12)(cid:12)f
(cid:12)(cid:12)(cid:12)E−
(cid:12)(cid:12)(cid:12)(cid:16)
(cid:17)
(cid:1) × (−a/2, a/2)
(cid:17) − H N−1(cid:16)
∂∗E ∩(cid:0)Gε × (−a/2, a/2)(cid:1)(cid:17)
(cid:17)(cid:12)(cid:12)(cid:12)f
≤ 2N +4 · 14M ρ(cid:0)aεγ(cid:1)N−1δ .
(cid:1) × (−a/2, a/2)
∂∗E ∩(cid:0)Aε ∪ Bε ∪ Γε
= H N−1(cid:16)
≤ 2N +4 · 14ρ(cid:0)aεγ(cid:1)N−1 ,
1 ∪ E+(cid:17) ∩(cid:16)
H N−1(cid:16)
which is the perfect analogous of (3.24). Thus, exactly as in (3.32), we obtain
Finally observe that, thanks to (3.47) and (3.50) and by the definition of G, we have
(Aε ∪ Bε ∪ Γ3,ε) × (−a/2, a/2)
∂∗E ∩ QN
ε
E−
=
(cid:12)(cid:12)(cid:12)(cid:0)E−
1 ∪ E+(cid:1) ∩(cid:16)(cid:0)Γ0,ε ∪ Γ1,ε ∪ Γ2,ε
Ff − Ef ≥ δ(cid:0)aεγ(cid:1)N−1
(cid:18) 1
(3.58)
We can finally write down the estimates for Ff − Ef . Indeed, on one hand, by (3.54), (3.58),
(3.56) and (3.53), and up to take ρ sufficiently small, we get
(cid:19)
(cid:0)1 − 2N +4 · 6ρ(cid:1) − 2N +4 · 14M ρ − 2N +8M 3ρ
≥ δ(cid:0)aεγ(cid:1)N−1
2M
.
(3.53)
x(cid:48) , and that
(3.54)
(3.55)
= 0 . (3.56)
(3.57)
On the other hand, putting together (3.55), (3.56), and (3.58), we also find
M
Ff − Ef ≤ E+f ≤ M δ(cid:0)aεγ(cid:1)N−1(cid:16)
(cid:17) ≤ 2M δ(cid:0)aεγ(cid:1)N−1 .
1 + 2N +4 · 14ρ
22
E. CINTI AND A. PRATELLI
2M ε1−(N−1)γ
aN−1
,
The same argument used in Step (v) of Part I ensures then that, if we define ¯δ =
then there exists an admissible δ such that
Ff − Ef = ε ,
ε1−(N−1)γ
2M aN−1 ≤ δ ≤ 2M ε1−(N−1)γ
aN−1
.
(3.59)
T +
T +
T −
T −
Let us now pass to study the perimeter of F . Exactly as in Step (iv) of Step I, we define
1 :=(cid:0)∂∗F \ ∂∗E(cid:1) ∩(cid:0)∂Qε × (−a/2, a/2)(cid:1) ,
3 := ∂∗E ∩(cid:0)Qε × (σ+, σ+ + δ)(cid:1) ,
4 := ∂∗E ∩(cid:0)Qε × (σ−, σ+)(cid:1) ,
2 := ∂∗(cid:0)E ∩ QN
(cid:1)σ− × (σ−, σ− + δ) ,
3 := π(cid:48)(cid:0)T −
(cid:1) ,
(cid:9) .
4 :=(cid:8)(x(cid:48), xN + δ) : (x(cid:48), xN ) ∈ T −
ε → Qε ×(cid:8)xN = σ+ + δ(cid:9) is the projection on the last variable. Then, we clearly
1 ) ≤ M δH N−2(cid:16)
where π(cid:48) : QN
still have the validity of (3.36), so we need to estimate the H N−1
The same argument which proved (3.37), keeping in mind (3.46) and using (3.45) in place
measures of the sets T ±
i .
of (3.6), gives now
H N−1
f
(3.60)
(T +
T +
T +
4
3
f
ε
Concerning T +
2 , (3.52) immediately gives
H N−1
(cid:1) ≤ M δH N−2(cid:0)∂∗(cid:0)E ∩ QN
(cid:0)T +
f
2
Exactly as in (3.39), a comparison between the H N−1(T +
from (3.51) and using the α-Holder property of f , since
H N−1
f
3 ) − H N−1
(T +
f
(T −
3 ) ≤
ε
∂∗E ∩(cid:0)∂Qε × (−a/2, a/2)(cid:1)(cid:17) ≤ 2N +2N M(cid:0)aεγ(cid:1)N−2δ .
(cid:1)σ−(cid:1) ≤ 3 · 2N +4M ρ(cid:0)aεγ(cid:1)N−2δ .
(cid:90)
(cid:17)
dH N−1(y) ≤ M δαH N−1(cid:0)T −
(cid:1) ≤ 2N +8M 3δα+1ρ(aεγ)N−1
≤ M δαH N−1(cid:0)∂∗E ∩ S¯i
(cid:90)
3 ) and H N−1(T −
f (π(cid:48)(y)) − f (y)
(cid:16)
T −
a
3
.
3
f (x(cid:48), xN + δ)) − f (x(cid:48), xN )
dH N−1(x)
4 ) ≤ M δα(cid:0)aεγ(cid:1)N−1(cid:0)1 + 2N +4 · 8ρ(cid:1) .
≤ M δαH N−1(T −
(cid:16)
T −
4
(cid:17)
(cid:1)
H N−1
f
4 ) − H N−1
(T +
f
(T −
4 ) =
Finally, using once again the α-Holder property of f as in (3.40), and recalling (3.47), we get
(3.61)
3 ) readily comes
(3.62)
(3.63)
Step (iv). Choice of γ and conclusion.
We are finally ready to conclude our proof.
In the preceding steps, we have shown that for
every 0 < ε (cid:28) 1 there exists some set F which equals E out of a small cube QN and such that
Ff −Ef = ε. Moreover, putting together (3.60), (3.61), (3.62) and (3.63), and recalling (3.59),
we also have the estimate
δεγ(N−2) + δαεγ(N−1)(cid:17) ≤ C(cid:48)(cid:48)(cid:16)
ε1−γ + εα+γ(N−1)(1−α)(cid:17)
Pf (F ) − Pf (E) ≤ C(cid:48)(cid:16)
,
(3.64)
being
C(cid:48)(cid:48) =
2N +4N M 2
a
+ (2M )1−αa(N−1)(1−α) .
Recall now that 0 < γ < 1
N−1 is a fixed constant, still to be chosen. It is then finally clear
what is the best choice for γ: indeed, notice that γ1 = 1 − γ is a decreasing function of γ, while
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
23
γ2 = α +γ(N −1)(1−α) is an increasing one, and notice also that, since 0 < α < 1, then γ1 > γ2
(resp., γ1 < γ2) for γ ≈ 0 (resp., γ ≈ 1
N−1 ). Therefore, the optimal choice of γ corresponds to
the situation when γ1 = γ2, which means that we can decide
γ :=
1 − α
α + N (1 − α)
.
Summarizing, we have been able to build a set F with Ff − Ef = ε and
with
Pf (F ) − Pf (E) ≤ 2C(cid:48)(cid:48)εβ
α + (N − 1)(1 − α)
α + N (1 − α)
,
β =
(3.65)
(3.66)
which corresponds to (1.2). To conclude the ε − εβ property, we only have to deal with the case
when ε < 0 and ε (cid:28) 1; however, the case of negative ε can be derived from the case of positive
ε exactly as we did in Step (v) of Part I. Hence, also this second part is concluded.
Part III. The case when f is continuous.
Let us conclude our proof by considering the case when f is only essentially bounded and
continuous. By the result of Part II, the essential boundedness of f , thus the essential α-Holder
N−1
property with α = 0, already tells us that the ε − ε
N property holds true with some constant
C(cid:48)(cid:48), since β = N−1
N when α = 0 by (3.66). What we want to do, is to show the validity of the
same property with any arbitrarily small constant.
To do so, recall that the estimate of Pf (F ) − Pf (E) comes from four terms, see (3.60) --
(3.63) above. Our strategy will be to slightly modify the definition of ¯δ in order to decrease as
desired the first two terms; unfortunately, while doing so the last two terms will correspondingly
increase. However, using the fact that f is continuous, instead of only essentially bounded, we
will be able to let also the last two terms become arbitrarily small.
Let us be more precise: we fix a small number c > 0 and we aim to show the ε − ε
N−1
N
property with constant C = c. To do so, we recall the construction and the estimates of Step II
with γ = 1/N , which corresponds to the case α = 0. The only difference now, is that we fix a
large constant L, to be specified later, and we want to build a set F such that
To do so, the only required change is to define
Ff − Ef = ε :=
ε
L
.
¯δ :=
2M ε1−(N−1)γ
LaN−1
=
1
N
2M ε
LaN−1 ,
which of course reduces to the old choice of ¯δ if L = 1. Then, we find again some δ satisfying
Ff − Ef = ε ,
1
ε
N
N−1
N aN−1
2M L
1
N
≤ δ ≤ 2M ε
N−1
N aN−1
L
.
(3.67)
24
E. CINTI AND A. PRATELLI
which still reduces to (3.59) if L = 1. By (3.60) and (3.61), recalling (3.67) and up to take ρ
small enough, we know that
1 ) + H N−1
(T +
2 ) ≤ 2N +3N M(cid:0)aεγ(cid:1)N−2δ = 2N +3N M aN−2 ε
H N−1
N−2
N L
N−2
N δ
(T +
f
f
≤ 2N +4N M 2
1
N
aL
ε
N−1
N ≤ c
3
N−1
N ,
ε
(3.68)
where the last inequality holds true up to choose a sufficiently large constant L.
We have now to evaluate H N−1
4 ). If we
just insert the new choice of δ into (3.62) and (3.63), these two estimates become worse because
of the presence of the big constant L: in fact, it is now time to use the continuity assumption
on f . Let us then call ω the standard continuity modulus of f on QN , i.e.,
f
f
f
f
(T −
ω(c) = sup
since f is continuous, thus uniformly continuous on QN , we have that ω (cid:38) 0 when c (cid:38) 0. We
can then easily modify the calculation of (3.62) as
H N−1
f
3 ) − H N−1
(T +
f
(T −
3 ) ≤
3 ) − H N−1
(T +
3 ) and H N−1
(T −
4 ) − H N−1
(T +
(cid:111)
(cid:110)(cid:12)(cid:12)f (x) − f (y)(cid:12)(cid:12) : x, y ∈ QN , x − y ≤ c
(cid:90)
(cid:17)
(cid:16)
dH N−1(y) ≤ ω(δ)H N−1(cid:0)T −
≤ ω(δ)H N−1(cid:0)∂∗E ∩ S¯i
(cid:1) ≤ 2N +7M ρ
(cid:90)
(cid:17)
f (π(cid:48)(y)) − f (y)
εω(δ) ≤ c
3
N−1
N ,
(cid:16)
f (x(cid:48), xN + δ)) − f (x(cid:48), xN )
T −
ε
a
3
;
3
dH N−1(x)
4 ) ≤ ω(δ)(cid:0)aεγ(cid:1)N−1(cid:0)1 + 2N +4 · 8ρ(cid:1)
(cid:18) 2M ε
N−1
N aN−1 ε
N−1
N ω
1
N
≤ 2L
N−1
N aN−1 ε
N−1
N ω(δ) ≤ 2L
≤ ω(δ)H N−1(T −
(cid:1)
N−1
N aN−1
L
(cid:18) 2M ε
1
N
ω
N−1
N aN−1
L
(cid:19)
≤
c
N−1
N aN−1
6L
(3.69)
(3.70)
(cid:19)
,
where the last inequality is true if ε is small enough. Finally, the estimate (3.63) now becomes
H N−1
f
4 ) − H N−1
(T +
f
(T −
4 ) =
T −
4
using also (3.67). We can select a sufficiently small ¯ε such that, whenever 0 < ε < ¯ε, one has
(recall that L has been already fixed). As a consequence, (3.70) yields
(3.71)
Summarizing, for any c > 0 we have found 0 < ¯ε (cid:28) 1 such that, for any 0 < ε < ¯ε, there exists
F which, by (3.68), (3.69) and (3.71), satisfies
ε
f
f
H N−1
4 ) − H N−1
(T +
(T −
4 ) ≤ c
3
N−1
N .
Ff − Ef = ε ,
Pf (F ) − Pf (E) ≤ cε
N−1
N .
N−1
N property with any constant c > 0.
Arguing in the usual way to treat the case of ε < 0, ε < ¯ε, we have then concluded the validity
of the ε − ε
(cid:3)
Remark 3.1. We underline that, in Theorem B, the validity of the ε − εβ property with any
small constant is a peculiarity of the case when f is continuous, but it cannot be inferred for the
general case of a α-Holder function f .
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
25
4. An example of an unbounded isoperimetric set
This section is devoted to show an example of an essentially bounded but discontinuous
function f which admits an unbounded isoperimetric set. This will show the sharpness of the
assumption in Theorem B.
i∈N be a sequence of disjoint balls of Euclidean volume(cid:12)(cid:12)Bi
(cid:9)
(cid:12)(cid:12)eucl = 1/2i, sufficiently
Let(cid:8)Bi
far from each other, and let f : RN → R be defined as
(cid:40)
1
M
if x ∈(cid:83)
otherwise ,
i ∂Bi ,
f (x) :=
where M > 1 is a constant, big enough, to be precised later. We will show that B = ∪iBi,
which is an unbounded set, is the unique isoperimetric set of volume M . To do so, let us pick a
smooth set E ⊆ RN with Ef = M ; we aim to show that, for a suitable constant ξ > 0, one has
Pf (E) ≥ Pf (B) + ξη
N−1
N ,
(4.1)
being B(cid:52)E = (B \ E) ∪ (E \ B) the symmetric difference between E and B. By the density of
smooth sets among sets of finite perimeter, (4.1) will show that B is the unique isoperimetric
set of volume M .
where
η :=
2
,
(cid:12)(cid:12)B(cid:52)E(cid:12)(cid:12)eucl
Claim 1. It is admissible to assume that every connected component of E intersects exactly one
of the balls Bi.
Proof. Assume first that a connected component of E intersects two different balls. We argue
which is far enough from each of the balls, and replacing it with a ball of the same volume (not
intersecting B ∪ E); the very same calculation done in the proof of Theorem A ensures that, if
exactly as in the proof of Theorem A: one can build a competitor (cid:101)E by cutting the part of E
the balls are chosen sufficiently far from each other, there will be such a set (cid:101)E having perimeter
smaller than that of E. Since by construction(cid:12)(cid:12)B(cid:52)E(cid:12)(cid:12)f =(cid:12)(cid:12)(cid:101)B(cid:52)E(cid:12)(cid:12)f , we can reduce ourselves to
show (4.1) for the set (cid:101)E; this shows that it is admissible to assume that no connected component
of E intersects two different balls.
Suppose now, instead, that there is a connected component of E which does not intersect
any ball; then, we can simply translate this component around RN until it touches one of the
balls Bi, or one of the other connected components which in turn touches some ball. Since the
density is constant in RN \ B, this translation does not effect the perimeter nor the volume of
E, hence it is admissible to assume that every connected component of E intersects at least one
(cid:3)
ball. The proof of the claim is then concluded.
Thanks to the above claim, for every i we can consider the (possibly empty) connected
component of E which intersects Bi, and subdivide it into two parts, the set Ei ⊆ Bi and the
remaining set Fi, as depicted in Figure 1. We call now
εi :=(cid:12)(cid:12)Bi \ Ei
(cid:12)(cid:12)eucl ,
δi :=(cid:12)(cid:12)Fi
(cid:12)(cid:12)eucl ,
26
and notice that
E. CINTI AND A. PRATELLI
(cid:88)
(cid:88)
(4.2)
by the definition of η in (4.1) and since Ef = Bf = M . Our next observation is the following.
δi = η
εi =
i
i
F1
E1
F2
E2
B2
E3
B3
B1
Figure 1. Sketch of the situation in the Example of Section 4; the set F3 is empty.
Claim 2. For every i ∈ N one has
Pf (Bi ∪ Fi) − Pf (Bi) ≥ M − 1
(cid:12)(cid:12) N−1
(cid:12)(cid:12)Fi
(cid:1) = Peucl(Fi) ≥ N ω1/N
2
H N−1(cid:0)∂Fi
N ω1/N
N δ
N−1
N
i
.
Proof. First of all, we know by the Euclidean isoperimetric inequality that
N−1
N
(4.3)
it is done by two parts, namely, Γ1 = ∂Fi \ Bi and Γ2 =
Consider now the boundary of Fi:
∂Fi ∩ ∂Bi. Call now π : RN \ Bi → ∂Bi the projection on Bi: since the ball in convex, π is
1-Lipschitz; therefore,
N
.
i
N
eucl = N ω1/N
N δ
As a consequence,
Pf (Bi ∪ Fi) − Pf (Bi) = H N−1
H N−1(Γ1) ≥ H N−1(cid:0)π(Γ1)(cid:1) ≥ H N−1(Γ2) .
(cid:1) = M H N−1(cid:0)Γ1
(cid:1) − H N−1
≥ (M − 1)H N−1(cid:0)Γ1
H N−1(cid:0)∂Fi
≥ M − 1
(cid:0)Γ2
(cid:1) ≥ M − 1
(cid:0)Γ1
N−1
N
N ω1/N
N δ
i
2
,
f
f
2
(cid:1)
(cid:1) − H N−1(cid:0)Γ2
(cid:1)
recalling that ∂Fi = Γ1 ∪ Γ2 and (4.3). The proof of the claim is then concluded.
As an immediate corollary, just adding over i ∈ N, using the concavity of t (cid:55)→ t
recalling (4.2), we get
(cid:88)
i∈N
Pf (Bi ∪ Fi) − Pf (B) =
Pf (Bi ∪ Fi) − Pf (Bi)
(cid:16)
(cid:88)
i∈N
≥ M − 1
2
(cid:17) ≥ M − 1
(cid:17) N−1
2
M − 1
N =
2
(cid:88)
i∈N
N ω1/N
N
N−1
N
δ
i
N ω1/N
N η
N−1
N .
(cid:16)(cid:88)
N ω1/N
N
i∈N δi
Now, since a quick observation tells us that
Pf (Ei ∪ Fi) − Pf (Bi ∪ Fi) ≥ Pf (Ei) − Pf (Bi) ,
(cid:3)
N−1
N , and
(4.4)
(4.5)
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
27
and thanks to (4.4), we are basically reduced to evaluate Pf (Ei) − Pf (Bi) in terms of εi. We
can start with an easy bound.
Claim 3. For every i ∈ N, one has
Pf (Ei) − Pf (Bi) ≥ −N ω1/N
N ε
N−1
N
i
.
Proof. The Euclidean isoperimetric inequality tells us that
Peucl(Ei) ≥ N ω1/N
N Ei N−1
N
eucl = N ω1/N
N−1
N , we have
N
therefore, again by the concavity of t (cid:55)→ t
Pf (Ei) − Pf (Bi) ≥ Peucl(Ei) − Peucl(Bi) ≥ N ω1/N
N
(cid:17) N−1
(cid:16)Bieucl − εi
(cid:18)(cid:16)Bieucl − εi
(cid:17) N−1
N − Bi N−1
N ;
N
eucl
≥ −N ω1/N
N ε
N−1
N
i
.
(cid:19)
(cid:3)
Unfortunately, we cannot simply conclude comparing (4.4) and last claim, because the
N now works against our estimate, since(cid:80)−ε
concavity of t (cid:55)→ t
this problem, we subdivide N into two parts, namely,
N−1
N−1
N
i
≤ −η
N−1
N . To overcome
(cid:26)
I :=
i ∈ N :
(cid:27)
(cid:12)(cid:12)Bi
(cid:12)(cid:12)eucl
εi
≤ 1
4
,
J :=
(cid:26)
i ∈ N :
(cid:12)(cid:12)Bi
(cid:12)(cid:12)eucl
εi
>
1
4
(cid:27)
;
in words, the indices belonging to I (resp., J) are those corresponding to sets Ei which contain
more (resp., less) than 75% of the corresponding ball Bi. The key point is that for indices in I
something much stronger than Claim 3 can be found.
Claim 4. For every i ∈ I, one has Pf (Ei) ≥ Pf (Bi).
Proof. First of all, we consider the spherical symmetrization E∗ of Ei, which satisfies Pf (E∗) ≤
Pf (Ei) by Theorem 1.4 (and since f is obviously radial in a neighborhood of Bi); notice that
it is also E∗eucl = Eeucl, since f is constant in Bi. Then, we write ∂E∗ = Γ1 ∪ Γ2, where
Γ1 = ∂E∗ ∩ Bi and Γ2 = ∂E∗ ∩ ∂Bi; in particular, Γ2 is a (possibly empty) spherical cap in Bi.
Let us then call P the (N − 1)-dimensional ball whose boundary coincides with the boundary
In other words, since Γ2 is a spherical cap then up to a rotation
of Γ2 as a subset of ∂Bi.
one has Γ2 = (cid:8)x ∈ ∂Bi : x · ν ≤ κ(cid:9) for suitable κ ∈ R and ν ∈ SN−1, we define the ball
P =(cid:8)x ∈ Bi : x · ν = κ(cid:9); the situation is depicted in Figure 2. We have now to distinguish two
cases.
Case I. H N−1(Γ2) ≥ H N−1(∂Bi)/2.
In this case, as in Claim 2 we call π the projection over P , which is 1-Lipschitz by the convexity
of P ; then, we can again estimate
H N−1(Γ1) ≥ H N−1(cid:0)π(Γ1)(cid:1) ≥ H N−1(P ) .
28
E. CINTI AND A. PRATELLI
Γ2
E∗
Γ1
P
Bi
Figure 2. Situation in Claim 4.
Moreover, since H N−1(Γ2) ≥ H N−1(∂Bi)/2, then
H N−1(cid:0)∂Bi \ Γ2
(cid:1) ≤ N ωN
2ωN−1
H N−1(P ) ≤ N H N−1(P ) .
Putting these two observation together, and assuming without loss of generality that M ≥ N ,
one directly gets
Pf (Ei) ≥ Pf (E∗) = M H N−1(Γ1) + H N−1(Γ2)
≥ M H N−1(P ) + H N−1(∂Bi) − H N−1(∂Bi \ Γ2) ≥ H N−1(∂Bi) = Pf (Bi) ,
hence we have concluded in this case (without even using the assumption that i ∈ I).
Case II. H N−1(Γ2) < H N−1(∂Bi)/2.
In this case, the Euclidean isoperimetric inequality and the fact that i ∈ I tell us that
Peucl(E∗) ≥ N ω1/N
N E∗ N−1
N
eucl = N ω1/N
N
N ≥ 3
4
N ω1/N
N Bi N−1
eucl =
N
3
4
Pf (Bi) .
(cid:0)Bieucl − εi
(cid:1) N−1
On the other hand, the assumption of this case gives
Peucl(E) = H N−1(Γ1) + H N−1(Γ2) ≤ H N−1(Γ1) +
1
2
Pf (Bi) .
The two preceding estimates imply H N−1(Γ1) ≥ 1
4 Pf (Bi), which in turn yields
Pf (E) ≥ M H N−1(Γ1) ≥ Pf (Bi) ,
as soon as M ≥ 4. The proof is then concluded also for this last case.
(cid:3)
The next step is to observe what happens for indices in J.
Claim 5. One has(cid:80)
(cid:16)
(cid:17) ≥ −C(N )η
Pf (Ei)−Pf (Bi)
i∈J
N−1
N , where C(N ) is a purely dimensional
constant.
Proof. The claim is emptily true if J = ∅, then we can directly suppose J (cid:54)= ∅ and call j the
smallest element of J. A simple calculation, recalling that εj > 1/(4 · 2j), yields
(cid:88)
N−1
N
i
ε
≤(cid:88)
i>j
i>j
(cid:0)2
N − 1(cid:1) · 2j N−1
1
N
N−1
1
2i N−1
N
=
≤ C1(N )ε
N−1
N
j
,
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
29
where
N−1
4
N
N−1
N − 1
Therefore, using Claim 3, Claim 4 and (4.2), we deduce
N−1
N
(cid:17) ≥ −N ω1/N
Pf (Ei) − Pf (Bi)
(cid:88)
C1(N ) =
(cid:16)
2
ε
i
N
.
(cid:19)
(cid:18)
N−1
N
j
ε
≥ −N ω1/N
N
i∈J
N−1
N
(cid:18)
(cid:88)
(cid:0)C1(N ) + 1(cid:1)ε
+
ε
j
j
i∈J, i>j
≥ −N ω1/N
= −C(N )η
N
N−1
N .
N−1
N
= −C(N )ε
N−1
N
j
≥ −C(N )
(cid:88)
(cid:16)(cid:88)
i>j
ε
+
(cid:19)
(cid:17) N−1
N
N−1
N
i
i∈N εi
(cid:3)
We are finally in position to conclude. In fact, putting together (4.4), (4.5), Claims 3, 4
and 5, and assuming without loss of generality that E has finite perimeter, we have
Pf (E) − Pf (B) =
Pf (Ei ∪ Fi) − Pf (B)
i∈N
(cid:88)
(cid:88)
≥(cid:88)
≥(cid:88)
i∈N
i∈N
=
i∈J
(cid:16)
(cid:16)
(cid:16)
Pf (Ei ∪ Fi) − Pf (Bi ∪ Fi)
Pf (Ei) − Pf (Bi)
Pf (Ei) − Pf (Bi)
(cid:17)
(cid:17)
(cid:88)
+
+
i∈N
M − 1
2
(cid:17)
+
(cid:88)
i∈N
Pf (Bi ∪ Fi) − Pf (B)
Pf (Bi ∪ Fi) − Pf (B)
N ω1/N
N η
N−1
N ≥ ξη
N−1
N ,
where ξ > 0 if M is big enough. We have then established the validity of (4.1), thus it is
definitively proved that B is the (unique) isoperimetric set of volume M , as desired.
5. Applications on existence and regularity
This final section is devoted to show two applications of Theorems A and B to the questions
of the existence and regularity of isoperimetric sets. Even if those two are only simple conse-
quences of the above theorems and of the known facts about existence and regularity, the results
that we can find are stronger than those which were previously known. More refined new results
concerning the regularity in the 2-dimensional case are contained in the forthcoming paper [11].
5.1. On the existence of isoperimetric sets. Let us start discussing the question of existence
of isoperimetric sets. As we explained in the Introduction, the existence is deeply connected
with the boundedness; in particular, some results in [24] provide existence of isoperimetric sets
(for a certain volume m) under the assumption that all the isoperimetric sets for volumes smaller
than m (if any) are bounded. Of course, in all these results the boundedness assumption can
be removed whenever it comes directly from Theorem A. Let us be more precise: we recall the
following result (which can be found in [24, Theorems 7.9, 7.11, 7.13]).
Theorem 5.1. Let f be a density on RN approaching a finite limit a > 0 at infinity, and assume
that the isoperimetric sets are bounded. Then there exist isoperimetric sets of all volumes if one
of the following properties holds:
30
E. CINTI AND A. PRATELLI
(i) for every V > 0 and for every R > 0, there is some ball B of volume V at distance from
the origin at least R such that
(cid:16)
(cid:17) N−1
N ;
f (x) ≤ a
1
N
sup
x∈B
f (x)
inf
x∈B
(ii) f is radial and, for any c > 0 and any ρ > 0, there exists some R ≥ ρ such that
(iii) for any V > 0, there exist balls B of volume V arbitrarily far from the origin satisfying
the mean inequality
f (R) ≤ a − e−cR ;
(cid:90)
--
∂B
(cid:18)
(cid:90)
f
--
B
(cid:19) N−1
N
.
f ≤ a
1
N
As an immediate application of Theorem A we can then strengthen the above existence
result as follows (notice that there is no need of requiring the essential boundedness to f , by the
assumption that f converges to a > 0 at infinity).
Theorem 5.2. Let f be a continuous density on RN approaching a finite limit a > 0 at infinity.
Then there exist isoperimetric sets of all volumes if one of the properties (i), (ii), or (iii) of
Theorem 5.1 holds true.
5.2. On the regularity of isoperimetric sets. We pass now to the regularity issue. To start,
we recall what is known up to now concerning the regularity of isoperimetric sets in RN with
density (see for instance [21, Proposition 3.5 and Corollary 3.8]).
Theorem 5.3. Let f be a smooth or Ck,α density on Rn, with k ≥ 1. Then the boundary of
any isoperimetric set is a smooth or Ck+1,α submanifold except on a singular set of Hausdorff
dimension at most n − 8.
Notice that, as we already explained in the Introduction, the known result covers only cases
in which the density is at least Lipschitz, while there are no results for lower regularity of the
density. Our theorems, instead, allow to obtain some regularity results for the isoperimetric sets
even with just bounded densities.
What we will do here, in fact, is just to put together our Theorem B and the well-known
regularity theory in the standard Euclidean setting. To do so, let us first briefly recall a couple
of important notions of minimality for the perimeter and the corresponding regularity properties
(whose proofs can be found for instance in [12, 19, 27], see also [18]); then, we will derive the
regularity results for our setting.
Definition 5.4. Let E ⊆ RN be a set of locally finite perimeter. We say that E is quasi-minimal
if, for some C > 0 and for every ball Br(x),
(cid:0)E, Br(x)(cid:1) ≤ CrN−1 .
Peucl
Moreover, we say that E is ω-minimal, for some continuous and increasing function ω : R+ →
R+ with ω(0) = 0, if, for every ball Br(x) and every set F such that F(cid:52)E ⊂⊂ Br(x), one has
(5.1)
(cid:0)F, Br(x)(cid:1) + ω(r) rN−1 .
(cid:0)E, Br(x)(cid:1) ≤ Peucl
Peucl
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
31
Definition 5.5. Let E be a Borel subset of RN . We say that E is porous if there exists a small
constant δ > 0 such that, for every x ∈ ∂E and every arbitrarily small ball Br(x) centered at x,
there are two balls B1, B2 ⊆ Br(x) of radius δr such that B1 ⊆ E and B2 ⊆ RN \ E.
Theorem 5.6. If a set E is quasi-minimal, then it is porous and the reduced and the topological
boundaries coincide (H N−1-a.e.). Moreover, if E is ω-minimal with ω(r) = Crη for some
0 < η ≤ 1, then ∂∗E is C1,η/2.
We will see that an isoperimetric set is always quasi-minimal if the density is even just
bounded from above and below, hence the porosity holds true in all these cases. Instead, the
ω-minimality holds true as soon as f is Holder continuous, thus in these more particular cases
we will get also some further regularity. More precisely, we obtain the following result.
Theorem 5.7. Let E be an isoperimetric set corresponding to the density f . If f is bounded
from above and below, then E is porous and ∂∗E = ∂E. If moreover f is α-Holder, then one
has also that ∂∗E ∈ C1,
2N (1−α)+2α .
α
We point out that the regularity given by the above theorem is surely not optimal. Indeed,
in the forthcoming paper [11] we will improve the above regularity result by showing that, in
dimension N = 2, if f ∈ C0,α then the boundary of any isoperimetric set is of class C1, α
3−2α ,
while Theorem 5.7 gives only C1, α
4−2α .
Before giving the proof of Theorem 5.7, a couple of remarks is in order.
Remark 5.8. We have claimed the regularity theorem under the assumption that the density
is bounded, or Holder, instead of essentially bounded or essentially Holder, just for simplicity
of notations. However, it is very easy to deduce the general claim. In fact, assume that f is
essentially bounded, or essentially α-Holder, and let Uδ be the open sets as in Definition 1.7.
Then, just arguing inside each open set Uδ, in the essentially bounded case we derive that an
δ>0 Uδ.
2N (1−α)+2α .
isoperimetric set E satisfies the porosity property on each Uδ, and that ∂E = ∂∗E on(cid:83)
Similarly, in the essentially α-Holder case, we deduce that ∂∗E∩(cid:83)
δ>0 Uδ is of class C1,
α
Remark 5.9. One could try to obtain some regularity for the isoperimetric sets also starting
from other standard regularity results (whose proofs can be found in [5, 20, 27]). For instance,
a set E is of class C1 if it has the uniform interior and exterior ball property. Recall that E is
said to satisfy the interior (resp., exterior) ball property if there is ¯r > 0 such that, for every
x ∈ ∂∗E, there exists a ball B¯r(y) contained in E (resp., in RN \ E) such that x ∈ ∂B¯r(y). In
addition, E is even of class C1,γ for every 0 < γ < 1 (and even C1,1 if N = 2) if it is Λ-minimal,
that is, there exists Λ ≥ 0 such that for every set F one has Peucl(E) ≤ Peucl(F ) + ΛF(cid:52)Eeucl.
But unfortunately, both these conditions seem to become useful only when f is at least
Lipschitz (while we are interested in the lower regularity case). To be more precise, concerning
the interior-exterior ball property one can easily observe that, if f is not Lipschitz, it is not even
true that an arc of circle of small radius is longer than the corresponding chord. And concerning
the Λ-minimality, one cannot hope to have it if f is α-Holder and α < 1, as one can derive
arguing as in the proof of Theorem 5.7 below, and as one can also guess because otherwise the
32
E. CINTI AND A. PRATELLI
regularity result (f of class C0,α would imply any isoperimetric set of class C1,1−ε) would be
excessively strong.
We are now ready to conclude the paper with the proof of the regularity Theorem 5.7.
Proof of Theorem 5.7. Keeping in mind Theorem 5.6, we are reduced to check that an isoperi-
metric set E is quasi-minimal if f is bounded from above and below, while E is ¯ω-minimal with
¯ω(r) = C r
N (1−α)+α if f is also α-Holder.
α
Let us start by assuming that 1/M < f < M , and suppose by contradiction that an
isoperimetric set E is not quasi-minimal. Hence, for any large constant K there exists a ball
Br(x) such that
Let then Br(x) be any such ball, and define the set F := E \ Br(x): provided that K is very
large, we deduce
Peucl
(cid:0)E, Br(x)(cid:1) > KrN−1 .
(cid:0)E, Br(x)(cid:1) + N ωN rN−1M
Pf (F ) ≤ Pf (E) − Pf
≤ Pf (E) − K
M
rN−1 + N ωN rN−1M ≤ Pf (E) − K
2M
rN−1 .
Thanks to Theorem B, we know that F fulfills the ε − ε
thus there exists a further set (cid:101)E satisfying (cid:101)Ef = Ef and with
Pf ((cid:101)E) ≤ Pf (F ) + C
≤ Pf (E) − K
2M
(cid:17) N−1
(cid:16)(cid:12)(cid:12)(cid:101)E(cid:12)(cid:12)f −(cid:12)(cid:12)F(cid:12)(cid:12)f
M ωN rN(cid:17) N−1
(cid:16)
rN−1 + C
N ≤ Pf (E) − K
2M
N−1
N property with some constant C,
(cid:16)(cid:12)(cid:12)E(cid:12)(cid:12)f −(cid:12)(cid:12)F(cid:12)(cid:12)f
(cid:17) N−1
N
rN−1 + C
N ≤ Pf (E) − K
3M
rN−1 < Pf (E) ,
where we are assuming again K to be large enough. Since this inequality is against the isoperi-
metric property of E, the contradiction shows the quasi-minimality of E.
Let us now assume that f is also α-Holder and E is an isoperimetric set: to conclude the
N (1−α)+α and some suitable C.
proof, we need to show that E is ¯ω-minimal with ¯ω(r) = C r
To do so, we pick 0 < η ≤ 1 and we investigate whether, for some suitable constant C, E is
ω-minimal with ω(r) = Crη (eventually, we will find that the best choice is ω = ¯ω).
α
Therefore we suppose that, for any large constant K, there exist a ball Br(x) and a set F
with F(cid:52)E ⊂⊂ Br(x), such that
Peucl(E, Br(x)) > Peucl(F, Br(x)) + Krη+N−1 .
(5.2)
By the first part of the proof we know that E is quasi-minimal, thus Peucl(E, Br(x)) ≤ C1rN−1,
and by (5.2) it is then also Peucl(F, Br(x)) ≤ C1rN−1. Let us assume, just for simplicity of
notations, that minBr(x) f = 1; hence, since f is α-Holder, it is maxBr(x) f ≤ 1 + M rα, so that
(cid:0)F, Br(x)(cid:1) ≤ (1 + M rα)Peucl
(cid:0)F, Br(x)(cid:1) .
Pf
(cid:0)E, Br(x)(cid:1) ≥ Peucl(E, Br(x)) ,
Pf
Recalling (5.2), we deduce
Pf (F ) − Pf (E) = Pf (F, Br(x)) − Pf (E, Br(x))
≤ (1 + M rα)Peucl(F, Br(x)) − Peucl(E, Br(x)) ≤ C2rα+N−1 − Krη+N−1 .
(5.3)
(cid:0)(cid:101)E(cid:1) ≤ Pf (F ) + C3
Pf
Combining this estimate with (5.3), we get
ε − εβ PROPERTY, BOUNDEDNESS OF ISOPERIMETRIC SETS AND APPLICATIONS
Applying now Theorem B, we can define a competitor set (cid:101)E with
(cid:16)(cid:12)(cid:12)Ff − (cid:101)Ef
(cid:12)(cid:12)(cid:17)β ≤ Pf (F ) + C4rN β .
(cid:12)(cid:12)(cid:101)E(cid:12)(cid:12)f =(cid:12)(cid:12)E(cid:12)(cid:12)f ,
Pf ((cid:101)E) − Pf (E) ≤ Pf (F ) − Pf (E) + C4rN β ≤ C2rα+N−1 − Krη+N−1 + C4rN β .
33
Since by the definition (1.2) of β one immediately checks that α + N − 1 ≥ N β for any α ∈ (0, 1],
we have a contradiction with the optimality of E -- with the choice of a sufficiently large K -- as
soon as η + N − 1 ≤ N β, that is, η ≤
N (1−α)+α . Summarizing, we have shown the ¯ω-minimality
(cid:3)
of E and hence the proof is concluded.
α
Acknowledgments
The authors wish to thank Luigi Ambrosio, Guido De Philippis, Nicola Fusco, Francesco
Maggi, Frank Morgan and Emanuele Spadaro for some fruitful discussions and for their sugges-
tions on a preliminary version of this paper. Both authors have been supported by the ERC
Starting Grant "AnOptSetCon" n. 258685, while AP has been also supported by the ERC
Advanced Grant "AnTeGeFI" n. 226234.
References
[1] W.K. Allard, A regularity theorem for the first variation of the area integrand, Bull. Amer. Math. Soc. 77
(1971), 772 -- 776.
[2] W.K. Allard, On the first variation of a varifold. Ann. of Math. (2) 95 (1972), 417 -- 491.
[3] F.J. Almgren, Jr., Existence and regularity almost everywhere of solutions to elliptic variational problems
with constraints, Mem. Amer. Math. Soc. 4 (1976), no. 165.
[4] F.J. Almgren, Jr., Almgren's big regularity paper. Q-valued functions minimizing Dirichlet's integral and
the regularity of area-minimizing rectifiable currents up to codimension 2. World Scientific Monograph Series
in Mathematics, Vol. 1: xvi+955 pp. (2000).
[5] L. Ambrosio, Corso introduttivo alla teoria geometrica della misura e alle superfici minime, Edizioni della
Normale (1996).
[6] L. Ambrosio, N. Fusco, D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems, Oxford
University Press (2000).
[7] M. Barchiesi, F. Cagnetti, N. Fusco, Stability of the Steiner symmetrization of convex sets, J. Eur. Math.
Soc., to appear.
[8] E. Bombieri, Regularity theory for almost minimal currents. Arch. Rational Mech. Anal. 78 (1982), no. 2,
99 -- 130.
[9] A. Canete, M. Jr. Miranda, D. Vittone, Some isoperimetric problems in planes with density, J. Geom. Anal.
20 (2010), no. 2, 243 -- 290.
[10] A. Cianchi, N. Fusco, F. Maggi, A. Pratelli, On the isoperimetric deficit in Gauss space, Amer. J. Math.
133 (2011), no. 1, 131 -- 186.
[11] E. Cinti, A. Pratelli, Regularity of isoperimetric sets with density in dimension 2, forthcoming.
[12] G. David, S. Semmes, Quasiminimal surfaces of codimension 1 and John domains, Pacific J. Math. 183
(1998), no. 2, 213 -- 277.
[13] A. D´ıaz, N. Harman, S. Howe, D. Thompson, Isoperimetric problems in sectors with density, Adv. Geom.,
to appear (available at http://arxiv.org/abs/1012.0450).
34
E. CINTI AND A. PRATELLI
[14] A. Figalli, F. Maggi, On the isoperimetric problem for radial log-convex densities, Calc. Var. Partial Differ-
ential Equations, to appear.
[15] N. Fusco, The classical isoperimetric Theorem, Rend. Acc. Sc. Fis. Mat. Napoli, 71 (2004), 63 -- 107.
[16] N. Fusco, F. Maggi, A. Pratelli, The sharp quantitative isoperimetric inequality, Ann. of Math. (2) 168
(2008), no. 3, 941 -- 980.
[17] N. Fusco, F. Maggi, A. Pratelli, On the isoperimetric problem with respect to a mixed Euclidean-Gaussian
density, J. Funct. Anal. 260 (2011), no. 12, 3678 -- 3717.
[18] M. Giaquinta, E. Giusti, Quasi-minima, Ann. Inst. H. Poincar´e Anal. Non Lin´eaire 1 (1984), no. 2, 79 -- 107.
[19] J. Kinnunen, R. Korte, A. Lorent, N. Shanmugalingam, Regularity of sets with quasiminimal boundary
surfaces in metric spaces, preprint (2011). Available at http://arxiv.org/abs/1105.3058.
[20] F. Maggi, Sets of Finite Perimeter and Geometric Variational Problems: an Introduction to Geometric
Measure Theory, Cambridge Studies in Advanced Mathematics no. 135, Cambridge University Press (2012).
[21] F. Morgan, Regularity of isoperimetric hypersurfaces in Riemannian manifolds, Trans. Amer. Math. Soc.
355 (2003), n. 12, 5041 -- 5052.
[22] F. Morgan, Manifolds with density, Notices Amer. Math. Soc. 52 (2005), no. 8, 853 -- 858.
[23] F. Morgan, Geometric Measure Theory: a Beginner's Guide, Academic Press, 4th edition, 2009.
[24] F. Morgan, A. Pratelli, Existence of isoperimetric regions in Rn with density, Ann. Global Anal. Geom, to
appear.
[25] C. Rosales, A. Canete, V. Bayle, F. Morgan, On the isoperimetric problem in Euclidean space with density,
Calc. Var. Partial Differential Equations 31 (2008), no. 1, 27 -- 46.
[26] R. Schoen, L. Simon, A new proof of the regularity theorem for rectifiable currents which minimize para-
metric elliptic functionals, Indiana U. Math. J. 31 (1982), 415 -- 434.
[27] I. Tamanini, Regularity results for almost-minimal oriented hypersurfaces in RN , Quaderni del Dipartimento
di Matematica dell'Universit`a del Salento (1984).
[28] V.A.I. Vol'pert, Spaces BV and quasilinear equations, Math. USSR Sb, 17 (1967), 225 -- 267.
|
1112.1632 | 1 | 1112 | 2011-12-07T17:18:32 | On a family of frames for Krein spaces | [
"math.FA"
] | A definition of frames for Krein spaces is proposed, which extends the notion of $J$-orthonormal basis of Krein spaces. A $J$-frame for a Krein space $(\HH, \K{\,}{\,})$ is in particular a frame for $\HH$ in the Hilbert space sense. But it is also compatible with the indefinite inner product $\K{\,}{\,}$, meaning that it determines a pair of maximal uniformly $J$-definite subspaces with different positivity, an analogue to the maximal dual pair associated to a $J$-orthonormal basis.
Also, each $J$-frame induces an indefinite reconstruction formula for the vectors in $\HH$, which resembles the one given by a $J$-orthonormal basis. | math.FA | math |
On a family of frames for Krein spaces
J. I. Giribet, A. Maestripieri, F. Mart´ınez Per´ıa and P. Massey
Abstract
A definition of frames for Krein spaces is proposed, which extends the notion of J-orthonormal
basis of Krein spaces. A J-frame for a Krein space (H, [ , ]) is in particular a frame for H in the
Hilbert space sense. But it is also compatible with the indefinite inner product [ , ], meaning that it
determines a pair of maximal uniformly J-definite subspaces with different positivity, an analogue to
the maximal dual pair associated to a J-orthonormal basis.
Also, each J-frame induces an indefinite reconstruction formula for the vectors in H, which re-
sembles the one given by a J-orthonormal basis.
keywords: Krein spaces,
MSC 2000: 46C20, 47B50, 42C15
frames, uniformly J-definite subspaces
1
Introduction
In recent years, frame theory for Hilbert spaces has been thoroughly developed, see e. g.
[6, 8, 9, 16].
Fixed a Hilbert space (H,h , i), a frame for H is a (generally overcomplete) family of vectors F = {fi}i∈I
in H which satisfies the inequalities
Akfk2 ≤Xi∈I
h f, fi i2 ≤ Bkfk2,
for every f ∈ H,
(1)
for positive constants 0 < A ≤ B. The (bounded, linear) operator S : H → H defined by
Sf =Xi∈I
h f, fi i fi,
f ∈ H,
(2)
is known as the frame operator associated to F . The inequalities in Eq. (1) imply that S is a (positive)
boundedly invertible operator, and it allows to reconstruct each vector f ∈ H in terms of the family F
as follows:
(3)
f =Xi∈I (cid:10) f, S−1fi(cid:11) fi =Xi∈I
h f, fi i S−1fi.
The above formula is known as the reconstruction formula associated to F . Notice that if F is a Parseval
if S = I, then the reconstruction formula resembles the Fourier series of f associated to an
frame, i.e.
orthonormal basis B = {bk}k∈K of H:
f = Xk∈K
h f, bk i bk,
but the frame coefficients {h f, fi i}i∈I given by F allow to reconstruct f even when some of these
coefficients are missing (or corrupted). Indeed, each vector f ∈ H may admit several reconstructions
in terms of the frame coefficients as a consequence of the redundancy of F . These are some of the
advantages of frames over (orthonormal, orthogonal or Riesz) bases in signal processing applications,
when noisy channels are involved, e.g. see [3, 17, 22].
Given a Krein space (H, [ , ]) with fundamental symmetry J, a J-orthonormalized system is a family
E = {ei}i∈I such that [ ei, ej ] = ±δij, for i, j ∈ I. A J-orthonormal basis is a J-orthonormalized system
which is also a Schauder basis for H. If E = {ei}i∈I is a J-orthonormal basis of H then the vectors in H
can be represented as follows:
(4)
σi [ f, ei ] ei,
f ∈ H,
f =Xi∈I
1
where σi = [ ei, ei ] = ±1.
J-orthonormalized systems (and bases) are intimately related to the notion of dual pair. In fact, each
J-orthonormalized system generates a dual pair, i.e. a pair (L+,L−) of subspaces of H such that L+ is
J-nonnegative, L− is J-nonpositive and L+ is J-orthogonal to L−, i.e.
[L+,L− ] = 0. Moreover, if E
is a J-orthonormal basis of H, the dual pair associated to E is maximal (with respect to the inclusion
preorder) and the subspaces L+ and L− are uniformly J-definite, see [18, Ch.1, §10]. Therefore the
dual pair (L+,L−) is a fundamental decomposition of H. Notice that, considering the Hilbert space
structure induced by the above fundamental decomposition, the J-orthonormal basis E turns out to be
an orthonormal basis in the associated Hilbert space. Therefore, each J-orthonormal basis can be realized
as an orthonormal basis of H (respect to an appropriate definite inner product).
Given a pair of maximal uniformly J-definite subspaces M+ and M− of a Krein space H, with
different positivity, if F± = {fi}i∈I± is a frame for the Hilbert space (M±,±[ , ]), it is easy to see that
F = F+ ∪ F−,
is a frame for H, which produces an indefinite reconstruction formula:
f =Xi∈I
σi[ f, gi ]fi =Xi∈I
σi[ f, fi ]gi,
f ∈ H,
where σi = sgn[ fi, fi ] and {gi}i∈I is some (equivalent) frame for H (see Example 2 and Proposition 5.3).
The aim of this work is to introduce and characterize a particular family of frames for a Krein space
(H, [ , ]) -hereafter called J-frames- that are compatible with the indefinite inner product [ , ]. Some
different approaches to frames for Krein spaces and indefinite reconstruction formulas are developed in
[14] and [21], respectively.
The paper is organized as follows: Section 2 contains some preliminaries results both in Krein spaces
and in frame theory for Hilbert spaces.
Section 3 presents the J-frames. Briefly, a J-frame for the Krein space (H, [ , ]) is a Bessel family
F = {fi}i∈I with synthesis operator T : ℓ2(I) → H such that the ranges of T+ := T P+ and T− :=
T (I − P+) are maximal uniformly J-positive and maximal uniformly J-negative subspaces, respectively,
where I+ = {i ∈ I : [ fi, fi ] > 0} and P+ is the orthogonal projection onto ℓ2(I+), as a subspace of ℓ2(I).
It is immediate that J-orthonormal bases are J-frames, because they generate maximal dual pairs
[18, Ch. 1, §10.12].
Also, if F is a J-frame for H, observe that R(T ) = R(T+) + R(T−) and recall that the sum of a
pair of maximal uniformly J-definite subspaces with different positivity coincides with H [2, Corollary
1.5.2]. Therefore, each J-frame is in fact a frame for H in the Hilbert space sense. Moreover, it is shown
that F+ = {fi}i∈I+ is a frame for the Hilbert space (R(T+), [ , ]) and F− = {fi}i∈I\I+ is a frame for
(R(T−),−[ , ]), i.e. there exist constants B− ≤ A− < 0 < A+ ≤ B+ such that
A±[ f, f ] ≤ Xi∈I±
[ f, fi ]2 ≤ B±[ f, f ]
for every f ∈ R(T±).
(5)
The optimal constants satisfying the above inequalities can be characterized in terms of T± and the
Gramian operators of their ranges.
This section ends with a geometrical characterization of J-frames, in terms of the (minimal) angles
between the uniformly J-definite subspace R(T±) and the cone of neutral vectors of the Krein space.
Section 4 is devoted to study the synthesis operators associated to J-frames. Fixed a Krein space H
and given a bounded operator T : ℓ2(I) → H, it is described under which conditions T is the synthesis
operator of a J-frame.
In Section 5 the J-frame operator is introduced. Given a J-frame F = {fi}i∈I , the J-frame operator
S : H → H is defined by
Sf =Xi∈I
σi [ f, fi ] fi,
f ∈ H,
where σi = sgn([ fi, fi ]). This operator resembles the frame operator for frames in Hilbert spaces (see Eq.
(2)), and it has similar properties, in particular S = T T # if T : ℓ2(I) → H is the synthesis operator of
2
F (see Proposition 5.1). Furthermore, each J-frame F = {fi}i∈I determines an indefinite reconstruction
formula, which depends on the J-frame operator S:
f =Xi∈I
σi [ f, S−1fi ] fi =Xi∈I
σi [ f, fi ] S−1fi,
for every f ∈ H.
(6)
In this case the family {S−1fi}i∈I turns out to be a J-frame too.
Finally, it will be shown that the J-frame operator of a J-frame F is intimately related to the
projection Q = PR(T+)//R(T−) determined by the decomposition H = R(T+) ∔ R(T−). In fact, fixed a
J-selfadjoint invertible operator S acting on a Krein space H, it is the J-frame operator for a J-frame
F if and only if there exists a projection Q with uniformly J-definite range and kernel such that QS is a
J-positive operator and (I − Q)S is a J-negative operator, see Theorem 5.5.
2 Preliminaries
If K is another Hilbert space then
Along this work H denotes a complex (separable) Hilbert space.
L(H,K) is the algebra of bounded linear operators from H into K and L(H) = L(H,H). The groups of
linear invertible and unitary operators acting on H are denoted by GL(H) and U(H), respectively. Also,
L(H)+ denotes the cone of positive semidefinite operators acting on H and GL(H)+ = GL(H) ∩ L(H)+.
If T ∈ L(H,K) then T ∗ ∈ L(K,H) denotes the adjoint operator of T , R(T ) stands for its range and
N (T ) for its nullspace. Also, if T ∈ L(H,K) has closed range, T † ∈ L(K,H) denotes the Moore-Penrose
inverse of T .
Hereafter, S ∔ T denotes the direct sum of two (closed) subspaces S and T of H. On the other hand,
S ⊕ T stands for the (direct) orthogonal sum of them and S ⊖ T := S ∩ (S ∩ T )⊥. If H = S ∔ T , the
oblique projection onto S along T is the unique projection with range S and nullspace T . It is denoted
by PS//T . In particular, PS := PS//S⊥ is the orthogonal projection onto S.
2.1 Krein spaces
In what follows we present the standard notation and some basic results on Krein spaces. For a complete
exposition on the subject (and the proofs of the results below) see the books by J. Bogn´ar [4] and T. Ya.
Azizov and I. S. Iokhvidov [18] and the monographs by T. Ando [2] and by M. Dritschel and J. Rovnyak
[13].
Given a Krein space (H, [ , ]) with a fundamental decomposition H = H+∔H−, the direct (orthogonal)
Observe that the indefinite metric and the inner product of H are related by means of a fundamental
sum of the Hilbert spaces (H+, [ , ]) and (H−,−[ , ]) is denoted by (H,h , i).
symmetry, i.e. a unitary selfadjoint operator J ∈ L(H) which satisfies:
[ x, y ] = h Jx, y i ,
x, y ∈ H.
If H and K are Krein spaces, L(H,K) stands for the vector space of linear transformations which are
bounded respect to the associated Hilbert spaces (H,h , iH) and (K,h , iK). Given T ∈ L(H,K), the
J-adjoint operator of T is defined by T # = JHT ∗JK, where JH and JK are the fundamental symmetries
associated to H and K, respectively. An operator T ∈ L(H) is J-selfadjoint if T = T #.
A vector x ∈ H is J-positive if [ x, x ] > 0. A subspace S of H is J-positive if every x ∈ S, x 6= 0, is a
J-positive vector. A subspace S of H is uniformly J-positive if there exists α > 0 such that
[ x, x ] ≥ αkxk2,
for every x ∈ S,
J-nonnegative, J-neutral, J-negative, J-nonpositive and uniformly J-negative vectors and subspaces
where k k stands for the norm of the associated Hilbert space (H,h , i).
are defined analogously.
Remark 2.1. If S+ is a closed uniformly J-positive subspace of a Krein space (H, [ , ]), observe that
(S+, [ , ]) is a Hilbert space. In fact, the forms [ , ] and h , i are equivalent inner products on S+, because
αkfk2 ≤ [ f, f ] ≤ kfk2,
for every f ∈ S+.
Analogously, if S− is a closed uniformly J-negative subspace of (H, [ , ]), (S−,−[ , ]) is a Hilbert space.
3
Proposition 2.2 ([18], Cor. 7.17). Let H be a Krein space with fundamental symmetry J and S a
J-nonnegative closed subspace of H. Then, S is the range of a J-selfadjoint projection if and only if S is
uniformly J-positive.
Recall that, given a closed subspace M of a Krein space H, the Gramian operator of M is defined by:
GM = PMJPM,
where PM is the orthogonal projection onto M and J is the fundamental symmetry of H.
If M is
J-semidefinite, then M ∩ M[⊥] coincides with N := {f ∈ M : [ f, f ] = 0}. Therefore, it is easy to see
that
GM = GM⊖N .
Given a subspace S of a Krein space H, the J-orthogonal companion to S is defined by
S[⊥] = {x ∈ H : [ x, s ] = 0 for every s ∈ S}.
A subspace S of H is J-non degenerated if S ∩ S[⊥] = {0}. Notice that if S is a J-definite subspace of H
then it is J-non degenerated.
2.2 Angles between subspaces and reduced minimum modulus
Given two closed subspaces S and T of a Hilbert space H, the cosine of the Friedrichs angle between S
and T is defined by
c(S,T ) = sup{h x, y i : x ∈ S ⊖ T ,kxk = 1, y ∈ T ⊖ S,kyk = 1}.
It is well known that
c(S,T ) < 1 ⇔ S + T is closed ⇔ c(S⊥,T ⊥) < 1.
Furthermore, if PS and PT are the orthogonal projections onto S and T , respectively, then c(S,T ) < 1
if and only if (I − PS)PT has closed range. See [10] for further details.
The next definition is due to T. Kato, see [19, Ch. IV, § 5].
Definition. The reduced minimum modulus γ(T ) of an operator T ∈ L(H,K) is defined by
γ(T ) = inf{kT xk : x ∈ N (T )⊥, kxk = 1}.
Observe that γ(T ) = sup{C ≥ 0 : Ckxk ≤ kT xk for every x ∈ N (T )⊥, kxk = 1}. It is well known
that γ(T ) = γ(T ∗) = γ(T ∗T )1/2. Also, it can be shown that an operator T 6= 0 has closed range if and
only if γ(T ) > 0. In this case, γ(T ) = kT †k−1.
If H and K are Krein spaces with fundamental symmetries JH and JK, respectively, and T ∈ L(H,K)
then
γ(T #) = γ(JHT ∗JK) = γ(T ∗) = γ(T ),
because JH (resp. JK) is a unitary operator on H (resp. K).
Remark 2.3. If M+ is a closed J-nonnegative subspace of a Krein space H then
γ(GM+ ) = α+,
(7)
where α+ ∈ [0, 1] is the supremum among the constants α ∈ [0, 1] such that αkfk2 ≤ [ f, f ] for every
f ∈ M+. From now on, the constant α+ is called the definiteness bound of M+. Notice that α+ is in
fact a maximum for the above set and M+ is uniformly J-positive if and only if α+ > 0.
Analogously, if M− is a J-nonpositive subspace then γ(GM− ) = α−, where α− is the definiteness
bound of M−, i.e.
α− = max{α ∈ [0, 1] :
[ f, f ] ≤ −αkfk2 for every f ∈ M−}.
4
2.3 Frames for Hilbert spaces
The following is the standard notation and some basic results on frames for Hilbert spaces, see [6, 8, 16].
A frame for a Hilbert space H is a family of vectors F = {fi}i∈I ⊂ H for which there exist constants
0 < A ≤ B < ∞ such that
A kfk2 ≤Xi∈I
hf, fii2 ≤ B kfk2 ,
for every f ∈ H.
(8)
The optimal constants (maximal for A and minimal for B) are known, respectively, as the upper and
lower frame bounds.
If a family of vectors F = {fi}i∈I satisfies the upper bound condition in (8), then F is a Bessel family.
For a Bessel family F = {fi}i∈I , the synthesis operator T ∈ L(ℓ2(I),H) is defined by
T x =Xi∈I
h x, ei i fi,
where {ei}i∈I is the standard orthonormal basis of ℓ2(I). It holds that F is a frame for H if and only if
T is surjective. In this case, the operator S = T T ∗ ∈ L(H) is invertible and is called the frame operator.
It can be easily verified that
(9)
h f, fi i fi,
for every f ∈ H.
Sf =Xi∈I
This implies that the frame bounds can be computed as: A = kS−1k−1 and B = kSk. From (9), it is
also easy to obtain the canonical reconstruction formula for the vectors in H:
f =Xi∈I (cid:10) f, S−1fi(cid:11) fi =Xi∈I
h f, fi i S−1fi,
for every f ∈ H,
and the frame {S−1fi}i∈I is called the canonical dual frame of F . More generally, if a frame G = {gi}i∈I
satisfies
(10)
h f, fi i gi,
for every f ∈ H,
f =Xi∈I
then G is called a dual frame of F .
h f, gi i fi =Xi∈I
3
J -frames: definition and basic properties
Let H be a Krein space with fundamental symmetry J. Given a Bessel family F = {fi}i∈I in H consider
the synthesis operator T ∈ L(ℓ2(I),H). If I+ = {i ∈ I :
[ fi, fi ] < 0},
consider the orthogonal decomposition of ℓ2(I) given by
[ fi, fi ] ≥ 0} and I− = {i ∈ I :
and denote by P± the orthogonal projection onto ℓ2(I±). Also, let T± = T P±. If M± = span{fi : i ∈ I±},
notice that span{fi : i ∈ I±} ⊆ R(T±) ⊆ M± and
ℓ2(I) = ℓ2(I+) ⊕ ℓ2(I−),
(11)
R(T ) = R(T+) + R(T−).
Definition. The Bessel family F = {fi}i∈I is a J-frame for H if R(T+) is a maximal uniformly J-positive
subspace of H and R(T−) is a maximal uniformly J-negative subspace of H.
Notice that, in particular, every J-orthogonalized basis of a Krein space H is a J-frame for H, because
it generates a maximal dual pair, see [18, Ch. 1, §10.12].
If F is a J-frame, as a consequence of its maximality, R(T±) is closed. So, R(T±) = M± and, by [2,
Corollary 1.5.2], M+ + M− = H. Then, it follows that F is a frame for the associated Hilbert space
(H,h , i) because
R(T ) = R(T+) + R(T−) = M+ + M− = H.
5
Given a Bessel family F = {fi}i∈I , consider the subspaces R(T+) and R(T−) as above. If K± : D± →
H∓ is the angular operator associated to R(T±), the operator of transition associated to the Bessel family
F is defined by
F = K+P + K−(I − P ) : D+ + D− → H,
2 (I + J) is the J-selfadjoint projection onto H+ and D± is a subspace of H± (the domain of
where P = 1
K±), see [15].
Proposition 3.1. Let F = {fi}i∈I be a Bessel family in H. Then, F is a J-frame if and only if F is
everywhere defined (i.e. D+ + D− = H) and kFk < 1.
Proof. Proof See [15, Proposition 2.6].
It follows from the definition that, given a J-frame F = {fi}i∈I for the Krein space H, [ fi, fi ] 6= 0 for
every i ∈ I, i.e. I± = {i ∈ I : ±[ fi, fi ] > 0}. This fact allows to endow the coefficients space ℓ2(I) with
a Krein space structure. Denote σi = sgn([ fi, fi ]) = ±1 for every i ∈ I. Then, the diagonal operator
J2 ∈ L(ℓ2(I)) defined by
(12)
J2 ei = σi ei,
for every i ∈ I,
is a selfadjoint involution on ℓ2(I). Therefore, ℓ2(I) with the fundamental symmetry J2 is a Krein space.
Now, if T ∈ L(ℓ2(I),H) is the synthesis operator of F , the J-adjoints of T , T+ and T− can be easily
calculated, in fact if f ∈ H:
T #
± f = ± Xi∈I±
[ f, fi ]ei,
and T #f = (T+ + T−)#f = T #
+ f + T #
− f =Pi∈I+
[ f, fi ]ei −Pi∈I− [ f, fi ]ei =Pi∈I σi[ f, fi ]ei.
Example 1. It is easy to see that not every frame of J-nonneutral vectors is a J-frame: given the Krein
space obtained by endowing C3 with the sesquilinear form
[(x1, x2, x3), (y1, y2, y3)] = x1y1 + x2y2 − x3y3,
consider f1 = (1, 0, 1√2
C3 because it is a (linear) basis for the space.
), f2 = (0, 1, 1√2
) and f3 = (0, 0, 1). Observe that F = {f1, f2, f3} is a frame for
On the other hand, M+ = span{f1, f2} and M− = span{f3}. If (a, b, 1√2
vector in M+ then
so M+ is a J-nonnegative subspace of C3. But M+ is not uniformly J-positive, because (1, 1,√2) ∈ M+
is a (non trivial) J-neutral vector. Therefore, F is not a J-frame for (C3, [ , ]).
[ f, f ] = a2 + b2 − 1
2a − b2 ≥ 0,
2a + b2 = 1
(a + b)) is an arbitrary
The following is a handy way to construct J-frames for a given Krein space. Along this section, it
will be shown that every J-frame can be realized in this way.
Example 2. Given a Krein space H with fundamental symmetry J, let M+ (resp. M−) be a maximal
uniformly J-positive (resp. J-negative) subspace of H. If F± = {fi}i∈I± is a frame for the Hilbert space
(M±,±[ , ]) then F = F+ ∪ F− is a J-frame for H.
Indeed, by Remark 2.1, F+ and F− are Bessel families in H. Hence, F is a Bessel family and, if
I = I+ ∪I− (the disjoint union of I+ and I−), the synthesis operator T ∈ L(ℓ2(I),H) of F is given by
T x = T+x+ + T−x−
if
x = x+ + x− ∈ ℓ2(I+) ⊕ ℓ2(I−) =: ℓ2(I),
where T± : ℓ2(I±) → M± is the synthesis operator of F±. Then, it is clear that R(T P±) = M± is a
maximal uniformly J-definite subspace of H.
6
Proposition 3.2. Let F = {fi}i∈I be a J-frame for H. Then, F± = {fi}i∈I± is a frame for the Hilbert
space (M±,±[ , ]), i.e. there exist constants B− ≤ A− < 0 < A+ ≤ B+ such that
for every f ∈ M±.
[ f, fi ]2 ≤ B±[ f, f ]
(13)
A±[ f, f ] ≤ Xi∈I±
Proof. Proof If F = {fi}i∈I is a J-frame for H, then R(T+) = M+ is a (maximal) uniformly J-positive
subspace of H. So, T+ is a surjection from ℓ2(I) onto the Hilbert space (M+, [ , ]). Therefore, F+ is a
frame for (M+, [ , ]). In particular, there exist constants 0 < A+ ≤ B+ such that Eq. (13) is satisfied
for M+. The assertion on F− follows analogously.
Now, assuming that F is a J-frame for a Krein space (H, [ , ]), a set of constants {B−, A−, A+, B+}
satisfying Eq. (13) is going to be computed. They depend only on the definiteness bounds for R(T±),
the norm and the reduced minimum modulus of T±.
Suppose that F is a J-frame for a Krein space (H, [ , ]) with synthesis operator T ∈ L(ℓ2(I),H).
Since R(T+) = M+ is a (maximal) uniformly J-positive subspace of H, there exists α+ > 0 such that
α+kfk2 ≤ [ f, f ] for every f ∈ M+. So,
Xi∈I+
where B+ = kT #
+ k2
α+
[ f, fi ]2 = kT #
+ fk2 ≤ kT #
+ k2kfk2 ≤ B+[ f, f ],
for every f ∈ M+,
= kT+k2
α+
. Furthermore, since N (T #
+ )⊥ = J(M+), if f ∈ M+,
Xi∈I+
[ f, fi ]2 = kT #
+ fk2 = kT #
+ PJ(M+)fk2 ≥ γ(T #
+ )2kPJ(M+)fk2 = γ(T+)2kPM+ Jfk2 =
= γ(T+)2kGM+ fk2 ≥ γ(T+)2γ(GM+ )2kfk2 ≥ A+[ f, f ],
where A+ = γ(T+)2γ(GM+ )2 = γ(T+)2α2
+, see Remark 2.3.
is the definiteness bound of the (maximal) uniformly J-negative subspace M−.
Analogously, A− = −γ(T−)2α2
Usually, the bounds A± = ±α2
− and B− = − kT−k2
±γ(T±)2 and B± = ± kT±k2
α±
α−
satisfy Eq (13) for every f ∈ R(T−) = M−, if α−
are not optimal for the J-frame F .
Definition. Let F be a J-frame for the Krein space H. The optimal constants B− ≤ A− < 0 < A+ ≤ B+
satisfying Eq. (13) are called the J-frame bounds of F .
In order to compute the J-frame bounds associated to a J-frame F = {fi}i∈I , consider the uniformly
J-definite subspaces M+ and M−. Recall that F+ = {fi}i∈I+ is a frame for the Hilbert space (M+, [ , ]).
Then, if G+ = GM+M+ ∈ GL(M+), the frame bounds for F+ are given by A+ = k(SG+ )−1k−1
+ and
B+ = kSG+k+, where SG+ = T+T ∗+G+ is the frame operator of F+ and kfk+ = [ f, f ]1/2 = kG1/2
+ fk,
f ∈ M+, is the operator norm associated to the inner product [ , ]. Therefore,
A+ = k(SG+)−1k−1
+ = kG1/2
+ (T+T ∗+G+)−1k−1 = kG−1/2
+
(T+T ∗+)−1k−1,
and B+ = kSG+k+ = kG1/2
Hilbert space (M−,−[ , ]). So, the frame bounds for F− are given by
and B− = kG1/2
A− = kG−1/2
−)−1k−1
(T−T ∗
−
− T−T ∗
−G−k,
+ T+T ∗+G+k. Analogously, it follows that F− = {fi}i∈I− is a frame for the
where G− = GM−M− ∈ GL(M−). Thus, the J-frame bound associated to F can be fully characterized
in terms of T± and the Gramian operators GM±.
3.1 Characterizing J -frames in terms of frame inequalities
Given a Bessel family F = {fi}i∈I in a Krein space H, the inequalities:
A [ f, f ] ≤Xi∈I
[ f, fi ]2 ≤ B [ f, f ]
for every f ∈ M = span{fi : i ∈ I},
(14)
7
with B ≥ A > 0, ensure that M is a J-nonnegative subspace of H. However, they do not imply that M
is uniformly J-positive, i.e. (M, [ , ]) is not necessarily a inner product space. See the example below.
Example 3. Consider again the Krein space (C3, [ , ]) as in Example 1. As it was mentioned before,
M = span{f1 = (1, 0, 1/√2), f2 = (0, 1, 1/√2)} is a J-nonnegative but not uniformly J-positive subspace
of C3.
In this case, the orthogonal basis
2 , 1√2
v1 = ( 1
2 , 1
) , v2 = ( 1√2
, −1√2
, 0) and v3 = ( 1√2
, 1√2
,−1),
is a basis of eigenvectors of GM, corresponding to the eigenvalues λ1 = 0, λ2 = 1 and λ3 = 0, respectively.
Moreover, M = span{v1, v2}. Thus, if f ∈ M there exists α, β ∈ C such that f = αv1 + βv2 and then,
since GMv1 = 0 ∈ C3, it is easy to see that
[ f, f1 ]2 + [ f, f2 ]2 = β2(h v2, f1 i2 + h v2, f2 i2) = β2 = [ f, f ].
Therefore, Eq. (14) holds with A = B = 1, but {f1, f2} cannot be extended to a J-frame, since M is
not a uniformly J-positive subspace.
The next result gives a complete characterization of the families satisfying Eq. (14) for B ≥ A > 0.
It is straightforward to formulate and prove analogues of all these assertions for a family satisfying Eq.
(14) for negative constants B ≤ A < 0.
Proposition 3.3. Given a Bessel family F = {fi}i∈I in a Krein space H, let M = span{fi : i ∈ I}
and N = M ∩ M[⊥]. If there exist constants 0 < A ≤ B such that
for every f ∈ M,
[ f, fi ]2 ≤ B [ f, f ]
A [ f, f ] ≤Xi∈I
then M ⊖ N is a (closed) uniformly J-positive subspace of M. Moreover, if F is a frame for the Hilbert
space (M,h , i), the converse holds.
Proof. Proof First, suppose that there exist 0 < A ≤ B such that Eq. (15) holds. So, M is a J-
nonnegative subspace of H, or equivalently, (M, [ , ]) is a semi-inner product space.
If T ∈ L(ℓ2(I),H) is the synthesis operator of the Bessel sequence F and C = kT ∗k2 > 0, then
T T ∗ ≤ CPM. So, using Eq. (15) it is easy to see that:
(15)
Ah GMf, f i ≤ kT #(PMf )k2 = h (PMJT T ∗JPM)f, f i ≤ C(cid:10) (GM)2f, f(cid:11) , f ∈ H.
A (GM)2. Applying Douglas' theorem [11] it is easy to see that
Thus, 0 ≤ GM ≤ C
(16)
R((GM)1/2) ⊆ R(GM) ⊆ R((GM)1/2).
Moreover, it follows that R(GM) is closed because R(GM) = R((GM)1/2).
R(GM) is closed, there exists α > 0 such that
Let M′ = M ⊖ N and notice that M′ is a closed uniformly J-positive subspace of H. In fact, since
[ f, f ] = h GMf, f i = k(GM)1/2fk2 ≥ αkfk2
for every f ∈ N (GM)⊥ = M ⊖ N .
Conversely, suppose that F is a frame for (M,h , i), i.e. there exist constants B′ ≥ A′ > 0 such that
A′PM ≤ T T ∗ ≤ B′PM,
where T ∈ L(ℓ2(I),M) is the synthesis operator of F . If M′ = M⊖N is a uniformly J-positive subspace
of H, then there exists α > 0 such that αPM′ ≤ GM′ ≤ PM′ . As a consequence of Douglas' theorem,
R((GM′ )1/2) = M′ = R(GM′ ). Since GM = GM′ it is easy to see that
A′(GM)2 = A′(GM′ )2 ≤ PMJT T ∗JPM ≤ B′(GM′ )2 = B′(GM)2.
Therefore, R(PMJT ) = R(GM′ ) = R((GM′ )1/2), or equivalently, there exist B ≥ A > 0 such that
AGM = AGM′ ≤ PMJT T ∗JPM ≤ BGM′ = BGM,
i.e. A [ f, f ] ≤Pi∈I [ f, fi ]2 ≤ B [ f, f ] for every f ∈ M.
8
Theorem 3.4. Let F = {fi}i∈I be a frame for H.
span{fi : i ∈ I±} then, F is a J-frame if and only if M± ∩ M[⊥]
A− < 0 < A+ ≤ B+ such that
If I± = {i ∈ I : ±[ fi, fi ] ≥ 0} and M± =
± = {0} and there exist constants B− ≤
A± [ f, f ] ≤ Xi∈I±
[ f, fi ]2 ≤ B± [ f, f ]
for every f ∈ M±.
(17)
Proof. Proof If F is a J-frame, the conditions on M± follow by its definition and by Proposition 3.2.
Conversely, if M+ is J-non degenerated and there exist constants 0 < A+ ≤ B+ such that
A+ [ f, f ] ≤ Xi∈I±
[ f, fi ]2 ≤ B+ [ f, f ]
for every f ∈ M+,
then, by Proposition 3.3, M+ is a uniformly J-positive subspace of H. Therefore, there exist constants
0 < A ≤ B such that
A kPM+fk2 ≤ kT #
But these inequalities can be rewritten as
+ PM+ fk2 ≤ BkPM+ fk2
for every f ∈ H.
A PM+ ≤ PM+ JT+T ∗+JPM+ ≤ B PM+ .
Then, by Douglas' theorem, R(PM+ JT+) = R(PM+ ) = M+. Furthermore, PJ(M+)(R(T+)) = J(M+)
because
J(M+) = J(R(PM+ JT+)) = R((JPM+ J)T+) = R(PJ(M+)T+) = PJ(M+)(R(T+)).
Therefore, taking the counterimage of PJ(M+)(R(T+)) by PJ(M+), it follows that
H = R(T+) ∔ J(M+)⊥ ⊆ M+ ∔ M[⊥]
+ = H.
Thus, R(T+) = M+ and F+ is a frame for M+. Analogously, F− = {fi}i∈I− is a frame for M−. Finally,
since F is a frame for H,
H = R(T ) = R(T+) + R(T−),
which proves the maximality of R(T±). Thus, F is a J-frame for H.
3.2 A geometrical characterization of J -frames
Let F = {fi}i∈I be a J-frame for H and consider F = F+ ∪ F+ the partition of F into J-positive and
J-negative vectors. Moreover, let M± be the (maximal) uniformly J-definite subspace of H generated
by F±.
The aim of this section is to show that it is possible to bound the correlation between vectors in F+
(resp. F−) and vectors in the cone of neutral vectors C = {n ∈ H :
[ n, n ] = 0}, in a strong sense:
h f, ni ≤ c± kfk knk , f ∈ M± , n ∈ C ,
(18)
√2
2 ≤ c± < 1. In order to make these ideas precise, consider the notion of minimal
for some constants
angle between a subspace M and the cone C.
Definition. Given a closed subspace M of the Krein space H, consider
c0(M,C) = sup{h m, ni : m ∈ M, n ∈ C, knk = kmk = 1} ,
(19)
Then, there exists a unique θ(M,C) ∈ [0, π
the minimal angle between M and C.
4 ] such that cos(θ(M,C)) = c0(M,C). In this case, θ(M,C) is
9
Observe that if the subspace M contains a non trivial J-neutral vector (e.g. if M is J-indefinite or
J-semidefinite) then c0(M,C) = 1, or equivalently, θ(M,C) = 0. On the other hand, it will be shown
that the minimal angle between a uniformly J-positive (resp. uniformly J-negative) subspace M and C
is always bounded away from 0.
Proposition 3.5. Let M be a J-semidefinite subspace of H with definiteness bound α. Then,
1
c0(M,C) =
√2 r 1 + α
+r 1 − α
In particular, M is uniformly J-definite if and only if c0(M,C) < 1.
Proof. Proof Let H = H+ ⊕ H− be a fundamental decomposition of H and suppose that M is a
J-nonnegative subspace of H.
Let m ∈ M with kmk = 1. Then, there exist (unique) m± ∈ H± such that m = m+ + m−. In this
case,
(21)
2 ! .
(20)
and
2
α ≤ [ m, m ] = km+k2 − km−k2.
1 = kmk2 = km+k2 + km−k2
Claim: Fixed m ∈ M with kmk = 1, sup{h m, ni : n ∈ C, knk = 1} = 1√2
(km+k + km−k).
Indeed, consider n ∈ C with knk = 1. Then, there exist (unique) n± ∈ H± such that n = n+ + n−.
In this case,
which imply that kn+k = kn−k = 1√2
0 = [ n, n ] = kn+k2 − kn−k2
. Therefore,
and
1 = knk2 = kn+k2 + kn−k2,
h m, ni ≤ (cid:10) m+, n+(cid:11) + (cid:10) m−, n−(cid:11) ≤
1
√2
(km+k + km−k).
On the other hand, if m− 6= 0 then let nm := 1√2
with z ∈ H−, kzk = 1. Now, it is easy to see that nm ∈ C and that h m, nm i = 1√2
which together with the previous facts prove the claim.
), otherwise consider nm = 1√2
(m + z),
(km+k + km−k)
( m+
km+k
+ m−
km−k
Now, let M1 = {m = m+ + m− ∈ M : m± ∈ H±, kmk = 1}. Using the claim above it follows that
c0(M,C) =
1
√2
sup
m∈M1
(km+k + km−k).
(22)
. Thus, in this
If α = 1 then M is a subspace of H+. Also, it is easy to see that c0(M,C) = 1√2
particular case, c0(M,C) = 1√2(cid:16)q 1+α
On the other hand, if α < 1, let k0 ∈ N be such that 1−α
of the definiteness bound, for every integer k ≥ k0 there exists mk = m+
α ≤ km+
2 +q 1−α
2 (cid:17).
k k2 − km−k k2 < α + 1
k . Then, it follows that
2 > 1
2k0
. Observe that, by the definition
k + m−k ∈ M1 such that
α + 1 ≤ 2km+
k k2 < α + 1 +
1
k
,
2 + 1
k k <q α+1
r 1 − α
2 −
or equivalently, q α+1
2 ≤ km+
2k . Moreover, km−k k =q1 − km+
k k2 implies that
< km−k k ≤r 1 − α
Therefore, for every integer k ≥ k0 there exists mk ∈ M1 such that
k k + km−k k <r α + 1
r 1 − α
2 −
+r α + 1
< km+
1
2k
1
2k
2
2
2
.
+
1
2k
+r 1 − α
2
.
10
2 +q 1−α
2 (cid:17).
Thus, c0(M,C) = 1√2(cid:16)q 1+α
Assume now that M is a J-nonpositive subspace of (H, [ , ]) with definiteness bound α, for 0 ≤ α ≤ 1.
Then, M is a J-nonnegative subspace of the antispace (H,−[ , ]), with the same definiteness bound α.
Furthermore, the cone of J-neutral vectors for the antispace is the same as for the initial Krein space
(H, [ , ]). Therefore, we can apply the previous arguments and conclude that Eq.(20) also holds for
J-nonpositive subspaces.
Finally, the last assertion in the statement follows from the formula in Eq. (20).
Let F be a J-frame for H as above. Notice that the Eq. (18) holds for some constant
√2
2 ≤ c± < 1
if and only if c0(M±,C) < 1, i.e. that the minimal angles θ(M±,C) are bounded away from 0. This is
intimately related with the fact that the aperture between the subspaces M+ (resp. M−) and H+ (resp.
H−) is bounded away from π
Remark 3.6. Given a Krein space H, fix a fundamental decomposition H = H+ ⊕ H−. Then, if M is a
J-nonnegative subspace of H the minimal angle between M and C is related with the aperture Φ(M,H+)
between the subspaces M and H+, see [1] and Exercises 3 -- 6 to [18, Ch. 1, §8]. In fact, if K ∈ L(H+,H−)
is the angular operator associated to M then, by [18, Ch. 1, §8 Exercise 4],
4 , whenever H = H+ ⊕ H− is a fundamental decomposition.
Φ(M,H+) =
p1 + kKk2
Also, if α is the definiteness bound of M then kKk = q 1−α
Φ(M,H+) =
between M and H+, it is easy to see that
kKk√1+kKk2 = q 1−α
kKk
.
1+α , see [18, Ch. 1, Lemma 8.4]. Therefore,
2 . Since Φ(M,H+) = sin ϕ(M,H+) for an angle ϕ(M,H+) ∈ [0, π
4 ]
cos ϕ(M,H+) =q1 − sin2 ϕ(M,H+) =r 1 + α
Therefore, if ϕ = ϕ(M,H+),
√2
2
i.e. ϕ(M,H+) + θ(M,C) = π
4 .
√2 r 1 + α
+r 1 − α
(cos ϕ + sin ϕ) =
cos( π
4 − ϕ) =
1
2
.
2
2 ! = c0(M,C),
The following result shows that, given a frame F = {fi}i∈I for H, the positivity of the angles θ(M±,C)
characterize it as a J-frame for H.
Proposition 3.7. Let F = {fi}i∈I be a frame for a Krein space H. Then, F is a J-frame for H if and
only if there exists a partition I = I1 ∪ I2 such that
θ(Mj ,C) > 0
for j = 1, 2,
(23)
where Mj = span{fi : i ∈ Ij}.
Proof. Proof If we assume that F is a J-frame then, consider I± and M± as usual. Then I = I+ ∪ I−
is a partition of I into disjoint sets and M± are uniformly J-definite subspaces associated to F . Hence,
by Proposition 3.5, we see that Eq. (23) holds in this case.
Conversely, assume that there exists a partition of I with the properties above. Notice that Proposition
3.5 implies that Mj is a uniformly J-definite subspace of H, for j = 1, 2. On the other hand, since F is a
frame, H ⊆ M1 +M2. Therefore, M1 and M2 have different positivity and they are maximal uniformly
J-definite subspaces. Suppose that M1 is uniformly J-positive and M2 is uniformly J-negative.
If
F = {fi}i∈I is a frame for H, its synthesis operator T ∈ L(ℓ2(I),H) is surjective. Therefore,
Then, consider the orthogonal projection Pj ∈ L(ℓ2(I)) onto the subspace ℓ2(Ij ), for j = 1, 2.
where Tj = T Pj, for j = 1, 2. Then, it is easy to see that R(Tj) = Mj for j = 1, 2 and F is a J-frame
for H.
R(T1) ∔ R(T2) = R(T ) = H,
11
Remark 3.8. Let (H,h , i) be a separable Hilbert space that models a signal space in which is con-
sidered a linear (robust and stable) encoding-decoding scheme for certain measurements, i.e. consider a
(redundant) frame G = {gi}i∈K for H.
Assume that the measurements of x ∈ H are given by y1 = P x and y2 = (I − P )x, where P ∈ L(H)
is an orthogonal projection (for instance, P and I − P are low pass and high pass filters, respectively).
Suppose that the signals having the same energy in R(P ) and R(I − P ) = N (P ) (i.e. signals x ∈ H such
that ky1k2 = ky2k2) are considered disturbances, see e.g. [5, 20].
Notice that, sampling the measurements y1, y2 with the frame G is the same as sampling y = (y1, y2) ∈
H × H with the frame F = {fi}i∈I = {(gi, 0)}i∈K ∪ {(0, gi)}i∈K for H × H.
It is easy to see that, the space K = H × H with the indefinite product [ y, z ] = h y1, z1 i − h y2, z2 i is
a Krein space, where y = (y1, y2), z = (z1, z2) ∈ K are the measurements of signals in H. Observe that
the set of disturbances is characterized as the set of J-neutral vectors C of K.
Also, notice that F is a J-frame for K. Hence, the (sampling) vectors of the frame F are away from
the disturbances set C.
Now, consider any (redundant) J-frame F = {fi}i∈I for (K, [ , ]). As usual, denote M+ and M− the
maximal uniformly J-definite subspaces generated by F . Since M± is uniformly J-definite, Proposition
3.5 shows that c0(M±,C) < 1, which is a bound for the correlation between the sampling vectors in F
and the distrubances of C because
h fi, ni ≤ c0(M±,C)kfik knk whenever i ∈ I± and n ∈ C.
(24)
That is, J-frames provide a class of frames for K with the desired properties. Moreover, later in Propo-
sition 5.3, it will be shown that the J-frame F admits a (canonical) dual J-frame that induces a linear
(indefinite) stable and redundant encoding-decoding scheme in which the correlation between both the
sampling and reconstructing vectors and the cone of neutral vectors is bounded from above. These re-
marks provide a quantitative measure of the advantage of considering J-frames with respect to usual
frames in this setting.
4 On the synthesis operator of a J -frame
If F is a J-frame with synthesis operator T , then QT = T+ = T P+, where Q = PM+//M− . Therefore,
Q = QT T † = T P+T †.
So, given a surjective operator T : ℓ2(I) → H, the idempotency of T P+T † is a necessary condition for T
to be the synthesis operator of a J-frame.
Lemma 4.1. Let T ∈ L(ℓ2(I),H) be surjective. Suppose that PS is the orthogonal projection onto a
closed subspace S of ℓ2(I) such that c(S, N (T )⊥) < 1. Then, T PST † is a projection if and only if
N (T ) = S ∩ N (T ) ⊕ S⊥ ∩ N (T ).
Proof. Proof Suppose that Q = T PST † is a projection. Then, if P = PN (T )⊥ , E = P PS P is an
orthogonal projection because it is selfadjoint and
E2 = (P PSP )2 = P PS P PSP = T †(T PST †)2T = T †(T PST †)T = P PSP = E.
Therefore, (P PS )k = Ek−1PS = EPS = (P PS )2 for every k ≥ 2. So, by [10, Lemma 18],
P PS = PS ∧ P = PS P.
Then, since PS and P commute, it follows that N (T ) = S ∩ N (T ) ⊕ S⊥ ∩ N (T ) (see [10, Lemma 9]).
Conversely, suppose that N (T ) = S ∩ N (T ) ⊕ S⊥ ∩ N (T ). Then, PS and P commute and
(T PST †)2 = T PS(T †T )PST † = T PSP PST † = T P PST † = T PST †.
12
Hereafter consider the set of possible decompositions of H as a (direct) sum of a pair of maximal
uniformly definite subspaces, or equivalently, the associated set of projections:
Q = {Q ∈ L(H) : Q2 = Q, R(Q) is uniformly J-positive and N (Q) is uniformly J-negative}.
Proposition 4.2. Let T ∈ L(ℓ2(I),H) be surjective. Then, T is the synthesis operator of a J-frame if
and only if there exists I+ ⊂ I such that ℓ2(I+) (as a subspace of ℓ2(I)) satisfies c(N (T )⊥, ℓ2(I+)) < 1
and
T P+T † ∈ Q,
where P+ ∈ L(ℓ2(I)) is the orthogonal projection onto ℓ2(I+).
Proof. Proof If T is the synthesis operator of a J-frame, the existence of such a subset I+ has already
been discussed before.
Conversely, suppose that there exists such a subset I+ of I. Then, since c(N (T )⊥, ℓ2(I+)) < 1 and
Q = T P+T † ∈ Q, it follows from Lemma 4.1 that P+ and P = PN (T )⊥ commute. Therefore,
QT = T P+P = T P P+ = T P+,
and (I−Q)T = T (I−P+). Hence, R(T P+) = R(Q) is (maximal) uniformly J-positive and R(T (I−P+)) =
N (Q) is (maximal) uniformly J-negative. Therefore F = {T ei}i∈I is by definition a J-frame for H.
Theorem 4.3. Given a surjective operator T ∈ L(ℓ2(I),H), the following conditions are equivalent:
1. There exists U ∈ U(ℓ2(I)) such that T U is the synthesis operator of a J-frame.
2. There exists Q ∈ Q such that
QT T ∗(I − Q)∗ = 0.
(25)
3. There exist closed range operators T1, T2 ∈ L(ℓ2(I),H) such that T = T1 + T2, R(T1) is uniformly
J-positive, R(T2) is uniformly J-negative and T1T ∗2 = T2T ∗1 = 0.
Proof. Proof 1. ⇒ 2.: Suppose that there exists U ∈ U(ℓ2(I)) such that V = T U is the synthesis
operator of a J-frame. If I± = {i ∈ I : ±[ V ei, V ei ] > 0} and P± ∈ L(ℓ2(I)) is the orthogonal projection
onto ℓ2(I±), define V± = V P±. Then, V = V+ + V− and M± = R(V±) is a maximal uniformly J-definite
subspace. So, considering Q = PM+//M− ∈ Q, it is easy to see that QV = V+, (I − Q)V = V− and
QT T ∗(I − Q)∗ = QV V ∗(I − Q)∗ = V+V ∗
− = V P+P−V ∗ = 0.
2. ⇒ 3.: Suppose that there exists Q ∈ Q such that QT T ∗(I − Q)∗ = 0. Defining T1 = QT and
T2 = (I − Q)T , it follows that T = T1 + T2, R(T1) = R(Q) is uniformly J-positive, R(T2) = N (Q) is
uniformly J-negative and
T1T ∗2 = T2T ∗1 = 0,
If there exist closed range operators T1, T2 ∈ L(ℓ2(I),H) satisfying the conditions of item 3.,
Consider the projection Q = PR(T1)//R(T2) ∈ Q and notice that QT = T1 and (I − Q)T = T2. If
i = T ui.
because Eq. (25) says that R(T ∗2 ) = R(T ∗(I − Q)∗) ⊆ N (QT ) = N (T1).
3. ⇒ 1.:
notice that T1T ∗2 = 0 implies that N (T2)⊥ ⊆ N (T1), or equivalently, N (T1)⊥ ⊆ N (T2).
B1 = {ui}i∈I1 is an orthonormal basis of N (T1)⊥, consider the family {f +
But, if i ∈ I1,
f +
i = QT ui + (I − Q)T ui = T1ui ∈ R(T1),
i }i∈I1 in H given by f +
because ui ∈ N (T1)⊥ ⊆ N (T2). Therefore, {f +
i }i∈I1 ⊆ R(T1). Since T1 is an isomorphism between
N (T1)⊥ and R(T1), it follows that R(T1) = span{f−i }i∈I1 .
Analogously, if B2 = {bi}i∈I2 is an orthonormal basis of N (T1), the family {f−i }i∈I2 defined by
f−i = T bi (i ∈ I2) lies in R(T2). Since T2 is an isomorphism between N (T2)⊥ and R(T2), it follows that
R(T2) = T2(N (T1)) ⊆ span{f−i }i∈I2 ⊆ R(T2).
13
Finally, consider U ∈ U(ℓ2(I)) which turns the standard orthonormal basis {ei}i∈I into B1 ∪ B2. Then,
if V = T U and F = {V ei}i∈I = {f +
i }i∈I1 ∪ {f−i }i∈I2, it is easy to see that
I+ = {i ∈ I : [ V ei, V ei ] > 0} = I1
and I− = {i ∈ I : [ V ei, V ei ] < 0} = I2.
So, R(V+) = R(T1) is maximal uniformly J-positive and R(V−) = R(T2) is maximal uniformly J-negative.
Therefore, F is a J-frame for H with synthesis operator V = T U .
5 The J -frame operator
Definition. Given a J-frame F = {fi}i∈I , the J-frame operator S : H → H is defined by
Sf =Xi∈I
σi[ f, fi ]fi,
for every f ∈ H,
where σi = sgn([ fi, fi ]).
The following proposition compiles some basic properties of the J-frame operator.
Proposition 5.1. Let F = {fi}i∈I be a J-frame with synthesis operator T ∈ L(ℓ2(I),H). Then, its
J-frame operator S ∈ L(H) satisfies:
1. S = T T #;
2. S = S+ − S−, where S+ := T+T #
3. S is an invertible J-selfadjoint operator;
+ and S− := −T−T #
− are J-positive operators;
4. ind±(S) = dimH±, where ind±(S) are the indices of S.
Proof. Proof If F = {fi}i∈I is a J-frame with synthesis operator T ∈ L(ℓ2(I),H), then T #f =
Pi∈I σi[ f, fi ]ei for f ∈ H. So,
T T #f = T Xi∈I
σi[ f, fi ]ei! =Xi∈I
for every f ∈ H.
σi[ f, fi ]fi = Sf,
Furthermore, if I± = {i ∈ I : ±[ fi, fi ] > 0}, consider T± = T P± as usual. Then,
because T+T #
operator because
− = T−T #
T T # = (T+ + T−)(T+ + T−)# = T+T #
− = T+T #
+ = 0. Therefore, S = S+ − S− if S± := ±T±T #
±J = T±T ∗
±J.
± = ±T±J2T ∗
S± = ±T±T #
+ + T−T #
+ − (−T−T #
− ),
± . Notice that S± is a J-positive
+ ∔ M[⊥]
+ ) + S+(M[⊥]
± ⊆ N (S±). So, S(M[⊥]
− . But it is easy to see that M[⊥]
± ) = S∓(M[⊥]
− ) = R(S−) + R(S+) = M+ + M− = H. Therefore, S is invertible.
To prove the invertibility of S observe that, if Sf = 0 then S+f = S−f . But R(S+) ∩ R(S−) ⊆
− ) because
± ) and R(S) =
R(T+) ∩ R(T−) = {0}. Thus, S is injective. On the other hand, R(S) = S(M[⊥]
H = M[⊥]
S−(M[⊥]
Finally, the identities ind±(S) = dimH± follow from the indices definition. Recall that if A ∈ L(H)
is a J-selfadjoint operator, ind+(A) is the supremum of all positive integers r such that there exists a
positive invertible matrix of the form ([ Axj , xk ])j,k=1,...,r, where x1, . . . , xr ∈ H (if no such r exists,
ind−(A) = 0). Similarly, ind−(A) = ind+(−A) is the supremum of all positive integers m such that there
exists a negative invertible matrix of the form ([ Ayj, yk ])j,k=1,...,m, where y1, . . . , ym ∈ H, see [13].
Corollary 5.2. Let F = {fi}i∈I be a J-frame for H with J-frame operator S ∈ L(H). Then, R(S±) =
M± and N (S±) = M[⊥]
± . Furthermore, if Q = PM+//M−,
+ ) + S(M[⊥]
S+ = QSQ# and S− = −(I − Q)S(I − Q)#.
(26)
14
Proof. Proof Recall that S+ := T+T #
+ = T+(J2T ∗+J) = T+T ∗+J. Then, R(S+) = R(T+T ∗+J) =
R(T+T ∗+) = R(T+) = M+ because R(T+) is closed. Since S+ is J-selfadjoint, it follows that N (S+) =
R(S+)[⊥] = M[⊥]
+ . Analogously, R(S−) = M− and N (S−) = M[⊥]
− .
Since S = S+ − S−, if Q = PM+//M− then
QS = Q(S+ − S−) = S+,
by the characterization of the range and nullspace of S+. Therefore, SQ# = QS = QSQ#. Analogously,
S(I − Q)# = (I − Q)S = (I − Q)S(I − Q)#.
The above corollary states that S is the diagonal block operator matrix
S = (cid:18) S+
according to the (oblique) decompositions H = M[⊥]
−
0 −S− (cid:19) ,
∔ M[⊥]
0
codomain of S, respectively.
(27)
+ and H = M+ ∔ M− of the domain and
5.1 The indefinite reconstruction formula associated to a J -frame
Given a J-frame F = {fi}i∈I with synthesis operator T , there is a duality between F and the frame
G = {gi}i∈I given by gi = S−1fi: if f ∈ H,
f = SS−1f = T T #(S−1f ) = T Xi∈I
σi[ S−1f, fi ]ei! =Xi∈I
σi[ S−1f, fi ]fi =Xi∈I
σi[ f, S−1fi ]fi.
Analogously,
f = S−1Sf = S−1(T T #f ) = S−1 Xi∈I
σi[ f, fi ]fi! =Xi∈I
σi[ f, fi ]S−1fi.
Therefore, for every f ∈ H, there is an indefinite reconstruction formula associated to F :
f =Xi∈I
σi[ f, gi ]fi =Xi∈I
σi[ f, fi ]gi.
(28)
The following question arises naturally: is G = {S−1fi}i∈I also a J-frame for H?
Proposition 5.3. If F = {fi}i∈I is a J-frame for a Krein space H with J-frame operator S, then
G = {S−1fi}i∈I is also a J-frame for H.
Proof. Proof Given a J-frame F = {fi}i∈I for H with J-frame operator S, observe that the synthesis
operator of G = {S−1fi}i∈I is V := S−1T ∈ L(ℓ2(I),H). Furthermore, by Corollary 5.2, S(M[⊥]
∓ ) = M±.
Then, S−1(M±) = M[⊥]
∓ and it follows that
[ S−1fi, S−1fi ] > 0
if and only if
[ fi, fi ] > 0.
Thus, V± = V P± = S−1T± and R(V+) (resp. R(V−)) is a maximal uniformly J-positive (resp. J-
negative) subspace of H. So, G is a J-frame for H.
If F = {fi}i∈I is a frame for a Hilbert space H with synthesis operator T ∈ L(ℓ2(I),H), then the
family {(T T ∗)−1fi}i∈I is called the canonical dual frame because it is a dual frame for F (see Eq. (10))
and it has the following optimal property: Given f ∈ H,
Xi∈I
(cid:10) f, (T T ∗)−1fi(cid:11)2 ≤Xi∈I
ci2, whenever
cifi,
f =Xi∈I
(29)
for a family (ci)i∈I ∈ ℓ2(I). In other words, the above representation has the smallest ℓ2-norm among
the admissible frame coefficients representing f (see [12]).
15
Remark 5.4. If F = {fi}i∈I is a J-frame for H then F± = {fi}i∈I± is a frame for the Hilbert space
(M±,±[ , ]). Furthermore, the frame operator associated to F+ is S+ = T+T #
+ and its canonical dual
+ fi}i∈I+ . Analogously, the frame operator associated to F− is S− = −T−T #
frame is given by G+ = {S−1
−
and its canonical dual frame is given by G− = {S−1
Then, since H = M+ ∔M−, H can be seen as the (outer) direct sum of the Hilbert spaces (M+, [ , ])
and (M−,−[ , ]), i.e. the inner product given by
− fi}i∈I−.
h f, g iF
= [ f+, g+ ] − [ f−, g− ],
) into a Hilbert space and the projection Q = PM+//M− is selfadjoint in this Hilbert
f+, g+ ∈ M+, f−, g− ∈ M−,
f = f+ + f−, g = g+ + g−,
[ (I − Q)f, S−1
+ fi ]2 + Xi∈I−
[ Qf, S−1
i 2 + Xi∈I−
c+
[ f, S−1fi ]2 = Xi∈I+
≤ Xi∈I+
i fi and f− = (I − Q)f = Pi∈I− c−i fi, for families (c±i )i∈I± ∈ ℓ2(I±).
− fi ]2 ≤
c−i 2,
c+
whenever f+ = Qf = Pi∈I+
Therefore,
turns (H,h , iF
space. So, if f ∈ H,
Xi∈I
Xi∈I
[ f, S−1fi ]2 ≤Xi∈I
ci2,
whenever f = Pi∈I cifi for some (ci)i∈I ∈ ℓ2(I). In other words, the J-frame G = {S−1fi}i∈I is the
canonical dual frame of F in the Hilbert space (H,h , iF
).
5.2 Characterizing the J -frame operators
In a Hilbert space H, it is well known that every positive invertible operator S ∈ L(H) can be realized
as the frame operator of a frame F = {fi}i∈I for H, see [16]. Indeed, if B = {xi}i∈I is an orthonormal
basis of H, consider T : ℓ2(I) → H given by T ei = S1/2xi for i ∈ I. Then, for every f ∈ H,
T T ∗f =Xi∈I D f, S1/2xiE S1/2xi = S1/2 Xi∈I D S1/2f, xiE xi! = Sf
Therefore, F = {S1/2xi}i∈I is a frame for H and its frame operator is given by S.
The following paragraphs are devoted to characterize the set of J-frame operators.
Theorem 5.5. Let S ∈ GL(H) be a J-selfadjoint operator acting on a Krein space H with fundamental
symmetry J. Then, the following conditions are equivalent:
1. S is a J-frame operator, i.e. there exists a J-frame F with synthesis operator T such that S = T T #.
2. There exists a projection Q ∈ Q such that QS is J-positive and (I − Q)S is J-negative.
3. There exist J-positive operators S1, S2 ∈ L(H) such that S = S1 − S2 and R(S1) (resp. R(S2)) is
a uniformly J-positive (resp. J-negative) subspace of H.
Proof. Proof 1. → 2. follows from Proposition 5.1 and Corollary 5.2.
2. → 3.: If there exists a projection Q ∈ Q such that QS is J-positive and (I − Q)S is J-negative,
consider the J-positive operators S1 = QS and S2 = −(I − Q)S. Then, S = S1 − S2 and, by hypothesis,
R(S1) = R(Q) is uniformly J-positive and R(S2) = R(I − Q) = N (Q) is uniformly J-negative.
3. → 1.: Suppose that there exist J-positive operators S1, S2 ∈ L(H) such that S = S1 − S2 and
R(S1) (resp. R(S2)) is a uniformly J-positive (resp. J-negative) subspace of H. Denoting Kj = R(Sj)
for j = 1, 2, observe that Aj = SjJKj ∈ GL(Kj)+. Therefore, there exists a frame Fj = {fi}i∈Ij ⊂ Kj
for Kj such that Aj = TjT ∗j if Tj ∈ L(ℓ2(I1),Kj ) is the synthesis operator of Fj, for j = 1, 2.
16
Then, consider ℓ2(I) := ℓ2(I1) ⊕ ℓ2(I2) and T ∈ L(ℓ2(I),H) given by
T x = T1x1 + T2x2,
if x ∈ ℓ2(I), x = x1 + x2, xj ∈ ℓ2(Ij ) for j = 1, 2.
It is easy to see that T is the synthesis operator of the frame F = F1 ∪ F2. Furthermore F is a J-frame
such that I+ = I1 and I− = I2.
Finally, endow ℓ2(I) with the indefinite inner product defined by the diagonal operator J2 ∈ L(ℓ2(I))
given by
J2 ei = σi ei,
where σi = 1 if i ∈ I1 and σi = −1 if i ∈ I2. Notice that T1J2 = T1 and T2J2 = −T2. Furthermore,
T1T ∗2 = T2T ∗1 = 0 because R(T ∗2 ) = N (T2)⊥ ⊆ ℓ2(I1)⊥ = ℓ2(I2) ⊆ N (T1). Thus,
T T # = T J2T ∗J = (T1 + T2)(T ∗1 − T ∗2 )J = T1T ∗1 J − T2T ∗2 J = A1J − A2J = S1 − S2 = S.
that
Given a J-frame F = {fi}i∈I for H with J-frame operator S ∈ L(H), it follows from Corollary 5.2
(30)
S(M[⊥]
− ) = M+ and S(M[⊥]
+ ) = M−.
i.e. S maps a maximal uniformly J-positive (resp. J-negative) subspace into another maximal uniformly
J-positive (resp. J-negative) subspace. The next proposition shows under which hypotheses the converse
holds.
Proposition 5.6. Let S ∈ GL(H) be a J-selfadjoint operator. Then, S is a J-frame operator if and only
if the following conditions hold:
1. there exists a maximal uniformly J-positive subspace T of H such that S(T ) is also maximal uni-
formly J-positive;
2. [ Sf, f ] ≥ 0 for every f ∈ T ;
3. [ Sg, g ] ≤ 0 for every g ∈ S(T )[⊥].
Proof. Proof If S is a J-frame operator, consider T = M[⊥]
subspace T of H. Then, S(T ) = M+ is also maximal uniformly J-positive. Furthermore,
[ Sf, f ] = [ SQ#f, Q#f ] = [ QSQ#f, f ] = [ S+f, f ] ≥ 0 for every f ∈ T ,
− which is a maximal uniformly J-positive
+ = N (Q#) = R((I − Q)#). So,
where Q = PM+//M−. Also, S(T )[⊥] = M[⊥]
[ Sg, g ] = [ S(I − Q)#g, (I − Q)#g ] = [ (I − Q)S(I − Q)#g, g ] = [−S−g, g ] ≤ 0
for every g ∈ S(T )[⊥].
Conversely, suppose that there exists a maximal uniformly J-positive subspace T satisfying the hy-
pothesis. Let M = S(T ), which is maximal uniformly J-positive. Then, consider Q = PM//T [⊥]. It is
well defined because T [⊥] is maximal uniformly J-negative, see [2, Corollary 1.5.2]. Moreover, Q ∈ Q.
Notice that R(S(I − Q)#) = S(M[⊥]) = S(S(T )[⊥]) = S(S−1(T [⊥])) = T [⊥]. Therefore, QS(I −
Q)# = 0 and
QS = QSQ# + QS(I − Q)# = QSQ#.
Furthermore, if [ Sf, f ] ≥ 0 for every f ∈ T then QS is J-positive. Analogously, if [ Sg, g ] ≤ 0 for every
g ∈ S(T )[⊥] then (I − Q)S is J-negative. Then, by Theorem 5.5, S is a J-frame operator.
As it was proved in Proposition 5.1, if an operator S ∈ L(H) is a J-frame operator then it is an
invertible J-selfadjoint operator satisfying ind±(S) = dim(H±). Unfortunatelly, the converse is not true.
Example 4. Consider the Krein space obtained by endowing C2 with the sesquilinear form
and the invertible J-selfadjoint operator S, whose matrix in the standard orthonormal basis is given by
[(x1, x2), (y1, y2)] = x1y1 − x2y2,
Then, S satisfies ind±(S) = dim(H±), but it maps each J-positive vector into a J-negative vector. Then,
by Proposition 5.6, S cannot be a J-frame operator.
S = (cid:18) 0
i
i
0 (cid:19) .
17
6 Final remarks
The following are some simple consequences of the material studied in the previous sections. Nevertheless,
they are not going to be thoroughly developed in this notes.
Synthesis operators of J -frames as sums of plus and minus operators
If F = {fi}i∈I is a J-frame for the Krein space (H, [ , ]), it is easy to see that T+ and T #
+ are plus operators
(considering ℓ2(I) as a Krein space with the fundamental symmetry J2 defined in (12)). Furthermore,
T #
+ is strict, and, T+ is a strict plus operator if and only if N (T ) ∩ ℓ2(I+) = {0}.
analogously that T− and T #
if and only if N (T ) ∩ ℓ2(I−) = {0} (see [18, Ch. 2] for the terminology).
− . Indeed, it follows
− is always strict, and, T− is a strict minus operator
Also, these conditions have a natural counterpart for the operators T− and T #
− are minus operators; T #
Frames for regular subspaces of a Krein space
Given a Krein space (H, [ , ]), recall that a subspace S of H is regular if there exists a (unique) J-
selfadjoint projection onto S. Since a regular subspace S, endowed with the restriction of the indefinite
inner product [ , ] to S, is a Krein space (see [18, Ch. 1,Theorem 7.16]) the definition of J-frames applies
for regular subspaces of H too. Therefore, it is easy to infer a notion of "J-frames for regular subspaces"
of a Krein space.
References
[1] N. I. Akhiezer and I. M. Glazman, Theory of Linear operators in Hilbert space, Dover Publ.
Inc., 1993.
[2] T. Ando, Linear operators on Krein spaces, Hokkaido University, Sapporo, Japan, 1979.
[3] B.G. Bodmann and V.I. Paulsen, Frames, graphs and erasures, Linear Algebra Appl. 404
(2005) 118-146.
[4] J. Bogn´ar, Indefinite Inner Product Spaces, Springer-Verlag, 1974.
[5] A. Buades, B. Coll and J.-M. Morel, Nonlocal image and movie denoising, Int. J. Comput.
Vis. 76 (2008), no. 2, 123 -- 140.
[6] P. G. Casazza, The art of frame theory, Taiwanese J. Math. 4 (2000), no. 2, 129-201.
[7] P. G. Casazza and O. Christensen, Frames containing a Riesz basis and preservation of this
property under perturbations, SIAM J. Math. Anal. 29 (1998), no. 1, 266-278.
[8] O. Christensen, An introduction to frames and Riesz bases, Applied and Numerical Harmonic
Analysis, Birkhauser, Boston, 2003.
[9] O. Christensen, Recent Developments in Frame Theory, Modern Mathematical Models,
Methods and Algorithms for Real World Systems. A.H. Siddiqi, I.S. Duff and O. Chris-
tensen (Editors), 2006, Anamaya Publishers, New Delhi, India.
[10] F. Deutsch, The angle between subspaces of a Hilbert space, Approximation theory, wavelets
and applications (Maratea, 1994), 107 -- 130, NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci.,
454, Kluwer Acad. Publ., Dordrecht, 1995.
[11] R. G. Douglas, On majorization, factorization and range inclusion of operators in Hilbert
space, Proc. Amer. Math. Soc. 17 (1966), 413-416.
[12] R. J. Duffin and A. C. Schaeffer, A class of nonharmonic Fourier series, Trans. Amer. Math.
Soc. 72 (1952), 341-366.
18
[13] M. A. Dritschel and J. Rovnyak, Operators on indefinite inner product spaces, Fields Insti-
tute Monographs no. 3, Amer. Math. Soc. Edited by Peter Lancaster 1996, 3, 141-232.
[14] K. Esmeral Garc´ıa and E. Wagner, Frames in Krein spaces, preprint.
[15] A. Grod, S. Kuzhel and V. Sudilovskaya, On operators of transition in Krein spaces, Opuscula
Math. 31 (2010), no. 1, 49-59.
[16] D. Han and D. R. Larson, Frames, bases and group representations, Mem. Amer. Math. Soc.
147 (2000), no. 697.
[17] R.B. Holmes and V.I. Paulsen, Optimal frames for erasures, Linear Algebra Appl. 377 (2004),
31-51.
[18] I. S. Iokhvidov and T. Ya. Azizov, Linear Operators in spaces with an indefinite metric, John
Wiley and sons, 1989.
[19] T. Kato, Perturbation theory for linear operators, Springer, New York, 1966.
[20] M.A.S. Masoum, S. Jamali and N. Ghaffarzadeh, Detection and classification of power quality
disturbances using discrete wavelet transform and wavelet networks, IET Sci. Meas. Technol.
4 (2010), 193 -- 205.
[21] I. Peng and S. Waldron, Signed frames and Hadamard products of Gram matrices, Linear
Algebra Appl. 347 (2002), no. 1-3, 131-157.
[22] T. Strohmer and R.W Heath Jr., Grassmannian frames with applications to coding and
communication, Appl. Comput. Harmon. Anal. 14, (2003) 257-275.
Juan I. Giribet
[email protected]
Departamento de Matem´atica, FI-UBA, Buenos Aires, Argentina
and
IAM-CONICET.
Alejandra Maestripieri
[email protected]
Departamento de Matem´atica, FI-UBA, Buenos Aires, Argentina
and
IAM-CONICET.
Francisco Mart´ınez Per´ıa
[email protected]
Departamento de Matem´atica, FCE-UNLP, La Plata, Argentina
and
IAM-CONICET.
Pedro G. Massey
[email protected]
Departamento de Matem´atica, FCE-UNLP, La Plata, Argentina
and
IAM-CONICET, Saavedra 15, Piso 3, (1083) Buenos Aires, Argentina.
19
|
1902.09206 | 1 | 1902 | 2019-02-25T11:45:55 | Extended Gevrey regularity via the short-time Fourier transform | [
"math.FA"
] | We study the regularity of smooth functions whose derivatives are dominated by sequences of the form $M_p^{\tau,\s}=p^{\tau p^{\s}}$, $\tau>0$, $\s\geq1$. We show that such functions can be characterized through the decay properties of their short-time Fourier transforms (STFT), and recover \cite[Theorem 3.1]{CNR} as the special case when $ \t>1$ and $\s = 1$, i.e. when the Gevrey type regularity is considered. These estimates lead to a Paley-Wiener type theorem for extended Gevrey classes. In contrast to the related result from \cite{PTT-05, PTT-04}, here we relax the assumption on compact support of the observed functions. Moreover, we introduce the corresponding wave front set, recover it in terms of the STFT, and discuss local regularity in such context. | math.FA | math |
EXTENDED GEVREY REGULARITY VIA THE
SHORT-TIME FOURIER TRANSFORM
NENAD TEOFANOV, FILIP TOMI ´C
σ
Abstract. We study the regularity of smooth functions whose
derivatives are dominated by sequences of the form M τ,σ
,
p = pτ p
τ > 0, σ ≥ 1. We show that such functions can be characterized
through the decay properties of their short-time Fourier transforms
(STFT), and recover [5, Theorem 3.1] as the special case when τ >
1 and σ = 1, i.e. when the Gevrey type regularity is considered.
These estimates lead to a Paley-Wiener type theorem for extended
Gevrey classes.
In contrast to the related result from [24, 25],
here we relax the assumption on compact support of the observed
functions. Moreover, we introduce the corresponding wave front
set, recover it in terms of the STFT, and discuss local regularity
in such context.
1. Introduction
Classes of extended Gevrey functions and the corresponding wave
front sets are introduced and investigated in [21, 22, 23, 32]. Such
classes consist of smooth function, and they are larger than any Gevrey
class. This turned out to be important e.g.
in the study of strictly
hyperbolic equations, see [1]. Paley-Wiener type theorem for compactly
supported extended Gevrey regular functions is given in [24, 25], and
it turns out that the Fourier-Laplace transform of such functions have
certain logarithmic decay at infinity which can be expressed in terms
of Lambert W function. This fact is used to resolve the wave front sets
in the context of extended Gevrey regularity. We refer to [22, 23] for
related theorems on propagation of singularities.
The aim of this paper is twofold. Firstly, we give another version of
the Paley-Wiener theorem for extended Gevrey regularity and formu-
late the result in terms of the short time Fourier transform (STFT)
(cf. [12]). More precisely, we prove a generalization of [5, Theorem 3.1],
where the STFT estimates are related to Gevrey type regularity, and
2010 Mathematics Subject Classification. Primary 46E10, 35A18, Secondary
46F05, 42B10.
Key words and phrases. Gevrey classes, Paley-Wiener theorem, modulation
spaces, Wave front sets, ultradistributions.
1
2
NENAD TEOFANOV, FILIP TOMI ´C
obtain the Paley-Wiener type result as its corollary. Secondly, we give
a description of (micro)local regularity related to the extended Gevrey
regularity by the means of the STFT. This result is inspired by recent
characterization of the C ∞ wave front sets via the STFT, given in [19].
The paper is organized as follows: In subsection 1.1 we fix some
notation and in Section 2 we collect the main notions and tools for
our analysis: Subsection 2.1 contains basic facts concerning the ex-
tended Gevrey classes. In Subsection 2.2 we introduce the notion of
extended associated function which appears in the formulation of our
main results. The correct asymptotic behavior of the extended asso-
ciated function is given by the means of the Lambert W function, see
Theorem 2.1.
In subsection 2.3 we introduce the short-time Fourier
transform and modulation spaces defined by the means of decay and
integrability conditions of the STFT of ultradistributions. We also
recall some basic properties of modulation spaces.
In Section 3 we prove Theorem 3.1.
It is a generalization of [5,
Theorem 3.1] which turned out to be important for the properties
of pseudodifferential operators with symbols of Gevrey, analytic and
ultra-analytic regularity, see [5] for details. As a corrolary of Theorem
3.1 we obtain a Paley-Wiener type theorem for element of modulation
spaces related to the extended Gevrey classes. This result extends [25,
Theorem 3.1] in the sense that the condition on compact support is
replaced by appropriate decay property given by a modulation space
norm, when the Fourier-Laplace transform is replaced by the STFT.
In Section 4 we recall the notion of wave front sets related to ex-
tended Gevrey regularity. We prove that such wave front sets can be
characterized by the decay properties of the STFT of a distribution
with respect to a suitably chosen window function, Theorem 4.1. As
a consequence we derive a result on local extended Gevrey regularity,
Theorem 4.1.
Our results are proved for the so-called Roumieu case, and we note
that proofs for the Beurling case are similar and therefore omitted.
1.1. Basic notions and notation. We denote by N, Z+, R, C the
sets of nonnegative integers, positive integers, real numbers and com-
plex numbers, respectively. For x ∈ Rd we put hxi = (1 +x2)1/2. The
integer parts (the floor and the ceiling functions) of x ∈ R+ are denoted
by ⌊x⌋ := max{m ∈ N : m ≤ x} and ⌈x⌉ := min{m ∈ N : m ≥ x}.
For a multi-index α = (α1, . . . , αd) ∈ Nd we write ∂α = ∂α1 . . . ∂αd,
Dα = (−i)α∂α, and α = α1 + . . .αd. Open ball of radius r > 0
centered at x0 is denoted by Br(x0). As usual, C ∞(Rd) is the space of
smooth functions, the Schwartz space of rapidly decreasing functions
EXTENDED GEVREY REGULARITY ...
3
is denoted by S(Rd), and S ′(Rd) denotes its dual space of tempered
distributions. Lebesgue spaces over an open set Ω ⊆ Rd are denoted
by Lp(Ω), 1 ≤ p < ∞, and the norm of f ∈ Lp(Ω) is denoted by kfkLp.
The Fourier transform is normalized to be
f (ω) = F f (ω) =Z f (t)e−2πitωdt.
hf, gi = R f (t)g(t)dt on L2(Rd) to the dual pairing between a test
We use the brackets hf, gi to denote the extension of the inner product
function space A and its dual A′: h·,·i = A′h·,·iA.
acting on f ∈ L2(Rd) are defined by
Translation and modulation operators, T and M respectively, when
Txf (·) = f (· − x)
and Mxf (·) = e2πix·f (·), x ∈ Rd.
Then for f, g ∈ L2(Rd) the following relations hold:
MyTx = e2πix·yTxMy, (Txf )= M−x f , (Mxf )= Tx f , x, y ∈ Rd.
These operators are extended to other spaces of functions and distri-
butions in a natural way.
Throughout the paper, A . B denotes A ≤ cB for a suitable con-
stant c > 0, whereas A ≍ B means that c−1A ≤ B ≤ cA for some c ≥ 1.
The symbol B1 ֒→ B2 denotes the continuous and dense embedding of
the topological vector space B1 into B2.
2. Preliminaries
In this section we collect the main tools and auxiliary results which
will be used in the sequel. More precisely, we introduce the test function
p = pτ pσ, p ∈ Z+, for a
spaces related to the sequences of the form M τ,σ
given τ > 0 and σ > 1. Notice that, when τ > 1 and σ = 1 M τ,1
p =
pτ p, p ∈ Z+ is (equivalent to) the Gevrey sequence. Then we discuss
associated functions to such sequences, which are the main tool of our
analysis. To describe precise asymptotic behavior of those associated
functions at infinity appears to be a nontrivial problem, which can be
resolved by the use of Lambert's W functions. Finally, we recall the
definition and some elementary properties of the STFT and modulation
spaces defined by mixed weighted Lebesgue norm conditions on the
STFT.
2.1. Extended Gevrey regularity. In this section we introduce ex-
tended Gevrey classes and discuss their basic properties. We employ
Komatsu's approach [17] to spaces of ultradifferentiable functions, and
p = pτ pσ, p ∈ N, depending
consider defining sequences of the form M τ,σ
on parameters τ > 0 and σ > 1, [22].
4
NENAD TEOFANOV, FILIP TOMI ´C
Essential properties of the defining sequences are listed in the fol-
lowing lemma. We refer to [21] for the proof. In the general theory
of ultradistributions (see [17] different properties of defining sequences
give rise to particular structural properties of the corresponding spaces
of ultradifferentiable functions, see [26] for a detalied survey.
p = pτ pσ , p ∈ Z+, M τ,σ
Lemma 2.1. Let τ > 0, σ > 1 and M τ,σ
0 = 1.
Then there exists an increasing sequence of positive numbers Cq, q ∈ N,
and a constant C > 0 such that:
(M.1) (M τ,σ
p
(M.2) M τ,σ
(M.2)′ M τ,σ
p+1, p ∈ Z+
)2 ≤ M τ,σ
, p, q ∈ N,
p−1M τ,σ
M τ 2σ−1,σ
q
p
q M τ,σ
p+q ≤ C pσ+qσ M τ 2σ−1,σ
p+q ≤ C pσ
∞Xp=1
M τ,σ
p−1
M τ,σ
, p, q ∈ N,
< ∞. Moreover,
p
p
(M.3)′
M τ,σ
p−1
M τ,σ
p ≤
1
(2p)τ (p−1)σ−1 , p ∈ N.
Let τ, h > 0, σ > 1 and let K ⊂⊂ Rd be a regular compact set. By
Eτ,σ,h(K) we denote the Banach space of functions φ ∈ C ∞(K) such
that
kφkEτ,σ,h(K) = sup
α∈Nd
sup
x∈K
∂αφ(x)
hασM τ,σ
α
< ∞.
(2.1)
The set of functions φ ∈ Eτ,σ,h(K) whose support is contained in K
is denoted by DK
Let U be an open set Rd and K ⊂⊂ U. We define families of spaces
by introducing the following projective and inductive limit topologies:
τ,σ,h.
lim−→h→∞Eτ,σ,h(K),
lim←−h→0Eτ,σ,h(K),
E{τ,σ}(U) = lim←−K⊂⊂U
E(τ,σ)(U) = lim←−K⊂⊂U
D{τ,σ}(U) = lim−→K⊂⊂U DK
D(τ,σ)(U) = lim−→K⊂⊂U DK
τ,σ,h).
τ,σ,h) ,
{τ,σ} = lim−→K⊂⊂U
(τ,σ) = lim−→K⊂⊂U
( lim−→h→∞DK
(lim←−h→0DK
We will use abbreviated notation τ, σ for {τ, σ} (the Roumieu case) or
(τ, σ) (the Beurling case) . The spaces Eτ,σ(U), DK
τ,σ and Dτ,σ(U) are
nuclear, cf. [21]. We refer to [21, 22, 23, 31, 32, 35] for other properties
of those spaces.
Remark 2.1. If τ > 1 and σ = 1, then E{τ,1}(U) = E{τ }(U) is the Gevrey
class, and D{τ,1}(U) = D{τ }(U) is its subspace of compactly supported
functions in E{τ }(U).
EXTENDED GEVREY REGULARITY ...
5
In particular,
lim−→t→∞E{t}(U) ֒→ Eτ,σ(U) ֒→ C ∞(U),
τ > 0, σ > 1,
so that the regularity in Eτ,σ(U) can be thought of as an extended
Gevrey regularity.
In
particular, Dτ,1(U) = {0} when 0 < τ ≤ 1, and E{1,1}(U) = E{1}(U) is
the space of analytic functions on U.
If 0 < τ ≤ 1, then Eτ,1(U) consists of quasianalytic functions.
0 ≤ φ ≤ 1 andRRd φ dx = 1, see [21] for a construction of a compactly
The non-quasianalyticity condition (M.3)′ provides the existence of
partitions of unity in E{τ,σ}(U), i.e.
for any given τ > 0 and σ > 1,
there exists a compactly supported function φ ∈ E{τ,σ}(U) such that
supported φ ∈ D{τ,σ}(U) \ D{t}(U), t > 1.
Note that the additional exponent σ, which appears in the power
of term h in (2.1), makes the definition of Eτ,σ(U) different from the
definition of Carleman classes, cf.
[16]. This difference appears to be
essential in many calculations, and in particular when dealing with the
operators of "infinite order", cf. [22].
2.2. The associated function to the sequence M τ,σ
. In
this subsection we recall the definition and asymptotic proeprties of
p = pτ pσ , p ∈ N,
extended associated function to the sequence M τ,σ
τ > 0, σ > 1, cf. [25]. We also recall the Paley-Wiener theorem related
to the extended Gevrey regularity.
p = pτ pσ
Definition 2.1. Let τ > 0, σ > 1 and M τ,σ
The extended associated function related to the sequence M τ,σ
by
p = pτ pσ, p ∈ Z+, M τ,σ
0 = 1.
, is given
p
hpσkp
M τ,σ
where ln+ A = max{0, ln A}, for A > 0.
Tτ,σ,h(k) = sup
p∈N
ln+
p
, h, k > 0,
Obviously Tτ,σ,h(k), τ, h > 0, σ > 1, is positive for sufficiently large
k > 0.
In fact, for any sequence of positive numbers Mp, p ∈ N, such that
is bounded from below and M0 = 1, its associated function is
p
M 1/p
defined to be
T (k) = sup
p∈N
ln
kp
Mp
,
k > 0.
Therefore, for τ > 0 and σ = 1, Tτ,1,h(k) := Tτ (hk) is the associated
function to the Gevrey sequence pτ p, p ∈ N (we may assume h = 1
without loosing generality).
6
NENAD TEOFANOV, FILIP TOMI ´C
It is well known (cf. [11, 27]) that
Ak1/τ − B ≤ Tτ (k) ≤ Ak1/τ ,
k > 0,
(2.2)
for suitable A, B > 0. In particular, the growth of eTτ (k) for τ > 1 is
subexponential.
Moreover, for any t, τ > 0 and σ > 1, by [25, Lemma 2.3] it follows
that there is a constant C > 0 such that
Tτ,σ,1(k) < Ck1/t,
k > 0.
Therefore the function eTτ,σ,h(k) has a less rapid growth at infinity than
any subexponential function.
The precise asymptotic behavior of Tτ,σ,h(k) at infinity is a chal-
lenging problem. We use an auxiliary special function to resolve that
problem.
The Lambert W function is defined as the inverse function of zez,
z ∈ C, wherefrom the following property holds:
x ≥ 0.
x = W (x)eW (x),
We denote its principal (real) branch by W (x), x ≥ 0 (see [6]). It is
a continuous, increasing and concave function on [0,∞), W (0) = 0,
W (e) = 1, and W (x) > 0, x > 0.
It can be shown that W can be represented in the form of the abso-
lutely convergent series
W (x) = ln x − ln(ln x) +
∞Xk=0
∞Xm=1
ckm
(ln(ln x))m
(ln x)k+m ,
x ≥ x0 > e,
with suitable constants ckm and x0, wherefrom the following estimates
hold:
ln x − ln(ln x) ≤ W (x) ≤ ln x −
1
2
ln(ln x),
x ≥ e.
(2.3)
The equality in (2.3) holds if and only if x = e. We refer to [15, 6] for
more details about the Lambert W function.
Theorem 2.1. ([25]) Let there be given τ, h > 0, σ > 1 and let Cτ,σ,h =
h− σ−1
τ e
σ−1
σ σ−1
τ σ . Then
σ−1(cid:16) σ − 1
expn(2σ−1τ )− 1
σ (cid:17) σ
τ σ (cid:17) 1
. expn(cid:16)σ − 1
σ−1
σ−1
W − 1
σ−1 (Cτ,σ,h ln k) ln
W − 1
σ−1 (Cτ,σ,h ln k) ln
σ
σ−1 ko . eTτ,σ,h(k)
k > e.
(2.4)
σ
σ−1 ko,
EXTENDED GEVREY REGULARITY ...
7
If, moreover 1 < σ < 2, then we have the precise asymptotic formula
τ σ (cid:17) 1
eTτ,σ,h(k) ≍ expn(cid:16)σ − 1
σ−1
W − 1
σ−1 (Cτ,σ,h ln k) ln
k > e.
σ
σ−1 ko,
The hidden constants in (2.4) and (2.1) depend on τ, σ and h.
Remark 2.2. Note that, in the view of (2.3) we have
W − 1
σ−1 (C ln k) ln
σ
σ−1 k ≍
ln
σ
σ−1 k
ln
σ−1 (C ln k) ≍
1
σ
σ−1 k
ln
ln
1
σ−1 (ln k)
,
k → ∞,
(2.5)
for any given σ > 1, and the last behavior follows from ln(C ln k) ≍
ln(ln k), k → ∞, for any given C > 0.
Since limk→∞(ln k)1/(σ−1)(ln(C ln k))−1/(σ−1) = ∞, for every C > 0,
(2.5) implies that for every M > 0 there exists B > 0 (depending on h
and M) such that
W − 1
σ−1 (C ln k)) ln
σ
σ−1 k > M ln k,
k > B.
Next we recall the Paley-Wiener theorem for DK
τ,σ when 1 < σ < 2.
For the proof we refer to [24], and a more general case when σ ≥ 2 is
proved in [25].
Theorem 2.1. Let τ > 0, 1 < σ < 2, U be open set in Rd and
K ⊂⊂ U. If ϕ ∈ DK
(τ,σ)) then its Fourier-Laplace
transform is an entire function and there exists constants A, B > 0
(resp. for every B > 0 there exists A > 0) such that
{τ,σ} (resp. ϕ ∈ DK
τ σ (cid:17) 1
bϕ(η) ≤ A expn−(cid:16) σ − 1
where HK(η) = sup
y∈K
Im(y · η).
σ−1
W − 1
σ−1(cid:16)B ln(e + η)(cid:17)ln
σ
σ−1 (e+η)+HK(η)o
h > 0, η ∈ Cd,
(2.6)
Conversely, if there exists A, B > 0 (resp.
for every B > 0 there
exists A > 0) such that an entire function bϕ(η) satisfies (2.6) then bϕ(η)
is the Fourier-Laplace transform of ϕ ∈ DK
The following corollary is an immediate consequence of Theorem 3.1
{2σ−1τ,σ} (resp. DK
(2σ−1τ,σ)).
and (2.5).
Corollary 2.1. Let 1 < σ < 2, U be open set in Rd and K ⊂⊂ U.
of
Then the entire function bϕ(η), η ∈ Cd, is the Fourier-Laplace transform
τ,σ)
ϕ ∈ lim−→τ →∞DK
τ,σ (resp. ϕ ∈ lim←−τ →0DK
8
NENAD TEOFANOV, FILIP TOMI ´C
if and only if there exist constant A, B > 0 (resp.
there exists A > 0) such that
for every B > 0
bϕ(η) ≤ A exp(−B
ln
Im(x · η).
where HK(η) = sup
x∈K
σ
1
σ−1 (e + η)
ln
σ−1 (ln(e + η))
+ HK(η)) , η ∈ Cd,
2.3. Modulation Spaces. The modulation spaces were initially (and
systematically) introduced in [7]. See also[12, Ch. 11-13] and the orig-
inal literature quoted there for various properties and applications of
the so called standard modulation spaces. It is usually sufficient to ob-
serve weighted modulation spaces with weights which may grow at most
polynomially at infinity. However, for the study of ultra-distributions
a more general approach which includes weights of exponential or even
superexponential growth is needed, cf. [4, 34]. We refer to [8, 9] for re-
lated but even more general constructions, based on the general theory
of coorbit spaces.
For our purposes it is sufficient to consider weights of exponential
growth. Therefore we begin with the Gelfand-Shilov space of analytic
functions S (1)(Rd) given by
f ∈ S (1)(Rd) ⇐⇒ sup
x∈Rd f (x)eh·x < ∞ and sup
ω∈Rd f (ω)eh·ω < ∞,
for every h > 0. Any f ∈ S (1)(Rd) can be extended to a holomorphic
function f (x + iy) in the strip {x + iy ∈ Cd :
y < T} some T > 0,
[11, 18]. The dual space of S (1)(Rd) will be denoted by S (1)′(Rd). In
fact, S (1)(Rd) is isomorphic to the Sato test function space for the space
of Fourier hyperfunctions S (1)′(Rd), see [2].
Let there be given f, g ∈ L2(Rd). The short-time Fourier transform
(STFT) of f with respect to the window g is given by
Vgf (x, ω) =Z e−2πitωf (t)g(t − x)dt, x, ω ∈ Rd.
(2.7)
It restricts to a mapping from S (1)(Rd) × S (1)(Rd) to S (1)(R2d), which
is proved in the next Lemma.
Lemma 2.1. Let f, g ∈ S (1)(Rd), and let the short-time Fourier trans-
form (STFT) of f with respect to g be given by (2.7). Then Vgf (x, ω) ∈
S (1)(R2d), that is Vgf (x, ω) < Ce−sk(x,ω)k, x, ω ∈ Rd, for every s > 0.
Proof. The proof is standard, see e.g. [12] for the proof in the context
of S(Rd). We use the arguments based on the structure of S (1)(Rd) as
follows. Let f ⊗ g be the tensor product f ⊗ g(x, t) = f (x) · g(t), let
EXTENDED GEVREY REGULARITY ...
9
T denote the asymmetric coordinate transform T F (x, t) = F (t, t− x),
and let F2 be the partial Fourier transform
F (x, t)e−2πitωdt,
x, ω ∈ Rd,
F2F (x, ω) =ZRd
of a function F on R2d. Then
Vgf (x, ω) = F2T (f ⊗ g)(x, ω),
(x, ω) ∈ R2d.
Since S (1)(R2d) ∼= S (1)(Rd) ⊗S (1)(Rd) (see e.g.
[30] for the kernel the-
orem in Gelfand-Shilov spaces) and since S (1)(R2d) is invariant under
the action of T and F2, we conclude that Vgf (x, ω) < Ce−sk(x,ω)k,
x, ω ∈ Rd, for every s > 0.
(cid:3)
Weight Functions.
In the sequel v will always be a continuous,
positive, even, submultiplicative function (submultiplicative weight),
i.e., v(0) = 1, v(z) = v(−z), and v(z1 + z2) ≤ v(z1)v(z2), for all
z, z1, z2 ∈ R2d. Moreover, v is assumed to be even in each group of co-
ordinates, that is, v(x, ω) = v(−ω, x) = v(−x, ω), for any (x, ω) ∈ R2d.
Submultipliciativity implies that v(z) is dominated by an exponential
function, i.e.
∃ C, k > 0 such that
(2.8)
and kzk is the Euclidean norm of z ∈ R2d. For example, every weight
of the form
v(z) ≤ Cekkzk,
z ∈ R2d,
v(z) = eskzkb
(1 + kzk)a logr(e + kzk)
for parameters a, r, s ≥ 0, 0 ≤ b ≤ 1 satisfies the above conditions.
Associated to every submultiplicative weight we consider the class of
so-called v-moderate weights Mv. A positive, even weight function m
on R2d belongs to Mv if it satisfies the condition
m(z1 + z2) ≤ Cv(z1)m(z2) ∀z1, z2 ∈ R2d .
We note that this definition implies that 1
v . m . v, m 6= 0 everywhere,
and that 1/m ∈ Mv.
The widest class of weights allowing to define modulation spaces is
the weight class N . A weight function m on R2d belongs to N if it is
a continuous, positive function such that
m(z) = o(ecz2
),
for z → ∞,
∀c > 0,
with z ∈ R2d. For instance, every function m(z) = eszb, with s > 0
and 0 ≤ b < 2, is in N . Thus, the weight m may grow faster than
exponentially at infinity. For example, the choice m ∈ N \ Mv is
related to the spaces of quasianalytic functions, [3]. We notice that
there is a limit in enlarging the weight class for modulation spaces,
10
NENAD TEOFANOV, FILIP TOMI ´C
imposed by Hardy's theorem: if m(z) ≥ Cecz2
, for some c > π/2, then
the corresponding modulation spaces are trivial [14]. We refer to [13]
for a survey on the most important types of weights commonly used in
time-frequency analysis.
Definition 2.1. Let v be a submultiplicative weight v, m ∈ Mv, and
let g be a non-zero window function in S (1)(Rd). For 1 ≤ p, q ≤ ∞
m (Rd) consists of all f ∈ S (1)′ (Rd) such that
the modulation space M p,q
Vgf ∈ Lp,q
m (Rd)
is
m (Rd) (weighted mixed-norm spaces). The norm on M p,q
kfkM p,q
m = kVgfkLp,q
m = ZRd(cid:18)ZRd Vgf (x, ω)pm(x, ω)p dx(cid:19)q/p
dω!1/q
(with obvious changes if either p = ∞ or q = ∞). If p, q < ∞, the
modulation space M p,q
m (Rd) is the norm completion of S (1)(Rd) in the
M p,q
m (Rd) is the completion of
S (1)(Rd) in the weak∗ topology.
m -norm. If p = ∞ or q = ∞, then M p,q
Note that for f, g ∈ S (1)(Rd) the above integral is convergent so that
m (Rd). Namely, in view of (2.8), for a given m ∈ Mv
S (1)(Rd) ⊂ M p,q
there exist l > 0 such that m(x, ω) ≤ Celk(x,ω)k and therefore
ZRd(cid:18)ZRd Vgf (x, ω)pm(x, ω)p dx(cid:19)q/p
≤ C(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ZRd(cid:18)ZRd Vgf (x, ω)pelpk(x,ω)k dx(cid:19)q/p
dω(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
< ∞,
If p = q, we write M p
we write M p,q and M p for M p,q
m instead of M p,p
m and M p,p
In the next proposition we show that M p,q
since by Lemma 2.1 it follows that Vgf (x, ω) < Ce−sk(x,ω)k for every
s > 0.
m , and if m(z) ≡ 1 on R2d, then
m , and so on.
m (Rd) are Banach spaces
whose definition is independent of the choice of the window g ∈ M 1
v (Rd)\
{0}. In order to do so, we need the adjoint of the short-time Fourier
transform.
m (R2d)
For a given window g ∈ S (1)(Rd) and a function F (x, ξ) ∈ Lp,q
we define V ∗
g F by
whenever the duality is well defined.
hV ∗
g F, fi := hF, Vgfi,
Then [12, Proposition 11.3.2] (see also [4]) can be rewritten as fol-
lows.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
dω(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
EXTENDED GEVREY REGULARITY ...
11
Proposition 2.1. Fix m ∈ Mv and g, ψ ∈ S (1), with hg, ψi 6= 0. Then
(1) V ∗
g : Lp,q
m (Rd), and
m ≤ CkVψgkL1
(2) The inversion formula holds: IM p,q
m (R2d) → M p,q
g FkM p,q
kV ∗
stands for the identity operator.
vkFkLp,q
m .
m = hg, ψi−1V ∗
g Vψ, where IM p,q
m
(3) M p,q
m (Rd) are Banach spaces whose definition is independent on
the choice of g ∈ S (1) \ {0}.
to M 1
v (Rd).
(4) The space of admissible windows can be extended from S (1)(Rd)
When m is a polynomial weight of the form m(x, ω) = hxithωis we
s,t (Rd) for the modulation spaces which consists
will use the notation M p,q
of all f ∈ (S (1))′(Rd) such that
s,t ≡ ZRd(cid:18)ZRd Vφf (x, ω)hxithωisp dx(cid:19)q/p
dω!1/q
< ∞
kfkM p,q
(with obvious interpretation of the integrals when p = ∞ or q = ∞).
The following theorem lists some basic properties of modulation
spaces. We refer to [7, 12, 20, 29, 33] for its proof.
Theorem 2.2. Let p, q, pj, qj ∈ [1,∞] and s, t, sj, tj ∈ R, j = 1, 2.
Then:
s,t (Rd) are Banach spaces, independent of the choice of φ ∈
(1) M p,q
S(Rd) \ 0;
(2) if p1 ≤ p2, q1 ≤ q2, s2 ≤ s1 and t2 ≤ t1, then
(3) ∩s,tM p,q
(4) Let 1 ≤ p, q ≤ ∞, and let ws(z) = ws(x, ω) = esk(x,ω)k, z =
s2,t2 (Rd) ⊆ S ′(Rd);
s,t (Rd) = S ′(Rd);
(x, ω) ∈ R2d. Then
S(Rd) ⊆ M p1,q1
s1,t1 (Rd) ⊆ M p2,q2
s,t (Rd) = S(Rd), ∪s,tM p,q
S (1)(Rd) =\s≥0
(Rd) =[s≥0
S (1)′
ws (Rd) = \m∈∩Mws
(Rd) = [m∈∩Mws
(5) For p, q ∈ [1,∞), the dual of M p,q
M p,q
1/ws
M p,q
1
p + 1
p′ = 1
q + 1
q′ = 1.
M p,q
m (Rd),
M p,q
1/m(Rd);
s,t (Rd) is M p′,q′
−s,−t(Rd), where
Modulation spaces include the following well-know function spaces:
(1) M 2(Rd) = L2(Rd), and M 2
t,0(Rd) = L2
t (Rd);
12
NENAD TEOFANOV, FILIP TOMI ´C
(2) The Feichtinger algebra: M 1(Rd) = S0(Rd);
(3) Sobolev spaces: M 2
(4) Shubin spaces: M 2
0,s(Rd) = H 2
s (Rd) = L2
s (Rd) = {f f (ω)hωis ∈ L2(Rd)};
s (Rd) = Qs(Rd), cf. [28].
s(Rd) ∩ H 2
3. Decay properties of the STFT
In this section we characterize certain regularity properties related
to the classes Eτ,σ, τ > 0, σ ≥ 1, by the rate of decay of the STFT. In
particular, we extend [5, Theorem 3.1], which is formulated in terms of
Gevrey sequences and the corresponding spaces of test functions.
In the proof of Theorem 3.1 in several occasions we will use the
following simple inequalities:
ασ + βσ ≤ α + βσ ≤ 2σ−1(ασ + βσ), α, β ∈ Nd, σ > 1, (3.1)
and
(1/√d)ξα ≤ ξα ≤ ξα, α ∈ Nd, ξ ∈ Rd.
(3.2)
Theorem 3.1. Let τ > 0, σ ≥ 1, let v be a submultiplicative weight,
m ∈ Mv, and let g ∈ M 1,1
k∂αgkL1
v⊗1(Rd)\{0} such that for some Cg > 0,
v(Rd) . C ασ
, α ∈ Nd.
(3.3)
g
ατ ασ
For a smooth function f the following conditions are equivalent:
i) There exists a constant Cf > 0 such that
ατ ασ
k∂αfkL∞(Rd) . m(x)C ασ
f
, α ∈ Nd;
(3.4)
ii) There exists a constant Cf,g > 0 such that
ξαVgf (x, ξ) . m(x)C ασ
f,g ατ ασ
, x, ξ ∈ Rd, α ∈ Nd;
iii) There exists a constant C > 0 such that
Vgf (x, ξ) . m(x)e−Tτ,σ,C (ξ), x, ξ ∈ Rd.
Proof. When σ = 1 we obtain [5, Theorem 3.1], where the function
Cx1/τ appears instead of Tτ,σ,C(ξ). However, this makes no differ-
ence, since from Tτ,1,C(ξ) := Tτ (Cξ) (see also (2.2)) it follows that
iii) is equivalent to
Vgf (x, ξ) . m(x)e−Cξ1/τ
,
x, ξ ∈ Rd,
for some C > 0. Note also that due to (3.2) the condition (ii) on
ξαVgf (x, ξ) given by (38) in [5] is equivalent to ii).
modifications, since we consider a more general situation.
Let σ > 1. We follow the proof of [5, Theorem 3.1], with necessary
EXTENDED GEVREY REGULARITY ...
13
i) ⇒ ii) Since Vgf (x, ξ) = F (f Txg)(ξ), we can formally write
ξαVgf (x, ξ) =
(2πi)αF f (∂α(f Txg))(ξ)
1
=
1
(2πi)αXβ≤α(cid:18)α
β(cid:19)F (∂α−βf ∂β(Txg))(ξ),
x, ξ ∈ Rd
(where we used the Leibnitz formula), and the formalism can be justi-
fied as follows.
ξαVgf (x, ξ) .
1
(2π)αXβ≤α(cid:18)α
β(cid:19)kF (∂α−βf Tx(∂βg))kL∞
(2π)αXβ≤α(cid:18)α
.
1
β(cid:19)k∂α−βf Tx(∂βg)kL1.
Since m is a positive v−moderate weight, from (3.3), (3.4), and
Holder's inequality we obtain
k∂α−βf Tx(∂βg)kL1 ≤ k∂α−βfkL∞
α − βτ α−βσ
. C α−βσ
f
1/mkm(x)(∂βTxg)kL1
m(x)kv(· − x)∂βg(· − x)kL1
f
. m(x)C α−βσ
α − βτ α−βσ
βτ βσ
. m(x) C ασ
where we used the fact that the sequence Mp = pτ pσ satisfies
p, q ∈ N,
q ≤ M τ,σ
(M.1)′ : M τ,σ
· C βσ
q ≤ p,
p−qM τ,σ
,
p
g
which follows from (M.1), see Lemma 2.1, and also [17].
Thus
f,g ατ ασ
,
ξαVgf (x, ξ) . (√d)αξαVgf (x, ξ)
m(x)Xβ≤α(cid:18)α
. √d
2π!α
m(x)ατ ασ C ασ
π !α
. √d
ii) ⇒ i) Note that (3.4) means that f ∈ M ∞,1
where we used (3.2), and ii) follows.
f,g = C ασ
β(cid:19) C ασ
f,g ατ ασ
,
f,g ατ ασ
may use the inversion formula for STFT (cf. Proposition 11.3.2.
m−1⊗1(Rd). Hence we
in
x, ξ ∈ Rd, α ∈ Nd,
14
NENAD TEOFANOV, FILIP TOMI ´C
[12]), and since f is a smooth function, we may assume that it holds
everywhere. So we formally write
∂αf (t) =
Vgf (x, ξ)∂α(MξTxg)(t) dxdξ
1
kgk2
1
kgk2
L2ZR2d
L2Xβ≤α(cid:18)α
β(cid:19)ZR2d
=
Vgf (x, ξ)(2πiξ)βMξTx(∂α−βg)(t) dxdξ,
α ∈ Nd, t ∈ Rd, where we used the Leibnitz formula. The estimates be-
low also justify the exchange of the order of derivation and integration.
Therefore,
∂αf (t) .
1
kgk2
L2Xβ≤α(cid:18)α
β(cid:19)(2π)βZR2d Vgf (x, ξ)ξβTx(∂α−βg)(t) dxdξ,
.
1
kgk2
L2Xβ≤α(cid:18)α
β(cid:19)(2π)βIα,β(t),
t ∈ Rd,
where we put
Iα,β(t) =ZR2d Vgf (x, ξ)ξβTx(∂α−βg)(t) dxdξ,
t ∈ Rd.
We note that ii) is equivalent with
hξiαVgf (x, ξ) . m(x)C ασ
f,g ατ ασ
, x, ξ ∈ Rd, α ∈ Nd,
and estimate Iα,β(t) as follows:
Iα,β(t) ≤ZR2dhξiβVgf (x, ξ)hξid+1
.ZR2d Vgf (x, ξ)hξiβ+d+1
β + d + 1τ β+d+1σZR2d
m(t)ZRd
. C · C βσ
f,g βτ βσ
= C · C βσ
. C β+d+1σ
m(x)
f,g
hξid+1g(α−β)(t − x) dxdξ
m(x)
1
hξid+1g(α−β)(t − x) dxdξ
hξid+1 m(x)g(α−β)(t − x) dxdξ
v(t − x)g(α−β)(t − x) dx
t ∈ Rd,
where C depends on τ, σ and d, and we used (M.2)′ property of the
sequence pτ pσ, p ∈ N, τ > 0, σ > 1, cf. Lemma 2.1.
m(t)kg(α−β)kL1
v,
f,g βτ βσ
EXTENDED GEVREY REGULARITY ...
15
Therefore, by (3.3) we obtain
∂αf (t) .
. C
m(t)
kgk2
m(t)
kgk2
L2Xβ≤α(cid:18)α
L2Xβ≤α(cid:18)α
β(cid:19)(2π)β C · C βσ
β(cid:19)(2π)βC βσ
f,g βτ βσ
f,g βτ βσ
kg(α−β)kL1
v,
C α−βσ
g
α − βτ α−βσ
,
. C
m(t)
kgk2
L2
C ασ
f,g ατ ασ
,
t ∈ Rd,
which gives (3.4). Here above we used (3.1) in several occasions.
The equivalence between ii) and iii) follows immediately from Defi-
(cid:3)
nition 2.1. Details are left for the reader.
Note that the condition (3.3) is weaker than the corresponding con-
dition in [5, Theorem 3.1], so by [5, Proposition 3.2] one can choose
elements from Gelfand-Shilov spaces as window functions (both in
Roumieu and Beurling case).
We note that if f ∈ DK
τ,σ, τ > 1, σ ≥ 1, then it obviously satisfies
the condition i) in Theorem 3.1, i.e.
∂αf (x) . m(x)C ασατ ασ
, α ∈ Nd,
so that Theorem 3.1 gives the decay properties of the STFT of elements
from DK
τ,σ.
We use this remark to extend Theorem 3.1. Recall the Paley-Wiener
type result for f ∈ DK
τ,σ describes the decay properties of the Fourier-
Laplace transform in the context of the extended Gevrey regularity.
The role of compact support in Paley-Wiener type theorems is essential.
In the following Corollary we weaken the assumptions from [25] (see
also [24, Corollary 3.2]) and allow the global growth condition given
by (3.4). Then, instead of cut-off functions, which are usually used in
localization procedures, we take a window in M 1,1
v⊗1(Rd)\{0}, and give
a Paley-Wiener type result by using the STFT.
Corollary 3.1. Let τ > 0, 1 < σ < 2, let v be a submultiplicative
weight, m ∈ Mv, and let g ∈ M 1,1
v⊗1(Rd)\{0} such that (3.3) holds for
some Cg > 0. Then a smooth function f satisfies (3.4) if and only if
Vgf (x, ξ) . m(x) expn−(cid:16)σ − 1
τ σ (cid:17) 1
σ−1 (ln(e + ξ))o,
σ−1 (e + ξ)
x, ξ ∈ Rd.
ln
ln
1
σ
σ−1
The proof is an immediate consequence of Theorems 2.1 and 3.1, and
Remark 2.2.
16
NENAD TEOFANOV, FILIP TOMI ´C
We finish this section with a version of Theorem 3.1 and Corollary
3.1 in the context of Beurling type ultradifferentiable functions. The
proofs are left as an exercise.
Theorem 3.2. Let τ > 0, σ ≥ 1, let v be a submultiplicative weight,
m ∈ Mv, and let g ∈ M 1,1
v⊗1(Rd)\{0} such that for some C > 0,
v(Rd) ≤ C ασ+1ατ ασ
, α ∈ Nd.
k∂αgkL1
For a smooth function f the following conditions are equivalent:
i) For every h > 0 there exists A > 0 such that
k∂αfkL∞(Rd) ≤ m(x)Ahασ
ατ ασ
, α ∈ Nd;
ii) For every h > 0 there exists A > 0 such that
(3.5)
ξαVgf (x, ξ) ≤ m(x)Ahασ
ατ ασ
, x, ξ ∈ Rd, α ∈ Nd;
iii) For every h > 0 there exists A > 0 such that
Vgf (x, ξ) ≤ m(x)Ae−Tτ,σ,h(ξ), x, ξ ∈ Rd.
Corollary 3.2. Let τ > 0, 1 < σ < 2, and let g ∈ M 1,1(Rd)\{0}
satisfies condition (3.3). Then a smooth function f satisfies (3.5) if
and only if for every H > 0 there exists A > 0 such that
σ−1 (e + ξ)
Vgf (x, ξ) ≤ m(x)A expn−(cid:16) σ − 1
τ σ (cid:17) 1
σ−1 (H ln(e + ξ))o, x, ξ ∈ Rd.
W
ln
σ−1
σ
1
Remark 3.1. A more general versions of Corollaries 3.1 and 3.2 when
σ ≥ 2 can be proved by using the asymptotic formulas (2.4) from
Theorem 2.1. This will give different necessary and sufficient conditions
for f in terms of the decay properties of the STFT. We leave details
for the reader.
4. Wave front sets WFτ,σ and STFT
In this section we characterize wave front sets related to the classes
introduced in Subsection 2.1, by the means of the STFT and extended
associated function from Subsection 2.2. We start with the following
definition of the wave front set WFτ,σ(u) of a distribution u with respect
to the extended Gevrey regularity, see also [22, 23, 25, 24, 32] for details.
Definition 4.1. Let U ⊆ Rd be open, τ > 0, σ > 1 or τ > 1 and
σ = 1, u ∈ D′(U), and let (x0, ξ0) ∈ Rd × Rd\{0}. Then (x0, ξ0) 6∈
WF{τ,σ}(u) (resp. (x0, ξ0) 6∈ WF(τ,σ)(u)) if and only if there exists a
conic neighborhood Γ of ξ0, a compact neighborhood K of x0, and
φ ∈ DK
(τ,σ) ) such that φ = 1 on some neighborhood of
{τ,σ} (resp. φ ∈ DK
EXTENDED GEVREY REGULARITY ...
17
x0, and there exists A, h > 0 (for every h > 0 there exists A > 0 such
that)
cφu(ξ) ≤ A
hN σ N τ N σ
ξN
, N ∈ N , ξ ∈ Γ .
τ,σ, see [25].
By using the Paley-Wiener theorem for DK
τ,σ it can be proved that
Definition 4.1 does not depend on the choice of the cut-off function
φ ∈ DK
Note that when τ > 1 and σ = 1 we have WF{τ,1}(u) = WFτ (u),
where WFτ (u) denotes Gevrey wave front set, cf.
[27]. We refer to
[22] for a relation between WFτ,σ(u) from Definition 4.1 and classical,
analytic and Gevrey wave front sets.
By using the ideas presented in [19] we resolve WFτ,σ(u) of a distri-
bution u via decay estimates of its STFT as follows.
Theorem 4.1. Let u ∈ D′(Rd), τ > 0, σ > 1. The following assertions
are equivalent:
i) (x0, ξ0) 6∈ WF{τ,σ}(u) (resp. (x0, ξ0) 6∈ WF(τ,σ)(u)) .
ii) There exists a conic neighborhood Γ of ξ0, a compact neighbor-
hood K of x0 such that for every φ ∈ DK
(τ,σ) )
there exists A, h > 0 (resp. for every h > 0 there exists A > 0)
such that
{τ,σ} (resp. φ ∈ DK
cφu(ξ) ≤ A
hN σ N τ N σ
ξN
, N ∈ N, ξ ∈ Γ ;
(4.1)
iii) There exists a conic neighborhood Γ of ξ0, a compact neigh-
(resp. φ ∈
for every h > 0 there
borhood K of x0 such that for every φ ∈ DK−{x0}
DK−{x0}
exists A > 0) such that
) there exists A, h > 0 (resp.
{τ,σ}
(τ,σ)
Vφu(x, ξ) ≤ Ae−Tτ,σ,h(ξ),
x ∈ K, ξ ∈ Γ,
(4.2)
where K − {x0} = {y ∈ Rd y + x0 ∈ K}.
Proof. We give the proof for the Roumieu case and leave the Beurling
case to the reader.
The equivalence i) ⇔ ii) is proved in Theorem 4.2. in [25].
ii) ⇒ iii) Without loss of generality we may assume that there exists
a conic neighborhood Γ of ξ0, compact neighborhood K1 = Br(x0),
r > 0, such that for every φ ∈ DK1
DK−{x0}
and x ∈ K, then Txφ(t) is a function in DK1
τ,σ.
Set K = Br/2(x0) and note that if φ(t) is an arbitrary function
τ,σ (4.1) holds.
τ,σ
18
NENAD TEOFANOV, FILIP TOMI ´C
Using the definition of STFT and (4.1) we have that
Vφu(x, ξ) = F (uTxφ) ≤ A inf
N ∈N
= Ae−Tτ,σ,1/h(ξ),
hN σ N τ N σ
ξN
x ∈ K, ξ ∈ Γ, for some constant A > 0, and iii) follows.
iii) ⇒ i) Since u ∈ D′(Rd), we may choose the window function
φ ∈ Dτ,σ(Rd) in iii) to be centered near any point in Rd. Let φ be
centered near 0, then clearly ψ = Tx0φ is centered near x0 and (4.2)
implies
cψu(ξ) = Vφu(x0, ξ) . e−Tτ,σ,h(ξ) . inf
for some h > 0 and the proof is finished.
N ∈N
(1/h)N σN τ N σ
ξN
,
ξ ∈ Γ,
(cid:3)
Next we discuss local extended Gevrey regularity via the STFT. To
that end we introduce the singular support as follows (cf. [31]).
Definition 4.2. Let there be given x0 ∈ Rd, u ∈ D′(U), τ > 0 and
σ > 1. Then x0 6∈ singsuppτ,σ(u) if and only if there exists open
neighborhood Ω ⊂ U of x0 such that u ∈ Eτ,σ(Ω).
The local regularity is related to the wave front set as follows.
Proposition 4.1. Let τ > 0 and σ > 1, u ∈ D′(U). Let π1 : U ×
Rd\{0} → U be the standard projection given by π1(x, ξ) = x. Then
singsuppτ,σ(u) = π1(WFτ,σ(u)) .
We refer to [31] for the proof, see also [22, Theorem 3.1] and [10,
Proposition 11.1.1].
Theorem 4.2. Let there be given x0 ∈ Rd, u ∈ D′(U), τ > 0 and
σ > 1. Then x0 6∈ singsuppτ,σ(u) if and only if there exists a compact
neighborhood K of x0 such that for every φ ∈ DK−{x0}
(resp. φ ∈
DK−{x0}
for every h > 0 there exists
A > 0) such that
) there exists A, h > 0 (resp.
{τ,σ}
(τ,σ)
Vφu(x, ξ) ≤ Ae−Tτ,σ,h(ξ),
x ∈ K, ξ ∈ Rd\{0},
(4.3)
where K − {x0} = {y ∈ Rd y + x0 ∈ K}.
Proof. We give the proof for the Roumieu case only. If x0 6∈ singsupp{τ,σ}(u)
then by Proposition 4.1 it follows that (x0, ξ) 6∈ WF{τ,σ}(u) for any
ξ ∈ Rd\{0}, so that (4.3) follows from Theorem 4.1 iii).
Now assume that (4.3) holds, then it holds for any cone Γ. By
Theorem 4.1 we conclude that (x0, ξ) 6∈ WF{τ,σ}(u) for any ξ ∈ Rd\{0}.
Now Proposition 4.1 implies that x0 6∈ singsupp{τ,σ}(u), and the proof
is completed.
(cid:3)
EXTENDED GEVREY REGULARITY ...
19
Acknowledgement
This work is supported by MPNTR through Project 174024.
References
[1] M. Cicognani, D. Lorenz, Strictly hyperbolic equations with coefficients low-
regular win time and smooth in space, J. Pseudo-Differ. Oper. Appl., 9 (2018),
643 -- 675.
[2] J. Chung, S.-Y., Chung, D., Kim, A characterization for Fourier hyperfunctions,
Publ. Res. Inst. Math. Sci., 30 (1994), 203 -- 208
[3] E. Cordero, S., Pilipovi´c, L., Rodino, N., Teofanov, Localization operators and
exponential weights for modulation spaces, Mediterranean Journal of Mathe-
matics 2 (2005) , 381 -- 394
[4] E. Cordero, S., Pilipovi´c, L., Rodino, N., Teofanov, Quasianalytic Gelfand-
Shilov spaces with application to localization operators, Rocky Mountain Jour-
nal of Mathematics, 40 (2010), 1123-1147
[5] E. Cordero, F. Nicola, L. Rodino, Gabor representations of evolution operators,
Trans. Am. Math. Soc., 367 (2015), 7639-7663.
[6] R.M. Corless, G.H. Gonnet, D.E.G. Hare, D.J. Jeffrey, D.E. Knuth, On the
Lambert W function, Adv. Comput. Math., 5 (1996), 329 -- 359.
[7] H. G., Feichtinger, Modulation spaces on locally compact abelian groups, Tech-
nical Report, University Vienna, 1983. and also in M. Krishna, R. Radha, S.
Thangavelu (eds.), Wavelets and Their Applications, Allied Publishers, 99 -- 140
(2003)
[8] H. G.,Feichtinger, K. Grochenig, Banach spaces related to integrable group
representations and their atomic decompositions I, J. Funct. Anal., 86 (1989),
307 -- 340
[9] H. G.,Feichtinger, K. Grochenig, Banach spaces related to integrable group rep-
resentations and their atomic decompositions II. Monatsh. f. Math. 108 (1989),
129 -- 148
[10] G. Friedlander, M. Joshi, The Theory of Distributions, Cambridge University
Press, Cambridge, 1998.
[11] I.M. Gelfand, G.E. Shilov, Generalized Functions II, Academic Press, New
York, 1968.
[12] K. Grochenig, Foundations of Time-frequency analysis, Birkhauser, Boston,
2001.
[13] K. Grochenig, Weight functions in time-frequency analysis, in: L. Rodino,
B.-W., Schulze, , M. W. Wong (eds.) Pseudodifferential Operators: Partial
Differential Equations and Time-Frequency Analysis, Fields Institute Comm.,
52 (2007), 343 -- 366
[14] K. Grochenig, G., Zimmermann, Hardy's theorem and the short-time Fourier
transform of Schwartz functions, J. London Math. Soc., 63 (2001), 205 -- 214
[15] A. Hoorfar, M. Hassani, Inequalities on the Lambert W function and hyper-
power function, J. Inequalities in Pure and Applied Math., 9 (2008), 5pp.
[16] L. Hormander, The Analysis of Linear Partial Differential Operators. Vol. I:
Distribution Theory and Fourier Analysis, Springer-Verlag, 1983.
20
NENAD TEOFANOV, FILIP TOMI ´C
[17] H. Komatsu, Ultradistributions, I: Structure theorems and a characterization.
J. Fac. Sci. Univ. Tokyo, Sect. IA Math., 20 (1973), 25 -- 105.
[18] F. Nicola, L. Rodino, Global Pseudo-differential calculus on Euclidean spaces,
Pseudo-Differential Operators. Theory and Applications 4, Birkhauser Verlag,
2010.
[19] S. Pilipovi´c, B. Prangoski, On the characterizations of wave front sets via
short-time Fourier transform, preprint (arXiv:1801.05999).
[20] S. Pilipovi´c, N., Teofanov, Wilson bases and ultra-modulation spaces, Math.
Nachr., 242 (2002), 179 -- 196
[21] S. Pilipovi´c, N. Teofanov, and F. Tomi´c, On a class of ultradifferentiable func-
tions, Novi Sad Journal of Mathematics, 45 (2015), 125 -- 142.
[22] S. Pilipovi´c, N. Teofanov, F. Tomi´c, Beyond Gevrey regularity, J. Pseudo-
Differ. Oper. Appl., 7 (2016), 113 -- 140.
[23] S. Pilipovi´c, N. Teofanov, and F. Tomi´c, Superposition and propagation of
singularities for extended Gevrey regularity, Filomat, 32 (2018), 2763 -- 2782.
[24] S. Pilipovi´c, N. Teofanov, and F. Tomi´c, Regularities for a new class of spaces
between distributions and ultradistributions, Sarajevo Journal of Mathematics,
14 (2) (2018), 251264.
[25] S. Pilipovi´c, N. Teofanov, and F. Tomi´c, A Paley-Wiener theorem in extended
Gevrey regularity, submitted (arXiv:1901.00698).
[26] A. Rainer, G. Schindl, Composition in ultradifferentiable classes, Studia Math,
224 (2) (2014), 97 -- 131.
[27] L. Rodino, Linear Partial Differential Operators in Gevrey Spaces, World Sci-
entific, 1993.
[28] M. A., Shubin, Pseudodifferential Operators and Spectral Theory, Springer-
Verlag, Berlin, second edition (2001)
[29] N. Teofanov, Modulation spaces, Gelfand-Shilov spaces and pseudodifferential
operators, Sampl. Theory Signal Image Process, 5 (2006), 225 -- 242
[30] N. Teofanov, Gelfand-Shilov spaces and localization operators, Funct. Anal.
Approx. Comput. 7 (2015), 135-158
[31] N. Teofanov, F. Tomi´c, Inverse closedness and singular support in extended
Gevrey regularity, J. Pseudo-Differ. Oper. Appl., 8 (3) (2017), 411 -- 421.
[32] N. Teofanov, F. Tomi´c, Ultradifferentiable functions of class M τ,σ
and microlo-
cal regularity, in: Oberguggenberger M., Toft J., Vindas J., Wahlberg P. (eds)
Generalized Functions and Fourier Analysis. Operator Theory: Advances and
Applications, vol 260. Birkhuser, Cham (2017), 193 -- 213.
p
[33] 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
[34] J. Toft, Images of function and distribution spaces under the Bargmann trans-
form, J. Pseudo-Differ. Oper. Appl. 8 (2017), 83 -- 139
[35] F. Tomi´c, A microlocal property of PDOs in E(τ,σ)(U ), in The Second Con-
ference on Mathematics in Engineering: Theory and Applications, Novi Sad,
(2017), 7 -- 12.
EXTENDED GEVREY REGULARITY ...
21
Nenad Teofanov, University of Novi Sad, Faculty of Sciences, De-
partment of Mathematics and Informatics, Trg Dositeja Obradovi´ca
4, 21000 Novi Sad, Serbia
E-mail address: [email protected]
Filip Tomi´c, University of Novi Sad, Faculty of Technical Sciences,
Department of Fundamental Sciences, Trg Dositeja Obradovi´ca 6,
21000 Novi Sad, Serbia
E-mail address: [email protected]
|
1703.04773 | 2 | 1703 | 2017-07-28T17:36:54 | Hypercyclic homogeneous polynomials on $H(\mathbb C)$ | [
"math.FA"
] | It is known that homogeneous polynomials on Banach spaces cannot be hypercyclic, but there are examples of hypercyclic homogeneous polynomials on some non-normable Fr\'echet spaces. We show the existence of hypercyclic polynomials on $H(\mathbb C)$, by exhibiting a concrete polynomial which is also the first example of a frequently hypercyclic homogeneous polynomial on any $F$-space.
We prove that the homogeneous polynomial on $ H(\mathbb C)$ defined as the product of a translation operator and the evaluation at 0 is mixing, frequently hypercyclic and chaotic. We prove, in contrast, that some natural related polynomials fail to be hypercyclic. | math.FA | math |
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
RODRIGO CARDECCIA, SANTIAGO MURO
Abstract. It is known that homogeneous polynomials on Banach spaces cannot be hypercyclic, but
there are examples of hypercyclic homogeneous polynomials on some non-normable Fr´echet spaces. We
show the existence of hypercyclic polynomials on H(C), by exhibiting a concrete polynomial which is also
the first example of a frequently hypercyclic homogeneous polynomial on any F -space. We prove that the
homogeneous polynomial on H(C) defined as the product of a translation operator and the evaluation
at 0 is mixing, frequently hypercyclic and chaotic. We prove, in contrast, that some natural related
polynomials fail to be hypercyclic.
1. Introduction
Let X be an F -space. A function T : X → X is said to be hypercyclic if there exists x ∈ X such
that its orbit, OrbT (x) := {T n(x) : n ∈ N}, is dense in X. In this case, x is called a hypercyclic vector.
The space H(C) of entire functions, with the compact open topology, was of crucial importance since
the beginnings of the theory of hypercyclic linear operators. Indeed, the first example of a hypercyclic
operator was found by Birkhoff in [11]. There, he showed that there exists an entire function g ∈ H(C)
whose translations by natural numbers approximate uniformly on compact sets any other entire function,
i.e. the translation operator τ1f (z) = f (z +1) acting on the space of entire functions H(C) is hypercyclic.
Later, MacLane [19] exhibited the second example of a hypercyclic operator, also on H(C), proving that
the differentiation operator Df (z) = f′(z) is also hypercyclic.
At the beginning of the 1990 decade, the theory of hypercyclic operators began to have a great
development. An article that inspired much of the subsequent work was the seminal paper of Godefroy
and Shapiro [14], where the authors proved (among other things) an important generalization of the
results of Birkhoff and MacLane. More recently, the concept of frequently hypercyclic operator was
introduced in [4], and shortly after, the operators considered by Birkhoff, MacLane, Godefroy and
Shapiro were shown to be also frequently hypercyclic [5, 12]. For a systematic treatment of hypercyclic
operators and related topics see the recent books [16, 6] and the references therein.
As a natural extension of the linear theory, one may study orbits of (non-linear) polynomial operators
on F -spaces. The first results were obtained by Bernardes in [8], in the context of homogeneous poly-
nomials acting on Banach spaces. Maybe surprisingly, he showed that no homogeneous polynomial, of
2010 Mathematics Subject Classification. 47H60 37F10, 47A16, 30D20, 30K99 .
Key words and phrases. frequently hypercyclic operators, homogeneous polynomials, entire functions, universal functions.
Partially supported by ANPCyT PICT 2015-2224, UBACyT 20020130200052BA and CONICET.
1
2
RODRIGO CARDECCIA, SANTIAGO MURO
degree ≥ 2, acting on a Banach space can be hypercyclic. In contrast, if the F -space is not normable,
it may support hypercyclic homogeneous polynomials. The first to realize this fact was Peris [23, 24].
As it is natural, the space where he sought a homogeneous hypercyclic polynomial was H(C). Unfortu-
nately, the example he gave was not well defined. However, he was able to construct another example,
this time on the space CN, the Fr´echet space of all complex sequences. He showed that the polynomial
(an) 7→ (a2
n+1) is not only hypercyclic but also chaotic on CN.
After the example of Peris, some other hypercyclic homogeneous polynomials were presented, on some
Kothe echelon spaces (including the space H(D), see [21]) and on some spaces of differentiable functions
on the real line [3]. But there are, up to our knowledge, no examples of hypercyclic homogeneous
polynomials on H(C). There are also no examples of frequently hypercyclic homogeneous polynomials
on any F -space. Given the key role of H(C) in the theory of linear dynamics, we believe it is desirable
to exhibit examples of hypercyclic homogeneous polynomials on H(C).
There are also some other articles investigating the dynamics of non-homogeneous polynomials ([7, 9,
17, 18, 20, 21, 25]) and of multilinear mappings ([10, 15]) on infinite dimensional spaces. For example, in
the recent paper [9], the existence of hypercyclic (non-homogeneous) polynomials of arbitrary positive
degree is shown on any infinite dimensional Fr´echet space.
In this note we show that the 2-homogeneous polynomial P (f ) = f (0) · τ1f defined on H(C) is mixing,
chaotic and frequently hypercyclic. In contrast, we prove that the polynomial P (f ) = f (0) · f′ is not
hypercyclic on H(C).
2. A hypercyclic polynomial on H(C)
In this section we prove Theorem 2.1, our main result, which states that there is a very natural
hypercyclic homogeneous polynomial on H(C). Let us first recall some definitions. If T is a mapping
acting on a topological space X, T is said to be transitive if for each nonempty open sets U, V ⊂ X there
exists n ∈ N such that T n(U ) ∩ V 6= ∅. If there exists n0 such that T n(U ) ∩ V 6= ∅ for every n ≥ n0, the
mapping is said to be mixing. Clearly a mixing map is transitive and by Birkhoff's Transitivity Theorem,
if the map is continuous and the underlying space is a complete separable metric space without isolated
points, then the map is transitive if and only if it is hypercyclic.
A set A ⊆ N is said to have positive lower density if
lim inf
n
#{x ∈ A : 0 ≤ x ≤ n}
n
> 0,
where # denotes the cardinality of the set. We say that a map T is frequently hypercyclic if there exists
some x ∈ X such that for every nonempty open set U , the set {n ∈ N : T n(x) ∈ U } has positive lower
density. Finally T is said to be chaotic if it is hypercyclic and has a dense set of periodic vectors.
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
3
Let X be an F -space. A mapping P : X → X is said to be a d-homogeneous polynomial, P ∈ P(dX),
if P is the restriction to the diagonal of some d-multilinear map L ∈ L(dX; X), that is,
d
P (x) = L(
x, ..., x).
z } {
We will be dealing with homogeneous polynomials acting on the space H(C) of entire functions, which
endowed with the compact open topology is a Fr´echet space. The seminorms
kf kK := sup
z∈K
f (z),
where K is a compact set, define the topology in H(C). Thus, the sets
Uǫ,f,R = {h ∈ H(C) : kh − f kB(0,R) < ǫ},
with ǫ, R > 0 form a basis of open neighborhoods of f ∈ H(C).
Theorem 2.1. The polynomial P ∈ P(2H(C)) defined by
P (f )(z) = f (0) · f (z + 1)
is mixing, chaotic and frequently hypercyclic.
Observe that P n(f )(z) = cn(f )f (z + n) where
(1)
cn(f ) = f (0)2n−1
· f (1)2n−2
· . . . · f (n − 1).
Proof that P is mixing. Let U and V be open sets. We can suppose that U = Uǫ,f,R and V = Uǫ,g,R.
Also we may suppose that R /∈ N and that f and g do not have zeros in Z.
By Runge's Theorem we can find, for n large enough, a polynomial p such that p ∈ U and p(·+n) ∈ V .
We assert that a more careful application of Runge's Theorem allows us to obtain a polynomial p that
also satisfies cn(p) ∼ 1.
Let n0 ∈ N such that n0 > 2R + 2, and fix n ≥ n0. This implies that B(0, R) ∩ B(n, R) = ∅ and that
we can define open balls B1, B2 ⊆ C such that {B(0, R), B1, B2, B(n, R)} are pairwise disjoint and such
that ⌊R⌋ + 1 ∈ B1, and ⌊R⌋ + 2, . . . , n − ⌊R⌋ − 1 ∈ B2, where ⌊R⌋ denotes the integer part of R. See
Fig. 1.
Define g(z) = g(z − n) and α any 2n−⌊R⌋−2th-root of the number
f (0)2n−1
· . . . · f (⌊R⌋)2n−⌊R⌋−1
· 12n−⌊R⌋−3
· . . . · 12⌊R⌋
· g(n − ⌊R⌋)2⌊R⌋−1
· . . . · g(n − 1).
4
RODRIGO CARDECCIA, SANTIAGO MURO
0
R ⌊R⌋ + 1
⌊R⌋ + 2
n − ⌊R⌋ − 1
n − R
n
B1
B(0, R)
B(n, R)
B2
Figure 1. The open sets B(0, R), B1, B2 and B(n, R).
Also consider the perturbed open sets in H(C),
Uk =nh ∈ H(C) : kf − hkB(0,R) <
Vk =nh ∈ H(C) : kg − hkB(n,R) <
z∈B1(cid:12)(cid:12)(cid:12)(cid:12)
α(cid:12)(cid:12)(cid:12)(cid:12)
k =(cid:26)h ∈ H(C) : sup
k =(cid:26)h ∈ H(C) : sup
h(z) −
z∈B2
1
h(z) − 1 <
W 1
W 2
ǫ
ǫ
ko ,
ko ,
k(cid:27) and
k(cid:27) .
1
1
<
By Runge's Theorem we can find, for each k, a polynomial pk in Uk ∩ W 1
k ∩ W 2
k ∩ Vk.
Observe that for j ∈ N, we have
pk(j) →
f (j)
if j ≤ ⌊R⌋;
1
α
1
if j = ⌊R⌋ + 1;
if ⌊R⌋ + 1 < j ≤ n − ⌊R⌋ − 1;
g(j)
if n − ⌊R⌋ − 1 < j ≤ n − 1,
as k → ∞. Also, by definition of α, cn(pk) → 1 as k → ∞. Thus, for large k we have
kcn(pk)pk − gkB(n,R) ≤
ǫ
2
+ cn(pk) − 1kpkkB(n,R) ≤
ǫ
2
+ cn(pk) − 1(cid:16) ǫ
2
+ kgkB(n,R)(cid:17) < ǫ.
Therefore, we can find a polynomial pk with
kf − pkkB(0,R) < ǫ
and
kg − P n(pk)kB(0,R) = kg − cn(pk)pkkB(n,R) < ǫ.
This proves that P is mixing.
Proof that P is chaotic. Observe that a periodic vector for P is a quasiperiodic function, that is,
there exist α ∈ C and n ∈ N such that f (z + n) = αf (z). If this happens, then the homogeneity of P
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
5
forces
(2)
2n−1
cn(f )α(cid:19) 1
(cid:18) 1
f
to be an n-periodic vector for P . Note also that if f is a periodic vector for P , then λf is not necessarily
a periodic vector for P .
show that the set of periodic functions satisfying that (cid:16) 1
It is known that the set of periodic functions is dense in H(C). To prove that P is chaotic we will
2n−1 ∼ 1 is also dense in H(C). So, it
will be useful to have a good characterization of the periodic functions. Define an infinite segment
L, beginning at zero, so that L ∩ T is not a root of the unity and θ = arg(L)
4 , 3
4 ), and define
z maps any band
Ω = C − L (so that a branch of the logarithm may be defined on Ω). Then, since e
cn(f )(cid:17) 1
∈ ( 1
2π
2πi
n
(n(θ + k), n(1 + θ + k)) × iR to Ω, we have the following.
Lemma 2.2. If an entire function f is n-periodic then there exist g holomorphic on Ω such that
for all z belonging to any band of the form (n(θ + k), n(1 + θ + k)) × iR. Reciprocally, if f (z) = g(cid:16)e
for all z ∈ C, then f is n-periodic.
2πi
n
z(cid:17)
f (z) = g(e
2πi
n
z)
We now begin with our proof of the chaoticity of P .
Let U = {h ∈ H(C) : kh − gkB(0,R) < ǫ} be a nonempty open set of H(C) with R 6∈ N. Our goal is
−1
2n−1 f ∈ U . By (2), this implies
to find, for some n ∈ N, an n-periodic function f ∈ U so that also cn(f )
that cn(f )
−1
2n−1 f is a periodic vector for P and therefore, the set of periodic vectors is dense in H(C).
Take n0 ∈ N so that n0 > 4R. Since the periodic functions with period greater than n0 are dense in
H(C) [2, Sublemma 7] , there exists, for some n > n0, an n-periodic function f with kf − gkB(0,R) < ǫ
2 .
We may also suppose that f (j) 6= 0 for every j ∈ Z.
Now take k ∈ Z such that B(0, R) is contained in the band (θ + k, n + θ + k) × iR. Thus, by the
previous Lemma, f (z) = h(e
2πiz
n ) for every z ∈ B(0, R) for an appropriate holomorphic function h on Ω.
Instead of applying Runge's Theorem to the function f we will apply it to h. The function e
n maps
N0 to Gn, the n-th roots of the unity, which we will denote ω0, . . . , ωn−1. Thus, h(ωj) 6= 0 for every
ωj ∈ Gn and
2πiz
P n(h ◦ e
2πiz
n ) = cn(h) · h ◦ e
2πiz
n ,
2πiz
n ). Consider B1 = {e
. . . h(ωn−1) = cn(h ◦ e
where cn(h) := h(ω0)2n−1
that ω0, ω1, . . . , ω⌊R⌋, and ωn−⌊R⌋, . . . ωn−1 are all in B while ω⌊R⌋+1, . . . , ωn−⌊R⌋−1 are in (B1
). Also,
4 ), B1 ⊆ Ω and h is holomorphic on B1. Runge's Theorem allows us to find h
since n > 4R and θ ∈ ( 1
such that h is close to h on B1 and at the same time cn(h) is close to 1. Indeed, choose B2 and B3 open
sets so that ω⌊R⌋+1 ∈ B2, ω⌊R⌋+2, . . . , ωn−⌊R⌋−1 are in B3 and B1, B2, B3 are pairwise disjoint. See Fig.
2.
: z ∈ B(0, R)} and observe
4 , 3
c
2πiz
n
6
RODRIGO CARDECCIA, SANTIAGO MURO
B2
ω⌊R⌋+2
ω⌊R⌋+1
ω⌊R⌋
B3
ω0
B1
ω−⌊R⌋
ω−⌊R⌋−1
Lθ
Figure 2. Shape and location of the sets B1, B2, B3, Lθ and Gn.
Now define U l
1, U l
2, U l
3 open sets in H(C) as
U l
U l
U l
1 =ng ∈ H(C) : kg − hkB1 <
z∈B2(cid:12)(cid:12)(cid:12)(cid:12)
2 =(cid:26)g ∈ H(C) : sup
3 =(cid:26)g ∈ H(C) : sup
g(z) −
z∈B3
ǫ
1
lo ,
α(cid:12)(cid:12)(cid:12)(cid:12)
<
g(z) − 1 <
ǫ
l(cid:27) ,
l(cid:27) ;
ǫ
where α is any 2n−⌊R⌋−2 th-root of the number
h(ω0)2n−1
· . . . · h(ω⌊R⌋)2n−⌊R⌋−1
· 12n−⌊R⌋−3
· . . . · 12⌊R⌋
· h(ωn−⌊R⌋)2⌊R⌋−1
· . . . · h(ωn).
By Runge's Theorem we can find, for every l, a polynomial hl ∈ U l
1 ∩ U l
2 ∩ U l
3. By the choice of hl and
α, cn(hl) → 1 and khl − hkB1 → 0 as l tends to infinity. Thus,
kcn(hl ◦ e
2πiz
n )
−1
2n−1 hl ◦ e
2πiz
n − f kB(0,R) = kcn(hl)
−1
2n−1 hl ◦ e
2πiz
n − h ◦ e
2πiz
n kB(0,R) → 0.
Therefore, for large enough l, cn(hl ◦ e
n-periodic and by (2) cn(hl ◦ e
n )
−1
2n−1 hl ◦ e
n )
2πiz
2πiz
n
is a periodic vector for P .
2πiz
−1
2n−1 hl ◦ e
2πiz
n ∈ U . Finally by Lemma 2.2, hl ◦ e
2πiz
n
is
Proof that P is frequently hypercyclic. To prove the existence of frequently hypercyclic vectors we
will use the following result [13, Lemma 2.5].
Lemma 2.3. There exist pairwise disjoint subsets An,m of N, each having positive lower density such
that for any k ∈ An,m, k′ ∈ An′,m′ we have k > m, and k − k′ > m + m′ if k 6= k′.
We will now prove that P supports a frequently hypercyclic vector. Our proof follows Example 9.6 in
[16] together with a careful use of Runge's theorem.
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
7
Let An,m be the subsets given by the above lemma and consider (kj)j ⊆ N the increasing sequence
m is a non natural radius. It
follows from the above lemma that the Bj are pairwise disjoint. Let (pn)n be a dense sequence in H(C)
such that pn(l) 6= 0 for every l ∈ Z, n ∈ N.
formed byS An,m. If kj ∈ An,m we define Bj = B(kj, rj), where rj = m
2 + 1
Applying Runge's Theorem recursively we will find (fj)j ⊆ H(C) such that fj approximates pj(z) :=
pn(z − kj) on Bj, where n is the only natural number such that kj ∈ An,m, such that ckj (f ) is close
to 1, and such that fj(l) 6= 0 for every l ∈ Z. To achieve this, let (ǫj)j ∈ ℓ1 be a sequence of positive
numbers such that ǫj < 1
m whenever kj ∈ An,m. We will define inductively a sequence of entire functions
(fj)j ⊂ H(C) and a sequence of positive numbers (δj)j satisfying
(a) kfj+1 − fjkB(0,kj + 1
kj
(b) kckj+1(fj+1)fj+1 − pj+1kBj+1 < δj,
(c) δj+1 < min{ǫj+1, ǫj+2/2, γj+1},
) < δj,
(d) δj+1 < γl0 −Pj
(e) fj+1 has no zero in Z,
l=l0
δl, for l0 = 1, . . . , j and
where γj is a positive number that depends on fj as follows. For any g ∈ H(C) and j ∈ N, let
Φj : Ckj → C defined as
Thus, if we set
Φj(cid:0)x0, . . . , xkj−1(cid:1) := x2kj −1
0
· . . . · xkj−1.
Kg,j = sup
g(z),
z<kj+1
we have that Φj is uniformly continuous on the product of the closed discs Π
and
kj
l=1B(0, Kg,j + kǫk1) ⊂ Ckj
ckj (g) = Φj(xg,j),
ǫj
where xg,j is the vector (g(0), . . . , g(kj − 1)). Since Φj is uniformly continuous, given the number
2(Kg,j +kǫk1) > 0 there exists γg,j > 0 such that for every x, y ∈ B(0, Kg,j + kǫk1) × · · · × B(0, Kg,j + kǫk1)
we have that
(3)
if kx − yk∞ < γg,j then Φj(x) − Φj(y) <
ǫj
2(Kg,j + kǫk1)
.
Once fixed the function fj, γj will be defined as γj := γfj ,j.
We start setting f1(z) = p1(z) (thus we have defined γ1 := γf1,1). We define δ1 > 0 such that
δ1 < min{ǫ1, ǫ2/2, γ1}.
Suppose now that f1, . . . fj ∈ H(C) and δ1, . . . , δj ∈ R>0 have been constructed and satisfy (a)-(e).
We will now define fj+1 and δj+1.
Consider B1
j+1 and B2
j+1 disjoint open sets so that kj +1 ∈ B1
j+1, {kj +2, . . . , kj+1 −⌊rj+1⌋−1} ⊆ B2
j+1,
and such that {z < kj + 1
kj
}, B1
j+1, B2
j+1, Bj+1 are all disjoint. See Fig. 3.
8
RODRIGO CARDECCIA, SANTIAGO MURO
B(0, kj + 1
kj
)
B1
j+1
B2
j+1
Bj+1
kj + 1
kj
kj + 1
kj + 2
kj+1 − ⌊rj+1⌋ − 1
kj+1 − ⌊rj+1⌋
kj+1
Figure 3. The open sets B(0, kj + 1
kj
), B1
j+1, B2
j+1 and Bj+1.
Now by Runge's Theorem we can find auxiliary entire functions gl satisfying
(i) kgl − fjkB(0,kj + 1
kj
(ii) supz∈B1
(iii) supz∈B2
(iv) kgl − pj+1kBj+1 < ǫj
(v) gl has no zero in Z;
) < ǫj
l ,
gl(z) − dj < ǫj
l ,
gl(z) − 1 < ǫj
l ,
l and
j+1
j+1
where 1/dj is a (2kj+1−kj−2)-th root of the number
fj(0)2kj+1 −1
· . . . · fj(kj )2kj+1−kj −1
· 12kj+1−kj −3
· . . . · 12⌊rj+1⌋
· pj+1(kj+1 − ⌊rj+1⌋)2⌊rj+1⌋−1
· . . . · pj+1(kj+1 − 1),
so that ckj+1(gl) approaches to 1 as l → ∞. Take now l large enough so that ǫj+1
l < δj and such that
kckj+1(gl)gl − pj+1kBj+1 ≤ ckj+1(gl) − 1 · kglkBj+1 + kgl − pj+1kBj+1
≤ ckj+1(gl) − 1 ·(cid:16)kpj+1kBj+1 +
ǫj
l (cid:17) +
ǫj
l
< δj.
For such an l, we set fj+1 := gl, hence determining as above the number γj+1 > 0. Finally we set
δj+1 > 0 such that
δj+1 < min(ǫj+1,
ǫj+2
2
, γj+1, γj − δj, γj−1 − δj − δj−1, . . . , γ1 −
δl) .
j
Xl=1
This concludes the construction of (fj)j and (δj)j satisfying (a)-(e).
Now we define f as
Note that (d) implies that
f := f1 +
(fj+1 − fj).
∞Xj=1
and in particular, the sequence (δj)j ∈ ℓ1. Thus, (a) and the fact that kj + 1
kj
entire function and that f = limj→∞ fj. Moreover,
δn ≤ γj,
Xn≥j
ckj (f ) = Φ(xf,j),
→ ∞ imply that f is an
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
9
where xf,j =(cid:16)fj(0) +Pn≥j(fn+1(0) − fn(0)), . . . , fj(kj − 1) +Pn≥j(fn+1(kj − 1) − fn(kj − 1))(cid:17) . Also,
by (a) and (c), both xf,j and xfj ,j belong to B(0, Kfj ,j + kǫk1) × · · · × B(0, Kfj ,j + kǫk1) and since
kxf,j − xfj ,jk∞ ≤ sup
z<kj−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn≥j
fn+1(z) − fn(z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
δn ≤ γj,
≤Xn≥j
by (3) we obtain
(4)
ckj (f ) − ckj (fj) = Φj(xf,j) − Φj(xfj ,j) ≤
ǫj
2(Kj + kǫk1)
.
Let z ∈ Bj, then using (a),(b),(c),(4),
ckj (f )f (z) − pn(z − kj) ≤ ckj (fj)f (z) − pn(z − kj) + (cid:0)ckj (f ) − ckj (fj)(cid:1) f (z)
ǫj
f (z)
≤ δj−1 +
2(Kj + kǫk1)
≤
≤
ǫj
2
ǫj
2
+
+
ǫj
2(Kj + kǫk1)
ǫj
2(Kj + kǫk1)
Therefore, for kj ∈ An,m,
fn+1(z) − fn(z))
(fj(z) +Xn≥j
(Kj +Xn≥j
ǫn) ≤ ǫj.
sup
z< m
2 + 1
m
P kj f (z) − pn(z) = sup
z∈Bj−kj
P kj f (z) − pn(z) = sup
z∈Bj
ckj (f )f (z) − pn(z − kj) ≤ ǫj <
1
m
.
Note that the sets
Un,m :=(h ∈ H(C) :
h(z) − pn(z) <
1
m) ,
sup
z< m
2 + 1
m
with n, m ∈ N, form a basis of open sets of H(C). Finally since for k ∈ An,m, P k(f ) ∈ Un,m and each
An,m has positive lower density, we conclude that f is a frequently hypercyclic vector for P .
3. Examples of non-hypercyclic polynomials on H(C)
The purpose of this section is to show that many natural homogeneous polynomials on H(C) fail to
be hypercyclic. In view of what we have proved in the previous section, and the fact that translation
and differentiation operators on H(C) share many dynamical properties, a natural candidate to be
hypercyclic is the homogeneous polynomial P (f ) := f (0) · f′. Another favorable motivation comes from
the study of bilinear hypercyclic operators on H(C). B`es and Conejero considered in [10, Section 4]
the bilinear operator M (f, g) = f (0)g′, and showed that it is hypercyclic (in the sense defined by the
authors). Since M (f, f ) = P (f ) it is reasonable to expect that P is also hypercyclic. Surprisingly, the
polynomial fails to be hypercyclic.
Proposition 3.1. The homogeneous polynomial P (f ) := f (0)f′(z) is not hypercyclic.
10
RODRIGO CARDECCIA, SANTIAGO MURO
Proof. The iterates of a function f are of the form P n(f ) = cn(f )f (n), with
cn(f ) = f (0)2n−1
f′(0)2n−2
. . . f (n−1)(0).
This fact can be easily proven by induction. Also the functions cn(f ) can be constructed recursively as
c1(f ) = f (0);
cn(f ) = cn−1(f )2f (n−1)(0).
Let us define X ⊆ H(C) as X := {f ∈ H(C) : lim sup cn(f )(n!)2Rn = ∞, for some R > 1}. The proof
of the proposition will be divided in three steps:
(1) X is P -invariant.
(2) 0 is not in the closure of X.
(3) If f /∈ X then f is not a hypercyclic vector for P .
If we prove (1), (2) and (3) it clearly follows that P is not hypercyclic.
Proof of (1). Note that ck+1(f ) = f (0)ck(P (f )). Take f ∈ X and let R > 1 such that lim sup cn(f )(n!)2Rn =
∞. Then P f ∈ X, because
lim sup(cid:12)(cid:12)cn(P f )(n!)2(R + 1)n(cid:12)(cid:12) = lim sup(cid:12)(cid:12)(cid:12)(cid:12)
cn+1(f )((n + 1)!)2Rn+1
(R + 1)n
f (0)(n + 1)2Rn+1(cid:12)(cid:12)(cid:12)(cid:12)
= ∞.
Proof of (2). Suppose that (fk)k ⊆ X is a null sequence. Since cn is continuous, there exists (kn)n ⊆ N
with cn(fkn) < 1
22n . Taking a subsequence, we may suppose that
cn(fn) <
1
22n .
We claim that for each n, there exists j ≥ 0 such that f (j+n)
n
(0) > (j + n)j+n. Indeed, if f (j+n)
n
(0) ≤
(j + n)j+n for every j ≥ 0 then we show by induction that cj+n(fn) < 1
22
n+
already know it for j = 0. Suppose it is true for some j. Note that for every n, j, we have
for every j ≥ 0. We
j
2
Thus
2(n+j) log2(n+j)
2(√2−1)2n+
1+j
2
≤ 1.
cn+j+1(fn) = c2
n+j(fn)f (j+n)
n
(0) ≤
1
22n+1+
j
2
f (j+n)
n
(0)
≤
1
22n+1+
j
2
2(n+j) log2(n+j)
2(n+j) log2(n+j)
2 2(√2−1)2n+
This implies that fn is not in X, which is a contradiction.
22n+
=
+ 1
j
2
j
2
≤
+ 1
2
1
22n+
.
j+1
2
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
11
Therefore f (j+n)
n
(0) > (j + n)j+n for some j ≥ 0. Recall that the seminorms given by
kf kk = sup
f (j)(0)
j
kj
j!
,
define the topology of H(C) (see for example [22, Example 27.27] or [23]).
For each n, let jn ≥ 0 be such that f (jn+n)
n
(0) > (jn + n)jn+n. Thus,
kfnkk = sup
j
f (j)
n (0)
kj
j!
> f (jn+n)
n
(0)
kjn+n
(n + jn)!
> 1.
This contradicts the fact that fn → 0.
Proof of (3). Suppose that cn(f )(n!)2 ≤ L < ∞ for every n ≥ 0. Then, by the Cauchy inequalities,
we have for some M, r > 0,
δ0(P n+1f ) = cn+1(f )f (n+1)(0) ≤
L
(n!)2
M (n + 1)!
rn+1 → 0.
Therefore f is not a hypercyclic vector.
(cid:3)
Aron and Miralles [3] showed that the polynomial P ∈ P(2C k(R)) defined as P (f )(z) = f (z + 1)2
is hypercyclic. However, if we consider the analogous map, but in H(C), the polynomial fails resound-
ingly to be hypercyclic. The rigidity of the holomorphic functions obstructs our search of hypercyclic
homogeneous polynomials. In particular, Hurwitz's Theorem impose several restrictions to this kind of
problem. This was already noted in [1], as the authors were looking for algebras of hypercyclic vectors.
Proposition 3.2. Let a, b ∈ C and let P ∈ P(2H(C)) be the polynomial defined by
If f is an accumulation point of an orbit of P then either f is identically zero or f (z) 6= 0 for every
P (g)(z) = g(z + a)g(z + b).
z ∈ C. In particular, P is not hypercyclic.
Proof. Note that if g ∈ H(C), then
(5)
P k(g)(z) =
k
Yj=0
g(z + ja + (k − j)b)(k
j).
Let f ∈ H(C) and suppose that f has a zero of order m ≥ 1 at z0. If P kl(g) converges uniformly to f on
B(z0, 2(a + b)), by Hurwitz's Theorem, for each sufficiently small δ > 0, there is some l0 such that for
l ≥ l0 the number of zeros of P kl(g) in B(z0, δ) is exactly m. Thus for each l ≥ l0 there is some jl ≤ kl
such that g(· + jla + (kl − jl)b) has a zero of positive order in B(z0, δ). But this implies, by (5), that
P kl(g) must have another zero of order ≥ kl in B(z0 + a − b, δ) (or in B(z0 − a + b, δ) if jl = kl), and
therefore f must be identically zero.
(cid:3)
12
RODRIGO CARDECCIA, SANTIAGO MURO
It was proved in [1] that, in contrast with the translation operator, the differentiation operator does
admit an algebra of hypercyclic vectors. However, Hurwitz's Theorem also prevents powers of the
differentiation operator to be hypercyclic.
Proposition 3.3. Let P ∈ P(2H(C)) be one of the following polynomials
(i) P (g)(z) = g′(z)2,
(ii) P (g)(z) = g(z)g′(z).
Then, P is not hypercyclic.
Proof. We only prove (i), the proof of (ii) is analogous. Note that if g′(z0) = 0 then (P (g))′(z0) = 0.
Suppose that P nk (g) → z2. Then P nk (g)′ → 2z. By Hurwitz's Theorem, there exists k0 such that for
every k ≥ k0, P nk (g)′ has a zero of order 1 in B(0, 1). Thus P n(g)′ has a zero of order at least 1 in
B(0, 1) for every n ≥ nk0. Therefore g is not hypercyclic for P .
(cid:3)
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 and Operator Theory, 58(4):591 -- 596, 2007.
[2] R. M. Aron and D. Markose. On universal functions. J. Korean Math. Soc., 41(1):65 -- 76, 2004. Satellite Conference on
Infinite Dimensional Function Theory.
[3] R. M. Aron and A. Miralles. Chaotic polynomials in spaces of continuous and differentiable functions. Glasg. Math.
J., 50(2):319 -- 323, 2008.
[4] F. Bayart and S. Grivaux. Hypercyclicity: the role of the unimodular point spectrum. (Hypercyclicit´e : Le role du
spectre ponctuel unimodulaire.). C. R., Math., Acad. Sci. Paris, 338(9):703 -- 708, 2004.
[5] F. Bayart and S. Grivaux. Frequently hypercyclic operators. Trans. Amer. Math. Soc., 358(11):5083 -- 5117 (electronic),
2006.
[6] F. Bayart and E. Matheron. Dynamics of linear operators. Cambridge Tracts in Mathematics 179. Cambridge: Cam-
bridge University Press. xiv, 337 p., 2009.
[7] L. Bernal-Gonz´alez. Backward φ-shifts and universality. Journal of mathematical analysis and applications, 306(1):180 --
196, 2005.
[8] N. C. Bernardes. On orbits of polynomial maps in Banach spaces. Quaest. Math., 21(3-4):311 -- 318, 1998.
[9] N. C. Bernardes and A. Peris. On the existence of polynomials with chaotic behaviour. Journal of Function Spaces and
Applications, 2013, 2013.
[10] J. B`es and J. A. Conejero. An extension of hypercyclicity for N -linear operators. In Abstract and Applied Analysis,
volume 2014. Hindawi Publishing Corporation, 2014.
[11] G. D. Birkhoff. D´emonstration d'un th´eor`eme ´el´ementaire sur les fonctions enti`eres. C. R Acad. Sci. Paris, 189:473 -- 475,
1929.
[12] A. Bonilla and K.-G. Grosse-Erdmann. On a theorem of Godefroy and Shapiro. Integral Equations Operator Theory,
56(2):151 -- 162, 2006.
[13] A. Bonilla and K.-G. Grosse-Erdmann. Frequently hypercyclic operators and vectors. Ergodic Theory Dyn. Syst.,
27(2):383 -- 404, 2007.
[14] G. Godefroy and J. H. Shapiro. Operators with dense, invariant, cyclic vector manifolds. J. Funct. Anal., 98(2):229 -- 269,
1991.
HYPERCYCLIC HOMOGENEOUS POLYNOMIALS ON H(C).
13
[15] K.-G. Grosse-Erdmann and S. G. Kim. Bihypercyclic bilinear mappings. Journal of Mathematical Analysis and Appli-
cations, 399(2):701 -- 708, 2013.
[16] K.-G. Grosse-Erdmann and A. Peris Manguillot. Linear chaos. Universitext. Berlin: Springer. xii, 386 p. EUR 53.45 ,
2011.
[17] A. Jung. Mixing, simultaneous universal and disjoint universal backward φ-shifts. Journal of Mathematical Analysis
and Applications, 452(1):246 -- 257, 2017.
[18] S. G. Kim, A. Peris, and H. G. Song. Numerically hypercyclic polynomials. Archiv der Mathematik, pages 1 -- 10, 2012.
[19] G. R. MacLane. Sequences of derivatives and normal families. J. Analyse Math., 2:72 -- 87, 1952.
[20] F. Mart´ınez-Gim´enez and A. Peris. Existence of hypercyclic polynomials on complex fr´echet spaces. Topology and its
Applications, 156(18):3007 -- 3010, 2009.
[21] F. Mart´ınez-Gim´enez and A. Peris. Chaotic polynomials on sequence and function spaces. International Journal of
Bifurcation and Chaos, 20(09):2861 -- 2867, 2010.
[22] R. Meise and D. Vogt. Introduction to functional analysis. Clarendon Press, 1997.
[23] A. Peris. Chaotic polynomials on Fr´echet spaces. Proc. Am. Math. Soc., 127(12):3601 -- 3603, 1999.
[24] A. Peris. Erratum to: "Chaotic polynomials on Fr´echet spaces". Proc. Am. Math. Soc., 129(12):3759 -- 3760, 2001.
[25] A. Peris. Chaotic polynomials on banach spaces. Journal of mathematical analysis and applications, 287(2):487 -- 493,
2003.
DEPARTAMENTO DE MATEM ´ATICA - PAB I, FACULTAD DE CS. EXACTAS Y NATURALES, UNIVER-
SIDAD DE BUENOS AIRES, (1428) BUENOS AIRES, ARGENTINA AND CONICET
E-mail address: [email protected]
E-mail address: [email protected]
|
1504.01263 | 2 | 1504 | 2015-12-22T19:22:52 | Uniqueness of Banach space valued graphons | [
"math.FA",
"math.CO",
"math.PR"
] | A Banach space valued graphon is a function $W:(\Omega, \mathcal{A},\pi)^2\to\mathcal{Z}$ from a probability space to a Banach space with a separable predual, measurable in a suitable sense, and lying in appropriate $L^p$-spaces. As such we may consider $W(x,y)$ as a two-variable random element of the Banach space. A two-dimensional analogue of moments can be defined with the help of graphs and weak-* evaluations, and a natural question that then arises is whether these generalized moments determine the function $W$ uniquely -- up to measure preserving transformations. The main motivation comes from the theory of multigraph limits, where these graphons arise as the natural limit objects for convergence in a generalized homomorphism sense. Our main result is that this holds true under some Carleman-type condition, but fails in general even with $\mathcal{Z}=\mathbb{R}$, for reasons related to the classical moment-problem. In particular, limits of multigraph sequences are uniquely determined - up to measure preserving transformations - whenever the tails of the edge-distributions stay small enough. | math.FA | math |
Uniqueness of Banach space valued graphons
MTA Alfréd Rényi Institute of Mathematics, Budapest, Hungary∗
Dávid Kunszenti-Kovács
Mathematics Subject Classification: 46G10, and 28B05, 05C60
July 30, 2018
Abstract
A Banach space valued graphon is a function W : (Ω, A, π)2 → Z from a probability
space to a Banach space with a separable predual, measurable in a suitable sense, and
lying in appropriate Lp-spaces. As such we may consider W (x, y) as a two-variable random
element of the Banach space. A two-dimensional analogue of moments can be defined
with the help of graphs and weak-* evaluations, and a natural question that then arises is
whether these generalized moments determine the function W uniquely -- up to measure
preserving transformations.
The main motivation comes from the theory of multigraph limits, where these graphons
arise as the natural limit objects for convergence in a generalized homomorphism sense.
Our main result is that this holds true under some Carleman-type condition, but fails in
general even with Z = R, for reasons related to the classical moment-problem. In partic-
ular, limits of multigraph sequences are uniquely determined - up to measure preserving
transformations - whenever the tails of the edge-distributions stay small enough.
1 Introduction
Moment determinacy deals with the question of whether a given type of probability measure
is uniquely determined by its moments. In the classical settings, the theory is rich and well
understood. For instance, if the probability measure lives on a bounded interval (Hausdorff
problem), then knowledge of the moments is enough to recover the measure. The same holds
true for the vector-valued version, where the measure is supported in a bounded domain of
Rk for some finite k. The notion of moments has to be slightly adapted though: to guarantee
, where Xj is the j-th
uniqueness, we need mixed moments, i.e., the expectations of Qk
coordinate of the random vector (1 ≤ j ≤ k), and the αj-s are nonnegative integers.
However, if the support is unbounded (cf. Stieltjes and Hamburger problems), there is no
general positive or negative answer to moment determinacy, and additional conditions (Car-
leman, Krein, etc.) are needed to prove or disprove uniqueness, see e.g. the monograph [1] by
j=1 X αj
j
∗The research leading to these results has received funding from the European Research Council under the
◦617747, and
European Union's Seventh Framework Programme (FP7/2007-2013) / ERC grant agreement n
from the MTA Rényi Institute Lendület Limits of Structures Research Group.
1
Akhiezer.
In a somewhat more general setting, given a measurable function f : [0, 1] → R, we have an
induced probability measure on R. Knowing the moments of this probability measure, we may
wish to recover f itself. Clearly this is not possible, but if the measure is moment determinate,
we may still recover a great deal of information about f . For instance, we may find a canonical
monotone increasing representation f ′ : [0, 1] → R, where f ′ is determined up to a null-set.
Also, we may find measure preserving transformations of [0, 1] that transform f and f ′ into
a.e. equal functions. A similar result holds for f and f ′ taking its values in Rk, where the
monotone reordering is extended to the lexicographic ordering.
In a recent paper by Borgs, Chayes and Lovász ([3]), the authors considered a variant of the
above Hausdorff question, involving an extra dimension in the domain of f . Namely, they
investigated bounded symmetric measurable functions f : [0, 1]2 → R. Such two-variable func-
tions also induce a probability measure on [0, 1], but the structure of the domain means that
there is an added spatial correlation in the function values. They proved that with an ap-
propriate notion of generalized moments that are adapted to the extra spatial dimension, all
such functions are uniquely determined by their generalized moments, up to measure preserv-
ing transformation of the variables. Note however, that there unfortunately is no canonical
reordering of the interval that would yield a "monotone" function in this two-variable setting.
The motivation for studying such functions comes from the theory of limits of simple dense
graphs, developed by Borgs, Chayes, Lovász, Sós and Vesztergombi [4, 5] and Lovász and
Szegedy [12]. Symmetric measurable functions W : [0, 1]2 → [0, 1] represent limit objects for
graph sequences under a combinatorial/probabilistic notion of convergence connected to homo-
morphism densities. The deep connection between analysis and limit theories of combinatorial
objects is further highlighted in the paper [13] by Lovász and Szegedy, where Szemerédi's
Regularity Lemma is reformulated and given analytic interpretations. We also note that the
question of uniqueness in the limit theory for hypergraphs was treated by Elek and Szegedy in
[6], but their methods were of a fundamentally different nature, making use of ultraproducts.
Further developments in this field have led to the investigation of limits of multigraphs with
no bound on the number of multiple edges between nodes, and more generally to limits of dec-
orated graphs. For the multigraph setting, this at a first glance simply corresponds to passing
from bounded to unbounded functions. But it turns out that for combinatorial reasons one
expects the limit functions to take measures as their values rather than simply a real number.
In [9], Lovász, Szegedy and the author developed a general functional analytic framework that
allows one to handle the various possible combinatorial interpretations opened up by the multi-
graph/decorated graph setting and compare the corresponding convergence notions.
The limit objects/graphons this generalized setting leads to are symmetric, weak-* measurable
functions W : [0, 1]2 → Z, where Z is a Banach space with separable predual (typically a space
of measures that depends on the specific combinatorial interpretation(s) studied). As such they
are two-variable random elements of said Banach space. The homomorphism densities that
are used to define convergence are integrals of products of weak-* evaluations of the graphon
(cf. Definition 10). These will be the two-dimensional generalizations of moments that are
adapted to the added structure of the domain of the graphons. For technical reasons, we shall
not restrict ourselves to graphons W with domain [0, 1]2, but rather more generally consider
domains that are the product of a probability measure space with itself, (Ω, A, π)2.
After one defines a class of objects that is rich enough to capture the whole limit theory, an
2
important question still remains, namely whether the class is too big or not. This was partly
answered in [9], where it was shown that every graphon is a limit of a sequence of decorated
graphs. Our aim is to address the remaining part of the question: how much redundancy is
there in the space of graphons? In other words, given two graphons, under what conditions
do they represent the same limit?
The paper follows the approach of [3], building on and refining its ideas and proofs and
combining them with functional analytic methods related to weak-* integrable functions to
extend the results to a much more general setting that also includes limits of multigraphs with
unbounded edge multiplicities. We show that if the generalized moments of a graphon satisfy
a Carleman-type condition, then, similarly to the one-variable case, the graphon is uniquely
determined, up to measure preserving transformations of the underlying space (Ω, A, π), see
(Theorem 11). In particular, bounded Banach space valued graphons are always moment
determinate.
However, as for the classical moment problems, one may not forgo some type of bounds on
the moments to guarantee uniqueness, and using moment indeterminate measures on N, one
can construct graphons the are not isomorphic in any sense, but have identical generalized
moments (cf. Section 7).
2 Preliminaries
As mentioned in the introduction, we are interested in functions in two variables taking values
in a Banach space Z, and a corresponding notion of moments. This involves taking integrals,
but there is no unique "natural" integral notion in Banach spaces.
The Bochner, or strong integral, corresponds to integrability "in norm", and is defined as a
limit of integrals of simple functions. The Pettis integral, or weak integral, uses duality to
reduce integrability of a Z-valued function to that of real valued ones through weak evalua-
tions. Finally, if Z possesses a predual, weak-* integrability can also be defined in a similar
way, as shall be done below. Of the three integral notions, weak-* integrability is the weak-
est property (provided it exists), and Bochner integrability the strongest, though some or all
properties coincide under certain conditions on the Banach space Z (e.g. separability).
In this paper we are interested in the largest class, that of weak-* integrable functions, as these
are the ones that arise naturally as limits of multigraph sequences (see [9]). After introducing
this class of functions and some of its properties, we shall turn our attention to the combi-
natorial structures that allow us to define a notion of moments adapted to the two-variable
setting and the added geometric structure that comes with it.
2.1 Weak-* integrable functions
Let Φ be a separable Banach space, and let Z denote its dual. The elements of Φ act on Z
as bounded linear functionals in the canonical way. Let further Ψ ⊂ Φ be a countable dense
subset.
Definition 1. Let (Ω, A, π) be a probability space. A function W : (Ω, A, π) → Z is called
weak-* measurable if for any ϕ ∈ Φ, the function hϕ, W i is measurable.
The weak-* measurable function W is called weak-* scalarly integrable, if for any ϕ ∈ Φ, the
3
function hϕ, W i lies in L1(π).
The weak-* measurable function W is called weak-* integrable (or Gelfand integrable), if there
exists a mapping µW : A → Z such that for any A ∈ A and ϕ ∈ Φ we have
ZA
hϕ, W i = ϕ(µ(A)).
Remark 2. Note that the standard definition of weak-* measurability only requires that the
weak-* evaluations be measurable with respect to the completion of the measure π. In this
paper, however, we shall need to differentiate between functions that are measurable with
respect to a measure, and those that are measurable only with respect to its completion.
Clearly a weak-* integrable function is also weak-* scalarly integrable. The following classi-
cal result shows that the converse is also true. We include its proof for the readers' conve-
nience.
Proposition 3. Each weak-* scalarly integrable function W : (Ω, A, π) → Z is weak-* inte-
grable.
Proof. For a given A ∈ A, consider the linear map WA : Φ → L1(Ω, A, π) given by WA(ϕ) :=
1A · hϕ, W i. Note that this map has a closed graph.
Indeed, given any convergent sequence ϕn → ϕ ∈ Φ with WA(ϕn) → f ∈ L1(Ω, A, π), we
can find a subsequence such that WA(ϕnk ) converges almost everywhere to f . But pointwise
WA(ϕn) converges to WA(ϕ), and so f = WA(ϕ) in L1(Ω, A, π).
Thus by the Closed Graph Theorem WA is bounded, and so
(cid:12)(cid:12)(cid:12)(cid:12)
ZA
hϕ, W i dπ(cid:12)(cid:12)(cid:12)(cid:12) ≤ZA
hϕ, W i dπ = kWA(ϕ)k ≤ kWAk · kϕk,
which means that the map ϕ 7→RAhϕ, W i dπ is a continuous linear functional on Φ. Therefore
there exists a representing element µW (A) ∈ Z, completing the proof.
The next result shows that the existence of a Radon-Nikodym derivative extends to the setting
of weak-* integrals, and it is a variant of a theorem due to Rybakov ([16, Thm. 2]). In his
paper Rybakov assumes the underlying measure space to be complete, as the proof makes
use of lifting on L∞(Ω, A). In our setting, however, the predual Φ is separable, and therefore
lifting can be avoided through a different approach, and the assertion holds even when the
measure space is not complete.
Proposition 4. Let (Ω, A, π) be a probability space, and suppose that the vector-valued mea-
sure µ : A → Z is of σ-finite variation, and µ ≪ π. Then there exists a weak-* integrable
function W : Ω → Z such that
hϕ, µ(A)i =ZA
hϕ, W i dπ
for every ϕ ∈ Φ and A ∈ A.
Proof. First, assume that there exists a c > 0 such that kν(A)k ≤ cµ(A), for every A ∈ A.
According to the classical Radon-Nikodym theorem for each ϕ ∈ Φ there exists a function
w′
ϕ ∈ L1(Ω, A, π) such that for all A ∈ A we have
4
hϕ, µ(A)i =ZA
w′
ϕ dπ
Clearly w′
ϕ ≤ ckϕk a.e., for each ϕ separately. In the general case, this is the point where a
lifting on L∞(Ω, A, π) would be used to assemble all these derivatives into a single Z-valued
function.
Instead, now let Ψ′ := linQ Ψ, and consider an enumeration ψ0, ψ1, ψ2, . . . of the countable set
Ψ′, where ψ0 = 0. Let us recursively do the following. Let wψ0 ≡ 0 and wψ1 := w′
, and for
ψ1
each n ≥ 2, if ψn is not in the linear hull of the previous ψi-s, then let wψn := w′
. If on the
ψn
i=1 aiwψi. Since
i=1 aiψi for some a1, a2, . . . , an−1 ∈ R, then let wψn :=Pn−1
other hand ψn =Pn−1
Radon-Nikodym derivation is linear, this is well-defined, and
hψn, µ(A)i =ZA
wψn dπ
for all A ∈ A and 1 ≤ n. Also, this can be linearly extended to all ψ ∈ linR Ψ =: Ψ′′.
Since Ψ′ ⊂ Φ is countable, there exists a set N ∈ A with π(N ) = 0 such that wψ ≤ ckψk for
all ω ∈ Ω\N and ψ ∈ Ψ′, and then by construction actually for all ψ ∈ Ψ′′. For each ω ∈ Ω\N
define the functional W ′
ω : Ψ′′ → R by
W ′
ω(ψ) := wψ(ω).
Clearly W ′
Wω ∈ Z to Φ. Now we can define a function W : (Ω, A, π) → Z through
ω is linear and bounded by c on Ψ′′, and hence has a unique bounded extension
W (ω) :=(cid:26) Wω
0
if ω ∈ Ω\N ;
otherwise.
Notice that W is automatically A-measurable, and we have kW (ω)k ≤ c for all ω ∈ Ω. Using
again that Ψ′′ = Φ, we may thus conclude that
hϕ, µ(A)i =ZA
hϕ, W i dπ
for every ϕ ∈ Φ and A ∈ A.
To finish the proof in the general case, we need that Ω can be written as the union of a set
of measure zero, and sets An ∈ A (n ∈ N+) such that for each n, we have kµ(A)k ≤ nπ(A)
for all An ⊃ A ∈ A. Since µ is of σ-finite variation, the Radon-Nikodym theorem still applies
to the measures ϕ ◦ µ, ϕ ∈ Φ. Define the functions wψ (ψ ∈ Ψ′) as above. For n ∈ N+, let
on Z, each An is measurable and satisfies kµ(A)k ≤ nπ(A) for all An ⊃ A ∈ A. It thus only
remains to be shown that Ω\ ∪n∈N+ An =: S is a set of measure zero.
An :=nω ∈ Ω(cid:12)(cid:12)(cid:12)wψ(ω) ≤ nkψk ∀ψ ∈ Ψ′o. Then, since Ψ′ is countable and generates the norm
Since µ is of σ-finite variation, we can write S = S∞
the variation of µ on Sk, and q := (cid:4) s
i=1 Si with Si ∈ A, and µ being of finite
variation on each Si. Suppose that there exists an Sk with d := π(Sk) > 0, and let then s be
k,j ∈ A
d(cid:5) + 1. For j ≥ 1, define the sets S+
k,j ∈ A and S−
through
S+
k,j :=nω ∈ Sk(cid:12)(cid:12)(cid:12)wψa(ω) < qkψak ∀ 1 ≤ a < j, and wψj (ω) ≥ qkψjko ,
5
and
S−
Then kµ(S±
s ≥
k,j)k ≥ qπ(S±
k,j :=nω ∈ Sk(cid:12)(cid:12)(cid:12)wψa(ω) < qkψak ∀ 1 ≤ a < j, and wψj (ω) ≤ −qkψjko .
∞Xj=1
k,j)j∈N+ form a partition of Sk, hence
k,j), and the sets (S±
k,j) + π(S−
k,j)k ≥ q
kµ(S+
k,j)k +
∞Xj=1
∞Xj=1
π(S+
kµ(S−
k,j) = qπ(Sk) = qd > s,
leading to a contradiction. Thus S is the countable union of sets of measure zero, and the
proof is complete.
The following lemma is an easy corollary.
Lemma 5. Let (Ω, A, π) be a probability space, and W : (Ω, A, π) → Z a weak-* integrable
function with kW k ∈ L1(π). Let further A′ ⊂ A be a sub-σ-algebra. Then there exists a weak-*
integrable function W ′ : (Ω, A′, πA′ ) → Z such that
hϕ, µ(A′)i =ZA′
hϕ, W i dπ
for every ϕ ∈ Φ and A′ ∈ A′. If W ′
everywhere with respect to πA′.
If for some 1 ≤ p < ∞ we have kW k ∈ Lp, then also kW ′k ∈ Lp with kW kp ≥ kW ′kp.
2 are two such functions, then W ′
1 and W ′
1 = W ′
2 almost
Proof. Let the vector-valued measure µ : A′ → Z be defined through the weak-* inte-
gral µ(A′) := RA′ W . Then µ is clearly absolutely continuous with respect to πA′ and
has σ-finite variation. Hence by Proposition 4, there exists a weak-* integrable function
W ′ : (Ω, A′, πA′ ) → Z satisfying the required equality. If W ′
2 are two such func-
tions, then for each ψ ∈ Ψ we have
1 and W ′
hψ, W ′
1i = E(hψ, W i(cid:12)(cid:12)A′ ) = hψ, W ′
2i
πA′-almost everywhere. Since Ψ is countable and separates Z, the assertion follows.
The inequality between the norms follows from the fact that hϕ, W (·)i ≤ kW (·)k · kϕk, and
Radon-Nikodym derivation is order-preserving, hence a contraction on every Lp space.
This allows us to extend the notion of conditional expectation to weak-* integrable functions
with values in Z.
Definition 6. Let (Ω, A, π) be a probability space, and W : (Ω, A, π) → Z a weak-* inte-
grable function. Let further A′ ⊂ A be a sub-σ-algebra. Then the πA′ -almost everywhere
unique function W ′ given in Lemma 5 is called the conditional expectation of W with respect
to the σ-algebra A′, and will be denoted by E(W A′ ).
The function W is said to be almost A′-measurable, if W = E(W A′ ) holds π-almost every-
where.
Definition 7. A symmetric function W : (Ω, A, π)2 → Z is called a Z-graphon if it is weak-*
measurable with respect to the completion A × A of the underlying σ-algebra, and the function
(x, y) 7→ kW (x, y)kZ lies in L(Ω,A,π) := T1≤p<∞ Lp(cid:16)(Ω, A, π)2(cid:17). Note that this function is
6
measurable with respect to the completed σ-algebra, since Φ is separable, and for a countable
dense subset Ψ ⊂ Φ we have
kW (x, y)kZ = sup
ψ∈Ψ\{0}
hf, W (x, y)i
kψkΦ
.
Let the space of Z-graphons on (Ω, A, π)2 be denoted by W(Ω,A,π). We set
(i.e., we take the Z-norm of W (x, y) for every x, y ∈ [0, 1], and then take the Lp-norm of the
resulting function).
kW kp :=(cid:13)(cid:13)kW (., .)kZ(cid:13)(cid:13)p.
The following notions are related to how "nice" a graphon and the underlying measure space
are.
Definition 8. A graphon W ∈ W(Ω,A,π) is called strong if it is also measurable with respect
to the σ-algebra A × A.
A graphon W ∈ W(Ω,A,π) is called Lebesguian if the measure space (A, π) is a standard (or
Lebesgue) measure space (for a definition of standard probability spaces see e.g. [15, Section
2.2]).
The graphon W ∈ W(Ω,A,π) is complete if (A, π) is a complete measure space, and the com-
pletion of a graphon W ∈ W(Ω,A,π) is the complete graphon W ∈ W(Ω,A,π) obtained through
completing the measure space (A, π).
Two points x1, x2 ∈ Ω are called twins with respect to the graphon W ∈ W(Ω,A,π) if W (x1, y) =
W (x2, y) for almost all y. The graphon W is called almost twin-free if there exists a null-set
N ⊂ Ω such that no two points in Ω\N are twins.
For a ϕ ∈ Φ and a W ∈ W(Ω,A,π) let the function Wϕ
: Ω2 → R be defined by
Note that we always have Wϕ
∈ L(Ω,A,π).
Wϕ
(x, y) := ϕ(W (x, y)).
The following lemma lets us prove weak-* measurability using only a countable dense subset
of Φ.
Lemma 9. Let Ψ ⊂ Φ be a countable dense subset, and let W : (Ω, A, π) → Z be a function
such that Wψ
is measurable for each ψ ∈ Ψ. Then W is weak-* measurable.
Proof. Since Ψ is countable and dense in Φ, for any ϕ ∈ Φ there exists a sequence (ψn) ⊂ Ψ
that converges to ϕ in norm. But then Wϕ
is the pointwise limit of the measurable functions
Wψn
, and hence itself measurable.
2.2 Decorated graphs and graph densities
We now turn our attention to decorated graphs, and the "moments" they induce for Banach
space valued graphons.
If X is any set, an X -decorated graph is a graph where every edge ij is decorated by an element
Xij ∈ X . An X -decorated graph will be denoted by (G, g), where G is a simple graph, and
g : E(G) → X .
7
Definition 10. For a Φ-decorated graph F = (F, f ) on k vertices and a Z-graphon W , let
Z
x1,...,xk∈(Ω,A,π) Yij∈E(F )
t(F, W ) :=
Wfij
(xi, xj)dx1 . . . xk.
Note that since kW kZ lies in all Lp spaces for 1 ≤ p < ∞, this integral is always finite.
Our aim is to investigate to what degree graph densities determine a Z-graphon. Just as in
the real valued case, there is an inherent indeterminacy related to the choice of the underlying
measure space. We therefore recall some further definitions from measure theory, and from the
extension of these notions to graphons (cf. [3, Section 2]). Given a graphon W ∈ W(Ω,A,π), we
may obtain a graphon with the exact same graph densities by deleting a null-set from Ω. To
this type of indeterminacy corresponds an equivalence relation between measure spaces. Let
(Ω, A, π) and (Ω′, A′, π′) be two probability spaces. They are called isomorphic mod 0 if there
exist null-sets N ⊂ Ω and N ′ ⊂ Ω′ and a bijection µ : Ω\N → Ω′\N ′ such that both µ and
µ−1 are measure preserving. The map µ itself is called an isomorphism mod 0.
Another way of obtaining a new graphon with the same graph densities is by applying a
"pull-back" using a measure preserving map. Let η : (Ω′′, A′′, π′′) → (Ω, A, π) be a measure
preserving map. The pull-back W ′′ := (W )η ∈ W(Ω′′,A′′,π′′) of the graphon W ∈ W(Ω,A,π)
under η is defined through (W )η(x, y) := W (η(x), η(y)). Then W ′′ again clearly has the same
graph densities as W .
Let U ∈ W(Ω1,A1,π1) and V ∈ W(Ω2,A2,π2) be two graphons, and suppose that we have a map
η : Ω1 → Ω2 that is measure preserving from the completion A1 into A2 and satisfies V η = U
almost everywhere. Then we say that η is a weak isomorphism from U to V .
These isomorphism notions can be extended to the graphons themselves. We say that the
graphons W1 ∈ W(Ω1,A1,π1) and W2 ∈ W(Ω2,A2,π2) are isomorphic mod 0 if there exists an
isomorphism mod 0 µ : Ω1 → Ω2 such that (W2)µ = W1 almost everywhere. As a short-hand
notation we write W1 ∼= W2.
We say that W1 and W2 are weakly isomorphic if there exists a graphon W3 ∈ W(Ω3,A3,π3) and
weak isomorphisms from each of W1 and W2 into W3. Note that it is not immediately clear
that being weakly isomorphic is an equivalence relation.
2.3 Main result
Using the notations and definitions introduced above, the main result of this paper can be
formulated as follows.
Theorem 11.
(i) Let W ∈ W(Ω,A,π) and W ′ ∈ W(Ω′,A′,π′) be almost twin-free strong Lebesguian graphons.
Further assume that the p-norms of W satisfy:
for all k ∈ N+. Then
∞Xn=1
kW k−k
2nk = ∞
t(F, W ) = t(F, W ′)
8
for every Ψ-decorated simple graph F if and only if W and W ′ are isomorphic mod 0.
(ii) Let W ∈ W(Ω,A,π) and W ′ ∈ W(Ω′,A′,π′) be general graphons. Further assume that the
p-norms of W satisfy:
for all k ∈ N+. Then
∞Xn=1
kW k−k
2nk = ∞
t(F, W ) = t(F, W ′)
for every Ψ-decorated simple graph F if and only if W and W ′ are weakly isomorphic.
Here we note that the family of Carleman-type conditions required can be viewed as requiring
that not only the function/random variable kW (·, ·)k, but all of its powers should satisfy the
Carleman condition in order to obtain a moment determinacy result. This may seem super-
fluous, but the bounds are actually used for determinacy of new random variables induced by
partially decorated graphs (see Section 5). These moments do not correspond to expectations
of simple powers of the same random variable. In addition, random variables whose moments
are bounded from above by (or even equal to) the moments of a power of a moment determi-
nate random variable need not be moment determinate themselves, even for the most common
distributions we know (see, e.g., the papers [2] by Berg, and [10] by Lin and Huang, or the
monograph [17] by Stoyanov).
The above theorem is a generalization of [3, Theorem 2.1], since in the case of bounded real
valued functions W , the Carleman conditions are automatically satisfied.
3 Graphon constructions
To prove the second part of our main theorem, we shall - following Borgs, Chayes and Lovász [3]
- be transforming our original graphon by changing the underlying measure space in several
steps until we end up with a standard Lebesguian space, thereby reducing the problem to
part (i). This will be achieved mainly through manipulating the corresponding σ-algebra and
adapting the graphon to these successive changes. We therefore briefly recall the necessary
notions from measure theory.
A set S of subsets of Ω induces a partition P[S] of Ω through the natural equivalence relation
ω1 ∼ ω2 ⇔ [∀S ∈ S : (ω1 ∈ S ∧ ω2 ∈ S) ∨ (ω1 6∈ S ∧ ω2 6∈ S)]. This is the finest partition for
which S separates the classes. A graphon W ∈ W(Ω,A,π) is said to be separating if A separates
Ω, or in other words if W = W/P[A].
A σ-algebra A is said to be countably generated if there is a countable set S ⊂ A such that
the generated σ-algebra satisfies σ(S) = A. A set S ⊂ A is said to be a basis of the measure
space (Ω, A, π) if σ(S) is dense in A, that is, if for every A1 ∈ A there exists an A2 ∈ σ(S)
such that π(A1△A2) = 0. A graphon W ∈ W(Ω,A,π) is said to be countably generated if A
itself is.
A probability space (Ω′, A′, π′) is said to be a full subspace of the probability space (Ω, A, π)
if Ω′ ⊂ Ω has outer measure 1 (it need not be measurable) and (A′, π′) is the restriction of
9
(A, π) to Ω′, i.e., A′ = {A ∩ Ω′A ∈ A} and π′(A ∩ Ω′) = π(A) for all A ∈ A.
A measure preserving map µ : (Ω, A, π) → (Ω′, A′, π′) between two probability spaces is called
an embedding of (Ω, A, π) into (Ω′, A′π′) if it is an isomorphism between the former and a
full subspace of the latter. Given two graphons W ∈ W(Ω,A,π) and W ′ ∈ W(Ω′,A′,π′), the
embedding µ is said to be an embedding of W into W ′ if in addition (W ′)µ = W almost
everywhere.
Let us start with the following Lemma, that allows us to change the σ-algebra we work with
to a countably generated one one, without losing measurability.
Lemma 12. Let (Ω, A) and (Ω′, A′) be measurable spaces, and let W : Ω × Ω′ → Z be a
weak-* measurable function with respect to the σ-algebra (A × A′). Then there exist countably
0 ⊂ A′ such that W is weak-* measurable with respect to
generated σ-algebras A0 ⊂ A and A′
(A0 × A′
0).
Proof. Choose a dense countable subset Ψ ⊂ Φ. For each ψ ∈ Ψ we can find countable sets
ψ ⊂ A′ such that ψ ◦ W is measurable w.r.t. the generated σ-algebra (cf.
Sψ ⊂ A and S ′
[3, Lemma 3.4], boundedness is actually not needed, since we can compose with the arctan
function to reduce to the bounded case). Thus taking A0 and A′
0 to be the sub-σ-algebras
generated by (Sψ)ψ∈Ψ and (S ′
ψ)ψ∈Ψ, respectively, Lemma 9 ensures the required measurability
of W .
Next we shall introduce two further constructions that given a graphon allow us to create a
new graphon, preserving some of its essential properties.
Remark 13. For ease of notation, using Proposition 4, the integrals we write from here on
are to be understood in the weak-* sense rather than the strong/Bochner sense.
Lemma 14. Let (Ω, A, π) and (Ω′, A′, π′) be probability spaces, let τ : Ω → Ω′ be a measure
preserving map, and let W ∈ W(Ω,A,π).
(i) There exists a strong graphon W ′ ∈ W(Ω′,A′,π′) that satisfies
ZA′
Wτ (x′, y′)dπ′(x′)dπ′(y′) =Zτ −1(A′
1)×τ −1(A′
2)
for all A′
1, A′
1×A′
2
2 ∈ A′.
(ii) If τ is an embedding, then (Wτ )τ = W almost everywhere.
W (x, y)dπ(x)dπ(y)
(1)
Proof. First let A′
Then define a measure µ on A′
τ := τ −1(A′) ⊂ A and define fW := E(W A′
τ through
τ × A′
τ × A′
τ ). Also let πτ := πA′
τ .
µ(A1 × A2) =ZA1×A2
W (x, y)dπ(x)dπ(y)
for all A1, A2 ∈ A′
τ . Since kW k lies in L1, we have that µ ≪ π × π. Let µτ be the push-
forward of µ. Since ϕ is measure-preserving, µϕ is a Z-valued measure on A′ that is absolutely
continuous with respect to π′ × π′.
By Lemma 5, there exists a weak-* integrable function Wτ : Ω′ × Ω′ → Z such that for all
A′
1, A′
2 ∈ A′,
µ(A′
1 × A′
2) =ZA′
1×A′
2
W ′(x, y)dπ′(x)dπ′(y).
10
This W ′ then clearly satisfies equation 1, and since the push-forward µ′ of the symmetric
measure µ is itself symmetric, it follows that W ′ is too. By the norm inequality in Lemma 5,
we actually have that W ′ ∈ W(Ω′,A′,π′), completing the proof of part (i).
By construction (Wτ )τ = fW = E(W A′
A′
τ .
τ × A′
τ ), so part (ii) follows from the definition of
Definition 15. The function Wτ defined in 14 is called the push-forward of W under τ .
Given a graphon W ∈ W(Ω,A,π) and a partition P of Ω, we can use the push-forward con-
struction to define the factor graphon W/P of W under P. Consider the surjection µ : ω 7→
[ω] from Ω to Ω/P, where [ω] denotes the partition class of ω. Define the measure space
(Ω/P, A/P, π/P) as the push-forward of (Ω, A, π) under µ. Then µ is automatically measure
preserving, and we let W/P := Wµ ∈ W(Ω/P,A/P,π/P).
Note that push-forwards being a type of conditional expectations, the moments of a graphon
are usually not preserved. However, if the underlying measure space is carefully manipulated,
this problem can be avoided, as illustrated by the next theorem, which sums up the different
steps that will allow us to pass from part (i) to part (ii) of Theorem 11.
Theorem 16. Let W ∈ W(Ω,A,π) be a graphon.
(i) One can change the value of W on a set of π × π-measure 0 to get a strong Z1-valued
graphon.
(ii) Suppose that W is a strong graphon. Then there exists a countably generated σ-algebra
A′ ⊂ A such that W is weak-* measurable with respect to (A′ × A′).
(iii) The graphon W/P[A] is separating. If W is countably generated, then so is W/P[A].
(iv) Suppose that W is a separating graphon on a probability space with a countable basis.
Then the completion of W can be embedded in a Lebesguian graphon.
(v) Suppose that W is a strong graphon, and let P be the partition into the twin-classes of
W . Then W/P is almost twin-free. If W is Lebesguian, then W/P is Lebesguian as well.
Furthermore the projection W → W/P is a weak isomorphism.
For part (i), consider the conditional expectation W ′ := E(W A × A ). Since W and the
underlying σ-algebra both are symmetric, we may assume that so is W ′. By construction, W ′
is measurable with respect to A × A, so it is enough to show that W = W ′ almost everywhere.
(W ′ − W ) = 0 for all A1, A2 ∈ A, and therefore also for all
To this end note that RA1×A2
S ∈ A × A we have RS(W ′ − W ) = 0, implying W ′ − W = 0 almost everywhere.
Part (ii) is an easy consequence of Lemma 12.
For part (iii), note that by identifying elements in the came class of the partition P[A], we
obtain a σ-algebra that is isomorphic to A under the natural map.
For part (iv), let S be a countable set generating A. Since any separating complete probability
space with a countable basis can be embedded into a Lebesgue space (see e.g. [15, Section 2.2])
and S is a countable basis for (A, π), there exists an embedding η : (Ω, A, π) → (Ω′, A′, π′)
where (Ω′, A′, π′) is a Lebesgue space. Consider the push-forward W ′ := Wη. By Lemma
14(ii), the mapping η is an embedding of the completion W of W into the Lebesguian graphon
W ′.
11
For part (v), we may by part (ii) assume that A is countably generated, since the relation of
being twins is the same for any underlying σ-algebra B × B that leaves W weak*-measurable
(and hence weak-* integrable). Let AP denote the sub-σ-algebra of A consisting of the sets
that do not separate any pair of twins. Note that by the construction in Lemma 14, we then
have that (W/P)τ = E(W AP × AP ) ∈ W(Ω,AP ,π). Thus the projection W → W/P is a weak
isomorphism if W = E(W AP × AP ) almost everywhere.
Let cW := E(W AP × AP ). Since Ψ ⊂ Φ is countable and dense, hence separates elements in
Z, it is sufficient to show that for any A, B ∈ A and ψ ∈ Ψ we have
ZA×B
Wψ
(x, y)dπ(x)dπ(y) =ZA×B cWψ
(x, y)dπ(x)dπ(y).
This, and the fact that W/P is twin-free, can easily be proven, and we refer to [3, Section
3.3.5] for the details.
We now wish to show that W/P is Lebesguian under the assumption that W itself is. By [15,
Section 3.2] the measure space (Ω/P, A/P, π/P) is a Lebesgue space if there exists a countable
set S ⊂ A that separates points if and only if they are from different partition classes. Let T
be a countable set generating A, closed under finite intersections. For each A ∈ A and x ∈ Ω,
let
µx(A) =ZA
W (x, y)dπ(y).
Since W is weak-* integrable with respect to A × A, the mapping A 7→ µx(A) is a Z-valued
measure for all x ∈ Ω, and x 7→ µx(A) is an A-measurable function on Ω for each A ∈ A.
Note that then x, x′ are twins if and only if µx(A) = µx′(A) for all A ∈ A, and since each
measure µx(·) is uniquely determined by the values taken on T , x and x′ are twins if and only
if µx(T ) = µx′(T ) for all T ∈ T .
Let Ψ ⊂ Φ be a countable dense set, and for every ψ ∈ Ψ, T ∈ T and rational number r let
Sψ,T,r := {x ∈ Ω : ψ(µx(T )) ≥ r}. These are countably many and clearly do not separate
twins. If however x and x′ are not twins, then for some T ∈ T we have µx(T ) 6= µx′(T ). Since
Ψ separates Z, we can then find ψ ∈ Ψ and r ∈ Q such that Sψ,T,r separates x and x′.
Corollary 17. Every graphon W ∈ W(Ω,A,π) has a weak isomorphism into an almost twin-free
strong Lebesguian Z-valued graphon fW . In addition kW kp =(cid:13)(cid:13)(cid:13)fW(cid:13)(cid:13)(cid:13)p
Proof. First let us change the graphon W ∈ W(Ω,A,π) on a null-set to obtain a strong graphon
W1 applying (i) of the previous theorem. Then the identity map on the underlying probability
space (Ω, A, π) will be a A−A-measurable weak isomorphism τ1 from W to W1. Now by points
(ii)-(iv) the completion of the graphon W1 has an embedding τ2 into a Lebesguian graphon
W3, which in turn by point (v) can be projected onto its almost twin-free Lebesguian form
fW ∈ W( eΩ, eA,eπ) through a weak isomorphism τ3. Since the composition τ := τ3 ◦ τ2 ◦ τ1 is a
A − eA-measurable mapping between (Ω, A, π) and the Lebesgue space (eΩ, eA,eπ), it is indeed
a weak isomorphism from W into fW . Concerning the last assertion, note that none of the
transformations in the previous theorem actually changed the norm of the graphon.
for all 1 ≤ p < ∞.
12
4 Anchored graphons
As mentioned in the introduction, two-variables functions on a product measure space Ω × Ω
lack a canonical form, or a canonical reordering of the underlying product space. To counter
this, Borgs, Chayes and Lovász made in [3] use of countable sequences of random elements of Ω
and defined what they called canonical ensembles. In essence this means that instead of having
a single well-defined canonical form, one rather has a whole family of functions, together with
a probability measure on said family. This pair contains all the relevant information on the
original function, and is easier to handle than the original single function. In addition these
new functions all live on the same σ-algebra, independently of the measure space of the original
graphon, making it possible to compare these random representations.
First we shall need a few results that guarantee that with probability 1, the randomness we
wish to introduce does not interfere with measurability.
Lemma 18. Let (Ω, A, π) and (Ω′, A′, π′) be probability spaces, and let W : Ω × Ω′ → Z be a
weak-* measurable function with respect to A × A′ such that kW k ∈ L2(Ω × Ω′). Let further
Y1, Y2, . . . be independent random points from Ω′, and A0 ⊆ A the random σ-algebra generated
by the functions W (·, Yk). Then with probability 1, W is almost weak-* measurable with respect
to A0 × A′.
Proof. By Lemma 12, we may assume that both A and A′ are countably generated. Let
A1 ⊂ A2 ⊂ . . . and A′
2 ⊂ . . . be sequences of finite σ-algebras such that σ(∪nAn) = A
and σ(∪nA′
n. For each
y ∈ S ∈ P ′
n denote the partition of Ω′ into the atoms of A′
n) = A′, and let P ′
n with π′(S) > 0 let
1 ⊂ A′
Un,m(x, y) :=
1
mπ′(S) Xj≤m
Yj∈S
W (x, Yj),
whilst Un,m(x, y) := 0 whenever π′(S) = 0. Let Ψ ⊂ Φ be a countable dense subset of the
predual of Z.
We first wish to prove that for every ψ ∈ Ψ, n ≥ 1, every A ∈ ∪nAn and A′ ∈ A′
with probability 1 that
n we have
ZA×A′
Uψ
n,m dπdπ′ m→∞
−−−−→ZA×A′
Wψ
dπdπ′.
(2)
n is atomic, it is enough to show this for sets A′ = S ∈ P ′
n with π′(S) > 0. For every
Since A′
y0 ∈ S we then have
ZA
Uψ
n,m(x, y0) dπ(x) =
1
mπ′(S) Xj≤m
Yj∈S
ZA
Wψ
(x, Yj)dπ(x).
Since kW k lies in L2 ⊂ L1, the function Wψ
Numbers to obtain
lies in L1, hence we may apply the Law of Large
ZA
Uψ
n,m(x, y0) dπ(x)
m→∞
−−−−→
1
π′(S)ZA×S
Wψ
dπdπ′.
13
Since both sides are independent of the choice of y0 ∈ S, we may integrate it out over S to
obtain equation (2).
We have a countable number of choices for ψ ∈ Ψ, n ∈ N+, A′ ∈ A′ and A ∈ ∪kAk, hence
with probability 1 equation (2) holds for all ψ ∈ Ψ, all n ∈ N+, A′ ∈ A′
n and A ∈ ∪kAk. Also,
note that
whence by AM-QM,
kUn,m(x, y)k ≤
kUn,m(x, y)k2 ≤
1
mπ′(S) Xj≤m
Yj ∈S
1
mπ′(S)2 Xj≤m
Yj ∈S
kW (x, Yj)k,
kW (x, Yj)k2.
For each y0 ∈ S ∈ P ′
n with π′(S) > 0 we then have
ZΩ
kUn,m(x, y0)k2 dπ(x) ≤
1
π′(S)2
1
m Xj≤m
Yj∈S
ZΩ
kW (x, Yj)k2dπ(x).
(3)
Again inequality (3) is independent of the choice of y0 ∈ S, so we may integrate over it to
obtain
ZΩ×S
kUn,m(x, y0)k2 dπ(x)dπ′(y0) ≤
1
π′(S)
1
m Xj≤m
Yj∈S
ZΩ
kW (x, Yj)k2dπ(x).
Since by assumption kW k lies in L2, we may again apply the Law of Large Numbers to the
right hand side to obtain that with probability 1
1
π′(S)
1
m Xj≤m
Yj∈S
ZΩ
kW (x, Yj)k2dπ(x)
m→∞
−−−−→
1
π′(S)ZΩ×S
kW k2dπdπ′ < ∞.
Since there are finitely many S ∈ P ′
probability 1, the sequence (Un,m)m∈N+ is uniformly bounded in L2-norm for each n ∈ N+.
n with π′(S) > 0, summing over all such S yields that with
From now on assume that the choice of the Yj is such that equation (2) holds for all ψ ∈ Ψ, all
n ∈ N+, A′ ∈ A′
n and A ∈ ∪nAn, and that the sequence (Un,m)m∈N+ is uniformly bounded in
L2-norm for each n ∈ N+. For a fixed n, by compactness and since there are countably many
elements in Ψ, we may by a diagonal argument choose a subsequence (mk)k∈N+ ⊂ N+ such
that for each ψ ∈ Ψ the sequence ( Uψ
n, R). Denote
the weak limits by Un,ψ. Since by construction each Uψ
n-measurable, the same
is true for the weak limits Un,ψ.
n,mk )k∈N+ converges weakly in L2(A × A′
n,mk is A0 × A′
By equation (2) we have that for each n ≥ 1 and ψ ∈ Ψ,
ZA×A′
Un,ψ dπdπ′ = lim
k→∞ZA×A′
Uψ
n,mk dπdπ′ =ZA×A′
Wψ
dπdπ′
= ZA×A′
E( Wψ
A × A′
n) dπdπ′
14
n) lie in L2(A ×
n, R), and since ∪nAn is dense in A, the above equality is in fact true for all A ∈ A. This
n) represent the same element in L2(A ×
for all A ∈ ∪nAn and A′ ∈ A′
A′
means that the functions Un,ψ and E( Wψ
A′
n. But both functions Un,ψ and E( Wψ
A × A′
n, R). Since Un,ψ is A0 × A′
n-measurable, this in turn means that for every ψ ∈ Ψ
A × A′
E( Wψ
A0 × A′
n) = E( Wψ
A × A′
n)
(4)
in L2(A × A′
to E( Wψ
for each ψ ∈ Ψ we have that
n, R). By Levy's Upward Theorem, however, the left hand side of (4) tends a.e.
A × A′). Thus
A0 × A′) as n → ∞, whilst the right hand side tends a.e. to E( Wψ
E( Wψ
A0 × A′) = E( Wψ
A × A′) = Wψ
in L2(A × A′, R), meaning that Wψ
is almost A0 × A′-measurable for all ψ ∈ Ψ. By Lemma
9, and since Ψ is countable, it then follows that W is almost weak-* measurable with respect
to A × A′.
The following is an immediate consequence of the previous theorem.
Corollary 19. Let (Ω, A, π) and (Ω′, A′, π′) be probability spaces, and let W : Ω × Ω′ → Z be
a weak-* measurable function with respect to A × A′ with kW k ∈ L2. Let further A0 ⊂ A be
a sub-σ-algebra. If W (·, y) is weak-* measurable with respect to A0 for almost all x ∈ Ω, then
W is almost weak-* measurable with respect to A0 × A′.
Applying the previous corollary twice, we obtain the following result, which is the key element
in our randomization.
Corollary 20. Let W ∈ W(Ω,A,π) be a strong graphon, and let X1, X2, . . . be independent ran-
dom points from Ω. Let A0 ⊂ A be the random σ-algebra generated by the functions W (·, Xk),
k ∈ N+. Then with probability 1, W is almost weak-* measurable with respect to A0 × A0.
Proof. The proof of [3, Corollary 4.3] works also in this more general setting, with minor
natural modifications.
Let now W ∈ W(Ω,A,π) be a strong graphon and α = (α1, α2, . . .) an infinite sequence of points
in Ω. Consider the map x 7→ Γα(x) defined by
Γα(x) := (( Wψ
(x, α1))ψ∈Ψ, ( Wψ
=: R.
(5)
(x, α2))ψ∈Ψ, . . .) ∈(cid:0)RΨ(cid:1)N+
Since each Wψ
(·, αj ) is measurable by assumption, Γα is a measurable map from Ω into R
with respect to the standard Borel σ-algebra K on the product space, and thereby defines a
push-forward measure κα on K through
κα(S) = π(Γ−1
α (S)),
(6)
for S ∈ K. Denote by L the completion of K with respect to κα, and λα the extension of κα
to L.
Let further WΓα be the push-forward of W under Γα, and denote by Wα ∈ W(R,L,λα) its
completion.
15
Definition 21. The graphon Wα is called the anchored graphon with respect to the anchor
sequence α. An anchor sequence α is called regular if W = (WΓα)Γα almost everywhere.
Remark 22. Since it is the product of countably many copies of R, R itself is a complete
Polish space, hence Wα is always a Lebesguian graphon.
Lemma 23. Almost all α ∈ ΩN+
are regular.
Proof. Let Aα denote the pullback of the σ-algebra K under the map Γα. Since Γα is mea-
surable, we have Aα ⊂ A. Also, by construction Aα is the smallest σ-algebra such that all
of the functions Wψ
(·, αj) (ψ ∈ Ψ, 1 ≤ j) are measurable. We may thus apply Lemma 18 to
obtain that for almost all α, W is almost weak-* measurable with respect to Aα × Aα. But by
the construction given in the proof of Lemma 14 it immediately follows that W = (WΓα)Γα
a.e..
5 Densities and coupling of anchored graphons
Now that we have introduced random canonical forms for graphons, our aim is to show that
on the one hand two weakly isomorphic graphons have essentially the same random canonical
form, and on the other hand that this random canonical form is determined by the moment
functions. The main helping tool here are densities stemming from so-called partially labeled
graphs. These provide the necessary randomness that will act as a bridge between the usual
deterministic densities and the random anchored graphons, and allow us to find appropriate
couplings between the anchors. We present the necessary notions and properties necessary in
the context of decorated graphs, and refer to [11] for futher properties and applications of
partial labeling.
Let F = (F, f ) be a decorated graph. A partial labeling of F is an injective map from a subset
of the vertices of F into the set N+ of positive integers. If the injective map has as image the
set {1, 2, . . . , k}, the partially labeled decorated graph is called k-labeled. To simplify notation,
the case k = 0 corresponds to unlabeled decorated graphs.
Two partially labeled decorated graphs F1 and F2 are isomorphic if there exists a graph
isomorphism between F1 and F2 that preserves both the labels and the decorations. The
product F1F2 of two partially labeled decorated graphs F1 and F2 is itself a partially labeled
decorated graph, defined as follows: take the disjoint union of F1 and F2, then merge the
vertices that have identical labels, whilst keeping the labels and decorations as well as any
multiple edges that may arise.
Next, we define marginals induced by partial labelings. Supose F = (F, f ) is a partially
labeled Φ-decorated graph with vertex set V (F ) = {v1, . . . , vk}, where the vertices v1, . . . , vr
are labeled by the positive integers ℓ1, . . . , ℓr, and the remaining vertices are unlabeled. Given
a graphon W ∈ W(Ω,A,π) and an infinite sequence β = (β1, β2, . . .) in Ω, for 1 ≤ j ≤ r set
Xj := βℓj , and let Xr+1, . . . , Xk be independent random points of Ω from the distribution π.
Then we may define the marginal
tβ(F, W ) := E
Yvivj ∈E(F )
Wfvivj
(Xi, Xj )
.
16
This marginal only depends on the finitely many elements of β whose index appears as label,
so it will be convenient to sometimes omit the tail of β containing no labels.
If F1 and F2 are two k-labeled graphs, it can easily be seen that
t(F1F2) =ZΩk
tx1...xk (F1, W )tx1...xk (F2, W )dπ(x1) . . . dπ(xk).
We shall show how anchor sequences can be used to prove almost everywhere equality of
appropriate graphons. Having to involve densities of labeled multigraphs is a technical necessity
of the approach introduced in [3], as we wish to prove equality of measures through equality
of moments, and higher mixed moments in this context naturally correspond to densities of
multigraphs rather than those of simple graphs. As we shall later see, this is not going to be
a hindrance.
Given a countably generated dense subspace Ψ ∈ Φ of decorations, let F ∗
k denote the set
of k-labeled Ψ-decorated multigraphs with no edge between labeled vertices, and let F ∗ :=
∪n=0F ∗
k .
Lemma 24. Let W ∈ W(Ω,A,π) and W ′ ∈ W(Ω′,A′,π′) be two strong graphons, and let α and
β be regular anchor sequences in Ω for W and Ω′ for W ′, respectively. Suppose that for every
partially labeled Φ-decorated multigraph F ∈ F ∗, we have
tα(F, W ) = tβ(F, W ′).
Further suppose that for some countable dense subset Ψ ⊂ Φ the p-norms of W and of its
αj-sections satisfy:
∞Xn=1
k Wψ
(X, αj )k−1
2n = ∞ and
∞Xn=1
k Wψ
k−1
2n = ∞
(7)
for all j ∈ N+ and ψ ∈ Ψ. Then the anchored graphons Wα ∈ W(R,L,λα) and W ′
satisfy λα = λ′
β almost everywhere with respect to the common measure.
β, and Wα = W ′
β ∈ W(R,L,λ′
β)
Proof. Let us first show that λα = λ′
β. Recall that λα is the distribution measure of the
(Y, α1))ψ∈Ψ, ( Wψ
(X, α1))ψ∈Ψ, ( Wψ
point with distribution π, whilst λα is the distribution measure of the random variable sequence
random variable vector (cid:0)( Wψ
(cid:0)( Wψ
(X, α2))ψ∈Ψ, . . .(cid:1), where X ∈ Ω is a random
(Y, α2))ψ∈Ψ, . . .(cid:1), where Y ∈ Ω′ is a random point with distribution π′.
Each of these random variables is real-valued, and we shall first show that their mixed mo-
ments are all equal. Let therefore (kn,ψ)n∈N+,ψ∈Ψ be a double-indexed sequence of nonnegative
integers, with only finitely many non-zero elements. Let m ∈ N+ be such that kn,ψ = 0 for all
n > m, and construct the partially labeled Ψ-decorated multigraph F ∈ F ∗ on m + 1 vertices
as follows. First label all but one vertex with the help of the labels 1, . . . , m. Then for each
1 ≤ j ≤ m and ψ ∈ Ψ consider the vertex j and the unlabeled vertex, and add kj,ψ edges
decorated by ψ between them. Then by construction
E
Yn∈N+,ψ∈Ψ
Wψ
(X, αn)knψ
= tα(F, W ).
17
Similarly
E
Yn∈N+,ψ∈Ψ
Wψ
′(Y, βn)kn,ψ
= tβ(F, W ′).
These two are by assumption equal, so all mixed moments are indeed the same. If these were
all bounded variables, it would immediately follow that they are equal, as the mixed moments
would uniquely determine their joint distribution. In the unbounded case, however, we need
an extra property to guarantee uniqueness. By [8, Cor. 3a], it is enough that the variables
pertaining to W each separately satisfy the Carleman condition, i.e., the family of conditions
(7). Hence λα = λ′
β.
Now we proceed to show that Wα(x, y) = W ′
β(x, y) almost everywhere. To this end we wish to
show that the random variables U1 = (X, Y, Wα(X, Y )) and U2 = (X, Y, W ′
β(X, Y )) have the
same distribution, where X and Y are independent random points in (R, λα). By definition
of the distributions on (R, λα) we can generate X and Y by taking the random independent
points X ′, Y ′ from (Ω, π) and letting X := Γα(X ′) and Y := Γα(Y ′). Since α is regular for
W , we have with probability 1 that Wα(X, Y ) = W (X ′, Y ′), and hence
U1 = (( Wψ
( Wψ
W (X ′, Y ′)).
(X ′, α1))ψ∈Ψ, ( Wψ
(Y ′, α1))ψ∈Ψ, ( Wψ
(X ′, α2))ψ∈Ψ, . . . ,
(Y ′α2))ψ∈Ψ, . . . ,
For W ′ we correspondingly take X ′′ and Y ′′ from (Ω′, π′) instead, and obtain with probability
1 that
U2 = (( Wψ
( Wψ
W ′(X ′′, Y ′′)).
′(X ′′, β1))ψ∈Ψ, ( Wψ
′(Y ′′, β1))ψ∈Ψ, ( Wψ
′(X ′′, β2))ψ∈Ψ, . . . ,
′(Y ′′β2))ψ∈Ψ, . . . ,
To compare these two vectors, however, we have to replace their last coordinate with a sequence
of real-valued variables with the help of the elements of Ψ.
Let therefore
and
bU1 := (( Wψ
( Wψ
( Wψ
(X ′, α1))ψ∈Ψ, ( Wψ
(Y ′, α1))ψ∈Ψ, ( Wψ
(X ′, Y ′))ψ∈Ψ)
(X ′, α2))ψ∈Ψ, . . . ,
(Y ′α2))ψ∈Ψ, . . . ,
bU2 := (( Wψ
( Wψ
( Wψ
′(X ′′, β1))ψ∈Ψ, ( Wψ
′(Y ′′, β1))ψ∈Ψ, ( Wψ
′(X ′′, Y ′′))ψ∈Ψ).
′(X ′′, β2))ψ∈Ψ, . . . ,
′(Y ′′β2))ψ∈Ψ, . . . ,
Each mixed moment is determined by nonnegative integers (an,ψ)n∈N+,ψ∈Ψ, (bn,ψ)n∈N+,ψ∈Ψ
and (cψ)ψ∈Ψ, such that only a finite number of them is non-zero. Assume for instance that
18
an,ψ = bn,ψ = 0 for all n > m for some appropriate integer m. Let us define a partially labeled
Ψ-decorated multigraph F ∈ F ∗ on m + 2 vertices as follows. First label all but two vertices
with the labels 1, . . . , m, and denote the remaining two by vx and vy. Then for each ψ ∈ Ψ
add cψ parallel edges between vx and vy, decorating them with ψ. Finally for each 1 ≤ n ≤ m
and ψ ∈ Ψ add an,ψ parallel edges decorated with ψ between vx and the vertex with label n,
and bn,ψ parallel edges decorated with ψ between vy and the vertex with label n. Then we
have
E
Yn∈N+,ψ∈Ψ
= tα(F, W ),
Wψ
(X ′, αn)an,ψ · Yn∈N+,ψ∈Ψ
Wψ
(Y ′, αn)bn,ψ · Yψ∈Ψ
Wψ
and
E
Yn∈N+,ψ∈Ψ
= tβ(F, W ′),
Wψ
′(X ′′, βn)an,ψ · Yn∈N+,ψ∈Ψ
Wψ
′(Y ′′, βn)bn,ψ · Yψ∈Ψ
Wψ
(X ′, Y ′)cψ
′(X ′′, Y ′′)cψ
which are by assumption equal, and using the conditions in (7), we are done.
Remark 25. Note that in condition (7), the second part implies the first for almost all
sequences α. Indeed, for 0 < c let Sc ⊂ Ω be the set of points y ∈ Ω such that
∞Xn=1
k Wψ
(·, y)k−1
2n ≤ c.
Suppose that π(Sc) > 0. We then for n ≥ 1 have by the Hölder inequality that
k Wψ
k2n ≥ (cid:18)ZSc(cid:18)ZΩ(cid:12)(cid:12)(cid:12) Wψ
1
≥
≥
=
Thus by convexity
dπ(y)
dπ(y)
2n
2n
dπ(x)(cid:19) dπ(y)(cid:19)1/2n
(x, y)(cid:12)(cid:12)(cid:12)
dπ(x)(cid:19)1/2n
π(Sc)1−1/2n ZSc(cid:18)ZΩ(cid:12)(cid:12)(cid:12) Wψ
(x, y)(cid:12)(cid:12)(cid:12)
dπ(x)(cid:19)1/2n
π(Sc)1/2 ZSc(cid:18)ZΩ(cid:12)(cid:12)(cid:12) Wψ
(x, y)(cid:12)(cid:12)(cid:12)
π(Sc)1/2 ZSc(cid:13)(cid:13)(cid:13) Wψ
(·, y)(cid:13)(cid:13)(cid:13)2n
pπ(Sc)
dπ(y).
≤
1
1
k Wψ
(·, y)k2n dπ(y)
2n
1
(·, y)k2n
k Wψ
dπ(y),
1
k Wψ
k2n
RSc
≤ pπ(Sc)ZSc
19
and so for all N ∈ N
NXn=1
1
k Wψ
k2n
ZSc
NXn=1
1
(·, y)k2n
1
(·, y)k2n
k Wψ
k Wψ
dπ(y)
dπ(y)
NXn=1
≤ pπ(Sc)
= pπ(Sc) ZSc
≤ cpπ(Sc)
3
,
leading to a contradiction. But if π(Sc) = 0 for all c = 0, then the sum
∞Xn=1
k Wψ
(·, y)k−1
2n
is infinite for almost all y ∈ Ω, and therefore also for almost every infinite sequence α.
Our next lemma shows that under a Carleman-type set of conditions, equality of multigraph
homomorphism densities, the random canonical forms can be coupled in such a way as to have
the corresponding marginals all equal.
The reason for us not wanting to have edges between labeled vertices in the test-graphs involved
in the coupling is that their absence significantly simplifies and improves the upper bounds on
the mixed moments, and thus weaker Carleman-type conditions will suffice. Luckily this is not
a restriction here, as the multigraph constructions arising from the mixed moments preserve
this property.
Lemma 26. Let W ∈ W(Ω,A,π) and W ′ ∈ W(Ω′,A′,π′) be two strong Lebesguian graphons such
that
t(F, W ) = t(F, W ′)
for every unlabeled Ψ-decorated multigraph F. Further suppose that the p-norms of W satisfy
the Carleman type conditions:
∞Xn=1
kW k−k
2nk = ∞
for all k ∈ N+. Then we can couple sequences α ∈ ΩN+
(α, β) is taken from the joint distribution, then with probability 1
with sequences β ∈ Ω′N+
such that if
for every partially labeled Ψ-decorated multigraph F ∈ F ∗.
tα(F, W ) = tβ(F, W ′)
Proof. We shall recursively define a coupling of sequences γ ∈ Ωk and δ ∈ Ω′k such that almost
k have that tγ(F, W ) = tδ(F, W ′). This is trivial to do for k = 0. Let
surely we for all F ∈ F ∗
us now assume we have such a coupling for sequences of length k, and let (γ1, . . . , γk) and
(δ1, . . . , δk) be chosen from this coupled distribution. Let further X be a random point from
(Ω, π), and Y a random point from (Ω′, π′), and define the random variables
C := (tγ1...γkX(F, W ))F∈F ∗
k+1
20
and
D := (tδ1...δkY (F, W ′))F∈F ∗
k+1
with values in RF ∗
k+1. Our aim is to show that they have the same distribution.
To prove that the joint distributions are equal, we first show that their mixed moments coin-
cide, and then prove that the coordinates of C satisfy the Carleman condition.
Let m ∈ N+, F1, . . . , Fm ∈ F ∗
j ≤ m define F
defined as the product Fq1
Then we have that the corresponding moment of C satisfies
1 . . . Fqm
k+1 as the qj-fold product of Fj with itself. Let further F ∈ F ∗
k+1, and let q1, . . . , qm be non-negative integers. Then for 1 ≤
k be
m with label k + 1 removed from the corresponding vertex.
qj
j ∈ F ∗
tγ1...γkX (Fj, W )qj
E
mYj=1
= E (tγ1...γkX(Fq1
1 . . . Fqm
m , W )) = tγ1...γk (F, W ).
Similar arguments yield that the corresponding mixed moment of D is tδ1...δk (F, W ′), which
by hypothesis is the same.
We now have to show that each coordinate of C satisfies the Carleman condition. Let F ∗
k+1 ∋
F = (F, f ) and consider CF(X) := tγ1...γkX(F, W ). Since F does not have any edges between
labeled vertices, we have
CF(X) = Z V (F )Ya=k+2
Y1≤j≤k
k+2≤a≤V (F )
mult(vk+1va)Yi=1
f(vk+1va)i
W
mult(vj va)Yi=1
f(vj va)i
f(vavb)i
(X, xa)
(γj, xa)
W
V (F )Ya,b=k+2
mult(vavb)Yi=1
dxk+2 . . . dxV (F )
W
(xa, xb)
For almost all anchor sequences γ we have thatRΩ W (γj, x)dx is finite for all j. By the Hölder
inequality (using that π(Ω) = 1) we thus have
CF(X) ≤ C
V (F )Ya=k+2
mult(vk+1va)Yi=1
(cid:18)ZΩ
kW (X, xa)kE(F ) dxa(cid:19) 1
E(F )
,
where C is a finite constant depending on W , F and γ1, . . . , γk. Consequently, by further
21
applications of Hölder's inequality we have
kCFkn
≤ C n
(cid:18)ZΩ
n ≤ C nZΩ
mult(vk+1va)Yi=1
V (F )Ya=k+2
ZΩ(cid:18)ZΩ
mult(vk+1va)Yi=1
V (F )Ya=k+2
(cid:18)ZΩ(cid:18)ZΩ
mult(vk+1va)Yi=1
V (F )Ya=k+2
(cid:18)ZΩZΩ
mult(vk+1va)Yi=1
V (F )Ya=k+2
mult(vk+1va)Yi=1
V (F )Ya=k+2
≤ C n
= C n
= C n
1
E(F )
dX
E(F )
E(F )
kW (X, xa)kE(F ) dxa(cid:19) n
kW (X, xa)kE(F ) dxa(cid:19) nE(F )
kW (X, xa)kE(F ) dxa(cid:19)n
kW (X, xa)knE(F ) dxa dX(cid:19) 1
dX
dX(cid:19) 1
E(F )
E(F )
kW kn
nE(F ) ≤ C n kW knE(F )
nE(F ) .
Therefore with K := E(F ) we have
∞Xn=1
kCFk−1
2n ≥
1
C
∞Xn=1
kW k−K
2nK = ∞,
and so the Carleman condition is indeed satisfied.
Thus C and D have the same distribution, and by [3, Lemma 6.2] we can then couple X and
Y so that with probability 1 we have C = D. Thus there exist random variables X ′ from
(Ω, π) and Y ′ from (Ω′, π′) such that the joint distribution of (X ′, Y ′) ∈ (Ω, Ω′) satisfies with
probability 1 that
tγ1...γkX ′(F, W ) = tδ1...δkY ′(F, W ′)
for every F ∈ Fk+1. This extends our coupling to one between Ωk+1 and Ω′k+1. Iterating, we
obtain the desired coupling between ΩN+
and Ω′N+
.
Our final lemma is to show that graphons having equal simple graph densities also have equal
multigraph densities, and hence no generality was lost in the assumptions of the previous
results. Note that labeled graphs play in the below proof a different role than above, and as
such it is not an issue that we here allow (multiple) edges between labeled vertices.
Lemma 27. Let W1 ∈ W(Ω1,A1,π1) and W2 ∈ W(Ω2,A2,π2) be two countably generated graphons,
and assume that t(F, W1) = t(F, W2) for every simple Φ-decorated graph F. Then t(F, W1) =
t(F, W2) for every Φ-decorated multigraph F = (F, f ).
Proof. We shall proceed by induction on the number of parallel edges in F . The base case is
given by the assumption. Let vi and vj be two vertices connected by more than one edge, and
let ϕ ∈ Φ be the decoration on one of them. Denote by F′ the decorated multigraph obtained
by deleting one ϕ-decorated edge between vi and vj. Let further Fk denote the decorated
multigraph obtained by adding a path of length k between vi and vj in F′, decorating each
edge in the path with ϕ. Thus F1 = F, but for each k > 1 the multigraph Fk has fewer
22
parallel edges than F. Hence by the inductive assumption we have that t(Fk, W1) = t(Fk, W2)
for each k > 1. Let us now label all the multigraphs Fk and F′ such that vi receives the label
1 whilst vj receives the label 2. Then Fk is the product of F′ with the path Pk+1 of length
(k + 1) with its two endpoints labeled 1 and 2 respectively and each edge decorated with ϕ,
and we may write
t(F, W1) =ZΩ2
t(Fk, W1) =ZΩ2
1
and
Note that
W (x, y)txy(F′, W1)dπ1(x)dπ1(y),
txy(Pk+1, W )txy(F′, W1)dπ1(x)dπ1(y).
1
txy(Pk+1, W ) =ZΩk−1
1
Wϕ
(x, x1) · · · Wϕ
(xk−1, y)dπ(x1) . . . dπ(xk−1).
Since Wϕ
and thus has a spectral decomposition
1 ∈ L2(Ω2
1, R), it is a self-adjoint Hilbert-Schmidt integral operator on L2(Ω1, C),
Wϕ
1(x, y) =
∞Xn,m=1
λn,mζn(x)ζm(y)
in the L2 sense, where (ζn) is an orthonormal system in L2(Ω1, C).
Then we obtain by induction on k that for every k > 1,
txy(Pk+1, W1) =
∞Xn,m=0
λk
n,mζn(x)ζm(y)
in L2, whereby
t(Fk, W1) =
∞Xn,m=1
nZΩ2
λk
ζn(x)ζm(y)txy(F′, W1)dπ1(x)dπ1(y),
since txy(F′, W1) lies in L2 by the assumption on W1.
Similarly, with the spectral decomposition
Wϕ
2(x, y) =
∞Xn,m=1
µn,mηn(x)ηm(y),
we obtain
0 = t(Fk, W1) − t(Fk, W2) =
∞Xn,m=1
for every k ≥ 2, where the parameters
an,mλk
n,m − bn,mµk
n,m
(8)
an,m :=ZΩ2
1
ζn(x)ζm(y)txy(F′, W1)dπ1(x)dπ1(y)
23
and
bn,m :=ZΩ2
2
are independent of k.
ηn(x)ηm(y)txy(F′, W2)dπ2(x)dπ2(y)
Since each non-zero eigenvalue has finite multiplicity and the only possible accumulation point
is 0, the asymptotic behaviour of the right hand side in (8) dictates that all the terms have
to cancel, i.e., for each c ∈ R\{0} we have that
Xλn,m=c
an,m = Xµn,m=c
bn,m.
Then
t(F, W1) =
∞Xn,m=1
λn,mZΩ2
similarly
ζn(x)ζm(y)txy(F′, W1)dπ1(x)dπ1(y) =
∞Xn,m=1
an,mλn,m,
t(F, W2) =
∞Xn,m=1
bn,mµn,m,
and the claim follows.
Remark 28. By Lemma 12, countable generation is not actually needed.
6 Proof of main theorem
Having extended the intermediate steps of [3] to the significantly more general context of our
investigations, we are now ready to bring together the elements of the previous sections to
prove our main result.
Proof of Theorem 11. Part (i): By Lemma 27, equality of simple graph densities implies
equality of multigraph densities, and thus we may apply Lemma 26 to our two graphons. If
we choose the anchor sequences α ∈ ΩN+
from the joint distribution given by
Lemma 26, almost all such choices satisfy
and β ∈ Ω′N+
tα(F, W ) = tβ(F, W ′)
for every partially labeled Ψ-decorated multigraph F ∈ F ∗, and but for a further null-set they
yield anchor sequences that are regular (Lemma 23). Hence, taking into consideration Remark
25, we can choose sequences α and β that satisfy the conditions of Lemma 24. Consequently
the anchored graphons Wα and W ′
β are isomorphic mod 0 through the identity map. If we now
could show that Γα is an isomorphism mod 0 between W and Wα, and that similarly Γβ is an
isomorphism mod 0 between W ′ and W ′
β, our proof would be complete. Due to symmetry we
shall only show the first isomorphism.
24
First note that since (Ω, A, π) is a Lebesgue space, the mapping Γα is not only measurable and
measure preserving as a mapping (Ω, A, π) → (R, K, κα), but also as a mapping (Ω, A, π) →
(R, L, λα). Let
S := {x ∈ Ω : Wα(Γα(x), Γα(y)) = W (x, y) for almost all y}.
Since α is a regular anchor sequence, we have π(S) = 1. Also, because W is almost twin-
free, we can find a null-set N ⊂ Ω such that each twin-class of W has at most one point
in T := S\N . Let Γ′
α be the restriction of Γα to T . Then it can easily be seen that T is
injective. By [15, Section 2.5], injective measure preserving maps between Lebesgue spaces
have an almost everywhere defined measurable inverse. Thus Γ′
α is an isomorphism mod 0,
and then so is Γα.
Part (ii): First, by Corollary 17 we can find two almost twin-free strong Lebesguian graphons
U ∈ W(O,B,ρ) and U ′ ∈ W(O′,B′,ρ′) and corresponding weak isomorphisms γ and γ′ from W and
W ′ to U and U ′, respectively. By part (i) we then have that U and U ′ are isomorphic mod 0,
hence for some measure preserving map η : O → O′ we have U = (U ′)η. Now let µ : Ω → O′
be defined through µ(x) := η(γ(x)). Then clearly W = (U ′)µ almost everywhere, and the
maps µ and γ′ are measure preserving from the completions of W and W ′, respectively, into
U ′.
7 A moment-indeterminate graphon
In this last section we wish to provide an example of two real-valued graphons that possess
the same homomorphism densities, but are not weakly isomorphic. In fact, in our example
they are not inducing the same probability measure on R, so they are even distinguishable
when forgetting about the geometry coming from the underlying product space.
Let σ and τ be two probability distributions on N with finite moments and having the same
moments (such distributions exist, see e.g. [14, Cor. 3.4]). Denote their n-th moments by Mn
(n ≥ 0).
Let further {Si}i∈N and {Tj}j∈N be two partitions of [0, 1] into measurable sets such that
λ(Si) = σ({i}) and λ(Tj) = τ ({j}) for all i, j ∈ N. Consider the functions fσ, fτ : [0, 1] → R
defined by
fσ(x) := nx whenever
fτ (x) := mx whenever
x ∈ Snx,
x ∈ Tmx,
respectively, and let Wσ, Wτ : [0, 1]2 → R be defined by Wσ(x, y) := fσ(x)fσ(y) and Wτ (x, y) =
fτ (x)fτ (y), respectively.
Let F be an R-decorated graph with each edge decorated with 1. By linearity, it is enough
to show that t(F, Wσ) = t(F, Wτ ) for every such F. Let the elements of V (F ) be denoted
by v1, v2, . . . , vk, and let di denote the degree of vertex vi. It can then easily be seen that we
25
have
t(F, Wσ) =
Z
x1,...,xk∈[0,1] Yvivj ∈E(F )
Z
x1,...,xk∈[0,1] Yvivj ∈E(F )
Z[0,1]
kYi=1
Similar calculations yield t(F, Wσ) =Qk
exact same generalized moments.
=
=
W (xi, xj)dx1 . . . xk
fσ(xi)fσ(xj)dx1 . . . xk
fσ(xi)didxi =
kYi=1
Mdi.
i=1 Mdi, and so the two graphons do indeed have the
It remains to be shown that Wσ and Wτ yield different probability measures on R, but this
easily follows from their product form, and the fact that σ 6= τ .
References
[1] N.I. Akhiezer, The Classical Moment Problem and Some Related Questions in Analysis,
Oliver & Boyd (1965).
[2] C. Berg, The cube of a normal distribution is indeterminate, Ann. Probab. 16 (1988),
910 -- 913.
[3] C. Borgs, J. Chayes, L. Lovász, Moments of Two-Variable Functions and the Uniqueness
of Graph Limits, Geom. funct. anal. 19 (2010) 1597 -- 1619.
[4] C. Borgs, J.T. Chayes, L. Lovász, V.T. Sós and K. Vesztergombi: Convergent Graph
Sequences I: Subgraph frequencies, metric properties, and testing, Advances in Math.
219 (2008), 1801 -- 1851.
[5] C. Borgs, J.T. Chayes, L. Lovász, V.T. Sós and K. Vesztergombi: Convergent Graph
Sequences II: Multiway Cuts and Statistical Physics, Annals of Math. 176 (2012), 151 --
219.
[6] G. Elek and B. Szegedy: A measure-theoretic approach to the theory of dense hyper-
graphs, Adv. in Math. 231 (2012), 1731 -- 1772.
[7] P. R. Halmos: Measure Theory, Graduate texts in Mathematics 18, Springer, New York,
Heidelberg, Berlin (1991).
[8] C. Kleiber, J. Stoyanov, Multivariate distributions and the moment problem, Journal of
Multivariate Analysis 113 (2013) 7 -- 18
[9] D. K.-K., L. Lovász, B. Szegedy, Multigraph limits, unbounded kernels, and Banach space
decorated graphs. Submitted, http://arxiv.org/abs/1406.7846
[10] G. Lin and J. Huang, The cube of the logistic distribution is indeterminate, Austral. J.
Statist. 39 (1997), 247 -- 252.
26
[11] L. Lovász, Large graphs, graph homomorphisms and graph limits, AMS (2012).
[12] L. Lovász and B. Szegedy: Limits of dense graph sequences, J. Combin. Theory B 96
(2006), 933 -- 957.
[13] L. Lovász and B. Szegedy: Szemerédi's Lemma for the analyst, Geom. funct. anal. 17
(2007), 252 -- 270.
[14] H. L. Pedersen: On Krein's Theorem for Indeterminacy of the Classical Moment Problem
J. Approx. Theory 95 (1998), 90 -- 100.
[15] V. A. Rohlin, On the fundamental ideas of measure theory, Translations of the American
Mathematical Society, Series 1, Vol. 10 (1962) 1 -- 54
(Russian original in Math Sb. 25 (1949) 107 -- 150).
[16] V. I. Rybakov, Vector-valued measures (in Russian), Izv. Vyssh. Uchebn. Zaved. Mat. 12
(1968), 92 -- 101.
[17] J. Stoyanov, Counterexamples in Probability. Wiley & Sons, 2nd ed. (1997)
27
|
1304.0157 | 1 | 1304 | 2013-03-31T05:54:14 | Operator inequalities of Jensen type | [
"math.FA",
"math.OA"
] | We present some generalized Jensen type operator inequalities involving sequences of self-adjoint operators. Among other things, we prove that if $f:[0,\infty) \to \mathbb{R}$ is a continuous convex function with $f(0)\leq 0$, then {equation*} \sum_{i=1}^{n} f(C_i) \leq f(\sum_{i=1}^{n}C_i)-\delta_f\sum_{i=1}^{n}\widetilde{C}_i\leq f(\sum_{i=1}^{n}C_i) {equation*} for all operators $C_i$ such that $0 \leq C_i\leq M \leq \sum_{i=1}^{n} C_i $ \ $(i=1,...,n)$ for some scalar $M\geq0$, where $ \widetilde{C_i} = 1/2 - |\frac{C_i}{M}- 1/2 |$ and $\delta_f = f(0)+f(M) - 2 f(\frac{M}{2})$. | math.FA | math |
OPERATOR INEQUALITIES OF JENSEN TYPE
M. S. MOSLEHIAN1, J. MI ´CI ´C2 AND M. KIAN3
Abstract. We present some generalized Jensen type operator inequalities
involving sequences of self-adjoint operators. Among other things, we prove
that if f : [0, ∞) → R is a continuous convex function with f (0) ≤ 0, then
nXi=1
f (Ci) ≤ f nXi=1
2 −(cid:12)(cid:12) Ci
Ci!
nXi=1 eCi ≤ f nXi=1
Ci! − δf
2(cid:12)(cid:12) and δf = f (0) + f (M ) − 2f(cid:0) M
2(cid:1).
for all operators Ci such that 0 ≤ Ci ≤ M ≤Pn
scalar M ≥ 0, wherefCi = 1
M − 1
i=1 Ci (i = 1, . . . , n) for some
1. Introduction and Preliminaries
Let B(H ) be the C ∗-algebra of all bounded linear operators on a complex Hilbert
space H and I denote the identity operator. If dim H = n, then we identify
B(H ) with the C ∗-algebra Mn(C) of all n × n matrices with complex entries.
Let us endow the real space Bh(H ) of all self-adjoint operators in B(H ) with
the usual operator order ≤ defined by the cone of positive operators of B(H ).
If T ∈ Bh(H ), then m = inf{hT x, xi : kxk = 1} and M = sup{hT x, xi :
kxk = 1} are called the bounds of T . We denote by σ(J) the set of all self-
adjoint operators on H with spectra contained in J. All real-valued functions
are assumed to be continuous in this paper. A real valued function f defined on an
interval J is said to be operator convex if f (λA+(1−λ)B) ≤ λf (A)+(1−λ)f (B)
for all A, B ∈ σ(J) and all λ ∈ [0, 1]. If the function f is operator convex, then
the so-called Jensen operator inequality f (Φ(A)) ≤ Φ(f (A)) holds for any unital
positive linear map Φ on B(H ) and any A ∈ σ(J). The reader is referred to
[3, 4, 8] for more information about operator convex functions and other versions
of the Jensen operator inequality. It should be remarked that if f is a real convex
function, but not operator convex, then the Jensen operator inequality may not
hold. To see this, consider the convex (but not operator convex) function f (t) = t4
defined on [0, ∞) and the positive mapping Φ : M3(C) → M2(C) defined by
2010 Mathematics Subject Classification. 47A63, 47A64, 15A60.
Key words and phrases. convex function, positive linear map, Jensen -- Mercer operator in-
equality, Petrovi´c operator inequality.
1
2
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
Φ((aij)1≤i,j≤3) = (aij)1≤i,j≤2 for any A = (aij)1≤i,j≤3 ∈ M3(C). If
A =
46 59 !
f (Φ(A)) = 36 46
then there is no relationship between
2 2 0
2 3 1
0 1 3
and
,
48 68 !
Φ(f (A)) = 36 48
in the usual operator order.
Recently, in [6] a version of the Jensen operator inequality was given without
operator convexity as follows:
Theorem A. [6, Theorem 1] Let (A1, . . . , An) be an n-tuple of operators Ai ∈
Bh(H ) with bounds mi and Mi, mi ≤ Mi, and let (Φ1, . . . , Φn) be an n-tuple of
positive linear mappings Φi on B(H ) such thatPn
(mC, MC) ∩ [mi, Mi] = Ø
i=1 Φi(I) = I. If
(1.1)
for all 1 ≤ i ≤ n, where mC and MC with mC ≤ MC are bounds of the self-adjoint
operator C =Pn
i=1 Φi(Ai), then
f nXi=1
Φi(Ai)! ≤
nXi=1
Φi (f (Ai))
(1.2)
holds for every convex function f : J → R provided that the interval J contains
all mi, Mi; see also [7].
Another variant of the Jesnen operator inequality is the so-called Jensen --
Mercer operator inequality [5] asserting that if f is a real convex function on
an interval [m, M], then
f M + m −
nXi=1
Φi(Ai)! ≤ f (M) + f (m) −
nXi=1
where Φ1, · · · , Φn are positive linear maps on B(H ) with Pn
A1, · · · , An ∈ σ([m, M]).
Φi(f (Ai)),
i=1 Φi(I) = I and
Recently, in [9] an extension of the Jensen -- Mercer operator inequality was
presented as follows:
Theorem B.[9, Corollary 2.3] Let f be a convex function on an interval J. Let
Ai, Bi, Ci, Di ∈ σ(J) (i = 1, · · · , n) such that Ai + Di = Bi + Ci and Ai ≤
m ≤ Bi, Ci ≤ M ≤ Di. Let Φ1, · · · , Φn be positive linear maps on B(H ) with
OPERATOR INEQUALITIES OF JENSEN TYPE
3
i=1 Φi(I) = I. Then
Pn
f nXi=1
Φi(Bi)! + f nXi=1
Φi(Ci)! ≤
nXi=1
Φi(f (Ai)) +
nXi=1
Φi(f (Di)).
(1.3)
The authors of [9] used inequality (1.3) to obtain some operator inequalities.
In particular, they gave a generalization of the Petrovi´c operator inequality as
follows:
Theorem C.[9, Corollary 2.5] Let A, D, Bi ∈ σ(J) (i = 1, · · · , n) such that
i=1 Bi and A ≤ m ≤ Bi ≤ M ≤ D (i = 1, · · · , n) for two real
numbers m < M. If f is convex on J, then
A + D = Pn
f (Bi) ≤ (n − 1)f(cid:18) 1
n − 1
A(cid:19) + f (D).
nXi=1
If f : [0, ∞) → R is a convex function such that f (0) = 0, then
f (a) + f (b) ≤ f (a + b)
(1.4)
for all scalars a, b ≥ 0. However, if the scalars a, b are replaced by two positive
operators, this inequality may not hold. For example if f (t) = t2 and A, B are
the following two positive matrices
A =
1 0 0
0 1 1
0 1 2
and B =
1 1 1
1 2 0
1 0 2
,
then a straightforward computation reveals that there is no relationship between
A2 + B2 and (A + B)2 under the operator order. Many authors tried to obtain
some operator extensions of (1.4). In [10], it was shown that
f (A + B) ≤ f (A) + f (B)
for all non-negative operator monotone functions f : [0, ∞) → [0, ∞) if and only
if AB + BA is positive.
Another operator extension of (1.4) was established in [9]
Theorem D. [9, Corollary 2.9] If f : [0, ∞) → [0, ∞) is a convex function with
f (0) ≤ 0, then f (A) + f (B) ≤ f (A + B) for all invertible positive operators A, B
such that A ≤ MI ≤ A + B and B ≤ MI ≤ A + B for some scalar M ≥ 0.
Some other operator extensions of (1.4) can be found in [1, 2, 11]. In this paper,
as a continuation of [9], we extend inequality (1.3), refine (1.3) and improve some
4
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
of our results in [9]. Some applications such as further refinements of the Petrovi´c
operator inequality and the Jensen -- Mercer operator inequality are presented as
well.
2. Results
To presenting our results, we introduce the abbreviation:
δf = f (m) + f (M) − 2f(cid:18)m + M
2 (cid:19)
for f : [m, M] → R, m < M.
We need the following lemma may be found in [7, Lemma 2]. We give a proof
for the sake of completeness.
Lemma 2.1. Let A ∈ σ([m, M]), for some scalars m < M . Then
f (A) ≤
M − A
M − m
f (m) +
A − m
M − m
(2.1)
holds for every convex function f : [m, M] → R, where
f (M) − δfeA
(cid:12)(cid:12)(cid:12)(cid:12) .
2
m + M
1
2
eA =
1
−
M − m(cid:12)(cid:12)(cid:12)(cid:12)A −
+ (1 − 2λ)b(cid:19)
2 (cid:19) + (1 − 2λ)f (b)
2
If f is concave on [m, M], then inequality (2.1) is reversed.
Proof. First assume that a, b ∈ [m, M] and λ ∈ [0, 1/2] so that λ ≤ 1 − λ. Then
f (λa + (1 − λ)b) = f(cid:18)2λ
a + b
≤ 2λf(cid:18)a + b
= λf (a) + (1 − λ)f (b) − λ(cid:18)f (a) + f (b) − 2f(cid:18)a + b
2 (cid:19)(cid:19) .
It follows that
f (λa + (1 − λ)b)
≤ λf (a) + (1 − λ)f (b) − min{λ, 1 − λ}(cid:18)f (a) + f (b) − 2f(cid:18)a + b
2 (cid:19)(cid:19)
(2.2)
OPERATOR INEQUALITIES OF JENSEN TYPE
5
for all a, b ∈ [m, M] and all λ ∈ [0, 1]. If t ∈ [m, M], then by using (2.2) with
λ = M −t
M −m, a = m and b = M we obtain
M − t
M − m
f (m) +
t − m
M − m
f (M)
f (t) = f(cid:18) M − t
M − m
m +
M(cid:19) ≤
t − m
M − m
t − m
− min(cid:26) M − t
M − m
,
f (t) ≤
(2.3) that
for any t ∈ [m, M]. Since min(cid:8) M −t
M − m(cid:12)(cid:12)(cid:12)(cid:12)t −
−(cid:18)1
t − m
M − m
M − t
M − m
f (m) +
−
2
1
(2.3)
2 − 1
M −m, t−m
M − m(cid:27)(cid:18)f (m) + f (M) − 2f(cid:18) m + M
2 (cid:19)(cid:19)
(cid:12)(cid:12), we have from
M −m(cid:12)(cid:12)t − m+M
M −m(cid:9) = 1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19)(cid:18)f (m) + f (M) − 2f(cid:18) m + M
2 (cid:19)(cid:19) ,
m + M
f (M)
(2.4)
2
2
for all t ∈ [m, M]. Now if A ∈ σ([m, M]), then by utilizing the functional calculus
(cid:3)
to (2.4) we obtain (2.1).
In the next theorem we present a generalization of [9, Theorem 2.1].
Pn1
i=1 Φi(I) = α I, Pn2
Theorem 2.2. Let Φi, Φi, Ψi, Ψi be positive linear mappings on B(H ) such that
i=1 Ψi(I) = δ I for
some real numbers α, β, γ, δ > 0. Let Ai (i = 1, . . . , n1), Di (i = 1, . . . , n2), Ci
(i = 1, . . . , n3) and Bi (i = 1, . . . , n4) be operators in σ(J) such that Ai ≤ m ≤
Bi, Ci ≤ M ≤ Di for two real numbers m < M . If
i=1 Ψi(I) = γ I, Pn4
i=1 Φi(I) = β I, Pn3
1
α
Φi(Ai) +
n1Xi=1
Ψi(Ci)! + f 1
β
n3Xi=1
1
δ
Φi(Di) =
n2Xi=1
Ψi(Bi)! ≤
n4Xi=1
≤
then
f 1
γ
Ψi(Ci) +
1
β
n4Xi=1
Ψi(Bi),
(2.5)
Φi (f (Ai)) +
Φi (f (Ai)) +
1
δ
1
δ
n2Xi=1
n2Xi=1
Φi (f (Di)) − δf eX
Φi (f (Di))
(2.6)
holds for every convex function f : J → R, where
1
γ
1
α
1
α
n3Xi=1
n1Xi=1
n1Xi=1
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1
β
n4Xi=1
Ψi(Bi) −
m + M
2
1
γ
n3Xi=1
Ψi(Ci) −
m + M
2
If f is concave, then the reverse inequalities are valid in (2.6).
eX = 1 −
1
M − m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
! .
6
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
Proof. We prove only the case when f is convex. Let [m, M] ⊆ J. It follows from
the convexity of f on J that
f (t) ≥
M − t
M − m
f (m) +
t − m
M − m
f (M)
(2.7)
for all t ∈ J \ [m, M]. Hence, by Ai ≤ m and Di ≥ M we have
f (Ai) ≥
M − Ai
M − m
f (m) +
Ai − m
M − m
f (M)
(i = 1, · · · , n1)
(2.8)
and similarly
f (Di) ≥
M − Di
M − m
f (m) +
Di − m
M − m
f (M)
(i = 1, · · · , n2).
(2.9)
Applying the positive linear mappings Φi and Φi, respectively, to both sides of
(2.8) and (2.9) and summing we get
1
αPn1
i=1 Φi(Ai) − m
M − m
f (M) (2.10)
Φi(f (Ai)) ≥
1
α
n1Xi=1
and
1
δ
n2Xi=1
Φi(f (Di)) ≥
M − 1
i=1 Φi(Ai)
αPn1
M − m
f (m) +
M − 1
i=1 Φi(Di)
δPn2
M − m
f (m) +
On the other hand, taking into account that m ≤ 1
and using Lemma 2.1 we obtain
f 1
β
n4Xi=1
Ψi(Bi)! ≤
and
M − 1
i=1 Ψi(Bi)
βPn4
M − m
f (m) +
1
δPn2
βPn4
βPn4
1
i=1 Φi(Di) − m
M − m
f (M). (2.11)
i=1 Ψi(Bi), 1
i=1 Ψi(Ci) ≤ M
γPn3
f (M ) − δfeB
(2.12)
i=1 Ψi(Bi) − m
M − m
γ
2 − 1
f 1
n3Xi=1
where eB = 1
eX = 1 −
Ψi(Ci)! ≤
M −m(cid:12)(cid:12)(cid:12) 1
βPn4
M − m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1
M − 1
i=1 Ψi(Ci)
γPn3
M − m
i=1 Ψi(Bi) − m+M
2
1
β
Ψi(Bi) −
n4Xi=1
1
i=1 Ψi(Ci) − m
f (m) +
M − m
2 − 1
γPn3
M −m(cid:12)(cid:12)(cid:12) 1
(cid:12)(cid:12)(cid:12) and eC = 1
γPn3
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
n3Xi=1
Ψi(Ci) −
1
γ
2
m + M
f (M ) − δfeC,
(2.13)
i=1 Ψi(Ci) − m+M
2
(cid:12)(cid:12)(cid:12).
m + M
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
!
Adding two inequalities (2.12) and (2.13) and putting
OPERATOR INEQUALITIES OF JENSEN TYPE
7
we obtain
f 1
β
n4Xi=1
Ψi(Bi)! + f 1
βPn4
M − m
i=1 Ψi(Bi) − 1
Ψi(Ci)!
n3Xi=1
γPn3
γ
i=1 Ψi(Ci)
i=1 Ψi(Bi) + 1
i=1 Ψi(Ci) − 2m
γPn3
δPn2
M − m
i=1 Φi(Ai) − 1
M − m
αPn1
i=1 Φi(Di)
i=1 Φi(Ai) + 1
i=1 Φi(Di) − 2m
f (m)
f (M ) − δf eX
f (m)
2M − 1
+
1
βPn4
2M − 1
≤
=
1
+
1
α
αPn1
n1Xi=1
(cid:12)(cid:12) ≤ M −m
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
n4Xi=1
1
β
2
M − m
δPn2
n2Xi=1
1
δ
f (M ) − δf eX (by (2.5))
(by (2.10) and (2.11))
≤
Φi(f (Ai)) +
Φi(f (Di)) − δf eX,
γPn3
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
n3Xi=1
1
γ
which is the first inequality in (2.6).
βPn4
i=1 Ψi(Bi), 1
Furthermore, m ≤ 1
i=1 Ψi(Ci) ≤ M . The numerical inequality
2
(m ≤ t ≤ M ) yields that
(cid:12)(cid:12)t − m+M
Therefore eX ≥ 0. Moreover, f is convex on [m, M ]. Hence δf ≥ 0. So the second
inequality in (2.6) holds.
≤ M − m.
Ψi(Bi) −
Ψi(Ci) −
m + M
m + M
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:3)
2
2
Remark 2.3. We can conclude some other versions of inequality (2.6). In fact,
under the assumptions in Theorem 2.2 the following inequalities hold true:
(1)
1
γ
n3Xi=1
Ψi(f (Ci)) +
1
β
n4Xi=1
(2) f 1
γ
n3Xi=1
Ψi(Ci)! +
1
β
in which
eX2 = 1 −
eX3 = 1 −
γ
1
M − m" 1
M − m"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1
1
γ
δ
δ
α
α
α
1
δ
n4Xi=1
Φi(Di)! − δf eX2
Ψi(f (Bi)) ≤ f 1
Φi(Ai)! + f 1
n2Xi=1
n1Xi=1
Φi(Di)! ;
≤ f 1
Φi(Ai)! + f 1
n2Xi=1
n1Xi=1
Ψi(f (Bi)) ≤ f 1
Φi(Ai)! +
n2Xi=1
n1Xi=1
Φi(f (Di)) − δf eX3
≤ f 1
Φi(Ai)! +
n1Xi=1
n2Xi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) +
Ψi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Bi −
n4Xi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Ψi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Bi −
n4Xi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19)# ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19)# .
Ψi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Ci −
n3Xi=1
n3Xi=1
Φi(f (Di)),
Ψi(Ci) −
M + m
M + m
M + m
M + m
1
β
1
β
1
δ
+
α
2
2
2
2
8
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
Before giving an example, we present some special cases of Theorem 2.2 which
are useful in our applications. The next corollary provides a refinement of [9,
Theorem 2.1].
Corollary 2.4. Let f be a convex function on an interval J . Let A, B, C, D ∈
σ(J) such that A + D = B + C and A ≤ m ≤ B, C ≤ M ≤ D for two real
numbers m < M . If Φ is a unital positive linear map on B(H ), then
where
In particular,
1
≤ Φ(f (A)) + Φ(f (D)),
f (Φ(B)) + f (Φ(C)) ≤ Φ(f (A)) + Φ(f (D)) − δfeX
M − m(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Φ(B) −
eX = 1 −
f (B) + f (C) ≤ f (A) + f (D) − δfeX ≤ f (A) + f (D).
(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)Φ(C) −
m + M
m + M
2
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) .
(2.14)
(2.15)
If f is concave on J , then inequalities (2.14) and (2.15) are reversed.
Another special case of Theorem 2.2 leads to a refinement of [9, Corollary 2.3].
Corollary 2.5. Let f be a convex function on an interval J . Let Ai, Bi, Ci, Di ∈
σ(J) (i = 1, · · · , n) such that Ai + Di = Bi + Ci and Ai ≤ m ≤ Bi, Ci ≤ M ≤
(i = 1, · · · , n). Let Φ1, · · · , Φn be positive linear mappings on B(H ) with
Di
i=1 Φi(I) = I. Then
Pn
(1) f nXi=1
Φi(Bi)! + f nXi=1
(2)
(3)
nXi=1
nXi=1
Φi(f (Bi)) +
nXi=1
Φi(f (Bi)) + f nXi=1
≤
Φi(Ci)! ≤
Φi(f (Ai)) +
Φi(f (Ai)) +
nXi=1
nXi=1
nXi=1
nXi=1
Φi(f (Ci)) ≤ f nXi=1
Φi(Ai)! + f nXi=1
≤ f nXi=1
Φi(Ai)! + f nXi=1
Φi(Ci)! ≤ f nXi=1
Φi(Di)! +
nXi=1
≤ f nXi=1
Φi(Di)! +
nXi=1
Φi(f (Di));
Φi(f (Di)) − δfeX1
Φi(Di)! − δfeX2
Φi(Di)! ;
Φi(f (Ai)) − δfeX3
Φi(f (Ai));
OPERATOR INEQUALITIES OF JENSEN TYPE
9
where
eX1 = 1 −
eX2 = 1 −
eX3 = 1 −
1
1
Φi(Bi) −
M − m"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
M − m" nXi=1
Φi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Bi −
M − m" nXi=1
Φi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Bi −
1
m + M
2
m + M
2
2
m + M
Φi(Ci) −
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) +
Φi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Ci −
nXi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
Φi(Ci) −
m + M
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
# ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19)# ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
# .
m + M
m + M
2
Now we give an example to show that how Theorem 2.2 works.
Example 2.6. Let ni = 1 for i = 1, 2, 3, 4 and let f (t) = t4. The function f is
convex but not operator convex[3]. Let Φ, Ψ, Ψ = Φ in which
Φ : M3(C) → M2(C), Φ((aij)1≤i,j≤3) = (aij)1≤i,j≤2.
If
then Φ(A) + Φ(D) = Φ(C) + Φ(B) and A ≤ 2.2I ≤ C, B ≤ 8I ≤ D. Also
1
2
15
9
1
1 10
2
1
1
1 −1
2
−1
1
2 −5
1
, C =
, D =
A =
0.345 0.655 !, whence
δf = 2766.4 and eX = 0.655 0.345
(Φ(C))4 + (Φ(B))4 = 1891 −859
3022 !
5281
5318
6202
6239
8758 ! = (Φ(A))4 + (Φ(D))4 − δf eX (cid:8) 7093
8867 ! = Φ(A4) + (Φ(D))4 − δf eX (cid:8) 7130
12263 ! = (Φ(A))4 + Φ(D4) − δf eX (cid:8) 8014
12372 ! = Φ(A4) + Φ(D4) − δf eX (cid:8) 8051
4373.5
2576.5
4311.5
2514.5
4373.5
2514.5
2576.5
4311.5
−859
(cid:8)
This shows that inequalities in (2.6) can be strict.
4 1 2
1 4 1
2 1 5
, B =
6 −1 1
−1
1
5
1
7
1
,
3531
3469
3469 10570 ! = (Φ(A))4 + (Φ(D))4
3531 10679 ! = Φ(A4) + (Φ(D))4
5266 14075 ! = (Φ(A))4 + Φ(D4)
5328 14184 ! = Φ(A4) + Φ(D4)
5328
5266
10
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
Moreover,
(Φ(A))4 + Φ(B4) − Φ(A4) − (Φ(B))4 = 884
(Φ(A))4 + Φ(B4) − Φ(A)4 − (Φ(B))4 = 921
Φ(A4) + Φ(B4) − (Φ(A))4 − Φ(B4) = 921
1735 3396 ! 6T 0
1797 3505 ! 6T 0
1797 3505 ! 6T 0.
1797
1735
1797
Hence there is no relationship between the right hand sides of inequalities in Corol-
lary 2.5.
Corollary 2.7. Let f be a convex function on an interval J . Let Ai, Bi, Ci, Di,
i = 1, . . . , n, be operators in σ(J). If Ai ≤ m ≤ Ci, Bi ≤ M ≤ Di, i = 1, . . . , n,
for two real numbers m < M and
then
and
(Ai + Di) =
(Ci + Bi),
f nXi=1
nXi=1
nXi=1
Di! − δf,neXn,
Ai! + f nXi=1
Bi! ≤ f nXi=1
Ci! + f nXi=1
≤ f nXi=1
Ai! + f nXi=1
Di!
f (Di) − δf nXi=1
nXi=1
nXi=1
nXi=1
nXi=1
(cid:1) and
in which δf,n = f (nm) + f (nM) − 2f(cid:0) nm+nM
+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
nM − nm"(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
If f is concave, then inequalities (2.17) and (2.18) are reversed.
eXn = 1 −
(eCi + eBi)!
nXi=1
nXi=1
f (Ci) +
f (Bi) ≤
f (Ai) +
f (Di)
# .
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nM + nm
Ci −
nXi=1
nM + nm
Bi −
2
f (Ai) +
(2.16)
(2.17)
(2.18)
≤
1
2
2
Proof. We prove only inequality (2.17) in the convex case. It follows from Ai ≤
m ≤ Ci, Bi ≤ M ≤ Di, (i = 1, . . . , n) that
Ai ≤ mnI ≤
nXi=1
Ci,
nXi=1
nXi=1
Bi ≤ MnI ≤
Di.
nXi=1
OPERATOR INEQUALITIES OF JENSEN TYPE
11
Using the same reasoning as in the proof of Theorem 2.2 we get
≤
=
i=1(Ci + Bi)
Mn − mn
i=1(Ai + Di)
i=1(Ci + Bi) − 2mn
Mn − mn
Ci! + f nXi=1
f nXi=1
2Mn −Pn
2Mn −Pn
≤ f nXi=1
which give the first inequality in (2.17). It is easy to see that δf,neXn ≥ 0, whence
Bi!
f (mn) +Pn
f (mn) +Pn
Di! − δf,neXn,
f (Mn) − δf,neXn
f (Mn) − δf,neXn
Mn − mn
Ai! + f nXi=1
the second inequality derived.
i=1(Ai + Di) − 2mn
Mn − mn
(cid:3)
(by (2.16))
3. Applications
Using the results in Section 2, we provide some applications which are refine-
ments of some well-known operator inequalities. As the first, we give a refinement
of the operator Jensen -- Mercer inequality.
Corollary 3.1. Let Φ1, · · · , Φn be positive linear maps on B(H ) withPn
I and B1, · · · , Bn ∈ σ([m, M]) for two scalars m < M . If f is a convex function
on [m, M], then
i=1 Φi(I) =
f m + M −
nXi=1
Φi(Bi)! ≤ f (m) + f (M) −
≤ f (m) + f (M) −
Φi(f (Bi)) − δfeB
Φi(f (Bi)),
Φi(Bi) −
m + M
2
nXi=1
nXi=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
# .
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
where eB = 1 −
1
M − m" nXi=1
Φi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)Bi −
m + M
2
Proof. Clearly m ≤ Bi ≤ M (i = 1, · · · , n). Set Ci = M + m − Bi
(i =
1, · · · , n). Then m ≤ Ci ≤ M and Bi + Ci = m + M (i = 1, · · · , n). Applying
inequality (3) of Corollary 2.5 when Ai = mI and Di = MI we obtain the desired
(cid:3)
inequalities.
The next result provides a refinement of the Petrovi´c inequality for operators.
12
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
Corollary 3.2. If f : [0, ∞) → R is a convex function and B1, · · · , Bn are
i=1 Bi = MI for some scalar M > 0, then
positive operators such thatPn
Bi! + (n − 1)f (0) − δfeB ≤ f nXi=1
f (Bi) ≤ f nXi=1
nXi=1
nXi=1(cid:12)(cid:12)(cid:12)(cid:12)
2(cid:12)(cid:12)(cid:12)(cid:12) .
where eB =
Proof. It follows from 0 ≤ Bi ≤ M that
Bi
M
n
2
−
−
1
f (Bi) ≤
M − Bi
M − 0
f (0) +
Bi − 0
M − 0
Summing above inequalities over i we get
f (M) − δffBi
Bi! + (n − 1)f (0),
(i = 1, · · · , n).
f (Bi) ≤
nXi=1
f (M) − δf
i=1 Bi
M
nXi=1fBi
Bi! − δfeB (byPn
Bi! (by δfeB ≥ 0).
M
i=1 Bi
f (0) +Pn
nM −Pn
= (n − 1)f (0) + f nXi=1
≤ (n − 1)f (0) + f nXi=1
nXi=1(cid:12)(cid:12)(cid:12)(cid:12)
2(cid:12)(cid:12)(cid:12)(cid:12) .
Bi
M
−
1
i=1 Bi = M )
(cid:3)
n
2
−
where eB =
As another consequence of Theorem 2.2, we present a refinement of the Jensen
operator inequality for real convex functions. The authors of [9] introduce a
subset Ω of Bh(H ) × Bh(H ) defined by
2
A + B
≤ M ≤ B,
We have the following result.
Corollary 3.3. Let f be a convex function on an interval J containing m, M .
i=1 Φi(I) = I.
for some m, M ∈ R(cid:27) .
Ω =(cid:26)(A, B)(cid:12)(cid:12) A ≤ m ≤
Let Φi, i = 1, . . . , n, be positive linear mappings on B(H ) withPn
(cid:19) − δfeX
(cid:19) ,
Φi(cid:18) f (Ai) + f (Di)
Φi(cid:18) f (Ai) + f (Di)
Φi(cid:18)Ai + Di
If (Ai, Di) ∈ Ω, i = 1, . . . , n, then
f nXi=1
(cid:19)! ≤
(3.1)
≤
2
2
2
nXi=1
nXi=1
OPERATOR INEQUALITIES OF JENSEN TYPE
13
where
1
2
−
eX =
1
M − m(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
Φi(cid:18) Ai + Di
2
(cid:19) −
m + M
2
.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
If f is concave, then inequalities in (3.1) are reversed.
Proof. Putting Bi = Ci = Ai+Di
conclude the desired result.
2
and using inequality (1) of Corollary 2.5, we
(cid:3)
Note that utilizing Corollary 2.5, we even be able to obtain a converse of the
Jensen operator inequality. For this end, under the assumptions in the Corol-
lary 3.3 we have
where
Φi(cid:18)f(cid:18) Ai + Di
2
(cid:19)(cid:19) ≤
nXi=1
1
1
2"f nXi=1
Φi(Ai)! + f nXi=1
2"f nXi=1
Φi(Ai)! + f nXi=1
Φi(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)
nXi=1
Ai + Di
−
2
2
Φi(Di)!# − δfeX
Φi(Di)!# ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) .
m + M
≤
1
(3.2)
1
2
−
M − m
eX =
Note that the function f need not to be operator convex. Let us give an example
to illustrate these inequalities.
Example 3.4. Let n = 1 and the unital positive linear map Φ : M3(C) → M2(C)
be defined by
Φ((aij)1≤i,j≤3) = (aij)1≤i,j≤2
for each A = (aij)1≤i,j≤3 ∈ M3(C). Consider the convex function f (t) = et on
[0, ∞). If
7 −1 0
−1
0
7
0
6
0
,
1
0
1 −1 0
−1
0
1
0
A =
2 (cid:19)(cid:19) = 79.8 −50.5
−50.5
D =
54.6 ! (cid:8) 759.2 −399
(cid:8) 768.2 −408
−399
−408
then 0 ≤ A ≤ 2I ≤ A+D
that
f(cid:18)Φ(cid:18) A + D
344 ! = Φ(cid:18) f (A) + f (D)
362 ! = Φ(cid:18) f (A) + f (D)
2
2
(cid:19) − δf eX
(cid:19) ,
2 ≤ 5I ≤ D, i.e., (A, D) ∈ Ω. Hence it follows from (3.1)
14
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
in which δf = 89.6 and eX = 0.1 0.1
0.1 0.2 !.
It should be mentioned that in the case when f is operator convex, under the
assumptions in Corollary 3.3 we have even more:
f nXi=1
Φi(cid:18)Ai + Di
2
(cid:19)! ≤
≤
≤
1
2
Φi(cid:18)f(cid:18) Ai + Di
(cid:19)(cid:19) (by the Jensen inequality)
nXi=1
2"f nXi=1
Φi(Ai)! + f nXi=1
Φi(Di)!# − δfeX (by (3.2))
2" nXi=1
Φi(f (Ai) + f (Di))# − δfeX (by the Jensen inequality)
Φi(cid:18)f (Ai) + f (Di)
nXi=1
2
1
Corollary 3.5. If f is a convex function on an interval J containing m, M , then
≤
(cid:19) (since δf eX ≥ 0).
f (λA + (1 − λ)D) ≤ λf (A) + (1 − λ)f (D) − δfeX
M − m(cid:12)(cid:12)(cid:12)(cid:12)
≤ λf (A) + (1 − λ)f (D)
1
2
If f is concave, then inequality (3.3) is reversed.
for all (A, D) ∈ Ω and all λ ∈ [0, 1], where eX =
−
1
2
Proof. Put n = 1 and let Φ be the identity map in Corollary 3.3 to get
A + D − M − m
f(cid:18)A + D
2 (cid:19) ≤
f (A) + f (D)
2
f (A) + f (D)
2
− δf eX ≤
for any (A, D) ∈ Ω, which implies (3.3) by the continuity of f .
(3.3)
(cid:12)(cid:12)(cid:12)(cid:12) .
(cid:3)
Regarding to obtain an operator version of (3.4), it is shown in [9] that if
f : [0, ∞) → [0, ∞) is a convex function with f (0) ≤ 0, then
f (A) + f (B) ≤ f (A + B)
(3.4)
for all strictly positive operators A, B for which A ≤ M ≤ A + B and B ≤ M ≤
A + B for some scalar M. We give a refined extension of this result as follows.
Theorem 3.6. If f : [0, ∞) → R is a convex function with f (0) ≤ 0 then
f (Ci) ≤ f nXi=1
nXi=1
Ci! − δf
nXi=1 eCi ≤ f nXi=1
Ci!
(3.5)
OPERATOR INEQUALITIES OF JENSEN TYPE
15
for all positive operators Ci such that Ci ≤ M ≤Pn
scalar M ≥ 0. If f is concave, then the reverse inequality is valid in (3.5).
i=1 Ci (i = 1, . . . , n) for some
In particular, if f is convex, then
for all positive operators A, B such that A ≤ MI ≤ A + B and B ≤ MI ≤ A + B
Proof. Without loss of generality let M > 0. Lemma 2.1 implies that
for some scalar M ≥ 0, where eX = 1 −(cid:12)(cid:12)(cid:12)(cid:12)
f (Ci) ≤
MI − Ci
M − 0
1
−
1
−
B
M
A
M
f (A) + f (B) ≤ f (A + B) − δfeX ≤ f (A + B)
2(cid:12)(cid:12)(cid:12)(cid:12) −(cid:12)(cid:12)(cid:12)(cid:12)
2(cid:12)(cid:12)(cid:12)(cid:12) .
f (M) − δfeCi
f (M) − δfeCi =
nXi=1 eCi.
f (Ci) ≤ Pn
i=1 Ci
M
f (M) − δf
f (0) +
M − 0
Ci
M
Ci
since f (0) ≤ 0. Summing the above inequalities over i we get
nXi=1
Since the spectrum ofPn
f nXi=1
MI −Pn
≥ Pn
Ci! ≥
i=1 Ci
M
M − 0
i=1 Ci
f (M)
f (M)
i=1 Ci
M − 0
f (0) +Pn
(since MI ≤Pn
(i = 1, . . . , n)
(3.6)
i=1 Ci is contained in [M, ∞) ⊂ [0, ∞) \ [0, M), we have
i=1 Ci and f (0) ≤ 0).
(3.7)
Combining two inequalities (3.6) and (3.7), we reach to the desired inequality
(cid:3)
(3.5).
Theorem 3.7. Let A, B, C, D ∈ σ(J) such that A ≤ m ≤ B, C ≤ M ≤ D for
two real numbers m < M . If f is a convex function on J and any one of the
following conditions
(i) B + C ≤ A + D and f (m) ≤ f (M)
(ii) A + D ≤ B + C and f (M) ≤ f (m)
is satisfied, then
1
f (B) + f (C) ≤ f (A) + f (D) − δfeX ≤ f (A) + f (D),
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) .
M − m(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)B −
(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)C −
(iii) B + C ≤ A + D and f (M) ≤ f (m)
M + m
M + m
2
2
If f is concave and any one of the following conditions
where eX = 1 −
(iv) A + D ≤ B + C and f (m) ≤ f (M)
is satisfied, then inequality (3.8) is reversed.
(3.8)
16
M.S. MOSLEHIAN, J. MI ´CI ´C, M. KIAN
Proof. Let f be convex and (i) is valid. It follows from Lemma 2.1 that
f (B) ≤
and
f (C) ≤
f (M) − f (m)
M − m
B +
f (m)M − f (M)m
M − m
f (M) − f (m)
M − m
C +
f (m)M − f (M)m
M − m
Summing above inequalities we get
2
− δf(cid:18)1
−δf(cid:18)1
2
−
1
−
2
2
1
M + m
M + m
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) .
M − m(cid:12)(cid:12)(cid:12)(cid:12)B −
M − m(cid:12)(cid:12)(cid:12)(cid:12)C −
− δf eX
− δf eX (by (i))
f (B) + f (C) ≤
≤
=
f (M) − f (m)
M − m
f (M) − f (m)
M − m
f (M) − f (m)
M − m
(B + C) + 2
(A + D) + 2
f (m)M − f (M)m
f (m)M − f (M)m
M − m
M − m
A +
f (m)M − f (M)m
M − m
f (m)M − f (M)m
M − m
− δf eX
(by (2.8) and (2.9))
(by δfeX ≥ 0 )
+
f (M) − f (m)
M − m
D +
≤ f (A) + f (D) − δfeX
≤ f (A) + f (D)
The other cases can be verified similarly.
(cid:3)
Applying the above theorem to the power functions we get
Corollary 3.8. Let A, B, C, D ∈ Bh(H ) be such that I ≤ A ≤ m ≤ B, C ≤
M ≤ D for two real numbers m < M . If one of the following conditions
(i) B + C ≤ A + D and p ≥ 1
(ii) A + D ≤ B + C and p ≤ 0
is satisfied, then
Proof. Let (i) be valid. Applying Theorem 3.7 for f (t) = tp, it follows
for each q ≥ p, where
Bp + C p ≤ Aq + Dq − δpeX ≤ Aq + Dq
M − m(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)B −
δp = mp+M p−2(cid:18) m + M
2 (cid:19)p
Bp + C p ≤ Ap + Dp − δpeX
≤ Aq + Dq − δpeX
, eX = 1−
≤ Aq + Dq
1
The other cases may be verified similarly.
(by q ≥ p)
(by δpeX ≥ 0 )
M + m
2
(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12)C −
M + m
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) .
(cid:3)
OPERATOR INEQUALITIES OF JENSEN TYPE
17
Acknowledgement. The third author would like to thank the Tusi Mathe-
matical Research Group (TMRG), Mashhad, Iran.
References
1. T. Ando and X. Zhan, Norm inequalities related to operator monotone functions, Math.
Ann. 315 (1999), 771 -- 780.
2. K. M.R. Audenaert and J.S. Aujla On norm sub-additivity and super-additivity inequalities
for concave and convex functions , arXiv:1012.2254v2.
3. T. Furuta, J. Mi´ci´c Hot, J. Pecari´c and Y. Seo, Mond -- Pecari´c Method in Operator Inequal-
ities, Zagreb, Element, 2005.
4. M. Kian and M.S. Moslehian, Operator inequalities related to Q-class functions, Math.
Slovaca, (to appear).
5. A. Matkovi´c, J. Pecari´c and I. Peri´c, A variant of Jensen's inequality of Mercer's type for
operators with applications, Linear Algebra Appl. 418 (2006), 551 -- 564.
6. J. Mi´ci´c, Z. Pavi´c and J. Pecari´c, Jensen's inequality for operators without operator convex-
ity, Linear Algebra Appl. 434 (2011), 1228 -- 1237.
7. J. Mi´ci´c, J. Pecari´c and J. Peri´c, Refined Jensen's operator inequality with condition on
spectra, Oper. Matrices 7 (2013), 293 -- 308.
8. M.S. Moslehian, Operator extensions of Hua's inequality, Linear Algebra Appl. 430 (2009),
no. 4, 1131-1139.
9. M.S. Moslehian, J. Mi´ci´c and M. Kian, An operator inequality and its consequences, Linear
Algebra Appl. (2012), .
10. M.S. Moslehian and H. Najafi. Around operator monotone functions, Integral Equations
Operator Theory 71 (2011), 575 -- 582.
11. M. Uchiyama, Subadditivity of eigenvalue sums, Proc. Amer. Math. Soc. 134 (2006), 1405 --
1412.
1 Department of Pure Mathematics, Center of Excellence in Analysis on Al-
gebraic Structures (CEAAS), Ferdowsi University of Mashhad, P.O. Box 1159,
Mashhad 91775, Iran
E-mail address: [email protected] and [email protected]
2 Faculty of Mechanical Engineering and Naval Architecture, University of
Zagreb, Ivana Luci´ca 5, 10000 Zagreb, Croatia
E-mail address: [email protected]
3 Department of Mathematics, Faculty of Basic Sciences, University of Bo-
jnord, Bojnord, Iran
E-mail address: [email protected]
|
1104.5195 | 1 | 1104 | 2011-04-27T17:20:32 | A hyponormal weighted shift on a directed tree whose square has trivial domain | [
"math.FA"
] | It is proved that, up to isomorphism, there are only two directed trees that admit a hyponormal weighted shift with nonzero weights whose square has trivial domain. These are precisely those enumerable directed trees, one with root, the other without, whose every vertex has enumerable set of successors. | math.FA | math |
A hyponormal weighted shift on a directed tree whose
square has trivial domain
Zenon Jan Jab lo´nski, Il Bong Jung, and Jan Stochel
Abstract. It is proved that, up to isomorphism, there are only two directed
trees that admit a hyponormal weighted shift with nonzero weights whose
square has trivial domain. These are precisely those enumerable directed trees,
one with root, the other without, whose every vertex has enumerable set of
successors.
1. Introduction
In a recent paper [4] a question of subnormality of unbounded weighted shifts on
directed trees has been investigated. A criterion for subnormality of such operators
whose C∞-vectors are dense in the underlying Hilbert space has been established
(cf. [4, Theorem 5.2.1]). It has been written in terms of consistent systems of Borel
probability measures. The assumption that the operator in question has a dense set
of C∞-vectors diminishes the class of weighted shifts on directed trees to which this
criterion can be applied (note that the set of all C∞-vectors of a classical, unilateral
or bilateral, weighted shift is always dense in the underlying Hilbert space). Unfor-
tunately, there is no general criterion for subnormality of densely defined operators
that have small set of C∞-vectors. The known characterizations of subnormality of
unbounded Hilbert space operators require the existence of additional objects (like
semispectral measures, elementary spectral measures or sequences of unbounded
operators) that have to satisfy appropriate, more or less complicated, conditions
(cf. [3, 7, 20, 21]). Among subnormal operators having small set of C∞-vectors,
the symmetric ones (which are always subnormal, see [1, Theorem 1 in Appen-
dix I.2]) play an essential role. According to [13] (see also [5]) there are closed
symmetric operators whose squares have trivial domain. Unfortunately, symmetric
weighted shifts on directed trees are automatically bounded; the same is true for
formally normal weighted shifts on directed trees (cf. [9, Proposition 3.1]).
The above discussion leads to the following problem.
2010 Mathematics Subject Classification. Primary 47B37, 47B20; Secondary 47A05.
Key words and phrases. Directed tree, weighted shift on a directed tree, hyponormal opera-
tor, trivial domain of square.
Research of the first and the third authors was supported by the MNiSzW (Ministry of Science
and Higher Education) grant NN201 546438 (2010-2013). The second author was supported by the
National Research Foundation of Korea (NRF) grant funded by the Korea government (MEST)
(No. R01-2008-000-20088-0).
1
2
Z. J. JAB LO ´NSKI, I. B. JUNG, AND J. STOCHEL
Question. Does there exist a subnormal weighted shift on a directed tree with
nonzero weights whose square has trivial domain?
At present, this question is unanswered (the reason for this is explained partially
in the previous paragraph). However, as is shown in Theorem 4.2, there are injective
hyponormal weighted shifts on directed trees with nonzero weights whose squares
have trivial domain. What is more, it is proved in Theorem 3.1 that the only
directed trees admitting densely defined weighted shifts with nonzero weights whose
squares have trivial domain are those enumerable directed trees whose every vertex
has enumerable set of successors (children).
2. Preliminaries
In what follows, C stands for the set of all complex numbers. Let A be an
operator in a complex Hilbert space H (all operators considered in this paper are
linear). Denote by D(A) and A∗ the domain and the adjoint of A (in case it
exists). A closed densely defined operator N in H is called normal if N∗N =
N N∗. A densely defined operator S in H is said to be subnormal if there exists a
complex Hilbert space K and a normal operator N in K such that H ⊆ K (isometric
embedding) and Sh = N h for all h ∈ D(S). Finally, a densely defined operator S in
H is called hyponormal if D(S) ⊆ D(S∗) and kS∗fk 6 kSfk for all f ∈ D(S). It is
well-known that subnormal operators are hyponormal (but not conversely) and that
hyponormal operators are closable and their closures are hyponormal (subnormal
operators have an analogous property). We refer the reader to [2, 22] for basic
facts on unbounded operators, [6, 16, 17, 18, 19] for the foundations of the theory
of (bounded and unbounded) subnormal operators and [14, 10, 11, 12, 15] for
elements of the theory of unbounded hyponormal operators.
Let T = (V, E) be a directed tree (V and E always stand for the sets of vertices
and edges of T , respectively). If T has a root, which will always be denoted by
root, then we write V ◦ := V \ {root}; otherwise, we put V ◦ = V . Set Chi(u) =
{v ∈ V : (u, v) ∈ E} for u ∈ V . If for a given vertex u ∈ V there exists a unique
vertex v ∈ V such that (v, u) ∈ E, then we denote it by par(u). The correspondence
u 7→ par(u) is a partial function from V to V . For an integer n > 1, the n-fold
composition of the partial function par with itself will be denoted by parn. Let par0
stand for the identity map on V . We call T leafless if V = {u ∈ V : Chi(u) 6= ∅}.
If W ⊆ V , we put Chi(W ) = Sv∈W Chi(v) and Des(W ) = S∞n=0 Chihni(W ), where
Chih0i(W ) = W and Chihn+1i(W ) = Chi(Chihni(W )) for all integers n > 0. For u ∈
V , we set Chihni(u) = Chihni({u}) and Des(u) = Des({u}). Combining equalities
(2.1.3), (6.1.3) and (2.1.10) of [8] with [8, Corollary 2.1.5], we obtain
(2.1)
(2.2)
(2.3)
(2.4)
(2.5)
Chi(u),
V ◦ = Gu∈V
Chihn+1i(u) = Gv∈Chihni(u)
Chi(v),
u ∈ V, n = 0, 1, 2, . . . ,
∞
Des(u) =
Gn=0
Des(u1) ∩ Des(u2) = ∅,
Chihni(u),
u ∈ V,
u1, u2 ∈ Chi(u), u1 6= u2, u ∈ V,
V = Des(root) provided that T has a root,
A HYPONORMAL WEIGHTED SHIFT WHOSE SQUARE HAS TRIVIAL DOMAIN
3
where the symbol F is reserved to denote pairwise disjoint union of sets.
Let ℓ2(V ) be the Hilbert space of all square summable complex functions on
V equipped with the standard inner product. For u ∈ V , we define eu ∈ ℓ2(V ) to
be the characteristic function of the one point set {u}. The family {eu}u∈V is an
orthonormal basis of ℓ2(V ). Denote by EV the linear span of {eu : u ∈ V }. Given
λ = {λv}v∈V ◦ ⊆ C, we define the operator Sλ in ℓ2(V ) by
D(Sλ) = {f ∈ ℓ2(V ) : ΛT f ∈ ℓ2(V )},
Sλf = ΛT f,
where ΛT is the map defined on functions f : V → C via
(2.6)
f ∈ D(Sλ),
(ΛT f )(v) =(λv · f(cid:0) par(v)(cid:1)
0
if v ∈ V ◦,
if v = root .
Sλ is called a weighted shift on the directed tree T with weights {λv}v∈V ◦ . Note
that any weighted shift Sλ on T is a closed operator (cf. [8, Proposition 3.1.2]).
Combining Propositions 3.1.3, 3.1.7 and 3.1.10 of [8], we get the following fact
(hereafter we adopt the convention that Pv∈∅ xv = 0).
Proposition 2.1. Let Sλ be a weighted shift on a directed tree T with weights
λ = {λv}v∈V ◦ . Then the following assertions hold:
(i) eu is in D(Sλ) if and only if Pv∈Chi(u) λv2 < ∞; if eu ∈ D(Sλ), then
(2.7)
λvev
Sλeu = Xv∈Chi(u)
and kSλeuk2 = Xv∈Chi(u)
λv2,
(ii) Sλ is densely defined if and only if EV ⊆ D(Sλ),
(iii) Sλ is injective if and only if T is leafless and Pv∈Chi(u) λv2 > 0 for
(iv) if D(Sλ) = ℓ2(V ) and λv 6= 0 for all v ∈ V ◦, then V is at most countable.
every u ∈ V ,
3. Directed trees admitting Sλ's with D(S2
λ) = {0}
The proof of Theorem 3.1 below contains a method of constructing densely
defined weighted shifts Sλ on directed trees with nonzero weights such that D(S2
λ) =
{0}. By imposing carefully tailored restrictions on weights, we will show in Theorem
4.2 below how to use this method to construct hyponormal weighted shifts on
directed trees with the aforesaid properties.
Theorem 3.1. Let T be a directed tree. Then the following assertions are equiv-
alent:
D(Sλ) = ℓ2(V ) and D(S2
(i) there exists a family λ = {λv}v∈V ◦ of nonzero complex numbers such that
λ) = {0},
(ii) card(Chi(u)) = ℵ0 for every u ∈ V .
Moreover, if Sλ is as in (i), then Sλ is injective.
Proof. Fix λ = {λv}v∈V ◦ ⊆ C. We show that
(†) a complex function f on V belongs to D(S2
λ) if and only if1
(3.1)
Xu∈V (cid:16)1 + ζ2
u + Xv∈Chi(u)
ζ2
vλv2(cid:17)f (u)2 < ∞,
1 with the convention that 0 · ∞ = 0
4
Z. J. JAB LO ´NSKI, I. B. JUNG, AND J. STOCHEL
where ζu :=qPv∈Chi(u) λv2 for u ∈ V .
if and only if f ∈ ℓ2(V ) and Pu∈V ζ2
Pu∈V ζ2
tion f on V belongs to D(S2
Indeed, by [8, Proposition 3.1.3], a complex function f on V belongs to D(Sλ)
uf (u)2 < ∞. Hence a complex func-
u)f (u)2 < ∞ and
u (Sλf )(u)2 < ∞. Since the following equalities hold for f ∈ D(Sλ),
λ) if and only if Pu∈V (1 + ζ2
ζ2
u λu2f (par(u))2
u (Sλf )(u)2 (2.6)
ζ2
Xu∈V
(2.1)
= Xu∈V ◦
= Xu∈V Xv∈Chi(u)
= Xu∈V (cid:16) Xv∈Chi(u)
vλv2f (par(v)2
ζ2
vλv2(cid:17)f (u)2,
ζ2
we see that a complex function f on V belongs to D(S2
λ) if and only if (3.1) holds.
(i)⇒(ii) Let Sλ be as in (i). By Proposition 2.1(iv), V is countable. Thus each
Chi(u) is countable. Suppose that, contrary to our claim, (ii) does not hold. Then
there exists u0 ∈ V such that Chi(u0) is finite. Since Sλ is densely defined, we infer
from assertions (i) and (ii) of Proposition 2.1 that ζv < ∞ for all v ∈ V . Hence
1 + ζ2
u0 + Xv∈Chi(u0)
ζ2
vλv2 < ∞.
This, combined with (†), implies that f = eu0 ∈ D(S2
λ), which contradicts (i).
(ii)⇒(i) First, we show that
(‡) for each (ϑ, u) ∈ (0,∞) × V there exists {λv}v∈Des(u) ⊆ (0,∞) such that
(3.2)
(3.3)
(cid:16) Xw∈Chi(v)
λ2
u = ϑ,
v = 1,
λ2
w(cid:17)λ2
v ∈ Des(u).
To do so, we fix u ∈ V and set Xn = Chih0i(u) ⊔ ··· ⊔ Chihni(u) for n > 1,
and X0 = Chih0i(u). Since, by (2.3), Des(u) = S∞n=1 Xn, we can construct the
required family inductively. For n = 1, we put λu = √ϑ and choose a family
{λv}v∈Chi(u) ⊆ (0,∞) such that(cid:0)Pv∈Chi(u) λ2
v(cid:1)ϑ = 1 (this is possible because Chi(u)
is nonempty and countable). Fix n > 1, and assume that we already have a family
{λv}v∈Xn ⊆ (0,∞) such that λ2
v = 1 for all v ∈ Xn−1.
Then for every v ∈ Chihni(u) we can choose a family {λw}w∈Chi(v) ⊆ (0,∞) such that
(cid:0)Pw∈Chi(v) λ2
v = 1. In view of (2.2), this gives us the family {λv}v∈Chihn+1i(u)
such that (cid:0)Pw∈Chi(v) λ2
v = 1 for all v ∈ Xn. Now by induction we are done.
u = ϑ and (cid:0)Pw∈Chi(v) λ2
does the job (the number λroot can be chosen arbitrarily).
If T has a root, then combining (†) and (‡) with (2.5) and Proposition 2.1(i)
Suppose now that T is rootless. Take u1 ∈ V and set u2 = par(u1). By
, which sat-
In the next step we construct a new fam-
such that the extended family
(‡), there exists a family {λv}v∈Des(u1) ⊆ (0,∞) with λu1 = 1√2
isfies (3.3) with u1 in place of u.
ily {λv}v∈Des(u2)\Des(u1) ⊆ (0,∞) with λu2 = 1√2
w(cid:1)λ2
w(cid:1)λ2
w(cid:1)λ2
A HYPONORMAL WEIGHTED SHIFT WHOSE SQUARE HAS TRIVIAL DOMAIN
5
{λv}v∈Des(u2) satisfies (3.3) with u2 in place of u. For this, note that
(3.4)
Des(u).
(2.4)
Des(u2) \ Des(u1)
= {u2} ⊔
Gu∈Chi(u2)\{u1}
Set λu2 = 1√2
and choose a family {ϑu}u∈Chi(u2)\{u1} ⊆ (0,∞) such that
(3.5)
(cid:16) Xu∈Chi(u2)\{u1}
ϑu + λ2
u2 = 1.
u1(cid:17)λ2
Applying (‡) to u ∈ Chi(u2) \ {u1} and ϑ = ϑu, we get the family {λv}v∈Des(u) ⊆
(0,∞) satisfying (3.2) and (3.3) with ϑ = ϑu. This, together with (3.5), leads
to (cid:0)Pu∈Chi(u2) λ2
u2 = 1. In view of (3.4), our construction is complete. Ap-
plying an induction argument (with λun = 1√2
for n > 2) and using the fact
that V = S∞k=0 Des(park(u1)) (cf. [8, Proposition 2.1.6]), we construct a family
λ = {λv}v∈V ⊆ (0,∞) such that ζ2
v = 1 for all v ∈ V . This, combined with (†)
and Proposition 2.1(i), gives (i).
u(cid:1)λ2
v λ2
The "moreover" part follows from (ii) and Proposition 2.1(iii).
(cid:3)
Our method enables us to construct Sλ with the additional property that
D(Sλ) * D(S∗λ), which is opposite to what happens in Theorem 4.2 below.
Theorem 3.2. If T is a directed tree such that card(Chi(u)) = ℵ0 for every u ∈ V ,
then there exists a family λ = {λv}v∈V ◦ of nonzero complex numbers such that Sλ
is injective and densely defined, D(Sλ) * D(S∗λ) and D(S2
λ) = {0}.
Proof. To achieve this, we proceed as in the proof of implication (ii)⇒(i) of
Theorem 3.1 with one exception, namely, we strengthen (‡) by requiring, in addition
to (3.2) and (3.3), that
(3.6)
sup
v∈Chi(u) Xw∈Chi(v)
λ4
w
1 + λ2
w
= ∞.
This in turn can be deduced from the following fact:
(3.7)
n = α2.
for every real number α > 0, there exists a sequence {λn}∞n=1 ⊆ (0,∞)
such that λ1 − α < 1 and P∞n=1 λ2
v(cid:1)ϑ = 1. Then evidently supv∈Chi(u) 1/λ2
Indeed, arguing as in the proof of (‡), we find a family {λv}v∈Chi(u) ⊆ (0,∞) such
that (cid:0)Pv∈Chi(u) λ2
v = ∞. In the next step,
using (3.7), we construct a family {λw}w∈Chih2i(u) such that(cid:0)Pw∈Chi(v) λ2
w(cid:1) = 1/λ2
for every v ∈ Chi(u) and supw∈Chih2i(u) λ2
w = ∞. This, combined with (2.2), implies
(3.6). The rest of the proof goes through as for (‡), with hardly any changes.
It follows from (2.7) and (3.3) that kSλewk2 = 1/λ2
w for all w ∈ Des(u), which
λw2
together with (3.6) implies that supv∈V Pw∈Chi(v)
1+kSλewk2 = ∞. By applying
[8, Theorem 4.1.1], we deduce that D(Sλ) * D(S∗λ). Obviously, such Sλ is never
hyponormal.
(cid:3)
v
4. Hyponormal weighted shifts Sλ with D(S2
λ) = {0}
Let us recall a characterization of hyponormality of weighted shifts on directed
trees with nonzero weights (in view of [4, Proposition 5.3.1], there is no loss of
generality in assuming that underlying directed trees are leafless).
6
Z. J. JAB LO ´NSKI, I. B. JUNG, AND J. STOCHEL
Theorem 4.1 ([8, Theorem 5.1.2 and Remark 5.1.5]). Let Sλ be a densely defined
weighted shift on a leafless directed tree T with nonzero weights λ = {λv}v∈V ◦ .
Then Sλ is hyponormal if and only if
Xv∈Chi(u)
λv2
kSλevk2
6 1,
u ∈ V.
Now we show that there are hyponormal weighted shifts Sλ with D(S2
λ) = {0}.
Theorem 4.2. If T is a directed tree such that card(Chi(u)) = ℵ0 for every u ∈ V ,
then there exists a family λ = {λv}v∈V ◦ of nonzero complex numbers such that Sλ
is injective and hyponormal, and D(S2
λ) = {0}.
1
λ2
Proof. We modify the proof of implication (ii)⇒(i) of Theorem 3.1. First we
note that for each positive real number r, there exists a sequence {rn}∞n=1 ⊆ (0, 1)
such that (P∞j=1 rj) r = 1 and P∞j=1 r2
rn for 1 6 j 6 n − 1, and
6 1). This fact, when
rj =
incorporated to the proof of (‡), leads to
rn2j−n+1 for j > n, where n > 2 is chosen so that
j 6 1 (e.g., rj = 1
1
(‡‡) for each (ϑ, u) ∈ (0,∞) × V there exists {λv}v∈Des(u) ⊆ (0, 1) such that
w(cid:1)λ2
u = ϑ, (cid:0)Pw∈Chi(v) λ2
w 6 1 for all v ∈ Des(u).
If T has a root, then applying (‡‡) to u = root and ϑ = 1 we get a family
λ = {λv}v∈V ◦ ⊆ (0, 1) such that
w(cid:1)λ2
(4.1)
v = 1 and Pw∈Chi(v) λ4
v = 1 and Xw∈Chi(v)
λ4
w 6 1 for all v ∈ V .
(cid:0) Xw∈Chi(v)
λ2
r2n
Suppose now that T is rootless. It is easily seen that for every r ∈ (0, 1), there
exists a sequence {rj}∞j=1 ⊆ (0, 1) such that r +P∞j=1 rj = 2 and r2 +P∞j=1 r2
j 6 1.
This fact combined with the proof of Theorem 3.1 (use (‡‡) in place of (‡)) enables
us to construct a family λ = {λv}v∈V ⊆ (0, 1) that satisfies (4.1).
Since card(Chi(u)) = ℵ0 for all u ∈ V , we infer from assertions (i) and (iii)
of Proposition 2.1, (4.1) and (†) that Sλ is injective and densely defined, and
λ) = {0}. It follows from (2.7) and the equality in (4.1) that λ2
D(S2
v = kSλevk−2
for all v ∈ V ◦, and thus
Xv∈Chi(u)
kSλevk2 = Xv∈Chi(u)
u ∈ V,
(4.1)
6 1,
λ4
v
λ2
v
which in view of Theorem 4.1 completes the proof.
(cid:3)
Remark 4.3. In view of Theorems 3.2 and 4.2, the weighted shift Sλ constructed in
the proof of implication (ii)⇒(i) of Theorem 3.1 may satisfy either of the following
two conditions: D(Sλ) * D(S∗λ) or D(Sλ) ⊆ D(S∗λ). It turns out that this general
construction always guarantees that D(S∗λ) * D(Sλ).
Indeed, since for a fixed
u ∈ V , kSλevk2 = 1/λ2
v < ∞,
we deduce that the function φ : Chi(u) ∋ v 7→ kSλevk ∈ C is unbounded, and
thus the operator Mu in ℓ2(Chi(u)) of multiplication by φ is unbounded (note that
the function λu : Chi(u) ∋ v 7→ λv ∈ C does not belong to D(Mu), and so the
definition [8, (4.2.2)] makes no sense). Applying [8, Theorem 4.2.2], we conclude
that D(S∗λ) * D(Sλ).
v for all v ∈ Des(u) (cf. (3.3)) and Pv∈Chi(u) λ2
A HYPONORMAL WEIGHTED SHIFT WHOSE SQUARE HAS TRIVIAL DOMAIN
7
Remark 4.4. It is worth pointing out that if T is a directed tree such that
card(Chi(u)) = ℵ0 for every u ∈ V , Sλ is a densely defined weighted shifts on T
with nonzero weights λ = {λv}v∈V ◦ such that D(S2
λ) = {0} (cf. Theorem 3.1) and
v0 ∈ V ◦, then the weighted shift S λ on T with nonzero weights λ = {λv}v∈V ◦
given by
λv =(λv
p1 + kSλevk2
for v 6= v0,
for v = v0,
is densely defined, D(Sλ) = D(S λ) (use [8, Proposition 3.1.3(i)]), D(S∗λ) = D(S∗λ)
(use [8, Proposition 3.4.1(iv)]), S λ is not hyponormal (use Theorem 4.1) and
D(S2
λ) = {0} (use (3.1)). Hence, if Sλ is constructed as in the proof of Theo-
rem 4.2, then by Remark 4.3 we have D(S λ) D(S∗λ).
Acknowledgement. The substantial part of this paper was written while
the first and the third authors visited Kyungpook National University during the
autumn of 2010 and the spring of 2011. They wish to thank the faculty and the
administration of this unit for their warm hospitality.
References
[1] N. I. Akhiezer, I. M. Glazman, Theory of linear operators in Hilbert space, Vol. II, Dover
Publications, Inc., New York, 1993.
[2] M. Sh. Birman, M. Z. Solomjak, Spectral theory of selfadjoint operators in Hilbert space, D.
Reidel Publishing Co., Dordrecht, 1987.
[3] E. Bishop, Spectral theory for operators on a Banach space, Trans. Amer. Math. Soc. 86
(1957), 414-445.
[4] P. Budzy´nski, Z. Jab lo´nski, I. B. Jung, J. Stochel, Unbounded subnormal weighted shifts on
directed trees, preprint 2011.
[5] P. R. Chernoff, A semibounded closed symmetric operator whose square has trivial domain,
Proc. Amer. Math. Soc. 89 (1983), 289-290.
[6] J. B. Conway, The theory of subnormal operators, Mathematical Surveys and Monographs,
Providence, Rhode Island, 1991.
[7] C. Foia¸s, D´ecompositions en op´erateurs et vecteurs propres. I., ´Etudes de ces d`ecompositions
et leurs rapports avec les prolongements des op´erateurs, Rev. Roumaine Math. Pures Appl.
7 (1962), 241-282.
[8] Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Weighted shifts on directed trees, to appear in Mem.
Amer. Math. Soc.
[9] Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Normal extensions escape from the class of weighted
shifts on directed trees, preprint 2011.
[10] J. Janas, On unbounded hyponormal operators, Ark. Mat. 27 (1989), 273-281.
[11] J. Janas, On unbounded hyponormal operators. II, Integr. Equat. Oper. Th. 15 (1992), 470-
478.
[12] J. Janas, On unbounded hyponormal operators. III, Studia Math. 112 (1994), 75-82.
[13] M. Naimark, On the square of a closed symmetric operator, Dokl. Akad. Nauk SSSR 26
(1940), 866-870; ibid. 28 (1940), 207-208.
[14] S. Ota, K. Schmudgen, On some classes of unbounded operators, Integr. Equat. Oper. Th.
12 (1989), 211-226.
[15] J. Stochel, An asymmetric Putnam-Fuglede theorem for unbounded operators, Proc. Amer.
Math. Soc. 129 (2001), 2261-2271.
[16] J. Stochel, F. H. Szafraniec, On normal extensions of unbounded operators. I, J. Operator
Theory 14 (1985), 31-55.
[17] J. Stochel and F. H. Szafraniec, On normal extensions of unbounded operators. II, Acta Sci.
Math. (Szeged) 53 (1989), 153-177.
[18] J. Stochel, F. H. Szafraniec, On normal extensions of unbounded operators. III, Spectral
properties, Publ. RIMS, Kyoto Univ. 25 (1989), 105-139.
8
Z. J. JAB LO ´NSKI, I. B. JUNG, AND J. STOCHEL
[19] J. Stochel, F. H. Szafraniec, The complex moment problem and subnormality: a polar de-
composition approach, J. Funct. Anal. 159(1998), 432-491.
[20] F. H. Szafraniec, Sesquilinear selection of elementary spectral measures and subnormality,
Elementary operators and applications (Blaubeuren, 1991), 243-248, World Sci. Publ., River
Edge, NJ, 1992.
[21] F. H. Szafraniec, On normal extensions of unbounded operators. IV. A matrix construction,
Operator theory and indefinite inner product spaces, 337-350, Oper. Theory Adv. Appl., 163,
Birkhauser, Basel, 2006.
[22] J. Weidmann, Linear operators in Hilbert spaces, Springer-Verlag, Berlin, Heidelberg, New
York, 1980.
Instytut Matematyki, Uniwersytet Jagiello´nski, ul. Lojasiewicza 6, PL-30348 Kra-
k´ow, Poland
E-mail address: [email protected]
Department of Mathematics, Kyungpook National University, Daegu 702-701, Ko-
rea
E-mail address: [email protected]
Instytut Matematyki, Uniwersytet Jagiello´nski, ul. Lojasiewicza 6, PL-30348 Kra-
k´ow, Poland
E-mail address: [email protected]
|
1806.10321 | 1 | 1806 | 2018-06-27T07:20:00 | On unitary equivalence of bilateral operator valued weighted shifts | [
"math.FA"
] | We establish a characterization of unitary equivalence of two bilateral operator valued weighted shifts with quasi-invertible weights by an operator of diagonal form. We also present an example of unitary equivalence between shifts defined on $\mathbb{C}^2$ which cannot be given by any unitary operator of diagonal form. The paper is concluded with investigation of unitary operators than can give unitary equivalence of bilateral operator valued weighted shifts. | math.FA | math |
ON UNITARY EQUIVALENCE OF BILATERAL OPERATOR
VALUED WEIGHTED SHIFTS
JAKUB KO´SMIDER
Abstract. We establish a characterization of unitary equivalence of two bi-
lateral operator valued weighted shifts with quasi-invertible weights by an
operator of diagonal form. We also present an example of unitary equivalence
between shifts defined on C2 which cannot be given by any unitary operator of
diagonal form. The paper is concluded with investigation of unitary operators
than can give unitary equivalence of bilateral operator valued weighted shifts.
1. Introduction and preliminaries
Classical weighted shift operators and their properties have already been studied
for a long time by many authors (see, e.g., [15, 2, 14, 5]). By classical weighted shifts
we understand both unilateral and bilateral weighted shifts defined on C. There
are many papers devoted to problems of weighted shifts in more general context in
which these operators are defined on arbitrary Hilbert spaces (see [11, 6, 8, 12, 3, 9]).
In some of them authors give or use results concerning unitary equivalence (see
[11, 6, 13, 12, 9]). Jab lo´nski, Jung and Stochel introduced in [10] the class of
weighted shifts on directed trees, which generalizes unilateral and bilateral shifts
with classical weights.
Unitary equivalence of unilateral operator valued weighted shifts with invertible
weights defined on arbitrary Hilbert space was characterized by Lambert in [11,
Corollary 3.3]. Orovcanec provided in [13, Theorem 1] characterization in case
shifts have quasi-invertible weights. This result was later proved with weaker as-
sumptions, namely, for unilateral shifts with weights having dense ranges by Anand,
Chavan, Jab lo´nski, Stochel in [1, Theorem 2.3]. Jab lo´nski proved in [9, Proposi-
ton 2.2] that unilateral operator valued weighted shift with invertible weights is uni-
tarily equivalent to unilateral operator valued weighted shift with weights {Tn}∞n=0
such that product Tn . . . T0 is a positive operator for all n ∈ N.
The are some partial results regarding unitary equivalence of bilateral operator
valued weighted shifts. Li, Ji and Sun proved that each bilateral operator val-
ued weighted shift with invertible weights defined on Cm for m ≥ 2 is unitarily
equivalent to a shift with upper triangular weights (see [12, Theorem 2.1]). Shields
provided in [15] characterization of unitary equivalence in case of classical bilateral
shifts. Guyker proved in [6] a result regarding unitary equivalence of bilateral oper-
ator valued weighted shift with the one having positive weights. The proof required
additional assumption i.e., normality and commutativity of weights.
2010 Mathematics Subject Classification. Primary 47B37; Secondary 47A62.
Key words and phrases. unitary equivalence, bilateral shift, partial isometry, quasi-invertible
weights.
1
2
JAKUB KO´SMIDER
In what follows, we denote by N, N+, Z, R, R+ and C the sets of non-negative
integers, positive integers, integers, real numbers, non-negative real numbers and
complex numbers, respectively. Throughout the paper by H we denote a nonzero
complex Hilbert space. The symbol B(H) stands for the C∗-algebra of all bounded
operators defined on H. All operators considered in this paper are assumed to be
linear. By R(A), N (A) and σ(A) we understand the range, the kernel and the
spectrum of operator A ∈ B(H), respectively. As usual, I ∈ B(H) stands for
the identity operator. Unitary equivalence of operators A and B ∈ B(H) will be
denoted by A ∼= B. We also write A ∼=U B to emphasize that unitary equivalence
is given by U . For a closed subspace M of H, by M⊥ we denote its orthogonal
complement. If M and N are two closed subspaces of H, which are orthogonal,
then we write M ⊥ N . We say that an operator A ∈ B(H) is quasi-invertible, if
A is injective and R(A) = H. The reader can verify that, if A ∈ B(H) is quasi-
invertible, then so is A∗. For a positive operator A ∈ B(H) we denote by A
2 the
(positive) square root of A. Operator A ∈ B(H) is called a partial isometry if
Ax = x for all x ∈ N (A)⊥. The following result is well known and it can be
found in [4, Exercise VIII.3.15].
Lemma 1.1. Let A ∈ B(H). Then the following are equivalent:
1
(i) A is a partial isometry,
(ii) A∗ is a partial isometry,
(iii) A∗A is the orthogonal projection onto N (A)⊥,
(iv) AA∗ is the orthogonal projection onto R(A),
(v) AA∗A = A,
(vi) A∗AA∗ = A∗.
Despite of the fact that the following lemma is definitely folklore, we will state
it for the reader's convenience, as we will refer to it later.
Lemma 1.2. Assume that S, T ∈ B(H) have dense ranges. If Sx = T x for
all x ∈ H, then there exists unitary operator V on H such that V S = T .
We define a Hilbert space ℓ2(Z,H) as the space ⊕n∈ZH equipped with the inner
product defined by hx, yi =P∞i=−∞hxi, yiiH for x, y ∈ ℓ2(Z,H). This space consists
of all vectors x = (. . . , x−1, x0 , x1, . . . ) satisfying P∞n=−∞ xn2
< ∞, where ·
denotes the 0th element of x. Operator U ∈ B(ℓ2(Z,H)) can be expressed as
infinite matrix [Ui,j]i,j∈Z, where Ui,j ∈ B(H) for all i, j ∈ Z.
We say that S ∈ B(ℓ2(Z,H)) is a diagonal operator if there exists a two-sided
sequence of operators {Sn}n∈Z ⊆ B(H) such that {Sn}n∈Z is bounded and
x ∈ ℓ2(Z,H).
S(. . . , x−1, x0 , x1, . . . ) = (. . . , S−1x−1, S0x0 , S1x1, . . . ),
Let {Sn}n∈Z ⊆ B(H) be a two-sided sequence of nonzero operators such that
H
{Sn}n∈Z is bounded. We define S ∈ B(ℓ2(Z,H)) by
S(..., x−1, x0 , x1, ...) = (..., S−1x−2, S0x−1 , S1x0, ...),
x ∈ ℓ2(Z,H).
Operator S is called a bilateral operator valued weighted shift on H with operator
weights {Sn}n∈Z and it will be denoted by S ∼ {Sn}n∈Z. Denote by F the unitary
bilateral operator valued weighted shift with all weights being identity operators
on H. We say that an operator S ∈ B(ℓ2(Z,H)) is of diagonal form if there exist
k ∈ Z and a diagonal operator T ∈ B(ℓ2(Z,H)) such that S = F kT .
UNITARY EQUIVALENCE OF BILATERAL SHIFTS
3
Let S ∼ {Sn}n∈Z. We can represent S by the following infinite matrix
S =
. . .
. . .
. . . S0
. . .
0
. . .
. . .
0
. . .
. . .
. . .
0
S1
0
. . .
0
0
S2
. . .
. . .
. . .
. . .
. . .
. . .
where ·
to [11], we do not assume that weights of S are invertible.
indicates the element indexed by (0, 0). It is worth noting that, as opposed
In this paper we focus on the problem of unitary equivalence of bilateral oper-
ator valued weighted shifts with quasi-invertible weights. The paper is organized
as follows. In Section 2 we investigate unitary equivalence given by operators of
diagonal form. Corollary 2.4 establishes the characterization of unitary equivalence
of bilateral operator valued weighted shifts with quasi-invertible weights given by
an operator of diagonal form. In Theorem 2.5 we prove that each bilateral operator
valued weighted shift with quasi-invertible weights is unitarily equivalent to a bilat-
eral weighted shift having positive weights. We conclude this section with proving
that bilateral operator valued weighted shift having normal and commuting weights
defined on Cm for m ≥ 2 is unitarily equivalent to a bilateral weighted shift with
weights being diagonal operators (see Proposition 2.9).
Section 3 is devoted to the problem of unitary equivalence given by operators
that are not of diagonal form and to investigation of unitary operators on ℓ2(Z,H)
that can give unitary equivalence of weighted shifts. We begin it with Example 3.1
that shows two bilateral operator valued weighted shifts defined on C2 which are
unitarily equivalent, but the unitary equivalence is not given by any operator of
diagonal form. Proposition 3.2 states that, if U ∈ B(ℓ2(Z,H)) contains exactly
two nonzero diagonals and all other elements of U are zero operators, then the
operators on these diagonals are partial isometries. We also investigate unitary
operators that give unitary equivalence of bilateral weighted shifts defined on Cm
for m ≥ 2. Proposition 3.6 states that under some additional assumptions, if
U ∈ B(ℓ2(Z, C2)) is unitary and all elements of U except for three diagonals are
zero operators, then one of the diagonals contains only zero operators.
Finally, Section 4 contains final remarks and concludes some open problems
related to unitary equivalence of bilateral weighted shifts.
2. Unitary equivalence given by an operator of diagonal form
In this section we present results related to unitary equivalence of bilateral oper-
ator valued weighted shifts given by an operator of diagonal form. It contains also
some general facts which usage is not limited to this section.
We will begin with stating the following key lemma required for further refer-
ences, which is a two-sided counterpart of [13, Lemma] (see also [11, Lemma 2.1]
and [15, Proposition 5 (a)]). Its proof is left to the reader.
Lemma 2.1. Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z and Sn, Tn be quasi-invertible for
each n ∈ Z. Assume that A ∈ B(ℓ2(Z,H)). Then the following are equivalent:
(i) AS = T A,
4
JAKUB KO´SMIDER
(ii) Ai+1,j+1Sj = TiAi,j for each i, j ∈ Z.
There is a significant difference between [13, Lemma] and the one presented
above. In the case of unilateral weighted shifts every vector in the range of a shift
has the zero as the first element. Hence, each operator intertwining two unilateral
weighted shifts has a triangular matrix.
In the case of bilateral weighted shifts
equality AS = T A does not imply triangularity of A (see Example 3.1 below).
Lemma 2.1 gives the following important result.
Corollary 2.2. Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z and Sn, Tn be quasi-invertible for
each n ∈ Z. Assume that A ∈ B(ℓ2(Z,H)) be such that AS = T A. If Ai,j 6= 0 for
some i, j ∈ Z, then Ai+n,j+n 6= 0 for all n ∈ Z.
It follows from Corollary 2.2 that the unitary operator
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
I
0
0
0
0
. . .
I
I
0
1√2
0
1√2
0
. . .
0
I
I
0
0
1√2
0
0 − 1√2
0
0
. . .
. . .
I
0
0
0
0
I
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
does not give unitary equivalence between any two bilateral operator valued weight-
ed shifts with quasi-invertible weights.
The following theorem gives a necessary and sufficient condition for two bilateral
operator valued weighted shifts with quasi-invertible weights to be unitarily equiv-
alent by an operator of diagonal form. Proof of this fact is based on the proof of
similar result for unilateral operator valued weighted shifts from [1, Theorem 2.3]
(see also [13, Theorem 1]).
Theorem 2.3. Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z and m ∈ Z be such that Sm+n, Tn,
S∗m−n−1 and T ∗
−n−1 have dense ranges for n ∈ N. Then the following are equivalent
(i) there exists U ∈ B(ℓ2(Z,H)) of diagonal form such that S ∼=U T and
U0,m 6= 0,
(ii) there exists unitary operator U0,m ∈ B(H) such that the following hold:
(a) Sm+n−1 . . . Smx = Tn−1 . . . T0U0,mx for all x ∈ H and n ∈ N+,
(b) S∗m−n . . . S∗m−1x = T ∗
−1U0,mx for all x ∈ H and n ∈ N+.
Proof. (i) ⇒ (ii). Assume that S ∼=U T , where U ∈ B(ℓ2(Z,H)) is of diagonal
form. Let n ∈ N+. Then, by Lemma 2.1,
−n . . . T ∗
Un,m+nSm+n−1 . . . Sm = Tn−1 . . . T0U0,m.
which implies (a). Let us now check that (b) also holds. Let n ∈ N+. Again, by
Lemma 2.1,
U0,mSm−1 . . . Sm−n = T−1 . . . T−nU−n,m−n
which is equivalent to the following
Sm−1 . . . Sm−nU∗
−n,m−n = U∗0,mT−1 . . . T−n.
UNITARY EQUIVALENCE OF BILATERAL SHIFTS
5
After taking adjoints we get that for all x ∈ H and n ∈ N+ it is true that
S∗m−n . . . S∗m−1x = T ∗
−n . . . T ∗
−1U0,mx,
which proves (b).
(ii) ⇒ (i). We will construct U ∈ B(ℓ2(Z,H)) of diagonal form with unitary
Un+1,m+n+1Sm+n = TnUn,m+n, n ∈ Z.
operators Un,m+n ∈ B(H) on its diagonal, which satisfy the following
(2.1)
In order to simplify formulas we introduce notation Vn := Un,m+n for n ∈ Z.
We will begin with constructing operators Vn for n ∈ N+. Since Sm and T0V0
have dense ranges and (a) holds with n = 1, then, by Lemma 1.2, there exists
unitary V1 such that V1Sm = T0V0. Now, assume that n > 1 and unitary operators
V1, . . . , Vn are already defined to be such that Vi+1Sm+i = TiVi for i ∈ {1, . . . , n−1}.
Again, we use Lemma 1.2 for operators Sm+n . . . Sm and Tn . . . T0V0, which have
dense ranges and get that there exists a unitary operator Vn+1 such that
By the above we see that
Vn+1Sm+n . . . Sm = Tn . . . T0V0.
(Vn+1Sm+n − TnVn)Sm+n−1 . . . Sm = Vn+1Sm+n . . . Sm − Tn . . . T0U0 = 0.
(2.2)
Since Sm+n−1 . . . Sm has dense range, (2.2) implies that Vn+1Sm+n = TnVn.
We will now focus on finding operators V−n for n ∈ N+. We begin with definition
of V−1. Since S∗m−1 and T ∗
−1U0,m have dense ranges and (b) holds for n = 1, by
Lemma 1.2, there exists a unitary V−1 such that V−1S∗m−1 = T ∗
−1U0. This implies
that V0Sm−1 = T−1V−1. Let n > 1. Assume that V−1, . . . , V−n+1 are already
defined unitary operators on H such that V−i+1Sm−i = T−iV−i for i ∈ {1, . . . , n−1}.
We will construct V−n such that V−n+1Sm−n = T−nV−n. It is enough to find V−n
so that the following holds
V−nS∗m−n . . . S∗m−1 = T ∗
because then we will get the following equality
−n . . . T ∗
−1V0,
V0Sm−1 . . . Sm−n = T−1 . . . T−nV−n.
We get V−n by using Lemma 1.2 for operators S∗m−n . . . S∗m−1 and T ∗
−1V0
with dense ranges. Now, we only need to show that V−n+1Sm−n = T−nV−n. We will
do this by proving that V−nS∗m−n = T ∗
−nV−n+1, which is an equivalent condition.
Let us consider the following:
(V−nS∗m−n−T ∗
Since S∗m−n+1 . . . S∗m−1 has dense range, V−nS∗m−n = T ∗
We constructed sequence {Un,m+n}n∈Z of unitary operators such that (2.1) holds.
By Lemma 2.1 it is true that S ∼=U T , where U is of diagonal form. This completes
the proof.
−nV−n+1)S∗m−n+1 . . . S∗m−1 = V−nS∗m−n . . . S∗m−1−T ∗
−nV−n+1.
−n . . . T ∗
−1V0 = 0.
−n . . . T ∗
(cid:3)
It is worth noting that, if we additionally assume that S and T have quasi-
invertible weights in Theorem 2.3, then we can choose any other operator Uk,m+k
instead of U0,m for k ∈ Z and modify the statement. In this way we get the following
result.
Corollary 2.4. Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z have quasi-invertible weights and
let m ∈ Z. Then the following are equivalent
6
JAKUB KO´SMIDER
(i) there exists U ∈ B(ℓ2(Z,H)) of diagonal form such that S ∼=U T and
U0,m 6= 0,
(ii) there exist k ∈ Z and unitary operator Uk,m+k ∈ B(H) such that the
following hold:
(a) Sm+n+k−1 . . . Sm+kx = Tn+k−1 . . . TkUk,m+kx for all x ∈ H and
(b) S∗m−n+k . . . S∗m−1+kx = T ∗
−1+kUk,m+kx for all x ∈ H
−n+k . . . T ∗
n ∈ N+,
and n ∈ N+.
Next result that we will prove is the unitary equivalence of bilateral operator
valued weighted shift with the one having positive weights. Shields proved in [15]
that each bilateral weighted shift with weights {an}n∈Z ⊆ C is unitarily equiva-
lent to the shift with weights {an}n∈Z. This fact follows from [15, Theorem 1].
Pietrzycki used it to prove that each bounded injective classical bilateral weighted
shift S satisfying S∗nSn = (S∗S)n for any n ≥ 2 is quasinormal (see [14, Theo-
rem 3.3]). Jab lo´nski, Jung and Stochel generalized Shields' result to the class of
weighted shifts on directed trees (see [10, Theorem 3.2.1]).
In the case of bilateral operator valued weighted shifts the situation is more
complicated. Guyker proved in [6, Theorem 1] that, if S ∼ {Sn}n∈Z has weights that
are commuting and normal operators, then S is unitarily equivalent to the bilateral
operator valued weighted shift with weights of the form (S∗nSn)
2 . This result is
similar to the one of Shields for shifts with classical weights. Ivanovski mentioned
in [8] that, without loss of generality, we can assume that each bilateral operator
valued weighted shift can be assumed to have positive weights. He referenced [11].
However, in [11] there is only a proof of unitary equivalence of shifts with those
of positive weights for unilateral operator valued weighted shifts with invertible
weights.
1
We will now prove the fact that each bilateral operator valued weighted shift
is unitarily equivalent to a bilateral shift with positive weights. We use argument
which is based on similar results from [13, 11] for unilateral operator valued weighted
shifts.
Theorem 2.5. Let S ∼ {Sn}n∈Z and Sn be quasi-invertible for all n ∈ Z. Then
S ∼= T , where T ∼ {Tn}n∈Z and each Tn is positive.
Proof. It follows from the polar decomposition that for each n ∈ Z there exist
unitary Un and positive Pn such that Sn = UnPn. Let P and U be diagonal
operators on ℓ2(Z,H) such that Pn = Pn+1 and Un = Un+1 for all n ∈ Z. Simple
calculation can prove that S = F U P . It is easy to verify that condition (ii) from
Theorem 2.3 is satisfied as F and F U have unitary weights. Thus there exists a
diagonal operator V ∈ B(ℓ2(Z,H)) such that V F = F U V . It is true that
S = F U P = V V ∗F U P = V F V ∗ P = V (F V ∗ P V )V ∗.
Observe that V ∗ P V is a diagonal operator. This implies that F V ∗ P V is a bilateral
operator valued weighted shift. Since unitary equivalence preserves positivity and
elements of V ∗ P V are unitarily equivalent to elements of P , the proof is completed.
(cid:3)
Now we will state a useful fact which gives necessary conditions of unitary equiv-
alence of bilateral operator valued weighted shifts given by operator of diagonal
form.
UNITARY EQUIVALENCE OF BILATERAL SHIFTS
7
Lemma 2.6. Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z have quasi-invertible weights. Sup-
pose that S ∼=U T , where U is of diagonal form and U0,k 6= 0 for some k ∈ Z. Then
Sn+k = Tn for each n ∈ Z.
Proof. Define Vn = Un,n+k for all n ∈ Z. By Lemma 2.1, Vn+1Sn+k = TnVn for each
n ∈ Z, where operators Un are unitary. Therefore, we see that Tn = Vn+1Sn+kV ∗n
and Sn+k = V ∗n+1TnVn for each n ∈ Z. This completes the proof.
(cid:3)
In the following proposition we provide necessary condition of unitary equiva-
lence given by a diagonal operator for H = C2.
Proposition 2.7. Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z be defined on C2 and have
normal weights. Assume that S ∼=U T where U is a diagonal operator. Then the
modulus of eigenvalues of corresponding weights are equal.
Proof. Since all weights are normal matrices, then they are diagonalizable. There-
fore, it is easy to see that we can diagonalize (using unitary operator) one of the
weights in each shifts. Let n ∈ Z. By the above we can assume that Sn and Tn
are diagonal matrices. By Corollary 2.4 there exists unitary V ∈ B(C2) such that
Snx = TnV x for all x ∈ C2. Let us now assume that
V =(cid:20)v1
Sn =(cid:20)s1
Tn =(cid:20)t1
s2(cid:21) ,
t2(cid:21) ,
v2
v4(cid:21)
Taking x = (1, 0) and y = (0, 1), by the previous property, we get the following
v3
0
0
0
0
system of equations:
s12 = v1t12 + v3t22,
s22 = v2t12 + v4t22.
We see that both equations are convex combinations. Also, by Lemma 2.6, it is
true that max{s1,s2} = max{t1,t2}. Since V is unitary, it must be true that
(s1 = t1
s2 = t2
or (s1 = t2
s2 = t1
which is exactly our claim.
(cid:3)
We can now use the above result to determine whether two bilateral operator
valued weighted shifts on C2 are unitarily equivalent by a diagonal operator. First,
we use Theorem 2.5 to transform both shifts to their forms with positive weights.
Then we compare the eigenvalues of the corresponding weights and check whether
their modulus are equal. It there is at least one pair of two corresponding weights
with at least one different eigenvalue, then it means that eventual unitary equiva-
lence of the shifts cannot be given by a diagonal operator. It is important to note
that moving to form with positive weights is achieved by using a diagonal operator
and, therefore, the argument presented above is correct.
Unfortunately, there is no clear dependency between spectra of weights of original
shift and the one with positive weights. Another problem is that the condition
provided in Proposition 2.7 is not sufficient. To see this let us consider the following
Example 2.8. Let H = C2. We will set Sn = Tn = I to be identity operators on
H for n ∈ Z \ {0, 1}. For n ∈ {0, 1} we define
Sn =(cid:20)s1,n
0
0
s2,n(cid:21) ,
Tn =(cid:20)t1,n
0
0
t2,n(cid:21) .
8
JAKUB KO´SMIDER
Let us fix s1,0 = t2,0 and s2,0 = t1,0 and s1,0 > s2,0 > 0. For S1 and T1 we
choose s1,1 = t1,1 and s2,1 = t2,1 and s1,1 > s2,1 > 1. Now, by Theorem 2.3,
for S and T to be unitarily equivalent by a diagonal operator we need a unitary
operator U ∈ B(H) such that:
S0x =T0U x,
S1S0x =T1T0U x,
x ∈ H.
But the above cannot be true as first equation determines that U must be equal
where u = 1 and v = 1. In this case, the second equation is not satisfied.
U =(cid:20)0 u
0(cid:21) ,
v
♦
Li, Ji and Sun proved in [12, Theorem 2.1] that bilateral weighted shift defined
on H = Ck is unitarily equivalent to the one with upper triangular weights. We
will see that, under some additional assumptions, it is possible to prove that some
bilateral operator valued weighted shifts are unitarily equivalent to the ones with
diagonal weights.
Proposition 2.9. Let S ∼ {Sn}n∈Z be a shift in ℓ2(Z, Ck) for k ≥ 2 with normal
and commuting weights. Then there is a D ∼ {Dn}n∈Z such that S ∼= D and Dn
is a diagonal operator for each n ∈ Z.
Proof. It is a well-known fact that any set of normal matrices {Ta}a∈A which com-
mutes with each other can be simultaneously diagonalized i.e. there exists a unitary
matrix V such that V TaV ∗ is diagonal for each a ∈ A (see [7, Theorem 1.3.19]).
Now, we see that a diagonal operator consisting of operators V on its diagonal gives
unitary equivalence between S and D ∼ {Dn}n∈Z where each Dn is diagonal for
every n ∈ Z.
(cid:3)
3. Unitary equivalence - the non-diagonal case
In this section we focus on investigation of unitary operators that can give unitary
equivalence of bilateral operator valued weighted shifts. Most of the results concern
only finite-dimensional Hilbert spaces.
In [15, Theorem 1] one can find a proof of the fact that in case of bilateral shifts
on ℓ2(Z, C) unitary equivalence is always given by an operator of diagonal form.
We will now see that there are bilateral operator valued weighted shifts which are
unitarily equivalent, but the unitary equivalence is not given by any operator of
diagonal form.
Example 3.1. Assume H = C2, w = 1
sn =(1,
1
n , otherwise.
2 i and define
2 − 1
if n = 0,
(3.1)
Let S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z have weights
Sn :=(cid:20) sn
−sn
sn
sn(cid:21) ,
Tn :=(cid:20) sn−1w + sn+1 ¯w sn−1 ¯w + sn+1w
−sn−1 ¯w − sn+1w sn−1w + sn+1 ¯w(cid:21)
for n ∈ Z. It is easy to see that weights of S and T are invertible, bounded and
normal. We construct unitary operator with two nonzero diagonals which gives
UNITARY EQUIVALENCE OF BILATERAL SHIFTS
9
unitary equivalence of S and T . Let us define the following operators
A =
1
2(cid:20)1 −i
1(cid:21) , B =
i
i
1
2(cid:20) 1
−i 1(cid:21) .
Both A and B are orthogonal projections onto one-dimensional subspaces. More-
over, AB = BA = 0 and A + B = I. Define
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
The reader can check that U is unitary and U S = T U .
. . .
. . .
0
. . . B
. . .
. . .
0
. . .
0
. . .
B
. . .
U =
A
A
0
0
Now, we show that it is not possible to find unitary operator of diagonal form
which would give unitary equivalence of S and T . First, one can easily verify that
Sn = √2sn for each n ∈ Z. Let us now compute the norms of operators Tn.
We find the eigenvalues of T ∗n Tn using the characteristic polynomial
The roots of W are 2s2
Now, it follows from (3.1) that
W (λ) = λ2 − 2λ(s2
n+1 and 2s2
n+1) + 4s2
n−1 + s2
n−1, hence Tn = max{√2sn−1,√2sn+1}.
n−1s2
n+1.
(3.2)
( Si = 1 if and only if i ∈ {−1, 0, 1},
Ti = 1 if and only if i ∈ {−2,−1, 0, 1, 2}.
Suppose that, contrary to our claim, S and T are unitarily equivalent by an
operator of diagonal form. By Lemma 2.6, there exists k ∈ Z that Sn+k = Tn
for all n ∈ Z. This contradicts (3.2).
♦
We presented the example of two bilateral operator valued weighted shifts that
are unitarily equivalent by an operator that is not of diagonal form. We also proved
that there is no operator of diagonal form that would give this unitary equivalence.
Let us note that, by Proposition 2.9, this example can be significantly simplified
if we diagonalize all weights before performing any computations. We leave the
details to the reader.
Example 3.1 shows even more. Let us first recall some known results. Shields
showed in [15, Theorem 1] that, if two bilateral shifts with complex weights are
unitarily equivalent, then there exists k ∈ Z such that sn = tn+k for each n ∈ Z.
Moreover, it follows from [13, Theorem 1] that, if two unilateral shifts S ∼ {Sn}n∈N,
T ∼ {Tn}n∈N with quasi-invertible weights are unitarily equivalent, then the unitary
equivalence is given by a diagonal operator. It follows from similar argument as in
Lemma 2.6 that Sn = Tn for each n ∈ N. This is not true for bilateral operator
valued weighted shifts defined on a Hilbert space of dimension greater then one.
We will now investigate unitary operators on ℓ2(Z,H) that can give unitary
equivalence of bilateral weighted shifts defined on H. Note that, as in the case
of finite-dimensional Hilbert space quasi-invertibility is the same property as in-
vertibility, Corollary 2.2 already gives us the information that, if any element in
a matrix representation of a unitary operator is nonzero, then the entire diagonal
10
JAKUB KO´SMIDER
containing this element is nonzero. Therefore, we will focus only on number of
nonzero diagonals in unitary operators.
Let U be a unitary operator acting on ℓ2(Z,H) with two nonzero diagonals.
Then there exist k1, k2 ∈ Z such that k1 6= k2 and operators Un,n+k1 , Un,n+k2
are nonzero elements from these diagonals for all n ∈ Z. From now we identify
nonzero diagonals of such operators with k1 and k2 and denote U (1)
:= Un,n+k1,
U (2)
:= Un,n+k2 for all n ∈ Z. Without loss of generality we can assume that
k2 > k1. We generalize this notation to an arbitrary number of diagonals in U .
Next proposition states that, if there are exactly two nonzero diagonals in a
n
n
unitary operator, then both diagonals contain only partial isometries.
Proposition 3.2. Suppose that U ∈ B(ℓ2(Z,H)) is a unitary operator that has
exactly two nonzero diagonals and all other elements are zero operators. Then the
elements on these diagonals are partial isometries such that elements in each row
of U have orthogonal ranges.
n
(3.3c)
(3.3a)
(3.3b)
Proof. Let us fix k = k2 − k1 > 0. For simplicity let us set An := U (1)
n , Bn := U (2)
for all n ∈ Z. Both {An}n∈Z and {Bn}n∈Z are sequences of bounded operators.
Note that U is unitary if and only if conditions
I = AnA∗n + BnB∗n,
I = A∗n+kAn+k + B∗nBn,
0 = An+kB∗n,
0 = A∗nBn,
(3.3d)
hold for all n ∈ Z. Now, we can multiply (3.3a) by An from the right and get
(3.4)
Now, by (3.3d) and (3.4), we see that BnB∗nAn = 0 and thus, by Lemma 1.1, An
is a partial isometry for all n ∈ Z. It is clear that operators Bn are also partial
isometries. From (3.3d) we deduce that R(An) is orthogonal to R(Bn) for all n ∈ Z.
This completes the proof.
An = AnA∗nAn + BnB∗nAn.
(cid:3)
It is worth noting that, using the property (3.3c), we can deduce that R(B∗n) ⊥
Next example shows that sequences {An}n∈Z, {Bn}n∈Z do not need to be se-
R(A∗n+k) for all n ∈ Z.
quences of orthogonal projections.
Example 3.3. Let H = C2. We define the following
0(cid:21) , Bn =(cid:20) 0
An =(cid:20)0 an
bn
0
0
0(cid:21) ,
where an = bn = 1 for all n ∈ Z. The reader can verify that these operators
satisfy conditions (3.3a) - (3.3d) from the proof of Proposition 3.2 and form a
unitary operator U with k1 = −1 and k2 = 1. Now we define S ∼ {Sn}n∈Z in the
following way
Sn =(cid:20)s1,n
0
0
s2,n(cid:21) ,
It is easy to check that U SU∗ is a bilateral
where s1,ns2,n 6= 0 for all n ∈ Z.
operator valued weighted shift and neither An nor Bn are orthogonal projections
for any n ∈ Z.
♦
UNITARY EQUIVALENCE OF BILATERAL SHIFTS
11
Now we will prove useful lemma that we will use later in the paper. We provide
more general version than we need, which is true for an arbitrary nonzero Hilbert
space.
Lemma 3.4. Let H be a Hilbert space and n ∈ N+. Assume that Ai ∈ B(H) are
positive operators for i ∈ {1, . . . , n} such that
(3.5)
and dim R(C) = 1. Then Ai = aiC, where ai ∈ R+ for all i ∈ {1, . . . , n} and
Pn
Proof. Let M := R(C) = lin{¯e} for some normalized ¯e ∈ H. Set B to be an
orthonormal basis of H containing ¯e. Then, by (3.5), we know that
C := A1 + ··· + An
i=1 ai = 1.
hAie, ei = 0,
e ∈ B \ {¯e}, i ∈ {1 . . . , n}.
This, Cauchy-Schwarz inequality and the square root theorem imply that
hAie, e′i2 = hA
2
i e, A
2
i e′i2
1
1
≤ hAie, eihAie′, e′i = 0,
e, e′ ∈ B(H), (e, e′) 6= (¯e, ¯e).
Thus Aie = 0 for all i ∈ {1 . . . , n} and e ∈ B \ {¯e}. Hence for all i ∈ {1 . . . , n},
Ai = aiC for some ai ∈ R+. Now, it follows from (3.5) that Pn
We will state another lemma which gives an equivalent condition for an operator
U ∈ B(ℓ2(Z,H)) with three nonzero diagonals to be a unitary operator. It is worth
noting that this result can be generalized to arbitrary diagonals, but then it is
significantly more complicated. Thus we present it only for the case in which the
three diagonals are located next to each other in the center of the matrix of U (see
(3.6) below). We leave its proof to the reader.
Lemma 3.5. Assume that U ∈ B(ℓ2(Z,H)) is an operator of the form
i=1 ai = 1.
(cid:3)
(3.6)
U =
. . .
. . .
. . .
. . . B−1 C−1
. . . A0
. . .
. . .
0
. . .
B0
A1 B1
. . .
. . .
. . .
0
C0
,
. . .
. . .
. . .
. . .
. . .
where {An}n∈Z, {Bn}n∈Z, {Cn}n∈Z ⊆ B(H). Then U is unitary if and only if the
following
(3.7a)
(3.7b)
(3.7c)
(3.7d)
(3.7e)
(3.7f)
hold for all n ∈ Z.
I = AnA∗n + BnB∗n + CnC∗n,
0 = CnA∗n+2,
0 = An+1B∗n + Bn+1C∗n,
I = A∗n+2An+2 + B∗n+1Bn+1 + C∗nCn,
0 = A∗nCn,
0 = C∗nBn + B∗n+1An+1,
12
JAKUB KO´SMIDER
The next proposition states that, under some additional assumptions, a unitary
operator U defined by (3.6) may consist of at most two nonzero diagonals.
Proposition 3.6. Assume that H is two-dimensional, U ∈ B(ℓ2(Z,H)) is a uni-
tary operator of the form (3.6) and 1 ∈ σ(CkC∗k ) ∩ σ(Ck+1C∗k+1) for some k ∈ Z.
Let S and U SU∗ be bilateral operator valued weighted shifts with invertible weights.
Then at least one of the sequences {An}n∈Z, {Bn}n∈Z or {Cn}n∈Z consists of zero
operators only.
Proof. Note that, by Corollary 2.2, if any element of any of the three sequences is
the zero operator, then all the operators in this sequence are zero operators.
First, we assume that Cn 6= 0 for all n ∈ Z. Otherwise, the proof is completed.
Now, let dim R(Cn) = 2 for some n ∈ Z. Since Cn is invertible, then, by (3.7e), we
get that A∗n = 0 which means that An = 0.
Now assume that dimR(Cn) = 1 for all n ∈ Z. Let n ∈ {k, k + 1}. Since
1 ∈ σ(CnC∗n), then Zn := I − CnC∗n is not invertible positive operator. Then, by
(3.7a), we see that
AnA∗n + BnB∗n = Zn
so, by Lemma 3.4, we have AnA∗n = anZn and BnB∗n = bnZn, where an + bn = 1. If
anbn = 0 for any n ∈ {k, k + 1}, then the proof is completed. Otherwise, by (3.7e),
for n ∈ {k, k + 1},
(3.8)
Hence (3.7f) with n = k implies 0 = B∗k+1Ak+1 which, together with (3.8) for
n = k + 1 and the fact that R(Ak+1) 6= {0}, lead to contradiction.
(cid:3)
R(Cn) ⊥ R(An) = R(AnA∗n) = R(BnB∗n) = R(Bn).
Observe that Proposition 3.6 remains true, if we replace the assumption that
1 ∈ σ(CkC∗k ) ∩ σ(Ck+1C∗k+1) by the assumption that 1 ∈ σ(A∗kAk) ∩ σ(A∗k+1Ak+1)
for some k ∈ Z.
Proposition 3.7. Let H be two-dimensional Hilbert space and U be a unitary
operator defined as in (3.6). Assume that S and U SU∗ are bilateral operator valued
weighted shifts with invertible weights and dimR(Bn) ≤ 1 for any n ∈ Z. Then for
all n ∈ Z either An or Cn is a partial isometry.
Proof. Note that the same argument as in Proposition 3.6 can be used to cover the
case when dimR(Cn) 6= 1 for any n ∈ Z.
Let us now assume that dim R(Cn) = 1 for all n ∈ Z. Fix n ∈ Z. By the above it
us true that dimR(BnB∗nCn) is equal to 0 or 1. If it is equal to 1, then we see that
R(BnB∗nCn) = R(BnB∗n), as dimR(BnB∗n) ≤ 1. Therefore A∗nBnB∗n = 0. Hence
An is a partial isometry. Now assume dimR(BnB∗nCn) = 0. This implies that Cn
is a partial isometry.
(cid:3)
The next result states that there cannot be more then m nonzero diagonals which
contain partial isometries in unitary operator giving unitary equivalence of bilateral
operator valued weighted shifts defined on m-dimensional Hilbert space for m ≥ 2.
Proposition 3.8. Let H be a m-dimensional Hilbert space for m ≥ 2 and let S ∼
{Sn}n∈Z, T ∼ {Tn}n∈Z have quasi-invertible weights. Assume U ∈ B(ℓ2(Z,H)) is
unitary and its matrix representation consists of partial isometries only. If U S =
T U , then U has at most m nonzero diagonals and all other elements of U are zero
operators.
UNITARY EQUIVALENCE OF BILATERAL SHIFTS
13
It follows from the Lemma 1.1 that P (j)
Proof. By the fact that U U∗ = I we get that Pj∈Z U (j)
for all j ∈ Z. It is a well-known fact that, if Pj∈Z P (j)
then R(P (i)
from the fact that H is m-dimensional and from Corollary 2.2.
n )∗ = I for all n ∈ Z.
n )∗ is an orthogonal projection
is an orthogonal projection,
n ) for all i,j ∈ Z such that i 6= j. The rest follows directly
n ) ⊥ R(P (j)
n (U (j)
n := U (j)
n (U (j)
(cid:3)
n
The next example shows that it is possible to find unitary operator with three
nonzero diagonals that give unitary equivalence between bilateral operator valued
weighted shifts defined on C2.
Example 3.9. Assume H = C2. First, let us define unitary operator U of the form
(3.6) with three nonzero diagonals, where
1√2# , Cn+1 = −An,
A2n =(cid:20) 1√2
for all n ∈ Z. It can be verified that conditions (3.7a) - (3.7f) are satisfied. Hence
U is a unitary operator. Let us now define S ∼ {Sn}n∈Z in the following way
0 − 1√2(cid:21) , Bn =" 1√2
0(cid:21) , A2n+1 =(cid:20)0
0
0
0
0
0
Sn =(cid:20) 0
(−1)n
The reader can check that U S = SU .
(−1)n
0 (cid:21) , n ∈ Z.
♦
4. Further remarks
Example 3.9 shows that it is possible to find unitary operator with three nonzero
diagonals, which gives unitary equivalence of bilateral weighted shifts defined on
C2, however, this unitary equivalence can be given by the identity operator.
It
is an open question, whether for S ∼ {Sn}n∈Z, T ∼ {Tn}n∈Z, where H is a two-
dimensional Hilbert space, S ∼= T implies that there exists U that has at most
two nonzero diagonals with all other elements of U being zero operators such that
S ∼=U T .
If one proves that any unitary equivalence of bilateral shifts defined
on finite-dimensional Hilbert space can be given by an operator consisting only of
partial isometries, then Proposition 3.8 gives the positive answer.
Another interesting problem for further investigation, which comes up naturally,
is the problem of characterization of unitary equivalence of bilateral shifts defined
on finite-dimensional Hilbert space. Corollary 2.4 gives characterization of unitary
equivalence given only by an operator of diagonal form. Example 3.1 shows that
there is a rich class of unitary operators in ℓ2(Z, Ck) which are not of diagonal
form and can give unitary equivalence of bilateral weighted shifts. Clearly, we see
that the problem of complete characterization is more complicated than in case
of unilateral operator valued weighted shifts and bilateral weighted shifts having
classical weights.
Acknowledgments
The author would like to thank Zenon Jab lo´nski for insightful discussions con-
cerning the subject of the paper.
14
JAKUB KO´SMIDER
References
[1] A. Anand, S. Chavan, Z. J. Jab lo´nski, J. Stochel, Complete systems of unitary invariants for
some classes of 2-isometries, https://arxiv.org/abs/1806.03229
[2] A. Athavale, On completely hyperexpansive operators, Proc. Amer. Math. Soc., 124 (1996),
3745-3752.
[3] A. Bourhim, C. E. Chidume, The single-valued extension property for bilateral operator
weighted shifts, Proc. Amer. Math. Soc., 133 (2004), 485-491.
[4] J. B. Conway, A Course in Functional Analysis, Springer-Verlag, New York, Inc., 1990.
[5] G. P. Geh´er, Bilateral weighted shift operators similar to normal operators, Oper. Matrices,
10(2), 2016.
[6] J. Guyker, On reducing subspaces of normally weighted bilateral shifts, Houston J. Math., Vol.
11, No. 4, 1985.
[7] R. A. Horn, C. R. Johnson, Matrix analysis, Cambridge University Press, 1990.
[8] N. Ivanovski, Similiarity and quasisimiliarity of bilateral operator valued weighted shifts, Mat.
Bilten, 17, 1993, 33-37.
[9] Z. J. Jab lo´nski, Hyperexpansive operator valued unilateral weighted shifts, Glasg. Math. J., 46
(2004) 405-416.
[10] Z. J. Jab lo´nski, I. B. Jung, J. Stochel, Weighted Shifts on Directed Trees, Mem. Amer. Math.
Soc., 1017, 107 (2012).
[11] A. Lambert, Unitary equivalence and reducibility of invertibly weighted shifts, Bull. Aust.
Math. Soc., 5 (1971), 157-173.
[12] J. X. Li, Y. Q. Ji, S. L. Sun, The essential spectrum and Banach reducibility of operator
weighted shifts, Acta Math. Sin., English Series, Vol. 17, 3 (2001) 413-424.
[13] M. Orovcanec, Unitary equivalence of unilateral operator valued weighted shifts with quasi-
invertible weights, Mat. Bilten, 17, 1993, 45-50.
[14] P. Pietrzycki, The single equality A∗nAn = (A∗A)n does not imply the quasinormality of
weighted shifts on rootless directed trees, J. Math. Anal. Appl., 435, (2016), 338-348.
[15] A. Shields, Weighted shift operators, analytic function theory, Math. Surveys, Amer. Math.
Soc., Providence, 1974, 49-128.
Instytut Matematyki, Uniwersytet Jagiello´nski, ul.
Lojasiewicza 6, PL-30348 Kra-
k´ow, Poland
E-mail address: [email protected]
|
1808.09711 | 2 | 1808 | 2018-09-21T10:09:28 | Fine properties of functions with bounded variation in Carnot-Carath\'eodory spaces | [
"math.FA",
"math.MG"
] | We study properties of functions with bounded variation in Carnot-Ca\-ra\-th\'eo\-do\-ry spaces. We prove their almost everywhere approximate differentiability and we examine their approximate discontinuity set and the decomposition of their distributional derivatives. Under an additional assumption on the space, called property $\mathcal R$, we show that almost all approximate discontinuities are of jump type and we study a representation formula for the jump part of the derivative. | math.FA | math |
FINE PROPERTIES OF FUNCTIONS WITH BOUNDED VARIATION IN
CARNOT-CARATHÉODORY SPACES
SEBASTIANO DON AND DAVIDE VITTONE
Abstract. We study properties of functions with bounded variation in Carnot-Carathéodory
spaces. We prove their almost everywhere approximate differentiability and we examine their
approximate discontinuity set and the decomposition of their distributional derivatives. Under
an additional assumption on the space, called property R, we show that almost all approximate
discontinuities are of jump type and we study a representation formula for the jump part of
the derivative.
1. Introduction
A lot of effort was devoted in the last decades to the development of Analysis and Geometry in
general metric spaces and, in particular, to the study of functions with bounded variation (BV ).
Carnot-Carathéodory (CC) spaces are among the most fruitful settings where BV functions have
been introduced ([10, 20]), see also [8, 12, 19, 22, 23, 24] and the more recent [3, 5, 6, 9, 11,
15, 31, 35, 44]. The aim of this paper is to give some contributions to this research lines by
establishing "fine" properties of BV functions in CC spaces. A non-trivial part of our work
consists in fixing the appropriate language in a consistent and robust manner.
A CC space is the space Rn endowed with the Carnot-Carathéodory distance d (see (1))
arising from a fixed family X = (X1, . . . , Xm) of smooth, linearly independent vector fields
(called horizontal) in Rn satisfying the Hörmander condition, see (2). As customary in the
literature, we always assume that metric balls are bounded with respect to the Euclidean
topology. Moreover, we work in equiregular CC spaces, where a homogeneous dimension Q,
usually larger than the topological dimension n, can be defined; recall that any CC space can
be lifted to an equiregular one, see e.g. [42].
The space BVX of function with bounded X-variation consists of those functions u whose
derivatives X1u, . . . , Xmu in the sense of distributions are represented by a vector-valued mea-
sure DX u with finite total variation DXu. These functions have been extensively studied in
the literature and important properties have been proved, like coarea formulae, approximation
theorems, Poincaré inequalities.
We now describe some of the results we prove in this paper. The first one, Theorem 1.1
below, concerns the almost everywhere approximate X-differentiability (see Section 2.3) of BVX
functions; its classical counterpart is very well-known, see e.g. [2, Theorem 3.83]. As customary,
we denote by Da
Xu, respectively, the absolutely continuous and singular part of DX u
with respect to the Lebesgue measure L n.
Xu and Ds
Date: September 24, 2018.
2010 Mathematics Subject Classification. 26B30, 53C17, 49Q15, 28A75.
Key words and phrases. Functions with bounded variation, Carnot-Carathéodory spaces.
The authors are supported by the University of Padova Project Networking and STARS Project "Sub-
Riemannian Geometry and Geometric Measure Theory Issues: Old and New" (SUGGESTION), and by
GNAMPA of INdAM (Italy) project "Campi vettoriali, superfici e perimetri in geometrie singolari". The second
named author wishes to ackowledge the support and hospitality of FBK-CIRM (Trento), where part of this
paper was written.
1
2
DON AND VITTONE
Theorem 1.1. Let (Rn, X) be an equiregular CC space, let Ω ⊆ Rn be an open set and let
u ∈ BVX (Ω; Rk). Then u is approximately X-differentiable at L n-almost every point of Ω.
Moreover, the approximate X-gradient coincides L n-almost everywhere with the density of
Da
X u with respect to L n.
The proof of Theorem 1.1 is based on Lemma 3.12, that is, on a suitable extension to CC
spaces of the inequality
B(p,r)
u(q) − u(p)
q − p
dL n(q) ≤ C 1
0
Du(B(p, tr))
tn
dt
valid for a classical BV function u on Rn. Lemma 3.12 answers an open problem stated in [5]
and it is new even in Carnot groups. We only recall that Carnot groups are connected, simply
connected and nilpotent Lie groups whose Lie algebra is stratified, and we refer to [18, 38, 31, 30]
for more detailed introduction to the subject. Carnot groups possess a canonical CC structure
obtained by fixing a basis X1, . . . , Xm of the first layer of the Lie algebra of left-invariant vector
fields; their importance in the theory stems from the fact that they constitute the infinitesimal
models of equiregular CC spaces, a fact that we heavily use in this paper.
Theorem 1.1 was proved in the setting of Carnot groups in [5] together with the following
result, which we also extend to our more general setting. We denote by H Q−1 the Hausdorff
measure of dimension Q − 1 and by Su the set of points where a function u does not possess
an approximate limit in the sense of Definition 2.19.
Theorem 1.2. Let (Rn, X) be an equiregular CC space, let Ω ⊆ Rn be an open set and let
u ∈ BVX(Ω; Rk). Then Su is contained in a countable union of sets with finite H Q−1 measure.
In the classical theory, an important object associated with a BV function u is its jump
set: roughly speaking, this is the set of points p for which there exist u+(p) 6= u−(p) and a
unit direction νu(p) such that, for small r > 0, u is approximately equal to u+(p) on half of
B(p, r) and to u−(p) on the complementary half of B(p, r), the two halves being separated
by an hyperplane orthogonal to νu(p). In this paper we introduce the notion of approximate
X-jumps, see Definition 2.24: this requires a certain amount of preliminary work, expecially
about "fine" local properties of hypersurfaces with intrinsic C 1 regularity (C 1
We denote by Ju ⊆ Su the set of X-jump points of u and by (u+(p), u−(p), νu(p)) the
X).
approximate X-jump triple (see Definition 2.24) at a point p ∈ Ju. The measures
Dj
X u := Ds
Xu Ju,
Dc
X u := Ds
Xu (Ω \ Ju),
are called, respectively, jump part and Cantor part of DXu. We want to study some further
properties of DXu and its decomposition
DXu = Da
Xu + Ds
Xu = Da
Xu + Dc
Xu + Dj
Xu.
We state some of them in the following result, which is a consequence of Theorem 3.16 and
Proposition 3.3.
Theorem 1.3. Let (Rn, X) be an equiregular CC space and consider an open set Ω ⊆ Rn, a
function u ∈ BVX (Ω; Rk) and a Borel set B ⊆ Ω. Then the following facts hold:
(i) there exists λ : Rn → (0, +∞) (not depending on Ω nor u) locally bounded away from 0
such that DXu ≥ λu+ − u−S Q−1 Ju;
(ii) if H Q−1(B) = 0, then DXu(B) = 0;
(iii) if H Q−1(B) < +∞ and B ∩ Su = ∅, then DXu(B) = 0;
(iv) Da
Xu = DXu (Ω \ S) and Ds
Xu = DXu S, where
DXu(B(p, r))
S :=(cid:26)p ∈ Ω : lim
r→0
= +∞(cid:27) ;
rQ
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
3
(v) Ju ⊆ Θu, where Θu ⊆ S is defined by
Θu :=(cid:26)p ∈ Ω : lim inf
r→0
DXu(B(p, r))
rQ−1
> 0(cid:27) .
However, for classical BV functions much stronger results than Theorems 1.1 and 1.3 are
indeed known: some of them are proved in the present paper also for BVX functions under the
additional assumption that the space (Rn, X) satisfies the following natural condition.
Definition 1.4 (Property R). Let (Rn, X) be an equiregular CC space with homogeneous
dimension Q. We say that (Rn, X) satisfies the property R if, for every open set Ω ⊆ Rn and
every E ⊆ Rn with locally finite X-perimeter in Ω, the essential boundary ∂∗E ∩ Ω of E is
countably X-rectifiable, i.e., there exists a countable family (Si)i∈N of C 1
X hypersurfaces such
that H Q−1(∂∗E ∩ Ω \ ∪i∈NSi) = 0.
Recall that a measurable set E ⊆ Rn has locally finite X-perimeter in Ω if its characteristic
function χE has locally bounded X-variation in Ω, while we refer to Definition 2.21 for the
essential boundary ∂∗E.
It was proved in the fundamental paper [1] that the X-perimeter
measure DXχE of E can be represented as θH Q−1 ∂∗E for a suitable positive function θ
that is locally bounded away from 0, see Theorem 2.39.
The validity of property R ("rectifiability") for general equiregular CC spaces is an interesting
open question even in Carnot groups (see [4] for a partial result). However, property R is
satisfied, besides in Euclidean spaces ([13]), in several interesting situations like Heisenberg
groups [22], Carnot groups of step 2 [23] and Carnot groups of type ⋆ [35]:
in particular,
Theorems 1.5, 1.6 and 1.7 below hold is such classes. We conjecture that property R holds also
in all CC spaces of step 2, see [3]. Building on the results of [14], we prove in Section 4 the
validity of the weaker property LR ("Lipschitz rectifiability", see Definition 3.4) in all Carnot
groups satisfying the algebraic property (47) below; in particular, a weaker version of Theorem
1.5 holds in such groups, see Theorem 3.5.
The first result we are able to prove assuming property R is a refinement of Theorem 1.2
and, roughly speaking, it states that H Q−1-almost all singularities of a BVX function are of
jump type.
Theorem 1.5. Let (Rn, X) be an equiregular CC space satisfying property R, let Ω ⊆ Rn be an
open set and let u ∈ BVX (Ω; Rk). Then Su is countably X-rectifiable and H Q−1(Su \ Ju) = 0.
Assuming property R, also Theorem 1.3 can be refined as follows.
Theorem 1.6. Under the assumption and notation of Theorem 1.3, assume that (Rn, X) sat-
isfies property R. Then
(i) H Q−1(Θu \ Ju) = 0 and Dj
(ii) Dc
(iii) if B ⊆ Ω is such that H Q−1 B is σ-finite, then Dc
Xu = DXu (S \ Θu);
X u = DXu Θu;
Xu(B) = Da
Xu(B) = 0.
Theorem 1.6 is part of Theorem 3.16. We also mention that, assuming property R, one can
define a precise representative up of u (see (45)) and prove that the convergence of the mean
values fflB(p,r) u dL n to up(p) holds, as r → 0, for H Q−1-almost every p. See Theorem 3.14.
Eventually, a further natural assumption, property D ("density", see Definition 3.8), concern-
ing the local behavior of the spherical Hausdorff measure S Q−1 of C 1
X hypersurfaces, allows to
obtain a stronger result about the jump part Dju, see Theorem 1.7. Property D is satisfied
in Riemannian manifolds (trivially), Heisenberg groups, Carnot groups of step 2 and Carnot
groups of type ⋆, see section 4; its validity in more general settings is an interesting open prob-
lem that will be object of future investigations. Theorem 1.7 follows from the more general
4
DON AND VITTONE
Theorem 3.10, which deals with a representation of the restriction of DXu to any countably
X-rectifiable set R.
Theorem 1.7. Let (Rn, X) be an equiregular CC space satisfying properties R and D; then,
there exists a function σ : Rn × Sm−1 → (0, +∞) such that, for every open set Ω ⊆ Rn and
every u ∈ BVX (Ω; Rk), one has
Dj
Xu = σ(·, νu)(u+ − u−) ⊗ νu S Q−1 Ju.
The paper is structured as follows.
In Section 2 we introduce the preliminary material
about CC spaces and their nilpotent approximation (Section 2.1), C 1
X hypersurfaces and X-
rectifiable sets (Section 2.2), approximate X-jumps and X-differentiability (Section 2.3) and
BVX functions (Section 2.4). Most of the material in Sections 2.2 and 2.3 is original. Section
3 contains the proof of our results, while in Section 4 we discuss some classes of Carnot groups
satisfying properties R, LR and/or D. Eventually, we collected in Appendix A some useful
result from Geometric Measure Theory in metric spaces and in Appendix B the proofs of some
(new but) technical results (Borel regularity, etc.) about the approximate X-jump and the
approximate X-differentiability sets.
Acknowledgements. It is a pleasure to thank V. Magnani, R. Monti, D. Morbidelli and D.
Pallara for their interest in this paper and for several stimulating discussions.
2. Preliminaries
2.1. Carnot-Carathéodory spaces and nilpotent approximation. In what follows Ω will
denote an open set in Rn and X = (X1, . . . , Xm) an m-tuple (m ≤ n) of smooth and linearly
independent vector fields on Rn, with 2 ≤ m ≤ n. We say that an absolutely continuous curve
γ : [0, T ] → Rn is an X-subunit path joining p and q if γ(0) = p, γ(T ) = q and there exist
j=1 h2
j ≤ 1 and for almost every t ∈ [0, T ] one has
h1, . . . , hm ∈ L∞([0, T ]; R) such thatPm
For every p, q ∈ Rn, we define the quantity
γ(t) =
h(t)Xj(γ(t)).
mXj=1
d(p, q) := inf {T > 0 : ∃ a X-subunit path γ joining p and q} ,
(1)
where we agree that inf ∅ = +∞.
A sufficient condition that makes d a metric on Rn is the following
Theorem 2.1 (Chow-Rashevsky). Suppose that
∀ p ∈ Rn
Lie{X1, . . . , Xm}(p) = TpRn ∼= Rn,
(2)
where Lie{X1, . . . , Xm}(p) denotes the linear span of all iterated commutators of the vector
fields X1, . . . , Xm computed at p. Then d is a distance.
We will refer to (2) as Hörmander condition. When (2) holds, the couple (Rn, X) is said to
be a Carnot-Carathéodory space of rank m. We denote by B(p, r) the d-ball of center p ∈ Rn
and radius r > 0.
For every p ∈ Rn and for every i ∈ N we denote by Li(p) the linear span of all the commutators
of X1, . . . , Xm up to order i computed at p. Notice that Lie{X1, . . . , Xm}(p) =Si∈N Li(p). We
say that (Rn, X) is equiregular if there exist natural numbers n0, n1, . . . , ns such that
0 = n0 < n1 < · · · < ns = n
and
∀ p ∈ Rn dim Li(p) = ni.
The natural number s is called step of the Carnot-Carathéodory space.
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
5
In the following theorem we resume some well-known facts about the geometry of equiregular
CC spaces, see e.g. [41, 36]. Recall that a Radon measure µ on a metric space (M, d) is doubling
if there exists C > 0 such that
µ(B(x, 2r)) ≤ Cµ(x, r)
for every x ∈ M and every r > 0.
Theorem 2.2. Let (Rn, X) be an equiregular CC space of step s. Then the following facts
hold.
(i) For every compact set K ⊆ Rn there exists M ≥ 1 such that
1
M
p − q ≤ d(p, q) ≤ Mp − q
1
s
for any p, q ∈ K.
(ii) The Hausdorff dimension of the metric space (Rn, d) is Q :=Ps
(iii) The metric measure space (Rn, d, L n) is locally Ahlfors Q-regular, i.e., for every com-
pact set K ⊆ Rn there exist R > 0 and C > 1 such that for every p ∈ K and for every
r ∈ (0, R)
i=1 i(ni − ni−1).
In particular, (Rn, d, L n) is locally doubling.
1
C
rQ ≤ L n(B(p, r)) ≤ CrQ.
(3)
As customary, we assume from now on that the metric balls B(p, r) are bounded with respect
to the Euclidean metric in Rn; this implies that the CC space (Rn, X) is geodesic, i.e., that
for every p, q ∈ Rn there exists a X-subunit curve realizing the infimum in (1). The existence
of length minimizing curves implies that, for every p ∈ Rn and for every r > 0, one has
L n(∂B(p, r)) = 0; see Proposition A.9.
Definition 2.3 (Adapted exponential coordinates). Let (Rn, X) be an equiregular CC space
and let p ∈ Rn be fixed; choose an open neighborhood V ⊆ Rn of p and smooth vector fields
Y1, . . . , Yn such that
• Yi = Xi for any i = 1, . . . , m;
• for every k = 1, . . . , s the vector fields Ynk−1+1, . . . , Ynk are chosen among the k-order
commutators of X1, . . . , Xm;
• for every q ∈ V and every k = 1, . . . , s the set {Y1(q), . . . , Ynk(q)} is a basis of Lk(q).
Then there exists a neighborhood U of 0 in Rn for which the map
F : U → Rn
x 7→ exp(x1Y1 + · · · + xnYn)(p)
(4)
is well defined. We say that (x1, . . . , xn) are adapted exponential coordinates around p.
The definition of F depends on the point p; when confusion may arise, we underline this
dependence by using the notation Fp to denote (for any x ∈ Rn for which it is defined) the
map Fp(x) := exp(x1Y1 + · · · + xnYn)(p). When needed, we will also write F (p, x) to denote
exp(x1Y1 + · · · + xnYn)(p); notice that, for every bounded set V ⊆ Rn, one can find an open
neighborhood U of 0 in Rn such that F is well defined in V × U.
For every p ∈ Rn and every j = 1, . . . , m we define
p (Xj ◦ Fp).
eXj := dF −1
It is readily seen that if X satisfies the Hörmander condition, then also eX does and we denote
by ed the CC distance in (a suitable open subset of) Rn associated with the m-tuple of vector
fields eX = (eX1, . . . , eXm), and by eB(x, r) the metric balls associated with ed. Again, when
confusion may arise we shall use the notation eBp(x, r) to specify that the metric ball is induced
6
DON AND VITTONE
it is easy to verify that for every p ∈ Rn and every sufficiently small r > 0 one has
by the map Fp. Since dFp(0)ej = Yj(p), we have eXj(0) = ej for every j = 1, . . . , m. Moreover
in particular, Fp(eB(x, r)) = B(Fp(x), r).
d(Fp(x1), Fp(x2)) = ed(x1, x2)
p )#L n, i.e., the measure defined for every Borel set A
∀ x1, x2 ∈ eB(0, r);
Remark 2.4. Let us consider µp := (F −1
in Rn by
µp(A) = L n (Fp(A)) = A
det ∇Fp dL n.
It is easy to see that, whenever 0 < ε < det ∇Fp(0), there exists an open neighborhood U of
0 such that
(det ∇Fp(0) − ε) L n U ≤ µp U ≤ (det ∇Fp(0) + ε) L n U.
(5)
Definition 2.5 (Degree, dilations and pseudo-norm). If (Rn, X) is an equiregular CC space
and p, Y1, . . . , Yn are as in Definition 2.3, we define the degree wj of the coordinate j by Yj(p) ∈
Lwj (p) \ Lwj−1(p) or, equivalently, by nwj−1 < j ≤ nwj . For every r > 0, the anisotropic dilation
δr : Rn → Rn is defined by
δr(x) := (x1, . . . , rwixi, . . . , rsxn) .
(6)
We say that a function f : Rn → R is δ-homogeneous of degree w ∈ N if for every p ∈ Rn and
every λ > 0 one has f (δλp) = λwf (p). We also introduce the pseudo-norm
kxk :=
nXj=1
xj1/wj ,
x ∈ Rn
and the pseudo-balls
A(r) := {x ∈ Rn : kxk ≤ r} .
(7)
Clearly, δr (A(1)) = A(r).
The following result is proved in [41].
Theorem 2.6. Let K ⊆ Rn be a compact set in an equiregular CC space (Rn, X) and let U be
a neighborhood of 0 such that, for every p ∈ K, the map Fp is well-defined in U. Then there
exists C > 1 such that for every x ∈ U and every p ∈ K we have
The following theorem is classical, see e.g. [7] or [39]. For an introduction to Carnot groups
(also known as stratified groups) see for instance [18, 38, 31, 30].
1
C
kxk ≤ edp(0, x) ≤ Ckxk.
Theorem 2.7. Let (Rn, X) be an equiregular CC space and let p ∈ Rn be fixed; then, there
exists a family bX := (bX1, . . . , bXm) of polynomial vector fields in Rn such that
(i) for every j = 1, . . . , m, bXj is 1-homogeneous, i.e. (dδr)[bXj] = rbXj ◦ δr for all r > 0;
(ii) for every j = 1, . . . , m we have r(dδr−1)[eXj ◦ δr] → bXj in C ∞
(iii) the couple (Rn, bX) is associated with a Carnot group structure on Rn;
(iv) ([39, Remark 2.6]) bX can be completed to a basis bX1, . . . , bXn of the Lie algebra of the
Carnot group in such a way that x = exp(Pn
j=1 xjbXj)(0) for any x ∈ Rn.
loc(Rn);
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
7
on the point p.
The vector fields bX1, . . . , bXm introduced in Theorem 2.7 are known in the literature as the
nilpotent approximation of X1, . . . , Xm at the point p; we will say that the structure (Rn, bX)
is tangent to (Rn, X) at p. We shall denote by bd the Carnot-Carathéodory distance associated
with bX and by bB the corresponding balls; recall that bd(δrx, δry) = rbd(x, y) for any r > 0 and
x, y ∈ Rn. When confusion may arise, we shall use the notation bBp,bdp to specify the dependence
By the Carnot group structure there exists bC = bCp > 0 such that
The constant bC depends on p; however, given a compact set K ⊆ Rn, there exists M > 0 such
that 1/M ≤ bCp ≤ M for any p ∈ K. See Remark 2.10 below.
L n(bBp(x, r)) = bCrQ
Proposition 2.8. Let (Rn, X) be an equiregular CC space, and let r > 0. Then for every
p ∈ Rn one has
We will need later the following simple result.
∀ x ∈ Rn, r > 0.
(8)
Proof. By well-known properties of Carnot groups and Theorem 2.7 (iv) we have
−x = exp −
x ∈ bBp(0, r) ⇐⇒ −x ∈ bBp(0, r).
xjbXj! (0) ="exp nXj=1
nXj=1
xjbXj! (0)#−1
bd(0, −x) = bd(0, x−1) = bd(x · 0, x · x−1) = bd(x, 0).
which combined with the left invariance of bd with respect to the group operation implies
This concludes the proof.
= x−1,
We recall for future references the following well-known result, for which we refer e.g. to
[7, 36].
(cid:3)
(9)
Theorem 2.9. Let (Rn, X) be an equiregular CC space and let p ∈ Rn be fixed; then
lim
r→0 sup(edp(x, y) −bdp(x, y)
r
: x, y ∈ eBp(0, r))! = 0.
In particular, for any ε > 0, there exists R > 0 such that
Remark 2.10. Let K ⊆ Rn be a compact set; then there exists M ≥ 1 such that the constant
bBp(0, (1 − ε)r) ⊆ eBp (0, r) ⊆ bBp(0, (1 + ε)r)
for any r ∈ (0, R).
bC = bCp appearing in (8) satisfies
1
M
≤ bCp ≤ M
∀ p ∈ K.
This follows because, by Theorem 2.9, for any p ∈ K
r→0
bCp = lim
=
L n(bBp(0, r))
rQ
1
det ∇Fp(0)
lim
r→0
= lim
r→0
L n(B(p, r))
L n(eBp(0, r))
rQ
rQ
= lim
r→0
L n(F −1
p (B(p, r)))
rQ
and one can conclude by using Theorem 2.2 (iii) and the smoothness of F (p, x).
8
DON AND VITTONE
2.2. Hypersurfaces of class C 1
X. This section is devoted to the study of hypersurfaces with
intrinsic C 1 regularity; we work in a fixed equiregular CC space (Rn, X). As customary, given
an open set Ω ⊆ Rn we denote by C 1
X(Ω) the space of continuous functions f : Ω → R such
that the derivatives X1f, . . . , Xmf are represented, in the sense of distributions, by continuous
functions.
Definition 2.11 (Hypersurface of class C 1
every p ∈ S there exist R > 0 and f ∈ C 1
X). We say that S ⊆ Rn is a C 1
X hypersurface if for
X(B(p, R)) such that the following facts hold
(i) S ∩ B(p, R) = {q ∈ B(p, R) : f (q) = 0};
(ii) Xf 6= 0 on B(p, R).
In this case, for every p in S we define the horizontal normal νS(p) ∈ Sm−1 to S at p letting
νS(p) :=
Xf (p)
Xf (p)
.
The horizontal normal is well-defined up to a sign and, in particular, it does not depend on
the choice of f : this is a consequence, for instance, of Corollary 2.14, below.
We will also use the notion of intrinsic Lipschitz regularity for hypersurfaces introduced in
[46]. In the next definition, the Lipschitz continuity of f is understood with respect to the CC
distance; recall that f : Ω → R is locally Lipschitz on an open set Ω ⊆ Rn if and only if it is
continuous and its distributional derivatives X1f, . . . , Xmf belong to L∞
loc(Ω); see [21, 25].
Definition 2.12 (X-Lipschitz hypersurface). We say that S ⊆ Rn is an X-Lipschitz hyper-
surface if for every p ∈ S there exist R > 0 and a Lipschitz map f : B(p, R) → R such
that
(i) B(p, R) ∩ S = {q ∈ B(p, R) : f (q) = 0};
(ii) there exist C > 0 and 1 ≤ j ≤ m such that Xjf ≥ C L n-a.e. on B(p, R).
Hypersurfaces with X-Lipschitz or C 1
X regularity have locally finite (Q − 1)-dimensional
Hausdorff measure, see [46].
Given ν ∈ Rm we define eLν : Rn → R letting
eLν(x) :=
νixi.
mXi=1
(10)
This notation will be extensively used throughout the paper. The following proposition shows
Proposition 2.13. Let p ∈ Rn, R > 0 and f ∈ C 1
that the maps eLν provide a sort of first-order "linear" approximation for C 1
: x ∈ eB(0, r))! = 0.
r→0 sup(f (Fp(x)) − f (p) −eLXf (p)(x)
X(B(p, R)) be fixed; then
lim
r
X functions.
Proof. It is not restrictive to assume that f (p) = 0. Let r ≤ R and take x ∈ eB(0, r). Set
d := ed(x, 0) and take a geodesic γ ∈ Lip([0, d]; Rn) such that γ(0) = 0, γ(d) = x and there
exists h : [0, d] → Rm such that for L 1-a.e. t ∈ [0, d] we have
h(t) = 1
and
γ(t) =
mXj=1
hj(t)eXj(γ(t)).
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
9
0
0
0
0
Notice that eXj(0) = ej, hence there exists C > 0 such that eXj(y) − ej ≤ Cr for every
y ∈ eB(0, r) and every j = 1, . . . , m. Therefore, for every k = 1, . . . , m
mXj=1
d
dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
hk(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
γ(t)dt(cid:19)k
− d
hj(t)(cid:16)eXj(γ(t))(cid:17)k
d
hj(t)(cid:16)eXj(γ(t)) − ej(cid:17)k
d
hk(t)dt(cid:12)(cid:12)(cid:12)(cid:12) : x ∈ eB(0, r), k = 1, . . . , m(cid:27)(cid:19) = 0.
Hence, if for every x ∈ eB(0, r) we set d := ed(x, 0) and we denote by h a control associated with
hj(t) (ej)k dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)xk − d
hk(t)dt(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:18) d
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
mXj=1
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
mXj=1
r(cid:12)(cid:12)(cid:12)(cid:12)xk − d
Notice also that for every x ∈ eB(0, r)
r→0(cid:18)sup(cid:26) 1
the geodesic γ joining 0 and x, we have
f (Fp(x)) = f (Fp(x)) − f (Fp(0))
≤ mCrd ≤ mCr2.
dt −
(11)
lim
0
0
0
= f (Fp(γ(d))) − f (Fp(γ(0))) = d
Xjf (Fp(γ(t)))hj(t)dt.
Let ε > 0 be fixed. By (11) and the continuity of Xf we can choose r0 ∈ (0, R) such that
0
mXj=1
h(t)dt(cid:12)(cid:12)(cid:12)(cid:12) : x ∈ eB(0, r)(cid:27) <
∀ r ∈ (0, r0)
∀ x ∈ eB(0, r0).
ε
0
ε
2
2Xf (p)
Xf (Fp(x)) − Xf (p) <
r(cid:12)(cid:12)(cid:12)(cid:12)(x1, . . . , xm) − d
sup(cid:26)1
For any r ∈ (0, r0) and x ∈ eB(0, r) we have
f (Fp(x)) −eLXf (p)(x)
= (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xjf (p)xj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
mXj=1
d
h(t)Xf (Fp(γ(t))) − Xf (p)dt + Xf (p)(cid:12)(cid:12)(cid:12)(cid:12)(x1, . . . , xm) − d
≤ d
h(t)dt(cid:12)(cid:12)(cid:12)(cid:12) .
+ Xf (p)(cid:12)(cid:12)(cid:12)(cid:12)(x1, . . . , xm) − d
The result follows dividing both sides by r and taking into account that d ≤ r.
hh(t), Xf (Fp(γ(t)))i dt −
0
ε
2
< d
0
0
0
h(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
An immediate consequence of Proposition 2.13 is Corollary 2.14, where we start using the
following very convenient notation: given t ∈ R and a function f : I → R defined on some set
I, we denote by {f > t}, {f = t}, etc. the sets {x ∈ I : f (x) > t}, {x ∈ I : f (x) = t}, etc. This
notation will be extensively used in the paper.
Corollary 2.14. Let p ∈ Rn and f ∈ C 1
Xf 6= 0 in B(p, R) and consider the C 1
every ε > 0, there exists r0 > 0 such that, for every r ∈ (0, r0)
X(B(p, R)) for some R > 0; suppose that f (p) = 0,
X hypersurface S := {q ∈ B(p, R) : f (q) = 0}. Then, for
F −1
p (S) ∩ eB(0, r) ⊆ {x ∈ eB(0, r) : −εr ≤eLXf (p)(x) ≤ εr}.
(12)
(cid:3)
10
Moreover
lim
r→0
DON AND VITTONE
L n({x ∈ eB(0, r) : f (Fp(x))eLXf (p)(x) < 0})
rQ
= 0.
(13)
Proof. Fix ε > 0 and apply Proposition 2.13 to get r0 > 0 such that for every 0 < r < r0 and for
2εr}, we also get
every x ∈ eB(0, r) we have f (Fp(x)) −eLXf (p)(x) ≤ εr. Then, if we take x ∈ eB(0, r) ∩ {eLXf (p) ≥
Reasoning in the same way with the set {eLXf (p) ≤ −2εr} we readily get (12). The previous
argument shows that for any ε > 0 there exists r0 > 0 such that for any r ∈ (0, r0) we have
f (Fp(x)) ≥ εr.
The proof of (13) follows by noticing that, by Theorem 2.6
eB(0, r) ∩ {(f ◦ Fp)eLXf (p) ≤ 0} ⊆ eB(0, r) ∩ {−εr ≤eLXf (p) ≤ εr}.
for a suitable constant C independent of r.
L n(eB(0, r) ∩ {−εr ≤eLXf (p) ≤ εr}) ≤ CεrQ,
We point out for future references the following observation.
Remark 2.15. Let (Rn, X) be an equiregular CC space, p ∈ Rn, R > 0 and suppose that
f1, f2 ∈ C 1
X(B(p, R)) are such that f1(p) = f2(p) = 0 and Xf1(p) = Xf2(p); then one has
Indeed, taking into account (12) we observe that
lim
r→0
1
rQ
L n(B(p, r) ∩ {f1f2 ≤ 0}) = 0.
lim
r→0
1
rQ
L n({ξ ∈ B(p, r) : f1(ξ)f2(ξ) = 0}) = 0.
On the other hand, since eLXf1(p) =eLXf2(p) the set B(p, r) ∩ {f1f2 < 0} is contained in
p ≤ 0}(cid:17),
B(p, r) ∩(cid:16){f1f2 < 0 and eLXf1(p) ◦ F −1
p > 0} ∪ {f1f2 < 0 and eLXf1(p) ◦ F −1
that combined with (13) completes the proof.
We can now introduce the notion of intrinsic rectifiability in equiregular CC spaces. We
denote by H k and S k, respectively, the k-dimensional Hausdorff and spherical Hausdorff
measures in (Rn, d), see e.g. Definition A.3.
Definition 2.16 (X-rectifiability). Let (Rn, X) be an equiregular CC space of homogeneous
dimension Q ∈ N and let R ⊆ Rn. We say that R is countably X-rectifiable (respectively,
countably X-Lipschitz rectifiable) if there exists a family {Sh : h ∈ N} of C 1
X hypersurfaces
(resp., X-Lipschitz hypersurfaces) such that
(cid:3)
(14)
H Q−1 R \
Sh! = 0.
∞[h=0
Moreover we say that R is X-rectifiable (resp., X-Lipschitz rectifiable) if R is countably X-
rectifiable (resp., countably X-Lipschitz rectifiable) and H Q−1(R) < +∞.
Definition 2.17 (Horizontal normal). Let R ⊆ Rn be countably X-rectifiable and let (Sh) be
C 1
X hypersurfaces such that (14) holds. Then the horizontal normal νR : R → Sm−1 to R is
defined by
νR(p) := νSh(p)
if p ∈ R ∩ Sh \[k<h
Sk .
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
11
The horizontal normal νR is well-defined, up to a sign, H Q−1-almost everywhere on R: this
is a standard consequence of the following result.
Proposition 2.18. Let (Rn, X) be an equiregular CC space and let S1, S2 ⊆ Rn be two hyper-
surfaces of class C 1
X. Then the set
E := {p ∈ S1 ∩ S2 : νS1(p) /∈ {±νS2(p)}}
is H Q−1-negligible.
Proof. By a localization argument we can suppose without loss of generality that S1 is bounded
in Rn and that H Q−1(E) ≤ H Q−1(S1) < +∞. For every δ > 0 define
Eδ := {p ∈ E : hνS1(p), νS2(p)i ≤ 1 − δ}.
Then we have E =S{Eδ : δ ∈ (0, +∞) ∩ Q}.
the following three properties hold for every r ≤ 2R
Fix ε ∈ (0, 1/4) and define for every R > 0 the set Eδ,R of all the points p of Eδ such that
(a) if C > 0 is the constant appearing in Theorem 2.6, for every x ∈ A(Cr) we have
(b) for i = 1, 2 we have F −1
bBp(x, εr) ⊆ eBp(x, 2εr);
(c) diam B(p, r) = diameBp(0, r) ≥ r.
By Theorems 2.9 and 2.14 and the fact1 that diambBp(0, r) = 2r we deduce that Eδ,R ր Eδ as
p (Si ∩ B(p, 2r)) ⊆ {eLνSi (p) < εr};
2 ). Then there exist a sequence (qh) in Rn and a sequence (rh) in (0, η) such
Fix now η ∈ (0, R
R → 0.
that
We can suppose without loss of generality that for every h ∈ N there exists ph ∈ B(qh, rh)∩Eδ,R.
Therefore for every h ∈ N one has B(qh, rh) ⊆ B(ph, 2rh) and consequently
Taking into account Theorem 2.6, we can find C > 0 such that for every h ∈ N one has
Eδ,R ⊆
B(ph, 2rh).
∞[h=0
F −1
ph (Eδ,R ∩ B(ph, 2rh)) ⊆ Ah
where
We prove now that L n(Ah) ≤ Cδε2rQ
hνS1(ph), νS2(ph)i ≤ 1 − δ, we have (up to an orthogonal change of coordinates)
Ah :=nx ∈ Rn : kxk ≤ Crh and eLνSi (ph)(x) ≤ εrh, for i = 1, 2o .
nx ∈ Rn : eLνSi (ph)(x) < εrh for i = 1, 2o ⊆ Q2(0, Cδεrh) × Rn−2,
h for some Cδ > 0 depending on δ.
is well-known to experts, even though the only reference we are aware of is [45, Proposizione 7.4].
1This is an easy consequence of the fact that the curve t 7→ exp(tbX1) is globally length minimizing. This fact
In fact, since
Eδ,R ⊆
B(qh, rh)
and
∞[h=0
(rh)Q−1 ≤
(diam B(qh, rh))Q−1 ≤ S Q−1
η
(Eδ,R) + 1.
∞Xh=0
∞Xh=0
12
DON AND VITTONE
where the notation Q2(z, s) denotes a 2-dimensional cube of center z and size s. Hence
Ah ⊆ Q2(0, Cδεrh) ∩(x ∈ Rm :
xj ≤ Crh)! ×(x ∈ Rn−m :
mXj=1
1
dj ≤ Crh)
nXj=m+1
xj
and consequently L n(Ah) ≤ Cδε2rQ
h .
For every h ∈ N, combining Theorem A.2 and the fact that Ah is compact, we can find Nh ∈ N
and a family {xh,j : j = 1, . . . , Nh} of points of Ah such that {bBph(xh,j, εrh) : j = 1, . . . , Nh}
covers Ah and {bBph(xh,j, ε rh
5 ) : j = 1, . . . , Nh} is pairwise disjoint. Reasoning as above, it is
easy to see that
Therefore we can estimate
h .
εrh
L n(cid:16)nx ∈ Rn : bdph(x, Ah) <
5 o(cid:17) ≤ eCδε2rQ
5 o(cid:17)
L n(cid:16)nx ∈ Rn : bdph(x, Ah) < εrh
≤ bCδε2−Q
5 )(cid:17)
L n(cid:16)bBph(xh,j, εrh
Nh ≤
for some bCδ > 0 that, by Remark 2.10, depends only on δ. By property (a) we have also
bBph(xh,j, εrh) ⊆ eBph(xh,j, 2εrh), hence the family {eBph(xh,j, 2εrh) : j = 1, . . . , Nh} is a covering
of Ah, that is also a covering of F −1
ph (Eδ,R ∩ B(ph, rh)). Hence the family {B(F −1
j ∈ N} is a covering of Eδ,R ∩ B(ph, 2rh) In particular, since ε ∈ (0, 1/4) we have
ph (xh,j), 2εrh) :
S Q−1
η
(Eδ,R) ≤ S Q−1
4εη (Eδ,R ∩ B(ph, 2rh))
≤
≤
S Q−1
4εη (Eδ,R) ≤
∞Xh=0
NhXj=1(cid:0)diam B(F −1
∞Xh=0
ph (xh,j), 2εrh)(cid:1)Q−1 ≤
∞Xh=0 bCδεrQ−1
S Q−1(Eδ) ≤ bCδε(S Q−1(Eδ) + 1).
h ≤ bCδε(S Q−1
(Eδ,R) + 1).
η
Nh(4εrh)Q−1
∞Xh=0
Letting η → 0 we get S Q−1(Eδ,R) ≤ bCδε(S Q−1(Eδ,R) + 1), which gives, letting R → 0
Letting now ε → 0 we get, for any δ > 0, that S Q−1(Eδ) = 0 , i.e., S Q−1(E) = 0. This
concludes the proof.
(cid:3)
2.3. Approximate notions of continuity, X-jumps and X-differentiability. In this sec-
tion we introduce the notions of approximate continuity, approximate X-jumps and approxi-
mate X-differentiability; we keep on working in a fixed equiregular CC space (Rn, X). We use
the notation
and, in what follows, we denote by Ω an open subset of Rn.
A
u dL n :=
1
L n(A) A
u dL n
Definition 2.19 (Approximate Limit). Let u ∈ L1
z ∈ Rk is the approximate limit of u at p if
loc(Ω; Rk), z ∈ Rk and p ∈ Ω. We say that
lim
r→0 B(p,r)
u − zdL n = 0.
We denote by u⋆(p) the approximate limit of u at p and by Su the set of points in Ω where u
does not admit an approximate limit.
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
13
If the approximate limit of u at a point p exists, it is also unique. By the generalized
Lebesgue's differentiation theorem (see e.g. [28, Section 2.7]), we have L n(Su) = 0 and u⋆ = u
a.e. on Ω. Moreover it can be easily proved (adapting e.g. [2, Proposition 3.64]) that Su is a
Borel set and that u⋆ : Ω \ Su → Rk is a Borel map.
Remark 2.20. Let Ω, u, z and p be as in Definition 2.19. Then u has approximate limit z at p
if and only if, working in adapted exponential coordinates Fp around p, as r → 0 the functions
u ◦ Fp ◦ δr converge in L1
loc(Rn; Rk) to the constant function z. This is an easy exercise left to the
reader; alternatively, it is enough to follow the proof of Proposition 2.26 below with a = b = z.
Definition 2.21 (Essential boundary). Given a measurable set E ⊆ Rn and t ∈ [0, 1] we denote
by Et the set of points with density t for E, i.e., the set of all p ∈ Rn satisfying
L n(E ∩ B(p, r))
L n(B(p, r))
lim
r→0
= t.
The essential boundary of E is ∂∗E := Rn \ (E0 ∪ E1).
The following proposition is standard; for the reader's convenience we prove it later in Propo-
sition A.1.
Proposition 2.22. Let u ∈ L1
loc(Ω), p ∈ Ω \ Su and t 6= u⋆(p); then p /∈ ∂∗{u > t}.
We now introduce the notion of X-jump points; this requires a certain amount of work,
one of the reasons being that there is no canonical way of separating a CC ball B(p, r) into
complementary "half-balls" B+
ν (p, r). We will use as separating sets an arbitrary hy-
persurface S of class C 1
X such that νS(p) = ν, and one of the issues (Remark 2.27 below) is
proving well-posedness of our definition independently of the choice of S.
ν (p, r), B−
For any fixed p ∈ Rn, ν ∈ Sm−1 and r > 0 we introduce the notation B+
ν (p, r)
X (B(p, R)) such that2 f (p) = 0 and Xf (p)/Xf (p) = ν, we
ν (p, r) and B−
as follows. Given R > 0 and f ∈ C 1
set for r ∈ (0, R)
B+
ν (p, r) := B(p, r) ∩ {f > 0}
and
B−
ν (p, r) := B(p, r) ∩ {f < 0}.
(15)
These objects are well-defined only if r is small enough. Moreover, there is a clear abuse of
notation, since B±
ν (p, r) depend on the choice of f . However, this will not effect the validity of
our results.
Before introducing the notion of approximate X-jumps we state some properties of the "half-
balls" B±
ν (p, r). Proposition 2.23 is used in the proof of Theorem 3.14.
Proposition 2.23. Let (Rn, X) be an equiregular CC space and let Ω ⊆ Rn be an open set.
Then, for any p ∈ Ω and ν ∈ Sm−1.
L n (B+
ν (p, r))
L n (B(p, r))
lim
r→0
= lim
r→0
L n (B−
ν (p, r))
L n (B(p, r))
=
1
2
Proof. Let U be a neighborhood of p and let f ∈ C 1
X (U) be such that f (p) = 0 and Xf (p) = ν.
Choose ε ∈ (0, 1). By Proposition 2.13 and Theorem 2.9 we can suppose without loss of
F −1
p (B+
Analogously
generality that for every small enough r one has Fp(eB(0, r)) = B(p, r) and
bB(0, (1 − ε)r) ∩ {eLν ≥ εr} ⊆ eB(0, r) ∩ {f ◦ Fp > 0} = F −1
ν (p, r)) = eB(0, r) ∩ {f ◦ Fp > 0} ⊆ bB(0, (1 + ε)r) ∩ {eLν ≥ −εr}.
2One can consider for instance f =eLν ◦ Fp.
p (B+
ν (p, r)).
(16)
(17)
14
DON AND VITTONE
Applying δ1/r to both sides of (16) and evaluating the Lebesgue measure we get
p (B+
L n(cid:0)F −1
rQ
ν (p, r))(cid:1)
Taking the lim sup as r → 0 and letting ε → 0 we infer
p (B+
L n(cid:0)F −1
rQ
ν (p, r))(cid:1)
lim sup
r→0
where the last equality follows from Proposition 2.8. With the same argument, from (17) we
get
L n(bB(0, 1)),
(18)
(19)
p (B+
ν (p, r))(cid:1)(cid:1)
= L n(cid:0)δ1/r(cid:0)F −1
≤ L n(cid:16)bB(0, 1 + ε) ∩ {eLν ≥ −ε}(cid:17) .
≤ L n(cid:16)bB(0, 1) ∩ {eLν ≥ 0}(cid:17) =
ν (p, r))(cid:1)
L n(bB(0, 1)),
ν (p, r))(cid:1)
L n(bB(0, 1)).
1
2
1
2
1
2
rQ
rQ
≥
=
p (B+
p (B+
= lim
r→0
L n(δ1/r(eB(0, r))) = L n(bB(0, 1)),
L n(cid:0)F −1
ν (p, r))(cid:1)
L n(eB(0, r))
(B+
1
2
=
.
p
lim
r→0
hence
By Theorem 2.9
lim
r→0
r→0
lim inf
L n(cid:0)F −1
L n(cid:0)F −1
L n(eB(0, r))
lim
r→0
rQ
and combining (18) and (19) we get
If c := det ∇F (0) > 0, using (5) we notice that for every 0 < ε < c and every sufficiently
small r > 0 we have
p
(B+
(c − ε) L n(cid:0)F −1
ν (p, r))(cid:1)
(c + ε) L n(eB(0, r))
≤
L n (B+
ν (p, r))
L n(B(p, r))
≤
p
(B+
(c + ε)L n(cid:0)F −1
ν (p, r))(cid:1)
(c − ε)L n(eB(0, r))
.
The result follows passing to the limit as r → 0, letting ε → 0 and, eventually, using a similar
argument for B−
ν .
(cid:3)
We can now introduce the notion of X-jump points.
Definition 2.24 (Approximate X-jumps). Let u ∈ L1
an approximate X-jump at p if there exist a, b ∈ Rk with a 6= b and ν ∈ Sm−1 such that
loc(Ω; Rk) and p ∈ Ω. We say that u has
lim
r→0 B+
ν (p,r)
u − adL n = lim
r→0 B−
ν (p,r)
u − bdL n = 0.
(20)
In this case we say that (a, b, ν) is an approximate X-jump triple of u at p. We shall denote
by Ju the set of approximate X-jump points of u and by (u+(p), u−(p), νu(p)) the approximate
X-jump triple for u at p ∈ Ju.
Remark 2.25. Using e.g. Proposition 2.23 one easily proves that Ju ⊆ Su.
Notice that, if u has an approximate X-jump at p associated with (a, b, ν), then it is also
associated with the triple (b, a, −ν). For this reason, it will be sometimes convenient to consider
the space of triples endowed with the equivalence relation (a, b, ν) ≡ (a′, b′, ν′) if and only if
(a, b, ν) = (a′, b′, ν′) or (a, b, ν) = (b′, a′, −ν′). The following Proposition 2.26 shows that the
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
15
X-jump triple (u+(p), u−(p), νu(p)) is unique up to equivalence, for the map Rk × Rk × Sm−1 ∋
(a, b, ν) → wa,b,ν ∈ L1
loc(Rn; Rk) defined by (21) below satisfies
wa,b,ν = wa′,b′,ν ′ ⇐⇒ (a, b, ν) ≡ (a′, b′, ν′).
In the theory of classical BV functions a jump point can be detected, via a blow-up procedure,
in terms of L1
loc-convergence to a function taking two different values on complementary half-
spaces; this is the content of the next statement, which also gives an equivalent definition of
approximate X-jump points.
Proposition 2.26. Let (Rn, X) be an equiregular CC space, Ω an open set, u ∈ L1
loc(Ω; Rk),
p ∈ Ω and let a, b ∈ Rk with a 6= b and ν ∈ Sm−1 be fixed. Then the following statements are
equivalent:
(i) p ∈ Ju and (u+(p), u−(p), νu(p)) ≡ (a, b, ν);
(ii) working in adapted exponential coordinates Fp around p, as r → 0 the functions eur :=
u ◦ Fp ◦ δr converge in L1
loc(Rn; Rk) to
(21)
wa,b,ν(y) :=(a
b
if eLν(y) > 0
if eLν(y) < 0.
Proof. We can assume without loss of generality that k = 1.
We prove the implication (i)⇒(ii); we can assume that (u+(p), u−(p), νu(p)) = (a, b, ν) and,
writing w := wa,b,ν, we prove that for any fixed R > 0 one has
lim
r→0 bB(0,R)
u ◦ Fp ◦ δr − w dL n = 0.
By a change of variables, this is equivalent to proving that
lim
r→0
1
rQ bB(0,r)
u ◦ Fp − w dL n = 0.
(22)
Let f be the real function of class C 1
the half-balls B±
ν (p, r) appearing in (20); we set for brevity
X defined on a neighborhood of p used to define, as in (15),
By Theorem 2.9 there exists an increasing function ω : (0, +∞) → (0, +∞) such that
bB+
ν (0, r) := bB(0, r) ∩ {eLν > 0},
eB+
ν (0, r) := eB(0, r) ∩ {f ◦ Fp > 0},
lim
r→0
ω(r)
r
= 0
and
for any sufficiently small r. Therefore
bB−
ν (0, r) := bB(0, r) ∩ {eLν < 0}
eB−
ν (0, r) := eB(0, r) ∩ {f ◦ Fp < 0}.
bB(0, r) ⊆ eB(0, r + ω(r))
=
≤
1
1
rQ bB(0,r)
rQ(cid:18) bB+
rQ(cid:18) eB+
+ eB−
1
ν (0,r)
u ◦ Fp − w dL n
u ◦ Fp − a dL n + bB−
ν (0,r)
u ◦ Fp − b dL n(cid:19)
ν (0,r+ω(r))
ν (0,r+ω(r))
u ◦ Fp − a dL n + bB+
u ◦ Fp − b dL n + bB−
ν (0,r)\ eB+
ν (0,r)\ eB−
ν (0,r+ω(r))(cid:0)u ◦ Fp − b + a − b(cid:1) dL n
ν (0,r+ω(r))(cid:0)u ◦ Fp − a + a − b(cid:1) dL n(cid:19)
16
DON AND VITTONE
1
≤
ν (0, r) \ eB±
ν (0, r + ω(r)) ⊆ eB(0, r + ω(r)) \ eB±
and using bB±
rQ(cid:18)2 eB+
u ◦ Fp − a dL n + 2 eB−
+ a − bL n(cid:0)eB(0, r + ω(r)) ∩ {(f ◦ Fp)eLν ≤ 0}(cid:1)(cid:19)
ν (0,r+ω(r))
ν (0,r+ω(r))
ν (0, r + ω(r)) ⊆ eB∓
u ◦ Fp − b dL n
ν (0, r + ω(r))
and (22) follows from (20) and Corollary 2.14 taking also Theorem 2.2 into account.
For the converse implication one has to prove that, if (ii) holds and f is a C 1
X real function
on a neighborhood of p such that f (p) = 0 and Xf (p)/Xf (p) = ν, then (20) holds with
B±
ν (p, r) defined (see (15)) in terms of f . By Theorem 2.2 and a change of variables, proving
(20) amounts to proving that
lim
r→0
1
rQ eB+
ν (0,r)
u ◦ Fp − a dL n = lim
r→0
1
rQ eB−
ν (0,r)
u ◦ Fp − b dL n = 0
and this can be done by a boring adaptation, that we omit, of the previous argument.
Remark 2.27. The proof of Proposition 2.26 implicitly shows that the validity of (20) does
not depend on the choice of the function f used in (15) to define B±
(cid:3)
ν (p, r).
The proof of the following result is standard and we postpone it to the Appendix B.
Proposition 2.28. Let (Rn, X) be an equiregular CC space, Ω be an open set and let u ∈
L1
loc(Ω; Rk). Then the following facts hold:
(i) Ju is a Borel set and, up to a choice of a representative for X-jump triples, the function
Ju → Rk × Rk × Sm−1
p 7→ (u+(p), u−(p), νu(p))
is Borel;
(ii) for every f ∈ Lip(Rk; Rh) and p ∈ Ju we have
p ∈ Jf ◦u ⇐⇒ f (u+(p)) 6= f (u−(p))
and in this case ((f ◦ u)+(p), (f ◦ u)−(p), νf ◦u(p)) ≡ (f (u+(p)), f (u−(p)), νu(p)). Other-
wise, p /∈ Sf ◦u and (f ◦ u)⋆(p) = f (u+(p)) = f (u−(p)).
We now pass to he introduction of approximate X-differentiability.
Definition 2.29 (Approximate X-differentiability). Let u ∈ L1
loc(Ω; Rk) and p ∈ Ω \ Su. We
say that u is approximately X-differentiable at p if there exist a neighborhood U of p and
f ∈ C 1
X(U; Rk) such that f (p) = 0 and
r→0 B(p,r)
lim
u − u⋆(p) − f
r
dL n = 0.
(23)
The subset of points of Ω in which u is approximately X-differentiable is denoted by Du.
If f is as in Definition 2.29 we will call Xf (p) ∈ Rk×m the approximate X-gradient of u at
p. By the following proposition the approximate X-gradient of u at p is uniquely determined,
and we denote it by Dap
Proposition 2.30 (Uniqueness of approximate X-gradient). Let (Rn, X) be an equiregular
CC space, Ω ⊆ Rn an open set, u ∈ L1
loc(Ω; Rk) and p ∈ Ω \ Su. Let R > 0 and f1, f2 ∈
X (B(p, R); Rk); suppose that formula (23) holds for both f = f1 and f = f2. Then p ∈ Du,
C 1
f1(p) = f2(p) = 0 and Xf1(p) = Xf2(p).
X u(p).
Conversely, if f1(p) = f2(p) = 0 and Xf1(p) = Xf2(p), then formula (23) holds for f = f1
if and only if it holds for f = f2.
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
17
Proof. It is not restrictive to assume that k = 1. Define for i = 1, 2 the functions Li :=eLXfi(p).
Suppose first that both f1, f2 satisfy (23). Fix ε > 0 and by Proposition 2.13 choose r > 0 such
that for every ∈ (0, r)
fi ◦ Fp − Li
<
ε
2
Then for such values of we have
on eB(0, ).
eB(0,)
L1 − L2
f1 ◦ Fp − f2 ◦ Fp
dL n + ε
f1 − f2
dL n + ε
dL n ≤ eB(0,)
≤ C B(p,)
≤ C B(p,)
u − u⋆(p) − f1 + u − u⋆(p) − f2
dL n + ε.
It follows that
lim
→0 eB(0,)
L1 − L2
dL n = 0.
If Xf1(p) 6= Xf2(p), by Theorem 2.6 one would get, for some C1 > 0
eB(0,)
L1 − L2 dL n =
≥
1
L n(eB(0, )) eB(0,)
L n (A(C1)) A(/C1)
1
L1 − L2 dL n
L1 − L2dL n = C
Q+1
Q = C,
a contradiction. This proves the first part of the statement.
Suppose now that Xf1(p) = Xf2(p) and that f1 satisfies (23). Then we have L1 = L2 and
B(p,)
≤ B(p,)
u − u⋆(p) − f2
dL n
f1 − L1 ◦ F −1
p
+ u(y) − u⋆(p) − f1 + f2 − L2 ◦ F −1
p
dL n.
By Proposition 2.13 this completes the proof.
(cid:3)
As for X-jump points, also approximate X-differentiability points can be detected by a blow-
up procedure.
Proposition 2.31. Let (Rn, X) be an equiregular CC space, Ω be an open subset of Rn, u ∈
L1
loc(Ω; Rk) and let p ∈ Ω \ Su. Then u is approximate X-differentiable at p if and only if there
exists z = (z1, . . . , zk) ∈ Rk×m such that
u ◦ Fp ◦ δr − u⋆(p)
r
In this case we have Dap
X u(p) = z.
→ (eLz1, . . . ,eLzk)
in L1
loc(Rn; Rk) as r → 0.
18
DON AND VITTONE
Proof. We assume without loss of generality that k = 1. Assume first that p ∈ Du and let f be
as in (23); set z := Dap
X u(p) ∈ Rm. Given R > 0, by Theorem 2.9 one has for small enough r
r
r
r
1
C
u ◦ Fp ◦ δr − u⋆(p)
−eLz(cid:12)(cid:12)(cid:12)(cid:12) dL n
bB(0,R)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
rQ bB(0,rR)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
u ◦ Fp − u⋆(p) −eLz
rQ B(p,2rR)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
u − u⋆(p) −eLz ◦ F −1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
B(p,r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
u − u⋆(p) −eLz ◦ F −1
rQ bB(0,2r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
u ◦ Fp − u⋆(p) −eLz
dL n ≤
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
C
r
r
p
=
≤
≤
which allows to conclude.
(we used Proposition 2.30), which proves the first part of the statement.
Conversely, for any small enough r > 0 we have
dL n ≤
1
rQ eB(0,2rR)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
dL n → 0 as r → 0,
u ◦ Fp − u⋆(p) −eLz
r
dL n
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
C
rQ eB(0,r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
dL n = C bB(0,2)(cid:12)(cid:12)(cid:12)(cid:12)
u ◦ Fp − u⋆(p) −eLz
u ◦ Fp ◦ δr − u⋆(p)
r
r
dL n
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
−eLz(cid:12)(cid:12)(cid:12)(cid:12) dL n,
(cid:3)
The proofs of the following two results are postponed to Appendix B.
Proposition 2.32 (Properties of approximate differentiability points). Let (Rn, X) be an
equiregular CC space, Ω be an open set in Rn and let u ∈ L1
loc(Ω; Rk). Then Du is a Borel
set and the map Dap
Proposition 2.33 (Locality). Let (Rn, X) be an equiregular CC space, Ω an open set in Rn
and u, v ∈ L1
loc(Ω; Rk). Suppose that p ∈ Ω is of density 1 for the set {q ∈ Ω : u(q) = v(q)}.
Then the following facts hold.
X u : Du → Rm×k is a Borel map.
(i) If p ∈ Ω \ (Su ∪ Sv), then u⋆(p) = v⋆(p).
(ii) If p ∈ Ju ∩ Jv, then (u+(p), u−(p), νu(p)) ≡ (v+(p), v−(p), νv(p)).
(iii) If p ∈ Du ∩ Dv then Dap
X u(p) = Dap
X v(p).
2.4. Functions with bounded X-variation. In this section we review the definition and
basic properties of BVX functions. We keep on working in a fixed equiregular CC space (Rn, X),
while Ω denotes a fixed open subset of Rn.
Definition 2.34 (Functions with bounded X-variation). We say that u ∈ L1
loc(Ω) is a function
of locally bounded X-variation in Ω, and we write u ∈ BVX,loc(Ω), if there exists a Rm-valued
Radon measure DXu = (DX1u, . . . , DXmu) in Ω such that for every open set A ⋐ Ω and for
every ϕ ∈ C 1
c (A) we have
∀ i = 1, . . . , m
A
ϕ d(DXiu) = −A
uX ∗
i ϕ dL n,
(24)
where X ∗
i denotes the formal adjoint of Xi. If u ∈ L1(Ω), we say that u has bounded X-variation
in Ω, and we write u ∈ BVX(Ω), if, moreover, the total variation DXu of DXu is finite on Ω.
As customary, we write BVX(Ω; Rk) := (BVX(Ω))k, and similarly for BVX,loc(Ω; Rk). It can
be useful to observe that if u ∈ BVX (Ω; Rk), the following inequalities hold
max
1≤i≤k
DXui(Ω) ≤ DXu(Ω) ≤
kXi=1
DXui(Ω).
(25)
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
19
The following approximation result is proved in [20, 24].
Theorem 2.35. Let u ∈ BVX(Ω; Rk). Then there exists a sequence (uh) in C ∞(Ω; Rk) such
that
lim
h
kuh − ukL1(Ω;Rk) = 0
and
lim
h
DXuh(Ω) = DXu(Ω).
We now state and prove a simple but useful result.
Let also X1, . . . , Xm be vector fields on Ω and define for every i = 1, . . . , m the vector fields
Proposition 2.36. Let Ω,eΩ be two open sets in Rn and let G : Ω → eΩ be a diffeomorphism.
Yi := dG(Xi) on eΩ. Then
(26)
More precisely, for every open set U ⋐ Ω and setting V := G(U), one has for every u ∈
BVX,loc(Ω) that
u ∈ BVX,loc(Ω) ⇐⇒ v := u ◦ G−1 ∈ BVY,loc(eΩ).
mDXu(U) ≤ DY v(V ) ≤ MDXu(U)
(27)
for m := inf U det ∇G and M := supU det ∇G.
Proof. We claim that, for any open set U ⋐ Ω and any u ∈ BVX,loc(Ω), one has
v := u ◦ G−1 ∈ BVY (V )
and
DY v(V ) ≤ MDX u(U).
This would be enough to conclude: indeed, the claim would imply both the ⇒ implication in
(26) and the second inequality in (27), while the ⇐ implication in (26) and the first inequality
in (27) simply follow by replacing X, U, u, G with (respectively) Y, V, v, G−1 and noticing that
m = (supV det ∇(G−1))−1.
Let us prove the claim. First we assume that u ∈ C ∞(U), so that also v is smooth on V . For
every ϕ ∈ C 1
c (V ; Rm) with ϕ ≤ 1, by a change of variable we have that
V
hY v, ϕidL n = U
hXu, (ϕ ◦ G)i det ∇GdL n,
which gives
In case u ∈ BVX(U) is not smooth, consider a sequence (uh) in C ∞(U) that converges to u in
L1(U) and such that
DY v(V ) ≤ MDX u(U).
lim
h
DXuh(U) = DXu(U).
Defining vh := uh ◦ G−1, we easily get that vh converges to v in L1(V ) as h → +∞. Therefore
DY v(V ) ≤ lim inf
h
DY vh(V ) ≤ M lim inf
h
DXuh(U) = MDX u(U)
and the proof is accomplished.
(cid:3)
Definition 2.37 (Sets with finite X-perimeter). A measurable set E ⊆ Rn has locally finite
X-perimeter (resp., finite X-perimeter) in Ω if χE ∈ BVX,loc(Ω) (resp., χE ∈ BVX (Ω)). In such
a case we define the X-perimeter measure P E
X of E by P E
X := DXχE.
It will sometimes be useful to write PX(E, ·) instead of P E
X .
Definition 2.38 (Measure theoretic horizontal normal). If E is a set with locally finite X-
perimeter, then by Riesz representation theorem there exists a P E
X -measurable function νE :
Rn → Sm−1 such that
We call νE the measure theoretic horizontal normal to E.
DXχE = νEP E
X .
The following result is proved in [1] and it will be of capital importance in the following.
20
DON AND VITTONE
Theorem 2.39. Let (Rn, X) be an equiregular CC space of homogeneous dimension Q; let
E ⊆ Rn be a set with finite X-perimeter in an open set Ω ⊆ Rn. Then
X Ω = θH Q−1 (Ω ∩ ∂∗E)
P E
(28)
for a suitable positive function θ that is locally bounded away from 0. Moreover
lim sup
r→0
P E
X (B(p, 2r))
P E
X (B(p, r))
< ∞
for P E
X -a.e. p ∈ Ω ∩ ∂∗E.
The proofs of the following well-known result can be found, for instance, in [20].
Theorem 2.40 (Coarea Formula for BVX functions). Let (Rn, X) be a CC space, let Ω be an
open set in Rn and let u ∈ BVX (Ω). Then, if we define Es := {p ∈ Ω : u(p) > s}, we have
DX u(Ω) = +∞
−∞
PX(Es; Ω)ds.
The next result is essentially [10, Theorem 1.2]; note, however, that the dimension Q appear-
ing in [10, Theorem 1.2] is slightly different from the homogeneous dimension we are considering.
See also [29].
Theorem 2.41. Let Ω be an open subset of an equiregular CC space (Rn, X) of homogeneous
dimension Q and let K ⊆ Ω be compact. Then there exist C > 0 and R > 0 such that, for
every p ∈ K, r ∈ (0, R) and u ∈ BVX,loc(Ω; Rk), the inequality
(cid:18) B(p,r)
u − up,r
Q
Q−1 dL n(cid:19) Q−1
Q
≤
C
rQ−1 DXu(B(p, r)),
where up,r := fflB(p,r) u dL n, holds.
Proof. It is clearly enough to consider the case k = 1. The proof then easily follows by [27,
Theorem 5.1] on taking into account Theorem 2.2, [10, Theorem 1.1], [27, Corollary 9.8 and
Theorem 10.3] and Theorem 2.35.
(cid:3)
An easy consequence of Theorem 2.41 is the following isoperimetric inequality.
Theorem 2.42 (Isoperimetric inequality in CC spaces). Let (Rn, X) be an equiregular CC
space and let K ⊆ Rn be a compact set. Then there exist C > 0 and R > 0 such that, for every
p ∈ K, r ∈ (0, R) and every L n-measurable set E ⊆ Rn, one has
min {L n(E ∩ B(p, r)), L n(B(p, r) \ E)}
Q−1
Q ≤ CPX(E, B(p, r)).
We conclude this section with some auxiliary results. The first one is proved in [16].
Theorem 2.43. Let X = (X1, . . . , Xm) and X j = (X j
m), j ∈ N, be m-tuples of linearly
independent smooth vector fields on Rn such that X satisfies the Hörmander condition and its
CC balls are bounded in Rn; assume that, for every i = 1, . . . , m, X j
loc(Rn) as
j → ∞. Let uj ∈ BVX j ,loc(Rn) be a sequence of functions that is locally uniformly bounded in
BVX j , i.e., such that for any compact set K ⊆ Rn there exists M > 0 such that
i → Xi in C ∞
1, . . . , X j
∀j ∈ N
kujkL1(K) + DX j uj(K) ≤ M < ∞.
Then, there exist u ∈ BVX,loc(Rn) and a subsequence (ujh) of (uj) such that ujh → u in L1
as h → ∞. Moreover, for any bounded open set Ω ⊆ Rn one has
loc(Rn)
DXu(Ω) ≤ lim inf
j→∞
DX j uj(Ω).
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
21
The proof of Theorem 2.43 given in [16] implicitly contains also the following result's proof,
that we however provide for the sake of completeness.
Proposition 2.44. Let X = (X1, . . . , Xm) and X j = (X j
linearly independent smooth vector fields on Rn such that, for every i = 1, . . . , m, X j
C ∞
for any open bounded set Ω ⊆ Rn one has
loc(Rn) be a sequence converging in L1
loc(Rn) as j → ∞. Let (uj) ⊆ L1
m), j ∈ N, be m-tuples of
i → Xi in
loc(Rn) to some u; then,
1, . . . , X j
Proof. For any i = 1, . . . , m and any j ∈ N we write
DXu(Ω) ≤ lim inf
j→∞
DX j uj(Ω)
Xi(x) =
ai,k(x)∂k
and
X j
aj
i,k(x)∂k
i (x) =
nXk=1
mXi=1
j→∞Ω
uj
i,k. Then, for any test function ϕ ∈ C 1
c (Ω; Rm) we have
∂k(ai,kϕi) dL n = lim
∂k(aj
i,kϕi) dL n
nXk=1
(29)
for suitable smooth functions ai,k, aj
Ω
i ϕi dL n =Ω
X ∗
u
mXi=1
nXk=1
mXi=1
j→∞Ω
u
nXk=1
mXi=1
= lim
uj
∗
X j
i
ϕi dL n ≤ kϕkL∞(Ω) lim inf
j→∞
DX j uj(Ω).
The proof is accomplished.
Remark 2.45. Let X, X j, uj, u be as in Proposition 2.44 and assume that DX j uj are locally
uniformly bounded in Rn, i.e., for any compact set K ⊆ Rn there exists CK < ∞ such that
DX j uj(K) < CK for all j. Then DX j uj weakly∗ converges to DX u in Rn.
(cid:3)
Indeed, one can reason as in (29) to show that for any test function ϕ ∈ C 1
c (Rn) and any
i = 1, . . . , m
and the density of C 1
c in C 0
lim
j→∞ ϕ dDX j
c allows to conclude.
i
uj = ϕ dDXiu
3. Fine properties of BV functions
This section is devoted to the proof of our main results.
Lemma 3.1. Let (Rn, X) be an equiregular CC space, let Ω ⊆ Rn be open and let (Eh) be a
sequence of measurable sets in Ω such that
L n(Eh) = 0
PX(Eh; Ω) = 0.
lim
h
and lim
h
Then for every α ∈ (0, 1) we have
Proof. Set
H Q−1 ∞\h=1(cid:26)p ∈ Ω : lim sup
h :=(cid:26)q ∈ Ω : lim sup
Eα
r→0
r→0
L n(Eh ∩ B(p, r))
L n(B(p, r))
L n(Eh ∩ B(q, r))
L n(B(q, r))
≥ α(cid:27)! = 0.
≥ α(cid:27) ,
and suppose without loss of generality that L n(Eh) > 0 for every h ∈ N. Let K ⋐ Ω. By
Theorem 2.2 there exist C > 1 and R > 0 such that for every q ∈ K, for every 0 < r < 2R we
have
1
C
rQ ≤ L n(B(q, r)) ≤ CrQ.
22
DON AND VITTONE
For any sufficiently large h ∈ N we have
Fix now p ∈ Eα
(cid:18)2CL n(Eh)
(cid:19) < RQ.
(cid:17)1/Q
h ∩ K and define δh =(cid:16) 4CL n(Eh)
; then
α
α
L n(Eh ∩ B(p, δh))
CL n(Eh)
L n(B(p, δh))
≤
δQ
h
=
α
4
.
On the other hand, by definition of Eα
h we can find arbitrarily small radii r > 0 such that
L n(Eh ∩ B(p, r))
L n(B(p, r))
≥
α
2
.
Taking into account Proposition A.9, a continuity argument allows us to find 0 < ≤ δh such
that
L n(Eh ∩ B(x, )) =
L n(B(x, )).
α
2
By the 5r-covering Lemma, we can find a family {B(pj, j) : j ∈ N} of pairwise disjoint balls
in Ω such that, for every j ∈ N,
pj ∈ Eα
h ∩ K
L n(Eh ∩ B(pj, j)) =
α
2
L n(B(pj, j))
Eα
h ∩ K ⊆
B(pj, 5j).
(30)
∞[j=0
Since L n(Eh) is finite, by Theorem 2.42 we get M > 0 such that
α
2C
Q
j ≤
α
2
L n(B(pj, j)) = L n(Eh ∩ B(pj, j)) ≤(cid:0)M PX (Eh; B(pj, j))(cid:1) Q
Q−1 .
Therefore we have that for every j ∈ N
Finally
H Q−1
10δh (cid:16)K ∩
Taking the limit for h → ∞ we get
PX(Eh; B(pj, j)).
Q
10δh
Eα
Q−1
(K ∩ Eα
h )
α (cid:19) Q−1
j ≤ M(cid:18) 2C
i(cid:17) ≤ H Q−1
∞\i=0
≤ ωQ−15Q−1M(cid:18)2C
≤ ωQ−15Q−1M(cid:18)2C
H Q−1 K ∩
∞\i=0
Eα
∞Xj=0
Q
Q
α (cid:19) Q−1
α (cid:19) Q−1
i! = 0.
(30)
≤ ωQ−15Q−1
Q−1
j
∞Xj=0
PX(Eh; B(pj, j))
PX(Eh; Ω).
By the arbitrariness of K, the proof is complete.
(cid:3)
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
23
Before passing to the next result, we introduce some notation that we are going to use fre-
quently in what follows. Let p ∈ Rn be fixed and let Fp denote adapted exponential coordinates
as in (4), for a fixed choice of a basis Y1, . . . , Yn as in (4). Given r > 0 and i ∈ {1, . . . , m},
define
If edr,eBr(x, ) denote, respectively, distance and balls with respect to the metric induced by the
vector fields (eX r
m), it is easy to see that the dilations δr satisfy
1, . . . , eX r
1
By Theorem 2.9, the convergence
i := r(dδr−1)[eXi ◦ δr].
eX r
red(δrξ, δrη).
edr(ξ, η) =
r→0eBr(0, ) = bB(0, )
lim
(31)
(32)
p. Moreover, given u ∈ BVX,loc(Rn; Rk) we set
holds in the Gromov-Hausdorff sense, bB(0, ) denoting a ball in the tangent Carnot group at
notice that D eX reur(eBr(0, )) = r1−QD eXeu(eB(0, r)).
Lemma 3.2. Let (Rn, X) be an equiregular CC space, let Ω ⊆ Rn be open and consider u ∈
BVX (Ω; Rk). Then
eur :=eu ◦ δr;
eu := u ◦ Fp
We can now prove the following lemma.
(33)
and
H Q−1(cid:18)(cid:26)p ∈ Ω : lim sup
r→0 B(p,r)
u
Q
Q−1 dL n = +∞(cid:27)(cid:19) = 0.
Proof. We can suppose without loss of generality that k = 1. Possibly considering u instead
of u, we can suppose that u ≥ 0; we also assume without loss of generality that Ω is bounded
in Rn. Define the set
By Proposition A.4 we have that H Q−1(D) = 0. For every h ∈ N we can find th ∈ (h, h + 1)
such that
D =(cid:26)p ∈ Ω : lim sup
r→0
DXu(B(p, r))
rQ−1
= +∞(cid:27) .
PX({u > th}, Ω) ≤ h+1
h
PX({u > t}, Ω)dt.
Define Eh = {u > th}. Since u ∈ L1(Ω) we have that limh L n(Eh) = 0 and applying the
Coarea Formula of Theorem 2.40 we get
PX (Eh, Ω) ≤ +∞
0
PX ({u > t}, Ω)dt = DXu(Ω) < +∞,
∞Xh=0
and therefore limh PX(Eh, Ω) = 0. We are in a position to apply Lemma 3.1. Defining for every
h ∈ N
Fh =(cid:26)p ∈ Ω : lim sup
r→0
L n(Eh ∩ B(p, r))
L n(B(p, r))
≥ α(cid:27) ,
where α > 0 will be chosen later depending on Ω only, we have that H Q−1 (T∞
is then sufficient to prove the inclusion
L :=(cid:26)p ∈ Ω : lim sup
r→0 B(p,r)
u
Q
Q−1 dL n = +∞(cid:27) ⊆ D ∪
Fh.
∞\h=0
h=0 Fh) = 0. It
24
DON AND VITTONE
Applying Theorem 2.41 we get C > 0 and R > 0 such that for every q ∈ Ω and all 0 < r < R
h=0 Fh and we prove that p /∈ L. Define up,r := fflB(p,r) udL n.
To this aim, we fix p /∈ D ∪T∞
B(q,r)
u(y) − uq,r
Q
Q−1 dL n(y) ≤ C(cid:18)DXu(B(q, r))
rQ−1
Q−1
(cid:19) Q
.
(34)
It is enough to prove that lim supr→0 up,r < +∞: in this case, in fact, inequality (34) and the
definition of D would imply that p /∈ L.
By contradiction we find an infinitesimal sequence (rj) such that limj up,rj = +∞. Define
DXu(B(p, rj)) is uniformly bounded with
Since p /∈ D, for any > 0 the sequence r1−Q
respect to j ∈ N; by Proposition 2.36, the same is true for the sequence
eu,eurj as in (33) (with r = rj) andevj :=eurj − up,rj; set also
eX j := (eX j
1, . . . , eX j
D eXeu(eB(0, rj)),
i := eX rj
eX j
D eX jevj(eBj(0, )) = r1−Q
and
m).
j
j
i
where eBj(0, ) := eBrj (0, ) according to the notation introduced after (31). Taking also (32) into
account, this proves that, for any compact set K ⊆ Rn, the sequence D eX jevj(K) is bounded;
by (34), also kevjkL1(K) is bounded.
By Theorem 2.43 (recalling also Theorem 2.7) there exists w ∈ L1(bB(0, 1)) such that, possibly
extracting a subsequence, evj → w in L1(bB(0, 1)). Consequently, for almost every x ∈ bB(0, 1)
u(Fp(δrj x)) = +∞
we have
lim
j
and then, for every h ∈ N,
L n(bB(0, 1)) = lim
= lim
j
j
= lim
j
r−Q
j
L n({x ∈ eBj(0, 1) : u(Fp(δrj x)) > th})
L n({x ∈ eB(0, rj) : u(Fp(x)) > th})
B(p,rj)∩Eh
det ∇F −1
dL n
1
rQ
j
p
≤ det ∇F −1
p (p) lim sup
r→0
L n(Eh ∩ B(p, r))
L n(B(p, r))
L n(B(p, r))
rQ
≤
C
det ∇Fp(0)
lim sup
r→0
L n(Eh ∩ B(p, r))
L n(B(p, r))
where C > 0 is given by Theorem 2.2 with K = Ω. Notice that L n(bB(0, 1)) depends on p.
Using (32) we obtain
lim sup
r→0
L n(Eh ∩ B(p, r))
L n(B(p, r))
det ∇Fp(0)
C
det ∇Fp(0)
C
L n(bB(0, 1)) =
lim
r→0
1
rQ
det ∇Fp(0)
C
lim
r→0
L n(eBr(0, 1))
rQ B(p,r)
lim
r→0
1
det ∇Fp(0)
C
L n(eB(0, r)) =
≥
1
C 2 .
rQ
h=0 Fh for α := 1/C 2, a contradiction.
≥
=
≥
L n(B(p, r))
1
C
lim inf
r→0
This proves that p ∈T∞
det ∇F −1
p
dL n
(cid:3)
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
25
The following proposition contains some of the first "fine" properties of BVX functions we
are interested in.
Proposition 3.3. Let (Rn, X) be an equiregular CC space. Then there exists λ : Rn → (0, +∞)
locally bounded away from 0 such that, for every open set Ω ⊆ Rn and every u ∈ BVX(Ω; Rk)
DXu ≥ λu+ − u−S Q−1 Ju
and for every Borel set B ⊆ Ω the following implications hold:
H Q−1(B) = 0 ⇒ DXu(B) = 0;
H Q−1(B) < +∞ and B ∩ Su = ∅ ⇒ DXu(B) = 0.
(35)
(36)
Proof. Let us prove the first part of the statement; we assume without loss of generality that
k = 1. Consider p ∈ Ju. By Proposition 2.26 the sequence eur := u ◦ Fp ◦ δr converges in
L1(bB(0, 1)) as r → 0 to the function
Defining eX r
i as in (31) and using Propositions 2.44 and 2.36 we obtain for any positive ε that
if eLν(p)(y) ≥ 0
if eLν(p)(y) < 0.
wp(y) :=(u+(p)
≥ det ∇Fp(0) lim inf
DX u(B(p, r))
u−(p)
lim inf
r→0
rQ−1
r
r
≥ det ∇Fp(0) lim inf
D eX reur(eBr(0, 1))
D eX reur(bB(0, 1 − ε))
≥ det ∇Fp(0) D bXwp(bB(0, 1 − ε)),
≥ det ∇Fp(0) D bXwp(bB(0, 1))
≥ det ∇Fp(0)u+(p) − u−(p)H n−1
e
(37)
(ν⊥ ∩ bB(0, 1))
whence
lim inf
r→0
DXu(B(p, r))
rQ−1
where ν := (ν1, . . . , νm, 0, . . . , 0) ∈ Rn and H n−1
denotes the Euclidean Hausdorff measure in
Rn. It is easily seen that, for any p ∈ Rn, there exist c > 0 and a neighborhood U of p such that
the function λ(q) := det ∇Fq(0)H n−1
A.5, this proves the first part of the statement.
(ν⊥ ∩ bBq(0, 1)) is such that λ ≥ c on U. By Corollary
By Theorem 2.39, the implication (35) is true in case k = 1 and u = χE for some E ⊆ Rn
with finite X-perimeter. If k = 1 and u ∈ BVX (Ω), we define Es := {u > s} and we apply
Theorem 2.40 (and, again, Theorem 2.39) to get
e
e
DXu(B) = +∞
−∞
PX(Es; B)ds = +∞
−∞ (cid:18)B∩∂∗Es
θsdH Q−1(cid:19) ds
for suitable positive functions θs. This allows to infer (35). In the general case k ≥ 1, it is
sufficient to recall inequality (25).
In order to prove (36) we consider u ∈ BVX(Ω; Rk) and a Borel subset B of Ω such that
B ∩ Su = ∅. If k = 1, by Theorem 2.40 we obtain again
DXu(B) = +∞
−∞ (cid:18)B∩∂∗Es
θs dH Q−1(cid:19) ds = B R
θs(p)χ∂∗Es(p) ds dH Q−1(p) = 0,
the last equality following from Proposition 2.22. In the case u ∈ BVX(Ω; Rk) with k ≥ 2, it
is sufficient to notice that B ∩ Su = ∅ implies B ∩ Suα = ∅ for every α = 1, . . . , k, and one
concludes using inequality (25).
(cid:3)
26
DON AND VITTONE
We now prove some of our main results.
Proof of Theorems 1.2 and 1.5. It is not restrictive to suppose k = 1. We first prove Theorem
1.2.
By the Coarea Formula we get a countable and dense set D ⊆ R such that for every t ∈ D
the level set {u > t} has finite X-perimeter. We prove that
Su \ L ⊆ [t∈D
∂∗{u > t}
(38)
where, as in Lemma 3.2, L denotes the H Q−1-negligible set
(cid:26)p ∈ Ω : lim sup
r→0 B(p,r)
u
Q
Q−1 dL n = +∞(cid:27) .
Theorem 1.2 is immediately implied by formula (38). In order to prove the latter, take p /∈ L
and suppose that p /∈St∈D ∂∗{u > t}; we will prove that p /∈ Su. By definition, p is either a
point of density 1 or a point of density 0 in {u > t} for every t ∈ D. Notice that for every
t ∈ D ∩ (0, +∞) one has
L n ({u > t} ∩ B(p, r))
L n(B(p, r))
≤
1
t B(p,r)
udL n ≤
1
t(cid:18) B(p,r)
u
Q
Q−1 dL n(cid:19) Q−1
Q
and therefore, if t ∈ D ∩ (0, +∞) is large enough, p is a point of density 0 for {u > t}.
Analogously, if t ∈ D ∩ (−∞, 0) and −t is large enough, p is a point of density 1 for {u > t}.
Hence we can find a real number
z = z(p) := sup {t ∈ D : {u > t} has density 1 at p} .
By the density of D in R we get that, for every t > z, {u > t} has density 0 at p and, for every
t < z, {u > t} has density 1 at p.
We prove now that z is the approximate limit of u at p. To this end define Eε := {u − z > ε}
and estimate
1
rQ B(p,r)
u − zdL n ≤ εC +
≤ εC +
1
1
u − zdL n
rQ Eε∩B(p,r)
rQ (L n(Eε ∩ B(p, r)))1/Q(cid:18)B(p,r)
rQ B(p,r)
(cid:19)1/Q(cid:18) 1
= εC +(cid:18) L n(Eε ∩ B(p, r))
rQ
u − z
u − z
Q
Q
Q−1 dL n(cid:19) Q−1
Q−1 dL n(cid:19) Q−1
Q
Q
.
Since both {u > z + ε} and {u < z − ε} have density 0 at p, one has
and, since p /∈ L, we get
L n(Eε ∩ B(p, r))
rQ
lim
r→0
= 0
lim sup
r→0
1
rQ B(p,r)
u − zdL n ≤ Cε,
from which we deduce that p /∈ Su, as desired.
We now prove Theorem 1.5. When property R holds, the countable X-rectifiability of Su
immediately follows from (38). We have to prove that H Q−1(Su \ Ju) = 0. Let ν = νSu be the
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
27
horizontal normal to Su and recall the notation B±
3.7 below, for H Q−1-almost every p ∈ Su there exist u+(p) and u−(p) in Rk such that
ν (p, r) introduced in (15). By Proposition
lim
r→0
1
rQ B+
ν(p)(p,r)
u − u+(p)dL n = lim
r→0
1
rQ B−
ν(p)(p,r)
u − u−(p)dL n = 0.
Notice that u+(p) 6= u−(p), for otherwise u would have an approximate limit at p. This implies
that p is an approximate X-jump point associated with the triple (u+(p), u−(p), ν(p)), and this
concludes the proof.
(cid:3)
A milder version of Theorem 1.5 holds when (Rn, X) satisfies the weaker property LR, that
we now introduce.
Definition 3.4 (Property LR). Let (Rn, X) be an equiregular CC space with homogeneous
dimension Q ∈ N. We say that (Rn, X) satisfies the property LR if, for every open set Ω ⊆ Rn
and every E ⊆ Rn with locally finite X-perimeter in Ω, the essential boundary ∂∗E ∩ Ω is
countably X-Lipschitz rectifiable.
The proof of the following result is an immediate consequence of (38).
Theorem 3.5. Let (Rn, X) be an equiregular CC space satisfying property LR and let u ∈
BVX (Ω; Rk). Then Su is countably X-Lipschitz rectifiable.
Before proving Proposition 3.7, that we used in the proof of Theorem 1.5, we state the
following theorem, which is a consequence of some results contained in [46]. We use the notation
B±
f (p, r) := {q ∈ B(p, r) : ±f (q) > 0}.
X(Ω) be such that Xf 6= 0 on Ω; let S be the C 1
Theorem 3.6. Let (Rn, X) be an equiregular CC space, let Ω ⊆ Rn be an open set and let
f ∈ C 1
X hypersurface S := {p ∈ Ω : f (p) = 0}.
Then there exist linear operators T +, T − : BVX,loc(Ω; Rk) → L1
loc(S, H Q−1) such that, for any
u ∈ BVX,loc(Ω; Rk), one has for H Q−1-a.e. p ∈ S
lim
r→0
1
rQ B+
f (p,r)
u − T +u(p)dL n = lim
r→0
1
rQ B−
f (p,r)
u − T −u(p)dL n = 0.
In particular, for H Q−1-a.e. p ∈ S
T ±u(p) = lim
r→0
1
rQ B±
f (p,r)
u dL n.
We can now prove the following proposition, where we implicitly use Remark 2.27.
Proposition 3.7. Let (Rn, X) be an equiregular CC space and let Ω ⊆ Rn be an open set.
Let R ⊆ Ω be a countably X-rectifiable set with horizontal normal νR. Then, for every u ∈
BVX (Ω; Rk) and for H Q−1-almost every p ∈ R there exists a couple (u+
R(p)) ∈ Rk × Rk
such that
R(p), u−
lim
r→0
1
rQ Ω∩B+
u − u+
R(p)dL n = lim
r→0
1
rQ Ω∩B−
νR(p)(p,r)
νR (p)(p,r)
3, then (u+
Moreover, if (Rn, X) satisfies property R and R = Ju
approximate X-jump triple for u at p in the sense of Definition 2.24.
3The jump set Ju is countably X-rectifiable by Theorem 1.5 and Remark 2.25.
u − u−
R(p)dL n = 0.
(39)
Ju(p), u−
Ju(p), νJu(p)) is an
28
DON AND VITTONE
Proof. We can assume without loss of generality that k = 1. Let u ∈ BVX(Ω) be fixed. By
definition of countable X-rectifiability we can find a family {Si : i ∈ N} of C 1
X hypersurfaces in
Rn such that
H Q−1 R \
Si! = 0.
∞[i=0
For every i ∈ N we can write, at least locally, Si = {fi = 0} and we can suppose that Xfi 6= 0
on Si. Formula (39) easily follows (with u±
R(p) = T ±u(p)) from Theorem 3.6 for H Q−1-a.e.
p ∈ R such that #{i ∈ N : p ∈ Si} = 1. It is then enough to show that, for any fixed couple
i, j ∈ N with i 6= j and for H Q−1-almost every p ∈ Si ∩ Sj, the equivalence
(T +
i u(p), T −
i u(p), νSi(p)) ≡ (T +
j u(p), T −
j u(p), νSj (p))
(40)
holds. Here, T ±
i , T ±
j are the trace operators provided by Theorem 3.6 with f = fi, fj.
Fix a point p ∈ Si ∩ Sj where νSi(p) = ±νSj (p); recall that this fact occurs at H Q−1-a.e.
Xfj(p) ; by Theorem 3.6 we have for
p ∈ Si ∩ Sj. Assume that νSi(p) = νSj (p), i.e., Xfi(p)
H Q−1-a.e. such p that
Xfi(p) = Xfj (p)
T ±
i (p) − T ±
j (p) = lim
r→0
≤ lim
r→0
≤ lim
r→0
1
1
u dL n −{±fj>0}∩B(p,r)
rQ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
{±fi>0}∩B(p,r)
rQ {fifj≤0}∩B(p,r)
L n({fifj ≤ 0} ∩ B(p, r))1/Q(cid:18)B(p,r)
udL n
1
rQ
udL n(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Q−1 dL n(cid:19) Q−1
Q
Q
u
.
By Remark 2.15 we have
lim
r→0
1
rQ
L n({fifj ≤ 0} ∩ B(p, r)) = 0,
while by Lemma 3.2 we also have that for H Q−1-almost every p ∈ Ω
lim sup
r→0
1
rQ B(p,r)
u
Q
Q−1 dL n < +∞.
This proves that T ±
argument shows that T ±
This proves (40), while the last statement of the proposition follows from Theorem 3.6.
j (p) for H Q−1-a.e. p ∈ Si ∩ Sj such that νSi(p) = νSj (p). A similar
j (p) holds for H Q−1-a.e. p ∈ Si ∩ Sj with νSi(p) = −νSj (p).
(cid:3)
i (p) = T ∓
i (p) = T ±
The problem of studying "intrinsic" measures of submanifolds of a CC space goes back to
M. Gromov [26, 0.6.b]: the interested reader might consult [32, 33, 34, 40] and the references
therein. Since we do not intend to dwell on such questions, we follow a different ("axiomatic")
path; this is based on the following definition, where we chose to work with the spherical
Hausdorff measure S Q−1, rather than the standard one, because the results mentioned above
(as well as [22, 23]) suggest S Q−1 to be more natural than the standard measure H Q−1.
Definition 3.8 (Property D). Let (Rn, X) be an equiregular CC space with homogeneous
dimension Q ∈ N. We say that (Rn, X) satisfies the property D if there exists a function
ζ : R × Sm−1 → (0, +∞) such that, for every C 1
X hypersurface S ⊆ Rn and every p ∈ S, one
has
S Q−1(S ∩ B(p, r))
= ζ(p, νS(p)).
lim
r→0
rQ−1
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
29
Remark 3.9. If (Rn, X) is an equiregular CC space satisfying property D and R ⊆ Rn is
X-rectifiable, then we have
S Q−1(R ∩ B(p, r))
lim
r→0
rQ−1
where ζ is as in Definition 3.8.
= ζ(p, νR(p))
for S Q−1-a.e. p ∈ R,
Let us prove this fact. Let Si, i ∈ N, be a family of C 1
X hypersurfaces such that S Q−1(R \
∪i∈NSi) = 0; it is enough to show that, for any fixed i ∈ N, we have
S Q−1(R ∩ B(p, r))
rQ−1
lim
r→0
= ζ(p, νR(p))
for S Q−1-a.e. p ∈ R ∩ Si.
Setting R∆Si := (R \ Si) ∪ (Si \ R), by Remark A.6 (applied with µ := S Q−1 (R∆Si)) we
obtain
S Q−1((R∆Si) ∩ B(p, r))
rQ−1
lim
r→0
= 0
for S Q−1-a.e. p ∈ R ∩ Si,
which gives for S Q−1-a.e. p ∈ R ∩ Si
S Q−1(R ∩ B(p, r))
rQ−1
= lim
r→0
S Q−1(Si ∩ B(p, r))
rQ−1
lim
r→0
= ζ(p, νSi(p)) = ζ(p, νR(p))
as desired.
Assuming properties R and D we are able to prove the following result, where we use the
notation u+
R, u−
R of Proposition 3.7.
Theorem 3.10. Let (Rn, X) be an equiregular CC space satisfying properties R and D; then,
there exists a function σ : Rn × Sm−1 → (0, +∞) such that the following holds. For every open
set Ω ⊆ Rn, u ∈ BVX(Ω; Rk) and every countably X-rectifiable set R ⊆ Rn one has
DXu R = σ(·, νR)(u+
R − u−
R) ⊗ νR S Q−1 R.
In particular, Dj
X u = σ(·, νu)(u+ − u−) ⊗ νu S Q−1 Ju.
Proof. We can assume without loss of generality that k = 1 and S Q−1(R) < ∞. By Theorem
1.5 and Proposition 3.3 we can also assume that R ⊆ Ju. Given p ∈ Rn we work in adapted
exponential coordinates Fp around p and we define
σ(p, ν) :=
det ∇Fp(0)H n−1
ζ(p, ν)
e
(ν⊥ ∩ bBp(0, 1))
where ζ is as in Definition 3.8 and, as in the proof of Proposition 3.3, H n−1
Euclidean Hausdorff measure in Rn.
e
denotes the
Let µR := DXu R; by Proposition 3.3 we have µR ≪ S Q−1 R. By Remark 3.9 we can
use [17, Theorem 2.9.8] (joint with [17, Theorem 2.8.17]) and it is enough to prove that for
S Q−1-a.e. p ∈ R
lim
r→0
µR(B(p, r))
S Q−1(R ∩ B(p, r))
= σ(p, νR(p))(u+
R(p) − u−
R(p))νR(p);
notice that the limit above exists S Q−1-almost everywhere. Taking into account Remark 3.9
and the fact that (by Remark A.6)
lim
r→0
DXu − µR(B(p, r))
rQ−1
= 0
for S Q−1-a.e. p ∈ R,
30
DON AND VITTONE
it suffices to prove that, for S Q−1-a.e. p ∈ R, there exists an infinitesimal sequence (ri) such
that
lim
i→+∞
DXu(B(p, ri))
rQ−1
i
= det ∇Fp(0)H n−1
e
(ν ⊥ ∩ bBp(0, 1))(u+
R(p) − u−
R(p))νR(p).
We prove that such a sequence exists at all points where lim supr→0
holds for S Q−1-a.e. p ∈ R due to Remark A.6.
DX u(B(p,r))
rQ−1
< ∞, which
L1
loc(Rn) to
Let then such a p ∈ R be fixed; since R ⊆ Ju, the functions eur := u ◦ Fp ◦ δr converge in
wp(y) :=(u+(p)
u−(p)
if eLνR(p)(y) ≥ 0
if eLνR(p)(y) < 0,
where we used the fact that νR = νJu = νu S Q−1-a.e. on R. Let eu := u ◦ Fp; since (recall
notation (31)) D eX reur(eBr(0, )) = D eXeu(eB(0, r))/rQ−1 is bounded as r → 0 for any positive
, by Remark 2.45 the sequence D eX reur weakly∗ converges in Rn to D bX wp as r → 0. Let si be
an infinitesimal sequence such that D eX sieusi weakly∗ to some measure λ in Rn; let ∈ (0, 1) be
such that λ(∂bBp(0, )) = 0 (which holds for all except at most countably many ) and define
ri := si. Proposition 2.36 gives
DX u(B(p, ri))
lim
i→∞
rQ−1
i
= det ∇Fp(0)
= det ∇Fp(0)
lim
i→∞
lim
i→∞
rQ−1
i
D eXeu(eB(0, ri))
D eX sieusi(eBsi(0, ))
Q−1
.
We prove in a moment that
lim
i→∞
assuming this to be true, we have
D eX sieusi(eBsi(0, ))
Q−1
=
D bXwp(bBp(0, ))
Q−1
;
lim
i→∞
DXu(B(p, ri))
rQ−1
i
= det ∇Fp(0)
D bX wp(bBp(0, ))
Q−1
= det ∇Fp(0)H n−1
e
R(p) − u−
R(p))νR(p).
(41)
(42)
and the proof would be concluded.
Let us prove (41). Defining
µi := D eX sieusi
µi
∗⇀ µ
eBsi(0, ),
and
and taking into account [2, Proposition 1.62 (b)], it will suffice to show that
(ν⊥ ∩ bBp(0, 1))(u+
µ := D bX wp bBp(0, )
µi ∗⇀ λ bBp(0, ).
Concerning the first statement in (42), fix a test function ϕ ∈ C 0
c (Rn); then
lim
i→∞ ϕ dµi = lim
i→∞ bBp(0,)
i→∞ bBp(0,)
i→∞ eBsi (0,)
ϕ dD eX sieusi
ϕ dD eX sieusi + eBsi (0,)\ bBp(0,)
ϕ dD bXwp,
= lim
= lim
ϕ dD eX sieusi − bBp(0,)\ eBsi (0,)
ϕ dD eX sieusi
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
31
that (denoting by ∆ the symmetric difference of sets)
where the last equality follows from the weak∗ convergence of D eX sieusi to D bXwp and the fact
that, in turn, can be proved as follows. For any ε > 0 there exists δ ∈ (0, ) such that
lim
i→∞
D eX sieusi(eBsi(0, )∆bBp(0, )) = 0
λ(cid:0)bBp(0, + δ) \ bBp(0, − δ)(cid:1) < ε;
by Theorem 2.9 we obtain
lim sup
i→∞
D eX sieusi(eBsi(0, )∆bBp(0, )) ≤ lim sup
i→∞
D eX sieusi(cid:0)bBp(0, + δ) \ bBp(0, − δ)(cid:1)
≤λ(cid:0)bBp(0, + δ) \ bBp(0, − δ)(cid:1) < ε,
where we used [2, Proposition 1.62 (a)].
The first statement in (42) is proved; we are left with the second one, which can be easily
(cid:3)
proved by the very same argument taking into account that µi = D eX sieusi eBsi(0, ).
Let us recall once more the notation up,r := fflB(p,r) u dL n.
Lemma 3.11. Let (Rn, X) be an equiregular CC space of homogeneous dimension Q and let
Ω ⊆ Rn be an open bounded set. Then there exist C = C(Ω) > 0 and R = R(Ω) > 0 such that,
for every p ∈ Ω, every u ∈ BVX(Ω; Rk) and every 0 < r < min{R, 1
2d(p, ∂Ω)}, one has
up,2r − up,r ≤ Cr1−QDXu(B(p, 2r)).
Proof. We use Theorems 2.2 and 2.41 to estimate
up,2r − up,r =(cid:12)(cid:12)(cid:12)(cid:12) B(p,r)
≤C(cid:18) B(p,2r)
(u − up,2r) dL n(cid:12)(cid:12)(cid:12)(cid:12) ≤ C B(p,2r)
Q−1 dL n(cid:19) Q−1
u − up,2r
Q
Q
u − up,2r dL n
≤ Cr1−QDXu(B(p, 2r)).
As mentioned in the Introduction, the next lemma possesses its own interest and it is the
key tool in the proof of Theorem 1.1.
(cid:3)
Lemma 3.12. Let (Rn, X) be an equiregular CC space of homogeneous dimension Q and let
Ω ⊆ Rn be an open bounded set. Then there exist C = C(Ω) > 0 and R = R(Ω) > 0 such that
the following holds: for every u ∈ BVX(Ω; Rk), p ∈ Ω \ Su and 0 < r < min{R, 1
2 d(p, ∂Ω)} one
has
B(p,r)
u(q) − u⋆(p)
d(p, q)
In particular
dL n(q) ≤ C(cid:18)DXu(B(p, r)) + 1
0
DXu(B(p, tr))
tQ
dt(cid:19) .
B(p,r)
u(q) − u⋆(p)
d(p, q)
dL n(q) ≤ C 2
0
DX u(B(p, tr))
tQ
dt.
32
DON AND VITTONE
Proof. Let u, p, r be as in the statement; we introduce the compact notation ui := up,2−ir, i ∈ N.
Since ui → u⋆(p) as i → ∞ we estimate
and use Lemma 3.11 and Theorem 2.41 to get
≤
≤
2i
2i
u(q) − u⋆(p)
2−ir
d(p, q)
dL n(q)
dL n(q)
u(q) − u⋆(p)
∞Xj=i−1
B(p,r)
∞Xi=1
B(p,2−i+1r)\B(p,2−ir)
r B(p,2−i+1r)\B(p,2−ir)(cid:18)u(q) − ui−1 +
uj − uj+1(cid:19)dL n(q)
∞Xi=1
r 2−irDXu(B(p, 21−ir)) +
∞Xi=1
∞Xj=i−1(cid:0)21−ir(cid:1)Q(cid:0)2−(j+1)r(cid:1)1−Q
∞Xi=1 DXu(B(p, 21−ir)) +
2(j−i+1)(Q−1)DXu(B(p, 2−jr))!
∞Xj=i−1
∞Xk=0(cid:16)1 + 1 + 2Q−1 + (2Q−1)2 + · · · + (2Q−1)k(cid:17)DXu(B(p, 2−kr))
∞Xk=0
DXu(B(p, 2−kr)).
2(k+1)(Q−1) − 1
2Q−1 − 1
≤C
≤C
=C
≤C
Since Q ≥ 2 we have 2Q−1 − 1 ≥ 2Q−1
2
, hence
DXu(B(p, 2−jr))!
B(p,r)
u(q) − u⋆(p)
d(p, q)
as desired.
dL n(q) ≤ C
2k(Q−1)DXu(B(p, 2−kr))
∞Xk=0
= C DXu(B(p, r)) +
∞Xk=1
= C DXu(B(p, r)) +
∞Xk=1
≤ C DXu(B(p, r)) +
∞Xk=1
= C(cid:18)DXu(B(p, r)) + 1
0
2k(Q−1)DXu(B(p, 2−kr))!
21−k
21−k
2kQDXu(B(p, 2−kr))dt!
DXu(B(p, tr))
dt!
tQ
2−k
2−k
DXu(B(p, tr))
tQ
dt(cid:19) ,
(cid:3)
We can now prove one of our main results; recall that we denote by Dap
X u(p) the approximate
X-gradient of u at p.
Proof of Theorem 1.1. We can assume without loss of generality that k = 1. Suppose that
DX u = vL n + Ds
Xu is the Radon-Nykodým decomposition of the measure DX u with respect
to L n. By the Radon-Nykodým Theorem in doubling metric spaces (see e.g. [43, Theorem 4.7
and Remark 4.5]), at L n-almost every p ∈ Ω we have
X u(B(p, r))
Ds
(43)
lim
r→0
rQ
= 0.
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
33
It is sufficient to prove that, for every p ∈ Ω \ (Su ∪ Sv) for which (43) holds, u is approximately
X-differentiable at p with Dap
X u(p) = v⋆(p).
Let R > 0 and f ∈ C 1(B(p, R)) be such that f (p) = 0 and Xf (p) = v⋆(p) and define
w(q) := u(q) − u⋆(p) − f (q)
Then w ∈ BVX(B(p, R)), p /∈ Sw and w⋆(p) = 0. We are in a position to apply Lemma 3.12 to
the function w and get C > 0 so that, for small enough r,
1
rQ B(p,r)
u(q) − u⋆(p) − f (q)
d(p, q)
dL n(q) ≤
C
rQ 2
0
DXw(B(p, tr))
tQ
dt
≤ C sup
t∈(0,2)
DXw(B(p, tr))
(tr)Q
.
It is then enough to show that limr→0 r−QDXw(B(p, r)) = 0. Taking into account that
DX w = (v − Xf )L n + Ds
Xu and (43), it suffices to check that
lim
r→0
1
rQ B(p,r)
v − Xf dL n = 0,
which follows by the generalized Lebesgue's differentiation theorem (see e.g. [28, Section 2.7])
and the inequality v − Xf ≤ v − v⋆(p) + v⋆(p) − Xf .
(cid:3)
As for classical BV functions (see e.g.
[2, pag. 177], the (approximate) convergence of
u ∈ BVX to u⋆(p) at points p /∈ Su can be improved in a L1∗
-sense, as we now state.
Proposition 3.13. Let (Rn, X) be an equiregular CC space, Ω ⊆ Rn an open set and let
u ∈ BVX(Ω). Then
lim
r→0 B(p,r)
u − u⋆(p)
Q
Q−1 dL n = 0
for H Q−1-a.e. p ∈ Ω \ Su.
Proof. We first prove that
lim
r→0
DXu(B(p, r))
rQ−1
= 0
for H Q−1-a.e. p ∈ Ω \ Su.
(44)
Let t > 0 be fixed and consider the set
Et :=(cid:26)p ∈ Ω \ Su : lim sup
r→0
DXu(B(p, r))
rQ−1
> t(cid:27) .
By Proposition A.4 one has H Q−1(Et) < ∞; Proposition 3.3 then implies that DXu(Et) = 0
and again Proposition A.4 gives H Q−1(Et) = 0. Since this is true for all positive t, formula
(44) immediately follows.
Combining Theorem 2.41 and (44) we immediately get that for H Q−1-a.e. p ∈ Ω
lim
r→0 B(p,r)
u − up,r
Q
Q−1 dL n = 0.
The conclusion follows by
u − u⋆(p)
Q
Q−1 ≤ 2
together with u⋆(p) = limr→0 up,r.
1
Q−1(cid:16)up,r − u⋆(p)
Q
Q−1 + u − up,r
Q
Q−1(cid:17) .
(cid:3)
34
DON AND VITTONE
When (Rn, X) satisfies property R, Ω ⊆ Rn is open and u ∈ BVX(Ω, Rk), by Theorem 1.5
the precise representative up
if p ∈ Ω \ Su
if p ∈ Ju
(45)
up(p) :=
u⋆(p)
u+(p) + u−(p)
2
is defined H Q−1-a.e. on Ω. We have the following result.
Theorem 3.14. Let (Rn, X) be an equiregular CC space satisfying property R, Ω ⊆ Rn an
open set and let u ∈ BVX(Ω; Rk). Then
lim
r→0 B(p,r)
u dL n = up(p)
for H Q−1-a.e. p ∈ Ω.
Proof. The statement easily follows for H Q−1-a.e. p ∈ Ω\ Su by Proposition 3.13. By Theorem
1.5 it suffices to prove the statement for all p ∈ Ju, which directly follows from Proposition
2.23 and Definition 2.24.
(cid:3)
Remark 3.15. When (Rn, X) satisfies property R, then Dc
is enough to combine Proposition 3.3 and Theorem 1.5.
Xu = Ds
Xu (Ω \ Su): to see this, it
We now want to study the properties of the decomposition DXu = Da
Xu + Dc
Xu + Dj
Xu; recall
that H 1
e denotes the Euclidean Hausdorff measure in Rn.
Theorem 3.16 (Properties of Cantor part and jump part). Let u ∈ BVX (Ω; Rk). Then the
following facts hold:
(a) Da
Xu = DXu (Ω \ S) and Ds
Xu = DXu S, where
S :=(cid:26)p ∈ Ω : lim
r→0
DXu(B(p, r))
rQ
= +∞(cid:27) .
Moreover, if E ⊆ Rk is such that H 1
e (E) = 0, then Dap
X u = 0 L n-a.e. in (u⋆)−1(E).
(b) Let Θu ⊆ S be defined by
Θu :=(cid:26)p ∈ Ω : L(p) := lim inf
r→0
DXu(B(p, r))
rQ−1
> 0(cid:27) .
Then Ju ⊆ Θu.
Moreover, if (Rn, X) satisfies property R, then
(c) H Q−1(Θu \ Ju) = 0 and Dj
Xu = DXu Θu. More generally, for every Borel set Σ
containing Ju and σ-finite with respect to H Q−1 we have Dj
Xu = DX u Σ.
Xu = DXu (S \ Θu).
(d) Dc
(e) if B ⊆ Ω is such that either H Q−1 B is σ-finite or B = (u⋆) −1(E) for some H 1
e -
negligible set E ⊆ Rk, then Dc
Xu(B) = 0.
Proof. In order to prove the first part of statement (a) it is sufficient to apply Radon-Nykodým
Theorem in doubling metric spaces (see e.g. [43, Theorem 4.7 and Remark 4.5]). Concerning
the second part, assume first that k = 1 and let B := (u⋆)−1(E). By Proposition 2.22, for any
t /∈ E we have B ∩ ∂∗{u > t} = ∅. By Theorems 2.40 and 2.39 we obtain
DXu(B) = R
PX({u > t} ∩ B) dt = 0 = R\E ∂∗{u>t}∩B
θtdH Q−1 dt = 0,
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
35
where θt denote suitable positive functions. When k ≥ 1 and j = 1, . . . , k we set Ej := {t ∈ R :
t = zj for some z ∈ E}; the set Ej is such that L 1(Ej) = 0 and by (25)
DXu(B) ≤
DXuj(B) ≤
kXj=1
kXj=1
DXuj(((uj)⋆)−1(Ej)) = 0.
We then conclude by Theorem 1.1.
By (37) in the proof of Proposition 3.3 we have Ju ⊆ Θu, and statement (b) follows.
We now prove (c). Applying Proposition A.4 we get that for every h ∈ N \ {0}
DXu {L ≥ 1
h } ≥
1
h
ωQ−1H Q−1 {L ≥ 1
h},
(46)
where L is defined in statement (b). In particular H Q−1(cid:0){L ≥ 1
h }(cid:1) < +∞. By (36)
DXu(cid:0){L ≥ 1
h} \ Su(cid:1) = 0
and consequently (by (46)) also H Q−1({L ≥ 1
h } ր Θu, on passing
to the limit for h → +∞ we get H Q−1(Θu \ Su) = 0. Taking Theorem 1.5 into account, we
conclude that H Q−1(Θu \ Ju) = 0.
Let now Σ be as in statement (c). Then, taking into account Proposition 3.3 and the fact that
H Q−1(Su \ Ju) = 0, we have
h} \ Su) = 0. Since {L ≥ 1
DX u Σ = DXu Ju + DXu (Σ \ Ju)
= Dj
= Dj
Xu + DXu (Σ \ Su) + DX u (Σ ∩ Su \ Ju)
Xu + DXu (Σ \ Su).
Since Σ is σ-finite with respect to H Q−1, using (36) we get that DXu (Σ \ Su) = 0, and so
DX u Σ = Dj
X u.
Statement (d) follows from (a), (b), (c) and the decomposition DXu = Da
Xu + Dc
Xu + Dj
X u,
which immediately give that Dc
Xu = DXu (S \ Θu).
We prove (e) in case H Q−1 B is σ-finite; we can assume (see e.g. [2, Theorem 1.43]) that
B is a Borel set. Using Proposition 3.3 and Theorem 1.5 we get that DXu(B \ Ju) = 0, which
gives (Da
Concerning the second part of statement (e), suppose first that k = 1 and let B = (u⋆) −1(E)
with L 1(E) = 0. By Proposition 2.22 we know that ∂∗{u > t} ∩ B = ∅ for every t /∈ E.
Applying the Coarea Formula of Theorem 2.40 we get
X u) B = 0.
X u + Dc
DXu(B) = E ∂∗{u>t}∩B
θtdH Q−1dt = 0
In the general case k ≥ 2 define for every α = 1, . . . , k the sets
for suitable functions θt.
Eα := πα(E), where πα denotes the canonical projection πα(x1, . . . , xk) = xα. Noticing that
L 1(Eα) ≤ H 1
e (E) = 0, we can use (25) to estimate
DXu((u⋆) −1(E)) ≤
kXα=1
DXuα((u⋆) −1(E)) ≤
kXα=1
DXuα(((uα)⋆) −1(Eα)) = 0.
(cid:3)
36
DON AND VITTONE
4. Applications to some classes of Carnot groups
Some of the main results of this paper rely on properties R, LR or D; in this section we
show how they can be in some meaningful CC spaces and, in particular, in some large classes
of Carnot groups.
We start by introducing the reduced boundary FXE of a set E with finite X-perimeter.
Recall that the reduced boundary was the object originally considered by E. De Giorgi in the
seminal paper [13] about the rectifiability of sets with finite (Euclidean) perimeter in Rn.
Definition 4.1 (Reduced boundary). Let E ⊆ Rn be a set with locally finite X-perimeter.
The X-reduced boundary FXE of E is the set of points p ∈ Rn such that PX (E, B(p, r)) > 0
for any r > 0 and the limit
DX χE(B(p, r))
DXχE(B(p, r))
r→0
eνE(p) := lim
exists with eνE(p) = 1.
For sets with finite (Euclidean) perimeter in Rn the symmetric difference between the essential
boundary and the reduced one is H n−1
-negligible, see e.g. [2, Theorem 3.61]. In our setting we
have the following result, which is a known consequence of Theorem 2.39, see e.g. [22, Theorem
7.3] for the Heisenberg group case and [23, Lemma 2.26] for step 2 Carnot groups.
e
Theorem 4.2. Let (Rn, X) be an equiregular CC space of homogeneous dimension Q and let
E ⊆ Rn be a set of locally finite X-perimeter. Then H Q−1(∂∗E \ FXE) = 0.
Proof. By Theorem 2.39 we have DXχE = θνEH Q−1 ∂∗E for a suitable positive function θ.
Therefore it is enough to prove that, for H Q−1-almost every p ∈ ∂∗E, one has
lim
r→0
DX χE(B(p, r))
DχE(B(p, r))
= νE(p).
This fact directly follows from [17, Theorem 2.9.8] taking into account Theorem 2.39 and [17,
Theorem 2.8.17].
(cid:3)
The proof of Theorem 4.2 also shows thateνE = νE H Q−1-a.e. on FXE.
The papers [22, 23, 35] prove the countable X-rectifiability of the reduced boundary of sets
with locally finite X-perimeter in, respectively, Heisenberg groups, Carnot groups of step 2, and
Carnot groups of type ⋆. These results, in conjunction with Theorem 4.2, show that property
R is satisfied in these settings.
Actually, Theorem 4.2 and the results about blow-up and representation of the X-perimeter
available in Heisenberg groups ([22, Theorems 4.1 and 7.1]), step 2 Carnot groups ([23, Theo-
rems 3.1 and 3.9]) and Carnot groups of type ⋆ [35, Theorems 4.12 and 4.13] imply that also
property D is satisfied in these settings.
Using also the left-invariance of the structure we can conclude what follows.
Theorem 4.3. Heisenberg groups, Carnot groups of step 2 and Carnot groups of type ⋆ satisfy
properties R and D. In particular, Theorems 1.5, 1.6, 1.7 and 3.14 hold in these settings.
Moreover, the function σ(p, ν) appearing in 1.7 and 3.14 does not depend on the point p ∈ Rn.
In the paper [14] the class of Carnot groups G satisfying the following assumption (see e.g.
[37] for the notion of abnormal curve)
there exists at least one direction V in the first layer of the stratified Lie algebra
of G such that t 7→ exp(tV ) is not an abnormal curve
(47)
is considered. This class includes, for instance, the Engel group, which is the simplest example
where the rectifiability problem for sets with finite X-perimeter is open. One of the main
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
37
results of [14] is the following one: for any set E with finite X-perimeter in a Carnot group G
satisfying (47), the reduced boundary FXE is countably X-Lipschitz rectifiable. Together with
Theorem 4.2, this gives the following result.
Theorem 4.4. The property LR is satisfied in all Carnot groups G such that (47) holds; in
particular, Theorem 3.5 holds in such groups.
Appendix A. Some tools from Geometric Measure Theory in metric spaces.
Proposition A.1. Let u ∈ L1
loc(Ω; Rk). If p ∈ Ω \ Su, then, for any ε > 0, the set
Eε := {q ∈ Ω : u(q) − u⋆(p) > ε}
has density 0 at p. Conversely, if u ∈ L∞
set
loc(Ω; Rk) and z ∈ Rk are such that, for any ε > 0, the
has density 0 at p, then p ∈ Ω \ Su and z = u⋆(p).
In particular, if k = 1 and p ∈ Ω \ Su and t 6= u⋆(p), then p /∈ ∂∗{q ∈ Ω : u(q) > t}.
Eε := {q ∈ Ω : u(q) − z > ε}
Proof. Suppose p ∈ Ω \ Su. By Chebychev inequality we have
ε
L n(Eε ∩ B(p, r))
L n(B(p, r))
≤ B(p,r)
u − u⋆(p)dL n → 0
as r → 0.
Conversely, suppose that u and z are as in the statement. Then we have for any r ∈ (0, 1)
B(p,r)
u − zdL n ≤ (kukL∞(B(p,1);Rk) + z)
L n(B(p, r) ∩ Eε))
L n(B(p, r))
+ ε
L n(B(p, r) \ Eε))
L n(B(p, r))
,
which is infinitesimal as r → 0.
Finally, consider p ∈ Ω \ Su and let t 6= u⋆(p). We already know that both {u > u⋆(p) + ε}
and {u < u⋆(p) − ε} have density 0 at p for every ε > 0. If t > u⋆(p), then choosing ε =
t − u⋆(p) we have that {u > t} has density 0 at p. If t < u⋆(p) then choose η > 0 such that
ε = u⋆(p) − t − η > 0 to infer that {u < t + η} has density 0 at p, and consequently {u ≥ t + η}
has density 1 at p. This implies that also {u > t} has density 1 at p.
(cid:3)
The following result is classical, see e.g. [43] or [28].
Theorem A.2 (5r-Covering Lemma). Let (M, d) be a separable metric space and let B a family
of closed balls in M such that
sup {diam B : B ∈ B} < +∞.
Denote by 5B the closed metric ball with same center as B and radius 5 times larger than that
of B. Then there exists a countable and pairwise disjoint subfamily F ⊆ B such that
[ B ⊆ [B∈F
5B.
Definition A.3 (Hausdorff measures). Let (M, d) be a metric space and k ≥ 0. For any δ > 0
and any E ⊆ M we define
H k
δ (E) :=
S k
δ (E) :=
ωk
ωk
2k inf( ∞Xh=0
2k inf( ∞Xh=0
(diam Eh)k : E ⊆
(diam Bh)k : E ⊆
Eh, diam Eh < δ)
∞[h=0
Bh, Bh balls with diam Bh < δ) ,
∞[h=0
38
DON AND VITTONE
where ωα := πα/2Γ(1+α/2)−1 and Γ(t) := ´ +∞
measure H k(E) and the spherical Hausdorff measure S k(E) of E are
0
st−1e−sds is the Euler Γ function. The Hausdorff
H k(E) := sup
δ>0
S k(E) := sup
δ>0
S k
δ (E) = lim
δ→0
H k
δ (E) = lim
δ→0
H k
δ (E)
S k
δ (E).
The useful inequalities H k ≤ S k ≤ 2kH k are classical.
If (M, d, µ) is a metric measure space, k ≥ 0 and x ∈ M, we define the upper k-density
k(µ, x) and the lower k-density Θ∗k(µ, x) of µ at x as
Θ∗
Θ∗
k(µ, x) := lim sup
r→0
µ(B(x, r))
ωkrk
,
Θ∗k(µ, x) := lim inf
r→0
µ(B(x, r))
ωkrk
.
k(H k E, x) and Θ∗k(E, x) :=
For every Borel set E ⊆ Rn we will also write Θ∗
Θ∗k(H k E, x). If Θ∗
k(µ, x) = Θ∗k(µ, x), then the common value is denoted by Θk(µ, x) and it
will be called k-density of µ at x. Hausdorff measures and densities are linked by Propositions
A.4 and A.5 below. A proof of Proposition A.4 can be found for instance in [43, Theorem 3.2];
in the latter reference, statement (i) below is stated with H k in place of S k, but the careful
reader will notice that the proof is indeed provided for this stronger version.
k(E, x) := Θ∗
Proposition A.4. Let (M, d) be a separable metric space, let µ be a Borel regular Radon
measure on M, let E ⊆ M be a Borel set and let t > 0. Then the following facts hold.
(i) If Θ∗
(ii) If Θ∗
k(µ, x) ≥ t for every x ∈ E, then µ ≥ tS k E.
k(µ, x) ≤ t for every x ∈ E, then µ ≤ 2ktH k E.
In particular, for H k-almost every x ∈ Rn we have Θ∗
k(µ, x) < +∞.
Corollary A.5. Let (M, d) be a separable metric space, let µ be a Borel regular Radon measure
on M, let E ⊆ M be a Borel set and let f : E → R be a strictly positive function. Then the
following facts hold.
(i) If Θ∗
(ii) If Θ∗
k(µ, x) ≥ f (x) for every x ∈ E , then µ ≥ f S k E.
k(µ, x) ≤ f (x) for every x ∈ E , then µ ≤ 2kf H k E.
Proof. (i) Let ε > 0 and define for every j ∈ Z the set
Ej := {x ∈ E : (1 + ε)j < f (x) ≤ (1 + ε)j+1}.
Suppose that Θ∗
k(µ, x) ≥ f (x) for every x ∈ E. Then, using (i) of Proposition A.4 we get
µ =Xj∈Z
µ Ej ≥Xj∈Z
(1 + ε)jS k Ej ≥Xj∈Z
f
1 + ε
S k Ej =
f
1 + ε
S k E.
The statement follows by the arbitrariness of ε.
(ii) Using (ii) of Proposition A.4 we have
µ Ej ≤Xj∈Z
µ =Xj∈Z
≤Xj∈Z
2k(1 + ε)j+1H k Ej
2k(1 + ε)f S k Ej = 2k(1 + ε)f S k E.
The statement follows by the arbitrariness of ε.
(cid:3)
As a consequence of the Corollary A.5 we have the following remark.
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
39
Remark A.6. Under the same assumptions of Corollary A.5, for H k-almost every x ∈ Rn we
have Θ∗
k(µ, x) < +∞ and for any Borel set B ⊆ Rn the implication
µ(B) = 0 =⇒ Θk(µ, x) = 0 for H k-a.e. x ∈ B
holds. In particular, if µ = gH k E we have Θk(µ, x) = 0 for H k-almost every x ∈ Rn \ E.
Definition A.7 (Porous sets). Let (M, d) be a metric space and let E ⊆ M be a Borel set.
Then E is said to be porous if there esist α ∈ (0, 1) and R > 0 such that for every x ∈ M and
every r ∈ (0, R) there exists y ∈ M such that
B(y, αr) ⊆ B(x, r) \ E.
Proposition A.8. Let (M, d) be a locally compact and separable metric space, µ a doubling
Radon measure on M and let E ⊆ M be a porous set. Then E has no points of density 1 and,
in particular, µ(E) = 0.
Proof. Let α and R be as in Definition A.7. Suppose by contradiction there exists x ∈ E1. For
every r ∈ (0, R) there exists y ∈ M such that B(y, αr) ⊆ B(x, r) \ E. This implies that
µ(B(x, r) \ E)
µ(B(x, r))
≥
µ(B(y, αr))
µ(B(x, r))
≥ C,
where C > 0 depends on α and the doubling constant of µ. Letting r → 0 and taking into
account that x ∈ (M \ E)0, we get a contradiction. The last part of the statement follows from
the generalized Lebesgue Theorem, see e.g. [28, Theorem 1.8].
(cid:3)
Proposition A.9. Let (Rn, X) be a geodesic equiregular CC space; then, for every p ∈ Rn and
for every r > 0 one has L n(∂B(p, r)) = 0.
Proof. By Proposition A.8 it is sufficient to prove that ∂B(p, r) is a porous set. Take q ∈
∂B(p, r) and consider a length minimizing absolutely continuous path γ : [0, r] → Rn such that
γ(0) = p, γ(r) = q and for every t ∈ [0, r] one has d(p, γ(t)) = t. Consider ε ∈ (0, 2r] and set
y := γ(r − ε
2) ∩ ∂B(p, r) = ∅, i.e., ∂B(p, r) is
porous.
(cid:3)
2) ⊆ B(q, ε), hence B(y, ε
2 ) ∈ B(p, r). Then B(y, ε
Appendix B. Proofs of some results about jumps and approximate
differentiability points
Proof of Proposition 2.28. (i) We can without loss of generality assume that k = 1. Consider
a countable dense subset {(ah, bh, νh) : h ∈ N} of R × R × Sm−1 and, for every h ∈ N, define
wh : Rn → R by
bh
wh(y) :=(ah
∞[h=0(cid:26)p ∈ Ω : lim sup
if eLνh(y) ≥ 0,
if eLνh(y) < 0.
r→0 A(r)
∞\ℓ=1
1
ℓ(cid:27) .
(48)
We first prove that (recalling the notation (7))
(Ω \ Su) ∪ Ju =
u ◦ Fp − whdL n <
The inclusion ⊆ in (48) is straightforward by Remark 2.20 and Proposition 2.26. In order to
prove the opposite inclusion, consider p ∈ Ω such that for every ℓ ∈ N \ {0} there exists whℓ
such that
lim sup
r→0 A(r)
u ◦ Fp − whℓdL n <
1
ℓ
.
40
DON AND VITTONE
We prove that, possibly passing to a subsequence, there exist a, b and ν such that (whℓ) is
convergent in L1(A(1)) to
Up to subsequences we can suppose that (νhℓ) converges to some ν. Define C := L n (A(1))
and let ℓ ∈ N be such that for every ℓ, k ≥ ℓ the set
A+(1) :=ny ∈ A(1) :eLνhℓ
(y) > 0o
is such that L n(A+(1)) ≥ 1
4 C. By a change of variables, for such h and k one has
b
w(y) :=(a
if eLν(y) ≥ 0,
if eLν(y) < 0.
(y) > 0 and eLνhk
C A+(1)
CrQ A(r)
whℓ − whkdL n =
u ◦ Fp − whℓdL n + 4 A(r)
ahℓ − ahkdL n ≤
4
4
whℓ − whkdL n
whℓ − whkdL n
u ◦ Fp − whkdL n.
ahℓ − ahk = A+(1)
C A(1)
≤ 4 A(r)
≤
4
Passing to the lim sup as r → 0 we get that (ahℓ) is Cauchy and therefore convergent to some
a ∈ R. Using the same technique we also get that (bhℓ) is convergent to some b ∈ R, and whℓ
converges in L1(A(1)) to w. Now, for sufficiently large ℓ ∈ N and for sufficiently small r > 0,
from
A(r)
u ◦ Fp ◦ δr − wdL n ≤ A(r)
u ◦ Fp ◦ δr − whℓdL n +A(1)
whℓ − wdL n
we deduce the remaining inclusion ⊇ in (48).
Notice that the right-hand side of (48) is a Borel set if, for any h ∈ N, and any small enough
r, the function
p 7−→ A(r)
u ◦ Fp − whdL n
(49)
is continuous. This is clearly true if u is of class C ∞. For general u, fix p ∈ Ω, r > 0
and ε > 0 and consider v ∈ C ∞(Ω) such that ku − vkL1(B(p,C1r)) < ε, where C1 is such that
Fp(A(r)) ⋐ B(p, C1r). Applying the triangular inequality, we find
u ◦ Fp − u ◦ FqdL n
A(r)
≤ A(r)(cid:0)u ◦ Fp − v ◦ Fp + v ◦ Fp − v ◦ Fq + v ◦ Fq − u ◦ Fq(cid:1)dL n < Cε,
for some C > 0, for every sufficiently small r and for every q sufficiently close to p. This proves
that the function in (49) is continuous. It follows that (Ω \ Su) ∪ Ju is a Borel set: then, also
Ju is a Borel set, for Ω \ Su is Borel and it is disjoint from Ju.
Select now for any p ∈ Ju an X-jump triple (u+(p), u−(p), ν(p)) according to Definition 2.24.
Define φ : Ju → Rm by φ(p) := (u+(p) − u−(p))ν(p). We prove that φ is Borel, so that also ν
is Borel up to re-defining it as ν(p) = φ(p)/φ(p). Set
wp(y) :=(u+(p)
u−(p)
if eLν(p)(y) > 0;
if eLν(p)(y) < 0,
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
41
and
eA(r) :=(y ∈ Rn : (y1, . . . , ym) +
nXj=m+1
1
wj ≤ r) .
yj
Notice that the sets eA(r) are invariant under rotations of the first m coordinates. By Proposition
2.26 we have that for every ψ ∈ C ∞
eA(1)
wp∂iψdL n = lim
c (eA(1)) and every i = 1, . . . , n
ε→0 eA(1)
(u ◦ Fp ◦ δε)∂iψdL n
= lim
ε→0
1
εQ eA(ε)
u(Fp(y))∂iψ(δε−1(y))dL n(y).
Hence, for every ψ ∈ C ∞
c (eA(1)) and for every i = 1, . . . , n the function
wp∂iψdL n
p 7−→ eA(1)
is Borel. Fix p ∈ Ju and consider a sequence (ψh) in C ∞
the (Euclidean) measure derivative of wp we obtain that for every i = 1, . . . , n
(cid:0)eA(1) ∩ {eLν(p) = 0}(cid:1)
h eA(1)
ψhdDiwp = − lim
φi(p)H n−1
e
= Diwp(eA(1)) = lim
c (eA(1)) converging to χ eA(1). Computing
h eA(1)
wp∂iψhdL n.
e
Since H n−1
is a Borel function, and therefore ν is Borel.
(eA(1) ∩ {eLν(p) = 0}) does not depend on p we deduce by the previous step that φ
Finally, by Proposition 2.26 we have
u+(p) = lim
ε→0
1
εQ A(ε)
χ{ eLν(p)>0}u ◦ FpdL n
and this concludes the proof.
The proof of (ii) is completely analogous to the Euclidean case, see [2].
(cid:3)
Proof of Proposition 2.32. We can assume without loss of generality that k = 1. Consider a
dense subset {zi : i ∈ N} of Rm. Reasoning as in the proof of Proposition 2.28 one can prove
that
Du =
∞\h=1
∞[i=0(cid:26)p ∈ Ω \ Su : lim sup
→0
which implies that Du is a Borel set.
We now prove that Dap
1
rQ+1 A(r)(cid:12)(cid:12)(cid:12)u ◦ Fp − u⋆(p) −eLzi(cid:12)(cid:12)(cid:12) dL n <
1
h(cid:27)
X u is Borel. Using Theorem 2.6, for any p ∈ Du one has
lim
ε→0
1
εQ+1 δεP(cid:12)(cid:12)(cid:12)u ◦ Fp − u⋆(p) −eLDap
X u(p)(cid:12)(cid:12)(cid:12) dL n = 0,
where for every n-tuple of positive real numbers (ℓ1, . . . , ℓn)
P = P (ℓ1, . . . , ℓn) := {ξ ∈ Rn : 0 ≤ ξ1/dj
j ≤ ℓj for any j = 1, . . . , n}
is the anisotropic box with axis that are parallel to the coordinate ones (e1, . . . , en). By a
change of variables we get
1
L n(P ) P eLDap
X u(p)dL n =
1
L n(P )
lim
ε→0
1
εQ+1 δεP
(u ◦ Fp − u⋆(p)) dL n.
42
DON AND VITTONE
From this we deduce that, for any n-tuple (ℓ1, . . . , ℓn) the function
is Borel. Now, for every i = 1, . . . , m and every h ∈ N \ {0} define the rectangles P i
P (1/h, . . . , 1/h, 1, 1/h, . . . , 1/h). A simple computation shows that
h :=
X u(p)dL n
(50)
p 7−→
1
L n(P ) P eLDap
h) P i
X u(p)dL n =
heLDap
lim
h
which completes the proof.
1
L n(P i
1
2
(Dap
X u(p))i ,
(cid:3)
Proof of Proposition 2.33. We can assume without loss of generality that k = 1.
(i) By Remark 2.20, the functionseur := u ◦ Fp ◦ δr andevr := u ◦ Fp ◦ δr converge, respectively,
loc(Rn) as r → 0. In particular, as r → 0 the families (eur) and (evr)
to u⋆(p) and v⋆(p) in L1
converge (locally) in measure to u⋆(p) and v⋆(p) respectively. By a change of variables we have
for any R > 0
It follows that (eur) and (evr) have the same measure limit, hence u⋆(p) = v⋆(p).
(ii) Using Proposition 2.26 and the same argument used in (i) we obtain that the functions
r→0
lim
r→0
L n(bB(0, R) ∩ {evr 6=eur}) = lim
if eLνu(p)(y) > 0
if eLνu(p)(y) < 0
u−(p)
U(y) :=(u+(p)
r−QL n(bB(0, rR) ∩ {u ◦ Fp 6= v ◦ Fp}) = 0.
if eLνv(p)(y) > 0
if eLνv(p)(y) < 0
V (y) :=(v+(p)
v−(p)
and
coincide for L n-almost every y, hence (u+(p), u−(p), νu(p)) ≡ (v+(p), v−(p), νv(p)).
(iii) By point (i) we already know that u⋆(p) = v⋆(p). Since
u(Fp(δr(y))) − u⋆(p)
r
6=
v(Fp(δr(y))) − v⋆(p)
r
⇐⇒ u(Fp(δr(y))) 6= v(Fp(δr(y))),
the statement follows using Proposition 2.31 and an argument similar to part (i) above.
(cid:3)
References
[1] Ambrosio, L. Some fine properties of sets of finite perimeter in Ahlfors regular metric measure spaces.
Adv. Math. 159, 1 (2001), 51 -- 67.
[2] Ambrosio, L., Fusco, N., and Pallara, D. Functions of bounded variation and free discontinuity
problems. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York,
2000.
[3] Ambrosio, L., Ghezzi, R., and Magnani, V. BV functions and sets of finite perimeter in sub-
Riemannian manifolds. Ann. Inst. H. Poincaré Anal. Non Linéaire 32, 3 (2015), 489 -- 517.
[4] Ambrosio, L., Kleiner, B., and Le Donne, E. Rectifiability of sets of finite perimeter in Carnot
groups: existence of a tangent hyperplane. J. Geom. Anal. 19, 3 (2009), 509 -- 540.
[5] Ambrosio, L., and Magnani, V. Weak differentiability of BV functions on stratified groups. Math. Z.
245, 1 (2003), 123 -- 153.
[6] Ambrosio, L., and Scienza, M. Locality of the perimeter in Carnot groups and chain rule. Ann. Mat.
Pura Appl. (4) 189, 4 (2010), 661 -- 678.
[7] Bellaïche, A. The tangent space in sub-Riemannian geometry. In Sub-Riemannian geometry, vol. 144 of
Progr. Math. Birkhäuser, Basel, 1996, pp. 1 -- 78.
[8] Biroli, M., and Mosco, U. Sobolev and isoperimetric inequalities for Dirichlet forms on homogeneous
spaces. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei (9) Mat. Appl. 6, 1 (1995), 37 -- 44.
[9] Bramanti, M., Miranda, Jr., M., and Pallara, D. Two characterization of BV functions on Carnot
groups via the heat semigroup. Int. Math. Res. Not. IMRN, 17 (2012), 3845 -- 3876.
[10] Capogna, L., Danielli, D., and Garofalo, N. The geometric Sobolev embedding for vector fields and
the isoperimetric inequality. Comm. Anal. Geom. 2, 2 (1994), 203 -- 215.
FINE PROPERTIES OF BV FUNCTIONS IN CC SPACES
43
[11] Comi, G. E., and Magnani, V. The Gauss-Green theorem in stratified groups. Preprint 2018,
arXiv:1806.04011.
[12] Danielli, D., Garofalo, N., and Nhieu, D.-M. Trace inequalities for Carnot-Carathéodory spaces
and applications. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 27, 2 (1998), 195 -- 252 (1999).
[13] De Giorgi, E. Nuovi teoremi relativi alle misure (r − 1)-dimensionali in uno spazio ad r dimensioni.
Ricerche Mat. 4 (1955), 95 -- 113.
[14] Don, S., Le Donne, E., Moisala, T., and Vittone, D. In preparation.
[15] Don, S., Massaccesi, A., and Vittone, D. Rank-one theorem and subgraphs of BV functions in Carnot
groups. Preprint 2017, arXiv:1712.02242.
[16] Don, S., and Vittone, D. A compactness result for BV functions in metric spaces. Ann. Acad. Sci.
Fenn. Math., to appear. arXiv:1803.07545.
[17] Federer, H. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153.
Springer-Verlag New York Inc., New York, 1969.
[18] Folland, G. B., and Stein, E. M. Hardy spaces on homogeneous groups, vol. 28 of Mathematical Notes.
Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1982.
[19] Franchi, B., Gallot, S., and Wheeden, R. L. Sobolev and isoperimetric inequalities for degenerate
metrics. Math. Ann. 300, 4 (1994), 557 -- 571.
[20] Franchi, B., Serapioni, R., and Serra Cassano, F. Meyers-Serrin type theorems and relaxation of
variational integrals depending on vector fields. Houston J. Math. 22, 4 (1996), 859 -- 890.
[21] Franchi, B., Serapioni, R., and Serra Cassano, F. Approximation and imbedding theorems for
weighted Sobolev spaces associated with Lipschitz continuous vector fields. Boll. Un. Mat. Ital. B (7) 11,
1 (1997), 83 -- 117.
[22] Franchi, B., Serapioni, R., and Serra Cassano, F. Rectifiability and perimeter in the Heisenberg
group. Math. Ann. 321, 3 (2001), 479 -- 531.
[23] Franchi, B., Serapioni, R., and Serra Cassano, F. On the structure of finite perimeter sets in step
2 Carnot groups. J. Geom. Anal. 13, 3 (2003), 421 -- 466.
[24] Garofalo, N., and Nhieu, D.-M. Isoperimetric and Sobolev inequalities for Carnot-Carathéodory spaces
and the existence of minimal surfaces. Comm. Pure Appl. Math. 49, 10 (1996), 1081 -- 1144.
[25] Garofalo, N., and Nhieu, D.-M. Lipschitz continuity, global smooth approximations and extension
theorems for Sobolev functions in Carnot-Carathéodory spaces. J. Anal. Math. 74 (1998), 67 -- 97.
[26] Gromov, M. Carnot-Carathéodory spaces seen from within. In Sub-Riemannian geometry, vol. 144 of
Progr. Math. Birkhäuser, Basel, 1996, pp. 79 -- 323.
[27] Hajłasz, P., and Koskela, P. Sobolev met Poincaré. Mem. Amer. Math. Soc. 145, 688 (2000), x+101.
[28] Heinonen, J. Lectures on analysis on metric spaces. Universitext. Springer-Verlag, New York, 2001.
[29] Jerison, D. The Poincaré inequality for vector fields satisfying Hörmander's condition. Duke Math. J. 53,
2 (1986), 503 -- 523.
[30] Le Donne, E. Lecture notes on sub-Riemannian geometry. sites.google.com/site/enricoledonne/, 2017.
[31] Magnani, V. Elements of geometric measure theory on sub-Riemannian groups. Scuola Normale Superiore,
Pisa, 2002.
[32] Magnani, V. Characteristic points, rectifiability and perimeter measure on stratified groups. J. Eur. Math.
Soc. (JEMS) 8, 4 (2006), 585 -- 609.
[33] Magnani, V., Tyson, J. T., and Vittone, D. On transversal submanifolds and their measure. J. Anal.
Math. 125 (2015), 319 -- 351.
[34] Magnani, V., and Vittone, D. An intrinsic measure for submanifolds in stratified groups. J. Reine
Angew. Math. 619 (2008), 203 -- 232.
[35] Marchi, M. Regularity of sets with constant intrinsic normal in a class of Carnot groups. Ann. Inst.
Fourier (Grenoble) 64, 2 (2014), 429 -- 455.
[36] Mitchell, J. On Carnot-Carathéodory metrics. J. Differential Geom. 21, 1 (1985), 35 -- 45.
[37] Montgomery, R. A tour of subriemannian geometries, their geodesics and applications, vol. 91 of Math-
ematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2002.
[38] Monti, R. Distances, boundaries and surface measures in Carnot-Carathéodory spaces, 2001. PhD thesis,
cvgmt.sns.it/paper/3706/.
[39] Monti, R., Pigati, A., and Vittone, D. On tangent cones to length minimizers in Carnot-Carathéodory
spaces. SIAM J. Control Optim., to appear. cvgmt.sns.it/paper/3529/.
[40] Monti, R., and Serra Cassano, F. Surface measures in Carnot-Carathéodory spaces. Calc. Var. Partial
Differential Equations 13, 3 (2001), 339 -- 376.
44
DON AND VITTONE
[41] Nagel, A., Stein, E. M., and Wainger, S. Balls and metrics defined by vector fields. I. Basic properties.
Acta Math. 155, 1-2 (1985), 103 -- 147.
[42] Rothschild, L. P., and Stein, E. M. Hypoelliptic differential operators and nilpotent groups. Acta
Math. 137, 3-4 (1976), 247 -- 320.
[43] Simon, L. Lectures on geometric measure theory, vol. 3 of Proceedings of the Centre for Mathematical
Analysis, Australian National University. Australian National University, Centre for Mathematical Analy-
sis, Canberra, 1983.
[44] Song, Y. Q., and Yang, X. P. BV functions in the Heisenberg group H n. Chinese Ann. Math. Ser. A
24, 5 (2003), 541 -- 554.
[45] Vittone, D. Regolarità delle geodetiche nei gruppi di Carnot. Master thesis, 2003. Available at
cvgmt.sns.it/paper/972/.
[46] Vittone, D. Lipschitz surfaces, perimeter and trace theorems for BV functions in Carnot-Carathéodory
spaces. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 11, 4 (2012), 939 -- 998.
(Don and Vittone) Università di Padova, Dipartimento di Matematica "T. Levi-Civita", via Tri-
este 63, 35121 Padova, Italy
E-mail address: [email protected]
E-mail address: [email protected]
|
1302.0597 | 1 | 1302 | 2013-02-04T06:57:10 | Decompositions of Weighted Conditional Expectation Type Operators | [
"math.FA"
] | In this paper we investigate boundedness, polar decomposition and spectral decomposition of weighted conditional expectation type operators on L^2(\Sigma). | math.FA | math |
DECOMPOSITIONS OF WEIGHTED CONDITIONAL
EXPECTATION TYPE OPERATORS
Y. ESTAREMI
Abstract. In this paper we investigate boundedness, polar decomposition
and spectral decomposition of weighted conditional expectation type operators
on L2(Σ).
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, measurable 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 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, S(f ) ⊆ S(E(f )).
q , where 1
p E(gq)
p + 1
1
1
q = 1 (Holder inequality).
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 a non-atomic measure space. It
is well-known fact that every σ-finite measure space (X, A, µA ) can be partitioned
2000 Mathematics Subject Classification. 47B47.
Key words and phrases. Conditional expectation, Polar decomposition, Spectral decomposi-
tion, Aluthge transformation.
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 [12]).
Combinations of conditional expectation operators and multiplication opera-
tors appear often in the study of other operators such as multiplication operators,
weighted composition operators and integral operators. Specifically, in [9], 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, [6], 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, [3], extended these characterizations to the setting of function ideals
and vector lattices. J. Herron presented some assertions about the operator EMu
on Lp spaces in [7, 8].
In [4, 5] we investigated some classic properties of multiplication conditional ex-
pectation 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 we present some
results on the boundedness, polar decomposition and spectral decomposition of this
operators in L2(Σ), using different methods than those employed in [5].
2. Polar decomposition
Theorem 2.1. The operator T = MwEMu : L2(Σ) → L2(Σ) is bounded if and
only if (E(w2)
1
2 )(E(u2)
1
2 ) ∈ L∞(A) and in this case kT k = k(E(w2)
1
2 )(E(u2)
1
2 )k∞.
Proof Suppose that (E(w2)
1
2 )(E(u2)
1
2 ) ∈ L∞(A). Let f ∈ L2(Σ). Then
kT (f )k2
2 =ZX
wE(uf )2dµ =ZX
E(w2)E(uf )2dµ ≤ZX
Since E(uf ) ≤ (E(u2)
1
2 (E(f 2)
1
2 . Thus
E(w2)E(u2)E(f 2)dµ.
kT k ≤ kE(w2)
1
2 (E(u2))
1
2 k∞.
To prove the converse, let T be bounded on L2(Σ) and consider the case that
µ(X) < ∞. Then for all f ∈ L2(Σ) we have
kT (f )k2
2 =ZX
wE(uf )2dµ =ZX
≤ kT k2ZX
f 2dµ.
E(w2)E(uf )2dµ
For each n ∈ N, define
En = {x ∈ X : u(x)(E(w2))
1
2 (x) ≤ n}.
Each En is Σ-measurable and En ↑ X. Define Gn = En ∩ S for each n ∈ N, where
S = S(u(E(w2))
2 ). Let A ∈ A and define
1
fn = ¯u(E(w2))
1
2 χGn∩A
WEIGHTED CONDITIONAL EXPECTATION TYPE OPERATORS
3
for each n ∈ N. It is clear that fn ∈ L∞(Σ) for all n (which in our case implies
fn ∈ L2(Σ)). For each n,
kT (fn)k2
E(w2)E(ufn)2dµ
2 =ZX
=ZX
=ZA
≤ kT k2ZX
= kT k2ZA
(E(w2))2(E(u2χGn.χA))2dµ
[E(w2)E(u2χGn )]2dµ
fn2dµ
E(u2χGn)E(w2)dµ.
Since A is an arbitrary A-measurable set and the integrands are both A-measurable
functions, we have
[E(w2)E(u2χGn )]2 ≤ kT k2E(u2χGn )E(w2)
almost everywhere. That is
[E(((E(w2))
1
2 uχEn)2χS)]2 ≤ kT k2E((uχEn (E(w2))
1
2 )qχS).
Since
and
we have
Thus
S = σ(u(E(w2))
1
2 ) = σu2E(w2)
u2E(w2)χS = u2E(w2),
E(((E(w2))
1
2 uχEn)2χS) ≤ kT k2.
This implies that (Ew2)
1
(Ew2)
2 (Eu2χEn )
1
2 (Eu2χEn )
1
2 ≤ kT k.
1
2 ∈ L∞(A) and
k(Ew2)
1
2 (Eu2)
1
2 k∞ ≤ kT k.
Moreover, since Em ↑ X, the conditional expectation version of the monotone
convergence theorem implies k(E(w2))
✷
2 (E(u2))
2 k∞ ≤ kT k.
1
1
proposition 2.2. Let g ∈ L∞(A) and let T = MwEMu : L2(Σ) → L2(Σ) be
bounded. If MgT = 0, then g = 0 on σ(E(w2)E(u2)).
Proof. Let f ∈ L2(Σ). Then gwE(uf ) = MgT (f ) = 0. Now, by Theorem 2.1
0 = kMgT k2 = kg2E(w2)E(u2)k∞,
which implies that g2E(w2)E(u2) = 0, and so g = 0 on σ(E(w2)E(u2)). ✷
Theorem 2.3. The bounded operator T = MwEMu is a partial isometry if and
only if E(w2)E(u2) = χA for some A ∈ A.
Proof. Suppose T is partial isometry. Then T T ∗T = T , that is T f = E(w2)E(u2)T f ,
and hence (E(w2)E(u2) − 1)T f = 0 for all f ∈ L2(Σ). Put S = S(E(u2)) and
4
Y. ESTAREMI
G = S(E(w2)). By Proposition 2.2. we get that E(w2)E(u2) = 1 on S ∩ G,
which implies that E(w2)E(u2) = χA, where A = S ∩ G.
Conversely, suppose that E(w2)E(u2) = χA for some A ∈ A. It follows that
A = S ∩ G, and we have
T T ∗T (f ) = E(w2)E(u2)T f = χS∩GwE(uf ) = wE(uf ),
where we have used the fact that S(T f ) = S(T f 2) ⊆ S ∩ G, which this is a
consequence of Holder's inequality for conditional expectation E.
✷
The spectrum of an operator A is the set
σ(A) = {λ ∈ C : A − λI
is not invertible}.
It is well known that any bounded operator A on a Hilbert space H can be expressed
in terms of its polar decomposition: A = U P , where U is a partial isometry and P
is a positive operator. (An operator is positive if hP f, f i ≥ 0, for all f ∈ H.) This
representation is unique under the condition that kerU = kerP = kerA. Moreover,
P = A = (A∗A)
1
2 .
Let q(z) be a polynomial with complex coefficients: q(z) = PN
α0I +PN
If
T is a bounded operator on L2(Σ), then the operator q(T ) is defined by q(z) =
n=1 αnT n. Let Mϕ be a bounded multiplication operator on L2(Σ), then
q(Mϕ) is also bounded and q(Mϕ) = Mq◦ϕ. By the continuous functional calculus,
for any f ∈ C(σ(Mϕ)), we have g(Mϕ) = Mg◦ϕ.
n=0 αnzn.
Proposition 2.4. Let S = S(E(u2)) and G = S(E(w2)). If f ∈ C(σ(ME(u2)))
and g ∈ C(σ(ME(w2))), Then
and
f (T ∗T ) = f (0)I + M(E(u2))−1.χS(cid:0)Mf ◦(E(u2)E(w2)) − f (0)I(cid:1) M¯uEMu
g(T T ∗) = g(0)I + M(E(w2))−1.χG(cid:0)Mg◦(E(u2)E(w2)) − g(0)I(cid:1) MwEM ¯w.
Proof. For all f ∈ L2(Σ), T ∗T (f ) = ¯uE(w2)E(uf ) and T T ∗(f ) = wE(u2)E( ¯wf ).
By induction, for each n ∈ N,
(T ∗T )n(f ) = ¯u(E(w2))n(E(u2))n−1E(uf ),
(T T ∗)n(f ) = w(E(u2))n(E(w2))n−1E( ¯wf ).
q(T ∗T ) = q(0)I + M(E(u2))−1.χS(cid:0)Mq◦(E(u2)E(w2)) − q(0)I(cid:1) M¯uEMu
q(T T ∗) = q(0)I + M(E(w2))−1.χG(cid:0)Mq◦(E(u2)E(w2)) − q(0)I(cid:1) MwEM ¯w.
By the Weierstrass approximation theorem we conclude that, for every f ∈ C(σ(ME(u2)))
and g ∈ C(σ(ME(w2))),
So
and
and
f (T ∗T ) = f (0)I + M(E(u2))−1.χS(cid:0)Mf ◦(E(u2)E(w2)) − f (0)I(cid:1) M¯uEMu
g(T T ∗) = g(0)I + M(E(w2))−1.χG(cid:0)Mg◦(E(u2)E(w2)) − g(0)I(cid:1) MwEM ¯w.
Theorem 2.5. The unique polar decomposition of T = MwEMu is U T , where
WEIGHTED CONDITIONAL EXPECTATION TYPE OPERATORS
5
E(u2)(cid:19) 1
T (f ) =(cid:18) E(w2)
2
χS ¯uE(uf ),
U (f ) =(cid:18)
for all f ∈ L2(Σ).
Proof. By Proposition 2.4 we have
E(w2)E(u2)(cid:19) 1
χS∩G
2
wE(uf ),
T (f ) = (T ∗T )
1
2 (f ) =(cid:18) E(w2)
E(u2)(cid:19) 1
2
χS ¯uE(uf ).
Define a linear operator U whose action is given by
U (f ) =(cid:18)
E(w2)E(u2)(cid:19) 1
χS∩G
2
wE(uf ),
f ∈ L2(Σ).
Then T = U T and by Theorem 2.3, U is a partial isometry. Also, it is easy
to see that N (T ) = N (U ). Since for all f ∈ L2(Σ), kT f k2 = k T f k2, hence
N (T ) = N (U ) and so this decomposition is unique.
✷
Theorem 2.6. The Aluthge transformation of T = MwEMu is
χSE(uw)
E(u2)
¯uE(uf ),
bT (f ) =
f ∈ L2(Σ).
Proof. Define operator V on L2(Σ) as
V f =(cid:18) E(w2)
(E(u2))3(cid:19) 1
4
χS ¯uE(uf ),
f ∈ L2(Σ).
Then we have V 2 = T and so by direct computation we obtain
bT (f ) = T
1
2 U T
1
2 (f ) =
χSE(uw)
E(u2)
¯uE(uf ).
✷
3. Spectral decomposition
The normal operators form one of the best understood and most tractable of
classes of operators. The principal reason for this is the spectral theorem, a powerful
structure theorem that answers many (not all) questions about these operators. In
this section we explore spectral measure and spectral decomposition corresponding
to a normal weighted conditional expectation operator EMu on L2(Σ).
Definition 3.1. If X is a set, Σ a σ-algebra of subsets of X and H a Hilbert
space, a spectral measure for (X, Σ, H) is a function E : Σ → B(H) having the
following properties.
(a) E(S) is a projection.
(b) E(∅) = 0 and E(X) = I.
(c) If S1, S2 ∈ Σ. E(S1 ∩ S2) = E(S1)E(S2).
(d) If {Sn}∞
n=0 is a sequence of pairwise disjoint sets in Σ, then
E(∪∞
n=0Sn) = Σ∞
n=0E(Sn).
6
Y. ESTAREMI
The spectral theorem says that: For every normal operator T on a Hilbert
space H, there is a unique spectral measure E relative to (σ(T ), H) such that
T =Rσ(T ) zdE, where z is the inclusion map of σ(T ) in C.
J. Herron showed that σ(EMu) = ess range(Eu) ∪ {0}, [8]. Also, He has proved
If T = EMu is normal, then
that: EMu is normal if and only if u ∈ L∞(A).
T n = MunE and (T ∗)n = M¯un E. So (T ∗)nT m = M(¯u)num E and
αn,mT m(T ∗)n =
N,MXn,m=0
αn,m ¯unumE = P (u, ¯u)E = EP (u, ¯u).
P (T, T ∗) =
N,MXn,m=0
Where p(z, t) =PN,M
n,m=0 αn,mzmtn. If q(z) =PN
n=0 αnzn, then q(T ) =PN
Hence by the Weierstrass approximation theorem we have f (T ) = Mf (u)E, for all
f ∈ C(σ(EMu)). Thus φ : C(σ(EMu)) → C∗(EMu, I), by φ(f ) = Mf (u)E, is a
unital ∗-homomorphism. Moreover, by Theorem 2.1.13 of [10], φ is also a uniqe
∗-isomorphism such that φ(z) = EMu, where z : σ(EMu) → C is the inclusion map.
n=0 αnunE.
If EMu is normal and compact, then σ(EMu) = {0} ∪ {λn}n∈N where λn 6= 0
for all n ∈ N. So, for each n ∈ N
En = {0 6= f ∈ L2(Σ) : E(uf ) = λnf }
= {0 6= f ∈ L2(A) : uf = λnf }
= L2(An, An, µn),
where An = {x ∈ X : u(x) = λn}, E0 = {f ∈ L2(Σ) : E(uf ) = 0}, An = {An ∩ B :
B ∈ A} and µn ≡ µ An . It is clear that for all n, m ∈ N ∪ {0}, En ∩ Em = ∅. This
implies that the spectral decomposition of EMu is as follows:
EMu =
∞Xn=0
λnPEn ,
where PEn is the orthogonal projection onto En. Since EMu is normal, then
σ(EMu) = ess range(u) ∪ {0}. So {λn}∞
n=0 is a resolution of the identity on
X.
Suppose that W = {u ∈ L0(Σ) : E(u2) ∈ L∞(A)}. If we set kuk = k(E(u2))
then W is a complete ∗-subalgebra of L∞(Σ).
1
2 k∞,
In the sequel we assume that, ϕ : X → X is nonsingular transformation i.e,
the measure µ ◦ ϕ−1 is absolutely continuous with respect to the measure µ, and
ϕ−1(Σ) is a sub-σ-finite algebra of Σ. Put h = dµ ◦ ϕ−1/dµ and Eϕ = Eϕ−1(Σ).
For S ∈ Σ, let E(S) : L2(Σ) → L2(Σ) be defined by
E(S)(f ) = EϕMχϕ−1 (S)
(f ),
i.e,
E(S) = EϕMχϕ−1 (S)
.
E defines a spectral measure for (X, Σ, L2(µ)). If EϕMu is normal on L2(Σ), then
by Theorem 2.5.5 of [10], E is the unique spectral measure corresponding to ∗-
homomorphism φ that is defined as follows:
φ : C(σ(EϕMu)) → C∗(EMu, I),
φ(f ) = EϕMf (u).
WEIGHTED CONDITIONAL EXPECTATION TYPE OPERATORS
7
So, for all f ∈ C(σ(EϕMu)) we have
φ(f ) =ZX
f dE.
In [1] it is explored that which sub-σ-algebras of Σ are of the form ϕ−1(Σ) for some
nonsingular transformation ϕ : X → X. These observations establish the following
theorem.
Theorem 3.2. Let (X, Σ, µ) be a σ-finite measure space, ϕ : X → X be a
nonsingular transformation and let u be in L∞(ϕ−1(Σ)). Consider the operator
EϕMu on L2(Σ). Then the set function E that is defined as: E(S) = EϕMχϕ−1(S)
for S ∈ Σ, is a spectral measure. Also, E has compact support and
EϕMu =Z zdE.
References
[1] J. T. Campbell, A. Lambert and B. M. Weinstock, Localizing Sets and the Structure of
Sigma-Algebras, Indiana University Mathematics Journal 47 (1998) 913-938.
[2] P.G. Dodds, C.B. Huijsmans and B. De Pagter, characterizations of conditional
expectation-type operators, Pacific J. Math 141 (1990) 55-77.
[3] P.G. Dodds, C.B. Huijsmans and B. De Pagter, characterizations of conditional expectation-
type operators, Pacific J. Math. 141(1) (1990), 55-77.
[4] Y. Estaremi, Essential norm of weighted conditional type operators on Lp-spaces, to appear
in positivity.
[5] Y. Estaremi and M.R. Jabbarzadeh, Weighted lambert type operators on Lp-spaces, Oper.
Matrices 1 (2013), 101-116..
[6] R. G. Douglas, Contractive projections on an L1 space, Pacific J. Math. 15 (1965), 443-462.
[7] J. Herron, Weighted conditional expectation operators on Lp spaces, UNC Charlotte Doctoral
Dissertation, 2004.
[8] J. Herron, Weighted conditional expectation operators, Oper. Matrices 1 (2011), 107-118.
[9] Shu-Teh Chen, Moy, Characterizations of conditional expectation as a transformation on
function spaces, Pacific J. Math. 4 (1954), 47-63
[10] G. J. Murphy, C ∗-algebras and operator theory, Academic Press, Boston San Diego, 1990.
[11] M. M. Rao, Conditional measure and applications, Marcel Dekker, New York, 1993.
[12] A. C. Zaanen, Integration, 2nd ed., North-Holland, Amsterdam, 1967.
Y. Estaremi
E-mail address: [email protected] - [email protected]
Department of Mathematics, University of Payame noor, p. o. box: 19395-3697,
Tehran, Iran.
|
1207.3818 | 2 | 1207 | 2012-10-30T13:11:40 | Large structures made of nowhere $L^p$ functions | [
"math.FA"
] | We say that a real-valued function $f$ defined on a positive Borel measure space $(X,\mu)$ is nowhere $q$-integrable if, for each nonvoid open subset $U$ of $X$, the restriction $f|_U$ is not in $L^q(U)$. When $(X,\mu)$ satisfies some natural properties, we show that certain sets of functions defined in $X$ which are $p$-integrable for some $p$'s but nowhere $q$-integrable for some other $q$'s ($0<p,q<\infty$) admit a variety of large linear and algebraic structures within them. The presented results answer a question from Bernal-Gonz\'alez, improve and complement recent spaceability and algebrability results from several authors and motivates new research directions in the field of spaceability. | math.FA | math | LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
SZYMON G LA¸ B, PEDRO L. KAUFMANN, AND LEONARDO PELLEGRINI
Abstract. We say that a real-valued function f defined on a positive Borel measure space
(X, µ) is nowhere q-integrable if, for each nonvoid open subset U of X, the restriction f U
is not in Lq(U ). When (X, µ) satisfies some natural properties, we show that certain sets
of functions defined in X which are p-integrable for some p's but nowhere q-integrable
for some other q's (0 < p, q < ∞) admit a variety of large linear and algebraic structures
within them. The presented results answer a question from Bernal-Gonz´alez, improve
and complement recent spaceability and algebrability results from several authors and
motivates new research directions in the field of spaceability.
2
1
0
2
t
c
O
0
3
]
.
A
F
h
t
a
m
[
2
v
8
1
8
3
.
7
0
2
1
:
v
i
X
r
a
1. Introduction
This work is a contribution to the study of large linear and algebraic structures within
essentially nonlinear sets of functions which satisfy special properties; the presence of such
structures is often described using the terminology lineable, algebrable and spaceable. Recall
that a subset S of a topological vector space V is said to be lineable (respectively, spaceable)
if S ∪ {0} contains an infinite dimensional vector subspace (respectively, a closed infinite
dimensional vector subspace) of V . Though results in this field date back to the sixties1,
this terminology was not introduced until recently: it first appeared in unpublished notes
by Enflo and Gurariy and was firstly published in [1]. We should mention that Enflo's and
Gurariy's unpublished notes were completed in collaboration with Seoane-Sep´ulveda and
will finally be published in [10]. It is current to say also that S is dense-lineable if S ∪ {0}
contains an dense infinite dimensional vector subspace of V . The adjective maximal is often
added to dense-lineable or spaceable when the corresponding space contained in S ∪ {0}
has the same dimension of V . We propose in Section 3 a notion of spaceability which is
more restrictive than the "maximal" spaceability in terms of dimension. For this reason,
we choose to use the notation maximal-dimension spaceable, maximal-dimension lineable
and so on when the maximality concerns dimension of the subspace found in S ∪ {0}.
The term algebrability was introduced later in [2]; if V is a linear algebra, S is said to
be κ-algebrable if S ∪ {0} contains an infinitely generated algebra, with a minimal set of
2010 Mathematics Subject Classification. 46G12, 15A03.
Key words and phrases. nowhere Lq functions, spaceability, algebrability.
The first named author was supported by the Polish Ministry of Science and Higher Education Grant
No. N N201 414939 (2010-2013).
The second named author was supported by CAPES, Research Grant PNPD 2256-2009, and by the
Institut de Math´ematiques de Jussieu.
1In [14], Gurariy showed that there exists in C([0, 1]) a closed infinite-dimensional subspace consisting,
except for the null function, only on nowhere differentiable functions - see also [15] for a version in english.
1
2
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
generators of cardinality κ (see [2] for details). We shall work with a strenghtened notion
of κ-algebrability, namely, strong κ-algebrability. The definition follows:
Definition 1.1. We say that a subset S of an algebra A is strongly κ-algebrable, where
κ is a cardinal number, if there exists a κ-generated free algebra B contained in S ∪ {0}.
We recall that, for a cardinal number κ, to say that an algebra A is a κ-generated free
algebra, means that there exists a subset Z = {zα : α < κ} ⊂ A such that any function f
from Z into some algebra A′ can be uniquely extended to a homomorphism from A into A′.
The set Z is called a set of free generators of the algebra A. If Z is a set of free generators
of some subalgebra B ⊂ A, we say that Z is a set of free generators in the subalgebra A.
If A is commutative, a subset Z = {zα : α < κ} ⊂ A is a set of free generators in A if for
each polynomial P and for any zα1, zα2, . . . , zαn ∈ Z we have
P (zα1, zα2, . . . , zαn) = 0 if and only if P = 0.
The definition of strong κ-algebrability was introduced in [3], though in several papers, sets
which are shown to be algebrable are in fact strongly algebrable, and that is seen clearly by
the proofs. See [2], [4] and [12], among others. Strong algebrability is in effect a stronger
condition than algebrability: for example, c00 is ω-algebrable in c0 but it is not strongly
1-algebrable (see [3]).
1.1. Results on large structures of non-integrable functions: recent and new.
Our object of study will be the quasi-Banach spaces Lp(X, M, µ). For a clear notation,
when there cannot be any confusion or ambiguity, we shall write Lp, Lp(X, µ) or Lp(X)
instead of Lp(X, M, µ). Our main focus will be on functions which are p-integrable but
not q-integrable, for some 0 < p, q ≤ ∞, and specially on functions which are p-integrable
but nowhere q-integrable. The notion of nowhere-q-integrability we consider is connected
to open sets:
Definition 1.2. Let 0 < q ≤ ∞. A scalar-valued function f defined on a Borel measure
space X is said to be nowhere q-integrable (or nowhere Lq) if, for each nonvoid open subset
U of X, the restriction f U is not in Lq(U).
In our context it would be pointless to substitute "for each nonvoid open subset U of X"
by "for each Borel subset U of positive measure of X" in the definition above; the reason
is that, if 0 < p, q ≤ ∞ and f ∈ Lp(X), there is always a Borel subset of X with positive
measure and on which f is q-integrable. This follows from a simple argument (see e.g. the
final remarks in [5]). Of course, not all Borel measure spaces (X, µ) admit Lp-nowhere-Lq
functions, but there is a large class of such spaces which admit plenty of such functions,
as we will see.
Let us start by mentioning some recent results and open questions on large structures
within sets of functions which are p-integrable but not q-integrable. For a survey on the
evolution of the results in this direction, we recommend [7].
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
3
Theorem 1.3 (Bernal-Gonz´alez, Ord´onez Cabrera [6]). Let (X, M, µ) be a measure space,
and consider the conditions
(α) inf{µ(A) : A ∈ M, µ(A) > 0} = 0, and
(β) sup{µ(A) : A ∈ M, µ(A) < ∞} = ∞.
Then the following assertions hold:
(1) if 1 ≤ p < ∞, then Lp \ ∪q>pLq is spaceable if and only if (α) holds;
(2) if 1 < p ≤ ∞, then Lp \ ∪q<pLq is spaceable if and only if (β) holds;
(3) if 1 < p < ∞, then Lp \ ∪q6=pLq is spaceable if and only if both (α) and (β) hold;
(4) if 1 < p < ∞ and Lp is separable, then Lp \ ∪q<pLq is maximal-dimension dense-
lineable if and only if (β) holds.
Note that any of the conditions (α), or (β), or (α) and (β) is enough to guarantee the
existence of nowhere q-integrable functions in Lp (just note that, if X contains an open
singleton {x} of positive measure, then each function from Lp(x) is q-integrable in {x},
for all q). Bernal-Gonz´alez et. al. use the convenient terminology '(left, right) strict order
integrability' when a function is p-integrable but not q-integrable for q 6= p (q < p, q > p).
We refer to [7] for improvements on item (2) of Theorem (1.3) above. And in [9] there is
a version of that same item which includes quasi-Banach spaces:
Theorem 1.4 (Botelho, F´avaro, Pellegrino, Seoane-Sep´ulveda [9]). Lp[0, 1] \ ∪q>pLq[0, 1]
is spaceable for every p > 0.
When it comes to nowhere integrable functions, Bernal-Gonz´alez gave the first initial
result:
Theorem 1.5 (Bernal-Gonz´alez [5]). Let (X, M, µ) be a measure space such that X is a
Hausdorff first-countable separable locally compact perfect topological space and that µ is
a positive Borel measure which is continuous, regular and has full support. Let 1 ≤ p <
∞.Then the set
{f ∈ Lp : f is nowhere q-integrable, for each q > p}
(1.1)
is dense in Lp.
It is clear that µ having full support (that is, µ(U) > 0 for every nonvoid open subset
U ⊂ X) is a necessary condition for the existence of nowhere q-integrable functions. Based
on the above result, Bernal-Gonz´alez rose the following question:
Problem 1. Is the set (1.1) lineable/maximal-dimension lineable/dense-lineable?
It is quite natural to seek for other large structures within (1.1).
The authors of this work have also presented some results on large structures of nowhere
integrable functions, and among them we mention the following:
Theorem 1.6 (G l¸ab, Kaufmann, Pellegrini [11]). The set of nowhere essentially bounded
functions in L1[0, 1] is
4
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
(1) spaceable and
(2) strongly c-algebrable.
In this landscape, we present a few new results which solve Problem 1 and, under
quite mild conditions on the measure space where our functions are defined, comple-
ment/generalize the results mentioned above. We summarize these results in Theorem
1.7 below.
Theorem 1.7. Suppose that X is a topological space admitting a countable π-base (that
is, a family (Un)n of nonvoid open subsets of X such that, for each nonvoid open subset A
of X, Uj ⊂ A for some j) and that µ a positive Borel measure on X. Let 0 < p < ∞ and
consider the sets
Sp(X)
.
= Sp
.
= {f ∈ Lp : f is nowhere Lq, for each p < q ≤ ∞},
.
= Sp \ ∪0<q<pLq, and
S′
p
G
.
=(f ∈ \0<q<∞
Lq : f is nowhere L∞) .
Then we have the following:
(a) if µ is atomless, outer regular and has full support, then Sp∪{0} contains a ℓp-isometric
subspace of Lp, which is in addition complemented if p ≥ 1;
(b) if µ infinite and σ-finite, then Lp \ ∪0<q<pLq contains a ℓp-isometric subspace of Lp,
which is in addition complemented if p ≥ 1;
(c) if µ is atomless, infinite, outer regular and has full support, then S′
p ∪ {0} contains a
ℓp-isometric subspace of Lp, which is in addition complemented if p ≥ 1;
(d) if µ is atomless, outer regular and has full support, then Sp is maximal-dimension
dense-lineable;
(e) if µ is atomless, outer regular and has full support, then G is strongly c-algebrable.
See Section 6 for comments on the choice of working with π-bases instead of the more
usual bases of open sets. In addition to Theorem 1.7 we also prove that, for a special classes
of positive Borel measure spaces, Sp contains an isomorphic copy of ℓ2 (see Theorems 3.5
and 3.6, and Corollary 3.7). This motivates a new investigation direction concerning space-
ability (see Section 3).
Remark 1. Referring to items (a) -- (c), it is worth recalling that for p < 1, Lp contains
no complemented copy of ℓp. This is easily seen if one recalls that, for p < 1, ℓp admits
nontrivial continuous linear forms (e.g. the evaluation functionals), while every nontrivial
linear form on Lp is discontinuous.
Remark 2.
In any measurable space which admits a set of strictly positive finite
measure (in particular for (X, µ) under the conditions in (e)) and 0 < p < q < ∞, the
set of Lp functions which are not Lq is not algebrable; to see this, just note that if f is p-
integrable but not q-integrable on some set of finite measure U, then f n is not p-integrable
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
5
if we choose a large enough power n. There is therefore no hope in looking for algebraic
structures of strict-order integrable functions in many cases. One exception is given by:
Theorem 1.8 (Garc´ıa-Pacheco, P´erez-Eslava, Seoane-Sep´ulveda [13]). If (X, M, µ) is a
measure space in which there exists and infinite family of pairwise disjoint measurable sets
An satisfying µ(An) ≥ ǫ for some ǫ > 0, then
is spaceable in L∞ and algebrable.
L∞ \ ∩∞
p=1Lp
Note that Theorem 1.7(e) complements, in some sense, the algebrability part of Theorem
1.8.
Remark 3. Theorem 1.7 relates to what was mentioned previously in the following
way:
• (a) generalizes Theorem 1.4, Theorem 1.6(1) and, under our assumptions, also
Theorem 1.3(1);
• it is not hard to adapt Theorem 1.3(2) for p < 1 and to see that the space guar-
anteeing the spaceability can be isometric to ℓp and complemented in case p ≥ 1;
since condition (β) from Theorem 1.3 is milder that the conditions in (b), it follows
that (b) does not really add much. But the construction in the proof we present is
used to prove also (c), thus we include (b) for completeness and clearness;
• under our assumptions, (c) improves Theorem 1.3(3);
• (d) improves Theorem 1.5 and gives a positive answer to Bernal-Gonz´alez's Problem
1;
• (e) improves Theorem 1.6(2).
The remaining sections will be organized as follows. In Section 2 we will prove The-
orem 1.7(a) -- (c), that is, its spaceability part. In Section 3, we introduce the notion of
V -spaceability (see Definition 3.1) and provide the results on ℓ2-spaceability. Section 4
contains the proof of the dense-lineability result (Theorem 1.7(d)), and Section 5 is on the
algebrability result (Theorem 1.7(e)). In Section 6 we briefly discuss conditions on positive
Borel measure spaces under which there exist, or not, functions p-nowhere-q integrable in
the corresponding Lp spaces. We include related open problems throughout the text.
2. Spaceability: proof of Theorem 1.7(a) -- (c)
Recall the following standard result from functional analysis on Banach spaces:
Theorem 2.1. Suppose that (X, µ) is a Borel measure space. Let 1 ≤ p < ∞, and suppose
that (fn) is a sequence of norm-one, disjointly supported functions in Lp(µ). Then (fn) is
a complemented basic sequence isometrically equivalent to the canonical basis of ℓp.
It is not hard to see that the same holds for 0 < p < 1, though the complementability
is lost, as we previously pointed out. Our strategy to prove Theorem 1.7(a) -- (c) will be
6
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
to find sequences of norm-one, disjointly supported functions in Sp, Lp \ ∪0<q<pLq and S′
p,
under the corresponding assumptions.
Lemma 2.2. Let X be a topological space with a countable π-base. Suppose that µ is an
atomless and outer-regular positive Borel measure on X with full support. Let U be an
open set such that µ(U) > 0 and let ε ∈ (0, 1). Then there is a nowhere-dense Borel subset
N of U such that µ(N) > µ(U)ε.
Proof. Let (Un) be a π-base of U. Since µ is atomless, there are Borel sets Bn ⊂ Un
such that µ(Bn) < εµ(U)/2n. Since µ is outer-regular, there are open set V ′
n ⊃ Bn with
µ(V ′
dense open subset of U. Therefore N = U \ V is nowhere dense subset of U with measure
greater than µ(U)ε.
(cid:3)
n ∩ U and put V =Sn Vn. Then µ(V ) < εµ(U) and V is a
n) < εµ(U)/2n. Let Vn = V ′
Lemma 2.3. Suppose that µ is an atomless positive Borel measure on X with full support.
Let A be a measurable set in X such that µ(A) > 0 and let (an) be a sequence in (0, +∞).
Then there is a sequence (An) of pairwise disjoint measurable subsets of A such that 0 <
µ(An) < ∞ and µ(An+1) ≤ anµ(An).
Proof. We may assume that an ≤ 1/2 for all n. Since µ is atomless, there is a Borel set
A1 ⊂ A such that
1
2
Likewise, there is Borel set A2 ⊂ A \ A1 such that
0 < µ(A1) <
µ(A).
0 < µ(A2) < a1µ(A1) ≤
1
2
µ(A1).
1
2
µ(An−1);
0 < µ(An) < an−1µ(An−1) ≤
Proceeding this way, we can find inductively An ⊂ A \Sk<n Ak such that
this is possible since µ(A \Sk<n Ak) > 0.
function hA in Lp \Sq>p Lq.
Lemma 2.4. Suppose that µ is an atomless positive Borel measure on X with full support.
Then for any given Borel set A in X such that µ(A) > 0 there is a norm-one, A-supported
Proof. Let A ⊂ X be measurable and µ(A) > 0; by Lemma 2.3 there exists a family {An,m :
n, m ∈ N} of pairwise disjoint subsets of A of positive measure such that µ(An,m+1) ≤
1
2µ(An,m). Let (rn) be a strictly decreasing sequence of real numbers tending to p. Put
(cid:3)
where arn
n,mµ(An,m) = 1/m. Then hnrn = ∞ and
hn =
an,mχAn,m,
∞
Xm=1
hnpdµ =
ZX
∞
Xm=1
ap
n,mµ(An,m) =
1
arn−p
n,m m
.
∞
Xm=1
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
7
Since
lim sup
m→∞
1
arn−p
n,m+1(m+1)
1
arn−p
n,m m
= lim sup
µ(An,m) (cid:19)
m→∞ (cid:18)µ(An,m+1)
rn−p
rn
2(cid:19)
≤(cid:18) 1
rn−p
rn
< 1,
then by the ratio test for series we obtain that hn ∈ Lp. Put
hA =
hn
hn2n .
∞
Xn=1
(cid:3)
Then hA ∈ Lp \Sq>p Lq and khAk = 1.
Proof of Theorem 1.7(a). Let (Un) be a π-base of X. Since µ is atomless and outer-
regular, we may assume that µ(Un) < ∞ for each n. (Indeed, suppose that µ(Un) = ∞.
Hence Un 6= ∅ and there is x ∈ Un. Since µ is atomless, then µ({x}) = 0. By the outer-
regularity of µ, there is an open neighborhood V of x with arbitrarily small µ-measure.
Since µ does not vanish on open sets, then 0 < µ(V ∩ Un) < ∞. We may replace Un ∩ V
with Un.)
By Lemma 2.2, there is a nowhere dense Borel set N1 ⊂ U1 with 0 < µ(N1) < 1
2. Since
N1 is nowhere dense we can find a nonempty open set U ⊂ U2 \ N1, and again by Lemma
2.2 there is a nowhere dense Borel set N2 ⊂ U ⊂ U2 with 0 < µ(N2) < 1
22 . We can then
inductively define a pairwise disjoint sequence of nowhere dense Borel sets (Nn) such that
Nn ⊂ Un and 0 < µ(Nn) < 1/2n. Decompose each Nn into µ-positive and pairwise disjoint
Borel sets Nn,m. For each n, m there exists, by Lemma 2.4, a norm-one, Nn,m-supported
function hNn,m in Lp \Sq>p Lq. If we put
fm =
hNn,m
2n ,
∞
Xn=1
then (fm) will form a norm-one basic sequence of elements from Sp with pairwise disjoint
supports, and by Theorem 2.1 our proof is concluded.
(cid:3)
Lemma 2.5. Suppose that µ is an infinite and σ-finite positive Borel measure on X. Then
for any given Borel set B ⊂ X of infinite measure, there exists a function gB ∈ Lp \Sq<p Lq
which is zero outside of B.
Proof. Let B ⊂ X be Borel of infinite measure, and let {Bn,m : n, m ∈ N} be a family of
pairwise disjoint subsets of B of positive finite measure such that 2µ(Bn,m) ≤ µ(Bn,m+1).
Let (rn) be a strictly increasing sequence of (strictly positive) real numbers tending to p.
Put
∞
where brn
n,mµ(Bn,m) = 1/m. Then
bn,mχBn,m,
gn =
Xm=1
gnpdµ =
ZX
∞
Xm=1
bp
n,mµ(Bn,m) =
bp−rn
n,m
m
,
∞
Xm=1
8
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
and since
lim sup
m→∞
bp−rn
n,m+1
(m+1)
bp−rn
n,m
m
= lim sup
µ(Bn,m+1)(cid:19)
m→∞ (cid:18) µ(Bn,m)
p−rn
rn
2(cid:19)
≤(cid:18)1
p−rn
rn
< 1,
by ratio test for series we obtain that gn ∈ Lp. Letting
.
=
gB
gn
kgnk2n ,
∞
Xn=1
we have that gB ∈ Lp. It suffices to show now that gB 6∈ Lq for any q < p. Fix such q; for
a large enough n, rn > q, and then
(kgnk2n)qZ gBq ≥Z gnq =Xm (cid:18)
q
rn
1
m.µ(Bn,m)(cid:19)
µ(Bn,m)
q
rn
m(cid:19)
=Xm (cid:18) 1
µ(Bn,m)
rn−q
rn = µ(Bn,1)
q
rn
= ∞.
rn−q
m(cid:19)
rn Xm (cid:18) 1
(cid:3)
Proof of Theorem 1.7(b). Since µ is infinite and σ-finite, then each Borel set of infinite
measure D can be written as an infinite disjoint union of Borel sets of infinite measure.
To see this it is enough to verify that D contains an infinite disjoint union of Borel sets
of infinite measure Dn. In effect, we can define inductively Borel sets Ck ⊂ D such that
1 ≤ µ(Ck) < ∞; let (Mn) be a pairwise disjoint family of infinite subsets of N. Then
Dn
result then follows from Lemma 2.5 and the same argument that was used in the proof of
Theorem 1.7(a).
(cid:3)
= Sk∈Mn Ck is a family of Borel sets satisfying the desired properties. The desired
.
The proof of Theorem 1.7(c) is a combination of the constructions from the proofs of
parts (a) and (b):
Proof of Theorem 1.7(c). Consider Un, Nn and fm as in the proof of Theorem 1.7(a).
of infinite measure Dm. Then by Lemma 2.5, for each m there is a norm-one function
Since µ(X \Sn Nn) = ∞, X \ Sn Nn can be written as a disjoint union of Borel sets
gDm ∈ Lp \Sq<p Lq which is zero outside of Dm. Then the norm-one functions
fm + gDm
, m ∈ N
are in S′
p and have almost disjoint supports.
2
(cid:3)
We have shown that, under special circumstances, the sets Sp, Lp \ ∪0<q<pLq and S′
p,
united to {0}, admit copies of ℓp. This suggests the following definition:
3. V -spaceability
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
9
Definition 3.1. Let V be a topological vector space and S be a subset of V . Given a
subspace W of V , we say that S is W -spaceable if S ∪ {0} contains a W -isomorphic
subspace of V .
There are plenty of examples where V -spaceability (of a subset of a topological vector
space V ) is a strictly more restrictive condition than maximal-dimension spaceability; for
instance, L1[0, 1] admits a subspace isomorphic to ℓ2, which turns to be maximal-dimension
spaceable but not L1[0, 1]-spaceable. As usual, we can add adjectives like "isometrically",
"complementably" and so on to "W -spaceable", depending on how nicely placed in V is
the copy of W we have found in S ∪ {0}. For example, Theorem 1.7(a) says that, under
our assumptions, Sp is isometrically (and complementably, if p ≥ 1) ℓp-spaceable. Note
that, given a topological vector space V , the notion of V -spaceability of some subset S
of V is quite strong; in particular, it implies that S is maximal-dimension spaceable, and
that S ∪ {0} contains copies of all subspaces of V . We have this phenomenon occurring,
for example, for the set of nowhere differentiable functions in C([0, 1]). The main theorem
from [21] can be reformulated as follows:
Theorem 3.2 (Rodr´ıguez-Piazza [21]). In C([0, 1]), the set of nowhere differentiable func-
tions is isometrically C([0, 1])-spaceable.
One step further is due to Hencl:
Theorem 3.3 (Hencl [16]). In C([0, 1]), the set of nowhere approximatively differentiable
and nowhere Holder functions is isometrically C([0, 1])-spaceable.
Another example derives from Theorem 1.7(b):
Corollary 3.4. In ℓp, the set ℓp \ ∪0<q<pℓq is isometrically ℓp-spaceable, and if p ≥ 1, it is
isometrically and complementably ℓp-spaceable.
To see this, just notice that the positive integers with the counting measure satisfy the
conditions in Theorem 1.7(b). This example is less interesting than the two previous ones,
since all closed infinite-dimensional subspaces of ℓp are isomorphic to ℓp, while C([0, 1])
contains isometric copies of all separable Banach spaces.
This remarks naturally motivate new directions of investigation concerning spaceability.
In our context of nowhere p-integrable functions, we can pose the following:
Problem 2. Under appropriate assumptions, for which subspaces V of Lp, is Sp (or
Lp \ ∪0<q<pLq, or S′
p) (isometrically, complementably...) V -spaceable?
The same could be asked when studying the spaceability of any other subset of a topo-
logical vector space. In the remaining of this section, we present some initial results in the
direction of solving this problem (Theorems 3.5 and 3.6, and Corollary 3.7):
Theorem 3.5. Suppose that 1 ≤ p < ∞ and that (X, µ) is a positive Borel measure space
such that Sp(X) is nonvoid. Then Sp(X × [0, 1]) is ℓ2-spaceable in Lp(X × [0, 1]).
Theorem 3.6. Sp([0, 1]) is ℓ2-spaceable in Lp[0, 1], for each 1 ≤ p < ∞.
10
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
Corollary 3.7. Sp([0, 1]n) is ℓ2-spaceable in Lp([0, 1]n), for each 1 ≤ p < ∞ and each
n ∈ N.
Note that Corollary 3.7 follows easily by induction from Theorems 3.5 and 3.6. Note
also that Theorem 3.6 is another improvement of Theorem 1.4 (for the p ≥ 1 case), in a
different direction if we compare to the improvement provided by Theorem 1.7(a).
To prove Theorem 3.5, we shall need some auxiliary results. The first one is a corollary
from the following:
Theorem 3.8 (Kitson, Timoney [18], Theorem 3.3). Let (En) be a sequence of Banach
spaces and F be a Fr´echet space. Let Tn : En → F be continuous linear operators and
W
.
= span{Sn Tn(En)}.
If W is not closed in F , then F \ W is spaceable.
Corollary 3.9. Let E be an infinite-dimensional Banach space and V be the linear span
of a sequence of elements of E. Then E \ V is spaceable.
In particular, if E is a sequence space, the set of elements x = (xn) of E such that
xn 6= 0 for infinitely many n is spaceable.
Proof. Note that, for the first part, it suffices to show that E \ V is spaceable in the
case where E is the closed linear span of (xn) and V is the linear span of (xn), for some
linearly independent sequence (xn) in E. But in this case, defining Tn : Rn → E by
Tn(λ1, . . . , λn)
E, we can apply Theorem 3.8 and conclude the proof of the first part.
j=1 λjxj, we have that V = span{Sn Tn(Rn)}. Since V is not closed in
For the second part, just apply the first part to V = span{en : N}, where {en : N} is the
(cid:3)
=Pn
.
canonical basis of E.
The sequence space we will be interested in will be ℓ2. Recall that,
Rademacher functions defined on [0, 1] by rn(t)
(rn), as a sequence in Lp[0, 1], is equivalent to the canonical basis of ℓ2.
if rn are the
.
= sign(sin(2nπt)) and 0 < p < ∞, then
Lemma 3.10. Let (an) be an element of ℓ2 having infinitely many nonzero an's. Then for
all open ∅ 6= U ⊂ [0, 1] we have that (Panrn)U 6≡ 0, where Panrn is a series in Lp[0, 1]
(0 < p < ∞).
Proof. Let (an) be an element of ℓ2 having infinitely many nonzero an's, and let U be a
nonempty open subset of [0, 1]. Let us denote
anrn, f<j
.
=
anrn, and f≥j
.
=
anrn.
.
=
f
∞
Xn=1
j−1
Xn=1
∞
Xn=j
U contains an interval of the form I = [ k
2N ], for some N ∈ N and k = 0, . . . , N −1. Note
that f<N is constant in I, but since we have infinitely nonzero an's, f≥N is not constant in
I. Thus f cannot be constant in I.
(cid:3)
2N , k+1
Proof of Theorem 3.5. By Corollary 3.9, there is a closed infinite-dimensional subspace
n=1 anrn from F , we have
F of span{rn : n ∈ N} such that, for each nonzero element P∞
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
11
that infinitely many elements from (an) are nonzero. By Lemma 3.10, for each nonzero
element h of F and each nonvoid open subset U from [0, 1], we have hU 6≡ 0. Note that,
as a closed infinite-dimensional subspace of a space isomorphic to ℓ2, F is isomorphic to
ℓ2.
Let f be a norm-one element from Sp(X), and define Φ : F → Lp(X × [0, 1]) by
∞
Note that the support of Φ(h) is σ-finite for each h ∈ F , and Φ is clearly an isometric
isomorphism onto its range. By Fubini's theorem,
p
p
∞
f (x)
anrn(t).
p
dt dx
.
= f (x)
anrn! (x, t)
Φ ∞
Xn=1
anrn!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Φ ∞
Xn=1
Xn=1
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
anrn(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=ZXZ 1
Xn=1
anrn(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f (x)p dx(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=Z 1
0 (cid:18)ZX
Xn=1
anrn(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
anrn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
dt =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
=Z 1
Xn=1
Xn=1
for allP∞
It remains to show that Φ(F ) ⊂ Sp(X ×[0, 1])∪{0}. LetP∞
anrn(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
anrn!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
a (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=Z b
dtZU
ZU ×(a,b)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Φ ∞
Xn=1
∞
Xn=1
p
∞
q
q
∞
∞
p
dt
p
p
,
f (x)qdx
n=1 anrn ∈ F . It follows that Φ(F ) is a ℓ2-isomorphic subspace from Lp(X ×[0, 1]).
n=1 anrn be a nonzero element
from F , U × (a, b) be a nonvoid basic open subset of X × [0, 1], and p ≤ q < ∞. Then
converges if q = p (by Khinchine's inequality and since f is p-integrable), and does not
converge if q > p (since the first factor is strictly positive by Lemma 3.10 and f is not
q-integrable). This concludes our proof.
(cid:3)
Before proceeding to the proof of Theorem 3.6, we point out that it is not hard to prove,
using Theorem 3.5, that Lp[0, 1] \Sq>p Lq[0, 1] is ℓ2-spaceable in Lp[0, 1], for 1 ≤ p < ∞.
In effect, recall that there exists a measure-preserving Borel isomorphism ψ from [0, 1]2
onto [0, 1], which in turn induces an isometric isomorphism Ψ from Lp([0, 1]2) onto Lp[0, 1],
.
= f ◦ ψ−1. It is easy to verify that, for each f ∈ Lp([0, 1]2) and each
defined by Ψ(f )
q > p, f is q-integrable if and only if Ψ(f ) is q-integrable; in particular, all functions
in Ψ(Sp([0, 1]2)) are not q-integrable for q > p, and the claim follows. But the nowhere
part is lost, since ψ is not an homeomorphism. We need thus to provide a finer construction.
Proof of Theorem 3.6. Let µ be the Lebesgue measure on [0, 1], that is, the unique Borel
measure such that µ([k/2n, (k + 1)/2n]) = 1/2n for every n ∈ N and k = 0, 1, . . . , 2n − 1.
By λ denote the Lebesgue measure on {0, 1}N, that is, the unique Borel measure such that
12
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
λ(hsi) = 1/2s for every finite sequence s of zeros and ones where s is the length of s
and hsi stands for the set of all x ∈ {0, 1}N such that x(k) = s(k) for k = 1, 2, . . . , s.
2n . Note that g−1(k/2n) consists of two
elements x, y such that x is a binary representation of k/2n with x(m) = 0 for m > n,
and y is a binary representation of k/2n with x(m) = 1 for m > n. Moreover g−1(t) is a
singleton if t is not of the form k/2n.
Let g : {0, 1}N → [0, 1] be given by g(x) =P∞
x(n)
n=1
Claim 1. µ(A) = λ(g−1(A)) for every Borel set A in [0, 1] and λ(B) = µ(g(B)) for every
Borel set B in {0, 1}N.
It is enough to show this for A = [k/2n, (k + 1)/2n] and B = hsi. Let s be a finite
sequence of zeros and ones which is a binary representation of the number k/2n. Then
s = n and g−1(A) = hsi. Thus λ(g−1(A)) = λ(hsi) = 1/2n = µ(A).
Let n = s and define k = s(n) + 2s(n − 1) + 22s(n − 2) + · · · + 2n−1s(1). Then g(B) =
g(hsi) = [k/2n, (k + 1)/2n]. Thus µ(g(B)) = µ([k/2n, (k + 1)/2n]) = 1/2n = 1/2s = λ(B).
Claim 1 is then proved.
Claim 2. F : Lp[0, 1] → Lp({0, 1}N) given by F (f )
between Lp[0, 1] and Lp({0, 1}N), with inverse L : Lp({0, 1}N) → Lp[0, 1] given by
.
= f ◦ g is an isometric isomorphism
L(h)(t)
.
=(cid:26) h(g−1(t)), if t ∈ [0, 1] is not of the form k/2n;
h(x), if t = k/2n and x is the binary representation of t with x(m) = 0, m > n.
Moreover, F Lq[0,1] = Lq({0, 1}N) for each q > p.
If A is a Borel set in [0, 1] and f = χA is the characteristic function of a set A, then by
Claim 1 we have
It is easily seen that (3.1) also holds if f is a step function, and it follows that
χg−1(A)dλ =Z{0,1}N
f ◦ gdλ.
(3.1)
Z[0,1]
f dµ = µ(A) = λ(g−1(A)) =Z{0,1}N
f dµ =Z{0,1}N
Z[0,1]
f ◦ gdλ.
holds for each f ∈ L1[0, 1]. It follows easily that kf kp = kf ◦ gkp, for f ∈ Lp[0, 1]. This
shows that F is norm-preserving and that F Lq[0,1] = Lq({0, 1}N) for each q > p.
It is clear that L is a left inverse for F . Note that, for a given h ∈ Lp({0, 1}N), F (L(h))
eventually differs from h on a countable set of elements x ∈ {0, 1}N with x(m) = 1
for almost every m. Since λ is a continuous measure, then λ({x ∈ {0, 1}N : h(x) 6=
F (L(h))(x)}) = 0. This means that h and F (L(h)) are the same element of Lp({0, 1}N).
Thus Claim 2 is proved.
Claim 3. F (Sp([0, 1])) = Sp({0, 1}N) and L(Sp({0, 1}N)) = Sp([0, 1]).
Let f ∈ Sp([0, 1]) and fix a basic set hsi in {0, 1}N. Note that there exists a positive
integer k such that g(hsi) = [k/2n, (k + 1)/2n], where n = s. Since
F (f )χhsi = (f ◦ g)χhsi = (f χ[k/2n,(k+1)/2n]) ◦ g = F (f χ[k/2n,(k+1)/2n])
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
13
and f χ[k/2n,(k+1)/2n 6∈ Lq[0, 1], it follows by the previous claim that F (f )χhsi 6∈ Lq({0, 1}N).
That means that F (f ) is nowhere Lq. We have then proved that F (Sp([0, 1])) ⊂ Sp({0, 1}N)
and, using the previous Claim, that L(Sp({0, 1}N)) ⊃ Sp([0, 1]).
Now let h ∈ Sp({0, 1}N) and fix a set [k/2n, (k + 1)/2n]. Let s be a finite set which is a
binary representation of k/2n. Since hχhsi is not in Lq for q > p, then
L(hχhsi) = (h ◦ g−1)χ[k/2n,(k+1)/2n] = L(h)χ[k/2n,(k+1)/2n].
Thus L(h) is nowhere Lq and the proof of Claim 3 is complete.
Given two positive Borel measure spaces X and Y and 0 < p < ∞, we shall say that
an application ϕ : Lp(X) → Lp(Y ) preserves Sp if ϕ(Sp(X)) ⊂ Sp(Y ). Claim 3 asserts in
particular that both F and L preserve Sp.
Claim 4. There is an isometric isomorphism G from Lp({0, 1}N) onto Lp({0, 1}N×{0, 1}N)
which preserves Sp.
Let ϕ : {0, 1}N × {0, 1}N → {0, 1}N be defined by
ϕ((x(1), x(2), x(3), . . . ), (y(1), y(2), y(3), . . . ))
.
= (x(1), y(1), x(2), y(2), x(3), y(3), . . . ).
It is well known that ϕ is a homeomorphism of {0, 1}N × {0, 1}N and {0, 1}N. Fix two finite
sequences s and s′ of zeros and ones. Note that
and
1
2s
1
2s′ = λ(hsi)λ(hs′i) = λ × λ(hsi × hs′i)
λ(ϕ(hsi × hs′i)) =
1
2s+s′ .
The last equality follows from the fact that the set ϕ(hsi × hs′i) is a subset of {0, 1}N such
that exactly s + s′ of its coordinates are fixed.
Using this we obtain that λ×λ(A) = λ(ϕ(A)) for any Borel subset A of {0, 1}N ×{0, 1}N.
.
Then G : Lp({0, 1}N × {0, 1}N) → Lp({0, 1}N) defined by G(f )
= f ◦ ϕ has inverse given by
G−1(h) = h ◦ ϕ−1 and satisfies the desired properties. This completes the proof of Claim
4.
Mimicking the reasoning used to prove Claims 1 -- 3, we can show that T : Lp({0, 1}N ×
[0, 1]) → Lp({0, 1}N × {0, 1}N) defined by
T (f )(x, y)
.
= f (x, g(y))
is an onto isometric isomorphism which preserves Sp. Hence, we have built the following
chain of Sp-preserving isometric isomorphisms:
Lp({0, 1}N×[0, 1])
T−→ Lp({0, 1}N×{0, 1}N) G−→ Lp({0, 1}N) L−→ Lp[0, 1].
Theorem 1.7 implies that Sp({0, 1}N) is nonempty, and from Theorem 3.5 we then ob-
tain that Sp({0, 1}N × [0, 1]) is ℓ2-spaceable in Lp({0, 1}N × [0, 1]). The conclusion follows
immediately.
(cid:3)
14
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
4. Dense-lineability: proof of Theorem 1.7(d)
We start by establishing some notation before proceeding to the proof of Theorem 1.7(d),
which will also be via Lemma 2.3. First, recall that a family {Ai : i ∈ I} of infinite subsets
of N is said to be almost disjoint if Ai ∩ Aj is finite for any distinct i, j ∈ I. It is well
known that there is a family of almost disjoint subsets of N of cardinality continuum. Let
{A′
α : α < c} be such family.
Fix a sequence of integers 1 = n0 < n1 < n2 < n3 < . . . such that
and consider Mk
Note that {Aα : α < c} is an almost disjoint family and that
.
= {nk, nk+1, . . . , nk+1−1}. Define, for each α < c, Aα
nk+1−1
Xi=nk
1
i
≥ 1,
.
=S{Mk : k ∈ A′
α}.
1
i
= ∞
Xi∈Aα
for each α < c. We shall fix the family {Aα : α < c} and use it in the following.
Proof of Theorem 1.7(d). For a fixed Borel set A of positive finite measure and α < c
we define a function hα
A as follows. Let {An,m : n, m ∈ N} be a family of pairwise disjoint
subsets of A of positive measure such that µ(An,m) ≥ 2µ(An,m+1), and let (rn) be a strictly
decreasing sequence of reals tending to p. Put
hα
n
.
= Xm∈Aα
am,nχAn,m,
(4.1)
where arn
that the A-supported, norm-one function
m,nµ(An,m) = 1/m. Then a similar argument as used in Lemma 2.4 leads us to
.
=
hα
A
hα
n
khα
nk2n
∞
Xn=1
(4.2)
is in Lp \Sq>p Lq.
As in the proof of Theorem 2.2, fix a basis (Un) for X, and let Nn ⊂ Un be a sequence
of pairwise disjoint nowhere dense Borel sets satisfying 0 < µ(Nn) < 1/2n. For each α < c,
by defining hα
Nn as in (4.2) and putting
f α .
=
we obtain that f α ∈ Sp and has norm one.
hα
Nn
2n ,
∞
Xn=1
Note that any ordinal number α < c is of the form β + n, where β is a limit ordinal and
n = 0, 1, 2, . . . . Let {Bβ : β < c} be an indexation of all Borel subsets of X. Then the set
{(Bβ, n) : β < c, n ∈ N} has cardinality c, thus there is a bijection (Bβ, n) 7→ α(β, n) onto
all ordinals less than c.
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
15
Consider, for β < c and n ∈ N, the functions
gβ,n .
= gα(β,n)
.
= χBβ +
1
n
f α(β,n).
(4.3)
By our construction, the linear span of {gα(β,n) : β < c, n ∈ N} is dense in the set of all
simple functions on X, and therefore it is also dense in Lp. We will show show that any
nontrivial linear combination of functions of the form (4.3) is in Sp. Let (β1, n1), . . . , (βk, nk)
be distinct and consider b1, . . . , bk ∈ R which are not all zero, and write
.
= b1gβ1,n1 + · · · + bkgβk,nk = (b1χBβ1
g
+ · · · + bkχBβk
) +
b1
n1
f α(β1,n1) + · · · +
bk
nk
f α(βk,nk).
Consider αi
then write
.
= α(βi, ni), and note that α1, . . . , αk are distinct ordinal numbers. We can
g = (b1χBβ1
+ · · · + bkχBβk
) +
= (b1χBβ1
+ · · · + bkχBβk
) +
b1
n1
b1
n1
f α1 + · · · +
bk
nk
f αk
hα1
Nn
2n + · · · +
bk
nk
∞
Xn=1
hαk
Nn
2n .
∞
Xn=1
Consider the family {Al,m : l, m ∈ N} of pairwise disjoint subsets of Nn of positive
Nn as in (4.1), using these sets and
.
=
.
= Aαk \ {1, 2, . . . , N} are disjoint; this is possible since {Aα :
measure such that µ(Al,m) ≥ 2µ(Al,m+1), and construct hαi
the corresponding al,m. Consider N ∈ N such that the sets C1
Aα2 \ {1, 2, . . . , N}, . . . , Ck
α < c} is almost disjoint. Then we have
.
= Aα1 \ {1, 2, . . . , N}, C2
hαi
l = Xm∈Aαi
amχAl,m = Xm∈Aαi T{1,...,N }
amχAl,m + Xm∈Ci
amχAl,m,
∞
1
hαi
Xl=1
l 2l
hαi
l
l 2l =
hαi
l 2l Xm∈Aαi T{1,...,N }
1
hαi
Xm∈Aαi T{1,...,N }
Xl=1
amχAl,m +
∞
amχAl,m + Xm∈Ci
l 2l Xm∈Ci
1
hαi
amχAl,m.
amχAl,m
amχAl,m for each i = 1, . . . , k, by our construction we
and thus
∞
hαi
Nn =
=
Writing wi
∞
Xl=1
Xl=1
= P∞
.
l=1
h
1
αi
l
2l Pm∈Ci
=SmSl∈Ci
.
have that each wi is in Lp \Sq>p Lq and w1, . . . , wk have disjoint supports; more precisely,
the support of each wi is N i
n} ⊂ Sp. The fact
n
that g ∈ Sp follows then from the fact that adding a simple function to a function from Sp
results in a function from Sp.
Al,m. Note that span{f αiχSn N i
Since Lp(X, µ) is separable, it has dimension c, as does span{gα(β,n) : β < c, n ∈ N},
(cid:3)
which concludes our proof.
16
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
5. Algebrability: proof of Theorem 1.7(e)
Proof of Theorem 1.7(e) Let (Un) be a basis for X. Similarly to the construction held
at the beginning of the Proof of Theorem 1.7(a), one can find pairwise disjoint nowhere
dense Borel sets Nn such that Nn ⊂ Un and 0 < µ(Nn) < 1
2n . Using Lemma 2.3, we can
find for each n a pairwise disjoint family (Nn,j)j of Borel subsets of Nn satisfying
µ(Nn,j+1) ≤
1
j + 1
µ(Nn,j).
Note that, for each n, j, we have µ(Nn,j) ≤ 1
open subsets of X intercept each Bj in non-null sets, and by the other hand µ(Bj) =
= Sn Nn,j. Then all nonvoid
j!2n . Let Bj
.
Let {θα : α < c} be a set of real numbers strictly greater than 1 such that the set
{ln(θα) : α < c} is linearly independent over the rational numbers. For each α < c, define
Pn µ(Nn,j) ≤ 1
j!.
.
=
gα
θj
αχBj .
∞
Xj=1
θpj
j! converges, thus each gα ∈ Lp, for each α < c and each
α
For each α the series Pj
0 < p < ∞.
Let us show that {gα : α < c} is a set of free generators, and the algebra generated
by this set is contained in G ∪ {0}. It suffices to show that, for every m and n positive
integers, for every matrix (kil : i = 1, . . . , m, l = 1, . . . , n) of non-negative integers with
non-zero and distinct rows, for every α1, . . . , αn < c and for every β1, . . . , βm ∈ R which
do not vanish simultaneously, the function
.
= β1gk11
α1 . . . gk1n
αn + · · · + βmgkm1
α1
g
. . . gkmn
αn
∞
(β1(θk11
α1 . . . θk1n
αn )j + · · · + βm(θkm1
α1
. . . θkmn
αn )j)χBj
=
Xj=1
is in G. First, let us show that it is in T0<p<∞ Lp. Fix p and put, for each i = 1, . . . , m,
α1 · · · θkin
αn . Then
.
= θki1
θi
Z gp ≤Z " ∞
Xj=1
(β1θj
1 + · · · + βmθj
m)pχBj# ≤
Q(θj
1, . . . , θj
m)
j!
∞
Xj=1
,
(5.1)
where Q : (x1, . . . , xm) 7→ (β1θj
such that Q(θj
1, . . . , θj
converges and g ∈ Lp.
m)p. It is straightforward to find C, b > 0
m) < C + bj for all j. Thus the sum on the right handside of (5.1)
1 + · · · + βmθj
Since ln(θi) = ln(θki1
αn ) = ki1 ln θα1 + · · · + kin ln θαn and ln θα1, . . . , ln θαn are
Q-linearly independent, the numbers ln(θ1), . . . , ln(θm) are distinct. Then by the strict
monotonicity of the logarithmic function we may assume that
α1 · · · θkin
θ1 > · · · > θm;
(5.2)
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
17
we also may assume β1 6= 0. Then we can write
∞
From (5.2) and since β1 is assumed to be nonzero, we can find j0 ∈ N such that
g =
(β1θj
1 + · · · + βmθj
m)χBj .
Xj=1
β2θj
2 + · · · + βmθj
m <
1
2
β1θj
1
for all j ≥ j0. Then for those j
1 + · · · + βmθj
β1θj
m ≥ β1θj
2 + · · · + βmθj
≥ β1θj
2 + ... + βmθj
1 −(cid:12)(cid:12)β2θj
1 −(cid:0)β2θj
m(cid:12)(cid:12)
m(cid:1) >
1
2
β1θj
1.
Since each nonvoid open subset of X intercepts all Bj in non-null sets, the inequality above
shows that g is nowhere essentially bounded.
(cid:3)
5.1. Comments and open problems. As a Corollary from Theorem 1.7(e) we have the
following:
Corollary 5.1. If µ is an atomless and outer regular positive Borel measure on X with
full support and 0 < p < ∞, then
.
= {f ∈ Lp(µ) : f is nowhere L∞(µ)}
Gp
is strongly c-algebrable.
It is a straightforward exercise for the reader to show, using a construction similar to
the one used to prove Theorem 1.7(a), that Gp is spaceable in Lp. To finish this section we
pose the following problem:
Problem 3. Does Gp ∪ {0} admit dense or closed subalgebras of Lp?
6. When are there nowhere q-integrable functions in Lp?
We conclude this work with a couple of remarks and questions on necessary/sufficient
conditions on a positive Borel measure space (X, µ), so that there exist nowhere q-integrable
Borel functions in Lp(X). An obvious necessary condition is that µ has full support, so
we will always assume that. It is not hard to see that it is also necessary that X has the
countable chain condition, as Proposition 6.1 below shows. Recall that X is said to have
the countable chain condition (or ccc, in short) if any family consisting of open non-empty
pairwise disjoint subsets of X is countable.
Proposition 6.1. Let X be a topological space without the ccc, assume that µ positive
Borel measure on X with full support, fix 0 < q < ∞ and let f : X → R be a Borel
function.
If f U is not in Lq(U) for any nonvoid open set U, then f is not in Lp(X) for any
0 < p < ∞.
18
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
Proof. Let (Us)s∈S be an uncountable family of pairwise disjoint non-empty open sets.
Since f Us is not in Lq(Us) for any Us, then f does not vanish on Us. Thus for each
0 < p < ∞ and each s ∈ S, kf Uskp > 0. Fix 0 < p < ∞. Since S is uncountable, then at
least one of the sets
Sn
.
=(cid:26)s ∈ S :ZUs
is uncountable. Hence RX f pdµ = ∞.
The next natural step is pose the following:
f pdµ ≥
1
n(cid:27)
(cid:3)
Problem 4. Suppose that X has the ccc, and that µ is a positive Borel measure on
X with full support. Given 0 < p < ∞, does there exist a Borel p-integrable function
f : X → R which is nowhere q-integrable for q > p?
We provide a partial answer to the problem above, through a consistency result. Recall
first that the product of two spaces with the ccc do not need to have the ccc, however this
statement is independent of ZFC. Under Martin's axiom, the product of two ccc spaces
has the ccc, but in some models of ZFC there exists a topological space called Suslin line,
which has the ccc but its square does not have the ccc (cf. [19]).
Theorem 6.2. It is consistent with ZFC that there is a topological space X satisfying the
ccc such that, for any positive Borel measure µ on X with full support and any 0 < p < ∞,
there is no Borel function f : X → R in Lp(µ) but nowhere Lq(µ) for q > p.
Proof. It is consistent with ZFC that there exists a Suslin line X. Suppose that there is
a Borel function f : X → R in Lp(µ) but nowhere Lq(µ) for q > p. Let f : X 2 → R
be defined by f (x, y) = f (x)f (y). Clearly f is Borel, and since supp f is σ-finite, so is
supp f = (supp f )2. Fubini's theorem then implies that
and for any two nonvoid open sets U, V ⊂ X and q > p we have that
f qZV
f q = ∞.
fpd(µ×µ) = kf k2p
p ,
ZX 2
fqd(µ×µ) =ZU
ZU ×V
Hence f is a Borel function in Lp(µ × µ) but nowhere Lq(µ × µ) for q > p and µ × µ
is a positive Borel measure with full support. Since X 2 does not have the ccc, we get a
contradiction.
(cid:3)
Finally we turn our attention to the presence of countable π-bases. First, note that
there exist topological spaces X with countable π-bases but admitting no countable bases,
and with positive Borel measures with full support defined on them: take for example the
Sorgenfrey line (the set of real numbers with the topology generated by intervals of the
form [a, b)) with the Lebesgue measure. The Sorgenfrey line RS has a countable π-basis
LARGE STRUCTURES MADE OF NOWHERE Lq FUNCTIONS
19
thus we can apply Theorem 1.7 to show that Sp(RS) is ℓp-spaceable, but any basis of the
Sorgenfrey line has cardinality c.
It turns out that the presence of a countable π-basis in X is also not necessary for the
existence of nowhere q-integrable functions in Lp(X). In fact, we have more.
Theorem 6.3. Let X be a topological space with a countable π-basis. Suppose that µ is an
atomless and outer-regular Borel probability measure with full support. Assume that κ is
an uncountable cardinal number. Let Y = X κ be the Tychonoff product of κ many copies
of X, and let λ = µκ be the product of κ many copies of µ. Then the Sp(Y ) is spaceable in
Y .
Proof. Let (Un) be a countable π-base in X. By a construction used to prove Theorem
1.7(a) applied to X and µ, there is a norm-one basic sequence f1, f2, . . . of elements of
k=1 akχAk where Ak are
.
Borel subsets of X. For (xα)α<κ ∈ X κ, put fi((xα)α<κ)
= fi(x0). Then ( fi) is a norm-one
basic sequence with pairwise disjoint supports. We need to show that each of them is in
Sp(X) with pairwise disjoint supports, and each fi is of the formP∞
Sp(Y ). Note that fi is of the form P∞
Let V be an nonempty open subset of Y . We may assume that V is of the formQα<κ Wα
if α ∈ κ \ F . Let F0 = F \ {0}. We have V = W0 ×Q1≤α<κ Wα and
ZV
where Wα are nonempty open subsets of X and there is finite set F ⊂ κ such that Wα = X
= Ak ×Q1≤α<κ X.
akqλ( Ak ∩ V )
, where Ak
k=1 akχ Ak
∞
∞
.
∞
=
dλ =
Xk=1
akqχ Ak
fiqdλ =ZV
Xk=1
akqλ (Ak ∩ W0) × Y1≤α<κ
Xk=1
µ(Wα)ZW0
= Yα∈F0
ZY
fipdλ =ZX
fiqdµ = ∞.
Similarly we get that
and our proof is concluded.
Wα! =
∞
Xk=1
akqµ(Ak ∩ W0) Yα∈F0
µ(Wα)
f pdµ < ∞,
(cid:3)
Acknowledgement: we would like to thank prof. Gilles Godefroy for several comments
that lead to improvements of this work.
References
[1] R. Aron, V. I. Gurariy, J. B. Seoane, Lineability and spaceability of set of functions on R, Proc. Amer.
Math. Soc., 133 (2005), 795 -- 803.
[2] R. Aron, J. B. Seoane-Sep´ulveda, Algebrability of the set of everywhere surjective functions on C,
Bull. Belg. Math. Soc. Simon Stevin, 14 (2007), 25 -- 31.
[3] A. Bartoszewicz and Sz. G l¸ab, Strong algebrability of sets of sequences and functions, Proc. Amer.
Math. Soc., DOI:10.1090/S0002-9939-2012-11377-2
[4] F. Bayart and L. Quarta, Algebras in sets of queer functions. Israel J. Math. 158 (2007), 285 -- 296.
20
SZ. G LA¸ B, P. L. KAUFMANN, AND L. PELLEGRINI
[5] L. Bernal-Gonz´alez, Algebraic genericity and strict-order integrability, Studia Math. 199(3) (2010),
279 -- 293.
[6] L. Bernal-Gonz´alez, M. Ord´onez Cabrera, Spaceability of strict order integrability, J. Math. Anal.
Appl. 385 (2012), 303 -- 309.
[7] G. Botelho, D. Cariello, V. F´avaro, D. Pellegrino, and J. B. Seoane-Sep´ulveda, Subspaces of maximal
[8] G. Botelho, D. Diniz, V. F´avaro, and D. Pellegrino, Spaceability in Banach and quasi-Banach sequence
dimension contained in Lp(Ω) −Sq<p Lq(Ω), preprint, 2012.
spaces, Linear Algebra Appl. 434 (2011), no. 5, 1255 -- 1260.
[9] G. Botelho, V. F´avaro, D. Pellegrino, and J. B. Seoane-Sep´ulveda, Lp[0, 1] \Sq>p Lq[0, 1] is spaceable
for every p > 0, Linear Algebra Appl. 436 (2012), no. 9, 2963 -- 2965.
[10] P. Enflo, V. I. Gurariy, J. B. Seoane-Sep´ulveda, On lineability and spaceability of sets in function
spaces, Trans. Amer.Math. Soc., to appear.
[11] Sz. G l¸ab, P. L. Kaufmann, and L. Pellegrini, Spaceability and algebrability of sets of nowhere integrable
functions, Proc. Amer. Math. Soc., to appear.
[12] F. J. Garc´ıa-Pacheco, N. Palmberg and J. B. Seoane-Sep´ulveda, Lineability and algebrability of patho-
logical phenomena in analysis, J. Math. Anal. Appl., 326 (2007), 929 -- 939.
[13] F. J. Garc´ıa-Pacheco, C. P´erez-Eslava and J. B. Seoane-Sep´ulveda, Moduleability, algebraic structures
and nonlinear properties J. Math. Anal. Appl., 370 (2010), 159 -- 167.
[14] V. I. Gurariy, Subspaces and bases in spaces of continuous functions (Russian), Dokl. Akad. Nauk
SSSR 167 (1966), 971 -- 973.
[15] V. I. Gurariy, Linear spaces composed of nondifferentiable functions, C. R. Acad. Bulgare Sci. 44
(1991), 4 13 -- 16.
[16] V. Hencl, Isometrical embeddings of separable Banach spaces into the set of nowhere approximatively
and nowhere Holder functions, Proc. Amer. Math. Soc., 128 (2000), no. 12, 3505 -- 3511.
[17] N. J. Kalton, Compact and strictly singular operators on Orlicz spaces, Israel J. Math. 26 (1977),
126 -- 136.
[18] D. Kitson and R. M. Timoney, Operator ranges and spaceability, J. Math. Anal. Appl. 378 (2011),
680 -- 686.
[19] K. Kunen, Set theory. An introduction to independence proofs, vol. 102, North-Holland Publishing
Co., Amsterdam, 1983.
[20] G. A. Munoz-Fern´andez, N. Palmberg, D. Puglisi, and J. B. Seoane-Sep´ulveda, Lineability in subsets
of measure and function spaces, Linear Algebra Appl. 428 (2008), 2805 -- 2812.
[21] L. Rodr´ıguez-Piazza, Every separable Banach space is isometric to a space of nowhere differentiable
functions, Proc. Amer. Math. Soc. 123 (1995), no. 12, 3649 -- 3654.
Institute of Mathematics, Technical University of L´od´z, W´olcza´nska 215, 93-005 L´od´z,
Poland
E-mail address: [email protected]
Instituto de matem´atica e estat´ıstica, Universidade de Sao Paulo, Rua do Matao, 1010,
CEP 05508-900, Sao Paulo, Brazil
E-mail address: [email protected]
Instituto de matem´atica e estat´ıstica, Universidade de Sao Paulo, Rua do Matao, 1010,
CEP 05508-900, Sao Paulo, Brazil
E-mail address: [email protected]
|
1605.06389 | 1 | 1605 | 2016-05-20T14:59:41 | On horizontal Hardy, Rellich, Caffarelli-Kohn-Nirenberg and $p$-sub-Laplacian inequalities on stratified groups | [
"math.FA",
"math.AP"
] | In this paper, we present a version of horizontal weighted Hardy-Rellich type and Caffarelli-Kohn-Nirenberg type inequalities on stratified groups and study some of their consequences. Our results reflect on many results previously known in special cases. Moreover, a new simple proof of the Badiale-Tarantello conjecture [2] on the best constant of a Hardy type inequality is provided. We also show a family of Poincar\'e inequalities as well as inequalities involving the weighted and unweighted $p$-sub-Laplacians. | math.FA | math |
ON HORIZONTAL HARDY, RELLICH,
CAFFARELLI-KOHN-NIRENBERG AND p-SUB-LAPLACIAN
INEQUALITIES ON STRATIFIED GROUPS
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
Abstract. In this paper, we present a version of horizontal weighted Hardy-
Rellich type and Caffarelli-Kohn-Nirenberg type inequalities on stratified groups
and study some of their consequences. Our results reflect on many results previ-
ously known in special cases. Moreover, a new simple proof of the Badiale-Tarantello
conjecture [2] on the best constant of a Hardy type inequality is provided. We also
show a family of Poincar´e inequalities as well as inequalities involving the weighted
and unweighted p-sub-Laplacians.
1. Introduction
Consider the following inequality
(cid:13)(cid:13)(cid:13)(cid:13)
f (x)
kxk(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
≤
p
n − p
k∇f kLp(Rn) ,
1 ≤ p < n,
(1.1)
p
0 (Rn\{0}), kxk = px2
where ∇ is the standard gradient in Rn, f ∈ C ∞
1 + ... + x2
n,
n−p is known to be sharp. The one-dimensional version of (1.1)
and the constant
for p = 2 was first discovered by Hardy in [28], and then for other p in [29], see
also [29] for the story behind these inequalities. Since then the inequality (1.1) has
been widely analysed in many different settings (see e.g.
[1]-[10], [12], [13], [16],
[17], [20], [30], [31]). Nowadays there is vast literature on this subject, for example,
the MathSciNet search shows about 5000 research works related to this topic. On
homogeneous Carnot groups (or stratified groups) inequalities of this type have been
also intensively investigated (see e.g.
[14], [25], [26], [27], [33], [34], [35], [36], [38]).
In this case inequality (1.1) takes the form
(cid:13)(cid:13)(cid:13)(cid:13)
f (x)
d(x)(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
≤
p
Q − p
k∇Hf kLp(G) , Q ≥ 3, 1 < p < Q,
(1.2)
where Q is the homogeneous dimension of the stratified group G, ∇H is the horizontal
gradient, and d(x) is the so-called L-gauge, which is a particular homogeneous quasi-
norm obtained from the fundamental solution of the sub-Laplacian, that is, d(x)2−Q
is a constant multiple of Folland's [22] (see also [23]) fundamental solution of the
sub-Laplacian on G. For a short review in this direction and some further discussions
2010 Mathematics Subject Classification. 22E30, 43A80.
Key words and phrases. Hardy inequality, Rellich inequality, Caffarelli-Kohn-Nirenberg inequal-
ity, p-sub-Laplacian, horizontal estimate, stratified group.
The authors were supported in parts by the EPSRC grant EP/K039407/1 and by the Leverhulme
Grant RPG-2014-02, as well as by the MESRK grant 5127/GF4.
1
2
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
we refer to our recent papers [40, 41, 42, 43, 44] and [39] as well as to references
therein.
The main aim of this paper is to give analogues of Hardy type inequalities on
stratified groups with horizontal gradients and weights. Actually we obtain more than
that, i.e., we prove general (horizontal) weighted Hardy, Rellich and Caffarelli-Kohn-
Nirenberg type inequalities on stratified groups. Our results extend known Hardy
type inequalities on abelian and Heisenberg groups, for example (see e.g.
[2] and
[11]). For the convenience of the reader let us now briefly recapture the main results
of this paper. Let G be a homogeneous stratified group of homogeneous dimension
Q, and let X1, . . . , XN be left-invariant vector fields giving the first stratum of the
Lie algebra of G, ∇H = (X1, . . . , XN ), with the sub-Laplacian
L =
X 2
k .
NXk=1
Denote the variables on G by x = (x′, x′′) ∈ G, where x′ corresponds to the first
stratum. For precise definitions we refer to Section 2.
Thus, to summarise briefly, in this paper we establish the following results:
• (Hardy inequalities) Let G be a stratified group with N being the dimen-
sion of the first stratum, and let α, β ∈ R. Then for all complex-valued
0 (G\{x′ = 0}) and 1 < p < ∞, we have the following
functions f ∈ C ∞
Lp-Caffarelli-Kohn-Nirenberg type inequality
1
x′α ∇Hf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p−1
f
β
p−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
x′
Lp(G)
,
(1.3)
where γ = α + β + 1 and · is the Euclidean norm on RN . If γ 6= N then the
constant N −γ
is sharp. In the special case of α = 0, β = p − 1 and γ = p,
inequality (1.3) implies
p
≤ k∇Hf kLp(G) ,
1 < p < ∞,
(1.4)
γ
p
Lp(G)
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤(cid:13)(cid:13)(cid:13)(cid:13)
f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
N − γ
p
f
x′
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N − p
p
1
x′
(cid:13)(cid:13)(cid:13)(cid:13)
f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
1
x′
(cid:13)(cid:13)(cid:13)(cid:13)
where the constant N −p
is sharp for p 6= N. One novelty of this is that we
do not require that p < N. In turn, for 1 < p < N, the inequality (1.4) gives
a stratified group version of Lp-Hardy inequality
p
≤
p
N − p
k∇Hf kLp(G) ,
1 < p < N,
(1.5)
again with p
N −p being the best constant.
• (Badiale-Tarantello conjecture) Let x = (x′, x′′) ∈ RN × Rn−N . In [2]
Badiale and Tarantello proved that for 2 ≤ N ≤ n and 1 ≤ p < N there
exists a constant Cn,N,p such that
1
x′
(cid:13)(cid:13)(cid:13)(cid:13)
f(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
≤ Cn,N,p k∇f kLp(Rn) ,
(1.6)
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES
3
where ∇ is the standard Euclidean gradient. Clearly, for N = n this gives the
classical Hardy's inequality with the best constant
Cn,p =
p
n − p
.
It was conjectured in [2, Remark 2.3] that the best constant in (1.6) is given
by
CN,p =
.
(1.7)
p
N − p
This conjecture was proved in [45]. As a consequence of our techniques, we
give a new proof of the Badiale-Tarantello conjecture.
• (Critical Hardy inequality) For p = N, the inequality (1.5) fails.
In
this case the Hardy inequality (1.1) is replaced by a logarithmic version, an
analogue of which we establish on stratified groups as well. For a bounded
domain Ω ⊂ G with 0 ∈ Ω and f ∈ C ∞
0 (Ω\{x′ = 0}) we have
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
f
x′log R
≤
x′(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)LN (Ω)
where R = sup
x∈Ω
x′
x′
N
N − 1(cid:13)(cid:13)(cid:13)(cid:13)
· ∇Hf (x)(cid:13)(cid:13)(cid:13)(cid:13)LN (Ω)
x′. In the abelian case of G = Rn being the Euclidean space,
, N ≥ 2,
(1.8)
inequality (1.8) reduces to the logarithmic Hardy inequality of Edmunds and
Triebel [19].
• (p-sub-Laplacian) Let G be a stratified group with N being the dimension
q = 1 and α, β ∈ R be such
of the first stratum, and let 1 < p < ∞ with 1
that p−N
p−1 ≤ γ := α + β + 1 ≤ 0. Then for all f ∈ C ∞
0 (G\{x′ = 0}) we have
p + 1
N + γ(p − 1) − p
p
∇Hf
x′
γ
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Lp(G)
≤(cid:13)(cid:13)(cid:13)(cid:13)
1
x′α Lpf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)(cid:13)
∇Hf
x′β(cid:13)(cid:13)(cid:13)(cid:13)Lq(G)
where · is the Euclidean norm on RN and Lp is the p-sub-Laplacian operator
defined in (2.3).
• (Higher order Hardy-Rellich inequalities) Let 1 < p < ∞. For any
,
(1.9)
k, m ∈ N we have
γ
p
f
p
1
Lp(G)
x′
N − γ
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤ eAα,meAβ,k(cid:13)(cid:13)(cid:13)(cid:13)
such thatQm−1
eAα,m := pm"m−1Yj=0
x′α−m ∇m+1
j=0 N − p(α − j) 6= 0, and
x′
H f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N − p(α − j)#−1
,
1
β
p−1 −k
∇k
p−1
,
Lp(G)
Hf(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
for any real-valued function f ∈ C ∞
0 (G\{x′ = 0}), γ = α + β + 1, and α ∈ R
4
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
as well as β ∈ R such thatQk−1
j=0(cid:12)(cid:12)(cid:12)N − p( β
eAβ,k := pk(p−1)"k−1Yj=0(cid:12)(cid:12)(cid:12)(cid:12)N − p(cid:18) β
p−1 − j)(cid:12)(cid:12)(cid:12) 6= 0, and
#−(p−1)
− j(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
p − 1
.
[46]) for the horizontal gradient is proved:
• (LN -Poincar´e inequality) The following LN -Poincar´e inequality (see e.g.
kf kLN (Ω) ≤ R k∇Hf kLN (Ω) ,
where R = sup
x∈Ω
x′, for f ∈ C ∞
0 (G) and any bounded domain Ω ⊂ G. Note
that the inequality (1.4) implies
N − p
Rp
kf kLp(Ω) ≤ k∇H f kLp(Ω) ,
1 < p < ∞,
(1.10)
for f ∈ C ∞
0 (Ω\{x′ = 0}) and R = sup
x∈Ω
inequality when N = p.
x′. However, (1.10) gives a trivial
• (Weighted p-sub-Laplacian) Let 0 ≤ F ∈ C ∞(G) and 0 ≤ η ∈ L1
loc(G) be
such that
ηF p−1 ≤ −Lp,ρF,
a.e. in G,
(1.11)
where Lp,ρ is a weighted p-sub-Laplacian defined in (5.8). Then we have
for all real-valued functions f ∈ C ∞
0 (G) and 2 ≤ p < ∞. Here Cp is a positive
constant. For 1 < p < 2 the inequality (1.12) is replaced by an analogous
one while for p = 2 it becomes an identity, see Remark 5.6 and Remark 5.5,
respectively.
In Section 2 we very briefly recall the main concepts of stratified groups and fix
the notation. In Section 3 we derive versions of Lp-Caffarelli-Kohn-Nirenberg type
inequalities on stratified groups and discuss their consequences including higher order
cases as well as a new proof of the Badiale-Tarantello conjecture. An analogue of the
critical Hardy inequality is proved in Section 4. Hardy-Rellich type inequalities and
their weighted versions on stratified groups are presented and analysed in Section 5.
2. Preliminaries
A Lie group G = (Rn, ◦) is called a stratified group (or a homogeneous Carnot
group) if it satisfies the following conditions:
(a) For some natural numbers N + N2 + ... + Nr = n, that is N = N1, the
decomposition Rn = RN × ... × RNr is valid, and for every λ > 0 the dilation δλ :
Rn → Rn given by
δλ(x) ≡ δλ(x′, x(2), ..., x(r)) := (λx′, λ2x(2), ..., λrx(r))
is an automorphism of the group G. Here x′ ≡ x(1) ∈ RN and x(k) ∈ RNk for k =
2, ..., r.
kη
1
p f kp
1
p F ∇H
Lp(G) + Cp(cid:13)(cid:13)(cid:13)(cid:13)ρ
f
F(cid:13)(cid:13)(cid:13)(cid:13)
p
Lp(G)
≤ kρ
1
p ∇Hf kp
Lp(G),
(1.12)
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES
5
(b) Let N be as in (a) and let X1, ..., XN be the left invariant vector fields on G
such that Xk(0) = ∂
∂xk
0 for k = 1, ..., N. Then
rank(Lie{X1, ..., XN }) = n,
for every x ∈ Rn, i.e. the iterated commutators of X1, ..., XN span the Lie algebra of
G.
That is, we say that the triple G = (Rn, ◦, δλ) is a stratified group. See also e.g.
[21] for discussions from the Lie algebra point of view. Here r is called a step of G
and the left invariant vector fields X1, ..., XN are called the (Jacobian) generators of
G. The number
is called the homogeneous dimension of G. The second order differential operator
Q =
kNk, N1 = N,
rXk=1
L =
X 2
k ,
(2.1)
NXk=1
is called the (canonical) sub-Laplacian on G. The sub-Laplacian L is a left invariant
homogeneous hypoelliptic differential operator and it is known that L is elliptic if
and only if the step of G is equal to 1. We also recall that the standard Lebesque
measure dx on Rn is the Haar measure for G (see, e.g. [21, Proposition 1.6.6]). The
left invariant vector field Xj has an explicit form and satisfies the divergence theorem,
see e.g. [40] for the derivation of the exact formula: more precisely, we can write
Xk =
∂
∂x′
k
+
rXl=2
NlXm=1
a(l)
k,m(x′, ..., x(l−1))
∂
∂x(l)
m
,
(2.2)
see also [21, Section 3.1.5] for a general presentation. We will also use the following
notations
for the horizontal gradient,
for the horizontal divergence,
∇H := (X1, . . . , XN )
divHv := ∇H · v
Lpf := divH(∇Hf p−2∇Hf ),
1 < p < ∞,
(2.3)
for the horizontal p-Laplacian (or p-sub-Laplacian), and
for the Euclidean norm on RN .
x′ =qx′2
1 + . . . + x′2
N
The explicit representation (2.2) allows us to have the identities
∇Hx′γ = γx′γ−1,
and
divH(cid:18) x′
x′γ(cid:19) = PN
for all γ ∈ R, x′ 6= 0.
j=1 x′γXjx′
j=1 x′
jγx′γ−1Xjx′
j −PN
x′2γ
(2.4)
(2.5)
=
N − γ
x′γ
6
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
3. Horizontal Lp-Caffarelli-Kohn-Nirenberg type inequalities and
consequences
In this section and in the sequel we adopt all the notation introduced in Section 2
concerning stratified groups and the horizontal operators.
3.1. Caffarelli-Kohn-Nirenberg inequalities. In this section we establish the fol-
lowing horizontal Lp-Caffarelli-Kohn-Nirenberg type inequalities on the stratified
group G and then discuss their consequences and proofs. The proof is analogous
to [39] in the case of homogeneous groups, but here we rely on the divergence theo-
rem rather on the polar decomposition which is less suitable for the stratified setting.
We refer e.g. to [6] and [7] for Euclidean settings of Caffarelli-Kohn-Nirenberg in-
equalities.
Theorem 3.1. Let G be a homogeneous stratified group with N being the dimension
of the first stratum, and let α, β ∈ R. Then for any f ∈ C ∞
0 (G\{x′ = 0}), and all
1 < p < ∞, we have
N − γ
p
f
x′
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
γ
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Lp(G)
≤(cid:13)(cid:13)(cid:13)(cid:13)
1
x′α ∇Hf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p−1
f
β
p−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
x′
Lp(G)
where γ = α + β + 1 and · is the Euclidean norm on RN . If γ 6= N then the constant
N −γ
is sharp.
p
,
(3.1)
In the abelian case G = (Rn, +), we have N = n, ∇H = ∇ = (∂x1, . . . , ∂xn), so
(3.1) implies the Lp-Caffarelli-Kohn-Nirenberg type inequality (see e.g. [10] and [18])
for G ≡ Rn with the sharp constant:
p−1
Lp(Rn)
f
kxk
β
p−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
n. In the case
,
(3.2)
for all f ∈ C ∞
N − p(α + 1)
f
p
p
1
1
Lp(Rn)
f
γ
kxk
n − γ
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1 + . . . + x2
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kxkα ∇f(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤(cid:13)(cid:13)(cid:13)(cid:13)
0 (Rn\{0}), and kxk =px2
p(cid:19) ,
β = γ(cid:18)1 −
≤(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)
x′α ∇Hf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
x′α+1(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
0 (G\{x′ = 0}) and all α ∈ R.
1
x′
(cid:13)(cid:13)(cid:13)(cid:13)
N − p
≤
1
p
p
that is, taking β = (α + 1)(p − 1) and γ = p(α + 1), the inequality (3.1) implies that
,
1 < p < ∞,
(3.3)
for any f ∈ C ∞
When α = 0 and 1 < p < N, the inequality (3.3) gives the following stratified
group version of Lp-Hardy inequality
k∇Hf kLp(G) , 1 < p < N,
(3.4)
p
N −p being the best constant (see [11] and [47] for the version on the
again with
Heisenberg group). In the abelian case G = (Rn, +), n ≥ 3, (3.4) implies the classical
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES
7
Hardy inequality for G ≡ Rn:
for all f ∈ C ∞
1 + . . . + x2
n.
The inequality (3.4) implies the following Heisenberg-Pauli-Weyl type uncertainly
[9], [42], [40] and [39] for different settings):
principle on stratified groups (see e.g.
For each f ∈ C ∞
0 (G\{x′ = 0}), using Holder's inequality and (3.4), we have
p
n − p
k∇f kLp(Rn) ,
f
≤
(cid:13)(cid:13)(cid:13)(cid:13)
kxk(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
0 (Rn\{0}), and kxk =px2
f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
kx′f k
L
1
x′
p−1 (G)
≤
p
p
N − p
kf k2
L2(G) ≤(cid:13)(cid:13)(cid:13)(cid:13)
k∇Hf kLp(G) kx′f k
p
p−1 (G)
,
1 < p < N,
(3.5)
L
that is,
kf k2
L2(G) ≤
p
N − p
k∇H f kLp(G) kx′f k
L
p
p−1 (G)
,
1 < p < N.
(3.6)
In the abelian case G = (Rn, +), taking N = n, we obtain that (3.6) with p = 2
0 (Rn\{0}), we
implies the classical uncertainty principle for G ≡ Rn: for all f ∈ C ∞
have
(cid:18)ZRn
f (x)2dx(cid:19)2
≤(cid:18) 2
n − 2(cid:19)2ZRn
∇f (x)2dxZRn
kxk2f (x)2dx,
which is the Heisenberg-Pauli-Weyl uncertainly principle on Rn.
On the other hand, directly from the inequality (3.1), using the Holder inequality,
we can obtain a number of Heisenberg-Pauli-Weyl type uncertainly inequalities which
have various consequences and applications. For instance, when αp = α + β + 1, we
get
and if 0 = α + β + 1 and α = −p, then
N − αp
p
(cid:13)(cid:13)(cid:13)(cid:13)
1
p
f
Lp(G)
∇H f
≤(cid:13)(cid:13)(cid:13)(cid:13)
x′α(cid:13)(cid:13)(cid:13)(cid:13)
x′α(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)x′
p−1 −αf(cid:13)(cid:13)(cid:13)
x′(cid:13)(cid:13)(cid:13)(cid:13)
Lp(G) ≤ kx′p∇Hf kLp(G)(cid:13)(cid:13)(cid:13)(cid:13)
Lp(G)
p−1
f
,
N
p
kf kp
p−1
Lp(G)
,
(3.7)
(3.8)
both with sharp constants.
Proof of Theorem 3.1. We may assume that γ 6= N since for γ = N the inequality
(3.1) is trivial. By using the identity (2.5), the divergence theorem and Schwarz's
8
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
inequality one calculates
ZG
f (x)p
x′γ
dx =
pf (x)f (x)p−2 x′ · ∇Hf
dx
x′γ
x′ · ∇Hf dx
x′γ(cid:19) dx
p
p
1
1
x′γ
= −
N − γ
f (x)p−1
f (x)pdivH(cid:18) x′
N − γZG
ReZG
≤(cid:12)(cid:12)(cid:12)(cid:12)
N − γ(cid:12)(cid:12)(cid:12)(cid:12)ZG
N − γ(cid:12)(cid:12)(cid:12)(cid:12)ZG
≤(cid:12)(cid:12)(cid:12)(cid:12)
≤(cid:12)(cid:12)(cid:12)(cid:12)
N − γ(cid:12)(cid:12)(cid:12)(cid:12)(cid:18)ZG
dx ≤(cid:18)ZG
f (x)p−1
x′α+β
∇Hf (x)p
∇Hf (x)p
x′αp
x′αp
p
∇Hf (x) dx
p ZG
dx(cid:19) 1
p ZG
dx(cid:19) 1
f (x)p
βp
p−1
x′
p
dx! p−1
dx! p−1
.
p
f (x)p
βp
p−1
x′
.
(3.9)
Here we have used Holder's inequality in the last line. Thus, we arrive at
(cid:12)(cid:12)(cid:12)(cid:12)
N − γ
p
f (x)p
x′γ
(cid:12)(cid:12)(cid:12)(cid:12)ZG
This proves (3.1). Now it remains to show the sharpness of the constant. Let us
examine the equality condition in above Holder's inequality as in the abelian case
(see [18]). For this we consider the function
(3.10)
(3.11)
p−1 + 1 6= 0,
1
,
λ x′λ
x′C , α − β
λ := α − β
p−1 + 1 = 0,
g(x) =( e− C
(cid:12)(cid:12)(cid:12) and γ 6= N. Then it can be checked that
(cid:12)(cid:12)(cid:12)(cid:12)
N − γ(cid:12)(cid:12)(cid:12)(cid:12)
x′αp =
p ∇Hg(x)p
g(x)p
x′
βp
p−1
p
,
p
where C =(cid:12)(cid:12)(cid:12) N −γ
p (cid:12)(cid:12)(cid:12) is sharp.
C =(cid:12)(cid:12)(cid:12) N −γ
which satisfies the equality condition in Holder's inequality. It shows that the constant
(cid:3)
3.2. Badiale-Tarantello conjecture. The proof of Theorem 3.1 gives the following
similar statement in Rn.
Proposition 3.2. Let x = (x′, x′′) ∈ RN × Rn−N , 1 ≤ N ≤ n, and α, β ∈ R. Then
for any f ∈ C ∞
0 (Rn\{x′ = 0}), and all 1 < p < ∞, we have
N − γ
p
f
x′
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
γ
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Lp(Rn)
≤(cid:13)(cid:13)(cid:13)(cid:13)
1
x′α ∇f(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
f
β
p−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
x′
p−1
Lp(Rn)
,
(3.12)
where γ = α + β + 1 and x′ is the Euclidean norm on RN . If γ 6= N then the
constant N −γ
is sharp.
p
The proof is similar to the proof of Theorem 3.1. However, for the sake of com-
pleteness here we give the details.
pf (x)f (x)p−2 x′ · ∇N f
dx
x′γ
x′γ(cid:19) dx
x′ · ∇N f dx
0 · ∇f dx
∇f (x) dx
p
p
p
1
1
x′
x′γ
x′γ
= −
N − γ
f (x)p−1
f (x)p−1
f (x)pdivN(cid:18) x′
N − γZRn
ReZRn
N − γ(cid:12)(cid:12)(cid:12)(cid:12)ZRn
≤(cid:12)(cid:12)(cid:12)(cid:12)
N − γ(cid:12)(cid:12)(cid:12)(cid:12)ZRn
=(cid:12)(cid:12)(cid:12)(cid:12)
≤(cid:12)(cid:12)(cid:12)(cid:12)
N − γ(cid:12)(cid:12)(cid:12)(cid:12)ZRn
N − γ(cid:12)(cid:12)(cid:12)(cid:12)(cid:18)ZRn
≤(cid:12)(cid:12)(cid:12)(cid:12)
dx ≤(cid:18)ZRn
f (x)p−1
x′α+β
∇f (x)p
∇f (x)p
x′αp
x′αp
p
p ZRn
dx(cid:19) 1
f (x)p
βp
p−1
x′
p
dx! p−1
,
p ZRn
dx(cid:19) 1
f (x)p
βp
p−1
x′
p
dx! p−1
.
(3.13)
0 = (x′, 0) ∈ Rn, that is x′
where x′
0 = x′, ∇N is the standard gradient on RN , as
well as ∇ is the gradient on Rn. Here we have used Holder's inequality in the last
line. Thus, we arrive at
(cid:12)(cid:12)(cid:12)(cid:12)
N − γ
p
f (x)p
x′γ
(cid:12)(cid:12)(cid:12)(cid:12)ZRn
This proves (3.12). Now it remains to show the sharpness of the constant. Let us
examine the equality condition in above Holder's inequality. Consider
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES
9
Proof of Proposition 3.2. We may assume that γ 6= N since for γ = N the inequality
(3.12) is trivial. By using the identity
divN
x′
x′γ =
N − γ
x′γ
,
for all γ ∈ R and x′ ∈ RN with x′ 6= 0, where divN is the standard divergence on
RN , the divergence theorem and Schwarz's inequality one calculates
ZRn
f (x)p
x′γ
dx =
p−1 + 1 6= 0,
p
1
,
λ x′λ
x′C , α − β
λ := α − β
p−1 + 1 = 0,
g(x) =( e− C
where C =(cid:12)(cid:12)(cid:12) N −γ
(cid:12)(cid:12)(cid:12) and γ 6= N. Then it can be checked that
N − γ(cid:12)(cid:12)(cid:12)(cid:12)
x′αp =(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
p (cid:12)(cid:12)(cid:12) is sharp.
C =(cid:12)(cid:12)(cid:12) N −γ
N − γ(cid:12)(cid:12)(cid:12)(cid:12)
p ∇N g(x)p
p ∇g(x)p
p
p
x′αp =
that
N − p(α + 1)
p
(cid:13)(cid:13)(cid:13)(cid:13)
f
x′α+1(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
≤(cid:13)(cid:13)(cid:13)(cid:13)
1
x′α ∇f(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
g(x)p
βp
p−1
x′
,
which satisfies the equality condition in Holder's inequality. It shows that the constant
(cid:3)
As above, taking β = (α + 1)(p − 1) and γ = p(α + 1) the inequality (3.12) implies
(3.14)
(3.15)
, 1 < p < ∞,
(3.16)
10
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
for any f ∈ C ∞
α = 0 and 1 < p < N, 2 ≤ N ≤ n, the inequality (3.16) implies that
0 (Rn\{x′ = 0}) and for all α ∈ R with the sharp constant. When
(cid:13)(cid:13)(cid:13)(cid:13)
1
x′
f(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
≤
p
N − p
k∇f kLp(Rn) ,
(3.17)
again with p
N −p being the best constant. This proves the Badiale-Tarantello conjec-
ture, which is stated in the introduction (see also [2, Remark 2.3] for the original
statement).
3.3. Horizontal higher order versions. In this subsection we show how by it-
erating the established Lp-Caffarelli-Kohn-Nirenberg type inequalities one can get
inequalities of higher order. Putting ∇Hf instead of f and α − 1 instead of α in
(3.3) we consequently have
,
x′α−1 ∇2
Hf = ∇H∇Hf , that is, ∇m
Hf =
for each α ∈ R such that α 6= N
p − 1 and α 6= N
p . This iteration process gives
for α 6= N
∇H∇m−1
H f , m ∈ N. Combining it with (3.3) we get
Hf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
Hf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
1
x′α−1 ∇2
(cid:13)(cid:13)(cid:13)(cid:13)
f
f
1
p
p
p
1
≤
≤
N − p(α + 1)
x′θ+1−k ∇k
∇Hf
x′α(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
p . Here and after we understand ∇2
(cid:13)(cid:13)(cid:13)(cid:13)
N − pα(cid:13)(cid:13)(cid:13)(cid:13)
N − pα)(cid:13)(cid:13)(cid:13)(cid:13)
x′α+1(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
Hf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
x′θ+1(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
(cid:13)(cid:13)(cid:13)(cid:13)
0 (G\{x′ = 0}) and all θ ∈ R such thatQk−1
N − p(θ + 1 − j)#−1
(cid:13)(cid:13)(cid:13)(cid:13)
x′ϑ+1(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
H f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)
0 (G\{x′ = 0}) and all ϑ ∈ R such thatQm−1
N − p(ϑ + 1 − j)#−1
≤ Aθ,k(cid:13)(cid:13)(cid:13)(cid:13)
Aθ,k := pk"k−1Yj=0
≤ Aϑ,m(cid:13)(cid:13)(cid:13)(cid:13)
Aϑ,m := pm"m−1Yj=0
x′ϑ+1−m ∇m+1
∇Hf
1
.
.
for any f ∈ C ∞
and
Similarly, we have
for any f ∈ C ∞
and
,
(3.18)
, 1 < p < ∞,
(3.19)
j=0 N − p(θ + 1 − j) 6= 0,
, 1 < p < ∞,
(3.20)
j=0 N − p(ϑ + 1 − j) 6= 0,
Now putting ϑ + 1 = α and θ + 1 = β
(3.1) we obtain
p−1 into (3.20) and (3.19), respectively, from
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES 11
Proposition 3.3. Let 1 < p < ∞. For any k, m ∈ N we have
∇k
Hf(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p−1
,
Lp(G)
(3.21)
j=0 N − p(α − j) 6=
N − γ
p
for any f ∈ C ∞
0, and
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
γ
f
1
x′
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
0 (G\{x′ = 0}), γ = α+β+1, and α ∈ R such thatQm−1
x′α−m ∇m+1
1
β
p−1 −k
Lp(G)
x′
≤ eAα,meAβ,k(cid:13)(cid:13)(cid:13)(cid:13)
eAα,m := pm"m−1Yj=0
j=0(cid:12)(cid:12)(cid:12)N − p( β
eAβ,k := pk(p−1)"k−1Yj=0(cid:12)(cid:12)(cid:12)(cid:12)N − p(cid:18) β
H f(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
N − p(α − j)#−1
p−1 − j)(cid:12)(cid:12)(cid:12) 6= 0, and
#−(p−1)
− j(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
p − 1
4. Horizontal critical Hardy type inequality
.
,
as well as β ∈ R such thatQk−1
For p = N the inequality (3.4) fails for any constant (see, e.g., [19] and [32] for
discussions in Euclidean cases). However, we state the following theorem for the
(critical) case p = N.
Theorem 4.1. For a bounded domain Ω ⊂ G with 0 ∈ Ω and all f ∈ C ∞
we have
0 (Ω\{x′ = 0})
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
f (x)
x′log R
x′(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)LN (Ω)
≤
x′
x′
N
N − 1(cid:13)(cid:13)(cid:13)(cid:13)
· ∇Hf (x)(cid:13)(cid:13)(cid:13)(cid:13)LN (Ω)
where R = sup
x∈Ω
x′.
, 1 < N < ∞,
(4.1)
Note that below we give the proof of (4.1) for real-valued functions, the same
inequality follows for all complex-valued functions by using the identity (cf. Davies
[15, p. 176])
∀z ∈ C :
zp =(cid:18)Z π
−π
cos θpdθ(cid:19)−1Z π
−π
Re(z) cos θ + Im(z) sin θp dθ,
(4.2)
which follows from the representation z = r(cos φ+i sin φ) by some manipulations. To
prove (4.1) we follow the Euclidean setting from [46]. First let us prove the following
more abstract theorem, and then the proof of Theorem 4.1 will follow easily from
this.
Theorem 4.2. Let 0 ∈ Ω ⊂ G be a bounded domain. Let g : (1, ∞) → R be a
C 2-function such that
for all t > 1 and
g′(t) < 0,
g′′(t) > 0
(−g′(t))2(N −1)
(g′′(t))N −1 ≤ C < ∞,
∀t > 1.
(4.3)
(4.4)
12
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
Then we have
(cid:18) N − 1
N (cid:19)NZΩ
f (x)N
x′N (cid:18)−g′(cid:18)log
Re
Re
x′(cid:19)(cid:19)N −2
≤ZΩ(cid:16)−g′(cid:16)log Re
(cid:16)g′′(cid:16)log Re
g′′(cid:18)log
x′(cid:17)(cid:17)2(N −1)
x′(cid:17)(cid:17)N −1
x′(cid:19) dx
(cid:12)(cid:12)(cid:12)(cid:12)
x′
x′
x′.
N
· ∇H f (x)(cid:12)(cid:12)(cid:12)(cid:12)
dx,
(4.5)
for all f ∈ C ∞
0 (Ω\{x′ = 0}). Here R = sup
x∈Ω
Proof of Theorem 4.1. If we take
g(t) = −log(t − 1),
for t > 1, then we see that this function satisfies all assumptions of Theorem 4.2.
That is,
g′(t) = −
1
t − 1
< 0,
g′′(t) =
1
(t − 1)2
> 0,
(−g′(t))2(N −1)
(g′′(t))N −1 = 1,
∀t > 1.
(cid:3)
.
and
Therefore, putting
and
in (4.5) we obtain (4.1).
Remark 4.3. Taking
in (4.5) we obtain
g′(cid:18)log
g′′(cid:18)log
1
Re
log R
x′
x′(cid:19) = −
x′(cid:19) =
(cid:16)log R
x′(cid:17)2
Re
1
g(t) = e
N t
1−N , t > 1,
kf kLN (Ω) ≤(cid:13)(cid:13)(cid:13)(cid:13)x′
x′
x′
· ∇H f(cid:13)(cid:13)(cid:13)(cid:13)LN (Ω)
x′ using Schwarz's inequality we get
Since R = sup
x∈Ω
which is LN -Poincare inequality for the horizontal gradient.
kf kLN (Ω) ≤ R k∇Hf kLN (Ω) ,
Proof of Theorem 4.2. Let us introduce notations
Rǫ := sup
Fǫ(x) := log
x∈Ωpx′2 + 2ǫ2,
px′2 + ǫ2
Rǫe
,
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES 13
and
Gǫ(x) = g(Fǫ(x)),
for, say, ǫ > 0. Then a direct calculation shows
∇HGǫ(x)N −2∇HGǫ(x) = (−g′(Fǫ(x)))N −1(cid:18) x′N −2x′
(x′2 + ǫ2)N −1(cid:19)
and since g′(t) < 0, with LN as in (2.3),
LN Gǫ(x) = divH(∇HGǫ(x)N −2∇H Gǫ(x))
= (N − 1) (−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
x′N
(x′2 + ǫ2)N
+(N − 1) (−g′(Fǫ(x)))N −1 2ǫ2x′N −2
(x′2 + ǫ2)N
.
The divergence theorem gives
ZΩ
f N LN Gǫ(x)dx =ZΩ
Now on the one hand,
f N divH(∇HGǫ(x)N −2∇HGǫ(x))dx
= −ZΩ
∇Hf N · (∇HGǫ(x)N −2∇HGǫ(x))dx.
(4.6)
ZΩ
f N LN Gǫ(x)dx
= (N − 1)ZΩ
+ (N − 1)ZΩ
≥ (N − 1)ZΩ
f N (−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
x′N
(x′2 + ǫ2)N
dx
f N (−g′(Fǫ(x)))N −1 2ǫ2x′N −2
(x′2 + ǫ2)N
dx
f N (−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
x′N
(x′2 + ǫ2)N
dx.
(4.7)
14
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
On the other hand, using
(cid:12)(cid:12)(cid:12)(cid:12)−ZΩ
∇Hf N · (∇HGǫ(x)N −2∇H Gǫ(x))dx(cid:12)(cid:12)(cid:12)(cid:12)
f f N −2∇Hf · (∇HFǫ(x)N −2∇HFǫ(x))dx(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)NZΩ
(x′2 + ǫ2)N −1 (cid:19) dx(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)NZΩ
f (x)N −2f (x) (−g′(Fǫ(x)))N −1(cid:18) x′N −2x′ · ∇Hf
f (x)N −1 (−g′(Fǫ(x)))N −1(cid:18) x′N −2 x′ · ∇Hf
(x′2 + ǫ2)N −1 (cid:19) dx
= NZΩ
≤ N(cid:18)ZΩ
dx(cid:19) N −1
ZΩ
· ∇Hf(cid:12)(cid:12)(cid:12)(cid:12)
(−g′(Fǫ(x)))2(N −1) (g′′(Fǫ(x)))−(N −1)(cid:12)(cid:12)(cid:12)(cid:12)
(−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
x′N f (x)N
(x′2 + ǫ2)N
x′
x′
N
N
Combining all (4.6), (4.7) and (4.8) we arrive at
(N − 1)ZΩ
f N (−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
x′N
(x′2 + ǫ2)N
dx
x′N f (x)N
(x′2 + ǫ2)N
(−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
≤ N(cid:18)ZΩ
ZΩ
(−g′(Fǫ(x)))2(N −1) (g′′(Fǫ(x)))−(N −1)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18) N − 1
N (cid:19)NZΩ
f N (−g′(Fǫ(x)))N −2 g′′(Fǫ(x))
x′N
that is,
(x′2 + ǫ2)N
dx.
N
dx(cid:19) N −1
· ∇Hf(cid:12)(cid:12)(cid:12)(cid:12)
N
x′
x′
N
dx! 1
.
(4.8)
N
dx! 1
,
(4.9)
≤ZΩ
(−g′(Fǫ(x)))2(N −1) (g′′(Fǫ(x)))−(N −1)(cid:12)(cid:12)(cid:12)(cid:12)
x′
x′
N
· ∇Hf(cid:12)(cid:12)(cid:12)(cid:12)
dx,
(4.10)
(cid:3)
Now letting ǫ → 0 we obtain (4.5).
5. Horizontal Hardy-Rellich type inequalities and weighted versions
5.1. Hardy-Rellich type inequalities. We prove the following Hardy-Rellich type
inequalities on the stratified group G:
Theorem 5.1. Let G be a stratified group with N being the dimension of the first
stratum, and let 1 < p < N with 1
q = 1 and α, β ∈ R be such that
p + 1
p − N
p − 1
≤ γ := α + β + 1 ≤ 0.
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES 15
Then for all f ∈ C ∞
0 (G\{x′ = 0}) we have
N + γ(p − 1) − p
p
∇Hf
x′
γ
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
p
p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Lp(G)
≤(cid:13)(cid:13)(cid:13)(cid:13)
1
x′α Lpf(cid:13)(cid:13)(cid:13)(cid:13)Lp(G)(cid:13)(cid:13)(cid:13)(cid:13)
∇Hf
x′β(cid:13)(cid:13)(cid:13)(cid:13)Lq(G)
where · is the Euclidean norm on RN and Lp is the p-sub-Laplacian operator defined
by (2.3).
,
(5.1)
Corollary 5.2. When β = 0, α = −1 and q = p
stratified group Rellich type inequality for Lp:
p−1, the inequality (5.1) gives a
k∇Hf kLp(G) ≤
p
N − p
kx′Lpf kLp(G) , 1 < p < N,
(5.2)
for all f ∈ C ∞
0 (G\{x′ = 0}).
Corollary 5.3. When α = 0, β = −1, the inequality (5.1) implies the following
Heisenberg-Pauli-Weyl type uncertainly principle for Lp, 1 < p < N: for each f ∈
C ∞
0 (G\{x′ = 0}) we have
k∇Hf kp
Lp(G) ≤
p
N − p
kLpf kLp(G) kx′∇Hf kLq(G) ,
1
p
+
1
q
= 1.
(5.3)
Proof of Theorem 5.1. As in the proof of Theorem 3.1 we have
ZG
∇Hf (x)p
x′γ
dx =
1
N − γZG
N − γZG
1
p
2
= −
∇Hf (x)pdivH(cid:18) x′
x′γ(cid:19) dx
∇Hf (x)p−2 x′ · ∇H∇Hf (x)2
dx
x′γ
=
We also have
p
2(γ − N)ZG
∇Hf (x)p−2 x′ · ∇H ∇Hf (x)2
x′γ
dx.
(5.4)
ZG
Lpf
x′γ
x′ · ∇H f (x)dx =ZG
= −ZG
∇Hf (x)p−2 ∇Hf (x)2
x′γ
divH(∇Hf (x)p−2∇Hf (x))
x′ · ∇H f (x)dx
x′γ
∇Hf (x)p−2∇H f (x) · ∇H(cid:18) x′ · ∇H f (x)
x′γ
+
x′ · ∇H∇Hf (x)2
2x′γ
−
γ x′ · ∇Hf (x)2
x′γ+2
(cid:19) dx
! dx,
= −ZG
that is,
∇Hf (x)p−2
ZG
∇Hf (x)p−2 x′ · ∇Hf (x)2
x′γ
x′γ+2
= 2γZG
x′ · ∇H ∇Hf (x)2dx
dx − 2ZG
∇Hf (x)p
x′γ
dx
−2ZG
Lpf
x′γ
x′ · ∇H f (x)dx.
16
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
Putting this in the right hand side of (5.4) we obtain
ZG
−
∇Hf (x)p
x′γ
p
γ − N ZG
dx =
pγ
γ − N ZG
∇Hf (x)p
x′γ
dx −
∇Hf (x)p−2 x′ · ∇Hf (x)2
γ − N ZG
Lpf
x′γ
x′γ+2
x′ · ∇Hf (x)dx.
p
dx
Thus,
ZG
Lpf
x′γ
ZG
Lpf
x′γ
=
x′ · ∇H f (x)dx
=
N − p − γ
p
x′ · ∇H f (x)dx
ZG
∇Hf (x)p
x′γ
dx + γZG
∇Hf (x)p−2 x′ · ∇H f (x)2
x′γ+2
dx.
(5.5)
Since γ ≤ 0, using Schwarz's inequality to the last integrants we get
N − p − γ
∇Hf (x)p
ZG
p
p
≥
x′γ
ZG
N − p − γ
∇Hf (x)p
dx + γZG
x′γ
=
∇Hf (x)p−2 x′ · ∇Hf (x)2
dx + γZG
∇Hf (x)p
x′γ+2
x′γ
dx
N + γ(p − 1) − p
∇Hf (x)p
dx
p
dx.
(5.6)
x′γ
ZG
On the other hand, again using Schwarz's inequality and Holder's inequality we have
ZG
Lpf
x′γ
x′ · ∇H f (x)dx ≤ZG
Lpf
x′γ−1 ∇Hf (x) dx
Lpf
p
≤(cid:18)ZG(cid:12)(cid:12)(cid:12)(cid:12)
x′α(cid:12)(cid:12)(cid:12)(cid:12)
p(cid:18)ZG(cid:12)(cid:12)(cid:12)(cid:12)
dx(cid:19) 1
∇H f
x′β(cid:12)(cid:12)(cid:12)(cid:12)
q
q
dx(cid:19) 1
.
(5.7)
(cid:3)
Combining it with (5.6) we prove Theorem 5.1.
5.2. Weighted versions. To give an idea for obtaining more general improved
weighted Hardy type inequalities let us conclude this paper with the following very
short discussion of techniques from [47] (see also [24] and [42]), now in the setting of
stratified groups.
Consider the following weighted p-sub-Laplacian
where 0 ≤ ρ ∈ C 1(G).
Lp,ρf = divH(cid:0)ρ(x)∇H f p−2∇Hf(cid:1) ,
1 < p < ∞,
(5.8)
Theorem 5.4. Let 2 ≤ p < ∞. Let 0 ≤ F ∈ C ∞(G) and 0 ≤ η ∈ L1
that
loc(G) be such
ηF p−1 ≤ −Lp,ρF,
a.e. in G.
Then we have
kη
1
p f kp
Lp(G) + Cp(cid:13)(cid:13)(cid:13)(cid:13)ρ
1
p F ∇H
p
Lp(G)
≤ kρ
f
F(cid:13)(cid:13)(cid:13)(cid:13)
1
p ∇Hf kp
Lp(G),
(5.9)
(5.10)
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES 17
for all real-valued functions f ∈ C ∞
0 (G). Here Cp is a positive constant.
Proof of Theorem 5.4. For all x, y ∈ Rn there exists a positive number Cp such that
(5.11)
xp + Cpyp + pxp−2x · y ≤ x + yp,
2 ≤ p < ∞.
Thus,
gp∇HF p + CpF p∇Hgp + F ∇HF p−2∇HF · ∇Hgp
≤ g∇HF + F ∇Hgp = ∇Hf p,
(5.12)
where g = f
F . It follows that
ρ(x)∇H F (x)pg(x)pdx
ρ(x)∇Hg(x)pF (x)pdx
Using (5.9) this implies that
ZG
ρ(x)∇Hf (x)pdx ≥ZG
+ CpZG
−ZG
≥ CpZG
+ZG
η(x)g(x)pF (x)pdx + CpZG
Lp(G) + Cp(cid:13)(cid:13)(cid:13)(cid:13)ρ
F(cid:13)(cid:13)(cid:13)(cid:13)
F we arrive at
(cid:13)(cid:13)(cid:13)(cid:13)ρ
2 F ∇H
p f kp
L2(G)
kη
f
2
1
1
ZG
Since g = f
proving (5.10).
divH (ρ(x)F (x)∇HF (x)p−2∇HF (x))g(x)pdx
ρ(x)∇H g(x)pF (x)pdx
−divH(ρ(x)∇H F (x)p−2∇H F (x))F (x)g(x)pdx.
ρ(x)∇Hg(x)pF (x)pdx ≤ZG
ρ(x)∇Hf (x)pdx.
1
p F ∇H
p
Lp(G)
≤ kρ
f
F(cid:13)(cid:13)(cid:13)(cid:13)
1
p ∇Hf kp
Lp(G),
(5.13)
(5.14)
(cid:3)
Remark 5.5. For p = 2 there is equality in (5.11) with C2 = 1, that is, the above
proof gives the following remainder formula
= kρ
1
2 ∇Hf k2
L2(G) − kη
1
2 f k2
L2(G).
(5.15)
Remark 5.6. For 1 < p < 2 the inequality (5.11) can be stated as for all x, y ∈ Rn
there exists a positive number Cp (see e.g. [37, Lemma 4.2]) such that
xp + Cp
yp
(x + y)2−p + pxp−2x · y ≤ x + yp,
1 < p < 2.
(5.16)
In turn, from the proof it follows that
kη
1
p f kp
Lp(G) + Cp(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ρ
f
F
1
2(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)
∇HF(cid:12)(cid:12)(cid:12)(cid:12) + F(cid:12)(cid:12)(cid:12)(cid:12)∇H
2
f
F(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) p−2
F ∇H
≤ kρ
2
L2(G)
f
F(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1
p ∇Hf kp
Lp(G)
(5.17)
18
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
for all real-valued f ∈ C ∞
0 (G).
Proposition 5.7. For f ∈ C ∞
0 (G) we have
≤
p
θ − p − 2
k∇H f kLp ,
1 < p < θ − 2, θ ≤ 2 + N, θ ∈ R.
(5.18)
(cid:13)(cid:13)(cid:13)(cid:13)
f
x′(cid:13)(cid:13)(cid:13)(cid:13)Lp
,
2p
p−2∇Hx′
n + ǫ)2(cid:1)− θ−p−2
p (cid:17)
− θ + 2 + N(cid:19) x′
(−θ + 2 + N)! x′
ǫ− θ−p−2
p−2
ǫ− (θ−p−2)(p−1)
p
−p
ǫ− (θ−p−2)(p−1)
p
−p.
Proof of Proposition 5.7. In Theorem 5.4 taking ρ = 1 and
for a given ǫ > 0, using the identity (2.4) we get
1 + ǫ)2 + . . . + (x′
p
p
p
p
p
=
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
θ − p − 2
θ − p − 2
θ − p − 2
Fǫ = x′
ǫ− θ−p−2
ǫ− θ−p−2
p =(cid:0)(x′
− Lp,1Fǫ = −divH(cid:0)∇HFǫp−2∇H Fǫ(cid:1)
= −divH(cid:16)∇Hx′
(cid:12)(cid:12)(cid:12)(cid:12)
p−2(cid:18) θ − p − 2
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
− Lp,1Fǫ ≥(cid:12)(cid:12)(cid:12)(cid:12)
η(x) =(cid:12)(cid:12)(cid:12)(cid:12)
= (cid:12)(cid:12)(cid:12)(cid:12)
θ − p − 2
θ − p − 2
+
p
p
p
p
p
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
If 1 < p < θ − 2 and θ ≤ 2 + N, then (5.19) gives
p
It shows that (5.10) (and (5.17)) implies (5.18).
θ − p − 2
1
x′
ǫp
F p−1
ǫ
,
θ − p − 2
p
1
x′
ǫp
.
that is, according to the assumption in Theorem 5.4, we can put
(5.19)
(5.20)
(cid:3)
Note that in the case of the Heisenberg group (5.18) was proved by D'Ambrosio in
[11]. Here it is worth to recall that on the Heisenberg group we have Q = 2 + N.
References
[1] Adimurthi and A. Sekar. Role of the fundamental solution in Hardy-Sobolev-type inequalities.
Proc. Roy. Soc. Edinburgh Sect. A, 136(6):1111 -- 1130, 2006.
[2] N. Badiale and G. Tarantello. A Sobolev-Hardy inequality with applications to a nonlinear
elliptic equation arising in astrophysics. Arch. Ration. Mech. Anal., 163:259 -- 293, 2002.
[3] H. Brezis and E. Lieb. Inequalities with remainder terms. J. Funct. Anal., 62:73 -- 86, 1985.
[4] H. Brezis and L. Nirenberg. Positive solutions of nonlinear elliptic equations involving critical
Sobolev exponents. Comm. Pure Appl. Math., 36:437 -- 477, 1983.
[5] H. Brezis and M. Marcus. Hardy's inequalities revisited. Ann. Scuola Norm. Sup. Pisa Cl. Sci.,
25(4):217 -- 237, 1997.
[6] L. Caffarelli, R. Kohn, and L. Nirenberg. First order interpolation inequalities with weights.
Compositio Mathematica, 53:259 -- 275, 1984.
[7] F. Catrina and Z. Q. Wang. On the Caffarelli-Kohn-Nirenberg inequalities: sharp constants,
existence (and non existence), and symmetry of extremals functions. Comm. Pure Appl. Math.,
54:229 -- 258, 2001.
HORIZONTAL HARDY, RELLICH AND CAFFARELLI-KOHN-NIRENBERG INEQUALITIES 19
[8] J. Chern and C. Lin. Minimizers of Caffarelli-Kohn-Nirenberg inequalities with the singularity
on the boundary. Arch. Ration. Mech. Anal., 197:401 -- 432, 2010.
[9] P. Ciatti, M. G. Cowling, and F. Ricci. Hardy and uncertainty inequalities on stratified Lie
groups. Adv. Math., 227:365 -- 387, 2015.
[10] D. G. Costa. Some new and short proofs for a class of Caffarelli-Kohn-Nirenberg type inequal-
ities. J. Math. Anal. Appl., 337:311 -- 317, 2008.
[11] L. D'Ambrosio. Some Hardy inequalities on the Heisenberg group. Differential Equations,
40(4):552 -- 564, 2004.
[12] L. D'Ambrosio. Hardy-type inequalities related to degenerate elliptic differential operators.
Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 4(3):451 -- 486, 2005.
[13] L. D'Ambrosio and S. Dipierro. Hardy inequalities on Riemannian manifolds and applications.
Ann. Inst. H. Poincar´e Anal. Non Lin´eaire, 31(3):449 -- 475, 2014.
[14] D. Danielli, N. Garofalo, and N. C. Phuc. Hardy-Sobolev type inequalities with sharp constants
in Carnot-Carath´eodory spaces. Potential Anal., 34(3):223 -- 242, 2011.
[15] E. B. Davies. One-parameter semigroups, volume 15 of London Mathematical Society Mono-
graphs. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], London-New York, 1980.
[16] E. B. Davies. A review of Hardy inequalities. In The Maz'ya anniversary collection, Vol. 2
(Rostock, 1998), volume 110 of Oper. Theory Adv. Appl., pages 55 -- 67. Birkhauser, Basel, 1999.
[17] E. B. Davies and A. M. Hinz. Explicit constants for Rellich inequalities in Lp(Ω). Math. Z.,
227(3):511 -- 523, 1998.
[18] Y. Di, L. Jiang, S.Shen, and Y. Jin. A note on a class of Hardy-Rellich type inequalities. J.
Inequal. Appl., 84:1 -- 6, 2013.
[19] D. E. Edmunds and H. Triebel. Sharp Sobolev embeddings and related Hardy inequalities: the
critical case. Math. Nachr., 207:79 -- 92, 1999.
[20] T. Ekholm, H. Kovar´ık, and A. Laptev. Hardy inequalities for p-Laplacians with Robin bound-
ary conditions. Nonlinear Anal., 128:365 -- 379, 2015.
[21] V. Fischer and M. Ruzhansky. Quantization on nilpotent Lie groups, volume 314 of Progress in
Mathematics. Birkhauser, 2016.
[22] G. B. Folland. Subelliptic estimates and function spaces on nilpotent Lie groups. Ark. Mat.,
13(2):161 -- 207, 1975.
[23] G. B. Folland and E. M. Stein. Hardy spaces on homogeneous groups, volume 28 of Mathematical
Notes. Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1982.
[24] R. Frank and R. Seiringer. Non-linear ground state representations and sharp Hardy inequali-
ties. J. Funct. Anal., 225(12):3407 -- 3430, 2008.
[25] N. Garofalo and E. Lanconelli. Frequency functions on the Heisenberg group, the uncertainty
principle and unique continuation. Ann. Inst. Fourier (Grenoble), 40(2):313 -- 356, 1990.
[26] J. A. Goldstein and I. Kombe. The Hardy inequality and nonlinear parabolic equations on
Carnot groups. Nonlinear Anal., 69(12):4643 -- 4653, 2008.
[27] G. Grillo. Hardy and Rellich-type inequalities for metrics defined by vector fields. Potential
Anal., 18(3):187 -- 217, 2003.
[28] G. H. Hardy. Notes on some points in the integral calculus. Messenger Math., 48:107 -- 112, 1919.
[29] G. H. Hardy. Note on a theorem of Hilbert. Math. Z., 6(3-4):314 -- 317, 1920.
[30] M. Hoffmann-Ostenhof, T. Hoffmann-Ostenhof, A. Laptev, and J. Tidblom. Many-particle
Hardy inequalities. J. Lond. Math. Soc. (2), 77(1):99 -- 114, 2008.
[31] T. Hoffmann-Ostenhof and A. Laptev. Hardy inequalities with homogeneous weights. J. Funct.
Anal., 268(11):3278 -- 3289, 2015.
[32] N. Ioku, M. Ishiwata, and T. Ozawa. Sharp remainder of a critical Hardy inequality. Arch.
Math. (Basel), 106(1):65 -- 71, 2016.
[33] Y. Jin and S. Shen. Weighted Hardy and Rellich inequality on Carnot groups. Arch. Math.
(Basel), 96(3):263 -- 271, 2011.
[34] A. E. Kogoj and S. Sonner. Hardy type inequalities for ∆λ-Laplacians. Complex Var. Elliptic
Equ., 61(3):422 -- 442, 2016.
20
MICHAEL RUZHANSKY AND DURVUDKHAN SURAGAN
[35] I. Kombe and M. Ozaydin. Hardy-Poincar´e, Rellich and uncertainty principle inequalities on
Riemannian manifolds. Trans. Amer. Math. Soc., 365(10):5035 -- 5050, 2013.
[36] B. Lian. Some sharp Rellich type inequalities on nilpotent groups and application. Acta Math.
Sci. Ser. B Engl. Ed., 33(1):59 -- 74, 2013.
[37] P. Lindqvist. On the equation div(∇up−2∇u)+λup−2u = 0, Proc. Amer. Math. Soc., 109:157 --
164, 1990.
[38] P. Niu, H. Zhang, and Y. Wang. Hardy type and Rellich type inequalities on the Heisenberg
group. Proc. Amer. Math. Soc., 129(12):3623 -- 3630, 2001.
[39] T. Ozawa, M. Ruzhansky and D. Suragan. Lp-Caffarelli-Kohn-Nirenberg type inequalities on
homogeneous groups. arXiv:1605.02520, 2016.
[40] M. Ruzhansky and D. Suragan. Layer potentials, Kac's problem, and refined Hardy inequality
on homogeneous Carnot groups. arXiv:1512.02547, 2015.
[41] M. Ruzhansky and D. Suragan. Critical Hardy inequality on homogeneous groups. arXiv:
1602.04809, 2016.
[42] M. Ruzhansky and D. Suragan. Local Hardy and Rellich inequalities for sums of squares of
vector fields. arXiv: 1601.06157, 2016.
[43] M. Ruzhansky and D. Suragan. Hardy and Rellich inequalities, identities, and sharp remainders
on homogeneous groups. arXiv:1603.06239, 2016.
[44] M. Ruzhansky and D. Suragan. Uncertainty relations
on nilpotent Lie
groups.
arXiv:1604.06702, 2016.
[45] S. Secchi, D. Smets and M. Willen. Remarks on a Hardy-Sobolev inequality. C. R. Acad. Sci.
Paris, Ser. I, 336:811 -- 815, 2003.
[46] F. Takahashi. A simple proof of Hardy's inequality in a limiting case. Arch. Math. (Basel),
104:77 -- 82, 2015.
[47] A. Yener. Weighted Hardy type inequalities on the Heisenberg group Hn. Math. Ineq. Appl.,
19(2):671 -- 683, 2016.
Michael Ruzhansky:
Department of Mathematics
Imperial College London
180 Queen's Gate, London SW7 2AZ
United Kingdom
E-mail address [email protected]
Durvudkhan Suragan:
Institute of Mathematics and Mathematical Modelling
125 Pushkin str.
050010 Almaty
Kazakhstan
and
Department of Mathematics
Imperial College London
180 Queen's Gate, London SW7 2AZ
United Kingdom
E-mail address [email protected]
|
1608.06720 | 2 | 1608 | 2016-10-13T07:10:46 | Orthogonal projectors onto spaces of periodic splines | [
"math.FA"
] | The main result of this paper is a proof that for any integrable function $f$ on the torus, any sequence of its orthogonal projections $(\widetilde{P}_n f)$ onto periodic spline spaces with arbitrary knots $\widetilde{\Delta}_n$ and arbitrary polynomial degree converges to $f$ almost everywhere with respect to the Lebesgue measure, provided the mesh diameter $|\widetilde{\Delta}_n|$ tends to zero. We also give a proof of the fact that the operators $\widetilde{P}_n$ are bounded on $L^\infty$ independently of the knots $\widetilde{\Delta}_n$. | math.FA | math |
ORTHOGONAL PROJECTORS ONTO SPACES OF PERIODIC
SPLINES
MARKUS PASSENBRUNNER
Abstract. The main result of this paper is a proof that for any integrable
gree converges to f almost everywhere with respect to the Lebesgue measure,
function f on the torus, any sequence of its orthogonal projections (ePnf ) onto
periodic spline spaces with arbitrary knots e∆n and arbitrary polynomial de-
provided the mesh diameter e∆n tends to zero. We also give a proof of the fact
that the operators ePn are bounded on L∞ independently of the knots e∆n.
1. Introduction
1.1. Splines on an interval. In this article we prove some results about the
periodic spline orthoprojector. In order to achieve this, we rely on existing results
for the non-periodic spline orthoprojector on a compact interval, so we first de-
scribe some of those results for the latter operator. Let k ∈ N and ∆ = (ti)r+k
i=ℓ a
knot sequence satisfying
ti < ti+k,
ti ≤ ti+1,
tℓ = · · · = tℓ+k−1,
tr+1 = · · · = tr+k.
Associated to this knot sequence, we define (Ni)r
normalized B-spline functions of order k on ∆ that have the properties
i=ℓ as the sequence of L∞-
supp Ni = [ti, ti+k],
Ni ≥ 0,
Ni ≡ 1.
rXi=ℓ
We write ∆ = maxℓ≤j≤r(tj+1−tj) for the maximal mesh width of the partition ∆.
Then, define the space Sk(∆) as the set of polynomial splines of order k (or at
most degree k−1) with knots ∆, which is the linear span of the B-spline functions
(Ni)r
i=ℓ. Moreover, let P∆ be the orthogonal projection operator onto the space
f (x)g(x) dx,
Sk(∆) with respect to the ordinary (real) inner product hf, gi =R tr+1
i.e.,
tℓ
The operator P∆ is also given by the formula
hP∆f, si = hf, si
for all s ∈ Sk(∆).
(1.1)
P∆f =
rXi=ℓ
hf, NiiN ∗
i ,
Date: May 11, 2018.
2010 Mathematics Subject Classification. 40A05, 41A15, 46E30.
Key words and phrases. Periodic splines, Almost everywhere convergence.
The author is supported by the FWF, project number P27723.
1
2
M. PASSENBRUNNER
i )r
i=ℓ denotes the dual basis to (Ni) defined by the relations hN ∗
where (N ∗
i , Nji = 0
when j 6= i and hN ∗
i , Nii = 1 for all i = ℓ, . . . , r. A famous theorem by A. Shadrin
states that the L∞ norm of this projection operator is bounded independently of
the knot sequence ∆:
Theorem 1.1 ([8]). There exists a constant ck depending only on the spline order
k such that for all knot sequences ∆ = (ti)r+k
i=ℓ as above,
kP∆ : L∞[tℓ, tr+1] → L∞[tℓ, tr+1]k ≤ ck.
We are also interested in the following equivalent formulation of this theorem,
which is proved in [1]: for a knot sequence ∆, let (aij) be the matrix (hN ∗
j i),
which is the inverse of the banded matrix (hNi, Nji). Then, the assertion of The-
orem 1.1 is equivalent to the existence of two constants K0 > 0 and γ0 ∈ (0, 1)
only depending on the spline order k such that
i , N ∗
(1.2)
aij ≤
,
ℓ ≤ i, j ≤ r,
0
K0γi−j
max{κi, κj}
where κi denotes the length of supp Ni. The proof of this equivalence uses Demko's
theorem [4] on the geometric decay of inverses of band matrices and de Boor's
stability (see [2] or [5, Chapter 5, Theorem 4.2]) which states that for 0 < p ≤
∞, the Lp norm of a B-spline series is equivalent to a weighted ℓp norm of its
coefficients, i.e. there exists a constant Dk only depending on the spline order k
such that:
Dkk−1/p(cid:16)Xj
cjpκj(cid:17)1/p
≤(cid:13)(cid:13)(cid:13)Xj
cjNj(cid:13)(cid:13)(cid:13)Lp ≤(cid:16)Xj
cjpκj(cid:17)1/p
.
In fact, for aij, we actually have the following improvement of (1.2) (see [6]):
There exist two constants K > 0 and γ ∈ (0, 1) that depend only on the spline
order k such that
(1.3)
aij ≤
Kγi−j
hij
,
ℓ ≤ i, j ≤ r,
where hij denotes the length of the convex hull of supp Ni∪supp Nj. This inequal-
ity can be used to obtain almost everywhere convergence for spline projections of
L1 functions:
Theorem 1.2 ([6]). For all f ∈ L1[tℓ, tr+1] there exists a subset A ⊂ [tℓ, tr+1] of
full Lebesgue measure such that for all sequences (∆n) of partitions of [tℓ, tr+1]
such that ∆n → 0, we have
lim
n→∞
P∆nf (x) = f (x),
x ∈ A.
Our aim in this article is to prove an analogue of Theorem 1.2 for orthopro-
jectors on periodic spline spaces. In this case, we do not have a periodic version
of (1.3) at our disposal, since the proof of this inequality does not carry over to
the periodic setting. However, by comparing orthogonal projections onto periodic
spline spaces to suitable non-periodic projections, we are able to obtain a periodic
version of Theorem 1.2.
In the course of the proof of the periodic version of Theorem 1.2, we also need
a periodic version of Theorem 1.1, which can be proved by first establishing the
ORTHOGONAL PROJECTORS ONTO SPACES OF PERIODIC SPLINES
3
same assertion for infinite point sequences and then by viewing periodic functions
as defined on the whole real line [A. Shadrin, private communication]. The proof of
Theorem 1.1 for infinite point sequences is announced in [8] and carried out [3]. In
this article we give a different proof of the periodic version of Shadrin's theorem by
employing a similar comparison of periodic and non-periodic projection operators
as in the proof of the periodic version of Theorem 1.2. This proof directly passes
from the interval case to the periodic result without recourse to infinite point
sequences.
1.2. Periodic splines. Let n ≥ k be a natural number and e∆ = (sj)n−1
j=0 be a
sequence of distinct points on the torus T = R/Z identified canonically with [0, 1),
such that for all j we have
and we extend (sj)n−1
sj ≤ sj+1,
j=0 periodically by
sj < sj+k,
srn+j = r + sj
for r ∈ Z \ {0} and 0 ≤ j ≤ n − 1.
Now, the main result of this article reads as follows:
0, we have
lim
n→∞ePnf (x) = f (x),
x ∈ eA,
In order to prove this result, we also need a periodic version of Theorem 1.1:
Theorem 1.3. For all functions f ∈ L1(T) there exists a set eA of full Lebesgue
measure such that for all sequences of partitions (e∆n) on T as above with e∆n →
where ePn denotes the orthogonal projection operator onto the periodic spline space
of order k with knots e∆n.
k such that for all knot sequences e∆ = (sj)n−1
projection operator eP satisfies the inequality
between the periodic projection operator eP and the non-periodic projection op-
erator P for certain non-periodic point sequences associated to e∆ = (si)n−1
Theorem 1.4. There exists a constant ck depending only on the spline order
j=0 on T, the associated orthogonal
The article is organized as follows. In Section 2, we prove a simple lemma on the
growth behaviour of linear combinations of non-periodic B-spline functions which
is frequently needed later in the proofs of both Theorem 1.3 and Theorem 1.4.
Section 3 is devoted to the proof of Theorem 1.4, which is needed for the proof
of Theorem 1.3 in Section 4. Finally, in Section 5, we also apply our method of
proof to recover Shadrin's theorem for infinite point sequences (see [3, 8]).
keP : L∞(T) → L∞(T)k ≤ ck.
The idea of the proofs of Theorems 1.3 and 1.4 is to estimate the difference
i=0 .
4
M. PASSENBRUNNER
2. A simple upper estimate for B-spline sums
Let A be a subset of [tℓ, tr+1]. Then, define the set of indices i(A) whose B-
spline supports intersect with A as
i(A) := {i : A ∩ supp Ni 6= ∅}.
We also write i(x) for i({x}). If we have two subsets U, V of indices, we write
d(U, V ) for the distance between U and V induced by the metric d(i, j) = i− j.
We will use the notation A(t) . B(t) to indicate the existence of a constant C
that depends only on the spline order k such that for all t we have A(t) ≤ CB(t),
where t denotes all explicit or implicit dependencies that the expressions A and
B might have.
The fact that B-spline functions are localized, so a fortiori the set i(x) is
localized for any x ∈ [tℓ, tr+1], can be used to derive the following lemma:
Lemma 2.1. Let J be a subset of the index set {ℓ, ℓ + 1, . . . , r − 1, r}, f =
j and p ∈ [1,∞]. Then, for all x ∈ [tℓ, tr+1], we have the estimate
Pj∈Jhh, NjiN ∗
κ1/p′
j
hjm
m∈i(x),j∈J
f (x) . γd(i(x),J)khkp max
≤ γd(i(x),J)khkp max
≤ γd(i(x),J)khkp · I(x)−1/p,
m∈i(x),j∈J(cid:0) max{κm, κj}(cid:1)−1/p
1 ≤ p ≤ ∞,
where γ ∈ (0, 1) is the constant appearing in (1.3), I(x) is the interval I = [ti, ti+1)
containing the point x and the exponent p′ is such that 1/p + 1/p′ = 1.
Proof. Since N ∗
j =Pm ajmNm,
This implies
f (x) =Xj∈J Xm∈i(x)
m∈i(x)(cid:16)Xj∈J
f (x) . max
ajmhh, NjiNm(x).
γj−m
hjm khkpkNjkp′(cid:17),
where we used inequality (1.3) for ajm, Holder's inequality with the conjugate
exponent p′ = p/(p − 1) to p and the fact that the B-spline functions Nm form a
partition of unity. Using again the uniform boundedness of Nj, we obtain
f (x) . max
m∈i(x)(cid:16)Xj∈J
j (cid:17).
γj−m
hjm khkpκ1/p′
Summing the geometric series now yields the first estimate. The second and the
third estimate are direct consequences of the first one.
(cid:3)
Remark 2.2. We note that we directly obtain the second estimate in the above
lemma if we use the weaker inequality (1.2) instead of (1.3).
ORTHOGONAL PROJECTORS ONTO SPACES OF PERIODIC SPLINES
5
3. The periodic spline orthoprojector is uniformly bounded on L∞
In this section, we give a direct proof of Theorem 1.4 on the boundedness of
periodic spline projectors without recourse to infinite knot sequences. Here, we
will only use the geometric decay of the matrix (ajm) defined above for splines
on an interval.
A vital tool in the proofs of both Theorem 1.1 and Theorem 1.2 are B-spline
functions. We will also make extensive use of them and introduce their periodic
version, cf. [7]. Associated to the periodic point sequence (sj)n−1
j=0 and its periodic
extension as in Section 1.2 we define the non-periodic point sequence
tj = sj,
for j = −k + 1, . . . , n + k − 1
and denote the corresponding non-periodic B-spline functions by (Nj)n−1
supp Nj = [tj, tj+k]. Then we define for x ∈ [0, 1)
j=−k+1 with
if we canonically identify T with [0, 1). Moreover, for j = n − k + 1, . . . , n − 1,
eNj(x) = Nj(x),
eNj(x) =(Nj−n(x),
Nj(x),
j = 0, . . . , n − k,
if x ∈ [0, sj],
if x ∈ (sj, 1).
splines of order k with knots (sj)n−1
We denote by eP the orthogonal projection operator onto the space of periodic
functions (eNj)n−1
j=0 and similarly to the non-periodic case we define
j=0 , which is the linear span of the B-spline
A ⊂ T.
Lemma 3.1. Let fi be a function on T with supp fi ⊂ [si, si+1] for some index i
in the range 0 ≤ i ≤ n − 1. Then, for any x ∈ T,
i(A) = {0 ≤ j ≤ n − 1 : A ∩ supp eNj 6= ∅},
ed(i(x),i(supp fi))kfik∞,
eP fi(x) . γ
where ed is the distance function induced by the canonical metric in Z/nZ and
γ ∈ (0, 1) is the constant appearing in inequality (1.3).
Proof. We assume that the index i is chosen such that si < si+1, since if si = si+1,
the function fi is identically zero in L∞.
Given a function f on T, we associate a non-periodic function T f defined on
[si, si+n+1] given by
T f (t) = f (π(t)),
t ∈ [si, si+n+1],
where π(t) is the quotient mapping from R to T. We observe that T is a linear oper-
ator, kT : L2(T) → L2([si, si+n+1])k = √2 and kT : L∞(T) → L∞([si, si+n+1])k =
1. Moreover, for x ∈ T, let r(x) be the representative of x in the interval [si, si+n).
We want to estimate eP fi(x). In order to do this, we first decompose
eP fi(x) = TeP fi(r(x)) = P T fi(r(x)) + (TeP fi − P T fi)(r(x)),
(3.1)
6
M. PASSENBRUNNER
where P is the orthogonal projection operator onto the space of splines of order
k corresponding to the point sequence ∆ = (tj)n+k
j=−k+1 associated to the non-
periodic grid points in the interval [si, si+n+1], i.e.,
tj = si+j,
j = 0, . . . , n + 1,
t−k+1 = · · · = t−1 = si,
i=−k+1 be the L∞-normalized B-spline basis corresponding to this
tn+2 = · · · = tn+k = si+n+1.
Also, let (Nj)n
point sequence.
We estimate the first term P T fi(r(x)) from the decomposition in (3.1) of
eP fi(x). Since P is a projection operator onto splines on an interval, we use rep-
resentation (1.1) to get
P T fi(r(x)) =
hT fi, NjiN ∗
j (r(x)),
nXj=−k+1
and, since supp T fi ⊂ [si, si+1]∪ [si+n, si+n+1] = [t0, t1]∪ [tn, tn+1] by definition of
fi and T and supp Nj ⊂ [tj, tj+k] for all j = −k + 1, . . . , n,
j (r(x)),
P T fi(r(x)) =Xj∈J1
hT fi, NjiN ∗
with J1 = {−k + 1, . . . , 0} ∪ {n − k + 1, . . . , n}. Employing now Lemma 2.1 with
p = ∞ to this sum, we obtain
(3.2)
P T fi(r(x)) . γd(i(r(x)),J1)kT fik∞ . γ
ed(i(x),i(supp fi))kfik∞.
Now we turn to the second term on the right hand side of (3.1). Let g :=
(TeP − P T )fi. Observe that g ∈ Sk(∆) since the range of both TeP and P is
contained in Sk(∆). Moreover,
h(TeP − T )fi, Nji = heP fi − fi, eNj+ii,
where we take the latter subindex j + i to be modulo n. This equation is true in
the given range of the parameter j, since in this case, the functions Nj and eNj+i
coincide on their supports. The fact that eP is an orthogonal projection onto the
span of the functions (eNj)n−1
j=0 then implies
j = 0, . . . , n − k + 1.
hTeP fi − T fi, Nji = heP fi − fi, eNj+ii = 0,
Combining this with the fact
j = 0, . . . , n − k + 1,
hP T fi − T fi, Nji = 0,
j = −k + 1, . . . , n,
since P is an orthogonal projection onto a spline space as well, we obtain that
Therefore, we can expand g as a B-spline sum
hg, Nji = 0,
j = 0, . . . n − k + 1.
g =Xj∈J2
hg, NjiN ∗
j ,
ORTHOGONAL PROJECTORS ONTO SPACES OF PERIODIC SPLINES
7
with J2 = {−k + 1, . . . ,−1}∪ {n− k + 2, . . . , n}. Now, we employ Lemma 2.1 on
the function g with the parameter p = 2 to get for the point y = r(x)
g(y) . γd(i(y),J2)kgk2 max
j∈J2 supp Nj−1/2.
get
Since g = (TeP − P T )fi and the operator TeP − P T has norm ≤ 2√2 on L2, we
[si+n, si+n+1] = [tn, tn+1] for j ∈ J2. Since d(i(y), J2) ≥ ed(i(x), i(supp fi)) and
where we also used the fact that supp Nj ⊃ [si, si+1] = [t0, t1] or supp Nj ⊃
kfik2 ≤ kfik∞ supp fi1/2, we finally get
g(y) . γd(i(y),J2)kfik2 supp fi−1/2,
g(y) . γ
ed(i(x),i(supp fi))kfik∞.
Looking at (3.1) and combining the latter estimate with (3.2), the proof is com-
pleted.
(cid:3)
This lemma can be used directly to prove Theorem 1.4 on the uniform bound-
edness of periodic orthogonal spline projection operators on L∞:
after summation of a geometric series.
i=0 f · 1[si,si+1)
(cid:3)
Proof of Theorem 1.4. We just decompose the function f as f =Pn−1
and apply Lemma 3.1 to each summand and the assertion keP fk∞ . kfk∞ follows
Remark 3.2. (i) Since eP is a selfadjoint operator, Theorem 1.4 also implies that
eP is bounded as an operator from L1(T) to L1(T) by the same constant ck as in
the above theorem. Moreover, by interpolation, eP is also bounded by ck as an
operator from Lp(T) to Lp(T) for any p ∈ [1,∞].
(ii) In the proof of Lemma 3.1, we only use the second inequality of Lemma 2.1
which follows from inequality (1.2) on the inverse of the B-spline Gram matrix and
does not need its stronger form (1.3). Similarly to the equivalence of Shadrin's
theorem and (1.2) in the non-periodic case, we can derive the equivalence of
Theorem 1.4 and the estimate
eaij ≤
,
0 ≤ i, j ≤ n − 1,
ed(i,j)
Kγ
max(eκi,eκj)
where (eaij) denotes the inverse of the Gram matrix (heNi, eNji), K > 0 and γ ∈
(0, 1) are constants only depending on the spline order k, eκi denotes the length
of the support of eNi and ed is the canonical distance in Z/nZ. The proof of this
equivalence uses the same tools as the proof in the non-periodic case: a periodic
version of both Demko's theorem and de Boor's stability.
4. Almost everywhere convergence
In this section we prove Theorem 1.3 on the a.e. convergence of periodic spline
projections.
8
M. PASSENBRUNNER
Proof of Theorem 1.3. Without loss of generality, we assume thate∆n has n points.
Let e∆n = (s(n)
tions. Associated to it, define the non-periodic point sequence ∆n = (t(n)
with the boundary points 0 and 1 as
)n−1
j=0 be the corresponding periodic B-spline func-
j )n+k−1
j=0 and (eN (n)
j )n−1
j=−m
j
j
,
t(n)
j = s(n)
t(n)
−m = · · · = t(n)
−1 = 0,
j = 0, . . . , n − 1,
n = · · · = t(n)
t(n)
n+k−1 = 1.
We choose the integer m such that the multiplicity of the point 0 in ∆n is k and
denote by (N (n)
)n−1
j=−m the non-periodic B-spline functions corresponding to this
point sequence and by Pn the orthogonal projection operator onto the span of
(N (n)
j
)n−1
j=−m.
j
We will show that ePnf (x) → f (x) for all x in the set A from Theorem 1.2 of
full Lebesgue measure such that lim PnT f (x) = T f (x) for all x ∈ A, where T
is just the operator that canonically identifies a function defined on T with the
corresponding function defined on [0, 1) and we write x for a point in T as well
as for its representative in the interval [0, 1).
(4.1)
So, choose an arbitrary (non-zero) point x ∈ A and decompose ePnf (x):
ePnf (x) = TePnf (x) = PnT f (x) +(cid:0)TePnf (x) − PnT f (x)(cid:1).
For the first term of (4.1), PnT f (x), we have that limn→∞ PnT f (x) = T f (x) =
Moreover,
It remains to estimate the second term gn(x) = TePnf (x) − PnT f (x) =
TePnf (x) − T f (x) + T f (x) − PnT f (x) of (4.1). First, note that gn ∈ Sk(∆n).
hTePnf − T f, N (n)
since ePn is the projection operator onto the span of the B-spline functions (eN (n)
i = hePnf − f, eN (n)
j = 0, . . . , n − k − 1,
i = 0,
and
),
j
j
j
hT f − PnT f, N (n)
j
i = 0,
j = −m, . . . , n − 1,
since Pn is the projection operator onto the span of the functions (N (n)
fore, gn ∈ Sk(∆n) can be written as
j
). There-
f (x) since x ∈ A.
gn =Xj∈Jn
hgn, N (n)
j
iN (n)∗
j
,
with Jn = {−m, . . . ,−1}∪{n− k, . . . , n− 1} and (N (n)∗
(N (n)
). We now apply Lemma 2.1 with p = 1 to gn and get
j
j
) being the dual basis to
gn(x) . γd(in(x),Jn)kgnk1 max
ℓ∈in(x),j∈Jn
ℓj denotes the length of the convex hull of supp N (n)
where h(n)
is the set of indices i such that x is in the support of N (n)
i
,
1
h(n)
ℓj
ℓ ∪supp N (n)
and in(x)
. Since for ℓ ∈ in(x), the
j
ORTHOGONAL PROJECTORS ONTO SPACES OF PERIODIC SPLINES
9
point x is contained in supp N (n)
is contained in supp N (n)
ℓ
j
, we can further estimate
and for j ∈ Jn either the point 0 or the point 1
gn(x) . γd(in(x),Jn)kgnk1
1
.
min(x, 1 − x)
Now, kgnk1 = k(TePn − PnT )fk1 . kfk1, since the operator T has norm one
on L1 and ePn and Pn are both bounded on L1 uniformly in n by Theorem 1.4
(cf. Remark 3.2) and Theorem 1.1, respectively. Since e∆n tends to zero, and
point x and decomposition (4.1), limePnf (x) = f (x). Since x ∈ A was arbitrary
a fortiori the same is true for ∆n, we have that d(in(x), Jn) tends to infinity
as n → ∞. This implies limn→∞ gn(x) = 0, and therefore, by the choice of the
and A is a set of full Lebesgue measure, we obtain
lim
n→∞ePnf (y) = 0,
for a.e. y ∈ T,
and the proof is completed.
(cid:3)
5. The case of infinite point sequences
In this last section, we use the methods introduced in the previous sections to
recover Shadrin's theorem for infinite point sequences (see [8, 3]).
Let (si)i∈Z be a biinfinite point sequence in R satisfying
si ≤ si+1,
si < si+k,
with the corresponding B-spline functions (eNi)i∈Z satisfying supp eNi = [si, si+k].
Furthermore, we denote by eP the orthogonal projection operator onto the closed
linear span of the functions (eNi)i∈Z .
Lemma 5.1. Let f be a function on (inf si, sup si) with compact support. Then,
for any x ∈ (inf si, sup si),
eP f (x) . γd(i(x),i(supp f ))kfk∞,
where γ ∈ (0, 1) is the constant appearing in inequality (1.3).
Proof. For notational simplicity, we assume in this proof that the sequence (si) is
strictly increasing. Let x ∈ (inf si, sup si) and let I(x) be the interval I = [si, si+1)
containing x. Since f has compact support and the sequence (si) is biinfinite, we
can choose the indices ℓ and r such that {x} ∪ supp f ⊂ [sℓ, sr+1) and with
J = {ℓ − k + 1, . . . , ℓ − 1} ∪ {r − k + 2, . . . , r}, the inequality
γd(i(x),J) supp f1/2I(x)−1/2 ≤ γd(i(x),i(supp f ))
is true.
Next, define the point sequence ∆ = (ti)r+k
i = ℓ, . . . , r + 1,
ti = si,
i=ℓ−k+1 by
a = tℓ−k+1 = · · · = tℓ = sℓ,
b = tr+k = · · · = tr+1 = sr+1,
and let the collection (Ni)r
i=ℓ−k+1 be the corresponding B-spline functions and
P the associated orthogonal projector. Let T be the operator that restricts a
10
M. PASSENBRUNNER
(5.1)
we decompose
function defined on (inf si, sup si) to the interval [a, b]. In order to estimate eP f (x),
Observe that P T f = Pn∈Fhf, NniN ∗
eP f (x) = TeP f (x) = P T f (x) +(cid:0)TeP f (x) − P T f (x)(cid:1).
supp Ni 6= ∅}. Applying Lemma 2.1 with the exponent p = ∞, we obtain
n, where F = i(supp f ) = {i : supp f ∩
P T f (x) . γd(i(x),F )kfk∞.
We now consider the second part of the decomposition (5.1), the function
moreover,
g = (TeP − P T )f = (TeP − T + T − P T )f . We observe that g ∈ Sk(∆) and,
hTeP f − T f, Nji = heP f − f, eNji = 0,
by definition of the projection operator eP , and,
hT f − P T f, Nji = 0,
j = ℓ, . . . , r − k + 1,
j = ℓ − k + 1, . . . , r,
by definition of the projection operator P . Therefore, we can write the function
g as
g =Xj∈J
hg, NjiN ∗
j
with J = {ℓ − k + 1, . . . , ℓ − 1} ∪ {r − k + 2, . . . , r} as defined above. Now, by
Lemma 2.1 with the exponent p = 2, we get
g(x) . γd(i(x),J)kgk2 · I(x)−1/2 . γd(i(x),J)kfk2 · I(x)−1/2
≤ γd(i(x),J) supp f1/2I(x)−1/2kfk∞.
Finally, due to the choice of ℓ and r,
γd(i(x),J) supp f1/2I(x)−1/2 ≤ γd(i(x),i(supp f )),
which proves the lemma.
(cid:3)
We can now use this lemma to define eP f for functions f ∈ L∞(inf si, sup si)
that are not necessarily in L2(inf si, sup si) if inf si = −∞ or sup si = +∞. If we
let fi := f 1[si,si+1), then fi has compact support and the above lemma implies
that the series
eP f (x) :=Xi∈Z eP fi(x),
x ∈ (inf si, sup si),
is absolutely convergent and, moreover, there exists a constant C only depending
on the spline order k such that
This operator enjoys the characteristic property of an orthogonal projection:
Remark 5.2. One can combine the proofs of Lemma 5.1 and Lemma 3.1 to also
obtain the uniform boundedness of the spline orthoprojector on L∞ for one-sided
infinite point sequences.
keP fk∞ ≤ Ckfk∞.
heP f − f, eNii = 0,
i ∈ Z.
ORTHOGONAL PROJECTORS ONTO SPACES OF PERIODIC SPLINES
11
References
[1] Z. Ciesielski. Orthogonal projections onto spline spaces with arbitrary knots. In Function
spaces (Pozna´n, 1998), volume 213 of Lecture Notes in Pure and Appl. Math., pages 133 -- 140.
Dekker, New York, 2000.
[2] C. de Boor. The quasi-interpolant as a tool in elementary polynomial spline theory. In
Approximation theory (Proc. Internat. Sympos., Univ. Texas, Austin, Tex., 1973), pages
269 -- 276. Academic Press, New York, 1973.
[3] C. de Boor. On the (bi)infinite case of Shadrin's theorem concerning the L∞-boundedness
of the L2-spline projector. Proc. Steklov Inst. Math., 277:s73 -- s78, 2012.
[4] S. Demko. Inverses of band matrices and local convergence of spline projections. SIAM J.
Numer. Anal., 14(4):616 -- 619, 1977.
[5] R. A. DeVore and G. G. Lorentz. Constructive approximation, volume 303 of Grundlehren
der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences].
Springer-Verlag, Berlin, 1993.
[6] M. Passenbrunner and A. Shadrin. On almost everywhere convergence of orthogonal spline
projections with arbitrary knots. J. Approx. Theory, 180:77 -- 89, 2014.
[7] L. L. Schumaker. Spline functions: basic theory. Cambridge Mathematical Library. Cam-
bridge University Press, Cambridge, third edition, 2007.
[8] A. Shadrin. The L∞-norm of the L2-spline projector is bounded independently of the knot
sequence: a proof of de Boor's conjecture. Acta Math., 187(1):59 -- 137, 2001.
Institute of Analysis, Johannes Kepler University Linz, Austria, 4040 Linz,
Altenberger Strasse 69
E-mail address: [email protected]
|
1010.0461 | 3 | 1010 | 2011-07-31T14:14:42 | New upper bounds for the constants in the Bohnenblust-Hille inequality | [
"math.FA"
] | A classical inequality due to Bohnenblust and Hille states that for every positive integer $m$ there is a constant $C_{m}>0$ so that $$(\sum\limits_{i_{1},...,i_{m}=1}^{N}|U(e_{i_{^{1}}},...,e_{i_{m}})| ^{\frac{2m}{m+1}}) ^{\frac{m+1}{2m}}\leq C_{m}| U|$$ for every positive integer $N$ and every $m$-linear mapping $U:\ell_{\infty}^{N}\times...\times\ell_{\infty}^{N}\rightarrow\mathbb{C}$, where $C_{m}=m^{\frac{m+1}{2m}}2^{\frac{m-1}{2}}.$ The value of $C_{m}$ was improved to $C_{m}=2^{\frac{m-1}{2}}$ by S. Kaijser and more recently H. Qu\'{e}ffelec and A. Defant and P. Sevilla-Peris remarked that $C_{m}=(\frac{2}{\sqrt{\pi}})^{m-1}$ also works. The Bohnenblust--Hille inequality also holds for real Banach spaces with the constants $C_{m}=2^{\frac{m-1}{2}}$. In this note we show that a recent new proof of the Bohnenblust--Hille inequality (due to Defant, Popa and Schwarting) provides, in fact, quite better estimates for $C_{m}$ for all values of $m \in \mathbb{N}$. In particular, we will also show that, for real scalars, if $m$ is even with $2\leq m\leq 24$, then $$C_{\mathbb{R},m}=2^{1/2}C_{\mathbb{R},m/2}.$$ We will mainly work on a paper by Defant, Popa and Schwarting, giving some remarks about their work and explaining how to, numerically, improve the previously mentioned constants. | math.FA | math |
IMPROVING THE CONSTANTS FOR THE REAL AND COMPLEX
BOHNENBLUST-HILLE INEQUALITY
DANIEL PELLEGRINO AND JUAN B. SEOANE-SEP ´ULVEDA*
Abstract. A classical inequality due to Bohnenblust and Hille states that for every N ∈ N and
every m-linear mapping U : ℓN
∞
× · · · × ℓN
∞
→ C we have
N
X
i1,...,im=1
(cid:12)(cid:12)(cid:12)
U (ei1 , ..., eim )(cid:12)(cid:12)(cid:12)
2m
m+1
m+1
2m
≤ Cm kU k ,
m−1
2
m−1
(in fact a recent remark of A. Defant and P. Sevilla-Peris indicates that
). Bohnenblust-Hille inequality is also true for real Banach spaces with the
where Cm = 2
√π(cid:17)m−1
Cm ≤ (cid:16) 2
. In this note we show that an adequate use of a recent new proof of
constants Cm = 2
Bohnenblust-Hille inequality, due to Defant, Popa and Schwarting, combined with the optimal
constants of Khinchine's inequality (due to Haagerup) provides quite better estimates for the
constants involved in both real and complex Bohnenblust-Hille inequalities. For instance, in
the real case, for 2 ≤ m ≤ 14, we show that the constants Cm = 2
can be replaced by
m−1
2
2
m2+6m−8
m2+6m−7
8m
2
of Cm. In both complex and real cases, the new constants are asymptotically better.
if m is odd, improving, in this way, the known values
if m is even and by 2
8m
In 1931, Bohnenblust and Hille ([2], or the more recent [8, 9]) asserted that for every positive
1. Preliminaries and background
integer N and every m-linear mapping U : ℓN
N
Xi1,...,im=1(cid:12)(cid:12)U (ei1 , ..., eim)(cid:12)(cid:12)
∞ → C we have
m+1
2m
≤ Cm kUk ,
∞ × ··· × ℓN
m+1
2m
m−1
2
where Cm = 2
(actually this result also holds for real Banach spaces). The case m = 2 is a
famous result known as Littlewood's 4/3-inequality. It seems that the Bohnenblust-Hille inequality
was overlooked and was only re-discovered several decades later by Davie [6] and Kaijser [13].
While the exponent
is not. Very recently, Defant and
Sevilla-Peris [8, Section 4] indicated that by using Sawa's estimate for the constant of the complex
2m
m+1 is optimal, the constant Cm = 2
m−1
2
Khinchine's inequality in Steinhaus variables (see [16]) it is possible to prove that Cm ≤(cid:16) 2
in the complex case (this is a strong improvement on the previous constants and it seems that
these are the best known estimates for the complex case).
√π(cid:17)m−1
The (complex and real) Bohnenblust-Hille inequality can be re-written in the context of multiple
summing multilinear operators, as we will see next. Multiple summing multilinear mappings
between Banach spaces is a recent, very important and useful nonlinear generalization of the
concept of absolutely summing linear operators. This class was introduced, independently, by
Matos [15] (under the terminology fully summing multilinear mappings) and Bombal, P´erez-Garc´ıa
and Villanueva [3]. The interested reader can also refer to [5, 4] for other Bohnenblust-Hille type
results.
Throughout this paper X1, . . . , Xm and Y will stand for Banach spaces over K = R or C, and
X′ stands for the dual of X. By L(X1, . . . , Xm; Y ) we denote the Banach space of all continuous
*Supported by the Spanish Ministry of Science and Innovation, grant MTM2009-07848.
1
2
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
m-linear mappings from X1 × ··· × Xm to Y with the usual sup norm. For x1, ..., xn in X, let
k(xj )n
j=1kw,1 := sup{k(ϕ(xj))n
j=1k1 : ϕ ∈ X′,kϕk ≤ 1}.
If 1 ≤ p < ∞, an m-linear mapping U ∈ L(X1, . . . , Xm; Y ) is multiple (p; 1)-summing (denoted
Π(p;1)(X1, . . . , Xm; Y )) if there exists a constant Lm ≥ 0 such that
(1.1)
U (x(1)
j1 , . . . , x(m)
N
Xj1,...,jm=1(cid:13)(cid:13)(cid:13)
1
p
≤ Lm
p
jm )(cid:13)(cid:13)(cid:13)
m
Yk=1(cid:13)(cid:13)(cid:13)
(x(k)
j
)N
j=1(cid:13)(cid:13)(cid:13)w,1
for every N ∈ N and any x(k)
jk ∈ Xk, jk = 1, . . . , N , k = 1, . . . , m. The infimum of the constants
satisfying (1.1) is denoted by kUkπ(p;1). For m = 1 we have the classical concept of absolutely
(p; 1)-summing operators (see, e.g. [7, 10]).
A simple reformulation of Bohnenblust-Hille inequality asserts that every continuous m-linear
(or Lm =
m+1 ; 1)-summing with Lm = Cm = 2
m−1
2
form T : X1 × ··· × Xm → K is multiple ( 2m
√π(cid:17)m−1
(cid:16) 2
the real case the best constants known seem to be Cm = 2
m−1
2
.
for the complex case, using the estimates of Defant and Sevilla-Peris, [8]). However, in
The main goal of this paper is to obtain better constants for the Bohnenblust-Hille inequality
in the real and complex case. For this task we will use a recent proof of a general vector-valued
version of Bohnenblust-Hille inequality ([9, Theorem 5.1]). The Bohnenblust-Hille inequality is
stated in [9, Corollary 2] as a consequence of [9, Theorem 5.1]. The procedure of the proof of [9,
Corollary 2] allows us to obtain much better values than Cm = 2
. However, in this note we
explore the ideas of [9] in a different way, in order to obtain even better estimates for the constants
that can be derived from [9, Corollary 2]. The constants we obtain here can be derived from [9,
Theorem 5.1] via an adequate choice of variables.
m−1
2
Let us recall some results that we will need in this note. The first result is a well-known
inequality due to Khinchine (see [10]):
Theorem 1.1 (Khinchine's inequality). For all 0 < p < ∞, there exist constants Ap and Bp such
that
(1.2)
Ap N
Xn=1
1
2
an2!
N
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ Z 1
Xn=1
anrn (t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
1
p
dt!
≤ Bp N
Xn=1
1
2
an2!
for every positive integer N and scalars a1, ..., an (here, rn denotes the n-Rademacher function).
Above, it is clear that B2 = 1. From (1.2) it follows that
(1.3)
N
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Z 1
Xn=1
anrn (t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
1
p
dt!
≤ BpA−1
N
Xn=1
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
r Z 1
anrn (t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
r
1
r
dt!
and the product of the constants BpA−1
r will appear later on in Theorem 1.3. The notation Ap
and Bp will be kept along the paper. Next, let us recall a variation of an inequality due to Blei
(see [9, Lemma 3.1]).
Theorem 1.2 (Blei, Defant et al.). Let A and B be two finite non-void index sets, and (aij)(i,j)∈A×B
a scalar matrix with positive entries, and denote its columns by αj = (aij )i∈A and its rows by
βi = (aij )j∈B. Then, for q, s1, s2 ≥ 1 with q > max(s1, s2) we have
s1
Xj∈B
X(i,j)∈A×B
q
kαjks2
≤ Xi∈A
q !
kβiks1
aw(s1,s2)
ij
w(s1 ,s2)
f (s2 ,s1)
f (s1 ,s2)
s2
,
1
ON THE CONSTANTS FOR THE BOHNENBLUST-HILLE INEQUALITY
3
with
w : [1, q)2 → [0,∞), w(x, y) :=
f : [1, q)2 → [0,∞), f (x, y) :=
q2(x + y) − 2qxy
q2 − xy
q2x − qxy
q2(x + y) − 2qxy
,
.
The following theorem is a particular case of [9, Lemma 2.2] for Y = K using that the cotype 2
constant of K is 1, i.e., C2(K) = 1 (following the notation from [9]):
Theorem 1.3 (Defant et al). Let 1 ≤ r ≤ 2, and let (yi1,...,im)N
Xi1,...,im=1
yi1...im2
≤ (A2,r)m
Xi1,...,im=1
1/2
N
N
where
Z[0,1]m(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ri1 (t1)...rim (tm)yi1...im(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
A2,r ≤ A−1
r B2 = A−1
r
(since B2 = 1).
1/r
,
r
dt1...dtm
i1,...,im=1 be a matrix in K. Then
The meaning of A2,r, w and f from the above theorems will also be kept in the next section
and KG will denote the complex Grothendieck constant.
2. Improved constants for the Bohnenblust-Hille theorem: The real case
The main results from [9], Theorem 5.1 and Corollary 5.2, are very interesting vector-valued
generalizations of the Bohnenblust-Hille inequality. In this note we explore the proof of [9, Theorem
5.1] in such a way that the constants obtained are better than those that can be derived from [9,
Corollary 5.2]. We will use here a modification of the proof of [9, Corollary 5.2] for the particular
case of the Bohnenblust-Hille inequality applying some changes, partly inspired by [9, Theorem
5.1], improving the constants. By doing this, we will avoid some technicalities from the arguments
from [9, Theorem 5.1] that are not needed here.
As we said, following the proof of [9, Corollary 5.2] and using the optimal values for the constants
of Khinchine's inequality (due to Haagerup), the following estimates can be calculated for Cm:
(2.1)
(2.2)
CR,2 = √2,
CR,m = 2
m−1
2m CR,m−1
A 2m−2
m !1− 1
m
for m ≥ 3.
In particular, if 2 ≤ m ≤ 13,
(2.3)
CR,m ≤ 2
m2+m−2
4m .
Remark 2.1. It is worth mentioning that the above constants are not explicitly calculated in [9].
Since our procedure below will provide better constants, we will not give much detail on the above
estimates.
A different approach on some of the ideas from [9] can give better estimates for the real case,
as we see in the following result.
Theorem 2.2. For every positive integer m and real Banach spaces X1, ..., Xm,
Π( 2m
m+1 ;1)(X1, ..., Xm; R) = L(X1, ..., Xm; R) and k.kπ( 2m
m+1 ;1) ≤ CR,m k.k
with
(2.4)
In particular, if 2 ≤ m ≤ 14,
(2.5)
CR,2 = 2
1
2 and CR,3 = 2
5
6 ,
CR,m ≤ 2
m−2
m
CR,m−2
A2
2m−4
m−1
1
2
for m > 3.
CR,m ≤ 2
m2+6m−8
8m
if m is even, and
4
(2.6)
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
CR,m ≤ 2
m2+6m−7
8m
if m is odd.
Proof. The case m = 2 is Littlewood's 4/3-inequality. For m = 3 we have CR,3 = 2
6 from (2.3).
We proceed by induction, but the case m is obtained as a combination of the cases 2 with m − 2
instead of 1 and m − 1 as in [9, Corollary 5.2].
Suppose that the result is true for m−2 and let us prove for m. Let U ∈ L(X1, ..., Xm; R) and N
be a positive integer. For each 1 ≤ k ≤ m consider x(k)
1 , ..., x(k)
(x(k)
)N
j
5
N ∈ Xk so that (cid:13)(cid:13)(cid:13)
j=1(cid:13)(cid:13)(cid:13)w,1 ≤ 1,
k = 1, .., m.
Consider, in the notation of Theorem 1.2,
q = 2, s1 =
4
3
, and s2 =
2(m − 2)
(m − 2) + 1
=
2m − 4
m − 1
.
Thus,
and, from Theorem 1.2, we have
w(s1, s2) =
2m
m + 1
(m+1)/2m
N
N
2m
≤
U (x(1)
i1 , ..., x(m)
Xi1,...,im=1(cid:12)(cid:12)(cid:12)
im )(cid:12)(cid:12)(cid:12)
≤
Xi1,...,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
≤
Xim−1,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
m+1
im−1,im=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
i1 , ..., x(m)
2
i1....,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)
)(cid:17)N
, ..., x(m)
im
i1
N
4
3
2
.
f ( 4
3 ,s2)
2(m−2)
(m−2)+1
f (s2, 4
3 )/ 2(m−2)
(m−2)+1
Now we need to estimate the two factors above. We will write dt := dt1...dtm−2.
For each im−1, im fixed, we have (from Theorem 1.3),
ri1 (t1)...rim−2 (tm−2)U (x(1)
i1 , ..., x(m)
4
3
dt
im )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N
Xim−2=1
rim−2 (tm−2)x(m−2)
im−2 , x(m−1)
4
3
dt.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
im
im−1 , x(m)
4
3
N
N
2, 4
2, 4
2 ≤
i1 , ...,
ri1 (t1)x(1)
i1 , ..., x(m)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
3 (cid:17)4/3
≤(cid:16)Am−2
3 (cid:17)4/3
=(cid:16)Am−2
Summing over all im−1,im = 1, ..., N we obtain
i1....,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
[0,1]m−2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xi1,...,im−2=1
R
[0,1]m−2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
U
Xi1=1
R
i1....,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
U
Xi1=1
Xim−1,im=1
Xim−1,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
3 (cid:17)4/3
(cid:16)Am−2
R
i1 , ..., x(m)
ri1 (t1)x(1)
2 ≤
[0,1]m−2
2, 4
N
N
N
4
3
i1 , ...
N
Xim−2=1
rim−2 (tm−2)x(m−2)
im−1 , x(m−1)
im
dt.
, x(m)
im
4
3
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ON THE CONSTANTS FOR THE BOHNENBLUST-HILLE INEQUALITY
5
Using the case m = 2 we thus have
4
3
2 ≤
i1 , ...
N
Xim−2=1
rim−1 (tm−2)x(m−2)
4
3
dt
π( 4
3 ;1)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
im−2 , ., .
Next we obtain the other estimate. For each i1, ..., im−2 fixed, and dt := dtm−1dtm, we have
3
4
≤ Am−2
3 kUk
2, 4
√2.
i1....,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)
)(cid:17)N
4
3
2
N
N
N
N
i1
4
3
.
2, 4
2, 4
2, 4
4
3
4
3
4
Hence
4
3
dt
≤ A2
, ..., x(m)
im
ri1 (t1)x(1)
i1 , ..., x(m)
(from Theorem 1.3):
Xim−1,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
i1....,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
[0,1]m−2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
U
3 (cid:17)
≤(cid:16)Am−2
Xi1=1
R
√2(cid:17)
≤(cid:16)Am−2
3 (cid:17)
[0,1]m−2(cid:16)kUk
R
√2(cid:17)
3 (cid:16)kUk
=(cid:16)Am−2
3 (cid:17)
Xim−1,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
im−1,im=1(cid:13)(cid:13)(cid:13)(cid:13)2 ≤
im )(cid:17)N
i1 , ..., x(m)
R[0,1]2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2,s2
Xim−1,im=1
R[0,1]2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2,s2
U
x(1)
Xi1,...,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
Xi1,...,im−2=1
2,s2 R[0,1]2
Xim−1=1
Summing over all i1, ...., im−2 = 1, ..., N we get:
im=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
i1 , ..., x(m)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
U
x(1)
Xi1,...,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
2,s2(cid:1)s2 ≤ R[0,1](cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤(cid:0)A2
2,s2(cid:1)s2 R[0,1]
≤(cid:0)A2
im=1(cid:13)(cid:13)(cid:13)(cid:13)
)(cid:17)N
U ., ...,
Xim−1=1
R,m−2 kUks2 dt =(cid:0)A2
We thus have, by the induction step,
i1 , ..., x(m−2)
im−2 ,
i1 , ..., x(m−2)
im−2 ,
≤ A2s2
, ..., x(m)
im
2 ≤
2 ≤
= A2
Cs2
s2
s2
i1
N
N
N
N
N
and so
rim−1 (tm−1)rim (tm)U (x(1)
i1 , ..., x(m)
s2
1/s2
im )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
dt
Xim=1
N
rim−1 (tm−1)x(m−1)
im−1 ,
rim (tm)x(m)
1/s2
.
s2
dt
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
im
N
Xim−1=1
rim−1 (tm−1)x(m−1)
im−1 ,
N
Xim=1
rim (tm)x(m)
im
s2
dt.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
im−1
rim−1 (tm−1)x(m−1)
2,s2(cid:1)s2 Cs2
Xim=1
R,m−2 kUks2
N
rim (tm)x(m)
s2
im !(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
dt
π(s2;1)
N
Xi1,...,im−2=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
i1 , ..., x(m)
s2
2
im=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
1/s2
≤(cid:0)A2
2,s2(cid:1) CR,m−2 kUk .
Hence, combining both estimates, we obtain
(m+1)/2m
N
Xi1,...,im=1(cid:12)(cid:12)(cid:12)
Also,
U (x(1)
3 ,s2)
2,s2(cid:1) CR,m−2 kUk(cid:3)f (s2, 4
(cid:2)(cid:0)A2
3 )
.
6
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
2
m
,
= 1 −
2
m
,
=
f (
2, 4
3
4
3
4
3
) =
2m
m+1
f (s2,
, s2) =
i1 , ..., x(m)
im )(cid:12)(cid:12)(cid:12)
≤hAm−2
4(cid:16) 4
4(cid:0) 2m−2
4(cid:0)1 + 2m−2
√2kUkif ( 4
m−1 (cid:17)
3(cid:16) 2m−4
4 4
3 − 2 4
3(cid:16) 2m−4
m−1 (cid:17) − 4 4
m−1 (cid:17)
3 + 2m−4
m (cid:1) − 2(cid:0) 2m−2
m (cid:1)
m (cid:1) − 4(cid:0) 2m−2
m (cid:1)
√2kUki
≤hAm−2
√2i
=hAm−2
m (cid:16)A2
√2i
m (cid:16)A2
=hAm−2
m−1 (cid:17)
m (cid:16)A2, 4
p whenever p ≤ 1.847. So, for 2 ≤ m ≤ 14 we have
m+1
im )(cid:12)(cid:12)(cid:12)
i1 , ..., x(m)
A2, 2m−4
(m+1)/2m
= 2
2, 4
3
2, 4
3
2, 4
3
2m
2
3
1
2
2
U (x(1)
A 2m−4
m−1
= 2
1
2 −
m−1
2m−4 .
and, therefore
N
Xi1,...,im=1(cid:12)(cid:12)(cid:12)
1
2 − 1
that Ap = 2
(2.7)
m−1 (cid:17) CR,m−2 kUki1− 2
m
m
1− 2
C
R,m−2 kUk
2, 2m−4
m h(cid:16)A2
m−1 (cid:17)1− 2
2, 2m−4
m
m
CR,m−2(cid:17)1− 2
m−2
kUk
2, 2m−4
m−1
2m−4
m
(CR,m−2)
m kUk .
Now let us estimate the constants CR,m. We know that B2 = 1 and, from [11], we also know
Hence, from (1.3) and using the best constants of Khinchine's inequality from [11], we have
and
1
m (cid:16)(cid:16)2
4 − 1
obtaining that, if 2 ≤ m ≤ 14,
CR,m ≤ 2
3
2(cid:17)(cid:16)2
(CR,m−2)1− 2
m = 2
m+2
2m (CR,m−2)1− 2
m
2m−4
m−1
m
4
3
m−1
2m−4
2m−4 − 1
3 ≤ A−1
A2, 4
m−1 ≤ A−1
A2, 2m−4
2(cid:17)(cid:17)
CR,m ≤ 2
CR,m ≤ 2
8m
8m
m2+6m−7
m2+6m−8
= 2
3
4 − 1
2 ,
m−1
2m−4 − 1
2 ,
= 2
if m is even,
if m is odd.
In general we easily get
The numerical values of CR,m, for m > 14, can be easily calculated by using the exact values of
CR,m ≤ 2
CR,m−2
A2
2m−4
m−1
1
2
m−2
m
.
A 2m−4
m−1
(see [11]):
A 2m−4
m−1
2 (cid:19)
Γ(cid:18) 2m−4
m−1 +1
√π
= √2
(m−1)/(2m−4)
In the below table we compare the first constants Cm = 2
derived from [9, Cor. 5.2] with the new constants CR,m:
.
m−1
2
and the constants that can be
(cid:3)
ON THE CONSTANTS FOR THE BOHNENBLUST-HILLE INEQUALITY
7
18
36
40
54
48
64
84
104
m CR,m (using (2.5) and (2.6)) Constants from [9, Cor. 5.2])
3
4
5
6
7
8
9
10
11
12
13
220/24 ≈ 1.782
232/32 = 2
2
40 ≈ 2.298
48 ≈ 2.520
2
56 ≈ 2.828
2
2
64 ≈ 3.084
72 ≈ 3.429
2
2
80 ≈ 3.732
88 ≈ 4.128
2
96 ≈ 4.490
2
2
104 ≈ 4.951
112 ≈ 5.383
25/6 ≈ 1.782
2
16 ≈ 2.18
2
20 ≈ 2.639
2
24 ≈ 3.17
28 ≈ 3.807
2
2
32 ≈ 4.555
36 ≈ 5.443
2
2
40 ≈ 6.498
44 ≈ 7.752
2
48 ≈ 9.243
2
2
52 ≈ 11.016
14 (cid:19)1− 1
28 (cid:18) 2
≈ 13.126
180
52
A 26
14
128
152
180
70
88
208
272
13
2
240
108
130
154
180
2
14
Cm = 2
m−1
2
22 = 4
22/2 = 2
23/2 ≈ 2.828
25/2 ≈ 5.656
26/2 = 8
27/2 ≈ 11.313
28/2 = 16
29/2 ≈ 22.627
210/2 = 32
211/2 ≈ 45.254
212/2 = 64
213/2 ≈ 90.509
In the column at the center of the previous table we have used equations (2.7) and (2.2) for
3 ≤ m ≤ 13. In the last line of this same column (m = 14) we have used equation (2.2) together
14/26
with the fact that that A 26
14
≈ 0.9736.
= √2
14
+1
√π
Γ(cid:18) 26
2 (cid:19)
Remark 2.3. In this section we have actually shown that the new constants obtained present a
better asymptotic behavior than the previous ones (including those derived from [9, Cor. 5.2]).
Indeed, we have previously seen that
CR,m ≤ 2
CR,m−2
A2
2m−4
m−1
1
2
m−2
m
.
As m → ∞ we know that A2, 2m−4
m−1
increases to 1. So,
For the original constants Cm = 2
m−1
2 we have
lim sup
CR,m
(CR,m−2)
m−2
m ≤ 2
1
2 .
and thus
Cm
(Cm−2)
m−2
m
lim
Cm
(Cm−2)
2m−3
m
= 2
= 4.
m−2
m
2 is replaced by 2
Also, for the constants from [9, Cor. 5.2], a similar calculation shows us that 2
in (2.3). To summarize, these new constants, although smaller than the "old ones", have the best
asymptotic behavior.
1
3. Improved constants for the Bohnenblust-Hille theorem: The complex case
As in the real case, following the proof of [9, Corollary 5.2] and using the optimal values for
the constants of Khinchine's inequality (due to Haagerup) and using that KG = CC,2 (see [1]), the
following estimates can be calculated for Cm:
CC,2 = KG ≤ 1.4049 < √2,
m !1− 1
2m CC,m−1
A 2m−2
CC,m = 2
m−1
m
for m ≥ 3,
8
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
In particular, if 2 ≤ m ≤ 13,
The above estimates are much better than CC,m = 2
CC,m ≤ 2
m2+m−6
4m K 2/m
G
m−1
2 but worst than the constants CC,m =
obtained by Defant and Sevilla-Peris [8]. However, our approach will provide even better
√π(cid:17)m−1
(cid:16) 2
constants.
The following lemma is essentially the main result from the previous section which comes from
[9], although now we will obtain different constants, since we will be dealing with the complex case.
Lemma 3.1. For every positive integer m and complex Banach spaces X1, . . . , Xm,
m+1 ;1) ≤ CC,m k.k
m+1 ;1)(X1, ..., Xm; C) = L(X1, ..., Xm; C) and k.kπ( 2m
Π( 2m
with
√π(cid:19)m−1
π1/m
In particular, if 4 ≤ m ≤ 14 we have
CC,m =(cid:18) 2
CC,m ≤
A2
m+2
2
2m
for m = 2, 3,
m−2
m
1
2m−4
m−1
π1/m(cid:19) 2
(CC,m−2)
m−2
m for m ≥ 4.
m+4
2m (CC,m−2)
m−2
m .
CC,m ≤(cid:18) 1
The proof of this result is essentially in the same spirit as that of Theorem 2.2. The cases of
CC,2 and CC,3 are already known and the proof is (also) done by induction, using the cases m − 2
and 2 in order to achieve the case m. By proceeding in this way one obtains, at the end, that
CC,m ≤
and for 2 ≤ m ≤ 14 we have
2
m+2
2m
π1/m
m−2
m
1
A2
2m−4
m−1
(CC,m−2)
m−2
m ,
A 2m−4
m−1
= 2
1
2 −
m−1
2m−4 = 2
−1
2m−4 ,
which leads to
CC,m ≤
m+2
2
2m
π1/m (cid:16)2
1
m−2(cid:17)
m−2
m
(CC,m−2)
m−2
m =(cid:18) 1
π1/m(cid:19) 2
m+4
2m (CC,m−2)
m−2
m
for 4 ≤ m ≤ 14.
3.1. Comparing the "first" constants. The first constants Dm =(cid:16) 2
than the constants that we have obtained in the previous lemma. However
√π(cid:17)m−1
from [8] are better
lim
Dm
(Dm−2)
=(cid:18) 2
√π(cid:19)4
m−2
m
≈ 1.621 > √2 = lim sup
CC,m
(CC,m−2)
.
m−2
m
So, our constants are asymptotically better, and from a certain level m, they will be better than
Dm =(cid:16) 2
√π(cid:17)m−1
Below we compare the first constants:
For the case m = 4, notice that we have
. We will show that this happens when m ≥ 8.
CC,4 ≤
2
π1/4 (CC,2)
√π(cid:17)3
but this constant is worst than (cid:16) 2
√π(cid:17)3
it would be better to consider (cid:16) 2
1
2 =
1
2
2
√π(cid:19)
π1/4 (cid:18) 2
=
23/2
π1/2 ≈ 1.5957
≈ 1.437. So, in order to improve the constants that follow,
instead of
2 for the value of CC,4.
2
π1/4 (CC,2)
1
ON THE CONSTANTS FOR THE BOHNENBLUST-HILLE INEQUALITY
9
For instance, for m = 8 the situation is different. We have
Similarly, for the cases 5 ≤ m ≤ 7 we also have that CC,m is slightly worst than (2/√π)m−1
but, for m ≥ 8 our constants are better than the old ones.
16 (cid:18) 2
π1/8(cid:19) 2
≈ 2.329. Also, as we announced, for m > 8 our
236/8
π2 ≈ 2.293
√π(cid:19)5!
π1/8(cid:19) 2
CC,8 ≤(cid:18) 1
8 =(cid:18) 1
√π(cid:17)7
and now this constant is better than (cid:16) 2
constants are better.
16 (CC,6)
=
6
8
12
12
6
In the next section we state the previous lemma using the previous information
3.2. Comparing the "remaining" constants (m > 8). Now it is time to state the last lemma
adding the better constants:
Theorem 3.2. For every positive integer m and every complex Banach spaces X1, ..., Xm,
Π( 2m
m+1 ;1)(X1, ..., Xm; C) = L(X1, ..., Xm; C) and k.kπ( 2m
m+1 ;1) ≤ CC,m k.k
with
√π(cid:19)m−1
π1/m
In particular, for 8 ≤ m ≤ 14 we have
CC,m =(cid:18) 2
CC,m ≤
A2
m+2
2
2m
for m = 2, 3, 4, 5, 6, 7,
m−2
m
1
2m−4
m−1
π1/m(cid:19) 2
(CC,m−2)
m−2
m for m ≥ 8.
m+4
2m (CC,m−2)
m−2
m .
CC,m ≤(cid:18) 1
Keeping in mind that for m > 14, the evaluation of the precise values of Ap need the use of
Gamma function (see [11])
we have that (assuming some slight rounding error, for high values of m, due to computer calculus):
m New Constants CC,m =(cid:16) 2
(from [8])
2 (cid:1)√π !1/p
,
Ap = √2 Γ(cid:0) p+1
√π(cid:17)m−1
≈ 2.329
≈ 2.628
≈ 2.965
≈ 3.346
≈ 3.775
≈ 4.260
≈ 4.807
≈ 5.425
≈ 6.121
≈ 372
≈ 155, 973
m−1
2
32
16
Cm = 2
≈ 11.313
≈ 22.627
≈ 45.425
≈ 90.509
≈ 181.019
≈ 23, 726, 566
128
64
≈ 7.96131459 · 1014
8
9
10
11
12
13
14
15
16
50
100
≈ 2.293
≈ 2.552
≈ 2.814
≈ 3.059
≈ 3.417
≈ 3.711
≈ 4.125
≈ 4.479
≈ 4.963
≈ 100
≈ 7, 761
Acknowledgements. The authors thank A. Defant and P. Sevilla-Peris for important remarks
on the complex case.
References
[1] O. Blasco, G. Botelho, D. Pellegrino and P. Rueda, Summability of multilinear mappings: Littlewood, Orlicz
and beyond, Monatshefte fur Mathematik, to appear.
[2] H. F. Bohnenblust and E. Hille, On the absolute convergence of Dirichlet series, Ann. of Math. 32 (1931),
600-622.
10
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
[3] F. Bombal, D. P´erez-Garc´ıa and I. Villanueva, Multilinear extensions of Grothendieck's theorem, Q. J. Math.
55 (2004), 441-450.
[4] G. Botelho, H.-A. Braunss, H. Junek, D. Pellegrino, Inclusions and coincidences for multiple summing multi-
linear mappings, Proc. Amer. Math. Soc. 137 (2009), no. 3, 991 -- 1000.
[5] G. Botelho, C. Michels and D. Pellegrino, Complex interpolation and summability properties of multilinear
operators, Rev. Mat. Complut. 23 (2010), no. 1, 139 -- 161.
[6] A. M. Davie, Quotient algebras of uniform algebras, J. London Math. Soc. 7 (1973), 31-40.
[7] A. Defant and K. Floret, Tensor norms and operator ideals, North Holland Math. stud., vol 176, 1993.
[8] A. Defant and P. Sevilla-Peris, A new multilinear insight on Littlewood's 4/3-inequality, J. Funct. Anal. 256
(2009), 1642-1664.
[9] A. Defant. D. Popa, U. Schwarting, Coordinatewise multiple summing operators in Banach spaces, J. Funct.
Anal. 259 (2010), 220-242.
[10] J. Diestel, H. Jarchow and A. Tonge, Absolutely summing operators, Cambridge University Press, Cambridge,
1995.
[11] U. Haagerup, The best constants in Khinchine inequality, Studia Math. 70 (1982), 231-283.
[12] U. Haagerup, A new upper bound for the complex Grothendieck constant, Israel J. Math. 60 (1987), 199-224.
[13] S. Kaijser, Some results in the metric theory of tensor products, Studia Math. 63 (1978), 157-170.
[14] J. Lindenstrauss and A. Pelczynski, Absolutely summing operators in Lp-spaces and their applications, Studia
Math. 29 (1968), 275-325.
[15] M. C. Matos. Fully absolutely summing and Hilbert-Schmidt multilinear mappings, Collectanea Math. 54
(2003), 111-136.
[16] J. Sawa, The best constant in the Khinchine inequality for complex Steinhaus variables, the case p = 1, Studia
Math. 81 (1985), 107-126.
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]
1
1
0
2
l
u
J
1
3
]
.
A
F
h
t
a
m
[
3
v
1
6
4
0
.
0
1
0
1
:
v
i
X
r
a
NEW UPPER BOUNDS FOR THE CONSTANTS IN THE
BOHNENBLUST-HILLE INEQUALITY
DANIEL PELLEGRINO* AND JUAN B. SEOANE-SEP ´ULVEDA**
Abstract. A classical inequality due to Bohnenblust and Hille states that for every positive
integer m there is a constant Cm > 0 so that
N
X
i1,...,im=1
(cid:12)(cid:12)(cid:12)
U (ei1 , . . . , eim )(cid:12)(cid:12)(cid:12)
m+1
2m
2m
m+1
≤ Cm kU k
for every positive integer N and every m-linear mapping U : ℓN
∞
m−1
m−1
. The value of Cm was improved to Cm = 2
2
m+1
2m 2
m
2
× · · · × ℓN
∞
→ C, where Cm =
by S. Kaijser and more recently H.
Qu´effelec and A. Defant and P. Sevilla-Peris remarked that Cm = (cid:16) 2
.
Bohnenblust -- Hille inequality also holds for real Banach spaces with the constants Cm = 2
In this note we show that a recent new proof of the Bohnenblust -- Hille inequality (due to Defant,
Popa and Schwarting) provides, in fact, quite better estimates for Cm for all values of m ∈ N.
In particular, we will also show that, for real scalars, if m is even with 2 ≤ m ≤ 24, then
also works. The
m−1
2
√π(cid:17)m−1
We will mainly work on a paper by Defant, Popa and Schwarting, giving some remarks about
their work and explaining how to, numerically, improve the previously mentioned constants.
CR,m = 2
1
2 CR,m/2.
In 1930, Littlewood proved that
1. Preliminaries and background
3
4
4
N
U (ei, ej)
Xi,j=1
∞ × ℓN
∞ → C and every positive integer N. This is the well-known
One year later, in 1931, Bohnenblust and Hille ([2]) improved this result to multilinear forms
(see also [8, 9] for recent approaches). More precisely, the Bohnenblust -- Hille inequality asserts
that for every positive integer m there is a Cm > 0 so that
for every bilinear form U : ℓN
Littlewood's 4/3 inequality [16].
√2kUk
3
≤
N
Xi1,...,im=1(cid:12)(cid:12)U (ei1 , . . . , eim)(cid:12)(cid:12)
2m
m+1
m+1
2m
≤ Cm kUk
m+1
m−1
2m 2
∞ × ··· × ℓN
for every m-linear mapping U : ℓN
∞ → C and every positive integer N (for polynomial
versions of Bohnenblust -- Hille inequality we refer to [10]). The original upper estimate for Cm
, but several improvements have been obtained since then. For instance, as an
is m
illustration for the complex case we compare, below, and for some values of m, the original constants
with the improvements obtained by S. Kaijser [14] and H. Qu´effelec [18], Defant, P. Sevilla-Peris
[8]:
2
Key words and phrases. Absolutely summing operators, Bohnenblust -- Hille Theorem.
2010 Mathematicss Subject Classification: 46G25, 47L22, 47H60.
**Supported by CNPq Grant 620108/2008-8 (Edital Casadinho).
**Supported by the Spanish Ministry of Science and Innovation, grant MTM2009-07848.
1
2
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
([18, 8], 1995) Cm = 2
([14], 1978) Cm = m
([2],1931)
m Cm =(cid:16) 2
3
4
5
6
7
8
9
10
15
50
100
√π(cid:17)m−1
≈ 1.273
≈ 1.437
≈ 1.621
≈ 1.829
≈ 2.064
≈ 2.330
≈ 2.628
≈ 2.965
≈ 5.425
≈ 372
≈ 155, 973
m−1
2
2
8
4
≈ 2.828
≈ 5.657
≈ 11.314
≈ 22.627
16
128
≈ 23, 726, 566
≈ 7.96131459 · 1014
m+1
m−1
2
2m 2
≈ 4.160
≈ 6.726
≈ 10.506
≈ 16.088
≈ 24.322
≈ 36.442
≈ 54.232
≈ 80.283
≈ 542.574
≈ 174, 465, 512
≈ 8.14675743 · 1015
The Bohnenblust -- Hille inequality also holds for real Banach spaces but sharper estimates for
Cm, in this case, seem to be Cm = 2
m−1
2
.
The aim of this paper is show that improved values for Cm are essentially contained in [9] for
both real and complex cases.
The (complex and real) Bohnenblust -- Hille inequality can be re-written in the context of multiple
summing multilinear operators, as we will see next. Multiple summing multilinear mappings
between Banach spaces is a recent, very important and useful nonlinear generalization of the
concept of absolutely summing linear operators. This class was introduced, independently, by
Matos [17] (under the terminology fully summing multilinear mappings) and Bombal, P´erez-Garc´ıa
and Villanueva [3]. The interested reader can also refer to [4, 5] for other Bohnenblust -- Hille type
results.
Throughout this paper X1, . . . , Xm and Y will stand for Banach spaces over K = R or C, and
X′ stands for the dual of X. By L(X1, . . . , Xm; Y ) we denote the Banach space of all continuous
m-linear mappings from X1 × ··· × Xm to Y with the usual sup norm. For x1,··· , xn in X, let
k(xj )n
j=1kw,1 := sup{k(ϕ(xj))n
j=1k1 : ϕ ∈ X′,kϕk ≤ 1}.
If 1 ≤ p < ∞, an m-linear mapping U ∈ L(X1, . . . , Xm; Y ) is multiple (p; 1)-summing (denoted
Π(p;1)(X1, . . . , Xm; Y )) if there exists a constant Lm ≥ 0 such that
(1.1)
N
Xj1,...,jm=1(cid:13)(cid:13)(cid:13)
U (x(1)
j1 , . . . , x(m)
≤ Lm
1
p
p
jm )(cid:13)(cid:13)(cid:13)
m
Yk=1(cid:13)(cid:13)(cid:13)
(x(k)
j
)N
j=1(cid:13)(cid:13)(cid:13)w,1
for every N ∈ N and any x(k)
jk ∈ Xk, jk = 1, . . . , N , k = 1, . . . , m. The infimum of the constants
satisfying (1.1) is denoted by kUkπ(p;1). For m = 1 we have the classical concept of absolutely
(p; 1)-summing operators (see, e.g. [7, 11]).
A simple reformulation of the Bohnenblust -- Hille inequality asserts that every continuous m-
m−1
linear form T : X1 × ··· × Xm → K is multiple ( 2m
(or Lm = (cid:16) 2
√π(cid:17)m−1
although in the real case the best known constants seem to be Cm = 2
m−1
2
.
for the complex case, using the estimates of Defant and Sevilla-Peris, [8]),
m+1 ; 1)-summing with Lm = Cm = 2
2
The main goal of this paper is to calculate better constants for the Bohnenblust -- Hille inequality
in the real and complex case (which are derived from [9]). For this task we will explore the proof
of a general vector-valued version of Bohnenblust -- Hille inequality ([9, Theorem 5.1]).
Let us recall some results that we will need in this note. The first one is a well-known inequality
due to Khinchine (see [11]):
Theorem 1.1 (Khinchine's inequality). For all 0 < p < ∞, there exist constants Ap and Bp such
that
(1.2)
Ap N
Xn=1
1
2
an2!
N
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ Z 1
Xn=1
anrn (t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
1
p
dt!
≤ Bp N
Xn=1
1
2
an2!
NEW UPPER BOUNDS FOR THE CONSTANTS IN THE BOHNENBLUST-HILLE INEQUALITY
3
for every positive integer N and scalars a1, . . . , an (here, rn denotes the n − th Rademacher
function).
Above, it is clear that B2 = 1. From (1.2) it follows that
(1.3)
N
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Z 1
Xn=1
anrn (t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
1
p
dt!
≤ BpA−1
N
Xn=1
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
r Z 1
anrn (t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
r
1
r
dt!
and the product of the constants BpA−1
r will appear later on in Theorem 1.3. The notation Ap
and Bp will be kept along the paper. Next, let us recall a variation of an inequality due to Blei
(see [9, Lemma 3.1]).
Theorem 1.2 (Blei, Defant et al.). Let A and B be two finite non-void index sets, and (aij)(i,j)∈A×B
a scalar matrix with positive entries, and denote its columns by αj = (aij )i∈A and its rows by
βi = (aij )j∈B. Then, for q, s1, s2 ≥ 1 with q > max(s1, s2) we have
s1
Xj∈B
X(i,j)∈A×B
q
kαjks2
≤ Xi∈A
q !
kβiks1
aw(s1,s2)
ij
w(s1 ,s2)
f (s2 ,s1)
f (s1 ,s2)
s2
,
1
with
w : [1, q)2 → [0,∞), w(x, y) :=
f : [1, q)2 → [0,∞), f (x, y) :=
q2(x + y) − 2qxy
q2 − xy
q2x − qxy
q2(x + y) − 2qxy
,
.
The following theorem is a particular case of [9, Lemma 2.2] for Y = K using that the cotype 2
constant of K is 1, i.e., C2(K) = 1 (following the notation from [9]):
Theorem 1.3 (Defant et al). Let 1 ≤ r ≤ 2, and let (yi1,...,im )N
Xi1,...,im=1
yi1,...,im2
≤ (A2,r)m
Xi1,...,im=1
1/2
N
N
where
Z[0,1]m(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ri1 (t1) . . . rim (tm)yi1...im(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
A2,r ≤ A−1
r B2 = A−1
r
(since B2 = 1).
i1,...,im=1 be a matrix in K. Then
1/r
r
dt1 . . . dtm
,
The meaning of A2,r, w and f from the above theorems will also be kept in the next section
and KG will denote the complex Grothendieck constant. Also, and throughout the paper, CK,m
will denote our estimates on the constants for the real or complex case (K = R or C respectively).
2. Improved constants for the Bohnenblust -- Hille theorem: The real case
From the proof of [9, Corollary 5.2] and using the optimal values for the constants of Khinchine's
inequality (due to Haagerup), the following estimates can be calculated for Cm:
(2.1)
(2.2)
CR,2 = √2,
CR,m = 2
m−1
2m CR,m−1
A 2m−2
m !1− 1
m
for m ≥ 3.
In particular, if 2 ≤ m ≤ 13,
(2.3)
If we change a little bit the induction process by obtaining the case m from the cases m − 2 and 2
we obtain even smaller constants. For example,
m2+m−2
4m .
CR,m = 2
(2.4)
CR,m = 2
m−2
m
CR,m−2
A2
2m−4
m−1
1
2
for m > 3.
4
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
In particular, if 2 ≤ m ≤ 14 a careful calculation gives us
(2.5)
CR,m = 2
m2+6m−8
8m
if m is even, and
(2.6)
CR,m = 2
m2+6m−7
8m
if m is odd.
However, in the next theorem a different induction approach leads us to even smaller (and, thus,
sharper) constants.
Remark 2.1. It is worth mentioning that the values for the above constants are not explicitly
calculated in [9] but, of course, are derived from [9]. Since our procedure below will improve the
constants, we will not give much detail on the above estimates.
The paper [9] provides, in fact, a family of constants Cm for Bohnenblust -- Hille inequality. In this
sense the following result is essentially contained in [9]. However, for the sake of completeness we
prefer to sketch the proof from [9]. Our particular approach is chosen to obtain sharper constants:
Theorem 2.2. For every positive integer m and real Banach spaces X1, . . . , Xm,
Π( 2m
m+1 ;1)(X1, . . . , Xm; R) = L(X1, ..., Xm; R) and k.kπ( 2m
m+1 ;1) ≤ CR,m k.k
with
CR,2 = 2
1
2 and CR,3 = 2
5
6 ,
CR,m =
CR,m/2
Am/2
2m
m+2
for m even and
CR,m =
for m odd. In particular,
CR, m−1
2
m+1
A
2
2m−2
m+1
f ( 2m−2
m+1 , 2m+2
m+3 )
.
CR, m+1
2
m−1
A
2
2m+2
m+3
CR,m = 2
1
2 CR,m/2
f ( 2m+2
m+3 , 2m−2
m+1 )
if m is even and 2 ≤ m ≤ 24.
Proof. We start with the case m even. The cases m = 2 (Littlewood's 4/3-inequality) and m = 3
are known. We proceed by induction, obtaining the case m as a combination of the cases m/2 and
m/2 (instead of 1 and m − 1 as in [9, Corollary 5.2]).
Suppose that the result is true for m/2 and let us prove for m. Let U ∈ L(X1, . . . , Xm; R) and N
be a positive integer. For each 1 ≤ k ≤ m consider x(k)
1 , . . . , x(k)
(x(k)
)N
j
N ∈ Xk so that (cid:13)(cid:13)(cid:13)
j=1(cid:13)(cid:13)(cid:13)w,1 ≤ 1,
k = 1, .., m.
Consider, following the notation of Theorem 1.2,
Thus,
q = 2, s1 = s2 =
2 · ( m
2 )
m
2 + 1
=
2m
m + 2
.
w(s1, s2) =
2m
m + 1
NEW UPPER BOUNDS FOR THE CONSTANTS IN THE BOHNENBLUST-HILLE INEQUALITY
5
and, from Theorem 1.2, we have
(m+1)/2m
N
N
2m
≤
U (x(1)
i1 , . . . , x(m)
im )(cid:12)(cid:12)(cid:12)
Xi1,...,im=1(cid:12)(cid:12)(cid:12)
≤
Xi1,...,im/2=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
Xi(m/2)+1,...,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
m+1
2
i(m/2)+1,...,im=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
i1 , . . . , x(m)
2
i1,...,im/2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
i1 , . . . , x(m)
s1
s2
N
f (s2,s1)/s2
f (s1,s2)/s1
.
Now we need to estimate the two factors above. We will write dt := dt1 . . . dtm/2. For each
i(m/2)+1, ..., im fixed, we have (from Theorem 1.3),
ri1 (t1) . . . rim/2 (tm/2)U (x(1)
i1 , . . . , x(m)
dt
ri1 (t1)x(1)
i1 , . . . ,
rim/2 (tm/2)x(m/2)
im/2
, x(m/2)+1
i(m/2)+1
s1
im )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
. . . , x(m−1)
im
im−1 , x(m)
s1
dt.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N
Xim/2=1
rim/2 (tm/2)x(m/2)
im/2
, . . . , x(m−1)
im
s1
dt.
, x(m)
im
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
rim/2 (tm/2)x(m/2)
im/2
, . . .
dt
s1
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
π(s1;1)
Summing over all i(m/2)+1, . . . , im = 1, . . . , N we obtain
N
Xim/2=1
s1
2 ≤
N
i1,...,im/2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
U
Xi1=1
ri1 (t1)x(1)
i1 , . . . ,
s1
2 ≤
i1,...,im/2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
Xim/2=1
ri1 (t1)x(1)
i1
, . . . ,
N
s1
2 ≤
N
N
i1 , . . . , x(m)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
2,s1(cid:17)s1
≤(cid:16)Am/2
2,s1(cid:17)s1
=(cid:16)Am/2
i1,...,im/2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Z[0,1]m/2
Xi1,...,im/2=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
U
Z[0,1]m/2
Xi1=1
Xi(m/2)+1,...,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
2,s1(cid:17)s1
(cid:16)Am/2
Xi(m/2)+1,...,im=1
Z[0,1]m/2
i1 , . . . , x(m)
N
N
Using the case m/2 we thus have
Hence
N
N
Xi(m/2)+1,...,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
i1 , . . . , x(m)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
U
2,s1(cid:17)s1
Z[0,1]m/2
≤(cid:16)Am/2
Xi1=1
2,s1(cid:17)s1
≤(cid:16)Am/2
(cid:0)kUk CR,m/2(cid:1)s1 .
Xi(m/2)+1,...,im=1(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)U (x(1)
m+1
im )(cid:12)(cid:12)(cid:12)
i1 , . . . , x(m)
U (x(1)
2m
N
N
Xi1,...,im=1(cid:12)(cid:12)(cid:12)
The other estimate is exactly the same. Hence, combining both estimates, we obtain
2,s1 kUk CR,m/2.
i1 , . . . , x(m)
1
s1
s1
≤ Am/2
2
i1,...,im/2=1(cid:13)(cid:13)(cid:13)(cid:13)
im )(cid:17)N
2,s1 kUk CR,m/2i1/2hAm/2
≤hAm/2
(m+1)/2m
2,s1 kUk CR,m/2i1/2
.
6
and
Hence
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
N
Xi1,...,im=1(cid:12)(cid:12)(cid:12)
U (x(1)
i1 , . . . , x(m)
(m+1)/2m
2m
m+1
im )(cid:12)(cid:12)(cid:12)
2,s1 kUk CR,m/2
≤ Am/2
= Am/2
2, 2m
m+2 kUk CR,m/2.
Now let us estimate the constants CR,m. We know that B2 = 1 and, from [12], we also know
CR,m = Am/2
2, 2m
m+2
CR,m/2.
that Ap = 2
(2.7)
1
2 − 1
p whenever p ≤ 1.847. So, for 2 ≤ m ≤ 24 we have
A 2m
m+2
= 2
1
2 −
m+2
2m = 2
−1
m .
Hence, from (1.3) and using the best constants of Khinchine's inequality from [12], we have
and
A2, 2m
m+2 ≤ A−1
2m
m+2
1
m
= 2
CR,m ≤(cid:16)2
1
m(cid:17)m/2
CR,m/2 = 2
1
2 CR,m/2
for m even, 2 ≤ m ≤ 24.
The numerical values of CR,m, for m > 24, can be easily calculated by using the exact values of
A 2m
m+2
(see [12]):
A 2m
m+2
= √2
Γ(cid:16) 2m
m+2 +1
√π
2
(cid:17)
(m+2)/2m
.
For the case m odd we proceed by induction, but the case m is obtained as a combination of
the cases m−1
2 with m+1
2
instead of 1 and m − 1 as in [9, Corollary 5.2].
Consider, in the notation of Theorem 1.2,
q = 2, s1 =
Thus,
and a similar proof gives us
2(cid:0) m−1
2 (cid:1)
2 + 1
m−1
=
2m − 2
m + 1
and s2 =
w(s1, s2) =
2m
m + 1
2(cid:0) m+1
2 (cid:1)
2 + 1
m+1
=
2m + 2
m + 3
.
m+1
2
2, 2m−2
m+1
CR, m−1
CR,m =(cid:18)A
≤
CR, m−1
2
m+1
2
2m−2
m+1
A
f ( 2m−2
m+1 , 2m+2
m+3 )
m+1 , 2m+2
m+3 )
2 (cid:19)f ( 2m−2
.
.(cid:18)A
2
CR, m+1
2
m−1
A
2m+2
m+3
m+3 , 2m−2
m+1 )
m−1
2
2, 2m+2
m+3
CR, m+1
f ( 2m+2
m+3 , 2m−2
m+1 )
2 (cid:19)f ( 2m+2
(cid:3)
In the below table we compare the first constants Cm = 2
derived from [9, Cor. 5.2] with the new constants CR,m:
m−1
2
and the constants that can be
NEW UPPER BOUNDS FOR THE CONSTANTS IN THE BOHNENBLUST-HILLE INEQUALITY
7
m−1
2
([14], 1978)
18
28
40
2
≈ 1.782
m New constants - CR,m Cm derived from ([9, Corollary 5.2], 2010) Cm = 2
3
4
5
6
7
8
9
10
11
12
25/6 ≈ 1.782
2
16 ≈ 2.18
2
20 ≈ 2.639
2
24 ≈ 3.17
28 ≈ 3.807
2
2
32 ≈ 4.555
36 ≈ 5.443
2
2
40 ≈ 6.498
44 ≈ 7.752
2
48 ≈ 9.243
2
≈ 2.298
≈ 2.520
≈ 2.6918
≈ 2.8284
≈ 3.055
≈ 3.249
≈ 3.4174
≈ 3.563
22/2 = 2
23/2 ≈ 2.828
25/2 ≈ 5.656
26/2 = 8
27/2 ≈ 11.313
28/2 = 16
29/2 ≈ 22.627
210/2 = 32
211/2 ≈ 45.254
22 = 4
108
130
154
54
70
88
3. Improved constants for the Bohnenblust -- Hille theorem: The complex case
As in the real case, following the proof of [9, Corollary 5.2] and using the optimal values for
the constants of Khinchine's inequality (due to Haagerup) and using that KG = CC,2 (see [1]), the
following estimates can be calculated for Cm:
CC,2 = KG ≤ 1.4049 < √2,
m !1− 1
2m CC,m−1
A 2m−2
CC,m = 2
m−1
m
for m ≥ 3,
In particular, if 2 ≤ m ≤ 13,
CC,m ≤ 2
m2+m−6
4m K 2/m
G .
The above estimates improve the values CC,m = 2
m−1
2 but are worst than the constants CC,m =
obtained by Defant and Sevilla-Peris [8]. However, our approach will provide even better
√π(cid:17)m−1
(cid:16) 2
constants.
The following lemma is essentially the main result from the previous section which comes from
[9], although now we will obtain different constants, since we will be dealing with the complex case.
Lemma 3.1. For every positive integer m and complex Banach spaces X1, . . . , Xm,
Π( 2m
m+1 ;1)(X1, . . . , Xm; C) = L(X1, . . . , Xm; C) and k.kπ( 2m
m+1 ;1) ≤ CC,m k.k
with
CC,m =(cid:18) 2
√π(cid:19)m−1
for m = 2, 3,
CC,m =
CC,m/2
Am/2
2m
m+2
for m even and
for m odd.
CC,m =
CC, m−1
2
m+1
A
2
2m−2
m+1
f ( 2m−2
m+1 , 2m+2
m+3 )
f ( 2m+2
m+3 , 2m−2
m+1 )
CC, m+1
2
m−1
A
2
2m+2
m+3
The proof of this result is essentially in the same spirit as that of Theorem 2.2. The cases of
CC,2 and CC,3 are already known and the proof is (also) done by induction.
8
D. PELLEGRINO AND J. B. SEOANE-SEP ´ULVEDA
from [8] are better
than the constants that we have obtained in the previous lemma. However our constants present a
3.1. Comparing the "first" constants. The first constants Dm =(cid:16) 2
smaller asymptotical growth and, from a certain level m on, they are better than Dm =(cid:16) 2
As we see next, this occurs when m ≥ 7.
Here below we compare the first constants. For the case m = 4, notice that we have
√π(cid:17)m−1
√π(cid:17)m−1
.
CC,4 = 21/2 · CC,2 = 21/2 ·(cid:18) 2
√π(cid:19) ≈ 1.5957
but, for m = 7 we have
≈ 1.437. So, in order to improve the constants that follow,
instead of 21/2 · CC,2 for the value of CC,4.
√π(cid:17)3
but this constant is worst than (cid:16) 2
√π(cid:17)3
it would be better to consider (cid:16) 2
Similarly, for the cases 5 ≤ m ≤ 6 we also have that CC,m is slightly worst than (2/√π)m−1
√π(cid:17)2
12(cid:17)4
(cid:16) 2
(cid:16)2 1
√π(cid:17)3
16(cid:17)3
(cid:16) 2
(cid:16)2 1
CC,7 =
= 1.9293 <(cid:18) 2
and our constants are better than the old ones. Also, as we announced, for m ≥ 7 our constants
also improve the old ones.
√π(cid:19)6
·
2 − 10
2 − 8
3/7
4/7
In the next section we state the previous lemma using the information we just obtained.
3.2. Comparing the "remaining" constants (m > 7). Now it is time to state the last lemma
adding the better constants:
Theorem 3.2. For every positive integer m and every complex Banach spaces X1, . . . , Xm,
Π( 2m
m+1 ;1)(X1, . . . , Xm; C) = L(X1, . . . , Xm; C) and k.kπ( 2m
m+1 ;1) ≤ CC,m k.k
with
for m ∈ {2, 3, 4, 5, 6},
for m > 6 even, and
√π(cid:19)m−1
CC,m/2
Am/2
2m
m+2
• CC,m =(cid:18) 2
• CC,m =
• CC,m =
A
CC, m−1
2
m+1
2
2m−2
m+1
f ( 2m−2
m+1 , 2m+2
m+3 )
f ( 2m+2
m+3 , 2m−2
m+1 )
for m > 5 odd.
The following table compares these new constants with the previous ones:
CC, m+1
2
m−1
A
2
2m+2
m+3
([18, 8], 1995)
·
√π(cid:17)m−1
≈ 2.329
≈ 2.628
≈ 2.965
≈ 3.346
≈ 3.775
≈ 4.260
≈ 4.807
≈ 5.425
≈ 6.121
m New Constants - CC,m (cid:16) 2
8
9
10
11
12
13
14
15
16
≈ 2.031
≈ 2.172
≈ 2.292
≈ 2.449
≈ 2.587
≈ 2.662
≈ 2.728
≈ 2.805
≈ 2.873
m−1
2
2
([14],1978) m
32
16
≈ 11.313
≈ 22.627
≈ 45.425
≈ 90.509
≈ 181.019
128
64
([2],1931)
m+1
m−1
2
2m 2
≈ 36.442
≈ 54.232
≈ 80.283
≈ 118.354
≈ 173.869
≈ 254.680
≈ 372.128
≈ 542.574
≈ 789.612
NEW UPPER BOUNDS FOR THE CONSTANTS IN THE BOHNENBLUST-HILLE INEQUALITY
9
Acknowledgements. The authors thank A. Defant and P. Sevilla-Peris for important remarks
on the complex case. The authors are also very grateful to the referee, whose thorough analysis
and insightful remarks helped to obtain even sharper constants.
References
[1] O. Blasco, G. Botelho, D. Pellegrino and P. Rueda, Summability of multilinear mappings: Littlewood, Orlicz
and beyond, Monatsh. Math. 163 (2011), 131-147.
[2] H.F. Bohnenblust and E. Hille, On the absolute convergence of Dirichlet series, Ann. of Math. 32 (1931),
600-622.
[3] F. Bombal, D. P´erez-Garc´ıa and I. Villanueva, Multilinear extensions of Grothendieck's theorem, Q. J. Math.
55 (2004), 441-450.
[4] G. Botelho, H.A. Braunss, H. Junek, D. Pellegrino, Inclusions and coincidences for multiple summing multilinear
mappings, Proc. Amer. Math. Soc. 137 (2009), 991 -- 1000.
[5] G. Botelho, C. Michels and D. Pellegrino, Complex interpolation and summability properties of multilinear
operators, Rev. Mat. Complut. 23 (2010), no, 139 -- 161.
[6] A.M. Davie, Quotient algebras of uniform algebras, J. London Math. Soc. 7 (1973), 31-40.
[7] A. Defant and K. Floret, Tensor norms and operator ideals, North Holland Math. stud., vol. 176, 1993.
[8] A. Defant and P. Sevilla-Peris, A new multilinear insight on Littlewood's 4/3-inequality, J. Funct. Anal. 256
(2009), 1642-1664.
[9] A. Defant. D. Popa, U. Schwarting, Coordinatewise multiple summing operators in Banach spaces, J. Funct.
Anal. 259 (2010), 220-242.
[10] A. Defant, L. Frerick, J. Ortega-Cerd´a, M. Ounaıes, K. Seip, The Bohnenblust -- Hille inequality for homogeneous
polynomials is hypercontractive, to appear in Ann. Math (2).
[11] J. Diestel, H. Jarchow and A. Tonge, Absolutely summing operators, Cambridge University Press, Cambridge,
1995.
[12] U. Haagerup, The best constants in Khinchine inequality, Studia Math. 70 (1982), 231-283.
[13] U. Haagerup, A new upper bound for the complex Grothendieck constant, Israel J. Math. 60 (1987), 199-224.
[14] S. Kaijser, Some results in the metric theory of tensor products, Studia Math. 63 (1978), 157-170.
[15] J. Lindenstrauss and A. Pe lczy´nski, Absolutely summing operators in Lp-spaces and their applications, Studia
Math. 29 (1968), 275-325.
[16] J.E. Littlewood, On bounded bilinear forms in an infinite number of variables, Q. J. Math. 1 (1930), 164-174.
[17] M.C. Matos. Fully absolutely summing and Hilbert-Schmidt multilinear mappings, Collectanea Math. 54
(2003), 111-136.
[18] H. Queff´elec, H. Bohr's vision of ordinary Dirichlet series: old and new results, Journal of Analysis 3 (1995),
43-60.
[19] J. Sawa, The best constant in the Khinchine inequality for complex Steinhaus variables, the case p = 1, Studia
Math. 81 (1985), 107-126.
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]
|
1901.03853 | 1 | 1901 | 2019-01-12T11:48:24 | Caristi-Kirk and Oettli-Th\'era Ball Spaces and applications | [
"math.FA"
] | Based on the theory of ball spaces introduced by Kuhlmann and Kuhlmann we introduce and study Caristi-Kirk and Oettli-Th\'era ball spaces. We show that if the underlying metric space is complete, then these have a very strong property: every ball contains a singleton ball. This fact provides quick proofs for several results which are equivalent to the Caristi-Kirk Fixed Point Theorem, namely Ekeland's Variational Principles, the Oettli-Th\'era Theorem, Takahashi's Theorem and the Flower Petal Theorem. | math.FA | math | CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
AND APPLICATIONS
PIOTR BŁASZKIEWICZ, HANNA ĆMIEL,
ALESSANDRO LINZI, PIOTR SZEWCZYK
Abstract. Based on the theory of ball spaces introduced by Kuhlmann
and Kuhlmann we introduce and study Caristi -- Kirk and Oettli -- Théra
ball spaces. We show that if the underlying metric space is complete,
then these have a very strong property: every ball contains a single-
ton ball. This fact provides quick proofs for several results which are
equivalent to the Caristi -- Kirk Fixed Point Theorem, namely Ekeland's
Variational Principles, the Oettli -- Théra Theorem, Takahashi's Theorem
and the Flower Petal Theorem.
9
1
0
2
n
a
J
2
1
]
.
A
F
h
t
a
m
[
1
v
3
5
8
3
0
.
1
0
9
1
:
v
i
X
r
a
1. Introduction
1.1. General setting. The literature on complete metric spaces contains
remarkable results such as the Theorem of Caristi and Kirk ([2] and [4]),
Ekeland's Principle ([3]), Takahashi's Theorem ([12]) and the Flower Petal
Theorem ([11]). These theorems are known to be equivalent (see, e.g., [11],
[10]). Their statements can be found in Section 4.
The concept of a ball space was first introduced by F.-V. and K. Kuhlmann
in [6],[7]. In [8] they connected it with the Caristi -- Kirk Fixed Point Theo-
rem (FPT) by providing a way to prove it using ball spaces techniques. In
this paper we further develop this connection by proving Theorem 2 which
provides a generic method to obtain simple proofs of all the results men-
tioned in the previous paragraph and, possibly, related ones in the future.
In [10], Oettli and Théra introduced an alternative approach to the Caristi --
Kirk Theorem and showed it to be equivalent to what was later (in publica-
tions such as [9]) called Oettli -- Théra Theorem. Our method can be applied
to easily prove this theorem as well as the theorems equivalent to it, which
are stated in [10] (see Section 3).
1.2. Ball spaces. As in [8], by a ball space we mean a pair (X, B), where
X is a nonempty set and B ⊆ P(X) is a nonempty family of nonempty
subsets of X. An element B ∈ B is called a ball. If no confusion arises, we
will write B in place of (X, B) when speaking of a ball space.
2010 Mathematics Subject Classification. Primary 54H25; Secondary 47H09, 47H10.
Key words and phrases. metric space, ball space, Caristi -- Kirk Fixed Point Theorem,
Ekeland's Variational Principle, Oettli -- Théra Theorem, Takahashi's Theorem, Flower
Petal Theorem.
1
2
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
A nest of balls in a ball space B is a nonempty family N of balls from
B which is totally ordered by inclusion. We say that a ball space B is
spherically complete if for every nest of balls N ⊆ B we have T N 6= ∅.
Further details about ball spaces may be found in [5].
Definition 1. A ball space (X, B) is strongly contractive if there is a func-
tion that associates to every x ∈ X some ball Bx ∈ B such that, for every
x, y ∈ X, the following conditions hold:
(1) x ∈ Bx;
(2) if y ∈ Bx then By ⊆ Bx;
(3) if y ∈ Bx \ {x} then By ( Bx.
This particular type of ball spaces has a remarkable property, stated in
the following theorem.
Theorem 2. In every spherically complete, strongly contractive ball space
every ball Bx contains a singleton ball. In other words, there exists a ∈ Bx
such that Ba = {a}.
Proof. Let B be a strongly contractive, spherically complete ball space and
Bx ∈ B any ball. Consider the family
A = {N ⊆ P(Bx) N is a nest of balls in B}.
This family is partially ordered by inclusion and nonempty since {Bx} ∈ A.
If we have a chain of nests in A, the union of that chain is again a nest of
balls in A, hence an upper bound of the chain. By Zorn's Lemma we obtain
the existence of a maximal nest M ∈ A. Since the space is spherically
by condition (2) of Definition 1 also Ba ⊆ B for every B ∈ M and so
complete, there exists an element a ∈T M. Since a ∈ B for every B ∈ M,
Ba ⊆ T M. This means that M ∪ {Ba} is a nest of balls in A which
contains M. By maximality of M we get that M ∪ {Ba} = M, i.e.,
Ba ∈ M. Now we wish to show that Ba is a singleton. Suppose that there
exists an element b ∈ Ba \ {a}. Then Bb ( Ba (in particular, Ba 6⊆ Bb)
and so Bb /∈ M. But this means that M ∪ {Bb} is a nest of balls that
properly contains M, which contradicts the maximality of M. Therefore,
Ba = {a}.
(cid:3)
2. Caristi -- Kirk and Oettli -- Théra ball spaces
In this section, we will be working with a nonempty metric space (X, d).
2.1. Caristi -- Kirk ball spaces. Consider a function ϕ : X → R, a point
x ∈ X and the following set:
Bϕ
x = {y ∈ X d(x, y) ≤ ϕ(x) − ϕ(y)}.
x ), we may think of this set as a ball and
x 6= ∅ (because x ∈ Bϕ
Since Bϕ
consider the ball space (X, Bϕ) where
Bϕ := {Bϕ
x x ∈ X} .
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
3
We will call the function ϕ a Caristi -- Kirk function on X if it is lower
semicontinuous, that is,
∀
y∈X
lim inf
x→y
ϕ(x) ≥ ϕ(y),
and bounded from below, that is,
inf
x∈X
ϕ(x) > −∞.
The corresponding balls Bϕ
x have been introduced in [8] as the Caristi -- Kirk
balls and Bϕ is the induced Caristi -- Kirk ball space. For brevity, we will
write CK in place of Caristi -- Kirk.
A number of remarkable properties of the balls defined above, given in
the following lemma, can be found in [8].
Lemma 3. Take a metric space (X, d) and any function ϕ : X → R. Then
the following assertions hold.
(1) For every x ∈ X, x ∈ Bϕ
x .
(2) If y ∈ Bϕ
y ⊆ Bϕ
x ; if in addition x 6= y, then Bϕ
x then Bϕ
y ( Bϕ
x and
ϕ(y) < ϕ(x).
(3) If ϕ is lower semicontinuous, then all CK balls Bx are closed in the
topology induced by the metric.
Lemma 3 immediately yields the following result.
Corollary 4. The CK ball space Bϕ is strongly contractive.
Another important fact about CK ball spaces may also be found in [8]:
Proposition 5. Let (X, d) be a metric space. Then the following statements
are equivalent:
(i) The metric space (X, d) is complete.
(ii) Every CK ball space (X, Bϕ) is spherically complete.
(iii) For every continuous function ϕ : X → R bounded from below, the
CK ball space (X, Bϕ) is spherically complete.
2.2. Oettli -- Théra ball spaces.
Definition 6. A function φ : X × X → (−∞, +∞] is an Oettli -- Théra
function on X if the following properties hold:
(a) φ(x, ·) : X → (−∞, +∞] is lower semicontinous for all x ∈ X;
(b) φ(x, x) = 0 for all x ∈ X;
(c) φ(x, y) ≤ φ(x, z) + φ(z, y) for all x, y, z ∈ X;
φ(x0, x) > −∞.
there exists x0 ∈ X s.t.
(d)
inf
x∈X
If an element x0 ∈ X satisfies property (d), we will call it an Oettli -- Théra
element for φ in X. If it is clear which space is considered, we will say that
4
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
φ is an Oettli -- Théra function and that x0 is an Oettli -- Théra element for φ.
For brevity, we will write OT in place of Oettli -- Théra.
Definition 7. Let φ be an OT function on X.
(i) The OT ball of x ∈ X is:
Bφ
x := {y ∈ X d(x, y) ≤ −φ(x, y)}.
If no confusion arises as to which OT function is considered, we will
x . This gives rise to a ball space (X, Bφ),
write Bx in place of Bφ
where
Bφ := {Bx x ∈ X}.
(ii) The OT ball space generated by an OT element x0 is (Bx0, Bφ
x0) where
Bφ
x0 := {Bx x ∈ Bx0}.
In this subsection, if an OT element x0 ∈ X has been fixed, we
will write for brevity B0 in place of Bx0.
It is worth noting that if we are given a CK function ϕ, we may define φ
by:
(1)
φ(x, y) := ϕ(y) − ϕ(x).
The following fact is straightforward to prove.
Fact 8. If ϕ is a CK function, then the function φ : X × X → R defined
in (1) is an OT function. Moreover, every x ∈ X is an OT element for φ.
As we know from Corollary 4, the CK ball space is strongly contractive.
A similar result can be shown for the OT ball space.
Lemma 9. Take a metric space (X, d) and φ : X × X → R a function
satisfying (b) and (c) in Definition 6. Then the following assertions hold,
for every x ∈ X.
(1) x ∈ Bx.
(2) If y ∈ Bx then By ⊆ Bx.
(3) If y ∈ Bx \ {x} then By ( Bx and φ(x, y) < φ(y, x).
Proof.
(1): Indeed, d(x, x) = −φ(x, x) = 0.
(2): Take y ∈ Bx, i.e.,
Take any z ∈ By, then
d(x, y) ≤ −φ(x, y).
d(y, z) ≤ −φ(y, z).
By condition (c) for an OT function we get
d(x, z) ≤ d(x, y) + d(y, z) ≤ −φ(x, y) − φ(y, z) ≤ −φ(x, z),
so z ∈ Bx and, as a result, By ⊆ Bx.
(3): Let y ∈ Bx and y 6= x. We wish to show that x /∈ By. Suppose that
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
5
x ∈ By. Then d(y, x) ≤ −φ(y, x) and by conditions (b) and (c) for an OT
function we get
0 < d(y, x) + d(x, y) ≤ −φ(y, x) − φ(x, y) ≤ −φ(y, y) = 0,
contradiction. Thus x /∈ By and so By ( Bx. Clearly, this also implies
−φ(y, x) < d(x, y) ≤ −φ(x, y).
(cid:3)
Lemma 9 instantly yields the following corollary.
Corollary 10. For an OT function φ on X, the ball space Bφ is strongly
contractive. Furthermore, for a fixed OT element x0 for φ in X the OT ball
space Bφ
x0 is also strongly contractive and all of its balls are contained in B0.
As stated in Fact 8, for the OT function φ defined in (1) every x ∈ X is
an OT element. While this doesn't have to be true in general for any OT
function φ, this property turns out to be 'hereditary' in the following sense.
Lemma 11. Let φ be an OT function. If x0 ∈ X is an OT element for φ
in X and x ∈ B0 then also x is an OT element for φ in X.
Proof. Let r ∈ R be such that
inf
y∈X
φ(x0, y) ≥ r.
Take any x ∈ B0. Note that 0 ≤ d(x0, x) ≤ −φ(x0, x). For every y ∈ X we
have
r ≤ φ(x0, y) ≤ φ(x0, x) + φ(x, y),
so
In particular,
φ(x, y) ≥ r − φ(x0, x).
inf
y∈X
φ(x, y) ≥ r − φ(x0, x) ≥ r.
(cid:3)
As stated in Proposition 5, there is an equivalence between complete-
ness of a metric space and spherical completeness of the respective CK ball
spaces. A similar result can be shown for the OT ball spaces. For that we
will need to state an auxiliary lemma first.
Lemma 12. Let (X, d) be a metric space, φ an OT function on X and x0
an OT element for φ in X. Moreover, let N ⊆ Bφ
x0 be a nest of balls and
write N = {Bx x ∈ A} for some set A ⊆ B0. Then for every x, y ∈ A we
have
(2)
d(x, y) ≤ φ(x0, x) − φ(x0, y).
Moreover, the following statements are equivalent for every x, y ∈ A:
(i) y ∈ Bx,
6
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
(ii) φ(x, y) ≤ φ(y, x),
(iii) φ(x0, y) ≤ φ(x0, x).
Proof. For every x, y ∈ A either x ∈ By or y ∈ Bx since N is a nest, so
(3)
d(x, y) ≤ max{−φ(x, y), −φ(y, x)}.
If the above maximum is equal to −φ(x, y), we have
d(x, y) ≤ −φ(x, y) ≤ φ(x0, x) − φ(x0, y) ≤ φ(x0, x) − φ(x0, y).
Similarly, if the maximum is equal to −φ(y, x), we have
d(x, y) ≤ −φ(y, x) ≤ φ(x0, y) − φ(x0, x) ≤ φ(x0, x) − φ(x0, y).
Either way, we deduce (2).
To prove (i) ⇔ (ii) assume y ∈ Bx. If y = x then (ii) is trivial. If y 6= x
then, by assertion (3) of Lemma 9 we have
−φ(y, x) < −φ(x, y).
Hence (ii) follows. Conversely, if y /∈ Bx (in particular, y 6= x) then x ∈ By \
{y}. As a result, again by assertion (3) of Lemma 9, −φ(x, y) < −φ(y, x).
To prove (i) ⇔ (iii) we proceed as follows. If y ∈ Bx then
0 ≤ d(x, y) ≤ −φ(x, y) ≤ −φ(x0, y) + φ(x0, x),
thus φ(x0, x) ≥ φ(x0, y). For the converse, if y /∈ Bx then x ∈ By and so
0 < d(x, y) ≤ −φ(y, x) ≤ −φ(x0, x) + φ(x0, y),
hence φ(x0, x) < φ(x0, y).
(cid:3)
Proposition 13. A metric space (X, d) is complete if and only if the OT
ball space (Bφ
x0) is spherically complete for every OT function φ on X
and every OT element x0 for φ in X.
x0, Bφ
Proof. Suppose that for every OT function φ and every OT element x0 for
φ in X the ball space (B0, Bφ
x0) is spherically complete. We wish to show
that the ball space (X, Bϕ) is spherically complete for every CK function ϕ
on X, which by Proposition 5 will yield the completeness of the space X.
Take any CK function ϕ on X, consider the ball space (X, Bϕ) and fix
any nest of balls N in Bϕ. Pick some Bϕ
x0 ∈ N and consider the nest
N0 = {B ∈ N B ⊆ Bϕ
x0}.
By Fact 8 x0 is an OT element for the OT function φ defined as in (1).
x ⊆ Bϕ
Moreover, for every x ∈ X such that Bϕ
x , hence
x0, Bφ
N0 is a nest in the OT ball space (Bφ
x0). By assumption we then obtain
x0, we have Bϕ
x = Bφ
that ∅ 6=T N0 =T N .
For the converse, assume that X is complete. Fix any OT function φ on
X and any OT element x0 for φ in X. Take a nest of balls N in the ball
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
7
space Bφ
on x0 there exists
x0 and write N = {Bx x ∈ A} for some set A ⊆ B0. By assumption
Let (xn)n∈N be a sequence of elements in A such that
r := inf
x∈A
φ(x0, x) ∈ R.
lim
n→∞
φ(x0, xn) = r.
Then (φ(x0, xn))n∈N is a Cauchy sequence because it converges to r. By
Lemma 12 the sequence (xn)n∈N is also Cauchy. Since X is complete, (xn)n∈N
converges to an element z ∈ X. We want to show that z ∈ T N or,
equivalently, that z ∈ Bx for every x ∈ A. Fix an arbitrary element x ∈ A.
If φ(x0, x) = r (in particular, the infimum is achieved) then by Lemma 12
we get that x = z, because
d(xn, x) ≤ φ(x0, xn) − φ(x0, x) = φ(x0, xn) − r → 0,
showing that x is a limit of (xn)n∈N. Hence in this case we obtain that
z ∈ Bx trivially. Therefore we may assume that φ(x0, x) > r. Then from the
definition of (xn)n∈N we obtain the existence of N ∈ N such that, for every
n ≥ N, we have φ(x0, xn) ≤ φ(x0, x). This, by Lemma 12, is equivalent to
φ(x, xn) ≤ φ(xn, x). Therefore, for every n ≥ N,
d(x, xn) ≤ max{−φ(x, xn), −φ(xn, x)} = −φ(x, xn),
where the first inequality is deduced similarly to (3). Taking lim sup on
both sides we get
d(x, z) ≤ lim sup
−φ(x, xn) ≤ −φ(x, z),
n→∞
claimed.
(cid:3)
so that z ∈ Bx. Since x ∈ A was an arbitrary element, we get z ∈ T N as
Remark 14. Proposition 13 does in general not hold for the ball space Bφ
in place of Bφ
x0. Take the complete metric space R, where d(x, y) = x − y,
and consider φ : R × R → R defined as:
φ(x, y) =(x − y
0
if x 6= 0
if x = 0
.
This is an OT function and yields balls of the form
Bx = {y ∈ X x − y ≤ y − x} = [x, ∞).
In the corresponding ball space we have a nest of balls with empty intersec-
tion, namely,
{[n, +∞) n ∈ N}.
Armed with the theory introduced so far, we can prove an important
property of OT (and as a result, also CK) ball spaces in a complete metric
space.
8
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
Proposition 15. Let (X, d) be a complete metric space.
(1) If φ is an OT function on X then for every OT element x0 for φ in
X there exists an element a ∈ B0 such that Bφ
a = {a}.
(2) If ϕ is a CK function on X then for every x ∈ X there exists a ∈ Bϕ
x
such that Bϕ
a = {a}.
Proof. Assertion (2) follows from assertion (1) by Fact 8. To prove assertion
(1) let φ, x0 and B0 be as in the assumption of the Proposition. By Propo-
sition 13 the OT ball space Bφ
x0 is spherically complete, and by Corollary
10 it is strongly contractive. Theorem 2 yields the result.
(cid:3)
2.3. Generalized Caristi -- Kirk ball spaces. Consider a function ϕ :
X → (−∞, +∞] which is lower semicontinuous, bounded from below and
not constantly equal to +∞. We will call such ϕ a CK ∞ function on X.
In this setting we may define the CK ∞ balls as follows:
B ϕ
x := {x ∈ X ϕ(y) + d(x, y) ≤ ϕ(x)}.
If an element x0 ∈ X satisfies ϕ(x) < +∞, we will call it a CK element for
ϕ in X (or simply a CK element).
An easy observation is that every CK function is a CK∞ function. How-
ever, for a CK∞ function ϕ, setting
(4)
φ(x, y) := ϕ(y) − ϕ(x)
as we did in the CK case (1), may not make sense.
ϕ(x) = +∞ = ϕ(y) there is no natural choice for the value of φ(x, y).
Indeed, in the case
In this subsection, if a CK element x0 is fixed, we will write B0
in place of B ϕ
x0.
For a CK element x0 we define the CK∞ ball space generated by x0 as the
ball space (B0, B ϕ
x0), where:
B ϕ
x0 := {B ϕ
x x ∈ B0}.
Note that in general the ball space {B ϕ
x x ∈ X} is not strongly contractive.
Indeed, if x, y ∈ X, x 6= y, satisfy ϕ(x) = ϕ(y) = +∞, then B ϕ
x = X = B ϕ
y .
However, if we work inside a CK∞ ball space, strong contractiveness holds,
as stated in the following lemma.
Lemma 16. Take a metric space (X, d), any function ϕ : X → (−∞, +∞].
Then the following assertions hold for every x ∈ X.
(1) x ∈ B ϕ
x .
(2) If y ∈ B ϕ
(3) Let x ∈ X be such that ϕ(x) < +∞ and let y ∈ B ϕ
x and ϕ(y) ≤ ϕ(x).
x then B ϕ
y ⊆ B ϕ
x \ {x}. Then
B ϕ
y ( B ϕ
x and ϕ(y) < +∞.
Proof.
(1): Indeed, ϕ(x) + d(x, x) = ϕ(x).
(2): If ϕ(x) = +∞ then B ϕ
x = X and B ϕ
y ⊆ B ϕ
x as well as ϕ(y) ≤ ϕ(x)
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
9
trivially. Now assume that ϕ(x) < +∞ and y ∈ B ϕ
because
x . Then also ϕ(y) < +∞
ϕ(y) ≤ ϕ(x) − d(x, y) ≤ ϕ(x) < +∞.
Take any z ∈ B ϕ
ϕ(z) < +∞ and therefore we may write
y . Through the same reasoning as above, we can see that
d(x, z) ≤ d(x, y) + d(y, z)
≤ ϕ(x) − ϕ(y) + ϕ(y) − ϕ(z)
= ϕ(x) − ϕ(z),
y ⊆ B ϕ
x .
x . Since z ∈ B ϕ
which shows that z ∈ B ϕ
B ϕ
(3): Assume that ϕ(x) < +∞ and y ∈ B ϕ
lemma we know that ϕ(y) < +∞ and B ϕ
that case
y was arbitrary, we deduce that
x \ {x}. From assertion (2) of our
y ⊆ B ϕ
y . In
x . Suppose that x ∈ B ϕ
and
ϕ(y) + d(x, y) ≤ ϕ(x)
ϕ(x) + d(x, y) ≤ ϕ(y).
Adding up these inequalities, taking into account that ϕ(x) < +∞ and
ϕ(y) < +∞, we obtain
0 < 2d(x, y) ≤ ϕ(x) − ϕ(y) + ϕ(y) − ϕ(x) = 0,
contradiction. So we must have x /∈ B ϕ
y , hence B ϕ
y ( B ϕ
x .
(cid:3)
From assertion (3) of Lemma 16, we obtain:
Corollary 17. Let x0 be a CK element for a CK∞ function ϕ. Then for
every y ∈ B0, y is also a CK element for ϕ. Further, ϕB0 is a CK function
and (B0, B ϕ
x0) is a CK ball space in the sense of Section 2.1.
Before we state another property of CK∞ balls, it is worth noting that
the proofs of Lemma 16 and Lemma 18 are similar (or, at times, identical)
to the original proof of Lemma 3, which can be found in [8]. This comes
from the fact that for a CK element x0, the set ϕ(B0) does not contain
infinity, so these balls keep the properties of the 'original' CK balls.
Lemma 18. For every x ∈ X and every CK∞ function ϕ, the ball B ϕ
closed in the topology induced by the metric.
x is
Proof. The complement {y ∈ X d(x, y) + ϕ(y) > ϕ(x)} of B ϕ
x is the
preimage of the final segment ( ϕ(x), +∞] of (−∞, +∞], which is open in
the Scott topology, under the function
X ∋ y
ψ
7−→ d(x, y) + ϕ(y).
Whenever ϕ is lower semicontinuous, then so is ψ and this preimage is open
in X.
(cid:3)
10
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
We are now ready to prove a result analogous to Propositions 5 and 15.
Proposition 19. Let (X, d) be a complete metric space and ϕ be a CK∞
function on X. If x0 ∈ X then there exists a ∈ B ϕ
x0 such that B ϕ
a = {a}.
Proof. Consider a complete metric space (X, d), fix any element x0 ∈ X and
consider the ball B0 := B ϕ
x0.
Assume first that x0 is a CK element for ϕ in X. As we know from
Lemma 18, B0 is closed, hence complete. Moreover, the function ϕ := ϕB0
is a CK function. Note that for every x ∈ B0 we have
x = {y ∈ B0 d(x, y) ≤ ϕ(x) − ϕ(y)}
Bϕ
⊆ B ϕ
x = {y ∈ X ϕ(y) + d(x, y) ≤ ϕ(x)}.
We wish to show that the above sets are equal. By assertion (2) of Lemma 16
we know that B ϕ
x ⊆ B0. On B0 we have ϕ = ϕ so that the values of ϕ are
finite and ϕ(y) + d(x, y) ≤ ϕ(x) is equivalent to d(x, y) ≤ ϕ(x) − ϕ(y). This
yields B ϕ
x ⊆ Bϕ
x .
Since ϕ is a CK function on a complete metric space B0, we may apply
x0), to acquire
assertion (2) of Proposition 15 to the CK ball space (B0, Bϕ
a ∈ B0 such that
{a} = Bϕ
a = B ϕ
a .
Assume now that x0 ∈ X is not a CK element for ϕ. Then we obtain
that B0 = X. Inside the ball B0 we may thus find a CK element x1 for ϕ.
From what we have proved above there exists a ∈ B ϕ
x1 ⊆ X = B0 such that
B ϕ
(cid:3)
a = {a}.
3. Applications of Proposition 15
In this section we give simple proofs for a number of known theorems,
in versions that involve OT functions, by applying Proposition 15. Note
that the multivalued Caristi -- Kirk FPT, Ekeland's Principle and Takahashi's
Theorem have already been proved in the OT form in [10] using the Oettli --
Théra Theorem. The original versions of these theorems are listed in Sec-
tion 4.
Theorem 20 (Caristi -- Kirk FPT, OT form). Let (X, d) be a complete met-
ric space and φ an OT function on X. If a function f : X → X satisfies:
(5)
∀
x∈X
d(x, f (x)) ≤ −φ(x, f (x)),
then f has a fixed point on X, i.e., there exists an element a ∈ X such that
f (a) = a.
Proof. Condition (5) implies that for every x ∈ X we have
Proposition 15 gives us the existence of a ∈ X such that Ba = {a}.
particular, since f (a) ∈ Ba, we have f (a) = a.
In
(cid:3)
f (x) ∈ Bx.
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
11
Theorem 21 (Caristi -- Kirk FPT, multivalued version, OT form). Let (X, d)
be a complete metric space and φ an OT function on X.
If a function F : X → P(X) satisfies:
(6)
∀
x∈X
∃
y∈F (x)
d(x, y) ≤ −φ(x, y),
then F has a fixed point on X, i.e., there exists a ∈ X such that a ∈ F (a).
Proof. Condition (6) means that for every x ∈ X there exists y ∈ F (x)∩Bx.
In particular, for x = a with a ∈ X given by Proposition 15 we obtain
y ∈ F (a) ∩ Ba ⊆ {a},
whence a = y ∈ F (a).
(cid:3)
Theorem 22 (Basic Ekeland's Principle, OT form). Let (X, d) be a com-
plete metric space and φ an OT function on X. There exists a ∈ X such
that
(7)
∀
x∈X\{a}
− φ(a, x) < d(a, x).
Proof. Property (7) is equivalent to Ba = {a} and the existence of such
a ∈ X follows from Proposition 15.
(cid:3)
Theorem 23 (Altered Ekeland's Principle, OT form). Let (X, d) be a com-
plete metric space and φ an OT function on X. For any γ > 0 and any OT
element x0 for φ in X there exists a ∈ X such that
(8)
and
(9)
∀
x∈X\{a}
− φ(a, x) < γd(a, x)
− φ(x0, a) ≥ γd(x0, a).
Proof. Since γ > 0, the function ψ := γ −1φ is an OT function on X, so we
can work with ψ and the respective ball space Bψ.
We apply Proposition 15 to the given complete metric space X, the func-
x0. This gives us the existence of an element
tion ψ and the ball B0 := Bψ
a ∈ B0 such that Bψ
a = {a}. Now, the assertion a ∈ B0 means that
d(x0, a) ≤ −ψ(x0, a) = −γ −1φ(x0, a),
which is equivalent to property (9). Similarly, Bψ
a = {a} implies
∀
x∈X\{a}
d(a, x) > −ψ(a, x) = −γ −1φ(a, x),
which is equivalent to property (8).
(cid:3)
12
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
Theorem 24 (Ekeland's Usual Variational Theorem, OT form). Let (X, d)
be a complete metric space and φ an OT function on X. Fix ε ≥ 0 and
x0 ∈ X such that −ε ≤ inf x∈X φ(x0, x). Then for any γ > 0 and δ ≥ 0
with γδ ≥ ε there exists a ∈ X such that d(a, x0) ≤ δ and a is the strict
minimum point of the function φγ : X → (−∞, +∞] defined as
φγ(x) = φ(a, x) + γd(x, a).
Proof. Take ε ≥ 0 and x0 as in the assumptions of the theorem, and fix
arbitrary real numbers γ > 0 and δ ≥ 0 such that γδ ≥ ε. The function
ψ := γ −1φ : X × X → (−∞, +∞] is an OT function on X, so we can
apply Proposition 15 with the function ψ and B0 := Bψ
x0 (note that x0 is
an OT element for ψ in X). We deduce the existence of a ∈ B0 such that
Bψ
a = {a}. Now, the property a ∈ B0 means that
or in other words:
d(x0, a) ≤ −ψ(x0, a),
γd(x0, a) ≤ −φ(x0, a) ≤ − inf
x∈X
φ(x0, x) ≤ ε ≤ γδ.
Thus,
d(a, x0) = d(x0, a) ≤ δ.
The property Ba = {a} means that for every x ∈ X \ {a} we have that
d(x, a) > −ψ(a, x) = −γ −1φ(a, x).
From this we obtain that
φγ(x) = φ(a, x) + γd(x, a) > 0 = φγ(a),
which means that a is the strict minimum point of the function φγ.
(cid:3)
Definition 25. Let (X, d) be a metric space. Take γ ∈ (0, ∞) and a, b ∈ X.
The petal associated with γ and a, b is the subset Pγ(a, b) of X defined as
follows:
Pγ(a, b) = {y ∈ X γd(y, a) + d(y, b) ≤ d(a, b)}.
Theorem 26 (Flower Petal Theorem). Let M be a complete subset of a
metric space (X, d). Take x0 ∈ M and b ∈ X \ M. Then for each γ > 0
there exists a ∈ Pγ(x0, b) ∩ M such that
Pγ(a, b) ∩ M = {a}.
Proof. We use the notation from the assertion of the theorem. As γ > 0,
the function ϕ : M → R given by
ϕ(x) := γ −1d(x, b)
is a CK function on M. In this setting we have, for every x ∈ M,
Pγ(x, b) ∩ M = {y ∈ M d(x, y) ≤ ϕ(x) − ϕ(y)} = Bϕ
x .
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
13
To conclude we use assertion (2) of Proposition 15 with M in place of X
and x := x0, which yields the existence of a ∈ Bϕ
x0 = Pγ(x0, b) ∩ M such
that
{a} = Bϕ
a = Pγ(a, b) ∩ M.
(cid:3)
Theorem 27 (Takahashi, OT form). Let (X, d) be a complete metric space,
φ an OT function on X and x0 ∈ X an OT element for φ in X. Assume
that for every u ∈ Bx0 with inf x∈X φ(u, x) < 0 there exists v ∈ X such
that v 6= u and d(u, v) ≤ −φ(u, v). Then there exists a ∈ Bx0 such that
inf x∈X φ(a, x) = 0.
Proof. Proposition 15 gives us the existence of a ∈ Bx0 such that Ba = {a}.
If inf x∈X φ(a, x) < 0, then by assumption there would exist v ∈ X \ {a}
such that d(a, v) ≤ −φ(a, v), which would mean that Ba is not a singleton,
contradiction. So inf x∈X φ(a, x) ≥ 0, but φ(a, a) = 0 which proves the
claim.
(cid:3)
Theorem 28 (Oettli-Théra). Let (X, d) be a complete metric space, φ an
OT function on X and x0 ∈ X an OT element for φ in X. Let Ψ ⊆ X
have the property that
(10)
∀
x∈Bx0 \Ψ
∃
y∈X\{x}
d(x, y) ≤ −φ(x, y).
Then there exists a ∈ Bx0 ∩ Ψ.
Proof. From Proposition 15 there exists a ∈ Bx0 such that Ba = {a}. If
a /∈ Ψ then, by assumption, Ba would contain another element y 6= a, which
would mean that Ba is not a singleton, contradiction.
(cid:3)
4. Applications of Proposition 19
Many of the theorems mentioned in the previous section have been orig-
inally stated and proved using the CK function ϕ. By Fact 8, proving the
version involving φ, through (1), will also automatically prove the version
involving ϕ. However, many sources (e.g., [11], [1]) cite the theorems in a
CK∞ form. As already remarked, we cannot directly define an OT function
from a CK∞ function. Nevertheless, we can use Proposition 19 to prove
these versions in the same way we did in the previous section using Propo-
sition 15. Since the proofs are analogous to the ones stated in Section 3, we
will leave them to the reader.
Note that here we do not include the Oettli-Théra Theorem (since it has
originally been stated in the OT form) nor the Flower Petal Theorem (since
it does not include either of the functions).
For the following theorems, fix a complete metric space (X, d) and a CK∞
function ϕ on X.
14
BŁASZKIEWICZ, ĆMIEL, LINZI, SZEWCZYK
Theorem 29 (Caristi-Kirk FPT, CK∞ form). If a function f : X → X
satisfies
∀
x∈X
ϕ(f (x)) + d(x, f (x)) ≤ ϕ(x)
then f has a fixed point on X, i.e., there exists an element a ∈ X such that
f (a) = a.
Theorem 30 (Caristi-Kirk FPT, multivalued version, CK∞ form). If a
function F : X → P(X) satisfies
(11)
∀
x∈X
∃
y∈F (x)
ϕ(y) + d(x, y) ≤ ϕ(x)
then F has a fixed point on X, i.e., there exists a ∈ X such that a ∈ F (a).
Theorem 31 (Basic Ekeland's Principle, CK∞ form). There exists a ∈ X
such that
∀
x∈X\{a}
ϕ(a) < ϕ(x) + d(a, x).
Theorem 32 (Altered Ekeland's Principle, CK∞ form). For all γ > 0 and
any x0 ∈ X there exists a ∈ X such that
and
∀
x∈X\{a}
ϕ(a) < ϕ(x) + γd(a, x)
ϕ(a) ≤ ϕ(x0) − γd(a, x0).
Theorem 33 (Ekeland's Usual Variational Theorem, CK∞ form). Let ε ≥ 0
and x0 ∈ X be such that ϕ(x0) ≤ inf ϕ(X) + ε. Then for any γ > 0 and
δ ≥ 0 with γδ ≥ ε there exists a ∈ X such that d(a, x0) ≤ δ and a is the
strict minimum point of the function
ϕγ(x) = ϕ(x) + γd(x, a).
Theorem 34 (Takahashi, CK∞ form). Suppose that for each u ∈ X with
inf x∈X ϕ(x) < ϕ(u) there exists v ∈ X such that v 6= u and ϕ(v) + d(u, v) ≤
ϕ(u). Then there exists a ∈ X such that inf x∈X ϕ(x) = ϕ(a).
CARISTI -- KIRK AND OETTLI -- THÉRA BALL SPACES
15
References
[1] Y. Araya, On generalizing Takahashi's nonconvex minimization theorem, Appl.
Math. Lett., 22 (2009), pp. 501 -- 504.
[2] J. Caristi, Fixed point theorems for mappings satisfying inwardness conditions,
Trans. Amer. Math. Soc., 215 (1976), pp. 241 -- 251.
[3] I. Ekeland, On the variational principle, J. Math. Anal. Appl., 47 (1974), pp. 324 --
353.
[4] W. A. Kirk, Caristi's fixed point theorem and metric convexity, Colloq. Math., 36
(1976), pp. 81 -- 86.
[5] F.-V. Kuhlmann and K. Kuhlmann, Ball spaces -- a basic framework for fixed
point theorems: ball spaces and spherical completeness, in preparation.
[6]
[7]
, A common generalization of metric and ultrametric fixed point theorems, Fo-
rum Math., 27 (2015), pp. 303 -- 327.
, Correction to "a common generalization of metric, ultrametric and topological
fixed point theorems", Forum Math., 27 (2015), pp. 329 -- 330.
[8] F.-V. Kuhlmann, K. Kuhlmann, and M. Paulsen, The Caristi-Kirk fixed point
theorem from the point of view of ball spaces, J. Fixed Point Theory Appl., 20 (2018),
pp. Art. 107, 9.
[9] I. Meghea, Ekeland variational principle, Old City Publishing, Philadelphia, PA;
Éditions des Archives Contemporaires, Paris, 2009.
[10] W. Oettli and M. Théra, Equivalents of Ekeland's principle, Bull. Austral.
Math. Soc., 48 (1993), pp. 385 -- 392.
[11] J.-P. Penot, The drop theorem, the petal theorem and Ekeland's variational prin-
ciple, Nonlinear Anal., 10 (1986), pp. 813 -- 822.
[12] W. Takahashi, Existence theorems generalizing fixed point theorems for multival-
ued mappings, in Fixed point theory and applications (Marseille, 1989), vol. 252 of
Pitman Res. Notes Math. Ser., Longman Sci. Tech., Harlow, 1991, pp. 397 -- 406.
Institute of Mathematics, University of Szczecin
ul. Wielkopolska 15, 70-451 Szczecin, Poland
E-mail address: [email protected], [email protected],
[email protected], [email protected]
|
1704.01825 | 1 | 1704 | 2017-04-06T13:16:46 | On the mean summability of series by nonlinear basis | [
"math.FA"
] | The nonlinear signal processing has achieved a rapid process in the recent years. A family of nonlinear Fourier bases, as a typical family of mono-component signals, has been constructed and applied to signal processing. In this paper, the approximation properties of the partial sums and Ces?aro summability of series by the nonlinear Fourier basis are investigated in the Lp(T). Furthermore, these results are applied to the prove of Bernstein's inequalities for nonlinear trigonometric polynomials. | math.FA | math |
ON THE MEAN SUMMABILITY OF
SERIES BY NONLINEAR FOURIER BASIS
HATICE ASLAN AND ALI GUVEN
family
typical
of mono-component
Abstract. The nonlinear signal processing has achieved a rapid
process in the recent years. A family of nonlinear Fourier bases, as
a
has been
constructed and applied to signal processing.
In this paper,
the approximation properties of the partial sums and Ces`aro
summability of
series by the nonlinear Fourier basis are
investigated in the Lp(T). Furthermore, these results are applied
to the prove of Bernstein's inequalities for nonlinear trigonometric
polynomials.
signals,
1. Introduction
Let 0 < p < ∞ and T := R/2πZ. Let f is a periodic function
on T, then we denote Lp(T) the set of Lebesgue measurable functions
f : T → R (or C)) such that
(cid:18) 1
2π ZT
f (x) p dx(cid:19)1/p
< ∞
where the integral is a Lebesgue integral, and we identify functions that
differ on a a set of measure zero. We define Lp-norm of f by
k f kp=(cid:18) 1
2πZT
f (x) p dx(cid:19)1/p
.
For p = ∞ the space L∞(T) consists of the Lebesgue measurable
functions f : T → R (or C) that are essentially bounded on T,
meaning that f is bounded on a subset of T whose complement has
measure zero. The norm on is essential supremum
k f k∞= inf{M
f (x) ≤ M a.e.
in T}.
Note that k f kp may take the value ∞.
2010 Mathematics Subject Classification. 41A25, 41A10, 41E30.
Key words and phrases. Nonlinear Fourier basis, Partial sum, Ces`aro mean,
Bernstein inequality .
1
2
HATICE ASLAN AND ALI GUVEN
The another important concept is the modulus of smoothness is
defined by
ω (f ; t)p := sup
0≤h≤t
k f (. + h) − f (.) kp
If p = ∞, then it is called modulus of contiunity. And this
nondecreasing continuous function on the interval
[0, 2π] having
properties:
ω(0) = 0, ω(t1 + t2) ≤ ω(t1) + ω(t2).
A family of nonlinear Fourier bases as the extension of the classical
Fourier basis, have been constructed and applied to signal processing
[1, 2, 7, 9, 8]. For any complex number a = a eita, a < 1, the
nonlinear phase function θa(t) is defined by the radical boundary value
of the Mobuis transformation
τa =
z − a
1 − az
that is,
eiθa(t) := τa(eit) =
eit − a
1 − aeit
It is easily seen that
(1.1)
θa(t + 2π) = θa(t) + 2π
and its derivative is the Poisson kernel
θ′
a(t) = pa(t) =
1 − a2
1 − 2 a cos(t − ta) + a2
which satisfies
(1.2)
0 <
1 − a
1 + a
≤ θ′
a(t) ≤
1 + a
1 − a
Hence, θa(t) is a strictly monotonic increasing function, which makes
cosθa(t) be a special mono-component signal [7, 8]. It has been shown
that for any sequence {ck}k∈Z of finite nonzero terms, there holds
ck 2=
ckeikθa(x) 2 dx =
1
2πZTXk∈Z
1
2πZTXk∈Z
ckeikθa(t) 2 pa(t)dt
which combining with (1.2) implies that the so-called nonlinear Fourier
Xk∈Z
basis (cid:8)einθa(t)(cid:9)n∈Z forms a Riesz basis for L2(T) with the upper bound
q 1+a
1−a and the lower bound q 1−a
1+a . When a = 0, (cid:8)einθa(t)(cid:9)n∈Z is
simply the Fourier basis {eint}n∈Z.
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS3
Let τ a
n be the space of all the nonlinear trigonometric polynomials of
degree less than or equal to n, that is,
The approximation error of f ∈ Lp(T),
τ a
n := span(cid:8)eikθa(t) : k ≤ n(cid:9) .
Ea
n(f )p = inf
T ∈τ a
n
k f − T kLp(T) .
Let us recall some known lemmas (see [5]) which will be used in the
sequel of paper.
Lemma 1. Let f ∈ C(T). We have
Ea
n(f )∞ ≤
24
1− a
ω(cid:18)f,
1
n(cid:19)∞
.
Lemma 2. Let f ∈ C(T). We have
1− a
2
ω(f, t)∞ ≤ ω(f ◦ θ−1
a , t)∞ ≤
2
1− a
ω(f, t)∞.
In the present paper first we deal with some properties of
nonlinear Fourier series. Then we discuss n-th partial sums and Ces`aro
sum of nonlinear Fourier series. Also we prove the necessary and
sufficient condition for nonlinear Fourier series which governs the (C, 1)
summability in Lp(T) for arbitrary function f from Lp(T). This
result is applied to the prove of Bernstein's inequality for nonlinear
trigonometric polynomials.
2. Convergence of Nonlinear Fourier Series
Let f ∈ L1(T). Then 1 ≤ p ≤ ∞ by nonlinear basis denoted by
∞
be its series by nonlinear Fourier basis where
f (x) ∼
ckeikθa(x)
Xk=−∞
ck = ck(f ) =
f (t)e−ikθa(t)pa(t)dt,
k ∈ Z.
1
2πZ π
−π
For simplicity throughtout the present paper we write nonlinear Fourier
series as series by nonlinear Fourier basis. Now we can begin to give
some properties of nonlinear Fourier series:
Assume f, g ∈ L1(T) and
f (x) ∼
∞
Xk=−∞
ckeikθa(x),
g(x) ∼
dkeikθa(x).
∞
Xk=−∞
4
HATICE ASLAN AND ALI GUVEN
For linearity and convolution we have the following properties
respectively.
∞
(Af + Bg)(x) ∼
(f ∗ g)(x) ∼
(Ack + Bdk)eikθa(x)
Xk=−∞
1− a (cid:19) (ckdk)eikθa(x)
Xk=−∞(cid:18) 1+ a
∞
Furthermore remember Sobolev space i.e.
W p
r (T) =(cid:8)f ∈ X p(T) : f, ..., f (r−1) ∈ AC(T), f (r) ∈ X p(T)(cid:9) .
1 (T). It can be easily seen that property i.e.
Assume that f ∈ W p
f ′(x) ∼
holds.
∞
Xk=−∞(cid:18) 1+ a
1− a (cid:19) (ikck)eikθa(x)
We wish to examine the convergence of nonlinear Fourier series. To
discuss the convergence, pointwise or uniform, of nonlinear Fourier
series, we need to discuss the convergence of the sequence {Sa
n} of
partial sums. We have
n
where
ck = ck(f ) =
f (t)e−ikθa(t)pa(t)dt.
Sa
n(f )(x) =
ckeikθa(x)
Xk=−n
2πZ π
−π
1
Proposition 1. Let Sa
n(f ) be the sequence of partial sums of the
nonlinear Fourier series of f . Let f ∈ L1(T) which is 2π-periodic.
Then
(2.1)
Sa
n(f )(x) =
(2.2)
=
−π
1
2πZ π
2π Z π
−π
1
f (t)Dn(θa(x) − θa(t))pa(t)dt.
F (θa(x) + t)Dn(t)dt,
(2.3)
=
{F (θa(x) − t) + F (θa(x) + t)} Dn(t)dt.
1
2π Z π
0
Proof. By the expression for the ck and considering the equality
Sa
n(f )(x) = Sn(F )(θa(x)) where F = f ◦θ−1
a we have already established
(2.1). By a change of variable u − θa(x) = y and the (1.1) equality of
θa(x) phase function's 2π-periodicity, it follows that the integral does
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS5
not change as long as the length of the interval of integration is 2π we
get
Sa
n(f )(x) =
F (θa(x) + y)Dn(y)dy.
1
2πZ π
−π
This proves (2.2). Finally, we split the integral in (2.2) as the sum of
integrals over [−π, 0] and [0, π]. Now
Z 0
−π
F (θa(x) + t)Dn(t)dt =Z π
0
F (θa(x) + t)Dn(t)dt,
using the change of variable y = −t and the evenness of Dn. This
proves (2.3).
(cid:3)
We can begin with properties of the operators Sa
n of partial sums of
nonlinear Fourier series. For this first we need following lemmas.
Lemma 3. Let F := f ◦ θ−1
a . Then we have
k F kp≤( k f kp,
1−a(cid:17)1/p
(cid:16) 1+a
p = 1, ∞, f ∈ C(T)
if
p = ∞;
k f kp 0 < p < ∞;
Proof. Let f ∈ C(T) ve F := f ◦ θ−1
a . Therefore for p = 1, ∞ we have
k f k∞:= supx∈T f (x) = supθa(x)∈T F (θa(x)) = kF k∞.
And for 0 < p < ∞ the following inequality holds.
k f kp:=(cid:18) 1
2π ZT
f (t) p dt(cid:19)1/p
.
By change of variable θa(t) = u and using phase function's property
which is giving in (1.2), we have the following equality.
k f kp≥(cid:18)1 − a
1 + a(cid:19)1/p(cid:18) 1
2πZT
F (u) p du(cid:19)1/p
=(cid:18)1 − a
1 + a(cid:19)1/p
Therefore we have result that we wanted.
kF kp.
(cid:3)
For the spaces L1(T), C(T), one can evaluate the norms k Sa
n k by
direct computation.
Theorem 1. One has
k Sa
n(f ) k∞≤ Λn k f k∞ .
6
HATICE ASLAN AND ALI GUVEN
Proof. Let consider Sa
n be an operator C(T) to C(T). By using Lemma
2.1 and equality (2.2), we see that each of the norms (2.2) is equal to
Sa
n(f )(x) =(cid:12)(cid:12)(cid:12)(cid:12)
2πZ π
≤
−π
1
≤
1
2πZ π
−π
F (θa(x) + t) Dn(t) dt
1
2πZ π
−π
F (θa(x) + t)Dn(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
2πZ π
−π
1
k F k∞ Dn(t) dt =
k f k∞ Dn(t) dt
From the Lebesgue constant's definition (see in [4])
k Sa
n(f ) k∞≤ Λn k f k∞
holds.
Theorem 2. For all f ∈ C(T),
(cid:3)
k f − Sa
n(f ) k∞≤ c log nEa
n(f )
holds.
Proof. Let f ∈ C(T). Consider linear operator Sa
n = 0, 1, .... The best approximation by t ∈ τ a
n is
n : C(T) → C(T),
Thus we can write the following equality.
Ea
n(f ) =k f − t k∞ .
k f − Sa
n(f ) k∞=k f − t + t − Sa
n(f ) k∞=k f − t + Sa
n(f )(t − f ) k∞
≤k f − t k∞ + k Sa
n(f )(t − f ) k∞= Ea
n(f )+ k Sa
n(f ) k∞k f − t k∞
holds. Therefore from Theorem 2.1
k f − Sa
n(f ) k∞≤ (1 + Λn)Ea
n(f )
holds. Hence by using theorem for Lebesgue constant (see in [4]) we
have
k Sa
n(f ) − f k∞≤ c log nEa
n(f ).
This completes the proof.
(cid:3)
Theorem 3. Let consider Sa
sequence, where n = 0, 1, ... and 1 ≤ p < ∞. Then
n : Lp(T) → Lp(T) linear operator
k Sa
n(f ) kp≤(cid:18) 1 + a
1 − a(cid:19)2/p
Λn k f kp
holds.
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS7
Proof. Let 1 ≤ p < ∞. Consider operator Sa
n = 0, 1, .... Thus if we consider the inequality (2.2)
n : Lp(T) → Lp(T),
holds. Hence by using Minkowski integration inequality (see in [10]),
we have
k Sa
=(cid:18) 1
k Sa
n(f ) kp≤
1
p
1
−π
−π
Sa
n(f )(x) p dx(cid:19)1/p
2πZ π
n(f ) kp=(cid:18) 1
dx(cid:19)1/p
2πZ π
2πZ π
−π(cid:12)(cid:12)(cid:12)(cid:12)
F (θa(x) − t)Dn(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
2πZ π
2πZ π
−π(cid:18) 1
× Dn(t) (cid:18) 1
2πZ π
2πZ π
1 − a(cid:19)1/p 1
≤(cid:18) 1 + a
F (θa(x) + t) p pa(x)dx(cid:19)
pa(x)(cid:19)1/p
F (θa(x) + t) p pa(x)dx)
k F kp Dn(t) dt.
2π Z π
−π
dt
1
(
−π
−π
−π
1
p
≤(cid:18) 1 + a
1 − a(cid:19)1/p 1
1
p Dn(t) dt
Therefore considering Lemma 2.2 and Dirichlet kernel's definition (see
in [3]) is giving result that we wanted.
(cid:3)
Now let examine the convergence of partial sum of the nonlinear
Fourier series in Lp(T) space.
Theorem 4. Let 1 ≤ p < ∞. Then for all f ∈ Lp(T)
(2.4)
k f − Sa
n(f ) kp→ 0, n → ∞
holds if and only if, there exist a M > 0 constant that only depend on
p such that
(2.5)
k Sa
n(f ) kp≤ M a
p k f kp .
Proof. For necessity let f ∈ Lp(T) and k f − Sa
n(f ) kp→ 0, n → ∞.
Thus {Sa
n(f )} is bounded since it converges in Lp(T) norm space for
f ∈ Lp(T). Therefore there exists Mf > 0 for all f ∈ Lp(T). Such that
k Sa
n(f ) kp< Mf . Thus considering uniform bounded principle (e.g.
[6]) we have
sup {k Sa
n k: n = 0, 1, 2, ...} < ∞.
If we consider Theorem 2.3 and choose
M = sup {k Sa
n k: n = 0, 1, 2, ...} < ∞
8
HATICE ASLAN AND ALI GUVEN
for f ∈ Lp(T), then we can write for (2.5) inequality.
For sufficiency let k Sa
n(f ) kp≤ M a
Ea
n(f )p =k f − ta
p k f kp. For
n kp
where f ∈ Lp(T) and n = 0, 1, 2, ... . Therefore we can prove (2.4) with
the help of the nonlinear polynomials ta
n of best approximation of the
function:
k f − Sa
n(f ) kp=k f − ta
n(f − ta
n kp + k Sa
n + ta
n + Sa
n) kp≤ (M a
n(f )(x) kp
p + 1)Ea
n(f )p.
≤k f − ta
Since
Ea
n(f )p → 0, n → ∞,
this completes the proof.
(cid:3)
Corollary 1. The norms of the operators Sa
space f ∈ Lp(T), 1 < p < ∞.
n(f ) are bounded in each
From Theorem 2.2, Theorem 2.3 and Theorem 2.4 we now derive
k Sa
n(f )−f kp≤( C log nEa
1−a(cid:17)2/p
Cp(cid:16) 1+a
n(f )p,
p = 1, ∞, f ∈ C(T), if p = ∞;
Ea
n(f )p, 1 < p < ∞.
We see that the partial sums Sa
n approximate almost as well as its
polynomial of best approximation. This is true even for f ∈ C(T), if
the factor log n is not essential for the problem considered.
For the partial sums Sa
n(f ) → f in C(T)
n of the Fourier series of f we do not have
Sa
But we do have fast
convergence of Sa
n for smooth functions f . Actually, the convergence
k Sa
n(f ) − f k∞→ 0 can be arbitrarily fast for some f (that are not
trigonometric polynomials): The following theorems shows that it is
k=0[akcos(kθa(x)) + bksin(kθa(x))], where
for each f .
ak, bk converge to zero sufficiently fast without being zero.
sufficient to take f (t) = P∞
Theorem 5. For f ∈ L1(T)
ca
n(f ) = lim
n→−∞
lim
n→∞
ca
n(f ) = 0
holds.
Proof. Let ε > 0. In this case there exists a t ∈ τ a
n such that
In this inequality let t has degree N. Therefore we can write
k f − t k< ε2π.
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS9
m=N
t(x) =
dmeimθa(x)
Xm=−N
2πZ π
−π
1
Thus for k > N,
ca
k(f ) =
f (x)e−ikθa(x)dx
1
2π Z π
−π
t(x)e−ikθa(x)dx
=
1
2πZ π
−π
f (x)e−ikθa(x)dx −
=
1
2πZ π
−π
[f (x) − t(x)]e−ikθa(x)dx
holds. Hence we obtain the result from following inequality.
ca
k(f ) ≤k f (x) − t(x) k1< ε.
Corollary 2. Let f ∈ L1(T). Then for all f ,
lim
n→∞
aa
n(f ) = lim
n→−∞
ba
n(f ) = 0
holds.
Proof. By writing aa
Theorem 2.5, we can complete the proof.
n(f ) = ca
n(f ) + ca
−n(f ) ve ba
n(f ) = ca
n(f )+ca
−n(f )
i
Theorem 6. Let f ∈ L1(T). If
(cid:3)
in
(cid:3)
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
t
dt < ∞
then for n → ∞, Sa
n(f )(x) → f (x) holds.
Proof. Let assume f ∈ L1(T). By using (2.3) equality, we can write
Z π
0
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
1
2πZ π
0
Sa
n(f )(x)−f (x) =
{F (θa(x) + t) + F (θa(x) − t) − 2f (x)} Dn(t)dt.
Therefore considering Dirichlet kernel's property (see in [10]), we obtain
the following.
Sa
n(f )(x) − f (x) =
1
0 (F (θa(x) + t) + F (θa(x) − t) − 2f (x)
2πZ π
× sin(cid:18)(cid:18)2n + 1
2 (cid:19) t(cid:19) dt
sin(cid:0) t
2(cid:1)
)
10
HATICE ASLAN AND ALI GUVEN
Now let think the φa
x(t) as following.
φa
x(t) = F (θa(x) + t) + F (θa(x) − t) − 2f (x).
In this case the last equality written as
Sa
n(f )(x) − f (x) =
From hypothesis φa
x(t) ∈ L1(T). Hence
sin(cid:18)2n + 1
2 (cid:19) t) dt.
holds and by using Jordan inequality (see [10]), we obtain
1
2(cid:19) t(cid:19) dt
1
0
0
φa
x(t)
x(t)
dt < ∞
dt < ∞.
Z π
φa
x(t)
t
0 ( φa
2πZ π
sin(cid:0) t
2(cid:1)
Z π
(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)
sin(cid:0) t
2(cid:1)
0 ( φa
2πZ π
sin(cid:0) t
2(cid:1)
) sin(nt) cos(cid:18) t
0 ( φa
sin(cid:0) t
2(cid:1)
0 ( φa
) cos(nt) sin(cid:18) t
sin(cid:0) t
2(cid:1)
x(t) =( φa
sin(cid:0) t
2(cid:1)
) cos(cid:18) t
2(cid:19) .
x(t)
x(t)
x(t)
x(t)
1
) sin(cid:18)(cid:18)n +
2(cid:19) dt
2(cid:19) dt
=
+
1
2πZ π
2πZ π
1
ga
Therefore
Sa
n(f )(x) − f (x) =
holds. Here if we say that
We find that ga
equalities
x(t) ∈ L1(T).
In this case, considering the following
bn (ga
x(t)) =
an (φa
x(t)) =
we see that
−π
1
2πZ π
2πZ π
1
−π
ga
x(t) sin(nt)dt
φa
x(t) cos(nt)dt
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS11
Sa
n(f ) − f (x) =
(bn (ga
x(t))) + an (φa
x(t))
1
4
holds. Thus by considering Corollary 2.2 and Riemann-Lebesgue Lemma
(see in [10]) for n → ∞, we have
So we obtain Sa
x(t)) → 0 and an (φa
bn (ga
n(f )(x) → f (x) as n → ∞. This completes the proof.
(cid:3)
x(t)) → 0.
Corollary 3. Let f ∈ L1(T). If F (x) − F (y) ≤ M x − y α, then
we have
Sa
n(f )(x) → f (x), n → ∞.
Proof. Let assume that f ∈ L1(T). Therefore
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
t
F (θa(x) + t) − f (x) + F (θa(x) − t) − f (x)
≤
≤
2Mtα
t
= 2Mtα−1
t
0 tα−1dt = πα
(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)
holds. Thus if we consider R π
following by using comparison test.
α equality, we have the
1
2πZ π
0
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
t
dt < ∞
Therefore from Theorem 2.6, we obtain Sa
n(f )(x) → f (x), n → ∞.
(cid:3)
Corollary 4. Let f ∈ L1(T). If f is differentable and F := f ◦ θ−1
all x ∈ T, then we have
a
for
Sa
n(f )(x) → f (x), n → ∞.
Proof. Let f ∈ L1(T). Hence
F ′(θa(x)) = lim
t→0
F (θa(x) + t) − F (θa(x))
t
holds. For ∃ δ > 0 ∋ 0 < t < δ
holds. Thus we have
F (θa(x) + t) − F (θa(x))
t
(cid:12)(cid:12)(cid:12)(cid:12)
< 1
− F ′(θa(x))(cid:12)(cid:12)(cid:12)(cid:12)
12
HATICE ASLAN AND ALI GUVEN
F (θa(x) + t) − F (θa(x))
t
< F ′(θa(x)) .
Therefore
F (θa(x) + t) − F (θa(x))
t
<
F (θa(x) + t) − F (θa(x))
δ
holds for t > δ. Hence we have the following inequality.
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
F (θa(x) + t) − f (x)
F (θa(x) − t) − f (x)
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
t
t
t
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
(cid:12)(cid:12)(cid:12)(cid:12)
dt +Z0<t<δ(cid:12)(cid:12)(cid:12)(cid:12)
t F ′(θa(x))dt +Z0<t<δ
t
F (θa(x) + t) + F (θa(x) − t) − 2f (x)
t F ′(θa(x))dt
t
t
δ
≤Z0<t<δ(cid:12)(cid:12)(cid:12)(cid:12)
0
Z π
(cid:12)(cid:12)(cid:12)(cid:12)
=Z0<t<δ(cid:12)(cid:12)(cid:12)(cid:12)
+Zδ≤t≤π(cid:12)(cid:12)(cid:12)(cid:12)
+Zδ≤t≤π(cid:12)(cid:12)(cid:12)(cid:12)
≤Z0<t<δ
+Zδ≤t≤π(cid:12)(cid:12)(cid:12)(cid:12)
dt
(cid:12)(cid:12)(cid:12)(cid:12)
dt
dt
dt
dt
(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)
dt
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:3)
So by considering Theorem 2.6, we obtain
Sa
n(f )(x) → f (x), n → ∞.
3. Boundness and Ces`aro Mean Summability for
Nonlinear Fourier Series
Precisely, we prove the necessary and sufficient condition for non-
linear Fourier series which governs the (C, 1) summability in Lp for
arbitrary function f from Lp. This result is applied to the prove
of Bernsteins inequality for nonlinear trigonometric polynomials. Let
f ∈ L1(T) and Sa
n(f ) is partial sum of nonlinear Fourier series. Define
n-th Fej´er (Ces`aro) for nonlinear Fourier series defined by
σa
n(f )(x) :=
1
n + 1
n
Xk=0
Sa
k (f )(x).
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS13
Now let find new expressions for Ces`aro mean which are very useful.
Proposition 2. Let σa
n(f ) be the sequence of partial sums of
the nonlinear Fourier series of f . Let f ∈ L1(T) which is 2π-periodic.
Then
(3.1)
σa
n(f )(x) =
(3.2)
=
−π
1
2π Z π
2π Z π
−π
1
f (t)Kn(θa(x) − θa(t))pa(t)dt,
F (θa(x) + t)Kn(t)dt,
(3.3)
=
{F (θa(x) − t) + F (θa(x) + t)} Kn(t)dt.
1
2π Z π
0
:=
n(f )(x)
Proof. By the expression for the ck, phase function θa(x) and
considering σa
k(f )(x), we have already
established (3.1). By a change of variable y = θa(t) − θa(x) and the
(1.1) equality of θa(x)'s 2π-periodicity, it follows that the integral does
not change as long as the length of the interval of integration is 2π we
get
n+1Pn
k=1 σa
1
σa
n(f )(x) =
F (θa(x) + y)Dn(y)dy.
1
2πZ π
−π
This proves (3.2). Finally, we split the integral in (3.2) as the sum of
integrals over [−π, 0] and [0, π]. Now
Z 0
−π
F (θa(x) + t)Kn(t)dt =Z π
0
F (θa(x) + t)Dn(t)dt,
using the change of variable y = −t and the evenness of Kn. This
proves (3.3).
(cid:3)
Theorem 7. Let σa
sequence. Then we have
n : C(T) → C(T) for n = 0, 1, ... be linear operator
for all f ∈ C(T).
k σa
n(f ) k∞≤k f k∞
Proof. Let consider σa
n : C(T) → C(T) as a linear operator for
n = 0, 1, .... Thus by using (3.2) inequality and Fej´er kernel's
positivity (see in [3]), we have
σa
n(f )(x) =(cid:12)(cid:12)(cid:12)(cid:12)
2π Z π
≤
−π
1
1
2πZ π
−π
F (θa(x) − t)Kn(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
F (θa(x) − t) Kn(t) dt,
14
HATICE ASLAN AND ALI GUVEN
≤
1
2πZ π
−π
k F k∞ Kn(t)dt.
If we consider Lemma 2.1 for last equality, we obtain that
σa
n(f )(x) =
k f k∞ Kn(t)dt.
1
2πZ π
−π
Finally by using Fej´er kernel's property (see in [10]) we obtain the
result.
k σa
n(f ) k∞≤k f k∞
(cid:3)
Theorem 8. Let f ∈ C(T). Then
k f − σa
n(f ) k∞→ 0, n → ∞.
Proof. Let σa
n : C(T) → C(T), n = 0, 1, .... Then we have
f (x) − σa
1
n(f )(x) =(cid:12)(cid:12)(cid:12)(cid:12)
2π Zt<δ
2πZδ≤t≤π
1
≤
+
1
2πZ π
−π
[f (x) − F (θa(x) + t)]Kn(t)dt(cid:12)(cid:12)(cid:12)(cid:12)
f (x) − F (θa(x) + t) Kn(t)dt
f (x) − F (θa(x) + t) Kn(t)dt
by (3.2) equality. Therefore since f ∈ C(T) and θa continuous, we can
write
f (x) − σa
n(f )(x) ≤
1
2π Zt<δ
ε
2
Kn(t)dt
f (x) − F (θa(x) + t) Kn(t)dt
≤
ε
4πZt<δ
+
1
2πZδ≤t≤π
Kn(t)dt +
1
2πZδ≤t≤π
2πZδ≤t≤π
1
≤
ε
4π
+
f (x) + F (θa(x) + t) Kn(t)dt
[k f k∞ + k F k∞]Kn(t)dt.
Now if we use Lemma 2.1 and Fej´er kernel's property (see in [10]), we
have
f (x) − σa
n(f )(x) ≤
ε
4π
+
1
2π
2 k f k∞ Zδ≤t≤π
Kn(t)dt
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS15
≤
ε
4π
+
1
π
≤
+
ε
2
k f k∞ Zδ≤t≤π
k f k∞
1
1
n + 1
sin2 δ
2
dt
2 Zδ≤t≤π
dt
π(n + 1)sin2 δ
ε
2
Since there exists ∃N ∈ N :
≤
+
2 k f k∞
.
π(n + 1)sin2 δ
2
n ≥ N
2 k f k∞
→ 0
as n → ∞, we obtain
π(n + 1)sin2 δ
2
2 k f k∞
π(n + 1)sin2 δ
2
<
ε
2
.
Therefore
f (x) − σa
n(f )(x) <
ε
2
ε
2
+
= ε
holds for ∃N ∈ N : n ≥ N ve ∀x ∈ T. Thus σa
n → ∞
n(f ) → f uniformly as
(cid:3)
Theorem 9. Let consider σa
n = 0, 1, ... linear operator sequence. Then
n : Lp(T) → Lp(T), 1 ≤ p < ∞ and
k σa
n(f ) kp≤(cid:18) 1 + a
1 − a(cid:19)2/p
holds for all f ∈ Lp(T).
k f kp
Proof. Let consider σa
linear operator sequence. Then we have
n : Lp(T) → Lp(T),1 ≤ p < ∞ and n = 0, 1, ...
Now by using Minkowski integral inequality (see in [10]),
σa
n(f )(x) p dx(cid:19)1/p
.
.
F (θa(x) − t)Kn(t)dt p dx(cid:19)1/p
F (θa(x) − t) p dx(cid:19)1/p
Kn(t)dt
k F kp Kn(t)dt
Hence by (3.2) equality, we have
k σa
n(f ) kp=
k σa
n(f ) kp=
k σa
−π
1
−π(cid:18) 1
2πZ π
n(f ) kp=(cid:18) 1
2πZ π
2πZ π
2πZ π
2πZ π
−π(cid:18) 1
2πZ π
=
−π
−π
−π
1
1
16
HATICE ASLAN AND ALI GUVEN
holds. Therefore we obtain the following result by using Lemma 2.1.
k σa
n(f ) kp≤(cid:18) 1 + a
1 − a(cid:19)2/p
k f kp
(cid:3)
Theorem 10. Let 1 ≤ p < ∞.Then we have
k f − σa
n(f ) kp→ 0, n → ∞
for all f ∈ Lp(T).
Proof. Let σa
Lp(T). Take ε > 0.Then there exists ∃ g ∈ C(T) such that
n : Lp(T) → Lp(T), n = 0, 1, ... linear operator and f ∈
k f − g kp<
ε
3
.
Let g ∈ C(T). Then for n → ∞, we have σa
n(g) → g converges
uniformly by Theorem 3.2. Thus by using the Lusin Theorem (see in
[3]), there exists ∃ n ≥ N such that k σa
k f − σa
n(f ) kp=k f − g + g − σa
n(g) − g kp< ε
3.
n(g) − σa
n(f ) kp
n(g) + σa
≤k f − g kp + k g − σa
Therefore by Theorem 3.3, we have
n(g) kp + k σa
n(f − g) kp
k f − σa
n(f ) kp≤k f − g kp + k g − σa
n(g) kp +(cid:18)1 + a
1 − a(cid:19)1/p
k f − g kp
≤ (cid:18) 1 + a
1 − a(cid:19)2/p
+ 1! k f − g kp +
ε
3
< ε.
This completes the proof.
(cid:3)
4. Bernstein's inequality for Nonlinear Fourier Series
Applying the inequalities for the Ces`aro means of nonlinear
trigonometric series derived in the previous section, we can prove the
nonlinear version of the well-known Bernstein's inequality. For any
trigonometric polynomial tn(x) of order ≤ n, for every 1 ≤ p ≤ ∞, we
have
(cid:18)Z π
−π
t′
n(x) p dx(cid:19)1/p
≤ cn(cid:18)Z π
−π
tn(x) p dx(cid:19)1/p
.
(4.1)
The last inequality is known as integral Bernstein's inequality. The
following extension of (4.1) is true.
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS17
Theorem 11. Let 1 < p < ∞ and assume that ta ∈ τ a
inequality
n . Then the
Z π
−π
(4.2)
holds.
t′
n,a(x) p dx ≤ C(cid:18) 1+ a
1− a (cid:19)2(1+1/p)Z π
−π
tn,a(x) p dx
Proof. It is well known from the (2.1) equality that
1
π Z π
−π
where
tn,a(u) p Dn(θa(u) − θa(x))pa(u)du
Dn(u) =
1
2
+
cos ku
n
Xk=1
is the Dirichlet's kernel of order n. Let T = tn,a◦θ−1
we obtain
a . By the derivation,
t′
n,a(x) =
tn,a(u) p D′
n(θa(u) − θa(x))(−pa(x))pa(u)du
1
−π
π Z π
πZ π
1
−π
= pa(x)
= pa(x)
Tn,a(θa(u))D′
n(θa(u) − θa(x))pa(u)du
n(y)dy
1
−π
Tn,a(y + θa(x))D′
π Z π
Tn,a(θa(x) + y) n
Xk=1
1
πZ π
−π
= pa(x)
= pa(x)
= pa(x)
(k sin ky) +
1
−π
π Z π
π Z π
1
−π
Tn,a(θa(x) + y) n
Xk=1
Tn,a(θa(x) + y)2n sin ny 1
2
+
n−1
k sin ky! dy
(k sin(2n − k)y)! dy
Xk=1
cos ky(cid:19)! dy
Xk=1(cid:18)1 −
k
n
n−1
= pa(x)
Tn,a(θa(x) + y)2n sin nyKn−1(y)dy
1
π Z π
−π
where Kn−1 is the Fej´er's kernel of order n − 1. By taking the absolute
values, we get
T ′
n,a ≤(cid:18) 1+ a
1− a (cid:19) 2n sin ny
πZ π
1− a (cid:19) 2n
≤(cid:18) 1+ a
−π
1
1
πZ π
−π
Tn,a(θa(x) + y)Kn−1(y)dy
Tn,a(θa(x) + y)Kn−1(y)dy
18
HATICE ASLAN AND ALI GUVEN
If we use Theorem 3.3, we get that
1
2π(cid:18)Z π
−π
t′
1
≤
−π
≤(cid:18) 1+ a
1− a (cid:19) 2nσa
a(x) p dx(cid:19)1/p
2π(cid:18)Z π
2πZ π
≤(cid:18) 1+ a
1− a (cid:19) 2n(cid:18) 1
1− a (cid:19) 2nσa
≤(cid:18) 1+ a
1− a (cid:19) 2n(cid:18) 1+ a
1− a (cid:19)2+1/p
≤ 2n(cid:18)1+ a
n,a kp≤ cn(cid:18) 1+ a
≤(cid:18) 1+ a
k t′
n−1(tn,a, x).
n−1(tn,a, x) p dx(cid:19)1/p
−π
1− a (cid:19) 2nσa
(cid:18) 1+ a
n−1(tn,a, x) p dx(cid:19)1/p
σa
n−1(tn,a, x) kp
1− a (cid:19)2/p
k tn,a kp
k tn,a kp
From this we obtain
and this completes the proof.
1− a (cid:19)2+1/p
k ta kp
(cid:3)
References
[1] Chen Q. H.,Li L. Q., Qian T., Stability of frames generated by nonlinear fourier
atoms, Int J Wavelets Multiresolut Inf Process, 3 (2005) 465-476.
[2] Chen Q. H. and Li L. Q.,Qian T., Two families of unit analytic signals with
nonlinear phase, Phys D, 221 (2006) 1-12.
[3] DeVore, R. A., Lorentz G. G.,Constructive Approximation, Springer-Verlag,
New York, 1993.
[4] Giuseppe M., Gradimir V. M., Interpolation Process Basic Theory and
Applications, Springer, Berlin, 2008.
[5] Huang, C., Yang, L H., Approximation by the nonlinear Fourier basis, Sci
China Math, 54 (2011) 1207-1214.
[6] Reed M., Simon B., Methods of Modern Mathematical Physics, Academic
Press, New York, 1972.
[7] Qian T., Analytic signals and harmonic measures,Math Anal Appl,314
(2006)526-536.
[8] Qian T., Analytic unit quadrature signals with nonlinear phase, Phys
D,(303)(2005)80-87.
[9] Qian T., Chen Q H. , Characterization of analytic phase signals, Comp. Math
Appl., 51(2006) 1471-1482.
[10] Zygmund A., Trigonometric series, 2nd ed.,Cambridge Univ. Press.,New York,
1959.
ON THE MEAN SUMMABILITY OF SERIES BY NONLINEAR FOURIER BASIS19
Department of Mathematics, Firat University, Faculty of Science,
23119, Elazig, Turkey
E-mail address: [email protected] (Corresponding author)
Department of Mathematics Faculty of Art and Science Balikesir
University 10145, Balikesir Turkey
E-mail address: [email protected]
|
1302.2422 | 2 | 1302 | 2013-02-12T08:05:47 | Operator theoretic differences between Hardy and Dirichlet-type spaces | [
"math.FA"
] | For $0<p<\infty $, the Dirichlet-type space $\Dp$ consists of those analytic functions $f$ in the unit disc $\D$ such that $\int_\D|f'(z)|\sp p(1-|z|)^{p-1}\,dA(z)<\infty$. Motivated by operator theoretic differences between the Hardy space $H^p$ and $\Dp$, the integral operator {displaymath} T_g(f)(z)=\int_{0}^{z}f(\zeta)\,g'(\zeta)\,d\zeta,\quad z\in\D, {displaymath} acting from one of these spaces to another is studied. In particular, it is shown, on one hand, that $T_g:\Dp\to H^p$ is bounded if and only if $g\in\BMOA$ when $0<p\le 2$, and, on the other hand, that this equivalence is very far from being true if $p>2$. Those symbols $g$ such that $T_g:\Dp\to H^q$ is bounded (or compact) when $p<q$ are also characterized. Moreover, the best known sufficient $L^\infty$-type condition for a positive Borel measure $\mu$ on $\D$ to be a $p$-Carleson measures for $\Dp$, $p>2$, is significantly relaxed, and the established result is shown to be sharp in a very strong sense. | math.FA | math |
OPERATOR THEORETIC DIFFERENCES BETWEEN
HARDY AND DIRICHLET-TYPE SPACES
JOS´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
Abstract. For 0 < p < ∞, the Dirichlet-type space Dp
p−1 consists of
those analytic functions f in the unit disc D such that RD f ′(z)p(1 −
z)p−1 dA(z) < ∞. Motivated by operator theoretic differences between
the Hardy space H p and Dp
p−1, the integral operator
Tg(f )(z) =Z z
0
f (ζ) g ′(ζ) dζ,
z ∈ D,
In particular, it
acting from one of these spaces to another is studied.
is shown, on one hand, that Tg : Dp
p−1 → H p is bounded if and only if
g ∈ BMOA when 0 < p ≤ 2, and, on the other hand, that this equivalence
is very far from being true if p > 2. Those symbols g such that Tg : Dp
p−1 →
H q is bounded (or compact) when p < q are also characterized. Moreover,
the best known sufficient L∞-type condition for a positive Borel measure µ
on D to be a p-Carleson measures for Dp
p−1, p > 2, is significantly relaxed,
and the established result is shown to be sharp in a very strong sense.
1. Introduction and main results
Let H(D) denote the algebra of all analytic functions in the unit disc D =
{z : z < 1} of the complex plane C. Let T be the boundary of D. The
Carleson square associated with an interval I ⊂ T is the set S(I) = {reit :
eit ∈ I, 1 − I ≤ r < 1}, where E denotes the normalized Lebesgue measure
of the set E ⊂ T. For our purposes it is also convenient to define for each
a ∈ D \ {0} the interval Ia = (cid:8)eiθ : arg(ae−iθ) ≤ π(1 − a)(cid:9), and denote
S(a) = S(Ia). For 0 < p ≤ ∞, the Hardy space H p consists of those f ∈ H(D)
for which
kf kH p = lim
r→1−
Mp(r, f ) < ∞,
Date: October 3, 2018.
2010 Mathematics Subject Classification. Primary 47G10; Secondary 30H10.
Key words and phrases. Operator theoretic differences, Hardy spaces, Spaces of Dirichlet
type, Integral operators, Carleson measures.
This research was supported in part by the Ram´on y Cajal program of MICINN (Spain),
Ministerio de Educaci´on y Ciencia, Spain, (MTM2011-25502), from La Junta de Andaluc´ıa,
(FQM210) and (P09-FQM-4468), MICINN- Spain ref. MTM2011-26538.
2
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
where
and
Mp(r, f ) =(cid:18) 1
2πZ 2π
0
1
p
f (reiθ)p dθ(cid:19)
,
0 < p < ∞,
M∞(r, f ) = max
0≤θ≤2π
f (reiθ).
For the theory of the Hardy spaces, see [9, 11].
For 0 < p < ∞ and −1 < α < ∞, the Dirichlet space Dp
α consists of those
f ∈ H(D) such that
kf kp
Dp
α
where dA(z) = dx dy
π
f ′(z)p(1 − z2)α dA(z) + f (0)p < ∞,
=ZD
is the normalized Lebesgue area measure on D.
The purpose of this study is to underscore operator theoretic differences
between the closely related spaces Dp
p−1 and H p. Before going to that, it
is appropriate to recall inclusion relations between these spaces. The classical
Littlewood-Paley formula implies D2
1 = H 2. Moreover, it is well known [10, 17]
that
(1.1)
and
Dp
p−1 ( H p,
0 < p < 2,
(1.2)
It is also worth mentioning that there are no inclusion relations between Dp
and Dq
2 < p < ∞.
q−1 when p 6= q [13].
p−1,
p−1
H p ( Dp
A natural way to illustrate differences between two given spaces is to consider
classical operators acting on them. For example, if 0 < p < 2, then the
behavior of the composition operator Cϕ(f ) = f ◦ ϕ reveals that Dp
p−1 is in
a sense a much smaller space than H p. Namely, it follows from Littlewood's
subordination theorem that Cϕ : H p → H p is bounded for each 0 < p < ∞
and all analytic self-maps ϕ of D, but in contrast to this, there are symbols ϕ
which induce unbounded operators Cϕ : Dp
p−1 when 0 < p < 2 [8,
Theorem 1.1(b)]. As in the case of Hardy spaces, any composition operator
Cϕ : Dp
p−1 is bounded when 2 ≤ p < ∞.
p−1 → Dp
p−1 → Dp
There are operators which do not distinguish between Dp
p−1 and H p. For a
given g ∈ H(D), the generalized Hilbert operator Hg is defined by
(1.3)
Hg(f )(z) =Z 1
0
f (t)g′(tz) dt,
p−1 → Dp
for any f ∈ H(D) such that R 1
If 1 < p < ∞, then Hg :
Dp
p−1 is bounded
(compact) by [12]. Moreover, the same condition, depending on g and p,
describes the boundedness (compactness) of the operators Hg : Dp
p−1 → H p
p−1 is bounded (compact) if and only if Hg : H p → Dp
0 f (t) dt < ∞.
OPERATOR THEORETIC DIFFERENCES
3
and Hg : H p → H p when 1 < p ≤ 2. As far as we known, the problem of
characterizing those symbols g for which Hg : Dp
p−1 → H p and Hg : H p → H p
are bounded when 2 < p < ∞ remains unsolved.
We will next study operator theoretic differences between Dp
p−1 and H p by
considering the integral operator
Tg(f )(z) =Z z
0
f (ζ) g′(ζ) dζ,
z ∈ D.
The bilinear operator (f, g) →R f g′ was introduced by Calder´on in harmonic
analysis in the 60's [5]. After his research on commutators of singular integral
operators, this bilinear form and its different variations, usually called "para-
products", have been extensively studied and they have become a fundamental
tool in harmonic analysis. Pommerenke was probably one of the first complex
function theorists to consider the operator Tg. He used it in late 70's to study
the space BMOA, which consists of those functions in the Hardy space H 1 that
have bounded mean oscillation on the boundary T [20]. The space BMOA can
be equipped with several different equivalent norms [11], here we will use the
one given by
kgk2
BMOA = sup
a∈D RS(a) g′(z)2(1 − z2) dA(z)
1 − a
+ g(0)2.
Two decades later, in late 90's, the pioneering works by Aleman and Siskakis [2,
3] lead to an abundant research activity on the operator Tg. In particular, those
analytic symbols g such that Tg : H p → H q is bounded were characterized by
Aleman, Cima and Siskakis [1, 2]. Their result in the case p = q says that
Tg : H p → H p is bounded if and only if g ∈ BMOA. Our first result shows
that whenever 0 < p ≤ 2, the domain space H p can be replaced by Dp
p−1.
Theorem 1. Let 0 < p ≤ 2 and g ∈ H(D). Then the following are equivalent:
p−1 → H p is bounded;
(i) Tg : Dp
(ii) Tg : H p → H p is bounded;
(iii) g ∈ BMOA.
The implication (ii)⇒(i) is a direct consequence of (1.1), so our contribution
here consists of showing (i)⇒(iii). The proof of the implication (ii)⇒(iii) in [1,
2] relies on several powerful properties of BMOA and H p such as the conformal
invariance of BMOA. Our proof is based on a circle of ideas developed in [19,
Chapter 4], and does not rely on these properties. Instead, the Fefferman-Stein
formula [22], which states that
(1.4)
kf kp
H p ≍ZT
Sp
f (ζ) dζ + f (0)p,
plays an important role in the reasoning. Here, dζ denotes the arclength
measure on T, Sf denotes the usual square function, also called the Lusin area
4
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
function,
(1.5)
Sf (ζ) =(cid:18)ZΓσ(ζ)
f ′(z)2 dA(z)(cid:19)1/2
,
ζ ∈ T,
where Γσ(ζ) denotes a nontangential approach region (a Stolz angle) with
vertex at ζ and of aperture σ.
We also show that the statement in Theorem 1 drastically fails for p > 2.
In order to give the precise statement, we will need to fix the notation. The
disc algebra A is the space of all analytic functions on D which are continuous
on the boundary T. For 0 < α ≤ 1, the Lipschitz space Λ(α) consists of those
g ∈ H(D), having a non-tangential limit g(eiθ) almost everywhere on T, such
that
sup
θ∈[0,2π], 0<t<1
tα
g(ei(θ+t)) − g(eiθ)
< ∞.
The "little oh"counterpart of this space is denoted by λ(α). The following
chain of strict inclusions is known:
λ(α) ( Λ(α) ( A ( H ∞ ( BMOA ( B,
0 < α ≤ 1.
Here, as usual, B stands for the Bloch space which consists of those f ∈ H(D)
such that kf kB = supz∈D f ′(z)(1 − z2) + f (0) < ∞.
Theorem 2. Let 2 < p < ∞ and g ∈ H(D).
(i) If Tg : Dp
(ii) There exist g ∈ A and f ∈ Dp
p−1 → H p is bounded, then g ∈ BMOA.
p−1 such that Tg(f ) /∈ H p.
Part (ii) shows that Dp
p−1 is in a sense a much larger space than H p when
p > 2. This is true because we may choose the inducing symbol g to be as
smooth as continuous on the boundary, but still a suitably chosen f ∈ Dp
p−1
establishes Tg(f ) /∈ H p. In contrast to this, when the inducing index of the
domain space is strictly smaller than the one of the target space, that is p < q,
then Tg does not distinguish between Dp
p−1 and H p.
Theorem 3. Let 0 < p < q < ∞ and g ∈ H(D).
q ≤ 1, then the following are equivalent:
(a) If 1
p − 1
(i) Tg : Dp
(ii) Tg : H p → H q is bounded;
(iii) g ∈ Λ( 1
p−1 → H q is bounded;
p − 1
q ).
(b) If 1
p − 1
constant.
q > 1, then Tg : Dp
p−1 → H q is bounded if and only if g is
Part (a) allows us to deduce a strengthened version of the classical result
of Hardy-Littlewood which states that a primitive of each function f ∈ H p,
0 < p < 1, belongs to H
1−p .
p
OPERATOR THEORETIC DIFFERENCES
5
+ 1
p2
1
Proposition 4. Let p, p1 and p2 be positive numbers such that p < 1 < p2 and
p = 1
1
p1−1 and f2 ∈ H(D)
p1
p
1−p .
. If f ∈ H(D) such that f = f1 · f2 where f1 ∈ Dp1
(1−z)1/p2(cid:17), then f is the derivative of a function in H
satisfies f2(z) = O(cid:16)
The statement in Proposition 4 with H p1 in place of Dp1
p1−1 was proved by
Aleman and Cima [1, p. 158]. The strict inclusions (1.1) and (1.2) show that
their result is better when p1 < 2 meanwhile the situation is another way round
when p1 > 2.
An important ingredient in the proofs of both Theorems 1 and 3 is the
following result on a Hormander-type maximal function
M(ϕ)(z) = sup
I: z∈S(I)
1
IZI
ϕ(ζ)
dζ
2π
,
z ∈ D,
defined for each 2π-periodic function ϕ(eiθ) ∈ L1(T)
Theorem A. Let 0 < p ≤ q < ∞ and 0 < α < ∞ such that pα > 1. Let µ be
α )]α : Lp(T) → Lq(µ) is bounded
a positive Borel measure on D. Then [M((·)
q
if and only if there exists a constant C > 0 such that µ(S(I)) ≤ CI
p for all
I ⊂ T Moreover,
1
k[M((·)
1
α )]αkq ≍ sup
I⊂T
µ (S(I))
q
p
I
.
q
p are known as q
This result follows by the well-known works by Carleson [6, 7], and hence
p -Carleson measures.
the measures µ for which µ(S(I)) ≤ CI
For more recent references, see either [9, Section 9.5], or the proof of [19,
Theorem 2.1] for a similar result. Theorem A has been used to characterize
so-called q-Carleson measures for Hardy spaces. Recall that, for a given Banach
space (or a complete metric space) X of analytic functions on D, a positive
Borel measure µ on D is called a q-Carleson measure for X if the identity
operator Id : X → Lq(µ) is bounded.
In nowadays these measures are a
standard tool in the operator theory in spaces of analytic functions in D.
p−1 to Dp
Let us now turn back to the two remaining cases that are not covered by
Theorems 1 and 2. They are the ones in which the operator Tg acts from either
H p or Dp
p−1. It is easy to see that, in terms of the language of the
previous paragraph, Tg : H p → Dq
q−1 is bounded if and only if µg,q = g′(z)q(1−
z2)q−1 dA(z) is a q-Carleson measure for H p. Therefore, in this case the
symbols g that induce bounded operators get characterized by [9, Theorem 9.5],
when q ≥ p, and [18] if q < p. Analogously, it follows that Tg : Dp
q−1
bounded if and only if µg,q is a q-Carleson measure for Dp
p−1. Unfortunately,
as far as we know, the existing literature does not offer a characterization of
these measures for the full range of parameter values in terms of a condition
depending on µ only. It is known that they coincide with q-Carleson measures
p−1 → Dq
6
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
for H p and can therefore be described by the condition
(1.6)
µ (S(I))
Iq/p < ∞,
sup
I⊂T
provided q > p [16, Theorem 1(a)]. This statement remains valid also in the
diagonal case q = p, if p ≤ 2, but fails for p > 2 [15, 21]. In more general terms,
the p-Carleson measures for Dp
α are known excepting the case α = p − 1 for
p > 2 [4, 21]. This corresponds to the diagonal case q = p > 2 which interests
us in particular. What is known with respect to this case, is that µ being a
1-Carleson measure is a necessary but not a sufficient condition for µ to be a
p-Carleson measure for Dp
p−1 [15], and that the more restrictive condition
sup
I⊂T
µ (S(I))
I(cid:16)log e
I(cid:17)−p/2 < ∞
is a sufficient condition for Id : Dp
p−1 → Lp(µ) to be bounded [14]. Our next
result shows that this best known sufficient condition can be relaxed by one
logarithmic term.
Theorem 5. Let 2 < p < ∞, and let µ be a positive Borel measure on D. If
(1.7)
sup
I⊂T
then µ is a p-Carleson measure for Dp
p−1.
µ (S(I))
I(cid:16)log e
I(cid:17)−p/2+1 < ∞,
We will see in Proposition 12 that the statement in Theorem 5 is sharp in a
very strong sense.
The remaining part of the paper is organized as follows. In Section 2 we
state and prove some preliminary results. Theorems 1 and 3 and their ex-
pected analogues for compact operators as well as Proposition 4 are proved
in Section 3. In Section 4 we will deal with the growth of integral means of
functions f ∈ Dp
p−1, p > 2, and we will prove Theorem 2.
Before proceeding further, a word about notation to be used. We will write
kT k(X,Y ) for the norm of an operator T : X → Y , and if no confusion arises
with regards to X and Y , we will simply write kT k. Moreover, for two real-
valued functions E1, E2 we write E1 ≍ E2 or E1 . E2, if there exists a positive
constant k, independent of the argument, such that 1
k E1 ≤ E2 ≤ kE1 or
E1 ≤ kE2, respectively.
2. Preliminaries
We begin with a straightforward but useful estimate that will be used in
proofs of Theorems 1 and 3.
OPERATOR THEORETIC DIFFERENCES
7
Lemma 6. Let 0 < q, p < ∞ and g ∈ H(D). If Tg : Dp
then
p−1 → H q is bounded,
(2.1)
M∞(r, g′) .
Proof. The functions
kTgk(Dp
(1 − r)1− 1
p−1,H q)
p + 1
q
,
0 ≤ r < 1.
Fa,p,γ(z) =(cid:18) 1 − a2
1 − az (cid:19)
1+γ
p
,
0 < γ < ∞,
a ∈ D,
satisfy
(2.2)
Fa,p,γ(z) ≍ 1,
z ∈ S(a),
and a calculation shows that shows that
kFa,p,γkp
Dp
p−1
≍ (1 − a),
a ∈ D.
Since Tg : Dp
p−1 → H q is bounded by the assumption, the well known relations
M∞(r, f ) . Mq(cid:0) 1+r
2 , f(cid:1) (1 − r)− 1
for all f ∈ H(D) (see [9, Chapter 5]), yield
2 , f(cid:1) (1 − r)−1, valid
g′(a) = (Tg(Fa,p,γ))′(a) .
q and Mq(r, f ′) . Mq(cid:0) 1+r
, (Tg(Fa,p,γ))′(cid:17)
Mq(cid:16) 1+a
2
1
q
(1 − a)
Mq(cid:16) 3+a
4
, Tg(Fa,p,γ)(cid:17)
(1 − a)1+ 1
q
.
kTgk(Dp
p−1,H q)kFa,p,γkDp
(1 − a)1+ 1
q
p−1
.
.
q
kTg(Fa,p,γ)kH q
(1 − a)1+ 1
kTgk(Dp
(1 − a)1+ 1
.
p−1,H q)
q − 1
p
,
a ∈ D,
and the assertion follows.
(cid:3)
We next recall some suitable reformulations of Lipschitz spaces Λ(α) [9].
Lemma B. Let 0 < α ≤ 1. Then the following are equivalent:
(i) g ∈ Λ(α);
(ii) M∞(r, g′) = O(cid:16)
1
(1−r)1−α(cid:17) ,
r → 1−;
(iii) dµg(z) = g′(z)2(1 − z2) dA(z) satisfies sup
I⊂T
µg (S(I))
I2α+1 < ∞.
We will also need the following result [16, Theorem 1(i)].
Theorem C. Let 0 < p < q < ∞ and µ be a positive Borel measure on D.
Then µ is a q-Carleson measure for Dp
p -Carleson
p−1 if and only if µ is a q
8
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
measure. Moreover,
kId(Dp
p−1,Lq(µ))kq ≍ sup
I⊂T
µ (S(I))
q
p
I
.
3. Integral operators from Hardy to Dirichlet type spaces Dq
q−1
Proof of Theorem 1. It is known that Tg : H p → H p is bounded if and only
if g ∈ BMOA [1], and therefore (ii) and (iii) are equivalent. Moreover, since
Dp
p−1 ⊂ H p for 0 < p ≤ 2, (ii) implies (i). To complete the proof we will
show that g ∈ BMOA, whenever Tg : Dp
p−1 → H p is bounded. To see this,
note first that kgkB . kTgk(Dp
p−1,H p) by Lemma 6, and thus g ∈ B. Let now
1 < α, β < ∞ such that β/α = p/2 < 1, and let α′ and β′ be the conjugate
indexes of α and β. Assume for a moment that g′ is continuous on D. Then
(2.2), Fubini's theorem and Holder's inequality yield
g′(z)2(1 − z2) dA(z)
ZS(a)
≍ZT(cid:18)ZS(a)∩Γσ (ζ)
≤ ZT(cid:18)ZΓσ(ζ)
·
ZT(cid:18)ZΓσ(ζ)∩S(a)
≍ kTg(Fa,p,γ)k
p
β
g′(z)2Fa,p,γ(z)2 dA(z)(cid:19)
g′(z)2Fa,p,γ(z)2 dA(z)(cid:19)
β
α
dζ
1
β
1
α + 1
α′
dζ!
1
β′
β′
α′
g′(z)2 dA(z)(cid:19)
dζ
a ∈ D,
H pkSg(χS(a))k
1
α′
β′
α′ (T)
L
,
(3.1)
where
Sg(ϕ)(ζ) =ZΓσ(ζ)
ϕ(z)2g′(z)2 dA(z),
ζ ∈ T,
α′(cid:17)′
for any bounded function ϕ on D. Now (cid:16) β′
≤1(cid:12)(cid:12)(cid:12)(cid:12)
ZT
kSg(χS(a))k
β′
α′ (T)
(3.2)
α−β (T)
sup
β(α−1)
=
L
khk
L
= β(α−1)
α−β > 1, and hence
h(ζ)Sg(χS(a))(ζ) dζ(cid:12)(cid:12)(cid:12)(cid:12)
by the duality. To estimate the right hand side, we shall write I(z) for the
arc {ζ ∈ T : z ∈ Γσ(ζ)} with I(z) ≍ 1 − z. Then Fubini's theorem, Holder's
OPERATOR THEORETIC DIFFERENCES
9
inequality and Theorem A yield
(3.3)
h(ζ)ZΓσ(ζ)∩S(a)
g′(z)2 dA(z) dζ
h(ζ) dζ(cid:19) dA(z)
g′(z)2(1 − z2)M(h)(z) dA(z)
h(ζ)Sg(χS(a))(ζ) dζ(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
ZT
≤ZT
g′(z)2(1 − z2)(cid:18) 1
1 − z2 ZI(z)
≍ZS(a)
.ZS(a)
≤(cid:18)ZS(a)
·(cid:18)ZD
g′(z)2(1 − z2) dA(z)(cid:19)
.(cid:18)ZS(a)
a∈D RS(a) g′(z)2(1 − z2) dA(z)
· sup
g′(z)2(1 − z2) dA(z)(cid:19)
α′ (cid:17)′
M(h)(cid:16) β′
1 − a
α′
β′
α′
β′
g′(z)2(1 − z2) dA(z)(cid:19)1− α′
β′
β′
!1− α′
khk
L( β′
α′ )′
(T)
.
Since any dilated function gr(z) = g(rz), 0 < r < 1, is analytic on D(cid:0)0, 1
r(cid:1),
by replacing g by gr in (3.1) -- (3.3), we deduce
ZS(a)
g′
r(z)2(1 − z2) dA(z)
(3.4)
. kTgr (Fa,p,γ)k
p
β
H p(cid:18)ZS(a)
g′
r(z)2(1 − z2) dA(z)(cid:19)
a∈D RS(a) g′
· sup
1 − a
r(z)2(1 − z2) dA(z)
1
α′ (cid:16)1− α′
β′ (cid:17)
.
!
1
β′
We claim that there exists γ and a constant C = C(p, γ) > 0 such that
(3.5)
sup
0<r<1
kTgr(Fa,p,γ)kp
H p ≤ CkTgkp
(Dp
p−1,H p)(1 − a),
a ∈ D,
the proof of which is postponed for a moment. Now this combined with (3.4)
and Fatou's lemma yield
sup
a∈D RS(a) g′(z)2(1 − z2) dA(z)
1 − a
and so g ∈ BMOA.
. kTgk2
(Dp
p−1,H p),
10
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
It remains to prove (3.5). To see this fix γ > p. Recall that
kTgr(Fa,p,γ)kp
H p ≍ZT(cid:18)ZΓσ(ζ)
r2g′(rz))2Fa,p,γ(z)2dA(z)(cid:19)p/2
dζ.
If a < 1
2, then
kTgr(Fa,p,γ)kp
H p . (1 − a)γ+1ZT(cid:18)ZΓσ(ζ)
r2g′(rz)2dA(z)(cid:19)
p
2
dζ
≍ (1 − a)γ+1kgr − g(0)kp
= (1 − a)kTg(1)kp
H p ≤ (1 − a)kg − g(0)kp
H p
H p . (1 − a)kTgkp
(Dp
p−1,H p).
Let now 1
2 ≤ a < 1
kTgr(Fa,p,γ)kp
2−r . Then 1 − arz ≤ 21 − az for all z ∈ D, and hence
H p . k(Tg(Fa,p,γ))rkp
≤ kTgkp
p−1,H p)kFa,p,γkp
(Dp
Dp
p−1
H p ≤ kTg(Fa,p,γ)kp
H p
≍ kTgkp
(Dp
p−1,H p)(1 − a).
In the remaining case 1
hence
2−r ≤ a < 1 we have r ≤ 2 − 1
a ≤ a. Now γ > p, and
kTgr(Fa)kp
H p . M p
p !p/2
2(γ+1)
dζ
dA(z)
. M p
∞(r, g′)(1 − a)γ+1ZT ZΓσ(ζ)
∞(a, g′)(1 − a)γ+1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
1
(1 − az)
≍ (M∞(a, g′)(1 − a))p (1 − a) ≤ kgkp
. kTgkp
p−1,H p)(1 − a).
(Dp
1 − az
p
γ+1
p −1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
H p
B(1 − a)
By combining these three separate cases we deduce (3.5).
(cid:3)
Next, we will prove Theorem 3 by using similar ideas that were employed in
the proof of Theorem 1.
Proof of Theorem 3. It is known that (ii) and (iii) are equivalent [1]. Further,
Lemma 6 and Lemma B give (i)⇒(iii) and (b). Moreover, if 0 < p ≤ 2, then
Dp
p−1 ⊂ H p and hence, in this case, (ii) implies (i). To complete the proof,
we show that (iii) implies (i) when 2 < p < ∞. Since q > 2, Lq/2(T) can be
identified with the dual of L
. Therefore,
Tg : Dp
p−1 → H q is bounded if and only if
q−2 (T)(cid:17)⋆
q
q
q−2 (T), that is, Lq/2(T) =(cid:16)L
f (z)2g′(z)2 dA(z)(cid:19) dζ(cid:12)(cid:12)(cid:12)(cid:12)
. khk
h(ζ)(cid:18)ZΓσ(ζ)
(cid:12)(cid:12)(cid:12)(cid:12)
ZT
q
q−2 (T)
kf k2
Dp
p−1
L
OPERATOR THEORETIC DIFFERENCES
11
q
q−2 (T) and f ∈ Dp
p−1. To see this, we use first Fubini's theorem
for all h ∈ L
to obtain
h(ζ)(cid:18)ZΓσ(ζ)
f (z)2g′(z)2(cid:18)ZI(z)
f (z)2g′(z)2 dA(z)(cid:19) dζ(cid:12)(cid:12)(cid:12)(cid:12)
h(ζ) dζ(cid:19) dA(z)
f (z)2M(h)(z)g′(z)2(1 − z2) dA(z).
(cid:12)(cid:12)(cid:12)(cid:12)
ZT
≤ZD
.ZD
Since g′(z)2(1−z2) dA(z) is a(cid:16)2(cid:16) 1
q(cid:17) + 1(cid:17)-Carleson measure by Lemma B,
q )/p, we may estimate the last integral upwards
and 2( 1
by Holder's inequality, Theorem C and Theorem A to
q ) + 1 = (2 + p − 2p
p − 1
p − 1
f (z)2+p− 2p
q g′(z)2(1 − z2) dA(z)(cid:19)
2q
(2+p)q−2p
(M(h)(z))1+ 2q
p(q−2) g′(z)2(1 − z2) dA(z)(cid:19)
1
2q
p(q−2)
1+
(cid:18)ZD
·(cid:18)ZD
. kf k2
Dp
khk
q
q−2 (T)
L
.
p−1
These estimates give the desired inequality for all h ∈ L
and thus Tg : Dp
p−1 → H q is bounded.
We now prove Proposition 4.
q
q−2 (T) and f ∈ Dp
p−1,
(cid:3)
Proof of Proposition 4. Let F2 be such that F ′
2 = f2. Then F ′
2(z) =
) by Lemma B.
1
O(cid:16)
R z
(1−z)1/p2(cid:17) by the assumption, and hence F2 ∈ Λ(1 − 1
2(ζ)f1(ζ) dζ =R z
Now Theorem 3 implies that the integral operator TF2 : Dp1
bounded, and since f1 ∈ Dp1
0 F ′
We finish this section by proving the expected versions of Theorems 1 and 3
for compact operators. The next auxiliary result is standard, and therefore its
proof is omitted.
1−p is
p1−1 by the assumption, we deduce TF2(f1)(z) =
(cid:3)
1−p , which gives the assertion.
0 f (ζ) dζ ∈ H
p1−1 → H
p2
p
p
Lemma 7. Let 0 < p, q < ∞ and g ∈ H(D). Then the following are equivalent:
(i) Tg : Dp
(ii) For any sequence of analytic functions {fn}∞
p−1 → H q is compact;
uniformly to 0 on compact subsets of D and satisfies supn∈N kfnkDp
∞, we have limn→∞ kTg(fn)kH q = 0.
p−1
n=1 on D that converges
<
Obviously the statement in this lemma remains valid if H p is replaced
by Dp
p−1.
12
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
The space VMOA consists of those functions in the Hardy space H 1 that
have vanishing mean oscillation on the boundary T. It is known that this space
is the closure of polynomials in BMOA and is characterized by the condition
a→1− RS(a) g′(z)2(1 − z2) dA(z)
1 − a
lim
= 0.
Theorem 8. Let 0 < p ≤ 2 and g ∈ H(D). Then the following are equivalent:
p−1 → H p is compact;
(i) Tg : Dp
(ii) Tg : H p → H p is compact;
(iii) g ∈ VMOA.
Proof. It is known that (ii) and (iii) are equivalent by [1]. Moreover, by bearing
in mind Lemma 7 and (1.1), we see that (ii) implies (i). It remains to show
that g ∈ VMOA, whenever Tg : Dp
p−1 → H p is compact. Since the proof of
this implication is similar to its counterpart in the proof of Theorem 1, we
only show in detail those steps that are significantly different. First observe,
that g ∈ BMOA by Theorem 1. Let fa,p,γ = Fa,p,γ
(1−a)1/p , where γ > 0 and
Fa,p,γ are those functions defined in the proof of Lemma 6.
It is clear that
≍ 1 and fa,p → 0, as a → 1−, uniformly in compact subsets
kfa,p,γkDp
of D. Therefore kTg(fa,p,γ)kH p → 0, as a → 1−, by Lemma 7. Now, let
1 < α, β < ∞ such that β/α = p/2 < 1. Arguing as in (3.1), we deduce
p−1
1
2
p ZS(a)
g′(z)2(1 − z2) dA(z) . kTg(fa,p,γ)k
H pkSg(χS(a)fa,p,γ)k
(1 − a)
for all a ∈ D. Following the reasoning in the proof of Theorem 1 and bearing
in mind that g ∈ BMOA, we obtain
L
p
β
1
α′
β′
α′ (T)
RS(a) g′(z)2(1 − z2) dA(z)
(1 − a)
2
p
. kTg(fa,p,γ)k
p
β
H p(cid:16)RS(a) g′(z)2(1 − z2) dA(z)(cid:17)
(1 − a)
p · 1
α′
2
which is equivalent to
RS(a) g′(z)2(1 − z2) dA(z)
(1 − a)
. kTg(fa,p,γ)kp
H p.
Therefore g ∈ VMOA.
α′
β′ · 1
α′
,
(cid:3)
It is known that the "little oh"analogue of Lemma B is valid. This together
with appropriate modifications in the proofs of Lemma 6 and Theorem 3 give
the next result.
Theorem 9. Let 0 < p < q < ∞, 1
are equivalent:
(i) Tg : Dp
(ii) Tg : H p → H q is compact;
p−1 → H q is compact;
p − 1
q ≤ 1, and g ∈ H(D). The following
OPERATOR THEORETIC DIFFERENCES
13
(iii) g ∈ λ( 1
p − 1
q ).
4. Growth of integral means of functions in Dp
p−1.
In this section we will prove sharp estimates for the growth of Mp(r, f )
p−1 and 0 < p < 2, then Mp(r, f ) is
p−1 and 2 < p < ∞. If f ∈ Dp
when f ∈ Dp
uniformly bounded due to (1.1).
Lemma 10. Let 2 < p < ∞ and Φ : [0, 1) → (1, ∞) be a differentiable
increasing unbounded function such that Φ′(r)
Φ(r) (1 − r) is decreasing. Then the
following hold:
(i) For any f ∈ Dp
(ii) there exists f ∈ Dp
(4.1)
Mq(r, f ) &(cid:18)log
1
2 − 1
p(cid:17) ,
p−1 such that
p−1, Mp(r, f ) = o(cid:16)(cid:0)log e
1−r(cid:1)
(1 − r)(cid:19)
2 (cid:18) Φ′(r)
e
1 − r(cid:19)
Φ2(r)
1
r → 1−;
1
p
,
0 < r < 1,
for any fixed 0 < q < ∞.
Part (i) is essentially known, but we include a proof for the sake of com-
pleteness. Part (ii), apart from showing that (i) is sharp in a very strong, will
be used to prove Theorem 2(ii) and the sharpness of Theorem 5. It is also
worth noticing that each function
(4.2)
ΦN,α(r) =(cid:18)logN
expN 2
1 − r (cid:19)α
, N ∈ N = {1, 2, . . .},
0 < α < ∞,
satisfies both hypotheses on the auxiliary function Φ in Lemma 10. Here,
as usual, logn x = log(logn−1 x), log1 x = log x, expn x = exp(expn−1 x) and
exp1 x = ex.
Proof of Lemma 10. (i) First observe that [14, Theorem 1.4] yields
(4.3)
where Ap
v p
2
Dp
p−1 ⊂ Ap
v p
2
,
kf kp
Dp
p−1
& kf kp
Ap
v p
2
,
f ∈ H(D),
denotes the weighted Bergman space induced by the rapidly increas-
ing weight v p
2
(z) =
1
(1−z)(log
,
p
2
e
1−z )
z ∈ D, see [19, Section 1.2]. Therefore,
kf kp
Dp
p−1
& kf kp
Ap
v p
2
r
≥Z 1
p (r, f )(cid:18)log
sM p
p (s, f )v p
2
2
e
1 − r(cid:19)1− p
≍ M p
(s) ds ≥ M p
p (r, f )Z 1
r
sv p
2
(s) ds
,
0 < r < 1,
and (i) follows.
14
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
(ii) Let Φ be as in the lemma. Consider the lacunary series
(4.4)
f (z) =
∞
Xk=1(cid:18) h(rk) − h(rk−1)
Φ(rk)
1
p
(cid:19)
z2k
,
rk = 1 − 2−k,
k ∈ N,
where h(r) = log Φ(r) is a positive function such that h′(r)(1 − r) is decreasing
by the assumptions. By [15, Proposition 3.2],
kf kp
Dp
p−1
.
=
and thus f ∈ Dp
p−1.
On the other hand
∞
Φ(rk)
Xk=1(cid:18) h(rk) − h(rk−1)
(cid:19)
Xk=1 R rk
≤Z 1
h′(t) dt
Φ(rk)
rk−1
∞
0
h′(t)
Φ(t)
dt = Φ(0)−1 < 1,
2
p
2
p
(cid:19)
(cid:19)
r2k+1
N
r2k+1
N
∞
N
Φ(rk)
Φ(rk)
Xk=1(cid:18)h(rk) − h(rk−1)
Xk=1(cid:18)h(rk) − h(rk−1)
Xk=1 Z rk
Xk=1
(Φ(rN ))
r2N+1
N
(Φ(rN ))
r2N+1
N
(log 2)
rk−1
N
N
2
p
2
p
2
p
M 2
2 (rN , f ) =
≥
≥
≥
&
h′(s)(1 − s)
(h′(rk)(1 − rk))
2
p
ds
1 − s!
2
p
1
(Φ(rN ))
2
p
(h′(rN )(1 − rN ))
2
p N
Let r ∈ [ 1
[22, Theorem 8.20 in p. 215 Vol I] yields
2, 1) be given, and choose N ∈ N such that rN ≤ r < rN +1. Then
M 2
q (r, f ) ≍ M 2
2 (r, f ) ≥ M 2
2 (rN , f ) &
1
(Φ(rN ))
2
p
(h′(rN )(1 − rN ))
2
p N
&
1
(Φ(r))
2
p
(h′(r)(1 − r))
2
p log
e
1 − r
≍(cid:18)log
e
1 − r(cid:19)(cid:18) Φ′(r)
Φ2(r)
2
p
,
(1 − r)(cid:19)
which finishes the proof.
(cid:3)
With these preparations we are ready to prove Theorem 2.
OPERATOR THEORETIC DIFFERENCES
15
Proof of Theorem 2 (i) If Tg : Dp
bounded because H p ( Dp
p−1 → H p is bounded, then Tg : H p → H p is
p−1 for 2 < p < ∞ by (1.2), and hence g ∈ BMOA.
(ii) In this part we use ideas from the proof of [15, Theorem 2.1]. Take a
function Φ as in Lemma 10 and let f ∈ Dp
p−1 the lacunary series associated
to Φ via (4.4). By using [22, Theorem 8.25, Chap. V , Vol. I], we find two
constants A > 0 and B > 0 such that for every r ∈ (0, 1) the set
(4.5)
Er = {t ∈ [0, 2π] : f (reit) > BM2(r, f )}
has the Lebesgue measure greater than or equal to A. Let now g be a lacunary
series. By using [22, Lemma 6.5, Chap. V, Vol. I] we find a constant C1 > 0
such that
(4.6)
ZEr
g′(reit)2 dt ≥ C1AM 2
2 (r, g′) = C2M 2
2 (r, g′),
0 < r < 1,
where C2 = C1A. Bearing in mind the definition (4.5) of the sets Er and using
(4.6), we obtain
kTg(f )k2
H p ≥ kTg(f )k2
f (z)2g′(z)2(1 − z2) dA(z)
(4.7)
r(1 − r)M 2
g′(reit)2 dt dr
f (reit)2g′(reit)2 dt dr
H 2 &ZD
2 (r, f )ZEr
r(1 − r)M 2
2 (r, f )M 2
2 (r, g′) dr
r(1 − r)ZEr
0
0
≥Z 1
≥ B2Z 1
≥ B2C2Z 1
≥ B2C2CZ 1
0
0
Choose now Φ(r) =(cid:0)log e
(cid:18)log
e
1 − r(cid:19)(cid:18) Φ′(r)
Φ2(r)
e
Φ2(r)
1 − r(cid:19)(cid:18) Φ′(r)
r(1 − r)(cid:18)log
1−r(cid:1)ε, where 0 < ε < p
≍(cid:18)log
(1 − r)(cid:19)
2
p
2 − 1, so that
e
1 − r(cid:19)1− 2
p (1+ε)
.
2
p
(1 − r)(cid:19)
M 2
2 (r, g′) dr.
Further, let
g(z) =
∞
Xj=0
1
(j + 1) (log j + 1)α z22j
,
1 < α < ∞.
Then clearly g ∈ A. Moreover, since ω(r) = (1 − r)(cid:0)log e
so-called regular weight, we deduce
1−r(cid:1)1− 2
p (1+ε)
is a
r2n+1ω(r) ≍ n−1ω(1 − n−1), n ∈ N,
Z 1
0
16
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
by [19, Lemma 1.3 and (1.1)]. This together with (4.7) yields
2 (r, g′) dr
e
M 2
p (1+ε)
1 − r(cid:19)1− 2
r22j +1−1(1 − r)(cid:18)log
0
22j+1
H p &Z 1
r(1 − r)(cid:18)log
(j + 1)2 (log j + 1)2α Z 1
(j + 1)2 (log j + 1)2α = ∞,
2(j+1)(1− 2
p (1+ε))
0
kTg(f )k2
≍
≍
∞
∞
Xj=1
Xj=1
and finishes the proof.
p (1+ε)
e
1 − r(cid:19)1− 2
dr!
(cid:3)
5. Carleson measures for the Dirichlet space Dp
p−1
The statement in Theorem 5 follows directly by (4.3) and [19, Theorem 2.1]
with ω = vp/2. We next show that this result is sharp in a very strong sense.
For this purpose, the following lemma is needed.
Lemma 11. Let 2 < p < ∞, and let Φ : [0, 1) → (0, ∞) be a differentiable
increasing function such that
p
2 −1 → 0,
r → 1−,
Φ(r)
(cid:0)log e
1−r(cid:1)
m = − lim inf
r→1−
Φ′(r)
Φ(r)
(1 − r) log
e
1 − r
> 1 −
p
2
.
(5.1)
and
(5.2)
Then
Proof. By the Bernouilli-l'Hopital theorem,
Z 1
r
Φ(s) ds
(1 − s)(cid:0)log e
1−s(cid:1)
r→1− R 1
Φ(s) ds
(1−s)(log e
r
lim sup
1−r )
Φ(r)
1−r )
(log e
p
2 −1
.
p
2
Φ(r)
(cid:0)log e
1−r(cid:1)
,
p
2 −1
r ∈ (0, 1).
p
2
≤(cid:16)m +
p
2
− 1(cid:17)−1
∈ (0, ∞),
and the assertion follows.
(cid:3)
If Φc(r) =(cid:0)log e
1−r(cid:1)c and c > 0, then
e
Φ′
c(r)
Φc(r)
(1 − r) log
= c,
0 < r < 1,
1 − r
OPERATOR THEORETIC DIFFERENCES
17
and thus Φc satisfies both (5.1) and (5.2) if c < p
Φn(r) = logn
, n ∈ N, satisfies
expn(2)
1−r
2 − 1. Further, each function
Φ′
n(r)
Φn(r)
(1 − r) log
e
1 − r
→ 0,
r → 1,
and hence satisfies all hypotheses of the next result.
Proposition 12. Let 2 < p < ∞, and let Φ : [0, 1) → (1, ∞) be a differentiable
increasing unbounded function such that Φ′(r)
Φ(r) (1−r) is decreasing and (5.1) and
(5.2) are satisfied. Then there exists a positive Borel measure µ on D such that
(5.3)
sup
I⊂T
µ (S(I))
I(cid:16)log e
I(cid:17)−p/2+1
Φ(1 − I)
but µ is not a p-Carleson measure for Dp
p−1.
< ∞,
z ∈ D.
Proof. The radial measure
dµ(z) =
Φ(z) dA(z)
(1 − z)(cid:16)log e
1−z(cid:17)p/2 ,
satisfies (5.3) by Lemma 11. To see that µ is not a p-Carleson measure for
Dp
p−1, consider the lacunary series associated to Φ via (4.4). By Lemma 10,
f ∈ Dp
p−1 and
kf kp
0
Lp(µ) =Z 1
&Z 1
0
M p
p (r, f )Φ(r)
1−r(cid:1)p/2 r dr
(1 − r)(cid:0)log e
rΦ′(r)
Φ(r)
dr & lim
t→1−
log Φ(t) = ∞,
(cid:3)
which finishes the proof.
References
[1] A. Aleman, J. A. Cima, An integral operator on H p and Hardy's inequality, J. Anal.
Math. 85 (2001), 157 -- 176.
[2] A. Aleman, A. Siskakis, An integral operator on H p, Complex Variables 28 (1995),
149 -- 158.
[3] A. Aleman, A. Siskakis, Integration operators on Bergman spaces, Indiana Univ. Math.
J. 46 (1997), 337 -- 356.
[4] N. Arcozzi, R. Rochberg and E. Sawyer, Carleson measures for analytic Besov spaces,
Rev. Mat. Iberoamericana 18 (2002), 443 -- 510.
[5] A. Calder´on, Commutators of singular integral operators, Proc. Nat. Acad. Sci. 53,
(1965) 1092 -- 1099.
[6] L. Carleson, An interpolation problem for bounded analytic functions, Amer. J. Math.
80 (1958), 921 -- 930.
18
JOS ´E ´ANGEL PEL ´AEZ, F. P´EREZ-GONZ ´ALEZ, AND JOUNI R ATTY A
[7] L. Carleson, Interpolations by bounded functions and the corona problem, Ann. of
Math. 76 (1962), 547 -- 559.
[8] B. R. Choe, H. Koo and W. Smith, Composition operators acting on Sobolev holomor-
phic spaces, Trans. Amer. Math. Soc (2003), 355, n.7, 2829 -- 2855.
[9] P. Duren, Theory of H p Spaces, Academic Press, New York-London 1970.
[10] T. M. Flett, The dual of an inequality of Hardy and Littlewood and some related
inequalities, J. Math. Anal. Appl. 38 (1972), 756 -- 765.
[11] J. Garnett, Bounded analytic functions, Academic Press, New York, 1981.
[12] P. Galanopoulos, D. Girela, J. A. Pel´aez and A. Siskakis, Generalized Hilbert operators
on classical spaces of analytic functions, preprint (submitted).
[13] D. Girela and J. A. Pel´aez, Growth properties and sequences of zeros of analytic func-
tions in spaces of Dirichlet type, J. Australian Math. Soc. 80 (2006), 397 -- 418.
[14] D. Girela, M. Pavlovi´c and J. A. Pel´aez, Spaces of analytic functions of Hardy-Bloch
type, J. Anal. Math. 100 (2006), 53 -- 81.
[15] D. Girela and J. A. Pel´aez, Carleson measures for spaces of Dirichlet type, Integral
Equations and Operator Theory 55 (2006), n. 3, 415 -- 427.
[16] D. Girela and J. A. Pel´aez, Carleson measures, multipliers and integration operators
for spaces of Dirichlet type, J. Funct. Analysis 241 (2006), n. 1, 334 -- 358.
[17] J. E. Littlewood and R. E. A. C. Paley, Theorems on Fourier series and power series.
II, Proc. London Math. Soc. 42 (1936), 52 -- 89.
[18] D. H. Luecking, Embedding derivatives of Hardy spaces into Lebesgue spaces. Proc.
London Math. Soc. 63 (1991), 565 -- 619.
[19] J. A. Pel´aez and J. Rattya, Weighted Bergman spaces induced by rapidly increasing
weights, Memoirs Amer. Math. Soc. (to appear).
[20] C. Pommerenke, Schlichte funktionen und analytische funktionen von beschrankter mit-
tlerer oszillation, Comment. Math. Helv. 52 (1977), 591 -- 602.
[21] Z. Wu, Carleson measures and multipliers for Dirichlet spaces, J. Funct. Anal. 169
(1999), 148 -- 163.
[22] Zygmund A., Trigonometric Series, Cambridge University Press, London, 1959.
Departamento de Analisis Matematico, Universidad de Malaga, Campus de
Teatinos, 29071 Malaga, Spain
E-mail address: [email protected]
Departamento de An´alisis Matem´atico, Universidad de La Laguna, 38271 La
Laguna, Tenerife, Spain
E-mail address: [email protected]
University of Eastern Finland, P.O.Box 111, 80101 Joensuu, Finland
E-mail address: [email protected]
|
1604.06323 | 2 | 1604 | 2016-07-18T18:09:52 | Optimal constants for a mixed Littlewood type inequality | [
"math.FA"
] | For $p\in\lbrack2,\infty]$ a mixed Littlewood-type inequality asserts that there is a constant $C_{(m),p}\geq1$ such that \[ \left( \sum_{i_{1}=1}^{\infty}\left( \sum_{i_{2},...,i_{m}=1}^{\infty }|T(e_{i_{1}},...,e_{i_{m}})|^{2}\right) ^{\frac{1}{2}\frac{p}{p-1}}\right) ^{\frac{p-1}{p}}\leq C_{(m),p}\Vert T\Vert \] for all continuous real-valued $m$-linear forms on $\ell_{p}\times c_{0} \times\dots\times c_{0}$ (when $p=\infty$, $\ell_{p}$ is replaced by $c_{0})$. We prove that for $p>2.18006$ the optimal constants $C_{(m),p}$ are $\left( 2^{\frac{1}{2}-\frac{1}{p}}\right) ^{m-1}.$ When $p=\infty,$ we recover the best constants of the mixed $\left( \ell_{1},\ell_{2}\right) $-Littlewood inequality. | math.FA | math | OPTIMAL CONSTANTS FOR A MIXED LITTLEWOOD TYPE INEQUALITY
TONY NOGUEIRA, DANIEL N ´U NEZ-ALARC ´ON, AND DANIEL PELLEGRINO
Abstract. For p ∈ [2, ∞] a mixed Littlewood-type inequality asserts that there is a constant C(m),p ≥ 1 such that
∞(cid:88)
∞(cid:88)
i1=1
i2,...,im=1
T (ei1 , ..., eim )2
≤ C(m),p(cid:107)T(cid:107)
p−1
p
p
p−1
2
1
(cid:17)m−1
(cid:16)
for all continuous real-valued m-linear forms on (cid:96)p × c0 × · · · × c0 (when p = ∞, (cid:96)p is replaced by c0). We prove
. When p = ∞, we recover the best constants of
that for p > 2.18006 the optimal constants C(m),p are
the mixed ((cid:96)1, (cid:96)2)-Littlewood inequality.
2 − 1
2
p
1
The Hardy -- Littlewood inequality ([17], 1934) is a continuation of famous works of Littlewood ([18], 1930) and
1. Introduction
Bohnenblust and Hille ([9], 1931) and can be stated as follows:
• [17, Theorems 2 and 4] If p, q ≥ 2 are such that
1
p
1
2
then there is a constant Cp,q ≥ 1 such that
<
+
< 1
1
q
pq−q−p
pq
A(ej, ek)
pq
pq−p−q
≤ Cp,q (cid:107)A(cid:107)
6
1
0
2
l
u
J
8
1
]
.
A
F
h
t
a
m
[
2
v
3
2
3
6
0
.
4
0
6
1
:
v
i
X
r
a
(1)
(2)
(3)
∞(cid:88)
j,k=1
∞(cid:88)
for all continuous bilinear forms A : (cid:96)p × (cid:96)q → R (or C). Moreover the exponent
• [17, Theorems 1 and 4] If p, q ≥ 2 are such that
pq
pq−p−q is optimal.
then there is a constant Cp,q ≥ 1 such that
1
p
+
1
q
≤ 1
2
3pq−2p−2q
4pq
A(ej, ek)
4pq
3pq−2p−2q
≤ Cp,q (cid:107)A(cid:107)
for all continuous bilinear forms A : (cid:96)p × (cid:96)q → R (or C). Moreover the exponent
4pq
3pq−2p−2q is optimal.
j,k=1
Above and henceforth, as usual in this field, when p and/or q is infinity, we consider c0 instead of (cid:96)p and/or (cid:96)q.
As mentioned in [20, Theorem 1] an unified version of the above two results of Hardy and Littlewood asserts
that there is a constant Cp,q ≥ 1 such that
∞(cid:88)
(cid:32) ∞(cid:88)
2 1
(cid:33) λ
λ
A(ej, ek)2
≤ Cp,q (cid:107)A(cid:107)
pq−p−q , for all continuous bilinear forms A : (cid:96)p × (cid:96)q → R (in fact, in [20, Theorem 1] just the complex
with λ = pq
case is considered, but for a general approach including the real case we refer to [11]; moreover the exponents are
optimal). The recent years witnessed an increasing interest in the study of summability of multilinear operators
j=1
k=1
2010 Mathematics Subject Classification. 11Y60, 46G25.
Key words and phrases. Absolutely summing operators; Hardy -- Littlewood inequality; Bohnenblust -- Hille inequality; Multiple sum-
ming operators.
T. Nogueira was supported by Capes, D. N´unez-Alarc´on was supported by Capes, Grant 000785/2015-06, and D. Pellegrino was
supported by CNPq.
1
2
T. NOGUEIRA, D. N ´U NEZ, AND D. PELLEGRINO
(see, for instance, [10, 23, 24]) and in estimating constants of the multilinear and polynomial Hardy -- Littlewood
and related inequalities (see [2, 3, 4, 6, 14, 15, 26]). Perhaps the main motivations are potential applications (see,
for instance, [19] for applications of the real-valued case of the estimates of the Bohnenblust -- Hille inequality and
[7, 12] for applications of the complex-valued case).
One of the most for reaching generalizations of the Hardy -- Littlewood inequality is the following theorem (see
(cid:21)
also [25]):
Theorem 1.1. (See Albuquerque, Araujo, N´unez, Pellegrino and Rueda [1]) Let m ≥ 2 be a positive integer,
(cid:17) , 2
1 ≤ k ≤ m and n1, . . . , nk ≥ 1 be positive integers such that n1 + ··· + nk = m. If q1, ..., qk ∈
and 0 ≤ 1
2 , then the following assertions are equivalent:
+ ··· + 1
1−(cid:16) 1
1
+···+ 1
≤ 1
(cid:20)
pm
p1
pm
p1
(a) There is a constant Ck = C(k, p1, ..., pm, q1, ..., qk) such that
∞(cid:88)
i1=1
...
(cid:32) ∞(cid:88)
ik=1
T (en1
i1
, ..., enk
ik
)qk
...
for all continuous m-linear forms T : (cid:96)p1 × ··· × (cid:96)pm → R.
(b) The numbers q1, ..., qk satisfy
1
q1
+ ··· +
1
qk
≤ k + 1
2
−
+ ··· +
1
pm
q1
q2 1
q1
(cid:19)
≤ Ck(cid:107)T(cid:107)
.
(cid:33) qk−1
qk
(cid:18) 1
p1
Above, the notation enj
j
represents the nj-tuple (ej, ..., ej). The optimal constants of the previous inequalities
are essentially unknown. Recent works have shown that in general these constants have a sublinear growth (see
[5, 6, 7], and references therein). One of the few cases in which the optimal constants are known for all m is the
case of mixed ((cid:96)1, (cid:96)2)-Littlewood inequality (see [21]):
• The optimal constants C(m),∞ satisfying
∞(cid:88)
∞(cid:88)
1
2
T (ei1 , ..., eim)2
≤ C(m),∞(cid:107)T(cid:107)
i1=1
i2,...,im=1
for all continuous real m-linear forms T : c0 × ··· × c0 → R are 2
m−1
2
.
From now on p0 ≈ 1.84742 is the unique real number satisfying
√
π
2
=
Γ
2
(cid:18) p0 + 1
(cid:19)
.
(5)
(4)
(6)
Our main result provides the optimal constants of a Hardy -- Littlewood-type inequality that encompasses (4); as
far as we know this is the first time in which a Hardy -- Littlewood type inequality (except for the case of mixed
((cid:96)1, (cid:96)2)-Littlewood inequality) is proved to have optimal constants with exponential growth:
Theorem 1.2. Let m ≥ 2 be a positive integer and p ≥ p0
p0−1 ≈ 2.18006. The optimal constant C(m),p such that
∞(cid:88)
∞(cid:88)
i1=1
i2,...,im=1
T (ei1, ..., eim)2
≤ C(m),p(cid:107)T(cid:107),
2
p
p
p−1 p−1
1
(cid:17)m−1
(cid:16)
1
2− 1
p
2
.
for all continuous m-linear forms T : (cid:96)p × c0 × ··· × c0 → R is
Note that the above Hardy -- Littlewood type inequality holds for p ≥ 2 (see Theorem 1.1). When p = 2 it is
simple to prove that the optimal constants are C(m),p = 1. As a consequence of the arguments of our proof of
Theorem 1.2 we remark that for 2 < p < p0
p0−1 the optimal constants still have exponential growth; so an eventual
decrease on the order of the growth when p → 2 does not happen. Moreover, for 2 < p < p0
p0−1 ≈ 2.18006, the
difference between the bases in the exponential upper and lower estimates of C(m),p is not bigger than 4 · 10−4 (see
the figures 1 and 2).
OPTIMAL CONSTANTS FOR A MIXED LITTLEWOOD TYPE INEQUALITY
3
In the final section we also provide upper and lower estimates for the sharp constants Cp,∞ of the real case of
(2), showing that
2
1
2− 1
p ≤ Cp,∞ ≤ 2
1
2− 1
2p
for all p ≥ p0
the Littlewood's 4/3 inequality obtained in [15].
p0−1 ≈ 2.18006. This result recovers, in particular, the optimality of the constant
√
2 of the real case of
2. The proof of Theorem 1.2
The Khinchine inequality (see [13]) asserts that, for any 0 < q < ∞, there are positive constants Aq, Bq such
that regardless of the scalar sequence (aj)n
j=1 we have
Aq
aj2
≤
ajrj(t)
dt
≤ Bq
aj2
1
q
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)q
n(cid:88)
j=1
1
2
,
n(cid:88)
j=1
1
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n(cid:88)
(cid:33) 1
(cid:1)
j=1
q
(cid:90) 1
Γ(cid:0) q+1
0
√
2
π
(cid:32)
where rj are the Rademacher functions. For real scalars, U. Haagerup [16] proved that if p0 is the number defined
in (5) then
√
2
Aq =
and
for 1.84742 ≈ p0 < q < 2
,
Let T : (cid:96)p × c0 × ··· × c0 → R be a continuous m-linear form. By the Khinchine inequality for multiple sums (see
[22]) we know that
Aq = 2
1
2− 1
q ,
for 1 ≤ q ≤ p0 ≈ 1.84742.
∞(cid:88)
∞(cid:88)
i1=1
i2,...,im=1
≤ (A−1
p−1
p
)m−1
= (A−1
p−1
p
)m−1
≤ (A−1
p
p−1
≤ (A−1
p
p−1
)m−1
)m−1
T (ei1 , ..., eim )2
[0,1]m−1
(cid:90)
i1=1
∞(cid:88)
(cid:90)
(cid:90)
[0,1]m−1
[0,1]m−1
2
p
p
i2,...,im
p−1 p−1
1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∞(cid:88)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T
(cid:32)
∞(cid:88)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)T
(cid:32)
∞(cid:88)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)T
(cid:32)
∞(cid:88)
ei1,
i1=1
i2=1
i2=1
·,
·,
i2=1
ri2(t2)··· rim (tm)T (ei1 , ..., eim )
∞(cid:88)
∞(cid:88)
ri2 (t2)ei2, ...,
rim (tm)eim
ri2(t2)ei2, ...,
im=1
rim (tm)eim
∞(cid:88)
∞(cid:88)
im=1
im=1
ri2(t2)ei2, ...,
rim(tm)eim
p
p−1
p−1
p
p
p−1
dt2 ··· dtm
(cid:33)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) p
p−1
dt2 ··· dtm
p−1
p
p−1
dt2 ··· dtm
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:33)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p
(cid:33)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
sup
t2,...,tm∈[0,1]
2− 1
1
p
p−1
)m−1(cid:107)T(cid:107) = (2
≤ (A−1
p0−1 ≈ 2.18006. Now let us show that (2
p )m−1(cid:107)T(cid:107)
whenever p ≥ p0
and T x2
2
: (cid:96)2
p → R be given by
(7)
and
for each x2 ∈ (cid:96)2∞. Observe that
(8)
T2 (x1, x2) =(cid:0)x1
2 + x2
2
(cid:1) x1
1 +(cid:0)x1
2 − x2
2
(cid:1) x2
1,
T x2
2
(x1) = T2 (x1, x2) ,
(cid:110)(cid:107)T x2
(cid:111)
2 (cid:107) : (cid:107)x2(cid:107)(cid:96)2∞ = 1
.
(cid:107)T2(cid:107) = sup
1
2− 1
p )m−1 is the best possible constant. Let T2 : (cid:96)2
p × (cid:96)2∞ → R
4
T. NOGUEIRA, D. N ´U NEZ, AND D. PELLEGRINO
Let us estimate (8). Since ((cid:96)p)
(9)
∗
= (cid:96) p
p−1
(cid:107)T2(cid:107) = sup
(x1) : (cid:107)x2(cid:107)(cid:96)2∞ = 1
(cid:12)(cid:12) : (cid:107)x2(cid:107)(cid:96)2∞ = 1
(cid:1) x2
(cid:1) x1
1 +(cid:0)x1
2 − x2
(cid:111)
2, 0, 0, ...(cid:1)(cid:13)(cid:13) p
(cid:27)
(cid:17) 1
: (cid:107)x2(cid:107)(cid:96)2∞ = 1
p−1
p−1 : x ∈ [−1, 1]
2 − x2
= 2.
2
1
p
= sup
= sup
= sup
= sup
p
p
2
T x2
2 + x2
2
x1∈B(cid:96)2
x1∈B(cid:96)2
, we have
(cid:111)
2 (cid:107) : (cid:107)x2(cid:107)(cid:96)2∞ = 1
(cid:110)(cid:107)T x2
sup
sup
(cid:12)(cid:12)(cid:0)x1
(cid:110)(cid:13)(cid:13)(cid:0)x1
(cid:26)(cid:16)1 + x p
(cid:26)(cid:16)1 + x1 + 1 − x1(cid:17)1
(cid:27)
(cid:17) p−1
2, x1
p−1 + 1 − x p
p−1
2 + x2
; x ∈ [−1, 1]
p
sup
for p ∈ [2,∞) , we have (cid:107)·(cid:107)(cid:96) p
p−1
≤ sup
. Therefore, for p ∈ [2,∞) we have
(cid:27)
= 2,
; x ∈ [−1, 1]
≤ (cid:107)·(cid:107)(cid:96)1
(cid:26)(cid:16)1 + x1 + 1 − x1(cid:17)1
≥(cid:16)1 + 1 p
p−1 + 1 − 1 p
p−1
(cid:27)
; x ∈ [−1, 1]
= 2.
(cid:17) p−1
p
= 2.
In order to verify the last equality, note that since
by the norm inclusion (cid:96)1 ⊂ (cid:96) p
p−1
(cid:26)(cid:16)1 + x p
sup
p−1 + 1 − x p
p−1
On the other hand, it is obvious that
(cid:26)(cid:16)1 + x p
(cid:16)
1
2− 1
p
sup
p−1 + 1 − x p
p−1
(cid:17)m−1
(cid:17) p−1
p
(cid:27)
; x ∈ [−1, 1]
In order to show that
2
is the best possible constant satisfying (6), let T2 be as in (7) and define for
all m ≥ 3 the m-linear operator Tm : (cid:96)2m−1
m + x2
Tm(x1, ...., xm) =(x1
p
× (cid:96)2m−1∞ × ··· × (cid:96)2m−1∞ → R by
m)Tm−1(x1, ..., xm−1)
+ (x1
m − x2
m)Tm−1(S2m−2
p
(x1), S2m−2
0
(x2), S2m−3
0
(x3)..., S2
0 (xm−1)),
where x1 ∈ (cid:96)2m−1
induction on m ≥ 2 we shall show that
p
, xk ∈ (cid:96)2m−1∞ for all k = 2, ..., m, and Sp : (cid:96)p → (cid:96)p and S0 : c0 → c0 are the backward shifts. By
(cid:107)Tm(cid:107) = 2m−1.
The case m = 2 is already done in (9). Let us suppose that (cid:107)Tm−1(cid:107) = 2(m−1)−1. Therefore,
Tm(x1, . . . , xm) ≤x1
mTm−1(x1, . . . , xm−1)
m + x2
+ x1
m − x2
≤2m−2[x1
mTm−1(S2m−2
m(cid:107)x1(cid:107)(cid:96)2m−1
p
m + x2
p
(x1), S2m−2
···(cid:107)xm−1(cid:107)(cid:96)2m−1∞
0
(x2), S2m−3
0
(x3)..., S2
0 (xm−1))
+ x1
m − x2
≤2m−2[x1
m + x2
=2m−1(cid:107)x1(cid:107)(cid:96)2m−1
≤2m−1(cid:107)x1(cid:107)(cid:96)2m−1
p
p
p
(x1)(cid:107)(cid:96)2m−1
m − x2
m(cid:107)S2m−2
m + x1
···(cid:107)xm−1(cid:107)(cid:96)2m−1∞
···(cid:107)xm(cid:107)(cid:96)2m−1∞
.
0
(cid:107)S2m−2
m](cid:107)x1(cid:107)(cid:96)2m−1
max{x1
(cid:107)S2m−3
(x2)(cid:107)(cid:96)2m−1∞
···(cid:107)xm−1(cid:107)(cid:96)2m−1∞
m,x2
m}
0
p
(x3)(cid:107)(cid:96)2m−1∞
···(cid:107)S2
0 (xm−1)(cid:107)(cid:96)2m−1∞
]
We thus have (cid:107)Tm(cid:107) ≤ 2m−1. Now consider am = e1 + e2 and note that
p
(cid:110)Tm (x1, . . . , xm−1, am) : x1 ∈ B(cid:96)2m−1
p
(cid:107)Tm(cid:107) ≥ sup
, x2 ∈ B(cid:96)2m−1∞
, ..., xm−1 ∈ B(cid:96)2m−1∞
(cid:111)
= 2(cid:107)Tm−1(cid:107) = 2m−1
and hence (cid:107)Tm(cid:107) = 2m−1.
Since
the proof is done.
OPTIMAL CONSTANTS FOR A MIXED LITTLEWOOD TYPE INEQUALITY
5
(cid:18)(cid:80)
(cid:16)(cid:80)
i2,...,im
i1
Tm(ei1,...,eim)2(cid:17) 1
2
(cid:107)Tm(cid:107)
p−1(cid:19) p−1
p
p
(cid:18)
=
2
− 1
p
1
2
(cid:19)m−1
,
3. Final remarks
The same argument used in the proof of Theorem 1.2 shows that for 2 < p < p0
p0−1 ≈ 2.18006 the optimal constants
also have exponential growth; curiously, for p = 2 the situation is quite different and the optimal constants are 1.
In fact, note that the second part of the proof (the optimality proof) holds for all p ≥ 2. Moreover, the first part
p0−1 ≈ 2.18006, the following
of the proof gives us the estimate C(m),p ≤(cid:16)
. We thus have, for 2 ≤ p < p0
A−1
p−1
(cid:17)m−1
1√
(cid:17)m−1 ≤ C(m),p ≤
p
2
(cid:16)
1
2− 1
p
2
Γ
(cid:16) 2p−1
2p−2
√
π
(cid:17)
1−p
p
m−1
.
inequalities
Figure 1. Plots of the functions A−1
x−1
x
Figure 2. Plot of the function
(cid:16)
A−1
x−1
x
and 2 1
2− 1
x , for x ∈ [2, p0
p0−1 ]
(cid:17)
, for x ∈ [2, p0
p0−1 ]
− 2 1
2− 1
x
6
T. NOGUEIRA, D. N ´U NEZ, AND D. PELLEGRINO
For p ≥ 2, we know that
(10)
∞(cid:88)
j=1
(cid:32) ∞(cid:88)
λ 2 1
(cid:33) 1
2
√
≤
2(cid:107)A(cid:107)
A(ej, ek)λ
p−1 , for all continuous bilinear forms A : (cid:96)p× c0 → R (see, for instance, [2, Theorem 1.2 and Remark 5.1]).
with λ = p
By interpolating (10) and the result of Theorem 1.2 for m = 2 in the sense of [2] or using the Holder inequality for
mixed sums ([8]) we obtain, for p ≥ p0
p0−1 ≈ 2.18006,
2p (cid:107)A(cid:107)(cid:17)1/2
p−2
2
1
2− 1
2p (cid:107)A(cid:107) .
= 2
∞(cid:88)
j,k=1
A(ej, ek) 4p
3p−2
k=1
4p
3p−2
(cid:32) 2(cid:80)
≤(cid:16)√
2(cid:107)A(cid:107)(cid:17)1/2(cid:16)
(cid:33) 3p−2
4p
T2(ej, ek) 4p
3p−2
(cid:107)T2(cid:107)
1
2− 1
p ≤ Cp,∞ ≤ 2
1
2− 1
2p .
2
Using the approach of the previous section we obtain the lower estimate
Cp,∞ ≥
j,k=1
and thus
3p−2
4p
4
2
=
= 2
1
2− 1
p
When p = ∞ we recover the well known optimal estimate of the famous Littlewood's 4/3 that can be found in [15].
Acknowledgement. The authors are indebted to the two anonymous referees for their important contributions
to the final version of this paper.
References
[1] N. Albuquerque, G. Ara´ujo, D. N´unez-Alarc´on, D. Pellegrino, and P. Rueda, Bohnenblust -- Hille and Hardy -- Littlewood inequalities
by blocks, arXiv:1409.6769v6 [math.FA].
[2] N. Albuquerque, F. Bayart, D. Pellegrino, J. Seoane-Sepulveda, Sharp generalizations of the multilinear Bohnenblust-Hille inequal-
ity, J. Funct. Anal. 266 (2014), no. 6, 3726 -- 3740.
[3] N. Albuquerque, F. Bayart, D. Pellegrino and J. B. Seoane-Sep´ulveda, Optimal Hardy-Littlewood type inequalities for polynomials
and multilinear operators, Isr. J. Math. 211 (2016), 197 -- 220.
[4] G. Ara´ujo, D. Pellegrino, Lower bounds for the constants of the Hardy-Littlewood inequalities, Linear Algebra Appl. 463 (2014),
10 -- 15.
[5] G. Ara´ujo, D. Pellegrino, On the constants of the Bohnenblust -- Hille and Hardy -- Littlewood inequalities, to appear in Bull. Braz.
Math. Soc.
[6] G. Ara´ujo, D. Pellegrino, D. Diniz P. Silva e Silva, On the upper bounds for the constants of the Hardy-Littlewood inequality, J.
[7] F. Bayart, D. Pellegrino and J. B. Seoane-Sep´ulveda, The Bohr radius of the n -- dimensional polydisk is equivalent to(cid:112)(log n)/n,
Funct. Anal. 267 (2014), no. 6, 1878 -- 1888.
Adv. Math. 264 (2014), 726 -- 746.
[8] A. Benedek, R. Panzone, The space Lp, with mixed norm, Duke Math. J. 28 1961 301 -- 324.
[9] H. F. Bohnenblust, E. Hille, On the absolute convergence of Dirichlet series, Ann. of Math. 32 (1931), 600 -- 622.
[10] G. Botelho, J. Santos, A Pietsch domination theorem for ((cid:96)s
[11] W. Cavalcante, D. N´unez-Alarc´on, Remarks on the Hardy -- Littlewood inequality for m-homogeneous polynomials and m-linear
p, (cid:96)p)-summing operators, Arch. Math. (Basel) 104 (2015), 47 -- 52.
forms, to appear in Quaest. Math.
[12] A. Defant, L. Frerick, J. Ortega-Cerd´a, M. Ounaıes, K. Seip, The Bohnenblust-Hille inequality for homogeneous polynomials is
hypercontractive, Ann. of Math. (2), 174 (2011), 485 -- 497.
[13] J. Diestel, H. Jarchow, A. Tonge, Absolutely summing operators, Cambridge University Press, Cambridge, 1995.
[14] V. Dimant and P. Sevilla -- Peris, Summation of coefficients of polynomials on (cid:96)p spaces, Publ. Mat. 60 (2016), 289 -- 310.
[15] D. Diniz, G. Munoz-Fern´andez, D. Pellegrino, J. Seoane-Sep´ulveda, Lower bounds for the constants in the Bohnenblust -- Hille
inequality: the case of real scalars, Proc. Amer. math. Soc. 142 (2014), 575 -- 580.
[16] U. Haagerup, The best constants in the Khinchine inequality, Studia Math. 70 (1982) 231 -- 283.
[17] G. Hardy and J. E. Littlewood, Bilinear forms bounded in space [p, q], Quart. J. Math. 5 (1934), 241 -- 254.
[18] J. E. Littlewood, On bounded bilinear forms in an infinite number of variables, Quart. J. Math. 1 (1930), 164 -- 174.
[19] A. Montanaro, Some applications of hypercontractive inequalities in quantum information theory. J. Math. Phys. 53 (2012), no.
12, 122206, 15 pp.
[20] B. Osikiewicz, A. Tonge, An interpolation approach to Hardy -- Littlewood inequalities for norms of operators on sequence spaces,
Linear Algebra Appl. 331 (2001), 1 -- 9.
[21] D. Pellegrino, The optimal constants of the mixed ((cid:96)1, (cid:96)2)-Littlewood inequality, J. Number Theory 160 (2016), 11 -- 18.
[22] D. Popa, Multiple Rademacher means and their applications, J. Math. Anal. Appl. 386 (2012), 699 -- 708.
[23] D. Popa, Multiple summing operators on (cid:96)p-spaces, Studia Math. 225 (2014), no. 1, 9 -- 28.
[24] P. Rueda, E.A. S´anchez-P´erez, Factorization of p-dominated polynomials through Lp-spaces, Michigan Math. J. 63 (2014), no. 2,
345 -- 353.
OPTIMAL CONSTANTS FOR A MIXED LITTLEWOOD TYPE INEQUALITY
7
[25] J. Santos and T. Velanga, A note on the Bohnenblust -- Hille inequality for multilinear forms, arXiv:1604.00040v2 [math.FA].
[26] D. M. Serrano-Rodr´ıguez, Improving the closed formula for subpolynomial constants in the multilinear Bohnenblust -- Hille inequal-
ities, Linear Algebra Appl. 438 (2013) 3124 -- 3138.
(T. Nogueira) Departamento de Matem´atica,
E-mail address: [email protected]
Universidade Federal da Para´ıba,
58.051-900 - Joao Pessoa, Brazil.
(D. N´unez-Alarc´on) Departamento de Matem´atica,
& Department of Mathematical Sciences,
Brazil,
Universidade Federal de Pernambuco,
50.740-560 - Recife,
Kent State University,
Kent, Ohio 44242, USA
E-mail address: [email protected]
(D. Pellegrino) Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900 - Joao Pessoa, Brazil.
E-mail address: [email protected] and [email protected]
|
1609.01049 | 2 | 1609 | 2017-01-01T15:43:38 | Noncommutative probability of type D | [
"math.FA"
] | We construct a deformed Fock space and a Brownian motion coming from Coxeter groups of type D. The construction is analogous to that of the $q$-Fock space (of type A) and the $(\alpha,q)$-Fock space (of type B). | math.FA | math |
Noncommutative probability of type D
Marek Bożejko∗, Wiktor Ejsmont†and Takahiro Hasebe‡
Abstract
We construct a deformed Fock space and a Brownian motion coming from Coxeter
groups of type D. The construction is analogous to that of the q-Fock space (of type A)
and the (α, q)-Fock space (of type B).
1
Introduction
Several deformations of boson, fermion and full Fock spaces and Brownian motion have been
proposed so far. Bożejko and Speicher used Coxeter groups of type A to construct a deformed
Fock space and Brownian motion [BSp91, BSp94]. Bożejko and Speicher also considered general
(mainly finite) Coxeter groups in [BSp94]. We followed this idea in [BEH15] and constructed
an (α, q)-Brownian motion on an (α, q)-Fock space using Coxeter groups of type B. The com-
mutation relation satisfied by the creation and annihilation operators on the (α, q)-Fock space
reads
bα,q(x)b∗α,q(y) − qb∗α,q(y)bα,q(x) = hx, yiI + αhx, yi q2N
(1.1)
where α, q ∈ [−1, 1], N is the number operator, x, y are vectors in an underlying Hilbert space
H and y 7→ y is a selfadjoint involution on H. The orthogonal polynomials associated to
the distribution of the (α, q)-Gaussian operator bα,q(x) + b∗α,q(x) are called q-Meixner-Pollaczek
polynomials satisfying the recurrence relation
tP (α,q)
n
(t) = P (α,q)
n+1 (t) + [n]q(1 + αqn−1)P (α,q)
n−1 (t),
n = 0, 1, 2, . . .
(1.2)
(t) = 0, P (α,q)
where P (α,q)
(t) = 1. When α = 0 then we get q-Hermite orthogonal polynomials
and when q = 0 then we get the orthogonal polynomials associated to a symmetric free Meixner
distribution [BB06]. The moments of the Brownian motion of type B are given by
−1
0
hΩ, (bα,q(x1) + b∗α,q(x1))· · · (bα,q(x2n) + b∗α,q(x2n))Ωiα,q
αnp(π,f )qcr(π)+2cnp(π,f ) Y{i<j}∈π
= X(π,f )∈PB
2 (2n)
f ({i<j})=1
hxi, xji Y{i<j}∈π
f ({i<j})=−1
hxi, xji,
(1.3)
where PB
2 (2n) is the set of pair partitions of type B, np(π, f ) is the number of negative pairs,
cr(π) denotes the number of the crossings of π and cnp(π, f ) is the number of covered negative
pairs of π. For further information the reader is referred to [BEH15] and the references therein.
∗Instytut Matematyczny, Polska Akademia Nauk, Ulica Śniadeckich 8, 00-956 Warszawa, Poland. Email:
[email protected]
†Department of Mathematics and Cybernetics, Wrocław University of Economics, ul. Komandorska 118/12
53-345 Wrocław, Poland. Email: [email protected]
‡Department of Mathematics, Hokkaido University, North 10 West 8, Kita-ku, Sapporo 060-0810 Japan.
Email: [email protected]
1
The goal of this paper is to introduce a new deformation of the full Fock space, creation
and annihilation operators and Brownian motion in terms of Coxeter groups of type D. Our
strategy is to replace the Coxeter groups of type A or B in the previous works [BSp91, BEH15]
by Coxeter groups of type D. Given this background, it seems natural to try to extend this
theory to Coxeter groups of type C. But it is known that Coxeter groups of type B and type
C are isomorphic, and one can check that for Coxeter groups of type C we get the orthogonal
polynomials (1.2) (i.e. we get the same probability measure as in the type B). This also follows
ℓ denotes the length function on the Coxeter groups.
from the fact thatPσ∈B(n) qℓ(σ) =Pσ∈C(n) qℓ(σ) (see Carter [Car89, Proposition 10.2.5]), where
The plan of the paper is following. First we define a q-Fock space, a creation operator
d∗q(x) and an annihilation operator dq(x) of type D in Section 2. Next, natural properties
of creation and annihilation operators, including norm estimates and the commutation rela-
tions, are presented in Section 3. The probability distribution of a Brownian motion of type
D, Gq(x) = dq(x) + d∗q(x), x ∈ H, is studied in the Section 4.1. The associated orthogonal
polynomials satisfy the recurrence relation
n (t) = P (q)
tP (q)
P (q)
0 (t) = 1, P (q)
n+1(t) + [n]q(1 + qn−1)P (q)
1 (t) = t, P (q)
2 (t) = t2 − 1,
n−1(t),
n = 2, 3, 4, . . . ,
(1.4)
n
(t), but they are different since P (α,q)
which is not known in the literature to the authors' knowledge. This polynomial P (q)
n (t) looks
like P (1,q)
(t) = t2 − (1 + α) from (1.2). This is because
the first Jacobi parameter for P (q)
(t). This difference of first
Jacobi parameter comes from the fact that the Coxeter group of type D can be realized as a
subgroup of the Coxeter group of type B with index 2. The main theorem is placed in Section
4.4 where we show a Wick formula of type D
n (t) is different from that for P (1,q)
2
n
hΩ, Gq(x1)· · · Gq(x2n)Ωiq = X(π,f )∈PD
2 (2n)
qcr(π)+2cnp(π,f ) Y{i<j}∈π
f ({i<j})=1
hxi, xji Y{i<j}∈π
f ({i<j})=−1
hxi, xji.
(1.5)
It is described by pair partitions of type D which are introduced and studied in Sections 4.2
and 4.3. It turns out that our Wick formula of type D generalizes the t-transformed classical
Wick formula for t = 2 (q = 1) as well as the free Wick formula (q = 0). Using this formula we
show that the vacuum vector is not tracial with respect to the von Neumann algebra generated
by our Brownian motion Gq(x), where x runs over a real Hilbert space.
Our deformed Fock space (both of type B and type D) raises natural interesting questions.
Studying the von Neumann algebra generated by the Brownian motions of type D is one
direction. In the case of type A, the q-deformed von Neumann algebra shares many properties
with the free group factor (e.g. factoriality [Ric05]), and it is even isomorphic to the free group
factor for sufficiently small q [GS14]. The von Neumann algebra of type B contains the so-
called t-deformed von Neumann algebra, which is well investigated by Wysoczanski [Wys06]
and Ricard [Ric06]. For further information see references in [BEH15]. In our von Neumann
algebra of type D, the vacuum state is not a trace (Proposition 4.18). The first basic question
would be asking if it is cyclic separating. Factoriality of the von Neumann algebra is also a
natural question.
Another possible future direction is to search for a connection between our pair partitions
of type D and noncrossing partitions of type D, the latter of which has been developed in
the literature, e.g. in [Rei97, AR04]. Note that our pair partitions of type D are related to
Coxeter groups of type D in a natural way (see Remark 4.6). Yet another direction is classical
probability.
It is known that q-Brownian motion of type A has a classical Markov process
realization [BKS97], and its probabilistic properties and extensions have been studied by several
2
authors (see e.g. [BW05, BW14]). It is natural to ask if such a realization also exists in the
type B and type D cases.
Combining [BSp91, BEH15] and the present paper, we have constructed deformations of a
full Fock space by the natural three families of finite Coxeter groups, of type A,B and D (type
C is isomorphic to type B as already mentioned). We wonder if a similar construction exists
for infinite Coxeter groups, e.g. affine Coxeter groups.
The above questions are also listed in the end of this paper as open problems, together with
some other problems.
2 Type D deformation of full Fock space
2.1 Full Fock space
Let H be a complex Hilbert space with inner product h·,·i linear on the right component and
anti-linear on the left. Let Ffin(H) be the algebraic full Fock space over H
Ffin(H) :=
H⊗n
(2.1)
∞Mn=0
with convention that H⊗0 = CΩ is a one-dimensional normed space along a unit vector Ω. Note
that elements of Ffin(H) are finite linear combinations of the elements from H⊗n, n ∈ N ∪ {0}
and we do not take the completion. We equip Ffin(H) with the inner product
hx1 ⊗ · · · ⊗ xm, y1 ⊗ · · · ⊗ yni0 = δm,n
hΩ, y1 ⊗ · · · ⊗ yni0 = hx1 ⊗ · · · ⊗ xm, Ωi0 = 0,
hΩ, Ωi0 = 1,
hxi, yii,
nYi=1
(2.2)
(2.3)
(2.4)
where m, n ≥ 1 and xi, yi ∈ H.
by [Voi85]
For x ∈ H the free left creation and annihilation operators l∗(x), l(x) on Ffin(H) are defined
l∗(x)(x1 ⊗ · · · ⊗ xn) := x ⊗ x1 ⊗ · · · ⊗ xn,
l∗(x)Ω = x,
l(x)(x1 ⊗ · · · ⊗ xn) := hx, x1i x2 ⊗ · · · ⊗ xn,
l(x)x1 := hx, x1i Ω,
l(x)Ω = 0.
(2.5)
(2.6)
(2.7)
(2.8)
(2.9)
It then holds that l∗(x)∗ = l(x) and l∗ : H → B(Ffin(H)) is linear, but l : H → B(Ffin(H)) is
anti-linear.
n ≥ 2,
n ≥ 1
2.2 Coxeter groups of type D
Let S(±n) be the group of all permutations of the 2n numbers ±1,· · · ,±n. The Coxeter group
of type B, denoted by B(n), is the set of permutations σ in S(±n) such that σ(−k) = −σ(k), k =
1, . . . , n. The Coxeter group B(n) is a subgroup of S(±n) generated by {π1, . . . , πn−1, πn},
where πn = (n,−n), πi = (i, i+1)(−i,−i−1), i = 1, . . . , n−1. Our strategy is to define a Coxeter
group of type D(n) as a subgroup of the Coxeter group B(n) of index 2. This is well studied
in the literature (see [Bou02]). First we define D(1) = {e} ⊂ B(1) = grp{π1}. Next we define,
Note that D(2) is isomorphic to Z2 ⊕ Z2. We define D(3) by the Coxeter-Dynkin diagram
as subgroup of B(2) with index 2, D(2) = grp{π1,bπ1} ⊂ B(2), where bπ1 = π2π1π2 = (1,−2).
3
π1
bπ2
π2
where bπ2 = π3π2π3 so that the group D(3) is isomorphic to the permutation group S(4). We
define D(n) for n ≥ 4 as a subgroup of B(n) generated by π1, . . . , πn−1,bπn−1 where bπn−1 =
(πn−2bπn−1)3 = (πiπi+1)3 = e, 1 ≤ i < n − 1 and (πiπj)2 = (πkbπn−1)2 = e if 1 ≤ i, j, k ≤
πnπn−1πn. The Coxeter-Dynkin diagram for D(n), n ≥ 4 is described in Fig. 1, which says
that the generators satisfy the generalized braid relations π2
n−1 = e, 1 ≤ i ≤ n − 1,
n − 1,i − j ≥ 2, k 6= n − 2. Note that {πi i = 1, . . . , n − 1} generates the symmetric group
S(n).
i = bπ2
π1
π2
. . .
πn−2
bπn−1
πn−1
Fig. 1: Coxeter-Dynkin diagram for D(n).
2.3 Fock space of type D
We deform the full Fock space Ffin(H) and creation l∗(x) and annihilation operators l(x) on it.
A selfadjoint involution on H is a selfadjoint linear bounded operator on H such that the
double application of it becomes the identity operator (see Example 2.2 below). Suppose that
x 7→ x, x ∈ H is a selfadjoint involution. We then define an action of D(n) on H⊗n by
πi(x1 ⊗ · · · ⊗ xn) = x1 ⊗ · · · ⊗ xi−1 ⊗ xi+1 ⊗ xi ⊗ xi+2 ⊗ · · · ⊗ xn,
bπn−1(x1 ⊗ · · · ⊗ xn) = x1 ⊗ x2 ⊗ · · · ⊗ xn−2 ⊗ xn ⊗ xn−1,
action of (the left version of) πn defined in [BEH15] is
Remark 2.1. The action ofbπn−1 is defined in the above way since bπn−1 = πnπn−1πn and the
n ≥ 2,
n ≥ 2.
(2.10)
(2.11)
πn(x1 ⊗ · · · ⊗ xn) = x1 ⊗ x2 ⊗ · · · ⊗ xn−1 ⊗ xn,
n ≥ 1.
(2.12)
A second way to define a subgroup of B(2) with index 2 is grp{π2π1π2π1}, but in this situation
we have π2π1π2π1(x1 ⊗ x2) = x1 ⊗ x2, which is not compatible with D(n) for n ≥ 3.
coincide.
Example 2.2. (1) The identity involution x 7→ x on H. Then the actions of bπn−1 and πn−1
(2) If H is spanned by an orthonormal basis {ei}i∈{±1,...,±n} (or {ei}i∈Z\{0} in the infinite-
dimensional case), then the map
extends to an involution on H.
ei = e−i,
i ∈ {±1, . . . ,±n}
4
Let ℓ be the length function on the Coxeter groups: ℓ(σ) is the minimal number k such that
σ can be written as the product of k generators, allowing multiple use of each generator. For
q ∈ [−1, 1] we define the symmetrizer of type D on H⊗n,
q = Xσ∈D(n)
bP (n)
bP (0)
q = IH ⊗0
q
qℓ(σ) σ,
n ≥ 1,
(2.13)
(2.14)
is a special case of the operator defined in [BSz03].
0 = IH ⊗n. Moreover let
with convention 00 = 1. Our operator bP (n)
Note that we have bP (n)
bPq =
be the symmetrizer of type D on the algebraic full Fock space Ffin(H). From Bożejko and
∞Mn=0 bP (n)
Speicher [BSp94, Theorem 2.2], the operator bP (n)
and hence bPq is positive for −1 ≤ q ≤ 1. If
q < 1 then bP (n)
hf, giq := hf,bPqgi0,
is a strictly positive operator meaning that it is positive and Ker(bP (n)
which is a semi-inner product from the positivity of bPq for q ∈ [−1, 1]. We restrict the pa-
rameters to the case q ∈ (−1, 1) so that the deformed semi-inner product becomes an inner
product.
The symmetrizer of type D allows us to define the deformation
f, g ∈ Ffin(H),
) = {0}.
(2.15)
q
q
q
q
Definition 2.3. For q ∈ (−1, 1), the algebraic full Fock space Ffin(H) equipped with the
inner product h·,·iq is called the (q-) Fock space of type D and is denoted by F(q)
fin (H). Let
d∗q(x) := l∗(x) and dq(x) be its adjoint in F(q)
fin (H). The operators d∗q(x) and dq(x) are called
(q-) creation and annihilation operators of type D, respectively.
More precisely, one can show that d∗q(x) is a bounded operator from H⊗n to H⊗(n+1) for
each n ≥ 0, and so dq(x) : H⊗(n+1) → H⊗n is defined to be its adjoint. They can then be
extended to linear operators on F(q)
fin (H) by direct sum. We see in Proposition 3.8 that they are
in fact bounded operators for q ∈ (−1, 1) on F(q)
fin (H))
is linear and dq : H → B(F(q)
Since bP0 is identity, our q-Fock space of type D is the full Fock space when q = 0, and
d∗0(x) = l∗(x) and d0(x) = l(x).
fin (H). One can check that d∗q : H → B(F(q)
fin (H)) is anti-linear, similarly to the free case l∗, l on Ffin(H).
Remark 2.4. In the type B case [BEH15] we defined the symmetrizer with two parameters
α, q ∈ [−1, 1] by
(2.16)
P (n)
α,q = Xσ∈B(n)
αℓ1(σ)qℓ2(σ)σ,
where ℓ1(σ) is the number of πn that appear in an irreducible form of σ and ℓ2(σ) = ℓ(σ)−ℓ1(σ).
However, in the type D case the number ofbπn−1 in an irreducible form of σ ∈ D(n) is not well
defined since πn−2bπn−1πn−2 =bπn−1πn−2bπn−1. Therefore, we introduce only one parameter q in
the type D case.
5
3 Creation and annihilation operators of type D
3.1 Recursive formula for symmetrizer of type D
is defined by
A natural embedding of D(n− 1) = grphπ1, . . . , πn−2,bπn−2i into D(n) = grphπ1, . . . , πn−1,bπn−1i
k = 1, 2, . . . , n − 2,
πk 7→ πk+1,
bπn−2 7→bπn−1.
(3.1)
(3.2)
. This decom-
Corresponding to the quotient D(n − 1)\D(n), we decompose the operator bP (n)
position is important throughout the paper.
q
Proposition 3.1. We have the decomposition
q
q )∗(I ⊗ bP (n−1)
),
n ≥ 2,
(3.3)
n−1Xk=1
qkπn−1 · · · πn−k! , n ≥ 3,
(3.4)
n = 2,
n = 1.
is a bounded linear operator on H⊗n defined by
q
where bR(n)
q
q
,
I +
q = (bR(n)
)bR(n)
q = (I ⊗ bP (n−1)
bP (n)
bP (1)
q = bR(1)
n−1Xk=1
+qn−1π1 · · · πn−2bπn−1 I +
I + qπ1 + qbπ1(I + qπ1),
qkπ1 · · · πk
q =
bR(n)
I,
w(k) =
Proof. Let n ≥ 2. It is known that there exist unique left coset representatives {w(k) 0 ≤
k ≤ 2n − 1} for D(n − 1)\D(n) with minimal lengths [Hum90, p. 19]. Due to Stumbo [Stu00,
Theorem 4], these coset representatives are given by
π1 · · · πk
π1 · · · πn−2bπn−1
π1 · · · πn−2bπn−1πn−1 . . . π2n−k
if 0 ≤ k ≤ n − 1,
if k = n,
if n + 1 ≤ k ≤ 2n − 1.
(3.5)
Therefore every σ ∈ D(n) decomposes into σ = σ′w(k) for some unique σ′ ∈ D(n − 1) and
unique k, and in this case it is known that ℓ(σ) = ℓ(σ′) + ℓ(w(k)) (see [Hum90, p. 19]). Since
q
q =
qℓ(w(k))w(k),
2n−1Xk=0
bR(n)
is written in the form
bR(n)
we have the first identity (3.3). The second identity follows by the selfadjointness of bP (n)
The operator bR(n)
operator in terms of bR(n)
πk.
Lemma 3.2. For x ∈ H and n ≥ 2 we have the following relations on H⊗n:
plays a central role in this paper. Firstly we compute the annihilation
. For this purpose we need a commutation relation between l∗(x) and
(3.6)
.
q
q
q
This implies l∗(x)bP (n)
q = (I ⊗ bP (n)
q
l∗(x)πk = πk+1l∗(x),
l∗(x)bπn−1 =bπnl∗(x).
)l∗(x) for n ≥ 1.
6
1 ≤ k ≤ n − 1,
Proof. Let n ≥ 2 and f = x1 ⊗ · · · ⊗ xn ∈ H⊗n. For 1 ≤ k < n we have
l∗(x)πkf = x ⊗ x1 ⊗ · · · ⊗ xk+1 ⊗ xk ⊗ · · · ⊗ xn = πk+1l∗(x)f
and
l∗(x)bπn−1f = x ⊗ x1 ⊗ · · · ⊗ xn ⊗ xn−1 =bπnl∗(x)f.
3.2 Formula for annihilation operator
Now we are ready to compute the annihilation operator in terms of bR(n)
Proposition 3.3. For n ≥ 1, we have
q
.
dq(x) = l(x)bR(n)
q
on H⊗n.
(3.7)
q
)l∗(x) (see Lemma
3.2) we get
Proof. Let f ∈ H⊗(n−1), g ∈ H⊗n. Recalling that l∗(x)bP (n−1)
= (I ⊗ bP (n−1)
hf, dq(x)giq = hd∗q(x)f, giq = hl∗(x)f, giq = hbP (n)
l∗(x)f, gi0
= hl∗(x)f, (I ⊗ bP (n−1)
q gi0 = h(I ⊗ bP (n−1)
)bR(n)
)l∗(x)f,bR(n)
= hl∗(x)bP (n−1)
f,bR(n)
q gi0 = hbP (n−1)
f, l(x)bR(n)
= hf, l(x)bR(n)
the conclusion.
q giq,
q gi0
q
q
q
q
q
q
q gi0
(3.8)
We compute the annihilator dq(x) more explicitly. Let N be the number operator on F(q)
fin (H)
defined by
N(f ) = nf,
and let J be the operator 0 ⊕L∞n=1 πn on F(q)
f ∈ H⊗n, n ∈ N ∪ {0}
fin (H), that is,
J(x1 ⊗ · · · ⊗ xn) = x1 ⊗ · · · ⊗ xn−1 ⊗ xn,
J(Ω) = 0.
(3.9)
(3.10)
(3.11)
n ≥ 1,
The operator J is in fact a selfadjoint involution on F(q)
fin (H), which is proved in Corollary 3.7.
Let lq and rq be left and right q-derivatives respectively: for x, x1, . . . , xn ∈ H, n ≥ 1,
lq(x)(x1 ⊗ · · · ⊗ xn) =
rq(x)(x1 ⊗ · · · ⊗ xn) =
lq(x)(Ω) = rq(x)(Ω) = 0,
nXk=1
nXk=1
qk−1hx, xki x1 ⊗ · · · ⊗ xk ⊗ · · · ⊗ xn,
qk−1hx, xn−k+1i x1 ⊗ · · · ⊗ xn−k+1 ⊗ · · · ⊗ xn,
(3.12)
(3.13)
(3.14)
where xk means that xk is removed from the tensor, e.g. x ⊗ y ⊗ z = x ⊗ z.
Theorem 3.4. For x ∈ H we have
dq(x) = lq(x) + qN Jrq(x) = lq(x) + Jrq(x)qN−1.
(3.15)
Note that d0(x) = l0(x) = l(x) is the free left annihilation operator.
7
Proof. Let n ≥ 2. From Propositions 3.1 and 3.3 we have
q (x1 ⊗ · · · ⊗ xn) = L + R,
where
dq(x)(x1 ⊗ · · · ⊗ xn) = l(x)bR(n)
qkπ1 · · · πk! (x1 ⊗ · · · ⊗ xn),
n−1Xk=1
L = l(x) I +
R = qn−1l(x)π1 · · · πn−2bπn−1 I +
n−1Xk=1
After some computations, we get
qkπn−1 · · · πn−k! (x1 ⊗ · · · ⊗ xn).
(3.16)
(3.17)
(3.18)
(3.19)
(3.20)
(3.21)
(3.22)
L = lq(x)(x1 ⊗ · · · ⊗ xn),
R = qn−1
nXk=1
qk−1hx, xn−k+1i J(x1 ⊗ · · · ⊗ xn−k+1 ⊗ · · · ⊗ xn).
By the selfadjointness of the involution we have hx, xn−k+1i = hx, xn−k+1i, and hence we get
the conclusion (3.15) on H⊗n for n ≥ 2. For n = 0, 1 we can directly check the formula.
3.3 Commutation relations and norm estimate
The commutation relations on the one particle space H⊗1 and on the other spaces H⊗n, n 6= 1
look different.
Proposition 3.5. For x, y ∈ H we have the commutation relation
dq(x)d∗q(y) − qd∗q(y)dq(x) = hx, yiI + hx, yi q2nJ
on H⊗n, n = 0, 2, 3, 4, . . .
Remark 3.6. For n = 1, the commutation relation has the form
dq(x)d∗q(y)z − qd∗q(y)dq(x)z = hx, yiz + qhx, ziy + q2hx, yi z
for z ∈ H.
The commutation relation for n 6= 1 looks similar to the type B case [BEH15]
bα,q(x)b∗α,q(y) − qb∗α,q(y)bα,q(x) = hx, yiI + αhx, yi q2N ,
while n = 1 case appears quite different.
Proof. Let n ≥ 2. From Theorem 3.4, it holds that
dq(x)d∗q(y)(x1 ⊗ · · · ⊗ xn) =
and
qd∗q(y)dq(x)(x1 ⊗ · · · ⊗ xn) =
and the conclusion follows.
qkhx, xki y ⊗ x1 ⊗ · · · ⊗ xk ⊗ · · · ⊗ xn
nXk=1
nXk=1
+ hx, yi x1 ⊗ · · · ⊗ xn + q2nhx, yi x1 ⊗ · · · ⊗ xn
nXk=1
nXk=1
qkhx, xki y ⊗ x1 ⊗ · · · ⊗ xk · · · ⊗ xn
+
+
qn+k−1hx, xn−k+1i J(y ⊗ x1 ⊗ · · · ⊗ xn−k+1 ⊗ · · · ⊗ xn)
qn+k−1hx, xn−k+1i J(y ⊗ x1 ⊗ · · · ⊗ xn−k+1 ⊗ · · · ⊗ xn),
8
Corollary 3.7. The operator J is a selfadjoint involution on (the completion of ) F(q)
particular, kJkq = 1.
Proof. The operator J is an involution since so is πn. When q = 0, the Fock space of type D is
the full Fock space and the selfadjointness of πn implies the selfadjointness of J. When q 6= 0,
take x ∈ H such that kxk = 1, x = ±x. The commutation relation (3.21) for y = x reads
fin (H). In
J = ±q−2n(dq(x)d∗q(x) − qd∗q(x)dq(x) − 1) on H⊗n, n = 0, 2, 3, . . . ,
(3.23)
so J is selfadjoint on (H⊗n,k · kq), n 6= 1. Finally, since JH ⊗1 = π1H ⊗1, the restriction of J to
Any selfadjoint involution A on a Hilbert space has the operator norm 1 since kAk2 =
H⊗1 is selfadjoint. Note that bP (1)
q = I and so the inner product is not deformed on H⊗1.
kA∗Ak = kA2k = kAk.
We study the norm of the creation operator of type D. Let [n]q be the q-number
[n]q := 1 + q + · · · + qn−1,
n ≥ 1.
(3.24)
Proposition 3.8. Suppose that x ∈ H.
(1) For −1 < q < 1, we have kxk√1−q ≤ kd∗q(x)kq;
(2) For 0 ≤ q < 1, we have kd∗q(x)kq ≤q 2
(3) For −1 < q < 0, we have kd∗q(x)kq ≤p1 + q + q2kxk.
kx⊗nk2
Proof. (1) Lower bound. For n ≥ 2 it follows that
1−qkxk;
q
q = hx⊗n,bP (n)
q x⊗ni0
)x⊗n,bR(n)
= h(I ⊗ bP (n−1)
q x⊗ni0
= [n]qh(I ⊗ bP (n−1)
)x⊗n, x⊗ni0 + qn−1[n]qh(I ⊗ bP (n−1)
x⊗n−1, x⊗(n−1)i0 + hx, xi qn−1[n]qhbP (n−1)
= [n]qkxk2hbP (n−1)
= [n]qkxk2kx⊗(n−1)k2
q + hx, xi qn−1[n]qhx⊗(n−1), x⊗(n−2) ⊗ xiq,
q
q
q
q
)x⊗n, x ⊗ x⊗(n−2) ⊗ xi0
x⊗(n−1), x⊗(n−2) ⊗ xi0
and so
kd∗q(x)x⊗(n−1)k2
q
q = kx⊗nk2
= [n]qkxk2kx⊗(n−1)k2
q + hx, xi qn−1[n]qhx⊗(n−1), x⊗(n−2) ⊗ xiq.
(3.25)
The limit n → ∞ gives the lower bound.
(2) Upper bound for 0 ≤ q < 1. The proof follows the line of [BSp91, Lemma 4]. As al-
q
is positive, so
q
ready noted, [BSp94, Theorem 2.2] guarantees that the operator bP (n)
q )(bR(n)
q )∗(I ⊗ bP (n−1)
)∗(I ⊗ bP (n−1)
)2 = (bP (n)
)∗bP (n)
)∗(bR(n)
= (I ⊗ bP (n−1)
0(I ⊗ bP (n−1)
≤ kbR(n)
q k2
= kbR(n)
0(I ⊗ (bP (n−1)
q k2
(bP (n)
)2).
)
q
q
q
q
q
q
q
)
(3.26)
9
q k0 ≤ (1 + qn−1)[n]q, for n ≥ 1. By taking
q
)
)
q
the square root of operators one gets the inequality
≤
Therefore we have for f ∈ H⊗n that
q ≤ kbR(n)
bP (n)
≤ (1 + qn−1)[n]q(I ⊗ bP (n−1)
Since kπik0 = 1 and kbπik0 = 1 for all i, we have kbR(n)
q k0(I ⊗ bP (n−1)
I ⊗ bP (n−1)
hd∗q(x)f, d∗q(x)fiq = hbP (n+1)
1 − qhx ⊗ (bP (n)
1 − qhbP (n)
q f, fi0hx, xi
qkxk2.
1 − qkfk2
2
1 − q
q
2
2
2
≤
=
=
q
.
(x ⊗ f ), x ⊗ fi0
q f ), x ⊗ fi0
(3.27)
(3.28)
This inequality holds true for f in F(q)
fin (H) which is the direct sum of H⊗n. Thus we obtain
kd∗q(x)fk ≤q 2
1−qkfkqkxk for any f ∈ F(q)
fin (H).
(3) Upper bound for −1 < q < 0. The operators dq(x)d∗q(x) and −qd∗q(x)dq(x) are positive
and map H⊗n into itself for n ≥ 0. So for ξ ∈ H⊗n,kξkq = 1, n ≥ 2,
kd∗q(x)ξk2
q = hξ, dq(x)d∗q(x)ξiq ≤ hξ, dq(x)d∗q(x)ξ − qd∗q(x)dq(x)ξiq
≤ kxk2 + hx, xiq2nkJkq ≤ (1 + q2)kxk2,
where we used the commutation relation (3.21) and kJkq = 1,kxk = kxk on the second line.
For ξ ∈ H,kξkq = kξk = 1, a similar estimate shows that
kd∗q(x)ξk2
q ≤ hξ, dq(x)d∗q(x)ξ − qd∗q(x)dq(x)ξiq
≤ kxk2 + qkxk2 + q2kxk2,
where we used equation (3.22). Finally for Ω ∈ H⊗0 we have kd∗q(x)Ωk2
estimates prove the conclusion.
q = kxk2. These
Remark 3.9. For q = 1, our operators dq(x) are unbounded.
4 Gaussian operator of type D
4.1 Orthogonal polynomials
Definition 4.1. The bounded selfadjoint operator
Gq(x) = dq(x) + d∗q(x),
x ∈ H
(4.1)
on F(q)
the (q-) Brownian motion of type D.
fin (H) is called the (q-) Gaussian operator of type D. The family {Gq(x) x ∈ H} is called
10
In this section we study the probability distribution of the Gaussian operator of type D
with respect to the vacuum state.
Let µ be a probability measure on R with finite moments of all orders. The Gram-Schmidt
method applied to the sequence 1, t, t2, t3, . . . in the Hilbert space L2(R, µ) yields a sequence of
orthogonal polynomials P0(t), P1(t), P2(t), . . . with deg Pn(t) = n. We take the normalization
of Pn(t) such that it becomes monic, i.e., the coefficient of tn is 1. It is known that they satisfy
a recurrence relation
tPn(t) = Pn+1(t) + βnPn(t) + γn−1Pn−1(t),
n = 0, 1, 2, . . .
(4.2)
with the convention that P−1(t) = 0. The coefficients βn and γn are called Jacobi parameters
and they satisfy βn ∈ R, γn ≥ 0 and
γ0 · · · γn−1 =ZR Pn(t)2µ(dt),
n ≥ 1.
(4.3)
Conversely, given a sequence βn ∈ R and γn ≥ 0, n = 0, 1, 2, . . . , there exists a probability
measure µ with finite moments such that it associates orthogonal polynomials determined by
the recursion (4.2). More details are found in [HO07].
Let (P (q)
n (t))∞n=0 be polynomials determined by the recursion relation
n (t) = P (q)
tP (q)
P (q)
0 (t) = 1, P (q)
n+1(t) + [n]q(1 + qn−1)P (q)
1 (t) = t, P (q)
2 (t) = t2 − 1.
n−1(t),
n = 2, 3, 4, . . . ,
(4.4)
This recursion means that βn = 0 for n ≥ 0 and γ0 = 1, γn−1 = [n]q(1 + qn−1) for n ≥ 2.
There exists a probability measure µq which associates the orthogonal polynomials P (q)
n (t).
The distribution of the q-Gaussian operator of type D coincides with µq.
Theorem 4.2. Suppose that q ∈ (−1, 1) and x ∈ H,kxk = 1, x = ±x. The probability
distribution of Gq(x) with respect to the vacuum state is given by µq.
Remark 4.3. We do not know what is the distribution of Gq(x) when x is not an eigenvector
of the involution ¯.
Proof. Using the formula dq(x) = l(x)bR(n)
q
on H⊗n, we have
Gq(x)Ω = x,
Gq(x)x = x⊗2 + Ω
and, using x = ±x,
Gq(x)x⊗n = x⊗(n+1) + [n]qx⊗(n−1) + qn−1[n]qhx, xi x⊗(n−2) ⊗ x
= x⊗(n+1) + [n]q(1 + qn−1)x⊗(n−1),
n ≥ 2.
These formulas are of the same form as (4.4).
Let P (q)
n (t) be the orthogonal polynomials as above. We next show by induction that
n (Gq(x))Ω = x⊗n,
P (q)
n ≥ 0.
(4.5)
(4.6)
(4.7)
This is true for n = 1, 2 since P (q)
induction hypothesis for n − 1, n, (4.4) and (4.6) shows that, for n ≥ 2,
1 (Gq(x))Ω = Gq(x)Ω = x and P (q)
2 (Gq(x))Ω = x⊗2. Using the
P (q)
n+1(Gq(x))Ω = Gq(x)P (q)
n (Gq(x))Ω − [n]q(1 + qn−1)P (q)
n−1(Gq(x))Ω = x⊗(n+1).
11
Since P (q)
n (t) and P (q)
0 (t) = 1 are orthogonal in L2(R, µq) for n ≥ 1, we get
hΩ, P (q)
n (t) dµq(t).
P (q)
n (Gq(x))Ωiq = δ0,n =ZR
hΩ, Gq(x)nΩiq =ZR
0 (t), . . . , P (q)
tn dµq(t).
(4.8)
(4.9)
Writing tn as the linear combination of P (q)
n (t), we obtain by induction on n that
Since the operator Gq(x) is bounded, its vacuum distribution is compactly supported. Hence
the moment problem is determinate and we conclude that Gq(x) has the distribution µq.
Remark 4.4. (1) The t-transformation (or Boolean convolution power by t) of a probability
measure [BW01] changes only the first Jacobi parameters (β0, γ0) into (tβ0, tγ0) and keep
the other Jacobi parameters unchanged. This shows that µq is the 1/2-transformation of
the Gaussian distribution of type B when α = 1, cf. (1.2).
(2) Expanding P (q)
m (t)P (q)
n (t) as the linear combination of tk, k = 0, 1, . . . , m + n and using (4.9)
show that
hΩ, P (q)
m (Gq(x))P (q)
n (Gq(x))Ωiq =ZR
P (q)
m (t)P (q)
n (t) dµq(t),
(4.10)
which generalizes (4.8). The left hand side is equal to hP (q)
hx⊗m, x⊗niq by selfadjointness of P (q)
m (Gq(x))Ω, P (q)
n (Gq(x))Ωiq =
n (Gq(x)). In particular, we obtain from (4.3) that
kx⊗nk2
q = γ0γ1 · · · γn−1,
(4.11)
where γ0 = 1 and γn−1 = [n]q(1 + qn−1) for n ≥ 2. Since kx⊗nk2
equation (4.11) is equivalent to
q
q = hx⊗n,bP (n)
(x⊗n)i0, the
(4.12)
Xσ∈D(n)
qℓ(σ) = γ0γ1 · · · γn−1,
n ≥ 1,
which equals [2]q[4]q . . . [2n−2]q[n]q. This is exactly a formula in the book of Carter [Car89,
Theorem 10.2.3 and Proposition 10.2.5] and confirms our good choice of D(2).
(3) In the limit q → 1 we get the t-transformed normal distribution N(0, 2) with t = 1/2.
(4) If q = 0 then we get the standard semicircle law (1/2π)√4 − t21(−2,2)(t) dt. The orthogonal
polynomials P (0)
n (t) are Chebyshev polynomials of the second kind.
4.2 Set partitions and partition statistics
From now on we study Wick's formula, that is, a formula for correlation functions of Brownian
motion, or more generally, creation and annihilation operators. Combinatorics of set partitions
is needed to describe that formula.
Let [n] be the set {1, . . . , n}. A pair (or a pair block) V of a set partition is a block with
cardinality 2 and a singleton of a set partition is a block with cardinality 1. The set of partitions
of [n] whose blocks have cardinality 1 or 2 is denoted by P1,2(n). When n is even, a set partition
of [n] is called a pair partition if every block is a pair. The set of pair partitions of [n] is denoted
by P2(n).
For subsets A, B of [n], we say that A is on the left of B or B is on the right of A if
min A < min B. We say that A is on the strict left of B or B is on the strict right of A if
max A < min B.
12
Fig. 2: {2, 4} is on the right of {1, 3}, but
not in the strict sense.
Fig. 3: {3, 4} is on the strict right of {1, 2}.
For two subsets A, B of [n], we say that A covers B if there are i, j ∈ A such that i < k < j
for any k ∈ B.
We introduce several statistics of set partitions π. Let Pair(π) be the set of pair blocks and
Sing(π) be the set of singletons of π. When writing pairs we sometimes simplify the notation
into {i < j} ∈ Pair(π) instead of {i, j} ∈ Pair(π), i < j.
Let cover(V ) be the number of blocks of π which cover V :
Let cs(π) be the number of covered singletons, counting multiplicity of covers:
cover(V ) = #{W ∈ π W covers V }.
cs(π) = #{(S, W ) ∈ Sing(π) × π W covers S}.
The number of singletons on the right of V ∈ π is denoted by sr(V ),
sr(V ) = #{S ∈ Sing(π) S is on the right of V }.
Similarly, the number of singletons on the strict right of V ∈ π is denoted by ssr(V ),
ssr(V ) = #{S ∈ Sing(π) S is on the strict right of V }.
The number of left crossings of V ∈ π is defined by
lcr(V ) = #{W ∈ π ∃i, j ∈ V,∃k, l ∈ W, k < i < l < j}.
Let cr(π) be the number of all crossings of π defined by
cr(π) =XV ∈π
lcr(V ).
We then define connected components, in particular outer connected components of a parti-
tion π of [n]. Given two blocks V, W of a partition we write V cr∼ W when V and W cross. Then
we write V ∼ W when V = W or there exist blocks V0 = V, V1, . . . , Vk−1, Vk = W of π such that
cr∼ Vi+1 for all i = 0, 1, . . . , k − 1. The equivalence relation ∼ splits the partition π, regarded
Vi
as a set, into equivalence classes π1, . . . , πm. Then Ck := SV ∈πk
V (k = 1, . . . , m) is called a
connected component of π. By definition, two blocks which cross each other are contained in
the same connected component. For example, the partition π = {{1, 3, 7},{2, 8},{4, 5, 6}} has
two connected components {1, 2, 3, 7, 8} and {4, 5, 6}.
Given two connected components, either one is on the strict right of the other, or one
covers the other. A connected component is said to be outer if it is not covered by any
other connected components.
In the partition π = {{1, 3, 7},{2, 8},{4, 5, 6}} the connected
component {1, 2, 3, 7, 8} is outer.
We want to exclude singletons: the number of outer connected components which are not
singletons is denoted by out(π). Moreover, out\sr(π) denotes the number of connected com-
ponents which are not singletons and which do not have singletons of π on the right. In the
partition π = {{1, 3, 6},{2, 5},{4},{7, 8}} the outer connected component {1, 2, 3, 5, 6} has a
singleton {4} on the right, so it is not counted into out\sr(π). The other outer connected
component {7, 8} is counted, so out\sr(π) = 1.
13
4.3 Set partitions of type D and colored partition statistics
A pair (π, f ) is called a colored set partition if π is a set partition and f : π → {±1} is a map,
which means a coloring of the blocks of π [BEH15]. A block colored by −1 is called a negative
block and a block colored by 1 is called a positive block.
Sometimes we want to think of partitions which are partially colored. When at least the
pairs of π are colored, namely a map f : Pair(π) → {±1} is given, we denote the set of negative
pairs by NPair(π, f ),
and by np(π, f ) its cardinality,
NPair(π, f ) = {W ∈ Pair(π) : f (W ) = −1},
np(π, f ) = #NPair(π, f ).
For simplicity we often drop the dependency on the coloring f from the notation, like np(π).
The number of negative pairs on the right of V ∈ π is denoted by npr(V ):
npr(V ) = #{W ∈ NPair(π) : W is on the right of V }.
Let cnp(π, f ) be the number of covered negative pairs counting multiplicity of covers:
cnp(π, f ) = #{(V, W ) ∈ NPair(π) × Pair(π) W covers V }.
Let npssr(π, f ) be the number of pairs formed by a negative pair and a singleton on the strict
right:
npssr(π, f ) = #{(V, S) ∈ NPair(π) × Sing(π) S is on the strict right of V }.
Definition 4.5. Suppose that π ∈ P1,2(n) and f is a coloring of π. A colored set partition
(π, f ) is called a partition of type D if f satisfies the following conditions (A) and (B).
(A) Coloring of pair blocks V ∈ Pair(π) should start from the rightmost pair (= the pair having
the largest left leg) and then go to left ones.
(1) If sr(V ) = lcr(V ) = cover(V ) = 0, then V must be colored by (−1)npr(V ). For example
take π = {{1, 3},{2, 4}}. If the block {2, 4} has color −1, then V = {1, 3} must be
If the block V = {2, 4} has color 1 then the block {1, 3} must be
colored by −1.
colored by 1.
(2) Otherwise, V can be colored by any of −1 and 1. For example, for π = {{1, 3},{2}}
the block {1, 3} has a singleton {2} on the right and so it can be colored by any of −1
and 1.
(B) After coloring all pairs, we assign a unique color to each singleton S as follows.
(1) If S is the rightmost singleton, then S must be colored by (−1)np(π). For example take
π = {{1, 3},{2}}. If {1, 3} is colored by −1 then S = {2} must be colored by −1. If
{1, 3} is colored by 1 then S = {2} must be colored by 1.
(2) Otherwise S must be colored by 1.
We denote by PD
1,2(n) the set of all partitions of type D in P1,2(n). We call (π, f ) ∈ PD
1,2(n)
a pair partition of [n] of type D if π is a pair partition. The set of pair partitions of [n] of
2 (n). Below we need colored partitions of a finite totally ordered set T ,
type D is denoted by PD
denoted e.g. by PD
1,2(T ), which is defined naturally from the unique isomorphism T ≃ [n] for
some n.
14
Remark 4.6. (1) The set of type B partitions PB
that π ∈ P1,2(n) and the singletons of π must be colored by 1 (see [BEH15]).
definition, PD
singleton.
1,2(n) consists of colored partitions (π, f ) such
In our
1,2(n) since a partition of type D may have a negative
1,2(n) is not a subset of PB
(2) About pair partitions, PD
2 (n) which is now the set of all colored pair
partitions. The relationship between them can be written as follows. Take π ∈ P2(n). Let
C1, . . . , Ck be the outer connected components of π and let
2 (n) is a subset of PB
πi = {V ∈ π min(Ci) ≤ min(V ) < max(V ) ≤ max(Ci)},
(4.13)
which is the set of pairs contained in or covered by the outer connected component Ci of
π. Then π =Si∈[k] πi. The leftmost block Li of πi satisfies lcr(Li) = cover(Li) = 0, and
the other blocks V of πi satisfy lcr(V ) 6= 0 or cover(V ) 6= 0. Hence, when we construct a
coloring of π of type D, we can choose arbitrary colors ±1 of the blocks in πi \ {Li} for
each i, and we must choose a unique color of Li. The definition (A1) of type D colorings
says that this unique color of L is determined so that the number of negative blocks of πi
becomes even (cf. Proposition 4.9(1)). This argument gives, with the notation (4.13), a
characterization of pair partitions of type D:
PD
2 (n) = {(π, f ) ∈ PB
2 (n) np(πi, fπi) is even for every i}.
(4.14)
(3) In some sense, the above observation is also compatible with the relation between the
Coxeter groups of type B and D. The Coxeter group of type B can be written as B(n) =
⋊S(n) and hence it can be defined as all signed permutations of the numbers ±1, . . . ,±n.
Zn
2
The group D(n) is a subgroup of B(n) consisting of all signed permutations having an even
number of negative entries in their window notation. Thus we can assign arbitrary signs
±1 to the points 2, . . . , n and then we must assign a unique sign to the first element 1
depending on the signs of the other points 2, . . . , n. The construction of colorings of type
D on each πi in (4.14) is exactly the above construction of D(n) ⊂ B(n).
Example 4.7. Let π = {{1},{2, 5},{3},{4, 6}} ∈ P1,2(6). Then the rightmost pair is {4, 6},
which has the left crossing {2, 5}, so it can be painted by any of ±1. Then the next rightmost
pair is {2, 5} which has the singleton {3} on the right, so it can also be painted by ±1. So
there are four possible colorings of π. The singleton {1} must be painted by 1, and {3} must
be painted by either 1 or −1 depending on np(π).
Example 4.8. There are five pair partitions of {1, 2, 3, 4} of type D. For example if π =
{{1, 4},{2, 3}}, then in our terminology V = {2, 3} is on the right of W = {1, 4} and cover(V ) =
1. According to (A2), V can be painted by any of 1 and −1. If V is positive then (A1) says
that W must be colored by 1, and if V is negative then W must be colored by −1.
1
1
1
1
-1
-1
1
1
-1
-1
1 2 3 4
1 2 3 4
1 2 3 4
1 2 3 4
1 2 3 4
Fig. 4: PD
2 (4).
The colors of blocks of type D partitions must satisfy some properties.
Proposition 4.9. Suppose that π ∈ PD
1,2(n).
15
(1) If π is a pair partition of type D, then np(π) is even (maybe zero).
(2) If π contains the singleton {1}, then {1} has color 1.
Proof. The proofs can be given by contradiction.
(1) This is a statement weaker than (4.14), but we give an independent proof. Suppose that
the number of negative pairs is odd and consider the leftmost pair V = {1, i} for some i ∈ [n].
Then lcr(V ) = 0, cover(V ) = 0, and the other pairs are on the right of V . If V has color 1
then npr(V ) = np(π) is odd, which contradicts (A1) in the definition of partitions of type D.
Otherwise, V has color −1, then npr(V ) = np(π) − 1 is even, so we again get contradiction
with the definition of partitions of type D.
(2) Suppose that {1} is colored by −1. The condition (B) shows that {1} must be the
rightmost singleton, and so it is the unique singleton of π. The restriction π{2,...,n} is also a
(pair) partition of type D since removing the singleton {1} does not matter in the condition (A)
for pairs. Then, what we have proved in (1) says that π{2,...,n} has an even number of negative
pairs, which contradicts the assumption and (B1).
Type D partitions behave nicely with respect to the restriction of partitions on {2, . . . , n}.
In fact they can be characterized by some recursion.
Lemma 4.10. Suppose that (π, f ) ∈ PD
from one of the following procedures.
1,2({2, . . . , n}). We get new partitions (π, f ) ∈ PD
1,2(n)
(1) Add the singleton {1} to (π, f ) and color it by 1.
(2) If π has singletons, choose a singleton {i} of π, create a pair {1, i}, color it and, if necessary,
recolor the rightmost singleton of π\{i} so that we obtain a partition of [n] of type D. More
precisely:
(i) If π has at least two singletons then {1, i} can be painted by any of ±1. After choosing
a color of {1, i}, the rightmost singleton of π \ {i} must be repainted by (−1)np(π);
(ii) If π has the unique singleton {i} then {1, i} must be colored by (−1)np(π).
When (π, f ) runs over PD
above procedures.
1,2({2, . . . , n}), every partition in PD
1,2(n) appears exactly once in the
Proof. We first prove that a resulting partition (π, f ) is indeed a partition of type D. In the
procedure (1), since the leftmost singleton does not matter in coloring the other pairs and
singletons, so (π, f ) is a partition of type D. In the procedure (2), suppose that π has a
singleton at i (≥ 2), pairs V1, . . . , Vt on the left of i and pairs W1, . . . , Wu on the strict right
of i. Suppose that we take i and create a pair {1, i}. The new pair {1, i} is a left crossing
or a covering of V1, . . . , Vt, but not of W1, . . . , Wu. Hence the old colors of W1, . . . , Wu given
in (π, f ) satisfy the conditions (A2),(A1) in the new partition (π, f ) too. For V1, . . . , Vt, they
have the left crossing or covering {1, i} and then according to (A2) any color is allowed, so the
old colors are valid. Thus the old colors of all pairs V1, . . . , Vt, W1, . . . , Wu satisfy the required
conditions (A) on blocks of the type D partition (π, f ).
For singletons, the color of the rightmost singleton of π is uniquely determined according
to (B2). The other singletons of π may keep the old colors unchanged. The above discussions
show that (π, f ) is a partition of type D.
One can check that all the partitions (π, f ) appearing in the above are distinct. Thus we
only need to show that all partitions in PD
1,2(n), the
first case is that 1 is a singleton of σ. Note then that {1} has color 1 by Proposition 4.9. In
this case the restriction (π, f ) := (σ \ {1}, gπ) is a partition of {2, . . . , n} of type D, and then
1,2(n) appear. Given a partition (σ, g) ∈ PD
16
after procedure (1) we get (π, f ) = (σ, g). The second case is that {1, i} is a pair in σ for some
point i. The block {1, i} may cover or cross (from the left) other pairs, say V1, . . . , Vt. Let π
be the restriction of σ to {2, . . . , n}, namely only {1, i} is replaced by {i}. Then V1, . . . , Vt ∈ π
have the singleton i on the right, so the original colors of them satisfy the conditions for type
D partitions of {2, . . . , n}. Keep the original colors of the other pairs too, and (re)color the
rightmost singleton (if any) uniquely to get a partition (π, f ) of type D. Now we can revert this
procedure from (π, f ) to (π, f ) = (σ, g).
4.4 Wick's formula of type D
Given ε = (ε(1), . . . , ε(n)) ∈ {1,∗}n, a set partition π ∈ P1,2(n), written in the form
π = {{i1 < j1}, . . . ,{ik < jk},{s1}, . . . ,{sm}},
k, m ∈ N ∪ {0},
is said to be ε-compatible if ε(ip) = 1 and ε(jp) = ∗ for all 1 ≤ p ≤ k and ε(sp) = ∗ for all
1 ≤ p ≤ m. The set of ε-compatible partitions in P1,2(n) is denoted by P1,2;ε(n). We also let
P2;ε(n) := P1,2;ε(n) ∩ P2(n). Let PD
1,2(n) such
that π is ε-compatible, and similarly PD
1,2;ε(n) be the set of type D partitions (π, f ) ∈ PD
2;ε(n) be defined.
Note that Lemma 4.10 can be extended to PD
1,2;ε(n). More precisely, given ε = (ε(2), . . . , ε(n)) ∈
{1,∗}n−1 and (π, f ) ∈ PD
partition (π, f ) ∈ PD
Similarly, the procedure (2) yields a colored partition in PD
in PD
({2, . . . , n}), the procedure (1) in Lemma 4.10 yields a colored
1,2;(∗,ε)(n) appears in this way.
1,2;(1,ε)(n) and every colored partition
1,2;(∗,ε)(n), and every colored partition of PD
1,2;ε{2,...,n}
1,2;(1,ε)(n) appears in this way.
We establish a vector-version of Wick's formula.
Theorem 4.11. For any x1, . . . , xn ∈ H and any ε = (ε(1), . . . , ε(n)) ∈ {1,∗}n, we have
dε(1)
q
(xn)Ω
q
(x1)· · · dε(n)
= X(π,f )∈PD
× Y{i<j}∈Pair(π)
1,2;ε(n)
f ({i<j})=1
qcr(π)+cs(π)+2cnp(π,f )+2npssr(π,f )
hxi, xji Y{i<j}∈Pair(π)
f ({i<j})=−1
hxi, xji Oi∈Sing(π,f )
xi,
does not have a negative singleton, and xv1 ⊗· · ·⊗ xvm if the rightmost singleton vm is negative.
whereNi∈Sing(π,f ) xi denotes the tensor product xv1 ⊗ · · · ⊗ xvm if Sing(π, f ) = {v1 < · · · < vm}
In the above formula, we use the conventionNi∈∅
Remark 4.12. (1) Usually people working on Fock spaces take x1, . . . , xn from a real Hilbert
subspace of H, and then the order i < j does not matter since hxi, xji = hxj, xii. Our
version is more general in this sense.
xi = Ω.
(2) If #{i ∈ {j, j + 1, . . . , n} ε(i) = 1} > #{i ∈ {j, j + 1, . . . , n} ε(i) = ∗} for some
(xn)Ω = 0. This case is also covered by Theorem 4.11 if we
j ∈ [n], then dε(1)
understand the sum over the empty set is 0 since PD
(x1)· · · dε(n)
q
q
1,2;ε(n) = ∅ in this case.
Proof. The proof is given by induction and is based on Lemma 4.10. When n = 1, dq(x1)Ω = 0
and d∗q(x1)Ω = x1 and hence the formula is true. Suppose that the formula is true for n − 1.
17
Then for any (ε(2),· · · , ε(n)) ∈ {1,∗}n−1, we get
dε(2)
q
(xn)Ω
q
(x2)· · · dε(n)
X(π,f )∈PD
=
× Y{i<j}∈Pair(π)
1,2;ε({2,...,n})
qcr(π)+cs(π)+2cnp(π,f )+2npssr(π,f )
hxi, xji Y{i<j}∈Pair(π)
f ({i<j})=−1
hxi, xji Ok∈Sing(π,f )
(4.15)
xk,
q
We apply the operator dε(1)
ε(1) = 1 from Theorem 3.4.
1,2;ε({2, . . . , n}), to yield the new type D partition (π, f ) ∈ PD
f ({i<j})=1
(x1), which equals d∗q(x1) if ε(1) = ∗ and lq(x1) + Jrq(x1)qN−1 if
Case I: ε(1) = ∗. The operator d∗q(x1) creates a tensor component x1 on the left. In terms
of partitions, this corresponds to procedure (1) in Lemma 4.10: to add the singleton {1} (with
1,2;ε(n). This
color 1) to (π, f ) ∈ PD
map (π, f ) 7→ (π, f ) does not change the numbers cr, cnp, npssr or cs, which is compatible with
the fact that the action of d∗q(x1) does not change the coefficient. Note that if ε(1) = ∗, then
any partition in PD
1,2;ε(n) has the singleton {1}. Hence Theorem 4.11 holds for n and ε(1) = ∗.
1,2;ε({2,· · · , n}) and suppose that π has singletons k1 <
· · · < kp < i < m1 < · · · < mr, negative pair blocks V1, . . . , Vt on the strict left of i and
pair blocks W1, . . . , Wu which cover i. There may be pair blocks on the strict right of i or
positive pair blocks on the strict left of i, but they do not matter. Then we discuss three cases
separately: (i) (p, r) 6= (0, 0), (ii) (p, r) = (0, 0) and (iii) π does not have a singleton.
Case II: ε(1) = 1. Fix (π, f ) ∈ PD
Case II(i): (p, r) 6= (0, 0). In this situation we have at least one singleton other than the
singleton i. In equation (4.15) we have two situations
Ok∈Sing(π,f )
xk = xk1 ⊗ · · · ⊗ xmr
and
xk1 ⊗ · · · ⊗ xmr ,
and discuss these two cases separately.
Case II(i)1: Suppose that np(π) is even, or equivalently, all singletons are colored by 1 (see
(B) in the definition of partitions of type D). In this case
Ok∈Sing(π,f )
xk = xk1 ⊗ · · · ⊗ xkp ⊗ xi ⊗ xm1 ⊗ · · · ⊗ xmr .
We discuss the left and right annihilation operators separately.
Case II(i)1(a): The left action. The q-derivative lq(x1) creates new p + r + 1 terms.
In
the ith term the inner product hx1, xii appears with coefficient qp. In terms of partitions this
corresponds to a case of Lemma 4.10(2i): to get a set partition (π, f ) ∈ PD
1,2;ε(n) by adding the
pair {1, i} with color 1 to π \ {i}. This pair crosses the blocks W1, . . . , Wu and so increases the
crossing number by u but decreases the number of covered singletons by u. The new covered
singletons {k1}, . . . ,{kp} and new inner negative blocks V1, . . . , Vt appear. Because i is not a
singleton in (π, f ), the number of singletons on the strict right of negative blocks decreases by
t. Altogether we have: cr(π) = cr(π) + u, cs(π) = cs(π) − u + p, cnp(π, f ) = cnp(π, f ) + t and
npssr(π, f ) = npssr(π, f ) − t. So the exponent of q increases by p. This factor qp is exactly the
factor appearing in the q-derivative formula (3.12) when lq(x1) acts on xi, see Figure 5.
Case II(i)1(b). When Jrq(x1)qN−1 acts on the tensor, then new p + r + 1 terms appear by
using the right q-derivative formula (3.20). The ith term has the coefficient qr+(p+r)hx1, xii. In
terms of partitions, this corresponds to a case of Lemma 4.10(2i): to create the new pair {1, i}
with color −1 and repaint the rightmost singleton mr or kp by −1 because now the number of
18
{1, i} colored by 1
-
-
-
-
1-th
∗
k1-th
. . .
∗
kp-th
∗
∗
i-th
1
∗
∗
∗
. . .
m1-th
mr-th
∗
∗
Fig. 5: The visualization of the action lq(x1) on the tensor product xk1 ⊗ · · · ⊗ xmr.
{1, i} colored by -1
-
-
-
-
1-th
∗
k1-th
. . .
∗
kp-th
∗
∗
i-th
-1
∗
∗
∗
. . .
m1-th
mr-th
∗
∗
Fig. 6: The visualization of the action rq(x1) on the tensor product xk1 ⊗ · · · ⊗ xmr .
negative pairs of the new partition π is odd – see Figure 6. Similarly to Step 1(a), we count the
change of numbers and get cr(π) = cr(π) + u, cs(π) = cs(π) − u + p, cnp(π, f ) = cnp(π, f ) + t
and npssr(π, f ) = npssr(π, f ) − t + r. Altogether, when moving from (π, f ) to (π, f ), a new
factor q2r+phx1, xii appears, which coincides with the coefficient appearing in the action of
Jrq(x1)qN−1.
Case II(i)2: Suppose that np(π) is odd, or equivalently, the rightmost singleton is negative.
This means that
Ok∈Sing(π,f )
xk = xk1 ⊗ · · · ⊗ xkp ⊗ xi ⊗ xm1 ⊗ · · · ⊗ xmr .
Case II(i)2(a): The left action. The action of lq(x1) creates p + r + 1 terms. Two situations
are possible.
• If r > 0 then we have a situation similar to Step 1(a) i.e. the ith term has the new
coefficient qphx1, xii. In terms of partitions this corresponds to a case of Lemma 4.10(2i):
we create the positive pair {1, i} and then np(π) is still odd so mr is still colored by −1.
• If r = 0 then the ith term creates the new factor qphx1, xii. In terms of partitions this
corresponds to a case of Lemma 4.10(2i): we create the negative pair {1, i}, so now np(π)
is even – see Figure 7(a). The rightmost singleton kp of π still has color 1.
In both situations the new partition (π, f ) ∈ PD
1,2;ε(n) satisfies cr(π) = cr(π) + u, cs(π) =
cs(π) − u + p, cnp(π, f ) = cnp(π, f ) + t and npssr(π, f ) = npssr(π, f ) − t, so the exponent of q
increases by p.
Case II(i)2(b): The right action. The operator Jrq(x1)qN−1 acting on the tensor product
creates new p + r + 1 terms. We have two situations.
• If r > 0 then the ith term creates the new factor qr+(p+r)hx1, xii. Pictorially this corre-
sponds to Lemma 4.10(2i): the new negative pair {1, i} is created and the singleton mr
is repainted by 1 because now np(π) is even.
19
(a)
(b)
Color -1
Color 1
Color 1
Color -1
-
-
. . .
1-th
∗
∗
∗
i-th
-
-
. . .
1-th
∗
kp-th
∗
i-th
∗
Fig. 7: The main structure of (π, f ) in Steps 2(a) and 2(b) when r = 0.
• If r = 0 then the ith term has the new coefficient qphx1, xii = qphx1, xii.
In terms of
partitions the new pair {1, i} is created with color 1, so the number of negative pairs is
still odd and we must repaint the singleton kp by −1 – see Figure 7(b). This is compatible
with the fact that under this action we obtain xk1 ⊗ · · · ⊗ xkp.
Altogether, when moving from (π, f ) to (π, f ), the exponent of q increases by 2r + p, which
coincides with the coefficient appearing in the action of Jrq(x1)qN−1, creating the inner product
hx1, xii if r > 0 and hx1, xii if r = 0. Similarly to Case II(ii)1(b), we get cr(π) = cr(π) + u,
cs(π) = cs(π) − u + p, cnp(π, f ) = cnp(π, f ) + t and npssr(π, f ) = npssr(π, f ) − t + r.
and hence Nk∈Sing(π,f ) xk = xi or xi, according to the color of the singleton. We recall that
Jrq(x1)xi = Jrq(x1)xi = 0, and so we only need to discuss the left q-derivative lq(x1). The
action creates the pair {1, i}. We have two situations.
Case II(ii): ε(1) = 1 and p = r = 0. In this case the partition (π, f ) has a unique singleton,
• If (π, f ) has the unique positive singleton i, then we have an even number of negative
pairs. The ith term of lq(x1)xi is hx1, xii.
In terms of partitions, this corresponds to
Lemma 4.10(2ii): we create the pair {1, i} with color 1 to get a new pair partition (π, f ).
• If (π, f ) has the unique negative singleton i, then we have an odd number of negative
pairs. The ith term of lq(x1)xi is hx1, xii.
In terms of partitions, this corresponds to
Lemma 4.10(2ii): we create the negative pair {1, i} to get a new pair partition (π, f ).
We emphasize that the number of negative pairs in this new partition is even, which is
compatible with Proposition 4.9 part (1).
Altogether, this pair creates a new partition (π, f ) but it does not change the number cr + cs +
2cnp + 2npssr by the same argument as in Case II(i)1(a). This is compatible with the fact that
the action lq(x1) does not change the exponent of q.
Case II(iii): π does not have an singleton. Then the action of dq(x1) on the vacuum gives
zero. This is compatible with the observation that this situation does not appear in Lemma
4.10, i.e. we cannot pass from PD
1,2(n) by creating a new
pair.
2 ({2, . . . , n}) (= vacuum vector) to PD
Through Case II(i) – Case II(iii) and by (the ε-compatible version of) Lemma 4.10, we
1,2;ε(n) with the desired coefficients.
(x1) yields all partitions in PD
conclude that the action of dε(1)
q
Hence we complete the proof.
20
Example 4.13. We have the formula
dq(x1)dq(x2)d∗q(x3)d∗q(x4)Ω
= qhx1, x3ihx2, x4i + qhx1, x3ihx2, x4i + hx1, x4ihx2, x3i + q2hx1, x4ihx2, x3i,
which corresponds to the colored partitions in Figure 8.
1
1
-1
-1
1
1
-1
-1
1 2 3 4
1 2 3 4
1 2 3 4
Fig. 8: Pair partitions of type D – PD
1 2 3 4
2;(1,1,∗,∗)(4).
When the involution is identity, our vector-version of Wick's formula can be written in terms
of set partitions (of type A).
Proposition 4.14. Suppose that xi = xi ∈ H, i = 1, 2, . . . , n. Then
q
dε(1)
q
(xn)Ω
2−out\sr(π)qcr(π)+cs(π) Y{i<j}∈Pair(π)(cid:0)(cid:0)1 + q2cover({i<j})+2ssr({i<j})(cid:1)hxi, xji(cid:1) Oi∈Sing(π)
(x1)· · · dε(n)
= Xπ∈P1,2;ε(n)
whereNi∈Sing(π) xi denotes xi1 ⊗ · · · ⊗ xik when Sing(π) = {i1 < · · · < ik}.
Proof. Since we have xi = xi, coloring of singletons is not important. Take π ∈ P1,2;ε(n) and
suppose that Pair(π) = {V1, . . . , Vk}. Then
xi,
kYi=1(cid:0)1 + q2(cover(Vi)+ssr(Vi))(cid:1)
= X(n1,...,nk)∈{1,−1}k
= X(n1,...,nk)∈{1,−1}k
= 2#{V ∈Pair(π) lcr(V ) = cover(V ) = sr(V ) = 0}Xf
kYi=1(cid:0)q2(cover(Vi)+ssr(Vi))(cid:1)δni ,−1
i=1(cover(Vi)+ssr(Vi))δni ,−1
q2 Pk
q2cnp(π,f )+2npssr(π,f ),
(4.16)
where f runs over all possible type D colorings of Pair(π). The last formula follows from the
observation that a pair block V ∈ π must be painted by a unique color (1 or −1) if and only if
(4.17)
lcr(V ) = cover(V ) = sr(V ) = 0
by the definition of partitions of type D. So the sum
X(n1,...,nk)∈{1,−1}k
q2 Pk
i=1(cover(Vi)+ssr(Vi))δni ,−1
(4.18)
counts the colorings of V twice as many as the type D colorings of V whenever V satisfies
(4.17).
From Theorem 4.11 it is sufficient to show that the number of pairs V which satisfy condition
(4.17) is equal to out\sr(π), the number of outer connected components with size at least 2
without singletons on the right. This follows from the fact that each connected component
C ∈ out\sr(π) with size ≥ 2 contains a unique block of π which does not have a left crossing
or a covering. In fact it is the block of π having the minimal point of C.
21
The main theorem is the Wick formula of type D, which is similar to the type B case when
α = 1, cf. (1.1), but now we use completely different partitions.
q
(x1)· · · dε(n)
q
Theorem 4.15. Suppose that x1, . . . , xn ∈ H and ε ∈ {1,∗}n.
(xn)Ωiq = X(π,f )∈PD
(1) hΩ, dε(1)
(2) hΩ, Gq(x1)· · · Gq(xn)Ωiq = X(π,f )∈PD
qcr(π)+2cnp(π,f ) Y{i<j}∈π
qcr(π)+2cnp(π,f ) Y{i<j}∈π
2;ε(n)
2 (n)
f ({i<j})=1
hxi, xji.
hxi, xji Y{i<j}∈π
hxi, xji Y{i<j}∈π
f ({i<j})=−1
f ({i<j})=−1
hxi, xji.
f ({i<j})=1
Proof. (1) is clear from Theorem 4.11 and (2) follows from (1) by taking the sum over all ε.
Corollary 4.16. Assume that xi = xi ∈ H for i = 1, . . . , 2m. In the limit q → 1, we recover
the t-transformed classical Brownian motion [BW01, Theorem 9.2] with t = 2:
hΩ, G1(x1)· · · G1(x2m)Ωi1 = Xπ∈P2(2m)
2m−#out(π) Y{i<j}∈π
hxi, xji.
Proof. The formula follows from Proposition 4.14 and taking the sum over ε.
Corollary 4.17. Assume that xi ∈ H for i = 1, . . . , 2m. When q = 0 we recover the moments
for a semicircular system
hΩ, G0(x1)· · · G0(x2m)Ωi0 = Xπ∈NC2(2m) Y{i<j}∈π
hxi, xji,
where NC2(2m) := {π ∈ P2(2m) cr(π) = 0}. Note that the involution − does not appear in
the formula.
Proof. The formula follows from Corollary 4.15(2). Note that 0cr(π)+2cnp(π,f ) gives a nonzero
value (=1) only when cr(π) + 2cnp(π, f ) = 0, which implies that all blocks are positive by the
definition of type D partitions.
4.5 Traciality of the vacuum state
In the context of the von Neumann algebra generated by a free Brownian motion (q = 0 case),
it is common to consider a Brownian motion indexed by a real Hilbert subspace; otherwise the
vacuum state would not be a trace. We follow this strategy and assume that HR is a real Hilbert
subspace of H such that H = HR ⊕ iHR. When considering elements in HR, it holds true that
hx, yi = hy, xi. Let vN(Gq(HR)) be the von Neumann algebra generated by {Gq(x) x ∈ HR}
acting on the completion of F(q)
fin (H).
Proposition 4.18. Let q ∈ (−1, 1). Suppose that dim(HR) ≥ 2. Then the vacuum state is a
trace on vN(Gq(HR)) if and only if q = 0.
Proof. Corollary 4.15(2) for n = 4 reads (see Example 4.8)
hΩ, Gq(x1)Gq(x2)Gq(x3)Gq(x4)Ωiq
= hx1, x2ihx3, x4i + qhx1, x3ihx2, x4i + qhx1, x3ihx2, x4i
+ hx1, x4ihx2, x3i + q2hx1, x4ihx2, x3i
(4.19)
22
and by permuting x1, x2, x3, x4,
hΩ, Gq(x2)Gq(x3)Gq(x4)Gq(x1)Ωiq
= hx2, x3ihx1, x4i + qhx2, x4ihx1, x3i + qhx2, x4ihx1, x3i
+ hx1, x2ihx3, x4i + q2hx1, x2ihx3, x4i,
where the assumption x1, . . . , x4 ∈ HR was used. Hence
hΩ, Gq(x1)Gq(x2)Gq(x3)Gq(x4)Ωiq − hΩ, Gq(x2)Gq(x3)Gq(x4)Gq(x1)Ωiq
= q2(hx1, x4ihx2, x3i − hx1, x2ihx3, x4i).
(4.20)
(4.21)
Therefore the vacuum state is not a trace when q 6= 0, because when dim(HR) ≥ 2 there are
two orthogonal unit eigenvectors e1, e2 of the involution −, and we take x1 = x2 = e1 and
x3 = x4 = e2. When q = 0, the von Neumann algebra becomes the free von Neumann algebra
and the traciality of the vacuum state is well known. It actually follows from the free Wick
formula Corollary 4.17.
Open problems
(1) Study the von Neumann algebra vN(Gq(HR)), in particular, injectivity, completely bounded
approximation property, cyclic separating property of the vacuum, factoriality and type.
(2) Prove the existence of a classical Markov process realization of the Brownian motion of
type D (see [BKS97] for the type A case).
(3) Find a connection between noncrossing partitions of type D in [AR04, Rei97] and our
pair partitions of type D. This problem may be related to the problem of finding a "free
probability of type D" in the spirit of Biane, Goodman and Nica [BGN03].
(4) Construct a Fock space deformed by affine Coxeter groups which are infinite groups.
(5) Compute the explicit form of µq, the distribution of the Gaussian operator of type D. Is it
absolutely continuous with respect to the Lebesgue measure?
(6) Describe the distribution of Gq(x) when x is not an eigenvector of the involution.
(7) Find the exact values of the norms of creation and the Gaussian operators of type D.
Acknowledgments
The work was supported by the MAESTRO grant DEC-2011/02/A/ ST1/00119 (M. Boże-
jko), Austrian Science Fund (FWF) Project No P 25510-N26 (W. Ejsmont), grant number
2014/15/B/ST1/00064 from the Narodowe Centrum Nauki (W. Ejsmont), Wymianę osobową z
Austrią Project No DWM.ZWB.183.1.2016 (M. Bożejko, W. Ejsmont) and JSPS Grant-in-Aid
for Young Scientists (B) 15K17549 (T. Hasebe).
References
[AR04]
C.A. Athanasiadis and V. Reiner, Noncrossing partitions for the group Dn, SIAM
J. Discrete Math. 18, No. 2 (2004), 397–417.
[BGN03] P. Biane, F. Goodman and A. Nica, Non-crossing cumulants of type B, Trans. Amer.
Math. Soc. 355 (2003), 2263–2303.
23
[Bou02]
N. Bourbaki, Lie Groups and Lie Algebras, Chapters 4-6, Springer, New York, 2002.
[BB06]
M. Bożejko, and W. Bryc, On a class of free Lévy laws related to a regression
problem, J. Funct. Anal. 236(1) (2006), 59–77.
[BEH15] M. Bożejko, W. Ejsmont and T. Hasebe, Fock space associated to Coxeter groups
of type B, J. Funct. Anal. 269(6) (2015), 1769–1795.
[BKS97] M. Bożejko, B. Kümmerer and R. Speicher, q-Gaussian processes: non-commutative
and classical aspects, Comm. Math. Phys., 185(1), (1997), 129–154.
[BSp91] M. Bożejko and R. Speicher, An example of a generalized Brownian motion, Com-
mun. Math. Phys. 137 (1991), 519–531.
[BSp94] M. Bożejko and R. Speicher, Completely positive maps on Coxeter groups, deformed
commutation relations, and operator spaces, Math. Ann. 300 (1994), 97–120.
[BSz03] M. Bożejko and R. Szwarc, Algebraic length and Poincaré series on reflection groups
with applications to representations theory, in Asymptotic Combinatorics with Ap-
plications to Mathematical Physics, Lecture Notes in Math. 1815 (2003), 201–221.
[BW01] M. Bożejko and J. Wysoczański, Remarks on t-transformations of measures and
convolutions, Ann. Inst. Henri Poincaré-PR 37(6) (2001), 737–761.
[BW05] W. Bryc and J. Wesołowski, Conditional moments of q-Meixner processes, Probab.
Theory Related Fields 131 (2005), no. 3, 415–441.
[BW14] W. Bryc and J. Wesołowski, Infinitesimal generators of q-Meixner processess,
Stochastic Process. Appl. 124 (2014), no. 1, 915–926.
[Car89]
[GS14]
[HO07]
R.W. Carter, Simple Groups of Lie Type, John Wiley & Sons, London-New York-
Sydney, 1989.
A. Guionnet and D. Shlyakhtenko, Free monotone transport, Invent. Math. 196
(2014), 613–661.
A. Hora and N. Obata, Quantum Probability and Spectral Analysis of Graphs,
Theoretical and Mathematical Physics, Springer, Berlin, 2007. xviii+371 pp.
[Hum90]
J.E. Humphreys, Reflection Groups and Coxeter Groups, Cambridge studies in
advances math. 29, Cambridge University Press, Cambridge, 1990.
[Rei97]
[Ric05]
[Ric06]
[Stu00]
[Wys06]
V. Reiner, Non-crossing partitions for classical reflection groups, Discrete Math.
177 (1997), 195–222.
É. Ricard, Factoriality of q-Gaussian von Neumann Algebras, Comm. Math. Phys.
257 (2005),659–665.
É. Ricard, The von Neumann algebras generated by t-Gaussians, Ann. Inst. Fourier
56 (2006), no. 2, 475–498.
F. Stumbo, Minimal length coset representatives for quotients of parabolic sub-
groups in Coxeter groups, Bul. Un. Math. Ital. Serie 8, 3-B (2000), No. 3, 699–715.
J. Wysoczański, The von Neumann algebra associated with an infinite number of
t-free noncommutative Gaussian random variables, Quantum probability, 435–438,
Banach Center Publ. 73, Polish Acad. Sci. Inst. Math., Warsaw, 2006.
24
[Voi85]
D. Voiculescu, Symmetries of some reduced free product C∗ algebras, Operator
algebras and their connections with topology and ergodic theory, Lect. Notes in
Math. 1132, Springer, Berlin (1985), 556–588.
25
|
1609.02534 | 1 | 1609 | 2016-09-08T19:04:45 | Joint functional calculus in algebra of polynomial tempered distributions | [
"math.FA"
] | In this paper we develop a functional calculus for a countable system of generators of contraction strongly continuous semigroups. As a symbol class of such calculus we use the algebra of polynomial tempered distributions. We prove a differential property of constructed calculus and describe its image with the help of the commutant of polynomial shift semigroup. As an application, we consider a function of countable set of second derivative operators. | math.FA | math | Methods of Functional Analysis and Topology
Vol. 22 (2016), no. 1, pp. 62 -- 73
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL
TEMPERED DISTRIBUTIONS
6
1
0
2
p
e
S
8
]
S. V. SHARYN
Abstract. In this paper we develop a functional calculus for a countable system
of generators of contraction strongly continuous semigroups. As a symbol class of
such calculus we use the algebra of polynomial tempered distributions. We prove
a differential property of constructed calculus and describe its image with the help
of the commutant of polynomial shift semigroup. As an application, we consider a
function of countable set of second derivative operators.
.
A
F
h
t
a
m
[
1
v
4
3
5
2
0
.
9
0
6
1
:
v
i
X
r
a
Introduction
A functional calculus is a theory that studies how to construct functions depending on
operators (roughly speaking, how to "substitute" an operator instead of the variable in
a function). Also it is said that the functional calculus for some (not necessary bounded)
operator A on a Banach space is a method of associating an operator f (A) to a function
f belonging to a topological algebra A of functions.
If we have such a method then,
actually, we have a continuous homomorphism from the algebra A to a topological algebra
of operators. So, in this terminology the functional calculus can be identified with the
above-mentioned homomorphism (but as a theory the functional calculus studies such
homomorphisms).
There are many different approaches to construct a functional calculus for one ope-
rator acting on a Banach space. For Riesz-Dunford functional calculus, based on the
Cauchy formula, we refer the reader to the book [11]. Such a functional calculus has
applications, in particular, in the spectral theory of elliptic differential equations and
maximal regularity of parabolic evolution equations (see e.g. [14, 18]). Another method,
based on the Laplace transformation, was developed in [13]. This method is known as
the Hille-Phillips functional calculus. It has many helpful applications, in particular, in
hydrology (see [2] and the references given there). Such type of calculus is the main
object of investigation in this article.
The Hille-Phillips functional calculus for functions of several variables is well developed
(see e.g. [19, 21]). The case of functions of infinitely many variables is less studied. We
mention the book [23] that is devoted to spectral questions (among them there is a
functional calculus) of countable families of self-adjoint operators on a Hilbert space.
The main goal of this article is the construction of Hille-Phillips type functional calculus
for countable set of generators of contraction strongly continuous semigroups, acting on
a Banach space.
The Borchers-Uhlmann algebra, i.e. the tensor algebra over the space of rapidly de-
creasing functions with tensor product as a multiplication was effectively used in quan-
tum field theory (see e.g.
[6, 7, 27]). Such algebras have an equivalent structure of
2010 Mathematics Subject Classification. Primary 46H30; Secondary 47A60, 46F05.
Key words and phrases. Functional calculus for generators of operator semigroups, polynomials on
locally convex spaces, infinite parameter operator semigroups.
62
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL DISTRIBUTIONS
63
polynomials with pointwise multiplication [9]. It was an incitement to research the prob-
lems connected with the polynomially extended cross-correlation of (ultra)distributions,
differentiations and the corresponding functional calculus [20]. Elements of the Borchers-
Uhlmann algebra can be treated as functionals on spaces of smooth functions of infinite
many variables. So, we can understand this algebra as space of polynomial distributions
with tensor structure. In this paper we would like to consider this structure for a special
case.
A Fr´echet-Schwartz space (briefly, (F S) space) is one that is Fr´echet and Schwartz
simultaneously (see [29]). Let S+ be the space of rapidly decreasing functions on [0, +∞)
and S ′
+ be its dual. It is known (see e.g. [16, 26]) that these spaces are nuclear Fr´echet-
Schwartz and dual Fr´echet-Schwartz spaces ((DF S) for short), respectively. These facts
are crucial for our investigation. The main objects of investigation are the algebras
P(S ′
+) of polynomial test and generalized functions, which have the tensor
+) and P ′(S ′
structures of the forms Lfin S ⊗n
+ and ××× S ′ ⊗n
+ , respectively.
Using the Grothendieck technique [10], we introduce the polynomial extension of cross-
correlation and prove the Theorem 1.3 about isomorphic representation of the algebra of
polynomial distributions onto the commutant of polynomial shift semigroup (see (3)) in
+ . In Proposition 1.4 we prove the
differential property of polynomial cross-correlation, which is essentially used in main
Theorem 3.2.
the space of linear continuous operators on Lfin S ⊗n
In the section 2 we extend the generalized Fourier transformation onto the spaces of
polynomial test and generalized functions. Images of this map we understand as functions
and functionals of infinite many variables (see Remarks 2.1 and 2.2), respectively.
The constructed polynomial test and generalized functions we apply to an operator
semigroup with infinitely many parameters. Namely, we construct the functional calculus
for countable system of generators of contraction C0-semigroups and prove its properties
(see Remark 3.1 and Theorem 3.2). This calculus is an infinite-dimensional analogue of
the one constructed in [19]. As an example we consider the infinite-dimensional Gaussian
semigroup, which is generated by a countable set of second derivative operators.
Finally we note that there are other known and widely used infinite-dimensional ge-
neralizations of classical spaces of distributions [3, 4]. For example, white noise analysis
is based on an infinite dimensional analogue of the Schwartz distribution theory (see e.g.
[12, 15, 17, 22]).
1. Background on polynomial tempered distributions
In what follows L(X, Y ) denotes the space of all continuous linear operators from a
locally convex space X into another such space Y , endowed with the topology of uniform
convergence on bounded subsets of X. Let L(X) := L(X, X) and IX denotes the identity
operator in L(X). The dual space X ′ := L(X, C) is endowed with strong topology. The
pairing between elements of X ′ and X we denote by h · , · i.
Let X ⊗n, n ∈ N, be the symmetric nth tensor degree of X, completed in the projective
∈ X ⊗n, n ∈ N. Set
tensor topology. For any x ∈ X we denote x⊗n := x ⊗ · · · ⊗ x
{z
n
}
X ⊗0 := C, x⊗0 := 1 ∈ C.
For any A ∈ L(X) its tensor power A⊗n ∈ L(X ⊗n), n ∈ N, is defined as a linear
continuous extension of the map x⊗n 7−→ (Ax)⊗n, where x ∈ X. It follows from results
of [5] that such an extension exists if X is a projective or inductive limit of separable
Hilbert spaces. In this article we consider only such spaces.
Let S+ be the Schwartz space of rapidly decreasing functions on R+ := [0, +∞) and S ′
+
be its dual space of tempered distributions supported by R+. Note that strong topology
64
S. V. SHARYN
on S ′
+ coincides with the Mackey topology and topology of inductive limit (see [24, IV.4,
IV.5]). Let δt be the Dirac delta functional concentrated at a point t ∈ R+. It is known
[28] that S ′
+ is a topological algebra with the unit δ := δ0 under the convolution, defined
as
hf ∗ g, ϕi = hf (s), hg(t), ϕ(s + t)ii,
f, g ∈ S ′
+, ϕ ∈ S+.
Note that here and everywhere the notation f (t) shows that a functional f acts on a test
function in the variable t.
From the duality theory as well as from the theory of nuclear spaces it follows that
S+ is a nuclear (F S) space, and S ′
+ is a nuclear (DF S) space.
To define the locally convex space P(nS ′
+) of n-homogeneous polynomials on S ′
+ we
use the canonical topological linear isomorphism
P(nS ′
+) ≃ (S ′ ⊗n
+ )′
described in [9]. Namely, given a functional pn ∈ (S ′ ⊗n
polynomial Pn ∈ P(nS ′
+) by Pn(f ) := pn(f ⊗n), f ∈ S ′
locally convex topology b of uniform convergence on bounded sets in S ′
C. The space P(S ′
linear span of all P(nS ′
strong dual of P(S ′
+ )′, we define an n-homogeneous
+) with the
+) :=
+ is defined to be the complex
+) mean the
+), n ∈ Z+, endowed with the topology b. Let P ′(S ′
+) of all continuous polynomials on S ′
+. We equip P(nS ′
+. Set P(0S ′
+) and
Γ(S+) :=Mfin
n∈Z+
S ⊗n
+ ⊂ Mn∈Z+
S ⊗n
+
and Γ(S ′
+) := ×××
S ′ ⊗n
+ .
n∈Z+
Note that we consider only the case when elements of the direct sum consist of a finite
but not fixed number of terms. For simplicity of notation we write Γ(S+) instead of
commonly used Γfin(S+).
We have the following assertion (see also [20, Proposition 2.1]).
Proposition 1.1. There exist linear topological isomorphisms
Υ : P(S ′
+) −→ Γ(S+), Ψ : P ′(S ′
+) −→ Γ(S ′
+).
Elements of the spaces P(S ′
+) we call the polynomial test functions and
polynomial distributions, respectively. In what follows elements of the spaces Γ(S+) and
Γ(S ′
+) and P ′(S ′
+) will be written as
and ×××
n∈Z+
fn = (f0, f1, . . . , fn, . . . ) ∈ Γ(S ′
+)
for all n ∈ Z+. To simplify, we write
for elements from the total subset (1) of the space Γ(S ′
space by linearity and continuity.
It is obvious that Γ(S ′
+) and extend it to the whole
+) is an algebra relative to
embedded into Γ(S ′
space Γ(S+) becomes an algebra with respect to ⊛.
the operation ⊛ with the unit element (cid:0)δ⊗n(cid:1). Since Γ(S+) is continuously and densely
For any K ∈ L(S+) let us define an operator K ⊗ ∈ L(cid:0)Γ(S+)(cid:1) as follows:
+) (see [20]) and the space S+ is a convolution algebra (see [28]), the
K ⊗ :=(cid:0)K ⊗n(cid:1) : p = (pn)
7−→ K ⊗p :=(cid:0)K ⊗npn(cid:1),
(2)
pn = (p0, p1, . . . , pm, 0, . . . ) ∈ Γ(S+)
mMn=0
(cid:0)pn(cid:1) and (cid:0)fn(cid:1) instead of Lm
for some m ∈ N, where pn ∈ S ⊗n
Note that the following systems of elements
+ and fn ∈ S ′ ⊗n
+
(1)
are total in Γ(S+) and Γ(S ′
Let us define the operation
n=0 pn and ×××n∈Z+ fn, respectively.
+), respectively.
(cid:8)(cid:0)ϕ⊗n(cid:1) : ϕ ∈ S+(cid:9),
+(cid:9)
(cid:8)(cid:0)f ⊗n(cid:1) : f ∈ S ′
(cid:0)f ⊗n(cid:1) ⊛(cid:0)g⊗n(cid:1) :=(cid:0)(f ∗ g)⊗n(cid:1)
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL DISTRIBUTIONS
65
where K ⊗0 := IC and each operator K ⊗n ∈ L(S ⊗n
extension of the map ϕ⊗n 7−→ (Kϕ)⊗n, with ϕ ∈ S+, n ∈ N.
+ ) is defined as a linear continuous
Consider the one-parameter C0-semigroup of shifts,
T : R+ ∋ s 7−→ Ts ∈ L(S+),
Tsϕ(t) := ϕ(t + s),
t ∈ R+, ϕ ∈ S+.
Hence, the map T ⊗n : R+ ∋ s 7−→ T ⊗n
T ⊗n is a one-parameter semigroup of operators. Denote T ⊗
s
mapping
s ∈ L(S ⊗n
+ ) is well defined. It easy to check that
:= (cid:0)T ⊗n
s
(cid:1), s ∈ R+. The
(3)
T ⊗ : R+ ∋ s 7−→ T ⊗
s ∈ L(cid:0)Γ(S+)(cid:1)
is called the polynomial shift semigroup.
The cross-correlation of a distribution f ∈ S ′
+ and a function ϕ ∈ S+ is defined to be
the function
(f ⋆ ϕ)(s) := hf, Tsϕi = hf (t), ϕ(t + s)i.
Similarly to [25] it is easy to prove that
(4)
f ⋆ ϕ ∈ S+,
f ⋆ Tsϕ = Ts(f ⋆ ϕ) and (f ∗ g) ⋆ ϕ = f ⋆ (g ⋆ ϕ)
for any s ∈ R+, f, g ∈ S ′
defined by
+ and ϕ ∈ S+. It follows that the cross-correlation operator
belongs to L(S+) for any f ∈ S ′
+. From (2) it follows that
Kf : ϕ 7−→ f ⋆ ϕ
(5)
where f := (fn) ∈ Γ(S ′
K ⊗
f :=(cid:0)K ⊗n
+) with fn ∈ S ′ ⊗n
fn (cid:1) ∈ L(cid:0)Γ(S+)(cid:1)
+ , n ∈ Z+.
and K ⊗n
fn
∈ L(S ⊗n
+ ),
The cross-correlation of a polynomial distribution f = (fn) ∈ Γ(S ′
+) and a polynomial
test function p = (pn) ∈ Γ(S+) is given by
f ⋆ p := K ⊗
f p =(cid:0)K ⊗n
fn
pn(cid:1).
Proposition 1.2. For any f ∈ Γ(S ′
polynomial test function belonging to Γ(S+).
+) and p ∈ Γ(S+) the cross-correlation f ⋆ p is a
+) and p = (pn) ∈ Γ(S+). Since f ⋆ p = (cid:0)K ⊗n
Proof. Let f = (fn) ∈ Γ(S ′
definition, we only need to check that K ⊗n
fn
this is obvious.
(4)). Consider the case n > 1. Since the operators K ⊗n
fn
sufficient to prove the statement for fn = f ⊗n and pn = ϕ⊗n with f ∈ S ′
Then the function
+ for all n ∈ Z+. In the case n = 0
If n = 1 we obtain that hf1, Tsp1i = (f1 ⋆ p1)(s) belongs to S+ (see
are linear and continuous, it is
+ and ϕ ∈ S+.
pn(cid:1) by
pn ∈ S ⊗n
fn
belongs to S ⊗n
+ as the n-th tensor power of a function from S+.
K ⊗n
fn
pn =(cid:10)f ⊗n, T ⊗n
s ϕ⊗n(cid:11) =(cid:10)f ⊗n, (Tsϕ)⊗n(cid:11) =(cid:10)f, Tsϕ(cid:11)⊗n
⊂ L(cid:0)Γ(S+)(cid:1) of the polynomial shift semigroup T ⊗ is defined
= (f ⋆ ϕ)⊗n
(cid:3)
The commutant (cid:2)T ⊗(cid:3)c
to be the set
(cid:2)T ⊗(cid:3)c
:=(cid:8)K ⊗ ∈ L(cid:0)Γ(S+)(cid:1) : K ⊗ ◦ T ⊗
s = T ⊗
s ◦ K ⊗, ∀ s ∈ R+(cid:9),
where K ⊗ is defined by (2).
Theorem 1.3. The mapping
Γ(S ′
+) ∋ f 7−→ K ⊗
f ∈ L(cid:0)Γ(S+)(cid:1)
66
S. V. SHARYN
is an algebraic isomorphism from the algebra (cid:8)Γ(S ′
the semigroup T ⊗ in the algebra (cid:8)L(cid:0)Γ(S+)(cid:1), ◦(cid:9).
holds:
+), ⊛(cid:9) onto the commutant (cid:2)T ⊗(cid:3)c
of
In particular, the following relation
K ⊗
f ⊛g = K ⊗
f ◦ K ⊗
g ,
f , g ∈ Γ(S ′
+).
Proof. Since the operator K ⊗
f is linear and continuous, it is sufficient to consider only
elements from the total subsets (1). Let p = (ϕ⊗n) ∈ Γ(S+) with ϕ ∈ S+ and f =
(f ⊗n), g = (g⊗n) ∈ Γ(S ′
+ be given. From definitions of operations ⊛ and
⋆, as well as from (4), we obtain
+) with f, g ∈ S ′
K ⊗
f ⊛gp =(cid:16)(cid:10)(f ∗ g)⊗n, T ⊗n
s ϕ⊗n(cid:11)(cid:17) =(cid:16)(cid:0)(f ∗ g) ⋆ ϕ(cid:1)⊗n(cid:17)
=(cid:16)(cid:0)f ⋆ (g ⋆ ϕ)(cid:1)⊗n(cid:17) =(cid:16)(cid:10)f, Ts(g ⋆ ϕ)(cid:11)⊗n(cid:17) =(cid:16)(cid:10)f ⊗n, T ⊗n
s
(g ⋆ ϕ)⊗n(cid:11)(cid:17) = K ⊗
f K ⊗
g p.
Using (4), we obtain
K ⊗
f T ⊗
s p =(cid:0)f ⊗n(cid:1) ⋆(cid:0)T ⊗n
s ϕ⊗n(cid:1) =(cid:0)f ⊗n(cid:1) ⋆(cid:0)(Tsϕ)⊗n(cid:1)
=(cid:0)(f ⋆ Tsϕ)⊗n(cid:1) =(cid:0)(Ts(f ⋆ ϕ))⊗n(cid:1) =(cid:0)T ⊗n
ϕ⊗n(cid:1) = T ⊗
=(cid:0)T ⊗n
s K ⊗n
fn
s K ⊗
f p
s
(f ⋆ ϕ)⊗n(cid:1)
for all s ∈ R+. Hence, the operator K ⊗
Conversely, let K ∈ L(S+) be an operator such that K ⊗ ∈ (cid:2)T ⊗(cid:3)c
. Let us show that
h . Such an element is h := (1, h, . . . , h⊗n, . . . ),
+ is defined by the relation hh, ϕi := (Kϕ)(0), ϕ ∈ S+. Since
there exists h ∈ Γ(S ′
where the distribution h ∈ S ′
(h ⋆ ϕ)(s) = hh, Tsϕi = (KTsϕ)(0) = (Kϕ)(s), we obtain
+) such that K ⊗ = K ⊗
for all f ∈ Γ(S ′
+).
f belongs to (cid:2)T ⊗(cid:3)c
K ⊗
h p =(cid:0)(h ⋆ ϕ)⊗n(cid:1) =(cid:0)(Kϕ)⊗n(cid:1) =(cid:0)K ⊗nϕ⊗n(cid:1) = K ⊗p.
Thus, K ⊗ = K ⊗
h . So, the range of the mapping Γ(S ′
+) ∋ f 7−→ K ⊗
coincides with the commutant (cid:2)T ⊗(cid:3)c
.
f ∈ L(cid:0)Γ(S+)(cid:1)
(cid:3)
Let D mean the differential operator on S+. We use the same letter D to denote the
operator of generalized differentiation on S ′
Let us define the operator D ∈ L(cid:0)Γ(S ′
+, i.e. hDf, ϕi = −hf, Dϕi.
D :
Γ(S ′
+)
(1, f, . . . , f ⊗n, . . . )
−→
+)(cid:1) as follows
7−→ (cid:16)0, Df, . . . ,
nXj=1
Γ(S ′
+)
f ⊗(j−1) ⊗ Df ⊗ f ⊗(n−j), . . .(cid:17).
Its restriction onto Γ(S+) acts as
D :
Γ(S+)
(1, ϕ, . . . , ϕ⊗n, . . . )
−→
7−→ (cid:16)0, Dϕ, . . . ,
nXj=1
Γ(S+)
ϕ⊗(j−1) ⊗ Dϕ ⊗ ϕ⊗(n−j), . . .(cid:17).
Analogically as in [20] it is easy to prove that D is a continuous derivative.
Proposition 1.4. For any f ∈ Γ(S ′
+) and p ∈ Γ(S+) the following equality holds:
(Df ) ⋆ p = −f ⋆ (Dp).
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL DISTRIBUTIONS
67
we have
+) with f ∈ S ′
+ and p = (cid:0)ϕ⊗n(cid:1) ∈ Γ(S+) with ϕ ∈ S+
Proof. For any f = (cid:0)f ⊗n(cid:1) ∈ Γ(S ′
(f ⋆ ϕ)⊗(j−1) ⊗ (Df ⋆ ϕ) ⊗ (f ⋆ ϕ)⊗(n−j), . . .(cid:17)
(Df ) ⋆ p =(cid:16)0, Df ⋆ ϕ, . . . ,
nXj=1
= −(cid:16)0, f ⋆ Dϕ, . . . ,
(f ⋆ ϕ)⊗(j−1) ⊗ (f ⋆ Dϕ) ⊗ (f ⋆ ϕ)⊗(n−j), . . .(cid:17)
nXj=1
(Tsϕ)⊗(j−1) ⊗ (TsDϕ) ⊗ (Tsϕ)⊗(n−j)(cid:11), . . .(cid:17)
= −(cid:16)0,(cid:10)f, TsDϕ(cid:11), . . . ,(cid:10)f ⊗n,
nXj=1
= −(cid:16)0,(cid:10)f, TsDϕ(cid:11), . . . ,(cid:10)f ⊗n, T ⊗n
nXj=1
ϕ⊗(j−1) ⊗ Dϕ ⊗ ϕ⊗(n−j)(cid:11), . . .(cid:17)
s
= − f ⋆ (Dp).
The proposition is proved.
(cid:3)
2. Fourier transform of polynomial tempered distributions
Since each element of the space S+ may be considered as a function ϕ ∈ L1(0, ∞) ∩
L2(0, ∞), we define the Fourier transform and its inverse, as follows:
F+ : S+ ∋ ϕ 7−→ bϕ(ξ) :=ZR+
2π ZR
+ : bϕ 7−→ ϕ(t) =
F −1
1
eitξbϕ(ξ) dξ,
e−itξϕ(t) dt,
ξ ∈ R,
t ∈ R+.
+. The space bS ′
+ is a
Let bS+ := F+[S+] stand for the range of S+ under the map F+.
that bS+ ⊂ L2(R). Using the injectivity of F+, we endow the space bS+ with a topology
induced by the topology in S+. Therefore, bS+ is a nuclear (F ) space (see [24]). For the
strong duals the appropriate adjoint transform (F −1
mapping
+ is well defined. The
It is known [1]
+ )′ : S ′
+ 7−→ bS ′
is called the generalized Fourier transform of distributions from S ′
F ′
+ := 2π(F −1
+ )′ : S ′
+ ∋ f 7−→ bf ∈ bS ′
+
Since delta functional is a unit element in the convolution algebra S ′
nuclear (DF S) space as a strong dual of the nuclear (F S) space bS+ (see [24]).
obtain [δ ∗ f = bf = [f ∗ δ and the space bS ′
the unit bδ with respect to the multiplication bf · bh := [f ∗ h, f, h ∈ S ′
+ (see [28]), we
+ is a commutative multiplicative algebra with
+. The following
bilinear form
hF ′
+f, F+ϕi = h2π(F −1
+ )′f, F+ϕi = 2πhf, F −1
+ F+ϕi = 2πhf, ϕi,
+, ϕ ∈ S+, defines the new duality hbS ′
+) := ×××
with f ∈ S ′
+ and Γ(bS ′
+) with f, h ∈ S ′
n∈Z+ bS ⊗n
Denote Γ(bS+) := Lfin
bh =(cid:0)bh⊗n(cid:1) ∈ Γ(bS ′
and extend it to the whole space Γ(bS ′
Γ(bS ′
+ we define the operation
+, bS+i.
+ . For any elements bf = (cid:0)bf ⊗n(cid:1),
n∈Z+ bS ′ ⊗n
bf b⊛bh :=(cid:0)(bf · bh)⊗n(cid:1)
+) by linearity and continuity. It is obvious, that
+) is an algebra relative to the operation b⊛. Similarly as above, we can induce this
68
S. V. SHARYN
operation on the space Γ(bS+) that becomes an algebra too. From [20, Proposition 2.1]
it follows that there exist the linear topological isomorphisms of algebras
bΥ : P(bS ′
+) −→ Γ(bS+) and bΨ : P ′(bS ′
+) −→ Γ(bS ′
+).
Using Proposition 1.1 we can extend the map F+ onto the space Γ(S+) as follows.
+ by the relations
+ with ϕ ∈ S+ we define the operation F ⊗n
First of all, for any ϕ⊗n ∈ S ⊗n
F ⊗n
+ : ϕ⊗n 7−→ bϕ⊗n
and F ⊗0
+ := IC.
F ⊗
+ to the whole space S ⊗n
+ by linearity and continuity,
Next, we extend the mapping F ⊗n
so F ⊗n
+ is defined to be the mapping
+ , bS ⊗n
+ =(cid:0)F ⊗n
+ (cid:1). Finally, F ⊗
+ (cid:1) : Γ(S+) ∋ p =(cid:0)pn(cid:1)
+ ∈ L(cid:0)S ⊗n
where bpn := F ⊗n
Remark 2.1. Note that bϕ⊗n for any n ∈ N may be treated as a function of n variables
Rn ∋ (ξ1, . . . , ξn) 7−→ bϕ(ξ1) · . . . · bϕ(ξn) ∈ C and may be written in the following way:
7−→ bp :=(cid:0)bpn(cid:1) ∈ Γ(bS+),
+ is a homomorphism of the corresponding
+ pn. It is easy to check that F ⊗
algebras.
e−i(t,ξ)n ϕ(t1) · . . . · ϕ(tn) dt,
bϕ⊗n(ξ1, . . . , ξn) =ZRn
+
(6)
as functions of infinitely many variables
where (t, ξ)n := t1ξ1+· · ·+tnξn, dt := dt1 . . . dtn. So, elements of Γ(bS+) can be considered
bp : (ξ1, . . . , ξn, . . . ) 7−→(cid:0)bp0,bp1(ξ1),bp2(ξ2, ξ3), . . . ,bpn(ξbn , . . . , ξen ), . . .(cid:1),
. But we note that actually eachbp ∈ Γ(bS+) depends
on a finite (depending on bp) number of variables, because for each bp the sequence in the
2 + 1, en := n(n+1)
right-hand side of (6) is finite.
where bn := n(n−1)
2
Define the operator F ′⊗
+ as follows
F ′⊗
+ ) : Γ(S ′
+ := (F ′⊗n
+) ∋ f =(cid:0)fn(cid:1)
+ fn ∈ bS ′ ⊗n
+),
: S ′ ⊗n
n ∈ N, is defined as a linear and continuous extension of the map f ⊗n 7−→ (F ′
f ∈ S ′
where bfn := F ′⊗n
Remark 2.2. From Remark 2.1 it follows that bfn ∈ bS ′ ⊗n
+ ≃ (bS ⊗n
:= IC, and each operator F ′⊗n
7−→ bf :=(cid:0)bfn(cid:1) ∈ Γ(bS ′
+ , F ′⊗0
"variables"
+.
+
+
+ −→ bS ′ ⊗n
+ ,
+f )⊗n with
+ )′ is a functional of n
"variables" in the following sense (cf. (6)):
with bpn ∈ bS ⊗n
bf :
+) we consider as a functional of infinitely many
bpn(ξ1, . . . , ξn) 7−→ hbfn,bpni := hbfn(ξ1, . . . , ξn),bpn(ξ1, . . . , ξn)i ∈ C
+ . So, any bf =(cid:0)bfn(cid:1) ∈ Γ(bS ′
bp(ξ1, . . . , ξn, . . . ) =(cid:0)bpn(ξbn , . . . , ξen )(cid:1)
7−→ hbf ,bpi := Xn∈Z+
Γ(bS+)
−→
C
hbfn,bpni.
3. Infinite parameter operator semigroups
Let E be a complex Banach space. Let A := (A1, A2, . . . , An, . . . ) be a countable
It is convenient for us to rewrite this system as
system of operators, acting on E.
follows. Denote An := (Abn , . . . , Aen ), where bn := n(n−1)
. Let by
definition A0 := ∅. Then the countable system of operators A can be represented as
2 + 1, en := n(n+1)
2
A = (A0, A1, A2, . . . , An, . . . ) or A =(cid:0)An(cid:1) for short.
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL DISTRIBUTIONS
69
For any t ∈ Rn
(see [8, 13]) of n-parameter C0-semigroup Rn
condition
+ let us denote t · An := t1Abn + · · · + tnAen . Let An be a generator
+ ∋ t 7−→ e−it·An ∈ L(E), satisfying the
(7)
ke−it·AnkL(E) ≤ 1.
sup
t∈Rn
+
In what follows we assume that operators of the set An for all n ∈ N commute with
each other. Note that in this case the semigroup can be represented (see [8, 13]) as a
composition of commuting one-parameter marginal semigroups
e−it·An = e−it1Abn ◦ · · · ◦ e−itnAen .
Let G be the set of countable systems of such generators. For all n ∈ N let Gn be
a set of collections of some n generators of one-parameter C0-semigroups satisfying the
condition (7), and let G0 := {∅} by definition.
Define the mapping
(8)
(9)
where eS :=Pn∈Z+ eSn. Here each eSn, n ∈ Z+, is defined to be the space of functions
e−it·An pn(t) dt ∈ L(E)
L := (Ln) : Γ(S+) ∋ p =(cid:0)pn(cid:1)
epn : Gn ∋ An 7−→ epn(An) :=ZRn
+
7−→ ep := Xn∈Z+epn ∈ eS,
in the sense of Bochner.
for n ∈ N, and ep0 : G0 ∋ A0 7−→ ep0(A0) := p0IE ∈ L(E), where the integral is understood
If the assumption (7) holds, then all the mappings Ln : pn 7−→ epn, n ∈ Z+, are
Indeed, the semigroups {e−i(λ,t)IE :
+ satisfy the condition (7). Therefore, their generators
isomorphisms by virtue of [13, Theorem 15.2.1].
t ∈ Rn
(−iλ1IE, . . . , −iλnIE) belong to Gn. Note that
+} with − Im λ ∈ int Rn
epn(−iλ1IE , . . . , −iλnIE) =ZRn
+
e−λ·tpn(t) dt
and an algebra of operator valued functions defined on G. On the other hand, the map
F ⊗
is the Laplace transform of a function pn ∈ S ⊗n
then pn ≡ 0, i.e., Ker Ln = {0}, n ∈ N. Hence, Ker L = {0} and the map L is an
isomorphism.
+ . Particularly, it follows that if epn ≡ 0,
Remark 3.1. The mapping L : Γ(S+) −→ eS is a homomorphism of the algebra(cid:8)Γ(S+), ⊛(cid:9)
+ : Γ(S+) −→ Γ(bS+) is a homomorphism too. So, we can treat the mapping
ep(A) = Pn∈Z+ epn(An) ∈ L(E) as a "value" of a function bp of infinitely many vari-
s ∈ L(cid:0)eS(cid:1) on the space eS,
Consider the one-parameter semigroup eT ⊗ : R+ ∋ s 7−→ eT ⊗
ables (see (6)) at a countable system A := (A1, A2, . . . , An, . . . ) ∈ G of generators of
contraction C0-semigroups.
+ )−1 : Γ(bS+) −→ eS
In other words, we understand the operator
as an "elementary" functional calculus.
L ◦ (F ⊗
where
s
(cid:1) : ep = Xn∈Z+epn
s :=(cid:0)eT ⊗n
eT ⊗
7−→ eT ⊗
s epn ∈ eSn is defined to be the map
The function eT ⊗n
s epn(An) :=ZRn
s epn : An 7−→ eT ⊗n
eT ⊗n
+
s ep := Xn∈Z+ eT ⊗n
s epn.
e−it·Anpn(t + s) dt.
70
S. V. SHARYN
Γ(S+) with pn = ϕ⊗n ∈ S ⊗n
Here the function epn of operator argument is defined by (9).
Using Bochner's integral properties (see [13, 3.7]), we obtain that for any p =(cid:0)pn(cid:1) ∈
s p(A) = L(cid:2)(cid:0)T ⊗n
gT ⊗
+ , ϕ ∈ S+, the following equalities
e−it·Anϕ(t1 + s) · . . . · ϕ(tn + s) dt
s pn(cid:1)(cid:3)(A) = L(cid:2)(cid:0)(Tsϕ)⊗n(cid:1)(cid:3)(A)
= IE +Xn∈N
= ep0(A0) +Xn∈N eT ⊗n
ZRn
+
s epn(An) = eT ⊗
s ep(A)
where fn := f ⊗n ∈ S ′ ⊗n
Define the mapping
(10)
Φ :=(cid:0)Φn(cid:1) : Γ(S ′
s = L ◦ T ⊗
s ◦ L−1. Continuity
of the mappings T ⊗
hold for all s ∈ R+ and A :=(cid:0)An(cid:1) ∈ G.
s can be represented as follows: eT ⊗
Hence, the operator eT ⊗
s ∈ L(cid:0)eS(cid:1) has the C0-property.
s 7−→ eT ⊗
We define commutant of the semigroup eT ⊗ to be the set
s and L as well as openness of L imply that the semigroup eT ⊗ : R+ ∋
(cid:2)eT ⊗(cid:3)c
:=(cid:8)eT ∈ L(cid:0)eS(cid:1) : eT ◦ eT ⊗
+) ∋ f =(cid:0)fn(cid:1)
+. Here Φfn ∈ L(cid:0)eSn(cid:1), n ∈ Z+, is defined by the following
s = eT ⊗
7−→ Φf := Xn∈Z+
s ◦ eT , ∀s ∈ R+(cid:9).
Φfn ∈ L(cid:0)eS(cid:1),
+ , f ∈ S ′
e−it·AnK ⊗n
defined by (2) and (5).
f pn(t) dt, n ∈ N.
formulas: (Φf0ep0)(A0) := IE and
Φfn : epn 7−→ eqn := Φfnepn, where eqn(An) :=ZRn
Here the function epn of operator argument is defined by (9), and the operator K ⊗n
(cid:8)Γ(S ′
of operators of the form eK ⊗ =
L ◦ K ⊗ ◦ L−1 ∈ L(cid:0)eS(cid:1), where K ∈ L(S+). In particular, the equality Φf ⊛g = Φf ◦ Φg
+), ⊛(cid:9) and the subalgebra in the commutant (cid:2)eT ⊗(cid:3)c
Theorem 3.2. The map Φ defined by (10) is an algebraic isomorphism of the algebra
+) and Φδ is the identity in L(cid:0)eS(cid:1), where δ =(cid:0)δ⊗n(cid:1).
Moreover, differential the property
holds for all f , g ∈ Γ(S ′
is
f
+
ΦDfep = −ΦffDp
(11)
holds for any f ∈ Γ(S ′
+) and p ∈ Γ(S+).
the equalities
Proof. For any f =(cid:0)fn(cid:1) ∈ Γ(S ′
(Φfep)(A) = Xn∈Z+
(12)
+), where fn := f ⊗n with f ∈ S ′
+, and p =(cid:0)pn(cid:1) ∈ Γ(S+)
(Φfnepn)(An) = IE +Xn∈N
= L(cid:2)(cid:0)K ⊗n
ZRn
+
e−it·AnK ⊗n
f pn(t) dt
f pn(cid:1)(cid:3)(A) =
]
K ⊗
f p(A)
are valid for all A := (cid:0)An(cid:1) ∈ G. It follows that the map Φ can be represented in the
f ◦ L−1. Continuity of the mappings K ⊗
f and L as well as openness of
form Φf = L ◦ K ⊗
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL DISTRIBUTIONS
71
+). It follows that the equalities
L imply that Φf ∈ L(cid:0)eS(cid:1) for all f ∈ Γ(S ′
+
+
s K ⊗n
f T ⊗n
s pn(t) dt
e−it·AnT ⊗n
e−it·AnK ⊗n
(Φf eT ⊗
(ΦfneT ⊗n
s ep)(A) = Xn∈Z+
ZRn
s epn)(An) = IE +Xn∈N
ZRn
= IE +Xn∈N
s ZRn
= IE +Xn∈N eT ⊗n
= Xn∈Z+
s Φfnepn)(An) = (eT ⊗
(eT ⊗n
s Φfep)(A)
hold for all s ∈ R+, ep = Pn∈Z+ epn ∈ eS and A := (cid:0)An(cid:1) ∈ G. Hence, for all f ∈ Γ(S ′
the operator Φf belongs to the commutant (cid:2)eT ⊗(cid:3)c
Conversely, let eK ⊗ = L ◦ K ⊗ ◦ L−1 ∈ L(cid:0)eS(cid:1) with K ∈ L(S+) belong to the commutant
(cid:2)eT ⊗(cid:3)c
e−it·AnK ⊗n
f pn(t) dt
f pn(t) dt
s ◦ L−1 =L ◦ K ⊗ ◦ L−1 ◦ L ◦ T ⊗
L ◦ K ⊗ ◦ T ⊗
. Then
+)
.
+
=L ◦ T ⊗
s ◦ L−1 ◦ L ◦ K ⊗ ◦ L−1 = L ◦ T ⊗
s ◦ L−1 = eK ⊗ ◦ eT ⊗
s = eT ⊗
s ◦ eK ⊗
s ◦ K ⊗ ◦ L−1,
therefore the operator K ⊗ belongs to the commutant of the semigroup T ⊗
proof of Theorem 1.3 it follows that there exists a unique f ∈ S ′
K ⊗ = K ⊗
s . From the
+ such that K = Kf and
f with f =(cid:0)f ⊗n(cid:1). Hence, eK ⊗ = eK ⊗
f .
The proved above property, Kf ⊛g = Kf ◦ Kg, implies the equality K ⊗
f ⊛g = K ⊗
f ◦ K ⊗
g .
Therefore,
Φf ⊛g = L ◦ K ⊗
f ⊛g ◦ L−1 = L ◦ K ⊗
= L ◦ K ⊗
g ◦ L−1
f ◦ K ⊗
f ◦ L−1 ◦ L ◦ K ⊗
g ◦ L−1 = Φf ◦ Φg.
As a consequence, we obtain the equalities Φδ ◦ Φf = Φδ⊛f = Φf = Φf ⊛δ = Φf ◦ Φδ,
so, Φδ ∈ L(cid:0)eS(cid:1) acts as the identity operator.
It remains to prove the differential property (11). From (12) it follows Φfep = ]f ⋆ p.
So, using the Proposition 1.4, we obtain
Thus, the theorem is proved.
ΦDfep = ^(Df ) ⋆ p = − ^f ⋆ (Dp) = −ΦffDp.
(cid:3)
Remark 3.3. For any fixed p ∈ Γ(S+) the map Γ(S ′
on G. Therefore we can treat this map as a functional calculus in the algebra of polyno-
+) ∋ f 7−→ Φfep ∈ eS is a homomor-
+), ⊛(cid:9) and the algebra of operator-valued functions defined
phism of the algebra (cid:8)Γ(S ′
mial tempered distributions. It is easy to see that a function Φfep of operator argument
can be represented as Φfep = ]f ⋆ p (see (8)). From (12) it follows that the operator
Φfep(A) = ]f ⋆ p(A) ∈ L(E) can be understood as a "value" of a function [f ⋆ p of infi-
nite many variables at a countable system A := (A1, A2, . . . , An, . . . ) ∈ G of generators
of contraction C0-semigroups.
sym(Rn) ≃ L2(R)
Example. Let us consider the case of a countable set of second derivative operators. Let
⊗n, n ∈ N, be the space of complex valued square integrable
Hn := L2
symmetric functions y(ξ) = y(ξ1, . . . , ξn). Set H0 := C. It is known that the symmetric
2 + 1,
Hn is a Hilbert space (see e.g. [22]). As above, let bn = n(n−1)
Fock space H := Mn∈Z+
72
S. V. SHARYN
en = n(n+1)
2
. Define the operators D2
n,m : H −→ H, n ∈ N, bn ≤ m ≤ en, as follows
D2
n,m := 0H0 ⊗ · · · ⊗ 0Hn−1 ⊗
∂2
∂ξ2
m
⊗ 0Hn+1 ⊗ . . . ,
where 0Hn , n ∈ Z+, denote zero operators of the corresponding spaces.
Let us define an "elementary" functional calculus in the algebra of polynomial test
functions for the countable set of operators
D2 := (D2
1,1, D2
2,1, D2
2,2, . . . , D2
n,bn
. . . ).
, . . . , D2
n, n ∈ N, generates the semigroup
n,en
Let D2
n := (D2
n,bn
, . . . , D2
n,en ). It is easy to see that D2
+ ∋ t = (t1, . . . , tn) 7−→ e−it·D2
Rn
n ∈ L(H),
where
e−it·D2
n := IH0 ⊗ . . . ⊗ IHbn−1 ⊗ e
−it1
∂2
∂ξ2
bn ◦ . . . ◦ e
−itn
∂2
∂ξ2
en ⊗ IHen +1 ⊗ . . .
Denote
gn(t, ζ) :=
nYj=1
ζ2
j
4tj .
−
e
1
p4πtj
From [19, Example 2] it follows that the semigroup e−it·D2
n acts as
e−it·D2
n y =(cid:0)y0, . . . , yn−1, gn(−it, · ) ∗ yn, yn+1, . . .(cid:1)
for any y = (y0, y1, . . . , yn, . . . ) ∈ H.
Let p =(cid:0)pn(cid:1) ∈ Γ(S+) be given. If we "substitute" the countable set D2 of operators
instead of variables of a function bp (see (6)) we obtain the operator acting as
n)yn(ξbn , . . . , ξen)
ep(D2)y(ξ1, ξ2, . . . ) =y0 +Xn∈Nepn(D2
ZRn
=y0 +Xn∈N
+
(gn(−it, · ) ∗ yn)(ξbn , . . . , ξen )pn(t) dt,
where y(ξ1, ξ2, . . . ) =(cid:0)y0, y1(ξ1), y2(ξ2, ξ3), . . . , yn(ξbn , . . . , ξen ), . . .(cid:1) ∈ H is a function of
infinite many variables.
Acknowledgments. The author would like to thank the referee for valuable comments
which helped to improve the manuscript.
References
1. N. I. Akhiezer, Lectures on Integral Transforms, Translations of Mathematical Monographs,
vol. 70, Amer. Math. Soc., Providence, RI, 1988. (Russian edition: Vishcha Shkola, Khar'kov,
1984)
2. B. Baeumer, M. Haase, M. Kov´acs, Unbounded functional calculus for bounded groups with
applications, J. Evol. Eqv. 9 (2009), no. 1, 171 -- 195. doi:10.1007/s00028-009-0012-z
3. Yu. M. Berezansky, Yu. G. Kondratiev, Spectral Methods in Infinite-Dimensional Analysis,
Vols. 1, 2, Kluwer Academic Publishers, Dordrecht -- Boston -- London, 1995. (Russian edition:
Naukova Dumka, Kiev, 1988)
4. Yu. M. Berezansky, Yu. S. Samoilenko, Nuclear spaces of functions of infinitely many variables,
Ukrain. Mat. Zh. 25 (1973), no. 6, 723 -- 737. (Russian); English transl. Ukrainian Math. J. 25
(1973), no. 6, 599 -- 609.
5. Yu. M. Berezansky, Z. G. Sheftel, G. F. Us, Functional Analysis, Vol. 2, Birkhauser Verlag,
Basel -- Boston -- Berlin, 1996; 3rd ed., Institute of Mathematics NAS of Ukraine, Kyiv, 2010.
(Russian edition: Vyshcha Shkola, Kiev, 1990)
6. H. Borchers, Algebras of unbounded operators in quantum fields theory, Physica 124A (1984),
127-144.
7. H. Borchers, On structure of the algebra of field operators, Nuovo Cimento 24 (1962), 1418 --
1440.
JOINT FUNCTIONAL CALCULUS IN ALGEBRA OF POLYNOMIAL DISTRIBUTIONS
73
8. P. L. Butzer, H. Berens, Semi-groups of Operators and Approximation, Springer-Verlag, New-
York, 1967.
9. S. Dineen, Complex Analysis on Infinite Dimensional Spaces, Springer-Verlag, Berlin --
Gottingen -- Heidelberg, 1999.
10. A. Grothendieck, Produits tensoriels topologues et espases nucl´eaire, Mem. Amer. Math. Soc.
16 (1955), no. 11, 1 -- 140.
11. M. Haase, The Functional Calculus for Sectorial Operators, Operator Theory: Advances and
Applications, vol. 169, Birkhauser-Verlag, Basel, 2006.
12. T. Hida, H. H. Kuo, J. Potthoff, L. Streit, White Noise: an Infinite Dimensional Calculus,
Kluwer Academic Publishers, Dordrecht, 1993.
13. E. Hille, R. Phillips, Functional Analysis and Semi-groups, Amer. Math. Soc. Coll. Publ.,
vol. 31, Providence, RI, 1957.
14. N. Kalton, P. Kunstmann, L. Weis, Perturbation and interpolation theorems for the H∞-
calculus with applications to differential operators, Math. Ann. 336 (2006), no. 4, 747 -- 801.
15. N. A. Kachanovsky, Elements of a non-gaussian analysis on the spaces of functions of infinitely
many variables, Ukrain. Mat. Zh. 62 (2010), no. 9, 1220 -- 1246. English transl. Ukrainian
Math. J. 62 (2011), no. 9, 1420 -- 1448.
16. H. Komatsu, An Introduction to the Theory of Generalized Functions, University Publ., Tokyo,
2000.
17. Y. G. Kondratiev, L. Streit, W. Westerkamp, J. Yan, Generalized functions in infinite dimen-
sional analysis, Hiroshima Math. J. 28 (1998), no. 2, 213 -- 260.
18. C. Kriegler, Functional calculus and dilation for C0-groups of polynomial growth, Semigroup
Forum 84 (2012), 393 -- 433.
19. O. V. Lopushansky, S. V. Sharyn, Generalized Hille-Phillips type functional calculus for mul-
tiparameter semigroups, Siberian Math. J. 55 (2014), no. 1, 105 -- 117.
20. O. V. Lopushansky, S. V. Sharyn, Polynomial ultradistributions on cone Rd
+, Topology 48
(2009), no. 2 -- 4, 80 -- 90.
21. A. R. Mirotin, On some properties of the multidimensional Bochner-Phillips functional calcu-
lus, Siberian Math. J. 52 (2011), no. 6, 1032 -- 1041.
22. N. Obata, White Noise Calculus and Fock Space, Lect. Notes in Math., vol. 1577, Springer-
Verlag, New-York, 1994.
23. Yu. S. Samoilenko, Spectral Theory of Families of Self-Adjoint Operators, Kluwer Academic
Publishers, Dordrecht -- Boston -- London, 1991. (Russian edition: Naukova Dumka, Kiev,
1984)
24. H. Schaefer, Topological Vector Spaces, Springer-Verlag, New-York, 1971.
25. S. V. Sharyn, The cross-correlation operation of Schwartz distributions, J. Math. Sci. 107
(2001), no. 1, 3604 -- 3609.
26. A. G. Smirnov, On topological tensor products of functional Frechet and DF spaces, Integral
Transforms Spec. Funct. 20 (2009), no. 3 -- 4, 309 -- 318.
27. A. Uhlmann, Uber die definition der quantenfelder nach Wightman und Haag, Wiss. Zeitschr.
Karl-Marx Univ., Leipzig 11 (1962), 213 -- 217.
28. V. S. Vladimirov, Generalized Functions in Mathematical Physics, Mir, Moscow, 1979. (Rus-
sian edition: Nauka, Moscow, 1976)
29. V. V. Zharinov, Compact families of locally convex topological vector spaces, Fr´echet-Schwartz
and dual Fr´echet-Schwartz spaces, Uspekhi Mat. Nauk 34 (1979), no. 4, 97 -- 131. (Russian);
English transl. Russian Math. Surveys 34 (1979), no. 4, 105 -- 143.
Department of Mathematics and Computer Sciences, Precarpathian National University,
57 Shevchenka str., Ivano-Frankivsk, 76018, Ukraine
E-mail address: [email protected]
Received 29/12/2014; Revised 21/04/2015
|
1810.13208 | 1 | 1810 | 2018-10-31T10:49:14 | Spectrum of composition operators on ${\mathcal S}({\mathbb R})$ with polynomial symbols | [
"math.FA"
] | We study the spectrum of operators in the Schwartz space of rapidly decreasing functions which associate each function with its composition with a polynomial. In the case where this operator is mean ergodic we prove that its spectrum reduces to 0, while the spectrum of any non mean ergodic composition operator with a polynomial always contains the closed unit disc except perhaps the origen. We obtain a complete description of the spectrum of the composition operator with a quadratic polynomial or a cubic polynomial with positive leading coefficient. | math.FA | math |
Spectrum of composition operators on S(R)
with polynomial symbols
Carmen Fern´andez, Antonio Galbis, Enrique Jord´a∗
November 1, 2018
Abstract
We study the spectrum of operators in the Schwartz space of
rapidly decreasing functions which associate each function with its
composition with a polynomial.
In the case where this operator is
mean ergodic we prove that its spectrum reduces to {0}, while the
spectrum of any non mean ergodic composition operator with a poly-
nomial always contains the closed unit disc except perhaps the origen.
We obtain a complete description of the spectrum of the composi-
tion operator with a quadratic polynomial or a cubic polynomial with
positive leading coefficient.
1
Introduction
Composition operators on Fr´echet spaces of smooth functions on the reals
have attracted the attention of several authors recently ([7, 5, 8, 9, 10, 11])
but to our knowledge very little is known about the spectra of composition
operators in this setting. See for instance [3] or [2], where the spectrum of
composition operators and other classical operators on spaces of real smooth
functions is investigated. This contrasts with the large number of existing ar-
ticles studying spectral properties of composition operators in Banach spaces
of analytic functions on the unit disc.
We study the spectrum of composition operators defined in the Schwartz
space S(R) of smooth rapidly decreasing functions, Cϕ : S(R) → S(R), f 7→
∗The present research was partially supported by the projects MTM2016-76647-P,
ACOMP/2015/186 (Spain). The third author was partially supported by GVA, Project
AICO/2016/054.
1
f ◦ ϕ, in the case that ϕ is a non constant polynomial. A smooth function
ϕ : R → R is said to be a symbol for S(R) if Cϕ maps S(R) continuously into
itself. The symbols for S(R) were completely characterized in [5, Theorem
2.3]. It follows from that characterization that any non constant polynomial ϕ
is a symbol for S(R). The results of the present paper complement our study
in [7] where we investigate dynamics and the spectrum of some particular
composition operators. Concrete examples were given where the spectrum
coincides with the open unit disc, the unit circle or C \ {0}. In particular, the
spectrum of translation and dilation operators was analyzed in [7, Examples
5-6]:
Example 1.1.
(a) Let ϕ(x) = x + 1. Then σ (Cϕ) = {λ ∈ C :
λ = 1}.
(b) Let ϕ(x) = ax where a 6= 0 and a 6= 1. Then σ (Cϕ) = C \ {0}.
However, in [7] we did not obtain any result concerning the spectrum of
the composition operator with a polynomial of degree greater than one. This
is precisely the objective of this work.
It turns out that some dynamical
properties of the composition operator are characterized by the spectrum of
the operator. For example, the spectrum of a mean ergodic composition op-
erator is always contained in the closed unit disk ([7, Corollary 4.5]). This
result can be improved when the symbol is a polynomial. As we prove in
Theorem 2.5, the spectrum of a composition operator which is mean ergodic
and whose symbol is a polynomial with degree greater than one coincides
with {0} while the behavior of non mean ergodic composition operators with
polynomial symbols is different (Theorem 2.8). For strictly decreasing sym-
bols, not necessarily polynomials, the containment of ∂D in the spectrum of
the operator is equivalent to its mean ergodicity.
In [3] the spectrum of composition operators on A(R), the space of all real
analytic functions, is investigated for the case that the symbol is a quadratic
polynomial. For quadratic polynomials, we have a complete characterization
of the spectrum of the corresponding composition operator depending on the
number of fixed points of the polynomial. As expected, the spectrum of the
operator depends on the space where it is considered. To give an example,
when ϕ is a quadratic polynomial without fixed points then σA(R)(Cϕ) = C,
whereas σS(R)(Cϕ) = {0}.
Theorem 4.1 contains a complete description of the spectrum of a com-
position operator with a polynomial of degree three whose leading coefficient
is positive. For polynomials with negative leading coefficient some partial
results are available but we lack a complete characterization.
2
The final section contains some results concerning the spectra of compo-
sition operators with (non polynomial) monotone symbols.
We recall that given an operator T : X → X on a Fr´echet space X,
σ(T ), the spectrum of T, is the set of all λ ∈ C such that T − λI : X → X
does not admit a continuous linear inverse. T is said to be power bounded
if {T n(x) : n ∈ N} is bounded for each x ∈ X. A closely related concept
to power boundedness is that of mean ergodicity. Given T ∈ L(X), the
k=1 T k/n. T is said to be mean
ergodic when T[n] converges to an operator P , which is always a projection,
in the strong operator topology, i.e. if (T[n](x)) is convergent to P (x) for each
x ∈ X.
Ces`aro means of T are defined as T[n] = Pn
From now on ϕn = ϕ ◦ . . . ◦ ϕ denotes the n-th iteration of ϕ.
The following results will be used in what follows.
Lemma 1.2. [7, 3.10] Let ϕ be a polynomial of even degree without fixed
points. Then there is N ∈ N such that ψ = ϕN has neither zeros nor fixed
points. Moreover, for every K > 0 there is m0 ∈ N such that
ψm+1(t) ≥ K (ψm(t))2 ∀m ≥ m0,
∀t ∈ R.
Theorem 1.3. [7, 3.11] Let ϕ be a polynomial with degree greater than or
equal to two. Then, the following are equivalent:
(1) Cϕ is power bounded.
(2) Cϕ is mean ergodic.
(3) The degree of ϕ is even and it has no fixed points.
2 Polynomial symbols
Two polynomials ϕ, ψ ∈ R[x] are linearly equivalent if there exists ℓ(x) =
ax + b with a, b ∈ R and a 6= 0 such that ψ = ℓ−1 ◦ ϕ ◦ ℓ. Then, for every
λ ∈ C,
Cϕ − λI = C−1
ℓ ◦ (Cψ − λI) ◦ Cℓ,
from where it follows that σ (Cψ) = σ (Cϕ) .
The first result follows immediately from this observation and [7, Exam-
ples 5-6] as each polynomial of degree one other than the identity is linearly
equivalent to a translation or to a dilation.
3
Proposition 2.1. Let ϕ(x) = ax + b a polynomial with a, b ∈ R and a 6= 0.
Then
(a) For a 6= 1, ϕ is linearly equivalent to ψ(x) = ax. Hence σ (Cϕ) = C\{0}
for a 6= ±1 while σ (Cϕ) = {−1, 1} for a = −1.
(b) For a = 1 and b 6= 0, ϕ is linearly equivalent to ψ(x) = x + 1. Hence
σ (Cϕ) = {λ ∈ C : λ = 1} .
From now on we will consider only polynomials of degree greater than
one.
We observe that the following version of the spectral theorem holds in
our setting.
Proposition 2.2. For every symbol ϕ and N ∈ N,
Proof. For every µ ∈ C \ {0} let λ1, . . . , λN denote its N-roots. Then
σ(CϕN ) =(cid:8)λN ∈ C : λ ∈ σ(Cϕ)(cid:9) .
CϕN − µI = C N
ϕ − µI = (Cϕ − λ1I) · · · (Cϕ − λnI) ,
from where the conclusion follows.
We also recall the following elementary properties, which will be used in
what follows.
Proposition 2.3. Let ϕ be a symbol for S(R).
(a) If ϕ admits fixed points then 1 ∈ σ(Cϕ).
(b) If a is a fixed point of ϕ then ϕ′(a) ∈ σ(Cϕ).
(c) If ϕ is a polynomial then 0 ∈ σ(Cϕ) if and only if ϕ′ vanishes at some
point.
Proof. (a) All the functions in the range of Cϕ − I vanish at fixed points of
ϕ, hence the conclusion.
(b) In fact, the derivative of any function in the range of Cϕ − ϕ′(a)I
vanishes at point a, hence Cϕ − ϕ′(a)I is not surjective.
(c) If ϕ′ does not vanish then inf x∈R ϕ′(x) > 0 and Cϕ is surjective by
[5, 4.2]. Moreover Cϕ is injective as ϕ(R) = R. Conversely, if ϕ′(x0) = 0 then
the derivative of any function in the range of Cϕ vanishes at point x0. Hence
0 ∈ σ(Cϕ).
4
We observe that an arbitrary symbol ϕ for S(R) satisfies conditions (i)
and (ii) in the next lemma with r = 1 if, and only if, Cϕ is power bounded.
This is the content of [7, Proposition 3.9].
Lemma 2.4. Let ϕ be a polynomial of degree ≥ 2. Assume that for all
r > 1, n ∈ N there exist C > 0, q ∈ N such that the following conditions
hold for each x ∈ R and m ∈ N:
(i) ϕ(n)
m (x) ≤ Crm(1 + ϕm(x))q
(ii) x ≤ (1 + ϕm(x))q.
Then the series
µmf ◦ ϕm
Xm
is convergent in S(R) for each µ < 1. If in addition we assume that (i)
happens with r = 1 then the series is convergent for each µ ∈ C.
Proof. For each n, m ∈ N0, by Fa´a de Bruno formula
(f ◦ ϕm)(n)(x) =
n
Xj=1
f (j)(ϕm(x))Bn,j(cid:0)ϕ′m(x) . . . ϕ(n−j+1)
m
(x)(cid:1) ,
where Bn,j are the Bell polynomials. Thus, from (i) and (ii), given f ∈ S(R),
n ∈ N0, r > 1 and a polynomial P we find another polynomial P such that
n
Xj=1
f (j)(ϕm(x))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
.
(1)
P (ϕm(x))
(cid:12)(cid:12)P (x)(f ◦ ϕm)(n)(x)(cid:12)(cid:12) ≤ rm(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
convergence in S(R) of P µmf ◦ ϕm follows.
Since f ∈ S(R) there is M > 0 such that (1) can be estimated by Mrm.
Therefore, given µ ∈ D we choose r > 1 such that rµ < 1 and the
If in addition (i) is satisfied with r = 1 then Cϕ is power bounded [7,
Proposition 3.9] and we have the estimate (1) with r = 1. By [7, Theorem
3.1] ϕ has even degree and lacks fixed points.
First we assume that
(a) ϕ(x) > x2 for every x ∈ R
and
(b) inf {ϕ(x) : x ∈ R} = a > 1.
5
This implies that ϕm(x) > a2m−1 and the series
µm
ϕm(x)
Xm
converges absolutely and uniformly in R for every µ ∈ C. Using (1) with
r = 1 we immediately have, for every polynomial P and n ∈ N, that
(cid:12)(cid:12)µmP (x)(f ◦ ϕm)(n)(x)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ M
for some M > 0. Consequently
µm
1 + ϕm(x)
(1 + ϕm(x)) P (ϕm(x))
n
Xj=1
f (j)(ϕm(x))(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
µm
1 + ϕm(x)
µmf ◦ ϕm
Xm
converges in S(R) for each µ ∈ C and f ∈ S(R).
In the general case, we may find N ∈ N such that ϕN satisfies conditions
(a) and (b) ([7, Lemma 3.10]). Hence
µmf ◦ ϕm =
Xm
N−1
Xj=0
µj Xm (cid:0)µN(cid:1)m
(f ◦ ϕj) ◦ ϕN m!
converges in S(R).
Theorem 2.5. Let ϕ be a polynomial with even degree and without fixed
points. Then σ(Cϕ) = {0}.
Proof. From Lemma 1.2 we find N ∈ N such that if ψ = ϕN ,
min{ψ(x) : x ∈ R} = a > 1
and
ψm+1(x) ≥ (ψm(x))2 ∀m,
∀x ∈ R.
In particular, this gives
for all x and every m.
ψm(x) > a2m−1
,
(2)
6
Since the range of ϕ is a proper (unbounded) interval then Cϕ is not
injective and 0 ∈ σ(Cϕ). To finish the proof it suffices to show that σ(Cψ) ⊂
{0}.
To this end, we fix λ ∈ C, λ 6= 0, and check that Cψ − λI is a bijection,
hence a topological isomorphism by the open mapping theorem.
(i) Injectivity. Let us assume Cψf = λf for some f ∈ S(R). Then, for
every m ∈ N,
Since
f (x) = λ−mf (ψm(x)) =
ψm(x)f (ψm(x))
λmψm(x).
λmψm(x) ≥ λm · a2m−1
→ ∞
then f (x) = 0 for all x ∈ R.
(ii) Surjectivity. From [7, Theorem 3.11 and Proposition 3.9] Cϕ is power
bounded and the hypothesis in Lemma 2.4 are satisfied with r = 1. Then,
for every g ∈ S(R) and λ 6= 0 the series
f = −
1
λk+1 g ◦ ψk
∞
Xk=0
(3)
converges in S(R) and clearly Cψf − λf = g.
According to Proposition 2.3 the behavior of composition operators with
polynomials having fixed points is different. In order to obtain more infor-
mation we first we need an auxiliary result.
Lemma 2.6. Let ϕ be a polynomial with degree greater than one. Then,
there is M > 0 such that for each x > M,
lim
n
1
λn f (ϕn(x)) = 0,
for 0 < λ ≤ 1 and every f ∈ S(R).
Proof. We fix 2 > p > 1 and take M > 1 such that ϕ(x) > xp whenever
x > M. Then x > M implies
Finally
ϕn(x) > M pn
∀n ∈ N.
lim
n (cid:12)(cid:12)(cid:12)(cid:12)
1
λn f (ϕn(x))(cid:12)(cid:12)(cid:12)(cid:12)
≤ lim
n (cid:12)(cid:12)(cid:12)(cid:12)
ϕn(x)f (ϕn(x))
λnM pn
= 0.
(cid:12)(cid:12)(cid:12)(cid:12)
7
Lemma 2.7. Let ϕ be a polynomial with odd degree greater than one such
that
ϕ(x) = −∞. Let a be a fixed point of ϕ. If a is the largest fixed
point of ϕ2 then ϕ′(a) ≤ −1.
lim
x→+∞
Proof. We first observe that
x > a we get
lim
x→+∞
ϕ2(x) = +∞. From ϕ2(x) > x for every
(ϕ′(a))2 = ϕ′2(a) = lim
x→a
ϕ2(x) − a
x − a
≥ 1.
Since a is also the largest fixed point of ϕ then ϕ(x) < x for all x > a, from
where it follows
ϕ′(a) = lim
x→a
ϕ(x) − a
x − a
≤ 1.
Consequently ϕ′(a) ≤ −1 or ϕ′(a) = 1. Finally we check that ϕ′(a) ≤ 0.
Otherwise there is δ > 0 such that ϕ is strictly increasing on [a, a + δ]. Since
a < ϕ(a+ δ) ≤ a+ δ then ϕ ([a, a + δ]) ⊂ [a, a+ δ]. This is a contradiction. In
fact, for every x > a the sequence (ϕ2n(x))n is increasing and unbounded.
Theorem 2.8. Let ϕ be a polynomial with degree greater than one and
having fixed points. Then,
D \ {0} ⊂ σ(Cϕ).
Proof. (a) First we consider the case that
ϕ(x) = +∞. We fix λ ∈
lim
x→+∞
D \ {0} and assume that λ /∈ σ(Cϕ), hence λ 6= 1. Let a ∈ R be given with
the property that ϕ(a) = a and ϕ(x) > x for x > a. Since ϕ′(a) ≥ 1 then ϕ is
strictly increasing in some interval [a, a + δ]. Let ψ : [a, ϕ(a + δ)] → [a, a + δ]
be the inverse of ϕ : [a, a + δ] → [a, ϕ(a + δ)]. Since ϕ(a + δ) > a + δ then
ψk, the k-th iterate of ψ, is well defined for every k ∈ N.
We fix x0 ∈ (a, a + δ) and define xk = ψk(x0). Then (xk)k is a decreasing
sequence converging to a. Let J0 be a closed interval contained in (x1, x0)
and take a smooth function g whose support is contained in (x1, x0) and
satisfying g(x) = 1 for all x ∈ J0. Then, there is a unique f ∈ S(R) such that
f (ϕ(x)) − λf (x) = g(x), x ∈ R.
After iterating this identity we obtain
f (ϕn(x)) = λnf (x) +
λn−1−kg (ϕk(x)) .
n−1
Xk=0
(4)
(5)
8
Let M > 0 be as in Lemma 2.6. For each x > a the sequence (ϕn(x))n
diverges to infinity, so there is m ∈ N with ϕn(x) > M and it easily follows
that
lim
n
1
λn f (ϕn(x)) = 0.
We conclude
f (x) = −
1
λk+1 g(ϕk(x))
(6)
∞
Xk=0
for all x > a. Finally, we fix y0 ∈ J0 and define yk = ψk(y0) ∈ (xk+1, xk).
We have ϕm(ym) = y0 ∈ J0, while ϕk(ym) = ϕk−m(y0) > x0 for k > m and
ϕk(ym) = ψm−k(y0) < x1 for k < m. Consequently
f (ym) = −λ−m−1 while f (a) = 0.
The last identity follows from (4) using λ 6= 1. Since
lim
m
f (ym) 6= f (a)
we get a contradiction.
(b) To deal with the case that
ϕ(x) = −∞ we have to consider two
lim
x→+∞
possibilities, depending on whether the degree of the polynomial is even or
odd.
(i) First case: the degree of ϕ is even. Since ϕ is linearly equivalent to
ψ(x) = −ϕ(−x) and lim
x→+∞
ψ(x) = +∞ then
D \ {0} ⊂ σ(Cψ) = σ(Cϕ)
and we are done.
(ii) Second case: the degree of ϕ is odd. Then lim
x→+∞
ϕ2(x) = +∞. As
above, ϕ2 is strictly increasing in some interval [a, a + δ], where a is the
greatest fixed point of ϕ2. Moreover, we can take δ small enough so that
ϕ(x) < a for every x ∈ (a, a + δ]. This is obvious in the case that ϕ(a) < a.
Otherwise, ϕ(a) = a and ϕ′(a) < 0 (Lemma 2.7) and we can take δ so that
ϕ is decreasing on [a, a + δ], hence ϕ(x) < ϕ(a) = a for every x ∈ (a, a + δ].
Now, we denote by ψ the inverse of ϕ2 : [a, a + δ] → [a, ϕ2(a + δ)].
Proceeding as in (a), we fix x0 ∈ (a, a + δ) and define xk = ψk(x0).
Let J0 be a closed interval contained in (x1, x0) and take a compactly sup-
ported smooth function g whose support is contained in (x1, x0) and satisfy-
ing g(x) = 1 for all x ∈ J0. As in (a), there is f ∈ S(R) such that equation (6)
holds for x > a. Finally, we fix y0 ∈ J0 and define yk = ψk(y0) ∈ (xk+1, xk).
9
We have ϕ2m(ym) = y0 ∈ J0, ϕ2k(ym) > x0 for k > m, ϕ2k(ym) < x1 for
k < m. Moreover ϕ2k+1(ym) /∈ (a, a + δ] since otherwise ϕ2k+2(ym) < a,
which is a contradiction. Consequently
f (ym) = −λ−2m−1 while f (a) = 0.
The same argument as in case (a) gives a contradiction.
Corollary 2.9. Let ϕ be a polynomial of degree greater than one. Then Cϕ
is mean ergodic if and only if σ(Cϕ) = {0}.
Proof. Apply [7, Theorem 3.11] and Theorems 2.5 and 2.8.
Theorem 2.10. Let ϕ be a polynomial of degree greater than one and having
a fixed point a such that ϕ′(a) > 1 and ϕ(n)(a) ≥ 0 for all n ≥ 2. Then
C \ {0} ⊂ σ (Cϕ) .
Proof. Since
ϕ(x) = a +
ϕ(n)(a)
n!
(x − a)n
∞
Xn=1
then ϕ and all its derivatives are increasing in [a, +∞). An inductive ar-
gument using Fa`a di Bruno formula implies that also ϕ(n)
is increasing in
[a, +∞) for every k, n ∈ N0. We observe that ϕ(x) > a + ϕ′(a)(x − a) > x
for any x > a. Hence a is the largest fixed point of ϕ.
k
We already know that D \ {0} ⊂ σ (Cϕ) . We now fix λ > 1 and assume
that λ /∈ σ(Cϕ). We fix x0 > a and define xk+1 = ψ(xk), where ψ stands for
the inverse of ϕ : [a, +∞) → [a, +∞). Then (xk)k is a decreasing sequence
converging to a. We put Ik = (xk+1, xk) , so that Ik = ψk(I0), and let J0 be
a closed subinterval of I0 and Jk := ψk(J0). Finally, we consider a compactly
supported smooth function g whose support is contained in I0 and such that
g(x) = x for every x ∈ J0. Then there is f ∈ S(R) satisfying Cϕf − λf = g.
Hence
f (ϕn(x)) = λnf (x) +
n−1
Xk=0
λn−1−kg (ϕk(x)) ∀n ∈ N, x ∈ R.
Since λ > 1 and (f (ϕn(x)))n is a bounded sequence then
f (x) = −
1
λ
∞
Xj=0
λ−jg (ϕj(x)) ∀x ∈ R.
10
For every x ∈ Jk we have ϕk(x) ∈ J0 while ϕn(x) /∈ I0 for every n 6= k.
Consequently
f (x) = −λ−k−1g (ϕk(x)) = −
ϕk(x)
λk+1 ∀x ∈ Jk.
In order to obtain a contradiction we proceed as follows. Our hypothesis and
Fa`a di Bruno formula permit to conclude
ϕ(n)
m+1(a) ≥ ϕ(n)
m (a) · ϕ′(a)n.
(7)
We select n0 ∈ N so that ϕ′(a)n0 > λ. We can find n > n0 and m ∈ N such
that ϕ(n)
m (a) 6= 0. Otherwise, every iterate ϕm would have degree less than or
equal n0, which is a contradiction. From (7) we get
ϕ(n)
k+m(a) ≥ ϕ(n)
m (a) · ϕ′(a)kn ∀k ∈ N.
Finally, for every x ∈ Jk+m we obtain, with C = ϕ
(n)
m (a)
λ1+m ,
(cid:12)(cid:12)f (n)(x)(cid:12)(cid:12) =
ϕ(n)
k+m(a)
λk+m+1
ϕ(n)
k+m(x)
λk+m+1 ≥
λ (cid:19)k
≥ C(cid:18)ϕ′(a)n
.
We conclude that f (n) is not a bounded function, which is a contradiction.
The following result will be useful in the proof of Proposition 4.2.
Proposition 2.11. Let η be a polynomial with odd degree and negative
leading coefficient such that ϕ = η ◦ η satisfies the hypothesis in Theorem
2.10. Then C \ {0} ⊂ σ(Cη).
Proof. Since η has odd degree then it has fixed points and we can apply
Theorem 2.8 to get D \ {0} ⊂ σ (Cη) . We now fix λ > 1 and assume that
λ /∈ σ(Cη). We observe that a is the largest fixed point of ϕ. From Lemma
2.7 we get δ > 0 such that η(x) < a for all x ∈ (a, a + δ). Now we fix
x0 ∈ (a, a + δ) and define g as in the proof of Theorem 2.10. Then there is
f ∈ S(R) such that
f (x) = −
1
λ
∞
Xn=0
λ−ng (ηn(x)) ∀x ∈ R.
11
f (x) = −
1
λ
λ−2jg (ϕj(x)) ∀x ≥ a.
∞
Xj=0
We observe that η2j = ϕj and g (η2j+1(x)) = 0 for all j ∈ N0 and x ≥ a
(otherwise η2j+1(x) ∈ (a, a + δ) and η2j+2(x) < a, which is a contradiction).
Then
Now we proceed as in Theorem 2.10 to get a contradiction.
As an application of Theorem 2.10 and Proposition 2.3 we have the fol-
lowing.
Example 2.12. Let ϕ(x) = xp, p ≥ 2. Then σ(Cϕ) = C.
Example 2.13. Let ϕ be a polynomial of degree N > 1 with positive leading
coefficient and complex fixed points z1, . . . , zN such that
1. zN ∈ R and the multiplicity of zN as a fixed point is 1.
2. Re(zk) ≤ zN for k < N.
Then C \ {0} ⊂ σ(Cϕ).
In fact, we can apply Theorem 2.10 taking a = zN .
3 Quadratic polynomials
Next we apply the previous results to discuss the spectrum of Cϕ in the case
that ϕ is a quadratic polynomial. Such a polynomial ϕ(x) = a0 + a1x + a2x2
2 − a2
(a2 6= 0) is linearly equivalent to ψ(x) = x2 + c where c = a0a2 + a1
4 .
In fact, take ℓ(x) = ax + b where a = a2, b = a1
2 . It is routine to check that
ϕ = ℓ−1 ◦ ψ ◦ ℓ. We observe that 0 ∈ σ(Cψ) = σ(Cϕ) since the range of Cψ
consists of even functions.
1
c > 1
4 implies that ϕ and ψ lack fixed points, hence σ(Cϕ) = {0} (Theorem
2.5). In the case c < 1
4 we have that ϕ (and also ψ) has two different fixed
points and we can apply Theorem 2.10 (see also Example 2.13) to conclude
that σ(Cϕ) = C.
Our next aim is to discuss the case c = 1
4.
Lemma 3.1. Let ϕ(x) = x2 + 1
C > 0 and p ∈ N such that
4 be given. Then, for every r > 1 there exist
ϕ′m(x) ≤ Crm (1 + ϕm(x))p .
12
Proof. We first observe that ϕ′(x) ≤ 2ϕ(x) with equality for x = ± 1
also
2 and
Proceeding by recurrence we conclude
ϕ′m+1(x) = 2ϕm(x)ϕ′m(x).
ϕ′m(x) = 2m
ϕj(x),
m−1
Yj=0
(8)
(9)
where ϕ0(x) = x.
Since every ϕm (m ≥ 1) is even and ϕ′m is odd we only need to consider
the case x ≥ 0. Now we proceed in several steps.
(i) For 0 ≤ x ≤ 1
2 we have ϕm(x) ≤ ϕm( 1
2 ) = 1
2 . An induction argument
gives
ϕ′m(x) ≤ ϕ′m(
1
2
) = 1.
(ii) For x ≥ x0 := 1 +
√3
2 we have ϕ′(x) ≤ ϕ(x). Since ϕ(x) > x then also
2ϕm(x) ≤ ϕm+1(x) for every m ∈ N. We check that
ϕ′m(x) ≤ ϕ2
m(x) ∀m ∈ N, x ≥ x0.
In fact, this inequality is obvious for m = 1 and assuming that it is true for
m we obtain
ϕ′m+1(x) = 2ϕm(x)ϕ′m(x) ≤ ϕm+1(x)ϕ2
m(x) ≤ ϕ2
m+1(x).
(iii) Take m0 ∈ N such that ϕm0( r
2) ≥ x0. Then, for every x ≥ r
2 and
m ≥ m0 we put
where ϕm0(x) ≥ x0. Hence
ϕm(x) = ϕm−m0 (ϕm0(x)) ,
ϕ′m(x) = ϕ′m−m0 (ϕm0(x)) · ϕ′m0(x) ≤ (1 + ϕm−m0 (ϕm0(x)))2 · ϕ′m0(x).
From (9) we obtain ϕ′m0(x) ≤ 2m0 (1 + ϕm(x))m0 . Finally, for p = m0 + 2 we
conclude
ϕ′m(x) ≤ 2m0 (1 + ϕm(x))p ∀m ≥ m0, x ≥
Hence we can find C > 0 such that
r
2
.
ϕ′m(x) ≤ C (1 + ϕm(x))p ∀m ∈ N, x ≥
r
2
.
13
(iv) We now consider 1
2 ≤ x < r
2 and select nx ≥ 1 with the property that
ϕj(x) ≥ r
2 whenever j ≥ nx while ϕj(x) < r
2 for 0 ≤ j < nx.
If m < nx+1 then, from (9), we get ϕ′m(x) ≤ rm. Otherwise we decompose
ϕ′m(x) =
(2ϕj(x)) ·
nx−1
Yj=0
(2ϕj(x))
m−1
Yj=nx
The first factor is dominated by rnx ≤ rm, while the second one coincides
with
2m−nx
ϕk (ϕnx(x)) = ϕ′m−nx (ϕnx(x)) .
m−1−nx
Yk=0
Since ϕnx(x) ≥ r
2 we can use the estimates in (iii) to conclude
ϕ′m(x) ≤ Crm (1 + ϕm−nx (ϕnx(x)))p = Crm (1 + ϕm(x))p .
Lemma 3.2. Let ϕ(x) = x2 + 1
there exist C > 0 and p ∈ N such that
4 be given. Then, for every r > 1 and n ∈ N
m (x)(cid:12)(cid:12) ≤ Crm (1 + ϕm(x))p .
(cid:12)(cid:12)ϕ(n)
Proof. It is enough to consider x ≥ 0. The case n = 1 is the content of the
previous Lemma. Let us now consider n = 2. From (9) we obtain, for every
x ≥ 1
2,
ϕ′′m(x) =
Hence
m−1
Xj=0
2ϕi(x) = ϕ′m(x)
2ϕ′j(x)Yi6=j
ϕ′j(x)
ϕj(x)
.
(10)
m−1
Xj=0
ϕ′′m(x) ≤ ϕ′m(x)
2ϕ′j(x).
m−1
Xj=0
We now fix r > 1 and take C > 0 and p ∈ N such that
Then, for every x ≥ 1
2,
(cid:12)(cid:12)ϕ′j(x)(cid:12)(cid:12) ≤ Crj (1 + ϕj(x))p ∀j ∈ N0, x ∈ R.
ϕ′′m(x) ≤ 2C 2rm (1 + ϕm(x))2p
rj
m−1
Xj=0
≤ 2
C 2
r − 1
r2m (1 + ϕm(x))2p .
14
Now, an application of Leibnitz rule and Fa`a di Bruno formula permits to
proceed by induction in order to prove the desired result for x ≥ 1
2 .
On the other hand, as ϕ(n)
m is increasing in [0, +∞) then for all x ∈ [0, 1
2]
we have
0 ≤ ϕ(n)
m (x) ≤ ϕ(n)
m (
1
2
) ≤ Crm(cid:18)1 + ϕm(
1
2
)(cid:19)p
Since r > 1 is arbitrary we conclude that for every r > 1 there exist C > 0
and q ∈ N such that ϕ′′m(x) ≤ Crm (1 + ϕm(x))q whenever x ≥ 1
2.
For n > 2 we apply (10) to get
m (x)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)ϕ(n)
(ϕ′′m)(n−2) (x)(cid:12)(cid:12)(cid:12)
≤
n−2
Xk=0(cid:18)n − 2
k (cid:19)(cid:12)(cid:12)ϕ(k+1)
m
(x)(cid:12)(cid:12)
m−1
Xj=0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ϕj(cid:19)(n−2−k)
(cid:18)ϕ′j
.
(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
for some p and C > 0 which only depend on n. Since ϕm( 1
done.
2) = 1
2 we are
Theorem 3.3. Let ϕ(x) = x2 + 1
4 be given. Then σ (Cϕ) = D.
Proof. Since ϕ admits a fixed point and Cϕ is not injective then D is contained
in σ (Cϕ) by Theorem 2.8. To finish we show that Cϕ − λI is invertible for
every λ > 1.
(a) Cϕ − λI is injective for λ > 1.
In fact, Cϕ(f ) = λf implies
f (ϕn(x)) = λnf (x) for every x ∈ R and n ∈ N. Since the left hand side
is bounded and λ > 1 then f (x) = 0 for every x ∈ R.
(b) Cϕ − λI is surjective for λ > 1. It suffices to show that
f ◦ ϕm
λm
∞
Xm=0
converges in S(R) for every f ∈ S(R) and λ ∈ C with λ > 1. Obviously
x ≤ 1 + ϕm(x) for all x ∈ R and m ∈ N. By Lemmas 3.1 and 3.2, ϕ satisfies
the hypothesis in Lemma 2.4 and we conclude.
Summarizing, we get the following.
Theorem 3.4. Let ϕ(x) = a0 + a1x + a2x2 be a quadratic polynomial with
real coefficients and take c = a0a2 + a1
1
2 − a2
4 .
(a) c > 1
4 implies σ (Cϕ) = {0} .
(b) c = 1
4 implies σ (Cϕ) = D.
(c) c < 1
4 implies σ (Cϕ) = C.
15
4 Cubic polynomials
Let us now consider polynomials ϕ of degree 3 with lim
x→+∞
ϕ(x) = +∞.
Theorem 4.1. Let ϕ be a polynomial of degree 3 with positive leading
coefficient. Then C \ {0} ⊂ σ(Cϕ) unless ϕ has a fixed point of multiplicity
3 in which case σ(Cϕ) = D \ {0}.
Proof. According to its fixed points the following cases can occur:
(i) ϕ has three different real fixed points,
(ii) ϕ has two different real fixed points, one with multiplicity two and the
other is simple,
(iii) ϕ has only one real fixed point, the other two being complex conjugate
numbers,
(iv) ϕ has one real fixed point of multiplicity 3.
In the first three cases, using that ϕ is linearly equivalent to ψ(x) =
−ϕ(−x) if necessary, we may apply Theorem 2.10 (see Example 2.13) to
conclude that C \ {0} ⊂ σ(Cϕ). In the forth case, ϕ is linearly equivalent to
ψ(x) = x + x3.
So, to complete the proof we discuss the spectrum of ϕ(x) = x + x3. From
Theorem 2.8 and Proposition 2.3 we already know that σ(Cϕ) ⊃ D \ {0} and
that 0 /∈ σ(Cϕ).
We will show that given n ≥ 1 and R > 1 there are C > 0 and q ∈ R
such that
ϕ(n)
m (x) ≤ CRm(1 + ϕm(x))q
(11)
for each x ∈ R and each m ∈ N. Since ϕ is an odd function, it suffices to
consider x ≥ 0. First, we check the inequality (11) for the first derivative.
Observe that ϕ′m+1(x) = ϕ′m(x)ϕ′(ϕm(x)). Then,
ϕ′m(x) =
m−1
Yk=0
ϕ′(ϕk(x)).
For x = 0 we have the inequality with q = C = 1. We will proceed by
induction on m. For m = 1 we clearly have the inequality for some C > 0
and q = 1. Also, we have ϕ′(x) = 1 + 3x2 ≥ 1. We take x0 > 0 such that
for x ≥ x0, we have ϕ′(x) < ϕ(x). Observe that this implies x0 > 1 therefore
1 + 3x2 < (1 + x2)2. As ϕ(x) > x for x > 0, we also have
1 + 3ϕm(x)2 <(cid:0)1 + ϕm(x)2(cid:1)2
16
for x ≥ x0. Hence, if we assume that for x ≥ x0 we have ϕ′m(x) ≤ (ϕm(x))2,
we obtain
ϕ′m+1(x)
(ϕm+1(x))2 =
ϕ′m(x)
(ϕm(x))2 ·
1 + 3ϕm(x)2
(1 + ϕm(x)2)2 ≤ 1.
Consequently
ϕ′m(x) ≤ (ϕm(x))2 ∀m ∈ N, x ≥ x0.
We now claim that for each K > 0 there is q ∈ N such that
ϕ′m(x) ≤ (1 + ϕm(x))q,
whenever x > K. In fact, for each K > 0 we take ℓ ∈ N such that ϕℓ(K) > x0.
Then, for x ≥ K and m > ℓ we have
ϕ′m(x) = ϕ′m−ℓ (ϕℓ(x)) · ϕ′ℓ(x)
ℓ−1
≤ (ϕm−ℓ (ϕℓ(x)))2 ·
Yj=0(cid:0)1 + 3ϕj(x)2(cid:1)
≤ 3ℓ (ϕm(x))2 · (1 + ϕm(x))2ℓ .
We take p ∈ N such that 3ℓ ≤ (1 + ϕ(K))p and put q = 2ℓ + p + 2. Then
ϕ′m(x) ≤ (1 + ϕm(x))q
for all x > K. The claim is proved. To finish the proof of (11) for n = 1 we
fix R > 0 and take K > 0 such that x > 0 and ϕ′(x) > R imply x > K. For
any x > 0 let nx ∈ N be the first n ∈ N with the property that ϕn(x) > K.
Then m ≤ nx implies
ϕ′m(x) =
m−1
Yj=0
ϕ′ (ϕj(x)) ≤ Rm,
while for m > nx we have
ϕ′m(x) =
nx−1
Yj=0
ϕ′ (ϕj(x)) ·
ϕ′ (ϕj(x))
m−1
Yj=nx
≤ Rnx
m−1−nx
Yj=0
ϕ′ (ϕj(ϕnx(x)))
= Rnxϕ′m−nx (ϕnx(x)) ≤ Rm (1 + ϕm(x))q .
17
For the second derivative we have,
ϕ′′m(x) =
m−1
Xk=0
6ϕk(x)2ϕ′(x)Yj6=k
ϕ′(ϕj(x)) = ϕ′m(x)Xk
ϕ′k(x)
6ϕk(x)2
1 + 3ϕk(x)2 .
From here, we argue as in the case of p(x) = x2 + 1
4 to conclude.
We observe that each polynomial ϕ of degree 3 is linearly equivalent to
some polynomial of the form
ψ(x) = ± x3 + Ax + B.
For ϕ(x) = x3 + Ax + B the spectrum of Cϕ is already discussed in
Theorem 4.1. Next we include some partial results concerning the spectrum
of Cϕ for ϕ(x) = −x3 + Ax + B. In the special case that B = 0 we have a
complete characterization.
Proposition 4.2. Let η(x) = −x3 + Ax be given. Then
(a) A = −1 implies σ(Cη) = D \ {0}.
(b) A < 0, A 6= −1, implies σ(Cη) = C \ {0}.
(c) A ≥ 0 implies σ(Cη) = C.
Proof. Since η has fixed points we can apply Theorem 2.8 to conclude that
D \ {0} ⊂ σ(Cη). From the fact that η is an odd function we have η2 = ω2
where ω(x) = −η(x) = x3 − Ax.
(a) In the case A = −1 we have ω(x) = x3 + x and
σ(Cη2) = σ(Cω2) = D \ {0}.
The last identity follows from Theorem 4.1 and spectral theorem. Conse-
quently σ(Cη) ⊂ D \ {0}.
By Proposition 2.3, in order to show (b) and (c) it suffices to check that
C \ {0} ⊂ σ(Cη) for A 6= −1. We first consider the case A > −1. Then
ω admits a fixed point a > 0 such that ω′(a) > 1 and ω(n)(a) ≥ 0. Hence,
ϕ = η◦η = ω ◦ω satisfies the hypothesis of Theorem 2.10. Since η(x) < 0 < a
for all x ≥ a we can apply Proposition 2.11 to conclude C \ {0} ⊂ σ(Cη).
In the case A < −1 the fixed point a = 0 satisfies ω′(a) > 1 and ω(n)(a) ≥
0. Hence we can proceed as before.
18
For polynomials ϕ(x) = −x3 + Ax + B with B 6= 0 we can provide some
examples.
Proposition 4.3. Let ψ(x) = x3 + Ax+ B be given with B 6= 0 and consider
ϕ(x) = −x3 −Ax−B. We assume the ψ has three different (real) fixed points.
Then
C \ {0} ⊂ σ (Cϕ) .
Proof. We first assume that B > 0. We have ϕ2(x) = ψ2(x) − 2B for every
x ∈ R. Let x = α be the greatest fixed point of ψ. Then ψ′(α) > 1 and
ψ′′(x) ≥ 0 for every x ≥ α. Then x = α is a fixed point of ψ2 and it satisfies
ψ′2(α) > 1 and ψ(n)
2 (x) ≥ 0 for every x ≥ α. Since B > 0 then the equation
ϕ2(x) = x, equivalently ψ2(x) = x + 2B
admits a solution β > α. Then
while
Consequently
ϕ′2(β) = ψ′2(β) ≥ ψ′2(α) > 1,
ϕ(n)
2 (x) = ψ(n)
2 (x) ≥ 0 ∀x ≥ β.
C \ {0} ⊂ σ (Cϕ2) ,
from where the conclusion follows after applying Proposition 2.11.
We now consider the case that B < 0. We put ϕ(x) = −ϕ(−x) and
ψ(x) = −ψ(−x). Then ψ(x) = x3 + Ax − B and ϕ(x) = − ψ(x). Since also ψ
admits three different fixed points and ψ(0) > 0 we conclude
C \ {0} ⊂ σ (C ϕ) = σ (Cϕ) .
Remark 4.4. Let ψ(x) = x3 + Ax + B be given with B > 0. If ψ has a
unique real fixed point then
where c < 0 < α.
ψ(x) = x +(cid:0)(x − α)2 + β2(cid:1) (x − c),
This means that we cannot adapt the previous argument to the case that
there is a unique fixed point. Nor can we adapt the argument in the case
where there is a simple fixed point and a double fixed point (the condition
B > 0 forces that the double fixed point is the greatest).
19
Remark 4.5. If X is a locally convex space and T ∈ L(X), the Waellbrock
spectrum σ∗(T ) is defined as the smallest set containing σ(T ) such that the
resolvent mapping R(·, T ) : C\σ∗(T ) → Lb(X), z 7→ R(z, T ) = (zI −T )−1 is
holomorphic (see [13]). Here Lb(X) stands for the space of continuous linear
operators on X endowed with the topology of convergence on bounded sets. If
X is a Fr´echet space and U ⊂ C is open then F : U → Lb(X) is holomorphic
if and only if the map U → C, z 7→ hu, F (z)(x)i is holomorphic for every
u ∈ X′, x ∈ X (see [6, Theorem 1], [4, corollary 10, Remark 11]). Hence,
from Lemma 2.4 we can get easily that the resolvent map z 7→ R(·, Cϕ) is
holomorphic in C \ {0} when ϕ does not have fixed points (Theorem 2.5),
and also in C \ D when ϕ is a polynomial of degree 2 or 3 with a unique fixed
point (Theorem 3.4, Theorem 4.1 and Propoposition 4.2 a)).
In all cases
then we have σ∗(Cϕ) = σ(Cϕ). Contrary to what happens for operators in
Banach spaces, where the Waellbrock spectrum equals the spectrum which is
always closed, this is not always true for operators defined on Fr´echet spaces,
even when the spectrum is bounded as one can check in [1, Remark 3.5 (vi)].
5 Monotone symbols
We recall that the symbols for S(R) were completely characterized in [5, The-
orem 2.3]. The aim of this section is to provide some information regarding
the spectrum of composition operators defined by monotone symbols. Then,
let us assume that the symbol ϕ is strictly monotone and let us denote by
ψ its inverse and by ψn its n-th iterate. For λ 6= 0, and f, g ∈ S(R), the
relation Cϕf − λf = g implies that (5) holds for every n, that is
f (ϕn(x)) = λnf (x) +
λn−1−kg (ϕk(x)) ,
n−1
Xk=0
which implies
f (x) = λnf (ψn(x)) +
= λnf (ψn(x)) +
n−1
Xk=0
Xj=1
n
λn−1−kg (ψn−k(x))
λj−1g (ψj(x)) .
(12)
Proposition 5.1. Let ϕ be a strictly increasing symbol other than the iden-
tity. Then σ(Cϕ) always contains {λ ∈ C : λ = 1}.
20
Proof. Let λ satisfies λ = 1 and assume that λ /∈ σ(Cϕ). Then, for every
g ∈ S(R) there is a unique f ∈ S(R) such that (4), (5) and (12) hold. Now
we discuss the following possibilities, covering all possible cases
(1) ϕ lacks fixed points.
(2) There exist a < b such that ϕ(a) = a, ϕ(b) = b and ϕ(x) 6= x for every
x ∈ (a, b) .
(3) There exists a ∈ R such that ϕ(a) = a and ϕ(x) 6= x for every x ∈
(a, +∞) .
(4) There exists a ∈ R such that ϕ(a) = a and ϕ(x) 6= x for every x ∈
(−∞, a) .
(1) Since, for every x ∈ R, the sequences (ϕn(x))n and (ψn(x))n diverge,
we obtain from (5) and (12),
f (x) = −
1
λ
∞
Xk=0
λ−kg (ϕk(x)) =
λkg (ψk(x)) .
1
λ
∞
Xk=1
This implies that
λ−kg(ϕk(x)) +
∞
Xk=0
∞
Xk=1
λkg(ψk(x)) = 0
for each x ∈ R. This cannot happen if g is a smooth function whose support
contains 0 and is contained in the open interval determined by ϕ(0) and ψ(0).
In the case that the symbol admits some fixed point then 1 ∈ σ(Cϕ), so
we will take in what follows λ 6= 1.
(2) Either
ϕn(x) ↓ a, ψn(x) ↑ b ∀x ∈ (a, b) (if ϕ(x) < x)
or
ϕn(x) ↑ b, ψn(x) ↓ a ∀x ∈ (a, b) (if ϕ(x) > x).
We fix x0 ∈ (a, b) and define xk+1 = ϕ(xk). Let I0 denote the open interval
with extremes (x0, x1) and let J0 be a closed interval contained in I0 and g a
smooth function with support contained in I0 such that g(x) = 1 for x ∈ J0.
The identity (4) implies that f (a) = f (b) = 0. Then (12) gives
f (x) =
∞
Xj=1
λj−1g (ψj(x)) ∀x ∈ (a, b).
21
Finally we fix y0 ∈ J0, define yk = ϕk(y0) and put c = limk→∞ yk (c = a or
c = b). Then f (yk) = λk−1g (ψk(yk)) = λk−1 and
f (yk) 6= f (c) ,
lim
k→∞
which is a contradiction.
(3) In the case ϕ(x) < x for every x > a we have ϕn(x) ↓ a ∀x > a and
we can proceed as in (2) to get a contradiction. We will discuss the case that
ϕ(x) > x for every x > a. Then
ψn(x) ↓ a while ϕn(x) ↑ +∞ ∀x > a.
From (5) we obtain
f (x) = −
∞
Xk=0
λ−k−1g (ϕk(x)) ∀x > a.
We fix x0 > a and define xk+1 = ψ(xk). Let I0 denote the open interval
(x1, x0) and let J0 be a closed interval contained in I0 and g a smooth function
with support contained in I0 such that g(x) = 1 for x ∈ J0. Finally we fix
y0 ∈ J0, define yk = ψk(y0). Then f (yk) = −λ−k−1g (ϕk(yk)) = −λ−k−1 and
we can proceed as in case (2) to get a contradiction.
(4) is analogous to (3).
For ϕ(x) = x + e−x2, the composition operator Cϕ is not power bounded
but we do not know whether it is mean ergodic or not (see [7, Remark 1]).
According to Proposition 5.1, the spectrum of Cϕ contains the unit circle.
We recall that a fixed point a of ϕ is said to be attracting if ϕ′(a) < 1
and repelling if ϕ′(a) > 1.
Proposition 5.2. Let us assume that a is an attracting fixed point of the
strictly increasing symbol ϕ. Then
{λ ∈ C : ϕ′(a) < λ < 1} ⊂ σ(Cϕ).
Proof. Let us denote by ψ the inverse of the bijection ϕ : [a, +∞) → [a, +∞).
Given λ ∈ D, and f, g ∈ S(R), the relation Cϕf − λf = g, implies that
f (x) =
∞
Xj=1
λj−1g (ψj(x)) , ∀x > a.
(13)
22
We take ϕ′(a) < ε < λ and choose 0 < δ so that ϕ′(x) < ε on (a, a + δ). For
every x ∈ (a, a + δ), by the mean value theorem, we have ϕ(x) < x hence the
sequence (ϕn(x))n decreases to a. In fact,
ϕn(x) − a ≤ εnx − a.
(14)
We fix a < b < a + δ and let J be a closed interval contained in (ϕ(b), b).
We consider a smooth function g whose support is contained in (ϕ(b), b)
and such that g(x) = 1 for every x ∈ J. We check that g cannot be in the
range of Cϕ − λI. We proceed by contradiction and assume that there is is
f ∈ S(R) such that Cϕf − λf = g. We take y0 ∈ J and yk := ϕk(y0). Then
ψj(yk) ∈ (ϕ(b), b) if and only if k = j. Hence, by (13), f (yk) = λk−1. Finally,
using (14),
f (yk) − f (a)
yk − a
≥
λk−1
y0 − aεk ,
(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
which goes to ∞ as k → ∞. This is a contradiction since (yk)k decreases to
a.
Corollary 5.3. Let us assume ϕ(a) = a, ϕ′(a) = 0, ϕ strictly increasing.
Then
{λ ∈ C :
λ ≤ 1} ⊂ σ(Cϕ).
With obvious modifications, one can show
Proposition 5.4. Let us assume that a is a repelling fixed point of ϕ and
ϕ strictly increasing. Then
{λ ∈ C : 1 ≤ λ < ϕ′(a)} ⊂ σ(Cϕ).
Proposition 5.5. Let ϕ be a strictly decreasing symbol. Then
{λ ∈ C :
λ = 1} ⊂ σ(Cϕ)
if and only if ϕ ◦ ϕ 6= I.
Proof. Let us assume ϕ ◦ ϕ 6= I and let a denote the unique fixed point of ϕ.
We proceed by contradiction, so we assume there is λ /∈ σ(Cϕ) with λ = 1.
Several possibilities can occur.
(1) There exist a ≤ b < c such that ϕ2(b) = b, ϕ2(c) = c and ϕ2(x) 6= x
for every x ∈ (b, c). For every smooth function g whose support is contained
in (b, c) there is f ∈ S(R) with f (ϕ(x)) − λf (x) = g(x). Then
f (x) = λ2nf (ψ2n(x)) +
λj−1g (ψj(x)) .
2n
Xj=1
23
Since ψ ([b, c]) ⊂ ψ ([a, +∞)) = (−∞, a] and ψ2k ([b, c]) ⊂ [b, c] we obtain
ψ2k+1 ([b, c]) ⊂ (−∞, a]. Hence
f (x) = λ2nf (ψ2n(x)) +
n
Xj=1
λ2j−1g (ψ2j(x)) ∀x ∈ (b, c).
Now we can argue as in the proof of Proposition 5.1 (case (2)) to get a
contradiction.
(2) There exists b ≥ a such that ϕ2(b) = b and ϕ2(x) 6= x for every x > b.
Since ψ ([b, +∞)) ⊂ ψ ([a, +∞)) ⊂ (−∞, a] and ψ2k ([b, +∞)) ⊂ [b, +∞) we
obtain ψ2k+1 ([b, +∞)) ⊂ (−∞, a]. Hence
f (x) = λ2nf (ψ2n(x)) +
n
Xj=1
λ2j−1g (ψ2j(x)) ∀x > b.
Now we can argue as in the proof of Proposition 5.1 (case (3)) to get a
contradiction. The other possibilities can be treated as (1) or (2).
Finally, let us assume that the unit circle is contained in σ(Cϕ). Since
ϕ) ⊃ (σ(Cϕ))2 then the unit circle is contained in σ(C 2
ϕ), which implies
σ(C 2
ϕ ◦ ϕ 6= I.
If ϕ ◦ ϕ = I then σ(Cϕ) = {−1, 1}. According to [7, Proposition 3.7] and
λ = 1} ⊂ σ(Cϕ) characterizes the
Proposition 5.5, the condition {λ ∈ C :
decreasing symbols ϕ such that Cϕ is mean ergodic.
References
[1] A. Albanese, J. Bonet, W. J. Ricker, Montel resolvents and uniformly
mean ergodic semigroups of linear operators. Quaest. Math. 36 (2013),
2, 253 -- 290.
[2] A. Albanese, J. Bonet, W. J. Ricker, Dynamics and spectrum of the
Ces`aro operator on C∞(R+). Monatsh. Math. 181 (2016), 267 -- 283.
[3] J. Bonet, P. Doma´nski, A note on the spectrum of composition operators
on spaces of real analytic functions. Complex Anal. Oper. Theory 11
(2017), 161 -- 174.
[4] J. Bonet, L. Frerick, E. Jord´a, Extension of vector-valued holomorphic
and harmonic functions. Studia Math. 183 (2007), no. 3, 225 -- 248.
24
[5] A. Galbis, E. Jord´a, Composition operators on the Schwartz space. Rev.
Mat. Iberoam. 34 (2018), 397 -- 412.
[6] K. G. Grosse-Erdmann, A weak criterion for vector-valued holomorphy.
Math. Proc. Cambridge Philos. Soc. 136 (2004), no. 2, 399 -- 411.
[7] C. Fern´andez, A. Galbis, E. Jord´a, Dynamics and spectra of composition
operators on the Schwartz space. J. Funct. Anal. 274 (2018), 3503 -- 3530.
[8] N. Kenessey, J. Wengenroth, Composition operators with closed range
for smooth injective symbols R → Rd. J. Funct. Anal. 260 (2011), 2997 --
3006.
[9] A. Przestacki, Composition operators with closed range for one-
dimensional smooth symbols. J. Math. Anal. Appl. 399 (2013), 225 -- 228.
[10] A. Przestacki, Characterization of composition operators with closed
range for one-dimensional smooth symbols. J. Funct. Anal. 266 (2014),
5847 -- 5857.
[11] A. Przestacki, Corrigendum to "Characterization of composition oper-
ators with closed range for one-dimensional smooth symbols" J. Funct.
Anal. 266 (2014) 5847 -- 5857]. J. Funct. Anal. 269 (2015), 2665 -- 2667.
[12] A. Przestacki, Dynamical properties of weighted composition operators
on the space of smooth functions. J. Math. Anal. Appl. 445 (2017),
1097 -- 1113.
[13] F. H Vasilescu, Analytic functional calculus and spectral decompositions.
Translated from the Romanian. Mathematics and its Applications (East
European Series), 1. D. Reidel Publishing Co., Dordrecht (1982).
25
|
1811.06865 | 2 | 1811 | 2019-05-30T10:08:10 | When are full representations of algebras of operators on Banach spaces automatically faithful? | [
"math.FA"
] | We examine the phenomenon when surjective algebra homomorphisms between algebras of operators on Banach spaces are automatically injective. In the first part of the paper we shall show that for certain Banach spaces $X$ the following property holds: For every non-zero Banach space $Y$ every surjective algebra homomorphism $\psi: \, \mathcal{B}(X) \rightarrow \mathcal{B}(Y)$ is automatically injective. In the second part of the paper we consider the question in the opposite direction: Building on the work of Kania, Koszmider and Laustsen \textit{(Trans. London Math. Soc., 2014)} we show that for every separable, reflexive Banach space $X$ there is a Banach space $Y_X$ and a surjective but not injective algebra homomorphism $\psi: \, \mathcal{B}(Y_X) \rightarrow \mathcal{B}(X)$. | math.FA | math |
WHEN ARE FULL REPRESENTATIONS OF ALGEBRAS
OF OPERATORS ON BANACH SPACES
AUTOMATICALLY FAITHFUL?
BENCE HORVÁTH
Abstract. We examine the phenomenon when surjective algebra ho-
momorphisms between algebras of operators on Banach spaces are au-
tomatically injective. In the first part of the paper we shall show that
for certain Banach spaces X the following property holds: For every
non-zero Banach space Y every surjective algebra homomorphism ψ :
B(X) → B(Y ) is automatically injective. In the second part of the pa-
per we consider the question in the opposite direction: Building on the
work of Kania, Koszmider, and Laustsen (Trans. London Math. Soc.,
2014) we show that for every separable, reflexive Banach space X there
is a Banach space YX and a surjective but not injective algebra homo-
morphism ψ : B(YX ) → B(X).
1. Introduction and preliminaries
1.1. Introduction. A classical result of Eidelheit (see for example [11, The-
orem 2.5.7]) asserts that if X, Y are Banach spaces then they are isomorphic
if and only if their algebras of operators B(X) and B(Y ) are isomorphic as
Banach algebras, in the sense that there exists a continuous bijective al-
gebra homomorphism ψ : B(X) → B(Y ). It is natural to ask whether for
some class of Banach spaces X this theorem can be strengthened in the
following sense: If Y is a non-zero Banach space and ψ : B(X) → B(Y ) is a
continuous, surjective algebra homomorphism, is ψ automatically injective?
It is easy find an example of a Banach space with this property. Indeed,
let X be a finite-dimensional Banach space, let Y be a non-zero Banach
space and let ψ : B(X) → B(Y ) be a surjective algebra homomorphism.
Since B(X) ≃ Mn(C) for some n ∈ N, simplicity of Mn(C) implies that
Ker(ψ) = {0}. One can also obtain an infinite-dimensional example: If H
be a separable, infinite-dimensional Hilbert space, Y is a non-zero Banach
space, and let ψ : B(H) → B(Y ) is a surjective algebra homomorphism,
then ψ is automatically injective; see the paragraph before the proof of
2010 Mathematics Subject Classification. Primary 46H10, 47L10; Secondary 46B03,
46B07, 46B10, 46B26, 47L20.
Key words and phrases. Banach space, Semadeni space, bounded linear operator,
ideal, semisimple, algebra homomorphism, automatically injective, SHAI property.
1
2
B. HORVÁTH
Theorem 1.3. These simple observations ensure that the following definition
is not vacuous.
Definition 1.1. A Banach space X has the SHAI property (Surjective Ho-
momorphisms Are Injective) if for every non-zero Banach space Y every
surjective algebra homomorphism ψ : B(X) → B(Y ) is automatically in-
jective.
The purpose of this paper is to initiate the study of this property. The
paper is structured as follows.
In the second part of Section 1 we establish our notations and introduce
the necessary background. We begin Section 2 by giving a list of examples of
Banach spaces which lack the SHAI property, see Example 2.4. We continue
by extending our list of examples of Banach spaces with the SHAI property.
Since ℓ2 possesses this property, it is therefore natural to ask the same
question for other classical sequence spaces. We obtain the following result:
Proposition 1.2. Suppose X is one of the Banach spaces c0 or ℓp for
1 ≤ p ≤ ∞. Then X has the SHAI property.
Another way of generalising the ℓ2-case is to ask whether all, not neces-
sarily separable Hilbert spaces have the SHAI property. As we will demon-
strate, the answer is affirmative:
Theorem 1.3. A Hilbert space of arbitrary density character has the SHAI
property.
We shall also provide more "exotic" examples of Banach spaces with
the SHAI property, including Schlumprecht's arbitrarily distortable Banach
space S, constructed in [42]:
Theorem 1.4. Let X be a complementably minimal Banach space such that
it has a complemented subspace isomorphic to X ⊕X. Then X has the SHAI
property. In particular, Schlumprecht's arbitrarily distortable Banach space
S has the SHAI property.
When studying the SHAI property of a Banach space X, understand-
ing the complemented subspaces of X and the lattice of closed two-sided
ideals of B(X) appears to be immensely helpful. For the Banach space
2 )Y , where Y is c0 or ℓ1, the complemented subspace structure
was studied by Bourgain, Casazza, Lindenstrauss, and Tzafriri in [6] and
the ideal lattice of B(X) by Laustsen, Loy, and Read in [27] and later by
Laustsen, Schlumprecht, and Zsák in [28]. Their results allow us to show
the following:
X = (Ln∈N ℓn
SURJECTIVE REPRESENTATIONS OF B(X)
3
Theorem 1.5. Let X := (Ln∈N ℓn
SHAI property.
2 )Y , where Y is c0 or ℓ1. Then X has the
Finally, in Section 2 we establish a permanence property:
Proposition 1.6. Let E be a Banach space and let F, G be closed subspaces
of E with E = F ⊕ G. If both F and G have the SHAI property then E has
the SHAI property.
We remark in passing that the stability of the SHAI property under finite
sums is of interest to us since B(F ⊕G) can have a very complicated lattice of
closed two-sided ideals even if B(F ) and B(G) themselves have the simplest
possible ideal structure, we refer the reader to [14] and [44]. We do not know
however if Lp[0, 1] possesses the SHAI property for p ∈ [1, ∞)\{2}.
Section 3 is devoted entirely to construct Banach spaces which fail the
SHAI property in a rather non-trivial manner; for every separable, reflexive
Banach space X we find a Banach space YX and a surjective but not injective
algebra homomorphism Θ : B(YX) → B(X). More precisely, we prove the
following:
Theorem 1.7. Let X be a non-zero, separable, reflexive Banach space. For
every S ∈ B(YX) there exists a unique Θ(S) ∈ B(X) and there exists a club
subset D ⊆ [0, ω1) such that for all α ∈ D and all ψ ∈ X ∗:
(1.1)
S ∗(δα ⊗ ψ) = δα ⊗ Θ(S)∗ψ.
Moreover, the map Θ : B(YX) → B(X); S 7→ Θ(S) is a non-injective alge-
bra homomorphism of norm one; and there exists an algebra homomorphism
Λ : B(X) → B(YX) of norm one with Θ ◦ Λ = idB(X). In particular Θ is
surjective.
All necessary terminology and notation will be explained in the subse-
quent sections.
1.2. Preliminaries. Our notations and terminology are standard. The set
of natural numbers not including zero will be denoted by N, and N0 :=
N ∪ {0}. The fields of real and complex numbers are denoted by R and C,
respectively.
1.2.1. Banach spaces, their algebras of operators and ideals thereof, Banach
algebras. In what follows, all Banach spaces and Banach algebras are as-
sumed to be over the complex scalar field C. Most of our results extend
however verbatim to Banach spaces over the real scalar field R. Whenever
4
B. HORVÁTH
an argument of ours holds only in the complex case, we emphasize the im-
portance of the choice of the scalar field.
If X is a Banach space then its dual space is X ∗ and h· , ·i is the duality
bracket between X and X ∗. The symbol IX is the identity operator on
X. The symbol B(X, Y ) stands for the Banach space of bounded linear
operators between the Banach spaces X and Y , we let B(X) := B(X, X).
For T ∈ B(X, Y ) its adjoint is T ∗ ∈ B(Y ∗, X ∗). If W, Z are closed linear
subspaces of X and Y , respectively, then for a T ∈ B(X, Y ) we denote the
restriction of T to W by T W , clearly T W ∈ B(W, Y ). If Ran(T ) ⊆ Z then
T Z denotes T considered as a bounded linear operator between X and Z,
that is, T Z ∈ B(X, Z).
The direct sum of Banach spaces X and Y will be denoted by X ⊕ Y .
Two Banach spaces X and Y are said to be isomorphic if there is a linear
homeomorphism between X and Y , it will be denoted by X ≃ Y . If X ≃
X ⊕ X we say that X is isomorphic to its square. Throughout this paper,
whenever two Banach spaces are isometrically isomorphic we shall identify
them when it does not cause any confusion. By an isomorphism of Banach
algebras A and B we understand that there is an algebra homomorphism
between A and B which is also a homeomorphism. This will also be denoted
by A ≃ B.
The symbols A(X), K(X), S(X), E(X), W(X) and X (X) stand for
the closed two-sided ideals of operators which are approximable, compact,
strictly singular, inessential, weakly compact and have separable range, re-
spectively. We recall that T ∈ B(X) is an inessential operator (see [34,
page 489]) if for every S ∈ B(X) it follows that dim(Ker(IX + ST )) < ∞
and codimX (Ran(IX + ST )) < ∞; this is equivalent to saying that IX + ST
is a Fredholm operator for every S ∈ B(X). It is well-known that A(X) ⊆
K(X) ⊆ S(X) ⊆ E(X) and K(X) ⊆ W(X) ∩ X (X) hold, see for example
[8].
A character on a unital Banach algebra A is a unit-preserving algebra
homomorphism from A to C. Any such character is necessarily of norm at
most 1 and therefore continuous.
1.2.2. Idempotents, projections. Let R be a ring. We say that p ∈ R is an
idempotent if p2 = p. If p, q ∈ R are idempotents then we say that they
are mutually orthogonal and write p ⊥ q if pq = 0 = qp. For p, q ∈ R
idempotents we write p ∼ q if there exist a, b ∈ R such that p = ab and
q = ba, in this case we say that p and q are equivalent. If p, q ∈ R are
idempotents, then we write q ≤ p whenever pq = q and qp = q hold. This is
SURJECTIVE REPRESENTATIONS OF B(X)
5
a partial order on the set of idempotents of R. We say that an idempotent
p ∈ R is minimal if it is minimal in the set of non-zero idempotents of R
with respect to this partial order. We write q < p if both q ≤ p and q 6= p
hold.
In a C ∗-algebra A an idempotent p ∈ A is called a projection if it is
self-adjoint. A projection is minimal if it is minimal in the set of non-zero
projections of A with respect to the partial order ≤.
1.2.3. Simple and semisimple algebras. We say that a unital algebra A is
simple if the only non-trivial two-sided ideal in A is A. If A is a unital
algebra, the Jacobson radical of A, denoted by rad(A), is the intersection of
all maximal left ideals in A, and it is a two-sided ideal in A. If there are no
proper left ideals in A we put rad(A) := A. A unital algebra is semisimple if
its Jacobson radical is trivial. For any Banach space X, the Banach algebra
B(X) is well-known to be semisimple but it is not simple whenever X is
infinite-dimensional, since A(X) is a proper non-trivial closed two-sided
ideal in B(X).
2. When surjective algebra homomorphisms are
automatically injective
A classical deep result of B. E. Johnson asserts the following.
Theorem 2.1 (Johnson). If A, B are Banach algebras such that B is semisim-
ple, then every surjective algebra homomorphism ψ : A → B is automati-
cally continuous.
For a modern discussion of this result we refer the reader to [11, The-
orem 5.1.5]. In what follows we shall use this fundamental result without
explicitly mentioning it.
We first observe that there is a large class of Banach spaces which obvi-
ously lack the SHAI property.
Lemma 2.2. Let X be an infinite-dimensional Banach space such that
Mn(C) is a quotient of B(X) for some n ∈ N. Then X does not have the
SHAI property.
Proof. Let ϕ : B(X) → Mn(C) be a surjective algebra homomorphism.
Since B(Cn) ≃ Mn(C) we immediately obtain that that there is a surjective
algebra homomorphism ψ : B(X) → B(Cn) which cannot be injective, since
X is infinite-dimensional.
(cid:3)
6
B. HORVÁTH
Remark 2.3. For any n ∈ N one can easily find an infinite-dimensional
Banach space X such that it satisfies the conditions of Lemma 2.2, that
is, Mn(C) is a quotient of B(X). Indeed, let E be an infinite-dimensional
Banach space such that B(E) has a character ϕ : B(E) → C. (Examples of
i=1 E, then there
is an isomorphism between the Banach algebras B(X) and Mn(B(E)), this
latter being the Banach algebra of (n × n)-matrices with entries in B(E).
Since every element A ∈ B(E) can be written uniquely as A = λIE + T for
some λ ∈ C and T ∈ Ker(ϕ), it is straightforward to check that
such spaces are given below in Example 2.4.) Let X :=Ln
(2.1)
ψ : Mn(B(E)) → Mn(C);
(λi,jIE + Ti,j)n
i,j=1 7→ (λi,j)n
i,j=1
defines surjective algebra homomorphism. So in particular Mn(C) is a quo-
tient of B(X).
In fact, something much stronger can be said then the above: It was
observed by Kania and Laustsen in [23, page 1022] that every complex,
semisimple, finite-dimensional, unital algebra is isomorphic to B(X)/K(X)
for a suitable Banach space X.
We recall that an infinite-dimensional Banach space X is indecomposable,
if there are no closed, infinite-dimensional subspaces Y, Z of X such that
X ≃ Y ⊕Z. A Banach space X is hereditarily indecomposable if every closed,
infinite-dimensional subspace of X is indecomposable.
The next example collects a variety of examples from the literature where
B(X) is known to have a character, so X does not have the SHAI property
by Lemma 2.2. In examples (1) -- (3) this character is shown explicitly and
in examples (4) -- (7) the character is obtained from a commutative quotient
on B(X). It is not intended to be an exhaustive list.
Example 2.4. None of the following spaces X have the SHAI property:
(1) X is a complex hereditarily indecomposable Banach space, since by
[17, Theorem 18] B(X) has a character whose kernel is S(X),
(2) X = Jp where 1 < p < ∞ and Jp is the pth James space, since by
[13, Paragraph 8], B(X) has a character whose kernel is W(X), see
also [24, Theorem 4.16],
(3) X = C[0, ω1], where ω1 is the first uncountable ordinal, since by [13,
Paragraph 9] B(X) has a character, see also [30, Proposition 3.1],
(4) X = C[0, ωη], where η is a regular cardinal, and ωη is the small-
est ordinal of cardinality ℵη, since by [32, Section 4] B(X) has a
character,
SURJECTIVE REPRESENTATIONS OF B(X)
7
(5) X = X∞, where X∞ is the indecomposable but not hereditarily inde-
composable Banach space constructed by Tarbard in [46, Chapter 4],
since B(X)/K(X) ≃ ℓ1(N0), where the right-hand side is endowed
with the convolution product,
(6) X = XK, where K is a countable compact Hausdorff space and XK is
the Banach space construced by Motakis, Puglisi, and Zisimopoulou
in [31, Theorem B], since B(X)/K(X) ≃ C(K),
(7) X = C(K0), where is K0 is the compact Hausdorff connected
"Koszmider" space constructed by Plebanek in [35, Theorem 1.3],
since B(X)/W(X) ≃ C(K0), as shown in [41, Proposition 3.3], and
it also follows from [12, Theorem 6.5(i)],
(8) X = G, where G is the Banach space constructed by Gowers in [16],
since B(X)/S(X) ≃ ℓ∞/c0, as shown in [24, Corollary 8.3].
The purpose of the following lemma is to show for a certain "nice" class
of Banach spaces, when studying the SHAI property it is enough to restrict
our attention to infinite-dimensional spaces Y .
Lemma 2.5. Let X be a Banach space such that X contains a comple-
mented subspace isomorphic to X ⊕ X. Then the following are equivalent:
(1) X has the SHAI property,
(2) for any Y infinite-dimensional Banach space any surjective algebra
homomorphism ψ : B(X) → B(Y ) is automatically injective.
Proof. Let Y be a non-zero Banach space and let ψ : B(X) → B(Y ) be
a surjective algebra homomorphism, we show that Y must be infinite-
dimensional. For assume towards a contradiction it is not; then clearly
B(Y ) is finite-dimensional, thus by B(X)/ Ker(ψ) ≃ B(Y ) we have that
Ker(ψ) is finite-codimensional in B(X). But X has a complemented sub-
space isomorphic to X ⊕ X therefore by successively applying [25, Proposi-
tions 1.9 and 2.3] and [11, Proposition 1.3.34] it follows that B(X) has no
proper ideals of finite codimension, a contradiction.
(cid:3)
We recall that if A, B are unital algebras and θ : A → B is a surjective
algebra homomorphism then θ[rad(A)] ⊆ rad(B).
Lemma 2.6. Let X be a Banach space, let B be a unital Banach algebra
and let ψ : B(X) → B be a continuous, surjective, non-injective algebra
homomorphism. Then ψ[E(X)] ⊆ rad(B). In particular, if B is semisimple
then E(X) ⊆ Ker(ψ).
8
B. HORVÁTH
Proof. Since ψ is not injective A(X) ⊆ Ker(ψ) holds and therefore there
exists a unique surjective algebra homomorphism θ : B(X)/A(X) → B
with θ ◦ π = ψ and kψk = kθk, where π : B(X) → B(X)/A(X) is the
quotient map. Thus θ[rad(B(X)/A(X))] ⊆ rad(B), which by Kleinecke's
theorem [8, Theorem 5.5.9] is equivalent to θ[π[E(X)]] ⊆ rad(B). This is
equivalent to ψ[E(X)] ⊆ rad(B), as required.
(cid:3)
Lemma 2.7. Let X be a Banach space such that E(X) is a maximal ideal
in B(X) and X has a complemented subspace isomorphic to X ⊕ X. Then
X has the SHAI property.
Proof. Let Y be an infinite-dimensional Banach space and let ψ : B(X) →
B(Y ) be a surjective algebra homomorphism. Assume towards a contradic-
tion that ψ in not injective. Since B(Y ) is semisimple in view of Lemma 2.6
it follows that E(X) ⊆ Ker(ψ) must hold. Since ψ is surjective, Ker(ψ)
is a proper ideal thus by maximality of E(X) in B(X) it follows that
Ker(ψ) = E(X). Thus B(X)/E(X) ≃ B(Y ), where the right-hand side
is simple, due to maximality of E(X) in B(X), which is a contradiction.
Therefore ψ must be injective thus by Lemma 2.5 the claim is proven. (cid:3)
Remark 2.8. We observe that the condition "X has a complemented sub-
space isomorphic to X ⊕ X" in the previous lemma cannot be dropped in
general. Indeed, let X be a hereditarily indecomposable Banach space, then
E(X) = S(X) is a maximal ideal in B(X) but by Example 2.4 (1) the space
X does not have the SHAI property.
Proof of Proposition 1.2. Let X be c0 or ℓp for 1 ≤ p < ∞. Gohberg,
Markus, and Feldman showed in [15] that A(ℓp) = K(ℓp) = S(ℓp) = E(ℓp)
is the only closed, non-trivial, proper, two-sided ideal in B(ℓp). In [26,
page 253], Loy and Laustsen deduced that W(ℓ∞) = X (ℓ∞) = S(ℓ∞) =
E(ℓ∞) is the unique maximal ideal in B(ℓ∞). Thus in both cases the result
follows from Lemma 2.7.
(cid:3)
We remark in passing that it was recently shown by W. B. Johnson, G.
Pisier and G. Schechtman in [20, Theorem 4.2] that B(ℓ∞) has continuum
many distinct closed two-sided ideals, thus the use of Lemma 2.7 in the
proof of Proposition 1.2 is essential.
We recall that a Banach space X is called complementably minimal if
every closed, infinite-dimensional subspace of X contains a subspace which
is complemented in X and isomorphic to X.
Proof of Theorem 1.4. Since X is complementably minimal, it follows from
[47, Theorem 6.2] that S(X) is the largest proper two-sided ideal in B(X).
SURJECTIVE REPRESENTATIONS OF B(X)
9
In particlar E(X) = S(X) is maximal in B(X), thus Lemma 2.7 yields the
claim.
We recall that Schlumprecht's space S is isomorphic to it is square and
it is complementably minimal, as shown, for example, in [43], thus the first
part of the theorem applies.
(cid:3)
In the following we show that for a Hilbert space H of arbitrary density
character, the projections lift from any quotient of B(H). In what follows,
if (X, µ) is a measure space and f ∈ L∞(X, µ) then
(2.2)
Mf : L2(X, µ) → L2(X, µ);
g 7→ f g
is called the multiplication operator by f and is clearly a bounded linear
operator.
Lemma 2.9. Let H be a Hilbert space and let J be a closed, two-sided
ideal in B(H). For any projection p ∈ B(H)/J there exists a projection
P ∈ B(H) such that p = π(P ), where π : B(H) → B(H)/J denotes the
quotient map.
Proof. Let p ∈ B(H)/J be a projection. There exists a self-adjoint A ∈
B(H) such that p = π(A). By the spectral theorem for bounded self-adjoint
operators [10, Chapter IX., Theorem 4.6] there exists a measure space
(X, µ), a µ-almost everywhere bounded, real-valued function f on X and
an isometric isomorphism U : H → L2(X, µ) such that A = U −1 ◦ Mf ◦ U.
Consequently
(2.3)
π(U −1 ◦ Mf ◦ U) = π(A) = p = p2 = π(A2) = π(U −1 ◦ Mf 2 ◦ U),
which is equivalent to
(2.4)
U −1 ◦ Mf −f 2 ◦ U = U −1 ◦ (Mf − Mf 2) ◦ U ∈ J .
Let f be a representative of the class f and let h be the class of 1
[ f ≥1/2],
the indicator function of the set [ f ≥ 1/2] := {x ∈ X : f (x) ≥ 1/2}.
Clearly h ∈ L∞(X, µ) is well-defined and P := U −1 ◦ Mh ◦ U ∈ B(H) is
a projection. We show that p = π(P ), which is equivalent to showing that
U −1 ◦ Mf −h ◦ U ∈ J . We first observe that it is enough to find g ∈ L∞(X, µ)
such that g(f − f 2) = h − f. Indeed, if such a function g were to exist then
Mg ◦ Mf −f 2 = Mh−f and consequently
U −1 ◦ Mh−f ◦ U = U −1 ◦ Mg ◦ Mf −f 2 ◦ U
(2.5)
= (U −1 ◦ Mg ◦ U) ◦ (U −1 ◦ Mf −f 2 ◦ U) ∈ J
holds by Equation (2.4) and the fact that J is an ideal in B(H).
10
B. HORVÁTH
Thus let g : X → R be the following function:
(2.6)
g(x) :=(cid:26) 1/( f (x) − 1)
1/ f (x)
if f (x) < 1/2
otherwise.
Let g be the class of g, clearly g is µ-almost everywhere bounded by 2. A
simple calculation shows that
(2.7)
g(x)( f (x) − f 2(x)) =(cid:26) ( f (x) − f 2(x)/( f (x) − 1)
so g(x)( f (x) − f 2(x)) = 1
quently g(f − f 2) = h − f, which proves the claim.
( f (x) − f 2(x)/ f (x)
[ f ≥1/2](x) − f (x) holds for every x ∈ X. Conse-
(cid:3)
if f (x) < 1/2
otherwise,
We recall that in a ring R if I E R is a two-sided ideal and p, q ∈ R are
idempotents with p ∼ q then p ∈ I if and only if q ∈ I. In a C ∗-algebra A
an idempotent e ∈ A is a projection if and only if kek ≤ 1.
The following lemma is straightforward, we omit its proof.
Lemma 2.10.
(1) Let X be a Banach space and suppose Q ∈ B(X) is an idempotent
such that Ran(Q) is isomorphic to its square. Then there exist mu-
tually orthogonal idempotents Q1, Q2 ∈ B(X) with Q1, Q2 ∼ Q and
Q1 + Q2 = Q.
(2) Let H be a Hilbert space and suppose Q ∈ B(H) is a projection
with infinite-dimensional range. Then there exist mutually orthogo-
nal projections Q1, Q2 ∈ B(H) with Q1, Q2 ∼ Q and Q1 + Q2 = Q.
Corollary 2.11.
(1) Let X be a Banach space and let J E B(X) be a closed, two-sided
ideal. Suppose Q ∈ B(X) is an idempotent such that Ran(Q) is
isomorphic to its square and Q /∈ J . Then there exist mutually
orthogonal idempotents Q1, Q2 ∈ B(X) with Q1, Q2 /∈ J such that
π(Q1), π(Q2) < π(Q), where π : B(X) → B(X)/J is the quotient
map.
(2) Let H be a Hilbert space and let J E B(H) be a closed, two-sided
ideal. Suppose Q ∈ B(H) projection such that Q /∈ J . Then there
exist mutually orthogonal projections Q1, Q2 ∈ B(H) with Q1, Q2 /∈
J such that π(Q1), π(Q2) < π(Q), where π : B(H) → B(H)/J is
the quotient map.
Proof. (1) By Lemma 2.10 (1) there exist mutually orthogonal idempotents
Q1, Q2 ∈ B(X) with Q1 + Q2 = Q and Q1, Q2 ∼ Q. For i ∈ {1, 2} we
SURJECTIVE REPRESENTATIONS OF B(X)
11
immediately get Qi ≤ Q and thus π(Qi) ≤ π(Q). Since Qi ∼ Q, the con-
dition Q /∈ J is equivalent to Qi /∈ J . Also, for i, j ∈ {1, 2} if i 6= j then
Qj = Q − Qi thus π(Qi) 6= π(Q).
(2) Immediate from the first part of this corollary and Lemma 2.10 (2).
(cid:3)
We recall a folklore lifting result for "Calkin" algebras of Banach spaces,
this will be essential in the proof of Theorem 1.5. A convenient reference
for the proof of this lemma is [4, Lemma 2.6]. It also follows from the much
more general result [2, Theorem C].
Lemma 2.12. Let X be a Banach space and let p ∈ B(X)/K(X) be an
idempotent. Then there exists an idempotent P ∈ B(X) with p = π(P )
where π : B(X) → B(X)/K(X) is the quotient map.
Proposition 2.13.
(1) Let X be a Banach space such that every infinite-dimensional com-
plemented subspace of X is isomorphic to its square.
Then B(X)/K(X) does not have minimal idempotents.
(2) Let H be a Hilbert space and let J E B(H) a non-zero, closed,
two-sided ideal. Then B(H)/J does not have minimal projections.
Proof. (1) Let p ∈ B(X)/K(X) be a non-zero idempotent. By Lemma 2.12
there exists an idempotent P ∈ B(X) with p = π(P ), where π : B(X) →
B(X)/K(X) is the quotient map. Clearly P /∈ K(X), equivalently Ran(P ) is
infinite-dimensional. Thus by the hypothesis it is isomorphic to its square,
consequently Corollary 2.11 (1) implies that there exists an idempotent
Q ∈ B(X) such that Q /∈ I and π(Q) < π(P ).
(2) Let p ∈ B(H)/J be a non-zero projection. By Lemma 2.9 there
exists a projection P ∈ B(H) with p = π(P ), where π : B(H) → B(H)/J
is the quotient map. Clearly P /∈ J thus by Corollary 2.11 (2) there exists
a projection Q ∈ B(H) such that Q /∈ J and π(Q) < π(P ).
(cid:3)
We show that Proposition 2.13 (2) can be strengthened with the aid of
the following simple observation. It is certainly well known among experts,
however, we could not locate its proof in the literature, thus we include it
here for the convenience of the reader.
Lemma 2.14. If a C ∗-algebra has minimal idempotents then it has minimal
projections.
Proof. Let A be a C ∗-algebra and suppose e ∈ A is a minimal idempotent.
By [40, Exercise 3.11(i)] there exists a projection p ∈ A with p ∼ e. Thus
12
B. HORVÁTH
there exist a, b ∈ A such that ab = p and ba = e, consequently ae = pa
and bp = eb. We show that p ∈ A is a minimal projection. Since e 6= 0 it
is clear that p 6= 0. Let q ∈ A be a non-zero projection with q ≤ p, this
is, pq = q and qp = q. We define f := bqa, and observe that f ∈ A is a
non-zero idempotent. Indeed, f 2 = bqabqa = bqpqa = bqa = f and f 6= 0
otherwise 0 = af b = abqab = pqp = q which is impossible. Let us observe
that f ≤ e. Indeed, ef = ebqa = bpqa = bqa = f and similarly f e = f holds.
Since e ∈ A is a minimal idempotent it follows that e = f and consequently
aeb = af b holds, equivalently pab = abqab equivalently p = pqp which is
just p = q. This shows that p ∈ A is a minimal projection.
(cid:3)
Corollary 2.15. Let H be a Hilbert space and let J E B(H) be a non-zero,
closed, two-sided ideal. Then B(H)/J does not have minimal idempotents.
Before we prove Theorem 1.3, let us remark here that the case where
H is separable immediately follows from well-known facts. Indeed, let Y
be a non-zero Banach space and let ψ : B(H) → B(Y ) be a continuous,
surjective algebra homomorphism. Since Ker(ψ) is a non-trivial, closed, two-
sided ideal in B(H), by the ideal classification result due to Calkin ([7]),
Ker(ψ) = {0} or Ker(ψ) = K(H) must hold. In the latter case, Cal(H) :=
B(H)/K(H) ≃ B(Y ). (We remark in passing that the ideal of compact
operators K(H) coincides with the operator norm-closure of the ideal of
finite-rank operators on H, since H has a Schauder basis.) Clearly Cal(H)
is simple and infinite-dimensional. If Y is infinite-dimensional, then B(Y ) is
not simple, which is impossible; if Y is finite-dimensional then so is B(Y ),
a contradiction. Thus ψ must be injective.
Proof of Theorem 1.3. Let H be a Hilbert space. Let Y be a Banach space
and assume towards a contradiction that there exists a surjective, non-
injective algebra homomorphism ψ : B(H) → B(Y ). Then Ker(ψ) is non-
zero and B(H)/ Ker(ψ) is isomorphic to B(Y ). This is a contradiction since
B(H)/ Ker(ψ) has no minimal idempotents by Corollary 2.15, whereas B(Y )
clearly does.
(cid:3)
Remark 2.16. In the proof of Theorem 1.3 the spectral theorem played a
key role, hence the use of complex Hilbert spaces was essential. We show
now that the theorem remains true for real Hilbert spaces.
In order to to this, we shall need the notion of the complexification of
real Banach and Hilbert spaces, and real Banach algebras. We refer the
interested reader to [5, Section 13] and [36, Chapter I, Section 3] for the
necessary background information.
SURJECTIVE REPRESENTATIONS OF B(X)
13
Let H be a real Hilbert space of arbitrary density character, we show that
H has the SHAI property. Assume towards a contradiction that there is a
non-zero real Banach space Y and a surjective, non-injective homomorphism
ϕ : B(H) → B(Y ) of real Banach algebras. Let \B(H) and [B(Y ) denote the
complexifications of the real Banach algebras B(H) and B(Y ), respectively.
We define the map
(2.8)
ψ : \B(H) → [B(Y );
(T, S) 7→ (ϕ(T ), ϕ(S)),
this is easily seen to be a surjective homomorphism of complex Banach
algebras. Since ϕ is not injective, there is a non-zero S ∈ B(H) with ϕ(S) =
0. Thus ψ(S, S) = (ϕ(S), ϕ(S)) = (0, 0), hence ψ is not injective. However,
\B(H) ≃ B( H) and [B(Y ) ≃ B( Y ) as complex Banach algebras, thus there
is a surjective, non-injective algebra homomorphism θ : B( H) → B( Y )
of complex Banach algebras. Since H is a complex Hibert space this is
impossible in view of Theorem 1.3. Therefore H has the SHAI property, as
required.
We recall the following piece of notation: If ℓn
2 denotes the n-dimensional
Banach space Cn with the ℓ2-norm, then
(2.9)
ℓn
2!ℓ1
Mn∈N
:=((xn)n∈N : (∀n ∈ N)(xn ∈ ℓn
is a Banach space with the norm k(xn)n∈Nk := Pn∈N
:=n(xn)n∈N : (∀n ∈ N)(xn ∈ ℓn
ℓn
2!c0
Mn∈N
Similarly,
(2.10)
is a Banach space with the norm k(xn)n∈Nk := sup
n∈N
kxnk.
kxnk < ∞)
2 ), Xn∈N
kxnk.
2 ),
lim
n→∞
kxnk = 0o
Example 2.17. For the following (non-Hilbertian) Banach spaces X every
infinite-dimensional complemented subspace of X is isomorphic to its square
therefore by Proposition 2.13 (1) the Calkin algebra B(X)/K(X) does not
have minimal idempotents:
(1) X = c0(λ), where λ is an infinite cardinal, since by [1, Proposi-
tion 2.8] every infinite-dimensional complemented subspace of c0(λ)
is isomorphic to c0(κ) for some infinite cardinal κ ≤ λ, and c0(κ) ≃
c0(κ) ⊕ c0(κ),
(2) X = ℓp where p ∈ [1, ∞)\{2}, since by Pełczyński's theorem ([33])
every infinite-dimensional complemented subspace of ℓp is isomor-
phic to ℓp and ℓp ≃ ℓp ⊕ ℓp,
14
B. HORVÁTH
(3) X = ℓ∞, since every infinite-dimensional complemented subspace of
ℓ∞ is isomorphic to ℓ∞ by Lindenstrauss' theorem ([29]) and ℓ∞ ≃
ℓ∞ ⊕ ℓ∞,
(4) X = ℓc
∞(λ), where λ is an infinite cardinal, since by [19, Theo-
∞(λ)
∞(κ) for some infinite cardinal κ ≤ λ, and
rem 1.4] every infinite-dimensional complemented subspace of ℓc
is isomorphic to ℓ∞ or ℓc
∞(κ),
ℓc
∞(κ) ≃ ℓc
∞(κ) ⊕ ℓc
(5) X = C[0, ωω], where ω is the first infinite ordinal, since by [3, Theo-
rem 3] every infinite-dimensional complemented subspace of C[0, ωω]
is isomorphic to c0 or C[0, ωω] and C[0, ωω] ≃ C[0, ωω] ⊕ C[0, ωω] by
[37, Remark 2.25 and Lemma 2.26],
(6) X = (Ln∈N ℓn
2 )Y where Y is c0 or ℓ1, since by [6, Corollary 8.4 and
Theorem 8.3] every infinite-dimensional complemented subspace of
X is isomorphic to Y or X and X ≃ X ⊕ X by [9, Corollary 7(i)].
Before we recall two important results of Laustsen -- Loy -- Read and Laustsen --
Schlumprecht -- Zsák, let us remind the reader of the following terminology.
For Banach spaces X and Y the symbol G Y (X) denotes the closed, two-
sided ideal of operators on X which factor through Y approximately, that
is, the closed linear span of the set {ST : S ∈ B(Y, X), T ∈ B(X, Y )}.
Theorem 2.18. [27, Corollary 5.6], [28, Theorem 2.12] Let X = (Ln∈N ℓn
2 )Y
where Y is c0 or ℓ1. Then the lattice of closed, two-sided ideals in B(X) is
given by
(2.11)
{0} ( K(X) ( G Y (X) ( B(X).
Proof of Theorem 1.5. Let Z be a Banach space and let ψ : B(X) → B(Z)
be a surjective algebra homomorphism. Since X ≃ X ⊕X, by Lemma 2.5 we
may suppose that Z is infinite-dimensional. Since B(X)/ Ker(ψ) ≃ B(Z),
by Theorem 2.18 it is enough to show that neither Ker(ψ) = K(X) nor
Ker(ψ) = G Y (X) can hold. The case Ker(ψ) = G Y (X) is not possible, since
G Y (X) is a maximal two-sided ideal in B(X) by Theorem 2.18 and therefore
B(X)/G Y (X) is simple as a Banach algebra whereas B(Z) is not, since Z
is infinite-dimensional. To see that Ker(ψ) = K(X) cannot hold we observe
that B(X)/K(X) does not have minimal idempotents by Example 2.17 (6)
whereas B(Z) has continuum many. Consequently Ker(ψ) = {0} must hold,
thus proving the claim.
(cid:3)
Finally in this section we shall establish some permanence properties of
Banach spaces with the SHAI property. We recall a trivial observation:
SURJECTIVE REPRESENTATIONS OF B(X)
15
Remark 2.19. If X is an infinite-dimensional Banach space and J is a
closed, two-sided ideal of B(X) such that A2 = 0 for all A ∈ J then J = {0}.
This follows from the fact that A(X) is the smallest non-trivial, closed,
two-sided ideal in B(X) and A(X) has an abundance of non-zero rank-one
idempotents.
Proof of Proposition 1.6. Let P, Q ∈ B(E) be idempotents with F = Ran(P )
and G = Ran(Q). Then P + Q = IE and P Q = 0 = QP . Now let X be a
non-zero Banach space and let ψ : B(E) → B(X) be a surjective algebra
homomorphism. Then Y := Ran(ψ(P )) and Z := Ran(ψ(Q)) are closed
(complemented) subspaces of X. Let us fix T ∈ B(F ), we observe that
ψ(P F ◦ T ◦ P F )Y ∈ B(Y ) holds. The only thing we need to check is that
the range of ψ(P F ◦ T ◦ P F )Y is contained in Y which is clearly true since
ψ(P ) ◦ ψ(P F ◦ T ◦ P F ) ◦ ψ(P ) = ψ(P F ◦ T ◦ P F ). Consequently the map
(2.12)
ϕ : B(F ) → B(Y );
T 7→ ψ(P F ◦ T ◦ P F )Y
is well-defined. It is immediate to see that ϕ is a linear map. To see that
it is multiplicative, it is enough to observe that P F ◦ P F = IF thus by
multiplicativity of ψ, for any T, S ∈ B(F ) we obtain ϕ(T )◦ϕ(S) = ϕ(T ◦S).
We show that ϕ is surjective. To see this we fix an R ∈ B(Y ). Then
ψ(P )Y ◦ R ◦ ψ(P )Y ∈ B(X) so by surjectivity of ψ it follows that there
exists A ∈ B(E) such that ψ(A) = ψ(P )Y ◦ R ◦ ψ(P )Y . Consequently
ψ(P ◦ A ◦ P ) = ψ(P ) ◦ ψ(A) ◦ ψ(P ) = ψ(P )Y ◦ R ◦ ψ(P )Y and thus by the
definition of ϕ we obtain
ϕ(P F ◦ A ◦ P F ) = ψ(P F ◦ P F ◦ A ◦ P F ◦ P F )Y = ψ(P ◦ A ◦ P )Y
(2.13)
This proves that ϕ is a surjective algebra homomorphism. Similarly we can
show that
=(cid:0)ψ(P )Y ◦ R ◦ ψ(P )Y(cid:1)(cid:12)(cid:12)(cid:12)Y
= R.
(2.14)
θ : B(G) → B(Z);
T 7→ ψ(QG ◦ T ◦ QG)(cid:12)(cid:12)Z
is a well-defined, surjective algebra homomorphism. Assume first that Y
and Z are both non-trivial subspaces of X. Since both F and G have the
SHAI property it follows that ϕ and θ are injective. Now let A ∈ Ker(ψ) be
arbitrary. Then ψ(A) = 0 implies
ϕ(P F ◦ A ◦ P F ) = ψ(P F ◦ P F ◦ A ◦ P F ◦ P F )(cid:12)(cid:12)Y = ψ(P ◦ A ◦ P )Y
= ψ(P ) ◦ ψ(A) ◦ ψ(P )Y = 0.
(2.15)
Since ϕ is injective it follows that P F ◦ A ◦ P F = 0. Using the injectivity of
θ a similar argument shows that QG ◦A◦QG = 0. We recall that E ≃ F ⊕G
16
B. HORVÁTH
and thus every A ∈ B(E) can be represented as the (2 × 2)-matrix
(cid:20)P F ◦ A ◦ P F P F ◦ A ◦ QG
QG ◦ A ◦ P F QG ◦ A ◦ QG(cid:21) .
From the previous we obtain that whenever A ∈ Ker(ψ) then A has the
off-diagonal matrix form
(2.16)
(2.17)
0
0
A =(cid:20)
QG ◦ A ◦ P F
P F ◦ A ◦ QG
0
(cid:21) .
On the one hand, since Ker(ψ) is an ideal in B(X), we obviously have that
A2 ∈ Ker(ψ) whenever A ∈ Ker(ψ), thus A2 also has the off-diagonal form
(2.18)
A2 =(cid:20)
QG ◦ A2 ◦ P F
P F ◦ A2 ◦ QG
0
(cid:21) .
On the other hand, the product of two (2 × 2) off-diagonal matrices is
diagonal and therefore by Equation (2.17)
(2.19)
P F ◦ A2 ◦ QG = 0,
QG ◦ A2 ◦ P F = 0
must also hold. Consequently A2 = 0, thus by Remark 2.19 the equality
Ker(ψ) = {0} must hold, equivalently, ψ is injective.
Let us observe that both Y = {0} and Z = {0} cannot hold. Indeed, if
both ψ(Q) and ψ(P ) were zero, then we had 0 = ψ(P + Q) = ψ(IE) = IX,
contradicting that X is non-zero. Thus without loss of generality we may
assume Y = {0} and Z 6= {0}. Hence ψ(P ) = 0, thus ψ(Q) = ψ(P )+ψ(Q) =
ψ(P + Q) = ψ(IE) = IX. This is equivalent to Z = Ran(ψ(Q)) = X, and
thus B(Z) = B(X). Therefore θ : B(G) → B(X), defined in Equation (2.14)
is a surjective algebra homomorphism. Since G has the SHAI property and
X is non-zero, it follows that θ is injective. Let A ∈ B(E) be such that
A ∈ Ker(ψ). Then
θ(QG ◦ A ◦ QG) = ψ(QG ◦ QG ◦ A ◦ QG ◦ QG)
(2.20)
= ψ(Q ◦ A ◦ Q) = ψ(Q) ◦ ψ(A) ◦ ψ(Q) = 0.
Since θ is injective, this is equivalent to QG ◦ A ◦ QG = 0 which in turn
is equivalent to Q ◦ A ◦ Q = 0. We observe that Q 6= 0, otherwise IX =
ψ(Q) = 0 which contradicts the fact that X is non-zero. Hence we can choose
x ∈ Ran(Q) and ξ ∈ E ∗ norm one vectors with hx, ξi = 1. Assume towards
a contradiction that ψ is not injective. Then in particular x ⊗ ξ ∈ F (E) ⊆
Ker(ψ), consequently Q ◦ (x ⊗ ξ) ◦ Q = 0. Thus 0 = (Q ◦ (x ⊗ ξ) ◦ Q)x =
hQx, ξiQx = hx, ξix = x, a contradiction. Hence ψ is injective, and therefore
we conclude that E has the SHAI property.
(cid:3)
SURJECTIVE REPRESENTATIONS OF B(X)
17
From Proposition 1.6 we immediately obtain the following corollary.
Corollary 2.20. If N ∈ N and {Ei}N
i=1 is set of Banach spaces with the
SHAI property then LN
i=1 Ei has the SHAI property.
3. Constructing surjective, non-injective homomorphisms
3.1. First remarks.
from B(YX) to B(X)
3.1.1. Ordinals as topological spaces and spaces of continuous functions thereof.
If α is an ordinal, then α+ denotes its ordinal successor. Equipped with the
order topology, α and α+ are locally compact and compact Hausdorff spaces,
respectively. It is well-known that the one-point (or Alexandroff) compact-
ification of α is α+. In line with the general convention, we let [0, α) := α
and [0, α] := α+. We recall that the first uncountable ordinal is denoted by
ω1.
If K is a compact Hausdorff space then C(K) denotes Banach space
of complex-valued functions on K, with respect to the supremum norm.
The Banach space C[0, ω1] is called the Semadeni space, since he showed in
[45] that C[0, ω1] is not isomorphic to its square. If L is a locally compact
Hausdorff space, and L := L ∪ {∞} is its one-point compactification, then
we introduce C0(L) := {g ∈ C( L) : g(∞) = 0}, the Banach space of
continuous functions vanishing at infinity, with respect to the supremum
norm. In this notation
(3.1)
C0[0, ω1) = {g ∈ C[0, ω1] : g(ω1) = 0}.
For a countable ordinal α let 1[0,α] denote indicator function of the interval
[0, α]. Since [0, α] is clopen, it follows that 1[0,α] ∈ C0[0, ω1). Also, by a the-
orem of Rudin [38, Theorem 6], the Banach space C[0, ω1]∗ is isometrically
isomorphic to the Banach space
(3.2)
ℓ1(ω+
1 ) =(f : [0, ω1] → C : Xα≤ω1
f (α) < ∞) .
The following definition is essential for our purposes:
A subset D ⊆ [0, ω1) is called a club subset if D is a closed and unbounded
subset of [0, ω1).
The following elementary lemma plays a crucial role in the main theorem
of this section, it can be found for example in [18, Lemma 3.4].
Lemma 3.1. A countable intersection of club subsets is a club subset.
18
B. HORVÁTH
We recall that for Banach spaces X and Y , whenever u ∈ E ⊗ F
(3.3)
n
kukǫ := sup((cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xi=1
hxi, ϕiyi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
: u =
n
Xi=1
xi ⊗ yi, ϕ ∈ X ∗, kϕk ≤ 1)
denotes the injective tensor norm on X⊗Y . The vector space X⊗Y endowed
with the norm k · kǫ is denoted by X ⊗ǫ Y . The completion of X ⊗ǫ Y with
respect to k · kǫ is called the injective tensor product of X and Y and it is
denoted by X ⊗ǫY . It is well-known (see e.g. [39, Proposition 3.2]) that for
Banach spaces X, Y , W , Z if S ∈ B(X, W ) and T ∈ B(Y, Z) then there
exists a unique S ⊗ǫ T ∈ B(X ⊗ǫY, W ⊗ǫZ) such that for every x ∈ X, y ∈ Y
the identity (S ⊗ǫ T )(x ⊗ y) = (Sx) ⊗ (T y) holds. Then kS ⊗ǫ T k = kSkkT k.
It follows from [39, Section 3.2] that for any Banach space X the Banach
space C([0, ω1]; X) of continuous functions on [0, ω1] with values in X is
isometrically isomorphic to the Banach space C[0, ω1] ⊗ǫX. The isometric
isomorphism
(3.4)
is given by
J : C[0, ω1] ⊗ǫX → C([0, ω1]; X)
(3.5)
(J(f ⊗ x))(α) = f (α)x (f ∈ C[0, ω1], x ∈ X, α ∈ [0, ω1]).
Definition 3.2. Let X be a non-zero Banach space. We define
(3.6)
YX := {F ∈ C([0, ω1]; X) : F (ω1) = 0}.
Although we shall not need this, we remark in passing that it follows
from the Hahn -- Banach Separation Theorem that C0[0, ω1) ⊗ǫX and YX are
isometrically isomorphic.
Lemma 3.3. Let X be a non-zero Banach space. Then YX is a comple-
mented subspace of C([0, ω1]; X).
Proof. For a fixed x0 ∈ X let us define the constant function
(3.7)
cx0 : [0, ω1] → X; α 7→ x0,
obviously cx0 ∈ C([0, ω1]; X). Thus we can define the map
(3.8)
Q : C([0, ω1]; X) → C([0, ω1]; X); F 7→ F − cF (ω1).
It is clear that Q is a bounded linear map with kQ(F )k ≤ 2kF k. Now
we observe that for any F ∈ C([0, ω1]; X) we clearly have Q(F )(ω1) = 0,
showing that Q(F ) ∈ YX. Also, for any F ∈ YX and any α ∈ [0, ω1] we have
(Q(F ))(α) = F (α), consequently Q is an idempotent with Ran(Q) = YX
thus proving the claim.
(cid:3)
SURJECTIVE REPRESENTATIONS OF B(X)
19
With the notations of the proof of Lemma 3.3, we define
(3.9)
P : C[0, ω1] → C[0, ω1],
g 7→ g − cg(ω1).
In particular, Ran(P ) = C0[0, ω1).
Remark 3.4. Clearly for any g ∈ C[0, ω1], x ∈ X and α ∈ [0, ω1] we have
(Q(g ⊗ x))(α) = (P g ⊗ x)(α). From this it follows that (P ⊗ǫ IX )Q(g ⊗ x) =
P g ⊗ x = Q(g ⊗ x), thus by linearity and continuity we obtain
(3.10)
IYX = (P ⊗ǫ IX )YX .
Lemma 3.5. Let X be a non-zero Banach space and suppose µ, ξ ∈ (YX)∗
satisfy hf ⊗ x, ξi = hf ⊗ x, µi for all f ∈ C0[0, ω1) and x ∈ X. Then ξ = µ.
Proof. The definition of P and the hypothesis of the lemma ensure that
for any x ∈ X and g ∈ C[0, ω1] the equality hP g ⊗ x, ξi = hP g ⊗ x, µi
holds. By Remark 3.4 we have hQ(g ⊗ x), ξi = hQ(g ⊗ x), µi, equivalently,
hg ⊗ x, (QYX )∗ξi = hg ⊗ x, (QYX )∗µi and thus by linearity and continuity of
(QYX )∗µ and (QYX )∗ξ we obtain that for all u ∈ C([0, ω1]; X) the identity
hu, (QYX )∗ξi = hu, (QYX )∗µi holds. Thus for any u ∈ C([0, ω1]; X) we have
hQu, ξi = hQu, µi consequently by Lemma 3.3 for all v ∈ YX we have that
hv, ξi = hv, µi, proving the claim.
(cid:3)
Remark 3.6. Let X be a non-zero Banach space. It is easy to see that YX
is not separable. Indeed, let x0 ∈ X be such that kx0k = 1 and let us define
the map
(3.11)
ι : C0[0, ω1) → YX;
f 7→ f ⊗ x0.
This is clearly a linear isometry, thus, since separability passes to subsets it
follows that YX cannot be separable.
In the following, if α ≤ ω1 is an ordinal, then δα ∈ C[0, ω1]∗ denotes the
Dirac measure centred at α; that is, the bounded linear functional defined
by δα(g) := g(α) for g ∈ C[0, ω1].
Remark 3.7. Let X be a non-zero Banach space and let α ∈ [0, ω1] and
ψ ∈ X ∗ be fixed. We can define a map by
(3.12)
δα ⊗ ψ : C([0, ω1]; X) → C;
u 7→ hu(α), ψi,
clearly δα ⊗ ψ ∈ C([0, ω1]; X)∗.
Let us observe that C[0, ω1] has the approximation property. By [38, The-
orem 6] we know that C[0, ω1]∗ is isometrically isomorphic to ℓ1(ω+
1 ), which
has the Radon-Nikodým property, consequently by [39, Theorem 5.33], the
Banach space (C[0, ω1] ⊗ǫX)∗ is isometrically isomorphic to C[0, ω1]∗ ⊗πX ∗,
20
B. HORVÁTH
the projective tensor product of C[0, ω1]∗ and X ∗ (see for example [39, Sec-
tion 2.1]). Equivalently, C([0, ω1]; X)∗ is isometrically isomorphic to ℓ1(ω+
1 with entries in
the Banach space of summable transfinite sequences on ω+
X ∗. This justifies the tensor notation in the definition of the functional
δα ⊗ ψ.
1 ; X ∗),
3.2. The construction. Our main theorem relies on the following result
of Kania, Koszmider, and Laustsen:
Theorem 3.8. [21, Theorem 1.5] For every T ∈ B(C0[0, ω1)) there exists a
unique
ϕ(T ) ∈ C such that there exists a club subset D ⊆ [0, ω1) such that for all
f ∈ C[0, ω1) and α ∈ D:
(3.13)
(T f )(α) = ϕ(T )f (α).
Moreover, ϕ : B(C0[0, ω1)) → C; T 7→ ϕ(T ) is a character.
In [21] the character ϕ : B(C0[0, ω1)) → C of the previous theorem
is termed the Alspach -- Benyamini character and its kernel the Loy -- Willis
ideal of B(C0[0, ω1)), and is denoted by MLW . Partial structure of the lat-
tice of closed two-sided ideals of B(C0[0, ω1)) is given in [22], in particular
E(C0[0, ω1)) = K(C0[0, ω1)) ( MLW .
Proof of Theorem 1.7. Fix S ∈ B(YX), x ∈ X and ψ ∈ X ∗. For any f ∈
C0[0, ω1) we can define the map
(3.14)
[0, ω1] → C; α 7→ h(S(f ⊗ x))(α), ψi.
x f :
x f is a continuous map, moreover by S(f ⊗ x) ∈ YX we
x f ∈ C0[0, ω1). This allows us to
x f )(ω1) = 0, consequently Sψ
Sψ
It is clear that Sψ
also have (Sψ
define the map
(3.15)
Sψ
x : C0[0, ω1) → C0[0, ω1);
x is a linear map with kSψ
f 7→ Sψ
x f.
x k ≤ kSkkxkkψk. Consequently,
It is clear that Sψ
by Theorem 3.8 there exists a club subset Dx,ψ ⊆ [0, ω1) such that for all
α ∈ Dx,ψ the equality
(3.16)
x )∗δα = ϕ(Sψ
x )δα
(Sψ
holds. We also have ϕ(Sψ
to define the map
x ) ≤ kSkkxkkψk, since kϕk = 1. This allows us
ΘS : X × X ∗ → C;
(3.17)
and we have for any x ∈ X and ψ ∈ X ∗ that ΘS(x, ψ) ≤ kSkkxkkψk.
Now we show that ΘS is bilinear. We only check that it is linear in the first
(x, ψ) 7→ ϕ(Sψ
x ),
SURJECTIVE REPRESENTATIONS OF B(X)
21
variable, linearity in the second variable follows by an analogous argument.
Let x, y ∈ X, ψ ∈ X ∗ and λ ∈ C be arbitrary. Fix f ∈ C0[0, ω1) and
α ∈ [0, ω1], then using linearity of the tensor product in the second variable,
of S and of the functional ψ it follows that
(Sψ
x+λyf )(α) = h(S(f ⊗ (x + λy)))(α), ψi
(3.18)
= h(S(f ⊗ x))(α), ψi + λh(S(f ⊗ y))(α), ψi
= (Sψ
x f )(α) + λ(Sψ
y . Since ϕ is linear, ΘS(x + λy, ψ) = ϕ(Sψ
y f )(α),
x ) + λϕ(Sψ
x+λy = Sx + λSψ
proving Sψ
x+λy) =
y ) = ΘS(x, ψ) + λ ΘS(y, ψ) readily follows, proving linearity
ϕ(Sψ
of ΘS in the first variable. Consequently ΘS is a bounded bilinear form
on X × X ∗. If κX : X → X ∗∗ denotes the canonical embedding then by
reflexivity of X the map
(3.19)
ΘS : X → X;
x 7→ κ−1
X ( ΘS(x, ·))
defines a bounded linear operator on X with kΘSk = k ΘSk and hΘS(x), ψi =
ΘS(x, ψ) = ϕ(Sψ
(3.20)
x ) for all x ∈ X, ψ ∈ X ∗. Thus we can define the map
Θ : B(YX ) → B(X); S 7→ ΘS.
Since X is separable and reflexive it follows that X ∗ is separable too. Let
Q ⊆ X and R ⊆ X ∗ be countable dense subsets. Let us fix S ∈ B(YX ),
x,ψ ⊆ [0, ω1) such
x ∈ Q and ψ ∈ R. As above, there exists a club subset DS
that for any α ∈ DS
x )f (α) and
hence
x,ψ and any f ∈ C0[0, ω1): (Sψ
x f )(α) = ϕ(Sψ
hS(f ⊗ x), δα ⊗ ψi = h(S(f ⊗ x))(α), ψi = (Sψ
= f (α)ϕ(Sψ
x f )(α)
x ) = hf (α)Θ(S)x, ψi
(3.21)
= hf ⊗ (Θ(S)x), δα ⊗ ψi.
By Lemma 3.1 it follows that
(3.22)
DS := \(x,ψ)∈Q×R
DS
x,ψ
is a club subset of [0, ω1). Consequently for any α ∈ DS, any f ∈ C0[0, ω1)
and any x ∈ Q, ψ ∈ R, Equation (3.21) holds. It is clear that for a fixed
S ∈ B(YX ), f ∈ C0[0, ω1) and α ∈ DS the maps
X × X ∗ → C;
(x, ψ) 7→ hS(f ⊗ x), δα ⊗ ψi,
(3.23)
X × X ∗ → C;
(x, ψ) 7→ hf ⊗ (Θ(S)x), δα ⊗ ψi
are continuous functions between metric spaces and thus by density of Q×R
in X × X ∗, Equation (3.21) holds everywhere on X × X ∗. In other words,
22
B. HORVÁTH
for any S ∈ B(YX ) there exists a club subset DS ⊆ [0, ω1) such that for any
α ∈ DS, f ∈ C0[0, ω1) and x ∈ X, ψ ∈ X ∗
(3.24)
hf ⊗ x, S ∗(δα ⊗ ψ)i = hf ⊗ x, δα ⊗ (Θ(S)∗ψ)i
holds. Therefore by Lemma 3.5 we obtain that for all α ∈ DS and ψ ∈ X ∗:
(3.25)
S ∗(δα ⊗ ψ) = δα ⊗ (Θ(S)∗ψ).
We show that for any S ∈ B(YX) the operator Θ(S) is determined by this
property. Indeed, suppose Θ1(S), Θ2(S) ∈ B(X) are such that there exist
club subsets DS
i and all
ψ ∈ X ∗
2 ⊆ [0, ω1) such that for i ∈ {1, 2}, all α ∈ DS
1 , DS
(3.26)
S ∗(δα ⊗ ψ) = δα ⊗ (Θi(S)∗ψ).
Let α ∈ DS
1 ∩ DS
2 , x ∈ X and ψ ∈ X ∗ be fixed. Then
hΘ1(S)x, ψi = h1[0,α] ⊗ x, δα ⊗ (Θ1(S)∗ψ)i
= h1[0,α] ⊗ x, S ∗(δα ⊗ ψ)i
= h1[0,α] ⊗ x, δα ⊗ (Θ2(S)∗ψ)i
(3.27)
= hΘ2(S)x, ψi
and thus Θ1(S) = Θ2(S). We are now prepared to prove that Θ is an algebra
homomorphism. To see this let S, T ∈ B(YX) be fixed. Let DT , DS, DT S ⊆
[0, ω1) be club subsets satisfying Equation (3.25). To see multiplicativity,
let α ∈ DT ∩ DS ∩ DT S, x ∈ X and ψ ∈ X ∗ be arbitrary. Then we obtain:
δα ⊗ (Θ(T S)∗ψ) = (T S)∗(δα ⊗ ψ) = S ∗T ∗(δα ⊗ ψ)
(3.28)
= S ∗(δα ⊗ (Θ(T )∗ψ))
= δα ⊗ (Θ(S)∗Θ(T )∗ψ)
= δα ⊗ ((Θ(T )Θ(S))∗ψ),
hence Θ(T S)∗ψ = (Θ(T )Θ(S))∗ψ, so Θ(T S)∗ = (Θ(T )Θ(S))∗, equivalently
Θ(T S) = Θ(T )Θ(S).
Linearity can be shown with analogous reasoning.
For any S ∈ B(YX ) we have kΘ(S)k = k ΘSk ≤ kSk, thus kΘk ≤ 1.
We now show that Θ has a right inverse. Let P ∈ B(C[0, ω1]) be the idem-
potent operator as in Equation (3.9). Let us fix an A ∈ B(X). We observe
that S := (P ⊗ǫ A)YX belongs to B(YX ). Indeed, for any g ∈ C[0, ω1] and
x ∈ X the identity ((P ⊗ǫ A)(g ⊗ x))(ω1) = (P g)(ω1)Ax = 0 holds plainly
because P g ∈ C0[0, ω1); thus by linearity and continuity of P ⊗ǫ A in fact
((P ⊗ǫ A)u)(ω1) = 0 for all u ∈ C[0, ω1] ⊗ǫX. This shows that S ∈ B(YX)
and therefore there exists a club subset DS ⊆ [0, ω1) such that Equation
SURJECTIVE REPRESENTATIONS OF B(X)
23
(3.25) is satisfied for all α ∈ DS and all ψ ∈ X ∗. Fix α ∈ DS. For any x ∈ X
and ψ ∈ X ∗
hAx, ψi = h1[0,α] ⊗ (Ax), δα ⊗ ψi = h(P ⊗ǫ A)(1[0,α] ⊗ x), δα ⊗ ψi
= h1[0,α] ⊗ x, S ∗(δα ⊗ ψ)i
= h1[0,α] ⊗ x, δα ⊗ (Θ(S)∗ψ)i
= hx, Θ(S)∗ψi
(3.29)
= hΘ(S)x, ψi,
and thus Θ(S) = A. In particular, we obtain Θ(IYX ) = IX, with kΘk ≤ 1
this yields kΘk = 1. Also, the above shows that the map
(3.30)
Λ : B(X) → B(YX ); A 7→ (P ⊗ǫ A)YX
satisfies Θ ◦ Λ = idB(X). It is immediate that Λ is linear with kΛk ≤ 1.
Also, Λ(IX) = IYX holds by Equation (3.10), consequently kΛk = 1. The
map Λ is an algebra homomorphism plainly because P ∈ B(C0[0, ω1)) is an
idempotent, therefore (P ⊗ǫ A)(P ⊗ǫ B) = P ⊗ǫ (AB) for every A, B ∈ B(X).
It remains to prove that Θ is not injective. For assume towards a con-
tradiction it is; then B(YX) and B(X) are isomorphic as Banach algebras.
By Eidelheit's Theorem this is equivalent to saying that YX and X are iso-
morphic as Banach spaces. This is clearly nonsense, since for example, X is
separable whereas by Remark 3.6 the Banach space YX is not.
(cid:3)
Remark 3.9. With the notations established in the proof of Theorem 1.7
we clearly have in fact
(3.31)
Ker(Θ) = {S ∈ B(YX ) : (∀x ∈ X)(∀ψ ∈ X ∗)(Sψ
x ∈ MLW )},
where Sψ
x is defined by (3.14).
If X is an infinite-dimensional Banach space then Ker(Θ) is of course not
maximal in B(YX ), however, it is not the smallest possible ideal in B(YX ).
To see this, we need some preliminary observations.
In the following, let P ∈ B(C[0, ω1]) be as in Equation (3.9). If X is a
non-zero Banach space, we fix x0 ∈ X and ξ ∈ X ∗ such that kx0k = kξk =
hx0, ξi = 1 and consider the linear isometry
(3.32)
ι : C0[0, ω1) → YX;
f 7→ f ⊗ x0.
We also consider the norm one linear map
(3.33)
ρ : C[0, ω1] ⊗ǫX → C[0, ω1]
24
B. HORVÁTH
which is unique with the property that for any g ∈ C[0, ω1] and x ∈ X the
identity
ρ(g ⊗ x) = hx, ξig holds. With this we obtain the following:
Lemma 3.10. Let X be a non-zero Banach space. Then
(3.34)
Ξ : B(C0[0, ω1)) → B(YX); S 7→(cid:0)(P C0[0,ω1) ◦ S ◦ P C0[0,ω1)) ⊗ǫ IX(cid:1)YX
(3.35)
Υ : B(YX) → B(C0[0, ω1));
T 7→ P C0[0,ω1) ◦ ρYX ◦ T ◦ ι
define norm one linear maps with Υ ◦ Ξ = idB(C0[0,ω1)). Moreover, Ξ is an
algebra homomorphism such that (Ξ(S))ψ
x = hx, ψiS for every x ∈ X and
ψ ∈ X ∗.
Proof. It is clear that (cid:0)(P C0[0,ω1) ◦ S ◦ P C0[0,ω1)) ⊗ǫ IX(cid:1)YX ∈ B(YX) holds
for any S ∈ B(C0[0, ω1)), thus Ξ is well-defined. It is easy to see that Ξ is
linear with kΞk ≤ 1. From Equation (3.10) it follows that Ξ(IC0[0,ω1)) = IYX ,
thus kΞk = 1. The map Ξ is multiplicative simply by the defining property
of injective tensor products of operators. Let S ∈ B(C0[0, ω1)), x ∈ X and
ψ ∈ X ∗ be fixed. Then for any f ∈ C0[0, ω1) and α ≤ ω1 ordinal
(3.36)
(cid:0)(Ξ(S))ψ
thus (Ξ(S))ψ
x f(cid:1)(α) = h(Ξ(S)(f ⊗ x))(α), ψi = h(Sf )(α)x, ψi = (Sf )(α)hx, ψi,
x = hx, ψiS indeed.
Linearity of Υ is immediate, so is kΥk ≤ 1. Since Υ(IYX ) = IC0[0,ω1)
follows from the definition of Υ, we obtain kΥk = 1 as required.
It remains to show that Υ ◦ Ξ = idB(C0[0,ω1)). For any S ∈ B(C0[0, ω1))
and f ∈ C0[0, ω1)
Υ(Ξ(S))f = (P C0[0,ω1) ◦ ρYX ◦ Ξ(S) ◦ ι)f
= (P C0[0,ω1) ◦ ρYX ◦ Ξ(S))(f ⊗ x0)
= (P C0[0,ω1) ◦ ρYX )(Sf ⊗ x0)
= P C0[0,ω1)(hx0, ξiSf )
(3.37)
= Sf,
consequently Υ(Ξ(S)) = S, which proves the claim.
(cid:3)
Corollary 3.11. The containment E(YX) ( Ker(Θ) holds.
Proof. By Lemma 2.6 it follows that E(YX) ⊆ Ker(Θ), we show that the
containment is proper. For assume towards a contradiction that Ker(Θ) =
E(YX). If S ∈ MLW then by Lemma 3.10 for all x ∈ X and ψ ∈ X ∗ in
fact (Ξ(S))ψ
x = hx, ψiS ∈ MLW , thus by Remark 3.9 then Ξ(S) ∈ Ker(Θ)
SURJECTIVE REPRESENTATIONS OF B(X)
25
follows. Thus Ξ(S) ∈ E(YX) by the indirect assumption and since E is an
operator ideal in the sense of Pietsch, it follows from Lemma 3.10 that
(3.38)
S = Υ(Ξ(S)) = P C0[0,ω1) ◦ ρYX ◦ Ξ(S) ◦ ι ∈ E(C0[0, ω1)).
This yields MLW = E(C0[0, ω1)), which is a contradiction.
(cid:3)
Acknowledgements. The majority of the research presented herein was
carried out during the author's Ph.D. studies, he is grateful to his supervi-
sors Dr Yemon Choi and Dr Niels J. Laustsen (Lancaster) for their invalu-
able advice during the preparation of this paper. He is indebted to Dr Saeed
Ghasemi (Prague), Dr György Pál Gehér (Reading), Professor Lajos Molnár
(Szeged), Dr Tamás Titkos, and Dr Zsigmond Tarcsay (Budapest) for many
enlightening conversations. We would like to thank the anonymous referee
for his/her many insightful comments, which helped to improve the presen-
tation of the paper a great deal, and for drawing our attention to [41]. The
author acknowledges the financial support from the Lancaster University
Faculty of Science and Technology and acknowledges with thanks the par-
tial funding received from GAČR project 19-07129Y; RVO 67985840 (Czech
Republic).
References
[1] S. A. Argyros, J. F. Castillo, A. S. Granero, M. Jimenéz, and J. P. Moreno, Com-
plementation and embeddings of c0(I) in Banach spaces, Proc. London Math. Soc.
85 (2002), 742 -- 768.
[2] B. Aupetit, E. Makai, M. Mbekhta, and J. Zemánek, Local and global liftings of
analytic families of idempotents in Banach algebras, Acta Sci. Math. (Szeged) 80
(2014), 149 -- 174.
[3] Y. Benyamini, An extension theorem for separable Banach spaces, Israel. J. Math.
29 (1978), 24 -- 30.
[4] M. T. Boedihardjo and W. B. Johnson, On mean ergodic convergence in the Calkin
algebras, Proc. Amer. Math. Soc. 143 (2015), 2451 -- 2457.
[5] F. F. Bonsall and J. Duncan, Complete Normed Algebras, Springer-Verlag, New
York, 1973.
[6] J. Bourgain, P. G. Casazza, J. Lindenstrauss, and L. Tzafriri, Banach spaces with a
unique unconditional basis, up to permutation, Mem. Amer. Math. Soc. 54 (1985).
[7] J. W. Calkin, Two-sided ideals and congruences in the ring of bounded operators in
Hilbert space, Annals of Mathematics 42 (1941), 839 -- 873.
[8] S. R. Caradus, W. E. Pfaffenberger, and B. Yood, Calkin Algebras and Algebras of
Operators on Banach Spaces, Marcel Dekker, Inc., New York, 1974.
26
B. HORVÁTH
[9] P. G. Casazza, C. A. Kottman, and B.-L. Lin, On some classes of primary Banach
spaces, Canad. J. Math. 29 (1977), 856 -- 873.
[10] J. B. Conway, A Course in Functional Analysis, Springer-Verlag, New York, 1990.
[11] H. G. Dales, Banach Algebras and Automatic Continuity, Oxford University Press
Inc., New York, 2000.
[12] H. G. Dales, T. Kania, P. Koszmider, T. Kochanek, and N. J. Laustsen, Maximal left
ideals of the Banach algebra of bounded linear operators on a Banach space, Studia
Math. 218 (2013), 245 -- 286.
[13] I. S. Èdel'šteın and B. S. Mitjagin, The homotopy type of linear groups of two classes
of Banach spaces, Funkcional. Anal. i Priložen 4 (1970), 61 -- 72 (in Russian).
[14] D. Freeman, Th. Schlumprecht, and A. Zsák, Closed ideals of operators between
classical sequence spaces, Bull. Lond. Math. Soc. 49 (2017), 859 -- 876.
[15] I. C. Gohberg, A. S. Markus, and I. A. Feldman, Normally solvable operators and
ideals associated with them, American Math. Soc. Translat. 61 (1967), 63 -- 84.
[16] W. T. Gowers, A solution to Banach's hyperplane problem, Bull. Lond. Math. Soc.
26 (1994), 523 -- 530.
[17] W. T. Gowers and B. Maurey, The unconditional basic sequence problem, J. Amer.
Math. Soc. (1993), 851 -- 874.
[18] K. Hrbacek and T. Jech, An introduction to Set Theory, 3rd ed., Marcel Dekker Inc.,
New York, 1999.
[19] W. B. Johnson, T. Kania, and G. Schechtman, Closed ideals of operators on and
complemented subspaces of Banach spaces of functions with countable support, Proc.
Amer. Math. Soc. 144 (2016), 4471 -- 4485.
[20] W. B. Johnson, G. Pisier, and G. Schechtman, Ideals in L(L1), arXiv:1811.06571
(2018)
[21] T. Kania, P. Koszmider, and N. J. Laustsen, A weak∗-topological dichotomy with
applications in operator theory, Trans. London Math. Soc. 1 (2014), 1 -- 28.
[22] T. Kania and N. J. Laustsen, Operators on two Banach spaces of continuous func-
tions on locally compact spaces of ordinals, Proc. Amer. Math. Soc. 143 (2015),
2585 -- 2596.
[23] T. Kania and N. J. Laustsen, Ideal structure of the algebra of bounded operators
acting on a Banach space, Indiana Univ. Math. J. 66 (2017), 1019 -- 1043.
[24] N. J. Laustsen, Maximal ideals in the algebra of operators on certain Banach spaces,
Proc. Edinb. Math. Soc. 45 (2002), 523 -- 546.
[25] N. J. Laustsen, On ring-theoretic (in)finiteness of Banach algebras of operators on
Banach spaces, Glasgow Math. J. 45 (2003), 11 -- 19.
SURJECTIVE REPRESENTATIONS OF B(X)
27
[26] N. J. Laustsen and R. J. Loy, Closed ideals in the Banach algebra of operators on a
Banach space, in: Topological algebras, their applications, and related topics, 245 --
264, Banach Center Publ. 67, Polish Acad. Sci. Inst. Math., Warsaw, 2005.
[27] N. J. Laustsen, R. J. Loy, and C. J. Read, The lattice of closed ideals in the Banach
algebra of operators on certain Banach spaces, J. Funct. Anal. 214 (2004), 106 -- 131.
[28] N. J. Laustsen, Th. Schlumprecht, and A. Zsák, The lattice of closed ideals in the
Banach algebra of operators on a certain dual Banach space, J. Operator Theory 56
(2006), 391 -- 402.
[29] J. Lindenstrauss, On complemented subspaces of m, Israel J. Math. 5 (1957) 153 -- 156.
[30] R. J. Loy and G. A. Willis, Continuity of derivations on B(E) for certain Banach
spaces E, J. London Math. Soc. 40 (1989), 327 -- 346.
[31] P. Motakis, D. Puglisi, and D. Zisimopoulou, A hierarchy of Banach spaces with
C(K) Calkin algebras, Indiana Univ. Math. J. 65 (2016), 39 -- 67.
[32] C. P. Ogden, Homomorphisms from B(C(ωη)), J. London Math. Soc. 54 (1996),
346 -- 358.
[33] A. Pełczyński, Projections in certain Banach spaces, Studia Math. 19 (1960) 206 --
228.
[34] A. Pietsch, Inessential operators in Banach spaces, Integral Equations and Operator
Theory 1 (1978), 589 -- 591.
[35] G. Plebanek, A construction of a Banach space C(K) with few operators, Topology
and its Applications 143 (2004), 217 -- 239.
[36] C. E. Rickart, General Theory of Banach Algebras, Robert E. Krieger Publishing
Co., Inc., New York, 1960.
[37] H. P. Rosenthal, The Banach Spaces C(K) (Handbook of the Geometry of Banach
Spaces, Vol.2), North Holland, Amsterdam, 2003.
[38] W. Rudin, Continuous functions on compact spaces without perfect subsets, Proc.
Amer. Math. Soc. 8 (1957) 39 -- 42.
[39] R. A. Ryan, Introduction to Tensor Products of Banach Spaces, Springer-Verlag,
London, 2002.
[40] M. Rørdam, F. Larsen, and N. J. Laustsen, An Introduction to K-theory for C ∗-
algebras, Cambridge University Press, Cambridge, 2000.
[41] I. Schlackow, Centripetal operators an Koszmider spaces, Topology Appl. 155 (2008)
1227 -- 1236.
[42] Th. Schlumprecht, An arbitrarily distortable Banach space, Israel J. Math. 76 (1991),
81 -- 95.
[43] Th. Schlumprecht, A complementably minimal Banach space not containing c0 or ℓp,
in: Seminar Notes in Functional Analysis and Partial Differential Equations, Baton
Rouge, LA, 1992.
28
B. HORVÁTH
[44] Th. Schlumprecht and A. Zsák, The algebra of bounded linear operators on ℓp ⊕ ℓq
has infinitely many closed ideals, J. Reine Angew. Math. 735 (2018), 225 -- 247.
[45] Z. Semadeni, Banach spaces non-isomorphic to their Cartesian squares, II. Bull.
Acad. Polon. Sci. Sér. Sci. Math. Astr. Phys. 8 (1960) 81 -- 84.
[46] M. Tarbard, Operators on Banach Spaces of Bourgain-Delbaen Type, Ph.D. Thesis,
University of Oxford, 2013.
[47] R. J. Whitley, Strictly singular operators and their conjugates, Trans. Amer. Math.
Soc. 113 (1964) 252 -- 261.
Institute of Mathematics, Czech Academy of Sciences, Žitná 25, 115 67
Prague 1, Czech Republic
E-mail address: [email protected], [email protected]
|
1210.1944 | 1 | 1210 | 2012-10-06T11:56:06 | Hyperbolic wavelet transform: an efficient tool for multifractal analysis of anisotropic textures | [
"math.FA"
] | Global and local regularities of functions are analyzed in anisotropic function spaces, under a common framework, that of hyperbolic wavelet bases. Local and directional regularity features are characterized by means of global quantities constructed upon the coefficients of hyperbolic wavelet decompositions. A multifractal analysis is introduced, that jointly accounts for scale invariance and anisotropy. Its properties are studied in depth. | math.FA | math |
Hyperbolic wavelet transform: an efficient tool for multifractal
analysis of anisotropic textures∗
P. Abry†, M. Clausel‡, S. Jaffard§, S.G. Roux†and B.Vedel¶
Abstract
Global and local regularities of functions are analyzed in anisotropic function spaces,
under a common framework, that of hyperbolic wavelet bases. Local and directional
regularity features are characterized by means of global quantities constructed upon the
coefficients of hyperbolic wavelet decompositions. A multifractal analysis is introduced,
that jointly accounts for scale invariance and anisotropy.
Its properties are studied in
depth.
Keywords : Hyperbolic wavelet analysis, Anisotropic Besov Spaces, Pointwise Holder Reg-
ularity, Anisotropic Multifractal Analysis.
2010 Mathematics Subject Classification : 42C40, 46E35.
1
Introduction
Natural images often display various forms of anisotropy. For a wide range of applications,
anisotropy has been quantified through regularity characteristics and features that strongly
differ when measured in different directions. This is, for instance, the case in medical imaging
(osteoporosis, muscular tissues, mammographies,...), cf. e.g. [15, 14], hydrology [44], fracture
surfaces analysis [22],. . . . For such images, a key issue consists first in describing, within a
suitable framework, the anisotropy of the texture, and then in defining regularity anisotropy
parameters that can actually and efficiently be measured via numerical procedures and fur-
ther involved into e.g., classification schemes. This requires the design of a mathematical
framework that allows to define and estimate these parameters. Such a program can be split
into several questions, some of them having already been either solved or, at least, patially
addressed.
A first issue lies in introducing global and local notions of anisotropic regularity, which
emcompass and extend (isotropic) regularity spaces, such as Sobolev or Besov spaces, and
the classical notion of pointwise Holder regularity. To model anisotropy, the particular set-
ting of an anisotropic self–similar field driven by two parameters (an anisotropy matrix and a
self–similarity index) has been introduced and studied in [14], where it is used as a relevant
∗This work has been supported by the ANR grant AMATIS (ANR2011 BS01 011 02) and the CNRS,
Groupe de Recherche Analyse Multifractale.
†Physics Dept., ENS Lyon, CNRS, UMR5672, Lyon, France.
‡University of Grenoble, CNRS, Laboratoire Jean Kuntzmann UMR 5224, Saint Martin d'H`eres, France.
§Universit´e Paris Est, LAMA, UMR 8050, Cr´eteil, France.
¶LMBA, Universit´e de Bretagne Sud, European University of Bretagne, Vannes, France.
1
model to describe osteoporosis. In [21], the question of defining in a proper way the concept
of anisotropy of an image in relation to its global regularity has been addressed. It has no-
tably been shown that these two parameters can be recovered without a–priori knowledge of
the characteristics of the model, by studying the global smoothness properties of the process.
Furthermore, some of the properties characterizing anisotropy are revealed by the regular-
ity of the sample paths when analyzed with functional spaces well-adapted to anisotropy:
Anisotropic Besov spaces.
This preliminary study thus showed the central role that such spaces should play in the
mathematical modeling of random anisotropic textures. The introduction of these spaces
traces back to the study of some PDEs, cf. e.g. [51], for the study of semi-elliptic pseudo-
differential operators whose symbols have different degrees of smoothness along different di-
rections, the reader is also referred to [2], and references therein, for a recent use of such spaces
for optimal regularity results for the heat equation. Other types of directional function spaces
have also been considered, cf. e.g. [16] for the variant supplied by Hardy directional spaces.
A second crucial issue consists in obtaining a simple characterization of these spaces on
a "dictionary".
Indeed, the challenge is to measure the critical exponent of any image in
anisotropic Besov spaces for different anisotropies using the same analysis tool. Wavelet
analysis is well–known to be an efficient tool for measuring smoothness in a large range of
functional spaces (cf. [39] for details). Here, however, the main point is that the anisotropy of
the analyzing spaces must not be set a priori to a fixed value but instead be allowed to vary.
Specific bases are thus looked for, which would serve as a common dictionary for anisotropic
Besov spaces with different anisotropies.
It is natural that one should use some form of
anisotropic wavelets such as curvelets, bandelets, contourlets, shearlets, ridgelets, or wedgelets
(see e.g. [34] for an thorough review of these representation systems and a comparison of their
properties for image processing); natural criteria of choice being, on the mathematical side,
that these variations on isotropic wavelets supply bases for the corresponding anisotropic
spaces, and, on the applied side, that practically tractable procedure can be devised and
implemented to permit the characterization of real-world data according to these function
spaces.
Many authors addressed this problem, and proposed different solutions, depending on the
precise definition of anisotropic space they started with, as well as on the anisotropic basis
they used, cf. e.g. [23, 33, 32, 38, 52] and also the recent papers [28] by G. Garrig´os and
A. Tabacco, and [31] by D. Haroske and E. Tam´asi, which contain numerous references on
the subject. Note also that, in several cases, a particular type of anisotropy was considered:
Parabolic anisotropy (where a contraction by λ in one direction is associated with a contrac-
tion by λ2 along the orthogonal direction, [37, 46, 42], and references therein, in particular
for applications to directional regularity), the corresponding dictionaries being in that case
curvelets or contourlets (corresponding to the Hart–Smith decomposition in the continuous
setting, see [49]). This particular choice of anisotropy was motivated from application to
PDEs (see e.g., [29, 20] where curvelets and ridgelets are used for the study of Fourier inte-
gral operators, with applications to the wave equation) but is no longer justified when dealing
with images, where no particular form of anisotropy can be postulated a priori. To the op-
posite, figuring out the precise form of anisotropy present in data is part of the issue. This
argument also implies that one should not restrict analysis to tools that match one specific
type of anisotropy, but rather that to tools embracing all of them simultaneously, in order to
be able to detect that that suits data.
From now on, two possible solutions for this problem will be focused on:
2
• One is supplied by anisotropic Triebel bases, see [52], that are constructed from the stan-
dard wavelet case through a multiresolution procedure, tailored to a specific anisotropy.
The collection of these bases does not constitute a frame. However, for a fixed anisotropy,
simple characterizations of anisotropic Besov spaces have been supplied within this sys-
tem. Such characterizations can thus be used as a building step to construct a multi-
fractal formalism [?]. This is further detailed in Section 3.
Triebel bases provide a powerful tool to deduce results on anisotropic Besov spaces, for
a fixed anisotropy. In particular, it enables to show that these spaces are isomorphic to
the corresponding isotropic Besov spaces. Further, some results such as embeddings or
profiles of Besov characteristics can be obtained, via the transference method proposed
by H. Triebel. However, when it comes to understand the link between different forms
of anisotropy - in term of function spaces by example - this tool remains of limited
interest. Indeed, the knowledge of the expansion of a function in one basis gives a priori
neither information about its expansion in an other basis nor about its belonging to all
anisotropic Besov spaces.
• Another possible decomposition system is supplied by hyperbolic wavelets, introduced
in various settings under different denominations (standard, rectangular or hyperbolic
wavelet analysis) notably in image coding (see [54]), numerical analysis (see [13], [12])
and in [23],[33] for the purpose of approximation theory. They are simply defined as
tensor products of 1D wavelets, yet allowing different dilations factors along different
directions, as opposed to the classical discrete wavelet transform that relies on a single
isotropic fixation factor. This key difference enables the study of anisotropy. Hyperbolic
wavelet basis form a non–redundant system by construction, and contain all possible
anisotropies. Hyperbolic wavelet bases have thus been used in statistics for the purpose
of adaptive estimation of multidimensional curves. Notably, it has been proven in two
seminal articles [41] and [40] that nonlinear thresholding of noisy hyperbolic wavelet
coefficients leads to (near)–optimal minimax rates of convergence over a wide range of
anisotropic smoothness classes. The reader is also referred to the recent work of F.
Autin, G. Claeskens, J.M. Freyermuth [3] where this problem is considered from the
maxiset point of view. Other interesting applications of hyperbolic analysis can also be
founded in [5],[4],[8],[6] and [7] where hyperbolic wavelet decompositions of Fractional
Brownian Sheets and Linear Fractional Stable Sheets are given and are used to prove
many sample paths properties of these random fields (smoothness properties, Hausdorff
dimension of the graph).
The key feature of hyperbolic wavelet bases is that they provide a common dictionary
for anisotropic Besov spaces. This result is stated in Theorem 2.2 of Section 2: The
critical exponent in anisotropic Besov spaces will be related to some ℓp norms of the
hyperbolic wavelet coefficients. These mathematical results yield an efficient method
for the detection of anisotropy, as detailed in a companion article, where numerical
investigations are conducted, [45] .
In the present article, it has been chosen to explore the possibilities supplied by the hyper-
bolic wavelet transform to investigate directional regularity, both in global (anisotropic Besov
spaces) and local (directional pointwise regularity) forms. The underlying motivation is to
develop a multifractal formalism relating these two notions (just as the standard multifractal
formalism relates the usual Besov spaces with the notion of (anisotropic) Holder pointwise
3
smoothness, see [35] and references therein). It also aims at obtaining a numerically stable
procedure that thus permits to extract the anisotropic features existing in natural images as
well as information related to the size (fractional dimensions) of the corresponding geometrical
sets.
Before proceeding further, let us motivate the choice of hyperbolic wavelets against Triebel
bases. For a fixed anisotropy, one can argue that Triebel bases display slightly better mathe-
matical advantages: An exact characterization of anisotropic Besov spaces, as shown in [52],
and a characterization of pointwise smoothness as sharp as in the isotropic case, as shown
by H. Ben Braiek and M. Ben Slimane in [10]. However, a first purpose of the present con-
tribution is to show that these two important properties hold almost as well for hyperbolic
wavelets: In Section 2, "almost characterizations" (i.e., necessary and sufficient conditions
that differ by a logarithmic correction) of anisotropic Besov spaces are obtained. Further-
more, if one is not only interested in analysis, but also in simulation, this slight disadvantage
(a logarithmic loss, which in applications can not be detected) is overcompensated by the
advantage of using a basis instead of an overcomplete system. Indeed, generating a random
field with prescribed regularity properties requires the use of a basis (using an overcomplete
system cannot guarantee a priori that the simulated field with coefficients of specific sizes has
the expected properties, since nontrivial linear combinations of the building blocks may van-
ish). A contrario, with the hyperbolic wavelet basis, one can easily provide toy examples with
different multifractal spectra depending on the anisotropy. Our being jointly motivated by
analysis and synthesis motivates the choice of a system that permits an interesting trade-off
among directional wavelets, in terms of mathematical efficiency and numerical simplicity and
robustness, both on the analysis and synthesis sides. The practical relevance of the mathe-
matical tools introduced and studied here are assessed in a companion paper [45].
Let us now further compare Triebel and hyperbolic wavelet bases in terms of pointwise
directional smoothness. First, note that this notion has been the subject of few investigations
so far: To our knowledge, the natural definition which allows for a wavelet characterization
was first introduced by M. Ben Slimane in the 90s, see [11], in order to investigate the mul-
tifractal properties of anisotropic selfsimilar functions. Partial results when using parabolic
basis (i.e., curvelets and Hart–Smith transform) have been obtained by J. Sampo and S.
Sumetkijakan see [37, 46, 42] and references therein. A generalization and implications in
terms of sizes of coefficients on directional wavelets (the so-called "anisets", which are a mix-
ture of of the wavelet and Gabor transform, where the wavelets can be arbitrarily shrunk
in certain directions) were also worked out in [36]. Finally, an "almost " characterization of
pointwise directional regularity was recently obtained by H. Ben Braiek and M. Ben Slimane
in [?] on the Triebel basis coefficients, where the basis is picked so that its anisotropy pa-
rameter is fitted to the type of directional regularity considered. In Section 2, we will obtain
a similar result, but relying on the coefficients of the hyperbolic wavelet basis, thus paving
the way to the construction of a multifractal formalism. An important difference with [?] is
that, here, a single basis fits all anisotropies. Therefore, as in the case of Besov spaces, the
advantage is that no a priori needs to be assumed on the particular considered anisotropy.
This thus can be used as a way to detect the specific anisotropy which exists in data at hand,
rather than assuming a priori its particular form beforehand. Note that other decomposition
systems have also been used for the detection of local singularities, see for instance [24, 30]
where shearlets and wedgelets are used for the detection of discontinuities along smooth edges.
4
Let us now come back to the anisotropic self–similar fields considered in [45, 21]. Such
exactly selfsimilar models are somewhat toy examples, and, though testing regularity indices
on their realizations is an important validation step, their study could prove misleadingly
simple (just as, in 1D, fractional Brownian motion is too simple a model to fit the richness of
situations met in real-world data). Natural images are indeed likely to consist of patchworks
of different kinds of deformed pieces and therefore, can be expected to exhibit more complex
scale invariance properties, and only in an approximate way. A natural setting to describe
such properties, where different kinds of singularities are mixed up, is supplied by multifractal
analysis. The next step is therefore to combine both anisotropy and multifractal analyses. To
this end, a new form of multifractal analysis is introduced, based on the hyperbolic wavelet
coefficients, and relating the global and local characterizations of regularity. It allows to take
into account both scale invariance properties and local anisotropic features of an image. Thus,
it provides a new tool for image classification, seen as a refinement of texture classification
based on the usual isotropic multifractal analysis, as proposed for instance in [1],[35]. Section 3
is devoted to the introduction of this new framework: A new multifractal formalism, referred
to as the hyperbolic multifractal formalism. It allows to relate local anisotropic regularity of
the analyzed image to global quantities called hyperbolic structure functions as commonly
done in multifractal analysis. Note that alternative multifractal analysis and multifractal
formalism were introduced by H. Ben Braiek and M. Ben Slimane in [?], based on Triebel
basis coefficients. In their approach, a particular anisotropy is picked, and the corresponding
basis is used. As above, the main difference between our point of view and theirs is that
we do not pick beforehand a particular anisotropy: Therefore, the approach proposed here
does not rely on any a priori assumptions on data, and can thus be used when anisotropies
of several types are simultaneously present in data.
Finally, detailed proofs of all the results stated in Sections 2 and 3 are provided in Section 4.
2 Anisotropic global regularity and hyperbolic wavelets
We first focus on the measure of anisotropic global regularity using a common analyzing dic-
tionary: hyperbolic wavelet bases. Here, we start by providing the reader with a brief account
of the corresponding functional spaces. Thereafter, we recall some well-known facts about
hyperbolic wavelet analysis (cf. Section 2.2). The main result of the present section con-
sists of Theorem 2.2, proven in Section 4.1, which allows to determine the critical directional
Besov indices of data by regressions on log-log plot of quantities based on hyperbolic wavelet
coefficients (see Section 2.2 for a precise statement).
2.1 Anisotropic Besov spaces
Anisotropic Besov spaces generalize classical (isotropic) Besov spaces, and many results con-
cerning isotropic spaces have been extended in this setting, see [18, 17] for a complete account
on the results used in this section, and [16, 53] for detailed overviews on anisotropic spaces.
Note in particular that these spaces are invariant by smooth diffeomorphisms on each coor-
dinate, an important requirement for image processing.
Anisotropic Besov spaces verify (asymptotically in the limit of small scales) norm invari-
ances with respect to anisotropic scaling, we, therefore, start by recalling this notion. Let
α = (α1, α2) denote a fixed couple of parameters, with α1, α2 ≥ 0 and α1 + α2 = 2. In the
remainder, such couples will be referred to as admisible anisotropies. Such couples quantify
5
the degree of anisotropy of the space (α1 = α2 = 1 corresponding to the isotropic case). For
any t ≥ 0 and ξ = (ξ1, ξ2) ∈ R2, we define anisotropic scaling by tαξ = (tα1ξ1, tα2ξ2). Note
that, in this definition and in the following, the coordinate axes are chosen as anisotropy di-
rections. This particular choice can of course be modified by the introduction of an additional
rotation (as envisaged e.g., in[45]).
ysis, which we now recall. Let ϕα
that
Anisotropic Besov spaces may be introduced using an anisotropic Littlewood Paley anal-
0 ≥ 0 belong to the Schwartz class S(R2) and be such
0 (x) = 1 if
ϕα
i=1,2ξi ≤ 1 ,
sup
and
For j ∈ N, we define
Then,
ϕα
0 (x) = 0 if
i=1,22−αi ξ ≥ 1 .
sup
ϕα
j (x) = ϕα
0 (2−(j−1)αξ) .
0 (2−jαξ) − ϕα
+∞Xj=0
ϕα
j ≡ 1 ,
and (ϕα
j )j≥0 is called an anisotropic resolution of the unity. It satisfies
supp (ϕα
0 ) ⊂ Rα
1 ,
supp (ϕα
k ) ⊂ Rα
j+1 \ Rα
j ,
where
For f ∈ S′(R2) let
i=1,2ξℓ ≤ 2αik} .
j = {ξ = (ξ1, ξ2) ∈ R2; sup
Rα
j f = F−1(cid:16)ϕα
j bf(cid:17) .
∆α
The sequence (∆α
Besov spaces are then defined as follows (see [18, 17]).
j f )j≥0) is called an anisotropic Littlewood–Paley analysis of f . The anisotropic
Definition 2.1 The Besov space Bs,α
defined by
p,q, log β (R2), for 0 < p ≤ +∞, 0 < q ≤ +∞, s, β ∈ R, is
Bs,α
p,q, log β (R2) = {f ∈ S′(R2);Xj≥0
=Xj≥0
kfkBs,α
p,q, log β
j−βq2jsqk∆α
j fkq
1/q
< +∞} .
p
j−βq2jsqk∆α
j fkq
1/q
,
p
6
This definition does not depend on the resolution of the chosen unity ϕα
0 and the quantity
is a norm (resp., quasi-norm) on Bs,α
p,q (R2) for 1 ≤ p, q ≤ +∞ (resp., 0 < p, q < 1).
As in the isotropic case, anisotropic Besov spaces encompass a large class of classical
anisotropic functional spaces (see [53] for details). For example, when p = q = 2 and (α1, α2) ∈
Q2 is an admissible anisotropy, let us consider s > 0 such that s/α1 and s/α2 are both integers,
then the anisotropic Sobolev space
H s,α(R2) = {f ∈ L2(R2) such that
∂s/α1f
∂x1 ∈ L2(R2) and
∂s/α2f
∂x2 ∈ L2(R2)} ,
coincides with the Besov space Bs,α
2,2 (R2).
spaces and are denoted Cs,α
In the special case where p = q = ∞, the spaces Bs,α
∞,∞(R2) are called anisotropic Holder
log u(R2). These spaces also admit a finite difference characterization
Let (e1, e2) denote the canonical basis of R2. For a function f : R2 → R, ℓ ∈ {1, 2} and
that we now recall (see also [53] for details).
t ∈ R one defines
∆1
t,ℓf (x) = f (x + teℓ) − f (x) .
The difference of order M , M ≥ 2, of function f , along direction eℓ, is then iteratively defined
as
∆M
t,ℓf (x) = ∆t,ℓ∆M−1
t,ℓ
f (x) .
One then has:
Proposition 2.1 Let α = (α1, α2) ∈ (R+
f : R2 → R. The function f belongs to the anisotropic Holder space Cs,α
∗ )2 such that α1 + α2 = 2, s > 0, u ∈ R and
log u(R2) if
kfkL∞(R2) +
2Xℓ=1
sup
t>0
k∆Mℓ
t,ℓ f (x)kL∞(R2)
ts/αℓ log(t)u
< +∞ ,
where for any ℓ ∈ {1, 2}, Mℓ = [s/αℓ] + 1.
2.2 Hyperbolic wavelet characterization of anisotropic Besov spaces
We state our first main result which consists of an hyperbolic wavelet caracterization of
anisotropic Besov spaces.
We first need to recall the definition of the hyperbolic wavelet bases as tensorial products
of two unidimensional wavelet bases (see [23]) and second state Theorem 2.2, further proven
in Section 4.1.
Definition 2.2 Let ψ denote the unidimensional Meyer wavelet and ϕ the associated scaling
function. The hyperbolic wavelet basis is defined as the system {ψj1,j2,k1,k2, (j1, j2) ∈ (Z+ ∪
{−1})2, (k1, k2) ∈ Z2} where
• if j1, j2 ≥ 0,
ψj1,j2,k1,k2(x1, x2) = ψ(2j1x1 − k1)ψ(2j2 x2 − k2) .
• if j1 = −1 and j2 ≥ 0
ψ−1,j2,k1,k2(x1, x2) = ϕ(x1 − k1)ψ(2j2 x2 − k2) .
7
• if j1 ≥ 0 and j2 = −1
ψj1,−1,k1,k2(x1, x2) = ψ(2j1 x1 − k1)ϕ(x2 − k2) .
• if j1 = j2 = −1
ψ−1,−1,k1,k2(x1, x2) = ϕ(x1 − k1)ϕ(x2 − k2) .
For any f ∈ S′(R2), one then defines its hyperbolic wavelet coefficients as follows:
cj1,j2,k1,k2 = 2j1+j2 < f, ψj1,j2,k1,k2 > if j1, j2 ≥ 0 ,
cj1,−1,k1,k2 = 2j1 < f, ψj1,j2,k1,k2 > if j1 ≥ 0 and j2 = −1 ,
c−1,j2,k1,k2 = 2j2 < f, ψj1,j2,k1,k2 > if j1 = −1 and j2 ≥ 0 ,
c−1,−1,k1,k2 = < f, ψj1,j2,k1,k2 > if j1 = j2 = −1 .
Remark 2.1 We chose a L1-normalization for the wavelet coefficients, known to be well-
matching scale invariance.
The main result of this section is an hyperbolic wavelet characterization of the spaces
Bs,α
p,q, log β (R2), up to a logarithmic correction. In the sequel, some notations are needed. For
any j = (j1, j2) ∈ (N ∪ {−1})2, let us define:
kcj1,j2,·,·kℓp = X(k1,k2)∈Z2 cj1,j2,k1,k2p
1/p
.
Let α = (α1, α2) be an admissible anisotropy, one also defines the following subsets of N2
Γ(HL)
j
Γ(LH)
j
Γ(HH)
j
and
(1)
(α) = {(j1, j2) ∈ N2, [(j − 1)α1] − 1 ≤ j1 ≤ [jα1] + 1 and 0 ≤ j2 ≤ [(j − 1)α2] − 1},
(α) = {(j1, j2) ∈ N2, 0 ≤ j1 ≤ [(j − 1)α1] − 1 and [(j − 1)α2] − 1 ≤ j2 ≤ [jα2] + 1},
(α) = {(j1, j2) ∈ N2, [(j − 1)α1] − 1 ≤ j1 ≤ [jα1] + 1 and [(j − 1)α2] − 1 ≤ j2 ≤ [jα2] + 1} ,
Γj(α) = Γ(HL)
j
(α) ∪ Γ(LH)
j
(α) ∪ Γ(HH)
j
(α) .
Let us now state our hyperbolic wavelet characterization of anisotropic Besov spaces:
Theorem 2.2 Let α = (α1, α2) be an admissible anisotropy, (s, β) ∈ R2 and (p, q) ∈ (0, +∞]2.
Let f ∈ S′(R2).
1. Set β(p, q) = max(1/p − 1, 0) + max(1 − 1/q, 0). If
Xj∈N0
jqβ(p,q)−βq2jsq X(j1,j2)∈Γj (α)
2− (j1+j2)q
p
kcj1,j2,·,·kq
1/q
< +∞ ,
ℓp
then f ∈ Bs,α
p,q, log β (R2) (with usual modifications when q = ∞).
8
2. Conversely,
(a) If q < ∞ and f ∈ Bs,α
p,q, log β (R2) then
j−βq−12jsq X(j1,j2)∈Γj (α)
Xj∈N0
2− (j1+j2)q
p
kcj1,j2,·,·kq
ℓp
1/q
< +∞ .
(b) If f ∈ Bs,α
p,∞, log β (R2) then
max
j∈N0
j−β2js max
(j1,j2)∈Γj (α)
2− (j1+j2)
p
kcj1,j2,·,·kℓp < +∞ .
Proof. Theorem 2.2 is proven in Section 4.1.
(cid:3)
In particular, for p = q = ∞, the following "almost" characterization of anisotropic Holder
spaces by means of hyperbolic wavelets holds:
Proposition 2.3 Let α = (α1, α2) an admissible anisotropy, (s, β) ∈ R2 and f ∈ S′(R2).
(i) If f ∈ Cs,α(R2) then there exists some C > 0 such that for all j ∈ N ∪ {−1} and any
(j1, j2) ∈ Γj(α),
(2)
kcj1,j2,·,·kℓ∞ ≤ C2−js .
(ii) Conversely, assume that there exists some C > 0 such that for all j ∈ N∪{−1} and any
(j1, j2) ∈ Γj(α)
(3)
then f ∈ Cs,α(R2).
kcj1,j2,·,·kℓ∞ ≤
2−js
j
,
In the special case where p = q = 2, that is if we consider anisotropic Sobolev spaces, there
is no logarithmic correction:
Theorem 2.4 Let α = (α1, α2) an admissible anisotropy, s ∈ R. Let f ∈ S′(R2). The two
following assertions are equivalent:
(i) f ∈ H s,α(R2) = Bs,α
(ii)
2,2 (R2).
Xj∈N0
22js X(j1,j2)∈Γj (α)
2−(j1+j2) kcj1,j2,·,·k2
1/2
< +∞ .
ℓ2
Proof. Theorem 2.4 is proven in Section 4.1.
(cid:3)
9
3 Hyperbolic multifractal analysis
We are now interested in the simultaneous analysis of local regularity properties and of
anisotropic features of a function. To that end, we construct a new multifractal analysis,
referred to as the hyperbolic multifractal analysis. Recall that in the isotropic case, the pur-
pose of multifractal analysis is to provide information on the the pointwise singularities of
functions. Multifractal functions are usually such that their local regularity strongly vary
from point to point, so that it is not possible to estimate the regularity index (defined below)
of a function at a given point. Instead, the relevant information consists of the "sizes" of the
sets of points where the regularity index takes the same value. This "size" is mathematically
formalized as the Hausdorff dimension. The function that associates the dimension of the set
of points sharing the same regularity index with this index is referred to as the spectrum of
singularities (or multifractal spectrum). The goal of a multifractal formalism is to provide
a method that allows to measure the spectrum of singularities from quantities that can ac-
tually be computed on real-world data. We extend this approach to the anisotropic setting.
Let us first recall that, in the case where the anisotropy of the analyzing space is fixed, this
has already been achieved: See [?] for anisotropic pointwise regularity analysis using Triebel
bases and [?] for the corresponding anisotropic multifractal formalism. Here, we generalize
these two previous works, providing a multifractal analysis which does not rely on the a priori
knowledge of the regularity and takes into account all possible anisotropies. Note that for a
fixed anisotropy, both formalims coincide: Indeed they are derived from wavelet characteriza-
tions of the same functional spaces. Nevertheless, the formalism based on hyperbolic wavelet
allows to deal simultaneously with all possible anisotropies, thus providing more useful algo-
rithms for analyzing real-world data. In addition, the use of hyperbolic wavelet bases offers
the possibility to define and synthesize deterministic and stochastic mathematical objects
with prescribed anisotropic behavior.
In Section 3.1, the different concepts related to pointwise regularity are first recalled. An
hyperbolic wavelet criterion is then devised in Section 3.1.2. Our main result, Theorem 3.1,
is stated in Section 3.1.2 and proven in Section 4. Hyperbolic wavelet analysis is defined
in Section 3.2.2 and the validity of the proposed multifractal formalism is investigated in
Theorem 3.2.
3.1 Anisotropic pointwise regularity and hyperbolic wavelet analysis
3.1.1 Definitions
Let us start by recalling the usual notion of pointwise regularity (cf. [35] for a complete
review).
Definition 3.1 Let f be in L∞loc(R2) and s > 0. The function f belongs to the space Cs
if there exist some C > 0, δ > 0 and Px0 a polynomial with degree less than s such that
log β (x0)
if x − x0 ≤ δ, f (x) − Px0(x) ≤ Cx − x0s · log(x − x0)β ,
where · is the usual Euclidean norm of R2. If β = 0, the space Cs
Cs(x0).
log 0(x0) is simply denoted
10
Anisotropic pointwise regularity is further defined as follows. Let P denote a polynomial
of the form:
P (t1, t2) = X(β1,β2)∈N2
aβ1,β2tβ1
1 tβ2
2 ,
and let α = (α1, α2) be an admissible anisotropy. The α–homogeneous degree of the polyno-
mial P is defined as:
dα(P ) = sup{α1β1 + α2β2, aβ1,β2 6= 0}; .
Finally, for any t = (t1, t2) ∈ R2, the α–homogeneous norm reads:
tα = t11/α1 + t21/α2 .
We can now define the spaces Cs,α
Definition 3.2 Let f ∈ L∞loc(R2), α = (α1, α2) such that α1 + α2 = 2, ·α, s > 0 and β ∈ R.
The function f belongs to Cs,α
log β (x0) if there exists some C > 0, δ > 0 and Px0 a polynomial
with α–homogeneous degree less than s such that
log β (x0).
if x − x0α ≤ δ, f (x) − Px0(x) ≤ Cx − x0s
log 0(x0) is simply denoted Cs,α(x0).
If β = 0, the space Cs,α
The anisotropic pointwise exponent of a locally bounded function f at x0 can be thus be
defined as:
α · log(x − x0α)β .
(4)
hf,α(x0) = sup{s, f ∈ Cs,α(x0)} .
The reader is referred to [10],[36] for more details about the material of this section.
3.1.2 An hyperbolic wavelet criterion
As in the usual anisotropic setting (see [35]), the anisotropic pointwise Holder regularity of
a function is closely related to the decay rate of decay of its wavelet leaders. The usual
definition of wavelet leaders needs to be tuned to to the hyperbolic setting:
For any (j1, j2, k1, k2), let λ(j1, j2, k1, k2) denote the hyperbolic dyadic cube:
(5)
λ = λ(j1, j2, k1, k2) = [
k1
2j1
,
k1 + 1
2j1
k2
2j2
,
[×[
k2 + 1
2j2
[ ,
and let cλ stand for cj1,j2,k1,k2. The hyperbolic wavelet leaders dλ, associated with the hyper-
bolic cube λ, can now be defined as:
dλ = sup
λ′⊂λcλ′ .
For any x0 = (a, b) ∈ R2, let
3λj1,j2(x0) = [
[2j1 a] − 1
2j1
,
[2j1a] + 2
2j1
[2j2b] − 1
2j2
,
[×[
[2j2b] + 2
2j2
[ ,
(where [·] denotes the integer part) and
dj1,j2(x0) =
sup
λ′⊂3λj1 ,j2 (x0)cλ′ .
The hyperbolic wavelet leaders criterion for pointwise regularity can now be stated as:
11
Theorem 3.1 Let s > 0 and α = (α1, α2) ∈ (R∗+)2 such that α1 + α2 = 2.
1. Assume that f ∈ Cs,α(x0). There exists some C > 0 such that for any j1, j2 ∈ N∪{−1}
one has
(6)
dj1,j2(x0) ≤ C2− max( j1
α1
, j2
α2
)s .
2. Conversely, assume that f is uniformly Holder and that (6) holds, then f ∈ Cs,α
log 2(x0).
Proofs are postponed to Section 4.
3.2 Anisotropic multifractal analysis
3.2.1 Two notions of dimension
In multifractal analysis, two different notions of dimension are mainly used: the Hausdorff
dimension and the packing dimension, whose definitions are now recalled.
The Hausdorff dimension is defined through the Hausdorff measure (see [25] for details).
The best covering of a set E ⊂ Rd with sets subordinated to a diameter ε can be estimated
as follows,
Hδ
ε(E) = inf{
Eiδ, E ⊂
Ei,Ei ≤ ε}.
∞Xi=1
∞[i=1
ǫ as ε goes to 0.
ε is an outer measure. The Hausdorff measure is defined as the (possibly infinite
Clearly, Hδ
or vanishing) limit Hδ
Hδ′
Definition 3.3 The Hausdorff dimension dimH(E) of a set E ⊂ Rd is defined as follows,
The Hausdorff measure is decreasing as δ goes to infinity. Moreover, Hδ(E) > 0 implies
(E) = ∞ if δ′ < δ. The following definition is thus meaningful.
dimH(E) = sup{δ : Hδ(E) = ∞} .
The box dimension (or Minkowski dimension) is simpler to define and to use than the
With this definition, dimH(∅) = −∞.
Hausdorff dimension.
Definition 3.4 Let E ⊂ Rd. If ε > 0, let Nε(E) be the smallest number of sets of radius ε
required to cover E. The upper box dimension is
The lower box dimension is
dimB(E) = lim sup
ε→0
log Nε(E)
− log ε
.
dimB(E) = lim inf
ε→0
log Nε(E)
− log ε
.
If these two quantities are equal, they define the box dimension dimB(E) of E.
A significant limitation of box dimension is that a set and its closure have the same dimension.
The packing dimension (introduced by Tricot, see [50]) has better mathematical properties
(e.g., the packing dimension of a countable union of sets is the supremum of their dimensions).
12
Definition 3.5 The packing dimension dimP (E) of a set E ⊂ Rd is defined by
dimP (E) = inf{sup
i {dimB(Ei)} : E ⊂
∞[i=1
Ei}.
The following inequality holds for any set E ⊂ Rd,
dimH(E) ≤ dimP (E).
We now define the hyperbolic spectrum of singularities of a locally bounded function using
the Hausdorff dimension.
Definition 3.6 Let f be a locally bounded function and α = (α1, α2) ∈ (R∗+)2 such that
α1 + α2 = 2. The iso–anisotropic–Holder set are defined as
Ef (H, α) = {x ∈ R2, hf,α(x) = H}
where the anisotropic pointwise Holder hf,α(x) has been defined in (4).
The hyperbolic spectrum of singularities of f is then defined as:
d : (R+ ∪ {∞}) × (0, 2) → R+ ∪ {−∞}
(H, a) 7→ dimH(Ef (H, (a, 2 − a))) .
3.2.2 The hyperbolic wavelet leader multifractal formalism
It is not always possible to compute theoretically the spectrum of singularities of a given func-
tion. A multifractal formalism thus consists of a practical procedure that yields (a convex
hull of) the function d, through the construction of structure functions and the use of the Leg-
endre transform. In the classical case, these formalisms are variants of a seminal derivation,
proposed by Parisi and Frisch in the context of the study of hydrodynamic turbulence [43].
The hyperbolic wavelet leader multifractal formalism described below aimed at extending the
procedure to where both anisotropy and singularities are studied jointly.
Let us define hyperbolic partition functions of a locally bounded function as:
(7)
S(j, p, α) = 2−2j X(j1,j2)∈Γj (α) X(k1,k2)∈Z2
dp
j1,j2,k1,k2
,
where Γj(α) has already been defined in Section 2.2 with Eq. (1).
From the definition of an anisotropic scaling function (or scaling exponents)
(8)
ωf (p, α) = lim inf
j→∞
log S(j, p, α)
log 2−j
,
let us further define the Legendre hyperbolic spectrum:
(9)
Lf (H, α) = inf
p∈R∗{Hp − ωf (p, (α, 2 − α)) + 2} .
Qualitatively, the Legendre hyperbolic spectrum and the hyperbolic spectrum of singular-
ities df (H, a) are expected to coincide, while the theorem below actually provides an upper
bound relationship.
13
Theorem 3.2 Let f a uniform Holder function. The following inequality holds
(10)
∀(H, a) ∈ (R∗+) × (0, 2),
df (H, a) ≤ Lf (H, a).
Definition 3.7 Let f a uniform Holder function, H > 0 and a ∈ (0, 2). If Eq. (10) simplifies
into an equality, i.e.,
∀(H, a) ∈ (R∗+) × (0, 2),
d(H, a) = Lf (H, α),
then f satisfies the hyperbolic multifractal formalism.
From an applied perspective, Eqs. (7) , (8) and (9) constitute the core of the practical
procedure enabling to compute the Legendre hyperbolic spectrum from the hyperbolic wavelet
leaders computed on the data to be analyzed. Practical implementation show preliminary
satisfactory results, notably, for isotropic function, it is clearly observed that the measured
Lf (H, α) does not depend on α.
4 Proofs
4.1 Proof of Theorem 2.2
4.1.1 Hyperbolic Littlewood-Paley characterization of Bs,α
Let θ0 ∈ S(R, R+) be supported on [−2, 2] such that θ0 = 1 on [−1, 1]. For any j ∈ N, let us
define
p,q (R2)
θj = θ0(2j·) − θ0(2j−1·) .
such thatPj≥0 θj(·) = 1 is a 1–D resolution of the unity.
Observe that, for any j ≥ 1, supp (θj) ⊂ {2j−1 ≤ ξ ≤ 2j+1}.
Remark 4.1 In the following, the function θ0 can be chosen with an arbitrary compact sup-
port. It does not change the main results even if technical details of proofs and lemmas have
to be adapted. It allows to chose the Fourier transform of a Meyer scaling function for θ0.
Definition 4.1
1. For any j, ℓ ≥ 0, and any ξ = (ξ1, ξ2) ∈ R2 set
φj1,j2(ξ1, ξ2) = θj1(ξ1)θj2(ξ2) .
For any j1, j2 ≥ 0, the function φj1,j2 belongs to S(R2) and is compactly supported
on {2ℓ1 ≤ ξ1 ≤ 2ℓ1+1} × {2ℓ2 ≤ ξ2 ≤ 2ℓ2+1]. Further Pj1≥0Pj2≥0 φj1,j2 = 1 and
(φj1,j2)(j1,j2)∈N2 is called an hyperbolic resolution of the unity.
2. For f ∈ S′(R2) and j1, j2 ≥ 0 set
The sequence ((∆j1,j2f )j1,j2≥0) is called an hyperbolic Littlewood–Paley analysis of f .
∆j1,j2f = F−1(cid:16)φj1,j2bf(cid:17) .
14
In the remainder of the section, we are given α = (α1, α2) a fixed anisotropy satisfying
j )j≥0 an anisotropic resolution of the unity. One then defines the following
α1 + α2 = 2 and (ϕα
functions for any j ≥ 0,
(11)
gα
j = Xj1,j2∈Γj (α)
φj1,j2 ,
where the sets Γj(α) have been defined in (1).
Remark 4.2 Hyperbolic Littlewood–Paley analysis is used in the definition of spaces of mixed
smoothness. We refer to [47] for a study of these spaces and to [48] for their link with tensor
products of Besov spaces and their hyperbolic wavelet characterizations.
We now provide the reader with the following hyperbolic Littlewood–Payley characterization
of anisotropic Besov spaces:
Theorem 4.1 Let s ∈ R and (p, q) ∈ (0, +∞]2.
1.
2.
(a) If q < ∞ and
(12)
Xj≥0
(13)
(b) If
then f ∈ Bs,α
jq max(1/p−1,0)+max(q−1,0) · j−βq2jsq X(j1,j2)∈Γj (α)
p,q, log β (R2).
j≥0 (cid:18)jmax(1/p−1,0)+1 · j−β2js max
p,∞, log β (R2).
(a) If q < ∞ and f ∈ Bs,α
then f ∈ Bs,α
max
k∆j1,j2(f )kq
p
1/q
< +∞ ,
(j1,j2)∈Γj (α) k∆j1,j2(f )kp(cid:19) < +∞ ,
(b) If f ∈ Bs,α
p,∞, log β (R2) then
p,q, log β (R2) then
Xj≥0
j−1 · j−βq2jsq X(j1,j2)∈Γj (α)
j≥0 j−β2js X(j1,j2)∈Γj (α)
max
k∆j1,j2(f )kq
p
1/q
< +∞ .
k∆j1,j2(f )kp < +∞ .
The proof of Theorem 4.1 consists of several steps, beginning with
Lemma 4.1
1. For any j ≥ 0 and any (j1, j2) 6∈ Γj(α), one has
(14)
supp(ϕα
j ) ∩ supp(φj1,j2) = ∅ .
15
2. For any j ≥ 0 and any ℓ 6∈ {j − 1, j, j + 1}, one has
j ) ∩ supp(ϕα
supp(gα
(15)
ℓ ) = ∅ .
Proof. Point 1 of the lemma is first proved, that is if (ℓ1 ≥ L(1)
(ℓ1 ≤ L(1)
min − 1 and ℓ2 ≤ L(2)
min − 1), then supp(ϕα
j ), then ξ ∈ Rα
j+1 \ Rα
j ) ∩ supp(φj1,j2) = ∅.
j . Hence, if ℓ1 ≥ L(1)
Indeed, if ξ ∈ supp(ϕα
max + 1,
one has for i = 1, 2, 2−ℓiξi ≤ 2αi(j+1)−ℓi ≤ 2αi−1, by assumptions on ℓ1, ℓ2. Since α1 or
α2 necessarily belongs to (0, 1), one has ξ 6∈ supp(φℓ1,ℓ2). Hence, φℓ1,ℓ2(ξ) = 0. The same
approach leads to φℓ1,ℓ2(ξ) = 0 if ℓ1 ≤ L(1)
min − 1 and Point (1) of Lemma 4.2
min − 1 and ℓ2 ≤ L(2)
max + 1 and ℓ2 ≥ L(2)
max + 1 and ℓ2 ≥ L(2)
max + 1) or
is obtained.
Point (2) of Lemma 4.2 can be obtained similarly.
(cid:3)
From Lemma 4.2 an intermediate hyperbolic Littelwood Paley characterization of anisotropic
Besov spaces is obtained.
Proposition 4.2 Let (p, q) ∈ (0, +∞]2, s, β ∈ R. the two following assertions are equivalent:
Let us first show that f ∈ Bs,α
p,q, log β (R2) implies Inequality (16) of Proposition 4.2. To
this end, Point 1 of Lemma 4.2 is used to deduce that for any j
1/q
q
p
< +∞ .
X(j1,j2)∈Γj (α)
j−βq2jsq(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
φj1,j2bf = ϕα
[∆j1,j2(f )](cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
j X(j1,j2)∈Γj (α)
φj1,j2bf = ϕα
j (cid:16)gα
j bf(cid:17) ,
1. f ∈ Bs,α
2.
p,q, log β (R2).
(16)
Xj≥0
Proof of Proposition 4.2.
j X(j1,j2)∈N2
ϕα
j bf = ϕα
where gα
j is defined by Equation (11). Observe now that replacing the usual dilation with an
anisotropic one gives an anisotropic version of Equation (13) in Section 1.5.2 in [51]. More
precisely assume that we are given p ∈ (0, +∞], b > 0 and M ∈ S(R2). There exists some
C > 0 not depending on b nor M such that for any h ∈ Lp(R2) such that supp(bh) ⊂ {ξ ∈
R2, supi ξi ≤ bαi}, one has
(17)
kF−1 (MFh)kLp(R2) ≤ CkM (bα·)kH s
2 (R2)khkLp(R2)
where H s
2 is the usual Bessel potential space and s > 2(1/ min(p, 1) − 1/2).
Set now b = 2j, M = ϕα
depending on j such that for any p ∈ (0, +∞] and any f ∈ Lp(R2)
∆j1,j2fkLp = Ck(F−1gα
j andbh = gα
j fkLp ≤ Ck X(j1,j2)∈Γj (α)
k∆α
j bf . Since ϕα
j (2jα·) = ϕα
1 , there exists some C > 0 not
j ) ∗ fkLp .
16
Then
(18) kfkBs,α
p,q, log β
=Xj≥0
which gives Point 1 of Proposition 4.2.
j−βq2jsqk∆α
j fkq
Lp
1/q
≤ CXj≥0
j−βq2jsqk(F−1gα
j ) ∗ fkq
Lp
1/q
,
Let us now prove that if Equation (16) of Proposition 4.2 holds then f ∈ Bs,α
Point 2 of Lemma 4.2 gives for any j ≥ 0
gα
j (cid:0)ϕα
j bf = gα
j−1 + ϕα
Hence, Inequality (17) applied with b = 2j, M = gα
existence of some C > 0 not depending on j nor f such that for any p ∈ (0, +∞]
2k(F−1ϕα
j (2jα·)kH s
j+1(cid:1)bf .
j andbh = (ϕα
j−1 + F−1ϕα
j + F−1ϕα
j−1 + ϕα
j + ϕα
j + ϕα
j+1) ∗ fkLp
j+1)bf gives the
p,q, log β (R2).
Since k · kLp is either a norm or a quasi–norm (according to the value of p), there exists some
C > 0 such that
k(cid:0)F−1gα
j(cid:1) ∗ fkLp ≤ ckgα
j(cid:1) ∗ fkLp ≤ C kgα
j (2jα·)kH s
k(cid:0)F−1gα
Let us first bound kgα
j−1) ∗ fkLp + k(F−1ϕα
2 . To this end, observe that
j ) ∗ fkLp + k(F−1ϕα
j+1) ∗ fkLp(cid:1) .
F[gα
j (2jα·)](ξ) = 2−j(α1+α2)bgj(2−jαξ) = 2−2jbgj(2−jαξ)
Since θ1 ∈ S(R), for any M > 1 there exists some C > 0 such that
2
j (2jα·)kH s
(cid:0)k(F−1ϕα
= X(j1,j2)∈Γj (α)
(1 + ξ2)s X(j1,j2)∈Γj (α)
(1 + ξ2)s X(j1,j2)∈Γj (α)
bθ1(ζ) ≤
× X(j1,j2)∈Γj (α)
× X(j1,j2)∈Γj (α)
(1 + ξ2)s
(1 + ξ2)s
Hence
kgα
j (2jα·)k2
H s
2
= ZR2
≤ ZR2
Finally
kgα
j (2jα·)k2
H s
2 ≤ CMZR2
≤ CMZR2
2j1+j2−2jbθ1(2j1−jα1 ξ1)bθ1(2j2−α2j ξ2) .
2j1+j2−2jbθ1(2j1−jα1 ξ1)bθ1(2j2−α2j ξ2)
2j1+j2−2jbθ1(2j1−jα1 ξ1)bθ1(2j2−α2j ξ2)
2
dξ
2
dξ .
CM
(1 + ζ)2M .
2
dξ
2j1+j2−2j
(1 + 2j1−jα1 ξ1)2M · (1 + 2j2−jα2 ξ2)2M
(2jα1−j1 + ξ1)2M · (2jα2−j2 + ξ2)2M
1
2
dξ .
17
By the inequality
(a + b)2 ≥ a max(b, 1) ,
valid for any a > 1, b > 0 and applied successively with a = 2jα1−j1 and b = ξ1, a = 2jα2−j2
and b = ξ2, it comes
kgα
(1 + ξ2)s
j (2jα·)k2
H s
2 ≤ CMZR2
× X(j1,j2)∈Γj (α)
2(j1−jα1)M 2(j2−jα2)M ×
max(1,ξ1)M max(1,ξ2)M
1
2
dξ .
With a M sufficiently large it follows that
sup
2(cid:1) < +∞ .
j (cid:0)kgα
Going back to an upper bound of k(cid:16)F−1gα
k(cid:0)F−1gα
j(cid:1) ∗ fkLp ≤ Cj(cid:0)k(F−1ϕα
(19) Xj≥0
Lp
j ) ∗ fkq
j−βq2jsqk(F−1gα
j (2jα·)kH s
j(cid:17) ∗ fkLp, there exists some C > 0 such that
j+1) ∗ fkLp(cid:1)
Lp
j fkq
=Xj≥0
j ) ∗ fkLp + k(F−1ϕα
j−1) ∗ fkLp + k(F−1ϕα
j−βq2jsqk∆α
≤ CkfkBs,α
p,q, log β
and
1/q
1/q
the last shows that if (16) holds then f ∈ Bs,α
p,q, log β (R2).
,
Proof of Theorem 4.1. Let us first recall that:
• For any p ∈ (0, +∞], n ∈ N, and (f1,··· , fn) ∈ Lp(R2)n
(20)
kf1 + ··· + fnkLp ≤ nmax(1/p−1,0) (kf1k + ··· + kfnk)
• For any q ∈ (0, +∞), n ∈ N, and (a1,··· , an) ∈ (R+)n
(a1 + ··· + an)q ≤ nmax(q−1,0) (aq
(21)
1 + ··· + aq
n) .
Let us now prove the first point of the theorem in the case where q 6= ∞. For this, assume
p,q, log β (R2). By Inequalities (20), (21) and the
that (12) holds and let us prove that f ∈ Bs,α
fact that Card(Γj(α)) ≤ Cj there exists C > 0 such that
Hence,
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
q
Lp
≤ Cjq max(1/p−1,0)+max(q−1,0) X(j1,j2)∈Γj (α)
∆j1,j2f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
j−βq2jsq(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
jq max(1/p−1,0)+max(q−1,0) · j−βq2jsq X(j1,j2)∈Γj (α)
∆j1,j2f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X(j1,j2)∈Γj (α)
1/q
Lp
q
X(j1,j2)∈Γj (α)
Xj≥0
≤ CXj≥0
18
k∆j1,j2fkq
Lp .
k∆j1,j2fkq
1/q
.
Lp
It proves that if (12) holds, one has
∆j1,j2fkq
< ∞ .
1/q
Lp
p,q, log β (R2).
jmax(1/p−1,0)j−β2js X(j1,j2)∈Γj (α)
k∆j1,j2fkLp .
Finally, by Point (1) of Proposition 4.2, it comes that f ∈ Bs,α
We now deal with the case q = ∞. In this case, we have
max
j≥0
j−β2js(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X(j1,j2)∈Γj (α)
Hence if (16) holds, f ∈ Bs,α
Xj≥0
j−βq2jsqk X(j1,j2)∈Γj (α)
∆j1,j2f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp
φj1,j2bf = φj1,j2(cid:0)gα
p,∞, log β (R2).
≤ C max
j≥0
To prove the converse assertion, let us assume f ∈ Bs,α
p,q, log β (R2). Observe that for any
j ≥ 0 and any (j1, j2) ∈ Γj(α), one has
j−1 + gα
j + gα
j+1(cid:1)bf .
Remark that φj1,j2(2jα·) is bounded in H s
(17), there exists C > 0 not depending on j nor f such that for any (j1, j2) ∈ Γj(α)
k(F−1φj1,j2) ∗ fkLp ≤ C(cid:0)k(F−1gα
2(R2) independently of (j1, j2) ∈ Γj(α). Hence, by
j+1) ∗ fkLp(cid:1) .
j ) ∗ fkLp + k(F−1gα
Again, two cases have to be distinguished according whether q 6= ∞ or q = ∞.
j−1) ∗ fkLp + k(F−1gα
Let us consider the case q < ∞. Observing that Card(Γj(α)) ≤ Cj, we deduce that
X(j1,j2)∈Γj (α)
k(F−1φj1,j2) ∗ fkq
Lp ≤ Cj
k(F−1gα
l ) ∗ fkq
Lp .
So
(22) Xj
j−1j−βq2jsq X(j1,j2)∈Γj (α)
k(F−1φj1,j2) ∗ fkq
j · j−1j−βq2jsqk(F−1gα
j ) ∗ fkq
Lp .
Since in addition the function f is assumed to belong to Bs,α
p,q, log β (R2), one has
Xj
j−βq2jsqk(F−1gα
j ) ∗ fkq
Lp =Xj
j · j−1j−βq2jsqk(F−1gα
j ) ∗ fkq
Lp < ∞ ,
which directly yields the required inequality using (22).
In the case q = ∞, we have
max
(j1,j2)∈Γj (α)k(F−1φj1,j2) ∗ fkq
Lp ≤ C
which leads for some C > 0 to
max
ℓ=j−1,j,j+1k(F−1gα
ℓ ) ∗ fkLp .
j≥0 (cid:18)j−β2js max
(j1,j2)∈Γj (α)k(F−1φj1,j2) ∗ fkLp(cid:19) ≤ C max
max
j≥0 (cid:16)j−β2jsk(F−1gα
j ) ∗ fkLp(cid:17) ,
19
j+1Xl=j−1
Lp ≤Xj
that is
(23)
j1,j2≥0(cid:18)max(
max
j1
α1
,
j2
α2
)(cid:19)−β
2max( j1
α1
, j2
α2
Since in addition f is assumed to belong to Bs,α
)sk(F−1φj1,j2) ∗ fkLp ≤ C max
j≥0
p,∞, log β (R2), it comes
j ) ∗ fkLp < ∞ .
max j−β2jsk(F−1gα
j−β2jsk(F−1gα
j ) ∗ fkLp .
Finally, the required conclusion is obtained by an approach similar to the one used for the
previous case.
4.1.2 The special case p = 2
In that case, an exact hyperbolic Littlewood–Paley characterization of anisotropic Besov
spaces is provided:
2,q, log β (R2).
Proposition 4.3 Let f ∈ S′(R2), s > 0 and q ∈ (0, +∞]. The two following assertions are
equivalent
(i) f ∈ Bs,α
(ii) Pj≥0
j−βq2jsq P(j1,j2)∈Γj (α)k∆j1,j2(f )k2
L2! q
j−βq2jsq P(j1,j2)∈Γj (α)kφj1,j2bfk2
L2! q
= Pj≥0
<
2
2
+∞.
In particular the following exact hyperbolic Littlewood–Paley characterization of anisotropic
Sobolev spaces can be stated:
Theorem 4.4 Let f ∈ S′(R2), s > 0 and q ∈ (0, +∞]. The two following assertions are
equivalent
(i) f ∈ H s,α
(ii) Pj≥0
log β (R2) = Bs,α
j−2β22js P(j1,j2)∈Γj (α) kφj1,j2bfk2
j−2β22js P(j1,j2)∈Γj (α) k∆j1,j2(f )k2
To prove Proposition 4.4, let us first precise the relation between anisotropic and hyperbolic
resolutions of the unity.
L2 = Pj≥0
L2 < +∞.
2,2 (R2).
Lemma 4.2 For any j ≥ 0, the following inequality holds on R2
(24)
ϕα
φj1,j2 ≤ ϕα
j−1 + ϕα
j + ϕα
j+1 .
j ≤ gj = X(j1,j2)∈Γj (α)
Proof. Let us first observe that
ϕα
j ≤ X(j1,j2)∈N2
φj1,j2 .
20
To get the left hand side of inequality (24), we just have to prove that if ξ = (ξ1, ξ2) ∈
supp(ϕα
max + 1)
or (ℓ1 ≤ L(1)
j ), one has φj1,j2(ξ) = 0 if (j1, j2) 6∈ Γj, that is if (ℓ1 ≥ L(1)
max + 1 and ℓ2 ≥ L(2)
min − 1 and ℓ2 ≤ L(2)
min − 1).
j+1 \ Rα
j ), then ξ ∈ Rα
Indeed, if ξ ∈ supp(ϕα
max + 1,
one has for i = 1, 2, 2−ℓiξi ≤ 2αi(j+1)−ℓi ≤ 2αi−1, by assumptions on ℓ1, ℓ2. Since α1 or
α2 necessarily belongs to (0, 1), one has ξ 6∈ supp(φℓ1,ℓ2). Hence, φℓ1,ℓ2(ξ) = 0. The same
approach leads to φℓ1,ℓ2(ξ) = 0 if ℓ1 ≤ L(1)
min − 1. The left hand side of
Let us now prove that the right hand side of inequality (24) holds. It comes from the
min − 1 and ℓ2 ≤ L(2)
inequality (24) is obtained.
max + 1 and ℓ2 ≥ L(2)
j . Hence, if ℓ1 ≥ L(1)
obvious equality
X(j1,j2)∈Γj
φj1,j2 ≤Xj≥0
ϕα
j ≡ 1 .
and if ξ ∈ supp(φj1,j2) for some (j1, j2) ∈ Γj(α) then for any ℓ ∈ {j − 1, j, j + 1}, ϕα
ℓ (ξ) = 0.
(cid:3)
Before proving Proposition 4.4, let us give some background about quasi–orthogonal systems.
Definition 4.2 Let a Hilbert space H and h·,·i the associated scalar product. A system
{fk, k ∈ Z} of H is said to be quasi-orthogonal if there exists some ℓ ∈ N such that
(25)
Lemma 4.3 Let H be a Hilbert space and k · k,h·,·i the associated norm and scalar product.
Let {fm, m ∈ Z} a quasi–orthogonal system of H and let ℓ ∈ N satisfying (25). Then
Proof. Observe that for any m′ ∈ Z, hfm, fm′i = 0 except if m′ − ℓ ≤ m ≤ m′ + ℓ. Hence
(26)
2
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xm∈Z
fm(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2
kfmk2 .
≤ (2ℓ + 1)Xm∈Z
∀(k, k′) ∈ Z2,(cid:0)k′ − k ≥ ℓ ⇒ hfk, fk′i = 0(cid:1) .
fm(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xm∈Z
≤ Xm′∈ZXm∈Z
m′+ℓXm=m′−ℓ
≤ Xm′∈Z
√2ℓ + 1 m′+ℓXm=m′−ℓ
≤ Xm′∈Z
√2ℓ + 1 Xm′∈Z m′+ℓXm=m′−ℓ
≤ (2ℓ + 1) Xm′∈Z
kfmk2! 1
kf′mk
kfmk2!! 1
kfmkkfm′k
hfm, fm′i
kfm′k2 .
≤
2
21
2
. Xm′∈Z
2
kfm′k2! 1
Proof of Proposition 4.4
By Plancherel Theorem and by Lemma 4.2, one has
j (ξ)2bf (ξ)2dξ ≤ C0Z gj(ξ)2bf (ξ)2dξ ,
k∆α
L2 =ZR2 ϕα
j fk2
j−1)2 + (ϕα
and
g2
ZR2
j )2 + (ϕα
j (ξ)bf (ξ)2dξ ≤ZR2(cid:2)(ϕα
j+1fk2
where gj is defined by (24). Proposition 4.4 is then a straightforward consequence of the
following lemma :
Lemma 4.4 There exists some C > 0 such that for any j ≥ 0, one has
(27)
j )2(cid:3)bf (ξ)2 ≤ k∆α
L2+k∆α
L2+k∆α
j−1fk2
j fk2
L2 ,
L2 ≤ C X(j1,j2)∈Γj (α)
L2 .
kφj1,j2bfk2
Proof. Since supp(φj1,j2) ∩ supp(φm1,m2) = ∅ if max(m1 − j1,m2 − j2) ≥ 3, the system
(φj1,j2bf ) is quasi–orthogonal. Hence, by Lemma 4.3, there exists some K > 0 such that
2
X(j1,j2)∈Γj (α)
C−1 X(j1,j2)∈Γj (α)
L2 ≤ kgjbfk2
kφj1,j2bfk2
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
φj1,j2bf(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤ K X(j1,j2)∈Γj (α)
For the converse inequality, observe that each term hφj1,j2bf , φj′
L2 ≤ X(j1,j2)∈Γj (α)
X(j1,j2)∈Γj (α)
hφj1,j2bf , φj1,j2bfiL2 +
kφj1,j2bfk2
= (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
φj1,j2bf(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
X(j1,j2)∈Γj (α)
is positive. Hence
L2
L2
2
.
(cid:3)
(cid:3)
1,j′
L2 .
kφj1,j2bfk2
2bfi =R φj1,j2(ξ)φj′
X(j1,j2)6=(j′
2)∈Γj (α)
1,j′
1,j′
2
(ξ)f (ξ)2dξ
2bfiL2
1,j′
hφj1,j2bf , φj′
4.1.3 Proof of the hyperbolic wavelet characterization of anisotropic Besov spaces
Let us first consider the general case where (p, q) ∈ (0, +∞]2,β, s ∈ R and α = (α1, α2) a
fixed anisotropy. Intermediate spaces E s,α
p,q, log β (R2) are defined as the collection of functions
k∆j1,j2fkq
p < +∞.
f of S′(R2) such as Xj≥0
A norm on E s,α
j−βq2jsq X(j1,j2)∈Γj (α)
p,q, log β (R2) is defined as follows
=Xj≥0
kfkE s,α
p,q, log β
such that the embeddings
j−βq2jsq X(j1,j2)∈Γj (α)
1/q
k∆j1,j2fkq
p
22
• if q < ∞
E s,α
p,q, log β−max(1/p−1,0)−max(1−1/q,0) (R2) ֒→ Bs,α
p,q, log β (R2) ֒→ E s,α
p,q, log β+1/q(R2) .
• if q = ∞
E s,α
p,∞, log β−max(1/p−1,0)−1(R2) ֒→ Bs,α
p,∞, log β (R2) ֒→ E s,α
p,q, log β (R2) .
are an exact rewriting of Theorem 4.1.
In the special case where p = 2, we proved in Proposition 4.4 that H s,α
2,2, log β (R2) and E s,α
Bs,α
In the following proposition, an hyperbolic wavelet characterization of spaces E s,α
p,q (R2)
is given. Combining Proposition 4.5, Theorems 4.1 and ?? directly implies Theorems 2.2
and 2.4.
2,2, log β (R2) coincide.
log β (R2) =
Proposition 4.5 Let (p, q) ∈ (0, +∞]2, s, β ∈ R2. The following assertions are equivalent
1. f ∈ E s,α
p,q, log β (R2)
2. (cid:18)Pj≥0 j−βq2jsqP(j1,j2)∈Γj
p(cid:19) 1
2−(j1+j2)q/p(cid:16)P(k1,k2)∈Z2 cj1,j2,k1,k2p(cid:17) q
3. (cid:18)P(j1,j2)∈N2
)(cid:17)−βq
Let us prove Proposition 4.5. The equivalence between assertions (2) and (3) holds since for
) + 2 − 2 ≤ j ≤ max( j1
any (j1, j2) ∈ Γj(α), one has max( j1
0. The
α1
crucial point is the equivalence between assertions (1) and (2) .
p(cid:19) 1
p (cid:17)q(cid:16)P(k1,k2)∈Z2 cj1,j2,k1,k2p(cid:17) q
) + 2 and ∪Γj = N2
0(cid:16)max( j1
2(cid:16)max( j1
< +∞.
)s− j1+j2
, j2
α2
, j2
α2
, j2
α2
α1
, j2
α2
α1
α1
q
q
< +∞.
Proof of implication (1) ⇒ (2) of Proposition 4.5
The proof of this implication relies on the following sampling lemma which is a adaptation of
Lemma 2.4 of [27] in the case of rectangular support.
Lemma 4.5 Let p ∈ (0, +∞] and j = (j1, j2) ∈ N2
{ξ, ξ1 ≤ 2j1+1 and ξ2 ≤ 2j2+1}. Then there exists C > 0 such that
X(k1,k2)∈Z2
2−(j1+j2)(cid:12)(cid:12)(cid:12)(cid:12)g(cid:18) k1
2j1
,
1/p
0. Suppose g ∈ S′(R2) and supp(bg) ⊂
p
2j2(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
≤ CkgkLp .
k2
Proof. Let ψ ∈ S(R2) be such that supp(bψ) ⊂ {ξ, max(ξ1,ξ2) ≤ π} and bψ ≡ 1 on [−2, 2]2.
Set ψj(x) = 2j1+j2ψ(2j1 x1, 2j2x2). One hascψj ≡ 1 on [−2j1+1, 2j1+1] × [−2j2+1, 2j2+1].
By assumption supp(bg) ⊂ [−2j1+1, 2j1+1] × [−2j2+1, 2j2+1], so that for any x = (x1, x2) ∈
R2 and any fixed y = (y1, y2) ∈ R2
g(x + y) = (ψj ⋆ g)(x + y) = (2π)−2Z 2j1+1
ξ2=−2j2+1cψj(ξ)bg(ξ)eixξeiyξdξ .
ξ1=−2j1+1Z 2j2+1
23
One has
Denote bhj the periodic extension of cψj with period 2ji+1π for each variable ξi (i = 1, 2).
(28)
g(x + y) = (2π)−2Z 2j1+1
ξ1=−2j1+1Z 2j2+1
ξ2=−2j2+1(cid:16)bhj(ξ)eixξ(cid:17)(cid:16)bg(ξ)eiyξ(cid:17) dξ .
Using an expansion of bhjeixξ in two dimensional Fourier series, it comes
bhj(ξ)eixξ
= X(ℓ1,ℓ2)∈Z2 Z 2j1+1π
= X(ℓ1,ℓ2)∈Z2 Z 2j1+1π
= 2−(j1+j2) X(ℓ1,ℓ2)∈Z2
ξ2=−2j2 +1π bhj(ξ)eixξe−i2−j1 ℓ1ξ1e−i2−j2 ℓ2ξ2! ei2−j1 ℓ1ξ1ei2−j2 ℓ2ξ2
ξ2=−2j2 +1πcψj(ξ)eixξe−i2−j1 ℓ1ξ1e−i2−j2 ℓ2ξ2! ei2−j1 ℓ1ξ1ei2−j2 ℓ2ξ2
ξ1=−2j1+1πZ 2j2 +1π
ξ1=−2j1+1πZ 2j2 +1π
ψj(x − 2−jℓ)ei2−j1 ℓ1ξ1ei2−j2 ℓ2ξ2 ,
where for j = (j1, j2) and ℓ = (ℓ1, ℓ2), the notation 2−jℓ = (2−j1ℓ1, 2−j2ℓ2) is used. Replacing
bhj(ξ)eixξ with the last sum in Equation (28) yields that for any x = (x1, x2) ∈ R2 and any
fixed y = (y1, y2) ∈ R2
ψj(x − 2−jℓ)ei2−j1 ℓ1ξ1ei2−j2 ℓ2ξ2(cid:16)bg(ξ)eiyξ(cid:17) dξ!
4π2 X(ℓ1,ℓ2)∈Z2 Z 2j1+1
ξ1=−2j1+1Z 2j2+1
g(2−j ℓ + y)ψj(x − 2−jℓ) .
g(x + y)
2−(j1+j2)
ξ2=−2j2+1
=
= 2−(j1+j2) X(ℓ1,ℓ2)∈Z2
Hence for all y ∈ λj1,j2,k1,k2 = [2−j1k1, 2−j1(k1 + 1)) × [2−j2k2, 2−j2 (k2 + 1))
x1≤2−j1√2,x2≤2−j2√2g(x + y)
z1−2−j1 k1≤2−j1 ,z2−2−j2 k2≤2−j2 g(z) ≤
≤ 2−(j1+j2) X(ℓ1,ℓ2)∈Z2 g(2−j ℓ + y) ·
≤ 2−(j1+j2) X(ℓ1,ℓ2)∈Z2 g(2−j ℓ + y) ·
sup
sup
1
(1 + ℓ)M
sup
max(2j1x1,2j2x2)≤√2ψj(x − 2−jℓ)
where the last inequality follows from the fast decay of ψ. Take M sufficiently large and use
either the p triangular inequality either the Holder inequality according whether p ∈ (0, 1) or
p ∈ [1, +∞]. Hence, one has
g(2−j1 k1, 2−j2k2)p ≤
z1−2−j1 k1≤2−j1 ,z2−2−j2 k2≤2−j2 g(z)p ≤ C2−(j1+j2) X(ℓ1,ℓ2)∈Z2 g(2−j ℓ+y)p·
sup
1
(1 + ℓ)M ′ ,
for some M′ > 1. An integration over y ∈ λj1,j2,k1,k2 leads to
2−(j1+j2)g(2−j1 k1, 2−j2k2)p ≤ X(ℓ1,ℓ2)∈Z2
1
(1 + ℓ)M ′ Zλj1,j2,k1,k2
g(y)pdy
24
and a sum over k ∈ Z2 gives
Xk
2−(j1+j2)g(2−j1 k1, 2−j2k2)p ≤Xk X(ℓ1,ℓ2)∈Z2
1
(1 + ℓ)3Zλj1,j2,k1,k2
g(y)pdy
which ends the proof of Lemma 4.5.
(cid:3)
Now, observe that cj1,j2,k1,k2 = ∆j1,j2f (2−j1k1, 2−j2k2). By Lemma 4.5 applied to g =
∆j1,j2f ∈ S(R2), one has
X(k1,k2)∈Z2 cj1,j2,k1,k2p = X(k1,k2)∈Z2 ∆j1,j2f (2−j1k1, 2−j2k2)p ≤ C2j12j2k∆j1,j2fkp
p ,
which is the desired wavelet characterization.
Proof of implication (2) ⇒ (1) of Proposition 4.5
To obtain the converse implication, the same approach as in the proof of Theorem 3.1 of [27]
is followed.
Since φj1,j2 and ψm1,m2,k1,k1 are both defined as a tensorial product, Lemma 3.3 of
[27]
can be applied: there exists some C > 0 such that for any α > 0 and for all x = (x1, x2) ∈ R2
one has
(29)
φj1,j2 ⋆ ψm1,m2,k1,k1(x) ≤ C
where M denotes the number of vanishing moments of the wavelets.
(1 + 2inf(j1,m1)x1 − 2−m1 k1)α(1 + 2inf(j2,m2)x2 − 2−m2 k2)α
2−(j1−m1+j2−m2)(M +3)
,
A lemma analogous to Lemma 3.4 of [27]
is now proved:
Lemma 4.6 Let p ∈ [1, +∞], ℓ1, ℓ2, m1, m2 integers such that ℓ1 ≤ m1 and ℓ2 ≤ m2. We are
also given some functions gk1,k2 satisfying the following inequality for some C > 0 :
∀x = (x1, x2) ∈ R2, gk1,k2(x) ≤
(1 + 2ℓ1x1 − 2−m1 k1)2(1 + 2ℓ2x2 − 2−m2k2)2 .
C
F = Xk=(k1,k2)∈Z2
dk1,k2gk1,k2
(30)
Set
Then
(31)
Proof. By definition of the Lp–norm, one has :
kFkLp ≤ C2−(m1+m2)/p2m1−ℓ12m2−ℓ2 · Xk=(k1,k2)∈Z2 dk1,k2p
Lp = ZR2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ Xk′=(k′
Xk=(k1,k2)∈Z2
2)∈Z2Zλm1,m2,k′
dk1,k2gk1,k2(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)
Xk=(k1,k2)∈Z2
dk1,k2gk1,k2(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,k′
2
1
1,k′
25
p
dx
kFkp
1/p
.
p
dx ,
1,k′
2
1,k′
2
sup
x∈λm1,m2,k′
where the hyperbolic dyadic cube λm1,m2,k′
are defined in (5). Observe now that, by the
usual triangular inequality and by inequality (30), there exists some C > 0 such that for any
(k1, k2) ∈ Z2, (k′1, k′2) ∈ Z2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=(k1,k2)∈Z2
≤ Xk=(k1,k2)∈Z2
Qi=1,2(1 + 2ℓi2−mik′i − 2−miki)2
(1 + 2ℓ1−m1k′1 − k1)2(1 + 2ℓ2−m2k′2 − k2)2
dk1,k2gk1,k2(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2)∈Z2 X(k1,k2)∈Z2
Lp ≤ C2−(m1+m2) X(k′
kFkp
Let us recall the usual convolution inequality, valid for any sequences s, s′ in ℓp(Z2) for p ≥ 1,
dk1,k2
Hence one has
dk1,k2
1,k′
p
,
ks ∗ s′kp
ℓp(Z2) ≤ kskp
ℓp(Z2)ks′kp
ℓ1(Z2) .
Applied to s = dk1,k2 and s′ = (1 + 2ℓ1−m1k′1 − k1)−2(1 + 2ℓ2−m2k′2 − k2)−2, it gives
kFkp
Lp ≤ C2−(m1+m2) X(k1,k2)∈Z2 dk1,k2p X(k′
1,k′
2)∈Z2
(1 + 2ℓ1−m1k′1)2(1 + 2ℓ2−m2k′2)2
1
p
,
Recall now the classical result :
Hence
1
1,k′
2)∈Z2
(1 + 2ℓ1−m1k′1)2(1 + 2ℓ2−m2k′2)2 ≤ C2m1−ℓ12m2−ℓ2
Xk′=(k′
Lp ≤ C2−(m1+m2)2(m1−ℓ1)p2(m2−ℓ2)p Xk=(k1,k2)∈Z2 dk1,k2p
p
kFkp
× Xk′=(k′
1,k′
2)∈Z2
1
Qi=1,2(1 + 2ℓi−mik′i)2
,
which directly yields the required result. It ends the proof of Lemma 4.6.
Let us now go back to Implication (2) ⇒ (1) of Proposition 4.5. Two cases are considered:
p ∈ (0, 1) and p ∈ [1, +∞].
Let us first assume that p ∈ (0, 1).
We have to bound k∆j1,j2fkLp = kφj1,j2 ⋆ fkLp. Observe that
(cid:3)
φj1,j2 ⋆ f = Xm1,m2 Xk1,k2
cm1,m2,k1,k2 (φj1,j2 ⋆ ψm1,m2,k1,k2)
By the p–triangular inequality, it comes
∀x = (x1, x2) ∈ R2, φj1,j2 ⋆ f (x)p ≤ Xm1,m2 Xk1,k2
26
cm1,m2,k1,k2p (φj1,j2 ⋆ ψm1,m2,k1,k2)(x)p
By Inequality (29), for all x = (x1, x2) ∈ R2, one has
φj1,j2⋆f (x)p ≤ Xm1,m2 Xk1,k2
cm1,m2,k1,k2p×
An integration over R2 implies that :
2−p(j1−m1+j2−m2)(M +3)
(1 + 2inf(j1,m1)x1 − 2−m1k1)pα(1 + 2inf(j2,m2)x2 − 2−m2k2)pα
kφj1,j2 ⋆ f (x)kp
cm1,m2,k1,k2p2−p(j1−m1+j2−m2)(M +3) .
Hence
kfkq
E s,α
p,q, log β
Lp ≤ Xm1,m2 Xk1,k2
)(cid:19)−βq
kcm1,m2,·,·kp
j1
α1
j2
α2
,
= Xj1,j2(cid:18)max(
≤ Xj1,j2 Xm1,m2
Set for any t ∈ R, (t)+ = max(t, 0) and
α1
Lp
, j2
α2
2qs max( j1
)kφj1,j2 ⋆ f (x)kq
ℓp2−p(j1−m1+j2−m2)(M +3)(cid:18)max(
sgn(t) =
1 if t > 0,
0 if t = 0,
−1 if t < 0.
j1
α1
,
j2
α2
)(cid:19)−βp
2ps max( j1
α1
, j2
α2
)!q/p
Observe now that for any integers j, m
and that for any integers j1, j2, m1, m2
m − (m − j)+ ≤ j ≤ (j − m)+ + m ,
max( m1
α1
, m2
α1
)
max(
(m1−j1)+
α1
max( m1
α1
1 −
,
, m2
α1
α1
)
(m2−j2)+
≤ max(
)
j1
α1
,
j2
α1
) ≤ max(
m1
α1
,
m2
α1
)(cid:20)1 + max(
(j1 − m1)+
α1
,
(j2 − m2)+
α1
)(cid:21) ,
(except in the case m1 = m2 = 0 which can be treated separately). Hence
E s,α
p,q, log β ≤Xj1,j2 Xm1,m2
)(cid:19)−βp
kfkq
sm1,m2 =(cid:18)max(
m2
α2
m1
α1
,
um1,m2vj1−m1,j2−m2·!q/p
,
2ps max( m1
α1
, m2
α2
)kcm1,m2,·,·kp
ℓp ,
with
and
s′j1,j2 = 2−p(j1+j2)(M +3)[1 + max(
(j1)+
α1
,
(j2)+
α2
)]−βp2sgn(s)ps max(
(j1)+
α1
,
(j2)+
α2
)
If q/p > 1 Young's inequality can be applied, which states that for any sequences s, s′,
ks ∗ s′kℓq/p(Z2) ≤ kskℓq/p(Z2)ks′kℓ1(Z2) ,
27
whereas if q/p ≤ 1 the usual (q/p)–triangle inequality and the usual inequality ks∗ s′kℓ1(Z2) ≤
kskℓ1(Z2)ks′kℓ1(Z2) valid for any sequence s, s′ are applied. In any case, the following inequality
is obtained
kfkq
E s,α
p,q, log β
,
m1
α1
m2
α2
≤ Xm1,m2(cid:18)max(
)(cid:19)−βp
×Xj1,j2 2−p(j1+j2)(M +3)(cid:18)max(
p,q, log β ≤ C Xm1,m2(cid:18)max(
m1
α1
E s,α
,
kfkq
2qs max( m1
α1
, m2
α2
j1
α1
,
j2
α2
)(cid:19)−βp
ℓp!
)kcm1,m2,·,·p
)!max(q/p,1)
2ps max( j1
, j2
α2
α1
.
)(cid:19)−βp
2qs max( m1
α1
, m2
α2
)kcm1,m2,·,·p
ℓp! ,
m2
α2
If the wavelets have sufficiently vanishing moments, we get that
which is the required result.
We now consider the case p ∈ [1, +∞]. In this case, observe that
∆j1,j2f = Xk1,k2
dk1,k2gk1,k2
with
and
gk1,k2 = 2(j1−m1+j2−m2)(M +3)(φj1,j2 ⋆ ψm1,m2,k1,k2) ,
dk1,k2 = 2−(j1−m1+j2−m2)(M +3)cj1,j2,k1,k2 .
We set ℓ1 = inf(j1, m1) and ℓ2 = inf(j2, m2). Lemma 4.6 gives
k∆j1,j2fkLp ≤ C2−p(j1−m1+j2−m2)(M +3)2−(m1+m2)/pkcm1,m2,·,·p
ℓp2m1−ℓ12m2−ℓ2
Again two cases q ≤ 1 and q > 1 are distinguished and the same approach than in the case
p ∈ (0, 1) is followed. It leads to the required conclusion.
4.2 Proof of Theorem 3.1
First a two–microlocal criterion is proved.
Proposition 4.6
1. Assume that f ∈ Cs,α(x0). Then there exists some C > 0 such that
for any (j1, j2, k1, k2) ∈ (N ∪ {−1})2 × Z2,
k1
cj1,j2,k1,k2 ≤ C min(2− j1 s
(32)
α1 +(cid:12)(cid:12)(cid:12)(cid:12)
2j1 − a(cid:12)(cid:12)(cid:12)(cid:12)
s
α1
, 2− j2 s
s
α2
) .
α2 +(cid:12)(cid:12)(cid:12)(cid:12)
k2
2j2 − b(cid:12)(cid:12)(cid:12)(cid:12)
2. Conversely, assume that f is uniformly Holder and that (32) holds, then f ∈ Cs,α
log 2(x0).
28
(f (x1, x2) − Px0(a, x2))ψ(2j1 x1 − k1)ψ(2j2x2 − k2)dx1dx2
(f (x1, x2) − Px0(x1, b))ψ(2j1 x1 − k1)ψ(2j2 x2 − k2)dx1dx2
ψ(2j1 x1 − k1)ψ(2j2 x2 − k2)dx1dx2
s/α1Z ψ(u1)ψ(u2)du1du2! .
Proof. Let us first assume that f ∈ Cs,α(x0) with x0 = (a, b). Assume that j1 6= −1 and
j2 6= −1. By definition of the hyperbolic wavelet coefficients one has
f (x1, x2)ψ(2j1 x1 − k1)ψ(2j2 x2 − k2)dx1dx2
Since ψ admits at least one vanishing moment, the two following equalities hold
and
(34)
(33)
cj1,j2,k1,k2 = 2j1+j2ZR2
cj1,j2,k1,k2 = 2j1+j2ZR2
cj1,j2,k1,k2 = 2j1+j2ZR2
Equality (33) and the assumption f ∈ Cs,α(x0) imply that
cj1,j2,k1,k2 ≤ 2j1+j2Z x1 − as
2j1 − as/α1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ 2j1+j2ZR2 (cid:12)(cid:12)(cid:12)(cid:12)x1 −
α1 ZR2 u1s/α1ψ(u1)ψ(u2)du1du2 +(cid:12)(cid:12)(cid:12)(cid:12)
2j1 − a(cid:12)(cid:12)(cid:12)(cid:12)
α1 +(cid:12)(cid:12)(cid:12)(cid:12)
2j1 − a(cid:12)(cid:12)(cid:12)(cid:12)
α2 +(cid:12)(cid:12)(cid:12)(cid:12)
2j2 − b(cid:12)(cid:12)(cid:12)(cid:12)
We now set u1 = 2j1x1 − k1, u2 = 2j2x2 − k2 and deduce that
cj1,j2,k1,k2 ≤ 2− j1s
Hence for some C depending only on ψ, s and α one has
cj1,j2,k1,k2 ≤ C(2− j2s
cj1,j2,k1,k2 ≤ C(2− j1 s
A similar approach yields that
2j1(cid:12)(cid:12)(cid:12)(cid:12)
+(cid:12)(cid:12)(cid:12)(cid:12)
k1
k2
k1
s/α1
s/α2
)
)
αψ(2j1 x1 − k1)ψ(2j2 x2 − k2)dx1dx2
k1
s/α1
k1
This shows that (32) can be read as a necessary condition for pointwise regularity of function
f .
Let us now prove the converse result. Assuming that (32) holds, the aim first consists in
defining a polynomial approximation of f at x0. To that end, a Taylor expansion is used to
investigate the differentiability of f at x0. Let us define fj as:
fj = X(j1,j2)∈Γj (α) X(k1,k2)∈Z2
cj1,j2,k1,k2ψj1,j2,k1,k2 .
where the notations are the same as in the proof of Proposition 2.3. One has
min(2−j1s/α1 + k1
fj(x) ≤ X(j1,j2)∈Γj X(k1,k2)∈Z2
2−js + k2
(1 + 2j1x1 − k1)N (1 + 2j2x2 − k2)N +Xj2≤j
2j − x2s/α2 + x2 − bs/α2
≤ Xj1≤j Xk1,k2
(1 + 2j1x1 − k1)N (1 + 2j2x2 − k2)N
2j1 − as/α1, 2−j2s/α2 + k2
2j2 − as/α2)
2−js + k1
2j1 − x1s/α1 + x1 − as/α2
(1 + 2j1x1 − k1)N (1 + 2j2x2 − k2)N
29
Then
(35)
fj(x) ≤ C(j2−js + jx1 − as/α1 + jx2 − bs/α2) .
In the same way, if β = (β1, β2), one has
∂βfj ≤ X(j1,j2)∈Γj
2j1β1+j2β2 X(k1,k2)∈Z2
min(2−j1s/α1 + k1
2j1 − as/α1, 2−j2s/α2 + k2
2j2 − as/α2 )
(1 + 2j1x1 − k1)N (1 + 2j2x2 − k2)N
.
Then
(36)
∂βfj(x) ≤ C2j(β1α1+β2α2)(2−js + x1 − as/α1 + x2 − bs/α2) .
So, the function f is β–differentiable at x0 provided that β1α1 + β2α2 ≤ s. The Taylor
polynomial of f at x0 is defined by
and
Pj,x0(x) = Xβ1α1+β2α2≤s
Px0(x) =Xj
(x − x0)β
β!
∂βfj(x0)
Pj,x0(x) .
We shall now bound f (x)−Px0(x) in the neighborhood of x0. Recall that f is assumed to
be uniformly Holder, namely there exists some ε∗0 > 0 such that f ∈ Cε∗
0 (R2). The inclusions
between Holder spaces with different anisotropies (see [53]) leads to the existence of ε0 such
that f ∈ Cε0,α(R2). Set J1 = [αJ/ε0]. Observe that
f (x) − Px0(x) ≤Xj≤J
fj(x) + Xj>J1
fj(x) +Xj>J
fj(x) − Pj,x0(x) +
J1Xj=J+1
Pj,x0(x) .
Let us now bound each term of the right hand side of this inequality.
We first deal with the term corresponding to j ≤ J. In this case we shall use an anisotropic
version of Taylor inequality which can be found in [19], [26] and recalled in [9]. It gives the
existence of some C > 0 such that
fj(x) − Pj,x0(x) ≤ C
X
β1+β2≤k+1, α1β1+α2β2>s
x − x0α1β1+α2β2
α
sup
z=(z1,z2)∈R2 ∂βfj .
with k = [max(s/α1, s/α2)]. The bound (36) implies that there exists some C > 0 such that
fj(x)−Pj,x0(x) ≤ C
X
β1+β2≤k+1, α1β1+α2β2>s
x−x0α1β1+α2β2
α
2j(β1α1+β2α2)(2−js+x1−as/α1+x2−bs/α2)
Hence,
Xj≤J
fj(x)−Pj,x0(x) ≤ C
X
β1+β2≤k+1, α1β1+α2β2>s
x−x0α1β1+α2β2
α
(2J(β1α1+β2α2−s)+2J(β1α1+β2α2)x−x0s
α) .
30
Since x − x0α ≤ 2−J it comes
Xj≤J
(37)
Let us now bound the sumPJ1
J1Xj=J
J1Xj=J+1
fj(x) ≤
on J, one has
(38)
fj(x) − Pj,x0(x) ≤ Cx − x0s
α
j=J+1 fj(x). By (35) and the definition of J1 which depends
(j2−js + jx − x0s
α) ≤ J2−Js + J 2x − x0s
α .
To bound the sumPj>J1 fj(x) the uniform regularity of f is used, leading to
(39)
Xj>J1
fj(x) ≤ C2−J1ε0 ≤ C2−Js
2j(β1α1+β2α2−s)
has
the last equality following from the definition of J1.
β! Xj>J
(x − x0)β
Finally, by (36), the sumPj>J Pj,x0(x) can be bounded. Indeed, for some C > 0, one
Xj>J
Pj,x0(x) ≤ Xβ1α1+β2α2<s
Xj>J
x1 − aβ1x2 − bβ2
Since x1 − a ≤ x − x0α1
(40) Xj>J
Finally, Inequalities (37), (38), (39) and (40) yield that f ∈ Cs,α
α ≤ 2−Jα1 and x2 − b ≤ x − x0α2
2−J(β1α1+β2α2)Xj>J
α ≤ 2−Jα2 it comes
2j(β1α1+β2α2−s) ≤ C2−Js .
∂βfj(x0) ≤ C Xβ1α1+β2α2<s
Pj,x0(x) ≤ C Xβ1α1+β2α2<s
log 2(x0).
β!
(cid:3)
Theorem 3.1 is a straightforward consequence of the two–microlocal criterion and of the
following lemma:
Lemma 4.7 The two following properties are equivalent:
(i) Inequality (32) holds.
(ii) Inequality (6) holds.
Proof. Assume that (32) holds. If λ′ ⊂ 3λj1,j2(x0), then
j′1 ≥ j1, j′2 ≥ j2 ,
and
Condition (32) implies
k′1
1 − a ≤ 2.2−j′
2j′
1 and
k′2
2 − b ≤ 2.2−j′
2j′
2 .
cλ′ ≤ min(2− j1s
α1 , 2− j2s
α2 ) = 2− max( j1
α1
, j2
α2
)s .
31
Conversely, assume that (6) holds. Let λ′ = λ(j′1, j′2, k′1, k′2) an hyperbolic dyadic cube.
Set
and
j1 = sup{ℓ1, 2−j′
k′1
1 − a ≤ 2−ℓ1}
2j′
k′2
2 − b ≤ 2−ℓ2}
2 +
2j′
We have λ′ ⊂ 3λj1,j2(x0). Since (6) holds one has
j2 = sup{ℓ2, 2−j′
1 +
cλ′ ≤ min(2− j1 s
α1 , 2− j2 s
α2 ) ≤ C min(2−
that is 32 holds.
4.3 Proof of Theorem 3.2
j′
1 s
α1 +(cid:12)(cid:12)(cid:12)(cid:12)
k′1
2j′
1 − a(cid:12)(cid:12)(cid:12)(cid:12)
s/α1
, 2−
s
α2
) ,
j′
2 s
α2 +(cid:12)(cid:12)(cid:12)(cid:12)
k′2
2j′
2 − b(cid:12)(cid:12)(cid:12)(cid:12)
(cid:3)
The proof of Theorem 3.2 is based on the two following lemmas, analogous to Propositions 7
and 8 of [35]:
Lemma 4.8 Set α = (a, 2 − a) and define
G(H, α) = {x ∈ R2, f 6∈ CH,α
log 2(x)} .
Let p > 0 and s ∈ (0, ω(p, α)/p]. Then for any H ≥ s − 2/p
dimH(G(H, α)) ≤ Hp − sp + 2 .
If H < s − 2/p, dimH(G(H, α)) = ∅.
Lemma 4.9 Set α = (a, 2 − a) and define
B(H, α) = {x ∈ R2, f ∈ CH,α(x)} .
Let p < 0 and s ∈ (0, ω(p, α)/p]. Then
dimH(B(H, α)) ≤ dimP (B(H, α)) ≤ Hp − sp + 2 .
The proof of Lemma 4.8 in the case H ≥ s−2/p is exactly the same as this of Proposition 7
of [35], except that the set Gj,H are replaced with the sets
G(j, H, α) = {λ = λ(j1, j2, k1, k2), (j1, j2) ∈ Γj(α),dλ ≥ 2−jHp} .
Lemma 4.8 in the case H < s − 2/p, comes from the hyperbolic wavelet characterization of
anisotropic Besov spaces stated in Theorem 2.2 and the Sobolev embeddings which can be
proved in the anisotropic case as in the isotropic one (see [53]).
The proof of Lemma 4.9 is exactly the same as this of Proposition 8 of [35], except that
the set BH are replaced with the sets B(H, α).
Lemmas 4.8 and 4.9 then imply Theorem 3.2, since for any α = (α1, α2) such that α1+α2 =
2 one has
E(H, α) ⊂(cid:0)∩H ′>HG(H′, α)(cid:1) ∩(cid:0)∪H ′<H B(H′, α)(cid:1) .
Acknowledgements. We warmly thank Florent Autin and Jean Marc Freyermuth for
many stimulating discussions about applications of non parametric statistics to the analysis
of anisotropic textures as well as Laurent Duval for giving us very interesting additional
references about hyperbolic wavelet analysis.
32
References
[1] P. Abry, S. Jaffard, and W. Wendt, Irregularities and scaling in signal and Image
processing: Multifractal analysis, "Benoit Mandelbrot: A Life in Many Dimensions"
World Scientific M. Frame Ed., 2012.
[2] H. Aimar and I. G´omez, Parabolic Besov regularity for the heat equation., Constr.
Approx., 36 (2012), pp. 145–159.
[3] F. Autin, G. Claeskens, and J.-M. Freyermuth, Hyperbolic wavelet thresholding
rules: the curse of dimensionality through the maxiset approach., Submitted, (2012).
[4] A. Ayache, Hausdorff dimension of the graph of the Fractional Brownian Sheet., Rev.
Mat. Iberoamericana, 20 (2004), pp. 395–412.
[5] A. Ayache, S. Leger, and M. Pontier, Les ondelettes `a la conquete du Drap Brown-
ien Fractionnaire., C. R. Math. Acad. Sci. Paris, 335 (2002), pp. 1063–1068.
[6] A. Ayache, F. Roueff, and Y. Xiao, Local and asymptotic properties of Linear
Fractional Stable Sheets., C. R. Math. Acad. Sci. Paris, 344 (2007), pp. 389–394.
[7]
, Linear Fractional Stable Sheets: Wavelet expansion and sample path properties.,
Stochastic Processes and their Applications, 119 (2009), pp. 1168–1197.
[8] A. Ayache and Y. Xiao, Asymptotic properties and Hausdorff dimensions of Fractional
Brownian Sheets., J. Fourier Anal. Appl., 11 (2005), pp. 407–439.
[9] H. Ben Braiek and M. Benslimane, Baire generic anisotropic multifractal formalism
in criteria in anisotropic function spaces, Submitted, (2011).
[10]
, Directional and anisotropic regularity and irregularity criteria in triebel wavelet
bases, Submitted, (2011).
[11] M. Benslimane, Multifractal formalism and anisotropic selfsimilar functions, Math.
Proc. Camb. Philos. Soc., 124 (1998), pp. 329–363.
[12] G. Beylkin, Wavelets and fast numerical algorithms, in Proceedings of Symposia of
Applied Math., vol. 47, 1993.
[13] G. Beylkin, R. Coifman, and V. Rokhlin, Fast wavelet transforms and numerical
algorithms I, J. Comm. Pure Appl. Math., 44 (1991), pp. 141–183.
[14] H. Bierm´e, M. Meerschaert, and H. Scheffler, Operator scaling stable random
fields., Stoch. Proc. Appl., 117 (2009), pp. 312–332.
[15] A. Bonami and A. Estrade, Anisotropic analysis of some Gaussian models., J. Fourier
Analysis and Applications, 9 (2003), pp. 215–236.
[16] M. Bownik, Anisotropic Hardy spaces and wavelets, Mem. AMS, 781 (2003).
[17]
, Atomic and molecular decomposition of anisotropic Besov spaces, Math. Z., 250
(2005), pp. 539–571.
33
[18] M. Bownik and K. P. Ho, Atomic and molecular decomposition of anisotropic Triebel–
Lizorkin spaces, Trans of Amer Math Soc, 385 (2005), pp. 1469–1510.
[19] A. Calder´on and A. Torchinsky, Parabolic maximal functions associted with a dis-
tribution, Adv. Math., 24 (1977), pp. 101–171.
[20] E. Candes and L. Demanet, The curvelet representation of wave propagators is opti-
mally sparse, Comm. Pure Appl. Math., 58 (2005), pp. 1472– 1508.
[21] M. Clausel and B. Vedel, Two optimality results about sample paths properties of
Operator Scaling Gaussian Random Fields, Submitted, (2010).
[22] S. Davies and P. Hall, Fractal analysis of surface roughness by using spatial data
(with discussion)., J. Roy. Statist. Soc. Ser., B 61 (1999), pp. 3–37.
[23] R. A. DeVore, S. V. Konyagin, and V. N. Temlyakov, Hyperbolic wavelet approx-
imation, Constr. Approx., 14 (1998), pp. 1–26.
[24] D. Donoho, Wedgelets: Nearly minimax estimation of edges, Ann. Stat., 27 (1999),
pp. 353–382.
[25] K. Falconer, Fractal Geometry: Mathematical Foundations and Applications, Wiley
and sons, 1990.
[26] G. Folland and E. Stein, Hardy spaces on homogeneous groups., Mathematical Notes,
28,Princeton University Press and University of Tokyo Press, 1982.
[27] M. Frazier and B. Jawerth, Decomposition of besov spaces, Indiana University Math-
ematics Journal, 34 (1985), pp. 777–799.
[28] G. Garrig´os and A. Tabacco, Wavelet decompositions of anisotropic Besov spaces,
Math. Nachr., 239–240 (2002), pp. 80–102.
[29] K. Guo and D. Labate, Representation of Fourier integral operators using shearlets,
The Journal of Fourier Analysis and Applications, 14 (2008), pp. 327–370.
[30]
, Analysis and detection of surface discontinuities using the 3d continuous shearlet
transform, Appl.Comp.Harm. Anal., 30 (2011), pp. 231–242.
[31] D. Haroske and E. Tam´asi, Wavelet frames for distributions in anisotropic Besov
spaces, Georg. Math. J, 12 (2005), pp. 637–658.
[32] R. Hochmuth, N-term approximation in anisotropic function spaces., Math. Nachr.,
244 (2002), pp. 131–149.
[33]
, Wavelet characterizations for Anisotropic Besov Spaces, Appl. Comput. Harmon.
Anal., 12 (2002), pp. 179–208.
[34] L. Jacques, L. Duval, C. Chaux, and G. Peyr´e, A panorama on Multiscale Geo-
metric Representations, Intertwining Spatial, Directional and Frequency Selectivity, Sub-
mitted, (2012).
34
[35] S. Jaffard, Wavelet techniques in multifractal analysis, fractal geometry and applica-
tions, Proc. Symp. Pure Math., 72 (2004), pp. 91–151.
[36]
, Pointwise and directional regularity of nonharmonic Fourier series, Appl. Comp.
Harm. Anal., 28 (2010), pp. 251–266.
[37] P. Lakhonchai, J. Sampo, and S. Sumetkijakan, Shearlet transforms and Holder
regularities, International Journal of Wavelets, Multiresolution and Information Process-
ing, To appear (2012).
[38] T. Long and H. Triebel, Equivalent norms and Schauder bases in anisotropic besov
spaces, Proc. Roy. Soc. Endinb., 84 A (1979), pp. 177–183.
[39] Y. Meyer, Ondelettes et Op´erateurs : vol. I., Paris, Hermann, 1990.
[40] M. Neumann, Multivariate wavelet thresolding in anisotropic function spaces, Statistica
sinica, 10 (2000), pp. 399–431.
[41] M. Neumann and R. von Sachs, Wavelet thresolding in anisotropic function classes
and application to adaptative estimation of evolutionary spectra, Annals of Statistics, 25
(1997), pp. 38–76.
[42] K. Nualtong and S. Sumetkijakan, Analysis of Holder regularities by wavelet–like
transforms with parabolic scaling,, Thai Journal of Mathematics, (2005), pp. 275–283.
[43] G. Parisi and U. Frisch, On the singularity spectrum of fully developped turbulence,
in Turbulence and Predictability in Geophysical Fluid Dynamics, Proceedings of the
International Summer School in Physics Enrico Fermi,North Holland, (1985), pp. 84–87.
[44] L. Ponson, D. Bonamy, H. Auradou, G. Mourot, S. Morel, E. Bouchaud,
C. Guillot, and J. Hulin, Anisotropic self-affine properties of experimental fracture
surfaces., Int.Journ. of fracture, 140 (2006), pp. 27–37.
[45] S. Roux, M. Clausel, B. Vedel, S. Jaffard, and P. Abry, The Hyperbolic Wavelet
Transform for self-similar anisotropic texture analysis, Submitted, (2012).
[46] J. Sampo and S. Sumetkijakan, Estimations of Holder regularities and direction
of singularity by hart smith and curvelet transforms, Journal of Fourier Analysis and
Applications, 15 (2009), pp. 58–79.
[47] H. Schmeisser and H. Triebel, Spaces of functions of mixed smoothness and their
relations to approximation from hyperbolic crosses, Journal of Approximation Theory,
128 (2004), pp. 115–150.
[48] W. Sickel and T. Ullrich, Tensor products of Sobolev–Besov spaces and applications
to approximation from the hyperbolic cross, Journal of Approximation Theory, 161 (2009),
pp. 748–786.
[49] H. Smith, A Hardy space for Fourier integral operators, J. Geom. Anal., 8 (1998),
pp. 629–653.
[50] C. Tricot, Two definitions of fractional dimension, Math. Proc. Cambridge Philos.
Soc., 18 (1991), pp. 54–74.
35
[51] H. Triebel, Interpolation theory, function spaces, differential operators., Amsterdam,
North-Holland, 1978.
[52]
, Wavelet bases in anisotropic function spaces., Function Spaces, Differential Opera-
tors and Nonlinear Analysis, FSDONA-04. Milovy, Czech Republic., 18 (2004), pp. 529–
550.
[53]
, Theory of function spaces III, Birhauser,Basel, 2006.
[54] P. H. Westerink, Subband coding of images, PhD thesis, Delft University of Technol-
ogy, Delft, The Netherlands, 1989.
36
|
1905.05152 | 1 | 1905 | 2019-05-13T17:10:00 | Characterizing compact families via the Laplace transform | [
"math.FA"
] | In 1985, Robert L. Pego characterized compact families in $L^2(\reals)$ in terms of the Fourier transform. It took nearly 30 years to realize that Pego's result can be proved in a wider setting of locally compact abelian groups (works of G\'orka and Kostrzewa). In the current paper, we argue that the Fourier transform is not the only integral transform that is efficient in characterizing compact families and suggest the Laplace transform as a possible alternative. | math.FA | math | Characterizing compact families via the Laplace
transform
Mateusz Krukowski
Łódź University of Technology, Institute of Mathematics,
Wólczańska 215, 90-924 Łódź, Poland
9
1
0
2
y
a
M
3
1
]
.
A
F
h
t
a
m
[
1
v
2
5
1
5
0
.
5
0
9
1
:
v
i
X
r
a
May 14, 2019
Abstract
In 1985, Robert L. Pego characterized compact families in L2(R) in
terms of the Fourier transform. It took nearly 30 years to realize that
Pego's result can be proved in a wider setting of locally compact abelian
groups (works of Górka and Kostrzewa).
In the current paper, we ar-
gue that the Fourier transform is not the only integral transform that
is efficient in characterizing compact families and suggest the Laplace
transform as a possible alternative.
Keywords : Laplace transform, Pego theorem, compactness
AMS Mathematics Subject Classification: 44A10 (primary), 42A38 (sec-
ondary)
1 Introduction
Characterizing compact families is a vital topic in function spaces' theory
at least since the end of the 19-th century. Around 1883, two Italian math-
ematicians Cesare Arzelà (1847-1912) and Giulio Ascoli (1843-1896) provided
the necessary and sufficient conditions under which every sequence of a given
1
family of real-valued continuous functions (defined on a closed and bounded
interval), has a uniformly convergent subsequence (this is called sequential com-
pactness). A couple of decades later (in 1931), Andrey Kolmogorov (1903-1987)
succeeded in characterizing the compact families in Lp(RN ), when 1 < p < ∞
and all the functions are supported in a common bounded set (comp. [13]). A
year later, Jacob David Tamarkin (1888-1945) got rid of the second restriction
(comp. [19]) and in 1933, Marcel Riesz (1886-1969), a younger brother of Frigyes
Riesz, proved the general case for Lp(RN ), where 1 6 p < ∞. In 1940, a French
mathematician and one of the leaders of the Bourbaki group, André Weil (1906-
1998) wrote a book 'L'intégration dans les groupes topologique' (comp. [20]), in
which he proved the Kolmogorov-Riesz theorem for a locally compact Hausdorff
group G instead of RN .
The next major contribution came over 40 years later (1985), when Robert
L. Pego characterized compact families in L2(R) via the Fourier transform. This
innovative idea was the cornerstone for the works of two Polish mathematicians
Przemysław Górka and Tomasz Kostrzewa. In [4] and [5], they proved that a
counterpart of Pego theorem holds for locally compact abelian groups (this is
reminiscent of Weil's contribution). Obviously, there are other works related to
the topic, which are worth-mentioning: [6], [7], [8], [9], [10] or [11] just to name
a few.
In the current paper, we argue that the Fourier transform is not the only one
that can be used to characterize compact families. In Section 2 we introduce
the basic definitions and discuss the necessary notation. We also prove the
fundamental theorems, which are very well-known in the context of the Fourier
transform, and probably less known in the context of the Laplace transform. In
Section 3 we prove the main results. Theorem 11, which is a counterpart of the
Pego's result, is the climax of the paper.
2 Preliminary results
Throughout the paper, by R+ we understand the open set (0,∞) and C
stands for the field of complex numbers. For a measurable, complex-valued
function f : R+ −→ C and a real number x > 0 we denote
fx(t) = f (t)e−xt.
We say that f : R+ −→ C is a Laplace-Pego function of order x > 0 if
fx ∈ L1(R+) ∩ L2(R+).
2
The norms in L1(R+) and L2(R+) are denoted by k · k1 and k · k2, respectively.
Moreover, if F is a family of Laplace-Pego functions with a common order x > 0,
then we denote
Let f be a Laplace-Pego function of order x > 0. The Laplace transform
Fx = {fx : f ∈ F}.
L{f} of the function f is defined by
L{f}(z) =Z ∞
0
f (t)e−zt dt.
A natural question arises: when does the above integral exist? To answer this
question, observe that if Re(z) > x, then
L{f}(z) =(cid:12)(cid:12)(cid:12)(cid:12)Z ∞
=Z ∞
0
0
f (t)e−Re(z)te−iIm(z)t dt(cid:12)(cid:12)(cid:12)(cid:12) 6Z ∞
f (t)e−xte(x−Re(z))t dt 6 kfxk1 < ∞.
0
f (t)e−Re(z)t dt
In other words, the Laplace transform L{f} exists in the half-plane Re(z) > x.
An important special case of the Laplace transform is the Fourier transform,
which we define by
Let us formulate a crucial theorem regarding the Laplace transform, which
we will use multiple times throughout the paper:
bf (y) = L{f}(2πiy).
Theorem 1. (Plancherel theorem for the Laplace transform)
If f is a Laplace-Pego function of order x > 0, then
1
2π Z ∞
−∞ L{f}(x + iy)2 dy =Z ∞
0
e−2xtf (t)2 dt.
Proof. At first, observe that
(1)
(2)
∀y∈R L{f}(x + iy) =Z ∞
=Z ∞
0
0
f (t)e−xte−iyt dt
f (t)e−xte−2πi y
2π
t dt = cfx(cid:18) y
2π(cid:19) .
By the classical Plancherel theorem (Theorem 3.5.2 in [1], p. 53 or Theorem 1.1
in [18], p. 208) we have
2
2π(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
−∞ (cid:12)(cid:12)(cid:12)(cid:12)cfx(cid:18) y
Z ∞
d
y
2π
−∞ fx(t)2 dt =Z ∞
=Z ∞
0
e−2xtf (t)2 dt.
3
Upon observing that
2
2π(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
−∞ (cid:12)(cid:12)(cid:12)(cid:12)cfx(cid:18) y
Z ∞
we conclude the proof.
(2)
=
d
y
2π
1
2π Z ∞
−∞ L{f}(x + iy)2 dy
The theorem, which we present below, is (again) a counterpart of a well-know
result in the theory of Fourier transform:
Theorem 2. (Riemann-Lebesgue lemma for the Laplace transform)
If f is a Laplace-Pego function of order x > 0, then
lim
y→±∞ L{f}(x + iy) = 0.
(3)
Proof. At first, let f = 1(a,b) where (a, b) ⊂ R+. Then
∀y∈R L{f}(x+iy) =Z ∞
1(a,b)(t)e−(x+iy)t dt =Z b
0
a
e−(x+iy)t dt =
e−(x+iy)a − e−(x+iy)b
x + iy
,
so (3) holds. By linearity of the Laplace transform, the result is also true for all
simple functions.
Finally, let f be an arbitrary Laplace-Pego function and let ε > 0. Since
simple functions are dense in L1(R+) (comp. Proposition 6.7 in [3], p. 183)
there exists a simple function g such that
Z ∞
0
f (t)e−xt − g(t) dt < ε.
(4)
Hence
y→±∞ L{f}(x + iy) = lim
lim
6Z ∞
f (t)e−xte−iyt dt(cid:12)(cid:12)(cid:12)(cid:12)
y→±∞ (cid:12)(cid:12)(cid:12)(cid:12)Z ∞
g(t)e−iyt dt(cid:12)(cid:12)(cid:12)(cid:12)
y→±∞ (cid:12)(cid:12)(cid:12)(cid:12)Z ∞
(cid:12)(cid:12)(cid:12)f (t)e−xt − g(t)(cid:12)(cid:12)(cid:12) dt + lim
2π(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) = ε,
y→±∞ (cid:12)(cid:12)(cid:12)(cid:12)bg(cid:18) y
y→±∞ L{g}(iy) = ε + lim
(4)
< ε + lim
0
0
where the last equality follows from the classical Riemann-Lebesgue lemma for
the Fourier transform (comp. Theorem 1.7 in [12], p. 136). Since ε > 0 was
chosen arbitrarily, we conclude the proof.
0
4
We will now recall the prominent fact that the Laplace trasnform 'changes
the convolution of two functions to multiplication'. A convolution of two Laplace-
Pego functions f, g with a common order x > 0 is defined by
∀t>0 f ⋆ g(t) =Z t
0
f (s)g(t − s) ds.
In order to prove that the convolution is well-defined, let us invoke a general
version of Tonelli's theorem (comp. Theorem B.3.3 in [2], p. 289):
Theorem 3. (Tonelli's theorem)
Let (X, µX), (Y, µY ) be two measure spaces and let F : X × Y −→ C be a
measurable function such that
n(x, y) ∈ X × Y : F (x, y) 6= 0o
is σ−finite. If one of the integrals
ZX ZY F (x, y) dµX(x) dµY (y) or
ZX ZY F (x, y) dµX(y) dµY (x)
is finite, then
ZX ZY
F (x, y) dµX(x) dµY (y) =ZX ZY
F (x, y) dµY (y) dµX(x).
In our case, both X and Y are the space R+ and both measures µX and µY
are the standard Lebesgue measure. Consequently, the assumption of
n(x, y) ∈ R+ × R+ : F (x, y) 6= 0o
being σ−finite becomes obsolete, since R+ × R+ is σ−finite.
Theorem 4. If f, g are Laplace-Pego functions with a common order x > 0,
then (f ⋆ g)x ∈ L1(R+). In particular, it exists almost everywhere.
Proof. At first, let us observe that
∀t>0 e−xtf ⋆ g(t) =Z t
=Z t
0
0
e−xtf (s)g(t − s) ds
e−xsf (s)e−x(t−s)g(t − s) ds.
(5)
5
Furthermore, by Proposition 3.9, p. 86 in [18], we note that the function
F : R+ × R+ −→ C defined by
F (t, s) = e−xsf (s)e−x(t−s)g(t − s)
is measurable, so we are in position to apply Tonelli's theorem. Consequently,
we obtain
Z ∞
0
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
e−xtf ⋆ g(t) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Tonelli's thm=
0
(5)
6Z ∞
Z ∞
0
0
Z t
Z ∞
s
e−xsf(s)e−x(t−s)g(t − s) ds dt
e−xsf(s)e−x(t−s)g(t − s) dt ds = kfxk1 kgxk1 < ∞,
which ends the proof.
Theorem 5. (convolution theorem for the Laplace transform, comp. Theorem
2.39 in [16], p. 92)
If f and g are Laplace-Pego functions with a common order x > 0, then
L{f ⋆ g}(z) = L{f}(z) · L{g}(z)
for Re(z) > x.
Proof. By Proposition 3.9, p. 86 in [18], we note that the function F : R+ × R+ −→ C
defined by
is measurable, so we are in position to apply Tonelli's theorem. We have
F (t, s) = f (s)g(t − s)e−zt
L{f ⋆ g}(z) =Z ∞
0
Tonelli's thm=
f ⋆ g(t)e−zt dt =Z ∞
Z ∞
Z ∞
0
0
Z t
0
f (s)g(t − s)e−zt ds dt
f (s)g(t − s)e−zt dt ds = L{f}(z) · L{g}(z),
s
which ends the proof.
3 Main results
A family F of Laplace-Pego functions with a common order x > 0 is said to
be exponentially L2−equivanishing at x, if the family Fx is L2−equivanishing,
i.e.
∀ε>0 ∃T >0 ∀f ∈F Z ∞
T
6
e−2xtf (t)2 dt < ε.
(6)
Furthermore, we say that a family F is Laplace equicontinuous at x, if
−∞ L{f}(x + iy + δ) − L{f}(x + iy)2 dy < ε.
∀ε>0 ∃δ>0 ∀f ∈F
We will now relate the concepts of Laplace equicontinuity and exponential
2π Z ∞
(7)
1
L2−equivanishing.
Theorem 6. Let F be a Laplace-Pego family with a common order x > 0. If F
is Laplace equicontinuous at x, then it is exponentially L2−equivanishing at x.
Furthermore, if Fx is L2−bounded, then the implication can be reversed.
Proof. We divide the proof into two parts:
Part 1.
At first, we assume that F is Laplace equicontinuous at x, so for a fixed
ε > 0 we may choose δ > 0 according to (7). Let T > 0 be such that
Consequently, for every f ∈ F we obtain
2
(cid:12)(cid:12)(cid:12)e−δT − 1(cid:12)(cid:12)(cid:12)
>
1
2
.
(8)
dt
2
ε >
=
Z T
0
Theorem 1
=
1
1
−∞ L{f}(x + iy + δ) − L{f}(x + iy)2 dy
2π Z ∞
f (t)(cid:16)e−δt − 1(cid:17) e−(x+iy)t dt(cid:12)(cid:12)(cid:12)(cid:12)
−∞ (cid:12)(cid:12)(cid:12)(cid:12)Z ∞
2π Z ∞
e−2xtf (t)2(cid:12)(cid:12)(cid:12)e−δt − 1(cid:12)(cid:12)(cid:12)
2Z ∞
dt +Z ∞
e−2xtf (t)2 dt,
e−2xtf (t)2(cid:12)(cid:12)(cid:12)e−δt − 1(cid:12)(cid:12)(cid:12)
2
dy
(8)
>
1
T
T
0
2
which ends the first part of the proof.
Part 2.
At this point, we assume that Fx is L2−bounded, so there exists M > 0
such that
∀f ∈F Z ∞
0
e−2xtf (t)2 dt < M.
We will show that if F is exponentially L2−equivanishing at x, then it is Laplace
equicontinuous at x.
7
Fix ε > 0 and choose T > 0 as in the definition of the exponential L2−equivanishing
(6). Let δ > 0 be such that
2
∀f ∈F (cid:12)(cid:12)(cid:12)e−δT − 1(cid:12)(cid:12)(cid:12)
M < ε.
(9)
We have
Theorem 1
=
1
2π Z ∞
Z T
e−2xtf (t)2(cid:12)(cid:12)(cid:12)e−δt − 1(cid:12)(cid:12)(cid:12)
M +Z ∞
6 (cid:12)(cid:12)(cid:12)e−δT − 1(cid:12)(cid:12)(cid:12)
(9)
T
2
0
2
which ends the proof.
−∞ L{f}(x + iy + δ) − L{f}(x + iy)2 dy
dt +Z ∞
e−2xtf (t)2 dt
e−2xtf (t)2(cid:12)(cid:12)(cid:12)e−δt − 1(cid:12)(cid:12)(cid:12)
(6)
< 2ε,
T
2
dt
A family F of Laplace-Pego functions with a common order x > 0 is said to
be exponentially L2−equicontinuous at x, if
∀ε>0 ∃δ>0 ∀s∈(0,δ)
f ∈F (cid:18)Z ∞
0
e−2xtf (t) − f (t − s)2 dt(cid:19) 1
2
< ε.
(10)
Furthermore, we say that a family F is Laplace equivanishing at x, if
∀ε>0 ∃T >0 ∀f ∈F ZR\[−T,T ] L{f}(x + iy)2 dy < ε.
(11)
We study the relationship between the novel notion of the exponential L2−equicontinuity
of F and the classical equicontinuity of Fx in the lemma below:
Lemma 7. Let F be a family of Laplace-Pego functions with a common order
x > 0. If Fx is L2−bounded then F is exponentially L2−equicontinuous at x if
and only if Fx is L2−equicontinuous, i.e.
∀ε>0 ∃δ>0 ∀s∈(0,δ),
f ∈F
Z ∞
0
(cid:12)(cid:12)(cid:12)e−x(t+s)f (t + s) − e−xtf (t)(cid:12)(cid:12)(cid:12)
2
dt < ε.
(12)
Proof. Since Fx is L2−bounded, there exists M > 0 such that
∀f ∈F (cid:18)Z ∞
0
e−2xtf (t)2 dt(cid:19) 1
2
< M.
8
We divide the proof of the lemma into two parts:
Part 1.
In the first part of the proof, we assume that the family F is exponentially
L2−equicontinuous at x. We fix ε > 0 and choose δ > 0 such that
• (12) is satisfied, and
• for every s ∈ (0, δ) we have
Z s
0
e−2xtf (t)2 dt < ε,
which is possible due to Theorem 8 in [14], p. 148, and
• for every s ∈ (0, δ) we have
(13)
(14)
2
0
−s
−s
−s
2
e−2x(t+s)f (t + s) − f (t)2 dt(cid:19) 1
+(cid:18)Z ∞
e−2x(t+s)exsf (t) − f (t)2 dt(cid:19) 1
−s
2
2
(cid:12)(cid:12)(cid:12)e−xs − 1(cid:12)(cid:12)(cid:12) M < ε.
=(cid:18)Z ∞
e−2xtf (t) − f (t − s)2 dt(cid:19) 1
e−2x(t+s)f (t + s) − exsf (t)2 dt(cid:19) 1
e−2x(t+s)f (t + s)2 dt +Z ∞
e−2x(t+s)f (t)2exs − 12 dt(cid:19) 1
e−2xtf (t)2 dt +Z ∞
Consequently, for every s ∈ (0, δ) and f ∈ F we have
(cid:18)Z ∞
6(cid:18)Z ∞
6 Z 0
+(cid:18)Z ∞
6 Z s
+ e−xs − 1M
Since ε > 0 was chosen arbitrarily, the above estimates end the first part of the
proof.
e−x(t+s)f (t + s) − e−xtf (t)2 dt! 1
e−2x(t+s)f (t + s) − exsf (t)2 dt! 1
0
(12),(13),(14)
2 + ε.
1
(2ε)
0
0
0
2
6
2
2
Part 2.
In this part of the proof, we assume that Fx is L2−equicontinuous. Again,
we fix ε > 0 and let δ > 0 be such that
9
• (10) is satisfied, and
• for every s ∈ (0, δ) we have
1 − exs M < ε and exs 6 2.
(15)
0
For every s ∈ (0, δ) and f ∈ F we have
(cid:18)Z ∞
dt(cid:19) 1
6(cid:18)Z ∞
dt(cid:19) 1
6 1 − exs(cid:18)Z ∞
(cid:12)(cid:12)(cid:12)e−x(t+s)f (t + s) − e−xtf (t)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)e−xtf (t) − e−x(t−s)f (t)(cid:12)(cid:12)(cid:12)
e−2xtf (t)2 dt(cid:19) 1
2
2
0
s
2
2
+ exsε
2
s
2
=(cid:18)Z ∞
+(cid:18)Z ∞
s
(15)
6 1 − exs M + 2ε < 3ε,
(cid:12)(cid:12)(cid:12)e−xtf (t) − e−x(t−s)f (t − s)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)e−x(t−s)f (t) − e−x(t−s)f (t − s)(cid:12)(cid:12)(cid:12)
2
dt(cid:19) 1
dt(cid:19) 1
2
2
which ends the proof.
We will now study the relationship between the exponential L2−equicontinuity
and the Laplace equivanishing. First, let us recall the Minkowski inequality:
Theorem 8. (Minkowski integral inequality, comp. [3], p. 194 or [17], p. 271)
Let (X, µX), (Y, µY ) be σ−finite measure spaces, 1 6 p < ∞ and let F : X × Y −→ C
be a measurable function. Then
(cid:18)ZX (cid:18)ZY F (x, y) dy(cid:19)p
p
dx(cid:19) 1
6ZY (cid:18)ZX F (x, y)p dx(cid:19) 1
p
dy.
(16)
Theorem 9. Let F be a Laplace-Pego family with a common order x > 0. Expo-
nential L2−equicontinuity at x implies Laplace equivanishing at x. Furthermore,
if Fx is L2−bounded, then the implication can be reversed.
Proof. We divide the proof into two parts:
Part 1.
We assume that the family F is exponentially L2−equicontinuous at x, so for
a fixed ε > 0 we can choose δ > 0 according to the exponential L2−equicontinuity
(10). Let g be a compactly supported, continuous function on R+, satisfying
the following conditions:
• g is nonnegative, and
10
• supp(g) ⊂ (0, δ), and
• R ∞
g(s) ds = 1.
0
Naturally, g is a Laplace-Pego function of order x.
By Theorem 2, let T > 0 be such that
∀y∈[−T,T ] L{g}(x + iy) 6
1
2
.
(17)
Consequently, we have
∀f ∈F ZR\[−T,T ] L{f}(x + iy)2 dy! 1
2
(17)
6 ZR\[−T,T ] (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)L{f}(x + iy)(cid:16)1 − L{g}(x + iy)(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
6 ZR\[−T,T ] (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)L{f}(x + iy)(cid:16)1 − L{g}(x + iy)(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+ ZR\[−T,T ] L{f}(x + iy)L{g}(x + iy)2 dy! 1
2 ZR\[−T,T ] L{f}(x + iy)2 dy! 1
dy! 1
+
1
2
2
2
2
dy! 1
2
,
11
which implies
∀f ∈F ZR\[−T,T ] L{f}(x + iy)2 dy! 1
2
2
2
2
0
2(cid:18)ZR L{f}(x + iy) − L{f ⋆ g}(x + iy)2 dy(cid:19) 1
2√2π(cid:18)Z ∞
6 2ZR\[−T,T ] (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)L{f}(x + iy)(cid:16)1 − L{g}(x + iy)(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
e−2xtf (t) − f ⋆ g(t)2 dt(cid:19) 1
f (t − s)g(s) ds(cid:12)(cid:12)(cid:12)(cid:12)
e−2xt(cid:12)(cid:12)(cid:12)(cid:12)f (t) −Z ∞
(cid:12)(cid:12)(cid:12)(cid:12)Z ∞
e−xt(f (t) − f (t − s))g(s) ds(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18)Z ∞
e−2xtf (t) − f (t − s)2g(s)2 dt(cid:19) 1
2√2πZ ∞
e−2xtf (t) − f (t − s)2 dt(cid:19) 1
g(s)(cid:18)Z ∞
e−2xtf (t) − f (t − s)2 dt(cid:19) 1
g(s)(cid:18)Z ∞
g(s) ds = 2√2πε.
dt! 1
dt! 1
ds
ds
ds
0
2
2
2
0
0
2
2
0
0
0
Theorem 5
6
Theorem 1
=
= 2√2π Z ∞
= 2√2π Z ∞
0
0
Minkowski ineq.
6
0
= 2√2πZ ∞
= 2√2πZ δ
6 2√2πεZ δ
0
0
2
1
2
dy
2
Let us remark that the use of Minkowski inequality in hte above estimates is
justified, because the function F (t, s) = e−2xtf (t) − f (t − s)2g(s)2 is measur-
able due to Proposition 3.9, p. 86 in [18]. Since ε > 0 was chosen arbitrarily,
the above estimates end the first part of the proof.
Part 2.
For this part of the proof, we assume that Fx is L2−bounded, so there exists
M1 > 0 such that
∀f ∈F Z ∞
0
e−2xtf (t)2 dt < M1.
(18)
We will show that if F is Laplace equivanishing at x then it is exponentially
L2−equicontinuous at x. For convenience, we denote T−sf (t) = f (t − s). We
12
observe the following equalities
f ∈F Z ∞
∀ s>0
0
e−2xtf (t) − f (t − s)2 dt =Z ∞
0
Theorem 1
=
1
−∞ L{f − T−sf}(x + iy)2 dy
e−2xtf (t) − T−sf (t)2 dt
2π Z ∞
−∞ L{f}(x + iy) − e−s(x+iy)L{f}(x + iy)2 dy
−∞ (cid:12)(cid:12)(cid:12)1 − e−s(x+iy)(cid:12)(cid:12)(cid:12)
L{f}(x + iy)2 dy.
(19)
2
1
2π Z ∞
2π Z ∞
1
=
=
Fix ε > 0 and choose T > 0 according to Laplace equivanishing (11). Let
δ > 0 be such that
y∈R
2
.
2
< ε,
and put
∀ s∈(0,δ)
M2 = max
s∈[0,δ],
y∈[−T,T ] (cid:12)(cid:12)(cid:12)1 − e−s(x+iy)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)1 − e−s(x+iy)(cid:12)(cid:12)(cid:12)
Finally, for every s ∈ (0, δ) and f ∈ F , we have
2π Z T
−T (cid:12)(cid:12)(cid:12)1 − e−s(x+iy)(cid:12)(cid:12)(cid:12)
2π ZR\[−T,T ] (cid:12)(cid:12)(cid:12)1 − e−s(x+iy)(cid:12)(cid:12)(cid:12)
2π ZR\[−T,T ] L{f}(x + iy)2 dy
2π(cid:19) ε.
ε 6(cid:18)M1 +
−T L{f}(x + iy)2 dy+
ε Z ∞
e−2xtf (t) − f (t − s)2 dt
e−2xtf (t)2 dt+
2π Z T
Theorem 1, (11)
Z ∞
(20), (21)
M2
2π
(19)
=
1
1
M2
M2
ε
6
+
0
6
0
(20)
(21)
2
L{f}(x + iy)2 dy
2
L{f}(x + iy)2 dy
Since ε > 0 was chosen arbitrarily, we conclude the proof.
Before we present the final theorem of the paper, let us recall the celebrated
Riesz-Kolmogorov theorem:
Theorem 10. (Riesz-Kolmogorov theorem, comp. [10])
A family A ⊂ L2(R+) is relatively compact if and only if
13
• A is L2−bounded, and
• A is L2−equicontinuous, and
• A is L2−equivanishing.
The final theorem, which is the climax of the paper, should be juxtaposed
with Pego theorem in [4], [5] and [15].
Theorem 11. Let F be the family consisting of Laplace-Pego functions with a
common order x and such that Fx is L2−bounded. The family Fx is relatively
compact in L2(R+) if and only if
• F is Laplace equicontinuous at x, and
• F is Laplace equivanishing at x.
Proof. The proof is divided into two parts:
Part 1.
We assume that F is Laplace equicontinuous and equivanishing at x. At
first, we note that Laplace equicontinuity of F at x implies that this family
is exponentially L2−equivanishing at x (Theorem 6).
In other words, Fx is
L2−equivanishing.
Furthermore, Laplace equivanishing of F at x implies that this family is ex-
ponentially L2−equicontinuous (Theorem 9). In other words, Fx is L2−equicontinuous
(Lemma 7). By Theorem 10, we conclude that Fx is relatively compact in
L2(R+).
Part 2.
For the second part of the proof, we assume that Fx is relatively com-
pact in L2(R+). By Theorem 10, the family Fx is L2−equicontinuous and
L2−equivanishing.
L2−equicontinuity of Fx implies that F is Laplace equivanishing at x (Lemma
7 and Theorem 9). Moreover, L2−equivanishing of Fx implies that F is Laplace
equicontinuous at x (Theorem 6), which ends the proof.
14
References
[1] Deitmar A. : A First Course in Harmonic Analysis, Springer-Verlag, New
York, 2005
[2] Deitmar A., Echterhoff S. : Principles of Harmonic Analysis, Springer, New
York, 2009
[3] Folland G. B. : Real Analysis: Modern techniques and their applications,
John Wiley and Sons, New York, 1999
[4] Górka, P. : Pego theorem on locally compact abelian groups, Journal of
Algebra and Its Applications, Vol. 14, No. 4 (2014)
[5] Górka P., Kostrzewa T. : Pego everywhere, Journal of Algebra and Its Ap-
plications, Vol. 15, No. 4 (2016)
[6] Górka P., Macios A. : The Riesz-Kolmogorov theorem on metric spaces,
Miskolc Mathematical Notes, Vol. 15 (2014)
[7] Górka P., Macios A. : Almost everything you need to know about relatively
compact sets in variable Lebesgue spaces, Journal of Functional Analysis,
Vol. 269, No. 7 (2015)
[8] Górka P., Rafeiro H. : From Arzelà-Ascoli to Riesz-Kolmogorov, Nonlinear
Analysis, Vol. 144 (2016)
[9] Górka P., Rafeiro H. : Light side of compactness in Lebesgue spaces: Sudakov
theorem, Annales Academiae Scientiarum Fennicae Mathematica, Vol. 42
(2017)
[10] Hanche-Olsen H., Holden H. : The Kolmogorov-Riesz compactness theorem,
Expositiones Mathematicae, Vol. 28, No. 4 (2010)
[11] Hanche-Olsen H., Holden H., Malinnikova E. : An improvement of the
Kolmogorov-Riesz compactness theorem, Expositiones Mathematicae (2018)
[12] Katznelson Y. : An introduction to harmonic analysis, Cambridge Univer-
sity Press, Cambridge, 2004
15
[13] Kolmogorov A. N. : Über Kompaktheit der Funktionenmengen bei der Kon-
vergenz im Mittel, Nachrichten von der Gesellschaft der Wissenschaften zu
Göttingen, Vol. 9, p. 60-63 (1931)
[14] Natanson I. P. : Theory of functions of a real variable, Dover Publications,
New York, 2016
[15] Pego R. L. : Compactness in L2 and the Fourier transform, Proceedings of
the American Mathematical Society, Vol. 95, No. 2 (1985)
[16] Schiff J. L. : The Laplace Transform. Theory and applications, Springer-
Verlag, New York, 1999
[17] Stein E. : Singular integrals and differentiability properties of functions,
Princeton University Press, Princeton, 1970
[18] Stein E., Shakarchi R. : Real analysis: measure theory, integration and
Hilbert spaces, Princeton University Press, Princeton 2005
[19] Tamarkin J. D. : On the compactness of the space Lp, Bulletin of the
American Mathematical Society, Vol. 32, p. 79-84 (1932)
[20] Weil A. : L'intégration dans les groupes topologiques et ses applications,
Hermann, Paris, 1965
16
|
1407.4632 | 2 | 1407 | 2015-01-08T14:25:25 | Weak factorization and Hankel forms for weighted Bergman spaces on the unit ball | [
"math.FA",
"math.CV"
] | We establish weak factorizations for a weighted Bergman space $A^p_{\a}$, with $1<p<\infty$, into two weighted Bergman spaces on the unit ball of $\C^n$. To obtain this result, we characterize bounded Hankel forms on weighted Bergman spaces on the unit ball of $\C^n$. | math.FA | math |
WEAK FACTORIZATION AND HANKEL FORMS FOR WEIGHTED
BERGMAN SPACES ON THE UNIT BALL
JORDI PAU AND RUHAN ZHAO
ABSTRACT. We establish weak factorizations for a weighted Bergman space Ap
α, with
1 < p < ∞, into two weighted Bergman spaces on the unit ball of Cn. To obtain this
result, we characterize bounded Hankel forms on weighted Bergman spaces on the unit
ball of Cn.
1. INTRODUCTION
A classical theorem of Riesz asserts that any function in the Hardy space H p on the
unit disk can be factored as f = Bg with kf kHp = kgkHp, where B is a Blaschke
product and g is an H p-function with no zeros on the unit disk. An immediate consequence
of that result is that any function in the Hardy space H p admits a "strong" factorization
f = f1f2 with f1 ∈ H p1, f2 ∈ H p2 and kf1kHp1 · kf2kHp2 = kf kHp, for any p1 and
p2 determined by the condition 1/p = 1/p1 + 1/p2. In [12], C. Horowitz obtained strong
factorizations of functions in a weighted Bergman space on the unit disk into functions
of two weighted Bergman spaces with the same weight (again using Blaschke products).
These strong factorization results are no longer possible to obtain [11] in the setting of
Hardy and Bergman spaces in the unit ball of the complex euclidian space Cn of dimension
n when n ≥ 2, but it is still possible to obtain some "weak" factorizations for functions in
these spaces.
For two Banach spaces of functions, A and B, defined on the same domain, the weakly
factored space A ⊙ B is defined as the completion of finite sums
kf kA⊙B = inf(Xk
kϕkkAkψkkB : f =Xk
ϕkψk) .
When 0 < p ≤ 1, weak type factorizations for the Hardy spaces H p and the weighted
Bergman spaces Ap
α on the unit ball of Cn are well known (see [6] and [9] for Hardy spaces;
and [5], [20] or [24, Corollary 2.33] for Bergman spaces). However, when 1 < p < ∞,
even for unweighted Bergman spaces the problem was still open (see, for example [4]).
In this paper we completely solve the above problem for Bergman spaces by establishing
β, with 1 < q < ∞ and β > −1, into
weak factorizations for a weighted Bergman space Aq
2010 Mathematics Subject Classification. 32A36, 47B35, 47B38.
Key words and phrases. Weak factorization, Hankel operators, Bergman spaces.
This work started when the second named author visited the University of Barcelona in 2013. He thanks
the support given by the IMUB during his visit. The first author was supported by DGICYT grant MTM2011-
27932-C 02-01 (MCyT/MEC) and the grant 2014SGR289 (Generalitat de Catalunya).
1
f =Xk
with the following norm:
ϕkψk,
{ϕk} ⊂ A, {ψk} ⊂ B,
2
J. PAU AND R. ZHAO
two weighted Bergman spaces with non necessarily the same weight, on the unit ball Bn
of Cn. The following is our main result.
Theorem 1. Let 1 < q < ∞ and β > −1. Then
Aq
β(Bn) = Ap1
α1 (Bn) ⊙ Ap2
α2 (Bn)
for any p1, p2 > 0 and α1, α2 > −1 satisfying
(1)
1
p1
+
1
p2
=
1
q
,
α1
p1
+
α2
p2
=
β
q
.
In this context, by "=" we mean equality of the function spaces and equivalence of the
is a
norms. The inclusion Ap1
direct consequence of Minkowski and Holder inequalities, so that the interesting part is the
other inclusion with the corresponding estimates for the norms.
β with the estimate kf kq,β . kf kA
α2 ⊂ Aq
α1 ⊙ Ap2
p1
α1 ⊙A
p2
α2
Now we are going to recall the definition of the weighted Bergman spaces. First we
need some notations. For any two points z = (z1, ..., zn) and w = (w1, ..., wn) in Cn, we
use
to denote the inner product of z and w, and
hz, wi = z1 ¯w1 + · · · + zn ¯wn
z =phz, zi =pz12 + · · · + zn2
to denote the norm of z in Cn. Let Bn = {z ∈ Cn : z < 1} be the unit ball in Cn
and Sn = {z ∈ Cn : z = 1} be the unit sphere in Cn. Let H(Bn) be the space of
all analytic functions on Bn. We use dv to denote the normalized volume measure on
Bn and dσ to denote the normalized area measure on Sn. For −1 < α < ∞, we let
dvα(z) = cα(1 − z2)α dv(z) denote the normalized weighted volume measure on Bn,
where cα = Γ(n + α + 1)/[n!Γ(α + 1)].
For 0 < p < ∞ and −1 < α < ∞, let Lp(Bn, dvα) be the weighted Lebesgue space
which contains measurable functions f on Bn such that
kf kp
p,α =ZBn
f (z)p dvα(z) < ∞.
Denote by Ap
α = Lp(Bn, dvα) ∩ H(Bn), the weighted Bergman space on Bn, with the
same norm as above. If α = 0, we simply write them as Lp(Bn, dv) and Ap respectively
and kf kp for the norm of f in these spaces.
It is a well-known fact that to obtain weak factorization results is equivalent to give
a "good" description of the boundedness of certain Hankel forms. A Hankel form is a
bilinear form B on a space of analytic functions such that for any f and g, B(f, g) is a
linear function of f g. These forms have been extensively studied on Hardy spaces and on
Bergman spaces. For the case of the Hardy space on the unit disk, a classical result by
Nehari [18] says that the Hankel form
Bb(f, g) := hf g, bi
(under the usual integral pair for Hardy spaces) with an analytic symbol b is bounded on
H 2 × H 2 if and only if b ∈ BM OA, the space of analytic functions of bounded mean
oscillation. The proof used the fact that a function in H 1 can be factored into product of
two functions in H 2. Unfortunately, such strong factorization is not possible (see [11]) for
Hardy spaces in the unit ball Bn of Cn. However, Coifman, Rochberg and Weiss [6] were
WEAK FACTORIZATION AND HANKEL OPERATORS
3
able to generalize Nehari's result to the unit ball Bn by using a weak factorization of H 1.
Namely, they proved that
H 2(Bn) ⊙ H 2(Bn) = H 1(Bn).
Our approach to the problem for weighted Bergman spaces on the unit ball is the oppo-
site to the one of Coifman, Rochberg and Weiss in [6]. We first characterize boundedness
of the Hankel forms on weighted Bergman spaces, and with this result the weak factoriza-
tion easily follows.
Given α > −1 and a holomorphic symbol function b we define the associated Hankel
type bilinear form T α
b for polynomials f and g by
T α
b (f, g) = hf g, biα,
where the integral pair h , iα is defined as
(2)
hϕ, ψiα =ZBn
ϕ(z) ψ(z) dvα(z).
Since the polynomials are dense in the weighted Bergman spaces, the Hankel form T α
densely defined on Ap1
is bounded on Ap1
b is
α2 for any p1, p2 > 0 and any α1, α2 > −1. We say that T α
b
α2 if there exists a positive constant C such that
α1 × Ap2
α1 × Ap2
The norm of T α
b is given by
T α
b (f, g) ≤ Ckf kp1,α1kgkp2,α2 .
kT α
b k = kT α
b kAp1
α1 ×Ap2
α2
:= sup{T α
The next result characterizes boundedness of the Hankel form T α
We will see in Section 3 that this implies the weak factorization in Theorem 1.
b (f, g) : kf kp1,α1 = kgkp2,α2 = 1 }.
b acting on Ap1
α2.
α1 × Ap2
Theorem 2. Let 1 < p1, p2 < ∞, and α, α1, α2 > −1 satisfy
(3)
1
p1
+
1
p2
< 1,
1 + α1
1 + α2
+
< 1 + α.
p1
p2
α2 if and only b ∈ Aq′
b is bounded on Ap1
Then T α
satisfying (1), and q′ and β′ are determined by the condition
α1 × Ap2
β ′, where q and β are real numbers
(4)
1
q
+
1
q′ = 1,
β
q
+
β′
q′ = α.
Furthermore, we have kT α
b k ≍ kbkq′,β ′
Remarks. Note that, condition (3) guarantees that q > 1 and β′ > −1. When q and β
satisfy condition (1), automatically we would have β > −1 (to see this, simply add two
equations in (1) together). By a general duality theorem for weighted Bergman spaces (see
Theorem A in Section 2), the condition b ∈ Aq′
β ′ means that the symbol b belongs to the
dual space of Aq
β under the pairing given by (2).
It turns out that boundedness of the Hankel form T α
b is equivalent to boundedness of a
(small) Hankel operator, which we are going to introduce in a moment. Let α > −1. It is
well-known that, the integral operator
Pαf (z) =ZBn
f (w)
(1 − hz, wi)n+1+α d vα(w)
4
J. PAU AND R. ZHAO
is the orthogonal projection from L2(Bn, dvα) onto the weighted Bergman space A2
α. The
above formula can be used to extend Pα to a linear operator from L1(Bn, dvα) into H(Bn).
For 1 < p < ∞, Pα is a bounded operator from Lp(Bn, dvα) onto Ap
α.
Denote by Ap
α the conjugate analytic functions f on Bn that are in Lp(Bn, dvα). Clearly,
Ap
α = {f : f ∈ Ap
α}.
Let Qα denote the orthogonal projection from L2(Bn, dvα) onto A2
α. Clearly one has
Qαf (z) = Pαf (z) =ZBn
f (w)
(1 − hw, zi)n+1+α dvα(w).
Given f ∈ L1(Bn, dvα) and a polynomial g, the weighted (small) Hankel operator is
defined by
hα
f g = Qα(f g).
Due to the density of polynomials, the small Hankel operator hα
weighted Bergman space Ap
Hankel operator with conjugate analytic symbols, that is, hα
f
to Ap2
f is densely defined on the
α for 1 ≤ p < ∞. We will study boundedness of the small
with f ∈ H(Bn), from Ap1
α1
α2 with 0 < p2 < p1 < ∞.
Theorem 3. Let 1 < p2 < p1 < ∞ and α1, α2 > −1 such that
(5)
Let f ∈ H(Bn) and α such that
(6)
1 + α1
p1
<
1 + α2
p2
.
1 + α >
1 + α2
p2
.
¯f : Ap1
α1 → Ap2
α2 is bounded if and only if f ∈ Aq
β, where q and β are real numbers
Then hα
such that
Moreover, we have khα
=
1
q
1
p2
f k ≍ kf kq,β.
−
1
p1
,
β
q
=
α2
p2
−
α1
p1
.
α1 ⊂ Ap2
Remarks. Condition (5) guarantees that β > −1. It is known that, when 0 < p2 <
p1 < ∞, Ap1
α2 if and only if (5) is true (see [22, Theorem 70]). Hence the above
result concerns the boundedness of hα
f from a smaller space to a larger space. Also, by [24,
Theorem 2.11], condition (6) means that the integral operator Pα is a bounded projection
from Lp2(Bn, dvα2 ) onto Ap2
α2.
If one considers the operator
Sα
f g = h ¯f g(z) = Pα(f g),
clearly, the boundedness of hα
f
and the norms of hα
theorem we easily obtain
f and Sα
is equivalent to the boundedness of Sα
f are equivalent. Now, if g ∈ Ap1
α1 and h ∈ Ap2
f from Ap1
α1 to Ap2
α2,
α2, by Fubini's
f (g, h) = hgh, f iα = hh, Pα(f ¯g)iα = hh, Sα
T α
f giα.
WEAK FACTORIZATION AND HANKEL OPERATORS
5
Hence, for p2 > 1, by duality (see Theorem A in Section 2), the Hankel form T α
bounded on Ap1
to Ap′
α′
α2 if and only if the small Hankel operator hα
f
, with equivalent norms. Here, the numbers α′
f is
is bounded from Ap1
α1
α1 × Ap2
2
2
1
p2
+
1
p′
2
= 1,
α =
2 are defined by the relation
2 and p′
α′
2
p′
2
α2
p2
+
.
Comparing Theorem 2 with Theorem 3, notice that the first inequality in (3) is equivalent
to condition 1 < p′
2, con-
dition (6) turns out to be equivalent to α2 > −1, and therefore is always satisfied; and
the second inequality in (3) is equivalent to condition (5). Therefore, Theorem 3 implies
Theorem 2.
2 < p1 < ∞. Also, when p2 and α2 are replaced by p′
2 and α′
In the case of the same weights, that is, when α1 = α2 = β = α, all the restrictions in
Theorem 3 reduces to p2 > 1. We isolate this case here, since it proves a conjecture in [4].
Corollary 4. Let α > −1, 1 < p2 < p1 < ∞ and f ∈ H(Bn). Then hα
f
bounded if and only if f ∈ Aq
α, with q = p1p2
p1−p2
.
: Ap1
α → Ap2
α is
The paper is organized as follows: in Section 2 we give some necessary concepts and
recall some key results which are needed in our proof of the main result. In Section 3 we
give in detail the connection between weak factorizations and Hankel forms. The proof of
Theorem 3 is given in Section 4.
In the following, the notation A . B means that there is a positive constant C such that
A ≤ CB, and the notation A ≍ B means that both A . B and B . A hold.
2. PRELIMINARIES
We need the following duality theorem. In this generality the result is due to Luecking
[16] (see also, Theorem 2.12 in [24]).
Theorem A. Suppose β, β′ > −1 and 1 < q < ∞. Then
(Aq
β)∗ = Aq′
β ′
(with equivalent norms) under the integral pair h , iα given by (2), where
1
q
+
1
q′ = 1,
α =
β
q
+
β′
q′ .
We need the following well known integral estimate that can be found, for example, in
[24, Theorem 1.12].
Lemma B. Let t > −1 and s > 0. There is a positive constant C such that
ZBn
(1 − w2)t dv(w)
1 − hz, win+1+t+s ≤ C (1 − z2)−s
for all z ∈ Bn.
For any a ∈ Bn with a 6= 0, we denote by ϕa(z) the Mobius transformation on Bn that
exchanges 0 and a. It is known that, for any z ∈ Bn
ϕa(z) =
a − Pa(z) − saQa(z)
1 − hz, ai
,
6
J. PAU AND R. ZHAO
where sa = 1 − a2 , Pa is the orthogonal projection from Cn onto the one dimensional
subspace [a] generated by a, and Qa is the orthogonal projection from Cn onto the orthog-
onal complement of [a]. When a = 0, ϕa(z) = −z. ϕa has the following properties:
ϕa ◦ ϕa(z) = z, and
1 − ϕa(z)2 =
(1 − a2)(1 − z2)
1 − hz, ai2
.
For z, w ∈ Bn, the pseudo-hyperbolic distance between z and w is defined by
ρ(z, w) = ϕz(w),
and the hyperbolic distance on Bn between z and w induced by the Bergman metric is
given by
For z ∈ Bn and r > 0, the Bergman metric ball at z is given by
β(z, w) = tanh ρ(z, w).
It is known that, for a fixed r > 0, the weighted volume
D(z, r) = {w ∈ Bn : β(z, w) < r}.
vα(D(z, r)) ≍ (1 − z2)n+1+α.
We refer to [24] for all of the above facts.
A sequence {ak} of points in Bn is a separated sequence (in Bergman metric) if there
exists a positive constant δ > 0 such that β(zi, zj) > δ for any i 6= j. We need a well-
known result on decomposition of the unit ball Bn. The following version is Theorem 2.23
in [24]
Lemma C. There exists a positive integer N such that for any 0 < r < 1 we can find a
sequence {ak} in Bn with the following properties:
(i) Bn = ∪kD(ak, r).
(ii) The sets D(ak, r/4) are mutually disjoint.
(iii) Each point z ∈ Bn belongs to at most N of the sets D(ak, 4r).
Any sequence {ak} satisfying the conditions of the above lemma is called a lattice (or
an r-lattice if one wants to stress the dependence on r) in the Bergman metric. Obviously
any r-lattice is separated.
For convenience, we will denote by Dk = D(ak, r) and Dk = D(ak, 4r). Then
k=1Dk and there is an positive integer N such that every
Lemma C says that Bn = ∪∞
point z in Bn belongs to at most N of sets Dk.
We also need the following atomic decomposition theorem for weighted Bergman spaces.
This turns out to be a powerful theorem in the theory of Bergman spaces. The result is ba-
sically due to Coifman and Rochberg [5], and can be found in Chapter 2 of [24].
Theorem D. Suppose p > 0, α > −1, and
b > n max(cid:18)1,
1
p(cid:19) +
1 + α
p
.
Then we have
(i) For any separated sequence {ak} in Bn and any sequence λ = {λk} ∈ ℓp, the
function
f (z) =
λk
∞Xk=1
(1 − ak2)b−(n+1+α)/p
(1 − hz, aki)b
WEAK FACTORIZATION AND HANKEL OPERATORS
7
belongs to Ap
α and
kf kp,α . kλkℓp.
(ii) There is an r-lattice {ak} in Bn such that, for any f ∈ Ap
α, there is a sequence
λ = {λk} ∈ ℓp with
and
f (z) =
λk
∞Xk=1
(1 − ak2)b−(n+1+α)/p
(1 − hz, aki)b
.
kλkℓp . kf kp,α.
In the proof given in [24], part (i) requires that the sequence {ak} is an r-lattice for some
r ∈ (0, 1], but it is well known that only the separation of the sequence {ak} is needed.
3. WEAK FACTORIZATIONS AND HANKEL FORMS
It is well known to specialists, but difficult to find in the literature, that the obtention
of weak factorizations is equivalent to estimates for small Hankel operators or Hankel
forms (in our case, estimates with loss). The equivalence between boundedness of the
b and weak factorization can be formulated as the following result.
bilinear Hankel form T α
Since this is the basis for our obtention of the weak factorization for Bergman spaces, for
completeness, we offer the proof here of the implication that gives the factorization.
Proposition 5. Let 1 < q < ∞ and α, β > −1. Let p1, p2 and α1, α2 satisfy (3) and (1),
and let q′ and β′ satisfy (4). The following are equivalent:
β ⊂ Ap1
(i) Aq
(ii) For any analytic function b, if T α
α2 with kf kAp1
α1 ⊙ Ap2
α1 ⊙Ap2
α2
. kf kq,β for f ∈ Aq
β.
α2, then b ∈ Aq′
α1 × Ap2
b is bounded on Ap1
β ′ with
kbkq′,β ′ . kT α
b k.
Proof. We will prove (ii) implies (i). The other implication is easier, and the interested
reader can follow the argument in [1, Corollary 1.2] for a proof. By the atomic decom-
position in Theorem D, we have the inclusion Aq
α2 . In order to have the
corresponding estimate for the norms, we will show that, for any bounded linear functional
F on Ap1
β ′ with kbF kq′,β ′ . kF k such that
F (f ) = hf, bF iα for f ∈ Aq
α2, there is a unique function bF ∈ Aq′
β ⊂ Ap1
α1 ⊙ Ap2
α1 ⊙ Ap2
β. This would give
kf kA
p1
α1 ⊙A
p2
α2
F (f ) ≤ sup
kF k=1
kbF kq′,β ′ · kf kq,β . kf kq,β.
Thus, suppose F ∈ (Ap1
α2 )∗ with norm kF k. Then for all ϕ ∈ Ap2
α2 we have
= sup
kF k=1
α1 ⊙ Ap2
F (ϕ) = F (1 · ϕ) ≤ kF k · k1kp1,α1 · kϕkp2,α2 = kF k · kϕkp2,α2 .
Hence F ∈ (Ap2
such that F (ϕ) = hϕ, biα for all ϕ ∈ Ap2
α2 )∗, and so, by Theorem A, there is an unique function b = bF ∈ Ap′
2
α′
2
2 and α′
2 satisfy
α2, where p′
α2
p2
+
= 1,
α′
2
p′
2
= α.
1
p2
+
1
p′
2
Now, for polynomials g and h we have
T α
b (g, h) = hgh, biα = F (gh)
≤ kF k · kghkA
≤ kF k · kgkp1,α1 · khkp2,α2 ,
p1
α1 ⊙A
p2
α2
8
J. PAU AND R. ZHAO
b is bounded on Ap1
which shows that T α
that b ∈ Aq′
linear form on Aq
F (f ) = Λ(f ) = hf, biα. The proof is complete.
b k ≤ kF k. Therefore, we know
b k . kF k. Hence Λ(f ) = hf, biα defines a bounded
β, we have
(cid:3)
β, and coincides with F on polynomials. Thus, for f ∈ Aq
β ′ with kbkq′,β ′ . kT α
α2 with kT α
α1 × Ap2
4. PROOF OF THEOREM 3
In this section we prove Theorem 3, from which Theorem 2 follows. Before that, for
s ≥ 0 and α > −1, let Rα,s denote the unique continuous linear operator on H(Bn)
satisfying
1
(1 − hz, wi)n+1+α(cid:19) =
Rα,s(cid:18)
Rα,sf (z) =ZBn
f (w)
1
(1 − hz, wi)n+1+α+s
for all w ∈ Bn. If f ∈ A1
α, then Rα,sf is given by the following integral expression
(7)
(1 − hz, wi)n+1+α+s dvα(w).
More properties of the "differential type" operators Rα,s can be found in [24, Section 1.4].
Now we are ready to prove Theorem 3.
Proof of Theorem 3. As we noticed before, we just need to prove that Sα
is bounded if and only if f ∈ Aq
β.
f : Ap1
α1 → Ap2
α2
Suppose first that f ∈ Aq
β. We need to show Sα
α1. If p2 > 1 then Pα : Lp2 (Bn, dvα2 ) → Ap2
g ∈ Ap1
inequality the result follows. Indeed,
f : Ap1
α1 → Ap2
α2 is bounded. Let
α2 is bounded, and then from Holder's
kSα
f gkp2,α2 = kPα(f ¯g)kp2,α2 ≤ Ckf gkp2,α2 ≤ Ckf kq,β · kgkp1,α1
f : Ap1
α1 → Ap2
which shows that Sα
α2 is bounded with
f k . kf kq,β.
α2 is bounded, we are going to show that f ∈ Aq
β.
We begin with using an argument of Luecking (see, e.g., [17]). Let rk(t) be a sequence of
Rademacher functions (see [8, Appendix A]). Let b be large enough so that
kSα
α1 → Ap2
Conversely, suppose Sα
f : Ap1
(8)
b > n +
1 + α1
p1
.
Fix any r > 0, and let {ak} be an r-lattice and {Dk} be the associated sets in Lemma C.
By Theorem D, we know that, for any sequence of real numbers λ = {λk} ∈ ℓp1, the
function
gt(z) =
λkrk(t)
(1 − ak2)b−(n+1+α1)/p1
(1 − hz, aki)b
∞Xk=1
belongs to Ap1
α1 with kgtkp1,α1 . kλkℓp1 for almost every t in (0, 1). Denote by
gk(z) =
(1 − ak2)b−(n+1+α1)/p1
(1 − hz, aki)b
.
Since Sα
f : Ap1
α1 → Ap2
α2 is bounded, we get that
kSα
f gtkp2
λkrk(t)Sα
dvα2 (z)
p2,α2 = ZBn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
. kSα
∞Xk=1
p2
f gk(z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f kp2 · kgtkp2
p1,α1 . kSα
f kp2 · kλkp2
ℓp1
WEAK FACTORIZATION AND HANKEL OPERATORS
9
for almost every t in (0, 1). Integrating both sides with respect to t from 0 to 1, and using
Fubini's Theorem and Khinchine's inequality (see [21, p.12]), we get
dvα2 (z) . kSα
f kp2 · kλkp2
ℓp1 .
λkp2 Sα
f gk(z)p2χ Dk
· p2
2
p2
(z)! 2
dvα2 (z).
(9)
Now we estimate
(10)
λkp2ZDk
∞Xk=1
λk2Sα
f gk(z)2!p2/2
ZBn ∞Xk=1
f gk(z)p2 dvα2 (z) =ZBn ∞Xk=1
∞Xk=1
Sα
Sα
If p2 ≥ 2, then 2/p2 ≤ 1, and from (10) we have
λkp2Z Dk
≤ZBn ∞Xk=1
≤ZBn ∞Xk=1
f gk(z)p2 dvα2 (z)
(z)!p2/2
dvα2 (z)
λk2Sα
f gk(z)2χ Dk
λk2Sα
f gk(z)2!p2/2
dvα2 (z).
If 1 < p2 < 2, then 2/p2 > 1, from (10), by Holder's inequality we get
f gk(z)p2 dvα2 (z)
∞Xk=1
Sα
λkp2Z Dk
≤ZBn ∞Xk=1
≤ N 1−p2/2ZBn ∞Xk=1
λk2Sα
(z)!1−p2/2
dvα2 (z)
f gk(z)2!p2/2 ∞Xk=1
f gk(z)2!p2/2
λk2Sα
χ Dk
dvα2 (z),
since each z ∈ Bn belongs to at most N of the sets Dk. Combining the above two inequal-
ities, and applying (9) we have
∞Xk=1
Sα
f gk(z)p2 dvα2 (z)
λkp2Z Dk
≤ min{1, N 1−p2/2}ZBn ∞Xk=1
. kSα
f kp2 · kλkp2
ℓp1 .
λk2Sα
f gk(z)2!p2/2
dvα2 (z)
By subharmonicity we know that,
Sα
f gk(ak)p2 .
From this we obtain
1
(1 − ak2)n+1+α2 Z Dk
Sα
f gk(z)p2 dvα2 (z).
(11)
∞Xk=1
λkp2 (1 − ak2)n+1+α2(cid:12)(cid:12)Sα
f gk(ak)(cid:12)(cid:12)p2 . kSα
f kp2 · kλkp2
ℓp1 .
10
J. PAU AND R. ZHAO
Let Rα,b be the integral operator defined in (7). Then
Sα
f (w)gk(w)
f gk(ak) = ZBn
= ZBn
= (1 − ak2)b−(n+1+α1)/p1ZBn
(1 − hak, wi)n+1+α dvα(w)
f (w)(1 − ak2)b−(n+1+α1)/p1
(1 − hak, wi)n+1+α(1 − hak, wi)b dvα(w)
= (1 − ak2)b−(n+1+α1)/p1 Rα,bf (ak).
f (w)
(1 − hak, wi)n+1+α+b dvα(w)
Thus (11) becomes
λkp2 (1 − ak2)(n+1+α2)+[b−(n+1+α1)/p1]p2 Rα,bf (ak)p2 . kSα
f kp2 ·kλkp2
ℓp1 .
the equation (12) is the same as
n + 1 + α1
(cid:19) p2 =(cid:18)b +
(n + 1 + α2) +(cid:18)b −
λkp2h(1 − ak2)b+(n+1+β)/qRα,bf (ak)ip2
∞Xk=1
p1
n + 1 + β
q
(cid:19) p2,
. kSα
f kp2 · kλkp2
ℓp1 .
Since {λk} was an arbitrary sequence in ℓp1, we know that {λp2
k } is an arbitrary sequence
in ℓp1/p2. Since the conjugate exponent of p1/p2 is (p1/p2)′ = p1/(p1 − p2), by duality
we obtain that
n(1 − ak2)b+(n+1+β)/qRα,bf (ak)o ∈ ℓp1p2/(p1−p2) = ℓq,
∞Xk=1
(12)
Since
(13)
and
(14)
(1 − ak2)bq+(n+1+β)Rα,bf (ak)q . kSα
f kq.
∞Xk=1
This is the discrete version of what we want. Now, we will deduce that f ∈ Aq
β with
f k using duality and the atomic decomposition for Bergman spaces. Indeed,
kf kq,β . kSα
choose β′ = q′(α − β/q) (which means α = β/q + β′/q′). Note that condition (6) implies
that 1 + α > (1 + β)/q, and this guarantees that β′ > −1. Hence, by the duality result in
Theorem A,
(15)
Observe that
kf kq,β ≍ sup
hh, f iα.
khkq′,β′ =1
n + 1 + α + b > n +
1 + β′
q′
.
Then, we can apply (14) with the r-lattice {ak} for which the atomic decomposition in
Theorem D for Aq′
β ′, there exists a sequence µ = {µk} ∈ ℓq′
with kµkℓq′ . khkq′,β ′ such that
β ′ holds. That is, for any h ∈ Aq′
h(z) =
µk
∞Xk=1
(1 − ak2)n+1+α+b−(n+1+β ′)/q′
(1 − hz, aki)n+1+α+b
.
WEAK FACTORIZATION AND HANKEL OPERATORS
11
Since
n + 1 + α + b − (n + 1 + β′)/q′ = b + (n + 1 + β)/q,
then (15), Holder's inequality and (14) gives
kf kq,β ≍
≤
sup
∞Xk=1
khkq′,β′ =1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
khkq′,β′ =1(cid:13)(cid:13)µ(cid:13)(cid:13)ℓq′" ∞Xk=1
sup
µk(1 − ak2)b+(n+1+β)/q Rα,bf (ak)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(1 − ak2)bq+(n+1+β)Rα,bf (ak)q#1/q
. kSα
f k.
Hence f ∈ Aq
β with
This finishes the proof.
kf kq,β . kSα
f k.
5. FURTHER RESULTS
(cid:3)
: Ap1
α1 → Ap2
5.1. Compactness. Under the assumptions of Theorem 3, actually one has that the small
Hankel operator hα
α2 is bounded if and only if it is compact. This is from
f
a general result of Banach space theory. It is known that, for 0 < p2 < p1 < ∞, every
bounded operator from ℓp1 to ℓp2 is compact (see, for example Theorem I.2.7, p.31 in [15]).
α is isomorphic to ℓp (see, Theorem 11, p.89 in [21],
Since the weighted Bergman space Ap
note that the same proof there works for weighted Bergman spaces on the unit ball Bn),
we get directly the above result.
5.2. Small Hankel operators with the same weights. Concerning the boundedness of
α (the case when all the weights are the same)
the small Hankel operator hα
f
for all possible choices of 0 < p1, p2 < ∞, we mention here that the case p1 = p2 > 1 is
by now classical (see [13], [23] and [4]), and in this case the boundedness is equivalent to
the symbol f being in the Bloch space B, that consists of those holomorphic functions f
on Bn with
α → Ap2
: Ap1
kf kB = f (0) + sup
z∈Bn
(1 − z2)Rf (z) < ∞.
Here, Rf denotes the radial derivative of f , that is,
Rf (z) =
zk
∂f
∂zk
(z),
z = (z1, . . . , zn) ∈ Bn.
nXk=1
The Bloch space also admits an equivalent norm in terms of the invariant gradiente∇f (z) :=
∇(f ◦ ϕz)(0) as follows
kf kB ≍ f (0) + sup
z∈Bn
e∇f (z).
The case 0 < p1 ≤ p2 is completely settled in [4] (actually the results are stated for the
unweighted Bergman spaces Ap, but the proofs works also for the weighted case). The
description for the case p1 = p2 = 1 is that f must belong to the so called logarithmic
Bloch space, a result that goes back to the one dimensional result of Attele [2]. Concerning
estimates with loss, in [4] Bonami and Luo obtained a description for the case 0 < p2 < p1
with p2 < 1 (again the result in [4] is stated for the unweighted Bergman spaces). Thus, in
view of Corollary 4, to complete the picture it remains to deal with the case p1 > p2 = 1
12
J. PAU AND R. ZHAO
(this problem is also open for the unit disk). In that case, also in [4], some partial results are
obtained (again for the unweighted case). Mainly, they provide a pointwise estimate that is
necessary for the small Hankel operator to be bounded, and they show that the condition
(16)
f (z) log
2
1 − z
∈ Lp′
1(Bn, dvα)
is sufficient. Moreover, they conjecture that the previous condition is also necessary. We
have not been able to prove the conjecture, but we are going to shed some light on that
problem.
Theorem 6. Let f ∈ H(Bn), α > −1 and p1 > 1. Let p′
Then hα
f
is bounded.
1 be the conjugate exponent of p1.
α is bounded if and only if the multiplication operator Mf : B → Ap′
α
α → A1
: Ap1
1
Before going to the proof we need first some preparation. First of all, recall that the
α under the integral pairing h , iα (see [24, Theorem 3.17]).
Bloch space is the dual of A1
We also need the following lemma, whose one dimensional analogue is essentially proved
in [3].
Lemma 7. Let 1 < p < ∞, σ > −1, and n + 1 + σ < b. Then
ZBn
f (z) − f (a)p
1 − ha, zib dvσ(z) .ZBn
dvσ(z)
1 − ha, zib
e∇f (z)p
for any f ∈ H(Bn) and a ∈ Bn.
Proof. We are going to prove first that, for 0 ≤ t < n + 1 + σ,
From [24, p.51], for β big enough, say β ≥ 1 + σ, we have
(17)
ZBn
f (z) − f (0)p
1 − ha, zit dvσ(z) .ZBn
f (z) − f (0) ≤ CZBn
(1 − w2)p Rf (w)p
1 − ha, wit
dvσ(w).
(1 − w2) Rf (w) dvβ−1(w)
.
1 − hz, win+β
Take a small number ε > 0 with σ − ε max(p, p′) > −1, where p′ denotes the conjugate
exponent of p, and t < n + 1 + σ − εp. An application of Holder's inequality and Lemma
B yields
f (z) − f (0)p . (1 − z2)−εpZBn
(1 − w2)p Rf (w)p dvβ−1+εp(w)
1 − hz, win+β
.
This together with Fubini's theorem and [19, Lemma 2.5] gives
ZBn
f (z) − f (0)p
1 − ha, zit dvσ(z)
.ZBn
.ZBn
(1 − w2)p Rf (w)p(cid:18)ZBn
(1 − w2)p Rf (w)p
1 − ha, wit
dvσ(w)
1 − ha, zit 1 − hz, win+β(cid:19) dvβ−1+εp(w)
dvσ−εp(z)
proving (17). Now, a change of variables z = ϕa(ζ) gives (see [24, Proposition 1.13])
ZBn
f (z)−f (a)p
1 − ha, zib dvσ(z) =ZBn
(f ◦ ϕa)(ζ)−(f ◦ ϕa)(0)p
1 −ha, ϕa(ζ)ib
(1 −a2)n+1+σ
1−ha, ζi2(n+1+σ) dvσ(ζ).
WEAK FACTORIZATION AND HANKEL OPERATORS
13
From [24, Lemma 1.3] we have
1 − ha, ϕa(ζ)i = 1 − hϕa(0), ϕa(ζ)i =
1 − a2
1 − ha, ζi
.
Therefore we obtain
ZBn
f (z)−f (a)p
1 − ha, zib dvσ(z) = (1 −a2)n+1+σ−bZBn
Due to our condition b > n + 1 + σ, we have
(f ◦ ϕa)(ζ)−(f ◦ ϕa)(0)p
1 −ha, ζi2(n+1+σ)−b
and we can apply (17) to get
t = 2(n + 1 + σ) − b < n + 1 + σ
ZBn
f (z)−f (a)p
1 − ha, zib dvσ(z) . (1 −a2)n+1+σ−bZBn
(1 − ζ2)p R(f ◦ ϕa)(ζ)p
1 − ha, ζi2(n+1+σ)−b
Since
another change of variables w = ϕa(ζ) finally gives
(1 − ζ2) R(f ◦ ϕa)(ζ) ≤ e∇(f ◦ ϕa)(ζ) = e∇f (ϕa(ζ)),
ZBn
1 − ha, zib dvσ(z) .ZBn
1 − ha, wib dvσ(w).
f (z)−f (a)p
e∇f (w)p
completing the proof of the lemma.
dvσ(ζ).
dvσ(ζ).
(cid:3)
After these preparations, we are now ready for the proof of Theorem 6.
Proof of Theorem 6. Assume first that the small Hankel operator hα
f
bounded. Let g ∈ Ap1
α . From the pointwise estimate for Bergman spaces, we get
: Ap1
α → A1
α is
hg, f iα = hα
f g(0) ≤ Ckhα
f gk1,α ≤ Ckhα
f k · kgkp1,α.
Therefore, by duality, we have that f ∈ Ap′
(18)
kf kp′
α with
1,α ≤ Ckhα
1
f k.
Recall that hα
f
f k ≍ khα
f
with kSα
(19)
α → A1
: Ap1
k. Also, since for any g ∈ Ap1
α is bounded, if and only if, Sα
α and h ∈ B,
f : Ap1
α → A1
α is bounded,
hSα
f g, hiα = hf, ghiα = hSα
f h, giα
we know that Sα
f : B → Ap′
1
α is bounded, and moreover, we have
kSα
f k
B→A
p′
1
α
. khα
f k.
For g in the Bloch space B, one has
kMf gkp′
p′
(20)
1
1,α =ZBn
.ZBn
f (z) g(z) p′
1 dvα(z)
Sα
f g(z) p′
1 dvα(z) +ZBn
f (z) g(z) − Sα
f g(z)p′
1 dvα(z).
14
J. PAU AND R. ZHAO
Due to the boundedness of Sα
f : B → Ap′
α ,
1
(21)
ZBn
Sα
f g(z) p′
1 dvα(z) ≤ kSα
f kp′
1
B→A
· kgkp′
1
B . khα
f kp′
1 · kgkp′
B .
1
p′
1
α
On the other hand, by the reproducing formula for Bergman spaces and Holder's inequality,
f (z) g(z) − Sα
p′
1
1
f (w)p′
f g(z)p′
f (w) (g(z) − g(w))
1 − hw, zin+1+α dvα+εp′
1 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ZBn
(1 − hz, wi)n+1+α dvα(w)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1 (w)!(cid:18)ZBn
1 − hw, zin+1+α dvα−εp1 (w) .ZBn
g(z) − g(w)p1
≤ ZBn
ZBn
where ε > 0 satisfies α − ε max(p1, p′
1) > −1. Using Lemma 7 and Lemma B we get
g(z) − g(w)p1
1 − hw, zin+1+α dvα−εp1 (w)(cid:19) p′
1
p1
,
1 − hw, zin+1+α dvα−εp1 (w)
e∇g(w)p1
Therefore, this together with Fubini's theorem, Lemma B and the estimate (18) gives
. kgkp1
B (1 − z2)−εp1 .
ZBn
(22)
f (z) g(z) − Sα
f g(z)p′
1 dvα(z)
≤ Ckgkp′
f (w)p′
1
B ZBn
1(cid:18)ZBn
dvα−εp′
1 − hw, zin+1+α(cid:19) dvα+εp′
1 (z)
1 (w)
≤ Ckgkp′
1
B · kf kp′
p′
1
1,α ≤ Ckhα
f kp′
1 · kgkp′
B .
1
Putting together the estimates (20), (21) and (22) it follows that Mf : B → Ap′
with kMf k
k.
B→A
p′
1
α
. khα
f
1
α is bounded
Conversely, suppose that Mf : B → Ap′
projection Pα : Lp′
and so obviously, Sα
well-known that the dual space of B0 is A1
Chapter 3 of [24]), from (19) we know that Sα
1 (Bn, dvα) → Ap′
f : B0 → Ap′
α is bounded. By the boundedness of the
α is also bounded,
α is bounded, where B0 is the little Bloch space, and it is
α under the integral pair h , iα (see, for example,
(cid:3)
α one deduces that Sα
f : B → Ap′
α is bounded.
f : Ap1
α → A1
1
1
1
1
As a consequence of Theorem 6 we can easily obtain the sufficient and necessary con-
ditions given in [4] as well as another relevant necessary condition for the boundedness of
hα
f
α.
α → A1
: Ap1
Corollary 8. Let f ∈ H(Bn), α > −1 and p1 > 1.
(i) If (16) holds, then hα
f
(ii) If hα
f
α → A1
α → A1
α is bounded, then
: Ap1
: Ap1
α is bounded.
(23)
sup
z∈Bn
(1 − z2)(n+1+α)/p′
1 f (z)(cid:16) log
2
1 − z2(cid:17) < ∞
WEAK FACTORIZATION AND HANKEL OPERATORS
15
and
(24)
ZBn
f (z)p′
1(cid:16) log
1
2
2
1 − z2(cid:17) p′
dvα(z) < ∞.
Proof. Part (i) follows directly from Theorem 6 and the pointwise estimate for Bloch func-
tions
To prove part (ii), for each z ∈ Bn, the function
g(z) ≤ kgkB log
2
1 − z2 .
gz(w) = log
2
1 − hw, zi
is in the Bloch space with kgzkB ≤ C with the constant C independent of the point z.
Therefore, from the pointwise estimate for functions in Bergman spaces, we get
(1 − z2)n+1+α(cid:16)f (z) log
1
2
1 − z2(cid:17)p′
= (1 − z2)n+1+αf (z) gz(z)p′
1
. kf gzkp′
1
1,α = kMf gzkp′
p′
1
p′
1,α
≤ kMf k
B→A
p′
1
α
· kgzkp′
1
B . kMf k
,
p′
1
α
B→A
Indeed, clearly Mf : B → Ap′
and (23) follows due to Theorem 6. The necessity of (24) is also a consequence of
Theorem 6.
α is bounded if and only if the measure
dµf (z) = f (z)p′
1- Carleson measure for the Bloch space (see [7, 10]
for the definition); and by Proposition 1.4 in [7] (the one dimensional case appears in [10]
and [14]) this implies (24), finishing the proof.
(cid:3)
1 dvα(z) is a p′
1
The established connection between Hankel operators on Bergman spaces and Carleson
measures for the Bloch space makes even more interesting the problem (as far as we know,
still open) of describing those measures.
Acknowledgments: The authors would like to thank the referee for valuable comments
that improved the final version of the paper.
REFERENCES
[1] N. Arcozzi, R. Rochberg, E. Sawyer and B. D. Wick, Bilinear forms on the Dirichlet space, Anal. PDE 3
(2010), 21 -- 47.
[2] K. Attele, Toeplitz and Hankel operators on Bergman one space, Hokkaido Math. J. 21 (1992), 279 -- 293.
[3] D. Blasi and J. Pau, A characterization of Besov type spaces and applications to Hankel type operators,
Michigan Math. J. 56 (2008), 401-417.
[4] A. Bonami and L. Luo, On Hankel operators between Bergman spaces on the unit ball, Houston J. Math.
31 (2005), 815 -- 828.
[5] R. Coifman and R. Rochberg, Representation theorems for holomorphic and harmonic functions in Lp,
Asterisque 77 (1980), 11-66.
[6] R. Coifman, R. Rochberg and G. Weiss, Factorization theorems for Hardy spaces in several variables, Ann.
of Math. (2) 103 (1976), 611 -- 635.
[7] E. Doubtsov, Carleson-Sobolev measures for weighted Bloch spaces, J. Funct. Anal. 258 (2010), 2801 --
2816.
[8] P.L. Duren, 'Theory of H p Spaces', Academic Press, New York-London 1970. Reprint: Dover, Mineola,
New York 2000.
[9] J. Garnett and R. Latter, The atomic decomposition for Hardy spaces in several complex variables, Duke
Math. J. 45 (1978), 815 -- 845.
16
J. PAU AND R. ZHAO
[10] D. Girela, J. A. Pel´aez, F. P´erez-Gonz´alez and J. Rattya, Carleson measures for the Bloch space, Integral
Equations and Operator Theory 61 (2008), 511 -- 547.
[11] M. Gowda, Nonfactorization theorems in weighted Bergman and Hardy spaces on the unit ball of Cn, Trans.
Amer. Math. Soc. 277 (1983), 203 -- 212.
[12] C. Horowitz, Factorization theorems for functions in the Bergman spaces, Duke Math. J., 44 (1977), 201 --
213.
[13] S. Janson, J. Peetre and R. Rochberg, Hankel forms and the Fock space, Revista Mat. Iberoamericana 3
(1987), 61 -- 138.
[14] T.G. Limperis, 'Embedding theorems for the Bloch space', PhD thesis, University of Arkansas, 1998.
[15] J. Lindenstrauss and L. Tzafriri, 'Classical Banach Spaces', Lecture Notes in Math. 338, Springer-Verlag,
Berlin, 1973.
[16] D. H. Luecking, Representations and duality in weighted spaces of analytic functions, Indiana Univ. Math.
J. 34 (1985) 319 -- 336.
[17] D. H. Luecking, Embedding theorems for spaces of analytic functions via Khinchine's inequality, Michigan
Math. J. 40 (1993), 333 -- 358.
[18] Z. Nehari, On bounded bilinear forms, Ann. of Math. (2) 65 (1957), 153 -- 162.
[19] J.M. Ortega and J. Fabrega, Pointwise multipliers and corona type decomposition in BM OA, Ann. Inst.
Fourier (Grenoble) 46 (1996), 111 -- 137.
[20] R. Rochberg, Decomposition theorems for Bergman spaces and their applications, in 'Operators and Func-
tion Theory', D. Reidel, 1985, 225 -- 277.
[21] P. Wojtaszczyk, 'Banach Spaces for Analysts', Cambridge Studies in Advanced Mathematics 25. Cambridge
University Press, 1991.
[22] R. Zhao and K. Zhu, 'Theory of Bergman spaces in the unit ball of Cn', Mem. Soc. Math. Fr. (N.S.), 115
(2008), vi+103 pp.
[23] K. Zhu, Hankel operators on the Bergman spaces of bounded symmetric domains, Trans. Amer. Math. Soc.
324 (1991), 707 -- 730.
[24] K. Zhu, 'Spaces of Holomorphic Functions in the Unit Ball', Springer-Verlag, New York, 2005.
JORDI PAU, DEPARTMENT DE MATEM `ATICA APLICADA I ANALISI, UNIVERSITAT DE BARCELONA,
08007 BARCELONA, CATALONIA, SPAIN
E-mail address: [email protected]
RUHAN ZHAO, DEPARTMENT OF MATHEMATICS, SUNY BROCKPORT, BROCKPORT, NY 14420, USA
E-mail address: [email protected]
|
1910.10538 | 1 | 1910 | 2019-10-23T12:58:33 | Similarity Invariants of Essentially normal Cowen-Douglas Operators and Chern Polynomials | [
"math.FA"
] | In this paper, we systematically study a class of essentially normal operators by using the geometry method from the Cowen-Douglas theory and prove a Brown-Douglas-Fillmore theorem in the Cowen-Douglas theory. More precisely, the Chern polynomials and the second fundamental forms are the similarity invariants (in the sense of Herrero) of this class of essentially normal operators. | math.FA | math |
SIMILARITY INVARIANTS OF ESSENTIALLY NORMAL COWEN-DOUGLAS
OPERATORS AND CHERN POLYNOMIALS
IN MEMORY OF R. G. DOUGLAS
CHUNLAN JIANG, KUI JI AND JINSONG WU
Abstract. In this paper, we systematically study a class of essentially normal operators by using the
geometry method from the Cowen-Douglas theory and prove a Brown-Douglas-Fillmore theorem in
the Cowen-Douglas theory. More precisely, the Chern polynomials and the second fundamental forms
are the similarity invariants (in the sense of Herrero) of this class of essentially normal operators.
1. Introduction
People have long sought to establish connections between operator theory on the one hand, and
geometry and algebraic topology on the other. The Brown-Douglas-Fillmore theory [4] and the Cowen-
Douglas operator theory [10] together provide a powerful suite of tools for building such connections.
The Brown-Douglas-Fillmore theorem states that any two essentially normal operators are unitarily
equivalent up to a compact perturbation (which is called essentially unitary equivalence) if and only
if their essential spectrums and Fredholm indexes are the same. In 1991, Berg and Davidson [5] give
a constructive proof of the Brown-Douglas-Fillmore theorem. The Brown-Douglas-Fillmore theorem
shows that the essentially unitary invaritants of an essentially normal operators are essential spectrums
and Fredholm indexes. It is highly nontrivial to obtain unitary (similarity) invariants of operators.
An alternative way is to study the closure of unitary (similarity) orbits. In general, the closure of
unitary orbit of a Hilbert space operator was completely determined by Hadwin [25]. The closure of
similarity orbit of a Hilbert space operator was "essentially" solved by Apostol, Herrero and Voiculescu
([3], Chapter 9 of [2]). The similarity orbit of a normal operator was characterized by Fialkow [21].
In 1986, Herrero [27] introduce (U + K)-orbits between unitary orbits and similarity orbits in which
the unitary operator is replaced by an invertible operator of the form "unitary operator + compact
operator". Herrero [27] ask if we can find a simple characterization of the closure of (U + K)-orbit
of a Hilbert space operator. In 1993, Guinand and Marcoux [23, 24] completely characterized the
the closure of (U + K)-orbits of normal operators, compact operators and certain weighted shifts
respectively. In 1998, the first author and Wang [42] completely characterize the closure of (U + K)-
Inspired by Herrero's (U + K)-equivalence and the
orbits of certain essentially normal operators.
Brown-Douglas-Fillmore theorem, one could ask
Question 1.1. What are the (U + K)-invariants for essentially normal operators?
2000 Mathematics Subject Classification. Primary 47C15, 47B37; Secondary 47B48, 47L40.
The first and the second authors were supported by National Natural Science Foundation of China (Grant No.
11831006 and 11922108). The third author was supported by National Natural Science Foundation of China (Grant No.
11771413).
1
2
CHUNLAN JIANG, KUI JI AND JINSONG WU
We have not found any (partial) results to this question in any literature. Moreover, it is challenging
to characterize the (U + K)-orbits or invariants of arbitrary operators.
The Cowen-Douglas theory introduced geometry operators for a given open subset Ω ⊂ C and a
natural number n. These operators are called Cowen-Douglas operators over Ω with index n. (The
set of these operators is denoted by Bn(Ω).) Cowen and Douglas showed that each operator in Bn(Ω)
gives rise a rank n Hermitian holomorphic vector bundle over Ω. The curvature of the vector bundle
defines the curvature of the operator. It is an elegant and deep result that the unitary invariants of
Cowen-Douglas operators are their curvatures and their covariant partial derivatives. Based on their
results, it is natural to ask what the similarity invariants of Cowen-Douglas operators are. Cowen and
Douglas [10] conjectured that the Cowen-Douglas operators over the open unit disk with index one are
similar to each other if and only if the limit of the quotient of the two curvatures is one. The conjecture
turns out to be false by Clark and Misra [12, 13]. The failure of the conjecture leads mathematicians
to find similarity invariants for Cowen-Douglas operators with additional structures. In 2013, Douglas,
Kwon and Treil [16] studied the similarity of hypercontractions and backward Bergman shifts. Hou, Ji
and Kown [30] showed that the geometry invariant is good for similarity invariants. In 2017, the first
two authors, Keshari and Misra introduced and studied a class of Cowen-Douglas operators with flag
structures, denoted by FBn(Ω). Later, the first two authors and Misra [37] introduced and studied
a class Cowen-Douglas operators with a little more structures. In 2019, the first two authors and
Keshari [36] found that the curvatures and the second fundamental forms can completely characterize
the similarity invariant for a norm dense class of Cowen-Douglas operators denoted by CFBn(Ω).
Thanks to the Brown-Douglas-Fillmore theory and the Cowen-Douglas theory, problems in geom-
etry and algebraic topology can be approached using techniques from operator theory. Inspired by
this, Douglas, Tang and Yu [18] used operator-theoritic methods to prove an analytic Grothendieck-
Riemann-Roch theorem for certain spaces with singularities. In this paper, we proceed in the reverse
direction, using methods from complex geometry to find a complete set of invariants for the simi-
larity problem for essentially normal operators. Besides the curvature, the Chern polynomial is also
important in geometry. We partially answer Questions 1.1 for certain essentially normal Cowen-
Douglas operators. (The set of these operators is denoted by NCFBn(Ω).) We also show that the
set NCFBn(Ω) contains enough nontrivial operators (see Theorem 3.14). It is surprising for us to
find that the Chern polynomials and the second fundamental forms can completely characterize the
(U + K)-invariants for NCFBn(Ω) (see Theorem 4.8 and 5.1). This result is a version of the Brown-
Douglas-Fillmore theorem in the Cowen-Douglas theory.
It is important to notice that the Chern
polynomials can not be taken as similarity invariants for CFBn(Ω). Our results show the Chern poly-
nomial can significantly simplify the (U + K)-invariants for NCFBn(Ω). It is also worth to point out
that the previous study on (U + K)-orbits requires that the spectrum of the operator has to be simply
connected. However, our results do not have assumption on the spectrum. In a word, our approach
offers a natural complement to that of Douglas, Tang and Yu [18].
The paper is organized as follows. In Section 2 we recall the definition and some properties of Cowen-
Douglas operators and the subclasses FBn(Ω),CFBn(Ω) of Cowen-Douglas operators.
In Section
3 we introduce the class NCFBn(Ω) of essentially normal Cowen-Douglas operators and study the
structures of the operators and present some examples. In Section 4 we obtain completely the (U + K)-
invariants of NCFBn(Ω). In Section 5 we obtain a complete set of unitary invariants of in NCFBn(Ω).
In Section 6 we suggest some further research on (U + K)-orbits and (U + K)-homogeneous.
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
3
2. Preliminaries
Throughout this paper, we denote H by an infinite-dimensional separable Hilbert space, L(H) the
algebra of all bounded operators on H, K(H) the algebra of all compact operators on H, and A(H)
the Calkin algebra L(H)/K(H). Let Q : L(H) → A(H) be the quotient map onto the Calkin algebra.
An operator T on H is essentially normal if Q(T ) is normal, i.e. T ∗T − T T ∗ ∈ K(H).
We are interested in the following equivalences of operators: for any T1, T2 ∈ L(H), we say
(1) T1 is unitarily equivalent to T2 if there is a unitary operator U ∈ L(H) such that T1U = U T2,
(2) T1 is (U +K)-equivalent to T2 if there are a unitary operator U ∈ L(H) and a compact operator
K ∈ K(H) such that U + K is invertible and T1(U + K) = (U + K)T2, denoted by T1 ∼U +K T2;
(3) T1 is similar to T2 if there is an invertible operator S ∈ L(H) such that T1S = ST2, denoted
denoted by T1 ∼u T2;
by T1 ∼s T2;
Remark 2.1. Suppose that T1, T2 ∈ L(H) are essentially unitary equivalent. Usually, we do not have
T1 ∼U +K T2. Suppose that {en}n≥0 is an orthogonal normal basis of H and T1 is the shift on H, i.e.
en+1, n ≥ 0. Then T1 − T2 is
en+1 and T1 is not similar to T2. In
T1en = en+1, n ≥ 0 and T2 is a Bergman shift such that T2en =r n
a compact operator such that (T1 − T2)en =
fact, if T1 ∼s T2, i.e. T1X = XT2 for some invertible operator X, we have that
n + 1
hXen, eni =r n + 1
n hXT2en−1, eni =r n + 1
1 eni =r n + 1
n hXen−1, en−1i,
1
n + 1 +pn(n + 1)
n hXen−1, T ∗
i.e.
n→∞hXen, eni = lim
lim
n→∞
√n + 1hXe0, e0i = ∞.
We see that X is not bounded, which leads a contradiction and T1 is not similar to T2.
We now recall the setup for Cowen-Douglas operators (geometry operators) in the following. Let H
be a separable infinite-dimensional Hilbert space and Gr(n,H) the set of all n-dimensional subspaces
of H for any n ∈ N. Suppose Ω ⊂ C is an open bounded connected subset and n ∈ N. A map
t : Ω → Gr(n,H) is a holomorphic curve if there exist n (point-wise linearly independent) holomorphic
functions γ1, γ2,··· , γn on Ω taking values in H such that t(w) =W{γ1(w),··· , γn(w)}, w ∈ Ω. Each
holomorphic curve t : Ω → Gr(n,H) gives rise to a rank n Hermitian holomorphic vector bundle Et
over Ω, namely,
Et = {(x, w) ∈ H × Ω x ∈ t(w)}
and π : Et → Ω, where π(x, w) = w.
The set {γ1 . . . , γn} of holomorphic functions is a holomorphic frame for t, denoted by γ. An operator
T acting on H is said to be a Cowen-Douglas operator with index n associated with an open bounded
subset Ω, if
(1) T − w is surjective,
(2) dim Ker (T − w) = n for all w ∈ Ω,
Ker (T − w) = H.
(3) Ww∈Ω
4
CHUNLAN JIANG, KUI JI AND JINSONG WU
The set of Cowen-Douglas operators is denoted by Bn(Ω).
Cowen and Douglas[10] showed that each operator T in Bn(Ω) gives also rise to a rank n Hermitian
holomorphic vector bundle ET over Ω,
ET = {(x, w) ∈ H × Ω x ∈ Ker (T − w)}
Two holomorphic curves t and t are said to be
(1) unitarily equivalent (denoted by t ∼u t), if there exists a unitary operator U ∈ L(H) such that
U (w)t(w) = t(w), where U (w) := UEt(w) is the restriction of the unitary operator U to the
fiber Et(w) = π−1(w).
and π : ET → Ω where π(x, w) = w.
(2) similar (denoted by t ∼s t) if there exists an invertible operator X ∈ L(H) such that
In
X(w)t(w) = t(w), where X(w) := XEt(w) is the restriction of X to the fiber Et(w).
this case, we say that the vector bundles Et and Et are similar.
For an open bounded connected subset Ω of C, a Cowen-Douglas operator T with index n induces
a non-constant holomorphic curve t : Ω → Gr(n,H), namely, t(w) = Ker (T − w), w ∈ Ω and hence a
Hermitian holomorphic vector bundle Et (here vector bundle Et is same as ET ). Unitary and similarity
invariants for the operator T are obtained from the vector bundle ET .
The metric of the Hermitian holomorphic vector bundle ET with respect to a holomorphic frame γ
is given by the n × n matrix
hγ(w) = (hγj(w), γk(w)i)n
k,j=1 , w ∈ Ω.
Note that {γ1(w), . . . , γn(w)} is linearly independent. The matrix hγ(w) is invertible. The curvature
of the Hermitian holomorphic vector bundle ET is
(2.1)
¯∂(cid:0)h−1
γ (w)∂hγ (w)(cid:1) = −
∂
∂w(cid:18)h−1
γ (w)
∂
∂w
hγ(w)(cid:19) dw ∧ d ¯w,
where ∂ is the differential operator with respect to the complex variable w and ¯∂ is with respect to the
conjugation ¯w. The curvature KT of the Cowen-Douglas operator T over Ω is given by the coefficient
of Equation (2.1):
The total Chern form of the holomorphic bundle ET is defined to be
KT (w) := −
∂
γ (w)
∂w(cid:18)h−1
CET = det(cid:18)I +
CET (w) = det(cid:18)I + w
i
2π
∂
∂w
hγ(w)(cid:19) .
KT(cid:19) .
KT(cid:19) .
i
2π
The Chern polynomial CET (w) of ET with respect to a complex variable w is defined to be
Next, we will describe the structure of Cowen-Douglas operators in Bn(Ω).
Theorem 2.2 (Upper triangular representation theorem, [42]). Let T ∈ Bn(Ω). Then there exists an
orthogonal decomposition H = H1 ⊕ H2 ⊕ ··· ⊕ Hn and operators T1,1, T2,2,··· , Tn,n in B1(Ω) such
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
5
that T has the following form
(2.2)
T =
···
···
T1,n
T2,n
T1,1 T1,2 T1,3
0 T2,2 T2,3
...
0 Tn−1,n−1 Tn−1,n
···
...
...
...
0
0
···
···
0
Tn,n
...
.
Let {γ1, γ2,··· , γn} be a holomorphic frame of ET with H = W{γj(w), w ∈ Ω, 1 ≤ j ≤ n}, and
tj : Ω → Gr(1,Hj) be a holomorphic frame of ETj,j , 1 ≤ j ≤ n. Then we can find certain relation
between {γj}n
j=1 as the following equations:
j=1 and {tj}n
(2.3)
γ1 = t1
γ2 = φ1,2(t2) + t2
γ3 = φ1,3(t3) + φ2,3(t3) + t3
··· = ··· ··· ··· ··· ··· ··· ··· ···
γj = φ1,j(tj) + ··· + φi,j(tj) + ··· + tj
··· = ··· ··· ··· ··· ··· ··· ··· ··· ··· ··· ···
γn = φ1,n(tn) + ··· + φi,n(tn) + ··· + tn,
where φi,j, (i, j = 1, 2,··· , n) are certain holomorphic bundle maps.
In the rest of the paper, we always assume that a Cowen-Douglas operator T ∈ Bn(Ω) has a
k,j=1 with respect to an orthogonal decomposition of the Hilbert space such that
matrix form (Tk,j)n
the diagonals are in B1(Ω).
Recall that the commutant {T}′ of T on H is the set of operators in L(H) commuting with T . The
operator T is strongly irreducible if there is no non trivial idempotent in {T}′.
An operator T in L(H) is said to be strongly irreducible, if there is no non-trivial idempotent
operator in {T}′, where {T}′ denotes the commutant of T , i.e., {T}′ = {B∈L(H) : T B = BT}. It
is known that for any T ∈ B1(Ω), T is strongly irreducible. Every Cowen-Douglas operator can be
written as the direct sum of finitely many strongly irreducible Cowen-Douglas operators (see [42],
chapter 3). By the results in [44], the similarity of Cowen-Douglas operators can be reduced to the
similarity of strongly irreducible Cowen-Douglas operators.
Now, we recall some classes of Cowen-Douglas operators with additional structures.
Definition 2.3. Let FBn(Ω) ⊂ Bn(Ω) be the set of T ∈ Bn(Ω) on H = H1 ⊕ ··· ⊕ Hn, having the
following form
T1,1 T1,2
T2,2
T =
··· T1,n
··· T2,n
...
. . .
Tn,n
such that
(1) Tj,j ∈ B1(Ω) for 1 ≤ j ≤ n;
(2) Tj,jTj,j+1 = Tj,j+1Tj+1,j+1 for any 1 ≤ j ≤ n − 1, i.e. Tj,j+1 intertwines Tj,j and Tj+1,j+1.
Definition 2.4. Let T1, T2 ∈ L(H). Then the Rosenblum operators τT1,T2 is the map from L(H) to
L(H) defined by:
τT1,T2(X) = T1X − XT2, X ∈ L(H).
If T1 = T2 = T , we let δT = τT,T .
6
CHUNLAN JIANG, KUI JI AND JINSONG WU
Definition 2.5 (Property (H)). Let T1, T2 ∈ L(H). The ordered pair (T1, T2) has the Property (H)
if the following condition holds: if there exists X ∈ L(H) such that
(1) T1X = XT2,
(2) X = T1Z − ZT2,
then X = 0.
for some Z ∈ L(H),
Definition 2.6 (Definition 2.7 in [36]). Let CFBn(Ω) be the subset of operators T ∈ Bn(Ω) satisfies
the following properties:
(1) T can be written as an n × n upper-triangular matrix from with respect to a topological direct
decomposition of the Hilbert space and Diag T ∈ {T}′, where Diag T is the diagonal of T .
(2) For 1 ≤ k ≤ j ≤ n, Tk,j = Ck,jTk,k+1Tk+1,k+2 ··· Tj−1,j for some Ck,j ∈ {Tk,k}′ on Hk;
(3) The ordered pair (Tj,j, Tj+1,j+1), 1 ≤ j ≤ n − 1 satisfies the Property (H), that is,
Ker τTj,j ,Tj+1,j+1 ∩ Ran τTj,j ,Tj+1,j+1 = {0},
1 ≤ j ≤ n − 1.
Remark 2.7. According to 1.20 in [10], there exists a holomorphic function φk,j such that Ck,j =
φk,j(Tk,k) in Definition 2.6.
Lemma 2.8. [36] Let T ∈ CFBn(Ω). Then T is strongly irreducible if and only if Tj,j+1 6= 0 for any
j = 1, 2··· , n − 1.
In the end of this section, we recall some results for B1(Ω) and some technique lemmas.
Proposition 2.9 (Proposition 4.12 in [36]). Let T,eT ∈ B1(Ω). Then T ∼U +K eT if and only if there
exists X ∈ L(H) such that Q(X) = αQ(I), 0 < α < 1 and
KT (w) − K eT (w) =
∂w∂w
∂2
Ψ(w), Ψ(w) = ln(cid:18)kX(t(w))k2
kt(w)k2 + (1 − α2)(cid:19)
where t is a non zero section of ET .
Thoughout this paper, we denoted by σ(T ) the spectrum of an operator T ∈ L(H).
Lemma 2.10. [26] Let X, T ∈ L(H). If X ∈ Ker δT ∩ Ran δT , then σ(X) = {0}.
Lemma 2.11. Let X, T ∈ L(H). If Q(X) ∈ Ker δQ(T ) ∩ Ran δQ(T ), then σ(Q(X)) = {0}.
Proof. It follows from Lemma 2.10 directly.
(cid:3)
Let T ∈ FBn(Ω). For a 2× 2 block Ti,i
0
Ti,i+1
Ti+1,i+1!, if Ti,iTi,i+1 = Ti,i+1Ti+1,i+1, the corresponding
second fundamental form θi,i+1, which is obtained by R. G. Douglas and G. Misra (see in [17]), is
(2.4)
θi,i+1(T )(w) =
KTi,i(w) d ¯w
(cid:0) kTi,i+1(ti+1(w))k2
kti+1(w)k2
− KTi,i(w)(cid:1)1/2
.
Let T,eT have upper-triangular representation as in (1.1) and assume that Ti,i, Ti+1,i+1 and Ti,i, Ti+1,i+1
have intertwines Ti,i+1 and Ti,i+1, respectively. If KTi,i = K eTi,i
θi,i+1(T )(w) = θi,i+1( T )(w) ⇔ kTi,i+1(ti+1(w))k2
, then from (1.3) it is easy to see that
.
kti+1(w)k2
= keTi,i+1(eti+1(w))k2
keti+1(w)k2
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
7
3. Essentially normal Cowen-Douglas Operators
In this section, we introduce certain essentially normal Cowen-Douglas operators and study their
basic properties.
Definition 3.1. Let EN be the set of all essentially normal operators on the Hilbert space H. For an
open subset Ω in C and n ∈ N, we define NCFBn(Ω) to be a subset of EN ∩ CFBn(Ω) subject to the
following condition: for any T ∈ NCFBn(Ω), there exist idempotents P1, . . . , Pn satisfying
Pj = I,
PkPj = 0, k 6= j and Q(Pj)∗ = Q(Pj ) for j = 1, . . . , n such that T ∈ CFBn(Ω) with respect to the
topological direct decomposition Ran P1 ∔ ··· ∔ Ran Pn.
Remark 3.2. In Definition 2.6, the topological direct decomposition can be perturbed by an in-
vertible operator to an orthogonal decomposition. To study (U + K)-invariant, we require that the
topological direct decomposition can perturbed by a (U + K)-invertible operator to an orthogonal
decomposition. The additional condition in Definition 3.1 is to meet the requirement above (See the
following Proposition 3.3 for a proof).
nPj=1
Proposition 3.3. Suppose that H = H1 ∔ H2 ∔ ··· ∔ Hn is a topological direct decomposition of H
Pj = I, PkPj = 0, k 6= j where Ran Pj = Hj.
and there exist idempotents P1, P2, . . . , Pn such that
If there exists a compact operator Kj such that Pj = (cid:16) IHj Kj
decomposition Hj ⊕ H⊥
U + K is invertible, Qj = (U + K)Pj(U + K)−1 is an (orthogonal) projection, and I =Pn
Proof. We assume that P1 has the matrix form with respect to the orthogonal decomposition H =
H1 ⊕ H⊥
0 (cid:17) with respect to the Hilbert space
j . Then there exist a unitary operator U and a compact operator K such that
nPj=1
j=1 Qj.
1 as
0
1
,
0
1,2
H⊥
P1 = I P (1)
0 ! H1
1,2 = −P (1)
I ! = I K (1)
1,2
0
where P (1)
1,2 is a compact operator. Suppose K (1)
1,2 , then we have that
0
0
1,2
1,2
I P (1)
0 ! I K (1)
I ! I 0
0 0! .
I ! and Q1 = (U1 + K1)P1(U1 + K1)−1, then
Pj (U1 + K1)−1.
I = Q1 ⊕ (U1 + K1)
Pj! (U1 + K1)−1 is a decomposition of IRan (I−Q1), repeating the argument
nXj=2
Set U1 + K1 = I K (1)
1,2
0
Since (U1 + K1) nPj=2
above for (U1 + K1)P2(U1 + K1)−1, there exists U2 + K2 such that
(I ⊕ (U2 + K2))(0 ⊕ (U2 + K2)P2(U2 + K2)−1)(I ⊕ (U2 + K2)−1)
8
CHUNLAN JIANG, KUI JI AND JINSONG WU
is an orthogonal projection (denoted by Q2). Then we have that
I = Q1 ⊕ Q2 ⊕ (I ⊕ (U2 + K2))(U1 + K1) nXi=3
= (I ⊕ (U2 + K2))(U1 + K1) nXi=1
Pi! (U1 + K1)−1(I ⊕ (U2 + K2)−1)
Pi! (U1 + K1)−1(I ⊕ (U2 + K2)−1)
By an inductive proof, we can find unitary operators Uj and compact operators Kj, j = 1, 2··· , n
such that
Q1⊕Q2⊕··· Qn = (I⊕I⊕··· (Un+Kn))··· (U1+K1)
This finishes the proof of this lemma.
nXj=1
Pj (U1+K1)−1 ··· (I⊕I⊕··· (Un+Kn)−1).
(cid:3)
For any topological direct decomposition of H, H = H1 ∔ H2 ∔ ··· ∔ Hn, if there exist n idem-
k 6= j and Ran Pj = Hj. Then we can
j=1 =
j=1 be a set of orthogonal projections with QkQj = 0, i 6= j. Furthermore,
H = (U + K)H1 ⊕ (U + K)H2 ⊕ ··· ⊕ (U + K)Hn,
potents P1, P2, . . . , Pn such that
find a unitary operator U and a compact operator K such that U + K is invertible and {Qj}n
(cid:8)(U + K)Pj(U + K)−1(cid:9)n
Pj = I, PkPj = 0,
nPj=1
where (U + K)Hj = Ran Qj.
3.1. Structure. We characterize the diagonals and off-diagonals of the operators in NCFBn(Ω) in
the following theorem:
Theorem 3.4. Let T ∈ NCFBn(Ω). Then we have that
(1) Tk,k, 1 ≤ k ≤ n are essentially normal;
(2) Tj,k are compact for 1 ≤ j < k ≤ n.
Proof. By the assumption that T is essentially normal, we have Q(T )Q(T ∗) = Q(T ∗)Q(T ). Expanding
it in the matrix form, we obtain that
Q(T1,1) Q(T1,2) ··· Q(T1,n)
Q(T2,2) ··· Q(T2,n)
...
...
Q(Tn,n)
(3.1)
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
1,n) Q(T ∗
2,n) ··· Q(T ∗
Recall that T ∈ CFBn(Ω), we see that
...
Q(T ∗
n,n)
=
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
2,n) ··· Q(T ∗
1,n) Q(T ∗
...
Q(T ∗
n,n)
Q(T1,1) Q(T1,2) ··· Q(T1,n)
Q(T2,2) ··· Q(T2,n)
...
...
Q(Tn,n)
.
and
Tk,kTk,j = Tk,jTj,j,
1 ≤ k < j ≤ n,
Q(Tk,k)Q(Tk,j) = Q(Tk,j)Q(Tj,j),
1 ≤ k < j ≤ n.
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
9
Without loss of generality, we can assume that
Tk,j = Tk,k+1Tk+1,k+2 ··· Tj−1,j, 1 ≤ k < j ≤ n.
By Equation (3.1), we have that
Q(T ∗
1,n)Q(T1,1) = Q(Tn,n)Q(T ∗
1,n),
and
Thus, we have that
Q(T ∗
1,n)Q(T1,2) + Q(T ∗
2,n)Q(T2,2) = Q(Tn,n)Q(T ∗
2,n).
Q(T ∗
1,n)Q(T1,2)Q(T2,n) + Q(T ∗
1,n)Q(T1,n) + Q(T ∗
Q(T ∗
2,n)Q(T2,2)Q(T2,n) = Q(Tn,n)Q(T ∗
2,n)Q(T2,n)Q(Tn,n) = Q(Tn,n)Q(T ∗
2,n)Q(T2,n),
2,n)Q(T2,n).
This indicates that
Q(T ∗
1,n)Q(T1,n) = Q(Tn,n)Q(T ∗
2,n)Q(T2,n) − Q(T ∗
2,n)Q(T2,n)Q(Tn,n).
On the other hand, since Q(T ∗
Q(T ∗
1,n)Q(T1,1) = Q(Tn,n)Q(T ∗
1,n)Q(T1,n)Q(Tn,n) = Q(Tn,n)Q(T ∗
1,n), we have that
1,n)Q(T1,n).
That means Q(T ∗
1,n)Q(T1,n) ∈ Ker δQ(Tn,n) ∩ Ran δQ(Tn,n). By Lemma 2.11, we have that
Since Q(T ∗
compact operator.
1,n)Q(T1,n) is positive, it follows that Q(T ∗
1,n)Q(T1,n) = 0. Hence Q(T1,n) = 0, i.e. T1,n is a
σ(cid:0)Q(T ∗
1,n)Q(T1,n)(cid:1) = {0}.
Now Equation (3.1) becomes:
(3.2)
=
Q(T1,1) Q(T1,2) ···
0
Q(T2,2) ··· Q(T2,n)
...
...
Q(Tn,n)
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
2,n) ··· Q(T ∗
...
Q(T ∗
0
n,n)
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
2,n) ··· Q(T ∗
...
Q(T ∗
0
n,n)
Q(T1,1) Q(T1,2) ···
0
Q(T2,2) ··· Q(T2,n)
...
...
Q(Tn,n)
.
By comparing the (n, 2), (n, 3)-entries in Equation (3.2), we have that
Q(T ∗
2,n)Q(T2,2) = Q(Tn,n)Q(T ∗
2,n),
and
Q(T ∗
2,n)Q(T2,3) + Q(T ∗
3,n)Q(T3,3) = Q(Tn,n)Q(T ∗
3,n).
10
CHUNLAN JIANG, KUI JI AND JINSONG WU
Repeating the same argument above, we also have that Q(T2,n) = 0, i.e. T2,n is a compact operator.
Similarly, we can obtain that Q(Tk,n) = 0, k < n. It follows that Equation (3.2) becomes:
=
Q(T1,1) Q(T1,2) ··· Q(T1,n−1)
Q(T2,2) ··· Q(T2,n−1)
...
...
Q(Tn−1,n−1)
0
0
...
0
Q(Tn,n)
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
Q(T ∗
1,n−1)
0
···
0
...
··· Q(T ∗
0
n−1,n−1)
0
Q(T ∗
n,n)
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
Q(T ∗
1,n−1)
0
···
0
...
··· Q(T ∗
0
n−1,n−1)
0
Q(T ∗
n,n)
Q(T1,1) Q(T1,2) ··· Q(T1,n−1)
Q(T2,2) ··· Q(T2,n−1)
...
...
Q(Tn−1,n−1)
0
0
...
0
Q(Tn,n)
.
Then we could focus on the (n − 1) × (n − 1)-matrix forms:
1,1)
1,2) Q(T ∗
2,2)
...
...
2,n−1) ··· Q(T ∗
1,n−1) Q(T ∗
Q(T1,1) Q(T1,2) ··· Q(T1,n−1)
Q(T2,2) ··· Q(T2,n−1)
Q(Tn−1,n−1)
Q(T ∗
Q(T ∗
...
...
Q(T ∗
...
n−1,n−1)
(3.3)
Q(T ∗
Q(T ∗
1,1)
1,2) Q(T ∗
2,2)
...
...
2,n−1) ··· Q(T ∗
1,n−1) Q(T ∗
...
Q(T ∗
n−1,n−1)
Q(T1,1) Q(T1,2) ··· Q(T1,n−1)
Q(T2,2) ··· Q(T2,n−1)
...
...
Q(Tn−1,n−1)
=
.
Notice that when n = 2, the conclusions holds by the same argument. Thus, by the inductive proof,
we have that Q(Tk,j) = 0, for any j > k. This finishes the proof of statement (1). Furthermore, we
have that
Q(T ) =
Q(T1,1)
Q(T2,2)
...
Q(Tn,n)
.
Thus, Q(T )Q(T ∗) = Q(T ∗)Q(T ) implies that Q(Tk,k)Q(T ∗
finishes the proof of statement (2).
k,k) = Q(T ∗
k,k)Q(Tk,k), 1 ≤ k ≤ n. This
(cid:3)
3.2. Examples. In this subsection, we show that there exists a strongly irreducible operator in
NCFBn(Ω) for any simply connected open subset Ω ⊂ C.
Let dµ denote the normalized Lebesgue area measure on the unit disk D. For any λ > −1, the
weighted Bergman space A2
λ is the space of analytic functions on D which are square-integrable with
respect to the measure dµλ = (λ + 1)(1 − z2)λdµ(z) (See [45] for more details on weighted Bergman
spaces). Let K(z, w) =
1
(1 − zw)λ be its positive definite kernel.
Proposition 3.5. Let T be a homogenous operator in B1(D). Then T ∈ NCFB1(D).
Proof. Let T ∈ B1(D) be a homogenous operator. Then there exists λ > 0 and shifts Mz on A2
λ
such that T ∼u M ∗
z . By a direct computation, there exist rank-one operators K0 and K1 such that
z Mz = I − K1. That means M ∗
z = I − K0, M ∗
MzM ∗
z , then T is
(cid:3)
also essentially normal.
z is essentially normal. Since T ∼u M ∗
Definition 3.6. Let T1, T2 ∈ B1(Ω). We call T1 ≺ T2 if Ker τT1,T2 6= {0}, and Ker τT2,T1 = {0}.
0 q 1
λ1
0
0 q 2
λ1+1
. . .
···
···
. . .
T1 =
0
0
...
0
···
···
···
···
. . .
. . .
0 q k
k+λ1−1
, T2 =
0 q 1
λ2
0
0 q 2
λ2+1
. . .
···
···
. . .
0
0
...
0
0 q k
k+λ2−1
···
···
···
···
. . .
. . .
.
(cid:3)
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
11
Proposition 3.7. [59] The relation "≺" is a strictly partially ordered relation for the operators in
B1(Ω).
Lemma 3.8. [36] Let Tj, j = 1, 2 be the adjoint of the multiplication operator M (λj )
Then T1 ≺ T2 if and only if λ1 < λ2.
Lemma 3.9. [36] Let Tk = M (λk)∗
A2
λk
, k = 1, 2 be the adjoint of the multiplication operator M (λk)
. If λ2 − λ1 < 2, then (T1, T2) satisfies Property (H).
acting on A2
λj
on
z
z
z
.
Proof. Suppose that there exists an non-zero operator Y such that T1Y = Y T2 and Y = T1X − XT2.
Notice that
By Lemma 4.3 and Lemma 4.3 in [37], we know that lim
k→∞
xk,k+1 = ∞ in the case of λ2 − λ1 < 2.
Theorem 3.10. Suppose that Ω is simply connected. Then there exists T ∈ NCFBn(Ω) such that
(1) T is strongly irreducible;
(2) σ(T ) = Ω, ρF (T ) ∩ σ(T ) = Ω;
(3) ind(w − T ) = n, w ∈ Ω,
where ρF (T ) denotes the Fredholm domain of T and ind(T ) is the Fredholm index of T . In particular,
NCFBn(Ω) 6= ∅.
Proof. First, we assume that Ω = D. Let Tk,k = M (λk)∗
in Lemma 3.8,
and 0 < λk+1 − λk < 2, 1 ≤ k ≤ n − 1. It is well known that Tk,k is a backward shift operator with
, 1 ≤ k ≤ n acting on A2
λk
z
weighted sequence((cid:18)
n
2)∞
n + λk − 1(cid:19) 1
n=1
. We define Tk,k+1, 1 ≤ k ≤ n − 1 to be
Tk,k+1(e(k+1)
n
) =
2
j
nQj=1(cid:18)
j + λk+1 − 1(cid:19) 1
nQj=1(cid:18)
j + λk − 1(cid:19) 1
j
2
e(k)
n ,
where {e(k)
n }∞
n=1 be an orthogonal normal basis of the Hilbert space A2
λk
. Let
Tk,j = Tk,k+1Tk+1,k+2 ··· Tj−1,j, j > k,
and Tk,j = 0, j < k. Then by a directly computation, we can see that
Tk,kTk,k+1 = Tk,k+1Tk+1,k+1, Tk,kTk,j = Tk,jTj+1,j+1, k < j.
12
CHUNLAN JIANG, KUI JI AND JINSONG WU
By Lemma 3.9, we can see that (Tk,k, Tj,j) satisfy the Property (H), for any k < j. Thus, we can see
that T ∈ CFBn(D). Notice that each Tk,k+1 is non-zero. By Lemma 2.8, T is a strongly irreducible
operator.
Notice that Tk,kT ∗
2 , where K (k)
1
and K (k)
2
are both
rank-one operators, we can see that each Tk,k is essentially normal operator. Since
k,k = IHk − K (k)
1 , and T ∗
k,kTk,k = IHk − K (k)
lim
n→∞
2
j
nQj=1(cid:16)
j+λk+1−1(cid:17) 1
j+λk−1(cid:17) 1
nQj=1(cid:16)
j
2
= 0, when λk < λk+1,
then each Tk,k+1 is compact operator. Thus each Tk,j is also compact. We can see that T ∈
NCFBn(D).
For simply connected region Ω ⊂ C, by Riemann mapping theorem, there exists an analytic function
f : Ω → D such that f is one-to-one and f (z0) = 0, f ′(z0) > 0 for some z0 ∈ Ω. By a similar argument,
we can see that when one replace f (T ) instead of T in the previous proceed, the same conclusion is
also valid. In fact, let T ∈ NCFBn(D) defined above. Then we have that
f (T ) =
f (T1,1) f ′(T1,1)T1,2 f1,3(f (T1,1))T1,2T2,3
0
...
0
0
f (T2,2)
...
···
···
f ′(T2,2)T2,3
...
0
···
···
···
f1,n(f (T1,1))T1,2···Tn−1,n
f2,n(f (T2,2))T2,3···Tn−1,n
...
f (Tn−1,n−1)
...
f ′(Tn−1,n−1)Tn−1,n
0
f (Tn,n)
,
where fk,j ∈ H∞(Ω). Since T is essentially normal, that means Q(T )Q(T ∗) = Q(T ∗)Q(T ). It follows
that
Q(f (T ))Q(f ∗(T )) = Q(f ∗(T ))Q(f (T )).
Thus, f (T ) is also essentially normal. Since each Tk,j, k < j is a compact operator, then we have
that fk,j(f (Tk,k)Tk,k+1 ··· Tj−1,j is also a compact operator. Since T ∈ NCFBn(D), then it is easy to
see that the condition (2) in Definition 2.6 is also satisfied. Thus, we only need to prove that f (T )
satisfies the Property (H) when 0 < λk+1 − λk < 2.
Without loss of generality, we only prove that if f (T1,1)Y = Y f (T2,2) and Y = f (T1,1)X − Xf (T2,2),
akzk and a0 = 0. In fact, if a0 6= 0, then we can consider
then Y = 0. Suppose that f (z) =
f (T1,1) − a0I and f (T2,2) − a0I. Thus, we can assume that
∞Pk=0
(3.4)
f (T1,1) =
0 a1,2 a1,3
a2,3
. . .
0
···
···
. . .
0
a1,k
a2,k
...
ak−1,k
0
···
···
···
···
. . .
. . .
, f (T2,2) =
b1,3
b2,3
. . .
···
···
. . .
0
b1,k
b2,k
...
bk−1,k
0
0 b1,2
0
···
···
···
···
. . .
. . .
,
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
13
where ak,k+1 = a1q k
k+λ1−1 and bk,k+1 = a1q k
k+λ2−1 . By Lemma 4.3, we have that
xk+1,k =
bi,i+1
=
ai,i+1
n
kQi=1
k+1Qi=1
i+λ2−1
.
n
a1q i
kQi=1
a1q i
k+1Qi=1
i+λ1−1
By Lemma 3.9 and Lemma 4.3 in [37], we know that lim
k→∞
xk,k+1 = ∞. Thus, Y = 0.
(cid:3)
3.3. (U + K)-Orbits. In [36], the first two authors and Keshari prove that CFBn(Ω) is norm dense
in Bn(Ω). Since it is quite difficult to describe (U + K)-orbit of an arbitrary operator, we would be
interested in obtain a similar result for essentially normal operators. But by the results in [42] and
Theorem 3.10, we see that the set NCFBn(Ω) contains many nontrivial operators.
Definition 3.11. Suppose Ω is simply connected. Let An(Ω) denote the class of operators S, each of
which satisfies that
(1) S is essentially normal;
(2) σ(S) = Ω, ρF (S) ∩ σ(S) = Ω,
(3) ind(S − w) = n, w ∈ Ω, dim Ker (S − w) = 0, w ∈ Ω.
Remark 3.12. By the definition of the Cowen-Douglas operator, it is clear that EN ∩Bn(Ω) ⊂ An(Ω).
Proposition 3.13. [42] Suppose Ω is simply connected. Let S ∈ An(Ω). Then the closure of (U + K)-
orbits of S consists of all operators A satisfying
(1) A is essentially normal;
(2) σ(A) = Ω, ρF (A) ∩ σ(A) = Ω,
(3) ind(w − A) = n, w ∈ Ω,
where ρF (A) denotes the Fredholm domain of A.
Theorem 3.14. Suppose that Ω is simply connected and T is essentially normal. If
(1) σ(T ) = Ω, ρF (T ) ∩ σ(A) = Ω,
(2) ind(w − T ) = n, w ∈ Ω,
then T ∈ NCFBn(Ω).
Proof. It is directly from Theorem 3.10 and Proposition 3.13.
4. Similarity Invariants
(cid:3)
In this section, we will study the similarity invariants of the class NCFBn(Ω) of essentially normal
Cowen-Douglas operators. We begin with some technique lemmas.
Lemma 4.1. Let T = (Tj,k)n
there exists K ∈ K(H) such that X = I + K is invertible and XT = eT X.
Proof. To find K, we need to solve the equation
j,k=1, eT = (eTj,k)n
j,k=1 ∈ NCFBn(Ω). If Tj,j = eTj,j, Tj,j+1 = eTj,j+1, then
(4.1)
(I + K)T = eT (I + K).
14
CHUNLAN JIANG, KUI JI AND JINSONG WU
Set X := I + K, where
From Equation (4.1), we have
K =
0 K1,2 K1,3 ··· K1,n
0 K2,3 ··· K2,n
0
...
...
...
...
...
0 Kn−1,n
0
0
0 ···
0 ···
0
···
.
1 K1,2 K1,3 ··· K1,n
1 K2,3 ··· K2,n
0
...
...
...
...
...
1 Kn−1,n
0
1
0
···
0 ···
0 ···
T1,1 T1,2 eT1,3
0 T2,2 T2,3
...
...
0 Tn−1,n−1 Tn−1,n
···
eT1,n
eT2,n
...
...
...
···
···
···
···
Tn,n
0
0
0
=
···
···
T1,n
T2,n
T1,1 T1,2 T1,3
0 T2,2 T2,3
...
0 Tn−1,n−1 Tn−1,n
···
...
...
···
···
Tn,n
0
0
...
...
0
1 K1,2 K1,3 ··· K1,n
1 K2,3 ··· K2,n
0
...
...
...
...
...
1 Kn−1,n
0
1
0 ···
0 ···
0
···
.
To find Kj,k for 1 ≤ j < k ≤ n, we need the following steps.
Step I. For 1 ≤ j ≤ n − 1, by equating the (j, j + 1) entry of Equation (4.2), we have
Tj,j+1 + Kj,j+1Tj+1,j+1 = Tj,jKj,j+1 + Tj,j+1,
i.e. Kj,j+1Tj+1,j+1 = Tj,jKj,j+1.
For 1 ≤ j ≤ n − 2, by comparing (j, j + 2) entry of Equation (4.2), we have
(4.3)
Tj,j+2 + Kj,j+1Tj+1,j+2 + Kj,j+2Tj+2,j+2 = Tj,jKj,j+2 + Tj,j+1Kj+1,j+2 + eTj,j+2.
We assume that
Kj,j+2Tj+2,j+2 = Tj,jKj,j+2,
1 ≤ j ≤ n − 2.
(4.2)
(4.4)
Then
(4.5)
Tj,j+2 + Kj,j+1Tj+1,j+2 = Tj,j+1Kj+1,j+2 + eTj,j+2.
By the definition of T and eT , we have
(4.6)
Cj,j+2Tj,j+1Tj+1,j+2 + Kj,j+1Tj+1,j+2 = Tj,j+1Kj+1,j+2 + Cj,j+2Tj,j+1Tj+1,j+2.
Let Kj,j+1 = C (1)
Cj,j+2Tj,j+1 + Tj,j+1C (1)
for Kj,j+2 in the next step.
j,j+1Tj,j+1, where C (1)
j+1,j+2. We also assume that Kn−1,n = Tn−1,n, i.e. C (1)
j,j+1 ∈ {Tj,j}′. Then we may assume that Kj,j+1 = Cj,j+2Tj,j+1 −
n−1,n = 1 and we will solve
Step II. By comparing (j, j + 3)-entry of Equation (4.2), we get
(4.7)
We assume that
(4.8)
Tj,j+3 + Kj,j+1Tj+1,j+3 + Kj,j+2Tj+2,j+3 + Kj,j+3Tj+3,j+3
=Tj,jKj,j+3 + Tj,j+1Kj+1,j+3 + eTj,j+2Kj+2,j+3 + eTj,j+3.
Tj,jKj,j+3 = Kj,j+3Tj+3,j+3, 1 ≤ j ≤ n − 3,
then from Equation (4.7) we have
(4.9)
Tj,j+3 + Kj,j+1Tj+1,j+3 + Kj,j+2Ti+2,i+3
=Ti,i+1Ki+1,i+3 + eTi,i+2Ki+2,i+3 + eTj,j+3.
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
15
Assume that Kj,j+2 = C (2)
I., we can solve for Kj,j+2 and assume that Kn−2,n = Tn−2,n−1Tn−1,n.
j,j+2Tj,j+1Tj+1,j+2, where C (2)
j,j+2 ∈ {Tj,j}′. By using the argument in Step
Iterating the arguments in Step II., we can solve for Kj,k, 1 ≤ j < k ≤ n and it is clear that Kj,k
(cid:3)
are compact.
Lemma 4.2. [42] Let T ∈ B1(Ω). For any a0 ∈ Ω, there exists an orthogonal normal basis {en}∞
of H and r > 0 such that T admits the upper-triangular matrix representation
n=0
(4.10)
T =
a0 a1,2 a1,3
a2,3
. . .
a0
···
···
. . .
a0
a1,k
a2,k
...
ak−1,k
a0
···
···
···
···
. . .
. . .
n=0 and ak−1,k > r, for all k = 2, 3,··· .
with respect to {en}∞
Lemma 4.3. Let T1 and T2 ∈ B1(Ω). Suppose T1 and T2 admit the following upper-triangular matrix
representation mentioned in Lemma 4.2,
(4.11)
T1 =
, T2 =
0 a1,2 a1,3
a2,3
. . .
0
···
···
. . .
0
a1,k
a2,k
...
ak−1,k
0
···
···
···
···
. . .
. . .
0 b1,2
0
b1,3
b2,3
. . .
···
···
. . .
0
b1,k
b2,k
...
bk−1,k
0
···
···
···
···
. . .
. . .
.
Suppose X, Y ∈ L(H) and admit the matrix representation X = (xk,j)∞
0, k > j. If T1Y = Y T2 and Y = T1X − XT2 both hold, then
k,j=0, Y = (yk,j)∞
k,j=0, yk,j =
(4.12)
xk+1,k =
bℓ,ℓ+1
.
aℓ,ℓ+1
n
kQℓ=1
k+1Qℓ=1
Proof. Since T1Y = Y T2 and Y = (yk,j), yk,j = 0, k > j. Then we have that yk,k =
T1X − XT2 = Y , then we have that xk,j = 0, k − j > 1, and
kQ
ℓ=1
kQ
ℓ=1
bℓ,ℓ+1
aℓ,ℓ+1
y1,1. Since
a1,2x2,1 = y1,1,
a2,3x3,2 − x2,1b1,2 = y2,2,
··· ,
ak,k+1xk+1,k − xk,k−1bk−1,k = yk,k,··· .
It follows that xk+1,k =
n
kQ
l=1
bl,l+1
k+1Q
l=1
al,l+1
.
Lemma 4.4. Let T ∈ NCFBn(Ω). Then we have that Ker τTj,j,Tℓ,ℓ = {0}, j > ℓ.
(cid:3)
16
CHUNLAN JIANG, KUI JI AND JINSONG WU
Proof. By Lemma 4.2, without loss of generality, we assume that 0 ∈ Ω and there exists {ek}∞
k=1 be
an orthogonal normal basis of H such that Tℓ,ℓ and Tj,j have the following the matrix representations
according to {ek}∞
k=1.
(4.13)
Tℓ,ℓ =
0 aℓ
aℓ
1,k
aℓ
2,k
1,2 aℓ
1,3 ···
0 aℓ
2,3 ···
...
...
0 aℓ
···
···
...
···
k−1,k ···
...
...
0
, Tj,j =
0 aj
1,2 aj
1,3 ···
0 aj
2,3 ···
...
...
0 aj
···
···
aj
1,k
aj
2,k
...
···
k−1,k ···
...
...
0
.
Furthermore, aℓ
k,k+1 > r and aj
k,k+1 > r for some r > 0 and any k > 2.
Since Tℓ,ℓTℓ,j = Tℓ,jTj,j, 1 ≤ ℓ ≤ j − 1, by comparing the elements in the matrix on both sides of
the equation, we have that
t1,1
t1,2
t2,2
···
···
. . .
t1,k
t2,k
...
tk,k
Ti,j =
···
···
···
. . .
. . .
,
aj
l,l+1
aℓ
l,l+1
t1,1. Without loss of generality, assume that t1,1 6= 0. By Theorem 3.4, each
where tk,k =
kQ
l=1
kQ
l=1
operator Ti,j is a compact operator. Thus, we can assume that
lim
k→∞
tk,k = lim
k→∞
aj
l,l+1
aℓ
l,l+1
kQl=1
kQl=1
t1,1 = 0.
Now assume there exists a bounded operator Y such that Tj,jY = Y Tℓ,ℓ, then by a similar computation,
we have that
Y =
y1,1 y1,2
y2,2
···
···
. . .
y1,k
y2,k
...
yk,k
···
···
···
. . .
. . .
,
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
17
aℓ
l,l+1
aj
l,l+1
y1,1. Thus, we have that lim
k→∞
yk,k = ∞. This means y1,1 = 0 and also yk,k =
where yk,k =
kQ
l=1
kQ
l=1
0, 1 ≤ k. Then we have that
0 y1,2
0
···
···
. . .
=
0 aj
1,2 aj
1,3 ···
0 aj
2,3 ···
...
...
0 aj
0 y1,2
0
···
···
. . .
y1,k
y2,k
...
0
0
···
···
aj
1,k
aj
2,k
...
···
k−1,k ···
...
...
···
···
···
. . .
. . .
y1,k
y2,k
...
0
···
···
···
. . .
. . .
···
···
.
0 aℓ
aℓ
1,k
aℓ
2,k
1,2 aℓ
1,3 ···
0 aℓ
2,3 ···
...
...
0 aℓ
...
···
k−1,k ···
...
...
0
By comparing the coefficients of matrices in both sides, we have that
y2,3 = y1,2
aℓ
2,3
aj
1,2
,
y3,4 = y2,3
aℓ
3,4
aj
2,3
,··· ,
yk,k+1 = yk−1,k
aℓ
k,k+1
aj
k−1,k
.
k−1Ql=1
k−1Ql=1
Thus, yk,k+1 =
aℓ
l,l+1
aj
l,l+1
aℓ
1,2 ! y1,2. Since lim
k,k+1
aℓ
k→∞
k−1Q
l=1
k−1Q
l=1
aℓ
l,l+1
aj
l,l+1
= 0 and aℓ
k,k+1 > r > 0. It follows that
y1,2 = 0. Thus, we have that yk,k+1 = 0, for any k > 0.
By an inductive proof, we can assume that yk,k+k′ = 0, k′ < s, k = 1, 2··· . By a similar argument,
we have that
yk,k+s =
aℓ
l,l+1
aj
l,l+1
k−1Ql=1
k−1Ql=1
sQm=1
sQm=1
ai
k,k+m
aℓ
m,m+1
y1,1+s.
Then we also have that y1,1+s = 0 and yk,k+s = 0, k ≥ 1. That means Y = 0. Thus, Ker τTj,j,Tℓ,ℓ =
{0}.
(cid:3)
Proposition 4.5. Let T,eT ∈ NCFBn(Ω). Suppose that X ∈ L(H) is invertible such that XT = eT X.
Then X and X −1 are upper triangular.
Proof. Let X be the invertible operator which intertwines T and T . Set Y = X −1 and X =
(Xi,j)n
i,j=1. Suppose that
i,j=1, Y = (Yi,j)n
(4.14)
X1,1 X1,2 ··· X1,n
X2,1 X2,2 ··· X2,n
...
...
...
...
Xn,1 Xn,2 ··· Xn,n
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
...
Tn,n
=
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
Tn,n
...
X1,1 X1,2 ··· X1,n
X2,1 X2,2 ··· X2,n
...
...
...
...
Xn,1 Xn,2 ··· Xn,n
,
18
(4.15)
CHUNLAN JIANG, KUI JI AND JINSONG WU
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
...
Tn,n
Y1,1 Y1,2 ··· Y1,n
Y2,1 Y2,2 ··· Y2,n
...
...
...
...
Yn,1 Yn,2 ··· Yn,n
=
Y1,1 Y1,2 ··· Y1,n
Y2,1 Y2,2 ··· Y2,n
...
...
...
...
Yn,1 Yn,2 ··· Yn,n
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
Tn,n
...
.
By Equations (4.14) and (4.15), we have that
Xn,1T1,1 = Tn,nXn,1, Tn,nYn,1 = Yn,1 T1,1.
Then we have that Xn,1T1,nYn,1 ∈ Ker τ Tn,n, T1,1
= {0}. Since T1,1T1,n = T1,nTn,n, then we know that
T1,n have dense range. Furthermore, since Xn,1 and Yn,1 intertwine two operators in B1(Ω), we know
Xn,1 and Yn,1 are zero operators or the operators with dense range. Thus, we can see that Xn,1 = 0
or Yn,1 = 0. Now assume that Xn,1 = 0, Yn,1 6= 0, then we have that Ker τTn,n, T1,1 6= {0}. Since
= {0}, by a similar
Ker τT2,2,Tn,n 6= {0}, Ker τTn,n,T2,2 = {0} and Ker τ T1,1, Tn,n 6= {0}, Ker τ Tn,n, T1,1
argument, we have that
(4.16)
Ker τ Tn,n,T2,2
= {0}.
Since Xn,1 = 0, then we have that
(4.17)
X1,1 X1,2 ··· X1,n
X2,1 X2,2 ··· X2,n
...
...
...
...
0 Xn,2 ··· Xn,n
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
...
Tn,n
=
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
Tn,n
...
X1,1 X1,2 ··· X1,n
X2,1 X2,2 ··· X2,n
...
...
...
...
0 Xn,2 ··· Xn,n
.
Thus, we have that Xn,2T2,2 = Tn,nXn,2. By Equation (4.16), Xn,2 = 0. Going to the next step, we
will also have Xn,i = 0, 1 ≤ i ≤ n. However, this is impossible since X is invertible. Thus, we will
have Yn,1 = 0. On the other hand, we can also prove that the assumption Xn,1 6= 0, Yn,1 = 0 does not
hold. By a direction calculation,
Thus, we have
Xn,1 = Yn,1 = 0.
and
X1,1 X1,2 ··· X1,n
X2,1 X2,2 ··· X2,n
...
...
...
...
0 Xn,2 ··· Xn,n
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
...
Tn,n
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
...
Tn,n
Y1,1 Y1,2 ··· Y1,n
Y2,1 Y2,2 ··· Y2,n
...
...
...
...
0 Yn,2 ··· Yn,n
=
=
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
Tn,n
...
Y1,1 Y1,2 ··· Y1,n
Y2,1 Y2,2 ··· Y2,n
...
...
...
...
0 Yn,2 ··· Yn,n
X1,1 X1,2 ··· X1,n
X2,1 X2,2 ··· X2,n
...
...
...
...
0 Xn,2 ··· Xn,n
T1,1 T1,2 ··· T1,n
T2,2 ··· T2,n
...
Tn,n
...
,
.
Thus, we have that Xn,2T2,2 = Tn,nXn,2, Yn,2 T2,2 = Tn,nYn,2. By a proof similar to the proof mentioned
above, we can also see that the statements Xn,2 = 0, Yn,2 6= 0 and Xn,2 6= 0, Yn,2 = 0 both can not
hold. Thus, we also have that Xn,2 = Yn,2 = 0. Keep doing this, we will have that
Xn,1 = Xn,2 = ··· = Xn,n−1 = Yn,1 = Yn,2 = ··· = Yn,n−1 = 0.
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
19
...
and
...
...
...
...
Y1,2
Y2,2
...
...
...
...
Tn−1,n−1
Tn−1,n−1
X1,1
X2,1
X1,2
X2,2
T1,n−1
T2,n−1
T1,n−1
T2,n−1
X1,1 X1,2
X2,1 X2,2
Xn−1,1 Xn−1,2 ··· Xn−1,n−1
··· X1,n−1
··· X2,n−1
...
··· X1,n−1
··· X2,n−1
...
T1,1 T1,2 ···
T2,2 ···
...
T1,1 T1,2 ···
T2,2 ···
...
Thus, Xn,nTn,n = Tn,nXn,n, Tn,nYn,n = Yn,n Tn,n. Since Xn,n and Yn,n are both non-zero operators,
we know that Tn,n ∼ Tn,n. Thus, we have that
Xn−1,1 Xn−1,2 ··· Xn−1,n−1
T1,1 T1,2 ···
T2,2 ···
...
T1,n−1
T2,n−1
=
=
,
.
Lemma 4.6. Let T,eT ∈ NCFBn(Ω). Suppose that T ∼U +K eT and T = (Tj,k)n
j,k=1, eT = (eTj,k)n
eT (U + K) = (U + K)T.
Proof. Since T ∼U +K eT , there exists an invertible operator X ∈ L(H) such that XT = eT X and X
Repeating this inductive proof, we can show that Xk,j = Yk,j = 0, whenever k > j. This finishes the
(cid:3)
proof of this lemma.
j,k=1.
Then there exists a diagonal unitary operator U and a compact upper triangular operator K such that
is a sum of a unitary operator and a compact operator. Then Q(X) is unitary in the Calkin algebra
A(H). By Proposition 4.5, X is upper-triangular operator, i.e.
Yn−1,1 Yn−1,2 ··· Yn−1,n−1
T1,n−1
T2,n−1
T1,1 T1,2 ···
T2,2 ···
...
Yn−1,1 Yn−1,2 ··· Yn−1,n−1
Y1,n−1
Y2,n−1
Y1,n−1
Y2,n−1
···
···
...
···
···
...
Y1,1
Y2,1
...
Y1,2
Y2,2
...
Y1,1
Y2,1
...
Tn−1,n−1
...
...
...
...
Tn−1,n−1
X =
···
···
X1,1 X1,2 X1,3
0 X2,2 X2,3
...
0 Xn−1,n−1 Xn−1,n
···
X1,n
X2,n
...
...
Xn,n
···
···
0
0
...
...
0
.
It follows that Q(X1,n)Q(X ∗
n,n) is invertible, then we have Q(X1,n) = 0. Thus,
we will have that Q(X1,n−1)Q(Xn−1,n−1) = 0, and then Q(X1,n−1) = 0. Repeating this arguments,
we will have that Q(Xi,j) = 0, i < j. That means Q(X) is a diagonal matrix and Q(Xi,i)Q(Ti,i) =
n,n) = 0. Since Q(X ∗
Q(eTj,j)Q(Xj,j eT ). Let Uj be the polar part of the polar decomposition of Xj,j. Then
Q(Xj,j) = Q (UjXj,j) = Q(Uj)Q(X ∗
j,jXj,j)1/2 = Q(Uj),
Let U be a diagonal matrix whose (j, j)-entry is Uj and K = X − U . Then U is unitary and K is
compact and the proof is complete.
(cid:3)
It is then clear that Xj,j intertwines Tj,j and eTj,j and Xj,j are invertible. Notice that each Q(Xj,j) is
invertible and Q(X) is unitary in A(H), we have that Q(X)Q(X ∗) = Q(X ∗)Q(X) = I, i.e.
Q(X1,1) Q(X1,2) ··· Q(X1,n)
Q(X2,2) ··· Q(X2,n)
...
...
Q(Xn,n)
Q(X ∗
Q(X ∗
2,2)
1,1)
1,2) Q(X ∗
...
...
2,n) ··· Q(X ∗
1,n) Q(X ∗
...
Q(X ∗
n,n)
= I.
20
CHUNLAN JIANG, KUI JI AND JINSONG WU
Lemma 4.7. Let T,eT ∈ CFBn(Ω). Suppose X ∈ L(H) is invertible and XeT = T X. We assume that
j,k=1 are upper triangular. Then
eT1,1 eT1,2
0
...
0
eT2,2 eT2,3
...
...
0
0
···
···
0
···
···
···
0
0
...
...
0
eTn−1,n−1 Tn−1,n
eTn,n
0
0
···
···
X1,1
0
0
0 X2,2 0
...
0 Xn−1,n−1
···
...
...
···
···
0
0
...
0
...
0
Xn,n
.
0
0
0
0
0
...
...
...
...
0
0
0
0
0
0
0
0
···
···
···
···
···
···
···
···
Tn,n
X1,1
Xn,n
...
...
...
...
T1,1 T1,2
T = (Tj,k)n
j,k=1, and X = (Xj,k)n
0
0 X2,2 0
...
0 Xn−1,n−1
···
0 T2,2 T2,3
...
0 Tn−1,n−1 Tn−1,n
···
j,k=1, eT = (eTj,k)n
=
Proof. By equating the entries of XeT = T X, we get
Xj,jeTj,j = Tj,jXj,j,
Y =
From T = XeT X −1, we get
Let Y := XDiag (X −1). Then
2,2 X1,3X −1
X2,3X −1
I X1,2X −1
0
...
...
···
···
0
···
0
0
...
I
1 ≤ j ≤ n.
3,3 ··· X1,nX −1
3,3 ··· X2,nX −1
n,n
n,n
...
...
I Xn−1,nX −1
n,n
0
I
.
i.e.
(4.19)
Tj,j+1 − Xj,jeTj,j+1X −1
j+1,j+1 = Xj,j+1X −1
T Y = T XDiag (X −1) = XeT Diag (X −1) = Y Diag (X)eT Diag (X −1),
T1,1 T1,2 T1,3
0 T2,2 T2,3
...
0 Tn−1,n−1 Tn−1,n
···
3,3 ··· X1,nX −1
3,3 ··· X2,nX −1
2,2 X1,3X −1
X2,3X −1
I X1,2X −1
0
T1,n
T2,n
...
...
...
Tn,n
···
···
···
···
0
0
0
0
n,n
n,n
...
...
...
...
0
I
I
I X1,2X −1
0
2,2 X1,3X −1
X2,3X −1
3,3 ··· X1,nX −1
3,3 ··· X2,nX −1
I
n,n
n,n
...
0
0
...
···
···
...
0
···
...
...
I Xn−1,nX −1
n,n
0
I
which is equivalent to
(4.18)
=
Comparing the (j, j + 1)-entries of Equation (4.18), we obtain that
Tj,j+1 + Tj,jXj,j+1X −1
j+1,j+1 = Xj,j+1X −1
j+1,j+1.
··· X1,1 eT1,nX −1
3,3 ··· X2,2 eT2,nX −1
n,n
n,n
...
Tn,n
.
0
0
...
2,2
···
···
0
···
...
...
...
T2,2
...
X2,2 eT2,3X −1
...
I Xn−1,nX −1
···
n,n
···
0
T1,1 X1,1 eT1,2X −1
0
j+1,j+1Tj+1,j+1 + Xj,jeTj,j+1X −1
j+1,j+1Tj+1,j+1 − Tj,jXj,j+1X −1
0
j+1,j+1,
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
21
denoted by Yj,j+1. Then we consider
j+1,j+1)
Tj,jYj,j+1 = Tj,j(Tj,j+1 − Xj,jeTj,j+1X −1
=Tj,j+1Tj+1,j+1 − Xj,jeTj,jeTj,j+1X −1
=Tj,j+1Tj+1,j+1 − Xj,jeTj,j+1eTj+1,j+1X −1
=Tj,j+1Tj+1,j+1 − Xj,jeTj,j+1X −1
=(Tj,j+1 − Xj,jeTj,j+1X −1
j+1,j+1)Tj+1,j+1
=Yj,j+1Tj+1,j+1,
j+1,j+1
j+1,j+1
j+1,j+1Tj+1,j+1
i.e. Yj,j+1 ∈ Ker τTj,j,Tj+1,j+1 ∩ Ran τTj,j ,Tj+1,j+1. Recall that T satisfies the Property (H), we see
Yj,j+1 = 0, i.e.
Finally, we have the following matrix equation:
Tj,j+1 = Xj,jeTj,j+1X −1
j+1,j+1,
=
X1,1
0
···
···
0
0 X2,2 0
...
0 Xn−1,n−1
···
...
...
···
···
0
0
...
0
0
0
...
0
T1,1 T1,2
0
···
···
...
...
0 T2,2 T2,3
...
0 Tn−1,n−1
···
...
···
···
0
0
0
1 ≤ j ≤ n − 1.
eT1,1 eT1,2
0
...
0
eT2,2 eT2,3
...
...
0
0
···
···
0
···
···
···
...
0
eTn−1,n−1
X1,1
0
···
···
0
0 X2,2 0
...
0 Xn−1,n−1
···
...
...
···
···
0
0
...
0
Xn,n
0
0
...
0
Tn,n
0
0
...
0
eTn,n
0
0
...
0
Xn,n
.
(cid:3)
Theorem 4.8. Let T,eT ∈ NCFBn(Ω). Then T ∼U +K eT if and only if the following statements hold:
(w)) ⊕ A(w)], for any w ∈ Ω, λ ∈ C;
2π KEdiag T
(1) Cλ(EdiagT )(w) = det[(I + iλ
(2)
φj
φj+1
θj,j+1(T ) = θj,j+1(eT ), for any 1 ≤ j ≤ n − 1,
nLj=1
∂w∂w
∂2
where A(w) =
iλ
2π ln(φj(w)),and φj(w) are functions appeared in Proposition 2.9.
Remark 4.9. In the main theorem above, when A(w) = 0, w ∈ Ω, then the equation in statement
(1) turns out to be
Cλ(EdiagT )(w) = Cλ(Ediag T )(w).
In the following, we give the proof of Theorem 4.8.
In this case, T ∼u T . (See in Theorem 5.1)
Proof. We assume that 0 ∈ Ω. Suppose that Cλ(EdiagT ) = deth(cid:16)I + iλ
w ∈ Ω, λ ∈ C. By the definition, we obtain that
2π KEdiag T
(w)(cid:17) ⊕ A(w)i for any
i
2π
nYj=1(cid:18)1 + λ
KTj,j (w)(cid:19) =
where tj is a nonzero section of bundle E eTj,j
j ) (See in Proposition
2.9). By the fundamental theorem of algebra, there exists a permutation Ξ on {1, 2,··· , n} such that
nYj=1(cid:18)1 − λ
, and φj(w) = (cid:13)(cid:13)Xj(tj(w))(cid:13)(cid:13)2
ln(φj(w)ktj (w)k2)(cid:19) ,
+ (1− α2
ktj(w)k2
∂w∂w
i
2π
∂2
22
CHUNLAN JIANG, KUI JI AND JINSONG WU
for any 1 ≤ j ≤ n, we have
Note that
KTj,j (w) = −
∂2
∂w∂w
ln(cid:0)φΞ(j)(w)ktΞ(j)(w)k2(cid:1) .
KTj,j = −
= −
= −
∂2
∂w∂w
∂2
∂w∂w
∂2
∂w∂w
ln(cid:0)φΞ(j)(w)ktΞ(j)(w)k2(cid:1)
∂2
ln φΞ(j)(w) −
∂w∂w
ln(ktΞ(j)(w)k2)
ln φΞ(j)(w) + K eTΞ(j),Ξ(j)
,
∂2
i.e. K eTΞ(j),Ξ(j) − KTj,j =
exist invertible operators Yj = Uj + Kj such that Uj are unitary, Kj are compact and
ln φΞ(j)(w). By Lemma 2.9, we have that Tj ∼U +K eTΞ(j),Ξ(j), i.e. there
∂w∂w
(Uj + Kj)eTΞ(j),Ξ(j) = Tj(Uj + Kj),
1 ≤ j ≤ n.
Now we will prove that Ξ(j) = j, 1 ≤ j ≤ n. We assume that Ξ(n) < n. Since (Un +Kn)eTΞ(n),Ξ(n) =
Tn,n(Un + Kn), then we have that
(Un + Kn)eTΞ(n),Ξ(n)eTΞ(n),n = Tn,n(Un + Kn)eTΞ(n),n,
(Un + Kn)eTΞ(n),neTn,n = Tn,n(Un + Kn)eTΞ(n),n.
(UΞ−1(n) + KΞ−1(n))eTn,n = TΞ−1(n),Ξ−1(n)(UΞ−1(n) + KΞ−1(n)),
and Ξ−1(n) < n (Recall Ξ(n) < n). Then we have that
Notice that
(Un + Kn)eTΞ(n),neTn,n(UΞ−1(n) + KΞ−1(n))−1 = Tn,n(Un + Kn)eTΞ(n),n(UΞ−1(n) + KΞ−1(n))−1.
It follows that
(Un + Kn)eTΞ(n),n(UΞ−1(n) + KΞ−1(n))−1TΞ−1(n),Ξ−1(n) = Tn,n(Un + Kn)eTΞ(n),n(UΞ−1(n) + KΞ−1(n))−1.
Thus,
Since Ξ−1(n) < n, by Lemma 4.4, we know that Ker τTn,n,TΞ−1(n),Ξ−1(n)
(Un + Kn)eTΞ(n),n(UΞ−1(n) + KΞ−1(n))−1 ∈ Ker τTn,n,TΞ−1(n),Ξ−1(n)
.
= {0}. Thus we have that
eTΞ(n),n = 0. This is a contradiction and Ξ(n) = n. Iterating the process, we can obtain that Ξ(j) = j
for 1 ≤ j ≤ n. Moreover
Let T = Y −1T Y , where Y is a diagonal matrix whose (j, j) entry is Yj = Uj + Kj. The operator
T has the following form:
j = 1, 2, . . . , n.
Tj,j = (Uj + Kj)eTj,j(Uj + Kj)−1,
T =
eT1,1 Y −1
0
...
Y −1
2 T2,3Y3
eT2,2
...
1 T1,2Y2
...
···
···
...
···
···
0
0
0
0
···
0
Y −1
1 T1,nYn
Y −1
2 T2,nYn
...
eTn,n
.
eTn−1,n−1 Y −1
n−1Tn−1,nYn
It is clear that T ∼U +K T .
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
23
We define operators A, B ∈ L(H) by using matrices:
A :=
and
It is easy to see that
0
0
0
0
B :=
0
0
...
eTn,n
.
1 T1,2Y2
eT1,1 Y −1
0
...
eT2,2
...
0
Y −1
2 T2,3Y3
···
0
...
···
···
...
0
eTn−1,n−1 Y −1
n−1Tn−1,nYn
,
eT1,1 eT1,2
0
...
0
0
···
0
0
eT2,2 eT2,3
...
...
··· eTn−1,n−1 eTn−1,n
eTn,n
···
...
0
0
...
0
0
0
θj,j+1(T ) = θj,j+1(A),
θj,j+1(eT ) = θj,j+1(B),
1 ≤ j ≤ n − 1.
We claim that A is unitarily equivalent to B. In fact, by Theorem 3.6 in [34], we only need to
j+1(ti+1(·)) is a non zero section
prove the second fundamental forms of A and B are same. Clearly Y −1
of E eTj+1,j+1
,
θj,j+1(A)(w) = kY −1
j Tj,j+1Yj+1Y −1
j+1(ti+1(w))k2
= kY −1
kY −1
j+1(ti+1(w))k2
j Ti,i+1(ti+1(w))k2
kY −1
j+1(ti+1(w))k2
j+1(tj+1(w))k2
.
j+1(tj+1(w))k2
θj,j+1(B)(w) = keTj,j+1Y −1
kY −1
φj(w) = kY −1
(tj(w))k2
,
j
ktj(w)k2
1 ≤ j ≤ n,
and
Since
and the condition (2) is satisfied, we then have
θj,j+1(B) = θj,j+1(eT )
= kY −1
φj(w)
φj+1(w)
=
j
θj,j+1(T )
(Tj,j+1(tj+1(w)))k2
= kY −1
kTj,j+1(ti+1(w))k2
j Tj,j+1(tj+1(w))k2
kY −1
= θj,j+1(A).
j+1(tj+1(w))k2
ktj+1(w)k2
j+1(tj+1(w))k2
kY −1
kTj,j+1(tj+1(w))k2
ktj+1(w)k2
24
CHUNLAN JIANG, KUI JI AND JINSONG WU
Hence there exists a diagonal unitary operator V whose (j, j)-entry is denoted by Vj such that V ∗BV =
A. Consider
V Y −1T Y V ∗ = V T V ∗
=
0
...
0
0
0
...
0
0
0
eT1,1 eT1,2 V1 Y
0
−1
1
eT2,2
0
−1
∗
3 V1 Y
1
∗
T1,4 Y4 V
4
∗
V2X2T2,4 X4V
4
T1,3 Y3 V
eT2,3
eT3,3
...
···
0
0
eT3,4
...
eTn−2,n−2
···
···
T1,n Yn V
−1
V1 Y
1
−1
V2 Y
2 T2,n Yn V
−1
T3,n Yn V
V3 Y
3
∗
n
∗
n
∗
n
...
eTn−1,n
eTn,n
.
eTn−2,n−1 Vn−2 Y
eTn−1,n−1
−1
n−2
Tn−2,n Yn V
∗
n
···
···
···
...
0
By Lemma 4.1, there exists eK ∈ K(H) such that I + eK is invertible and
(I + eK)V Y −1T Y V ∗(I + eK)−1 = eT .
Thus T ∼U +K eT .
On the other hand, suppose that T ∼U +K eT . Then there is a unitary operator U and compact
operator K such that T (U + K) = (U + K)eT . By Proposition 4.5, U + K is upper triangular. By
Lemma 4.6, we can assume that
U + K =
U1,1+K1,1
···
···
K1,2
K1,3
U2,2+K2,2 K2,3
...
0 Un−1,n−1+Kn−1,n−1 Kn−1,n
···
K1,n
K2,n
...
...
···
···
...
0
Un,n+Kn,n
0
...
0
0
,
where U = Diag {U1,1, U2,2,··· , Un,n}, K = (Kℓ,j)n×n, Kℓ,j = 0, ℓ > j.
By Lemma 4.7, we have
0
0
0
...
...
0
0
0
0
···
···
···
···
Tn,n
...
...
T1,1 T1,2
U1,1+K1,1
0 T2,2 T2,3
...
0 Tn−1,n−1 Tn−1,n
···
=
Tj,j(Uj,j + Kj,j) = (Uj,j + Kj,j)eTj,j,
...
···
···
0
0
...
U1,1+K1,1
0
0
0
...
...
0
0
0
0
0
0
···
···
···
···
...
...
Un,n+Kn,n
0
eT1,1 eT1,2
0
...
0
U2,2+K2,2 0
...
0 Un−1,n−1+Kn−1,n−1
···
(Uj,j + Kj,j)Tj,j+1 = eTj,j+1(Uj+1,j+1 + Kj+1,j+1).
eTn−1,n−1 eTn−1,n
eTn,n
0
eT2,2 eT2,3
...
...
Un,n+Kn,n
...
···
···
···
···
0
···
0
0
0
0
...
...
0
0
0
···
···
0
U2,2+K2,2 0
...
0 Un−1,n−1+Kn−1,n−1
···
...
0
.
It follows that
Furthermore, Uj,j + Kj,j is invertible and Uj,j is a unitary operator and Kj,j is a compact operator.
Let
φj(w) := k(Uj,j + Kj,j)−1(tj(w))k2
ktj(w)k2
,
1 ≤ j ≤ n.
Following the same argument as in the sufficient part and from Proposition 2.9, we can conclude
that
KTj,j − K eTj,j
=
∂2
∂w∂w
ln(φj).
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
25
Thus, we have that
Cλ(EDiag T ) =det(cid:18)I − i
=det(cid:20)(cid:18)I +
λ
2π
iλ
2π
Φ(cid:19)
KEdiag T
(w)(cid:19) ⊕ A(w)(cid:21) ,
where Φ(w) =
∂2
∂w∂w
nMj=1
θj,j+1(T ) = k(Uj + Kj)Tj,j+1(tj+1)k2
ln(cid:0)φj(w)ktj (w)k2)(cid:1), tj is a non zero section of bundle E eTj,j
k(Uj+1 + Kj+1)(tj+1)k2 = keTj,j+1(Uj+1 + Kj+1)(tj+1)k2
k(Uj+1 + Kj+1)(tj+1)k2
φj
φj+1
. Furthermore,
= θj,j+1(eT ).
(cid:3)
This finishes the proof of the necessary part.
5. Unitary Invariants
In this section, we completely determined the unitary invariants of this class NCFBn(Ω) of essen-
tially normal Cowen-Douglas operators.
Theorem 5.1. Let T, T ∈ NCFBn(Ω). Suppose that Tℓ,j = φℓ,j(Tℓ,i)Tℓ,ℓ+1Tℓ+1,ℓ+2 ··· Tj−1,j and
Tℓ,j = φℓ,j( Tℓ,ℓ) Tℓ,ℓ+1 Tℓ+1,ℓ+2 ··· Tj−1,j Then T ∼u eT if and only if the following statements hold:
(1) Cλ(EDiag T ) = Cλ(EDiag T ) ;
(2) θℓ,j(T ) = θℓ,j( T ),
(3) φℓ,j = φℓ,j, 1 ≤ ℓ ≤ j ≤ n.
Proof. Suppose T ∼u T in NCFBn(Ω), i.e. there is a unitary operator U such that U T = T U . By
Theorem 3.5 in [34], such an intertwining unitary must be diagonal, that is, U = U1 ⊕ ··· ⊕ Un, for
some choice of n unitary operators U1, . . . , Un.
Since UℓTℓ = TℓUℓ, 1 ≤ ℓ ≤ n, and UℓTℓ,j = Tℓ,jUj, 1 ≤ ℓ < j ≤ n. In particular, UℓTℓ,ℓ+1 =
Tℓ,ℓ+1Uℓ+1, 1 ≤ ℓ ≤ n− 1. Suppose that there exists a holomorphic function φℓ such that Uℓ(tℓ) = φℓtℓ.
Then we have that
UℓTℓ,ℓ+1(tℓ+1) =Uℓ(tℓ) = φℓtℓ
= Tℓ,ℓ+1Uℓ+1(tℓ+1)
=φℓ+1tℓ.
Thus, we have φℓ = φℓ+1. We then set φ := φℓ, ℓ = 1, 2··· , n. That means
(5.1)
Uℓ(tℓ(w)) = φ(w)tℓ(w), 1 ≤ ℓ ≤ n, w ∈ Ω,
where φ is some non zero holomorphic function. Thus
or
KTℓ = K Tℓ
and
ktℓ+1(w)k
ktℓ+1(w)k
= ktℓ(w)k
ktℓ(w)k
,
1 ≤ ℓ ≤ n − 1.
KTℓ = K Tℓ
and
θℓ−1,ℓ(T ) = θℓ−1,ℓ( T ) 1 ≤ ℓ ≤ n.
It follows that Cλ(EDiag T ) = Cλ(EDiag T ).
26
CHUNLAN JIANG, KUI JI AND JINSONG WU
On the other hand, we have that Tℓ,j(tj) = φℓ,jtℓ and Tℓ,j(tj) = φℓ,j tℓ. By the equations UℓTℓ,j =
Tℓ,jUj, 1 ≤ ℓ < j ≤ n − 1, we have that
Since ktℓ+1k
ktℓ+1k
= ktℓk
ktℓk
, 1 ≤ ℓ ≤ n − 1, we also have that
φℓ,j = φℓ,j,
1 ≤ ℓ < j ≤ n
ktjk
ktjk
= ktℓk
ktℓk
, 1 ≤ ℓ < j ≤ n.
Since φℓ,j = φℓ,j, then it follows that
θℓ,j(T ) = kTℓ,j(tj)k
ktjk
= kφℓ,jtℓk
ktjk
= k φℓ,jtℓk
ktjk
= k Tℓ,j(tj)k
ktjk
= θℓ,j( T ), 1 ≤ ℓ < j ≤ n.
This finishes the proof of the necessary part.
Conversely assume that T and T are operators in NCFBn(Ω) for which these invariants are the
same. Since Cλ(EDiag T ) = Cλ(EDiag T ), then we have that
nYk=1(cid:18)1 + λ
i
2π
KTk,k(cid:19) =
nYk=1(cid:18)1 + λ
i
2π
K Tk,k(cid:19) .
Thus, there exists a permutation Ξ on {1, 2,··· , n} such that
KTℓ,ℓ = K TΞ(ℓ),Ξ(ℓ)
,
1 ≤ ℓ ≤ n.
That means there exist unitary operators Uℓ such that
Uℓ TΞ(ℓ),Ξ(ℓ) = TℓUℓ,
1 ≤ ℓ ≤ n.
Now we will prove that Ξ(ℓ) = ℓ, 1 ≤ ℓ ≤ n. Otherwise, without loss of generality, we can assume
that Ξ(n) < n. In fact, if Ξ(n) = n, we can assume Ξ(n − 1) < n − 1. Since Un TΞ(n),Ξ(n) = Tn,nUn,
then we have that
Notice that
Un TΞ(n),Ξ(n)
TΞ(n),n =TnUn TΞ(n),n
Un TΞ(n),n
Tn,n =Tn,nUn TΞ(n),n.
UΞ−1(n)
Tn,n = TΞ−1(n),Ξ−1(n)UΞ−1(n),
and Ξ−1(n) < n (Recall Ξ(n) < n). Then we have that
Un TΞ(n),n
It follows that
Tn,nU ∗
Ξ−1(n) = Tn,nUn TΞ(n),nU ∗
Ξ−1(n).
Un TΞ(n),nU ∗
Ξ−1(n)TΞ−1(n),Ξ−1(n) = Tn,nUn TΞ(n),nU ∗
Ξ−1(n).
Thus, Un TΞ(n),nU ∗
Ker τTn,n,TΞ−1(n),Ξ−1(n)
that Ξ(ℓ) = ℓ, 1 ≤ ℓ ≤ n, and
Ξ−1(n) ∈ Ker τTn,n,TΞ−1(n),Ξ−1(n)
. Since Ξ−1(n) < n, by Lemma 4.4, we know that
= {0}. Thus we have that TΞ(n),n = 0. This is a contradiction. So we have
Tℓ,ℓ = Uℓ Tℓ,ℓU ∗
ℓ , KTℓ,ℓ = K Tℓ,ℓ
ℓ = 1, 2, . . . , n.
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
27
Equality of the two curvatures KTℓ = K Tℓ
together with the equality of the second fundamental
ktℓk , 1 ≤ ℓ ≤ n − 1 implies that there exist a non-zero holomorphic function φ defined
forms ktℓ+1k
on Ω (if necessary, one may choose a domain Ω0 ⊆ Ω such that φ is non zero on Ω0) such that
ktℓ+1k = ktℓk
ktℓ(w)k = φ(w)ktℓ(w)k,
For 1 ≤ ℓ ≤ n, we define Uℓ : Hℓ → eHℓ by the formula
Uℓ(tℓ(w)) = φ(w)tℓ(w), w ∈ Ω.
1 ≤ ℓ ≤ n.
and extend to the linear span of these vectors. For 1 ≤ ℓ ≤ n,
kUℓ(ti(w))k = kφ(w)tℓ(w)k
= φ(w)ktℓ(w)k
= ktℓ(w)k.
Thus each Uℓ extends to an isometry from Hℓ to eHℓ. Since Uℓ is isometric and UℓTℓ = TℓUℓ. Let Ω0
be arbitrary open connected subset of Ω. Since Ww∈Ω0{tℓ(w)} = Hℓ, then we can see each Uℓ is unitary.
By a direct computation, we can see that UℓTℓ,ℓ+1 = Tℓ,ℓ+1Uℓ+1 for 1 ≤ ℓ ≤ n − 1 also. For any
w ∈ Ω and 0 ≤ ℓ < j ≤ n − 2 with j − ℓ ≥ 2, we have
Uℓ(Tℓ,j(tj(w))) = Uℓ(φℓ,j(w)tℓ(w)) = φℓ,j(w)φ(w)tℓ(w) = φℓ,j(w)φ(w)tℓ(w) = Tℓ,j(Uj(tj(w)))
which implies that
UℓTℓ,j = Tℓ,jUj.
Hence setting U = U1 ⊕ ··· ⊕ Un, we see that U is unitary and U T = T U . This completes the proof
of the Theorem.
(cid:3)
6. Summary and Further Research
In Section 3, it is critical to find an operator in NCFBn(Ω) (Theorem 3.10) to show that NCFBn(Ω)
contains many nontrivial operators when Ω is simply connected. It is natural to ask the following
questions:
1. Can we construct an operator in NCFBn(Ω) when Ω is connected?
2. Can we generalize Proposition 3.13 to the case Ω is (2-)connected?
In 1984, Misra [50] introduced homogeneous operators. An operator T ∈ L(H) is homogeneous
if σ(T ) ⊂ D and φα(T ) is unitarily equivalent to T for all M´'obius transformation φα; Misra also
introduce homogeneous for similarity. An operator T ∈ L(H) is weakly homogeneous if σ(T ) ⊂ D and
φα(T ) is similar to T for all M´'obius transformation φα; In the framework of Herrero, we introduce
(U + K)-homogeneous between homogeneous and weakly homogeneous.
Definition 6.1. An operator T ∈ L(H) with kTk ≤ 1 is (U + K)-homogenous if there exists a unitary
operator Uα and a compact operator Kα such that Uα + Kα is invertible and
for all Mobius transformation φα of the unit disk.
(Uα + Kα)T = φα(T )(Uα + Kα), α ∈ D.
In [50], Misra showed that an operator T ∈ B1(Ω) with kTk ≤ 1 is homogeneous if and only if its
curvature is −λ(1 − w2)−2 for some positive number λ.
28
CHUNLAN JIANG, KUI JI AND JINSONG WU
Question 6.2. Can we give a simple characterization of (U + K)-homogeneous operators?
References
[1] D. Alpay, A remark on the Cowen-Douglas classes Bn(Ω). Arch. Math. (Basel) 51 (1988), no. 6, 539-546.
[2] C. Apostol, L. A. Fialkow, D. A. Herrero and D. Voiculescu, Approximation of Hilbert Space Operators II, Research
Notes in Math. Vol. 102, Pitman Book Ltd., 1984.
[3] C. Apostol, D. A. Herrero and D. Voiculescu, The closure of the similarity orbit of a Hilbert space operator Bull.
Amer. Math. Soc. (New Series), 6 (1982), 421-426.
[4] L. G. Brown, R. G. Douglas, P. A. Fillmore, Extensions of C∗-algebras and K-homology. Ann. of Math. (2) 105
(1977), no. 2, 265-324.
[5] I. D. Berg and K. R. Davidson Almost commuting matrices and a quantitative version of the Brown-Douglas-Fillmore
theorem. Acta Math. 166 (1991), 121-161.
[6] C. Apostol, M. Martin, C ∗-algebra approach to the Cowen-Douglas theory. Topics in modern operator theory
(Timioara/Herculane, 1980), pp. 45-51, Operator Theory: Adv. Appl., 2, Birkhauser, Basel-Boston, Mass., 1981.
[7] M. Carlsson, On the Cowen-Douglas class for Banach space operators. Integral Equations Operator Theory 61
(2008), no. 4, 593-598.
[8] J.B. Cowen, The theory of subnormal operators. Mathematical Surveys and Monographs, 36. American Mathematical
Society, Providence, RI, 1991.
[9] A. Connes, Noncommutative differential geometry, Publ. Math. Inst. Hautes ´Etudes Sci. 62(1985), 257-360.
[10] M. J. Cowen and R. G. Douglas, Complex geometry and operator theory, Acta Math. 141 (1978), 187 - 261.
[11] L. Chen and R. G. Douglas, A local theory for operator tuples in the Cowen-Douglas class. Adv. Math. 307 (2017),
754-779.
[12] D. N. Clark and G. Misra, On curvature and similarity, Michigan Math. J., 30 (1983), 361 - 367.
[13] D. N. Clark and G. Misra, On weighted shifts, curvature and similarity, J. London Math. Soc., 31(1985), 357 - 368.
[14] D. N. Clark and G. Misra, The curvature function and similarity of operators, International symposium on complex
analysis and applications (Arandjelovac, 1984) Mat. Vesnik 37(1985), no. 1, 21-32.
[15] R. Curto and N. Salinas, Generalized Bergman kernels and the Cowen-Douglas theory, Amer. J. Math. 106 (1984),
447 - 488.
[16] R. G. Douglas, H. Kwon and S. Treil, Similarity of n-hypercontractions and backward Bergman shifts., J. Lond.
Math. Soc. (2) 88 (2013), no. 3, 637-648.
[17] R. G. Douglas and G. Misra, Equivalence of quotient Hilbert modules. II, Trans. Amer. Math. Soc., 360 (2008),
2229-2264.
[18] R. G. Douglas, X. Tang, and G. Yu, An analytic Grothendieck Riemann Roch theorem, Adv. Math. 294(2016),
307-331.
[19] J. Eschmeier and S. Langendorfer, Cowen-Douglas tuples and fiber dimensions. J. Operator Theory 78 (2017), no.
1, 21-43.
[20] J. Eschmeier and S. Langend´'orfer, Cowen-Douglas operators and dominating sets. J. Operator Theory 72 (2014),
no. 1, 277-290.
[21] L. A. Fialkow, The similarity orbit of a normal operator, Trans. Amer. Math. Soc. 210(1975), 129-137.
[22] S. Ghara, The orbit of a bounded operator under the Mobius group modulo similarity equivalence, arXiv:1811.05428
[23] P. Guinand and L. W. Marcoux, Between the unitary and similarity orbits of normal operators. Pacific Journal of
Mathematics, 159, no. 2 (1993), 299-335.
[24] P. Guinand and L. W. Marcoux, On the U + K-orbits of certain weighted shifts. Integr. Equat. Oper. Th., 17 (1993),
516-543.
[25] D. W. Hadwin, An operator-valued spectrum, Indiana Univ. Math. J. 26(1977), 329-340.
[26] P. R. Halmos, Ten problems in Hilbert space, Bull. Amer. Math. Soc., 76 (1970) 887 - 933.
[27] D. A. Herrero, A trace obstruction to approximation by block-diagonal nilpotents, American Journal of Mathematics,
108, no. 2 (1986), 451-484
[28] D. A. Herrero, Spectral pictures of operators in the Cowen-Douglas class Bn(Ω) and its closure. J. Operator Theory
18 (1987), no. 2, 213-222.
ESSENTIALLY NORMAL COWEN-DOUGLAS OPERATORS
29
[29] B. Hou, P. Cui and Y. Cao, Chaos for Cowen-Douglas operators. Proc. Amer. Math. Soc. 138 (2010), no. 3, 929-936.
[30] Y. Hou, K. Ji and H. Kwon, The Trace of the Curvature determines Similarity. Studia Math., 236 (2017), No. 2,
193-200.
[31] G. Ji, Generalized Cowen-Douglas operators over Hilbert C*-modules. Integral Equations Operator Theory 20
(1994), no. 4, 395-409.
[32] K. Ji and J. Sarkar, Similarity of Quotient Hilbert modules in the Cowen-Douglas Class, arXiv:1212.2707 .
[33] K. Ji, C. Jiang, D. K. Keshari and G. Misra, Flag structure for operators in the Cowen-Douglas class, Comptes
rendus - Math´ematique, 352 (2014), 511 - 514.
[34] K. Ji, C. Jiang, D. K. Keshari and G. Misra, Rigidity of the flag structure for a class of Cowen-Douglas operators,
J. Func. Anal., 272 (2017), 2899 - 2932.
[35] Y. Ji, J. Li, Y. Yang A characterization of bilateral operator weighted shifts being Cowen-Douglas operators. Proc.
Amer. Math. Soc. 129 (2001), no. 11, 3205-3210.
[36] C. Jiang, K. Ji and D. K. Keshari, Geometric similarity classification of Cowen-Douglas operators, preprint, 2019.
[37] C.Jiang, K. Ji and G. Misra, Classification of quasi-homogeneous holomorphic curves and operators in the Cowen-
Douglas class, J. Func. Anal., 274(2017), 2399-2445.
[38] C. Jiang, Similarity, reducibility and approximation of the Cowen-Douglas operators, J. Operator Theory, 32 (1994),
77 - 89.
[39] C. Jiang, Similarity classification of Cowen-Douglas operators, Canad. J. Math., 56 (2004), 742 - 775.
[40] C. Jiang and K. Ji, Similarity classification of holomorphic curves, Adv. Math., 215 (2007), 446 - 468.
[41] C. Jiang, X. Guo and K. Ji, K-group and similarity classification of operators, J. Funct. Anal., 225 (2005), 167 -
192.
[42] C. Jiang and Z. Wang, Strongly irreducible operators on Hilbert space, Pitman Research Notes in Mathematics
Series, 389. Longman, Harlow, 1998.
[43] C. Jiang and Z. Wang, Structure of Hilbert Space Operators, World Scientific Publishing Co. Pte. Ltd., Hackensack,
NJ, 2006.
[44] C. Jiang and Z. Wang, The spectral picture and the closure of the similarity orbit of strongly irreducible operators,
Integral Equations Operator Theory, 24 (1996), 81 - 105.
[45] C. Jiang and D. Zheng, Similarity of analytic Toeplitz operator on the Bergman space. J. Funct. Anal., 258 (2010),
2961-2982 .
[46] A. Koranyi and G. Misra, Homogeneous Hermitian holomorphic vector bundles and the Cowen-Douglas class over
bounded symmetric domains, Comptes rendus - Math`ematique, 354 (2016), 291- 295.
[47] A. Kor´anyi and G. Misra, A classification of homogeneous operators in the Cowen-Douglas class, Adv. Math., 226
(2011), 5338 - 5360.
[48] A. Kor´anyi and G. Misra, Multiplicity-free homogeneous operators in the Cowen-Douglas class, Perspectives in
mathematical sciences. II, 83-101, Stat. Sci. Interdiscip. Res., 8, World Sci. Publ., Hackensack, NJ, 2009.
[49] A. Kor´anyi and G. Misra, Homogeneous operators on Hilbert spaces of holomorphic functions, J. Func. Anal., 254
(2008), 2419 - 2436.
[50] G. Misra, Curvature and the backward shift operators, Proc. Amer. Math. Soc., 91 (1984), 105 - 107.
[51] S. Krantz and N. Salinas, Proper holomorphic mappings and the Cowen-Douglas class. Proc. Amer. Math. Soc.
117 (1993), no. 1, 99-105.
[52] H. Kwon and S. Treil, Similarity of operators and geometry of eigenvector bundles, Publ. Mat., 53. (2009), 417 -
438.
[53] Q. Lin and N. Salinas A characterization of coanalytical Toeplitz operators in Cowen-Douglas class. Analysis and
partial differential equations, 519-527, Lecture Notes in Pure and Appl. Math., 122, Dekker, New York, 1990.
[54] M. Martin and N. Salinas, Flag manifolds and the Cowen-Douglas theory. J. Operator Theory 38 (1997), no. 2,
329-365.
[55] J. McCarthy, Boundary values and Cowen-Douglas curvature. J. Funct. Anal. 137 (1996), no. 1, 1-18.
[56] P. H. Naeini and B. Yousefi, On some properties of Cowen-Douglas class of operators. J. Funct. Spaces 2018, Art.
ID 6078594, 6 pp.
30
CHUNLAN JIANG, KUI JI AND JINSONG WU
[57] R. Reza, Curvature inequalities for operators in the Cowen-Douglas class of a planar domain. Indiana Univ. Math.
J. 67 (2018), no. 3, 1255-1279.
[58] A. L. Shields, Weighted Shift Operators and Analytic Function Theory, Mathematical Survey, 13 (1974), 49-128.
[59] L. Tian, W. Guo and K. Ji, A note on a subclass of Cowen-Douglas operators, To appear.
[60] S. Treil and B. D. Wick, Analytic projections, corona problem and geometry of holomorphic vector bundles, Journal
of the American Mathematical Society 22 (2009), no. 1, 55-76.
[61] D. Voiculescu A non-commutative Weyl-von Neumann theorem, Rev. Roum. Math. Pures et Appl. 21 (1976) 97-113.
[62] K. Wang and G. Zhang Curvature inequalities for operators of the Cowen-Douglas class. Israel J. Math. 222 (2017),
no. 1, 279-296.
[63] R. O. Wells Differential Analysis on Complex Manifolds, Graduate Texts in Mathematics, Springer; Softcover
reprint of hardcover 3rd ed. 2008.
[64] D. Wilkins, Homogeneous vector bundles and Cowen-Douglas operators. Internat. J. Math. 4 (1993), no. 3, 503-520.
[65] G. Yu, Reducing properties and unitary invariants of a class of operators. Chinese Science Bulletin 19 (1985).
[66] K. Zhu, Operators in Cowen-Douglas classes. Illinois J. Math. 44 (2000), no. 4, 767-783.
Current address, C. Jiang and K. Ji: Department of Mathematics, Hebei Normal University, Shijiazhuang, Hebei
050016, China
Current address, J. Wu: Institute for Advanced Study in Mathematics, Harbin Institute of Technology, Heilongjiang
150001, China
E-mail address, C. Jiang: [email protected]
E-mail address, K. Ji: [email protected], [email protected]
E-mail address, J. Wu: [email protected]
|
1908.01320 | 1 | 1908 | 2019-08-04T11:46:14 | A Sharp Inequality of Hardy-Littlewood Type Via Derivatives | [
"math.FA",
"math.CV"
] | In this paper we consider a generalized version of Carleman's inequality. An equivalent version of it states that $\|f\|_{A_\alpha^{2\alpha}}\leq\|f\|_{H^2}$, where $f$ is a holomorphic function and $\alpha>1$. If the norms $\|f\|_{A_\alpha^{2\alpha}}$ are decreasing in $\alpha$, then the inequality holds for $f$. For a dense set of functions, we calculate the derivative of the norms $\|f\|_{A_\alpha^{2\alpha}}$ in $\alpha$ and give sufficient conditions for this derivative to be non-positive. As an application, we prove the inequality for linear combinations of two reproducing kernels. Some numerical evidences are also provided. | math.FA | math | A Sharp Inequality of Hardy-Littlewood Type Via
Derivatives
Hui Dan, Kunyu Guo and Yi Wang
August 6, 2019
Abstract:
In this paper we consider a generalized version of Carleman's inequality. An equivalent
α ≤ (cid:107)f(cid:107)H 2, where f is a holomorphic function and α > 1. If
version of it states that (cid:107)f(cid:107)A2α
the norms (cid:107)f(cid:107)A2α
α are decreasing in α, then the inequality holds for f . For a dense set of
functions, we calculate the derivative of the norms (cid:107)f(cid:107)A2α
α in α and give sufficient conditions
for this derivative to be non-positive. As an application, we prove the inequality for linear
combinations of two reproducing kernels. Some numerical evidences are also provided.
Keywords: Carleman's inequality, inequality of Hardy and Littlewood type
2010 AMS Subject Classification: 30H20, 30H10
1
Introduction
In this paper we consider a sharp inequality concerning the weighted Bergman norms and
the Hardy norm on the unit disc D. Recall that for 0 < p < ∞, the Hardy space H p consists
of all holomorphic functions f on D such that
9
1
0
2
g
u
A
4
]
.
A
F
h
t
a
m
[
1
v
0
2
3
1
0
.
8
0
9
1
:
v
i
X
r
a
(cid:18)(cid:90) 2π
(cid:107)f(cid:107)H p := sup
0<r<1
f (reiθ)p dθ
2π
0
(cid:19) 1
p
< ∞.
(cid:19) 1
p
α consists of all holomorphic functions f on D such
For α > 1, the weighted Bergman space Ap
that
(cid:18)(cid:90)
α :=
where dµ(z) = (1−z2)−2 dxdy
we are considering in this paper is the following.
Conjecture 1. For any 0 < p ≤ 2 and any f ∈ H p,
D
π
(cid:107)f(cid:107)Ap
f (z)p(α − 1)(1 − z2)αdµ(z)
< ∞,
is the Mobius invariant measure of the unit disc. The inequality
(cid:107)f(cid:107)A2
2/p
≤ (cid:107)f(cid:107)H p.
1
(1.1)
In the case when p = 1, (1.1) is the Carleman's inequality (cf. [14]). For p = 1/k where k
is any positive integer, Burbea [9] showed that (1.1) holds true. If one releases the restriction
on the controlling constants, that is, if one asks whether
≤ C(cid:107)f(cid:107)H p
(cid:107)f(cid:107)A2
(1.2)
2/p
for some constant C, then using interpolation techniques, Brevig, Ortega-Cerd`a, Seip and
Zhao has proved that (1.2) holds for 0 < p < 1 and C as close to 1 as C = (2/(e log 2))1/2 =
1.030279 . . ..
In [6], the authors also gave and discussed about several interesting related
conjectures and questions.
In the case when p = 1, inequality (1.1) becomes
≤ (cid:107)f(cid:107)H 1.
(cid:107)f(cid:107)A2
2
This is known as the Carleman's inequality. In 1921, Carleman [10] proved this inequality
and used it to give the first complex-analytic proof of the famous isoperimetric theorem. For
a different purpose, in 1932, Hardy and Littlewood showed that H p ⊂ A2p
(in particular,
H 1 ⊂ A2
2) in [11]. See [14] for an excellent exposition of the relation between the two
problems. Various generalizations were proved, for example, in [1][7][8][9][13][15].
2
In recently years, Inequality (1.1) has regained attention because of its application in num-
ber theory. Via an iterating process [2] [12], contractive inequalities like (1.1) may "lift" multi-
plicatively to interesting inequalities for Hardy spaces on the infinite-dimensional torus, which
in turn, by the Bohr transform, translates into inequalities of Dirichlet polynomials[3][4][5].
Next, let us go to the technical side. An immediate observation is that for Conjecture 1,
it suffices to consider any outer function f , because multiplying an inner function on f does
not make a difference on the right hand side of (1.1), but makes the left hand side smaller.
For an outer function, one can consider its powers. By replacing f with f α where α = 2/p,
it is easy to show that Conjecture 1 is equivalent to the following (cf. [6]).
Conjecture 2. For any α > 1 and any f ∈ H 2,
(cid:107)f(cid:107)A2α
α ≤ (cid:107)f(cid:107)H 2.
(1.3)
Then Burbea's result [9] is equivalent to that (1.3) holds when α is any integer that is
greater than 1. A straight-forward proof was given in [6, Corollary 3].
In the case when
α is not an integer, the problem becomes very hard. The following computation may give
us a clue. In the case when α > 1 is an integer, and suppose that f is an outer function,
f =(cid:80)∞
n=0 anzn and a0 = 1. Then one can compute that
H 2 − (cid:107)f(cid:107)2α
(cid:107)f(cid:107)2α
∞(cid:88)
(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)an1 ··· ank − am1 ··· aml
(cid:12)(cid:12)(cid:12)(cid:12)2
(cid:1) and(cid:0)α
equation (under some convergence assumption). However, the coefficients(cid:0)α
This gives an alternative proof of Burbea's result. For non-integer valued α, we have the same
(cid:18)N + α − 1
(cid:19)−1 N(cid:88)
(cid:18)α
(cid:19)(cid:18)α
n1+···+nk=N,ni≥1
m1+···+ml=N,mi≥1
(cid:1) may
(cid:88)
.
(1.4)
A2α
α
N
=
1
2
N =0
k,l=1
k
l
2
k
l
be negative. Similar obstructions occur when one tries to extend other proofs of Burbea's
result to a non-integer valued α.
In [6], the authors gave several related conjectures (including the Conjectures 1 and 2
above) and questions. In particular, in [6, Question 1], they asked whether (cid:107)f(cid:107)2α
is non-
increasing in the parameter α, for an outer function f with (cid:107)f(cid:107)H 2 = 1. A positive answer
to the question above will lead to a positive answer to Conjecture 2.
In this paper, we
will mainly consider the following similar question, which allows us to drop the assumption
"(cid:107)f(cid:107)H 2 = 1" (see Remark 2.6).
Question 3. Suppose that f is an outer function. For α > 1, denote
A2α
α
Nf (α) = (cid:107)f(cid:107)A2α
α =
f (z)2α(α − 1)(1 − z2)αdµ(z)
.
(cid:19) 1
2α
(cid:18)(cid:90)
D
Is it true that
for all α > 1?
Nf (α) ≤ 0
∂
∂α
In Section 2, we will first show that a positive answer to Question 3 implies Conjecture
∂α Nf (α) for a dense set of functions. Based on
∂α Nf (α) to be non-positive.
2. Then we will give a discrete formula of ∂
the formula, in Section 3, we give some sufficient conditions for ∂
As an application, in Section 4, we obtain the following result.
Theorem 1.1 (Theorem 4.1 ). Suppose f ∈ H 2 and f = ηF , where η is inner and F has
no zeros in D. Suppose
for some α > 1, and c ∈ C2, w ∈ D2. Then for any 1 ≤ β ≤ α, we have
F α = c1Kw1,α + c2Kw2,α.
(1.5)
Equality holds if and only if F α = cKw,α for some c ∈ C and w ∈ D. As a consequence, we
have
.
β
(cid:107)F(cid:107)A2α
α ≤ (cid:107)F(cid:107)A2β
Equality holds if and only if f = cKw,1 for some c ∈ C and w ∈ D.
(cid:107)f(cid:107)A2α
α ≤ (cid:107)f(cid:107)H 2.
Some further remarks and numerical evidences are provided in Section 5.
2 A Discrete Formula
It is well-known that
for p > 0 and f ∈ H p [16]. Similarly, we have the following lemma.
(cid:107)f(cid:107)Ap
α = (cid:107)f(cid:107)H p
lim
α→1+
3
(1.6)
(2.1)
Lemma 2.1. Let O∗ denote the set of holomorphic functions f defined in some open neigh-
borhood of D such that f (z) (cid:54)= 0 for all z ∈ D. Then for any f ∈ O∗ and any p > 0,
(cid:107)f(cid:107)Aαp
α = (cid:107)f(cid:107)H p.
lim
α→1+
∂α Nf (α) ≤ 0, ∀α > 1, then (cid:107)f(cid:107)A2α
As a consequence, if f ∈ O∗ and ∂
Proof. Without loss of generality, we assume (cid:107)f(cid:107)H p = 1. Since f ∈ O∗, there exists C > c > 0
such that c ≤ f (z) ≤ C for z ∈ D.
It is easy to find a constant M > 0 such that
xα − x ≤ M (α − 1) for x ∈ [cp, C p] and α ∈ (1, 2). By (2.1),
α ≤ (cid:107)f(cid:107)H 2, ∀α > 1.
(2.2)
(cid:90)
lim
α→1+
D
(cid:12)(cid:12)(cid:12)(cid:12)(cid:90)
D
(cid:90)
≤ M (α − 1)
= M (α − 1) → 0, α → 1 + .
(α − 1)(1 − z2)αdµ(z)
D
f (z)p(α − 1)(1 − z2)αdµ(z) = 1.
On the other hand, since f (z)p ∈ [cp, C p] for all z ∈ D, we have
f (z)p(α − 1)(1 − z2)αdµ(z) −
f (z)αp(α − 1)(1 − z2)αdµ(z)
(cid:90)
D
(cid:12)(cid:12)(cid:12)(cid:12)
Therefore
(cid:107)f(cid:107)pα
Apα
α
→ 1, α → 1 + .
α = 1. The rest of the lemma is obvious. This completes the proof.
So limα→1+ (cid:107)f(cid:107)Apα
Remark 2.2. From Lemma 2.1, it is easy to see that a positive answer to Question 3 implies
Conjecture 2. The statement that ∂
jecture 2. However, we are still optimistic enough to expect a positive answer. One of the
evidences is the following. In [6], the authors proved a lemma ([6, Lemma 2]) which implies
(cid:107)f(cid:107)A2kα
α for any positive integer k, and used this lemma to prove (1.3) in the case
when α > 1 is an integer. In Section 5, we will also provide some numerical evidences that
support a positive answer to Question 3.
∂α Nf (α) ≤ 0 looks like a stronger statement than Con-
≤ (cid:107)f(cid:107)A2α
kα
Denote Logz the single-valued branch of log z on C\{z ∈ R : z ≤ 0} such that Log1 = 0.
By direct computation, we have
Proposition 2.3. If f ∈ O∗, then
∂
∂α
Nf (α) =
α − 1
2α
N 1−2α
f
where
If (α) =
(cid:90)
(α)
D
− 2
(cid:18)
f (z)2α(1 − z2)αLog(cid:0)f (z)2(1 − z2)(cid:1)dµ(z).
(α − 1)2 N 2α
f (α)LogNf (α) +
α − 1
N 2α
1
f (α) + If (α)
(cid:19)
, (2.3)
(2.4)
4
Suppose f is an outer function, then f α makes sense and is also an outer function. We
have the equation
= (cid:107)f α(cid:107)A2
α.
Also, from (2.3) and (2.4), we see that the value of ∂
∂α Nf (α) depends only on the function
f α. This allows us to consider f α instead of f and take advantage of the reproducing kernel
Hilbert space structure of A2
α.
(cid:107)f(cid:107)α
A2α
α
The main goal of this section is to prove the following theorem.
Theorem 2.4. Suppose f ∈ O∗ and α > 1. Suppose there exists k points, w1,··· , wk in D
and k numbers c1,··· , ck ∈ C such that
k(cid:88)
i=1
f α(z) =
ci
1
(1 − wiz)α ,
z ∈ D.
Nf (α) =
1
2α2 N 1−2α
f
(α)Df (α),
Then
where
Df (α) =
k(cid:88)
i,j=1
cicj
1
(1 − wiwj)α
∂
∂α
(cid:18)
log f α(wi) + log f α(wj) + αLog(1 − wiwj) − 2αLogNf (α)
.
(2.5)
(cid:19)
(2.6)
For log f α(wi), we fix a holomorphic function g such that f α(z) = eg(z) and let log f α(wi) =
g(wi). (since f α is outer, such function g exists.)
Before proving Theorem 2.4, let us use an example to illustrate our idea.
Example 2.5. It is well-known (and also implied by the proof of [6, Corollary 3]) that for
integer-valued α, the equation in (1.3) holds if and only if f (z) = c
w ∈ D. Indeed, if f (z) = 1
1−wz , then one can compute directly that ∂
1−wz for some c ∈ C and
∂α Nf (α) ≡ 0.
By Proposition 2.3, in order to compute ∂
the Mobius transform λ = ϕw(z) = w−z
1−wz , we get
∂α Nf (α), one needs to find out If (α). Applying
If (α) =
(cid:90)
(cid:90)
D
(cid:19)α
(cid:18) 1 − z2
(cid:18) 1 − λ2
(cid:19)α
(cid:90)
1 − wz2
D
=
1 − w2
= (1 − w2)−α
D
1 − z2
1 − wz2 dµ(z)
1 − λ2
1 − w2 dµ(λ)
Log
Log
(1 − λ2)α−2Log(1 − λ2)
dm(λ)
π
(cid:90)
−(1 − w2)−αLog(1 − w2)
(1 − λ2)α−2 dm(λ)
π
.
D
Here dm denotes the Lebesgue measure. Using the polar coordinates and applying an integra-
tion by parts, we have(cid:90)
D
(1 − λ2)α−2Log(1 − λ2)
dm(λ)
π
= −(α − 1)−2.
5
Similar computations give
and
Thus
(cid:90)
D
(1 − λ2)α−2 dm(λ)
π
=
1
α − 1
f (α) = (1 − w2)−α.
N 2α
If (α) = −(α − 1)−2(1 − w2)−α − (α − 1)−1(1 − w2)−αLog(1 − w2).
From this and Proposition 2.3, it is easy to see that ∂
∂α Nf (α) ≡ 0.
By Example 2.5, ∂
α. The linear span of reproducing kernels form a dense set in A2
∂α Nf (α) = 0 whenever f α is a constant multiple of a reproducing kernel
α. This explains our
of A2
reason of considering such functions in Theorem 2.4.
Next, let us give the proof of Theorem 2.4.
Proof of Theorem 2.4. Suppose f ∈ O∗ and
k(cid:88)
i=1
f α(z) =
ci
1
(1 − wiz)α .
As in Example 2.5, in order to calculate ∂
∂α Nf (α), we need to find out If (α) as defined in
(2.4). Compared with Example 2.5, the main difficulty here is that we can not use the Mobius
transform. We will get around by applying the Stoke's Theorem and the Residue Theorem.
Let
and
II =
III =
1
α
(cid:90)
By (2.4), it is easy to see that
(cid:90)
D
D
f α(z)2(1 − z2)αLogf α(z)2dµ(z)
f α(z)2(1 − z2)αLog(1 − z2)dµ(z).
(2.7)
(2.8)
(2.9)
Taking advantage of the fact that log f α(z) is a holomorphic function in D, we have
If (α) = II + III.
(cid:19)
II =
2
α
Re
D
f α(z)2(1 − z2)α log f α(z)dµ(z)
1
(1 − wjz)α f α(z) log f α(z)(α − 1)(1 − z2)αdµ(z)
D
(cid:18)(cid:90)
(cid:90)
Re
cj
j=1
k(cid:88)
k(cid:88)
k(cid:88)
j=1
Re
i,j=1
=
=
=
2
α(α − 1)
2
α(α − 1)
1
α(α − 1)
(cid:19)
log f α(wi) + log f α(wj)
.
(2.10)
cjf α(wj) log f α(wj)
cicj
1
(1 − wiwj)α
(cid:18)
6
Next, we calculate III.
f α(z)2(1 − z2)αLog(1 − z2)dµ(z)
(cid:90)
cicjIIIij.
III =
=
=
Here
cicj
1
(1 − wiz)α
1
(1 − zwj)α (1 − z2)αLog(1 − z2)dµ(z)
D
IIIij =
1
(1 − wiz)α
D
1
(1 − zwj)α (1 − z2)αLog(1 − z2)dµ(z).
For i, j = 1,··· , k, define
IVij =
D
Vij =
1
(1 − wiz)α
1
(1 − zwj)α (1 − z2)αLog
1 − z2
1 − zwj2 dµ(z)
1
(1 − wiz)α
1
(1 − zwj)α (1 − z2)αLog1 − zwj2dµ(z).
IIIij = IVij + Vij.
and
Then
D
i,j=1
(cid:90)
k(cid:88)
k(cid:88)
(cid:90)
(cid:90)
(cid:90)
i,j=1
D
(cid:90)
(cid:90)
D
(2.11)
(2.12)
(2.13)
(2.14)
(2.15)
Since Logz = Logz, we have
1
(1 − wiz)α
1
(cid:18)
Vij =
1
1
(1 − zwj)α Log(1 − zwj)(1 − z2)αdµ(z)
(1 − zwj)α Log(1 − wjz)(1 − z2)αdµ(z)
(cid:18)
(cid:19)
(cid:19)
(1 − wiwj)α Log(1 − wj2)
Log(1 − wiwj) + Log(1 − wj2)
1
+
=
D
1
(1 − wiz)α
(1 − wiwj)α Log(1 − wiwj) +
1
(1 − wiwj)α
It remains to calculate IVij. Let
α − 1
α − 1
=
1
1
.
(2.16)
ϕij(z) =
1
α − 1
1
wj − z
and
By direct computation, we have
ψij(z) = Log
(1 − z2)α−1
(1 − zwj)α
1
(1 − wiz)α ,
1 − z2
1 − zwj2 .
z ∈ D, z (cid:54)= wj.
¯∂ϕij(z) =
1
(1 − wiz)α
(1 − z2)α−2
(1 − zwj)α
7
and
Therefore
¯∂ψij(z) =
wj − z
(1 − z2)(1 − zwj)
.
(cid:90)
1
π
IVij =
¯∂ϕij(z)ψij(z)dxdy.
(2.17)
D
For any ε > 0 sufficiently small, define
Dε,j := {z ∈ D : z − wj > ε}.
Define the one-form ω = ϕij(z)ψij(z)dz. Then
dω = −( ¯∂ϕijψij + ϕij
¯∂ψij)dz ∧ d¯z = 2
Applying the Stokes's Theorem on Dε,j, we get
√−1( ¯∂ϕijψij + ϕij
¯∂ψij)dx ∧ dy.
( ¯∂ϕijψij + ϕij
¯∂ψij)dx ∧ dy
(cid:90)
ϕijψijdz −
{z:z−wj=ε}
ϕijψijdz.
{z:z−wj=ε}
(cid:19)
ϕijψijdz
(cid:90)
1
π
(cid:18)(cid:90)
(cid:90)
T
Dε,j
1
=
2π
√−1
√−1
= − 1
2π
¯∂ϕijψijdx ∧ dy
The second equality is because ϕijψij = 0 on the unit circle T.
(cid:90)
(cid:18) 1
Dε,j
(cid:90)
Therefore
IVij = lim
ε→0
1
π
= − lim
ε→0
(cid:90)
= −
D
− lim
ε→0
= −
(α − 1)2
where
V Iij = − lim
ε→0
1
√−1
2π
(cid:90)
1
2π
Dε,j
π
1
(cid:90)
¯∂ψijdx ∧ dy +
ϕij
(1 − z2)α
√−1
α − 1
(1 − wiz)α(1 − zwj)α dµ(z)
√−1
1
(1 − wiwj)α + V Iij,
{z:z−wj=ε}
1
wj − z
α − 1
2π
1
1
1
(cid:90)
{z:z−wj=ε}
(cid:19)
ϕijψijdz
(1 − z2)α−1
(1 − zwj)α−1
1
(1 − wiz)α Log
1 − z2
1 − zwj2 dz
(2.18)
1
α − 1
1
wj − z
(1 − z2)α−1
(1 − zwj)α−1
1
(1 − wiz)α Log
1 − z2
1 − zwj2 dz.
{z:z−wj=ε}
To calculate V Iij, notice that
(1 − z2)α−1
(1 − zwj)α−1
1
(1 − wiz)α Log
1 − z2
1 − zwj2 → −
1
(1 − wiwj)α Log(1 − wj2),
z → wj.
8
Standard estimates will give us
V Iij =
1
α − 1
= − 1
α − 1
1
(1 − wiwj)α Log(1 − wj2)Res(
(1 − wiwj)α Log(1 − wj2).
1
1
wj − z
, wj)
By (2.18) and (2.19), we have
IVij = −
1
(α − 1)2
1
(1 − wiwj)α − 1
α − 1
1
(1 − wiwj)α Log(1 − wj2).
By (2.15), (2.20) and (2.19), we get
IIIij = −
1
(α − 1)2
1
(1 − wiwj)α +
1
α − 1
1
(1 − wiwj)α Log(1 − wiwj).
(2.19)
(2.20)
(2.21)
Then combining (2.9), (2.10), (2.11) and (2.21), we have
1
α(α − 1)
cicj
1
(1 − wiwj)α
log f α(wi) + log f α(wj)
(cid:19)
(cid:18)
(cid:18)
1
cicj
(1 − wiwj)α
(1 − wiwj)α Log(1 − wiwj)
1
cicj
If (α) =
=
i,j=1
k(cid:88)
k(cid:88)
k(cid:88)
k(cid:88)
i,j=1
i,j=1
i,j=1
−
1
(α − 1)2
− 1
α − 1
1
α(α − 1)
−
1
(α − 1)2 N 2α
f (α).
(cid:19)
(2.22)
(2.23)
cicj
1
(1 − wiwj)α
log f α(wi) + log f α(wj) − αLog(1 − wiwj)
The last equality is because
f (α) = (cid:107)f α(cid:107)2
N 2α
A2
α
= (cid:104)f α, f α(cid:105)A2
α =
k(cid:88)
i,j=1
cicj
1
(1 − wiwj)α .
Finally, plugging in (2.22) and (2.23) into (2.3), we get (2.5) and (2.6). This completes
the proof.
Remark 2.6. In [4], the authors raised the question whether Nf (α)2α = (cid:107)f(cid:107)2α
is non-
increasing in α given that (cid:107)f(cid:107)H 2 = 1. Using our method, we can also compute the derivative
∂α Nf (α)2α. In fact, by direct computation, we get
A2α
α
∂
Nf (α)2α = (α − 1)−1Nf (α)2α + (α − 1)If (α).
∂
∂α
9
By (2.22), if f ∈ O∗ and f α(z) =(cid:80)k
If (α) =
Thus
i=1 ci
(1−wiz)α , we have
1
(cid:18)
k(cid:88)
i,j=1
cicj
1
(1 − wiwj)α
1
α(α − 1)
−
1
(α − 1)2 N 2α
f (α).
k(cid:88)
(cid:18)
log f α(wi) + log f α(wj) − αLog(1 − wiwj)
(cid:19)
(cid:19)
1
1
α
cicj
Nf (α)2α =
∂
∂α
Using (2.24), one can easily check that if we drop the condition (cid:107)f(cid:107)H 2 = 1, then there exists
f such that Nf (α)2α is increasing.
log f α(wi) + log f α(wj) − αLog(1 − wiwj)
(1 − wiwj)α
(2.24)
i,j=1
.
An immediate consequence of Theorem 2.4 is the following.
Theorem 2.7. Suppose f ∈ O∗, α > 1 and f α =(cid:80)k
1
of points {w1,··· , wk} belong to a single real line, and that ci ≥ 0, i = 1··· , k. Then
i=1 ci
(1−wiz)α . Suppose further that set
Proof. The proof simply an application of the Jensen's inequality. Note that under our
assumption,
cicj
1
(1 − wiwj)α ≥ 0,
f α(wj) =
ci
1
(1 − wiwj)α ≥ 0,
∀i, j = 1,··· , k.
Nf (α) ≤ 0.
∂
∂α
k(cid:88)
j=1
Without loss of generality, let us assume that Nf (α) = 1. That is
k(cid:88)
i,j=1
cicj
1
(1 − wiwj)α = 1.
Then by the Jensen's inequality, we have
Df (α) =
1
(1 − wiwj)α Log
f α(wi)f α(wj)(1 − wiwj)α
k(cid:88)
i,j=1
cicj
(cid:18) k(cid:88)
(cid:18)
(cid:19)
≤ Log
cicjf α(wi)f α(wj)
i,j=1
f (α)
= LogN 4α
= 0.
10
(cid:19)
(2.25)
The second equality is because
k(cid:88)
k(cid:88)
ci
k(cid:88)
cj
cif α(wi) =
i=1
i=1
j=1
1
(1 − wiwj)α = N 2α
f (α).
By (2.23) and (2.25), we have ∂
∂α Nf (α) ≤ 0. This completes the proof.
From the proof of Theorem 2.7 we know that the inequality ∂
∂α Nf (α) ≤ 0 holds true if we
can "formally" apply the Jensen's inequality. However, in general, the coefficients involved
are not positive, and one needs to find other ways.
3 Sufficient Conditions
In this section, we give some other sufficient conditions for ∂
∂α Nf (α) to be non-positive. We
want to consider the right hand side of (2.6) under a suitable general setting. For this, let
us first discuss about how Theorem 2.4 can be used to answer Question 3.
In (2.6), the term log f α(wi) depends on the fact that f is an outer function: the imaginary
part of log f α(wi) depends on the formula (assuming f (0) > 0)
log f (z) =
1
2π
eiθ + z
eiθ − z
Logf (eiθ)dθ.
(3.1)
(cid:90) π
−π
However, it is unclear how this formula could enter the estimates. Things are relatively easy
if we are able to apply the single-valued branch Logz to all f α(wi). It turns out that such
special cases are enough for our purpose (See Proposition 3.3). Before going into details, let
us fix some notations.
Notations:(1) In the rest of this paper, we use k to denote a positive integer. If not otherwise
specified, c denotes a k-tuple of complex numbers, and w denotes a k-tuple of points in D.
that is, c = (c1, c2,··· , ck), ci ∈ C, w = (w1, w2,··· , wk), wi ∈ D. Given c, w and α > 0,
we use Wα to denote the k × k matrix with entry
(1−wiwj )α in the i-th row and j-th column.
Thinking of c as a row vector, we reserve the notation fα = (f1,α,··· , fk,α) for the row vector
(1−wiwj )α . If α is specified, then we
defined by fα = cWα. Denote Nα = cWαc∗ =(cid:80)k
i,j=1 cicj
1
1
drop the subscription α.
(2) It is well-known that for α > 0 and w ∈ D, the functions
(1 − wz)α , w ∈ D
define a unique reproducing kernel Hilbert space on D [1].
α = 1, it is H 2 = A2
1. In this paper, for any α > 0, we use A2
reproducing kernel Hilbert space determined by {Kw,α : w ∈ D}.
We find it convenient to consider the following general setting.
Kw,α(z) =
1
If α > 1, the space is A2
α; if
α to denote the uniquely defined
11
Definition 3.1. (1) Let H := {z ∈ C : Rez > 0} denote the right half plane. Suppose α > 0,
c ∈ Ck and w ∈ Dk satisfies fi,α ∈ H, i = 1,··· , k, where fα = (f1,α,··· , fk,α) is defined as
above. Define
Dα(c, w) =
1
(1 − wiwj)α
Logfi,α + Logfj,α + αLog(1 − wiwj)
k(cid:88)
i,j=1
cicj
k(cid:88)
i=1
(cid:18)
k(cid:88)
i,j=1
(cid:19)
− NαLogNα
(3.2)
(3.3)
(cid:27)
= 2Re
cifi,αLogfi,α + α
cicj
1
(1 − wiwj)α Log(1 − wiwj) − NαLogNα.
(2) For any α > 0 and 0 < ε ≤ 1, define
(cid:26)
Λα,ε =
and
(cid:26) k(cid:88)
(c, w) : c ∈ Ck, w ∈ (−ε, ε)k, fi,α ∈ H, i = 1,··· , k, k is a positive integer.
(cid:27)
.
(cid:27)
.
Kα,ε =
ci
We use Λα, Kα to denote Λα,1, Kα,1.
i=1
1
(1 − wiz)α : (c, w) ∈ Λα,ε
(3) Define
Γ =
(cid:26)
(c, w) : c ∈ Rk, w ∈ (−1, 1)k, k is a positive integer
For (c, w) ∈ Γ and α > 0, define
Dα(c, w) = 2
cifi,αLogfi,α +
k(cid:88)
i=1
k(cid:88)
i,j=1
cicj
1
(1 − wiwj)α Log(1 − wiwj)α − NαLogNα.
(3.4)
Note that since xLogx tends to 0 as x tends to 0, the definition above makes sense even if
fi = 0 for some i = 1,··· , k. It is also easy to see that (3.4) coincides with (3.2) when
fi,α > 0,∀i = 1,··· , k.
Remark 3.2. Suppose (c, w) ∈ Γ and f α(z) =(cid:80)k
it is easy to see that Df (α) = Dα(c, w). If (c, w) ∈ Λα and f α =(cid:80)k
i=1 ciKw,α(z) ∈ O∗, then by (2.6) and (3.4),
i=0 ciKw,α ∈ O∗, then
it is not necessarily true that Df (α) = Dα(c, w). However, if one knows that {w1,··· , wk}
is contained in a connected open subset Ω of D which is mapped, by f α, into H, then by
standard argument, the function Logf α(z)Ω differs from the function given in (3.1), by an
integer multiple of 2πi. Then from the expression of (2.6) one can see that Df (α) = Dα(c, w).
We will use this fact later.
It turns out that we only need to consider the case when f α ∈ Kα,ε for ε small enough.
12
Proposition 3.3. Suppose for any α > 1 there exists 0 < ε ≤ 1 such that Dα(c, w) ≤ 0 for
∂α Nf (α) ≤ 0 for all f ∈ O∗. As a consequence, Conjecture 2 holds.
all (c, w) ∈ Λα,ε. Then ∂
The proof is based on the following two lemmas.
Lemma 3.4. Suppose α > 1 and 0 < ε ≤ 1. Then for any g ∈ O∗ such that g(0) = 1, there
exists 0 < δ ≤ ε and a sequence {gn} ∈ Kα,δ such that gn converges uniformly on D to g.
Moreover, gn(z) ∈ H for all n and all z ∈ D with z < δ.
Proof. Since g(0) = 1, we can choose 0 < δ ≤ ε such that Reg(z) > 1
z < δ. Choose r > 1 such that g is defined on {z ∈ C : z ≤ r}. Define
2 for any z ∈ D with
gr(z) = g(rz),
z ∈ D.
Obviously, gr ∈ A2
α. The subspace
span{Kw,α : w ∈ (−δ, δ)}
α. Choose a sequence {gn} ⊂ span{Kw,α : w ∈ (−δ, δ)} such that gn → gr in A2
α
is dense in A2
norm. Then gn converge uniformly to gr on {z ∈ C : z ≤ 1
), n = 1, 2,··· .
gn(z) = gn(
r}. Define
Then gn converge uniformly to g on D. By construction, each gn is of form
z
r
kn(cid:88)
where wn,i ∈ (−δ, δ), ∀i. Therefore
gn =
cn,iKwn,i,α,
i=1
cn,i
1
(1 − wn,i
r z)α =
kn(cid:88)
i=1
cn,iK wn,i
r
,α.
kn(cid:88)
i=1
gn(z) =
r ∈ (−δ, δ). Also, since gn converge uniformly to g and Reg(z) > 1
Since r > 1 we have wn,i
2 if
z < δ, by passing to a subsequence, we have gn(z) ∈ H for any n and any z with z < δ.
In particular, gn( wn,i
r ) ∈ H. Therefore gn ∈ Kα,δ for any n. This completes the proof.
The following lemma is simply a consequence of the fact that, for f ∈ H 2, fr(z) := f (rz)
converges to f in H 2 norm as r → 1−.
Lemma 3.5. Suppose f is an outer function in H 2. Then there exists a sequence {fn} ⊂ O∗
such that fn tends to f in the Hardy norm (cid:107) · (cid:107)H 2.
Proof of Proposition 3.3. Assume that for some α > 1 and 0 < ε ≤ 1 we have Dα(c, w) ≤
0 for all (c, w) ∈ Λα,ε. For any f ∈ O∗, we want to show ∂
∂α Nf (α) ≤ 0. Without loss of
generality we can assume f (0) = 1. By Lemma 3.4, there exists 0 < δ ≤ ε and a sequence
{gn} in Kα,δ such that gn converges uniformly to f α on D and gn maps {z ∈ D : z < δ}
13
n . By Remark 3.2, we have ∂
to H. Also, since f α is bounded away from 0 on D, for n large enough, gn is outer and we
∂α Nfn(α) = Dα(cn, wn) ∈ Λα,δ, where gn
can define fn = g1/α
∂α Nfn(α) ≤ 0 for n large enough.
corresponds to (cn, wn). In particular, we have ∂
On the other hand, since f is bounded away from 0 we also have Logfn converging
uniformly to Logf on D. By (2.3) and (2.4) it is easy to see that ∂
∂α Nfn(α) → ∂
∂α Nf (α) as
Aα ≤ (cid:107)f(cid:107)H 2 for all
n tend to infinity. Thus ∂
f ∈ O∗. By Lemma 3.5, the inequality also holds for all outer functions. Suppose f = ηg
α ≤ (cid:107)g(cid:107)H 2 = (cid:107)f(cid:107)H 2. This completes
where η is inner and g is outer. Then (cid:107)f(cid:107)A2α
the proof.
∂α Nf (α) ≤ 0. Then by Lemma 2.1, we have (cid:107)f(cid:107)2α
α ≤ (cid:107)g(cid:107)A2α
Now we are ready to give some sufficient conditions for Dα(c, w) to be non-positive.
Theorem 3.6. Suppose α > 0 and (c, w) ∈ Λα ∪ Γ satisfy the following conditions.
(1) w1 < w2 < ··· < wk, where k is the number of entries in w;
(2) either {c2,··· , ck} or {c1,··· , ck−1} are real and have the same sign.
Then we have Dα(c, w) ≤ 0.
We will need the following lemma in the proof of Theorem 3.6.
Lemma 3.7. Suppose α > 0, c ∈ Ck and A = [aij] is semi-positive definite, aij > 0. Let
f = cA and N = cAc∗. Then for any x1,··· , xk ≥ 0 we have
k(cid:88)
i,j=1
xixjaijLog
fifj
aijN
≤ 0.
(3.5)
Without loss of generality, we can also assume that(cid:80)k
Proof. The proof is, again, an application of the Jenson's Inequality. If some fi equals zero
then the left hand side is −∞ and the inequality always holds. Assume fi are all non-zero.
i,j=1 xixjaij = 1. Applying the Jenson's
Inequality, we get
k(cid:88)
i,j=1
(cid:18) k(cid:88)
xixjaijLog
fifj
aijN
≤ Log
xixjaij
fifj
aijN
≤ Log
(cid:19)
(cid:0)(cid:80)
i xifi(cid:1)2
N
.
(3.6)
i,j=1
(cid:1)2
(cid:0)(cid:80)k
, which is less than or equal to Log(cid:80)k
(cid:80)k
semi-positive definite. Since xi ≥ 0 and aij > 0, we have(cid:80)k
Choose ei ∈ C such that ei = 1 and eifi = fi. Then the right hand side of (3.6) becomes
i,j=1 xieixjejaij by the fact that A is
Log
i,j=1 xixjaij =
i,j=1 xieixjejaij ≤(cid:80)k
i,j=1 xieicj aij
i,j=1 cicj aij
1. Therefore the left hand side of (3.5) is less than or equal to 0. This completes the proof.
Proof of Theorem 3.6. We will prove the theorem in the case when (c, w) ∈ Λα. The
proof when (c, w) ∈ Γ is similar. First, we notice that if we let −w = (−w1,··· ,−wk), then
Dα(c, w) = Dα(c,−w). From this, it is easy to see that it suffices to consider the case when
w1 < w2 < ··· < wk and {c2,··· , ck} are real and have the same sign.
14
Let us further reduce the cases. Suppose w1 < w2 < ··· < wk. Let zi = −ϕw1(wi),
i = 1,··· , k. Then it is easy to check that 0 = z1 < z2 < ··· < zk. Suppose (c, w) ∈ Λα. Let
di = ci(1−ziw1)α
(1−zizj )α ], f = cW
and g = dΛ. Using the well-known formula
(1−wiwj )α ], Λ = [
(1−w2
1
1
1)α/2 and consider the pair (d, z). Write W = [
(1 − za)(1 − aw)
(1 − a2)(1 − zw)
1 − ϕa(z)ϕa(w)
=
1
it is easy to check that
g = f diag
(cid:18) (1 − w2
1)α/2
(1 − ziw1)α
,
(cid:19)
z, w, a ∈ D,
.
Then (d, z) ∈ Λα. From the equation above,
it is also straight-forward to check that
Dα(c, w) = Dα(d, z). Therefore, in order to prove Theorem 3.6, we only need to consider
the case when 0 = w1 < ··· < wk and {c2,··· , ck} are real and have the same sign.
Assume that (c, w) ∈ Λα, 0 = w1 < ··· < wk and ci ≥ 0,∀i = 2,··· , k. The case when
c2 = ··· = ck = 0 is trivial. Thus we can assume that ci > 0 for some i = 2,··· , k. Define f
and W as before. The idea is to find a non-increasing function that takes value Dα(c, w) at
α and 0 at 0.
(1−wiwj )t , i, j = 1,··· , k. Define Wt = [aij,t], Nt = cWtc∗
and ft = cWt. Since only c1 may have imaginary part, the signs of the imaginary part of
each fi,t depend only on that of c1. Assume, without loss of generality, that Imc1 ≥ 0. Then
Imfi,t ≥ 0,∀i = 1,··· , k. Define
For 0 ≤ t ≤ α, define aij,t =
1
Dt = 2Re
cicjaij,tLogaij,t − NtLogNt,
0 ≤ t ≤ α.
k(cid:88)
i=1
cifi,tLogfi,t − k(cid:88)
dt fi,t =(cid:80)k
i,j=1
> 0,∀i = 1,··· , k. So
Notice that since w1 = 0, we have d
the points t such that fi,t = 0 for some i, are isolated. Also, zLogz → 0 if z tends to 0. From
this we can see that Dt is a continuous, piecewise differentiable function.
(1−wiwj )t Log
1−wiwj
j=2 cj
1
1
Next, we show that Dt is non-increasing. By the previous argument, it suffices to show
dt Dt ≤ 0 at the points where each fi,t is non-zero. By direct computation, we get
that d
Logfi,t + Logfj,t − Log(aij,tNt)
.
(cid:19)
Since w1 = 0 and ci ≥ 0,∀i = 2,··· , k, we have
1
1
Dt =
cicj
d
dt
(cid:1)(cid:18)
Logfi,t + Logfj,t − Log(aij,tNt)
(cid:19)
dt
i,j=1
cicj
aij,t
(cid:18) d
(cid:19)(cid:18)
k(cid:88)
(1 − wiwj)t Log(cid:0)
(cid:0) ∞(cid:88)
(cid:1)Log
k(cid:88)
i wn
j
1
n
wn
n=1
ciwn
i cjwn
j aij,tLog
1 − wiwj
fi,tfj,t
aij,tNt
fi,tfj,t
aij,tNt
d
dt
Dt =
i,j=2
k(cid:88)
k(cid:88)
∞(cid:88)
i,j=2
=
cicjaij,t
1
n
=
≤ 0.
n=1
i,j=2
15
Here the last inequality is by Lemma 3.7. If {c2,··· , ck} are all non-positive, simply replace
ci with −ci in the above argument. Thus in either case we have that Dt is non-increasing.
It is obvious that Dα = Dα(c, w). By straight-forward computation it is also easy to
show that D0 = 0. Therefore Dα(c, w) = Dα ≤ D0 = 0. This completes the proof.
Theorem 3.8. Suppose α > 0, c ∈ C2, w ∈ D2 and (c, w) ∈ Λα ∪ Γ. Then
Moreover, Dα(c, w) = 0 if and only if c1 = 0, or c2 = 0, or w1 = w2.
Dα(c, w) ≤ 0.
Proof. The proof is similar as that of Theorem 3.6. Define W = [aij] = [
For 0 ≤ t ≤ 1, define
1
(1−wiwj )α ] as before.
(cid:40)
aij,t =
(1 − t)a11 + t a2
aij,
12
a22
,
i = j = 1
otherwise.
Write Wt = [aij,t], Nt = cWtc∗ and ft = cWt. It is easy to check that the following hold.
(i) f1,t = (1 − t)f1 + t a12
a22
f2, f2,t ≡ f2;
(ii) d
dt a11,t ≤ 0;
(iii) each Wt is semi-positive definite;
(iv) W1 has rank 1.
By (i), the paths f1,t and f2,t stay in H. Define
2(cid:88)
cifi,tLogfi,t − 2(cid:88)
i=1
i,j=1
Dt := 2Re
cicjaij,tLogaij,t − NtLogNt,
Then Dt is a differentiable function on (0, 1). From (iv) it is easy to compute that D1 = 0.
By direct computation, we have
(cid:19)
(cid:18) d
dt
Dt = c12
d
dt
a11,t
Log
f1,t2
a11,tNt
.
Since Wt is positive definite, by Lemma 3.7, it is easy to see that d
D0 ≤ D1 = 0. Equality holds if and only if d
w1 (cid:54)= w2, then d
dt a11,t (cid:54)= 0. Thus either c1 = 0 or
f1,t2
a11,tNt
dtDt ≥ 0. Thus Dα(c, w) =
dt Dt ≡ 0. This always holds when w1 = w2. If
≡ 1. In particular, if c1 (cid:54)= 0, then
f1,02 = cW0e∗
12 = a11,0N0 = (e1W0e∗
1)(cW0c∗).
Here e1 = (1, 0,··· , 0). Since W0 is positive definite, this occurs only when c2 = 0. This
completes the proof.
In terms of ∂
∂α Nf (α), we summarize our results as follows.
16
Theorem 3.9. Suppose f ∈ O∗, α > 1, f α =(cid:80)k
i=1 ciKwi,α and (c, w) ∈ Λα ∪ Γ. Suppose
one of the following holds.
(1) ci ≥ 0, i = 1,··· , k.
(2) w1 < ··· < wk, and either {c2,··· , ck} or {c1,··· , ck−1} are real and have the same
sign.
(3) k = 2.
Then we have ∂
∂α Nf (α) ≤ 0.
4 Norm Inequalities for Linear Combinations of Two
Reproducing Kernels
∂α Nf (α) ≤ 0,∀α > 1, for some f ∈ O∗, then
Recall that in Lemma 2.1, we showed that if ∂
Conjecture 2 holds for f . In this section, we provide an alternative way of proving results on
Conjecture 2, using results obtained in Section 3. As a consequence, we prove the following
theorem.
Theorem 4.1. Suppose f ∈ H 2 and f = ηF , where η is inner and F has no zeros in D.
Suppose
for some α > 1, and c ∈ C2, w ∈ D2. Then for any 1 ≤ β ≤ α, we have
F α = c1Kw1,α + c2Kw2,α.
(4.1)
Equality holds if and only if F α = cKw,α for some c ∈ C and w ∈ D. As a consequence, we
have
.
β
(cid:107)F(cid:107)A2α
α ≤ (cid:107)F(cid:107)A2β
Equality holds if and only if f = cKw,1 for some c ∈ C and w ∈ D.
(cid:107)f(cid:107)A2α
α ≤ (cid:107)f(cid:107)H 2.
(4.2)
1
The proof is based on a different way of viewing Dα(c, w). Recall that in Definition 3.1,
for α > 0 and (c, w) ∈ Λα ∪ Γ, we defined Wα = [
(1−wiwj )α ] and fα = cWα. Then Dα(c, w)
is defined using fα and Wα. In the case when {wi : i = 1,··· , k} are distinct points, the
matrix Wα is invertible. Therefore we have c = fαW−1
α . This means we can define Dα(c, w)
using fα.
Definition 4.2. Suppose α > 0, k is a positive integer, and w ∈ Dk is such that {wi : i =
1,··· , k} are distinct. Define Wα as usual. Suppose either f ∈ Hk or w ∈ (−1, 1)k, f ∈ Rk.
Let cα = fW−1
α . Define
(cid:98)Dα(f , w) = Dα(cα, w).
(4.3)
17
Definition 4.3. Suppose w ∈ Dk, {w1,··· , wk} are distinct, and α > 0. Define
Kw,α = span{Kwi,α : i = 1,··· , k} ⊂ A2
α
and Pw,α the orthogonal projection from A2
(f (w1),··· , f (wk)), then it is easy to compute that
α onto Kw,α. For f ∈ A2
α, if we denote f (w) =
(cid:107)Pw,α(f )(cid:107)2
A2
α
= f (w)W−1
α f (w)∗.
Proof of Theorem 4.1. It is easy to see that (4.2) follows from (4.1). Thus we only need
to prove (4.1).
(1) First, we prove (4.1) under the following conditions.
(i) F α = c1 + c2Kw,α, w ≥ 0;
(ii) There is a connected open neighborhood Ω, of {0, w}, such that F α(Ω) ⊂ H.
Assume the above, then we can choose log F α such that log F αΩ = LogF αΩ. As a conse-
quence, F βΩ ⊂ H for any β ∈ [1, α].
For 1 ≤ β ≤ α, consider
(cid:18)
Nβ :=
F β(w)W−1
where F β(w) = (F β(w1), F β(w2)). Then by straight-forward computation, we have
= (cid:107)Pw,βF β(cid:107)2/β
,
A2
β
β F β(w)∗(cid:19)1/β
(cid:98)Dβ(F β(w), w) ≤ 0.
d
dβ
Nβ =
1
β2 N 1−β
β
The last inequality is because of Theorem 3.8. Therefore we have
(cid:107)F(cid:107)A2α
α = (cid:107)F α(cid:107)1/α
A2
α
= N 1/2
α ≤ N 1/2
β = (cid:107)Pw,βF β(cid:107)1/β
A2
β
≤ (cid:107)F β(cid:107)1/β
A2
β
= (cid:107)F(cid:107)A2β
β
.
If we assume that w (cid:54)= 0, then by Theorem 3.8, we also know that the equality holds if and
only if for each β, either c1,β = 0 or c2,β = 0. In particular, either c1 = 0 or c2 = 0. On the
other hand, if either c1 = 0 or c2 = 0, then it is easy to check that the equality in (4.1) holds.
This completes the proof for case (1).
(2) Next, we consider the case when F α = c1 + c2Kw,α, w ∈ D, c1, c2 ∈ C. Choose θ ∈ [0, 2π]
so that eiθw ≥ 0. Let Fθ(z) = F (e−iθz). Then F α
θ = c1 + c2Keiθw,α. Inequality (4.1) for F
follows from (4.1) for Fθ. Thus we may assume that w ≥ 0 in the beginning. Suppose F has
no zeros in D and F α = c1 + c2Kw,α with w ≥ 0. We will show that after multiplying F by
a non-zero constant, the condition (ii) in case (1) will be satisfied. This will lead to (4.1) for
case (2). We may as well assume that F α = 1 + cKw,α with w ≥ 0. It is easy to see that F α
(1−w2)α ,
maps the interval [0, w] onto the (complex valued) interval between 1 + c and 1 +
which we denote by I. Since F has no zeros in D, 0 /∈ I. Thus I must be contained in some
half plane eiθ1H. By standard trick we can find a connected open neighborhood Ω of [0, w]
c
18
such that F α(Ω) ⊂ eiθ1H. Therefore(cid:0)e−iθ1/αF(cid:1)α(Ω) ⊂ H. This completes the proof for case
(2).
(3) In general, suppose F has no zeros in D and F α = c1Kw1,α + c2Kw2,α. Denote
kw,β(z) =
(1 − w2)β/2
(1 − wz)β , w ∈ D, β > 0.
Let G = F ◦ ϕw1 · kw1,1. Then G also has no zeros in D and Gβ = F β ◦ ϕw1 · kw1,β. It is
standard to check that (cid:107)Gβ(cid:107)A2
, ∀β ∈ [1, α]. Notice that (4.1) is equivalent to
β
= (cid:107)F β(cid:107)A2
(cid:107)F α(cid:107)1/α
β
A2
α
≤ (cid:107)F β(cid:107)1/β
β ∈ [1, α].
,
A2
β
So it suffices to prove
(cid:107)Gα(cid:107)1/α
which is in turn, equivalent to (cid:107)G(cid:107)A2α
we have
A2
α
≤ (cid:107)Gβ(cid:107)1/β
α ≤ (cid:107)G(cid:107)A2β
A2
β
β
c1
Gα =
(1 − w12)α/2 + c2
β ∈ [1, α],
,
, β ∈ [1, α]. By straight-forward computation
(1 − w12)α/2
(1 − w2w1)α Kϕw1 (w2),α.
Thus G satisfies case (2). This completes the proof of (4.1) in the general case. Tracing back
to the proof of (1), we also see that the equality in (4.1) holds if and only if F α = cKw,α for
some c ∈ C and w ∈ D. Then (4.2) follows immediately. This completes the proof.
In terms of Conjecture 1, (4.2) becomes the following.
Corollary 4.4. Suppose 0 < p ≤ 2 and α = 2
and F has no zeros in D. If
p. Suppose f ∈ H p, f = ηF , where η is inner
F = c1Kw1,α + c2Kw2,α,
then
Equality holds if and only if f = cKw,α for some c ∈ C and w ∈ D.
(cid:107)f(cid:107)A2
2/p
≤ (cid:107)f(cid:107)H p.
5 Remarks and Numerical Evidences
5.1 Some Further Remarks
(1) In Proposition 3.3, we give a sufficient condition for Conjecture 2 to hold. In its most
general form, we list the conjecture below.
Conjecture 4. Suppose α > 0 and (c, w) ∈ Λα. Then Dα(c, w) defined as in (3.2) and (3.4)
is non-positive.
Another interesting question to ask is the following.
19
A2
α
(cid:20)
(cid:21)
are non-
Logfi + Logfj − Log(aij,αNα)
Question 5. Suppose Conjecture 4 holds. Does it imply that the norms (cid:107)f α(cid:107)1/α
increasing for α > 0, for some set of functions f ?
(2) An immediate observation from (3.2) is that Dα(c, w) is the bi-linear form cWα◦ Bc,wc∗,
and A ◦ B denotes the Hadmadard
where Bc,w =
product of A and B. Thus a sufficient condition of Conjecture 4 would be that −Bc,w is
semi-positive definite. However, this is not true, even in the simplest case. Take, for example,
c1 = c2 = 1, w1 = 0, and w2 ∈ (0, 1) such that
(3) In Section 4, we also defined the function (cid:98)Dα(f , w). Then Conjecture 4 is equivalent to
Conjecture 6. Suppose α > 0, w ∈ (−1, 1)k and f ∈ Hk. Then (cid:98)Dα(f , w) ≤ 0.
2)α = 1.1.
the following.
i,j=1,··· ,k
1
(1−w2
(4) There is , yet another way of looking at (3.2). Suppose A = [aij] is a semi-positive definite
k × k matrix, aij ∈ H, and c ∈ Ck is such that each entry of f := cA belongs to H. Then we
can define N (c, A) = cAc∗ and
(cid:101)D(c, A) = 2Re
k(cid:88)
cifiLogfi − k(cid:88)
cicjaijLogaij − N (c, A)LogN (c, A).
i=1
i,j=1
are secretly using this definition: we let the matrix A to vary from Wα to a rank 1 matrix.
Then we can ask whether (cid:101)D(c, A) ≤ 0. In the proofs of Theorem 3.6 and Theorem 3.8, we
Then we used the fact that if A has rank 1, then (cid:101)D(c, A) = 0. An interesting observation is
the following: if we define (cid:101)A = [(cid:101)aij] to be the block matrix
and(cid:101)c = (−1, c), then (cid:101)A is semi-positive definite and (cid:101)D(c, A) = (cid:101)D((cid:101)c,(cid:101)A). Moreover, (cid:101)D((cid:101)c,(cid:101)A)
(cid:20)cAc∗
f
f∗ A
(cid:101)A =
(cid:21)
,
has the simple expression
(cid:101)D((cid:101)c,(cid:101)A) = −(cid:88)(cid:101)ci(cid:101)cj(cid:101)aijLog(cid:101)aij.
In general, we ask the following question.
Question 7. Suppose A = [aij] is semi-positive definite and Reaij ≥ 0, ∀i, j = 1,··· , k.
Suppose c ∈ Ck is such that cA = 0. Do we have
(cid:101)D(c, A) = − k(cid:88)
cicjaijLogaij ≤ 0?
5.2 Numerical Evidences
i,j=1
One of the advantages that Theorem 2.4 offers is that we can now test Question 3 using
numerical methods. We have tested for a wide range of values of c and w. We list a few
graphs for the interested readers.
20
Figure 1: k = 4, c, w as indicated
(1) In the proof of Theorem 3.6, we showed that Dα(c, w) is non-increasing in α under the
given conditions. However, Figure 1 shows that this is not always true.
(2) If we adopt the definition (cid:98)Dα(f , w), then by fixing f2 and f3 and let f1 vary, we get
Figures 2 and 3.
Figure 2: k = 3, Imf1, f2, f3 and w as indicated
(3) In the special case when all entries of c are real, we can use the definition (3.4). By
letting one of the coefficients vary, we get Figure 4.
Acknowledgment: The first and second author is partially supported by the Natural Sci-
ence Foundation of China. The third author is partially supported by the Natural Science
Foundation.
21
Figure 3: k = 3, Ref1, f2, f3 and w as indicated
Figure 4: k = 3, c1, c2 and w as indicated
References
[1] N. Aronszajn, Theory of reproducing kernels, Trans. Amer. Math. Soc.,
68(1950), 337-404.
[2] F. Bayart, Hardy spaces of Dirichlet series and their composition operators.
Monatsh. Math. 136 (2002), no. 3, 203-236.
[3] F. Bayart, O. Brevig, A. Haimi, J. Ortega-Cerd`a, K. Perfekt, Contractive in-
equalities for Bergman spaces and multiplicative Hankel forms. Trans. Amer.
Math. Soc. 371 (2019), no. 1, 681-707.
[4] A. Bondarenko, O. Brevig, E. Saksman, K. Seip, J. Zhao, Pseudomoments of
the Riemann zeta function. Bull. Lond. Math. Soc. 50 (2018), no. 4, 709-724.
[5] A. Bondarenko, W. Heap, K. Seip, An inequality of Hardy-Littlewood type for
Dirichlet polynomials. J. Number Theory 150 (2015), 191-205.
22
[6] O. Brevig, J. Ortega-Cerd`a, K. Seip, J. Zhao, Contractive inequalities for Hardy
spaces. Funct. Approx. Comment. Math. 59 (2018), no. 1, 41-56.
[7] J. Burbea, Inequalities for reproducing kernel spaces. Illinois J. Math. 27 (1983),
no. 1, 130-137.
[8] J. Burbea, Inequalities for holomorphic functions of several complex variables.
Trans. Amer. Math. Soc. 276 (1983), no. 1, 247-266.
[9] J. Burbea, Sharp inequalitites for holomorphic functions. Illinois J. Math. 31
(1987), 248-264
[10] T. Carleman, Zur theorie der minimalflachen, Math. Zeit., 9(1921) 154-160.
[11] G. H. Hardy, J. E. Littlewood, Some properties of fractional integrals. II. Math.
Zeit. 34 (1932), no. 1, 403-439.
[12] H. Helson, Hankel forms and sums of random variables. Studia Math. 176 (2006),
no. 1, 85-92.
[13] M. Mateljevi´c and M. Pavlovi´c, New proofs of the isoperimetric inequality and
some generalizations, J. Math. Anal. Appl. 98(1984) 25-30.
[14] D. Vukoti´c, The isoperimetric inequality and a theorem of Hardy and Littlewood.
Amer. Math. Monthly 110 (2003), no. 6, 532-536.
[15] S. E. Warschawski, On the differentiability at the boundary in conformal map-
ping. Proc. Amer. Math. Soc. 12 (1961), 614-620.
[16] K. Zhu, Translating inequalities between Hardy and Bergman spaces. Amer.
Math. Monthly 111 (2004), no. 6, 520-525.
Hui Dan, School of Mathematical Sciences, Fudan University, Shanghai, 200433, China, E-
mail: [email protected]
Kunyu Guo, School of Mathematical Sciences, Fudan University, Shanghai, 200433, China,
E-mail: [email protected]
Yi Wang, Department of Mathematics, SUNY Buffalo, 244 Mathematics Building Buffalo,
NY 14260-2900 E-mail: [email protected]
23
|
1705.07792 | 1 | 1705 | 2017-05-22T15:02:03 | Fourier multipliers in Banach function spaces with UMD concavifications | [
"math.FA",
"math.CA"
] | We prove various extensions of the Coifman-Rubio de Francia-Semmes multiplier theorem to operator-valued multipliers on Banach function spaces. Our results involve a new boundedness condition on sets of operators which we call $\ell^{r}(\ell^{s})$-boundedness, which implies $\mathcal{R}$-boundedness in many cases. The proofs are based on new Littlewood-Paley-Rubio de Francia-type estimates in Banach function spaces which were recently obtained by the authors. | math.FA | math |
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
WITH UMD CONCAVIFICATIONS
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Abstract. We prove various extensions of the Coifman -- Rubio de Francia --
Semmes multiplier theorem to operator-valued multipliers on Banach function
spaces. Our results involve a new boundedness condition on sets of opera-
tors which we call ℓr(ℓs)-boundedness, which implies R-boundedness in many
cases. The proofs are based on new Littlewood -- Paley -- Rubio de Francia-type
estimates in Banach function spaces which were recently obtained by the au-
thors.
1. Introduction
In [46] Rubio de Francia proved a surprising extension of the classical Littlewood --
Paley square function estimate: for all p ∈ [2, ∞) there exists a constant Cp > 0
such that for any collection I of mutually disjoint intervals in R, the estimate
(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)XI∈I
s > (cid:12)(cid:12) 1
SI f 2(cid:1)1/2(cid:13)(cid:13)(cid:13)(cid:13)Lp(R)
2(cid:12)(cid:12), then every m : R → C of bounded s-variation
holds for all Schwartz functions f ∈ S(R), where SI is the Fourier projection onto
I. As a consequence, in [14] Coifman, Rubio de Francia, and Semmes showed that
if p ∈ (1, ∞) and 1
uniformly on dyadic intervals induces a bounded Fourier multiplier Tm on Lp(R).
This is analogous to the situation for the Marcinkiewicz multiplier theorem (the
s = 1 case of the Coifman -- Rubio de Francia -- Semmes theorem), which follows from
the classical Littlewood -- Paley theorem.
(1.1)
≤ Cpkf kLp(R)
p − 1
Consider a Banach space X. We are interested in analogues of the results
above for operator-valued multipliers on X-valued functions; that is, for multipliers
m : R → Lb(X), where Lb(X) denotes the space of bounded linear operators on X,
and where we consider a natural extension of the Fourier transform which acts on
X-valued functions. A necessary condition for boundedness of the Fourier multi-
plier Tm on some Bochner space Lp(R; X) is that the range m(R) is R-bounded (see
Remark 5.9). R-boundedness is a probabilistic strengthening of uniform bounded-
ness which holds automatically for bounded scalar-valued multipliers. Following the
breakthrough papers [12, 51] there has been an extensive study of operator-valued
multiplier theory, in which R-boundedness techniques are central. For example,
2010 Mathematics Subject Classification. Primary: 42B15 Secondary: 42B25; 46E30, 47A56.
Key words and phrases. Fourier multipliers, UMD Banach function spaces, bounded s-
variation, Littlewood -- Paley -- Rubio de Francia inequalities, Muckenhoupt weights, Complex
interpolation.
The authors are supported by the VIDI subsidy 639.032.427 of the Netherlands Organisation
for Scientific Research (NWO).
1
2
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Marcinkiewicz-type theorems were obtained in [2, 4, 7, 12, 21, 48, 51]. We refer to
[22] for a more detailed historical description.
An operator-valued analogue of the Coifman -- Rubio de Francia -- Semmes theorem
was obtained in [24]. There the Banach space X was assumed to satisfy the so-
called LPRp (Littlewood -- Paley -- Rubio de Francia) property, which was previously
studied in [5, 19, 24, 25, 45]. This is a generalisation of the square function estimate
(1.1) which may be formulated for all Banach spaces, but which may not hold.
Naturally, R-boundedness assumptions play a role in the results of [24]. In [1] we
proved a range of Littlewood -- Paley -- Rubio de Francia-type estimates for Banach
function spaces, including the LPRp property, under assumptions involving the
UMD property and convexity (generalising a key result of [45]). The main goal
of this paper is to prove Coifman -- Rubio de Francia -- Semmes type results for such
Banach function spaces.
The following multiplier theorem is the fundamental result of this paper. Let ∆ =
{±[2k, 2k+1), k ∈ Z} denote the standard dyadic partition of R. Let X and Y be
Banach function spaces, and for a set of bounded linear operators T ⊂ Lb(X, Y ) let
V s(∆; T ) denote the space of functions m : R → span(T ) with bounded s-variation
uniformly on dyadic intervals J ∈ ∆, measured with respect to the Minkowski norm
on span(T ) (see below Definition 4.1). Denote the q-concavification of a Banach
function space X by X q (see Section 2.2).
Theorem 1.1. Let q ∈ (1, 2], p ∈ (q, ∞), s ∈ [1, q), and let w be a weight in
the Muckenhoupt class Ap/q. Let X and Y be Banach function spaces such that
X q and Y have the UMD property. Let T ⊂ Lb(X, Y ) be absolutely convex and
ℓ2(ℓq′
)-bounded, and suppose that m ∈ V s(∆; T ). Then the Fourier multiplier Tm
is bounded from Lp(w; X) to Lp(w; Y ).
This is proven as part of Theorem 5.8. The assumptions on X imply Littlewood --
Paley -- Rubio de Francia-type estimates that are used in the proof. In this theorem
a condition called 'ℓ2(ℓq′
)-boundedness' appears where one would usually expect
an R-boundedness condition. This is a new notion which arises naturally from
the proof; it turns out to imply R-boundedness. We investigate the more general
notion of ℓr(ℓs)-boundedness in Section 3.
The case q = 2 and w = 1 of Theorem 1.1 was considered in [24, Theorem 2.3]
for Banach spaces X = Y with the LPRp property. Our approach only works for
Banach function spaces (and closed subspaces thereof), but these are currently the
only known examples of Banach spaces with LPRp. As the parameter q decreases
we assume less of X, but more of T and m. In Section 5 we prove Theorem 1.1,
along with various other extensions and modifications of this result. In particular
we obtain the following improvement of Theorem 1.1 for Lebesgue spaces.
Theorem 1.2. Let s ∈ [2, ∞). Suppose that m : R → L(Lr(w)) for some r ∈ (1, ∞)
and all w ∈ Ar(Rd), and that the following Holder-type condition is satisfied:
sup
x∈R
km(x)kL(Lr(w)) + sup
J∈∆
J
1
s [m]C 1/s(J;Lb(Lr(w))) ≤ φr([w]Ar ).
Then the Fourier multiplier Tm is bounded on Lp(R; Lr(Rd)) in each of the following
cases:
(i) r ∈ [2, ∞) and 1
(ii) r ∈ (1, 2] and 1
s > max(cid:8) 1
s > max(cid:8) 1
2 − 1
p , 1
2 − 1
r , 1
p − 1
r },
r − 1
2 , 1
r − 1
2 , 1
p − 1
p }.
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
3
Here φr([w]Ar ) denotes an unspecified non-decreasing function of the Mucken-
houpt characteristic [w]Ar . The result follows from the combination of Proposition
5.11 and Example 5.16. The Holder assumption allows for the construction of a
suitable set T as in Theorem 1.1. The condition on s becomes less restrictive as the
numbers p, r, and 2 get closer. Taking p = r or r = 2 is particularly illustrative: the
condition on s is then 1
theorem. However, even if p = r, the operator-valued nature of the symbol m pre-
vents us from simply deducing the boundedness of Tm from the scalar-valued case
by a Fubini argument. Using the same techniques, one could also deduce versions
of Theorem 1.2 with Muckenhoupt weights in the R- and Rd-variables.
2(cid:12)(cid:12), as in the Coifman -- Rubio de Francia -- Semmes
p − 1
s > (cid:12)(cid:12) 1
In Section 5.4 we present some new Coifman -- Rubio de Francia -- Semmes-type
theorems on UMD Banach spaces (not just Banach functon spaces) which are
complex interpolation spaces between a Hilbert space and a UMD space. Typi-
cal examples which are not Banach function spaces include the space of Schatten
class operators, and more generally non-commutative Lp-spaces. Our results in
this context are weaker than those that we obtain for Banach function spaces, but
nonetheless they seem to be new even for scalar multipliers.
Overview.
• In Section 2 we present some preliminaries on Muckenhoupt weights, UMD
Banach function spaces, and Rubio de Francia extrapolation.
• In Section 3 the notion of ℓr(ℓs)-boundedness of a set of operators is defined
and investigated.
• In Section 4 we discuss the class V s of functions of bounded s-variation,
and a related atomic space Rs.
• In Section 5 we present our main results, which are several operator-valued
Fourier multiplier theorems. We cover results for Hilbert spaces, UMD
Banach function spaces, 'intermediate' UMD Banach function spaces, and
general 'intermediate' UMD Banach spaces.
Notation. Throughout the paper we consider complex Banach spaces, but every-
thing works just as well for real Banach spaces.
If Ω is a measure space (we omit reference to the measure unless it is needed)
and X is a Banach space, we let L0(Ω; X) denote the vector space of measurable
functions modulo almost-everywhere equality, and we let Σ(Ω; X) denote the vector
space of all simple functions f : Ω → X. When X = C we write L0(Ω) and Σ(Ω).
For vector spaces V and W , L(V, W ) denotes the vector space of linear operators
from V to W . For Banach spaces X and Y , Lb(X, Y ) denotes the bounded linear
operators from X to Y and kT kL(X,Y ) the operator norm.
Throughout the paper we write φa,b,... to denote a non-decreasing function
[1, ∞) → [1, ∞) which depends only on the parameters a, b, . . ., and which may
change from line to line. Nondecreasing dependence on the Muckenhoupt charac-
teristic of weights is used in applications of extrapolation theorems. We do not
obtain sharp dependence on Muckenhoupt characteristics in our results. In [1, Ap-
pendix A] it is shown that monotone dependence on the Muckenhoupt characteristic
can be deduced from a more general estimate in terms of the characteristic.
For p, q ∈ [1, ∞] and θ ∈ [0, 1], we define the interpolation exponent [p, q]θ by
1
[p, q]θ
=
1 − θ
p
+
θ
q
4
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
with the interpretation 1/0 := ∞. This lets us write interpolation results such as
[Lp, Lq]θ = L[p,q]θ in a pleasing compact form.
Occasionally we will work with Rd for a fixed dimension d ≥ 1. Implicit constants
in estimates will depend on d, but we will not state this.
2. Preliminaries
2.1. Muckenhoupt weights. A locally integrable function w ∈ L1
loc(Rd) is called
a weight if it is non-negative almost everywhere. For p ∈ [1, ∞) the space Lp(w) =
Lp(Rd, w) consists of all f ∈ L0(Rd) such that
kf kLp(Rd,w) :=(cid:16)Rd
f (x)pw(x) dx(cid:17)1/p
< ∞.
The Muckenhoupt Ap class is the set of all weights w such that
[w]Ap := sup
B
1
B B
w(x) dx ·(cid:16) 1
B B
w(x)− 1
p−1(cid:17)p−1
< ∞,
where the supremum is taken over all balls B ⊂ Rd, and where the second factor is
replaced by kw−1kL∞(B) when p = 1. Define A∞ =Sp≥1 Ap. For 1 < p ≤ q ≤ ∞
we say that a weight w is in the αp,q class if w1−p′
∈ Ap′/q′ , and we write
[w]αp,q := [w1−p′
]Ap′ /q′ .
This class naturally arises in duality arguments. The αp,2 class is used in [26],
where it is denoted by αp.
We will need the following properties of the Ap classes.
Proposition 2.1.
(i) The Ap classes are increasing in p, with [w]Aq ≥ [w]Ap when 1 ≤ q ≤ p.
(ii) For all w ∈ Ap with p ∈ (1, ∞) there is an ε > 0 such that w ∈ Ap−ε.
(iii) For all w ∈ Ap with p ∈ [1, ∞) there is a δ > 0 such that w1+δ ∈ Ap.
For proofs and further details on Muckenhoupt weights see [20, Chapter 9].
2.2. The UMD property. We say that a Banach space X has the UMD property if
the Hilbert transform extends to a bounded operator on Lp(R; X) for all p ∈ (1, ∞).
This is equivalent to the original definition in terms of martingale differences [9, 6].
For a detailed account of the theory of UMD spaces we refer the reader to [10]
the reflexive Lp spaces, Sobolev
and [22]. The "classical" reflexive spaces (i.e.
spaces, Besov spaces, Triebel -- Lizorkin spaces and Schatten classes) have the UMD
property. The UMD property implies reflexivity, so for example L1 and L∞ do not
have the UMD property.
Most of our results are stated in terms of Banach function spaces that are p-
convex for some p ∈ (1, ∞), and whose p-concavifications X p are also Banach
function spaces, where X p = {f : f 1/p ∈ X} with norm
kxkX p = kx1/pkp
X .
For an introduction to these notions see [1, Section 2.1]. We write 'X p ∈ UMD'
as shorthand notation for 'X p is a Banach space which has the UMD property'.
If p ≥ 1 this therefore includes the assumption that X is p-convex. The condition
that X p ∈ UMD is open in p: in fact, if X p ∈ UMD, then there exists ε > 0 such
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
5
that X q ∈ UMD for all 0 < q < p + ε [47, Theorem 4]. In particular, X p ∈ UMD
for some p > 1 if and only if X is UMD.
2.3. Extrapolation. The following Rubio de Francia-type vector-valued extrapo-
lation result was obtained by the authors in [1, Theorem 3.2].
Theorem 2.2. Fix p0 ∈ (0, ∞) and let X be a Banach function space over (Ω, µ)
with X p0 ∈ UMD. Suppose that F ⊂ L0
+(Rd; X) and that for all
p > p0, (f, g) ∈ F , and w ∈ Ap/p0 we have
+(Rd; X) × L0
kf (·, ω)kLp(w) ≤ φp,p0 ([w]Ap/p0
)kg(·, ω)kLp(w)
µ-a.e. ω ∈ Ω.
Then for all p > p0, (f, g) ∈ F , and w ∈ Ap/p0 we have
kf kLp(w;X) ≤ φX,p,p0 ([w]Ap/p0
)kgkLp(w;X).
This theorem implies the following corollary for operators, which is also proved
in [1], where it is formulated more generally. For the definition of the extension eT
see [1, Lemma 2.4].
Theorem 2.3. Fix p0 ∈ (0, ∞), and let T ∈ Lb(Lp(w)) for all p > p0 and w ∈
Ap/p0 , with
Then for all Banach function spaces X with X p0 ∈ UMD, the operator T has an
kT kL(Lp(w)) ≤ φp,p0 ([w]Ap/p0
).
extension eT on Lp(w; X) for all p > p0 and w ∈ Ap/p0 , with
keT kL(Lp(w;X)) ≤ φX,p,p0 ([w]Ap/p0
).
We used these results in [1] to deduce Littlewood -- Paley -- Rubio de Francia-type
estimates, and we use them here to prove ℓr(ℓs)-boundedness of families of opera-
tors.
3. ℓr(ℓs)-boundedness
Our operator-valued multiplier theorems involve a new condition on sets of
bounded operators T ⊂ Lb(X, Y ), which we call ℓr(ℓs)-boundedness. This gen-
eralises the more familiar notions of R-boundedness and ℓs-boundedness. In this
section we introduce and explore the concept.
3.1. Definitions and basic properties.
Definition 3.1. Let X and Y be Banach spaces and T ⊂ Lb(X, Y ).
• Let (εk)∞
k=1 be a Rademacher sequence on a probability space Ω. We say
j=1 in
that T is R-bounded if for all finite sequences (Tj)n
X,
j=1 in T and (xj)n
nXk=1
(cid:13)(cid:13)(cid:13)
εkTkxk(cid:13)(cid:13)(cid:13)L2(Ω;Y )
.(cid:13)(cid:13)(cid:13)
nXk=1
εkxk(cid:13)(cid:13)(cid:13)L2(Ω;X)
.
The least admissible implicit constant is called the R-bound of T , and
denoted [T ]R.
6
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
• Suppose that X and Y are Banach function spaces and suppose s ∈ [1, ∞].
j=1 in T and
We say that T is ℓs-bounded if for all finite sequences (Tj)n
(xj )n
j=1 in X,
(cid:13)(cid:13)(cid:13)(cid:16) nXk=1
Tkxks(cid:17)1/s(cid:13)(cid:13)(cid:13)Y
.(cid:13)(cid:13)(cid:13)(cid:16) nXk=1
xks(cid:17)1/s(cid:13)(cid:13)(cid:13)X
.
The least admissible implicit constant is called the ℓs-bound of T , and
denoted [T ]ℓs.
For a detailed treatment of R-boundedness we refer the reader to [23, 29], and
for ℓs-boundedness see [28, 50].
Definition 3.2. Let X and Y be Banach function spaces, T ⊂ Lb(X, Y ) and r, s ∈
[1, ∞]. We say that T is ℓr(ℓs)-bounded if for all finite doubly-indexed sequences
(Tj,k)n,m
j,k=1 in T and (xj,k)n,m
j,k=1 in X,
(cid:13)(cid:13)(cid:13)(cid:16) nXj=1(cid:16) mXk=1
Tj,kxj,ks(cid:17)r/s(cid:17)1/r(cid:13)(cid:13)(cid:13)Y
.(cid:13)(cid:13)(cid:13)(cid:16) nXj=1(cid:16) mXk=1
xj,ks(cid:17)r/s(cid:17)1/r(cid:13)(cid:13)(cid:13)X
.
The least admissible implicit constant is called the ℓr(ℓs)-bound of T , and denoted
[T ]ℓr(ℓs).
For R- and ℓ2-boundedness it suffices to consider subsets of T in the defining
inequality (see [12, 31]). For ℓs- and ℓr(ℓs)-boundedness with r, s 6= 2 this is not
the case: one must consider sequences, allowing for repeated elements. A singleton
{T } can fail to be ℓs-bounded, as the defining estimate may fail for arbitrarily long
constant sequences (T, . . . , T ) (see [28, Example 2.16]). We say that an operator
T ∈ Lb(X, Y ) is ℓs- or ℓr(ℓs)-bounded if the singleton {T } is.
If a set T ⊂ Lb(X, Y ) is R-, ℓs-, or ℓr(ℓs)-bounded, then so is its closure in the
strong operator topology, and likewise its absolutely convex hull absco(T ). This was
proven in [29] for R-boundedness and [28] for ℓs-boundedness; the proof generalises
to ℓr(ℓs)-boundedness.
It is immediate from the definition that ℓs-boundedness and ℓs(ℓs)-boundedness
are equivalent. The following proposition encapsulates a few other connections be-
tween R-, ℓr-, and ℓr(ℓs)-boundedness. For a thorough discussion on the connection
between R and ℓ2-boundedness we refer to [31].
Proposition 3.3. Let X and Y be Banach function spaces and T ⊂ Lb(X, Y ).
(i) If Y is p-concave for some p < ∞ and T is R-bounded, then T is ℓ2-bounded
with [T ]ℓ2 . [T ]R.
(ii) If X is p-concave for some p < ∞ and T is ℓ2-bounded, then T is R-bounded
with [T ]R . [T ]ℓ2.
(iii) Let p, s ∈ [1, ∞]. If X is p-concave, Y is p-convex, and T is ℓs-bounded, then
T is ℓp(ℓs)-bounded with [T ]ℓp(ℓs) ≤ [T ]ℓs.
(iv) Let r, s ∈ [1, ∞]. If T is ℓr(ℓs)-bounded, then T is ℓr- and ℓs-bounded with
[T ]ℓr ≤ [T ]ℓr(ℓs) and [T ]ℓs ≤ [T ]ℓr(ℓs).
Proof. Statements (i) and (ii) follow from the Khintchine-Maurey inequalities (see
[36, Theorem 1.d.6]). For (iii), consider doubly-indexed finite sequences (Tj,k)m,n
j,k=1
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
7
in T and (xj,k)m,n
j,k=1 in X. Then we have
(cid:13)(cid:13)(cid:13)(cid:16) mXj=1(cid:16) nXk=1
Tj,kxj,ks(cid:17)p/s(cid:17)1/p(cid:13)(cid:13)(cid:13)Y
p
≤(cid:16) mXj=1(cid:13)(cid:13)(cid:13)(cid:16) nXk=1
≤ [T ]ℓs(cid:16) mXj=1(cid:13)(cid:13)(cid:13)(cid:16) nXk=1
≤ [T ]ℓs(cid:13)(cid:13)(cid:13)(cid:16) mXj=1(cid:16) nXk=1
Tj,kxj,ks(cid:17)1/s(cid:13)(cid:13)(cid:13)
X(cid:17)1/p
xj,ks(cid:17)1/s(cid:13)(cid:13)(cid:13)
Y(cid:17)1/p
xj,ks(cid:17)p/s(cid:17)1/p(cid:13)(cid:13)(cid:13)X
p
,
so [T ]ℓp(ℓs) ≤ [T ]ℓs. Finally, (iv) follows by taking one index to be a singleton. (cid:3)
Proposition 3.3 shows in particular that if T is ℓ2(ℓs)- or ℓs(ℓ2)-bounded for
some s ∈ [1, ∞], then T is ℓ2-bounded, and hence R-bounded if Y is p-concave for
some p < ∞.
Consider the situation of Theorem 2.3. If a family of linear operators T satisfies
the hypothesis of the theorem uniformly, then the family of extensions eT is auto-
matically ℓr(ℓs)-bounded for r, s > p0. This observation is a convenient source of
ℓr(ℓs)-bounded families.
Proposition 3.4. Fix p0 ∈ (1, ∞), and suppose that T ⊂ Lb(Lp(w)) for all p ∈
(p0, ∞) and w ∈ Ap/p0 . In addition suppose that for each T ∈ T and f ∈ Lp(w),
kT f kLp(w) ≤ φp0,p([w]Ap/p0
)kf kLp(w).
the set of extensions obtained in Theorem 2.3. Then for all p, r, s ∈ (p0, ∞) and all
Let X be a Banach function space with X p0 ∈ UMD, and let eT = {eT : T ∈ T } be
w ∈ Ap/p0 , eT is ℓr(ℓs)-bounded on Lp(w; X) and
Proof. Consider doubly-indexed finite sequences (Tj,k)m,n
Σ(Rd; X). Let Ω be the underlying measure space of X, and define
[eT ]ℓr (ℓs) ≤ φp0,p,r,s,X ([w]Ap/p0
).
j,k=1 in T and (gj,k)m,n
j,k=1 in
F, G : Rd × Ω × {1, . . . , m} × {1, . . . , n} → R+
by
F (·, ω, j, k) = Tj,kgj,k(·, ω)
and G(·, ω, j, k) = gj,k(·, ω).
Then from the assumption on T we see that for all p > p0 and all w ∈ Ap/p0 ,
kF (·, ω, j, k)kLp(w) ≤ φp0,p([w]Ap/p0
)kG(·, ω, j, k)kLp(w).
Letting Y := X(ℓr
)) is
UMD, with UMD constants independent of m, n ∈ N. Hence Theorem 2.2 implies
that for all p ∈ (p0, ∞) and w ∈ Ap/p0 ,
n)), it follows from [47, p. 214] that Y p0 = X p0(ℓr/p0
m (ℓs/p0
m(ℓs
n
kF kLp(w;Y ) ≤ φX,p0,p,r,s([w]Ap/p0
)kGkLp(w;Y ).
This, combined with [1, Lemma 2.4], implies the claimed result.
(cid:3)
Taking X to be the scalar field C, so that X p0 = X for any p0, we obtain the
following special case. Note that in this case a more direct proof may be given as
in [18, Theorem 2.3].
8
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Proposition 3.5. Fix p0 ∈ (1, ∞), and suppose that T ⊂ Lb(Lp(w)) for all p ∈
(p0, ∞) and w ∈ Ap/p0 , and in addition suppose that for all T ∈ T and f ∈ Lp(w),
kT f kLp(w) ≤ φp0,p([w]Ap/p0
)kf kLp(w).
Then for all p, r, s ∈ (p0, ∞) and all w ∈ Ap/p0 , T is ℓr(ℓs)-bounded on Lp(w) and
[T ]ℓr (ℓs) ≤ φp0,p,r,s([w]Ap/p0
).
Duality and interpolation may be used to establish ℓr(ℓs)-boundedness, as shown
in the following two propositions.
Proposition 3.6. Let X, Y be Banach function spaces, and let T ⊂ Lb(X, Y ). Let
r, s ∈ [1, ∞]. If T is ℓr(ℓs)-bounded, then the adjoint family
T ∗ = {T ∗ : T ∈ T } ⊂ Lb(Y ∗, X ∗)
is ℓr′
(ℓs′
)-bounded with [T ∗]ℓr′ (ℓs′ ) = [T ]ℓr (ℓs).
Proof. This follows from the duality relation X(ℓr
Section 1.d]).
n )) (see [36,
(cid:3)
m(ℓs
n))∗ = X ∗(ℓr′
m(ℓs′
To exploit interpolation we must assume order continuity, which holds auto-
matically for reflexive spaces and thus in particular for UMD spaces ([37, Section
2.4]).
Proposition 3.7. Let X and Y be order continuous Banach function spaces and
T ⊂ Lb(X, Y ). Let rk, sk ∈ [1, ∞] for k = 0, 1. If T is ℓrk (ℓsk )-bounded for k = 0, 1,
then T is ℓrθ (ℓsθ )-bounded for all θ ∈ (0, 1), where rθ := [r0, r1]θ and sθ := [s0, s1]θ.
Moreover we have the estimate
[T ]ℓrθ (ℓsθ ) ≤ [T ]θ
ℓr0 (ℓs0 )[T ]1−θ
ℓr1 (ℓs1 ) ≤ max{[T ]ℓr0 (ℓs0 ), [T ]ℓr1 (ℓs1 )}.
Proof. This follows from Calder´on's theory of complex interpolation for order con-
tinuous vector-valued function spaces [11].
(cid:3)
Combining Proposition 3.3(iv) with Proposition 3.7 we deduce the following.
Corollary 3.8. Let X and Y be order continuous Banach function spaces and
T ⊂ Lb(X, Y ). Fix r, s ∈ [1, ∞] and suppose that T is ℓr(ℓs)-bounded. If
then T is ℓu(ℓv)-bounded with [T ]ℓu(ℓv) ≤ [T ]ℓr(ℓs).
r ≤ u ≤ v ≤ s
or
s ≤ v ≤ u ≤ r,
To end this section we present a technical lemma on the ℓr(ℓs)-boundedness of
the closure of a family of operators on spaces other than that in which the closure
was taken. It is used in our multiplier result for intermediate spaces, where several
Lebesgue spaces are used simultaneously. A similar result can be proved with
general order continuous Banach function spaces in place of Lebesgue spaces.
Lemma 3.9. Let (Ω, ρ, µ) be a metric measure space, and assume µ is finite on
bounded sets. Let p ∈ (1, ∞) and T ⊂ L(Σ(Ω), L0(Ω)) be such that T ⊂ L(Lp(Ω))
is uniformly bounded and absolutely convex. Let T denote the closure of T in
L(Lp(Ω)). Suppose q ∈ (1, ∞), and let w be a weight on Ω which is integrable
on bounded sets. Suppose also that T ⊂ L(Lq(w)) is ℓr(ℓs)-bounded for some
r, s ∈ [1, ∞]. Then T is ℓr(ℓs)-bounded on Lq(w) with [T ]ℓr (ℓs) = [T ]ℓr (ℓs).
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
9
Note that we take the closure T of T in one space, and then establish ℓr(ℓs)-
boundedness of T considered as a set of operators on a different space.
m=1,n=1 in eT and (fm,n)M,N
Proof. Fix (Tm,n)M,N
m=1,n=1 in Lq(w). By a density argument
we may assume each for each m, n that fm,n is bounded and supported on a bounded
subset of Ω, which implies fm,n ∈ Lp(Ω). For each m, n choose (T (k)
m,n)k≥1 in
T such that T (k)
m,nfm,n → Tm,nfm,n in
Lp(Ω). By passing to subsequences we may suppose that for all m, n we have
T (k)
m,nfm,n → Tm,nfm,n, µ-a.e. Therefore, by Fatou's lemma,
m,n → Tm,n in L(Lp(Ω)). Then also T (k)
(cid:13)(cid:13)(cid:13)(cid:16) MXm=1(cid:16) NXn=1
r(cid:13)(cid:13)(cid:13)Lq(w)
Tm,nfm,ns(cid:17) r
s(cid:17) 1
T (k)
≤ lim inf
k→∞ (cid:13)(cid:13)(cid:13)(cid:16) MXm=1(cid:16) NXn=1
≤ [T ]ℓr(ℓs)(cid:13)(cid:13)(cid:13)(cid:16) MXm=1(cid:16) NXn=1
r(cid:13)(cid:13)(cid:13)Lq(w)
s(cid:17) 1
m,nfm,ns(cid:17) r
r(cid:13)(cid:13)(cid:13)Lq(w)
s(cid:17) 1
fm,ns(cid:17) r
,
with the appropriate adjustment if r = ∞ or s = ∞. So T is indeed ℓr(ℓs)-bounded
on Lq(w).
(cid:3)
3.2. ℓr(ℓs)-boundedness of single operators. As noted before, a single operator
T ∈ Lb(X, Y ) can fail to be ℓr(ℓs)-bounded. For positive operators we have the
following result, which is an adaptation of [39, Lemma 4].
Proposition 3.10. Let X and Y be Banach function spaces and let P ∈ Lb(X, Y )
be a positive operator. Then P is ℓr(ℓs)-bounded for all r, s ∈ [1, ∞], and we have
the ℓr(ℓs)-bound [{P }]ℓr(ℓs) ≤ kP kL(X.Y ).
Proof. Let (xj,k)m,n
j,k=1 be a doubly-indexed sequence in X, and note that by posi-
tivity of P we may take the elements of the sequence to be positive. By positivity
of P we can estimate
(cid:13)(cid:13)(cid:13)(cid:16) mXj=1(cid:16) nXk=1
P xj,ks(cid:17)r/s(cid:17)1/r(cid:13)(cid:13)(cid:13)Y
so [{P }]ℓr (ℓs) ≤ kP kL(X,Y ).
sup
k)k
≤1
bj
k(aj
k(bj )k
sup
ℓr′
m
=(cid:13)(cid:13)(cid:13)
mXj=1
≤(cid:13)(cid:13)(cid:13)P(cid:16)
mXj=1
≤ kP kL(X,Y )(cid:13)(cid:13)(cid:13)(cid:16) mXj=1(cid:16) nXk=1
sup
ℓr′
m
k(bj )k
bj
≤1
≤1
ℓs′
n
aj
kP xj,k(cid:13)(cid:13)(cid:13)Y
nXk=1
kxj,k(cid:17)(cid:13)(cid:13)(cid:13)Y
nXk=1
xj,ks(cid:17)r/s(cid:17)1/r(cid:13)(cid:13)(cid:13)X
sup
k)k
aj
ℓs′
n
≤1
,
k(aj
(cid:3)
For an ℓ1-bounded operator on a Lebesgye space one has ℓr(ℓs)-boundedness for
all r, s ∈ [1, ∞] (see [22, Theorem 2.7.2]). The result below actually holds with
Lp(Ω) replaced by any Banach lattice X with a Levi norm (see [8] and [35, Fact
2.5]). A duality argument implies a similar result for ℓ∞-boundedness.
Proposition 3.11. Let p ∈ [1, ∞) and T ∈ L(Lp(Ω)). If T is ℓ1-bounded, then
{T } is ℓr(ℓs)-bounded for all r, s ∈ [1, ∞].
10
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Remark 3.12. Even on Lp it can be quite hard to establish the ℓr(ℓs)-boundedness
of a single operator. By using i.i.d. s-stable random variables ξ1, . . . , ξn : Ω → R
(see [33, Section 5]), for p ∈ (0, s) one can linearise the estimate by writing
(cid:16) nXj=1
T xjs(cid:17)1/s
= Cp,s(cid:13)(cid:13)(cid:13)T
nXj=1
ξjxj(cid:13)(cid:13)(cid:13)Lp(Ω)
.
By using Fubini's theorem and Minkowski's inequality, one can deduce that any
T ∈ L(Lp) is ℓr(ℓs)-bounded if p ≤ r ≤ s ≤ 2 or 2 ≤ s ≤ r ≤ p. Most of the
remaining cases seem to be open (see [30, Problem 2] and [16, Corollary 1.44]).
3.3. Non-examples. We end this section with two examples to demonstrate that
ℓr(ℓs)-boundedness is not just the conjunction of ℓr- and ℓs-boundedness. Consider
the class of kernels
K = {k ∈ L1(R) : k ∗ f ≤ M f a.e. for all simple f : R → R},
where M is the Hardy -- Littlewood maximal operator. For k ∈ K and f ∈ Lp(R)
with p ∈ (1, ∞) define an operator Tk by
Tkf (t) = R
k(t − s)f (s) ds,
and set T = {Tk : k ∈ K}.
Example 3.13. Let p ∈ (1, ∞). The family of operators T ⊂ Lb(Lp(R)) defined
above is ℓs-bounded for all s ∈ [1, ∞], but not ℓ1(ℓs)- or ℓ∞(ℓs)- bounded for any
s ∈ (1, ∞).
Proof. The ℓs-boundedness of T for s ∈ [1, ∞] is proved in [40, Theorem 4.7].
Since T = T ∗, Proposition 3.6 says that ℓ1(ℓs)-boundedness of T on Lp(R) implies
ℓ∞(ℓs′
(R), so it suffices to show that T is not ℓ∞(ℓs)-bounded
on Lp(R) for any s ∈ (1, ∞). We follow the proof of [40, Proposition 8.1].
)-boundedness on Lp′
Fix n ∈ N and for i, j ∈ N define fi,j ∈ Lp(R) by
fi,j(t) = 1(0,1](t)1(2−j ,2−j+1](t − (i − 1)2−n)
so that
(3.1)
(cid:13)(cid:13)(cid:13) sup
1≤i≤2n(cid:16) nXj=1
fi,j(t)s(cid:17)1/s(cid:13)(cid:13)(cid:13)Lp(R)
≤(cid:13)(cid:13)(cid:13) sup
1≤i≤2n
1(0,1](cid:13)(cid:13)(cid:13)Lp(R)
= 1.
Next, for i, j ∈ N define
ki,j(t) =
1
2−j+2 1(−2−j+1,2−j+1)(t)
and Ti,j = Tki,j . Then Ti,j ∈ T , as for any simple function f we have
Ti,jf (t) = ki,j ∗ f (t) =
1(−2−j+1,2−j+1)(t − τ )f (τ ) dτ(cid:12)(cid:12)(cid:12)
f (τ ) dτ(cid:12)(cid:12)(cid:12) ≤ M f (t).
1
1
=
2−j+2 (cid:12)(cid:12)(cid:12)R
2−j+2 (cid:12)(cid:12)(cid:12) t+2−j+1
2−j+2 t+2−j+1−(i−1)2−n
t−2−j+1−(i−1)2−n
t−2−j+1
1
Ti,jfi,j(t) =
1(2−j ,2−j+1](τ ) dτ
Furthermore, for any 1 ≤ j ≤ n, t ∈ (0, 1] and 1 ≤ i ≤ 2n with t ∈ ((i−1)2−n, i2−n],
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
11
1(2−j,2−j+1](τ ) dτ =
2−j
2−j+2 =
1
4
.
≥(cid:13)(cid:13)(cid:13)(cid:16) n
4s(cid:17)1/s
1(0,1](cid:13)(cid:13)(cid:13)Lp(R)
=
n1/s
4
1
≥
2−j
2−j+2 2−j+1
Ti,jfi,j(t)s(cid:17)1/s(cid:13)(cid:13)(cid:13)Lp(R)
Therefore
(cid:13)(cid:13)(cid:13) sup
1≤i≤2n(cid:16) nXj=1
which tends to ∞ as n → ∞. Combining this with (3.1) disproves the ℓ∞(ℓs)-
boundedness of T on Lp(R).
(cid:3)
The previous example can be modified to construct examples without ℓ2(ℓs)-
boundedness, by using stochastic integral operators. For k ∈ K and f ∈ Lp(R+)
with p ∈ (2, ∞), define
Skf (t) := t
0
k(t − s)
1
2 f (s) dW (s),
where W is a standard Brownian motion on a probability space (Ω, F , P). Define
S := {Sk : k ∈ K}.
Example 3.14. Let p ∈ (2, ∞). The family of operators S from Lp(R+) to Lp(R+×
Ω) is ℓr-bounded for all r ∈ [2, ∞), but not ℓ2(ℓr)-bounded for any r ∈ (2, ∞).
Proof. Let r ∈ [2, ∞) and X = ℓr. Take f ∈ Lp(R+; X) and k ∈ L1(R+; X) such
that kj ∈ K for all j ∈ N. By [41, Corollary 2.10] and the Kahane -- Khintchine
inequalities (see for example [33]), we know that
(cid:16)E(cid:13)(cid:13)(cid:13) t
0
k(t − s)
1
2 f (s) dW (s)(cid:13)(cid:13)(cid:13)
p
X(cid:17)1/p
≃(cid:13)(cid:13)(cid:13)(cid:16) t
0
2(cid:13)(cid:13)(cid:13)X
k(t − s)f (s)2 ds(cid:17) 1
for any t ∈ R+. This implies that S is ℓr-bounded from Lp(R+) to Lp(R+ × Ω) if
and only if T restricted to R+ is ℓr/2-bounded on Lp/2(R+), so S is ℓr-bounded for
all r ∈ [2, ∞) by Example 3.13. Repeating the argument with X = ℓ2(ℓr), we also
get from Example 3.13 that S is not ℓ2(ℓr)-bounded for any r ∈ (2, ∞).
(cid:3)
4. The function spaces V s(J ; Y ) and Rs(J ; Y )
The multipliers we consider are members of the space of functions of bounded
s-variation, which we denote by V s(J , Y ) for s ≥ 1. This space contains the class
of 1/s-Holder continuous functions. In our arguments we will also use the atomic
function space Rs(J , Y ), which was introduced in the scalar case in [14].
Definition 4.1.
(i) Let Y be a Banach space, J = [J−, J+] ⊂ R a bounded interval and s ∈
[1, ∞). A function f : R → Y is said to be of bounded s-variation on J, or
f ∈ V s(J; Y ), if
kf kV s(J;Y ) := kf k∞ + [f ]Vs(J;Y ) < ∞,
where
[f ]Vs(J;Y ) :=
J−=t0<···<tN =J+(cid:16) NXi=1
sup
kf (ti−1) − f (ti)ks
Y(cid:17)1/s
.
Furthermore we define V ∞(J; Y ) = L∞(J; Y ).
12
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
(ii) When J is a collection of mutually disjoint bounded intervals in R, the space
V s(J ; Y ) ⊂ L∞(R; Y ) consists of all f ∈ L∞(R; Y ) such that
kf kV s(J ;Y ) := sup
J∈J
kf J kV s(J;Y ) < ∞.
If J = (Jk)k∈N is ordered, we define V s
subspace consisting of f ∈ V s(J ; Y ) with limk→∞kf Jk kV s(J;Y ) = 0.
0 (J ; Y ) ⊂ V s(J ; Y ) to be the closed
Clearly V s(J ; Y ) ֒→ V t(J ; Y ) contractively when 1 ≤ s ≤ t ≤ ∞, and V s(J ; Y )
is complete when Y is complete.
In our applications the space Y is usually the span of a bounded and abso-
lutely convex subset B of a normed space Z (i.e. a disc in Z), equipped with the
Minkowski norm
kxkB := inf{λ > 0 : x
λ ∈ B},
and we write V s(J ; B) := V s(J ; span B). Clearly kxkZ ≤ CBkxkB for x ∈ Y . If
the Minkowski norm on span B is complete, then B is called a Banach disc. If Z
is a Banach space and B is closed, then B is a Banach disc [42, Proposition 5.1.6],
but this is not a necessary condition [42, Proposition 3.2.21].
Definition 4.2.
(i) Let Y be a normed space, J ⊂ R a bounded interval, and s ∈ [1, ∞). Say
that a function a : J → Y is an Rs(J; Y )-atom, written a ∈ Rs
at(J; Y ), if
there exists a set I of mutually disjoint subintervals of J and a set of vectors
(cI )I∈I ⊂ Y such that
a =XI∈I
cI 1I
and (cid:16)XI∈I
kcI ks
Y(cid:17)1/s
≤ 1.
Define Rs(J; Y ) ⊂ L∞(J; Y ) by
Rs(J; Y ) :=nf ∈ L∞(J; Y ) : f =
where the series f = P∞
Rs(J; Y ) by
∞Xk=1
λkak, (λk) ∈ ℓ1, (ak) ⊂ Rs
at(J; Y )o,
k=1 λkak converges in L∞(J; Y ). Define a norm on
kf kRs(J;Y ) := infnkλkkℓ1 : f =
∞Xk=1
λkak as aboveo.
Furthermore we define R∞(J; Y ) := L∞(J; Y ).
(ii) When J is a collection of mutually disjoint bounded intervals in R, the space
Rs(J ; Y ) ⊂ L∞(R; Y ) consists of all f ∈ L∞(R; Y ) such that
kf kRs(J ;Y ) := sup
J∈J
kf J kRs(J;Y ) < ∞.
If J = (Jk)k∈N is ordered, we define Rs
subspace consisting of f ∈ Rs(J ; Y ) with limk→∞kf JkkRs(Jk;Y ) = 0.
0(J ; Y ) ⊂ Rs(J ; Y ) to be the closed
Clearly Rs(J ; Y ) ֒→ Rt(J ; Y ) contractively when 1 ≤ s ≤ t ≤ ∞, and Rs(J ; Y )
is complete when Y is complete. As with the classes V s, when B is a disc in a
normed space Z, we put the Minkowski norm on the linear span of B and write
Rs(J ; B) := Rs(J ; span B).
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
13
For α ∈ (0, 1] and an interval J ⊂ R we let Cα(J; Y ) denote the space of α-Holder
continuous functions with kf kC α(J;Y ) = max{kf k∞, [f ]C α(J;Y )}, where
[f ]C α(J;Y ) := sup
x,y∈J
kf (x) − f (y)kY
x − yα
.
Lemma 4.3. Let s ∈ [1, ∞), let Y be a Banach space and fix a bounded interval
J ⊂ R.
(i) If q ∈ (s, ∞), then Rs(J; Y ) ⊂ V s(J; Y ) ⊂ Rq(J; Y ) and for all f ∈ L∞(J; Y )
we have
kf kRq(J;Y ) .q,s kf kV s(J;Y ) . kf kRs(J;Y ).
(ii) We have C1/s(J; Y ) ⊂ V s(J; Y ), and for all f ∈ V s(J; Y ),
kf kV s(J;Y ) ≤ kf k∞ + J1/s[f ]C 1/s(J;Y ).
Proof. For part (i) we note that both Rs(J; Y ) ⊂ V s(J; Y ) and the second norm
estimate follow directly from the fact that for any atom a ∈ Rs
at(J; Y ) with
cI 1I
a =XI∈I
kcI − cJ ks(cid:17)1/s
≤ 1 + 2(cid:16)XI∈I
kcI ks(cid:17)1/s
≤ 3.
we have by Minkowski's inequality that
kakV s(J;Y ) ≤ sup
I∈I
kcI kY +(cid:16) XI,J∈I
I6=J
The embedding V s(J; Y ) ⊂ Rq(J; Y ) with the first norm estimate is shown in [14,
Lemme 2] for scalar functions, and the argument extends to the general case. Part
(ii) is straightforward to check.
(cid:3)
We end this section with complex interpolation containments for the V s- and
Rs-classes. It is an open problem whether complex interpolation of the V s-classes
as below can be proved with ε = 0 (see [43, Chapter 12]). It is also not clear whether
converse inclusions hold, but since we don't need them we leave the question open.
Theorem 4.4. Suppose 1 ≤ q0 ≤ q1 ≤ ∞, θ ∈ (0, 1), ε > 0 and let Y be a Banach
space. Then for all bounded intervals J ⊂ R we have continuous inclusions
(4.1)
(4.2)
V [q0,q1]θ −ε(J; Y ) ֒→ [V q0 (J; Y ), V q1 (J; Y )]θ,
R[q0,q1]θ (J; Y ) ֒→ [Rq0 (J; Y ), Rq1 (J; Y )]θ,
q1 6= ∞.
Furthermore, if J = (Jk)k∈N is an ordered collection of mutually disjoint bounded
intervals in R, then we have continuous inclusions
(4.3)
(4.4)
V [q0,q1]θ−ε
0
R[q0,q1]θ
0
(J ; Y ) ֒→ [V q0
(J ; Y ) ֒→ [Rq0
0 (J ; Y ), V q1
0 (J ; Y ), Rq1
0 (J ; Y )]θ
0 (J ; Y )]θ,
q1 6= ∞.
Proof. For q0 = 1 and q1 = ∞ we have (4.1) by applying subsequently [43, Lemma
12.11], [3, Theorem 3.4.1], and [3, Theorem 4.7.1],
V [q0,q1]θ −ε(J; Y ) ֒→(cid:0)V 1(J; Y ), L∞(J; Y )(cid:1)θε,∞
֒→(cid:0)V 1(J; Y ), L∞(J; Y )(cid:1)θ,1
֒→(cid:2)V 1(J; Y ), L∞(J; Y )(cid:3)θ
14
with
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
θε = 1 −
1
1
1−θ − ε
< θ.
The intermediate cases follow from the reiteration theorem for complex interpola-
tion [3, Theorem 4.6.1].
In the remainder of the proof we will need the following notation: when Ik is
a collection of intervals for each k ∈ N and I ∈ Ik, let πI,k denote the canonical
projection ℓ∞(Ik; Y ) → Y . We abbreviate Banach couples (X0, X1) by X•, and
use this shorthand for expressions like
[ℓp•(N; X)]θ = [ℓp0 (N; X), ℓp1(N; X)]θ.
We let F (X•) denote the space of bounded analytic functions from the closed strip
S := {z ∈ C : ℜz ∈ [0, 1]} to the sum X0 + X1 whose restrictions to the sets
{z ∈ C : ℜz = 0} and {z ∈ C : ℜz = 1} map continuously into X0 and X1
respectively, equipped with the norm
kF kF (X•) := max(cid:18)sup
t∈R
kF (it)kX0
, sup
t∈R
kF (1 + it)kX1(cid:19)
as in [3, §4.1].
For (4.2) let 1 ≤ q0 ≤ q1 ≤ ∞ and write qθ := [q0, q1]θ for brevity. Suppose
f ∈ Rqθ (J; Y ), with atomic decomposition
f =
∞Xk=1
λkak =
∞Xk=1
λk XI∈Ik
1I πI,k(ck),
where ck ∈ ℓqθ (Ik; Y ) for each k ∈ N.
Let ε > 0. For each k ∈ N we have ℓqθ (Ik; Y ) = [ℓq• (Ik; Y )]θ with equal
norms [49, Theorem 1.18.1], hence there exists a function Ck ∈ F (ℓq• (Ik; Y )) with
Ck(θ) = ck and kCkkF (ℓq• (Ik;Y )) ≤ (1 + ε)kckkℓqθ (Ik;Y ) ≤ 1 + ε. For all z ∈ S and
t ∈ J, define
Ak(z)(t) := XI∈Ik
1I (t)πI,k(Ck(z)),
noting that for each t there is at most one non-zero term in the sum. It follows
from kCkkF (ℓq• (Ij ;Y )) ≤ 1 + ε that kAkkF (Rq• (J;Y )) ≤ 1 + ε for all z ∈ S.
We will show that each Ak : S → Rq0(J; Y ) + Rq1 (J; Y ) is analytic on S, using
that Rq0 (J; Y ) + Rq1(J; Y ) = Rq1(J; Y ) and ℓq0 (Ik; Y ) + ℓq1(Ik; Y ) = ℓq1(Ik; Y ).
Fix z0 ∈ S. Since Ck is analytic with values in ℓq1(Ik; Y ), there exists a Taylor
expansion
Ck(z) =
∞Xn=0
(z − z0)nβk,n
for z in a neigbourhood of z0, where (βk,n)∞
Thus for such z we have
n=0 ⊂ ℓq0 (Ik; Y ) is a bounded sequence.
Ak(z) = XI∈Ik
1I πI,k(Ck(z)) =
∞Xn=0
(z − z0)n XI∈Ik
1I πI,k(βk,n) =:
∞Xn=0
(z − z0)nγk,n
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
15
using the mutual disjointness of Ik to interchange the sums. The functions γk,n are
in Rq1 (J; Y ) as we can write
kγk,nkRq1 (J;Y ) =(cid:13)(cid:13)(cid:13)XI∈Ik
1IπI,k(βk,n)(cid:13)(cid:13)(cid:13)Rq1 (J;Y )
Similarly we can show that each Ak : S → Rq1 (J; Y ) is continuous.
≤ kβk,nkℓq1 (Ik;Y ) < ∞.
Now for z ∈ S and j ∈ N define
F (z) :=
∞Xk=1
λkAk(z).
Since the functions Ak : S → Rq0 (J; Y ) + Rq1 (J; Y ) are bounded uniformly in k,
continuous on S, and analytic on S, and since λ ∈ ℓ1(N), and each Ak maps into
Rq0(J; Y ) + Rq1 (J; Y ), we find that each F ∈ F (Rq• (J; Y )). Furthermore we have
F (θ) =
∞Xk=1
λkAk(θ) =
∞Xk=1
λk XI∈Ik
1I πI,k(Ck(θ)) = f
and
kF kF (Rq• (J;Y )) ≤ kλkkℓ1(N) sup
k∈N
kAkkF (Rq• (J;Y )) ≤ (1 + ε)kλkkℓ1(N).
Since ε > 0 was arbitrary, taking the infimum over all atomic decompositions of f
and all possible F ∈ F (Rq• (J; Y )) with F (θ) = f completes the proof.
Now consider a collection J of mutually disjoint bounded intervals in R. We
will only prove (4.3), as the proof of (4.4) is similar. We introduce the following
if J = [J−, J+) ⊂ R is a bounded interval and f ∈ L0(J; Y ), we let
notation:
fJ ∈ L0([0, 1); Y ) be the function
fJ (x) := f ((J+ − J−)x + J+)
x ∈ [0, 1).
Then for each s ∈ [1, ∞] the map τJ : V s(J; Y ) → V s([0, 1); Y ) defined by τJ (f ) :=
fJ is an isometry. Consequently we can write
kf kV s(J ;Y ) = sup
J∈J
0 (J ; Y ) → c0(J ; V s([0, 1); Y )) defined by
and therefore the map Φ : V s
kf J kV s(J;Y ) = sup
J∈J
kτJ (f J )kV s([0,1);Y ),
Φ(f ) := (τJ (f J ))J∈J
is an isometry. Since the intervals in J are mutually disjont, Φ is an isometric
isomorphism. Thus Φ−1 induces an isometric isomorphism
Φ−1 : c0(cid:0)J ; [V q• ([0, 1); Y )]θ(cid:1) =(cid:2)c0(J ; V q• ([0, 1); Y ))(cid:3)θ → [V q•
using [49, Remark 3, §1.18.1]. By (4.1) we have
0 (J ; Y )]θ,
V [q0,q1]θ−ε([0, 1); Y ) ֒→ [V q• ([0, 1); Y )]θ,
so that Φ−1 yields an embedding
c0(J ; V [q0,q1]θ −ε([0, 1); Y )) ֒→ [V q•
0 (J ; Y )]θ.
Precomposing with Φ gives the bounded inclusion
(J ; Y ) ֒→ [V q•
V [q0,q1]θ−ε
0
0 (J ; Y )]θ
and completes the proof.
(cid:3)
16
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
5. Fourier multipliers
The Fourier transform and operator-valued Fourier multipliers on vector-valued
functions are defined similarly to the scalar-valued case. Here we just mention that
our normalisation of the Fourier transform is
bf (ξ) = F f (ξ) := Rd
f (t)e−2πit·ξ dt,
f ∈ L1(Rd; X), ξ ∈ Rd,
and that since S(Rd) ⊗ X is dense in Lp(w; X) for every p ∈ (1, ∞) and w ∈ A∞
(see [20, Ex. 9.4.1] for the scalar case), the Lp(w; X) → Lp(w; Y )-boundedness of
a Fourier multiplier Tm : S(Rd) ⊗ X → S′(Rd; Y ) reduces to the estimate
kTmf kLp(w;Y ) . kf kLp(w;X),
f ∈ S(Rd) ⊗ X.
Our goal is to find conditions on Banach function spaces X and Y which imply
this estimate for m ∈ V s(∆; L(X, Y )) and w in a suitable Muckenhoupt class. We
will only consider multipliers m defined on R; extensions to multipliers defined on
Rd can be obtained by an induction argument as in [27, Section 4], [32] and [52],
and extensions to multipliers on the torus T can be obtained by transference, see
[1, Proposition 4.1]. In this case one must consider multipliers defined on T = Z,
where bounded s-variation for a function on Z is defined analogously to Definition
4.1.
We start with a result that is well-known in the unweighted setting (see [21, 48]).
This is not so important to our main results; it will only be used in the proof
of Theorem 5.18. Recall that ∆ = {±[2k, 2k+1), k ∈ Z} is the standard dyadic
partition of R.
Theorem 5.1 (Vector-valued Marcinkiewicz multiplier theorem). Let X and Y
be UMD Banach spaces, and suppose T ⊂ Lb(X, Y ) is absolutely convex and R-
bounded. Suppose m ∈ V 1(∆; T ). Then for all p ∈ (1, ∞) and w ∈ Ap,
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φX,Y,p([w]Ap )[T ]RkmkV 1(∆;T ).
Proof. To prove the result one can repeat the argument in [21, Theorem 4.3] using
weighted Littlewood -- Paley inequalities with sharp cut-off functions, which can be
found for instance in [17] (see also [34]).
(cid:3)
Our starting point for multiplier theorems for m ∈ V s with s > 1 is an estimate
of Littlewood -- Paley -- Rubio de Francia type. For an interval I ⊂ R let SI denote
the Fourier projection onto I, defined by SI f := (1I f )∨ for Schwartz functions
f ∈ S(R) ⊗ X. The following result was obtained in [1, Theorem 6.5]. Related
results have been obtained in [27, 45].
Theorem 5.2. Suppose q ∈ [2, ∞) and let X be a Banach function space such that
X q′
∈ UMD. Let I be a collection of mutually disjoint intervals in R. Then for all
p > q′, all w ∈ Ap/q′ , and all f ∈ Lp(w; X),
(cid:13)(cid:13)(cid:13)(cid:16)XJ∈∆(cid:16)XI∈I
I⊂J
SI f q(cid:17)2/q(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;X)
≤ φX,p,q([w]Ap/q′ )kf kLp(w;X).
For Hilbert spaces the following variant holds (see [1, Proposition 6.6 and Remark
6.7]).
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
17
Proposition 5.3. Suppose q ∈ [2, ∞) and let X be a Hilbert space. Let I be a
collection of mutually disjoint intervals in R. Then for all p > q′, all w ∈ Ap/q′
and all f ∈ Lp(w; X),
(cid:13)(cid:13)(cid:13)(cid:16)XJ∈∆(cid:16)XI∈I
I⊂J
kSI f kq
X(cid:17)2/q(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w)
≤ φp,q([w]Ap/q′ )kf kLp(w;X).
5.1. Multipliers in Hilbert spaces. The first part of the following theorem is
an analogue of [27, Theorem A(i)], and the second part is an unweighted analogue
of [27, Theorem A(ii)]. The second part is also proved in [24, Proposition 3.3]. The
exponents (p, s) for which each part of the theorem applies are pictured in Figure
1.
Theorem 5.4. Let X and Y be Hilbert spaces, p, s ∈ (1, ∞), and consider a mul-
tiplier m ∈ V s(∆; Lb(X, Y )).
(i) If s ≤ 2 and p ≥ s, then for all w ∈ Ap/s we have
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φp,s([w]Ap/s)kmkV s(∆;Lb(X,Y )).
(ii) If 1
p − 1
kTmkL(Lp(R;X),Lp(R;Y )) .p,s kmkV s(∆;Lb(X,Y )).
s >(cid:12)(cid:12) 1
2(cid:12)(cid:12) we have
Figure 1. Allowable exponents for Theorem 5.4: the weighted
case (i) dark shaded, the unweighted case (ii) light shaded.
1/s
1
1
2
0
0
1
2
1/p
1
1
To prove Theorem 5.4 we use the following proposition, which is a version of
the first part for R-class multipliers. The techniques used to prove this proposition
are strongly related to those used in the proof of our main result for UMD Banach
function spaces, Theorem 5.8.
18
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Proposition 5.5. Let X and Y be Hilbert spaces, s ∈ (1, 2], and consider a mul-
tiplier m ∈ Rs(∆; Lb(X, Y )). Then for all p > s and w ∈ Ap/s we have
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φp,s([w]Ap/s )kmkRs(∆;Lb(X,Y )).
Proof. We only consider the case s < 2. The case s = 2 is similar, but simpler.
Fix ε > 0 and let f ∈ Lp(w; X). By approximation we may assume that the
dyadic Littlewood -- Paley decomposition of f has finitely many nonzero terms and
set ∆f = {J ∈ ∆ : SJ f 6= 0}. For each J ∈ ∆f let
mJ =
NXk=1
λkaJ
k ,
aJ
k = XI∈J J
k
cJ,k
I 1I
be an Rs(J; Lb(X, Y ))-atomic decomposition of the restriction mJ with λk inde-
pendent of J and
NXk=1
λk ≤ (1 + ε)kmkRσ(∆;Lb(X,Y ))
as in [24, Theorem 2.3].
Note that SJ Tm = TmSJ , where we abuse notation by letting SJ denote either
the X- or Y -valued Fourier projection. By the Littlewood -- Paley estimate (see [38,
Proposition 3.2]), Holder's inequality, Proposition 5.3, and w ∈ Ap/s ⊂ Ap, we have
kTmf kLp(w;Y ) ≤ φp([w]Ap )(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f
≤ φp([w]Ap )(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:16) NXk=1
≤ φp([w]Ap )
≤ φp([w]Ap )
≤ φp,s([w]Ap/s)
kcJ,k
kTmSJ f k2
Y(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w)
λk XI∈J J
λk(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:0) XI∈J J
NXk=1
λk(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:16) XI∈J J
NXk=1
NXk=1
λk kf kLp(w;X).
kcJ,k
kSI f ks′
k
k
k
I SI f kY(cid:17)2(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w)
2(cid:13)(cid:13)(cid:13)Lp(w)
s′(cid:17) 1
s(cid:0) XI∈J J
X(cid:1) 2
I ks(cid:1) 2
X(cid:17)2/s′(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w)
kSI f ks′
k
Since ε > 0 was arbitrary this implies
kTmf kLp(w;Y ) ≤ φp,s([w]Ap/s)kmkRs(∆;Lb(X,Y ))kf kLp(w;X)
for all w ∈ Ap/s and f ∈ Lp(w; X).
(cid:3)
Proof of Theorem 5.4. Part (i): We first consider the case s < p and s < 2. Let
w ∈ Ap/s and take σ ∈ (s, 2] such that w ∈ Ap/σ, which is possible by Proposition
2.1(ii). By Lemma 4.3 we know that m ∈ Rσ(∆; Lb(X, Y )) with
kmkRσ (∆;Lb(X,Y )) .s,σ kmkV s(∆;Lb(X,Y )),
so by Proposition 5.5 we obtain
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φp,s([w]Ap/s)kmkV s(∆;Lb(X,Y )).
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
19
Next we consider the case p > s = 2. Observe that by [22, Proposition 5.3.16] it
suffices to prove the result for the truncated multipliers
mN := 1SN
n=1 Jn m,
n=1 is an arbitrary ordering of ∆. Since mN ∈ V s
where ∆ = (Jn)∞
0 (∆; Lb(X, Y ))
uniformly, without loss of generality we may work with an arbitrary decaying mul-
tiplier m ∈ V s
0 (∆; Lb(X, Y )). Fix w ∈ Ap/2. Then by Proposition 2.1(iii) there
exists a δ > 0 such that w1+δ ∈ Ap/2. Take
θ =
p0 = (1 + δ)(1 − θ)p,
and σ = 2 − θ.
2
p(cid:16)1 −
1
1 + δ(cid:17),
Then θ ∈ (0, 1), σ ∈ (1, 2) and p0 = p + (p − 2)δ > p, so by the first case we have
kTmkL(Lp0 (w;X),Lp0 (w;Y )) ≤ φp0,σ([w]Ap/2 )kmkV σ
0 (∆;Lb(X,Y )).
Moreover by Plancherel's theorem (which is valid since X and Y are Hilbert spaces)
we know that
(5.1)
Since
kTmkL(L2(R;X),L2(R;Y )) ≤ kmkL∞(R;Lb(X,Y )).
1
[p0, 2] θ
=
1
p(1 + δ)
+
1
p
−
1
p(1 + δ)
=
1
p
,
we know by [49, Theorem 1.18.5] that Lp(w; X) = [Lp0(w1+δ, X), L2(R; X)]θ, and
likewise with X replaced by Y . Moreover since [σ, ∞]θ = 2−θ
1−θ > 2 we have the
continuous inclusions
V 2(∆; Lb(X, Y )) ֒→ [V σ
֒→ [V σ
0 (∆; Lb(X, Y )), V ∞
0 (∆; Lb(X, Y ))]θ
0 (∆; Lb(X, Y )), L∞(R; Lb(X, Y ))]θ
by Theorem 4.4. By bilinear complex interpolation [3, §4.4] applied to the bilinear
map (m, f ) 7→ Tmf we have boundedness of Tm : Lp(w; X) → Lp(w; Y ) with the
required norm estimate.
Finally we consider the case p = s ≥ 2; we will use another interpolation ar-
gument. Fix w ∈ A1. Then by Proposition 2.1(iii) there exists a δ > 0 such that
w1+δ ∈ A1. Fix p1 ∈ (s, s + (s − 1)δ). By the argument of the previous cases we
have
kTmkL(Lp1 (w1+δ;X),Lp1 (w1+δ;Y )) ≤ φp1,s([w]A1 )kmkV s(∆;Lb(X,Y )).
Let θ ∈ (0, 1) be such that θ(1 + δ)s = p1; such a θ exists since p1 < s + (s − 1)δ.
Choose p0 ∈ (1, s) such that [p0, p1]θ = s. Such a p0 exists since p1 > s and
[1, p1]θ < s. Indeed, the latter follows from
s
[1, p1]θ
= s(1 − θ) + s
θ
p1
= s −
p1
1 + δ
+
1
1 + δ
> 1.
Since p0 < s ≤ 2 we have by duality with the previous cases (taking w = 1) that
kTmkL(Lp0 (R;X),Lp0 (R;Y )) .p0,s kmkV s(∆;Lb(X,Y )).
As before our choice of θ yields Ls(w; X) = [Lp0(R, X), Lp1(w1+δ; X)]θ, and likewise
with X replaced by Y . Therefore by complex interpolation we have boundedness
of Tm : Ls(w; X) → Ls(w; Y ) with the required norm estimate.
Part (ii): The case p = 2 is clear from (5.1) and the embedding of the V s
classes in L∞. For p > 2 we may assume without loss of generality that m ∈
20
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
V s
0 (∆; Lb(X, Y )) as in part (i). Moreover, by embedding of the V s classes, we may
assume that s > 2.
= p. Such a t exists
Let σ ∈(cid:0)s,(cid:0) 1
since p > 2 and
2 − 1
p(cid:1)−1(cid:1) and fix t ∈ (2, ∞) such that [2, t] σ
2
1
p
=
1
[2, t] σ
2
=
1
2
−
1
σ
+
2
σ
1
t
,
which implies that
1
t
=
2
s(cid:16) 1
p
+
1
σ
−
1
2(cid:17) > 0.
Using the boundedness properties
V ∞
0 (∆; Lb(X, Y )) × L2(R; X) → L2(R; Y )
V 2
0 (∆; Lb(X, Y )) × Lt(R; X) → Lt(R; Y )
and
of the bilinear map (m, f ) 7→ Tmf , which follow from (5.1) and part (i) respectively,
we have boundedness of Tm : Lp(w; X) → Lp(w; Y ) with the required norm estimate
by bilinear complex interpolation [3, §4.4]. Here we use [49, Theorem 1.18.4] and
Theorem 4.4 to identify the interpolation spaces as before. The case p < 2 follows
by a duality argument.
(cid:3)
Remark 5.6.
(1) If the multiplier is scalar-valued and X = Y , then Theorem 5.4 follows
simply from the scalar case and a standard Hilbert space tensor extension
argument (see [22, Theorem 2.1.9]).
(2) As in [27, Theorem A], a weighted version of Theorem 5.4(ii) can be proved,
but we omit it to prevent things from getting too complicated.
5.2. Multipliers in UMD Banach function spaces. We now turn to our main
result (Theorem 5.8). Its proof is inspired by that of [24, Theorem 2.3], which is
a generalisation of the Hilbert space result in Theorem 5.4. Besides the regularity
assumption on the multiplier as in the Hilbert space case, we will need an ℓ2(ℓq)-
boundedness assumption. We first prove a result for R-class multipliers, analogous
to Proposition 5.5.
Proposition 5.7. Let q ∈ (1, 2], p ∈ (q, ∞), and w ∈ Ap/q. Let X and Y be
Banach function spaces with X q ∈ UMD and Y ∈ UMD. Let T ⊂ Lb(X, Y ) be
absolutely convex and ℓ2(ℓq′
)-bounded, and suppose m ∈ Rq(∆; T ). Then
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φX,Y,p,q([w]Ap/q )[T ]ℓ2(ℓq′ )kmkRq(∆;T ).
Proof. Fix ε > 0 and let f ∈ Lp(w; X). We begin as in the proof of Proposition
5.5, which began as in the proof of [24, Theorem 2.3]: we assume that the dyadic
Littlewood -- Paley decomposition of f has finitely many nonzero terms and set ∆f =
{J ∈ ∆ : SJ f 6= 0}. For each J ∈ ∆f let
mJ =
NXk=1
λkaJ
k ,
aJ
k = XI∈J J
k
cJ,k
I 1I
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
21
be a Rq(J; T )-atomic decomposition of the restriction mJ with λk independent of
J, with each J J
k finite, and with
NXk=1
λk ≤ (1 + ε)kmkRq (∆;Lb(X,Y )).
As before, SJ Tm = TmSJ . By the Littlewood -- Paley theorem for UMD Banach
function spaces (see [1, Proposition 6.1]), using that Y ∈ UMD and w ∈ Ap/q ⊂ Ap,
we have
We estimate the sum on the right hand side by
k
k
cJ,k
cJ,k
TmSJ f 2(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;Y )
I SI f(cid:12)(cid:12)(cid:12)
NXk=1
λk XI∈J J
λk(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:12)(cid:12)(cid:12) XI∈J J
I SI f(cid:12)(cid:12)(cid:12)
2(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;Y )
I kT (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
T(cid:1)1/q(cid:0) XI∈J J
q′(cid:17)2/q′(cid:19)1/2(cid:13)(cid:13)(cid:13)(cid:13)Lp(w;Y )
SI f(cid:12)(cid:12)(cid:12)
cJ,k
I SI f
kcJ,k
.
k
2(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;Y )
2(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;Y )
.
q′(cid:1)1/q′(cid:17)2(cid:19)1/2(cid:13)(cid:13)(cid:13)(cid:13)Lp(w;Y )
kTmf kLp(w;Y ) ≤ φY,p([w]Ap )(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f
= φY,p([w]Ap )(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:12)(cid:12)(cid:12)
NXk=1
≤ φY,p([w]Ap )
k
≤
cJ,k
I SI f(cid:12)(cid:12)(cid:12)
λk(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:12)(cid:12)(cid:12) XI∈J J
NXk=1
λk(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18) XJ∈∆f(cid:16)(cid:0) XI∈J J
NXk=1
λk(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18) XJ∈∆f(cid:16) XI∈J J
(cid:12)(cid:12)(cid:12)
NXk=1
≤
k
k
kcJ,k
I kq
cJ,k
I
kcJ,k
I kT
By the definition of the Minkowski norm, the operators cJ,k
so by ℓ2(ℓq′
)-boundedness of T we have
I /kcJ,k
I kT all lie in T ,
kTmf kLp(w;Y )
≤ φY,p([w]Ap )[T ]ℓ2(ℓq′ )
By Theorem 5.2,
NXk=1
λk(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:0) XI∈J J
k
SI f q′(cid:1)2/q′(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;X)
.
(cid:13)(cid:13)(cid:13)(cid:16) XJ∈∆f(cid:0) XI∈J J
k
SI f q′(cid:1)2/q′(cid:17)1/2(cid:13)(cid:13)(cid:13)Lp(w;X)
≤ φX,p,q([w]Ap/q )kf kLp(w;X).
k=1 λk ≤ (1 + ε)kmkRq(∆;T ) and ε > 0 was arbitrary, this finishes the
(cid:3)
Since PN
proof.
Our main multiplier theorem follows easily. Recall that w ∈ αp,q if and only if
w1−p′
∈ Ap′/q′ with [w]αp,q := [w1−p′
]Ap′/q′ .
Theorem 5.8. Let X and Y be Banach function spaces, and let T ⊂ Lb(X, Y ) be
absolutely convex. Let q ∈ (1, 2], s ∈ [1, q) and m ∈ V s(∆; T ).
22
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
(i) Suppose that X q ∈ UMD, Y ∈ UMD, and T is ℓ2(ℓq′
)-bounded. Then for all
p ∈ (q, ∞) and w ∈ Ap/q we have
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φX,Y,p,q([w]Ap/q )[T ]ℓ2(ℓq′ )kmkV s(∆;T ).
(ii) Suppose that X ∈ UMD, (Y ∗)q ∈ UMD, T is ℓ2(ℓq)-bounded, and m ∈
V s(∆; T ). Then for all p ∈ (1, q′) and w ∈ αp,q′ we have
kTmkL(Lp(w;X),Lp(w;Y )) ≤ φX,Y,p,q([w]αp,q )[T ]ℓ2(ℓq)kmkV s(∆;T ).
Proof. The first part follows directly from Proposition 5.7 and Lemma 4.3. For the
second part a standard duality argument shows that
kTmkL(Lp(w;X),Lp(w;Y )) ≤ kTm∗ kL(Lp′ (w1−p′ ;Y ∗),Lp′ (w1−p′ ;X ∗)),
with m∗ : R → span(T ∗) defined by m∗(t) = m(t)∗ for all t ∈ R. Applying the first
part to m∗, using Proposition 3.6 to show that T ∗ is ℓ2(ℓq′
)-bounded and noting
that m∗ ∈ V q(∆; T ∗), completes the proof.
(cid:3)
If q = 2 and w = 1 in Theorem 5.8, we recover [24, Corollary 2.5] for Banach
function spaces, except for the endpoint p = 2, which is missing since we worked in
the weighted setting. If the multiplier is scalar-valued and X = Y , the result was
proved in [1] using vector-valued extrapolation.
Remark 5.9. The ℓ2(ℓq′
)-boundedness assumption in Theorem 5.8 arises naturally
from the proof. It is known that boundedness of Tm implies R-boundedness -- and
thus ℓ2-boundedness if X has finite cotype -- of the image of the Lebesgue points of
m (see [13] or [22, Theorem 5.3.15]). However, ℓ2(ℓq′
)-boundedness is not necessary,
as may be seen by considering m = nS where n ∈ Rq(∆) is a scalar multiplier and
S : X → Y is a bounded linear operator. In this case Tm will be bounded, but {S}
need not be ℓ2(ℓq′
)-bounded for q 6= 2 (see [28, Example 2.16]).
Using complex interpolation, the reverse Holder inequality, and the openness of
the UMD property, we can obtain a result for the endpoint p = q = s in Theorem
5.8.
Proposition 5.10. Let X and Y be Banach function spaces. Let q, r ∈ (1, 2)
and suppose that X q ∈ UMD and (Y ∗)r ∈ UMD. Let T ⊂ L(X, Y ) be absolutely
convex and both ℓ2(ℓq′
)- and ℓ2(ℓr)-bounded. Let s = min{q, r} and suppose that
m ∈ V s(∆; T ). Then for all w ∈ A1,
kTmkL(Lq(w;X),Lq(w;Y )) ≤ φX,Y,q,r([w]A1 ) max{[T ]ℓ2(ℓq′ ), [T ]ℓ2(ℓr)}kmkV s(∆;T ).
Proof. Fix w ∈ A1, so that by Proposition 2.1(iii) there exists an δ > 0 such
that w1+δ ∈ A1. By the openness of the UMD property we know that there exist
q0 ∈ (q, q + (q − 1)δ) and r0 ∈ (r, ∞) such that X q0, (Y ∗)r0 ∈ UMD. By Corollary
3.8 we know that T is ℓ2(ℓq′
0)- and T is ℓ2(ℓr′
0)-bounded with
(5.2)
[T ]ℓ2(ℓq′
0 ) ≤ [T ]ℓ2(ℓq′ )
and [T ]ℓ2(ℓr′
0 ) ≤ [T ]ℓ2(ℓr′ ).
Fix p1 ∈ (q0, q + (q − 1)δ). By Theorem 5.8(i) and (5.2) we know that
kTmkL(Lp1 (w1+δ;X),Lp1 (w1+δ ;Y )) ≤ φX,Y,p1,q0 ([w]A1 )[T ]ℓ2(ℓq′ )kmkV s(∆;T ).
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
23
Let θ ∈ (0, 1) be such that θ(1 + δ)q = p1, and fix p0 ∈ (1, q) such that [p0, p1]θ = q.
These parameters exist by the same argument as in Theorem 5.4(i). Since p0 < r′
0,
we know by Theorem 5.8(ii) and (5.2) that
kTmkL(Lp0 (R;X),Lp0 (R;Y )) .X,Y,p0,r0 [T ]ℓ2(ℓr)kmkV s(∆;T ).
Therefore by complex interpolation as in Theorem 5.4(i) we have boundedness of
Tm : Lq(w; X) → Lq(w; Y ) with the required norm estimate.
(cid:3)
When dealing with operator-valued multipliers m, to check the hypotheses of our
results, one needs an ℓ2(ℓq′
)-bounded subset T ⊂ Lb(X, Y ) whose span contains
m(R), such that m has the appropriate regularity when measured with respect to
the Minkowski norm induced by T . An obvious naıve choice is to assume that m(R)
is ℓ2(ℓq′
)-bounded and to take T = m(R), but m may not be sufficiently regular
with respect to the T -Minkowski norm. By making T larger m becomes more
regular in the T -Minkowski norm, but enlarging T may violate ℓ2(ℓq′
)-boundedness.
Constructing such a set T given a general multiplier m is quite subtle (except of
course in the scalar case, where the Minkowski norm on the one-dimensional span
of m is equivalent to the absolute value on C). Below we give an example where
these problems may be surmounted using extrapolation techniques.
Proposition 5.11. Let α ∈ (0, 1]. Suppose that m : R → L(Σ(Rd), L0(Rd)) and
that for some p0 ∈ (1, ∞) and all w ∈ Ap0 the following Holder-type condition is
satisfied:
(5.3)
sup
x∈R
km(x)kL(Lp0 (w)) + sup
J∈∆
Jα[m]C α(J;Lb(Lp0 (w))) ≤ φ([w]Ap0
).
Then there exists a subset T ⊂ L(Σ(Rd), L0(Rd)) such that m ∈ V 1/α(∆; T ) and
T is ℓu(ℓv)-bounded on Lp(w) for all p, u, v ∈ (1, ∞) and w ∈ Ap, with
Proof. For each J ∈ ∆ define
[T ]ℓu(ℓv) ≤ φp,u,v([w]Ap ).
T (J) := m(J) ∪n m(x) − m(y)
x − yα
Jα : x 6= y ∈ Jo,
and set T := SJ∈∆ T (J). Note that m(R) ⊂ T . We will show that T has the
desired properties.
Since m(x) ∈ T and m(x)−m(y)
Jα ∈ T for all J ∈ ∆ and all x 6= y ∈ J,
by the definition of the Minkowski and Holder norms, we have km(x)kT ≤ 1 and
Jα[m]C α(J;T ) ≤ 1, from which it follows directly that m ∈ V 1/α(∆; T ).
x−yα
By scalar extrapolation (see [15, Theorems 3.9 and Corollary 3.14]), we have
(5.3) for all p ∈ (1, ∞), which implies that
kT f kLp(w) ≤ φp([w]Ap )kf kLp(w)
for all p ∈ (1, ∞), w ∈ Ap(Rd), f ∈ Lp(w), and T ∈ T . Thus the ℓu(ℓv)-
boundedness result follows directly from Proposition 3.5.
(cid:3)
In the next example we specialise to the case X = Y = Lr and s ∈ (1, 2). Results
for s ∈ [2, ∞) will be presented in Example 5.16. Note that the ℓ2-boundedness or
ℓ2(ℓs)-boundedness assumptions can be deduced for instance from weight-uniform
Holder estimates as in Proposition 5.11.
24
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Example 5.12. Let p, r ∈ (1, ∞) and let T ⊂ Lb(Lr) be absolutely convex. Let
s ∈ (1, 2) and m ∈ V s(∆; T ). Then Tm is bounded on Lp(w; Lr) in each of the
following cases:
(i) If r = 2,
(a) p ∈ [s, ∞) and w ∈ Ap/s.
(b) p ∈ (1, s′] and w ∈ αp,s′ .
(ii) If r ∈ (2, ∞),
(a) p ∈ (2, ∞), w ∈ Ap/2 and T is ℓ2-bounded.
(b) p ∈ (1, r), s ∈ (1, r′), w ∈ αp,s′ and T is ℓ2(ℓs)-bounded.
(iii) If r ∈ (1, 2),
(a) p ∈ (1, 2), w ∈ αp,2 and T is ℓ2-bounded.
(b) p ∈ (r, ∞), s ∈ (1, r), w ∈ Ap/s and T is ℓ2(ℓs′
)-bounded.
Proof. The case (i)(a) follows from Theorem 5.4 and the case (i)(b) from a duality
argument. The cases (ii)(a) and (iii)(a) follow from Theorem 5.8(i) and (ii) with
q = 2. For (iii)(b) choose q ∈ (s, r) such that w ∈ Ap/q. By Corollary 3.8, T
is ℓ2(ℓq′
)-bounded, and therefore Theorem 5.8(i) applies. Similarly, (ii)(b) follows
from Theorem 5.8(ii).
(cid:3)
There is some overlap between the cases (a) and (b) in Example 5.12, but the
classes of weights considered are difficult to compare. For X = Lr, we can exploit
that we always have either X 2 ∈ UMD or (X ∗)2 ∈ UMD. This is not possible for
general UMD Banach function spaces, which restricts the class of multipliers that
can be handled by our results, as shown in the following example.
Example 5.13. Let p ∈ (1, ∞), r ∈ (1, 2), and let T ⊂ Lb(Lr ⊕ Lr′
) be absolutely
convex. Let s ∈ (1, r) and m ∈ V s(∆; T ). Then Tm is bounded on Lp(w; Lr ⊕ Lr′
)
in each of the following cases:
(i) p ∈ (r, ∞), w ∈ Ap/s and T is ℓ2(ℓs′
)-bounded.
(ii) p ∈ (1, r′), w ∈ αp,s′ and T is ℓ2(ℓs)-bounded.
The result follows from Theorem 5.8 in the same way as in Example 5.12.
5.3. Multipliers in intermediate UMD Banach function spaces. We can
prove stronger results, allowing for multipliers of lower regularity, if we consider
'intermediate' spaces X = [Y, H]θ where Y q ∈ UMD for some q ∈ (1, 2] and H is
a Hilbert space. For example, when r ∈ (2, ∞), we have Lr = [Lr0, L2]θ for some
r0 ∈ (r, ∞) and θ ∈ (0, 1). In this case Y = Lr0 satisfies the conditions of Theorem
5.8(i) with q = 2 and with H = L2 we can use Theorem 5.4.
In order to use interpolation methods we will need that span(T ) with the Minkow-
ski norm is a Banach space, i.e. that T is a Banach disc (see below Definition 4.1).
Theorem 5.14. Let p ∈ (1, ∞), q ∈ (1, 2] and θ ∈ (0, 1). Let Y and H be Banach
function spaces over the same measure space, with Y q ∈ UMD, H a Hilbert space,
and Y ∩ H dense in both Y and H. Let X = [Y, H]θ. Suppose T ⊂ Lb(Y ∩ H)
is a Banach disc which is ℓ2(ℓq′
)-bounded on Y and uniformly bounded on H. Let
s ∈ (1, ∞) and suppose that m ∈ V s(∆; T ).
(i) If s < min{p, [q, 2]θ} and s ≥ [q, 1]θ, then
kTmkL(Lp(w;X)) ≤ φY,p,q,s,θ([w]Ap/s )kmkV s(∆;T )[T ]ℓ2(ℓq′ )
for all w ∈ Ap/s.
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
25
(ii) If
1
s
and p > [q, 1]θ, then
> maxn 1
[q, 2]θ
−
1
p
,
1 − θ
q
,
1
p
−
θ
2o
kTmkL(Lp(R;X)) .Y,p,q,s,θ kmkV s(∆;T )[T ]ℓ2(ℓq′ ).
The allowable exponents (p, s) in Theorem 5.14 are shown in Figure 2. The
symmetry in Figure 2 is due to the equalities
θ
2
=
1
[∞, 2]θ
− 0 =
1
[q, 1]θ
−
1
[q, 2]θ
=
1
[q, 2]θ
−
1
[q, ∞]θ
and
1 − θ
q
=
1
[q, ∞]θ
− 0 =
1
[q, 2]θ
−
1
[∞, 2]θ
.
Figure 2. Allowable exponents for Theorem 5.14: the weighted
case (i) dark shaded, the unweighted case (ii) light shaded.
1/s
1
1
[q,1]θ
1
[q,2]θ
1
[q,∞]θ
0
0
1
1
[∞,2]θ
[q,2]θ
1
[q,1]θ
1
1/p
Proof. As in the proof of Theorem 5.4, it suffices to consider decaying multipliers
m ∈ V s
0 (∆; T ). Moreover, by Lemma 4.3, Proposition 2.1(ii) and the openness of
the upper bound assumptions on s, it suffices to consider m ∈ Rs
0(∆; T ). Through-
out the proof we let rs,θ,q ∈ [1, ∞) be the unique number such that
which exists if [q, 1]θ ≤ s < [q, ∞]θ.
[q, rs,θ,q]θ = s,
Part (i): First assume s 6= [q, 1]θ, so that rs,θ,q > 1. Fix a weight w ∈ A1. Take
t > q and define σ = [t, rs,θ,q]θ > s. By Proposition 5.7 we have boundedness of
the bilinear map
Rq
0(∆; T ) × Lt(w; Y ) → Lt(w; Y ),
(m, f ) 7→ Tmf
26
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
using that T is ℓ2(ℓq′
rs,θ,q ≤ 2, so we have by Theorem 5.4(i) and Lemma 4.3 that the bilinear map
)-bounded on Y . Moreover, since s ≤ [q, 2]θ, we know that
Rrs,θ,q
0
(∆; T ) × Lrs,θ,q (w; H) → Lrs,θ,q (w; H),
(m, f ) 7→ Tmf
is bounded, using
(5.4)
kmkRs(∆;Lb(H)) . kmkRs(∆;T )
by the uniform boundedness of T on H.
We define a bilinear map
0(∆; T ) ∩ Rrs,θ,q
0
(cid:0)Rs
(∆; T )(cid:1)×(cid:0)Lt(w; Y ) ∩ Lrs,θ,q (w; H)(cid:1)
→ Lr(w; Y ) ∩ Lrs,θ,q (w; H),
(m, f ) 7→ Tmf.
This is well-defined as it is the extension of the map (m, f ) 7→ Tmf defined for
m ∈ Rs∧rs,θ,q
(∆; T ) and f ∈ S(R; Y ∩ H). Here we use that Y ∩ H is dense in both
Y and H. By bilinear complex interpolation [3, §4.4] we have boundedness of
0
[Rq
0(∆; T ), Rrs,θ,q
0
(∆; T )]θ×[Lt(w; Y ), Lrr,θ,q (w; H)]θ
→ [Lt(w; Y ), Lrs,θ,q (w; H)]θ,
(m, f ) 7→ Tmf.
Here we use that the Minkowski norm on the linear span of T is complete, i.e. that
T ⊂ Lb(Y ∩ H) is a Banach disc.
By Theorem 4.4 we have
R[q,rs,θ,q]θ
0
(∆; T ) ֒→ [Rq
0(∆; T ), Rrs,θ,q
0
(∆; T )]θ.
Using this embedding and complex interpolation of weighted Bochner spaces (see
[49, Theorem 1.18.5]; note that the proof simply extends to the case X0 6= X1), we
get boundedness of
Rs
0(∆; T ) × Lσ(w; X) → Lσ(w; X),
(m, f ) 7→ Tmf
with norm estimate
kkTmf kX kLσ(w) ≤ φY,q,s,t,σ,θ([w]A1 )kmkRs(∆;T )[T ]ℓ2(ℓq′ )kkf kX kLσ(w)
for all w ∈ A1 and all simple functions f : R → X. By scalar-valued extrapolation
(see [15, Theorems 3.9 and Corollary 3.14]) and density of the simple functions we
deduce
kTmf kLp(w;X) ≤ φY,p,q,s,t,σ,θ([w]Ap/σ )kmkRs(∆;T )[T ]ℓ2(ℓq′ )kf kLp(w;X)
for all p ∈ [σ, ∞) and all w ∈ Ap/σ. Taking t arbitrarily close to q and using
Proposition 2.1(ii) proves the case [q, 1]θ 6= s.
Next if [q, 1]θ = s and w ∈ Ap/s, then by Proposition 2.1(ii) we can choose
t ∈ (s, [q, 2]θ) such that w ∈ Ap/t. By the previous case Tm is bounded on Lp(w; X)
for all m ∈ Rt(∆; T ) and hence also for m ∈ Rs(∆; T ), which completes the proof.
Part (ii): Without loss of generality we may assume that s > [q, 2]θ by embed-
ding of the Rs-spaces and the fact that
1
[q, 2]θ
> maxn 1
[q, 2]θ
−
1
p
,
1 − θ
q
,
1
p
−
θ
2o
for p > [q, 1]θ. Note that this implies that rs,θ,q > 2. We will consider three cases:
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
27
Case 1: p ≥ [∞, 2]θ. Since
1
1
pθ
>
θ(cid:16) θ
2
+
1 − θ
q
−
1
s(cid:17) =
1
2
−
1
rs,θ,q
2 − 1
rs,θ,q
we can find a p1 > pθ ≥ 2 such that p1 < p and p1 < ( 1
know by Theorem 5.4(ii), using (5.4), that the bilinear map
)−1. Therefore we
Rrs,θ,q
0
(∆; T ) × Lp1(R; H) → Lp1(R; H),
(m, f ) 7→ Tmf
is bounded. Since p < [∞, p1]θ we can find a p0 ∈ (p, ∞) such that p = [p0, p1]θ.
By Proposition 5.7 we have boundedness of the bilinear map
Rq
0(∆; T ) × Lp0(R; Y ) → Lp0(R; Y ),
(m, f ) 7→ Tmf,
using that T is ℓ2(ℓq′
)-bounded on Y . We can now finish the proof using bilinear
complex interpolation, Theorem 4.4 and complex interpolation of Bochner spaces
as in the first part.
(∆; T ) ֒→ L∞(R; T ). Therefore
Case 2: [q, 2]θ < p < [∞, 2]. Note that Rrs,θ,q
by Plancherel's theorem and (5.4) the bilinear map
(∆; T ) × L2(R; H) → L2(R; H),
Rrs,θ,q
0
0
(m, f ) 7→ Tmf
is bounded. Since [q, 2]θ < p < [∞, 2]θ we can find a p0 ∈ (q, ∞) such that
p = [p0, 2]θ. By Proposition 5.7 we have boundedness of the bilinear map
Rq
0(∆; T ) × Lp0(R; Y ) → Lp0(R; Y ),
(m, f ) 7→ Tmf,
using that T is ℓ2(ℓq′
)-bounded on Y . The proof can now be finished as before.
Case 3: [q, 1]θ < p ≤ [q, 2]. Let p ∈ (1, 2] be such that p = [q, p]θ. Then since
1
θ(cid:16) θ
2
+
<
1
p
1 − θ
1
2
2 + 1
we can find a 1 < p1 < p such that p1 > ( 1
rs,θ,q
Theorem 5.4(ii), using (5.4), that the bilinear map
(∆; T ) × Lp1(R; H) → Lp1(R; H),
Rrs,θ,q
1
s
−
+
q (cid:17) =
0
1
rs,θ,q
)−1. Therefore we know by
(m, f ) 7→ Tmf
is bounded. Since p1 < p we can find a p0 ∈ (q, ∞) such that p = [p0, p1]θ. By
Proposition 5.7 we have boundedness of the bilinear map
Rq
0(∆; T ) × Lp0(R; Y ) → Lp0(R; Y ),
(m, f ) 7→ Tmf,
again using that T is ℓ2(ℓq′
before.
)-bounded on Y . The proof can again be finished as
(cid:3)
The conditions on m in Theorem 5.14(ii) with q = 2 are less restrictive than
the conditions of [24, Theorem 3.6], which allows for Banach spaces with the LPRp
property. The proof of Theorem 5.14(ii) can also be used to improve the conditions
of [24, Theorem 3.6]
Remark 5.15. A weighted variant of part (ii) of Theorem 5.14 holds for an ap-
propriate class of weights, by using a weighted variant of Theorem 5.4(ii) (see [27,
Theorem A(ii)]) and limited range extrapolation (see [15, Theorem 3.31]). However
this involves a reverse Holder assumption on the weight or the dual weight, so the
technical details are therefore left to the interested reader.
We continue with an application to X = Lr for s ∈ [2, ∞). Results for s ∈ (1, 2)
have been previously covered by Example 5.12.
28
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
Example 5.16. Let (Ω, µ) be a measure space and let p, r ∈ (1, ∞). Let T ⊂
L(Σ(Ω), L0(Ω)) be absolutely convex and ℓ2-bounded on Lt(Ω) for all t ∈ (1, ∞).
Let s ∈ [2, ∞) and assume m ∈ V s(∆; T ). Then Tm is bounded on Lp(R; Lr(Ω))
in each of the following cases:
(i) r ∈ [2, ∞) and 1
(ii) r ∈ (1, 2] and 1
s > max(cid:8) 1
s > max(cid:8) 1
p − 1
r , 1
2 − 1
p , 1
2 − 1
r }.
r − 1
2 , 1
r − 1
2 , 1
p − 1
p }.
Proof. It suffices to prove (i), as (ii) follows from a duality argument. Let T be
the closure of T in Lb(L2(Ω)). Then T is a Banach disc. Moreover, by Lemma 3.9
we know that T ⊆ Lb(Lt(Ω)) is ℓ2-bounded for all t ∈ (1, ∞). We will check the
conditions of Theorem 5.14(ii) with T , q = 2, Y = Lt(Ω) for an appropriate t > r
and H = L2(Ω). Choose θ ∈ (0, 2
r ) such that
1
s
> maxn 1
2
−
1
p
,
1 − θ
2
,
1
p
−
θ
2o.
Since s ≥ 2 it follows that p > [2, 1]θ. Now the result follows by choosing t > r
such that r = [t, 2]θ.
(cid:3)
In a similar way we obtain the following from Theorem 5.14(i) and duality. This
partly improves Example 5.12.
s ≤ 1
r + 1
2 and w ∈ Ap/s.
Example 5.17. Let (Ω, µ) be a measure space and let p, r ∈ (1, ∞). Let T ⊂
L(Σ(Ω), L0(Ω)) be absolutely convex and ℓ2-bounded on Lt(Ω) for all t ∈ [2, ∞).
Let s ∈ (1, 2) and assume m ∈ V s(∆; T ). Then Tm is bounded on Lp(w; Lr(Ω)) if
p < 1
1
5.4. Multipliers in intermediate UMD Banach spaces. In this section we
consider general UMD Banach spaces (not just Banach function spaces) and use
interpolation to improve the conditions of Theorem 5.1 considerably, assuming X
is an interpolation space between a UMD space and a Hilbert space, and using the
same interpolation scheme as in Theorem 5.14. This result is new even for scalar-
valued multipliers, and it implies sufficient conditions for Fourier multipliers on the
space of Schatten class operators.
Theorem 5.18. Let p ∈ (1, ∞) and θ ∈ (0, 1). Let Y and H be an interpolation
couple, with Y ∈ UMD, H a Hilbert space, and Y ∩ H dense in both Y and H. Let
X = [Y, H]θ. Suppose T ⊂ Lb(Y ∩ H) is a Banach disc which is R-bounded on Y
and uniformly bounded on H. Let s ∈ (1, ∞) and suppose that m ∈ V s(∆; T ).
(i) If 1/s > min{1/p, 1 − (θ/2)}, then
kTmkL(Lp(w;X)) ≤ φY,p,s,θ([w]Ap/s )kmkV s(∆;T )[T ]R
for all w ∈ Ap/s.
(ii) If
then
1
s
> maxn1 −
θ
2
−
1
p
, 1 − θ,
1
p
−
θ
2o,
kTmkL(Lp(R;X)) .Y,p,s,θ kmkV s(∆;T )[T ]R.
The allowable exponents (p, s) above are shown in Figure 3.
Proof. To prove the result one can argue as in Theorem 5.14 with q = 1, and using
Theorem 5.1 instead of Proposition 5.7.
(cid:3)
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
29
Figure 3. Allowable exponents for Theorem 5.18: the weighted
case (i) dark shaded, the unweighted case (ii) light shaded.
1/s
1
1 − θ
2
1 − θ
0
0
θ
2
1 − θ
2
1/p
1
(i) r ∈ [2, ∞) and 1
s > maxn 1
s > maxn 1
p′ − 1
r − 1
r ,(cid:12)(cid:12) 1
p′ ,(cid:12)(cid:12) 1
r − 1
r − 1
p − 1
r − 1
ro.
po.
r′(cid:12)(cid:12), 1
r′(cid:12)(cid:12), 1
In the next example we apply Theorem 5.18 to operator-valued multipliers on
the Schatten class operators C r ⊆ Lb(ℓ2) for r ∈ [1, ∞]. This is potentially
useful for Schur multipliers (see [22, Theorem 5.4.3] and [44, Theorem 4]). For
r ∈ (1, ∞) these spaces have the UMD property, and for p, q ∈ [1, ∞] one has
C [p,q]θ = [C p, C q]θ (see [22, Propositions 5.4.2 and D.3.1]).
Example 5.19. Let X = C r with p, r ∈ (1, ∞) and T ⊆ L(C t) be absolutely
convex and R-bounded for all t ∈ (1, ∞). Let s ∈ (1, ∞) and assume m ∈ V s(∆; T ).
Then Tm is bounded on Lp(R; C r) in each of the following cases:
(ii) r ∈ (1, 2] and 1
In particular, if p ∈ [r ∧ r′, r ∨ r′] then Tm is bounded on Lp(R; C r) if r ∈ (1, ∞)
and 1
s > 1
r − 1
r′ .
Proof. The result follows from Theorem 5.18(ii) by arguing as in Example 5.16. A
similar result can be derived on Lp(w; C r) by Theorem 5.18(i).
(cid:3)
References
[1] A. Amenta, E. Lorist, and M. Veraar. Rescaled extrapolation for vector-valued functions.
arXiv:1703.06044, 2017.
[2] W. Arendt and S. Bu. The operator-valued Marcinkiewicz multiplier theorem and maximal
regularity. Math. Z., 240(2):311 -- 343, 2002.
[3] J. Bergh and J. Lofstrom. Interpolation spaces. An introduction. Springer-Verlag, Berlin-New
York, 1976. Grundlehren der Mathematischen Wissenschaften, No. 223.
[4] E. Berkson and T. A. Gillespie. Spectral decompositions and harmonic analysis on UMD
spaces. Studia Math., 112(1):13 -- 49, 1994.
[5] E. Berkson, T. A. Gillespie, and J. L. Torrea. Vector valued transference. In P. Liu, editor,
Functional Space Theory and its applications (Wuhan, 2003), pages 1 -- 27, 2003.
30
ALEX AMENTA, EMIEL LORIST, AND MARK VERAAR
[6] J. Bourgain. Some remarks on Banach spaces in which martingale difference sequences are
unconditional. Ark. Mat., 21(2):163 -- 168, 1983.
[7] J. Bourgain. Vector-valued singular integrals and the H 1-BMO duality. In Probability theory
and harmonic analysis (Cleveland, Ohio, 1983), volume 98 of Monogr. Textbooks Pure Appl.
Math., pages 1 -- 19. Dekker, New York, 1986.
[8] A. V. Buhvalov. The duality of functors that are generated by spaces of vector-valued func-
tions. Izv. Akad. Nauk SSSR Ser. Mat., 39(6):1284 -- 1309, 1438, 1975.
[9] D. L. Burkholder. A geometric condition that implies the existence of certain singular integrals
of Banach-space-valued functions. In Conference on harmonic analysis in honor of Antoni
Zygmund, Vol. I, II (Chicago, Ill., 1981), Wadsworth Math. Ser., pages 270 -- 286. Wadsworth,
Belmont, CA, 1983.
[10] D. L. Burkholder. Martingales and singular integrals in Banach spaces. In Handbook of the
geometry of Banach spaces, Vol. I, pages 233 -- 269. North-Holland, Amsterdam, 2001.
[11] A.-P. Calder´on. Intermediate spaces and interpolation, the complex method. Studia Math.,
24:113 -- 190, 1964.
[12] P. Cl´ement, B. de Pagter, F. Sukochev, and H. Witvliet. Schauder decompositions and mul-
tiplier theorems. Studia Math., 138(2):135 -- 163, 2000.
[13] P. Cl´ement and J. Pruss. An operator-valued transference principle and maximal regularity
on vector-valued Lp-spaces. In Evolution equations and their applications in physical and
life sciences (Bad Herrenalb, 1998), volume 215 of Lecture Notes in Pure and Appl. Math.,
pages 67 -- 87. Dekker, New York, 2001.
[14] R. Coifman, J. L. Rubio de Francia, and S. Semmes. Multiplicateurs de Fourier de Lp(R) et
estimations quadratiques. C. R. Acad. Sci. Paris S´er. I Math., 306(8):351 -- 354, 1988.
[15] D. V. Cruz-Uribe, J. M. Martell, and C. P´erez. Weights, extrapolation and the theory of Rubio
de Francia, volume 215 of Operator Theory: Advances and Applications. Birkhauser/Springer
Basel AG, Basel, 2011.
[16] H. G. Dales, N. J. Laustsen, T. Oikhberg, and V. G Troitsky. Multi-norms and banach
lattices. Dissertationes Mathematicae (Rozprawy Matematyczne), 2016.
[17] S. Fackler, T.P. Hytonen, and N. Lindemulder. Weighted estimates for operator-valued mul-
tipliers. Preprint.
[18] C. Gallarati, E. Lorist, and M. C. Veraar. On the ℓs-boundedness of a family of integral
operators. Rev. Mat. Iberoam., 32(4):1277 -- 1294, 2016.
[19] T. A. Gillespie and J. L. Torrea. Transference of a Littlewood-Paley-Rubio inequality and
dimension free estimates. Rev. Un. Mat. Argentina, 45(1):1 -- 6 (2005), 2004.
[20] L. Grafakos. Modern Fourier analysis, volume 250 of Graduate Texts in Mathematics.
Springer, New York, second edition, 2009.
[21] R. Haller, H. Heck, and A. Noll. Mikhlin's theorem for operator-valued Fourier multipliers in
n variables. Math. Nachr., 244:110 -- 130, 2002.
[22] T. P. Hytonen, J. M. A. M. van Neerven, M. C. 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. Folge. Springer International Publishing, 2016.
[23] T. P. Hytonen, J. M. A. M. van Neerven, M. C. Veraar, and L. Weis. Analysis in Banach
Spaces. Volume II: Probabilistic Methods and Operator Theory. In preparation., 2017.
[24] T. P. Hytonen and D. Potapov. Vector-valued multiplier theorems of Coifman-Rubio de
Francia-Semmes type. Arch. Math. (Basel), 87(3):245 -- 254, 2006.
[25] T. P. Hytonen, J. L. Torrea, and D. V. Yakubovich. The Littlewood-Paley-Rubio de Francia
property of a Banach space for the case of equal intervals. Proc. Roy. Soc. Edinburgh Sect.
A, 139(4):819 -- 832, 2009.
[26] S. V. Kislyakov. The Littlewood-Paley theorem for arbitrary intervals: weighted estimates.
J. Math. Sci (N. Y.), 156(5):824 -- 833, 2009.
[27] S. Kr´ol. Fourier multipliers on weighted Lp spaces. Math. Res. Lett., 21(4):807 -- 830, 2014.
[28] P. Kunstmann and A. Ullmann. R
s-sectorial operators and generalized Triebel-Lizorkin
spaces. J. Fourier Anal. Appl., 20(1):135 -- 185, 2014.
[29] P. Kunstmann and L. Weis. Maximal Lp-regularity for parabolic equations, Fourier mul-
tiplier theorems and H∞-functional calculus. In Functional analytic methods for evolution
equations, volume 1855 of Lecture Notes in Math., pages 65 -- 311. Springer, Berlin, 2004.
FOURIER MULTIPLIERS IN BANACH FUNCTION SPACES
31
[30] S. Kwapie´n. On operators factorizable through Lp space. In Actes du Colloque d'Analyse
Fonctionnelle de Bordeaux (Univ. de Bordeaux, 1971), pages 215 -- 225. Bull. Soc. Math.
France, M´em. No. 31 -- 32. Soc. Math. France, Paris, 1972.
[31] S. Kwapie´n, M. C. Veraar, and L. Weis. R-boundedness versus γ-boundedness. Ark. Mat.,
54(1):125 -- 145, 2016.
[32] M.T. Lacey. Issues related to Rubio de Francia's Littlewood-Paley inequality, volume 2 of New
York Journal of Mathematics. NYJM Monographs. State University of New York, University
at Albany, Albany, NY, 2007.
[33] M. Ledoux and M. Talagrand. Probability in Banach spaces, volume 23 of Ergebnisse der
Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)].
Springer-Verlag, Berlin, 1991. Isoperimetry and processes.
[34] N. Lindemulder. Parabolic initial-boundary value problems with inhomogeneous data. Mas-
ter's thesis, Utrecht University, Utrecht, the Netherlands, 2014.
[35] N. Lindemulder. Banach space-valued extensions of linear operators on L∞. In Ordered Struc-
tures and Applications (Positivity VII, Zaanen Centennial Conference). Birkhauser Verlag,
2016.
[36] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. II, volume 97 of Ergebnisse
der Mathematik und ihrer Grenzgebiete. Springer-Verlag, Berlin-New York, 1979. Function
spaces.
[37] P. Meyer-Nieberg. Banach lattices. Universitext. Springer-Verlag, Berlin, 1991.
[38] M. Meyries and M. C. Veraar. Pointwise multiplication on vector-valued function spaces with
power weights. J. Fourier Anal. Appl., 21(1):95 -- 136, 2015.
[39] S. Montgomery-Smith. Stability and dichotomy of positive semigroups on Lp. Proc. Amer.
Math. Soc., 124(8):2433 -- 2437, 1996.
[40] J. M. A. M. van Neerven, M. C. Veraar, and L. Weis. On the R-boundedness of stochastic
convolution operators. Positivity, 19(2):355 -- 384, 2015.
[41] J. M. A. M. van Neerven and L. Weis. Stochastic integration of functions with values in a
Banach space. Studia Math., 166(2):131 -- 170, 2005.
[42] P. P´erez Carreras and J. Bonet. Barrelled locally convex spaces, volume 131 of North-Holland
Mathematics Studies. North-Holland Publishing Co., Amsterdam, 1987. Notas de Matem´atica
[Mathematical Notes], 113.
[43] G. Pisier. Martingales in Banach spaces. Cambridge: Cambridge University Press, 2016.
[44] D. Potapov and F. Sukochev. Operator-Lipschitz functions in Schatten-von Neumann classes.
Acta Math., 207(2):375 -- 389, 2011.
[45] D. Potapov, F. Sukochev, and Q. Xu. On the vector-valued Littlewood-Paley-Rubio de Fran-
cia inequality. Rev. Mat. Iberoam., 28(3):839 -- 856, 2012.
[46] J. L. Rubio de Francia. A Littlewood-Paley inequality for arbitrary intervals. Rev. Mat.
Iberoam., 1(2):1 -- 14, 1985.
[47] J. L. Rubio de Francia. Martingale and integral transforms of Banach space valued functions.
In Probability and Banach spaces (Zaragoza, 1985), volume 1221 of Lecture Notes in Math.,
pages 195 -- 222. Springer, Berlin, 1986.
[48] Z. Strkalj and L. Weis. On operator-valued Fourier multiplier theorems. Trans. Amer. Math.
Soc., 359(8):3529 -- 3547 (electronic), 2007.
[49] H. Triebel. Interpolation theory, function spaces, differential operators, volume 18 of North-
Holland Mathematical Library. North-Holland Publishing Co., Amsterdam-New York, 1978.
[50] L. Weis. A new approach to maximal Lp-regularity. In Evolution equations and their appli-
cations in physical and life sciences (Bad Herrenalb, 1998), volume 215 of Lecture Notes in
Pure and Appl. Math., pages 195 -- 214. Dekker, New York, 2001.
[51] L. Weis. Operator-valued Fourier multiplier theorems and maximal Lp-regularity. Math.
Ann., 319(4):735 -- 758, 2001.
[52] Q. Xu. Fourier multipliers for Lp(Rn) via q-variation. Pacific J. Math., 176(1):287 -- 296, 1996.
Delft Institute of Applied Mathematics, Delft University of Technology, P.O. Box
5031, 2600 GA Delft, The Netherlands
E-mail address: [email protected]
E-mail address: [email protected]
E-mail address: [email protected]
|
1810.01260 | 1 | 1810 | 2018-10-02T14:01:06 | H\"ormander condition for pseudo-multipliers associated to the harmonic oscillator | [
"math.FA",
"math.AP",
"math.SP"
] | In this paper we prove H\"ormander-Mihlin multiplier theorems for pseudo-multipliers associated to the harmonic oscillator (also called the Hermite operator). Our approach can be extended to also obtain the $L^p$-boundedness results for multilinear pseudo-multipliers. By using the Littlewood-Paley theorem associated to the harmonic oscillator we also give $L^p$-boundedness and $L^p$-compactness properties for multipliers. $(L^p,L^q)$-estimates for spectral pseudo-multipliers also are investigated. | math.FA | math |
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
ASSOCIATED TO THE HARMONIC OSCILLATOR
DUV ´AN CARDONA AND MICHAEL RUZHANSKY
Abstract. In this paper we prove Hormander-Mihlin multiplier theorems
for pseudo-multipliers associated to the harmonic oscillator (also called the
Hermite operator). Our approach can be extended to also obtain the Lp-
boundedness results for multilinear pseudo-multipliers. By using the Littlewood-
Paley theorem associated to the harmonic oscillator we also give Lp-boundedness
and Lp-compactness properties for multipliers. (Lp, Lq)-estimates for spectral
pseudo-multipliers also are investigated.
Contents
Introduction
1.
2. Boundedness of pseudo-multipliers associated to the harmonic
(Lp, Lq)-boundedness of spectral pseudo-multipliers
oscillator, Hormander condition
2.1. Hermite functions in Lp spaces
2.2. Hormander condition for pseudo-multipliers on Lp spaces
2.3.
2.4. Lower bounds for the operator norm of multipliers on Lp spaces
3. Compactness of pseudo-multipliers
3.1. L2-compactness of multipliers
3.2. Lp-compactness and Lp-boundedness for multipliers via Littlewood-
Paley theory
4. Lp-boundedness for multilinear pseudo-multipliers
References
1
6
7
9
16
20
21
21
22
24
32
1. Introduction
In this paper we are interested in the Lp-boundedness of pseudo-multipliers
associated to the harmonic oscillator (also called Hermite pseudo-multipliers) on
Lp(Rn)-spaces. The harmonic oscillator is the fundamental operator of quantum
mechanics defined by
Hψ := (−∆x + x2)ψ,
(1.1)
Date: October 3, 2018.
2010 Mathematics Subject Classification. Primary 81Q10; Secondary 42C10, 35J10, 33C45.
Key words and phrases. Pseudo-multiplier; Harmonic oscillator; Hermite functions;
Hormander condition; Multilinear operator; Fourier multipliers.
The second author was supported in parts by the FWO Odysseus Project, by the EPSRC
Grant EP/R003025/1 and by the Leverhulme Research Grant RPG-2017-151.
1
2
D. CARDONA AND M. RUZHANSKY
i=1 x2
with x2 := Pn
i . The harmonic oscillator extends to an unbounded self-
adjoint operator on L2(Rn), and its spectrum consists of the discrete set λν :=
2ν + n, ν ∈ Nn
0 (called Hermite
functions) which provide an orthonormal basis of L2(Rn). Each Hermite function
φν on Rn has the form
0 , with a set of real eigenfunctions φν, ν ∈ Nn
φν := Πn
j=1φνj , φνj (xj) := (2νj νj!√π)− 1
2 Hνj (xj)e− 1
2 x2
j ,
(1.2)
where x = (x1, . . . , xn) ∈ Rn, ν = (ν1, . . . , νn) ∈ Nn
0 , and
Hνj (xj) := (−1)νj ex2
j
dk
dxk
j
(e−x2
j )
denotes the Hermite polynomial of order νj. By the spectral theorem, for every
f ∈ D(Rn) we have
where bf (φν) is the Fourier-Hermite transform of f at ν defined by
f (x)φν(x) dx.
0
Hf (x) = Xν∈Nn
λνbf (φν)φν(x),
(FHf )(ν) ≡ bf (φν) = (f, φν)L2(Rn) :=ZRn
m(ν)bf (φν)φν(x),
Tmf (x) := Xν∈Nn
0
A multiplier associated to the harmonic oscillator (or Hermite multiplier) is a
linear operator Tm of the form
(1.3)
(1.4)
(1.5)
for every function f ∈ D(Rn). The discrete function m is called the symbol
of the operator Tm. In particular, if m is a measurable function, the symbol
of the spectral multiplier m(H) (defined by the functional calculus) is given by
m(ν) := m(λν), so that the spectral multipliers are natural examples of multipli-
ers associated to the harmonic oscillator. We can refer to e.g. Prugovecki [19] for
the quantum mechanical aspects of the harmonic oscillators.
Now, we present some historical results on the analysis of multipliers. If we
denote by Pℓ the orthogonal projection to the subspace generated by the set
{φν : ν = ℓ}, and m is a radial function in the sense that m(ν) = m(ν′) when
ν = ν′, then the multiplier Tm can be written as
Tm ≡ Tµf (x) :=
µ(ℓ)(Pℓf )(x),
(1.6)
∞Xℓ=0
where µ(ν) = m(ν). An earlier result by G. Mauceri [16] (by using methods of
Bonami-Clerc [2] and R. Strichartz [23]) states that the condition
sup
j
2j(k−1) X2j ≤N ≤2j+1 ∆k
N µ(N) < ∞,
(1.7)
where 0 ≤ k ≤ n + 1, implies the boundedness of Tµ for all 1 < p < ∞. As it was
pointed out in [29], the number of discrete derivatives k above can be taken in
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
3
the range 0 ≤ k ≤ [ 3n−2
[30]) states that if m satisfies the discrete Marcienkiewicz condition
] + 2. A remarkable result proved by S. Thangavelu (see
6
∆α
ν m(ν) ≤ Cα(1 + ν)−α, α ∈ Nn
0 , α ≤ [
n
2
] + 1,
(1.8)
where ∆ν is the usual difference operator, then the corresponding multiplier Tm :
Lp(Rn) → Lp(Rn) extends to a bounded operator for all 1 < p < ∞. This result
is a discrete analogue of the result proved by Mihlin [17] for Fourier multipliers
of the form
a(ξ)F f (ξ)e−2πix·ξdξ,
(1.9)
Taf (x) =ZRn
where F is the Fourier transform on Rn. The Mihlin condition states that if a is
a function on Rn satisfying
ξ a(ξ) ≤ Cαξ−α, ξ 6= 0, α ∈ Nn
∂α
0 , α ≤ [
n
2
] + 1,
(1.10)
then Ta : Lp(Rn) → Lp(Rn) extends to a bounded operator for all 1 < p < ∞.
In [14] Hormander generalised the Mihlin condition (1.10) to the condition of the
form
kakl.u.H s := sup
r>0 ka(r·)η(·)kH s(Rn) = sup
r>0
rs− n
2 ka(·)η(r−1·)kH s(Rn) < ∞,
(1.11)
where η ∈ D(0,∞) and s > n
multiplier Ta on Lp(Rn) for all 1 < p < ∞.
harmonic oscillator is quite different. In fact, for all s and ε > 0 with
As it was pointed out in [3], the situation for multipliers associated to the
2 , in order to guarantee the boundedness of a Fourier
n
2
< s ≤
n
2
+
1
6 − ε
(1.12)
we can not guarantee the Lp-boundedness of Riesz means operators satisfying
(1.11), for all 1 < p < ∞. However, it was proved in [3] that there exists p0 ∈ [1, 2]
such that a general operator a(H) satisfying so-called Plancherel estimates can
be extended to a bounded operator on Lp(Rn) for all p0 < p < p′
0, provided that
a satisfies (1.11) for s > n+1
2 . Hormander conditions for Hermite operators were
established in [24], see also [5, Theorem III.9].
An extension of Fourier multipliers is given by so-called pseudo-multipliers (see
0 the associated pseudo-multiplier Tm
[1]). If m is a bounded function on Rn × Nn
is the operator defined by
Tmf (x) := Xν∈Nn
0
m(x, ν)bf (φν)φν(x)
(1.13)
for every function f ∈ D(Rn). We refer to the function m as the symbol of the
operator Tm. If m(ν) = µ(ν) (as in the Maceuri result mentioned previously), it
was proved among other things by S. Bagchi and S. Thangavelu [1] (see also J.
Epperson [10]), that for n ≥ 2, the condition
x∈Rn ∆jµ(x, k) ≤ Cj(2k + n)−j, 0 ≤ j ≤ n + 1,
sup
(1.14)
4
D. CARDONA AND M. RUZHANSKY
implies that the pseudo-multiplier Tµ is of weak type (1,1) and bounded on Lp(Rn)
provided that Tµ is bounded on L2(Rn). The reference [1] provides several condi-
tions for the boundedness of pseudo-multipliers including continuity in Lp-spaces
with weights.
From the point of view of the theory of pseudo-differential operators, pseudo-
multipliers would be the special case of the symbolic calculus developed in the
works of the second author and N. Tokmagambetov [20, 21].
The main result of this paper is the Hormander type condition for pseudo-
multiplier operators (1.13) and for their multilinear versions. In order to classify
the order of regularity in our Hormander conditions, we use the following norms,
kmkl.u.,H s := sup
k>0, y∈Rn
2k(s− n
x ) < ∞,
(1.15)
2 )khxisF [m(y,·)ψ(2−k · )](x)kL2(Rn
2 )khxisF −1
H [m(y,·)ψ(2−k · )](x)kL2(Rn
2k(s− n
1
k>0
sup
y∈Rn
x ) < ∞, (1.16)
kmkl.u.,Hs := sup
defined by the Fourier transform and the Fourier-Hermite transform, respectively,
with hxi := (1 + x2)
2 . In (1.15) we consider functions m on Rn × Rn, but to
these functions we associate a pseudo-multiplier with symbol {m(x, ν)}x∈Rn,ν∈Nn
0 .
Our main results for pseudo-multipliers can be summarised in the following two
theorems.
Theorem 1.1. Let us assume that 2 ≤ p < ∞. If Tm is a pseudo-multiplier with
symbol m satisfying (1.15), then under one of the following conditions,
n+1 , and s > sn,p := 3n
n+1 , and s > sn,p := 3n
2 + n−1
n+1 < p ≤ 2n
n−2, and s > sn,p := 3n
n−2 ≤ p < ∞, and s > sn,p := 3n−1
• n ≥ 2, 2 ≤ p < 2(n+3)
• n ≥ 2, p = 2(n+3)
• n ≥ 2, 2(n+3)
• n ≥ 2,
• n = 1, 2 ≤ p < 4, s > s1,p := 3
2,
• n = 1, p = 4, s > s1,4 := 2,
• n = 1, 4 < p < ∞, s > s1,p := 4
2 + n−1
2(n+3) ,
2 − 1
2 +n( 1
2 − 1
p),
3 + 2
3( 1
2n
2 ( 1
2 − 1
p),
2 − 1
p),
3 ( 1
6 + 2n
2 − 1
p),
the operator Tm extends to a bounded operator on Lp(Rn). For 1 < p ≤ 2, under
one of the following conditions
2n
• n ≥ 2, 2(n+3)
• n ≥ 2,
• n ≥ 2, 1 < p ≤ 2n
• n = 1, 4
• n = 1, 1 < p < 4
n+5 ≤ p ≤ 2, and s > sn,p := 3n
n+2 ≤ p ≤ 2(n+3)
3 ≤ p < 2, s > s1,p := 3
2,
3+ 2
3, s > s1,p := 4
n+2, and s > sn,p := 3n−1
n+5 , and s > sn,p := 3n
2 − 1
p ),
2 + n−1
2 − 1
2 +n( 1
3( 1
2 ( 1
2 − 1
p),
6 + 2n
2 − 1
3 ( 1
p),
2 − 1
p ),
the operator Tm extends to a bounded operator on Lp(Rn). However, in general:
implies the Lp-
3 < p < 4 and every n, the condition s > 3n
2
• for every 4
boundedness of Tm.
If the symbol m of the pseudo-multiplier Tm satisfies the Hormander condition
(1.16), in order to guarantee the Lp-boundedness of Tm, in every case above we
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
5
can take s > sn,p − 1
boundedness of Tm for all 4
3 < p < 4.
12. Moreover, the condition s > 3n
12 implies the Lp-
2 − 1
Now we discuss some important facts concerning the results of this paper.
• It is usual to assume the L2-boundedness of a pseudo-multiplier Tm in
order to provide its Lp-boundedness (see [1] and [10]). Indeed, as it was
pointed out in [1], the problem of finding satisfactory conditions for the L2-
boundedness of pseudo-multipliers remains open. However, in our main
theorem we solve such problem by considering symbols m(x, ν) satisfying
the Hormander condition (1.15) of order s > 3n
2 , uniformly in y ∈ Rn, or
the condition (1.16) for s > 3n
• A function m belongs to the Kohn-Nirenberg class S0,ρ(Rn × Rn) if it
12 , uniformly in y ∈ Rn.
2 − 1
satisfies the symbol inequalities
ξ m(x, ξ) ≤ Cα(1 + ξ)−α,
∂α
(1.17)
uniformly in x ∈ Rn. Symbols in the class S0,2n+1 are functions satisfying
(1.15) and they provide bounded pseudo-multipliers in Lp-spaces for all
1 < p < ∞. In particular symbols in the class S0, [ 3n
2 ]+1 provide bounded
pseudo-multipliers in L2(Rn). These facts will be proved in Proposition
2.11. Moreover, (see Corollary 2.12) if we assume the condition,
α ≤ ρ,
∆α
0 , α ≤ ρ,
ν m(x, ν) ≤ Cα(1 + ν)−α, α ∈ Nn
(1.18)
for ρ = [3n/2] + 1, then Tm extends to a bounded operator on L2(Rn), and
for ρ = 2n + 1 we have the Lp(Rn)-boundedness of Tm for all 1 < p < ∞.
• For n = 1 and by assuming the L2-boundedness of a pseudo-multiplier
Tm, it was proved by Epperson [10] that (1.18) is a sufficient condition
for the Lp-boundedness of Tm provided that ρ = 5. In constrast, we only
require derivatives up to order ρ = 3. For spectral pseudo-multipliers
m(x, H) and n ≥ 2, and newly by assuming the L2-boundedness, Bagchi
and Thangavelu proved the Lp-boundedness provided that (1.18) holds
true for ρ = n + 1. Although we impose for n ≥ 2, ρ = [3n/2] + 1, we
do not assume the L2-boundedness for these operators. We also include
general pseudo-multipliers and particularly spectral pseudo-multipliers.
• The (Lp, Lq)−boundedness of pseudo-multipliers will be investigated in
• By using the Littlewood-Paley theorem associated to the harmonic os-
cillator, we give a Lp-multiplier theorem and a Lp-compactness theorem
for multipliers (see Theorem 3.3), the sufficient condition imposed is how-
ever, different from the Hormander condition. The L2-compactness of
multipliers will be characterised in Theorem 3.1.
Theorem 2.13 and Theorem 2.14.
In this paper we introduce the notion of multilinear pseudo-multipliers, which,
in analogy with the definition of multilinear Fourier multipliers, are operators of
the form
Tm(f1,· · · , fκ) := Xν:=(ν1,··· ,νκ)∈Nnκ
0
m(x, ν)bf1(φν1) · · ·bfκ(φνκ)φν1 · · · φνκ , x ∈ Rn,
(1.19)
6
D. CARDONA AND M. RUZHANSKY
for all f1, f2,· · · , fκ ∈ D(Rn). In this setting, by imposing discrete multilinear
Hormander conditions on the symbol m, of the type
kmkl.u.,Hs := sup
k>0, x∈Rn
2k(s− nκ
2 )khzisF −1
H [m(x,·)ψ(2−k · )](z)kL2(Rnκ
z ) < ∞,
(1.20)
z ) < ∞, (1.21)
kmkl.u.,H s := sup
we want to guarantee the boundedness of Tm. Thus, we establish the following
multilinear result.
2 )khzisF [m(x,·)ψ(2−k · )](z)kL2(Rnκ
2k(s− nκ
k>0, x∈Rn
Theorem 1.2. Let us consider a multilinear pseudo-multiplier Tm defined on
D(Rn)κ with symbol satisfying (1.20) or (1.21) for
+ (κ − 1)γ∞,
with γ∞, defined as in (2.11). Then the operator
s > sn,κ,p := max{
2
3nκ
3nκ
2
+
(κ − 1)n
4
},
(1.22)
extends to a bounded multilinear operator provided that 1 ≤ pj ≤ ∞, 1 ≤ p ≤ 2,
and 1
Tm : Lp1 × Lp2 × · · · × Lpκ−1 × Lpκ → Lp(Rn)
. If m satisfies the condition (1.20) or (1.21) for
p = 1
p1
pκ
+ · · · + 1
s > max{
3nκ
2
+
(κ − 1)n
4
,
3nκ
2
+
(n − 1)(κ − 1)
2
+ γp},
with γp defined as in (2.11), then (1.22) holds true for all 2 ≤ p ≤ ∞ and
p = 1
pκ
.
1
+ · · · + 1
p1
Let us note that 3nκ
2 + (κ−1)n
4
and 3nκ
2 + (n−1)(κ−1)
2
+ γp cannot be compared
immediately because the sign of γp depends on the values of p.
This multilinear theorem for pseudo-multipliers is analogous to ones obtained
in the framework of multilinear multipliers. Although the literature for the mul-
tilinear analysis is extensive, we refer the reader to [11, 12, 13] and to the seminal
work of R. Coifman and Y. Meyer where the multilinear harmonic analysis was
originated.
This paper is organised as follows. In Section 2 we present the proof of our
main theorem. In Section 3 we discuss the compactness properties. Finally, in
Section 4 we prove the result mentioned above for multilinear pseudo-multipliers.
2. Boundedness of pseudo-multipliers associated to the harmonic
oscillator, Hormander condition
Throughout this paper the function ψ ∈ D(0,∞) will be supported in [ 1
2 , 4]
with ψ ≡ 1 on [1, 2]. In this section we will use functions in a (locally uniformly)
Sobolev space of order s > 0, which consists of all functions m on Rn × Rn
satisfying
kmkl.u.,H s := sup
k>0, y∈Rn
2k(s− n
2 )khxisF [m(y,·)ψ(2−k · )](x)kL2(Rn
x ) < ∞,
(2.1)
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
7
in order to establish the Lp-boundedness of Hermite pseudo-multipliers. We have
denoted by F the Fourier transform on Rn defined by
(F f )(ξ) =ZRn
e−2πix·ξf (x)dx.
(2.2)
Another option that we can use in order to define (local) discrete Sobolev spaces
come from the norm
kmkl.u.,Hs := sup
We recall that the Fourier-Hermite transform FH is defined for every f ∈ D(Rn)
by the formula
H [m(y,·)ψ(2−k · )](x)kL2(Rn
2 )khxisF −1
x ) < ∞.
2k(s− n
sup
y∈Rn
(2.3)
k>0
f (x)ψν(x)dx, ν ∈ Nn
0 .
If we denote the inverse Fourier-Hermite transform by F −1
(FHf )(ν) :=ZRn
(2.4)
H which is defined by
u(ν)φν(x),
(2.5)
where u is a function with compact support on Nn
inversion formula is given by
0 , then the Fourier-Hermite
(FH f )(ν)φν(x).
(2.6)
(F −1
H u)(x) := Xν∈Nn
0
f (x) = Xν∈Nn
0
Now, a pseudo-multiplier Tm with symbol m has, in terms of the transformation
FH, the alternative representation
Tmf (x) = F −1
H [m(x, ν)(FH f )](x).
(2.7)
For properties and basics of the Fourier-Hermite transform and Hermite expan-
sions we refer the reader to Thangavelu [30].
2.1. Hermite functions in Lp spaces. The main tool in the formulation of
our results will be estimates of the Lp-norms of Hermite functions. Our starting
point is the following lemma for one-dimensional Hermite functions (see Lemma
4.5.2 of Thangavelu [30]).
Lemma 2.1. Let us denote by φν, ν ∈ Nn
these functions satisfy the estimates
4 , 1 ≤ p < 4.
8 ln(ν).
6p − 1
0 , the Hermite functions. As ν → ∞,
2p − 1
• kφνkLp(R) ≍ ν
• kφνkL4(R) ≍ ν− 1
• kφνkLp(R) ≍ ν− 1
12 , 4 < p ≤ ∞.
1
Now, we present a lemma on the behaviour of Lp(Rn)- norms of Hermite func-
tions on Rn for all 1 ≤ p ≤ 2.
Lemma 2.2. Let φν, ν ∈ Nn
we have
0 , be Hermite functions on Rn. Then for 1 ≤ p ≤ 2
kφνkLp(Rn) . ν
n
2 ( 1
p − 1
2 ).
(2.8)
8
D. CARDONA AND M. RUZHANSKY
Proof. We will use the first equivalence in Lemma 2.1. Every Hermite function
on Rn has the form φν = φν1 × · · · × φνn and as a consequence we have
kφνjkLp(R).
kφνkLp(Rn) =Yj
νj! 1
≤(cid:18)Pj νj
4 ≥ 0 and
2p − 1
4
Now, if 1 ≤ p ≤ 2 then 1
2p − 1
kφνkLp(Rn) ≍ Yj
. ν
where we have used the inequality x1 × · · · × xn ≤ ( x1+···+xn
n
2p − 1
4 )
n (cid:19)n( 1
n
2 ( 1
p − 1
2 ),
)n for xi > 0.
We now recall the following sharp lemma on the Lp-norms of Hermite functions
for 2 ≤ p ≤ ∞ (see H. Koch and D. Tataru [15]).
Lemma 2.3. Let us consider a Hermite function φ = φν on Rn which, as an
eigenfunction of the harmonic oscillator on Rn, has the associated eigenvalue
λ2 = (2ν + n). Then for n ≥ 2 we have,
(2.9)
(2.10)
(cid:3)
2n
n−2, then kφνkLp(Rn) . (2ν + n)− 1
n+1 , then kφνkLp(Rn) . (2ν + n)
• if 2 ≤ p < 2(n+3)
• if 2(n+3)
n+1 < p ≤ 2n
n−2 ≤ p ≤ ∞, then kφνkLp(Rn) . (2ν + n)− 1
• if
• if 2 ≤ p < 4, kφνkLp(R) . (2ν + n)− 1
• if 4 < p ≤ ∞, kφνkLp(R) . (2ν + n)− 1
2 − 1
2 ( 1
p ),
p − 1
6 ( 1
6 + 1
2 + n
2 ).
1
2p − 1
4 ,
6 + n
6 ( 1
2 − 1
p ),
2 ( 1
2 − 1
p ),
and for n = 1,
It is important to mention that in the previous lemma we denote 2n
n−2 = ∞,
when n = 2. We adopt this convention in the whole paper. Let us mention that,
curiously, the proof of the lemma above is a consequence of some dispersive and
Strichartz estimates for the corresponding Schrodinger equation for the harmonic
oscillator. In our further analysis, we will need the following lemma.
Lemma 2.4. Let us assume that 2 ≤ p ≤ ∞ and n ≥ 2. Then, the Hermite
functions satisfy the following estimates as ν → ∞ :
• if 2 ≤ p < 2(n+3)
n+1 , then
• if 2(n+3)
n+1 < p ≤ 2n
• if
2n
n−2 ≤ p ≤ ∞, then
kφνkLp(Rn)kφνkLp′ (Rn) . ν
n−2, then
kφνkLp(Rn)kφνkLp′ (Rn) . ν− 1
n−1
2 ( 1
2 − 1
p ),
6 + 2n
3 ( 1
2 − 1
p ),
kφνkLp(Rn)kφνkLp′ (Rn) . ν− 1
2 +n( 1
2 − 1
p ).
Let us recall that we have denoted 2n
n−2 = ∞ when n = 2. For n = 1 we have
• if 2 ≤ p < 4,
kφνkLp(R)kφνkLp′ (R) . 1,
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
9
• if 4 < p ≤ ∞,
In general:
kφνkLp(R)kφνkLp′ (R) . ν− 1
6 + 2
3 ( 1
2 − 1
p ).
• for every 4
3 < p < 4 and every n, kφνkLp(Rn)kφνkLp′ (Rn) = O(1).
Proof. Except for the last item, the proof is a straightforward computation by
replacing p in Lemma 2.2 and the estimates in Lemma 2.3. The last item can be
proved by using that p, p′ ∈ ( 4
3, 4) and the first estimate in Lemma 2.1, in fact
kφνkLp(Rn)kφνkLp′ (Rn) =Yj
≍Yj
kφνjkLp(R)kφνjkLp′ (R)
2p − 1
ν
j
2p′ − 1
ν
j
4
4
1
1
1
2 ( 1
ν
j
=Yj
p + 1
p′ )− 1
2
1 = 1,
=Yj
(cid:3)
completing the proof.
Remark 2.5. Because kφνjkL∞(R) . νj− 1
kφνkL∞(Rn) . ν− 1
and from the inequality ν ≤ nνi we obtain ν
desired estimate.
12 . Indeed, when ν → ∞, then νi
i ≤ n
12 when νj → ∞, we can estimate
:= max1≤j≤n νj → ∞,
12 which implies the
− 1
12
1
12ν− 1
2.2. Hormander condition for pseudo-multipliers on Lp spaces. Now, we
analyse the boundedness of pseudo-multipliers with symbols in (locally uniform)
Sobolev spaces. We denote by γp the exponent that according to Lemma 2.4
satisfies
kφνkLp(Rn)kφνkLp′ (Rn) . νγp.
(2.11)
Remark 2.6. Since
n+1 , implies 0 ≤ γp := n−1
• n ≥ 2, 2 ≤ p < 2(n+3)
• n ≥ 2, 2(n+3)
n+1 < p ≤ 2n
• n ≥ 2, 2n
n−2 ≤ p ≤ ∞, implies 1
• n = 1, 2 ≤ p < 4, implies γp = 0,
• n = 1, 4 < p ≤ ∞, implies 1
n−2, implies − 1
6 + 2n
2 ≤ γp := − 1
2 − 1
4 < γp := 1
p ≤ 1
2,
2 ( 1
2 − 1
3(n+3) ≤ γp := − 1
2 + n( 1
2 − 1
p) < n−1
2(n+3) ,
3 ( 1
6 + 2n
p) ≤ n−1
2 ,
2 − 1
p ) < 1
2,
we have that γp ≥ 0, for all 2 ≤ p ≤ ∞. This lower bound will be useful in our
further analysis.
Proposition 2.7. Let us consider 1 < p < ∞ and s > 3n
pseudo-multiplier with symbol m satisfying
2 )kh · isF −1
H [m(x,·)ψ(2−k · )]kL2(Rn) < ∞,
kmkl.u.Hs = sup
2 + γp − 1
12 . If Tm is a
2k(s− n
(2.12)
k>0,x∈Rn
then Tm extends to a bounded operator on Lp(Rn). Moreover,
kTmkB(Lp(Rn)) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
(2.13)
10
D. CARDONA AND M. RUZHANSKY
Proof. In order to prove Proposition 2.7 we will decompose the symbol m as
m(x, ν) = m(x, 0) +
∞Xk=0
mk(x, ν), mk(x, ν) := m(x, ν) · 1{2k≤ν<2k+1}.
(2.14)
Let us denote by Tm(j) the pseudo-multiplier associated to mj, for j ≥ 0, and
by T0 the operator with symbol σ ≡ m(x, 0)δν,0. Then we want to show that the
operator series
T0 +Xk
Tm(k)
(2.15)
converges to Tm in the strong topology on B(Lp(Rn)) and
kTmkB(Lp(Rn)) ≤ kT0kB(Lp(Rn)) +Xk
kTm(k)kB(Lp(Rn)).
(2.16)
So, we want to estimate every norm kTm(j)kB(Lp(Rn)). For this, we will use the
fact that for f ∈ C ∞
0 (Rn),
kTm(j)fkLp(Rn) = sup{(Tm(j)f, g)L2(Rn) : kgkLp′ (Rn) = 1}.
(2.17)
In fact, for f and g as above we have
(Tm(k)f, g)L2(Rn) =ZRn
Tm(k)f (x)g(x)dx
=ZRn X2k≤ν<2k+1
=ZRnZRn X2k≤ν<2k+1
m(x, ν)bf (φν)φν(x)g(x)dx
m(x, ν)f (y)φν(y)φν(x)g(x)dydx.
For every x ∈ Rn let us denote the inverse Fourier-Hermite transform of the
sequence {m(x, ν)ψ(2−kν)}ν by F −1
H [m(x,·)ψ(2−k · )]. So, we have
mk(x, ν) = FH(F −1
H [m(x,·)ψ(2−k·)])(ν) =ZRn
F −1
H [m(x,·)ψ(2−k·)](z)φν(z)dz.
(2.18)
Consequently, we can write
(Tm(k)f, g)L2(Rn) =
ZRnZRn X2k≤ν<2k+1ZRn
F −1
H [m(x,·)ψ(2−k · )](z)φν(z)dz
× f (y)φν(y)φν(x)g(x)dydx.
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
11
Now, we have
(Tm(k)f, g)L2(Rn)
sup
≤ X2k≤ν<2k+1
. X2k≤ν<2k+1
x∈RnZRn F −1
x∈RnZRn F −1
sup
H [m(x,·)ψ(2−k · )](z)φν(z)dz
× kfkLpkgkLp′kφνkLpkkφνkLp′
H [m(x,·)ψ(2−k · )](z)φν(z)dz
× kfkLpkgkLp′νγp.
So, we can estimate the operator norm of Tm(k) by
kTm(k)kB(Lp)
. X2k≤ν<2k+1
. X2k≤ν<2k+1
sup
x∈RnZRn F −1
x∈Rn(cid:18)ZRnhzi2sF −1
H [m(x,·)ψ(2−k · )](z)φν(z)dzνγp
H [m(x,·)ψ(2−k · )](z)2dz(cid:19) 1
sup
2
kφν(·)h·i−skL2νγp.
If we denote by θ∞ some real number satisfying kφνkL∞(Rn) . νθ∞, we can
estimate
if we require s > n
12 ≥ 3n
s > 3n
2 + γp − 1
kφν(·)h·i−skL2 ≤ kφνkL∞kh·i−skL2 . νθ∞,
2 . By Remark 2.6, the condition s > n
2 − 1
2 holds true because
2 . Now, if additionally we consider the hypothesis
12 > n
then we have
sup
x∈Rn(cid:18)ZRnhzi2sF −1
H [m(x,·)ψ(2−k · )](z)2dz(cid:19) 1
kTm(k)kB(Lp) . X2k≤ν<2k+1 kmkl.u.Hs · 2−k(s− n
2
2 )νγp+θ∞
2 )+kγp+kθ∞kmkl.u.Hs = 2−k(s− 3n
2 −γp−θ∞).
. 2kn−k(s− n
≤ kmkl.u.Hs · 2−k(s− n
2 ),
(2.19)
Taking into account that
kT0fkLp(Rn) . km(·, 0)kL∞(Rn)kfkLp(Rn),
we obtain the boundedness of T0 on Lp. It is clear that if we want to end the proof,
we need to estimate I :=Pk≥0 kTm(k)kB(Lp(Rn)). As a consequence we obtain
0 ≤ I . kT0kB(Lp) +
2−k(s− 3n
2 −γp−θ∞)kmkl.u.,Hs < ∞,
∞Xk=1
for s > 3n
2 + γp + θ∞. So, we have
kTmkB(Lp) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
12
D. CARDONA AND M. RUZHANSKY
From Remark 2.5 we end the proof because we can take θ∞ = − 1
12 .
Proposition 2.8. Let us consider 1 < p < ∞ and s > 3n
function, and Tm is a pseudo-multiplier with symbol {m(x, ν)}x∈Rn,ν∈Nn
2 )kh · isF [m(x,·)ψ(2−k · )]kL2(Rn) < ∞,
kmkl.u.H s = sup
2k(s− n
k>0,x∈Rn
2 +γp. If m : R2n → C is a
0 satisfying
(cid:3)
then Tm extends to a bounded operator on Lp(Rn). Moreover,
kTmkB(Lp) ≤ C(kmkl.u.,H s + km(·, 0)kL∞(Rn)).
Proof. By following the notation in the proof of Proposition 2.7 we have
(Tm(k)f, g)L2(Rn) =ZRnZRn X2k≤ν<2k+1
m(x, ν)f (y)φν(y)φν(x)g(x)dydx.
For every x ∈ Rn let us write
mk(x, ν) = F −1(F [m(x,·)ψ(2−k · )])(ν) =ZRn
So, we have
F [m(x,·)ψ(2−k · )](z)e2πiν·zdz.
(2.22)
(2.20)
(2.21)
(Tm(k)f, g)L2(Rn)
sup
≤ X2k≤ν<2k+1
. X2k≤ν<2k+1
x∈RnZRn F [m(x,·)ψ(2−k · )](z)dz
x∈RnZRn F [m(x,·)ψ(2−k · )](z)dz
sup
× kfkLpkgkLp′kφνkLpkkφνkLp′
× kfkLpkgkLp′νγp.
So, we can estimate the operator norm of Tm(k) by
kTm(k)kB(Lp)
. X2k≤ν<2k+1
. X2k≤ν<2k+1
sup
x∈RnZRn F [m(x,·)ψ(2−k · )](z)dzνγp
x∈Rn(cid:18)ZRnhzi2sF [m(x,·)ψ(2−k · )](z)2dz(cid:19) 1
sup
2
kh · i−skL2νγp.
By Remark 2.6 we have s > n
by the hypothesis
2 , together with the estimate kh · i−skL2 < ∞, and
x∈Rn(cid:18)ZRnhzi2sF [m(x,·)ψ(2−k · )](z)2dz(cid:19) 1
sup
2
≤ kmkl.u.H s · 2−k(s− n
2 ),
(2.23)
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
13
we deduce that
Since
kTm(k)kB(Lp) . X2k≤ν<2k+1 kmkl.u.H s · 2−k(s− n
2 )νγp
≍ 2kn−k(s− n
2 )+kγp = 2−k(s− 3n
2 −γp).
kT0fkLp(Rn) . km(·, 0)kL∞(Rn)kfkLp(Rn),
we have the boundedness of T0 on Lp. It is clear that if we want to end the proof,
we need to estimate I :=Pk≥0 kTm(k)kB(Lp(Rn)). As a consequence we obtain
0 < I . kT0kB(Lp) +
2−k(s− 3n
2 −γp) sup
x∈Rn kmkl.u.,H s < ∞,
∞Xk=1
for s > 3n
2 + γp. So, we have
kTmkB(Lp) ≤ C(km(x,·)kl.u.,H s + km(·, 0)kL∞(Rn)).
The proof is complete.
(cid:3)
Now, we record explicitly the degree of regularity s considered in the proposi-
tions above.
Theorem 2.9. Let us assume 2 ≤ p < ∞. If Tm is a pseudo-multiplier with
symbol m satisfying (2.20), then under one of the following conditions,
n+1 , and s > sn,p := 3n
n+1 , and s > sn,p := 3n
2 + n−1
n+1 < p ≤ 2n
n−2, and s > sn,p := 3n
n−2 ≤ p < ∞, and s > sn,p := 3n−1
• n ≥ 2, 2 ≤ p < 2(n+3)
• n ≥ 2, p = 2(n+3)
• n ≥ 2, 2(n+3)
• n ≥ 2,
• n = 1, 2 ≤ p < 4, s > s1,p := 3
2,
• n = 1, p = 4, s > s1,4 := 2,
• n = 1, 4 < p < ∞, s > s1,p := 4
2 + n−1
2(n+3) ,
2 − 1
2 +n( 1
2 − 1
p),
3 + 2
3( 1
2n
2 ( 1
2 − 1
p),
2 − 1
p),
3 ( 1
6 + 2n
2 − 1
p),
the operator Tm extends to a bounded operator on Lp(Rn). For 1 < p ≤ 2, under
one of the following conditions
2n
• n ≥ 2, 2(n+3)
• n ≥ 2,
• n ≥ 2, 1 < p ≤ 2n
• n = 1, 4
• n = 1, 1 < p < 4
n+5 ≤ p ≤ 2, and s > sn,p := 3n
n+2 ≤ p ≤ 2(n+3)
3 ≤ p < 2, s > s1,p := 3
2,
3, s > s1,p := 4
3+ 2
n+2, and s > sn,p := 3n−1
n+5 , and s > sn,p := 3n
2 − 1
p ),
2 + n−1
2 − 1
2 +n( 1
3( 1
2 − 1
2 ( 1
p),
6 + 2n
3 ( 1
2 − 1
p),
2 − 1
p ),
the operator Tm is Lp-bounded. Moreover, in general:
• for every 4
boundedness of Tm.
3 < p < 4 and every n, the condition s > 3n
2
implies the Lp-
14
D. CARDONA AND M. RUZHANSKY
If the symbol m of the pseudo-multiplier Tm satisfies the Hormander condition
(2.12), in order to guarantee the Lp-boundedness of Tm, in every case above we
can take s > sn,p − 1
12 implies the Lp-boundedness of Tm
for all 4
12. However, s > 3n
2 − 1
3 < p < 4.
Proof. In view of the Propositions 2.7 and 2.8 and considering the following values
for γp: (according to Lemma 2.4),
p), if n ≥ 2, 2 ≤ p < 2(n+3)
2 − 1
2 ( 1
• γp = n−1
n+1 ,
• γp = − 1
6 + 2n
3 ( 1
2 − 1
2 + n( 1
2 − 1
• γp = − 1
• γp = 0, if n ∈ N, 4
2 − 1
• γp = − 1
p), if n ≥ 2, 2(n+3)
p), if n ≥ 2, 2n
3 < p < 4,
p), if n = 1, 4 < p < ∞,
6 + 2
n+1 < p ≤ 2n
n−2,
n−2 ≤ p < ∞,
3( 1
the proof ends if we take into account that γp = γp′ and
n−2 ⇒ 2n
n+2 ≤ p′ ≤ 2(n+3)
n+5 ,
• n ≥ 2, 2(n+3)
• n ≥ 2, 2(n+3)
• n ≥ 2, 2n
• 4
• 1 < p < 4
• 4
n+5 ≤ p ≤ 2 ⇒ 2 ≤ p′ < 2(n+3)
n+1 ,
n+1 < p ≤ 2n
n−2 ≤ p < ∞ ⇒ 1 < p′ ≤ 2n
n+2,
3 ≤ p ≤ 2 ⇒ 2 ≤ p′ ≤ 4,
3 ⇒ 4 < p′ < ∞,
3 < p < 4 ⇔ 4
3 < p′ < 4.
The proof is complete.
Remark 2.10. Let us note that for n ≥ 2 and p = 2(n+3)
2 + n−1
provided that m satisfies (2.20). Indeed, from Remark 2.6, if 2 ≤ r < 2(n+3)
ω ≤ 2n
n+1 , the condition s > sn,p =
n+1 (Rn)
n+1 <
2(n+3) , implies the boundedness of a pseudo-multiplier Tm on L
n−2, then
2(n+3)
(cid:3)
3n
lim
rր 2(n+3)
n+1
γr = lim
ωց 2(n+3)
n+1
γω =
n − 1
2(n + 3)
.
(2.24)
From the real interpolation we obtain γ 2(n+3)
2(n+3) . So, by Proposition 2.8
the condition s > sn,p = 3n
2(n+3) implies the Lp(Rn)-boundedness of Tm when
p = 2(n+3)
2 and the condition
s > s1,4 = 2 implies the boundedness of Tm on L4(Rn). So, this remark and
Theorem 2.9 proves Theorem 1.1.
n+1 . Now, if n = 1, a similar analysis shows that γ4 < 1
2 + n−1
n+1
= n−1
In the following proposition we exhibit a class of symbols providing Lp-pseudo-
multipliers.
Proposition 2.11. Let us consider a complex-valued function m on Rn× Rn and
a pseudo-multiplier Tm with symbol {m(x, ν)}x∈Rn,ν∈Nn
0 . If m satisfies the symbol
inequalities
(2.25)
for ρ = [3n/2] + 1, then Tm extends to a bounded operator on L2(Rn). Moreover,
for ρ = 2n + 1 we have the Lp(Rn)-boundedness of Tm for all 1 < p < ∞.
ξ m(x, ξ) ≤ Cα(1 + ξ)−α,
∂α
α ≤ ρ,
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
15
Proof. For the proof, we will use that the Sobolev space H s(Rn) defined by those
functions g satisfying kgkH s(Rn) := khzis(F g)kL2(Rn) < ∞, has the equivalent
norm
(2.26)
H s(Rn) := Xβ≤s
kgk′
k∂β
ξ gkL2(Rn),
when s is an integer (see, e.g. [9], p. 163). We will show that
sup
k>0,x∈Rn
2k(ρ− n
2 )km(x,·)ψ(2−k · )kH ρ = sup
k>0,x∈Rn km(x, 2k·)ψ( · )kH ρ < ∞,
(2.27)
provided that ρ is an integer. From the estimate
km(x, 2k·)ψ( · )kH ρ ≍ km(x, 2k·)ψ( · )k′
ξ (m(x, 2k·)ψ( · ))kL2(Rn),
(2.28)
we will estimate the L2-norms of the derivatives ∂β
ξ (m(x, 2k·)ψ( · ))(ξ). Because
the function ψ is supported in some closed interval not containing the origin, the
function ψ( · ) is smooth. By the Leibniz rule we have
H ρ = Xβ≤s
k∂β
So, we obtain
∂β
ξ (m(x, 2kξ)ψ(ξ)) = Xα≤β
ξ (m(x, 2k·)ψ( · ))kL2 ≤ Xα≤ρ
k∂β
2kα(∂α
ξ m)(x, 2kξ)∂β−α
ξ ψ(ξ).
Cαk∂β−α
ξ ψ( · )kL2,
(2.29)
where we have used that (2.25) implies the estimate 2kα(∂α
ξ m)(x, 2k·) ≤ Cα,
for k large enough. Now, (2.27) follows by summing both sides of (2.29) over
β ≤ ρ. We finish the proof by observing that every sn,p defined in Theorem 2.9,
satisfies the upper bound sn,p ≤ 2n, and we can obtain the Lp-boundedness of Tm
by taking ρ > sn,p with ρ = 2n + 1. A similar analysis shows that ρ = [3n/2] + 1
implies the L2−boundedness of Tm.
(cid:3)
Corollary 2.12. Let us consider a complex-valued function m on Rn × Zn and a
0 . If m satisfies the discrete
pseudo-multiplier Tm with symbol {m(x, ν)}x∈Rn,ν∈Nn
difference conditions
∆α
(2.30)
for ρ = [3n/2] + 1, then Tm extends to a bounded operator on L2(Rn). Moreover,
for ρ = 2n + 1 we have the Lp(Rn)-boundedness of Tm for all 1 < p < ∞.
Proof. Let us define for every z0 ∈ Rn, the function mz0 given by mz0(ν) =
m(z0, ν). Then we have the estimates
ν m(x, ν) ≤ Cα(1 + ν)−α, α ∈ Nn
0 , α ≤ ρ,
∆α
ν mz0(ν) ≤ Cα(1 + ν)−α, α ∈ Nn
(2.31)
From Corollary 4.5.7 of [22], there exists a suitable function mz0 defined on Rn
such that mz0Zn = mz0 and additionally satisfying the conditions,
0 , α ≤ ρ.
ξ mz0(ξ) ≤ Cα(1 + ξ)−α, α ∈ Nn
∂α
0 , α ≤ ρ.
(2.32)
16
D. CARDONA AND M. RUZHANSKY
The function m defined by m(z0, ξ) := mz0(ξ) satisfies (2.25), and by Proposition
2.11 we obtain the L2-boundedness of the pseudo-multiplier T m with symbol
0 , if ρ = [3n/2] + 1 (or the Lp-boundedness, for all 1 < p < ∞
{ m(x, ν)}x∈Rn,ν∈Nn
if ρ = 2n + 1). We finish the proof by observing that T m = Tm in view of the
equality sets { m(x, ν)}x∈Rn,ν∈Nn
(cid:3)
2.3. (Lp, Lq)-boundedness of spectral pseudo-multipliers. Let us assume
n ∈ N, arbitrary but fixed. Let us define the set 2N0 + n := {2m + n : m ∈ N0}.
We will consider continuous functions m(x, ξ) defined on Rn
x × Rξ and we will de-
note by m(x, ℓ) the restriction of m(x, ξ) to the set Rn×(2N0 + n), so that x ∈ Rn
and ℓ ∈ 2N0+n. If we set F : L2(−∞,∞) → L2(−∞,∞) for the one-dimensional
Fourier transform, we will consider symbols m(x, ℓ) := m(x, ξ)Rn×(2N0+n) satisfy-
ing the, so called, Hormander condition of order s > 0,
0 .
0 = {m(x, ν)}x∈Rn,ν∈Nn
kmkl.u.,hs := sup
x∈Rn,k>0
2k(s− n
2 )
∞Z−∞
hti2sF [m(x,·)ψ(2−k · )](t)2dt
1
2
< ∞,
(2.33)
where the function ψ ∈ D(0,∞) satisfies ψ(t) = 1 for all t ∈ [1, 2]. With the
previous notation we want to investigate the Hormander condition for pseudo-
multipliers of the form
m(x, H)f (x) :=
m(x, ℓ)Pℓf (x),
(2.34)
∞Xℓ=0
where we have denoted by Pℓ the orthogonal projection to the subspace generated
by the set {φν : ν = ℓ}. For symbols m(x, ℓ) = m(ℓ) depending only on the ℓ
variables we have used in (1.6) the term radial symbols. If it depends on x we can
talk about them as spectral pseudo-multipliers. In the next theorem, we prove
that for symbols satisfying the Hormander condition of order s, for s suitable,
the corresponding spectral pseudo-multipliers are bounded operators from Lp(Rn)
into Lq(Rn) when 1
p = 1. We will denote δ(p) := n 1
2 and q = p′.
q + 1
p − 1
2 − 1
Theorem 2.13. Let us consider a function m satisfying (2.33). Let m(x, H) be
a spectral pseudo-multiplier with symbol {m(x, ℓ)}x∈Rn,ℓ∈2N0+n. Under one of the
following conditions
2n
• n ≥ 2, 1 ≤ p ≤ 2n
• n ≥ 2,
• n = 1, 4
• n = 1, p = 4
• n = 1, 1 < p < 4
n+2 and s > n+1
2 + δ(p),
n+2 < p ≤ 2 and s > 3n
2 ,
3 < p ≤ 2, s > 2 − 1
p ,
3, s > 1 + 1
3p ,
3, s > 3
2,
the operator m(x, H) extends to a bounded operator from Lp(Rn) into Lp′(Rn).
Proof. In order to prove Theorem 2.13 we will split the symbol m as
m(x, ℓ) =
∞Xk=0
mk(x, ℓ), mk(x, ℓ) := m(x, ℓ) · 1{2k≤ℓ<2k+1}.
(2.35)
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
17
Let us denote by Tm(j) the pseudo-multiplier associated to mj, for j ≥ 0. Then
the operator series
T0 +Xk
Tm(k)
(2.36)
converges to Tm in the strong topology on B(Lp(Rn), Lp′(Rn)) and
km(x, H)kB(Lp(Rn),Lp′ (Rn)) ≤Xk
kTm(k)kB(Lp(Rn),Lp′ (Rn)).
(2.37)
So, we want to estimate every norm kTm(j)kB(Lp(Rn),Lp′ (Rn)). For this, we will
use the fact that for f ∈ C ∞
0 (Rn),
kTm(j)fkLp′ (Rn) = sup{(Tm(j)f, g)L2(Rn) : kgkLp(Rn) = 1}.
(2.38)
In fact, for f and g as above we have
(Tm(k)f, g)L2(Rn) =ZRn
Tm(k)f (x)g(x)dx
=ZRn X2k≤ℓ<2k+1
= X2k≤ℓ<2k+1ZRn
m(x, ℓ)Pℓf (x)g(x)dx
m(x, ℓ)Pℓf (x)g(x)dx.
For every x ∈ Rn let us denote the one-dimensional Fourier transform of
m(x,·)ψ(2−k · ) by F [m(x,·)ψ(2−k · )]. So, we have
mk(x, ℓ) = F −1(F [m(x,·)ψ(2−k · )])(ℓ) =
Consequently, we have
∞Z−∞
F [m(x,·)ψ(2−k · )](t)ei2πℓ·tdt.
(2.39)
(Tm(k)f, g)L2(Rn)
≤ X2k≤ℓ<2k+1
∞Z−∞
sup
x∈Rn
F [m(x,·)ψ(2−k · )](t)dt
× kPℓfkLp′kgkLp.
Now, let us fix n ∈ N, n ≥ 2. By taking into account the Karadzhov's estimate
(see Thangavelu [31], p. 268)
kPℓfkLp′ (Rn) ≤ Cpℓδ(p)− 1
2kfkLp, 1 ≤ p ≤
2n
n + 2
, n ≥ 2,
(2.40)
18
D. CARDONA AND M. RUZHANSKY
we can estimate the operator norm of Tm(k) by
kTm(k)kB(Lp,Lp′ )
if we impose s > 1
2
sup
sup
x∈Rn
hti2sF [m(x,·)ψ(2−k · )](t)2dt
∞Z−∞
. X2k≤ℓ<2k+1
F [m(x,·)ψ(2−k · )](t)dtℓδ(p)− 1
x∈Rn
∞Z−∞
. X2k≤ℓ<2k+1
hti2sF [m(x,·)ψ(2−k · )](t)2dt
x∈Rn
∞Z−∞
kTm(k)kB(Lp,Lp′ ) . X2k≤ℓ<2k+1 kmkl.u.,hs · 2−k(s− n
sup
1
2
then we have
2 )+k(δ(p)− 1
2 )
2. If additionally we consider the hypothesis (2.33), that is,
1
2
kh·i−skL2(−∞,∞)ℓδ(p)− 1
2 ,
≤ kmkl.u.,hs · 2−k(s− n
2 ),
(2.41)
≍ 2k−k(s− n
As a consequence we obtain
2 )+k(δ(p)− 1
0 ≤
∞Xk=1
kTm(k)kB(Lp,Lp′ ) .
for s > n+1
2 + δ(p). So, we have
2 −δ(p))kmkl.u.,hs.
2 )kmkl.u.,hs = 2−k(s− n+1
∞Xk=1
2−k(s− n+1
2 −δ(p))kmkl.u.,hs < ∞,
2n
2n
n + 2
n+2 → L
.
2n
km(x, H)kB(Lp,Lp′ ) ≤ Ckmkl.u.,hs, 1 ≤ p ≤
n+2) = 1
2 , and consequently Tm : L
Let us note that δ( 2n
n−2 extends to a
bounded operator provided that s > n+2
2 then from Theorem
2.9 we obtain the L2-boundedness of m(x, H). Thus, by the real interpolation we
obtain the boundedness of m(x, H) from Lp into Lp′ for 2n
n+2 ≤ p ≤ 2 provided
that s > 3n
2 .
Now, if n = 1, by the Cauchy-Schwarz inequality we have for every ℓ ∈ N0 the
2 . If we assume s > 3n
estimate,
kPℓfkLp′ (R) = k(f, φℓ)L2(R)φℓkLp′ (R) ≤ kφℓk2
Lp′ (R)kfkLp(R).
(2.42)
By Thangavelu's Lemma 2.1, we can estimate the Lp′(R)-norm of the function
Pℓf as follows:
2p′ − 1
kPℓfkLp′ (R) . ℓ2( 1
kPℓfkLp′ (R) . ℓ2(− 1
6p′ − 1
1
2 − 1
4 )kfkLp(R) = ℓ
12 )kfkLp(R) = ℓ− 1
pkfkLp(R),
2 + 1
3pkfkLp(R), 1 < p <
< p ≤ 2,
4
3
(2.43)
.
(2.44)
4
3
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
19
Recalling that for all n ∈ N we have,
(Tm(k)f, g)L2(Rn)
≤ X2k≤ℓ<2k+1
∞Z−∞
sup
x∈Rn
F [m(x,·)ψ(2−k · )](t)dt
we obtain, with n = 1,
kTm(k)fkLp′ (Rn) ≤ X2k≤ℓ<2k+1
× kPℓfkLp′ (Rn)kgkLp(Rn),
2−k(s− n
2 )kmkl.u.,hskPℓfkLp′ (Rn)
∞Xk=1
∞Xk=1
∞Xk=1
∞Xk=1
Thus, by (2.43) and the estimate above we have,
. 2−k(−1+s− n
2 )kmkl.u.,hskPℓfkLp′ (Rn).
0 ≤ km(x, H)kB(Lp,Lp′ ) ≤
=
kTm(k)kB(Lp,Lp′ ) .
2−k(−1+s− n
2 − 1
2 + 1
p )kmkl.u.,hs
∞Xk=1
p )kmkl.u.,hs < ∞,
2−k(s−2+ 1
for s > 2 − 1
p, when 4
3 < p ≤ 2, and
0 ≤ km(x, H)kB(Lp,Lp′ ) ≤
=
kTm(k)kB(Lp,Lp′ ) .
2−k(−1+s− n
2 + 1
2 − 1
3p )kmkl.u.,hs
∞Xk=1
3p )kmkl.u.,hs < ∞,
2−k(s−1− 1
4
3p , and 1 < p < 4
3, in view of (2.44). The (L
3 satisfying s > 2− 1
for s > 1 + 1
m(x, H) now follows by the real interpolation for s > 3
there exists p1 > 4
1 < p0 < 4
from Lp0(Rn) into Lp′
inequality p0 < 4
a bounded extension from L
proof.
3 , L4)-boundedness of
2 . In fact, if we fix s > 3
2
2 . We also have the existence of p0 > 0,
p1
. Thus, m(x, H) admits bounded extensions
1(Rn) respectively. By the
3 < p1 and the real interpolation we deduce that m(x, H) has
3 (Rn) into L4(Rn). Thus, we have completed the
(cid:3)
0(Rn) and from Lp1(Rn) into Lp′
> 3
3 such that s > 3
2 > 1 + 1
3p0
4
Now, by the real interpolation we give the following general (Lp, Lq) bounded-
ness theorem.
Theorem 2.14. Let us consider a function m = m(x, ℓ) satisfying (2.33). Let
m(x, H) be a spectral pseudo-multiplier with symbol {m(x, ℓ)}x∈Rn,ℓ∈2N0+n. Under
one of the following conditions
• n ≥ 2, 1 < p ≤ 2n
• n ≥ 2,
• n = 1, 4
n+2 and s > 3n−1
2 + n( 1
n+2 < p ≤ 2 and s > 3n
2 ,
3 ≤ p < 2, s > 3
2,
2n
2 − 1
p),
20
D. CARDONA AND M. RUZHANSKY
• n = 1, 1 < p < 4
3, s > 1 + 1
3p ,
the operator m(x, H) extends to a bounded operator from Lp(Rn) into Lq(Rn), for
all p ≤ q ≤ p′.
Proof. Let us observe that δ( 2n
2 − 1
1
2.13, under one of the following conditions
n+2 and that
p ≤ 0 for 1 < p ≤ 2. With these inequalities in mind, from Theorems 2.9 and
• n ≥ 2, 1 < p ≤ 2n
n+2,
1
s >
2 −
for 1 ≤ p ≤ 2n
2 ≤ δ(p) ≤ n−1
) = max{
n+2) = 1
3n − 1
1
2 −
1
p
)},
3n − 1
2
+ δ(p),
n + 1
2
2
+ n(
2
+ n(
1
p
n+5 ≤ p ≤ 2,
• n ≥ 2, 2(n+3)
3n
2
s >
= max{
n+2 ≤ p ≤ 2(n+3)
n+5 ,
• n ≥ 2, 2n
3n
2
,
3n
2
+
n − 1
2
(
1
2 −
1
p
)},
s >
3n
2
3 ≤ p < 2,
• n = 1, 4
= max{
3n
2
,
3n
2 −
1
6
+
2n
3
(
1
2 −
1
p
)},
s >
3
2
= max{
3
2
, 2 −
1
p},
• n = 1, 1 < p < 4
3,
s > 1 +
1
3p
= max{1 +
1
3p
,
4
3
+
2
3
(
1
2 −
1
p
)},
the spectral pseudo-multiplier m(x, H) extends to a bounded operator from Lp(Rn)
into Lp′(Rn) and also we have its boundedness from Lp(Rn) into Lp(Rn). So, by
the Riesz-Thorin interpolation theorem we deduce the boundedness of m(x, H)
from Lp(Rn) into Lq(Rn) for all p ≤ q ≤ p′. So, we finish the proof.
(cid:3)
2.4. Lower bounds for the operator norm of multipliers on Lp spaces.
Now, we estimate from below the operator norm of multipliers associated to the
harmonic oscillator.
Theorem 2.15. Let 1 ≤ p ≤ ∞. Let us assume that Tm is a multiplier associated
to the harmonic oscillator. If Tm is a bounded operator on Lp(Rn), then we have
the following lower bound for the Lp-operator norm of Tm,
kTmkB(Lp(Rn)) ≥ sup
ν∈Nn m(ν).
Proof. For the proof we can take advantage of the orthogonality properties of the
Hermite functions φν, ν ∈ Nn
0 . By definition we have
Tm(φν) = m(ν)φν.
(2.45)
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
21
As consequence we obtain
kTmkB(Lp(Rn)) ≥ kTm(
φν
kφνkLp
)kLp(Rn) = m(ν).
Thus, we end the proof.
(cid:3)
3. Compactness of pseudo-multipliers
3.1. L2-compactness of multipliers. Now, we use the Fourier analysis pro-
duced by the harmonic oscillator in order to characterise the L2-compactness of
multipliers. The following is an analogue of a criterion very well known in other
settings.
Theorem 3.1. Let us assume that Tm is a bounded multiplier on L2(Rn). Then,
Tm is a compact operator on L2(Rn) if and only if limν→∞ m(ν) = 0.
Proof. In order to prove the theorem, let us first assume that Tm is an L2-compact
operator. If f ∈ L2(Rn), by the Plancherel theorem we have
L2(Rn) = Xν∈Nn
kfk2
0
(f, φν)L22.
(3.1)
Consequently, we have (f, φν) → 0 as ν → ∞. So, we conclude that in L2(Rn),
0 converges weakly to zero. By the compactness of Tm the
the sequence {φν}ν∈Nn
0 converges to zero in the L2-norm. So,
sequence {T (φν)}ν∈Nn
lim
ν→∞kTmφνkL2(Rn) = lim
ν→∞m(ν) = 0.
(3.2)
For the proof of the converse assertion, let us assume that the sequence {m(ν)}ν∈Nn
tends to zero as ν → ∞. In order to show that Tm is compact, we will approxi-
mate it with operators of finite rank. So, let us define the sequence of finite rank
operators Tm(k), k ∈ N, by
0
Tm(k)f := Xν≤k
m(ν)bf (φν)φν.
m(ν)2(f, φν)L22 ≤ sup
ν≥k m(ν)2kfk2
L2(Rn).
(3.3)
(3.4)
(3.5)
(cid:3)
By the orthogonality of the Hermite functions, we have
kTm(k)f − Tmfk2
L2(Rn) = Xν≥k
So, we obtain
k→∞kTm(k) − TmkB(L2) ≤ lim
lim
k→∞
ν≥k m(ν) = 0.
sup
With the last line we finish the proof.
22
D. CARDONA AND M. RUZHANSKY
3.2. Lp-compactness and Lp-boundedness for multipliers via Littlewood-
Paley theory. In the preceding subsection we have characterised the compact-
ness on L2(Rn) of multipliers with the Plancherel theorem as a fundamental tool.
In order to investigate the Lp-compactness of multipliers for 1 < p < ∞, but
p 6= 2, we will use the Littlewood-Paley theorem (which is a partial substitute of
the Plancherel theorem on Lp-spaces) associated to dyadic decompositions of the
spectrum of the harmonic oscillator. The main notion in the Littlewood-Paley
theory is the concept of a dyadic decomposition. Here, the sequence {ψl}l∈N0 is a
dyadic decomposition, defined as follows: we choose a function ψ ∈ D(0,∞) sup-
ported in [1/2, 1], ψ = 1 on [2/3, 4/5]. Denote by ψl the function ψl(t) = ψ(2−lt),
t ∈ R. For some smooth compactly supported function ψ0 we have
ψl(λ) = 1,
for every λ > 0.
(3.6)
Xl∈N0
Now we present the Littlewood-Paley Theorem in the form of the following result
(see Theorems 7.1 and 7.3 of [4] and Proposition 5 of [18]).
Theorem 3.2. Let 1 < p < ∞ and for every l ∈ N0, let us consider the multipliers
Tψl given by
Then there exist constants 0 < cp, Cp < ∞ depending only on p such that
Tψlf (x) := X2l≤hνi<2l+1
ψl(hνi)φν(x)bf (φν), hνi := (1 + ν2)
cpkfkLp(Rn) ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
Tψlf (x)2! 1
∞Xl=0
≤ CpkfkLp(Rn),
holds for all f ∈ Lp(Rn).
1
2 .
(3.7)
(3.8)
The following Lp multiplier theorem provides sufficient conditions for the Lp-
boundedness of multipliers (different from the Hormander-Mihlin condition) and
their Lp-compactness.
Theorem 3.3. Let us assume that Tm is a multiplier and let 1 < p < ∞. Let us
assume that there exists a sequence {νl} satisfying: 2l ≤ νl < 2l+1, m(νl) 6= 0
for every l ∈ N0 and
= K 6= 0,
(3.9)
for every sequence {ν′
sequences νl and ν′
• if kmkL∞(Nn
l < 2l+1 (the constant K depends on the
l). Then,
0 ) < ∞, then the operator Tm extends to a bounded operator
on Lp(Rn) and
(3.10)
• if m(ν) → 0 as ν → ∞, then the operator Tm extends to a compact
kTmkB(Lp(Rn)) ≤ CkmkL∞(Nn
0 ).
operator on Lp(Rn).
lim
l→∞
m(ν′
l)
m(νl)
l} where 2l ≤ ν′
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
23
Proof. Let us assume that f ∈ Lp(Rn), 1 < p < ∞. By the Littlewood-Paley
Theorem (see Theorem 3.2 above) we have
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
∞Xl=0
Taking into account both, that m is bounded and the condition (3.9) we have
1
1
.
1
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
kTmfkLp(Rn) .(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
TψlTmf (x)2! 1
∞Xl=0
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ψl(hνi)m(ν)φν(x)bf (φν)2
∞Xl=0
X2l≤hνi<2l+1
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
ψl(hνi)m(ν)φν(x)bf (φν)2
X2l≤hνi<2l+1
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
φν(x)bf (φν)2
∞Xl=0
m(νl)2 X2l≤hνi<2l+1
0 )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
φν(x)bf (φν)2
∞Xl=0
X2l≤hνi<2l+1
2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Lp(Rn)
0 )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ψl(hνi)φν(x)bf (φν)2
∞Xl=0
X2l≤hνi<2l+1
≤ kmkL∞(Nn
. kmkL∞(Nn
0 ) kfkLp(Rn) ,
ψl(hνi)
m(ν)
m(νl)
ψl(hνi)
m(ν)
m(νl)
≍ kmkL∞(Nn
1
1
where in the last line we have used the Littlewood-Paley Theorem 3.2 again. So,
we have proved the first part of the theorem. Now, if in addition m(ν) → 0 as
ν → ∞, we will prove that Tm can be approximated by rank finite operators
and consequently we obtain the compactness of Tm. Let us define for every k ∈ N
the operator,
(3.11)
Tm(k)f := Xhνi≤k
m(ν)bf (φν)φν.
A similar argument as in the proof of the first assertion shows us that the estimate
kTmf − Tm(k)fkLp(Rn) . sup
hνi≥2k m(ν)kfkLp,
holds true. Consequently we have the norm estimates
kTm − Tm(k)kB(Lp(Rn)) . sup
hνi≥2k m(ν) → 0 as ν → ∞.
So, we finish the proof.
(3.12)
(3.13)
(cid:3)
24
D. CARDONA AND M. RUZHANSKY
Remark 3.4. Let us note that m(ν) := (1 + ν)iτ , τ ∈ R, satisfies (3.9) and
clearly it is a bounded symbol. By the preceding theorem we conclude that Tm
extends to a bounded operator on Lp(Rn), 1 < p < ∞. Also, is easy to see that
mκ(ν) := (1 + ν)−κ for κ > 0 satisfies (3.9) and mκ(ν) → 0 as ν → ∞.
Consequently every operator Tmκ extends to a compact operator on Lp(Rn) for
all 1 < p < ∞.
4. Lp-boundedness for multilinear pseudo-multipliers
In this section we analyse the boundedness of multilinear pseudo-multipliers
on Lebesgue spaces which are operators defined by
Tm(f1,· · · , fκ)(x) := Xν∈Nnκ
0
m(x, ν)bf1(φν1) · · ·bfκ(φνκ )φν1(x) · · · φνκ (x),
(4.1)
0 , and x ∈ Rn.
for all (f1, f2,· · · , fκ) ∈ D(Rn)κ, where ν := (ν1,· · · , νκ), νi ∈ Nn
In order to prove a general theorem on the boundedness of these operators, we
establish the following proposition.
Proposition 4.1. Let us consider a pseudo-multiplier Tm defined on D(Rn)κ and
let m : Rn × Nnκ
0 → C be its symbol. Let us assume that for s > 0, m satisfies
the condition
kmkl.u.,Hs := sup
and that κ ≥ 2. Then
H [m(x,·)ψ(2−k·)](z)kL2(Rnκ
z ) < ∞, (4.2)
2 )khzisF −1
2k(s− nκ
k>0, x∈Rn
2 + (κ − 1)γ∞, the operator Tm extends to a bounded multilinear
1. If s > 3nκ
operator from L1 × L∞ × · · · × L∞ × L∞ into L1(Rn), and
kTmkB(L1×(L∞)κ−1, L1) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
2. If s > 3nκ
2 + (κ−1)n
4 − 1
12 , the operator Tm extends to a bounded multilinear
operator from L2 × L∞ × · · · × L∞ × L∞ into L2(Rn), and
kTmkB(L2×(L∞)κ−1, L2) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
3. If s > 3nκ
2 + (n−1)(κ−1)
+ γp, γp defined as in (2.11), the operator Tm
extends to a bounded multilinear operator from Lp × L∞ × · · ·× L∞ × L∞
into Lp(Rn), and
2
(4.3)
(4.4)
kTmkB(Lp×(L∞)κ−1, Lp) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)),
for all 2 < p ≤ ∞.
Proof. We proceed with the proof of the first statement. Since
kTmfkL1(Rn) = sup
kgkL∞ =1(Tmf, g),
(4.5)
(4.6)
similar to the previous section we will estimate (Tmf, g) for kgkL∞ = 1. Now,
for f := (f1, f2,· · · , fκ) ∈ D(Rn)κ we have
(Tmf, g) ≤ (T0f, g) +
(Tm(k)f, g),
(4.7)
∞Xk=0
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
25
where Tm(k) is the pseudo-multiplier associated to the symbol
mk(x, ν) = m(x, ν) · 1[2k,2k+1)(ν),
and T0 is the operator with symbol m(x, 0)δν,0. For zj ∈ Rn, z = (z1, z2,· · · , zκ) ∈
Rnκ, and φν(z) = φν1(z1) · · · φνκ (zκ), the inversion formula for the Fourier-Hermite
transform gives
(Tm(k)f, g) =(cid:12)(cid:12)(cid:12)(cid:12)ZRn
= ZRn X2K ≤ν<2k+1
≤ X2k≤ν<2k+1 ZRn
= X2k≤ν<2k+1 ZRnZRnκ
≤ X2k≤ν<2k+1
sup
Tm(k)f (x)g(x)dx(cid:12)(cid:12)(cid:12)(cid:12)
mk(x, ν)bf1(φν1) · · ·bfκ(φνκ)φν1(x) · · · φνκ (x)g(x)dx
mk(x, ν)bf1(φν1) · · ·bfκ(φνκ)φν1(x) · · · φνκ (x)g(x)dx
φν(z)F −1
H [mk(x,·)](z)dz
× bf1(φν1) · · ·bfκ(φνκ)φν1(x) · · · φνκ (x)g(x)dx
H [mk(x,·)](z)dz
x∈RnZRnκ φν(z)F −1
κYj=2
× kf1kL1kφν1kL∞
kfjkL∞kφνjkL1 · kgkL∞ZRn φν1(x) · · · φνκ (x)dx.
Taking into account that κ ≥ 2 we write,
ZRn φν1(x) · · · φνκ (x)dx ≤ kφν1kL2kφν2kL2
κYj6=1,2
kφνjkL∞ . 1,
(4.8)
where we have used Remark 2.5 for the terms in the products. Consequently,
(Tm(k)f, g) ≤ X2k≤ν<2k+1
H [mk(x,·)](z)dz
sup
x∈RnZRnκ φν(z)F −1
κYj=2
× kf1kL1kφν1kL∞
kfjkL∞kφνjkL1 · kgkL∞.
26
D. CARDONA AND M. RUZHANSKY
Hence we have the following estimate for the norm of Tm(k),
kTm(k)kB(L1×(L∞)κ−1, L1)
. X2k≤ν<2k+1
≤ X2k≤ν<2k+1
sup
x∈RnZRnκ φν(z)F −1
x∈Rn khzisF −1
sup
H [mk(x,·)]kL2(Rnκ )
H [mk(x,·)](z)dz × kφν1kL∞
κYj=2
kφνjkL1
× khzi−skL2(Rnκ )kφνkL∞ × kφν1kL∞
κYj=2
kφνjkL1.
Since
kφνkL∞ × kφν1kL∞
κYj=2
kφνjkL1 ≤ kφν1k2
L∞
kφνjkL1kφνjkL∞
(4.9)
κYj=2
and by using the following estimate in Lemma 2.4 for p = ∞,
kφνjkL1(Rn)kφνjkL∞(Rn) . νjγ∞, 2 ≤ j ≤ κ,
(4.10)
(γ∞ = n−1
2
for n ≥ 2 and for n = 1, γ∞ = 1/6), we obtain
kφνkL∞ × kφν1kL∞
κYj=2
kφνjkL1 . kφν1k2
L∞
κYj=2
νjγ∞ . νγ∞(κ−1).
(4.11)
Let us note that in the last estimates we have used that Remark 2.5 implies
kφν1kL∞ = O(1). Consequently we have for s > nκ
2 ,
kTm(k)kB(L1×(L∞)κ−1, L1)
H [mk(x,·)]kL2(Rnκ )khzi−skL2(Rnκ )ν(κ−1)γ∞
2 )2k(κ−1)γ∞(κ−1)kmkl.u.Hs ≍ 2−k(s− nκ
2 )+k(κ−1)γ∞+knκkmkl.u.Hs
x∈Rn khzisF −1
sup
2−k(s− nκ
. X2k≤ν<2k+1
. X2k≤ν<2k+1
= 2−k(s− 3nκ
2 −(κ−1)γ∞)kmkl.u.Hs.
So, we obtain the following upper bound for the series
∞Xk=1
kTm(k)kB(L1×(L∞)κ−1, L1) ≤ kmkl.u.Hs ×
2−k(s− 3nκ
2 −(κ−1)γ∞)
∞Xk=1
which converges provided that s > 3nκ
2 + (κ − 1)γ∞. Now, it is easy to see that
kT0kB(L1×(L∞)κ−1, L1) . km(·, 0)kL∞(Rn).
As a consequence we get
kTmkB(L1×(L∞)κ−1, L1) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
27
So, we finish the proof of the first statement. For the proof of the second state-
ment, we observe that
(Tm(k)f, g)
≤ X2k≤ν<2k+1
. X2k≤ν<2k+1
sup
x∈RnZRnκ φν(z)F −1
κYj=2
x∈RnZRnκ φν(z)F −1
× kf1kL2kφν1kL2
sup
H [mk(x,·)](z)dz
kfjkL∞kφνjkL1 · kgkL2 · (ZRn φν1(x) · · · φνκ(x)2dx)
1
2 .
H [mk(x,·)](z)dz
κYj=2
12 ,
× kf1kL2
kfjkL∞kφνjkL1 · kgkL2ν− 1
where we have estimated (RRn φν1(x) · · · φνκ(x)2dx)
12 . ν− 1
kφνikL∞(Rnκ ) . νi− 1
1
12 ,
be obtained as follows. If νi := max1≤j≤κ νj, similar to Remark 2.5 we have
2 . ν− 1
12 . This estimate can
when ν is large enough. On the other hand, if k 6= i, it follows that
(ZRn φν1(x) · · · φνκ(x)dx)
1
2 . kφνkkL2kφνikL∞
κYj6=i,k
kφνjkL∞ . ν− 1
12 ,
(4.12)
where we have used the crude estimate kφνjkL∞ = O(1) for j 6= k, i and that the
L2−norm of the function φνk is normalised. By using this and Lemma 2.2 for
p = 1 we obtain
kTm(k)kB(L2×(L∞)κ−1, L2)
. X2k≤ν<2k+1
. X2k≤ν<2k+1
. X2k≤ν<2k+1
sup
x∈RnZRnκ φν(z)F −1
x∈Rn khzisF −1
sup
2−k(s− nκ
2 )kmkl.u.Hsν
H [mk(x,·)](z)dz
kφνjkL1ν− 1
12
κYj=2
H [mk(x,·)]kL2(Rnκ )khzi−skL2(Rnκ )
12 ≍ 2−k(s− 3nκ
4 (κ−1)ν− 1
n
κYj=2
νj
n
4 ν− 1
12
2 − n
4 (κ−1)+ 1
12 ).
Now, we only need to proceed as in the first part, in order to obtain the estimate
kTmkB(L2×(L∞)κ−1, L2) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)),
(4.13)
28
D. CARDONA AND M. RUZHANSKY
2 + n
for s > 3nκ
4 (κ − 1) − 1
(Tm(k)f, g) ≤ X2k≤ν<2k+1
sup
x∈RnZRnκ φν(z)F −1
H [mk(x,·)](z)dz
12 . The last statement can be proved by observing that
× kf1kLpkφν1kLp′
kfjkL∞kφνjkL1kgkLp′ (ZRn φν1(x) · · · φνκ (x)pdx)
1
p
κYj=2
. X2k≤ν<2k+1
x∈RnZRnκ φν(z)F −1
sup
H [mk(x,·)](z)dz
× kf1kLpkφν1kLp′kφν1kLp
kfjkL∞kφνjkL1kgkLp′ ,
κYj=2
where in the last line we have used again that the L∞-norm of Hermite functions
is O(1). Now, if we denote by γp the exponent that according to Lemma 2.4
satisfies
(4.14)
and we assume that ν1 := max1≤j≤κ νj (which can be obtained by a simple
permutation of the ν′
kφν1kLpkφν1kLp′ . ν1γp,
js) we obtain ν1 ≍ ν, and the estimate
kφν1kLpkφνkLp′
kφνjkL1kφνjkL∞ . νγp+ (κ−1)(n−1)
2
.
κYj=2
In the last line according to Lemma 2.4 we have used the estimate kφνjkL1kφνjkL∞ .
ν
2 . Now, if we repeat the argument of the first part we obtain
n−1
kTm(k)kB(Lp×(L∞)κ−1, Lp) . X2k≤ν<2k+1
≍ 2−k(s− 3nκ
and consequently the estimate
2−k(s− nκ
2 )kmkl.u.Hs2k (n−1)(κ−1)
2
+kγp
2 − (n−1)(κ−1)
2
−γp)kmkl.u.Hs,
kTmkB(Lp×(L∞)κ−1, Lp) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)),
for s > 3nκ
2 + (n−1)(κ−1)
2
+ γp. Thus we conclude the proof.
(cid:3)
Remark 4.2. The different regularity orders s imposed to obtain the boundedness
of multilinear pseudo-multipliers in Proposition 4.1 for p = 1, 2 or other values of
p, lie in the slight variations that we use for the proof of every specific case. To
be more precise, these differences appear as consequence of the conclusions (4.8)
for p = 1, (4.12) for p = 2, (where we have used strongly that the L2-norm of
every Hermite functions is normalised) and the estimate (4.14) when 2 < p < ∞.
Let us mention that our main strategy in the proof of Proposition 4.5 will be to
use the real interpolation for p between p0 = 1 and p1 = 2 or p between p1 = 2
and arbitrary p with 2 < p < ∞, together with the different regularity orders
imposed in Proposition 4.1.
With a similar proof, as in the previous result, we present the following propo-
sition.
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
29
Proposition 4.3. Let us consider a pseudo-multiplier Tm defined on D(Rn)κ
0 where m : Rn × Rnκ → C satisfies the
with symbol m = {m(x, ν)}x∈Rn,ν∈Nnκ
condition
kmkl.u.,H s := sup
Then
2 )khzisF [m(x,·)ψ(2−k · )](z)kL2(Rnκ
z ) < ∞. (4.15)
2k(s− nκ
k>0, x∈Rn
1. If s > 3nκ
2 + (κ − 1)γ∞, the operator Tm extends to a bounded multilinear
operator from L1 × L∞ × · · · × L∞ × L∞ into L1(Rn), and
kTmkB(L1×(L∞)κ−1, L1) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
(4.16)
2. If s > 3nκ
2 + (κ−1)n
4
, the operator Tm extends to a bounded multilinear
operator from L2 × L∞ × · · · × L∞ × L∞ into L2(Rn), and
kTmkB(L2×(L∞)κ−1, L2) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)).
(4.17)
3. If s > 3nκ
2 + (n−1)(κ−1)
+γp, γp defined as in (2.11), the operator Tm extends
to a bounded multilinear operator from Lp × L∞ × · · · × L∞ × L∞ into
Lp(Rn), and
2
kTmkB(Lp×(L∞)κ−1, Lp) ≤ C(kmkl.u.,Hs + km(·, 0)kL∞(Rn)),
for all 2 < p ≤ ∞.
(4.18)
Let us note that the second assertion of Proposition 4.1 requires symbols with
, instead of its analogue condition in (4.3) where
12 . This difference is consequence of the different
regularity order s > 3nκ
we only need s > 3nκ
Fourier transforms that we use to classify the regularity of symbols.
2 + (κ−1)n
4 − 1
2 + (κ−1)n
4
In order to present our multilinear result, we will need the following interpo-
lation theorem which is valid for general measure spaces, but for simplicity we
record it on Rn.
Proposition 4.4 (Riesz-Thorin Interpolation). Let us assume that a linear opera-
tor T can be extended to a bounded operator T : Lpi(Rn) → Lqi(Rn) for i ∈ {0, 1}.
If 0 < θ < 1 and p, q are defined by
(4.19)
the operator T can be extended to a bounded operator T : Lp(Rn) → Lq(Rn) with
operator norm estimated by
1/p = (1 − θ)/p0 + θ/p1, 1/q = (1 − θ)/q0 + θ/q1,
kTkB(Lp,Lq) ≤ kTk1−θ
B(Lp0 ,Lq0 )kTkθ
B(Lp1 ,Lq1 ).
(4.20)
Although the Riesz-Thorin Interpolation Theorem is a well known result, we
will use strongly the control on the norms given in (4.20). In the following result
we will consider multilinear symbols satisfying Hormander conditions of order
3nκ
(κ − 1)n
,
+
3nκ
(n − 1)(κ − 1)
+
s > sn,κ,p := max{
2
2 + (κ−1)n
with γp defined as in (2.11). Let us note that 3nκ
+γp
can be not compared immediately because the sign of γp depends on the values
of p.
2 + (n−1)(κ−1)
2
and 3nκ
2
4
4
2
+ γp},
30
D. CARDONA AND M. RUZHANSKY
Proposition 4.5. Let 2 ≤ κ < ∞, κ ∈ N0. Let us consider a multilinear pseudo-
multiplier Tm defined on D(Rn)κ with symbol satisfying (4.2) or (4.15) for
s > sn,κ,p := max{
3nκ
2
+ (κ − 1)γ∞,
3nκ
2
+
(κ − 1)n
4
}, 1 ≤ p ≤ 2,
with γ∞, defined as in (2.11). Then the operator
(4.21)
extends to a bounded multilinear operator provided that 1 ≤ pj ≤ ∞, 1 ≤ p ≤ 2,
and 1
Tm : Lp1 × Lp2 × · · · × Lpκ−1 × Lpκ → Lp(Rn)
. If m satisfies the condition (4.2) or (4.15) for
p = 1
p1
+ · · · + 1
pκ
s > sn,κ,p := max{
3nκ
2
+
(κ − 1)n
4
,
3nκ
2
+
(n − 1)(κ − 1)
2
+ γp},
1
.
pκ
+ · · · + 1
with γp defined as in (2.11), then (4.21) holds true for all 2 ≤ p ≤ ∞ and
p = 1
p1
Proof. In order to prove the statement we will use real interpolation together with
induction on κ. Let us define the set
M := {κ ∈ N : κ ≥ 2, and from (4.2) or (4.15) we deduce (4.21) for s > sn,κ,p}.
(4.22)
First, we will prove that κ = 2 ∈ M. Then, let us assume that a bilinear operator
Tm satisfies (4.2) or (4.15). By Proposition 4.1 we have that Tm ∈ B(Lr×L∞, Lr)
for r = 1, 2, provided that
s > sn,κ,p := max{
2
Now, if we fix g0 ∈ L∞ and we consider the operator
2
+ (κ − 1)γ∞,
3nκ
3nκ
+
(κ − 1)n
4
}.
Tm,0 := Tm(·, g0),
then Tm,0 ∈ B(Lr) for r = 1, 2, and by real interpolation for all 1 ≤ r ≤ 2.
Moreover, if r is given by
1
r
=
1 − θ
1
+
θ
2
,
for some 0 < θ < 1, and we taking into account the norm estimates
kTm,0kB(L1) ≤ kTmkB(L1×L∞,L1)kg0kL∞, kTm,0kB(L2) ≤ kTmkB(L2×L∞,L2)kg0kL∞,
(4.23)
by application of (4.20) we have
B(L1×L∞,L1)kTmkθ
kTm,0kB(Lr) ≤ kTmk1−θ
(4.24)
Consequently we deduce the boundedness of Tm from Lp × L∞ into Lp, 1 ≤
p ≤ 2. Similarly we obtain the boundedness of Tm from L∞ × Lp into Lp. Now if
we repeat the argument for every entry of Tm, i.e., first fixing the first argument
and later fix the second argument, by interpolation we have the boundedness of
Tm from Lp1 × Lp2 into Lp, 1 ≤ p ≤ 2, with p1 and p2 satisfying the relation
B(L2×L∞,L2)kg0kL∞.
1
p1
=
θ
p
and
1
p2
=
1 − θ
p
(4.25)
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
31
p1
+ 1
p2
for some θ ∈ [0, 1]. Clearly, 1
= 1. Now, we will assume that every integer
number s less that κ belongs to M. So, let us assume now that we have a mul-
tilinear operator Tm on D(Rm)κ. By Proposition 4.1 the operator Tm extends to
a bounded multilinear operator from Lr × L∞ × · · · × L∞ × L∞ into Lr(Rn), for
r = 1, 2. If we consider g0 ∈ L∞ and similarly, as in the bilinear case, we define
Tm,0 := Tm(·,·,· · · ,·, g0),
fixing the last argument of Tm we obtain a multilinear operator on D(Rm)κ−1,
and by considering that κ − 1 ∈ M, we have that
(4.26)
extends to a bounded multilinear operator provided that 1 ≤ pj ≤ ∞, 1 ≤ p ≤ 2
and 1
Tm,0 : Lp1 × Lp2 × · · · × Lpκ−1 → Lp(Rn)
. As a consequence we obtain that
p = 1
p1
+ · · · + 1
pκ−1
Tm : Lp1 × Lp2 × · · · × Lpκ−1 × L∞ → Lp(Rn),
(4.27)
is bounded. Because
Tm : L∞ × L∞ × · · · × L∞ × Lp → Lp(Rn)
(4.28)
is bounded for all 1 ≤ p ≤ 2, we can fix every argument of Tm and apply the real
interpolation in order to provide the boundedness of Tm from L p1 ×L p2 ×· · ·×L pκ
Lp(Rn), where
θ
pi
=
1
pi
for 1 ≤ i ≤ κ − 1, and
1 − θ
p
=
1
pκ
,
for some 0 ≤ θ ≤ 1. Now, we finish the proof of induction by observing that
1
pi
=
θ
p
+
1 − θ
p
=
1
p
, and κ ∈ M,
(4.29)
κXi=1
so, we have proved that M = {κ ∈ N : κ ≥ 2}. Now, in a similar way we can use
statements 2 and 3 of Proposition 4.1 and the real interpolation (by repetition
of the arguments above) in order to provide the boundedness of Tm for the case
when 2 ≤ p < ∞.
(cid:3)
Remark 4.6. Taking into account that γ∞ = n−1
for n ≥ 2 and for n = 1,
2
γ∞ = 1/6, we can compute explicitly the regularity order sn,κ,p, 1 ≤ p ≤ 2,
defined in the previous proposition. Indeed, if n ≥ 2,
(κ − 1)n
sn,κ,p := max{
and
(κ − 1)(n − 1)
+ (κ − 1)γ∞,
} =
3nκ
3nκ
3nκ
+
+
2
2
2
2
4
,
3nκ
3nκ
(κ − 1)n
3nκ
(κ − 1)n
2
2
+
+(κ − 1)γ∞,
sn,κ,p := max{
4
for n = 1. Let us note that these regularity orders cannot be applied to κ = 1,
in order to recover those regularity orders given in the linear case, because our
Proposition 4.5 is a consequence of Proposition 4.1 whose proof uses strongly that
κ ≥ 2.
} =
+
=
+
4
4
2
3κ
2
κ − 1
,
32
D. CARDONA AND M. RUZHANSKY
References
1. Bagchi, S. Thangavelu, S. On Hermite pseudo-multipliers. J. Funct. Anal. 268 (1) (2015),
140 -- 170.
2. Bonami, A., Clerc, J.L. Somes de Cesaro et multiplicaturs des developments en harmonics
spheriques. Trans. Amer. Math. Soc. 183 (1973), 223 -- 263.
3. Blunck, S. A Hormander-type spectral multiplier theorem for operators without heat kernel.
Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 2 (2003), no. 3, 449 -- 459.
4. Bui, T. A., Duong, X. T. Besov and Triebel-Lizorkin spaces associated to Hermite operators.
J. Fourier Anal. Appl. 21 (2015), no. 2, 405 -- 448.
5. Chen, P., Ouhabaz, E. M., Sikora, A., Yan, L. Restriction estimates, sharp spectral multi-
pliers and endpoint estimates for Bochner-Riesz means. J. Anal. Math. 129 (2016), 219 -- 283.
6. Coifman, R., Meyer, Y. On commutators of singular integrals and bilinear singular integrals,
Trans. Amer. Math. Soc. 212 (1975), 315 -- 331.
7. Coifman, R., Meyer, Y. Au del`a des op´erateurs pseudo-diff´erentiels, Ast´erisque 57 (1978),
1 -- 185.
8. Coifman, R., Meyer, Y. Commutateurs dint´egrales singuli`eres et op´erateurs multilin´eaires,
Ann. Inst. Fourier (Grenoble) 28 (1978), 177 -- 202.
9. Duoandikoetxea, J. Fourier Analysis, Amer. Math. Soc. (2001)
10. Epperson, J. Hermite multipliers and pseudo-multipliers, Proc. Amer. Math. Soc. 124
(1996), no. 7, 2061 -- 2068.
11. Grafakos, L., Miyachi, A., Tomita, N. On multilinear Fourier multipliers of limited smooth-
ness. Canad. J. Math. 65 (2013), no. 2, 299 -- 330.
12. Grafakos, L., Nguyen, H. V. Multilinear Fourier multipliers with minimal Sobolev regularity,
I. Colloq. Math. 144 (2016), no. 1, 1 -- 30.
13. Grafakos, L., Miyachi, A., Nguyen, H. V., Tomita, N. Multilinear Fourier multipliers with
minimal Sobolev regularity, II. J. Math. Soc. Japan 69 (2017), no. 2, 529 -- 562.
14. Hormander, L. (1960) Estimates for translation invariant operators in Lp spaces. Acta
Math., 104, 93 -- 140.
15. Koch, H., Tataru, D. Lp-eigenfunction bounds for the Hermite operator. Duke Math. J. 128
(2005), 369 -- 392.
16. Mauceri, G. The Weyl transform and bounded operators on Lp(Rn). J. Funct. Anal. 39(3),
(1980), 408 -- 429.
17. Mihlin, S.G. On the multipliers of Fourier integrals. Dokl. Akad. Naulc SSSR (N. S.), 109
(1956), 701 -- 703 (Russian).
18. Petrushev, P., Xu, Y. Decomposition of spaces of distributions induced by Hermite expan-
sions. J. Fourier Anal. Appl. 14 (2008), no. 3, 372 -- 414.
19. Prugovecki, E. Quantum mechanics in Hilbert space. Second edition. Pure and Applied
Mathematics, 92. Academic Press, Inc, New York-London, 1981.
20. Ruzhansky M., Tokmagambetov N., Nonharmonic analysis of boundary value problems,
Int. Math. Res. Notices, (2016) 2016 (12), 3548 -- 3615.
21. Ruzhansky M., Tokmagambetov N., Nonharmonic analysis of boundary value problems
without WZ condition, Math. Model. Nat. Phenom., 12 (2017), 115 -- 140.
22. Ruzhansky, M., Turunen, V., Pseudo-differential Operators and Symmetries: Background
Analysis and Advanced Topics. Birkhauser-Verlag, Basel, (2010).
23. Strichartz, R. Multipliers for spherical harmonics expansions. Trans. Amer. Math. Soc. 167
(1972), 115-124.
24. Sikora, A., Yan, L., Yao, X. Sharp spectral multipliers for operators satisfying generalized
Gaussian estimates. J. Funct. Anal. 266 (2014), no. 1, 368 -- 409.
25. Simon, B. Distributions and their Hermite expansions. J. Math. Phys. 12 (1971), 140 -- 148.
26. Stempak, K. Multipliers for eigenfunction expansions of some Schrdinger operators, Proc.
Amer. Math. Soc. 93 (1985), 477 -- 482.
H ORMANDER CONDITION FOR PSEUDO-MULTIPLIERS
33
27. Stempak, K., Torre, J.L. On g-functions for Hermite function expansions, Acta Math. Hung.
109 , 99 -- 125.
28. Stempak, K., Torre, J.L. BMO results for operators associated to Hermite expansions,
Illinois J. Math. 49 (2005), 1111 -- 1132.
29. Thangavelu, S. Multipliers for Hermite expansions, Revist. Mat. Ibero. 3 (1987), 1 -- 24.
30. Thangavelu, S. Lectures on Hermite and Laguerre Expansions, Math. Notes, vol. 42, Prince-
ton University Press, Princeton, 1993.
31. Thangavelu, S. Hermite and special Hermite expansions revisited, Duke Math. J., 94(2)
(1998), 257 -- 278.
Duv´an Cardona:
Department of Mathematics
Pontificia Universidad Javeriana
Bogot´a
Colombia
E-mail address [email protected]; [email protected]
Michael Ruzhansky:
Department of Mathematics
Ghent University, Belgium
and
School of Mathematics
Queen Mary University of London
United Kingdom
E-mail address [email protected]
|
1604.01872 | 1 | 1604 | 2016-04-07T04:23:09 | Contractivity and complete contractivity for finite dimensional Banach Spaces | [
"math.FA"
] | Choose an arbitrary but fixed set of $n\times n$ matrices $A_1, \ldots, A_m$ and let $\Omega_\mathbf A\subset \mathbb C^m$ be the unit ball with respect to the norm $\|\cdot\|_{\mathbf A},$ where $\|(z_1,\ldots ,z_m)\|_{\mathbf A}=\|z_1A_1+ \cdots+z_mA_m\|_{\rm op}.$ It is known that if $m\geq 3$ and $\mathbb B$ is any ball in $\mathbb C^m$ with respect to some norm, say $\|\cdot\|_{\mathbb B},$ then there exists a contractive linear map $L:(\mathbb C^m,\|\cdot\|^*_{\mathbb B})\to \mathcal M_k$ which is not completely contractive. The characterization of those balls in $\mathbb C^2$ for which contractive linear maps are always completely contractive thus remains open. We answer this question for balls of the form $\Omega_\mathbf A$ in $\mathbb C^2.$ | math.FA | math |
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE
DIMENSIONAL BANACH SPACES
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
Abstract. Choose an arbitrary but fixed set of n × n matrices A1, . . . , Am and let ΩA ⊂ Cm be
the unit ball with respect to the norm k · kA, where k(z1, . . . , zm)kA = kz1A1 + · · · + zmAmkop.
It is known that if m ≥ 3 and B is any ball in Cm with respect to some norm, say k · kB, then
there exists a contractive linear map L : (Cm, k · k∗
B) → Mk which is not completely contractive.
The characterization of those balls in C2 for which contractive linear maps are always completely
contractive thus remains open. We answer this question for balls of the form ΩA in C2.
1. Introduction
In 1936 von Neumann (see [17, Corollary 1.2]) proved that if T is a bounded linear operator on
a separable complex Hilbert space H, then, for all complex polynomials p,
if and only if kTk ≤ 1. Or, equivalently, the homomorphism ρT induced by T on the polynomial
ring P [z] by the rule ρT (p) = p(T ) is contractive if and only if T is contractive.
kp(T )k ≤ kpk∞,D := sup{p(z) : z < 1}
The original proof of this inequality is intricate. A couple of decades later, Sz.-Nazy (see [17,
Theorem 4.3]) proved that a bounded linear operator T admits a unitary (power) dilation if and
only if there exists a unitary operator U on a Hilbert space K ⊇ H such that
PH p(U )H = p(T ),
for all polynomials p. The existence of such a dilation may be established by actually constructing
a unitary operator U dilating T. This construction is due to Schaffer (cf. [14]). Clearly, the von
Neumann inequality follows from the existence of a power dilation via the spectral theorem for
unitary operators.
Let P = ((pij)) be a k × k matrix valued polynomial in m variables. Let
kPk∞,Ω = sup{k ((pij(z))) kop : z ∈ Ω},
where Ω ⊆ Cm is a bounded open and connected set. Define P (T ) to be the operator ((pij(T ))) ,
1 ≤ i, j ≤ k. The homomorphism ρT is said to be completely contractive if
kP (T )k ≤ kPk∞,Ω, k = 1, 2, . . . .
A deep theorem due to Arveson (cf.
[1]) says that T has a normal boundary dilation if and
only if ρT is completely contractive. Clearly, if ρT is completely contractive, then it is contractive.
The dilation theorems due to Sz.-Nazy and Ando (cf.
[17]) give the non-trivial converse in the
case of the disc and the bi-disc algebras.
However, Parrott (cf. [15]) showed that there are three commuting contractions for which it is
impossible to find commuting unitaries dilating them. In view of Arveson's theorem this naturally
leads to the question of finding other algebras O(Ω) for which all contractive homomorphisms are
necessarily completely contractive. At the moment, this is known to be true of the disc, bi-disc
The work of G.Misra was supported, in part, through the J C Bose National Fellowship and UGC-CAS. The work of A. Pal was
supported, in part, through the UGC-NET and the IFCAM Research Fellowship. The results of this paper are taken from his PhD
thesis, after significant simplifications, submitted to the Indian Institute of Science in 2014.
1
2
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
(cf. [17]), symmetrized bi-disc (cf. [3]) and the annulus algebras (cf. [2]). Counter examples are
[8]) and any ball in Cm, m ≥ 3, as we will explain
known for domains of connectivity ≥ 2 (cf.
below.
Neither Ando's proof of the existence of a unitary dilation for a pair of commuting contractions,
nor the counter example to such an existence theorem due to Parrott involved the notion of
complete contractivity directly.
In the papers [10, 11, 12], it was shown that the examples of
Parrott are not even 2 -- contractive. In these papers, for any bounded, connected and open set
Ω ⊂ Cm, the homomorphism ρV : O(Ω) → Mp+q, induced by an m-tuple of p × q matrices V =
(V1, . . . , Vm), modeled after the examples of Parrott, was introduced. This was further studied,
in depth, by V. Paulsen [18], where he showed that the question of "contractive vs completely
contractive" for Parrott like homomorphisms ρV is equivalent to the question of "contractive vs
completely contractive" for the linear maps LV from some finite dimensional Banach space X to
Mn(C). The existence of linear maps of the form LV which are contractive but not completely
contractive for m ≥ 5 were found by him. A refinement (see remark at the bottom of p. 76 in
[16]) includes the case m = 3, 4, leaving the question of what happens when m = 2 open. This is
Problem 1 on page 79 of [16] in the list of "Open Problems".
For the normed linear space (C2,k·kA), we show, except when the pair A1, A2 is simultaneously
diagonalizable, that there is a contractive linear map on (C2,k·kA) taking values in p×q matrices,
which is not completely contractive.
We point out that the results of Paulsen used deep ideas from geometry of finite dimensional
Banach spaces. In contrast, our results are elementary in nature, although the computations, at
times, are somewhat involved.
2. Preliminaries
The norm kzkA = kz1A1 + ··· + zmAmkop, z ∈ Cm, is obtained from the embedding of the
linear space Cm into the C∗ algebra of n × n matrices via the map PA(z) := z1A1 + ··· + zmAm.
Let ΩA ⊂ Cm be the unit ball with respect to the norm k · kA. Let O(ΩA) denote the algebra
of functions each of which is holomorphic on some open set containing the closed unit ball ¯ΩA.
Given p × q matrices V1, . . . , Vm and a function f ∈ O(ΩA), define
(2.1)
i=1 ∂if (w) Vi
(cid:17) for a fixed w ∈ ΩA.
ρV(f ) :=(cid:16) f (w)Ip Pm
0
f (w)Iq
Clearly, ρV : (O(ΩA),k · k∞) → (Mp+q(C),k · kop) defines an algebra homomorphism.
At the outset we point out the interesting and useful fact that ρV is contractive on O(ΩA) if
and only if it is contractive on the subset of functions which vanish at w. This is the content of
the following lemma. The proof is reproduced from [18, Lemma 5.1], a direct proof appears in
[10, Lemma 3.3].
Lemma 2.1. supkfk∞=1{kρV(f )kop : f ∈ O(ΩA)} ≤ 1 if and only if supkgk∞=1{kρV(g)kop : g ∈
O(ΩA), g(w) = 0} ≤ 1.
Proof. The implication in one direction is obvious. To prove the converse, assume that kρV(g)k ≤
1 for every g such that g(w) = 0 and kgk∞ = 1.
For f ∈ O(ΩA) with kfk∞ = 1 let φf (w) be the Mobius map of the disc which maps f (w) to 0.
We let g = φf (w) ◦ f . Then g(w) = 0,kgk∞ = 1 and, from our assumption, kρV(g)k ≤ 1. So
kρV(f )k = kρV(φ−1
f (w) ◦ gk
f (w)(cid:0)ρV(g)(cid:1)k since ρV is a homomorphism
= kφ−1
≤ 1.
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES
3
In the last step we use the von Neumann inequality since φ−1
disc to itself.
f (w) is a rational function from the
(cid:3)
Note: For the rest of this work, we restrict to the case where w = 0 in the definition (2.1) of ρV
above.
The following lemma provides a characterization of the unit ball Ω∗A with respect to the dual
norm k · k∗A in Cm, that is Ω∗A = (Cm,k · k∗A)1.
Lemma 2.2. The dual unit ball
Ω∗A =(cid:8)(cid:0)∂1f (0), ∂2f (0),··· , ∂mf (0)(cid:1) : f ∈ Hol(ΩA, D), f (0) = 0(cid:9).
Proof. Given z ∈ Cm such that kzkA = 1 and f ∈ Hol(ΩA, D), f (0) = 0, we define gz : D → ΩA
by
Then f ◦ gz : D → D with (f ◦ gz)(0) = 0. Applying the Schwarz Lemma to the function (f ◦ gz)
we get
gz(λ) = λz, λ ∈ D.
1 ≥ (f ◦ gz)′(0) = f′(gz(0)) · g′z(0) = f′(0) · g′z(0) = f′(0) · z.
In the above, f′(0) · z =Pm
Hence(cid:0)∂1f (0), ∂2f (0),··· , ∂mf (0)(cid:1) ∈ Ω∗A.
Conversely, given w ∈ Ω∗A, we define fw(z) = w · z so that ∂ifw(0) = wi.
i=1(∂if (0))zi, etc.
2.1. The Maps L(k)
V : From Lemma 2.1 above it follows that
(cid:3)
(2.2)
kρVk ≤ 1 if and only if
sup
kfk∞=1,f (0)=0 k
∂if (0) Vikop ≤ 1.
Considering Lemma 2.2 and the equivalence (2.2) above it is natural to consider the induced linear
map LV : (Cm,k · k∗A) → Mp,q(C) given by
mXi=1
It follows from (2.2) above that
LV(w) = w1V1 + ··· + wmVm.
kρVk ≤ 1 if and only if kLVk ≤ 1.
We will show now that the complete contractivity of ρV and LV are also related similarly.
For a holomorphic function F : ΩA → Mk with kFk = supz∈ΩA kF (z)k, we define
(2.3)
ρ(k)
V (F ) := (ρV(Fij))m
i,j=1 =(cid:16) F (0)⊗I Pm
0
i=1(∂iF (0))⊗Vi
F (0)⊗I
(cid:17) .
Using a method similar to that used for ρV it can be shown that
kρ(k)
V k ≤ 1 if and only if sup
(∂iF (0)) ⊗ Vik : F ∈ Hol(ΩA, (Mk)1), F (0) = 0} ≤ 1,
that is, (by repeating the argument used for ρV) we have
F {k
mXi=1
V k ≤ 1 if and only if kL(k)
kρ(k)
V k ≤ 1,
where
is the map
L(k)
V : (Cm ⊗ Mk,k · k∗A,k) → (Mk ⊗ Mp,q,k · kop)
L(k)
V (Θ1, Θ2,··· , Θm) = Θ1 ⊗ V1 + Θ2 ⊗ V2 + ··· + Θm ⊗ Vm for Θ1, Θ2,··· , Θm ∈ Mk.
4
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
2.2. The polynomial PA. A very useful construct for our analysis is the matrix valued polyno-
mial PA with PA(ΩA) ⊆ (Mn,k · kop)1 defined by
PA(z1, z2,··· , zm) = z1A1 + z2A2 + ··· + zmAm,
with the norm kPAk∞ = sup(z1,··· ,zm)∈ΩA kPA(z1,··· , zm)kop. Note that kPAk∞ = 1 by definition.
The typical procedure used to show the existence of a homomorphism which is contractive but
not completely contractive is to construct a contractive homomorphism ρV (by a suitable choice
of V) and to then show that its evaluation on PA, that is, ρ(n)
V (PA), has norm greater than 1.
2.3. Homomorphisms induced by m-vectors. We now consider the special situation when
the matrices V1,··· , Vm are vectors in Cm realized as row m-vectors. For w = (w1, . . . , wm) in
some bounded domain Ω ⊆ Cm, the commuting m-tuple of (m + 1)× (m + 1) matrices of the form
0 wiIm(cid:17), 1 ≤ i ≤ m, induce the homomorphism ρV via the usual functional calculus, that is,
(cid:16) wi Vi
ρV (f ) := f(cid:0)(cid:16) w1 V1
0 w1Im(cid:17), . . . ,(cid:16) wm Vm
0 wmIm(cid:17)(cid:1), f ∈ O(Ω),
see (2.1). The localization of a commuting m - tuple T of operators in the class B1(Ω), introduced
in ([5, 6]), is also a commuting m - tuple of (m + 1) × (m + 1) matrices, which is exactly of the
form described above. The vectors V1, . . . , Vm appearing in such localizations are given explicitly
in terms of the curvature of the holomorphic Hermitian vector bundle corresponding to T as
shown in [6]. The contractivity of the homomorphism ρV then results in curvature inequalities
(see [9, 11, 12, 13]).
Let Vi = (cid:0)vi1 vi2
contractivity and complete contractivity in this special case, where, as before, we assume that
Ω = ΩA and w = 0.
vim(cid:1) , i = 1,··· , m. The propositions below are useful to study
···
Proposition 2.3. The following are equivalent:
(i) ρV is contractive,
(ii) supPm
j=1 zj2≤1 kPm
j=1 zj Bjk2
op ≤ 1, where Bj =Pm
i=1 vijAi.
Proof. We have shown that the homomorphisms k ρV kO(ΩA)→Mm+1(C) is contractive if and only if
A)→(Cm,k · k2) is contractive (equivalently if kL∗Vk(Cm,k· k2)→(Cm,k· kA) is
the linear map kLVk(Cm,k · k∗
contractive).
...
...
... v1m
Hence the contractivity of L∗V is given by the condition that
vm1 ... vmm(cid:19) .
The matrix representation of L∗V is(cid:18) v11
j=1 zj2≤1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
zm(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)A
vm1 ... vmm(cid:19)(cid:18) z1
(cid:18) v11
From the definition of k · kA it follows that
... v1m
Pm
sup
...
...
...
...
...
≤ 1.
kL∗Vk(Cm,k · k2)→(Cm,k · kA) ≤ 1 if and only if
i=1 vijAi.
where Bj =Pm
In particular, if V1 =(cid:0)u 0(cid:1) and V2 =(cid:0)0 v(cid:1) , the condition (ii) above becomes
sup
j=1 zj2≤1kz1uA1 + z2vA2k2 ≤ 1,
P2
which is equivalent to the following two conditions:
sup
j=1 zj2≤1k
Pm
mXj=1
zj Bjk2
op ≤ 1
(cid:3)
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES
5
inf
V (PA)k ≤ 1,
···
Proposition 2.4. The following are equivalent:
1k2 or v2 ≤ 1
(i) u2 ≤ 1
2k2
kA∗
kA∗
(ii)
β∈C2,kβk=1n1 − u2kA∗1βk2 − v2kA∗2βk2 + uv2(cid:16)kA∗1βk2kA∗2βk2 − hA1A∗2β, βi 2(cid:17)o ≥ 0.
(i) kρ(n)
(ii) T he n × mn matrix(cid:0)B1 B2
Proof. Since PA(0) = 0, it follows from the definition (2.3) that kρ(n)
For Vi =(cid:0)vi1
Thus kρ(n)
In particular if V1 =(cid:0)u 0(cid:1) and V2 =(cid:0)0 v(cid:1) the condition (ii) above becomes
vim(cid:1) , we have
A1 ⊗ V1 + . . . + Am ⊗ Vm =(cid:0)B1 B2
··· Bm(cid:1) is contractive, where Bj =Pm
V (PA)k ≤ 1 if and only if k(cid:0)B1 B2
··· Bm(cid:1)k ≤ 1.
kA1 ⊗ V1 + . . . + Am ⊗ Vmk ≤ 1.
··· Bm(cid:1)
V (PA)k ≤ 1 if and only if
i=1 vijAi.
β∈C2,kβk=1n1 − u2kA∗1βk2 − v2kA∗2βk2o ≥ 0.
inf
(cid:3)
Note: For most of this paper we will restrict to the two dimensional case. That is, we consider
C2 with the norm defined by a matrix pair (A1, A2). In fact, for the most part, we even restrict
to the situation where A1, A2 are 2 × 2 matrices. This is adequate for our primary purpose of
constructing homomorphisms of O(ΩA) which are contractive but not completely contractive.
Many of the results can be adapted to higher dimensional situation.
3. Defining Function and Test Functions
Recall the matrix valued polynomial PA : ΩA → (M2,k · kop)1 defined earlier by
PA(z1, z2) = z1A1 + z2A2,
where (M2,k · kop)1 is the matrix unit ball with respect to the operator norm. For (z1, z2) in ΩA,
the norm
by definition of the polynomial PA.
kPAk∞ := sup
(z1,z2)∈ΩA kPA(z1, z2)kop = 1
Let B2 be the unit ball in C2. For (α, β) ∈ B2 × B2, define p(α,β)
A
map
: ΩA → D to be the linear
p(α,β)
A (z1, z2) = hPA(z1, z2)α, βi = z1hA1α, βi + z2hA2α, βi.
A k∞, for any pair of vectors (α, β) in B2 × B2, is at most 1 by definition. Let
The sup norm kp(α,β)
PA denote the collection of linear functions {p(α,β)
A
: (α, β) ∈ B2 × B2}.
The map PA which we call the 'Defining Function' of the domain and the collection of functions
PA which we call a family of 'Test Functions' encode a significant amount of information relevant
to our purpose about the homomorphism ρV. For instance ρV is contractive if its restriction to
PA is contractive. Also the lack of complete contractivity can often be shown by evaluating ρ(2)
on PA. Some of the details are outlined in the lemma below.
V
6
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
Lemma 3.1. In the notation fixed in the preceding discussion, we have
(i)
sup
kαk=kβk=1 kρV(p(α,β)
A )k ≤ kρ(2)
V (PA)k,
(ii) ρV is contractive if and only if
sup
kαk=kβk=1 kρV(p(α,β)
A )k ≤ 1.
Proof of (i). Since
by definition, it follows that
ρV(p(α,β)
A ) =(cid:18) 0 (∂1p(α,β)
A )k = k(∂1p(α,β)
0
kρV(p(α,β)
A (0)) V1 + (∂2p(α,β)
A (0)) V2
0
(cid:19)
A (0)) V1 + (∂2p(α,β)
A (0)) V2kop
= khA1α, βi V1 + hA2α, βi V2kop
=
sup
kuk=kvk=1 hA1α, βihV1u, vi + hA2α, βihV2u, vi.
Hence
(3.1)
sup
kαk=kβk=1 kρV(p(α,β)
A )k =
=
sup
kαk=kβk=1
sup
kαk=kβk=1
=
sup
kαk=kβk=1
≤ kρ(2)
V (PA)k.
sup
sup
kuk=kvk=1 hA1α, βihV1u, vi + hA2α, βihV2u, vi
kuk=kvk=1 h(A1 ⊗ V1 + A2 ⊗ V2)α ⊗ u, β ⊗ vi
kuk=kvk=1 hρ(2)
V (PA)α ⊗ u, β ⊗ vi
sup
Proof of (ii). As indicated earlier the contractivity of ρV is equivalent to the contractivity of
given by the formula
LV : (C2,k · k∗A) → (Mp,q,k · kop)
LV(ω1, ω2) = ω1V1 + ω2V2.
So we identify the conditions for the contractivity of LV :
A≤1 kω1V1 + ω2V2kop
k(ω1 ,ω2)k∗
sup
kLVk =
=
sup
k(ω1 ,ω2)k∗
A≤1
sup
kuk=kvk=1 ω1hV1u, vi + ω2hV2u, vi.
Hence, since (ω1, ω2) lies in the dual of ΩA,
kLVk ≤ 1 ⇐⇒ (hV1u, vi,hV2u, vi) ∈ ΩA ∀u, v such that kuk = kvk = 1
⇐⇒ sup
⇐⇒ sup
kuk=kvk=1 khV1u, viA1 + hV2u, viA2kop ≤ 1
kαk=kβk=1
kuk=kvk=1 hA1α, βihV1u, vi + hA2α, βihV2u, vi ≤ 1
sup
⇐⇒ sup
kαk=kβk=1 kρV(p(α,β)
A )k ≤ 1 from (3.1) above.
(cid:3)
(cid:3)
As mentioned earlier, by choosing a pair (V1, V2) such that the inequality in (i) above is strict,
we can often construct a contractive homomorphism which is not completely contractive. We
illustrate below choices of (V1, V2) for the Euclidean ball for which the inequality is strict.
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES
7
Example 3.2. (Euclidean Ball) Choosing A = (( 1 0
clidean ball B2 in C2. Choose V1 = (v11 v12), V2 = (v21 v22). We will prove that
0 0 ) , ( 0 1
0 0 )) , we see that ΩA defines the Eu-
if V1 and V2 are linearly independent.
sup
kαk=kβk=1 kρV(p(α,β)
A )k < kρ(2)
V (PA)kop,
V (PA)kop >
1. This example of a contractive homomorphism of the ball algebra which is not completely
contractive was found in [10, 11].
In fact we can choose (V1, V2) such that supkαk=kβk=1 kρV(p(α,β)
Theorem 3.3. For ΩA = B2, let V1 =(cid:0)v11 v12(cid:1) , V2 =(cid:0)v21 v22(cid:1). Then
A )k ≤ 1 and kρ(2)
(i)
sup
kαk=kβk=1 kρV(p(α,β)
V (PA)k2
A )k2 = k ( v11 v12
op
v21 v22 ) k2
HS (HS represents the Hilbert − Schmidt norm)
(ii) kρ(2)
op = k ( v11 v12
Consequently, supkαk=kβk=1 kρV(p(α,β)
v21 v22 )k2
A )k < kρ(2)
Proof. By the definition of ρV we have
V (PA)kop if V1 and V2 are linearly independent.
sup
kαk=kβk=1 kρV(p(α,β)
A )k2 =
=
=
sup
sup
sup
kαk=kβk=kuk=kvk=1 hA1α, βihV1u, vi + hA2α, βihV2u, vi2
kαk=kβk=kuk=1 α1(v11u1 + v12u2) + α2(v21u1 + v22u2)2β12
kαk=kuk=1 α1(v11u1 + v12u2) + α2(v21u1 + v22u2)2
kuk=1 v11u1 + v12u22 + v21u1 + v22u22
=(cid:13)(cid:13) ( v11 v12
v21 v22 )(cid:13)(cid:13)2
= sup
op.
On the other hand, we have
kρ(2)
V (PA)k2
op = kV1k2 + kV2k2 = k ( v11 v12
v21 v22 )k2
HS.
If V1 and V2 are linearly independent
and we have
(cid:13)(cid:13) ( v11 v12
v21 v22 )(cid:13)(cid:13)2
kαk=kβk=1 kρV(p(α,β)
sup
op < k ( v11 v12
v21 v22 ) k2
HS
A )k < kρ(2)
V (PA)kop.
(cid:3)
Now choose V1 =(cid:0)1 0(cid:1) and V2 =(cid:0)0 1(cid:1). From Lemma 3.1 and Theorem 3.3 it follows that
ρV is contractive but kρ(2)
V (PA)k = √2.
4. Unitary Equivalence and Linear Equivalence
If U and W are 2 × 2 unitary matrices and eA = (U A1W, U A2W ), then
k(z1, z2)kA = kz1A1 + z2A2kop = kz1(U A1W ) + z2(U A2W )kop = k(z1, z2)keA.
There are, therefore, various choices of the matrix pair (A1, A2) related as above which give rise
to the same norm. We use this freedom to ensure that A1 is diagonal. Consider the invertible
linear transformation (z1, z2) 7→ (z1, z2) on C2 defined as follows:
8
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
For z = (z1, z2) in C2, let
where p, q, r, s ∈ C. Then
where eA is related to A as follows:
z1 = pz1 + q z2
z2 = rz1 + sz2,
k(z1, z2)kA = k(z1, z2)keA,
eA1 = pA1 + rA2
eA2 = qA1 + sA2.
kT zkA = kzkA(T⊗I).
More concisely, if T is the linear transformation above on C2, then
In particular T maps ΩeA onto ΩA.
Lemma 4.1. For k = 1, 2, . . . , the contractivity of the linear maps L(k)
(C2⊗ Mk,k·k∗
eA,k
and conversely, where eA = A(T ⊗ I) and eV = (T ⊗ I)V.
)→(Mk⊗Mp,q,k·kop) ≤ 1 ⇐⇒ kL(k)
) determine the contractivity of the linear maps L(k)
eV
Proof. For k = 1, 2, . . . , we have to show that
V k(C2⊗Mk,k·k∗
eV k(C2⊗Mk,k·k∗
kL(k)
eA,k
V defined on
defined on (C2⊗ Mk,k·k∗A,k)
A,k)→(Mk⊗Mp,q,k·kop) ≤ 1.
We prove this result for the case k = 1, that is, for the map LV. The proof for the general case
is similar.
Consider the bijection between the spaces {f ∈ Hol(ΩA, D), f (0) = 0} and
{ef ∈ Hol(ΩeA, D),ef (0) = 0} defined as follows:
f 7→ ef = f ◦ T, ef 7→ f = ef ◦ T −1
Using this bijection
k LV k(C2,k·k∗
eA
)→(Mp,q ,k·kop) ≤ 1 ⇐⇒ sup
⇐⇒ sup
⇐⇒ sup
⇐⇒ sup
⇐⇒ k L(T⊗I)V k(C2,k·k∗
ef {kDef (0) · Vkop : ef ∈ Hol(ΩeA, D),ef (0) = 0} ≤ 1
f {kD(f ◦ T )(0) · Vkop : f ∈ Hol(ΩA, D), f (0) = 0} ≤ 1
f {kDf (0) T · Vkop : f ∈ Hol(ΩA, D), f (0) = 0} ≤ 1
f {kDf (0) · (T ⊗ I)Vkop : f ∈ Hol(ΩA, D), f (0) = 0} ≤ 1
A)→(Mp,q ,k·kop) ≤ 1
i=1 XiYi.
It follows that, in our study of the existence of contractive homomorphisms which are not
In the above, Df is a row vector, T is a 2 × 2 matrix and by an expression of the form X · Y we
meanP2
completely contractive, two sets of matrices A = (A1, A2) and eA = (eA1, eA2) which are related
through linear combinations as above yield the same result. We can, therefore, restrict our
attention to a subcollection of matrices.
(cid:3)
Since A1 has already been chosen to be diagonal, we consider transformations as above with
r = 0 to preserve the diagonal structure of A1. By choosing the parameters p, q, s suitably we can
ensure that one diagonal entry of A1 is 1 and the diagonal entries of A2 are 1 and 0. By further
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES
9
conjugating with a diagonal unitary and a permutation matrix it follows that we need to consider
only the following three families of matrices:
Table 1. Cases modulo unitary and linear equivalence
A1
A2
(cid:18) 1 0
0 d (cid:19) d ∈ C (cid:18) 1 b
(cid:18) d 0
0 1 (cid:19) d ∈ C (cid:18) 1 b
(cid:18) 1 0
0 d (cid:19) d ∈ C (cid:18) 0 b
c 0 (cid:19) c ∈ C, b ∈ R+
c 0 (cid:19) c ∈ C, b ∈ R+
c 0 (cid:19) c ∈ C, b ∈ R+
In the above, R+ represents the set of non-negative real numbers.
4.1. Simultaneously Diagonalizable Case. For the study of contractivity and complete con-
tractivity in this situation we consider two possibilities. The first when A1 and A2 are simultane-
ously diagonalizable and the second when they are not. The simultaneously diagonalizable case
reduces to the case of the bi-disc where we know that any contractive homomorphism is completely
contractive. In all the other cases (when A1 and A2 are not simultaneously diagonalizable) we
show that there exists a contractive homomorphism which is not completely contractive.
Consider first the case when A1 and A2 are simultaneously diagonalizable. Based on the
discussion of linear equivalence above we need to study only the following possibilities:
Table 2. Simultaneously diagonalizable cases
A1
A2
0 d (cid:19) d ∈ C (cid:18) 1 0
(cid:18) 1 0
0 0 (cid:19)
(cid:18) d 0
0 1 (cid:19) d ∈ C (cid:18) 1 0
0 0 (cid:19)
Applying linear transformations as before, both cases can be reduced to A = (( 1 0
0 1 ))
which represents the bi-disc. As mentioned earlier, it is known that any contractive homomorphism
is completely contractive in this case. We now study the situation when A1 and A2 are not
simultaneously diagonalizable.
0 0 ) , ( 0 0
10
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
5. Contractivity, Complete Contractivity and Operator Space Structures
We recall some notions about operator spaces which are relevant to our purpose.
Definition 5.1. (cf. [17, Chapter 13, 14]) An abstract operator space is a linear space X together
with a family of norms k · kk defined on Mk(X), k = 1, 2, 3, . . . , where k · k1 is simply a norm on
the linear space X. These norms are required to satisfy the following compatibility conditions:
(1) kT ⊕ Skp+q = max{kTkp,kSkq} and
(2) kASBkp ≤ kAkopkSkqkBkop
for all S ∈ Mq(X), T ∈ Mp(X) and A ∈ Mp,q(C), B ∈ Mq,p(C).
Two such operator spaces (X,k · kk) and (Y,k · kk) are said to be completely isometric if there
is a linear bijection T : X → Y such that T ⊗ Ik : (Mk(X),k·kk) → (Mk(Y ),k·kk) is an isometry
for every k ∈ N. Here we have identified Mk(X) with X ⊗ Mk in the usual manner. We note
that a normed linear space (X,k · k) admits an operator space structure if and only if there is an
isometric embedding of it into the algebra of operators B(H) on some Hilbert space H. This is
the well-known theorem of Ruan (cf. [16]).
We recall here the notions of MIN and MAX operator spaces and a measure of their distance,
α(X), following [17, Chapter 14].
Definition 5.2. The MIN operator structure M IN (X) on a (finite dimensional) normed linear
space X is obtained by isometrically embedding X in the C∗ algebra C(cid:0)(X∗)1(cid:1), of continuous
functions on the unit ball (X∗)1 of the dual space. Thus for ((vij)) in Mk(X), we set
where the norm of a scalar matrix ((f (vij))) in Mk is the operator norm.
k((vij ))kM IN = k((cvij ))k = sup{k((f (vij)))k : f ∈ (X∗)1},
For an arbitrary k × k matrix over X, we simply write k((vij ))kM IN (X) to denote its norm in
Mk(X). This is the minimal way in which we represent the normed space as an operator space.
There is also a 'maximal' representation which is denoted M AX(X).
Definition 5.3. The operator space M AX(X) is defined by setting
k((vij ))kM AX = sup{k((T (vij )))k : T : X → B(H)},
and the supremum is taken over all isometries T and all Hilbert spaces H.
Every operator space structure on a normed linear space X 'lies between' M IN (X) and
M AX(X). The extent to which the two operator space structures M IN (X) and M AX(X)
differ is characterized by the constant α(X) introduced by Paulsen (cf.[17, Chapter 14]), which
we recall below.
Definition 5.4. The constant α(X) is defined as
α(X) = sup{k((vij ))kM AX : k((vij ))kM IN ≤ 1, ((vij)) ∈ Mk(X), k ∈ N}.
Thus α(X) = 1 if and only if the identity map is a complete isometry from M IN (X) to
M AX(X). Equivalently, we conclude that there exists a unique operator space structure on X
whenever α(X) is 1. Therefore, those normed linear spaces for which α(X) = 1 are rather
special. Unfortunately, there aren't too many of them! The familiar examples are (C2,k · k∞),
and consequently C2 with the ℓ1 norm.
It is pointed out in [16, pp. 76]) that α(X) > 1 for
dim(X) ≥ 3, refining an earlier result of Paulsen that α(X) > 1 whenever dim(X) ≥ 5. This
leaves the question open for normed linear spaces whose dimension is 2.
Returning to the space (C2,k · kA) with k(z1, z2)kA = kz1A1 + z2A2kop, we show below that
α(ΩA) > 1 in a large number of cases. From [18, Theorem 4.2], it therefore follows that, in all these
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES 11
cases, there must exist a contractive homomorphism of O(ΩA) into the algebra B(H) which is
not completely contractive. In the remaining cases, the existence of a contractive homomorphism
which is not completely contractive is established by a careful study of certain extremal problems.
The norm k(z1, z2)kA = kz1A1 + z2A2kop defines a natural isometric embedding into M2(C)
given by (z1, z2) 7→ z1A1 + z2A2. However, note that
k(z1, z2)kA = kz1A1 + z2A2kop = kz1At
1 + z2At
2kop = k(z1, z2)kAt .
We, therefore, get another isometric embedding into M2(C) given by (z1, z2) 7→ z1At
In a variety of cases the operator spaces determined by these two embeddings are distinct and
the parameter α > 1 in these cases. Therefore, the existence of contractive homomorphisms which
are not completely contractive is established in these cases. We present the details below.
1 + z2At
2.
Recall the map PA defined earlier by PA(z1, z2) = z1A1 + z2A2. Let P (2)
A = PA ⊗ I2. For the
three families of matrices A = (A1, A2) characterized in Table 1 we show that A and At define
distinct operator space structures unless d = 1 or b = c.
Theorem 5.5. Let Z1 = ( 1 0
kP (2)
Proof. We illustrate the proof for the case A1 =(cid:0) 1 0
0 0 ). If d 6= 1 and b 6= c then kP (2)
A (Z1, Z2)kop 6=
0 0 ) and Z2 = ( 0 1
At (Z1, Z2)kop.
similarly.
0 d(cid:1) , A2 =(cid:0) 1 b
c 0(cid:1). The other cases can be proved
1 (cid:17)(cid:13)(cid:13)(cid:13)op
¯cZ ∗
2
¯dZ ∗
¯c(Z1+Z2)Z ∗
c2Z2Z ∗
2 +b ¯dZ2Z ∗
2 +d2Z1Z ∗
.
2
op. Using the form of (Z1, Z2) this is equivalent to
For this case
kP (2)
A (Z1, Z2)k2
(5.1)
Similarly we have
(5.2)
At (Z1, Z2)k2
kP (2)
A (Z1, Z2)k2
op = kP (2)
Assume kP (2)
bZ2
cZ2
bZ ∗
2
cZ2(Z1+Z2)∗+bdZ1Z ∗
2
op =(cid:13)(cid:13)(cid:13)(cid:16) (Z1+Z2)
dZ1(cid:17)(cid:16) (Z1+Z2)∗
=(cid:13)(cid:13)(cid:13)(cid:16) (Z1+Z2)(Z1+Z2)∗+b2Z2Z ∗
op =(cid:13)(cid:13)(cid:13)(cid:16) (Z1+Z2)(Z1+Z2)∗+c2Z2Z ∗
(cid:13)(cid:13)(cid:13)(cid:16) 2+b2
=(cid:13)(cid:13)(cid:13)(cid:16) 2+c2
c2+d2(cid:17)(cid:13)(cid:13)(cid:13)op
At (Z1, Z2)k2
bZ2(Z1+Z2)∗+¯cdZ1Z ∗
2
¯c
c
b
2
1
1 (cid:17)(cid:13)(cid:13)(cid:13)op
1 (cid:17)(cid:13)(cid:13)(cid:13)op
.
b(Z1+Z2)Z ∗
2 +c ¯dZ2Z ∗
1
b2Z2Z ∗
2 +d2Z1Z ∗
b
b2+d2(cid:17)(cid:13)(cid:13)(cid:13)op
i.e. (b2 − c2)(1 − d2) = 0 (note that the matrices on the left and right have the same trace),
from which the result follows.
(cid:3)
Since α(ΩA) = 1 if and only if the two operator spaces MIN(ΩA) and MAX(ΩA) are completely
isometric, it follows from the Theorem we have just proved that if d 6= 1 and b 6= c, then
α(X) > 1. Consequently, there exists a contractive homomorphism of O(ΩA) into B(H), which is
not completely contractive.
Example 5.6. (Euclidean Ball) The Euclidean ball B2 is characterized by A1 = ( 1 0
0 0 ) , A2 = ( 0 1
0 0 ).
So, in Theorem 5.5, we have d 6= 1 and b 6= c. Hence A and At give rise to distinct operator space
structures and, consequently, there exists a contractive homomorphism which is not completely
contractive.
6. Cases not Amenable to the Operator Space Method
Theorem 5.5 shows that there is a contractive homomorphism which is not completely con-
tractive for all the choices of (A1, A2) listed in Table 1 except when d = 1 or b = c. We are,
therefore, left with the following families of (A1, A2) to be considered:
12
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
Table 3. Cases not covered by the operator space approach
A1
A2
0
0
0
(i)
(ii)
0 eiθ (cid:19) θ ∈ R (cid:18) 1 b
(cid:18) 1
0 eiθ (cid:19) θ ∈ R (cid:18) 0 b
(cid:18) 1
(iii) (cid:18) eiθ 0
1 (cid:19) θ ∈ R (cid:18) 1 b
(cid:18) 1 0
0 d (cid:19) d ∈ C
(cid:18) 1 0
0 d (cid:19) d ∈ C
(cid:18) d 0
0 1 (cid:19) d ∈ C
c 0 (cid:19) c ∈ C, b ∈ R+
c 0 (cid:19) c ∈ C, b ∈ R+
c 0 (cid:19) c ∈ C, b ∈ R+
(cid:18) 1 c
0 (cid:19) c ∈ C
(cid:18) 0 c
0 (cid:19) c ∈ C
(cid:18) 1 c
0 (cid:19) c ∈ C
(iv)
(vi)
(v)
c
c
c
These six families are not disjoint and have been classified as such on the basis of the method
of proof used.
6.1. Dual norm method. We first consider a special case of type (ii) in Table 3 with A1 =
( 1 0
0 1 ) , A2 = ( 0 1
0 0 ). Although this case is covered by the more general method to be outlined later
we present an alternate, interesting procedure for this example since it is possible to explicitly
calculate the dual norm k · k∗A in this case. Equipped with the information about the dual norm
we can directly construct a pair V = (V1, V2) such that kLVk ≤ 1 and kL(2)
V (PA)k > 1.
Note that in this case
and the unit ball
k(z1, z2)kA = z2 +pz22 + 4z12
ΩA = {(z1, z2) : z12 + z2 < 1}.
2
Lemma 6.1. Let A1 = ( 1 0
0 1 ) , A2 = ( 0 1
0 0 ). If (ω1, ω2) ∈ (C2,k · k∗A) then the dual norm
if ω2 ≥ ω12 ;
if ω2 ≤ ω12 .
Proof. Let fω1,ω2 be the linear functional on (C2,k · kA) defined by
k(ω1, ω2)k∗A =( ω12+4ω22
ω1
4ω2
fω1,ω2(z1, z2) = ω1z1 + ω2z2.
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES 13
Then
k(ω1, ω2)k∗A = sup
sup
=
(z1,z2)∈ΩA fω1,ω2(z1, z2)
z2≤1−z12 ω1z1 + ω2z2
z2≤1−z12
z1≤1(cid:0)ω1z1 + ω2(1 − z12)(cid:1).
(ω1z1 + ω2z2)
sup
=
= sup
If ω2 ≥ ω12
the expression on the right attains its maximum at z1 = ω1
2ω2 ≤ 1 and the
maximum value is ω12+4ω22
.
4ω2
If ω2 ≤ ω12
the expression on the right is monotonic in z1 and the maximum is attained at
z1 = 1. The maximum value in this case is ω1.
(cid:3)
Theorem 6.2. Let A1 = ( 1 0
0 1 ) , A2 = ( 0 1
0(cid:17) , V2 =(cid:0)0 1(cid:1) . Then
0 0 ) and V1 =(cid:16) 1√2
V (PA)k =r 3
(i) kLVk(C2,k·k∗
(ii) kL(2)
2
.
A)→(C2,k·k2) = 1
Consequently ρV, for this choice of V = (V1, V2), is contractive on O(ΩA) but not completely
contractive.
Proof of (i).
kLVk2
(C2,k·k∗
A)→(C2,k·k2) =
=
We now consider two cases:
sup
k(ω1,ω2)k∗
sup
k(ω1,ω2)k∗
2
A=1 kω1V1 + ω2V2k2
+ ω22(cid:17).
A=1(cid:16)ω12
2
Case (a): ω2 ≥ ω12 and 1 = k(ω1, ω2)k∗A = ω12+4ω22
These two constraints together can be seen to be equivalent to the constraints 1
from Lemma 6.1.
4ω2
2 ≤ ω2 ≤ 1
and ω12 = 4ω2(1 − ω2).
Hence the supremum above for this range of (ω1, ω2) is given by
sup
2≤ω2≤1ω2(cid:0)2 − ω2(cid:1) = 1.
1
Case (b): ω2 ≤ ω12 and 1 = k(ω1, ω2)k∗A = ω1
The supremum for this range of (ω1, ω2) is given by
from Lemma 6.1.
sup
ω2≤ 1
2(cid:0) 1
2
+ ω22(cid:1) =
3
4
.
Taking the larger of the supremums in Case (a) and Case (b) we get that kLVk = 1.
(cid:3)
14
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
Proof of (ii).
kL(2)
V (PA)k2 = kA1 ⊗ V1 + A2 ⊗ V2k2
=(cid:13)(cid:13)(cid:13)(cid:16) 1√2
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:16) 1√2
=(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 3
2
0
=
.
3
2
2
A∗1
1√2
A1 A2(cid:17)(cid:13)(cid:13)(cid:13)
A1 A2(cid:17)
2(cid:19)(cid:13)(cid:13)(cid:13)(cid:13) using the form of A1, A2
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
A∗2
0
1
(cid:3)
6.2. General cases not amenable to the operator space method. The various families
of (A1, A2) listed in Table 3 require a case by case analysis to show that there is a contractive
homomorpism which is not completely contractive. We first present a general outline of the
method used.
We choose the pair V = (V1, V2) to be of the form V1 = (cid:0)u 0(cid:1) , V2 = (cid:0)0 v(cid:1) , u, v ∈ R+.
LV : (C2,k · k∗A) → (C2,k · k2) then becomes the linear map (z1, z2) 7→ (z1u, z2v).
while kL(2)
We show, in each case, that by a suitable choice of u and v we can ensure that LV is contractive
V (PA)k > 1 although kPAk = 1 by definition.
We list the contractivity conditions (see Propositions 2.3 and 2.4 for details).
(a) LV is contractive if and only if the following two conditions are satisfied:
(i) u ≤ 1
kA∗
1k
(ii)
β∈C2,kβk=1n1 − u2kA∗1βk2 − v2kA∗2βk2 + u2v2(cid:16)kA∗1βk2kA∗2βk2 − hA1A∗2β, βi 2(cid:17)o ≥ 0.
or v ≤ 1
kA∗
2k
and
inf
(6.1)
(b) kL(2)
V (PA)k ≤ 1 if and only if
(6.2)
β∈C2,kβk=1n1 − u2kA∗1βk2 − v2kA∗2βk2o ≥ 0.
inf
Note that the term in parenthesis in (6.1) is non-negative by the Schwarz inequality and that
the expression (6.2) is the same as the first three terms in (6.1).
We show that, in each case, we can choose (u, v) such that the infimum in (6.1) is exactly 0.
Also that this infimum is attained at β = β0 such that the term in parenthesis in (6.1) is positive
(that is, the Schwarz inequality referred to above is a strict inequality at β0).
It then follows
that the expression in braces in (6.2) is negative when β = β0 and, consequently, the infimum
in (6.2) is negative. Taken together it follows that LV (and consequently ρV) is contractive but
V (PA)k > 1 and, as a result, ρ(2)
Let η(i), i = 1, 2, be the vectors such that A∗1η(i) and A∗2η(i) are linearly dependent. That is,
kL(2)
the term in parenthesis in (6.1) vanishes when β = η(i).
V is not contractive.
We now provide the details of the argument which proceeds in two steps.
Step 1: Show that there are certain ranges of the parameters (u, v) such that the infimum in
(6.1) is not attained at η(1) or η(2) for those values of (u, v).
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES 15
Let
gu,v(β) = 1 − u2kA∗1βk2 − v2kA∗2βk2 + u2v2(cid:16)kA∗1βk2kA∗2βk2 − hA1A∗2β, βi 2(cid:17).
We need to show that there exists β such that
when (u, v) take values in a range of interest. That is, we need to find β such that
gu,v(β) < gu,v(η(i)), i = 1, 2,
(6.3)
Here
(6.4)
gu,v(η(i)) − gu,v(β) = ai(β)u2 + bi(β)v2 − c(β)u2v2 > 0.
ai(β) = kA∗1βk2 − kA∗1η(i)k2
bi(β) = kA∗2βk2 − kA∗2η(i)k2
c(β) = kA∗1βk2kA∗2βk2 − hA1A∗2β, βi 2 ≥ 0.
Consider the functions
fi(u, v, β) = ai(β)u2 + bi(β)v2 − c(β)u2v2 with c(β) ≥ 0, i = 1, 2.
The following result is evident from the nature of the functions fi(u, v, β).
Lemma 6.3. (i) Assume ai(β) > 0 for some fixed β and i = 1, 2. Then, given any u0 > 0, there
exists v0 > 0 (depending on u0) such that fi(u, v, β) > 0 in the region u < u0, v < v0
u, that is,
u0
inside the triangle with vertices (0, 0), (u0, 0) and (u0, v0).
(ii) Assume bi(β) > 0 for some fixed β and i = 1, 2. Then, given any v0 > 0, there exists
v , that is,
u0 > 0 (depending on v0) such that fi(u, v, β) > 0 in the region v < v0, u < u0
v0
inside the triangle with vertices (0, 0), (0, v0) and (u0, v0).
(iii) If fi(u0, v0, β) > 0 then fi(tu0, tv0, β) > 0 for 0 < t < 1.
We will show below that, in each of the six cases in Table 3, it is possible to ensure the positivity
of ai(β), i = 1, 2 or bi(β), i = 1, 2 for some choice of β. Consequently, it will follow that the
inequality (6.2) will be true for that vector β with (u, v) in the region characterized in Lemma
6.3 above. Hence, for (u, v) in this range, the infimum is not attained at η(i), i = 1, 2.
Consider first the cases (i), (ii) and (iii).
A1
A2
(i)
(ii)
0
0
c 0 (cid:19) c ∈ C, b ∈ R+
c 0 (cid:19) c ∈ C, b ∈ R+
c 0 (cid:19) c ∈ C, b ∈ R+
0 eiθ (cid:19) θ ∈ R (cid:18) 1 b
(cid:18) 1
0 eiθ (cid:19) θ ∈ R (cid:18) 0 b
(cid:18) 1
1 (cid:19) θ ∈ R (cid:18) 1 b
0 e−iθ(cid:1) so that A1 becomes the identity matrix. In case (iii)
(iii) (cid:18) eiθ 0
on the left by the unitary matrix (cid:0) 1
0
0
We use the unitary equivalence described in Section 4. In cases (i) and (ii) multiply A1 and A2
16
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
multiply A1 and A2 on the left by the unitary matrix (cid:0) e−iθ 0
matrix.
1(cid:1) so that A1 becomes the identity
Now conjugate A1 and A2 by the unitary which makes A2 upper triangular so that cases (i),(ii)
0
and (iii) reduce to the situation
A1 = ( 1 0
0 1 ) and A2 = ( µ σ
0 ν ) with µ ≥ ν, σ 6= 0
In this case ai(β) = 0 for all β but it is possible to choose β such that bi(β) > 0.
η(i) satisfies the equation (A∗2 − λiA∗1)η(i) = 0. So in this case η(i) is a (unit) eigenvector of A∗2
with eigenvalue λi. Since the eigenvalues of A∗2 are ¯µ and ¯ν it follows that kA∗2η(i)k2 = µ2 or ν2.
Hence we can take β = ( 1
0 ) so that bi(β) ≥ σ2 > 0.
Now consider cases (iv) and (v).
A1
A2
0 d (cid:19) d ∈ C,d 6= 1 (cid:18) 1 c
(iv) (cid:18) 1 0
(cid:18) 1 0
0 d (cid:19) d ∈ C,d 6= 1 (cid:18) 0 c
0 (cid:19) c ∈ C
0 (cid:19) c ∈ C
(v)
c
c
In cases (iv) and (v) we have, in Equation (6.4),
ai(β) = β12 + d2β22 − η(i)
= (β12 − η(i)
= (1 − d2)(η(i)
1 2 − d2η(i)
2 2
1 2) + d2(β22 − η(i)
2 2)
2 2 − β22).
If η(i)
we can choose β such that β2 < η(i)
ai(β) > 0 for i = 1, 2.
2 = 0 or 1 then c = 0 and it reduces to the simultaneously diagonalizable case. If η(i)
2 6= 0, 1
2 ) if d < 1 (resp. d > 1) to ensure that
2 (resp. β2 > η(i)
The methods used in cases (iv) and (v) can be adapted to the last case (vi):
A1
A2
(vi) (cid:18) d 0
0 1 (cid:19) d ∈ C,d 6= 1 (cid:18) 1 c
0 (cid:19) c ∈ C
c
Step 2: Show that, in each case, there is a choice of (u, v) in the region characterized in Lemma
6.3 for which the infimum in (6.1) is, in fact, zero.
We choose β to ensure that ai( β) or bi( β) is positive as described in Step 1.
Note that gu,v( β) vanishes at the two points (u, v) = (
, 0), (u, v) = (0,
1
1( β)k
kA∗
1
2( β)k
kA∗
) and also
along a curve joining these two points.
We now consider two cases:
Case (i): ai( β) > 0
Choose (u0, v0) such that 0 < v0 < 1
kA∗
2k
using Lemma 6.3 and the above note about the vanishing of gu,v( β).
, fi(u0, v0, β) > 0 and gu0,v0( β) = 0. This is possible
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES 17
Let
Note that x2
0 ≥
1
1k2+λ2
kA∗
gu,λ0u(β) ≤ 0} where λ0 =
x0 = inf{u : inf
2k2 . Also, from Lemma 6.3, it is clear that fi(x0, λ0x0, β) > 0.
β
.
0kA∗
v0
u0
We now show that inf β gx0,λ0x0(β) = 0.
To prove this we first show that g(x0, λ0x0)(β) ≥ 0 for all β (with kβk2 = 1) as follows. Assume
there exists β = µ such that g(x0,λ0x0)(µ) < 0. Then there exists a neighborhood U of x0 such that
gu,λ0u(µ) < 0 for all u ∈ U. For any u ∈ U, inf β gu,λ0u(β) < 0, since gu,λ0u(µ) < 0 for all u ∈ U.
Since U is a neighborhood of x0 there exists a u ∈ U such that u < x0. By the previous assertion,
inf β gu,λ0u(β) ≤ 0 for this smaller value of u, which is a contradiction.
Since inf β gx0,λ0x0(β) ≤ 0 by the definition of x0 it follows that inf β gx0,λ0x0(β) = 0.
Case (ii): bi( β) > 0
The arguments in this case are similar to Case (i). This time choose (u0, v0) such that 0 <
, fi(u0, v0, β) > 0 and gu0,v0( β) = 0.
u0 < 1
kA∗
1k
Let
y0 = inf{v : inf
β
gλ0v,v(β) ≤ 0} where λ0 =
u0
v0
.
As in Case (i) we can see that y2
0 ≥
1
λ2
0kA∗
1k2+A∗
2k2 and (from Lemma 6.3) that fi(λ0y0, y0, β) > 0.
Using a procedure similar to that used in Case (i) it follows that inf β gλ0y0,y0(β) = 0.
We have therefore shown that for all the cases in Table 3 which were not covered by the
operator space approach it is possible to choose (u, v) such that the infimum in (6.1) is zero and
this infimum is attained at a vector β not equal to η(1) or η(2), so that the last term in parenthesis
in (6.1) is positive at β.
It follows that, in each of these cases, there exists a contractive homomorphism which is not
completely contractive.
7. An Interesting Operator Space Computation
In Section 5 the existence of contractive homomorphisms which are not completely contractive
was shown in many cases by studying different isometric embeddings of the space (C2,k · kA)
into (M2,k · kop) which led to distinct operator space structures. The two embeddings considered
there were (z1, z2) 7→ z1A1 + z2A2 and (z1, z2) 7→ z1At
2. In this section we show that we
can, for some choices of (A1, A2), construct large collections of isometric embeddings of the space
(C2,k · kA) into various matrix spaces. Although the embeddings are into very distinct matrix
spaces, we show that the operator space structures thus obtained are equivalent.
A result which is very useful in this context is the following proposition due to Douglas, Muhly
1 + z2At
and Pearcy (cf. [7, Prop. 2.2]).
into H1 such that
Proposition 7.1. For i = 1, 2, let Ti be a contraction on a Hilbert space Hi and let X be an
operator mapping H2 into H1. A necessary and sufficient condition that the operator on H1 ⊕ H2
defined by the matrix (cid:16) T1 X
The operator norm of the block matrix(cid:0) αIm B
form(cid:16) α1Im B
α2In(cid:17) , for arbitrary α1, α2 ∈ C.
0 T2(cid:17) be a contraction is that there exist a contraction C mapping H2
0 αIn(cid:1) , where B is an m × n matrix and α ∈ C, is
X =p1H1 − T1T ∗1 C p1H2 − T ∗2 T2.
not hard to compute (cf. [10, Lemma 2.1]). The result can be easily extended to a matrix of the
0
18
GADADHAR MISRA, AVIJIT PAL AND CHERIAN VARUGHESE
Lemma 7.2. If B is an m × n matrix and α1, α2 ∈ C then
Proof. Consider the following two sets
and
0
0
B
(cid:13)(cid:13)(cid:13)(cid:18)α1Im
α2In(cid:19)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:18)α1 kBk
α2 (cid:19)(cid:13)(cid:13)(cid:13).
α2In(cid:17)(cid:13)(cid:13) ≤ 1(cid:9)
S1 =(cid:8)(cid:0)(α1, α2); B(cid:1) :(cid:13)(cid:13)(cid:16) α1Im B
0 α2 (cid:17)(cid:13)(cid:13) ≤ 1(cid:9).
S2 =(cid:8)(cid:0)(α1, α2); B(cid:1) :(cid:13)(cid:13)(cid:16) α1 kBk
0
To prove the lemma, it is sufficient to show that these unit balls are the same.
From Proposition 7.1 the condition for the contractivity of the elements of S1 and S2 is the
same, that is,
kBk2 ≤ (1 − α12)(1 − α22)
(cid:3)
0
The important observation from the lemma above is that, for fixed α1, α2, the norm of the
α2In(cid:17) depends only on kBk.
matrix(cid:16) α1Im B
0 0(cid:1). Given any m × n matrix
Now consider the pair A = (A1, A2) with A1 =(cid:0) α1 0
B with kBk = β we have the following isometric embedding of (C2,k · kA) into (Mm+n,k · kop)
(z1, z2) 7→(cid:18)z1α1Im
0 α2(cid:1) , A2 =(cid:0) 0 β
z1α2In(cid:19) .
For various choices of the dimensions m, n and the matrix B, this represents a large collection of
isometric embeddings.
z2B
0
For fixed α1, α2, we let XB represent the above embedding of (C2,k · kA) into (Mm+n,k · kop).
We now show that the operator space structures determined by these embeddings depend only
If VA is the space (C2,k · kA), then (XB ⊗ Ik) gives the embedding of Mk(VA) into
on kBk.
Mk(Mm+n(C)). An element of Mk(VA) is defined by a pair of k × k matrices Z1, Z2, and the
corresponding embedding into Mk(Mm+n(C)) has the form
Z2 ⊗ B
It now remains to show that the operator norm of this matrix depends only on kBk. Using
Proposition 7.1 it can be shown that
0
(cid:18)α1Z1 ⊗ Im
α2Z1 ⊗ In(cid:19) .
α2Z1 ⊗ In(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1 if and only if (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Z2 ⊗ B
(cid:18)α1Z1 ⊗ Im
0
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18)α1Z1 Z2kBk
0
α2Z1 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1.
Hence it follows that these two norms are in fact equal. We have therefore proved the following
theorem.
Theorem 7.3. For all m × n matrices B with the same (operator) norm, the operator space
structures on Cm+n, determined by the different embeddings
(z1, z2) 7→ z1(cid:16) α1Im
0
0
α2In (cid:17) + z2(cid:16) 0 B
0 (cid:17), α1, α2 ∈ C,
0
are completely isometric irrespective of the particular choice of B. Moreover all of them are com-
pletely isometric to the MIN space.
Acknowledgment. The authors gratefully acknowledge the help they have received from Sayan
Bagchi, Michael Dritschel and Dmitry Yakubovich.
CONTRACTIVITY AND COMPLETE CONTRACTIVITY FOR FINITE DIMENSIONAL BANACH SPACES 19
References
[1] W. Arveson, Subalgebras of C ∗-algebras II, Acta Math., 128 (1972), 271 - 308.
[2] J. Agler, Rational dilation on an annulus, Ann. of Math., 121 (1985), 537 - 563.
[3] J. Agler and N. J. Young, Operators having the symmetrized bidisc as a spectral set, Proc. Edinburgh Math.
Soc., 43 (2000), 195 - 210.
[4] B. Bagchi and G. Misra, Contractive homomorphisms and tensor product norms, Integral Equations and Op-
erator Theory, 21 (1995), 255 - 269.
[5] M. J. Cowen and R. G. Douglas, Complex geometry and operator theory, Acta Math., 141 (1978), 187 - 261.
[6] M. J. Cowen and R. G. Douglas, Operators possessing an open set of eigenvalues, Functions, series, operators,
Vol. I, II (Budapest, 1980), 323 - 341, Colloq. Math. Soc. Janos Bolyai, 35, North-Holland, Amsterdam, 1983.
[7] R. G. Douglas, P. S. Muhly and Carl Pearcy, Lifting Commuting Operators, Michigan Math. Journal, 15 (1968),
385 - 395.
[8] M. Dritschel and S. McCullough, The failure of rational dilation on a triply connected domain, J. Amer. Math.
Soc., 18 (2005), 873 - 918.
[9] G. Misra, Curvature inequalities and extremal properties of bundle shifts, J. Operator Th., 11 (1984), 305 -
317.
[10] G. Misra, Completely contractive Hilbert modules and Parrott's example, Acta Math. Hungar., 63 (1994), 291
- 303.
[11] G. Misra and N. S. N. Sastry, Contractive modules, extremal problems and curvature inequalities, J. Funct.
Anal., 88 (1990), 118 - 134.
[12] G. Misra and N. S. N. Sastry, Completely contractive modules and associated extremal problems, J. Funct.
Anal., 91 (1990), 213 - 220.
[13] G. Misra and A. Pal, Contractivity, complete contractivity and curvature inequalities, to appear, Journal
d'Analyse Mathematique.
[14] B. Sz.-Nagy and C. Foias, Harmonic analysis of Hilbert space operators, NorthHolland, 1970.
[15] S. Parrott, Unitary dilations for commuting contractions, Pac. J. Math., 34 (1970), 481 - 490.
[16] G. Pisier, An Introduction to the Theory of Operator Spaces, Cambridge University Press, 2003.
[17] V. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge University Press, 2002.
[18] V. Paulsen, Representations of function algebras, abstract operator Spaces and Banach space geometry, J. Funct.
Anal., 109 (1992), 113 - 129.
(G. Misra) Department of mathematics, Indian Institute of Science, Bangalore - 560 012, India
E-mail address, G. Misra: [email protected]
(A. Pal) Department of Mathematics and Statistics, Indian Institute of Science Education And
Research Kolkata, Mohanpur - 741 246
E-mail address, A. Pal: [email protected]
(C. Varughese) Renaissance Communications, Bangalore - 560 058
E-mail address, C. Varughese: [email protected]
|
1804.10349 | 1 | 1804 | 2018-04-27T06:00:46 | Measures of Noncompactness in $\left(\bar{N}_{\Delta^{-}}^{q}\right)$ Summable Difference Sequence Spaces | [
"math.FA"
] | In the given paper we first introduce $\bar{N}_{\Delta^{-}}^{q}$ summable difference sequence spaces and prove some properties of these spaces. We then obtain the necessary and sufficient conditions for infinite matrices $A$ to map these sequence spaces on the spaces $c_0, c$ and $\ell_\infty$, the Hausdorff measure of noncompactness is then used to obtain the necessary and sufficient conditions for the compactness of the linear operators defined on these spaces. | math.FA | math |
Measures of Noncompactness in (cid:0) ¯N q
∆−(cid:1)
Summable Difference Sequence Spaces
Tanweer Jalal ∗
National Institute of Technology, Srinagar
Ishfaq Ahmad Malik
National Institute of Technology, Srinagar
Abstract. In the given paper we first introduce ¯N q
∆− summable dif-
ference sequence spaces and prove some properties of these spaces. We
then obtain the necessary and sufficient conditions for infinite matrices
A to map these sequence spaces on the spaces c0, c and ℓ∞, the Haus-
dorff measure of noncompactness is then used to obtain the necessary
and sufficient conditions for the compactness of the linear operators
defined on these spaces.
AMS Subject Classification: 40H05, 46A45, 47B07
Keywords and Phrases: Difference sequence space, BK spaces, ma-
trix transformations, Measures of noncompactness
1
Introduction and Preliminaries
We write w for the set of all complex sequences x = (xk)∞
k=0 and
φ, c0, c and ℓ∞ for the sets of all finite, convergent sequences and se-
quences convergent to zero, and bounded respectively. The sequence
e is given by e = (1, 1, 1, . . .) and e(n) is the sequence with 1 as only
nonzero term at the nth place for each n ∈ N, where N = {0, 1, 2, . . .}.
Further by cs and ℓ1 we denote the spaces of all sequences whose series
is convergent and absolutely convergent respectively.
The β−dual of a subset X of w is defined by
X β = {a ∈ w : ax = (akxk) ∈ cs for all x = (xk) ∈ X}
If A is an infinite matrix with complex entries ank n, k ∈ N, we write
An for the sequence in the nth row of A, An = (ank)∞
k=0 n ∈ N . The
∗Corresponding Author
2
Tanweer Jalal , Ishfaq Ahmad Malik
A−transform of any x = (xk) ∈ w is given by Ax = (An(x))∞
k=0, where
An(x) =
∞
Xk=0
ankxk
n ∈ N
the series on right must converge for each n ∈ N.
If X and Y are subsets of w, we denote by (X, Y ), the class of all
infinite matrices that map X into Y . So A ∈ (X, Y ) if and only if
An ∈ X β , n = 0, 1, 2, . . . and Ax ∈ Y for all x ∈ X. The matrix
domain of an infinite matrix A in X is defined by
XA = {x ∈ w : Ax ∈ X}
If X and Y are Banach Spaces, then by B(X, Y ) we denote the set of
all bounded (continuous) linear operators L : X → Y , which is itself a
Banach space with operator norm kLk = supx {kL(x)kY : kxk = 1} for
all L ∈ B(X, Y ). The linear operator L : X → Y is said to be compact
if the domain of L is all of X and every bounded sequence (xn) ∈ X ,
the sequence (L(xn)) has a sub-sequence which converges in Y . The op-
erator L ∈ B(X, Y ) is said to be of finite rank if dim R(L) < ∞, where
R(L) denotes the range space of L. A finite rank operator is clearly
compact.
The concept of difference sequence spaces was first introduced by Kiz-
maz [1] and later several authors studeid new sequence spaces defined
by using difference operators like Mursaleen and Nouman [2] and many
more.
In the past, several authors studied matrix transformations on sequence
spaces that are the matrix domains of the difference operator, or of the
matrices of the classical methods of summability in spaces such as ℓp,
c0, c,ℓ∞ or others. For instance, some matrix domains of the difference
operator were studied in ([1],[3] ), of the Riesz matrices in [4], and so
on. In this paper, we first define a new summable difference sequence
space as the matrix domains XT of arbitrary triangles ¯Nq and ∆− and
obtain it basis, β dual of the new sequence spaces. We then find out
the necessary and sufficient condition for the exists of matrix transfor-
mations and finally obtain the results related to the compactness of the
linear operators on these new sequence spaces.
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
3
2
¯N q
∆− Summable Difference Sequence Spaces
Define the difference operator as
∆−xk = xk−1 − xk , k = 0, 1, 2, . . . where x−1 = 0
(1)
The ∆− = (δnk)∞
n,k=0 is a triangular matrix written as
−1
1
0
k = n
k = n − 1
k > n
δnk =
The inverse of this matrix is S = (snk) given as
snk =(cid:26)−1 0 ≤ k ≤ n
k > n
0
k=0 be positive sequences and (Qn)∞
i=0 qi. The ( ¯N , q) transform of of the sequence (xk)∞
n=0 be the sequence defined
k=0 is
Let (qk)∞
as Qn = Pn
defined as
tn =
1
Qn
qixi
∞
Xi=0
The matrix ¯Nq for this transformation can be written as
( ¯Nq)nk =(cid:26) qk
Qn
0
0 ≤ k ≤ n
k > n
The inverse of this matrix is [5]
( ¯Nq)−1
nk =((−1)n−k Qk
qn
0
n − 1 ≤ k ≤ n
0 ≤ k ≤ n − 2, k > n
We define the ¯N q
∆− summable difference sequence spaces as
( ¯N q
( ¯N q
( ¯N q
∆−)0
= (c0, ∆−) ¯Nq =(x ∈ w : ¯Nq∆−x = 1
= (c, ∆−) ¯Nq =(x ∈ w : ¯Nq∆−x = 1
∆−)∞ = (ℓ∞, ∆−) ¯Nq =(x ∈ w : ¯Nq∆−x = 1
∆−)
Qn
Qn
Qn
n
n
Xk=0
Xk=0
Xk=0
n
qk∆−xk!∞
qk∆−xk!∞
qk∆−xk!∞
n=0
n=0
n=0
∈ c0)
∈ c)
∈ ℓ∞)
4
Tanweer Jalal , Ishfaq Ahmad Malik
For any sequence x = (xk)∞
term given by
k=0, define τ = τ (x) as the sequence with nth
τn = ( ¯N q
∆−)n(x) =
1
Qn
n
Xk=0
qk∆−xk
(n = 0, 1, 2, . . .)
(2)
For any two sequences x and y, the product xy = (xkyk)∞
k=0.
2.1 Basis for the new sequence spaces
Proposition 2.1. [ [6], 1.4.8, p.9]
Every triangle T has a unique inverse S = (snk)∞
triangle, and x = T (S(x)) = S(T (x)) for all x ∈ w.
n,k=0 which is also a
Proposition 2.2. [[7], Theorem 2.3] If (cid:0)b(n)(cid:1)∞
metric space (X, d), then (cid:0)S(b(n))(cid:1)∞
n=0 is a basis of the linear
n=0 is a basis of Z = XT with metric
dT defined by dT (z, ¯z) = d(T (z), T (¯z)) for all z, ¯z ∈ Z. Where S is the
inverse of the matrix T .
It is obvious that (c0, ∆−) ¯Nq = (c0) ¯Nq ·∆−, So the basis for new spaces
of the elements of (c0, ∆−) ¯Nq as
∆−(cid:1) x(cid:1)k for all k ∈ N. Define the sequence
are given by (cid:0) ¯Nq · ∆−(cid:1)−1(cid:0)e(n)(cid:1) = (∆−)−1 ·(cid:0) ¯Nq(cid:1)−1(cid:0)e(n)(cid:1)we have
Theorem 2.3. Let λk =(cid:0)(cid:0) ¯N q
n on∈N
s(k) =ns(k)
j=1 Qj(cid:16) 1
n =
Pk
i) The sequence(cid:8)s(k)(cid:9)k∈N is a basis for the space (c0, ∆−) ¯Nq and any
x ∈ (c0, ∆−) ¯Nq can be uniquely represented in the form
qj(cid:17) 0 ≤ k < n
for every fixed k ∈ N. Then
k−1
Xj=1
,
s(−1)
n =
k = n
k > n
− 1
qj+1
− Qk
qk
0
sk
n
Xk=0
Qj(cid:18) 1
qj+1
−
1
qj(cid:19) +
Qk
qk
.
λks(k)
x =Xk
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
5
ii) The set (cid:8)e, s(k)(cid:9) is a basis for the spaces (c, ∆−) ¯Nq and any x ∈
(c, ∆−) ¯Nq has a unique representation in the form
x = ls(−1)
n +Xk
where for all k ∈ N, l = limk→∞(cid:0)(cid:0) ¯N q
∆−(cid:1) x(cid:1)k.
(λk − l)s(k)
Proof. Since (X, ∆−) ¯Nq = (X) ¯Nq ·∆− for X = c0, c, ℓ∞. Now e =
(e(k))∞
Now ¯Nq is a triangle and ∆− is triangle so ¯Nq · ∆− is also a triangle and
k= is the standard basis for c and by
·(cid:0) ¯Nq(cid:1)−1 =
=(cid:0)∆−(cid:1)−1
(cid:0) ¯Nq · ∆−(cid:1)−1
Hence (cid:8)s(k)(cid:9)k∈N is a basis for the space (c0, ∆−) ¯Nq and the results i)
and ii) are obvious to follow.
Note: We consider the standard basis to find the general results related
to our sequence spaces.
k = n
k > n
(cid:3)
− 1
qk(cid:17) 0 ≤ k < n
Qk(cid:16) 1
qk+1
− Qn
qn
0
∆−(cid:1) and (cid:0) ¯N q
∆−(cid:1)∞ are
∆−
kxk ¯N q
given by
BK-spaces with norm k k ¯N q
∆−(cid:1)0, (cid:0) ¯N q
Theorem 2.4. The sequence spaces (cid:0) ¯N q
n (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
qk∆−xk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=0
x = le +Xk
If Qn → ∞ (n → ∞), then ( ¯N q
(xk)∞
∆−) has unique representation
k=0 ∈ ( ¯N q
(λk − l)e(k)
1
Qn
= sup
∆−
n
where l ∈ C is such that x − le ∈ ( ¯N q
∆−)0
∆−)0 has AK, and every sequence x =
(3)
Proof. Since (X, ∆−) ¯Nq
= X ¯Nq ·∆− for all X = c0, c, ℓ∞ and the spaces
c0, c, ℓ∞ are BK spaces with respect to natural norm [[8], p.217-218] and
The space ( ¯N q
the matrix ¯Nq · ∆− is a triangle so by Theorem 4.3.12, [6], gives(cid:0) ¯N q
∆−(cid:1) and (cid:0) ¯N q
(cid:0) ¯N q
(cid:0) ¯N q
∆−(cid:1) are simply followed from Theorem 2 of [9] and [10].
∆−(cid:1)∞ are BK spaces
∆−)0 has AK and the unique representation of elements of
∆−(cid:1)0,
(cid:3)
6
Tanweer Jalal , Ishfaq Ahmad Malik
2.2
β dual of the new spaces
In order to find the β dual we need the results of [11] which are
Lemma 2.5. A ∈ (c0 : l1) if and only if
Lemma 2.6. A ∈ (c0 : c) if and only if
< ∞
sup
ank(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
K ∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk∈K
n X ank < ∞,
sup
ank − αk = 0.
lim
n→∞
Lemma 2.7. A ∈ (c0 : ℓ∞) if and only if
Theorem 2.8. Let (qk)∞
a = (ak) ∈ w we define a matrix C = (cnk)∞
sup
n X ank < ∞,
k=0 be positive sequences, Qn = Pn
n,k=0 as
i=0 qi and
j=k+1 aj
qk(cid:17)Pn
− 1
− Qkak
qk
0
0 ≤ k < n
k = n
k > n
qk+1
Qk(cid:16) 1
cnk =
cnk < ∞)
n Xk
cnk =Xk (cid:12)(cid:12)(cid:12)
n→∞Xk
= c1 ∩ c2 , (cid:2)(cid:0) ¯N q
= c2 ∩ c3.
lim
and consider the sets
c1 =(a ∈ w : sup
c3 =(a ∈ w :
∆−(cid:1)0iβ
Then h(cid:0) ¯N q
∆−(cid:1)∞iβ
h(cid:0) ¯N q
; c2 =na ∈ w :
) ; c4 =(a ∈ w :
cnk(cid:12)(cid:12)(cid:12)
lim
n→∞
∆−(cid:1)(cid:3)β = c1 ∩ c2 ∩ c4 and
lim
n→∞
cnk exists for each k ∈ No
cnk exists )
lim
n→∞Xk
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
7
Proof. We prove the result for h(cid:0) ¯N q
procedure can be followed. Let x ∈ (cid:0) ¯N q
∆− x.
that y = ¯N q
Hence
for the other two same
∆−(cid:1)0iβ
∆−(cid:1)0 then there exists a y such
akxk =
n
Xk=0
=
=
n
n
Xk=0
Xk=0
Xk=0
n
yk
k−1
∆−(cid:1)−1
ak(cid:0) ¯N q
ak
Qj(cid:18) 1
Xj=0
Qk−1(cid:18) 1
qk
−
= (Cy)n
qj+1
1
−
qj(cid:19) yj −
qk−1(cid:19) n
Xj=k+1
1
aj −
Qk
qk
yk
qk
yk
Qkak
∆−(cid:1)0 if and only if Cy ∈ cs
show the other two results as well.
By Theorem 7.2.9, [6] we know that if X is a BK-space and a ∈ w then
= c1 ∩ c2 In the same way we can
(cid:3)
∞
whenever y ∈ c0.
So ax = (anxn) ∈ cs whenever x ∈ (cid:0) ¯N q
∆−(cid:1)0iβ
Using Lemma 2.6 we get h(cid:0) ¯N q
kak∗ = sup((cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
akxk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=0
∆−(cid:1)∞iβ
∆−(cid:1)0iβ
∆−(cid:1)(cid:3)β and h(cid:0) ¯N q
Theorem 2.9. For h(cid:0) ¯N q
Qk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
aj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
qk(cid:19) n
qn (cid:12)(cid:12)(cid:12)(cid:12)
+(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18) 1
Xj=k+1
: kxk = 1)
, (cid:2)(cid:0) ¯N q
n
k k∗ is defined as
kak∗ = sup
Xk=0
Qnan
qk+1
n−1
−
1
provided the term on the right side exists and is finite, which is the case
whenever a ∈ X β.
the norm
8
Tanweer Jalal , Ishfaq Ahmad Malik
Proof. If x[n] denotes the nth section of the sequence x ∈(cid:0) ¯N q
using (2) we have
∆−(cid:1)0 then
τ [n]
k = τk(x[n]) =
1
Qk
k
Xj=0
qj∆−x[n]
j
Let a ∈ h(cid:0) ¯N q
sequence d[n] as
∆−(cid:1)0iβ
k =
d[n]
, then for any non-negative integer n define the
Qk(cid:16) 1
qk+1
j=k+1 aj
qk(cid:17)Pn
− 1
− Qkak
qk
0
0 ≤ k < n
k = n
k > n
k=0 d[n]
k (cid:17) where Π = (cid:2)(cid:0) ¯N q
∆−(cid:1)(cid:3)β.
is obvious.
The inequality kakΠ ≤ kak∗
Also
Let kakΠ = supn kd[n]k1 = supn(cid:16)P∞
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=0
k (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
akx[n]
1
qj
n−1
1
∞
k
n
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=0
≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=0
≤ sup
k
n
−
(Qjτ [n]
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
j−1)
ak
j − Qj−1τ [n]
Xj=0
k (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
qk(cid:19)
aj
+(cid:12)(cid:12)(cid:12)(cid:12)
Qk(cid:18) 1
τ [n]
Xj=k+1
k ·
qk(cid:19) n
aj +(cid:12)(cid:12)(cid:12)(cid:12)
Qk(cid:18) 1
Xj=k+1
kd[n]k1
qk+1
qk+1
−
1
τ [n]
= kx[n]k ¯N q
= kakΠkx[n]k ¯N q
∆−
∆−
τ [n]
n
anQn
qn (cid:12)(cid:12)(cid:12)(cid:12)
qn (cid:12)(cid:12)(cid:12)(cid:12)
anQn
Hence kak∗ ≤ kakΠ
From the above inequalities we get the required conclusion.
Following are some well known results
(cid:3)
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
9
Proposition 2.10. (cf. [12], Theorem 7) Let X and Y be BK spaces,
then (X, Y ) ⊂ B(X, Y ) that is every matrix A from X into Y defines an
element LA of B(X, Y ) where
LA(x) = A(x)
∀ x ∈ X
Also A ∈ (X, ℓ∞) if and only if
kAk∗ = sup
n
kAnk∗ = kLAk < ∞
k=0 is a basis of X, Y and Y1 are FK spaces with Y1 a closed
If (cid:0)b(k)(cid:1)∞
subspace of Y , then A ∈ (X, Y1) if and only if A ∈ (X, Y ) and A(cid:0)b(k)(cid:1) ∈
Y1 for all k = 0, 1, 2, . . ..
Proposition 2.11. (cf. [13], Proposition 3.4) Let T be a triangle
(i) If X and Y are subsets of w, then A ∈ (X, YT ) if and only if
B = T A ∈ (X, Y ).
(ii) If X and Y are BK spaces and A ∈ (X, YT ), then
kLAk = kLBk
Using Proposition 2.10 and Theorem 2.9 we can easily conclude that
following:
Corollary 2.12. Let (qk)∞
∆− be the difference operator as defined in (1), then
k=0 be a positive sequence, Qn =Pn
k=0 qk and
m−1
, ℓ∞(cid:17) if and only if
i) A ∈(cid:16)(cid:0)N q
∆−(cid:1)∞
Qk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
m,n
qk(cid:19) m
(cid:18) 1
Xk=0
Xj=k+1
qk+1
and
sup
1
−
anj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Qmanm
qm (cid:12)(cid:12)(cid:12)(cid:12)
< ∞ (4)
+(cid:12)(cid:12)(cid:12)(cid:12)
AnQ
q
∈ c0
∀ n = 0, 1, . . .
(5)
10
Tanweer Jalal , Ishfaq Ahmad Malik
ii) A ∈(cid:0)(cid:0)N q
∆−(cid:1) , ℓ∞(cid:1) if and only if condition (4) holds and
∈ c
∀ n = 0, 1, 2, . . .
AnQ
q
iii) A ∈(cid:16)(cid:0)N q
iv) A ∈(cid:16)(cid:0)N q
∆−(cid:1)0
∆−(cid:1)0
v) A ∈(cid:16)(cid:0)N q
∆−(cid:1)0
, ℓ∞(cid:17) if and only if condition (4) holds.
, c0(cid:17) if and only if condition (4) holds and
lim
n→∞
ank = 0
for all k = 0, 1, 2 . . .
, c(cid:17) if and only if condition (4) holds and
lim
n→∞
ank = αk
for all k = 0, 1, 2 . . .
(6)
(7)
(8)
vi) A ∈ (cid:0)(cid:0)N q
and
∆−(cid:1) , c0(cid:1) if and only if conditions (4), (5) and (7) holds
∞
ank = 0
for all k = 0, 1, 2 . . .
(9)
lim
n→∞
Xk=0
lim
n→∞
Xk=0
vii) A ∈ (cid:0)(cid:0)N q
and
∆−(cid:1) , c(cid:1) if and only if conditions (4), (5) and (8) holds
∞
ank = α
for all k = 0, 1, 2 . . .
(10)
Again Using Proposition 2.11 and Theorem 2.4 we have the following
corollary:
3 Hausdorff Measure of Noncompactness
From Mursaleen et. al. [14] we let S and M be the subsets of a metric
space (X, d) and ǫ > 0. Then S is called an ǫ−net of M in X if for every
x ∈ M there exists s ∈ S such that d(x, s) < ǫ. Further, if the set S is
finite, then the ǫ−net S of M is called finite ǫ−net of M . A subset of
a metric space is said to be totally bounded if it has a finite ǫ−net for
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
11
every ǫ > 0.
If MX denotes the collection of all bounded subsets of metric space
(X, d). If Q ∈ MX then the Hausdorff Measure of Noncompactness of
the set Q is defined by
χ(Q) = inf {ǫ > 0 : Q has a finite ǫ − net in X}
The function χ : MX → [0, ∞) is called Hausdorff Measure of Noncom-
pactness [15]
The basic properties of Hausdorff Measure of Noncompactness can be
found in ([5], [16], [15]). Some of those properties are
If Q, Q1 and Q2 are bounded subsets of a metric space (X, d), then
χ(Q) = 0 ⇔ Q is totally bpunded set,
χ(Q) = χ( ¯Q),
Q1 ⊂ Q2 ⇒ χ(Q1) ≤ χ(Q2),
χ(Q1 ∪ Q2) = max {χ(Q1), χ(Q2)} ,
χ(Q1 ∩ Q2) = min {χ(Q1), χ(Q2)} .
Further if X is a normed space the χ has the additional properties con-
nected with the linear structure.
χ(Q1 + Q2) ≤ χ(Q1) + χ(Q2)
χ(ηQ) = ηχ(Q)
η ∈ C
If X and Y are normed space, then for A ∈ B(X, Y ) the Hausdorff
Measure of Noncompactness of A, is denoted by kAkχ and is defined as
kAkχ = χ(AB)
Where B = {x ∈ X : kxk = 1} is the unit ball in X.
Also A is said to be compact if and only if kAkχ = 0 and kAkχ ≤ kAk.
Proposition 3.1. ([15], Theorem 6.1.1, X = c0) Let Q ∈ Mc0 and Pr :
c0 → c0 (r ∈ N be the operator defined by Pr(x) = (x0, x1, . . . , xr, 0, 0, . . .)
for all x = (xk) ∈ c0. Then, we have
χ(Q) = lim
r→∞ sup
x∈Q
k(I − Pr)(x)k!
12
Tanweer Jalal , Ishfaq Ahmad Malik
where I is the identity operator on c0.
Proposition 3.2. ([15], Theorem 6.1.1) Let X be a Banach space with
a Schauder basis {e1, e2, . . .}, and Q ∈ MX and Pn : X → X (n ∈ N be
the projector onto the linear span of {e1, e2, . . . , en}. Then, we have
1
a
lim
n→∞
sup sup
x∈Q
k(I − Pn)(x)k! ≤ χ(Q) ≤ inf
n sup
x∈Q
k(I − Pn)(x)k! ≤ lim
n→∞
sup sup
x∈Q
k(I − Pn)(x)k!
where a = limn→∞ sup kI − Pnk. If X = c then a = 2. (see [15], p.22).
,(cid:0)N q
∆−(cid:1)
Theorem 4.1. Consider the matrix A as in Corollary 2.12, and for any
integers n,s, n > s set
4 Compact operators on the spaces(cid:0)N q
∆−(cid:1)0
and (cid:0)N q
∆−(cid:1)∞
ani(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Qj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
m
+(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18) 1
Xj=0
∆−(cid:1)0 or (cid:0)N q
If X be either (cid:0)N q
∆−(cid:1) and A ∈ (X, c0). Then
qj(cid:19) m
Xi=j+1
qm (cid:12)(cid:12)(cid:12)(cid:12)
kAk(s) = sup
n>p
Qmanm
qj+1
sup
m−1
−
1
kLAkχ = lim
s→∞
kAk(s).
If X be either (cid:0)N q
∆−(cid:1)0 or (cid:0)N q
∆−(cid:1) and A ∈ (X, c). Then
· lim
s→∞
kAk(s) ≤ kLAkχ ≤ lim
r→∞
kAk(s).
1
2
(11)
(12)
(13)
and if X be either (cid:0)N q
∆−(cid:1)0 , (cid:0)N q
∆−(cid:1) or (cid:0)N q
0 ≤ kLAkχ ≤ lim
s→∞
∆−(cid:1)∞ and A ∈ (X, ℓ∞). Then
kAk(s).
(14)
Proof. Let F = {x ∈ X : kxk ≤ 1} if A ∈ (X, c0) and X is one of the
spaces (cid:0)N q
∆−(cid:1)0 or (cid:0)N q
∆−(cid:1), then by Proposition 3.1
kLAkχ = χ(AF ) = lim
s→∞(cid:20)sup
x∈F
k(I − Ps)Axk(cid:21)
(15)
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
13
Again using Proposition 2.10 and Corollary 2.12 we have
kAks = sup
x∈F
k(I − Ps)Axk
(16)
From (15) and (16) we get
kLAkχ = lim
s→∞
kAk(s).
Since every sequence x = (xk)∞
k=0 ∈ c has a unique representation
x = le +
∞
(xk − l)e(k)
Xk=0
where l ∈ C is such that x − le ∈ c0
We define Ps : c → c by Ps(x) = le +Ps
Then kI − Psk = 2 and using (16) and Proposition 3.2 we get
k=0(xk − l)e(k), s = 0, 1, 2, . . ..
1
2
· lim
s→∞
kAk(s) ≤ kLAkχ ≤ lim
s→∞
kAk(s)
Finally we define Ps : ℓ∞ → ℓ∞ by Ps(x) = (x0, x1, . . . , xs, 0, 0 . . .),
x = (xk) ∈ ℓ∞.
Clearly AF ⊂ Ps(AF ) + (I − Ps)(AF )
So using the properties of χ we get
χ(AF ) ≤ χ[Ps(AF )] + χ[(I − Ps)(AF )]
= χ[(I − Ps)(AF )]
≤ sup
x∈F
k(I − Ps)A(x)k
Hence by Proposition 2.10 and and Corollary 2.12 we get
0 ≤ kLAkχ ≤ lim
s→∞
kAk(s)
(cid:3)
A direct corollary of the above theorem is
Corollary 4.2. Consider the matrix A as in Corollary 2.12, and X =
∆−(cid:1)0 or X =(cid:0)N q
(cid:0)N q
∆−(cid:1) then if A ∈ (X, c0) or A ∈ (X, c) we have
LA is compact if and only if
kAk(s) = 0
lim
s→∞
14
Tanweer Jalal , Ishfaq Ahmad Malik
Further, for X = (cid:0)N q
(X, ℓ∞) then we have
∆−(cid:1)0 , X = (cid:0)N q
∆−(cid:1) or X = (cid:0)N q
∆−(cid:1)∞, if A ∈
LA is compact if
kAk(s) = 0
lim
s→∞
(17)
In (17) it is possible for LA to be compact although lims→∞ kAk(s) 6= 0,
that is the condition is only sufficient condition for LA to be compact.
For example, let the matrix A be defined as An = e(1) n = 0, 1, 2, . . .
and qn = 3n , n = 0, 1, 2, . . . .
Then by (4) we have
1
−
+
3
1
2
(1 − 3−n)(cid:19) < 2
But
Qmanm
m−1
Xk=0
(cid:18) 1
qk+1
sup
m,n
Qk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
qk(cid:19) m
Xj=k+1
qm (cid:12)(cid:12)(cid:12)(cid:12)
n (cid:18)2
= sup
, ℓ∞(cid:17) .
∆−(cid:1)∞
anj(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+(cid:12)(cid:12)(cid:12)(cid:12)
Now by Corollary 2.12 we know A ∈(cid:16)(cid:0)N q
(1 − 3−n)(cid:21) =
∆−(cid:1)∞, so LA is compact operator.
Since A(x) = x1 for all x ∈(cid:0)N q
Which gives kAk(s) = 7
n>s(cid:20) 2
3
kAk(s) = sup
2 · 3r+1
6 6= 0.
+
1
2
7
6
−
1
∀ r
References
[1] H Kizmaz. Certain sequence spaces. Can. Math. Bull., 24(2):169 --
176, 1981.
[2] Mohammad Mursaleen and Abdullah K Noman. On some new
difference sequence spaces of non-absolute type. Mathematical and
Computer Modelling, 52(3-4):603 -- 617, 2010.
[3] SM Sirajudeen. Matrix transformations of bv into. Indian J. pure
appl. Math, 23(1):55 -- 61, 1992.
[4] Bilal Altay and Feyzi Basar. On the paranormed riesz sequence
spaces of non-absolute type. Southeast Asian Bull. Math, 26(5):701 --
715, 2002.
Measures of Noncompactness in (cid:0) ¯N q
Spaces
∆−(cid:1) Summable Difference Sequence
15
[5] Jozef Banas and Mohammad Mursaleen. Sequence spaces and mea-
sures of noncompactness with applications to differential and inte-
gral equations. Springer, 2014.
[6] Albert Wilansky. Summability through functional analysis, vol-
ume 85. Elsevier, 2000.
[7] Abdullah M Jarrah and Eberhard Malkowsky. Ordinary, absolute
and strong summability and matrix transformations. Filomat, pages
59 -- 78, 2003.
[8] Ivor John Maddox. Elements of functional analysis. CUP Archive,
1988.
[9] AM Al-Jarrah and E Malkowsky. Bk spaces, bases and linear oper-
ators. Rend. del Circ. Mat. di Palermo. Serie II. Suppl, 52:177 -- 191,
1998.
[10] Eberhard Malkowsky and V Rakocevic. Measure of noncompact-
ness of linear operators between spaces of sequences that are (n,
q) summable or bounded. Czechoslovak Mathematical Journal,
51(3):505 -- 522, 2001.
[11] Michael Stieglitz and Hubert Tietz. Matrixtransformationen von
folgenraumen eine ergebnisubersicht. Mathematische Zeitschrift,
154(1):1 -- 16, 1977.
[12] E Malkowsky and V Rakocevic. The measure of noncompactness of
linear operators between certain sequence spaces. Acta Scientiarum
Mathematicarum, 64(1):151 -- 170, 1998.
[13] E Malkowsky and V Rakocevic. The measure of noncompactness of
linear operators between spaces of mth-order difference sequences.
Studia Scientiarum Mathematicarum Hungarica, 35(3-4):381 -- 396,
1999.
[14] Mohammad Mursaleen, Vatan Karakaya, Harun Polat, and
N Sim¸sek. Measure of noncompactness of matrix operators on some
difference sequence spaces of weighted means. Computers & Math-
ematics with Applications, 62(2):814 -- 820, 2011.
16
Tanweer Jalal , Ishfaq Ahmad Malik
[15] J. Banas and K. Goebl. Measures of noncompactness in Banach
spaces. Lecture Notes in Pure and Appl. Math. Number 60. 1980.
[16] Eberhard Malkowsky and Vladimir Rakocevi´c. An introduction
into the theory of sequence spaces and measures of noncompactness.
Number 17. Matematicki institut SANU, 2000.
Tanweer Jalal
Department of Mathematics
Associate Professor of Mathematics
National Institute of Technology, Srinagar
Srinagar, India
E-mail: [email protected]
Ishfaq Ahmad Malik
Department of Mathematics
Research Scholar
National Institute of Technology, Srinagar
Srinagar, India
E-mail: ishfaq [email protected]
|
1312.7587 | 1 | 1312 | 2013-12-29T20:51:50 | An embedding result | [
"math.FA"
] | In unbounded subset $\Omega$ in $R^n$ we study the operator $u\rightarrow gu$ as an operator defined in the Sobolev space $W^{r,p}(\Omega)$ and which takes values in $L^p(\Omega)$. The functions $g$ belong to wider spaces of $L^p$ connected with the Morrey type spaces. The main result is an embedding theorem from which we can deduce a Fefferman type inequality. | math.FA | math |
AN EMBEDDING RESULT
A. CANALE
Abstract. In unbounded subset Ω in Rn we study the operator u → gu as an operator
defined in the Sobolev space W r,p(Ω) and which takes values in Lp(Ω). The functions g
belong to wider spaces of Lp connected with the Morrey type spaces. The main result is
an embedding theorem from which we can deduce a Fefferman type inequality.
AMS Subject Classifications: 35J25, 46E35
Key Words: elliptic equations, multiplication operator, embedding theorems.
1. Introduction
Let Ω be an unbounded open subset in Rn.
In literature there are different results about the study of multiplication operator for a
suitable function g : Ω → C
u −→ gu,
(1.1)
as an operator defined in a Sobolev space (with or without weight) and which takes values
in a Lp(Ω) space.
In W 1,p
0
(Ω) or in W 1,p(Ω) with Ω regular enough, reference results are some well-known
inequalities which state the boundedness of (1.1): Hardy type inequalities (see H.Brezis
[2], A.Kufner [11], J.Necas [12]) when g(x) is an appropriate power of the distance of x
from a subset of ∂Ω, C.Fefferman inequality [10] (see, e.g. F.Chiarenza-M.Franciosi [8],
F.Chiarenza-M.Frasca [9]) obtained when g belongs to a suitable Morrey space.
In this paper we study the operator (1.1) in the Sobolev space W r,p(Ω), r ∈ N ,1 ≤ p ≤
+∞ when g belongs to suitable spaces Sp,s defined in Section 3.
One of the aspects of our interest in this type of inequality lies in the fact that the
embedding results are useful tools to prove a priori bounds when studying elliptic equations.
For applications in the study of the a priori bounds see [3], [4], [5], [6], [7]. The spaces
considered in some of these papers are connected to the spaces Sp,s.
These spaces are wider than Lp spaces, than classical Morrey space and are connected,
for suitable values of s, to the well known Morrey spaces defined when Ω is an unbounded
open subset in Rn.
We characterize the classes of functions in Sp,s and their inclusion properties in Section
3. In Section 4 we state the main result of this paper. We prove that if g ∈ Sq,s(Ω) for an
appropriate q ∈ [p, +∞[ and s ≥ 0, then (1.1) defines a bounded operator from W r,p(Ω) in
Lp(Ω).
We remark that our results imply that, for any function g : Ω → C such that
sup
x∈Ω
ρ∈]0,d]
ρq−nZΩ∩B(x,ρ)
gq < +∞,
d > 0,
q ≥ p, q ≥ n, q > n se n = p > 1,
2
A. CANALE
the following inequality holds
kgukLp(Ω) ≤ ckgk
q, s
p (Ω)
S
k∇ukLp(Ω)
∀u ∈ W 1,p
0
(Ω).
2. Notations
Let Rn be the n-dimensional real euclidean space. We set
B(x, r) = {y ∈ Rn :
y − x < r}, Br = B(0, r)
∀x ∈ Rn, ∀r ∈ R+.
For any x ∈ Rn, we call open infinite cone having vertex at x every set of the type
{x + λ(y − x) : λ ∈ R+, y − z < r},
where r ∈ R+ and z ∈ Rn are such that z − x > r.
For all θ ∈]0, π/2[ and for all x ∈ Rn we denote by Cθ(x) an open infinite cone having
vertex at x and opening θ.
For a fixed Cθ(x), we set
Cθ(x, h) = Cθ(x) ∩ B(x, h) ,
∀h ∈ R+.
Let Ω be an open set in Rn. We denote by Γ(Ω, θ, h) the family of open cones C ⊂⊂ Ω of
opening θ and height h.
We assume that the following hypothesis is satisfied:
h1) There exists θ ∈]0, π/2[ such that
∀x ∈ Ω
∃Cθ(x)
such that Cθ(x, ρ) ⊂ Ω.
3. Spaces Sp,s(Ω)
Let (Ω(x))x∈Ω be the family of open sets in Rn defined as
Ω(x) = B(x, ρ) ∩ Ω,
x ∈ Ω,
ρ > 0.
If 1 ≤ p < +∞ and s ∈ R, we denote by Sp,s(Ω) the space of functions g ∈ Lp
that
loc(Ω) such
kgkSp,s(Ω) = sup
ρ∈]0,d](cid:16)ρs−n/p kgkLp(Ω(x))(cid:17) < +∞,
x∈Ω
d > 0,
(3.1)
with the norm defined by (3.1).
We remark that
L∞(Ω) ֒→ Sp,s(Ω)
∀p ∈ [1, +∞[ and ∀s ≥ 0.
Indeed, if g ∈ L∞(Ω), we get
kgkSp,s(Ω) = sup
ρ∈]0,d](cid:16)ρs−n/p kgkLp(Ω(x))(cid:17) ≤
x∈Ω
≤ kgkL∞(Ω)
sup
x∈Ω
ρ+∈]0,d]
ρs(cid:16)ρ−n/pΩ(x)1/p(cid:17) = ckgkL∞(Ω)
where c = c(n, p).
(3.2)
EMBEDDING RESULT
The following inclusions hold
Lr(Ω) ֒→ Sq,s(Ω),
s ≥
n
q
,
1 ≤ q ≤ r ≤ +∞,
and
In particular
Indeed
Sq,s(Ω) ֒→ Sp,s(Ω)
1 ≤ p ≤ q ≤ +∞,
Lp(Ω) ֒→ Sp, n
p (Ω).
kgkSq,s(Ω) ≤ sup
x∈Ω
(ρs− n
q kgkLr(Ω(x))Ω(x)
1
q − 1
r ) ≤
ρ∈]0,d]
≤ c1 sup
x∈Ω
ρ∈]0,d]
ρs− n
r kgkLr(Ω(x)) ≤ c1kgkLr(Ω),
where c1 is a constant independent of g. From which we deduce also that
kgkSp,s(Ω) ≤ c2 sup
x∈Ω
ρ∈]0,d]
ρs− n
q kgkLq(Ω(x)),
where c2 is a constant independent of g.
3
(3.3)
(3.4)
(3.5)
(3.6)
Regard to the inclusion (3.5) we note that, for example, the constant functions belong to
p (Ω) and do not belong to Lp(Ω). Furthermore the function
p for
1+xα belongs to Sp, n
Sp, n
any α > 0 but does not belong to Lp if α ∈]0, n
1
p [.
Remark 3.1. When Ω = Rn the space Sp,s includes the classical Morrey spaces Lp,n−sp,
0 ≤ s ≤ n
p , defined as the space of functions g ∈ Lp
loc(Rn) such that
kgkLp,n−sp(Rn) = sup
ρ>0 (cid:16)ρs−n/p kgkLp(B(x,ρ)(cid:17) < +∞.
x∈Ω
4. Embedding result
Let us consider the function
φ : (x, y) ∈ Ω × Ω −→(1
0
if
if
y ∈ Ω(x)
y 6∈ Ω(x).
and, for any x ∈ Ω, we set
E(x) = {y ∈ Ω : x ∈ Ω(y)}.
(3.7)
(4.1)
(4.2)
Lemma 4.1. For any x ∈ Ω, E(x) is a measurable set and there exist c1, c2 ∈ R+ such
that
c1ρn ≤ E(x) ≤ c2ρn
∀x ∈ Ω.
(4.3)
4
A. CANALE
Proof. Clearly the function φ defined by (4.1) is a measurable function. Then, for any
fixed y ∈ Ω, the function φy : x ∈ Ω → φ(x, y) is measurable. Since φy is the characteristic
function of E(y), we have that E(y) is measurable.
Now we prove that (4.3) holds. The inequality on the right is easily proved. We will
prove the inequality on the left.
Let us consider Cθ(x, ρ) such that Cθ(x, ρ) ⊂ Ω. We get
Cθ(x, ρ) ⊂ E(x).
In fact let y ∈ Cθ(x, ρ). Then there exists a cone C ∈ Γ(Ω, θ, h) such that x, y ∈ C. So
Thus the inequality (4.3) is stated.
x ∈ B(y, ρ) ∩ Ω =⇒ y ∈ E(x).
(cid:3)
Now we state a Lemma which we will use in the proof of the embedding result.
Lemma 4.2. If h1) holds, then, for any s ≥ 0, we have v ∈ L1(Ω) if and only if the map
x ∈ Ω → ρ−nvL1(Ω(x)) belongs to L1(Ω). Therefore there exist c1, c2 ∈ R+ such that
c1kvkL1(Ω) ≤ZΩ
ρ−nkvkL1(Ω(x))dx ≤ c2kvkL1(Ω)
∀v ∈ L1(Ω).
(4.4)
Proof. The result is a consequence of the relation
ZΩ
ρ−nkvkL1(Ω(x))dx =ZΩ
=ZΩ
ρ−nZΩ
v(y)dyZE(y)
and of the Lemma 4.1.
v(y)φ(x, y)dxdy =
ρ−ndx
(4.5)
(cid:3)
Let r, s, p, q be real number with the condition
h2)
r ∈ N,
s ≥ 0,
1 ≤ p ≤ q < +∞,
q ≥
Let u ∈ W r,p(Ω). For any x ∈ Ω we set
n
r
,
q >
n
r
if
n
r
= p > 1.
Ψx : y ∈ Ω → x +
y − x
ρ
,
Ω∗(x) = Ψx(Ω(x)) ,
u∗ = (ux)∗ : z ∈ Ω∗(x) −→ u(x + ρ(z − x)) .
We note that
u∗ ∈ W r,p(Ω∗(x)).
We also note that, in consequence of h1) Ω∗(x) has the cone property, with the characteristic
cone having height and opening independent of x.
On the other hand, if τ = q/p, from h2) we get
τ ≥ 1,
τ > 1
if
n
r
= p > 1,
τ − 1
pτ
≥
1
p
−
r
n
.
From well-known imbedding theorems of Sobolev spaces (see, e.g., R.A. Adams [1]), we
deduce that
u∗ ∈ L
τ −1 (Ω∗(x))
pτ
EMBEDDING RESULT
and the following bound holds
τ −1 ,Ω∗(x) ≤ c0ku∗kW r,p(Ω∗(x)),
where c0 = c0(p, q, r, n) is a constant independent of x and u∗.
u∗ pτ
From (4.6) easily it follows that
ρ−n (τ −1)
pτ
u pτ
τ −1 ,Ω(x) ≤ c0 Xα≤r
ρα− n
p ∂αup,Ω(x).
5
(4.6)
(4.7)
Theorem 4.3. If h1) and h2) hold, then for any g ∈ Sq, s
p (Ω), s ≤ p, and for any
u ∈ W r,p(Ω) we get gu ∈ Lp(Ω) and
where the constant c = c(p, q, r, n) is independent of g and u.
kgukLp(Ω) ≤ ckgk
q, s
p (Ω)
S
kukW r,p(Ω),
(4.8)
Proof. Let u ∈ W r,p(Ω) and g ∈ Sq, s
p (Ω). By (4.4) and by Holder inequality it follows
that
ZΩ
gupdx ≤ c1ZΩ
ρ−nZΩ(x)
gupdy dx ≤ c1ZΩ
ρ−nkgkp
Lpτ (Ω(x))kukp
L
pτ
τ −1 (Ω(x))
dx ≤
(4.9)
≤ c1kgkp
S
q, s
p (Ω)
ρ−(s+n τ −1
kukp
L
pτ
τ −1 (Ω(x))
dx.
τ )ZΩ
On the other hand from (4.7) and Lemma 4.2 we obtain
ρ−n (τ −1)
τ ZΩ
kukp
L
pτ
τ −1 (Ω(x))
ραp−nZΩ
k∂αukp
Lp(Ω(x))dx ≤
ραpk∂αukp
Lp.
dx ≤ c0 Xα≤r
≤ c1 Xα≤r
From (4.9), (4.10) the inequality (4.8) follows.
(4.10)
(cid:3)
The following theorem is a consequence of the embedding result stated in the
Theorem 4.3 (see result of Fefferman [10] and also [9] for a simplified proof).
Theorem 4.4. If h1) and h2) hold, then for any g ∈ Sq, s
p (Ω), s ≤ p, we get
kgukLp(Ω) ≤ ckgk
q, s
p (Ω)
S
k∇ukLp(Ω)
∀u ∈ W 1,p
0
(Ω).
where the constant c = c(p, q, n) is independent of g and u.
Proof. Taking in mind the Hardy inequality, the proof is a direct consequence of the
(cid:3)
Theorem 4.3 when r = 1.
6
A. CANALE
References
[1] R.A.Adams, Sobolev spaces, Academic Press, 1975.
[2] H.Brezis, Analisi funzionale, Liguori Ed., 1990.
[3] A. Canale, A priori bounds in weighted spaces, J. Math. Anal. Appl. Vol. 287 (4) (2003), 101-117.
[4] A. Canale On some results in weighted spaces under Cordes type conditions, J. Interdiscip. Math.,
Vol. 10 (2)(2007), 245-261.
[5] A. Canale, On some results in weighted spaces under Chicco type conditions, Int. J. Pure Appl. Math.,
Vol. 31 (2) (2006), 185-202.
[6] A. Canale, Bounds in spaces of Morrey under Cordes type conditions, J. Appl. Funct. Anal., Vol. 3
(1) (2008), 11-32.
[7] A. Canale, Bounds in spaces of Morrey under Chicco type conditions, Math. Ineq. and Appl., Vol. 12
(2) (2009), 265-278.
[8] F.Chiarenza - M.Franciosi, A generalization of a theorem by C.Miranda, Ann. Mat. Pura Appl.,(IV)
(1992), 285-297.
[9] F.Chiarenza - M.Frasca, A remark on a paper by C.Fefferman, Proc. Amer. Math. Soc.,(2) (108)
(1990), 407-409.
[10] C.Fefferman, The uncertainty principle, Bull. Am. Math. Soc., (9) (1983), 129-206.
[11] A.Kufner, Weighted Sobolev spaces, Teubner Texte zur Math., Band (31) (1980).
[12] J.Necas, Les m´ethodes directes en theorie des ´equations elliptiques, Masson et C.ie Editeurs, Paris,
1967.
Dipartimento di Matematica, Universit`a degli Studi di Salerno, Via Giovanni Paolo II n. 132,
84084 FISCIANO (Sa), Italy.
E-mail address: [email protected]
|
1605.03411 | 2 | 1605 | 2016-12-13T13:45:51 | Lattice sub-tilings and frames in LCA groups | [
"math.FA"
] | Given a lattice $\Lambda$ in a locally compact abelian group $G$ and a measurable subset $\Omega$ with finite and positive measure, then the set of characters associated to the dual lattice form a frame for $L^2(\Omega)$ if and only if the distinct translates by $\Lambda$ of $\Omega$ have almost empty intersections. Some consequences of this results are the well-known Fuglede theorem for lattices, as well as a simple characterization for frames of modulates. | math.FA | math |
Lattice sub-tilings and frames in LCA groups
D. Barbieri∗, E. Hern´andez† and A. Mayeli‡
July 8, 2018
Abstract
Given a lattice Λ in a locally compact abelian group G and a measur-
able subset Ω with finite and positive measure, then the set of characters
associated to the dual lattice form a frame for L2(Ω) if and only if the
distinct translates by Λ of Ω have almost empty intersections. Some con-
sequences of this results are the well-known Fuglede theorem for lattices,
as well as a simple characterization for frames of modulates.
Keywords: Tiling sets, frames of exponentials, systems of translates.
MSC 2010 : 43A25, 42C15, 52C22
1 Introduction
Let G denote a locally compact and second countable abelian group (LCA
group). A closed subgroup Λ of G is called a lattice if it is discrete and co-
compact, i.e, the quotient group G/Λ is compact. Recall that, since G is second
countable, then any discrete subgroup of G is also countable (see e.g. [19, Sec-
tion 12, Example 17]). Assume that G is abelian, and denote the dual group by
bG. The dual lattice of Λ is defined as follows:
Λ⊥ = {χ ∈ bG : hχ, λi = 1 ∀λ ∈ Λ},
(1)
where hχ, λi indicates the action of character χ on the group element λ.
We recall that, by the duality theorem between subgroups and quotient
over, since G/Λ is compact, the dual lattice Λ⊥ is discrete. Notice also that
groups (see e.g. [21, Lemma 2.1.2]), the dual lattice Λ⊥ is a subgroup of bG that
is topologically isomorphic to the dual group of G/Λ, i.e., Λ⊥ ∼= dG/Λ. More-
bG/Λ⊥ ∼= bΛ, which implies that Λ⊥ is co-compact, hence it is a lattice.
∗Universidad Aut´onoma de Madrid, 28049 Madrid. E-mail : [email protected]
†Universidad Aut´onoma de Madrid, 28049 Madrid. E-mail : [email protected]
‡City University of New York, Queensborough and the Graduate Center. E-mail :
[email protected]
1
Let dg denote a Haar measure on G. For a function f in L1(G), the Fourier
transform of f is defined by
FG(f )(χ) =ZG
f (g)hχ, gi dg ,
χ ∈ bG,
where hχ, gi denotes the action of the character χ on g. By the inversion theorem
[21, Section 1.5.1], a Haar measure dχ can be chosen on bG so that the Fourier
transform FG is an isometry from L2(G) onto L2(bG). More precisely,
hFG(f ), FG(g)iL2( bG,dχ) = hf, giL2(G,dg)
f, g ∈ L2(G) .
(2)
For any χ ∈ bG, we define the exponential function eχ by
eχ : G → C,
eχ(g) := hχ, gi.
For any measurable subset Ω of G, we let Ω denote the Haar measure of Ω.
Throughout this paper, we let 1Ω denote the characteristic function of the set
Ω. We shall also use the addition symbol '+' for the group action, and 0 for the
neutral element, since G is abelian.
Definition 1 (Sub-Tiling). Let Ω ⊂ G be a measurable set with finite and
positive Haar measure, and let Λ be a lattice subgroup of G. We say that (Ω, Λ)
is a sub-tiling pair for G if
Xλ∈Λ
1Ω(g − λ) ≤ 1 a.e. g ∈ G .
(3)
By replacing the inequality with an equality, the definition is that of a tiling
pair. In this weaker form, it is equivalent to say that the translates of Ω by
elements of Λ are a.e. disjoint, i.e. (Ω, Λ) is a sub-tiling pair for G if and only if
Ω ∩ (Ω + λ) = 0 ∀ λ ∈ Λ , λ 6= 0.
Observe also that any sub-tiling set is a subset of a tiling set.
Our main result is the following. Recall that a cross section QΛ ⊂ G for a
group G and a lattice Λ is a measurable set of representatives of G/Λ.
Theorem 2 (Main Result). Let Λ be a lattice in G, let Ω ⊂ G be a set with
finite and positive measure, and let QΛ ⊂ G be a cross section for G/Λ. Then
the following are equivalent.
1) The pair (Ω, Λ) is sub-tiling for G.
2) For a.e. χ ∈ bG it holds
Xλ∈Λ⊥
FG(1Ω)(χ + λ)2 = QΛ Ω.
2
3) The system of translates {pΩ
L2(G).
−1
1Ω(· − λ) : λ ∈ Λ} is orthonormal in
4) The exponential set EΛ⊥ = {eλ : λ ∈ Λ⊥} is a frame for L2(Ω).
Moreover, if any of the above conditions holds, then the frame in point 4) is
tight, with constant QΛ.
As a first corollary we can obtain the following result, which was proved by
B. Fuglede in the Euclidean setting [6], and in the present setting by S. Pedersen
with a different approach ([18]).
Corollary 3. A set of finite and positive measure Ω tiles G with translations
by Λ if and only if the exponential set EΛ⊥ is an orthogonal basis for L2(Ω).
Let us now denote with M : Λ⊥ → U(L2(Ω)) the modulations Mλf (g) =
eλ(g)f (g). As a second consequence of Theorem 2 we obtain the following.
Corollary 4. Conditions 1) - 4) of Theorem 2 are equivalent to
5) The system of modulates ΨΛ⊥ = {Mλψ : λ ∈ Λ⊥} is a frame for L2(Ω),
with frame bounds 0 < AQΛ ≤ BQΛ < ∞, for any ψ ∈ L2(Ω) satisfying
0 < A ≤ ess inf ψ2 ≤ ess sup ψ2 ≤ B < ∞.
The novelty of this paper is that relates subtilings with frames of exponen-
tials. Moreover, the proof of 2) ⇒ 3) in Theorem 2 is shown by using the bracket
map of a system of translates (Corollary 8) that have been introduced to study
properties of translation invariant spaces (see [7] and the references therein.)
The setting of LCA groups allows to prove simultaneously results for a large
variety of groups, namely Rn, Zd, Tk, and all finite groups F with discrete topol-
ogy, as well as the so called elementary LCA groups G = Rn × Zd × Tk × F .
Observe that knowing Theorem 2 and Corollary 3 for every factor group does
not immediately provide the corresponding results for G.
The motivation for this paper comes from the problem of studying the re-
lationship between spectrum sets and tiling pairs, whose roots dates back to a
1974 paper of B. Fuglede ([6]). There he proved that a set E ⊂ Rd, d ≥ 1, of
positive Lebesgue measure, tiles Rd by translations with a lattice Λ if and only
if L2(E) has an orthogonal basis of exponentials indexed by the annihilator of
Λ. A more general statement in Rd, which says that if E ⊂ Rd, d ≥ 1, has
positive Lebesgue measure, then L2(E) has an orthogonal basis of exponentials
(not necessary indexed by a lattice) if and only if E tiles Rd by translations,
has been known as the Fuglede Conjecture.
A variety of results were proved establishing connections between tiling and
orthogonal exponential bases. See, for example, [17], [11], [16], [12] and [13]. In
2001, I. Laba proved that the Fuglede conjecture is true for the union of two
In 2003, A. Iosevich, N. Katz and T. Tao ([9])
intervals in the plane ([15]).
proved that the Fuglede conjecture holds for convex planar domains.
It was
3
also proved in [11] and [17] that Λ tiles Rd by the unit cube Qd if and only if Λ
is a spectral set for Qd. In 2004, T. Tao ([22]) disproved the Fuglede Conjecture
in dimension d = 5 and larger, by exhibiting a spectral set in R5 which does
not tile the space by translations. In [14], M. Kolountzakis and M. Matolcsi
also disproved the reverse implication of the Fuglede Conjecture for dimensions
d = 4 and higher. In [4] and [3], the dimension of counter-examples was further
reduced.
In fact, B. Farkas, M. Matolcsi and P. Mora show in [3] that the
Fuglede conjecture is false in R3. The general feeling in the field is that sooner
or later the counter-examples of both implications will cover all dimensions.
However, in [10] the authors showed that the Fuglede Conjecture holds in two-
dimensional vector spaces over prime fields. Then, in [2] the authors prove that
tiling implies spectral in Z3
p, p prime, and Fuglede conjecture is true for Z3
2
and Z3
3. Very recently, important developments in LCA groups, with crucial
implications on sampling theory, have been developed by E. Agora, J. Antezana
and C. Cabrelli in [1], where the authors could obtain a full characterization of
Riesz bases for multi-tiling sets in LCA groups.
Acknowledgements: D. Barbieri was supported by a Marie Curie Intra Euro-
pean Fellowship (626055) within the 7th European Community Framework Pro-
gramme. D. Barbieri and E. Hern´andez were supported by Grant MTM2013-
40945-P (Ministerio de Econom´ıa y Competitividad, Spain). A. Mayeli was
supported by PSC-CUNY-TRADB-45-446, and by the Postgraduate Program
of Excellence in Mathematics at Universidad Aut´onoma de Madrid from June
19 to July 17, 2014, when this paper was started. The authors wish to thank
Alex Iosevich for several interesting conversations regarding this paper and his
expository paper on the Fuglede conjecture for lattices [8].
2 Notations and Preliminaries
Let Λ be a lattice in an LCA group G. Denote by QΛ ⊂ G a measurable cross
section of G/Λ. By definition, a cross section is a set of representatives of all
cosets in G/Λ, so that the intersection of QΛ with any coset g + Λ has only one
element. The existence of a Borel measurable cross section is guaranteed by [5,
Theorem 1]. Moreover, it is evident that (QΛ, Λ) is a tiling pair for G, while
any tiling set Ω differs from a cross section at most for a zero measure set.
Let d g be a normalized Haar measure for G/Λ. Then the relation between
Haar measure on G and Haar measure for G/Λ is given by Weil's formula: for
g ∈ G/Λ
any function f ∈ L1(G), the periodization map Φ( g) =Pλ∈Λ f (g +λ),
is well defined almost everywhere in G/Λ, belongs to L1(G/Λ), and
ZG
f (g)dg = QΛZG/ΛXλ∈Λ
f (g + λ)d g.
(4)
This formula is a special case of [20, Theorem 3.4.6]. The constant QΛ, called
the lattice size, appears in (4) because G/Λ is equipped with the normalized
Haar measure d g.
4
Definition 5 (Dual integrable representations ([7])). Let G be an LCA group,
and let π be a unitary representation of G on a Hilbert space H. We say π is
dual integrable if there exists a sesquilinear map [·, ·]π : H × H → L1(bG), called
bracket map for π, such that
hφ, π(g)ψiH =Z bG
[φ, ψ]π(χ)e−g(χ)dχ ∀ g ∈ G , ∀ φ, ψ ∈ H.
Example 6. Let Λ be a lattice in an LCA group G. For any λ ∈ Λ, define
us denote with QΛ⊥ a cross section for the annihilator lattice Λ⊥. Thus, by
Plancherel formula (2) and Weil's formula (4) we have
Tλφ(g) = φ(g − λ) on φ ∈ L2(G) and Mλh(χ) = eλ(χ)h(χ) on h ∈ L2(bG). Let
hφ, TλψiL2(G) = hFG(φ), MλFG(ψ)iL2( bG) =Z bG
FG(φ)(χ)FG(ψ)(χ)e−λ(χ)dχ
= QΛ⊥Z bG/Λ⊥ Xλ∈Λ⊥
= QΛ⊥Z bG/Λ⊥ Xλ∈Λ⊥
FG(φ)( χ + λ)FG(ψ)( χ + λ)e−λ( χ + λ)d χ
FG(φ)( χ + λ)FG(ψ)( χ + λ)e−λ( χ)d χ.
Since FG(φ)FG(ψ) ∈ L1(bG), we have that
[φ, ψ]T ( χ) := QΛ⊥ Xλ∈Λ⊥
χ ∈ bG/Λ⊥
defines a sesquilinear map [·, ·]T : L2(G) × L2(G) → L1(bG/Λ⊥), so T is a dual
integrable representation of Λ on H = L2(G).
FG(φ)( χ + λ)FG(ψ)( χ + λ)
a.e.
A relevant application of dual integrable representations is the possibility to
characterize bases of unitary orbits in terms of their associated bracket maps.
The following result has been proved in [7, Proposition 5.1].
Theorem 7. Let G be a countable abelian group, let π be a dual integrable
representation of G on a Hilbert space H, and let φ ∈ H. The system {π(g)φ :
g ∈ G} is orthonormal in H if and only if [φ, φ]π(χ) = 1 for almost every χ ∈ bG.
As a consequence of Theorem 7 and Example 6, and of the basic fact
QΛQΛ⊥ = 1 (for completeness, we have provided a proof in the appendix),
we have the following result.
Corollary 8. Let T and Λ be as in Example 6, and let φ ∈ L2(G). Then the
system of translates {Tλφ : λ ∈ Λ} is an orthonormal system in L2(G) if and
only if
Xλ∈Λ⊥
FG(φ)(χ + λ)2 = QΛ a.e. χ ∈ bG .
5
3 Proof of Theorem 2
In this section we shall prove Theorem 2 and its corollaries.
Proof of Theorem 2.
1) ⇒ 4) It is well-known ([21]) that, for any cross section QΛ, the exponential
set EΛ⊥ is an orthogonal basis for L2(QΛ). Thus, for all f ∈ L2(QΛ),
Xλ∈Λ⊥
hf,
1
pQΛ
eλiL2(QΛ)2 = kf k2
L2(QΛ) .
(5)
Since condition (1) says that Ω is contained in some cross section QΛ, then the
previous identity still holds for all f ∈ L2(Ω). Hence EΛ⊥ is a tight frame for
L2(Ω) with constant QΛ.
4) ⇒ 1) Suppose, by contradiction, that Ω is not a subtiling set. Then we
claim that for all cross section QΛ there exist λ1, λ2 ∈ Λ, λ2 6= 0, such that
(QΛ + λ1) ∩ Ω ∩ (Ω + λ2) > 0.
(6)
If this is true, then let Ω1 = (QΛ + λ1) ∩ Ω ∩ (Ω + λ2), and Ω2 = Ω1 − λ2.
Both are subsets of Ω with positive measure and, since λ2 6= 0, they are disjoint
because Ω1 ⊂ QΛ + λ1 and Ω2 ⊂ Qλ + λ1 − λ2. Therefore, the function
f = 1Ω1 − 1Ω2
is nonzero and belongs to L2(Ω). Then, for all λ ∈ Λ⊥ we have
hf, eλiL2(Ω) =ZΩ1
eλ(g)dg −ZΩ2
eλ(g)dg =ZΩ1(cid:0)eλ(g) − eλ(g − λ2)(cid:1)dg = 0.
This implies that the system EΛ⊥ can not be a frame for L2(Ω).
In order to prove (6), let us proceed by contradiction and suppose that for
all λ ∈ Λ and all λ∗ ∈ Λ, λ∗ 6= 0 we have
Now take λ′ ∈ Λ, λ′ 6= 0. By definition of cross section, we have
(QΛ + λ) ∩ Ω ∩ (Ω + λ∗) = 0.
Ω ∩ (Ω + λ′) = Gλ∈Λ
(QΛ + λ) ∩ Ω ∩ (Ω + λ′)
which implies that Ω ∩ (Ω + λ′) = 0. Hence, Ω would be a subtiling set of G
by Λ, which is a contradiction.
1) ⇒ 2) Since (5) holds, we can obtain 2) by choosing f = eχ 1Ω.
2) ⇒ 3) This follows as an application of Corollary 8.
3) ⇒ 1) By orthogonality, we have that for all λ ∈ Λ, λ 6= 0
0 = h1Ω, 1Ω(· − λ)iL2(G) = Ω ∩ (Ω + λ)
so Ω is sub-tiling.
6
Proof of Corollary 3. If (Ω, Λ) is a tiling pair then it is well-known that EΛ⊥ is
an orthogonal basis for L2(Ω). To prove the converse, assume by contradiction
that Ω is not tiling. Then one of the following cases holds
i. Ω is a strictly sub-tiling set, i.e. there exists a cross section QΛ such that
Ω ⊂ QΛ and QΛ \ Ω > 0.
ii. Ω is not a sub-tiling set, so that (6) holds.
For case i., observe that the assumption of EΛ⊥ being an orthogonal basis for
L2(Ω) implies
Xλ∈Λ⊥
hf,
1
pΩ
eλiL2(Ω)2 = kf k2
L2(Ω) ∀ f ∈ L2(Ω).
On the other hand, since EΛ⊥ is an orthogonal basis for L2(QΛ), we have
Xλ∈Λ⊥
hf 1Ω,
1
pQΛ
eλiL2(QΛ)2 = kf 1Ωk2
L2(QΛ) ∀ f ∈ L2(Ω)
so that Ω = QΛ, which contradicts i.
For case ii., in Theorem 2 we already proved that EΛ⊥ can not even be a
frame.
Proof of Corollary 4. Assume 4) holds, i.e. that EΛ⊥ is a tight frame for L2(Ω)
with constant QΛ. Then
Xλ∈Λ⊥
hf, MλψiL2(Ω)2 = Xλ∈Λ⊥
hf ψ, eλiL2(Ω)2 = QΛ kf ψk2
L2(Ω) ∀ f ∈ L2(Ω) .
Since Akf k2
5) holds. Then, since A > 0, for all f ∈ L2(Ω) we can write
L2(Ω) ≤ Bkf k2
L2(Ω) ≤ kf ψk2
L2(Ω), this proves 5). Conversely, assume
Xλ∈Λ⊥
hf, eλiL2(Ω)2 = Xλ∈Λ⊥
hf /ψ, MλψiL2(Ω)2 ,
so that, by the hypotheses on ψ, we get
A
B
QΛ kf k2
L2(Ω) ≤ Xλ∈Λ⊥
hf, eλiL2(Ω)2 ≤
B
A
QΛ kf k2
L2(Ω) ∀ f ∈ L2(Ω).
Thus EΛ⊥ is a frame, hence proving 4). Observe that, by Theorem 2, this implies
that EΛ⊥ is a tight frame with constant QΛ, hence improving the inequalities
above.
7
4 Comments on related work
The statement of Theorem 2 relating sutilings pairs (Ω, Λ) for a lattice Λ and
frames of exponentials for L2(Ω) is new. However, the proof has similarities with
existing proofs of the similar statement for tiling sets by lattices and orthonormal
bases of exponentials.
The result in Corollary 3 is proved by B. Fuglede ([6], Lemma 6) for G = Rn
and by S. Pedersen ([18], Theorem 3.6) for LCA groups.
In both papers, [6] and [18], as well as in the present work, the implication
1) ⇒ 4) is done in the same way by observing that EΛ⊥ is an orthogonal basis
of exponentials of L2(QΛ). As for the implication 4) ⇒ 1) both papers give a
direct proof of the fact that Ω ∩ (Ω + λ) have measure zero for all λ ∈ Λ and
(Ω + λ) has measure zero. In the present paper we give
that also the set G \ [λ∈Λ
an argument by contradiction assuming that 1) does not hold and exhibiting a
non-zero function in L2(Ω) which is perpendicular to all eλ, λ ∈ Λ⊥.
The manuscript [8] by A. Iosevich gives a proof of the equivalence of 1) and
2) in Theorem 2 for a tiling set Ω ∈ Rn by a lattice Λ. It is then stated without
proof that 2) and 4) are equivalent by a density argument. The proof of 1) ⇔ 2)
in [8] goes as follows. Consider the functions
f (x) ≡ Xλ∈Λ
1Ω(x + λ) ,
x ∈ Rn
and
H(ξ) ≡ Xλ∈Λ⊥
FRn (1Ω)(ξ + λ)2 .
Using Fourier Analysis it can be shown that, as a periodic function in L2(QΛ⊥ ),
the Fourier coefficients of H are
Ω ∩ (Ω + λ)
QΛ⊥
,
λ ∈ Λ ,
bH(λ) =
and the Fourier coefficients of f , as a periodic function in L2(QΛ), are
FRn(1Ω)(λ)
QΛ
,
λ ∈ Λ .
bf (λ) =
(7)
(8)
Assuming that 1) of Theorem 2 holds for a tiling set Ω, equation (7) shows,
using the Fourier inversion theorem, that H is constant with value Ω QΛ a.e.,
since QΛ QΛ⊥ = 1. This shows 2) of Theorem 2. Conversely, assuming 2)
of Theorem 2 holds, equation (7) shows that Ω is a subtiling set of Rn by Λ.
Equation (8) is then used to show that f (x) = 1 a. e., which shows 1).
A proof along the lines described above can be designed for LCA groups and
subtiles. In the proof given in the present work, we have proved the equivalence
of 1), 2) and 3) in Theorem 2 by using the notion of bracket map (see [7] and the
references therein) and the characterization of frame sequences of translates of a
single function along lattices stated in Theorem 7 and Corollary 8. As in [8] our
paper also uses Fourier Analysis to compute Fourier coefficients (see Example
6).
8
A Appendix
We provide here a self-contained proof of the following basic result.
transform be an isometry, i.e.
Lemma 9. Let G be an LCA group, let bG be its dual group, and let the Haar
measures dg on G and dχ on bG be chosen in such a way that the group Fourier
subgroup of G, and denote with QΛ ⊂ G a cross section for G/Λ, let Λ⊥ ⊂ bG
be the dual lattice of Λ, and denote with QΛ⊥ ⊂ bG a cross section for bG/Λ⊥.
such that (2) holds. Let Λ ⊂ G be a lattice
Then
QΛQΛ⊥ = 1
i.e. the product of the size of the lattice times the size of the dual lattice, com-
puted with respect to Haar measures satisfying (2), is 1.
Proof. Observe first that, by (2), we have
QΛ =ZG
1QΛ(g)2dg =Z bG
FG(1QΛ )(χ)2dχ
=ZQΛ⊥ Xλ∈Λ⊥
FG(1QΛ )(χ + λ)2dχ
(9)
where the last identity is due to the fact that (QΛ⊥ , Λ⊥) is a tiling pair for bG by
definition of cross-section. We can thus apply the same argument used to prove
point 2) of Theorem 2: since {eλ : λ ∈ Λ⊥} is an orthogonal basis of L2(QΛ),
then it satisfies (5), so by choosing f = eχ1QΛ we get
Xλ∈Λ⊥
FG(1QΛ)(χ + λ)2 = QΛ2
a.e. χ ∈ bG.
The claim then follows by inserting this identity in (9).
References
[1] E. Agora, J. Antezana, C. Cabrelli, Multi-tiling sets, Riesz bases, and sam-
pling near the critical density in LCA groups. Adv. Math. 285 (2015) 454-
477.
[2] C. Aten et al, Tiling sets and spectral sets over finite fields. Preprint,
arxiv.org/abs/1509.01090.
[3] B. Farkas, M. Matolcsi and P. M´ora, On Fuglede's conjecture and the exis-
tence of universal spectra. J. Fourier Anal. Appl. 12 (2006), 483-494.
[4] B. Farkas and S. Revesz, Tiles with no spectra in dimension 4. Math. Scand.
98 (2006), 44-52.
9
[5] J. Feldman and F. P. Greenleaf, Existence of Borel transversals in groups.
Pacific J. Math. 25 (1968), 455-461.
[6] B. Fuglede: Commuting self-adjoint partial differential operators and a
group theoretic problem. J. Funct. Anal. 16 (1974), 101-121.
[7] E. Hern´andez, H. Sikic, G. Weiss, E. Wilson, Cyclic subspaces for unitary
representation of LCA groups: generalized Zak transforms. Colloq. Math.
118 (2010), 313 - 332.
[8] A.
Iosevich, Fuglede Conjecture for Lattices. Preprint available at
www.math.rochester.edu/people/faculty/iosevich/expository/FugledeLattice.pdf
[9] A. Iosevich, N. Katz and T. Tao, The Fuglede spectral conjecture holds for
convex planar domains. Math. Res. Lett. 10 (2003), no. 5-6, 559-569.
[10] A. Iosevich, A. Mayeli, J. Pakianathan, The Fuglede Conjecture holds in
Zp × Zp. To appear on Anal. PDE.
[11] A. Iosevich and S. Pedersen, Spectral and tiling properties of the unit cube.
Internat. Math. Res. Notices (1998), no. 16, 819-828.
[12] S. Konyagin and I. Laba, Spectra of certain types of polynomials and tiling
of integers with translates of finite sets. J. Number Theory 103 (2003), no.
2, 267-280.
[13] M. Kolountzakis and I. Laba, Tiling and spectral properties of near-cubic
domains. Studia Math. 160 (2004), no. 3, 287-299.
[14] M. Kolountzakis and M. Matolcsi, Tiles with no spectra. Forum Math. 18
(2006), 519-528.
[15] I. Laba, Fuglede's conjecture for a union of two intervals. Proc. Amer.
Math. Soc. 129 (2001), 2965-2972.
[16] I. Laba, The spectral set conjecture and multiplicative properties of roots of
polynomials. J. London Math. Soc. 65 (2002), 661-671.
[17] J. Lagarias, J. Reed and Y. Wang, Orthonormal bases of exponentials for
the n-cube. Duke Math. J. 103 (2000), 25-37.
[18] S. Pedersen, Spectral Theory of Commuting Self-Adjoint Partial Differential
Operators. Journal of Functional Analysis 73 (1987), 122-134 .
[19] L.S. Pontryagin, Topological Groups. Princeton Univ. Press (1946)
[20] H. Reiter, J.D. Stegeman, Classical Harmonic Analysis on Locally Compact
groups. Clarendon Press, Oxford (2000).
[21] W. Rudin, Fourier Analysis on Groups. John Wiley & Sons, 1990.
[22] T. Tao, Fuglede's conjecture is false in 5 and higher dimensions. Math. Res.
Lett. 11 (2004), 251-258.
10
|
1704.01768 | 1 | 1704 | 2017-04-06T10:05:35 | $\Gamma$-flatness and Bishop-Phelps-Bollob\'as type theorems for operators | [
"math.FA"
] | The Bishop-Phelps-Bollob\'{a}s property deals with simultaneous approximation of an operator $T$ and a vector $x$ at which $T$ nearly attains its norm by an operator $T_0$ and a vector $x_0$, respectively, such that $T_0$ attains its norm at $x_0$. In this note we extend the already known results about {the} Bishop-Phelps-Bollob\'{a}s property for Asplund operators to a wider class of Banach spaces and to a wider class of operators. Instead of proving a BPB-type theorem for each space separately we isolate two main notions: $\Gamma$-flat operators and Banach spaces with ACK$_\rho$ structure. In particular, we prove a general BPB-type theorem for $\Gamma$-flat operators acting to a space with ACK$_\rho$ structure and show that uniform algebras and spaces with the property $\beta$ have ACK$_\rho$ structure. We also study the stability of the ACK$_\rho$ structure under some natural Banach space theory operations. As a consequence, we discover many new examples of spaces $Y$ such that the Bishop-Phelps-Bollob\'{a}s property for Asplund operators is valid for all pairs of the form ($X,Y$). | math.FA | math | Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE
THEOREMS FOR OPERATORS
BERNARDO CASCALES, ANTONIO J. GUIRAO, VLADIMIR KADETS,
AND MARIIA SOLOVIOVA
ABSTRACT. The Bishop -- Phelps -- Bollob´as property deals with simulta-
neous approximation of an operator T and a vector x at which T nearly
attains its norm by an operator T0 and a vector x0, respectively, such that
T0 attains its norm at x0. In this note we extend the already known results
about the Bishop -- Phelps -- Bollob´as property for Asplund operators to a
wider class of Banach spaces and to a wider class of operators. Instead
of proving a BPB-type theorem for each space separately we isolate two
main notions: Γ-flat operators and Banach spaces with ACKρ structure.
In particular, we prove a general BPB-type theorem for Γ-flat operators
acting to a space with ACKρ structure and show that uniform algebras
and spaces with the property β have ACKρ structure. We also study the
stability of the ACKρ structure under some natural Banach space the-
ory operations. As a consequence, we discover many new examples of
spaces Y such that the Bishop -- Phelps -- Bollob´as property for Asplund
operators is valid for all pairs of the form (X, Y ).
1. INTRODUCTION
In this paper X, Y are Banach spaces (real or complex), K stands for the
field of scalars R or C, L(X, Y ) is the space of all bounded linear operators
T : X → Y , L(X) = L(X, X), BX and SX denote the closed unit ball and
the unit sphere of X, respectively and acoA stands for the absolute convex
hull of the set A.
According to [1], a pair (X, Y ) has the Bishop -- Phelps -- Bollob´as property
(BPB property) for operators if for every ε > 0 there exists δ(ε) > 0 such
that for every operator T ∈ L(X, Y ) of norm 1, if x0 ∈ SX is such that
(cid:107)T (x0)(cid:107) > 1 − δ(ε), then there exist u0 ∈ SX and S ∈ SL(X,Y ) satisfying
(cid:107)S(u0)(cid:107) = 1, (cid:107)x0 − u0(cid:107) < ε, and (cid:107)T − S(cid:107) < ε.
7
1
0
2
r
p
A
6
]
.
A
F
h
t
a
m
[
1
v
8
6
7
1
0
.
4
0
7
1
:
v
i
X
r
a
Date: VERSION: 31 December, 2016.
2010 Mathematics Subject Classification. Primary: 46B20, 46E25 Secondary: 47B07,
Key words and phrases. Bishop -- Phelps -- Bollob´as, Asplund operators, norm attaining,
47B48.
uniform algebra.
The research of first, second and third authors was partially supported by MINECO
grant MTM2014-57838-C2-1-P and Fundaci´on S´eneca, Regi´on de Murcia grant
19368/PI/14. The research of the third author is done in frames of Ukranian Ministry
of Science and Education Research Program 0115U000481. The research of fourth author
has been partially performed during her stay in University of Murcia in frames of Erasmus+
program.
1
2
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
If an analogous definition is valid for operators T , S from a subspace
I ⊂ L(X, Y ), then we say that (X, Y ) has the Bishop -- Phelps -- Bollob´as
property for operators from I.
With this terminology, the original Bishop -- Phelps -- Bollob´as theorem [8]
says that for every X, the pair (X, K) has the BPB property for operators.
Also, thanks to Acosta, Aron, Garc´ıa, and Maestre [1, Theorem 2.2], if Y
has the Lindenstrauss' property β (Definition 4.8), then for every Banach
space X the pair (X, Y ) has the Bishop -- Phelps -- Bollob´as property for op-
erators.
In 2011 Aron, Cascales, and Kozhushkina [4, Theorem 2.4] showed that
for every X and every compact Hausdorff space K the pair (X, C(K)) has
the BPB property for Asplund operators (Definition 2.2). In 2013 Cascales,
Guirao and Kadets [9] extended this result to uniform algebras A ⊂ C(K).
The exact statement of the last result is given below.
Theorem 1.1 ([9, Theorem 3.6]). Let A ⊂ C(K) be a uniform algebra
and T : X → A be an Asplund operator with (cid:107)T(cid:107) = 1. Suppose that
2 . Then there exist
0 < ε <
u0 ∈ SX and an Asplund operator S ∈ SL(X,A) satisfying that:
2 and x0 ∈ SX are such that (cid:107)T x0(cid:107) > 1 − ε2
√
(cid:107)Su0(cid:107) = 1,
(cid:107)x0 − u0(cid:107) ≤ ε
and
(cid:107)T − S(cid:107) < 2ε.
In the same vein, Acosta, Becerra Guerrero, Garc´ıa, Kim, and Maestre [2]
generalized [4, Theorem 2.4] to some spaces of continuous vector-valued
functions (see Theorem 4.13 below).
The aim of this paper is to extend all these results to a wider class of Ba-
nach spaces and to a wider class of operators. The main difference of our
approach is that instead of proving a Bishop -- Phelps -- Bollob´as kind theorem
for each space separately (and thus repeating essential parts of the proof
many times), we introduce a new Banach space property (called ACKρ
structure) which extracts all the useful technicalities for the BPB type of
approximation. We prove a general Bishop -- Phelps -- Bollob´as type theorem
for Γ-flat operators (see Definition 2.8) acting to a space with ACKρ struc-
ture and show that uniform algebras and spaces with the property β have
ACKρ structure. After that, we study the stability of the ACKρ structure
under some natural Banach space theory operations which as a consequence
gives us a wide collection of examples of pairs (X, Y ) possessing the BPB
property for Asplund operators.
The structure of the paper is as follows. In section 2 we collect the nec-
essary definitions (in particular that of Asplund operators and of Γ-flat op-
erators) and prove an important Basic Lemma. In section 3 we introduce
the central concept of ACKρ structure and prove a general BPB type the-
orem for this class of Banach spaces. Finally, in section 4 we perform the
announced study of spaces with ACKρ structure which, on the one hand,
gives a unified proof of several results from [1, 2, 4] and [9], and on the
other hand, leads to new BPB type theorems in concrete spaces.
For the non-defined notions used through this article, we refer to [12].
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
3
2. Γ-FLAT OPERATORS AND THE BASIC LEMMA
Let (B, τ ) be a topological space, ρ be a metric on B (possibly, not re-
lated with τ). B is said to be fragmented by ρ, if for every non-empty subset
A ⊂ B and for every ε > 0 there exists a τ-open U such that U ∩ A (cid:54)= ∅
and diam(U∩A) < ε. Some important examples of fragmented topological
spaces come from Banach space theory. For instance, every weakly com-
pact subset of a Banach space is fragmented by the norm (i.e., by the metric
ρ(x, y) = (cid:107)x − y(cid:107)), see [16].
A Banach space X is called an Asplund space if, whenever f is a convex
continuous function defined on an open subset U of X, the set of all points
of U where f is Fr´echet differentiable is a dense Gδ-subset of U. This def-
inition is due to Asplund [3] under the name strong differentiability space.
This concept has multiple characterizations via topology or measure theory,
as in the following:
Theorem 2.1 ([17, 21, 22]). Let X be a Banach space. Then the following
conditions are equivalent:
(i) X is an Asplund space;
(ii) every w∗-compact subset of (X, w∗) is fragmented by the norm;
(iii) each separable subspace of X has separable dual;
(iv) X∗ has the Radon-Nikod´ym property.
According to the above, every reflexive space and every separable space
whose dual is separable is an Asplund space. Classical example of Asplund
spaces are Lp and (cid:96)p with 1 < p < ∞, and also c0; examples of spaces
that are not Asplund are C[0, 1], (cid:96)1, (cid:96)∞, L1[0, 1] and L∞[0, 1], see for in-
stance [11].
Definition 2.2 ([23]). An operator T ∈ L(X, Y ) is said to be an Asplund
operator if it factors through an Asplund space, i.e., there exist an Asplund
Banach space Z and operators T1 ∈ L(X, Z), T2 ∈ L(Z, Y ) such that T =
T2 ◦ T1.
Compact and weakly compact operators are Asplund operators (every
weakly compact operator factorizes through a reflexive space).
Theorem 2.1 yields the following result:
Remark 2.3 ([23]). If T is an Asplund operator, then its adjoint T ∗ sends
the unit ball of Y ∗ into a w∗-compact subset of (X, w∗) that is norm frag-
mented.
Definition 2.4. Let Y be a Banach space. Y is said to have the Bishop --
Phelps -- Bollob´as property for Asplund operators (A-BPBp for short) if for
every ε > 0 there exists δ(ε) > 0, such that for every Banach space X and
every Asplund operator T ∈ SL(X,Y ), if x0 ∈ SX is such that (cid:107)T (x0)(cid:107) >
1 − δ(ε), then there exist u0 ∈ SX and S ∈ SL(X,Y ) satisfying
(cid:107)S(u0)(cid:107) = 1,(cid:107)x0 − u0(cid:107) < ε and (cid:107)T − S(cid:107) < ε.
4
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
Definition 2.5 ([13]). Let A and B be topological spaces. A function
f : A → B is said to be quasi-continuous, if for every non-empty open sub-
set U ⊂ A, every z ∈ U and every neighborhood V of f (z) there exists a
non-empty open subset W ⊂ U such that f (W ) ⊂ V .
Let us introduce some new terminology. Note that a similar concept of
fragmentability of maps was introduced in [14].
Definition 2.6. Let A be a topological space and (M, d) be a metric space.
A function f : A → M is said to be openly fragmented, if for every non-
empty open subset U ⊂ A and every ε > 0 there exists a non-empty open
subset V ⊂ U with d-diam(f (V )) < ε.
Every continuous or quasi-continuous function f : A → M is openly
fragmented. In particular, if A is a discrete topological space then every
f : A → M is openly fragmented. For every metric space M, every left-
continuous f : [0, 1] → M and every right-continuous function f : [0, 1] →
M are openly fragmented. Every f : A → M with a dense set of continuity
points is openly fragmented. Every separately continuous function of two
variables f : [0, 1] × [0, 1] → M is quasi-continuous [6] and, consequently,
openly fragmented. Some other easy but useful examples are given in the
following theorem:
Theorem 2.7. Let A, B be topological spaces, ρ be a metric on B (possibly,
not related with the original topology), and f : A → B be a function.
(i) If B is fragmented by ρ, and f is continuous in the original topolo-
gies, then f : A → (B, ρ) is openly fragmented.
(ii) If A is fragmented by some metric ρ1 and f : (A, ρ1) → (B, ρ) is
uniformly continuous, then f : A → (B, ρ) is openly fragmented.
Let, moreover, (B,(cid:107)·(cid:107)) be a Banach space. Then
(iii) If f, g : A → (B,(cid:107)·(cid:107)) are openly fragmented then f + g : A →
(iv) If f : A → (B,(cid:107)·(cid:107)) and g : A → K are openly fragmented then
(B,(cid:107)·(cid:107)) is openly fragmented.
gf : A → (B,(cid:107)·(cid:107)) is openly fragmented.
The statements (ii), (iii) and (iv) are routine.
Proof. (i) For a given non-empty open subset U ⊂ A consider f (U ) ⊂ B.
By ρ-fragmentability of B, for every ε > 0 there exits an open subset W
of B with f (U ) ∩ W (cid:54)= ∅ and diam(f (U ) ∩ W ) < ε. By continuity of f
the set f−1(W ) is open and V := f−1(W ) ∩ U will be the non-empty open
subset V ⊂ U we need.
(cid:3)
Definition 2.8. Let X, Y be Banach spaces and Γ ⊂ Y ∗. An operator
T ∈ L(X, Y ) is said to be Γ-flat, if T ∗Γ : (Γ, w∗) → (X∗,(cid:107)·(cid:107)X∗) is openly
fragmented. In other words, for every w∗-open subset U ⊂ Y ∗ with U∩Γ (cid:54)=
∅ and every ε > 0 there exists a w∗-open subset V ⊂ U with V ∩ Γ (cid:54)= ∅
such that diam(T ∗(V ∩ Γ)) < ε. The set of all Γ-flat operators in L(X, Y )
will be denoted by FlΓ(X, Y ).
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
5
Statements (iii) and (iv) of the previous theorem imply that FlΓ(X, Y ) is
a linear subspace of L(X, Y ). Let us list some examples of Γ-flat operators.
Example A. Every Asplund operator T ∈ L(X, Y ) is Γ-flat for every Γ ⊂
BY ∗. This follows from Remark 2.3 and Theorem 2.7, (i).
Example B. If (Γ, w∗) ⊂ Y ∗ is norm fragmented, then every bounded op-
erator in L(X, Y ) is Γ-flat (Theorem 2.7, (ii)). In particular, we have the
next concrete example.
Example C. If (Γ, w∗) ⊂ Y ∗ is discrete, then every operator T ∈ L(X, Y )
is Γ-flat.
ε
The notion of Γ-flat generalizes the property of Asplund operators that
allowed to prove [4, Lemma 2.3]. The immediate generalization of that
lemma is the following result:
Lemma 2.9 (Basic Lemma). Let X, Y be Banach spaces, Γ ⊂ BY ∗ be
a 1-norming set, T ∈ FlΓ(X, Y ) be a Γ-flat operator with (cid:107)T(cid:107) = 1 ,
0 < ε < 2/3, and x0 ∈ SX be such that (cid:107)T x0(cid:107) > 1 − ε. Then for every
r > 0 and for every k ∈ [
(i) a w∗-open set Ur ⊂ Y ∗ with Ur ∩ Γ (cid:54)= ∅, and
(ii) points x∗
(cid:107)x0 − ur(cid:107) ≤ ε
k
2(1−ε), 1) there exist:
r ∈ SX∗ and ur ∈ SX with x∗
and (cid:107)T ∗z∗ − x∗
r(cid:107) ≤ r+2k for every z∗ ∈ Ur∩Γ. (2.1)
The proof of this fact is a modification of that of [4, Lemma 2.3]. First,
r(ur) = 1 such that
we use the following fact:
Proposition 2.10 ([19, Corollary 2.2]). Let X be a real Banach space, z∗ ∈
SX∗, z ∈ SX, η > 0 and z∗(z) ≥ 1 − η. Then for every k ∈ (0, 1) there
exist y∗ ∈ SX∗ and u ∈ SX such that
y∗(u) = 1,
(cid:107)z − u(cid:107) ≤ η
k
,
(cid:107)z∗ − y∗(cid:107) ≤ 2k.
In the next proposition, we relax the condition z∗ ∈ SX allowing (cid:107)z∗(cid:107) to
be smaller than 1. Note that x∗ plays the role of z∗.
Proposition 2.11. Let X be a Banach space, ε ∈ (0, 2/3), x ∈ SX, x∗ ∈
2(1−ε) , 1) there exist y∗ ∈
BX∗ and x∗(x) ≥ 1 − ε. Then, for every k ∈ [
SX∗ and u ∈ SX such that
ε
y∗(u) = 1,
(cid:107)x − u(cid:107) ≤ ε
k
,
(cid:107)x∗ − y∗(cid:107) ≤ 2k.
Proof. Without loss of generality we can assume that x∗(x) ≥ 1 − ε. Then
(cid:107)x∗(cid:107) ≥ 1 − ε. Set z∗ := x∗/(cid:107)x∗(cid:107) , z := x. Then z∗(z) ≥ 1 − η for
η = 1 − (1 − ε)(cid:107)x∗(cid:107)−1 ∈ [0, ε]. If η = 0, then z∗(z) = 1, so we can take
y∗ = z∗ and u = x, which satisfy the inequalities we want. So we may
6
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
assume that 0 < η ≤ ε. Set k0 := kη
2.10, there exist y∗ ∈ SX∗ and u ∈ SX such that
ε ∈ (0, 1). So, according to Proposition
y∗(u) = 1,
(cid:107)z − u(cid:107) ≤ η
k0
,
Therefore, (cid:107)x − u(cid:107) ≤ η/k0 = ε/k. Also, we have
(cid:107)x∗ − y∗(cid:107) ≤ (cid:107)x∗ − z∗(cid:107) + (cid:107)z∗ − y∗(cid:107) ≤
(cid:107)z∗ − y∗(cid:107) ≤ 2k0.
(cid:13)(cid:13)(cid:13)(cid:13) + 2k0
(cid:13)(cid:13)(cid:13)(cid:13)x∗ − x∗
(cid:18)
(cid:107)x∗(cid:107)
1 − 1 − ε
(cid:107)x∗(cid:107)
ε (1 − 1−ε
(cid:19)
.
0,
ε
(cid:19)
(cid:18)
. So, if k ≥ ε
(cid:113) 2k(1−ε)
= 1 − (cid:107)x∗(cid:107) + 2k0 = 1 − (cid:107)x∗(cid:107) +
2k
ε
t ) is increasing when
2(1−ε), we have ψ((cid:107)x∗(cid:107)) ≤ ψ(1) = 2k. In
(cid:3)
0 ∈ Γ such
Observe that the function ψ(t) = 1 − t + 2k
t ∈
this case, we get our conclusion.
Proof of Lemma 2.9. Use that Γ ⊂ BY ∗ is 1-norming and pick y∗
that
0 ∈ U ∩ Γ ⊂ BY ∗.
Set U := {y∗ ∈ Y ∗ : T ∗y∗(x0) > 1− ε}. We have that y∗
Since U is w∗-open in Y ∗ and U ∩ Γ (cid:54)= ∅, according to Definition 2.8, for
every r > 0 there exists a w∗-open subset Ur ⊂ U with Ur ∩ Γ (cid:54)= ∅ such
that diam(T ∗(Ur ∩ Γ)) < r.
1(x0) >
Fix some y∗
1 − ε which, by applying Proposition 2.11 to any
2(1−ε) ≤ k < 1, gives
r ∈ SX∗ and ur ∈ SX with x∗
x∗
(cid:107)x0 − ur(cid:107) ≤ ε
k
r(ur) = 1 and such that
1 ∈ Ur ∩ Γ and set x∗
0)(x0) = y∗
0(T x0) > 1 − ε.
1. Then, 1 ≥ (cid:107)x∗
1(cid:107) ≥ x∗
1 − x∗
r(cid:107) ≤ 2k.
1 = T ∗y∗
T ∗(y∗
(cid:107)x∗
and
ε
Finally, let z∗ ∈ Ur ∩ Γ be arbitrary. Then,
(cid:107)T ∗z∗ − x∗
r(cid:107) ≤ (cid:107)T ∗z∗ − x∗
1(cid:107) + (cid:107)x∗
1 − x∗
r(cid:107) ≤ r + 2k,
(cid:3)
which finishes the proof.
3. THE ACK STRUCTURE
In the definition below we extract the structural properties of C(K) and
its uniform subalgebras that were essential in the proof of [9, Th. 3.6]. The
name "ACK structure" comes from the words "Asplund" and "C(K)".
Definition 3.1. Let X be a Banach space and O be a non-emtpy subset of
L(X). We will say that X has O-ACK structure with parameter ρ, for some
ρ ∈ [0, 1) (X ∈ O-ACKρ, for short) whenever there exists a 1-norming set
Γ ⊂ BX∗ such that for every ε > 0 and every non-empty relatively w∗-open
subset U ⊂ Γ there exist a non-empty subset V ⊂ U, vectors x∗
1 ∈ V ,
e ∈ SX and an operator F ∈ O with the following properties:
(I) (cid:107)F e(cid:107) = (cid:107)F(cid:107) = 1;
(II) x∗
1(F e) = 1;
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
7
1 = x∗
1;
then v∗(F e) ≤ ρ for every x∗ ∈ Γ \ V1;
(III) F ∗x∗
(IV) denoting V1 = {x∗ ∈ Γ : (cid:107)F ∗x∗(cid:107)+(1−ε)(cid:107)(IX∗ − F ∗)(x∗)(cid:107) ≤ 1},
(V) dist(F ∗x∗, aco{0, V }) < ε for every x∗ ∈ Γ; and
(VI) v∗(e) − 1 ≤ ε for every v∗ ∈ V .
The Banach space X is said to have simple O-ACK structure (X ∈ O-
ACK) if V1 = Γ. In other words, for X ∈ O-ACK the above definition
holds true with the following modification: the property (IV) becomes
(IV)' (cid:107)F ∗x∗(cid:107) + (1 − ε)(cid:107)(IX∗ − F ∗)(x∗)(cid:107) ≤ 1 for every x∗ ∈ Γ.
In case of O = L(X), we will simply say ACKρ (and simple ACK) struc-
ture.
Remark 3.2. If X belongs to the class ACKρ, then X also belongs to
ACKσ for every σ ∈ [ρ, 1). Moreover, ACK ⊂ ACKρ for every ρ ∈ [0, 1).
Definition 3.3. A linear subspace I ⊂ L(X, Y ) is said to be a Γ-flat ideal,
if all elements of I are Γ-flat operators, I contains all operators of finite
rank, and for every T ∈ I and every F ∈ L(Y ) their composition F ◦ T
belongs to I.
Observe that the subspace of Asplund operators in L(X, Y ) is an example
of Γ-flat ideal. The theorem below motivates the above definition.
Theorem 3.4. Let X be a Banach space, Y ∈ ACKρ , Γ ⊂ Y ∗ be the
corresponding 1-norming set from Defintion 3.1 and T ∈ L(X, Y ) be a Γ-
flat operator with (cid:107)T(cid:107) = 1. Let 0 < ε ≤ 1/2 and let x0 ∈ SX be such that
(cid:107)T x0(cid:107) > 1 − ε. Then there exist u0 ∈ SX and an operator S ∈ SL(X,Y )
with (cid:107)Su0(cid:107) = 1 such that
max{(cid:107)x0 − u0(cid:107) ,(cid:107)T − S(cid:107)} <
√
2
1 − ρ +
2ε
Moreover, if Y ∈ ACK then the estimate can be improved to
√
1 +
2ε
max{(cid:107)x0 − u0(cid:107) ,(cid:107)T − S(cid:107)} <
2ε.
Additionally, S can be chosen from I whenever T belongs to a Γ-flat ideal
I. In particular, every Y ∈ ACKρ (ACK) has the A-BPBp.
Before proving the theorem, we need a preliminary result.
Lemma 3.5. Under the conditions of Theorem 3.4 above, for every k ∈
(ε/(2(1 − ε)), 1) and for every
(cid:18)
√
(cid:19)
.
(cid:18)
(cid:19)
ν > 2k
1 +
2
1 − ρ + 2k
,
there exist u0 ∈ SX and S ∈ SL(X,Y ) satisfying (cid:107)Su0(cid:107) = 1, (cid:107)x0 − u0(cid:107) ≤ ε
and (cid:107)T − S(cid:107) < ν. In the case of Y ∈ ACK the same is true for every
ν > 2k.
If, moreover, T belongs to a Γ-flat ideal I, then S can be chosen from I
k
as well.
8
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
Proof. First, consider the more involved case of Y ∈ ACKρ. Fix r > 0 and
0 < ε(cid:48) < 2/3. Now, we can apply Lemma 2.9 with Y , Γ, r and ε > 0. We
r ∈ SX∗ and
produce a w∗-open set Ur ⊂ Y ∗ with Ur ∩ Γ (cid:54)= ∅, and points x∗
ur ∈ SX with x∗
Since Ur ∩ Γ (cid:54)= ∅, we can apply Definition 3.1 to U = Ur ∩ Γ and ε(cid:48)
1 ∈ V , e ∈ SY , F ∈ L(Y ) and V1 ⊂ Γ
and obtain a non-empty V ⊂ U, y∗
which satisfy properties (I) -- (VI). In particular, for every z∗ ∈ V ⊂ Ur ∩ Γ
according to (2.1) we have
r(ur) = 1 such that (2.1) holds true.
(cid:107)T ∗z∗ − x∗
r(cid:107) ≤ r + 2k.
(3.1)
Define now the linear operator S : X → Y by the formula
r(x)F e + (1 −(cid:101)ε)(IY − F )T x,
S(x) := x∗
where the value of(cid:101)ε ∈ [ε(cid:48), 1) will be specified below in such a way that
(cid:107)S(cid:107) ≤ 1. In order to do this, bearing in mind the fact that Γ is 1-norming,
we can write
(3.2)
(cid:107)S(cid:107) = (cid:107)S∗(cid:107) = sup{(cid:107)S∗y∗(cid:107) : y∗ ∈ Γ} .
So our first goal is to estimate
r + (1 −(cid:101)ε)T ∗(IY ∗ − F ∗)(y∗)(cid:107)
(cid:107)S∗y∗(cid:107) = (cid:107)y∗(F e)x∗
(3.3)
from above for all y∗ ∈ Γ. For y∗ ∈ V1, the sought estimate (cid:107)S∗y∗(cid:107) ≤ 1
follows immediately from the definition of V1 (see property (IV)). So, it
remains to consider the case y∗ ∈ Γ \ V1. Thanks to (V), for every y∗ ∈ Γ,
k=1 λkv∗
k with
(cid:107)F ∗y∗ − v∗(cid:107) < ε(cid:48)
(3.4)
k=1 λk ≤ 1. According to (3.1) we have
there exists an element v∗ =(cid:80)n
k=1 ⊂ V , and(cid:80)n
r − T ∗v∗(cid:107) ≤ n(cid:88)
k}n
r(cid:107) ≤ r + 2k, consequently
λk(cid:107)v∗
n(cid:88)
such that {v∗
(cid:107)T ∗v∗
k − x∗
(cid:107)v∗(e)x∗
k=1
(VI)≤ ε(cid:48) +
k(e)x∗
r − T ∗v∗
k(cid:107)
λk(cid:107)x∗
r − T ∗v∗
k(cid:107) ≤ ε(cid:48) + r + 2k.
(3.5)
k=1
Now, for every y∗ ∈ Γ \ V1
(cid:107)S∗y∗(cid:107) ≤(cid:101)εy∗(F e) + (1 −(cid:101)ε)(cid:107)y∗(F e)x∗
r + T ∗y∗ − T ∗F ∗y∗(cid:107)
(IV)≤ (cid:101)ερ + (1 −(cid:101)ε)(cid:107)T ∗y∗(cid:107) + (1 −(cid:101)ε)(cid:107)(F ∗y∗)(e)x∗
(3.4)≤ (cid:101)ερ + (1 −(cid:101)ε) + 2ε(cid:48)(1 −(cid:101)ε) + (1 −(cid:101)ε)(cid:107)v∗(e)x∗
(3.5)≤ (cid:101)ερ + (1 −(cid:101)ε) + 2ε(cid:48)(1 −(cid:101)ε) + (1 −(cid:101)ε)(ε(cid:48) + r + 2k)
≤(cid:101)ερ + (1 −(cid:101)ε)(1 + 3ε(cid:48) + r + 2k).
r − T ∗F ∗y∗(cid:107)
r − T ∗v∗(cid:107)
(cid:107)S − T(cid:107) = (cid:107)S∗ − T ∗(cid:107) = sup
y∗∈Γ
≤ sup
y∗∈Γ
(cid:107)y∗(F e)x∗
(cid:107)S∗y∗ − T ∗y∗(cid:107)
r − T ∗F ∗y∗(cid:107) + 2(cid:101)ε.
(3.6)
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
This means, that if we choose(cid:101)ε = (3ε(cid:48) + r + 2k)/(1− ρ + 3ε(cid:48) + r + 2k),
9
then we have (cid:107)S(cid:107) ≤ 1. In this case,
1 = x∗
r(ur) (II)
= y∗
1(x∗
r(ur)F e) (III)
= y∗
1(S(ur)) ≤ (cid:107)S(ur)(cid:107) ≤ 1.
Therefore, (cid:107)S(cid:107) = 1 and S attains the norm at the point u0 := ur ∈ SX for
which by (2.1) we already had that (cid:107)u0 − x0(cid:107) ≤ ε
k.
Now, let us estimate
For every y∗ ∈ Γ we can proceed the same way as before. Namely,
that
(cid:107)(F ∗y∗)(e)x∗
r − T ∗F ∗y∗(cid:107) (3.4)≤ 2ε(cid:48) + (cid:107)v∗(e)x∗
(3.5)≤ 3ε(cid:48) + r + 2k.
r − T ∗v∗(cid:107)
Combining this with the inequalities (3.6) and the value of(cid:101)ε we conclude
3ε(cid:48) + r + 2k
.
(cid:107)T − S(cid:107) ≤ 3ε(cid:48) + r + 2k + 2
(3.7)
Since r > 0 and 0 < ε(cid:48) < 2/3 are arbitrary, for suitable values we will
To finish the proof in the case of Y ∈ ACKρ we observe that if T belongs
have the desired estimate (cid:107)T − S(cid:107) < ν.
to a Γ-flat ideal I then S ∈ I.
1 − ρ + 3ε(cid:48) + r + 2k
Now the simpler case of Y ∈ ACK. In this case (cid:107)S∗y∗(cid:107) ≤ 1 for all
y∗ ∈ Γ thanks to (IV)'. So, (cid:107)S(cid:107) ≤ 1 for all values of(cid:101)ε ∈ [ε(cid:48), 1) and we can
simply take(cid:101)ε = ε(cid:48). With such a choice of(cid:101)ε the estimate (3.7) changes to
ε0 instead of ε and substitute k = (cid:112)ε0/2. In the case of Y ∈ ACKρ we
take ν ∈(cid:16)√
(cid:107)T − S(cid:107) ≤ 5ε(cid:48) + r + 2k, which again for small values of r and ε(cid:48) gives us
(cid:107)T − S(cid:107) < ν for the ν which corresponds to this case.
(cid:3)
Proof of Theorem 3.4. First, select ε0 ∈ (0, ε) in such a way that the in-
equality (cid:107)T x0(cid:107) > 1 − ε0 is still valid. Now we apply Lemma 3.5 with
, and in the case of
Y ∈ ACK we take ν ∈ (
(cid:3)
Remark 3.6. The statements of Lemma 3.5 and Theorem 3.4 remain correct
if in the definition of ACKρ and ACK the property (IV) is substituted by the
following weaker one, in which V1 is larger than in the original definition:
Denote V1 = {y∗ ∈ Γ : y∗(F e)+(1−ε(cid:48))(cid:107)(IY ∗ − F ∗)(y∗)(cid:107) ≤ 1}. Then
v∗(F e) ≤ ρ for every v∗ ∈ Γ \ V1.
Also, a look at the proof of Lemma 2.9 shows that the condition of T
being Γ-flat can be weaken in the following way: for every y ∈ BY and
every δ > 0 if the w∗-slice S(Γ, x, δ) := {y∗ ∈ Γ : Re y∗(y) > 1 − δ} is
√
2
√
1−ρ+
2ε0
2ε0,
,
√
2ε).
(cid:17)(cid:17)
2
1−ρ+
(cid:16)
(cid:16)
(cid:17)
1 +
√
2ε0
1 +
√
2ε
2ε
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
10
not empty, then for every ε > 0 there exists a non-empty relatively w∗-open
subset V ⊂ S(Γ, x, δ) such that diam(T ∗(V )) < ε.
There are two reasons why we have selected the more restrictive variants.
Firstly, with the restrictive definition of (IV) we are able to prove a nice
stability result (Theorem 4.12 below), and secondly, all the examples with
"relaxed" versions of (IV) and of Γ-flatness that we have in hand, satisfy
the restrictive variant of (IV) and of Γ-flatness.
4. BANACH SPACES WITH ACK STRUCTURE
The aim of this section is presenting those natural examples of Banach
spaces having ACK structure as well as showing the stability of the ACK
structure under some operations, such us (cid:96)∞-sums or injective tensor prod-
ucts.
First of all, let us introduce the first natural class of Banach spaces with
ACK structure. As commented above, Definition 3.1, comes from an anal-
ysis of the proofs in [9]. We shall show next that, indeed, every uniform
algebra A has simple ACK structure. The key tool is Lemma 4.2, that was
proved in [9, Lemma 2.5 and Lemma 2.7], and is about the existence of
peak functions f ∈ SA whose range is contained in the Stolz's region
Stε = {z ∈ C : z + (1 − ε)1 − z ≤ 1}.
For a topological space (T, τ ), we denote by Cb(T ) the space of bounded
continuous functions f : T → K equipped with the sup-norm.
Definition 4.1. Let (T, τ ) be a topological space. A subalgebra A ⊂ Cb(T )
is said to be an ACK-subalgebra, if for every non-empty open set W ⊂ T
and 0 < ε < 1, there exist f ∈ A and t0 ∈ W such that f (t0) = (cid:107)f(cid:107)∞ = 1,
f (t) < ε for every t ∈ T \ W and f (T ) ⊂ Stε.
Lemma 4.2. Let A ⊂ C(K) be a uniform algebra. Then there exists a
topological space ΓA such that A is isometric to an ACK-subalgebra of
Cb(ΓA). In the case of K being the space of multiplicative functionals on A
the corresponding ΓA can be selected as a topological subspace of K.
We will use the following elementary property of Stε.
Lemma 4.3. If z belongs to the Stolz region Stε, then zn ∈ Stε.
Proof. For every z ∈ Stε it holds
zn + (1 − ε)1 − zn = zn + (1 − ε)1 − z1 + z + . . . + zn−1
≤ zn + (1 − z)1 + z + . . . + zn−1
≤ zn + (1 − z)(1 + z + . . . + zn−1)
= zn + (1 − zn) = 1,
which finishes the proof.
(cid:3)
The following simple lemma gives an essential property that turns uni-
form algebras into Banach spaces with simple ACK structure.
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
11
Lemma 4.4. Let A ⊂ Cb(ΓA) be an ACK-subalgebra. Then, for every
non-empty open set W ⊂ ΓA and 0 < ε < 1, there exist a non-emtpy subset
W0 ⊂ W , functions f, e ∈ A, and t0 ∈ W0 such that f (t0) = (cid:107)f(cid:107) = 1,
e(t0) = (cid:107)e(cid:107) = 1, f (t) < ε for every t ∈ ΓA \ W0, 1 − e(t) < ε for every
t ∈ W0 and f (ΓA) ⊂ Stε.
Proof. By using Definition 4.1 for the open set W ⊂ ΓA and ε, we get a
function e ∈ A and t0 ∈ W such that e(t0) = (cid:107)e(cid:107) = 1, e(t) < ε for
every t ∈ ΓA \ W and e(ΓA) ⊂ Stε. Let W0 := {t ∈ W : 1 − e(t) < ε}.
Define the function fn : ΓA → K by fn(t) := (e(t))n whose range, by
Lemma 4.3, is contained in Stε. From the very definition of W0 and the
fact that e(ΓA) ⊂ Stε, we deduce that e(t) ≤ 1 − ε(1 − ε) < 1 for
every t ∈ ΓA \ W0. Thus, taking a suitable n0 ∈ N, we can assume that
fn0(t) = e(t)n0 < ε on ΓA \ W0. Therefore, f := fn0 ∈ A gives the
(cid:3)
conclusions of the lemma.
Theorem 4.5. Let A ⊂ Cb(ΓA) be an ACK-subalgebra, and let X be a
subspace A ⊂ X ⊂ Cb(ΓA) that has the following property: f x ∈ X
for every x ∈ X and f ∈ A. Then X ∈ ACK with the corresponding
1-norming subset of BX∗ being Γ = {δt : t ∈ ΓA}.
Proof. Fix ε > 0 and a non-emtpy relatively w∗-open subset U = {δt :
t ∈ W ⊂ ΓA} ⊂ Γ. Observe that W ⊂ ΓA is open. Now, by applying
Lemma 4.4 to W with ε we obtain the corresponding W0 ⊂ ΓA, t0 ∈ W0,
f, eA ∈ A. Let us define V ⊂ U, x∗
1 ∈ V , e ∈ SX and F ∈ L(X) as
follows:
V := {δt : t ∈ W0},
x∗
1 := δt0,
e := eA, F x := f x, for x ∈ X.
Then, F ∗x∗ = f (t)x∗ for every x∗ = δt ∈ Γ. We shall show that properties
(I) -- (VI) are satisfied. First, (cid:107)F(cid:107) ≤ 1 and (cid:107)F e(cid:107) = e(t0)f (t0) = 1,
which proves (I). Property (II) is straightforward from x∗
1(f e) =
e(t0)f (t0) = 1. From (F ∗x∗
1(x) we deduce
1, which is (III). To show (IV)', take x∗ = δt ∈ Γ and estimate
that F ∗x∗
(cid:107)F ∗x∗(cid:107) + (1 − ε)(cid:107)(IX∗ − F ∗)(x∗)(cid:107)
1)(x) = x(t0)f (t0) = x(t0) = x∗
1(F e) = x∗
1 = x∗
≤ f (t) + (1 − ε)1 − f (t) ≤ 1.
Let us show now (V). Take x∗ = δt ∈ Γ. In case t belongs to ΓA \ W0,
then (cid:107)F ∗x∗(cid:107) = f (t) < ε. Otherwise, t ∈ W0 (that is, x∗ ∈ V ), using
that F ∗x∗ = f (t)x∗ and that f ∈ SX, we deduce that f (t)x∗ ∈ aco{0, V }.
Hence, in both cases
dist(F ∗x∗, aco{0, V }) < ε.
Finally, for every v∗ ∈ V we have that v∗(e) = e(t) for some t ∈ W0. So,
v∗(e) − 1 = e(t) − 1 ≤ ε,
which shows (VI) and finishes the proof.
(cid:3)
12
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
From Lemma 4.2 and Theorem 4.5 taking X = A we obtain the promised
example.
Corollary 4.6. Every uniform algebra A has simple ACK structure.
i.e., A(T) is the closure in C(T) of the set {(cid:80)m
of the set {(cid:80)m
Theorem 4.5 gives more examples of spaces with simple ACK structure.
For instance, let T be the unit disk in C, A(T) ⊂ C(T) be the disc-algebra,
k=0 akzk : ak ∈ C, m ∈ N}
of all polynomials. For a given n ∈ N denote An(T) the closure in C(T)
k=−n akzk : ak ∈ C, m ∈ N}. Then A(T) and X = An(T)
satisfy all the conditions of Theorem 4.5, so An(T) ∈ ACK, but An(T) is
not an algebra. Another example: let c0 ⊂ X ⊂ (cid:96)∞. Then X ∈ ACK.
The first example is of illustrative character, because the space An(T) is
isometric to the algebra A(T). In contrast, the second example gives a big
variety of mutually non-isomorphic spaces with ACK structure. Observe
that the simple ACK structure of those X such that c0 ⊂ X ⊂ (cid:96)∞ can be
also deduced from Theorem 4.9 below.
Remark 4.7. In general, it is not clear whether for a given T ∈ FlΓ(X, Y )
the formula (3.2) gives a Γ-flat operator S. But, under the conditions of
Theorem 4.5, we have an additional property F ∗x∗ = f (t)x∗. Combining
this property with (iv) of Theorem 2.7, we get S ∈ FlΓ(X, Y ). In particular,
in the case of uniform algebras the Bishop -- Phelps -- Bollob´as type approxi-
mation of Γ-flat operators can be made by operators that are Γ-flat as well.
Now we show that Banach spaces with Lindenstrauss' property β (see
for instance [18]) have ACK structure.
Definition 4.8. A Banach space X is said to have the property β if there
α : α ∈ Λ} ⊂ SX∗ and ρ ∈ [0, 1)
exist two sets {xα : α ∈ Λ} ⊂ SX, {x∗
such that the following conditions hold:
α(xα) = 1;
α(xγ) ≤ ρ < 1 if α (cid:54)= γ; and
(i) x∗
(ii) x∗
(iii) (cid:107)x(cid:107) = sup{x∗
α(x) : α ∈ Λ}, for all x ∈ X.
Theorem 4.9. Let X have the property β. Then X ∈ ACKρ with the same
value of ρ as in Definition 4.8 and with Γ = {x∗
α : α ∈ Λ} from that
definition. Moreover, if X has property β with ρ = 0, then X ∈ ACK.
Proof. Since X has property β, the set Γ = {x∗
α : α ∈ Λ} is a 1-norming
subset of BX∗. Observe that property β implies that (Γ, w∗) is a discrete
topological space. Fix ε > 0 and a non-empty relatively w∗-open subset
U ⊂ Γ. Take x∗
1 ∈ V , e ∈ SX,
and F ∈ L(X) as follows:
1 := x∗
x∗
α0,
α0 for every x∗ ∈ X∗. We shall show that
It is clear that F ∗x∗ = x∗(xα0)x∗
properties (I) -- (VI) of Definition 3.1 hold true. Properties (I) -- (III) are
α0 ∈ U. Let us define the corresponding V , x∗
α0} ⊂ U,
e := xα0, F (x) := x∗
V := {x∗
α0(x)xα0.
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
13
routine. To show (IV) observe first that
(cid:13)(cid:13)F ∗x∗
α0
(cid:13)(cid:13) + (1 − ε)(cid:13)(cid:13)(IX∗ − F ∗)(x∗
α0)(cid:13)(cid:13) =(cid:13)(cid:13)x∗
that is, x∗
and thus v∗(F e) = x∗
α0 ∈ V1. Consequently, whenever v∗ = x∗
α(xα0) ≤ ρ.
In case that ρ = 0, we have that F ∗x∗
(cid:107)F ∗x∗
α(cid:107) + (1 − ε)(cid:107)(IX∗ − F ∗)x∗
α = 0 for every α (cid:54)= α0, so
α(cid:107) < 1,
α(cid:107) = (1 − ε)(cid:107)x∗
(cid:13)(cid:13) = 1,
α0(xα0)x∗
α ∈ Γ \ V1, then α (cid:54)= α0
α0
α(xα0)x∗
α ∈ Γ, because F ∗x∗ = x∗
α0, so v∗(e) − 1 = 0 ≤ ε.
i.e., V1 = Γ.
Property (V) is a consequence of the fact that F ∗x∗ ∈ aco{0, V } for
every x∗ = x∗
α0. Finally, property (VI)
and in turn our conclusions are consequence of the fact that the unique
v∗ ∈ V is v∗ = x∗
(cid:3)
Corollary 4.10 ([1, Theorem 2.2]). Let Y have property β. Then, for every
Banach space X, the pair (X, Y ) has the Bishop -- Phelps -- Bollob´as property
for operators.
Proof. In the proof of Theorem 4.9, (Γ, w∗) is a discrete topological space.
Therefore every operator T ∈ L(X, Y ) is Γ-flat (Example C after Definition
(cid:3)
2.8). Now the application of Theorem 3.4 completes the proof.
Now we show the stability of the ACK structure with respect to the oper-
ations of (cid:96)∞-sum and injective tensor product of two spaces (Theorem 4.11
and Theorem 4.12)
Theorem 4.11. Let X, Y be Banach spaces having ACK structure with
ρ = max{ρX, ρY }. Moreover, Z ∈ ACK whenever X, Y ∈ ACK.
Proof. Observe that both X and Y have ACK structure with parameter ρ.
Let ΓX ⊂ BX∗ and ΓY ⊂ BY ∗ be the corresponding 1-norming subsets in
Definition 3.1. Then, the set
parameters ρX and ρY respectively. Then Z := X(cid:76)∞ Y ∈ ACKρ with
Γ := {(x∗, 0) : x∗ ∈ ΓX} ∪ {(0, y∗) : y∗ ∈ ΓY }
is a 1-norming subset of BZ∗. Take a non-empty relatively w∗-open subset
U ⊂ Γ. Then, there exist relatively w∗-open subsets UX ⊂ ΓX and UY ⊂
ΓY that are not both empty and such that (UX × {0}) ∪ ({0} × UY ) ⊂ U.
Without loss of generality we may assume that UX (cid:54)= ∅.
empty subset VX ⊂ UX, x∗
erties (I) -- (VI). Thus, we can define the corresponding V ⊂ U, z∗
e ∈ SZ and F ∈ L(Z) as follows:
Fix ε > 0. By using Definition 3.1 for X, ε, and UX we obtain a non-
1 ∈ VX, eX ∈ SX, FX ∈ L(X) with the prop-
1 ∈ V ,
V := {(x∗, 0) : x∗ ∈ VX} ⊂ U,
1 := (x∗
z∗
1, 0),
e := (eX, 0),
and for (x, y) ∈ Z,
Let us check the required properties. It is clear that (cid:107)F(cid:107) = 1 and that
(cid:107)F e(cid:107) = (cid:107)FX(eX)(cid:107) = 1, which shows (I). (II) follows easily; z∗(F e) =
F (x, y) := (FX(x), 0).
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
14
x∗
1(FXeX) = 1. Due to the fact that (FXx∗
F ∗z∗
x∗ ∈ VX,1 we have
1 = z∗
1, 0), we deduce that
1, showing that (III) holds. Now, for every z∗ = (x∗, 0) ∈ V with
1, 0) = (x∗
(cid:107)F ∗z∗(cid:107) + (1 − ε)(cid:107)(IZ∗ − F ∗)(z∗)(cid:107)
Xx∗(cid:107) + (1 − ε)(cid:107)(IX∗ − F ∗
X)(x∗)(cid:107)
= (cid:107)F ∗
≤ 1,
which can be easily deduced from F ∗z∗ = (F ∗
Xx∗, 0). Consequently, for
every x∗ ∈ VX,1 we have z∗ = (x∗, 0) ∈ V1. (Observe that in the case of
simple ACK structure we have already proved (IV)'). Let v∗ ∈ Γ\ V1. Then,
either v∗ = (0, y∗), or v∗ = (x∗, 0) with x∗ ∈ ΓX \ VX,1. On the one hand,
when v∗ = (0, y∗), we have v∗(F e) = 0 ≤ ρ. On the other hand, whenever
v∗ = (x∗, 0) with x∗ ∈ ΓX \ VX,1, then v∗(F e) = x∗(FXeX) ≤ ρ, which
proves (IV). Now, let z∗ ∈ Γ. Whenever z∗ = (0, y∗) we have F ∗z∗ = 0.
Xx∗, aco{0, VX}) < ε. Thus,
Otherwise, z∗ = (x∗, 0) and we have dist(F ∗
in both cases
dist(F ∗z∗, aco{0, V }) < ε.
Finally, for every v∗ = (x∗, 0) ∈ V we have v∗(e)− 1 = x∗(eX)− 1 ≤ ε,
(cid:3)
which proves (VI) and concludes our proof.
Recall, that given two normed spaces X and Y , one can define their
injective tensor product X ⊗ε Y , as the completion of (X ⊗ Y,(cid:107)·(cid:107)ε), where
(cid:107)z(cid:107)ε := sup{(cid:104)x∗ ⊗ y∗, z(cid:105) : x∗ ∈ BX∗, y∗ ∈ BY ∗},
for every z ∈ X ⊗ Y and (cid:104)x∗ ⊗ y∗, x⊗ y(cid:105) := x∗(x) y∗(y), for every x⊗ y ∈
X ⊗ Y and for every x∗ ∈ X∗ and y∗ ∈ Y ∗.
An important example of such a product is the Banach space C(K) ⊗ε Y ,
which can be naturally identified with C(K, Y ), that is, the Banach space
of continuous (Y,(cid:107)·(cid:107))-valued functions defined on K, endowed with the
supremum norm (cid:107)f(cid:107) = sup{(cid:107)f (t)(cid:107) : t ∈ K}.
Note that it follows from the definition of the injective norm that if X0 ⊂
BX∗ and Y0 ⊂ BY ∗ are 1-norming, then for every z ∈ X ⊗ε Y the following
equality holds:
(cid:107)z(cid:107)ε = sup{(cid:104)x∗ ⊗ y∗, z(cid:105) : x∗ ∈ X0, y∗ ∈ Y0}.
Recall also that (cid:107)x∗ ⊗ y∗(cid:107)(X ⊗εY )∗ = (cid:107)x∗(cid:107) · (cid:107)y∗(cid:107) for every x∗ ∈ X∗ and
y∗ ∈ Y ∗.
This is all the information about tensor products that will be used in The-
orem 4.12 below. We refer to Ryan's book [20] for tensor products theory
in general and the above definitions and statements in particular.
Theorem 4.12. Let X and Y be Banach spaces both of which have ACK
(resp. ACKρ) structure. Then, X ⊗ε Y has ACK (resp. ACKρ) structure.
Proof. Since X and Y have ACK (resp. ACKρ) structure, there exist 1-
norming sets ΓX ⊂ SX∗ and ΓY ⊂ SY ∗ satisfying Definition 3.1. Define the
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
15
First, we shall show that the map φ is continuous. Let {(x∗
map φ : (BX∗, w∗) × (BY ∗, w∗) → (B(X ⊗εY )∗, w∗) by φ(x∗, y∗) = x∗ ⊗ y∗,
for every x∗ ∈ BX∗ and for every y∗ ∈ BY ∗.
α)}α∈Λ be
a convergent net to (x∗, y∗) ∈ BX∗ × BY ∗. Then, for every x⊗ y ∈ X ⊗ Y ,
we can estimate
α, y∗
α) − φ(x∗, y∗), x ⊗ y(cid:105) = x∗
α(x)y∗
(cid:104)φ(x∗
α, y∗
α(y) − x∗(x)y∗(y)
α(y) − y∗(y))
≤ (x∗
≤ x∗
≤ x∗
α(x) − x∗(x))y∗
α(x) − x∗(x)(cid:107)y∗
α(x) − x∗(x)(cid:107)y(cid:107) + (cid:107)x(cid:107)y∗
α(y) + x∗(x)(y∗
α(cid:107)(cid:107)y(cid:107) + (cid:107)x∗(x)(cid:107)y∗
α(y) − y∗(y),
α(y) − y∗(y)
which tends to zero. This argument extends easily to every element in X⊗Y
and, in turn, to every z ∈ X ⊗ε Y (due to the boundedness of the range of
the map φ).
The 1-norming set Γ that we need for our theorem can be introduced as
follows:
Γ := {x∗ ⊗ y∗ : x∗ ∈ ΓX, y∗ ∈ ΓY } = φ(ΓX × ΓY ).
0, y∗
0 ∈ WX, y∗
1 ∈ VX and y∗
0 ∈ ΓY be such that φ(x∗
0 ∈ WY and φ(WX × WY ) ⊂ U.
Let ε > 0 and U be a non-empty relatively w∗-open subset of Γ. Let
0 ∈ ΓX and y∗
0) ∈ U. The continuity of φ
x∗
ensures that there exist non-empty relatively w∗-open subsets WX ⊂ ΓX,
WY ⊂ ΓY such that x∗
We can apply Definition 3.1 to X and Y , to the former with ε/2 and WX
and to the latter with ε/2 and WY , to find two non-empty sets VX ⊂ WX
1 ∈ VY , two points eX ∈ SX
and VY ⊂ WY , two functionals x∗
and eY ∈ SY and finally, two operators FX ∈ L(X) and FY ∈ L(Y ),
satisfying respectively the properties (I) -- (VI), or with their corresponding
modifications for the the simple ACK structure. Denote also by VX,1 and
VY,1 the corresponding variants for X and Y of the set V1 from property
(IV) of Definition 3.1.
1 ∈ V , e ∈
SX⊗εY , F ∈ L(X ⊗ε Y ) as follows: V := φ(VX × VY ) ⊂ U, z∗
1 :=
1, e := eX ⊗ eY , and F (x ⊗ y) := FX(x) ⊗ FY (y) for
φ(x∗
every x ⊗ y ∈ X ⊗ Y . It remains to check the properties (I) -- (VI). First,
observe that F ∗(x∗ ⊗ y∗) = F ∗
Y y∗ for every x∗ ∈ X∗ and y∗ ∈ Y ∗.
(I) Let z belong to BX ⊗εY , then
(cid:107)F z(cid:107)ε = sup
x∗∈ΓX
= sup
x∗∈ΓX
≤ (cid:107)F ∗
Now, define the non-emtpy set V ⊂ U and corresponding z∗
1, y∗
(cid:104)x∗ ⊗ y∗, F z(cid:105) = sup
sup
y∗∈ΓY
x∗∈ΓX
(cid:104)F ∗
sup
y∗∈ΓY
X(cid:107)(cid:107)F ∗
Y (cid:107) ≤ 1,
(cid:104)F ∗(x∗ ⊗ y∗), z(cid:105)
Xx∗(cid:107)(cid:107)F ∗
sup
y∗∈ΓY
Y y∗, z(cid:105) ≤ sup
x∗∈ΓX
Xx∗ ⊗ F ∗
Xx∗ ⊗ F ∗
1) = x∗
1 ⊗ y∗
Y y∗(cid:107)
sup
y∗∈ΓY
(cid:107)F ∗
which implies that (cid:107)F(cid:107) = 1, since
(cid:107)F e(cid:107) = (cid:107)FXeX ⊗ FY eY (cid:107) = (cid:107)FXeX(cid:107)(cid:107)FY eY (cid:107) = 1.
(II) z∗
1(F e) = (x∗
1 ⊗ y∗
1)(FXeX ⊗ FY eY ) = x∗
1(FXeX)y∗
1(FY eY ) = 1.
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
16
(III) F ∗z∗
(F ∗z∗
1 ⊗ y∗
1 = z∗
1)(x ⊗ y) = (x∗
1, since for every x ⊗ y ∈ X ⊗ Y we have
1)(FXx ⊗ FY y) = (F ∗
Xx∗
1)(x)(F ∗
1(y) = z∗
1(x)y∗
1)(x ⊗ y) = x∗
which, in turn, implies that (F ∗z∗
(IV) For (x∗, y∗) ∈ ΓX × ΓY , denote z∗ = x∗ ⊗ y∗. Firstly, let us show that
for every x∗ ∈ VX,1 and y∗ ∈ VY,1 the functional z∗ belongs to V1, i.e., that
(cid:107)F ∗z∗(cid:107) + (1 − ε)(cid:13)(cid:13)(I(X ⊗εY )∗ − F ∗)(z∗)(cid:13)(cid:13) ≤ 1.
Y y∗
1)(y),
1(x ⊗ y).
First of all, observe that
(cid:107)x∗ ⊗ y∗ − F ∗
Xx∗ ⊗ F ∗
= (cid:107)x∗ ⊗ (y∗ − F ∗
≤ (cid:107)y∗ − F ∗
Y y∗(cid:107) =
Y y∗) − (x∗ − F ∗
Y y∗(cid:107) + (cid:107)F ∗
Y y∗(cid:107)(cid:107)(x∗ − F ∗
Y y∗(cid:107)
Xx∗) ⊗ F ∗
Xx∗)(cid:107) .
Xx∗ ⊗ F ∗
Xx∗(cid:107) (cid:107)F ∗
Y y∗(cid:107)(cid:0)(cid:107)F ∗
Y y∗(cid:107) + (1 − ε)(cid:107)x∗ ⊗ y∗ − F ∗
Xx∗(cid:107) + (1 − ε)(cid:107)(x∗ − F ∗
Y y∗(cid:107) ≤ 1.
Therefore,
(cid:107)F ∗
= (cid:107)F ∗
≤ (cid:107)F ∗
This implies that for every z∗ = x∗ ⊗ y∗ ∈ Γ\ V1 we have two possibilities:
either x∗ /∈ VX,1 or y∗ /∈ VY,1. By symmetry, it is sufficient to consider
x∗ /∈ VX,1. In this case x∗(FXeX) ≤ ρ, so
Xx∗)(cid:107)(cid:1) + (1 − ε)(cid:107)y∗ − F ∗
Y y∗(cid:107) + (1 − ε)(cid:107)y∗ − F ∗
Y y∗(cid:107)
Y y∗(cid:107)
z∗(F e) = x∗(FXeX)y∗(FY eY ) ≤ x∗(FXeX) ≤ ρ.
(V) We shall show that dist(F ∗z∗, aco{0, V }) < ε for every z∗ = x∗ ⊗
Xx∗, aco{0, VX}) < ε/2 and that
y∗ ∈ Γ. Due to the facts that dist(F ∗
Y ∈
dist(F ∗
Y (cid:107) < ε/2. Then
aco{0, VY } such that (cid:107)F ∗
v∗ := v∗
Y y∗, aco{0, VY }) < ε/2, there exist v∗
X ⊗ v∗
X ∈ aco{0, VX} and v∗
X(cid:107) < ε/2 and (cid:107)F ∗
Xx∗ − v∗
Y y∗ − v∗
Y belongs to aco{0, V } and
X) ⊗ F ∗
X(cid:107)(cid:107)F ∗
(VI) For every v∗ = x∗ ⊗ y∗ ∈ V we get
(cid:107)F ∗z∗ − v∗(cid:107) ≤ (cid:107)(F ∗
≤ (cid:107)F ∗
Xx∗ − v∗
Xx∗ − v∗
Y y∗(cid:107) + (cid:107)v∗
Y y∗(cid:107) + (cid:107)v∗
X ⊗ (F ∗
X(cid:107)(cid:107)F ∗
Y y∗ − v∗
Y y∗ − v∗
Y )(cid:107)
Y (cid:107) ≤ ε.
v∗(e) − 1 = x∗(eX)y∗(eY ) − 1 ≤ x∗(eX)y∗(eY ) − y∗(eY )
+ y∗(eY ) − 1 ≤ ε
2
y∗(eY ) +
≤ ε.
ε
2
This finishes the proof.
(cid:3)
4.1. Sup-normed spaces of vector-valued functions. As we mentioned
in the introduction, Acosta, Becerra Guerrero, Garc´ıa, Kim, and Maestre
considered A-BPBp in spaces of continuous vector-valued functions. Let
us recall their result explicitly. Here, as usual, σ(Z, ∆) denotes the weakest
topology on Z in which all elements of ∆ ⊂ Z∗ are continuous.
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
17
Theorem 4.13 ([2, Theorem 3.1]). Let X, Z be Banach spaces, K be a com-
pact Hausdorff topological space. Let Z satisfy property β for the subset of
α : α ∈ ∆}. Let τ ⊇ σ(Z, ∆) be a linear topology on
functionals ∆ = {z∗
Z dominated by the norm topology. Then for every closed operator ideal I
contained in the ideal of Asplund operators, we have that (X, C(K, (Z, τ )))
has the Bishop -- Phelps -- Bollob´as property for operators from I.
The next proposition together with Theorem 3.4 generalize Theorem 4.13
for the case of Z endowed with its strong topology.
Proposition 4.14. Let K be a compact Hausdorff topological space. Then,
(Y ∈ ACKρ) ⇒ (C(K, Y ) ∈ ACKρ);
(Y ∈ ACK) ⇒ (C(K, Y ) ∈ ACK).
Proof. Bearing in mind Corollary 4.6 and Theorem 4.12, the fact that the
space C(K) ⊗ε Y is isometric to C(K, Y ) concludes the proof.
(cid:3)
Our aim now is showing a generalization of Theorem 4.13 in the spirit of
the ACK structure, that covers all topologies τ from that theorem. In order
to do this we need some terminology.
For a topological space T and a Banach space Z denote by Cbof(T, Z)
the space of all bounded openly fragmented (see Definiton 2.6) functions
f : T → Z equipped with the sup-norm. For a topology τ on Z denote
by Cb(T, (Z, τ )) the space of bounded τ-continuous functions f : T → Z
equipped with the sup-norm.
Definition 4.15. Let Z ∈ ACKρ and let Γ ⊂ BZ∗ be the corresponding
1-norming set. A linear topology τ on Z is said to be Γ-acceptable, if it is
dominated by the norm topology and dominates σ(Z, Γ).
The following result simultaneously generalizes our Theorem 4.5 and
Theorem 4.13. We state the result in the most general settings, which makes
the statement bulky. Some "elegant" partial cases will be given as corollar-
ies.
Theorem 4.16. Let A ⊂ Cb(ΓA) be an ACK-subalgebra. Let Z be a Ba-
nach space and O ⊂ L(Z) such that Z ∈ O-ACKρ (Z ∈ O-ACK) with
ΓZ ⊂ BZ∗ being the corresponding 1-norming set. Finally, let τ be a ΓZ-
acceptable topology on Z. Let X ⊂ Cb(ΓA, (Z, τ )) be a Banach space
satisfying the following properties:
(i) For every x ∈ X and f ∈ A the function f x belongs to X.
(ii) X contains all functions of the form f ⊗ z, f ∈ A, z ∈ Z.
(iii) F ◦ x ∈ X for every x ∈ X and F ∈ O.
(iv) For every finite collection {xk}n
k=1 ⊂ X the corresponding func-
tion of two variables ϕ : ΓA×(ΓZ, w∗) → Kn, defined by ϕ(t, z∗) =
(z∗(xk(t)))n
k=1, is quasi-continuous.
18
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
Then X ∈ ACKρ (X ∈ ACK, respectively) with the corresponding 1-
norming subset of BX∗ being Γ = {δt ⊗ z∗ : t ∈ ΓA, z∗ ∈ ΓZ}, where the
functional δt ⊗ z∗ ∈ X∗ acts as follows: (δt ⊗ z∗)(x) = z∗(x(t)).
Proof. Fix ε > 0 and a non-empty relatively w∗-open subset U ⊂ Γ. Let
0 ∈ U. Since U is relatively
t0 ∈ ΓA and z∗
k=1 ⊂ X such that δt ⊗ z∗ ∈ Γ belongs to U
w∗-open, there exist {xk}n
whenever
0 ∈ ΓZ be such that δt0 ⊗ z∗
(cid:104)(δt0 ⊗ z∗
Consider the non-emtpty open set
max
1≤k≤n
0) − (δt ⊗ z∗), xk(cid:105) < 1.
max
1≤k≤n
0(xk(t)) − z∗
B := {t ∈ ΓA : z∗
0(xk(t0)) < 1 for 1 ≤ k ≤ n},
and define the following non-empty relatively w∗-open subset of ΓZ:
0(xk(t0)) < 1 for 1 ≤ k ≤ n}.
D := {z∗ ∈ ΓZ : z∗(xk(t0)) − z∗
Using property (iv) for {xk}n
k=1 ⊂ X we can find a non-empty open subset
B1 ⊂ B and a non-empty relatively w∗-open subset D1 ⊂ D such that for
every t ∈ B1 and every z∗ ∈ D1 it holds
z∗(xk(t)) − z∗
0(xk(t0)) < 1.
Define the non-empty subset W := {δt ⊗ z∗ : t ∈ B1, z∗ ∈ D1} ⊂ Γ. It is
clear that W ⊂ U.
By applying Definition 3.1 to Z, ΓZ, D1 and (ε/2), we get VZ ⊂ D1,
1 ∈ VZ, eZ ∈ SZ and FZ ∈ O satisfying (I) -- (VI). Denote also VZ,1 ⊂ ΓZ,
z∗
the subset that appears in property (IV) (in the case of Z ∈ ACK we have
VZ,1 = ΓZ). By applying Lemma 4.4 to A, ΓA, the non-empty open set
B1 and (ε/2), we find a non-empty subset B2 ⊂ B1, functions f0, eA (both
belonging to A) and s0 ∈ B2, satisfying its conclusions.
Finally, let us define the requested non-empty subset V ⊂ U and corre-
sponding x∗
1 ∈ V , e ∈ SX, F ∈ L(X) as follows:
V := {δt ⊗ z∗ : t ∈ B2, z∗ ∈ VZ} ⊂ W ⊂ U,
1 := δs0 ⊗ z∗
x∗
1,
e(t) := eA(t)eZ, for every t ∈ ΓA
(condition (ii) implies e ∈ X), and
(F x)(t) := f0(t)FZ(x(t)),
for every x ∈ X and for every t ∈ ΓA. Conditions (i) and (iii) ensure that
F (x) ∈ X. Observe that for every x∗ = δt ⊗ z∗ ∈ Γ
Zz∗) .
F ∗x∗ = f0(t) (δt ⊗ F ∗
It remains to check the properties (I) -- (VI).
(I) It is clear that (cid:107)F(cid:107) = (cid:107)FZ(cid:107) = 1 and (cid:107)F e(cid:107) = (cid:107)f0eA(cid:107)(cid:107)FZ(eZ)(cid:107) = 1.
(II) x∗
(III) F ∗x∗
(F ∗x∗
1(f0(s0)eA(s0)FZ(eZ)) = 1.
1, since for every x ∈ X we have
1 (f0(s0)FZx(s0)) = (F ∗
1(F e) = z∗
1 = x∗
1)(x) = z∗
1)(x(s0)) = z∗
1(x(s0)) = x∗
Zz∗
1(x).
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
19
(IV) For every x∗ ∈ Γ, we have x∗ = δt ⊗ z∗, t ∈ ΓA and z∗ ∈ ΓZ. First,
consider the case z∗ ∈ VZ,1 and observe that
Zz∗(cid:107)
(cid:107)(IX∗ − F ∗)(x∗)(cid:107) = (cid:107)z∗ − f0(t)F ∗
≤ 1 − f0(t)(cid:107)z∗(cid:107) + f0(t) · (cid:107)(IZ∗ − F ∗
Z)(z∗)(cid:107) + 1 − f0(t).
= f0(t) · (cid:107)(IZ∗ − F ∗
Z)(z∗)(cid:107)
Therefore, in this case
(cid:107)F ∗x∗(cid:107) + (1 − ε)(cid:107)(IX∗ − F ∗)(x∗)(cid:107)
= f0(t) · (cid:107)F ∗
Zz∗(cid:107) + (1 − ε)(cid:107)z∗ − f0(t)F ∗
Zz∗(cid:107) + (1 − ε)(cid:107)(IZ∗ − F ∗
≤ f0(t)(cid:0)(cid:107)F ∗
Zz∗(cid:107)
Z)(z∗)(cid:107)(cid:1) + (1 − ε)1 − f0(t)
≤ f0(t) + (1 − ε)1 − f0(t) ≤ 1.
Whenever Z ∈ ACK, then VZ,1 = ΓZ, so the above inequality holds for
every z∗ ∈ ΓZ. Thus, we have proved (IV)'. If Z ∈ ACKρ we still must
consider those x∗ belonging to Γ \ V1. The above inequality implies that
z∗ /∈ VZ,1 and, consequently, z∗(FZeZ) ≤ ρ which, in turn, implies that
x∗(F e) = f0(t)eA(t)z∗(FZeZ) ≤ ρ.
(V) Let x∗ = δt ⊗ z∗ ∈ Γ. Recall that F ∗x∗ = f0(t)δt ⊗ F ∗
Zz∗. Set VA :=
{δt : t ∈ B2}. In the proof of Theorem 4.5 it was proved that for every
t ∈ ΓA it holds
Thus, there exist a∗ ∈ aco{0, VA} and b∗ ∈ aco{0, VZ} such that
(cid:107)f (t)δt − a∗(cid:107) <
and (cid:107)F ∗
Zz∗ − b∗(cid:107) <
ε
2
ε
2
.
In particular, since a∗ ⊗ b∗ belongs to aco{0, V }, we can deduce that
dist(F ∗x∗, aco{0, V }) ≤ (cid:107)f0(t)δt ⊗ F ∗
≤ (cid:107)f0(t)δt ⊗ F ∗
+ (cid:107)f0(t)δt ⊗ b∗ − a∗ ⊗ b∗(cid:107)
≤ (cid:107)F ∗
Zz∗ − a∗ ⊗ b∗(cid:107)
Zz∗ − f0(t)δt ⊗ b∗(cid:107) +
Zz∗ − b∗(cid:107) + (cid:107)f0(t)δt − a∗(cid:107) < ε.
2 and z∗(eZ) − 1 ≤ ε
(VI) For every x∗ = δt⊗z∗ ∈ V we have t ∈ B2 and z∗ ∈ VZ. Consequently,
eA(t) − 1 ≤ ε
x∗(e) − 1 = eA(t)z∗(eZ) − 1 = eA(t)(z∗(eZ) − 1) + (eA(t) − 1) ≤ ε,
(cid:3)
which completes the proof.
2. From this we get
On the other hand, by our construction, we deduce that
dist(f (t)δt, aco{0, VA}) <
dist (F ∗
Zz∗, aco{0, VZ}) <
ε
2
.
ε
2
.
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
20
Remark 4.17. Under the hypothesis of the previous theorem, given F ∈
L(Z) and f ∈ A we can consider the operators CF : X → X and Pf : X →
X defined, respectively, by CF (x) = F ◦ x and Pf (x) = f x, for every
x ∈ X. Then, if we set O(cid:48) := {CF ◦ Pf : F ∈ O, f ∈ A}, then X has
O(cid:48)-ACKρ (resp. O(cid:48)-ACK) structure.
Conditions (i) -- (iii) in Theorem 4.16 are easily verified in concrete ex-
amples. In contrast, condition (iv) looks technical. So, in order to make
Theorem 4.16 more applicable, we shall present easy-to-verify sufficient
conditions for (iv).
Before passing to these sufficient conditions, observe that the function
of two variables ϕ : ΓA × (ΓZ, w∗) → Kn from condition (iv) is separately
continuous. Therefore, the role of sufficient condition for (iv) can be played
by any theorem about quasi-continuity of a separately continuous function
f : U × V → W . There is a number of such theorems (see Encyclopedia of
Mathematics article "Separate and joint continuity" or the introduction to
[7]). For example, according to Namioka's theorem [15] this (and a much
stronger result) occurs for U being a regular, strongly countably complete
topological space, V being a locally compact σ-compact space and W being
a pseudo-metric space. The results of the kind "separate continuity implies
quasi-continuity" that we list and apply below do not pretend to be new.
Proposition 4.18. Let U, V , W be topological spaces, V be discrete and
f : U × V → W be separately continuous. Then, f is continuous (and
consequently quasi-continuous).
If Z has property β, the corresponding (ΓZ, w∗) is a discrete topolog-
ical space. Thus, the above proposition guaranties the validity of (iv) of
Theorem 4.16 in this case.
Corollary 4.19. Under the conditions of Theorem 4.13, C(K, (Z, τ )) ∈
ACKρ, where ρ is the parameter from the property β of Z. If β = 0, then
C(K, (Z, τ )) ∈ ACK. In particular, this implies the conclusion of Theorem
4.13.
Proposition 4.18 also guaranties (iv) of Theorem 4.16 in the case of ΓA =
N (just change the roles of U and V in Proposition 4.18).
If we apply
Theorem 4.16 with A = c0 ⊂ Cb(N) = (cid:96)∞, this leads to the following
result:
Corollary 4.20. Let Z ∈ ACKρ (Z ∈ ACK), c0(Z) ⊂ X ⊂ (cid:96)∞(Z),
and X has the following property: (F z1, F z2, . . .) ∈ X for every x =
(z1, z2, . . .) ∈ X and F ∈ L(Z). Then X ∈ ACKρ (X ∈ ACK respec-
tively).
This corollary is applicable to c0(Z) and (cid:96)∞(Z) themselves and also for
some intermediate spaces like c0(Z, w) of weakly null sequences in Z.
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
21
Proposition 4.21. Let Z be a Banach space, (ΓA, τ ) be a topological space,
ΓZ ⊂ (BZ∗, w∗), and xk : ΓA → Z for k ∈ {1, 2, . . . , n} be τ-σ(Z, ΓZ)-
continuous and τ-(cid:107)·(cid:107)-openly fragmented functions. Then, the function
ϕ : (ΓA, τ ) × (ΓZ, w∗) → Kn given by ϕ(t, z∗) = (z∗(xk(t)))n
k=1 is quasi-
continuous.
0) ∈ ΓA × ΓZ. Let UA ⊂ ΓA, UZ ⊂ ΓZ be open and
Proof. Fix (t0, z∗
0 respectively. Set U := UA × UZ. We
w∗-open neighborhoods of t0 and z∗
have to show that, for a given ε > 0, there exist a non-empty open subset
WA ⊂ UA and a non-empty relatively w∗-open subset WZ ⊂ UZ such that
for every t ∈ WA and every z∗ ∈ WZ
z∗(xk(t)) − z∗
0(xk(t0)) < ε.
max
1≤k≤n
Fix δ < ε/4 and define
(cid:26)
(4.1)
(cid:27)
.
VA :=
t ∈ UA : max
1≤k≤n
0(xk(t)) − z∗
z∗
0(xk(t0)) < δ
The set VA ⊂ UA is a non-emtpy open neighborhood of t0 because of the
0 ◦ xk is a K-valued τ-continuous
τ-σ(Z, ΓZ) continuity of xk (the map z∗
function). Applying inductively the definition of openly fragmented func-
tion, we define a non-empty open set WA ⊂ (VA, τ ) in such a way that for
all k = 1, . . . , n it holds
diam(xk(WA)) < δ.
(cid:26)
(cid:27)
Fix a t1 ∈ WA and define the non-empty relatively w∗-open subset WZ ⊂
UZ as follows:
WZ :=
z∗ ∈ UZ : max
1≤k≤n
z∗(xk(t1)) − z∗
0(xk(t1)) < δ
.
Let us show, for every t ∈ WA and every z∗ ∈ WZ, the validity of
inequality (4.1):
z∗
0(xk(t0)) − z∗(xk(t)) ≤ z∗
0(xk(t))
0(xk(t0)) − z∗
0(xk(t)) − z∗
0(xk(t1))
+ z∗
+ z∗
0(xk(t1)) − z∗(xk(t1))
+ z∗(xk(t1)) − z∗(xk(t)).
The first summand in the right-hand side of the previous inequality does
not exceed δ since t ∈ VA. Accordingly, the second and fourth summands
0, z∗ ∈ BZ∗ and (cid:107)xk(t) − xk(t1)(cid:107) < δ since
are both bounded by δ since z∗
t, t1 ∈ WA and diam(xk(WA)) < δ. Finally, the corresponding third sum-
mand is bounded by δ since z∗ ∈ WZ. Therefore,
z∗
0(xk(t0)) − z∗(xk(t)) ≤ 4δ < ε,
which completes the proof of (4.1) and that of the proposition.
(cid:3)
22
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
As an application of the previous proposition we get the following corol-
laries which contain as a particular case the space Cw(K, Z) of Z-valued
weakly continuous functions for Z ∈ ACKρ (or Z ∈ ACK).
Corollary 4.22. Let Z ∈ O-ACKρ (or Z ∈ O-ACK) and A ⊂ C(K)
be a uniform algebra with K being the space of multiplicative functionals
on A. Fix ΓZ ⊂ H ⊂ Z∗, where ΓZ is the 1-norming set given by the
ACK structure of Z. Denote by Aσ(Z,H)(K, Z) the following subspace of
C(K, (Z, σ(Z, H))):
Aσ(Z,H)(K, Z) =(cid:8)f ∈ Z K : z∗ ◦ f ∈ A for all z∗ ∈ H(cid:9) .
Let us assume that
(i) F ∗H ⊂ H for every F ∈ O.
(ii) (f (K), σ(Z, H)) is fragmented by the norm for every f belonging
to Aσ(Z,H)(K, Z).
Then, Aσ(Z,H)(K, Z) ∈ ACKρ (resp. Aσ(Z,H)(K, Z) ∈ ACK).
Sketch of the proof: It relays on the use of Theorem 4.16. Let ΓA ⊂ K
be the corresponding subset from Lemma 4.2. Then, restrictions of ele-
ments of A to ΓA form an ACK-subalgebra Cb(ΓA) isometric to A (that
we identify with A) and restrictions of elements of Aσ(Z,H)(K, Z) to ΓA
form a subspace X ⊂ Cb(ΓA, (Z, σ(Z, H))) isometric to Aσ(Z,H)(K, Z).
The conditions (i) and (ii) of Theorem 4.16 follow from the definition of
Aσ(Z,H)(K, Z). The condition (iii) of Theorem 4.16 is reduced to the present
condition (i). And, finally, the condition (iv) of Theorem 4.16 is reduced to
(cid:3)
the present (ii) by using Proposition 4.21.
The condition (i) above could be quite demanding, for instance, when
O = L(Z) in which case H is forced to be Z∗. However, in all concrete
examples that we know of ACK structure, the family O can be taken really
small. Thus, for concrete examples of Z, the condition (i) could be easily
satisfied for every election of H.
By using the results from [5] it can be shown that condition (ii) above
is satisfied for every H whenever (Z, w) is Lindelof. Indeed, given f be-
longing to Aσ(Z,H)(K, Z), f (K) ⊂ Z is σ(Z, H)-compact, thus, it is also
Lindelof. A straightforward application of [5, Corollary E] ensures that
(f (K), σ(Z, H)) is norm-fragmented. Hence, in this case, Corollary 4.22
can be simplified as follows:
Corollary 4.23. Let Z ∈ O-ACKρ (or Z ∈ O-ACK) such that (Z, w) is
Lindelof and A ⊂ C(K) be a uniform algebra with K being the space of
multiplicative functionals on A. Fix ΓZ ⊂ H ⊂ Z∗ such that F ∗H ⊂ H
for every F ∈ O, where ΓZ is the 1-norming set given by the ACK structure
of Z. Then, Aσ(Z,H)(K, Z) ∈ ACKρ (resp. Aσ(Z,H)(K, Z) ∈ ACK).
Observe that when Z has property β, the set O coincides with the set
α(·) xα : α ∈ Λ}. Therefore, in this case, F ∗H ⊂ H for every H and for
{x∗
every F ∈ O. Thus, we have proved the following corollary.
Γ-FLATNESS AND BISHOP -- PHELPS -- BOLLOB ´AS TYPE THEOREMS
23
Corollary 4.24. Let Z be a Banach space with property β such that (Z, w)
is Lindelof and A ⊂ C(K) be a uniform algebra with K being the space of
α : α ∈
multiplicative functionals on A. Fix ΓZ ⊂ H ⊂ Z∗, where ΓZ = {x∗
Λ}. Then, Aσ(Z,H)(K, Z) ∈ ACKρ.
However, this technique can not fully generalize Theorem 4.13 by Acosta
et al. to the case of vector-valued uniform algebras, since here the Lindelof
property is essential and property β does not imply in general weak Lin-
delof. Observe that nevertheless the original statement of Theorem 4.13 is
covered completely by our Corollary 4.19.
REFERENCES
[1] M.D. Acosta, R.M. Aron, D. Garc´ıa, and M. Maestre, The Bishop-Phelps-Bollob´as
theorem for operators, J. Funct. Anal. 254 (2008), no. 11, 2780 -- 2799.
[2] M.D. Acosta, J. Becerra Guerrero, D. Garc´ıa, S.K. Kim, and M. Maestre, Bishop --
Phelps -- Bollob´as property for certain spaces of operators, J. Math. Anal. Appl. 414
(2014), No. 2, 532 -- 545.
[3] E. Asplund, Fr´echet differentiability of convex functions, Acta Math. 121 (1968), 31 --
[4] R.M. Aron, B. Cascales, and O. Kozhushkina, The Bishop-Phelps-Bollob´as theorem
and Asplund operators, Proc. Amer. Math. Soc. 139 (2011), no. 10, 3553 -- 3560.
[5] B. Cascales, I. Namioka, and G. Vera, The Lindelof property and fragmentability,
Proc. Amer. Math. Soc. 128 (2000), 3301 -- 3309.
[6] R. Baire, Sur les functions des variables reelles, Ann. Mat. Pura Appl. 3 (1899),
47.
1 -- 122.
[7] T. Banakh, Quasicontinuous and separately continuous functions with values in
Maslyuchenko spaces, arXiv:1506.01661v4 [math.GN].
[8] B. Bollob´as, An extension to the theorem of Bishop and Phelps, Bull. London Math.
Soc. 2 (1970), 181 -- 182.
[9] B. Cascales, A.J. Guirao, and V. Kadets, A Bishop -- Phelps -- Bollob´as type theorem for
uniform algebras, Advances in Mathematics 240 (2013), 370 -- 382.
[10] B. Cascales, I. Namioka, and J. Orihuela, The Lindelof property in Banach spaces,
Studia Mathematica 150 (2003), 165 -- 192.
[11] J. Diestel and J.J. Uhl, Vector measures, American Mathematical Society, Providence,
R.I., 1977.
[12] 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/Ouvrages
de Math´ematiques de la SMC, Springer, New York, 2011.
[13] S. Kempisty, Sur les fonctions quasi-continues, Fund. Math. 19 (1932), 184 -- 197.
[14] G. Koumoullis, A generalization of functions of the first class, Topology and its Ap-
plications 50, no. 3 (1993), 217 -- 239.
[15] I. Namioka, Separate and joint continuity Pacific J. Math. , 51 (1974) pp. 515 -- 531.
[16] I. Namioka, Radon-Nikod´ym compact spaces and fragmentability, Mathematika 34
[17] I. Namioka and R.R. Phelps, Banach spaces which are Asplund spaces, Duke Math.
[18] J. Lindenstrauss, On operators which attain their norm, Israel J. Math. 1 (1963),
[19] R.R. Phelps, Support Cones in Banach Spaces and Their Applications, Adv. Math.
no. 2 (1987), 258 -- 281.
J. 42 no. 4, (1975), 735 -- 750.
139 -- 148.
13 (1974), 1 -- 19.
24
CASCALES, GUIRAO, KADETS, AND SOLOVIOVA
[20] R.A. Ryan, Introduction to tensor products of Banach spaces, Springer Monographs
in Mathematics. London: Springer, xiv, 225 p. (2002).
[21] Ch. Stegall, The Radon-Nikod´ym property in conjugate Banach spaces, Trans. Amer.
Math. Soc. 206 (1975), 213 -- 223.
[22] Ch. Stegall, The duality between Asplund spaces and spaces with the Radon-Nikod´ym
property, Israel J. Math. 29 no. 4, (1978), 408 -- 412.
[23] Ch. Stegall, The Radon-Nikod´ym property in conjugate Banach spaces. II, Trans.
Amer. Math. Soc. 264 no. 2 (1981), 507 -- 519.
DEPTO DE MATEM ´ATICAS, UNIVERSIDAD DE MURCIA, 30100 ESPINARDO, MUR-
CIA, SPAIN
E-mail address: [email protected]
IUMPA, UNIVERSITAT POLIT `ECTNICA DE VAL `ENCIA, 46022, VALENCIA, SPAIN
E-mail address: [email protected]
DEPARTMENT OF MATHEMATICS AND INFORMATICS, V.N. KARAZIN KHARKIV
NATIONAL UNIVERSITY, 61022 KHARKIV, UKRAINE
E-mail address: [email protected]
E-mail address: [email protected]
|
1510.01653 | 3 | 1510 | 2017-02-11T23:54:40 | Frame Scalings: A Condition Number Approach | [
"math.FA"
] | Scaling frame vectors is a simple and noninvasive way to construct tight frames. However, not all frames can be modifed to tight frames in this fashion, so in this case we explore the problem of finding the best conditioned frame by scaling, which is crucial for applications like signal processing. We conclude that this problem is equivalent to solving a convex optimization problem involving the operator norm, which is unconventional since this problem was only studied in the perspective of Frobenious norm before. We also further study the Frobenious norm case in relation to the condition number of the frame operator, and the convexity of optimal scalings. | math.FA | math |
FRAME SCALINGS: A CONDITION NUMBER APPROACH
PETER CASAZZA AND XUEMEI CHEN
Abstract. Scaling frame vectors is a simple and noninvasive way to construct tight frames. How-
ever, not all frames can be modifed to tight frames in this fashion, so in this case we explore the
problem of finding the best conditioned frame by scaling, which is crucial for applications like signal
processing. We conclude that this problem is equivalent to solving a convex optimization problem
involving the operator norm, which is unconventional since this problem was only studied in the
perspective of Frobenius norm before. We also further study the Frobenius norm case in relation
to the condition number of the frame operator, and the convexity of optimal scalings.
Keywords: Scalable frames, Condition number, Tight frame
AMS subject classification: 15B48, 42C15, 65F35
1. Introduction
A family of vectors Φ = {ϕi}M
i=1 is a frame in an N -dimensional Hilbert space HN if there are
constants 0 < A ≤ B < ∞ so that for all x ∈ HN ,
Akxk2 ≤
hx, ϕii2 ≤ Bkxk2.
N
Xi=1
The largest A and smallest B satisfying these inequalities are called lower and upper frame bounds,
respectively. One often also writes Φ for the N × M matrix whose ith column is the vector ϕi.
When A = B, the frame is called an A-tight frame. Furthermore, A = B = 1 produces a Parseval
frame. In the sequel, the set of frames with M vectors in HN will be denoted by F(M, N ). It is
well known that Φ is A-tight if and only if
(1)
S := ΦΦ∗ =
ϕiϕ∗i = AIN ,
M
Xi=1
where IN is the identity matrix in HN , and S is the frame operator of the frame. We refer to [15]
for an introduction to frame theory and to [8] for an overview of the current research in the field.
Frames have traditionally played a significant role in the theory of signal processing, but today
they have found application to packet based network communication [7, 18], wireless sensor networks
[9, 10, 11, 12], distributed processing [7], quantum information theory, bio-medical engineering
[2, 25], compressed sensing [3, 14], fingerprinting [26], spectral theory [6, 19, 20], and much more.
Some of the applications of frames result from their ability to deliver redundant, yet stable
expansions. The redundancy of a frame is typically utilized by applications which may require
robustness of the frame coefficients to noise, erasures, quantization, etc. In this setting tight frames
can give fast convergence and recovery. It is known that unit norm tight frames are characterized
in terms of the frame potential [1]. There have been various works on constructions of tight frames
[4, 6, 10, 11, 24, 27]. However, it is desirable to construct tight frames by just scaling each frame
vector as it is noninvasive, and frame properties such as erasure resilience or sparse expansions are
left untouched by this modification.
The authors were supported by NSF DMS 1307685; NSF ATD 1321779.
1
2
PETER CASAZZA AND XUEMEI CHEN
This procedure is called frame scaling. To be specific, a frame Φ = {ϕi}M
i=1 for HN is called
scalable if there exist scalars {si}M
i=1 is a tight frame for HN . By the nature of
scaling, if a frame is scalable, there exist scalars such that the scaled frame is a Parseval frame. So
by (1), a frame is scalable if and only if there exists ci ≥ 0 such that
i=1 such that {siϕi}M
IN =
ciϕiϕ∗i .
M
Xi=1
The ci here corresponds to each scalar as ci = si2 ≥ 0. The notion of a scalable frame was first
introduced in [22]. In [22], characterizations of scalable frames, both of functional analytic and
geometric type were derived in the infinite as well as finite dimensional settings. The work [5]
considers the complex case as well, and it was shown that the set of all possible sequences of scalars
is the convex hull of minimal scalars. The paper [16] focuses on the numerical algorithms to find
different scalings with different purposes. Recently, the work [13] studies the case when a frame is
not scalable by measuring
(2)
min
ci≥0 kIN −
M
Xi=1
ciϕiϕ∗i kF
using the minimal ellipsoid of the convex hull of the frame vectors, where k · kF is the Frobenius
norm. The theory of frame scalings has also been extended to matrices with Laurent polynomials
(with applications to construct tight wavelet filter banks) [21], and probabilistic frames [23].
When a frame is not scalable, we wish to find scalars such that {siϕi}M
i=1 is as tight as possible.
This should naturally mean that {siϕi}M
i=1 is the best conditioned in the sense that the ratio of
upper and lower frame bounds is the closest to 1. However, it is not clear whether solving (2) gives
the best conditioned frame.
In this note, we study the scalability of frames by si so that {siϕi}M
a given square matrix T , we define
i=1 is best conditioned. For
cond(T ) :=
biggest singular value of T
smallest singular value of T
.
We include the singular case where condition number is ∞. We wish to solve
(3)
min
ci≥0
cond M
Xi=1
ciϕiϕ∗i! ,
i=1 ciϕiϕ∗i(cid:17) is exactly the ratio of the upper and lower frame bounds of the scaled
i=1. For convenience of discussion, we include "frames" that do not span, whose
The first contribution of this paper is to establish an equivalence between (3) and the following
where cond(cid:16)PM
frame {siϕi}M
condition number is ∞.
problem:
(4)
min
ci≥0 kIN −
M
Xi=1
ciϕiϕ∗i k2,
where k · k2 is the operator norm of a matrix. We prove this equivalence in Theorem 6 of Section
3. This means that rather than solving the scalability problem with the Frobenius norm, using the
operator norm is more efficient if one wants to find the best conditioned frame by scaling, which is
the original goal of scaling. Moreover, it is not very clear how one can solve problem (3) at first
glance. With Theorem 6, we have converted it to a convex programming Problem. Some properties
of the minimizer of (4) are also established in Section 3.
FRAME SCALINGS: A CONDITION NUMBER APPROACH
3
The second contribution of the paper is to study how solving (2) is related to solving (3).
Unfortunately, we show in Section 4.1 that the best scaled "frame" from (2) may not even span
the space (hence is not a frame). Theorem 17 lists a sufficient condition for when the best scaling
corresponds to a frame in HN . In the end, we further study the convexity of all the optimal scalings
(a polytope), and show that the vertices of this polytope are the so called minimal optimal scalings.
This is interesting in its own right.
2. Notations and convexity of optimal scalings
Given Φ = {ϕi}M
frame of the outer products {ϕiϕ∗i}M
squared.
i=1 ∈ F(M, N ), we define F = (cid:0)hϕi, ϕji2(cid:1), which is the Gram matrix of the
i=1. Let g = (kϕ1k2,kϕ2k2,··· ,kϕMk2) be the vector of norms
Some of the notations are common for both (2) and (4), so k · k could be either the operator
norm or the Frobenius norm in the following. Let
M
M
min
Xi=1
ci≥0 kIN −
c′iϕiϕ∗i k = kIN − TΦk.
i=1 : minci≥0 kIN −PM
We call (c′i)M
{(c′i)M
we define (c′i)M
span TΦ with nonnegative coefficients.
Xi=1
ciϕiϕ∗i k = kIN −
i=1 an optimal scaling and TΦ = PM
i=1 c′iϕiϕ∗i an optimal operator. We call OΦ =
i=1 ciϕiϕ∗i k = kIN −PM
i=1 c′iϕiϕ∗i k} the set of optimal scalings. Moreover,
i=1 ∈ OΦ to be a minimal optimal scaling if no proper subset of {ϕiϕ∗i : c′i > 0} can
TΦ is not necessarily unique for the operator norm (See Example 2), in which case TΦ will mean
the set of optimal operators. So in the definition of minimal optimal scaling above, "span TΦ"
means span any operator in TΦ.
In fact, the concept of minimal optimal scaling is much more
meaningful in the Frobenius norm case as we will see in Section 4.3.
The following theorem states that OΦ is a convex set. But for the Frobenius case, we can actually
say much more, see Section 4.3.
Theorem 1. With either operator norm or Frobenius norm, the set of optimal scalings OΦ is a
convex set.
Proof. Let (c′i)M
be the corresponding optimal operators. For any a, b ≥ 0, a + b = 1, we can prove (ac′i + bd′i)M
an optimal scaling too since
i=1 ∈ OΦ be two optimal scalings, and T1 =PM
i=1 c′iϕiϕ∗i , T2 =PM
i=1, (d′i)M
i=1 d′iϕiϕ∗i
i=1 is
M
kIN −
Xi=1
(ac′i + bd′i)ϕiϕ∗i k = kIN − (aT1 + bT2)k
≤akIN − T1k + bkIN − T2k = min
ci≥0 kIN −
M
Xi=1
ciϕiϕ∗i k.
(cid:3)
3. Minimizing with the operator norm
This section focuses on problem (4), which uses the operator norm for the scaling problem. Once
again, let
(5)
min
ci≥0 kIN −
M
Xi=1
The optimal operator TΦ does not need to be unique here.
ciϕiϕ∗i k2 = kIN − TΦk2.
4
PETER CASAZZA AND XUEMEI CHEN
, 0), ϕ2 = ( 1√2
Example 2. Let ϕ1 = (1, 1√2
i=1 is a frame of R3. For
i=1 ciϕiϕ∗i . One
can obtain, without much effort, that (5) is minimized when c1 = c2 = 2/3, and c3 takes on
i=1 ciϕiϕ∗i have eigenvalues
any ci ≥ 0, it is straightforward to calculate the eigenvalues of the operator P3
any value in [1 − 2√2/3, 1 + 2√2/3]. With such selection of c′is, P3
1 − 2√2/3, c3, 1 + 2√2/3, which result in different operators with different c3.
, 1, 0), ϕ3 = (0, 0, 1), so {ϕi}3
The following proposition is obvious.
Proposition 3. If a positive semi-definite matrix T has eigenvalues λ1 ≥ λ2 ≥ ··· ≥ λN ≥ 0, then
kIN − Tk2 = max{1 − λ1,1 − λN}.
The next proposition is essential in proving that solving (4) optimizes the condition number. It
essentially says that the optimal operator comes to a balance when the largest and the smallest
eigenvalue have equal distance to 1.
Proposition 4. Let Φ = {ϕi}M
eigenvalues λ1 ≥ λ2 ≥ ··· ≥ λN ≥ 0 and c = 2
i=1 ∈ F(M, N ) be a frame for HN .
If T = PM
i=1 ciϕiϕ∗i has
λ1+λN
, then
Moreover, if T = TΦ then
and consequently λ1 + λN = 2.
Proof. We check
and
Also,
kI − cTk2 = cλ1 − 1 = 1 − cλN .
kIN − Tk2 = λ1 − 1 = 1 − λN ,
cλN =
2
λ1 + λN
λN =
2
+ 1 ≤ 1,
λ1
λN
cλ1 =
2
λ1 + λN
λ1 =
2
1 + λN
λ1
≥ 1,
cλ1 + cλN = c(λ1 + λN ) = 2,
so that cλ1 − 1 = 1 − cλN . Now,
kIN − cTk2 = max{1 − cλ1,1 − cλN} = cλ1 − 1 = 1 − cλN .
Now assume T = TΦ and we check two cases.
Case 1: λ1 + λN > 2.
In this case, c < 1 and λ1 > 1. So
kIN − cTk2 = cλ1 − 1 < λ1 − 1 ≤ kIN − Tk2,
which is a contradiction to (5).
Case 2: λ1 + λN < 2.
In this case, c > 1 and λN < 1. So
kIN − cTk2 = 1 − cλN < 1 − λN ≤ kIN − Tk2,
which is a contradiction to (5).
So we must have c = 1 and therefore λ1 ≥ 1 ≥ λN .
The proof of Proposition 4 immediately implies
(cid:3)
FRAME SCALINGS: A CONDITION NUMBER APPROACH
5
Corollary 5. If T has eigenvalues λ1 ≥ ··· ≥ λN ≥ 0, let c = 2
λ1+λN
, then
and equality holds if and only if c = 1.
kIN − cTk2 ≤ kIN − Tk2,
Now we are ready to prove the main theorem.
Theorem 6. Problem (3) and Problem (4) are equivalent in the sense that
(a) Solving (4) gives the minimal condition number among {PM
(b) If {di}M
i=1 ciϕiϕ∗i(cid:17), then the scalars {cdi}M
i=1 = arg minci≥0 cond(cid:16)PM
mum in (4) for some c > 0.
i=1 ciϕiϕ∗i : ci ≥ 0}.
i=1 achieves the mini-
As a consequence of (a), if
min
ci≥0 kIN −
M
Xi=1
ciϕiϕ∗i k2 = kIN − TΦk2,
then TΦ has the smallest condition number among {PM
of the optimal operator has an upper bound as
i=1 ciϕiϕ∗i : ci ≥ 0}, and the condition number
cond(TΦ) ≤ cond(ΦΦ∗).
Proof. (a) Assume TΦ has eigenvalues λ1 ≥ λ2 ≥ ··· ≥ λN ≥ 0. By Proposition 4, λ1 ≥ 1 ≥ λN and
λ1−1 = 1−λN . For arbitrary ci ≥ 0, let R =PM
i=1 ciϕiϕ∗i have eigenvalues µ1 ≥ µ2 ≥ ··· ≥ µN ≥ 0.
Letting c = 2
, we have by Proposition 4,
µ1+µN
Now,
kIN − cRk2 = cµ1 − 1 = 1 − cµN .
Hence, 1 ≤ λ1 ≤ cµ1. Similarly,
λ1 − 1 = kIN − TΦk2 ≤ kIN − cRk2 = cµ1 − 1.
1 − λN = kIN − TΦk2 ≤ kIN − cRk2 = 1 − cµN ,
and hence 1 ≥ λN ≥ cµN . It follows immediately that
cµ1
cµN
λ1
λN ≤
cond(TΦ) =
= cond(R).
This shows that solving (4) gives the minimal condition number among all scalings ci.
(b) Suppose T = PM
eigenvalues λ1 ≥ ··· ≥ λN > 0. For arbitrary ci ≥ 0, let R = PM
µ1 ≥ µ2 ≥ ··· ≥ µN ≥ 0, and let c = 2
i=1 diϕiϕ∗i has the smallest condition number after solving (3), and T has
i=1 ciϕiϕ∗i have eigenvalues
, d = 2
. So
µ1+µN
λ1+λN
cond(cT ) ≤ cond(dR)
cλ1
cλN ≤
dµ1
dµN ⇒
2 − cλN
cλN ≤
⇒
⇒1 − cλN ≤ 1 − dµN ⇒ kIN − cTk2 ≤ kIN − dRk2.
dµN
2 − dµN
By Corollary 5,
kIN − cTk2 ≤ kIN − dRk2 ≤ kIN − Rk2.
(cid:3)
Remark 7. From the proof of Proposition 4, Corollary 5 and Theorem 6, Theorem 6 also holds
for the case when Φ is not a frame (do not span HN ).
6
PETER CASAZZA AND XUEMEI CHEN
If we do start with an actual frame Φ, since the condition number of TΦ has a finite upper bound,
we immediately have
Corollary 8. Given Φ ∈ F(M, N ), using the operator norm, any optimal operator TΦ, as in
minci≥0 kIN −PM
of kIN − TΦk2 as
Being invertible means that the smallest eigenvalue λN > 0, so we can easily get an upper bound
i=1 ciϕiϕ∗i k2 = kIN − TΦk2, is invertible.
kIN − TΦk2 = 1 − λN < 1.
The following theorem shows that this upper bound is tight.
Theorem 9. For any fixed M, N , and any Φ ∈ F(M, N ),
min
ci≥0 kIN −
M
Xi=1
ciϕiϕ∗i k2 = kIN − TΦk2 < 1.
Moreover, given an arbitrary ǫ > 0, there exists a frame Ψ = {ψi}M
ci≥0 kIN −X ciψiψ∗i k2 = 1 − ǫ.
min
To prove the tightness part of this theorem, we need a few lemmas first.
i=1 such that
Lemma 10. Given M ≥ N , there is a universal constant K > 0 so that whenever {ϕi}M
unit norm frame in HN and
i=1 is a
M
kIN −
Xi=1
then c′i ≤ K for all i = 1, 2, . . . , M.
Proof. Assume the optimal operator TΦ has eigenvalues λ1 ≥ ··· ≥ λN . By Proposition 4,
ci≥0 kIN −X ciϕiϕ∗i k2,
c′iϕiϕ∗i k2 = min
Therefore
c′i ≤
M
Xi=1
c′i =
M
Xi=1
Lemma 11. The function
λ1 ≤ λ1 + λN = 2.
2 = Tr M
Xi=1
c′ikϕik2
c′iϕiϕ∗i! =
N
Xi=1
λi ≤ 2N,
(cid:3)
f (Φ) = f (ϕ1, ϕ2, . . . , ϕM ) = min
ci≥0 kIN −X ciϕiϕ∗i k2,
defined on N × M matrices with unit norm columns, is continuous with respect to Φ.
Proof. Given a sequence of N × M matrices with unit norm columns Φ(n) = {ϕi(n)}M
i=1 such that
Φ(n) = Φ,
lim
n→∞
we need to prove limn→∞ f (Φ(n)) = f (Φ).
i=1 so that
Choose {di(n)}M
Xi=1
kIN −
M
di(n)ϕi(n)ϕi(n)∗k2 = min
ci≥0 kIN −
ciϕi(n)ϕi(n)∗k2 = f (Φ(n)).
M
Xi=1
Since the {di(n)} are uniformly bounded by the previous Lemma, by switching to a subsequence
we may assume
di(n) = di for all i = 1, 2, . . . , M.
lim
n→∞
FRAME SCALINGS: A CONDITION NUMBER APPROACH
7
Now, if
(6)
we are done. For any scalar ci ≥ 0,
kIN −
M
Xi=1
diϕiϕ∗i k2 = f (Φ),
kIN −
M
Xi=1
ciϕiϕ∗i k2 = lim
n→∞kIN −
ciϕi(n)ϕi(n)∗k2
di(n)ϕi(n)ϕi(n)∗k2
M
M
Xi=1
Xi=1
diϕiϕ∗i k2,
≥ lim
n→∞kIN −
Xi=1
M
= kIN −
which means that the scalars (di)M
i=1 achieves the minimum, hence (6).
(cid:3)
Proof of Theorem 9. Let f (Φ) be as defined in Lemma 11.
Let e1 = [1, 0,··· , 0] ∈ HN , and Φ1 := {e1,··· , e1}, which is M copies of e1. Then f (Φ1) = 1.
Choose Ψ that is close enough to Φ1 but spans HN (a frame), then we have f (Ψ) is close to
f (Φ1) = 1.
(cid:3)
4. Minimizing with Frobenius norm
This section focuses on Problem (2). Notice for the Frobenius norm, the optimal operator TΦ is
i=1 ciϕiϕ∗i : ci ≥ 0}. Since CΦ is closed and convex,
the projection of IN onto the cone CΦ = {PM
we have the uniqueness of TΦ. To remind the reader of the notation, let
min
ci≥0 kIN −
M
Xi=1
ciϕiϕ∗i kF = kIN − TΦkF .
4.1. Invertibility of TΦ. We are still interested in the condition number of the optimal operator
TΦ. Surprisingly, TΦ does not need to be invertible, or equivalently, TΦ may not be a frame operator.
We need some basic facts to set up counterexamples.
ciϕiϕ∗i kF = kIN − PΦkF .
Notice we allow ci to be negative here, so PΦ is the projection of IN onto the subspace spanned by
{ϕiϕ∗i }M
Observe that
Let
(7)
M
min
Xi=1
i=1. Therefore PΦ 6= TΦ in general.
ci∈RkIN −
(8)
(9)
M
Xi=1
h(c) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
ciϕiϕ∗i − IN(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Taking the partial derivative with respect to ci, we get
ciϕiϕ∗i + I
M
Xi=1
cikϕik2
2 + N.
2
F
M
M
=
cicjϕiϕ∗i ϕjϕ∗j − 2
Xi=1
= tr
Xi,j=1
Xi,j=1
cicjhϕi, ϕji2 − 2
= 2Xj
cjhϕi, ϕji2 − 2kϕik2
M
2
∂h
∂ci
8
PETER CASAZZA AND XUEMEI CHEN
i=1 ciϕiϕ∗i , then the scalars should satisfy
Therefore if PΦ =PM
(10)
2,··· ,kϕNk2
where F is as defined in Section 2 as F (i, j) = hϕi, ϕji2, and g = (kϕ1k2
2)∗. Since F is
i=1, F is strictly positive definite if {ϕiϕ∗i }M
the Gram matrix of {ϕiϕ∗i }M
i=1 is linearly independent,
in which case c = F −1g is the global minimum of (7). Furthermore, if it so happens that the
coordinates of c = F −1g are all nonnegative, then naturally PΦ = TΦ.
F c = g,
Equation (10) is not a necessary condition for c to be an optimal scaling because we are mini-
mizing h(c) over the first orthant (rather than over the whole space). But given an optimal scaling,
we still have ∂h
= 0 for i ∈ {j : cj > 0} since it does not sit on the boundary of the first orthant.
∂ci
To summarize,
Proposition 12. Given Φ ∈ F(M, N ),
(1) If {ϕiϕ∗i }M
(2) If c = (ci)M
i=1 is linearly independent, then c = F −1g is the unique solution of (7). If further
c = F −1g ≥ 0 (component wise), c is the unique solution of both (2) and (7), and PΦ = TΦ.
i=1 is an optimal scaling, then
(11)
and
cjhϕi, ϕji2 = kϕik2
2,
∀i ∈ {j : cj > 0}
M
Xj=1
0 0 1
1 1 1
0 1 0
Example 13. For Φ =
, the optimal operator TΦ is not invertible. Because if TΦ is
invertible, the optimal scaling must have ci > 0, i = 1, 2, 3. By Proposition 12(2), it must hold that
the solution of (10) is all positive. However, solving (10) gives c = (−1, 1, 1).
Example 14. In RN (N ≥ 3) we define
ϕ1 = (1, 0, . . . , 0), ϕ2 = (a, b, 0, . . . , 0) where a2 + b2 = 1 and a < b,
ϕi = (c, c, 0, . . . ,p1 − 2c2
, 0, . . . , 0), where c2 + c2(a + b)2 = 1 + a2, i ≥ 3.
{z
}
i=1 is a frame of RN . We display the first two columns
ith
It is easy to check that 2c2 < 1. Now, {ϕi}N
of F :
1
a2
c2
...
c2
F =
a2
1
c2(a + b)2
...
c2(a + b)2
···
···
···
...
···
Once again, if TΦ were invertible, we require ci > 0, and F c = g. However, the solution of F c = g
is
a contradiction.
c = (
1
1 + a2 ,
1
1 + a2 , 0, . . . , 0),
Remark 15. The example above can be generalized to arbitrarily many frame vectors. Indeed, de-
i=1 ciϕiϕ∗i :
fine ϕi, i = 1, 2,··· , N as in Example 14, and add in ϕN +1,··· , ϕM such that ϕiϕ∗i ∈ {PN
ci ≥ 0}, i = N + 1,··· , M .
The following proposition shows that in R2, the optimal operator is always invertible.
FRAME SCALINGS: A CONDITION NUMBER APPROACH
9
Proposition 16. Given any frame Φ ∈ F(M, 2), the optimal operator TΦ for the Frobenius norm
is invertible.
Proof. Suppose the eigenvalues of TΦ are λ1 ≥ λ2, then kIN − TΦk2
order to prove λ2 > 0, it suffices to show that kIN − TΦkF < 1.
F = (1 − λ1)2 + (1 − λ2)2. In
Pick two independent vectors, say ϕ1, ϕ2 from the frame. It suffices to show that
min
ci≥0 kIN − c1ϕiϕ∗i − c2ϕ2ϕ∗2kF < 1,
since kIN − TΦkF ≤ minci≥0 kIN − c1ϕiϕ∗i − c2ϕ2ϕ∗2kF .
that
With rotation and adding a negative sign to the vectors, we can assume without loss of generality
We can further assume that they are both unit norm as a2 + b2 = 1 and a ≥ b.
A direct calculation shows that the eigenvalues of the frame operator S of {ϕ1, ϕ2} are {2a2, 2b2}.
ϕ1 = (a, b) and ϕ2 = (a,−b).
minkI −
2
Xi=1
Taking c = 1
2a2 , we get
ciϕiϕ∗i k2
F ≤ min
c≥0 kI − cSk2
F = min
c≥0 (cid:0)(1 − 2ca2)2 + (1 − 2cb2)2(cid:1)
minkI −
2
Xi=1
ciϕiϕ∗i k2
F ≤ (1 − b2/a2)2 < 1.
(cid:3)
4.2. When is TΦ invertible? It is not easy to find a condition on the frame so that its optimal
operator is guaranteed to be at least invertible. But the situation can be simplified when the frame
Φ is full spark (every N frame vectors from Φ span HN ), and the outer products {ϕiϕ∗i }M
i=1 are
linearly independent. In this case, if the solution c = F −1g happens to be
ci ≥ 0, and #(supp(c)) ≥ N,
then we are guaranteed to have an invertible TΦ because the scaled frame {√ciϕi} will for sure
span. Moreover, the condition that both Φ is full spark and {ϕiϕ∗i } is independent is not harsh
when M is no greater than the real dimension of the N × N symmetric matrices (N 2 for C, and
N (N + 1)/2 for R). In fact, such sets of frames are generic.
The following theorem also provides a sufficient condition for the invertibility of TΦ, when the
outer products are independent. Moreover, the advantage of this sufficient condition is that it is
directly on the frame.
Theorem 17. Let {ϕi}M
Let fi be the ith column of F =(cid:0)hϕi, ϕji2(cid:1). If
i=1 be a unit norm frame for HN and let {ϕiϕ∗i }M
i=1 be linearly independent.
kfik1 − kfjk1 < λ ∀i, j = 1, 2, . . . , M,
where λ is the smallest singular value of the the matrix F , then
TΦ = PΦ =
M
Xi=1
aiφiφ∗i where ai > 0, for all i = 1, 2, . . . , M.
Proof. By Proposition 12, it suffices to show that the solution of F a = g has positive coordinates.
This is equivalent to showing that det Fi > 0, where Fi denotes the matrix F with its ith column
replaced by g, since det F > 0.
10
PETER CASAZZA AND XUEMEI CHEN
First note that if
δ =
then
Let D =(cid:16)PM
j=1 fj(cid:17) − δg, then
1
2(cid:18) max
1≤i≤M kfik1 + min
1≤i≤M kfik1(cid:19) ,
kfik1 − δ < λ.
(12)
(13)
(14)
det F = det(f1, f2, . . . , fi−1,
fj, fi+1, . . . , fM )
M
Xj=1
= det(f1, . . . , fi−1, δg, fi+1, . . . , fM ) + det(f1, . . . , fi−1, D, fi+1, . . . , fM )
= δ · det Fi + det(f1, . . . , fi−1, D, fi+1, . . . , fM ).
Note that if D = (d1, d2, . . . , dM )T , then
di =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj
hϕi, ϕji2 − δ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
< λ, for all i = 1, 2, . . . , M.
Let Mij be the determinant of the matrix obtained by deleting the ith row and jth column of
F . For fixed i, let
D(i) = (d(i)
1 , x(i)
Let x(i) = (x(i)
1 , d(i)
2 , . . . , x(i)
(det F )x(i)
i
(15)
(16)
(17)
(18)
2 , . . . , d(i)
M )T where d(i)
j (−1)i+jMij = djMij for all j = 1, 2, . . . , N.
M )T = F −1D(i) and by Cramer's rule
= det(f1, f2, . . . , fi−1, D(i), fi+1, . . . , fM )
N
Xj=1
djMji
M
j Mji =
M
=
Xj=1
(−1)j+id(i)
(−1)j+idjMji(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≥ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xj=1
On the other hand, since λ is the smallest singular value of F ,
= det(f1, . . . , fi−1, D, fi+1, . . . , fN )
(19)
x(i)
i = (F −1D(i))i ≤ kF −1D(i)k∞ ≤ kF −1kkD(i)k∞ <
1
λ
λ
By (14), (18) and (19),
δ · det Fi = det F − det(f1,··· , fi−1, D,··· , fM )
≥ det F − (det F )x(i)
i > det F − det F · 1 = 0
(cid:3)
4.3. Convexity of OΦ revisited. The convexity of optimal scalings with the Frobenius norm is
thoroughly studied in [5], in the case when frames are scalable. We wish to generalize it to any
frames. Minimal optimal scalings play an important role in the structure of OΦ, so we make some
important observations of minimal optimal scalings first, which are missing in [5].
Proposition 18. Given Φ ∈ F(M, N ),
{ϕiϕ∗i : ci > 0}, are linearly independent.
(a) If {ci}M
(b) Different minimal optimal scalings have different supports.
i=1 is a minimal optimal scaling, then the outer products associated with it, i.e.,
FRAME SCALINGS: A CONDITION NUMBER APPROACH
11
Proof. (a) Let I = {i : ci > 0}. {ϕiϕ∗i }i∈I is linearly independent if and only if there is only one
point in the set SS = {(xi)i∈I : Pi∈I xiϕiϕ∗i = TΦ, xi ∈ R}. Suppose to the contrary that SS
is a nontrivial affine subspace (more than one point).
If we call {(xi)i∈I : xi ≥ 0} the positive
orthant, then the affine subspace SS intersects the positive orthant since we have a scaling to begin
with. Therefore, it will also intersect the boundary of the positive orthant, providing a solution of
{(xi)i∈I :Pi∈I xiϕiϕ∗i = TΦ, xi ≥ 0} with at least one xi to be 0, which contradicts to the fact that
{ci}M
i=1 is minimal.
(b) If they have the same support I, then having two minimal scalings means that {ϕiϕ∗i }i∈I is
linearly dependent, which is not true by (a).
(cid:3)
We already know that OΦ is convex. But for the Frobenius norm, it is a polytope, with minimal
optimal scalings as its vertices. The proof is similar to that of [5].
Theorem 19. With Frobenius norm, the set of optimal scalings OΦ is a polytope. Moreover, OΦ
is the convex hull of the minimal optimal scalings, i.e., the vertices of OΦ are minimal optimal
scalings.
i=1 : P xiϕiϕ∗i =
Proof. OΦ is a polytope because it is an intersection of half planes: OΦ = {(xi)M
TΦ, xi ≥ 0}.
To prove the vertices part, we first show any vertex must be a minimal optimal scaling of OΦ. Let
u ∈ OΦ be a vertex, and assume to the contrary that u is not minimal. Then there exists v ∈ OΦ
whose support is a proper subset of u's. Let w(t) = v + t(u − v). We observe that w(t) ∈ OΦ if
and only if every component of w(t), w(t)i ≥ 0. Pick
t0 =(cid:26) 2,
min{ vi
vi−ui
If for every i, vi ≤ ui
.
: vi > ui}, otherwise
We observe that t0 > 1 and w(t0)i ≥ 0, so w(t0) ∈ OΦ, which indicates that u lies on the line
segment with endpoints v and w(t0), hence not a vertex. This is a contradiction.
Now suppose we are given a minimal scaling w which is not a vertex of OΦ. Then we can write
w as a convex combination of vertices, say w =P tivi, where we know at least two ti's are nonzero,
say t1 and t2. Since both t1 and t2 are positive and all the entries of v1 and v2 are nonnegative, it
follows that supp(v1) ∪ supp(v2) ⊆ supp(w). Moreover, supp(v1) 6= supp(v2) by Proposition 18(b).
This contradicts to the fact the w is a minimal scaling.
(cid:3)
The theorem above does not work for the operator norm case because the optimal operator is
not necessarily unique.
We would like to thank the reviewers for their constructive comments and thorough review,
especially on the remark to allow condition number to be infinity.
Acknowledgements
References
[1] J. J. Benedetto and M. Fickus, Finite Normalized Tight Frames, Adv. Comput. Math., 18(2003), 357 -- 385.
[2] R. Bhagavatula, M. T. McCann, M. Fickus, C. A. Castro, J. A. Ozolek, J. Kovacevic, A vocabulary for the
identification of teratoma tissues in H&E-stained samples, J. Pathol. Inform. 19.5 (2014).
[3] B. Bodmann, J. Cahill and P.G. Casazza, Fusion frames and the restricted isometry property, In Numerical
Functional Analysis and Optimnization, Special Issue on Operator Algebras and Representation Theory:
Frames, Wavelets, Fractals, P. G. Casazza, P.E.T. Jorgensen, K.A. Kornelson, G. Kutynoik, D.R. larson, P.
Massopust, G. Olegsson, J.A. Packer, S. Silvertor and Q. Sun Eds. 33 No. 7-9 (2012)
[4] H. Broome, and S. Waldron. On the construction of highly symmetric tight frames and complex polytopes.
Linear Algebra and its Applications 439.12 (2013): 4135-4151.
[5] J. Cahill and X. Chen, A note on scalable frames, Proceedings of the 10th International Conference on Sampling
Theory and Applications, pp. 93 -- 96.
12
PETER CASAZZA AND XUEMEI CHEN
[6] J. Cahill, M. Fickus, D.G. Mixon, M.J. Poteet and N.K. Strawn, Constructing finite frames of a given sprectrum
and set of lengths, Applied and Computational Harmonic Analysis 35.1 (2013) 52-73.
[7] R. Calderbank, P.G. Casazza, A Heinecke, G. Kutyniok and A. Pezeshki, Sparse fusion frames: existence and
construction, Advances in Computational Mathematics, 35.1 (2011) 1-31.
[8] P. G. Casazza and G. Kutyniok, Finite Frame Theory, Eds., Birkhauser, Boston (2012).
[9] P.G. Casazza, M. Fickus, A. Heinecke, Y. Wang and Z. Zhou, Spectral tetris fusion frame constructions,
Journal of Fourier Analysis and Applications 18.4 (2012) 828-851.
[10] P.G. Casazza, M. Fickus, D. Mixon, Y. Wang and Z. Zhou, Constructing tight fusion frames, Applied and
Computational Harmonic Analysis, 30.2 (2011) 175-187.
[11] P.G. Casazza, A. Heinecke, K. Kornelson, Y. Wang and Z. Zhou, Necessary and sufficient conditions to perform
spectral tetris, Linear Algebra and Applications 438.5 (2013) 2239-2255.
[12] P.G. Casazza and J. Peterson, Weighted Fusion Frame Construction Via Spectral Tetris, Adv. Comput. Math
40.2 (2014) 335-351..
[13] X. Chen, G. Kutyniok, K. A. Okoudjou, F. Philipp, R. Wang, Measures of scalability, IEEE transactions on
information theory, 8(2015): 4410-4423
[14] X. Chen, H. Wang, and R. Wang. A null space analysis of the ℓ1-synthesis method in dictionary-based com-
pressed sensing Applied and Computational Harmonic Analysis 37.3 (2014): 492-515.
[15] O. Christensen, An introduction to frames and Riesz bases, Applied and Numerical Harmonic Analysis.
Birkhauser Boston, Inc., Boston, MA, 2003.
[16] C.A. Clark, and K. A. Okoudjou. On optimal frame conditioners. Sampling Theory and Applications
(SampTA), 2015 International Conference on. IEEE, 2015.
[17] M. Copenhaver, Y. Kim, C. Logan, K. Mayfield, S. K. Narayan, M. J. Petro, and J. Sheperd, Diagram vectors
and tight frame scaling in finite dimensions, Operators and Matrices, 8 (2014), 78-88.
[18] M. Fickus, D. G. Mixon, Numerically erasure-robust frames, Linear Algebra and Appl. 437.6 (2012) 1394-1407.
[19] M. Fickus, D. G. Mixon, M. J. Poteet, N. Strawn, Constructing all self-adjoint matrices with prescribed
spectrum and diagonal, Adv. Comput. Math. 39.3 (2013) 585-609.
[20] M. Fickus, D. G. Mixon, M. J. Poteet, Constructing finite frames with a given spectrum, in: "Finite Frames:
Theory and Applications", P. G. Casazza and G. Kutyniok eds., Birkhauser (2013).
[21] Y. Hur, K. A. Okoudjou. Scaling Laplacian Pyramids, SIAM Journal on Matrix Analysis and Applications
36.1 (2015): 348-365.
[22] G. Kutyniok, K. A. Okoudjou, F. Philipp, and E. K. Tuley, Scalable frames, Linear Algebra and its Applications
438 (2013), 2225 -- 2238.
[23] C. W. Lau and K. A. Okoudjou. Scalable probabilistic frames. arXiv preprint arXiv:1501.07321 (2015).
[24] J. Leng, D. Han, and T. Huang. Probability modelled optimal frames for erasures. Linear Algebra and its
Applications 438.11 (2013): 4222-4236.
[25] M. T. McCann, R. Bhagavatula, M. Fickus, J. A. Ozolek, J. Kovacevic, Automated colities detection from en-
doscopic biopsies as a tissue screening tool in diagnostic pathology, 19th IEEE conference on Image Processing
(ICIP) (2012) 2809-2812.
[26] D. G. Mixon, C. J. Quinn, N. Kiyavash, M. Fickus, Fingerprinting with equiangular tight frames, IEEE Trans.
Inform. Theory 59.3 (2013) 1855-1865.
[27] R. Vale, and S. Waldron. Tight frames generated by finite nonabelian groups. Numerical Algorithms 48.1-3
(2008): 11-27.
Department of Mathematics, University of Missouri, Columbia
E-mail address: [email protected]
Department of Mathematics and Statistics, University of San Francisco
E-mail address: [email protected]
|
1702.02689 | 3 | 1702 | 2017-12-23T21:00:06 | Supercharacters and the discrete Fourier, cosine, and sine transforms | [
"math.FA"
] | Using supercharacter theory, we identify the matrices that are diagonalized by the discrete cosine and discrete sine transforms, respectively. Our method affords a combinatorial interpretation for the matrix entries. | math.FA | math |
SUPERCHARACTERS AND THE DISCRETE FOURIER,
COSINE, AND SINE TRANSFORMS
STEPHAN RAMON GARCIA AND SAMUEL YIH
Abstract. Using supercharacter theory, we identify the matrices that are
diagonalized by the discrete cosine and discrete sine transforms, respectively.
Our method affords a combinatorial interpretation for the matrix entries.
1. Introduction
The theory of supercharacters was introduced by P. Diaconis and I.M. Isaacs in
2008 [14], generalizing earlier seminal work of C. Andr´e [2–4]. The original aim of
supercharacter theory was to provide new tools for handling the character theory
of intractable groups, such as the unipotent matrix groups Un(q). Since then,
supercharacters have appeared in the study of combinatorial Hopf algebras [1],
Schur rings [26, 29] and their combinatorial properties [15, 35, 36], and exponential
sums from number theory [9, 18, 20].
Supercharacter techniques permit us to identify the algebra of matrices that are
diagonalized by the discrete Fourier transform (DFT) and discrete cosine transform
(DCT), respectively. A natural modification handles the discrete sine transform
(DST). Although the matrices that are diagonalized by the DCT or DST have been
studied previously [7, 17, 31, 32], we further this discussion in several ways.
For the DCT, we produce a novel combinatorial description of the matrix entries
and obtain a basis for the algebra that has a simple combinatorial interpretation.
In addition to recapturing results presented from [31], we are also able to treat the
case in which the underlying cyclic group has odd order.
A similar approach for the DST runs into complications, but we can still charac-
terize the diagonalized matrices by considering the "orthocomplement" of the DCT
supercharacter theory. In special cases, the diagonalized matrices are T -class ma-
trices [7], which first arose in the spectral theory of Toeplitz matrices and have since
garnered significant interest because of their computational advantages [8,11,25,28].
For cyclic groups of even order, we recover results on [7]. However, our approach
also works if the underlying cyclic group has odd order. This is not as well studied
as the even order case. In addition, we produce a second natural basis equipped
with a novel combinatorial interpretation for the matrix entries.
For all of our results, we provide explicit formulas for the matrix entries of the
most general matrix diagonalized by the DCT or DST, respectively.
We hope that it will interest the supercharacter community to see that its tech-
niques are relevant to the study of matrix transforms that are traditionally the
Partially supported by a David L. Hirsch III and Susan H. Hirsch Research Initiation Grant.
First author partially supported by National Science Foundation Grant DMS-1265973.
1
2
STEPHAN RAMON GARCIA AND SAMUEL YIH
province of engineers, computer scientists, and applied mathematicians. Conse-
quently, this paper contains a significant amount of exposition since we mean to
bridge a gap between communities that do not often interact. We thank the anony-
mous referee for suggesting several crucial improvements to our exposition.
2. Preliminaries
The main ingredients in this work are the theory of supercharacters and the
discrete Fourier transform (DFT), along with its offspring (the DCT and DST). In
this section, we briefly survey some relevant definitions and ideas.
2.1. Supercharacters. The theory of supercharacters, which extends the classical
character theory of finite groups, was developed axiomatically by Diaconis–Isaacs
[14], building upon earlier important work of Andr´e [2–4]. It has since become an
industry in and of itself. We make no attempt to conduct a proper survey of the
literature on this topic.
Definition 1 (Diaconis–Isaacs [14]). Let G be a finite group, let X be a partition
of the set Irr G of irreducible characters of G, and let K be a partition of G. We
call the ordered pair (X ,K) a supercharacter theory if
(i) K contains {0}, where 0 denotes the identity element of G,
(ii) X = K,
(iii) For each X ∈ X , the function σX =(cid:80)
χ∈X χ(0)χ is constant on each K ∈ K.
The functions σX are supercharacters and the elements K of K are superclasses.
While introduced primarily to study the representation theory of non-abelian
groups whose classical character theory is largely intractable, recent work has re-
vealed that it is profitable to apply supercharacter theory to the most elementary
groups imaginable: finite abelian groups [5, 9, 10, 16, 18, 20, 21, 27].
We outline the approach developed in [9]. Although it is the "one-dimensional"
case that interests us here, there is no harm in discussing things in more general
terms. Let ζ = exp(−2πi/n), which is a primitive nth root of unity. Classical
character theory tells us that the set of irreducible characters of G = (Z/nZ)d is
Irr G = {ψx : x ∈ G},
in which
Here we write
ψx(y) = ζ x·y.
d(cid:88)
x · y :=
xiyi,
i=1
in which x = (x1, x2, . . . , xd) and y = (y1, y2, . . . , yd) are typical elements of G.
Since x · y is computed modulo n it causes no ambiguity in the expression that
defines ψx. We henceforth identify the character ψx with x. Although this identi-
fication is not canonical (it depends upon the choice of ζ), this potential ambiguity
disappears when we construct certain supercharacter theories on G.
Let Γ be a subgroup of GLd(Z/nZ) that is closed under the matrix transpose
operation. If d = 1, then Γ can be any subgroup of the unit group (Z/nZ)×. The
action of Γ partitions G into Γ-orbits; we collect these orbits in the set
K = {K1, K2, . . . , KN}.
SUPERCHARACTERS AND THE DFT, DCT, AND DST
3
For i = 1, 2, . . . , N , we define
σi :=
(cid:88)
x∈Ki
ψx.
The hypothesis that Γ is closed under the transpose operation ensures that σi is
constant on each Ki [9, p. 154] (this condition is automatically satisfied if d = 1).
For i = 1, 2, . . . , N , let Xi = {ψx : x ∈ Ki}. Then
X = {X1, X2, . . . , XN}
is a partition of Irr G and the pair (X ,K) is a supercharacter theory on G.
As an abuse of notation, we identify both the supercharacter and superclass
partitions as {X1, X2, . . . , XN} (such an identification is not always possible with
general supercharacter theories). Since the value of each supercharacter σi is con-
stant on each superclass Xj, we denote this common value by σi(Xj).
Maintaining the preceding notation and conventions, the following theorem links
supercharacter theory on certain abelian groups and combinatorial-flavored matrix
theory [9, Thm. 2].
Theorem 1 (Brumbaugh, et. al., [9]). For each fixed z in Xk, let ci,j,k denote the
number of solutions (xi, yj) ∈ Xi × Xj to x + y = z; this is independent of the
representative z in Xk which is chosen.
(a) For 1 ≤ i, j, k, (cid:96) ≤ N , we have
(b) The matrix
U =
1√
nd
is unitary (U∗ = U−1) and U 4 = I.
(c) The matrices T1, T2, . . . , TN , whose entries are given by
each satisfy TiU = U Di, in which
[Ti]j,k =
,
Di = diag(cid:0)σi(X1), σi(X2), . . . , σi(XN )(cid:1).
(d) Each Ti is normal (T ∗
for the algebra A of all N × N matrices T such that U∗T U is diagonal.
i Ti = TiT ∗
i ) and the set {T1, T2, . . . , TN} forms a basis
The quantities cijk are combinatorial in nature and are nonnegative integers that
relate the values of the supercharacters to each other. Of greater interest to us is
the unitary matrix U defined in (2). It is a normalized "supercharacter table" of
sorts. As in classical character theory, a suitable normalization of the rows and
columns of a character table yields a unitary matrix. This suggests that U encodes
an interesting "transform" of some type. Theorem 1 describes, in a combinatorial
manner, the algebra of matrices that are diagonalized by U .
This is the motivation for our work: we can select G and Γ appropriately so
that U is either the discrete Fourier or discrete cosine transform matrix. Conse-
quently, we can describe the algebra of matrices that are diagonalized by these
σi(X(cid:96))σj(X(cid:96)) =
ci,j,kσk(X(cid:96)).
(cid:34)
k=1
N(cid:88)
σi(Xj)(cid:112)Xj
(cid:112)Xi
(cid:112)Xk
(cid:112)Xj
ci,j,k
(cid:35)N
i,j=1
(2)
(3)
4
STEPHAN RAMON GARCIA AND SAMUEL YIH
transforms. The discrete sine transform can be obtained as a sort of "complement"
to the supercharacter theory corresponding to the DCT. To our knowledge, such
complementary supercharacter theories have not yet been explored in the literature.
2.2. The discrete Fourier transform. It is hallmark of an important theory
that even the simplest applications should be of wide interest. This occurs with the
theory of supercharacters, for its most immediate byproduct is the discrete Fourier
transform (DFT), a staple in engineering and discrete mathematics.
A few words about the discrete Fourier transform are in order. As before, let
G = Z/nZ and ζ = exp(−2πi/n). Let L2(G) denote the complex Hilbert space of
all functions f : G → C, endowed with the inner product
The space L2(G) hosts two familiar orthonormal bases. First of all, there is the
standard basis {δ0, δ1, . . . , δn−1}, which consists of the functions
n−1(cid:88)
j=0
(cid:40)
(cid:104)f, g(cid:105) =
f (j)g(j).
δj(k) =
1
0
if j = k,
if j (cid:54)= k.
We work here modulo n, which explains our preference for the indices 0, 1, . . . , n −
1. A second orthonormal basis of L2(G) is furnished by the exponential basis
{0, 1, . . . , n−1}, in which
The discrete Fourier transform of f ∈ L2(G) is the function (cid:98)f ∈ L2(G) defined by
j(ξ) =
n
.
e2πijξ/n√
n−1(cid:88)
(cid:98)f (ξ) =
1√
n
f (j)e−2πijξ/n = (cid:104)f, ξ(cid:105) .
n so that the map f (cid:55)→ (cid:98)f is a unitary operator from L2(G) to itself. Indeed,
√
The choice of normalization varies from field to field. We have selected the constant
1/
the unitarity of the DFT follows from the fact that
j=0
That is, the DFT is norm-preserving since it sends one orthonormal basis to another.
The matrix representation of the DFT with respect to the standard basis is
(cid:98)j = δj,
1
1
1
...
1
j = 0, 1, . . . , n − 1.
1
ζ
ζ 2
...
ζ n−1
1
ζ 2
ζ 4
...
ζ 2(n−1)
···
···
···
. . .
···
1
ζ n−1
ζ 2(n−1)
...
ζ (n−1)2
.
Fn =
1√
n
(4)
This is the DFT matrix of order n (also called the Fourier matrix of order n).
basis, then a short exercise with finite geometric series reveals that (cid:98)δj = j. A little
If we regard elements of L2(G) as column vectors, with respect to the standard
more work confirms that F 2
n = I. Thus, the eigenvalues of Fn
are among 1,−1, i,−i; the exact multiplicities can be deduced from the evaluation
of the quadratic Gauss sum, which is the trace of
n = −I and hence F 4
√
nFn [6].
SUPERCHARACTERS AND THE DFT, DCT, AND DST
5
There are many compelling reasons why the discrete Fourier transform arises
in both pure and applied mathematics. It would take us too far afield to go into
details, so we content ourselves with mentioning that the DFT arises in signal
processing, number theory (e.g., arithmetic functions), data compression, partial
differential equations, and numerical analysis (e.g., fast integer multiplication). A
particularly fast implementation of the DFT, the fast Fourier transform (FFT),
was named one of the Top 10 algorithms of the 20th century [34]. Although often
credited to Cooley–Tukey (1965) [12], the FFT was originally discovered by Gauss
in 1805 [24]. A valuable reference for all things Fourier-related is [30]. The recent
text of Stein and Shakarchi [33] is a new classic on the subject of Fourier analysis
and it highly recommended for its friendly and understandable approach.
How does the DFT relate to supercharacter theory? Consider the following
example, which was first worked out in [9].
Example 5 (Discrete Fourier transform). Let G = Z/nZ and let Γ = {1}, the triv-
ial subgroup of (Z/nZ)×, act upon G by multiplication. Then the Γ-orbits in G are
singletons: Xj = {j} for j = 0, 1, 2, . . . , n − 1. The corresponding supercharacters
are classical exponential characters:
(cid:88)
x∈Xj
σj(k) =
ζ xk = ζ jk
and hence the unitary matrix U from (2) is the DFT matrix. That is,
U = Fn.
Theorem 1 permits us to identify the matrices that are diagonalized by U . With a
little work, one can show that the matrices (3) are
(cid:40)
0
1
if k − j (cid:54)= i,
if k − j = i,
[Ti]j,k =
and they satisfy TiU = U Di, in which
The algebra A generated by the Ti is the algebra of all N × N circulant matrices
Di = diag(1, ζ i, ζ 2i, . . . , ζ (n−1)i).
c0
c1
...
cN−2
cN−1
cN−1
c0
c1
cN−2
···
cN−1
c0
. . .
···
c2
. . .
. . .
c1
c1
c2
...
cN−1
c0
.
More information about circulant matrices and their properties can be found in [19,
Sect. 12.5].
The preceding example shows that the discrete Fourier transform arises as the
simplest possible application of supercharacter theory. If the action of the trivial
group {1} on Z/nZ already produces items of great interest, it should be fruitful to
consider actions of slightly-less trivial groups as well. This motivates our exploration
of the discrete cosine transform.
6
STEPHAN RAMON GARCIA AND SAMUEL YIH
3. Discrete cosine transform
As before, we fix a positive integer n and let G = Z/nZ. Let
+(G) = {f ∈ L2(G) : f (x) = f (−x) ∀x ∈ G}
L2
and
L2−(G) = {f ∈ L2(G) : f (x) = −f (−x) ∀x ∈ G}
n−1(cid:88)
n−1(cid:88)
denote the subspaces of even and odd functions in L2(G), respectively. Observe
that L2
+(G) is invariant under the DFT, since, if f is even,
(cid:98)f (ξ) = (cid:104)f, ξ(cid:105) =
f (−k)e2πikξ/n = (cid:104)f, −ξ(cid:105) = (cid:98)f (−ξ),
and hence (cid:98)f is even as well. Since L2(G) is finite dimensional and the DFT is
f (j)e−2πijξ/n =
1√
n
j=0
1√
n
k=0
unitary, it follows that L2−(G) = L2
we have the orthogonal decomposition
+(G)⊥ is invariant under the DFT. Consequently,
L2(G) = L2
+(G) ⊕ L2−(G),
in which both subspaces on the right-hand side are DFT-invariant. The discrete
cosine transform (DCT) is the restriction of the DFT to L2
+(G). Being the restric-
tions of a unitary operator (on a finite-dimensional Hilbert space) to an invariant
subspace, the DCT is a unitary operator on L2
+(G). In a similar manner, the dis-
crete sine transform (DST) is the restriction of the DFT to L2−(G).
It too is a
unitary operator.
The DCT is a workhorse in engineering and software applications. The MP3 file
format, which contains compressed audio data, and the JPEG file format, which
contains compressed image data, make use of the DCT [22]. These "lossy" file
formats do not perfectly replicate the original source; that is, some information is
lost. However, by judiciously eliminating high-frequency components in the signal,
one is able to produce sounds or images that are, to human senses, virtually indis-
tinguishable from the source. Moreover, this can be done in such a way that the
final file size is much smaller than the original.
Why is the DCT more prevalent than the DST? Suppose that we have samples
s0, s1, . . . , sm−1 taken at times t = 0, 1, 2 . . . , m − 1. To employ Fourier-analytic
techniques, this signal must be extended to t ∈ Z in a periodic fashion. For many
applications, it behooves the user to make this extension "smooth" in the sense that
there are not large discrepancies between adjacent values. This suggests the use
of a reflection and even boundary conditions; see Figure 1. A standard dictum in
Fourier analysis is that greater smoothness of the input signal translates into more
rapid numerical convergence of associated algorithms. The periodic extension of
the sample that is used by the DCT is naturally "smoother" (for typical real-world
signals) than those utilized by the DFT or DST. Consequently, it is the DCT that
plays a central role in modern signal processing.
There are many subtle variants of "the" DCT that appear in the literature, along
with their multidimensional analogues. Our particular selection is the most suitable
from the viewpoint of supercharacter theory. Indeed, our DCT matrix is precisely
the U -matrix that arises from a particularly simple supercharacter theory on Z/nZ.
SUPERCHARACTERS AND THE DFT, DCT, AND DST
7
Figure 1. (top) Periodic extension of data s0, s1, . . . , s19 (red). The extension
belongs to L2(Z/20Z).
(middle) Periodic extension of the same data, but
with odd boundary conditions. This signal belongs to L2−(Z/40Z). (bottom)
Periodic extension of the same data, but with even boundary conditions at
both edges. This signal belongs to L2
Its "smoothness" suggests
that the DCT may be of more practical use than the DST, or even the DFT,
for the manipulation, storage, or compression of "natural" data.
+(Z/40Z).
Let G = Z/nZ and ζ = exp(−2πi/n). Consider the action of the subgroup
Γ = {±1} of (Z/nZ)× upon G. This produces the orbit decomposition
2}(cid:9) if n is even,
if n is odd.
(cid:8){0}, {±1}, {±2}, . . . ,{ n
(cid:8){0}, {±1}, {±2}, . . . ,{ n±1
2 }(cid:9)
2 ± 1}, { n
X =
-20-1010203040-551015-20-1010203040-15-10-551015-20-10102030-5510158
STEPHAN RAMON GARCIA AND SAMUEL YIH
Let N = X = (cid:98) n
2(cid:99). For j = 0, 1, . . . , N , we define the corresponding superclasses
Xj =
(cid:40){j,−j}
(cid:40)
{j}
if 2j (cid:54)= 0,
if 2j = 0.
ζ jk + ζ−jk
ζ jk
if 2j (cid:54)= 0,
if 2j = 0.
(cid:16) 2πjk
(cid:17)
For j = 0, 1, . . . , N , we have the supercharacters
σj(k) =
Euler's formula tells that
σj(±k) = Xj cos
(7)
in which Xj ∈ {1, 2} is the cardinality of Xj. We index the superclasses starting at
0 rather than 1. Doing so ensures that i ∈ Xi for all i ∈ G, and so we may consider
group elements and indices interchangeably. This convenience is more than enough
to justify what is a small burden of notation.
n
,
In the notation of Theorem 1, we have
√
[U ]j+1,k+1 =
(cid:112)XjXk
cos
n
or more explicitly,
(6)
√
2
√
2
2 cos 2π
n
2 cos 4π
n
...
−√
n
2
2 cos (n−2)π
2 cos 4π
n
2 cos 8π
n
...
√
2
2 cos 2(n−2)π
n
···
···
···
. . .
···
···
U =
1√
n
√
1
√
2
2
...
√
2
1
√
1
√
2
2
...
√
√
2
2
(cid:16) 2πjk
(cid:17)
,
n
√
2
2 cos (n−2)π
2 cos 2(n−2)π
n
n
...
2 −1)2π
2 cos 2( n
√
(−1)
2 −1
2
n
n
√
2
...
(8)
√
2
−√
1
√
2
2
(−1)
...
2 −1
(−1)
n
2
n
√
2
...
2 cos (n−3)π
2 cos 2(n−3)π
n
n
2 cos (n−1)π
2 cos 2(n−1)π
n
n
if n is even and
U =
1√
n
√
2
√
2
2 cos 2π
n
2 cos 4π
n
...
2 cos 4π
n
2 cos 8π
n
...
···
···
···
. . .
···
···
2 cos (n−3)π
2 cos (n−1)π
n
n
2 cos 2(n−3)π
2 cos 2(n−1)π
n
n
2 cos (n−3)2π
2 cos (n−3)(n−1)π
n
n
2 cos (n−3)(n−1)π
2 cos (n−1)2π
n
n
if n is odd. These so-called DCT matrices are real, symmetric, and unitary. They
belong to MN +1, the set of (N + 1) × (N + 1) matrices.
The main result of this section identifies the matrices diagonalized by the DCT
matrix (8). Let cijk denote the number of distinct solutions (x, y) ∈ Xi × Xj to
x + y = z, in which z ∈ Xk is fixed. As stated in Theorem 1, cijk is independent of
the particular representation z ∈ Xk that is chosen.
Theorem 9. Let G = Z/nZ, N = (cid:98) n
transform matrix (8). The matrices T0, T1, . . . , TN ∈ MN +1 defined by
2(cid:99), and let U ∈ MN +1 be the discrete cosine
[Ti]j+1,k+1 =
(10)
(cid:112)Xk
(cid:112)Xj
cijk
SUPERCHARACTERS AND THE DFT, DCT, AND DST
9
form a basis for the algebra A of matrices that are diagonalized by U . They are
real, symmetric, and satisfy
Ti = U DiU∗,
in which
Di = Xi diag(1, cos 2πi
n , cos 4πi
n , . . . , cos 2πN i
n ) ∈ MN +1.
Moreover, T0 = I and Ti generates A if and only if i is relatively prime to n. The
most general matrix T ∈ MN +1 diagonalized by U is
[T ]j,k =
tmin(n−k−j,k−j) + tmin(n−k−j+2,k+j−2)
Xj−1 1
X n
2Xk−1 1
2X n
2 tmin(j−1,k−1)
2 +1−k 1
2 +1−j 1
2 +1−j, n
2 tmax( n
2 + 1,
in which t0, t1, . . . , tN ∈ C are parameters (the last case only occurs if n is even).
2 + 1 or k = n
for j = n
2 +1−k)
for 1 < j, k < n
2 + 1,
for j = 1 or k = 1,
We defer the proof until Section 4. Instead, we focus on several examples.
Example 11. If n is even, then N = n/2 and T is
√
t0
√
2t1
√
2t2
2t3
...
2tN−1
tN
√
√
2t1
t0 + t2
t1 + t3
t2 + t4
...
√
2t2
t1 + t3
t0 + t4
t1 + t5
...
√
tN−2 + tN tN−3 + tN−1
2tN−1
√
2tN−2
√
2t3
t2 + t4
t1 + t5
t0 + t6
...
tN−4 + tN−2
√
2tN−3
···
···
···
···
. . .
···
···
√
2tN−1
tN−2 + tN
tN−3 + tN−1
tN−4 + tN−2
...
√
t0 + t2
2t1
√
√
2t1
t0 + t2
t1 + t3
t2 + t4
Example 12. If n is odd, then N = (cid:98)n/2(cid:99) and T is
···
···
···
···
. . .
···
···
√
t0
√
2t1
√
2t2
2t3
...
√
√
2tN−1
2tN
tN−2 + tN tN−3 + tN−1
tN−1 + tN tN−2 + tN−1
√
2t2
t1 + t3
t0 + t4
t1 + t5
2t3
t2 + t4
t1 + t5
t0 + t6
tN−4 + tN−2
tN−3 + tN−2
...
...
...
√
2tN−1
tN−2 + tN
tN−3 + tN−1
tN−4 + tN−2
...
t0 + t3
t1 + t2
.
√
tN
√
2tN−1
√
2tN−2
2tN−3
...
√
2t1
t0
√
2tN
tN−1 + tN
tN−2 + tN−1
tN−3 + tN−2
...
t1 + t2
t0 + t1
.
The combinatorial aspect of Theorem 9 deserves special attention.
Example 13. If n = 7, then
X0 = {0},
X1 = {1, 6},
and X3 = {3, 4}.
The only solution in X3 × X1 to x + y = 3 is (4, 6). Consequently, (10) produces
The two solutions in X3 × X3 to x + y = 0 are (3, 4) and (4, 3). Thus,
c313
X2 = {2, 5},
(cid:112)X3
(cid:112)X1 = 1.
(cid:112)X0
(cid:112)X3 =
√
c330
2.
[T3]2,4 =
[T3]4,1 =
10
STEPHAN RAMON GARCIA AND SAMUEL YIH
Computing the remaining entries in a similar fashion yields
0
0
√
0
2
0
0
1
1
0
1
1
0
.
√
2
1
0
0
T3 =
Example 14. If n = 8, then
X0 = {0},
and X4 = {4}.
X1 = {1, 7},
The solutions in X3 × X1 to x + y = 4 are (3, 1) and (5, 7). Thus, (10) produces
X3 = {3, 5},
X2 = {2, 6},
The only solution in X3 × X4 to x + y = 1 is (5, 4). Thus,
Computing the remaining entries in a similar fashion yields
Example 15. For n = 10, the most general matrix that is diagonalized by U is
in which t0, t1, t2, t3, t4, t5 are free parameters. It is a linear combination of
0
0
0
√
0
2
0
√
0
2
0
0
0
0
0
0
0
1
√
0
2
0
1
0
1
0
0
0
0
0
0
1
0
0
1
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
1
1
0
0
0
0
0
T0 =
T1 =
(cid:112)X4
(cid:112)X1 =
(cid:112)X1
(cid:112)X4 =
√
2
0
1
0
0
0
0
1
√
0
2
0
1
0
1
0
c314
[T3]2,5 =
c341
[T3]5,2 =
0
0
√
0
2
0
√
2t2
t1 + t3
t0 + t4
t1 + t5
√
t2 + t4
2t3
√
2t3
t2 + t4
t1 + t5
t0 + t4
√
t1 + t3
2t2
T3 =
√
2t1
t0 + t2
t1 + t3
t2 + t4
√
t3 + t5
2t4
√
2.
√
2.
√
0
2
0
0
0
√
.
2t4
t3 + t5
t2 + t4
t1 + t3
√
t0 + t2
2t1
√
2
0
1
0
0
0
,
,
0
1
0
0
√
0
2
0
1
0
1
0
0
0
0
1
0
1
0
√
2
0
1
0
0
0
0
0
√
0
2
0
0
√
0
2
0
0
0
0
T3 =
T5 =
0
0
√
0
2
0
0
0
0
0
0
0
1
0
0
1
0
1
0
0
0
0
0
1
0
0
0
0
1
0
0
0
1
0
0
√
0
2
0
0
1
0
0
0
√
t5
√
2t4
√
2t3
√
2t2
2t1
t0
,
,
0
√
0
2
0
0
0
0
0
1
0
1
0
√
2
0
0
0
1
0
0
1
0
0
0
0
0
1
0
1
0
0
.
1
0
0
0
0
0
√
t0
√
2t1
√
2t2
√
2t3
2t4
t5
0
0
0
1
0
0
√
2
0
0
0
1
0
0
0
1
0
1
0
0
1
0
1
0
0
0
0
0
1
√
0
2
0
√
0
2
0
0
0
0
0
0
√
0
2
0
T2 =
T4 =
SUPERCHARACTERS AND THE DFT, DCT, AND DST
11
Example 16. For n = 11, the most general matrix that is diagonalized by U is
√
t0
√
2t1
√
2t2
√
2t3
√
2t4
2t5
T =
√
2t1
t0 + t2
t1 + t3
t2 + t4
t3 + t5
t4 + t5
√
2t2
t1 + t3
t0 + t4
t1 + t5
t2 + t5
t3 + t4
√
2t3
t2 + t4
t1 + t5
t0 + t5
t1 + t4
t2 + t3
√
2t4
t3 + t5
t2 + t5
t1 + t4
t0 + t3
t1 + t2
√
2t5
t4 + t5
t3 + t4
t2 + t3
t1 + t2
t0 + t1
in which t0, t1, t2, t3, t4, t5 are free parameters. It is a linear combination of
,
0
0
0
0
0
1
1
0
0
0
0
0
0
√
0
2
0
0
0
0
0
0
√
0
2
0
T0 =
T2 =
T4 =
0
1
0
0
0
0
0
0
1
0
0
0
0
1
0
1
0
0
0
0
0
1
0
1
0
0
0
0
1
0
0
1
0
0
0
1
0
0
0
1
0
0
√
2
0
0
0
1
0
0
0
1
0
0
1
0
1
0
0
1
0
0
0
1
0
0
1
√
2
0
0
1
0
0
,
,
0
0
0
1
1
0
0
1
1
0
0
0
√
0
2
0
0
0
0
0
0
√
0
2
0
0
0
0
0
0
√
0
2
,
,
.
√
2
0
1
0
0
0
0
0
1
0
1
0
0
0
0
0
1
1
0
1
0
0
0
1
0
0
0
1
1
0
0
1
0
1
0
0
0
0
1
0
1
0
√
2
0
0
0
0
1
0
0
0
1
0
1
0
1
0
0
1
0
0
0
1
1
0
0
0
1
1
0
0
0
0
0
0
0
1
1
0
0
1
1
0
0
√
2
1
0
0
0
0
T1 =
T3 =
T5 =
The matrices above are analogous to those encountered by Feig and Ben-Or [17],
who considered the modified DCT matrix
in which ci =(cid:112)1/n for i = 1 and(cid:112)2/n otherwise.
[Cn]i,j = ci cos
4n
2π(2j − 1)(i − 1)
,
Example 17. Matrices diagonalized by the DCT have been studied before, but
with different techniques and sometimes with different DCT matrices [17,31]. The-
orem 9 recovers many established results. For example, the matrix
√
2
1/
. . .
0
√
2
1/
0
...
...
0
1/2
. . .
···
,
0
1/2
. . .
. . .
. . .
···
···
. . .
. . .
. . .
1/2
0
···
. . .
0
...
...
1/2
. . .
√
1/
2
0
√
2
1/
0
appears in [31]. In our notation, it corresponds to even n and parameters t0 = 0,
t1 = 1/2, and t2 = t3 = ··· = tN = 0; see Example 11.
12
STEPHAN RAMON GARCIA AND SAMUEL YIH
Example 18. For n odd, the bottom right N × N submatrix of any matrix diag-
onalized by U is a Toeplitz plus Hankel matrix:
+
t0
t1
...
tN−1
t1
. . .
. . .
···
···
. . .
. . .
t1
tN−1
...
t1
t0
···
t2
...
. . .
. . .
tN
tN tN−1
tN
. . .
. . .
···
tN
tN−1
...
t1
An analogous presentation exists when n is even if we also exclude the first and last
row and column. In [31] it is shown that the DCT-I matrix, obtained by replacing
Xk1/2 and Xj1/2 with Xk and Xj, respectively, in (8), diagonalizes matrices
that are genuinely Toeplitz plus Hankel. In [23] Grishin and Strohmer demonstrate
that it is simple to go from the DCT-I to U , and that there are advantages to both
matrices. While the DCT-I diagonalizes certain Toeplitz plus Hankel matrices, it
is not unitary like U .
Theorem 9 also provides an explanation for this Toeplitz plus Hankel structure.
The matrix entry [Ti]j+1,k+1 = cijk is nonzero if and only if ±i ± j = k, or equiva-
lently, when i ∈ Xk−j or i ∈ Xk+j. If i ∈ Xk−j, then i ∈ X(k+(cid:96))−(j+(cid:96)) for any (cid:96) ∈ G.
Thus, along the diagonal that contains (j + 1, k + 1), Ti is always nonzero; this gives
us one sub- or super-diagonal of a Toeplitz matrix. Similarly, if i ∈ Xk−j, then Ti
is nonzero along the entire anti-diagonal containing (j, k), giving us a component
of a Hankel matrix. See [13] for a displacement-rank approach to such matrices.
4. Proof of Theorem 9
Let A denote the commutative, complex algebra of matrices that are diagonalized
by U . The algebra of (N + 1) × (N + 1) diagonal matrices has dimension N + 1.
Thus, dimA = N + 1. The diagonal matrices D0, D1, D2, . . . , DN are linearly
independent because their diagonals
n cos 4πi
n ]T ∈ CN +1
n . . . cos 2πN i
are the columns of the matrix [σi−1(j − 1)]N +1
matrix U . Thus, {T0, T1, . . . , TN} is linearly independent and hence it spans A.
i,j=1, which is similar to the unitary
[1 cos 2πi
The eigenvalues
1,
cos 2πi
n ,
cos 4πi
n , . . . ,
cos 2πN i
n
of Ti are distinct if and only if i is relatively prime to n. In this case, the Lagrange
interpolation theorem ensures that for any diagonal matrix D ∈ MN +1, there is a
polynomial p so that p(Ti) = U DU∗. Thus, Ti generates A.
We claim that T0 = I. If i = 0, then Xi = {0}. Consequently, x + y ∈ Xk and
(x, y) ∈ Xi × Xj and imply j = k; moreover, cijj = 1. Thus, T0 = I.
We now consider Ti for i = 1, 2, . . . , N and identify the locations of all nonzero
entries in each matrix. First suppose that n is odd (if n is even then there are a
few additional cases to consider; we will do this later).
If j = 0, then the argument above implies that i = k. Since n is odd, −i = k
means i = 0. Thus, cijk = 1 and, since k = i (cid:54)= 0, we have Xk = 2. By symmetry,
[Ti]i+1,1 = [Ti]1,i+1 =
2.
√
An analogous approach applies if k = 0. In all other cases, i, j, k are nonzero and
hence Xj = Xk = 2.
SUPERCHARACTERS AND THE DFT, DCT, AND DST
13
(i) Suppose that cijk = 2. Without loss of generality, let (i, j) be one of the
solutions to x + y = k with (x, y) ∈ Xi × Xj. The other potential solution
must be one of (i,−j), (−i,−j), or (−i, j). These possibilities imply that
2j = 0, 2k = 0, or 2i = 0, respectively. Since i, j, k (cid:54)= 0, this is not possible.
(ii) Suppose that cijk = 1, with (i, j) as the solution. We see that ±i ± j = k if
and only if i ∈ Xj+k or i ∈ Xk−j. For such i,
Since cijk ∈ {0, 1, 2}, it follows that Ti is 0 elsewhere.
[Ti]j+1,k+1 = 1.
If T ∈ A, then T =(cid:80)N
implies that T equals
√
t0
√
2t1
√
2t2
2t3
...
√
√
2tN−1
2tN
i=0 tiTi for some t0, t1, . . . , tN ∈ C. The preceding analysis
√
2t1
t0 + t2
t−1 + t3
t−2 + t4
...
√
2t2
t1 + t3
t0 + t4
t−1 + t5
...
√
2t3
t2 + t4
t1 + t5
t0 + t6
...
t2−N + tN
t1−N + tN +1
t3−N + tN +1
t2−N + tN +2
t4−N + tN +2
t3−N + tN +3
· · ·
· · ·
· · ·
· · ·
. . .
· · ·
· · ·
√
2tN−1
tN−2 + tN
tN−3 + tN +1
tN−4 + tN +3
...
t0 + t2N−2
t−1 + t2N−1
√
2tN
tN−1 + tN +1
tN−2 + tN +2
tN−3 + tN +3
√
...
2t1
t0 + t2N
in which, for the sake of convenience, we let ti = t−i = tn−i for all i. The preceding
simplifies to the matrix presented in Example 12.
Now suppose that n is even. The preceding results largely carry over, but there
are now extra cases to consider.
(iii) Suppose that i = N = n
2 . Then the only solutions to x + y ∈ Xk with
(x, y) ∈ Xi × Xj are when, without loss of generality, k = N − j. Since
Xj = XN−j for all j, an appeal to (10) reveals that TN is the reversed
identity matrix.
(iv) Suppose that i (cid:54)= 0, i (cid:54)= N , and cijk = 2. In addition to the cases identified in
(ii), we now have the possibilities j = N or k = N . From (10) we obtain
√
[Ti]N +1,N−i+1 = [Ti]N−i+1,N +1 =
2.
If T ∈ A, then T = (cid:80)N
In all other cases, i, j, k (cid:54)∈ {0, N}, so the rest of our analysis from the odd case
i=0 tiTi for some t0, t1, . . . , tN ∈ C. The
carries over.
preceding analysis implies that T equals
√
√
√
√
2t1
t0 + t2
t−1 + t3
t−2 + t4
2t2
t1 + t3
t0 + t4
t−1 + t5
2t3
t2 + t4
t1 + t5
t0 + t6
...
...
...
√
t2−N + tN t3−N + tN−1
√
2tN−1
2tN−2
t4−N + tN−2
√
2tN−3
· · ·
· · ·
· · ·
· · ·
. . .
· · ·
· · ·
2tN−1
tN−2 + tN
tN−3 + tN−1
tN−4 + tN−2
...
t0 + t2N−2
√
2t1
√
√
√
tN
2tN−1
2tN +2
2tN +3
...
√
2t1
t0
√
t0
√
2t1
√
2t2
2t3
...
2tN−1
tN
√
.
This simplifies to the matrix presented in Example 11.
We can now obtain an explicit formula for the entries of the most general matrix
T = [Tj,k] ∈ MN +1 diagonalized by U . The first row and column (and the last row
and column if n is even) have a different structure from the rest of the matrix; one
14
STEPHAN RAMON GARCIA AND SAMUEL YIH
can see that
[T ]j,k =
Xj−1 1
X n
2Xk−1 1
2X n
2 +1−j 1
2 tmin(j−1,k−1)
2 +1−k 1
2 tmax( n
2 +1−j, n
2 +1−k)
for j = 1 or k = 1,
for j = n
2 + 1 or k = n
2 + 1,
holds for these sections of T ; the second case occurs only if n is even.
We direct our attention now to the remaining entries. First observe that
[T ]j,k = tk−1−(j−1) + tk−1+j−1 = tk−j + tk+j−2
To ensure that all of our subscripts are between 0 and n − 1, we take the absolute
2 , it follows that 0 ≤ k + j − 2 ≤
value of the first subscript. Since 1 < j, k < N < n
n − 1. Thus,
Finally, we need to ensure that our subscripts are between 0 and N . If N + 1 ≤ i ≤
n − 1, then n − i gives the correct index and is in the desired range. Consequently,
[T ]j,k = tk−j + tk+j−2.
[T ]j,k = tmin(n−k−j,k−j) + tmin(n−k−j+2,k+j−2). (cid:3)
5. Discrete sine transform
The discrete sine transform (DST) is the oft-neglected sibling of the DCT. Since
L2(G) is finite dimensional and L2−(G) = L2
it follows from the DFT-
+(G) that L2−(G) is also DFT-invariant (recall that the DFT is
invariance of L2
a unitary operator). As mentioned in Section 3, the DST is the restriction of the
DFT to L2−(G), the subspace of odd functions in L2(G). Here G = Z/nZ, as usual.
Let
+(G)⊥,
(cid:106) n − 1
4
(cid:107)
.
2
N =
As before, the sets Xj = {j,−j} for j = 1, 2, . . . , N , along with {0} (and {n/2} if
n is even), partition G.
In our consideration of the DCT, we saw that the supercharacters (7) are constant
on each superclass. In contrast, the corresponding "supercharacters" obtained by
replacing cosines with sines are no longer constant on each superclass. This is a
crucial distinction between the DCT and DST: the DST does not arise directly
from a supercharacter theory on G. Nevertheless, we are still able to obtain an
analogue of Theorem 9 for the DST by appealing to the DFT-invariance of L2−(G)
and considering the "orthogonal complement" of the DCT supercharacter theory.
j=1 is an orthogonal
Define τj(k) = ζ−jk − ζ jk for j = 0, 1, . . . , N . Then {τj}N
basis for L2−(G) and
τj(k) = −2i sin
(19)
in which i denotes the imaginary unit. Normalizing the τj yields
vj(k) =
τj(k)√
2n
=
2 sin( 2πjk
n )
√
i
n
.
Let Vn ∈ MN denote the matrix representation of the restriction of F to L2−(G) with
respect to the orthonormal basis {vj}N
j=1. Then Vn is unitary and a computation
confirms that
√
2
i
(cid:16) 2πjk
[Vn]j,k =
(cid:17)
(20)
sin
n
n
.
(cid:16) 2πjk
(cid:17)
,
n
√
SUPERCHARACTERS AND THE DFT, DCT, AND DST
15
Thus,
Vn =
√
2
i
n
sin 2π
n
sin 4π
n
...
sin 4π
n
sin 8π
n
...
sin 2N π
n
sin 4N π
n
···
···
. . .
···
sin 2N π
n
sin 4N π
n
...
sin 2N 2π
n
.
The matrices Vn are purely imaginary, complex symmetric, and unitary. If n is clear
from context, we often omit the subscript and write V . Although the DST cannot
be attacked directly via supercharacter theory, we can use the DFT invariance of
L2−(G) to obtain a satisfying analogue of Theorem 9.
Theorem 21. Let G = Z/nZ, N = (cid:98) n− 1
transform matrix (20). Let sgn x denote the sign of x; let sgn 0 = 0.
(a) The most general S ∈ MN diagonalized by V is given by
2 (cid:99), and let V ∈ MN be the discrete sine
4
sgn(cid:0) n
2 − k − j − 2(cid:96) + 1(cid:1)smin(n−k−j−2(cid:96)+1,k−j+2(cid:96)−1),
(22)
min(j,k)(cid:88)
(cid:96)=1
[S]j,k =
in which s0 = 0, and s1, s2, . . . , sN ∈ C are free parameters that correspond,
in that order, to the entries in the first row of S. For i = 1, 2, . . . , N , the
matrices Si obtained by setting sj = δi,j in (22) form a basis for the algebra
A diagonalized by V . In particular, S1 = I.
(b) Let xi,j = 1 if j ∈ Xi and 0 otherwise. The matrices T1, T2, . . . , TN ∈ MN
defined by
[Ti]j,k = xi,j−k − xi,j+k
(23)
are real, symmetric, and satisfy
Ti = V DiV ∗,
in which
Di = 2 diag(cid:0) cos 2πi
n , cos 4πi
n , . . . , cos 2πN i
n
(cid:1) ∈ MN .
Moreover, Ti generates A if and only i is relative prime to n.
(c) If n is odd, then {T1, T2, . . . , TN} is a basis for A. Another formula for the
entries for a general T ∈ A is given by
[Ti]j,k = tmin(n−j−k,j+k) − tmin(n−j+k,j−k),
in which t0 = 0, and t1, t2, . . . , tN ∈ C are free parameters.
(24)
For odd n, Theorem 21 provides two bases for A. The basis described in (a) is
obtained by a brute force method which, if applied to the DCT, yields the basis
in Theorem 9. However, it is cumbersome to work with; the following examples
illustrate its inelegance and unwieldiness. The basis obtained in (b) is superior
in several ways. Not only is it much simpler in appearance, it also has a nice
combinatorial explanation.
The matrices given by (22) and (24) are easier to grasp with examples. We defer
the proof of Theorem 21 until Section 6 and focus on some instructive examples.
16
STEPHAN RAMON GARCIA AND SAMUEL YIH
Example 25. If n is even, the most general matrix diagonalized by Vn is
s1
s2
s3
...
sN−1
sN
s2
s1 + s3
s2 + s4
...
s3
s2 + s4
s1 + s3 + s5
...
sN−2 + sN sN−3 + sN−1
sN−1
sN−2
···
···
···
. . .
···
···
sN−1
sN−2 + sN
sN−3 + sN−1
sN
sN−1
sN−2
...
s1 + s3
s2
...
s2
s1
in which s1, s2, . . . , sN are free parameters. Bini and Capovani were the first to call
the matrix above a T -class matrix, and referred to the algebra A as TN . This class
of matrices occurs in the study of Toeplitz matrices and is known to be diagonalized
by our DST matrix [7]. We recapture this result, and with our method we are able
to find an analogous basis for the case where n is odd, which has been much less
studied. These matrices also form a subspace of the Toeplitz plus Hankel matrices
[8]. There is a considerable amount of literature on T -class matrices because of
their desirable computational properties. For instance, a TN matrix system can
be solved in O(N log N ) time using algorithms for centrosymmetric Toeplitz plus
Hankel matrices [25]. This makes T -class matrices suitable as preconditioners for
banded Toeplitz systems [8, 11, 28].
From [28], T -class matrices may also be defined as the N × N matrices A =
i,j=1 whose entries satisfy the "cross-sum" condition
ai−1,j + ai+1,j = ai,j−1 + ai,j+1,
[aij]N
(26)
in which aN +1,j = ai,N +1 = a0,j = ai,0 = 0.
Example 27. If n is odd, the most general matrix that is diagonalized by Vn is
sN−2 + sN
sN−1
sN
sN−1 + sN +1 + sN +1
sN−3 + sN−1 + sN +1
sN−2 + sN + sN +2
j=1 s2j
j=1 s2j−1
in which s1, s2, . . . , sN are free parameters, and si = −sn−i. A glance at Example
25 confirms that the even and odd cases are strikingly different. Because of this
unexpected complexity, the odd case, as mentioned in the preceding example, does
not appear to have been addressed completely in the literature before.
j=1 s2j−1
j=1 s2j
However, these matrices enjoy many of the same properties T matrices do; they
are Toeplitz plus Hankel, symmetric, and diagonalized by the DST matrix (20).
Further, the same equation (22) used to obtain these matrices recovers the T
matrices if n is even, so we may consider (22) as providing a generalization of T
matrices. Using (24), a more transparent description is
t2
t3 − t1
t4 − t2
...
tN − tN−2
tN − tN−1
t3 − t1
t5 − t1
t4
...
tN − tN−3
tN−1 − tN−2
t4 − t2
t5 − t1
t6
...
tN−1 − tN−4
tN−2 − tN−3
···
···
···
. . .
···
···
tN − tN−2
tN − tN−3
tN−1 − tN−4
tN − tN−1
tN−1 − tN−2
tN−2 − tN−3
...
t3
t2 − t1
...
t2 − t1
t1
T =
s1
s2
...
...
s2
s1 + s3
...
...
s3
s2 + s4
. . .
...
sN−1
sN
sN−2 + sN
sN−1 + sN +1
···
···
...
. . .
...
...
... (cid:80)N−1
··· (cid:80)N−1
...
...
(cid:80)N−1
(cid:80)N
SUPERCHARACTERS AND THE DFT, DCT, AND DST
17
in which t1, t2, . . . , tN are free parameters. From this parameterization we see these
matrices even almost satisfy (26), failing to hold only at the right edge. For instance,
considering the (2, N ) entry,
[T ]1,N + T3,N = tN − tN−1 + tN−2 − tN−3 (cid:54)= [T ]2,N−1 + [T ]2,N +1 = tN − tN−3
since the cross-sum condition takes [T ]2,N +1 = 0.
Example 28. For n = 11, the most general matrix diagonalized by Vn is
in which s1, s2, s3, s4, s5 ∈ C are free parameters. It is a linear combination of
1
0
0
0
0
0
1
0
0
0
S1 =
s1
s2
s3
s4
s5
s2
s1 + s3
s2 + s4
s3 + s5
s4 − s5
s3
s2 + s4
s1 + s3 + s5
s2 + s4 − s5
s3 + s5 − s4
, S2 =
0
0
0
1
0
0
0
0
0
1
0
0
1
0
0
1
0
1
1 −1
1
0
1 −1
0
1
1 −1
1 −1
0
1
0
0
0
,
0
0
1
0
0
0
0
0
1
0
S4 =
and S5 =
s4
s3 + s5
s2 + s4 − s5
s1 + s3 + s5 − s4
s2 + s4 − s5 − s3
s5
s4 − s5
s3 + s5 − s4
s2 + s4 − s5 − s3
s1 + s3 + s5 − s4 − s2
.
(29)
,
0
1
0
1
0
1
0
0
0
1
0
1
0
1
1 −1
0
1 −1
1
0
0
1
0
0
, S3 =
0
0
0
0
1
1 −1
0
0
0
1 −1
0
0
1
1 −1
1 −1
0
1 −1
1 −1
1
1
0
1
0
0
0
1
0
1
0
0
0
0
0
1
0
0
1
1 −1
It is apparent each Si is Toeplitz plus Hankel; hence (29) is Toeplitz plus Hankel
as well. Using (24) we obtain the alternate parametrization
t5 − t4
t4 − t3
t3 − t2
t2 − t1
t2
t4
t3 − t1
t5 − t1
t5 − t2
t4 − t3
t4 − t2
t5 − t1
t4 − t1
t3 − t2
t5
t5 − t3
t5 − t2
t4 − t1
t2 − t1
t3
t3 − t1
t4 − t2
t5 − t3
t5 − t4
t1
in which t1, t2, t3, t4, t5 ∈ C are free parameters. It is a linear combination of
T1 =
0 −1
−1
0 −1
0
0
0
0
0
0 −1
0
0
0 −1
0
0 −1
0 −1
0 −1
1
0
,
0
0
1
0
−1
, T4 =
0 −1
1
0
0
0 −1
0
0
0
−1
0 −1
0
0
0 −1
1
0
0
0 −1
0
1
0
0 −1
1
0
0
1
0
0
0
0
, T5 =
0
0
0
1
1
0
0
1
1
0
0
1
1
0
0
T2 =
0
1
0
0
1
1
0
0
1
0
,
1
1
0
0
0
.
1
0
0
0
0
T3 =
0
1
1
0
0
0
−1
0
0 −1
0 −1
0
0
0
1
0
0 −1
1
0
1
0
0
0
This example highlights some of the advantages of working with either of the two
bases. The S-basis is analogous to the most natural basis for the T matrices, and
in particular S1 = I. However, the T -basis matrices tend to be sparser and can be
computed with purely combinatorial arguments.
18
STEPHAN RAMON GARCIA AND SAMUEL YIH
6. Proof of Theorem 21
(a) Let G = Z/nZ and N = (cid:98)(n − 1
denote the discrete sine transform matrix corresponding to the modulus n. For
j = 1, 2, . . . , N , define the diagonal matrices
4 )/2(cid:99) = dim L2−(G), and let V = Vn ∈ MN
Cj = diag(cid:0)τj(1), τj(2), . . . , τj(N )(cid:1) ∈ MN .
These matrices are linearly independent because their diagonals are scalar multiples
of the rows of the unitary matrix V . Thus, {V CjV ∗}N
j=1 is a basis for A.
The entries of V CjV ∗ are
[V CjV ∗]k,(cid:96) =
1
n
τj(m)τk(m)τ(cid:96)(m).
For supercharacter theories like that for the DCT and discussed in [9], resolving the
analogous quantity exploited supercharacter invariance on superclasses to simplify
the preceding into an inner product (cid:104)σjσk, σ(cid:96)(cid:105). We do not enjoy such a simplification
but we do have the identity
τj(x)τk(x) + τ1(x)τj+k+1(x) = τj+1(x)τk+1(x)
(30)
for all j, k, x ∈ G. Define
sj,k =
1
n
τj((cid:96))τ1((cid:96))τk((cid:96))
so that [sj,1 sj,2 . . . sj,N ] is the first row of V CjV ∗. Then by (30) we may rewrite
N(cid:88)
m=1
N(cid:88)
(cid:96)=1
N(cid:88)
(cid:16) N(cid:88)
m=1
1
n
[V CjV ∗]k+1,(cid:96)+1 =
1
n
τj(m)τk+1(m)τ(cid:96)+1(m)
N(cid:88)
(cid:17)
τj(m)τ1(m)τk+(cid:96)+1(m)
τj(m)τk(m)τ(cid:96)(m) +
=
= [V CjV ∗]k,(cid:96) + sj,k+(cid:96)+1.
m=1
m=1
s1
s2
...
Because τj = −τ−j for all j, we have sj,k = −sj,−k for all k. This condition forces
t0 = 0, and also tn/2 = 0 if n is even. Furthermore, V CjV ∗ is uniquely determined
by its first row. Since this holds for all j, any matrix in the span of these matrices
must enjoy the same relation among its entries. If [s1 s2 . . . sN ] is the first row of
some matrix in A, then that matrix is
s3
sN−1
sN
s2
· · ·
· · ·
. . .
sN−2 + sN
...
· · · (cid:80)N−1
· · · (cid:80)N−1
j=1 s2j−1
j=1 s2j
sN−1 + sN +1
...
(cid:80)N−1
(cid:80)N
j=1 s2j
j=1 s2j−1
s1 + s3
...
s2 + s4
...
sN−2 + sN
sN−1
sN
sN−3 + sN−1 + sN +1
sN−2 + sN + sN +2
in which we adopt the convention si = −sn−i.
sN−1 + sN +1 + sN +1
For each S ∈ A and some 1 < j, k ≤ N , we have
[S]j,k − sj+k−1 = [S]j−1,k−1.
Repeat this min(j, k) − 1 times, until j = 1 or k = 1. The other subscript will be
max(j, k) − (min(j, k) − 1) = max(j, k) − min(j, k) + 1 = k − j + 1.
SUPERCHARACTERS AND THE DFT, DCT, AND DST
19
From this starting subscript, going down the diagonal we increase the row and
column subscript simultaneously by 1 each time, hence increasing the subscript of
s by 2 in the summation:
min(j,k)(cid:88)
(cid:96)=1
min(j,k)(cid:88)
(cid:96)=1
[S]j,k =
sk−j+1+2((cid:96)−1) =
sk−j+2(cid:96)−1.
We must ensure that all subscripts are in {1, 2, . . . , N}. Since s(cid:96) = −s−(cid:96), we
2 . To achieve the former, an
reverse the sign of the s with indices larger than n
argument similar to that in the proof of Theorem 9 permits us to use the index
min(n − k − j − 2(cid:96) + 1,k − j + 2(cid:96) − 1).
For the latter, note that the proper sign of the term is the same as
sgn( n
2 − k − j + 2(cid:96) − 1).
since the sign is simply dependent on whether the index is larger than n
2 . Hence,
[S]j,k =
sgn( n
2 − k − j − 2(cid:96) + 1)smin(n−k−j−2(cid:96)+1,k−j+2(cid:96)−1).
min(j,k)(cid:88)
(cid:96)=1
(b) Let T1, T2, . . . , TN and D1, D2, . . . , DN be defined as in the statement of Theo-
rem 21. Let σj be as defined in (6) of the DCT section and note that
Di = diag(cid:0)σi(1), σi(2), . . . , σi(N )(cid:1) ∈ MN .
Since σ is real valued and V is symmetric,
[V DiV ∗]j,k =
1
n
τj((cid:96))σi((cid:96))τk((cid:96)) =
1
n
τj((cid:96))τk((cid:96))σi((cid:96)).
N(cid:88)
(cid:96)=1
N(cid:88)
(cid:96)=1
Here we may actually make a substantial simplification, since the product of two
odd functions is constant on each Xj. Hence we may rewrite this as an inner
product in L2
2 if n is even), then τj(x) = 0 and so
+(G). If x = 0 (and x = n
τj((cid:96))τk((cid:96))σi((cid:96)) =
1
2n
τj((cid:96))τk((cid:96))σi((cid:96)) =
(cid:104)τjτk, σi(cid:105).
1
2n
N(cid:88)
(cid:96)=1
1
n
Further,
(cid:88)
x∈G
τj(x)τk(x) = ζ (j−k)x + ζ (k−j)x − ζ (j+k)x − ζ−(j+k)x
for all j, k, x ∈ G. Consequently,
=
2
Xj−k σj−k(x) − 2
Xj+k σj+k(x)
[V DiV ∗]j,k =
1
nXj−k(cid:104)σj−k, σi(cid:105) −
1
nXj+k(cid:104)σj+k, σi(cid:105)
Since σ1, σ2, . . . , σN are orthogonal, we use the fact that (cid:107)σj(cid:107)2 = nXj to get (23).
Each Ti matrix with i relatively prime to n generates A again by an appeal to the
Lagrange interpolation theorem, as in the proof of Theorem 9.
(c) Suppose n is odd. By (23), we have [Ti]j,j = xi,0 − xi,2j for j = 1, 2, . . . , N .
Hence each Ti is nonzero along the main diagonal only if i ∈ X0 or i ∈ X2j. Since i
ranges from 1 to N , it follows that each Ti is zero along the main diagonal except at
20
STEPHAN RAMON GARCIA AND SAMUEL YIH
For some T =(cid:80)N
the 2ith index, in which 2 denotes the multiplicative inverse of 2 modulo n. Hence
Ti is the only matrix in {T1, T2, . . . , TN} that does not vanish at the (2i, 2i) entry.
Thus, {T1, T2, . . . , TN} is linearly independent and hence it is a basis for A.
i=1 tiTi, observe that Ti is nonzero precisely at the (j, k) entries
for which j + k ∈ Xi or j − k ∈ Xi.
If we agree that ti = t−i = tn−i, then
[T ]j,k = tj+k − tj−k . The techniques used in the proof of Theorem 9 to relabel the
indices so that the subscripts lie in {1, 2, . . . , N} can be used to obtain (24).
(cid:3)
References
[1] Marcelo Aguiar, Carlos Andr´e, Carolina Benedetti, Nantel Bergeron, Zhi Chen, Persi Diaco-
nis, Anders Hendrickson, Samuel Hsiao, I. Martin Isaacs, Andrea Jedwab, Kenneth Johnson,
Gizem Karaali, Aaron Lauve, Tung Le, Stephen Lewis, Huilan Li, Kay Magaard, Eric Mar-
berg, Jean-Christophe Novelli, Amy Pang, Franco Saliola, Lenny Tevlin, Jean-Yves Thibon,
Nathaniel Thiem, Vidya Venkateswaran, C. Ryan Vinroot, Ning Yan, and Mike Zabrocki, Su-
percharacters, symmetric functions in noncommuting variables, and related Hopf algebras,
Adv. Math. 229 (2012), no. 4, 2310–2337. MR 2880223
[2] Carlos A. M. Andr´e, The basic character table of the unitriangular group, J. Algebra 241
(2001), no. 1, 437–471. MR MR1839342 (2002e:20082)
[3]
, Basic characters of the unitriangular group (for arbitrary primes), Proc. Amer.
Math. Soc. 130 (2002), no. 7, 1943–1954 (electronic). MR MR1896026 (2003g:20075)
[4] Carlos A.M. Andr´e, Basic characters of the unitriangular group, J. Algebra 175 (1995), no. 1,
287–319. MR MR1338979 (96h:20081a)
[5] Samuel G. Benidt, William R. S. Hall, and Anders O. F. Hendrickson, Upper and lower
semimodularity of the supercharacter theory lattices of cyclic groups, Comm. Algebra 42
(2014), no. 3, 1123–1135. MR 3169622
[6] Bruce C. Berndt, Ronald J. Evans, and Kenneth S. Williams, Gauss and Jacobi sums, Cana-
dian Mathematical Society Series of Monographs and Advanced Texts, John Wiley & Sons,
Inc., New York, 1998, A Wiley-Interscience Publication. MR 1625181
[7] Dario Bini and Milvio Capovani, Spectral and computational properties of band symmetric
Toeplitz matrices, Linear Algebra Appl. 52/53 (1983), 99–126. MR 709346
[8] E. Boman and I. Koltracht, Fast transform based preconditioners for Toeplitz equations,
SIAM J. Matrix Anal. Appl. 16 (1995), no. 2, 628–645. MR 1321801
[9] J. L. Brumbaugh, Madeleine Bulkow, Patrick S. Fleming, Luis Alberto Garcia German,
Stephan Ramon Garcia, Gizem Karaali, Matt Michal, Andrew P. Turner, and Hong Suh,
Supercharacters, exponential sums, and the uncertainty principle, J. Number Theory 144
(2014), 151–175. MR 3239156
[10] J. L. Brumbaugh, Madeleine Bulkow, Luis Alberto Garcia German, Stephan Ramon Garcia,
Matt Michal, and Andrew P. Turner, The graphic nature of the symmetric group, Exp. Math.
22 (2013), no. 4, 421–442. MR 3171103
[11] Raymond H. Chan, Michael K. Ng, and C. K. Wong, Sine transform based preconditioners
for symmetric Toeplitz systems, Linear Algebra Appl. 232 (1996), 237–259. MR 1366587
[12] James W. Cooley and John W. Tukey, An algorithm for the machine calculation of complex
Fourier series, Math. Comp. 19 (1965), 297–301. MR 0178586
[13] Carmine Di Fiore and Paolo Zellini, Matrix decompositions using displacement rank
and classes of commutative matrix algebras, Linear Algebra Appl. 229 (1995), 49–99.
MR 1352839
[14] Persi Diaconis and I. M. Isaacs, Supercharacters and superclasses for algebra groups, Trans.
Amer. Math. Soc. 360 (2008), no. 5, 2359–2392. MR 2373317
[15] Persi Diaconis and Nathaniel Thiem, Supercharacter formulas for pattern groups, Trans.
Amer. Math. Soc. 361 (2009), no. 7, 3501–3533. MR MR2491890 (2010g:20013)
[16] William Duke, Stephan Ramon Garcia, and Bob Lutz, The graphic nature of Gaussian peri-
ods, Proc. Amer. Math. Soc. 143 (2015), no. 5, 1849–1863. MR 3314096
[17] Ephraim Feig and Michael Ben-Or, On algebras related to the discrete cosine transform,
Linear Algebra Appl. 266 (1997), 81–106. MR 1473195
[18] Christopher F. Fowler, Stephan Ramon Garcia, and Gizem Karaali, Ramanujan sums as
supercharacters, Ramanujan J. 35 (2014), no. 2, 205–241. MR 3266478
SUPERCHARACTERS AND THE DFT, DCT, AND DST
21
[19] Stephan Ramon Garcia and Roger A. Horn, A second course in linear algebra, Cambridge
Mathematical Textbooks, Cambridge University Press, New York, NY, 2017.
[20] Stephan Ramon Garcia, Trevor Hyde, and Bob Lutz, Gauss's hidden menagerie: from cyclo-
tomy to supercharacters, Notices Amer. Math. Soc. 62 (2015), no. 8, 878–888. MR 3379072
[21] Stephan Ramon Garcia and Bob Lutz, A supercharacter approach to Heilbronn sums, J.
Number Theory, in press.
[22] Rafael C. Gonzalez and Richard E. Woods, Digital image processing, Pearson, 2017, Fourth
Edition.
[23] Denis Grishin and Thomas Strohmer, Fast scattered data approximation with Neumann and
other boundary conditions, Linear Algebra Appl. 391 (2004), 99–123. MR 2094605
[24] Michael T. Heideman, Don H. Johnson, and C. Sidney Burrus, Gauss and the history of the
fast Fourier transform, Arch. Hist. Exact Sci. 34 (1985), no. 3, 265–277. MR 815154
[25] Georg Heinig, Chebyshev-Hankel matrices and the splitting approach for centrosymmet-
ric Toeplitz-plus-Hankel matrices, Linear Algebra Appl. 327 (2001), no. 1-3, 181–196.
MR 1823348
[26] Anders O. F. Hendrickson, Supercharacter theory constructions corresponding to Schur ring
products, Comm. Algebra 40 (2012), no. 12, 4420–4438. MR 2989654
[27] Anders Olaf Flasch Hendrickson, Supercharacter theories of cyclic p-groups, ProQuest LLC,
Ann Arbor, MI, 2008, Thesis (Ph.D.)–The University of Wisconsin - Madison. MR 2711764
[28] Jef Hendrickx and Marc Van Barel, Fast direct solution methods for symmetric banded
Toeplitz systems, based on the sine transform, Linear Algebra Appl. 343/344 (2002), 211–
232, Special issue on structured and infinite systems of linear equations. MR 1878943
[29] Stephen P. Humphries and Kenneth W. Johnson, Fusions of character tables and Schur
rings of abelian groups, Comm. Algebra 36 (2008), no. 4, 1437–1460. MR MR2406596
(2009b:20008)
[30] David W. Kammler, A first course in Fourier analysis, second ed., Cambridge University
Press, Cambridge, 2007. MR 2382058
[31] V. Sanchez, P. Garcia, A. Peinado, J. Segura, and Rubio A., Diagonalizing properties of
the discrete cosine transforms, IEEE Transactions on Signal Processing 43 (1995), no. 11,
2631–2641.
[32] Victoria Sanchez, Antonio M. Peinado, Jose C. Segura, Pedro Garcia, and Antonio J. Rubio,
Generating matrices for the discrete sine transforms, IEEE Transactions on Signal Processing
44 (1996), no. 10, 2644–2646.
[33] Elias M. Stein and Rami Shakarchi, Fourier analysis, Princeton Lectures in Analysis, vol. 1,
Princeton University Press, Princeton, NJ, 2003, An introduction. MR 1970295
[34] Francis Sullivan and Jack Dongarra, Guest editors' introduction: The top 10 algorithms,
Computing in Science & Engineering 2 (2000), 22–23.
[35] Nathaniel Thiem, Branching rules in the ring of superclass functions of unipotent upper-
triangular matrices, J. Algebraic Combin. 31 (2010), no. 2, 267–298. MR MR2592079
[36] Nathaniel Thiem and Vidya Venkateswaran, Restricting supercharacters of the finite group
of unipotent uppertriangular matrices, Electron. J. Combin. 16 (2009), no. 1, Research Paper
23, 32. MR MR2482091 (2010e:20024)
Department of Mathematics, Pomona College, 610 N. College Ave., Claremont, CA 91711
E-mail address: [email protected]
URL: http://pages.pomona.edu/~sg064747
|
1905.02603 | 2 | 1905 | 2019-06-06T23:51:19 | Weighted sampling and weighted interpolation on combinatorial graphs | [
"math.FA"
] | For Paley-Wiener functions on weighted combinatorial finite or infinite graphs we develop a weighted sampling theory in which samples are defined as inner products with weight functions (measuring devices). Three reconstruction methods are suggested. The first two of them are using language of dual Hilbert frames and the so-called frame algorithm respectively. The third one is using the so-called weighted variational interpolating splines which are constructed in the setting of combinatorial graphs. This development requires a new set of Poincar\'e-type inequalities which we prove for functions on combinatorial graphs. | math.FA | math |
WEIGHTED SAMPLING AND WEIGHTED INTERPOLATION
ON COMBINATORIAL GRAPHS
ISAAC Z. PESENSON
Abstract. For Paley-Wiener functions on weighted combinatorial finite or
infinite graphs we develop a weighted sampling theory in which samples are
defined as inner products with weight functions (measuring devices). Three
reconstruction methods are suggested. The first two of them are using language
of dual Hilbert frames and the so-called frame algorithm respectively. The
third one is using the so-called weighted variational interpolating splines which
are constructed in the setting of combinatorial graphs. This development
requires a new set of Poincar´e-type inequalities which we prove for functions
on combinatorial graphs.
1. Introduction and main results
During the last decade signal processing on graphs was developed in a number of
papers, for example, in [3], [6], [12]-[20]. Many of the papers on this list considered
what can be called as a "point-wise sampling". The goal of the present article is
to develop sampling on graphs which is based on weighted averages over relatively
small subgraphs. The idea to use local information (other than point values) for
reconstruction of bandlimited functions on graphs was already explored in [19].
However, the results and methods of [19] and of our paper are very different. We
also want to mention that results of the present paper are similar to results of our
papers [10] and [11] in which sampling by weighted average values was developed
in abstract Hilbert spaces and on Riemannian manifolds.
Let G denote an undirected weighted graph, with a finite or countable number
of vertices V (G) and weight function w : V (G)× V (G) → R+
0 . w is symmetric, i.e.,
w(u, v) = w(v, u), and w(u, u) = 0 for all u, v ∈ V (G). The edges of the graph are
the pairs (u, v) with w(u, v) 6= 0. Our assumption is that for every v ∈ V (G) the
following finiteness condition holds
(1.1)
w(v) = Xu∈V (G)
w(u, v) < ∞.
Let ℓ2(G) denote the space of all complex-valued functions with the inner product
and the norm
hf, gi = Xv∈V (G)
kfkG = kfk =
Xv∈V (G)
1
f (v)g(v)
1/2
.
f (v)2
2
ISAAC Z. PESENSON
Definition 1. The weighted gradient norm of a function f on V (G) is defined by
(1.2)
k∇fk =
Xu,v∈V (G)
1
2f (u) − f (v)2w(u, v)
1/2
.
The set of all f : G → C for which the weighted gradient norm is finite will be
denoted as D(∇).
Remark 1.1. The factor 1
2 makes up for the fact that every edge (i.e., every
unordered pair (u, v)) enters twice in the summation. Note also that loops, i.e.
edges of the type (u, u), in fact do not contribute.
We intend to prove Poincar´e-type estimates involving weighted gradient norm.
In the case of a finite graph and ℓ2(G)-space the weighted Laplace operator
L : ℓ2(G) → ℓ2(G) is introduced via
(Lf )(v) = Xu∈V (G)
(1.3)
(f (v) − f (u))w(v, u) .
This graph Laplacian is a well-studied object; it is known to be a positive-semidefinite
self-adjoint bounded operator.
According to Theorem 8.1 and Corollary 8.2 in [5] if for an infinite graph there
exists a C > 0 such that the degrees are uniformly bounded
(1.4)
w(v) = Xu∈V (G)
w(u, v) ≤ C,
then operator which is defined by (1.3) on functions with compact supports has
a unique positive-semidefinite self-adjoint bounded extension L which is acting ac-
cording to (1.3).
In section 2 we consider a finite connected graph G which contains more than
one vertex and a functional Ψ on ℓ2(G) which is defined by a function ψ ∈ ℓ2(G),
i.e.
Ψ(f ) = hf, ψi = Xv∈V (G)
f (v)ψ(v).
We will use notation χG for the characteristic function: χG(v) = 1 for all v ∈ V (G).
In these notions we prove (Theorem 2.2) that if Ψ(χG) is not zero then for any
f ∈ Ker(Ψ) the following inequality holds
(1.5)
θ
λ1k∇fk2, f ∈ Ker(Ψ),
kfk2 ≤
where λ1 is the first non zero eigenvalue of the Laplacian (1.3) and
(1.6)
θ = Gkψk2
Ψ(χG)2 ,
where G is cardinality of V (G).
In section 3 we extend this result to situations in which a cover by finite and
connected subgraphs of a finite or infinite graph G is given. Namely, we are
working under the following assumptions.
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
3
Assumptions 1. We assume that S = {Sj}j∈J form a cover of V (G)
(1.7)
Sj = V (G).
[j∈J
We don't assume that the sets Sj are disjoint but we assume that there is no any
edge in E(G) which belongs to two different subsets Sj, j ∈ J.
Let Lj be the Laplacian for the induced subgraph Sj. In order to insure that
Lj has at least one non zero eigenvalue, we assume that every Sj ⊂ V (G), j ∈ J,
is a finite and connected subset of vertices with more than one vertex. The first
nonzero eigenvalue of the operator Lj will be denoted as λ1,j. Let k∇jfjk be the
weighted gradient for the induced subgraph Sj. With every Sj, j ∈ J, we associate
a function ψj ∈ ℓ2(G) whose support is in Sj and introduce the functionals Ψj on
ℓ2(G) defined by these functions
Ψj(f ) = hf, ψji = Xv∈V (Sj )
f (v)ψj (v),
f ∈ ℓ2(G).
Notation χj will be used for characteristic function of Sj and use fj for f χj, f ∈
ℓ2(G).
As usual, the induced graph Sj has the same vertices as the set Sj but only such
edges of E(G) which have both ends in Sj.
The two inequalities below (1.9) and (1.10) are essentially the main inequalities
we prove in section 3. We call them generalized Poincar´e-type inequalities since
they contain an estimate of a function through its smoothness. Namely, we show
that if
Ψj(χj) = Xv∈Sj
ψj(v) 6= 0,
and
(1.8)
θj = Sjkψjk2
Ψj(χj)2 ,
then the following inequalities hold for every f ∈ ℓ2(G) and every ǫ > 0
(1.9)
(1.10)
kfk2 ≤ (1 + ǫ)Xj∈J
θj
λ1,j k∇jfjk2 +
kfk2 ≤ (1 + ǫ)
ΘΞ
ΛS kL1/2fk2 +
1 + ǫ
ǫ Xj∈J
ǫ Xj∈J
Sj2
Ψj(χj)2 Ψj(fj)2,
Sj2
Ψj(χj)2 Ψj(fj)2 .
1 + ǫ
where Ξ = (S, {Ψj}j∈J ), S = {Sj}j∈J ,
ΘΞ = sup
(1.11)
j∈J
θj < ∞,
where θj is computed according to (2.3) and
(1.12)
ΛS = inf
j∈J
λ1,j > 0.
4
ISAAC Z. PESENSON
Note, that an important situation occurs in (1.9) and (1.10) when f ∈ ∩j∈J KerΨj.
In this case one has
(1.13)
and
(1.14)
kfk2 ≤Xj∈J
θj
λ1,j k∇jfjk2,
f ∈ \j∈J
Ker Ψj,
kfk2 ≤
ΘΞ
ΛS kL1/2fk2,
f ∈ \j∈J
Ker Ψj.
Another interesting case occurs when for every j ∈ J the functional Ψj is a Dirac
measure δvj at a vertex vj ∈ Sj. In this case the condition f ∈ ∩j∈J Ker δvj means
that f (vj ) = 0,
j ∈ J, and one obtains (see (3.13) and (3.14) below)
kfk2 ≤Xj∈J
λ1,j k∇jfjk2, f ∈ \j∈J
Sj
Ker δvj ,
(1.15)
and
(1.16)
kfk2 ≤
supj∈J Sj
ΛS
kL1/2fk2, f ∈ \j∈J
Ker δvj .
A few more interesting particular situations will be discussed at the end of section
3. We also have similar inequalities for subgraphs (see formulas (3.17) and (3.18)
below). Let's note, that in the continuous case (see [9]-[11]) such inequalities play an
important role in the sampling and interpolation theories on Riemannian manifolds.
Remark 1.2. It is interesting to note that if one will rearrange and mutually
"local" inequalities (1.9), (1.13), (1.15) will stay the same but the "global" ones
connect subgraphs {Sj}j∈J in any other way to obtain a new graph eG then the
(1.10), (1.14), (1.16 ) will change since they will involve a new Laplacian eL which
corresponds to eG.
It is worth to stress that the "local" inequalities (1.9), (1.13), (1.15) are quite
informative and capture highly irregular local structures of graphs. Indeed, in the
case, say, of a Riemannian manifold a difference between two small neighborhoods
Si and Sj is essentially their diameter. However, in the case of a graph two different
even "small" sets can have very different structures. These differences are better
reflected by quantities like λ1,j and θj.
Let's also note that from the practical point of view, the averaging procedure
(which corresponds the case when ψj is the characteristic function χj of a subset
Sj) can be instrumental in reducing noise inherited into point wise measurements.
In section 4 we introduce Paley-Wiener spaces P Wω for finite and infinite graphs.
In section 5 using inequalities (1.9) and (1.10) and their variations we develop a
sampling theory for Paley-Wiener functions on finite and infinite graphs (Theorems
5.1 and 5.2). At this point for reconstruction of functions from weighted average
samples we adopt dual Hilbert frames and the so-called frame algorithm.
In section 6 by using inequality (1.14) we outline a construction of variational
interpolating splines which interpolate functions using their weighted average values
over subsets. It is shown that Paley-wiener functions can be reconstructed using
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
5
weighted average interpolating splines when smoothness of splines goes to infinity.
In section 7 we illustrate some of our results using infinite graph Z.
2. A Poincare-type inequality for finite graphs
The following lemma is important for us (see [7] for finite graphs, and [3] for
infinite).
Lemma 2.1. If a graph G is finite or the condition (1.4) is satisfied then one has
the equality
(2.1)
for all f ∈ ℓ2(G).
Proof. It is easy to verify that under assumption (1.4) the domain D(∇) coincides
with ℓ2(G). Let d(u) = wV (G)(u). Then we obtain
kL1/2fk = k∇fk
hf, Lfi = Xu∈V (G)
= Xu∈V (G)
In the same way
(f (u) − f (v)) w(u, v)
Xv∈V (G)
f (u)
f (u)2d(u) − Xv∈V (G)
f (u)f (v)w(u, v)
.
hf, Lfi = hLf, fi
= Xu∈V (G)
f (u)2d(u) − Xv∈V (G)
f (u)f (v)w(u, v)
.
Averaging these equations yields
hf, Lfi = Xu∈V (G)
f (u)f (v)w(u, v)
f (u)2d(u) − Re Xv∈V (G)
2 Xu,v∈V (G)
= Xu,v∈V (G)
=
1
1
2f (v) − f (u)2w(u, v) = k∇fk2 .
f (u)2w(u, v) + f (v)2w(u, v) − 2Ref (u)f (v)w(u, v)
Lemma is proved.
(cid:3)
For a finite connected graph G which contains more than one vertex let Ψ be a
functional on ℓ2(G) which is defined by a function ψ ∈ ℓ2(G), i.e.
Ψ(f ) = hf, ψi = Xv∈V (G)
f (v)ψ(v).
We will use notation χG for the characteristic function: χG(v) = 1 for all v ∈ V (G).
Using these notions we prove the following.
6
ISAAC Z. PESENSON
Theorem 2.2. Let G be a finite connected graph which contains more than one
vertex and Ψ(χG) is not zero. If f ∈ Ker(Ψ) then
(2.2)
θ
λ1k∇fk2, f ∈ Ker(Ψ),
kfk2 ≤
where λ1 is the first non zero eigenvalue of the Laplacian (1.3) and
(2.3)
θ = Gkψk2
Ψ(χG)2 ,
where G is cardinality of V (G).
Proof. If λ0 < λ1 ≤ ....λN −1, N = G is the set of eigenvalue and ϕ0, ϕ1, ..., ϕN −1
is a set of orthonormal eigenfunctions then {ck(f ) = hf, ϕki} is a set of Fourier
coefficients. One has
and then using Parseval equality and Schwartz inequality we obtain
and if f ∈ Ker(Ψ) then
0 = Ψ(f ) =
From here
kfk2 = c0(f )2 +
(2.4)
N −1Xk=1
G
ck(f )2 =
N −1Xk=1
ck(f )2
Ψ(χG)2
At the same time, since ϕ0 = χG√G
c0(f ) = − pG
Ψ(χG)
f =
N −1Xk
ck(f )ϕk
c0(f )Ψ(χG) +
N −1Xk=
ck(f )Ψ(ϕk).
1
pG
G
ck(f )Ψ(ϕk),
N −1Xk=1
Ψ(χG)2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N −1Xk=1
and hψ, ϕki = Ψ(ϕk) we have
ck(f )Ψ(ϕk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N −1Xk=1
Ψ(ϕk)2 +
N −1Xk=1
2
+
ck(f )2.
N −1Xk=1
ck(f )2 ≤
and from Parseval formula
N −1Xk=1
1
Ψ(ϕk)ϕk,
ψ =
Ψ(χG)ϕ0 +
pG
N −1Xk=1
Ψ(ϕk)2 = kψk2 − Ψ(χG)2
G
.
We plug the right-hand side of this formula into (2) and obtain the next inequality
in which θ is given by (2.3)
θ
λ1kL1/2k2.
To finish the proof one has to apply Lemma 2.1. Theorem is proven.
λ1/2
k ck(f )2 =
ck(f )2 ≤
kfk2 ≤ θ
θ
λ1
N −1Xk=1
N −1Xk=1
(cid:3)
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
7
Corollary 2.1. Let G be a finite connected graph which contains more than one
vertex and Ψ(χG) is not zero. Then one has for every f ∈ ℓ2(G)
(2.5)
2
Ψ(f )
Ψ(χG)
θ
λ1 k∇fk2,
≤
(cid:13)(cid:13)(cid:13)(cid:13)f −
χG(cid:13)(cid:13)(cid:13)(cid:13)
where θ as in (2.3).
hold:
The proof follows from the fact that for g = f − Ψ(f )
g ∈ Ker (Ψ), ∇g = ∇f.
Ψ(χG) χG the following properties
When ψ equals to the eigenfunction ϕ0 then for the corresponding functional Ψ0
the condition f ∈ Ker(Ψ0) is equivalent to hf, ϕ0i = 0. It is easy to see that in
this case θ = 1 and then (2.2) gives the following Corollary.
Corollary 2.2. If hf, ϕ0i = 0 then
(2.6)
kfk2 ≤
1
λ1 k∇fk2.
Note also, that this inequality immediately follows from Lemma 2.1 and from
the fact that the norm of the operator L−1/2 on the subspace of all functions which
are orthogonal to ϕ0 is 1/√λ1.
In another particular case when ψ = χG and
fG =
1
G Xv∈V (G)
f (v),
one has that f − fGχG belongs to the kernel of the corresponding functional Ψ and
it gives the next Corollary.
Corollary 2.3. For every finite graph G and for every f ∈ ℓ2(G) the following
holds
kf − fGχGk2 ≤
1
λ1 k∇fk2.
Theorem 2.3. Let G be a finite graph and Ψ be a functional on ℓ2(G) such that
Ψ(χG) is not zero. Then the following Poincare inequality holds for every f ∈ ℓ2(G)
and every ǫ > 0
(2.7)
kfk2 ≤
θ
λ1
(1 + ǫ)k∇fk2 +
1 + ǫ
ǫ
where θ is defined in (2.3).
Proof. One has
kfk2 ≤(cid:13)(cid:13)(cid:13)(cid:13)
(cid:18)f −
Ψ(f )
Ψ(χG)
Next, we apply the inequality
G2
Ψ(χG)2 Ψ(f )2,
χG(cid:13)(cid:13)(cid:13)(cid:13)
χG(cid:19) +
Ψ(f )
Ψ(χG)
2
f ∈ ℓ2(G), ǫ > 0,
(2.8)
A2 ≤ (1 + ǫ)A − B2 +
1 + ǫ
ǫ
B2 ,
which holds for every positive ǫ > 0. This inequality follows from two obvious
inequalities
A2 ≤ A − B2 + 2BA − B + B2
8
and
ISAAC Z. PESENSON
2BA − B ≤ ǫA − B2 + ǫ−1B2,
Choosing an ǫ > 0 and using inequality (2.8) one obtains
ǫ > 0.
(2.9)
kfk2 ≤ (1 + ǫ)(cid:13)(cid:13)(cid:13)(cid:13)f −
Ψ(f )
Ψ(χG)
2
χG(cid:13)(cid:13)(cid:13)(cid:13)
+
1 + ǫ
ǫ
G2
Ψ(χG)2 Ψ(f )2.
Now an application of Corollary 2.1 gives the result. Theorem is proved.
(cid:3)
In the case when Ψ is defined by ψ = χG one has that
Ψ(f )
Ψ(χG)
χG = fGχG,
fG =
1
G Xv∈V (G)
f (v).
Since in this case θ in (2.3) is 1, G2/Ψ(χG)2 = 1, and Ψ(f ) =Pv∈V (G) f (v) we
obtain
Corollary 2.4. For every connected and finite graph G which contains more than
one vertex the following Poincar´e inequality holds
(2.10)
kfk2 ≤ (1 + ǫ)
1
λ1 k∇fk2 +
1 + ǫ
ǫ
2
,
f ∈ ℓ2(G), ǫ > 0.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xv∈V (G)
f (v)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
3. A generalized Poincare-type inequality for finite and infinite
graphs
Let G be a finite or infinite and countable connected graph and S ⊂ V (G) is a
finite and connected subset of vertices which we will treat as an induced graph and
will denote by the same letter S. We remind that this means that the set of vertices
of such graph, which will be denoted as V (S), is exactly the set of vertices in S
and the set of edges is the set of edges in E(G) whose both ends belong to S. Let
LS and k∇S (fS)k be the Laplace operator and the weighted gradient constructed
according to (1.3) and (1.2) for the induced graph S. Let wS(u, v), u, v ∈ V (S),
and
wS(v) = Xu∈V (S)
wS(u, v), v ∈ V (S),
be the corresponding weight functions. We notice that for every induced subgraph
S one has the inequalities and every u, v ∈ V (S) one has w(u, v) = wS(u, v).
However, in general w(u) ≥ wS(u).
Below we consider a cover of V (G) by finite and connected sets of vertices Sj, j ∈
J. We are using the same assumptions and notations which were introduced in
Assumptions 1 in Introduction.
Theorem 3.1. Let G be a connected finite or infinite and countable graph. Suppose
that (1.7) holds true. Let Lj be the Laplace operator of the induced subgraph Sj
whose first nonzero eigenvalue is λ1,j. The following inequality holds for every
f ∈ ℓ2(G) and every ǫ > 0
kfk2 ≤ (1 + ǫ)Xj∈J
Sj2
Ψj(χj)2 Ψj(fj)2,
θj
λ1,j k∇jfjk2 +
ǫ Xj∈J
1 + ǫ
(3.1)
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
9
where Ψj(f ) = hf, ψji, function ψj ∈ ℓ2(G) has support in Sj,
Ψj(χj) = Xv∈Sj
ψj(v) 6= 0,
and
(3.2)
θj = Sjkψjk2
Ψj(χj)2 .
Xv∈V (Sj)
Proof. One has
(3.3)
kfk2 = Xv∈V (G)
fj(v)2
.
We apply Theorem 2.3 to have for every j ∈ J and every ǫ > 0,
Sj2
Ψj(χj)2Ψj(fj)2,
(3.4)
f (v)2 =Xj∈J
θj
λ1,j k∇jfjk2 +
fj(v)2 ≤ (1 + ǫ)
Xv∈V (Sj )
1 + ǫ
ǫ
and then we have for f ∈ ℓ2(G), ǫ > 0,
(3.5)
kfk2 ≤ (1 + ǫ)Xj∈J
θj
λ1,j k∇jfjk2 +
1 + ǫ
ǫ Xj∈J
Sj2
Ψj(χj)2 Ψj(fj)2.
Theorem is proved.
(cid:3)
As a consequence we obtain the following.
Theorem 3.2. If in addition to assumptions of Theorem 3.1 we have that
(3.6)
ΘΞ = sup
j∈J
θj < ∞,
Ξ = ({Sj}j∈J , {Ψj}j∈J ) ,
where θj is computed according to (2.3) and
(3.7)
ΛS = inf
j∈J
λ1,j > 0,
S = {Sj}j∈J ,
then the following inequality holds for every f ∈ ℓ2(G) and every ǫ > 0
Sj2
Ψj(χj)2 Ψj(fj)2 .
(3.8)
ΘΞ
ΛS kL1/2fk2 +
kfk2 ≤ (1 + ǫ)
ǫ Xj∈J
1 + ǫ
Proof of this statement follows from Theorem 3.1 and Lemma 2.1 according to
which
Xj∈J
k∇jfjk2 ≤Xj∈J
k∇jfjk2 ≤ k∇fk2 = kL1/2fk.
Let's consider a few interesting cases.
Corollary 3.1. If all the notations and conditions of Theorems 3.1 and 3.2 are
satisfied and if for every j the corresponding function ψj = χj is the characteristic
function of a subset of vertices Uj ⊆ Sj then the following inequalities hold
(3.9)
kfk2 ≤ (1 + ǫ)Xj∈J
Sj
λ1,jUjk∇jfjk2 +
1 + ǫ
ǫ
Xj∈J
Sj2
Uj2
Xv∈Uj
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
,
f (v)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
10
and
ISAAC Z. PESENSON
(3.10)
kfk2 ≤ (1 + ǫ)
Sj
UjkL1/2fk2 +
In particular, if Uj = Sj for every j ∈ J then
1
ΛS
sup
j∈J
1 + ǫ
ǫ
kfk2 ≤ (1 + ǫ)Xj∈J
1
λ1,j k∇jfjk2 +
1 + ǫ
(3.11)
and
(3.12)
kfk2 ≤ (1 + ǫ)
1
ΛS kL1/2fk2 +
1 + ǫ
ǫ Xj∈J
Xv∈Uj
2
.
f (v)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Sj2
Xj∈J
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xv∈Sj
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Uj2
f (v)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xv∈Sj
f (v)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
ǫ Xj∈J
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
.
2
,
Indeed, it follows from the fact that in this situation kψjk2 = Uj,
Ψj(χj)2 =
Uj2 and
Ψj(χj)2 = Sj
Uj
The condition (3.6) boils down to supj∈J Sj < ∞.
θj = Sjkψjk2
.
Corollary 3.2. Suppose that all the notations and conditions of Theorems 3.1 and
3.2 are satisfied. If for every j the corresponding function ψj is a Dirac measure
δvj at a vertex vj ∈ Sj then
kfk2 ≤ (1 + ǫ)Xj∈J
(3.13)
Sj
λ1,j k∇jfjk2 +
Sj2f (vj)2,
ǫ Xj∈J
1 + ǫ
and
(3.14)
kfk2 ≤ (1 + ǫ)
supj∈J Sj
ΛS
kL1/2fk2 +
1 + ǫ
ǫ Xj∈J
Sj2f (vj)2.
Proof. In this case one has kψjk = 1, Ψj(f ) = f (vj), Ψj(χj) = 1, θj = Sj for
every j ∈ J.
(cid:3)
The next corollary is about functions which annihilate all the functionals Ψj, j ∈
J.
Corollary 3.3. If all the notations and conditions of Theorems 3.1 and 3.2 are
satisfied and for a function f ∈ ℓ2(G) one has that
Ker Ψj
f ∈ \j∈J
then
(3.15)
and
(3.16)
kfk2 ≤Xj∈J
θj
λ1,j k∇jfjk2,
kfk2 ≤
ΘΞ
ΛS kL1/2fk2,
f ∈ \j∈J
f ∈ \j∈J
KerΨj,
KerΨj.
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
11
Remark 3.3. If J0 ⊂ J and G0 = ∪j∈J0 Gj then every inequality in this section
can be replaced by a similar one in which the term kfk2 on the left is replaced by
kfk2
G0 = Xv∈G0
kfk2,
and summation over J on the right is replaced by summation over J0. For example,
the last two inequalities (3.15) and (3.16) would take the form
(3.17)
and
(3.18)
kfk2
G0 ≤ Xj∈J0
θj
λ1,j k∇jfjk2,
f ∈ \j∈J0
KerΨj,
kfk2
G0 ≤
ΘΞ
ΛS kL1/2
G0 f0k2,
f0 = fG0, f0 ∈ \j∈J0
KerΨj,
where LG0 is the Laplacian of the induced graph G0.
Note, that in the case when {Ψj} is a set of "uniformly" distributed Dirac func-
tions the last inequality (3.18) is called sometimes "the inequality for functions with
many zeros".
4. Paley-Wiener vectors in ℓ2(G)
Our next goal is to introduce the so-called Paley-Wiener functions (bandlimited
functions) for which a sampling theory will be developed in the setting of combi-
natorial graphs. We use for this the self-adjoint positive definite operator L in a
Hilbert space ℓ2(G). In the case when L has discrete spectrum (which is always the
case with finite graphs) then the Paley-Wiener space P Wω(L) is simply the span of
eigenfunctions of L whose corresponding eigenvalues are not greater ω. However,
when graph is infinite and spectrum of L is continuous it takes a bigger effort to
define spaces P Wω(L).
Consider a self-adjoint positive definite operator L in a Hilbert space ℓ2(G).
According to the spectral theory [1] for self-adjoint non-negative operators there
exists a direct integral of Hilbert spaces H =R H(λ)dm(λ) and a unitary operator
F from ℓ2(G) onto H, which transforms the domains of Lk, k ∈ N, onto the sets
Hk = {x ∈ Hλkx ∈ H} with the norm
(4.1)
kx(λ)kHk = hx(λ), x(λ)i1/2
(cid:18)Z ∞
H(λ)dm(λ)(cid:19)1/2
λ2kkx(λ)k2
H(λ) =
0
.
and satisfies the identity F (Lkf )(λ) = λk(F f )(λ), if f belongs to the domain of
Lk. We call the operator F the Spectral Fourier Transform. As known, H is the
set of all m-measurable functions λ 7→ x(λ) ∈ H(λ), for which the following norm
is finite:
kxkH =(cid:18)Z ∞
0
H(λ)dm(λ)(cid:19)1/2
kx(λ)k2
For the characteristic function 1[0, ω] one can introduce the projector 1[0, ω](L) by
using the formula
(4.2)
1[0, ω](L)f = F −11[0, ω](λ)F f, f ∈ H.
12
ISAAC Z. PESENSON
Definition 2. The Paley-Wiener space P Wω(L) ⊂ ℓ2(G) is defined as the image
space of the projection operator 1[0, ω](L).
Many properties of Paley-Wiener spaces for general self-adjoint operators in
Hilbert spaces can be found in our papers [9]. The most important for us is the
following.
Theorem 4.1. A function f ∈ ℓ2(G) belongs to the spaces P Wω(L) if and only if
the following Bernstein inequalities holds true
(4.3)
kLsfk ≤ ωskfk
for all s ∈ R+;
5. A sampling theorem and a reconstruction methods using frames
5.1. A sampling theorem. Let's remind that a set of vectors {ξν} in a Hilbert
space H is called a Hilbert frame if there exist constants A, B > 0 (frame bounds)
such that for all f ∈ H
(5.1)
hf, ξνi2 ≤ Bkfk2.
Akfk2 ≤Xν
What is remarkable about frames is the fact that one can perfectly reconstruct a
vector f from its projections hf, ξνi. Namely, according to the general theory of
Hilbert frames [2], [4] the frame inequality (5.1) implies that there exists a dual
frame {Ων} (which is not unique in general) for which the following reconstruction
formula holds
(5.2)
f =Xv
hf, ξνi Ων.
In general it is not easy to find a dual frame. For this reason one can resort to
the following frame algorithm (see [4], Ch. 5) which performs reconstruction by
iterations. Given a relaxation parameter 0 < ρ < 2
B , set η = max{1 − ρA, 1 −
ρB} < 1. Let f0 = 0 and define recursively
(5.3)
where Φ is the frame operator which is defined on H by the formula Φf =Pν hf, ξνi ξν .
In particular, f1 = ρΦf = ρPj hf, ξνi ξν. Then limn→∞ fn = f with a geometric
fn = fn−1 + ρΦ(f − fn−1),
rate of convergence, that is,
(5.4)
In particular, for the choice ρ = 2
kf − fnk ≤ ηnkfk.
A+B the convergence factor is
η =
B − A
A + B
.
Let δsi , i ∈ I be the Dirac delta concentrated at the vertex si.
Theorem 5.1. If all the notations and conditions of Theorems 3.1 and 3.2 hold
then the set of functionals {Ψj}j∈J is a frame in any space P Wω(L) as long as
(1)
(5.5)
0 < ω <
ΛS
(1 + ǫ)ΘΞ
,
ǫ > 0,
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
13
(2) there exists a constant c = c({Sj},{Ψj}) such that for every j ∈ J the
following inequality holds
(5.6)
Sj2
Ψj(χj)2 ≤ c,
(3) there exists a constant C = C({Sj},{Ψj}) such that for every j ∈ J one
has
(5.7)
kψjk2 ≤ C.
In other words, if for an ǫ > 0 the following inequality holds
(5.8)
γ = (1 + ǫ)
ΘΞ
ΛS
ω < 1, ǫ > 0,
along with (5.6) and (5.7) then
(1 + ǫ)c kfk2 ≤Xj∈J
(1 − γ)ǫ
(5.9)
Ψj(f )2 ≤ Ckfk2.
Proof. We notice that since support of ψj is in Sj we have
Ψj(fj) = hf, ψji = Ψj(f ).
Now, if f ∈ P Wω(L) then by the Bernstein inequality (4.3) the (3.8) can be rewrit-
ten as
kfk2 ≤ (1 + ǫ)
ΘΞ
ΛS
ωkfk2 +
Sj2
Ψj(χj)2 Ψj(fj)2 .
1 + ǫ
ǫ Xj∈J
fj(v)ψj (v)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
If (5.6) and (5.8) hold then one obtains the left-hand side of (5.9). On the other
hand, we have
Xj∈J
Ψj(f )2 =Xj∈J
Theorem is proven.
Xv∈Sj
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤Xj∈J
kψjk2kfjk2 ≤ Ckfk2.
(cid:3)
Note, that for the classical Paley-Wiener spaces on the real line the inequalities
similar to (5.9) in the case when {ψj} are delta functions were proved by Plancherel
and Polya. Today they are better known as the frame inequalities. Now we can
formulate sampling theorem based on average values.
Theorem 5.2. Under the same conditions and notations as in Theorem 5.1 every
function f ∈ P Wω(L) is uniquely determined by the set of numbers {hf, ψji}j∈J
and can be reconstructed from this set of values in a stable way using dual frames
(5.2) or the iterative frame algorithm (5.3).
5.2. Important particular cases.
(1) (Sampling by averages-I). If for every j the corresponding function ψj = χj
is the characteristic function of a subset of vertices Uj ⊂ Sj then inequalities
(5.5)-(5.7) take the form respectively
0 < ω <
ΛS
(1 + ǫ) supj∈J Sj
,
Sj2
Uj2 ≤ c,
Uj ≤ C,
14
ISAAC Z. PESENSON
and the Plancherel-Polya inequalities (5.9) hold with the same constants c
and C. In particular, if Uj = Sj for every j ∈ J then (5.5) takes the form
(5.10)
0 < ω <
ΛS
(1 + ǫ) supj∈J Sj
,
the condition (5.6) is trivially satisfied with c = 1, and (5.7) becomes
Sj ≤ C. The (5.9) holds true with the corresponding constants C and
c = 1.
(2) (Sampling by averages-II). In the case Uj = Sj and
ψj =
χj,
1
pSj
every θj in (3.2) is one and it gives that ΘΞ in (3.6) is also one. Thus (5.5)
takes the form
(5.11)
(5.12)
(5.13)
0 < ω <
ΛS
1 + ǫ
,
ǫ > 0.
Moreover, in this case C = c = 1. After all the Plancherel-Polya inequality
(5.9) becomes
(1 + ǫ) kfk2 ≤Xj∈J
(1 − γ)ǫ
Ψj(f )2 ≤ kfk2, f ∈ P Wω(G),
where
γ =
1 + ǫ
ΛS
ω < 1,
ǫ > 0.
(3) (Point wise sampling).
If for every j the corresponding function ψj is a
Dirac measure δvj at a vertex vj ∈ Sj then the condition (5.5) takes the
form (5.10), the condition (5.6) will have form Sj2 ≤ c, the condition (5.7)
is trivially satisfied with C = 1. The (5.9) holds true with these constants.
5.3. Reconstruction algorithms in terms of frames. What we just proved
in the previous section is that under the same assumptions as above the set of
functionals f → hf, ψji is a frame in the subspace P Wω(L). This fact allows
to apply the well known result of Duffin and Schaeffer [2] which describes a stable
method of reconstruction of a function f ∈ P Wω(L) from a set of samples {hf, ψji}.
Theorem 5.3. If all the conditions of Theorem 3.1 are satisfied then there exists
a dual frame {Ωj} in P Wω(L) such that
f =Xj
hf, ψji Ωj =Xj
hf, Ωji Pωψj
where Pω is the orthogonal projection of ℓ2(G) onto P Wω(L).
Another possibility for reconstruction is to use frame algorithm (see section 5).
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
15
6. Weighted Average Variational Splines and a reconstruction
algorithm
6.1. Variational interpolating splines. As in the previous sections we assume
that G is a connected finite or infinite graph, S = {Sj}j∈J , is a disjoint cover of
V (G) by connected and finite subgraphs Sj and every ψj ∈ ℓ2(Sj),
j ∈ J, has
support in Sj.
For a given sequence a = {aj} ∈ l2 the set of all functions in ℓ2(G) such that
Ψj(f ) = hf, ψji = aj will be denoted by Za. In particular,
Z0 = \j∈J
Ker(Ψj)
corresponds to the sequence of zeros. We consider the following optimization prob-
lem:
For a given sequence a = {aj} ∈ l2 find a function f in the set Za ⊂ ℓ2(G) which
minimizes the functional
u → kLk/2uk, u ∈ Za.
Theorem 6.1. Under the above assumptions the optimization problem has a unique
solution for every k.
Proof. Using Theorem 3.1 one can justify the following algorithm (see [8], [10]):
(1) Pick any function f ∈ Za.
(2) Construct P0f where P0 is the orthogonal projection of f onto Z0 with
respect to the inner product
hf, gik =Xj
hf, ψjihg, ψji + hLk/2f, Lk/2gi.
(3) The function f − P0f is the unique solution to the given optimization
problem.
(cid:3)
Definition 3. For f ∈ ℓ2(G) the interpolating variational spline is denoted by
sk(f ) and it is the solution of the minimization problem such that sk(f ) − f ∈ Z0.
Clearly, "interpolation" is understood in the sense that
(6.1)
Ψj(sk(f )) = Ψj(f ).
One can easily prove the following characterization of variational splines.
Theorem 6.2. A function u ∈ ℓ2(G) is a variational spline if and only if Lku is
orthogonal to LkZ0.
6.2. Reconstruction using splines. The following Lemma was proved in [8],
[10].
Lemma 6.3. If A is a self-adjoint non-negative operator in a Hilbert space X and
for an ϕ ∈ X and a positive a > the following inequality holds true
then for the same ϕ ∈ H, and all k = 2l, l = 0, 1, 2, ... the following inequality holds
kϕk ≤ akAϕk,
kϕk ≤ akkAkϕk.
16
ISAAC Z. PESENSON
By using the same reasoning as in [8], [10] one can prove the following recon-
struction theorem. Below we are keeping notations of Theorem 5.1.
Theorem 6.4. Let's assume that G is a connected finite or infinite graph, {Sj}j∈J
is a disjoint cover of V (G) by connected and finite subgraphs Sj and every ψj ∈
ℓ2(Sj ), j ∈ J, has support in Sj. If
(6.2)
0 < ω <
ΛS
ΘΞ
,
(6.3)
(6.4)
ΘΞ = sup
j∈J
θj = θj = Sjkψjk2
Ψj(χj)2 ,
ΛS = inf
j∈J
λ1,j,
then any function f in P Wω(L), ω > 0, can be reconstructed from a set of values
{hf, ψji} using the formula
f = lim
k→∞
sk(f ),
k = 2l,
l = 0, 1, ...,
and the error estimate is
(6.5)
where
kf − sk(f )k ≤ 2γkkfk, k = 2l,
l = 0, 1, ...,
γ =
ΘΞ
ΛS
ω < 1.
Proof. For a k = 2l, l = 0, 1, 2, .... apply to the function f − sk(f ) inequality (3.8)
for any ǫ > 0:
kf − sk(f )k2 ≤ (1 + ǫ)
ΘΞ
ΛS kL1/2(f − sk(f ))k2+
(6.6)
1 + ǫ
ǫ Xj∈J
Sj2
Ψj(χj)2 Ψj(fj)2 .
Since sk(f ) interpolates f the last term here is zero. Because ǫ here is any positive
number it brings us to the next inequality
kf − sk(f )k2 ≤
and an application of Lemma 6.3 gives
ΘΞ
ΛS kL1/2(f − sk(f ))k2,
ΛS(cid:19)k
kf − sk(f )k2 ≤(cid:18) ΘΞ
kLk/2(f − sk(f ))k2.
Using minimization property of sk(f ) and the Bernstein inequality (4.3) for f ∈
P Wω(L) one obtains (6.5). Theorem is proved.
(cid:3)
One can formulate similar statements adapted to particular cases listed in sub-
section 5.2.
WEIGHTED SAMPLING AND INTERPOLATION ON GRAPHS
17
7. Example. Average sampling on Z
Let us consider a one-dimensional infinite lattice Z = {...,−1, 0, 1, ...} as an
unweighted graph. The dual group of the commutative additive group Z is the
one-dimensional torus. The corresponding Fourier transform F on the space ℓ2(Z)
is defined by the formula
F (f )(ξ) =Xk∈Z
f (k)eikξ, f ∈ ℓ2(Z), ξ ∈ [−π, π).
It gives a unitary operator from ℓ2(Z) on the space L2(T) = L2(T, dξ/2π), where T
is the one-dimensional torus and dξ/2π is the normalized measure. One can verify
the following formula
The next result is obvious.
F (Lf )(ξ) = 4 sin2 ξ
2F (f )(ξ).
Theorem 7.1. The spectrum of the Laplace operator L on the one-dimensional
lattice Z is the interval [0, 4]. A function f belongs to the space P Wω(Z), 0 ≤ ω ≤ 4,
if and only if the support of F f is a subset of [−π, π) on which 4 sin2 ξ
2 ≤ ω.
We consider the cover Ξ = {Sj} of Z by disjoint sets Sj = {j − 1, j, j + 1} where
j runs over all integers divisible by 3: {...,−3, 0, 3, ...} = 3Z. We treat every Sj as
an induced graph whose set of vertices is V (Sj) = {j − 1, j, j + 1}, j ∈ 3Z, and
which has two edges (j − 1, j) and (j, j + 1). Let' introduce functionals Ψj as
j ∈ 3Z, f ∈ ℓ2(Z).
(7.1) Ψj(f ) = hf, ψji =
One can check that spectrum of the Laplace operator Lj on Sj defined by (1.3)
contains just three values {0, 2, 4}. Thus ΛS = 2. For an 0 < ω < 4 and ǫ > 0
condition (6.2) takes form
(f (j − 1) + f (j) + f (j + 1)) ,
1
√3
(7.2)
γ = (1 + ǫ)
ω
2
< 1.
Note, that since 1 + ǫ can be arbitrary close to 1 the condition (7.2) implies that
0 < ω < 2. As an application of Theorem 5.2 we obtain the following result.
Theorem 7.2. If 0 < ω < 2 then every f ∈ P Wω(Z) is uniquely determined by its
average values {hf, ψji} defined in (7.1) and can be reconstructed from them in a
stable way.
In particular, if instead of infinite graph Z one would consider a path graph ZN
whose eigenvalues are given by formulas 2 − 2 cos kπ
N −1 , k = 0, 1, ..., N − 1, the last
Theorem would mean that any eigenfunction with eigenvalue from a lower half of
the spectrum is uniquely determined and can be reconstructed from averages (7.1).
References
[1] M. Birman and M. Solomyak, Spectral Theory of Selfadjoint Operators in Hilbert Space,
D. Reidel Publishing Co., Dordrecht, 1987.
[2] R. Duffin, A. Schaeffer, A class of nonharmonic Fourier series, Trans. AMS, 72, (1952),
341-366.
[3] H. Fuhr, Hartmut, I. Pesenson, Poincar´e and Plancherel-Polya inequalities in harmonic
analysis on weighted combinatorial graphs, SIAM J. Discrete Math. 27 (2013), no. 4, 2007-
2028.
[4] K. Grochenig, Foundations of Time-Frequency Analysis, Birkhauser, 2001.
18
ISAAC Z. PESENSON
[5] S. Haeseler, M. Keller, D. Lenz, R. Wojciechowski, Laplacians on infinite graphs: Dirichlet
and Neumann boundary conditions, J. Spectr. Theory 2 (2012), no. 4, 397-432.
[6] Madeleine
S. Kotzagiannidis,
Pier Luigi Dragotti,
Sampling
tion of Sparse Signals on Circulant Graphs
https://doi.org/10.1016/j.acha.2017.10.003.
- An Introduction to Graph-FRI
and Reconstruc-
,
[7] B. Mohar, Some applications of Laplace eigenvalues of graphs, in G. Hahn and G. Sabidussi,
editors, Graph Symmetry: Algebraic Methods and Applications (Proc. Montr´real 1996), vol-
ume 497 of Adv. Sci. Inst. Ser. C. Math. Phys. Sci., pp. 225-275, Dordrecht (1997), Kluwer.
[8] I. Pesenson, Sampling of Paley-Wiener functions on stratified groups, J. Four. Anal. Appl. 4
(1998), 269 -- 280.
[9] I. Pesenson, A sampling theorem on homogeneous manifolds, Trans. Amer. Math. Soc. 352/9
(2000), 4257 -- 4269.
[10] I. Pesenson, Sampling of band limited vectors, J. Fourier Anal. Appl. 7/1 (2001), 93-100.
[11] I. Pesenson, Poincar´e-type inequalities and reconstruction of Paley-Wiener functions on
manifolds, J. Geometric Anal. 4/1 (2004), 101-121.
[12] I. Pesenson, Sampling in Paley-Wiener spaces on combinatorial graphs, Trans. Amer. Math.
Soc. 360 (2008), no. 10, 5603-5627.
[13] I. Z. Pesenson, Variational splines and Paley-Wiener spaces on combinatorial graphs, Constr.
Approx. 29 (2009), no. 1, 1-21.
[14] I. Z. Pesenson, M. Z. Pesenson, Sampling, filtering and sparse approximations on combina-
torial graphs, J. Fourier Anal. Appl. 16 (2010), no. 6, 921-942.
[15] Chen, Siheng; Varma, Rohan; Sandryhaila, Aliaksei; Kovacevich, Jelena, Discrete signal
processing on graphs: sampling theory, IEEE Trans. Signal Process. 63 (2015), no. 24, 6510-
6523.
[16] David I Shuman ; Sunil K. Narang ; Pascal Frossard ; Antonio Ortega ; Pierre Vandergheynst
,The emerging field of signal processing on graphs: Extending high-dimensional data analysis
to networks and other irregular domains, IEEE Signal Processing Magazine, 30(3),2013, 83-
98.
[17] Strichartz, Robert S. Half sampling on bipartite graphs, J. Fourier Anal. Appl. 22 (2016), no.
5, 1157-1173.
[18] Tsitsvero, Mikhail; Barbarossa, Sergio; Di Lorenzo, Paolo, Signals on graphs: uncertainty
principle and sampling, IEEE Trans. Signal Process. 64 (2016), no. 18, 4845-4860.
[19] Xiaohan Wang, Pengfei Liu, Yuantao Gu Local-Set-Based Graph Signal Reconstruction, IEEE
Transactions on Signal Processing (2015)
[20] Huang, Weiyu; Marques, Antonio G.; Ribeiro, Alejandro R. Rating prediction via graph signal
processing, IEEE Trans. Signal Process. 66 (2018), no. 19, 5066-5081.
Department of Mathematics, Temple University, Philadelphia, PA 19122
E-mail address: [email protected]
|
1802.01973 | 1 | 1802 | 2018-02-06T14:42:10 | Shorted operators and minus order | [
"math.FA"
] | Let $\mathcal{H}$ be a Hilbert space, $L(\mathcal{H})$ the algebra of bounded linear operators on $\mathcal{H}$ and $W \in L(\mathcal{H})$ a positive operator. Given a closed subspace $\mathcal{S}$ of $\mathcal{H}$, we characterize the shorted operator $W_{/ \mathcal{S}}$ of $W$ to $\mathcal{S}$ as the maximum and as the infimum of certain sets, for the minus order $\stackrel{-}{\leq}.$ Also, given $A \in L(\mathcal{H})$ with closed range, we study the following operator approximation problem considering the minus order: $$ min_{\stackrel{-}{\leq}} \ \{(AX-I)^*W(AX-I) : X \in L(\mathcal{H}), \mbox{ subject to } N(A^*W)\subseteq N(X) \}. $$
We show that, under certain conditions, the shorted operator $W_{/R(A)}$ (of $W$ to the range of $A$) is the minimum of this problem and we characterize the set of solutions. | math.FA | math |
Shorted operators and minus order
M. Continoa,b J. I. Giribeta,b A. Maestripieria,b
a Instituto Argentino de Matem´atica "Alberto Calder´on" - CONICET. Saavedra 15, piso 3 (1083)
Ciudad Aut´onoma de Buenos Aires, Argentina.
b Facultad de Ingenier´ıa - Universidad de Buenos Aires. Paseo Col´on 850 (1063), Ciudad Aut´onoma de
Buenos Aires, Argentina.
ARTICLE HISTORY
Compiled November 9, 2018
ABSTRACT
Let H be a Hilbert space, L(H) the algebra of bounded linear operators on H and W ∈ L(H)
a positive operator. Given a closed subspace S of H, we characterize the shorted operator W/S
−
of W to S as the maximum and as the infimum of certain sets, for the minus order
≤. Also,
given A ∈ L(H) with closed range, we study the following operator approximation problem
considering the minus order:
min−
≤
{(AX − I)∗W (AX − I) : X ∈ L(H), subject to N (A∗W ) ⊆ N (X)}.
We show that, under certain conditions, the shorted operator of W/R(A) is the minimum of
this problem and we characterize the set of solutions.
KEYWORDS
Shorted operators; minus order; oblique projections
1. Introduction
The minus order was introduced by Hartwig [21] and independently by Nambooripad [29], in
both cases on semigroups, with the idea of generalizing some classical partial orders. In [5],
Antezana et al., extended the notion of the minus order to bounded linear operators acting
on an infinite-dimensional Hilbert space. In this case, if H is a Hilbert space and L(H) is the
algebra of bounded linear operators acting on H, for A, B ∈ L(H), A
−
≤ stands for the minus order of operators) if the Dixmier angle between R(A) and R(B − A)
and the Dixmier angle between R(A∗) and R(B∗ − A∗) are less than 1, where R(T ) stands for
the range of the operator T. Independently, in [31] Semrl gave another characterization of the
−
≤ B (where the symbol
−
≤ B if and only if there are bounded (oblique) projections, i.e.
minus order, he showed that A
idempotents, P and Q such that A = P B and A∗ = QB∗. In [18], a new characterization of
the minus order for operators acting on Hilbert spaces was given in terms of the so called range
−
≤ B is equivalent to the range of B being the
additivity property. Namely, it was proved that A
direct sum of ranges of A and B − A and the range of B∗ being the direct sum of ranges of A∗
and B∗ − A∗, which generalizes previous results presented in [31]. This plays an equivalent role
to the rank additivity characterization when A and B are matrices [21, 26].
In [5] the notion of shorted operator appears in relation with the minus order. Given a closed
subspace S of H and W ∈ L(H) a positive operator, in 1947, Krein [23], proved the existence
of a maximum (with respect to the order induced by the cone of positive operators) of the set
M(W, S) = {X ∈ L(H) : 0 ≤ X ≤ W and R(X) ⊆ S ⊥}.
Krein used this extremal operator in his theory of extension of symmetric operators. Years later,
Anderson and Trapp [2] studying the same problem, called this maximum the shorted operator
of W to S (in the following denoted W/S) and showed interesting properties of this operator
and its connections with electrical circuit theory. The shorted operator have shown to be useful
in many applications [5], [32].
The pair (W, S) is said to be compatible if there exists a bounded linear (not necessarily
selfadjoint) projection Q onto S such that W Q is selfadjoint. Thus, if
P(W, S) = {Q ∈ L(H) : Q2 = Q, R(Q) = S, W Q = Q∗W },
then (W, S) is compatible if and only if P(W, S) is not empty. In [15], it was shown that there
exists a strong relationship between compatibility, the projections of P(W, S) and the shorted
operator W/S . Later, in [5], the notion of shorted operator was generalized to that of bilateral
shorted operator, for an operator W ∈ L(H) (not necessarly positive) and a pair of closed
subspaces. The proposed definition comes from the notion of weak complementability, which is
a refinement of a finite dimensional notion due to T. Ando [3]. In that paper, it was proven that
given an operator W ∈ L(H) positive and a closed subspace S, the shorted operator W/S was
the maximum of the set
M−(W, S) = {X ∈ L(H) : X
−
≤ W, R(X) ⊆ S ⊥, R(X ∗) ⊆ S ⊥}
with the minus order when the pair (W, S) is compatible. One of the goals of this paper is to
prove that the shorted operator is the maximum of the set M−(W, S) with the minus order if
and only if the operator W is compatible with the nullspace of W/S , N (W/S ), which generalizes
the results given in [5].
Given W ∈ L(H)+ and A, B ∈ L(H) with closed ranges such that N (B) = N (A∗W ), in [10]
it was shown that, considering the order induced by the cone of positive operators,
W/R(A) = inf≤
X∈L(H)
(AXB − I)∗W (AXB − I).
Moreover, also in [10] it was proved the the minimum of the previous set exists if and only if
the pair (W, R(A)) is compatible.
In this work we study a similar problem considering the minus order: given W ∈ L(H) a
positive operator and A, B ∈ L(H) with closed ranges such that N (B) = N (A∗W ), we study
the existence of
(AXB − I)∗W (AXB − I),
(1.1)
min −
≤
X∈L(H)
and its connection with the shorted operator.
The paper is organized as follows. In section 2, some characterizations of the compatibility of
the pair (W, S) are given. Also the concept of W -inverse of an operator A, and some properties
2
are presented.
In section 3, we collect some useful known results about the minus order, and the connection
between compatibility and shorted operators is stated. Also, in this section we prove that the
shorted operator of W to S, denoted by W/S , is the maximum (in the minus order) of the set
M−(W, S) if and only if the pair (W, N (W/S )) is compatible.
In section 4, we study problem (1.1). We prove that W/R(A) is a lower bound for the set
{(AXB − I)∗W (AXB − I) : X ∈ L(H)} (in the minus order) if and only if the pair (W, N (W/S ))
is compatible and in this case, under additional hypothesis, in particular, if W is injective, then
W/R(A) is the infimum of that set.
We also prove that W/R(A) is the minimum of the previous set if and only if the pair (W, R(A))
is compatible. Moreover, a characterization of the set of solutions is given in terms of the W -
inverses of A.
Motivated by the concepts of the left and the right star orders (see [8]), in section 5 we define
the left and right weigthed star orders on L(H). These are equivalent to the minus order, with
an additional condition on the angle between the ranges of the operators involved imposed by
some positive weight. The last part of the section is devoted to appplications. We apply the new
characterization of the weighted star order to systems of equations and least squares problems.
We also give some formulas for the compression WS = W − W/S of W to the range of the sum
of two closed range operators.
2. Preliminaries
In the following H denotes a separable complex Hilbert space, L(H) is the algebra of bounded
linear operators from H to H, and L(H)+ the cone of semidefinite positive operators. GL(H)
is the group of invertible operators in L(H), CR(H) is the subset of L(H) of all operators
with closed range. For any A ∈ L(H), its range and nullspace are denoted by R(A) and N (A),
respectively. Finally, A† denotes the Moore-Penrose inverse of the operator A ∈ L(H).
Given two closed subspaces M and N of H, M +N denotes the direct sum of M and N and
M ⊖ N = M ∩ (M ∩ N )⊥. If H is decomposed as a direct sum of closed subspaces H = M +N ,
the projection onto M with nullspace N is denoted by PM//N , and the orthogonal projection
onto M is denoted PM = PM//M⊥. Also, Q denotes the subset of L(H) of oblique projections,
i.e. Q = {Q ∈ L(H) : Q2 = Q}.
Given W ∈ L(H)+ and a closed subspace S of H, the pair (W, S) is compatible if there exists
Q ∈ Q with R(Q) = S such that W Q = Q∗W. The last condition means that Q is W -Hermitian,
in the sense that h Qx, y iW = h x, Qy iW , for every x, y ∈ H, where h x, y iW = h W x, y i defines
a semi-inner product on H. Thus, if
P(W, S) = {Q ∈ L(H) : Q2 = Q, R(Q) = S, W Q = Q∗W },
then (W, S) is compatible if and only if P(W, S) is not empty.
The W -orthogonal companion of S is
S ⊥W = {x ∈ H : h W x, y i = 0, y ∈ S}.
We will use that S ⊥W = (W S)⊥ = W −1(S ⊥).
The concept of compatibility between a positive operator W ∈ L(H) and a closed subspace S,
has proved to be useful in several applications such as approximation theory, signal processing,
among others, see for instance [12]-[17].
As it was proved in [15, Prop. 3.3], the compatibility of the pair (W, S) is equivalent to a
3
decomposition of the space in terms of the subspace S and its W -orthogonal companion. This
is stated in the following theorem.
Theorem 2.1. Given W ∈ L(H)+ and a closed subspace S ⊆ H, the following conditions are
equivalent:
i) The pair (W, S) is compatible,
ii) H = S + S ⊥W ,
iii) (W, S + N (W )) is compatible,
iv) H = S + (S ⊥W ⊖ S).
Suppose that the pair (W, S) is compatible and let N = S ∩S ⊥W , observe that N = S ∩N (W );
we define the projection,
PW,S = PS//S ⊥W ⊖N .
The following theorem, proved in [7, Theorem 3.1], establishes conditions for the existence of
solutions of the equation AXB = C.
Theorem 2.2. Let A, B, C ∈ L(H). If R(A), R(B) or R(C) are closed, then the equation
AXB = C admits a bounded solution if and only if R(C) ⊆ R(A) and R(C ∗) ⊆ R(B∗).
In [25] S. K. Mitra and C. R. Rao introduced the notion of W -inverse of a matrix, following
these ideas, in [9] this notion has been extended to linear operators as follows.
Definition. Given A ∈ CR(H), B ∈ L(H) and W ∈ L(H)+, X0 ∈ L(H) is a W -inverse of A in
R(B), if for each x ∈ H, X0x is a weighted least squares solution of Az = Bx, i.e.
kAX0x − BxkW ≤ kAz − BxkW , for every x, z ∈ H.
When B = I, X0 is called a W -inverse of A, see [11]. The next theorem characterizes the
W -inverses.
Theorem 2.3. Given A ∈ CR(H), B ∈ L(H) and W ∈ L(H)+, the following conditions are
equivalent:
i) The operator A admits a W -inverse in R(B),
ii) R(B) ⊆ R(A) + R(A)⊥W ,
iii) the normal equation A∗W (AX − B) = 0 admits a solution.
In particular, the pair (W, R(A)) is compatible if and only if the normal equation A∗W (AX−I) =
0 admits a solution.
Proof. See [9, Theorem 2.4]
Theorem 2.4. Given A ∈ CR(H) and W ∈ L(H)+. If the pair (W, R(A)) is compatible then,
for X0 ∈ L(H), the following conditions are equivalent:
i) X0 is a W -inverse of A,
ii) (AX0 − I)∗W (AX0 − I) = min≤
X∈L(H)
(AX − I)∗W (AX − I) = W/R(A),
iii) X0 is a solution of the normal equation A∗W (AX − I) = 0.
Proof. See [9, Proposition 4.4].
4
Section 3 is devoted to establish a link between the minus order and the shorted operator.
For this purpose, the concept of compatibility will be used.
3. Shorted operator, compatibility and minus order
Given a positive operator W ∈ L(H)+ and a closed subspace S ⊆ H the notion of shorted
operator of W to S, was introduced by M. G. Krein in [23] and later rediscovered by W. N.
Anderson and G. E. Trapp, who established a new characterization for this operator [1]. As it
was proved in [2], the set
M(W, S) = {X ∈ L(H) : 0 ≤ X ≤ W and R(X) ⊆ S ⊥}
has a maximum element, with respect to the order ≤, induced in L(H) by the cone of positive
operators. Then, the shorted operator of W to S is defined by
The S-compression WS of W is the (positive) operator defined by
W/S = max≤ M(W, S).
WS = W − W/S .
For many results on shorted operators, the reader is referred to [1] and [2]. Next we collect
some results regarding W/S and WS which will be useful in the rest of this work. If W ∈ L(H)+,
W 1/2 denotes the (positive) square root of W.
Theorem 3.1. Let W ∈ L(H)+ and S ⊆ H a closed subspace. Then
i) W/S = inf≤ {E∗W E : E ∈ Q, N (E) = S},
ii) W/S = W 1/2PW −1/2(S ⊥)W 1/2,
iii) R(W ) ∩ S ⊥ ⊆ R(W/S ) ⊆ R(W 1/2) ∩ S ⊥, and N (W/S ) = W −1/2(W 1/2(S)),
iv) W (S) ⊆ R(WS) ⊆ W (S), and N (WS ) = W −1(S ⊥).
The formula in ii) was stated by Pekarev, see [27]. The inclusions in iii) and iv) can be strict,
see [16].
Since the infimum appearing in i) of Theorem 3.1 is not attained, it is useful to establish a
condition when it is. This is given in the following theorem, see [15] and [16].
Theorem 3.2. Let W ∈ L(H)+ and S ⊆ H be a closed subspace. The following conditions are
equivalent:
i) The pair (W, S) is compatible,
ii) W/S = min {E∗W E : E ∈ Q, N (E) = S},
iii) R(W/S ) = R(W ) ∩ S ⊥ and N (W/S ) = N (W ) + S.
If any (and then all) of the above conditions holds, then
W/S = W (I − Q), for any Q ∈ P(W, S).
5
Minus order and compatibility
Different (but equivalent) definitions where given for minus order, for example, using generalized
−
≤ B if the Dixmier
inverses in the matrix case, see [24]. For operators A, B ∈ L(H), in [5], A
angle between R(A) and R(B − A) and the Dixmier angle between R(A∗) and R(B∗ − A∗) are
less than 1. In this work we give the following definition, equivalent to those appearing in [31]
and [18], where in [18, Theorem 3.3], a characterization of the minus order in terms of the range
additivity property is given.
Definition. Consider A, B ∈ L(H), we write A
such that A = P B and A∗ = QB∗.
−
≤ B if there exist (oblique) projections P, Q ∈ Q
The projections P and Q can be taken such that R(P ) = R(A) and R(Q) = R(A∗). It was
proven in [31] and [18] that
−
≤ is a partial order, known as the minus order for operators.
Theorem 3.3. Consider A, B ∈ L(H). Then the following assertions are equivalent:
−
≤ B,
i) A
ii) R(B) = R(A) +R(B − A) and R(B∗) = R(A∗) +R(B∗ − A∗).
The left minus order was defined in [18] for operators in L(H). Using Theorem 3.3, it is easy
to see that this notion is weaker than the minus order, in the infinite dimensional setting.
Definition. Consider A, B ∈ L(H), we write A − ≤ B if R(B) = R(A) +R(B − A).
If R(B) is closed, the left minus order and the minus order are equivalent, in the sense that
A − ≤ B if and only if A
−
≤ B, as shows the following result, see [18, Theorem 3.14].
Theorem 3.4. Let A, B ∈ L(H) such that A − ≤ B. If R(B) is closed then R(A) and R(B − A)
are closed and A
−
≤ B.
Recall that M(W, S) = {X ∈ L(H) : 0 ≤ X ≤ W and R(X) ⊆ S ⊥}.
Lemma 3.5. Let W ∈ L(H)+ and S ⊆ H be a closed subspace. Then
i) M(W, N (W/S )) = M(W, S),
ii) W/S = W/N (W/S ),
iii) W (N (W/S ))⊥ = W (S)⊥.
i) : The inclusion S ⊆ N (W/S ) (see iii) of Theorem 3.1) implies that M(W, N (W/S )) ⊆
Proof.
M(W, S). To see the opposite inclusion, take X ∈ M(W, S), then 0 ≤ X ≤ W/S . In this case,
applying Douglas' theorem [19], R(X 1/2) ⊆ R(W 1/2
/S ) ⊆ R(W/S ) = N (W/S )⊥. But, R(X) ⊆
R(X 1/2) and then X ∈ M(W, N (W/S )).
ii) : Using item i), we have that
W/N (W/S ) = max≤ M(W, N (W/S )) = max≤ M(W, S) = W/S .
iii) : It follows from item ii) that WS = WN (W/S) so that, from iv) of Theorem 3.1,
W (N (W/S ))⊥ = N (WN (W/S )) = N (WS ) = W (S)⊥.
6
The next result shows that the inequality W/S
−
≤ W is equivalent to the range inclusion
R(W/S ) ⊆ R(W ) and also to a compatibility condition.
Proposition 3.6. Let W ∈ L(H)+ and S ⊆ H be a closed subspace. Then the following condi-
tions are equivalent:
−
≤ W,
i) W/S
ii) the pair (W, N (W/S )) is compatible,
iii) H = N (W/S ) + N (WS ),
iv) R(W/S ) ⊆ R(W ).
i) ⇔ ii) : Suppose that W/S
−
≤ W, then there exists E ∈ Q with R(E∗) = R(W/S ), such
Proof.
that
W/S = E∗W = W E.
Then (I − E∗)W = W (I − E) and R(I − E) = N (W/S ) Hence, the pair (W, N (W/S )) is
compatible.
Conversely, suppose that the pair (W, N (W/S )) is compatible. Let Q ∈ P(W, N (W/S )) then
by Theorem 3.2 and Lemma 3.5,
W (I − Q) = W/N (W/S) = W/S ,
and W/S
−
≤ W.
ii) ⇔ iii) : By Theorem 2.1, (W, N (W/S )) is compatible if and only if
H = N (W/S ) + N (W/S )⊥W = N (W/S ) + S ⊥W = N (W/S ) + N (WS ),
where we used Lemma 3.5 and Theorem 3.1.
−
≤ W , applying Theorem 3.3, we have that R(W/S ) ⊆ R(W ).
i) ⇔ iv) : Suppose that W/S
Conversely, suppose that R(W/S ) ⊆ R(W ), then R(W ) = R(W/S ) + R(WS ).
To see that R(W/S ) ∩ R(WS ) = {0}, consider x ∈ R(W/S ) ∩ R(WS ). Since R(WS ) ⊆ R(W ),
it follows from iii) of Theorem 3.1 that R(W/S ) = W ∩ S ⊥. Then there exists y such that
x = W y and x ∈ S ⊥. Also, from iv) of Theorem 3.1, x ∈ W (S) so that there exists a sequence
{sn}n∈N ⊆ S such that W sn →
h y, x i . Hence,
n→∞
0 = h y, x i = kW 1/2yk2, and then W 1/2y = 0 and x = W y = 0. Therefore R(W ) = R(W/S ) ∔
x. In this case 0 = h x, sn i = h y, W sn i →
n→∞
R(WS ), and applying Theorem 3.3, it follows that W/S
−
≤ W.
Corollary 3.7. Let W ∈ L(H)+ and S ⊆ H be a closed subspace. Then the pair (W, S) is
compatible if and only if W/S
−
≤ W and W 1/2(S) is closed in R(W 1/2).
Proof.
If (W, S) is compatible then, by Theorem 3.2, N (W/S ) = S + N (W ). By Theorem 2.1,
(W, N (W/S )) is also compatible. Applying Proposition 3.6, it follows that W/S
−
≤ W.
7
Also, (W, S) is compatible if and only if H = S + W −1(S ⊥) and applying W 1/2 to both sides
of the equality, we get
R(W 1/2) = W 1/2S⊕(W 1/2S)⊥∩R(W 1/2) ⊆ W 1/2S∩R(W 1/2)⊕(W 1/2S)⊥∩R(W 1/2) ⊆ R(W 1/2).
Therefore, W 1/2S = W 1/2S ∩ R(W 1/2), so that W 1/2(S) is closed in R(W 1/2). See also, [16,
Proposition 3.8].
Conversely, suppose that W/S
−
≤ W and W 1/2(S) is closed in R(W 1/2). Then, by Proposition
3.6, the pair (W, N (W/S )) is compatible. But, by Theorem 3.1,
N (W/S ) = W −1/2(W 1/2(S) ∩ R(W 1/2)) = W −1/2(W 1/2(S)) = S + N (W ).
Therefore, by Theorem 2.1, the pair (W, S) is compatible.
The shorted operator can also be characterized as the maximum of certain set when the minus
order is considered. More precisely, for W ∈ L(H)+ and S ⊆ H a closed subspace, define
M−(W, S) = {X ∈ L(H) : X
−
≤ W, R(X) ⊆ S ⊥, R(X ∗) ⊆ S ⊥}.
In [24], Mitra proved (for matrices in Cm×n) that the shorted operator is the maximum of the
set M−(W, S), where the partial ordering is the minus order. In [5], Antezana et al. proved a
similar result for operators when the pair (W, S) is compatible. The next results generalize this
fact.
Lemma 3.8. Let W ∈ L(H)+ and S ⊆ H be a closed subspace. Then
M−(W, N (W/S )) = M−(W, S).
Proof. Since S ⊆ N (W/S ), then N (W/S )⊥ ⊆ S ⊥ and M−(W, N (W/S )) ⊆ M−(W, S).
On the other hand, let X ∈ M−(W, S), then X
−
≤ W and so, by Theorem 3.3, R(X) ⊆ R(W ).
Therefore, if R(X) ⊆ S ⊥ then W −1(R(X)) ⊆ W −1(S ⊥) = W −1(N (W/S )⊥). Then
W (W −1(R(X))) ⊆ N (W/S )⊥,
but W (W −1(R(X))) = R(X) ∩ R(W ) = R(X). Hence, R(X) ⊆ N (W/S )⊥. Analogously,
R(X ∗) ⊆ N (W/S )⊥, because, since W is positive, X
M−(W, S) ⊆ M−(W, N (W/S )).
−
≤ W implies that X ∗
−
≤ W . Therefore
Theorem 3.9. Let W ∈ L(H)+ and S ⊆ H be a closed subspace. Then the pair (W, N (W/S ))
is compatible if and only if
W/S = max −
≤
M−(W, S).
−
≤ W. By
Proof. Suppose the pair (W, N (W/S )) is compatible. By Proposition 3.6, W/S
Lemma 3.5, W/S = W/N (W/S), therefore R(W/S ) = R(W/N (W/S )) ⊆ N (W/S )⊥. Hence W/S ∈
8
M−(W, N (W/S )).
On the other hand, given X ∈ M−(W, N (W/S )), from X
−
≤ W, there exists E ∈ Q such that
X = EW. Let Q ∈ P(W, N (W/S )) then by Theorem 3.2, W/N (W/S ) = W (I − Q). The inclusion
R(X ∗) ⊆ N (W/S )⊥ = R(Q)⊥ = N (Q∗), implies that Q∗X ∗ = 0, or equivalently X(I − Q) = X.
Then
X = X(I − Q) = EW (I − Q) = EW/N (W/S ) = EW/S .
In a similar way, there exists a projection F such that X ∗ = F W/S . Therefore X
−
≤ W/S .
Then, by Lemma 3.8,
W/S = max −
≤
M−(W, N (W/S )) = max −
≤
M−(W, S).
Conversely, if W/S = max −
≤
M−(W, S), in particular W/S
−
≤ W, and by Proposition 3.6, the
pair (W, N (W/S )) is compatible.
4. Shorted operator characterizations and the minus order
Let W ∈ L(H)+ and A, B ∈ CR(H) such that N (B) = R(A)⊥W .
In [10, Proposition 3.1] it was proved that the infimum (in the order induced by the cone of
positive operators) of the set {(AXB − I)∗W (AXB − I) : X ∈ L(H)} exists and
inf≤
X∈L(H)
(AXB − I)∗W (AXB − I) = W/R(A).
(4.1)
Also in [10, Theorem 3.2], it was proved that the minimum of this set exists if and only if the
pair (W, R(A)) is compatible.
In this section, we study a similar problem considering the minus order: for W ∈ L(H)+ and
A, B ∈ CR(H), analyze the existence of
(AXB − I)∗W (AXB − I).
(4.2)
min −
≤
X∈L(H)
Problem (4.2) can be restated as a minimization problem with a constraint. In fact, problem
(4.2) is equivalent to the following problem: analize the existence of
(AX − I)∗W (AX − I) subject to N (A∗W ) ⊆ N (X).
min −
≤
X∈L(H)
9
Proposition 4.1. Let A, B ∈ CR(H) and W ∈ L(H)+ such that N (B) = R(A)⊥W . Then the
shorted operator W/R(A) is a lower bound for {(AXB − I)∗W (AXB − I) : X ∈ L(H)} if and
only if the pair (W, N (W/R(A))) is compatible.
In this case, if R(A) + R(A)⊥W is closed, then
W/R(A) = inf −
≤
X∈L(H)
(AXB − I)∗W (AXB − I).
Proof. Let X ∈ L(H), then
F (X) := (AXB − I)∗W (AXB − I) = W/R(A) + (AXB − I)∗WR(A)(AXB − I).
By Lemma 3.5, we have that W/N (W/R(A)) = W/R(A) and N (W/R(A))⊥W = R(A)⊥W .
Suppose the pair (W, N (W/R(A))) is compatible, let Q ∈ P(W, N (W/R(A))), then
F (X)(I − Q) = W/R(A)(I − Q) + (AXB − I)∗WR(A)(AXB − I)(I − Q) = W/R(A),
where we used that R(Q) = N (W/R(A)), then W/R(A)(I − Q) = W/R(A), and the facts that
R(I − Q) = N (Q) ⊆ N (W/R(A))⊥W = R(A)⊥W = N (B) and N (WR(A)) = R(A)⊥W . Therefore
F (X)(I − Q) = W/R(A) = (I − Q∗)F (X), for every X ∈ L(H).
Then
W/R(A)
−
≤ F (X), for every X ∈ L(H).
Hence, W/R(A) is a lower bound for {F (X) : X ∈ L(H)}.
Conversely, suppose W/R(A) is a lower bound for {F (X) : X ∈ L(H)}, in particular W/R(A)
−
≤
W, and by Proposition 3.6, the pair (W, N (W/R(A))) is compatible.
Finally, suppose that (W, N (W/R(A))) is compatible and R(A) + R(A)⊥W is closed. Then
R(A) + N (B) is closed and by [22, Corollary 2.5] R(BA) is closed. Also, since N (B) = N (A∗W ),
we have that N (BA) = N (A∗W A). Therefore R(A∗W 1/2) ⊆ R(A∗W A) = R(A∗B∗) =
R(A∗B∗). Then, by Theorem 2.2, there exists X0 ∈ L(H) such that W 1/2AX0BA = W 1/2A.
Then R(A) ⊆ N (W 1/2(AX0B − I)) = N (F (X0)), or equivalently
R(F (X0)) ⊆ R(A)⊥.
Let D ∈ L(H) be any lower bound for {F (X) : X ∈ L(H)}, then
−
≤ F (X), for every X ∈ L(H).
D
In particular, D
−
≤ W. Since D
−
≤ F (X0), by Theorem 3.3, R(D) ⊆ R(F (X0)) ⊆ R(A)⊥. From
−
≤ F (X0). Therefore, in the
D
same way as before, we get R(D∗) ⊆ R(A)⊥. Then D ∈ M−(W, S) and since (W, N (W/R(A)))
−
≤ F (X0) and the fact that F (X0) is positive we get that D∗
is compatible, by Theorem 3.9, D
−
≤ W/R(A).
10
For example, if W is injective, then W/R(A) = inf −
≤
X∈L(H)
(AXB − I)∗W (AXB − I). In fact, if
W is injective and the pair (W, N (W/R(A))) is compatible, by Theorem 2.1, H = N (W/R(A)) ∔
W −1(N (W/R(A))⊥) = N (W/R(A)) ∔ W −1(R(A)⊥), and since R(A) ⊆ N (W/R(A)), it follows that
R(A) ∔ W −1(R(A)⊥) is also closed and by the last theorem we get the result.
The next proposition shows that the minimum of (4.2) in the minus order is W/R(A) if and
only if the pair (W, R(A)) is compatible. First, we need the following lemma which shows that
when W/R(A) is in the image of the function G(X) = (AX − I)∗W (AX − I), the pair (W, R(A))
is compatible. For its proof, we follow the same ideas as in [9, Theorem 3.2].
Lemma 4.2. Let A ∈ CR(H) and W ∈ L(H)+, there exists X0 ∈ L(H) such that
W/R(A) = (AX0 − I)∗W (AX0 − I),
if and only if the pair (W, R(A)) is compatible.
Proof. Suppose that W/R(A) = (AX0 − I)∗W (AX0 − I), for certain X0 ∈ L(H). Writing W =
W/R(A) + WR(A), it follows that
W/R(A) = (AX0 − I)∗W (AX0 − I) = W/R(A) + (AX0 − I)∗WR(A)(AX0 − I),
because R(A) ⊆ N (W/R(A)). Therefore
(AX0 − I)∗WR(A)(AX0 − I) = 0, or, equivalently, W 1/2
R(A)(AX0 − I) = 0.
Then, by iv) of Theorem 3.1,
R(AX0 − I) ⊆ N (WR(A)) = W −1(R(A)⊥), or W (R(AX0 − I)) ⊆ R(A)⊥ ∩ R(W ).
Then
R(W/R(A)) = R((AX0 − I)∗W (AX0 − I)) ⊆ (AX0 − I)∗(R(A)⊥ ∩ R(W )) = R(A)⊥ ∩ R(W ),
because A∗(R(A)⊥) = 0. Then, by iii) of Theorem 3.1, R(W/R(A)) = R(A)⊥ ∩ R(W ).
Also x ∈ N (W/R(A)) if and only if W 1/2(AX0 − I)x = 0, or equivalently (AX0 − I)x ∈ N (W ).
In this case x ∈ N (W ) + R(A), and then again applying iii) of Theorem 3.1,
N (W/R(A)) = N (W ) + R(A).
Therefore R(W/R(A)) = R(A)⊥ ∩ R(W ) and N (W/R(A)) = N (W ) + R(A) and by Theorem 3.2,
the pair (W, R(A)) is compatible.
Conversely, if the pair (W, R(A)) is compatible, by Theorem 3.2, W/R(A) = W (I − Q) =
(I − Q)∗W (I − Q), for any Q ∈ P(W, R(A)). Consider X0 = A†Q, then AX0 = Q and F (X0) =
W/R(A).
We denote Sp, for some 1 ≤ p < ∞, the p-Schatten class, see [30]. Let W ∈ L(H)+, such that
11
W 1/2 ∈ Sp for some 1 ≤ p < ∞, in the following we define
kXkp,W = kW 1/2Xkp,
for any X ∈ L(H). For A, B ∈ CR(H) such that N (B) = R(A)⊥W , the next proposition,
shows the equivalence between the existence of minimum of problem (4.2) and the existence of
min
X∈L(H)
kAXB − Ikp,W .
Proposition 4.3. Let A, B ∈ CR(H) and W ∈ L(H)+ such that N (B) = R(A)⊥W . Then the
following are equivalent:
i) W/R(A) = min −
≤
X∈L(H)
(AXB − I)∗W (AXB − I),
ii) the pair (W, R(A)) is compatible,
iii) W/R(A) = min≤
X∈L(H)
(AXB − I)∗W (AXB − I).
If W 1/2 ∈ Sp for some 1 ≤ p < ∞, conditions i), ii) and iii) are also equivalent to
iv) kW 1/2
/R(A)kp = min
X∈L(H)
kAXB − Ikp,W .
i) ⇔ ii) : Suppose the pair (W, R(A)) is compatible. Then, by Theorem 2.1 and The-
Proof.
orem 3.2, N (W/R(A)) = R(A) + N (W ) and the pair (W, N (W/R(A))) is compatible. Thus, by
Proposition 4.1, W/R(A) is a lower bound for {F (X) : X ∈ L(H)}, where F (X) := (AXB −
I)∗W (AXB − I). It remains to see that there exists X0 ∈ L(H) such that W/R(A) = F (X0).
Let Q ∈ P(W, R(A)), then by Theorem 3.2, W/R(A) = (I − Q)∗W (I − Q). Observe that
R(W 1/2Q) = R(W 1/2A). On the other hand, since B ∈ CR(H) and N (W 1/2Q) = N (W Q) =
N (Q∗W ) = W −1(N (Q∗)) = W −1(R(Q)⊥) = R(A)⊥W = N (B), we have that R(Q∗W 1/2) ⊆
N (W 1/2Q)⊥ = N (B)⊥ = R(B∗). Then, by Theorem 2.2, there exists X0 ∈ L(H), such that
W 1/2AX0B = W 1/2Q. Thus
(AX0B − I)∗W (AX0B − I) = (I − Q)∗W (I − Q) = W/R(A),
and W/R(A) ∈ {F (X) : X ∈ L(H)}. Hence,
W/R(A) = min −
≤
X∈L(H)
F (X).
Conversely, suppose W/R(A) = min −
≤
X∈L(H)
F (X). Then, there exists X0 ∈ L(H), such that
(AX0B − I)∗W (AX0B − I) = min −
≤
X∈L(H)
F (X) = W/R(A).
Then, by Lemma 4.2, the pair (W, R(A)) is compatible.
ii) ⇔ iii) and ii) ⇔ iv) : follow from the fact that N (B) = N (A∗W ) and [10, Theorem 4.3].
12
Proposition 4.4. Let A, B ∈ CR(H) and W ∈ L(H)+ such that N (B) = R(A)⊥W and the pair
(W, R(A)) is compatible. Then, for X0 ∈ L(H), the following conditions are equivalent:
i) (AX0B − I)∗W (AX0B − I) = min −
≤
X∈L(H)
ii) A∗W (AX0B − I) = 0,
iii) X0B is a W -inverse of A.
(AXB − I)∗W (AXB − I) = W/R(A),
Proof. The equivalence between i), ii), iii) follows from the fact that X0 ∈ L(H) satisfies
(AXB − I)∗W (AXB − I) = W/R(A) if and only if
(AX0B − I)∗W (AX0B − I) = min −
≤
X∈L(H)
X0 satisfies (AX0B − I)∗W (AX0B − I) = min≤
X∈L(H)
Proposition 4.3) and [10, Theorem 3.2].
(AXB − I)∗W (AXB − I) = W/R(A) (see
5. Weighted star order and applications
In [20], Drazin introduced the star order on semigroups with involutions and in [8], Baksalary
and Mitra defined the left and right star orders on the set of complex matrices. Later, Antezana
et al. studied the star order on the algebra of bounded operators on a Hilbert space [4].
Given A, B ∈ L(H), the star order A
order A ≤ ∗ B are defined, respectively, by A
A∗A = A∗B and R(A) ⊆ R(B) and A ≤ ∗ B if AA∗ = BA∗ and R(A∗) ⊆ R(B∗).
∗
≤ B, the left star order A ∗ ≤ B and the right star
∗
≤ B if A∗A = A∗B and AA∗ = BA∗; A ∗ ≤ B if
If A, B ∈ L(H), then A
∗
≤ B if and only if there exist orthogonal projections P, Q such that
A = P B and A∗ = QB∗, see [4, Proposition 2.3].
Observe that A
∗
≤ B if and only A
−
≤ B, R(B − A) ⊆ R(A)⊥ and R(B∗ − A∗) ⊆ R(A∗)⊥.
More generally, given a positive operator W, we now introduce the weighted star order, which is
the minus order with an additional condition on the angle between the ranges of the operators
involved, imposed by the weight W.
Definition. Given A, B ∈ L(H) and W ∈ L(H)+, we say A ∗W ≤ B if A − ≤ B and R(B − A) ⊆
R(A)⊥W .
Analogously, we say A ≤ ∗W B if A ≤ − B and R(B∗ − A∗) ⊆ R(A∗)⊥W .
Finally, we say A
∗W
≤ B if A ∗W ≤ B and A ≤ ∗W B.
Proposition 5.1. Let A, B ∈ L(H) and W ∈ L(H)+, then A ∗W ≤ B if and only if R(B) =
R(A) ∔ R(B − A) and A∗W A = A∗W B.
Proof. Suppose that R(B) = R(A) ∔ R(B − A) and A∗W A = A∗W B, then by the definition
of − ≤ , A − ≤ B and since A∗W A = A∗W B it follows that R(B − A) ⊆ N (A∗W ) = R(A)⊥W ,
then A ∗W ≤ B.
Conversely, if A ∗W ≤ B, then A − ≤ B, or R(B) = R(A) ∔ R(B − A), and since R(B − A) ⊆
R(A)⊥W = N (A∗W ), we have A∗W A = A∗W B.
Proposition 5.2. Let W ∈ L(H)+, then the relations ∗W ≤ , ≤ ∗W and
∗W
≤ are partial orders.
13
Proof. Observe that ∗W ≤ , ≤ ∗W and
−
≤, ≤ − and − ≤ are antisymmetric (see [18, Proposition 3.11]).
∗W
≤ are clearly reflexive and also antisymmetric, because
Finally, if A ∗W ≤ B and B ∗W ≤ C, then A − ≤ B and it follows that R(A) ⊆ R(B), therefore
R(W 1/2A) ⊆ R(W 1/2B). Also A∗W A = A∗W B, so that W 1/2A ∗ ≤ W 1/2B. In the same way
W 1/2B ∗ ≤ W 1/2C, then since ∗ ≤ is transitive, W 1/2A ∗ ≤ W 1/2C and therefore A∗W A =
A∗W C. On the other hand, if A ∗W ≤ B and B ∗W ≤ C, then A − ≤ B and B − ≤ C, and since
− ≤ is transitive (see [18, Proposition 3.11]), we also have that A − ≤ C. Then, by Proposition
5.1, A ∗W ≤ C and the relation ∗W ≤ is a partial order. In a similar way, it can be proven that
∗W
≤ , ≤ ∗W are transitive.
Lemma 5.3. Let A, B ∈ L(H) and W ∈ L(H)+, such that N (W ) ∩ R(A) = {0} and the pair
(W, R(A)) is compatible. Then the following are equivalent:
i) A ∗W ≤ B,
ii) R(A) ⊆ R(B) and A = PW,R(A)B.
iii) R(A) ⊆ R(B) and A∗W A = A∗W B.
i) ⇒ ii) : Suppose A ∗W ≤ B, then R(B) = R(A)∔R(B −A) and R(B −A) ⊆ R(A)⊥W =
Proof.
N (PW,R(A)), therefore R(A) ⊆ R(B). Hence 0 = PW,R(A)(B − A) = PW,R(A)B − A = 0 and
A = PW,R(A)B.
ii) ⇒ iii) : Suppose R(A) ⊆ R(B) and A = PW,R(A)B. Then R(B − A) ⊆ N (PW,R(A)) ⊆
R(A)⊥W = N (A∗W ), therefore A∗W A = A∗W B.
iii) ⇒ i) : Suppose R(A) ⊆ R(B) and A∗W A = A∗W B. The inclusion R(A) ⊆ R(B) is
equivalent to R(B) = R(A) + R(B − A), see [6, Proposition 2.4].
On the other hand, from A∗W A = A∗W B, it follows that R(B − A) ⊆ N (A∗W ) = R(A)⊥W .
Finally, if y ∈ R(A) ∩ R(B − A), then there exists x, z ∈ H such that
Multiplying the last equation by A∗W, we get
y = Ax = (B − A)z.
A∗W Ax = A∗W (B − A)z = 0,
then W Ax = 0, so that Ax ∈ R(A) ∩ N (W ) = {0}, hence y = Ax = 0 and R(B) = R(A) ∔
R(B − A). Then A ∗W ≤ B.
A similar result can be stated for ≤ ∗W .
Corollary 5.4. Let A ∈ CR(H), B ∈ L(H) and W ∈ L(H)+ such that N (W ) ∩ R(A) = {0},
N (W )∩R(A∗) = {0} and the pairs (W, R(A)) and (W, R(A∗)) are compatible. Then the following
are equivalent:
∗W
≤ B,
i) A
ii) A∗ = PW,R(A∗)B∗ and A = PW,R(A)B,
iii) AW A∗ = BW A∗ and A∗W A = A∗W B.
Proof. Straightforwards.
Theorem 5.5. Let A, B ∈ L(H) such that R(A + B) is closed and C ∈ L(H). Let W ∈ L(H)+
14
such that (W, R(A + B)) is compatible, W 1/2 ∈ Sp for some 1 ≤ p < ∞, and A ∗W ≤ A + B.
Then the following statements are equivalent:
i) X0 ∈ L(H) satisfies
k(A + B)X0 − Ckp,W = min
X∈L(H)
k(A + B)X − Ckp,W ;
ii) X0 ∈ L(H) satisfies
kAX0 − Ckp,W = min
X∈L(H)
kBX0 − Ckp,W = min
X∈L(H)
kAX − Ckp,W
kBX − Ckp,W .
Proof. First observe that if R(A + B) is closed and A ∗W ≤ A + B, by Proposition 5.1 and
Corollary 3.4, we have A∗W B = 0 and A
−
≤ A + B.
Now, suppose item i) holds, then by [9, Theorem 4.5], X0 is a solution of
(A∗ + B∗)W ((A + B)X − C) = 0.
Therefore, R(W (A + B)X0 − C) ⊆ N (A∗ + B∗) = N (A∗) ∩ N (B∗), because A
−
≤ A + B. Then
A∗W ((A + B)X0 − C) = B∗W ((A + B)X0) − C) = 0,
and since A∗W B = 0, it holds that
A∗W (AX0 − C) = B∗W (BX0 − C) = 0,
or equivalently by [9, Theorem 4.5], X0 ∈ L(H) satisfies
kAX0 − Ckp,W = min
X∈L(H)
kBX0 − Ckp,W = min
X∈L(H)
kAX − Ckp,W
kBX − Ckp,W .
Conversely, suppose item ii) holds, then by [9, Theorem 4.5], X0 is a solution of the system
(cid:26) A∗W (AX − C) = 0
B∗W (BX − C) = 0,
therefore, since A∗W B = 0, we have (A∗ + B∗)W ((A + B)X0 − C) = A∗W (AX0 − C) +
B∗W (BX0 − C) = 0. Then X0 is as solution of (A∗ + B∗)W ((A + B)X − C) = 0 and by [9,
Theorem 4.5], we have item i).
Corollary 5.6. Let A, B ∈ L(H) such that R(A + B) is closed. Let W ∈ L(H)+ such that the
pair (W, R(A + B)) is compatible and A ∗W ≤ A + B. Then X0 is a W -inverse of A + B in R(C)
if and only if X0 is a W -inverse of A in R(C) and a W -inverse of B in R(C),
Proof. By Theorem 2.3, X0 is W -inverse of A + B in R(C) if and only if X0 is a solution of
(A∗ + B∗)W ((A + B)X − C) = 0,
15
if and only if, by the proof of Theorem 5.5, X0 is a solution of the system
(cid:26) A∗W (AX − C) = 0,
B∗W (BX − C) = 0,
or equivalently, again by Theorem 2.3, X0 is a W -inverse of A in R(C) and a W -inverse of B in
R(C).
Corollary 5.7. Let A, B ∈ L(H) with R(A+B) closed. Let W ∈ L(H)+ such that (W, R(A+B))
is compatible and A ∗W ≤ A + B. Then
WR(A+B) = WR(A) + WR(B).
It follows from Theorem 2.3, for C = I, that (W, R(A + B)) is compatible if and only if
Proof.
there exists X0 such that
(A∗ + B∗)W ((A + B)X0 − I) = 0.
By Corollary 5.6, this is equivalent to
(cid:26) A∗W (AX0 − I) = 0,
B∗W (BX0 − I) = 0.
(5.1)
(5.2)
On the other hand, by Theorem 2.4, equation (5.1) and (5.2) are equivalent to the equalities
W/R(A+B) = ((A + B)X0 − I)∗W ((A + B)X0 − I),
W/R(A) = (AX0 − I)∗W (AX0 − I) and W/R(B) = (BX0 − I)∗W (BX0 − I).
But, using the fact that A ∗W ≤ A + B, it follows that
W/R(A+B) = ((A + B)X0 − I)∗W ((A + B)X0 − I) =
= (AX0 − I)∗W (AX0 − I) + (BX0 − I)∗W (BX0 − I) − W = W/R(A) + W/R(B) − W.
Therefore,
WR(A+B) = WR(A) + WR(B).
Acknowledgements
This research was partially supported by CONICET PIP 0168.
16
References
[1] Anderson W.N., Shorted Operators, SIAM J. Appl. Math, 20 (1971), 520-525.
[2] Anderson W.N., Trapp G.E, Shorted Operators II, SIAM J. Appl. Math, 28 (1975), 60-71.
[3] Ando T., Generalized Schur complements, Linear Algebra Appl. 27 (1979), 173-186.
[4] Antezana J. , Cano C., Mosconi I., Stojanoff D. , A note on the star order in Hilbert spaces,
Linear and Multilinear Algebra 58 8 (2010), 1037-1051.
[5] Antezana J., Corach G., Stojanoff D. Bilateral shorted operators and parallel sums, Linear
Algebra Appl. 414 2 (2006), 570-588.
[6] Arias M.L., Corach G., Maestripieri A. Range additivity, shorted operator and the
Sherman-Morrison-Woodbury formula, Linear Algebra Appl., 467 (2015), 86-99.
[7] Arias M.L., Gonzalez M.C., Positive solutions to operator equations AXB = C, Linear
Algebra and its Applications, 433 (2010), 1194-1202.
[8] Baksalary J.K., Mitra S. K., Left-star and right-star partial orderings, Linear Algebra
Appl. 149 (1991), 73-89.
[9] Contino M., Giribet J.I, Maestripieri A., Weighted Procrustes problems, J. Math. Anal.
Appl., 445 (2017), 443458.
[10] Contino M., Giribet J.I, Maestripieri A., Weighted least square solutions of the equation
AXB − C = 0, Linear Algebra Appl. 518 (2017), 177-197.
[11] Corach G., Fongi G., Maestripieri A., Weighted projections into closed subspaces, Studia
Mathematica, 216 (2013), 131-148.
[12] Corach G., Giribet J. I., Oblique projections and sampling problems. Integral Equations
and Operators Theory, 70, (2011), 307-322.
[13] Corach G., Giribet J. I., Sard's approximation processes and oblique projections. Studia
Mathematica 194 (2009), 65-80.
[14] Corach G., Maestripieri A., Weighted generalized inverses, oblique projections and least
squares problems, Numer. Funct. Anal. Optim., 6 (2005), 659-673.
[15] Corach G., Maestripieri A., Stojanoff D., Oblique projections and Schur complements, Acta
Sci. Math. (Szeged), 67 (2001), 439-459.
[16] Corach G., Maestripieri A., Stojanoff D., Generalized orthogonal projections and shorted
operators, Margarita Mathem´atica, Departamento de Matem´aticas y Computaci´on, Uni-
versidad de La Rioja, (2001), 607-625.
[17] Corach G., Maestripieri A., Stojanoff D., Oblique projections and abstract splines, Journal
of Approximation Theory, 117 (2002), 189-206.
[18] Djiki´c, M.S., Fongi G., Maestripieri A., The minus order and range additivity, Linear
Algebra Appl. 531 (2017), 234-256.
[19] Douglas R. G., On majorization, factorization and range inclusion of operators in Hilbert
space, Proc. Amer. Math. Soc., 17 (1966), 413-416.
[20] Drazin M.P., Natural structures on semigroups with involution, Bulletin of the American
Mathematical Society, 84, 1 (1978), 139-141.
[21] Hartwig R.E., How to partially order regular elements, Japanese Journal of Mathematics,
25 (1980), 1-13.
[22] Izumino S., The product of operators with closed range and an extension of the reverse
order law, Tohoku Math. J., 34 (1982), 43-52.
[23] Krein M.G., The theory of self-adjoint extensions of semibounded Hermitian operators and
its applications, Mat. Sb. (N.S.), 20, 62 (1947), 431-495.
[24] Mitra S.K., The minus partial order and the shorted matrix, Linear Algebra Appl. 83
(1986), 1-27.
[25] Mitra, S. K., Rao, C. R., Projections under seminorms and generalized Moore Penrose
inverses and operator ranges, Linear Algebra Appl., 9 (1974), 155-167.
17
[26] Mitra, S. K., Matrix partial order through generalized inverses: unified theory, Linear Al-
gebra Appl. 148 (1991), 237-263.
[27] Pekarev E. L., Shorts of operators and some extremal problems, Acta Sci. Math.(Szeged),
56 (1992), 147-163.
[28] Mitra, S. K., Rao, C. R., Theory and application of constrained inverse of matrices, SIAM
Journal on Applied Mathematics, 24 (1973), 473-488.
[29] Nambooripad S. K., The natural partial order on a regular semigroup, Proceedings of the
Edinburgh Mathematical Society (Series 2) 23 03 (1980), 249-260.
[30] Ringrose J.R., Compact non-self-adjoint operators, Van Nostrand Reinhold Co., 1971.
[31] Semrl P., Automorphisms of B(H) with respect to minus partial order, Journal of Mathe-
matical Analysis and Applications 369 1 (2010), 205-213.
[32] Zhang F., The Schur complement and its applications, Vol. 4., Springer Science & Business
Media, 2006.
18
|
1208.3699 | 1 | 1208 | 2012-08-17T22:04:29 | On discrete analytic functions: Products, Rational Functions, and some Associated Reproducing Kernel Hilbert Spaces | [
"math.FA",
"math.CV"
] | We introduce a family of discrete analytic functions, called expandable discrete analytic functions, which includes discrete analytic polynomials, and define two products in this family. The first one is defined in a way similar to the Cauchy-Kovalevskaya product of hyperholomorphic functions, and allows us to define rational discrete analytic functions. To define the second product we need a new space of entire functions which is contractively included in the Fock space. We study in this space some counterparts of Schur analysis. | math.FA | math |
ON DISCRETE ANALYTIC FUNCTIONS: PRODUCTS,
RATIONAL FUNCTIONS, AND SOME ASSOCIATED
REPRODUCING KERNEL HILBERT SPACES
DANIEL ALPAY, PALLE JORGENSEN, RON SEAGER, AND DAN VOLOK
Abstract. We introduce a family of discrete analytic functions, called ex-
pandable discrete analytic functions, which includes discrete analytic polyno-
mials, and define two products in this family. The first one is defined in a way
similar to the Cauchy-Kovalevskaya product of hyperholomorphic functions,
and allows us to define rational discrete analytic functions. To define the sec-
ond product we need a new space of entire functions which is contractively
included in the Fock space. We study in this space some counterparts of Schur
analysis.
Contents
Introduction
1.
2. Polynomials and rational functions on the set of integers
3. Discrete polynomials of two variables
4. Discrete analytic polynomials
5. Expandable discrete analytic functions
6. The Cauchy-Kovalevskaya product
7. Rational discrete analytic functions
8. The C∗-algebra associated to expandable discrete analytic functions
9. A reproducing kernel Hilbert space of entire functions
10. A reproducing kernel Hilbert space of expandable discrete analytic
function
References
1
4
7
10
12
14
16
18
21
26
28
1. Introduction
In this paper, we explore a spectral theoretic framework for representation of dis-
crete analytic functions. While the more familiar classical case of analyticity plays
1991 Mathematics Subject Classification. Primary 30G25, 30H20, 32A26, 43A22 , 46E22,
46L08, 47B32, 47B39; secondary: 20G43.
Key words and phrases. Discrete analytic functions, 2D lattice Z2, reproducing kernel Hilbert
space, Szego and Bergman, multipliers, Cauchy integral representation, difference operators, Lie
algebra of operators, Fourier transform, realizable linear systems, expandable functions, rational
functions, Cauchy-Riemann equations, Cauchy-Kovalevskaya theorem, Schur analysis, Fock space.
D. Alpay thanks the Earl Katz family for endowing the chair which supported his research.
The research of the authors was supported in part by the Binational Science Foundation grant
2010117.
1
2
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
an important role in such applications as the theory of systems and their realiza-
tions, carrying over this to the case of discrete analytic functions involves a number
of operator- and spectral theoretic subtleties; for example, we show that the repro-
ducing kernel, in the discrete case, behaves quite differently from the case of the
more familiar classical kernels of Szego and Bergman. We introduce a reproduc-
ing kernel of discrete analytic functions, which is naturally isomorphic to a Hilbert
space of entire functions contractively included in the Fock space. The pointwise
product of two discrete analytic functions need not be discrete analytic, and we
introduce two products, each taking into account the specificities of discrete ana-
lyticity.
The first product is determined by a solution to an extension question for rational
functions, extending from Z+ to the right half-plane in the 2D lattice Z2. Our
solution to the extension problem leads to a new version of the multiplication op-
erator Z. We further prove that the product in A will be defined directly from
Z. This in turn yields a representation of the multiplier problem for the repro-
ducing kernel Hilbert space H. While it is possible to think of the reproducing
kernel Hilbert space H as an extension of one of the classical Beurling-Lax theory
for Hardy space, the case for discrete analytic function involves a new and different
spectral analysis, departing from the classical case in several respects. For example,
we show that the new multiplication operator Z is part of an infinite-dimensional
non-Abelian Lie algebra of operators acting on the reproducing kernel Hilbert space
H. With this, we are able to find the spectral type of the operators described above.
The theory of discrete analytic functions has drawn a lot of attention recently, in
part because of its connections with electrical networks and random walks. In the
case of functions defined on the integer grid the notion of discrete analyticity was
introduced by J. Ferrand (Lelong) in [15]:
Definition 1.1. A function f : Z2 −→ C is said to be discrete analytic if
(1.1)
f (x + 1, y) − f (x, y + 1)
.
∀(x, y) ∈ Z2,
The properties of discrete analytic functions were extensively investigated by R. J.
Duffin in [13]. In this work it was shown that discrete analytic functions share many
important properties of the classical continuous analytic functions in the complex
domain, such as Cauchy integral representation and the maximum modulus prin-
ciple. More recently, the notion of discrete analyticity and accompanying results
were extended by C. Mercat to the case of functions defined on arbitrary graph
embedded in an orientable surface; see [22].
f (x + 1, y + 1) − f (x, y)
=
1 + i
1 − i
The concept of discrete analyticity seems to cause significant difficulties in the
following regard: the pointwise product of two discrete analytic functions is not
necessarily discrete analytic. For example, the functions z := x + iy and z2 are
discrete analytic in the sense of Definition 1.1, but z3 is not. Thus a natural ques-
tion arises, how to describe all complex polynomials in two variables x, y whose
restriction to the integer grid Z2 is discrete analytic, and, more generally, rational
discrete analytic functions. This problem was originally considered by R. Isaacs,
using a definition of discrete analyticity (the so-called monodiffricity), which is al-
gebraically simpler than Definition 1.1. In [17] R. Isaacs has posed a conjecture that
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
3
all monodiffric rational functions are polynomials. This conjecture was disproved
by C. Harman in [16], where an explicit example of a non-polynomial monodiffric
function, rational in one quadrant, was constructed.
The results of R. Isaacs and C. Harman suggest that in the setting of discrete an-
alytic functions the notion of rationality based on the pointwise product is not a
suitable one. In order to introduce a class of rational discrete analytic functions,
which would be sufficiently rich for applications, one needs a suitable definition of
the product. This is one of the main objective of the present paper to introduce
two products in the setting of discrete analytic functions.
Organization: The paper is organized as follows. Sections 2 through 4 cover our
preparation of the discrete framework: analysis and tools. In Definition 2.1, the
notion of discrete analyticity makes a key link between the representation in the
two integral variables (x, y) in the 2-lattice Z2 , thus making precise the interaction
between the two integral variables x and y implied by analyticity. The notion of
analyticity in the discrete case is a basic rule (Definition 1.1) from which one makes
precise contour-summations around closed loops in Z2. In section 2, we introduce a
basis system of polynomials (which will appear to be restriction to the positive real
axis of discrete analytic polynomials ζn defined in section 5), see equation (2.1). We
further introduce a discrete Fourier transform for functions of (x, y) in the 2-lattice
Z2, and in the right half-plane H+ = Z+×Z in Z2. The Fourier representation in H+
is then used in sections 3 and 4; where we study extensions from Z to Z2, and from
Z+ to H+. We begin our analysis in section 4 with some lemmas for the polyno-
mial case. Theorem 4.1 offers a discrete version of the Cauchy-Riemann equations;
we show that the discrete analytic functions are defined as the kernel of a del-bar
operator D; to be studied in detail in section 7, as part of a Lie algebra represen-
tation. In Theorem 4.2 we show that every polynomial function on Z has a unique
discrete analytic extension to Z2. Sections 5 and 6 deal with expandable functions
(Definition 5.5), and section 7 rational discrete analytic functions. The expand-
able functions are defined from a basis system of discrete analytic polynomials ζn
from section 2, and a certain Cauchy-estimate, equation (5.10). We shall need two
products defined on expandable functions, the first is our Cauchy-Kovalesvskaya
product (in section 6), and the second (section 10) is defined on algebra gener-
ated by the discrete analytic polynomials ζn. Its study makes use of realizations
from linear systems theory. The definition of the Cauchy-Kovalesvskaya product
relies on uniqueness of extensions for expandable functions (Corollary 5.6). Hence
it is defined first for expandable functions, and then subsequently enlarged; first to
the discrete analytic rational functions (section 7), and then to a new reproducing
kernel Hilbert space in section 8. The latter reproducing kernel Hilbert space has
its kernel defined from the discrete analytic polynomials ζn; see (8.1). The study
of the reproducing kernel Hilbert space in turns involves such tools from analysis
as representations of Lie algebras (Theorem 7.4), and of C∗-algebras (Theorem 8.4).
(2.3)
where
(2.4)
f (x) =
∞Xn=1bf (n)x[n],
n→∞ (cid:16)n!bf (n)(cid:17) 1
lim sup
n ≤ 1.
4
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
2. Polynomials and rational functions on the set of integers
In what follows, Ω stands for one of two sets: Z or Z+, and x[n] denotes the
polynomial of degree n defined by
(2.1)
(if n = 0, x[0] := 1).
x[n] :=
n−1Yj=0
(x − j)
They have the generating function
(2.2)
(1 + t)x =
x[k] tk
k!
,
xXk=0
and their discrete analytic extensions ζn(x, y) are studied in section 5.
The purpose of this section is to prove (see Theorem 2.9 below) that any rational
function f : Z+ −→ C (see Definition 2.7) has a unique representation
Definition 2.1. The linear difference operator δ on the space of functions f :
Ω −→ C is defined by
(δf )(x) = f (x + 1) − f (x),
x ∈ Ω.
Proposition 2.2. Let f : Z+ −→ C. Then, for every x ∈ Z+ the series
f (x) := Xn∈Z+
f (n)x[n]
has a finite number of non zero terms.
Proof. In view of (2.1),
∀x ∈ Z+, ∀n ∈ Z+,
x < n =⇒ x[n] = 0.
Therefore, for every x ∈ Z+ the series f (x) contains at most x + 1 non-zero terms.
(cid:3)
Proposition 2.3. Let f : Z+ −→ C. Then there exists a unique function f :
Z+ −→ C such that
(2.5)
f (n)x[n].
∀x ∈ Z+,
f (x) = Xn∈Z+
The function f (n) is given by
(2.6)
f (n) =
(δnf )(0)
n!
.
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
5
Proof. To show the uniqueness of f (n), assume that (2.5) holds for some function
f (n) and apply the identity
(2.7)
δx[n] = nx[n−1]
repeatedly to obtain (2.6).
Next, let the function f (n) be defined by (2.6). Then, according to Proposition 2.2,
the series
f (x) := Xn∈Z+
f (n)x[n]
converges absolutely for every x ∈ Z+ and defines a function f : Z+ −→ C. Hence
(2.6) holds with f replacing f and
∀n ∈ Z+,
(δnf )(0) = n! f (n) = (δn f )(0).
Now one can verify that
(2.8)
∀x ∈ Z+, ∀n ∈ Z+,
by induction on x: If
(δnf )(x) = (δn f )(x)
∀n ∈ Z+,
(δnf )(x) = (δn f )(x)
then
∀n ∈ Z+,
(δnf )(x + 1) = (δnf )(x) + (δn+1f )(x)
= (δn f )(x) + (δn+1 f )(x) = (δn f )(x + 1).
It remains to set n = 0 in (2.8) to obtain (2.5).
Definition 2.4. Let f : Z+ −→ C. Then the function f : Z+ −→ C, defined by
(2.6), is said to be the Fourier transform of f .
Suppose that a function f : Z −→ C is such that
(cid:3)
∀x ∈ Z,
f (x) = p(x),
where p(x) is a polynomial with complex coefficients. Then such a polynomial p(x)
is unique, and we shall call the function f itself a polynomial. If f (x) 6≡ 0, the
degree of f (x) is the same as the degree of p(x).
Proposition 2.5. Let f : Z −→ C be a polynomial, and let f : Z+ −→ C be the
. Then the function f has a finite support
Fourier transform of the restriction fZ+
and
∀x ∈ Z,
Proof. Note that
(if f = const, δf = 0). Hence
∀n ∈ Z+,
Since
f (x) = Xn∈Z+
f (n)x[n].
deg(δf ) = deg(f ) − 1
n > deg(f ) =⇒ δnf = 0.
(δf )Z+
= δ(fZ+
),
6
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
(2.6) implies that
It follows that
∀n > deg(f ),
f (n) = 0.
f (n)x[n]
Xn∈Z+
is, in fact, a polynomial, which coincides with the polynomial f (x) on Z+ and hence
on Z.
(cid:3)
It follows from Proposition 2.5 and identity (2.7) that if f : Z −→ C is a polynomial
then so is δf . A converse statement can be formulated as follows:
Proposition 2.6. Let f : Z −→ C be a polynomial. Then there exists a polynomial
g : Z −→ C such that f (x) ≡ (δg)(x). If f (x) 6≡ 0 then deg(g) = deg(f ) + 1.
Proof. By Proposition 2.5,
where f : Z+ −→ C is a function with finite support. Consider the polynomial
f (n)x[n],
f (x) = Xn∈Z+
g(x) = Xn∈Z+
(δg)(x) = Xn∈Z+
f (n)
n + 1
x[n+1],
f (n)x[n] = f (x).
(cid:3)
then, in view of (2.7),
∀x ∈ Z,
Definition 2.7. A function f : Z+ −→ C is said to be rational if there exist
polynomials p, q : Z+ −→ C such that
∀x ∈ Z+,
q(x) 6= 0,
and
∀x ∈ Z+,
f (x) =
p(x)
q(x)
.
Proposition 2.8. Let f : Z+ −→ C be given. Then f (x) is a rational function if
and only if there exist a polynomial p(x) and matrices A, B, C such that
(2.9)
and
(2.10)
σ(A) ∩ Z+ = ∅,
f (x) = p(x) + C(xI − A)−1B.
Proof. We first recall that a matrix-valued rational function of a complex variable
can always be written in the form
(2.11)
r(z) = p(z) + C(zI − A)−1B,
where the matrix-valued polynomial p takes care of the pole at infinity, and A, B
and C are matrices of appropriate sizes. Furthermore, when the dimension of A is
minimal, the (finite) poles of r coincide with the spectrum of A; see [8, 19]. Here,
we consider complex-valued functions, and thus C and B are respectively a row
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
7
and column vector.
Now, by Definition 2.7, f is a rational function if and only if it is the restriction
on Z+ of a rational function of a complex variable, with no poles on Z+, that is
if and only if it can be written as (2.10) with the matrix A satisfying furthermore
(2.9).
(cid:3)
Theorem 2.9. Let f : Z+ −→ C be rational, and let f : Z+ −→ C be the Fourier
transform of f . Then
lim sup
n→∞
( f (n)n!)1/n ≤ 1.
Proof. According to Proposition 2.8, there exist a polynomial p(x) and matrices
A, B, C such that σ(A) ∩ Z+ = ∅ and
Hence, for n > deg(p),
f (x) = p(x) + C(xI − A)−1B.
nYj=1
k(cid:18)I −
1
n
A(cid:19)−1
f (n)n! = (δnf )(0) ≤ kCk · kBk · kA−1k ·
A(cid:19)−1
n ≥ N =⇒ f (n)n! ≤ M (1 + ǫ)n,
n→∞ k(cid:18)I −
∀ǫ > 0, ∃M, N ∈ Z+, ∀n ∈ Z+
k = 1,
1
n
lim
k.
:
Since
and the conclusion follows.
(cid:3)
3. Discrete polynomials of two variables
Definition 3.1. The linear difference operators δx, δy on the space of functions
f : Ω1 × Ω2 −→ C are defined by
(δxf )(x, y) := f (x + 1, y) − f (x, y),
(δyf )(x, y) := f (x, y + 1) − f (x, y).
Note that the difference operators δx and δy commute:
(3.1)
(δxδyf )(x, y) = (δyδxf )(x, y) = f (x + 1, y + 1) − f (x, y + 1) − f (x + 1, y) + f (x, y).
Proposition 3.2. Let f : Z2
+ −→ C. Then for every (x, y) ∈ Z2
+, the series
(3.2)
f (x, y) := X(m,n)∈Z2
+
f (m, n)x[m]y[n]
contains finitely many non-zero terms.
Proof. In view of (2.1),
∀x ∈ Z+, ∀n ∈ Z+,
x < n =⇒ x[n] = 0.
Therefore, for every (x, y) ∈ Z2
non-zero terms.
+ the series f (x) contains at most (x + 1)(y + 1)
(cid:3)
Formula (3.2) can be viewed as a transform of a discrete function. The inverse
transform is calculated in the next proposition.
8
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
Proposition 3.3. Let f : Z2
Z2
+ −→ C such that
(3.3)
∀(x, y) ∈ Z2
+,
+ −→ C. Then there exists a unique function f :
f (x, y) = X(m,n)∈Z2
f (m, n)x[m]y[n].
+
The function f (m, n) is given by
(3.4)
f (m, n) =
(δm
x δn
y f )(0, 0)
m!n!
,
(m, n) ∈ Z2
+.
Proof. First, fix x ∈ Z+ and consider the function fx : Z+ −→ C given by
fx(y) = f (x, y),
y ∈ Z+.
Then, according to Proposition 2.3, there is a unique function fx : Z+ −→ C such
that
∀y ∈ Z+,
the function fx(n) is given by
fx(y) = Xn∈Z+
fx(n)y[n];
fx(n) =
(δn
y f )(x, 0)
n!
, n ∈ Z+.
Next, fix n ∈ Z+ and consider the function gn : Z+ −→ C given by
gn(x) = fx(n),
x ∈ Z+.
By the same Proposition 2.3, there is a unique function gn : Z+ −→ C such that
∀x ∈ Z+,
the function gn(m) is given by
gn(m) =
gn(x) = Xm∈Z+
gn(m)x[m];
(δm
x δn
y f )(0, 0)
m!n!
, m ∈ Z+.
Thus
∀(x, y) ∈ Z2
+,
f (x, y) = X(m,n)∈Z2
+
gn(m)x[m]y[n].
It remains to set f (m, n) = gn(m).
(cid:3)
Definition 3.4. Let f : Z2
(3.4), is said to be the Fourier transform of f .
+ −→ C. Then the function f : Z2
+ −→ C, defined by
Theorem 3.5. Let p(z, w) be a complex polynomial in two variables. Then, p(cid:12)(cid:12)Z2 =
0 if and only if p ≡ 0.
Proof. Write
(3.5)
p(z, w) =
pn(z)wn,
NXn=0
where the pn are polynomials in z. The equations p(z, w) ≡ 0 for w = 0, 1, . . . N
lead to (using a Vandermonde determinant) that p0(z), p1(z), . . . , pN (z) vanish on
Z and hence identically.
(cid:3)
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
9
Suppose that a function f : Z2 −→ C is such that
∀(x, y) ∈ Z2,
f (x, y) = p(x, y),
where p(x, y) is a polynomial with complex coefficients. Then in view of Theorem
3.5, such a polynomial p(x, y) is unique, and we shall call the function f itself a
polynomial. If f (x, y) 6≡ 0, the degree of f (x, y) is the same as the degree of p(x, y).
Proposition 3.6. Let f : Z2 −→ C be a polynomial, and let f : Z2
+ −→ C be the
. Then the function f has a finite support
Fourier transform of the restriction fZ2
and
+
∀(x, y) ∈ Z2,
f (x, y) = X(m,n)∈Z2
+
f (m, n)x[m]y[n].
Proof. First, in view of (3.4) and of the fact that
∀(m, n) ∈ Z2
+, ∀(x, y) ∈ Z2, m + n > deg(f ) =⇒ (δm
x δn
y f )(x, y) = 0,
the function f has a finite support. It follows that
X(m,n)∈Z2
+
f (m, n)x[m]y[n]
+ and
(cid:3)
is, in fact, a polynomial, which coincides with the polynomial f (x, y) on Z2
hence on Z2.
It follows from Proposition 3.6 and identity (2.7) that if f : Z2 −→ C is a polynomial
then so are δxf and δyf . A converse statement can be formulated as follows. It
will be used in the proof of Theorem 4.2,
Proposition 3.7. Let f, g : Z2 −→ C be two polynomials, such that
(δyf )(x, y) ≡ (δxg)(x, y).
(3.6)
Then there exists a polynomial h : Z2 −→ C such that
(3.7)
(δxh)(x, y) ≡ f (x, y),
(δyh)(x, y) ≡ g(x, y).
Proof. By Proposition 3.6, there exist functions f , g : Z2
such that
+ −→ C with finite support,
f (x, y) = X(m,n)∈Z2
+
f (m, n)x[m]y[n],
g(m, n)x[m]y[n].
g(x, y) = X(m,n)∈Z2
+
Then identity (3.6) implies that
(3.8)
∀(m, n) ∈ Z2
+,
Consider the polynomial
h(x, y) = X(m,n)∈Z2
+
(n + 1) f (m, n + 1) = (m + 1)g(m + 1, n).
f (m, n)
m + 1
x[m+1]y[n] + Xn∈Z+
g(0, n)
n + 1
y[n+1],
then
∀(x, y) ∈ Z2,
(δxh)(x, y) = X(m,n)∈Z2
+
f (m, n)x[m]y[n] = f (x, y).
10
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
On the other hand, in view of (3.8),
h(x, y) = X(m,n)∈Z2
+
g(m, n)
n + 1
x[m]y[n+1] + Xn∈Z+
f (m, 0)
m + 1
x[m+1],
hence
∀(x, y) ∈ Z2,
(δyh)(x, y) = X(m,n)∈Z2
+
g(m, n)x[m]y[n] = g(x, y).
4. Discrete analytic polynomials
(cid:3)
Difference operators: It is convenient to recast Definition 1.1 in terms of the
difference operators. In what follows, each of the sets Ω1, Ω2 is either Z or Z+.
Theorem 4.1. A function f : Ω1 × Ω2 −→ C is discrete analytic if and only if
∀(x, y) ∈ Ω1 × Ω2,
( ¯Df )(x, y) = 0,
where
(4.1)
Proof. In view of (3.1),
¯D := (1 − i)δx + (1 + i)δy + δxδy.
∀(x, y) ∈ Ω1 × Ω2,
f (x + 1, y + 1) − f (x, y)
1 + i
−
f (x + 1, y) − f (x, y + 1)
1 − i
1 − i
2
=
1 − i
2
=
(f (x + 1, y + 1) − f (x, y) − if (x + 1, y) + if (x, y + 1))
((δxδyf )(x, y) − 2f (x, y) + (1 − i)f (x + 1, y) + (1 + i)f (x, y + 1))
=
1 − i
2
((δxδyf )(x, y) + (1 − i)(δxf )(x, y) + (1 + i)(δyf )(x, y)).
(cid:3)
Extension: In view of Definition 1.1, given f0 : Z −→ C there are infinitely dis-
crete analytic functions f on Z2 such that f (x, 0) = f0(x). However, the following
theorem show that in the case when f0 is a polynomial, only one of these discrete
analytic extensions will be a polynomial in x, y. The result itself originates with
the work of Duffin [13], and we give a new proof.
Theorem 4.2. Let p : Z −→ C be a polynomial. Then there exists a unique discrete
analytic polynomial q : Z2 −→ C such that
∀x ∈ Z,
q(x, 0) = p(x).
In particular, q(x, y) ≡ 0 if and only if p(x) ≡ 0. If this is not the case,
deg(q) = deg(p).
Lemma 4.3. Let q : Z2 −→ C be a discrete analytic polynomial, such that q(x, 0) ≡
0. Then q(x, y) ≡ 0.
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
11
Proof. Assume the opposite, then deg(q) > 0, and q can be chosen so that deg q is
the smallest possible. Observe that (δxq)(x, y) is also a discrete analytic polynomial,
that (δxq)(x, 0) ≡ 0, and that deg(δxp) < deg(p). Hence (δxq)(x, y) ≡ 0. Since
q(x, y) is discrete analytic, Definition 4.1 implies that
∀(x, y) ∈ Z2,
(δyq)(x, y) =
i − 1
2
((1 − i + δy)δxq)(x, y) = 0.
Thus
and hence q = const - a contradiction.
(cid:3)
(δxq)(x, y) ≡ (δyq)(x, y) ≡ 0
Proof of Theorem 4.2. The uniqueness of the polynomial q(x, y) follows from Lemma
4.3. The existence in the case p = const is clear: it suffices to set
q(x, y) = p(0).
If p 6= const we proceed by induction on d = deg(p). According to Proposition
2.6, (δp)(x) is a polynomial and deg(δp) = d − 1. Therefore, by the induction
assumption, there is a discrete analytic polynomial f (x, y), such that
∀x ∈ Z,
f (x, 0) = (δp)(x)
and deg(f ) = d − 1. Let g : Z2 −→ C be defined by
g(x, y) = if (x, y) −
1 − i
2
(δyf )(x, y),
then g(x, y) is also a discrete analytic polynomial, deg(g) = d − 1. Furthermore,
since ( ¯Df )(x, y) ≡ 0,
∀(x, y) ∈ Z2,
(δxg)(x, y) = i(δxf )(x, y) −
1 − i
2
(δxδyf )(x, y)
= i(δxf )(x, y) +
1 − i
2
((1 − i)(δxf )(x, y) + (1 + i)(δyf )(x, y)) = (δyf )(x, y).
Hence, according to Proposition 3.7, there exists a polynomial h : Z2 −→ C such
that
(δxh)(x, y) ≡ f (x, y),
(δyh)(x, y) ≡ g(x, y).
Since
∀(x, y) ∈ Z2,
the polynomial h(x, y) is discrete analytic. Finally, since
( ¯Dh)(x, y) = (1 − i)f (x, y) + (1 + i)g(x, y) + (δyf )(x, y) = 0,
∀x ∈ Z,
(δxh)(x, 0) = f (x, 0) = (δp)(x),
h(x, 0) − p(x) is a constant function. Thus it suffices to set
q(x, y) = h(x, y) − h(0, 0) + p(0)
to complete the proof.
(cid:3)
12
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
5. Expandable discrete analytic functions
In view of Theorem 4.2, there exists a unique discrete analytic polynomial ζn(x, y)
determined by
Then (as follows from Proposition 3.6 and identities (2.7), (3.1)) (δxζn)(x, y) is also
a discrete analytic polynomial such that
ζn(x, 0) ≡ x[n].
Hence, by the uniqueness part of Theorem 4.2,
(δxζn)(x, 0) ≡ δx[n] ≡ nx[n−1].
(δxζn)(x, y) ≡ nζn−1(x, y)
(5.1)
(if n = 0, ζ0(x, y) ≡ 1 and (δxζ0)(x, y) ≡ 0).
Proposition 5.1. For each (x, y) ∈ Z2 the function
(5.2)
1 + i + z(cid:19)y
ex,y(z) = (1 + z)x(cid:18) 1 + i + iz
,
is analytic (in the usual sense) in the variable z in the open unit disk D, and admits
the Taylor expansion
Proof. The analyticity is clear because x, y ∈ Z and we have
∀z ∈ D.
zncn(x, y)
n!
,
znζn(x, y)
n!
,
∀z ∈ D.
ex,y(z) = Xn∈Z+
ex,y(z) = Xn∈Z+
(5.3)
where
Since
dn
cn(x, y) =
dzn ex,y(cid:12)(cid:12)(cid:12)z=0
ex+1,y(z) − ex,y(z) = zex,y(z) = Xn∈Z+
∀(x, y) ∈ Z2, ∀n ∈ Z+,
.
zn+1cn(x, y)
n!
,
(δxcn)(x, y) = ncn−1(x, y)
(if n = 0, c0(x, y) ≡ 1 and (δxc0)(x, y) ≡ 0). Similarly, since
(5.4)
ex+1,y+1(z) − ex,y(z)
ex+1,y(z) − ex,y+1(z)
=
1 + i
−
1 − i
=(cid:18)(1 + z)
1 + i + iz
1 + i + z − 1(cid:19) ex,y(z)
1 + i −(cid:18)1 + z −
1 + i + iz
1 + i + z(cid:19) ex,y(z)
1 − i
= 0,
we have
∀(x, y) ∈ Z2, and ∀n ∈ Z+,
( ¯Dcn)(x, y) = 0.
Thus, for every n ∈ Z+, the function cn : Z2 −→ C is discrete analytic. Next, we
show by induction that
(5.5)
Indeed, for n = 0,
∀n ∈ Z+,
cn(x, y) ≡ ζn(x, y).
c0(x, y) ≡ 1 ≡ ζ0(x, y).
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
13
Assume that, for some n ∈ Z+,
then, in view of (5.1),
cn(x, y) ≡ ζn(x, y),
(δxcn+1(x, y) ≡ (n + 1)cn(x, y) ≡ (n + 1)ζn(x, y) ≡ (δxζn+1(x, y),
hence
But the functions cn+1 and ζn+1 are discrete analytic, hence
(δx(ζn+1 − cn+1))(x, y) ≡ 0.
and
It follows that
since
one concludes that
( ¯D(ζn+1 − cn+1))(x, y) ≡ 0
(δy(ζn+1 − cn+1))(x, y) ≡ 0.
ζn+1 − cn+1 = const;
ζn+1(0, 0) = 0 = cn+1(0, 0),
cn+1(x, y) ≡ ζn+1(x, y),
and (5.5) follows.
Corollary 5.2. Let x, n ∈ Z+. Then,
(5.6)
ζn(x, 0) =(x(x − 1)··· (x − n + 1) = x[n], if n ≤ x
if n > x.
0,
Proof. Set y = 0 in ex,y(z) in (5.2). By (5.5) we get
(5.7)
(1 + z)x = Xn∈Z+
zn
n!
ζn(x, 0).
(cid:3)
(cid:3)
(5.6) follows by comparing the coefficients of zn in (5.7).
Theorem 5.3. It holds that
∀(x, y) ∈ Z+ × (Z \ {0}),
lim sup
n→∞ (cid:18)ζn(x, y)
n!
(cid:19)1/n
=
1
√2
.
Proof. When x ≥ 0 the function (5.2) is analytic in the variable z in the disk
centered at the origin and of radius √2, and has a pole on the boundary of this
disk. Hence the radius of convergence of the McLaurin series is precisely √2, and
lim sup
n→∞ (cid:18)ζn(x, y)
n!
(cid:19)1/n
=
1
√2
.
(cid:3)
Theorem 5.4. Let g : Z+ −→ C be such that for every (x, y) ∈ Z+ × Z the series
(5.8)
g(n)ζn(x, y)
converges absolutely. Then the function f : Z+ × Z −→ C, defined by (5.8), is
discrete analytic, and it holds that
f (x, y) = Xn∈Z+
g(n) ≡ f0(n),
14
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
where the function f0 : Z+ −→ C is given by
f0(x) = f (x, 0),
(5.9)
x ∈ Z+.
Proof. The discrete analyticity of f follows directly from the discrete analyticity of
the polynomials ζn. Furthermore, when y = 0 the formula (5.8) becomes
f0(x) = Xn∈Z+
g(n)x[n],
hence, according to Proposition 2.3, g = f0.
(cid:3)
Now we can introduce the main class of functions to be considered in this paper.
Definition 5.5. A function f : Z+ × Z −→ C is said to be expandable if:
(1) the Fourier transform f0 of the function f0 : Z+ −→ C, given by (5.9),
satisfies the estimate
(5.10)
lim sup
n→∞
( f0(n)n!)1/n < √2;
(2) the function f admits the representation
f (x, y) = Xn∈Z+
f0(n)ζn(x, y),
(x, y) ∈ Z+ × Z.
The class of expandable functions contains all discrete analytic polynomials, and
elements of this class are determined by their values on the positive horizontal axis.
Corollary 5.6. Suppose that f0 : Z+ −→ C is rational. Then there exists a unique
expandable function f : Z+ × Z −→ C such that
f (x, 0) ≡ f0(x).
Proof. This is a consequence of Theorem 2.9.
(cid:3)
6. The Cauchy-Kovalevskaya product
Theorem 5.4 and Corollary 5.6 allows us to define a (partially) defined product on
expandable functions, which is everywhere defined on rational functions. This prod-
uct will be denoted by ⊙ and called, for reasons to be explain later in the section,
the Cauchy-Kovalesvskaya product. Consider f1 and f2 two expandable functions,
and assume that the Fourier transform of the pointwise product f1(x, 0)f2(x, 0)
satisfy (5.10). Then there exists a unique discrete analytic expandable function g
such that
g(x, 0) = f1(x, 0)f2(x, 0)
(6.1)
g is called the Cauchy-Kovalesvskaya product of f1 and f2 and is denoted by f1⊙f2.
Note that the Cauchy-Kovalesvskaya product of two rational functions always exist.
We now give a more formal definition of the product:
Definition 6.1. Let f : Z2 −→ C be a polynomial, such that
f (x, 0) ≡ c0 + c1x + ··· + cnxn,
and let g : Ω × Z −→ C be given. The Cauchy-Kovalevskaya (C-K) product of f
and g is defined by
(6.2) (g ⊙ f )(x, y) = (f ⊙ g)(x, y) = c0g(x, y) + c1(Zg)(x, y) +··· + cn(Z ng)(x, y).
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
15
We shall abbreviate this as
(6.3)
f ⊙ g = f (Z)g.
The commutativity asserted in the definition is proved in the following theorem.
Theorem 6.2.
(1) The restriction of the C-K product to Z+ is the product of the restriction.
(2) The C-K product is the unique discrete analytic extension corresponding to the
product of the restrictions.
(3) The C-K product is commutative, i.e.
(6.4)
g ⊙ f = f ⊙ g
for all choices of discrete analytic polynomials
Proof.
(1) Setting y = 0 in (6.2) and taking into account the definition of Z we have
(g ⊙ f )(x, 0) = c0g(x, 0) + c1(Zg)(x, 0) + ··· + cn(Z ng)(x, 0)
(6.5)
= c0g(x, 0) + c1xg(x, 0) + ··· + cnxng(x, 0)
= f (x, 0)g(x, 0).
(2) For an expandable function the discrete Cauchy-Riemann equation Df = 0
with prescribed initial values on the horizontal positive axis has a unique solution
(see Theorem 5.4).
(3) is then clear from (1) and (2).
(cid:3)
We now explain the name given to this product. Recall that the classical Cauchy-
Kovalevskaya theorem concerns uniqueness of solutions of certain partial differential
equations with given initial conditions. See for instance [20]. In Clifford analysis,
where the pointwise product of hyperholomorphic functions need not be hyper-
holomorphic, this theorem was used by F. Sommen in [26] (see also [9]) to define
the product of hyperholomorphic quaternionic-valued functions in R4 by extending
the pointwise product from an hyperplane. In the present setting of expandable
functions the discrete Cauchy-Riemann equation Df = 0 with prescribed initial
values on the horizontal positive axis also has a unique solution. This is why the
pointwise product on the horizontal positive axis can be extended to a unique ex-
pandable function, which we call the C-K product.
If p is a discrete analytic polynomial and f is an expandable function, p ⊙ f is the
expandable function determined by
(6.6)
(p ⊙ f )(x, 0) = p(x, 0)f (x, 0).
However it is not true in general that the pointwise product of the restrictions of
two expandable functions, say f and g, is itself the restriction of an expandable
function, as is illustrated by the example
(6.7)
where ex,y(t) is defined by (5.2) and t >. Indeed,
ex,0(t) = (1 + t)x
(6.8)
f (x, y) = g(x, y) = ex,y(t),
16
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
(ex,0(t))2 = ex,0(2t + t2) will not be the restriction of an expandable function for
is the restriction of an expandable function whenever t < √2. On the other hand,
t >p1 + √2 − 1 which is strictly smaller than √2.
We note that the C-K product for hyperholomorphic functions was used in [3,
4, 5] to define and study rational hyperholomorphic functions, and some related
reproducing Hilbert spaces.
Proposition 6.3. For m, n ∈ Z+ and j ∈ {0, . . . , m + n}, set
(6.9)
cm,n
j =
x=0.
Then,
(6.10)
j!
δj(cid:0)x[m]x[n](cid:1)
m+nXj=0
ζm ⊙ ζn =
cm,n
j
ζj.
Proof. It suffices to note that (6.10) holds for y = 0, thanks to Proposition 2.3. (cid:3)
Systems like (6.10) occur in the theory of discrete hypergroups. See [21].
7. Rational discrete analytic functions
As we already mentioned, the pointwise product of two discrete analytic functions
need not be discrete analytic. In the sequel of the section we define a product on
discrete analytic functions when one of the terms is a polynomial, and show that a
rational function is a quotient of discrete analytic polynomials with respect to this
product. We first need the counterpart of multiplication by the complex variable.
Definition 7.1. The multiplication operator Z on the class of functions f : Ω ×
Z −→ C is given by
(Zf )(x, y) = xf (x, y) + iy
f (x, y + 1) + f (x, y − 1)
.
2
Proposition 7.2. Let f be a function from Ω × Z into C. Then
(1)
(7.1)
(Zf )(x, 0) ≡ xf (x, 0).
Furthermore:
(2) If f is a polynomial, then so is Zf.
(3) If f is discrete analytic, then so is Zf .
Proof. The proofs of (1) and (2) are clear from the definition. The proof of (3)
follows from the identity (7.6) in Theorem 7.4.
(cid:3)
Proposition 7.3. Let f : Z+ × Z −→ C be expandable. Then so is Zf .
particular,
In
(7.2)
Proof. By Proposition 7.2, for every n ∈ Z+, (Zζn)(x, y) is a discrete analytic
polynomial. In view of (7.1),
(Zζn)(x, y) ≡ ζn+1(x, y) + nζn(x, y).
∀n ∈ Z+,
hence formula (7.2) follows from Theorem 4.2.
(Zζn)(x, 0) = x · x[n] = x[n+1] + nx[n],
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
17
Let f0 : Z+ −→ C be defined by (5.9), then
(Zf )(x, 0) = xf0(x) = Xn∈Z+
(n f0(n) + f0(n − 1))x[n],
where f0(−1) := 0. Since
lim sup
n→∞
( f0(n)n!)1/n < √2,
lim sup
n→∞
(n f0(n) + f0(n − 1)n!)1/n < √2.
f (x, y) = Xn∈Z+
f0(n)(Zζn)(x, y) = Xn∈Z+
f0(n)ζn(x, y),
(n f0(n) + f0(n − 1))ζn(x, y).
Finally, since
where the convergence is absolute,
(Zf )(x, y) = Xn∈Z+
From the preceeding proof we note that
(7.3)
ζ1 ⊙ ζn = Zζn = nζn + ζn+1.
(cid:3)
Theorem 7.4. The operators δx, δy,Z and D generate a Lie algebra of linear
operators on the space of all functions from Z2 into C. The Lie bracket is [A, B] =
AB − BA and the relations on the generators are
(7.4)
(7.5)
(7.6)
(7.7)
[δx,Z] = 1 + δx,
[δy,Z] = i(1 + δy + δ2
y),
[D,Z] =(cid:18) 1 + i
i
2
+
2
δy(cid:19) D,
[D, δx] = [D, δy] = [δx, δy] = 0.
Proof. The identities (7.4)-(7.7) can be verified by the calculations in the proofs of
the two preceding propositions.
(cid:3)
Definition 7.5. A function f : Z+ × Z −→ C is said to be a rational discrete
analytic function if f (x, y) is expandable and f (x, 0) is rational.
Theorem 7.6. An expandable function f : Z+ × Z −→ C is rational if and only if
it is a C-K quotient of discrete analytic polynomials function that is, if and only if
there exist discrete analytic polynomials p(x, y) and q(x, y) such that
and
∀x ∈ Z+,
q(x, 0) 6= 0
(q ⊙ f )(x, y) ≡ p(x, y).
Proof. Suppose first that f is rational, and let f0(x) denote the restriction of f to
the horizontal positive axis. By definition there exists two polynomials p0(x) and
q0(x) such that q0(x) 6= 0 (on Z+) and
(7.8)
q0(x)f0(x) = p0(x),
x ∈ Z+.
18
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
Let p(x, y) and q(x, y) denote the discrete analytic polynomials extending p0(x) and
q0(x) respectively. Then both p and q⊙ f are expandable functions, which coincide
on Z+, and therefore everywhere.
Conversely, let p(x, y) and q(x, y) be the discrete analytic polynomials such that
and
Setting y = 0 leads to
which ends the proof.
∀x ∈ Z+,
q(x, 0) 6= 0
(q ⊙ f )(x, y) ≡ p(x, y).
∀x ∈ Z+,
q(x, 0)f (x, 0) = p(x, 0),
(cid:3)
Theorem 7.7. Let p(x, y) and q(x, y) be the discrete analytic polynomials such
that
Then there is a unique expandable rational function f such that
∀x ∈ Z+,
q(x, 0) 6= 0
Proof. Denote
(q ⊙ f )(x, y) ≡ p(x, y).
According to Proposition 7.3, the function g : Z+ × Z −→ C is expandable, and
therefore can be written as
g(x, y) = (q ⊙ f )(x, y).
g(x, y) = Xn∈Z+
g0(n)ζn(x, y),
where g0 is the Fourier transform of its restriction
g0(x) = g(x, 0).
In particular, by Theorem 5.4, g is discrete analytic. In view of (7.1),
g0(x) = q(x, 0)f (x, 0) = p(x, 0),
hence, by Proposition 2.5, g0 has finite support and g is a discrete analytic polyno-
mial. In view of Theorem 4.2,
g(x, y) ≡ p(x, y).
(cid:3)
8. The C∗-algebra associated to expandable discrete analytic
functions
. We denote by HDA the reproducing kernel Hilbert space with reproducing kernel
(8.1)
ζn(x1, y1)ζn(x2, y2)∗
K((x1, y1), (x2, y2)) =
,
∞Xn=0
(n!)2
and let en := 1
n! ζn be the corresponding ONB in HDA. Then,
Theorem 8.1.
(8.2)
δxe1 = 0 and δxen = en−1, n > 1,
i.e., δx is a copy of the backwards shift.
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
19
Proof. Using Proposition 5.1, we get
zn
n! ζn(x, y) into (8.3), we get δxζ1 = 0
(cid:3)
(8.3)
δxex,y(z) = ex+1,y(z) − ex,y(z) = zex,y(z).
and δxζn = nζn−1 if n > 1. The result (8.2) follows.
Proposition 8.2. In HDA we have
Substituting the expression ex,y(z) =Pn∈Z+
δy = δx(cid:18)I −
2 (cid:19)n
∞Xn=0(cid:18) i − 1
i − 1
2
(8.4)
δx(cid:19)−1
δn+1
x
,
=
where the convergence of the above series is in the operator norm
Proof. Recall that the operator D was defined in (4.1). Since the elements of HDA
are discrete analytic we have D = 0 in HDA, that is,
(1 − i)δx + (1 + i)δy + δxδy = 0,
δy ((1 + i)I + δx) = (1 − i)δx.
(8.5)
and thus
(8.6)
(8.7)
Since δx is an isometry, and has in particular norm 1, we can solve equation (8.5)
and obtain (8.4). The power expansion converges in the operator norm since
k(cid:18) i − 1
2 (cid:19) δxk =
1
√2
< 1.
(cid:3)
Theorem 8.3. The C∗-algebra generated by δx, or equivalently by δx and δy is the
Toeplitz C∗-algebra.
Proof. This follows from the preceding proposition and from [11].
Indeed it is
known [11] that the Toeplitz C∗-algebra T is the unique C∗-algebra generated by
the shift. Since δ∗
x is a copy of the shift, and δy ∈ C∗(δ∗
x), the result follows.
(cid:3)
We now consider
1
2
(8.8)
(Z + Z ∗) ,
A = Re Z =
where Z is defined from Definition 7.1.
Theorem 8.4.
(i) The operator A is essentially self-adjoint on the linear span D of the functions
ζn, n ∈ Z+.
(ii) On D it holds that
(8.9)
[δx, A] =
1
(iii) There exists a strongly continuous one parameter semi-group αt
such that
: T −→ T
x(cid:1) .
2(cid:0)I + δx + δ2
(8.10)
where T denotes the Toeplitz C∗-algebra.
(eitA)b(e−itA) = αt(b),
∀t ∈ R, ∀b ∈ T ,
20
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
Proof. In (iii), we denote by eitA the unitary one-parameter group generated by
the self-adjoint operator A from (i). The matrix representation of A with respect
to the ONB en := 1
n! ζn is:
(8.11)
0
n − 2 n − 1
n − 1 n − 1
n
0
0
0
n
n
0
n + 1
n + 1 n + 1 n + 2
n + 2 n + 2
0
0
It is therefore a banded infinite matrix with terms going to infinity linearly with n.
It follows from [18] that A is essentially self-adjoint.
(ii) By definition of Z and δx
[A, δ∗
x](en) = (Aδ∗
xA)(en)
x − δ∗
(en + en+1 + en+2)
=
=
1
2
1
2
(I + δ∗
x + δ∗2
x )en,
(8.12)
and hence the result.
(iii) We have
(8.13)
and
(8.14)
eitAbe−itA =
(it)n
n!
∞Xn=0
(ad A)n (b),
(ad A)n+1 (b) = [A, (ad A)n b],
We verify (8.13) on monomials of δx and δ∗
for more details regarding limits.
x using (8.14) and induction. See [10, 24]
(cid:3)
The next corollary deals with a flow. For more information on this topic, see [23].
0 −→ T /K −→ T −→ C(T) −→ 0,
Corollary 8.5. The one-parameter group {αt} ⊂ Aut (T ) passes to a flow on the
circle group T = {z ∈ C ; z = 1}.
Proof. By [11], the Toeplitz algebra T has a represnetation as a short exact sequence
(8.15)
where K is a copy of the C∗-algebra of all compact operators on HDA. Hence,
(8.16)
Since the left hand-side of (8.10) leaves K invariant, it follows that
(8.17)
passes to the quotient T /K ≃ C(T).
T /K ≃ C(T).
: T −→ T
αt
(cid:3)
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
21
9. A reproducing kernel Hilbert space of entire functions
As we stated in the previous section, the C-K product has the disadvantage of not
being defined for all pairs of expandable functions. In Section 10 we introduce a
different product which turns the space of expandable functions into a ring, and
consider a related reproducing kernel Hilbert space. In preparation we introduce
in the present section a reproducing kernel Hilbert space of entire functions of a
complex variable within which the results of Section 10 can be set in a natural way.
To set these results in a wider setting, let us recall a few facts on Schur analysis,
that is, on the study of functions analytic and contractive in the open unit disk. If
s0 is such a function (in the sequel, we write s0 ∈ S ), the operator of multiplication
by s0 is a contraction from the Hardy space of the open unit disk H2 into itself.
The kernel
(9.1)
1 − s0(z)s0(w)∗
1 − zw∗
is then positive definite in the open unit D, and its associated reproducing kernel
Hilbert space H(s0) was first studied by de Branges and Rovnyak. Spaces H(s0)
and their various generalizations play an important role in linear system theory
and in operator theory. See for instance [1, 2, 7, 14] for more information. Here we
replace H2 by two spaces, a space of entire functions in the present section and a
space of discrete analytic functions in the next section.
Thus, let H be a Hilbert space, and let O denote the space of L(H)-valued functions
analytic at the origin, and consider the linear operator T on O defined by
(9.2)
T (znAn) =
zn
n!
An, An ∈ L(H).
Then TO is a space of L(H)-valued entire functions. The operator T induces a
product ♦ of elements in TO via
(T f )♦(T g) = T (f g).
Theorem 9.1. Let H be a Hilbert space and let A is a bounded operator from H
into itself. Then the L(H)-valued entire function
znAn
∞Xn=0
(IH − zA)−♦ = (
) = ezA
n!
satisfies
(9.3)
and it is the only function in TO with this property.
Proof. This comes from the power expansion and norm estimates.
(IH − zA)♦(IH − zA)−♦ = IH,
(cid:3)
Take now H = C and let H2 denote the Hardy space of the unit disk. Then T is
a positive contractive injection from H2 into itself. Denote by H the space T H2
equipped with the range norm:
∀f ∈ H2,
kT fkH = kfk2.
22
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
Then T : H2 −→ H is unitary, and H is a reproducing kernel Hilbert space of entire
functions with the reproducing kernel
KH(z, w) =
(zw∗)n
(n!)2 .
∞Xn=0
Proposition 9.2. H is the Hilbert space of entire functions such that
ZC f (z)2K0(2z)dA(z) < ∞,
K0(r) =
exp(−r cosh t)dt
1
πZR
where
(9.4)
and
(9.5)
is the modified Bessel function of the second kind of order 0.
Proof. This follows from the fact that the Mellin transform of the square of the
function Γ is the function K0(2√x). See for instance [12, p. 50] for the latter. (cid:3)
We note that H is contractively included in the Fock space since the reproducing
kernel of the latter is
KF (z, w) =
,
n!
znw∗n
∞Xn=0
(zw∗)n(cid:18) 1
∞Xn=0
1
(n!)2(cid:19)
n! −
KF (z, w) − KH(z, w) =
is positive definite in C. See for instance [6, Theorem I, p. 354], [25] for differences
of positive definite functions.
In view of Liouville's theorem, the only multipliers on H in the sense of the usual
pointwise product are constants. The class of multipliers in the sense of the ♦
product is more interesting.
Theorem 9.3. A function s ∈ O is a contractive ♦-multiplier on H if and only if
it is of the form
where S denotes the Schur class of functions analytic and contractive in the open
unit disk.
s = T s0,
s0 ∈ S ,
Proof. Assume first that s ∈ O is a contractive ♦-multiplier on H. Then s =
s♦1 ∈ H and hence s = T s0 for some s0 ∈ H2. Furthermore, let f ∈ H2. Since
s♦(T f ) = T (s0f ) ∈ H, s0f ∈ H2. Since
kfkH2 = kT fkH ≥ ks♦(T f )kH = kT (s0f )kH = ks0fkH2,
s0 ∈ S .
Conversely, if s = T s0 where s0 ∈ S , and f ∈ H2 then s♦(T f ) = T (s0f ) ∈ H and
ks♦(T f )kH = kT (s0f )kH = ks0fkH2 ≤ kfkH2 = kT fkH.
Thus s is a contractive ♦-multiplier on H.
(cid:3)
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
23
Let s0 ∈ S . The operator Ms0 of pointwise multiplication is a contraction from
s0 endowed with the range norm is
called the associated de Branges-Rovnyak space. We denote it by H(s0). Similarly
one can associate with s ∈ T S a reproducing kernel
H2 into itself. The operator rangepI − Ms0 M ∗
Ks(z, w) = ((I − MsM ∗
s )KH(·, w)) (z),
where Ms denotes the operator of ♦-multiplication by s on H. The corresponding
s ) with the range norm; it will
reproducing kernel Hilbert space is ran(pI − MsM ∗
be denoted by H(s).
Theorem 9.4. The mapping f 7→ T f is unitary from de Branges - Rovnyak space
H(s0) onto H(s).
Proof. Since
MsT = T Ms0,
pI − MsM ∗
s T = TqI − Ms0M ∗
s0 .
(cid:3)
The H(s) spaces can be characterized in terms of ∂-invariance, where ∂ is the
differentiation operator:
∂f = f ′.
Lemma 9.5. The operator ∂ is bounded on H; moreover,
(9.6)
∂T = T R0,
where R0 is the backward shift operator, and
∂∗∂ = IH − C∗C,
∂∂∗ = IH,
where Cf := f (0). Furthermore, the reproducing kernel of H is given by
KH(z, w) = Cez∂ew∗∂ ∗
C∗.
Proof. The claims follow from the definition of the operator T in (9.2). We prove
only (9.6). Let f ∈ H2 with power series expansion
anzn.
(9.7)
f (z) =
(R0f )(z) =
anzn−1,
∞Xn=0
∞Xn=1
∞Xn=1
dz ∞Xn=0
d
an
(n − 1)!
an
n!
= (∂T f )(z).
Then,
(9.8)
and therefore
(9.9)
(T R0f )(z) =
=
zn−1
zn!
(cid:3)
24
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
Theorem 9.6. A closed subspace H of H is ∂-invariant if and only if
where s0(z) is an inner function.
H = H ⊖ MT s0 H,
Proof. Let H be a closed subspace of H then H = T H0 where H0 is a closed sub-
space of H2. H is ∂-invariant if and only if H0 is R0-invariant, which is equivalent
to H2 ⊖ H0 being invariant under multiplication by z. By the Beurling-Lax theo-
rem, the last condition holds if and only if H2 ⊖ H0 = Ms0 H2, where s0 is an inner
function.
(cid:3)
Theorem 9.7. Let s ∈ T S . Then s admits the representation
Cet∂Bdt,
s(z) = D +Z z
(cid:18) ∂ B
C D(cid:19) :(cid:18)H(s)
0
C (cid:19) −→(cid:18)H(s)
C (cid:19)
where
is a coisometry given by
∂f = f ′,
B1 = s′,
Cf = f (0),
D1 = s(0).
Proof. Write s = T s0, where s0 ∈ S . Then
where
s0(z) = D0 + zC0(I − zR0)−1B0,
(cid:18)R0 B0
C0 D0(cid:19) :(cid:18)H(s0)
C (cid:19) −→(cid:18)H(s0)
C (cid:19)
is a coisometry given by
R0f = (f − f (0))/z,
B01 = R0s0,
C0f = f (0),
D01 = s0(0) = s(0).
Hence
zn+1
(n + 1)!
zn+1
(n + 1)!
C0Rn
0 B0
C0T −1(T R0T −1)nT B0
s(z) = D0 +
= D0 +
∞Xn=0
∞Xn=0
∞Xn=0
= D +Z z
= D +
0
zn+1
(n + 1)!
C∂nB
Cet∂Bdt.
(cid:3)
be a coisometry. Then the function
(cid:18)A B
C D(cid:19) :(cid:18)H
s(z) = D +Z z
0
C(cid:19) −→(cid:18)H
C(cid:19)
CetABdt
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
25
Theorem 9.8. Let H be a Hilbert space and let
is a contractive ♦-multiplier on H, and the corresponding reproducing kernel is
given by
Proof. Set
Ks(z, w) = CezAew∗A∗
C∗.
s0(z) = D + zC(I − zA)−1B,
then s0 ∈ S and s = T s0. Since
Ks0(z, w) = C(I − zA)−1(I − wA)−∗C∗,
the formula for Ks(z, w) follows.
(cid:3)
Theorem 9.9. A reproducing kernel Hilbert space H of functions in O is of the
form H = H(s) for some s ∈ T S if and only if
(1) H is ∂-invariant;
(2) for every f ∈ H
k∂fk2
H ≤ kfk2
H − f (0)2.
Proof. One direction follows immediately from Theorem 9.4. The proof of the other
direction is modelled after the proof of [2, Theorem 3.1.2, p. 85] and is done as
follows: let H be ∂-invariant; then for every f ∈ H
Cez∂f = f (z),
where Cf = f (0). Hence the reproducing kernel of H is given by
L(z, w) = Cez∂ew∗∂ ∗
C∗.
Since
there exists a coisometry
But the the function
∂∗∂ + C∗C ≤ I,
C(cid:19) −→(cid:18)H
C(cid:19) .
(cid:18) ∂ B
C D(cid:19) :(cid:18)H
s(z) = D +Z z
0
Cet∂Bdt
is a contractive ♦-multiplier and the associated kernel Ks(z, w) coincides with
L(z, w). Hence H = H(s).
(cid:3)
It is also of interest to consider ♦-rational matrix valued functions.
Theorem 9.10. Tne following are equivalent:
(1) A function f ∈ TO is ♦-rational in the sense that for some polynomial
p(z), not vanishing at the origin, p♦f is also a polynomial.
26
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
(2) f (z) is of the form
f (z) = D +Z z
0
CetABdt
with A, B, C, D - matrices of suitable dimensions;
(3) the columns of ∂f belong to a finite-dimensional ∂-invariant space.
Proof. It suffices to observe that a function f ∈ TO is ♦- rational if and only if it
is of the form f = T f0 wheref0 ∈ O is rational in the usual sense.
(cid:3)
10. A reproducing kernel Hilbert space of expandable discrete
analytic function
In parallel with the previous section, we introduce the product ⊡ of expandable
discrete analytic functions by
(10.1)
ζn ⊡ ζm =
m!n!ζm+n
(m + n)!
.
The advantage of this product versus the C-K one is that the space of expandable
discrete analytic functions forms a ring.
Consider the linear mapping V : zn 7→ ζn. Then V T maps, in particular, the space
of functions analytic in a neighborhood of the closed disk {z : z ≤ 1/√2} onto the
space of expandable functions. Then V H with the range norm is the reproducing
kernel Hilbert space HDA with the reproducing kernel (8.1)
Note that
K((x1, y1), (x2, y2)) =
ζn(x1, y1)ζn(x2, y2)∗
(n!)2
.
∞Xn=0
V ∂ = δxV,
V (ezA) = ex,y(A).
Since V : H −→ HDA is unitary, the following theorems are direct consequences of
Theorems 9.3-9.10 in the previous section. We state them here in order to emphasize
the new product.
Theorem 10.1. A closed subspace H of HDA is δx-invariant if and only if
H = HDA ⊖ MV T s0HDA,
where s0(z) is an inner function.
Theorem 10.2. Let s ∈ V T S . Then s admits the representation
s(x, y) = D + Cex,y(δx) ⊡ (ζ1(x, y)B),
where
is a coisometry given by
(cid:18)δx B
C D(cid:19) :(cid:18)H(s)
C (cid:19) −→(cid:18)H(s)
C (cid:19)
B1 = δxs,
Cf = f (0),
D1 = s(0).
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
27
Theorem 10.3. Let H be a Hilbert space and let
(cid:18)A B
C D(cid:19) :(cid:18)H
C(cid:19) −→(cid:18)H
C(cid:19)
be a coisometry. Then the function
s(z) = D + Cex,y(A) ⊡ (ζ1(x, y)B),
is a contractive ⊡-multiplier on HDA, and the corresponding reproducing kernel is
given by
Ks((x1, y1), (x2, y2)) = Cex1,y1(A)(ex2,y2(A))∗C∗.
Theorem 10.4. A reproducing kernel Hilbert space H of expandable functions is
of the form H = H(s) for some s ∈ V T S if and only if
(1) H is δx-invariant;
(2) for every f ∈ H
kδxfk2
H ≤ kfk2
H − f (0, 0)2.
Theorem 10.5. The following are equivalent:
(1) An expandable function f is ⊡-rational in the sense that for some discrete
analytic polynomial p(x, y), not vanishing at the origin, p ⊡ f is also a
discrete analytic polynomial.
(2) f (x, y) is of the form
f (x, y) = D + Cex,y(A) ⊡ (ζ1(x, y)B),
with A, B, C, D - matrices of suitable dimensions, and kAk < √2, and
ex,y(A) is as in (5.2).
(3) the columns of δxf belong to a finite-dimensional δx-invariant space of ex-
pandable functions.
Theorem 10.6. Let s ∈ T S . Then s admits the representation
s(x, y) = D + Cex,y(δx) ⊡ (ζ1(x, y)B),
where
is a coisometry given by
(cid:18)δx B
C D(cid:19) :(cid:18)H(s)
C (cid:19) −→(cid:18)H(s)
C (cid:19)
B1 = s′,
Cf = f (0),
D1 = s(0).
Theorem 10.7. The following are equivalent:
(1) An expandable function f is ⊡-rational in the sense that for some discrete
analytic polynomial p(x, y), not vanishing at the origin, p ⊡ f is also a
discrete analytic polynomial.
(2) f (x, y) is of the form
f (x, y) = D + Cex,y(A) ⊡ (ζ1(x, y)B),
with A, B, C, D - matrices of suitable dimensions, and kAk < √2, and
ex,y(A) is as in (5.2).
(3) the columns of δxf belong to a finite-dimensional δx-invariant space.
28
D. ALPAY, P. JORGENSEN, R. SEAGER, AND D. VOLOK
References
[1] D. Alpay. The Schur algorithm, reproducing kernel spaces and system theory. American
Mathematical Society, Providence, RI, 2001. Translated from the 1998 French original by
Stephen S. Wilson, Panoramas et Synth`eses.
[2] D. Alpay, A. Dijksma, J. Rovnyak, and H. de Snoo. Schur functions, operator colligations,
and reproducing kernel Pontryagin spaces, volume 96 of Operator theory: Advances and
Applications. Birkhauser Verlag, Basel, 1997.
[3] D. Alpay, M. Shapiro, and D. Volok. Espaces de de Branges Rovnyak: le cas hyper -- analytique.
Comptes Rendus Math´ematiques, 338:437 -- 442, 2004.
[4] D. Alpay, M. Shapiro, and D. Volok. Rational hyperholomorphic functions in R4. J. Funct.
Anal., 221(1):122 -- 149, 2005.
[5] D. Alpay, M. Shapiro, and D. Volok. Reproducing kernel spaces of series of Fueter polyno-
mials. In Operator theory in Krein spaces and nonlinear eigenvalue problems, volume 162 of
Oper. Theory Adv. Appl., pages 19 -- 45. Birkhauser, Basel, 2006.
[6] N. Aronszajn. Theory of reproducing kernels. Trans. Amer. Math. Soc., 68:337 -- 404, 1950.
[7] M. Bakonyi and T. Constantinescu. Schur's algorithm and several applications, volume 261
of Pitman Research Notes in Mathematics Series. Longman Scientific & Technical, Harlow,
1992.
[8] H. Bart, I. Gohberg, and M.A. Kaashoek. Minimal factorization of matrix and operator
functions, volume 1 of Operator Theory: Advances and Applications. Birkhauser Verlag,
Basel, 1979.
[9] F. Brackx, R. Delanghe, and F. Sommen. Clifford analysis, volume 76. Pitman research notes,
1982.
[10] Ola Bratteli and Palle E. T. Jørgensen. Unbounded derivations tangential to compact groups
of automorphisms. J. Funct. Anal., 48(1):107 -- 133, 1982.
[11] L. A. Coburn. The C ∗-algebra generated by an isometry. Bull. Amer. Math. Soc., 73:722 -- 726,
1967.
[12] Serge Colombo. Les transformations de Mellin et de Hankel: Applications `a la physique
math´ematique. Monographies du Centre d'´Etudes Math´ematiques en vue des Applications:
B. -- M´ethodes de Calcul. Centre National de la Recherche Scientifique, Paris, 1959.
[13] R. J. Duffin. Basic properties of discrete analytic functions. Duke Math. J., 23:335 -- 363, 1956.
[14] H. Dym. J -- contractive matrix functions, reproducing kernel Hilbert spaces and interpolation.
Published for the Conference Board of the Mathematical Sciences, Washington, DC, 1989.
[15] Jacqueline Ferrand. Fonctions pr´eharmoniques et fonctions pr´eholomorphes. Bull. Sci. Math.
(2), 68:152 -- 180, 1944.
[16] C. J. Harman. A note on a discrete analytic function. Bull. Austral. Math. Soc., 10:123 -- 134,
1974.
[17] Rufus Isaacs. Monodiffric functions. Construction and applications of conformal maps. In
Proceedings of a symposium, National Bureau of Standards, Appl. Math. Ser., No. 18, pages
257 -- 266, Washington, D. C., 1952. U. S. Government Printing Office.
[18] Palle E. T. Jørgensen. Essential self-adjointness of semibounded operators. Math. Ann.,
237(2):187 -- 192, 1978.
[19] R. E. Kalman, P. L. Falb, and M. A. Arbib. Topics in mathematical system theory. McGraw-
Hill Book Co., New York, 1969.
[20] S. G. Krantz and H. P. Parks. A primer of real analytic functions, volume 4 of Basler
Lehrbucher [Basel Textbooks]. Birkhauser Verlag, Basel, 1992.
[21] Rupert Lasser and Eva Perreiter. Homomorphisms of l1-algebras on signed polynomial hy-
pergroups. Banach J. Math. Anal., 4(2):1 -- 10, 2010.
[22] Christian Mercat. Discrete Riemann surfaces and the Ising model. Comm. Math. Phys.,
218(1):177 -- 216, 2001.
[23] Edward Nelson. Topics in dynamics. I: Flows. Mathematical Notes. Princeton University
Press, Princeton, N.J., 1969.
[24] Gert K. Pedersen. C ∗-algebras and their automorphism groups, volume 14 of London Math-
ematical Society Monographs. Academic Press Inc. [Harcourt Brace Jovanovich Publishers],
London, 1979.
[25] S. Saitoh. Theory of reproducing kernels and its applications, volume 189. Longman scientific
and technical, 1988.
A PRODUCT FOR DISCRETE ANALYTIC FUNCTIONS
29
[26] F. Sommen. A product and an exponential function in hypercomplex function theory. Appli-
cable Anal., 12(1):13 -- 26, 1981.
(DA) Department of Mathematics
Ben Gurion University of the Negev
P.O.B. 653,
Be'er Sheva 84105,
ISRAEL
E-mail address: [email protected]
(PJ) Department of Mathematics
14 MLH
The University of Iowa, Iowa City,
IA 52242-1419 USA
E-mail address: [email protected]
(RS) and (DV) Mathematics Department
138 Cardwell Hall
Kansas State University,
Manhattan, KS 66506
E-mail address: [email protected]
|
1903.03912 | 1 | 1903 | 2019-03-10T03:36:27 | Multiplicative weak convergence in Banach $f$-algebras | [
"math.FA"
] | In this paper, we study multiplicative weak convergence in Banach f-algebra and multiplicative weak* convergence in its dual. | math.FA | math |
MULTIPLICATIVE WEAK CONVERGENCE IN
BANACH f -ALGEBRAS
ZHANGJUN WANG, ZILI CHEN, AND JINXI CHEN
Abstract. A net (xα) in an f -algebra E is called multiplicative
o−→ 0 for all u ∈ E+. A
order convergent to x ∈ E if xα − xu
net (xα) in a Banach f -algebra E is called multiplicative norm
convergent to x ∈ E if xα − xu → 0 for all u ∈ E+. In this paper,
we study this convergence in Banach f -algebra and its dual, A
net (xα) in a Banach f -algebra E is called multiplicative weak
convergent to x ∈ E if xα − xu
w−→ 0 for all u ∈ E+.
1. Introduction
A net (xα) in a Riesz space E is called order convergent to x ∈ E
if xα − x ≤ yβ for yβ ↓ 0 ∈ E+. A net (xα) in a Riesz space E is
o−→ 0 for
called unbounded order convergent to x ∈ E if xα − x ∧ u
all u ∈ E+[4,5]. A net (xα) in a Banach lattice E is called unbounded
kk
norm convergent to x ∈ E if xα − x ∧ u
−→ 0 for all u ∈ E+[6,7].
A net (xα) in a Banach lattice E is called unbounded absolute weak
convergent to x ∈ E if xα − x ∧ u
w−→ 0 for all u ∈ E+[8,9].
In [1],a vector lattice E under an associative multiplication is said
to be a Riesz algebra whenever the multiplication makes E an algebra
(with the usual properties).A Riesz algebra E is called commutative if
xy = yx for all x, y ∈ E. A Riesz algebra E is called f-algebra if E has
additionally property that x∧ y = 0 implies (xz) ∧ y = (zx) ∧ y = 0 and
xy ∈ E+ for every x, y ∈ E+.A vector lattice E is called Archimedean
whenever 1
n x ↓ 0 holds in E for each x ∈ E+.Every Archimedean
f -algebra is commutative.In this article, unless otherwise, all vector
lattices are assumed to be real and Archimedean, and so f -algebras
are commutative. An f -algebra E which is at the same time a Banach
Date: March 20, 2019.
2010 Mathematics Subject Classification. xxx,xxx.
Key words and phrases. Banach lattices, f -algebra, multiplicative weak conver-
gence, multiplicative order convergence, multiplicative weak convergence, multi-
plicative weak star convergence.
1
2
Z. WANG, Z. CHEN, AND J. CHEN
lattice is called a Banach f -algebra whenever kxyk ≤ kxkyk holds for
all x, y ∈ E.
A net (xα) in an f -algebra E is called multiplicative order convergent
o−→ 0 for all u ∈ E+ in[11]. A net (xα) in a
to x ∈ E if xα − xu
Banach f -algebra E is called multiplicative norm convergent to x ∈ E
if xα − xu → 0 for all u ∈ E+ in[12].
Definition 1.1. A net (xα) in a Banach f -algebra E is called multi-
w−→ 0 for all u ∈ E+
plicative weak convergent to x ∈ E if xα − xu
(xα
convergent to x
α) in a Banach f -algebra E
w∗
−→ 0 for all u
′ ∈ E if x
′u
α − x
′
′ is called multiplicative weak star
′
′ ∈ E
′
+ (x
α
mw∗
−−→ x
′).
mw−−→ x).
A net (x
′
mw−−→ x)
Remark 1.2. For a net (xα) in a Banach f -algebra E,(xα
mw−−→ xy)for all y ∈ E because of xαy − xyu = xα − xyu
implies (xαy
for all u ∈ E+. The converse holds true in Banach f -algebras with
mw−−→ xy)for each y ∈ E.
the multiplication unit. Indeed, assume(xαy
w−→ 0. Similarily, the mw∗-
Fix u ∈ E+, so, xα − xu = xαe − xeu
convergence has those properties.
Remark 1.3. In Banach f -algebras, the weak convergence does not
implies the mw-convergence, unless it has Schur property. The con-
verse holds true in Banach f -algebras with the multiplication unit.
w−→ 0, so
Assume(xαy
xα
mw−−→ xy)for each y ∈ E, then xα − x = xα − xe
w−→ x.
Remark 1.4. In order continuous Banach f -algebras, order conver-
gence and mo-convergence imply the mn-convergence.
2. Results
Lemma 2.1. Let (xα)and (yα) be two nets in a Banach f -algebra E.
Then the following holds true:
mw−−→ x iff xα − x
mw−−→ x then yβ
(1) xα
(2) if xα
(3) suppose xα
mw−−→ 0 iff xα − x mw−−→ 0;
mw−−→ x for each subnet (yβ) of (xα);
mw−−→ x and yβ
mw−−→ y, then axα + byβ
mw−−→ ax + by for
any a, b ∈ R;
mw−−→ x and xα
mw−−→ x then xα mw−−→ x.
(4) if xα
(5) if xα
The mw∗-convergence also has those properties.
mw−−→ y, then x = y;
Proof. We only need x − y ≤ x − xα + xα − y and (cid:12)
xα − x.
(cid:12)xα − x(cid:12)
(cid:12) ≤
(cid:3)
MW-CONVERGENCE
3
Lemma 2.2. For a Banach f -algebra E is order continuous, the mo-
convergence implies mw-convergence.
Proof. According to [12,remark 1.2], we have the conclusion.
(cid:3)
Lemma 2.3. Every disjoint decreasing sequence in a Banach f -algebra
mw-converges to zero.
Proof. Suppose (xn) is a disjoint and decreasing sequence in an Banach
f -algebra E. So, xnu is also a disjoint sequence in E for all u ∈ E+
by[2,definition 2.53 and 3,definition 3.1.8].Fix u ∈ E+, by[8.lemma 2],
w−→ 0 in E for all w ∈ E+. Thus,
we have xnu
in particular for fixed n0, taking w as xn0u. Then, for all n ≥ n0, we
get
uaw−−→ 0 in E. So, xnu∧w
xnu = xnu ∧ xn0u = xnu ∧ w
Since xnu ≤ xn0u, therefore, xn
mw−−→ 0 in E.
w−→ 0
(cid:3)
Lemma 2.4. Let E be an Banach f -algebra, B be a projection band
mw−−→ x in
of E and PB be the corresponding band projection. Then xα
E implies PB(xα) mw−−→ PB(x) in both E and B.
Proof. The proof is similarly to [11,proposition 2.7].
(cid:3)
Lemma 2.5. Let (xα) be a net in a Banach f -algebra E with order
continuous norm. Then we have that
mw−−→ x implies x ∈ E+;
(1) 0 ≤ xα
(2) if (xα) is monotone and xα
mw−−→ x then xα
w−→ x.
Proof. (1): Assume (xα) consists of non-zero elements and mw-converges
mw−−→ x+ ≥ 0.
to x ∈ E. Then, by Lemma 2.1(5), we have xα = x+
α
Therefore, we have x ∈ E+.
(2): For a fixed α ,we have xβ − xα ∈ E+ for β ≥ α. By (1), we
mw−−→ x − xα ∈ E+, so xα ↑ x, since E is order continuous,
w−→ x.
(cid:3)
have xβ − xα
therefore, xα
The lattice operations in Banach lattice f -algebras are mw-continuous
in the following sense.
Proposition 2.6. Let (xα)α∈A and (yβ)β∈B be two nets in a Banach
mw−−→
f -algebra E. If xα
x ∨ y.(∧ and are similarily)
mw−−→ y then (xα ∨ yβ)(α,β)∈A×B
mw−−→ x and yβ
Proof. Assume xα
mw−−→ x and yβ
mw−−→ y. Since we have
xα ∨ yβ − x ∨ yu ≤ xα ∨ yβ − xα ∨ yu + xα ∨ y − x ∨ yu
≤ yβ − yu + xα − xu
for every u ∈ E+.That is, (xα ∨ yβ)(α,β)∈A×B
w−→ 0
mw−−→ x ∨ y.
(cid:3)
4
Z. WANG, Z. CHEN, AND J. CHEN
The multiplication in f -algebra is mw-continuous in the following
sense.
Proposition 2.7. Let E be a Banach lattice f -algebra, and (xα)α∈A
mw−−→ y for some
and (yβ)β∈A be two nets in E. If xα
mw−−→ xy.
x, y ∈ E ,(xα) or (yβ) is monotone, then we have xαyβ
mw−−→ x and yβ
Proof. Assume xα
and yβ − xu
mw−−→ x and yβ
mw−−→ y, then we have xα − xu
w−→ 0
w−→ 0 for every u ∈ E+.Since
xαyβ − xyu = xαyβ − xαy + xαy − xyu
≤ xα − x + xyβ − yu + xα − xyu
≤ xα − xyβ − yu + yβ − yxu + xα − xyu
The second and the third terms are weak converges to zero, we show
first term is weak converges to zero. Assume (xα) is increasing, then
xα − x ≤ 2x and 2yβ − yxu is weak converges to zero, so we have
the conclusion.
(cid:3)
Theorem 2.8. Let E be a order continuous Banach f -algebra with a
mw−−→ 0 iff
multiplicative unit e and (xn) ↓ be a sequence in E. Then xn
xn(u ∧ n) w−→ 0 for all u ∈ E+.
Proof. For the forward implication, assume xn
for all u ∈ E+. Since xn(u∧e) ≤ xnu
mw−−→ 0, then xnu
w−→ 0
w−→ 0,therefore, xn(u∧n) w−→ 0.
For the reverse implication, by applying [2,theorem 2.57],we have
xnu ≤ xn(u − u ∧ ne) + xn(u ∧ ne) ≤
1
n
u2xn + nxn(u ∧ e)
Since (xn) ↓ and E is a order continuous Banach f -algebra, we have
the first term converges weakly to zero, and it is similarily to the proof
mw−−→
of [6,lemma 2.11], the second term weak convergent to zero, so xn
0.
(cid:3)
It is similarily to [12,proposition 2.4], we have the following result.
Theorem 2.9. Let 0 ≤ (xα)α∈A ↓ be a net in a Banach f -algebra E
with a quasi-interior point e. Then xα
mw−−→ 0 iff xαe
w−→ 0.
Corollary 2.10. Let 0 ≤ (xα)α∈A ↓ be a net in an order continuous
mw−−→ 0 iff
Banach f -algebra E with a weak order unit e. Then xα
xαe
w−→ 0.
Corollary 2.11. Let 0 ≤ (xα)α∈A ↓ be a net in a separable Banach
f -algebra E. Then xα
mw−−→ 0 iff xαe
w−→ 0.
MW-CONVERGENCE
5
Corollary 2.12. Let 0 ≤ (xα)α∈A ↓ be a net in a Banach f -algebra E
mw−−→ 0 iff xα(e + u) w−→ 0 for all
with a quasi-interior point e. Then xα
u ∈ E+.
Definition 2.13. A subset A of E is called a f -weak-almost order
bounded if for any ǫ > 0, there is u ∈ E+, f ∈ E
+ such that f (x −
ux) ≤ ǫ.
′
Next, we have the following work, it is similarily to [10,proposition
2.8].
Theorem 2.14. Let E be a Banach f -algebra. If (xα) is f -weak-almost
order bounded and mw-converges to x, then xα
′
σ(E,E
−−−−−→ x.
)
Proof. If (xα) is f -weak-almost order bounded net. Then the net (xα −
x) is also. For any ǫ > 0, there exists u > 0, f ∈ E
+ such that
′
f (xα − x − uxα − x) ≤ ǫ.
mw−−→ x, we have xα −xu
w−→ 0. Therefore, we have xα
Since xα
x.
′
σ(E,E
)
−−−−−→
(cid:3)
Definition 2.15. Let (xα) be a net in a Banach lattice f -algebra E.
Then
(1) (xα) is said to be mw-Cauchy if the net (xα − xα′ )(α,α′ )∈A×A
mw-converges to 0.
(2) E is called (σ)-mw-complete if every mw-Cauchy net(sequence)
in E is mw-convergent.
(3) E is called mw-continuous if xα
o−→ 0 implies that xα
mw−−→ 0.
mw−−→ 0.
Lemma 2.16. A Banach f -algebra is mw-continuous iff xα ↓ 0 implies
xα
Proof. Let (xα) o−→ 0, we have there exists yβ ↓ 0 and xα ≤ yβ. Since
yβ ↓ 0, so yβ
(cid:3)
mw−−→ 0, we have xα
mw−−→ 0.
Theorem 2.17. Let E be an mw-complete Banach f -algebra. Then
the following statements are quivalent:
(1) E is mw-continuous;
(2) if 0 ≤ xα ↑≤ x holds in E then (xα) is an mw-Cauchy net;
(3) xα ↓ 0 implies xα
mw−−→ 0 in E.
Proof. (1) ⇒ (2): By the proof of [2,lemma 1.37], we have (yβ −xα)α,β ↓
mw−−→ 0, so (xα) is an
0; according to lemma 2.16, we have (yβ − xα)α,β
mw-Cauchy net.
6
Z. WANG, Z. CHEN, AND J. CHEN
(2) ⇒ (3): Fix arbitrary α0, we have 0 ≤ (xα0 − xα)(α≥α0) ↑≤ xα0.
So (xα0 − xα)(α≥α0) is mw-Cauchy net ,so (xα′ − xα) mw−−→ 0. Since E
mw−−→ x. Because of lemma 2.5, we have x = 0,
is mw-complete, so xα
therefore, xα
mw−−→ 0.
(3) ⇒ (1): By lemma 2.16.
(cid:3)
Corollary 2.18. (σ)-mw-complete also has those properties of the last
theorem.
Corollary 2.19. Every mw-continuous and mw-complete Banach f -
algebra is Dedekind complete.
Proof. Suppose E is mw-continuous and mw-complete and 0 ≤ xα ↑≤
mw−−→ x and by the
y in E. By theorem 2.17, (xα) is mw-Cauchy, so xα
proof of lemma 2.5, we have xα ↑ x, so E is Dedekind complete.
(cid:3)
It was observed in [7,8,10], we now turn our attention to a topology
on Banach f -algebras. The sets of the form
Vu,ǫ,f = {x ∈ E : f (xu) < ǫ},
where u ∈ E+, ǫ > 0, f ∈ E
+ form a base of zero neighborhoods for a
Hausdorff topology, and the convergence in this topology is exactly the
mw-convergen.
′
Similarily to [7,8,10], we these conclusions.
Lemma 2.20. Vu,ǫ,f is either contained in [−u, u] or contains a non-
trivial ideal.
Lemma 2.21. If Vu,ǫ,f is contained in [−u, u], then u is a strong unit.
Similarily to [10]
Theorem 2.22. Let E be a Banach lattice.E
when one of the following conditions is valid:
(1) mw-topology agrees with norm topology.
(2) mw-topology agrees with weak topology.
′ has a strong order unit
Problem 2.23. How to describe the compactness of mw-topology.
When the mw-topology is metrizability.
Acknowledgement. xxxxxx
References
[1] W. A. J. Luxemburg and A. C. Zaanen, Riesz Spaces. II, North-Holland,
Amsterdam, 1983.
[2] C. D. Aliprantis and O. Burkinshaw, Positive Operators, Springer, 2006.
MW-CONVERGENCE
7
[3] P. Meyer-Nieberg, Banach lattices, Universitext, Springer-Verlag, Berlin,
1991.
[4] N. Gao and F. Xanthos, Unbounded order convergence and application to
martingales without probability, J. Math. Anal. Appl. 415(2), 2014, 931-947.
[5] H. Li, Z. Chen ,Some loose ends on unbounded order convergence, Positivity.
22(1), 2018, 83-90.
[6] Y. Deng, M. O'Brien and V. G. Troitsky, Unbounded norm convergence in
Banach lattices, to appear in Positivity.21(3),2018,963-974.
[7] M. Kandi´c, M. A. A. Marabeh and V. G. Troitsky, Unbounded norm topology
in Banach lattices, J. Math. Anal. Appl. 451(1), 2017, 259-279.
[8] Zabeti, Omid, Unbounded absolute weak convergence in Banach lattices, Pos-
itivity 22.2 (2018): 501-505.
[9] H.Li and Z.Chen, Some results on unbounded absolute weak Dunford-Pettis
operators, arXiv(2019):1902.00232.
[10] Z.Wang ,Z.Chen and J.Chen , Unbounded absolute weak* convergence in dual
Banach lattices, arXiv(2019):1903.02168v2.
[11] A. Aydin, Multiplicative
order
convergence
in
f -algebras
,
arXiv(2019):1901.04043v2.
[12] A. Aydin, Multiplicative norm convergence in Banach lattice f -algebras ,
arXiv(2019):1902.10927.
School of Mathematics, Southwest Jiaotong University, Chengdu,
Sichuan, China, 610000.
E-mail address: [email protected]
School of Mathematics, Southwest Jiaotong University, Chengdu,
Sichuan, China, 610000.
E-mail address: [email protected]
School of Mathematics, Southwest Jiaotong University, Chengdu,
Sichuan, China, 610000.
E-mail address: [email protected]
|
1708.07669 | 2 | 1708 | 2018-01-05T10:58:11 | The distance between two limit $q$-Bernstein operators | [
"math.FA"
] | For $q\in(0,1),$ let $B_q$ denote the limit $q$-Bernstein operator. In this paper, the distance between $B_q$ and $B_r$ for distinct $q$ and $r$ in the operator norm on $C[0,1]$ is estimated, and it is proved that $1\leqslant \|B_q-B_r\|\leqslant 2,$ where both of the equalities can be attained. To elaborate more, the distance depends on whether or not $r$ and $q$ are rational powers of each other. For example, if $r^j\neq q^m$ for all $j,m\in \mathbb{N},$ then $\|B_q-B_r\|=2,$ and if $r=q^m, m\in \mathbb{N},$ then $\|B_q-B_r\|=2(m-1)/m.$ | math.FA | math |
The distance between two limit q-Bernstein
operators
Sofiya Ostrovska and Mehmet Turan
Atilim University, Department of Mathematics, Incek 06836, Ankara, Turkey
e-mail: [email protected], [email protected]
Abstract
For q ∈ (0, 1), let Bq denote the limit q-Bernstein operator. In this paper, the distance
between Bq and Br for distinct q and r in the operator norm on C[0, 1] is estimated, and it
is proved that 1 6 kBq − Brk 6 2, where both of the equalities can be attained. To elaborate
more, the distance depends on whether or not r and q are rational powers of each other.
For example, if rj 6= qm for all j, m ∈ N, then kBq − Brk = 2, and if r = qm, m ∈ N, then
kBq − Brk = 2(m − 1)/m.
Keywords: Limit q-Bernstein operator, Peano kernel, positive linear operators
2010 MSC: 47A30, 47B30, 41A36
1
Introduction and statement of results
The limit q-Bernstein operator Bq can be viewed as an analogue of the Sz´asz-Mirakyan operator
pertinent the Euler distribution, also known as the q-deformed Poisson distribution, see [5, Ch. 3,
Sec. 3.4] and [9]. The latter is used in the q-boson theory, which delivers a q-deformation of the
quantum harmonic oscillator formalism [3]. Going into details, the q-deformed Poisson distribution
defines the distribution of energy in a q-analogue of the coherent state [3, 7]. The q-analogue of
the boson operator calculus is recognized as an indispensable area within theoretical physics. It
1
brings out explicit expressions for the representations of the quantum group SUq(2), which plays an
important role in various problems such as exactly solvable lattice models of statistical mechanics,
integrable model field theories, conformal field theory, only to mention a few. For additional
information the reader is referred to [4]. Therefore, linear operators related to the q-deformed
Poisson distribution, in particular the limit q-Bernstein operator, are of significant interest for
applications.
The operator Bq also emerges as a limit for a sequence of q-Bernstein polynomials in the case
0 < q < 1. Over the past years, the limit q-Bernstein operator has been studied widely from
different perspectives. Its approximation, spectral, and functional-analytic properties, probabilistic
interpretation, the behavior of iterates, and the impact on the analytic characteristics of functions
have been examined. See, for example, [8, 12, 13]. The review of obtained results along with
extensive bibliography can be found in [10].
Let q > 0. For any a ∈ C, as given in [2, Ch. 10], we denote:
(a; q)0 := 1,
(a; q)k :=
k−1
(1 − aqs),
Ys=0
(a; q)∞ :=
∞
Ys=0
(1 − aqs).
Definition 1.1. [6] For each q ∈ (0, 1), and f ∈ C[0, 1], the limit q-Bernstein operator is defined
by f 7→ Bqf where
f (1 − qk)pk(q; x)
if x ∈ [0, 1)
∞
Bq(f ; x) =
Xk=0
f (1)
if x = 1
in which
pk(q; x) =
xk(x; q)∞
(q; q)k
.
As it can be readily seen that Bq(f ; x) is defined by the values of f on the set
Jq := {1 − qk : k ∈ N0}.
Further, Euler's identity [2, page 490, Corollary 10.2.2]
1
(x; q)∞
=
∞
Xk=0
xk
(q; q)k
,
x < 1,
q < 1,
2
(1)
(2)
(3)
implies that
∞
Xk=0
pk(q; x) =
1, x ∈ [0, 1)
0, x = 1.
(4)
Formulae (1) and (4) show that Bq is a positive linear operator on C[0, 1] with kBqk = 1.
Recently, the continuity of the operator Bq with respect to the parameter has been examined in [1]
where the outcome below has been presented.
Theorem 1.2. For every f ∈ C[0, 1], one has:
and the convergence is uniform on [0, 1].
lim
q→a
Bq(f ; x) = Ba(f ; x)
This demonstrates the continuity of Bq in strong operator norm. The aim of the current paper is
to investigate whether the continuity persists with respect to the topology produced by the uniform
operator norm. It turns out that in this topology, {Bq}q∈(0,1) forms a discrete set of operators where
each Bq is an isolated point so that kBq − Brk > 1 whenever q 6= r.
The reasoning of the present paper is based essentially on the next theorem, which in itself can
be of interest. The idea of the proof is attributed to a statement made by a Mathoverflow user
under the nickname 'fedja', see [14].
Theorem 1.3. Let q ∈ (0, 1) and m > 2 be an integer. Then, for all x ∈ [0, 1] and all k ∈ N0, the
following inequality holds:
pmk(q; x) 6 pk(qm; x).
(5)
Obviously, by the triangle inequality, kBq − Brk 6 2. In some cases, the equality is attained, as
claimed by the below-mentioned result.
Theorem 1.4. Let q, r ∈ [0, 1]. If Jq ∩ Jr = {0}, then kBq − Brk = 2.
Now, comes the case Jq ∩ Jr 6= {0}. This situation occurs when rj = qm for some positive
integers j and m and reveals:
3
Theorem 1.5. Let 0 < r < q < 1 and rj = qm where j and m are mutually prime positive integers.
Then kBq − Brk >
2(m−1)
m .
In the case when j = 1, i.e., Jqm = Jr ⊂ Jq, the exact value of kBq − Brk has been obtained.
Theorem 1.6. Let q ∈ (0, 1) and r = qm for some integer m > 2. Then kBq − Brk = 2(m−1)
m .
Corollary 1.7. For any q 6= r, one has 1 6 kBq − Brk 6 2. The inequalities are sharp in the sense
that both equalities are attained.
2 Some auxiliary results
In this section, some results which will later contribute in proving our theorems, are presented. To
begin with, we point out the following:
Observation 2.1. For any positive integer m, one has:
lim
x→1−
∞
Xk=0
pmk(q; x) =
1
m
.
(6)
Proof. Clearly,
pmk(q; x) = (x; q)∞
∞
Xk=0
xmk
(q; q)mk
∞
Xk=0
6 (x; q)∞
xmk
(q; q)k
=
(x; q)∞
(xm; q)∞
∞
Xk=0
where the identity (3) is used. Therefore,
∞
Xk=0
pmk(q; x) 6
(x; q)
(xm; q)∞
=
1 − x
1 − xm
(qx; q)∞
(qxm; q)∞ →
1
m
as x → 1−.
On the other hand,
pmk(q; x) = (x; q)∞
> (x; q)∞
∞
Xk=0
xmk
(q; q)mk
∞
Xk=0
(x; q)∞
=
1 − x
1 − xm
=
(q; q)∞(1 − xm)
∞
xmk
(q; q)∞
Xk=0
(qx; q)∞
(q; q)∞ →
1
m
as x → 1−.
Lemma 2.2. If, for every k ∈ N, inequality (5) holds for x ∈ [0, 1 − qmk+m/2], then it holds for all
x ∈ [0, 1].
4
Proof. Clearly, [0, 1] = {1}S∞
(2), inequality (5) can be expressed as
k=1[0, 1 − qmk+m/2]. For x = 1, the inequality (5) is obvious. Now, by
xmk(x; q)∞
(q; q)mk
6
xk(x; q)∞
(qm; qm)k
or, equivalently,
Clearly,
uk(x) := x(m−1)k
k−1
m−1
Yj=0
Yℓ=1
1
1 − qℓ+mj
6
∞
m−1
Yj=0
Yℓ=1
1
1 − qℓ+mjx
.
(7)
max
j∈N
uj(x) = uk(x)
on [1 − αk−1, 1 − αk] where αk can be found from the equation uk+1(x) = uk(x). Therefore, if for
every k ∈ N, (7) holds on [0, 1 − αk], then it holds on [1 − αk−1, 1 − αk], and, as a result, on [0, 1].
That is why, (7) is going to be proved on [0, 1 − qmk+m/2] ⊇ [0, 1 − αk] for every k ∈ N. To justify
this inclusion, one has to show that αk > qmk+m/2. Indeed,
m−1
(1 − αk)m−1 =
m−1
m−1
(1 − qmk+ℓ)
[(1 − qmk+ℓ)(1 − qmk+m−ℓ)]
Yℓ=1
=vuut
Yℓ=1
Yℓ=1 p1 − qmk+ℓ − qmk+m−ℓ + q2mk+m
Yℓ=1 q(1 − qmk+m/2)2 = (1 − qmk+m/2)m−1
m−1
=
6
by virtue of the arithmetic-geometric mean inequality. This completes the proof.
Lemma 2.3. Let ρ(t) = 1/(et + e−t − 2), t > 0. Then, for all s, t > 0, the following inequality is
valid:
ρ(s + t) 6 e−sρ(t).
(8)
Proof. Equivalently, one may prove that 1/ρ(s + t) > es/ρ(t), that is, e−t − 2 6 e−2s−t − 2e−s. If
s = 0, then both sides are equal, while for s > 0, the derivative of the right hand side with respect
to s is positive, which yields the statement.
5
we set, for m > 2,
For the sequel, a special quadrature formula to approximateR b
b − a
m
Qm(f ; a, b) :=
m−1
b − a
m − 1
Xj=1
f(cid:18)a +
j(cid:19) .
(9)
a f (t)dt is needed. More precisely,
It is not difficult to see that the quadrature formula gives the exact value of the integral for
polynomials of degree at most 1. Denote by Ra,b(f ) the error in this approximation, i.e.,
Ra,b(f ) =Z b
a
f (t)dt − Qm(f ; a, b).
Lemma 2.4. The error (10) is given by
Ra,b(f ) =Z b
a
Ka,b(t)f ′′(t)dt
where
Ka,b(t) =
and h1 = b−a
m .
1
1
2(t − a)2
m−1(cid:1)2
2(cid:0)t − a − mh1k
2(b − t)2
1
+ mh2
1k(m−k−1)
2(m−1)2
if t ∈ [a, a + h1]
if t ∈ [a + kh1, a + (k + 1)h1], 1 6 k 6 m − 2
if t ∈ [b − h1, b]
(10)
(11)
(12)
Proof. By Peano's Theorem (see, for example [11, Theorem 3.2.3, page 123]), the error is expressed
by (11) where Ka,b(t) = Ra,b((x − t)+) and
(x − t)+ =
x − t
0
if x > t,
if x < t.
Here, (x− t)+ is considered as a function of x. Plain calculation of Ra,b((x− t)+) using (10) results
in (12).
In what follows, given h > 0, denote by K(t) the h-periodic function on R such that K(t) =
K0,h(t) for t ∈ [0, h] where Ka,b(t) is given by (12). In other words, K(t + h) = K(t) for all t ∈ R
and
K(t) =
1
2
t2 −
hk
m − 1
t +
h2k(k + 1)
2m(m − 1)
where h1 = h/m.
for t ∈ [kh1, (k + 1)h1], 0 6 k 6 m − 1
(13)
6
Lemma 2.5. For all m > 2, the following inequality holds:
Proof. For j = 0, 1, 2, set
Z h
0
K(t)
dt
t2
> 8Z 3h
h
K(t)
dt
t2 .
Ij :=Z h
0
K(t)
(t + jh)2 dt.
The inequality (14) is equivalent to I0 > 8(I1 + I2). Now,
I0 =Z h
0
=
h
2m
= h −
which gives
m−1
kh1
Xk=1Z (k+1)h1
k(cid:19) +
1
+
K(t)
0
m−1
t2 dt =Z h1
Xk=1(cid:20) h
hk
2m −
m − 1
k ln(cid:18)1 +
Xk=1
K(t)
t2 dt +
ln(cid:18)1 +
k(cid:19)
m−1
h
1
m − 1
K(t)
t2 dt
h2k(k + 1)
2m(m − 1)(cid:18) 1
kh1 −
(14)
1
(k + 1)h1(cid:19)(cid:21)
(m − 1)I0
h
= m − 1 − m ln m + ln(m!).
On the other hand, for j > 1, using (13) and the substitution x = t + jh, one obtains
Ij =Z h
0
m−1
m−1
kh1
K(t)
K(t)
Xk=0Z (k+1)h1
(t + jh)2 dt =
2 − h(cid:18)j +
(cid:20) 1
m − 1(cid:19) ln(cid:18)1 +
Xk=0Z jh+(k+1)h1
Xk=0(cid:18)j +
h
2 − h
jh+kh1
(t + jh)2 dt.
+ h2(cid:18) j2
m − 1(cid:19) 1
jm + k(cid:19) + Sj
m−1
x
k
k
2
1
=
=
where
+
jk
m − 1
+
k(k + 1)
2m(m − 1)(cid:19) 1
x2(cid:21) dx
Sj =
=
=
=
m−1
2
h
h2
h1
Xk=0(cid:18)j2
2 "j2(cid:18) 1
j −
2 ( j
2 (cid:20) j
j + 1
j + 1
+
h
h
+
+
+
jk
m − 1
j + 1(cid:19) +
1
1
m − 1
m
m − 1 −
k(k + 1)
2m(m − 1)(cid:19)(cid:18) 1
jm + k −
1
m − 1
m−1
Xk=0
[2mjk + k(k + 1)](cid:18) 1
1
jm + k + 1(cid:19)
jm + k −
jm + k + 1(cid:19)(cid:21))
1
1
jm + k + 1(cid:19)#
m−1
Xk=0(cid:20)1 − jm(jm + 1)(cid:18) 1
jm(jm + 1)
m − 1
(cid:18) 1
jm −
7
jm + k −
jm + m(cid:19)(cid:21) =
1
h
2
.
Therefore,
or
Ij = h − h
m−1
Xk=0(cid:18)j +
k
m − 1(cid:19) ln(cid:18)1 +
1
jm + k(cid:19)
Ij
h
= 1 − j ln(cid:18)1 +
1
j(cid:19) −
m
m − 1
ln(jm + m) +
ln[(jm + m)!] − ln[(jm)!]
m − 1
.
With the help of Stirling's formula
√2πnn+1/2e−n+1/(12n+1) < n! < √2πnn+1/2e−n+1/(12n),
one gets
I1 + I2
h
m[ln(2m) + ln(3m)]
m[ln(2m) + ln(3m)]
+
ln[(3m)!] − ln(m!)]
m − 1
= 2 + ln 2 − 2 ln 3 −
6 2 + ln 2 − 2 ln 3 −
m − 1(cid:20)(cid:18)3m +
+
1
m − 1
m − 1
1
2(cid:19) ln(3m) −(cid:18)m +
1
2(cid:19) ln(m) − 2m +
1
36m −
1
12m + 1(cid:21)
and as a result
while
(m − 1)(I1 + I2)
h
6 −2 +
5
2
ln 3 − ln 2 +
1
36m −
1
12m + 1
,
(m − 1)I0
h
> −1 + ln√2πm +
1
12m + 1
.
The needed inequality I0 > 8(I1 + I2) for all m > 2 follows from the fact that
−1 + ln√2πm +
1
12m + 1
> 8(cid:18)−2 +
5
2
ln 3 − ln 2 +
1
36m −
1
12m + 1(cid:19) ,
or equivalently,
To see this, let
ln√2πm +
9
12m + 1 −
2
9m
+ 15 − 20 ln 3 + 8 ln 2 > 0.
θ(x) = ln√2πx +
9
12x + 1 −
2
9x
+ 15 − 20 ln 3 + 8 ln 2.
8
Then,
θ′(x) =
1
2x −
108
(12x + 1)2 +
2
9x2 >
1
2x −
108
(12x)2 +
2
9x2 =
18x − 19
36x2 > 0 for x > 2.
Hence, for all m > 2, one has:
which completes the proof.
θ(m) > θ(2) ≈ 0.0073,
Remark 2.6. In the case m = 2, an alternative proof is presented in [14].
Corollary 2.7. Let ρ(t) be the function from Lemma 2.3. Then
Z h
0
K(t)ρ(t)dt > 8Z 3h
h
K(t)ρ(t)dt.
Proof. The statement is a consequence of the fact that t 7→ 1
for t > 0.
t2ρ(t) = et+e−t−2
t2
is an increasing function
Corollary 2.8. For h 6 h0 := ln 4,
Z h
0
K(t)ρ(t)dt > e3h/2Z 3h
h
K(t)ρ(t)dt.
Next, for a given f : [a, b] → R, denote by Ea,b the error in the composite quadrature formula
to approximate R b
a f (t)dt when the interval [a, b] is divided into n subintervals of equal length h
and the rule (9) is applied on each subinterval. That is,
Xj=1
where h = (b − a)/n. If b = ∞, we take n = ∞.
f (t)dt −
Ea,b =Z b
n
a
Qm(f ; a + (j − 1)h, a + jh),
(15)
Lemma 2.9. Let f (t) = − ln(1 − e−t), t > 0. Then, for all a > 0 and any step size, one obtains:
Es+a,∞ 6 e−sEa,∞.
Proof. By Peano's Theorem on the integral representation of the error term,
Es,∞ =Z ∞
s
f ′′(t)K(t − s)dt,
9
where K(t) is defined by (13). Since f ′′(t) = ρ(t), from Lemma 2.3, application of (8) yields:
Es+a,∞ =Z ∞
s+a
ρ(t)K(t − s − a)dt =Z ∞
a
ρ(t + s)K(t − a)dt 6 e−sZ ∞
a
ρ(t)K(t − a)dt = e−sEa,∞.
Lemma 2.10. If f (t) = − ln(1 − e−t), t > 0, then for all S, T > 0, there holds:
Z ∞
S
f (t)dt > −ST +Z T
0
f (t)dt.
Proof. It can be observed geometrically since f (t) is a decreasing continuous function symmetric
about the line y = x.
3 Proofs of the main results
Proof of Theorem 1.3. As it was done in the proof of Lemma 2.2, inequality (5) is equivalent to
x(m−1)k
k−1
m−1
Yj=0
Yℓ=1
1
1 − qℓ+mj
6
∞
m−1
Yj=0
Yℓ=1
1
1 − qℓ+mjx
.
Taking the logarithm of both sides leads to
−(m − 1)k ln(cid:18) 1
x(cid:19) +
k−1
m−1
Xj=0
Xℓ=1
ln(cid:18)
1
1 − qℓ+mj(cid:19) 6
∞
m−1
Xj=0
Xℓ=1
ln(cid:18)
1
1 − qℓ+mjx(cid:19) .
Set h = ln(1/qm), S = ln(1/x), T = kh and f (t) = − ln(1 − e−t). Then the inequality becomes
−ST +
k−1
Xj=0
h
m − 1
m−1
Xℓ=1
f (jh +
h
m
ℓ) 6
∞
Xj=0
h
m − 1
m−1
Xℓ=1
f (S + jh +
h
m
ℓ)
which can be written as
−ST +
k−1
Xj=0
Qm(f ; jh, (j + 1)h) 6
∞
Xj=0
Qm(f ; S + jh, S + (j + 1)h).
The sums in the last inequality can be viewed as the composite quadrature formulas for the integrals
R T
0 f (t)dt and R ∞
S f (t)dt, respectively. Therefore, by (15), one has
−ST +Z T
0
f (t)dt − E0,T 6Z ∞
S
f (t)dt − ES,∞.
10
Using Lemma 2.10, if one can show that
E0,T > ES,∞
(16)
for h 6 h0, the proof will be complete for q > 1/2. Due to Lemma 2.2, we need only to deal with
the case x ∈ [0, 1 − qmk+m/2] or
e−S 6 1 − e−T −h/2.
(17)
By Lemma 2.9, one derives that ES,∞ 6 e−SE0,∞ and also ET,∞ 6 e−T +hEh,∞ As f ′′(t) = ρ(t),
using Corollary 2.8 along with Peano's Theorem implies that Eh,3h 6 e−3h/2E0,h whenever h 6 h0.
Thence,
whence
Eh,∞ = Eh,3h + E3h,∞ 6 e−3h/2E0,h + e−2hEh,∞ 6 e−3h/2E0,∞,
E0,T = E0,∞ − ET,∞ >(cid:0)1 − e−T −h/2(cid:1) E0,∞ > e−SE0,∞ > ES,∞
due to (17). This is the desired inequality (16). To finish the proof, we observe that for all q ∈ (0, 1
2),
x ∈ [0, 1] and i ∈ N, it is true that
1
6
x
1 − qi
1 − qix
and thence, for q ∈ (0, 1
2) and x ∈ [0, 1]
Yℓ=1
Yj=0
Yj=0
1 − qℓ+mj
m−1
k−1
k−1
6
x
m−1
Yℓ=1
1
1 − qℓ+mjx
6
∞
m−1
Yj=0
Yℓ=1
1
1 − qℓ+mjx
.
The proof is complete.
Proof of Theorem 1.4. For given ε > 0, one can find δ > 0 such that p0(q; x) < ε/4 for x ∈
[1 − 2δ, 1 − δ]. Because of (4), for N ∈ N, one may write:
Xk=N +1
pk(q; x) = 1 − p0(q; x) −
x ∈ [0, 1).
Xk=1
pk(q; x),
∞
N
Notice that the series in (4) converges uniformly on any closed subinterval of [0, 1), in particular,
on [0, 1 − δ]. Therefore, there exists N0 ∈ N, such that
∞
Xk=N +1
pk(q; x) <
ε
4
,
for all x ∈ [0, 1 − δ], N > N0.
11
Hence, on [1 − 2δ, 1 − δ], there holds:
pk(q; x) > 1 −
ε
2
.
N
Xk=1
(18)
Apply this procedure to find N1, N2 and δ satisfying both
pk(q; x) > 1 −
ε
2
and
N1
Xk=1
pk(r; x) > 1 −
ε
2
N2
Xk=1
for x ∈ [1 − 2δ, 1 − δ]. Setting N = max{N1, N2}, one obtains
N
Xk=1
[pk(q; x) + pk(r; x)] > 2 − ε
for x ∈ [1 − 2δ, 1 − δ].
At this point, for every N ∈ N, consider a function fN ∈ C[0, 1] such that kfNk = 1 and
when k = 1, 2, . . . , N,
fN (1 − qk) = 1
fN (1 − rk) = −1
when k = 1, 2, . . . , N,
fN (1 − qk) = fN (1 − rk) = 0 when k 6= 1, 2, . . . , N, .
Then,
Since,
(Bq − Br) f (x) =
N
Xk=1
[pk(q; x) + pk(r; x)] > 2 − ε
for x ∈ [1 − 2δ, 1 − δ].
kBq − Brk > k (Bq − Br) f (x)k > 2 − ε
and as ε > 0 has been chosen arbitrarily, one concludes that kBq − Brk > 2. On the other hand,
by the triangle inequality, one has kBq − Brk 6 2, and the statement follows.
∞
Xk=0
=Xmk
(Bq − Br) f (x) =
Proof of Theorem 1.5. Let f ∈ C[0, 1]. Then, for x ∈ [0, 1),
Xk=0
f (1 − qk)pk(q; x) −
f (1 − qk)pk(q; x) +Xm∤k
−Xjk
Xk=0
f (1 − qmk) [pmk(q; x) − pjk(r; x)]
f (1 − rk)pk(r; x) −Xj∤k
∞
=
12
∞
f (1 − rk)pk(r; x)
f (1 − qk)pk(q; x)
f (1 − rk)pk(r; x)
+Xm∤k
f (1 − qk)pk(q; x) −Xj∤k
f (1 − rk)pk(r; x)
For each N ∈ N, choose fN ∈ C[0, 1] with kfNk = 1 in such a way that
for j ∤ k, k 6 N,
fN (1 − rk) = −1
fN (1 − qk) = −1
fN (1 − qk) = 1
fN (1 − qk) = fN (1 − rk) = 0 for k > N.
for mk, k 6 N,
for m ∤ k, k 6 N,
Then,
(Bq − Br) fN (x) =
N
N
⌊N/m⌋
pk(q; x) +
pk(r; x) − 2
pmk(q; x).
Xk=0
Xk=0
Xk=0
⌊N/m⌋
Xk=0
Following the line of reasoning in the preceding proof and bearing in mind (6), one may opt for
δ > 0 and N ∈ N such that, for every ε > 0,
N
Xk=0
N
Xk=0
The statement is now immediate.
pk(q; x) +
pk(r; x) − 2
pmk(q; x) > 2 −
2
m − ε when x ∈ [1 − 2δ, 1 − δ].
∞
∞
(Bq − Br) f (x) =
Proof of Theorem 1.6. Let f ∈ C[0, 1]. Then
Xk=0
f (1 − qk)pk(q; x) −
f (1 − qk)pk(q; x) +Xm∤k
f (1 − qmk) [pmk(q; x) − pk(qm; x)] +Xm∤k
Xk=0
=Xmk
Xk=0
f (1 − rk)pk(r; x)
f (1 − qk)pk(q; x) −
=
∞
Using Theorem 1.3 and (4), the last inequality becomes
∞
f (1 − qmk)pk(qm; x)
Xk=0
f (1 − qk)pk(qm; x)
(Bq − Br) f (x) 6 2kfk 1 −
pmk(q; x)! ,
for x ∈ [0, 1).
∞
Xk=0
Now, by (6), for all ε > 0 there exists δ > 0 such that
pmk(q; x) >
1
m −
ε
2
,
for x ∈ [1 − 2δ, 1 − δ].
∞
Xk=0
Therefore,
(Bq − Br) f (x) 6 2kfk(cid:18)1 −
1
m
+
ε
2(cid:19) ,
for x ∈ [1 − 2δ, 1 − δ].
This implies that kBq − Brk 6 2− 2/m. Together with Theorem 1.5, this yields the statement.
13
Acknowledgments
During the work on this paper, the authors were lucky to see the discussion on Mathoverflow [14]
concerning inequality (5) in the case m = 2. We would like to express our sincere gratitude to
all MO users, who participated in this fruitful discussion, especially to the user whose nickname
is 'fedja' and whose grasp of the subject was of a significant inspiration. We are also pleased to
thank Prof. Alexandre Eremenko (Purdue University, USA) for his encouragement and valuable
help throughout the entire process of our work.
References
[1] M. M. Almesbahi, On Properties of q-Bernstein Polynomials, Master's Thesis, Atilim Uniersity,
2017.
[2] G. E. Andrews, R. Askey, R. Roy, Special Functions, Encyclopedia of Mathematics and Its
Applications, The University Press, Cambridge, 1999, 664 pp.
[3] L.C.Biedenharn, The quantum group SUq(2) and a q-analogue of the boson operators,
J.Phys.A: Math. Gen., 22, 1989, L873-L878.
[4] L. Castellani , J. Wess (eds), Quantum Groups and Their Applications in Physics, IOS Press,
1996, 652 pages
[5] Ch. A. Charalambides, Discrete q-Distributions, Wiley, 2016.
[6] II'inskii A, Ostrovska S. Convergence of generalized Bernstein polynomials. J. Approx. Theory
2002; 116(1):100-112.
[7] S. Jing, The q-deformed binomial distribution and its asymptotic behaviour, J. Phys. A: Math.
Gen., 27, 1994, 493-499.
[8] N. I. Mahmudov, Higher order limit q-Bernstein operators, Math. Methods Appl. Sciences,
(2011) 34(13), 1618-1626.
14
[9] S. Ostrovska, Positive linear operators generated by analytic functions, Proc. Indian Acad.
Sci. (Math. Sci.) Vol. 117, No 4, November 2007, pp. 485-493.
[10] S. Ostrovska, A Survey of Results on the Limit q-Bernstein Operator, Journal of Applied
Mathematics, Volume 2013 (2013), Article ID 159720, 7 pages.
[11] J. Stoer, R. Bulirsch, Introduction to numerical analysis, Springer-Verlag, New York, 1980.
[12] V.S. Videnskii, On some classes of q-parametric positive operators, Operator Theory, Advances
and Applications, Vol. 158, (2005), 213-222.
[13] H. Wang, Properties of convergence for the q-Meyer-Konig and Zeller operators, J. Math. Anal.
Appl., 335 (2), (2007), 1360-1373.
[14] https://mathoverflow.net/questions/269740/inequality-for-functions-on-0-1
15
|
1309.1908 | 1 | 1309 | 2013-09-07T21:18:53 | On Borel structures in the Banach space C(\beta\omega) | [
"math.FA"
] | M. Talagrand showed that, for the Cech-Stone compactification \beta\omega\ of the space of natural numbers, the norm and the weak topology generate different Borel structures in the Banach space C(\beta\omega). We prove that the Borel structures in C(\beta\omega) generated by the weak and the pointwise topology are also different.
We also show that in C(\omega*), where \omega*=\beta\omega - \omega, there is no countable family of pointwise Borel sets separating functions from C(\omega*). | math.FA | math | ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
Abstract. M. Talagrand showed that, for the Cech-Stone compactification
βω of the space of natural numbers ω, the norm and the weak topology generate
different Borel structures in the Banach space C(βω). We prove that the Borel
structures in C(βω) generated by the weak and the pointwise topology are also
different.
We also show that in C(ω∗), where ω∗ = βω \ ω, there is no countable family
of pointwise Borel sets separating functions from C(ω∗).
3
1
0
2
p
e
S
7
]
.
A
F
h
t
a
m
[
1
v
8
0
9
1
.
9
0
3
1
:
v
i
X
r
a
1. Introduction
Given a compact space K, by C(K) we denote the Banach space of continuous
real-valued functions K, equipped with the standard supremum norm. If K = βω,
the Cech-Stone compactification of the space ω of natural numbers, then C(βω)
is isometric to the classical Banach space l∞.
One can consider three natural topologies on C(K):
τp ⊆ weak ⊆ norm,
where τp is the topology of pointwise convergence. Consequently, one has three
corresponding Borel σ-algebras
Borel(C(K), τp) ⊆ Borel(C(K), weak) ⊆ Borel(C(K), norm).
Those three σ-algebras are equal for many classes of nonmetrizable spaces K,
this is the case for all spaces K such that the space C(K) admits the so called
Date: September 30, 2018.
2010 Mathematics Subject Classification. Primary 46B26, 46E15, 54C35, 54H05.
Key words and phrases. weak topology, pointwise topology, C(K), Borel structure.
Research of the first author was partially supported by the National Science Center research
grant DEC-2012/07/B/ST1/03363.
Research of the second author was partially supported by MNiSW Grant N N201 418939
(2010 -- 2013).
1
2
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
pointwise Kadec renorming, see [Ed2] and [Ra], we also refer the reader to the
paper [MP, Sec. 3] for some comments concerning coincidence of these σ-algebras.
On the other hand, Talagrand [Ta] proved that
Borel(C(βω), weak) 6= Borel(C(βω), norm).
Marciszewski and Pol [MP] showed that Borel(C(S), τp) 6= Borel(C(S), weak)
for S being the Stone space of the measure algebra. Since, for the space S, the
Banach spaces C(S) and C(βω) are isomorphic, it follows that C(S) has three
different Borel structures. Let us note that the Borel structures in function spaces
(C(S), τp) and (C(βω), τp) are essentially different, see Remark 4.6.
We show in the present paper that Borel(C(βω), τp) 6= Borel(C(βω), weak);
our result and Talagrand's theorem mentioned above imply that, even though βω
is separable, the space C(βω) possesses three different Borel structures as well.
Proving our main result, stated below as Theorem 3.5, we build on ideas from
[MP] and show that in fact there is a measure bµ ∈ C(βω)∗ which is not pointwise
Borel measurable.
Recall that, if ϕ : K → L is a continuous surjection, then the map f 7→
f ◦ ϕ defines an embedding of C(L) into C(K) with respect to the norm, weak,
and pointwise topologies. Since βω is a continuous image of ω∗, it follows from
Theorem 3.5 that for ω∗ = βω \ ω one also has
Borel(C(ω∗), τp) 6= Borel(C(ω∗), weak).
This result was obtained in [MP, remark 6] under some additional set-theoretic
assumption.
We show in Section 4 that no sequence of pointwise Borel sets separates points
of C(ω∗). Section 5 contains some remarks concerning σ-fields of Baire sets in
function spaces on βω and ω∗.
2. Measures on ω and βω
We shall consider only nonnegative, finite measures. We will use the well-
known fact that any finitely additive measure µ on (ω, P(ω)) corresponds to a
uniquely determined Radon measure bµ on βω such that µ(A) = bµ(A), for any
A ∈ P(ω), where A is the closure of A in βω, cf. [Fr].
ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
3
In the sequel, we consider only measures µ on ω vanishing on singletons; then
for the corresponding measures bµ on βω, we have bµ(ω) = 0, and we may as well
treat such measures bµ as being defined on ω∗.
The following auxiliary result can be found in [BJ, Theorem 2.2.4].
Proposition 2.1. If (Gn)n is a sequence of dense open subsets of 2ω then there
is a sequence (In)n of pairwise disjoint finite subsets of ω and a sequence of
functions ϕn : In → 2 such that x ∈ Tn Gn for every x ∈ 2ω for which the set
{n ∈ ω : xIn = ϕn} is infinite.
Proposition 2.2. No nonzero measure on ω, vanishing on singletons, is mea-
surable with respect to the σ-algebra of subsets of 2ω having the Baire property.
Proof. Suppose, towards a contradiction, that µ, treated as a function on 2ω, is
measurable with respect to the σ-algebra of subsets of 2ω having the Baire prop-
erty. Without loss of generality, we can assume that µ(ω) = 1. The inverse image
µ−1(S) of any Borel subset S of the unit interval [0, 1] is a tail-set with the Baire
property, hence, by 0 -- 1 Law (see [Ox]) is either meager or comeager. Observe that
there exist (necessarily unique) t ∈ [0, 1] such that µ−1(t) is comeager. Indeed,
if µ−1(1) is comeager, then we are done. Otherwise, we can define inductively a
sequence of integers kn ≤ 2n − 1, such that µ−1([kn/2n, (kn + 1)/2n)) is comeager
for n ∈ ω. Then the required t is a unique element of Tn∈ω[kn/2n, (kn + 1)/2n).
The map h : P(ω) → P(ω), defined by h(A) = ω \ A, is a homeomorphism of
P(ω) such that h(µ−1(t)) = µ−1(1 − t). Therefore t = 1 − t, and t = 1/2.
By Proposition 2.1 , we have functions ϕn : In → 2 defined on pairwise disjoint
finite sets In such that µ(A) = 1/2 whenever χA agrees with infinitely many ϕn's.
Let N1, N2, N3 be a partition of ω consisting of infinite sets and let
Bi =[{{k : ϕn(k) = 1} : n ∈ Ni},
i = 1, 2, 3. Then, for each i ≤ 3, µ(Bi) = 1/2 and the sets Bi are pairwise disjoint,
a contradiction.
(cid:3)
For any subset A of ω we write
d(A) = lim sup
n
A ∩ n
n
,
4
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
for the outer asymptotic density of a set A and
d(A) = lim
n
A ∩ n
n
,
whenever the set A has the asymptotic density, i.e. when the above limit exists.
Given a bounded sequence (xn)n∈ω and an ultrafilter ℘ ∈ ω∗, by lim℘ xn we
denote the ℘-limit of (xn). For any ultrafilter ℘ ∈ ω∗, we define the measure d℘
on ω by the formula
d℘(A) = lim
℘
A ∩ n
n
,
for A ⊆ ω, cf. [Fr] or [BFPR].
Lemma 2.3. For any ultrafilter ℘ ∈ ω∗ and any ε > 0, there exists a set A ∈ ℘
having asymptotic density and such that d(A) < ε.
Proof. Take n ≥ 1 such that 1/n < ε and put Ak = {ni + k : i ∈ ω} for k =
0, 1, . . . , n − 1. Then there exists k < n such that Ak ∈ ℘. Clearly, d(Ak) =
1/n.
(cid:3)
We also recall the following standard fact concerning the outer density. Here,
for A, B ⊆ ω, A ⊆∗ B denotes, as usual, that A \ B is finite, and we denote by
A∗ the set A \ A, where A is the closure of A in βω.
Lemma 2.4. Let ε > 0 and An ⊆ ω, n ∈ ω, be such that An ⊆ An+1 and
d(An) < ε for every n ∈ ω. Then there exists A ⊆ Sn∈ω An such that d(A) ≤ ε
and An ⊆∗ A for every n ∈ ω.
Proof. By the definition of d we have
∀n ∈ ω ∃kn ∈ ω ∀k ≥ kn
An ∩ k
k
< ε .
Without loss of generality, we can assume that the sequence (kn) is increasing.
We define
A = [n∈ω
An ∩ (kn+1 \ kn) ⊆ [n∈ω
An .
Since An ⊆ An+1, we have (An \ kn) ⊆ A, and therefore A∗
n ⊆ A∗ for every n ∈ ω.
For any k > k0, we have k ∈ kn+1 \ kn, for some n ∈ ω, and A ∩ k ⊆ An ∩ k,
therefore A ∩ k/k < ε, and consequently d(A) ≤ ε.
(cid:3)
ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
5
Obviously, for the sets An and A as in the above lemma, we haveSn∈ω A∗
n ⊆ A∗
and d℘(A) ≤ ε for any ultrafilter ℘ ∈ ω∗. Let us note, however, that this does
not necessarily mean that for every increasing sequence A0 ⊆ A1 ⊆ . . . ⊆ ω such
that d℘(An) < ε there is A almost containing every An and such that d℘(A) ≤ ε.
Measures on P (ω) with such an approximation property may fail to exist, see
[Me] for details.
Corollary 2.5. For any ultrafilter ℘ ∈ ω∗, the measure bd℘ vanishes on separable
subsets of ω∗.
Proof. Let X be a subset of ω∗ contained in the closure of a set {Fn : n ∈ ω} ⊆
ω∗. Fix ε > 0. For any n ∈ ω, by Lemma 2.3, we can pick Bn ∈ Fn with
d(Bn) < ε/2n+1. Then, for An =Sk≤n Bk, we have d(An) < ε, and we can apply
Lemma 2.4 for the sequence (An), obtaining the set A satisfying d(A) ≤ ε. For
any n ∈ ω, we have Bn ⊆ An ⊆∗ A, hence A ∈ Fn. Therefore the closure in ω∗
of the set {Fn : n ∈ ω} is contained in A∗, and bd℘(X) ≤ bd℘(A∗) ≤ ε. Since ε was
arbitrarily chosen, it follows that bd℘(X) = 0.
3. Borel(C(βω), weak) and Borel(C(βω), τp) are different
(cid:3)
Let ℘ be a fixed ultrafilter from ω∗ and let µ = d℘ be the measure on P (ω)
respect to the σ-algebra of weakly Borel subsets of C(βω). In this section we
defined in section 2; we write bµ for the corresponding Radon measure on βω.
Then bµ is a continuous functional on C(βω) so in particular bµ is measurable with
shall show that the measure bµ is not pointwise Borel measurable and in this way
conclude our main result. The approach presented below builds on the technique
developed by Burke and Pol [BPo] and Marciszewski and Pol [MP].
We need to fix several pieces of notation. For a set X, by [X]<ω we denote the
family of all finite subsets of X, and X <ω stands for the set of all finite sequences
of elements of X. Given sequences s, t ∈ X <ω, s a t denotes their concatenation.
For functions f and g, f ≺ g means that the domain dom(f ) of f is contained
in the domain of g and g dom(f ) = f . We also use this notation for sequences,
treating them as functions.
Writing 2 = {0, 1}, we denote by Cp(βω, 2) the space of all continuous functions
f : βω → 2 equipped with the pointwise topology.
In the sequel we consider some subsets of (P (ω))2 = P (ω) × P (ω); a typical
element of such a set is a pair c = (A, B), where A, B ⊆ ω. Given some ci ∈
6
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
(P (ω))2, we shall use the convention for elements that every ci can be written as
ci = (Ai, Bi).
Let C be a subset of (P (ω) × P (ω))ω of those sequences c = (c0, c1, . . .) for
which the following conditions are satisfied for every i:
3.1(1) Ai ⊆ Ai+1, Bi ⊆ Bi+1, Ai ∩ Bi = ∅;
3.1(2) d(Ai), d(Bi) < 1/6.
We moreover denote by S the set of all finite sequences from (P (ω) × P (ω))<ω
satisfying conditions 3.1.
Given f ∈ C(βω), i ∈ {0, 1}, and A ⊆ ω, we write f A ≃ i if the equality
f (x) = i holds for bµ-almost all x ∈ A∗.
We equip P (ω) × P (ω) with the discrete topology and C with the product
topology inherited from (P (ω) × P (ω))ω. Finally, we define a topological space
E that is crucial for our considerations as follows
E = {(f, c) ∈ Cp(βω, 2) × C : f An ≃ 0, f Bn ≃ 1 for every n};
here c = (c0, c1, c2, . . .) and ci = (Ai, Bi).
Let Z be the set of all pairs
z = (z(0), z(1)) ∈ [ω∗]<ω × [ω∗]<ω,
such that z(0)∩z(1) = ∅. For z, z′ ∈ Z we write z ⊏ z′ to denote that z(0) ⊆ z′(0)
and z(1) ⊆ z′(1).
Basic open neighborhoods in E are of the form N(σ, z, s), where σ ∈ 2<ω,
z ∈ Z, s ∈ S, and N(σ, z, s) is the set of all (f, c) ∈ E such that
(a) f (x) = i for every x ∈ z(i), i = 0, 1;
(b) σ ≺ f and s ≺ c.
singletons.
Note that every set of the form N(σ, z, s) is nonempty, since bµ vanishes on
Let us say that s ∈ S captures z ∈ Z if, writing s = t a (A, B), we have
z(0) ⊆ A∗ and z(1) ⊆ B∗.
Lemma 3.2. Every basic open set N(σ, z, s) in E contains a neighborhood N(σ, z, s′)
where s′ captures z.
Proof. Indeed, if (A, B) is the final pair in s then for any ε > 0, using Lemma 2.3
we can find sets C, D ⊆ ω of asymptotic density < ε and such that z(0) ⊆ C ∗,
ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
7
z(1) ⊆ D∗. Then we can put s′ = s a (A′, B′), where A′ = A ∪ C, B′ = B ∪ D
and ε is small enough.
(cid:3)
Lemma 3.3. Let N(σ, z, s) be a basic open set in E, where σ ∈ 2l. If G is a
dense open subset of N(σ, z, s) then, for every k ≥ l, there are m > k, z′ ∈ Z
with z ⊏ z′, s′ ∈ S with s ≺ s′, and a function
ϕ : I = {i : k ≤ i < m} → 2,
such that for every τ ∈ 2k−l
N(σ a τ a ϕ, z′, s′) ⊆ G.
Proof. Given τ0 ∈ 2k−l, we have N(σ a τ0, z, s) ∩ G 6= ∅ so for some interval
I1 = {i : k ≤ i < m1} and ϕ1 : I1 → 2 there are z1 ⊐ z and s1 ≻ s such that
N(σ a τ0 a ϕ1, z1, s1) ⊆ G.
Take another τ1 ∈ 2k−l. Apply the same argument for N(σ a τ1 a ϕ1, z1, s1). It is
clear that we arrive at the conclusion after examining all τ ∈ 2k−l.
(cid:3)
Lemma 3.4. Let (Gn)n∈ω be a decreasing sequence of open subsets of E such
that G0 6= ∅ and every Gn is dense in G0. Then there exist a sequence c =
((An, Bn))n∈ω ∈ C, sets A, B ⊆ ω, countable sets Z(0), Z(1) ⊆ ω∗, and a sequence
ϕn : In → 2 of functions defined on pairwise disjoint finite sets In ⊆ ω such that
(i) Sn∈ω A∗
n ⊆ A∗, Sn∈ω B∗
n ⊆ B∗, A ∩ B = ∅;
(ii) µ(A), µ(B) ≤ 1
6 ;
(iii) Z(0) ⊆ A∗, Z(1) ⊆ B∗;
(iv) for every f ∈ Cp(βω, 2) satisfying
-- f A ≃ 0, f B ≃ 1,
-- f Z(i) = i for i = 0, 1,
-- f I0 = ϕ0,
-- f In = ϕn for infinitely many n ≥ 1,
we have (f, c) ∈Tn∈ω Gn.
Proof. Fix a basic neighborhood N(σ0, z0, s0) ⊆ G0; by Lemma 3.2 we can assume
that s0 captures z0. Take k0 such that σ ∈ 2k0, set I0 = {0, . . . , k0 − 1} and
ϕ0 = σ0.
8
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
We shall define inductively natural numbers k0 < k1 < k2 < . . ., functions
ϕn : In = {i : kn−1 ≤ i < kn} → 2, pairs zn ∈ Z with z0 ⊏ z1 ⊏ . . ., and
sequences s0 ≺ s1 ≺ . . . in S such that for every n ≥ 0
-- sn captures zn;
-- for every τ ∈ 2kn−k0 we have N(ϕ0 a τ a ϕn+1, zn+1, sn+1) ⊆ Gn+1.
Having kn, ϕn, . . . defined, we make the inductive step using Lemma 3.3 for the
neighborhood N(ϕ0, zn, sn) with G = Gn+1 ∩ N(ϕ0, zn, sn), l = k0, and k = kn
and we use m, z′, and s′ given by this lemma to define kn+1, zn+1, and sn+1. We
complete our choice applying Lemma 3.2.
The sequence sn ∈ S defines the unique element c = ((An, Bn))n∈ω ∈ C; we
take A, B applying Lemma 2.4 to sequences (An)n and (Bn)n (see also the remark
note that Z(0) ⊆ A∗ and Z(1) ⊆ B∗.
Now, if f satisfies (iv) then (f, c) ∈ Gn, for infinitely many n, so (f, c) ∈
(cid:3)
following the proof of Lemma 2.4). We put Z(0) =Sn zn(0) and Z(1) =Sn zn(1);
Tn Gn.
Theorem 3.5. The measure bµ is not measurable with respect to the pointwise
Borel sets in C(βω). In particular,
Borel(C(βω), τp) 6= Borel(C(βω), weak).
Proof. Suppose otherwise; then
F0 = {f ∈ Cp(βω, 2) : Z f dbµ < 1/2},
is pointwise Borel in Cp(βω, 2). Let F1 = Cp(βω, 2) \ F0.
Let π : E → Cp(βω, 2) denote the projection onto the first axis.
It follows
that the sets π−1(Fi) are Borel in E, so both π−1(F0) and π−1(F1) have the Baire
property in E. Therefore, for some i ∈ {0, 1}, there is a decreasing sequence (Gn)n
of open sets in E, where G0 6= ∅, every Gn is dense in G0 and Tn Gn ⊆ π−1(Fi).
Take c ∈ C, A, B ⊆ ω, Z(0), Z(1), ϕn : In → 2 as in Lemma 3.4.
Let R be an uncountable almost disjoint family of infinite subsets of ω. For
R ∈ R let
IR = I0 ∪ [n∈R
In;
then the family {IR : R ∈ R} is almost disjoint too. Therefore there is R ∈ R
such that
ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
9
-- (Z(0) ∪ Z(1)) ∩ I ∗
-- µ(IR) = 0.
R = ∅, and
Set A′ = A \ IR, B′ = B \ IR. Take any function f ∈ C(βω, 2) such that f ≡ 0
on A′, f ≡ 1 on B′ and f is defined on IR so that f In = ϕn for n ∈ R ∪ {0}.
Then f ≡ 0 on Z(0) and f ≡ 1 on Z(1), f A ≃ 0, f B ≃ 1. It follows from
Lemma 3.4 that (f, c) ∈Tn Gn ⊆ π−1(Fi).
On the other hand, f can be freely defined on the set
D = ω \ (A′ ∪ B′ ∪ IR),
where µ(D) ≥ 2/3, soR f dbµ can take values less than 1/2 and greater than 1/2,
a contradiction.
(cid:3)
4. On C-sets in (C(ω∗), τp)
Let us recall that in a topological space X, the elements of the smallest σ-
algebra in X containing open sets and closed under the Souslin operation are
called C-sets, cf. [Ke]. The C-sets are open modulo meager sets and any preimage
of a C-set under a continuous map is a C-set.
Theorem 4.1. No countable family of C-sets separates the functions in the space
(C(ω∗), τp).
Remark 4.2. From the fact that βω is separable, it follows easily that (C(βω), τp)
contains a countable family of open sets separating functions.
The above properties of function spaces imply immediately the following
Corollary 4.3. There is no Borel-measurable injection ϕ : (C(ω∗), τp) → (C(βω), τp).
We keep here a part of the notation introduced in Section 3; in particular, we
will use the space C and the sets S, Z, defined in that section.
By Cp(ω∗, 2) we denote the subspace of (C(ω∗), τp) consisting of 0 -1-valued
functions. The role of the space E from the previous section will be played by
the following space
F = {(f, c) ∈ Cp(ω∗, 2) × C : f A∗
n ≡ 0, f B∗
n ≡ 1 for every n};
where c = (c0, c1, c2, . . .) and ci = (Ai, Bi).
10
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
We will say that a pair z ∈ Z and a sequence s = ((A0, B0), . . . , (An, Bn)) ∈ S
n. Clearly, if s captures z (see Sec. 3),
n = ∅ = z(1) ∩ A∗
are consistent if z(0) ∩ B∗
then s and z are consistent.
Basic open neighborhoods in F are of the form O(z, s), where z ∈ Z and s ∈ S
are consistent, and O(z, s) is the set of all (f, c) ∈ F such that
(a) f (x) = i for every x ∈ z(i), i = 0, 1;
(b) s ≺ c.
Note that the condition that s and z are consistent implies that every set
O(z, s) is nonempty.
Repeating the proof of Lemma 3.2, one easily obtains the following
Lemma 4.4. For any consistent z ∈ Z and s ∈ S, the basic open set O(z, s) in
F contains a neighborhood O(z, s′), where s′ captures z.
The proof of Theorem 4.1 is based on the following auxiliary result.
Lemma 4.5. For any sequence (Xn)n∈ω of C-sets in F there exist a sequence
c = ((An, Bn))n∈ω ∈ C and sets A, B ⊆ ω such that
-- Sn∈ω A∗
n ⊆ A∗, Sn∈ω B∗
-- µ(A), µ(B) ≤ 1
6 ;
-- for any n ∈ ω, the set
n ⊆ B∗, A ∩ B = ∅,
{f ∈ Cp(ω∗, 2) : f A∗ ≡ 0, f B∗ ≡ 1} × {c}
is either contained in Xn or disjoint from Xn.
Proof. We inductively define a decreasing sequence (Vn)n of nonempty open sub-
sets of F and for every n we choose
• sequences (U n
k )k of open sets dense in Vn such that Tk∈ω U n
k is either
contained in Xn or disjoint from Xn;
• zn ∈ Z and sn ∈ S capturing zn such that zn−1 ⊏ zn, sn−1 ≺ sn, and
O(zn, sn) ⊆ \i,k≤n
U i
k.
Suppose that the construction has been carried out for i < n (or n = 0). Since
Xn ∩ O(zn−1, sn−1)
ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
11
is a C−set there is a nonempty open set Vn ⊆ O(zn−1, sn−1) and a sequence of its
dense open subsets (U n
k is either contained in Xn or disjoint
k is open and nonempty (because Vi are
k )k such thatTk∈ω U n
k are dense in Vi). Moreover, Gn ⊆ Vn ⊆ O(zn−1, sn−1).
from it. Then the set Gn = Ti,k≤n U i
decreasing and U i
Now, we can choose consistent zn ∈ Z and sn ∈ S such that zn−1 ⊏ zn,
sn−1 ≺ sn, and O(zn, sn) ⊆ Gn; by Lemma 4.4 we can additionally require
that sn captures zn. The sequence s0 ≺ s1 ≺ . . . defines the unique element
c = ((An, Bn))n∈ω ∈ C. We also obtain the sets A, B in the same way as in the
proof of Lemma 3.4, applying Lemma 2.4 to sequences (An)n and (Bn)n.
It follows that whenever the function f ∈ Cp(ω∗, 2) takes values 0 on A∗ and
1 on B∗, the pair (f, c) belongs to F and, for every n, (f, c) ∈ O(zn, sn), since
sn ≺ c and sn captures zn. Therefore
(f, c) ∈ \n∈ω \i,k≤n
U i
k = \n∈ω\k∈ω
U n
k ⊆ \k∈ω
U n
k ,
for every n, and the lemma follows.
(cid:3)
Theorem 4.1 can be easily derived from the above lemma. Indeed, if Yn are
C−sets in (C(ω∗), τp) then Zn = Yn ∩ Cp(ω∗, 2) are C−sets in Cp(ω∗, 2). Let
π : F → Cp(ω∗, 2) be the projection onto the first axis. Then Xn = π−1(Zn) are
C−sets in the space F. Applying Lemma 4.5 to such sets Xn we conclude that
there are two different functions g1, g2 with giA∗ ≡ 0, giB∗ ≡ 1 for i = 1, 2. It
follows that (gi, c) are not separated by Xn and hence gi are not separated by the
sets Yn.
Remark 4.6. Adjusting the argument from the proof of Theorem 4.1 for the
setting from the paper [MP], one can also prove that, for the Stone space S of
the measure algebra, the function space (C(S), τp) has no sequence of C−sets
separating points (this is an unpublished result of R. Pol). It follows that the
Banach spaces C(S) and C(βω) are isomorphic, but there is no Borel-measurable
injection ϕ : (C(S), τp) → (C(βω), τp).
5. Baire σ-algebras in function spaces on βω and ω∗
If (X, τ ) is any topological space then the Baire σ-algebra Ba(X, τ ) is defined
to be the smallest one making all the continuous functions on X measurable, cf.
[Ed1], [Ed2]. Recall that if (X, τ ) is a locally convex linear topological space then
12
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
Ba(X, τ ) is actually generated by all continuous functionals, see [Ed1], Theorem
2.3.
For a compact space K, we denote by Ba(C(K), weak), Ba(C(K), τp) the Baire
σ-algebras in C(K) endowed with the weak topology, or the pointwise topology,
respectively,
Theorem 3.5 implies directly the following result from [APR1, Theorem 3.4]
Theorem 5.1 (Avil´es-Plebanek-Rodr´ıguez).
Ba(C(βω), weak) 6= Ba(C(βω), τp) .
Using results from section 2 we can also give a simpler proof of the above
theorem:
Ba(C(βω), τp)-measurable. Assume the contrary. Then there exists a countable
Proof. We shall show that, for any ultrafilter ℘ ∈ ω∗, the measure bd℘ is not
subset X of βω such that bd℘ is measurable with respect to the σ-algebra of subsets
of C(βω) generated by {δx : x ∈ X}. Corollary 2.5 implies that bdp vanishes
bd℘(A) < 1. Let E = {f ∈ C(βω) : f A ≡ 0}. Observe that bd℘E is measurable
with respect to the σ-algebra generated by {δx : x ∈ X ∩ ω}, and for any subset
C of B = ω \ A the characteristic function χC : βω → R belongs to E. Then the
measure ν : P(ω) → [0, 1] defined by
on the closure of X ∩ ω∗ in βω. Take A ⊆ ω such that X ∩ ω∗ ⊆ A and
v(Z) = d℘(Z ∩ B) = bd℘(Z ∩ B) = bd℘(χZ∩B) ,
for Z ∈ P(ω), is nonzero, Borel-measurable and vanishes on points of ω, a con-
tradiction with Proposition 2.2.
(cid:3)
Note finally that for any compact space K we have the following inclusions
Ba(C(K), τp) ⊂ Borel(C(K), τp)
∩
∩
Ba(C(K), weak) ⊂ Borel(C(K), weak) ⊂ Borel(C(K), norm).
The space K = 2ω1 is an example of a nonmetrizable compactum K for which
all the five σ-algebras on C(K) are equal, see [APR2]. From previous results and
the proposition below it follows that all inclusions in the above diagram are strict
for the space K = βω. Since βω is a continuous image of ω∗, this is also the case
for K = ω∗, cf. [APR1, Corollary 3.3].
ON BOREL STRUCTURES IN THE BANACH SPACE C(βω)
13
Proposition 5.2. Bor(C(βω), τp) 6⊆ Ba(C(βω), weak).
Proof. Let {Aα : α < ω1} be a family of almost disjoint subsets of ω. For every
α < ω1 we pick Fα ∈ ω∗ such that Aα ∈ Fα. Let us consider the set
V = {f ∈ C(βω) : f (Fα) > 0 for some α < ω1}.
Then V is τp-open; we shall check that V /∈ Ba(C(βω), weak).
Suppose otherwise; then V lies in the σ-algebra generated by {δn : n ∈ ω} and
some family {µn : n ∈ ω}, where every µn is a probability measure on ω∗. There
is β < ω1 such that µn(Aβ) = 0 for every n. Let F be the set of all 0 -1-valued
functions in C(βω) which vanish outside Aβ. It follows that the set F ∩ V lies in
the σ-algebra of subsets of F which is generated by the restrictions of µn's and
δn's to F which is simply the σ-algebra generated by δn for n ∈ Aβ. On the other
hand, F ∩ V = {χN : N ∈ Fβ, N ⊆ Aβ}, a contradiction, since Fβ ∩ 2Aβ is not
Borel in the Cantor set 2Aβ .
(cid:3)
References
[APR1] A. Avil´es, G. Plebanek, and J. Rodr´ıguez, On Baire measurability in spaces of contin-
uous functions, J. Math. Anal. Appl. 398 (2013), 230 -- 238.
[APR2] A. Avil´es, G. Plebanek, J. Rodr´ıguez, Measurability in C(2κ) and Kunen cardinals,
Israel J. Math. 195 (2013), 1 -- 30.
T. Bartoszyski and H. Judah, Set theory. On the structure of the real line, A K Peters,
[BJ]
Ltd., Wellesley, MA, 1995.
[BFPR] A. Blass, R. Frankiewicz, G. Plebanek, C. Ryll-Nardzewski, A note on extensions of
asymptotic density, Proc. Amer. Math. Soc. 129 (2001), 3313 -- 3320.
[BPo] D.K. Burke and R. Pol, On Borel sets in function spaces with the weak topology, J.
London Math. Soc. (2) 68 (2003), no. 3, 725 -- 738.
[Ed1] G.A. Edgar, Measurability in a Banach space, Indiana Univ. Math. J. 26 (1977), no.
4, 663 -- 677.
[Ed2] G.A. Edgar, Measurability in a Banach space. II, Indiana Univ. Math. J. 28 (1979),
no. 4, 559 -- 579.
[Fr]
D.H. Fremlin, Measure Theory, Vol. 4: Topological Measure Theory, Torres Fremlin,
2003.
A.S. Kechris, Classical Descriptive Set Theory, Springer-Verlag, New York, 1995.
[Ke]
[MP] W. Marciszewski and R. Pol, On some problems concerning Borel structures in function
spaces, RACSAM Rev. R. Acad. Cienc. Exactas F´ıs. Nat. Ser. A Mat., 104 (2010),
327 -- 335.
14
[Me]
[Ox]
[Ra]
[Ta]
WITOLD MARCISZEWSKI AND GRZEGORZ PLEBANEK
A.H. Mekler, Finitely additive measures on N and the additivity property, Proc. Amer.
Math. Soc. 92 (1984), 439 -- 444.
J.C. Oxtoby, Measure and category, volume 2 of Graduate Texts in Mathematics.
Springer Verlag, New York-Berlin, 1980.
M. Raja, Kadec norms and Borel sets in a Banach space, Studia Math. 136 (1999),
1 -- 16.
M. Talagrand, Comparaison des boreliens dun espace de Banach pour les topologies
fortes et faibles, Indiana Univ. Math. J. 27 (1978), 1001 -- 1004.
Institute of Mathematics, University of Warsaw, Banacha 2
02 -- 097 Warszawa, Poland
E-mail address: [email protected]
Instytut Matematyczny, Uniwersytet Wroc lawski, pl. Grunwaldzki 2/4,
50 -- 384 Wroc law, Poland
E-mail address: [email protected]
|
1504.05547 | 2 | 1504 | 2015-06-26T01:57:05 | Asymptotic estimates on the von Neumann inequality for homogeneous polynomials | [
"math.FA",
"math.OA"
] | By the von Neumann inequality for homogeneous polynomials there exists a positive constant $C_{k,q}(n)$ such that for every $k$-homogeneous polynomial $p$ in $n$ variables and every $n$-tuple of commuting operators $(T_1, \dots, T_n)$ with $\sum_{i=1}^{n} \Vert T_{i} \Vert^{q} \leq 1$ we have \[ \|p(T_1, \dots, T_n)\|_{\mathcal L(\mathcal H)} \leq C_{k,q}(n) \; \sup\{ |p(z_1, \dots, z_n)| : \textstyle \sum_{i=1}^{n} \vert z_{i} \vert^{q} \leq 1 \}\,. \] For fixed $k$ and $q$, we study the asymptotic growth of the smallest constant $C_{k,q}(n)$ as $n$ (the number of variables/operators) tends to infinity. For $q = \infty$, we obtain the correct asymptotic behavior of this constant (answering a question posed by Dixon in the seventies). For $2 \leq q < \infty$ we improve some lower bounds given by Mantero and Tonge, and prove the asymptotic behavior up to a logarithmic factor. To achieve this we provide estimates of the norm of homogeneous unimodular Steiner polynomials, i.e. polynomials such that the multi-indices corresponding to the nonzero coefficients form partial Steiner systems. | math.FA | math | ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS
POLYNOMIALS
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
ABSTRACT. By the von Neumann inequality for homogeneous polynomials there exists a positive constant
Ck,q (n) such that for every k-homogeneous polynomial p in n variables and every n-tuple of commuting
operators (T1,... , Tn ) withPn
i=1 kTi kq ≤ 1 we have
kp(T1,... , Tn )kL (H ) ≤ Ck,q (n) sup{p(z1,... , zn ) :Pn
i=1 zi q ≤ 1}.
For fixed k and q, we study the asymptotic growth of the smallest constant Ck,q (n) as n (the number of
variables/operators) tends to infinity. For q = ∞, we obtain the correct asymptotic behavior of this con-
stant (answering a question posed by Dixon in the seventies). For 2 ≤ q < ∞ we improve some lower
bounds given by Mantero and Tonge, and prove the asymptotic behavior up to a logarithmic factor. To
achieve this we provide estimates of the norm of homogeneous unimodular Steiner polynomials, i.e.
polynomials such that the multi-indices corresponding to the nonzero coefficients form partial Steiner
systems.
5
1
0
2
n
u
J
6
2
]
.
A
F
h
t
a
m
[
2
v
7
4
5
5
0
.
4
0
5
1
:
v
i
X
r
a
1. INTRODUCTION
A classical inequality in operator theory, due to von Neumann [30], asserts that if T is a linear con-
traction on a complex Hilbert space H (i.e., its operator norm is less than or equal to one) then
kp(T )kL (H ) ≤ sup{p(z) : z ∈ C, z ≤ 1},
for every polynomial p in one (complex) variable. Note that, as a direct consequence of von Neumann's
inequality, we can define a functional calculus on the disk algebra. There are many other consequences
of this important inequality in functional analysis; we refer the reader to [25, Chapter 1] and the refer-
ences therein for a fuller treatment of this inequality and its applications.
For some time, it was very natural to ask whether the von Neumann inequality could be extended to
polynomials in two or more commuting contractions. For polynomials in two contractions Ando [2],
using "dilation theory" (see [28]), provided a positive answer. However, in the mid seventies, Varopou-
los [29] showed that von Neumann's inequality cannot be extended to three or more contractions. For
2000 Mathematics Subject Classification. Primary 47A13, 47A60, Secondary 28A78, 60G99, 46G25, 05B05.
Key words and phrases. Multivariable von Neumann inequality; Commuting contractions, unimodular homogeneous
polynomials, Steiner systems.
The first two named authors were supported by projects CONICET PIP 0624, PICT 2011-1456, UBACyT 20020130300057BA,
UBACyT20020130300052BA. The third named author was supported by project MTM2014-57838-C2-2-P.
1
2
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
this, he used the metric theory of tensor products together with probabilistic tools to construct a poly-
nomial and operators that violate the inequality. The work of Varopoulos has since been simplified and
extended by several authors [5, 9, 15, 21, 22].
It is an open problem of great interest in operator theory (see [6, 25]) to determine whether there
exists a constant K (n) that adjusts von Neumann's inequality. More precisely, it is unknown whether or
not for every n there exists a constant K (n) such that
(1)
kp(T1,..., Tn)kL (H ) ≤ K (n) sup{p(z1,... , zn) : zi ≤ 1},
for every polynomial p in n variables and every n-tuple (T1,... , Tn) of commuting contractions in L (H ).
Dixon in [15] gave lower estimates for the optimal K (n) and showed that, if such a constant verifying
(1) exists, then it must grow faster than any power of n. He did this by considering the problem in the
smaller class of k-homogeneous polynomials. More precisely, he studied the asymptotic behavior (as
n, the number of variables/operators, tends to infinity) of the smallest constant Ck,∞(n) such that
(2)
kp(T1,..., Tn)kL (H ) ≤ Ck,∞(n) sup{p(z1,..., zn) : zi ≤ 1},
for every k-homogeneous polynomial p in n variables and every n-tuple of commuting contractions
(T1,..., Tn). In [15, Theorem 1.2] he showed that
(3)
n
1
2£ k−1
2 ¤ ≪ Ck,∞(n) ≪ n
k−2
2
,
where [x] denotes the integer part of x. For the lower bound Dixon used probabilistic techniques (the
Kahane-Salem-Zygmund theorem) and combinatorial ideas (Steiner systems) along with an ingenious
construction of the operators and the Hilbert space involved.
This problem was taken up by Mantero and Tonge in [21]. Among other problems, for each 1 ≤ q < ∞
they consider Ck,q (n), the smallest constant such that
(4)
kp(T1,..., Tn)kL (H ) ≤ Ck,q (n) sup{p(z1,..., zn) :
n
Xj=1z jq ≤ 1},
for every k-homogeneous polynomial p in n variables and every n-tuple of commuting contractions
L (H ) ≤ 1. They give upper and lower estimates for the growth of Ck,q (n) [21,
q
i=1kTik
(T1,..., Tn) withPn
Propositions 11 and 17] (here q′ denotes the conjugate of q; see below):
(5)
(6)
n
n
k
2 − 1
k−1
q′ − 1
2£ k
2¤ ≪ Ck,q (n) ≪ n
2¡£ k
2¤+1¢ ≪ Ck,q (n) ≪ n
k−2
q′
k−2
2
for 1 ≤ q ≤ 2,
for 2 ≤ q < ∞.
It is worth noting that the upper bounds here hold for every n-tuple (T1,..., Tn) satisfyingPn
i=1kTikq ≤ 1
(and even a weaker condition), not necessarily commuting. If we do not ask the contractions to com-
mute, this bound is shown to be optimal in [21, Proposition 15].
ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS POLYNOMIALS
3
Based on the combinatorial methods from [15] (i.e., considering polynomials whose monomials are
determined by Steiner blocks) we change the construction of the Hilbert space and the operators given
there to find the exact asymptotic growth of Ck,∞(n), answering a question that was explicitly posed by
Dixon.
On the other hand, by applying some probabilistic tools used by Bayart in [3], we are able to control the
increments of a Rademacher process and in this way we in this way we manage to narrow the range in
(6), showing that the exponent in the power of n is indeed optimal. We collect this in our main result.
Theorem 1.1. For k ≥ 3 and 1 ≤ q ≤ ∞, let Ck,q(n) be the smallest constant such that
kp(T1,..., Tn)kL (H ) ≤ Ck,q (n) sup{p(z1,... , zn) : k(z j ) jkq ≤ 1},
for every k-homogeneous polynomial p in n variables and every n-tuple of commuting contractions
L (H ) ≤ 1. Then
q
i=1kTik
(T1,... , Tn) withPn
(i) Ck,∞(n) ∼ n
(ii) for 2 ≤ q < ∞ we have
k−2
2
log−3/q (n) n
In particular, n
k−2
2 −ε ≪ Ck,q (n) ≪ n
k−2
2
k−2
2
.
k−2
2 ≪ Ck,q (n) ≪ n
for every ε > 0.
The proof of this result will be given in Section 3.
2. STEINER UNIMODULAR POLYNOMIALS
The systematic study of norms of random homogeneous polynomials started with the Kahane-Salem-
Zygmund theorem [17, Chapter 6], which is found very useful in Fourier analysis. More recently, ap-
plications of norms of random polynomials with unimodular coefficients were found in complex and
functional analysis (see for example [7, 13, 8, 3]).
The philosophy in this problem and in many others of the same kind (e.g. to compute the Sidon
constant for polynomials [23, 11]) is to find polynomials which have "big" (or "many") coefficients, but
whose maximum modulus on the unit ball is "small".
In this section we are going to relax the number of terms appearing in the polynomials, by allowing
them to have some zero coefficients. In this way we will find a special class of tetrahedral unimodular
polynomials having many terms, but keeping the maximum modulus quite small.
Let us first start with some notation and preliminaries. As usual we will denote ℓn
q for Cn with the
norm k(z1,..., zn)kq =¡Pn
A k-homogeneous polynomial in n variables is a function p : Cn → C of the form
c J z j1 ··· z jk ,
i=1ziq¢1/q if 1 ≤ q < ∞ and k(z1,..., zn)k∞ = maxi=1,...,n zi for q = ∞.
p(z1,..., zn) = Xα∈Nn
n = XJ=( j1,..., jk)
1 ··· zαn
aαzα1
1≤ j1≤...≤ jk≤n
0
α=k
4
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
where aα ∈ C and α = α1 +··· + αn. Given α we have aα = c J where J = (1, α1...,1,..., n, αk..., n). We will
write zα1
q ) the Banach space of all
n = zα and z j1 ··· z jk = z J . For 1 ≤ q ≤ ∞ we denote by P (k ℓn
1 ··· zαn
k-homogeneous polynomials on n variables with the norm
kpkP (k ℓn
q ) = sup{p(z1,..., zn) : k(z1,..., zn)kq ≤ 1}.
It is a well known fact (see e.g.
[14, Chapter 1]) that for every k-homogeneous polynomial there is
a unique symmetric k-linear form L on Cn such that p(z) = L(z,... , z) for all z ∈ Cn. Also for each
1 ≤ q ≤ ∞ and k ≥ 2 there exists a constant λ(k, q) > 0 such that
(7)
q ) ≤ sup{L(z(1),..., z(k)): kz( j )kq ≤ 1, j = 1,... , k} ≤ λ(k, q)kpkP (kℓn
q ) .
k! but improvements in concrete cases include λ(k,2) = 1 and λ(k,∞) ≤ k
kpkP (kℓn
In general λ(k, q) ≤ k k
(see [14, Propositions 1.44, 1.43]).
If (an)n and (bn)n are two sequences of real numbers we will write an ≪ bn if there exists a constant
C > 0 (independent of n) such that an ≤ C bn for every n. We will write an ∼ bn if an ≪ bn and bn ≪ an.
Given a set A we will denote its cardinality by A.
For an index 1 < q < ∞ we denote by q′ its conjugate: 1 = 1
q + 1
q′ .
k+1
2
k
2 (k+1)
2k k!
Let C ⊂ Nn
0 denote any set of multi-indices α with α = k. Then as a consequence of the Kahane-
Salem-Zygmund theorem [17, Chapter 6] there exists a k-homogeneous polynomial, with unimodular
coefficients aα for α ∈ C and aα = 0 if α ∉ C , of small maximum modulus on the n-polydisk. More
precisely, let (εα)α∈C be independent Bernoulli variables on a probability space (Ω, Σ, P), then we have
P{ω ∈ Ω : k Xα∈C
εα(ω)zαkP (k ℓn
∞
) ≥ D¡n log(k)C ¢1/2} ≤
1
k 2e n ,
where D > 0 is an absolute constant which is less than 8. In particular there are signs (aα)α∈C such that
the k-homogeneous unimodular polynomial
(8)
satisfies
(9)
p(z) = Xα∈C
aαzα ,
kpkP (kℓn
∞
) ≤ D¡n log(k)C ¢1/2 .
We are going to work with polynomials with many zero coefficients, expecting that this will make the
norm of the polynomial small enough. The presence of C in (9) is sufficient for our needs when the
norm of the polynomial is computed in ℓn
p and then we need
∞
different tools. The relevant results we have to hand [7, 12, 13, 3] do not take into account the number
but not when we consider the norm in ℓn
of non-zero coefficients, so considering our tetrahedral polynomials does not improve these estimates.
We deal with polynomials with a particular combinatorial configuration in order to get useful estimates
for our purposes. We modify some arguments from [3], reflecting this configuration.
ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS POLYNOMIALS
5
To achieve our goal we consider special subsets of multi-indices: partial Steiner systems on the set
{1,... , n}. An S p(t, k, n) partial Steiner system is a collection of subsets of size k of {1,... , n} such that
every subset of t elements is contained in at most one member of the collection of subsets of size k.
Definition 2.1. A k-homogeneous polynomial of n variables, is a Steiner unimodular polynomial
if
there exists an S p(t, k, n) partial Steiner system S such that p(z1,..., zn) =PJ∈S c J z J and c J = ±1.
Observe that our Steiner unimodular polynomials are tetrahedral, i.e. in every term z J each variable
z j0 appears at most once. In other words, no term in the polynomial contains a factor of degree 2 or
higher in any of the variables z1,..., zn.
The first one to consider Steiner unimodular polynomials was Dixon [15], who used S p([(k − 1)/2], k, n)
partial Steiner systems. He used this to obtain lower bounds for (2). The combinatorial property was
only applied to define some Hilbert space operators that violate the inequality, but not to estimate the
norm of the polynomial, which he did using (9) and the number of non-zero coefficients.
In the following lemmas, in ℓn
q , 1 ≤ q < ∞, we will strongly use the fact that the multi-indices of
the non-zero coefficients form a partial Steiner system to estimate the maximum modulus. We use an
entropy argument due to Pisier to control the increments of a Rademacher process and subsequently
apply an interpolation argument.
Let us first recall some definitions and a result on regularity of random process. A complete account
on these can be found in [20, Chapters 4 and 11].
A Young function ψ is a convex increasing function defined on [0,∞[ such that limt→∞ ψ(t) = ∞ and
ψ(0) = 0. For a probability space (Ω, Σ, P), the Orlicz space Lψ = Lψ(Ω, Σ, P) is defined as the space of
all real-valued random variables Z for which there exists c > 0 such that E(ψ(Z/c)) < ∞. It is a Banach
space with the norm kZkLψ = inf{c > 0 : E(ψ(Z/c)) ≤ 1}.
Let (X , d) be a metric space. Given ε > 0, the entropy number N (X , d; ε) is defined as the smallest
number of open balls of radius ε in the metric d, which form a covering of the metric space X .
With this, the entropy integral of (X , d) with respect to ψ is given by
Jψ(X , d) :=Zdiam(X )
0
ψ−1(N (X , d; ε))d ε.
We are going to define a random process (Yz)z∈Bℓn
supz Yz. To do so, we use the following theorem due to Pisier [24] (see also [20, Theorem 11.1]) that
bounds this expectation with the entropy integral, provided that the random process satisfies a certain
and we will need to estimate the expectation of
2
contraction condition.
Theorem 2.2. Let Z = (Zx)x∈X be a random process indexed by (X , d) in Lψ such that, for every x, x′ ∈ X ,
kZx − Zx′kLψ ≤ d(x, x′).
6
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
Then, if Jψ(X , d) is finite, Z is almost surely bounded and
E¡ sup
x,x′∈X Zx − Zx′¢ ≤ 8Jψ(X , d).
Let now k ≥ 2 and let S be a S p(k − 1, k, n) partial Steiner system. We consider a family of indepen-
we define the following
dent Bernoulli variables (εJ )J∈S on the probability space (Ω, Σ, P). For z ∈ Bℓn
Rademacher process indexed by Bℓn
as
2
2
(10)
Yz =
1
k XJ∈S
εJ z J .
We view it as a random process in the Orlicz space defined by the Young function ψ2(t) = e t 2
− 1.
Lemma 2.3. The Rademacher process defined in (10) fulfils the following Lipschitz condition:
kYz − Yz′kLψ2 ≤ Ckz − z′k∞,
for some universal constant C ≥ 1 and every z, z′ ∈ Bℓn
2
.
Proof. As a consequence of Khintchine inequalities (see e.g. [10, Sect 8.5]), the ψ2-norm of a Rademacher
process is comparable to its L2-norm. Now,
kYz − Yz′kL2 =
=
≤
≤
k
1
1
k¡ZΩ XJ∈S
εJ (ω)(z J − z′J )2d P(ω)¢1/2
k¡ XJ∈S
Xu=1
)z′ju+1
z j1 ... z ju−1(z ju − z′ju
Xu=1¡ XJ∈S z j1 ... z ju−1(z ju − z′ju
)z′ju+1
Xu=1kz − z′k∞¡ XJ∈S z j1 ... z ju−1 z′ju+1
1
k
1
k
k
k
1
=
k¡ XJ∈S z J − z′J2¢1/2
... z′jk2)1/2
... z′jk2¢1/2
... z′jk2¢1/2
Since S is an S p(k − 1, k, n) partial Steiner system, given j1,..., ju−1, ju+1,... , jk for a fixed u, there is at
... z′jk2
most one index ju such that ( j1,... , jk) belongs to S . Therefore the sumPJ∈S z j1 ... z ju−1 z′ju+1
can be bounded by
n
n
n
(
Xlu−1=1zlu−12)(
and this is less than or equal to one (since z, z′ ∈ Bℓn
concludes the proof.
Xl1=1zl12)··· (
2
n
Xlu+1=1z′lu+12)··· (
Xlk=1z′lk2),
). This combined with the previous inequality
(cid:3)
We are now in a position to use Theorem 2.2 with Lψ2, X = Bℓn
and d = k · k∞ to bound the ex-
,k · k∞). Note that
pectation of the supremum. For this, we estimate the entropy integral Jψ2(Bℓn
2 (t) = log1/2(t + 1); we use instead log1/2(t), which does not change the computation of the inte-
ψ−1
gral. We estimate the integral in the following result, which is a version of [3, Lemma 2.1]; the proof is
2
2
essentially the same and we include it here for the sake of completeness.
ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS POLYNOMIALS
7
Lemma 2.4. There exists C > 0 such that for every n ≥ 2 we have
Jψ2 (Bℓn
2
,k · k∞) ≤ C log3/2(n).
Proof. We fix n and for each m we consider the number
By result of Schütt [27, Theorem 1] there exists a constant K , independent of n and m, such that
em = inf{σ > 0 : Bℓn
2 ⊂
xi + σBℓn
∞
}.
2m
[i=1
(11)
em ≤ K ×
1
m )
¡ log(1+ 2n
m
2− m
2n n− 1
2
¢1/2
if m ≤ log(n),
if log(n) ≤ m ≤ 2n,
if m ≥ 2n .
Let us note that Schütt's result is stated for real spaces. Since the (2n)-dimensional real euclidean space
2 we get (11).
2 ≤ ε < K 2− m
2n n− 1
2n n− 1
is isometrically isomorphic to ℓn
For m ≥ 2n, if K 2− m+1
ZK /(2pn)
log1/2¡N (Bℓn
0
2
With the same argument, if K
integral from K
2pn
to Kq log2
We define now εm =¡ log(1+ 2n
2m+1. Then
2 , then by (11) we have N (Bℓn
,k · k∞; ε) ≤ 2m+1 ≤ 2 K 2n
ε2n nn and
2
0
ε2n¢d ε =ZK /2
n1/2 log1/2¡ 2K 2
,k · k∞; ε)¢d ε ≤ZK /(2pn)
2pn ≤ εKq log2
¢1/2 for [log n] ≤ m < 2n. Again by (11), if εm+1 ≤ ε < εm then N (Bℓn
u2 ¢d u = K1 < ∞.
,k · k∞; ε) ≤ 22n and with this we can bound the
log1/2¡ 2C 2
2n by some K2.
m )
2n then N (Bℓn
m
0
2
2
,k ·k∞; ε) ≤
Zε[log(n)]
ε2n
log1/2¡N (Bℓn
2
We write
,k · k∞; ε)¢d ε ≤
(εm − εm+1) = Khlog1/2(1+ 2n
m1/2
m )
and we get
−
2n−1
Xm=[log(n)]
log1/2(1+ 2n
m )
(m + 1)1/2 +
(m + 1)1/2(εm − εm+1)log1/2(2).
log1/2(1+ 2n
m )
(m + 1)1/2 −
log1/2(1+ 2n
m+1 )
(m + 1)1/2
i
Zε[log(n)]
ε2n
2
log1/2¡N (Bℓn
≤ K³log1/2(n)
,k · k∞; ε)¢d ε
Xs=[log(n)]
2s−1
2n
s + 1
Finally, for the remaining subinterval we have that, by (11), if ε ≥ ε[log(n)], then N (Bℓn
Hence
)− log1/2(1+
Xs=[log(n)]
log1/2(1+
(s + 1)1/2
s3/2 +
2n
s
2s−1
2
)´ ≤ K3 log3/2(n).
,k · k∞; ε) ≤ 2log(n).
Z1
ε[log(n)]
log1/2¡N (Bℓn
2
,k · k∞; ε)¢d ε ≤ K4Z1
0
log1/2(n)d ε ≤ K4 log1/2(n).
(cid:3)
This completes the proof.
We can now find Steiner unimodular polynomials that have small norm in P (k ℓn
q ), for every 2 ≤ q ≤
∞ simultaneously.
8
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
Theorem 2.5. Let k ≥ 2 and S be an S p(k − 1, k, n) partial Steiner system. Then there exist signs (c J )J∈S
and a constant Ak,q > 0 independent of n such that the k-homogeneous polynomial p =PJ∈S c J z J satis-
fies
kpkP (kℓn
q ) ≤ Ak,q ×
log
3
q (n)n
k
2 ( q−2
q )
3q−3
q (n)
log
for 2 ≤ q < ∞
for 1 ≤ q ≤ 2.
Moreover, the constant Ak,q may be taken independent of k for q 6= 2.
Proof. To prove this theorem we will first find a polynomial with small norm both in P (k ℓn
2 ) and in
). For this we use an interesting technique borrowed from the proof of [8, Lemma 2.1], followed
P (k ℓn
∞
by an interpolation argument.
k¡ n
k−1¢. We use S to
Note first that any S p(k − 1, k, n) partial Steiner system S satisfies that S ≤ 1
as in (10). By Lemma 2.3, Theorem 2.2 and Lemma 2.4 there is
define a Rademacher process (Yz)z∈Bℓn
a constant K > 0 such that E(supz∈Bℓn
P{ω ∈ Ω : k XJ∈S
(12)
2
2 Yz) ≤ K log3/2(n). Therefore, by Markov's inequality we have
εJ (ω)z JkP (k ℓn
2 ) ≥ M kK log3/2(n)} ≤
1
M
,
where M is some constant to be determined. On the other hand, recall that by (8) we have
εJ (ω)z JkP (k ℓn
P{ω ∈ Ω : k XJ∈S
k 2en−1 (note that we can take M = 2 here) we have the following inequalities for ω
) ≥ D¡n log(k)S ¢1/2} ≤
∞
1
k 2e n ,
Therefore, if M > 1+ 1
in a positive measure set
kPJ∈S εJ (ω)z JkP (kℓn
kPJ∈S εJ (ω)z JkP (kℓn
2 ) ≤ M kK log3/2(n),
) ≤ D¡n log(k)S ¢1/2
∞
≤ D³ log(k)
k
k−1¢n´1/2
¡ n
≤ D³ log(k)
k! nk´1/2
.
There is a choice of signs (c J )J∈S such that the polynomial p(z) :=PJ∈S c J z J satisfies the inequalities in
q ) for 2 < q < ∞.
We consider the k-linear form associated to p then [4, Theorem 4.4.1], together with (7) and (13), give
(13). We now use an interpolation argument to obtain a bound of the norm of p in P (k ℓn
(13)
(14)
(15)
kpkP (kℓn
log1/2(k)
q
q ) ≤¡M kK¢2/q¡Dλ(k,∞)
≤ max{M K , D}³ k
¢ q−2
pk!
2 (k + 1) k+1
2 plog k
´ q−2
2k k!pk!
{z
Ak,q
k
}
log3/q (n)n
k
2 ( q−2
q )
q k
2
q
log3/q (n)n
k
2 ( q−2
q ) .
Note that for q > 2, Ak,q → 0 as k → ∞, and thus we may take a constant independent of k in this
case.
For q = 1, it is immediately seen that every Steiner unimodular polynomial has norm less than or
equal to one. Actually, more can be said. Let P(z) =Pα=k aαzα be any k-homogeneous polynomial.
ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS POLYNOMIALS
9
Then
(16)
P(z) ≤ Xα=kaαzα ≤ sup
α=knaα
α!
k!o Xα=k
k!
α!
1 and ℓn
In particular, the polynomial p considered above satisfies kpkP (kℓn
polation between the ℓn
2 cases we obtain that for 1 < q < 2,
q ) ≤³ k k
q ¡M kK log3/2(n)¢ 2q−2
q = Ak,q log
Note that also in this case, for every 1 ≤ q < 2 we have Ak,q → 0 as k → ∞.
(k!)2´ 2−q
kpkP (kℓn
zα = sup
α=knaα
1 ) ≤ 1
α!
k!o³ n
Xj=1z j´k
.
k! . Finally, proceeding by inter-
3q−3
q (n).
(cid:3)
As was already noted in [12, Corollary 6.5], the argument in (16) improves the estimates given in [7]
and [3, Corollary 3.2] for the q = 1 case.
2 ) less than or equal to 1
Remark 2.6. It is not difficult to prove that every 2-homogeneous Steiner unimodular polynomial has
norm in P (2ℓn
2. It would be interesting to know if there exists a constant C ,
perhaps depending on k ≥ 3 and not on n, such that given any S p(k − 1, k, n) partial Steiner system
S , we can find a k-homogeneous unimodular polynomial p(z) :=PJ∈S c J z J with kpkP (kℓn
2 ) ≤ C . An
affirmative answer to this question would in particular give that the upper bound given by Mantero
and Tonge (6) for Ck,q(n) with 2 ≤ q < ∞ is actually optimal.
The last ingredient we need for our applications is the existence of nearly optimal partial Steiner
systems, in the sense that they have many elements. This translates to many unimodular coefficients
of the Steiner polynomials. It is well known that any partial Steiner system S p(t, k, n) has cardinality less
t¢/¡k
than or equal to¡n
t¢. A conjecture of Erdos and Hanani [16], proved positively by Rödl [26], states
that there exist partial Steiner systems S p(t, k, n) of cardinality at least (1−o(1))¡n
t¢, where o(1) tends
to zero as n goes to infinity. This bound was improved in [1] (see also [19] for a panoramic overview of
the subject), where it is proved that there exists a constant c > 0 such that there exist partial Steiner
systems S p(k − 1, k, n) of cardinality at least
¡ n
k−1¢k ³1−
¡ n
k−1¢k ³1−
k−1´,
´,
for k = 3.
for k > 3,
t¢/¡k
c log3/2 n
1
k−1
n
c
1
n
(17)
Taking partial Steiner systems of this cardinality in Theorem 2.5 we have the following.
Corollary 2.7. Let k ≥ 3. Then there exists a k-homogeneous Steiner unimodular polynomial p of n com-
plex variables with at least ψ(k, n) (defined in (17)) coefficients satisfying the estimates in Theorem 2.5.
Note that in this case ψ(k, n)≫ nk−1.
Remark 2.8. Very recently, a longstanding open problem in combinatorial design theory was solved by
Keevash [18]. A Steiner system S(t, k, n) is a collection of subsets of size k of {1,... , n} such that every
10
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
subset of t elements is contained in exactly one member of the collection of subsets of size k. Keevash's
result implies the asymptotic existence of Steiner systems, that is, that given t < k, Steiner systems
S(t, k, n) exist for every sufficiently large n that satisfies some natural divisibility conditions. In partic-
ular, for an infinite number of n's we may take ψ(k, n)=¡ n
k−1¢/k in the above corollary.
3. ESTIMATES ON THE MULTIVARIABLE VON NEUMANN INEQUALITY
In this section we estimate the asymptotic failure of different versions of the multivariable von Neu-
mann inequality for homogeneous polynomials. Before we prove Theorem 1.1, let us observe that we
modify Dixon's original proof of the lower bound in (3) in several ways.
Dixon considered partial Steiner systems S p([(k − 1)/2], k, n), for which the number of non-zero coef-
ficients is of the order n[ k−1
2 ]. This is not enough to find a good lower bound. Instead, we use partial
Steiner systems S p(k − 1, k, n). This allows us to have more non-zero coefficients, but also forces us to
make a new construction of the Hilbert space and the operators which we feel is closer to that given by
Varopoulos in [29].
Proof of Theorem 1.1 -- (i). The upper bound was proved in [15, Theorem 1.2]. Thus we only have to
construct a polynomial, a Hilbert space and commuting contractions that show that the asymptotic
growth of this bound is optimal.
Let n ≥ k ≥ 3 and choose a partial Steiner system S p(k − 1, k, n), denoted by S , such that S = ψ(k, n)
as in (17). By Theorem 2.5, see also (13), there exists a k-homogeneous polynomial p(z) =PJ∈S c J z J ,
with c J = ±1 for every J ∈ S and such that
(18)
kpkP (k ℓn
∞
) ≤ D³ log(k)
k
à n
k − 1!n´1/2
.
Let H be the (finite dimensional) Hilbert space which has as orthonormal basis the following vectors
e;
e( j1,..., jm)
fi
g .
for 0 ≤ m ≤ k − 2 and 1 ≤ j1 ≤ ··· ≤ jm ≤ n;
for i = 1,... , n;
Given any subset {i1,..., ir } ⊂ {1,... , n}, we denote by [i1,..., ir ] its nondecreasing reordering.
We define, for l = 1,..., n, the operators that act as follows on the basis of H ,
Tl e = e(l )
Tl e( j1,..., jm) = e[l , j1,..., jm],
Tl e( j1,..., jk−2) =Pi γ{i ,l, j1,..., jk−2} fi ,
Tl fi = δl i g ,
Tl g = 0,
if 0 ≤ m < k − 2
ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS POLYNOMIALS
11
where
γ{i1,...,ik} =
c{i1,...,ik },
0
if {i1,..., ik} ∈ S
otherwise.
Since S is an S p(k − 1, k, n) partial Steiner system, kTlk = 1 for l = 1,..., n. It is easily checked that the
operators commute. We have
p(T1,..., Tn)e = X{i1,...,ik }∈S
c{i1,...,ik }Ti1Ti2 ... Tik e = X{i1,...,ik }∈S
c2
{i1,...,ik}g = S g = ψ(k, n)g .
Now, using (18) we get
kp(T1,... , Tn)kL (H ) ≥ kp(T1,..., Tn)ekH = ψ(k, n)
This gives the desired conclusion.
≥
1
D³
¡ n
k−1¢
nk log(k)´ 1
2¡1− o(1)¢kpkP (kℓn
∞
) ≫ n
k−2
2 kpkP (kℓn
∞
).
(cid:3)
T j
Proof of Theorem 1.1 -- (ii). The upper bound was proved in [21, Corollary 11]. For the lower bound, we
take the Hilbert space and the operators T1,..., Tn defined in the proof of Theorem 1.1 -- (i). Then R j =
n1/q for j = 1,..., n clearly satisfyPn
i=1kRikq ≤ 1. Taking the polynomial p given by Theorem 2.5 we have
1
nk/q kp(T1,..., Tn)ekH = S
kp(R1,..., Rn)kL (H ) ≥
nk/q
kpkP (kℓn
Ak,q log3/q (n)n
q )S
2 ( q−2
q )
k
≥
nk/q ≥ A−1
k,qCk log−3/q (n)n
k−2
2 ).
2 kpkP (kℓn
This concludes the proof of the theorem.
(cid:3)
3.1. Other possible extensions of the von Neumann inequality for homogeneous polynomials: some
particular cases. Mantero and Tonge [21, Proposition 17] also obtained lower bounds for Ck,q,r (n),
defined as the least constant C such that
(19)
kp(T1,... , Tn)kL (H ) ≤ C sup{p(z1,..., zn) :
n
Xj=1z jq ≤ 1},
for every k-homogeneous polynomial p in n variables and every n-tuple of commuting contractions
L (H ) ≤ 1. Proceeding as in the proof of Theorem 1.1 -- (ii), we can show the
i=1kTikr
(T1,... , Tn) withPn
following.
Proposition 3.1. Let k ≥ 3, then
r )−1
q − 1
(i) log−3/q (n)nk( 1
(ii) log−3/q′(n)n
k
2+ 1
r′ −1 ≪ Ck,q,r (n), for q ≤ 2 and 1 ≤ r ≤ ∞.
≪ Ck,q,r (n), for q ≥ 2 and 1 ≤ r ≤ ∞,
Remark 3.2. The above proposition improves the lower bounds for Ck,q,r given in [21, Proposition 17]
in all cases but q ≤ 2 and k = 3.
12
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
Another possible multivariable extension of the von Neumann inequality (also studied in [21]) is by
considering polynomials on commuting operators T1,..., Tn satisfying that for any pair h, g of norm
one vectors in the Hilbert space,
(20)
Xj=1〈T j h, g〉q ≤ 1,
or, equivalently, that for any vector α ∈ Cn such that kαkℓn
n
q′ = 1, we have
Let Dk,q (n), denote the smallest constant such that
n
Xj=1
°°°
α j T j°°° ≤ 1.
(21)
kp(T1,..., Tn)kL (H ) ≤ Dk,q (n) sup{p(z1,..., zn) :
n
Xj=1z jq ≤ 1},
for every k-homogeneous polynomial p in n variables and every n-tuple of commuting contractions
(T1,..., Tn) satisfying (20). The upper bound obtained in [21, Proposition 20] is
Dk,q(n) ≪
(k−1)( 1
(k−1)( 1
2+ 1
q )
2+ 1
q′
)
n
n
for q ≥ 2,
for q ≤ 2.
For k = 3 and q = 2 we show that this is optimal up to a logarithmic factor.
Proposition 3.3. We have the following asymptotic behavior:
n2
log15/4 n ≪ D3,2(n) ≪ n2.
Proof. Let p(z) =PJ∈S c J z J be a 3-homogeneous Steiner unimodular polynomial as in Theorem 2.5
,...,
and let T1,..., Tn be the operators defined in the proof of Theorem 1.1 -- (i). We prove first that
satisfy (20). Note that these operators are defined on a (2n + 2)-dimensional Hilbert space H with or-
thonormal basis {e, e1,..., en, f1,..., fn, g }.
T1
kpk1/2
P (3 ℓn
2 )
Tn
kpk1/2
P (3 ℓn
2 )
For α ∈ ℓn
2 and h ∈ H , (below we take some β in the unit ball of ℓn
2 )
°°°Xj
2
α j T j h°°°
2
2
= 〈h, e〉2kαk2
α j〈h, f j〉¯¯¯
+¯¯¯Xj
α j〈h, el〉a{i , j ,l}¯¯¯
=Xj α j〈h, e〉2+Xi ¯¯¯Xj ,l
α j〈h, el〉a{i , j ,l}´2
α j〈h, f j〉¯¯¯
+¯¯¯Xj
2 +³Xi
βiXj ,l
2 k(〈h, el〉)lk2
2 )kαk2
2 +kpkP (3ℓn
ℓn
2 khk2
H .
≤ 〈h, e〉2kαk2
2 )kαk2
≤ kpkP (3ℓn
2 +kαk2
ℓn
ℓn
ℓn
ℓn
ℓn
2
2 k(〈h, f j〉) jk2
ℓn
2
Therefore,
T1
kpk1/2
P (3ℓn
2 )
,... ,
Tn
kpk1/2
P (3ℓn
2 )
°°°p(
and this concludes the proof.
)°°°L (H ) ≥ kpk−3/2
P (3ℓn
2 )kp(T1,..., Tn)ekH
= kpk−3/2
P (3ℓn
2 )S ≫ kpkP (3ℓn
2 )
n2
log15/4 n
,
(cid:3)
ASYMPTOTIC ESTIMATES ON THE VON NEUMANN INEQUALITY FOR HOMOGENEOUS POLYNOMIALS
13
Acknowledgements. We wish to thank the referee for her/his comments. Also, we would like to warmly
thank our friend Michael Mackey for his careful reading and suggestions that improved considerably
the final presentation of the paper.
REFERENCES
[1] Noga Alon, Jeong-Han Kim, and Joel Spencer. Nearly perfect matchings in regular simple hypergraphs. Israel J. Math.,
100:171 -- 187, 1997.
[2] Tsuyoshi Andô. On a pair of commutative contractions. Acta Sci. Math. (Szeged), 24:88 -- 90, 1963.
[3] Frédéric Bayart. Maximum modulus of random polynomials. Q. J. Math., 63(1):21 -- 39, 2012.
[4] Jöran Bergh and Jörgen Löfström. Interpolation spaces. An introduction. Springer-Verlag, Berlin, 1976. Grundlehren der
Mathematischen Wissenschaften, No. 223.
[5] Ron Blei. Multidimensional extensions of the Grothendieck inequality and applications. Ark. Mat., 17(1):51 -- 68, 1979.
[6] Ron Blei. Analysis in integer and fractional dimensions, volume 71 of Cambridge Studies in Advanced Mathematics. Cam-
bridge University Press, Cambridge, 2001.
[7] Harold P. Boas. Majorant series. J. Korean Math. Soc., 37(2):321 -- 337, 2000. Several complex variables (Seoul, 1998).
[8] Daniel Carando and Verónica Dimant. Extension of polynomials and John's theorem for symmetric tensor products. Proc.
Amer. Math. Soc., 135(6):1769 -- 1773 (electronic), 2007.
[9] Michael J. Crabb and Alexander M. Davie. von Neumann's inequality for Hilbert space operators. Bull. London Math.
Soc., 7:49 -- 50, 1975.
[10] Andreas Defant and Klaus Floret. Tensor norms and operator ideals. North-Holland Mathematics Studies, 176. North-
Holland Publishing Co., Amsterdam, 1993.
[11] Andreas Defant, Leonhard Frerick, Joaquim Ortega-Cerdà, Myriam Ounaïes, and Kristian Seip. The Bohnenblust-Hille
inequality for homogeneous polynomials is hypercontractive. Ann. of Math. (2), 174(1):485 -- 497, 2011.
[12] Andreas Defant, Domingo García, and Manuel Maestre. Bohr's power series theorem and local Banach space theory. J.
Reine Angew. Math., 557:173 -- 197, 2003.
[13] Andreas Defant, Domingo García, and Manuel Maestre. Maximum moduli of unimodular polynomials. J. Korean Math.
Soc., 41(1):209 -- 229, 2004. Satellite Conference on Infinite Dimensional Function Theory.
[14] Seán Dineen. Complex analysis on infinite dimensional spaces. Springer Monographs in Mathematics. London: Springer,
1999.
[15] Peter G. Dixon. The von Neumann inequality for polynomials of degree greater than two. J. London Math. Soc. (2),
14(2):369 -- 375, 1976.
[16] Paul Erdos and Haim Hanani. On a limit theorem in combinatorial analysis. Publ. Math. Debrecen, 10:10 -- 13, 1963.
[17] Jean-Pierre Kahane. Some random series of functions. D. C. Heath and Co. Raytheon Education Co., Lexington, Mass.,
1968.
[18] Peter Keevash. The existence of designs. Preprint arXiv:1401.3665.
[19] Jeong Han Kim. Nearly optimal partial Steiner systems. In Brazilian Symposium on Graphs, Algorithms and Combina-
torics, volume 7 of Electron. Notes Discrete Math., page 4 pp. (electronic). Elsevier, Amsterdam, 2001.
[20] Michel Ledoux and Michel Talagrand. Probability in Banach spaces, volume 23 of Ergebnisse der Mathematik und ihrer
Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1991. Isoperimetry and processes.
[21] Anna Maria Mantero and Andrew Tonge. Banach algebras and von Neumann's inequality. Proc. London Math. Soc. (3),
38(2):309 -- 334, 1979.
14
DANIEL GALICER, SANTIAGO MURO, PABLO SEVILLA-PERIS
[22] Anna Maria Mantero and Andrew Tonge. The Schur multiplication in tensor algebras. Studia Math., 68(1):1 -- 24, 1980.
[23] Brian Maurizi and Hervé Queffélec. Some remarks on the algebra of bounded Dirichlet series. J. Fourier Anal. Appl.,
16(5):676 -- 692, 2010.
[24] Gilles Pisier. Some applications of the metric entropy condition to harmonic analysis. In Banach spaces, harmonic anal-
ysis, and probability theory (Storrs, Conn., 1980/1981), volume 995 of Lecture Notes in Math., pages 123 -- 154. Springer,
Berlin, 1983.
[25] Gilles Pisier. Similarity problems and completely bounded maps, volume 1618 of Lecture Notes in Mathematics. Springer-
Verlag, Berlin, expanded edition, 2001. Includes the solution to "The Halmos problem".
[26] Vojtech Rödl. On a packing and covering problem. European J. Combin., 6(1):69 -- 78, 1985.
[27] Carsten Schütt. Entropy numbers of diagonal operators between symmetric Banach spaces. J. Approx. Theory, 40(2):121 --
128, 1984.
[28] Béla Sz.-Nagy. Unitary dilations of Hilbert space operators and related topics. American Mathematical Society, Providence,
R.I., 1974. Expository Lectures from the CBMS Regional Conference held at the University of New Hampshire, Durham,
N.H., June 7-11, 1971, Conference Board of the Mathematical Sciences Regional Conference Series in Mathematics, No.
19.
[29] Nicholas Th. Varopoulos. On an inequality of von Neumann and an application of the metric theory of tensor products
to operators theory. J. Functional Analysis, 16:83 -- 100, 1974.
[30] Johann von Neumann. Eine Spektraltheorie für allgemeine Operatoren eines unitären Raumes. Math. Nachr., 4:258 -- 281,
1951.
DEPARTAMENTO DE MATEMÁTICA - PAB I, FACULTAD DE CS. EXACTAS Y NATURALES, UNIVERSIDAD DE
BUENOS AIRES, (1428) BUENOS AIRES, ARGENTINA AND CONICET
E-mail address: [email protected]
E-mail address: [email protected]
INSTITUTO UNIVERSITARIO DE MATEMÁTICA PURA Y APLICADA, UNIVERSITAT POLITÈCNICA DE DE VALÈN-
CIA, CMNO VERA S/N, 46022, VALENCIA, SPAIN
E-mail address: [email protected]
|
1307.5620 | 1 | 1307 | 2013-07-22T08:32:59 | On the spaces of $\lambda-$ convergent and bounded series | [
"math.FA"
] | The main purpose of this study is to introduce the spaces $cs^{\lambda}, cs_0^{\lambda}$ and $bs^{\lambda}$ which are $BK-$spaces of non-absolute type. We prove that these spaces are linearly isomorphic to the spaces $cs, cs_0$ and $bs$, respectively and derive some inclusion relations. Additionally, their Schauder bases have been constructed and the $\alpha-,\beta-$ and $\gamma-$ duals of these spaces have been computed. Finally, we characterize some matrix classes from the spaces $cs^{\lambda}, cs_0^{\lambda}$ and $bs^{\lambda}$ to spaces $\ell_p, c$ and $c_0$, where $1\leq p\leq\infty$ . | math.FA | math |
ON THE SPACES OF λ− CONVERGENT AND BOUNDED
SERIES
MELTEM KAYA AND HASAN FURKAN∗
Abstract. The main purpose of this study is to introduce the spaces csλ, csλ
and bsλ which are BK −spaces of non-absolute type. We prove that these
spaces are linearly isomorphic to the spaces cs, cs0 and bs, respectively and
derive some inclusion relations. Additionally, their Schauder bases have been
constructed and the α−, β− and γ− duals of these spaces have been computed.
Finally, we characterize some matrix classes from the spaces csλ, csλ
0 and bsλ
to spaces ℓp, c and c0, where 1 6 p 6 ∞ .
0
1. Preliminaries, Background and Notation
By a sequence space, we understand a linear subspace of the space w = CN,
where N = {0, 1, 2, . . . }. A sequence spaces E with a linear topology is called a
K− spaces provided each of the maps pi : E → C defined pi(x) = xi is continuous
for all i ∈ N, where C denotes the complex field and N = {0, 1, 2, . . . }. A K−
space is called an F K− space provided E is a complete linear metric space. An
F K− spaces whose topology is normable is called a BK− space (see [1, pages 272-
273]) which contains φ, the set of all finitely non -- zero sequences. We write ℓ∞, c
and c0 for the spaces of all bounded, convergent and null sequences, respectively.
Also by ℓp, we denote the space of all p -- absolutely summable sequences, where
1 6 p < ∞. Moreover, we write bs, cs and cs0 for the sequences spaces of all
bounded, convergent and null series, respectively.
Let µ and ν be two sequence spaces, and let A = (ank) be an infinite matrix
of complex numbers ank, where n, k ∈ N. Then we say that A defines a matrix
transformation from µ into ν, and we denote it by writing A : µ → ν if for every
sequence x = (xk) ∈ µ, the sequence Ax = {(Ax)}, the A- transform of x, is in ν,
where
ankxk , (n ∈ N, x ∈ D00(A)),
(1.1)
(Ax)n :=Xk
and by D00(A) denotes the subspace of w consisting of x ∈ w for which the sum
exists as a finite sum. For simplicity in notation, here and in what follows, the
summation without limits runs from 0 to ∞ and we shall use the convention that
any term with a negative subscript is equal to naught, for example, λ−1 = 0 and
x−1 = 0.
By (µ : ν), we denote the class of all matrices A such that A : µ → ν. Thus
A ∈ (µ : ν) if and only if the series on the right side of (1.1) convergens for each
2000 Mathematics Subject Classification. 40C05,40H05,46A45.
Key words and phrases. λ spaces, matrix domain of a sequence space, α, β and γ duals, matrix
mappings, the series spaces cs and bs.
*Corresponding author.
1
n ∈ N and each x ∈ µ and we have Ax = {(Ax)n}n∈N ∈ ν for all x ∈ µ . For an
arbitrary sequence space µ, the matrix domain µA of an infinite matrix A in µ is
defined by
µA := {x ∈ w : Ax ∈ µ},
(1.2)
which is a sequence space. If A is triangle, then one can easily observe that the
normed sequence spaces µA and µ are norm isomorphic, i.e., µA ∼= µ. If µ is a
sequence space, then the continuous dual µ∗
A of the space µA is defined by
µ∗
A := {f : f = g ◦ A, g ∈ µ∗}.
We denote the collection of all finite subsets of N by F . Also, we will write e(k)
for the sequence whose only non-zero term is 1 in the kth place for each k ∈ N.
Throughout this paper, let λ = (λk) be a strictly increasing sequence of positive
real tending to infinity; that is
0 < λ0 < λ1 < λ2 < ...,
lim
k→∞
λk = ∞.
We define the matrix Λ = (λnk) of weighted mean relative to the sequence λ by
λnk =(cid:26) λk−λk−1
0,
λn
,
0 6 k 6 n,
k > n,
for all k, n ∈ N. With a direct calculation we derive the equality
Λn(x) =
1
λn
n
Xk=0
(λk − λk−1)xk; (n ∈ N).
(1.3)
It is easy to show that the matrix Λ is regular and is reduced, in the special case
λk = k + 1 for all k ∈ N to the matrix C1 of Ces`aro means of order one. Introducing
the concept of Λ− strong convergence, several results on Λ− strong convergence of
numerical sequences and Fourier series were given by M´oricz [2]. Since we have
Qn =
n
Xk=0
qk = λn, rnk =
qk
Qn
=
λk − λk−1
λn
= λnk
in the special case qk = λk − λk−1 for all k ∈ N, the matrix Λ is also reduced to
the Riesz means Rq = (rnk) with respect to the sequence q = (qk).
∞, cλ, cλ
0 and ℓλ
We summarize the knowledge in the existing literature concerning with the
λ- matrices domain over some sequence spaces. Mursaleen and Noman [3, 4, 5, 6]
introduced the spaces ℓλ
p of lambda-bounded, lambda-convergent,
lambda-null and lambda-absolutely p−summable sequences and gave the inclusion
relations between these spaces and the classical sequence spaces ℓ∞, c and c0.
Later, Mursaleen and Noman [7] investigated the difference spaces cλ
0 (∆) and cλ(∆)
0 and cλ. Recently, on paranormed λ− sequence spaces
obtained from the spaces cλ
of non-absolute type has been studied by Karakaya, Noman and Polat [8]. More
recently, Sonmez and Ba¸sar [9] introduce the difference sequence spaces cλ
0 (B) and
0 (∆) and cλ(∆). Quite recently,
cλ(B), which are the generalization of the spaces cλ
on some new sequence spaces of non-absolute type and matrix transformations has
been studied by Ganie and Sheikh [10]. Same authors has been studied on spaces of
λ− convergent sequences and almost convergence [11]. Also, on the fine spectrum
of the operator defined by lambda matrix over the spaces of null and convergent
sequences has been studied by Ye¸silkayagil and Ba¸sar [12].
2
In this work, our purpose is to construct sequence spaces of csλ, csλ
0 and bsλ
by using a matrix domain over a normed space as done by [3].
We define the sequence y = (yk), which will be frequently used, as the Λ-
transform of a sequence x = (xk), i.e., y = Λ(x) and so we have
yk :=
1
λk
k
Xj=0
(λj − λj−1)xj ; (k ∈ N).
(1.4)
Also, we say that a sequence x = (xk) ∈ w is λ-convergent if Λx ∈ c. In particular,
we say that x is λ-null sequence if Λx ∈ c0 and we say that x is λ-bounded if
Λx ∈ ℓ∞.
2. The sequence spaces csλ, csλ
0 and bsλ
In the present section, we introduce the sequence spaces csλ, csλ
0 and bsλ
as the sets of all sequences whose Λ-transforms are in the spaces cs, cs0 and bs,
respectively, that is
and
m
m
csλ
lim
m→∞
csλ = (x = (xk) ∈ w :
0 = (x = (xk) ∈ w :
Xn=0
Xn=0
m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
bsλ = (x = (xk) ∈ w : sup
Xn=0
lim
m→∞
m
1
λn
1
λn
1
λn
n
n
Xk=0
Xk=0
n
Xk=0
(λk − λk−1)xk exists) ,
(λk − λk−1)xk = 0)
(λk − λk−1)xk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
< ∞) .
With the notation of (1.2), we can redefine the spaces csλ, csλ
0 and bsλ as the
csλ = (cs)Λ, csλ
matrix domains of the triangle Λ in the spaces cs, cs0 and bs by
0 = (cs0)Λ and bsλ = (bs)Λ.
Then, it is immediate by (2.1) that the sets csλ, csλ
(2.1)
0 and bsλ are linear spaces
0 and bsλ
with coordinatewise addition and scalar multiplication, that is, csλ, csλ
are the sequence spaces consisting of all sequences which are λ-convergent, λ-null
and λ-bounded series of type λ, respectively.
Now, we may begin with the following theorem which is essential in the text.
Theorem 2.1. The sequence spaces csλ, csλ
norm kxkcsλ = kxkcsλ
= kxkbsλ, that is,
0
0 and bsλ are BK-spaces with the same
kxkbsλ = kΛ(x)kbs = sup
< ∞.
m
Xn=0
m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Λn(x)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Proof. Since (2.1) holds and cs, cs0 and bs are BK- spaces with the respect to their
natural norms and the matrix Λ is a triangle, Theorem 4.3.12 of Wilansky [13, page
63] gives the fact that csλ, csλ
0 and bsλ are BK- spaces with the given norms. This
completes the proof.
(cid:3)
Remark 1. One can easily check that the absolute property does not hold on
the spaces csλ, csλ
0 and bsλ, that is, kxkcsλ 6= kxkcsλ , kxkcsλ
and
0 and bsλ, and this
kxkbsλ 6= kxkbsλ for at least one sequence in the spaces csλ, csλ
6= kxkcsλ
0
0
3
shows that csλ, csλ
x = (xk).
0 and bsλ are the sequence spaces of non-absolute type, where
Now, we give the final theorem of this section.
0 and bsλ of non-absolute type are
Theorem 2.2. The sequence spaces csλ, csλ
isometrically isomorphic to the spaces cs, cs0 and bs, respectively, that is csλ ∼= cs,
csλ
0
∼= cs0 and bsλ ∼= bs.
Proof. To prove this, we should show the existence of an isometric isomorphism
between the spaces csλ
0 and cs0. Consider the transformation T defined, with the
notation of (1.4), from csλ
0 to cs0 by x 7→ y(λ) = T x. Then, T (x) = y = Λ(x) ∈ cs0
for every x ∈ csλ
0 and the linearity of T is clear. Also, it is trivial that x = θ
whenever T x = θ and hence T is injective. Furthermore, let y = (yk) ∈ cs0 be
given and define the sequence x = (xk) by
xk :=
k
Xj=k−1
(−1)k−j
λj
λk − λk−1
yj; (k ∈ N).
(2.2)
Then, by using (1.3) and (2.2), we have for every n ∈ N that
Λn(x) =
=
=
1
λn
1
λn
1
λn
= yn.
n
n
Xk=0
Xk=0
Xk=0
n
(λk − λk−1)xk
(−1)k−j λjyj
k
Xj=k−1
(λkyk − λk−1yk−1)
This shows that Λ(x) = y and since y ∈ cs0, we obtain that Λ(x) ∈ cs0. Thus, we
deduce that x ∈ csλ
0 and T x = y. Hence T is surjective. Moreover, one can easily
see for every x ∈ csλ
0 that
kT xkcs0 = ky(λ)kcs0 = kλ(x)kcs0 = kxkcsλ
0
which means that T is norm preserving. Therefore T is isometry. Consequently T is
an isometric isomorphism which show that the spaces csλ
0 and cs0 are isometrically
isomorphic.
It is clear that if the spaces csλ
0 and cs0 are replaced by the respective one of
the spaces csλ and cs or bsλ and bs, then we obtain the fact that csλ ∼= cs and
bsλ ∼= bs. This completes the proof.
(cid:3)
3. The inclusion Relations
In the present section, we establish some inclusion relations concerning with
the spaces csλ, csλ
0 and bsλ. We may begin with the following lemma.
Lemma 3.1. For any sequence x = (xk) ∈ w, the equality
Sn(x) = xn − Λn(x); (n ∈ N)
(3.1)
4
holds, where S(x) = {Sn(x)} is the sequence defined by
S0(x) = 0 and Sn(x) =
1
λn
n
Xk=1
λk−1(xk − xk−1); (n > 1).
Theorem 3.2. The inclusions csλ
0 ⊂ csλ ⊂ bsλ strictly hold.
0 ⊂ csλ ⊂ bsλ hold.
Proof. It is obvious that the inclusions csλ
Let us consider the sequence x = (xk) defined by
xk =
λk
1
(k+2)2 − λk−1
λk − λk−1
1
(k+1)2
; (k ∈ N).
In the present case, we obtain for every n ∈ N that equality
Λn(x) =
1
λn
n
Xk=0(cid:18)λk
1
(k + 2)2 − λk−1
1
(k + 1)2(cid:19) =
1
(n + 2)2
which shows that Λ(x) ∈ cs\cs0. Thus, the sequence x is in csλ but not in csλ
0 .
Hence csλ
0 ⊂ csλ is a strict inclusion.
To show that csλ ⊂ bsλ inclusion is strict, we define the sequence y = (yk) by
yk = (−1)k(cid:18) λk + λk−1
λk − λk−1(cid:19) ; (k ∈ N).
Then, we have for every n ∈ N that
n
m
m
Xn=0
1
λn
Xn=0
Xk=0
Λn(y) =
(−1)k(λk + λk−1) =
(−1)n.
m
Xn=0
This shows Λ(y) ∈ bs\cs. Thus, the sequence y is in bsλ but not in csλ and hence
csλ ⊂ bsλ is a strict inclusion. This concludes the proof.
(cid:3)
Lemma 3.3. [3, Theorem 4.1.]The inclusions cλ
0 ⊂ cλ ⊂ ℓλ
∞ strictly hold.
Theorem 3.4. The inclusions csλ ⊂ cλ
0 and bsλ ⊂ ℓλ
∞ strictly hold.
Proof. It is clear that the inclusion csλ ⊂ cλ
and hence Λ(x) ∈ c0 which means that x ∈ cλ
defined by
0 holds, since x ∈ csλ implies Λ(x) ∈ cs
0 . Consider the sequence x = (xk)
xk =
1
k + 1
; (k ∈ N).
Then, x ∈ c0 and hence x ∈ cλ
hand, we have for every n ∈ N that
0 , since the inclusion c0 ⊂ cλ
0 holds. On the other
Λn(x) =
>
=
1
λn
n
Xk=0
1
λn(n + 1)
1
n + 1
λk − λk−1
k + 1
(λk − λk−1)
n
Xk=0
which shows that Λ(x) 6∈ cs and hence x 6∈ csλ. Thus, the sequence x is in cλ
not in csλ. Therefore, the inclusion csλ ⊂ cλ
0 is strict.
0 but
5
Similarly, it is also trivial that the inclusion bsλ ⊂ ℓλ
∞ holds. To show that this
inclusion is strict, we define the sequence y = (yk) by
In the present case, we have for every n ∈ N that
yk = e = (1, 1, 1, ...) ; (k ∈ N).
Λn(y) =
1
λn
n
Xk=0
(λk − λk−1) = 1
which shows that Λ(y) ∈ ℓ∞\bs. Thus, the sequence y is in ℓλ
hence bsλ ⊂ ℓλ
∞ is a strict inclusion. This completes the proof.
∞ but not in bsλ and
Theorem 3.5. The inclusion csλ ⊂ cs holds if and only if S(x) ∈ cs for every
sequence x ∈ csλ.
Proof. Suppose that the inclusion csλ ⊂ cs holds, and take any x = (xk) ∈ csλ.
Then Λx ∈ cs and x ∈ cs by the hypothesis. Thus, we deduce from (3.1) that
(cid:3)
m
m
m
(xn − Λn(x)) =
xn −
Λn(x) =
Xn=0
Xn=0
Xn=0
Hence, we obtain from (3.2) by letting m → ∞ that
Sn(x).
(3.2)
m
Xn=0
lim
m
Xn=0
Sn(x) =Xn
xn −Xn
Λn(x).
(3.3)
m
m
As (xn) ∈ cs and (Λn(x)) ∈ cs, the right hand side of the equality (3.3) is convergent
Conversely, let x ∈ csλ be given. Then, we have by the hypothesis that
n=0 Sn(x) converges and so, S(x) ∈ cs.
as m → ∞. Thereby, the series Pm
S(x) ∈ cs. Again, it follows by (3.2) that
lim
m
Xn=0
xn =Xn
Sn(x) +Xn
Λn(x)
which shows that x ∈ cs since Λ(x) ∈ cs and S(x) ∈ cs. Hence, the inclusion
csλ ⊂ cs holds and this concludes the proof.
(cid:3)
Theorem 3.6. The inclusion csλ
sequence x ∈ csλ
0 .
0 ⊂ cs0 holds if and only if S(x) ∈ cs0 for every
Proof. One can see by analogy that the inclusion csλ
if S(x) ∈ cs0 for every sequence x ∈ csλ
0 . This completes the proof.
0 ⊂ cs0 also holds if and only
(cid:3)
Theorem 3.7. The inclusion bsλ ⊂ bs holds if and only if S(x) ∈ bs for every
sequence x ∈ bsλ.
Proof. Suppose that the inclusion bsλ ⊂ bs holds, and take any x = (xk) ∈ bsλ.
Then, x ∈ bs by the hypothesis. Thus, we obtain from equality (3.1)
kS(x)kbs 6 kxkbs + kΛ(x)kbs = kxkbs + kxkbsλ < ∞
which yields that S(x) ∈ bs.
Conversely, assume that S(x) ∈ bs for every x ∈ bsλ. Again, we obtain from
equality (3.1)
kxkbs 6 kS(x)kbs + kΛ(x)kbs = kS(x)kbs + kxkbsλ < ∞.
6
This shows that x ∈ bs. Hence, the inclusion bsλ ⊂ bs holds. This completes the
proof.
(cid:3)
4. The Basis for the Spaces csλ, csλ
0 and bsλ
In the present section, we give a sequence of the points of the spaces csλ and
csλ
0 which forms a basis for these spaces. If a normed sequence space X contains a
sequence (bn) with the property that for every x ∈ X there is a unique sequence of
scalars (αn) such that
lim
n→∞
kx − (α0b0 + α1b1 + . . . + αnbn)k = 0
which has the sum x is then called the expansion of x with respect to (bn) and is
then (bn) is called a Schauder basis (or briefly basis) for X. The series Pk αkbk
written as x =Pk αkbk.
Now, since the transformation T defined from csλ
0 to cs0 in the proof of Theo-
rem 2.2 is an isomorphism. Therefore, we have the following theorem:
Theorem 4.1. Define the sequence e(n)
λn
λ = {(e(n)
{(e(n)
λ )k} =(cid:26) (−1)k−n
0
λk −λk−1
λ )k} for every fixed k ∈ N by
, n 6 k 6 n + 1,
, otherwise,
for all k ∈ N.
Then, we have
a) The sequence (e(0)
every x ∈ csλ or csλ
λ , . . .) is a Schauder basis for the spaces csλ and csλ
λ , e(1)
0 has a unique representation of the form
0 and
b) bsλ has no Schauder basis.
x =
∞
Xn=0
Λn(x)e(n)
λ .
(4.1)
Proof. a) ([13, Theorem 2.3.]) It is clear that e(n)
λ
basis for cs and Λ(e(n)
λ ) = e(n). Let x ∈ csλ be given. Then, y = Λ(x) ∈ cs and
is a basis for csλ since e(n) is a
y[m] =
yne(n) → y (m → ∞)
m
Xn=0
for a unique sequence (yn)∞
n=0 of scalars. Therefore, we obtain that
Since yn = Λn(x), we can write as
Λ−1(y[m]) =
m
Xn=0
ynΛ−1(e(n)) =
yne(n)
λ .
m
Xn=0
m
x[m] =
λn(x)e(n)
λ .
Xn=0
Consequently,
kx[m] −xkcsλ = kΛ(x[m] −x)kcs = kΛ(x[m])−Λ(x)kcs = ky[m] −ykcs → 0 (m → ∞).
Thus, we deduce that limm→∞ kx[m] − xk = 0, which shows that x ∈ csλ is repre-
sented as in (4.1).
7
Finally, let us show the uniqueness of the representation (4.1) of x ∈ csλ. For,
λ .
Since the linear transformation T defined from csλ to cs, in the proof of Theorem
2.2, is continuous, we have
suppose on the contrary that there exists another representation x = Pn αne(n)
Λk(x) =Xn
αnΛk(e(n)
λ ) =Xn
αnδkn = αk; (k ∈ N).
Therefore, the representation (4.1) of x ∈ csλ is unique. It can be proved similarly
for csλ
0 . This completes the proof.
b)As a direct consequence of Remark 2.2. of Malkowsky and Rakocevi´c [15],
(cid:3)
bsλ has no Schauder basis.
As a result, it easily follows from Theorem 2.1 that csλ and csλ
0 are the Banach
spaces with their natural norms. Then by Theorem 4.1 we obtain the the following
corollary:
Corollary 4.2. The sequence spaces csλ and csλ
0 of non-absolute type are separable.
5. The α−, β− and γ− duals of the spaces csλ, csλ
0 and bsλ
In this section, we state and prove the theorems determining the α−, β− and
0 and bsλ of non-absolute type. For arbitrary
γ− duals of the sequence spaces csλ, csλ
sequence spaces X and Y , the set M (X, Y ) defined by
M (X, Y ) = {a = (ak) ∈ w : ax = (akxk) ∈ Y ∀x = (xk) ∈ X}
(5.1)
is called the multiplier space of X and Y . One can easily observe for a sequence
space Z with Y ⊂ Z and Z ⊂ X that the inclusions M (X, Y ) ⊂ M (X, Z) and
M (X, Y ) ⊂ M (Z, Y ) hold, respectively. With the notation of (5.1), the α−, β−
and γ− duals of a sequence space X, which are respectively, denoted by X α, X β
and X γ are defined by
X α = M (X, ℓ1), X β = M (X, cs) and X γ = M (X, bs).
It is clear that X α ⊂ X β ⊂ X γ. Also, it can be obviously seen that the inclusions
X α ⊂ Y α, X β ⊂ Y β, and X γ ⊂ Y γ hold whenever Y ⊂ X.
The following known results [16] are fundamental for this section.
Lemma 5.1. A = (ank) ∈ (cs : ℓ1) if and only if
sup
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
sup
Xn∈N Xk∈K
Xn∈N Xk∈K
(ank − an,k−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(ank − an,k+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
< ∞.
(5.2)
< ∞.
(5.3)
Lemma 5.2. A = (ank) ∈ (cs0 : ℓ1) if and only if
Lemma 5.3. A = (ank) ∈ (bs : ℓ1) if and only if (5.3) holds and
lim
k
ank = 0, ∀n ∈ N.
8
(5.4)
Lemma 5.4. A = (ank) ∈ (cs : c) if and only if
and
ank − an,k+1 < ∞,
sup
n Xk
lim
n
ank exist for all k ∈ N.
Lemma 5.5. A = (ank) ∈ (cs0 : c) if and only if (5.5) holds
lim
n
(ank − an,k+1) exist for all k ∈ N.
Lemma 5.6. A = (ank) ∈ (bs : c) if and only if (5.4) and (5.6) hold and
ank − an,k−1 converges.
Xk
Lemma 5.7. A = (ank) ∈ (cs : ℓ∞) if and only if
ank − an,k−1 < ∞.
sup
n Xk
Lemma 5.8. A = (ank) ∈ (cs0 : ℓ∞) if and only if (5.5) holds.
Lemma 5.9. A = (ank) ∈ (bs : ℓ∞) if and only if (5.4) and (5.5) hold.
Now, we prove the following result.
(5.5)
(5.6)
(5.7)
(5.8)
(5.9)
Theorem 5.10. Define the sets mλ
1 and mλ
2 as follows:
and
mλ
1 =(a = (an) ∈ w :
mλ
2 =(a = (an) ∈ w :
sup
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
sup
(bλ
nk − bλ
Xn∈N Xk∈K
(bλ
nk − bλ
Xn∈N Xk∈K
n,k−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
n,k+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
< ∞)
< ∞) ;
(5.10)
(5.11)
where the matrix Bλ = (bλ
nk) is defined via the sequence a = (an) ∈ w by
bλ
nk =(cid:26) (−1)n−k
0
λk
λn−λn−1 an
if n − 1 6 k 6 n,
if n − 1 > k or k > n,
for all n, k ∈ N. Then {csλ}α = mλ
1 and {csλ
0 }α = {bsλ}α = mλ
2 .
Proof. Let a = (an) ∈ w. Then, by bearing in mind the relations (1.4) and (2.2),
it is immediate that the equality
anxn =
n
Xk=n−1
(−1)n−k
λk
λn − λn−1
anyk = Bλ
n(y)
(5.12)
holds for all n ∈ N. We therefore observe by (5.12) that ax = (anxn) ∈ ℓ1 whenever
x = (xk) ∈ csλ if and only if Bλy ∈ ℓ1 whenever y = (yk) ∈ cs. This means that
the sequence a = (an) ∈ {csλ}α if and only if Bλ ∈ (cs : ℓ1). Hence, we obtain by
Lemma 5.1 with Bλ instead of A that a = (an) ∈ {csλ}α if and only if
(bλ
nk − bλ
< ∞
sup
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn∈N Xk∈K
9
n,k−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
which yields the result that {csλ}α = mλ
1 .
Similarly, we deduce from Lemma 5.3 with (5.12) that a = (an) ∈ {bsλ}α if
and only if Bλ ∈ (bs : ℓ1). Then, it is clear that the columns of the matrix B are
in the space c0, since
for all k ∈ N. Therefore, we derive from (5.3) that
lim
n
bλ
nk = 0
This shows that {csλ
2 . This completes the proof.
Theorem 5.11. Define the sets mλ
3 and mλ
4 as follows :
nk − bλ
< ∞.
n,k+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(5.13)
(cid:3)
∞
(bλ
sup
Xn∈N Xk∈K
0 }α = {bsλ}α = mλ
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
3 = (a = (ak) ∈ w :
Xk=0(cid:12)(cid:12)(cid:12)(cid:12)
k (cid:12)(cid:12)(cid:12)(cid:12)
4 = (cid:26)a = (ak) ∈ w : sup
k→∞(cid:12)(cid:12)(cid:12)(cid:12)
5 = (cid:26)a = (ak) ∈ w :
λk − λk−1(cid:19) =
¯∆(cid:18)
lim
ak
mλ
mλ
mλ
where
λk − λk−1
< ∞) ,
¯∆(cid:18) ¯∆(
λk
λk − λk−1
λk
λk − λk−1
ak
)λk(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
ak(cid:12)(cid:12)(cid:12)(cid:12)
< ∞(cid:27) ,
ak(cid:12)(cid:12)(cid:12)(cid:12)
exists(cid:27) ,
ak+1
λk+1 − λk
ak
−
λk − λk−1
3 ∩ mλ
0 }β = mλ
for all k ∈ N. Then {csλ}β = {csλ
4 and {bsλ}β = mλ
3 ∩ mλ
5 .
Proof. Because of the proof may also be obtained for the space bsλ in the similar
way, we omit it. Take any a = (ak) ∈ w and consider the equation
(5.14)
anyn = T λ
n (y),
where the matrix T λ = (tλ
akxk =
n
Xk=0
ak
yj
λn
λn − λn−1
λk − λk−1
k
n
λj
ak
n−1
(−1)k−j
Xk=0
Xj=k−1
λk − λk−1(cid:19) λkyk +
¯∆(cid:18)
Xk=0
nk =
λk −λk−1(cid:17) λk
nk) is defined by
λn−λn−1
0
¯∆(cid:16)
an
λn
ak
=
tλ
if 0 6 k 6 n − 1,
if k = n,
if k > n.
for all n, k ∈ N. Thus, we deduce by (5.14) that ax = (akxk) ∈ cs where x =
(xk) ∈ csλ if and only if T λ(y) ∈ c whenever y = (yk) ∈ cs. This means that
a = (ak) ∈ {csλ}β if and only if T λ ∈ (cs : c). Therefore, by using Lemma 5.4, we
derive from (5.5) and (5.6) that
< ∞,
∞
Xk=0(cid:12)(cid:12)(cid:12)(cid:12)
¯∆(cid:18) ¯∆(
n (cid:12)(cid:12)(cid:12)(cid:12)
sup
ak
λk − λk−1
λn
λn − λn−1
10
)λk(cid:19)(cid:12)(cid:12)(cid:12)(cid:12)
an(cid:12)(cid:12)(cid:12)(cid:12)
< ∞
and
lim
n
tλ
nk = ¯∆(cid:18)
ak
λk − λk−1(cid:19) λk,
respectively. Thereby, we conclude that {csλ}β = {csλ
0 }β = mλ
3 ∩ mλ
4 .
(cid:3)
Theorem 5.12. The γ-dual of the space csλ, csλ
0 and bsλ is the set mλ
3 ∩ mλ
4 .
Proof. The proof of this result follows the same lines that in the proof of Theorem
5.11 using Lemma 5.7, Lemma 5.8 and Lemma 5.9 instead of Lemma 5.4.
(cid:3)
6. Certain matrix mappings on the spaces csλ, csλ
0 and bsλ
In this present section, we characterize the matrix classes (csλ : ℓp), (csλ
0 : ℓp),
0 : c) and (bsλ : c), where
(bsλ : ℓp), (csλ : c0), (csλ
1 6 p 6 ∞.
0 : c0), (bsλ : c0), (csλ : c), (csλ
For an infinite matrix A = (ank), we write for brevity that
ank = ¯∆(cid:18)
ank
λk − λk−1(cid:19) λk =(cid:18)
ank
λk − λk−1
−
an,k+1
λk+1 − λk(cid:19) λk (n, k ∈ N).
The following lemmas will be needed in proving our results.
Lemma 6.1. A = (ank) ∈ (cs : c0) if and only if (5.5) holds and
lim
n
ank = 0 (∀k ∈ N).
Lemma 6.2. A = (ank) ∈ (cs0 : c0) if and only if (5.5) holds and
lim
n
(ank − an,k+1) = 0 (∀k ∈ N).
Lemma 6.3. A = (ank) ∈ (bs : c0) if and only if (5.4) holds and
Lemma 6.4. A = (ank) ∈ (cs0 : ℓp) if and only if
lim
ank − an,k+1 = 0.
n Xk
(ank − an,k+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(ank − an,k−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
sup
k Xn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk∈K
k Xn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk∈K
sup
p
< ∞ (1 < p < ∞).
(6.4)
p
< ∞ (1 < p < ∞).
(6.5)
(6.1)
(6.2)
(6.3)
(6.6)
(6.7)
Lemma 6.5. A = (ank) ∈ (cs : ℓp) if and only if
Lemma 6.6. A = (ank) ∈ (bs : ℓp) if and only if (5.4) and (6.4) hold.
Now, we give the following results on the matrix transformations.
Theorem 6.7. (i) A = (ank) ∈ (csλ : ℓ∞) if and only if
∞
and
sup
ank − ak−1 < ∞
n
Xk=0
k (cid:12)(cid:12)(cid:12)(cid:12)
sup
λk
λk − λk−1
< ∞.
ank(cid:12)(cid:12)(cid:12)(cid:12)
11
(ii) A = (ank) ∈ (csλ
0 : ℓ∞) if and only if (6.7) holds and
(iii) A = (ank) ∈ (bsλ : ℓ∞) if and only if (6.8) holds and
∞
n
sup
Xk=0
k→∞(cid:12)(cid:12)(cid:12)(cid:12)
lim
ank − ak+1 < ∞.
λk
λk − λk−1
exists,
ank(cid:12)(cid:12)(cid:12)(cid:12)
lim
k
ank = 0
and
an < ∞.
sup
n
(6.8)
(6.9)
(6.10)
(6.11)
Proof. Suppose that conditions (6.6) and (6.7) hold and take any x = (xk) ∈ csλ.
k=0 ∈ (csλ)β for all n ∈ N and this
Then, we have by Theorem 5.11 that (ank)∞
implies the existence of the A-transform of x, i.e.; Ax exists. Further, it is clear
that the associated sequence y = (yk) is in the cs and hence y ∈ c0.
Let us now consider the following equality derived by using the relation (1.4)
from the mth partial sum of the series Pk ankxk :
m−1
m
λm
ankxk =
ankyk +
λm − λm−1
Xk=0
anmym, (∀n, m ∈ N).
(6.12)
Xk=0
Xk
Xk=0
Therefore, by using (6.6) and (6.7), from (6.12) as m → ∞ we obtain that equality
ankxk =Xk
ankyk for all n ∈ N.
(6.13)
Further, since the matrix A = (ank) is in the class (cs : ℓ∞) by Lemma 5.7 and
(6.6); we have Ay ∈ ℓ∞. Therefore, we deduce from (1.1) and (6.13) that Ax ∈ ℓ∞
and hence A ∈ (csλ, ℓ∞).
Conversely, suppose that A ∈ (csλ, ℓ∞). Then (ank)∞
k=0 ∈ (csλ)β for all n ∈ N
and this, with Theorem 5.11, implies both (6.7) and
∞
ank − ak+1 < ∞ for all n ∈ N
which together imply that relation (6.13) holds for all sequences x ∈ csλ and y ∈ cs.
Further, since Ax ∈ ℓ∞ by the hipothesis; we obtain by (6.13) that Ay ∈ ℓ∞ which
shows that A ∈ (cs : ℓ∞), where A = (ank). Hence, the necessity of (6.6) is
immediate by (5.7). This concludes the proof of part (i).
Since part (ii) can be proved similarly, we omit its proof.
(cid:3)
Corollary 6.8. (i) A = (ank) ∈ (csλ : c) if and only if (6.7), and (6.8) hold and
lim
n→∞
ank exists.
(ii) A = (ank) ∈ (csλ
0 : c) if and only if (6.7) and (6.8) hold and
lim
n
(ank − an,k+1) exists.
(iii) A = (ank) ∈ (bsλ : c) if and only if (6.9), (6.10) and (6.14)hold and
lim
n→∞Xk
ank − ak−1 exists,
12
(6.14)
(6.15)
(6.16)
Corollary 6.9. (i) A = (ank) ∈ (csλ : c0) if and only if (6.7) and (6.8) hold and
lim
n
an exist.
(6.17)
lim
n
ank = 0.
(ii) A = (ank) ∈ (csλ
0 : c0) if and only if (6.7) and (6.8) hold and
lim
n
(ank − an,k+1) = 0.
(iii) A = (ank) ∈ (bsλ : c0) if and only if (6.9) and (6.10) hold and
ank − an,k+1 = 0,
lim
n Xk
lim
n
an = 0.
Corollary 6.10. (i) A = (ank) ∈ (csλ : ℓ1) if and only if (6.7) holds and
(ii) A = (ank) ∈ (csλ
0 : ℓ1) if and only if (6.7) and (6.22) hold and
< ∞.
(6.23)
< ∞.
(6.24)
(iii) A = (ank) ∈ (bsλ : ℓ1) if and only if (6.9), (6.10), (6.22), and (6.24) hold and
ank − an,k+1 < ∞,
sup
Xk
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn∈N Xk∈K
N,K∈F(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn∈N Xk∈K
sup
(ank − an,k−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(ank − an,k+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
p
p
an < ∞.
Xn
(ank − an,k−1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(ank − an,k+1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xn
anp < ∞.
sup
K∈FXn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk∈K
K∈FXn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk∈K
sup
(6.18)
(6.19)
(6.20)
(6.21)
(6.22)
(6.25)
(6.28)
Corollary 6.11. (i) A = (ank) ∈ (csλ : ℓp) if and only if (6.7) and (6.22) hold
and
(ii) A = (ank) ∈ (csλ
0 : ℓp) if and only if (6.7) and (6.22) hold and
< ∞.
(6.26)
< ∞.
(6.27)
(iii) A = (ank) ∈ (bsλ : ℓp) if and only if (6.9), (6.10), (6.22) and (6.27) hold and
Since Corollary 6.8, Corollary 6.9, Corollary 6.10 and Corollay 6.11 can be
proved similarly with Theorem 6.7, we omit their proofs.
13
7. Conclusion
∞, cRt = rt
c and (c0)Rt = rt
0 in [20], (ℓp)Rt = rt
0 and cEr = er
c in [22], (ℓp)Er = er
p and (ℓ∞)Er = er
c in [24], [(c0)(u, p)]Ar = ar
In the literature, the approach of constructing a new sequence space by means
of the matrix domain of a particular limitation method has recently been employed
by several authors, for example, [17 − 35]. They introduced the sequence spaces
(ℓ∞)Nq and (cNq in [17], (ℓp)C1 = Xp and (ℓ∞)C1 = X∞ in [18], µG = Z(u, ν; µ)
in [19], (ℓ∞)Rt = rt
p in [21],
∞ in [23], (c0)Ar = ar
(c0)Er = er
0
and cAr = ar
c(u, p) in [25],
er
0(∆ : p) = (c0(p))Er ∆, er(∆ : p) = (c(p))Er ∆ and er
∞(∆ : p) = (ℓ∞(p))Er ∆
in [26], (ℓp)Ar = ar
∞ in [27], (c0)C1 = c0 and cC1 = c in [28],
(f0)C1 = f0 and fC1 = f in [29], (f0)Rt = f0 and fRt = f in [30], {ℓ∞}B(r,s) = ℓ∞,
{cB(r,s) = c, {c0}B(r,s) = c0 and {ℓp}B(r,s) = ℓp in [31], fB(r,s) = f , {f0}B(r,s) = f0
in [32], νB(r,s,t) = ν(B) in [33], and fB(r,s,t) = f (B) in [34]; where Nq, C1, Rt, and
Er denote Norlund, Ces`aro, Riesz, and Euler means, respectively. Ar, G, B(r, s)
and B(r, s, t) are , respectively, defined in [24, 19, 31, 33],µ ∈ {c0, c, ℓp}, ν ∈
{ℓ∞, c0, c, ℓp} and 1 6 p < ∞. Also co(u, p) and c(u, p) denote sequence spaces
genarated from the Maddox's spaces c0(p) and c(p) by Ba¸sarır [35].
0(u, p) and [c(u, p)]Ar = ar
p and (ℓ∞)Ar = ar
0 and ℓλ
Quite recently, Mursaleen and Noman have introduced the sequence spaces
∞, cλ, cλ
ℓλ
p in the case 1 6 p 6 ∞ in [3, 5], respectively. Although the matrix
Λ is used for obtaining some new sequence spaces by its domain from the classical
sequence spaces, the triangle matrix Λ over the sequence spaces cs, cs0 and bs is
not studied. So, working the domain of Λ matrix in the spaces cs, cs0 and bs is
meaningful, which is filling up a gap in the existing literature.
Finally, the domain of difference matrix in the cs, cs0, bs and csλ, csλ
0 , bsλ was
not studied. We conclude our work by expressing from now on that the aim of our
next papers is to investigate the difference spaces csλ(∆), csλ
0 (∆) and bsλ(∆).
References
[1] B. Choudhary, S. Nanda, Functional Analysis with Applications, John Wiley & Sons, New
Delhi, India 1989.
[2] F. M´oricz, On Λ -- strong convergence of numerical sequences and Fourier series, Acta Math-
ematica Hungarica, 54(3-4)(1989), 319 -- 327.
[3] M. Mursaleen, A.K. Noman, On the spaces of λ-convergent and bounded sequences, Thai J.
Math. 8(2)(2010), 311 -- 329.
[4] M. Mursaleen, A.K. Noman, Applications of the Hausdorff measure of noncompactness in
some sequence spaces of weighted means, Comput. Math. Appl. 60(5)(2010), 1245 -- 1258.
[5] M. Mursaleen, A. K. Noman, On some new sequence spaces of non-absolute type related to
the spaces ℓp and ℓ∞ I, Faculty of sciences and Math. Univ. of Nis 25(2)(2011), 33 -- 51.
[6] M. Mursaleen, A.K. Noman, On some new sequence spaces of non-absolute type related to
the spaces ℓp and ℓ∞ II, Math. Commun. 16(2011), 383 -- 398.
[7] M. Mursaleen, A.K. Noman, On some new difference sequence spaces of non-absolute type,
Math. Comput. Modelling 52(3-4)(2010), 603 -- 617.
[8] V. Karakaya, A.K. Noman, H. Polat, On paranormed λ− sequence spaces of non-absolute
type, Math. Comput. Modelling 54(2011), 1473 -- 1480.
[9] A. Sonmez, F. Ba¸sar, Generalized difference spaces of non-absolute type of convergent
and null sequences, Abstract and Applied Analysis 2012: Article ID 435076,20pages,
doi:10.1155/2012/435076.
[10] Ab.H. Ganie, N. A. Sheikh, On some new sequence spaces of non-absolute type and matrix
transformations, J. Egyptian Math. Society, 21(2013), 108 -- 114.
[11] N. A. Sheikh, Ab.H. Ganie, On spaces of λ− convergent sequences and almost convergence,
Thai J. Math., 11(2)(2013), 393-398.
14
[12] M. Ye¸silkayagil, F. Ba¸sar, On the fine spectrum of the operator defined by lambda matrix over
the spaces of null and convergent sequences, Abstract and Applied Analysis 2013: Article
ID687393,13pages,http://dx.doi.org/10.1155/2013/687393 .
[13] A. Wilansky, Summability Throught Functional Analysis, vol. 85 of North-Holland Mathe-
matics Studies, North-Holland, Oxford, UK, 1984.
[14] A. M.Jarrah, E. Malkowsky, Ordinary, absolute and strong summability and matrix trans-
formations, FILOMAT 17(2003), 59 -- 78.
[15] E. Malkowsky, V. Rakocevi´c, On matrix domains of triangles, App. Math. and Comp.
189(2007), 1146 -- 1168.
[16] M. Stieglitz, H. Tietz, Matrixtransformationen von Folgenraumen Eine Ergebnisubersict,
Math. Z., 154(1977), 1 -- 16.
[17] C. S. Wang, On Norlund sequence spaces, Tamkang J. Math. 9(1978), 269 -- 274.
[18] P. N. Ng, P. Y. Lee, Ces`a ro sequence spaces of non-absolute type, Comment. Math. Prace
Mat. 20(2)(1978), 429 -- 433.
[19] E. Malkowsky, E. Sava¸s, Matrix transformations between sequence spaces of generalized
weighted means, Appl. Math. Comput. 147(2004), 333 -- 345 .
[20] E. Malkowsky, Recent results in the theory of matrix transformations in sequence spaces,
mat. Vesnik, 49(1997), 187 -- 196.
[21] B. Altay, F. Ba¸sar,On the paranormed Riesz sequence spaces of non-absolute type, Southeast
Asian Bull. Math., 26(5)(2003), 701 -- 715.
[22] B. Altay, F. Ba¸sar,Some Euler sequence spaces of non-absolute type, Ukrainian Math. j.,
57(1)(2005), 1 -- 17.
[23] B. Altay, F. Ba¸sar, M. Mursaleen,On the Euler sequence spaces which include the spaces ℓp
and ℓ∞ I, Information Sciences, 176(10)(2006), 1450 -- 1462.
[24] C. Aydın, F. Ba¸sar,On the new sequence spaces which include the spaces c0 and c , Hokkaido
math. J.,33(2)(2004), 383 -- 398.
[25] V. Karakaya, H. Polat,Some new paranormed sequence spaces defined by Euler and difference
operators , Acta Sci. Math. (Szeged) ,76(1-4)(2010), 877 -- 100.
[26] C. Aydın, F. Ba¸sar,Some new paranormed sequence spaces , Information Science,160(1-
4)(2004), 27 -- 40.
[27] C. Aydın, F. Ba¸sar,Some new paranormed sequence spaces which include the spaces ℓp and
ℓ∞ , Demonstratio Mathematica,38(3)(2005), 641 -- 656.
[28] M. S¸engonul, F. Ba¸sar,Some new Ces`aro sequence spaces of non-absolute type which include
the spaces c0 and c , Soochow j. Math.,31(1)(2005), 107 -- 119.
[29] K.Kayaduman, M. S¸engonul, The spaces of Ces`aro almost convergent sequences and core
theorems, Acta Math. Sci., Ser. B, Engl. Ed. 32(6)(2012), 2265 -- 2278.
[30] M. S¸engonul, K.Kayaduman, On the Riesz almost convergent sequence spaces , Abstract and
Applied Analysis 2012: Article ID691694,18pages,http://dx.doi.org/10.1155/2012/691694 .
[31] M. Kiri¸s¸ci, F. Ba¸sar, Some new sequence spaces derived by the domain of genaralized differ-
ence matrix , Comput. Math. Appl.(2011) 60, 1299 -- 1309.
[32] F. Ba¸sar, M. Kiri¸s¸ci , Almost convergence and generalized difference matrix, Comput. Math.
Appl. 61(3)(2011), 602 -- 611.
[33] A. Sonmez, Some new sequence spaces derived by the domain of the triple band matrix,
Comput. Mat. Appl. 62(2)(2011), 641 -- 650.
[34] A. Sonmez,A Almost convergence and triple band matrix, Math. Comput. Model. 57(2013),
2393 -- 2402 .
[35] M. Ba¸sarır, On some new sequence spaces and related matrix transformation, Indian J. Pure
and App.Math., 26(10)(1995), 1003 -- 1010 .
(M. Kaya) Kahramanmaras¸ Sutc¸u Imam Universitesi, Fen Bilimleri Enstitusu , 46100 --
Kahramanmaras¸, Turkiye
E-mail address, M. Kaya: [email protected]
(H. Furkan) Kahramanmaras¸ Sutc¸u Imam Universitesi, Fen Edebiyat Fakultesi, 46100 --
Kahra-manmaras¸, Turkiye
E-mail address, H. Furkan: [email protected], [email protected]
15
|
1801.04469 | 1 | 1801 | 2018-01-13T17:48:13 | On the equivalence of the Mizoguchi-Takahashi locally contractive map to Nadler's locally contractive map | [
"math.FA"
] | In this article, we have proved the equivalence between the Mizoguchi-Takahashi uniformly~locally~contractive map to the multi-valued map satisfying the Nadler contractive condition uniformly~locally~on a metrically convex space. | math.FA | math | On the equivalence of the Mizoguchi-Takahashi locally
contractive map to Nadler's locally contractive map ‡
Asrifa Sultana ∗
Xiaolong Qin †
Abstract
In this article we have proved the equivalence between the Mizoguchi-Takahashi uniformly lo-
cally contractive map to the multi-valued map satisfying the Nadler contractive condition uni-
formly locally on a metrically convex space.
∗Department of Mathematics, Indian Institute of Technology Bhilai, Chhattisgarh 492015, India
†Institute for Fundamental and Frontier Science, University of Electronic Science and Technology of China, China
‡This article is an initial draft of the complete paper
8
1
0
2
n
a
J
3
1
]
.
A
F
h
t
a
m
[
1
v
9
6
4
4
0
.
1
0
8
1
:
v
i
X
r
a
1
1
Introduction
In 1969, Nadler [6] established a multi-valued extension of the famous Banach contraction principle.
Let (X, d) be a complete metric space and CB(X) denote the family of all non-empty closed and
bounded subsets of X. Let H be the Hausdorff metric with respect to d on CB(X). Nadler [6]
proved that any multi-valued map F from X to CB(X) has a fixed point if there exists some
k ∈ [0, 1) such that H(F (x), F (y)) ≤ kd(x, y) for all x, y ∈ X. An interesting extension of this
result was obtained in 1989, by Mizoguchi and Takahashi [5].
Theorem 1.1 [5] Let (X, d) be a complete metric space and let F : X → CB(X) be a map such
that
H(F (x), F (y)) ≤ α(d(x, y))d(x, y) ∀ x, y ∈ X, x 6= y,
(1)
where α : (0, ∞) → [0, 1) is such that lim sups→t+ α(s) < 1 for every t ∈ [0, ∞). Then F has a fixed
point.
It is worth to note that any Mizoguchi-Takahashi contraction on a metric space need not be a
Nadler contraction (for example, one can refer [7, 1]).
On the other hand, Edelstein [2] introduced the notion of uniformly locally contractive mapping on
a metric space. For a metric space (X, d), a mapping f : X → X is called an (ε, k)-uniformly lo-
cally contractive (where ε > 0 and k ∈ [0, 1)) if d(f x, f y) ≤ kd(x, y) for all x, y ∈ X with d(x, y) < ε.
It is worth to note that an (ε, k)-uniformly locally contractive need not be a Banach contraction
(for example, one can refer [2, 4]). Edelstein [2] established the following extension of Banach
contraction principle.
Theorem 1.2 [2] Let (X, d) be a complete metric space. An (ε, k)-uniformly locally contractive
map f : X → X has a unique fixed point if (X, d) is ε-chainable, that is, for any given a, b ∈ X,
there exist N ∈ N and a sequence (yi)N
i=0 in X such that y0 = a, yN = b and d(yi−1, yi) < ε for
1 ≤ i ≤ N .
In [6], Nadler extended this result to multi-valued mappings. Let (X, d) be a metric space and H be
the Hausdorff metric with respect to d on the family CB(X) of all non empty closed and bounded
subsets of X. Nadler [6] generalized the above result by deriving the following theorem.
Theorem 1.3 [6] Let (X, d) be a complete metric space and F : X → CB(X) be a mapping such
that for all x, y ∈ X with d(x, y) < ε,
H(F (x), F (y)) ≤ kd(x, y)
for some k ∈ [0, 1).
(2)
Then the map F has a fixed point if (X, d) is ε-chainable.
Recently, Eldred et al. [1] explored some spaces on which the Mizoguchi-Takahashi contraction is
equivalent to Nadler contraction. They have proved that Mizoguchi-Takahashi's condition reduces
to a multi-valued contraction in a metrically convex space. Also, they have derived the equivalence
in a compact metric space.
2
In this paper, we have shown that the multi-valued map satisfying the Mizoguchi-Takahashi's
contractive condition uniformly locally is equivalent to a uniformly locally contractive multi-valued
map due to Nadler [6] on a metrically convex space.
2 Preliminaries
In this section we give some definitions and notations which are useful and related to context of
our results.
Let (X, d) be a metric space and p, q be any two arbitrary points in X. A point r ∈ X is called
metrically between p and q if d(p, q) = d(p, r) + d(r, q) where p 6= q 6= r. We say the metric space
(X, d) metrically convex if for any two arbitrary points p and q, there is a point r in X which is
metrically between p and q.
A subset M of a metric space (X, d) is said to be a metric segment with joining points p, q ∈ M if
there is a closed and bounded interval [a, b] ⊆ R and an isometry φ from the set [a, b] onto the set
M such that φ(a) = p and φ(b) = q.
Now, we recall the following result due to Khamsi and Kirk [3] which will be used in our main
result.
Theorem 2.1 Let (X, d) be complete metrically convex metric space. Then any two arbitrary
points in X are the joining points of at least one metric segment in X.
For the given metric space (X, d), the notation CB(X) denotes the family of all non empty closed
and bounded subsets of X. For A, B ∈ CB(X), let
H(A, B) = max(cid:26)sup
a∈A
d(a, B), sup
b∈B
d(b, A)(cid:27) ,
where d(a, B) = infb∈Bd(a, b). Then the map H is a metric on CB(X) which is called the
Hausdorf f metric induced by d.
3 Main results
The following result is the main result proved by us in this article.
Theorem 3.1 Let (X, d) be a metrically convex complete metric space and F : X → CB(X) be
such that ∀x, y ∈ X with d(x, y) < ε (where ε > 0),
H(F (x), F (y)) ≤ α(d(x, y))d(x, y),
(3)
where α : [0, ∞) → [0, 1) is such that lim sups→t+ α(s) < 1 for every t ∈ [0, ∞). Then F is an
(ε, k)-uniformly locally contractive multi-valued mapping (for some k ∈ [0, 1)).
3
Proof. Our aim here is to prove that F is an (ε, k)-uniformly locally contractive multi-valued
mapping, that is, there exists k ∈ [0, 1) such that H(F (x), F (y)) ≤ kd(x, y) ∀x, y ∈ X with d(x, y) <
ε. Let us consider a subset P of real numbers where
P = {p > 0 : sup{α(d(x, y)) : 0 ≤ d(x, y) ≤ p} = 1}.
Then the following two cases occurs. Our aim is to show that F becomes an (ε, k)-uniformly lo-
cally contractive mapping in each case. Let x1, x2 be two arbitrary elements in X such that
d(x1, x2) < ε.
Case 1: P = ∅.
Therefore sup{α(d(x, y)) : 0 ≤ d(x, y) ≤ p} < 1 for any p > 0. Let q be a fixed positive real number.
Let us suppose that sup{α(d(x, y)) : 0 ≤ d(x, y) ≤ q} = k. Clearly, k < 1. Hence, it follows from
equation (3) that if x and y be two elements in X with d(x, y) < min{ε, q}, then
H(F (x), F (y)) ≤ α(d(x, y))d(x, y) ≤ kd(x, y).
(4)
If q ≥ ε, then min{ε, q} = ε. Hence d(x1, x2) < min{ε, q}. By (4), we have
H(F (x1), F (x2)) ≤ α(d(x1, x2))d(x1, x2) ≤ kd(x1, x2).
Suppose that q < ε. Then min{ε, q} = q. Since X is metrically convex metric space, we can find
a, b ∈ R and an isometry φ : [a, b] → X such that φ(a) = x1 and φ(b) = x2. For some r with
0 < r < q, there exists a positive integer m such that
d(x1, x2) = d(φ(a), φ(b)) = d(a, b)
= d(a, a + r) + d(a + r, a + 2r) + · · · + d(a + mr, b)
= d(x1, φ(a + r)) + d(φ(a + r), φ(a + 2r)) + · · · + d(φ(a + mr), x2),
where m is such that a + mr < b < a + (m + 1)r. Now, d(x1, φ(a + r)) = d(a, a + r) = r < min{ε, q}
and hence we have from equation (4) that
H(F (x1), F (φ(a + r)))) ≤ kd(x1, φ(a + r))).
Similarly, for any natural number n < m, we have d(φ(a + nr), φ(a + (n + 1)r)) < min{ε, q} and
hence by (4), we get
H(F (φ(a + nr)), F (φ(a + (n + 1)r)))) ≤ kd(φ(a + nr), φ(a + (n + 1)r))).
Moreover, d(φ(a + mr), x2) < r < min{ε, q} and thus by using (4) we have
H(F (φ(a + mr)), F (x2)) ≤ kd(φ(a + mr), x2).
Thus for any x1, x2 ∈ X with d(x1, x2) < ε, we have
H(F (x1), F (x2)) ≤ H(F (x1), F (φ(a + r))) + · · · + H(F (φ(a + mr)), F (x2))
≤ k[d(x1, φ(a + r)) + d(φ(a + r), φ(a + 2r)) + · · · + d(φ(a + mr), x2)]
= kd(x1, x2).
4
Case 2: P 6= ∅.
Let p0 = inf P . If p0 = 0, then we can find a sequence {pn}n of P such that pn → 0 as n → ∞.
Since pn ∈ P for all n, sup{α(d(x, y)) : 0 ≤ d(x, y) ≤ pn} = 1 for all n ∈ N. By the properties of
supremum, there exists a sequence {(xn, yn)}n such that d(xn, yn) ≤ pn and α(d(xn, yn)) > 1 − 1
n .
Hence 1 − 1
n ≤ α(d(xn, yn)) ≤ 1 for all n ∈ N. Thus d(xn, yn) → 0 and α(d(xn, yn)) → 1 as n → ∞.
This is a contradiction as α : [0, ∞) → [0, 1) with lim sups→t+ α(s) < 1 ∀t ∈ [0, ∞). Therefore
p0 6= 0. Thus for any 0 < q0 < p0, sup{α(d(x, y)) : 0 ≤ d(x, y) ≤ q0} = k0 < 1. Hence, it follows
from equation (3) that for any two elements x and y in X with d(x, y) < min{ε, q0}
H(F (x), F (y)) ≤ α(d(x, y))d(x, y) ≤ k0d(x, y).
(5)
If q0 ≥ ε, then d(x1, x2) < ε = min{ε, q0}. Hence from equation (5)
H(F (x1), F (x2)) ≤ α(d(x1, x2))d(x1, x2) ≤ k0d(x1, x2).
Let us assume that q0 < ε. Since X is metrically convex metric space, we can find a0, b0 ∈ R and
an isometry φ0 : [a0, b0] → X such that φ0(a0) = x1 and φ0(b0) = x2. Similar to case 1, it follows
that for some r0 with q0 > r0 > 0, there exists a positive integer m0 such that
d(x1, x2) = d(x1, φ0(a0 + r0)) + · · · + d(φ0(a0 + m0r0), x2),
where m0 is such that a0 + m0r0 < b0 < a0 + (m0 + 1)r0. Moreover, for any non-negative integer
0 ≤ n0 < m0, we have similar to Case 1 that d(φ0(a0 + n0r0), φ0(a0 + (n0 + 1)r0)) < min{ε, q0} and
hence by (5)
H(F (φ0(a0 + n0r0)), F (φ0(a0 + (n0 + 1)r0)))) ≤ k0d(φ0(a0 + n0r0), φ0(a0 + (n0 + 1)r0))).
Thus for any x1, x2 ∈ X with d(x1, x2) < ε, we have
H (F (x1), F (x2)) ≤ H(F (x1), F (φ0(a0 + r0))) + · · · + H(F (φ0(a0 + m0r0)), F (x2))
≤ k0 [d(x1, φ0(a0 + r0)) + · · · + d(φ0(a0 + m0r0), x2)] = k0d(x1, x2).
Therefore F is an (ε, k)-uniformly locally contractive multi-valued mapping in both cases. Hence
the proof is complete.
✷
References
[1] A. A. Eldred, J. Anuradha, P. Veeramani, On the equivalence of the Mizoguchi-Takahashi fixed
point theorem to Nadler's theorem, Applied Mathematics Letters 22 (2009) 1539-1542.
[2] M. Edelstein, An extension of BanachSs contraction principle, Proc. Amer. Math. Soc. 12 (1961)
7-10.
[3] M.A. Khamsi, W.A. Kirk, An Introduction to Metric Spaces and Fixed Point Theory, John
Wiley and Sons, Inc., 2001.
[4] L. M´at´e, The Hutchinson UBarnsley theory for certain non-contraction mappings, Period. Math.
Hungar. 27 (1993) 21-33.
5
[5] N. Mizoguchi, W. Takahashi, Fixed point theorems for multivalued mappings on complete metric
space, J. Math. Anal. Appl. 141 (1989) 177-188.
[6] S.B. Nadler Jr., Multi-valued contraction mappings, Pacific J. Math. 30 (1969) 475-488.
[7] T. Suzuki, Mizoguchi-Takahashi's fixed point theorem is a real generalization of Nadler's, J.
Math. Anal. Appl. 340 (2008) 752-755.
6
|
1804.07288 | 4 | 1804 | 2018-10-03T15:24:16 | On the Invertibility of the Sum of Operators | [
"math.FA",
"math.OA"
] | The primary purpose of this paper is to investigate the question of invertibility of the sum of operators. The setting is bounded and unbounded linear operators. Some interesting examples and consequences are given. As an illustrative point, we characterize invertibility for the class of normal operators. Also, we give a very short proof of the self-adjointness of a normal operator when the latter has a real spectrum. | math.FA | math |
ON THE INVERTIBILITY OF THE SUM OF OPERATORS
MOHAMMED HICHEM MORTAD
Abstract. The primary purpose of this paper is to investigate the question
of invertibility of the sum of operators. The setting is bounded and unbounded
linear operators. Some interesting examples and consequences are given. As an
illustrative point, we characterize invertibility for the class of normal operators.
Also, we give a very short proof of the self-adjointness of a normal operator
when the latter has a real spectrum.
Let H be a complex Hilbert space and let B(H) denote the algebra of all bounded
1. Introduction
linear operators on H. An A ∈ B(H) is called positive (symbolically A ≥ 0) if
hAx, xi ≥ 0 ∀x ∈ H.
By a square root of A ∈ B(H), we mean a B ∈ B(H) such that B2 = A. If A ≥ 0,
then there is one and only one B ≥ 0 such that B2 = A. This positive B is denoted
by √A.
Recall that any T ∈ B(H) is expressible as T = A + iB where A, B ∈ B(H) are
self-adjoint. Besides,
A = ReT =
T + T ∗
2
and B = ImT =
T − T ∗
2i
.
It is readily 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 [8] for proofs).
Theorem 1.1. Let A, B ∈ B(H). Then
0 ≤ A ≤ B =⇒
√A ≤
√B.
Lemma 1.2. Let A, B ∈ B(H) be such that AB = BA and A, B ≥ 0. Then
√A + B ≤ √A + √B.
Theorem 1.1 is known to hold for α ∈ (0, 1) instead of 1
Hence we may easily establish the analogue of Lemma 1.2 for nth roots.
Lemma 1.3. Let A, B ∈ B(H) be such that AB = BA and A, B ≥ 0. If n ∈ N,
then (A + B)
2 (the Heinz Inequality).
n + B
1
1
n .
1
n ≤ A
Let us say a few more words about the absolute value of an operator (that is,
A = √A∗A with A ∈ B(H)). It is well known that the properties of the absolute
value of complex numbers cannot all just be carried over to B(H) (even for self-
adjoint operators). This applies for example to the multiplicativity property and to
2010 Mathematics Subject Classification. Primary 47A05, Secondary 47B65, 47B25, 47A10.
Key words and phrases. Invertibility. Absolute value. Normal, self-adjoint and positive operators.
Square roots. Cartesian decomposition. Bounded and unbounded operators.
1
2
M. H. MORTAD
triangle inequalities. For counterexamples, readers may wish to consult [8]. See also
[9] to see when these results hold. The similar question on unbounded operators
may be found in the recent work [2]. Some results, however, do hold without any
special assumption. One of them is the following simple result.
Proposition 1.4. Let A, B ∈ B(H). Then
A + B2 ≤ 2A2 + 2B2.
The following known result is also primordial.
Proposition 1.5. (see [5] for a new proof ) Let A, B ∈ B(H). If 0 ≤ A ≤ B and
if A is invertible, then B is invertible and B−1 ≤ A−1.
We digress a little bit to notice a simple proof of the positiveness of the spectrum
of a positive operator using the previous proposition: If λ < 0, then −λI > 0 and so
A − λI ≥ −λI because A ≥ 0. Hence A − λI is invertible as −λI is, i.e. λ 6∈ σ(A).
A simple application of the Functional Calculus for self-adjoint operators is as
follows.
Example 1.6. Let A ∈ B(H) be such that 0 ≤ A ≤ I. If α ∈ [0, 1], then Aα ≥ A.
From [6] we recall the following result.
Proposition 1.7. Let A, B ∈ B(H) be commuting. Then
σ(A + B) ⊂ σ(A) + σ(B)
where
where
σ(A) + σ(B) = {λ + µ : λ ∈ σ(A), µ ∈ σ(B)}.
We call the result in the previous proposition the "subadditivity of the spec-
trum". There is also a "submultiplicativity of the spectrum", that is,
Proposition 1.8. ([6]) Let A, B ∈ B(H) be commuting. Then
σ(AB) ⊂ σ(A)σ(B)
σ(A)σ(B) = {λµ : λ ∈ σ(A), µ ∈ σ(B)}.
In Proposition 2.21, we show that Proposition 1.8 implies Proposition 1.7 in the
context of self-adjoint operators and that the backward implication also holds but
for positive and invertible operators.
Recall also the following definition.
Definition. Let T and S be unbounded positive self-adjoint operators. We say
that S ≥ T if D(S
2 ) ⊆ D(T
2 ).
1
1
The "natural but weak extension" is defined as in Definition 10.5 (Page 230) in
[11]: If S and T are non-necessarily bounded symmetric operators, then S (cid:23) T if
D(S) ⊂ D(T ) and
1
2 ) and(cid:13)(cid:13)(cid:13)S
1
2 x(cid:13)(cid:13)(cid:13) ≥(cid:13)(cid:13)(cid:13)T
1
2 x(cid:13)(cid:13)(cid:13) for all x ∈ D(S
hSx, xi ≥ hT x, xi ∀x ∈ D(S).
Notice that Proposition 1.5 remains valid for unbounded operators. Indeed, as
on Page 200 in [12], if S and T are self-adjoint, T is boundedly invertible and
S ≥ T ≥ 0, then S is boundedly invertible and S−1 ≤ T −1.
of Operator Theory.
Finally, we assume that readers are familiar with other basic notions and results
INVERTIBILITY OF SUMS
3
The sum of two invertible operators is not necessarily invertible even if strong
conditions are imposed. For instance, if we take A to be invertible and positive,
then setting B = −A, we see that AB = BA and that B is invertible. But plainly
A + B is not invertible. Positivity must also be avoided as it may make some of the
results evident. For instance, if A, B ∈ B(H) are such that A, B ≥ 0 and A say is
invertible, then obviously A + B (≥ A) is invertible by Proposition 1.5. These two
observations make the investigation of this question a little hopeless. However, the
approach considered by Bikchentaev in [1] deserves to be investigated further. This
is one aim of the paper. Another purpose is to treat some of these questions in an
unbounded setting. Some interesting consequences, examples and counterexamples
accompany our results.
2. Main Results
Theorem 2.1. Let A, B ∈ B(H).
(1) If A + B is invertible, then so is A2 + B2.
(2) If A + B is invertible, then so are A + B and A2n + B2n (n ∈ N) as
(3) Assume here that A, B ≥ 0 and let α, β ∈ C. Then
well.
αA + βB invertible =⇒ A + B invertible.
Remark. Most of the previous results appeared in [1], but our proof is simpler.
Proof. (1): It is clear that
A + B invertible =⇒ A + B2 invertible =⇒ A2 + B2 invertible
by Propositions 1.4 & 1.5.
(2): By the first property, A2 + B2 is invertible from which we readily get
that A4 + B4 is invertible and, by induction, we establish the invertibility of
A2n + B2n .
The invertibility of A + B is not hard to prove. WLOG we may assume that
kAk ≤ 1 and kBk ≤ 1. Hence A∗A ≤ I and B ∗B ≤ I, and so A ≤ I and B ≤ I
by Theorem 1.1. By the Functional Calculus, A2 ≤ A and B2 ≤ B. This
implies that
A + B ≥ A2 + B2 ≥ 0.
Proposition 1.5 allows us to confirm the invertibility of A + B, as desired.
(3): Since αA + βB is invertible, by the previous property, so is αA + βB
or merely αA + βB as A, B ≥ 0. Since we can assume α ≥ β > 0, we infer
that
α(A + B) ≥ αA + βB.
Consequently, α(A + B) or simply A + B is invertible.
Corollary 2.2. Let A ∈ B(H) be invertible. Then A − B + B is invertible for
every B ∈ B(H).
Proof. Since A is invertible and A = (A − B) + B, it follows that A − B + B too
is invertible by the previous result.
(cid:3)
(cid:3)
Remark. It is clear that if T is invertible, then the self-adjoint ReT + ImT need not
be invertible. For instance:
4
M. H. MORTAD
Example 2.3. Let A be self-adjoint and invertible and set B = −A. Then
is invertible while A + B = 0 is not.
A + iB = A − iA = (1 − i)A
Nonetheless, we have the following.
Corollary 2.4. Let A ∈ B(H) be invertible. Then ReA + ImA is invertible.
Proof. Just write A = ReA + iImA, then apply Theorem 2.1.
(cid:3)
The next corollary appeared in [1].
Corollary 2.5. Let A, B ∈ B(H) be such that A + B is invertible. If p, q ∈ (0,∞),
then Ap + Bq is invertible.
Proof. WLOG, we assume that A ≤ I and B ≤ I. We can always find an n ∈ N
such that p, q ≤ 2n. Then by Example 1.6, we have
Therefore,
Thus,
p
q
A
2n ≥ A and B
2n ≥ B.
Ap ≥ A2n
Ap + Bq ≥ A2n
and Bq ≥ B2n
+ B2n
.
.
Since A2n + B2n is already invertible (Theorem 2.1), we obtain the invertibility
of Ap + Bq from Proposition 1.5.
(cid:3)
Now, we give an interesting consequence on the invertibility of matrices of
bounded operators.
Corollary 2.6. Let A, B, C, D ∈ B(H) and define T ∈ B(H ⊕ H) by
T =(cid:18) A B
C D (cid:19) .
If T is invertible, then so are A + C and B + D.
In particular, if D is normal and B = 0 (resp. if A is normal and C = 0), then
Proof. Write
σ(D) (resp. σ(A)) ⊂ σ(T ).
T =(cid:18) A 0
0 D (cid:19)
+(cid:18) 0 B
C 0 (cid:19)
{z
}
{z
}
Since T is invertible, so is S + R. But
D (cid:19) and R =(cid:18) C
B + D (cid:19)
(cid:18) A + C
S =(cid:18) A
and so
0
0
0
0
0
S
R
.
0
B (cid:19)
is invertible. This means that A + C and B + D are invertible.
operator is invertible iff its absolute value is.
For the last claim just reason contrapositively by remembering that a normal
(cid:3)
INVERTIBILITY OF SUMS
5
Proposition 2.7. Let A, B ∈ B(H) be such that AB = BA and either A or B is
normal. Then
A + B is invertible ⇐⇒ A2 + B2 is invertible.
Proof. We already proved the implication "⇒" in Theorem 2.1.
Assume now that A2 + B2 is invertible. Since A commutes with B, it follows
by the Fuglede Theorem that A∗B = BA∗. Hence by known techniques, we may
get the commutativity of A2 and B2. Therefore, from Lemma 1.2 we obtain
Hence
pA2 +pB2 ≥pA2 + B2.
A + B ≥pA2 + B2.
SincepA2 + B2 is invertible, Proposition 1.5 gives the invertibility of A + B,
as wished.
(cid:3)
Remark. The invertibility of A2 + B2 does not yield that of A + B even in the event
of the self-adjointness of A and B. As a counterexample, just consider an invertible
and self-adjoint A such that A = −B.
Example 2.8. Let A ∈ B(H). We know that cos2 A + sin2 A = I. It then follows
that cos A + sin A is invertible.
Example 2.9. Let A be self-adjoint and invertible. Set B = iA. Then A2 +B2 = 0
is obviously not invertible.
The idea of the proof of Proposition 2.7 leads to the following generalization.
Proposition 2.10. Let A, B ∈ B(H) be such that AB = BA and either A or B is
normal. Then
A + B is invertible ⇐⇒ An + Bn is invertible
where n ∈ N.
The following example shows that both the real and imaginary parts of an in-
vertible operator may be not invertible.
Example 2.11. Consider the multiplication operators
Af (x) = cos xf (x) and Bf (x) = sin xf (x)
both defined on L2(R). Then A and B are not invertible. However,
T f (x) = (A + iB)f (x) = (cos x + i sin x)f (x) = eixf (x)
is invertible (even unitary!).
It is fairly easy to see that a normal T = A + iB is invertible iff A2 + B2
[8]). With this observation, we may state the following
is invertible (see e.g.
interesting characterization of invertibility for the class of normal operators.
Proposition 2.12. Let T = A + iB be normal in B(H). Then
T is invertible ⇐⇒ A + B is invertible.
In particular, if λ = α + iβ, then
λ ∈ σ(T ) ⇐⇒ A − αI + B − βI is not invertible.
6
M. H. MORTAD
Remark. (cf. [2]) Another way of establishing the previous result is as follows. By
[9], we know that if T ∈ B(H) is normal, then T ≤ ReT + ImT. Hence the
invertibility of T entails that ReT + ImT. Conversely, if T is normal (in fact
hyponormality suffices here), then ImT ≤ T and ReT ≤ T and so
ImT + ReT ≤ 2T.
Therefore, the invertibility of ImT + ReT implies that of T .
The following related version to Proposition 1.7 does not make use of the Gelfand
Transform.
Proposition 2.13. Let S, T ∈ B(H) be normal and such that ST = T S. Then
σ(S + T ) ⊂ σ(ReS + ReT ) + iσ(ImS + ImT ).
Proof. Write S = A + iB and T = C + iD. Since ST = T S and S and T are
normal, S + T is normal (see e.g. [10]). Hence, if we let λ = α + iβ ∈ C, then
S + T − λI = (A + C − αI) + i(B + D − βI)
becomes normal. So, if λ ∈ σ(S +T ), then Proposition 2.12 says that A+C−αI+
B + D − βI is not invertible. If either A + C − αI or B + D − βI is invertible,
then clearly A + C − αI + B + D − βI would be invertible! Therefore, both
A + C − αI and B + D − βI are not invertible, i.e. A + C − αI and B + D − βI
are not invertible. In other language, α ∈ σ(A+C) and β ∈ σ(B +D). Accordingly,
λ = α + iβ ∈ σ(A + C) + iσ(B + D), as wished.
Corollary 2.14. Let T = A + iB be normal in B(H). Then
(cid:3)
σ(T ) ⊂ σ(A) + iσ(B).
As another consequence, we have a new and shorter proof of a well known result.
Corollary 2.15. Let A ∈ B(H) be self-adjoint. Then σ(A) ⊂ R.
Proof. Let λ 6∈ R, i.e. λ = α + iβ (α ∈ R, β ∈ R∗). Since A − αI is self-adjoint, it
follows that A − αI − iβI is normal. By the invertibility of βI, it follows that of
A − αI + βI (by Proposition 1.5). By Proposition 2.12, this means that A − λI
is invertible, that is, λ 6∈ σ(A).
(cid:3)
The following result appeared in [9].
Proposition 2.16. 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.
As a consequence of the previous result, we get a very short proof concerning
the spectrum of unitary operators.
Corollary 2.17. Let A ∈ B(H) be unitary. Then σ(A) ⊂ {λ ∈ C :
Proof. We have A = I and so 1 − λI = I − λI ≤ A − λI. Thus, if λ 6= 1,
then λ 6∈ σ(A).
λ = 1}.
(cid:3)
It is known that a normal operator having a real spectrum is self-adjoint. The
proof is very simple if we know the very complicated Spectral Theorem for normal
operators. It would be interesting to prove this result along the lines of the proof of
Corollary 2.15. Notice also that this can very easily be established if the imaginary
INVERTIBILITY OF SUMS
7
part of T is a scalar operator. A new proof in the general case has not been obtained
yet. Nonetheless, as an application of Proposition 1.7, we have the following short
proof (which seems to have escaped notice) of this result.
Proposition 2.18. Let T = A + iB be normal in B(H) and such that σ(T ) ⊂ R.
Then T is self-adjoint.
Proof. Recall that A and B are self-adjoint. The normality of T is equivalent to
AB = BA. Hence T A = AT . Since T − A = iB, we have by Proposition 1.7
iσ(B) = σ(iB) = σ(T − A) ⊂ σ(T ) + σ(−A) ⊂ R.
Thus, necessarily σ(B) = {0}. Accordingly, the Spectral Radius Theorem gives us
B = 0 and so T = A, i.e. T is self-adjoint.
(cid:3)
What is also interesting is that since Proposition 1.7 holds in the context of
Banach algebras (see Theorem 11.23 in [10]), Proposition 2.18 becomes valid in the
context of C ∗-algebras. The proof is identical and so it is omitted.
Proposition 2.19. Let A be a C ∗-algebra and let a ∈ A be a normal element. If
σ(a) ⊂ R, then a is hermitian.
We also have the following related result.
Proposition 2.20. Let A be a C ∗-algebra and let a ∈ A be a normal element. If
σ(a) is purely imaginary, then a is skew-hermitian (i.e. a∗ = −a).
Proof. Write a = x + iy where x and y are commuting hermitian elements of A.
Write x = a − iy and proceed as above to force x = 0.
The last result about the spectrum is the following.
(cid:3)
Proposition 2.21. The "subadditivity of the spectrum" is equivalent to the "sub-
multiplicativity of the spectrum" in the context of bounded positive, commuting and
invertible operators.
Proof. Let A, B ∈ B(H) be self-adjoint and such that AB = BA. Assume that
Proposition 1.8 holds. Let λ ∈ σ(A + B). Then the Spectral Mapping Theorem
yields
eλ ∈ σ(eA+B) = σ(eAeB) ⊂ σ(eA)σ(eB),
that is, eλ = eαeβ for α ∈ σ(A) and β ∈ σ(B). Hence
λ = ln(eαeβ) = α + β ∈ σ(A) + σ(B).
Now, suppose that Proposition 1.7 holds. Suppose also here that A and B are
positive and invertible. Let λ ∈ σ(AB). Since AB is positive and invertible, by the
Spectral Mapping Theorem we get
ln λ ∈ σ[ln(AB)] = σ(ln A + ln B) ⊂ σ(ln A) + σ(ln B) = ln[σ(A)] + ln[σ(B)],
i.e. ln λ = ln α + ln β with α ∈ σ(A) and β ∈ σ(B). Thus, λ = αβ ∈ σ(A)σ(B), as
required.
(cid:3)
Let's go back to invertibility.
Proposition 2.22. Let A, B ∈ B(H) be such that AB ≥ 0 and AB is invertible.
Then A∗2 + B2 is invertible.
8
M. H. MORTAD
Proof. Let x ∈ H. Then
hABx, xi = hABx, xi = hBx, A∗xi
≤ kA∗xkkBxk ≤ 2(kA∗xk2 + kBxk2)
= 2(hAA∗x, xi + hB ∗Bx, xi) = h2(AA∗ + B ∗B)x, xi,
that is, AB ≤ 2(AA∗ + B ∗B) = 2(A∗2 + B2). Proposition 1.5 allows us to
establish the invertibility of A∗2 + B2, as required.
Remark. Notice that the result is obvious if either A or B is invertible. In order to
keep the result non-trivial we need also to avoid ker A = ker A∗ and ker B = ker B ∗
(see [4]). This remark applies to the unbounded case as well (treated in Theorem
2.25 below).
(cid:3)
Remark. The fact that we have assumed the invertibility of AB is essential as seen
by the following example.
Example 2.23. Let A be the positive operator defined by
Setting B = A, we see that ABf (x) = x2f (x) is positive. However,
Af (x) = xf (x), f ∈ L2(0, 1).
(A∗2 + B2)f (x) = (A2 + B2)f (x) = 2x2f (x)
is not invertible.
As a consequence of the foregoing proposition, we have the following.
Corollary 2.24. Let A ∈ B(H) be right (resp.
left) invertible with B ∈ B(H)
being its right (resp. left) inverse. Then A∗2 + B2 (resp. A2 + B ∗2) is always
invertible.
Proof. Since A is right invertible, for some B ∈ B(H), we have AB = I. Since
the latter is positive and invertible, the result follows immediately. The case of
left-invertibility is identical.
(cid:3)
We can generalize Proposition 2.22 to unbounded operators.
Theorem 2.25. Let A and B be two operators such that A is closed and B ∈ B(H).
(1) If BA is positive (i.e. hBAx, xi ≥ 0 for all x ∈ D(BA)) and invertible,
then A2 + B ∗2 is invertible. Besides,
h(A2 + B ∗2)−1x, xi ≤ h(BA)−1x, xi ∀x ∈ H.
(2) If AB is densely defined, positive and invertible, then A∗2 +B2 is invert-
ible. Moreover,
h(A∗2 + B2)−1x, xi ≤ h(AB)−1x, xi ∀x ∈ H.
Proof. (1): The first step is to show that BA ≤ 2A2 + 2B ∗2. Observe that
D(A2 + B ∗2) = D(A∗A + BB ∗) = D(A∗A) ⊂ D(A) = D(BA).
Now, let x ∈ D(A∗A). As in the bounded case, we may prove
hBAx, xi = hBAx, xi = hAx, B ∗xi ≤ kAxkkB ∗xk ≤ 2h(A2 + B ∗2)x, xi.
This means that BA (cid:22) 2A2 + 2B ∗2. Since BA is positive, it is (only) symmetric.
Since it is invertible, it follows that BA is actually self-adjoint (and positive).
Thus, by Lemma 10.10 in [11], "(cid:22)" becomes "≤", that is, we have established the
INVERTIBILITY OF SUMS
9
desired inequality BA ≤ 2A2 + 2B ∗2. Since BA is positive, invertible and BA
and A2 + B ∗2 are self-adjoint, it follows that A2 + B ∗2 is invertible by the
unbounded version of Proposition 1.5 (recalled also in the introduction), as wished.
(2): The idea is similar to the previous case. As AB is symmetric and invertible,
then it is self-adjoint (and positive). Hence B ∗A∗ ⊂ (AB)∗ = AB and so D(A∗) =
D(B ∗A∗) ⊂ D(AB). The main point is to show that
AB ≤ 2A∗2 + 2B2.
Clearly,
D(A∗2 + B2) = D(A∗2) = D(AA∗) ⊂ D(A∗) ⊂ D(AB).
Now, let x ∈ D(AA∗). Then
hABx, xi = hBx, A∗xi ≤ kBxkkA∗xk
≤ 2kBxk2 + kA∗xk2 = 2hBx, Bxi + 2hA∗x, A∗xi = h2(B2 + A∗2)x, xi.
As above, AB (cid:22) 2B2+2A∗2 becomes AB ≤ 2B2+2A∗2. Thus, the invertibility
of B2 + A∗2 follows from that of AB, as wished.
(cid:3)
Let's give an explicit and non-trivial application of the previous result.
Example 2.26. Let A be defined by Af (x) = f ′(x) on the domain
D(A) = {f ∈ L2(0, 1) : f ′ ∈ L2(0, 1)}.
Then A is densely defined and closed (but it is not normal, see e.g. [7]). Also,
A∗f (x) = −f ′(x) on D(A∗) = {f ∈ L2(0, 1) : f ′ ∈ L2(0, 1), f (0) = f (1) = 0}
so that A∗2 = AA∗f (x) = −f ′′(x) with
D(AA∗) = {f ∈ L2(0, 1) : f ′′ ∈ L2(0, 1), f (0) = f (1) = 0}.
Let V be the Volterra operator defined on L2(0, 1), i.e.
(V f )(x) =Z x
f (t)dt, f ∈ L2(0, 1).
Then (see e.g. [8]) V 2f = V ∗V f = P∞
n=1 λnhf, fnifn with (fn) being the eigen-
vectors corresponding to the distinct eigenvalues λn of V ∗V (and f ∈ L2(0, 1)).
Now, neither A nor V is invertible. However, A is right invertible for D(AV ) =
L2(0, 1) and AV f (x) = f (x) for f ∈ L2(0, 1). This means that AV is positive and
invertible. Therefore, the unbounded operator
0
d2
dx2 + V 2
is invertible on the domain D(AA∗) given above.
−
Next, we pass to the invertibility of finite sums.
Lemma 2.27. (cf. [3]) Let (Ak)k=1,...,n be in B(H) and let (ak)k=1,...,n be in C.
Then
2
n
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
akAkx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
≤
n
Xk=1
ak2* n
Xk=1
A∗
kAkx, x+
for all x ∈ H.
10
M. H. MORTAD
Proof. Clearly,
n
k=1 AkBk
k=1 A∗
n
n
n
n
(cid:3)
If
(cid:3)
k=1 Ak2.
≤
n
Xk=1
ak2! 1
2 n
Xk=1
all vanish simultaneously!).
2
kAkxk2! 1
akAkx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
kakAkxk ≤ n
Xk=1
k=1 akAk is invertible, it is bounded below, i.e., for some α > 0 and all
k=1 akAkxk ≥ αkxk. By Lemma 2.27 and by the self-adjointness
k=1 Ak2 is invertible (given that the ak cannot
(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)
Xk=1
for all x ∈ H.
Corollary 2.28. Let (Ak)k=1,...,n be in B(H) and let (ak)k=1,...,n be in C.
k=1 akAk is invertible, then so is Pn
Pn
Proof. AsPn
x ∈ H, we have kPn
kAk, it follows that Pn
of Pn
Theorem 2.29. Let (Ak)k=1,...,n and let (Bk)k=1,...,n be in B(H). If Pn
is positive and invertible, then Pn
k=1 Bk2 too is invertible.
Proof. Let x ∈ H. Then
AkBkx, x+ =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
* n
* n
Xk=1
Xk=1
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
kxi(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=1
Xk=1
hBkx, A∗
hBkx, A∗
≤
kBkxk2vuut
≤vuut
Xk=1
Xk=1
kA∗
kBkx, x+vuut* n
=vuut* n
Xk=1
Xk=1
2 * n
Bk2x, x+ +* n
Xk=1
Xk=1
≤
Since Pn
k2 +Pn
ofPn
k=1 A∗
Corollary 2.30. Let (Ak)k=1,...,n be in B(H). If Pn
invertible, then so is Pn
k2 +Pn
hAkBkx, xi(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
AkBkx, x+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Xk=1
Xk=1
kBkxkkA∗
kxi ≤
kxk
kBkx, xivuut
kxk2 =vuut
Xk=1
Xk=1
hB ∗
kx, x+
k2x, x+! =
A∗
k=1 AkBk is positive and invertible, Proposition 1.5 gives the invertibility
(cid:3)
k=1 Bk2, as needed.
k (or Pn
k2 +Pn
k ≥ 0 for every k. Hence clearlyPn
k=1 Ak2.
k=1 A∗
k=1 AkA∗
Proof. First, remember that AA∗
Now apply Theorem 2.29 to get the desired result.
k2)x, x+ .
1
2* n
Xk=1
k=1 AkA∗
k ≥ 0.
(cid:3)
(Bk2 + A∗
hAkA∗
kx, xi
n
n
n
n
k=1 A∗
kAk) is
B ∗
AkA∗
k=1 A∗
1
References
[1] A. M. Bikchentaev, On invertibility of some operator sums, Lobachevskii J. Math., 33/3
(2012) 216-222.
[2] I. Boucif, S. Dehimi and M. H. Mortad, On The Absolute Value of Unbounded Operators, J.
Operator Theory (to appear). arXiv:1804.00899v2
[3] C. Costara, D. Popa, Exercises in Functional Analysis, Kluwer Texts in the Mathematical
Sciences, 26, Kluwer Academic Publishers Group, Dordrecht, 2003.
[4] S. Dehimi, M. H. Mortad, Right (or left) invertibility of bounded and unbounded operators and
applications to the spectrum of products, Complex Anal. Oper. Theory, 12/3 (2018) 589-597.
INVERTIBILITY OF SUMS
11
[5] S. Dehimi, M. H. Mortad, Generalizations of Reid Inequality, Mathematica Slovaca (to ap-
pear). arXiv:1707.03320.
[6] R. V. Kadison, J. R. Ringrose, Fundamentals of the Theory of Operator Algebras. Vol. I. Ele-
mentary Theory. Reprint of the 1983 original. Graduate Studies in Mathematics, 15. American
Mathematical Society, Providence, RI, 1997.
[7] M. H. Mortad, A Criterion for the Normality of Unbounded Operators and Applications to
Self-adjointness, Rend. Circ. Mat. Palermo (2), 64 (2015) 149-156.
[8] M. H. Mortad, An Operator Theory Problem Book, World Scientific Publishing Co., (to appear
in 2018).
[9] M. H. Mortad, On The Absolute Value of The Product and the Sum of Linear Operators, Rend.
Circ. Mat. Palermo, II. Ser (to appear). DOI: 10.1007/s12215-018-0356-8. arXiv:1702.08671.
[10] W. Rudin, Functional Analysis, McGraw-Hill Book Co., Second edition, International Series
in Pure and Applied Mathematics, McGraw-Hill, Inc., New York, 1991.
[11] K. Schmüdgen, Unbounded Self-adjoint Operators on Hilbert Space, Springer. GTM 265
(2012).
[12] J. Weidmann, Linear Operators in Hilbert Spaces (translated from the German by J. Szücs).
Springer-Verlag. GTM 68 (1980).
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].
|
Subsets and Splits
No saved queries yet
Save your SQL queries to embed, download, and access them later. Queries will appear here once saved.