Optimal quantizers for Radon random vectors in a Banach space
Siegfried Graf Universität Passau, Fakultät für Mathematik und Informatik, D-94030 Passau, Germany. E-mail: graf@fmi.uni-passau.de
Harald LuschgyUniversität Trier, FB IV-Mathematik, D-54286 Trier, Germany. E-mail: luschgy@uni-trier.de
and Gilles Pages Laboratoire de Probabilités et Modèles aléatoires, UMR 7599, Université Paris 6, case 188, 4, pl. Jussieu, F-75252 Paris Cedex 5, France. E-mail:gpa@ccr.jussieu.fr
November 27, 2006
Abstract
For
and a Radon random vector
with values in a Banach space
let
, where the infimum is taken over all subsets
of
with card
(
-quantizers). We investigate the existence of optimal
-quantizers for this
-quantization propblem, derive their stationarity properties and establish for
-spaces
the pathwise regularity of stationary quantizers.
Key words: Functional quantization, optimal quantizer, stationary quantizer, stochastic process, intersection properties of balls.
2000 Mathematics Subject Classification: 41A46, 60B11, 94A29
1 Introduction
We investigate optimal quantizers and the quantization error in the functional
-quantization problem for stochastic processes viewed as random variables in a Banach (function) space. So let
be a real Banach space and consider a Radon random variable
which means that
is Borel measurable and its distribution
is a Radon probability measure on
.
For
and
, the
-quantization problem for
of level
consists in minimizing
over all subsets
with
. Such a set
is called
-codebook or
-quantizer. The minimal
th quantization error is then defined by
|
(1.1)
|
Under the integrability condition
the quantity
is finite.
For a given
-codebook
one defines an associated closest neighbour projection
and the induced
(or
) of
by
|
(1.3)
|
where
is a Voronoi partition induced by
, that is a Borel partition of
satisfying
for every
. Then one easily checks that, for any measurable random variable
,
so that finally
| |
|
(1.4)
|
Functional quantization of stochastic processes can thus be seen as a discretization of the path-space
of a process and the approximation (coding) of a stochastic process by finitely many deterministic functions from its path-space. Typical settings are
and
. Functional quantization is the natural extension to stochastic processes or Banach space valued random vectors of the so-called optimal vector quantization of random vectors in
which has been extensively investigated since the late 1940's in Signal processing and Information Theory (see [9] , [15] ). For the mathematical aspects of vector quantization in
, one may consult [13] and for algorithmic aspects see [25] .
Recently, the extension of optimal vector quantization to stochastic processes has given raise to many theoretical developments including the rate of convergence of the quantization errors
to zero as
and the construction of good or even rate optimal quantizers (see e.g. [6] , [7] , [8] [14] , [21] , [22] , [23] ). For a first promising application to the pricing of financial derivatives through numerical integration on path-spaces see [26] . In this paper we aim to develop general results on the existence of optimal quantizers and their properties.
The paper is organized as follows. In Section 2, a theorem about the existence of optimal
-quantizers for
-valued Radon random vectors lying in
or in some suitable superspace
is established under some very general assumptions. It relates existence to intersection properties of closed balls. This problem is connected with its bidual counterpart and enlightened by counterexamples.
Furthermore, bounds of the quantization errors
in terms of
for superspaces
and in terms of marginals of
for vector valued processes are derived. In Section 3 the stationarity property of optimal
-quantizers is investigated. This turns out to be an essential key for the functional quantization of 1-dimensional diffusion processes (see [23] ). For smooth Banach spaces stationary quantizers are defined as the critical points of the distortion function. In the case of
-spaces
which are natural path-spaces of processes some pathwise regularity for these stationary quantizers is established. The result applies e.g. to Gaussian processes,
-dimensional diffusion processes and certain Lévy processes.
2 Optimal quantizers and quantization errors
Let
be a Radon
-valued random variable with distribution
. The Radon property of
means inner regularity w.r.t. compact sets and on Banach spaces it is the same as tightness which in turn is equivalent to the existence of a separable Borel measurable set with
-probability 1. It is to be noticed that if
for some Banach subspace
of
is Radon when viewed as
-valued random variable. On the other hand, if
is a Banach subspace of some Banach space
then
is also Radon as
-valued random variable.
We will assume throughout this section that
satisfies the integrability condition (1.2) for some
. Then
|
(2.1)
|
As a matter of fact, the support of
being separable there exists a countable subset
everywhere dense in
. It is clear that
by the Lebesgue dominated convergence Theorem. On the other hand, the existence of optimal quantizers, i.e. the fact that
actually stands as a minimum needs much more care.
2.1 Existence of optimal quantizers
set
is called an
-optimal
-quantizer for
if
|
(2.2)
|
Let
denote the set of all
-optimal
-quantizers for
in
.
We first provide some interesting properties of
-optimal quantizers (they can be seen as necessary conditions for
-optimality). Their proofs are literally the same as those (established in finite-dimension) of Theorem 4.1 and Theorem 4.2 in [13] respectively. They are related with the Voronoi partitions induced by a
-quantizer
: these are the Borel partitions
of
which satisfy
|
(2.3)
|
Let us note that
is closed and star-shaped relative for
and for every
,
Furthermore, as soon as
is strictly convex (
), any Voronoi partition satisfies for every
|
(2.4)
|
and
Proposition 1
Assume that card
.
Let
. Then card
and for every
,
Assume that
is smooth (
) and strictly convex. If
and
then
|
(2.5)
|
Note that, under the strict convexity assumption, (2.5) is then equivalent to both
The first results of existence for optimal quantizers are due to Cuesta-Albertos and Matràn [5] and Pärna [24] ) for uniformly convex and reflexive Banach spaces, respectively. We provide an extension to Banach spaces having the property that the closed balls form a compact system. A system
of subsets of
is called compact if each subsystem
of
which has the finite intersection property
the intersection of each finite subsystem of
is not empty) has a nonempty intersection. Let
be the closed ball of radius
centered at
.
Definition 1
A pair
consisting of a Banach space
and a Banach subspace
of
is called admissible if
is a compact system in
.
is called admissible if
is admissible.
The level
-distortion function is defined by
|
(2.6)
|
Theorem 1
Assume that
for some Banach subspace
of
and that
is admissible. Then, for every
,
Proof. Fix
. Let
denote the topology on
generated by the system
and let
be the product topology on
(these topologies usually do not satisfy the Hausdorff axiom). The family
being a compact system in
, one checks that
is
-quasi-compact(
). Consequently,
is
-quasi-compact. It is obvious that any lower semi-continuous (l.s.c.) function defined on
then reaches a minimum. Hence, the proof amounts to showing that the distortion function
is
-lower semi-continuous.
For every
and
, set
. Then
is
-closed for every
. Hence,
is
-lower semi-continuous. In turn any convex combination of such functions are
-l.s.c. as well. This implies that
(and
) are
-lower semi-continuous provided card
.
For general
we will show that for every
is
-open. First note that from (1.4) and (2.1), there exists a sequence of quantizations
, card
, such that
Consider first the case
. It follows from Minkowski's inequality that, for every
,
is
-Lipschitz on
:
| |
| |
|
(2.7)
|
Let
. It follows from ( 2.7 ) that, the
-open set
is always contained in
. Furthermore, it contains
for large enough
, still by ( 2.7 ).
Hence
is
-open and
is
-l.s.c.
When
, one concludes the same way round, using now that
for every
, one derives that for every
,
In the non-quantization setting
, Theorem 1 with
is due to Herrndorf (see [16] ).
One easily checks that if
is a 1-complemented closed subspace of some Banach space
and
is admissible, then
is admissible. Here
is said to be
-complemented in
if there is a linear projection
from
onto
with
. An interesting case is
. One simply notes that the closed balls in the bidual
of
are
-compact and thus
is admissible.
The following characterization is a slight generalization of Theorem 5.9 in [19] .
Proposition 2
is admissible if and only if
In particular, if
is 1-complemented in its bidual
, then
is admissible.
An investigation of the admissibility feature of Banach spaces
and the ball topology
(with
) used in the proof of Theorem 1 can be found in [10] , [11] .
One derives for three main classes of Banach spaces the following corollary.
Corollary 1
In any of the following cases
is 1-complemented in
and hence, for every
.
(i)
is a
(Kantorovich-Banach)-space.
(ii)
is a dual space.
(iii)
is an order complete AM-space with unit.
Proof.
By definition, a Banach lattice that is a band in its bidual is a
-space. Since
is an order complete Banach lattice,
is a projection band in
and the band projection from
onto
has norm 1. (cf. [28] , Chap. II.5).
Dual spaces are clearly 1-complemented in their bidual.
See [28] , Chap. II.7.
The order complete AM-space without unit
and the AM-space with unit
which is not order complete admit random variables
without optimal
-quantizers even for
(see the subsequent counterexamples) In particular, both spaces are not admissible.
Example.
-spaces are equipped with the norm
if
and
-ess sup
if
, where
denotes the
-norm on
.
-spaces with respect to arbitrary measure spaces and are
-spaces and hence
-spaces.
-spaces,
, with respect to arbitrary measure spaces are reflective and hence dual spaces.
-spaces with respect to
-finite measure spaces are dual spaces and also order complete
-spaces with unit (cf. [28] , Chap. IV 7).
Remarks.
Concerning the Banach spaces
, the above theorem provides new existence results for the
-optimal quantizers in the cases
and
.
Any pathwise continuous process
is an
-Radon random variable since
is a Polish subspace of
(any probability on a Polish space is tight
Radon). The above existence theorem shows that if
for some
, then, for every
,
has at least one
-optimal
-quantizer for the
-norm. However, nothing is known about the pathwise regularity of these optimal quantizers. Surprisingly, we will see in Section 3 that, for the same process,
-optimal
-quantizers with
have much more regular paths (
considering
and
).
Optimal
-quantizers may not exist in
Let
where
denotes the set of real valued sequences
such that
and
.
Let
denote the canonical basis of
defined by
where
is for the Kronecker symbol. One considers an
-valued random vector
supported by
with a distribution
satisfying
for every
. Now
so that
. One checks that the assumption of Theorem 1 is not fulfilled either since the system
has an empty intersection whereas any finite subsystem has a nonempty intersection.
So let
and
. We will show that
More precisely we will show that the corresponding level
quantization problem extended to the Banach space
does have a unique solution
in
given by
, that is
which in turn implies that it admits no solution in
. In fact,
For an arbitrary
one gets the following: if
for some
, then, for every
,
Hence
|
(2.8)
|
| |
| |
| |
In case
for every
, one clearly obtains
According to the above reasoning, any
that achieves the infimum must satisfy
for every
which clearly implies
. Finally
On the other hand, as a minimizing sequence from
one may choose
. Then
Consequently,
and since
, it follows that
is empty.
This example is enlightened by the general Theorem 2 . This theorem solves the correspondence between the quantization problem in
and in
. It shows that the quantization error does not decrease when
is seen as random vector in the bidual
of
and that the set of its optimal
-quantizers as an
-valued random vector is made up with those of its optimal
-quantizers as an
-valued random vector that lie in
. In particular,
corresponds to the phenomenon that any optimal
-quantizer of
has at least one element in
: this is precisely what happens in the above example.
Theorem 2
We have for every
,
In particular,
If card
, then
.
Assume that
is admissible. Further assume
. Then
We first need the following equivariance properties contained in the lemma below.
Lemma 1
Let
and
be Banach spaces and let
be a Radon
-valued random vector satisfying
.
If
is a bounded linear operator, then
If
is a bijective linear isometry,
and
, then
Proof. Let us prove
the first assertion. For any
with
,
| |
| |
and thus the assertion.
Proof of Theorem 2 .
The inequality
is obvious. To prove the converse inequality assume first that
is finite. Let
and let
denote the linear subspace of
spanned by
. Since
is finite-dimensional, there exists by local reflexivity of
, for every
, a bounded linear operator
satisfying
and
for every
. (cf. [20] Lemma 1.e.6). Using Lemma 1, one derives
| |
| |
Hence
For general
and
, choose a quantization
of
, card
, for sufficiently large
such that
| |
Then,
and
Since card
, we have
. This yields
Hence
. Furthermore, since
by Corollary 1, it follows from Proposition 1
that
is strictly decreasing provided card
.
The inclusion
follows from
. To prove the converse inclusion, we may assume
. Let
. By Proposition 2, for every
there exists
such that for every
,
Setting
this implies
and that the closed set
satisfies
. Therefore,
and in particular,
. One obtains
for every
and hence,
. By Proposition 1
, we have card
which yields
. Hence
.
Remark. It is to be noticed that the situation
never occurs for Gaussian (Radon) random vectors
. In view of Lemma 1, we may assume without loss of generality that
is centered. Let
. It follows from the Anderson inequality ([18] ) that, for every
,
so that
. However, it remains an open question whether
may be empty for
or not.
An immediate consequence of Theorem 2(a) is as follows. Let us call a Banach subspace
of
locally
-complemented
if there is a linear operator
of norm
satisfying
for every
. Notice that local 1-complementation coincides with the notion of an ideal introduced in [12] .
Corollary 2
Assume that
for some Banach subspace
of
and that
is locally 1-complemented in
. Then, for every
,
In particular,
implies
.
Proof. It follows from Theorem 2(a) and Lemma 1 that
| |
| |
One observes that the preceding corollary contains Theorem 2(a) since
is obviously locally 1-complemented in
.
Example
AM-spaces
are locally 1-complemented as Banach subspace in any Banach space
. In fact, since
is an order complete AM-space with unit, this feature follows from Theorem II.7.10 in [28] . For instance, if
for some compact metric space
and
for some closed subset
of
, then
is a closed vector sublattice of the AM-space
and thus an AM-space.
AL-spaces
are 1-complemented as Banach sublattice in any Banach lattice
(see [28] , II.8). Finite dimensional subspaces of dimension
are admissible but not necessarily (locally) 1-complemented. In fact, it may happen that
for some 2-dimensional subspace
of
and
even for
. In particular,
is not admissible. The following example is taken from Herrndorf [16] .
A counterexample when dim
Let
be equipped with the
-norm
.
Let
be the canonical basis of
and set
and
. Consider the
-valued random variable
supported by
with
. Let
denote the linear span of
in
. So
and dim
.
Let
and
. First will show that
and
In fact,
On the other hand, once noticed that
for
, one shows like in the previous counterexample that for every
and that any
-optimal 1-quantizer
must satisfy
for every
which implies
. As for
, observe that
Any
can be written as
, so that
| |
| |
since
. This yields
.
Now we construct a Banach subspace
of
such that
and
Choose
such that
and
is strictly increasing with
. Define
as the hyperplane
Then
. For
, set
. One obtains
and
Consequently,
For an arbitray
one gets the following: if
for
, then
and hence
. If
for some
, a can be strictly improved. Set
One checks that
and for every
,
This implies
Consequently,
.
2.2 Optimal quantizers for continuous stochastic processes
Now we turn to
-valued pathwise continuous processes
indexed by a compact metric space
. The space
of
-valued continuous functions on
and the space
of bounded,
-valued, Borel measurable functions on
are Banach spaces under the norm
|
(2.9)
|
where
denotes the
-norm on
. Since
is separable,
is Radon when viewed as
-valued random variable. Consequently,
is Radon as
-random variable.
Theorem 3
Let
be compact metric space. Then the pair
is admissible under the norm (2.9). In particular, if
is a
-valued pathwise continuous process with
, then for every
,
The proof of Theorem 3 is based on the admissibility of
-spaces and the following “lifting property”.
Lemma 2
Let
be a finite Borel measurable on the compact metric space
with
.
Then for every
there exists
such that
-a.e. and
where
|
(2.10)
|
Proof. One notes that for
,
and
Therefore, it is enough to consider the case
. Set
and
. Let
be a countable dense subset of
. Observe that the norms
and
coincide on
. This is a consequence of the assumption
. Let
. For
, set
Then
and
are Borel subsets of
with
-measure zero. Consequently,
satisfies
. Since for every
and
,
one obtains
|
(2.11)
|
The construction of the function
is given in two steps.
Step 1. For
and
, let
where
denotes the open ball in
of radius
centered at
. Define the “upper limit function”
of
by
One easily checks that for any Borel subset
of
, the function
is Borel. Therefore, for every
,
| |
| |
is a Borel set and thus
is Borel measurable. The function
has the following property: for every
there exists a sequence
in
such that
and
.
In fact, let
and let
so that
. For every
, there exists
such that
This implies
Step 2. Define
by
We show that
has the required properties. Observe that
is Borel measurable,
-a.e. and
. Let
If
, then
and hence by (2.4),
.
By step 1, if
, there exists a sequence
in
such that
and
.
Therefore,
| |
| |
| |
Consequently,
|
(2.12)
|
Now let
. There exists a sequence
in
such that
. For every
,
| |
Since
one obtains
|
(2.13)
|
Conversely, we clearly have
Proof of Theorem 3. Let
be a system of closed balls in
with centers
satisfying the finite intersection property. Choose a finite Borel measure
on
such that
and consider the system
of corresponding closed balls in
under the norm
(see (2.10)) where
denotes the quotient map. It is obvious that
also has the finite intersection property. Since
is admissible by Proposition 2 and Corollary 1,
has a nonempty intersection. Let
be a member of this intersection. Lemma 2 implies that there is a function
such that
-a.e.
and
Consequently,
belongs to the intersection of
. This yields the required admissibility.
One derives from Corollary 2 that the quantization error does not decrease when
is seen as
-or even
-valued random variable.
Theorem 4 Assume that
is a
-valued pathwise continuous process indexed by a compact metric space
with
. . Let
be a finite Borel measure on
with
. Then for every
,
where
and
are equipped with the sup-norm (2.9) and
is equipped with the norm (2.10). In particular,
| |
| |
Proof.
is an AM-space so that Corollary 2 applies. We obtain
Since
can be considered as a subspace of
, the same argument yields
(The latter equality is also an immediate consequence of Lemma 2.)
We will exhibit a pathwise continuous process
having no
-optimal 1-quantizer in
. In particular, due to the lack of order completeness,
is not admissible.
Optimal 1-quantizer may not exist in
Let
. Define, for every
, a continuous function
by
One considers an E-valued random variable
supported by
with
satisfying
for every
and
. The assumption of Theorem 1 is not fulfilled since the system
has the finite intersection property whereas it has an empty intersection (see below).
Let
and
. We will show that
Recall that by Theorem 4,
where
equipped with
.
Set
. One checks that, for every
,
so that
On the other hand, one shows like in the
-counterexample preceding Theorem 2 that for every
and that any
-optimal 1-quantizer
must satisfy
for every
: one reproduces the string of inequalities starting at (2.8) once noticed that
for every
. This implies
and
.
Furthermore, no
can satisfy the condition
for every
. In fact, if
, then
with
and
large enough so that
If
, then
for
large enough so that
Consequently,
.
2.3 Bounds for quantization errors
As before let
be a Radon random variable in
satisfying the integrability condition (1.2). The following observation (a) is already contained in [4] .
Proposition 3
Assume that
for some Banach subspace
of
.
(a) For every
,
(b) If
is locally c-complemented in
, then for every
,
Proof. (a) We have to prove only the second inequality. Let
and
.
Choose
such that
. This implies that
for every
and hence
Consequently,
This yields the assertion.
(b) is an immediate consequence of Theorem 2(a) and Lemma 1.
It is to be noticed that the factor 2 in part (a) of the preceding proposition is sharp. It cannot be improved as universal constant. This is demonstrated in the subsequent Example. In view of (a), the cases of interest in part (b) are
.
The constant 2 is sharp We modify the setting of the counterexample following Corollary 2.
Let
and let
denote the canonical basis of E. Fix
and set
. One considers the
-valued random variable
supported by
with
. Let
denote the linear span of
. So
.
Let
and
. One checks like in the above mentioned counterexample that
We will show that
In fact, for
,
Any
can be written as
and hence
| |
| |
Using the elementary inequalities
and
, one obtains
Consequently,
Next we describe marginal bounds for
-valued stochastic processes. For
, let
,
finite measure, equipped with the norm
|
(2.14)
|
Assume that
is separable. Let
be a bi-measurable
-valued process such that
Then the process
can be seen as a (Radon) random vector taking its values in
. For the sake of simplicity, we consider the case
. As for bounds when constants are not important there will be no loss of generality since usual inequalities on
-norms imply for
Proposition 4
Let
. For every
such that
,
Proof. As for the upper estimate, let
be a
-optimal
-quantizer for
(see Corollary 1). Set
. Thus
consists of functions
with
and card
. One obtains
| |
| |
| |
| |
As for the lower estimate, let
with card
. Then
| |
| |
This yields the lower estimate.
Now let
be a compact metric space and assume that
is a
-valued continuous process. Let
equipped with the sup-norm (2.9). Assume
|
(2.16)
|
Proposition 5
Let
. Let
such that
. Then for every
such that
,
Proof. For
and
, choose
such that card
and
Set
. Then
, card
and
| |
| |
| |
| |
This yields the upper esxtimate. As for the lower estimate, let
with card
.
Then for every
,
which gives the lower estimate.
In the preceding proposition one may replace
and
by
and
respectively for any finite Borel measure
on
with
. This follows from Theorem 4.
3 Stationary quantizers
Let
be a Radon
-valued random variable satisfying condition (1.2). We will introduce a notion of
-stationary quantizer as the critical points of level
-distortion function
formerly defined by Equation (2.6). For a quantizer
let
and
.
Definition 2
A
-quantizer
of size
is called admissible for
if
A
-tuple
is admissible if its associated
-quantizer is.
Proposition 6
Assume that
is smooth. Let
. Then the
-distortion function
is Gateaux-differentiable at every admissible
-tuple
with a Gateaux differential given by
where
denotes any Voronoi partition induced by
. If the norm is Fréchet-differentiable at every
, then
is the Fréchet derivative.
Furthermore, if
is uniformly smooth, then
is continuous on the set of admissible
-tuples (where
is endowed with its norm).
When
, the above results extend to admissible
-tuples with
.
Remark. In case
, the above proposition as well as Proposition 1(b) do not apply since the
-norm is neither smooth nor strictly convex.
Proof. A straightforward adaptation of Lemma 4.10 in [13] yields both differentiability properties.
Then, if
is uniformly smooth, the mapping
is continuous (see [2]). One derives the continuity of
by the Lebesgue dominated convergence theorem using that
takes its values in the unit ball of
.
Definition 3
Let
be a Banach space and let
. A
-quantizer
of size
is called
-stationary for
if
and
|
(3.1)
|
where
denotes any Voronoi partition induced by
. (This requires that the Gateaux-differential
is defined
-
on
and, furthermore, that
when
).
This finally leads to the following proposition which makes the (expected) connection between optimality and stationarity.
Proposition 7
Assume that
is smooth and strictly convex. Let
. Assume that card
. Then any
-optimal
-quantizer
is
-stationary (and admissible) for
. This extends to
if
.
Proof. Any
-optimal
-quantizer
is admissible by Proposition 1
, hence the Gateaux-differential
does exist and is
which exactly means stationarity.
3.1 Stationarity for stochastic processes
Let
be a finite measure space, let
be a bi-measurable
-valued process defined on a probability space
and let
. Assume that
is separable and that
i.e.
|
(3.2)
|
Then, the process
can be seen as a (Radon) random vector taking its values in the Banach space
satisfying an
-integrability property, that is
. When
, the
-spaces are uniformly smooth and strictly convex, so the above abstract results apply.
Furthermore, if
denotes the conjugate Hölder exponent of
, for every
,
,
so that the
-stationarity condition reads for any Voronoi partition
with
, for every
,
|
(3.3)
|
with the convention
, where
. When
, the condition is formally the same. This may be written in a more synthetic way by introducing the
-quantization
of
defined by (1.3), namely:
|
(3.4)
|
When
,
(and
if
), Equation (3.3) looks simpler and reads
|
(3.5)
|
One derives from Proposition 7 and Proposition 1 the following corollary.
Corollary 3
Let
, let
. If
|
(3.6)
|
then, any
-optimal
-quantizer is
-stationary in the sense of (3.3).
Proof. It remains to consider the case
. The space
is not smooth. However,
is Gateaux-differentiable at every
such that
a.e. for every
. Now, by the Fubini Theorem, one has for every
i.e.
-a.e.)
-a.s. Let
be an
-optimal
-quantizer and
. This definition is consistent since
by Proposition 1 (a). It follows easily that
,
, is Gateaux differentiable with a Gateaux-differential given by
Now, still following Proposition
is a minimum for
so that its Gateaux differential is zero.
Hence, for every
,
Remark. Continuity of
a.e. for some
implies continuity of
.
3.2 Pathwise regularity of stationary quantizers (
)
As before, let
for some finite measure space
such that
is separable. We will derive from Equations (3.3) (and (3.5)) some pathwise continuity result for the
-stationary quantizers (which extends a result established in [21] in the purely quadratic case
). For
, if
for every
, define the “intrinsic” semimetric
on
by
Theorem 5 Let
,
. Let
be a bi-measurable
-valued process satisfying (3.2) and
Let
be an
-stationary
-quantizer (in the sense of (3.3)). Set
if
and
otherwise.
(a) Let
be a compact metric space and let
be a continuous finite Borel measure on
. If
, if
is pathwise continuous with
(in case
is viewed as a
-random vector) for some
and some closed subset
of
with
and if the distribution
is continuous on
for every
, then the components of
have
-versions consisting of continuous functions such that
.
(b) If
, then the components
of
have
-versions consisting of
-continuous functions. Furthermore, if
, then there are such versions with
.
(c) If
, then the components
of
have
-versions consisting of
-Lipschitz continuous functions.
Remarks.
If
is continuous then
.
If
and
, then
even lies in the reproducing kernel Hilbert space of
. This is a consequence of (3.5).
Let
be a separable metric space and
a finite Borel measure on
. If
and
from
into
is continuous that is
is majorized by the initial metric
on
, then the
-components of
have versions consisting of
-continuous functions. The
-continuity assumption is fulfilled e.g. if
is pathwise
-continuous and
.
Proof of Theorem 5. For every
, set
. The measure
is finite: if
, this is obvious, otherwise,
On the other hand,
is a nonzero measure equivalent to
since
and for
,
. Now, define on
the function
by
First note that the function
is real valued. If
, the Young inequality with
and
implies
| |
| |
so that
. When
(or
), the result is obvious.
For every fixed
and
,
is (strictly) increasing, hence
is strictly increasing too. The continuity of
on
for every
follows from the Lebesgue dominated convergence Theorem. Furthermore, for every
,
so that
by Fatou's Lemma. Similarly,
.
The proof reduces to providing an argument for the
-continuity of
for every
.
If
, one starts from the inequality
When
, the Hölder inequality applied with the conjugate exponents
and
yields
| |
| |
This still holds if
.
If
, one starts from
Since
the Holder Inequality applied with
and
yields
| |
| |
A new application of the Holder Inequality to the expectation in the right hand side of the above inequality yields
| |
| |
| |
Owing to these properties, one easily checks that for every
, the equation
admits a unique solution
and that the implicitly defined function
is
-continuous.
On the other hand the function
satisfies
-a.e.
so that
-a.e..
If
then
so that
.
Now let
be a compact metric space. When
,
The continuity of
on
for every
and the continuity of
at every point
for every
follows from the pathwise continuity of
and from the continuity of
, by the Lebesgue dominated convergence Theorem: the sign function is bounded and
. Similarly one shows that
. It is also obvious that
is nondecreasing
. To establish strict monotonicity
, one proceeds as follows: let us consider the subset of
defined by
where
It is a nonempty open subset of
since
is everywhere
-dense in
in view of
. Now, for
and every nonempty open interval
the set
is clearly everywhere dense in
since
so that
is a nonempty set. On the other hand,
and
are both continuous as functionals on
so that
is a (nonempty) open subset of
. Now, if
for some
, then
.
is equivalent to
. Consequently
This is impossible owing to the assumption on the support of
. Consequently
is strictly increasing for every
and one concludes like in the case
to the existence of a continuous version of
in
.
To be a bit more precise, the equation
has for
a unique solution
and for
, since
-a.s.,
is the unique solution. The function
is continuous at every
since
is strictly increasing on
and
is continuous at
for every
. One must consider the behaviour of
at
more carefully.
First note that
is continuous at
for every
since
is pathwise continous.
Now let
be a sequence in
going to
such that
for some
. Then,
for every
so that
which is impossible. Hence
. One shows similarly that
i.e.
.
It is a consequence of Equation (3.5):
When
, The Holder Inequality yields the announced result
with
When
, one sets accordingly
.
Examples First consider real or
-valued processes with
and
.
The
-stationary
-quantizers,
, of the standard Brownian motion and are made up with continuous functions which are null at
,
-Hölder if
. The same result holds for the Brownian bridge over
where any of its stationary quantizers are
at
and for the standard
-dimensional Brownian motion.
One considers a
-valued Brownian diffusion process
| |
where
is a
-dimensional standard Brownian motion and
,
are Borel functions with linear growth such that the above SDE admits at least one (weak) solution over
. This solution is pathwise continuous and it is classical background (see [17] ) that
for every
and
for every
. Thus the
-stationary
-quantizers,
, are made up with continuous functions which are
at
,
-Holder if
. The same holds if
for the homogeneous SDE with
and
independent of
and
provided
and
are bounded with bounded derivatives up to order 3 and
is uniformly elliptic. In fact, the assumptions imply that
has a Lebesgue density for every
and by the support theorem, in
,
(see [3], p. 11 and [1], p. 25).
The fractional Brownian motion
on
with Hurst exponent
is a centered continuous Gaussian process having the covariance function
and thus satisfies for every
Consequently,
-stationary
-quantizers,
are made up with continuous functions which are null at
, H-Hölder if
.
We consider some examples of (cadlag) real Lévy processes
restricted to
(without Brownian component). Since the increments of
are stationary and
,
so that the behaviour of the semimetric
reduces to the behaviour of
.
The
-stable Lévy motions indexed by
satisfy a self-similarity property, namely
Furthermore,
For this background see [27] . It follows that for every
Consequently, since the
-stable distributions
have a Lebesgue density, the
-stationary
-quantizers,
, are made up with continuous functions which are null at 0.
The
-processes are Lévy processes whose distribution
at
is a
-distribution
. So, for every
Consequently,
-stationary
-quantizers,
, are made up with continuous functions,
-Hölder if
.
The compound Poisson process is given by
, where
are i.i.d. real random variables with
and
is a standard Poisson process (with intensity
) independent of
. If
and
, easy computations show that
Assume
. Then the
-stationary
-quantizers,
, are made up with continuous functions,
-Hölder if
. Here it has to be noticed that the function
is the only atom of
in
.
Theorem 5
does not apply to the above examples because of the pathwise continuity assumption so that the case
remains open.
As for a real multiparameter process on
with
:
The
-stationary
-quantizers,
, of the standard Brownian sheet are made up with continous functions which are null on
and 1/2-Hölder if
.
As for an example with noncompact
consider
and
.
The stationary Ornstein-Uhlenbeck process
on
is a centered continuous Gaussian process having the covariance function
Clearly,
can be seen as
-valued random vector for every
. The process satisfies for every
Consequently,
-stationary
-quantizers,
have components consisting of continuous functions,
-Hölder if
.
A counterexample when
We will exhibit a bounded pathwise continuous process
on
having a discontinuous
-optimal
-quantizer. Consider functions
and
from the
-counterexample following Theorem 4. Then set
. One checks that in
, for every
,
so that
On the other hand,
by the
-counterexample and Theorem 4. Consequently, the
-norm being nondecreasing as a function of
,
with obvious definition of
. The function
is an
-optimal
-quantizer without continuous
-version of the pathwise continuous process
,
.
Note that
from
into
is continuous for any
since
is pathwise continuous and uniformly bounded by
. Consequently it follows from Theorem 5 that, as soon as
, any
optimal
-quantizer of
(has a
-version which) consists of continuous functions. However,
from
into
is not continuous (at
), so the pathwise regularity of an optimal
-optimal
-quantizer of an
-continuous process remains open. But the
-norm being nowhere Gateaux-differentiable, the very notion of
-stationary quantizer no longer exists. So this would require to develop a new approach.
References
-
Bass, R.F., Diffusions and Elliptic Operators, Springer, New York, 1998.
-
Beauzamy, B., Introduction to Banach Spaces and their Geometry (2nd edition 1985), North-Holland Mathematics Studies 68, North-Holland, Amsterdam, 1982.
-
Bichteler, K., Gravereaux, J.-B., Jacod, J., Malliavin Calculus for Processes with Jumps, Gordon and Breach Science Publishers, 1987 (2nd printing, 1994).
-
Creutzig, J., Approximation of Gaussian random vectors in Banach spaces, PhD thesis, Univ. Jena, 2002.
-
Cuesta-Albertos, J.A., Matran, C., The strong law of large numbers for k-means and best possible nets of Banach valued random variables, Probab. Theory Rel. Fields, 78, 1988 523-534.
-
Dereich, S., Fehringer, F., Matoussi, A., Scheutzow, M., On the link between small ball probabilities and the quantization problem for Gaussian measures on Banach spaces, J. Theoretical Probab., 16, 2003, 249-265.
-
Dereich, S., High resolution coding of stochastic processes and small ball probabilities, PhD thesis, TU Berlin, 2003.
-
Dereich, S., The quantization complexity of diffusion processes, Preprint, 2004.
-
Gersho, A., Gray, R.M., Vector Quantization and Signal Compression, Kluwer, Boston, 1992.
-
Godefroy, G., Application à la dualité d'une propriété d'intersection. Math. Z., 182, 1983, 233-236.
-
Godefroy, G., Kalton, N.J., The ball topology and its applications. Contemporary Mathematics, 85, 1989, 195-237.
-
Godefroy, G., Kalton, N.J., Saphar, P.D., Unconditional ideals in Banach spaces. Studia Mathematica, 104, 1993, 13-59.
-
Graf, S., Luschgy, H., Foundations of Quantization for Probability Distributions, Lecture Notes in Mathematics
1730, Springer, 2000.
-
Graf, S., Luschgy, H., Pages G., Functional Quantization and small ball probabilities for Gaussian processes, J. Theoretical Probab., 16, 2003, 1047-1062.
-
Gray, R.M., Neuhoff, D.L., Quantization, IEEE Trans. Inform. Theory, 44, 1998, 2325-2383.
-
Herrndorf, N., Approximation of vector-valued random variables by constants, J. Approx. Theory, 37, 1983, 175-181.
-
Karatzas, I., Shreve, S., Brownian Motion and Stochastic Calculus, Springer, New York, 1988 (
edition 1991).
-
Ledoux, M., Talagrand, M., Probability on Banach Spaces: Isoperimetry and Processes, Springer, Berlin, 1991 (
printing, 2002).
-
Lindenstrauss, J. Extension of Compact Operators, AMS, 1964
-
Lindenstrauss, J., Tzafriri, L., Classical Banach Spaces I, Springer, Berlin, 1977.
-
Luschgy, H., Pages, G., Functional quantization of Gaussian processes, J. Funct. Anal., 196, 2002, 486-531.
-
Luschgy, H., Pages, G., Sharp asymptotics of the functional quantization problem for Gaussian processes, Ann. Probab., 32, 2004, 1574-1599.
-
Luschgy, H., Pages, G., Functional quantization of 1-dimensional Brownian diffusion processes, Preprint, 2003.
-
Parna, K., On the existence and weak convergence of
-centers in Banach spaces, Tartu Ülikooli Toimetised, 893, 1990, 17-287.
-
Pages, G., Printems, J., Optimal quadratic quantization for numerics: the Gaussian case, Monte Carlo Methods and Applications, 9, 2003, 135-166.
-
Pages, G., Printems, J., Functional quadratic quantization for pricing derivatives, Preprint, LMPA-900, Univ. Paris 6, 2004.
-
Samorodnitsky, G., Taqqu, M.S., Stable Non-Gaussian Random Processes, Chapmann & Hall, 1994.
-
Schaefer, H.H., Banach Lattices and Positive Operators, Springer, Berlin, 1974.