1. Introduction
Let
be a measure space with
. Suppose we are given a symmetric Dirichlet form
in
. The associated Markovian semigroup is denoted by
and we assume that
. Here, 1 stands for a constant function of
M, taking the value 1. For any
, we use the notation
We also assume that 1 is the unique invariant function for the semigroup
. Then, as
, we have
in
. The semigroup
is called ergodic when Equation (
2) holds. We define the index
by
which is often called the spectral gap (see e.g., Theorem 4.2.5 of [
1]). Here,
m stands for a linear operator
and
stands for the operator norm from
to
. In connection to this index, we are interested in another index
defined by
Here,
Z is the Zygmund space (sometimes denoted by
). The space
Z is defined as follows. Set
and
. Then
We introduce norms in
Z later (see
Section 2).
On the other hand, the logarithmic Sobolev inequality is a powerful tool in the analysis of Markovian semigroups. The inequality takes the following form:
Here,
stands for the
-norm and the constant
is chosen to be maximal and is called the logarithmic Sobolev constant. The form of the inequality reminds us the notion of entropy:
An important application of the logarithmic Sobolev inequality is the following estimate of the entropy (see e.g., Chapter 6.1 of [
2]):
We are interested in the relation between
and
. In fact, we show the inequality
. This kind of estimate of
is given in [
3], but in this paper we give a direct connection to the constant
.
The organization of this paper is as follows. In
Section 2, we give several kinds of norms in the Zygmund space
Z. Using these notions, we show relations between the entropy and the norm in the space
Z and give a proof of the main result. As an example, we discuss the Laguerre operator in
Section 3. We give a precise expression of the resolvent kernel. In
Section 4, we introduce Orlicz spaces (
). We also discuss how to show the boundedness of operators in Orlicz spaces. Using these, we investigate the spectrum of the Laguerre operator in Orlicz spaces in
Section 5. We can completely determine the spectrum and can see the relation between the spectral gap and the logarithmic Sobolev constant.
2. Entropy and the Zygmund Space
2.1. The Zygmund Space
We start with the Zygmund space. Let be a measure space, and we assume that , i.e., m is a probability measure. All functions in the paper are assumed to be -measurable. We denote the integration with respect to m by . Of course, we assume the integrability of a function f. We also use the notation for .
The Zygmund space is the set of all measurable functions
f with
. We denote it by
Z or
. We can define a norm in this space. To do this, we introduce a function
on
defined by
and further, we define
is a convex function. Now, define
by
This norm is sometimes called the Luxemburg norm (see e.g., [
4]). The norm of the constant function 1 can be computed as
Z becomes a Banach with the norm .
The dual space of
Z is given as follows. Let
be the inverse function of
, i.e.,
The dual space of
Z can be identified with the space of all measurable functions
f with
for some
(see [
4]).
The following inequality is fundamental:
Using this, we can show that
In fact, if
, we have
The right-hand side takes its minimum
when
. Hence, we obtain Equation (
13). This shows that
.
2.2. Entropy
Now, we recall the notion of entropy. In this section, all functions are taken from
Z. For any non-negative function
f, the entropy
is defined by
We will discuss the relation between and . First, we show the following.
Proposition 1. For any non-negative function f, we have Proof. We note the following inequality
Using this, we can get
which is the desired result. ☐
If, in addition, we assume , we can get another estimate.
Proposition 2. If a non-negative function f satisfies ,
then we have Proof. Let us show the inequality
for any
. Set
(1) The case .
Hence,
. By differentiating the function
F, we have
and so we easily see that
. The second-order derivative is given by
Thus, we have for .
(2) The case .
In this case, we have
So
is clear. The derivative of
F is
and so we easily see that
. Furthermore, we have
Thus, we have and for .
Using the inequality Equation (
18), we have
which completes the proof. ☐
Now we are ready to show that the -norm is dominated by the entropy.
Proposition 3. For any non-negative function f, we have Proof. We note that since
is convex and
,
satisfies that for
and, for
The proof of (
19) is divided into two cases.
(1) The case .
Set
. Applying Proposition 1 to the function
, we have
We used Equation (
20) in the second line. Thus we have, in this case,
(2) The case .
Set
. Since
, we can apply Proposition 2 to
and obtain
Now,
follows since the left-hand side equals 1. Hence,
Combining both of them, we have
which completes the proof. ☐
In turn, we prove the inequality of the reversed direction.
Proposition 4. We have the following inequality:
Proof. Here,
. In fact, Equation (
23) is clear and so we only show Equation (
24). For
, we have
It is not hard to show that
for
. Hence, we have Equation (
24).
Since
, we have, for all
,
Substituting
in it and integrating both hands, we have
This completes the proof. ☐
When , we can show another inequality.
Proposition 5. If we assume ,
then we have Proof. Since
, we easily see
Using this, we have
Integrating both hands, we have
Now set
on
. Then,
for
. Hence,
g takes its maximum at
. Therefore,
Thus we get the desired result. ☐
We are ready to show that the entropy is dominated by the -norm.
Proposition 6. We have the following inequality:
Proof. Since the function
is convex,
satisfies following inequality. For
, we have
and, for
,
The proof is divided into two cases.
(1) The case .
For notational simplicity, we denote
by
N. Using Proposition 4 for
,
Here we used Equation (
29) in the second line.
(2) The case .
This time we use Proposition 5 and obtain
Since
, we have Equation (
27).
Let us recall the logarithmic Sobolev inequality:
which yields the following entropy estimate:
Now, we are in a position to prove the following main theorem.
Theorem 1. We have the following inequality.
Proof. We may assume
. Let
f be a non-negative function. If
, then we have for sufficiently large
tNext, we take a general
f. If
, then
),
NΦ(
f) ≤
NΦ(|
f|) =
NΦ(
f) ≤ 1 and so
Hence, this completes the proof. ☐
In Theorem 1, we have shown that
. We now connect the Logarithmic Sobolev constant
and the spectral gap. Let us denote the set of spectrum of
in the Zygmund space
Z by
. Then, the following inequality is known (see e.g., Chapter IV, Proposition 2.2 of Engel-Nagel [
5])
If we restrict ourselves to the mean 0 functions, we have
Now we set
and call it the spectral gap in
Z. Hence, we have the following inequalities:
Example 1. Let us consider the Ornstein–Uhlenbeck operator on
. The reference measure is
and the Dirichlet form is given by In this case, it is known that .
Moreover is an eigenfunction for the eigenvalue .
Hence, we havewhich shows .
In Section 5, we will give an example that holds.
3. Spectrum of the Laguerre Operator
In this section, we give an example.
3.1. The Laguerre Operator
We consider the following operator:
Since eigenfunctions of
are Laguerre polynomials (see e.g., Lebedev [
6]), we call the diffusion process generated by
the Laguerre process as in [
7]. It is also an interest rate model called the Cox-Ingersoll-Ross process in mathematical finance.
We assume that
. This is necessary to ensure that the invariant measure becomes a probability measure. Actually, the invariant probability measure is given by
which is the gamma distribution of the parameters
, 1.
There is another characterization of a diffusion process by a speed measure and a scale function. In our case, setting
the speed measure is
and the scale function
s is determined by
. Following Feller, the boundary 0 is classified as a non-exit, an entrance when
and exit, and an entrance when
. We impose the Neumann boundary condition when
to ensure that the associated diffusion process is conservative.
We can give the associated Dirichlet form
as
Here,
. Therefore,
It is well-known that the set of the spectrum of in is and eigenfunctions are Laguerre polynomials. We also give an expression of the resolvent. To do this, we need confluent hypergeometric functions.
3.2. Confluent Hypergeometric Functions
We recall confluent hypergeometric functions (see, e.g., Beals-Wong [
8] or Lebedev [
6]). They are defined by
Here,
is the Pochhammer symbol, i.e.,
A function defied by Equation (
38) converges for all
and is an analytic function. This function satisfies the following differential equation:
This equation is called the Kummer equation (or the confluent hypergeometric equation), and, of course, is closely related to our generator
in Equation (
35). Our interest is in the spectrum of
, and so confluent hypergeometric functions are candidates of eigenfunctions. If
belongs to
, it is an eigenfunction and it is so when
,
. In this case,
(−
n;
c;
x) is nothing but a Laguerre polynomial (up to constant) and is an eigenfunction. For simplicity, we introduce the following notation:
By the way, Equation (
42) is a second-order differential equation; there is another independent solution, which is given by
This function is called a confluent hypergeometric function of the second kind. Their Wronskian is
The Laguerre polynomial is written as
Our parameter
is chosen to be consistent with the parameter of the Laguerre polynomial. The asymptotic behavior of these function is given as follows (see e.g., Lebedev [
6]). When
,
However, when , should be replaced by .
Here, we assumed a, .
Now we can give an expression of the resolvent. Recall that we assumed
and
. The resolvent
has the following kernel expression:
where
Here,
W stands for the Wronskian in Equation (
45) and
. Hence, we have
is a bounded operator in if . We will discuss later what happens in the Zygmund space.
3.3. The Logarithmic Sobolev Inequality
We show that the logarithmic Sobolev inequality holds for the Laguerre operator
. You can also see the result in Chapter 2.7.3 of [
1] from the view point of the curvature dimension condition. Recall that the Dirichlet form associated with
is given by Equation (
39).
Theorem 2. We assume that .
Then, the following logarithmic Sobolev inequality holds for the Dirichlet form in (
36):
Proof. It is enough to check Bakry-Emery’s
-criterion. It is as follows. From Equation (
36), the square field
is given by
The generator is
. Hence, the
is computed as
From this we have under the condition . Due to Bakry-Emery’s -criterion, this implies that . ☐
Taking , we can see that is the best constant.
Remark 1. This result was shown in Korzeniowski-Stroock [
7]
when .
In that paper, it was emphasized that the logarithmic Sobolev constant differs from the spectral gap.
4. Orlicz Space
We start with the definition of the Orlicz space. Take any
and fix it. We introduce a norm in the space of all functions
f with
. Define a function
on
by
is a concave function. To get the behavior of
at
∞, we use the l’Hospital theorem and get
Therefore, when
, we can see
We define the space
by
Then,
becomes a Banach space with the norm
defined by
For instance, the norm of the constant function 1 is
If , then . In the sequel, the operator norm of linear operators from into is defined by using the norm .
4.1. Dual Space
The dual space of
is characterized as follows. Let
be the inverse function of
:
The Orlicz space associated with is the dual space of . Let us study the asymptotic behavior of at .
Proof. We use the l’Hôspital theorem.
Equation (
63) easily follows from this. ☐
The following Hausdorff-Young inequality plays a fundamental role in the later computation.
For example, if
, then we can show that for
This shows that
and there exists a constant
so that
4.2. Linear Operators in Orlicz Spaces
Orlicz space is a Banach space with the norm . The operator norm can also be defined in terms of this norm. However, since this norm is hard to calculate concretely, we take another way.
We introduce a new norm
, which is called the Orlicz norm, by
Here,
g runs over all functions satisfying
. Replacing
f with
,
Hence, we can rewrite Equation (
65) as follows:
We will rewrite the condition .
Proof. From Proposition 7, we have
We can take large constant
so that
Therefore, if
, then we have
which is the desired result. ☐
Then, by Proposition 8, we can see that if .
It is well-known that two norms,
and
, are equivalent (see e.g., p. 61, Chapter III. 3.3, Proposition 4 of Rao-Ren [
4]):
From this, we have the following
Proposition 9. A linear operator T on is bounded if there exist positive constants A and B so that Proof. If we assume Equation (
68), then we have
which implies
. The rest is easy from Equation (
67).
Corollary 1. Let be a constant defined by (
65).
Then a linear operator T on is bounded if there exist constants A and B so that for any non-negative function g with and any non-negative function ,
we have Proof. From Equation (58), we have
and hence we can find constants
a and
b so that
Now, from Proposition 9, T is bounded. ☐
We list up some inequalities which are necessary later. For
x,
,
There exists a positive constant
so that
This inequality is a modification of the following Hausdorff-Young inequality:
5. The Spectrum of the Laguerre Operator in
The kernel representation of the resolvent of Laguerre operator was given in Equation (
53). It is bounded in
. We will examine whether it is bounded in
. Recall that our reference measure is
. We assume
. From now on, we ignore the constant and consider the measure
.
We take any
. We also take a non-negative function
g satisfying
. Our aim is to show that there exist constants
A and
B so that
. The integrability is important, and we do not need the precise constant. Hence, we use the following notation:
This means that there exist constants
k and
l so that
Here, constants k and l are independent of functions f and g. This is important but we do not mention this each time.
We starts with an estimate of the defective Gamma function.
Proposition 10. Take any .
If ,
thenIf ,
then there exists a constant so that These are easily obtained by seeing that as .
Proposition 11. Assume that and .
Then, there exists a constant C depending on κ and λ so that Proof. This inequality can be reduced to the previous proposition. By the change of variable formula, we have
This completes the proof. ☐
We study integrals involving function g.
Proposition 12. For any ,
and ,
there exist constants ,
so that We have assumed that ,
so we have that there exists a constant c depending only on k, β, and α so that Proof. Set
,
. Then
F takes its maximum at
and for
,
F is decreasing. Hence, if
, then
and if
, then for
, we have
Hence, there exists a constant
depending on
k such that for
,
This completes the proof. ☐
Proposition 13. Assume .
Then, there exists a constant C so that Recalling that ,
we have When ,
we have Again by ,
Proof. By using Equation (
70), we have
When , in the computation above, we just need to note that the primitive function of is . ☐
Of course, when
, the left-hand side of Equation (
77) is bounded.
We have seen the asymptotic behavior of integrals as . We can also get the asymptotics as . We will prove this by reducing to the previous case.
Proposition 14. Suppose that ,
.
Then, there exist constants ,
so that By using ,
we have Proof. For
, we have
which is the desired result. ☐
Lastly, we show the estimate involving the function f.
Proposition 15. We have the following inequality for f.
Proof. From Young’s inequality
☐
We now investigate the spectrum. Let us start with the point spectrum.
Theorem 3. The point spectrum of is . Here, stands for the integer part of β.
Proof. We show that
a is an eigenvalue of
if
. We have seen that
satisfies the differential equation
(see Equation (
42)). We only need to show that
. The integrability of
on
is trivial since
M is bounded on
. We see the integrability on
:
The finiteness in the last line follows from .
It remains to be shown that () is an eigenvalue. In fact, is an polynomial of order n (a Laguerre polynomial up to normalization) and hence the integrability of follows easily.
Theorem 4. If , then a belongs to the resolvent set except when , , , .
Proof. We show that
in (
51) is bounded in
. To do this, we use Corollary 1. We recall the kernel
.
(1) The case .
Let us consider the resolvent kernel
(recall Equation (
53) but we ignore a constant multiplication). In the region of
, we have
When
,
in the above computation should be changed as follows:
In the region of
, we have
(2) The case .
We consider the resolvent kernel
. In the region
, we have
In the region
, we have
Thus, we have shown that is bounded in . Hence, the spectrum of is completely determined. ☐
The spectrum is shown as in
Figure 1. The case of
is the Zygmund space. Hence, the spectral gap equals 1. So we have
where
is defined by Equation (
33). Therefore, by Equation (
34), we have
and so it shows that
may happen. This is a well-known result in the case of
. Furthermore, in [
9], we have shown that assuming the logarithmic Sobolev inequality, the spectra in
(
) are all the same. In our case, the spectrum in
is 0,
,
The spectrum in
is quite different from that. Moreover, the logarithmic Sobolev inequality may not give a uniform estimate for spectral gaps among the Orlicz spaces
(
).