Fundamental Solution to 1D Degenerate Diffusion Equation with Locally Bounded Coefficients
Abstract.
In this work we study the degenerate diffusion equation for , equipped with a Cauchy initial data and the Dirichlet boundary condition at . We assume that the order of degeneracy at 0 of the diffusion operator is , and both and are only locally bounded. We adopt a combination of probabilistic approach and analytic method: by analyzing the behaviors of the underlying diffusion process, we give an explicit construction to the fundamental solution and prove several properties for ; by conducting a localization procedure, we obtain an approximation for for in a neighborhood of 0 and sufficiently small, where the error estimates only rely on the local bounds of and (and their derivatives). There is a rich literature on such a degenerate diffusion in the case of . Our work extends part of the existing results to cases with more general order of degeneracy, both in the analysis context (e.g., heat kernel estimates on fundamental solutions) and in the probability view (e.g., wellposedness of stochastic differential equations).
Key words and phrases:
degenerate diffusion equation, locally bounded coefficients, estimates on the fundamental solution, generalized Wright-Fisher equation, stochastic differential equation, wellposedness of stochastic differential equations2000 Mathematics Subject Classification:
35K20, 35K65, 35Q921. Introduction
In this article we consider the following Cauchy initial value problem with the Dirichlet boundary condition:
(1.1) |
where and .
Set .
We further impose the following assumptions on
and :
-
(H1):
, for every and .
, when , and when .
, , , and are all bounded on for every . -
(H2):
There exists such that for every ,
Our goal is to construct and to study the fundamental solution to (1.1), with a particular emphasis on the behaviors of when are near the boundary and when is small. Since is degenerate at , standard methods on strictly parabolic equation no longer apply in this case, and the degeneracy of does have an impact on the regularity of . Moreover, the assumptions (H1) and (H2) only guarantee local boundedness of , and their derivatives, so we need to conduct our analysis of only relying on the local bounds of the coefficients.
1.1. Background and motivation
Our work is primarily motivated by two previous works [8] and [7] on related problems. [8] treats the initial/boundary value problem for a degenerate diffusion equation similar to the one in (1.1), but under stronger conditions on the coefficients. To interpret the hypotheses adopted in [8] in terms of , and in our setting, we define the following two functions for :
(1.2) |
where is the constant such that . It is assumed in [8] that , as well as
(1.3) |
Under the assumption (1.3), through a series of transformations and perturbations, [8] completes a construction of the fundamental solution to (1.1), conducts a careful analysis of the regularity properties of near the boundary, and derive an approximations for in terms of explicitly formulated functions. In particular, if denotes the approximation for , then it is proven in [8] that there exists a constant , universal in all in a neighborhood of and all sufficiently small, such that
(1.4) |
Such an estimate is useful in multiple ways. First, while one expects to resemble the fundamental solution to a strictly parabolic equation for away from the boundary, (1.4) captures accurately the asymptotics of when are close to the boundary, and demonstrates the influence of the degeneracy of on . Second, if one could apply the general heat kernel estimates (see, e.g., of [32]) to , then one would get that for every and sufficiently small, there is constant such that
(1.5) |
where is the distance between and under the Riemannian metric corresponding to ; it is clear that (1.4) is a sharper estimate than (1.5) for small , and hence is a more accurate short-term approximation for compared with the general heat kernel approximation. In addition, in [8], is presented in an explicit formula (in terms of special functions) and “” in (1.4) can be replaced by an exact expression; therefore, (1.4) is easily accessible in computational applications that involve the fundamental solution to any degenerate diffusion equation in the form of .
We aim to generalize the results in [8], particularly the construction of and the short-term near-boundary approximation , to a more general family of degenerate diffusion equations. The hypotheses (H1) and (H2) proposed above are more relaxed compared with the assumption (1.3) adopted in [8]. For example, it can be checked with direct computations that in general (1.3) does not hold if , which means that, given (H1) and (H2), (1.3) is only satisfied when . Moreover, (1.3) clearly imposes strong global conditions on and , but with (H1) and (H2), we have to find an access to without relying on any global bound on the coefficients. To tackle this issue, we invoke a “localization” procedure, as inspired by [7].
[7] studies the following well known Wright-Fisher diffusion equation, which has its origin in population genetics:
(1.6) |
Different from (1.1), (1.6) has two-sided Dirichlet boundaries at 0 and 1, and the diffusion operator degenerates linearly at both boundaries. Set and let be the fundamental solution to (1.6). Since (1.6) is symmetric on , to study near the boundaries, it is sufficient to only consider the left boundary 0. In [7], a “localization” method is devised to construct near 0: since can be viewed as the density of the underlying diffusion process corresponding to , we can acquire information on by studying the behaviors of the process near 0; in particular, by tracking the excursions of the diffusion process near 0, we can “localize” and within a neighborhood of 0 where only the degeneracy at 0 has a substantial impact. Heuristically speaking, when restricted near 0, is close to the operator , and hence it is natural to expect that with near 0 is close to the fundamental solution to (with Dirichlet boundary 0). Indeed, it is established in [7] that, not only can be constructed in an explicit way via , is also well approximated by in the sense that satisfies (1.4) for near 0 and sufficiently small. In our work we want to adopt a similar localization procedure and start our investigation of (1.1) on a bounded set where the local bounds of the coefficients would be sufficient for our purposes.
In addition to treating directly the fundamental solutions, degenerate diffusion equations in the form of have also been discussed in many other contexts, with most of the existing literature concerning the case when . For example, Epstein et al ([15, 16, 17, 18, 19]) conduct an comprehensive study of the generalized Kimura operators, which can be viewed as a generalization of with in the manifold setting, obtaining results such as the Hölder space of the solutions, the maximum principle and the Harnack inequality. Related works on generalized Kimura diffusions include [18, 19, 30, 31]. From a probabilistic view, there are abundant theories on existence and uniqueness of solutions to stochastic differential equations with degenerate diffusion coefficients (see, e.g., [10, 20, 27, 28, 34, 36, 39] and the references therein); when , a series of works (see, e.g., [1, 3, 4, 9, 14, 38]) provide conditions on and that are sufficient for the stochastic differential equation corresponding to to be well posed, and some of the results will also be used later in our discussions.
1.2. Our main results
Our strategy in solving (1.1) and getting
is to combine the ideas and the techniques
from [8] and [7], and tackle the
two challenges we face: general order of degeneracy in at the
boundary, and lack of global bounds on the coefficients. Below we
briefly describe the main steps we will take to complete this work
(see Table 1 for an illustration).
1. Localization and transformation (). Since the coefficients in are locally bounded, we first consider a “localized” version of (1.1). Given , we study the diffusion equation in (1.1) on with an extra Dirichlet boundary at , i.e.,
Let be the fundamental solution to . To solve , we carry out a transformation that turns into a diffusion operator that degenerates linearly at . In fact, with a change of variable , solving becomes equivalent to solving the following problem:
where is the image of after the change of variable,
is a constant, and is a function on
(, and will be specified in ). If
we can find the fundamental solution to , denoted
by , then
can be obtained through via the transformation
(and its inverse) between and .
2. Model equation (). Our strategy for solving is to treat the operator as a perturbation of . We temporarily return to the “global” view, omit the potential and the right boundary , and consider the following model equation on the entire :
This model equation has the advantage that its fundamental solution
has an explicit formula in terms of a Bessel
function ([8]), and properties of the solutions
to the model equation are already known to us (Proposition2.4).
With in hand, we return to the local view of
the model equation (with the Dirichlet boundary condition “restored”
at ) and derive the fundamental solution
to the localized model equation on (Proposition
2.6).
3. Solving the localized equation ().
Upon getting , we can start the construction
of the fundamental solutions to and .
Viewing as a
perturbation of with a potential
function , we invoke Duhamel’s perturbation method
to construct using
as the “building block” (Proposition 3.2).
Although in general does not have a
closed-form formula and our representation of
is in the form of a series, by focusing on the first term of the series
expression we can show that is well
approximated by for sufficiently small
(Proposition 3.4).
4. Solving the global equation (). We finally return to (1.1) and produce from by “reversing” the localization procedure. More specifically, we establish the relation between (1.1) and its localized version with the help of the underlying diffusion process corresponding to . By analyzing the excursions of the diffusion process over , is achieved as the limit of as increases to infinity (Theorem 4.3). Again, although does not have a closed-form formula, we find an approximation for such that has an explicit and relatively simple expression, and is more accurate than the standard heat kernel estimate for (Theorem 4.5).
(transformation) | |||
localized equations: | |||
(convergence)(localization) | (perturbation) | ||
- - - - - - - - - - - - | - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - | ||
global equations: | (localization) | ||
(approximation) | |||
(transformation) |
In each of the steps above, in addition to the standard analytic methods from the study of parabolic equations, we also rely on a probabilistic point of view towards diffusion equations. Whenever applicable, we treat the fundamental solution as the transition probability density function of the underlying diffusion process corresponding to the concerned operator. In fact, the localization procedure (and the reverse of it) proposed above is possible because of the (strong) Markov properties of the diffusion process. We also invoke some classical tools in the study of stochastic processes, e.g., Itô’s formula and Doob’s stopping time theorem, in deriving probabilistic interpretations of the (fundamental) solutions to the involved diffusion equations. In we give a brief overview of the probabilistic components involved in this work.
In we consider a generalization of the classical Wright-Fisher equation (1.6), where we assume that the diffusion operator vanishes with a general order at both of the degenerate boundaries and . In particular, for and , we consider the equation
Although this problem is in a different setting from (1.1), our methods and results still apply. We can follow the same steps as above to study its fundamental solution and obtain similar estimates for near either of the boundaries (Proposition 5.1).
1.3. Stochastic differential equation, underlying diffusion process
This subsection gives a brief overview of the probabilistic foundation needed for our investigation. We start with the stochastic differential equation corresponding to the operator , and that is, given ,
(1.7) |
For a general stochastic differential equation, there are two notions of existence/uniqueness of a solution : strong existence/uniqueness and weak existence/uniqueness. Our work only requires the existence of a weak solution to (1.7) and the solution being unique in the weak sense. We will not expand on the general theory and refer interested readers to [25, 35] for a comprehensive exposition on these topics.
We say that (1.7) has a (weak) solution if, on some filtered probability space , there exist two adapted processes and such that, (i) is a standard Brownian motion; (ii) has continuous sample paths; (iii) almost surely satisfies that
In this case, we also refer to as the underlying diffusion process corresponding to starting from . We say that a solution is unique (in law), if whenever (i)-(iii) are satisfied by another triple , and , it must be that the distribution of under is identical with that of under . We say that the stochastic differential equation (1.7) is well posed if a solution exists and is unique.
In later discussions we will use an important corollary of the wellposedness property, and that is, if (1.7) is well posed and is the unique solution, then is a strong Markov process and for every ,
is a local martingale.
Now let us examine the wellposedness of (1.7) under the hypotheses (H1) and (H2). There is a rich literature on the wellposedness of a degenerate stochastic differential equation with a diffusion operator that degenerates linearly. While the diffusion coefficient in (1.7) may have nonlinear degeneracy, we can convert it into a linear degeneracy case simply through a change of variable. To be specific, we consider the following diffeomorphism on and its inverse:
(1.8) |
One can easily verify that is a solution to (1.1) if and only if is the solution to
(1.9) |
where , and
The stochastic differential equation corresponding to (1.9) is that, given ,
(1.10) |
Assuming (H1) and (H2), we get down to verifying the wellposedness of (1.10) where the diffusion operator degenerates linearly at . First, when , by (H1), (1.8) and direct computations, we see that both and are Lipschitz continuous on any bounded subset of . Furthermore, (H2) and (1.8) imply that there exists constant such that for every ,
(1.11) |
It follows from classical results (e.g., Yamada-Watanabe [38], Stroock-Varadhan [35], Engelbert-Schmidt [14], Cherny [9]) that (1.10) is well posed for every in this case. Next, when , we note that
This time (H1) and (1.8) guarantee that and are both Hölder continuous on any bounded subset of ; meanwhile, the growth control (1.11) on and still applies. Thus, the results of Bass-Perkins [3] lead to the wellposedness of (1.10) for every . Therefore, for every , (H1) and (H2) are sufficient for (1.10) to be well posed. Assume that is the unique solution to (1.10). By setting
we immediately get the following conclusion.
Lemma.
So far there is no constraint on the behavior of at the boundary 0. Returning to the original problem (1.1), to incorporate the Dirichlet boundary condition, we only need to focus on up to the time it hits . Intuitively speaking, if we set
then the probability density function of the conditional distribution of given should coincide with the fundamental solution to (1.1).
Notations.
For , we write and .
For every , denotes the indicator function of .
Let be a filtered probability space. For an integrable random variable on and a set , we write . For an adapted process with non-negative continuous sample paths, we set
i.e., is the hitting time at conditioning on the process starting from ; for , we set .
2. Model Equation
In this section, we will carry out the first three steps outlined in . Although we use similar transformations as those in [8], we need to adapt the method so that it applies to and that are under weaker conditions.
2.1. Localization and Transformation
Let and be fixed throughout this section. Our first step is to introduce an extra Dirichlet boundary to the equation at and to consider a localized version of (1.1) on . Namely, given , we look for such that
(2.1) |
Once restricted on , the coefficients (and their derivatives) in (2.1) are all bounded.
We want to find the fundamental solution to (2.1). Given , let be the unique solution to (1.7), as found in . We expect that coincides with the probability density function of , conditioning on . This probabilistic interpretation of is indeed correct and will be justified later. For now, let us conduct an analysis of (2.1) via standard perturbation methods.
As mentioned in , we will transform (1.1) into a diffusion equation that has linear degeneracy at . For , let and be defined as in (1.2). It is clear that , is strictly increasing, and . The constant in the definition of is chosen such that
Under (H1) and (H2), it is easy to verify that
and hence . Let , be the inverse function of and . We introduce two more functions on :
(2.2) |
and
or equivalently,
(2.3) |
Now we are ready to state the result on the transformation.
Proposition 2.1.
Given , we define
Then, is a solution to (2.1) if and only if
(2.4) |
where is a solution to the following problem:
(2.5) |
We omit the proof of Proposition 2.1 since it can be verified by direct computations. If is the fundamental solution to (2.5), then is connected with following the same relation as the one in (2.4). Set . Compared with , has a simpler structure consisting of a linear diffusion, a constant drift and a potential. In the next subsection we will solve (2.5) by treating as a perturbation of and invoking Duhamel’s perturbation method. As a preparation, we state below some technical results on and .
Lemma 2.2.
Let be defined as in (2.2). Then, for every ,
(2.6) |
Hence, there exists constant that can be made explicit (see (6.1) in the Appendix) such that
(2.7) |
Let be defined as in (2.3). Then, there exists constant such that for every ,
(2.8) |
The proof of Lemma 2.2 is left in the Appendix since it is based on straightforward computations that are lengthy and not crucial to our work. We note that when , the potential function may be singular at . This is a generalization of the case considered in [8] where is assumed to be bounded near 0.
2.2. From to
Let and be the same as above. As mentioned in the previous subsection, to solve (2.5), we will first consider the analogous problem with replaced by . Namely, given , we look for such that
(2.9) |
Let be the fundamental solution to (2.9).
We consider as our model equation. To solve (2.9), we temporarily return to the “global” view and study the model equation on instead of . That is, for , we consider the following problem:
(2.10) |
Let be the fundamental solution to (2.10). In fact, is the starting point of our “journey”, and from we will derive the (fundamental) solutions to all the concerned equations.
The stochastic differential equation corresponding to the model equation is that, given ,
(2.11) |
It follows from the discussions in that (2.11) is well posed, and hence there exists a unique solution that is also a strong Markov process.
Remark 2.3.
We want to remark that, independent of the Dirichlet boundary condition imposed in (2.10), the constant determines the attainability of the boundary . Under (H1) and (H2), we have that , and hence is either an exit boundary or a regular boundary. This is to say that, no matter what is, hits with a positive probability in finite time. For more details on the topic of boundary classification, we refer readers to of [26].
The operator , as well as (2.10) and (2.10), has been well studied in [8]. Below we will review some useful facts about , and their connections to . The details can be found in of [8].
Proposition 2.4.
(Proposition 2.1, 2.3 of [8]) The fundamental solution to (2.10) is
(2.12) |
for , where is the modified Bessel function. is smooth on , and for every ,
(2.13) |
and
(2.14) |
Given , if
then is the unique solution in to (2.10), and is smooth on . Moreover,
(2.15) |
which implies that for every Borel set ,
(2.16) |
Finally, satisfies the Chapman-Kolmogorov equation, i.e., for every and ,
(2.17) |
It is clear from (2.16) that, for every , is the probability density function of , provided that . Now we turn our attention to , the fundamental solution to (2.5) which has an extra Dirichlet boundary at . Intuitively speaking, to get , we need to remove from the “contribution” of once exists the interval . Based on this idea combined with the fact that is a strong Markov process, we define
(2.18) |
for every . Again, by (2.16), we see that for every Borel set ,
(2.19) |
In other words, is the probability density function of provided that .
To better analyze , we need the following probability estimates on the hitting times of .
Lemma 2.5.
For every ,
(2.20) |
For and , we have that
(2.21) |
Furthermore, almost surely
Proof.
Based on (2.11), one can apply Itô’s formula (see, e.g., of [25]) to check that, for every , is a local martingale, and hence by Doob’s stopping time theorem (see, e.g., of [33]), is a bounded martingale. Thus,
To show (2.21), we check that for every and every , if
then is a martingale. By a similar argument as above and Fatou’s lemma, we get that
(2.22) |
Set . Since , we have that and . Therefore, a simple application of Markov’s inequality leads to
Plugging the value of into the right hand side yields (2.21). The fact that converges to as almost surely follows from (2.21) and the monotonicity of in .
Finally, we observe that exists almost surely. Take such that . It follows from (2.22) and the reverse Fatou’s lemma that
which implies that and hence almost surely. ∎
Proposition 2.6.
Let be defined as in (2.18). Then, is continuous on , and for every , we have that
(2.23) |
satisfies the Chapman-Kolmogorov equation, i.e., for every and ,
(2.24) |
For every , is a smooth solution to the Kolmogorov backward equation corresponding to :
(2.25) |
for every , is a smooth solution to the Kolmogorov forward equation corresponding to :
(2.26) |
where is the formal adjoint of .
Proof.
We start with (2.24) since its proof is straightforward. Given , and Borel set , by (2.19) and the strong Markov property of , we can write
which implies (2.24). Next, given , we take any , any and . By (2.14), (2.19) and, again, the Markov property of , we have that
Since is almost surely continuous and are chosen arbitrarily, the relation above implies that for every measurable functional on ,
where for every . In particular, for arbitrary , , if is chosen such that for every ,
then we have that
This is sufficient for us to conclude (2.23).
Now we turn attention to (2.25) and (2.26). By (2.23), it suffices to prove only one of them, say, (2.26). To this end, we take and consider , which, according to (2.19), can be written as
As reviewed in , for every ,
is a bounded martingale. Thus,
and hence
This means that for every , solves the equation in the sense of distribution. Since is a hypoelliptic operator (see, e.g., of [32]), is a smooth solution to (2.26).
For , we set
(2.28) |
Then, for every , is smooth on . It is easy to see that is equicontinuous in from any bounded subset of , which implies that , as well as , is continuous on .
We proceed to the proof of the last statement. Again, by the hypoellipticity of , to show that is a smooth solution to the model equation, we only need to show that it solves the equation as a distribution. Let us take and consider, for every ,
By (2.25), we have that
The only remaining thing to do is to verify that satisfies the initial value and the boundary value conditions in (2.9). Given , by (2.15) and (2.19), we have that for every ,
which, according to (2.21), goes to as , and the convergence is uniformly fast for on any compact subset of . Therefore,
To verify that satisfies the boundary condition, it is sufficient to show that
We observe that, by (2.13), is bounded uniformly in by
where we used the fact that
Therefore, (2.20) implies that
Finally, the last statement in Lemma 2.5 and the dominated convergence theorem lead to
∎
We will close this subsection with a result on the comparison between and . Intuitively speaking, given sufficiently far from the boundary and sufficiently small, would not have exited by time , which means that and should be close to each other. We will make this statement rigorous by proving that, as , converges to 1 uniformly fast in away from .
Corollary 2.7.
Set . Then, for every ,
(2.29) |
3. Localized Equation
3.1. From to
Now we get down to solving (2.5) by the perturbation method of Duhamel. First we want to find a function on that solves the integral equation
(3.1) |
for every , and then verify that is the fundamental solution to (2.5). To this end, for every and , we define
(3.2) |
To state the technical results on , we need to introduce more notations. Set
(3.3) |
We have that , and if is the constant found in Lemma 2.2, then (2.8) can be rewritten as
For and , we define
(3.4) |
It follows from a simple application of Stirling’s formula that is summable in , and hence is well defined.
Lemma 3.1.
There exists a universal constant such that for every and ,
(3.5) |
and hence
(3.6) |
is well defined as an absolutely convergent series. Moreover, for every ,
(3.7) |
and satisfies (3.1).
Proof.
Without causing any substantial change, we will assume that is defined on with for When , since is bounded on with , (3.5)-(3.7) can be derived in exactly the same way as in [8] (Lemma 3.4) with
There is nothing we need to do in this case. Hence, we will assume for the rest of the proof, and only treat the case when has a singularity at .
First, we claim that there exists a universal constant such that
(3.8) |
for every and . To see this, we use (2.12) and (2.14) to write the integral in (3.8) as
where for every ,
Interchanging the order of summation and integration yields
Since , we have that for ,
where (with ) is the beta function and
(3.9) |
Therefore, we have that
which confirms the claim (3.8).
To proceed, we notice that by Lemma 2.2, (2.17) and (3.2),
for every . Assume that up to some , for every ,
For , we have that
Upon rewriting as , we immediately obtain (3.5)-(3.7). Finally, (3.1) can be verified by plugging the series representation of into the right hand side of (3.1) and integrating term by term. ∎
We are now ready to solve (2.5).
Proposition 3.2.
Let be defined as in (3.6). Then, is continuous on , and for every , we have that
(3.10) |
also satisfies the following integral equation:
(3.11) |
Proof.
To prove (3.10), we first note that if, for and , we define
(3.13) |
then . In other words, (3.13) is an equivalent recursive relation to (3.2). To see this, one can expand both the right hand side of (3.2) and that of (3.13) into two respective fold integrals, and confirm that the two integrals are identical. Next, we will show by induction that for every and ,
When , this relation is simply (2.23). Assume that this relation holds up to some . By (2.23) and the equivalence between (3.2) and (3.13), we have that
(3.10) follows immediately. To establish (3.11), we write its right hand side as
where, again, we used the equivalence between (3.2) and (3.13). By (3.1) and (3.7), is continuous for every , and by (3.11), is equicontinuous in from any bounded subset of . From here one can easily derives the continuity of in on .
3.2. Approximation of
In general we do not expect to find a closed-form formula for , but when is sufficiently small, the above construction does provide accurate approximations for whose exact formulas are explicit or even in closed forms. Intuitively speaking, when is small, the effect of the potential in has not become “substantial” so that is close to , and hence it is natural to expect that is close to which, as we have seen in Corollary 2.7, is well approximated by for sufficiently small . To make it rigorous, we take to be the same as in Corollary 2.7 and use (2.29) and (3.5) to derive that for every ,
Hence, for some constant uniformly in ( may depend on and ),
(3.15) |
(3.15) confirms that when is small, is indeed well approximated by . However, viewing from (3.6), is only the “first order” approximation to , since the error in (3.15) is generated by keeping only the first term in the series in (3.6). It is possible to derive a more general “th order” approximation for with , and obtain an analog of (3.15) with the error being . To achieve this purpose, we introduce a new sequence of functions. For and , we set
(3.16) |
where, again, we assume that for . By following the proof of (3.5) line by line with replaced by , we also get that for every and ,
(3.17) |
Clearly, is the “global” counterpart of , and we will justify that is close to when is sufficiently small.
Lemma 3.3.
Proof.
When , (3.18) simply becomes (2.28). Assume that (3.18) holds up to some . Following (3.2) and (3.16), we write
(3.19) |
We use Fubini’s theorem and (2.28) to rewrite the first term on the right hand side of (3.19) as
(3.20) |
which, by (3.17), is bounded by
By (3.8) and the fact that
we can further bound (3.20) from above by
According to the inductive assumption, the second term on the right hand side of (3.19) is bounded by
which, by (2.14) and (2.23), is equal to
We use Fubini’s theorem again to rewrite the expression above as
Thus, combining the estimates of the two terms on the right hand side of (3.19), we obtain that
∎
Proposition 3.4.
Let . Then, for every and ,
(3.21) |
where
(3.22) |
In particular, there exists uniformly in and ( may depend on and ) such that
(3.23) |
3.3. From to
Now we are ready to return to the localized equation (2.1). Recall that , and are functions on defined by (1.2), and and are related by ; with , is the inverse function of , and is as defined in (2.2). Guided by Proposition 2.1, we define
(3.24) |
for every . We immediately obtain several results on based on Proposition 2.1 and Proposition 3.2. In addition, we can establish the connection between and the unique solution to (1.7) and underlying diffusion process corresponding to .
Proposition 3.5.
is the fundamental solution to (2.1). Given ,
(3.26) |
is the unique solution in to (2.1), and is smooth on . Moreover,
(3.27) |
and hence for every Borel set ,
(3.28) |
Finally, satisfies the Chapman-Kolmogorov equation, i.e., for every and ,
(3.29) |
Proof.
(3.25) follows directly from (3.10). Given , we set for . By (3.12), it is straightforward to check that
and hence it follows from Proposition 3.2 that is a smooth solution to (2.1). Since
is a bounded martingale, by equating its expectation at and , we obtain (3.27), which further leads to (3.28). Since is the unique solution to (1.7), is the unique solution to (2.1). Finally, (3.29) follows from (3.28) and the strong Markov property of . ∎
Remark 3.6.
The approximations we obtained in Proposition 3.4 for can also be “transported” to in a straightforward way. To see this, we define, for ,
(3.30) |
and more generally for ,
(3.31) |
Then Proposition 3.4 can be rewritten as follows.
Corollary 3.7.
Remark 3.8.
We want to point out that, from now on, whenever , the constants , and that were introduced in will also be written as , and respectively. In addition, by plugging into (2.3) and (2.6), we get that
(3.33) |
and
(3.34) |
With (3.34) and (3.33), it is possible to rewrite some of the expressions that appeared above (e.g., (3.24) and (3.25)) in a more explicit way, see, e.g., (6.2) and (6.3) in the Appendix. Especially when , these expressions take much simpler forms than in the general case, as we will see with a concrete example in .
4. Global Equation
In the previous section we have solved the localized equation (2.1) and obtained its fundamental solution . Now we proceed with the last step to complete our project, which is to build the “link” between (2.1) and the original problem (1.1). To achieve this goal, we rely on the strong Markov property of and the probabilistic interpretations of the solutions found in the previous sections.
4.1. From to
We introduce two more notations for this section: given ,
Our first task is to derive probability estimates for the hitting times of .
Lemma 4.1.
We define, for ,
(4.1) |
Then, for every ,
(4.2) |
if is the constant found in Lemma 2.2 (upon identifying with for ), then
(4.3) |
Moreover, for every , and such that , we have that
(4.4) |
Proof.
Next, we consider as a family parametrized by , and for every , we want to find out the link between and , i.e., the fundamental solutions to (2.1) with the right boundary at and respectively. To this end, we choose a third constant and define for each a sequence of hitting times of where and for ,
(4.7) |
In other words, the sequence records the downward crossings of from to . With the help of and the strong Markov property of , we are able to connect and as follows.
Proposition 4.2.
For ,
(4.8) |
Proof.
Given , we use (3.27) to write
According to the number of downward crossings (from to ) completed by , we further decompose as
By the strong Markov property of , we have that for each ,
On one hand, by (2.13), (3.7) and (3.24),
(4.9) |
On the other hand, if , then it must be that (i) , (ii) during the time interval , the process starts from and hits before , and (iii) for each , during the time interval , the process starts from and hits before . Hence, by (4.2) and the strong Markov property of , we have that
(4.10) |
Combining the above, we obtain that for every ,
This guarantees that the series in the right hand of (4.8) is absolutely convergent. ∎
Theorem 4.3.
For every , we set
Given , let be the sequence of hitting times defined as in (4.7) (for the downward crossings of from to ). Then, for every ,
(4.11) |
is continuous on , and for every ,
(4.12) |
For every , is a smooth solution to the Kolmogorov backward equation corresponding to , i.e.,
for every , is a smooth solution to the Kolmogorov forward equation corresponding to , i.e.,
is the fundamental solution to (1.1). Given ,
is the unique solution in to (1.1), and is smooth on . Moreover, for every ,
and hence for every Borel set ,
Finally, satisfies the Chapman-Kolmogorov equation, i.e., for every and ,
(4.13) |
Proof.
It is clear from (4.8) that for every , by taking , we know that the family is non-decreasing, so as the limit of (as ) is well defined. Since is the unique solution to (1.7), almost surely as (see, e.g., of [35]). Thus, (4.11) follows from (4.8) by sending to infinity, and (4.12) follows from (3.25).
Now we examine the continuity of . First, (4.9) and (4.10) guarantee that the series in the right hand side of (4.11) converges uniformly on any bounded subset of , from where it is easy to see that for every , is continuous on . Furthermore in the proof of Proposition 3.2 we have seen that is equicontinuous in from any bounded subset of , which, combined with (4.12), leads to the continuity of in all three variables.
Next, we turn our attention to for . It is clear that
and further by (3.27),
which means that is indeed the probability density function of provided that . (4.13) follows from the strong Markov property of . Furthermore, by (4.11), if and are sufficiently large such that and , then
Let us re-examine the event involved in the series above. If , then we must have that , , and for each , . Thus,
By (4.2) and (4.4), we have that when ,
(4.14) |
Therefore, when is sufficiently small,
which tends to as or as . Therefore, we have that
The only remaining thing is to prove the statement on and being smooth solutions to the concerned equations, which, again, by the hypoellipticity of , is reduced to showing that they are distribution solutions. Take for instance. We observe that for any ,
which implies that
This confirms that is a solution to (1.1) as a distribution. The statements on follow from similar arguments. ∎
Remark 4.4.
We want to point out that the function defined in (4.1) has a specific role in the boundary classification for diffusion process. In fact, is the scale function for the underlying diffusion process corresponding to , and as approaches a boundary, whether remains bounded or not is a factor in boundary classification (see of [26]). In particular, when viewing as a boundary of , we introduce the escape probability at (escaping from to ) as
(4.15) |
Then, when , is non-attracting, in which case (4.2) implies that ; when , is attracting and .
4.2. Approximation of
In the previous sections, for the fundamental solutions that do not have explicit formulas, we provide approximations that are accessible and of high accuracy, at least for small time. These approximations can be useful in computational applications of degenerate diffusion equations studied in this work. Below we will present an approximation for in the same spirit. In particular, we find explicitly defined approximations to such that (i) these approximations are more accurate than the standard heat kernel estimates, and (ii) when is sufficiently small, these approximations are “close” to uniformly in in any compact set. Note that this result is a generalization of [8] for that the error estimates we derive here only depend on the local bounds of and .
Theorem 4.5.
Let and , , be defined as in (3.30) and (3.31) respectively. For any , set . Then, for every , and ,
(4.16) |
In particular, there exists constant uniformly in and ( may depend on and ) such that
where is the constant defined in (3.3).
Proof.
Only (4.16) requires proof. For every , we have that
By (3.32), we have that for every , the second term on the right hand side above is bounded uniformly in by
We define hitting times as in (4.7) (for the downward crossings from to ). Then, according to (4.11),
It follows from (2.13), (3.7), (3.24) and (4.10) that for every and ,
and further by (3.30) and (4.5) we have that
∎
We close this section with two variations of (4.16). First, by (4.5), we note that
Therefore, by sending to in (4.16), we get the following estimate.
Corollary 4.6.
Second, by making in Theorem 4.5 smaller if necessary, we can derive an estimate analogous to (4.16) but independent of . Intuitively speaking, when is sufficiently small, how well approximates should not depend on the probability of the process escaping to infinity. To make it rigorous, we first observe that (H2) guarantees the existence of such that
(4.17) |
then, by using (4.14) instead of (4.10) in the proof of (4.16), we get that for every and , is bounded from above by
It follows that for every ,
Therefore, we have the following estimate on the error between and , which is a potential improvement of (4.16) for small .
Corollary 4.7.
For every , let be such that (4.17) holds. Then, for every ,
5. Generalized Wright-Fisher Diffusion
As reviewed in , the classical Wright-Fisher diffusion equation given by (1.6) has two degenerate boundaries at and , and the localization method was adopted in [7] so that one only needs to focus on one boundary at a time. Although in our setting only degenerate diffusions with one-sided boundary are concerned, the framework developed in the previous sections can also be applied to degenerate diffusions with two-sided boundaries. In this section we discuss a variation of the Wright-Fisher diffusion where the diffusion operator has general order of degeneracy at both boundaries 0 and 1.
For two constants , we consider the following Cauchy problem with two-sided boundaries on , where, given , we look for such that
(5.1) |
Set . We want to apply the method developed in the previous sections to construct and study the fundamental solution to (5.1). has two degenerate boundaries 0 and 1 with (possibly distinct) general order of degeneracy, and both boundaries are attainable according to the boundary classification mentioned in Remark 2.3.
Although having a second degenerate boundary at , has the advantage that its coefficient is bounded on . Therefore, for every , the stochastic differential equation
(5.2) |
always has a solution in the sense described in (see, e.g., of [32]). Although we are not yet ready to claim the uniqueness of this solution, we can follow the theory in of [32] to extract a solution to (5.2) that has strong Markov property. In other words, (5.2) always has a solution that is a strong Markov process.
The existence of a strong Markovian solution to (5.2) enables us to follow the steps in to tackle (5.1). In particular, with the localization procedure, we have the option of placing our “focal point” in the neighborhood of either or while constructing . We will see that these two views are consistent and will lead to the same .
Let us start with the construction of with a focus only on the left boundary , and we will follow the steps in the previous sections with and . Here we only state the results of each step but leave the computational details in the Appendix (i.e., (6.4)-(6.6)). We add a superscript “(L)” to relevant quantities and functions to indicate that only the left boundary is “effective” in this construction.
We take and localize (5.1) onto . All the functions involved in the transformation are as follows:
where is the incomplete beta function; furthermore,
in addition, for every ,
and hence
This confirms that the statement in Lemma 2.2 still holds in this case.
Next, for the model equation discussed in , we plug in and obtain as in (2.12) and as in (2.18) accordingly. We then follow exactly the same steps as in to derive based on , and to obtain through reversing the transformation , i.e.,
To proceed, we follow the arguments in to obtain the fundamental solution to (5.1) as
By (5.2), itself is a martingale, and as in Lemma 4.1, we can derive probability estimates for the hitting times of as
(5.3) |
for every and , where
For every , if is the sequence of hitting times as in (4.7) (for the downward crossings of from to ), then for every ,
(5.4) |
where the series on the right hand side is absolutely convergent.
Let us rewrite the results in Theorem 4.3 for found above.
Proposition 5.1.
is smooth on , and for every ,
(5.5) |
For every , is a smooth solution to the Kolmogorov backward equation corresponding to , i.e.,
for every , is a smooth solution to the Kolmogorov forward equation corresponding to , i.e.,
is the fundamental solution to (5.1). Given ,
(5.6) |
is a smooth solution to (1.1). Moreover, for every ,
(5.7) |
and hence for every Borel set ,
(5.8) |
Finally, satisfies the Chapman-Kolmogorov equation, i.e., for every and ,
(5.9) |
Proof.
The only thing that requires proof is the smoothness of on . By Theorem 4.3, we know that is smooth, and at the same time solves the equation . It is easy to see from here that has all the partial derivatives in of all orders. ∎
The proposition above also leads to the wellposedness of the stochastic differential equation associated with .
Corollary 5.2.
Proof.
Next we briefly discuss the other way of constructing , which is to start with the localization of (5.1) in a neighborhood of the right boundary . It is easy to see that, by exchanging and , and at the same time exchanging and , we can follow the same steps as above to develop another construction of the fundamental solution to (5.1). We will not repeat the details but only specify quantities and functions that are necessary for the statement of the results. For example, in this case the transformation is given by
is the fundamental solution to the model equation with , and given , is the fundamental solution to the localization of the model equation on ; furthermore, we have that
and for every ,
we construct from via Duhamel’s method, and obtain the fundamental solution to (5.1) localized on as
finally, if is the sequence of hitting times that records the upward crossings of from to . Then, (5.3) and the strong Markov property of are sufficient for us to obtain another version of the fundamental solution, denoted by temporarily, as
(5.10) |
for . It is easy to see that also satisfies (5.5), (5.8) and (5.9), which implies that almost everywhere on , i.e., the two constructions of the fundamental solution to (5.1) are consistent and satisfies both (5.4) and (5.10).
Depending on near which boundary we are conducting our analysis, we can choose either (5.4) or (5.10) as the definition of . For example, when both and are close to one of the boundaries, we can develop approximations for similarly as in .
Corollary 5.3.
For , set
and
Let and, respectively, be defined as in (3.4) with , and, respectively, , . Fix , and set
Then, for every and such that ,
Similarly, for every and such that ,
Proof.
We only need to look at the statement involving . There is not much to be done since a similar estimate (4.16) has been proven in Theorem 4.5. We notice that is chosen such that the function is increasing on , and . Furthermore, in this case for every , and hence . Combining the proof of Corollary 4.6, Corollary 4.7, as well as (6.4) and (6.5) in the Appendix, we get that for every as described in the statement,
∎
6. Appendix
This Appendix contains detailed derivations involving and for . Assuming that , it is sufficient for us to look at and for , where the notations become simpler. Recall that
We also introduce two more notations:
According to (1.2) and (2.2), we have that for every ,
Notice that
and further, if , then
Plugging these two expressions back into the right hand side of leads to
which is exactly (2.6). Given (H1) and (H2), the integral in the exponential function above is well defined in both cases (when and ).
With the notations introduced above, we have that when , for every ,
and
when , for every ,
and
Hence, if we set
then for every ,
(2.7) follows from here by setting
(6.1) |
For every , we can follow the arguments above to get that
Moreover, if is as defined in (4.1), then
In addition, from (3.34), we can easily derive that, for every ,
(6.2) |
and
(6.3) |
Now we move onto and recall from (3.33) that for every ,
By (1.2), the choice of and (H1) and (H2), it is straightforward to verify that when ,
which implies that
In addition, we also have that
We notice that
and
Putting all the above together yields that when ,
when , is bounded for . Thus, we have proven all the claims in Lemma 2.2.
Next, we look at the case when , where most of the expressions above take simpler forms. For example,
In particular, if as in , then
(6.4) |
Furthermore,
Since
(6.5) |
we see that
and
where in the last line we used the fact that for every ,
Combining the upper bound and the lower bound of leads to
which, by (6.5), implies that when ,
Therefore, with this specific case of , we see that the constant as introduced Lemma 2.2 (identified with in for ) can be taken as
(6.6) |
References
- [1] S. Athreya, M. Barlow, R. Bass, and E. Perkins. Degenerate stochastic differential equations and super-markov chains. Probability Theory and Related Fields, 123:484–520, 2002.
- [2] V. Barbu, A. Favini, and S. Romanelli. Degenerate evolution equations and regularity of their associated semigroups. Funkcialaj Ekvacioj, 39:421–448, 1996.
- [3] R. Bass and E. Perkins. Degenerate stochastic differential equations with holder continuous coefficients and super-markov chains. Transactions of the American Mathematical Society, 355(1):373–405, 2002.
- [4] R. Bass and E. Perkins. Degenerate stochastic differential equations arising from catalytic branching networks. Electronic Journal of Probability, 13:1808–1885, 2008.
- [5] M. Campiti, G. Metafune, and D. Pallara. Degenerate self-adjoint evolution equations on the unit interval. Semigroup Forum, 57:1–36, 1998.
- [6] S. Cerrai and P. Clément. Well-posedness of the martingale problem for some degenerate diffusion processes occurring in dynamics of populations. Bull. Sci. Math., 128:355–389, 2004.
- [7] L. Chen and D. Stroock. The fundamental solution to the wright-fisher equation. SIAM J. Math. Anal., 42(2):539–567, 2010.
- [8] L. Chen and I. Weih-Wadman. The fundamental solution to one-dimensional degenerate equation, i. pages 1–33, 2019. arXiv:1905.12716.
- [9] A. S. Cherny. On the uniqueness in law and the pathwise uniqueness for stochastic differential equations. Theory of Probability and Its Applications, 46(3):483–497, 2001.
- [10] A. S. Cherny and H.-J. Engelbert. Singular stochastic differential equations. Springer, Lecture Notes in Mathematics Series No. 1858, 2005.
- [11] P. Clément and C. A. Timmermans. On -semigroups generated by differential operators satisfying ventcel’s boundary conditions. Indag. Math., 89:379–387, 1986.
- [12] D. A. Dawson. Measure-valued markov processes. Lecture Notes in Mathematics, Spring-Verlag, Berlin, 1451:417–472, 1993. Ecole d’Eté de Probabilités de Saint Flour 1991.
- [13] D. A. Dawson and P. March. Resolvent estimates for fleming-viot operators and uniqueness of solutions to related martingale problems. J. Funct. Anal., 132:417–472, 1995.
- [14] H.-J. Engelbert and W. Schmidt. Strong markov continuous local martingales and solutions of one-dimensional stochastic differential equations, i, ii, iii. Math. Machr., 143,144,151:167–184,241–281,149–197, 1989, 1991.
- [15] C. L. Epstein and R. Mazzeo. Degenerate diffusion operators arising in population biology. Annals of Mathematics Studies, 185, 2013. Princeton University Press, Princeton, NJ.
- [16] C. L. Epstein and R. Mazzeo. The geometric microlocal analysis of generalized kimura and heston diffusions. Analysis and Topology in Nonlinear Differential Equations. Progress in Nonlinear Differential Equations and Their Applications, 85:241–266, 2014.
- [17] C. L. Epstein and R. Mazzeo. Harnack inequalities and heat-kernel estimates for degenerate diffusion operators arising in population biology. Applied Mathematics Research eXpress, 2016(2):217–280, 2016.
- [18] C. L. Epstein and C. A. Pop. The feynman-kac formula and harnack inequality for degenerate diffusions. Annuals of Probability, 45(5):3336–3384, 2017.
- [19] C. L. Epstein and C. A. Pop. Transition probabilities for degenerate diffusions arising in population genetics. Probability Theory and Related Fields, 173:537–603, 2019.
- [20] S. N. Ethier. A class of degenerate diffusion processes occurring in population genetics. Communications on Pure and Applied Math, 29:483–493, 1976.
- [21] S. N. Ethier and T. Kurtz. Fleming-viot processes in populations genetics. SIAM J. Control Optim., 31:345–386, 1993.
- [22] W. H. Fleming and M. Viot. Some measure-valued markov processes in populations genetics theory. Indiana Univ. Math. J., 28:817–843, 1979.
- [23] J. A. Goldstein and C. Y. Lin. Degenerate parabolic problems and wentzel boundary condition. Semigroup Theory and Applications, Lecture Notes in Pure and Applied Mathematics No. 116, pages 189–199, 1989.
- [24] J. A. Goldstein and C. Y. Lin. Highly degenerate parabolic boundary value problems. Differential Integral Equations, 2:216–227, 1989.
- [25] I. Karatzas and S. Shreve. Brownian motion and stochastic calculus. Springer-Verlag New York, 1998.
- [26] S. Karlin and H. Taylor. A second course in stochastic processes. Academic Press, 1981.
- [27] N. V. Krylov. On some estimates of the distribution of stochastic integrals. Izv. Akad. Nank SSSR Ser. Math., 38:228–248, 1974.
- [28] K. S. Kumar. A class of degenerate stochastic differential equations with non-lipschitz coefficients. Proc. Indian Acad. Sci. (Math. Sci.), 123(3):443–454, 2013.
- [29] E. Perkins. Dawson-watanabe superprocesses and measure-valued diffusions. Lecture Notes in Mathematics, Spring-Verlag, Berlin, 1781, 2001. Ecole d’Eté de Probabilités de Saint Flour 1999.
- [30] C. A. Pop. -estimates and smoothness of solutions to the parabolic equation defined by kimura operators. Journal of Functional Analysis, 272(1):47–82, 2017.
- [31] C. A. Pop. Existence, uniqueness and the strong markov property of solutions to kimura diffusions with singular drift. Trans. Amer. Math. Soc., 369:5543–5579, 2017.
- [32] D. Stroock. Partial differential equations for probabilists. Cambridge Univ. Press, 2008.
- [33] D. Stroock. Probability, an analytic view, 2nd edition. Cambridge Univ. Press, 2011.
- [34] D. Stroock and S. R. S. Varadhan. Diffusion processes with continuous coefficients, i, ii. Commun. Math. Phys., 22:345–400,479–530, 1969.
- [35] D. Stroock and S. R. S. Varadhan. Multidimensional diffusion processes. Springer-Verlag, Grundlehren Series No. 233, 2006.
- [36] J. Swart. Pathwise uniqueness for a sde with non-lipschitz coefficients. Stoch. Proc. Appl., 98:131–149, 2002.
- [37] V. Vespri. Analytic semigroups, degenerate elliptic operators and applications to nonlinear cauchy problems. Ann. Mat. Pura Appl., CLV:353–388, 1989.
- [38] T. Yamada and S. Watanabe. On the uniqueness of solutions of stochastic differential equations. J. Math. Kyoto Univ., 11(1):155–167, 1971.
- [39] A. K. Zvonkin. A transformation of the phase space of a diffusion process that will remove the drift. Mat. Sb., 93:129–149, 1974.