This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Fundamental Solution to 1D Degenerate Diffusion Equation with Locally Bounded Coefficients

Linan Chen* and Ian Weih-Wadman *Department of Mathematics and Statistics, McGill University, 805 Sherbrooke St. West, Montreal, Quebec, H3A 0B9, Canada. †Department of Mathematics and Statistics, McMaster University, 1280 Main St. West, Hamilton, Ontario, L8S 4K1, Canada. *[email protected] [email protected]
Abstract.

In this work we study the degenerate diffusion equation t=xαa(x)x2+b(x)x\partial_{t}=x^{\alpha}a\left(x\right)\partial_{x}^{2}+b\left(x\right)\partial_{x} for (x,t)(0,)2\left(x,t\right)\in\left(0,\infty\right)^{2}, equipped with a Cauchy initial data and the Dirichlet boundary condition at 0. We assume that the order of degeneracy at 0 of the diffusion operator is α(0,2)\alpha\in\left(0,2\right), and both a(x)a\left(x\right) and b(x)b\left(x\right) 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 p(x,y,t)p\left(x,y,t\right) and prove several properties for p(x,y,t)p\left(x,y,t\right); by conducting a localization procedure, we obtain an approximation for p(x,y,t)p\left(x,y,t\right) for x,yx,y in a neighborhood of 0 and tt sufficiently small, where the error estimates only rely on the local bounds of a(x)a\left(x\right) and b(x)b\left(x\right) (and their derivatives). There is a rich literature on such a degenerate diffusion in the case of α=1\alpha=1. 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 equations
2000 Mathematics Subject Classification:
35K20, 35K65, 35Q92
The first author was partially supported by the NSERC Discovery Grant (No. 241023).

1. Introduction

In this article we consider the following Cauchy initial value problem with the Dirichlet boundary condition:

(1.1) tuf(x,t)=xαa(x)x2uf(x,t)+b(x)xuf(x,t) for (x,t)(0,)2,limt0uf(x,t)=f(x) for x(0,) and limx0uf(x,t)=0 for t(0,),\begin{array}[]{c}\partial_{t}u_{f}\left(x,t\right)=x^{\alpha}a\left(x\right)\partial_{x}^{2}u_{f}\left(x,t\right)+b\left(x\right)\partial_{x}u_{f}\left(x,t\right)\text{ for }\left(x,t\right)\in\left(0,\infty\right)^{2},\\ \lim_{t\searrow 0}u_{f}\left(x,t\right)=f\left(x\right)\text{ for }x\in\left(0,\infty\right)\text{ and }\lim_{x\searrow 0}u_{f}\left(x,t\right)=0\text{ for }t\in\left(0,\infty\right),\end{array}

where fCc((0,))f\in C_{c}\left(\left(0,\infty\right)\right) and α(0,2)\alpha\in\left(0,2\right). Set L:=xαa(x)x2+b(x)xL:=x^{\alpha}a\left(x\right)\partial_{x}^{2}+b\left(x\right)\partial_{x}. We further impose the following assumptions on a(x)a\left(x\right) and b(x)b\left(x\right):

  • (H1):

    aC([0,))C2((0,))a\in C\left([0,\infty)\right)\cap C^{2}\left(\left(0,\infty\right)\right), a(x)>0a\left(x\right)>0 for every x[0,)x\in[0,\infty) and a(0)=1a\left(0\right)=1.
    bC([0,))C1((0,))b\in C\left([0,\infty)\right)\cap C^{1}\left(\left(0,\infty\right)\right), b(0)[0,1)b\left(0\right)\in[0,1) when α1\alpha\leq 1, and b(0)=0b\left(0\right)=0 when α>1\alpha>1.
    a(x)a\left(x\right), a(x)a^{\prime}\left(x\right), a′′(x)a^{\prime\prime}\left(x\right), b(x)b\left(x\right) and b(x)b^{\prime}\left(x\right) are all bounded on (0,I](0,I] for every I>0I>0.

  • (H2):

    There exists C>0C>0 such that for every x[0,)x\in[0,\infty),

    a(x)C(1+x2α) and |b(x)|C(1+x).a\left(x\right)\leq C\left(1+x^{2-\alpha}\right)\text{ and }\left|b\left(x\right)\right|\leq C\left(1+x\right).

Our goal is to construct and to study the fundamental solution p(x,y,t)p\left(x,y,t\right) to (1.1), with a particular emphasis on the behaviors of p(x,y,t)p\left(x,y,t\right) when x,yx,y are near the boundary and when tt is small. Since LL is degenerate at 0, standard methods on strictly parabolic equation no longer apply in this case, and the degeneracy of LL does have an impact on the regularity of p(x,y,t)p\left(x,y,t\right). Moreover, the assumptions (H1) and (H2) only guarantee local boundedness of a(x)a\left(x\right), b(x)b\left(x\right) and their derivatives, so we need to conduct our analysis of p(x,y,t)p\left(x,y,t\right) 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 α\alpha, a(x)a\left(x\right) and b(x)b\left(x\right) in our setting, we define the following two functions for x>0x>0:

(1.2) ϕ(x):=14(0xdssα2a(s))2 and θ(x):=12ν+2b(x)(xαa(x))2xα2a(x)ϕ(x),\phi\left(x\right):=\frac{1}{4}\left(\int_{0}^{x}\frac{ds}{s^{\frac{\alpha}{2}}\sqrt{a\left(s\right)}}\right)^{2}\text{ and }\theta\left(x\right):=\frac{1}{2}-\nu+\frac{2b\left(x\right)-\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{2x^{\frac{\alpha}{2}}\sqrt{a\left(x\right)}}\sqrt{\phi\left(x\right)},

where ν\nu is the constant such that limx0θ(x)=0\lim_{x\searrow 0}\theta\left(x\right)=0. It is assumed in [8] that ν<1\nu<1, as well as

(1.3) supx(0,)|θ(x)|ϕ(x)< and supx(0,)|θ(x)|ϕ(x)<.\sup_{x\in(0,\infty)}\frac{\left|\theta\left(x\right)\right|}{\sqrt{\phi\left(x\right)}}<\infty\text{ and }\sup_{x\in\left(0,\infty\right)}\frac{\left|\theta^{\prime}\left(x\right)\right|}{\phi^{\prime}\left(x\right)}<\infty.

Under the assumption (1.3), through a series of transformations and perturbations, [8] completes a construction of the fundamental solution p(x,y,t)p\left(x,y,t\right) to (1.1), conducts a careful analysis of the regularity properties of p(x,y,t)p\left(x,y,t\right) near the boundary, and derive an approximations for p(x,y,t)p\left(x,y,t\right) in terms of explicitly formulated functions. In particular, if papprox.(x,y,t)p^{approx.}\left(x,y,t\right) denotes the approximation for p(x,y,t)p\left(x,y,t\right), then it is proven in [8] that there exists a constant C>0C>0, universal in all x,yx,y in a neighborhood of 0 and all tt sufficiently small, such that

(1.4) |p(x,y,t)papprox.(x,y,t)1|Ct.\left|\frac{p\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}-1\right|\leq Ct.

Such an estimate is useful in multiple ways. First, while one expects p(x,y,t)p\left(x,y,t\right) to resemble the fundamental solution to a strictly parabolic equation for x,yx,y away from the boundary, (1.4) captures accurately the asymptotics of p(x,y,t)p\left(x,y,t\right) when x,yx,y are close to the boundary, and demonstrates the influence of the degeneracy of LL on p(x,y,t)p\left(x,y,t\right). Second, if one could apply the general heat kernel estimates (see, e.g., §4\mathsection 4 of [32]) to p(x,y,t)p\left(x,y,t\right), then one would get that for every δ\delta and tt sufficiently small, there is constant Cδ,t>1C_{\delta,t}>1 such that

(1.5) Ct,δ1exp(d(x,y)22(1δ)t)p(x,y,t)Ct,δexp(d(x,y)22(1+δ)t),C_{t,\delta}^{-1}\exp\left(-\frac{d\left(x,y\right)^{2}}{2\left(1-\delta\right)t}\right)\leq p\left(x,y,t\right)\leq C_{t,\delta}\exp\left(-\frac{d\left(x,y\right)^{2}}{2\left(1+\delta\right)t}\right),

where d(x,y)d\left(x,y\right) is the distance between xx and yy under the Riemannian metric corresponding to LL; it is clear that (1.4) is a sharper estimate than (1.5) for small tt, and hence papprox.(x,y,t)p^{approx.}\left(x,y,t\right) is a more accurate short-term approximation for p(x,y,t)p\left(x,y,t\right) compared with the general heat kernel approximation. In addition, in [8], papprox.(x,y,t)p^{approx.}\left(x,y,t\right) is presented in an explicit formula (in terms of special functions) and “CtCt” 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 tL=0\partial_{t}-L=0.

We aim to generalize the results in [8], particularly the construction of p(x,y,t)p\left(x,y,t\right) and the short-term near-boundary approximation papprox.(x,y,t)p^{approx.}\left(x,y,t\right), 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 b(0)0b\left(0\right)\neq 0, which means that, given (H1) and (H2), (1.3) is only satisfied when 1α<21\leq\alpha<2. Moreover, (1.3) clearly imposes strong global conditions on a(x)a\left(x\right) and b(x)b\left(x\right), but with (H1) and (H2), we have to find an access to p(x,y,t)p\left(x,y,t\right) 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) tuf(x,t)=x(1x)x2uf(x,t) for (x,t)(0,1)×(0,),limt0uf(x,t)=f(x) for x(0,1) for some fCb((0,1)), and limx0uf(x,t)=limx1uf(x,t)=0 for t(0,).\begin{array}[]{c}\partial_{t}u_{f}\left(x,t\right)=x\left(1-x\right)\partial_{x}^{2}u_{f}\left(x,t\right)\text{ for }\left(x,t\right)\in\left(0,1\right)\times\left(0,\infty\right),\\ \lim_{t\searrow 0}u_{f}\left(x,t\right)=f\left(x\right)\text{ for }x\in\left(0,1\right)\text{ for some }f\in C_{b}\left(\left(0,1\right)\right),\\ \text{ and }\lim_{x\searrow 0}u_{f}\left(x,t\right)=\lim_{x\nearrow 1}u_{f}\left(x,t\right)=0\text{ for }t\in\left(0,\infty\right).\end{array}

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 LWF:=x(1x)x2L_{WF}:=x\left(1-x\right)\partial_{x}^{2} and let pWF(x,y,t)p_{WF}\left(x,y,t\right) be the fundamental solution to (1.6). Since (1.6) is symmetric on [0,1]\left[0,1\right], to study pWF(x,y,t)p_{WF}\left(x,y,t\right) near the boundaries, it is sufficient to only consider the left boundary 0. In [7], a “localization” method is devised to construct pWF(x,y,t)p_{WF}\left(x,y,t\right) near 0: since pWF(x,y,t)p_{WF}\left(x,y,t\right) can be viewed as the density of the underlying diffusion process corresponding to LWFL_{WF}, we can acquire information on pWF(x,y,t)p_{WF}\left(x,y,t\right) by studying the behaviors of the process near 0; in particular, by tracking the excursions of the diffusion process near 0, we can “localize” LWFL_{WF} and pWF(x,y,t)p_{WF}\left(x,y,t\right) within a neighborhood of 0 where only the degeneracy at 0 has a substantial impact. Heuristically speaking, when restricted near 0, LWFL_{WF} is close to the operator xx2x\partial_{x}^{2}, and hence it is natural to expect that pWF(x,y,t)p_{WF}\left(x,y,t\right) with x,yx,y near 0 is close to the fundamental solution p0(x,y,t)p_{0}\left(x,y,t\right) to txx2=0\partial_{t}-x\partial_{x}^{2}=0 (with Dirichlet boundary 0). Indeed, it is established in [7] that, not only can pWF(x,y,t)p_{WF}\left(x,y,t\right) be constructed in an explicit way via p0(x,y,t)p_{0}\left(x,y,t\right), pWF(x,y,t)p_{WF}\left(x,y,t\right) is also well approximated by p0(x,y,t)p_{0}\left(x,y,t\right) in the sense that pWF(x,y,t)/p0(x,y,t)p_{WF}\left(x,y,t\right)/p_{0}\left(x,y,t\right) satisfies (1.4) for x,yx,y near 0 and tt 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 tL=0\partial_{t}-L=0 have also been discussed in many other contexts, with most of the existing literature concerning the case when α=1\alpha=1. 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 LL with α=1\alpha=1 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 α=1\alpha=1, a series of works (see, e.g., [1, 3, 4, 9, 14, 38]) provide conditions on a(x)a\left(x\right) and b(x)b\left(x\right) that are sufficient for the stochastic differential equation corresponding to LL to be well posed, and some of the results will also be used later in our discussions.

Degenerate diffusions have also been treated in the context of the measure-valued process (see, e.g., [6, 12, 13, 21, 22, 29]), as well as via the semigroup approach (see, e.g., [2, 5, 11, 23, 24, 37]).

1.2. Our main results

Our strategy in solving (1.1) and getting p(x,y,t)p\left(x,y,t\right) is to combine the ideas and the techniques from [8] and [7], and tackle the two challenges we face: general order of degeneracy in LL 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 (§2.1\mathsection 2.1). Since the coefficients in LL are locally bounded, we first consider a “localized” version of (1.1). Given I>0I>0, we study the diffusion equation in (1.1) on (0,I)\left(0,I\right) with an extra Dirichlet boundary at II, i.e.,

tu(x,t)=Lu(x,t) for (x,t)(0,I)×(0,)with u(x,t)0 as x0 or xI for t(0,).()\begin{array}[]{c}\partial_{t}u\left(x,t\right)=Lu\left(x,t\right)\text{ for }\left(x,t\right)\in\left(0,I\right)\times\left(0,\infty\right)\\ \text{with }u\left(x,t\right)\rightarrow 0\text{ as }x\searrow 0\text{ or }x\nearrow I\text{ for }t\in\left(0,\infty\right).\end{array}\quad\left(\star\right)

Let pI(x,y,t)p_{I}\left(x,y,t\right) be the fundamental solution to ()\left(\star\right). To solve ()\left(\star\right), we carry out a transformation that turns LL into a diffusion operator that degenerates linearly at 0. In fact, with a change of variable xzx\mapsto z, solving ()\left(\star\right) becomes equivalent to solving the following problem:

tvV(z,t)=(zz2+νz+V(z))vV(z,t) for (z,t)(0,J)×(0,)with vV(z,t)0 as z0 or zJ for t(0,),()\begin{array}[]{c}\begin{array}[]{c}\partial_{t}v^{V}\left(z,t\right)=\left(z\partial_{z}^{2}+\nu\partial_{z}+V\left(z\right)\right)v^{V}\left(z,t\right)\text{ for }\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right)\\ \text{with }v^{V}\left(z,t\right)\rightarrow 0\text{ as }z\searrow 0\text{ or }z\nearrow J\text{ for }t\in\left(0,\infty\right),\end{array}\end{array}\quad\left(\dagger\right)

where JJ is the image of II after the change of variable, ν<1\nu<1 is a constant, and V(z)V\left(z\right) is a function on (0,J)\left(0,J\right) (JJ, ν\nu and VV will be specified in §2.1\mathsection 2.1). If we can find the fundamental solution to ()\left(\dagger\right), denoted by qJV(z,w,t)q_{J}^{V}\left(z,w,t\right), then pI(x,y,t)p_{I}\left(x,y,t\right) can be obtained through qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) via the transformation (and its inverse) between ()\left(\star\right) and ()\left(\dagger\right).

2. Model equation (§2.2\mathsection 2.2). Our strategy for solving ()\left(\dagger\right) is to treat the operator zz2+νz+V(z)z\partial_{z}^{2}+\nu\partial_{z}+V\left(z\right) as a perturbation of zz2+νzz\partial_{z}^{2}+\nu\partial_{z}. We temporarily return to the “global” view, omit the potential V(z)V\left(z\right) and the right boundary JJ, and consider the following model equation on the entire (0,)\left(0,\infty\right):

tv(z,t)=(zz2+νz)v(z,t) for every (z,t)(0,)2with v(z,t)0 as z0 for t(0,).\begin{array}[]{c}\partial_{t}v\left(z,t\right)=\left(z\partial_{z}^{2}+\nu\partial_{z}\right)v\left(z,t\right)\text{ for every }\left(z,t\right)\in\left(0,\infty\right)^{2}\\ \text{with }v\left(z,t\right)\rightarrow 0\text{ as }z\searrow 0\text{ for }t\in\left(0,\infty\right).\end{array}

This model equation has the advantage that its fundamental solution q(z,w,t)q\left(z,w,t\right) 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 q(z,w,t)q\left(z,w,t\right) in hand, we return to the local view of the model equation (with the Dirichlet boundary condition “restored” at JJ) and derive the fundamental solution qJ(z,w,t)q_{J}\left(z,w,t\right) to the localized model equation on (0,J)\left(0,J\right) (Proposition 2.6).

3. Solving the localized equation (§3\mathsection 3). Upon getting qJ(z,w,t)q_{J}\left(z,w,t\right), we can start the construction of the fundamental solutions to ()\left(\dagger\right) and ()\left(\star\right). Viewing zz2+νz+V(z)z\partial_{z}^{2}+\nu\partial_{z}+V\left(z\right) as a perturbation of zz2+νzz\partial_{z}^{2}+\nu\partial_{z} with a potential function V(z)V\left(z\right), we invoke Duhamel’s perturbation method to construct qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) using qJ(z,w,t)q_{J}\left(z,w,t\right) as the “building block” (Proposition 3.2). Although in general qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) does not have a closed-form formula and our representation of qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is in the form of a series, by focusing on the first term of the series expression we can show that qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is well approximated by q(z,w,t)q\left(z,w,t\right) for sufficiently small tt (Proposition 3.4).

4. Solving the global equation (§4\mathsection 4). We finally return to (1.1) and produce p(x,y,t)p\left(x,y,t\right) from pI(x,y,t)p_{I}\left(x,y,t\right) by “reversing” the localization procedure. More specifically, we establish the relation between (1.1) and its localized version ()\left(\star\right) with the help of the underlying diffusion process corresponding to LL. By analyzing the excursions of the diffusion process over (0,I)\left(0,I\right), p(x,y,t)p\left(x,y,t\right) is achieved as the limit of pI(x,y,t)p_{I}\left(x,y,t\right) as II increases to infinity (Theorem 4.3). Again, although p(x,y,t)p\left(x,y,t\right) does not have a closed-form formula, we find an approximation papprox.(x,y,t)p^{approx.}\left(x,y,t\right) for p(x,y,t)p\left(x,y,t\right) such that papprox.(x,y,t)p^{approx.}\left(x,y,t\right) has an explicit and relatively simple expression, and papprox.(x,y,t)p^{approx.}\left(x,y,t\right) is more accurate than the standard heat kernel estimate for p(x,y,t)p\left(x,y,t\right) (Theorem 4.5).

(transformation)
localized equations: pI(x,y,t)p_{I}\left(x,y,t\right) \longleftrightarrow qJV(z,w,t)q_{J}^{V}\left(z,w,t\right)
(convergence)\;\downarrow\quad\uparrow\;(localization) \quad\;\uparrow\;(perturbation)
qJ(z,w,t)q_{J}\left(z,w,t\right)
- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - -
global equations: p(x,y,t)p\left(x,y,t\right) \quad\;\uparrow\;(localization)
(approximation)\;\uparrow\quad\quad\quad\quad\quad\quad
papprox.(x,y,t)p^{approx.}\left(x,y,t\right) \longleftarrow q(z,w,t)q\left(z,w,t\right)
(transformation)
Table 1. Relation among the fundamental solutions.

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 §1.3\mathsection 1.3 we give a brief overview of the probabilistic components involved in this work.

In §5\mathsection 5 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 0 and 11. In particular, for fCc((0,1))f\in C_{c}\left(\left(0,1\right)\right) and α,β(0,2)\alpha,\beta\in\left(0,2\right), we consider the equation

tuf(x,t)=xα(1x)βx2uf(x,t) for (x,t)(0,1)×(0,),limt0uf(x,t)=f(x) for x(0,1) and limx0uf(x,t)=limx1uf(x,t)=0 for t(0,).\begin{array}[]{c}\partial_{t}u_{f}\left(x,t\right)=x^{\alpha}\left(1-x\right)^{\beta}\partial_{x}^{2}u_{f}\left(x,t\right)\text{ for }\left(x,t\right)\in\left(0,1\right)\times\left(0,\infty\right),\\ \lim_{t\searrow 0}u_{f}\left(x,t\right)=f\left(x\right)\text{ for }x\in\left(0,1\right)\text{ and }\\ \lim_{x\searrow 0}u_{f}\left(x,t\right)=\lim_{x\nearrow 1}u_{f}\left(x,t\right)=0\text{ for }t\in\left(0,\infty\right).\end{array}

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 p(x,y,t)p\left(x,y,t\right) and obtain similar estimates for p(x,y,t)p\left(x,y,t\right) 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 L=xαa(x)x2+b(x)xL=x^{\alpha}a\left(x\right)\partial_{x}^{2}+b\left(x\right)\partial_{x}, and that is, given x>0x>0,

(1.7) dX(x,t)=2Xα(x,t)a(X(x,t))dB(t)+b(X(x,t))dt for every t0 with X(0,x)x.dX\left(x,t\right)=\sqrt{2X^{\alpha}\left(x,t\right)a\left(X\left(x,t\right)\right)}dB\left(t\right)+b\left(X\left(x,t\right)\right)dt\text{ for every }t\geq 0\text{ with }X\left(0,x\right)\equiv x.

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 (Ω,,{t:t0},)\left(\Omega,\mathcal{F},\left\{\mathcal{F}_{t}:t\geq 0\right\},\mathbb{P}\right), there exist two adapted processes {B(t):t0}\left\{B\left(t\right):t\geq 0\right\} and {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} such that, (i) {B(t):t0}\left\{B\left(t\right):t\geq 0\right\} is a standard Brownian motion; (ii) {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} has continuous sample paths; (iii) almost surely {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} satisfies that

X(x,t)=x+0t2Xα(x,s)a(X(x,s))𝑑B(s)+0tb(X(x,s))𝑑s for every t0.X\left(x,t\right)=x+\int_{0}^{t}\sqrt{2X^{\alpha}\left(x,s\right)a\left(X\left(x,s\right)\right)}dB\left(s\right)+\int_{0}^{t}b\left(X\left(x,s\right)\right)ds\text{ for every }t\geq 0.

In this case, we also refer to {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} as the underlying diffusion process corresponding to LL starting from xx. We say that a solution {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} is unique (in law), if whenever (i)-(iii) are satisfied by another triple (Ω,,{t:t0},)\left(\Omega^{\prime},\mathcal{F}^{\prime},\left\{\mathcal{F}_{t}^{\prime}:t\geq 0\right\},\mathbb{P}^{\prime}\right), {B(t):t0}\left\{B^{\prime}\left(t\right):t\geq 0\right\} and {X(x,t):t0}\left\{X^{\prime}\left(x,t\right):t\geq 0\right\}, it must be that the distribution of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} under \mathbb{P} is identical with that of {X(x,t):t0}\left\{X^{\prime}\left(x,t\right):t\geq 0\right\} under \mathbb{P}^{\prime}. 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 {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} is the unique solution, then {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} is a strong Markov process and for every HC([0,)2)C2,1((0,)2)H\in C\left([0,\infty)^{2}\right)\cap C^{2,1}\left(\left(0,\infty\right)^{2}\right),

{H(X(x,t),t)0t(s+L)H(X(x,s),s)𝑑s:t0}\left\{H\left(X\left(x,t\right),t\right)-\int_{0}^{t}\left(\partial_{s}+L\right)H\left(X\left(x,s\right),s\right)ds:t\geq 0\right\}

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 (0,)\left(0,\infty\right) and its inverse:

(1.8) ξ=ξ(x):=x2α(2α)2 and x=x(ξ):=(2α)22αξ12α for x,ξ>0.\xi=\xi\left(x\right):=\frac{x^{2-\alpha}}{\left(2-\alpha\right)^{2}}\text{ and }x=x\left(\xi\right):=\left(2-\alpha\right)^{\frac{2}{2-\alpha}}\xi^{\frac{1}{2-\alpha}}\text{ for }x,\xi>0.

One can easily verify that uf(x,t)C2,1((0,))u_{f}\left(x,t\right)\in C^{2,1}\left(\left(0,\infty\right)\right) is a solution to (1.1) if and only if wg(ξ,t):=uf(x(ξ),t)w_{g}\left(\xi,t\right):=u_{f}\left(x\left(\xi\right),t\right) is the solution to

(1.9) twg(ξ,t)=ξc(ξ)ξ2wg(ξ,t)+d(ξ)ξwg(ξ,t) for (ξ,t)(0,)2,limt0wg(ξ,t)=g(ξ) for ξ(0,1) and limξ0wg(ξ,t)=0 for t(0,),\begin{array}[]{c}\partial_{t}w_{g}\left(\xi,t\right)=\xi c\left(\xi\right)\partial_{\xi}^{2}w_{g}\left(\xi,t\right)+d\left(\xi\right)\partial_{\xi}w_{g}\left(\xi,t\right)\text{ for }\left(\xi,t\right)\in\left(0,\infty\right)^{2},\\ \lim_{t\searrow 0}w_{g}\left(\xi,t\right)=g\left(\xi\right)\text{ for }\xi\in\left(0,1\right)\text{ and }\lim_{\xi\searrow 0}w_{g}\left(\xi,t\right)=0\text{ for }t\in\left(0,\infty\right),\end{array}

where g(ξ):=f(x(ξ))g\left(\xi\right):=f\left(x\left(\xi\right)\right), c(ξ):=a(x(ξ))c\left(\xi\right):=a\left(x\left(\xi\right)\right) and

d(ξ):=1α2αa(x(ξ))+(x(ξ))1α2αb(x(ξ)).d\left(\xi\right):=\frac{1-\alpha}{2-\alpha}a\left(x\left(\xi\right)\right)+\frac{\left(x\left(\xi\right)\right)^{1-\alpha}}{2-\alpha}b\left(x\left(\xi\right)\right).

The stochastic differential equation corresponding to (1.9) is that, given ξ>0\xi>0,

(1.10) dZ(ξ,t)=2Z(ξ,t)c(Z(ξ,t))dB(t)+d(Z(ξ,t))dt for every t0 with Z(ξ,0)ξ.dZ\left(\xi,t\right)=\sqrt{2Z\left(\xi,t\right)c\left(Z\left(\xi,t\right)\right)}dB\left(t\right)+d\left(Z\left(\xi,t\right)\right)dt\text{ for every }t\geq 0\text{ with }Z\left(\xi,0\right)\equiv\xi.

Assuming (H1) and (H2), we get down to verifying the wellposedness of (1.10) where the diffusion operator degenerates linearly at 0. First, when α(1,2)\alpha\in\left(1,2\right), by (H1), (1.8) and direct computations, we see that both c(ξ)c\left(\xi\right) and d(ξ)d\left(\xi\right) are Lipschitz continuous on any bounded subset of (0,)\left(0,\infty\right). Furthermore, (H2) and (1.8) imply that there exists constant C>0C>0 such that for every ξ[0,)\xi\in[0,\infty),

(1.11) |c(ξ)|+|d(ξ)|C(1+(x(ξ))2α)C(1+ξ).\left|c\left(\xi\right)\right|+\left|d\left(\xi\right)\right|\leq C\left(1+\left(x\left(\xi\right)\right)^{2-\alpha}\right)\leq C\left(1+\xi\right).

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 ξ>0\xi>0 in this case. Next, when α(0,1]\alpha\in(0,1], we note that

limξ0d(ξ)=limx0(1α2αa(x)+x1α2αb(x))0.\lim_{\xi\searrow 0}d\left(\xi\right)=\lim_{x\searrow 0}\left(\frac{1-\alpha}{2-\alpha}a\left(x\right)+\frac{x^{1-\alpha}}{2-\alpha}b\left(x\right)\right)\geq 0.

This time (H1) and (1.8) guarantee that c(ξ)c\left(\xi\right) and d(ξ)d\left(\xi\right) are both Hölder continuous on any bounded subset of (0,)\left(0,\infty\right); meanwhile, the growth control (1.11) on c(ξ)c\left(\xi\right) and d(ξ)d\left(\xi\right) still applies. Thus, the results of Bass-Perkins [3] lead to the wellposedness of (1.10) for every ξ>0\xi>0. Therefore, for every α(0,2)\alpha\in\left(0,2\right), (H1) and (H2) are sufficient for (1.10) to be well posed. Assume that {Z(ξ,t):t0}\left\{Z\left(\xi,t\right):t\geq 0\right\} is the unique solution to (1.10). By setting

X(x,t):=x(Z(ξ(x),t)) for t0,X\left(x,t\right):=x\left(Z\left(\xi\left(x\right),t\right)\right)\text{ for }t\geq 0,

we immediately get the following conclusion.

Lemma.

The stochastic differential equation (1.7) is well posed for every x>0x>0, {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} defined above is the unique solution to (1.7), and {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} is a strong Markov process. Moreover, if u(x,t)C2,1((0,)2)u\left(x,t\right)\in C^{2,1}\left(\left(0,\infty\right)^{2}\right) is a solution to tu(x,t)=Lu(x,t)\partial_{t}u\left(x,t\right)=Lu\left(x,t\right), then given any (x,t)(0,)2\left(x,t\right)\in\left(0,\infty\right)^{2}, {u(X(x,s),ts):s[0,t]}\left\{u\left(X\left(x,s\right),t-s\right):s\in\left[0,t\right]\right\} is a local martingale.

So far there is no constraint on the behavior of X(x,t)X\left(x,t\right) at the boundary 0. Returning to the original problem (1.1), to incorporate the Dirichlet boundary condition, we only need to focus on X(x,t)X\left(x,t\right) up to the time it hits 0. Intuitively speaking, if we set

ζ0X(x):=inf{s0:X(x,s)=0} for x>0,\zeta_{0}^{X}\left(x\right):=\inf\left\{s\geq 0:X\left(x,s\right)=0\right\}\text{ for }x>0,

then the probability density function of the conditional distribution of X(x,t)X\left(x,t\right) given {t<ζ0X(x)}\left\{t<\zeta_{0}^{X}\left(x\right)\right\} should coincide with the fundamental solution to (1.1).

Notations.

For α,β\alpha,\beta\in\mathbb{R}, we write αβ:=max{α,β}\alpha\vee\beta:=\max\left\{\alpha,\beta\right\} and αβ:=min{α,β}\alpha\wedge\beta:=\min\left\{\alpha,\beta\right\}.

For every Γ[0,)\Gamma\subseteq[0,\infty), 𝕀Γ\mathbb{I}_{\Gamma} denotes the indicator function of Γ\Gamma.

Let (Ω,,{t:t0},)\left(\Omega,\mathcal{F},\left\{\mathcal{F}_{t}:t\geq 0\right\},\mathbb{P}\right) be a filtered probability space. For an integrable random variable XX on Ω\Omega and a set AA\in\mathcal{F}, we write 𝔼[X;A]:=AX𝑑\mathbb{E}\left[X;A\right]:=\int_{A}Xd\mathbb{P}. For an adapted process {W(t):t0}\left\{W\left(t\right):t\geq 0\right\} with non-negative continuous sample paths, we set

ζyW(x):=inf{t0:W(t)=y|W(0)=x} for every x,y0,\zeta_{y}^{W}\left(x\right):=\inf\left\{t\geq 0:W\left(t\right)=y|W\left(0\right)=x\right\}\text{ for every }x,y\geq 0,

i.e., ζyW(x)\zeta_{y}^{W}\left(x\right) is the hitting time at yy conditioning on the process starting from xx; for y1,y20y_{1},y_{2}\geq 0, we set ζy1,y2W(x):=ζy1W(x)ζy2W(x)\zeta_{y_{1},y_{2}}^{W}\left(x\right):=\zeta_{y_{1}}^{W}\left(x\right)\wedge\zeta_{y_{2}}^{W}\left(x\right).

2. Model Equation

In this section, we will carry out the first three steps outlined in §1.2\mathsection 1.2. Although we use similar transformations as those in [8], we need to adapt the method so that it applies to a(x)a\left(x\right) and b(x)b\left(x\right) that are under weaker conditions.

2.1. Localization and Transformation

Let α(0,2)\alpha\in\left(0,2\right) and I>0I>0 be fixed throughout this section. Our first step is to introduce an extra Dirichlet boundary to the equation tL=0\partial_{t}-L=0 at x=Ix=I and to consider a localized version of (1.1) on (0,I)\left(0,I\right). Namely, given fCc((0,I))f\in C_{c}\left(\left(0,I\right)\right), we look for uf,(0,I)(x,t)C2,1((0,I)×(0,))u_{f,\left(0,I\right)}\left(x,t\right)\in C^{2,1}\left(\left(0,I\right)\times\left(0,\infty\right)\right) such that

(2.1) tuf,(0,I)(x,t)=Luf,(0,I)(x,t) for (x,t)(0,I)×(0,),limtuf,(0,I)(x,t)=f(x) for x(0,I) and limx0uf,(0,I)(x,t)=limxIuf,(0,I)(x,t)=0 for t(0,),\begin{array}[]{c}\partial_{t}u_{f,\left(0,I\right)}\left(x,t\right)=Lu_{f,\left(0,I\right)}\left(x,t\right)\text{ for }\left(x,t\right)\in\left(0,I\right)\times\left(0,\infty\right),\\ \lim_{t\searrow}u_{f,\left(0,I\right)}\left(x,t\right)=f\left(x\right)\text{ for }x\in\left(0,I\right)\\ \text{ and }\lim_{x\searrow 0}u_{f,\left(0,I\right)}\left(x,t\right)=\lim_{x\nearrow I}u_{f,\left(0,I\right)}\left(x,t\right)=0\text{ for }t\in\left(0,\infty\right),\end{array}

Once restricted on (0,I)\left(0,I\right), the coefficients (and their derivatives) in (2.1) are all bounded.

We want to find the fundamental solution pI(x,y,t)p_{I}\left(x,y,t\right) to (2.1). Given x>0x>0, let {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} be the unique solution to (1.7), as found in §1.3\mathsection 1.3. We expect that ypI(x,y,t)y\mapsto p_{I}\left(x,y,t\right) coincides with the probability density function of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\}, conditioning on {t<ζ0,IX(x)}\left\{t<\zeta_{0,I}^{X}\left(x\right)\right\}. This probabilistic interpretation of pI(x,y,t)p_{I}\left(x,y,t\right) 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 §1.2\mathsection 1.2, we will transform (1.1) into a diffusion equation that has linear degeneracy at 0. For x(0,I]x\in(0,I], let ϕ(x)\phi\left(x\right) and θ(x)\theta\left(x\right) be defined as in (1.2). It is clear that ϕC2((0,I))\phi\in C^{2}\left(\left(0,I\right)\right), ϕ\phi is strictly increasing, and θC1((0,I))\theta\in C^{1}\left(\left(0,I\right)\right). The constant ν\nu in the definition of θ(x)\theta\left(x\right) is chosen such that

ν=12+limx02b(x)(αxα1a(x)+xαa(x))2xα2a(x)ϕ(x).\nu=\frac{1}{2}+\lim_{x\searrow 0}\frac{2b\left(x\right)-\left(\alpha x^{\alpha-1}a\left(x\right)+x^{\alpha}a^{\prime}\left(x\right)\right)}{2x^{\frac{\alpha}{2}}\sqrt{a\left(x\right)}}\sqrt{\phi\left(x\right)}.

Under (H1) and (H2), it is easy to verify that

ν=1α2α𝕀(0,2)\{1}(α)+b(0)𝕀{1}(α),\nu=\frac{1-\alpha}{2-\alpha}\mathbb{I}_{\left(0,2\right)\backslash\left\{1\right\}}\left(\alpha\right)+b\left(0\right)\mathbb{I}_{\left\{1\right\}}\left(\alpha\right),

and hence ν<1\nu<1. Let J:=ϕ(I)J:=\phi\left(I\right), ψ:(0,J](0,I]\psi:(0,J]\rightarrow(0,I] be the inverse function of ϕ\phi and θ~:=θψ\tilde{\theta}:=\theta\circ\psi. We introduce two more functions on (0,J](0,J]:

(2.2) Θ:z(0,J]Θ(z):=exp(0zθ~(u)2u𝑑u),\Theta:\;z\in(0,J]\mapsto\Theta\left(z\right):=\exp\left(-\int_{0}^{z}\frac{\tilde{\theta}(u)}{2u}du\right),

and

V:z(0,J]V(z):=zΘ′′(z)Θ(z)+(ν+θ~(z))Θ(z)Θ(z),V:\;z\in(0,J]\mapsto V\left(z\right):=z\frac{\Theta^{\prime\prime}\left(z\right)}{\Theta\left(z\right)}+\left(\nu+\tilde{\theta}\left(z\right)\right)\frac{\Theta^{\prime}\left(z\right)}{\Theta\left(z\right)},

or equivalently,

(2.3) V(z)=θ~2(z)4zθ~(z)2+(1ν)θ~(z)2z.V\left(z\right)=-\frac{\tilde{\theta}^{2}\left(z\right)}{4z}-\frac{\tilde{\theta}^{\prime}\left(z\right)}{2}+\left(1-\nu\right)\frac{\tilde{\theta}\left(z\right)}{2z}.

Now we are ready to state the result on the transformation.

Proposition 2.1.

Given fCc((0,I))f\in C_{c}\left(\left(0,I\right)\right), we define

h(z):=fψ(z)Θ(z) for z(0,J).h\left(z\right):=\frac{f\circ\psi\left(z\right)}{\Theta\left(z\right)}\text{ for }z\in\left(0,J\right).

Then, uf,(0,I)(x,t)C2,1((0,I)×(0,))u_{f,\left(0,I\right)}\left(x,t\right)\in C^{2,1}\left(\left(0,I\right)\times\left(0,\infty\right)\right) is a solution to (2.1) if and only if

(2.4) uf,(0,I)(x,t)=Θ(ϕ(x))vh,(0,J)V(ϕ(x),t) for every (x,t)(0,I)×(0,),u_{f,\left(0,I\right)}\left(x,t\right)=\Theta\left(\phi\left(x\right)\right)v_{h,\left(0,J\right)}^{V}\left(\phi\left(x\right),t\right)\text{ for every }\left(x,t\right)\in\left(0,I\right)\times\left(0,\infty\right),

where vh,(0,J)V(z,t)C2,1((0,J)×(0,))v_{h,\left(0,J\right)}^{V}\left(z,t\right)\in C^{2,1}\left(\left(0,J\right)\times\left(0,\infty\right)\right) is a solution to the following problem:

(2.5) tvh,(0,J)V(z,t)=(zz2+νz+V(z))vh,(0,J)V(z,t) for (z,t)(0,J)×(0,),limt0vh,(0,J)V(z,t)=h(z) for z(0,J) and limz0vh,(0,J)V(z,t)=limzJvh,(0,J)V(z,t)=0 for t(0,).\begin{array}[]{c}\begin{array}[]{c}\partial_{t}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=\left(z\partial_{z}^{2}+\nu\partial_{z}+V\left(z\right)\right)v_{h,\left(0,J\right)}^{V}\left(z,t\right)\text{ for }\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right),\\ \lim_{t\searrow 0}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=h\left(z\right)\text{ for }z\in\left(0,J\right)\text{ and }\\ \lim_{z\searrow 0}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=\lim_{z\nearrow J}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=0\text{ for }t\in\left(0,\infty\right).\end{array}\end{array}

We omit the proof of Proposition 2.1 since it can be verified by direct computations. If qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is the fundamental solution to (2.5), then pI(x,y,t)p_{I}\left(x,y,t\right) is connected with qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) following the same relation as the one in (2.4). Set LV:=zz2+νz+V(z)L^{V}:=z\partial_{z}^{2}+\nu\partial_{z}+V\left(z\right). Compared with LL, LVL^{V} 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 LVL^{V} as a perturbation of L0:=zz2+νzL_{0}:=z\partial_{z}^{2}+\nu\partial_{z} and invoking Duhamel’s perturbation method. As a preparation, we state below some technical results on Θ(z)\Theta\left(z\right) and V(z)V\left(z\right).

Lemma 2.2.

Let Θ(z)\Theta\left(z\right) be defined as in (2.2). Then, for every z(0,J]z\in(0,J],

(2.6) Θ(z)={(ψ(z))α4(4z)α4(2α)(a(ψ(z)))14exp(0ψ(z)b(w)2wαa(w)𝑑w) if α1,(ψ(z)4z)14b02(a(ψ(z)))14exp(0ψ(z)12w(b(w)a(w)b(0))𝑑w) if α=1.\Theta\left(z\right)=\begin{cases}\left(\psi\left(z\right)\right)^{\frac{\alpha}{4}}\left(4z\right)^{-\frac{\alpha}{4\left(2-\alpha\right)}}\left(a\left(\psi\left(z\right)\right)\right)^{\frac{1}{4}}\exp\left(-\int_{0}^{\psi\left(z\right)}\frac{b\left(w\right)}{2w^{\alpha}a\left(w\right)}dw\right)&\text{ if }\alpha\neq 1,\\ \left(\frac{\psi\left(z\right)}{4z}\right)^{\frac{1}{4}-\frac{b_{0}}{2}}\left(a\left(\psi\left(z\right)\right)\right)^{\frac{1}{4}}\exp\left(-\int_{0}^{\psi\left(z\right)}\frac{1}{2w}\left(\frac{b\left(w\right)}{a\left(w\right)}-b\left(0\right)\right)dw\right)&\text{ if }\alpha=1.\end{cases}

Hence, there exists constant ΘJ>0\Theta_{J}>0 that can be made explicit (see (6.1) in the Appendix) such that

(2.7) supz(0,J)(Θ(z)1Θ(z))ΘJ.\sup_{z\in\left(0,J\right)}\left(\Theta\left(z\right)\vee\frac{1}{\Theta\left(z\right)}\right)\leq\Theta_{J}.

Let V(z)V\left(z\right) be defined as in (2.3). Then, there exists constant VJ>0V_{J}>0 such that for every z(0,J]z\in(0,J],

(2.8) |V(z)|{VJz12α if α(0,1) and b(0)0,VJz1α2α if α(0,1) and b(0)=0,VJ if α[1,2).\left|V\left(z\right)\right|\leq\begin{cases}V_{J}\cdot z^{-\frac{1}{2-\alpha}}&\text{ if }\alpha\in\left(0,1\right)\text{ and }b\left(0\right)\neq 0,\\ V_{J}\cdot z^{-\frac{1-\alpha}{2-\alpha}}&\text{ if }\alpha\in\left(0,1\right)\text{ and }b\left(0\right)=0,\\ V_{J}&\text{ if }\alpha\in[1,2).\end{cases}

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 α(0,1)\alpha\in\left(0,1\right), the potential function V(z)V\left(z\right) may be singular at 0. This is a generalization of the case considered in [8] where V(z)V\left(z\right) is assumed to be bounded near 0.

2.2. From q(z,w,t)q\left(z,w,t\right) to qJ(z,w,t)q_{J}\left(z,w,t\right)

Let II and JJ be the same as above. As mentioned in the previous subsection, to solve (2.5), we will first consider the analogous problem with LVL^{V} replaced by L0L_{0}. Namely, given gCc((0,I))g\in C_{c}\left(\left(0,I\right)\right), we look for vg,(0,J)(z,t)C2,1((0,J)×(0,))v_{g,\left(0,J\right)}\left(z,t\right)\in C^{2,1}\left(\left(0,J\right)\times\left(0,\infty\right)\right) such that

(2.9) tvg,(0,J)(z,t)=L0vg,(0,J)(z,t) for every (z,t)(0,J)×(0,)limt0vg,(0,J)(z,t)=g(z) for z(0,J) and limz0vg,(0,J)(z,t)=limzJvg,(0,J)(z,t)=0 for t(0,).\begin{array}[]{c}\partial_{t}v_{g,\left(0,J\right)}\left(z,t\right)=L_{0}v_{g,\left(0,J\right)}\left(z,t\right)\text{ for every }\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right)\\ \lim_{t\searrow 0}v_{g,\left(0,J\right)}\left(z,t\right)=g\left(z\right)\text{ for }z\in\left(0,J\right)\\ \text{ and }\lim_{z\searrow 0}v_{g,\left(0,J\right)}\left(z,t\right)=\lim_{z\nearrow J}v_{g,\left(0,J\right)}\left(z,t\right)=0\text{ for }t\in\left(0,\infty\right).\end{array}

Let qJ(z,w,t)q_{J}\left(z,w,t\right) be the fundamental solution to (2.9).

We consider tL0=0\partial_{t}-L_{0}=0 as our model equation. To solve (2.9), we temporarily return to the “global” view and study the model equation on (0,)\left(0,\infty\right) instead of (0,J)\left(0,J\right). That is, for gCc((0,))g\in C_{c}\left(\left(0,\infty\right)\right), we consider the following problem:

(2.10) tvg(z,t)=L0vg(z,t) for every (z,t)(0,)2limt0vg(z,t)=g(z) for z(0,) and limz0vg(z,t)=0 for t(0,).\begin{array}[]{c}\partial_{t}v_{g}\left(z,t\right)=L_{0}v_{g}\left(z,t\right)\text{ for every }\left(z,t\right)\in\left(0,\infty\right)^{2}\\ \lim_{t\searrow 0}v_{g}\left(z,t\right)=g\left(z\right)\text{ for }z\in\left(0,\infty\right)\text{ and }\lim_{z\searrow 0}v_{g}\left(z,t\right)=0\text{ for }t\in\left(0,\infty\right).\end{array}

Let q(z,w,t)q\left(z,w,t\right) be the fundamental solution to (2.10). In fact, q(z,w,t)q\left(z,w,t\right) is the starting point of our “journey”, and from q(z,w,t)q\left(z,w,t\right) we will derive the (fundamental) solutions to all the concerned equations.

The stochastic differential equation corresponding to the model equation is that, given z>0z>0,

(2.11) dY(z,t)=2Y(z,t)dB(t)+νdt for t0 with Y(z,0)z.dY\left(z,t\right)=\sqrt{2Y\left(z,t\right)}dB\left(t\right)+\nu dt\text{ for }t\geq 0\text{ with }Y\left(z,0\right)\equiv z.

It follows from the discussions in §1.2\mathsection 1.2 that (2.11) is well posed, and hence there exists a unique solution {Y(z,t):t0}\left\{Y\left(z,t\right):t\geq 0\right\} 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 ν\nu determines the attainability of the boundary 0. Under (H1) and (H2), we have that ν<1\nu<1, and hence 0 is either an exit boundary or a regular boundary. This is to say that, no matter what zz is, {Y(z,t):t0}\left\{Y\left(z,t\right):t\geq 0\right\} hits 0 with a positive probability in finite time. For more details on the topic of boundary classification, we refer readers to §15\mathsection 15 of [26].

The operator L0L_{0}, as well as (2.10) and (2.10), has been well studied in [8]. Below we will review some useful facts about q(z,w,t)q\left(z,w,t\right), vg(z,t)v_{g}\left(z,t\right) and their connections to {Y(z,t):t0}\left\{Y\left(z,t\right):t\geq 0\right\}. The details can be found in §2\mathsection 2 of [8].

Proposition 2.4.

(Proposition 2.1, 2.3 of [8]) The fundamental solution to (2.10) is

(2.12) q(z,w,t):=z1ν2wν12tez+wtI1ν(2zwt)=z1νt2νez+wtn=0(zw)nt2nn!Γ(n+2ν)q\left(z,w,t\right):=\frac{z^{\frac{1-\nu}{2}}w^{\frac{\nu-1}{2}}}{t}e^{-\frac{z+w}{t}}I_{1-\nu}\left(2\frac{\sqrt{zw}}{t}\right)=\frac{z^{1-\nu}}{t^{2-\nu}}e^{-\frac{z+w}{t}}\sum_{n=0}^{\infty}\frac{\left(zw\right)^{n}}{t^{2n}n!\Gamma\left(n+2-\nu\right)}

for (z,w,t)(0,)3\left(z,w,t\right)\in\left(0,\infty\right)^{3}, where I1νI_{1-\nu} is the modified Bessel function. q(z,w,t)q\left(z,w,t\right) is smooth on (0,)3\left(0,\infty\right)^{3}, and for every (z,w,t)(0,)3\left(z,w,t\right)\in\left(0,\infty\right)^{3},

(2.13) z1νt2νΓ(2ν)ez+wtq(z,w,t)(z1νt2ν)e(zw)2t,\frac{z^{1-\nu}}{t^{2-\nu}\Gamma\left(2-\nu\right)}e^{-\frac{z+w}{t}}\leq q\left(z,w,t\right)\leq\left(\frac{z^{1-\nu}}{t^{2-\nu}}\right)e^{-\frac{\left(\sqrt{z}-\sqrt{w}\right)^{2}}{t}},

and

(2.14) w1νq(z,w,t)=z1νq(w,z,t).w^{1-\nu}q\left(z,w,t\right)=z^{1-\nu}q\left(w,z,t\right).

Given gCc((0,))g\in C_{c}\left(\left(0,\infty\right)\right), if

vg(z,t):=0g(w)q(z,w,t)𝑑w for (z,t)(0,),v_{g}\left(z,t\right):=\int_{0}^{\infty}g\left(w\right)q\left(z,w,t\right)dw\text{ for }\left(z,t\right)\in\left(0,\infty\right),

then vg(z,t)v_{g}\left(z,t\right) is the unique solution in C2,1((0,)2)C^{2,1}\left(\left(0,\infty\right)^{2}\right) to (2.10), and vg(z,t)v_{g}\left(z,t\right) is smooth on (0,)2\left(0,\infty\right)^{2}. Moreover,

(2.15) vg(z,t)=𝔼[g(Y(z,t));t<ζ0Y(z)] for every (z,t)(0,)2,v_{g}\left(z,t\right)=\mathbb{E}\left[g\left(Y\left(z,t\right)\right);t<\zeta_{0}^{Y}\left(z\right)\right]\text{ for every }\left(z,t\right)\in\left(0,\infty\right)^{2},

which implies that for every Borel set Γ(0,)\Gamma\subseteq\left(0,\infty\right),

(2.16) Γq(z,w,t)𝑑w=(Y(z,t)Γ,t<ζ0Y(z)).\int_{\Gamma}q\left(z,w,t\right)dw=\mathbb{P}\left(Y\left(z,t\right)\in\Gamma,t<\zeta_{0}^{Y}\left(z\right)\right).

Finally, q(z,w,t)q\left(z,w,t\right) satisfies the Chapman-Kolmogorov equation, i.e., for every z,w>0z,w>0 and t,s>0t,s>0,

(2.17) q(z,w,t+s)=0q(z,ξ,t)q(ξ,w,s)𝑑ξ.q\left(z,w,t+s\right)=\int_{0}^{\infty}q\left(z,\xi,t\right)q\left(\xi,w,s\right)d\xi.

It is clear from (2.16) that, for every (z,t)(0,)2\left(z,t\right)\in\left(0,\infty\right)^{2}, wq(z,w,t)w\mapsto q\left(z,w,t\right) is the probability density function of Y(z,t)Y\left(z,t\right), provided that t<ζ0Y(z)t<\zeta_{0}^{Y}\left(z\right). Now we turn our attention to qJ(z,w,t)q_{J}\left(z,w,t\right), the fundamental solution to (2.5) which has an extra Dirichlet boundary at JJ. Intuitively speaking, to get qJ(z,w,t)q_{J}\left(z,w,t\right), we need to remove from q(z,w,t)q\left(z,w,t\right) the “contribution” of Y(z,t)Y\left(z,t\right) once Y(z,t)Y\left(z,t\right) exists the interval (0,J)\left(0,J\right). Based on this idea combined with the fact that {Y(z,t):t0}\left\{Y\left(z,t\right):t\geq 0\right\} is a strong Markov process, we define

(2.18) qJ(z,w,t):=q(z,w,t)𝔼[q(J,w,tζJY(z));ζJY(z)tζ0Y(z)]\begin{split}q_{J}\left(z,w,t\right)&:=q\left(z,w,t\right)-\mathbb{E}\left[q\left(J,w,t-\zeta_{J}^{Y}\left(z\right)\right);\zeta_{J}^{Y}\left(z\right)\leq t\wedge\zeta_{0}^{Y}\left(z\right)\right]\end{split}

for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right). Again, by (2.16), we see that for every Borel set Γ(0,J)\Gamma\subseteq\left(0,J\right),

(2.19) ΓqJ(z,w,t)𝑑w=(Y(z,t)Γ,t<ζ0Y(z))(Y(z,t)Γ,ζJY(z)t<ζ0Y(z))=(Y(z,t)Γ,t<ζ0,JY(z)).\begin{split}\int_{\Gamma}q_{J}\left(z,w,t\right)dw&=\mathbb{P}\left(Y\left(z,t\right)\in\Gamma,t<\zeta_{0}^{Y}\left(z\right)\right)-\mathbb{P}\left(Y\left(z,t\right)\in\Gamma,\zeta_{J}^{Y}\left(z\right)\leq t<\zeta_{0}^{Y}\left(z\right)\right)\\ &=\mathbb{P}\left(Y\left(z,t\right)\in\Gamma,t<\zeta_{0,J}^{Y}\left(z\right)\right).\end{split}

In other words, qJ(z,w,t)q_{J}\left(z,w,t\right) is the probability density function of Y(z,t)Y\left(z,t\right) provided that t<ζ0,JY(z)t<\zeta_{0,J}^{Y}\left(z\right).

To better analyze qJ(z,w,t)q_{J}\left(z,w,t\right), we need the following probability estimates on the hitting times of Y(z,t)Y\left(z,t\right).

Lemma 2.5.

For every z(0,J)z\in\left(0,J\right),

(2.20) (ζJY(z)ζ0Y(z))=z1νJ1ν.\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq\zeta_{0}^{Y}\left(z\right)\right)=\frac{z^{1-\nu}}{J^{1-\nu}}.

For t>0t>0 and Jz|ν|tJ-z\geq\left|\nu\right|t, we have that

(2.21) (ζJY(z)t)exp((Jztν)24tJ).\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq t\right)\leq\exp\left(-\frac{\left(J-z-t\nu\right)^{2}}{4tJ}\right).

Furthermore, almost surely

limJζJY(z)= and limzJζJY(z)=0.\lim_{J\nearrow\infty}\zeta_{J}^{Y}\left(z\right)=\infty\text{ and }\lim_{z\nearrow J}\zeta_{J}^{Y}\left(z\right)=0.
Proof.

Based on (2.11), one can apply Itô’s formula (see, e.g., §5\mathsection 5 of [25]) to check that, for every z>0z>0, {(Y(z,t))1ν:t0}\left\{\left(Y\left(z,t\right)\right)^{1-\nu}:t\geq 0\right\} is a local martingale, and hence by Doob’s stopping time theorem (see, e.g., §8\mathsection 8 of [33]), {(Y(z,tζ0,JY(z)))1ν:t0}\left\{\left(Y\left(z,t\wedge\zeta_{0,J}^{Y}\left(z\right)\right)\right)^{1-\nu}:t\geq 0\right\} is a bounded martingale. Thus,

z1ν=𝔼[(Y(z,ζ0,JY(z)))1ν]=(ζJY(z)ζ0Y(z))J1ν;z^{1-\nu}=\mathbb{E}\left[\left(Y\left(z,\zeta_{0,J}^{Y}\left(z\right)\right)\right)^{1-\nu}\right]=\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq\zeta_{0}^{Y}\left(z\right)\right)J^{1-\nu};

To show (2.21), we check that for every z(0,J)z\in\left(0,J\right) and every λ>0\lambda>0, if

E(z,t):=exp(λY(z,t)λνtλ20tY(z,s)𝑑s) for t0,E\left(z,t\right):=\exp\left(\lambda Y\left(z,t\right)-\lambda\nu t-\lambda^{2}\int_{0}^{t}Y\left(z,s\right)ds\right)\text{ for }t\geq 0,

then {E(z,tζJY(z)):t0}\left\{E\left(z,t\wedge\zeta_{J}^{Y}\left(z\right)\right):t\geq 0\right\} is a martingale. By a similar argument as above and Fatou’s lemma, we get that

(2.22) eλJ𝔼[e(λν+λ2J)ζJY(z);ζJY(z)<]lim inft𝔼[E(z,tζJY(z))]=eλz.e^{\lambda J}\mathbb{E}\left[e^{-\left(\lambda\nu+\lambda^{2}J\right)\zeta_{J}^{Y}\left(z\right)};\zeta_{J}^{Y}\left(z\right)<\infty\right]\leq\liminf_{t\nearrow\infty}\mathbb{E}\left[E\left(z,t\wedge\zeta_{J}^{Y}\left(z\right)\right)\right]=e^{\lambda z}.

Set λ:=Jzνt2tJ\lambda:=\frac{J-z-\nu t}{2tJ}. Since J>z+t|ν|J>z+t\left|\nu\right|, we have that λ>0\lambda>0 and λν+λ2J>0\lambda\nu+\lambda^{2}J>0. Therefore, a simple application of Markov’s inequality leads to

(ζJY(z)t)=(e(λν+λ2J)ζJY(z)e(λν+λ2J)t)e(λν+λ2J)t𝔼[e(λν+λ2J)ζJY(z);ζJY(z)<]exp(λ2tJλ(Jzνt)).\begin{split}\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq t\right)&=\mathbb{P}\left(e^{-\left(\lambda\nu+\lambda^{2}J\right)\zeta_{J}^{Y}\left(z\right)}\geq e^{-\left(\lambda\nu+\lambda^{2}J\right)t}\right)\\ &\leq e^{\left(\lambda\nu+\lambda^{2}J\right)t}\mathbb{E}\left[e^{-\left(\lambda\nu+\lambda^{2}J\right)\zeta_{J}^{Y}\left(z\right)};\zeta_{J}^{Y}\left(z\right)<\infty\right]\\ &\leq\exp\left(\lambda^{2}tJ-\lambda\left(J-z-\nu t\right)\right).\end{split}

Plugging the value of λ\lambda into the right hand side yields (2.21). The fact that ζJY(z)\zeta_{J}^{Y}\left(z\right) converges to \infty as JJ\nearrow\infty almost surely follows from (2.21) and the monotonicity of ζJY(z)\zeta_{J}^{Y}\left(z\right) in JJ.

Finally, we observe that ζ:=limzJζJY(z)\zeta:=\lim_{z\nearrow J}\zeta_{J}^{Y}\left(z\right) exists almost surely. Take λ\lambda\in\mathbb{R} such that λν0\lambda\nu\geq 0. It follows from (2.22) and the reverse Fatou’s lemma that

eλJ𝔼[eλνζ]𝔼[lim supzJE(z,ζJY(z))]lim supzJ𝔼[E(z,ζJY(z))]lim supzJlim supt𝔼[E(z,tζJY(z))]=eλJ,\begin{split}e^{\lambda J}\mathbb{E}\left[e^{-\lambda\nu\zeta}\right]&\geq\mathbb{E}\left[\limsup_{z\nearrow J}E\left(z,\zeta_{J}^{Y}\left(z\right)\right)\right]\geq\limsup_{z\nearrow J}\mathbb{E}\left[E\left(z,\zeta_{J}^{Y}\left(z\right)\right)\right]\\ &\geq\limsup_{z\nearrow J}\,\limsup_{t\nearrow\infty}\mathbb{E}\left[E\left(z,t\wedge\zeta_{J}^{Y}\left(z\right)\right)\right]=e^{\lambda J},\end{split}

which implies that 𝔼[eλνζ]=1\mathbb{E}\left[e^{-\lambda\nu\zeta}\right]=1 and hence ζ=0\zeta=0 almost surely. ∎

Proposition 2.6.

Let qJ(z,w,t)q_{J}\left(z,w,t\right) be defined as in (2.18). Then, qJ(z,w,t)q_{J}\left(z,w,t\right) is continuous on (0,J)2×(0,)\left(0,J\right)^{2}\times\left(0,\infty\right), and for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right), we have that

(2.23) w1νqJ(z,w,t)=z1νqJ(w,z,t).w^{1-\nu}q_{J}\left(z,w,t\right)=z^{1-\nu}q_{J}\left(w,z,t\right).

qJ(z,w,t)q_{J}\left(z,w,t\right) satisfies the Chapman-Kolmogorov equation, i.e., for every z,w(0,J)z,w\in\left(0,J\right) and t,s>0t,s>0,

(2.24) qJ(z,w,t+s)=0JqJ(z,ξ,t)qJ(ξ,w,s)𝑑ξ.q_{J}\left(z,w,t+s\right)=\int_{0}^{J}q_{J}\left(z,\xi,t\right)q_{J}\left(\xi,w,s\right)d\xi.

For every w(0,J)w\in\left(0,J\right), (z,t)qJ(z,w,t)\left(z,t\right)\mapsto q_{J}\left(z,w,t\right) is a smooth solution to the Kolmogorov backward equation corresponding to L0L_{0}:

(2.25) tqJ(z,w,t)=L0qJ(z,w,t);\partial_{t}q_{J}\left(z,w,t\right)=L_{0}q_{J}\left(z,w,t\right);

for every z(0,J)z\in\left(0,J\right), (w,t)qJ(z,w,t)\left(w,t\right)\mapsto q_{J}\left(z,w,t\right) is a smooth solution to the Kolmogorov forward equation corresponding to L0L_{0}:

(2.26) tqJ(z,w,t)=L0qJ(z,w,t)\partial_{t}q_{J}\left(z,w,t\right)=L_{0}^{*}q_{J}\left(z,w,t\right)

where L0=ww2+(2ν)wL_{0}^{*}=w\partial_{w}^{2}+\left(2-\nu\right)\partial_{w} is the formal adjoint of L0L_{0}.

Moreover, qJ(z,w,t)q_{J}\left(z,w,t\right) is the fundamental solution to (2.9). Given gCc((0,J))g\in C_{c}\left(\left(0,J\right)\right), if

(2.27) vg,(0,J)(z,t):=0Jg(w)qJ(z,w,t)𝑑w for (z,t)(0,J)×(0,),v_{g,\left(0,J\right)}\left(z,t\right):=\int_{0}^{J}g\left(w\right)q_{J}\left(z,w,t\right)dw\text{ for }\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right),

then vg,(0,J)(z,t)v_{g,\left(0,J\right)}\left(z,t\right) is the unique solution in C,2,1((0,J)×(0,))C^{,2,1}\left(\left(0,J\right)\times\left(0,\infty\right)\right) to (2.9), and vg,(0,I)(z,t)v_{g,\left(0,I\right)}\left(z,t\right) is smooth on (0,J)×(0,)\left(0,J\right)\times\left(0,\infty\right).

Proof.

We start with (2.24) since its proof is straightforward. Given z,w(0,J)z,w\in\left(0,J\right), t,s>0t,s>0 and Borel set Γ(0,J)\Gamma\subseteq\left(0,J\right), by (2.19) and the strong Markov property of Y(z,t)Y\left(z,t\right), we can write

ΓqJ(z,w,t+s)𝑑w=(Y(z,t+s)Γ,t+s<ζ0,JY(z))=Γ𝔼[qJ(Y(z,t),w,s);t<ζ0,JY(z)]𝑑w=Γ0JqJ(z,ξ,t)qJ(ξ,w,s)𝑑ξ𝑑w,\begin{split}\int_{\Gamma}q_{J}\left(z,w,t+s\right)dw&=\mathbb{P}\left(Y\left(z,t+s\right)\in\Gamma,t+s<\zeta_{0,J}^{Y}\left(z\right)\right)\\ &=\int_{\Gamma}\mathbb{E}\left[q_{J}\left(Y\left(z,t\right),w,s\right);t<\zeta_{0,J}^{Y}\left(z\right)\right]dw\\ &=\int_{\Gamma}\int_{0}^{J}q_{J}\left(z,\xi,t\right)q_{J}\left(\xi,w,s\right)d\xi dw,\end{split}

which implies (2.24). Next, given t>0t>0, we take any mm\in\mathbb{N}, any 0=s0<s1<s2<<sm1<sm=t0=s_{0}<s_{1}<s_{2}<\cdots<s_{m-1}<s_{m}=t and φ0,φ2,,φmCc((0,J))\varphi_{0},\varphi_{2},\cdots,\varphi_{m}\in C_{c}\left(\left(0,J\right)\right). By (2.14), (2.19) and, again, the Markov property of Y(z,t)Y\left(z,t\right), we have that

0J𝔼[k=0mφk(Y(z,sk));t<ζ0,JY(z)]dzz1ν=0J∫⋯∫(0,J)mφ0(z)k=1mφk(ξk)qJ(z,ξ1,s1)qJ(ξ1,ξ2,s2s1)qJ(ξm1,ξm,tsm1)dξmdξ1dzz1ν=0J∫⋯∫(0,J)mφ0(z)k=1mφk(ξk)qJ(ξ1,z,t(ts1))qJ(ξ2,ξ1,(ts1)(ts2))qJ(ξm,ξm1,tsm1)dξmξm1νdξm1dξ1dz=0J𝔼[k=0mφk(Y(ξm,tsk));t<ζ0Y(ξm)]dξmξm1ν.\begin{split}&\int_{0}^{J}\mathbb{E}\left[\prod_{k=0}^{m}\varphi_{k}\left(Y\left(z,s_{k}\right)\right);t<\zeta_{0,J}^{Y}\left(z\right)\right]\frac{dz}{z^{1-\nu}}\\ =&\int_{0}^{J}\idotsint_{\left(0,J\right)^{m}}\varphi_{0}\left(z\right)\prod_{k=1}^{m}\varphi_{k}\left(\xi_{k}\right)q_{J}\left(z,\xi_{1},s_{1}\right)q_{J}\left(\xi_{1},\xi_{2},s_{2}-s_{1}\right)\\ &\hskip 56.9055pt\hskip 56.9055pt\cdots q_{J}\left(\xi_{m-1},\xi_{m},t-s_{m-1}\right)d\xi_{m}\cdots d\xi_{1}\frac{dz}{z^{1-\nu}}\\ =&\int_{0}^{J}\idotsint_{\left(0,J\right)^{m}}\varphi_{0}\left(z\right)\prod_{k=1}^{m}\varphi_{k}\left(\xi_{k}\right)q_{J}\left(\xi_{1},z,t-\left(t-s_{1}\right)\right)q_{J}\left(\xi_{2},\xi_{1},\left(t-s_{1}\right)-\left(t-s_{2}\right)\right)\\ &\hskip 56.9055pt\hskip 56.9055pt\hskip 28.45274pt\cdots q_{J}\left(\xi_{m},\xi_{m-1},t-s_{m-1}\right)\frac{d\xi_{m}}{\xi_{m}^{1-\nu}}d\xi_{m-1}\cdots d\xi_{1}dz\\ =&\int_{0}^{J}\mathbb{E}\left[\prod_{k=0}^{m}\varphi_{k}\left(Y\left(\xi_{m},t-s_{k}\right)\right);t<\zeta_{0}^{Y}\left(\xi_{m}\right)\right]\frac{d\xi_{m}}{\xi_{m}^{1-\nu}}.\end{split}

Since tY(z,t)t\mapsto Y\left(z,t\right) is almost surely continuous and s0,,sm,φ0,,φms_{0},\cdots,s_{m},\varphi_{0},\cdots,\varphi_{m} are chosen arbitrarily, the relation above implies that for every measurable functional FF on C([0,t])C\left(\left[0,t\right]\right),

0J𝔼[F(Y(z,)|[0,t]);t<ζ0,JY(z)]dzz1ν=0J𝔼[F(Y(w,)|[0,t]);t<ζ0,JY(w)]dww1ν\int_{0}^{J}\mathbb{E}\left[F\left(\left.Y\left(z,\cdot\right)\right|_{\left[0,t\right]}\right);t<\zeta_{0,J}^{Y}\left(z\right)\right]\frac{dz}{z^{1-\nu}}=\int_{0}^{J}\mathbb{E}\left[F\left(\left.\overleftarrow{Y}\left(w,\cdot\right)\right|_{\left[0,t\right]}\right);t<\zeta_{0,J}^{Y}\left(w\right)\right]\frac{dw}{w^{1-\nu}}

where Y(z,s):=Y(z,ts)\overleftarrow{Y}\left(z,s\right):=Y\left(z,t-s\right) for every s[0,t]s\in\left[0,t\right]. In particular, for arbitrary φ\varphi, φCc((0,J))\varphi^{*}\in C_{c}\left(\left(0,J\right)\right), if FF is chosen such that for every y()C([0,t])y\left(\cdot\right)\in C\left(\left[0,t\right]\right),

F(y())={φ(y(0))φ(y(t)),if 0<y(s)<J for every s[0,t],0otherwise,F\left(y\left(\cdot\right)\right)=\begin{cases}\varphi\left(y\left(0\right)\right)\varphi^{*}\left(y\left(t\right)\right),&\text{if }0<y\left(s\right)<J\text{ for every }s\in\left[0,t\right],\\ 0&\text{otherwise},\end{cases}

then we have that

0J0Jφ(z)φ(w)qJ(z,w,t)dwdzz1ν=0J0Jφ(z)φ(w)qJ(w,z,t)dwdzw1ν.\int_{0}^{J}\int_{0}^{J}\varphi\left(z\right)\varphi^{*}\left(w\right)q_{J}\left(z,w,t\right)\frac{dwdz}{z^{1-\nu}}=\int_{0}^{J}\int_{0}^{J}\varphi\left(z\right)\varphi^{*}\left(w\right)q_{J}\left(w,z,t\right)\frac{dwdz}{w^{1-\nu}}.

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 φCc((0,J))\varphi\in C_{c}^{\infty}\left(\left(0,J\right)\right) and consider vφ,(0,J)(z,t)v_{\varphi,\left(0,J\right)}\left(z,t\right), which, according to (2.19), can be written as

vφ,(0,J)(z,t)=𝔼[φ(Y(z,t));t<ζ0,JY(z)] for every (z,t)(0,J)×(0,).v_{\varphi,\left(0,J\right)}\left(z,t\right)=\mathbb{E}\left[\varphi\left(Y\left(z,t\right)\right);t<\zeta_{0,J}^{Y}\left(z\right)\right]\text{ for every }\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right).

As reviewed in §1.2\mathsection 1.2, for every z(0,J)z\in\left(0,J\right),

{φ(Y(z,tζ0,JY(z)))0tζ0,JY(z)(L0φ)(Y(z,s))𝑑s:t0}\left\{\varphi\left(Y\left(z,t\wedge\zeta_{0,J}^{Y}\left(z\right)\right)\right)-\int_{0}^{t\wedge\zeta_{0,J}^{Y}\left(z\right)}\left(L_{0}\varphi\right)\left(Y\left(z,s\right)\right)ds:t\geq 0\right\}

is a bounded martingale. Thus,

φ(z)=𝔼[φ(Y(z,t));t<ζ0,JY(z)]0t𝔼[L0φ(Y(z,s));s<ζ0,JY(z)]𝑑s=vφ,(0,J)(z,t)0t0JL0φ(w)qJ(z,w,s)𝑑w𝑑s,\begin{split}\varphi\left(z\right)&=\mathbb{E}\left[\varphi\left(Y\left(z,t\right)\right);t<\zeta_{0,J}^{Y}\left(z\right)\right]-\int_{0}^{t}\mathbb{E}\left[L_{0}\varphi\left(Y\left(z,s\right)\right);s<\zeta_{0,J}^{Y}\left(z\right)\right]ds\\ &=v_{\varphi,\left(0,J\right)}\left(z,t\right)-\int_{0}^{t}\int_{0}^{J}L_{0}\varphi\left(w\right)q_{J}\left(z,w,s\right)dwds,\end{split}

and hence

t(0Jφ(w)qJ(z,w,t)𝑑w)=0JL0φ(w)qJ(z,w,t)𝑑w.\partial_{t}\left(\int_{0}^{J}\varphi\left(w\right)q_{J}\left(z,w,t\right)dw\right)=\int_{0}^{J}L_{0}\varphi\left(w\right)q_{J}\left(z,w,t\right)dw.

This means that for every z(0,J)z\in\left(0,J\right), (w,t)qJ(z,w,t)\left(w,t\right)\mapsto q_{J}\left(z,w,t\right) solves the equation (tL0)qJ(z,w,t)=0\left(\partial_{t}-L_{0}^{*}\right)q_{J}\left(z,w,t\right)=0 in the sense of distribution. Since tL0\partial_{t}-L_{0}^{*} is a hypoelliptic operator (see, e.g., §7.4\mathsection 7.4 of [32]), (w,t)qJ(z,w,t)\left(w,t\right)\mapsto q_{J}\left(z,w,t\right) is a smooth solution to (2.26).

For (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right), we set

(2.28) r(z,w,t):=q(z,w,t)qJ(z,w,t)=𝔼[q(J,w,tζJY(z));ζJY(z)tζ0Y(z)].r\left(z,w,t\right):=q\left(z,w,t\right)-q_{J}\left(z,w,t\right)=\mathbb{E}\left[q\left(J,w,t-\zeta_{J}^{Y}\left(z\right)\right);\zeta_{J}^{Y}\left(z\right)\leq t\wedge\zeta_{0}^{Y}\left(z\right)\right].

Then, for every w(0,J)w\in\left(0,J\right), (z,t)r(z,w,t)\left(z,t\right)\mapsto r\left(z,w,t\right) is smooth on (0,J)×(0,)\left(0,J\right)\times\left(0,\infty\right). It is easy to see that wr(z,w,t)w\mapsto r\left(z,w,t\right) is equicontinuous in (z,t)\left(z,t\right) from any bounded subset of (0,J)×(0,)\left(0,J\right)\times\left(0,\infty\right), which implies that r(z,w,t)r\left(z,w,t\right), as well as qJ(z,w,t)q_{J}\left(z,w,t\right), is continuous on (0,J)2×(0,)\left(0,J\right)^{2}\times\left(0,\infty\right).

We proceed to the proof of the last statement. Again, by the hypoellipticity of tL0\partial_{t}-L_{0}, to show that vg,(0,J)(z,t)v_{g,\left(0,J\right)}\left(z,t\right) is a smooth solution to the model equation, we only need to show that it solves the equation as a distribution. Let us take φCc((0,J))\varphi\in C_{c}^{\infty}\left(\left(0,J\right)\right) and consider, for every t>0t>0,

φ,vg,(0,J)(,t):=0Jφ(z)vg,(0,J)(z,t)𝑑z=0J0Jφ(z)qJ(z,w,t)𝑑zg(w)𝑑w.\left\langle\varphi,v_{g,\left(0,J\right)}\left(\cdot,t\right)\right\rangle:=\int_{0}^{J}\varphi\left(z\right)v_{g,\left(0,J\right)}\left(z,t\right)dz=\int_{0}^{J}\int_{0}^{J}\varphi\left(z\right)q_{J}\left(z,w,t\right)dzg\left(w\right)dw.

By (2.25), we have that

ddtφ,vg,(0,J)(,t)\displaystyle\frac{d}{dt}\left\langle\varphi,v_{g,\left(0,J\right)}\left(\cdot,t\right)\right\rangle =0J(0Jφ(z)(L0qJ(,w,t))(z)𝑑z)g(w)𝑑w\displaystyle=\int_{0}^{J}\left(\int_{0}^{J}\varphi\left(z\right)\left(L_{0}q_{J}\left(\cdot,w,t\right)\right)\left(z\right)dz\right)g\left(w\right)dw
=0J0JL0φ(z)qJ(z,w,t)g(w)𝑑w𝑑z\displaystyle=\int_{0}^{J}\int_{0}^{J}L_{0}^{*}\varphi\left(z\right)q_{J}\left(z,w,t\right)g\left(w\right)dwdz
=L0φ,vg,(0,J)(,t).\displaystyle=\left\langle L_{0}^{*}\varphi,v_{g,\left(0,J\right)}\left(\cdot,t\right)\right\rangle.

The only remaining thing to do is to verify that vg,(0,J)(z,t)v_{g,\left(0,J\right)}\left(z,t\right) satisfies the initial value and the boundary value conditions in (2.9). Given gCc((0,J))g\in C_{c}\left(\left(0,J\right)\right), by (2.15) and (2.19), we have that for every z(0,I)z\in\left(0,I\right),

|vg,(0,J)(z,t)vg(z,t)|𝔼[|g(Y(z,t))|;ζJY(z)t<ζ0Y(z)]gu(ζJY(z)t)\begin{split}\left|v_{g,\left(0,J\right)}\left(z,t\right)-v_{g}\left(z,t\right)\right|&\leq\mathbb{E}\left[\left|g\left(Y\left(z,t\right)\right)\right|;\zeta_{J}^{Y}\left(z\right)\leq t<\zeta_{0}^{Y}\left(z\right)\right]\\ &\leq\left\|g\right\|_{u}\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq t\right)\end{split}

which, according to (2.21), goes to 0 as t0t\searrow 0, and the convergence is uniformly fast for zz on any compact subset of (0,J)\left(0,J\right). Therefore,

limt0vg,(0,J)(z,t)=limt0vg(z,t)=0.\lim_{t\searrow 0}v_{g,\left(0,J\right)}\left(z,t\right)=\lim_{t\searrow 0}v_{g}\left(z,t\right)=0.

To verify that vg,(0,I)(z,t)v_{g,\left(0,I\right)}\left(z,t\right) satisfies the boundary condition, it is sufficient to show that

limz0r(z,w,t)=0 and limzJr(z,w,t)=q(J,w,t) for every (w,t)(0,J)×(0,).\lim_{z\searrow 0}r\left(z,w,t\right)=0\text{ and }\lim_{z\nearrow J}r\left(z,w,t\right)=q\left(J,w,t\right)\text{ for every }\left(w,t\right)\in\left(0,J\right)\times\left(0,\infty\right).

We observe that, by (2.13), q(J,w,tζJY(z))q\left(J,w,t-\zeta_{J}^{Y}\left(z\right)\right) is bounded uniformly in zz by

J1ν(Jw)2(ν2)(2νe)2νJ^{1-\nu}\left(\sqrt{J}-\sqrt{w}\right)^{2\left(\nu-2\right)}\left(\frac{2-\nu}{e}\right)^{2-\nu}

where we used the fact that

sups>0s2νes=(2νe)2ν.\sup_{s>0}\,s^{2-\nu}e^{-s}=\left(\frac{2-\nu}{e}\right)^{2-\nu}.

Therefore, (2.20) implies that

limz0r(z,w,t)J1ν((Jw)22νe)2νlimz0(ζJY(z)ζ0Y(z))=0.\lim_{z\searrow 0}r\left(z,w,t\right)\leq J^{1-\nu}\left(\left(\sqrt{J}-\sqrt{w}\right)^{2}\frac{2-\nu}{e}\right)^{2-\nu}\lim_{z\searrow 0}\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq\zeta_{0}^{Y}\left(z\right)\right)=0.

Finally, the last statement in Lemma 2.5 and the dominated convergence theorem lead to

limzJr(z,w,t)=q(J,w,t).\lim_{z\nearrow J}r\left(z,w,t\right)=q\left(J,w,t\right).

We will close this subsection with a result on the comparison between qJ(z,w,t)q_{J}\left(z,w,t\right) and q(z,w,t)q\left(z,w,t\right). Intuitively speaking, given z(0,J)z\in\left(0,J\right) sufficiently far from the boundary JJ and tt sufficiently small, Y(z,t)Y\left(z,t\right) would not have exited (0,J)\left(0,J\right) by time tt, which means that q(z,w,t)q\left(z,w,t\right) and qJ(z,w,t)q_{J}\left(z,w,t\right) should be close to each other. We will make this statement rigorous by proving that, as t0t\searrow 0, qJ(z,w,t)/q(z,w,t)q_{J}\left(z,w,t\right)/q\left(z,w,t\right) converges to 1 uniformly fast in (z,w)\left(z,w\right) away from JJ.

Corollary 2.7.

Set tJ:=4J9(2ν)t_{J}:=\frac{4J}{9\left(2-\nu\right)}. Then, for every t(0,tJ)t\in\left(0,t_{J}\right),

(2.29) sup(z,w)(0,19J)2|qJ(z,w,t)q(z,w,t)1|exp(2J9t).\sup_{\left(z,w\right)\in\left(0,\frac{1}{9}J\right)^{2}}\left|\frac{q_{J}\left(z,w,t\right)}{q\left(z,w,t\right)}-1\right|\leq\exp\left(-\frac{2J}{9t}\right).
Proof.

It is easy to verify that tJt_{J} is chosen such that the function ssν2exp(4J9s)s\mapsto s^{\nu-2}\exp\left(-\frac{4J}{9s}\right) is increasing on (0,tJ)\left(0,t_{J}\right). By (2.13) and (2.20), we have that for every (z,w)(0,19J)\left(z,w\right)\in\left(0,\frac{1}{9}J\right) and t(0,tJ)t\in\left(0,t_{J}\right),

|qJ(z,w,t)q(z,w,t)1|=|r(z,w,t)|q(z,w,t)J1ν(ζJY(z)ζ0Y(z))sups(0,t)sν2exp((Jw)2s)z1νtν2exp(z+wt)sups(0,t)sν2exp(4J9s)tν2exp(z+wt)exp(4J9t+z+wt)exp(2J9t).\begin{split}\left|\frac{q_{J}\left(z,w,t\right)}{q\left(z,w,t\right)}-1\right|=\frac{\left|r\left(z,w,t\right)\right|}{q\left(z,w,t\right)}&\leq\frac{J^{1-\nu}\mathbb{P}\left(\zeta_{J}^{Y}\left(z\right)\leq\zeta_{0}^{Y}\left(z\right)\right)\cdot\sup_{s\in\left(0,t\right)}s^{\nu-2}\exp\left(-\frac{\left(\sqrt{J}-\sqrt{w}\right)^{2}}{s}\right)}{z^{1-\nu}t^{\nu-2}\exp\left(-\frac{z+w}{t}\right)}\\ &\leq\frac{\sup_{s\in\left(0,t\right)}s^{\nu-2}\exp\left(-\frac{4J}{9s}\right)}{t^{\nu-2}\exp\left(-\frac{z+w}{t}\right)}\leq\exp\left(-\frac{4J}{9t}+\frac{z+w}{t}\right)\leq\exp\left(\frac{-2J}{9t}\right).\end{split}

3. Localized Equation

3.1. From qJ(z,w,t)q_{J}\left(z,w,t\right) to qJV(z,w,t)q_{J}^{V}\left(z,w,t\right)

Now we get down to solving (2.5) by the perturbation method of Duhamel. First we want to find a function qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) on (0,J)2×(0,)\left(0,J\right)^{2}\times\left(0,\infty\right) that solves the integral equation

(3.1) qJV(z,w,t)=qJ(z,w,t)+0t0JqJ(z,ξ,ts)qJV(ξ,w,s)V(ξ)𝑑ξ𝑑sq_{J}^{V}\left(z,w,t\right)=q_{J}\left(z,w,t\right)+\int_{0}^{t}\int_{0}^{J}q_{J}\left(z,\xi,t-s\right)q_{J}^{V}\left(\xi,w,s\right)V\left(\xi\right)d\xi ds

for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right), and then verify that qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is the fundamental solution to (2.5). To this end, for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right) and nn\in\mathbb{N}, we define

(3.2) qJ,0(z,w,t):=qJ(z,w,t) and qJ,n+1(z,w,t):=0t0JqJ(z,ξ,ts)qJ,n(ξ,w,s)V(ξ)𝑑ξ𝑑s.q_{J,0}\left(z,w,t\right):=q_{J}\left(z,w,t\right)\text{ and }q_{J,n+1}\left(z,w,t\right):=\int_{0}^{t}\int_{0}^{J}q_{J}\left(z,\xi,t-s\right)q_{J,n}\left(\xi,w,s\right)V\left(\xi\right)d\xi ds.

To state the technical results on {qJ,n(z,w,t):n0}\left\{q_{J,n}\left(z,w,t\right):n\geq 0\right\}, we need to introduce more notations. Set

(3.3) 𝔟:={ν if α(0,1) and b(0)0,1ν if α(0,1) and b(0)=0,1 if α[1,2).\mathfrak{b}:=\begin{cases}\nu&\text{ if }\alpha\in\left(0,1\right)\text{ and }b\left(0\right)\neq 0,\\ 1-\nu&\text{ if }\alpha\in\left(0,1\right)\text{ and }b\left(0\right)=0,\\ 1&\text{ if }\alpha\in[1,2).\end{cases}

We have that 0<𝔟10<\mathfrak{b}\leq 1, and if VJV_{J} is the constant found in Lemma 2.2, then (2.8) can be rewritten as

|V(z)|VJz𝔟1 for every z(0,J).\left|V\left(z\right)\right|\leq V_{J}\cdot z^{\mathfrak{b}-1}\text{ for every }z\in\left(0,J\right).

For nn\in\mathbb{N} and t>0t>0, we define

(3.4) mn(t):=Γn+1(𝔟)(𝔠t𝔟VJ)nΓ((n+1)𝔟) and M(t):=n=0mn(t).m_{n}\left(t\right):=\frac{\Gamma^{n+1}\left(\mathfrak{\mathfrak{b}}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{n}}{\Gamma\left(\left(n+1\right)\mathfrak{b}\right)}\text{ and }M\left(t\right):=\sum_{n=0}^{\infty}m_{n}\left(t\right).

It follows from a simple application of Stirling’s formula that mn(t)m_{n}\left(t\right) is summable in nn\in\mathbb{N}, and hence M(t)M\left(t\right) is well defined.

Lemma 3.1.

There exists a universal constant 𝔠1\mathfrak{c}\geq 1 such that for every nn\in\mathbb{N} and (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right),

(3.5) |qJ,n(z,w,t)|mn(t)q(z,w,t),\left|q_{J,n}\left(z,w,t\right)\right|\leq m_{n}\left(t\right)q\left(z,w,t\right),

and hence

(3.6) qJV(z,w,t):=n=0qJ,n(z,w,t)q_{J}^{V}\left(z,w,t\right):=\sum_{n=0}^{\infty}q_{J,n}\left(z,w,t\right)

is well defined as an absolutely convergent series. Moreover, for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right),

(3.7) |qJV(z,w,t)|M(t)q(z,w,t),\left|q_{J}^{V}\left(z,w,t\right)\right|\leq M\left(t\right)q\left(z,w,t\right),

and qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) satisfies (3.1).

Proof.

Without causing any substantial change, we will assume that V(z)V\left(z\right) is defined on (0,)\left(0,\infty\right) with V(z)0V\left(z\right)\equiv 0 for zJ.z\geq J. When 1α<21\leq\alpha<2, since V(z)V\left(z\right) is bounded on (0,)\left(0,\infty\right) with VJ=VuV_{J}=\left\|V\right\|_{u}, (3.5)-(3.7) can be derived in exactly the same way as in [8] (Lemma 3.4) with

mn(t)=(tVJ)nn! and M(t)=etVJ.m_{n}\left(t\right)=\frac{\left(tV_{J}\right)^{n}}{n!}\text{ and }M\left(t\right)=e^{tV_{J}}.

There is nothing we need to do in this case. Hence, we will assume α(0,1)\alpha\in\left(0,1\right) for the rest of the proof, and only treat the case when V(z)V\left(z\right) has a singularity at 0.

First, we claim that there exists a universal constant 𝔠>0\mathfrak{c}>0 such that

(3.8) 0q(z,ξ,t)q(ξ,w,s)ξ𝔟1𝑑ξ𝔠(t+sts)1𝔟q(z,w,t+s)\int_{0}^{\infty}q\left(z,\xi,t\right)q\left(\xi,w,s\right)\xi^{\mathfrak{b}-1}d\xi\leq\mathfrak{c}\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}q\left(z,w,t+s\right)

for every z,w(0,J)2z,w\in\left(0,J\right)^{2} and t,s>0t,s>0. To see this, we use (2.12) and (2.14) to write the integral in (3.8) as

z1ν(ts)2νeztws0e(t+s)ξtsξ𝔟ν(n=0(zξ)nn!Γ(n+2ν)t2n)(n=0(wξ)nn!Γ(n+2ν)s2n)𝑑ξ=z1ν(ts)2νeztws0e(t+s)ξtsξ𝔟νn=0ξ2nωn(z,w,t,s)dξ\begin{split}&\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\int_{0}^{\infty}e^{-\frac{(t+s)\xi}{ts}}\xi^{\mathfrak{b}-\nu}\left(\sum_{n=0}^{\infty}\frac{\left(z\xi\right)^{n}}{n!\Gamma\left(n+2-\nu\right)t^{2n}}\right)\left(\sum_{n=0}^{\infty}\frac{\left(w\xi\right)^{n}}{n!\Gamma\left(n+2-\nu\right)s^{2n}}\right)d\xi\\ =&\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\int_{0}^{\infty}e^{-\frac{(t+s)\xi}{ts}}\xi^{\mathfrak{b}-\nu}\sum_{n=0}^{\infty}\xi^{2n}\omega_{n}\left(z,w,t,s\right)d\xi\end{split}

where for every nn\in\mathbb{N},

ωn(z,w,t,s):=k=0nzkwnkk!(nk)!Γ(k+2ν)Γ(nk+2ν)t2ks2(nk).\omega_{n}\left(z,w,t,s\right):=\sum_{k=0}^{n}\frac{z^{k}w^{n-k}}{k!\left(n-k\right)!\Gamma\left(k+2-\nu\right)\Gamma\left(n-k+2-\nu\right)t^{2k}s^{2(n-k)}}.

Interchanging the order of summation and integration yields

z1ν(ts)2νeztwsn=0ωn(z,w,t,s)0e(t+s)ξtsξ2n+𝔟ν𝑑ξ=z1ν(ts)2νeztwsn=0ωn(z,w,t,s)(t+sts)ν𝔟12nΓ(2n+1+𝔟ν)=(t+sts)1𝔟z1ν(ts)2νeztwsn=0(t+sts)ν22nΓ(2n+1+𝔟ν)ωn(z,w,t,s).\begin{split}&\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\sum_{n=0}^{\infty}\omega_{n}\left(z,w,t,s\right)\int_{0}^{\infty}e^{-\frac{(t+s)\xi}{ts}}\xi^{2n+\mathfrak{b}-\nu}d\xi\\ =&\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\sum_{n=0}^{\infty}\omega_{n}\left(z,w,t,s\right)\left(\frac{t+s}{ts}\right)^{\nu-\mathfrak{b}-1-2n}\Gamma\left(2n+1+\mathfrak{b}-\nu\right)\\ =&\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\sum_{n=0}^{\infty}\left(\frac{t+s}{ts}\right)^{\nu-2-2n}\Gamma\left(2n+1+\mathfrak{b}-\nu\right)\omega_{n}\left(z,w,t,s\right).\end{split}

Since 0<𝔟10<\mathfrak{b}\leq 1, we have that for nn\in\mathbb{N},

Γ(2n+1+𝔟ν)Γ(2n+2ν)=B(2n+1+𝔟ν,1𝔟)Γ(1𝔟)1(1𝔟)Γ(1𝔟)=1Γ(2𝔟)𝔠,\frac{\Gamma\left(2n+1+\mathfrak{b}-\nu\right)}{\Gamma\left(2n+2-\nu\right)}=\frac{B\left(2n+1+\mathfrak{b}-\nu,1-\mathfrak{b}\right)}{\Gamma\left(1-\mathfrak{b}\right)}\leq\frac{1}{\left(1-\mathfrak{b}\right)\Gamma\left(1-\mathfrak{b}\right)}=\frac{1}{\Gamma\left(2-\mathfrak{b}\right)}\leq\mathfrak{c},

where B(u,v)B\left(u,v\right) (with u,v>0u,v>0) is the beta function and

(3.9) 𝔠:=1mins[1,2]Γ(s)1.12917.\mathfrak{c}:=\frac{1}{\min_{s\in\left[1,2\right]}\Gamma\left(s\right)}\thickapprox 1.12917.

Therefore, we have that

(t+sts)1𝔟z1ν(ts)2νeztwsn=0(t+sts)ν22nΓ(2n+1+𝔟ν)ωn(z,w,t,s)𝔠(t+sts)1𝔟z1ν(ts)2νeztwsn=0(t+sts)ν22nΓ(2n+2ν)ωn(z,w,t,s)=𝔠(t+sts)1𝔟z1ν(ts)2νeztws0e(t+s)ξtsn=0ξ2n+1νωn(z,w,t,s)dξ=𝔠(t+sts)1𝔟0q(z,ξ,t)q(ξ,w,s)𝑑ξ=𝔠(t+sts)1𝔟q(z,w,t+s),\begin{split}&\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\sum_{n=0}^{\infty}\left(\frac{t+s}{ts}\right)^{\nu-2-2n}\Gamma\left(2n+1+\mathfrak{b}-\nu\right)\omega_{n}\left(z,w,t,s\right)\\ \leq&\mathfrak{c}\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\sum_{n=0}^{\infty}\left(\frac{t+s}{ts}\right)^{\nu-2-2n}\Gamma\left(2n+2-\nu\right)\omega_{n}\left(z,w,t,s\right)\\ =&\mathfrak{c}\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}\frac{z^{1-\nu}}{\left(ts\right)^{2-\nu}}e^{-\frac{z}{t}-\frac{w}{s}}\int_{0}^{\infty}e^{-\frac{(t+s)\xi}{ts}}\sum_{n=0}^{\infty}\xi^{2n+1-\nu}\omega_{n}\left(z,w,t,s\right)d\xi\\ =&\mathfrak{c}\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}\int_{0}^{\infty}q\left(z,\xi,t\right)q\left(\xi,w,s\right)d\xi\\ =&\mathfrak{c}\left(\frac{t+s}{ts}\right)^{1-\mathfrak{b}}q\left(z,w,t+s\right),\end{split}

which confirms the claim (3.8).

To proceed, we notice that by Lemma 2.2, (2.17) and (3.2),

|qJ,1(z,w,t)|0t0qJ(z,ξ,ts)qJ(ξ,w,s)|V(ξ)|𝑑ξ𝑑sVJ0t0q(z,ξ,ts)q(ξ,w,s)ξ𝔟1𝑑ξ𝑑s𝔠VJ0tt1𝔟s1𝔟(ts)1𝔟𝑑sq(z,w,t)=𝔠t𝔟VJB(𝔟,𝔟)q(z,w,t)\begin{split}\left|q_{J,1}\left(z,w,t\right)\right|&\leq\int_{0}^{t}\int_{0}^{\infty}q_{J}\left(z,\xi,t-s\right)q_{J}\left(\xi,w,s\right)\left|V\left(\xi\right)\right|d\xi ds\\ &\leq V_{J}\int_{0}^{t}\int_{0}^{\infty}q\left(z,\xi,t-s\right)q\left(\xi,w,s\right)\xi^{\mathfrak{b}-1}d\xi ds\\ &\leq\mathfrak{c}V_{J}\int_{0}^{t}\frac{t^{1-\mathfrak{b}}}{s^{1-\mathfrak{b}}\left(t-s\right)^{1-\mathfrak{b}}}ds\cdot q\left(z,w,t\right)\\ &=\mathfrak{c}t^{\mathfrak{b}}V_{J}B\left(\mathfrak{b},\mathfrak{b}\right)q\left(z,w,t\right)\end{split}

for every (z,w,t)(0,)3\left(z,w,t\right)\in\left(0,\infty\right)^{3}. Assume that up to some n1n\geq 1, for every (z,w,t)(0,)3\left(z,w,t\right)\in\left(0,\infty\right)^{3},

|qJ,n(z,w,t)|(𝔠t𝔟VJ)n(j=1nB(𝔟,j𝔟))q(z,w,t).\left|q_{J,n}\left(z,w,t\right)\right|\leq\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{n}\left(\prod_{j=1}^{n}B\left(\mathfrak{b},j\mathfrak{b}\right)\right)q\left(z,w,t\right).

For n+1n+1, we have that

|qJ,n+1(z,w,t)|VJ0t0q(z,ξ,ts)|qJ,n(ξ,w,s)|ξ𝔟1𝑑ξ𝑑s𝔠nVJn+1(j=1nB(𝔟,j𝔟))0tsn𝔟0q(z,ξ,ts)q(z,w,t)ξ𝔟1𝑑ξ𝑑s(𝔠VJ)n+1(j=1nB(𝔟,j𝔟))0tsn𝔟t1𝔟s1𝔟(ts)1𝔟𝑑sq(z,w,t)=(𝔠t𝔟VJ)n+1(j=1n+1B(𝔟,j𝔟))q(z,w,t).\begin{split}\left|q_{J,n+1}\left(z,w,t\right)\right|&\leq V_{J}\int_{0}^{t}\int_{0}^{\infty}q\left(z,\xi,t-s\right)\left|q_{J,n}\left(\xi,w,s\right)\right|\xi^{\mathfrak{b}-1}d\xi ds\\ &\leq\mathfrak{c}^{n}V_{J}^{n+1}\left(\prod_{j=1}^{n}B\left(\mathfrak{b},j\mathfrak{b}\right)\right)\int_{0}^{t}s^{n\mathfrak{b}}\int_{0}^{\infty}q\left(z,\xi,t-s\right)q\left(z,w,t\right)\xi^{\mathfrak{b}-1}d\xi ds\\ &\leq\left(\mathfrak{c}V_{J}\right)^{n+1}\left(\prod_{j=1}^{n}B\left(\mathfrak{b},j\mathfrak{b}\right)\right)\int_{0}^{t}s^{n\mathfrak{b}}\frac{t^{1-\mathfrak{b}}}{s^{1-\mathfrak{b}}\left(t-s\right)^{1-\mathfrak{b}}}ds\cdot q\left(z,w,t\right)\\ &=\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{n+1}\left(\prod_{j=1}^{n+1}B\left(\mathfrak{b},j\mathfrak{b}\right)\right)q\left(z,w,t\right).\end{split}

Upon rewriting j=1nB(𝔟,j𝔟)\prod_{j=1}^{n}B\left(\mathfrak{b},j\mathfrak{b}\right) as (Γ(𝔟))n+1Γ((n+1)𝔟)\frac{\left(\Gamma\left(\mathfrak{b}\right)\right)^{n+1}}{\Gamma\left(\left(n+1\right)\mathfrak{b}\right)}, we immediately obtain (3.5)-(3.7). Finally, (3.1) can be verified by plugging the series representation of qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) 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 qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) be defined as in (3.6). Then, qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is continuous on (0,J)2×(0,)\left(0,J\right)^{2}\times\left(0,\infty\right), and for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right), we have that

(3.10) w1νqJV(z,w,t)=z1νqJV(w,z,t).w^{1-\nu}q_{J}^{V}\left(z,w,t\right)=z^{1-\nu}q_{J}^{V}\left(w,z,t\right).

qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) also satisfies the following integral equation:

(3.11) qJV(z,w,t)=qJ(z,w,t)+0t0qJV(z,ξ,ts)qJ(ξ,w,s)V(ξ)𝑑ξ𝑑s.q_{J}^{V}\left(z,w,t\right)=q_{J}\left(z,w,t\right)+\int_{0}^{t}\int_{0}^{\infty}q_{J}^{V}\left(z,\xi,t-s\right)q_{J}\left(\xi,w,s\right)V\left(\xi\right)d\xi ds.

Moreover, qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is the fundamental solution to (2.5). Given hCc((0,J))h\in C_{c}\left(\left(0,J\right)\right),

(3.12) vh,(0,J)V(z,t):=0JqJV(z,w,t)h(w)𝑑w for (z,t)(0,J)×(0,)v_{h,\left(0,J\right)}^{V}\left(z,t\right):=\int_{0}^{J}q_{J}^{V}\left(z,w,t\right)h\left(w\right)dw\text{ for }\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right)

is a smooth solution to (2.5).

Proof.

To prove (3.10), we first note that if, for (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right) and nn\in\mathbb{N}, we define

(3.13) q~J,0(z,w,t):=qJ(z,w,t) and q~J,n+1(z,w,t):=0t0Jq~J,n(z,ξ,ts)qJ(ξ,w,s)V(ξ)𝑑ξ𝑑s,\tilde{q}_{J,0}\left(z,w,t\right):=q_{J}\left(z,w,t\right)\text{ and }\tilde{q}_{J,n+1}\left(z,w,t\right):=\int_{0}^{t}\int_{0}^{J}\tilde{q}_{J,n}\left(z,\xi,t-s\right)q_{J}\left(\xi,w,s\right)V\left(\xi\right)d\xi ds,

then q~J,n(z,w,t)=qJ,n(z,w,t)\tilde{q}_{J,n}\left(z,w,t\right)=q_{J,n}\left(z,w,t\right). 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 2n2n-fold integrals, and confirm that the two integrals are identical. Next, we will show by induction that for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right) and nn\in\mathbb{N},

w1νqJ,n(z,w,t)=z1νqJ,n(w,z,t).w^{1-\nu}q_{J,n}\left(z,w,t\right)=z^{1-\nu}q_{J,n}\left(w,z,t\right).

When n=0n=0, this relation is simply (2.23). Assume that this relation holds up to some nn\in\mathbb{N}. By (2.23) and the equivalence between (3.2) and (3.13), we have that

w1νqJ,n+1(z,w,t)=z1ν0t0JqJ(ξ,z,ts)qJ,n(w,ξ,s)V(ξ)𝑑ξ𝑑s=z1ν0t0Jq~J,n(w,ξ,s)qJ(ξ,z,ts)V(ξ)𝑑ξ𝑑s=z1νq~J,n+1(w,z,t)=z1νqJ,n+1(w,z,t).\begin{split}w^{1-\nu}q_{J,n+1}\left(z,w,t\right)&=z^{1-\nu}\int_{0}^{t}\int_{0}^{J}q_{J}\left(\xi,z,t-s\right)q_{J,n}\left(w,\xi,s\right)V\left(\xi\right)d\xi ds\\ &=z^{1-\nu}\int_{0}^{t}\int_{0}^{J}\tilde{q}_{J,n}\left(w,\xi,s\right)q_{J}\left(\xi,z,t-s\right)V\left(\xi\right)d\xi ds\\ &=z^{1-\nu}\tilde{q}_{J,n+1}\left(w,z,t\right)=z^{1-\nu}q_{J,n+1}\left(w,z,t\right).\end{split}

(3.10) follows immediately. To establish (3.11), we write its right hand side as

qJ(z,w,t)+0t0qJV(z,ξ,ts)qJ(ξ,w,s)V(ξ)𝑑ξ𝑑s=qJ(z,w,t)+n=00t0q~J,n(z,ξ,ts)qJ(ξ,w,s)V(ξ)𝑑ξ𝑑s=qJ(z,w,t)+n=0q~J,n+1(z,w,t)=qJV(z,w,t),\begin{split}&q_{J}\left(z,w,t\right)+\int_{0}^{t}\int_{0}^{\infty}q_{J}^{V}\left(z,\xi,t-s\right)q_{J}\left(\xi,w,s\right)V\left(\xi\right)d\xi ds\\ =&q_{J}\left(z,w,t\right)+\sum_{n=0}^{\infty}\int_{0}^{t}\int_{0}^{\infty}\tilde{q}_{J,n}\left(z,\xi,t-s\right)q_{J}\left(\xi,w,s\right)V\left(\xi\right)d\xi ds\\ =&q_{J}\left(z,w,t\right)+\sum_{n=0}^{\infty}\tilde{q}_{J,n+1}\left(z,w,t\right)\\ =&q_{J}^{V}\left(z,w,t\right),\end{split}

where, again, we used the equivalence between (3.2) and (3.13). By (3.1) and (3.7), (z,t)qJV(z,w,t)\left(z,t\right)\mapsto q_{J}^{V}\left(z,w,t\right) is continuous for every w(0,J)w\in\left(0,J\right), and by (3.11), wqJV(z,w,t)w\mapsto q_{J}^{V}\left(z,w,t\right) is equicontinuous in (z,t)\left(z,t\right) from any bounded subset of (0,J)×(0,)\left(0,J\right)\times\left(0,\infty\right). From here one can easily derives the continuity of qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) in (z,w,t)\left(z,w,t\right) on (0,J)2×(0,)\left(0,J\right)^{2}\times\left(0,\infty\right).

Given hCc((0,J))h\in C_{c}\left(\left(0,J\right)\right), for every (z,t)(0,J)×(0,)\left(z,t\right)\in\left(0,J\right)\times\left(0,\infty\right), let vh,(0,J)V(z,t)v_{h,\left(0,J\right)}^{V}\left(z,t\right) and vh,(0,J)(z,t)v_{h,\left(0,J\right)}\left(z,t\right) be defined as in (3.12) and (2.27) respectively. It follows from (3.1) that

(3.14) vh,(0,J)V(z,t)=vh,(0,J)(z,t)+0t0JqJ(z,ξ,ts)vh,(0,J)V(ξ,s)V(ξ)𝑑ξ𝑑s.\begin{split}v_{h,\left(0,J\right)}^{V}\left(z,t\right)&=v_{h,\left(0,J\right)}\left(z,t\right)+\end{split}\int_{0}^{t}\int_{0}^{J}q_{J}\left(z,\xi,t-s\right)v_{h,\left(0,J\right)}^{V}\left(\xi,s\right)V\left(\xi\right)d\xi ds.

Let 𝔟\mathfrak{b} and 𝔠\mathfrak{c} be as in (3.3) and (3.9) respectively. By (3.7) and (3.8), we have that

|vh,(0,J)V(z,t)vh,(0,J)(z,t)|=|0t0JqJ(z,ξ,ts)vh,(0,J)V(ξ,s)V(ξ)𝑑ξ𝑑s|hu0J0t0JqJ(z,ξ,ts)|qJV(ξ,u,s)||V(ξ)|𝑑ξ𝑑s𝑑uhuM(t)0J0t0JqJ(z,ξ,ts)q(ξ,u,s)|V(ξ)|𝑑ξ𝑑s𝑑uhuM(t)𝔠t𝔟VJB(𝔟,𝔟)0Jq(z,u,t)𝑑u.\begin{split}\left|v_{h,\left(0,J\right)}^{V}\left(z,t\right)-v_{h,\left(0,J\right)}\left(z,t\right)\right|&=\left|\int_{0}^{t}\int_{0}^{J}q_{J}\left(z,\xi,t-s\right)v_{h,\left(0,J\right)}^{V}\left(\xi,s\right)V\left(\xi\right)d\xi ds\right|\\ &\leq\left\|h\right\|_{u}\int_{0}^{J}\int_{0}^{t}\int_{0}^{J}q_{J}\left(z,\xi,t-s\right)\left|q_{J}^{V}\left(\xi,u,s\right)\right|\left|V\left(\xi\right)\right|d\xi dsdu\\ &\leq\left\|h\right\|_{u}M\left(t\right)\int_{0}^{J}\int_{0}^{t}\int_{0}^{J}q_{J}\left(z,\xi,t-s\right)q\left(\xi,u,s\right)\left|V\left(\xi\right)\right|d\xi dsdu\\ &\leq\left\|h\right\|_{u}M\left(t\right)\mathfrak{c}t^{\mathfrak{b}}V_{J}B\left(\mathfrak{b},\mathfrak{b}\right)\int_{0}^{J}q\left(z,u,t\right)du.\end{split}

Since vh,(0,J)(z,t)v_{h,\left(0,J\right)}\left(z,t\right) is a solution to (2.9), the second last inequality implies that

limz0vh,(0,J)V(z,t)=limzJvh,(0,J)V(z,t)=0,\lim_{z\searrow 0}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=\lim_{z\nearrow J}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=0,

and the last inequality leads to limt0vh,(0,J)V(z,t)=h(z)\lim_{t\searrow 0}v_{h,\left(0,J\right)}^{V}\left(z,t\right)=h\left(z\right).

The only thing that remains to be proven is that vh,(0,J)V(z,t)v_{h,\left(0,J\right)}^{V}\left(z,t\right) is a smooth solution to the equation in (2.5), which, by the hypoellipticity of the operator tLV\partial_{t}-L^{V}, can be reduced to showing that vh,(0,J)V(z,t)v_{h,\left(0,J\right)}^{V}\left(z,t\right) is a solution in the sense of distribution. We take φCc((0,J))\varphi\in C_{c}^{\infty}\left(\left(0,J\right)\right) and consider

φ,vh,(0,J)V(,t):=0Jvh,(0,J)V(z,t)φ(z)𝑑z for t0,\left\langle\varphi,v_{h,\left(0,J\right)}^{V}\left(\cdot,t\right)\right\rangle:=\int_{0}^{J}v_{h,\left(0,J\right)}^{V}\left(z,t\right)\varphi\left(z\right)dz\text{ for }t\geq 0,

and use (3.14) to write it as

φ,vh,(0,J)V(,t)=φ,vh,(0,J)(,t)+0t0Jφ,qJ(,u,ts)vh,(0,J)V(u,s)V(u)𝑑u𝑑s.\left\langle\varphi,v_{h,\left(0,J\right)}^{V}\left(\cdot,t\right)\right\rangle=\left\langle\varphi,v_{h,\left(0,J\right)}\left(\cdot,t\right)\right\rangle+\int_{0}^{t}\int_{0}^{J}\left\langle\varphi,q_{J}\left(\cdot,u,t-s\right)\right\rangle v_{h,\left(0,J\right)}^{V}\left(u,s\right)V\left(u\right)duds.

Therefore,

ddtφ,vh,(0,J)V(,t)=ddtφ,vh,(0,J)(,t)+Vφ,vh,(0,J)V(,t)+0t0Jddtφ,qJ(,u,ts)vh,(0,J)V(u,s)V(u)𝑑u𝑑s=L0φ,vh,(0,J)(,t)+Vφ,vh,(0,J)V(,t)+0t0JL0φ,qJ(,u,ts)vh,(0,J)V(u,s)V(u)𝑑u𝑑s=(L0+V)φ,vh,(0,J)V(,t).\begin{split}\frac{d}{dt}\left\langle\varphi,v_{h,\left(0,J\right)}^{V}\left(\cdot,t\right)\right\rangle&=\frac{d}{dt}\left\langle\varphi,v_{h,\left(0,J\right)}\left(\cdot,t\right)\right\rangle+\left\langle V\varphi,v_{h,\left(0,J\right)}^{V}\left(\cdot,t\right)\right\rangle\\ &\qquad\qquad+\int_{0}^{t}\int_{0}^{J}\frac{d}{dt}\left\langle\varphi,q_{J}\left(\cdot,u,t-s\right)\right\rangle v_{h,\left(0,J\right)}^{V}\left(u,s\right)V\left(u\right)duds\\ &=\left\langle L_{0}^{*}\varphi,v_{h,\left(0,J\right)}\left(\cdot,t\right)\right\rangle+\left\langle V\varphi,v_{h,\left(0,J\right)}^{V}\left(\cdot,t\right)\right\rangle\\ &\qquad\qquad+\int_{0}^{t}\int_{0}^{J}\left\langle L_{0}^{*}\varphi,q_{J}\left(\cdot,u,t-s\right)\right\rangle v_{h,\left(0,J\right)}^{V}\left(u,s\right)V\left(u\right)duds\\ &=\left\langle\left(L_{0}^{*}+V\right)\varphi,v_{h,\left(0,J\right)}^{V}\left(\cdot,t\right)\right\rangle.\end{split}

3.2. Approximation of qJV(z,w,t)q_{J}^{V}\left(z,w,t\right)

In general we do not expect to find a closed-form formula for qJV(z,w,t)q_{J}^{V}\left(z,w,t\right), but when tt is sufficiently small, the above construction does provide accurate approximations for qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) whose exact formulas are explicit or even in closed forms. Intuitively speaking, when tt is small, the effect of the potential V(z)V\left(z\right) in LVL^{V} has not become “substantial” so that LVL^{V} is close to L0L_{0}, and hence it is natural to expect that qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is close to qJ(z,w,t)q_{J}\left(z,w,t\right) which, as we have seen in Corollary 2.7, is well approximated by q(z,w,t)q\left(z,w,t\right) for sufficiently small tt. To make it rigorous, we take tJt_{J} to be the same as in Corollary 2.7 and use (2.29) and (3.5) to derive that for every t(0,tJ)t\in\left(0,t_{J}\right),

supz,w(0,19J)2|qJV(z,w,t)q(z,w,t)1|\displaystyle\sup_{z,w\in\left(0,\frac{1}{9}J\right)^{2}}\left|\frac{q_{J}^{V}\left(z,w,t\right)}{q\left(z,w,t\right)}-1\right| supz,w(0,19J)2(|qJV(z,w,t)qJ(z,w,t)q(z,w,t)|+|qJ(z,w,t)q(z,w,t)1|)\displaystyle\leq\sup_{z,w\in\left(0,\frac{1}{9}J\right)^{2}}\left(\left|\frac{q_{J}^{V}\left(z,w,t\right)-q_{J}\left(z,w,t\right)}{q\left(z,w,t\right)}\right|+\left|\frac{q_{J}\left(z,w,t\right)}{q\left(z,w,t\right)}-1\right|\right)
M(t)1+exp(2J9t).\displaystyle\leq M\left(t\right)-1+\exp\left(-\frac{2J}{9t}\right).

Hence, for some constant C>0C>0 uniformly in t(0,tJ)t\in\left(0,t_{J}\right) (CC may depend on JJ and α\alpha),

(3.15) supz,w(0,19J)2|qJV(z,w,t)q(z,w,t)1|Ct𝔟.\sup_{z,w\in\left(0,\frac{1}{9}J\right)^{2}}\left|\frac{q_{J}^{V}\left(z,w,t\right)}{q\left(z,w,t\right)}-1\right|\leq Ct^{\mathfrak{b}}.

(3.15) confirms that when tt is small, qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) is indeed well approximated by q(z,w,t)q\left(z,w,t\right). However, viewing from (3.6), q(z,w,t)q\left(z,w,t\right) is only the “first order” approximation to qJV(z,w,t)q_{J}^{V}\left(z,w,t\right), since the error t𝔟t^{\mathfrak{b}} 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 “kk-th order” approximation for qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) with kk\in\mathbb{N}, and obtain an analog of (3.15) with the error being tk𝔟t^{k\mathfrak{b}}. To achieve this purpose, we introduce a new sequence of functions. For (z,w,t)(0,)3\left(z,w,t\right)\in\left(0,\infty\right)^{3} and nn\in\mathbb{N}, we set

(3.16) q0(z,w,t):=q(z,w,t) and qn(z,w,t):=0t0q(z,ξ,ts)V(ξ)qn(ξ,w,s)𝑑ξ𝑑s,q_{0}\left(z,w,t\right):=q\left(z,w,t\right)\text{ and }q_{n}\left(z,w,t\right):=\int_{0}^{t}\int_{0}^{\infty}q\left(z,\xi,t-s\right)V\left(\xi\right)q_{n}\left(\xi,w,s\right)d\xi ds,

where, again, we assume that V(z)0V\left(z\right)\equiv 0 for z>Jz>J. By following the proof of (3.5) line by line with qJ,n(z,w,t)q_{J,n}\left(z,w,t\right) replaced by qn(z,w,t)q_{n}\left(z,w,t\right), we also get that for every (z,w,t)(0,)3\left(z,w,t\right)\in\left(0,\infty\right)^{3} and nn\in\mathbb{N},

(3.17) |qn(z,w,t)|mn(t)q(z,w,t).\left|q_{n}\left(z,w,t\right)\right|\leq m_{n}\left(t\right)q\left(z,w,t\right).

Clearly, qn(z,w,t)q_{n}\left(z,w,t\right) is the “global” counterpart of qJ,n(z,w,t)q_{J,n}\left(z,w,t\right), and we will justify that qJ,n(z,w,t)q_{J,n}\left(z,w,t\right) is close to qn(z,w,t)q_{n}\left(z,w,t\right) when tt is sufficiently small.

Lemma 3.3.

For every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right) and nn\in\mathbb{N},

(3.18) |qJ,n(z,w,t)qn(z,w,t)|(2𝔠t𝔟B(𝔟,𝔟)VJ)nr(z,w,t),\left|q_{J,n}\left(z,w,t\right)-q_{n}\left(z,w,t\right)\right|\leq\left(2\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n}r\left(z,w,t\right),

where r(z,w,t)r\left(z,w,t\right) is as in (2.28).

Proof.

When n=0n=0, (3.18) simply becomes (2.28). Assume that (3.18) holds up to some n0n\geq 0. Following (3.2) and (3.16), we write

(3.19) qn+1(z,w,t)qJ,n+1(z,w,t)=0t0r(z,ξ,ts)V(ξ)qn(ξ,w,s)𝑑ξ𝑑s+0t0qJ(z,ξ,ts)V(ξ)(qn(ξ,w,s)qJ,n(ξ,w,s))𝑑ξ𝑑s.\begin{split}&q_{n+1}\left(z,w,t\right)-q_{J,n+1}\left(z,w,t\right)\\ =&\int_{0}^{t}\int_{0}^{\infty}r\left(z,\xi,t-s\right)V\left(\xi\right)q_{n}\left(\xi,w,s\right)d\xi ds\\ &\qquad\qquad+\int_{0}^{t}\int_{0}^{\infty}q_{J}\left(z,\xi,t-s\right)V\left(\xi\right)\left(q_{n}\left(\xi,w,s\right)-q_{J,n}\left(\xi,w,s\right)\right)d\xi ds.\end{split}

We use Fubini’s theorem and (2.28) to rewrite the first term on the right hand side of (3.19) as

(3.20) 𝔼[0tζJY(z)0q(J,ξ,tsζJY(z))V(ξ)qn(ξ,w,s)𝑑ξ𝑑s;ζJY(z)tζ0Y(z)],\mathbb{E}\left[\int_{0}^{t-\zeta_{J}^{Y}\left(z\right)}\int_{0}^{\infty}q\left(J,\xi,t-s-\zeta_{J}^{Y}\left(z\right)\right)V\left(\xi\right)q_{n}\left(\xi,w,s\right)d\xi ds;\zeta_{J}^{Y}\left(z\right)\leq t\wedge\zeta_{0}^{Y}\left(z\right)\right],

which, by (3.17), is bounded by

Γn+1(𝔟)(𝔠VJ)nΓ((n+1)𝔟)𝔼[0tζJY(z)0q(J,ξ,tsζJY(z))|V(ξ)|sn𝔟q(ξ,w,s)𝑑ξ𝑑s;ζJY(z)tζ0Y(z)].\frac{\Gamma^{n+1}\left(\mathfrak{b}\right)\left(\mathfrak{c}V_{J}\right)^{n}}{\Gamma\left(\left(n+1\right)\mathfrak{b}\right)}\mathbb{E}\left[\int_{0}^{t-\zeta_{J}^{Y}\left(z\right)}\int_{0}^{\infty}q\left(J,\xi,t-s-\zeta_{J}^{Y}\left(z\right)\right)\left|V\left(\xi\right)\right|s^{n\mathfrak{b}}q\left(\xi,w,s\right)d\xi ds;\zeta_{J}^{Y}\left(z\right)\leq t\wedge\zeta_{0}^{Y}\left(z\right)\right].

By (3.8) and the fact that

Γn+1(𝔟)Γ((n+1)𝔟)=j=1nB(𝔟,j𝔟)Bn(𝔟,𝔟),\frac{\Gamma^{n+1}\left(\mathfrak{b}\right)}{\Gamma\left(\left(n+1\right)\mathfrak{b}\right)}=\prod_{j=1}^{n}B\left(\mathfrak{b},j\mathfrak{b}\right)\leq B^{n}\left(\mathfrak{b},\mathfrak{b}\right),

we can further bound (3.20) from above by

𝔠nBn(𝔟,𝔟)VJn+1𝔼[0tζJY(z)sn𝔟0q(J,ξ,tsζJY(z))ξ𝔟1q(ξ,w,s)𝑑ξ𝑑s;ζJY(z)tζ0Y(z)]𝔠n+1Bn(𝔟,𝔟)VJn+1𝔼[q(J,w,tζJY(z))0tζJY(z)(tζJY(z))1𝔟sn𝔟(tsζJY(z))1𝔟s1𝔟𝑑s;ζJY(z)tζ0Y(z)](𝔠t𝔟B(𝔟,𝔟)VJ)n+1r(z,w,t).\begin{split}&\mathfrak{c}^{n}B^{n}\left(\mathfrak{b},\mathfrak{b}\right)V_{J}^{n+1}\mathbb{E}\left[\int_{0}^{t-\zeta_{J}^{Y}\left(z\right)}s^{n\mathfrak{b}}\int_{0}^{\infty}q\left(J,\xi,t-s-\zeta_{J}^{Y}\left(z\right)\right)\xi^{\mathfrak{b}-1}q\left(\xi,w,s\right)d\xi ds;\zeta_{J}^{Y}\left(z\right)\leq t\wedge\zeta_{0}^{Y}\left(z\right)\right]\\ \leq&\mathfrak{c}^{n+1}B^{n}\left(\mathfrak{b},\mathfrak{b}\right)V_{J}^{n+1}\mathbb{E}\left[q\left(J,w,t-\zeta_{J}^{Y}\left(z\right)\right)\int_{0}^{t-\zeta_{J}^{Y}\left(z\right)}\frac{\left(t-\zeta_{J}^{Y}\left(z\right)\right)^{1-\mathfrak{b}}s^{n\mathfrak{b}}}{\left(t-s-\zeta_{J}^{Y}\left(z\right)\right)^{1-\mathfrak{b}}s^{1-\mathfrak{b}}}ds;\zeta_{J}^{Y}\left(z\right)\leq t\wedge\zeta_{0}^{Y}\left(z\right)\right]\\ \leq&\left(\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n+1}r\left(z,w,t\right).\end{split}

According to the inductive assumption, the second term on the right hand side of (3.19) is bounded by

2n𝔠nBn(𝔟,𝔟)VJn+10t0q(z,ξ,ts)ξ𝔟1sn𝔟r(ξ,w,s)𝑑ξ𝑑s,2^{n}\mathfrak{c}^{n}B^{n}\left(\mathfrak{b},\mathfrak{b}\right)V_{J}^{n+1}\int_{0}^{t}\int_{0}^{\infty}q\left(z,\xi,t-s\right)\xi^{\mathfrak{b}-1}s^{n\mathfrak{b}}r\left(\xi,w,s\right)d\xi ds,

which, by (2.14) and (2.23), is equal to

2n𝔠nBn(𝔟,𝔟)VJn+1w1νzν10t0q(ξ,z,ts)ξ𝔟1sn𝔟r(w,ξ,s)𝑑ξ𝑑s.\frac{2^{n}\mathfrak{c}^{n}B^{n}\left(\mathfrak{b},\mathfrak{b}\right)V_{J}^{n+1}}{w^{1-\nu}z^{\nu-1}}\int_{0}^{t}\int_{0}^{\infty}q\left(\xi,z,t-s\right)\xi^{\mathfrak{b}-1}s^{n\mathfrak{b}}r\left(w,\xi,s\right)d\xi ds.

We use Fubini’s theorem again to rewrite the expression above as

2n𝔠nBn(𝔟,𝔟)VJn+1w1νzν1𝔼[ζJY(w)t0q(ξ,z,ts)ξ𝔟1sn𝔟q(J,ξ,sζJY(w))𝑑ξ𝑑s;ζJY(w)tζ0Y(w)]2n𝔠n+1Bn(𝔟,𝔟)VJn+1w1νzν1𝔼[q(J,z,tζJY(w))ζJY(w)t(tζJY(w))1𝔟sn𝔟(ts)1𝔟(sζJY(w))1𝔟𝑑s;ζJY(w)tζ0Y(w)]2n(𝔠t𝔟B(𝔟,𝔟)VJ)n+1w1νzν1r(w,z,t)=2n(𝔠t𝔟B(𝔟,𝔟)VJ)n+1r(z,w,t).\begin{split}&\frac{2^{n}\mathfrak{c}^{n}B^{n}\left(\mathfrak{b},\mathfrak{b}\right)V_{J}^{n+1}}{w^{1-\nu}z^{\nu-1}}\mathbb{E}\left[\int_{\zeta_{J}^{Y}\left(w\right)}^{t}\int_{0}^{\infty}q\left(\xi,z,t-s\right)\xi^{\mathfrak{b}-1}s^{n\mathfrak{b}}q\left(J,\xi,s-\zeta_{J}^{Y}\left(w\right)\right)d\xi ds;\zeta_{J}^{Y}\left(w\right)\leq t\wedge\zeta_{0}^{Y}\left(w\right)\right]\\ \leq&\frac{2^{n}\mathfrak{c}^{n+1}B^{n}\left(\mathfrak{b},\mathfrak{b}\right)V_{J}^{n+1}}{w^{1-\nu}z^{\nu-1}}\mathbb{E}\left[q\left(J,z,t-\zeta_{J}^{Y}\left(w\right)\right)\int_{\zeta_{J}^{Y}\left(w\right)}^{t}\frac{\left(t-\zeta_{J}^{Y}\left(w\right)\right)^{1-\mathfrak{b}}s^{n\mathfrak{b}}}{\left(t-s\right)^{1-\mathfrak{b}}\left(s-\zeta_{J}^{Y}\left(w\right)\right)^{1-\mathfrak{b}}}ds;\zeta_{J}^{Y}\left(w\right)\leq t\wedge\zeta_{0}^{Y}\left(w\right)\right]\\ \leq&\frac{2^{n}\left(\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n+1}}{w^{1-\nu}z^{\nu-1}}r\left(w,z,t\right)=2^{n}\left(\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n+1}r\left(z,w,t\right).\end{split}

Thus, combining the estimates of the two terms on the right hand side of (3.19), we obtain that

|qJ,n+1(z,w,t)qn+1(z,w,t)|(1+2n)(𝔠t𝔟B(𝔟,𝔟)VJ)n+1r(z,w,t)(2𝔠t𝔟B(𝔟,𝔟)VJ)n+1r(z,w,t).\begin{split}\left|q_{J,n+1}\left(z,w,t\right)-q_{n+1}\left(z,w,t\right)\right|&\leq\left(1+2^{n}\right)\left(\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n+1}r\left(z,w,t\right)\\ &\leq\left(2\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n+1}r\left(z,w,t\right).\end{split}

Proposition 3.4.

Let tJ:=4J9(2ν)t_{J}:=\frac{4J}{9\left(2-\nu\right)}. Then, for every t(0,tJ)t\in\left(0,t_{J}\right) and k\{0}k\in\mathbb{N}\backslash\left\{0\right\},

(3.21) supz,w(0,19J)2|qJV(z,w,t)n=0k1qn(z,w,t)q(z,w,t)|mk(t)M(t)+Dk(t)exp(2J9t),\sup_{z,w\in\left(0,\frac{1}{9}J\right)^{2}}\left|\frac{q_{J}^{V}\left(z,w,t\right)-\sum_{n=0}^{k-1}q_{n}\left(z,w,t\right)}{q\left(z,w,t\right)}\right|\leq m_{k}\left(t\right)M\left(t\right)+D_{k}\left(t\right)\exp\left(-\frac{2J}{9t}\right),

where

(3.22) Dk(t):=n=0k1(2𝔠t𝔟B(𝔟,𝔟)VJ)n.D_{k}\left(t\right):=\sum_{n=0}^{k-1}\left(2\mathfrak{c}t^{\mathfrak{b}}B\left(\mathfrak{b},\mathfrak{b}\right)V_{J}\right)^{n}.

In particular, there exists C>0C>0 uniformly in t(0,tJ)t\in\left(0,t_{J}\right) and k\{0}k\in\mathbb{N}\backslash\left\{0\right\} (CC may depend on JJ and ν\nu) such that

(3.23) supz,w(0,19J)2|qJV(z,w,t)n=0k1qn(z,w,t)q(z,w,t)|Ctk𝔟.\sup_{z,w\in\left(0,\frac{1}{9}J\right)^{2}}\left|\frac{q_{J}^{V}\left(z,w,t\right)-\sum_{n=0}^{k-1}q_{n}\left(z,w,t\right)}{q\left(z,w,t\right)}\right|\leq Ct^{k\mathfrak{b}}.
Proof.

Only (3.21) requires proof, since (3.23) follows from (3.21) trivially. By (3.5) and (3.6), we know that for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right) and k\{0}k\in\mathbb{N}\backslash\left\{0\right\}

|qJV(z,w,t)n=0k1qJ,n(z,w,t)q(z,w,t)|n=kmn(t)\left|\frac{q_{J}^{V}\left(z,w,t\right)-\sum_{n=0}^{k-1}q_{J,n}\left(z,w,t\right)}{q\left(z,w,t\right)}\right|\leq\sum_{n=k}^{\infty}m_{n}\left(t\right)

and we further derive that

n=kmn(t)=n=kΓn+1(𝔟)(𝔠t𝔟VJ)nΓ((n+1)𝔟)=Γk(𝔟)(𝔠t𝔟VJ)kl=0Γl+1(𝔟)(𝔠t𝔟VJ)lΓ((l+1)𝔟+k𝔟)=Γk(𝔟)(𝔠t𝔟VJ)kΓ(k𝔟)l=0Γl+1(𝔟)(𝔠t𝔟VJ)lΓ((l+1)𝔟)B((l+1)𝔟,k𝔟)Γk+1(𝔟)(𝔠t𝔟VJ)kΓ((k+1)𝔟)l=0Γl+1(𝔟)(𝔠t𝔟VJ)lΓ((l+1)𝔟)=mk(t)M(t),\begin{split}\sum_{n=k}^{\infty}m_{n}\left(t\right)&=\sum_{n=k}^{\infty}\frac{\Gamma^{n+1}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{n}}{\Gamma\left(\left(n+1\right)\mathfrak{b}\right)}\\ &=\Gamma^{k}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{k}\sum_{l=0}^{\infty}\frac{\Gamma^{l+1}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{l}}{\Gamma\left(\left(l+1\right)\mathfrak{b}+k\mathfrak{b}\right)}\\ &=\frac{\Gamma^{k}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{k}}{\Gamma\left(k\mathfrak{b}\right)}\sum_{l=0}^{\infty}\frac{\Gamma^{l+1}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{l}}{\Gamma\left(\left(l+1\right)\mathfrak{b}\right)}B\left(\left(l+1\right)\mathfrak{b},k\mathfrak{b}\right)\\ &\leq\frac{\Gamma^{k+1}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{k}}{\Gamma\left(\left(k+1\right)\mathfrak{b}\right)}\sum_{l=0}^{\infty}\frac{\Gamma^{l+1}\left(\mathfrak{b}\right)\left(\mathfrak{c}t^{\mathfrak{b}}V_{J}\right)^{l}}{\Gamma\left(\left(l+1\right)\mathfrak{b}\right)}\\ &=m_{k}\left(t\right)M\left(t\right),\end{split}

where we again used the fact that B((l+1)𝔟,k𝔟)B(𝔟,k𝔟)B\left(\left(l+1\right)\mathfrak{b},k\mathfrak{b}\right)\leq B\left(\mathfrak{b},k\mathfrak{b}\right) for every ll\in\mathbb{N}. Meanwhile, by (3.18), we have that for every (z,w,t)(0,J)2×(0,)\left(z,w,t\right)\in\left(0,J\right)^{2}\times\left(0,\infty\right),

|n=0k1qJ,n(z,w,t)n=0k1qn(z,w,t)|Dk(t)r(z,w,t),\left|\sum_{n=0}^{k-1}q_{J,n}\left(z,w,t\right)-\sum_{n=0}^{k-1}q_{n}\left(z,w,t\right)\right|\leq D_{k}\left(t\right)r\left(z,w,t\right),

which, combined with (2.29), leads to (3.21). ∎

3.3. From qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) to pI(z,w,t)p_{I}\left(z,w,t\right)

Now we are ready to return to the localized equation (2.1). Recall that I>0I>0, ϕ(x)\phi\left(x\right) and θ(x)\theta\left(x\right) are functions on (0,I)\left(0,I\right) defined by (1.2), and II and JJ are related by J=ϕ(I)J=\phi\left(I\right); with z(0,J)z\in\left(0,J\right), ψ(z)\psi\left(z\right) is the inverse function of ϕ\phi, θ~(z)=θ(ψ(z))\tilde{\theta}\left(z\right)=\theta\left(\psi\left(z\right)\right) and Θ(z)\Theta\left(z\right) is as defined in (2.2). Guided by Proposition 2.1, we define

(3.24) pI(x,y,t):=qJV(ϕ(x),ϕ(y),t)Θ(ϕ(x))Θ(ϕ(y))ϕ(y).p_{I}\left(x,y,t\right):=q_{J}^{V}\left(\phi\left(x\right),\phi\left(y\right),t\right)\frac{\Theta\left(\phi\left(x\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right).

for every (x,y,t)(0,I)2×(0,)\left(x,y,t\right)\in\left(0,I\right)^{2}\times\left(0,\infty\right). We immediately obtain several results on pI(x,y,t)p_{I}\left(x,y,t\right) based on Proposition 2.1 and Proposition 3.2. In addition, we can establish the connection between pI(x,y,t)p_{I}\left(x,y,t\right) and {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} the unique solution to (1.7) and underlying diffusion process corresponding to L=xαa(x)x2+b(x)xL=x^{\alpha}a\left(x\right)\partial_{x}^{2}+b\left(x\right)\partial_{x}.

Proposition 3.5.

Let pI(x,y,t)p_{I}\left(x,y,t\right) be defined as in (3.24). Then, pI(x,y,t)p_{I}\left(x,y,t\right) is continuous on (0,I)2×(0,)\left(0,I\right)^{2}\times\left(0,\infty\right) and

(3.25) (ϕ(y))1νϕ(y)Θ2(ϕ(y))pI(x,y,t)=(ϕ(x))1νϕ(x)Θ2(ϕ(x))pI(y,x,t)\frac{\left(\phi\left(y\right)\right)^{1-\nu}}{\phi^{\prime}\left(y\right)}\Theta^{2}\left(\phi\left(y\right)\right)p_{I}\left(x,y,t\right)=\frac{\left(\phi\left(x\right)\right)^{1-\nu}}{\phi^{\prime}\left(x\right)}\Theta^{2}\left(\phi\left(x\right)\right)p_{I}\left(y,x,t\right)

for every (x,y,t)(0,I)2×(0,)\left(x,y,t\right)\in\left(0,I\right)^{2}\times\left(0,\infty\right).

pI(x,y,t)p_{I}\left(x,y,t\right) is the fundamental solution to (2.1). Given fCc((0,I))f\in C_{c}\left(\left(0,I\right)\right),

(3.26) uf,(0,I)(x,t):=0If(y)pI(x,y,t)𝑑yu_{f,\left(0,I\right)}\left(x,t\right):=\int_{0}^{I}f\left(y\right)p_{I}\left(x,y,t\right)dy

is the unique solution in C2,1((0,I)×(0,))C^{2,1}\left(\left(0,I\right)\times\left(0,\infty\right)\right) to (2.1), and uf,(0,I)(x,t)u_{f,\left(0,I\right)}\left(x,t\right) is smooth on (0,I)×(0,)\left(0,I\right)\times\left(0,\infty\right). Moreover,

(3.27) uf,(0,I)(x,t)=𝔼[f(X(x,t));t<ζ0,IX(x)],u_{f,\left(0,I\right)}\left(x,t\right)=\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,I}^{X}\left(x\right)\right],

and hence for every Borel set Γ(0,I)\Gamma\subseteq\left(0,I\right),

(3.28) ΓpI(x,y,t)𝑑y=(X(x,t)Γ,t<ζ0,IX(x)).\int_{\Gamma}p_{I}\left(x,y,t\right)dy=\mathbb{P}\left(X\left(x,t\right)\in\Gamma,t<\zeta_{0,I}^{X}\left(x\right)\right).

Finally, pI(x,y,t)p_{I}\left(x,y,t\right) satisfies the Chapman-Kolmogorov equation, i.e., for every x,y(0,I)x,y\in\left(0,I\right) and t,s>0t,s>0,

(3.29) pI(x,y,t+s)=0IpI(x,ξ,t)pI(ξ,y,s)𝑑ξ.p_{I}\left(x,y,t+s\right)=\int_{0}^{I}p_{I}\left(x,\xi,t\right)p_{I}\left(\xi,y,s\right)d\xi.
Proof.

(3.25) follows directly from (3.10). Given fCc((0,I))f\in C_{c}\left(\left(0,I\right)\right), we set h(z):=fψ(z)Θ(z)h\left(z\right):=\frac{f\circ\psi\left(z\right)}{\Theta\left(z\right)} for z(0,J)z\in\left(0,J\right). By (3.12), it is straightforward to check that

uf,(0,I)(x,t)=Θ(ϕ(x))0If(y)qJV(ϕ(x),ϕ(y),t)ϕ(y)Θ(ϕ(y))𝑑y=Θ(ϕ(x))0Jf(ψ(w))qJV(ϕ(x),w,t)dwΘ(w)=Θ(ϕ(x))vh,(0,J)V(ϕ(x),t),\begin{split}u_{f,\left(0,I\right)}\left(x,t\right)&=\Theta\left(\phi\left(x\right)\right)\int_{0}^{I}f\left(y\right)q_{J}^{V}\left(\phi\left(x\right),\phi\left(y\right),t\right)\frac{\phi^{\prime}\left(y\right)}{\Theta\left(\phi\left(y\right)\right)}dy\\ &=\Theta\left(\phi\left(x\right)\right)\int_{0}^{J}f\left(\psi\left(w\right)\right)q_{J}^{V}\left(\phi\left(x\right),w,t\right)\frac{dw}{\Theta\left(w\right)}\\ &=\Theta\left(\phi\left(x\right)\right)v_{h,\left(0,J\right)}^{V}\left(\phi\left(x\right),t\right),\end{split}

and hence it follows from Proposition 3.2 that uf,(0,I)(x,t)u_{f,\left(0,I\right)}\left(x,t\right) is a smooth solution to (2.1). Since

{uf,(0,I)(X(x,sζ0,IX(x)),tsζ0,IX(x)):0st}\left\{u_{f,\left(0,I\right)}\left(X\left(x,s\wedge\zeta_{0,I}^{X}\left(x\right)\right),t-s\wedge\zeta_{0,I}^{X}\left(x\right)\right):0\leq s\leq t\right\}

is a bounded martingale, by equating its expectation at s=0s=0 and s=ts=t, we obtain (3.27), which further leads to (3.28). Since {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} is the unique solution to (1.7), uf,(0,I)(x,t)u_{f,\left(0,I\right)}\left(x,t\right) is the unique C2,1((0,I)×(0,))C^{2,1}\left(\left(0,I\right)\times\left(0,\infty\right)\right) solution to (2.1). Finally, (3.29) follows from (3.28) and the strong Markov property of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\}. ∎

Remark 3.6.

Note that the properties developed above for pI(x,y,t)p_{I}\left(x,y,t\right) and uf,(0,I)(x,t)u_{f,\left(0,I\right)}\left(x,t\right) also lead to corresponding results on qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) and vh,(0,J)V(z,t)v_{h,\left(0,J\right)}^{V}\left(z,t\right). For example, we see from (3.29) that qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) also satisfies the Chapman-Kolmogorov equation, i.e., for every z,w(0,J)z,w\in\left(0,J\right) and t,s>0t,s>0,

qJV(z,w,t+s)=0JqJV(z,ξ,t)qJV(ξ,w,s)𝑑ξ,q_{J}^{V}\left(z,w,t+s\right)=\int_{0}^{J}q_{J}^{V}\left(z,\xi,t\right)q_{J}^{V}\left(\xi,w,s\right)d\xi,

and the uniqueness of uf,(0,I)(z,t)u_{f,\left(0,I\right)}\left(z,t\right) implies that, given hCc((0,J))h\in C_{c}\left(\left(0,J\right)\right), vh,(0,J)V(z,t)v_{h,\left(0,J\right)}^{V}\left(z,t\right) is the unique C2,1((0,J)×(0,))C^{2,1}\left(\left(0,J\right)\times\left(0,\infty\right)\right) solution to (2.5).

The approximations we obtained in Proposition 3.4 for qJV(z,w,t)q_{J}^{V}\left(z,w,t\right) can also be “transported” to pI(x,y,t)p_{I}\left(x,y,t\right) in a straightforward way. To see this, we define, for (x,y,t)(0,I)2×(0,)\left(x,y,t\right)\in\left(0,I\right)^{2}\times\left(0,\infty\right),

(3.30) papprox.(x,y,t):=q(ϕ(x),ϕ(y),t)Θ(ϕ(x))Θ(ϕ(y))ϕ(y),p^{approx.}\left(x,y,t\right):=q\left(\phi\left(x\right),\phi\left(y\right),t\right)\frac{\Theta\left(\phi\left(x\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right),

and more generally for k\{0}k\in\mathbb{N}\backslash\left\{0\right\},

(3.31) pkapprox.(x,y,t):=n=0k1qn(ϕ(x),ϕ(y),t)Θ(ϕ(x))Θ(ϕ(y))ϕ(y).p^{k-approx.}\left(x,y,t\right):=\sum_{n=0}^{k-1}q_{n}\left(\phi\left(x\right),\phi\left(y\right),t\right)\frac{\Theta\left(\phi\left(x\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right).

Then Proposition 3.4 can be rewritten as follows.

Corollary 3.7.

There exists tI>0t_{I}>0 such that for every t(0,tI)t\in\left(0,t_{I}\right) and k\{0}k\in\mathbb{N}\backslash\left\{0\right\},

(3.32) sup(x,y)(0,ψ(19ϕ(I)))2|pI(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)|mk(t)M(t)+Dk(t)exp(2ϕ(I)9t)\sup_{\left(x,y\right)\in\left(0,\psi\left(\frac{1}{9}\phi\left(I\right)\right)\right)^{2}}\left|\frac{p_{I}\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|\leq m_{k}\left(t\right)M\left(t\right)+D_{k}\left(t\right)\exp\left(-\frac{2\phi\left(I\right)}{9t}\right)

where Dk(t)D_{k}\left(t\right) is as in (3.22). In particular,

sup(x,y)(0,ψ(19ϕ(I)))2|pI(x,y,t)papprox.(x,y,t)1|=M(t)1+Dk(t)exp(2ϕ(I)9t).\begin{split}\sup_{\left(x,y\right)\in\left(0,\psi\left(\frac{1}{9}\phi\left(I\right)\right)\right)^{2}}\left|\frac{p_{I}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}-1\right|&=M\left(t\right)-1+D_{k}\left(t\right)\exp\left(-\frac{2\phi\left(I\right)}{9t}\right).\end{split}
Remark 3.8.

We want to point out that, from now on, whenever J=ϕ(I)J=\phi\left(I\right), the constants ΘJ\Theta_{J}, VJV_{J} and tJt_{J} that were introduced in §3\mathsection 3 will also be written as ΘI\Theta_{I}, VIV_{I} and tIt_{I} respectively. In addition, by plugging ϕ(x)\phi\left(x\right) into (2.3) and (2.6), we get that

(3.33) V(ϕ(x))=θ2(x)4ϕ(x)θ(x)2ϕ(x)+1ν2θ(x)ϕ(x).V\left(\phi\left(x\right)\right)=-\frac{\theta^{2}\left(x\right)}{4\phi\left(x\right)}-\frac{\theta^{\prime}\left(x\right)}{2\phi^{\prime}\left(x\right)}+\frac{1-\nu}{2}\frac{\theta\left(x\right)}{\phi\left(x\right)}.

and

(3.34) Θ(ϕ(x))={xα4a14(x)2α2(2α)(ϕ(x))α4(2α)exp(0xb(w)2wαa(w)𝑑w) if α1,x14a14(x)212b0(ϕ(x))1412b0x12b0exp(0x12w(b(w)a(w)b(0))𝑑w) if α=1.\Theta\left(\phi\left(x\right)\right)=\begin{cases}\frac{x^{\frac{\alpha}{4}}a^{\frac{1}{4}}\left(x\right)}{2^{\frac{\alpha}{2\left(2-\alpha\right)}}\left(\phi\left(x\right)\right)^{\frac{\alpha}{4\left(2-\alpha\right)}}}\exp\left(-\int_{0}^{x}\frac{b\left(w\right)}{2w^{\alpha}a\left(w\right)}dw\right)&\text{ if }\alpha\neq 1,\\ \frac{x^{\frac{1}{4}}a^{\frac{1}{4}}\left(x\right)}{2^{\frac{1}{2}-b_{0}}\left(\phi\left(x\right)\right)^{\frac{1}{4}-\frac{1}{2}b_{0}}x^{\frac{1}{2}b_{0}}}\exp\left(-\int_{0}^{x}\frac{1}{2w}\left(\frac{b\left(w\right)}{a\left(w\right)}-b\left(0\right)\right)dw\right)&\text{ if }\alpha=1.\end{cases}

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 b(x)0b\left(x\right)\equiv 0, these expressions take much simpler forms than in the general case, as we will see with a concrete example in §5\mathsection 5.

4. Global Equation

In the previous section we have solved the localized equation (2.1) and obtained its fundamental solution pI(x,y,t)p_{I}\left(x,y,t\right). 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 {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} and the probabilistic interpretations of the solutions found in the previous sections.

4.1. From pI(x,y,t)p_{I}\left(x,y,t\right) to p(x,y,t)p\left(x,y,t\right)

We introduce two more notations for this section: given I>0I>0,

aI:=maxx[0,I]{1a(x),a(x)} and bI:=maxx[0,I]|b(x)|.a_{I}:=\max_{x\in\left[0,I\right]}\left\{\frac{1}{a\left(x\right)},a\left(x\right)\right\}\text{ and }b_{I}:=\max_{x\in\left[0,I\right]}\left|b\left(x\right)\right|.

Our first task is to derive probability estimates for the hitting times of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\}.

Lemma 4.1.

We define, for x0x\geq 0,

(4.1) S(x):=0ϕ(x)wνΘ2(w)𝑑w.S\left(x\right):=\int_{0}^{\phi\left(x\right)}w^{-\nu}\Theta^{2}\left(w\right)dw.

Then, for every 0<x<yI0<x<y\leq I,

(4.2) (ζyX(x)<ζ0X(x))=S(x)S(y);\mathbb{P}\left(\zeta_{y}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)=\frac{S\left(x\right)}{S\left(y\right)};

if ΘI\Theta_{I} is the constant found in Lemma 2.2 (upon identifying ΘJ\Theta_{J} with ΘI\Theta_{I} for J=ϕ(I)J=\phi\left(I\right)), then

(4.3) (ζyX(x)<ζ0X(x))ΘI4(ϕ(x)ϕ(y))1ν.\mathbb{P}\left(\zeta_{y}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)\leq\Theta_{I}^{4}\left(\frac{\phi\left(x\right)}{\phi\left(y\right)}\right)^{1-\nu}.

Moreover, for every G(0,I)G\in\left(0,I\right), x(0,G)x\in\left(0,G\right) and t>0t>0 such that IG>tbII-G>tb_{I}, we have that

(4.4) (ζIX(x)t)exp((IxtbI)24tIαaI).\mathbb{P}\left(\zeta_{I}^{X}\left(x\right)\leq t\right)\leq\exp\left(-\frac{\left(I-x-tb_{I}\right)^{2}}{4tI^{\alpha}a_{I}}\right).
Proof.

We use Itô’s formula to verify that, for every y>xy>x, {S(X(x,tζ0,yX(x)))}\left\{S\left(X\left(x,t\wedge\zeta_{0,y}^{X}\left(x\right)\right)\right)\right\} is a bounded martingale, and hence (4.2) follows immediately. Further, by (2.7), we have that for every x(0,I)x\in\left(0,I\right),

(4.5) ΘI2(ϕ(x))1ν1νS(x)ΘI2(ϕ(x))1ν1ν,\Theta_{I}^{-2}\frac{\left(\phi\left(x\right)\right)^{1-\nu}}{1-\nu}\leq S\left(x\right)\leq\Theta_{I}^{2}\frac{\left(\phi\left(x\right)\right)^{1-\nu}}{1-\nu},

which leads to (4.3).

Now we get down to proving (4.4), and the proof is similar to that of Lemma 2.5. For every λ0\lambda\geq 0 and t0t\geq 0,

{exp(λX(x,tζIX(x))λ0tζIX(x)b(X(x,s))𝑑sλ20tζIX(x)Xα(x,s)a(X(x,s))𝑑s):t0},\left\{\exp\left(\lambda X\left(x,t\wedge\zeta_{I}^{X}\left(x\right)\right)-\lambda\int_{0}^{t\wedge\zeta_{I}^{X}\left(x\right)}b\left(X\left(x,s\right)\right)ds-\lambda^{2}\int_{0}^{t\wedge\zeta_{I}^{X}\left(x\right)}X^{\alpha}\left(x,s\right)a\left(X\left(x,s\right)\right)ds\right):t\geq 0\right\},

is a bounded martingale, from where we get that

𝔼[exp(λ0ζIX(x)b(X(x,s))𝑑sλ20ζIX(x)Xα(x,s)a(X(x,s))𝑑s);ζIX(x)<]exp(λxλI).\mathbb{E}\left[\exp\left(-\lambda\int_{0}^{\zeta_{I}^{X}\left(x\right)}b\left(X\left(x,s\right)\right)ds-\lambda^{2}\int_{0}^{\zeta_{I}^{X}\left(x\right)}X^{\alpha}\left(x,s\right)a\left(X\left(x,s\right)\right)ds\right);\zeta_{I}^{X}\left(x\right)<\infty\right]\leq\exp\left(\lambda x-\lambda I\right).

Since

|λ0ζIX(x)b(X(x,s))𝑑s+λ20ζIX(x)Xα(x,s)a(X(x,s))𝑑s|(λbI+λ2IαaI)ζIX(x),\left|\lambda\int_{0}^{\zeta_{I}^{X}\left(x\right)}b\left(X\left(x,s\right)\right)ds+\lambda^{2}\int_{0}^{\zeta_{I}^{X}\left(x\right)}X^{\alpha}\left(x,s\right)a\left(X\left(x,s\right)\right)ds\right|\leq\left(\lambda b_{I}+\lambda^{2}I^{\alpha}a_{I}\right)\zeta_{I}^{X}\left(x\right),

we further have that

(4.6) 𝔼[exp((λbI+λ2IαaI)ζIX(x));ζIX(x)<]exp(λxλI).\mathbb{E}\left[\exp\left(-\left(\lambda b_{I}+\lambda^{2}I^{\alpha}a_{I}\right)\zeta_{I}^{X}\left(x\right)\right);\zeta_{I}^{X}\left(x\right)<\infty\right]\leq\exp\left(\lambda x-\lambda I\right).

By Markov’s inequality,

(ζIX(x)t)=(e(λbI+λ2IαaI)ζIX(x)e(λbI+λ2IαaI)t)eλ2tIαaIλ(IxtbI).\mathbb{P}\left(\zeta_{I}^{X}\left(x\right)\leq t\right)=\mathbb{P}\left(e^{-\left(\lambda b_{I}+\lambda^{2}I^{\alpha}a_{I}\right)\zeta_{I}^{X}\left(x\right)}\geq e^{-\left(\lambda b_{I}+\lambda^{2}I^{\alpha}a_{I}\right)t}\right)\leq e^{\lambda^{2}tI^{\alpha}a_{I}-\lambda\left(I-x-tb_{I}\right)}.

(4.4) is obtained by minimizing the right hand side above over λ0\lambda\geq 0. ∎

Next, we consider {pI(x,y,t):I>0}\left\{p_{I}\left(x,y,t\right):I>0\right\} as a family parametrized by II, and for every 0<I<H0<I<H, we want to find out the link between pI(x,y,t)p_{I}\left(x,y,t\right) and pH(x,y,t)p_{H}\left(x,y,t\right), i.e., the fundamental solutions to (2.1) with the right boundary at II and HH respectively. To this end, we choose a third constant G(0,I)G\in\left(0,I\right) and define for each x(0,G)x\in\left(0,G\right) a sequence of hitting times of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} where η0(x):=0\eta_{0}\left(x\right):=0 and for n\{0}n\in\mathbb{N}\backslash\left\{0\right\},

(4.7) η2n1(x):=inf{sη2n2(x):X(x,s)I},η2n(x):=inf{sη2n1(x):X(x,s)G}.\eta_{2n-1}\left(x\right):=\inf\left\{s\geq\eta_{2n-2}\left(x\right):X\left(x,s\right)\geq I\right\},\eta_{2n}\left(x\right):=\inf\left\{s\geq\eta_{2n-1}\left(x\right):X\left(x,s\right)\leq G\right\}.

In other words, the sequence {ηn(x):n}\left\{\eta_{n}\left(x\right):n\in\mathbb{N}\right\} records the downward crossings of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} from II to GG. With the help of {ηn(x):n}\left\{\eta_{n}\left(x\right):n\in\mathbb{N}\right\} and the strong Markov property of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\}, we are able to connect pH(x,y,t)p_{H}\left(x,y,t\right) and pI(x,y,t)p_{I}\left(x,y,t\right) as follows.

Proposition 4.2.

For (x,y,t)(0,G)2×(0,)\left(x,y,t\right)\in\left(0,G\right)^{2}\times\left(0,\infty\right),

(4.8) pH(x,y,t)=pI(x,y,t)+n=1𝔼[pI(G,y,tη2n(x));η2n(x)t,η2n(x)<ζ0,HX(x)].\begin{split}p_{H}\left(x,y,t\right)&=p_{I}\left(x,y,t\right)+\sum_{n=1}^{\infty}\mathbb{E}\left[p_{I}\left(G,y,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0,H}^{X}\left(x\right)\right]\end{split}.
Proof.

Given fCc((0,G))f\in C_{c}\left(\left(0,G\right)\right), we use (3.27) to write

0Gf(y)pH(x,y,t)𝑑y=𝔼[f(X(x,t));t<ζ0,HX(x)].\int_{0}^{G}f\left(y\right)p_{H}\left(x,y,t\right)dy=\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,H}^{X}\left(x\right)\right].

According to the number of downward crossings (from II to GG) completed by {X(x,s):0st}\left\{X\left(x,s\right):0\leq s\leq t\right\}, we further decompose 𝔼[f(X(x,t));t<ζ0,HX(x)]\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,H}^{X}\left(x\right)\right] as

𝔼[f(X(x,t));t<ζ0,IX(x)]+n=1𝔼[f(X(x,t));η2n(x)t<η2n+1(x)ζ0,HX(x),η2n(x)<ζ0,HX(x)].\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,I}^{X}\left(x\right)\right]+\sum_{n=1}^{\infty}\mathbb{E}\left[f\left(X\left(x,t\right)\right);\eta_{2n}\left(x\right)\leq t<\eta_{2n+1}\left(x\right)\wedge\zeta_{0,H}^{X}\left(x\right),\eta_{2n}\left(x\right)<\zeta_{0,H}^{X}\left(x\right)\right].

By the strong Markov property of X(x,t)X\left(x,t\right), we have that for each n1n\geq 1,

𝔼[f(X(x,t));η2n(x)t<η2n+1(x)ζ0,HX(x),η2n(x)<ζ0,HX(x)]=𝔼[0Gf(y)pI(G,y,tη2nX(x));η2n(x)t,η2n(x)<ζ0,HX(x)].\begin{array}[]{c}\mathbb{E}\left[f\left(X\left(x,t\right)\right);\eta_{2n}\left(x\right)\leq t<\eta_{2n+1}\left(x\right)\wedge\zeta_{0,H}^{X}\left(x\right),\eta_{2n}\left(x\right)<\zeta_{0,H}^{X}\left(x\right)\right]\\ \hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt=\mathbb{E}\left[\int_{0}^{G}f\left(y\right)p_{I}\left(G,y,t-\eta_{2n}^{X}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0,H}^{X}\left(x\right)\right].\end{array}

On one hand, by (2.13), (3.7) and (3.24),

pI(G,y,tη2n(x))\displaystyle p_{I}\left(G,y,t-\eta_{2n}\left(x\right)\right) M(t)(ϕ(G))1ν(tη2n(x))2νexp((ϕ(G)ϕ(y))2tη2n(x))Θ(ϕ(G))Θ(ϕ(y))ϕ(y)\displaystyle\leq M\left(t\right)\frac{\left(\phi\left(G\right)\right)^{1-\nu}}{\left(t-\eta_{2n}\left(x\right)\right)^{2-\nu}}\exp\left(-\frac{\left(\sqrt{\phi\left(G\right)}-\sqrt{\phi\left(y\right)}\right)^{2}}{t-\eta_{2n}\left(x\right)}\right)\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)
(4.9) M(t)(2νe)2ν(ϕ(G))1ν(ϕ(G)ϕ(y))2(ν2)Θ(ϕ(G))Θ(ϕ(y))ϕ(y).\displaystyle\leq M\left(t\right)\left(\frac{2-\nu}{e}\right)^{2-\nu}\left(\phi\left(G\right)\right)^{1-\nu}\left(\sqrt{\phi\left(G\right)}-\sqrt{\phi\left(y\right)}\right)^{2\left(\nu-2\right)}\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right).

On the other hand, if η2n(x)<ζ0,HX(x)\eta_{2n}\left(x\right)<\zeta_{0,H}^{X}\left(x\right), then it must be that (i) ζGX(x)<ζ0X(x)\zeta_{G}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right), (ii) during the time interval [ζGX(x),η1(x)]\left[\zeta_{G}^{X}\left(x\right),\eta_{1}\left(x\right)\right], the process starts from GG and hits II before 0, and (iii) for each j=0,,n1j=0,\cdots,n-1, during the time interval [η2j(x),η2j+1(x)]\left[\eta_{2j}\left(x\right),\eta_{2j+1}\left(x\right)\right], the process starts from GG and hits II before 0. Hence, by (4.2) and the strong Markov property of X(x,t)X\left(x,t\right), we have that

(4.10) (η2n(x)<ζ0,HX(x))(ζGX(x)<ζ0X(x))((ζIX(G)<ζ0X(G)))n=S(x)S(G)(S(G)S(I))n.\mathbb{P}\left(\eta_{2n}\left(x\right)<\zeta_{0,H}^{X}\left(x\right)\right)\leq\mathbb{P}\left(\zeta_{G}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)\left(\mathbb{P}\left(\zeta_{I}^{X}\left(G\right)<\zeta_{0}^{X}\left(G\right)\right)\right)^{n}=\frac{S\left(x\right)}{S\left(G\right)}\left(\frac{S\left(G\right)}{S\left(I\right)}\right)^{n}.

Combining the above, we obtain that for every (x,y,t)(0,G)2×(0,)\left(x,y,t\right)\in\left(0,G\right)^{2}\times\left(0,\infty\right),

n=1𝔼[pI(G,y,tη2n(x));η2n(x)t,η2n(x)<ζ0X(x)]M(t)(2νe)2ν(ϕ(G))1ν(ϕ(G)ϕ(y))2(ν2)Θ(ϕ(G))Θ(ϕ(y))ϕ(y)S(x)S(I)S(G).\begin{split}&\sum_{n=1}^{\infty}\mathbb{E}\left[p_{I}\left(G,y,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right]\\ \leq&M\left(t\right)\left(\frac{2-\nu}{e}\right)^{2-\nu}\left(\phi\left(G\right)\right)^{1-\nu}\left(\sqrt{\phi\left(G\right)}-\sqrt{\phi\left(y\right)}\right)^{2\left(\nu-2\right)}\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)\frac{S\left(x\right)}{S\left(I\right)-S\left(G\right)}.\end{split}

This guarantees that the series in the right hand of (4.8) is absolutely convergent. ∎

With Lemma 4.1 and Proposition 4.2, we are ready to prove our main result.

Theorem 4.3.

For every (x,y,t)(0,)3\left(x,y,t\right)\in\left(0,\infty\right)^{3}, we set

p(x,y,t):=limIpI(x,y,t).p\left(x,y,t\right):=\lim_{I\nearrow\infty}p_{I}\left(x,y,t\right).

Given 0<G<I<H0<G<I<H, let {ηn(x):n}\left\{\eta_{n}\left(x\right):n\in\mathbb{N}\right\} be the sequence of hitting times defined as in (4.7) (for the downward crossings of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} from II to GG). Then, for every (x,y,t)(0,G)2×(0,)\left(x,y,t\right)\in\left(0,G\right)^{2}\times\left(0,\infty\right),

(4.11) p(x,y,t)=pI(x,y,t)+n=1𝔼[pI(G,y,tη2n(x));η2n(x)t,η2n(x)<ζ0X(x)].\begin{split}p\left(x,y,t\right)&=p_{I}\left(x,y,t\right)+\sum_{n=1}^{\infty}\mathbb{E}\left[p_{I}\left(G,y,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right]\end{split}.

p(x,y,t)p\left(x,y,t\right) is continuous on (0,)3\left(0,\infty\right)^{3}, and for every (x,y,t)(0,)3\left(x,y,t\right)\in\left(0,\infty\right)^{3},

(4.12) (ϕ(y))1νϕ(y)Θ2(ϕ(y))p(x,y,t)=(ϕ(x))1νϕ(x)Θ2(ϕ(x))p(y,x,t).\frac{\left(\phi\left(y\right)\right)^{1-\nu}}{\phi^{\prime}\left(y\right)}\Theta^{2}\left(\phi\left(y\right)\right)p\left(x,y,t\right)=\frac{\left(\phi\left(x\right)\right)^{1-\nu}}{\phi^{\prime}\left(x\right)}\Theta^{2}\left(\phi\left(x\right)\right)p\left(y,x,t\right).

For every y>0y>0, (x,t)p(x,y,t)\left(x,t\right)\mapsto p\left(x,y,t\right) is a smooth solution to the Kolmogorov backward equation corresponding to LL, i.e.,

tp(x,y,t)=xαa(x)x2p(x,y,t)+b(x)xp(x,y,t);\partial_{t}p\left(x,y,t\right)=x^{\alpha}a\left(x\right)\partial_{x}^{2}p\left(x,y,t\right)+b\left(x\right)\partial_{x}p\left(x,y,t\right);

for every x>0x>0, (y,t)p(x,y,t)\left(y,t\right)\mapsto p\left(x,y,t\right) is a smooth solution to the Kolmogorov forward equation corresponding to LL, i.e.,

tp(x,y,t)=y2(yαa(y)p(x,y,t))y(b(y)p(x,y,t)).\partial_{t}p\left(x,y,t\right)=\partial_{y}^{2}\left(y^{\alpha}a\left(y\right)p\left(x,y,t\right)\right)-\partial_{y}\left(b\left(y\right)p\left(x,y,t\right)\right).

p(x,y,t)p\left(x,y,t\right) is the fundamental solution to (1.1). Given fCc((0,))f\in C_{c}\left(\left(0,\infty\right)\right),

uf(x,t):=0f(y)p(x,y,t)𝑑y for (x,t)(0,)2u_{f}\left(x,t\right):=\int_{0}^{\infty}f\left(y\right)p\left(x,y,t\right)dy\text{ for }\left(x,t\right)\in\left(0,\infty\right)^{2}

is the unique solution in C2,1((0,)2)C^{2,1}\left(\left(0,\infty\right)^{2}\right) to (1.1), and uf(x,t)u_{f}\left(x,t\right) is smooth on (0,)2\left(0,\infty\right)^{2}. Moreover, for every (x,t)(0,)2\left(x,t\right)\in\left(0,\infty\right)^{2},

uf(x,t)=𝔼[f(X(x,t));t<ζ0X(x)],u_{f}\left(x,t\right)=\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0}^{X}\left(x\right)\right],

and hence for every Borel set Γ(0,)\Gamma\subseteq\left(0,\infty\right),

Γp(x,y,t)𝑑y=(X(x,t)Γ,t<ζ0X(x)).\int_{\Gamma}p\left(x,y,t\right)dy=\mathbb{P}\left(X\left(x,t\right)\in\Gamma,t<\zeta_{0}^{X}\left(x\right)\right).

Finally, p(x,y,t)p\left(x,y,t\right) satisfies the Chapman-Kolmogorov equation, i.e., for every x,y>0x,y>0 and t,s>0t,s>0,

(4.13) p(x,y,t+s)=0p(x,ξ,t)p(ξ,y,s)𝑑ξ.p\left(x,y,t+s\right)=\int_{0}^{\infty}p\left(x,\xi,t\right)p\left(\xi,y,s\right)d\xi.
Proof.

It is clear from (4.8) that for every (x,y,t)(0,)3\left(x,y,t\right)\in\left(0,\infty\right)^{3}, by taking G>xyG>x\vee y, we know that the family I(G,)pI(x,y,t)I\in\left(G,\infty\right)\mapsto p_{I}\left(x,y,t\right) is non-decreasing, so p(x,y,t)p\left(x,y,t\right) as the limit of pI(x,y,t)p_{I}\left(x,y,t\right) (as II\nearrow\infty) is well defined. Since {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} is the unique solution to (1.7), ζ0,HX(x)ζ0X(x)\zeta_{0,H}^{X}\left(x\right)\rightarrow\zeta_{0}^{X}\left(x\right) almost surely as HH\nearrow\infty (see, e.g., §10\mathsection 10 of [35]). Thus, (4.11) follows from (4.8) by sending HH to infinity, and (4.12) follows from (3.25).

Now we examine the continuity of p(x,y,t)p\left(x,y,t\right). 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 (0,G)2×(0,)\left(0,G\right)^{2}\times\left(0,\infty\right), from where it is easy to see that for every x(0,G)x\in\left(0,G\right), (y,t)p(x,y,t)\left(y,t\right)\mapsto p\left(x,y,t\right) is continuous on (0,G)×(0,)\left(0,G\right)\times\left(0,\infty\right). Furthermore in the proof of Proposition 3.2 we have seen that xpI(G,x,s)x\mapsto p_{I}\left(G,x,s\right) is equicontinuous in ss from any bounded subset of (0,)\left(0,\infty\right), which, combined with (4.12), leads to the continuity of p(x,y,t)p\left(x,y,t\right) in all three variables.

Next, we turn our attention to uf(x,t)u_{f}\left(x,t\right) for fCc((0,))f\in C_{c}\left(\left(0,\infty\right)\right). It is clear that

uf(x,t)\displaystyle u_{f}\left(x,t\right) =limI0If(y)pI(x,y,t)𝑑y=limIuf,(0,I)(x,t),\displaystyle=\lim_{I\nearrow\infty}\int_{0}^{I}f\left(y\right)p_{I}\left(x,y,t\right)dy=\lim_{I\nearrow\infty}u_{f,\left(0,I\right)}\left(x,t\right),

and further by (3.27),

uf(x,t)=limI𝔼[f(X(x,t));t<ζ0,IX(x)]=𝔼[f(X(x,t));t<ζ0X(x)],u_{f}\left(x,t\right)=\lim_{I\nearrow\infty}\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,I}^{X}\left(x\right)\right]=\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0}^{X}\left(x\right)\right],

which means that p(x,y,t)p\left(x,y,t\right) is indeed the probability density function of X(x,t)X\left(x,t\right) provided that t<ζ0X(x)t<\zeta_{0}^{X}\left(x\right). (4.13) follows from the strong Markov property of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\}. Furthermore, by (4.11), if GG and II are sufficiently large such that x(0,G)x\in\left(0,G\right) and supp(f)(0,I)\text{supp}\left(f\right)\subseteq\left(0,I\right), then

uf(x,t)=uf,(0,I)(x,t)+n=1𝔼[uf,(0,I)(G,tη2n(x));η2n(x)t,η2n(x)<ζ0X(x)].u_{f}\left(x,t\right)=u_{f,\left(0,I\right)}\left(x,t\right)+\sum_{n=1}^{\infty}\mathbb{E}\left[u_{f,\left(0,I\right)}\left(G,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right].

Let us re-examine the event {η2n(x)t,η2n(x)<ζ0X(x)}\left\{\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right\} involved in the series above. If η2n(x)t\eta_{2n}\left(x\right)\leq t, then we must have that ζGX(x)<ζ0X(x)\zeta_{G}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right), η1(x)ζGX(x)t\eta_{1}\left(x\right)-\zeta_{G}^{X}\left(x\right)\leq t, and for each j=0,,n1j=0,\cdots,n-1, η2j+1(x)η2j(x)t\eta_{2j+1}\left(x\right)-\eta_{2j}\left(x\right)\leq t. Thus,

(η2n(x)t,η2n(x)<ζ0X(x))(ζGX(x)<ζ0X(x))((ζIX(G)<t))n.\mathbb{P}\left(\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)\leq\mathbb{P}\left(\zeta_{G}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)\left(\mathbb{P}\left(\zeta_{I}^{X}\left(G\right)<t\right)\right)^{n}.

By (4.2) and (4.4), we have that when IG>tbII-G>tb_{I},

(4.14) (η2n(x)t,η2n(x)<ζ0X(x))S(x)S(G)exp(n(IGtbI)24tIαaI).\mathbb{P}\left(\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)\leq\frac{S\left(x\right)}{S\left(G\right)}\exp\left(-\frac{n\left(I-G-tb_{I}\right)^{2}}{4tI^{\alpha}a_{I}}\right).

Therefore, when tt is sufficiently small,

|n=1𝔼[uf,(0,I)(G,tη2n(x));η2n(x)t,η2n(x)<ζ0X(x)]|fun=1(η2n(x)t,η2n(x)<ζ0X(x))fuS(x)S(G)exp((IGtbI)24tIαaI)4tIααI(IGtbI)2\begin{split}&\left|\sum_{n=1}^{\infty}\mathbb{E}\left[u_{f,\left(0,I\right)}\left(G,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right]\right|\\ \leq&\left\|f\right\|_{u}\sum_{n=1}^{\infty}\mathbb{P}\left(\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)\\ \leq&\left\|f\right\|_{u}\frac{S\left(x\right)}{S\left(G\right)}\exp\left(-\frac{\left(I-G-tb_{I}\right)^{2}}{4tI^{\alpha}a_{I}}\right)\frac{4tI^{\alpha}\alpha_{I}}{\left(I-G-tb_{I}\right)^{2}}\end{split}

which tends to 0 as t0t\searrow 0 or as x0x\searrow 0. Therefore, we have that

limx0uf(x,t)=limx0uf,(0,I)(x,t)=0 and limt0uf(x,t)=limt0uf,(0,I)(x,t)=f(x).\lim_{x\searrow 0}u_{f}\left(x,t\right)=\lim_{x\searrow 0}u_{f,\left(0,I\right)}\left(x,t\right)=0\text{ and }\lim_{t\searrow 0}u_{f}\left(x,t\right)=\lim_{t\searrow 0}u_{f,\left(0,I\right)}\left(x,t\right)=f\left(x\right).

The only remaining thing is to prove the statement on p(x,y,t)p\left(x,y,t\right) and uf(x,t)u_{f}\left(x,t\right) being smooth solutions to the concerned equations, which, again, by the hypoellipticity of tL\partial_{t}-L, is reduced to showing that they are distribution solutions. Take uf(x,t)u_{f}\left(x,t\right) for instance. We observe that for any φCc((0,))\varphi\in C_{c}^{\infty}\left(\left(0,\infty\right)\right),

φ,uf(,t)=limI0φ(x)uf,(0,I)(x,t)𝑑x=φ,f+limI0t0φ(x)Luf,(0,I)(x,s)𝑑x𝑑s=φ,f+limI0t0(Lφ)(x)uf,(0,I)(x,s)𝑑x𝑑s=φ,f+0t0(Lφ)(x)uf(x,t)𝑑x𝑑s,\begin{split}\left\langle\varphi,u_{f}\left(\cdot,t\right)\right\rangle&=\lim_{I\nearrow\infty}\int_{0}^{\infty}\varphi\left(x\right)u_{f,\left(0,I\right)}\left(x,t\right)dx\\ &=\left\langle\varphi,f\right\rangle+\lim_{I\nearrow\infty}\int_{0}^{t}\int_{0}^{\infty}\varphi\left(x\right)Lu_{f,\left(0,I\right)}\left(x,s\right)dxds\\ &=\left\langle\varphi,f\right\rangle+\lim_{I\nearrow\infty}\int_{0}^{t}\int_{0}^{\infty}\left(L^{*}\varphi\right)\left(x\right)u_{f,\left(0,I\right)}\left(x,s\right)dxds\\ &=\left\langle\varphi,f\right\rangle+\int_{0}^{t}\int_{0}^{\infty}\left(L^{*}\varphi\right)\left(x\right)u_{f}\left(x,t\right)dxds,\end{split}

which implies that

ddtφ,uf(,t)=Lφ,uf(,t).\frac{d}{dt}\left\langle\varphi,u_{f}\left(\cdot,t\right)\right\rangle=\left\langle L^{*}\varphi,u_{f}\left(\cdot,t\right)\right\rangle.

This confirms that uf(x,t)u_{f}\left(x,t\right) is a solution to (1.1) as a distribution. The statements on p(x,y,t)p\left(x,y,t\right) follow from similar arguments. ∎

Remark 4.4.

We want to point out that the function S(x)S\left(x\right) defined in (4.1) has a specific role in the boundary classification for diffusion process. In fact, S(x)S\left(x\right) is the scale function for the underlying diffusion process corresponding to LL, and as xx approaches a boundary, whether S(x)S\left(x\right) remains bounded or not is a factor in boundary classification (see §15.6\mathsection 15.6 of [26]). In particular, when viewing \infty as a boundary of (0,)\left(0,\infty\right), we introduce the escape probability at G>0G>0 (escaping from GG to \infty) as

(4.15) 𝔭G:=limI(ζIX(G)<ζ0X(G)).\mathfrak{p}_{G}:=\lim_{I\rightarrow\infty}\mathbb{P}\left(\zeta_{I}^{X}\left(G\right)<\zeta_{0}^{X}\left(G\right)\right).

Then, when limxS(x)=\lim_{x\rightarrow\infty}S\left(x\right)=\infty, \infty is non-attracting, in which case (4.2) implies that 𝔭G=0\mathfrak{p}_{G}=0; when limxS(x)<\lim_{x\rightarrow\infty}S\left(x\right)<\infty, \infty is attracting and 𝔭G>0\mathfrak{p}_{G}>0.

4.2. Approximation of p(x,y,t)p\left(x,y,t\right)

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 p(x,y,t)p\left(x,y,t\right) in the same spirit. In particular, we find explicitly defined approximations to p(x,y,t)p\left(x,y,t\right) such that (i) these approximations are more accurate than the standard heat kernel estimates, and (ii) when tt is sufficiently small, these approximations are “close” to p(x,y,t)p\left(x,y,t\right) uniformly in (x,y)\left(x,y\right) 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 a(x)a\left(x\right) and b(x)b\left(x\right).

Theorem 4.5.

Let papprox.(x,y,t)p^{approx.}\left(x,y,t\right) and pkapprox.(x,y,t)p^{k-approx.}\left(x,y,t\right), k\{0}k\in\mathbb{N}\backslash\left\{0\right\}, be defined as in (3.30) and (3.31) respectively. For any G>0G>0, set tG:=4ϕ(G)9(2ν)t_{G}:=\frac{4\phi\left(G\right)}{9\left(2-\nu\right)}. Then, for every t(0,tG)t\in\left(0,t_{G}\right), I>GI>G and k\{0}k\in\mathbb{N}\backslash\left\{0\right\},

(4.16) sup(x,y)(0,ψ(ϕ(G)9))2|p(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)|mk(t)M(t)+[Dk(t)+ΘG4M(t)(ϕ(G))1ν(1ν)(S(I)S(G))]exp(2ϕ(G)9t),\begin{split}&\sup_{\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2}}\left|\frac{p\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|\\ &\hskip 14.22636pt\hskip 14.22636pt\hskip 14.22636pt\leq m_{k}\left(t\right)M\left(t\right)+\left[D_{k}\left(t\right)+\frac{\Theta_{G}^{4}M\left(t\right)\left(\phi\left(G\right)\right)^{1-\nu}}{\left(1-\nu\right)\left(S\left(I\right)-S\left(G\right)\right)}\right]\exp\left(-\frac{2\phi\left(G\right)}{9t}\right),\end{split}

where mk(t)m_{k}\left(t\right), M(t)M\left(t\right) and Dk(t)D_{k}\left(t\right) are as in (3.4) and (3.22) respectively.

In particular, there exists constant C>0C>0 uniformly in t(0,tG)t\in\left(0,t_{G}\right) and k\{0}k\in\mathbb{N}\backslash\left\{0\right\} (CC may depend on GG and ν\nu) such that

sup(x,y)(0,ψ(ϕ(G)9))2|p(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)|Ctk𝔟,\sup_{\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2}}\left|\frac{p\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|\leq Ct^{k\mathfrak{b}},

where 𝔟\mathfrak{b} is the constant defined in (3.3).

Proof.

Only (4.16) requires proof. For every (x,y,t)(0,G)2×(0,)\left(x,y,t\right)\in\left(0,G\right)^{2}\times\left(0,\infty\right), we have that

|p(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)||p(x,y,t)pI(x,y,t)papprox.(x,y,t)|+|pI(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)|.\begin{split}\left|\frac{p\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|&\leq\left|\frac{p\left(x,y,t\right)-p_{I}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|+\left|\frac{p_{I}\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|.\end{split}

By (3.32), we have that for every t(0,tG)t\in\left(0,t_{G}\right), the second term on the right hand side above is bounded uniformly in (x,y)(0,ψ(ϕ(G)9))2\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2} by

mk(t)M(t)+Dk(t)exp(2ϕ(G)9t).m_{k}\left(t\right)M\left(t\right)+D_{k}\left(t\right)\exp\left(-\frac{2\phi\left(G\right)}{9t}\right).

We define hitting times {ηn(x):n}\left\{\eta_{n}\left(x\right):n\in\mathbb{N}\right\} as in (4.7) (for the downward crossings from II to GG). Then, according to (4.11),

p(x,y,t)pI(x,y,t)=n=1𝔼[pI(G,y,tη2n(x));η2n(x)t,η2n(x)<ζ0X(x)].p\left(x,y,t\right)-p_{I}\left(x,y,t\right)=\sum_{n=1}^{\infty}\mathbb{E}\left[p_{I}\left(G,y,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right].

It follows from (2.13), (3.7), (3.24) and (4.10) that for every t(0,tG)t\in\left(0,t_{G}\right) and (x,y)(0,ψ(ϕ(G)9))2\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2},

|p(x,y,t)pI(x,y,t)|M(t)(sups(0,t)sν2e4ϕ(G)9s)(ϕ(G))1νΘ(ϕ(G))Θ(ϕ(y))ϕ(y)S(x)S(I)S(G)M(t)tν2(ϕ(G))1νΘ(ϕ(G))Θ(ϕ(y))ϕ(y)exp(4ϕ(G)9t)S(x)S(I)S(G),\begin{split}&\left|p\left(x,y,t\right)-p_{I}\left(x,y,t\right)\right|\\ \leq&M\left(t\right)\left(\sup_{s\in\left(0,t\right)}s^{\nu-2}e^{-\frac{4\phi\left(G\right)}{9s}}\right)\left(\phi\left(G\right)\right)^{1-\nu}\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)\frac{S\left(x\right)}{S\left(I\right)-S\left(G\right)}\\ \leq&M\left(t\right)t^{\nu-2}\left(\phi\left(G\right)\right)^{1-\nu}\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)\exp\left(-\frac{4\phi\left(G\right)}{9t}\right)\frac{S\left(x\right)}{S\left(I\right)-S\left(G\right)},\end{split}

and further by (3.30) and (4.5) we have that

|p(x,y,t)pI(x,y,t)papprox.(x,y,t)|\displaystyle\left|\frac{p\left(x,y,t\right)-p_{I}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right| M(t)Θ(ϕ(G))Θ(ϕ(x))(ϕ(G)ϕ(x))1νexp(2ϕ(G)9t)S(x)S(I)S(G)\displaystyle\leq M\left(t\right)\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(x\right)\right)}\left(\frac{\phi\left(G\right)}{\phi\left(x\right)}\right)^{1-\nu}\exp\left(-\frac{2\phi\left(G\right)}{9t}\right)\frac{S\left(x\right)}{S\left(I\right)-S\left(G\right)}
ΘG4M(t)(ϕ(G))1ν(1ν)(S(I)S(G))exp(2ϕ(G)9t).\displaystyle\leq\frac{\Theta_{G}^{4}M\left(t\right)\left(\phi\left(G\right)\right)^{1-\nu}}{\left(1-\nu\right)\left(S\left(I\right)-S\left(G\right)\right)}\exp\left(-\frac{2\phi\left(G\right)}{9t}\right).

We close this section with two variations of (4.16). First, by (4.5), we note that

1S(I)S(G)=1S(G)S(G)/S(I)1S(G)/S(I)ΘG21ν(ϕ(G))1νS(G)/S(I)1S(G)/S(I).\frac{1}{S\left(I\right)-S\left(G\right)}=\frac{1}{S\left(G\right)}\frac{S\left(G\right)/S\left(I\right)}{1-S\left(G\right)/S\left(I\right)}\leq\Theta_{G}^{2}\frac{1-\nu}{\left(\phi\left(G\right)\right)^{1-\nu}}\frac{S\left(G\right)/S\left(I\right)}{1-S\left(G\right)/S\left(I\right)}.

Therefore, by sending II to \infty in (4.16), we get the following estimate.

Corollary 4.6.

For every G>0G>0, let tG>0t_{G}>0 be the same as in Theorem 4.5, and 𝔭G\mathfrak{p}_{G} be defined as in (4.15). Then, for every t(0,tG)t\in\left(0,t_{G}\right),

sup(x,y)(0,ψ(ϕ(G)9))2|p(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)|mk(t)M(t)+(Dk(t)+ΘG6M(t)𝔭G1𝔭G)exp(2ϕ(G)9t).\begin{split}&\sup_{\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2}}\left|\frac{p\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|\\ &\hskip 14.22636pt\hskip 14.22636pt\hskip 14.22636pt\hskip 14.22636pt\leq m_{k}\left(t\right)M\left(t\right)+\left(D_{k}\left(t\right)+\Theta_{G}^{6}M\left(t\right)\frac{\mathfrak{p}_{G}}{1-\mathfrak{p}_{G}}\right)\exp\left(-\frac{2\phi\left(G\right)}{9t}\right).\end{split}

Second, by making tGt_{G} in Theorem 4.5 smaller if necessary, we can derive an estimate analogous to (4.16) but independent of 𝔭G\mathfrak{p}_{G}. Intuitively speaking, when tt is sufficiently small, how well papprox.(x,y,t)p^{approx.}\left(x,y,t\right) approximates p(x,y,t)p\left(x,y,t\right) 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 tG>0t_{G}^{\prime}>0 such that

(4.17) IGtGbI>2tGIααI for every I>2G;I-G-t_{G}^{\prime}b_{I}>2\sqrt{t_{G}^{\prime}I^{\alpha}\alpha_{I}}\text{ for every }I>2G;

then, by using (4.14) instead of (4.10) in the proof of (4.16), we get that for every t(0,tG)t\in\left(0,t_{G}^{\prime}\right) and (x,y)(0,ψ(ϕ(G)9))2\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2}, |p(x,y,t)pI(x,y,t)|\left|p\left(x,y,t\right)-p_{I}\left(x,y,t\right)\right| is bounded from above by

M(t)tν2(ϕ(G))1νΘ(ϕ(G))Θ(ϕ(y))ϕ(y)exp(4ϕ(G)9t(IGtbI)24tIαaI)S(x)S(G)4tIααI(IGtbI)2M(t)tν2(ϕ(G))1νΘ(ϕ(G))Θ(ϕ(y))ϕ(y)S(x)S(G)exp(4ϕ(G)9t).\begin{split}&M\left(t\right)t^{\nu-2}\left(\phi\left(G\right)\right)^{1-\nu}\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)\exp\left(-\frac{4\phi\left(G\right)}{9t}-\frac{\left(I-G-tb_{I}\right)^{2}}{4tI^{\alpha}a_{I}}\right)\frac{S\left(x\right)}{S\left(G\right)}\frac{4tI^{\alpha}\alpha_{I}}{\left(I-G-tb_{I}\right)^{2}}\\ \leq&M\left(t\right)t^{\nu-2}\left(\phi\left(G\right)\right)^{1-\nu}\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)\frac{S\left(x\right)}{S\left(G\right)}\exp\left(-\frac{4\phi\left(G\right)}{9t}\right).\end{split}

It follows that for every (x,y,t)(0,ψ(ϕ(G)9))2×(0,tG)\left(x,y,t\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2}\times\left(0,t_{G}^{\prime}\right),

|p(x,y,t)pI(x,y,t)papprox.(x,y,t)|\displaystyle\left|\frac{p\left(x,y,t\right)-p_{I}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right| M(t)Θ(ϕ(G))Θ(ϕ(x))(ϕ(G)ϕ(x))1νS(x)S(G)exp(2ϕ(G)9t)\displaystyle\leq M\left(t\right)\frac{\Theta\left(\phi\left(G\right)\right)}{\Theta\left(\phi\left(x\right)\right)}\left(\frac{\phi\left(G\right)}{\phi\left(x\right)}\right)^{1-\nu}\frac{S\left(x\right)}{S\left(G\right)}\exp\left(-\frac{2\phi\left(G\right)}{9t}\right)
ΘG6M(t)exp(2ϕ(G)9t).\displaystyle\leq\Theta_{G}^{6}M\left(t\right)\exp\left(-\frac{2\phi\left(G\right)}{9t}\right).

Therefore, we have the following estimate on the error between pkapprox.(x,y,t)p^{k-approx.}\left(x,y,t\right) and p(x,y,t)p\left(x,y,t\right), which is a potential improvement of (4.16) for small tt.

Corollary 4.7.

For every G>0G>0, let tG>0t_{G}^{\prime}>0 be such that (4.17) holds. Then, for every t(0,tG)t\in\left(0,t_{G}^{\prime}\right),

sup(x,y)(0,ψ(ϕ(G)9))2|p(x,y,t)pkapprox.(x,y,t)papprox.(x,y,t)|mk(t)M(t)+(Dk(t)+ΘG6M(t))exp(2ϕ(G)9t).\begin{split}&\sup_{\left(x,y\right)\in\left(0,\psi\left(\frac{\phi\left(G\right)}{9}\right)\right)^{2}}\left|\frac{p\left(x,y,t\right)-p^{k-approx.}\left(x,y,t\right)}{p^{approx.}\left(x,y,t\right)}\right|\\ &\hskip 14.22636pt\hskip 14.22636pt\hskip 14.22636pt\hskip 14.22636pt\leq m_{k}\left(t\right)M\left(t\right)+\left(D_{k}\left(t\right)+\Theta_{G}^{6}M\left(t\right)\right)\exp\left(-\frac{2\phi\left(G\right)}{9t}\right).\end{split}

5. Generalized Wright-Fisher Diffusion

As reviewed in §1.1\mathsection 1.1, the classical Wright-Fisher diffusion equation given by (1.6) has two degenerate boundaries at 0 and 11, 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 α,β(0,2)\alpha,\beta\in\left(0,2\right), we consider the following Cauchy problem with two-sided boundaries on (0,1)\left(0,1\right), where, given fCb((0,1))f\in C_{b}\left(\left(0,1\right)\right), we look for uf(x,t)C2,1((0,1)×(0,))u_{f}\left(x,t\right)\in C^{2,1}\left(\left(0,1\right)\times\left(0,\infty\right)\right) such that

(5.1) tuf(x,t)=xα(1x)βx2uf(x,t) for every (x,t)(0,1)×(0,),limt0uf(x,t)=f(x) for every x(0,1),limx0uf(x,t)=limx1uf(x,t)=0 for every t(0,).\begin{array}[]{c}\partial_{t}u_{f}\left(x,t\right)=x^{\alpha}\left(1-x\right)^{\beta}\partial_{x}^{2}u_{f}\left(x,t\right)\text{ for every }\left(x,t\right)\in\left(0,1\right)\times\left(0,\infty\right),\\ \lim_{t\searrow 0}u_{f}\left(x,t\right)=f\left(x\right)\text{ for every }x\in\left(0,1\right),\\ \lim_{x\searrow 0}u_{f}\left(x,t\right)=\lim_{x\nearrow 1}u_{f}\left(x,t\right)=0\text{ for every }t\in\left(0,\infty\right).\end{array}

Set Lα,β:=xα(1x)βx2L_{\alpha,\beta}:=x^{\alpha}\left(1-x\right)^{\beta}\partial_{x}^{2}. We want to apply the method developed in the previous sections to construct and study the fundamental solution p(x,y,t)p\left(x,y,t\right) to (5.1). Lα,βL_{\alpha,\beta} 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 11, Lα,βL_{\alpha,\beta} has the advantage that its coefficient xα(1x)βx^{\alpha}\left(1-x\right)^{\beta} is bounded on (0,1)\left(0,1\right). Therefore, for every x(0,1)x\in\left(0,1\right), the stochastic differential equation

(5.2) dX(x,t)=2Xα(x,t)(1X(x,t))βdB(t) with X(x,0)xdX\left(x,t\right)=\sqrt{2X^{\alpha}\left(x,t\right)\left(1-X\left(x,t\right)\right)^{\beta}}dB\left(t\right)\text{ with }X\left(x,0\right)\equiv x

always has a solution in the sense described in §1.3\mathsection 1.3 (see, e.g., of [32]). Although we are not yet ready to claim the uniqueness of this solution, we can follow the theory in §12\mathsection 12 of [32] to extract a solution to (5.2) that has strong Markov property. In other words, (5.2) always has a solution {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} that is a strong Markov process.

The existence of a strong Markovian solution to (5.2) enables us to follow the steps in §2§4\mathsection 2-\mathsection 4 to tackle (5.1). In particular, with the localization procedure, we have the option of placing our “focal point” in the neighborhood of either 0 or 11 while constructing p(x,y,t)p\left(x,y,t\right). We will see that these two views are consistent and will lead to the same p(x,y,t)p\left(x,y,t\right).

Let us start with the construction of p(x,y,t)p\left(x,y,t\right) with a focus only on the left boundary 0, and we will follow the steps in the previous sections with a(x)=(1x)βa\left(x\right)=\left(1-x\right)^{\beta} and b(x)0b\left(x\right)\equiv 0. 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 0 is “effective” in this construction.

We take I(0,1)I\in\left(0,1\right) and localize (5.1) onto (0,I)\left(0,I\right). All the functions involved in the transformation are as follows:

ϕ(L)(x)=14(b(L)(x))2 and θ(L)(x)=α2(2α)ααxβx4xα22(1x)β22b(L)(x),\phi^{(L)}\left(x\right)=\frac{1}{4}\left(b^{(L)}\left(x\right)\right)^{2}\text{ and }\theta^{(L)}\left(x\right)=\frac{\alpha}{2\left(2-\alpha\right)}-\frac{\alpha-\alpha x-\beta x}{4}x^{\frac{\alpha-2}{2}}\left(1-x\right)^{\frac{\beta-2}{2}}b^{(L)}\left(x\right),

where b(L)(x):=0xsα/2(1s)β/2𝑑sb^{(L)}\left(x\right):=\int_{0}^{x}s^{-\alpha/2}\left(1-s\right)^{-\beta/2}ds is the incomplete beta function; furthermore,

Θ(ϕ(L)(x))=xα4(1x)β4(b(L)(x))α2(2α) with ΘI(L)=(22α)α2(2α)(1I)β2(2α);\Theta\left(\phi^{(L)}\left(x\right)\right)=\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{\left(b^{(L)}\left(x\right)\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}}\text{ with }\Theta_{I}^{(L)}=\left(\frac{2}{2-\alpha}\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}\left(1-I\right)^{-\frac{\beta}{2\left(2-\alpha\right)}};

in addition, for every x(0,I)x\in\left(0,I\right),

V(ϕ(L)(x))=α(α4)4(2α)2(b(L)(x))2+xα2(1x)β2((ααxβx)216α(1x)2+βx24),V\left(\phi^{(L)}\left(x\right)\right)=-\frac{\alpha\left(\alpha-4\right)}{4\left(2-\alpha\right)^{2}\left(b^{(L)}\left(x\right)\right)^{2}}+x^{\alpha-2}\left(1-x\right)^{\beta-2}\left(\frac{\left(\alpha-\alpha x-\beta x\right)^{2}}{16}-\frac{\alpha\left(1-x\right)^{2}+\beta x^{2}}{4}\right),

and hence

|V(ϕ(L)(x))|VI(L)(ϕ(L)(x))1α2α with VI(L)=β16(4β+2α)(1I)β2α2.\left|V\left(\phi^{(L)}\left(x\right)\right)\right|\leq V_{I}^{(L)}\left(\phi^{(L)}\left(x\right)\right)^{-\frac{1-\alpha}{2-\alpha}}\text{ with }V_{I}^{(L)}=\frac{\beta}{16}\left(4-\beta+2\alpha\right)\left(1-I\right)^{\frac{\beta}{2-\alpha}-2}.

This confirms that the statement in Lemma 2.2 still holds in this case.

Next, for the model equation discussed in §2.2\mathsection 2.2, we plug in ν(L):=1α2α\nu^{(L)}:=\frac{1-\alpha}{2-\alpha} and obtain q(L)(z,w,t)q^{\left(L\right)}\left(z,w,t\right) as in (2.12) and qϕ(L)(I)(L)(z,w,t)q_{\phi^{(L)}\left(I\right)}^{(L)}\left(z,w,t\right) as in (2.18) accordingly. We then follow exactly the same steps as in §3.1\mathsection 3.1 to derive qϕ(L)(I)(L),V(z,w,t)q_{\phi^{(L)}\left(I\right)}^{(L),V}\left(z,w,t\right) based on qϕ(L)(I)(L)(z,w,t)q_{\phi^{(L)}\left(I\right)}^{(L)}\left(z,w,t\right), and to obtain pI(L)(x,y,t)p_{I}^{(L)}\left(x,y,t\right) through reversing the transformation z=ϕ(L)(x)z=\phi^{(L)}\left(x\right), i.e.,

pI(L)(x,y,t)=qϕ(L)(I)(L),V(ϕ(L)(x),ϕ(L)(y),t)xα4(1x)β42y3α4(1y)3β4(b(L)(y))4α2(2α)(b(L)(x))α2(2α).p_{I}^{(L)}\left(x,y,t\right)=q_{\phi^{(L)}\left(I\right)}^{(L),V}\left(\phi^{(L)}\left(x\right),\phi^{(L)}\left(y\right),t\right)\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{2y^{\frac{3\alpha}{4}}\left(1-y\right)^{\frac{3\beta}{4}}}\frac{\left(b^{(L)}\left(y\right)\right)^{\frac{4-\alpha}{2\left(2-\alpha\right)}}}{\left(b^{(L)}\left(x\right)\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}}.

To proceed, we follow the arguments in §4\mathsection 4 to obtain the fundamental solution to (5.1) as

p(x,y,t)=limI1pI(L)(x,y,t) for (x,y,t)(0,1)2×(0,).p\left(x,y,t\right)=\lim_{I\nearrow 1}p_{I}^{(L)}\left(x,y,t\right)\text{ for }\left(x,y,t\right)\in\left(0,1\right)^{2}\times\left(0,\infty\right).

By (5.2), {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} itself is a martingale, and as in Lemma 4.1, we can derive probability estimates for the hitting times of X(x,t)X\left(x,t\right) as

(5.3) (ζyX(x)<ζ0X(x))=xy and (ζIX(x)t)exp((Ix)24Mα,βt)\mathbb{P}\left(\zeta_{y}^{X}\left(x\right)<\zeta_{0}^{X}\left(x\right)\right)=\frac{x}{y}\text{ and }\mathbb{P}\left(\zeta_{I}^{X}\left(x\right)\leq t\right)\leq\exp\left(-\frac{\left(I-x\right)^{2}}{4M_{\alpha,\beta}t}\right)

for every 0<x<y<I0<x<y<I and t>0t>0, where

Mα,β:=maxx[0,1]xα(1x)β=ααββ(α+β)α+β.M_{\alpha,\beta}:=\max_{x\in\left[0,1\right]}x^{\alpha}\left(1-x\right)^{\beta}=\frac{\alpha^{\alpha}\beta^{\beta}}{\left(\alpha+\beta\right)^{\alpha+\beta}}.

For every 0<G<I<H<10<G<I<H<1, if {ηn(x):n}\left\{\eta_{n}\left(x\right):n\in\mathbb{N}\right\} is the sequence of hitting times as in (4.7) (for the downward crossings of X(x,t)X\left(x,t\right) from II to GG), then for every (x,y,t)(0,G)2×(0,)\left(x,y,t\right)\in\left(0,G\right)^{2}\times\left(0,\infty\right),

(5.4) p(x,y,t)=pI(L)(x,y,t)+n=1𝔼[pI(L)(G,y,tη2n(x));η2n(x)t,η2n(x)<ζ0,1X(x)],p\left(x,y,t\right)=p_{I}^{(L)}\left(x,y,t\right)+\sum_{n=1}^{\infty}\mathbb{E}\left[p_{I}^{(L)}\left(G,y,t-\eta_{2n}\left(x\right)\right);\eta_{2n}\left(x\right)\leq t,\eta_{2n}\left(x\right)<\zeta_{0,1}^{X}\left(x\right)\right],

where the series on the right hand side is absolutely convergent.

Let us rewrite the results in Theorem 4.3 for p(x,y,t)p\left(x,y,t\right) found above.

Proposition 5.1.

p(x,y,t)p\left(x,y,t\right) is smooth on (0,1)2×(0,)\left(0,1\right)^{2}\times\left(0,\infty\right), and for every (x,y,t)(0,1)2×(0,)\left(x,y,t\right)\in\left(0,1\right)^{2}\times\left(0,\infty\right),

(5.5) yα(1y)βp(x,y,t)=xα(1x)βp(y,x,t).y^{\alpha}\left(1-y\right)^{\beta}p\left(x,y,t\right)=x^{\alpha}\left(1-x\right)^{\beta}p\left(y,x,t\right).

For every y(0,1)y\in\left(0,1\right), (x,t)p(x,y,t)\left(x,t\right)\mapsto p\left(x,y,t\right) is a smooth solution to the Kolmogorov backward equation corresponding to Lα,βL_{\alpha,\beta}, i.e.,

tp(x,y,t)=xα(1x)βx2p(x,y,t);\partial_{t}p\left(x,y,t\right)=x^{\alpha}\left(1-x\right)^{\beta}\partial_{x}^{2}p\left(x,y,t\right);

for every x(0,1)x\in\left(0,1\right), (y,t)p(x,y,t)\left(y,t\right)\mapsto p\left(x,y,t\right) is a smooth solution to the Kolmogorov forward equation corresponding to Lα,βL_{\alpha,\beta}, i.e.,

tp(x,y,t)=y2(yα(1y)βp(x,y,t)).\partial_{t}p\left(x,y,t\right)=\partial_{y}^{2}\left(y^{\alpha}\left(1-y\right)^{\beta}p\left(x,y,t\right)\right).

p(x,y,t)p\left(x,y,t\right) is the fundamental solution to (5.1). Given fCc((0,1))f\in C_{c}\left(\left(0,1\right)\right),

(5.6) uf(x,t):=0f(y)p(x,y,t)𝑑y for (x,t)(0,1)×(0,)u_{f}\left(x,t\right):=\int_{0}^{\infty}f\left(y\right)p\left(x,y,t\right)dy\text{ for }\left(x,t\right)\in\left(0,1\right)\times\left(0,\infty\right)

is a smooth solution to (1.1). Moreover, for every (x,t)(0,1)×(0,)\left(x,t\right)\in\left(0,1\right)\times\left(0,\infty\right),

(5.7) uf(x,t)=𝔼[f(X(x,t));t<ζ0,1X(x)],u_{f}\left(x,t\right)=\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,1}^{X}\left(x\right)\right],

and hence for every Borel set Γ(0,1)\Gamma\subseteq\left(0,1\right),

(5.8) Γp(x,y,t)𝑑y=(X(x,t)Γ,t<ζ0,1X(x))\int_{\Gamma}p\left(x,y,t\right)dy=\mathbb{P}\left(X\left(x,t\right)\in\Gamma,t<\zeta_{0,1}^{X}\left(x\right)\right)

Finally, p(x,y,t)p\left(x,y,t\right) satisfies the Chapman-Kolmogorov equation, i.e., for every x,y(0,1)x,y\in\left(0,1\right) and t,s>0t,s>0,

(5.9) p(x,y,t+s)=0p(x,ξ,t)p(ξ,y,s)𝑑ξ.p\left(x,y,t+s\right)=\int_{0}^{\infty}p\left(x,\xi,t\right)p\left(\xi,y,s\right)d\xi.
Proof.

The only thing that requires proof is the smoothness of p(x,y,t)p\left(x,y,t\right) on (0,1)2×(0,)\left(0,1\right)^{2}\times\left(0,\infty\right). By Theorem 4.3, we know that (y,t)p(x,y,t)\left(y,t\right)\mapsto p\left(x,y,t\right) is smooth, and at the same time (x,t)p(x,y,t)\left(x,t\right)\mapsto p\left(x,y,t\right) solves the equation (tLα,β)p(x,y,t)=0\left(\partial_{t}-L_{\alpha,\beta}\right)p\left(x,y,t\right)=0. It is easy to see from here that p(x,y,t)p\left(x,y,t\right) has all the partial derivatives in (x,y,t)\left(x,y,t\right) of all orders. ∎

The proposition above also leads to the wellposedness of the stochastic differential equation associated with Lα,βL_{\alpha,\beta}.

Corollary 5.2.

The stochastic differential equation (5.2) is well posed for every x(0,1)x\in\left(0,1\right) up to the hitting time at either 0 or 11 in the sense that if {X~(x,t):t0}\left\{\tilde{X}\left(x,t\right):t\geq 0\right\} is another solution to (5.2), then the distribution of X(x,t)X\left(x,t\right) conditioning on t<ζ0,1X(x)t<\zeta_{0,1}^{X}\left(x\right) is identical with that of X~(x,t)\tilde{X}\left(x,t\right) given t<ζ0,1X~(x)t<\zeta_{0,1}^{\tilde{X}}\left(x\right).

Proof.

It is sufficient to observe that, for every fCc((0,1))f\in C_{c}\left(\left(0,1\right)\right), if uf(x,t)u_{f}\left(x,t\right) is defined as in (5.6), then {uf(X~(x,s),ts):s[0,t]}\left\{u_{f}\left(\tilde{X}\left(x,s\right),t-s\right):s\in\left[0,t\right]\right\} is a martingale, which, by (5.7), implies that

𝔼[f(X(x,t));t<ζ0,1X(x)]=uf(x,t)=𝔼[f(X~(x,t));t<ζ0,1X~(x)].\mathbb{E}\left[f\left(X\left(x,t\right)\right);t<\zeta_{0,1}^{X}\left(x\right)\right]=u_{f}\left(x,t\right)=\mathbb{E}\left[f\left(\tilde{X}\left(x,t\right)\right);t<\zeta_{0,1}^{\tilde{X}}\left(x\right)\right].

Next we briefly discuss the other way of constructing p(x,y,t)p\left(x,y,t\right), which is to start with the localization of (5.1) in a neighborhood of the right boundary 11. It is easy to see that, by exchanging xx and 1x1-x, and at the same time exchanging α\alpha and β\beta, 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

z=ϕ(R)(x)=14(b(R)(x))2 where b(R)(x):=x1sα/2(1s)β/2𝑑s;z=\phi^{(R)}\left(x\right)=\frac{1}{4}\left(b^{(R)}\left(x\right)\right)^{2}\text{ where }b^{(R)}\left(x\right):=\int_{x}^{1}s^{-\alpha/2}\left(1-s\right)^{-\beta/2}ds;

q(R)(z,w,t)q^{\left(R\right)}\left(z,w,t\right) is the fundamental solution to the model equation with ν(R)=1β2β\nu^{(R)}=\frac{1-\beta}{2-\beta}, and given I(0,1)I\in\left(0,1\right), qϕ(R)(I)(R)(z,w,t)q_{\phi^{(R)}\left(I\right)}^{(R)}\left(z,w,t\right) is the fundamental solution to the localization of the model equation on (0,ϕ(R)(I))\left(0,\phi^{(R)}\left(I\right)\right); furthermore, we have that

Θ(ϕ(R)(x))=xα4(1x)β4(b(R)(x))β2(2β) with ΘI(R)=(22β)β2(2β)Iα2(2β),\Theta\left(\phi^{(R)}\left(x\right)\right)=\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{\left(b^{(R)}\left(x\right)\right)^{\frac{\beta}{2\left(2-\beta\right)}}}\text{ with }\Theta_{I}^{(R)}=\left(\frac{2}{2-\beta}\right)^{\frac{\beta}{2\left(2-\beta\right)}}I^{-\frac{\alpha}{2\left(2-\beta\right)}},

and for every x(I,1)x\in\left(I,1\right),

|V(ϕ(R)(x))|VI(R)(ϕ(R)(x))1β2β with VI(R)=α16(4α+2β)Iα2β2;\left|V\left(\phi^{(R)}\left(x\right)\right)\right|\leq V_{I}^{(R)}\left(\phi^{(R)}\left(x\right)\right)^{-\frac{1-\beta}{2-\beta}}\text{ with }V_{I}^{(R)}=\frac{\alpha}{16}\left(4-\alpha+2\beta\right)I^{\frac{\alpha}{2-\beta}-2};

we construct qϕ(R)(I)V,(R)(z,w,t)q_{\phi^{(R)}\left(I\right)}^{V,(R)}\left(z,w,t\right) from qϕ(R)(I)(R)(z,w,t)q_{\phi^{(R)}\left(I\right)}^{(R)}\left(z,w,t\right) via Duhamel’s method, and obtain the fundamental solution to (5.1) localized on (I,1)\left(I,1\right) as

pI(R)(x,y,t)=qϕ(R)(I)V,(R)(ϕ(R)(x),ϕ(R)(y),t)xα4(1x)β42y3α4(1y)3β4(b(R)(y))4β2(2β)(b(R)(x))β2(2β);p_{I}^{(R)}\left(x,y,t\right)=q_{\phi^{(R)}\left(I\right)}^{V,(R)}\left(\phi^{(R)}\left(x\right),\phi^{(R)}\left(y\right),t\right)\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{2y^{\frac{3\alpha}{4}}\left(1-y\right)^{\frac{3\beta}{4}}}\frac{\left(b^{(R)}\left(y\right)\right)^{\frac{4-\beta}{2\left(2-\beta\right)}}}{\left(b^{(R)}\left(x\right)\right)^{\frac{\beta}{2\left(2-\beta\right)}}};

finally, if {η~n(x):n}\left\{\tilde{\eta}_{n}\left(x\right):n\in\mathbb{N}\right\} is the sequence of hitting times that records the upward crossings of X(x,t)X\left(x,t\right) from II to HH. Then, (5.3) and the strong Markov property of {X(x,t):t0}\left\{X\left(x,t\right):t\geq 0\right\} are sufficient for us to obtain another version of the fundamental solution, denoted by p~(x,y,t)\tilde{p}\left(x,y,t\right) temporarily, as

(5.10) p~(x,y,t)=limG0pG(R)(x,y,t)=pI(R)(x,y,t)+n=1𝔼[pI(R)(H,y,tη~2n(x));η~2n(x)t,η~2n(x)<ζ0,1X(x)].\begin{split}\tilde{p}\left(x,y,t\right)&=\lim_{G\searrow 0}p_{G}^{(R)}\left(x,y,t\right)\\ &=p_{I}^{(R)}\left(x,y,t\right)+\sum_{n=1}^{\infty}\mathbb{E}\left[p_{I}^{(R)}\left(H,y,t-\tilde{\eta}_{2n}\left(x\right)\right);\tilde{\eta}_{2n}\left(x\right)\leq t,\tilde{\eta}_{2n}\left(x\right)<\zeta_{0,1}^{X}\left(x\right)\right].\end{split}

for (x,y,t)(H,1)2×(0,)\left(x,y,t\right)\in\left(H,1\right)^{2}\times\left(0,\infty\right). It is easy to see that p~(x,y,t)\tilde{p}\left(x,y,t\right) also satisfies (5.5), (5.8) and (5.9), which implies that p~(x,y,t)=p(x,y,t)\tilde{p}\left(x,y,t\right)=p\left(x,y,t\right) almost everywhere on (0,1)2×(0,)\left(0,1\right)^{2}\times\left(0,\infty\right), i.e., the two constructions of the fundamental solution to (5.1) are consistent and p(x,y,t)p\left(x,y,t\right) 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 p(x,y,t)p\left(x,y,t\right). For example, when both xx and yy are close to one of the boundaries, we can develop approximations for p(x,y,t)p\left(x,y,t\right) similarly as in §4.2\mathsection 4.2.

Corollary 5.3.

For (x,y,t)(0,1)2×(0,)\left(x,y,t\right)\in\left(0,1\right)^{2}\times\left(0,\infty\right), set

p(L)approx.(x,y,t):=q(L)(ϕ(L)(x),ϕ(L)(y),t)xα4(1x)β42y3α4(1y)3β4(b(L)(y))4α2(2α)(b(L)(x))α2(2α)p^{(L)-approx.}\left(x,y,t\right):=q^{(L)}\left(\phi^{(L)}\left(x\right),\phi^{(L)}\left(y\right),t\right)\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{2y^{\frac{3\alpha}{4}}\left(1-y\right)^{\frac{3\beta}{4}}}\frac{\left(b^{(L)}\left(y\right)\right)^{\frac{4-\alpha}{2\left(2-\alpha\right)}}}{\left(b^{(L)}\left(x\right)\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}}

and

p(R)approx.(x,y,t)=q(R)(ϕ(R)(x),ϕ(R)(y),t)xα4(1x)β42y3α4(1y)3β4(b(R)(y))4β2(2β)(b(R)(x))β2(2β).p^{(R)-approx.}\left(x,y,t\right)=q^{(R)}\left(\phi^{(R)}\left(x\right),\phi^{(R)}\left(y\right),t\right)\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{2y^{\frac{3\alpha}{4}}\left(1-y\right)^{\frac{3\beta}{4}}}\frac{\left(b^{(R)}\left(y\right)\right)^{\frac{4-\beta}{2\left(2-\beta\right)}}}{\left(b^{(R)}\left(x\right)\right)^{\frac{\beta}{2\left(2-\beta\right)}}}.

Let M(L)(t)M^{(L)}\left(t\right) and, respectively, M(R)(t)M^{(R)}\left(t\right) be defined as in (3.4) with ν=ν(L)\nu=\nu^{(L)}, VI=VI(L)V_{I}=V_{I}^{(L)} and, respectively, ν=ν(R)\nu=\nu^{(R)}, VI=VI(R)V_{I}=V_{I}^{(R)}. Fix 0<G<I<H<10<G<I<H<1, and set

t(L):=(4(2α)9(3α)ϕ(L)(G))(IG)24Mα,β and t(R):=(4(2β)9(3β)ϕ(R)(H))(HI)24Mα,β.t^{(L)}:=\left(\frac{4\left(2-\alpha\right)}{9\left(3-\alpha\right)}\phi^{(L)}\left(G\right)\right)\wedge\frac{\left(I-G\right)^{2}}{4M_{\alpha,\beta}}\text{ and }t^{(R)}:=\left(\frac{4\left(2-\beta\right)}{9\left(3-\beta\right)}\phi^{(R)}\left(H\right)\right)\wedge\frac{\left(H-I\right)^{2}}{4M_{\alpha,\beta}}.

Then, for every t(0,t(L))t\in\left(0,t^{(L)}\right) and (x,y)(0,G)2\left(x,y\right)\in\left(0,G\right)^{2} such that ϕ(L)(x)ϕ(L)(y)19ϕ(L)(G)\phi^{(L)}\left(x\right)\vee\phi^{(L)}\left(y\right)\leq\frac{1}{9}\phi^{(L)}\left(G\right),

|p(x,y,t)p(L)approx.(x,y,t)1|\displaystyle\left|\frac{p\left(x,y,t\right)}{p^{(L)-approx.}\left(x,y,t\right)}-1\right| M(L)(t)1\displaystyle\leq M^{(L)}\left(t\right)-1
+[1+(22α)α2αM(L)(t)(1G)2β2α(G1G1)]exp(2ϕ(L)(G)9t).\displaystyle\hskip 14.22636pt+\left[1+\left(\frac{2}{2-\alpha}\right)^{\frac{\alpha}{2-\alpha}}\frac{M^{(L)}\left(t\right)}{\left(1-G\right)^{\frac{2\beta}{2-\alpha}}}\left(\frac{G}{1-G}\wedge 1\right)\right]\exp\left(-\frac{2\phi^{(L)}\left(G\right)}{9t}\right).

Similarly, for every t(0,t(R))t\in\left(0,t^{(R)}\right) and (x,y)(H,1)2\left(x,y\right)\in\left(H,1\right)^{2} such that ϕ(R)(x)ϕ(R)(y)19ϕ(R)(H)\phi^{(R)}\left(x\right)\vee\phi^{(R)}\left(y\right)\leq\frac{1}{9}\phi^{(R)}\left(H\right),

|p(x,y,t)p(R)approx.(x,y,t)1|\displaystyle\left|\frac{p\left(x,y,t\right)}{p^{(R)-approx.}\left(x,y,t\right)}-1\right| M(R)(t)1\displaystyle\leq M^{(R)}\left(t\right)-1
+[1+(22β)β2βM(R)(t)H2α2β(1HH1)]exp(2ϕ(R)(H)9t).\displaystyle\hskip 14.22636pt+\left[1+\left(\frac{2}{2-\beta}\right)^{\frac{\beta}{2-\beta}}\frac{M^{(R)}\left(t\right)}{H^{\frac{2\alpha}{2-\beta}}}\left(\frac{1-H}{H}\wedge 1\right)\right]\exp\left(-\frac{2\phi^{(R)}\left(H\right)}{9t}\right).
Proof.

We only need to look at the statement involving p(L)approx.(x,y,t)p^{(L)-approx.}\left(x,y,t\right). There is not much to be done since a similar estimate (4.16) has been proven in Theorem 4.5. We notice that t(L)t^{(L)} is chosen such that the function ssν2exp(4ϕ(L)(G)9s)s\mapsto s^{\nu-2}\exp\left(-\frac{4\phi^{(L)}\left(G\right)}{9s}\right) is increasing on (0,t(L))\left(0,t^{(L)}\right), and (IG)24Mα,βt(L)\left(I-G\right)^{2}\geq 4M_{\alpha,\beta}t^{(L)}. Furthermore, in this case S(x)=xS\left(x\right)=x for every x(0,1)x\in\left(0,1\right), and hence 𝔭G=G\mathfrak{p}_{G}=G. 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 (x,y,t)(0,G)2×(0,t(L))\left(x,y,t\right)\in\left(0,G\right)^{2}\times\left(0,t^{(L)}\right) as described in the statement,

|p(x,y,t)pI(L)(x,y,t)p(L)approx.(x,y,t)|\displaystyle\left|\frac{p\left(x,y,t\right)-p_{I}^{(L)}\left(x,y,t\right)}{p^{(L)-approx.}\left(x,y,t\right)}\right| M(L)(t)Θ(ϕ(L)(G))Θ(ϕ(L)(x))(ϕ(L)(G)ϕ(L)(x))12αxGexp(2ϕ(L)(G)9t)G1G\displaystyle\leq M^{(L)}\left(t\right)\frac{\Theta\left(\phi^{(L)}\left(G\right)\right)}{\Theta\left(\phi^{(L)}\left(x\right)\right)}\left(\frac{\phi^{(L)}\left(G\right)}{\phi^{(L)}\left(x\right)}\right)^{\frac{1}{2-\alpha}}\frac{x}{G}\exp\left(-\frac{2\phi^{(L)}\left(G\right)}{9t}\right)\frac{G}{1-G}
(22α)α2α(1G)2β2αM(L)(t)exp(2ϕ(L)(G)9t)(G1G1).\displaystyle\leq\left(\frac{2}{2-\alpha}\right)^{\frac{\alpha}{2-\alpha}}\left(1-G\right)^{-\frac{2\beta}{2-\alpha}}M^{(L)}\left(t\right)\exp\left(-\frac{2\phi^{(L)}\left(G\right)}{9t}\right)\left(\frac{G}{1-G}\wedge 1\right).

6. Appendix

This Appendix contains detailed derivations involving Θ(z)\Theta\left(z\right) and V(z)V\left(z\right) for z(0,J)z\in\left(0,J\right). Assuming that J=ϕ(I)J=\phi\left(I\right), it is sufficient for us to look at Θ(ϕ(x))\Theta\left(\phi\left(x\right)\right) and V(ϕ(x))V\left(\phi\left(x\right)\right) for x(0,I)x\in\left(0,I\right), where the notations become simpler. Recall that

aI:=maxx[0,I]{1a(x),a(x)} and bI:=maxx[0,I]|b(x)|.a_{I}:=\max_{x\in\left[0,I\right]}\left\{\frac{1}{a\left(x\right)},a\left(x\right)\right\}\text{ and }b_{I}:=\max_{x\in\left[0,I\right]}\left|b\left(x\right)\right|.

We also introduce two more notations:

aI:=maxx[0,I]|a(x)| and bI:=maxx[0,I]|b(x)|.a_{I}^{\prime}:=\max_{x\in\left[0,I\right]}\left|a^{\prime}\left(x\right)\right|\text{ and }b_{I}^{\prime}:=\max_{x\in\left[0,I\right]}\left|b^{\prime}\left(x\right)\right|.

According to (1.2) and (2.2), we have that for every x(0,I)x\in\left(0,I\right),

Θ(ϕ(x))=exp(0xθ(w)2ϕ(w)ϕ(w)𝑑w)=exp(0x(12ν2ϕ(w)wαa(w)(wαa(w))4wαa(w)+b(w)2wαa(w))𝑑w).\begin{split}\Theta\left(\phi\left(x\right)\right)&=\exp\left(-\int_{0}^{x}\frac{\theta(w)}{2\phi\left(w\right)}\phi^{\prime}\left(w\right)dw\right)\\ &=\exp\left(-\int_{0}^{x}\left(\frac{\frac{1}{2}-\nu}{2\sqrt{\phi\left(w\right)w^{\alpha}a\left(w\right)}}-\frac{\left(w^{\alpha}a\left(w\right)\right)^{\prime}}{4w^{\alpha}a\left(w\right)}+\frac{b\left(w\right)}{2w^{\alpha}a\left(w\right)}\right)dw\right).\end{split}

Notice that

12ν2ϕ(w)wαa(w)(wαa(w))4wαa(w)=(ln(2ϕ(w))12ν(wαa(w))14),\frac{\frac{1}{2}-\nu}{2\sqrt{\phi\left(w\right)w^{\alpha}a\left(w\right)}}-\frac{\left(w^{\alpha}a\left(w\right)\right)^{\prime}}{4w^{\alpha}a\left(w\right)}=\left(\ln\frac{\left(2\sqrt{\phi\left(w\right)}\right)^{\frac{1}{2}-\nu}}{\left(w^{\alpha}a\left(w\right)\right)^{\frac{1}{4}}}\right)^{\prime},

and further, if α=1\alpha=1, then

b(w)2wa(w)=(ln(wb(0)2))+12w(b(w)a(w)b(0)).\frac{b\left(w\right)}{2wa\left(w\right)}=\left(\ln\left(w^{\frac{b\left(0\right)}{2}}\right)\right)^{\prime}+\frac{1}{2w}\left(\frac{b\left(w\right)}{a\left(w\right)}-b\left(0\right)\right).

Plugging these two expressions back into the right hand side of Θ(ϕ(x))\Theta\left(\phi\left(x\right)\right) leads to

Θ(ϕ(x))={xα4(4ϕ(x))α4(2α)(a(x))14exp(0xb(w)2wαa(w)𝑑w) if α1,(x4ϕ(x))14b(0)2(a(x))14exp(0x12w(b(w)a(w)b(0))𝑑w) if α=1,\Theta\left(\phi\left(x\right)\right)=\begin{cases}x^{\frac{\alpha}{4}}\left(4\phi\left(x\right)\right)^{-\frac{\alpha}{4\left(2-\alpha\right)}}\left(a\left(x\right)\right)^{\frac{1}{4}}\exp\left(-\int_{0}^{x}\frac{b\left(w\right)}{2w^{\alpha}a\left(w\right)}dw\right)&\text{ if }\alpha\neq 1,\\ \left(\frac{x}{4\phi\left(x\right)}\right)^{\frac{1}{4}-\frac{b\left(0\right)}{2}}\left(a\left(x\right)\right)^{\frac{1}{4}}\exp\left(-\int_{0}^{x}\frac{1}{2w}\left(\frac{b\left(w\right)}{a\left(w\right)}-b\left(0\right)\right)dw\right)&\text{ if }\alpha=1,\end{cases}

which is exactly (2.6). Given (H1) and (H2), the integral in the exponential function above is well defined in both cases (when α=1\alpha=1 and α1\alpha\neq 1).

With the notations introduced above, we have that when α1\alpha\neq 1, for every x(0,I)x\in\left(0,I\right),

(1α2)α2(2α)aI12(2α)xα4(a(x))14(4ϕ(x))α4(2α)(1α2)α2(2α)aI12(2α)\left(1-\frac{\alpha}{2}\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}a_{I}^{-\frac{1}{2\left(2-\alpha\right)}}\leq\frac{x^{\frac{\alpha}{4}}\left(a\left(x\right)\right)^{\frac{1}{4}}}{\left(4\phi\left(x\right)\right)^{\frac{\alpha}{4\left(2-\alpha\right)}}}\leq\left(1-\frac{\alpha}{2}\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}a_{I}^{\frac{1}{2\left(2-\alpha\right)}}

and

exp(0x|b(w)|2wαa(w)𝑑w)𝕀(0,1)(α)eaIbII1α2(1α)+𝕀(1,2)(α)eaIbII2α2(2α);\exp\left(\int_{0}^{x}\frac{\left|b\left(w\right)\right|}{2w^{\alpha}a\left(w\right)}dw\right)\leq\mathbb{I}_{\left(0,1\right)}\left(\alpha\right)\cdot e^{\frac{a_{I}b_{I}I^{1-\alpha}}{2\left(1-\alpha\right)}}+\mathbb{I}_{\left(1,2\right)}\left(\alpha\right)\cdot e^{\frac{a_{I}b_{I}^{\prime}I^{2-\alpha}}{2\left(2-\alpha\right)}};

when α=1\alpha=1, for every x(0,I)x\in\left(0,I\right),

2b(0)12aI12b(0)12(x4ϕ(x))14b(0)2(a(x))142b(0)12aI1212b(0)2^{b\left(0\right)-\frac{1}{2}}a_{I}^{\frac{1}{2}b\left(0\right)-\frac{1}{2}}\leq\left(\frac{x}{4\phi\left(x\right)}\right)^{\frac{1}{4}-\frac{b\left(0\right)}{2}}\left(a\left(x\right)\right)^{\frac{1}{4}}\leq 2^{b\left(0\right)-\frac{1}{2}}a_{I}^{\frac{1}{2}-\frac{1}{2}b\left(0\right)}

and

exp(0x12w|b(w)a(w)b(0)|𝑑w)e12aI2(aIbI+aIbI)I.\exp\left(\int_{0}^{x}\frac{1}{2w}\left|\frac{b\left(w\right)}{a\left(w\right)}-b\left(0\right)\right|dw\right)\leq e^{\frac{1}{2}a_{I}^{2}\left(a_{I}b_{I}^{\prime}+a_{I}^{\prime}b_{I}\right)I}.

Hence, if we set

AI:=𝕀(0,1)(α)eaIbII1α2(1α)+𝕀{1}(α)e12aI2(aIbI+aIbI)I+𝕀(1,2)(α)eaIbII2α2(2α).A_{I}:=\mathbb{I}_{\left(0,1\right)}\left(\alpha\right)\cdot e^{\frac{a_{I}b_{I}I^{1-\alpha}}{2\left(1-\alpha\right)}}+\mathbb{I}_{\left\{1\right\}}\left(\alpha\right)\cdot e^{\frac{1}{2}a_{I}^{2}\left(a_{I}b_{I}^{\prime}+a_{I}^{\prime}b_{I}\right)I}+\mathbb{I}_{\left(1,2\right)}\left(\alpha\right)\cdot e^{\frac{a_{I}b_{I}^{\prime}I^{2-\alpha}}{2\left(2-\alpha\right)}}.

then for every x(0,I)x\in\left(0,I\right),

(1α2)12νaI1ν2AI1Θ(ϕ(x))(1α2)12νaI1ν2AI.\left(1-\frac{\alpha}{2}\right)^{\frac{1}{2}-\nu}a_{I}^{-\frac{1-\nu}{2}}A_{I}^{-1}\leq\Theta\left(\phi\left(x\right)\right)\leq\left(1-\frac{\alpha}{2}\right)^{\frac{1}{2}-\nu}a_{I}^{\frac{1-\nu}{2}}A_{I}.

(2.7) follows from here by setting

(6.1) ΘI:=((1α2)ν122)aI1ν2AI.\Theta_{I}:=\left(\left(1-\frac{\alpha}{2}\right)^{\nu-\frac{1}{2}}\vee\sqrt{2}\right)a_{I}^{\frac{1-\nu}{2}}A_{I}.

For every x,y(0,I)x,y\in\left(0,I\right), we can follow the arguments above to get that

aI1ν2AI1Θ(ϕ(x))Θ(ϕ(y))aI1ν2AI.a_{I}^{-\frac{1-\nu}{2}}A_{I}^{-1}\leq\frac{\Theta\left(\phi\left(x\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\leq a_{I}^{\frac{1-\nu}{2}}A_{I}.

Moreover, if S(x)S\left(x\right) is as defined in (4.1), then

S(x)=22ν10xexp(0ub(w)wαa(w)𝑑w)𝑑u.S\left(x\right)=2^{2\nu-1}\int_{0}^{x}\exp\left(-\int_{0}^{u}\frac{b\left(w\right)}{w^{\alpha}a\left(w\right)}dw\right)du.

In addition, from (3.34), we can easily derive that, for every x,y(0,I)x,y\in\left(0,I\right),

(6.2) Θ(ϕ(x))Θ(ϕ(y))ϕ(y)=(ϕ(y)ϕ(x))14ν2ϕ12(y)xα4a14(x)y3α4a34(y)exp(yxb(w)2wαa(w)𝑑w),\begin{split}\frac{\Theta\left(\phi\left(x\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)&=\left(\frac{\phi\left(y\right)}{\phi\left(x\right)}\right)^{\frac{1}{4}-\frac{\nu}{2}}\frac{\phi^{\frac{1}{2}}\left(y\right)x^{\frac{\alpha}{4}}a^{\frac{1}{4}}\left(x\right)}{y^{\frac{3\alpha}{4}}a^{\frac{3}{4}}\left(y\right)}\exp\left(-\int_{y}^{x}\frac{b\left(w\right)}{2w^{\alpha}a\left(w\right)}dw\right),\end{split}

and

(6.3) (ϕ(x))1νϕ(x)Θ2(ϕ(x))=(ϕ(x))12νxαa(x)(4ϕ(x))α2(2α)exp(0xb(w)wαa(w)𝑑w)=22ν1xαa(x)exp(0xb(w)wαa(w)𝑑w).\begin{split}\frac{\left(\phi\left(x\right)\right)^{1-\nu}}{\phi^{\prime}\left(x\right)}\Theta^{2}\left(\phi\left(x\right)\right)&=\frac{\left(\phi\left(x\right)\right)^{\frac{1}{2}-\nu}x^{\alpha}a\left(x\right)}{\left(4\phi\left(x\right)\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}}\exp\left(-\int_{0}^{x}\frac{b\left(w\right)}{w^{\alpha}a\left(w\right)}dw\right)\\ &=2^{2\nu-1}x^{\alpha}a\left(x\right)\exp\left(-\int_{0}^{x}\frac{b\left(w\right)}{w^{\alpha}a\left(w\right)}dw\right).\end{split}

Now we move onto V(z)V\left(z\right) and recall from (3.33) that for every x(0,I)x\in\left(0,I\right),

V(ϕ(x))=θ(x)4ϕ(x)(θ(x)+22ν)θ(x)2ϕ(x) for every x(0,I).V\left(\phi\left(x\right)\right)=\frac{\theta\left(x\right)}{4\phi\left(x\right)}\left(-\theta\left(x\right)+2-2\nu\right)-\frac{\theta^{\prime}\left(x\right)}{2\phi^{\prime}\left(x\right)}\text{ for every }x\in\left(0,I\right).

By (1.2), the choice of ν\nu and (H1) and (H2), it is straightforward to verify that when x(0,I)x\in\left(0,I\right),

ϕ(x)=x2α(2α)2(1+𝒪(x)) and θ(x)=b(0)x1α2α𝕀(0,1)(α)+𝒪(x2α),\phi\left(x\right)=\frac{x^{2-\alpha}}{\left(2-\alpha\right)^{2}}\left(1+\mathcal{O}\left(x\right)\right)\text{ and }\theta\left(x\right)=\frac{b\left(0\right)x^{1-\alpha}}{2-\alpha}\mathbb{I}_{\left(0,1\right)}\left(\alpha\right)+\mathcal{O}\left(x^{2-\alpha}\right),

which implies that

θ(x)ϕ(x)=(2α)b(0)x𝕀(0,1)(α)+𝒪(1).\begin{split}\frac{\theta\left(x\right)}{\phi\left(x\right)}&=\frac{\left(2-\alpha\right)b\left(0\right)}{x}\mathbb{I}_{\left(0,1\right)}\left(\alpha\right)+\mathcal{O}\left(1\right).\end{split}

In addition, we also have that

θ(x)ϕ(x)=b(x)(xαa(x))′′2+2b(x)(xαa(x))2(12ϕ(x)xαa(x)(xαa(x))2xαa(x))=b(x)+b(x)2ϕ(x)xαa(x)b(x)(xαa(x))2xαa(x)(xαa(x))′′2(xαa(x))4ϕ(x)xαa(x)+((xαa(x)))24xαa(x).\begin{split}\frac{\theta^{\prime}\left(x\right)}{\phi^{\prime}\left(x\right)}&=b^{\prime}\left(x\right)-\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime\prime}}{2}+\frac{2b\left(x\right)-\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{2}\left(\frac{1}{2\sqrt{\phi\left(x\right)x^{\alpha}a\left(x\right)}}-\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{2x^{\alpha}a\left(x\right)}\right)\\ &=b^{\prime}\left(x\right)+\frac{b\left(x\right)}{2\sqrt{\phi\left(x\right)x^{\alpha}a\left(x\right)}}-\frac{b\left(x\right)\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{2x^{\alpha}a\left(x\right)}-\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime\prime}}{2}-\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{4\sqrt{\phi\left(x\right)x^{\alpha}a\left(x\right)}}+\frac{\left(\left(x^{\alpha}a\left(x\right)\right)^{\prime}\right)^{2}}{4x^{\alpha}a\left(x\right)}.\end{split}

We notice that

(xαa(x))′′2(xαa(x))4ϕ(x)xαa(x)+((xαa(x)))24xαa(x)=𝒪(xα1),-\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime\prime}}{2}-\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{4\sqrt{\phi\left(x\right)x^{\alpha}a\left(x\right)}}+\frac{\left(\left(x^{\alpha}a\left(x\right)\right)^{\prime}\right)^{2}}{4x^{\alpha}a\left(x\right)}=\mathcal{O}\left(x^{\alpha-1}\right),

and

b(x)2ϕ(x)xαa(x)b(x)(xαa(x))2xαa(x)=𝕀(0,1)(α)(b(0)(1α)x+𝒪(xα1))+𝒪(1).\frac{b\left(x\right)}{2\sqrt{\phi\left(x\right)x^{\alpha}a\left(x\right)}}-\frac{b\left(x\right)\left(x^{\alpha}a\left(x\right)\right)^{\prime}}{2x^{\alpha}a\left(x\right)}=\mathbb{I}_{\left(0,1\right)}\left(\alpha\right)\left(\frac{b\left(0\right)\left(1-\alpha\right)}{x}+\mathcal{O}\left(x^{\alpha-1}\right)\right)+\mathcal{O}\left(1\right).

Putting all the above together yields that when α(0,1)\alpha\in\left(0,1\right),

V(ϕ(x))=(1ν)(2α)b(0)2x(1α)b(0)2x+𝒪(xα1)=αb(0)2x+𝒪(xα1);V\left(\phi\left(x\right)\right)=\frac{\left(1-\nu\right)\left(2-\alpha\right)b\left(0\right)}{2x}-\frac{\left(1-\alpha\right)b\left(0\right)}{2x}+\mathcal{O}\left(x^{\alpha-1}\right)=\frac{\alpha b\left(0\right)}{2x}+\mathcal{O}\left(x^{\alpha-1}\right);

when α[1,2)\alpha\in[1,2), V(ϕ(x))V\left(\phi\left(x\right)\right) is bounded for x(0,I)x\in\left(0,I\right). Thus, we have proven all the claims in Lemma 2.2.

Next, we look at the case when b(x)0b\left(x\right)\equiv 0, where most of the expressions above take simpler forms. For example,

Θ(ϕ(x))=xα4a14(x)2α2(2α)(ϕ(x))α4(2α),V(ϕ(x))=α(α4)16(2α)2ϕ(x)+(xαa(x))′′43((xαa(x)))216,\Theta\left(\phi\left(x\right)\right)=\frac{x^{\frac{\alpha}{4}}a^{\frac{1}{4}}\left(x\right)}{2^{\frac{\alpha}{2\left(2-\alpha\right)}}\left(\phi\left(x\right)\right)^{\frac{\alpha}{4\left(2-\alpha\right)}}},\;V\left(\phi\left(x\right)\right)=-\frac{\alpha\left(\alpha-4\right)}{16\left(2-\alpha\right)^{2}\phi\left(x\right)}+\frac{\left(x^{\alpha}a\left(x\right)\right)^{\prime\prime}}{4}-\frac{3\left(\left(x^{\alpha}a\left(x\right)\right)^{\prime}\right)^{2}}{16},
Θ(ϕ(x))Θ(ϕ(y))ϕ(y)=(ϕ(y)ϕ(x))α4(2α)ϕ12(y)xα4a14(x)y3α4a34(y) and (ϕ(x))1νϕ(x)Θ2(ϕ(x))=2α2αxαa(x).\frac{\Theta\left(\phi\left(x\right)\right)}{\Theta\left(\phi\left(y\right)\right)}\phi^{\prime}\left(y\right)=\left(\frac{\phi\left(y\right)}{\phi\left(x\right)}\right)^{\frac{\alpha}{4\left(2-\alpha\right)}}\frac{\phi^{\frac{1}{2}}\left(y\right)x^{\frac{\alpha}{4}}a^{\frac{1}{4}}\left(x\right)}{y^{\frac{3\alpha}{4}}a^{\frac{3}{4}}\left(y\right)}\text{ and }\frac{\left(\phi\left(x\right)\right)^{1-\nu}}{\phi^{\prime}\left(x\right)}\Theta^{2}\left(\phi\left(x\right)\right)=2^{-\frac{\alpha}{2-\alpha}}x^{\alpha}a\left(x\right).

In particular, if a(x)=(1x)βa\left(x\right)=\left(1-x\right)^{\beta} as in §5\mathsection 5, then

(6.4) Θ(ϕ(x))=xα4(1x)β42α2(2α)(0xdssα(1s)β)α4(2α) with ΘI=(22α)α2(2α)(1I)β2(2α).\Theta\left(\phi\left(x\right)\right)=\frac{x^{\frac{\alpha}{4}}\left(1-x\right)^{\frac{\beta}{4}}}{2^{\frac{\alpha}{2\left(2-\alpha\right)}}}\left(\int_{0}^{x}\frac{ds}{\sqrt{s^{\alpha}\left(1-s\right)^{\beta}}}\right)^{\frac{-\alpha}{4\left(2-\alpha\right)}}\text{ with }\Theta_{I}=\frac{\left(\frac{2}{2-\alpha}\right)^{\frac{\alpha}{2\left(2-\alpha\right)}}}{\left(1-I\right)^{\frac{\beta}{2\left(2-\alpha\right)}}}.

Furthermore,

V(ϕ(x))=α(4α)4(2α)2(0xdssα(1s)β)2+α(α4)16xα2(1x)βαβ8xα1(1x)β1+β(β4)16xα(1x)β2.\begin{split}V\left(\phi\left(x\right)\right)&=\frac{\alpha\left(4-\alpha\right)}{4\left(2-\alpha\right)^{2}}\left(\int_{0}^{x}\frac{ds}{\sqrt{s^{\alpha}\left(1-s\right)^{\beta}}}\right)^{-2}+\frac{\alpha\left(\alpha-4\right)}{16}x^{\alpha-2}\left(1-x\right)^{\beta}\\ &\hskip 28.45274pt\hskip 28.45274pt\hskip 28.45274pt-\frac{\alpha\beta}{8}x^{\alpha-1}\left(1-x\right)^{\beta-1}+\frac{\beta\left(\beta-4\right)}{16}x^{\alpha}\left(1-x\right)^{\beta-2}.\end{split}

Since

(6.5) x2α(2α)2ϕ(x)(1x)βx2α(2α)2,\frac{x^{2-\alpha}}{\left(2-\alpha\right)^{2}}\leq\phi\left(x\right)\leq\left(1-x\right)^{-\beta}\frac{x^{2-\alpha}}{\left(2-\alpha\right)^{2}},

we see that

V(ϕ(x))xα2(1x)β2(αβx(1x)8+β(β4)16x2)V\left(\phi\left(x\right)\right)\geq x^{\alpha-2}\left(1-x\right)^{\beta-2}\left(-\frac{\alpha\beta x\left(1-x\right)}{8}+\frac{\beta\left(\beta-4\right)}{16}x^{2}\right)

and

V(ϕ(x))\displaystyle V\left(\phi\left(x\right)\right) α(4α)16xα2[1(1x)β]\displaystyle\leq\frac{\alpha\left(4-\alpha\right)}{16}x^{\alpha-2}\left[1-\left(1-x\right)^{\beta}\right]
+xα2(1x)β2(αβx(1x)8+β(β4)16x2)\displaystyle\hskip 28.45274pt\hskip 28.45274pt+x^{\alpha-2}\left(1-x\right)^{\beta-2}\left(-\frac{\alpha\beta x\left(1-x\right)}{8}+\frac{\beta\left(\beta-4\right)}{16}x^{2}\right)
xα1(1x)β2(α(4α)β16αβ(1x)8+β(β4)16x2).\displaystyle\leq x^{\alpha-1}\left(1-x\right)^{\beta-2}\left(\frac{\alpha\left(4-\alpha\right)\beta}{16}-\frac{\alpha\beta\left(1-x\right)}{8}+\frac{\beta\left(\beta-4\right)}{16}x^{2}\right).

where in the last line we used the fact that for every x(0,I)x\in\left(0,I\right),

1(1x)ββx(1x)β2.1-\left(1-x\right)^{\beta}\leq\beta x\left(1-x\right)^{\beta-2}.

Combining the upper bound and the lower bound of V(ϕ(x))V\left(\phi\left(x\right)\right) leads to

|V(ϕ(x))|xα1(1x)β2β16(4β+2α) for every x(0,I),\left|V\left(\phi\left(x\right)\right)\right|\leq x^{\alpha-1}\left(1-x\right)^{\beta-2}\frac{\beta}{16}\left(4-\beta+2\alpha\right)\text{ for every }x\in\left(0,I\right),

which, by (6.5), implies that when α(0,1)\alpha\in\left(0,1\right),

|V(ϕ(x))|β16(4β+2α)(2α)2α22α(1I)β2α2(ϕ(x))1α2α for every x(0,I).\left|V\left(\phi\left(x\right)\right)\right|\leq\frac{\beta}{16}\left(4-\beta+2\alpha\right)\left(2-\alpha\right)^{\frac{2\alpha-2}{2-\alpha}}\left(1-I\right)^{\frac{\beta}{2-\alpha}-2}\left(\phi\left(x\right)\right)^{-\frac{1-\alpha}{2-\alpha}}\text{ for every }x\in\left(0,I\right).

Therefore, with this specific case of a(x)=(1x)βa\left(x\right)=\left(1-x\right)^{\beta}, we see that the constant VIV_{I} as introduced Lemma 2.2 (identified with VJV_{J} in for J=ϕ(I)J=\phi\left(I\right)) can be taken as

(6.6) VI=β16(4β+2α)(1I)β2α2.V_{I}=\frac{\beta}{16}\left(4-\beta+2\alpha\right)\left(1-I\right)^{\frac{\beta}{2-\alpha}-2}.

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 c0c_{0}-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. c0c^{0}-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.