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

Time-changed Feller’s Brownian motions are birth-death processes

Liping Li Fudan University, Shanghai, China. [email protected]
Abstract.

A Feller’s Brownian motion refers to a Feller process on the interval [0,)[0,\infty) that is equivalent to the absorbing Brownian motion before reaching 0. It is fully determined by four parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}), reflecting its killing, reflecting, sojourn, and jumping behaviors at the boundary 0. On the other hand, a birth-death process is a continuous-time Markov chain on \mathbb{N} with a given transition density matrix QQ, and it is characterized by three parameters (γ,β,ν)(\gamma,\beta,\nu) that describe its killing, reflecting, and jumping behaviors at the boundary \infty. The primary objective of this paper is to establish a connection between Feller’s Brownian motion and birth-death process. We will demonstrate that any Feller’s Brownian motion can be transformed into a specific birth-death process through a unique time change transformation, and conversely, any birth-death process can be derived from Feller’s Brownian motion via time change. Specifically, the birth-death process generated by the Feller’s Brownian motion, determined by the parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}), through time change, has the parameters:

γ=p1,β=2p2,νn=𝔭n,n,\gamma=p_{1},\quad\beta=2p_{2},\quad\nu_{n}=\mathfrak{p}_{n},\;n\in\mathbb{N},

where {𝔭n:n}\{\mathfrak{p}_{n}:n\in\mathbb{N}\} is a sequence derived by allocating weights to the measure p4p_{4} in a specific manner. Utilizing the pathwise representation of Feller’s Brownian motion, our results provide a pathwise construction scheme for birth-death processes, addressing a gap in the existing literature.

Key words and phrases:
Feller’s Brownian motions, Birth-death processes, Continuous-time Markov chains, Time change, Boundary conditions, Local times, Dirichlet forms, Approximation.
2020 Mathematics Subject Classification:
Primary 60J27, 60J40, 60J46, 60J50, 60J74.
The author is a member of LMNS, Fudan University. He is partially supported by NSFC (No. 11931004 and 12371144).

1. Introduction

To study the boundary behavior of Markov processes, historically, two classical probabilistic models have been developed that seem different but are actually similar. The first model is based on Brownian motion on the half-line (0,)(0,\infty), allowing all possible behaviors of the process at the boundary 0 while ensuring the strong Markov property. This line of work began with Feller’s research on the theory of transition semigroup [5] and was later extended by Itô and McKean, who completed the pathwise construction of all such processes in [10]. These processes are referred to as Feller’s Brownian motion by Itô and McKean, with a detailed definition provided in [10, §5] (see also Definition 3.1). From an analytical perspective, the key to determining the domain of the generator of Feller’s Brownian motion is the boundary condition at 0:

p1f(0)p2f(0)+p32f′′(0)+(0,)[f(0)f(x)]p4(dx)=0,p_{1}f(0)-p_{2}f^{\prime}(0)+\frac{p_{3}}{2}f^{\prime\prime}(0)+\int_{(0,\infty)}[f(0)-f(x)]p_{4}(dx)=0, (1.1)

where p1,p2,p30p_{1},p_{2},p_{3}\geq 0 are constants, and p4p_{4} is a positive measure on (0,)(0,\infty) satisfying (0,)1xp4(dx)<\int_{(0,\infty)}1\wedge x\,p_{4}(dx)<\infty. From a probabilistic perspective, these four parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) represent the four possible types of boundary behavior of Feller’s Brownian motion at 0: killing, reflecting, sojourn, and jumping. The work of Itô and McKean [10] deeply explores the mechanisms underlying these four types of boundary behavior, providing a clear and intuitive pathwise representation of Feller’s Brownian motion; see §3.2 for a brief summarization.

The other type of model is a fundamental one in the theory of continuous-time Markov chains, known as the birth-death process. This process is induced by a standard transition matrix, whose derivative at time 0 is the density matrix QQ given by (2.1), on the discrete space \mathbb{N}. Its defining characteristic is that from any given point, the process can only jump to its immediate left or right neighbors, a feature akin to the continuity of Brownian motion trajectories. Since the birth-death process may reach \infty in finite time, describing its behavior at \infty and after reaching \infty is a fundamental problem in the theory of birth-death processes. This line of inquiry also began with Feller, who in [6] observed that the boundary behavior of the birth-death process at \infty bears some intrinsic similarity to the boundary behavior of the aforementioned Feller’s Brownian motion at 0. Feller’s approach was analytical; although he only examined partial birth-death processes, he derived the boundary condition at \infty for the resolvent function FF (i.e., the function in the domain of the generator) as follows:

γF()+β2F+()+k(F()F(k))νk=0,\gamma F(\infty)+\frac{\beta}{2}F^{+}(\infty)+\sum_{k\in\mathbb{N}}(F(\infty)-F(k))\nu_{k}=0, (1.2)

where F+F^{+} denotes the discrete gradient (see §2.3), γ,β0\gamma,\beta\geq 0 are constants, and ν=(νk)k\nu=(\nu_{k})_{k\in\mathbb{N}} is a positive measure on \mathbb{N}. Shortly thereafter, building on Feller’s work, Yang in 1965 (see [26, Chapter 7]) completed the analytical construction of all birth-death processes using the resolvent method, thereby finalizing the construction theory from an analytical perspective. In our recent article [18], we proved that all (non-Doob) birth-death processes satisfy the boundary condition (1.2) observed by Feller. Conversely, Wang, in his 1958 doctoral thesis, provided a probabilistic construction of all birth-death processes (see [26, Chapter 6]). However, it should be noted that this probabilistic construction differs from the pathwise construction of Feller’s Brownian motion by Itô and McKean, as it instead uses a series of Doob processes to approximate the original birth-death process. In fact, understanding the probabilistic intuition behind the parameters (γ,β,ν)(\gamma,\beta,\nu) through this construction is challenging, although, by comparing with (1.1), it becomes evident that these parameters should correspond to the killing, reflecting, and jumping behaviors of the birth-death process at \infty, respectively.

Apart from two minor differences, the structural consistency of the boundary conditions (1.1) and (1.2) is quite evident: First, the reflecting term in (1.1) is negative, while in (1.2) it is positive; this is because 0 and \infty are at opposite ends of their respective state spaces, thus causing the direction of the gradient to be exactly opposite. Second, (1.2) lacks a sojourn term; this is because in the definition of the birth-death process, the index set of the transition matrix is \mathbb{N}, which causes that the birth-death process can not visit \infty for a positive duration (correspondingly, the process that allows sojourn at \infty is called a generalized birth-death process, and its index set is {}\mathbb{\mathbb{N}}\cup\{\infty\}; see [6]). Naturally, we hope to find a deeper connection between the two beyond methodology.

For the symmetric case (from an analytical perspective, that is, when the transition semigroup satisfies both the Kolmogorov backward and forward equations; see [6]), there is no jumping behavior at the boundary for either process, and the boundary conditions reduce to the classic Dirichlet, Neumann, or Robin boundaries (see [18, §2.3]). At this point, Feller’s Brownian motion and birth-death process are both special cases of a more general class of processes known as quasidiffusions. These processes are also referred to as generalized diffusions or gap diffusions in some literature, such as [15, 14]. The study of quasidiffusions originated from Kac and Krein’s research on the spectral theory of a class of generalized second-order differential operators [12]. Soon after, these self-adjoint operators were realized as strong Markov processes on some closed subset of [,][-\infty,\infty] with trajectories that satisfy the skip-free property; see [13]. The skip-free property is a combination of the continuity of Brownian motion trajectories and the characteristic of birth-death process trajectories: the trajectory is continuous where the space is continuous and can only jump to adjacent positions where there is a gap in space. In addition to studying quasidiffusions within the classic Feller framework (see [24]), with the aid of symmetry, we can also use the Dirichlet form theory established by Fukushima (see [3, 8]) to study them, as demonstrated in several recent articles [17, 19, 16]. In general, the connection between symmetric Feller’s Brownian motions and symmetric birth-death processes is very clear: they can not only be unified under the more general quasidiffusion but also, the symmetric birth-death process is a time-changed process of a certain symmetric Feller’s Brownian motion. For example, the minimal birth-death process corresponds to the time-changed process of absorbing Brownian motion, and the (Q,1)(Q,1)-process corresponds to that of reflecting Brownian motion; see [18, §3].

However, for the non-symmetric case (especially the so-called ‘pathological’ case by Feller, where p4p_{4} or ν\nu is an infinite measure), the connection between Feller’s Brownian motion and birth-death process becomes very vague. Indeed, both theories have been extensively studied and developed within their respective fields. The former has given rise to a mature theory of diffusion processes, see, e.g., [11, 21], which is one of the most important classes of models in the general theory of Markov processes; the latter, originating from the discrete space, has also evolved a series of theories, such as continuous-time Markov chains (see, e.g., [4]) and Markov jump processes (see, e.g., [2]). However, during their parallel development, the connection between the two is more reflected in the mutual borrowing of research methods, and few people discuss the fit as reflected in the boundary conditions (1.1) and (1.2).

The goal of our paper is to construct the missing bridge between Feller’s Brownian motion and birth-death process. In summary, our main result is that any Feller’s Brownian motion can be transformed into a specific birth-death process through a unique time change transformation; conversely, any birth-death process can be derived through the time change transformation of some Feller’s Brownian motion.

To facilitate the explanation of our results, we first perform a spatial transformation on the birth-death process as described in §4.1, aligning it with the natural scale of Feller’s Brownian motion and reflecting the boundary point \infty so that it moves to 0, thereby matching the boundary point of Feller’s Brownian motion. The transformed space is denoted by (see (4.1))

E¯={c^n:n}{0},\overline{E}=\{\hat{c}_{n}:n\in\mathbb{N}\}\cup\{0\},

where the subscript nn of c^n\hat{c}_{n} represents the point before the transformation, and limnc^n=0\lim_{n\to\infty}\hat{c}_{n}=0 is topologically consistent with the one-point compactification {}\mathbb{N}\cup\{\infty\} of \mathbb{N}. Under this spatial transformation, aside from the change in the symbol of the state points, the transformed process, which we denote by X^\hat{X}, remains indistinguishable from the original birth-death process. For convenience, we refer to X^\hat{X} as a birth-death process on E¯\overline{E}. Additionally, due to the change in the direction of the gradient at the boundary point, the corresponding boundary condition for X^\hat{X} is modified to:

γF^(0)β2F^+(0)+k(F^(0)F^(c^k))νk=0,\gamma\hat{F}(0)-\frac{\beta}{2}\hat{F}^{+}(0)+\sum_{k\in\mathbb{N}}(\hat{F}(0)-\hat{F}(\hat{c}_{k}))\nu_{k}=0, (1.3)

where F^\hat{F} is the transformed function of FF in (1.2). After this modification, the aforementioned boundary condition closely resembles (1.1).

Our main results, namely Theorems 4.2 and 5.1, can be stated as follows: Given the birth-death density matrix QQ, for the Feller’s Brownian motion YY corresponding to the boundary condition (1.1), it can always (and uniquely, see Theorem 4.4) be transformed into a birth-death process on E¯\overline{E} through a time change transformation induced by a positive continuous additive functional A=(At)t0A=(A_{t})_{t\geq 0}. This positive continuous additive functional AA given by (4.6) is the integral of the local times of YY with respect to the speed measure μ\mu determined by QQ. Moreover, the parameters (γ,β,ν)(\gamma,\beta,\nu) exhibited in the boundary condition (1.3) of the birth-death process obtained by time change can be derived from the parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) of YY as follows:

γ=p1,β=2p2,νn=𝔭n,n,\gamma=p_{1},\quad\beta=2p_{2},\quad\nu_{n}=\mathfrak{p}_{n},\;n\in\mathbb{N}, (1.4)

where {𝔭n:n}\{\mathfrak{p}_{n}:n\in\mathbb{N}\} is the sequence obtained by allocating weights to the measure p4p_{4} in the manner of (5.1) and (5.2). Particularly, p4=n𝔭nδc^np_{4}=\sum_{n\in\mathbb{N}}\mathfrak{p}_{n}\delta_{\hat{c}_{n}} whenever p4p_{4} is supported on {c^n:n}\{\hat{c}_{n}:n\in\mathbb{N}\}. It should be noted that the sojourn parameter p3p_{3} of YY does not play any role in the expression (1.4). This is because any positive duration of stay at 0 by YY will be eliminated by the time change.

From the relationship between the two sets of parameters in (1.4), it is not difficult to further deduce that every birth-death process on E¯\overline{E} can be obtained by time change from a certain Feller’s Brownian motion; see Corollary 5.2. This conclusion also provides a pathwise construction for all birth-death processes: the trajectory of X^\hat{X} is the trace of the corresponding Feller’s Brownian motion on E¯\overline{E}; see §5.2.

Let us introduce some of the tools that will be used in the proof of our main results. For the symmetric case, we primarily employ the theory of Dirichlet forms. The concepts and notation we use are consistent with the foundational references [3, 8], so they will not be further elaborated upon in the text. For the non-symmetric case where p4p_{4} is a finite measure, it essentially involves the Ikeda-Nagasawa-Watanabe piecing out transformation (see [9]) of the symmetric case. This is very similar to the discussion of birth-death processes found in [18, §5 and §8]. Based on this piecing out transformation, we can directly construct the trajectories of the birth-death process without relying on Feller’s Brownian motion; see §6. The most challenging scenario arises when p4p_{4} is an infinite measure. In this case, the piecing out transformation becomes ineffective, and the jumps of Feller’s Brownian motion at the boundary become very frequent and complex. Notably, our previous study [18] also regrettably stops short of studying this case for the birth-death process. To address this situation, we draw on Wang’s idea of constructing an approximating sequence for the birth-death process (see [26]), and we similarly construct a sequence of approximating processes for Feller’s Brownian motion (see §7). The boundary behavior of these processes is relatively simple, and they can be linked to the approximating sequence of the corresponding birth-death process, thereby ultimately establishing the parameter relationship (1.4) between the target processes.

At the end of this section, we briefly describe the commonly used notation. The symbols contained in the Markov process Y=(Ω,𝒢,𝒢t,Yt,θt,𝐏x)Y=(\Omega,\mathscr{G},\mathscr{G}_{t},Y_{t},\theta_{t},\mathbf{P}_{x}) are standard, as seen in [1, 25]. For brevity, some elements are sometimes omitted, such as writing Y=(Ω,Yt,𝐏x)Y=(\Omega,Y_{t},\mathbf{P}_{x}). The expectation corresponding to the probability measure 𝐏x\mathbf{P}_{x} is denoted by 𝐄x\mathbf{E}_{x}. In this paper, the cemetery point for Markov processes is uniformly denoted by \partial. Given a topological space EE, (E),+(E),b(E),C(E)\mathcal{B}(E),\mathcal{B}_{+}(E),\mathcal{B}_{b}(E),C(E) and Cb(E)C_{b}(E) represent the space of all Borel measurable, non-negative Borel measurable, bounded Borel measurable functions, totality of continuous functions, and bounded continuous functions, respectively. If EE is an interval, Cc(E)C_{c}(E) and C0(E)C_{0}(E) denote the set of all continuous functions with compact support and the set of all continuous functions that equal 0 at the open endpoints, respectively. The notation for continuous functions with subscripts indicates differentiability, for example, C2C^{2} indicates twice continuously differentiable, while CC^{\infty} indicates infinitely continuously differentiable. The symbol δa\delta_{a} stands for the Dirac measure at a[0,)a\in[0,\infty).

2. Birth-death processes

In this paper, we will strive to use the same terminologies and symbols as in [18] for the birth-death processes. The following restates only some of the important content.

2.1. Elements of QQ-processes

We consider a birth-death density matrix as follows:

Q=(qij)i,j:=(q0b000a1q1b100a2q2b2),Q=(q_{ij})_{i,j\in\mathbb{N}}:=\left(\begin{array}[]{ccccc}-q_{0}&b_{0}&0&0&\cdots\\ a_{1}&-q_{1}&b_{1}&0&\cdots\\ 0&a_{2}&-q_{2}&b_{2}&\cdots\\ \cdots&\cdots&\cdots&\cdots&\cdots\end{array}\right), (2.1)

where ak>0a_{k}>0 for k1k\geq 1 and bk>0,qk=ak+bkb_{k}>0,q_{k}=a_{k}+b_{k} for k0k\geq 0. (Set a0=0a_{0}=0 for convenience.) A continuous-time Markov chain X=(Xt)t0X=(X_{t})_{t\geq 0} is called a birth-death QQ-process (or QQ-process for short) if its transition matrix (pij(t))i,j(p_{ij}(t))_{i,j\in\mathbb{N}} is standard and its density matrix is QQ, i.e., pij(0)=qijp^{\prime}_{ij}(0)=q_{ij} for i,ji,j\in\mathbb{N}. Readers are referred to [4, 26] for the terminologies concerning continuous-time Markov chains; see also [18]. In our consideration, two QQ-processes with the same transition matrix will not be distinguished. For convenience, we will also refer to such (pij(t))i,j(p_{ij}(t))_{i,j\in\mathbb{N}} as a QQ-process when there is no risk of confusion.

There always exists at least one QQ-process XminX^{\text{min}}, known as the minimal QQ-process. This process is killed at the first time it almost reaches \infty. Similar to a regular diffusion on an interval, the minimal QQ-process can be characterized by the scale function (on \mathbb{N}):

c0=0,c1=12b0,ck=12b0+i=2ka1a2ai12b0b1bi1,k2c_{0}=0,\quad c_{1}=\frac{1}{2b_{0}},\quad c_{k}=\frac{1}{2b_{0}}+\sum_{i=2}^{k}\frac{a_{1}a_{2}\cdots a_{i-1}}{2b_{0}b_{1}\cdots b_{i-1}},\;k\geq 2 (2.2)

and the speed measure μ\mu on \mathbb{N}:

μ({0}):=μ0=1,μ({k}):=μk=b0b1bk1a1a2ak,k1.\mu(\{0\}):=\mu_{0}=1,\quad\mu(\{k\}):=\mu_{k}=\frac{b_{0}b_{1}\cdots b_{k-1}}{a_{1}a_{2}\cdots a_{k}},\;k\geq 1. (2.3)

The transition matrix (pijmin(t))(p^{\text{min}}_{ij}(t)) of XminX^{\text{min}} is symmetric with respect to μ\mu in the sense that μipijmin(t)=μjpjimin(t)\mu_{i}p_{ij}^{\text{min}}(t)=\mu_{j}p_{ji}^{\text{min}}(t) for all i,ji,j\in\mathbb{N} and t0t\geq 0. For more details about this characterization, please refer to [18, §3.1].

2.2. Resolvent representation of QQ-processes

When \infty is regular or an exit (for XminX^{\text{min}}) in Feller’s sense (see, e.g., [18, Definition 3.3]), there exist other QQ-processes besides the minimal one, such as Doob processes (see, e.g., [18, §5]) and the (Q,1)(Q,1)-process (which is only applicable in the regular case; see, e.g., [18, §3.3]). In this case, we have c:=limkck<c_{\infty}:=\lim_{k\rightarrow\infty}c_{k}<\infty.

In what follows, let us present a well-known analytic approach to characterize all birth-death processes by solving their resolvents. For α>0\alpha>0 and ii\in\mathbb{N}, define

uα(i)=𝐄imineαζmin,u_{\alpha}(i)=\mathbf{E}_{i}^{\text{min}}e^{-\alpha\zeta^{\text{min}}},

where 𝐄imin\mathbf{E}^{\text{min}}_{i} stands for the expectation of XminX^{\text{min}} starting from ii and ζmin\zeta^{\text{min}} is the lifetime of XminX^{\text{min}}. Denote by (Rαmin)α>0(R^{\text{min}}_{\alpha})_{\alpha>0} the resolvent of the minimal QQ-process. Set

Φij(α):=Rαmin(i,{j}),α>0,i,j.\Phi_{ij}(\alpha):=R^{\text{min}}_{\alpha}(i,\{j\}),\quad\alpha>0,i,j\in\mathbb{N}.

Let ν\nu be a positive measure on \mathbb{N} and let γ,β0\gamma,\beta\geq 0 be two constants. Set |ν|:=k1νk|\nu|:=\sum_{k\geq 1}\nu_{k}. When both

k0νk(j=k(cj+1cj)i=0jμi)<,|ν|+β0,\sum_{k\geq 0}\nu_{k}\left(\sum_{j=k}^{\infty}(c_{j+1}-c_{j})\sum_{i=0}^{j}\mu_{i}\right)<\infty,\quad|\nu|+\beta\neq 0, (2.4)

and

β=0,if  is an exit\beta=0,\quad\text{if }\infty\text{ is an exit} (2.5)

are satisfied, define, for α>0\alpha>0 and i,ji,j\in\mathbb{N},

Ψij(α):=Φij(α)+uα(i)k0νkΦkj(α)+βμjuα(j)γ+k0νk(1uα(k))+βαk0μkuα(k).\Psi_{ij}(\alpha):=\Phi_{ij}(\alpha)+u_{\alpha}(i)\frac{\sum_{k\geq 0}\nu_{k}\Phi_{kj}(\alpha)+\beta\mu_{j}u_{\alpha}(j)}{\gamma+\sum_{k\geq 0}\nu_{k}(1-u_{\alpha}(k))+\beta\alpha\sum_{k\geq 0}\mu_{k}u_{\alpha}(k)}. (2.6)

The matrix (Ψij(α))i,j(\Psi_{ij}(\alpha))_{i,j\in\mathbb{N}} is called the (Q,γ,β,ν)(Q,\gamma,\beta,\nu)-resolvent matrix. For a constant M>0M>0, the (Q,Mγ,Mβ,Mν)(Q,M\gamma,M\beta,M\nu)-resolvent matrix is obviously the same as the (Q,γ,β,ν)(Q,\gamma,\beta,\nu)-resolvent matrix.

The following theorem is attributed to [26, §7.6]. Note that a QQ-process is called honest if its transition semigroup (pij(t))(p_{ij}(t)) satisfies jpij(t)=1\sum_{j\in\mathbb{N}}p_{ij}(t)=1 for all ii\in\mathbb{N} and t0t\geq 0.

Theorem 2.1.

The transition matrix (pij(t))i,j(p_{ij}(t))_{i,j\in\mathbb{N}} is a QQ-process that is not the minimal one, if and only if there exists a unique (up to a multiplicative positive constant) triple (ν,γ,β)0(\nu,\gamma,\beta)\geq 0 with (2.4) and (2.5) such that the resolvent of (pij(t))i,j(p_{ij}(t))_{i,j\in\mathbb{N}} is given by

Rα(i,{j})=Ψij(α),α>0,i,j,R_{\alpha}(i,\{j\})=\Psi_{ij}(\alpha),\quad\alpha>0,i,j\in\mathbb{N},

where (Ψij(α))i,j(\Psi_{ij}(\alpha))_{i,j\in\mathbb{N}} is the (Q,γ,β,ν)(Q,\gamma,\beta,\nu)-resolvent matrix. Furthermore,

  • (1)

    (pij(t))i,j(p_{ij}(t))_{i,j\in\mathbb{N}} is honest, if and only if γ=0\gamma=0.

  • (2)

    (pij(t))i,j(p_{ij}(t))_{i,j\in\mathbb{N}} is a Doob process, if and only if 0<|ν|<0<|\nu|<\infty and β=0\beta=0.

2.3. Ray-Knight compactification of QQ-processes

In a previous study [18], we obtained a càdlàg modification for each QQ-process XX, denoted by X¯=(X¯t)t0\bar{X}=(\bar{X}_{t})_{t\geq 0}, using the Ray-Knight compactification. The state space of X¯\bar{X} is {}\mathbb{N}\cup\{\infty\}, which corresponds to the Alexandroff compactification of \mathbb{N}. It is important to note that the Ray-Knight compactifications of Doob processes are not normal, whereas those of other QQ-processes are Feller processes on {}\mathbb{N}\cup\{\infty\}; see [18, Corollary 5.2]. If no ambiguity arises, we will not distinguish between the QQ-process and its Ray-Knight compactification. Additionally, we will refer to a non-minimal and non-Doob QQ-process as a Feller QQ-process.

Regarding a Feller QQ-process XX, we can further derive its infinitesimal generator by utilizing the resolvent representation in Theorem 2.1; see [18, Theorem 6.3]. The crucial fact is that every function FF in the domain of the infinitesimal generator satisfies the following boundary condition at \infty:

β2F+()+k(F()F(k))νk+γF()=0,\frac{\beta}{2}F^{+}(\infty)+\sum_{k\in\mathbb{N}}(F(\infty)-F(k))\nu_{k}+\gamma F(\infty)=0, (2.7)

where F():=limkF(k)F(\infty):=\lim_{k\rightarrow\infty}F(k) (if it exists), F+(k):=(F(k+1)F(k))/(ck+1ck)F^{+}(k):=(F(k+1)-F(k))/(c_{k+1}-c_{k}) for kk\in\mathbb{N} and F+():=limkF+(k)F^{+}(\infty):=\lim_{k\rightarrow\infty}F^{+}(k) (if it exists). (When deriving the boundary condition (2.7) in [18], we mistakenly wrote the parameter β2\frac{\beta}{2} as β\beta. This error occurred because the scale function used in [18], specifically (2.2), is half of the scale function in [6]. However, in the proof of [18, Theorem 6.3], when citing the result from [6], particularly the equation above [18, (6.8)], we forgot to multiply by two accordingly. As a result, the parameter β2\frac{\beta}{2} in [18, (6.10)] was incorrectly written as β\beta. Unfortunately, this mistake was not corrected before the publication of [18].)

3. Feller’s Brownian motions

The terminology of Feller’s Brownian motions is borrowed from the renowned article [10] by Itô and McKean. This class of Markov processes was first discovered by Feller in [5], with its detailed definition found in [10, §5]. Recall that a Markov process on [0,)[0,\infty) is called a Feller process if its semigroup (Tt)t0(T_{t})_{t\geq 0} satisfies the following conditions: T0T_{0} is the identify mapping, TtC0([0,))C0([0,))T_{t}C_{0}([0,\infty))\subset C_{0}([0,\infty)) and TtffT_{t}f\rightarrow f in C0([0,))C_{0}([0,\infty)) as t0t\rightarrow 0 for any fC0([0,))f\in C_{0}([0,\infty)), where C0([0,))C_{0}([0,\infty)) is the Banach space consisting of all continuous functions ff on [0,)[0,\infty) such that limxf(x)=0\lim_{x\rightarrow\infty}f(x)=0. We provide a definition for Feller’s Brownian motions in a modern manner as follows.

Definition 3.1.

A Feller process Y:=(Ω,𝒢,𝒢t,Yt,θt,(𝐏x)x[0,))Y:=(\Omega,{\mathscr{G}},{\mathscr{G}}_{t},Y_{t},\theta_{t},(\mathbf{P}_{x})_{x\in[0,\infty)}) with lifetime ζ\zeta on [0,)[0,\infty) is called a Feller’s Brownian motion if its killed process upon hitting 0 is identical in law to the absorbing Brownian motion on (0,)(0,\infty).

Remark 3.2.

This definition excludes the special cases discussed in [10, §6]. The ‘Brownian motions’ in these special cases do not satisfy the normal property at 0, i.e., 𝐏0(Y0=0)1\mathbf{P}_{0}(Y_{0}=0)\neq 1, or the quasi-left-continuity at the first time reaching 0, i.e., for x>0x>0, 𝐏x(Ylimnσn=0)<1\mathbf{P}_{x}(Y_{\lim_{n\rightarrow\infty}\sigma_{n}}=0)<1, where σn:=inf{t>0:Yt=εn}\sigma_{n}:=\inf\{t>0:Y_{t}=\varepsilon_{n}\} for any sequence εn0\varepsilon_{n}\downarrow 0.

The case 4a. in [10, §6] warrants special mention. Similar to Doob processes in the context of QQ-processes, it corresponds to the piecing out of the absorbing Brownian motion Y0Y^{0} with lifetime ζ0\zeta^{0} on (0,)(0,\infty) with respect to a certain probability measure λ\lambda on (0,){}(0,\infty)\cup\{\partial\}. The process obtained through this piecing out is a right process on (0,)(0,\infty), but it is not a Feller process. It can, however, be extended to a Ray process Y¯\bar{Y} on [0,)[0,\infty), where the initial transition function at 0 is given by T¯0(0,):=𝐏¯0(Y¯0)=λ()\bar{T}_{0}(0,\cdot):=\bar{\mathbf{P}}_{0}(\bar{Y}_{0}\in\cdot)=\lambda(\cdot). Analogous to [18, Theorem 5.1], the following expression for the resolvent of Y¯\bar{Y} can be easily obtained:

G¯αh(x)=Gα0h(x)+𝐄x(eαζ0)(0,)Gα0h(x)λ(dx)1𝐄λ(eαζ0),h+((0,)),\bar{G}_{\alpha}h(x)=G^{0}_{\alpha}h(x)+\mathbf{E}_{x}\left(e^{-\alpha\zeta^{0}}\right)\cdot\frac{\int_{(0,\infty)}G^{0}_{\alpha}h(x)\lambda(dx)}{1-\mathbf{E}_{\lambda}\left(e^{-\alpha\zeta^{0}}\right)},\quad\forall h\in\mathcal{B}_{+}((0,\infty)), (3.1)

where Gα0G^{0}_{\alpha} is the resolvent of Y0Y^{0} and 𝐄λ(eαζ0):=(0,)𝐄x(eαζ0)λ(dx)\mathbf{E}_{\lambda}\left(e^{-\alpha\zeta^{0}}\right):=\int_{(0,\infty)}\mathbf{E}_{x}\left(e^{-\alpha\zeta^{0}}\right)\lambda(dx).

Indeed, the reflecting Brownian motion on [0,)[0,\infty) is a classic example of a Feller’s Brownian motion. However, it’s important to note that not all Feller’s Brownian motions exhibit a.s. continuous sample paths. Some Feller’s Brownian motions may experience jumps into (0,)(0,\infty) from 0 or just before reaching 0.

3.1. Infinitesimal generators

It is well known that a function fC0([0,))f\in C_{0}([0,\infty)) is said to belong to the domain 𝒟()\mathcal{D}(\mathscr{L}) of the infinitesimal generator of YY if the limit

f:=limt0Ttfft\mathscr{L}f:=\lim_{t\downarrow 0}\frac{T_{t}f-f}{t}

exists in C0([0,))C_{0}([0,\infty)). The operator :𝒟()C0([0,))\mathscr{L}:\mathcal{D}(\mathscr{L})\rightarrow C_{0}([0,\infty)) thus defined is called the infinitesimal generator of the process YY. The following characterization of infinitesimal generators of Feller’s Brownian motions is attributed to Feller [5].

Theorem 3.3.

A Markov process Y=(Yt)t0Y=(Y_{t})_{t\geq 0} on [0,)[0,\infty) is a Feller’s Brownian motion, if and only if it is a Feller process on [0,)[0,\infty) whose infinitesimal generator on C0([0,))C_{0}([0,\infty)) is f=12f′′\mathscr{L}f=\frac{1}{2}f^{\prime\prime} with domain

𝒟()=\displaystyle\mathcal{D}(\mathscr{L})= {fC0([0,))C2([0,)):f′′C0([0,)),\displaystyle\bigg{\{}f\in C_{0}([0,\infty))\cap C^{2}([0,\infty)):f^{\prime\prime}\in C_{0}([0,\infty)), (3.2)
p1f(0)p2f(0)+p32f′′(0)+(0,)[f(0)f(x)]p4(dx)=0},\displaystyle\qquad p_{1}f(0)-p_{2}f^{\prime}(0)+\frac{p_{3}}{2}f^{\prime\prime}(0)+\int_{(0,\infty)}[f(0)-f(x)]p_{4}(dx)=0\bigg{\}},

where p1,p2,p3p_{1},p_{2},p_{3} are non-negative numbers and p4p_{4} is a positive measure on (0,)(0,\infty) such that

p1+p2+p3+(0,)(x1)p4(dx)=1p_{1}+p_{2}+p_{3}+\int_{(0,\infty)}(x\wedge 1)p_{4}(dx)=1 (3.3)

and

|p4|:=p4((0,))=+ifp2=p3=0.|p_{4}|:=p_{4}\left((0,\infty)\right)=+\infty\quad\text{if}\quad p_{2}=p_{3}=0. (3.4)

The parameters p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} are uniquely determined for each Feller’s Brownian motion.

Proof.

For the readers’ convenience, we provide necessary details regarding this proof. The uniqueness of (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) is evident.

The necessity can be established by demonstrating that

𝒟(){fC0([0,))C2([0,)):f′′C0([0,))},\displaystyle\mathcal{D}(\mathscr{L})\subset\{f\in C_{0}([0,\infty))\cap C^{2}([0,\infty)):f^{\prime\prime}\in C_{0}([0,\infty))\}, (3.5)
f=12f′′,f𝒟(),\displaystyle\mathscr{L}f=\frac{1}{2}f^{\prime\prime},\quad\forall f\in\mathcal{D}(\mathscr{L}),

and then applying [21, II§5, Theorem 2]. To achieve this, let (Gα)α>0(G_{\alpha})_{\alpha>0} denote the resolvent of YY. Fix α>0\alpha>0. By the Hille-Yosida theorem, 𝒟()=GαC0([0,))C0([0,))\mathcal{D}(\mathscr{L})=G_{\alpha}C_{0}([0,\infty))\subset C_{0}([0,\infty)), and for f=Gαhf=G_{\alpha}h with hC0([0,))h\in C_{0}([0,\infty)),

f=αGαhhC0([0,)).\mathscr{L}f=\alpha G_{\alpha}h-h\in C_{0}([0,\infty)). (3.6)

Set τ0:=inf{t>0:Yt=0}\tau_{0}:=\inf\{t>0:Y_{t}=0\}. It follows from Definition 3.1 and Dynkin’s formula that

Gαh(x)=Gα0h(x)+Gαh(0)𝐄xeατ0,x>0,G_{\alpha}h(x)=G_{\alpha}^{0}h(x)+G_{\alpha}h(0)\mathbf{E}_{x}e^{-\alpha\tau_{0}},\quad x>0, (3.7)

where Gα0G^{0}_{\alpha} is the resolvent of the absorbing Brownian motion on (0,)(0,\infty). Note that 𝐄xeατ0=e2αx\mathbf{E}_{x}e^{-\alpha\tau_{0}}=e^{-\sqrt{2\alpha}x} (see, e.g., [11, page 26, 5)]), and (see, e.g., [21, II§3, #7 and #8])

Gα0hC2((0,))C0((0,)),12(Gα0h)′′=αGα0hh.G^{0}_{\alpha}h\in C^{2}((0,\infty))\cap C_{0}((0,\infty)),\quad\frac{1}{2}\left(G^{0}_{\alpha}h\right)^{\prime\prime}=\alpha G^{0}_{\alpha}h-h. (3.8)

For any x>0x>0, it holds that

(Gαh)′′(x)\displaystyle\left(G_{\alpha}h\right)^{\prime\prime}(x) =(Gα0h)′′(x)+Gαh(0)2αe2αx\displaystyle=\left(G_{\alpha}^{0}h\right)^{\prime\prime}(x)+G_{\alpha}h(0)\cdot 2\alpha\cdot e^{-\sqrt{2\alpha}x} (3.9)
=2αGα0h(x)2h(x)+2αGαh(0)e2αx,\displaystyle=2\alpha G^{0}_{\alpha}h(x)-2h(x)+2\alpha G_{\alpha}h(0)e^{-\sqrt{2\alpha}x},

and thus

limx0(Gαh)′′(x)=2h(0)+2αGαh(0),limx(Gαh)′′(x)=0.\lim_{x\rightarrow 0}\left(G_{\alpha}h\right)^{\prime\prime}(x)=-2h(0)+2\alpha G_{\alpha}h(0),\quad\lim_{x\rightarrow\infty}\left(G_{\alpha}h\right)^{\prime\prime}(x)=0. (3.10)

Using (3.6), (3.7), (3.9), and (3.10), we conclude that f=GαhC0([0,))C2([0,))f=G_{\alpha}h\in C_{0}([0,\infty))\cap C^{2}([0,\infty)), f′′C0([0,)f^{\prime\prime}\in C_{0}([0,\infty), and f=12f′′\mathscr{L}f=\frac{1}{2}f^{\prime\prime}. In other words, (3.5) is established.

Regarding the sufficiency, we first apply [21, II§5, Theorem 3] to \mathscr{L}, showing that \mathscr{L} with domain (3.2) is indeed the infinitesimal generator of a Feller process YY on [0,)[0,\infty). It remains to demonstrate that the killed process

Yt0:={Yt,t<τ0,,tτ0Y^{0}_{t}:=\left\{\begin{aligned} &Y_{t},\quad&t<\tau_{0},\\ &\partial,\quad&t\geq\tau_{0}\end{aligned}\right. (3.11)

is identical in law to the absorbing Brownian motion on (0,)(0,\infty). In fact, the proof of [21, II§5, Theorem 3] shows that the resolvent kernel of Y0=(Yt0)t0Y^{0}=(Y^{0}_{t})_{t\geq 0} is given by [21, §3, #7] (with p(x)=2xp(x)=2x and mm being the Lebesgue measure). This is precisely the resolvent kernel of absorbing Brownian motion. ∎

Remark 3.4.

When p2=p3=|p4|=0p_{2}=p_{3}=|p_{4}|=0 and p1=1p_{1}=1, the condition (3.4) is not satisfied. However, this case corresponds to the absorbing Brownian motion on (0,)(0,\infty).

The triple (p1,p2,p4)(p_{1},p_{2},p_{4}) plays a role analogous to (γ,β,ν)(\gamma,\beta,\nu) for a QQ-process, with some heuristic explanations for the latter triple presented in [18, §2]. The additional parameter p3p_{3} measures the sojourn of the Feller’s Brownian motion at 0. One one hand, this can be observed by examining the paths of the Feller’s Brownian motion as constructed by Itô and McKean, as illustrated in [10, §15]; see also (3.25). On the other hand, we can examine the symmetric case, where the transition semigroup (Tt)t0(T_{t})_{t\geq 0} of YY is symmetric with respect to some σ\sigma-finite measure mm in the sense that

[0,)Ttf(x)g(x)m(dx)=[0,)f(x)Ttg(x)m(dx),t0,f,gb([0,)),\int_{[0,\infty)}T_{t}f(x)g(x)m(dx)=\int_{[0,\infty)}f(x)T_{t}g(x)m(dx),\quad\forall t\geq 0,f,g\in\mathcal{B}_{b}([0,\infty)),

as stated in the following corollary. In this special case, the parameter p3/(2p2)p_{3}/(2p_{2}) represents the mass of the symmetric measure mm at 0. It is well known that when this symmetric measure, also known as the speed measure of YY, is larger in a specific region, the motion of YtY_{t} will be slower when passing through that region.

Corollary 3.5.

Let Y=(Yt)t0Y=(Y_{t})_{t\geq 0} be a Feller’s Brownian motion on [0,)[0,\infty) with parameters p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} as described in Theorem 3.3. The process YY is symmetric if and only if p2>0,|p4|=0p_{2}>0,|p_{4}|=0. Furthermore, the symmetric measure must be

m(dx)=p32p2δ0(dx)+1(0,)(x)dxm(dx)=\frac{p_{3}}{2p_{2}}\delta_{0}(dx)+1_{(0,\infty)}(x)dx (3.12)

up to a multiplicative constant, where δ0\delta_{0} denotes the Dirac measure at 0, and the Dirichlet form associated with this symmetric Feller’s Brownian motion on L2([0,),m)L^{2}([0,\infty),m) is

\displaystyle{\mathscr{F}} =H1([0,)),\displaystyle=H^{1}([0,\infty)), (3.13)
(f,g)\displaystyle{\mathscr{E}}(f,g) =120f(x)g(x)𝑑x+p12p2f(0)g(0),f,g,\displaystyle=\frac{1}{2}\int_{0}^{\infty}f^{\prime}(x)g^{\prime}(x)dx+\frac{p_{1}}{2p_{2}}f(0)g(0),\quad f,g\in{\mathscr{F}},

where

H1([0,)):={fL2([0,)):f is absolutely continuous and fL2([0,))}.H^{1}([0,\infty)):=\{f\in L^{2}([0,\infty)):f\text{ is absolutely continuous and }f^{\prime}\in L^{2}([0,\infty))\}.
Proof.

We continue to denote by GαG_{\alpha} the resolvent of YY and by Y0Y^{0} the killed process of YY upon hitting 0. Let gα0(x,y)g^{0}_{\alpha}(x,y) represent the resolvent density of Y0Y^{0}, i.e., Gα0(x,dy)=gα0(x,y)dyG^{0}_{\alpha}(x,dy)=g^{0}_{\alpha}(x,y)dy. We have

gα0(x,y)={W1u(x)u+(y),xy,W1u(y)u+(x),yx,g_{\alpha}^{0}(x,y)=\left\{\begin{aligned} &W^{-1}u_{-}(x)u_{+}(y),\quad&x\leq y,\\ &W^{-1}u_{-}(y)u_{+}(x),\quad&y\leq x,\end{aligned}\right. (3.14)

where u(x)=sinh(2αx),u+(x)=e2αxu_{-}(x)=\sinh(\sqrt{2\alpha}x),u_{+}(x)=e^{-\sqrt{2\alpha}x}, and W:=W(u,u+)=2α2W:=W(u_{-},u_{+})=\frac{\sqrt{2\alpha}}{2} is the Wronskain of u,u+u_{-},u_{+} (with W1:=1/WW^{-1}:=1/W); see, e.g., [21, II§3, #7]. As established in [10, §15, 4.], for any hC0([0,))h\in C_{0}([0,\infty)), it holds that

Gαh(0)=2p2(0,)e2αxh(x)𝑑x+p3h(0)+(0,)Gα0h(x)p4(dx)p1+2αp2+αp3+(0,)[1e2αx]p4(dx).G_{\alpha}h(0)=\frac{2p_{2}\int_{(0,\infty)}e^{-\sqrt{2\alpha}x}h(x)dx+p_{3}h(0)+\int_{(0,\infty)}G^{0}_{\alpha}h(x)p_{4}(dx)}{p_{1}+\sqrt{2\alpha}p_{2}+\alpha p_{3}+\int_{(0,\infty)}[1-e^{-\sqrt{2\alpha}x}]p_{4}(dx)}. (3.15)

To demonstrate the necessity, let mm be a symmetric measure of YY. From [8, Lemma 4.1.3], it follows that m|(0,)m|_{(0,\infty)} is the symmetric measure of Y0Y^{0}. Hence, without loss of generality, we may assume that m|(0,)m|_{(0,\infty)} is the Lebesgue measure on (0,)(0,\infty). For any h1,h2C0((0,))h_{1},h_{2}\in C_{0}((0,\infty)), substituting (3.7) and (3.15) into

[0,)Gαh1(x)h2(x)m(dx)=[0,)h1(x)Gαh2(x)m(dx),\int_{[0,\infty)}G_{\alpha}h_{1}(x)h_{2}(x)m(dx)=\int_{[0,\infty)}h_{1}(x)G_{\alpha}h_{2}(x)m(dx), (3.16)

we obtain

(0,)×(0,)h1(x)h2(y)e2αxm(dx)(p4Gα0)(dy)\displaystyle\iint_{(0,\infty)\times(0,\infty)}h_{1}(x)h_{2}(y)e^{-\sqrt{2\alpha}x}m(dx)\left(p_{4}G^{0}_{\alpha}\right)(dy)
=(0,)×(0,)h1(y)h2(x)e2αxm(dx)(p4Gα0)(dy),h1,h2C0((0,)),\displaystyle\quad=\iint_{(0,\infty)\times(0,\infty)}h_{1}(y)h_{2}(x)e^{-\sqrt{2\alpha}x}m(dx)\left(p_{4}G^{0}_{\alpha}\right)(dy),\quad\forall h_{1},h_{2}\in C_{0}((0,\infty)),

where (p4Gα0)(dy)=(0,)p4(dx)Gα0(x,dy)=((0,)gα0(x,y)p4(dx))dy\left(p_{4}G^{0}_{\alpha}\right)(dy)=\int_{(0,\infty)}p_{4}(dx)G_{\alpha}^{0}(x,dy)=\left(\int_{(0,\infty)}g^{0}_{\alpha}(x,y)p_{4}(dx)\right)dy. This implies that

H(y):=e2αy(0,)gα0(x,y)p4(dx)H(y):=e^{\sqrt{2\alpha}y}\int_{(0,\infty)}g^{0}_{\alpha}(x,y)p_{4}(dx)

is constant. Substituting (3.14) into H(y)H(y), we have

WH(y)=(0,)Hy(x)p4(dx),W\cdot H(y)=\int_{(0,\infty)}H_{y}(x)p_{4}(dx),

where

Hy(x)={sinh(2αx),xy,sinh(2αy)e2α(xy),yx.H_{y}(x)=\left\{\begin{aligned} &\sinh(\sqrt{2\alpha}x),\quad&x\leq y,\\ &\sinh(\sqrt{2\alpha}y)e^{-\sqrt{2\alpha}(x-y)},&y\leq x.\end{aligned}\right.

It is straightforward to verify that Hy(x)H_{y}(x) is increasing in yy. Letting y0y\downarrow 0, we conclude that H(y)0H(y)\equiv 0. Since HyH_{y} is strictly positive for y>0y>0, it must hold that |p4|=0|p_{4}|=0. Substituting (3.7) and (3.15) into (3.16) again, but now taking h1,h2C0([0,))h_{1},h_{2}\in C_{0}([0,\infty)), we further obtain

(p32p2m({0}))(h1(0)0h2(x)e2αx𝑑xh2(0)0h1(x)e2αx𝑑x)=0.\left(p_{3}-2p_{2}m(\{0\})\right)\cdot\left(h_{1}(0)\int_{0}^{\infty}h_{2}(x)e^{-\sqrt{2\alpha}x}dx-h_{2}(0)\int_{0}^{\infty}h_{1}(x)e^{-\sqrt{2\alpha}x}dx\right)=0.

This clearly implies

p3=2p2m({0}).p_{3}=2p_{2}\cdot m(\{0\}).

Particularly, if p2=0p_{2}=0, then p3=0p_{3}=0. This contradicts the condition (3.4). Therefore, we conclude that p2>0,|p4|=0p_{2}>0,|p_{4}|=0, and the symmetric measure is precisely (3.12).

Next, we consider the case where p2>0p_{2}>0 and |p4|=0|p_{4}|=0. Note that (3.13) is, in fact, a regular Dirichlet form on L2([0,),m)L^{2}([0,\infty),m); see, e.g., [7, §5.3]. It suffices to show that

Gαh,α(Gαh,g)=(h,g)m,h,gCc([0,)).G_{\alpha}h\in{\mathscr{F}},\quad{\mathscr{E}}_{\alpha}(G_{\alpha}h,g)=(h,g)_{m},\quad\forall h,g\in C_{c}^{\infty}([0,\infty)).

Indeed, Gα0hH01((0,))={fH1([0,)):f(0)=0}G_{\alpha}^{0}h\in H^{1}_{0}((0,\infty))=\{f\in H^{1}([0,\infty)):f(0)=0\}, and hence (3.7) implies that GαhG_{\alpha}h\in{\mathscr{F}}. Using (3.7), we further have

α(Gαh,g)\displaystyle{\mathscr{E}}_{\alpha}(G_{\alpha}h,g) =120(Gα0h)(x)g(x)𝑑x+Gαh(0)20(e2αx)g(x)𝑑x\displaystyle=\frac{1}{2}\int_{0}^{\infty}\left(G^{0}_{\alpha}h\right)^{\prime}(x)g^{\prime}(x)dx+\frac{G_{\alpha}h(0)}{2}\int_{0}^{\infty}\left(e^{-\sqrt{2\alpha}x}\right)^{\prime}g^{\prime}(x)dx (3.17)
+p12p2Gαh(0)g(0)+α0Gα0h(x)g(x)𝑑x\displaystyle\quad+\frac{p_{1}}{2p_{2}}G_{\alpha}h(0)g(0)+\alpha\int_{0}^{\infty}G_{\alpha}^{0}h(x)g(x)dx
+αGαh(0)0e2αxg(x)𝑑x+αp32p2Gαh(0)g(0).\displaystyle\quad+\alpha G_{\alpha}h(0)\int_{0}^{\infty}e^{-\sqrt{2\alpha}x}g(x)dx+\frac{\alpha p_{3}}{2p_{2}}G_{\alpha}h(0)g(0).

Note that (see, e.g., [21, II§4, (23)])

W(Gα0h)(x)=u+(x)0xu(y)h(y)𝑑y+u(x)xu+(y)h(y)𝑑y.W\cdot\left(G^{0}_{\alpha}h\right)^{\prime}(x)=u^{\prime}_{+}(x)\int_{0}^{x}u_{-}(y)h(y)dy+u^{\prime}_{-}(x)\int_{x}^{\infty}u_{+}(y)h(y)dy.

This yields that

limx(Gα0h)(x)=0,(Gα0h)(0)=20h(y)e2αy𝑑y.\lim_{x\rightarrow\infty}\left(G^{0}_{\alpha}h\right)^{\prime}(x)=0,\quad\left(G^{0}_{\alpha}h\right)^{\prime}(0)=2\int_{0}^{\infty}h(y)e^{-\sqrt{2\alpha}y}dy. (3.18)

Substituting (3.8), (3.15), and (3.18) into (3.17), we can obtain that

α(Gαh,g)=0h(x)g(x)𝑑x+p32p2h(0)g(0)=(h,g)m.{\mathscr{E}}_{\alpha}(G_{\alpha}h,g)=\int_{0}^{\infty}h(x)g(x)dx+\frac{p_{3}}{2p_{2}}h(0)g(0)=(h,g)_{m}.

This completes the proof. ∎

It is worth noting that the L2L^{2}-generator of symmetric Feller’s Brownian motion, i.e., the L2L^{2}-generator of the Dirichlet form (3.13), is derived in, e.g., [7, Theorem 5.3].

3.2. Pathwise construction

The sample paths of a Feller’s Brownian motion were constructed by Itô and McKean in [10]. The symmetric cases characterized in Corollary 3.5 correspond to either reflecting Brownian motion (for p1=0p_{1}=0) or elastic Brownian motion (for p1>0p_{1}>0), up to a time change transformation induced by the speed measure (3.12). Their pathwise constructions are quite clear; see also [10, §10]. For the readers’ convenience, we will restate some necessary details of the pathwise construction for non-symmetric cases in this subsection.

3.2.1. p2=0,p3>0,|p4|<p_{2}=0,p_{3}>0,|p_{4}|<\infty

This case was examined in [10, §9]. The sample paths of Feller’s Brownian motion can be constructed as follows: Given a reflecting Brownian motion with sample paths tWt+t\mapsto W^{+}_{t} starting at a point x[0,)x\in[0,\infty), let Yt:=Wt+Y_{t}:=W^{+}_{t} up to the first hitting time τ0+:=inf{t>0:Wt+=0}\tau^{+}_{0}:=\inf\{t>0:W^{+}_{t}=0\} of 0. Then, make YY wait at 0 for an exponential holding time 𝔢\mathfrak{e} with the conditional law

𝐏x(𝔢>t|W+)=ep1+|p4|p3t;\mathbf{P}_{x}(\mathfrak{e}>t|W^{+})=e^{-\frac{p_{1}+|p_{4}|}{p_{3}}t}; (3.19)

at the end of this time (only applicable in the case where p1+|p4|>0p_{1}+|p_{4}|>0), let it jump to a point in (0,){}(0,\infty)\cup\{\partial\} according to the distribution λ\lambda given by

λ(dx)|(0,):=p4(dx)p1+|p4|,λ({}):=p1p1+|p4|.\lambda(dx)|_{(0,\infty)}:=\frac{p_{4}(dx)}{p_{1}+|p_{4}|},\quad\lambda(\{\partial\}):=\frac{p_{1}}{p_{1}+|p_{4}|}. (3.20)

If the reaching point is in (0,)(0,\infty), let it start afresh; if it jumps to \partial, let Yt:=Y_{t}:=\partial at all later times.

3.2.2. p2>0,0<|p4|<p_{2}>0,0<|p_{4}|<\infty

This case was addressed in [10, §12] using an increasing Lévy process, defined by (3.21). In fact, we can provide an alternative construction via the Ikeda-Nagasawa-Watanabe piecing out procedure (see [9]), similar to the approach in [18, Theorem 8.1] for QQ-processes with |ν|<+|\nu|<+\infty; see also [21, III§4].

To be precise, let us begin with a symmetric Feller’s Brownian motion Y1=(Yt1)t0Y^{1}=(Y^{1}_{t})_{t\geq 0} with parameters (p1+|p4|,p2,p3,0)(p_{1}+|p_{4}|,p_{2},p_{3},0). Let λ\lambda be the probability measure on (0,){}(0,\infty)\cup\{\partial\} defined as in (3.20). Then the Feller’s Brownian motion Y=(Yt)t0Y=(Y_{t})_{t\geq 0} with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) such that p2>0,0<|p4|<p_{2}>0,0<|p_{4}|<\infty is actually the piecing out of Y1Y^{1} with respect to λ\lambda, as described in [18, Appendix A]. Intuitively, YY repeatedly splices resurrection paths at the death times of Yt1Y^{1}_{t}, with the resurrection points randomly determined by the distribution λ\lambda.

3.2.3. |p4|=|p_{4}|=\infty

When |p4|=|p_{4}|=\infty, the pathwise construction becomes significantly more challenging. This was accomplished by Itô and McKean in [10, §12-§15].

We first consider the case where p1=p3=0p_{1}=p_{3}=0 and |p4|=+|p_{4}|=+\infty (p20)p_{2}\geq 0). Let Z=(Z(t))t0Z=(Z(t))_{t\geq 0} be a subordinator, i.e., an increasing Lévy process on \mathbb{R} with Z(0)=0Z(0)=0, whose distribution has the Laplace transform

𝐄exZ(t)=exp{t[p2x+(0,)(1exy)p4(dy)]}.\mathbf{E}e^{-xZ(t)}=\exp\left\{-t\left[p_{2}x+\int_{(0,\infty)}(1-e^{-xy})p_{4}(dy)\right]\right\}. (3.21)

In [10], this subordinator is referred to as an increasing differential process. For every s0s\geq 0, define

Z1(s):=inf{t>0:Z(t)>s}.Z^{-1}(s):=\inf\{t>0:Z(t)>s\}.

Further, let W+=(Wt+)t0W^{+}=(W^{+}_{t})_{t\geq 0} be a reflecting Brownian motion on [0,)[0,\infty), independent of ZZ, with local time (t)t0(\ell_{t})_{t\geq 0} at 0 given by

t=limε012ε0t1{Ws+<ε}𝑑s,t0.\ell_{t}=\lim_{\varepsilon\downarrow 0}\frac{1}{2\varepsilon}\int_{0}^{t}1_{\{W^{+}_{s}<\varepsilon\}}ds,\quad\forall t\geq 0. (3.22)

Note that the Revuz measure of \ell with respect to W+W^{+} is δ02\frac{\delta_{0}}{2}; see, e.g., [22, X. Proposition 2.4]. It was shown in [10, §13] that

Yt1:=Z(Z1(t))t+Wt+,t0Y^{1}_{t}:=Z(Z^{-1}(\ell_{t}))-\ell_{t}+W^{+}_{t},\quad t\geq 0 (3.23)

is indeed a Feller’s Brownian motion with parameters (0,p2,0,p4)(0,p_{2},0,p_{4}). For a detailed explanation of the sample paths defined by (3.23), please refer to [10, §12]. Note that this pathwise construction (3.23) also applies to the case p1=p3=0,p2>0p_{1}=p_{3}=0,p_{2}>0, and |p4|<+|p_{4}|<+\infty.

Regarding the general case, we note that

tY1:=Z1(t),t0\ell^{Y^{1}}_{t}:=Z^{-1}(\ell_{t}),\quad t\geq 0 (3.24)

is the local time of the process defined by (3.23) at 0, as established in [10, §14]. Define 𝔣(t):=t+p3tY1\mathfrak{f}(t):=t+p_{3}\ell^{Y^{1}}_{t} and 𝔣1(t):=inf{s>0:𝔣(s)>t}\mathfrak{f}^{-1}(t):=\inf\{s>0:\mathfrak{f}(s)>t\}. Then, the time-changed process of (3.23) with respect to 𝔣\mathfrak{f} is

Yt2:=Y𝔣1(t)1,t0.Y^{2}_{t}:=Y^{1}_{\mathfrak{f}^{-1}(t)},\quad t\geq 0. (3.25)

This process Y2Y^{2} is a Feller’s Brownian motion with parameters (0,p2,p3,p4)(0,p_{2},p_{3},p_{4}). The desired Feller’s Brownian motion YY with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) is obtained as the subprocess of (3.25) perturbed by the multiplicative functional

Mt:=ep1𝔣1(t)Y1,t0.M_{t}:=e^{-p_{1}\ell^{Y^{1}}_{\mathfrak{f}^{-1}(t)}},\quad t\geq 0. (3.26)

A rigorous construction of the subprocess can be found in, e.g., [1, III, §3]. Roughly speaking, one can take a random time ζ\zeta such that

𝐏x(ζ>t|Y1)=ep1𝔣1(t)Y1,t>0,\mathbf{P}_{x}(\zeta>t|Y^{1})=e^{-p_{1}\ell^{Y^{1}}_{\mathfrak{f}^{-1}(t)}},\quad\forall t>0, (3.27)

and then kill the process Y2Y^{2} at time ζ\zeta.

3.3. Local times of Feller’s Brownian motion

In [10, §14], Itô and McKean examined the local time (3.24) of the special Feller’s Brownian motion defined by (3.23) at 0. What we focus on here is the local time of a general Feller’s Brownian motion at a given point a>0a>0.

Consider a Feller’s Brownian motion on [0,)[0,\infty):

Y=(Ω,𝒢,𝒢t,Yt,θt,(𝐏x)x[0,))Y=(\Omega,{\mathscr{G}},{\mathscr{G}}_{t},Y_{t},\theta_{t},(\mathbf{P}_{x})_{x\in[0,\infty)})

with lifetime ζ\zeta, where (𝒢t)t0({\mathscr{G}}_{t})_{t\geq 0} is the augmented natural filtration on Ω\Omega. Let 𝒢:=t0𝒢t{\mathscr{G}}_{\infty}:=\bigvee_{t\geq 0}{\mathscr{G}}_{t} denote the σ\sigma-algebra generated by t0𝒢t\bigcup_{t\geq 0}{\mathscr{G}}_{t}. We adjoin to Ω\Omega the dead path [][\partial] with Yt([]):=Y_{t}([\partial]):=\partial for all t0t\geq 0. A positive continuous additive functional of YY is defined as follows. For detailed discussions, see, e.g., [1, IV§1] and [3, Definition A.3.1].

Definition 3.6.

A family A=(At)t0A=(A_{t})_{t\geq 0} of functions from Ω\Omega to [0,][0,\infty] is called a positive continuous additive functional (PCAF for short) of YY if there exists Λ𝒢\Lambda\in{\mathscr{G}}_{\infty} such that

𝐏x(Λ)=1 for x[0,) and θtΛΛ for t0,\mathbf{P}_{x}(\Lambda)=1\text{ for }x\in[0,\infty)\quad\text{ and }\quad\theta_{t}\Lambda\subset\Lambda\text{ for }t\geq 0,

and the following conditions are satisfied:

  • (A1)

    For each t0t\geq 0, At|Λ𝒢t|ΛA_{t}|_{\Lambda}\in{\mathscr{G}}_{t}|_{\Lambda}.

  • (A2)

    For every ωΛ\omega\in\Lambda, A(ω)A_{\cdot}(\omega) is continuous on [0,)[0,\infty), A0(ω)=0A_{0}(\omega)=0, At(ω)<A_{t}(\omega)<\infty for t<ζ(ω)t<\zeta(\omega), and At(ω)=Aζ(ω)(ω)A_{t}(\omega)=A_{\zeta(\omega)}(\omega) for tζ(ω)t\geq\zeta(\omega).

  • (A3)

    For ωΛ\omega\in\Lambda and every t,s0t,s\geq 0, At+s(ω)=At(ω)+As(θtω)A_{t+s}(\omega)=A_{t}(\omega)+A_{s}(\theta_{t}\omega).

The set Λ\Lambda is called a defining set of AA. We further make the convention At([])=0A_{t}([\partial])=0 for all t0t\geq 0.

The fine support of a PCAF (At)t0(A_{t})_{t\geq 0} is defined as

Supp(A):={x[0,):𝐏x(R=0)=1},\text{Supp}(A):=\{x\in[0,\infty):\mathbf{P}_{x}(R=0)=1\},

where R(ω):=inf{t>0:At(ω)>0}R(\omega):=\inf\{t>0:A_{t}(\omega)>0\}. According to [1, V. Theorem 3.13], for a given a>0a>0, there exists a PCAF (unique up to a multiplicative constant) of YY with fine support {a}\{a\}.

Recall that τ0=inf{t>0:Yt=0}\tau_{0}=\inf\{t>0:Y_{t}=0\}. The killed process Y0Y^{0} can be written as

Y0=(Ω,𝒢,𝒢t,Yt0,θt0,(𝐏x)x(0,)),Y^{0}=\left(\Omega,{\mathscr{G}},{\mathscr{G}}_{t},Y^{0}_{t},\theta^{0}_{t},(\mathbf{P}_{x})_{x\in(0,\infty)}\right),

where Yt0Y^{0}_{t} is defined as in (3.11), θt0(ω):=θt(ω)\theta^{0}_{t}(\omega):=\theta_{t}(\omega) for t<τ0(ω)t<\tau_{0}(\omega) and θt0(ω):=[]\theta^{0}_{t}(\omega):=[\partial] for tτ0(ω)t\geq\tau_{0}(\omega); see, e.g., [25, (12.21i)]. The lifetime of Y0Y^{0} is ζ0:=ζτ0(=τ0)\zeta^{0}:=\zeta\wedge\tau_{0}(=\tau_{0}). We can define the PCAFs for Y0Y^{0} and their fine supports analogously. The following fact is crucial to our discussion.

Lemma 3.7.

Given a>0a>0, let a=(ta)t0\ell^{a}=(\ell^{a}_{t})_{t\geq 0} be a PCAF of YY with fine support {a}\{a\}. Then

ta,0:={ta,t<ζ0,ζ0a,tζ0\ell^{a,0}_{t}:=\left\{\begin{aligned} &\ell^{a}_{t},\quad&t<\zeta^{0},\\ &\ell^{a}_{\zeta^{0}},\quad&t\geq\zeta^{0}\end{aligned}\right.

is a PCAF of Y0Y^{0} with fine support {a}\{a\}.

Proof.

Assuming without loss of generality that the defining set of a\ell^{a} is Ω\Omega, it is straightforward to verify that a,0\ell^{a,0} satisfies properties (A1) and (A2). To establish property (A3) for a,0\ell^{a,0}, it suffices to consider the case where t+sτ0(ω)t+s\geq\tau_{0}(\omega). If tτ0(ω)t\geq\tau_{0}(\omega), then ta,0(ω)=τ0(ω)a(ω)\ell^{a,0}_{t}(\omega)=\ell^{a}_{\tau_{0}(\omega)}(\omega) and sa,0(θt0ω)=sa,0([])=0a([])=0\ell^{a,0}_{s}(\theta^{0}_{t}\omega)=\ell^{a,0}_{s}([\partial])=\ell^{a}_{0}([\partial])=0. Hence, we have

t+sa,0(ω)=ta,0(ω)+sa,0(θt0ω).\ell^{a,0}_{t+s}(\omega)=\ell^{a,0}_{t}(\omega)+\ell^{a,0}_{s}(\theta^{0}_{t}\omega). (3.28)

If t<τ0(ω)t<\tau_{0}(\omega), (3.28) can be verified using τ0(θtω)=τ0(ω)ts\tau_{0}(\theta_{t}\omega)=\tau_{0}(\omega)-t\leq s. Thus, a,0\ell^{a,0} is indeed a PCAF of Y0Y^{0}. The fine support of a,0\ell^{a,0} is {a}\{a\} because ta,0ta\ell^{a,0}_{t}\leq\ell^{a}_{t} and 𝐏a(τ0>0)=1\mathbf{P}_{a}(\tau_{0}>0)=1. ∎

According to Definition 3.1, Y0Y^{0} is identical in law to the absorbing Brownian motion on (0,)(0,\infty). Therefore, we can define the Revuz measure νa\nu_{a} of a,0\ell^{a,0} with respect to Y0Y^{0} as follows:

(0,)f(x)νa(dx)\displaystyle\int_{(0,\infty)}f(x)\nu_{a}(dx) =limt01t(0,)(𝐄x[0tf(Yt0)𝑑ta,0])𝑑x\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\int_{(0,\infty)}\left(\mathbf{E}_{x}\left[\int_{0}^{t}f(Y^{0}_{t})d\ell^{a,0}_{t}\right]\right)dx
=limt01t(0,)(𝐄x[0tτ0f(Yt)𝑑ta])𝑑x.\displaystyle=\lim_{t\downarrow 0}\frac{1}{t}\int_{(0,\infty)}\left(\mathbf{E}_{x}\left[\int_{0}^{t\wedge\tau_{0}}f(Y_{t})d\ell^{a}_{t}\right]\right)dx.

Note that νa\nu_{a} is a constant multiple of δa\delta_{a}. Unlike the approach of normalizing local times by the values of their potentials as described in [1, V. Theorem 3.13], we opt to use the unique PCAF with fine support {a}\{a\} in the following sense.

Definition 3.8.

Given a>0a>0, a PCAF La=(Lta)t0L^{a}=(L^{a}_{t})_{t\geq 0} with Supp(La)={a}\text{Supp}(L^{a})=\{a\} is called the local time of YY at aa if the Revuz measure of La,0:=(Ltζ0a)t0L^{a,0}:=(L^{a}_{t\wedge\zeta^{0}})_{t\geq 0} with respect to Y0Y^{0} is δa\delta_{a}.

In the symmetric case, where p2>0p_{2}>0 and |p4|=0|p_{4}|=0, we can also define the Revuz measure of the local time LaL^{a} with respect to YY and the symmetric measure mm. According to [3, Proposition 4.1.10], this Revuz measure is also equal to the Dirac measure δa\delta_{a}. Therefore, the definition of local time provided here is consistent with the definition of local time in the theory of Dirichlet forms.

4. Time-changed Feller’s Brownian motions are birth-death processes

Consider the birth-death density matrix (2.1) and from now on, assume that \infty is regular for the minimal QQ-process. Particularly, the scale function (ck)k0(c_{k})_{k\geq 0} given by (2.2) satisfies

c=limkck<,c_{\infty}=\lim_{k\rightarrow\infty}c_{k}<\infty,

and the speed measure μ\mu, as defined in (2.3), is finite; see, e.g., [18, Remark 3.4].

The aim of this section is to demonstrate that any Feller’s Brownian motion can be converted into a QQ -process through a time change transformation and a spatial homeomorphism. The special cases of absorbing and reflecting Brownian motions have been analyzed in [18, §3]. In these cases, the transformed QQ-processes are the minimal QQ-process and the (Q,1)(Q,1)-process, respectively.

4.1. Spatial transformation

The formulation we will present encounters a significant issue because the boundary point 0 of Feller’s Brownian motion and the boundary point \infty of the QQ-process are located at opposite ends of their respective state spaces. Additionally, unlike Feller’s Brownian motion, the QQ-process is not on the natural scale. However, both issues can be resolved by applying a straightforward spatial transformation to the QQ-process. To address this, define c^n:=ccn\hat{c}_{n}:=c_{\infty}-c_{n} for nn\in\mathbb{N}. Let

E:={c^n:n},E¯=E{0}E:=\{\hat{c}_{n}:n\in\mathbb{N}\},\quad\overline{E}=E\cup\{0\} (4.1)

and

Ξ:E,c^nn.\Xi:E\rightarrow\mathbb{N},\quad\hat{c}_{n}\mapsto n. (4.2)

Clearly, Ξ\Xi can be extended to a homeomorphism between E¯\overline{E} and {}\mathbb{N}\cup\{\infty\}, where E¯\overline{E} is endowed with the relative topology of {\mathbb{R}}.

For each QQ-process X=(Xt)t0X=(X_{t})_{t\geq 0} on {}\mathbb{N}\cup\{\infty\},

X^t:=Ξ1(Xt),t0\hat{X}_{t}:=\Xi^{-1}(X_{t}),\quad t\geq 0

defines a continuous-time Markov chain on E¯\overline{E}. It is a Feller process on E¯\overline{E} whenever XX is a Feller QQ-process. For convenience, we also refer to X^=(X^t)t0\hat{X}=(\hat{X}_{t})_{t\geq 0} as a Doob process or a Feller QQ-process (on EE or E¯\overline{E}).

4.2. Time change

We first prepare the ingredient, specifically the PCAF, for the time change transformation on a Feller’s Brownian motion YY. For c^nE\hat{c}_{n}\in E, let Lc^n=(Ltc^n)t0L^{\hat{c}_{n}}=(L^{\hat{c}_{n}}_{t})_{t\geq 0} denote the local time of YY at c^n\hat{c}_{n} as defined in Definition 3.8. Define

At=nμnLtc^n,t0.A_{t}=\sum_{n\in\mathbb{N}}\mu_{n}L^{\hat{c}_{n}}_{t},\quad t\geq 0. (4.3)

We will show that A=(At)t0A=(A_{t})_{t\geq 0} is a PCAF of YY.

Lemma 4.1.

The family of functions A=(At)t0A=(A_{t})_{t\geq 0} defined as (4.3) is a PCAF of the Feller’s Brownian motion YY. Furthermore,

  • (1)

    If p2=0,p3>0p_{2}=0,p_{3}>0, and |p4|<|p_{4}|<\infty, then the fine support of AA is EE;

  • (2)

    Otherwise, the fine support of AA is E¯\overline{E}.

Proof.

We first demonstrate that

𝐄x0et𝑑At=nμn𝐄x0et𝑑Ltc^n<,x[0,).\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dA_{t}=\sum_{n\in\mathbb{N}}\mu_{n}\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t}<\infty,\quad\forall x\in[0,\infty). (4.4)

The main task is to estimate 𝐄x0et𝑑Ltc^n\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t} for each nn\in\mathbb{N}. Let Tn:=inf{t>0:Yt=c^n}T_{n}:=\inf\{t>0:Y_{t}=\hat{c}_{n}\}. According to [1, V. Theorem 3.13], there exists a positive constant knk_{n} such that

𝐄x0et𝑑Ltc^n=kn𝐄xeTn,x[0,).\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t}=k_{n}\mathbf{E}_{x}e^{-T_{n}},\quad\forall x\in[0,\infty). (4.5)

By the strong Markov property of YY and [1, IV, Proposition 1.13], we obtain

𝐄x0et𝑑Ltc^n=𝐄x0τ0et𝑑Ltc^n+𝐄xeτ0𝐄00et𝑑Ltc^n.\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t}=\mathbf{E}_{x}\int_{0}^{\tau_{0}}e^{-t}dL^{\hat{c}_{n}}_{t}+\mathbf{E}_{x}e^{-\tau_{0}}\cdot\mathbf{E}_{0}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t}.

Integrating both sides with respect to the Lebesgue measure 𝔪\mathfrak{m} on [0,)[0,\infty), we have

𝐄𝔪0et𝑑Ltc^n\displaystyle\mathbf{E}_{\mathfrak{m}}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t} =𝐄𝔪0τ0et𝑑Ltc^n+𝐄𝔪eτ0𝐄00et𝑑Ltc^n\displaystyle=\mathbf{E}_{\mathfrak{m}}\int_{0}^{\tau_{0}}e^{-t}dL^{\hat{c}_{n}}_{t}+\mathbf{E}_{\mathfrak{m}}e^{-\tau_{0}}\cdot\mathbf{E}_{0}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t}
=(1𝐄c^neτ0)+𝐄𝔪eτ0𝐄00et𝑑Ltc^n,\displaystyle=(1-\mathbf{E}_{\hat{c}_{n}}e^{-\tau_{0}})+\mathbf{E}_{\mathfrak{m}}e^{-\tau_{0}}\cdot\mathbf{E}_{0}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t},

where the second equality follows from [3, (4.1.3)]. Substituting (4.5) and 𝐄xeτ0=e2x\mathbf{E}_{x}e^{-\tau_{0}}=e^{-\sqrt{2}x} for x>0x>0 into the above equation, we obtain

kn=2(1e2c)2𝐄𝔪eTn𝐄0eTn.k_{n}=\frac{\sqrt{2}(1-e^{-\sqrt{2}c})}{\sqrt{2}\mathbf{E}_{\mathfrak{m}}e^{-T_{n}}-\mathbf{E}_{0}e^{-T_{n}}}. (4.6)

The strong Markov property also implies that for any 0<x<c^n0<x<\hat{c}_{n},

𝐄xeTn\displaystyle\mathbf{E}_{x}e^{-T_{n}} =𝐄x(eTn;Tn<τ0)+𝐄x(eTn;Tn>τ0)\displaystyle=\mathbf{E}_{x}\left(e^{-T_{n}};T_{n}<\tau_{0}\right)+\mathbf{E}_{x}\left(e^{-T_{n}};T_{n}>\tau_{0}\right) (4.7)
=𝐄x(eTn;Tn<τ0)+𝐄x(eτ0;Tn>τ0)𝐄0eTn.\displaystyle=\mathbf{E}_{x}\left(e^{-T_{n}};T_{n}<\tau_{0}\right)+\mathbf{E}_{x}\left(e^{-\tau_{0}};T_{n}>\tau_{0}\right)\mathbf{E}_{0}e^{-T_{n}}.

For x>c^nx>\hat{c}_{n}, it holds that

𝐄xeTn=e2(xc^n).\mathbf{E}_{x}e^{-T_{n}}=e^{-\sqrt{2}(x-\hat{c}_{n})}. (4.8)

Substituting (4.7), (4.8) into (4.6) and using [11, Problem 6 of page 29], we deduce that

kn=2(e2c^ne2c^n)2(e2c^n𝐄0eTn).k_{n}=\frac{\sqrt{2}(e^{\sqrt{2}\hat{c}_{n}}-e^{-\sqrt{2}\hat{c}_{n}})}{2(e^{\sqrt{2}\hat{c}_{n}}-\mathbf{E}_{0}e^{-T_{n}})}. (4.9)

Since 𝐄0eTn1\mathbf{E}_{0}e^{-T_{n}}\leq 1, it follows that

𝐄x0et𝑑Ltc^nkn2(e2c^ne2c^n)2(e2c^n1)=22(1+e2c^n)2.\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dL^{\hat{c}_{n}}_{t}\leq k_{n}\leq\frac{\sqrt{2}(e^{\sqrt{2}\hat{c}_{n}}-e^{-\sqrt{2}\hat{c}_{n}})}{2(e^{\sqrt{2}\hat{c}_{n}}-1)}=\frac{\sqrt{2}}{2}(1+e^{-\sqrt{2}\hat{c}_{n}})\leq\sqrt{2}.

Note that nμn<\sum_{n\in\mathbb{N}}\mu_{n}<\infty. Therefore, (4.4) can be concluded.

To prove that (4.3) is a PCAF of YY, we start by considering the defining sets Λn𝒢\Lambda_{n}\in{\mathscr{G}}_{\infty} of Lc^nL^{\hat{c}_{n}}. Define

Λ:=(nΛn){ωΩ:At(ω)<,t<ζ(ω)}.\Lambda:=\left(\bigcap_{n\in\mathbb{N}}\Lambda_{n}\right)\bigcap\left\{\omega\in\Omega:A_{t}(\omega)<\infty,\forall t<\zeta(\omega)\right\}.

Clearly, At|nΛn𝒢t|nΛnA_{t}|_{\bigcap_{n}\Lambda_{n}}\in{\mathscr{G}}_{t}|_{\bigcap_{n}\Lambda_{n}}. Note that

{ωΩ:At(ω)<,t<ζ(ω)}=t>0({At<,t<ζ}{ζt}).\left\{\omega\in\Omega:A_{t}(\omega)<\infty,\forall t<\zeta(\omega)\right\}=\bigcap_{t>0}\left(\{A_{t}<\infty,t<\zeta\}\cup\{\zeta\leq t\}\right). (4.10)

Thus, it is straightforward to verify that Λ𝒢\Lambda\in{\mathscr{G}}_{\infty} and that Λ\Lambda satisfies all the conditions in Definition 3.1 except for

𝐏x(Λ)=1,x[0,).\mathbf{P}_{x}(\Lambda)=1,\quad\forall x\in[0,\infty). (4.11)

To show (4.11), assume for contradiction that

𝐏x((t>0({At<,t<ζ}{ζt}))c)=𝐏x(t>0{At=,t<ζ})>0.\mathbf{P}_{x}\left(\left(\bigcap_{t>0}\left(\{A_{t}<\infty,t<\zeta\}\cup\{\zeta\leq t\}\right)\right)^{c}\right)=\mathbf{P}_{x}\left(\bigcup_{t>0}\{A_{t}=\infty,t<\zeta\}\right)>0.

This implies 𝐏x(t>0{At=})>0\mathbf{P}_{x}(\bigcup_{t>0}\{A_{t}=\infty\})>0. Since AtA_{t} is increasing in tt, {At=}\{A_{t}=\infty\} is also increasing in tt. Therefore, there exists t0>0t_{0}>0 such that 𝐏x(At0=)>0\mathbf{P}_{x}(A_{t_{0}}=\infty)>0. We have

𝐄x0et𝑑At\displaystyle\mathbf{E}_{x}\int_{0}^{\infty}e^{-t}dA_{t} 𝐄x(0t0et𝑑At;{At0=})\displaystyle\geq\mathbf{E}_{x}\left(\int_{0}^{t_{0}}e^{-t}dA_{t};\{A_{t_{0}}=\infty\}\right)
et0𝐄x(0t0𝑑At;{At0=})\displaystyle\geq e^{-t_{0}}\mathbf{E}_{x}\left(\int_{0}^{t_{0}}dA_{t};\{A_{t_{0}}=\infty\}\right)
=.\displaystyle=\infty.

This contradicts (4.5). Thus, (4.11) holds.

Finally, we examine the fine support of AA. Note that dLtc^ndL^{\hat{c}_{n}}_{t} (as a measure in tt) vanishes outside {t:Yt=c^n}\{t:Y_{t}=\hat{c}_{n}\}. Thus, dAtdA_{t} vanishes on {t:YtE¯}\{t:Y_{t}\notin\overline{E}\}. Consequently, by [1, V. Corollary 3.10], we have Supp(A)E¯\text{Supp}(A)\subset\overline{E}. On the other hand, since AtμnLtc^nA_{t}\geq\mu_{n}L^{\hat{c}_{n}}_{t}, it follows from the definition that c^nSupp(A)\hat{c}_{n}\in\text{Supp}(A). Therefore, ESupp(A)E\subset\text{Supp}(A).

If p2=0,p3>0p_{2}=0,p_{3}>0, and |p4|<|p_{4}|<\infty, then Yt=0Y_{t}=0 for all 0t<𝔢0\leq t<\mathfrak{e}, 𝐏0\mathbf{P}_{0}-a.s., where 𝔢\mathfrak{e} is given by (3.19). Note that 𝐏0(𝔢>0)=1\mathbf{P}_{0}(\mathfrak{e}>0)=1. Hence, 𝐏0(R𝔢>0)=1\mathbf{P}_{0}(R\geq\mathfrak{e}>0)=1, where R=inf{t>0:At>0}R=\inf\{t>0:A_{t}>0\}. By the definition of fine support, it follows that 0Supp(A)0\notin\text{Supp}(A). Therefore, Supp(A)=E\text{Supp}(A)=E.

It remains to show 0Supp(A)0\in\text{Supp}(A) for the remaining cases. We proceed by contradiction. Suppose 0Supp(A)0\notin\text{Supp}(A). Then 𝐏0(R>0)=1\mathbf{P}_{0}(R>0)=1. For 𝐏0\mathbf{P}_{0}-a.s. ωΩ\omega\in\Omega and 0<t<R(ω)0<t<R(\omega), we have Ltc^n(ω)=0L^{\hat{c}_{n}}_{t}(\omega)=0 for all nn\in\mathbb{N}. According to [1, V Theorem 3.8], Yt(ω)EY_{t}(\omega)\notin E for all 0<t<R(ω)0<t<R(\omega). Note that Yt(ω)Y_{t}(\omega) is càdlàg in tt. From the pathwise representation of YY (see §3.2), we can obtain that if Yt(ω)Yt(ω)Y_{t-}(\omega)\neq Y_{t}(\omega), then Yt(ω)=0Y_{t-}(\omega)=0. This fact, combined with Yt(ω)EY_{t}(\omega)\notin E for 0<t<R(ω)0<t<R(\omega) and Y0(ω)=0Y_{0}(\omega)=0, implies that Yt(ω)=0Y_{t}(\omega)=0 for all 0t<R(ω)0\leq t<R(\omega). Particularly, Yt(ω)Y_{t}(\omega) is continuous in t[0,R(ω))t\in[0,R(\omega)). This is impossible when p2>0,|p4|<p_{2}>0,|p_{4}|<\infty, because before the first jumping time, the sample paths of YY are those of a symmetric Feller’s Brownian motion with parameters (p1+|p4|,p2,p3,0)(p_{1}+|p_{4}|,p_{2},p_{3},0). This process, as a regular diffusion process on [0,)[0,\infty), can not stay at any point for an extended period; see, e.g., [23, V. (47.1)]. For the case where |p4|=|p_{4}|=\infty, the pathwise representation of YY in §3.2.3 indicates that Yt(ω)Y_{t}(\omega) is exactly a Brownian path Wt+(ω)W^{+}_{t}(\omega) (up to a transformation of time change (3.25)) for t[0,R(ω))t\in[0,R(\omega)). This also contradicts Yt(ω)=0Y_{t}(\omega)=0 for t[0,R(ω))t\in[0,R(\omega)). ∎

With the Feller’s Brownian motion YY and its PCAF A=(At)t0A=(A_{t})_{t\geq 0} given by (4.3), we can now introduce the time-changed process of YY with respect to AA. Define the right-continuous inverse of At(ω)A_{t}(\omega) for each ωΩ\omega\in\Omega as

γt(ω):={inf{s:As(ω)>t}for t<Aζ(ω)(ω),for tAζ(ω)(ω).\gamma_{t}(\omega):=\left\{\begin{aligned} &\inf\{s:A_{s}(\omega)>t\}\quad&\text{for }t<A_{\zeta(\omega)-}(\omega),\\ &\infty\quad&\text{for }t\geq A_{\zeta(\omega)-}(\omega).\end{aligned}\right.

Further, let

Yˇt(ω):=Yγt(ω)(ω),ζˇ(ω):=Aζ(ω)(ω)(=Aζ(ω)(ω)),t0,ωΩ.\check{Y}_{t}(\omega):=Y_{\gamma_{t}(\omega)}(\omega),\quad\check{\zeta}(\omega):=A_{\zeta(\omega)-}(\omega)(=A_{\zeta(\omega)}(\omega)),\quad t\geq 0,\omega\in\Omega.

Note that Yt(ω):=Y_{t}(\omega):=\partial for ζ(ω)t\zeta(\omega)\leq t\leq\infty, so Yˇt(ω)=\check{Y}_{t}(\omega)=\partial for tζˇ(ω)t\geq\check{\zeta}(\omega). According to [3, Proposition A.3.8(iv)], we may assume without loss of generality that Yˇt(ω)Supp(A){}\check{Y}_{t}(\omega)\in\text{Supp}(A)\cup\{\partial\} for all t0t\geq 0 and all ωΩ\omega\in\Omega. Set 𝒢ˇt:=𝒢γt\check{{\mathscr{G}}}_{t}:={\mathscr{G}}_{\gamma_{t}} and θˇt:=θγt\check{\theta}_{t}:=\theta_{\gamma_{t}}. It is well known that the time-changed process

Yˇ:=(Ω,𝒢,𝒢ˇt,Yˇt,θˇt,(𝐏x)xSupp(A))\check{Y}:=\left(\Omega,{\mathscr{G}},\check{{\mathscr{G}}}_{t},\check{Y}_{t},\check{\theta}_{t},\left(\mathbf{P}_{x}\right)_{x\in\text{Supp}(A)}\right) (4.12)

with lifetime ζˇ\check{\zeta} is a right process on Supp(A)\text{Supp}(A); see, e.g., [3, Theorem A.3.11].

Theorem 4.2.

Let YY be a Feller’s Brownian motion with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) as specified in Theorem 3.3, and let Yˇ\check{Y} be the time-changed process (4.12) of YY with respect to the PCAF (4.3). Then Ξ(Yˇ):=(Ξ(Yˇt))t0\Xi(\check{Y}):=\left(\Xi(\check{Y}_{t})\right)_{t\geq 0} is a QQ-process whose birth-death density matrix is (2.1), where Ξ\Xi is defined as (4.2). Furthermore,

  • (1)

    If p2=0,p3>0p_{2}=0,p_{3}>0, and |p4|=0|p_{4}|=0, then Ξ(Yˇ)\Xi(\check{Y}) is the minimal QQ-process;

  • (2)

    If p2=0,p3>0p_{2}=0,p_{3}>0, and 0<|p4|<0<|p_{4}|<\infty, then Ξ(Yˇ)\Xi(\check{Y}) is a Doob process;

  • (3)

    Otherwise, Ξ(Yˇ)\Xi(\check{Y}) is a Feller QQ-process.

Proof.

Denote the transition semigroup and resolvent of Yˇ\check{Y} by (Tˇt)t0(\check{T}_{t})_{t\geq 0} and (Gˇα)α>0(\check{G}_{\alpha})_{\alpha>0}, respectively. Define

pˇij(t):=Tˇt1{c^j}(c^i),t0,i,j,\check{p}_{ij}(t):=\check{T}_{t}1_{\{\hat{c}_{j}\}}(\hat{c}_{i}),\quad t\geq 0,i,j\in\mathbb{N},

and

Ψˇij(α):=0eαtpˇij(t)𝑑t,α>0,i,j.\check{\Psi}_{ij}(\alpha):=\int_{0}^{\infty}e^{-\alpha t}\check{p}_{ij}(t)dt,\quad\alpha>0,i,j\in\mathbb{N}.

The goal is to prove that (pˇij(t))i,j(\check{p}_{ij}(t))_{i,j\in\mathbb{N}} is a QQ-process.

We first demonstrate that (pˇij(t))i,j(\check{p}_{ij}(t))_{i,j\in\mathbb{N}} is a standard transition matrix. According to [26, §2.5, Theorem 1], it suffices to verify that Ψˇij(α)\check{\Psi}_{ij}(\alpha) satisfies the following conditions:

Ψˇij(α)0,αjΨˇij(α)1,\displaystyle\check{\Psi}_{ij}(\alpha)\geq 0,\quad\alpha\sum_{j\in\mathbb{N}}\check{\Psi}_{ij}(\alpha)\leq 1, (4.13)
Ψˇij(α)Ψˇij(β)+(αβ)kΨˇik(α)Ψˇkj(β)=0,α,β>0,\displaystyle\check{\Psi}_{ij}(\alpha)-\check{\Psi}_{ij}(\beta)+(\alpha-\beta)\sum_{k\in\mathbb{N}}\check{\Psi}_{ik}(\alpha)\check{\Psi}_{kj}(\beta)=0,\quad\forall\alpha,\beta>0,
limααΨˇij(α)=δij,\displaystyle\lim_{\alpha\rightarrow\infty}\alpha\check{\Psi}_{ij}(\alpha)=\delta_{ij},

where δij\delta_{ij} is the Kronecker delta. To accomplish this, note that

Ψˇij(α)=Gˇα1{c^j}(c^i).\check{\Psi}_{ij}(\alpha)=\check{G}_{\alpha}1_{\{\hat{c}_{j}\}}(\hat{c}_{i}).

Thus, the first condition in (4.13) is straightforward, and the third condition follows from the right continuity of Yˇ\check{Y}. To prove the second condition, it is sufficient to demonstrate

Gˇα1{0}(c^i)=0\check{G}_{\alpha}1_{\{0\}}(\hat{c}_{i})=0 (4.14)

and then apply the resolvent equation of Gˇα\check{G}_{\alpha}. In fact, it follows from the definition of Yˇ\check{Y}, [22, Proposition 4.9 of page 8], and the Fubini theorem that

Gˇα1{0}(c^i)\displaystyle\check{G}_{\alpha}1_{\{0\}}(\hat{c}_{i}) =𝐄c^i0eαAγt1{0}(Yγt)𝑑t=𝐄c^i0eαAt1{0}(Yt)𝑑At\displaystyle=\mathbf{E}_{\hat{c}_{i}}\int_{0}^{\infty}e^{-\alpha A_{\gamma_{t}}}1_{\{0\}}(Y_{\gamma_{t}})dt=\mathbf{E}_{\hat{c}_{i}}\int_{0}^{\infty}e^{-\alpha A_{t}}1_{\{0\}}(Y_{t})dA_{t}
=nμn𝐄c^i0eαAt1{0}(Yt)𝑑Ltc^n=0.\displaystyle=\sum_{n\in\mathbb{N}}\mu_{n}\mathbf{E}_{\hat{c}_{i}}\int_{0}^{\infty}e^{-\alpha A_{t}}1_{\{0\}}(Y_{t})dL^{\hat{c}_{n}}_{t}=0.

Thus, (4.14) is established.

Recall that Y0Y^{0}, with lifetime ζ0\zeta^{0}, is an absorbing Brownian motion on (0,)(0,\infty). According to Lemma 3.7 and Lemma 4.1, we have that

At0:=Atζ0,t0A^{0}_{t}:=A_{t\wedge\zeta^{0}},\quad t\geq 0

is a PACF of Y0Y^{0}, with Revuz measure nμnδc^n\sum_{n\in\mathbb{N}}\mu_{n}\delta_{\hat{c}_{n}}. Define

Y^0=(Ω,𝒢,𝒢^t,Y^t0,θ^t0,(𝐏x)xE),\hat{Y}^{0}=\left(\Omega,{\mathscr{G}},\hat{{\mathscr{G}}}_{t},\hat{Y}^{0}_{t},\hat{\theta}^{0}_{t},(\mathbf{P}_{x})_{x\in E}\right),

with lifetime ζ^0:=Aζ00=Aζ0\hat{\zeta}^{0}:=A^{0}_{\zeta^{0}}=A_{\zeta^{0}}, as the time-changed process of Y0Y^{0} with respect to the PACF A0A^{0}. Specifically,

𝒢^t:=𝒢γt0,θ^t0:=θγt00,Y^t0:=Yγt00,\hat{{\mathscr{G}}}_{t}:={\mathscr{G}}_{\gamma^{0}_{t}},\quad\hat{\theta}^{0}_{t}:=\theta^{0}_{\gamma^{0}_{t}},\quad\hat{Y}^{0}_{t}:=Y^{0}_{\gamma^{0}_{t}},

with

γt0(ω):={inf{s:As0(ω)>t}for t<ζ^0,for tζ^0.\gamma^{0}_{t}(\omega):=\left\{\begin{aligned} &\inf\{s:A^{0}_{s}(\omega)>t\}\quad&\text{for }t<\hat{\zeta}^{0},\\ &\infty\quad&\text{for }t\geq\hat{\zeta}^{0}.\end{aligned}\right.

It has been established in [18, Lemma 3.1] that Ξ(Y^0)\Xi(\hat{Y}^{0}) is exactly the minimal QQ-process.

Next, let us examine the killed process of Yˇ\check{Y} upon hitting 0E¯0\in\overline{E}. More precisely, let ηˇn:=inf{t>0:Yˇt=c^n}\check{\eta}_{n}:=\inf\{t>0:\check{Y}_{t}=\hat{c}_{n}\} for all n1n\geq 1. From the definition (4.12) of Yˇ\check{Y}, we find that for 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. ωΩ\omega\in\Omega, ηˇn(ω)\check{\eta}_{n}(\omega) is increasing in nn for n>in>i. Hence,

ηˇ(ω):=limnηˇn(ω)\check{\eta}_{\infty}(\omega):=\lim_{n\rightarrow\infty}\check{\eta}_{n}(\omega) (4.15)

is well defined for 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. ω\omega and all ii\in\mathbb{N}. We will show that ηˇ=Aζ0=ζ^0\check{\eta}_{\infty}=A_{\zeta^{0}}=\hat{\zeta}^{0}, and hence the killed process

Yˇt0:={Yˇt,t<ζˇηˇ,,tζˇηˇ,\check{Y}^{0}_{t}:=\left\{\begin{aligned} &\check{Y}_{t},\quad&t<\check{\zeta}\wedge\check{\eta}_{\infty},\\ &\partial,\quad&t\geq\check{\zeta}\wedge\check{\eta}_{\infty},\end{aligned}\right.

with lifetime ζˇ0:=ζˇηˇ(=ηˇ)\check{\zeta}^{0}:=\check{\zeta}\wedge\check{\eta}_{\infty}(=\check{\eta}_{\infty}), on EE is identical to Y^0\hat{Y}^{0}. To prove this, fix 𝐏c^i\mathbf{P}_{\hat{c}_{i}} and let ηn:=inf{t>0:Yt=c^n}\eta_{n}:=\inf\{t>0:Y_{t}=\hat{c}_{n}\} for n>in>i. Note that Yηn=c^nY_{\eta_{n}}=\hat{c}_{n} and Ytc^nY_{t}\neq\hat{c}_{n} for all t<ηnt<\eta_{n}, 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. It follows from [1, V. Theorem 3.8] that Aηn=Aηnε0A_{\eta_{n}}=A_{\eta_{n}-\varepsilon_{0}} for some ε0>0\varepsilon_{0}>0 and Aηn+ε>AηnA_{\eta_{n}+\varepsilon}>A_{\eta_{n}} for all ε>0\varepsilon>0. Consequently,

γAηn=inf{s>0:As>Aηn}=ηn,\gamma_{A_{\eta_{n}}}=\inf\{s>0:A_{s}>A_{\eta_{n}}\}=\eta_{n},

and

γtηnε0,t<Aηn.\gamma_{t}\leq\eta_{n}-\varepsilon_{0},\quad\forall t<A_{\eta_{n}}.

These yield

YˇAηn=YγAηn=Yηn=c^n\check{Y}_{A_{\eta_{n}}}=Y_{\gamma_{A_{\eta_{n}}}}=Y_{\eta_{n}}=\hat{c}_{n}

and

Yˇt=Yγtc^n,t<Aηn.\check{Y}_{t}=Y_{\gamma_{t}}\neq\hat{c}_{n},\quad\forall t<A_{\eta_{n}}.

In other words, ηˇn=Aηn\check{\eta}_{n}=A_{\eta_{n}}, 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. Noting that limnηn=ζ0\lim_{n\rightarrow\infty}\eta_{n}=\zeta^{0}, we obtain that ηˇ=limnηˇn=limnAηn=Aζ0\check{\eta}_{\infty}=\lim_{n\rightarrow\infty}\check{\eta}_{n}=\lim_{n\rightarrow\infty}A_{\eta_{n}}=A_{\zeta^{0}}.

According to the argument in the previous two paragraphs, we can conclude that pˇij0(t):=𝐏c^i(Yˇt=c^j,t<ηˇ)\check{p}^{0}_{ij}(t):=\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{t}=\hat{c}_{j},t<\check{\eta}_{\infty}) is the transition matrix of the minimal QQ-process. This implies that

limt0pˇij0(t)δijt=qij.\lim_{t\rightarrow 0}\frac{\check{p}^{0}_{ij}(t)-\delta_{ij}}{t}=q_{ij}. (4.16)

Mimicking the proof of [18, Proposition 3.6], we can also obtain that

limt0𝐏c^i(ηˇt)t=0.\lim_{t\rightarrow 0}\frac{\mathbf{P}_{\hat{c}_{i}}(\check{\eta}_{\infty}\leq t)}{t}=0. (4.17)

Combining (4.16) and (4.17), we obtain

limt0pˇij(t)δijt=qij.\lim_{t\rightarrow 0}\frac{\check{p}_{ij}(t)-\delta_{ij}}{t}=q_{ij}.

In other words, (pˇij(t))i,j(\check{p}_{ij}(t))_{i,j\in\mathbb{N}} is a QQ-process.

Finally, let us classify the QQ-process Ξ(Yˇ)\Xi(\check{Y}) for different cases. For the first case, where p2=0,p3>0p_{2}=0,p_{3}>0 and |p4|=0|p_{4}|=0, we observe that tAtt\mapsto A_{t} does not increase after time ζ0\zeta^{0} (according to the sample path representation in §3.2.1). Consequently, γt=\gamma_{t}=\infty for all tAζ0t\geq A_{\zeta^{0}}, which implies that Y^t=\hat{Y}_{t}=\partial for all tAζ0=ηˇt\geq A_{\zeta^{0}}=\check{\eta}_{\infty}. Thus, Ξ(Yˇ)\Xi(\check{Y}) aligns with the minimal QQ-process. In the second case, where p2=0,p3>0p_{2}=0,p_{3}>0 and 0<|p4|<0<|p_{4}|<\infty, tAtt\mapsto A_{t} may continue to increase after time ζ0\zeta^{0}. A similar argument shows that Ξ(Yˇ)\Xi(\check{Y}) is not the minimal QQ-process. However, according to Lemma 4.1, Yˇ\check{Y} is a right process with state space Supp(A)=E\text{Supp}(A)=E. Therefore, YˇηˇE{}\check{Y}_{\check{\eta}_{\infty}}\in E\cup\{\partial\}, 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. for all ii\in\mathbb{N}. Based on [18, Corollary 5.2], it follows that Ξ(Yˇ)\Xi(\check{Y}) is a Doob process.

For the remaining cases, it suffices to show that Yˇηˇ=0\check{Y}_{\check{\eta}_{\infty}}=0, 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. for all ii\in\mathbb{N}. To demonstrate this, consider

γηˇ=inf{t>0:At>ηˇ}=inf{t>0:At>Aζ0}.\gamma_{\check{\eta}_{\infty}}=\inf\{t>0:A_{t}>\check{\eta}_{\infty}\}=\inf\{t>0:A_{t}>A_{\zeta^{0}}\}.

If it were false that γηˇ=ζ0\gamma_{\check{\eta}_{\infty}}=\zeta^{0}, then there would exist some ε0>0\varepsilon_{0}>0 such that Aζ0+ε0=Aζ0A_{\zeta^{0}+\varepsilon_{0}}=A_{\zeta^{0}}. This implies that YtEY_{t}\notin E for all ζ0<t<ζ0+ε0\zeta^{0}<t<\zeta^{0}+\varepsilon_{0} by [1, V. Theorem 3.8]. Since Yζ0=0Y_{\zeta^{0}}=0 and YY is right continuous, it follows that Yt=0Y_{t}=0 for all ζ0t<ζ0+ε0\zeta^{0}\leq t<\zeta^{0}+\varepsilon_{0}. However, this contradicts the previous argument that YY cannot remain at 0 for any extended period, as noted in the last paragraph of the proof of Lemma 4.1. Therefore, we have Yˇηˇ=Yζ0=0\check{Y}_{\check{\eta}_{\infty}}=Y_{\zeta^{0}}=0. ∎

In the context of general Markov process, the Feller’s Brownian motion YY with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) is termed conservative if 𝐏x(ζ<)=0\mathbf{P}_{x}(\zeta<\infty)=0 for any x[0,)x\in[0,\infty), which is equivalent to p1=0p_{1}=0 (see [10, §15]). In terms of continuous-time Markov chains, a conservative QQ-process is also referred to as an honest QQ-process (see §2.2).

Corollary 4.3.

If the Feller’s Brownian motion YY is conservative, then for any a>0a>0 and x[0,)x\in[0,\infty),

𝐏x(La=)=1.\mathbf{P}_{x}(L^{a}_{\infty}=\infty)=1. (4.18)

Particularly, the QQ-process (Ξ(Yˇt))t0\left(\Xi(\check{Y}_{t})\right)_{t\geq 0}, obtained in Theorem 4.2, is also conservative.

Proof.

Write T:=Ta:=inf{t>0:Yt=a}T:=T_{a}:=\inf\{t>0:Y_{t}=a\}. We first show that

𝐏x(T<)=1.\mathbf{P}_{x}(T<\infty)=1. (4.19)

To demonstrate this, let σ:={t>0:Yta}T\sigma:=\{t>0:Y_{t}\geq a\}\leq T. Observe that

{σ=}={Yt<a,t>0}.\{\sigma=\infty\}=\{Y_{t}<a,\forall t>0\}.

The sample path of YY is described by (3.23) up to a time change transformation (see (3.25)). Given that p1=0p_{1}=0, it follows that

{σ=}{Wt+<a,t>0}.\{\sigma=\infty\}\subset\{W^{+}_{t}<a,\forall t>0\}.

Therefore, 𝐏x(σ=)𝐏x(Wt+<a,t>0)=0\mathbf{P}_{x}(\sigma=\infty)\leq\mathbf{P}_{x}(W^{+}_{t}<a,\forall t>0)=0. On the other hand, by the strong Markov property, we have

𝐏x(T=,σ<)=𝐄x(𝐏Yσ(T=);{σ<}).\mathbf{P}_{x}(T=\infty,\sigma<\infty)=\mathbf{E}_{x}\left(\mathbf{P}_{Y_{\sigma}}(T=\infty);\{\sigma<\infty\}\right).

Since YσaY_{\sigma}\geq a, 𝐏x\mathbf{P}_{x}-a.s. on {σ<}\{\sigma<\infty\}, it follows that 𝐏Yσ(T=)=0\mathbf{P}_{Y_{\sigma}}(T=\infty)=0, 𝐏x\mathbf{P}_{x}-a.s. on {σ<}\{\sigma<\infty\}. Therefore, 𝐏x(T=,σ<)=0\mathbf{P}_{x}(T=\infty,\sigma<\infty)=0. Combining this with 𝐏x(σ=)=0\mathbf{P}_{x}(\sigma=\infty)=0, we can eventually derive (4.19).

Next, by mimicking the computation in (4.9), we can show that for α>0\alpha>0,

𝐄a0eαt𝑑Lta=12αe2αae2αae2αa𝐄0eαT.\mathbf{E}_{a}\int_{0}^{\infty}e^{-\alpha t}dL^{a}_{t}=\frac{1}{\sqrt{2\alpha}}\frac{e^{\sqrt{2\alpha}a}-e^{-\sqrt{2\alpha}a}}{e^{\sqrt{2\alpha}a}-\mathbf{E}_{0}e^{-\alpha T}}.

It follows from (4.19) (with x=0x=0) that

limα0𝐄a0eαt𝑑Lta=.\lim_{\alpha\rightarrow 0}\mathbf{E}_{a}\int_{0}^{\infty}e^{-\alpha t}dL^{a}_{t}=\infty. (4.20)

Finally, by using (4.19) and (4.20), we can apply [1, V. Theorem 3.17] to conclude (4.18). ∎

4.3. Uniqueness of PCAF for time change

We continue to examine the time change transformation in Theorem 4.2. The goal is to demonstrate that (4.3) is indeed the unique PCAF of YY for which the corresponding time-changed process is a QQ-process with the given density matrix (2.3).

Theorem 4.4.

Let YY be a Feller’s Brownian motion and A1=(At1)t0A^{1}=(A^{1}_{t})_{t\geq 0} be a PCAF of YY such that Supp(A1)E¯\text{Supp}(A^{1})\subset\overline{E}. Denote by Yˇ1\check{Y}^{1} the time-changed process of YY with respect to the PCAF A1A^{1}. If X1:=Ξ(Yˇ1)X^{1}:=\Xi(\check{Y}^{1}) is a QQ-process with the given density matrix (2.3), then

At1=At,t0,A^{1}_{t}=A_{t},\quad\forall t\geq 0,

where A=(At)t0A=(A_{t})_{t\geq 0} is defined as (4.3).

Proof.

Let =(t)t0\ell=(\ell_{t})_{t\geq 0} denote the local time of YY at 0, as discussed in [10] or [1, V§3]. (We will soon see that the normalization of this local time is not necessary for our proof.) Define

A¯t:=At+t,t0.\bar{A}_{t}:=A_{t}+\ell_{t},\quad t\geq 0.

Clearly, A¯=(A¯t)t0\bar{A}=(\bar{A}_{t})_{t\geq 0} is a PCAF of YY. To verify the condition in [1, V§3, Proposition 3.11] for A1A^{1} and A¯\bar{A}, consider g+([0,))g\in\mathcal{B}_{+}([0,\infty)) such that

𝐄x0g(Yt)𝑑A¯t=0,x[0,).\mathbf{E}_{x}\int_{0}^{\infty}g(Y_{t})d\bar{A}_{t}=0,\quad\forall x\in[0,\infty).

Then g|E¯0g|_{\overline{E}}\equiv 0. Since Supp(A1)E¯\text{Supp}(A^{1})\subset\overline{E}, we have

𝐄x0g(Yt)𝑑At1=0,x[0,).\mathbf{E}_{x}\int_{0}^{\infty}g(Y_{t})dA^{1}_{t}=0,\quad\forall x\in[0,\infty).

Applying [1, V§3, Proposition 3.11] to A1A^{1} and A¯\bar{A}, we obtain that

At1=nμ~nLtc^n+μ~t,A^{1}_{t}=\sum_{n\in\mathbb{N}}\tilde{\mu}_{n}L^{\hat{c}_{n}}_{t}+\tilde{\mu}_{\infty}\ell_{t},

for some μ~n0\tilde{\mu}_{n}\geq 0 and μ~0\tilde{\mu}_{\infty}\geq 0.

The fact that X1X^{1} is a QQ-process implies that

𝐄x0eαt1{0}(Yˇt1)𝑑t=0\mathbf{E}_{x}\int_{0}^{\infty}e^{-\alpha t}1_{\{0\}}(\check{Y}^{1}_{t})dt=0 (4.21)

for α>0\alpha>0 and xSupp(A1)x\in\text{Supp}(A^{1}). However, the left-hand side of (4.21) can be expressed as

𝐄x0eαAt11{0}(Yt)𝑑At1=μ~𝐄x0eαAt1𝑑t.\mathbf{E}_{x}\int_{0}^{\infty}e^{-\alpha A^{1}_{t}}1_{\{0\}}(Y_{t})dA^{1}_{t}=\tilde{\mu}_{\infty}\cdot\mathbf{E}_{x}\int_{0}^{\infty}e^{-\alpha A^{1}_{t}}d\ell_{t}.

Hence, we conclude that μ~=0\tilde{\mu}_{\infty}=0.

On the other hand, it is straightforward (or follows from the proof of Theorem 4.2) that the time-changed process of the absorbing Brownian motion Y0Y^{0}, with lifetime ζ0\zeta^{0}, with respect to the PCAF (Atζ01)t0(A^{1}_{t\wedge\zeta^{0}})_{t\geq 0} corresponds precisely to the minimal QQ-process. Particularly, the Revuz measure of

Atζ01=nμ~nLtζ0c^nA^{1}_{t\wedge\zeta^{0}}=\sum_{n\in\mathbb{N}}\tilde{\mu}_{n}L^{\hat{c}_{n}}_{t\wedge\zeta^{0}}

with respect to Y0Y^{0} is actually the speed measure μ\mu. Therefore, we can conclude that μ~n=μn\tilde{\mu}_{n}=\mu_{n} for all nn\in\mathbb{N}. ∎

5. Parameters of birth-death processes obtained by time change

According to Theorem 2.1, the QQ-process Ξ(Yˇ)\Xi(\check{Y}) obtained in Theorem 4.2 (excluding the first case, which yields the minimal QQ-process) admits a resolvent representation given by a triple (γ,β,ν)(\gamma,\beta,\nu), which is unique up to a multiplicative constant. It is certainly interesting to explore how the parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) of Feller’s Brownian motion determine (γ,β,ν)(\gamma,\beta,\nu).

5.1. Main result

Recall that the sequence {c^n:n}\{\hat{c}_{n}:n\in\mathbb{N}\} is given in (4.1). Based on the measure p4p_{4} on (0,)(0,\infty), we define a sequence {𝔭n:n}\{\mathfrak{p}_{n}:n\in\mathbb{N}\} as follows:

𝔭0:=(c^1,c^0]xc^1c^0c^1p4(dx)+p4((c^0,))\mathfrak{p}_{0}:=\int_{(\hat{c}_{1},\hat{c}_{0}]}\frac{x-\hat{c}_{1}}{\hat{c}_{0}-\hat{c}_{1}}p_{4}(dx)+p_{4}\left((\hat{c}_{0},\infty)\right) (5.1)

and

𝔭n:=(c^n+1,c^n]xc^n+1c^nc^n+1p4(dx)+(c^n,c^n1)c^n1xc^n1c^np4(dx),n1.\mathfrak{p}_{n}:=\int_{(\hat{c}_{n+1},\hat{c}_{n}]}\frac{x-\hat{c}_{n+1}}{\hat{c}_{n}-\hat{c}_{n+1}}p_{4}(dx)+\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}p_{4}(dx),\quad n\geq 1. (5.2)

Note that if |p4|<|p_{4}|<\infty, then |p4|=n𝔭n|p_{4}|=\sum_{n\in\mathbb{N}}\mathfrak{p}_{n}. Our main result is as follows.

Theorem 5.1.

Let YY be a Feller’s Brownian motion with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) as described in Theorem 3.3 (excluding the first case, i.e., p2=0,p3>0p_{2}=0,p_{3}>0, and |p4|=0|p_{4}|=0, in Theorem 4.2). Then the parameters (γ,β,ν)(\gamma,\beta,\nu) that determine the resolvent matrix of the QQ-process Ξ(Yˇ)\Xi(\check{Y}), obtained in Theorem 4.2, are given by

γ=p1,β=2p2,νn=𝔭n,n\gamma=p_{1},\quad\beta=2p_{2},\quad\nu_{n}=\mathfrak{p}_{n},\;n\in\mathbb{N}

up to a multiplicative constant, where {𝔭n}\{\mathfrak{p}_{n}\} is the sequence defined by (5.1) and (5.2).

The proof of Theorem 5.1 will be completed in the the following three sections. Here, we present a consequence: not only is every time-changed Feller’s Brownian motion a QQ-process, as shown in Theorem 4.2, but also every QQ-process can be derived from a Feller’s Brownian motion through time change.

Corollary 5.2.

For every QQ-process XX, there exists a Feller’s Brownian motion YY such that the QQ-process obtained from YY through time change, as described in Theorem 4.2, is identical in law to XX.

Proof.

Let (γ,β,ν)(\gamma,\beta,\nu) be the triple determining the resolvent matrix of XX. When XX is the minimal QQ-process, we can choose YY with p2=0,p3>0p_{2}=0,p_{3}>0 and |p4|=0|p_{4}|=0. In this case, the first case of Theorem 4.2 applies. When XX is a Doob process, i.e., β=0,|ν|<\beta=0,|\nu|<\infty, we can choose

p1=γ,p2=0,p3>0,p4=nνnδc^n.p_{1}=\gamma,\quad p_{2}=0,\quad p_{3}>0,\quad p_{4}=\sum_{n\in\mathbb{N}}\nu_{n}\delta_{\hat{c}_{n}}.

Then, we apply the second case of Theorem 4.2 and Theorem 5.1. When XX is non-minimal and non-Doob, we can choose

p1=γ,p2=β2,p3=0,p4=nνnδc^n.p_{1}=\gamma,\quad p_{2}=\frac{\beta}{2},\quad p_{3}=0,\quad p_{4}=\sum_{n\in\mathbb{N}}\nu_{n}\delta_{\hat{c}_{n}}.

(To satisfy (3.3), it may be necessary to multiply all parameters by a positive constant.) ∎

5.2. Pathwise construction of QQ-processes

In this subsection, we discuss the pathwise construction of a Feller QQ-process XX based on Theorem 5.1. For simplicity, we focus on the honest case where γ=0\gamma=0, and consider the Feller process X^=Ξ1(X)\hat{X}=\Xi^{-1}(X) on E¯\overline{E} with parameters (ν,β,0)(\nu,\beta,0).

In the context of QQ-processes, the (Q,1)(Q,1)-process plays a role analogous to that of the reflecting Brownian motion in the context of Feller’s Brownian motion. We denote its corresponding QQ-process on E¯\overline{E} by W^+\hat{W}^{+}. This process is the time-changed process of the reflecting Brownian motion W+W^{+} with respect to At+:=nμnLt+,c^nA^{+}_{t}:=\sum_{n\in\mathbb{N}}\mu_{n}L^{+,\hat{c}_{n}}_{t}, where L+,c^nL^{+,\hat{c}_{n}} denotes the local time of W+W^{+} at c^n\hat{c}_{n} as defined in Definition 3.1. Let γ+\gamma^{+} be the right-continuous inverse of A+A^{+}. The local time of W^+\hat{W}^{+} at 0, denoted by ^\hat{\ell}, can be derived from the Brownian local time \ell through the corresponding time change transformation, specifically, ^t=γt+\hat{\ell}_{t}=\ell_{\gamma^{+}_{t}}.

We initially considered replacing W+W^{+} with W^+\hat{W}^{+} and constructing the general QQ-process similarly to (3.23), by defining

X^t1:=Z(Z1(^t))^t+W^t+,t0,\hat{X}^{1}_{t}:=Z(Z^{-1}(\hat{\ell}_{t}))-\hat{\ell}_{t}+\hat{W}^{+}_{t},\quad t\geq 0, (5.3)

where ZZ is the subordinator (3.21) with p2=β2p_{2}=\frac{\beta}{2} and p4=nνnδc^np_{4}=\sum_{n\in\mathbb{N}}\nu_{n}\delta_{\hat{c}_{n}}, which is independent of W^+\hat{W}^{+}. (A similar idea appears in [10, §20].) However, at a time tt when ^t\hat{\ell}_{t} increases to a discontinuous point xx (=^t=\hat{\ell}_{t}) of Z(Z1)Z(Z^{-1}), where h(x):=Z(Z1(x))x=lh(x):=Z(Z^{-1}(x))-x=l for some 0<lE0<l\in E and h(x)=0h(x-)=0, the term h(^t)h(\hat{\ell}_{t}) in (5.3) will continuously decrease from ll for a short period thereafter (see [10, §12]). Meanwhile, W^+\hat{W}^{+} ‘diffuses’ within the space E¯\overline{E} near 0. This discrepancy causes the process defined by (5.3) to move outside of E¯\overline{E}.

Note that (5.3) can be rewritten as

X^t1=Z(Z1(γt+))γt++Wγt++,t0.\hat{X}^{1}_{t}=Z(Z^{-1}(\ell_{\gamma^{+}_{t}}))-\ell_{\gamma^{+}_{t}}+W^{+}_{\gamma^{+}_{t}},\quad t\geq 0.

However, according to (3.23) and Theorem 5.1, the correct pathwise representation of X^\hat{X} should be

X^t=Z(Z1(γt))γt+Wγt+,t0,\hat{X}_{t}=Z(Z^{-1}(\ell_{\gamma_{t}}))-\ell_{\gamma_{t}}+W^{+}_{\gamma_{t}},\quad t\geq 0,

where γ\gamma is the right-continuous inverse of (4.3). In other words, (5.3) incorrectly reverses the order of adding jumps using ZZ and the transformation of time change.

6. Proof of Theorem 5.1 for |p4|<|p_{4}|<\infty

In this section, we will prove Theorem 5.1 for various cases where |p4|<|p_{4}|<\infty. Simultaneously, we will provide a more comprehensive characterization of the corresponding QQ-processes. From now on, we will denote Xt:=Ξ(Yˇt)X_{t}:=\Xi(\check{Y}_{t}) for t0t\geq 0 and use X:=(Xt)t0X:=(X_{t})_{t\geq 0} for convenience.

6.1. Doob processes

According to Theorem 4.2, XX is a Doob process (but not the minimal QQ-process) if and only if p2=0,p3>0p_{2}=0,p_{3}>0, and 0<|p4|<0<|p_{4}|<\infty.

Theorem 6.1.

If p2=0,p3>0p_{2}=0,p_{3}>0 and 0<|p4|<0<|p_{4}|<\infty, then the parameters (γ,β,ν)(\gamma,\beta,\nu) that determine the resolvent matrix of the Doob process X=Ξ(Yˇ)X=\Xi(\check{Y}) are given by

γ=p1,β=0,νn=𝔭n,n.\gamma=p_{1},\quad\beta=0,\quad\nu_{n}=\mathfrak{p}_{n},\quad n\in\mathbb{N}.

Particularly, the instantaneous distribution of XX is

π({})=p1p1+|p4|,π({n})=𝔭np1+|p4|,n.\pi(\{\partial\})=\frac{p_{1}}{p_{1}+|p_{4}|},\quad\pi(\{n\})=\frac{\mathfrak{p}_{n}}{p_{1}+|p_{4}|},\;n\in\mathbb{N}.
Proof.

The fact that β=0\beta=0 follows directly from Theorems 2.1 and 4.2. Fix 𝐏c^i\mathbf{P}_{\hat{c}_{i}} for some ii\in\mathbb{N}, and let ηˇ\check{\eta}_{\infty} be defined as (4.15), representing the first flying time (see [18, Corollary 4.7]) of Yˇ\check{Y}. According to [18, Theorem 5.1], it suffices to demonstrate that

𝐏c^i(Yˇηˇ=)=p1p1+|p4|,𝐏c^i(Yˇηˇ=c^n)=𝔭np1+|p4|,n.\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{\check{\eta}_{\infty}}=\partial)=\frac{p_{1}}{p_{1}+|p_{4}|},\quad\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{\check{\eta}_{\infty}}=\hat{c}_{n})=\frac{\mathfrak{p}_{n}}{p_{1}+|p_{4}|},\quad n\in\mathbb{N}. (6.1)

We will use the same symbols as those in the proof of Theorem 4.2. Define

σ1:=inf{t>ζ0:Yt0},σ2:=inf{t>σ1:YtE},\sigma_{1}:=\inf\{t>\zeta^{0}:Y_{t}\neq 0\},\quad\sigma_{2}:=\inf\{t>\sigma_{1}:Y_{t}\in E\},

and

σ:=inf{t>0:YtE}.\sigma:=\inf\{t>0:Y_{t}\in E\}.

Note that ηˇ=Aζ0\check{\eta}_{\infty}=A_{\zeta^{0}}. From the pathwise representation of YY in §3.2.1, it follows that

γηˇ=inf{t>0:At>Aζ0}=σ2.\gamma_{\check{\eta}_{\infty}}=\inf\{t>0:A_{t}>A_{\zeta^{0}}\}=\sigma_{2}.

It is straightforward to verify that Yσ2=Yσθσ1Y_{\sigma_{2}}=Y_{\sigma}\circ\theta_{\sigma_{1}} and that {Yσ2=c^n}{Yσ1(c^n+1,c^n1)}\{Y_{\sigma_{2}}=\hat{c}_{n}\}\subset\{Y_{\sigma_{1}}\in(\hat{c}_{n+1},\hat{c}_{n-1})\} (with the convention c^1:=\hat{c}_{-1}:=\infty). By the strong Markov property of YY, we have

𝐏c^i(Yˇηˇ=c^n)\displaystyle\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{\check{\eta}_{\infty}}=\hat{c}_{n}) =𝐏c^i(Yσ2=c^n)=𝐏c^i(Yσθσ1=c^n,Yσ1(c^n+1,c^n1))\displaystyle=\mathbf{P}_{\hat{c}_{i}}(Y_{\sigma_{2}}=\hat{c}_{n})=\mathbf{P}_{\hat{c}_{i}}\left(Y_{\sigma}\circ\theta_{\sigma_{1}}=\hat{c}_{n},Y_{\sigma_{1}}\in(\hat{c}_{n+1},\hat{c}_{n-1})\right)
=𝐏c^i(𝐏Yσ1(Yσ=c^n);Yσ1(c^n+1,c^n1)).\displaystyle=\mathbf{P}_{\hat{c}_{i}}\left(\mathbf{P}_{Y_{\sigma_{1}}}(Y_{\sigma}=\hat{c}_{n});Y_{\sigma_{1}}\in(\hat{c}_{n+1},\hat{c}_{n-1})\right).

Note that 𝐏c^i(Yσ1dx)=λ(dx)\mathbf{P}_{\hat{c}_{i}}\left(Y_{\sigma_{1}}\in dx\right)=\lambda(dx), where λ\lambda is given by (3.20). For x(c^n+1,c^n1)x\in(\hat{c}_{n+1},\hat{c}_{n-1}), we have

𝐏x(Yσ=c^n)={xc^n+1c^nc^n+1,x(c^n+1,c^n],c^n1xc^n1c^n,x(c^n,c^n1),\mathbf{P}_{x}(Y_{\sigma}=\hat{c}_{n})=\left\{\begin{aligned} &\frac{x-\hat{c}_{n+1}}{\hat{c}_{n}-\hat{c}_{n+1}},\quad&x\in(\hat{c}_{n+1},\hat{c}_{n}],\\ &\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}},\quad&x\in(\hat{c}_{n},\hat{c}_{n-1}),\end{aligned}\right.

with the convention :=1\frac{\infty}{\infty}:=1. Thus, a straightforward computation yields

𝐏c^i(Yˇηˇ=c^n)=(c^n+1,c^n1)𝐏x(Yσ=c^n)λ(dx)=𝔭np1+|p4|.\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{\check{\eta}_{\infty}}=\hat{c}_{n})=\int_{(\hat{c}_{n+1},\hat{c}_{n-1})}\mathbf{P}_{x}(Y_{\sigma}=\hat{c}_{n})\lambda(dx)=\frac{\mathfrak{p}_{n}}{p_{1}+|p_{4}|}.

Finally, 𝐏c^i(Yˇηˇ=)=1n𝐏c^i(Yˇηˇ=c^n)=p1p1+|p4|\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{\check{\eta}_{\infty}}=\partial)=1-\sum_{n\in\mathbb{N}}\mathbf{P}_{\hat{c}_{i}}(\check{Y}_{\check{\eta}_{\infty}}=\hat{c}_{n})=\frac{p_{1}}{p_{1}+|p_{4}|}. Therefore, (6.1) is established. This completes the proof. ∎

6.2. Symmetric case

The symmetric case, where p2>0p_{2}>0 and |p4|=0|p_{4}|=0, can be analyzed using Dirichlet form theory. In this framework, the time change of a Markov process corresponds to a trace Dirichlet form (see [3, Chapter 5]).

To state our results, we first define the following quadratic form for a function ff on \mathbb{N}:

𝒜(f,f):=k(f(k+1)f(k))2ck+1ck.{\mathscr{A}}(f,f):=\sum_{k\in\mathbb{N}}\frac{(f(k+1)-f(k))^{2}}{c_{k+1}-c_{k}}.

If 𝒜(f,f)<{\mathscr{A}}(f,f)<\infty, it follows from c=limkck<c_{\infty}=\lim_{k\rightarrow\infty}c_{k}<\infty and the Cauchy-Schwarz inequality that f():=limkf(k)f(\infty):=\lim_{k\rightarrow\infty}f(k) exists.

Theorem 6.2.

If p2>0p_{2}>0 and |p4|=0|p_{4}|=0, then the parameters (γ,β,ν)(\gamma,\beta,\nu) determining the resolvent matrix of the QQ-process X=Ξ(Yˇ)X=\Xi(\check{Y}) are

γ=p1,β=2p2,νn=0,n.\gamma=p_{1},\quad\beta=2p_{2},\quad\nu_{n}=0,\;n\in\mathbb{N}. (6.2)

Furthermore, XX is symmetric with respect to the speed measure μ\mu, and the associated Dirichlet form on L2({},μ)L^{2}(\mathbb{N}\cup\{\infty\},\mu) is given by

ˇ={fL2({},μ):𝒜(f,f)<}\displaystyle\check{{\mathscr{F}}}=\{f\in L^{2}(\mathbb{N}\cup\{\infty\},\mu):{\mathscr{A}}(f,f)<\infty\}
ˇ(f,g)=12k(f(k+1)f(k))(g(k+1)g(k))ck+1ck+p12p2f()g(),f,gˇ.\displaystyle\check{{\mathscr{E}}}(f,g)=\frac{1}{2}\sum_{k\in\mathbb{N}}\frac{(f(k+1)-f(k))(g(k+1)-g(k))}{c_{k+1}-c_{k}}+\frac{p_{1}}{2p_{2}}f(\infty)g(\infty),\quad f,g\in\check{{\mathscr{F}}}.
Proof.

In this case, the Revuz measure of the PCAF (4.5) with respect to YY is exactly μ~:=kμkδc^k\tilde{\mu}:=\sum_{k\in\mathbb{N}}\mu_{k}\delta_{\hat{c}_{k}}, by Definition 3.8 and [3, Proposition 4.1.10]. According to [3, Theorem 5.2.2], the time-changed process Yˇ\check{Y} is symmetric with respect to μ~\tilde{\mu}, and its associated Dirichlet form (~,~)(\tilde{\mathscr{E}},\tilde{\mathscr{F}}) on L2(E¯,μ~)L^{2}(\overline{E},\tilde{\mu}) is derived in [3, (5.2.4)].

In what follows, we will compute (~,~)(\tilde{{\mathscr{E}}},\tilde{{\mathscr{F}}}). Regarding ~\tilde{{\mathscr{F}}}, we note that the extended Dirichlet space of (3.13) is

He1([0,)):={f is absolutely continuous on [0,) and fL2([0,))},H^{1}_{e}([0,\infty)):=\{f\text{ is absolutely continuous on }[0,\infty)\text{ and }f^{\prime}\in L^{2}([0,\infty))\},

as detailed in [3, Theorem 2.2.11]. According to [3, (5.2.4)], it is straightforward to compute that

~=He1([0,))|E¯L2(E¯,μ~)={Ξ(f):fˇ},\tilde{{\mathscr{F}}}=H^{1}_{e}([0,\infty))|_{\overline{E}}\cap L^{2}(\overline{E},\tilde{\mu})=\{\Xi(f):f\in\check{{\mathscr{F}}}\}, (6.3)

where Ξ(f)(c^n):=f(n)\Xi(f)(\hat{c}_{n}):=f(n) for all nn\in\mathbb{N}. The quadratic form ~\tilde{{\mathscr{E}}} can be formulated using [3, Theorem 5.5.9]. The crucial step is to compute the Feller measure UU on E¯×E¯\overline{E}\times\overline{E} and the supplementary Feller measure VV on E¯\overline{E}, as defined in [3, (5.5.7)]. In fact, by mimicking the proof of [20, Theorem 2.1], we can show that the strongly local part of ~\tilde{{\mathscr{E}}} vanishes, and UU is supported on {(c^n,c^n+1),(c^n+1,c^n):n}E¯×E¯\{(\hat{c}_{n},\hat{c}_{n+1}),(\hat{c}_{n+1},\hat{c}_{n}):n\in\mathbb{N}\}\subset\overline{E}\times\overline{E} (with the notation (,)(\cdot,\cdot) indicating a pair of points, not an interval) with

U({(c^n,c^n+1)})=U({(c^n+1,c^n)})=12|c^nc^n+1|=12|cn+1cn|.U(\{(\hat{c}_{n},\hat{c}_{n+1})\})=U(\{(\hat{c}_{n+1},\hat{c}_{n})\})=\frac{1}{2|\hat{c}_{n}-\hat{c}_{n+1}|}=\frac{1}{2|c_{n+1}-c_{n}|}.

Regarding the measure VV, we note that 𝐏x(τζ)=0\mathbf{P}_{x}(\tau\geq\zeta)=0 for any x[0,)E¯x\in[0,\infty)\setminus\overline{E}, where τ:=inf{t[0,ζ]:Yt[0,)E¯}\tau:=\inf\{t\in[0,\zeta]:Y_{t}\notin[0,\infty)\setminus\overline{E}\}. Thus, by the definition [3, (5.5.7)] of VV, we have V0V\equiv 0. Therefore, applying [3, Theorem 5.5.9], we obtain that for u~u\in\tilde{{\mathscr{F}}},

~(u,u)=12k(u(c^k+1)u(c^k))2ck+1ck+p12p2u(0)2.\tilde{\mathscr{E}}(u,u)=\frac{1}{2}\sum_{k\in\mathbb{N}}\frac{(u(\hat{c}_{k+1})-u(\hat{c}_{k}))^{2}}{c_{k+1}-c_{k}}+\frac{p_{1}}{2p_{2}}u(0)^{2}. (6.4)

Applying the spatial transformation Ξ\Xi to (6.3) and (6.4) yields the expression for (ˇ,ˇ)(\check{{\mathscr{E}}},\check{{\mathscr{F}}}).

Finally, the parameters (γ,β,ν)(\gamma,\beta,\nu) are given by (6.2), as discussed in [18, Lemma 7.1 and Theorem 8.1]. ∎

6.3. Non-symmetric case with finite jumping measure

We turn to examine the non-symmetric case where p2>0p_{2}>0 and 0<|p4|<0<|p_{4}|<\infty. The sample path representation of Feller’s Brownian motion for this case is detailed in §3.2.2. Let π\pi be a probability measure on {}\mathbb{N}\cup\{\partial\} given by

π({}):=p1p1+|p4|,π({n})=πn:=𝔭np1+|p4|,n,\pi(\{\partial\}):=\frac{p_{1}}{p_{1}+|p_{4}|},\quad\pi(\{n\})=\pi_{n}:=\frac{\mathfrak{p}_{n}}{p_{1}+|p_{4}|},\quad n\in\mathbb{N}, (6.5)

where {𝔭n}\{\mathfrak{p}_{n}\} is specified by (5.1) and (5.2).

Theorem 6.3.

If p2>0p_{2}>0 and 0<|p4|<0<|p_{4}|<\infty, then the parameters (γ,β,ν)(\gamma,\beta,\nu) determining the resolvent matrix of the QQ-process X=Ξ(Yˇ)X=\Xi(\check{Y}) are

γ=p1,β=2p2,νn=𝔭n,n.\gamma=p_{1},\quad\beta=2p_{2},\quad\nu_{n}=\mathfrak{p}_{n},\quad n\in\mathbb{N}. (6.6)

Furthermore, XX can be obtained by piecing out X1X^{1} with respect to the probability measure π\pi on {}\mathbb{N}\cup\{\partial\}, where X1X^{1} is the symmetric QQ-process corresponding to the parameters (p1+|p4|,2p2,0)(p_{1}+|p_{4}|,2p_{2},0) and π\pi is defined as (6.5).

Proof.

Let Y1Y^{1}, with lifetime ζ1\zeta^{1}, be the symmetric Feller’s Brownian motion with parameters (p1+|p4|,p2,p3,0)(p_{1}+|p_{4}|,p_{2},p_{3},0), as discussed in §3.2.2. It represents the killed process of YY at ζ1\zeta^{1}, and YY can be obtained by piecing out Y1Y^{1} with respect to the probability measure λ\lambda given by (3.20). Particularly, we have

𝐏x(Yζ1)=λ(),x[0,).\mathbf{P}_{x}(Y_{\zeta^{1}}\in\cdot)=\lambda(\cdot),\quad\forall x\in[0,\infty).

By following the steps used to obtain Y^0=Yˇ0\hat{Y}^{0}=\check{Y}^{0} in the proof of Theorem 4.2, we can also show that the killed process Yˇ1\check{Y}^{1} of Yˇ\check{Y} at time ζˇ1:=Aζ1\check{\zeta}^{1}:=A_{\zeta^{1}} is identical in law to the time-changed process of Y1Y^{1} with respect to the PCAF At1:=Atζ1,t0A^{1}_{t}:=A_{t\wedge\zeta^{1}},t\geq 0. Note that the Revuz measure of A1A^{1} with respect to Y1Y^{1} is exactly nμnδc^n\sum_{n\in\mathbb{N}}\mu_{n}\delta_{\hat{c}_{n}}. Hence, according to Theorem 6.2, Ξ(Yˇ1)\Xi(\check{Y}^{1}) is identical in law to X1X^{1}.

Denote by R~α1\tilde{R}_{\alpha}^{1} and R~α\tilde{R}_{\alpha} the resolvents of Yˇ1\check{Y}^{1} and Yˇ\check{Y}, respectively. For a bounded function ff on E¯\overline{E}, Dynkin’s formula gives us:

R~αf(c^)=𝐄c^0eαtf(Yˇt)𝑑t=R~α1f(c^)+𝐄c^(eαζˇ1R~αf(Yˇζˇ1)),c^E¯.\tilde{R}_{\alpha}f(\hat{c})=\mathbf{E}_{\hat{c}}\int_{0}^{\infty}e^{-\alpha t}f(\check{Y}_{t})dt=\tilde{R}^{1}_{\alpha}f(\hat{c})+\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\tilde{R}_{\alpha}f(\check{Y}_{\check{\zeta}^{1}})\right),\quad\forall\hat{c}\in\overline{E}. (6.7)

Let σ:=inf{t>0:YtE}\sigma:=\inf\{t>0:Y_{t}\in E\} and σ1:=inf{t>ζ1:YtE}\sigma_{1}:=\inf\{t>\zeta^{1}:Y_{t}\in E\}. Then,

γζˇ1=inf{t>0:At>Aζ1}=σ1.\gamma_{\check{\zeta}^{1}}=\inf\{t>0:A_{t}>A_{\zeta^{1}}\}=\sigma_{1}.

Note that Yσ1=Yσθζ1Y_{\sigma_{1}}=Y_{\sigma}\circ\theta_{\zeta^{1}}. According to the construction procedures of piecing out as stated in [18, Appendix A], it is straightforward to see that Yζ1Y_{\zeta^{1}} (with distribution λ\lambda) is independent of Y1Y^{1}. Thus, Yζ1Y_{\zeta^{1}} is also independent of ζˇ1=Aζ1\check{\zeta}^{1}=A_{\zeta^{1}}. From this independence and the strong Markov property of YY, it follows that

𝐄c^(eαζˇ1R~αf(Yˇζˇ1))\displaystyle\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\tilde{R}_{\alpha}f(\check{Y}_{\check{\zeta}^{1}})\right) =𝐄c^(eαζˇ1R~αf(Yσθζ1))\displaystyle=\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\tilde{R}_{\alpha}f(Y_{\sigma}\circ\theta_{\zeta^{1}})\right) (6.8)
=𝐄c^(eαζˇ1𝐄Yζ1(R~αf(Yσ)))\displaystyle=\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\mathbf{E}_{Y_{\zeta^{1}}}\left(\tilde{R}_{\alpha}f(Y_{\sigma})\right)\right)
=𝐄c^(eαζˇ1)𝐄λ(R~αf(Yσ)).\displaystyle=\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\right)\cdot\mathbf{E}_{\lambda}\left(\tilde{R}_{\alpha}f(Y_{\sigma})\right).

Similar to (6.4), we can also deduce that 𝐏λ(Yσ)=π~()\mathbf{P}_{\lambda}(Y_{\sigma}\in\cdot)=\tilde{\pi}(\cdot), where π~({c^n}):=πn\tilde{\pi}(\{\hat{c}_{n}\}):=\pi_{n} for all nn\in\mathbb{N} and π~({}):=π({})\tilde{\pi}(\{\partial\}):=\pi(\{\partial\}). Substituting (6.8) into (6.7) yields

R~αf(c^)=R~α1f(c^)+𝐄c^(eαζˇ1)π~(R~αf).\tilde{R}_{\alpha}f(\hat{c})=\tilde{R}^{1}_{\alpha}f(\hat{c})+\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\right)\cdot\tilde{\pi}(\tilde{R}_{\alpha}f).

Integrating both sides with respect to π~\tilde{\pi}, we get

π~(R~αf)=π~(R~α1f)1𝐄π~(eαζˇ1).\tilde{\pi}(\tilde{R}_{\alpha}f)=\frac{\tilde{\pi}(\tilde{R}^{1}_{\alpha}f)}{1-\mathbf{E}_{\tilde{\pi}}\left(e^{-\alpha\check{\zeta}^{1}}\right)}.

It follows that

R~αf(c^)=R~α1f(c^)+π~(R~α1f)1𝐄π~(eαζˇ1)𝐄c^(eαζˇ1),c^E¯.\tilde{R}_{\alpha}f(\hat{c})=\tilde{R}^{1}_{\alpha}f(\hat{c})+\frac{\tilde{\pi}(\tilde{R}^{1}_{\alpha}f)}{1-\mathbf{E}_{\tilde{\pi}}\left(e^{-\alpha\check{\zeta}^{1}}\right)}\cdot\mathbf{E}_{\hat{c}}\left(e^{-\alpha\check{\zeta}^{1}}\right),\quad\forall\hat{c}\in\overline{E}.

Finally, repeating the argument in the proof of [18, Theorem 8.1] (the steps after (8.5)) and applying the spatial transformation to Yˇ\check{Y}, we can conclude that XX is the piecing out of X1X^{1} with respect to π\pi. Particularly, the parameters (γ,β,ν)(\gamma,\beta,\nu) for XX are given by (6.6). ∎

7. Approximation of Feller’s Brownian motion

We now turn to the ‘pathological’ case where |p4|=|p_{4}|=\infty. Our primary method involves constructing a sequence of Markov processes that converge to Feller’s Brownian motion, using the strategy outlined in the construction of QQ-processes in [26, §6.1-§6.6]. Each constituent process is the piecing out of the absorbing Brownian motion with respect to a specific probability measure λ(n)\lambda^{(n)} on (0,){}(0,\infty)\cup\{\partial\}, whose resolvent is expressed as (3.1) with λ=λ(n)\lambda=\lambda^{(n)}. We refer to these processes as Doob’s Brownian motions, with λ(n)\lambda^{(n)} termed the instantaneous distribution (of the piecing out transformation).

7.1. Approximating sequence of Doob’s Brownian motions

We begin by introducing an important transformation on the sample paths. Let x(t)x(t) be a right-continuous function on [0,){}[0,\infty)\cup\{\partial\}. Consider two sequences of positive constants (αm)(\alpha_{m}) and (βm)(\beta_{m}) such that

0(=:β0)<α1β1<α2β2<.0(=:\beta_{0})<\alpha_{1}\leq\beta_{1}<\alpha_{2}\leq\beta_{2}<\cdots.

(These sequences may consist of finite numbers.) We say that the function y(t)y(t) is obtained from x(t)x(t) by the C(αm,βm)C(\alpha_{m},\beta_{m})-transformation if

y(t)=x(t),\displaystyle y(t)=x(t), 0t<α1,\displaystyle\quad 0\leq t<\alpha_{1},
y(dm+t)=x(βm+t),\displaystyle y(d_{m}+t)=x(\beta_{m}+t), 0t<αm+1βm,\displaystyle\quad 0\leq t<\alpha_{m+1}-\beta_{m},

where d1:=α1d_{1}:=\alpha_{1} and dm+1:=dm+(αm+1βm)d_{m+1}:=d_{m}+(\alpha_{m+1}-\beta_{m}). Intuitively speaking, the C(αm,βm)C(\alpha_{m},\beta_{m})-transformation discards the path of x(t)x(t) corresponding to the interval [αm,βm)[\alpha_{m},\beta_{m}), keeps the segment [0,α1)[0,\alpha_{1}) unchanged, and shifts the remaining parts to the left, connecting them in the original order without intersection, thereby obtaining a new right-continuous path y(t)y(t).

Fix nn\in\mathbb{N}. Define η:=inf{t>0:Yt=0}\eta:=\inf\{t>0:Y_{t-}=0\} and σ(n):=inf{t>0:Ytc^n}ζ\sigma^{(n)}:=\inf\{t>0:Y_{t}\geq\hat{c}_{n}\}\wedge\zeta (inf:=\inf\emptyset:=\infty). We then define a sequence of stopping times as follows:

η1(n):=η,σ1(n):=inf{tη1(n):Y(t)c^n}ζ,\eta^{(n)}_{1}:=\eta,\quad\sigma^{(n)}_{1}:=\inf\{t\geq\eta_{1}^{(n)}:Y(t)\geq\hat{c}_{n}\}\wedge\zeta,

and if ηm1(n),σm1(n)\eta^{(n)}_{m-1},\sigma^{(n)}_{m-1} are already defined, we set

ηm(n):=inf{tσm1(n):Yt=0}ζ\eta^{(n)}_{m}:=\inf\{t\geq\sigma^{(n)}_{m-1}:Y_{t-}=0\}\wedge\zeta

and

σm(n):=inf{tηm(n):Y(t)c^n}ζ.\sigma^{(n)}_{m}:=\inf\{t\geq\eta^{(n)}_{m}:Y(t)\geq\hat{c}_{n}\}\wedge\zeta.

Note that if ηm(n)<ζ\eta^{(n)}_{m}<\zeta, then Yηm(n)=0Y_{\eta^{(n)}_{m}}=0 due to the quasi-left-continuity of YY. Particularly, η=τ0=inf{t>0:Yt=0}\eta=\tau_{0}=\inf\{t>0:Y_{t}=0\}. We define η\eta using the left limit of YY instead of directly using τ0\tau_{0} because, although our focus is on Feller’s Brownian motion, the discussion in this section also applies to Doob’s Brownian motion. For Doob’s Brownian motion, defining the ‘return’ time to 0 requires the use of the left limit.

For nn\in\mathbb{N} and every ωΩ\omega\in\Omega, by performing the C(ηm(n)(ω),σm(n)(ω))C(\eta^{(n)}_{m}(\omega),\sigma^{(n)}_{m}(\omega))-transformation on Yt(ω)Y_{t}(\omega), we obtain a new path, denoted by Yt(n)(ω)Y^{(n)}_{t}(\omega).

Lemma 7.1.

The process

Y(n):=(Ω,𝒢,Yt(n),(𝐏x)x(0,))Y^{(n)}:=\left(\Omega,{\mathscr{G}},Y^{(n)}_{t},(\mathbf{P}_{x})_{x\in(0,\infty)}\right)

is a Doob’s Brownian motion with instantaneous distribution

λ(n)()=𝐏0(Yσ(n))\lambda^{(n)}(\cdot)=\mathbf{P}_{0}(Y_{\sigma^{(n)}}\in\cdot)

supported on [c^n,){}[\hat{c}_{n},\infty)\cup\{\partial\}.

Proof.

Fix x(0,)x\in(0,\infty) and α>0\alpha>0. We aim to compute the resolvent of Y(n)Y^{(n)} for a positive and bounded Borel measurable function ff on (0,)(0,\infty):

Gα(n)f(x)=𝐄x0eαtf(Yt(n))𝑑t.G^{(n)}_{\alpha}f(x)=\mathbf{E}_{x}\int_{0}^{\infty}e^{-\alpha t}f(Y^{(n)}_{t})dt.

Let ζ1:=η1(n)=η\zeta_{1}:=\eta^{(n)}_{1}=\eta and ζm:=ηm(n)σm1(n)\zeta_{m}:=\eta^{(n)}_{m}-\sigma^{(n)}_{m-1} for m2m\geq 2. By the definition of Yt(n)Y^{(n)}_{t}, we have

Gα(n)f(x)=𝐄x0ζ1eαtf(Yt)𝑑t+m1𝐄xi=1mζii=1m+1ζieαtf(Yti=1mζi+σm(n))𝑑t.G^{(n)}_{\alpha}f(x)=\mathbf{E}_{x}\int_{0}^{\zeta_{1}}e^{-\alpha t}f(Y_{t})dt+\sum_{m\geq 1}\mathbf{E}_{x}\int_{\sum_{i=1}^{m}\zeta_{i}}^{\sum_{i=1}^{m+1}\zeta_{i}}e^{-\alpha t}f(Y_{t-\sum_{i=1}^{m}\zeta_{i}+\sigma^{(n)}_{m}})dt. (7.1)

Set

Jm\displaystyle J_{m} :=𝐄xi=1mζii=1m+1ζieαtf(Yti=1mζi+σm(n))𝑑t\displaystyle:=\mathbf{E}_{x}\int_{\sum_{i=1}^{m}\zeta_{i}}^{\sum_{i=1}^{m+1}\zeta_{i}}e^{-\alpha t}f(Y_{t-\sum_{i=1}^{m}\zeta_{i}+\sigma^{(n)}_{m}})dt
=𝐄xeαi=1mζi0ζm+1eαtf(Yt+σm(n))𝑑t.\displaystyle=\mathbf{E}_{x}e^{-\alpha\sum_{i=1}^{m}\zeta_{i}}\int_{0}^{\zeta_{m+1}}e^{-\alpha t}f(Y_{t+\sigma^{(n)}_{m}})dt.

Note that ζm+1=ηθσm(n)\zeta_{m+1}=\eta\circ\theta_{\sigma^{(n)}_{m}}. It follows from the strong Markov property that

Jm\displaystyle J_{m} =𝐄x(eαi=1mζi𝐄Yσm(n)0ηeαtf(Yt)𝑑t)\displaystyle=\mathbf{E}_{x}\left(e^{-\alpha\sum_{i=1}^{m}\zeta_{i}}\mathbf{E}_{Y_{\sigma^{(n)}_{m}}}\int_{0}^{\eta}e^{-\alpha t}f(Y_{t})dt\right)
=𝐄x(eαi=1mζiGα0f(Yσm(n))),\displaystyle=\mathbf{E}_{x}\left(e^{-\alpha\sum_{i=1}^{m}\zeta_{i}}G^{0}_{\alpha}f(Y_{\sigma^{(n)}_{m}})\right),

where Gα0G^{0}_{\alpha} is the resolvent of the absorbing Brownian motion.

Let us prove that

Jm=𝐄xeαη(𝐄λ(n)eαη)m10Gα0f(x)λ(n)(dx),m1.J_{m}=\mathbf{E}_{x}e^{-\alpha\eta}\cdot\left(\mathbf{E}_{\lambda^{(n)}}e^{-\alpha\eta}\right)^{m-1}\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx),\quad m\geq 1. (7.2)

For m=1m=1, we have Yσ1(n)=Yσ(n)θηY_{\sigma^{(n)}_{1}}=Y_{\sigma^{(n)}}\circ\theta_{\eta}, and thus, by the strong Markov property of YY,

J1\displaystyle J_{1} =𝐄x(eαηGα0(Yσ(n)θη))=𝐄x(eαη𝐄Yη(Gα0(Yσ(n))))\displaystyle=\mathbf{E}_{x}\left(e^{-\alpha\eta}G^{0}_{\alpha}(Y_{\sigma^{(n)}}\circ\theta_{\eta})\right)=\mathbf{E}_{x}\left(e^{-\alpha\eta}\mathbf{E}_{Y_{\eta}}\left(G^{0}_{\alpha}\left(Y_{\sigma^{(n)}}\right)\right)\right)
=𝐄xeαη𝐄0(Gα0(Yσ(n)))=𝐄xeαη0Gα0f(x)λ(n)(dx).\displaystyle=\mathbf{E}_{x}e^{-\alpha\eta}\cdot\mathbf{E}_{0}\left(G^{0}_{\alpha}\left(Y_{\sigma^{(n)}}\right)\right)=\mathbf{E}_{x}e^{-\alpha\eta}\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx).

In general, note that Yσi(n)=Yσ(n)θηi(n)Y_{\sigma^{(n)}_{i}}=Y_{\sigma^{(n)}}\circ\theta_{\eta^{(n)}_{i}} and ζi=ηi(n)σi1(n)=ηθσi1(n)\zeta_{i}=\eta^{(n)}_{i}-\sigma^{(n)}_{i-1}=\eta\circ\theta_{\sigma^{(n)}_{i-1}} for all i2i\geq 2. Then the case m=2m=2 can be deduced by the strong Markov property as follows:

J2\displaystyle J_{2} =𝐄x(eα(ζ1+ζ2)𝐄Yη2(n)(Gα0(Yσ(n))))\displaystyle=\mathbf{E}_{x}\left(e^{-\alpha(\zeta_{1}+\zeta_{2})}\mathbf{E}_{Y_{\eta^{(n)}_{2}}}\left(G^{0}_{\alpha}\left(Y_{\sigma^{(n)}}\right)\right)\right)
=𝐄xeα(η+ηθσ1(n))0Gα0f(x)λ(n)(dx)\displaystyle=\mathbf{E}_{x}e^{-\alpha(\eta+\eta\circ\theta_{\sigma^{(n)}_{1}})}\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx)
=𝐄x(eαη𝐄Yσ1(n)(eαη))0Gα0f(x)λ(n)(dx).\displaystyle=\mathbf{E}_{x}\left(e^{-\alpha\eta}\mathbf{E}_{Y_{\sigma^{(n)}_{1}}}\left(e^{-\alpha\eta}\right)\right)\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx).

Using Yσ1(n)=Yσ(n)θηY_{\sigma^{(n)}_{1}}=Y_{\sigma^{(n)}}\circ\theta_{\eta} and Yη=0Y_{\eta}=0, we obtain that

J2\displaystyle J_{2} =𝐄x(eαη𝐄Yσ(n)θη(eαη))0Gα0f(x)λ(n)(dx)\displaystyle=\mathbf{E}_{x}\left(e^{-\alpha\eta}\mathbf{E}_{Y_{\sigma^{(n)}}\circ\theta_{\eta}}\left(e^{-\alpha\eta}\right)\right)\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx)
=𝐄xeαη𝐄0(𝐄Yσ(n)eαη)0Gα0f(x)λ(n)(dx)\displaystyle=\mathbf{E}_{x}e^{-\alpha\eta}\cdot\mathbf{E}_{0}\left(\mathbf{E}_{Y_{\sigma^{(n)}}}e^{-\alpha\eta}\right)\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx)
=𝐄xeαη𝐄λ(n)(eαη)0Gα0f(x)λ(n)(dx).\displaystyle=\mathbf{E}_{x}e^{-\alpha\eta}\cdot\mathbf{E}_{\lambda^{(n)}}\left(e^{-\alpha\eta}\right)\cdot\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx).

The general case can be formulated similarly.

Finally, by substituting (7.2) into (7.1), we conclude that

Gα(n)f(x)=Gα0f(x)+𝐄xeαη0Gα0f(x)λ(n)(dx)1𝐄λ(n)(eαη).G^{(n)}_{\alpha}f(x)=G_{\alpha}^{0}f(x)+\mathbf{E}_{x}e^{-\alpha\eta}\frac{\int_{0}^{\infty}G^{0}_{\alpha}f(x)\lambda^{(n)}(dx)}{1-\mathbf{E}_{\lambda^{(n)}}\left(e^{-\alpha\eta}\right)}. (7.3)

This result corresponds precisely to the resolvent of the Doob’s Brownian motion with instantaneous distribution λ(n)\lambda^{(n)}. ∎

The sequence of Doob’s Brownian motions Y(n)Y^{(n)} approximates YY in the case where p3=0p_{3}=0 in the following sense.

Theorem 7.2.

Assume that p3=0p_{3}=0. Then, for any x(0,)x\in(0,\infty),

𝐏x(limnYt(n)=Yt)=1,t0.\mathbf{P}_{x}(\lim_{n\rightarrow\infty}Y^{(n)}_{t}=Y_{t})=1,\quad\forall t\geq 0. (7.4)
Proof.

A crucial consequence of the assumption p3=0p_{3}=0 is that

|{t0:Yt=0}|=0,𝐏x-a.s.,\left|\{t\geq 0:Y_{t}=0\}\right|=0,\quad\mathbf{P}_{x}\text{-a.s.},

where |||\cdot| denotes the Lebesgue measure of the time set; see, e.g., [10, §14]. This can also be observed from the pathwise representation (3.23), as Yt=0Y_{t}=0 always indicates Wt+=0W^{+}_{t}=0. Hence, in principle, we may repeat the proof of Theorem A.1 step by step to complete the proof. Below, we provide a concise proof using the pathwise representation of YtY_{t}.

Fix t0t\geq 0. Let ρt(n)(ω)\rho^{(n)}_{t}(\omega) denote the total duration discarded from the path of Y(ω)Y_{\cdot}(\omega) when constructing Ys(n)Y^{(n)}_{s} for 0st0\leq s\leq t. Specifically, we have Yt(n)(ω)=Yt+ρt(n)(ω)(ω)Y^{(n)}_{t}(\omega)=Y_{{t+\rho^{(n)}_{t}}(\omega)}(\omega). It suffices to show

𝐏x(limnρt(n)=0)=1.\mathbf{P}_{x}(\lim_{n\rightarrow\infty}\rho^{(n)}_{t}=0)=1. (7.5)

Let us first consider the special case where YY is the reflecting Brownian motion W+W^{+}. To differentiate this specific case from the general one, we denote η,σ(n),ηm(n),σm(n),ρt(n)\eta,\sigma^{(n)},\eta^{(n)}_{m},\sigma^{(n)}_{m},\rho^{(n)}_{t} by η~,σ~(n),η~m(n),σ~m(n),ρ~t(n)\tilde{\eta},\tilde{\sigma}^{(n)},\tilde{\eta}^{(n)}_{m},\tilde{\sigma}^{(n)}_{m},\tilde{\rho}^{(n)}_{t}, respectively. Clearly, the instantaneous distribution of the approximating Doob’s Brownian motion, denoted by W+,(n)W^{+,(n)}, is δc^n\delta_{\hat{c}_{n}}. Since W+,(n)W^{+,(n)} is conservative by the expression of its resolvent in (3.1), we have ρ~t(n)<\tilde{\rho}^{(n)}_{t}<\infty for all nn. Note that ρ~t(n)\tilde{\rho}^{(n)}_{t} is decreasing as nn\rightarrow\infty. It follows that

ρ~t(n)(ω)\displaystyle\tilde{\rho}^{(n)}_{t}(\omega) |{0st+ρ~t(n)(ω):Ws+<c^n}|\displaystyle\leq\left|\left\{0\leq s\leq t+\tilde{\rho}^{(n)}_{t}(\omega):W^{+}_{s}<\hat{c}_{n}\right\}\right|
|{0st+ρ~t(0)(ω):Ws+<c^n}|,\displaystyle\leq\left|\left\{0\leq s\leq t+\tilde{\rho}^{(0)}_{t}(\omega):W^{+}_{s}<\hat{c}_{n}\right\}\right|,

which implies that

limnρ~t(n)(ω)\displaystyle\lim_{n\rightarrow\infty}\tilde{\rho}^{(n)}_{t}(\omega) limn|{0st+ρ~t(0)(ω):Ws+<c^n}|\displaystyle\leq\lim_{n\rightarrow\infty}\left|\left\{0\leq s\leq t+\tilde{\rho}^{(0)}_{t}(\omega):W^{+}_{s}<\hat{c}_{n}\right\}\right| (7.6)
=|{0st+ρ~t(0)(ω):Ws+=0}|=0.\displaystyle=\left|\left\{0\leq s\leq t+\tilde{\rho}^{(0)}_{t}(\omega):W^{+}_{s}=0\right\}\right|=0.

Now, consider a general Feller’s Brownian motion YY. According to its pathwise representation (3.23) (with lifetime ζ\zeta given in (3.27)), Yt=0Y_{t}=0 or Yt=0Y_{t-}=0 always implies Wt+=0W^{+}_{t}=0. Since YtWt+Y_{t}\geq W^{+}_{t}, it follows that σ~(n)σ(n)\tilde{\sigma}^{(n)}\geq\sigma^{(n)}. (Recall that the notation with tilde is defined for W+W^{+}.) Consequently, each discarding interval [ηm(n),σm(n))[\eta^{(n)}_{m},\sigma^{(n)}_{m}) for Y(n)Y^{(n)} must be contained within some discarding interval for W+,(n)W^{+,(n)} (though the converse is not necessarily true). Particularly, ρt(n)ρ~t(n)\rho^{(n)}_{t}\leq\tilde{\rho}^{(n)}_{t}. Therefore, (7.5) follows from (7.6). ∎

Remark 7.3.

If we do not assume that p3=0p_{3}=0, then the sequence of Doob’s Brownian motions Y(n)Y^{(n)} will converge to a Feller’s Brownian motion Y~\tilde{Y} with parameters (p1,p2,0,p4)(p_{1},p_{2},0,p_{4}) in the sense of (7.4). This is because, according to Lemma 8.2, YY can be obtained from Y~\tilde{Y}, with lifetime ζ~\tilde{\zeta}, through a time change transformation corresponding to a strictly increasing PCAF. Therefore, they share the same sequence of distributions λ(n)\lambda^{(n)}, that is, 𝐏0(Yσ(n))=𝐏0(Y~σ~(n))\mathbf{P}_{0}(Y_{\sigma^{(n)}}\in\cdot)=\mathbf{P}_{0}(\tilde{Y}_{\tilde{\sigma}^{(n)}}\in\cdot), where σ~(n):=inf{t>0:Y~tc^n}ζ~\tilde{\sigma}^{(n)}:=\inf\{t>0:\tilde{Y}_{t}\geq\hat{c}_{n}\}\wedge\tilde{\zeta}.

7.2. Representation of instantaneous distributions

The following lemma describes the relationship between the instantaneous distributions.

Lemma 7.4.

For every m,nm,n\in\mathbb{N} with m>nm>n, the following relationship holds:

λ(n)=[c^m,c^n)xλ(m)(dx)c^nδc^n+λ(m)|[c^n,){}1[c^m,c^n)(1xc^n)λ(m)(dx).\lambda^{(n)}=\frac{\frac{\int_{[\hat{c}_{m},\hat{c}_{n})}x\lambda^{(m)}(dx)}{\hat{c}_{n}}\delta_{\hat{c}_{n}}+\lambda^{(m)}|_{[\hat{c}_{n},\infty)\cup\{\partial\}}}{1-\int_{[\hat{c}_{m},\hat{c}_{n})}\left(1-\frac{x}{\hat{c}_{n}}\right)\lambda^{(m)}(dx)}. (7.7)
Proof.

Take a positive and bounded Borel measurable function ff on (0,)(0,\infty). Note that Yσ(n)=Yσ(n)θσ(m)Y_{\sigma^{(n)}}=Y_{\sigma^{(n)}}\circ\theta_{\sigma^{(m)}}, since σ(n)σ(m)\sigma^{(n)}\geq\sigma^{(m)}. Using the strong Markov property of YY, we have

λ(n)(f)=𝐄0f(Yσ(n)θσ(m))=𝐄0(𝐄Yσ(m)f(Yσ(n))).\displaystyle\lambda^{(n)}(f)=\mathbf{E}_{0}f(Y_{\sigma^{(n)}}\circ\theta_{\sigma^{(m)}})=\mathbf{E}_{0}\left(\mathbf{E}_{Y_{\sigma^{(m)}}}f(Y_{\sigma^{(n)}})\right). (7.8)

For yc^ny\geq\hat{c}_{n}, it holds 𝐏y\mathbf{P}_{y}-a.s. that Yσ(n)=Y0=yY_{\sigma^{(n)}}=Y_{0}=y. Thus,

𝐄0(𝐄Yσ(m)f(Yσ(n));Yσ(m)c^n)\displaystyle\mathbf{E}_{0}\left(\mathbf{E}_{Y_{\sigma^{(m)}}}f(Y_{\sigma^{(n)}});Y_{\sigma^{(m)}}\geq\hat{c}_{n}\right) =𝐄0(f(Yσ(m));Yσ(m)c^n)\displaystyle=\mathbf{E}_{0}\left(f(Y_{\sigma^{(m)}});Y_{\sigma^{(m)}}\geq\hat{c}_{n}\right) (7.9)
=[c^n,)f(x)λ(m)(dx).\displaystyle=\int_{[\hat{c}_{n},\infty)}f(x)\lambda^{(m)}(dx).

For c^my<c^n\hat{c}_{m}\leq y<\hat{c}_{n}, we have

𝐄yf(Yσ(n))=𝐄y(f(Yσ(n));σ(n)<η)+𝐄y(f(Yσ(n));σ(n)>η).\mathbf{E}_{y}f(Y_{\sigma^{(n)}})=\mathbf{E}_{y}\left(f(Y_{\sigma^{(n)}});\sigma^{(n)}<\eta\right)+\mathbf{E}_{y}\left(f(Y_{\sigma^{(n)}});\sigma^{(n)}>\eta\right). (7.10)

Since Yσ(n)=c^nY_{\sigma^{(n)}}=\hat{c}_{n} on {σ(n)<η}\{\sigma^{(n)}<\eta\}, it follows that

𝐄y(f(Yσ(n));σ(n)<η)=f(c^n)yc^n.\mathbf{E}_{y}\left(f(Y_{\sigma^{(n)}});\sigma^{(n)}<\eta\right)=f(\hat{c}_{n})\cdot\frac{y}{\hat{c}_{n}}. (7.11)

Using Yσ(n)=Yσ(n)θηY_{\sigma^{(n)}}=Y_{\sigma^{(n)}}\circ\theta_{\eta} on {σ(n)>η}\{\sigma^{(n)}>\eta\} and applying the strong Markov property of YY, we get

𝐄y(f(Yσ(n));σ(n)>η)\displaystyle\mathbf{E}_{y}\left(f(Y_{\sigma^{(n)}});\sigma^{(n)}>\eta\right) =𝐄y(𝐄Yηf(Yσ(n));σ(n)>η)\displaystyle=\mathbf{E}_{y}\left(\mathbf{E}_{Y_{\eta}}f(Y_{\sigma^{(n)}});\sigma^{(n)}>\eta\right) (7.12)
=𝐄0f(Yσ(n))𝐄y(σ(n)>η)\displaystyle=\mathbf{E}_{0}f(Y_{\sigma^{(n)}})\cdot\mathbf{E}_{y}(\sigma^{(n)}>\eta)
=λ(n)(f)(1yc^n).\displaystyle=\lambda^{(n)}(f)\cdot\left(1-\frac{y}{\hat{c}_{n}}\right).

Substituting (7.11) and (7.12) into (7.10) yields

𝐄yf(Yσ(n))=f(c^n)yc^n+λ(n)(f)(1yc^n),c^my<c^n.\mathbf{E}_{y}f(Y_{\sigma^{(n)}})=f(\hat{c}_{n})\cdot\frac{y}{\hat{c}_{n}}+\lambda^{(n)}(f)\cdot\left(1-\frac{y}{\hat{c}_{n}}\right),\quad\forall\hat{c}_{m}\leq y<\hat{c}_{n}.

Therefore,

𝐄0\displaystyle\mathbf{E}_{0} (𝐄Yσ(m)f(Yσ(n));c^mYσ(m)<c^n)\displaystyle\left(\mathbf{E}_{Y_{\sigma^{(m)}}}f(Y_{\sigma^{(n)}});\hat{c}_{m}\leq Y_{\sigma^{(m)}}<\hat{c}_{n}\right)
=𝐄0(f(c^n)Yσ(m)c^n+λ(n)(f)(1Yσ(m)c^n);c^mYσ(m)<c^n)\displaystyle=\mathbf{E}_{0}\left(f(\hat{c}_{n})\cdot\frac{Y_{\sigma^{(m)}}}{\hat{c}_{n}}+\lambda^{(n)}(f)\cdot\left(1-\frac{Y_{\sigma^{(m)}}}{\hat{c}_{n}}\right);\hat{c}_{m}\leq Y_{\sigma^{(m)}}<\hat{c}_{n}\right)
=[c^m,c^n)(f(c^n)yc^n+λ(n)(f)(1yc^n))λ(m)(dy).\displaystyle=\int_{[\hat{c}_{m},\hat{c}_{n})}\left(f(\hat{c}_{n})\cdot\frac{y}{\hat{c}_{n}}+\lambda^{(n)}(f)\cdot\left(1-\frac{y}{\hat{c}_{n}}\right)\right)\lambda^{(m)}(dy).

Combining this with (7.8) and (7.9), we can verify (7.7). ∎

Based on this lemma, we can define a measure λ\lambda on (0,){}(0,\infty)\cup\{\partial\} as follows:

λ|(c^n,){}:=λ(n)|(c^n,){}Λn,n,\lambda|_{(\hat{c}_{n},\infty)\cup\{\partial\}}:=\frac{\lambda^{(n)}|_{(\hat{c}_{n},\infty)\cup\{\partial\}}}{\Lambda_{n}},\quad n\in\mathbb{N}, (7.13)

where Λn:=1[c^n,c^0)(1xc^0)λ(n)(dx)\Lambda_{n}:=1-\int_{[\hat{c}_{n},\hat{c}_{0})}\left(1-\frac{x}{\hat{c}_{0}}\right)\lambda^{(n)}(dx). Clearly, λ\lambda is a well-defined σ\sigma-finite measure on (0,){}(0,\infty)\cup\{\partial\}. The following lemma provides the representation of λ(n)\lambda^{(n)} in terms of this measure λ\lambda.

Lemma 7.5.

For each nn\in\mathbb{N}, the following holds:

1Λn=λ((c^n,){})+c^0(1λ((c^0,){}))(c^n,c^0]xλ(dx)c^n,\frac{1}{\Lambda_{n}}=\lambda\left((\hat{c}_{n},\infty)\cup\{\partial\}\right)+\frac{\hat{c}_{0}\left(1-\lambda\left((\hat{c}_{0},\infty)\cup\{\partial\}\right)\right)-\int_{(\hat{c}_{n},\hat{c}_{0}]}x\lambda(dx)}{\hat{c}_{n}}, (7.14)

and

λ(n)|(c^n,){}=Λnλ|(c^n,){}\displaystyle\lambda^{(n)}|_{(\hat{c}_{n},\infty)\cup\{\partial\}}=\Lambda_{n}\cdot\lambda|_{(\hat{c}_{n},\infty)\cup\{\partial\}} (7.15)
λ(n)({c^n})=Λnc^n(c^0(1λ((c^0,){}))(c^n,c^0]xλ(dx)).\displaystyle\lambda^{(n)}(\{\hat{c}_{n}\})=\frac{\Lambda_{n}}{\hat{c}_{n}}\left(\hat{c}_{0}\left(1-\lambda\left((\hat{c}_{0},\infty)\cup\{\partial\}\right)\right)-\int_{(\hat{c}_{n},\hat{c}_{0}]}x\lambda(dx)\right).

Particularly,

(0,)(1x)λ(dx)<.\int_{(0,\infty)}\left(1\wedge x\right)\lambda(dx)<\infty. (7.16)
Proof.

According to Lemma 7.4, we have

λ(n)({c^n})=ΛnΛn+1([c^n+1,c^n)xc^nλ(n+1)(dx)+λ(n+1)({c^n})).\lambda^{(n)}(\{\hat{c}_{n}\})=\frac{\Lambda_{n}}{\Lambda_{n+1}}\cdot\left(\int_{[\hat{c}_{n+1},\hat{c}_{n})}\frac{x}{\hat{c}_{n}}\lambda^{(n+1)}(dx)+\lambda^{(n+1)}(\{\hat{c}_{n}\})\right).

It follows that

c^nλ(n)({c^n})Λn\displaystyle\frac{\hat{c}_{n}\lambda^{(n)}(\{\hat{c}_{n}\})}{\Lambda_{n}} =c^n+1λ(n+1)({c^n+1})Λn+1+(c^n+1,c^n)xλ(n+1)(dx)+c^nλ(n+1)({c^n})Λn+1\displaystyle=\frac{\hat{c}_{n+1}\lambda^{(n+1)}(\{\hat{c}_{n+1}\})}{\Lambda_{n+1}}+\frac{\int_{(\hat{c}_{n+1},\hat{c}_{n})}x\lambda^{(n+1)}(dx)+\hat{c}_{n}\lambda^{(n+1)}(\{\hat{c}_{n}\})}{\Lambda_{n+1}}
=c^n+1λ(n+1)({c^n+1})Λn+1+(c^n+1,c^n]xλ(dx).\displaystyle=\frac{\hat{c}_{n+1}\lambda^{(n+1)}(\{\hat{c}_{n+1}\})}{\Lambda_{n+1}}+\int_{(\hat{c}_{n+1},\hat{c}_{n}]}x\lambda(dx).

Let hn:=c^nλ(n)({c^n})Λnh_{n}:=\frac{\hat{c}_{n}\lambda^{(n)}(\{\hat{c}_{n}\})}{\Lambda_{n}}. Then hnh_{n} is decreasing in nn, and

h0limnhn=n=0(hnhn+1)=n=0(c^n+1,c^n]xλ(dx)=(0,c^0]xλ(dx).h_{0}-\lim_{n\rightarrow\infty}h_{n}=\sum_{n=0}(h_{n}-h_{n+1})=\sum_{n=0}^{\infty}\int_{(\hat{c}_{n+1},\hat{c}_{n}]}x\lambda(dx)=\int_{(0,\hat{c}_{0}]}x\lambda(dx).

Particularly, (7.16) holds true. The induction also yields that

hn=hn+m+(c^n+m,c^n]xλ(dx),m1,h_{n}=h_{n+m}+\int_{(\hat{c}_{n+m},\hat{c}_{n}]}x\lambda(dx),\quad\forall m\geq 1,

and by letting mm\rightarrow\infty, we obtain

hn=limmhm+(0,c^n]xλ(dx)=h0(c^n,c^0]xλ(dx).h_{n}=\lim_{m\rightarrow\infty}h_{m}+\int_{(0,\hat{c}_{n}]}x\lambda(dx)=h_{0}-\int_{(\hat{c}_{n},\hat{c}_{0}]}x\lambda(dx). (7.17)

Therefore, the first identity in (7.15) is established. The second identity in (7.15) follows directly from (7.13). Finally, substituting (7.15) into the definition of Λn\Lambda_{n}, we obtain (7.14). ∎

7.3. Parameters of Feller’s Brownian motion

The aim of this subsection is to express the parameters of the original Feller’s Brownian motion in terms of the measure λ\lambda, which is defined based on the approximating Doob’s Brownian motions.

Theorem 7.6.

Let YY be a Feller’s Brownian motion with parameters (p1,p2,0,p4)(p_{1},p_{2},0,p_{4}), and let λ\lambda be the measure (7.13) on (0,){}(0,\infty)\cup\{\partial\} in terms of the approximating Doob’s Brownian motions Y(n)Y^{(n)}. Then, up to a certain multiplicative constant needed to satisfy (3.3), it holds that

p1=λ({}),p2=c^0(1λ({}))(0,)(xc^0)λ(dx),p4=λ|(0,).p_{1}=\lambda(\{\partial\}),\quad p_{2}=\hat{c}_{0}\left(1-\lambda\left(\{\partial\}\right)\right)-\int_{(0,\infty)}\left(x\wedge\hat{c}_{0}\right)\lambda(dx),\quad p_{4}=\lambda|_{(0,\infty)}. (7.18)
Proof.

Define p~2n:=hn=c^0λ(0)({c^0})(c^n,c^0]xλ(dx)\tilde{p}^{n}_{2}:=h_{n}=\hat{c}_{0}\lambda^{(0)}(\{\hat{c}_{0}\})-\int_{(\hat{c}_{n},\hat{c}_{0}]}x\lambda(dx) for nn\in\mathbb{N} (see (7.17)). This sequence is decreasing and converges to

p~2:=c^0(1λ({}))(0,)(xc^0)λ(dx).\tilde{p}_{2}:=\hat{c}_{0}\left(1-\lambda\left(\{\partial\}\right)\right)-\int_{(0,\infty)}\left(x\wedge\hat{c}_{0}\right)\lambda(dx). (7.19)

According to (7.3) and Lemma 7.5, the resolvent of Y(n)Y^{(n)} can be expressed as

Gα(n)\displaystyle G^{(n)}_{\alpha} f(x)\displaystyle f(x)
=Gα0f(x)+𝐄xeαη(c^n,)Gα0f(x)λ(dx)+p~2nGα0f(c^n)c^n1Λn(c^n,)e2αxλ(dx)p~2ne2αc^nc^n\displaystyle=G^{0}_{\alpha}f(x)+\mathbf{E}_{x}e^{-\alpha\eta}\cdot\frac{\int_{(\hat{c}_{n},\infty)}G^{0}_{\alpha}f(x)\lambda(dx)+\tilde{p}^{n}_{2}\cdot\frac{G^{0}_{\alpha}f(\hat{c}_{n})}{\hat{c}_{n}}}{\frac{1}{\Lambda_{n}}-\int_{(\hat{c}_{n},\infty)}e^{-\sqrt{2\alpha}x}\lambda(dx)-\tilde{p}^{n}_{2}\cdot\frac{e^{-\sqrt{2\alpha}\hat{c}_{n}}}{\hat{c}_{n}}}
=Gα0f(x)+𝐄xeαη(c^n,)Gα0f(x)λ(dx)+p~2nGα0f(c^n)c^nλ({})+(c^n,)(1e2αx)λ(dx)+p~2n1e2αc^nc^n\displaystyle=G^{0}_{\alpha}f(x)+\mathbf{E}_{x}e^{-\alpha\eta}\cdot\frac{\int_{(\hat{c}_{n},\infty)}G^{0}_{\alpha}f(x)\lambda(dx)+\tilde{p}^{n}_{2}\cdot\frac{G^{0}_{\alpha}f(\hat{c}_{n})}{\hat{c}_{n}}}{\lambda(\{\partial\})+\int_{(\hat{c}_{n},\infty)}(1-e^{-\sqrt{2\alpha}x})\lambda(dx)+\tilde{p}^{n}_{2}\cdot\frac{1-e^{-\sqrt{2\alpha}\hat{c}_{n}}}{\hat{c}_{n}}}

for all fCc((0,))f\in C_{c}((0,\infty)) and x(0,)x\in(0,\infty). Note that

limnGα0f(c^n)c^n=limx0Gα0f(x)x=2(0,)e2αxf(x)𝑑x\lim_{n\rightarrow\infty}\frac{G^{0}_{\alpha}f(\hat{c}_{n})}{\hat{c}_{n}}=\lim_{x\rightarrow 0}\frac{G^{0}_{\alpha}f(x)}{x}=2\int_{(0,\infty)}e^{-\sqrt{2\alpha}x}f(x)dx

by virtue of the expression of Gα0G^{0}_{\alpha} provided in the first paragraph of the proof of Corollary 3.5. Additionally,

limn1e2αc^nc^n=2α.\lim_{n\rightarrow\infty}\frac{1-e^{-\sqrt{2\alpha}\hat{c}_{n}}}{\hat{c}_{n}}=\sqrt{2\alpha}.

Given (7.16), we have

limn\displaystyle\lim_{n\rightarrow\infty} Gα(n)f(x)\displaystyle G^{(n)}_{\alpha}f(x)
=Gα0f(x)+𝐄xeαη(0,)Gα0f(x)λ(dx)+2p~2(0,)e2αxf(x)𝑑xλ({})+(0,)(1e2αx)λ(dx)+p~22α.\displaystyle=G^{0}_{\alpha}f(x)+\mathbf{E}_{x}e^{-\alpha\eta}\cdot\frac{\int_{(0,\infty)}G^{0}_{\alpha}f(x)\lambda(dx)+2\tilde{p}_{2}\cdot\int_{(0,\infty)}e^{-\sqrt{2\alpha}x}f(x)dx}{\lambda(\{\partial\})+\int_{(0,\infty)}(1-e^{-\sqrt{2\alpha}x})\lambda(dx)+\tilde{p}_{2}\cdot\sqrt{2\alpha}}.

By Theorem 7.2, this limit is exactly the resolvent of YY. Comparing this expression with (3.15), we can eventually derive (7.18). ∎

8. Proof of Theorem 5.1 for |p4|=|p_{4}|=\infty

8.1. No sojourn case p3=0p_{3}=0

In this subsection, we consider the case where |p4|=|p_{4}|=\infty and further assume that p3=0p_{3}=0. According to Theorem 7.2, we can construct a sequence of Doob’s Brownian motions Y(n)Y^{(n)} that approximates the Feller’s Brownian motion YY. Their instantaneous distributions λ(n)\lambda^{(n)} are characterized in Lemma 7.5 in terms of the measure λ\lambda defined by (7.13).

Regarding the time-changed Feller’s Brownian motion X^:=Yˇ\hat{X}:=\check{Y}, we can also construct a sequence of Doob processes X^(n)\hat{X}^{(n)} with instantaneous distribution λ^(n)\hat{\lambda}^{(n)}, as described in Appendix A, that approximates X^\hat{X}.

Lemma 8.1.

For all nn\in\mathbb{N}, the following holds:

λ^(n)({c^0})=Λn((c^1,c^0]xc^1c^0c^1λ(dx)+λ((c^0,))),\displaystyle\hat{\lambda}^{(n)}(\{\hat{c}_{0}\})=\Lambda_{n}\cdot\left(\int_{(\hat{c}_{1},\hat{c}_{0}]}\frac{x-\hat{c}_{1}}{\hat{c}_{0}-\hat{c}_{1}}\lambda(dx)+\lambda\left((\hat{c}_{0},\infty)\right)\right),
λ^(n)({c^i})=Λn((c^i+1,c^i]xc^i+1c^ic^i+1λ(dx)+(c^i,c^i1)c^i1xc^i1c^iλ(dx))\displaystyle\hat{\lambda}^{(n)}(\{\hat{c}_{i}\})=\Lambda_{n}\cdot\left(\int_{(\hat{c}_{i+1},\hat{c}_{i}]}\frac{x-\hat{c}_{i+1}}{\hat{c}_{i}-\hat{c}_{i+1}}\lambda(dx)+\int_{(\hat{c}_{i},\hat{c}_{i-1})}\frac{\hat{c}_{i-1}-x}{\hat{c}_{i-1}-\hat{c}_{i}}\lambda(dx)\right)

for 1in11\leq i\leq n-1, and

λ^(n)({c^n})=Λn(p~2nc^n+(c^n,c^n1)c^n1xc^n1c^nλ(dx)),\hat{\lambda}^{(n)}(\{\hat{c}_{n}\})=\Lambda_{n}\cdot\left(\frac{\tilde{p}^{n}_{2}}{\hat{c}_{n}}+\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx)\right),

where p~2n=c^0λ(0)({c^0})(c^n,c^0]xλ(dx)\tilde{p}^{n}_{2}=\hat{c}_{0}\lambda^{(0)}(\{\hat{c}_{0}\})-\int_{(\hat{c}_{n},\hat{c}_{0}]}x\lambda(dx).

Proof.

We consider 1in11\leq i\leq n-1 and address the other two cases in a similar manner. Note that

{X^σ^(n)=c^i}={Yσ(n)(c^i+1,c^i1),τc^iθσ(n)<(τc^i+1τc^i1)θσ(n)},\{\hat{X}_{\hat{\sigma}^{(n)}}=\hat{c}_{i}\}=\{Y_{\sigma^{(n)}}\in(\hat{c}_{i+1},\hat{c}_{i-1}),\tau_{\hat{c}_{i}}\circ\theta_{\sigma^{(n)}}<\left(\tau_{\hat{c}_{i+1}}\wedge\tau_{\hat{c}_{i-1}}\right)\circ\theta_{\sigma^{(n)}}\},

where τa:=inf{t>0:Yt=a}\tau_{a}:=\inf\{t>0:Y_{t}=a\} for a(0,)a\in(0,\infty). It follows from the strong Markov property of YY that

λ^(n)({c^i})\displaystyle\hat{\lambda}^{(n)}(\{\hat{c}_{i}\}) =𝐏0(𝐏Yσ(n)(τc^i<τc^i+1τc^i1);Yσ(n)(c^i+1,c^i1))\displaystyle=\mathbf{P}_{0}\left(\mathbf{P}_{Y_{\sigma^{(n)}}}(\tau_{\hat{c}_{i}}<\tau_{\hat{c}_{i+1}}\wedge\tau_{\hat{c}_{i-1}});Y_{\sigma^{(n)}}\in(\hat{c}_{i+1},\hat{c}_{i-1})\right)
=(c^i+1,c^i]xc^i+1c^ic^i+1λ(n)(dx)+(c^i,c^i1)c^i1xc^i1c^iλ(n)(dx).\displaystyle=\int_{(\hat{c}_{i+1},\hat{c}_{i}]}\frac{x-\hat{c}_{i+1}}{\hat{c}_{i}-\hat{c}_{i+1}}\lambda^{(n)}(dx)+\int_{(\hat{c}_{i},\hat{c}_{i-1})}\frac{\hat{c}_{i-1}-x}{\hat{c}_{i-1}-\hat{c}_{i}}\lambda^{(n)}(dx).

Then applying Lemma 7.5, we can obtain the desired expression of λ(n)({c^i})\lambda^{(n)}(\{\hat{c}_{i}\}). ∎

We are now prepared to prove Theorem 5.1 for the case where p3=0p_{3}=0 and |p4|=|p_{4}|=\infty.

Proof of Theorem 5.1 for p3=0,|p4|=p_{3}=0,|p_{4}|=\infty.

Let

𝔭~0:=(c^1,c^0]xc^1c^0c^1λ(dx)+λ((c^0,))\tilde{\mathfrak{p}}_{0}:=\int_{(\hat{c}_{1},\hat{c}_{0}]}\frac{x-\hat{c}_{1}}{\hat{c}_{0}-\hat{c}_{1}}\lambda(dx)+\lambda\left((\hat{c}_{0},\infty)\right)

and

𝔭~n:=(c^n+1,c^n]xc^n+1c^nc^n+1λ(dx)+(c^n,c^n1)c^n1xc^n1c^nλ(dx),n1.\tilde{\mathfrak{p}}_{n}:=\int_{(\hat{c}_{n+1},\hat{c}_{n}]}\frac{x-\hat{c}_{n+1}}{\hat{c}_{n}-\hat{c}_{n+1}}\lambda(dx)+\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx),\quad n\geq 1.

Substituting the expression of λ^(n)\hat{\lambda}^{(n)} from Lemma 8.1 into (2.6), we can formulate the resolvent matrix Ψ^ij(n)(α):=𝐄c^i0eαt1{c^j}(X^t(n))𝑑t\hat{\Psi}^{(n)}_{ij}(\alpha):=\mathbf{E}_{\hat{c}_{i}}\int_{0}^{\infty}e^{-\alpha t}1_{\{\hat{c}_{j}\}}(\hat{X}^{(n)}_{t})dt of X^(n)\hat{X}^{(n)} for α>0\alpha>0 and i,ji,j\in\mathbb{N} as follows:

Ψ^ij(n)(α)Φij(α)\displaystyle\hat{\Psi}^{(n)}_{ij}(\alpha)-\Phi_{ij}(\alpha)
=\displaystyle= uα(i)k=0n1𝔭~kΦkj(α)+(p~2n+c^n(c^n,c^n1)c^n1xc^n1c^nλ(dx))Φnj(α)c^n1Λnk=0n1𝔭~kuα(k)(p~2n+c^n(c^n,c^n1)c^n1xc^n1c^nλ(dx))uα(n)c^n\displaystyle u_{\alpha}(i)\frac{\sum_{k=0}^{n-1}\tilde{\mathfrak{p}}_{k}\Phi_{kj}(\alpha)+\left(\tilde{p}^{n}_{2}+\hat{c}_{n}\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx)\right)\frac{\Phi_{nj}(\alpha)}{\hat{c}_{n}}}{\frac{1}{\Lambda_{n}}-\sum_{k=0}^{n-1}\tilde{\mathfrak{p}}_{k}u_{\alpha}(k)-\left(\tilde{p}^{n}_{2}+\hat{c}_{n}\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx)\right)\frac{u_{\alpha}(n)}{\hat{c}_{n}}}
=\displaystyle= uα(i)k=0n1𝔭~kΦkj(α)+(p~2n+c^n(c^n,c^n1)c^n1xc^n1c^nλ(dx))Φnj(α)c^nλ({})+k=0n1𝔭~k(1uα(k))+(p~2n+c^n(c^n,c^n1)c^n1xc^n1c^nλ(dx))1uα(n)c^n.\displaystyle u_{\alpha}(i)\frac{\sum_{k=0}^{n-1}\tilde{\mathfrak{p}}_{k}\Phi_{kj}(\alpha)+\left(\tilde{p}^{n}_{2}+\hat{c}_{n}\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx)\right)\frac{\Phi_{nj}(\alpha)}{\hat{c}_{n}}}{\lambda(\{\partial\})+\sum_{k=0}^{n-1}\tilde{\mathfrak{p}}_{k}(1-u_{\alpha}(k))+\left(\tilde{p}^{n}_{2}+\hat{c}_{n}\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx)\right)\frac{1-u_{\alpha}(n)}{\hat{c}_{n}}}.

Note that

c^n(c^n,c^n1)c^n1xc^n1c^nλ(dx)(c^n,c^n1)xλ(dx)0\hat{c}_{n}\int_{(\hat{c}_{n},\hat{c}_{n-1})}\frac{\hat{c}_{n-1}-x}{\hat{c}_{n-1}-\hat{c}_{n}}\lambda(dx)\leq\int_{(\hat{c}_{n},\hat{c}_{n-1})}x\lambda(dx)\rightarrow 0

as nn\rightarrow\infty, due to (7.16). In addition, (see, e.g., [26, §7.10, (3) and (9)])

limnΦnj(α)c^n=2uα(j)μj,limn1uα(n)c^n=2αkμkuα(k).\lim_{n\rightarrow\infty}\frac{\Phi_{nj}(\alpha)}{\hat{c}_{n}}=2u_{\alpha}(j)\mu_{j},\quad\lim_{n\rightarrow\infty}\frac{1-u_{\alpha}(n)}{\hat{c}_{n}}=2\alpha\sum_{k\in\mathbb{N}}\mu_{k}u_{\alpha}(k).

Therefore,

limnΨ^ij(n)\displaystyle\lim_{n\rightarrow\infty}\hat{\Psi}^{(n)}_{ij} (α)=Φij(α)\displaystyle(\alpha)=\Phi_{ij}(\alpha)
+uα(i)k=0𝔭~kΦkj(α)+2p~2uα(j)μjλ({})+k=0𝔭~k(1uα(k))+2αp~2kμkuα(k),\displaystyle+u_{\alpha}(i)\frac{\sum_{k=0}^{\infty}\tilde{\mathfrak{p}}_{k}\Phi_{kj}(\alpha)+2\tilde{p}_{2}u_{\alpha}(j)\mu_{j}}{\lambda(\{\partial\})+\sum_{k=0}^{\infty}\tilde{\mathfrak{p}}_{k}(1-u_{\alpha}(k))+2\alpha\tilde{p}_{2}\sum_{k\in\mathbb{N}}\mu_{k}u_{\alpha}(k)},

where p~2\tilde{p}_{2} is defined as (7.19). According to Theorem A.1, this limit is precisely the resolvent matrix of X^\hat{X}. Hence, the parameters of X^\hat{X} are (up to a multiplicative constant):

γ=λ({}),β=2p~2,νk=𝔭~k,k.\gamma=\lambda(\{\partial\}),\quad\beta=2\tilde{p}_{2},\quad\nu_{k}=\tilde{\mathfrak{p}}_{k},\;k\in\mathbb{N}.

Using Theorem 7.6, we can eventually obtain the desired conclusion. ∎

8.2. Sojourn case p3>0p_{3}>0

Let us consider the final case where p3>0p_{3}>0 and |p4|=|p_{4}|=\infty. The basic idea is to transform this case into one with p3=0p_{3}=0. Recall that Y1Y^{1}, defined by (3.23), is a Feller’s Brownian motion with parameters (0,p2,0,p4)(0,p_{2},0,p_{4}) and Y1\ell^{Y^{1}}, defined by (3.24), is its local time at 0. According to §3.2.3, the subprocess Y3Y^{3} of Y1Y^{1}, perturbed by the multiplicative functional

Mt3:=ep1tY1,t0,M^{3}_{t}:=e^{-p_{1}\ell^{Y^{1}}_{t}},\quad t\geq 0,

is a Feller’s Brownian motion with parameters (p1,p2,0,p4)(p_{1},p_{2},0,p_{4}). Denote by ζ3\zeta_{3} the lifetime of Y3Y^{3}. Mimicking the proof of Lemma 3.7, we can easily show that

𝔣3(t):=tζ3+p3tζ3Y1,t0\mathfrak{f}_{3}(t):=t\wedge\zeta_{3}+p_{3}\ell^{Y^{1}}_{t\wedge\zeta_{3}},\quad t\geq 0

is a PCAF of Y3Y^{3} with Supp(𝔣3)=[0,)\text{Supp}(\mathfrak{f}_{3})=[0,\infty).

Lemma 8.2.

The time-changed process of Y3Y^{3} with respect to the PCAF 𝔣3\mathfrak{f}_{3} is a Feller’s Brownian motion with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}).

Proof.

Given Y1Y^{1} expressed as Y1=(Ω1,𝒢1,Yt1,(𝐏x1)x[0,))Y^{1}=(\Omega^{1},{\mathscr{G}}^{1},Y^{1}_{t},(\mathbf{P}^{1}_{x})_{x\in[0,\infty)}), according to [1, III, Theorem 3.3], we can write Y3=(Ω3,𝒢3,Yt3,𝐏x3)Y^{3}=(\Omega^{3},{\mathscr{G}}^{3},Y^{3}_{t},\mathbf{P}^{3}_{x}) as follows:

Ω3=Ω1×[0,],𝒢3:=𝒢1×([0,]),\displaystyle\Omega^{3}=\Omega^{1}\times[0,\infty],\quad{\mathscr{G}}^{3}:={\mathscr{G}}^{1}\times\mathcal{B}([0,\infty]), (8.1)
Yt3(ω,κ):=Yt1(ω) for t<κ, and Yt3(ω,κ):= for tκ,\displaystyle Y^{3}_{t}(\omega,\kappa):=Y^{1}_{t}(\omega)\text{ for }t<\kappa,\text{ and }Y^{3}_{t}(\omega,\kappa):=\partial\text{ for }t\geq\kappa,

and for Γ𝒢3\Gamma\in{\mathscr{G}}^{3} with Γω:={κ[0,]:(ω,κ)Γ}\Gamma^{\omega}:=\{\kappa\in[0,\infty]:(\omega,\kappa)\in\Gamma\} for ωΩ1\omega\in\Omega^{1},

𝐏x3(Γ):=𝐄x1(αω(Γω)),\mathbf{P}^{3}_{x}(\Gamma):=\mathbf{E}^{1}_{x}\left(\alpha_{\omega}(\Gamma^{\omega})\right),

where αω\alpha_{\omega} is a probability measure on [0,][0,\infty] such that αω((t,])=Mt3(ω)\alpha_{\omega}((t,\infty])=M^{3}_{t}(\omega) for all t0t\geq 0. Note that ζ3(ω,κ)=κ\zeta_{3}(\omega,\kappa)=\kappa.

Let 𝔣31(t)\mathfrak{f}^{-1}_{3}(t) denote the right-continuous inverse of 𝔣3(t)\mathfrak{f}_{3}(t). This inverse is continuous and strictly increasing up to ζ3\zeta_{3}. The time-changed process Y4=(Ω4,𝒢4,Yt4,𝐏x4)Y^{4}=(\Omega^{4},{\mathscr{G}}^{4},Y^{4}_{t},\mathbf{P}^{4}_{x}) of Y3Y^{3} with respect to 𝔣3\mathfrak{f}_{3} can be expressed as

Ω4=Ω3,𝒢4=𝒢3,𝐏x4=𝐏x3,\Omega^{4}=\Omega^{3},\quad{\mathscr{G}}^{4}={\mathscr{G}}^{3},\quad\mathbf{P}^{4}_{x}=\mathbf{P}^{3}_{x},

and

Yt4:=Y𝔣31(t)3 for t<ζ4:=𝔣3(ζ3), and Yt4:= for tζ4,Y^{4}_{t}:=Y^{3}_{\mathfrak{f}^{-1}_{3}(t)}\text{ for }t<\zeta_{4}:=\mathfrak{f}_{3}(\zeta_{3}),\text{ and }Y^{4}_{t}:=\partial\text{ for }t\geq\zeta_{4},

where ζ4\zeta_{4} is the lifetime of Y4Y^{4}. By substituting the expression (8.1) of Y3Y^{3} into this expression of Y4Y^{4}, we find that for (ω,κ)Ω1×[0,](\omega,\kappa)\in\Omega^{1}\times[0,\infty],

Yt4(ω,κ)=Y𝔣1(t)1(ω),t<ζ4(ω,κ)=κ+p3κY1(ω)Y^{4}_{t}(\omega,\kappa)=Y^{1}_{\mathfrak{f}^{-1}(t)}(\omega),\quad t<\zeta_{4}(\omega,\kappa)=\kappa+p_{3}\ell^{Y^{1}}_{\kappa}(\omega)

where 𝔣1\mathfrak{f}^{-1} is the right-continuous inverse of 𝔣(t)=t+p3tY1(ω)\mathfrak{f}(t)=t+p_{3}\ell^{Y^{1}}_{t}(\omega) (t0t\geq 0). In other words, Y4Y^{4} is the killed process of Y2Y^{2}, defined in (3.25), at time ζ4\zeta_{4}. Note that for all t0t\geq 0,

αω({κ[0,]:ζ4(ω,κ)t}))=αω({κ[0,]:κ𝔣1(t)})=ep1𝔣1(t)Y1(ω).\alpha_{\omega}\left(\{\kappa\in[0,\infty]:\zeta_{4}(\omega,\kappa)\geq t\}\right))=\alpha_{\omega}\left(\{\kappa\in[0,\infty]:\kappa\geq\mathfrak{f}^{-1}(t)\}\right)=e^{-p_{1}\ell^{Y^{1}}_{\mathfrak{f}^{-1}(t)}(\omega)}.

Thus, Y4Y^{4} is the subprocess of Y2Y^{2} perturbed by the multiplicative functional in (3.26). According to the pathwise representation in §3.2.3, we conclude that Y4Y^{4} is a Feller’s Brownian motion with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}). ∎

Now, we complete the proof of Theorem 5.1 as follows.

Proof of Theorem 5.1 for p3>0,|p4|=p_{3}>0,|p_{4}|=\infty.

Let Y3=(Ω3,Yt3,𝐏x3)Y^{3}=(\Omega^{3},Y^{3}_{t},\mathbf{P}^{3}_{x}), with lifetime ζ3\zeta_{3}, be a Feller’s Brownian motion with parameters (p1,p2,0,p4)(p_{1},p_{2},0,p_{4}). According to Lemma 8.2, the Feller’s Brownian motion Y=(Ω,Yt,𝐏x)Y=(\Omega,Y_{t},\mathbf{P}_{x}) with parameters (p1,p2,p3,p4)(p_{1},p_{2},p_{3},p_{4}) can be represented as the time-changed process of Y3Y^{3} with respect to 𝔣3\mathfrak{f}_{3}. The lifetime of YY is ζ=𝔣3(ζ3)\zeta=\mathfrak{f}_{3}(\zeta_{3}).

Denote by Yˇ3\check{Y}^{3} and Yˇ\check{Y} the time-changed process obtained in Theorem 4.2 for Y3Y^{3} and YY, respectively. The parameters determining the resolvent matrix of Yˇ3\check{Y}^{3} have been examined in §8.1. It suffices to show that Yˇ\check{Y} is identical in law to Yˇ3\check{Y}^{3}. Let Lt3,c^nL^{3,\hat{c}_{n}}_{t} denote the local time of Y3Y^{3} at c^n\hat{c}_{n} as defined in Definition 3.1, and set At3:=nμnLt3,c^nA^{3}_{t}:=\sum_{n\in\mathbb{N}}\mu_{n}L^{3,\hat{c}_{n}}_{t} for all t0t\geq 0. It is straightforward to verify that

Ltc^n:=L𝔣31(t)3,c^n,t0L^{\hat{c}_{n}}_{t}:=L^{3,\hat{c}_{n}}_{\mathfrak{f}^{-1}_{3}(t)},\quad t\geq 0

is the local time of YY at c^n\hat{c}_{n} in the sense of Definition 3.1. Thus, the PCAF of YY given by (4.3) is At=A𝔣31(t)3A_{t}=A^{3}_{\mathfrak{f}^{-1}_{3}(t)}, and the right-continuous inverse of AA is

γt=inf{s>0:As>t}=inf{s>0:A𝔣31(s)3>t},t<Aζ=Aζ33.\gamma_{t}=\inf\{s>0:A_{s}>t\}=\inf\{s>0:A^{3}_{\mathfrak{f}^{-1}_{3}(s)}>t\},\quad\forall t<A_{\zeta}=A^{3}_{\zeta_{3}}.

Since 𝔣3\mathfrak{f}_{3} is strictly increasing up to ζ3\zeta_{3}, it follows that

𝔣31(γt)=inf{s>0:As3>t}=:γt3,t<Aζ=Aζ33,\mathfrak{f}^{-1}_{3}(\gamma_{t})=\inf\{s>0:A^{3}_{s}>t\}=:\gamma^{3}_{t},\quad\forall t<A_{\zeta}=A^{3}_{\zeta_{3}},

where γt3\gamma^{3}_{t} is the right-continuous inverse of A3A^{3}. Particularly, for t<Aζ=Aζ33t<A_{\zeta}=A^{3}_{\zeta_{3}},

Yˇt=Yγt=Y𝔣31(γt)3=Yγt33=Yˇt3.\check{Y}_{t}=Y_{\gamma_{t}}=Y^{3}_{\mathfrak{f}^{-1}_{3}(\gamma_{t})}=Y^{3}_{\gamma^{3}_{t}}=\check{Y}^{3}_{t}. (8.2)

Note that AζA_{\zeta} is the lifetime of Yˇ\check{Y} and Aζ33A^{3}_{\zeta_{3}} is the lifetime of Yˇ3\check{Y}^{3}. Therefore, (8.2) establishes the identification between Yˇ\check{Y} and Yˇ3\check{Y}^{3}. ∎

Appendix A Approximation of birth-death process

For a Feller QQ-process XX, we can also construct a sequence of Doob processes X(n)X^{(n)} in the same manner as described in Lemma 7.1. This approach is precisely the probabilistic construction method for all QQ-processes presented in [26]. For the convenience of readers, we restate some necessary details as follows.

Let us consider the corresponding QQ-process X^t=Ξ1(Xt)\hat{X}_{t}=\Xi^{-1}(X_{t}) on E¯\overline{E}, with lifetime denoted by ζ^\hat{\zeta}. Define

η^:=inf{t>0:X^t=0}\hat{\eta}:=\inf\{t>0:\hat{X}_{t-}=0\}

and

σ^(n):=inf{t>0:X^t{c^0,,c^n}}ζ^,n.\hat{\sigma}^{(n)}:=\inf\left\{t>0:\hat{X}_{t}\in\{\hat{c}_{0},\cdots,\hat{c}_{n}\}\right\}\wedge\hat{\zeta},\quad n\in\mathbb{N}.

Analogous sequences of stopping times {η^m(n):m1}\{\hat{\eta}^{(n)}_{m}:m\geq 1\} and {σ^m(n):m1}\{\hat{\sigma}^{(n)}_{m}:m\geq 1\} can be defined by repeating the procedures used before Lemma 7.1 (see also [26, §6.3]). Applying the C(η^m(n),σ^m(n))C(\hat{\eta}^{(n)}_{m},\hat{\sigma}^{(n)}_{m})-transformation to X^\hat{X} yields a sequence of Doob processes X^(n)\hat{X}^{(n)} on EE, with instantaneous distribution λ^(n)()=𝐏0(X^σ^(n))\hat{\lambda}^{(n)}(\cdot)=\mathbf{P}_{0}(\hat{X}_{\hat{\sigma}^{(n)}}\in\cdot).

The main result of [26, §6] establishes the convergence of X^(n)\hat{X}^{(n)} to X^\hat{X} in the following sense.

Theorem A.1.

Let XX be a Feller QQ-process and let X(n)=Ξ(X^(n))X^{(n)}=\Xi(\hat{X}^{(n)}) denote the approximating sequence of Doob processes. Then, for any ii\in\mathbb{N},

𝐏i(limnXt(n)=Xt)=1,t0.\mathbf{P}_{i}\left(\lim_{n\rightarrow\infty}X^{(n)}_{t}=X_{t}\right)=1,\quad\forall t\geq 0. (A.1)
Proof.

The original proof in [26] is quite lengthy and only addresses the honest case. Here, we provide an alternative proof using the right continuity of XX. For convenience, we will consider X^\hat{X} and X^(n)\hat{X}^{(n)} instead of XX and X(n)X^{(n)}. Denote by ζ^(n)\hat{\zeta}^{(n)} the lifetime of X^(n)\hat{X}^{(n)}. According to the definition of the C(η^m(n),σ^m(n))C(\hat{\eta}^{(n)}_{m},\hat{\sigma}^{(n)}_{m})-transformation, it is straightforward to observe that

ζ^(1)ζ^(2)ζ^(n)ζ^.\hat{\zeta}^{(1)}\leq\hat{\zeta}^{(2)}\leq\cdots\leq\hat{\zeta}^{(n)}\leq\cdots\leq\hat{\zeta}.

Let ζ^:=limnζ^(n)\hat{\zeta}_{\infty}:=\lim_{n\rightarrow\infty}\hat{\zeta}^{(n)} (ζ^\leq\hat{\zeta}).

Firstly, we show that there exists an integer NN such that for any nNn\geq N,

𝐏c^i(σ^m(n)<)=1,m1,i.\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(n)}_{m}<\infty\right)=1,\quad\forall m\geq 1,i\in\mathbb{N}. (A.2)

It suffices to prove (A.2) for some integer n=Nn=N, due to σ^m(n+1)σ^m(n)\hat{\sigma}^{(n+1)}_{m}\leq\hat{\sigma}^{(n)}_{m} for all nn\in\mathbb{N}. To do this, note that we can find some t0t\geq 0 and N1N\geq 1 such that

𝐏0(X^tc^N or X^t=)>0,\mathbf{P}_{0}(\hat{X}_{t}\geq\hat{c}_{N}\text{ or }\hat{X}_{t}=\partial)>0, (A.3)

because otherwise

1=𝐏0(t+,n1{X^t<c^n})=𝐏0(X^t=0,t0),1=\mathbf{P}_{0}\left(\bigcap_{t\in\mathbb{Q}_{+},n\geq 1}\left\{\hat{X}_{t}<\hat{c}_{n}\right\}\right)=\mathbf{P}_{0}\left(\hat{X}_{t}=0,\forall t\geq 0\right),

where +\mathbb{Q}_{+} is the set of all positive rational numbers. To prove (A.2) for n=Nn=N with NN in (A.3), consider the case m=1m=1. We note that σ^1(N)=σ^(N)θη^\hat{\sigma}^{(N)}_{1}=\hat{\sigma}^{(N)}\circ\theta_{\hat{\eta}}, and η^<\hat{\eta}<\infty, 𝐏c^i\mathbf{P}_{\hat{c}_{i}}-a.s. It follows that

𝐏c^i(σ^1(N)<)=𝐏0(σ^(N)<)𝐏0(X^tc^N or X^t=)>0.\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{1}<\infty\right)=\mathbf{P}_{0}\left(\hat{\sigma}^{(N)}<\infty\right)\geq\mathbf{P}_{0}(\hat{X}_{t}\geq\hat{c}_{N}\text{ or }\hat{X}_{t}=\partial)>0.

Thus, there exists a constant T>0T>0 such that

𝐏c^i(σ^1(N)T)=𝐏0(σ^(N)T)=:α<1.\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{1}\geq T\right)=\mathbf{P}_{0}\left(\hat{\sigma}^{(N)}\geq T\right)=:\alpha<1.

For k2k\geq 2, it follows from the strong Markov property that

𝐏c^i(σ^1(N)kT)\displaystyle\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{1}\geq kT\right) =𝐏c^i(σ^1(N)θ(k1)TT;σ^1(N)(k1)T)\displaystyle=\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{1}\circ\theta_{(k-1)T}\geq T;\hat{\sigma}^{(N)}_{1}\geq(k-1)T\right)
=𝐏c^i(𝐏X^(k1)T(σ^1(N)T);σ^1(N)(k1)T)\displaystyle=\mathbf{P}_{\hat{c}_{i}}\left(\mathbf{P}_{\hat{X}_{(k-1)T}}\left(\hat{\sigma}^{(N)}_{1}\geq T\right);\hat{\sigma}^{(N)}_{1}\geq(k-1)T\right)
α𝐏c^i(σ^1(N)(k1)T).\displaystyle\leq\alpha\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{1}\geq(k-1)T\right).

By induction, we get 𝐏c^i(σ^1(N)kT)αk\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{1}\geq kT\right)\leq\alpha^{k} for all k0k\geq 0. Hence,

𝐄c^iσ^1(N)Tk0𝐏c^i(σ^1(N)kT)Tk0αk<.\mathbf{E}_{\hat{c}_{i}}\hat{\sigma}^{(N)}_{1}\leq T\sum_{k\geq 0}\mathbf{P}_{\hat{c}_{i}}(\hat{\sigma}^{(N)}_{1}\geq kT)\leq T\sum_{k\geq 0}\alpha^{k}<\infty.

This indicates (A.2) for the case m=1m=1. For general m2m\geq 2, (A.2) can also be established by using the strong Markov property. We provide details for the case m=2m=2, with the other cases handled by induction. In fact, we have

𝐏c^i(σ^2(N)<)\displaystyle\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{2}<\infty\right) =𝐏c^i(σ^2(N)<;X^σ^1(N)E)+𝐏c^i(σ^2(N)<;X^σ^1(N)=)\displaystyle=\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{2}<\infty;\hat{X}_{\hat{\sigma}_{1}^{(N)}}\in E\right)+\mathbf{P}_{\hat{c}_{i}}\left(\hat{\sigma}^{(N)}_{2}<\infty;\hat{X}_{\hat{\sigma}_{1}^{(N)}}=\partial\right)
=𝐏c^i(𝐏X^σ^1(N)(σ^1(N)<);X^σ^1(N)E)+𝐏c^i(X^σ^1(N)=)\displaystyle=\mathbf{P}_{\hat{c}_{i}}\left(\mathbf{P}_{\hat{X}_{\hat{\sigma}_{1}^{(N)}}}\left(\hat{\sigma}^{(N)}_{1}<\infty\right);\hat{X}_{\hat{\sigma}_{1}^{(N)}}\in E\right)+\mathbf{P}_{\hat{c}_{i}}\left(\hat{X}_{\hat{\sigma}_{1}^{(N)}}=\partial\right)
=𝐏c^i(X^σ^1(N)E)+𝐏c^i(X^σ^1(N)=)=1.\displaystyle=\mathbf{P}_{\hat{c}_{i}}\left(\hat{X}_{\hat{\sigma}_{1}^{(N)}}\in E\right)+\mathbf{P}_{\hat{c}_{i}}\left(\hat{X}_{\hat{\sigma}_{1}^{(N)}}=\partial\right)=1.

Next, we prove that for all ii\in\mathbb{N},

ζ^=ζ^,𝐏c^i-a.s.\hat{\zeta}_{\infty}=\hat{\zeta},\quad\mathbf{P}_{\hat{c}_{i}}\text{-a.s.}

If ζ^(ω)<\hat{\zeta}(\omega)<\infty, then according to the definition of X^(n)\hat{X}^{(n)}, we have

ζ^(n)(ω)ζ^(ω)|{0t<ζ^(ω):X^t(ω)<c^n}|,\hat{\zeta}^{(n)}(\omega)\geq\hat{\zeta}(\omega)-\left|\{0\leq t<\hat{\zeta}(\omega):\hat{X}_{t}(\omega)<\hat{c}_{n}\}\right|,

where |||\cdot| stands for the Lebesgue measure of the time set. Note that

|{0t<ζ^(ω):X^t(ω)=0}|=0,𝐏c^i-a.s.\left|\{0\leq t<\hat{\zeta}(\omega):\hat{X}_{t}(\omega)=0\}\right|=0,\quad\mathbf{P}_{\hat{c}_{i}}\text{-a.s.}

It follows that

ζ^(ω)\displaystyle\hat{\zeta}_{\infty}(\omega) =limnζ^(n)(ω)ζ^(ω)limn|{0t<ζ^(ω):X^t(ω)<c^n}|\displaystyle=\lim_{n\rightarrow\infty}\hat{\zeta}^{(n)}(\omega)\geq\hat{\zeta}(\omega)-\lim_{n\rightarrow\infty}\left|\{0\leq t<\hat{\zeta}(\omega):\hat{X}_{t}(\omega)<\hat{c}_{n}\}\right|
=ζ^(ω)limn|{0t<ζ^(ω):X^t(ω)=0}|=ζ^(ω).\displaystyle=\hat{\zeta}(\omega)-\lim_{n\rightarrow\infty}\left|\{0\leq t<\hat{\zeta}(\omega):\hat{X}_{t}(\omega)=0\}\right|=\hat{\zeta}(\omega).

Thus, ζ^(ω)=ζ^(ω)\hat{\zeta}_{\infty}(\omega)=\hat{\zeta}(\omega) whenever ζ^(ω)<\hat{\zeta}(\omega)<\infty. For the case ζ^(ω)=\hat{\zeta}(\omega)=\infty, we prove that ζ^(n)(ω)=\hat{\zeta}^{(n)}(\omega)=\infty for all nNn\geq N with NN satisfying (A.2). It suffices to consider the case where λ^(n)({})>0\hat{\lambda}^{(n)}(\{\partial\})>0, because otherwise X^(n)\hat{X}^{(n)} is honest by Theorem 2.1. If ζ^(n)<\hat{\zeta}^{(n)}<\infty, then the total duration discarded from X^(ω)\hat{X}_{\cdot}(\omega) in constructing X^(n)(ω)\hat{X}^{(n)}_{\cdot}(\omega) must be infinite. It follows from (A.2) that

{ζ^(n)<,ζ^=}{X^σ^m(n)E,m1}.\{\hat{\zeta}^{(n)}<\infty,\hat{\zeta}=\infty\}\subset\{\hat{X}_{\hat{\sigma}^{(n)}_{m}}\in E,\forall m\geq 1\}.

By the strong Markov property, for any integer M1M\geq 1,

𝐏c^i(X^σ^m(n)E,1mM)=(1λ^(n)({}))M.\mathbf{P}_{\hat{c}_{i}}\left(\hat{X}_{\hat{\sigma}^{(n)}_{m}}\in E,1\leq m\leq M\right)=(1-\hat{\lambda}^{(n)}(\{\partial\}))^{M}.

Therefore,

𝐏c^i(ζ^(n)<,ζ^=)limM(1λ^(n)({}))M=0.\mathbf{P}_{\hat{c}_{i}}\left(\hat{\zeta}^{(n)}<\infty,\hat{\zeta}=\infty\right)\leq\lim_{M\rightarrow\infty}(1-\hat{\lambda}^{(n)}(\{\partial\}))^{M}=0.

This implies ζ^(ω)=ζ^(ω)\hat{\zeta}_{\infty}(\omega)=\hat{\zeta}(\omega) for ζ^(ω)=\hat{\zeta}(\omega)=\infty.

We are now prepared to prove (A.1). Specifically, we need to establish that

limnX^t(n)=X^t,𝐏c^i-a.s.\lim_{n\rightarrow\infty}\hat{X}^{(n)}_{t}=\hat{X}_{t},\quad\mathbf{P}_{\hat{c}_{i}}\text{-a.s.}

for a fixed t0t\geq 0. For ωΩ\omega\in\Omega, let ρt(n)(ω)\rho^{(n)}_{t}(\omega) represent the total duration discarded from the path of X^(ω)\hat{X}_{\cdot}(\omega) in constructing X^s(n)(ω)\hat{X}^{(n)}_{s}(\omega) for 0st0\leq s\leq t, namely, X^t(n)(ω)=X^t+ρt(n)(ω)(ω)\hat{X}^{(n)}_{t}(\omega)=\hat{X}_{t+\rho^{(n)}_{t}(\omega)}(\omega). Obviously, ρt(n)(ω)\rho^{(n)}_{t}(\omega) is decreasing as nn\rightarrow\infty. If tζ^(ω)=ζ^(ω)t\geq\hat{\zeta}(\omega)=\hat{\zeta}_{\infty}(\omega), then

X^t(ω)=X^t(n)(ω)=,n.\hat{X}_{t}(\omega)=\hat{X}^{(n)}_{t}(\omega)=\partial,\quad\forall n\in\mathbb{N}.

If t<ζ^(ω)=ζ^(ω)t<\hat{\zeta}(\omega)=\hat{\zeta}_{\infty}(\omega), then there exists an integer n0n_{0} (which may depend on ω\omega) such that

t<ζ^(n0)(ω)ζ^(n)(ω),nn0.t<\hat{\zeta}^{(n_{0})}(\omega)\leq\hat{\zeta}^{(n)}(\omega),\quad\forall n\geq n_{0}.

This implies

ρt(n)(ω)ρt(n0)(ω)<.\rho^{(n)}_{t}(\omega)\leq\rho^{(n_{0})}_{t}(\omega)<\infty.

We have

ρt(n)(ω)\displaystyle\rho^{(n)}_{t}(\omega) |{0st+ρt(n)(ω):X^s(ω)<c^n}|\displaystyle\leq\left|\{0\leq s\leq t+\rho^{(n)}_{t}(\omega):\hat{X}_{s}(\omega)<\hat{c}_{n}\}\right|
|{0st+ρt(n0)(ω):X^s(ω)<c^n}|.\displaystyle\leq\left|\{0\leq s\leq t+\rho^{(n_{0})}_{t}(\omega):\hat{X}_{s}(\omega)<\hat{c}_{n}\}\right|.

Thus

limnρt(n)(ω)\displaystyle\lim_{n\rightarrow\infty}\rho^{(n)}_{t}(\omega) limn|{0st+ρt(n0)(ω):X^s(ω)<c^n}|\displaystyle\leq\lim_{n\rightarrow\infty}\left|\{0\leq s\leq t+\rho^{(n_{0})}_{t}(\omega):\hat{X}_{s}(\omega)<\hat{c}_{n}\}\right|
=|{0st+ρt(n0)(ω):X^s(ω)=0}|=0.\displaystyle=\left|\{0\leq s\leq t+\rho^{(n_{0})}_{t}(\omega):\hat{X}_{s}(\omega)=0\}\right|=0.

Therefore, by the right continuity of X^\hat{X}, we obtain

limnX^t(n)(ω)=limnX^t+ρt(n)(ω)(ω)=X^t(ω),\lim_{n\rightarrow\infty}\hat{X}^{(n)}_{t}(\omega)=\lim_{n\rightarrow\infty}\hat{X}_{t+\rho^{(n)}_{t}(\omega)}(\omega)=\hat{X}_{t}(\omega),

whenever t<ζ^(ω)t<\hat{\zeta}(\omega). ∎

Acknowledgement

The author wishes to thank Professor Patrick J. Fitzsimmons from the University of California, San Diego, for his invaluable suggestion, which inspired the exploration of this topic.

References

  • [1] R M Blumenthal and Ronald Getoor. Markov processes and potential theory. Pure and Applied Mathematics, Vol. 29. Academic Press, New York-London, 1968.
  • [2] Mu-Fa Chen. From Markov chains to non-equilibrium particle systems. World Scientific, March 2004.
  • [3] Zhen-Qing Chen and Masatoshi Fukushima. Symmetric Markov processes, time change, and boundary theory, volume 35 of London Mathematical Society Monographs Series. Princeton University Press, Princeton, NJ, 2012.
  • [4] Kai Lai Chung. Markov chains with stationary transition probabilities. Second edition. Die Grundlehren der mathematischen Wissenschaften, Band 104. Springer-Verlag New York, Inc., New York, 1967.
  • [5] William Feller. The parabolic differential equations and the associated semi-groups of transformations. Ann. of Math., 55(3):468–519, May 1952.
  • [6] William Feller. The birth and death processes as diffusion processes. J. Math. Pures Appl. (9), 38:301–345, 1959.
  • [7] Masatoshi Fukushima. On general boundary conditions for one-dimensional diffusions with symmetry. J. Math. Soc. Japan, 66(1):289–316, 2014.
  • [8] Masatoshi Fukushima, Yoichi Oshima, and Masayoshi Takeda. Dirichlet forms and symmetric Markov processes, volume 19 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, extended edition, 2011.
  • [9] Nobuyuki Ikeda, Masao Nagasawa, and Shinzo Watanabe. A construction of Markov processes by piecing out. Proc. Japan Acad., 42:370–375, 1966.
  • [10] K. Itô and H. P. McKean. Brownian motions on a half line. Illinois Journal of Mathematics, 7(2):181–231, 1963.
  • [11] Kiyosi Itô and Henry P Kean Jr. Diffusion processes and their sample paths. Springer-Verlag, Berlin-New York, 1974.
  • [12] I. S. Kac and M. G. Krein. On the spectral functions of the string, volume 103, page 19–102. American Mathematical Society, Providence, Rhode Island, 1974.
  • [13] Yuji Kasahara. Spectral theory of generalized second order differential operators and its applications to Markov processes. Japan. J. Math. (N.S.), 1(1):67–84, 1975.
  • [14] Frank B Knight. Characterization of the Lévy measures of inverse local times of gap diffusion. In Seminar on Stochastic Processes, 1981 (Evanston, Ill., 1981), pages 53–78. Birkhäuser, Boston, Mass., 1981.
  • [15] S Kotani and S Watanabe. Krein’s spectral theory of strings and generalized diffusion processes. In Functional analysis in Markov processes (Katata/Kyoto, 1981), pages 235–259. Springer, Berlin-New York, 1982.
  • [16] Liping Li. Dirichlet form approach to one-dimensional Markov processes with discontinuous scales. Journal of the Mathematical Society of Japan, to appear.
  • [17] Liping Li. On generalization of quasidiffusions. arXiv: 2203.15444, March.
  • [18] Liping Li. Ray–Knight compactification of birth and death processes. Stochastic Processes and their Applications, 177:104456, 2024.
  • [19] Liping Li and Ying Li. Resolvent approach to diffusions with discontinuous scale. (arXiv:2309.06211), September 2023. arXiv:2309.06211 [math].
  • [20] Liping Li and Jiangang Ying. On structure of regular Dirichlet subspaces for one-dimensional Brownian motion. Ann. Probab., 45(4):2631–2654, July 2017.
  • [21] Petr Mandl. Analytical treatment of one-dimensional Markov processes. Die Grundlehren der mathematischen Wissenschaften, Band 151. Academia Publishing House of the Czechoslovak Academy of Sciences, Prague; Springer-Verlag New York Inc., New York, 1968.
  • [22] Daniel Revuz and Marc Yor. Continuous martingales and Brownian motion, volume 293 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, Berlin, Heidelberg, third edition, 1999.
  • [23] L C G Rogers and David Williams. Diffusions, Markov processes, and martingales. Vol. 2. Wiley Series in Probability and Mathematical Statistics: Probability and Mathematical Statistics. John Wiley & Sons, Inc., New York, 1987.
  • [24] Dieter Schütze. One-dimensional diffusions with discontinuous scale. Z. Wahrsch. Verw. Gebiete, 49(1):97–104, 1979.
  • [25] Michael Sharpe. General theory of Markov processes, volume 133 of Pure and Applied Mathematics. Academic Press, Inc., Boston, MA, 1988.
  • [26] Zi Kun Wang and Xiang Qun Yang. Birth and death processes and Markov chains. Springer-Verlag, Berlin; Science Press Beijing, Beijing, 1992.