Time-changed Feller’s Brownian motions are birth-death processes
Abstract.
A Feller’s Brownian motion refers to a Feller process on the interval that is equivalent to the absorbing Brownian motion before reaching . It is fully determined by four parameters , reflecting its killing, reflecting, sojourn, and jumping behaviors at the boundary . On the other hand, a birth-death process is a continuous-time Markov chain on with a given transition density matrix , and it is characterized by three parameters that describe its killing, reflecting, and jumping behaviors at the boundary . 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 , through time change, has the parameters:
where is a sequence derived by allocating weights to the measure 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, 60J74.1. 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 , allowing all possible behaviors of the process at the boundary 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 :
(1.1) |
where are constants, and is a positive measure on satisfying . From a probabilistic perspective, these four parameters represent the four possible types of boundary behavior of Feller’s Brownian motion at : 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 is the density matrix given by (2.1), on the discrete space . 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 in finite time, describing its behavior at and after reaching 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 bears some intrinsic similarity to the boundary behavior of the aforementioned Feller’s Brownian motion at . Feller’s approach was analytical; although he only examined partial birth-death processes, he derived the boundary condition at for the resolvent function (i.e., the function in the domain of the generator) as follows:
(1.2) |
where denotes the discrete gradient (see §2.3), are constants, and is a positive measure on . 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 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 , 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 and 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 , which causes that the birth-death process can not visit for a positive duration (correspondingly, the process that allows sojourn at is called a generalized birth-death process, and its index set is ; 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 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 -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 or 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 so that it moves to , thereby matching the boundary point of Feller’s Brownian motion. The transformed space is denoted by (see (4.1))
where the subscript of represents the point before the transformation, and is topologically consistent with the one-point compactification of . Under this spatial transformation, aside from the change in the symbol of the state points, the transformed process, which we denote by , remains indistinguishable from the original birth-death process. For convenience, we refer to as a birth-death process on . Additionally, due to the change in the direction of the gradient at the boundary point, the corresponding boundary condition for is modified to:
(1.3) |
where is the transformed function of 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 , for the Feller’s Brownian motion corresponding to the boundary condition (1.1), it can always (and uniquely, see Theorem 4.4) be transformed into a birth-death process on through a time change transformation induced by a positive continuous additive functional . This positive continuous additive functional given by (4.6) is the integral of the local times of with respect to the speed measure determined by . Moreover, the parameters exhibited in the boundary condition (1.3) of the birth-death process obtained by time change can be derived from the parameters of as follows:
(1.4) |
where is the sequence obtained by allocating weights to the measure in the manner of (5.1) and (5.2). Particularly, whenever is supported on . It should be noted that the sojourn parameter of does not play any role in the expression (1.4). This is because any positive duration of stay at by 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 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 is the trace of the corresponding Feller’s Brownian motion on ; 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 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 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 are standard, as seen in [1, 25]. For brevity, some elements are sometimes omitted, such as writing . The expectation corresponding to the probability measure is denoted by . In this paper, the cemetery point for Markov processes is uniformly denoted by . Given a topological space , and 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 is an interval, and denote the set of all continuous functions with compact support and the set of all continuous functions that equal at the open endpoints, respectively. The notation for continuous functions with subscripts indicates differentiability, for example, indicates twice continuously differentiable, while indicates infinitely continuously differentiable. The symbol stands for the Dirac measure at .
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 -processes
We consider a birth-death density matrix as follows:
(2.1) |
where for and for . (Set for convenience.) A continuous-time Markov chain is called a birth-death -process (or -process for short) if its transition matrix is standard and its density matrix is , i.e., for . Readers are referred to [4, 26] for the terminologies concerning continuous-time Markov chains; see also [18]. In our consideration, two -processes with the same transition matrix will not be distinguished. For convenience, we will also refer to such as a -process when there is no risk of confusion.
There always exists at least one -process , known as the minimal -process. This process is killed at the first time it almost reaches . Similar to a regular diffusion on an interval, the minimal -process can be characterized by the scale function (on ):
(2.2) |
and the speed measure on :
(2.3) |
The transition matrix of is symmetric with respect to in the sense that for all and . For more details about this characterization, please refer to [18, §3.1].
2.2. Resolvent representation of -processes
When is regular or an exit (for ) in Feller’s sense (see, e.g., [18, Definition 3.3]), there exist other -processes besides the minimal one, such as Doob processes (see, e.g., [18, §5]) and the -process (which is only applicable in the regular case; see, e.g., [18, §3.3]). In this case, we have .
In what follows, let us present a well-known analytic approach to characterize all birth-death processes by solving their resolvents. For and , define
where stands for the expectation of starting from and is the lifetime of . Denote by the resolvent of the minimal -process. Set
Let be a positive measure on and let be two constants. Set . When both
(2.4) |
and
(2.5) |
are satisfied, define, for and ,
(2.6) |
The matrix is called the -resolvent matrix. For a constant , the -resolvent matrix is obviously the same as the -resolvent matrix.
The following theorem is attributed to [26, §7.6]. Note that a -process is called honest if its transition semigroup satisfies for all and .
Theorem 2.1.
The transition matrix is a -process that is not the minimal one, if and only if there exists a unique (up to a multiplicative positive constant) triple with (2.4) and (2.5) such that the resolvent of is given by
where is the -resolvent matrix. Furthermore,
-
(1)
is honest, if and only if .
-
(2)
is a Doob process, if and only if and .
2.3. Ray-Knight compactification of -processes
In a previous study [18], we obtained a càdlàg modification for each -process , denoted by , using the Ray-Knight compactification. The state space of is , which corresponds to the Alexandroff compactification of . It is important to note that the Ray-Knight compactifications of Doob processes are not normal, whereas those of other -processes are Feller processes on ; see [18, Corollary 5.2]. If no ambiguity arises, we will not distinguish between the -process and its Ray-Knight compactification. Additionally, we will refer to a non-minimal and non-Doob -process as a Feller -process.
Regarding a Feller -process , 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 in the domain of the infinitesimal generator satisfies the following boundary condition at :
(2.7) |
where (if it exists), for and (if it exists). (When deriving the boundary condition (2.7) in [18], we mistakenly wrote the parameter as . 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 in [18, (6.10)] was incorrectly written as . 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 is called a Feller process if its semigroup satisfies the following conditions: is the identify mapping, and in as for any , where is the Banach space consisting of all continuous functions on such that . We provide a definition for Feller’s Brownian motions in a modern manner as follows.
Definition 3.1.
A Feller process with lifetime on is called a Feller’s Brownian motion if its killed process upon hitting is identical in law to the absorbing Brownian motion on .
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 , i.e., , or the quasi-left-continuity at the first time reaching , i.e., for , , where for any sequence .
The case 4a. in [10, §6] warrants special mention. Similar to Doob processes in the context of -processes, it corresponds to the piecing out of the absorbing Brownian motion with lifetime on with respect to a certain probability measure on . The process obtained through this piecing out is a right process on , but it is not a Feller process. It can, however, be extended to a Ray process on , where the initial transition function at is given by . Analogous to [18, Theorem 5.1], the following expression for the resolvent of can be easily obtained:
(3.1) |
where is the resolvent of and .
Indeed, the reflecting Brownian motion on 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 from or just before reaching .
3.1. Infinitesimal generators
It is well known that a function is said to belong to the domain of the infinitesimal generator of if the limit
exists in . The operator thus defined is called the infinitesimal generator of the process . The following characterization of infinitesimal generators of Feller’s Brownian motions is attributed to Feller [5].
Theorem 3.3.
A Markov process on is a Feller’s Brownian motion, if and only if it is a Feller process on whose infinitesimal generator on is with domain
(3.2) | ||||
where are non-negative numbers and is a positive measure on such that
(3.3) |
and
(3.4) |
The parameters are uniquely determined for each Feller’s Brownian motion.
Proof.
For the readers’ convenience, we provide necessary details regarding this proof. The uniqueness of is evident.
The necessity can be established by demonstrating that
(3.5) | ||||
and then applying [21, II§5, Theorem 2]. To achieve this, let denote the resolvent of . Fix . By the Hille-Yosida theorem, , and for with ,
(3.6) |
Set . It follows from Definition 3.1 and Dynkin’s formula that
(3.7) |
where is the resolvent of the absorbing Brownian motion on . Note that (see, e.g., [11, page 26, 5)]), and (see, e.g., [21, II§3, #7 and #8])
(3.8) |
For any , it holds that
(3.9) | ||||
and thus
(3.10) |
Using (3.6), (3.7), (3.9), and (3.10), we conclude that , , and . In other words, (3.5) is established.
Regarding the sufficiency, we first apply [21, II§5, Theorem 3] to , showing that with domain (3.2) is indeed the infinitesimal generator of a Feller process on . It remains to demonstrate that the killed process
(3.11) |
is identical in law to the absorbing Brownian motion on . In fact, the proof of [21, II§5, Theorem 3] shows that the resolvent kernel of is given by [21, §3, #7] (with and being the Lebesgue measure). This is precisely the resolvent kernel of absorbing Brownian motion. ∎
Remark 3.4.
When and , the condition (3.4) is not satisfied. However, this case corresponds to the absorbing Brownian motion on .
The triple plays a role analogous to for a -process, with some heuristic explanations for the latter triple presented in [18, §2]. The additional parameter measures the sojourn of the Feller’s Brownian motion at . 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 of is symmetric with respect to some -finite measure in the sense that
as stated in the following corollary. In this special case, the parameter represents the mass of the symmetric measure at . It is well known that when this symmetric measure, also known as the speed measure of , is larger in a specific region, the motion of will be slower when passing through that region.
Corollary 3.5.
Let be a Feller’s Brownian motion on with parameters as described in Theorem 3.3. The process is symmetric if and only if . Furthermore, the symmetric measure must be
(3.12) |
up to a multiplicative constant, where denotes the Dirac measure at , and the Dirichlet form associated with this symmetric Feller’s Brownian motion on is
(3.13) | ||||
where
Proof.
We continue to denote by the resolvent of and by the killed process of upon hitting . Let represent the resolvent density of , i.e., . We have
(3.14) |
where , and is the Wronskain of (with ); see, e.g., [21, II§3, #7]. As established in [10, §15, 4.], for any , it holds that
(3.15) |
To demonstrate the necessity, let be a symmetric measure of . From [8, Lemma 4.1.3], it follows that is the symmetric measure of . Hence, without loss of generality, we may assume that is the Lebesgue measure on . For any , substituting (3.7) and (3.15) into
(3.16) |
we obtain
where . This implies that
is constant. Substituting (3.14) into , we have
where
It is straightforward to verify that is increasing in . Letting , we conclude that . Since is strictly positive for , it must hold that . Substituting (3.7) and (3.15) into (3.16) again, but now taking , we further obtain
This clearly implies
Particularly, if , then . This contradicts the condition (3.4). Therefore, we conclude that , and the symmetric measure is precisely (3.12).
Next, we consider the case where and . Note that (3.13) is, in fact, a regular Dirichlet form on ; see, e.g., [7, §5.3]. It suffices to show that
Indeed, , and hence (3.7) implies that . Using (3.7), we further have
(3.17) | ||||
Note that (see, e.g., [21, II§4, (23)])
This yields that
(3.18) |
Substituting (3.8), (3.15), and (3.18) into (3.17), we can obtain that
This completes the proof. ∎
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 ) or elastic Brownian motion (for ), 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.
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 starting at a point , let up to the first hitting time of . Then, make wait at for an exponential holding time with the conditional law
(3.19) |
at the end of this time (only applicable in the case where ), let it jump to a point in according to the distribution given by
(3.20) |
If the reaching point is in , let it start afresh; if it jumps to , let at all later times.
3.2.2.
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 -processes with ; see also [21, III§4].
To be precise, let us begin with a symmetric Feller’s Brownian motion with parameters . Let be the probability measure on defined as in (3.20). Then the Feller’s Brownian motion with parameters such that is actually the piecing out of with respect to , as described in [18, Appendix A]. Intuitively, repeatedly splices resurrection paths at the death times of , with the resurrection points randomly determined by the distribution .
3.2.3.
When , the pathwise construction becomes significantly more challenging. This was accomplished by Itô and McKean in [10, §12-§15].
We first consider the case where and (. Let be a subordinator, i.e., an increasing Lévy process on with , whose distribution has the Laplace transform
(3.21) |
In [10], this subordinator is referred to as an increasing differential process. For every , define
Further, let be a reflecting Brownian motion on , independent of , with local time at given by
(3.22) |
Note that the Revuz measure of with respect to is ; see, e.g., [22, X. Proposition 2.4]. It was shown in [10, §13] that
(3.23) |
is indeed a Feller’s Brownian motion with parameters . 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 , and .
Regarding the general case, we note that
(3.24) |
is the local time of the process defined by (3.23) at , as established in [10, §14]. Define and . Then, the time-changed process of (3.23) with respect to is
(3.25) |
This process is a Feller’s Brownian motion with parameters . The desired Feller’s Brownian motion with parameters is obtained as the subprocess of (3.25) perturbed by the multiplicative functional
(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 such that
(3.27) |
and then kill the process at time .
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 . What we focus on here is the local time of a general Feller’s Brownian motion at a given point .
Consider a Feller’s Brownian motion on :
with lifetime , where is the augmented natural filtration on . Let denote the -algebra generated by . We adjoin to the dead path with for all . A positive continuous additive functional of is defined as follows. For detailed discussions, see, e.g., [1, IV§1] and [3, Definition A.3.1].
Definition 3.6.
A family of functions from to is called a positive continuous additive functional (PCAF for short) of if there exists such that
and the following conditions are satisfied:
-
(A1)
For each , .
-
(A2)
For every , is continuous on , , for , and for .
-
(A3)
For and every , .
The set is called a defining set of . We further make the convention for all .
The fine support of a PCAF is defined as
where . According to [1, V. Theorem 3.13], for a given , there exists a PCAF (unique up to a multiplicative constant) of with fine support .
Recall that . The killed process can be written as
where is defined as in (3.11), for and for ; see, e.g., [25, (12.21i)]. The lifetime of is . We can define the PCAFs for and their fine supports analogously. The following fact is crucial to our discussion.
Lemma 3.7.
Given , let be a PCAF of with fine support . Then
is a PCAF of with fine support .
Proof.
Assuming without loss of generality that the defining set of is , it is straightforward to verify that satisfies properties (A1) and (A2). To establish property (A3) for , it suffices to consider the case where . If , then and . Hence, we have
(3.28) |
If , (3.28) can be verified using . Thus, is indeed a PCAF of . The fine support of is because and . ∎
According to Definition 3.1, is identical in law to the absorbing Brownian motion on . Therefore, we can define the Revuz measure of with respect to as follows:
Note that is a constant multiple of . 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 in the following sense.
Definition 3.8.
Given , a PCAF with is called the local time of at if the Revuz measure of with respect to is .
In the symmetric case, where and , we can also define the Revuz measure of the local time with respect to and the symmetric measure . According to [3, Proposition 4.1.10], this Revuz measure is also equal to the Dirac measure . 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 is regular for the minimal -process. Particularly, the scale function given by (2.2) satisfies
and the speed measure , 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 -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 -processes are the minimal -process and the -process, respectively.
4.1. Spatial transformation
The formulation we will present encounters a significant issue because the boundary point of Feller’s Brownian motion and the boundary point of the -process are located at opposite ends of their respective state spaces. Additionally, unlike Feller’s Brownian motion, the -process is not on the natural scale. However, both issues can be resolved by applying a straightforward spatial transformation to the -process. To address this, define for . Let
(4.1) |
and
(4.2) |
Clearly, can be extended to a homeomorphism between and , where is endowed with the relative topology of .
For each -process on ,
defines a continuous-time Markov chain on . It is a Feller process on whenever is a Feller -process. For convenience, we also refer to as a Doob process or a Feller -process (on or ).
4.2. Time change
We first prepare the ingredient, specifically the PCAF, for the time change transformation on a Feller’s Brownian motion . For , let denote the local time of at as defined in Definition 3.8. Define
(4.3) |
We will show that is a PCAF of .
Lemma 4.1.
The family of functions defined as (4.3) is a PCAF of the Feller’s Brownian motion . Furthermore,
-
(1)
If , and , then the fine support of is ;
-
(2)
Otherwise, the fine support of is .
Proof.
We first demonstrate that
(4.4) |
The main task is to estimate for each . Let . According to [1, V. Theorem 3.13], there exists a positive constant such that
(4.5) |
By the strong Markov property of and [1, IV, Proposition 1.13], we obtain
Integrating both sides with respect to the Lebesgue measure on , we have
where the second equality follows from [3, (4.1.3)]. Substituting (4.5) and for into the above equation, we obtain
(4.6) |
The strong Markov property also implies that for any ,
(4.7) | ||||
For , it holds that
(4.8) |
Substituting (4.7), (4.8) into (4.6) and using [11, Problem 6 of page 29], we deduce that
(4.9) |
Since , it follows that
Note that . Therefore, (4.4) can be concluded.
To prove that (4.3) is a PCAF of , we start by considering the defining sets of . Define
Clearly, . Note that
(4.10) |
Thus, it is straightforward to verify that and that satisfies all the conditions in Definition 3.1 except for
(4.11) |
To show (4.11), assume for contradiction that
This implies . Since is increasing in , is also increasing in . Therefore, there exists such that . We have
Finally, we examine the fine support of . Note that (as a measure in ) vanishes outside . Thus, vanishes on . Consequently, by [1, V. Corollary 3.10], we have . On the other hand, since , it follows from the definition that . Therefore, .
If , and , then for all , -a.s., where is given by (3.19). Note that . Hence, , where . By the definition of fine support, it follows that . Therefore, .
It remains to show for the remaining cases. We proceed by contradiction. Suppose . Then . For -a.s. and , we have for all . According to [1, V Theorem 3.8], for all . Note that is càdlàg in . From the pathwise representation of (see §3.2), we can obtain that if , then . This fact, combined with for and , implies that for all . Particularly, is continuous in . This is impossible when , because before the first jumping time, the sample paths of are those of a symmetric Feller’s Brownian motion with parameters . This process, as a regular diffusion process on , can not stay at any point for an extended period; see, e.g., [23, V. (47.1)]. For the case where , the pathwise representation of in §3.2.3 indicates that is exactly a Brownian path (up to a transformation of time change (3.25)) for . This also contradicts for . ∎
With the Feller’s Brownian motion and its PCAF given by (4.3), we can now introduce the time-changed process of with respect to . Define the right-continuous inverse of for each as
Further, let
Note that for , so for . According to [3, Proposition A.3.8(iv)], we may assume without loss of generality that for all and all . Set and . It is well known that the time-changed process
(4.12) |
with lifetime is a right process on ; see, e.g., [3, Theorem A.3.11].
Theorem 4.2.
Let be a Feller’s Brownian motion with parameters as specified in Theorem 3.3, and let be the time-changed process (4.12) of with respect to the PCAF (4.3). Then is a -process whose birth-death density matrix is (2.1), where is defined as (4.2). Furthermore,
-
(1)
If , and , then is the minimal -process;
-
(2)
If , and , then is a Doob process;
-
(3)
Otherwise, is a Feller -process.
Proof.
Denote the transition semigroup and resolvent of by and , respectively. Define
and
The goal is to prove that is a -process.
We first demonstrate that is a standard transition matrix. According to [26, §2.5, Theorem 1], it suffices to verify that satisfies the following conditions:
(4.13) | ||||
where is the Kronecker delta. To accomplish this, note that
Thus, the first condition in (4.13) is straightforward, and the third condition follows from the right continuity of . To prove the second condition, it is sufficient to demonstrate
(4.14) |
and then apply the resolvent equation of . In fact, it follows from the definition of , [22, Proposition 4.9 of page 8], and the Fubini theorem that
Thus, (4.14) is established.
Recall that , with lifetime , is an absorbing Brownian motion on . According to Lemma 3.7 and Lemma 4.1, we have that
is a PACF of , with Revuz measure . Define
with lifetime , as the time-changed process of with respect to the PACF . Specifically,
with
It has been established in [18, Lemma 3.1] that is exactly the minimal -process.
Next, let us examine the killed process of upon hitting . More precisely, let for all . From the definition (4.12) of , we find that for -a.s. , is increasing in for . Hence,
(4.15) |
is well defined for -a.s. and all . We will show that , and hence the killed process
with lifetime , on is identical to . To prove this, fix and let for . Note that and for all , -a.s. It follows from [1, V. Theorem 3.8] that for some and for all . Consequently,
and
These yield
and
In other words, , -a.s. Noting that , we obtain that .
According to the argument in the previous two paragraphs, we can conclude that is the transition matrix of the minimal -process. This implies that
(4.16) |
Mimicking the proof of [18, Proposition 3.6], we can also obtain that
(4.17) |
Combining (4.16) and (4.17), we obtain
In other words, is a -process.
Finally, let us classify the -process for different cases. For the first case, where and , we observe that does not increase after time (according to the sample path representation in §3.2.1). Consequently, for all , which implies that for all . Thus, aligns with the minimal -process. In the second case, where and , may continue to increase after time . A similar argument shows that is not the minimal -process. However, according to Lemma 4.1, is a right process with state space . Therefore, , -a.s. for all . Based on [18, Corollary 5.2], it follows that is a Doob process.
For the remaining cases, it suffices to show that , -a.s. for all . To demonstrate this, consider
If it were false that , then there would exist some such that . This implies that for all by [1, V. Theorem 3.8]. Since and is right continuous, it follows that for all . However, this contradicts the previous argument that cannot remain at for any extended period, as noted in the last paragraph of the proof of Lemma 4.1. Therefore, we have . ∎
In the context of general Markov process, the Feller’s Brownian motion with parameters is termed conservative if for any , which is equivalent to (see [10, §15]). In terms of continuous-time Markov chains, a conservative -process is also referred to as an honest -process (see §2.2).
Corollary 4.3.
If the Feller’s Brownian motion is conservative, then for any and ,
(4.18) |
Particularly, the -process , obtained in Theorem 4.2, is also conservative.
Proof.
Write . We first show that
(4.19) |
To demonstrate this, let . Observe that
The sample path of is described by (3.23) up to a time change transformation (see (3.25)). Given that , it follows that
Therefore, . On the other hand, by the strong Markov property, we have
Since , -a.s. on , it follows that , -a.s. on . Therefore, . Combining this with , we can eventually derive (4.19).
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 for which the corresponding time-changed process is a -process with the given density matrix (2.3).
Theorem 4.4.
Proof.
Let denote the local time of at , 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
Clearly, is a PCAF of . To verify the condition in [1, V§3, Proposition 3.11] for and , consider such that
Then . Since , we have
Applying [1, V§3, Proposition 3.11] to and , we obtain that
for some and .
The fact that is a -process implies that
(4.21) |
for and . However, the left-hand side of (4.21) can be expressed as
Hence, we conclude that .
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 , with lifetime , with respect to the PCAF corresponds precisely to the minimal -process. Particularly, the Revuz measure of
with respect to is actually the speed measure . Therefore, we can conclude that for all . ∎
5. Parameters of birth-death processes obtained by time change
According to Theorem 2.1, the -process obtained in Theorem 4.2 (excluding the first case, which yields the minimal -process) admits a resolvent representation given by a triple , which is unique up to a multiplicative constant. It is certainly interesting to explore how the parameters of Feller’s Brownian motion determine .
5.1. Main result
Recall that the sequence is given in (4.1). Based on the measure on , we define a sequence as follows:
(5.1) |
and
(5.2) |
Note that if , then . Our main result is as follows.
Theorem 5.1.
Let be a Feller’s Brownian motion with parameters as described in Theorem 3.3 (excluding the first case, i.e., , and , in Theorem 4.2). Then the parameters that determine the resolvent matrix of the -process , obtained in Theorem 4.2, are given by
up to a multiplicative constant, where 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 -process, as shown in Theorem 4.2, but also every -process can be derived from a Feller’s Brownian motion through time change.
Corollary 5.2.
For every -process , there exists a Feller’s Brownian motion such that the -process obtained from through time change, as described in Theorem 4.2, is identical in law to .
Proof.
Let be the triple determining the resolvent matrix of . When is the minimal -process, we can choose with and . In this case, the first case of Theorem 4.2 applies. When is a Doob process, i.e., , we can choose
Then, we apply the second case of Theorem 4.2 and Theorem 5.1. When is non-minimal and non-Doob, we can choose
(To satisfy (3.3), it may be necessary to multiply all parameters by a positive constant.) ∎
5.2. Pathwise construction of -processes
In this subsection, we discuss the pathwise construction of a Feller -process based on Theorem 5.1. For simplicity, we focus on the honest case where , and consider the Feller process on with parameters .
In the context of -processes, the -process plays a role analogous to that of the reflecting Brownian motion in the context of Feller’s Brownian motion. We denote its corresponding -process on by . This process is the time-changed process of the reflecting Brownian motion with respect to , where denotes the local time of at as defined in Definition 3.1. Let be the right-continuous inverse of . The local time of at , denoted by , can be derived from the Brownian local time through the corresponding time change transformation, specifically, .
We initially considered replacing with and constructing the general -process similarly to (3.23), by defining
(5.3) |
where is the subordinator (3.21) with and , which is independent of . (A similar idea appears in [10, §20].) However, at a time when increases to a discontinuous point () of , where for some and , the term in (5.3) will continuously decrease from for a short period thereafter (see [10, §12]). Meanwhile, ‘diffuses’ within the space near . This discrepancy causes the process defined by (5.3) to move outside of .
6. Proof of Theorem 5.1 for
In this section, we will prove Theorem 5.1 for various cases where . Simultaneously, we will provide a more comprehensive characterization of the corresponding -processes. From now on, we will denote for and use for convenience.
6.1. Doob processes
According to Theorem 4.2, is a Doob process (but not the minimal -process) if and only if , and .
Theorem 6.1.
If and , then the parameters that determine the resolvent matrix of the Doob process are given by
Particularly, the instantaneous distribution of is
Proof.
The fact that follows directly from Theorems 2.1 and 4.2. Fix for some , and let be defined as (4.15), representing the first flying time (see [18, Corollary 4.7]) of . According to [18, Theorem 5.1], it suffices to demonstrate that
(6.1) |
We will use the same symbols as those in the proof of Theorem 4.2. Define
and
Note that . From the pathwise representation of in §3.2.1, it follows that
It is straightforward to verify that and that (with the convention ). By the strong Markov property of , we have
Note that , where is given by (3.20). For , we have
with the convention . Thus, a straightforward computation yields
Finally, . Therefore, (6.1) is established. This completes the proof. ∎
6.2. Symmetric case
The symmetric case, where and , 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 on :
If , it follows from and the Cauchy-Schwarz inequality that exists.
Theorem 6.2.
If and , then the parameters determining the resolvent matrix of the -process are
(6.2) |
Furthermore, is symmetric with respect to the speed measure , and the associated Dirichlet form on is given by
Proof.
In this case, the Revuz measure of the PCAF (4.5) with respect to is exactly , by Definition 3.8 and [3, Proposition 4.1.10]. According to [3, Theorem 5.2.2], the time-changed process is symmetric with respect to , and its associated Dirichlet form on is derived in [3, (5.2.4)].
In what follows, we will compute . Regarding , we note that the extended Dirichlet space of (3.13) is
as detailed in [3, Theorem 2.2.11]. According to [3, (5.2.4)], it is straightforward to compute that
(6.3) |
where for all . The quadratic form can be formulated using [3, Theorem 5.5.9]. The crucial step is to compute the Feller measure on and the supplementary Feller measure on , 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 vanishes, and is supported on (with the notation indicating a pair of points, not an interval) with
Regarding the measure , we note that for any , where . Thus, by the definition [3, (5.5.7)] of , we have . Therefore, applying [3, Theorem 5.5.9], we obtain that for ,
(6.4) |
Applying the spatial transformation to (6.3) and (6.4) yields the expression for .
6.3. Non-symmetric case with finite jumping measure
We turn to examine the non-symmetric case where and . The sample path representation of Feller’s Brownian motion for this case is detailed in §3.2.2. Let be a probability measure on given by
(6.5) |
Theorem 6.3.
If and , then the parameters determining the resolvent matrix of the -process are
(6.6) |
Furthermore, can be obtained by piecing out with respect to the probability measure on , where is the symmetric -process corresponding to the parameters and is defined as (6.5).
Proof.
Let , with lifetime , be the symmetric Feller’s Brownian motion with parameters , as discussed in §3.2.2. It represents the killed process of at , and can be obtained by piecing out with respect to the probability measure given by (3.20). Particularly, we have
By following the steps used to obtain in the proof of Theorem 4.2, we can also show that the killed process of at time is identical in law to the time-changed process of with respect to the PCAF . Note that the Revuz measure of with respect to is exactly . Hence, according to Theorem 6.2, is identical in law to .
Denote by and the resolvents of and , respectively. For a bounded function on , Dynkin’s formula gives us:
(6.7) |
Let and . Then,
Note that . According to the construction procedures of piecing out as stated in [18, Appendix A], it is straightforward to see that (with distribution ) is independent of . Thus, is also independent of . From this independence and the strong Markov property of , it follows that
(6.8) | ||||
Similar to (6.4), we can also deduce that , where for all and . Substituting (6.8) into (6.7) yields
Integrating both sides with respect to , we get
It follows that
Finally, repeating the argument in the proof of [18, Theorem 8.1] (the steps after (8.5)) and applying the spatial transformation to , we can conclude that is the piecing out of with respect to . Particularly, the parameters for are given by (6.6). ∎
7. Approximation of Feller’s Brownian motion
We now turn to the ‘pathological’ case where . 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 -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 on , whose resolvent is expressed as (3.1) with . We refer to these processes as Doob’s Brownian motions, with 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 be a right-continuous function on . Consider two sequences of positive constants and such that
(These sequences may consist of finite numbers.) We say that the function is obtained from by the -transformation if
where and . Intuitively speaking, the -transformation discards the path of corresponding to the interval , keeps the segment unchanged, and shifts the remaining parts to the left, connecting them in the original order without intersection, thereby obtaining a new right-continuous path .
Fix . Define and (). We then define a sequence of stopping times as follows:
and if are already defined, we set
and
Note that if , then due to the quasi-left-continuity of . Particularly, . We define using the left limit of instead of directly using 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 requires the use of the left limit.
For and every , by performing the -transformation on , we obtain a new path, denoted by .
Lemma 7.1.
The process
is a Doob’s Brownian motion with instantaneous distribution
supported on .
Proof.
Fix and . We aim to compute the resolvent of for a positive and bounded Borel measurable function on :
Let and for . By the definition of , we have
(7.1) |
Set
Note that . It follows from the strong Markov property that
where is the resolvent of the absorbing Brownian motion.
Let us prove that
(7.2) |
For , we have , and thus, by the strong Markov property of ,
In general, note that and for all . Then the case can be deduced by the strong Markov property as follows:
Using and , we obtain that
The general case can be formulated similarly.
The sequence of Doob’s Brownian motions approximates in the case where in the following sense.
Theorem 7.2.
Assume that . Then, for any ,
(7.4) |
Proof.
A crucial consequence of the assumption is that
where 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 always indicates . 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 .
Fix . Let denote the total duration discarded from the path of when constructing for . Specifically, we have . It suffices to show
(7.5) |
Let us first consider the special case where is the reflecting Brownian motion . To differentiate this specific case from the general one, we denote by , respectively. Clearly, the instantaneous distribution of the approximating Doob’s Brownian motion, denoted by , is . Since is conservative by the expression of its resolvent in (3.1), we have for all . Note that is decreasing as . It follows that
which implies that
(7.6) | ||||
Now, consider a general Feller’s Brownian motion . According to its pathwise representation (3.23) (with lifetime given in (3.27)), or always implies . Since , it follows that . (Recall that the notation with tilde is defined for .) Consequently, each discarding interval for must be contained within some discarding interval for (though the converse is not necessarily true). Particularly, . Therefore, (7.5) follows from (7.6). ∎
Remark 7.3.
If we do not assume that , then the sequence of Doob’s Brownian motions will converge to a Feller’s Brownian motion with parameters in the sense of (7.4). This is because, according to Lemma 8.2, can be obtained from , with lifetime , through a time change transformation corresponding to a strictly increasing PCAF. Therefore, they share the same sequence of distributions , that is, , where .
7.2. Representation of instantaneous distributions
The following lemma describes the relationship between the instantaneous distributions.
Lemma 7.4.
For every with , the following relationship holds:
(7.7) |
Proof.
Take a positive and bounded Borel measurable function on . Note that , since . Using the strong Markov property of , we have
(7.8) |
For , it holds -a.s. that . Thus,
(7.9) | ||||
For , we have
(7.10) |
Since on , it follows that
(7.11) |
Using on and applying the strong Markov property of , we get
(7.12) | ||||
Substituting (7.11) and (7.12) into (7.10) yields
Therefore,
Based on this lemma, we can define a measure on as follows:
(7.13) |
where . Clearly, is a well-defined -finite measure on . The following lemma provides the representation of in terms of this measure .
Lemma 7.5.
For each , the following holds:
(7.14) |
and
(7.15) | ||||
Particularly,
(7.16) |
Proof.
According to Lemma 7.4, we have
It follows that
Let . Then is decreasing in , and
Particularly, (7.16) holds true. The induction also yields that
and by letting , we obtain
(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 , 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 , which is defined based on the approximating Doob’s Brownian motions.
Theorem 7.6.
Proof.
Define for (see (7.17)). This sequence is decreasing and converges to
(7.19) |
According to (7.3) and Lemma 7.5, the resolvent of can be expressed as
for all and . Note that
by virtue of the expression of provided in the first paragraph of the proof of Corollary 3.5. Additionally,
Given (7.16), we have
By Theorem 7.2, this limit is exactly the resolvent of . Comparing this expression with (3.15), we can eventually derive (7.18). ∎
8. Proof of Theorem 5.1 for
8.1. No sojourn case
In this subsection, we consider the case where and further assume that . According to Theorem 7.2, we can construct a sequence of Doob’s Brownian motions that approximates the Feller’s Brownian motion . Their instantaneous distributions are characterized in Lemma 7.5 in terms of the measure defined by (7.13).
Regarding the time-changed Feller’s Brownian motion , we can also construct a sequence of Doob processes with instantaneous distribution , as described in Appendix A, that approximates .
Lemma 8.1.
For all , the following holds:
for , and
where .
Proof.
We consider and address the other two cases in a similar manner. Note that
where for . It follows from the strong Markov property of that
Then applying Lemma 7.5, we can obtain the desired expression of . ∎
We are now prepared to prove Theorem 5.1 for the case where and .
Proof of Theorem 5.1 for .
Let
and
Substituting the expression of from Lemma 8.1 into (2.6), we can formulate the resolvent matrix of for and as follows:
Note that
as , due to (7.16). In addition, (see, e.g., [26, §7.10, (3) and (9)])
Therefore,
where is defined as (7.19). According to Theorem A.1, this limit is precisely the resolvent matrix of . Hence, the parameters of are (up to a multiplicative constant):
Using Theorem 7.6, we can eventually obtain the desired conclusion. ∎
8.2. Sojourn case
Let us consider the final case where and . The basic idea is to transform this case into one with . Recall that , defined by (3.23), is a Feller’s Brownian motion with parameters and , defined by (3.24), is its local time at . According to §3.2.3, the subprocess of , perturbed by the multiplicative functional
is a Feller’s Brownian motion with parameters . Denote by the lifetime of . Mimicking the proof of Lemma 3.7, we can easily show that
is a PCAF of with .
Lemma 8.2.
The time-changed process of with respect to the PCAF is a Feller’s Brownian motion with parameters .
Proof.
Given expressed as , according to [1, III, Theorem 3.3], we can write as follows:
(8.1) | ||||
and for with for ,
where is a probability measure on such that for all . Note that .
Let denote the right-continuous inverse of . This inverse is continuous and strictly increasing up to . The time-changed process of with respect to can be expressed as
and
where is the lifetime of . By substituting the expression (8.1) of into this expression of , we find that for ,
where is the right-continuous inverse of (). In other words, is the killed process of , defined in (3.25), at time . Note that for all ,
Thus, is the subprocess of perturbed by the multiplicative functional in (3.26). According to the pathwise representation in §3.2.3, we conclude that is a Feller’s Brownian motion with parameters . ∎
Now, we complete the proof of Theorem 5.1 as follows.
Proof of Theorem 5.1 for .
Let , with lifetime , be a Feller’s Brownian motion with parameters . According to Lemma 8.2, the Feller’s Brownian motion with parameters can be represented as the time-changed process of with respect to . The lifetime of is .
Denote by and the time-changed process obtained in Theorem 4.2 for and , respectively. The parameters determining the resolvent matrix of have been examined in §8.1. It suffices to show that is identical in law to . Let denote the local time of at as defined in Definition 3.1, and set for all . It is straightforward to verify that
is the local time of at in the sense of Definition 3.1. Thus, the PCAF of given by (4.3) is , and the right-continuous inverse of is
Since is strictly increasing up to , it follows that
where is the right-continuous inverse of . Particularly, for ,
(8.2) |
Note that is the lifetime of and is the lifetime of . Therefore, (8.2) establishes the identification between and . ∎
Appendix A Approximation of birth-death process
For a Feller -process , we can also construct a sequence of Doob processes in the same manner as described in Lemma 7.1. This approach is precisely the probabilistic construction method for all -processes presented in [26]. For the convenience of readers, we restate some necessary details as follows.
Let us consider the corresponding -process on , with lifetime denoted by . Define
and
Analogous sequences of stopping times and can be defined by repeating the procedures used before Lemma 7.1 (see also [26, §6.3]). Applying the -transformation to yields a sequence of Doob processes on , with instantaneous distribution .
The main result of [26, §6] establishes the convergence of to in the following sense.
Theorem A.1.
Let be a Feller -process and let denote the approximating sequence of Doob processes. Then, for any ,
(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 . For convenience, we will consider and instead of and . Denote by the lifetime of . According to the definition of the -transformation, it is straightforward to observe that
Let ().
Firstly, we show that there exists an integer such that for any ,
(A.2) |
It suffices to prove (A.2) for some integer , due to for all . To do this, note that we can find some and such that
(A.3) |
because otherwise
where is the set of all positive rational numbers. To prove (A.2) for with in (A.3), consider the case . We note that , and , -a.s. It follows that
Thus, there exists a constant such that
For , it follows from the strong Markov property that
By induction, we get for all . Hence,
This indicates (A.2) for the case . For general , (A.2) can also be established by using the strong Markov property. We provide details for the case , with the other cases handled by induction. In fact, we have
Next, we prove that for all ,
If , then according to the definition of , we have
where stands for the Lebesgue measure of the time set. Note that
It follows that
Thus, whenever . For the case , we prove that for all with satisfying (A.2). It suffices to consider the case where , because otherwise is honest by Theorem 2.1. If , then the total duration discarded from in constructing must be infinite. It follows from (A.2) that
By the strong Markov property, for any integer ,
Therefore,
This implies for .
We are now prepared to prove (A.1). Specifically, we need to establish that
for a fixed . For , let represent the total duration discarded from the path of in constructing for , namely, . Obviously, is decreasing as . If , then
If , then there exists an integer (which may depend on ) such that
This implies
We have
Thus
Therefore, by the right continuity of , we obtain
whenever . ∎
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.