by Jim Pitman
Introduction
When I first met Marc Yor around 1977, the theory of excursions of Brownian motion was at an early stage of development, with few applications. The theory had been initiated by Lévy [◊], and developed further by Itô-McKean [◊], Section 2.9. David Williams [◊] made the key observation that a Brownian excursion of length \( t \) is identical in law to a \( \operatorname{BES}(3) \) bridge from 0 to 0 of length \( t \). Here \( \operatorname{BES}(\delta) \) for \( \delta = 1,2, \ldots \) stands for a \( \delta \)-dimensional Bessel process, meaning the radial part of a \( \delta \)-dimensional Brownian motion. Williams [◊] introduced the key idea of decomposing the path of a one-dimensional diffusion at the time of its maximum. This identified various path fragments in Brownian motion with other path fragments derived from a \( \operatorname{BES}(3) \) process. \( \operatorname{BES}(3) \) then came to be understood as a convenient representation of Brownian motion on the positive half-line conditioned never to return to 0, in a sense made precise by Doob’s theory of \( h \)-transforms. Then Chung [◊] provided a detailed development of Lévy’s theory of excursions straddling a fixed time, emphasising parallels with the theory of last exit decompositions for Markov chains. But it was not until some years later that Williams [◊] explained the importance of his path decompositions for the theory of Brownian excursions. Around this time, I was interested in the theory of path decompositions and excursions, in the general Markovian setting of Itô [◊] and Maisonneuve [◊], while Marc was engaged with Jacques Azéma in work on local times of semi-martingales, the balayage formula associated with last exit times, and the Skorokhod embedding problem. Marc and I were both deeply impressed by the work of David Williams. We sensed there was much to be done to flesh out Williams’ ideas, which related various path decompositions of Brownian motion to the Ray–Knight representations of Brownian local time processes as squares of Bessel processes. This shared interest in the Brownian world was the starting point of a collaboration which continued for over 25 years. This collaboration also involved numerous other students and researchers, both in Berkeley, where Marc visited regularly each summer, and in Paris where I spent a number of sabbaticals.
In the early years of our collaboration, I provided expertise in Markovian excursion theory, while Marc was the master of martingale calculus. There was some friendly competition across the Atlantic, whether the laws of Brownian functionals were best computed by stochastic calculus or excursion theory. But it was not long before Marc was the master and I was the apprentice in both frameworks.
The perspectives of excursion theory and of martingale calculus are very different. In excursion theory, the Brownian path is viewed as a concatenation of a countable collection of random path fragments, each with its own internal dynamics, with timing of the concatenation by a local time clock. The fragments are dotted around some space of excursion paths, in which they can readily be subjected to various operations such as time-reversal, time-changes or other transformations. By comparison, to apply martingale calculus, you start with a given filtration, work with processes adapted to that filtration, enlarging the filtration when necessary to accommodate random times that are not stopping times. It is not so easy in this framework to cut up pieces of path and reassemble them, and have any idea what process has been created. Still, as Marc showed over and over again, each of these frameworks has much to offer the other. The key to many of Marc’s deepest contributions to the theory of Brownian motion and related processes was his mastery of both martingale calculus and excursion theory.
This article offers some higlights of Marc Yor’s work in the theory of excursions of Brownian motion and related processes, especially Bessel processes. See the articles of Bertoin [◊] and Le Gall [◊] for brief introductions to the context of this work. I follow quite closely the presentations of Pitman–Yor [◊] and [◊]. For much more extensive accounts of this work, see these articles and the monographs Revuz–Yor [◊], Yor [◊], Yen–Yor [◊], and Mallein–Yor [◊].
The agreement formula
In his fundamental paper
[◊],
Itô showed how to construct
a Poisson point process of excursions of a strong Markov process
\( X \) over time intervals when \( X \) is away from a recurrent point \( a \) of its statespace.
The point process is parameterized by the local time process of \( X \) at \( a \).
Each point of the excursion process is a path in a suitable space of
possible excursions of \( X \), starting at \( a \) at time 0, and returning
to \( a \) for the first time at some strictly positive time \( \zeta \), called
the lifetime of the excursion.
The intensity measure of the Poisson process of excursions is a \( \sigma \)-finite
measure \( \Lambda \) on the space of excursions, known as Itô’s excursion law.
For a reflecting Brownian motion \( X \) on \( [0,\infty) \),
there are the following two fundamental descriptions of Itô’s law \( \Lambda \) for excursions away from 0.
The first description is drawn
from Itô’s definition and observations of Lévy
[◊],
Itô-McKean
[◊],
and Williams
[◊].
The
second description is due to Williams
[◊]
and proved in Rogers
[◊].
Description I: Conditioning on the lifetime:
First pick a lifetime \( t \) according to the \( \sigma \)-finite density \( ({2 \pi})^{-1/2} t^{-3/2} \,dt \) on \( (0, \infty) \); thengiven \( t \),run a \( \operatorname{BES}(3) \) bridge from 0 to 0 over time \( t \).
Description II: Conditioning on the maximum:
First pick a maximum value \( m \) according to the \( \sigma \)-finite density \( m^{-2} \,dm \) on \( (0, \infty) \); thengiven \( m \),join back to back two independent \( \operatorname{BES}(3) \) processes, each started at 0 and run till it first hits \( m \).
Pitman–Yor [◊] generalized the equivalence of these two descriptions of Itô’s law of positive Brownian excursions as follows. The equivalence involving \( \operatorname{BES}(3) \) is extended to one involving \( \operatorname{BES}(\delta) \), as defined by Shiga–Watanabe [◊] for all real \( \delta \ge 0 \) using additivity properties of squares of Bessel processes, and as discussed further in Le Gall [◊], Section 5.
Description I: Conditioning on the lifetime:
First pick a lifetime \( t \) according to the \( \sigma \)-finite density \( 2^{- \delta/2} \Gamma ( \delta / 2 )^{-1} t^{- \delta/2} \,dt \) on \( (0, \infty) \); then given \( t \), run a \( \operatorname{BES}(\delta) \) bridge from 0 to 0 over time \( t \).
Description II: Conditioning on the maximum:
First pick a maximum value \( m \) according to the \( \sigma \)-finite density \( m^{1 - \delta } \,dm \) on \( (0, \infty) \); thengiven \( m \),join back to back two independent \( \operatorname{BES}(\delta) \) processes, each started at 0 and run till it first hits \( m \).
The measures \( \Lambda_{00}^\delta \) defined by Description II for \( \delta > 2 \) were considered already by Pitman–Yor [◊] and further studied by Biane–Yor [◊], who gave Description I in this case. It was shown in Pitman–Yor [◊] that for \( 2 < \delta < 4 \) the measure \( \Lambda_{00}^\delta \) is Itô’s excursion law for excursions of \( \operatorname{BES}(4-\delta) \) away from zero, up to multiplication by a constant depending on the normalization of local time. In particular, if \( n = n_{+} + n_{-} \) is the decomposition of Itô’s excursion law for excursions of standard Brownian motion into its parts for positive and negative excursions, with the usual normalization of local time as occupation density relative to Lebesgue measure, as in Revuz–Yor [◊], Section XII.4, then \[ \Lambda_{00}^3 = 2 n_{+} \] and Theorem 1 reduces to the preceding descriptions of Itô’s law \( \Lambda = 2 n_{+} \) for excursions of a reflecting Brownian motion.
For all \( \delta \ge 2 \) the measure \( \Lambda_{00}^\delta \) concentrates on excursion paths starting at 0 and first returning to 0 at their lifetime. But the measure with density \( t^{-\delta/2} \,dt \) on \( (0, \infty) \) is a Lévy measure only for \( 2 < \delta < 4 \). So for \( \delta \le 2 \) or \( \delta \ge 4 \) the measure \( \Lambda_{00}^\delta \) is not the excursion law of any Markov process. Nonetheless, these measures \( \Lambda_{00}^\delta \) are well defined for all \( \delta > 0 \), and have some interesting properties and applications. It was shown in in Pitman–Yor [◊] that the measure \( N \) defined as the distribution of the square of the path under \( 2 \Lambda_{00}^4 \) appears, due to the Ray–Knight description of Brownian local times, as the distribution of the total local time process of a path governed by the Itô’s law for positive Brownian excursions \( n_+ = \frac{1}{2} \Lambda_{00}^3 \). As a consequence, the same measure \( N \) appears in the Lévy–Khintchine representation of the infinitely divisible family of squares of Bessel processes and Bessel bridges found in Pitman–Yor [◊], and described in Le Gall [◊], Section 5.
For \( 0 < \delta < 2 \), the point 0 is a recurrent point for \( \operatorname{BES}(\delta) \), and the measure \( \Lambda_{00}^\delta \) concentrates on paths which, unlike excursions, return many times to 0 before finally being killed at 0. This \( \sigma \)-finite measure can be described in another way as follows: for a suitable normalization of local time, the total local time at 0 is distributed according to Lebesgue measure, and given that this local time is \( \ell \) the path is distributed like \( \operatorname{BES}(\delta) \) started at 0 and run until inverse local time \( \ell \). Such measures were first considered in Pitman–Yor [◊], Remark (3.9), and they played a role in the work of Biane–Yor [◊] and Biane–Le Gall–Yor [◊]. Closely related \( \sigma \)-finite measures on paths with infinite lifetime appear also in the study of limit laws obtained for Brownian penalisations by Najnudel–Roynette–Yor [◊], [◊]. See also the recent work by Marc’s students Najnudel–Nikeghbali [◊] for generalizations and variations of this theme, whereby for a suitable non-negative submartingale \( (X_t) \) relative to a filtered probability space \( (\Omega, \mathcal{F}, P, (\mathcal{F}_t)_{t \ge 0}) \) satisfying some technical conditions, there is a \( \sigma \)-finite measure \( Q \) on \( (\Omega, \mathcal{F}) \) such that for all \( t \ge 0 \) and all events \( F_t \in \mathcal{F}_t \), \[ Q( F_t, g \le t ) = P ( 1_{F_t} X_t ) \] where \( g \) is the last time that \( X \) hits 0.
In Pitman–Yor [◊] we established Theorem 1 for all \( \delta > 0 \) using a general formulation of Williams’ path decomposition at the maximum for one-dimensional diffusion bridges, due to Fitzsimmons [◊]. As an application of this decomposition, following the trail blazed for \( \delta =3 \) by Biane–Yor [◊], we described the law of the standard \( \operatorname{BES}(\delta) \) bridge by its density on path space relative to the law obtained by taking two independent \( \operatorname{BES}(\delta) \) processes started at 0 and run till they first hit 1, joining these processes back to back, and scaling the resultant process with a random lifetime and maximum 1 to have lifetime 1 and a random maximum.
Note especially the most important instance of the agreement formula \eqref{ran.5}, which arises with
\( F(R^{\mathrm{br}} ) \) and \( F( \tilde{R}^{\mathrm{br}} ) \) in \eqref{ran.5} replaced by
\( F(M^{\mathrm{br}} ) \) and \( F( \tilde{M}^{\mathrm{br}} ) \) where \( M^{\mathrm{br}} \) is the maximum of \( R^{\mathrm{br}} \), and the domain of \( F \) is now \( [0,\infty) \) instead of \( C[0,1] \).
This absolute continuity relation between the laws of the \( M^{\mathrm{br}} \) and \( \tilde{M}^{\mathrm{br}} \) is the heart of the matter, which is derived from the agreement formula for the \( \sigma \)-finite
measures by a Fubini argument. This style of argument and the scope of the agreement formula were gradually extended, starting from their first appearances in work of
Biane–Le Gall–Yor
[◊]
and Biane–Yor
[◊].
See Pitman–Yor
[◊],
[◊]
for variations of this splicing construction for Bessel processes, such as replacing first passage times by last passage times,
and for explicit descriptions by infinite series of the distribution of the maximum of a Bessel bridge for all real \( \delta > 0 \).
Brownian excursions and the Riemann Zeta function
Riemann [◊] showed that his zeta function, initially defined by the series \begin{equation} \label{zeta1} \zeta(s) :=\sum_{n=1}^{\infty}n^{-s} \quad (\Re s > 1) \end{equation} admits a meromorphic continuation to the entire complex plane, with only a simple pole at 1, and that the function \begin{equation} \label{xifn} \xi(s):= \textstyle\frac{1}{2} s ( s-1) \pi^{-s/2} \Gamma\bigl( \frac{1}{2} s \bigr) \zeta( s) \quad (\Re s > 1) \end{equation} is the restriction to \( (\Re s > 1) \) of a unique entire analytic function \( \xi \), which satisfies the functional equation \begin{equation} \label{xieq} \xi(s)= \xi (1-s) \end{equation} for all complex \( s \). Riemann showed that these basic properties of \( \zeta \) and \( \xi \) follow from a representation of \( 2 \xi \) as the Mellin transform of a function involving derivatives of Jacobi’s theta function. Biane–Yor [◊] identified this function with the probability density of the distribution of the maximum of a Brownian excursion, first found by Chung and Kennedy, to conclude that if \( M \) denotes the maximum of a Brownian excursion of length 1, and \( Y:= \sqrt{2/\pi}M \), then \begin{equation} \label{y1} E ( Y^s ) = 2 \xi (s ) \quad (s \in \mathbb{C} ) \end{equation} Biane–Yor showed that Riemann’s functional equation \eqref{xieq} is a consequence of the agreement formula \eqref{ran.5} for the maximum for \( \delta=3 \), combined with Chung’s remarkable identity in distribution that \begin{equation} \label{y2} Y^2 \stackrel{d}{=} \frac{\pi}{2} ( T + \hat{T} ) \end{equation} for \( T \) and \( \hat{T} \) the respective hitting times of 1 of two independent independent \( \operatorname{BES}(3) \) processes starting at 0, as in Corollary 2 for \( \delta = 3 \). See Yen–Yor [◊], Section 11.6, for a derivation of Chung’s identity via Brownian excursions, based on Yor [◊], Section 6.2. Many other constructions of random variables with the same distribution as \( Y \) have been discovered, involving functionals of the path of a Brownian motion or Brownian bridge in \( \mathbb{R}^d \) for \( d = 1,2,3 \) or 4. See Biane–Pitman–Yor [◊] for a survey of such relations between the distribution of Brownian functionals and the Riemann zeta function, and Pitman–Yor [◊] for further developments motivated by study of the distribution of ranked heights of excursions of a Brownian bridge.
Arcsine laws and interval partitions
For \( B \) a standard Brownian motion started at \( B_0 = 0 \), and \( t \ge 0 \), let \[ A_t := \int_0^t 1(B_t > 0 ) \,dt \] denote the amount of time that \( B \) spends positive up to time \( t \). It was noticed already by Lévy [◊] that distribution of \( A_T/T \) is the same, both for any fixed time \( T \), and for any inverse local time \( T = T_s := \inf \{t : L_t > s \} \), where \( L \) is the local time process of \( B \) at 0. It is easily shown, due to the independence of positive and negative Brownian excursions, and Brownian scaling, that distribution of \( A_{T_s}/T_s \) is identical to that of \( T_{s/2}/T_s \), where \( T_s = T_{s/2} + T_{s/2}^{\prime} \) for two independent and identically distributed stable\( (1/2) \) variables \( T_{s/2} \) and \( T_{s/2}^{\prime} \). It follows that this common distribution of \( A_{T_s}/T_s \) for all \( s > 0 \) is the arcsine or beta\( (1/2,1/2) \) distribution, where beta\( (a,b) \) for \( a > 0, b > 0 \) is the probability distribution on \( [0,1] \) whose density at \( 0 < u < 1 \) is proportional to \( u^{a-1}(1-u)^{b-1} \). Lévy showed by separate computations that this arcsine law was also the distribution of \( A_t/t \), as well as the distribution of \( G_t/t \), for each fixed time \( t \), where \( G_t:= \sup \{s \le t : B_s = 0 \} \), \( D_t:= \sup \{s > t : B_s = 0 \} \), so that \( (G_t, D_t) \) is the excursion interval of \( B \) that straddles \( t \), and \( G_t < t < D_t \) with probability one for each fixed \( t \).
Why there should be an identity in law between \( A_T/T \) between fixed times \( T \) and inverse local times \( T \) is not readily apparent. The point is that if \( T= T_s \) for some fixed \( s \), then \( B_T = 0 \) and \( G_T = T = D_T \) almost surely, so there is no excursion straddling \( T \). Considering the time spent positive \( A_T \) as a sum of times spent in positive excursions, for an inverse local time \( T \), the excursions are all completed. But for a fixed time \( T \) there is a collection of completed excursions up to time \( G_T \), followed by an incomplete excursion, known as a meander of length \( T- G_T > 0 \). There is no obvious way to relate the structure of the partition of \( [0,T] \) into excursion intervals between fixed and inverse local times \( T \). So this identity in distribution due to Lévy is not immediately explained by excursion theory.
To provide an adequate explanation of this identity in terms of Brownian excursions, Pitman–Yor [◊] considered the sequence \[ {\mathbf V}(t):=(V_1(t),V_2(t),\cdots), \quad V_1(t)\geq V_2(t)\geq\cdots, \] of ranked lengths of the maximal open subintervals of \( Z^c\cap(0,t) \), where \( Z \) is the closure of the range of a subordinator \( (T_s, s \ge 0) \). In the stable case of index \( 0 < \alpha < 1 \), we showed that \[ {\mathbf V}(t)/t \stackrel{d}{=} {\mathbf V}(T_s)/T_s \quad\text{for all} t > 0 \text{and} s > 0. \] In the case \( \alpha=\frac 12 \), when \( Z \) can be taken to be the zero set of a Brownian motion, Lévy’s identity in distribution of \( A_T/T \) for fixed and inverse local times \( T \) is an immediate consequence.
While for the range of a stable subordinator, the law of relative ranked lengths \( {\mathbf V}(T)/T \) is the same for both fixed and inverse local times \( T \), the way these intervals are arranged is not the same in the two cases. For \( T= T_s \), the arrangement forms an exchangeable interval partition of \( [0,T] \), meaning that (assuming for simplicity \( V_1(T) > V_2(T) > \cdots \)) the longest interval of length \( V_1(T) \) is equally likely to be to the right of the left of the second longest interval of length \( V_2(T) \), the longest 3 intervals are equally in any of the \( 3! \) possible orders, and so on. But for \( T \) a fixed time, there is the special meander interval \( (G_T,T] \), \( T- G_T = V_J(T) \) for some random index \( J \). Remarkably, for fixed times \( T \), the meander length \( V_J(T) \) turns out to be a sized biased pick from the sequence \( {\mathbf V}(T) \), meaning that \[ P(J=j \,|\, {\mathbf V}(T) ) = V_j(T)/T \quad ( j = 1,2, \ldots) \] Once this special length is selected from the ranked lengths to be placed on the extreme right, and even conditionally on the length of this interval, the order of remaining lengths, forming an interval partition of \( [0,G_T] \), is that of an exchangeable interval partition. In the Brownian case (\( \alpha = \frac{1}{2} \)) this is so because the interval partition of \( [0,G_T] \) is generated by a Brownian bridge of length \( G_T \), which conditionally given \( G_T \) has exchangeable increments. But the interval partitition generated by a stable subordinator has this property more generally for all \( 0 < \alpha < 1 \).
The two-parameter Poisson–Dirichlet distribution
This unexpected appearance of size-biased sampling in the analysis of excursion lengths of Brownian motion, and more generally for any Markov process whose zero set is the range of a stable subordinator, led to a sequence of papers which explored variations and extensions of this idea. The size factor proportional to \( t \) appears almost accidentally in this analysis, as the ratio of the mass of the Lévy measure to the right of \( t \) relative to the density of the Lévy measure at \( t \): \( t^{-\alpha}/t^{-\alpha - 1} = t \). But size-biased sampling is an operation on interval partitions, or on associated random discrete distributions, with natural interpretations in other contexts, in particular in Bayesian nonparametric statistics, in species sampling, in population genetics, and in modeling of time-evolutions of random partitions. In such contexts, it had been known for some time that it was often easiest to describe a size-biased random permutation of lengths, defined by putting interval lengths in the order they are discovered in a process of sampling by a sequence of independent uniform random points in the interval. This suggested that it should be fruitful to describe the size-biased random permutation of \( {\mathbf V}(T)/T \) derived as above from the range of a stable subordinator. This problem was solved by Perman–Pitman–Yor [◊] for a general subordinator, and led in the stable case to the two-parameter Poisson–Dirichlet distribution, denoted by \( \operatorname{PD}(\alpha, \theta \)), studied in detail in Pitman–Yor [◊]. This is a probability distribution for a sequence \( (V_n, n = 1,2, \ldots) \) with \( V_{1} > V_{2} > \cdots \) and \( \sum_{n}V_{n}=1 \). Let \( 0\leq \alpha < 1 \) and \( \theta > -\alpha \), and let \( W_n,\ n=1,2,\cdots, \) be independent random variables, with \( W_{n} \) having a beta(\( 1-\alpha,\theta+n\alpha \)) distribution. Put \[ \tilde{V}_{1}=W_{1}, \tilde{V}_{n}= (1-W_{1})\cdots (1-W_{n-1})W_{n},\quad n\geq 2, \] and let \( V_{1}\geq V_{2}\geq\cdots \) be the ranked values of \( \tilde{V}_{1},\tilde{V}_{2},\cdots \). Then \( (\tilde{V}_{1},\tilde{V}_{2},\cdots) \) is a size-biased random permutation of \( (V_{n}) \), whose distribution is \( \operatorname{PD}(\alpha,\theta \)). In particular, for \( 0 < \alpha < 1, \theta = 0 \), the \( \operatorname{PD}(\alpha,\theta \)) distribution is the distribution of ranked relative lengths \( {\mathbf V}(T)/T \) derived from the range of a stable\( (\alpha) \) subordinator \( (T_s) \), as above, either for \( T = t \) a fixed time, or for \( T= T_s \). And for \( 0 < \alpha < 1, \theta = \alpha \) this is the distribution of ranked relative lengths \( {\mathbf V}(G_T)/G_T \) in the same setting. When \( Z \) is the zero set of a Brownian motion or recurrent \( \operatorname{BES}(\delta) \) process \( B \), with \( \delta = 2 - 2 \alpha \), the path of \( B \) on \( [0,G_T] \) can be rescaled to a standard Brownian or \( \operatorname{BES}(\delta) \) bridge of length 1, from 0 to 0. So \( \operatorname{PD}(\alpha,\alpha \)) describes the distribution of ranked lengths of excursions of a standard Brownian or Bessel bridge of dimension \( 2 - 2 \alpha \). The key to this remarkably simple relation between ranked lengths of excursions of a standard Brownian or Bessel bridge and those of the corresponding unconditioned process is a basic absolute condinuity relation, due to Biane–Le Gall–Yor [◊] in the Brownian case, whereby the the bridge is represented by a change of measure relative to the process obtained by Brownian scaling of the unconditioned process up to an inverse local time. The density factor involved is just a constant times a power of the local time at 0. Such changes of measure provided a key to further analysis of the Poisson–Dirichlet family of random discrete distributions. For a review of developments and applications of this family in various settings see Pitman [◊]. The term Pitman–Yor process was introduced by Ishwaran–James [◊] for a random measure with \( \operatorname{PD}(\alpha,\theta) \) distributed atoms. Such random measures are now applied in the setting of Bayesian non-parametric statistics and related developments in machine learning, to model clustering phenomena exhibiting typical power-law behavior encountered in a number of contemporary applications. See [◊], [◊] for some recent developments with references to earlier work. The two-parameter Poisson–Dirichlet distribution has also been applied in the context of random fragmentation trees by [◊], and to fragmentation-coalescence processes by Bertoin [◊].
Last exit times
Hand in hand with excursion theory, a recurring theme of Marc’s work was the study of last exit times, especially for path fragments of Brownian motion and one-dimensional diffusions. His early work on last exits relied on his balayage formula for semi-martingales, which is closely related to the Azéma–Yor [◊] solution of Skorokhod’s problem. This problem is to find, for a suitable probability distribution \( \mu \) on \( \mathbb{R} \), a stopping time \( T \) of Brownian motion \( (B_t, t \ge 0 ) \) such that \( B_T \) has distribution \( \mu \), and \( T \) is in some sense as small as possible. The result of Azéma–Yor is that if \( \mu \) has mean 0, and \( T_\mu:= \inf \{t: S_t \ge \psi_\mu(B_t) \} \), where \( \psi_\mu(x) \) is the mean of \( \mu \) conditioned on \( [x,\infty) \), then \( B_{T_\mu} \) has distribution \( \mu \), and the martingale \( (B_{T_\mu} \wedge t, t \ge 0) \) is uniformly integrable. See Yen–Yor [◊], Sections 3.4 and 3.6, for a review. Over time, Marc’s view of balayage, last exits and the Azéma–Yor solution of Skorokhod’s problem evolved to a point where these notions were fully integrated with Itô’s theory applied to Brownian excursions, as in Obłój–Yor [◊]. See also Obłój [◊] for a survey of the large number of solutions to Skorokhod’s problem, amongst which the Azéma–Yor solution stands out as both the most elegant and most explicit.
In Pitman–Yor [◊], Theorem 6.1, we used martingale calculus to find a formula for the probability density of the last exit time of an upwardly transient diffusion process, and indicated a number of applications to Bessel processes and properties of Bessel functions. In principle, the idea is very simple. For a discrete time transient Markov chain, with transition probabilities \( p(t,x,y) \), every time the chain exits a state \( y \) could be the last time, if the process never returns. So starting at \( x \) the probability that the last time in state \( y \) is at time \( t \) is \( p(t,x,y) \rho(y) \) where \( \rho(y) = P_y( T_y = \infty) \) for \( T_y \) the time of first return to \( y \). For a continuous time diffusion, the idea is that every time the process returns to \( y \) (now a continuum of times) there is a risk of ultimately leaving \( y \) at rate \( \rho(y) \) per unit local time at \( y \), where \( \rho(y) \) is the intensity of excursions away from \( y \) with infinite lifetime. This is in the extension of Itô’s theory to excursions of a Markov process away from a transient state, indicated by Meyer [◊]. Consequently, under regularity conditions which ensure that a diffusion process has a nice transition density \( p(t,x,y) \) relative to its speed measure \( m(dy) \), the probability density at time \( t \) of the last exit time of state \( y \) for the diffusion started in state \( x \) is just \( p(t,x,y) \rho(y) \) for some \( \rho(y) \), which is just a normalization constant related to the normalization of local time at \( y \) and the rate of last exits from \( y \). Martingale calculus also allows \( \rho(y) \) to be expressed in terms of the scale function of the diffusion. In subsequent work, Marc returned often to this theme, developing some remarkable connections between last exit times of Bessel processes and functionals of geometric Brownian motion of interest in mathematical finance. See for instance Yor [◊], [◊]. Starting from such examples, and influenced by the general theory of enlargement of filtations, which Marc developed in collaboration with Thierry Jeulin in the 1980’s, Marc and his students and collaborators developed a general theory of last times defined as the ends of optional or predicable sets relative to a given filtration. They have also shown how this theory of last exit times has natural interpretations and applications in mathematical finance. See the text of Profeta–Roynette–Yor [◊], and the survey of Nikeghbali–Platen [◊] for recent overviews of this work.
En guise de conclusion
From “The Last Time” by Mick Jagger and Keith Richards (1965), a refrain we enjoyed for many years, until finally the last time came:
Well this could be the last time
This could be the last time
Maybe the last time
I don’t know. Oh no. Oh no