3.1. Laplace Exponent, the Reflected Process, Local Times and Excursions from the Supremum, Supremum Process and Long-Term Behaviour, Exponential Change of Measure
Since the Poisson process admits exponential moments of all orders, it follows that
and, in particular,
for all
. Indeed, it may be seen by a direct computation that for
,
,
, where
is the Laplace exponent of
X. Moreover,
is continuous (by the DCT) on
and analytic in
(use the theorems of Cauchy (
Rudin 1970, p. 206, 10.13 Cauchy’s theorem for triangle), Morera (
Rudin 1970, p. 209, 10.17 Morera’s theorem) and Fubini).
Next, note that tends to as over the reals, due to the presence of the atom of at h. Upon restriction to , is strictly convex, as follows first on by using differentiation under the integral sign and noting that the second derivative is strictly positive, and then extends to by continuity.
Denote then by the largest root of . Indeed, 0 is always a root, and due to strict convexity, if , then 0 and are the only two roots. The two cases occur, according as to whether or , which is clear. It is less obvious, but nevertheless true, that this right derivative at 0 actually exists, indeed . This follows from the fact that is nonincreasing as for and hence the monotone convergence applies. Continuing from this, and with a similar justification, one also gets the equality (where we agree if ). In any case, is continuous and increasing, it is a bijection and we let be the inverse bijection, so that .
With these preliminaries having been established, our first theorem identifies characteristics of the reflected process, the local time of
X at the maximum (for a definition of which see e.g., (
Kyprianou 2006, p. 140, Definition 6.1)), its inverse, as well as the expected length of excursions and the probability of an infinite excursion therefrom (for definitions of these terms see e.g., (
Kyprianou 2006, pp. 140–47); we agree that an excursion (from the maximum) starts immediately after
X leaves its running maximum and ends immediately after it returns to it; by its length we mean the amount of time between these two time points).
Theorem 1 (Reflected process; (inverse) local time; excursions)
. Let for and .
The generator matrix of the Markov process on is given by (with ): , unless , in which case we have .
For the reflected process Y, 0 is a holding point. The actual time spent at 0 by Y, which we shall denote L, is a local time at the maximum. Its right-continuous inverse , given by (for ; otherwise), is then a (possibly killed) compound Poisson subordinator with unit positive drift.
Assuming that to avoid the trivial case, the expected length of an excursion away from the supremum is equal to ; whereas the probability of such an excursion being infinite is .
Assume again to avoid the trivial case. Let N, taking values in , be the number of jumps the chain makes before returning to its running maximum, after it has first left it (it does so with probability 1
). Then the law of is given by (for ): In particular, has a killing rate of , Lévy mass and its jumps have the probability law on given by the length of a generic excursion from the supremum, conditional on it being finite, i.e., that of an independent N-fold sum of independent -distributed random variables, conditional on N being finite. Moreover, one has, for , , where the coefficients satisfy the initial conditions:the recursions:and may be interpreted as the probability of X reaching level 0 starting from level for the first time on precisely the k-th jump ().
Proof. Theorem 1-1 is clear, since, e.g., Y transitions away from 0 at the rate at which X makes a negative jump; and from to 0 at the rate at which X jumps up by s or more etc.
Theorem 1-2 is standard (
Kyprianou 2006, p. 141, Example 6.3 & p. 149, Theorem 6.10).
We next establish Theorem 1-3. Denote, provisionally, by
the expected excursion length. Furthermore, let the discrete-time Markov chain
W (on the state space
) be endowed with the initial distribution
for
,
; and transition matrix
P, given by
, whereas for
:
, if
;
, if
; and
otherwise (
W jumps down with probability
p, up
i steps with probability
,
, until it reaches 0, where it gets stuck). Further let
N be the first hitting time for
W of
, so that a typical excursion length of
X is equal in distribution to an independent sum of
N (possibly infinite)
-random variables. It is Wald’s identity that
. Then (in the obvious notation, where
indicates the sum is inclusive of ∞), by Fubini:
, where
is the mean hitting time of
for
W, if it starts from
, as in (
Norris 1997, p. 12). From the skip-free property of the chain
W it is moreover transparent that
,
, for some
(with the usual convention
). Moreover we know (
Norris 1997, p. 17, Theorem 1.3.5) that
is the minimal solution to
and
(
). Plugging in
, the last system of linear equations is equivalent to (provided
)
, where
. Thus, if
, the minimal solution to the system is
,
, from which
follows at once. If
, clearly we must have
, hence
and thus
.
To establish the probability of an excursion being infinite, i.e., , where , we see that (by the skip-free property) , , and by the strong Markov property, for , . It follows that , i.e., . Hence, by Theorem 2-2, whose proof will be independent of this one, (since , if and only if X drifts to ).
Finally, Theorem 1-4 is straightforward. ☐
We turn our attention now to the supremum process . First, using the lack of memory property of the exponential law and the skip-free nature of X, we deduce from the strong Markov property applied at the time , that for every , : In particular, for any : And since for , (-a.s.) one has (for ): . Therefore .
Next, to identify
,
, observe that (for
,
):
and hence
is an
-martingale by stationary independent increments of
X, for each
. Then apply the optional sampling theorem at the bounded stopping time
(
) to get:
Please note that
and
converges to
(
-a.s.) as
on
. It converges to
on the complement of this event,
-a.s., provided
. Therefore we deduce by dominated convergence, first for
and then also for
, by taking limits:
Before we formulate our next theorem, recall also that any non-zero Lévy process either drifts to
, oscillates or drifts to
(
Sato 1999, pp. 255–56, Proposition 37.10 and Definition 37.11).
Theorem 2 (Supremum process and long-term behaviour)
. The failure probability for the geometrically distributed is ().
X drifts to , oscillates or drifts to according as to whether is positive, zero, or negative. In the latter case has a geometric distribution with failure probability .
is a discrete-time increasing stochastic process, vanishing at 0 and having stationary independent increments up to the explosion time, which is an independent geometric random variable; it is a killed random walk.
Remark 2. Unlike in the spectrally negative case (Bertoin 1996, p. 189), the supremum process cannot be obtained from the reflected process, since the latter does not discern a point of increase in X when the latter is at its running maximum. Proof. We have for every
:
Thus Theorem 2-1 obtains.
For Theorem 2-2 note that letting
in (
2), we obtain
(
-a.s.), if and only if
, which is equivalent to
. If so,
is geometrically distributed with failure probability
and then (and only then) does
X drift to
.
It remains to consider the case of drifting to
(the cases being mutually exclusive and exhaustive). Indeed,
X drifts to
, if and only if
is finite for each
(
Bertoin 1996, p. 172, Proposition VI.17). Using again the nondecreasingness of
in
, we deduce from (
1), by the monotone convergence, that one may differentiate under the integral sign, to get
. So the
are finite, if and only if
(so that
-a.s.) and
. Since
is the inverse of
, this is equivalent to saying
.
Finally, Theorem 2-3 is clear. ☐
Table 1 briefly summarizes for the reader’s convenience some of our main findings thus far.
We conclude this section by offering a way to reduce the general case of an upwards skip-free Lévy chain to one which necessarily drifts to . This will prove useful in the sequel. First, there is a pathwise approximation of an oscillating X, by (what is again) an upwards skip-free Lévy chain, but drifting to infinity.
Remark 3. Suppose X oscillates. Let (possibly by enlarging the probability space to accommodate for it) N be an independent Poisson process with intensity 1 and () so that is a Poisson process with intensity ϵ, independent of X. Define . Then, as , converges to X, uniformly on bounded time sets, almost surely, and is clearly an upwards skip-free Lévy chain drifting to .
The reduction of the case when X drifts to is somewhat more involved and is done by a change of measure. For this purpose assume until the end of this subsection, that X is already the coordinate process on the canonical space , equipped with the -algebra and filtration of evaluation maps (so that coincides with the law of X on and , while for , , where , for ). We make this transition in order to be able to apply the Kolmogorov extension theorem in the proposition, which follows. Note, however, that we are no longer able to assume the standard conditions on . Notwithstanding this, remain -stopping times, since by the nature of the space , for , , .
Proposition 1 (Exponential change of measure)
. Let . Then, demanding:this introduces a unique measure on . Under the new measure, X remains an upwards skip-free Lévy chain with Laplace exponent , drifting to , if , unless . Moreover, if is the new Lévy measure of X under , then and λ-a.e. in . Finally, for every -stopping time T, on restriction to , and: Proof. That
is introduced consistently as a probability measure on
follows from the Kolmogorov extension theorem (
Parthasarathy 1967, p. 143, Theorem 4.2). Indeed,
is a nonnegative martingale (use independence and stationarity of increments of
X and the definition of the Laplace exponent), equal identically to 1 at time 0.
Furthermore, for all
,
and
:
An application of the Functional Monotone Class Theorem then shows that X is indeed a Lévy process on and its Laplace exponent under is as stipulated (that -a.s. follows from the absolute continuity of with respect to on restriction to ).
Next, from the expression for , the claim regarding follows at once. Then clearly X remains an upwards skip-free Lévy chain under , drifting to , if .
Finally, let
and
. Then
, and by the Optional Sampling Theorem:
Using the MCT, letting , we obtain the equality . ☐
Proposition 2 (Conditioning to drift to +∞)
. Assume and denote (see (3)). We then have as follows. For every , .
For every , the stopped process is identical in law under the measures and on the canonical space .
Proof. With regard to Proposition 2-1, we have as follows. Let
. By the Markov property of
X at time
t, the process
is identical in law with
X on
and independent of
under
. Thus, letting
(
), one has for
and
, by conditioning:
since
. Next, noting that
:
The second term clearly converges to
as
. The first converges to 0, because by (
2)
, as
, and we have the estimate
.
We next show Proposition 2-2. Note first that
X is
-progressively measurable (in particular, measurable), hence the stopped process
is measurable as a mapping into
(
Karatzas and Shreve 1988, p. 5, Problem 1.16).
Furthermore, by the strong Markov property, conditionally on , is independent of the future increments of X after , hence also of for any . We deduce that the law of is the same under as it is under for any . Proposition 2-2 then follows from Proposition 2-1 by letting tend to , the algebra being sufficient to determine equality in law by a /-argument. ☐
3.2. Wiener-Hopf Factorization
Definition 2. We define, for , , i.e., -a.s., is the last time in the interval that X attains a new maximum. Similarly we let be, -a.s., the last time on of attaining the running infimum ().
While the statements of the next proposition are given for the upwards skip-free Lévy chain
X, they in fact hold true for the Wiener-Hopf factorization of
any compound Poisson process. Moreover, they are (essentially) known in
Kyprianou (
2006). Nevertheless, we begin with these general observations, in order to (a) introduce further relevant notation and (b) provide the reader with the prerequisites needed to understand the remainder of this subsection. Immediately following Proposition 3, however, we particularize to our the skip-free setting.
Proposition 3. Let . Then:
Proof. These claims are contained in the remarks regarding compound Poisson processes in (
Kyprianou 2006, pp. 167–68) pursuant to the proof of Theorem 6.16 therein. Analytic continuations have been effected in part Proposition 3-3 using properties of zeros of holomorphic functions (
Rudin 1970, p. 209, Theorem 10.18), the theorems of Cauchy, Morera and Fubini, and finally the finiteness/integrability properties of
q-potential measures (
Sato 1999, p. 203, Theorem 30.10(ii)). ☐
Remark 4. (Kyprianou 2006, pp. 157, 168) is also the Laplace exponent of the (possibly killed) bivariate descending ladder subordinator , where is a local time at the minimum, and the descending ladder heights process (on ; otherwise) is X sampled at its right-continuous inverse : As for the strict ascending ladder heights subordinator (on ; otherwise), being the right-continuous inverse of , and denoting the amount of time X has spent at a new maximum, we have, thanks to the skip-free property of X, as follows. Since , X stays at a newly achieved maximum each time for an -distributed amount of time, departing it to achieve a new maximum later on with probability , and departing it, never to achieve a new maximum thereafter, with probability . It follows that the Laplace exponent of is given by:(where ). In other words, is a killed Poisson process of intensity and with killing rate .
Again thanks to the skip-free nature of X, we can expand on the contents of Proposition 3, by offering further details of the Wiener-Hopf factorization. Indeed, if we let and (, ) then clearly are the arrival times of a renewal process (with a possibly defective inter-arrival time distribution) and is the ‘number of arrivals’ process. One also has the relation: , (-a.s.). Thus the random variables entering the Wiener-Hopf factorization are determined in terms of the renewal process .
Moreover, we can proceed to calculate explicitly the Wiener-Hopf factors as well as
and
. Let
. First, since
is a geometrically distributed random variable, we have, for any
:
Note here that
for all
. On the other hand, using conditioning (for any
):
Now, conditionally on
,
is independent of
and has the same distribution as
. Therefore, by (
1) and the theorem of Fubini:
We identify from (
4) for any
:
and therefore for any
:
We identify from (
5) for any
:
Therefore, multiplying the last two equalities, for
and
, the equality:
obtains. In particular, for
and
, we recognize for some constant
:
. Next, observe that by independence and duality (for
and
):
Both sides of this equality are continuous in
and analytic in
. They agree on
, hence agree on
by analytic continuation. Therefore (for all
,
):
i.e., for all
and
for which
one has:
Moreover, for the unique
, for which
, one can take the limit
in the above to obtain:
. We also recognize from (
7) for
and
with
, and some constant
:
. With
one can take the limit in the latter as
to obtain:
.
In summary:
Theorem 3 (Wiener-Hopf factorization for upwards skip-free Lévy chains)
. We have the following identities in terms of ψ and Φ:
For every and :and(the latter whenever ; for the unique such that , i.e., for , one has the right-hand side given by ). For some and then for every and :and(the latter whenever ; for the unique such that , i.e., for , one has the right-hand side given by ).
☐
As a consequence of Theorem 3-1, we obtain the formula for the Laplace transform of the running infimum evaluated at an independent exponentially distributed random time:
(and
). In particular, if
, then letting
in (
8), one obtains by the DCT:
We obtain next from Theorem 3-2 (recall also Remark 4-1), by letting
therein, the Laplace exponent
of the descending ladder heights process
:
where we have set for simplicity
, by insisting on a suitable choice of the local time at the minimum. This gives the following characterization of the class of Laplace exponents of the descending ladder heights processes of upwards skip-free Lévy chains (cf. (
Hubalek and Kyprianou 2011, Theorem 1)):
Theorem 4. Let , , and , with . Then:There exists (in law) an upwards skip-free Lévy chain X with values in and with (i) γ being the killing rate of its strict ascending ladder heights process (see Remark 4-2), and (ii) , , being the Laplace exponent of its descending ladder heights process.
if and only if the following conditions are satisfied: .
Setting x equal to 1, when , or to the unique solution of the equation:on the interval , otherwise2; and then defining , , ; it holds:
Such an X is then unique (in law), is called the parent process, its Lévy measure is given by , and .
Remark 5. Condition Theorem 4-2 is actually quite explicit. When (equivalently, the parent process does not drift to ), it simply says that the sequence should be nonincreasing. In the case when the parent process X drifts to (equivalently, (hence )), we might choose first, then , and finally γ.
Proof. Please note that with
,
, and comparing the respective Fourier components of the left and the right hand-side, (
10) is equivalent to:
.
.
, .
Moreover, the killing rate of the strict ascending ladder heights processes expresses as , whereas (1) and (3) alone, together imply .
Necessity of the conditions. Remark that the strict ascending ladder heights and the descending ladder heights processes cannot simultaneously have a strictly positive killing rate. Everything else is trivial from the above (in particular, we obtain that such an X, when it exists, is unique, and has the stipulated Lévy measure and ).
Sufficiency of the conditions. The compound Poisson process
X whose Lévy measure is given by
(and whose Laplace exponent we shall denote
, likewise the largest zero of
will be denoted
) constitutes an upwards skip-free Lévy chain. Moreover, since
, unless
, we obtain either way that
with
,
. Substituting in this relation
, we obtain at once that if
(so
), that then
X drifts to
,
, and hence
is the killing rate of the strict ascending ladder heights process. On the other hand, when
, then
, and a direct computation reveals
. So
X does not drift to
, and
, whence (again)
. Also in this case, the killing rate of the strict ascending ladder heights process is
. Finally, and regardless of whether
is strictly positive or not, compared with (
10), we conclude that
is indeed the Laplace exponent of the descending ladder heights process of
X. ☐