1. Introduction
Lévy processes are used in many applications, including ones to finance, insurance, physics, and reliability theory, see, e.g., [
1,
2] and the references therein. A common approach to the simulation of a Lévy process is to use its shot-noise representation, which represents the Lévy process as an infinite sum. This idea goes back to the pioneering work of Khintchine in 1931 [
3], see also [
4] and the references therein for recent developments. To use a shot-noise representation for simulation, one must truncate the infinite series at some finite point. It has been argued (see, e.g., [
1] or [
2]) that this truncation point should not be deterministic but random, and should be chosen so as to control the magnitudes of the jumps that are excluded. In this paper, we introduce the remainder process, which represents the jumps that are removed due to the truncation. This process has two parameters, one representing time in the original Lévy process and the other representing the cut-off beyond which jumps are removed. Throughout, we focus on subordinators, which are Lévy processes with only positive jumps. They are the building blocks from which all finite variation Lévy processes are built, even in the multivariate case, see [
5]. We characterize when the remainder process of a subordinator is self-similar and show that the resulting class can be indexed by a parameter
. We further show that the remainder process can be written as a continuous function of
and that, under mild assumptions, this function is monotonically increasing. When
, it corresponds to the classes of
-stable and truncated
-stable subordinators. From this perspective, the distinction between these two is not relevant as they have the same remainder processes on bounded intervals. When
, the process corresponds to the class of generalized
h-Dickman distributions, and when
, it corresponds to certain compound Poisson distributions. In this context, generalized
h-Dickman distributions play the role of 0-stable distributions.
There has been some interest in finding a natural analogue of
-stable distributions for
. In [
6,
7], it is argued that gamma distributions are a natural analogue. The idea is as follows. First, apply an Esscher transform to an
-stable distribution to obtain a certain type of tempered
-stable distribution. Next, take the limit as
, which leads to a gamma distribution. We note that a similar approach can be made with other classes of tempered stable distributions, which lead to a variety of other distributions in the limit, see [
8] or [
9]. From a different perspective, ref. [
10] showed that, as
, the distribution of
, where
is an
-stable random variable and
is an appropriate shift, approaches a reciprocal exponential distribution. In [
11,
12], the Dickman distribution is called a “truncated 0-stable subordinator” due to the form of its Lévy measure. All of these approaches to thinking about 0-stable distributions are valid from different perspectives. From the perspective of the remainder process, generalized
h-Dickman distributions are the natural analogue.
The rest of this paper is organized as follows. In
Section 2, we review basic properties of infinitely divisible distributions and their associated Lévy processes. In
Section 3, we introduce the remainder process for subordinators, and in
Section 4, we give some examples. In
Section 5, we give our main results about continuity and monotonicity, and we fully characterize when remainder processes are self-similar. In
Section 6, we give two useful lemmas. In the interest of generality, we present these in the multivariate setting as they may be of independent interest.
Before proceeding, we introduce our notation. We write “w.p. 1” to mean “with probability 1”, “increasing” to mean non-decreasing, and “decreasing” to mean “non-increasing”. We write to denote the Borel sets on , to denote the indicator function of set A, and ∅ to denote the empty set. We write ∨ and ∧ to denote the maximum and minimum, respectively. For a distribution , we write to denote its characteristic function, to denote that X is a random variable with distribution , and to denote that are independent and identically distributed (iid) random variables with distribution . We write , , and to denote, respectively, a defining equality, an equality in distribution, and equality in finite dimensional distributions (fdd). We write to denote the uniform distribution on and to denote an exponential distribution with rate . Unless otherwise specified, throughout this paper, we let and be independent sequences of random variables and we set for .
2. Infinitely Divisible Distributions and Lévy Processes
The characteristic function of an infinitely divisible (ID) distribution
can be written in the form
where
is the Gaussian part,
is the shift, and
M is the Lévy measure, which is a Borel measure satisfying
The parameters
,
M, and
b uniquely determine this distribution and we write
. Associated with every ID distribution
is a Lévy process
, where
. Lévy processes are characterized by independent and stationary increments, càdlàg paths, stochastic continuity, and the initial condition
w.p. 1. Of these, only stochastic continuity and càdlàg paths cannot be determined by finite dimensional distributions. This leads to the following.
Lemma 1. If two stochastic processes are equal in fdd and one of them is a Lévy process, then the other is also a Lévy process if and only if it is stochastically continuous and has càdlàg paths.
The jumps of a Lévy process are governed by its Lévy measure. Specifically, for a Lévy process
with
and any
with
,
is the expected number of jumps that the process has in the time interval
that falls inside
B. A Lévy process has finite variation if and only if
and
M satisfies the additional condition
Through a slight abuse of terminology, we also say that the associated distribution
and the associated Lévy measure
M have finite variation. In this case, the characteristic function can be written in the form
where
is the drift, and we write
. If, in addition,
and
, then the associated Lévy process is increasing and is called a subordinator. Again, by a slight abuse of terminology, we call the associated distribution
a subordinator as well. For more on infinitely divisible distributions and Lévy processes see [
2] or [
13].
3. Shot-Noise and the Remainder Process
Let
be a subordinator. For simplicity and without loss of generality, we take
throughout. For
, define the truncated Lévy measure
. We refer to a subordinator with Lévy measure
as a truncated subordinator. For
, let
Note that these functions uniquely determine the corresponding Lévy measures, that they are decreasing, and that
For notational convenience, we write
,
, and
. Let
be the generalized inverse of
V. We use a similar definition for
. Note that
and
are decreasing.
Fix a finite time horizon
and consider the process
This is a Lévy process with
, and the infinite sum converges almost surely and uniformly for
, see, e.g., Proposition 6.3 in [
2]. We call
the jumps of the process. Note that they are ordered from largest to smallest. This series representation is sometimes called a shot-noise representation. It can be used to approximately simulate the Lévy process. To do so, we must first truncate the infinite series at some finite value. It has been argued, see e.g., [
1] or [
2], that the proper way to truncate is not at a deterministic point. Instead, we should choose some
and truncate at the random time
, where
. This way, we control the magnitudes of the jumps that we are removing. We take
for some
. We will see that this is equivalent to removing jumps of magnitudes greater that
s. The remainder in this approximation can be written as
We call this the remainder process with Lévy measure
L. We have not previously seen processes of this type in the literature. We now give a lemma that is useful when working with remainder processes. First, for
, we define the set
.
Lemma 2. 1. For any , we have
2. If and Y is a positive random variable with , then = w.p. 1.
3. There is at most a countable number of for which has more than one element.
Proof. The first part follows from the fact that
where the second line follows from the fact that
for
and the fourth from the fact that
for
. For the second part, if
then, since
V is decreasing, we have
. Similarly, if
then
. From here the results follow from the fact that
.
Next, we turn to the third part. Without loss of generality, take
. Assume that there are
with
. Let
be a sequence with
and
be a sequence with
. We have
It follows that
V has a discontinuity at
s. From here, the result follows from the fact that a decreasing function can have at most a countable number of discontinuities. The remaining parts are a special case of Lemma 5 given in
Section 6. □
For any
and
, if the set
has Lebesgue measure zero, then Lemma 2 implies that
where we use the fact that the number of terms in the sum is countable. Further, Lemma 2 tells us that there are at most a countable number of
s for which
has positive Lebesgue measure. This number reduces to 0 if
V is invertible. Note that (
5) justifies saying that
is obtained from
by removing the large jumps. Next, applying Lemma 2 gives
Lemma 3. For any fixed , is a Lévy process with .
Proof. Applying (
4) to
instead of
V shows that for fixed
,
is a Lévy process with
. It is readily checked that for fixed
s,
is stochastically continuous with càdlàg paths. Hence, the result follows by Lemma 1 and (
6). □
4. Classes of Subordinators
One of the best known classes of subordinators is the class of stable subordinators. Recall that a distribution
is said to be stable if for any
n, there exist
and
such that if
, then
In this sense, these distributions are stable under addition. It can be shown that
for some
and the corresponding distribution is said to be
-stable. The class of 2-stable distributions corresponds to the class of normal distributions. It is well-known that the Lévy processes associated with stable distributions are the only ones that are self-similar. Specifically,
is a Lévy process such that for any
, there exists a
and a function
with
if and only if
has a stable distribution. Further, in this case,
and
for some constant
, see Section 13 in [
13]. For more on stable distributions, see [
14] or [
15].
Only stable distributions with
can be subordinators. The Lévy measure of an
-stable subordinator is of the form
where
. In this case, for
, we have
The remainder process is given by
It is sometimes convenient to restrict
s to the set
for some
. In this case, Lemma 2 gives
where in the case
we take
.
When
, (
7) gives the remainder process with Lévy measure
Such distributions are called truncated stable subordinators, see [
16]. They can be seen as variants of tempered stable distributions, and in [
17] they were called tempered stable distributions with hard truncation. See [
8,
9] and the references therein for more on tempered stable distributions. Note that the remainder process for a truncated stable subordinator is just the remainder process for a stable subordinator with the parameter
s restricted to a bounded interval.
Lévy measures of the form (
8) can be extended to Lévy measures for any
and are subordinators so long as
. When
and
, we have
where
denotes the positive part, i.e.,
for
. The remainder process is given by
When
, for
,
and the remainder process is given by
where for
, we take
.
Truncated stable subordinators are no longer stable under addition. However, they satisfy the following property. For any
, if
, then
where
. This is readily checked by comparing the characteristic functions of the random variables on the two sides.
When
, the class of truncated stable subordinators coincides with the class of generalized
h-Dickman distributions. Recall that a random variable
X is said to have a generalized
h-Dickman distribution if
where
and
is independent of
X on the right side. We denote this distribution by
and note that it has Lévy measure
. When
, this is called a generalized Dickman distribution, and when
, it is just called the Dickman distribution. Many properties of these distributions are discussed in [
18,
19,
20,
21,
22,
23] and the references therein. It is easily shown that, for any
, if
, then
. The idea of the Dickman distribution being a truncated 0-stable subordinator was first published in [
11], see also [
12].
5. Main Results
In this section, we consider a process indexed by three parameters,
s,
t, and
, which combines all of the processes described in
Section 4 into one. We will derive various properties of this process, which will show how the processes in
Section 4 relate to each other. Recall that we need
to ensure that the measure in (
8) is a Lévy measure with finite variation. For
,
, and
, let
With this notation, all of the remainder processes discussed in
Section 4 can be written as
where
. When
, we can allow for
so long as we take
in this case. We now collect some basic properties of the function
.
Lemma 4. 1. For fixed and , the function is continuous.
2. We have for each choice of the parameters.
3. For any fixed and , the function is decreasing.
4. For any , , and , we have .
5. For and fixed, the function is increasing.
It can be shown that for , is neither increasing for every x nor decreasing for every x.
Proof. The first part is immediate for
and can be easily verified for
using L’Hôpital’s rule. The second and third parts are immediate. The fourth part follows directly from the third. We now turn to the fifth part. Assume either
or
and
. By Part 2, it suffices to show that
is increasing in
. We have
which is non-negative by 4.1.33 in [
24] and the fact that
. Next, by L’Hôpital’s rule,
where we note that for any
x and
close enough to 0, we have
. □
Lemma 4 shows that is continuous in . However, cannot be continuous in for all values of s. This is because the indicator function depends on both and s. Nevertheless, we will show that there is continuity in so long as s is bounded away from for each . Toward this end, let be such that , where is the closure of . Note that is a random set. Clearly, we can construct such a set, where its complement has an arbitrarily small (but positive) Lebesgue measure.
Theorem 1. Fix and . We haveand for any finite set , we have Proof. Fix
and, for simplicity, let
. Note that
and, thus, that there exists an
with
. For
close enough to
so that
and any
, we have
Combining this with Lemma 4 gives
Hence,
From here, the result follows by dominated convergence. To obtain a dominating function, we note that for any
and
close enough to
,
where we use Lemma 4. This is summable w.p. 1 by the shot-noise representation of the appropriate distribution. The second part follows by combining the fact that
with the first part. □
Next, we verify the self-similarity of the remainder process.
Theorem 2. If and , then for any , we haveFor , the result also holds when . Proof. For
, the fifth part of Lemma 2 implies that
For
, we have
where the equality in fdd follows from the fourth part of Lemma 2 with
. Here, we use the facts that for
, we have
⊂
and for
, we have
⊂
for
.
For
, Lemma 2 gives
where
. We note that for
, we have
and for
, we have
⊂
. □
Theorem 3. Let be a Borel measure on such that is a Lévy measure with finite variation for each . Let be a remainder process with Lévy measure . If there are functions such that for each and each ,then there exists an such that X is of the form (9). Proof. Fix
and note that
for large enough
h. For any
,
is the Lévy measure of
. Since
, we have
Now, let
be bounded away from infinity. There exists an
such that
for every
. Thus,
and
Hence,
L satisfies (
11) and the result follows by Lemma 6. □
We now give some implications of our results for Lévy processes. First, we recall a definition. Let
and
be an
-valued stochastic processes, each with index set
I. We say that
X is smaller than
Y in the usual stochastic order and write
if for any positive integer
n and any
, we have
for every increasing Borel function
for which both expectations exist. Here,
being increasing means that
whenever
for each
. For a monographic treatment of stochastic order, see [
25].
Theorem 4. For any and , set - 1.
If , then for any fixed α, is a increasing Lévy process and for any fixed , is a continuous and increasing process.
- 2.
For , we have
andIf, in addition, , then Proof. The first part follows from Theorem 1 with
and Lemma 4. The second part follows by combining Lemma 4 with (
10) and Lemma 2. □
6. Lemmas
In this section, we give two lemmas that are used in our proofs. We present these in the multivariate setting. For the first lemma, this is needed because it will be applied in the context of equality in fdd. For the second lemma, it is done in the interest of generality as the proof is no more difficult in the multivariate case, and the result may be of independent interest. We now introduce our notation for working in the multivariate setting. Let be the space of d-dimensional column vectors equipped with the usual norm . Let be the unit sphere in . We write and to denote the Borel sets in and , respectively. For and , we write . For , we write to denote component-wise multiplication.
Lemma 5. Let , let be iid -valued random vectors independent of the ’s, and set . Let I be a non-empty index set and for each , let be a Borel function. Assume that for each , - 1.
- 2.
If are independent of the ’s and the ’s, then for any and any deterministic sequence , we have
Proof. Let be an arbitrary finite collection of distinct elements in I. Consider the function , whose output is the outputs of stacked on top of each other. Similarly, we write to denote the random vector in , where the first m elements of the vector are , the next m elements are , and so forth.
We begin with the first part. Let
. For any integer
, we have
where
,
for
and
. By the memoryless property of the exponential distribution,
. Thus,
and
where the last equality follows by relabeling the
’s.
We now turn to the second part. Let
and, for integer
, recursively define
. For
, let
and note that these are iid random variables each having a geometric distribution with parameter
b. Let
Here, we use the well-known and easily verified fact that the distribution of a geometric sum of iid exponential random variables is exponential, see e.g., Proposition 2.3 in [
26]. Since
are iid random variables with
, it follows that
Next, note that the conditional distribution of
given
is the same as the distribution of
. We have
where the first equality in distribution is obtained by removing all terms in the sum on the left that have
, relabeling the
’s and
’s, and replacing the
’s with random variables that have the appropriate conditional distribution. □
Lemma 6. Let be a Borel measure on with and assume that, for every , we have If there exists a function ψ such that for every and every that is bounded away from 0 and infinity, we havethen there exists an and a finite Borel measure on such that andIn this case, D is a Lévy measure if and only if . Further, if for some , , then is a Lévy measure if and only if and it has finite variation if and only if . Proof. First, we show that (
11) holds for every
. Fix
and note that we can write
where the
’s are mutually disjoint and each
is bounded away from 0 and infinity. By countable additivity, we have
Next, fix
so that
B is bounded away from 0 and
∞ and satisfies
. Let
, let
with
, and note that
and
. By (
11) and the continuity of measures (see e.g., Problem 10.4 in [
27]), we have
which shows that
is continuous. Next, for any
, we have
and
, which implies that
. Now, let
for
and note that
for
. It follows (see e.g., Appendix A20 in [
27]) that
for some
. Hence,
,
, for some
.
For
, countable additivity implies
Let
for
and note that for any
,
Since the collection of sets of the form
for
and
is a
-system that generates the Borel sets, (
12) follows by Theorem 10.3 in [
27].
Next, for
, note that
Let
for
. For any
,
From here, (
12) follows as in the previous case.
Finally, we turn to the case
. Let
for
. Fix
. For any positive integers
m and
n, (
11) implies that
Applying (
11) again gives
By the continuity of measures (see Theorem 10.2 in [
27]) and the fact that rational numbers are dense in
, it follows that for
, we have
Equivalently, for
, we have
Noting that for
,
allows us to conclude that, for any
,
From here, the results follow as in the previous cases.
The results for when D and are Lévy measures or have finite variation are easily verified and, hence, the proof is omitted. □