3.1. Higher-Order EP Scale Mixture Representations and Related Topics
Proposition 1. Let , . Thenwhere is an r.v. such that: if , then for any and if , then is absolutely continuous with probability density Proof. From (2), it follows that the symmetric (
) strictly stable distribution has the characteristic function
Rewrite (3) with
in terms of characteristic functions:
Then, changing the notation
, by formal transformations of Equality (8), we obtain
It can be easily verified that
is the probability density of a nonnegative r.v. Indeed, since
is a probability density, for any
we have
Therefore, it follows from (9) that
The proposition is thus proved. □
Let
. Then the assertion of Proposition 1 can be rewritten as
It is easily seen that . Setting , from (10) we obtain
Corollary 1. Any symmetric EP distribution with is a scale mixture of normal laws [
16]:
Now let
. The r.v. having the Laplace distribution with variance 2 will be denoted
. As it has already been noted,
. It is well known that
On the other hand, from Corollary 1 it follows that
Therefore, by virtue of identifiability of scale mixtures of normal laws (see [
25] and details below), having compared (
11) and (
12), we obtain
that is, the r.v.
has the exponential distribution with parameter
whereas the r.v.
has the inverse exponential (Fréchet) distribution,
,
.
Corollary 2. Any symmetric EP distribution with is a scale mixture of Laplace laws: For
, from Corollary 2 we obtain one more representation of the EP distribution as a scale mixture of normal laws:
Corollary 3. If , then is infinitely divisible.
Proof. By virtue of identifiability of scale mixtures of normal laws, from (13) and Corollary 1 we obtain that if
, then the distribution of the r.v.
is mixed exponential:
Hence, in accordance with the result of [
26] which states that the product of two independent non-negative r.v.s is infinitely divisible, provided one of the two is exponentially distributed, from (14) it follows that, with
, the distribution of
is infinitely divisible. It remains to use Corollary 1 and a well-known result that a normal scale mixture is infinitely divisible, if the mixing distribution is infinitely divisible (see, e.g., [
27], Chapter XVII,
Section 3). □
The interval
does not cover all values of
providing the infinite divisibility of
. Another obvious value of
for which
is infinitely divisible is
: the distribution of
is normal and hence, infinitely divisible as well. Moreover, as is shown in [
14], with values of
, the EP distributions are not infinitely divisible.
From Proposition 1, as a by-product, we can obtain simple expressions for the moments of negative orders of one-sided strictly stable distributions.
Corollary 4. If and , then Proof. As it was made sure in the proof of Proposition 1, the function
is a probability density, that is,
Now consider some properties of the mixing r.v. in (7). First, we present some inequalities for the tail of the distribution of .
Proposition 2. For any and , we haveas . Let , . Then for any For any , and , we have Proof. With the account of the well-known relation
(e.g., see [
18], Chapter 7, Section 36) we conclude that for any
, there exists a
such that
Therefore, as .
With the account of (6), we have
Proposition 3. Let , . The moments of the r.v. of orders are infinite, whereas for , we have Let . Then, Proof. To prove (15), notice that, by the definition of
,
and use (6).
To prove (16), note that for arbitrary
and
Letting , from (17) we obtain (16). The proposition is proved. □
Now consider the property of identifiability of scale mixtures of EP distributions. Recall the definition of identifiability of scale mixtures. Let Q be an r.v. with the d.f. , and be two nonnegative r.v.s. The family of scale mixtures of is said to be identifiable, if the equality implies .
Lemma 1. Let be a d.f. such that . The family of scale mixtures of is identifiable, if the Fourier–Stieltjes transform of the d.f. is not identically zero in some nondegenerate real interval [
25].
Proposition 4. For any fixed , the family of scale mixtures of EP distributions is identifiable; that is, if and are two nonnegative r.v.s, then the equality implies .
Proof. First assume that
and prove that
. For this purpose use Lemma 1. Denote
,
,
. We obviously have
Therefore, by the chain differentiation rule we have
Hence, the Fourier–Stieltjes transform
of the d.f.
is
Multiplying the integrand in the last integral by
and changing the variables
so that
and
, we obtain
The reference to Lemma 1 proves that . Now assume that . Then, obviously, and the desired result follows from what has just been proved. □
Proposition 5. Let . Then, Proof. From Proposition 1 (see (10)) we have
and
Now the desired result follows from Proposition 4 which states that the family of scale mixtures of EP distributions is identifiable; that is, if and for some r.v.s and , then . In the case under consideration , . □
Proposition 5 relates the distributions of the r.v.s
with different values of
but with the same values of
. As regards the relation between the distributions of the r.v.s
with different values of
but with the same values of
, it can be easily seen that for any
and
In other words, for any
Consider some properties of the
one-sided EP distribution of the r.v.
. Obviously, the density
of
is given by (18), so that for
Lemma 2. A d.f. such that corresponds to a mixed exponential distribution, if and only if its complement is completely monotone: and for all .
Proof. This statement immediately follows from the Bernstein theorem [
28]. □
Proposition 6. The distribution of the r.v. can be represented as mixed exponential if and only if . In that case the mixing density is .
Proof. Let
. As is known, the Laplace–Stieltjes transform
of the nonnegative strictly stable r.v.
is
Hence, by formal transformation, we obtain
where the function
was introduced in Proposition 1 and proved to be a probability density function. Relation (20) means that if
, then the distribution of
is mixed exponential.
Now, let
. We have
and
It can be easily seen that for we have while for we have with strict inequalities for . Hence, by Lemma 2 the distribution of is not mixed exponential. The proposition is proved. □
In terms of r.v.s the statement of Proposition 6 can be formulated as
provided
(also see Corollary 1).
Corollary 5. Let . Then the d.f. is infinitely divisible.
Proof. This statement immediately follows from (21) and the result of [
26] which states that the product of two independent non-negative r.v.s is infinitely divisible, provided one of the two is exponentially distributed. □
3.2. Convergence of the Distributions of Maximum and Minimum Random Sums to One-Sided EP Laws
From Corollary 1 and (13), we obtain
Corollary 6. is a scale mixture of folded normal distributions: if , thenmoreover, if , then . In this section we demonstrate that the one-sided EP distribution can be limiting for maximum sums of a random number of independent r.v.s (maximum random sums), minimum random sums and absolute values of random sums. Convergence in distribution will be denoted by the symbol ⟹.
Consider independent not necessarily identically distributed r.v.s
with
and
,
. For
denote
,
,
,
. Assume that the r.v.s
satisfy the Lindeberg condition: for any
It is well known that under these assumptions (this is the classical Lindeberg central limit theorem) and , , and , , (this is one of manifestations of the invariance principle).
Let be a sequence of nonnegative r.v.s such that for each the r.v.s are independent. For let , , (for definiteness assume that ). Let be an infinitely increasing sequence of positive numbers. Here and in what follows convergence is meant as .
Lemma 3 Assume that the r.v.s and satisfy the conditions specified above. In particular, let Lindeberg condition (22) hold. Moreover, let in probability. Then the distributions of normalized random sums weakly converge to some distribution; that is, there exists an r.v. Y such that , if and only if any of the following conditions holds: - (i)
;
- (ii)
there exists an r.v. such that ;
- (iii)
there exists an r.v. such that ;
- (iv)
there exists a nonnegative r.v. U such that .
Moreover, , ; , ; , .
The
proof of Lemma 3 was given in [
29].
Lemma 3 and Corollary 6 imply the following statement.
Proposition 7. Let . Assume that the r.v.s and satisfy the conditions specified above. In particular, let the Lindeberg condition (22) hold. Moreover, let in probability. Then the following five statements are equivalent: 3.3. Extensions of Gleser’s Theorem for Gamma Distributions
In [
30], it was shown that a gamma distribution can be represented as mixed exponential if and only if its shape parameter is no greater than one. Namely, the density
of a gamma distribution with
can be represented as
where
In [
31], it was proved that if
,
and
and
are independent gamma-distributed r.v.s, then the density
defined by (
23) corresponds to the r.v.
where
is the r.v. with the Snedecor–Fisher distribution corresponding to the probability density
In other words, if
, then
A natural question arises: is there a product representation of
in terms of exponential r.v.s for
similar to (
26)? The results of the preceding section can give an answer to this question.
For simplicity, without loss of generality let .
Proof. As it has been already mentioned,
Therefore, with the account of (21), we obtain the desired result. □
Gamma distributions, as well as one-sided EP distributions, are particular representatives of the class of generalized gamma distributions (GG distributions), that was first described (under another name) in [
32,
33] in relation with some hydrological problems. The term “generalized gamma distribution” was proposed in [
34] by E. W. Stacy who considered a special family of lifetime distributions containing both gamma and Weibull distributions. However, these distributions are particular cases of a more general family introduced by L. Amoroso [
35]. A generalized gamma distribution is the absolutely continuous distribution defined by the density
with
,
,
. An r.v. with the density
will be denoted
. It is easy to see that
The following statement can be regarded as a generalization of (
27).
Proposition 9. Let , . Then, Proof. From Proposition 1 it follows that if
and
, then
Now let
,
and use (
28) to obtain the desired result. □
3.4. Alternative Mixture Representations
Let . By we will denote an r.v. with the uniform distribution on the segment .
Note that Lemma 4 with
yields an ‘unexpected’ uniform mixture representation for the normal distribution:
The following statement extends and generalizes a result of [
36] (see Lemma 4).
Proposition 10. For any , the EP distribution can be represented as a scale mixture of uniform distributions: the case is covered by Lemma 4 and if , thenfor any . Proof. Let
. From (10) with the positions of
and
switched and Lemma 4 we have
Now it remains to use (
29). □
Setting , we obtain
Now turn to other mixture representations. If in (
28)
, then
. From this fact and Gleser’s result (
26), we obtain the following statement.
Proposition 11. If , then the one-sided EP distribution is a scale mixture of Weibull distributions: Let
be an r.v. such that
. For
define the r.v.
as the symmetrized r.v.
:
With the account of (
25), it is easy to make sure that the probability density
of the r.v.
has the form
It is worth noting that the probability density of the r.v. is , . Then, from Proposition 11 we obtain one more mixture representation for the EP distribution with via the Weibull distribution.
As by-products of Proposition 10 and Corollary 7, consider some mixture representations for the exponential and normal distributions. Using Corollary 1 we obtain for
that
Here we use the notation
for the r.v. having the chi-squared distribution with
m degrees of freedom. Setting
in (
36), we obtain the following representation for the exponentially distributed r.v.:
Now on the left-hand side of (
37) use the easily verified relation
and on the right-hand side of (
37) use relation
(see (
26)). Then (
37) will be transformed into
and since the family of mixed exponential distributions is identifiable, this yields the following mixture representation for the folded normal distribution:
Along with (
32), from (
38) we obtain one more product representation for the normal r.v., this time in terms of the ‘scaling’ r.v.s in (
26) and Corollary 1:
Since the r.v.
has the discrete uniform distribution on the set
, relation (
39) can be regarded as one more uniform mixture representation for the normal distribution.
3.5. Some Limit Theorems for Extreme Order Statistics in Samples with Random Sizes
Proposition 11 declares that the one-sided EP distribution with
is a scale mixture of Weibull distribution with shape parameter
. In other words, relation (
35) can be expressed in the following form: for any
At the same time, Proposition 8 means that any gamma distribution with shape parameter
can also be represented as a scale mixture of the Weibull distribution with the same shape parameter. In other words, relation (
27) can be expressed in the following form: for any
From (
40) and (
41), it follows that one-sided EP distribution with
and the gamma distribution with
can appear as a limit distribution in limit theorems for extreme order statistics constructed from samples with random sizes. To illustrate this, we will consider the limit setting dealing with the min-compound doubly stochastic Poisson processes.
A doubly stochastic Poisson process (also called Cox process) is defined in the following way. A stochastic point process is called a doubly stochastic Poisson process, if it has the form , where , , is a time-homogeneous Poisson process with intensity equal to one and the stochastic process , , is independent of and has the following properties: , for any , the trajectories of are right-continuous and do not decrease. In this context, the Cox process is said to be lead or controlled by the process .
Now let
,
, be the a doubly stochastic Poisson process (Cox process) controlled by the process
. Let
be the points of jumps of the process
. Consider a marked Cox point process
, where
are independent identically distributed (i.i.d.) r.v.s assumed to be independent of the process
. Most studies related to the point process
deal with
compound Cox process which is a function of the marked Cox point process
defined as the sum of all marks of the points of the marked Cox point process which do not exceed the time
t,
. In
, the operation of summation is used for compounding. Another function of the marked Cox point process
that is of no less importance is the so-called max-compound Cox process which differs from
by that compounding operation is the operation of taking maximum of the marking r.v.s. The analytic and asymptotic properties of max-compound Cox processes were considered in [
37,
38]. Here we will consider the min-compound Cox process.
Let
be a Cox process. The process
defined as
, is called a
min-compound Cox process.
We will also use the conventional notation .
Lemma 5. Assume that there exist a positive infinitely increasing function and a positive r.v. L such thatas . Additionally assume that and the d.f. satisfies the condition: there exists a number such that for any Then there exist functions and such thatas , wherefor and for . Moreover, the functions and can be defined as Proof. This lemma can be proved in the same way as Theorem 2 in [
37] dealing with max-compound Cox processes using the fact that
. □
Proposition 12. Let . Assume that there exists a positive infinitely increasing function such thatas . Additionally assume that and the d.f. satisfies condition (42) with . Then there exist functions and such thatas . Moreover, the functions and can be defined by (43). Proof. This statement directly follows from Lemma 5 with the account of (
40). □
Proposition 13. Let . Assume that there exists a positive infinitely increasing function such thatas . In addition, assume that and the d.f. satisfies condition (42) with . Then there exist functions and such thatas . Moreover, the functions and can be defined by (43). Proof. This statement directly follows from Lemma 5 with the account of (
41). □
Propositions 11 and 12 describe the conditions for the convergence of the distributions of extreme order statistics to one-sided EP distributions with
and to gamma distributions with
, respectively. Using (21) and (
26) instead of (
40) and (
41) correspondingly, we can also cover the cases
and
.
Proposition 14. Let . Assume that there exists a positive infinitely increasing function such thatas . In addition, assume that and the d.f. satisfies condition (42) with . Then there exist functions and such that (44) holds as . Moreover, the functions and can be defined by (43). Proof. This statement directly follows from Lemma 5 with the account of (21). □
Proposition 15. Let . Assume that there exists a positive infinitely increasing function such thatas . In addition, assume that and the d.f. satisfies condition (42) with . Then there exist functions and such that (45) holds as . Moreover, the functions and can be defined by (43). Proof. This statement directly follows from Lemma 5 with the account of (
26). □
It is very simple to give examples of processes satisfying the conditions described in Propositions 11–14. Let and , , where U is a positive r.v. Then choosing an appropriately distributed U we can provide the validity of the corresponding condition for the convergence of . Moreover, the parameter t may not have the meaning of physical time. For example, it may be some location parameter of , so that the statements of this section concern the case of large mean intensity of the Cox process.