1. Introduction
According to the generalized Rényi theorem, a geometric random sum of independent identically distributed (i.i.d.) nonnegative random variables (r.v.’s), normalized by its mean, converges in distribution to the exponential law when the expectation of the geometric number of summands tends to infinity. Some numerical bounds for the exponential approximation to geometric random sums, as well as their various applications, can be found in the classical monograph of Kalashnikov [
1]. Peköz and Röllin [
2] developed Stein’s method for the exponential distribution and obtained moment-type estimates for the exponential approximation to geometric and non-geometric random sums with non-negative summands completing Kalashnikov’s bounds in the Kantorovich distance. Their method was substantially based on the equilibrium transform (stationary renewal distribution) of non-negative random variables, hence yielding the technical restriction on the support of the random summands under consideration. Moreover, Peköz and Röllin considered dependent random summands with constant conditional expectations and presented some error-bounds in this case. The present authors extended Stein’s method to alternating (i.e., taking values of both signs) random summands by generalizing the equilibrium transform to distributions with arbitrary support, and obtained moment-type estimates of the accuracy of the exponential approximation for geometric and non-geometric random sums of independent alternating random variables. The same paper [
3] contains a detailed overview of the estimates of the exponential approximation to geometric random sums.
The aim of the present work is to extend the results of [
3] to dependent random summands with constant conditional expectations, also generalizing the results of [
2] to alternating random summands.
Recall that the Kantorovich distance
between probability distributions of r.v.’s
X and
Y with distribution functions (d.f.’s)
F and
G is defined as a simple probability metric with
-structure (see [
1,
4]) as
where
and
If both
X and
Y are integrable, then
and the supremum in (
1) can be taken over a wider class
of Lipschitz functions. In this case, according to the Kantorovich–Rubinshtein theorem,
allows several alternative representations
where
and
are generalized inverse functions of
F and
G, respectively.
We will use a
generalized equilibrium transform that was introduced and studied in [
3]. Given a probability distribution of a r.v.
X with d.f.
F and finite
, its equilibrium transform is defined as a (signed) measure
on
with the d.f.
Observe that
is absolutely continuous (a.c.) with respect to (w.r.t.) the Lebesgue measure with the density
The characteristic function (ch.f.) of
can be expressed in terms of the original ch.f.
f of r.v.
X as
If X is nonnegative or nonpositive almost surely (a.s.), then is a probability distribution and it is possible to construct a r.v. .
Below, we list some other properties of the equilibrium transform (see ([
3], Theorem 1) for details and proofs) which will be used in the present work:
Homogeneity. For any r.v.
X with finite
and d.f.
we have
Moments. If
for some
, then for all
we have
We will also use the following inequality from ([
3], Theorem 3), which states that the Kantorovich distance to the exponential law is no more than twice greater than distance to the equilibrium transform.
Lemma 1. Let X be a square integrable r.v. with and . Then,where is the equilibrium transform of . The r.-h.s.’s of (
8), in turn, can be bounded from above by the second moment
in the following way.
Lemma 2 (see Theorem 2 and Remark 2 in [
3])
. For any square-integrable r.v. X with , Note the presence of the Kantorovich distance between
and possibly signed measure
on the r.-h.s.’s of (
8) and (
9), which requires some extra explanation. As described in [
3], it is defined in terms of d.f.’s in the same way as for probability measures in (
1) and allows an alternative representation as an area between d.f.’s of its arguments (similar to the last expression in (
2)). Moreover, this generalization retains the property of the homogeneity of order 1 (see ([
3], Lemma 1)). Namely, if
F and
G are d.f.’s of (signed) Borel measures on
with
and
,
,
,
, then
Using the above notation and techniques, we prove moment-type error bounds in the Kantorovich distance for the exponential approximation to random sums of possibly dependent r.v.’s with positive finite expectations (Theorem 1), which generalize the results of [
2] to alternating random summands and results of [
3] to dependent random summands.
Moreover, we extend the definitions of
new better than used in expectation (NBUE) and
new worse than used in expectation (NWUE) distributions to alternating random variables in terms of the corresponding d.f.’s and provide a criteria in terms of conditional expectations similar to the classical one (Theorem 2). Finally, we provide simplified error-bounds in cases of NBUE/NWUE conditional distributions of random summands, generalizing those obtained in [
2].
2. Main Results
Lemma 3. Let be a sequence of random variables, such that for every there exists a regular conditional probability with the constant conditional expectation . Let for , and N be a -valued r.v., independent of , with Then the characteristic function of iswhere M is an -valued r.v. withand is the characteristic function of the equilibrium transform of the conditional distribution . Or, in terms of (conditional) distribution functions,where denotes the joint d.f. of and denotes the conditional d.f. of , . Here, and in what follows, we assume that and for denote unconditional ch.f. and d.f. of . A similar notation will be used for other characteristics of distributions. Remark 1. If are independent, then (11)–(12) reduces to the single summand property of the equilibrium transform (see (Equation (30), [3]). Remark 2. If all a.e. and M is independent of , then (11)–(12) can be expressed in terms of random variables aswhere the sequence is independent of M and the conditional distribution of given coincides with . Proof. According to ([
3], Lemma 2), for every
and
we have
where
. By applying (
5) twice, we obtain for
Theorem 1. Let be a sequence of random variables, such that for every there exists a regular conditional probability with the constant conditional expectation . Let for , and N be a -valued r.v., independent of , with Let , , and M be a -valued r.v. with Then, for any joint distribution of N and M we havewhere the first term vanishes in case of andand both notations , stand for the short forms of , , respectively. Proof. By Lemma 1 and homogeneity of both the Kantorovich distance and the equilibrium transform (see (
6) and (
10)), we have
Let us bound from above.
For a given joint distribution
, let
,
,
. Denoting
for
, designating
and
the short forms of the conditional d.f.’s of
and
, respectively,
, and using Lemma 3 together with the representation of the Kantorovich distance between (signed) measures as an area between their distribution functions, we obtain
where
For the summands with
by Tonelli’s theorem we have
where
stands for the conditional joint d.f. of
given that
. By adding and subtracting
under the modulus sign and using further the triangle inequality, we obtain
where
is the Dirac measure concentrated in zero.
For the case of
by Tonelli’s theorem, we have
where
stands for the conditional d.f.
.
By adding and subtracting
in the integrand under the modulus sign and using further the triangle inequality, we obtain
Combining both
and
cases and using the fact that
, we get
where the first sum can be bounded from above by
Substituting the latter bound into (
14) yields (
13). If
, then we take a comonotonic pair
as
, which eliminates the first term on the r.-h.s. of (
15). □
Remark 3. Theorem 1 reduces to ([2], Theorem 3.1) in case of nonnegative and to ([3], Theorem 6) in case of independent . If both expectations
and
are finite, then
in (
13) can be replaced with
due to the dual representation of the
-metric as
Moreover, if
N and
M are stochastically ordered (that is,
for all
or vice versa), then
If, in addition, all
, then
and the first term on the r.-h.s of (
13) can be bounded from above as
Hence, we arrive at the following.
Corollary 1. Let, in addition to the conditions of Theorem 1, for all and the r.v.’s N and M be stochastically ordered with finite expectations. Then Remark 4. If , , that is , , then In this case, for every with we have Therefore, by the Cauchy–Bunyakovsky–Schwarz inequality, we have Thus, the first term on the r.-h.s of (13) can be bounded from above as This means that in case of and , the first term on the r.-h.s. of (13) is, at most, of order as . If , and for all , then , and thus, . Therefore, if , then the first term on the r.-h.s. of (13) vanishes. If , and for all , then as well, and thus, .
Next, let us simplify the second term
D in (
13).
Corollary 2. Let, in addition to the conditions of Theorem 1, for every and the r.v. M be independent of . Then Proof. By Lemma 2, we have
which proves the statement of the corollary. □
Recall that a nonnegative r.v.
X with finite
is said to be new better than used in expectation (NBUE), if
and new worse than used in expectation (NWUE), if
Using Tonelli’s theorem, it can be ascertained that X is NBUE if and only if X stochastically dominates its equilibrium transform , that is, for all . Similarly, X is NWUE if and only if stochastically dominates X. We will show that the same results hold true if we extend both NBUE and NWUE notions to the case of r.v.s without support constraints.
Definition 1. We say that a (possibly alternating) r.v. X with d.f. F and is NBUE, if for all , where is the equilibrium transform w.r.t. F. Similarly, we say that the r.v. X with d.f. F and is NWUE (new worse than used in expectation), if for all .
Theorem 2. A r.v. X with finite is NBUE if and only if Moreover, (16) implies that a.s. The r.v. X with finite is NWUE if and only if Proof. By Tonelli’s theorem, for every
, we have
Note that the same chain of equalities holds true with the event
in place of
. If
, then
and (
18) turns into
Let
X be NBUE, i.e.,
for all
. This implies that
due to the absolute continuity of
, and hence, with the account of (
19), we obtain (
16).
Conversely, let (
16) hold true. For
, we have
which is possible if and only if
, i.e.,
a.s. Hence,
for
. For positive
t, inequality (
16) together with Equation (
19) yields
. Therefore,
X is NBUE.
Let
X be NWUE, i.e.,
for all
. This yields
, and hence, with the account of (
19), we obtain (
17).
Conversely, let (
17) hold true. For
we have
, since
has negative density on the negative half-line. Finally, (
17) and (
19) yield
for positive
t. □
If
X is NBUE or NWUE with
and
, then
where the last equality follows from (
7). Hence, if for all
and
the conditional distribution
is NBUE or NWUE, then the second term on the r.-h.s. of (
13) takes the form
If
M is independent of
, then the latter expression can be bounded from above with the help of the conditional Jensen’s inequality
where we used the notation
as before. If all the r.v.’s
are independent, then (
20) may be simplified as
Hence, we get the following
Corollary 3. Let, in addition to the conditions of Theorem 1, for every , the conditional distributions be NBUE or NWUE for all and , and the r.v. M be independent of . Then Moreover, if all the r.v.’s are independent, then Corollary 3 reduces to ([
2], Corollary 3.1) in the case of all
being independent, all
and the r.v.s.
N,
M being stochastically ordered (cf. Corollary 1).