1. Introduction
The Markov branching process is a Markov process with non-negative integer states which models the temporal progression of the size of a population of individuals whose independent lifetimes have an exponential law with the parameter (split rate) . At the end of their lives, each individual begets j offspring with probability , and all reproduction events are mutually independent and independent of life lengths. The probability-generating function (pgf) of this offspring number law is , and we assume that .
Sometimes, it is assumed that
on the basis that begetting one offspring is unobservable, e.g., p. 118 in [
1]. This assumption entails no loss of generality but, given our aims, we do not adopt it. We do assume that
. The transition probabilities
, where
and
signal that
, are specified by the single pgf
together with
expressing the independent evolution of lines of descent.
The zero state is accessible and absorbing, and hence, there is no limiting distribution. The transition semigroup does, however, have a stationary measure
in the sense that, for
,
and this measure is unique up to multiplication by a constant; see p. 110 in [
2]. We write
. As shall be explained in §2, we lose no generality by assuming that the mean per capita of offspring
;
is subcritical (
) or it is critical (
). In this case, the generating function (gf) of the stationary measure is
where
is an arbitrary constant. In particular,
. Recalling that
if
, we see that
.
In the case
, we choose
and we write
to mean
f is a subcritical offspring-number pgf. The stationary measure is closely associated with discrete limiting conditional laws (LCLs) of the MBP. Let
denote the time to extinction, or hitting time of the zero state. Then,
, where
and
, i.e., extinction is (almost) certain. If
, then the limits
exist, they are independent of the initial population size
i,
, and their pgf is
The distribution (
2) is called a limiting conditional law (LCL) and we write
to mean that
is the pgf of the LCL of some subcritical MBP, or that
is an LCL, according to the context.
Until recently, very few explicit examples of LCLs were known. Theorem 2.1 in [
3] (see Theorem 9 below) collects sufficient conditions (which are necessary) for a probability mass function
satisfying
(written
), ensuring that
. Several explicit examples have thereby been constructed. One motivation for this stems from the fact that
satisfies the identity
where
. So, if
can explicitly be inverted, there results an explicit expression for
. Several such examples are given in [
3].
Not known to the authors at the time [
3] was written is the fact that the earlier paper [
4] presents a characterisation of LCLs in the form of a sequence of inequalities. This criterion seems less suited for constructing examples, though two are given. As an aside, the representations in [
4] imply that
.
If
, the critical case, then we set
in (
1) and denote with
a Borel subset of the positive reals with finite Lebesgue measure
. Theorem 7.1 in [
5] asserts the existence of
where
is a certain sequence such that
is geometrically fast as
and
; see (
50) below. Hence, the stationary measure looms large here too, and, in addition,
Hence, explicit expressions for
and its inverse will yield explicit forms of
.
The structure of this paper is as follows. In §2, we recall how the supercritical case can be reduced to the subcritical case. We then present characterisations of the stationary measure along the lines of Theorem 1 in [
4] and of Theorem 2.1 in [
3]. In §3, we begin by recalling a simple way of constructing laws concentrated on the positive integers which strictly contain the LCLs of subcritical MBPs; see Theorem 3. Thus, Theorem 3, in essence, is Theorem 1.1 in ([
6]) but stated in a stronger form: If
is a non-negative sequence satisfying (
13) below, then
defined by (
3) is an honest pgf if and only if (
14) holds. In particular, if
, then
is a pgf. Theorem 4 gives estimates for the right-hand tails generated by
.
The pgf
is the composition
, where
is the pgf of the so-called Sibuya law [
7]. This law is a shifted Poisson mixture whose infinitely divisible (infdiv) mixing law is known to be a generalised gamma convolution (see §3 for definitions) and, hence, is self-decomposable. This law also belongs to a larger family of infdiv laws whose Laplace exponents are complete-Bernstein functions. On the other hand, it has not been known (until now) whether they are Thorin–Bernstein. The remainder of §3 is a rather long detour (Theorems 5–8) which completely resolves this issue in the affirmative.
We remark that in this portion of the text only, the letter a is used to denote one parameter which indexes the generalised Mittag–Leffler function.
The purpose of §4 is to supplement the conditions of Theorem 3 with additional structure, ensuring that
. The discussion begins with general comments recalling the sense in which
generates a limiting conditional law if
. Then, Theorem 9 recalls Theorem 2.1 from ([
3]) stated here in a stronger form. It gives workable criteria ensuring that
. Examples 2 and 3 complete two examples in [
3] by exhibiting explicit expressions for the stationary measure. Theorem 10 expresses Theorem 9 in terms of sequences
. It is illustrated by an Example 5 which extends the treatment of the geometric offspring law in [
8].
As motivation for §§5,6, let
denote a critical Galton–Watson process with time to extinction
. With no further moment conditions, it is shown in [
9] (Theorem 1) that, for
,
Theorem 2 in [
9] asserts an evaluation of the limit which can be expressed as
where
denotes the (unique up to scaling) stationary measure of the process; see p. 68 in [
1]. In §5, we use (
5) to obtain an MBP version (Theorem 11) of these results by attaching a meaning to
, where
is fixed. This limit turns out to belong to the power series laws generated by the weights
.
Examples for Theorem 11 are presented in §6. If is a critical offspring-law pgf, then is a pgf and . The function generates a renewal sequence , a generalised one in the sense that . In addition the stationary measure is determined by . Theorem 12, a critical analogue of Theorems 9/10 is based on this relation. It draws from Example 6 which suggests that Hausdorff moment sequences may induce a law with the required monotonicity property. Examples 9 and 10 illustrate this idea. Alternatively, one may simply choose an admissible law to work back to a stationary measure. Example 11 exhibits this possibility.
3. A General Criterion
In this section, we discuss a general criterion for constructing a class of laws on the positive integers which strictly contain the LCLs of subcritical MBPs. This material draws on [
6], and the next result tightens Theorem 1.1 in that reference.
Theorem 3. If satisfiesthen defined by (3) is a pgf if and only if Proof. That these two conditions are sufficient is the content of Theorem 1.1 in [
6]. For the converse, it is obviously necessary that
for
. Assuming this condition and (
13), it follows that if (
14) does not hold, then we can choose
and then a natural number
such that
if
. Hence,
, where
is a polynomial of degree
, implying that
, i.e.,
is not a pgf. □
Remark 2. We assume in the sequel that (13) and (14) both hold and that for some . This ensures the existence of the limitLet denote the subset of satisfying (13) and (15). If , then , and hence, we can, and usually will, assume that the limit (15) equals unity, in which case, the resulting pgf induces a family of discrete laws whose pgf is Remark 3. The convention in Remark 2 implies the following representation: For ,Thus, choosing simple forms of provides an easy recipe for constructing pgfs. We illustrate this in the following simple example. Example 1. Let and . Then, andcorresponding to a shifted geometric law. This is the case in §4 of [3], where it is shown that is the pgf of the LCL of a linear birth-and-death process. See Example 5 below. Next, we examine the right-hand tail behaviour of the laws
. This draws on Theorem 2.3 in [
6] although Part (a) is new. We denote with
the set of measurable functions which are slowly varying at infinity, and with
the set of normalised members of
; see [
11] for these familiar notions. In addition, for
, we define the harmonic numbers
.
Theorem 4. (a) Suppose that and that . Then,where . The condition (18) is equivalent to(b) If , then, with ,This condition is equivalent toand Proof. Clearly
where
, and
is defined for
. Differentiating term-wise yields
But
, and hence, for small
, we can choose
such that
if
. Hence,
Hence,
. It follows that
and (
18) follows.
Next, it follows from the definition of
that
and
, and hence,
The first sum on the right-hand side equals
Similarly, the second sum is bounded above by
The estimate (
19) follows.
The equivalence of (
18) and (
19) and the assertions of Part (b) follow from Tauberian theory, as in the proof of Theorem 2.3 in [
6]. □
Remark 4. As observed in [6], the mean of is , and if , then is the pgf of a compound Poisson law. We deviate from the main theme to observe that if
, then the derived pgf
is a compounded pgf, because
is obviously the pgf of a positive-integer-valued random variable (
, for example) with the Sibuya distribution. Hence
is the pgf of the random sum
, where the summands are independent with pgf
B and independent of
.
The pgf of
is
where, for
,
The function
is known to be the Laplace–Stieltjes transform (LST) of a positive random variable (
X, for example),
and its distribution is a generalised gamma convolution law (abbreviated to GGC), i.e., it belongs to the proper subset of positive infinitely divisible (infdiv) laws, which is the smallest set containing all gamma laws and closed under convolution and weak limits. A GGC is characterised by the property that its Lévy measure has a density
such that
is completely monotone:
where
is the so-called Thorin measure, a positive measure satisfying
The representation (
23) implies that
is non-increasing, and hence, any GGC is self-decomposable.
See p. 86 in [
13] or p. 415 in [
14] for the GGC assertion about
, and these references and [
15] for the associated general concepts. In summary, these references show that
X has an exponential mixture law with the specific representation
, where
has the standard exponential law and
denotes a random variable with a standard gamma law with shape parameter
. The three components of
X are independent. This identification of the mixing law for the Sibuya distribution was first reported in [
16].
The quickest way to obtain this identification of
is by observing that
has the type-2 beta density
where the normalising constant
, a beta function. Hence
has the density
the form asserted in [
14]. The LST (
21) follows by writing
, where
has a type-2 beta law, and then, with
,
The GGC property directly arises from the evident fact that
is hyperbolically completely monotone, i.e., if
and
, then, for all
, the product
is a completely monotone function of
w. The density
appears on p. 87 of [
13] in a more complicated form than that obtained above using the form
. The details in that study are omitted with a reference to ‘tedious calculations’.
In what follows, we use properties of the generalised Mittag–Leffler function defined in terms of two parameters,
and
, as follows:
When
, it is understood that
. The original one-parameter version is
, which generalises the exponential function,
. We list known properties used in the sequel (p. 210 in [
17], or [
18]). First, there is an LST identity,
An asymptotic form, valid as
,
and integer
, is
The following result is an (almost) immediate consequence of (
26) and (
27). In particular, it gives a very different-looking representation of
, which is not at all evident from the above integral form.
Theorem 5. The density function of X specified by (21) and (22) has the representationand its asymptotic expansion as is, if , Clearly,
, and hence, the cumulant-generating function (cgf)
of the infinitely divisible random variable
X is, for
,
where
is the Lévy measure of the law of
X, which, by definition, satisfies the condition
.
Differentiation of the cgf yields the Laplace transform relation
The first assertion of the next theorem follows from the above discussion, (
26) and (
27). The second assertion follows from Theorem 7 below.
Theorem 6. (i) The law of X specified by (21) and (22) is a GGC, and its Lévy measure is absolutely continuous with density given by(ii) The corresponding Thorin measure is absolutely continuous, , where A function similar to arises as the density function of a quotient of independent positive stable random variables. Specifically, is the density of , where the components of the quotient are independent and they have a positive stable law with index .
The fact that
X is self-decomposable implies that it has the ‘in law’ stochastic integral representation
where
is a subordinator. Its cgf
(i.e.,
) is related to
by
. See [
19] for these connections. The left-hand side of this identity is the Laplace transform of the right-hand tail
of the Lévy measure of the subordinator. Hence, we have the identification
. In particular,
is the Laplace transform of the Thorin density
.
The pgf
(
20) has a canonical representation as the pgf of a generalised negative-binomial convolution (i.e., a GGC mixture of Poisson laws; see p. 388 in [
14]),
where
. Normalising the Poisson jump-size pgf as
, i.e.,
, we have the evaluations
The author was initially led to Theorems 5 and 6 through finding Entry 11 in the catalogue of complete Bernstein functions recorded in [
15]. This asserts that if
, then
is complete-Bernstein, meaning that it is the cgf of a positive infdiv law whose Lévy measure has a completely monotone density. In fact, the source for this identification [
20] allows the parameter range
. It follows that there is a subordinator
such that
and
. Furthermore, if
has the standard exponential law and is independent of
, then the exponentially stopped version
has the LST
and it has an infdiv law. The law of
X above is recovered by setting
and
.
The next theorem collects the infdiv properties of W.
Theorem 7. If , then the law of W specified by (30) is a GGC. Its pdf isandThe Lévy measure of W is absolutely continuous with the density given byand the corresponding Thorin measure is absolutely continuous with density (recalling (28)) Proof. We observe first that the case
reduces to the case
, as can be seen by defining
and observing that
. This implies the representation
, where
X is as above, and
is independent of
X and has the positive stable law with index
, with the understanding that
. Alternatively,
, where
is the stable subordinator with
. In particular, it follows from Theorem 3.3.2 in [
13]
that W has a GGC law.
Next, the density (
31) follows by rewriting (
30) as
and referring to (
26). The asymptotic estimate follows from (
27).
The identity (
32) follows by writing
and observing that
and referring, again, to (
26).
Finally, it follows from Theorem 3.1.4 in [
13] that the Thorin density
exists and that
, where
denotes the imaginary part of
. But
The identity (
33) follows □
Remark 5. The identity (28) follows from (33) with and , and we observe that . Returning to the complete Bernstein function (
29), denote its (completely monotone) Lévy density as
, and let
, whose Laplace transform is
It follows from (
26) that
The derivative identity
yields the explicit expression
with the understanding that any term in the series expansion (
25) for which the gamma function denominator has a non-positive integer argument contributes zero to the sum. It follows from (
27) that
The complete Bernstein function
has a Stieltjes transform representation
for a certain absolutely continuous measure whose density is listed in [
15] and derived in [
20]. On the other hand, the Lévy density has a Laplace–Stieltjes representation
, implying that
Hence,
, and, referring to the table entry in [
15],
, where
We are now able to state the following result.
Theorem 8. The cgf (29) is Thorin–Bernstein, and its Thorin measure has the density . Proof. The function
is differentiable, and
. Hence,
i.e.,
Hence, a necessary and sufficient condition for
to be Thorin–Bernstein is that
in
.
It suffices to consider only the case
, for if
, then defining
, we have
, where
and, appealing to Theorem 3.3.2 in [
13], we see that proving
is Thorin–Bernstein will prove it for
.
The corresponding form of (
34) is
Finally, setting
, it follows that
if and only if
, where
and
. Note that the denominator is positive.
Carrying out the differentiation yields
, where, suppressing the subscript
,
and
a convex function satisfying
and
. We look at three cases.
First, if
, then
. Next, if
, then
and
. Hence,
is convex increasing from
and
if
. Noting that
, it follows for
that
But
,
and
remembering that
.
The term in square brackets, , satisfies , , and . Hence, is concave increasing and then decreasing, and , if . Hence, for all .
Next, let
, implying that
. The discriminant of
is
because
decreases from unity to
across
. Hence,
if
, and
if
.
If
, then
Obviously,
is convex in
and
. In addition,
has a single minimum at
. The numerical computation shows that
increases from unity at
to
at
. In addition,
. Further numerical calculation shows that
increases from 0 at
to 2 at
. Hence,
if
.
It follows in all cases that
, as desired. Finally, the integrability condition (
24) with
is equivalent to
This condition is satisfied because
as
and
as
. It follows that
is the density of a Thorin measure. □
4. The Subcritical Case
In this section, , in which case, it follows from Theorem 1 that the stationary measure satisfies . Hence, the parts of Theorem 4 with apply to LCLs. In relation to Remark 2, we recall that that if and only if the so-called condition holds, i.e., .
A further general point is that if
, then the derived pgfs
arise as the pgfs of LCLs in the following extended sense. Let
denote an arbitrary initial law,
with
. If the limits
exist and not all of them are zero, then there is a constant
such that
solves the functional equation
This follows from the analogous result for the GWP, p. 65 in [
21]. Differentiating with respect to
t using the backward equation in the form
leads to the first equality in the identity
and the second equality results from (
3) and (
1). Hence,
, i.e.,
. Thus,
is a non-defective conditional limiting law in the above wider sense. In addition, the limits
exist if and only if for some
, the initial law satisfies
i.e., it is attracted to the positive stable law with index
.
With our chosen scaling, it follows from Theorem 1 that
satisfies (
13) and (
14). Conversely, if
satisfies these conditions and if the resulting pgf
satisfies the conditions of the next theorem, then
. The next theorem is simply a recasting of Theorem 2.1 in [
3] expressed in a stronger form.
Theorem 9. Suppose that is a pgf satisfying . Then, the function has a power series expansion converging for , . Let be a constant satisfyingand suppose that for all . Then,and is the pgf of the corresponding LCL. Conversely, if , then it is related to the pgf of its LCL by the identity (36) with as above, and (35) holds. Remark 6. The condition (35) is equivalent to . In addition, There is a very obvious corollary of Theorems 3 and 9, viz., if (13) and (14) both hold and if satisfies the conditions of Theorem 9, then , with f as in (36) and (35). Not surprisingly, there is a gap in that the distributions derived from Theorem 3 may not be the LCL of a MBP. One example suffices to show this. Example 2. Given constants and , let and if . Then, , and is a pgf. It follows that , and it is evident that the coefficients alternate in sign. Hence .
The next two examples complete the details for two examples of LCLs exhibited in [
3] by identifying the stationary measure.
Example 3. Let , , and denote the Cayley tree counting function, i.e., the positive-valued solution of the functional equation . This function is analytic in a disc of radius . It is shown in [3] that if , then is the pgf of a Borel law and it is an LCL. The gf of the corresponding stationary measure iswhere . Now, () andUsing the notation , it follows from the Leibniz rule thatLagrange reversion of the series thus leads to the evaluationIt is not at all obvious from mere inspection that the conditions of Theorem 3 are satisfied, but they are. Example 4. Let and . It is shown in [3] that the log-series law with pgf is the LCL generated by the offspring lawsprovided that A calculation yieldsObserve thatand expand to see thatIt follows that the stationary measure is given byThe last integral can be interpreted in terms of a moment of an exponentially stopped gamma process, , where is a standard gamma process (i.e., it is a subordinator, and ) and is independent of . Theorem 9 expresses the conditions on a pgf ensuring it is that of an LCL. The following result is an equivalent version which imposes conditions on a sequence . Extend the associated -sequence by defining , thus ensuring that when . Next, for , define and observe that is a pmf whose pgf is . In addition, and , implying that .
The following result expresses the conditions of Theorem 9 at a more fundamental level, i.e., in terms of a putative stationary measure. Theorem 10 may also be seen as a version of Theorem 2 restricted to the subcritical case.
Theorem 10. If and , then the functionif and only ifand the coefficients in the power series satisfy when . Equivalently, suppose that is a pmf with pgf and satisfying . Then,if and only ifand the coefficients of the power series satisfy for . It seems that, for some examples, it is a little easier to check the conditions of Theorem 10. Revisiting Example 4, it follows from (
37) that
and, for
,
In the next example, we choose a simple form for the sequence to generalise the case of a geometric offspring law.
Example 5. The geometric offspring law is examined in [8] where, with different notation, it is shown that the corresponding LCL has the pgfThe evaluation of yields , i.e., . We explore the generalisation
, where
and
, i.e.,
. Requiring that
, we assume that
, a condition that is necessary and sufficient for
. Hence,
implying that
is a pgf. In addition,
, as required. The expansion
leads to the evaluations
Computing
leads to the evaluation
whence, for
,
The coefficients
alternate in sign if
, but not if
. In this case,
and
We thus see that the conditions of Theorem 10 are satisfied only if
. Assuming this and observing that
and
, the condition (
38) is
Referring to (
41), we conclude that
if and only if
,
and (
44) holds. It follows from (
39) that the corresponding family of offspring laws is
and, for
,
The boundary case
entails
and, hence, (
44) becomes
and
if
, corresponding to a family of linear birth–death processes generalised in the sense that
if
.
In the case
, the geometric offspring law, it is shown in [
8] that the inverse
of the function
can be expressed in terms of a so-called generalised Wright function. The point of this is that, if such functional inversion is possible, then it follows from (
4) that the population size pgf has an ‘explicit’ representation:
The Wright function, as defined in [
8], is better viewed as a four-parameter extension of the confluent hypergeometric function
in that the definition
specialises to
. The actual Wright function
is the case
and
; see p. 211 in [
17].
It follows from (
41) that
and hence the inverse
satisfies the functional equation
. Lagrange reversion (or the binomial theorem) yields the expansion
where
The signs of the denominator gamma functions oscillate as
j increases. In addition, if
is rational, then
can take non-positive integral values. The corresponding coefficient
. The numerator gamma function equals
, so, substituting this into (
47) and comparing the resulting form of (
46) with (
45), we obtain
This evaluation agrees with the evaluation of
in [
8] (where
; see p. 363) after correcting a minor error: the factor
should be
. The error arises in going from line -10 to -6:
By following the line of argument in [
8], one evaluates the radius of convergence of the series (
46) as
.
6. Critical Case: Examples
A minimal desirable outcome for any example is obtaining explicit expressions for
and/or
, because this gives an identification of the limiting-power-series family whose pgf is
. Better still is obtaining an explicit expression for
and its inverse
, because this would, from (
49), yield the evaluation
We begin with two examples achieving the first aim.
Example 6. If , thenwhere is the principal Lambert function. See [8] for the ramifications of this critical geometric offspring law and, in particular, evaluation of and . We return to this example as Example 10 below. Example 7. If , the critical Poisson offspring law, then some algebra yields the evaluationIt does not seem possible to evaluate in an explicit form. We now pursue a general scheme which yields examples of varying degrees of explicitness. Recalling that
is a pgf, it is observed in [
5] (§7) that
is the generating function of a renewal sequence, but a generalised one in the sense that, because
, we have
. Here, it is more convenient to define the modified pgf
which satisfies
, and hence,
is the generating function of a renewal sequence
which satisfies the standard normalisation
. However,
and hence,
i.e.,
Example 8. Recall Example 6. We have , so , and hence, from (55), if , thenwhere is a probability measure on the closed unit interval . So, in this example, is the moment sequence of a probability law on . This motivates the following strategy. Let
be an arbitrary probability measure on
. It is known that its moments
comprise a renewal sequence, in fact, a Kaluza sequence, meaning that
. Evaluating the sum in (
55) yields
and hence,
i.e., compared with (
54), we see that
Hence, starting from (
56), we obtain the function
In order that
, we require that
a necessary condition. In addition,
But we require that
, i.e., a second necessary condition is that
Finally, since
is a renewal-generating function derived from a pgf,
with
, substituting it into (
59) and expanding it yields
This yields the following result, which is the analogue of Theorems 9 and 10 above for the subcritical MBP.
Theorem 12. Suppose that is a probability measure on , and is specified by (57) and (58). Then, is a critical pgf iff (59) and (60) both hold and the sequence is non-increasing. Remark 7. It is obvious that the assertion will hold if is a standard renewal sequence such that , the condition (60) holds, and the coefficients of are non-decreasing. Before looking at examples, observe that the integration of (
57) yields
and that, despite (
59), the integral is finite because the log-term equals
It is easier to treat specific cases individually rather than using the above general expression.
Example 9. The simplest possible example for (56) is the uniform law on . This yields and . Entry 140 in [22], in essence, states thatMore generally, expanding the integrand yieldsandas required for Theorem 12. The condition (59) is satisfied and, since , the condition (60) is satisfied if . It follows from (
55) that
, the
j-th Harmonic number. Hence,
where we use Entry 111 in [
22], and
is the dilogarithm function.
So, we obtain an explicit expression for
and its coefficients
, but the inverse
looks elusive. The measure
can be extended to
, where
and
. If
, then
where
. Hence,
, entailing a closed expression for
. However, it does not seem possible to go further, even if
or 2.
Example 10. Fix constants , , and choose , thus generalising Example 6. Hence, ,whence . Defining and noting that , some algebra leads toHence, andThus, if , as required. Referring to (58), we conclude thatis a critical pgf iff Continuing, we have
and hence,
Inverting this expression seems intractable except if
, i.e.,
.
In this case,
and
is a critical pgf iff
Writing
,
and
, the Equation (
62) can be recast as
, i.e.,
. This is the Lambert functional equation, whose solution is
, where, because
,
is the principal Lambert function. It follows that the inverse of
is
Thus, an explicit, though complicated, expression for
follows from (
53) and (
62). In particular, we have a complete identification of the limiting conditional law specified by (
52):
Example 6 is recovered when
.
We can say a little more about this case by observing that (
63) can be expressed as
Setting
, where
, this becomes
This family of critical offspring pgfs coincides with the critical theta-offspring laws which have a finite variance identified and explored as Case 2 in [
23].
A more direct way of achieving specifications of
and
is simply to select a mass function
with decreasing masses. We then have
and this is a critical offspring pgf iff
Example 11. Set , a shifted negative-binomial law. The decreasing property holds iff and we have . It does not seem possible to go much further unless σ assumes small integer values. The case is the particular instance of Example 10 with .
If
, then we require
. Defining
, some algebra yields
where
is a signed measure with unit total mass. Hence,
It follows that the stationary measure is given by
, and
Example 12. Another simple choice is , in which case we require that . Also,andHence, Let
and consider the function
The integrand can be expanded as a power series and the result integrated term-wise. If
, this yields
In particular,
and
, implying that