1. Introduction and Motivation
As a main difference with respect to deterministic (or classical) differential equations, solving a random differential equation (RDE) does not only consist of determining, exactly or approximately, its solution stochastic process (SP), say
, but also of computing its main statistical properties such as the mean and the variance. Even more, the approximation of the first probability density function (1-PDF), say
, of the solution SP is a more ambitious and useful goal since from it, one can calculate, via its integration, the mean and the variance of
,
as well as its higher one-dimensional statistical moments (such as asymmetry, kurtosis, etc.),
provided they exist. Moreover, given a fixed time instant
, the 1-PDF permits calculating key information as the probability that the solution varies in any specific set of interest, say
, just via its integration,
The computation of the 1-PDF in the setting of RDEs and their applications is currently a cutting-edge topic for which a number of interesting advances have been achieved in the recent literature. Here, we highlight the following contributions that are strongly related to the goal of the present paper [
1,
2,
3,
4]. The main aim of this paper is to contribute to advance the field of RDEs by extending important classical results to the stochastic context. Specifically, we address the problem of constructing reliable approximations for the 1-PDF of the solution SP to second-order linear differential equations whose coefficients are analytic SPs depending on a random variable (RV) about a regular-singular point and whose initial conditions (ICs) are also RVs. Therefore, we are dealing with what is often called RDEs having a finite degree of randomness [
5] (p. 37). In this manner, we complete the analysis performed in our previous contribution [
6], where we dealt with the ordinary point case. It is important to emphasize that the analysis of second-order linear random differential equations has been performed in previous contributions using the random mean square calculus [
5] (Ch. 4) (see [
7,
8,
9,
10,
11,
12] for the Airy, Hermite, Legendre, Laguerre, Chebyshev, and Bessel equations, respectively) and using other approaches like the random homotopy method [
13,
14], the Adomian method [
15], the differential transformation method [
16], the variational iteration method [
17], etc. However, in all these contributions, only approximations for the two first statistical moments (mean and variance) of the solution were calculated. In contrast, in [
6] and in the present paper, we deal with the general form of the aforementioned second-order linear RDEs, and furthermore, we provide additional key information of the solution via the computation of approximations for the 1-PDF that, as has been previously indicated, permits calculating not only the mean and the variance, but also higher moments of the solution, as well as further relevant information as the probability that the solution lies in specific sets of interest. To the best of our knowledge, these results for the 1-PDF of second-order linear RDEs about regular-singular points are new, and then, they contribute to advance the setting of this important class of RDEs. At this point, is important to underscore the main differences between our previous contribution [
6] and the present paper. In [
6], we provided a comprehensive study of second-order linear differential equations with random analytic coefficients about ordinary points. Now, we propose a natural continuation of [
6] by extending the analysis for the same class of differential equations, but about singular-regular points, whose mathematical nature is completely distinct. Our aim is to complete the stochastic analysis for this important type of differential equation inspired by the well-known extension of the deterministic setting, the first one dealing with ordinary points and secondly with singular-regular points, by applying the Fröbenius theorem. It is important to point out that apparently, the problems have a strong similarity, but they are applied to differential equations of a different nature. In this regard, this new contribution allows us to study, for example, the important randomized Bessel differential equation, which does not fall within the mathematical setting of the previous contribution [
6].
For the sake of completeness, below, we summarize the main results about the deterministic theory of second-order linear differential equations about ordinary and regular-singular points. These results will be very useful for the subsequent development. Let us then consider the second-order linear differential equation:
where coefficients
,
are analytic functions at a certain point, say
, i.e., they admit convergent Taylor series expansions about
(in practice, these coefficients are often polynomials, which are analytic everywhere). To study when this equation admits a power series solution centered at the point
, say
(where the coefficients
,
must be determined so that this series satisfies the differential equation), it is convenient to recast Equation (
2) in its standard or canonical form:
where
and
. The key question is how must we pick the center of the previous expansion,
, because this choice fully determines the region of convergence of the power series. To this end,
is classified as an ordinary or a singular point. Specifically,
is called an ordinary point of the differential Equation (
3) (or equivalently (
2)) if coefficients
and
are both analytic at
, otherwise
is termed a singular point. Recall that the quotient of analytic functions is also an analytic function provided the denominator is distinct from zero. Therefore, if
in (
2) (and using that their coefficients are analytic functions about
), then
is an ordinary point. In the particular case that
,
are polynomials without common factors, then
is an ordinary point of (
2) if and only if
. It is well known that when
is an ordinary point of differential Equation (
2), then its general solution can be expressed as
, where
are free constants and
and
are two linearly independent series solutions of (
2) centered at the point
. These series are also analytic about
, i.e., they converge in a certain common interval centered at
:
(
being
the radius of convergence of
,
, respectively). In the important case that coefficients
,
are polynomials, the radius of convergence
of both series is at least as great as the distance from
to the nearest root of
. As a consequence, if the leading coefficient
is constant, then a power series solution expanded about any point can be found, and this power series will converge on the whole real line. In [
6], we solved, in the probabilistic sense previously explained, the randomization of Equation (
3) by assuming that coefficients
and
depend on a common RV, denoted by
A, together with two random ICs fixed at the ordinary point
, namely
and
.
The study of Equation (
3) (or equivalently, (
2)) about a singular point requires further distinguishing between regular-singular points and irregular-singular points. In this paper, we shall deal with the former case, which happens when
approaches infinity no more rapidly than
and
no more rapidly than
, as
. In other words,
and
have only weak singularities at
, i.e., writing Equation (
3) in the form:
where:
then
and
are analytic about
. Otherwise, the point
is called an irregular-singular point. In the case that
and/or
defined in (
4) become indeterminate forms at
, the situation is determined by the limits:
If
, then
may be an ordinary point of the differential equation
(or equivalently, dividing it by
and taking into account (
5), of the differential Equation (
3), i.e.,
). Otherwise, if both limits in (
6) exist and are finite (and distinct form zero), then
is a regular-singular point, while if either limit fails to exist or is infinite, then
is an irregular-singular point. As has been previously underlined, the most common case in applications, in dealing with differential equations of the form (
4), is when
and
are both polynomials. In such a case,
and
are simply the coefficients of the terms
of these polynomials, if they are expressed in powers of
, so
is a regular-singular point. In dealing with the case that
is a regular-singular point, once the differential Equation (
2) is written in the form (
4), we are formally working in a deleted (or punctured) neighborhood of
, say
,
. The solution of (
4) (equivalently of (
2)) is then sought via generalized power series (often called Fröbenius series) centered at the point
,
,
, being
and where
r is a (real or complex) value to be determined, as well as coefficients
,
by imposing that this series satisfies the differential Equation (
4) (equivalently of (
2)). This leads to the fact that parameter
r must satisfy the following quadratic equation:
This equation is called the indicial equation of differential Equation (
8), and its two roots,
and
(possibly equal), are called the exponents of the differential equation at the regular-singular point
,
which is obtained after multiplying (
4) by
. The full solution of differential Equation (
8) (or equivalently, (
2)) is given in the following theorem in terms of the nature of
and
.
Theorem 1. (Fröbenius method) [18] (p. 240). Let us consider the differential Equation (2) whose coefficients are analytic about , and assume that and defined in (4) and (5) are analytic about . Let us assume that is a regular-singular point of the differential Equation (2). Let and be the roots of the indicial equation associated with (8) at the point :where and are defined in (6). Without loss of generality, we assume that , where stands for the real part. Then, the differential Equation (2) has two linearly independent solutions, and , in a certain deleted neighborhood centered at , of the following form: - (a)
If is not a non-negative integer (i.e., ),where and . - (b)
If is a positive integer (i.e., ),where , , and c is a constant, which may or may not be distinct from zero. - (c)
If ,where .
As was previously indicated, in this contribution, the objective is to extend the analysis performed in [
6] for the aforementioned randomization of differential Equation (
3) (equivalently of (
2)) about an ordinary point
, to the case that
is a regular-singular point. Specifically, we will consider the following second-order linear RDE:
together with the following random ICs:
fixed at an initial instant
belonging to a certain deleted interval centered at
,
,
. Then, we will construct approximations of the 1-PDF,
, of the SP,
, to this random initial value problem (IVP) via a random Fröbenius series centered at the regular-singular point
. In (
10),
and
are SPs satisfying certain hypotheses, which will be stated later, which depend on the common RV denoted by
A. Here, we assume that we are working on a probability space
, which is complete, to which the real RVs
A,
, and
belong. Furthermore, we assume that these three RVs have a joint density, i.e., the random vector
is absolutely continuous.
As usual, observe that in (
10) and (
11), we are writing RVs with capital letters, for instance,
(here,
is referred to as the range of
Z). Moreover, each realization of an RV, say
Z, will be denoted by
,
or simply
. To be as general as possible, in the subsequent analysis,
A,
, and
are assumed to be dependent absolutely continuous RVs, and
will denote their joint PDF. Notice that, if
A,
, and
are independent RVs, then their PDF can be factorized as the product of their respective marginal PDFs, i.e.,
Based on our previous contribution [
6], together with Theorem 1 and further reasons that will be apparent later, hereinafter, we will assume the following hypotheses:
H0: A is a bounded RV, i.e.,
H1:
is continuous in the second component and bounded, i.e.,
In addition, we will assume that the SPs and are analytic about for every , , i.e.,
H2: There exists a common neighborhood
where:
Here,
denotes a deleted interval centered at
that has been previously defined. Recall that here, the SP
is analytic about a point
,
, for all
, if the deterministic function
is analytic about
(equivalently to the SP
). To simplify notation, we shall assume that
and
are analytic in a common neighborhood that will be denoted by
. In practice, this neighborhood is determined intersecting the domains of analyticity of
and
. In [
5] (Th. 4.4.3), a characterization was stated of analyticity of second-order SPs (those having finite variance) in terms of the analyticity of the correlation function. Moreover, to ensure that the IVP (
10)–(
11) has a unique solution, both SPs
and
are assumed to satisfy all the needed conditions. (see [
5] (Th. 5.1.2), for instance).
According to the Fröbenius method (see Theorem 1) and under the analyticity condition assumed in Hypothesis H2, the solution SP of RDE (
10) about a regular-singular point,
, can be written as a linear combination of two uniformly convergent independent random series,
and
,
where
, and the coefficients
and
can be obtained in terms of the random ICs given in (
11), which are established at the time instant
. In the subsequent development, we will take advantage of the above representation along with the application of the random variable transformation (RVT) technique [
5] (p. 25), [
6] (Th. 1), to construct the 1-PDF,
, corresponding to approximations,
, of the solution SP,
, about the regular-singular point
. We shall provide sufficient conditions so that
approximate the 1-PDF,
, of
. The RVT method has been successfully applied to obtain exact or approximate representations of the 1-PDF of the solution SP to random ordinary/partial differential and difference equations and systems [
1,
2,
3,
4]. In particular, the RVT technique has also been applied to conduct the study for the random IVP (
10) and (
11) in the case that
is an ordinary point [
6]. Thus, the present contribution can be considered as a natural continuation of our previous paper [
6]. This justifies that in our subsequent development, we directly refer to [
6] when applying a number of technical results already stated in the above-mentioned contribution. In this manner, we particularly avoid repeating the multidimensional random variable transformation technique [
6] (Th. 1), Poincare’s expansion [
6] (Th. 2), as well as several results related to uniform convergence that can be inferred from classical real Analysis [
6] (Prop. 1-4). For the sake of clarity, we point out that we keep identical the notation in both contributions. These results were also extensively applied in [
6,
19,
20,
21].
The paper is organized as follows. In
Section 2, the 1-PDF,
, of the approximate solution SP,
, to the random IVP (
10) and (
11) is formally constructed. This function is obtained by applying the RVT method to
, which follows from truncating the random generalized power series solution derived after applying the Fröbenius method stated in Theorem 1.
Section 3 is devoted to rigorously proving the convergence of approximations
to the exact 1-PDF,
, associated with the solution SP,
. In
Section 4, several illustrative examples are shown to demonstrate the usefulness of the theoretical results established in
Section 2 and
Section 3. Our main conclusions are drawn in
Section 5.
3. Study of the Convergence
In this section, we give sufficient conditions to guarantee the convergence of the approximation,
, given in (
21), to the exact function
when
N tends to infinity, i.e., we study when:
being:
where
and
are the realizations of the random series
and
defined in (
17).
Now, we include here some remarks about the convergence and boundedness of the series involved in (
23), which will play a key role in our subsequent developments.
Remark 2. Expanding first the terms and as their corresponding series and interpreting the RV A as a parameter indexed by , , under the hypothesis H2, Poincaré’s theorem, stated in [6](Th. 2), permits representing the solution as a series of parameter , . Therefore, and do. On the other hand, taking into account the uniqueness of the solution of IVP (10)–(11), both series expansions (as powers of and as powers of ) match. Henceforth, the series and , given by (14), are convergent in -deleted neighborhoods and , respectively, for all , . In addition, uniform convergence takes place in every closed set contained in and . Notice that the domain of convergence, , of the series solution given in (13), satisfies for all , , where is defined in Hypothesis H1. Remark 3. Functions and are linear combinations of series and (see (17)), which, by Remark 2, are uniformly convergent in every closed set contained in . Then, and also converge uniformly in . Remark 4. Note that for every : Then, according to Remark 2, there exist a certain neighborhood, , and a positive constant, , such that for all : On the other hand, by Remark 3, , , converge uniformly in every closed set contained in for all , . This guarantees the existence of constants such that: Let
be fixed; below, we establish conditions to assure the convergence stated in (
22). First, let us take limits as
in Expression (
21):
To commute the limit as
and the above double integral, we apply [
6] (Prop. 1). We assume the following hypothesis:
H3:
is a Lebesgue measurable set of
with finite measure such that:
Then, we shall prove that
and
, being:
and:
By Remark 4, considering the lower bound given in (
24) for
, one gets:
Therefore, as
is a PDF, we conclude that
is Lebesgue integrable in
, i.e.,
. To prove that
uniformly in
, we will apply the same argument as in the previous contribution [
6]. We demonstrate that for every
, there exists
such that for all
:
Adding and subtracting the term
and applying the triangular inequality, one gets:
Let
be fixed and
be an arbitrary non-negative integer, and consider
, for all
,
. Then, we apply [
6] (Prop. 2) to
fixed,
arbitrary,
,
,
,
, and
. Note that, by Remark 4,
for all
. In addition, it is obvious that
converges uniformly to
on
D. Regarding the real function
, it converges uniformly to
, by the arguments shown in Remarks 2 and 3. As
, its absolute value is bounded. By Remark 4, there exists constants
and
, such that
,
. Taking into account the uniform convergences, the bounds, and the hypothesis H1 (boundedness of the PDF
), for every
, there exists
(which depends on
) such that:
independently of the values of
.
Now, we obtain an analogous result for the second term. First, assuming Hypothesis H3, we apply [
6] (Prop. 3) to
,
,
,
,
,
and
, with
fixed, and such that
,
. By Remark 3,
converges uniformly to
on
. Then, given
fixed,
converges uniformly to
in
. Secondly, we apply [
6] (Prop. 2) taking
,
,
,
,
, and
. Note that, previously we have proven, by applying [
6] (Prop. 3), that
converges uniformly to
. In addition, by Remark 3,
converges uniformly to
. Let
and
, these bounds being the ones established in Hypothesis H3 and Remark 4, then:
The next step is to apply [
6] (Prop. 4) with the following identification:
,
,
such that
for every
,
,
,
,
,
and
. Finally, as it was previously shown, the sequence
is uniformly convergent to
, and according to Hypothesis H1, the mapping
is continuous in
. Therefore, taking into account the lower bound of
given in Remark 4 and applying [
6] (Prop. 4), for every
there exists
, which depends on
, such that
independently of the values of
. Hence, we proved that
, and then, by [
6] (Prop. 1) and Hypothesis H3, we can commute the limit and the integral. Therefore, applying the continuity of the joint PDF
with respect to the second variable (see Hypothesis H1) and taking into account Expression (
23), one gets:
Summarizing, we establish the following result.
Theorem 2. Let us consider the random IVP (10) and (11), and assume that: - (i)
The RV A satisfies Hypothesis H0.
- (ii)
The joint PDF, , of the random input data satisfies Hypothesis H1.
- (iii)
The coefficients and satisfy Hypothesis H2.
- (iv)
The domain of the random vector satisfies Hypothesis H3.
Let and be the finite sum defined in (20). Then, , defined by (19)–(21), is the 1-PDF of the truncated solution SP given by (13)–(15) to the IVP (10) and (11). Furthermore, for each , 1-PDF converges to given in (17) and (23).