Skip to content
Publicly Available Published by De Gruyter January 29, 2021

On the asymptotics of wright functions of the second kind

  • Richard B. Paris , Armando Consiglio and Francesco Mainardi EMAIL logo

Abstract

The asymptotic expansions of the Wright functions of the second kind, introduced by Mainardi [see Appendix F of his book Fractional Calculus and Waves in Linear Viscoelasticity (2010)],

Fσ(x)=n=0(x)nn!Γ(nσ),Mσ(x)=n=0(x)nn!Γ(nσ+1σ)(0<σ<1)

for x → ± ∞ are presented. The situation corresponding to the limit σ → 1 is considered, where Mσ(x) approaches the Dirac delta function δ(x − 1). Numerical results are given to demonstrate the accuracy of the expansions derived in the paper, together with graphical illustrations that reveal the transition to a Dirac delta function as σ → 1.

1 Introduction

The particular Wright function under consideration (also known as a generalized Bessel function) is defined by

(1.1)Wλ,μ(z)=n=0znn!Γ(λn+μ),

where λ is supposed real and μ is, in general, an arbitrary complex parameter. The series converges for all finite z provided λ > −1 and, when λ = 1, it reduces to the modified Bessel function z(1μ)/2Iμ1(2z). The asymptotics of this function were first studied by Wright [14, 15] using the method of steepest descents applied to the integral representation

(1.2)Wλ,μ(z)=12πi(0+)tμet+ztλdt  (λ>1,μC).

The case corresponding to λ = −σ, 0 < σ < 1 arises in the analysis of time-fractional diffusion and diffusion-wave equations. The function with negative λ has been termed a Wright function of the second kind by Mainardi [3], with the function with λ > 0 being referred to as a Wright function of the first kind. In the former context, Mainardi [4, Appendix F] defined the auxiliary functions

(1.3)Fσ(z)=Wσ,0(z)=n=1(z)nn!Γ(nσ),  0<σ<1,
(1.4)Mσ(z)=Wσ,1σ(z)=n=0(z)nn!Γ(nσ+1σ),  0<σ<1.

These functions are interrelated by the following relation:

(1.5)Fσ(z)=σzMσ(z).

The case μ = 0 in (1.1) also finds application in probability theory and is discussed extensively in [13], where it is denoted by

(1.6)ϕ(λ,0;z)=Wλ,0(z)

and referred to as a ’reduced’ Wright function.

Plots of Mσ(x) for real x and varying σ are presented in [4, Appendix F] and [5]. These graphs illustrate the transition between the special values σ=0,12,1, where Mσ(x) has simple representations in terms of known functions. These are

(1.7)M0(x)=ex,M1/2(x)=1πex2/4,M1/3(x)=32/3Ai(x/31/3),

where Ai is the Airy function. As σ → 1, the function Mσ(x) tends to the Dirac delta function δ(x − 1).

In this paper we present the asymptotic expansions of Fσ(x) and Mσ(x) for x → ±∞ by exploiting the known asymptotics of the function ϕ(−σ, 0, x) discussed in [13]. The resulting expansions involve a combination of algebraic-type and exponential-type expansions, for which explicit representation of the coefficients in both types of expansion is given. In order to give a self-contained account, we describe the derivation of the expansion for Mσ(x) based on the asymptotics of integral functions of hypergeometric type described in [10] (see also [11, §4.2]). The asymptotic treatment of the function Wλ, μ(z) given by Wright [14], [15] did not give precise information about the coefficients appearing in the exponential expansions; see also [10] for a more detailed account.

2 The asymptotic expansions of Fσ(x) and Mσ(x) for x → ±∞

We define the quantities

(2.1)κ=1σ,ϑ=σ12,h=σσ,X=κ(hx)1/κ,A(σ)=2πσ(σκ)σ.

The connection between Fσ(x) and the function ϕ defined in (1.6) is

Fσ(x)=ϕ(σ,0,x).

The asymptotic expansions of ϕ(−σ, 0, x) for x → ±∞ when 0 < σ < 1 are given in [13, §5.2]. We therefore obtain the expansions stated in the following theorem:

Theorem 2.1

When 0 < σ < 1 we have the expansion of the auxiliary Wright functionFσ(x) given by[*]

(2.2)Fσ(x)~A(σ)2πX1/2eXj=0cj(σ)(X)j  (0<σ<1)

and

(2.3)Fσ(x)~{E(x)+H(x)(0<σ<12)H(x)(12<σ<1)

asx → +∞, whereA′(σ) = A(σ)(σ/κ)κandc0(σ) = 1. The formal exponential and algebraic expansionsE′(x) andH′(x) are defined by (see [13, (5.10), (5.11)])

E(x):=A(σ)πX1/2eXcosπσ/κj=0cj(σ)(X)jcos[Xsinπσκ+πκ(ϑj)]

and

H(x):=1σk=0x(k+1)/σk!Γ(1k+1σ).

The case σ=12 needs no special attention since

F1/2(x)=x2πex2/4,

but see the comment at the end of Section 3 as this case is associated with a Stokes phenomenon.

The coefficients cj(σ) appearing in the exponential expansions in Theorem 2.1 can be obtained [**] from [10, (4.6)] (when the parameter δ therein is replaced by σ). We have:

(2.4)cj(σ)=(2σ)(12σ)23j3jj!σjdj(σ)(j1),

where the first few coefficients dj(σ) are:

d1(σ)=1,d2(σ)=2+19σ+2σ2,d3(σ)=15(5561628σ9093σ21628σ3+556σ4),d4(σ)=15(4568+226668σ465702σ22013479σ3465702σ4+226668σ5+4568σ6),d5(σ)=17(262206412598624σ167685080σ2+302008904σ3+1115235367σ4+302008904σ5167685080σ612598624σ7+2622064σ8)d6(σ)=135(167898208+22774946512σ88280004528σ2611863976472σ3+1041430242126σ4+3446851131657σ5+1041430242126σ6611863976472σ788280004528σ8+22774946512σ9+167898208σ10).

These polynomial coefficients are related to the so-called Zolotarev polynomials, see [13].

From the relation (1.5), we have Mσx) = Fσx)/(±πx) and after a little algebra we deduce the expansion of Mσ(x) given by:

Theorem 2.2

When 0 < σ < 1 we have the expansion of the auxiliary Wright functionMσ(x) given by:

(2.5)Mσ(x)~A(σ)2πXϑeXj=0cj(σ)(X)j  (0<σ<1)

and

(2.6)Mσ(x)~{E^(x)+H^(x)(0<σ<12)H^(x)(12<σ<1)

asx → +∞, where the coefficientscj(σ) are as defined in Theorem 2.1. The formal exponential and algebraic expansionsE^(x)andH^(x)are defined by

E^(x):=A(σ)πXϑeXcosπσ/κj=0cj(σ)(X)jcos[Xsinπσκ+πκ(ϑj)]

and

H^(x):=1σk=1x(k+σ)/σk!Γ(kσ).

For x → +∞, the function Mσ(x) is exponentially small for all values of σ in the interval 0 < σ < 1. The case of Mσ(−x), however, is seen to be more structured. When 0<σ<13, the factor cos πσ/κ > 0 and Mσ(−x) is exponentially large (with an oscillation) as x → +∞, with the algebraic expansion H^(x) being subdominant. When σ=13, this factor is zero and E^(x) is oscillatory with an algebraically controlled amplitude and H^(x)0. When 13<σ<12, the expansion E^(x) is exponentially small and the behaviour of Mσ(−x) is controlled by the algebraic expansion. Finally, when 12<σ<1 the expansion of Mσ(−x) is purely algebraic in character.

3 The asymptotic expansion of Mσ(x) for x → ±∞

In order to make this paper more self contained we present in this section an alternative derivation of the expansion of Mσ(x) as x → ±∞. Define the function

(3.1)(z):=n=0Γ(nσ+σ)n!zn  (0<σ<1).

Then use of the reflection formula for the gamma function shows that the auxiliary Wright function Mσ(x) defined in (1.4) can be expressed in terms of (x) as

Mσ(x)=1πn=0Γ(σn+σ)n!(x)nsinπ(n+1)σ
(3.2)=12π{eπiϑ(xeπiκ)+eπiϑ(xeπiκ)},

and in a similar manner

(3.3)Mσ(x)=12π{eπiϑ(xeπiσ)+eπiϑ(xeπiσ)}.

From the discussion in [10, Section 2], the Stokes lines for (z), where its exponential expansion is maximally subdominant relative to its algebraic expansion, are situated on the rays arg z = ±πκ. An important distinction between (3.2) and (3.3) when x > 0 is that for Mσ(−x) the arguments of the functions (xe±πiσ) are only situated on the Stokes lines arg z = ±πκ when σ=12, since κ=1σ=12, whereas for Mσ(x) the arguments of (xe±πiκ) are situated on the Stokes lines for all values of σ in the range 0 < σ < 1.

From [10, §4.1] (see also [12, §2.3]), the asymptotic expansion of (z) is given by

(3.4)(z)~{E(z)+H(zeπi)(|argz|πκϵ)H(zeπi)(πκ+ϵ|argz|π)

as |z| → ∞. The upper or lower signs are chosen according as arg z > 0 or arg z < 0, respectively and ϵ denotes an arbitrarily small positive quantity. The formal exponential and algebraic expansions E(z) and H(z) are defined by

(3.5)E(z):=A(σ)ZϑeZj=0cj(σ)Zj,  Z:=κ(hz)1/κ,
(3.6)H(z):=1σk=0()kk!Γ(k+σσ)z(k+σ)/σ,

where the parameters κ, h, ϑ and A(σ) are defined in (2.1) and the coefficients cj(σ) are those appearing in Theorem 1; see Appendix A for an algorithm for the calculation of these coefficients.

The exponential expansion E(z) is dominant in the sector |argz|<12πκ and becomes exponentially small in the adjacent sectors 12πκ<|argz|πκ. On arg z = ±πκ, E(z) is maximally subdominant relative to the algebraic expansion and switches off in a smooth manner (at fixed |z|) across these Stokes lines. The expansion in this case is given in Subsection 3.1.

3.1 The expansion of Mσ(x) as x → +∞

To deal with this case we require the expansion of (xe±πiκ) for large x > 0. As stated above, the arguments of (z) are situated on the Stokes lines arg z = ±πκ, where the exponential expansion is in the process of switching off as |arg z| increases. From [10, (4.7)], we have the expansion

(3.7)(xe±πiκ)~e±πiσσk=0m1Γ(k+σσ)k!x(k+σ)/σ+(Xe±πi)ϑeXj=0(12Aj(σ)±iBj(σ)2πX)(X)j

as x → +∞, where Aj(σ) = A(σ)cj(σ) and m denotes the optimal truncation index (that is, truncation at, or near, the smallest term) of the algebraic expansion; see also [9, §4.2]. The coefficients Bj(σ) involve linear combinations of the Aj(σ); see [10, §4.1]. However, the precise values of m and Bj(σ) do not concern us here since in the combination (3.2) the algebraic expansion and the terms involving Bj(σ) all cancel.

The algebraic component of the right-hand side of (3.2) is then seen to be

12πσk=0m1()kΓ(k+σσ)k!{eπiϑ(xeπiκeπi)(k+σ)/σ+eπiϑ(xeπiκeπi)(k+σ)/σ}=cosπ(ϑσ)πσk=0m1Γ(k+σσ)k!x(k+σ)/σ0,

upon recalling that ϑ=σ12. The exponentially small contributions involving the coefficients Bj(σ) in (3.7) are also seen to cancel in the combination in (3.2), thereby yielding the expansion (2.5) stated in Theorem 2.2.

3.2 The expansion of Mσ(−x) as x → +∞ (when σ12)

The algebraic component in the expansion for Mσ(−x) is from (3.6) and (3.3)

(3.8)H^(x):=12π{eπiϑH(xeπiσeπi)+eπiϑH(xeπiσeπi)}=12πiσk=0Γ(k+σσ)k!{(xeπi)(k+σ)/σ(xeπi)(k+σ)/σ}=1σk=1x(k+σ)/σk!Γ(k/σ).

Note that H^(x)0 when σ = 1/p, p = 2, 3, 4, …. The exponential component (with ω := eπiσ/κ for brevity) is, from (3.5),

(3.9)E^(x):=12π{eπiϑE(xeπiσ)+eπiϑE(xeπiσ)}=Xϑ2π{eXω+πiϑ/κj=0Aj(σ)(Xω)j+eX/ωπiϑ/κj=0Aj(σ)(X/ω)j}=XϑπeXcosπσ/κj=0Aj(σ)(X)jcos[Xsinπσκ+πκ(ϑj)],

provided 0<σ<12. Then, from (3.4), we obtain the expansion (2.6) in Theorem 2.2.

Remark 3.1

The expansion (2.6) in Theorem 2.2 does not hold when σ=12 as this case requires a separate treatment on account of the Stokes phenomenon. However, this is not essential here since by (1.7) we have the exact value M1/2x) = π−1/2exp[−x2/4]. It is worth noting that when σ=12=κ, the algebraic expansion H^(x)0 and, since cj(12)=0 for j ≥ 1, the exponential expansion E^(x) in (3.9) reduces to 2π−1/2exp[−x2/4], which is twice the correct value. This is due to our not having taken into account the Stokes phenomenon present in the particular case of (2.6) in Theorem 2 corresponding to σ=12.

4 Numerical results

We present some numerical results to verify the expansions in Theorems 1 and 2. In Table 1 the values (accurate to 10dp) of the coefficients cj(σ) appearing in the exponential expansion are shown for two values of σ. Table 2 shows the absolute relative error in the computation of Mσ(x) as a function of the truncation index j with the expansion (2.5) in Theorem 2.2. Table 3 shows the same error in the computation of Mσ(−x) for different values of x with the expansion (2.6). Note that for σ = 1/4 and σ = 1/3 in Table 3 we have H^(x)0. For σ = 2/5, the algebraic expansion H^(x) has been optimally truncated, but for σ = 2/3 the truncation index was taken as k = 11.

The limit σ → 1 in Mσ(x) can be obtained by setting σ = 1 − ϵ, ϵ → 0+ so that the parameters in (2.1) become

κ=ϵ,ϑ=12ϵ,X=ϵ1ϵ(x(1ϵ))1/ϵ,A(σ)=2π1ϵ(1ϵϵ)1ϵ.

Table 1

Values of the coefficients cj(σ) for σ = 1/4 and σ = 3/4

jσ = 1/4σ = 3/4
0+1.0000000000+1.0000000000
1+0.1458333333−0.0347222222
2+0.0835503472−0.0167582948
3+0.0597617067−0.0224719333
4+0.0052249186−0.0510817883
5−0.2249669579−0.1651975373
6−1.1657705000−0.6952815250

Table 2

Values of the absolute relative error in the computation of Mσ(x) for different truncation index j

σ = 1/4σ = 3/4
x = 6x = 10x = 4x = 6
02.623 × 10−21.376 × 10−21.262 × 10−32.531 × 10−4
12.819 × 10−37.618 × 10−42.190 × 10−58.881 × 10−7
24.123 × 10−45.561 × 10−51.054 × 10−68.654 × 10−9
42.877 × 10−51.336 × 10−69.988 × 10−93.359 × 10−12
62.915 × 10−53.111 × 10−72.819 × 10−103.874 × 10−15

Table 3

Values of the absolute relative error in the computation of Mσ(−x) for varying x

xσ = 1/4σ = 1/3σ = 2/5σ = 2/3
45.260 × 10−23.447 × 10−46.825 × 10−26.130 × 10−4
62.176 × 10−41.570 × 10−52.863 × 10−22.988 × 10−6
86.088 × 10−62.510 × 10−65.153 × 10−43.365 × 10−9
103.787 × 10−63.111 × 10−74.993 × 10−56.279 × 10−11
121.048 × 10−71.508 × 10−81.431 × 10−72.397 × 10−12

Then from Theorem 2.2 we obtain the leading behavior:

(4.1)Mσ(x)~(x(1ϵ))1/(2ϵ)12πϵexp[ϵ1ϵ(x(1ϵ))1/ϵ],
(4.2)Mσ(x)~ϵx2ϵ(1ϵ)Γ(1+1σ){1+O(x1/σ)}

as x → +∞ and ϵ → 0. The above approximation for Mσ(x) agrees with that obtained in [6] by application of the saddle-point method applied to the integral (1.2). This argument is explained in Section 5.

Plots of Mσ(x) given by (4.1) are shown in Figs. 1, 2 and 3, and plots of Mσ(−x) given by (4.2) are shown in Fig. 4. These illustrate the transition to a Dirac delta function as ϵ → 0.

Figure 1 Plots of Mσ(x) for ϵ = 0.1 in linear (up) and semi-logarithmic scale (down)
Figure 1

Plots of Mσ(x) for ϵ = 0.1 in linear (up) and semi-logarithmic scale (down)

5 The Kreis-Pipkin Method

This section focuses on the argument introduced as a variant of the saddle-point method by Kreis and Pipkin in [2] (revisited by Mainardi and Tomirotti in [7] for a wave problem in fractional viscoelasticity) to deal with sharply peaked functions around x ~ 1, in the limit where ϵ → 0+. The method is of interest from a numerical point of view, allowing us to deal with functions that are also physically relevant such as, in seismology, the pulse response in the nearly elastic limit.

Figure 2 Plots of Mσ(x) for ϵ = 0.01 in linear (up) and semi-logarithmic scale (down)
Figure 2

Plots of Mσ(x) for ϵ = 0.01 in linear (up) and semi-logarithmic scale (down)

In this way it is possible, adapting the Kreis-Pipkin method to the M-Wright function, to study its asymmetric structure when it tends towards the Dirac delta function δ(x − 1).

We start by recalling the integral definition of the auxiliary Wright function Fσ(x) (compare (1.2))

(5.1)Fσ(x)=12πi(0+)etxtσdt,x>0,0<σ<1,

related to the function Mσ(x) by (1.5). Taking into account the procedure described in [2], we have with σ = 1 − ϵ that the exponent is stationary at the point:

t0ϵ=1x(1ϵ).
Figure 3 Plots of Mσ(x) for ϵ = 0.001 in linear (up) and semi-logarithmic scale (down)
Figure 3

Plots of Mσ(x) for ϵ = 0.001 in linear (up) and semi-logarithmic scale (down)

Figure 4 Plots of Mσ(−x) for different values of ϵ in linear (up) and semi-logarithmic scale (down)
Figure 4

Plots of Mσ(−x) for different values of ϵ in linear (up) and semi-logarithmic scale (down)

The next step is to expand tϵ in powers of ϵln t/t0, this being more accurate than expanding the exponent in powers of tt0, and using z = t/t0. The final result is:

(5.2)Fσ(x)~Λ2πiϵ(0+)eΛz(lnz1)dz,  Λ=ϵt0,

where we emphasize that this procedure is valid only in the limit ϵ → 0+. The relation (1.5) tells us that the expression of Mσ(x) can be simply obtained from knowledge of Fσ(x), and vice versa. The exponential factor appearing in (5.2) has a saddle point at z = 1 and the contour can be made to coincide with the steepest descent path, which is locally perpendicular to the real z-axis at the saddle. Then finally, by means of the steepest descent method, the function Mσ(x) as σ → 1 can be expressed via a real integral.

The results are presented in Figs. 5, 6 and 7; each figure shows a comparison in linear and semi-logarithmic scale between three curves obtained using different methods. These are respectively the Kreis-Pipkin method, (4.1) of this work and the classical saddle-point method used by Mainardi and Tomirotti [6] (denoted by M-T 1995 in the figures). Note that the curves obtained via (4.1) and M-T 1995 are equivalent, and indeed can be simply shown to be analytically equivalent. The plots for 0 ≤ x ⋍ 1 in the Kreis-Pipkin method were obtained via an integral representation for Mσ(x) combined with matching to the leading asymptotic behavior. The method proposed by Kreis and Pipkin is thus seen to be a useful tool to reproduce the asymmetric structure of Mσ(x) that would be impossible with the standard saddle-point method.

6 Conclusions

We have given asymptotic expansions as x → ±∞ for the auxiliary Wright functions Fσ(x) and Mσ(x) defined in (1.3) and (1.4) when 0 < σ < 1. These expansions consist of series of an exponential and algebraic character whose relative dominance depends on the parameter σ. An algorithm for determining the coefficients in the exponential expansion is discussed and explicit representation of the first few coefficients has been given.

Numerical results are presented to confirm the accuracy of the expansions. Of particular interest is the the limit σ → 1, where the function Mσ(x) approaches a Dirac delta function centered on x = 1. Graphical results based on the Kreis-Pipkin method are given that illustrate the leading asymptotic forms and the transition of Mσ(x) to a delta function.

Appendix A

Appendix A An algorithm for the computation of the coefficients cj(σ)

In this Appendix we describe an algorithm for the computation of the coefficients cj(σ) appearing in the exponential expansion of the function (z) in (3.1). A full account of this procedure is given in [10, Appendix A], where it is shown that the cj(σ) result from the inverse factorial expansion of the ratio of gamma functions Γ(σs + σ)/Γ(1 + s) for large |s|. This inverse factorial expansion takes the form

Figure 5 Comparison of the three different methods for the computation of Mσ(x) in linear (up) and semi-logarithmic (down) scale, for ϵ = 0.1
Figure 5

Comparison of the three different methods for the computation of Mσ(x) in linear (up) and semi-logarithmic (down) scale, for ϵ = 0.1

Figure 6 Comparison of the three different methods for the computation of Mσ(x) in linear (up) and semi-logarithmic (down) scale, for ϵ = 0.01
Figure 6

Comparison of the three different methods for the computation of Mσ(x) in linear (up) and semi-logarithmic (down) scale, for ϵ = 0.01

(A.1)Γ(σs+σ)Γ(κs+ϑ)Γ(1+s)=κA0(σ)(hκκ)s{j=0M1cj(σ)(κs+ϑ)j+O(1)(κs+ϑ)M}

for |s| → ∞ uniformly in |arg s| ≤ πϵ, where the parameters κ, h, ϑ, A0(σ) are defined in (2.1), with ϑ′ = 1 − ϑ.

Introduction of the scaled Gamma function Γ*(z)=Γ(z)(2π)12ezz12z leads to the representation

Γ(αs+a)=(2π)12eαs(αs)αs+a12e(αs;a)Γ*(αs+a),

where

e(αs;a):=ea(1+aαs)αs+a12=exp[(αs+a12)log(1+aαs)a].

Then, after some routine algebra we find that the left-hand side of (A.1) can be written as

(A.2)Γ(σs+σ)Γ(κs+ϑ)Γ(1+s)=κA0(hκκ)sR(s)ϒ(s),

where

ϒ(s):=Γ*(σs+σ)Γ*(κs+ϑ)Γ*(1+s),R(s):=e(σs;σ)e(κs;ϑ)e(s;1).
Figure 7 Comparison of the three different methods for the computation of Mσ(x) in linear (up) and semi-logarithmic (down) scale, for ϵ = 0.001
Figure 7

Comparison of the three different methods for the computation of Mσ(x) in linear (up) and semi-logarithmic (down) scale, for ϵ = 0.001

Substitution of (A.2) in (A.1) then yields the inverse factorial expansion in the alternative form

(A.3)R(s)ϒ(s)=j=0M1cj(σ)(κs+ϑ)j+O(1)(κs+ϑ)M

as |s| → ∞ in |arg s| ≤ πϵ.

We now expand R(s) and ϒ(s) for s → +∞ making use of the well-known expansion (see, for example, [12, p. 71])

Γ*(z)~k=0()kγkzk  (|z|;|argz|πϵ),

where γk are the Stirling coefficients with γ0 = 1, γ1=112, γ2=1288, γ3=13951840,. Then we find

Γ*(αs+a)=1γ1αs+O(s2),  e(αs;a)=1+a(a1)2αs+O(s2),

whence

R(s)=1+A2s+O(s2),  ϒ(s)=1+12s+O(s2),

where we have defined the quantities A and by

A=σ1ϑκ(1ϑ),=1σ+σκ.

Upon equating coefficients of s−1 in (A.3) we then obtain

(A.4)c1(σ)=12κ(A+16)=124σ(2σ)(12σ).

The higher coefficients are obtained by continuation of this expansion process in inverse powers of s. We write the product on the left-hand side of (A.3) as an expansion in inverse powers of κs in the form

R(s)ϒ(s)=1+j=1M1Cj(κs)j+O(sM)

as s → +∞, where the coefficients Cj are determined with the aid of Mathematica; see [10, Appendix A] for details. The coefficients cj(σ) are then obtained by a recursive process to yield the expressions given in (2.4). This procedure is found to work well in specific cases when the various parameters have numerical values, where up to a maximum of 100 coefficients have been so calculated.

Acknowledgements

The research activity of AC and FM has been carried out in the framework of the activities of the National Group of Mathematical Physics (GNFM, INdAM). The activity of AC, PhD student at the University of Würzburg, is carried out also in the Würzburg-Dresden Cluster of Excellence - Complexity and Topology in Quantum Matter (ct.qmat).

References

[1] D. O. Cahoy, (2012) Estimation and simulation for the M-Wright function. Communications in Statistics - Theory and Methods41, No 8 (2012), 1466–1477; DOI:10.1080/03610926.2010.543299.Search in Google Scholar

[2] A. Kreis and A.C. Pipkin, : Viscoelastic pulse propagation and stable probability distributions. Quart. Appl. Math. 44 (1986), 353–360.10.1090/qam/856190Search in Google Scholar

[3] V. Lipnevich and Yu. Luchko, The Wright function: Its properties, applications, and numerical evaluation. AIP Conference Proceedings1301 (2010), Art. 614; (2010) DOI: 10.1063/1.3526663.Search in Google Scholar

[4] F. Mainardi, Fractional Calculus and Waves in Linear Viscoelasticity. Imperial College Press, London, 2010.10.1142/p614Search in Google Scholar

[5] F. Mainardi and A. Consiglio, The Wright function of the second kind in mathematical physics. In: “SI on Special Functions with Applications in Mathematical Physics”, Mathematics8, No 6 (2020), Art. 884.10.3390/math8060884Search in Google Scholar

[6] F. Mainardi and M. Tomirotti, On a special function arising in the time fractional diffusion-wave equation. In: P. Rusev, I. Dimovski and V. Kiryakova (Eds), Transform Methods and Special Functions, Sofia 1994, Science Culture Technology, Singapore (1995), 171–183. [Proc. Int. Workshop Transform Methods and Special Functions, Sofia 12–17 August 1994].Search in Google Scholar

[7] F. Mainardi and M. Tomirotti, Seismic pulse propagation with constant Q and stable probability distributions. Annali di Geofisica40 (1997), 1311–1328; E-print: http://arxiv.org/abs/1008.1341.10.4401/ag-3863Search in Google Scholar

[8] F.W.J. Olver, D.W. Lozier, R.F. Boisvert and C.W. Clark (Editors), NIST Handbook of Mathematical Functions. Cambridge University Press, Cambridge, 2010.Search in Google Scholar

[9] R.B. Paris, Exponentially small expansions of the Wright function on the Stokes lines. Lithuanian Math. J. 54 (2014), 82–105.10.1007/s10986-014-9229-9Search in Google Scholar

[10] R.B. Paris, The asymptotics of the generalised Bessel function. Math. Aeterna7 (2017), 381–406.Search in Google Scholar

[11] R.B. Paris, Asymptotics of the special functions of fractional calculus. In: A. Kochubei, Y. Luchko (Eds). Handbook of Fractional Calculus with Applications, Vol. 1, 297–325, De Gruyter, Berlin, 2019.10.1515/9783110571622-012Search in Google Scholar

[12] R.B. Paris and D. Kaminski, Asymptotics and Mellin-Barnes Integrals. Cambridge University Press, Cambridge, 2001.10.1017/CBO9780511546662Search in Google Scholar

[13] R.B. Paris and V. Vinogradov, Asymptotic and structural properties of the Wright function arising in probability theory. Lithuanian Math. J. 56 (2016), 377–409.10.1007/s10986-016-9324-1Search in Google Scholar

[14] E.M. Wright, The asymptotic expansion of the generalized Bessel function. Proc. Lond. Math. Soc. (Ser. 2)38 (1934), 286–293.10.1112/plms/s2-38.1.257Search in Google Scholar

[15] E.M. Wright, The generalized Bessel function of order greater than one. Quart. J. Math. 11 (1940), 36–48.10.1093/qmath/os-11.1.36Search in Google Scholar

Received: 2020-08-28
Published Online: 2021-01-29
Published in Print: 2021-02-23

© 2021 Diogenes Co., Sofia

Downloaded on 24.4.2024 from https://www.degruyter.com/document/doi/10.1515/fca-2021-0003/html
Scroll to top button