Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Time-fractional telegraph equation of distributed order in higher dimensions with Hilfer fractional derivatives


  • In this paper, we consider the time-fractional telegraph equation of distributed order in higher spatial dimensions, where the time derivative is in the sense of Hilfer, thus interpolating between the Riemann-Liouville and the Caputo fractional derivatives. By employing the techniques of the Fourier, Laplace, and Mellin transforms, we obtain a representation of the solution of the Cauchy problem associated with the equation in terms of convolutions involving functions that are Laplace integrals of Fox H-functions. Fractional moments of the first fundamental solution are computed and for the special case of double-order distributed it is analyzed in detail the asymptotic behavior of the second-order moment, by application of the Tauberian Theorem. Finally, we exhibit plots of the variance showing its behavior for short and long times, and for different choices of the parameters along small dimensions.

    Citation: Nelson Vieira, M. Manuela Rodrigues, Milton Ferreira. Time-fractional telegraph equation of distributed order in higher dimensions with Hilfer fractional derivatives[J]. Electronic Research Archive, 2022, 30(10): 3595-3631. doi: 10.3934/era.2022184

    Related Papers:

    [1] Dewang Li, Meilan Qiu, Jianming Jiang, Shuiping Yang . The application of an optimized fractional order accumulated grey model with variable parameters in the total energy consumption of Jiangsu Province and the consumption level of Chinese residents. Electronic Research Archive, 2022, 30(3): 798-812. doi: 10.3934/era.2022042
    [2] María Guadalupe Morales, Zuzana Došlá, Francisco J. Mendoza . Riemann-Liouville derivative over the space of integrable distributions. Electronic Research Archive, 2020, 28(2): 567-587. doi: 10.3934/era.2020030
    [3] Li Tian, Ziqiang Wang, Junying Cao . A high-order numerical scheme for right Caputo fractional differential equations with uniform accuracy. Electronic Research Archive, 2022, 30(10): 3825-3854. doi: 10.3934/era.2022195
    [4] Ping Zhou, Hossein Jafari, Roghayeh M. Ganji, Sonali M. Narsale . Numerical study for a class of time fractional diffusion equations using operational matrices based on Hosoya polynomial. Electronic Research Archive, 2023, 31(8): 4530-4548. doi: 10.3934/era.2023231
    [5] Jin Li, Yongling Cheng . Barycentric rational interpolation method for solving time-dependent fractional convection-diffusion equation. Electronic Research Archive, 2023, 31(7): 4034-4056. doi: 10.3934/era.2023205
    [6] Xinguang Zhang, Yongsheng Jiang, Lishuang Li, Yonghong Wu, Benchawan Wiwatanapataphee . Multiple positive solutions for a singular tempered fractional equation with lower order tempered fractional derivative. Electronic Research Archive, 2024, 32(3): 1998-2015. doi: 10.3934/era.2024091
    [7] Ziqing Yang, Ruiping Niu, Miaomiao Chen, Hongen Jia, Shengli Li . Adaptive fractional physical information neural network based on PQI scheme for solving time-fractional partial differential equations. Electronic Research Archive, 2024, 32(4): 2699-2727. doi: 10.3934/era.2024122
    [8] Mehmet Ali Özarslan, Ceren Ustaoğlu . Extended incomplete Riemann-Liouville fractional integral operators and related special functions. Electronic Research Archive, 2022, 30(5): 1723-1747. doi: 10.3934/era.2022087
    [9] Yihui Xu, Benoumran Telli, Mohammed Said Souid, Sina Etemad, Jiafa Xu, Shahram Rezapour . Stability on a boundary problem with RL-Fractional derivative in the sense of Atangana-Baleanu of variable-order. Electronic Research Archive, 2024, 32(1): 134-159. doi: 10.3934/era.2024007
    [10] Jin Li, Yongling Cheng . Barycentric rational interpolation method for solving fractional cable equation. Electronic Research Archive, 2023, 31(6): 3649-3665. doi: 10.3934/era.2023185
  • In this paper, we consider the time-fractional telegraph equation of distributed order in higher spatial dimensions, where the time derivative is in the sense of Hilfer, thus interpolating between the Riemann-Liouville and the Caputo fractional derivatives. By employing the techniques of the Fourier, Laplace, and Mellin transforms, we obtain a representation of the solution of the Cauchy problem associated with the equation in terms of convolutions involving functions that are Laplace integrals of Fox H-functions. Fractional moments of the first fundamental solution are computed and for the special case of double-order distributed it is analyzed in detail the asymptotic behavior of the second-order moment, by application of the Tauberian Theorem. Finally, we exhibit plots of the variance showing its behavior for short and long times, and for different choices of the parameters along small dimensions.



    Distributed-order fractional calculus (DOFC) is a branch of fractional calculus important for the modeling of complex systems. It generalizes the constant fractional operators by integrating the fractional kernel of these operators over an extended range of orders. The fractional differential operator of distributed order, for orders not great than 2, is given by

    Dα=ulb(α)dαdtαdα,0l<u2,b(α)0

    where dαdtα stands for a single-order fractional derivative and b(α) is a non-negative weight function or generalized function. DOFC takes into account the superposition of orders and is a useful tool for modeling decelerating anomalous diffusion, ultraslow diffusive processes and strong anomaly (see e.g., [1,2]). DOFC models systems whose behavior stems from the complex interplay and superposition of nonlocal and memory effects occurring over a multitude of scales [3]. Over the last two decades, a significant number of papers appeared focusing on mathematical aspects and real-world applications of fractional partial differential equations with distributed order, see e.g., [2,4,5,6,7,8,9,10,11,12,13,14,15] for mathematical aspects, [16,17] for applications, and the review paper [18] about the mathematics of DOFC, including analytical, numerical methods, and the extensive overview of the recent applications of DOFC to fields like transport processes, and control theory. Moreover, DOFC was applied also in the study of composite materials [19,20] and viscoelastic materials having spatially varying properties [21].

    The classical telegraph equation was first derived by Lord Kelvin in the 19th century [22]. It is a hyperbolic partial differential equation of the form

    c22ttu(x,t)+c1tu(x,t)c202xxu(x,t)+du(x,t)=q(x,t),xR,t>0.

    This equation was proposed by Cattaneo in 1958 (see [23]) to overcome the problem of infinite propagation velocity in heat transmission. Over the years, this equation and its time-fractional versions appeared in the study of several phenomena such as transmission lines for all frequencies [24], random walks [25], solar particle transport [26], oceanic diffusion [27], wave propagation [28], damped small vibrations, anomalous diffusion and wave-like processes [29,30,31,32], scalar part of the Maxwell equations.

    The TFTE with time-fractional derivatives of orders α1]0,1] and α2]1,2] was studied from the analytical, numerical, and probabilistic points of view by several authors. In [33], Cascaval et al. discussed the well-posedness of some initial-boundary value problems for the TFTE as well as the asymptotic behaviour of their solutions. In [31], the authors studied the neutral case of the TFTE and obtained an explicit Fourier representation of the fundamental solution (FS) and made a probabilistic interpretation of the FS in terms of stable probability density functions. Particular attention was given to the case α1=1/2 and α2=1 due to its connection of the telegraph process with Brownian motion. Some of these results were generalized by Camargo et al. in [34] for general α1 and α2 and studied later by Boyadjiev and Luchko in [29]. In [35], the authors considered a generalized telegraph equation with time-fractional derivatives in the Hilfer and Hadamard senses and space-fractional derivatives are in the sense of Riesz-Feller. Górska et al., (see [36]) considered various types of generalized telegraph equations and determine the conditions under which solutions can be recognized as probability density distributions.

    The works [32,37,38,39,40,41] are examples of works devoted to the study of the TFTE in the multidimensional case with n space variables, where in some cases the second derivative in space is replaced by the Euclidean Laplace operator. In [32] the authors solved the multi-dimensional TFTE with multi-term time-fractional derivatives and proved that its fundamental solution is the law of a stable isotropic multi-dimensional process time-changed. Ovidio et al., [41] constructed compositions of vector processes whose distribution is related to space-time fractional n-dimensional telegraph equations. We refer also the works of Masoliver and his co-workers about the TFTE and its connections with random walks (see [42,43,44,45]), and the recent survey paper [30] where is presented a very complete review of the fractional telegraph process. In [38,40] were employed Fourier, Laplace and Mellin transform techniques to obtain the first and second FS. Moreover, the application of the Residue Theorem allows obtaining double series representation for the FS of the TFTE in higher dimensions. Connections of the TFTE with fractional Clifford analysis and Sturm-Liouville theory were presented in [39] and [37].

    In our recent paper [46] we studied the time-fractional telegraph equation with generalized distributed order in Rn×R+ finding a representation of the fundamental solution in terms of convolutions involving Fox H-functions. In this work, we extend our analysis to time-fractional telegraph equations of distributed order with Hilfer (or composite) time-fractional derivatives. Hilfer's derivative tDγ,ν0+ was defined by Hilfer as a two-parameter family of fractional derivatives of order γ>0 and type ν[0,1] given by

    (tDγ,ν0+f)(t)=(Iν(mγ)0+ddt(I(1ν)(mγ)0+f))(t),

    where Iγ0+ denotes the left Riemann-Liouville fractional integral of order γ>0 (see Eq (2.1) in Section 2). The Hilfer fractional derivative allows to interpolate smoothly between the Riemann-Liouville and the Caputo fractional derivatives (see [16,47,48]). These special cases are obtained when ν=0 and ν=1, respectively. The type-parameter produces more stationary states, provides an extra degree of freedom on the initial condition, and increases the flexibility for the description of complex data. It was first used by Hilfer to describe the dynamics in glass formers over an extremely large-frequency window [49]. During the last years fractional differential equations with composite derivatives were studied by several authors, see e.g., [35,50,51,52,53,54,55,56]. In these works the equations are considered in R×R+, i.e., one single space variable and one time variable. Here, we consider the telegraph equation of distributed order with Hilfer time-fractional derivatives in the higher dimensional case, i.e., Rn×R+.

    The paper is organized as follows. In Section 2 we recall some basic facts about fractional derivatives, integral transforms, and special functions, which are necessary for the development of this work. In Section 3 we formulate the problem of generalized distributed order telegraph equation for general density functions. Following the ideas presented in [46], we use a combination of Laplace, Fourier and Mellin transforms to obtain a representation of the solution of our equation via convolution integrals involving Fox H-functions. The key points to obtain our main result are the use of the classical Titchmarsh's Theorem to invert the Laplace transform, and the use of the Mellin transform to invert the Fourier transform. Some particular cases of our equation are analyzed by considering specific choices of the parameters of the equation. In Section 4 we compute the expression of the fractional moments of arbitrary order of the first fundamental solution in the Laplace domain. For the particular case of single-order derivatives, we invert the Laplace transform of the second-order moment obtaining an expression in terms of the three-parameter Mittag-Leffler function. For numerical purposes, we study the corresponding asymptotic behavior of the second-order moment in the time domain for the long and short time limit, by using the Tauberian Theorem. In the final part of the paper, we present and analyze some plots of the second-order moment for this particular case. As it will be shown the graphical representations support the analytical conclusions obtained via the Tauberian Theorem.

    Let a,bR with a<b and α>0. The left Riemann-Liouville fractional integral Iγa+ of order γ>0 is given by (see [57])

    (Iγa+f)(x)=1Γ(γ)xaf(w)(xw)1γdw,x>a.

    The left Hilfer (or composite) fractional derivative tDγ,ν0+ of order γ>0 and type 0ν1 is given by (see [16,47,48])

    (tDγ,ν0+f)(t)=(Iν(mγ)0+ddt(I(1ν)(mγ)0+f))(t), (2.1)

    where m=[γ]+1 and [γ] means the integer part of γ. We observe that in the case when ν=0 we recover the left Riemann-Liouville fractional derivative and in the case when ν=1 we have the left Caputo fractional derivative. The previous definitions of fractional integrals and derivatives can be naturally extended to Rn considering partial fractional integrals and derivatives (see Chapter 5 in [58]).

    In this work, some integral transforms are used, namely, the Laplace, the Fourier and the Mellin transforms. The Laplace transform of a real-valued function f(t) is defined by (see [57])

    L{f(t)}(s)=˜f(s)=+0estf(t)dt,Re(s)C

    and when it is applied to Eq (2.1) leads to (see [50])

    L{tDγ,ν0+f(t)}(s)=sγ˜f(s)m1j=0smjν(mγ)1[djdtj(tI(1ν)(mγ)0+f)](0+), (2.2)

    where the initial-value terms [dkdtk(tI(1ν)(mγ)0+f)](0+) are evaluated in the limit t0+. Concerning the inverse Laplace transform of functions involving a branch point, we have the following theorem from Titchmarsh (see [59]).

    Theorem 2.1. Let ˜f(s) be an analytic function which has a branch cut on the real negative semiaxis, which has the following properties

    ˜f(s)=O(1),|s|+,˜f(s)=O(1|s|),|s|0,

    for any sector |arg(s)|<πη, where 0<η<π. Then the inverse Laplace transform of ˜f(s) is given by

    f(t)=L1{˜f(s)}(t)=1π+0ertIm(˜f(reiπ))dr,

    where Im() denotes the imaginary part.

    The convolution of two integrable functions f and g with support in [0,+) is defined by

    (ftg)(t)=t0f(tw)g(w)dw,tR+ (2.3)

    and the Convolution Theorem for the Laplace transform is given by

    L{(ftg)(t)}(s)=L{f}(s)L{g}(s). (2.4)

    The n-dimensional Fourier transform of a real-valued integrable function f in Rn is defined by (see [57])

    F{f(x)}(κ)=ˆf(κ)=Rneiκxf(x)dx,κRn,

    while the corresponding inverse Fourier transform is given formally by

    f(x)=F1{ˆf(κ)}(x)=1(2π)nRneixκˆf(κ)dκ,xRn. (2.5)

    The convolution operator of two functions in Rn is defined by the integral

    (fxg)(x)=Rnf(xz)g(z)dz,xRn (2.6)

    and the Convolution Theorem for the Fourier transform is given by

    F{(fxg)(x)}(κ)=F{f}(κ)F{g}(κ). (2.7)

    For the n-dimensional Laplace operator Δx=ni=12x2i we have (see formula (1.3.32) in [57])

    F{Δxf(x)}(κ)=|κ|2F{f(x)}(κ). (2.8)

    Another integral transform that we use in this work is the Mellin transform. For f locally integrable on ]0,+[ it is defined by (see [57])

    M{f(w)}(s)=f(s)=+0ws1f(w)dw,sC, (2.9)

    and the inverse Mellin transform is given by

    f(w)=M1{f(s)}(w)=12πiγ+iγiwsf(s)ds,w>0,γ=Re(s). (2.10)

    The condition for the existence of Eq (2.9) is that p<γ<q (called the fundamental strip), where p, q are the order of f at the origin and , respectively. The integration in Eq (2.10) is performed along the imaginary axis and the result does not depend on the choice of γ inside the fundamental strip. For more information about this transform and its properties, see e.g., [57]. The Mellin convolution between two functions is defined by

    (fMg)(x)=+0f(xu)g(u)duu, (2.11)

    and satisfies the Mellin Convolution Theorem (see formula (1.4.40) in [57])

    M{fMg}(s)=M{f}(s)M{g}(s).

    The following relation holds (see (1.4.30) in [57])

    M{f(1x)}(s)=M{f}(s). (2.12)

    The solution of the time-fractional telegraph equation of distributed order obtained in this work involves the Fox H-function Hm,np,q, which is defined, via a Mellin-Barnes type integral, by (see [60])

    Hm,np,q[z(a1,α1),,(ap,αp)(b1,β1),,(bq,βq)]=12πiCmj=1Γ(bj+βjs)ni=1Γ(1aiαis)pi=n+1Γ(ai+αis)qj=m+1Γ(1bjβjs)zsds, (2.13)

    where ai,bjC, and αi,βjR+, for i=1,,p and j=1,,q and C is a suitable contour in the complex plane separating the poles of the two factors in the numerator (see [60]). The expression of the second-order moment in Section 4.1 is presented in terms of the three parameter Mittag-Leffler function Eβ3β1,β2(z) (see [61]), which is defined, in terms of power series, by

    Eγα,β(z)=k=0(γ)kzkk!Γ(αk+β),zC,α,β,γR,α>0, (2.14)

    where (γ)k is the Pochhammer symbol.

    Throughout the paper, we assume that all the involved functions are Laplace and Fourier transformable.

    In this work we consider the following generalized time-fractional telegraph equation of distributed order

    1021b2(β,ν)tβ,ν0+u(x,t)dβdν +a1010b1(α,μ)tα,μ0+u(x,t)dαdμc2Δxu(x,t)+d2u(x,t)=q(x,t), (3.1)

    for given weight functions b2(β,ν)>0 and b1(α,μ)>0, satisfying

    1021b2(β,ν)dβdν=C2,1010b1(α,μ)dαdμ=C1, (3.2)

    and subject to the following initial and boundary conditions

    (tI(1μ)(1α)0+u)(x,0+)=f(x),(tI(1ν)(2β)0+u)(x,0+)=g1(x) (3.3)
    [0.2cm][t(tI(1ν)(2β)0+u)](x,0+)=g2(x),lim|x|+u(x,t)=0, (3.4)

    where (x,t)Rn×R+, Δx is the classical Laplace operator in Rn, the partial time-fractional derivatives of order β]1,2] and α]0,1], and types μ,ν[0,1] are in the Hilfer sense and given by Eq (2.1), aR+0, cR{0}, dR, and C1,C2R+. The positive constants C1 and C2 can be taken as 1 if we assume the normalization condition for the integrals Eq (3.2). Moreover, q belongs to L1(Rn×I), and f,g1,g2L1(Rn). We look for solutions u(x,t) of our problem in the space C2(Rn)×C2(0,+) with possible exception at x=0.

    In order to analytically determine the solution of Eqs (3.1) and (3.2) in the space-time domain we start applying the Fourier and Laplace transforms to Eq (3.1) and solve the equation in the Fourier-Laplace domain. After that, there are two alternative strategies related to the order in carrying out the inversions of the Fourier and Laplace transforms are performed (see [2]):

    (S1) invert the Fourier transform, yielding ˜u(x,s), and then invert the Laplace transform of the result.

    (S2) invert the Laplace transform, yielding ˆu(κ,t), and then invert the Fourier transform of the result.

    In this work, we consider the strategy (S2) where the inversion of the Laplace transform is performed via the classical Titchmarsh's Theorem, and the inversion of the Fourier transform is performed via the Mellin transform.

    we start by applying in Eq (3.1) the Laplace transform with respect to the variable tR+ and the n-dimensional Fourier transform with respect to the variable xRn. Taking into account relations Eq (2.2) and Eq (2.8), and the initial conditions in Eqs (3.3) and (3.4), we obtain

    ˆ˜u(κ,s)1021b2(β,ν)sβdβdνˆg1(κ)1021b2(β,ν)s1ν(2β)dβdνˆg2(κ)1021b2(β,ν)sν(2β)dβdν+aˆ˜u(κ,s)1010b1(α,μ)sαdαdμaˆf(κ)1010b1(α,μ)sμ(1α)dαdμ+c2|κ|2ˆ˜u(κ,s)+d2ˆ˜u(κ,s)=ˆ˜q(κ,s),

    which is equivalent to

    ˆ˜u(κ,s)=ˆf(κ)B1(s)B2(s)+B1(s)+|κ|2+ˆg1(κ)(B2(s)d2c2)s1(B2(s)+B1(s)+|κ|2)+ˆg2(κ)(B2(s)d2c2)B2(s)+B1(s)+|κ|2+ˆ˜q(κ,s)c2(B2(s)+B1(s)+|κ|2), (3.5)

    where ˆf and ˆgi are the Fourier transforms of the functions f and gi, respectively, and

    B2(s)=1c2(1021b2(β,ν)sβdβdν+d2),B1(s)=ac21010b1(α,μ)sαdαdμ, (3.6)
    B2(s)=1c2(1021b2(β,ν)sν(2β)dβdν+d2),B1(s)=ac21010b1(α,μ)sμ(1α)dαdμ. (3.7)

    Remark 3.1. When μ=ν=1, i.e., the telegraph equation has only Caputo fractional derivatives, we have the following relations between Eqs (3.6) and (3.7)

    B1(s)=s1B1(s) and B2(s)d2c2=s2(B2(s)d2c2).

    The previous relations combined with the fact that Eqs (3.3) and (3.4) reduce to

    u(x,0)=f(x),u(x,0)=g1(x),ut(x,0)=g2(x),

    i.e., f=g1 when ν=μ=1, allow us to reduce the expression (3.5) to the correspondent one obtained in [46].

    In this section we perform the inversion of the Laplace and Fourier transforms in order to obtain our solution in the space-time domain. Let us consider the following auxiliary functions in the Laplace domain

    ˆ˜u1(κ,s)=B1(s)B2(s)+B1(s)+|κ|2, (3.8)
    [0.2cm]ˆ˜u2(κ,s)=B2(s)sp(B2(s)+B1(s)+|κ|2), (3.9)
    [0.2cm]ˆ˜u3(κ,s)=1sp(B2(s)+B1(s)+|κ|2), (3.10)

    with p=0 or p=1. To further proceed we make the following additional assumption:

    (H1): The functions ˆ˜uj(κ,reiπ),j=1,2,3 are in the conditions of Theorem 2.1. (3.11)

    Assumption (H1) holds for the particular cases we consider later on. Applying Theorem 2.1, we have

    ˆuj(κ,t)=1π+0ertIm(ˆ˜uj(κ,reiπ))dr,j=1,2,3. (3.12)

    To evaluate the imaginary parts of the functions ˆ˜uj(κ,reiπ), j=1,2,3, along the ray s=reiπ, with r>0, we consider the following polar decompositions

    Bl(reiπ)=ρl(cos(γlπ)+isin(γlπ)){ρl=|Bl(reiπ)|γl=1πarg(Bl(reiπ)),l=1,2 (3.13)
    [0.2cm]Bl(reiπ)=ρl(cos(γlπ)+isin(γlπ)){ρl=|Bl(reiπ)|γl=1πarg(Bl(reiπ)),l=1,2. (3.14)

    After straightforward calculations, we obtain the following expressions

    Im{ˆ˜u1(κ,reiπ)}=K1(|κ|,r)=ρ1[(A+|κ|2)sin(γ1π)Bcos(γ1π)][(A+|κ|2)2+B2], (3.15)
    [0.2cm]Im{ˆ˜u2(κ,reiπ)}=K2(p,|κ|,r)=ρ2[(A+|κ|2)sin(γ2π)Bcos(γ2π)](r)p[(A+|κ|2)2+B2], (3.16)
    [0.2cm]Im{ˆ˜u3(κ,reiπ)}=K3(p,|κ|,r)=B(r)p[(A+|κ|2)2+B2], (3.17)

    where

    A=ρ2cos(γ2π)+ρ1cos(γ1π) and B=ρ2sin(γ2π)+ρ1sin(γ1π). (3.18)

    Remark 3.2. Taking into account Remark 3.1, when μ=ν=1 we have, by straightforward calculations, the following relations

    {ρ2=r2ρ2γ2=γ2 and {ρ1=r1ρ1γ1=1+γ1.

    Applying the inverse Laplace transform to Eq (3.5) and taking into account Eqs (3.12), (3.15), (3.16) and (3.17), we obtain

    ˆu(κ,t)=ˆf(κ)π+0ertK1(|κ|,r)drˆg1(κ)π+0ert[K2(1,|κ|,r)d2c2K3(1,|κ|,r)]drˆg2(κ)π+0ert[K2(0,|κ|,r)d2c2K3(0,|κ|,r)]drˆq(κ,t)πc2t+0ertK3(0,|κ|,r)dr, (3.19)

    where t is given by Eq (2.3) and in the last term me made use of Eq (2.4). For the inversion of the Fourier transform, taking into account Eqs (2.5) and (2.7), we obtain

    u(x,t)=f(x)xF1{1π+0ertK1(|κ|,r)dr}(x,t)g1(x)xF1{1π+0ert[K2(1,|κ|,r)d2c2K3(1,|κ|,r)]dr}(x,t)g2(x)xF1{1π+0ert[K2(0,|κ|,r)d2c2K3(0,|κ|,r)]dr}(x,t)q(x,t)txF1{1πc2+0ertK3(0,|κ|,r)dr}(x). (3.20)

    Using the following formula presented in [58] for the inverse Fourier transform of L1-functions

    1(2π)nRneixκφ(|κ|)dκ=|x|1n2(2π)n2+0φ(w)wn2Jn21(|x|w)dw, (3.21)

    and since we are dealing with radial functions in κ, Eq (3.20) can be rewritten as

    u(x,t)=1πf(x)x[|x|1n2(2π)n2+0+0ertK1(w,r)drwn2Jn21(|x|w)dw]1πg1(x)x[|x|1n2(2π)n2+0+0ert[K2(1,w,r)d2c2K3(1,w,r)]drwn2Jn21(|x|w)dw]1πg2(x)x[|x|1n2(2π)n2+0+0ert[K2(0,w,r)d2c2K3(0,w,r)]drwn2Jn21(|x|w)dw]1πc2q(x,t)tx[|x|1n2(2π)n2+0+0ertK3(0,w,r)drwn2Jn21(|x|w)dw]=1πf(x)x[+0ert|x|1n2(2π)n2+0K1(w,r)wn2Jn21(|x|w)dwI1dr]1πg1(x)x[+0ert|x|1n2(2π)n2+0[K2(1,w,r)d2c2K3(1,w,r)]wn2Jn21(|x|w)dwI2dr]1πg2(x)x[+0ert|x|1n2(2π)n2+0[K2(0,w,r)d2c2K3(0,w,r)]wn2Jn21(|x|w)dwI3dr]1πc2q(x,t)tx[+0ert|x|1n2(2π)n2+0K3(0,w,r)wn2Jn21(|x|w)dwI4dr]. (3.22)

    To compute explicitly I1, I2, I3, and I4 in Eq (3.22) we are going to use the Mellin transform. First, we rewrite these integrals as a Mellin convolution Eq (2.11). In fact, considering the following auxiliary functions

    g1(w)=K1(w,r),g2(w)=K2(1,w,r)d2c2K3(1,w,r),g3(w)=K2(0,w,r)d2c2K3(0,w,r),g4(w)=K3(0,w,r),f(w)=1(2π)n2|x|nwn2+1Jn21(1w),

    we have for i=1,2,3,4

    Ii=(giMf)(1|x|)=+0gi(w)f(1|x|w)dww=+0gi(w)wn2+1|x|n2+1(2π)n2|x|nJn21(|x|w)dww=|x|1n2(2π)n2+0gi(w)wn2Jn21(|x|w)dw. (3.23)

    From the relations Eqs (2.12) and (2.11), we have for i=1,2,3,4

    M{Ii}(s)=M{(giMf)(1|x|)}(s)=M{gi}(s)M{f}(s)

    which is equivalent to

    M{Ii}(s)=M{gi}(s)M{f}(s),i=1,2,3,4. (3.24)

    Now, we compute the Mellin transforms that appear in (3.24). The Mellin transform of the function f was already calculated in [46] (see formula (43)):

    M{f}(s)=1πn12|x|n2n1Γ(ns)Γ(n+1s2)Γ(s2). (3.25)

    To compute the Mellin transform of the function g1, we take into account Eqs (2.9), (3.15) and (3.18), obtaining

    M{g1}(s)=+0ws1K1(w,r)dw=+0ws1ρ1[(A+w2)sin(γ1π)Bcos(γ1π)](A+w2)2+B2dw=ρ1sin(γ1π)+0ws+1(A+w2)2+B2dw+ρ1[Asin(γ1π)Bcos(γ1π)]+0ws1(A+w2)2+B2dw. (3.26)

    Considering the change of variables w2=z in Eq (3.26) we obtain

    M{g1}(s)=ρ1sin(γ1π)+0zs2z2+2Az+A2+B2dzI5+ρ1[Asin(γ1π)Bcos(γ1π)]+0zs21z2+2Az+A2+B2dzI6. (3.27)

    Integrals I5 and I6 were already calculated in [46] (see formulas (49) and (50)). Therefore, we have that

    I5=πsin(ψ)Γ(1+s2)Γ(1(1+s2))Γ(sψ2π)Γ(1sψ2π)(A2+B2)s412 (3.28)

    and

    I6=πsin(ψ)Γ(s2)Γ(1s2)Γ(ψπ(s21))Γ(1ψπ(s21))(A2+B2)s41, (3.29)

    where

    ψ=arccos(AA2+B2). (3.30)

    Hence, from Eqs (3.28) and (3.29) we conclude that Eq (3.27) takes the form

    M{g1}(s)=πρ1sin(γ1π)2sin(ψ)Γ(1+s2)Γ(1(1+s2))Γ(sψ2π)Γ(1sψ2π)(A2+B2)s412πρ1[Asin(γ1π)Bcos(γ1π)]2sin(ψ)Γ(s2)Γ(1s2)Γ(ψπ(s21))Γ(1ψπ(s21))(A2+B2)s41. (3.31)

    Now, we calculate the Mellin transform of g2. Taking into account Eqs (2.9), (3.16)–(3.18) and (3.30) we get

    M{g2}(s)=+0ws1K2(1,w,r)dwd2c2+0ws1K3(1,w,r)dw=+0ws1ρ2[(A+w2)sin(γ2π)Bcos(γ2π)](r)1[(A+w2)2+B2]dwd2c2+0ws1B(r)1[(A+w2)2+B2]dw (3.32)
    [0.2cm]=rρ2sin(γ2π)+0ws+1(A+w2)2+B2dwrρ2[Asin(γ2π)Bcos(γ2π)]+0ws1(A+w2)2+B2dwrd2Bc2+0ws1(A+w2)2+B2dw=rρ2sin(γ2π)+0ws+1(A+w2)2+B2dwrρ2c2[Asin(γ2π)Bcos(γ2π)]+rd2Bc2+0ws1(A+w2)2+B2dw. (3.33)

    Considering the change of variables w2=z in Eq (3.33), we obtain

    M{g2}(s)=rρ2sin(γ2π)2+0zs2z2+2Az+A2+B2dzrρ2c2[Asin(γ2π)Bcos(γ2π)]+rd2B2c2+0zs21z2+2Az+A2+B2dz. (3.34)

    The two integrals in Eq (3.34) correspond to I5 and I6. Hence, from Eqs (3.28) and (3.29) we arrive to

    M{g2}(s)=πrρ2sin(γ2π)2sin(ψ)Γ(1+s2)Γ(1(1+s2))Γ(sψ2π)Γ(1sψ2π)(A2+B2)s412+πr[ρ2c2[Asin(γ2π)Bcos(γ2π)]+d2B]2c2sin(ψ)Γ(s2)Γ(1s2)Γ(ψπ(s21))Γ(1ψπ(s21))(A2+B2)s41. (3.35)

    For the calculation of the Mellin transform of g3 we use Eqs (2.9), (3.16)–(3.18) and (3.30) to get

    M{g3}(s)=+0ws1K2(0,w,r)dwd2c2+0ws1K3(0,w,r)dw=+0ws1ρ2[(A+w2)sin(γ2π)Bcos(γ2π)](r)0[(A+w2)2+B2]dwd2c2+0ws1B(r)0[(A+w2)2+B2]dw. (3.36)

    Expression Eq (3.36) is very similar to Eq (3.32) with a difference in the power of r. However, this exponent does not affect any of the performed calculations in obtaining Eq (3.35). Then we get

    M{g3}(s)=πρ2sin(γ2π)2sin(ψ)Γ(1+s2)Γ(1(1+s2))Γ(sψ2π)Γ(1sψ2π)(A2+B2)s412
    π[ρ2c2[Asin(γ2π)Bcos(γ2π)]+d2B]2c2sin(ψ)Γ(s2)Γ(1s2)Γ(ψπ(s21))Γ(1ψπ(s21))(A2+B2)s41. (3.37)

    Finally, we calculate the Mellin transform of g4. Taking into account Eqs (2.9), (3.17), (3.18) and (3.30) we get

    M{g4}(s)=+0ws1K3(0,w,r)dw=B+0ws1(A+w2)2+B2dw. (3.38)

    Considering the change of variables w2=z in Eq (3.38), we obtain

    M{g4}(s)=B2+0zs21z2+2Az+A2+B2dz. (3.39)

    The integral in Eq (3.39) corresponds to the integral I6. Hence, from Eq (3.29) we arrive to

    M{g4}(s)=Bπ2sin(ψ)Γ(s2)Γ(1s2)Γ(ψπ(s21))Γ(1ψπ(s21))(A2+B2)s41. (3.40)

    Now, using the inverse Mellin transform Eq (2.10) applied to (3.24), we obtain the representation of the integrals I1, I2, I3, and I4 in terms of Mellin-Barnes integrals and, consequently, as Fox H-functions. For the integral I1, taking into account Eqs (2.10), (3.18), (3.31) and (3.25), we obtain

    I1=ρ1sin(γ1π)(A2+B2)12πn32(2|x|)nsin(ψ)12πiγ+iγiΓ(1+s2)Γ(ns)Γ(s2)Γ(s2)Γ(ψs2π)Γ(n+12s2)Γ(1ψs2π)((A2+B2)14|x|)sdsρ1[Asin(γ1π)Bcos(γ1π)](A2+B2)1πn32(2|x|)nsin(ψ)×12πiγ+iγiΓ(ns)Γ(1s2)Γ(ψπ+ψs2π)Γ(n+12s2)Γ(1+ψπψs2π)((A2+B2)14|x|)sds

    which is equivalent, by Eq (2.13), to the following expression in terms of Fox H-functions

    I1=ρ1sin(γ1π)(A2+B2)12πn32(2|x|)nsin(ψ)H1,24,3[(A2+B2)14|x|(1n,1),(1,12),(0,12),(0,ψ2π)(1,12),(1n2,12),(0,ψ2π)]
    ρ1[Asin(γ1π)Bcos(γ1π)](A2+B2)1πn32(2|x|)nsin(ψ)H0,23,2[(A2+B2)14|x|(1n,1),(0,12),(ψπ,ψ2π)(1n2,12),(ψπ,ψ2π)]. (3.41)

    For the integral I2, taking into account Eqs(2.10), (3.18), (3.35) and (3.25), we obtain

    I2=rρ2sin(γ2π)(A2+B2)12πn32(2|x|)nsin(ψ)12πiγ+iγiΓ(1+s2)Γ(ns)Γ(s2)Γ(s2)Γ(ψs2π)Γ(n+12s2)Γ(1ψs2π)((A2+B2)14|x|)sds+r[ρ2c2(Asin(γ2π)Bcos(γ2π))+Bd2](A2+B2)1c2πn32(2|x|)nsin(ψ)×12πiγ+iγiΓ(ns)Γ(1s2)Γ(ψπ+ψs2π)Γ(n+12s2)Γ(1+ψπψs2π)((A2+B2)14|x|)sds

    which is equivalent, by Eq (2.13), to the following expression in terms of Fox H-functions

    I2=rρ2sin(γ2π)(A2+B2)12πn32(2|x|)nsin(ψ)H1,24,3[(A2+B2)14|x|(1n,1),(1,12),(0,12),(0,ψ2π)(1,12),(1n2,12),(0,ψ2π)]+r[ρ2c2(Asin(γ2π)Bcos(γ2π))+Bd2](A2+B2)1c2πn32(2|x|)nsin(ψ)×H0,23,2[(A2+B2)14|x|(1n,1),(0,12),(ψπ,ψ2π)(1n2,12),(ψπ,ψ2π)]. (3.42)

    Due to the similarities between g2 and g3 (and consequently between I2 and I3) we have that

    I3=ρ2sin(γ2π)(A2+B2)12πn32(2|x|)nsin(ψ)H1,24,3[(A2+B2)14|x|(1n,1),(1,12),(0,12),(0,ψ2π)(1,12),(1n2,12),(0,ψ2π)][ρ2c2(Asin(γ2π)Bcos(γ2π))+Bd2](A2+B2)1c2πn32(2|x|)nsin(ψ)×H0,23,2[(A2+B2)14|x|(1n,1),(0,12),(ψπ,ψ2π)(1n2,12),(ψπ,ψ2π)]. (3.43)

    Finally, for the integral I4, taking into account Eqs (2.10), (3.18), (3.40) and (3.25), we obtain

    I4=B(A2+B2)1πn32(2|x|)nsin(ψ)12πiγ+iγiΓ(ns)Γ(1s2)Γ(ψπ+ψs2π)Γ(n+12s2)Γ(1+ψπψs2π)((A2+B2)14|x|)sds

    which is equivalent, by (2.13), to the following expression in terms of Fox H-functions

    I4=B(A2+B2)1πn32(2|x|)nsin(ψ)H0,23,2[(A2+B2)14|x|(1n,1),(0,12),(ψπ,ψ2π)(1n2,12),(ψπ,ψ2π)]. (3.44)

    From Eqs (3.41)–(3.44) we conclude that the representation (3.22) of the solution u(x,t) of Eqs (3.1) and (3.2) corresponds to the sum of convolution integrals involving Fox H-functions.

    In the next subsection we summarize our calculations in the main result of the paper.

    Taking into account Eqs (3.22) and (3.41)–(3.44) we obtain our main result.

    Theorem 3.3. The solution of the time-fractional telegraph equation of distributed order Eq (3.1) subject to the conditions Eqs (3.3) and (3.2) and the additional assumption Eq (3.11) is given, in terms of convolution integrals, by

    u(x,t)=Rnf(z)G1(xz,t)dz+Rng1(z)G2(xz,t)dz+Rng2(z)G3(xz,t)dz+Rnt0q(z,w)G4(xz,tw)dwdz, (3.45)

    where the functions G1, G2, G3 and G4 are given by

    G1(x,t)=1πn12(2|x|)n+0ρ1(A2+B2)12ertsin(ψ)×[sin(γ1π)H((A2+B2)14|x|)+[Asin(γ1π)Bcos(γ1π)](A2+B2)12H((A2+B2)14|x|)]dr,G2(x,t)=1πn12(2|x|)n+0r(A2+B2)12ertsin(ψ)[ρ2sin(γ2π)H((A2+B2)14|x|)+1c2[ρ2c2(Asin(γ2π)Bcos(γ2π))+Bd2](A2+B2)12H((A2+B2)14|x|)]dr,G3(x,t)=1πn12(2|x|)n+0(A2+B2)12ertsin(ψ)[ρ2sin(γ2π)H((A2+B2)14|x|)+1c2[ρ2c2(Asin(γ2π)Bcos(γ2π))+Bd2](A2+B2)12H((A2+B2)14|x|)]dr,G4(x,t)=1c2πn12(2|x|)n+0B(A2+B2)1ertsin(ψ)H((A2+B2)14|x|)dr,

    where ρ1, γ1, ρ2, and γ2, A and B, and ψ are given, respectively, by Eq (3.14), (3.18) and (3.30). Moreover, the functions H and H are expressed in terms of the following Fox H-functions

    H((A2+B2)14|x|)=H1,24,3[(A2+B2)14|x|(1n,1),(1,12),(0,12),(0,ψ2π)(1,12),(1n2,12),(0,ψ2π)],H((A2+B2)14|x|)=H0,23,2[(A2+B2)14|x|(1n,1),(0,12),(ψπ,ψ2π)(1n2,12),(ψπ,ψ2π)].

    Remark 3.4. If we consider

    f(x)=δ(x)=ni=1δ(xi),g(x)=q(x,t)=0,a=c=1,d=λ

    with λR+ in Eqs (3.1) and (3.2), then the solution u(x,t) given by Eq (3.45) corresponds to the eigenfunctions of the generalized time-fractional telegraph equation of distributed order in Rn×R+. Moreover, if additionally b2(β,ν)=0 (resp. b1(α,μ)=0) we obtain the representation of the generalized eigenfunctions of the time-fractional diffusion (resp. wave) equation of distributed order in Rn×R+.

    Putting a=0 in Theorem 3.3, we have the following simplifications

    B1(s)=B1(s)=0,A=ρcos(γπ),B=ρsin(γπ),A2+B2=ρ2,ψ=γπ,

    which give the following result.

    Corollary 3.5. The solution of the generalized time-fractional wave equation of distributed order in Rn×R+

    1021b2(β,ν)tβ,ν0+u(x,t)dβdνc2Δxu(x,t)+d2u(x,t)=q(x,t)

    for a given density function b2(β,ν), subject to the following initial and boundary conditions

    (tI(1ν)(2β)0+u)(x,0+)=g1(x),[t(tI(1ν)(2β)0+u)](x,0+)=g2(x),lim|x|+u(x,t)=0,

    is given, in terms of convolution integrals, by

    u(x,t)=Rng1(z)G2(xz,t)dz+Rng2(z)G3(xz,t)dz+Rnt0q(z,w)G4(xz,tw)dwdz,

    where the functions G2, G3, and G4 are given by

    G2(x,t)=1πn12(2|x|)n+0rertρsin(γπ)[ρsin(γπ)H(1|x|ρ)+d2c2sin(γπ)H(1|x|ρ)]dr,G3(x,t)=1πn12(2|x|)n+0ertρsin(γπ)[ρsin(γπ)H(1|x|ρ)+d2c2sin(γπ)H(1|x|ρ)]dr,G4(x,t)=1c2πn12(2|x|)n+0ertρH(1|x|ρ)dr

    with ρ and γ, ρ and γ given by Eqs (3.13) and (3.14), respectively, and the functions H and H are expressed in terms of the following Fox H-functions

    H(1|x|ρ)=H0,23,2[1|x|ρ(1n,1),(1,12),(0,γ2)(1n2,12),(0,γ2)],H(1|x|ρ)=H0,23,2[1|x|ρ(1n,1),(0,12),(γ,γ2)(1n2,12),(γ,γ2)].

    Remark 3.6. If we consider ν=μ=1 in Theorem 3.3 (i.e., the telegraph equation has only Caputo fractional derivatives) we obtain the main result in [46]. For that we need to take into account Remarks 3.1 and 3.2 and to combine the first two integrals in Eq (3.45) into a unique integral. Hence, the solution u(x,t) is given by

    u(x,t)=Rnf(z)(G1(xz,t)+G2(xz,t))dz+Rng2(z)G3(xz,t)dz+Rnt0q(z,w)G4(xz,tw)dwdz, (3.46)

    where for ν=μ=1

    G1(x,t)+G2(x,t)=1πn12(2|x|)n+0(A2+B2)12ertrsin(ψ){[ρ1sin(γ1π)+ρ2sin(γ2π)]H((A2+B2)14|x|)+[Aρ1sin(γ1π)Bρ1cos(γ1π)+Aρ2sin(γ2π)Bρ2cos(γ2π)](A2+B2)12H((A2+B2)14|x|)+d2Bc2(A2+B2)12H((A2+B2)14|x|)}dr. (3.47)

    By Eq (3.18) we have that

    Aρ1sin(γ1π)Bρ1cos(γ1π)+Aρ2sin(γ2π)Bρ2cos(γ2π)=A[ρ1sin(γ1π)+ρ2sin(γ2π)]B[ρ1cos(γ1π)+ρ2cos(γ2π)]=ABBA=0,

    and, hence, Eq (3.47) simplifies to

    G1(x,t)+G2(x,t)=1πn12(2|x|)n+0B(A2+B2)12ertrsin(ψ)[H((A2+B2)14|x|)+d2c2(A2+B2)12H((A2+B2)14|x|)]dr

    which corresponds to the function G1 presented in the main result of [46]. Moreover, in Eq (3.46)

    G3(x,t)=1πn12(2|x|)n+0(A2+B2)12ertr2sin(ψ)[ρ2sin(γ2π)H((A2+B2)14|x|)+1c2[ρ2c2(Asin(γ2π)Bcos(γ2π))+Bd2](A2+B2)12H((A2+B2)14|x|)]dr,G4(x,t)=1c2πn12(2|x|)n+0B(A2+B2)1ertsin(ψ)H((A2+B2)14|x|)dr

    which correspond, respectively, to the functions G2 and G3 that appear in the main result of [46]. Therefore, we can claim that there is consistency in our results.

    Remark 3.7. The telegraph equations studied in [38,40] are particular cases of the equation studied in this paper, for the choices ν=μ=1, b2(β,ν)=δ(ββ1), b1(α,μ)=δ(αα1), with 1<β12 and 0<α11 d=0, and q(x,t)=0.

    The numerical implementation of Eq (3.45) is possible, however, depends substantially on the study of the asymptotic behaviour of G1, G2, G3 and G4 through the study of the asymptotic behaviour of the associated Fox H-functions. We would like to remark also that Eq (3.45) is a very general solution, but for particular cases of the dimension, of the fractional parameters, and/or of the density functions, it is possible to get simpler expressions.

    In this section we obtain the expression for some fractional moments of the first fundamental solution G1 of the time-fractional telegraph equation of distributed order Eq (3.1) with d=q(x,t)=0, and subject to the initial and boundary conditions

    f(x)=g1(x)=δ(x)=nj=1δ(xj) and g2(x)=0. (4.1)

    Then G1 is given by G1(x,t)=G1(x,t)+G2(x,t), where G1 and G2 are given in Theorem 3.3.

    It is well known that the Mellin transform (2.9) can be interpreted as the fractional moment of order s1 of the function f (see [62]). Therefore, we can calculate the fractional moments Mα,μ;β,νn;,γ of arbitrary order γ>0 of ˜G1, where ˜G1 denotes the Laplace transform of G1. Denoting by s the variable in the Laplace domain and by r the radial quantity |x|, we have, from the definition of the Mellin transform, that

    ˜Mα,μ;β,νn;,γ(s)=+0rγ˜G1(r,s)dr=+0rγn+11rn˜G1(r,s)dr=M{rn˜G1(r,s)}(γn+1). (4.2)

    From Eq (3.5), we have that

    G1(r,t)=L1{F1{ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1}(r,s)}(r,t)

    which is equivalent to

    ˜G1(r,s)=L{G1(r,t)}(r,s)=F1{ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1}(r,s).

    To calculate the inverse Fourier transform we are going to use the Mellin transform, similarly as it was done in Section 3. Taking into account Eq (3.21), we have that

    F1{ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1}(r,s)=r1n2(2π)n2+0[ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1]wn2Jn21(|x|w)dw=(g5Mf)(1r), (4.3)

    where M denotes the Mellin convolution given by Eq (2.11) at the point 1r with

    g5(w)=ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1

    and

    f(w)=1(2π)n2|x|nwn2+1Jn21(1w).

    Denoting by I7 the integral in Eq (4.3), we have, by relations Eqs (2.12) and (2.11), that

    M{I7}(s)=M{(g5Mf)(1r)}(s)=M{g5}(s)M{f}(s)

    which is equivalent to

    M{I7}(s)=M{g5}(s)M{f}(s). (4.4)

    From Eq (3.25) we have that

    M{f}(s)=1πn12|x|n2n1Γ(ns)Γ(n+1s2)Γ(s2). (4.5)

    Now, we calculate the Mellin transform of the function g5. Taking into account Eqs (2.9) and (3.8) with p=0, Eqs (3.9) and (3.10) with p=1, we get

    M{g5}(s)=+0ws1[ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1]dw=[B1(s)+sc2[c2B2(s)d2]]+0ws1B2(s)+B1(s)+w2dw.

    The integral in the previous expression was already calculated in [46] (see formula (84)). Therefore, we have that

    M{g5}(s)=[B1(s)+sc2[c2B2(s)d2]]πΓ(1s)Γ(s)Γ(1+s2)Γ(1s2)(B2(s)+B1(s))s21. (4.6)

    From Eqs (4.5), (4.6), (4.4) and (4.2) we conclude that

    M{rnF1{ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1}(r,s)}(γn+1,s)=[B1(s)+sc2[c2B2(s)d2]](B2(s)+B1(s))s21πn322n1Γ(n+s)Γ(1+s)Γ(s)Γ(n+1+s2)Γ(s2)Γ(1s2)Γ(1+s2)|s=γn+1. (4.7)

    By the duplication formula of the Gamma function Γ(2z)=22z1πΓ(z)Γ(z+12), we have the following equalities for the Gamma functions that appear in (4.7)

    Γ(n+s)Γ(12+n+s2)=2n+s1πΓ(n+s2), (4.8)
    [0.2cm]Γ(1+s)Γ(1+s2)=2sπΓ(1+s2), (4.9)
    [0.2cm]Γ(s)Γ(12s2)=2s1πΓ(s2). (4.10)

    Taking into account Eqs (4.8)–(4.10), expression Eq (4.7) simplifies to

    M{rnF1{ˆ˜u1(κ,s)+ˆ˜u2(κ,s)|p=1d2c2ˆ˜u3(κ,s)|p=1}(r,s)}(γn+1,s)=[B1(s)+sc2[B2(s)d2]](B2(s)+B1(s))s21πn22s1Γ(n+s2)Γ(1+s2)|s=γn+1

    and, consequently, the fractional moments of arbitrary order γ in the Laplace domain are given by

    ˜Mα,μ;β,νn;γ(s)=[B1(s)+sc2[c2B2(s)d2]](B2(s)+B1(s))γ+n32πn22γnΓ(γ+12)Γ(3+γn2). (4.11)

    If we restrict (4.11) to the time-fractional telegraph equation of distributed order with Caputo fractional derivatives, i.e., if we consider μ=ν=1 (which implies that B1(s)=s1B1(s) and B2(s)d2c2=s2[B2(s)d2c2] by Remark 3.1), then (4.11) becomes equal to

    ˜Mα,1;β,1n;γ(s)=c2B1(s)+c2B2(s)d2πn2c2s(B2(s)+B1(s))γ+n322γnΓ(γ+12)Γ(3+γn2),

    which coincides with the correspondent expression deduced in [46]. Let us now analyse expression Eq (4.11) for some special cases:

    ● When γ=n2k3, with n>2k+3 and kN0, the correspondent moments in the Laplace domain become infinite.

    ● When γ=1 (mean value), we have

    ˜Mα,μ;β,νn;1(s)=[B1(s)+sc2[c2B2(s)d2]](B2(s)+B1(s))n22πn221nΓ(2n2), (4.12)

    which becomes infinite when n=4+2k, with kN0.

    ● When γ=2 (variance), we have

    ˜Mα,μ;β,νn;2(s)=[B1(s)+sc2[c2B2(s)d2]](B2(s)+B1(s))n52πn+1223nΓ(5n2), (4.13)

    which becomes infinite when n=5+2k, with kN0.

    When the fundamental solution is a positive function, it is possible to classify the diffusion process by analysing the correspondent second-order moment, also called the mean squared displacement of a particle. This is obtained by comparison with the variance in the normal diffusion process. In this subsection we consider the particular case of double-order distributed fractional derivatives, and we analyze the second-order moment for short and long-time. We separate our analysis between the diffusion and the wave cases. The following Laplace inversion formulas will be needed in the sequel:

    ● Formula (2.1.1.1) in [63]

    L1{1sν}(t)=tν1Γ(ν),ν>0. (4.14)

    ● Formula (5.1.26) in [61]

    L1{sαγβ(sαλ)γ}(t)=tβ1Eγα,β(λtα),Re(α),Re(β)>0,λC, (4.15)

    where Eγα,β is the three parameter Mittag-Leffler function given by (2.14).

    Here we consider b2(β,ν)=0, which implies that B2(s)=B2(s)=0. In this case, the second-order moment in the Laplace domain becomes

    ˜Mα,μ;,n;2(s)=˜Mα,μn;2(s)=21nΓ(5n2)πn12B1(s)(B1(s))n52,n5+2k,kN0. (4.16)

    Further, we assume

    b1(α,μ)=k1δ(αα1)δ(μμ1)+k2δ(αα2)δ(μμ2) (4.17)

    with 0<α1<α21, 0μ1,μ21, μ2<μ11α11α2, k1,k2>0, and k1+k2=1. For this b1(α,μ) we get

    B1(s)=ak1c2sα1+ak2c2sα2andB1(s)=ak1c2sμ1(1α1)+ak2c2sμ2(1α2). (4.18)

    Considering Eq (4.18) in Eq (4.16) we get

    ˜M(α1,α2),(μ1,μ2)n;2(s)=21nan32Γ(5n2)πn12cn3(k1sμ1(1α1)+k2sμ2(1α2))(k1sα1+k2sα2)n52. (4.19)

    To invert the Laplace transform of ˜M(α1,α2),(μ1,μ2)n;2 we first rearrange the expression Eq (4.19):

    ˜M(α1,α2),(μ1,μ2)n;2(s)=21n(ak2)n32Γ(5n2)πn12cn3k1k2s(α2α1)(5n)2+α2(n5)2μ1(1α1)(sα2α1(k1k2))5n2+21n(ak2)n32Γ(5n2)πn12cn3s(α2α1)(5n)2+α2(n5)2μ2(1α2)(sα2α1(k1k2))5n2. (4.20)

    Taking into account Eq (4.15) with

    α=α2α1,γ=5n2,λ=k1k2,β=α2(5n)2+μ1(1α1)(1st term),β=α2(5n)2+μ2(1α2)(2nd term),

    we get from Eq (4.20) that

    M(α1,α2),(μ1,μ2)n;2(t)=21n(ak2)n32Γ(5n2)πn12cn3k1k2tα2(5n)2+μ1(1α1)1E5n2α2α1,α2(5n)2+μ1(1α1)(k1k2tα2α1)+21n(ak2)n32Γ(5n2)πn12cn3tα2(5n)2+μ2(1α2)1E5n2α2α1,α2(5n)2+μ2(1α2)(k1k2tα2α1). (4.21)

    Remark 4.1. For n=1, the expression Eq (4.21) reduces to the expression (28) in [51] with suitable identification of the parameters, which indicates consistency in our results.

    The graphical representation of M(α1,α2),(μ1,μ2)n;2 as a function of t using Eq (4.21) is not an easy procedure due to the presence of the three parameter Mittag-Leffler function Eγα,β. The numerical implementation of this special function is possible for some range of the parameter α (see [64]), which do not include all the cases studied in this work. Therefore, we will make an asymptotic analysis of (4.19) using the Tauberian analysis. Since sμ2(1α2)+μ1(1α1)0 and sα2α10 as s0 then, we get the following asymptotic behaviour of ˜M(α1,α2),(μ1,μ2)n;2(s) as s0:

    ˜M(α1,α2),(μ1,μ2)n;2(s)=21nan32Γ(5n2)πn12cn3(k1sμ1(1α1)+k2sμ2(1α2))(k1sα1+k2sα2)n5221n(ak1)n32Γ(5n2)πn12cn3sμ1(1α1)+α1(n5)2. (4.22)

    Concerning the symbol in the previous and subsequent expressions, we say that f and g are asymptotically equivalent as w (resp. as w0), i.e., fg, if and only if limwf(w)g(w)=1 (resp. limw0f(w)g(w)=1).

    Using (4.14) to invert the Laplace transform in (4.22), we obtain the asymptotic behavior of M(α1,α2),(μ1,μ2)n;2 for t+, with 0μ11 and n and α1 according to the following cases:

    M(α1,α2),(μ1,μ2)n;2(t){21n(ak1)n32Γ(5n2)πn12cn3tμ1+α1(5n2μ1)21Γ(μ1+α1(5n2μ1)2),α1<α2,n=1,2,3,421n(ak1)n32Γ(5n2)πn12cn3tμ1+α1(5n2μ1)21Γ(μ1+α1(5n2μ1)2),α1<2μ12μ1+n5α1<α2,n=6,8,. (4.23)

    To classify the type of diffusion process we need to compare M(α1,α2),(μ1,μ2)n;2 with the moment of the normal diffusion process which is given by

    M(1,1),(μ1,μ2)n;2=21n(ak1)n32πn12cn3t3n2. (4.24)

    According with the dimension n we have the following cases:

    n=1,2,3: M(α1,α2),(μ1,μ2)n;2(t)/M(1,1),(μ1,μ2)n;2(t)0, as t+, for all 0μ1<1, and all admissible α1 and n, which corresponds to a subdiffusion process in the long time. In the limit case, μ1=1 (Caputo case), for n=1,2 we still have a subdiffusion process while for n=3 holds M(α1,α2),(μ1,μ2)3;2(t)/M(1,1),(μ1,μ2)3;2(t)k>0, as t+, thus the process coincides with the normal diffusion in the long time;

    n=4: The classification of the type of diffusion depends on the type μ1.

    ▶ For 0μ1<1/2 holds M(α1,α2),(μ1,μ2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)0, as t+, thus corresponding to a subdiffusion process in the long time;

    ▶ For μ1=1/2 holds M(α1,α2),(μ1,μ2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)k>0, as t+, thus the process coincides with the normal diffusion in the long time;

    ▶ For 1/2<μ11 holds M(α1,α2),(μ1,μ2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)+, as t+, thus corresponding to a superdiffusion process in the long time.

    n=6+2k: The moment M(α1,α2),(μ1,μ2)n;2(t) is positive when n=8+4k and negative when n=6+4k, kZ+. This is due to the change of sign of the gamma function in the numerator. In the second case, a probabilistic interpretation is no longer possible.

    Finally, we note a different behaviour of the monotonicity of the second-order moment along the dimensions, depending on the values assumed by the several parameters in the expression. When n=1,2,3,4, if we assume that c>0, then the following conclusions about M(α1,α2),(μ1,μ2)n;2 for large values of t can be drawn:

    n=1: M(α1,α2),(μ1,μ2)n;2 is an increasing function when 1μ12μ1<α11, is constant when α1=1μ12μ1, and is a decreasing function when 0<α1<1μ12μ10.5, with 0μ11.

    n=2: M(α1,α2),(μ1,μ2)n;2 is an increasing function when 22μ132μ1<α11, is constant when α1=22μ132μ1, and is a decreasing function when 0<α1<22μ132μ123, with 0μ11.

    n=3: M(α1,α2),(μ1,μ2)n;2 is a decreasing function when 0<α1<1 and 0μ1<1, and is constant when α1=1 or 0α11 and μ1=1.

    n=4: M(α1,α2),(μ1,μ2)n;2 is a decreasing function for all 0<α11 and 0μ1<1, and is constant when α1=0 and μ1=1.

    Now, we study the asymptotic behaviour of M(α1,α2),(μ1,μ2)n;2 for small values of t. From Eq (4.19) we have the following asymptotic behaviour when s+:

    ˜M(α1,α2),(μ1,μ2)n;2(s)=21nan32Γ(5n2)πn12cn3(k1sμ1(1α1)+k2sμ2(1α2))(k2sα1+k2sα2)n5221n(ak2)n32Γ(5n2)πn12cn3sμ2(1α2)+α2(n5)2. (4.25)

    Using Eq (4.14) to invert the Laplace transform in Eq (4.25), we obtain the asymptotic behavior of M(α1,α2),(μ1,μ2)n;2 for t0+, with 0μ21 and n and α2 according to the following cases:

    M(α1,α2),(μ1,μ2)n;2(t){21n(ak2)n32Γ(5n2)πn12cn3tμ2+α2(5n2μ2)21Γ(μ2+α2(5n2μ2)2),α2>α1,n=1,2,3,421n(ak2)n32Γ(5n2)πn12cn3tμ2+α2(5n2μ2)21Γ(μ2+α2(5n2μ2)2),α2<2μ22μ2+n5α2>α1,n=6,8,. (4.26)

    Comparing M(α1,α2),(μ1,μ2)n;2 with M(1,1),(μ1,μ2)n;2 when t0+, we have the following conclusions:

    n=1,2,3: M(α1,α2),(μ1,μ2)n;2(t)/M(1,1),(μ1,μ2)n;2(t)+, as t0+, for all 0μ1<1, and all admissible α1 and n, which corresponds to a superdiffusion process in the short time. In the limit case, μ1=1 (Caputo case), for n=1,2 we still have a superdiffusion process while for n=3 holds M(α1,α2),(μ1,μ2)3;2(t)/M(1,1),(μ1,μ2)3;2(t)k>0, as t+, thus the process coincides with the normal diffusion in the short time;

    n=4: The classification of the type of diffusion depends on the parameter μ1.

    ▶ For 0μ1<1/2 holds M(α1,α2),(μ1,μ2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)+, as t0+, thus corresponding to a superdiffusion process in the short time;

    ▶ For μ1=1/2 holds M(α1,α2),(μ1,μ2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)k>0, as t0+, thus the process coincides with the normal diffusion in the short time;

    ▶ For 1/2<μ11 holds M(α1,α2),(μ1,μ2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)0, as t0+, thus corresponding to a subdiffusion process in the short time.

    n=6+2k: As happen in the long time case the moment is not always positive, and hence also here it is not possible to perform a probabilistic interpretation for all the values of n.

    Here we consider b1(α,μ)=0. Hence, B1(s)=B1(s)=0 and the second-order moment in the Laplace domain becomes

    M,;β,νn;2(s)=Mβ,νn;2(s)=21nΓ(5n2)πn12sB2(s)(B2(s))n52,n5+2k,kN0. (4.27)

    Let us now assume that

    b2(β,ν)=k1δ(ββ1)δ(νν1)+k2δ(ββ2)δ(νν2)

    with 1<β1<β22, 0ν1,ν21, ν2<ν12β12β2, k1,k2>0, and k1+k2=1. For this b2(β,ν) we get

    B2(s)=k1c2sβ1+k2c2sβ2andB2(s)=k1c2sν1(2β1)+k2c2sν2(2β2). (4.28)

    Considering Eq (4.28) in Eq (4.27) we get

    ˜M(β1,β2),(ν1,ν2)n;2(s)=21nΓ(5n2)πn12cn3s(k1sν1(2β1)+k2sν2(2β2))(k1sβ1+k2sβ2)n52. (4.29)

    Inverting the Laplace transform of ˜M(β1,β2),(ν1,ν2)n;2(s) following the same steps of the deduction of Eq (4.21) we get

    M(β1,β2),(ν1,ν2)n;2(t)=21nkn322Γ(5n2)πn12cn3k1k2tβ2(5n)2+ν1(2β1)2E5n2β2β1,β2(5n)2+ν1(2β1)1(k1k2tβ2β1)+21nkn322Γ(5n2)πn12cn3tβ2(5n)2+ν2(2β2)2E5n2β2β1,β2(5n)2+ν2(2β2)1(k1k2tβ2β1). (4.30)

    First, we study the asymptotic behaviour of M(β1,β2),(ν1,ν2)n;2 for large values of t. From Eq (4.19) we have the following asymptotic behaviour when s0:

    ˜M(β1,β2),(ν1,ν2)n;2(s)=21nΓ(5n2)πn12cn3s(k1sν1(2β1)+k2sν2(2β2))(k1sβ1+k2sβ2)n5221nkn321Γ(5n2)πn12cn3s1ν1(2β1)+β1(n5)2. (4.31)

    Using Eq (4.14) to invert the Laplace transform in Eq (4.31), we obtain the asymptotic behavior of M(β1,β2),(ν1,ν2)n;2 for t+, with β1, ν1, and n according to the following cases:

    M(β1,β2),(ν1,ν2)n;2(t){21nkn321Γ(5n2)πn12cn3t2ν1+β1(5n2ν1)22Γ(2ν1+β1(5n2ν1)21),β1<β2,0ν11,n=1,2,321nkn321Γ(5n2)πn12cn3t2ν1+β1(5n2ν1)22Γ(2ν1+β1(5n2ν1)21),β1<β2,12<ν11,n=4. (4.32)

    To classify the type of diffusion-wave process we need to compare M(β1,β2),(ν1,ν2)n;2 with the normal diffusion process. The situation here depends on the dimension and we have the following cases:

    n=1: M(β1,β2),(ν1,ν2)1;2(t)/M(1,1),(μ1,μ2)1;2(t)+, as t+, for 1<32ν12ν1<β1<2 and 0ν11, which corresponds to a superdiffusion process in the long time and M(β1,β2),(ν1,ν2)1;2(t)/M(1,1),(μ1,μ2)1;20, as t+, for 1<β1<32ν12ν1<2 and 0ν11, which corresponds to a subdiffusion process in the long time. In the special case β1=32ν12ν1 the process coincides with the normal diffusion in the long time.

    n=2: M(β1,β2),(ν1,ν2)n;2(t)/M(1,1),(μ1,μ2)2;2(t)+, as t+, for 1<54ν132ν1<β1<2 and 0ν11, which corresponds to a superdiffusion process in the long time and M(β1,β2),(ν1,ν2)n;2(t)/M(1,1),(μ1,μ2)2;2(t)0, as t+, for 1<β1<54ν132ν1<2 and 0ν11, which corresponds to a subdiffusion process in the long time. In the special case β1=54ν132ν1 the process coincides with the normal diffusion in the long time.

    n=3 M(β1,β2),(ν1,ν2)3;2(t)/M(1,1),(μ1,μ2)3;2(t)0, as t+, for all 1<β12 and 0ν1<1, thus corresponding to a subdiffusion process. For ν1=1 the process coincides with the normal diffusion case.

    n=4: M(β1,β2),(ν1,ν2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)0, as t+, for all 1<β12 and 1/2<ν11, thus corresponding to a subdiffusion process.

    Finally, we study the asymptotic behaviour of M(β1,β2),(ν1,ν2)n;2 for small values of t knowing the asymptotic behaviour of Eq (4.29) when s+. From (4.29), as s+, we have

    ˜M(β1,β2),(ν1,ν2)n;2(s)=21nΓ(5n2)πn12cn3s(k2sν1(2β1)+k2sν2(2β2))(k1sβ1+k2sβ2)n5221nkn321Γ(5n2)πn12cn3s1ν2(2β2)+β2(n5)2. (4.33)

    Using Eq (4.14) to invert the Laplace transform in Eq (4.33), we obtain the asymptotic behavior of M(β1,β2),(ν1,ν2)n;2 for t0+, with β2, ν2, and n according to the following cases:

    M(β1,β2),(ν1,ν2)n;2(t){21nkn322Γ(5n2)πn12cn3t2ν2+β2(5n2ν2)22Γ(2ν2+β2(5n2ν2)21),β2>β1,0ν21,n=1,2,321nkn322Γ(5n2)πn12cn3t2ν2+β2(5n2ν2)22Γ(2ν2+β2(5n2ν2)21),β2>β1,12<ν21,n=4. (4.34)

    The analysis of Eq (4.34) is similar to the one performed for Eq (4.32). Regarding the classification of the diffusion-wave process the following conclusions can be taken:

    n=1: M(β1,β2),(ν1,ν2)1;2(t)/M(1,1),(μ1,μ2)1;2(t)0, as t0+, for 1<32ν12ν1<β2<2 and 0ν21, while M(β1,β2),(ν1,ν2)n;2(t)/M(1,1),(μ1,μ2)1;2(t)+, as t0+, for 1<β2<32ν12ν1<2 and 0ν21. Hence, in the short time, the process is subdiffusive in the first case and is superdiffusive in the second case. In the special case β1=32ν12ν1 the process coincides with the normal diffusion in the short time.

    n=2: M(β1,β2),(ν1,ν2)2;2(t)/M(1,1),(μ1,μ2)2;2(t)0, as t0+, for 1<54ν132ν1<β2<2 and 0ν21, which corresponds to a subdiffusion process in the short time, and M(β1,β2),(ν1,ν2)2;2(t)/M(1,1),(μ1,μ2)2;2(t)+, as t0+, for 1<β2<54ν132ν1<2 and 0ν21, thus corresponding to a superdiffusion process in the short time. In the special case β1=54ν132ν1 the process coincides with the normal diffusion in the short time.

    n=3: M(β1,β2),(ν1,ν2)3;2(t)/M(1,1),(μ1,μ2)3;2(t)+, as t0+, for all 1<β22 and 0ν2<1, thus corresponding to a superdiffusion process. For ν2=1 the process coincides with the normal diffusion case.

    n=4: M(β1,β2),(ν1,ν2)4;2(t)/M(1,1),(μ1,μ2)4;2(t)+, as t0+, for all 1<β22 and 1/2<ν21, thus corresponding to a superdiffusion process in the short time.

    Remark 4.2. If we consider the Caputo case in Sections 4.1.1 and 4.1.2, our analysis of the diffusion-wave process for the double-order case improve the results presented in [46]. Moreover, considering single order derivatives and the one-dimensional case, it was proved in [65] that the fundamental solution corresponds to a probability density function only when the fractional derivatives are in the Caputo sense.

    In this section we present and analyse the plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)n;2(t), when t+, for some of the cases studied previously separating the diffusion and the wave cases. The plots were generated using Mathematica software and the commands available in it.

    The diffusion case: In the following figures, we show the graphical representation of Eq (4.23) for n=1,2,3,4, α1=0.25,0.50,0.75, and different values μ1, using a logarithmic scale in the axes when needed.

    Looking at the plots we see that the classification of the diffusion process in each dimension agrees with the analysis of Eq (4.23) performed previously. The plots show an interpolation between the extreme cases μ1=0 and μ1=1, which correspond to the Riemann-Liouville (RL) and Caputo cases, respectively. The extreme cases have different behaviour, e.g., the slope of the variance is different and in the dimension, n=3 the variance is constant in the Caputo case. In contrast, in the RL case, the variance decreases for large values of t. Moreover, for α1=0.25 we can observe a different behaviour of the diffusion in dimensions n=1 and n=2: in the RL case the variance decreases for large values of t while in the Caputo case the variance increases.

    Figure 1.  Plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)1;2(t) when t+ for α1=0.25,0.50,0.75 (from left).
    Figure 2.  Plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)n;2(t) when t+ for n=2 and α1=0.25, n=3 and α1=0.50, n=4 and α1=0.75 (from left).

    The wave case: In the following figures, we show the graphical representation of Eq (4.32) for n=1,2,3,4, β1=1.25,1.50,1.75 and different values ν1.

    For each dimension and different values of the fractional parameters, the process classification agrees with the analysis of Eq (4.32). The range of the plots increases with the increase of β1 and ν1. Again it is interesting to observe the different behaviour of the variance for the extreme cases ν1=0 (RL case) and ν1=1 (Caputo case) for the dimension n=3. Also the slope of the variance is different in the other dimensions.

    Figure 3.  Plots of the asymptotic behaviour of M(β1,β2),(ν1,ν2)1;2(t) when t+ for β1=1.25,1.50,1.75 (from left).
    Figure 4.  Plots of the asymptotic behaviour of M(β1,β2),(ν1,ν2)n;2(t) when t+ for n=2 and β1=1.25, n=3 and β1=1.50, n=4 and β1=1.75 (from left).

    In this section we present and analyse the graphical representation of the asymptotic behaviour of M(α1,α2),(μ1,μ2)n;2(t), when t0+, for some of the cases studied previously separating the diffusion and the wave cases. The plots were generated using Mathematica software and the commands available in it.

    The diffusion case: In the following figures, we show the graphical representation of (4.26) for n=1,2,3,4, α2=0.25,0.50,0.75, and different values μ2.

    Looking at the plots we see that the range of the plots decreases with the increase of α2 and μ2. The type of process is in accordance with the conclusion made in the analysis of Eq (4.26). Again, we can observe a different behaviour of the variance for small values of t in the extreme cases ν2=0 and ν2=1, corresponding to the RL and Caputo cases, respectively. For α1=0.25 and the dimensions n=1 and n=2 the variance decreases in the RL case and increases in the Caputo case, for small values of t. In the case ν2=1 the plots coincide with the correspondent ones obtained in [46].

    Figure 5.  Plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)1;2(t) when t0+ for α2=0.25,0.50,0.75 (from left).
    Figure 6.  Plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)n;2(t) when t0+ for n=2 and α2=0.25, n=3 and α2=0.50, n=4 and α2=0.75 (from left).

    The wave case: In the following figures, we have the graphical representation of Eq (4.34) for n=1,2,3,4, β2=1.25,1.50,1.75, and different values ν2 and n.

    Looking at the plots we see that the conclusions are similar to those already exposed. The behaviour of the functions is in accordance with the conclusions made in the analysis of (4.34). When ν2=1 the plots coincide with those presented in [46].

    Figure 7.  Plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)1;2(t) when t0+ for β2=1.25,1.50,1.75 (from left).
    Figure 8.  Plots of the asymptotic behaviour of M(α1,α2),(μ1,μ2)n;2(t) when t0+ for n=2 and β2=1.25, n=3 and β2=0.50, n=4 and β2=0.75 (from left).

    The results presented here generalize those obtained in [46] by the introduction of the Hilfer derivative, that allows a smooth interpolation between the Riemann-Liouvile and the Caputo fractional derivatives. The solution of the Cauchy problem associated with the telegraph equation was expressed as convolutions with functions that are expressed by Laplace integrals involving Fox H-functions. For particular cases of the equation the solution can be simplified and we showed that we can recover known results presented in the literature, which reveals consistency of our results. The classification of the diffusion-wave process depends, not only on the spatial dimension, but also on the order and type of the derivatives. This is very different from previous works on the literature since in most cases the telegraph equation is studied only for Caputo or Riemann-Liouville fractional derivatives. It would be interesting to consider other types of density functions in addition to those considered in this article, but the calculations would become cumbersome.

    The work of the authors was supported by Portuguese funds through CIDMA–Center for Research and Development in Mathematics and Applications, and FCT–Fundação para a Ciência e a Tecnologia, within projects UIDB/04106/2020 and UIDP/04106/2020.

    N. Vieira was also supported by FCT via the 2018 FCT program of Stimulus of Scientific Employment - Individual Support (Ref: CEECIND/01131/2018).

    The authors thank the anonymous reviewers for the careful reading of the manuscript, their suggestions, and additional references that have improved the article.

    The authors declare there is no conflicts of interest.



    [1] A. V. Chechkin, R. Gorenflo, I. M. Sokolov, Retarding subdiffusion and accelerating superdiffusion governed by distributed-order fractional diffusion equations, Phys. Rev. E, 66 (2002), 046129. https://doi.org/10.1103/PhysRevE.66.046129 doi: 10.1103/PhysRevE.66.046129
    [2] F. Mainardi, A. Mura, G. Pagnini, R. Gorenflo, Time-fractional diffusion of distributed order, J. Vib. Control, 14 (2008), 1267–1290. https://doi.org/10.1177%2F1077546307087452
    [3] T. Sandev, A. Chechkin, N. Korabel, H. Kantz, I. M. Sokolov, R. Metzler, Distributed-order diffusion equations and multifractality: Models and solutions, Phys. Rev. E, 92 (2015), 04217. https://doi.org/10.1103/physreve.92.042117 doi: 10.1103/physreve.92.042117
    [4] M. Caputo, Distributed order differential equations modeling dielectric induction and diffusion, Fract. Calc. Appl. Annal., 4 (2001), 421–442.
    [5] A. Ansari, M. Moradi, Exact solutions to some models of distributed-order time fractional diffusion equations via the Fox H functions, ScienceAsia, 39 (2013), 57–66. http://dx.doi.org/10.2306/scienceasia1513-1874.2013.39S.057 doi: 10.2306/scienceasia1513-1874.2013.39S.057
    [6] W. Ding, S. Patnaik, F. Semperlotti, Multiscale nonlocal elasticity: A distributed order fractional formulation, Int. J. Mech. Sci. 226 (2021), 19. https://doi.org/10.1016/j.ijmecsci.2022.107381
    [7] J. Jia, H. Wang, Analysis of a hidden memory variably distributed-order space-fractional diffusion equation, Appl. Math. Lett., 124 (2022), 107617. https://doi.org/10.1016/j.aml.2021.107617 doi: 10.1016/j.aml.2021.107617
    [8] J. Jia, X. Zheng, H. Wang, Analysis and fast approximation of a steady-state spatially-dependent distributed-order space-fractional diffusion equation, Fract. Calc. Appl. Anal., 24 (2021), 1477–1506. https://doi.org/10.1515/fca-2021-0062 doi: 10.1515/fca-2021-0062
    [9] Y. Kumar, V. K. Singh, Computational approach based on wavelets for financial mathematical model governed by distributed order fractional differential equation, Math. Comput. Simul., 190 (2021), 531–569. https://doi.org/10.1016/j.matcom.2021.05.026 doi: 10.1016/j.matcom.2021.05.026
    [10] M. Naber, Distributed order fractional sub-diffusion, Fractals, 12 (2004), 23–32. https://doi.org/10.1142/S0218348X04002410 doi: 10.1142/S0218348X04002410
    [11] I. M. Sokolov, A. V. Chechkin, J. Klafter, Distributed order fractional kinetics, Acta Physica Polonica, 35 (2004), 1323–1341.
    [12] Yu. Luchko, Boundary value problems for the generalized time-fractional diffusion equation of distributed order, Fract. Calc. Appl. Anal., 12(2009), 409–422.
    [13] M. Al-Refai, Yu. Luchko, Analysis of fractional diffusion equations of distributed order: Maximum principles and their applications, Analysis, 36 (2016), 123–133. https://doi.org/10.1515/anly-2015-5011 doi: 10.1515/anly-2015-5011
    [14] S. Patnaik, F. Semperlotti, Application of variable- and distributed-order fractional operators to the dynamic analysis of nonlinear oscillators, Nonlinear Dyn., 100 (2020), 561–580. https://doi.org/10.1007/s11071-020-05488-8 doi: 10.1007/s11071-020-05488-8
    [15] R. Gorenflo, Yu. Luchko, M. Stojanović, Fundamental solution of a distributed order time-fractional diffusion-wave equation as probability density, Fract. Calc. Appl. Anal., 16 (2013), 297–316. https://doi.org/10.2478/s13540-013-0019-6 doi: 10.2478/s13540-013-0019-6
    [16] R. Hilfer, Fractional calculus and regular variation in thermodynamics, in Applications of Fractional Calculus in Physics (ed. R. Hilfer), World Scientific, Singapore, (2000), 429–463.
    [17] R. Metzler, J. Klafter, The random walk's guide to anomalous diffusion: A fractional dynamics approach, Phys. Rep., 339 (2000), 1–77. https://doi.org/10.1016/S0370-1573(00)00070-3 doi: 10.1016/S0370-1573(00)00070-3
    [18] W. Ding, S. Patnaik, S. Sidhardh, F. Semperlotti, Applications of Distributed-Order Fractional Operators: A Review, Entropy, 23 (2021), 110. https://doi.org/10.3390/e23010110 doi: 10.3390/e23010110
    [19] M. Caputo, M. Fabrizio, The kernel of the distributed order fractional derivatives with an application to complex materials, Fractal Fract., 1 (2017), 13. https://doi.org/10.3390/fractalfract1010013 doi: 10.3390/fractalfract1010013
    [20] G. Calcagni, Towards multifractional calculus, Front. Phys., 6 (2018), 58. https://doi.org/10.3389/fphy.2018.00058 doi: 10.3389/fphy.2018.00058
    [21] C. F. Lorenzo, T. T. Hartley, Variable order and distributed order fractional operators, Nonlinear Dyn. 29 (2002), 57–98. https://doi.org/10.1023/A:1016586905654
    [22] W. Thomson, On the theory of the electric telegraph, Proc. R. Soc. Lond, Ser. I, 7 (1854), 382–399. https://www.jstor.org/stable/111814
    [23] C. Cattaneo, Sur une forme de l'équation de la chaleur éliminant le paradoxe d'une propagation instantanée, C. R. Acad. Sci., Paris, 246 (1958), 431–433.
    [24] W. Hayt, Engineering Electromagnetics, 5th edition, McGraw-Hill, New York, 1989.
    [25] J. Banasiak, R. Mika, Singular perturbed telegraph equations with applications in random walk theory, J. Appl. Stoch. Anal., 11 (1998), 9–28. https://doi.org/10.1155/S1048953398000021 doi: 10.1155/S1048953398000021
    [26] F. Effenberger, Y. Litvinenko, The diffusion approximation versus the telegraph equation for modeling solar energetic particle transport with adiabatic focusing, Astrophys. J., 783 (2014), 15. https://doi.org/10.1088/0004-637X/783/1/15 doi: 10.1088/0004-637X/783/1/15
    [27] A. Okuko, Application of the telegraph equation to oceanic diffusion: another mathematical model, Technical Report No.69, Chesapeake Bay Institute, Johns Hopkins University, Baltimore, 1971.
    [28] V. H. Weston, S. He, Wave splitting of telegraph equation in R and its application to inverse scattering, Inverse Probl., 9 (1993), 789–812. https://doi.org/10.1088/0266-5611/9/6/013 doi: 10.1088/0266-5611/9/6/013
    [29] L. Boyadjiev, Y. Luchko, The neutral-fractional telegraph equation, Math. Model. Nat. Phenom., 12 (2017), 51–67. https://doi.org/10.1051/mmnp/2017064 doi: 10.1051/mmnp/2017064
    [30] J. Masoliver, Telegraphic transport processes and their fractional generalization: A review and some extensions, Entropy, 23 (2021), 364. https://doi.org/10.3390/e23030364 doi: 10.3390/e23030364
    [31] E. Orsingher, L. Beghin, Time-fractional telegraph equations and telegraph processes with Brownian time, Probab. Theory Relat. Fields, 128 (2004), 141–160. https://doi.org/10.1007/s00440-003-0309-8 doi: 10.1007/s00440-003-0309-8
    [32] E. Orsingher, B. Toaldo, Space-time fractional equations and the related stable processes at random time, J. Theor. Probab., 30 (2017), 1–26. https://doi.org/10.1007/s10959-015-0641-9 doi: 10.1007/s10959-015-0641-9
    [33] R. C. Cascaval, E. C. Eckstein, L. C. Frota, J. A. Goldstein, Fractional telegraph equations, J. Math. Anal. Appl., 276 (2002), 145–159. https://doi.org/10.1016/S0022-247X(02)00394-3 doi: 10.1016/S0022-247X(02)00394-3
    [34] R. F. Camargo, A. O. Chiacchio, E. C. de Oliveira, Differentiation to fractional orders and the fractional telegraph equation, J. Math. Phys., 49 (2008), 12. https://doi.org/10.1063/1.2890375 doi: 10.1063/1.2890375
    [35] R. K. Saxena, R. Garra, E. Orsingher, Analytical solution of space-time fractional telegraph-type equations involving Hilfer and Hadamard derivatives, Integr. Transform. Spec. Funct., 27 (2015), 30–42. https://doi.org/10.1080/10652469.2015.1092142 doi: 10.1080/10652469.2015.1092142
    [36] K. Górska, A. Horzela, E. K. Lenzi, G. Pagnini, T. Sandev, Generalized Cattaneo (telegrapher's) equations in modeling anomalous diffusion phenomena, Phy. Review E, 102 (2020), 13. https://doi.org/10.1103/PhysRevE.102.022128 doi: 10.1103/PhysRevE.102.022128
    [37] M. Ferreira, M. M. Rodrigues, N. Vieira, Application of the fractional Sturm-Liouville theory to a fractional Sturm-Liouville telegraph equation, Complex Anal. Oper. Theory, 15 (2021), 36. https://doi.org/10.1007/s11785-021-01125-3 doi: 10.1007/s11785-021-01125-3
    [38] M. Ferreira, M. M. Rodrigues, N. Vieira, First and second fundamental solutions of the time-fractional telegraph equation with Laplace or Dirac operators, Adv. Appl. Clifford Algebr., 28 (2018), 14. https://doi.org/10.1007/s00006-018-0858-7 doi: 10.1007/s00006-018-0858-7
    [39] M. Ferreira, M. M. Rodrigues, N. Vieira, Fundamental solution of the time-fractional telegraph Dirac operator, Math. Methods Appl. Sci., 40 (2017), 7033–7050. https://doi.org/10.1002/mma.4511 doi: 10.1002/mma.4511
    [40] M. Ferreira, M. M. Rodrigues, N. Vieira, Fundamental solution of the multi-dimensional time fractional telegraph equation, Fract. Calc. Appl. Anal., 20 (2017), 868–894. https://doi.org/10.1515/fca-2017-0046 doi: 10.1515/fca-2017-0046
    [41] M. D'Ovidio, E. Orsingher, B. Toaldo, Time-changed processes governed by space-time fractional telegraph equations, Stoch. Anal. Appl., 32 (2014), 1009–1045. https://doi.org/10.1080/07362994.2014.962046 doi: 10.1080/07362994.2014.962046
    [42] J. Masoliver, K. Lindenberg, Two-dimensional telegraphic processes and their fractional generalizations, Phys. Rev. E, 101 (2020). https://doi.org/10.1103/PhysRevE.101.012137
    [43] J. Masoliver, Three-dimensional telegrapher's equation and its fractional generalization, Phys. Rev. E, 96 (2017). https://link.aps.org/doi/10.1103/PhysRevE.96.022101
    [44] J. Masoliver, Fractional telegrapher's equation from fractional persistent random walks, Phys. Rev. E, 93 (2016). https://link.aps.org/doi/10.1103/PhysRevE.93.052107
    [45] J. Masoliver, J. M. Porrà, G. H. Weiss, Some two and three-dimensional persistent random walks, Phys. A: Stat. Mech. Appl., 193 (1993), 469–482. https://doi.org/10.1016/0378-4371(93)90488-P doi: 10.1016/0378-4371(93)90488-P
    [46] N. Vieira, M. M. Rodrigues, M. Ferreira, Time-fractional telegraph equation of distributed order in higher dimensions, Commun. Nonlinear Sci. Numer. Simulat., 102 (2021). https://doi.org/10.1016/j.cnsns.2021.105925
    [47] R. Hilfer, Applications of Fractional Calculus in Physics, World Scientific, Singapore, 2000.
    [48] R. Hilfer, Threefold introduction to fractional derivatives, Chapter 2, in: Anomalous Transport: Foundations and Applications (eds. R. Klages, G.Radons and I.M. Sokolov). Weinheim: Wiley-VCH, (2008), 17–74.
    [49] R. Hilfer, Experimental evidence for fractional time evolution in glass forming materials, Chem. Phys., 284 (2002), 399–408. https://doi.org/10.1016/S0301-0104(02)00670-5 doi: 10.1016/S0301-0104(02)00670-5
    [50] R. Hilfer, Y. Luchko, Z. Tomovski, Operational method for the solution of fractional differential equations with generalized Riemann-Liouville fractional derivatives, Fract. Calc. Appl. Anal., 12 (2009), 299–318.
    [51] T. Sandev, Z. Tomovski, B. Crnkovic, Generalized distributed order diffusion equations with composite time fractional derivative, Comput. Math. Appl., 73 (2017), 1028–1040. https://doi.org/10.1016/j.camwa.2016.07.009 doi: 10.1016/j.camwa.2016.07.009
    [52] R. K. Saxena, A. M. Mathai, H. J. Haubold, Space–time fractional reaction-diffusion equations associated with a generalized Riemann–Liouville fractional derivative, Axioms, 3 (2014), 320–334. https://doi.org/10.3390/axioms3030320 doi: 10.3390/axioms3030320
    [53] R. K. Saxena, Z. Tomovski, T. Sandev, Fractional Helmholtz and fractional wave equations with Riesz-Feller and generalized Riemann-Liouville fractional derivatives, Eur. J. Pure Appl. Math., 7 (2014), 312–334. http://www.ejpam.com/index.php/ejpam/article/view/2176
    [54] Z. Tomovski, Generalized Cauchy type problems for nonlinear fractional differential equations with composite fractional derivative operator, Nonlinear Anal., 75 (2012), 3364–3384. https://doi.org/10.1016/j.na.2011.12.034 doi: 10.1016/j.na.2011.12.034
    [55] Z. Tomovski, T. Sandev, R. Metzler, J. Dubbeldam, Generalized space-time fractional diffusion equation with composite fractional time derivative, Phys. A., 391 (2012), 2527–2542. 10.1016/j.physa.2011.12.035 doi: 10.1016/j.physa.2011.12.035
    [56] Z. Tomovski, T. Sandev, Distributed-order wave equations with composite time fractional derivative, Int. J. Comput. Math., 95 (2018), 1100–1113. https://doi.org/10.1080/00207160.2017.1366465 doi: 10.1080/00207160.2017.1366465
    [57] A. A. Kilbas, H. M. Srivastava, J. J. Trujillo, Theory and Applications of Fractional Differential Equations, North-Holland Mathematics Studies, Elsevier, Amsterdam, 2006.
    [58] S. G. Samko, A. A. Kilbas, O. I. Marichev, Fractional Integrals and Derivatives: Theory and Applications, Gordon and Breach, New York, 1993.
    [59] E. C. Titchmarsh, Introduction to the Theory of Fourier Integrals, Clarendon Press, Oxford, 1937.
    [60] A. A. Kilbas, M. Saigo, H-transforms. Theory and applications, Analytical Methods and Special Functions, Chapman & Hall/CRC, Boca Raton, 2004.
    [61] R. Gorenflo, A. A. Kilbas, F. Mainardi, S. V. Rogosin, Mittag-Leffler Functions, Related Topics and Applications, 2nd extended and updated edition, Springer Monographs in Mathematics, Springer, Berlin, 2020.
    [62] M. Ferreira, N. Vieira, Fundamental solutions of the time fractional diffusion-wave and parabolic Dirac operators, J. Math. Anal. Appl., 447 (2017), 329–353. https://doi.org/10.1016/j.jmaa.2016.08.052 doi: 10.1016/j.jmaa.2016.08.052
    [63] A. P. Prudnikov, Y. A. Brychkov, O. I. Marichev, Integrals and series. Volume 5: Inverse Laplace transforms, Gordon and Breach Science Publishers, New York etc., 1992.
    [64] R. Garrapa, Numerical evaluation of two and three parameter Mittag-Leffler functions, SIAM J. Numer. Anal., 53 (2015), 1350–1369. https://doi.org/10.1137/140971191 doi: 10.1137/140971191
    [65] N. Vieira, M. Ferreira, M. M. Rodrigues, Time-fractional telegraph equation with ψ-Hilfer derivatives, Comput. Appl. Math., 41 (2022).
  • This article has been cited by:

    1. Ravshan Ashurov, Rajapboy Saparbayev, Time-dependent identification problem for a fractional Telegraph equation with the Caputo derivative, 2024, 27, 1311-0454, 652, 10.1007/s13540-024-00240-0
    2. Mohammad Hossein Derakhshan, Stability analysis of difference-Legendre spectral method for two-dimensional Riesz space distributed-order diffusion-wave model, 2023, 144, 08981221, 150, 10.1016/j.camwa.2023.05.035
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3138) PDF downloads(143) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog