Return to Article Details A stable finite difference scheme for fractional viscoelastic wave propagation in space-time domain

A stable finite difference scheme for fractional viscoelastic wave propagation in space-time domain

Mohamed Ait Ichou and Abdelaaziz Ezziani
(Date: December 29, 2024; accepted: June 23, 2025; published online: June 30, 2025.)
Abstract.

This work focuses on the numerical approximation of the one-dimensional wave propagation equation in a viscoelastic medium modeled by the fractional Zener model. First, we propose an explicit centered finite difference scheme. Then, we prove a sufficient stability condition for this scheme using an energy technique. Furthermore, we demonstrate that this condition is numerically necessary. Finally, we present various numerical simulations to validate the proposed scheme.

Key words and phrases:
Finite-difference method, energy decay, fractional derivative, viscoelastic waves Zener’s model, stability condition, energy technique.
2005 Mathematics Subject Classification:
35L05, 65N06, 35R11, 74D05.
EMA Team, LRST Laboratory, ESEFA, Ibnou Zohr University of Agadir, Morocco, e-mail: moha.aitichou@gmail.com.
Laboratory MAEGE FSJES Aïn Sebaâ, Hassan II University of Casablanca, Morocco.

1. Introduction

This work is devoted to the mathematical modeling of wave propagation in dissipative media based on the fractional Zener model. A related study addressing the propagation of viscoelastic waves using an integer-order derivative was previously carried out in [1]. The study of these phenomena is important in the fields of seismology and geophysics, as well as in other applications such as medical imaging [2] and polymer modeling [3, 4].

In their current research, the researchers have demonstrated that mathematical models constructed using fractional operators are often more precise and reliable than those constructed using integer-order operators. Fractional-order models have a memory property that allows more historical knowledge to be incorporated, enabling them to predict and translate models more accurately. These concepts have found applications in the work of mathematicians and physicists, as documented in references [5, 6, 7, 8, 9].

Fractional derivatives are an excellent tool for describing the memory and hereditary properties of various materials and processes [10, 11]. There are many different approaches to defining fractional derivatives. The most commonly used are the Riemann-Liouville approach [12] and the Caputo approach [13]. In our study, we use the latter because it is more suitable for numerical simulation [14, 7].

Numerous authors have explored the numerical analysis of such problems. Konjik et al. [15] investigated the one-dimensional fractional viscoelastic wave equation using an explicit finite difference scheme, though without a comprehensive stability analysis. Xu and Stewart [16] also studied viscoelastic wave propagation, but did not detail the numerical method employed. Other contributions, such as those by Brono and Moczo [17, 18], focused on frequency-domain simulations, which are generally less appropriate for capturing the transient dynamics inherent to time-dependent problems. The Grünwald–Letnikov approximation has been widely used for discretizing Caputo derivatives in time (see, e.g., [19, 20]). However, this method can be computationally expensive for long time simulations due to the storage of historical data and the lack of energy-based stability analysis. Haddar et al. [21] proposed a method based on diffusive representations, offering strong modeling capabilities but involving significant analytical complexity. Afshari et al. [22] developed a finite difference scheme for a spatio-temporal fractional wave model and established its unconditional stability and convergence; however, their formulation does not account for relaxation parameters that are essential for accurately modeling viscoelastic media.

In recent years, several advanced numerical techniques have been proposed to solve time-fractional wave equations with improved accuracy and computational efficiency. Dehghan and Abbaszadeh [23] introduced a meshless collocation method based on radial basis functions (RBFs), which avoids the complexities of grid generation and is particularly well-suited for problems in irregular domains. More recently, Wang et al. [24] proposed a fast and accurate algorithm designed to reduce the computational cost associated with memory effects in time-fractional models, while maintaining the desired accuracy and stability properties. These contributions highlight the growing interest in efficient numerical methods for fractional wave equations and reinforce the relevance of our approach.

In this work, we address these limitations by proposing a simple and efficient explicit finite difference scheme for the space-time fractional viscoelastic wave equation in one spatial dimension. Unlike Grünwald–Letnikov-type methods, we reformulate the Caputo derivative using a weakly singular integral that we approximate via the trapezoidal rule, allowing for a compact and accurate implementation. Furthermore, we prove a discrete energy decay result and establish a sufficient stability condition, which closely matches that of the integer-order case. Our approach also includes the influence of relaxation parameters τ0 and τ1, which are not accounted for in several earlier works. Numerical experiments illustrate the impact of the fractional order α and the relaxation parameters on wave attenuation.

2. Model problem

We consider the fractional Zener model for waves propagation in dissipative media (see [25]):

(1) ρ2ut2(x,t)σx(x,t)=f(x,t),(x,t)×]0,T],σ(x,t)+τ0ασtα(x,t)=μ[ux(x,t)+τ1αtα(ux(x,t))],(x,t)×]0,T],u(x,0)=u0(x),ut(x,0)=u1(x),σ(x,0)=μux(x,0),x.

Where

  • Displacement field u(x,t): Represents movement of particles in the medium at position x and time t.

  • σ is the stress, represents the internal forces per unit area within the material caused by deformation.

  • f is the source terms or forcing functions: External inputs or forces applied to the system.

  • μ is the dynamic viscosity coefficient, represents the measures the material’s resistance to shear deformation.

  • ρ Represents the mass per unit volume of the medium.

  • τ0 and τ1 are the relaxation parameters, characterizing how quickly the material responds to and recovers from stress or deformation. (with the ratio θ=τ1/τ0>1 which guarantees the dissipation condition see [25]).

For all 0<α<1, the Caputo derivative of order α (see [12]) is defined as follows:

dαgdtα(t)=1Γ(1α)0t1(tτ)αgτ(τ)𝑑τ,

where Γ(1α) is the classical gamma function.

We rewrite the model problem (1) in a form that is more adaptable for numerical approximation. To do this, we introduce the following variable s=σμxu (the difference between the viscoelastic stress and the purely elastic one: τ0=τ1=0). Then, by deriving the second equation from the β-derivative with with β=1α, we obtain:

(2) ρ2ut2(x,t)σx(x,t)=f(x,t),(x,t)×]0,T],βtβs(x,t)+τ0st(x,t)μ(τ1τ0)t(ux(x,t))=0,(x,t)×]0,T],σ(x,t)=s(x,t)+μux(x,t),(x,t)×[0,T],u(x,0)=u0(x),ut(x,0)=u1(x),s(x,0)=0,x.

The following equation is added for the approach of βtβ(see [25]):

(3) φt(x,t,ξ)=ξφ(x,t,ξ)+s(x,t),(x,t,ξ)×]0,T]×[0,+[,φ(x,0,ξ)=0,
(4) βstβ(x,t)=0+φt(x,t,ξ)𝑑Mα(ξ),

with dMα(ξ)=sin(απ)πξαdξ.

Using the equations (3) and (4), we can rewrite the model problem as follows:

(5) ρ2ut2(x,t)σx(x,t)=f(x,t), (x,t)×]0,T],
τ0st(x,t)+0+φt(x,t,ξ)𝑑Mα(ξ)
μ(τ1τ0)t(ux(x,t))=0, (x,t)×]0,T],
φt(x,t,ξ)=ξφ(x,t,ξ)+s(x,t), (x,t)×]0,T],ξ[0,+[,
σ(x,t)=s(x,t)+μux(x,t), (x,t)×[0,T],
u(x,0)=u0(x),ut(x,0)=u1(x),
s(x,0)=0,φ(x,0,ξ)=0, (x,ξ)×[0,+[.

3. Finite difference scheme

We use a numerical approximation scheme based on the solution of ordinary differential equations to approximate the model problem (5).

We use the trapezium rule to approximate the integral in the second equation of the system (5). We introduce a geometric grid on the ξ axis, defined by the lower limit ξm, the upper limit ξM and the number of discretisation points Nξ. We note that:

Δξ=ξMξmNξ1,
0+φt(x,t,ξ)𝑑Mα(ξ)ξmξMφt(x,t,ξ)𝑑Mα(ξ)j=1Nξwjφt(x,t,ξj),

with

w1=sin(απ)2πξmαΔξ,wNξ=sin(απ)2πξMαΔξ

and

wj=sin(απ)πξjαΔξ,j=2,,Nξ1.

The system (5) then becomes:

(6a) ρ2ut2(x,t)σx(x,t)=f(x,t),
(6b) τ0st(x,t)+j=1Nξwjφt(x,t,ξj)=μ(τ1τ0)2uxt(x,t),
(6c) φt(x,t,ξj)=ξjφ(x,t,ξj)+s(x,t),j=1,,Nξ,
(6d) σ(x,t)=s(x,t)+μux(x,t),
(6e) u(x,0)=u0(x),ut(x,0)=u1(x),s(x,0)=0,φ(x,0,ξj)=0,

3.1. Space discretization

We introduce a regular mesh of the domain Ω=[a,b] with edges h=baNx consisting of the segments [xi,xi+1],i=1,,Nx,

xi=a+(i1)h.
Refer to caption
Figure 1. Space discretization.

We approach the equation (6a) in the point xi and the equation (6b), (6c) and (6d) in the point xi+1/2=(xi+xi+1)/2. We propose the centered scheme:

(7a) ρd2ui(t)dt2σi+12(t)σi12(t)h=fi(t),
(7b) τ0dsi+12(t)dt+j=1Nξwjdφi+12,j(t)dt=μ(τ1τ0)(ddt(ui+1(t)ui(t)h)),
(7c) dφi+12,j(t)dt=ξjφi+12,j(t)+si+12(t),j=1,,Nξ,
(7d) σi+12(t)=si+12(t)+μui+1(t)ui(t)h,
(7e) ui(0)=u0(xi),duidt(0)=u1(xi),si+12(0)=0,φi+12,j(0)=0,

with ui(t)u(xi,t), σi+12(t)σ(xi+1/2,t), si+12(t)s(xi+1/2,t), φi+12,j(t)φ(xi+1/2,t,ξj) and fi(t)f(xi,t).

3.2. Space-time discretization

We consider a regular time grid with a step size of Δt and we set

tn=nΔt,tn+1/2=tn+Δt2,n.

In order to approximate the system (7), we use a centered finite difference scheme in time by approaching the first equation at tn, the second and third equations at tn+1/2, and the last equation at tn. The resulting centred scheme is as follows:

(8a) ρuin+12uin+uin1Δt2σi+12nσi12nh=fin,
(8b) τ0si+12n+1si+12nΔt+j=1Nξwjφi+12,jn+1φi+12,jnΔt=
(8c) μ(τ1τ0)h(ui+1n+1ui+1nΔtuin+1uinΔt),
(8d) φi+12,jn+1φi+12,jnΔt=ξjφi+12,jn+1+φi+12,jn2+si+12n+1+si+12n2,j=1,,Nξ,
(8e) σi+12n=si+12n+μui+1nuinh,
(8f) ui0=u0(xi),ui1=ui0+Δtu1(xi)+Δt22ρ(fi0+σi+120σi120h),
(8g) si+120=0,φi+1/2,j0=0,

with uinu(xi,tn), σi+12nσ(xi+1/2,tn), si+12ns(xi+1/2,tn), φi+12,jnφ(xi+1/2,tn,ξj) and finf(xi,tn).

By using the equation (8d), we have

φi+12,jn+1=2ξjΔt2+ξjΔtφi+12,jn+Δt2+ξjΔt(si+12n+1+si+12n),

after injecting this last equality into (8b), we obtain:

ρuin+12uin+uin1Δt2σi+12nσi12nh=fin,
τ0si+12n+1si+12nΔt+λ(si+12n+1+si+12n)j=1Nξwj~φi+12,jn=μ(τ1τ0)h(ui+1n+1ui+1nΔtuin+1uinΔt),
φi+12,jn+1=2ξjΔt2+ξjΔtφi+12,jn+Δt2+ξjΔt(si+12n+1+si+12n),j=1,,Nξ,
σi+12n=si+12n+μui+1nuinh,
ui0=u0(xi),ui1=ui0+Δtu1(xi)+Δt22ρ(fi0+σi+120σi120h),si+120=0,φi+1/2,j0=0,

with

wj~=ξjwj22+ξjΔtandλ=j=1Nξwj2+ξjΔt.

Finally, we obtain the fully explicit scheme:

uin+1=Δt2ρ(fin+σi+12nσi12nh)+2uinuin1,
si+12n+1=Δtτ0+λΔt(μ(τ1τ0)h[ui+1n+1ui+1nΔtuin+1uinΔt]+j=1Nξwj~φi+12,jn)λΔtτ0λΔt+τ0si+12n,
φi+12,jn+1=2ξjΔt2+ξjΔtφi+12,jn+Δt2+ξjΔt(si+12n+1+si+12n),j=1,,Nξ,
σi+12n=si+12n+μui+1nuinh,
ui0=u0(xi),ui1=ui0+Δtu1(xi)+Δt22ρ(fi0+σi+120σi120h),
si+120=0,φi+1/2,j0=0.

We note that the number of unknowns can be reduced by replacing the expression for σi+12n in the first equation of the final system. This yields:

uin+1= Δt2ρ(fin+μui+1n2uin+ui1nh2+si+12nsi12nh)+2uinuin1,
si+12n+1= Δtτ0+λΔt(μ(τ1τ0)h[ui+1n+1ui+1nΔtuin+1uinΔt]+j=1Nξwj~φi+12,jn)λΔtτ0λΔt+τ0si+12n,
φi+12,jn+1= 2ξjΔt2+ξjΔtφi+12,jn+Δt2+ξjΔt(si+12n+1+si+12n),j=1,,Nξ,
ui0= u0(xi),ui1=ui0+Δtu1(xi)+Δt2ρ(fi0+σi+120σi120h),
si+120= 0,φi+1/2,j0=0.

3.3. Discrete energy and stability analysis

We will use an energy technique to prove the stability of the scheme (see (3.2)). We introduce discrete normed spaces.

(9) Lh,02={uh=(ui)i,i|ui|2<+},
(10) Lh,1/22={σh=(σi+1/2)i,i|σi+1/2|2<+}

With the scalar products:

(uh,u~h)0 =hiuiu~i,(σh,σ~h)1/2=hiσi+1/2σ~i+1/2,(φh,j,φ~h,j)1/2
=hiφi+1/2,jφ~i+1/2,j,j=1,,Nξ,

where

uh=(ui)iLh,02,σh=(σi+1/2)iLh,1/22
φh,j=(φi+1/2,j)iLh,1/22,j=1,,Nξ,

and associated norm:

uh02 =hi|ui|2,σh1/22=hi|σi+1/2|2,
φh,j1/22 =hi|φi+1/2,j|2,j=1,,Nξ.

We recall that the results for energy decay in the continuous case are given by (see [25]):

E(t)= 12(ρ|ut|2dx+μ|ux|2dx+τ0μ(τ1τ0)|s|2dx
(11) +1μ(τ1τ0)0+ξ|φ|2dMα(ξ)dx),

and it satisfies:

dE(t)dt=(f,ut)1μ(τ1τ0)0+|sξφ|2𝑑Mα(ξ)𝑑x,

with

s=σμux.

We write the system (3.2) in vector form:

(12) ρuhn+12uhn+uhn1Δt2Bσhn=0,τ0μ(τ1τ0)shn+1shnΔt+1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt+Buhn+1uhnΔt=0,j=1,,Nξ,φh,jn+1φh,jnΔt=ξjφh,jn+1+φh,jn2+shn+1+shn2,j=1,,Nξ,μ1σhnμ1shn+Buhn=0.

with

(13) {B:Lh,1/22Lh,02σh(σi+1/2σi1/2h)j
Lemma 1.

The norm of the operator B satisfies the inequality:

(14) B2h.
Proof.
(Bσh,uh)0=(σh,Buh)1/2=1hi𝕫hσi+1/2(ui+1ui).1hi𝕫hσi+1/2(ui+1ui).1h(i𝕫h|σi+1/2|2)1/2(i𝕫h|ui+1ui|2)1/2.

However, as i𝕫h|ui+1ui|2 verifies the inequality:

i𝕫h|ui+1ui|2=i𝕫h|ui+1|22i𝕫huiui+1+i𝕫h|ui|2,2i𝕫h|ui+1|2+2i𝕫h|ui|2=4uh02,

it implies that

(Bσh,uh)2huh0σh1/2,

and we get

B2h.

Definition 2.

We define the discrete energy by:

En+1/2= ρ2uhn+1uhnΔt02+μ2(Bun+1,Bun)0
+τ04μ(τ1τ0)[shn+11/22+shn1/22]
(15) +Δt24(Buhn+1uhnΔt,shn+1shnΔt)1/2
+14μ(τ1τ0)j=1Nξwjξj[φh,jn+11/22+φh,jn1/22].
Remark 3.

The discrete energy can be decomposed as the sum of four terms.

  • -

    The first term, ρ2uhn+1uhnΔt02+12(Bun+1,μBun)1/2 corresponds to the classical discrete energy in the purely elastic case (i.e., when τ0=τ1=0), which corresponds to the approximation of the first and second part of the continuous energy (3.3).

  • -

    The second term, τ04μ(τ1τ0)[shn+11/22+shn1/22] due to the viscoelasticity, which approaches the third parts of the continuous energy.

  • -

    The third term, Δt22(Buhn+1uhnΔt,shn+1shnΔt)1/2 is a small term of order 𝒪(Δt2), which is due to the finite difference approximation.

  • -

    The last term, 14μ(τ1τ0)j=1Nξwjξj[φh,jn+11/22+φh,jn1/22] depends on α the fractional derivative order corresponds to the approximation of the last part of the continuous energy.

The following theorem proves an energy decay result.

Theorem 4.

The discrete energy verifies:

(16) En+1/2En1/2Δt=12μ(τ1τ0)j=1Nξwj[φh,jn+1φh,jnΔt1/22+φh,jnφh,jn1Δt1/22].
Proof.

We use the variational formulation associated with the discrete scheme (12):

(ρuhn+12uhn+uhn1Δt2,u~h)(Bσhn,u~h)=0,
(17) (τ0μ(τ1τ0)shn+1shnΔt,s~h)+(1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt,s~h)+(Buhn+1uhnΔt,s~h)=0,
φh,jn+1φh,jnΔt=ξjφh,jn+1+φh,jn2+shn+1+shn2,
μ1σhnμ1shn+Buhn=0.

If we take u~h=uhn+1uhn12Δt (the centered approximation of duhdt in time tn) and s~h=shn+1+shn2 (the centered approximation of sh in time tn+1/2). Equation (17) becomes:

(18) (ρuhn+12uhn+uhn1Δt2,uhn+1uhn12Δt)(Buhn+1uhn12Δt,σhn)=0,(τ0μ(τ1τ0)shn+1shnΔt,shn+1+shn2)+(1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt,shn+1+shn2)++(Buhn+1uhnΔt,shn+1+shn2)=0,φh,jn+1φh,jnΔt+ξjφh,jn+1+φh,jn2=shn+1+shn2,μ1σhnμ1shn+Buhn=0.

Substituting shn+1+shn2 into equation (3) and then into equation (2) gives:

(19) i)(ρuhn+12uhn+uhn1Δt2,uhn+1uhn12Δt)(Buhn+1uhn12Δt,σhn)=0,ii)(τ0μ(τ1τ0)shn+1shnΔt,shn+1+shn2)+(Buhn+1uhnΔt,shn+1+shn2)++(1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt,φh,jn+1φh,jnΔt+ξjφh,jn+1+φh,jn2)=0.

which becomes:

(20) i)(ρuhn+12uhn+uhn1Δt2,uhn+1uhn12Δt)0(Buhn+1uhn12Δt,σhn)1/2=0,ii)(τ0μ(τ1τ0)shn+1shnΔt,shn+1+shn2)1/2+(1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt,φh,jn+1φh,jnΔt)1/2++(1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt,ξjφh,jn+1+φh,jn2)1/2+(Buhn+1uhnΔt,shn+1+shn2)1/2=0.

We get:

i)ρ2Δt[uhn+1uhnΔt02uhnuhn1Δt02](Buhn+1uhn12Δt,σhn)1/2=0.
ii)τ02Δtμ(τ1τ0)[shn+11/22shn1/22]+1μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt1/22++12Δtμ(τ1τ0)j=1Nξwjξj[φh,jn+11/22φh,jn1/22]+(Buhn+1uhnΔt,shn+1+shn2)1/2=0.

Using the last equation of (18), we obtain:

(Buhn+1uhn12Δt,σhn)1/2=(Buhn+1uhn12Δt,σhnshn)1/2+(Buhn+1uhn12Δt,shn)1/2=12Δt[(Buhn+1,μBuhn)1/2(Buhn1,μBuhn)1/2]+(Buhn+1uhn12Δt,shn)1/2.

We use that σhnshn=μBuhn, which allows us to rewrite (i) in the form:

(21) ρ2Δt[uhn+1uhnΔt02uhnuhn1Δt02]++12Δt[(Buhn+1,μBuhn)0(Buhn1,μBuhn)0](Buhn+1uhn12Δt,shn)1/2=0.

If we average the equation (ii) between the two moments tn+12 and tn12, we obtain:

(22) τ04Δtμ(τ1τ0)[shn+11/22shn1/22]+τ04Δtμ(τ1τ0)[shn1/22shn11/22]+12μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt1/22+12μ(τ1τ0)j=1Nξwjφh,jnφh,jn1Δt1/22+14Δtμ(τ1τ0)j=1Nξwjξj[φh,jn+11/22φh,jn1/22]+14Δtμ(τ1τ0)j=1Nξwjξj[φh,jn1/22φh,jn11/22]+12(Buhn+1uhnΔt,shn+1+shn2)1/2+12(Buhnuhn1Δt,shn+shn12)1/2=0.

Equation (22) is equivalent to:

(23) τ04Δtμ(τ1τ0)[shn+11/22+shn1/22]τ04Δtμ(τ1τ0)[shn1/22+shn11/22]
+12μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt1/22+12μ(τ1τ0)j=1Nξwjφh,jnφh,jn1Δt1/22
+14Δtμ(τ1τ0)j=1Nξwjξj[φh,jn+11/22+φh,jn1/22]
14Δtμ(τ1τ0)j=1Nξwjξj[φh,jn1/22+φh,jn11/22]
+12(Buhn+1uhnΔt,shn+1+shn2)1/2+12(Buhnuhn1Δt,shn+shn12)1/2 =0.

We decompose the last two terms of this formula as follows:

(24) 12(Buhn+1uhnΔt,shn+1+shn2)1/2=12(Buhn+1uhnΔt,shn+1shn2)1/2+12(Buhn+1uhnΔt,shn)1/2.12(Buhnuhn1Δt,shn+shn12)1/2=12(Buhnuhn1Δt,shnshn12)1/2+12(Buhnuhn1Δt,shn)1/2.

Substitute (24) in (23), we get

(25) (Buhn+1uhn12Δt,shn)1/2=12Δt(T1n+1/2T1n1/2)+12Δt(T2n+1/2T2n1/2)+14Δt(T3n+1/2T3n1/2)+Tn,

with

T1n+1/2= τ02μ(τ1τ0)[shn+11/22+shn1/22],
T2n+1/2= 12(Buhn+1uhn,shn+1shn)1/2,
T3n+1/2= 12μ(τ1τ0)j=1Nξwjξj[φh,jn+11/22+φh,jn1/22],
Tn= 12μ(τ1τ0)j=1Nξwjφh,jn+1φh,jnΔt1/22+12μ(τ1τ0)j=1Nξwjφh,jnφh,jn1Δt1/220.

After substitution of the quantity (25) in (21), we will find (4):

En+1/2En1/2Δt=Tn0,

which is the desired identity. ∎

Thanks to Theorem 4, in order to establish a sufficient stability condition it suffices to show that the energy En+1/2 is a positive quadratic form (see [26]). The same condition applies as in the integer derivative model discussed in [1] because the final term in the energy En+1/2 does not affect the positivity of the form.

Theorem 5.

The numerical scheme (12) is L2 stable under the sufficient CFL condition:

(26) Δt1ch,

where c=cτ1τ0, which corresponds to the velocity of viscolastic wave s in height frequency and c=μρ, which corresponds to the velocity of viscolastic waves in low frequency.

Proof.

We look for a condition in which the discrete energy is positive. To achieve this, we use the equivalent vector form of En+1/2:

(27) En+1/2=ρ2uhn+1uhnΔt02+μ2(Bun+1,Bun)0+τ04μ(τ1τ0)[shn+11/22+shn1/22]+Δt24(Buhn+1uhnΔt,shn+1shnΔt)1/2+14μ(τ1τ0)j=1Nξwjξj[φh,jn+11/22+φh,jn1/22].

As the following quantities verify:

{shn+11/22+shn1/22=12(shn+1+shn1/22+shn+1shn122)(Buhn+1,Buhn)0=14B(uhn+1+uhn)0214B(uhn+1uhn)02,

we can rewrite En+1/2 in the form: En+1/2=E1n+1/2+E2n+1/2, where

E1n+1/2= μ8Buhn+1+uhn202
+14μ(τ1τ0)(τ02shn+1+shn21/22+j=1Nξwjξj[φh,jn+11/22+φh,jn1/22])0.
𝑬2n+1/2= ρ2uhn+1uhnΔt02μΔt28Buhn+1uhnΔt02+τ0μ(τ1τ0)Δt28shn+1shnΔt1/22
+Δt24(Buhn+1uhnΔt,shn+1shnΔt)1/2.

Moreover, the quantity E2n+1/2 satisfies

𝑬2n+1/2
12([ρIμΔt24BB]uhn+1uhnΔt,uhn+1uhnΔt)+τ0μ(τ1τ0)Δt28shn+1shnΔt1/22
μ(τ1τ0)τ0Δt28(BBuhn+1uhnΔt,uhn+1uhnΔt)0τ0μ(τ1τ0)Δt28shn+1shnΔt1/22,
= 12(ρuhn+1uhnΔt,uhn+1uhnΔt)(μ+μ(τ1τ0)τ0)Δt28(BBuhn+1uhnΔt,uhn+1uhnΔt),
= 12(ρuhn+1uhnΔt,uhn+1uhnΔt)μτ1τ0Δt28(BBuhn+1uhnΔt,uhn+1uhnΔt).

The positivity of En+1/2 is ensured under the inequality

μτ1τ0Δt24(BBuh,uh)0(ρuh,uh)0,uh.

this is equivalent to:

(μτ1τ0Δt2)2(Buh,Buh)0 (ρuh,uh)0,uh,
(μτ1τ0Δt2)2B2uh02 ρuh02,
(μτ1τ0Δt2)2B2 ρ,

we use Lemma 1, we get the stability sufficient condition:

Δth/c,withc=cτ1τ0 and c=μρ.

Remark 6.

We observe that:

  • The stability condition (26) is also necessary. Indeed, if we set Δt=hc(1+103) the numerical scheme will be unstable.

4. Numerical simulation

We validate the numerical scheme suggested in the previous section, since we have an exact solution for a particular choice of initial conditions. We consider the model problem in the segment ]0,1[ with the following data:

(28) ρ2ut2σx=0,(x,t)]0,1[×]0,T],σ+τ0ασtα=μ[ux+τ1αtα(ux)],(x,t)]0,1[×]0,T],u(x,0)=sin(πx),ut(x,0)=0,σ(x,0)=μτ1τ0πcos(πx),x]0,1[,u(0,t)=u(1,t)=0t[0,T]

and the coefficients:

ρ=1, μ=1,τ0=1, τ1=1.2.

The exact solution (u,σ) of the problem:

(29) {u(x,t)=U(t)sin(πx),σ(x,t)=Σ(t)πcos(πx).

Substituting (29) in (28) we find U is solution of:

(30) dα+2Udtα+2+dUdt+π2τ1dαUdtα+π2U=0U(0)=1,dU(0)dt=0.

The solution of this system is of the form:

U(t)=eηt(cos(ωt)ηωsin(ωt)),

where ηiω is the solution of:

Xα+2+X2+τ1π2Xα+π2=0,

For the computation of the numerical solution, we use h=0.1 and ensure that the stability condition (Δt=h/c) is satisfied.

Fig. 2 shows a comparison of the exact solution (uex) and the numerical solution (unum) at the point x=0.5 as a function of time for different values of α.

Refer to caption Refer to caption
α=0.3 α=0.6
Refer to caption Refer to caption
α=0.7 α=0.9
Figure 2. The numerical and the exact solutions as a function of time for different α.

We observe that the numerical solution coincides with the exact solution for all values of α. The figure also shows that, for a large value of α, the waves are more damped.

Numerical stability: The stability condition is given by CFL=1c=1cτ1/τ0, c, where c denotes the velocity of a high-frequency wave and CFL is the Courant–Friedrichs–Lewy number.

For numerical stability, if we perturb the stability condition to be CFL+105 and even by changing the value of α, we will still have Fig. 3:

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3. The numerical solution as a function of time following the perturbation is stability 105.

Fig. 4 shows the evolution of the numerical error, uexaunum as a function of step h for two values of the parameter α, namely α=0.5 (left-hand panel) and α=0.9 (right-hand panel).

h uexaunum
101 3.2083 1003
5 102 7.1497 1004
2.5 102 1.7204 1004
1.25 102 4.2436 1005
6.25 103 1.0553 1005
3.125 103 2.6326 1006
h uexaunum
101 1.629 1003
5 102 3.0609 1004
2.5 102 6.8340 1005
1.25 102 1.6344 1005
6.25 103 4.01139 1006
3.125 103 9.94608 1007
Figure 4. Error for α=0.5 and α=0.9.

We observe that the error decreases steadily as h decreases, and that for the same step, the error is smaller when alpha is larger.

To study the convergence of our scheme, we change the discretisation parameters h with Δt=hc. Fig. 6 shows the L-norm of the corresponding error with respect to the space step h (on a log-log scale). Fig. 5Fig. 6 shows that there is convergence and that the error is of order 2, regardless of the values of α.

Refer to caption
Figure 5. Convergence for value of α=0.4; 0.5; 0.6.
Refer to caption
Figure 6. Convergence for value of α=0.7; 0.8; 0.9.

We conduct an experiment in the domain [5,5] with a point source located at the centre of the domain and verify:

f(t)={2π2f02(tt0)eπ2f02(tt0)2, if t2t00, otherwise. 
t0=1f0,f0=1 central frequency ,

with the following data:

u(x,0)=0,ut(x,0)=0,σ(x,0)=0,x]5,5[,
u(5,t)=u(5,t)=0,t[0,T],

and the coefficients:

ρ=1, μ=1.

The parameters of the discretization are h=5.102 and Δt=hc satisfies the stability condition (26).

To demonstrate the impact of relaxation parameters, such as τ0 and τ1, on the attenuation and propagation velocity of fractional viscoelastic waves, we present a numerical solution in Fig. 7, where we vary the ratio, θ=τ1/τ0, with values of: 4, 2 and 1.2.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7. Damping and propagation quality with α=0.8.

We can see that absorption and velocity of propagation are related to the ratio θ: the larger the value of θ, the greater the damping and the faster the velocity of propagation.

Fig. 8 shows that, after the source f is extinguished, the viscoelastic energy decreases exponentially, which confirms our predictions (see (4)).

Refer to caption
Figure 8. Discrete energy with θ=1.2 and α=0.5.

Fig. 9 shows the influence of the ratio θ on the discrete energy.

Refer to caption Refer to caption
θ=1.8 θ1
Figure 9. Discrete energy with α=0.5.

We observe that, when θ=1.8, the discrete energy converges more quickly to zero. However, when θ=1.0001 (θ1), energy conservation occurs and the viscoelastic model converges to the elastic one.

4.1. Parameter Selection for Diffusive Representation

The diffusive representation introduces a continuous variable, denoted by the symbol ξ, which must be discretised over the interval [ξm,ξM] with Nξ points. Through systematic numerical experimentation (see Fig. 10), we determine the optimal parameter ranges.

  • ξm0.01/T (characteristic timescale)

  • ξM10/Δt (inverse time step)

  • Nξ=50-100 points (logarithmic spacing)

Refer to caption
Figure 10. Sensitivity analysis showing the effect of (a) ξm, (b) ξM, and (c) Nξ on solution accuracy. Errors plateau when ξm102, ξM103, and Nξ50.

Fig. 10 illustrates the effect that diffusive scheme parameters have on the accuracy of the solution, as measured by the relative L error. Three clear trends emerge.

  1. (1)

    Sensitivity to ξm (see Fig. 10. on the left):

    • Error decreases exponentially as ξm0.

    • Stabilises for ξm102 (error <104 ).

    • Confirms that the slow dynamics must be captured by the value of the parameter.

  2. (2)

    Influence of ξM (middle figure in Fig. 10):

    • High-frequency truncation: Errors exceed 102 when ξM<102 due to loss of spectral resolution in the fractional kernel tα

    • Convergence threshold: Numerical convergence (error <105) is achieved for ξM103/Δt, where Δt is the time step

    • Optimal choice: Our selection ξM=103/Δt ensures:

      (31) ξMsin(απ)πξα𝑑ξ<0.1% of total weight
  3. (3)

    Effect of Nξ (see Fig 10. on the right):

    • Spectral convergence is typical of fractional methods.

    • Negligible gain beyond 50 Nξpoints.

    • An error platform at approximately 5×105.

5. Conclusion

In this study, we developed and analysed a second-order, explicit, finite-difference scheme for a fractional wave propagation model. Using energy methods, we derived a stability condition that is consistent with the classical (integer-order) case. Our analysis yielded a criterion that is both theoretically sufficient and numerically precise. A variety of numerical experiments demonstrated the impact of the fractional order α and the relaxation parameters τ0 and τ1 on wave attenuation. Future work will focus on extending the method to higher dimensions and improving the efficiency of adaptive time-stepping schemes. Furthermore, this approach can be applied to more complex models in poroelastic media, in which fractional effects are crucial for capturing memory and dissipation phenomena.

Data Availability

No data were used to support this study.

References