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

Averaging principle on infinite intervals for stochastic ordinary differential equations

  • Received: 01 August 2020 Revised: 01 January 2021 Published: 22 February 2021
  • 34C29, 60H10, 37B20, 34C27

  • In contrast to existing works on stochastic averaging on finite intervals, we establish an averaging principle on the whole real axis, i.e. the so-called second Bogolyubov theorem, for semilinear stochastic ordinary differential equations in Hilbert space with Poisson stable (in particular, periodic, quasi-periodic, almost periodic, almost automorphic etc) coefficients. Under some appropriate conditions we prove that there exists a unique recurrent solution to the original equation, which possesses the same recurrence property as the coefficients, in a small neighborhood of the stationary solution to the averaged equation, and this recurrent solution converges to the stationary solution of averaged equation uniformly on the whole real axis when the time scale approaches zero.

    Citation: David Cheban, Zhenxin Liu. Averaging principle on infinite intervals for stochastic ordinary differential equations[J]. Electronic Research Archive, 2021, 29(4): 2791-2817. doi: 10.3934/era.2021014

    Related Papers:

    [1] Hongxia Lin, Sabana, Qing Sun, Ruiqi You, Xiaochuan Guo . The stability and decay of 2D incompressible Boussinesq equation with partial vertical dissipation. Communications in Analysis and Mechanics, 2025, 17(1): 100-127. doi: 10.3934/cam.2025005
    [2] Chunyou Sun, Junyan Tan . Attractors for a Navier–Stokes–Allen–Cahn system with unmatched densities. Communications in Analysis and Mechanics, 2025, 17(1): 237-262. doi: 10.3934/cam.2025010
    [3] Zhigang Wang . Serrin-type blowup Criterion for the degenerate compressible Navier-Stokes equations. Communications in Analysis and Mechanics, 2025, 17(1): 145-158. doi: 10.3934/cam.2025007
    [4] Haifa El Jarroudi, Mustapha El Jarroudi . Asymptotic behavior of a viscous incompressible fluid flow in a fractal network of branching tubes. Communications in Analysis and Mechanics, 2024, 16(3): 655-699. doi: 10.3934/cam.2024030
    [5] Andrea Brugnoli, Ghislain Haine, Denis Matignon . Stokes-Dirac structures for distributed parameter port-Hamiltonian systems: An analytical viewpoint. Communications in Analysis and Mechanics, 2023, 15(3): 362-387. doi: 10.3934/cam.2023018
    [6] Rui Sun, Weihua Deng . A generalized time fractional Schrödinger equation with signed potential. Communications in Analysis and Mechanics, 2024, 16(2): 262-277. doi: 10.3934/cam.2024012
    [7] Hilal Essaouini, Pierre Capodanno . Analysis of small oscillations of a pendulum partially filled with a viscoelastic fluid. Communications in Analysis and Mechanics, 2023, 15(3): 388-409. doi: 10.3934/cam.2023019
    [8] İrem Akbulut Arık, Seda İğret Araz . Delay differential equations with fractional differential operators: Existence, uniqueness and applications to chaos. Communications in Analysis and Mechanics, 2024, 16(1): 169-192. doi: 10.3934/cam.2024008
    [9] Xiao Qing Huang, Jia Feng Liao . Existence and asymptotic behavior for ground state sign-changing solutions of fractional Schrödinger-Poisson system with steep potential well. Communications in Analysis and Mechanics, 2024, 16(2): 307-333. doi: 10.3934/cam.2024015
    [10] Siegfried Carl . Quasilinear parabolic variational-hemivariational inequalities in RN×(0,τ) under bilateral constraints. Communications in Analysis and Mechanics, 2025, 17(1): 41-60. doi: 10.3934/cam.2025003
  • In contrast to existing works on stochastic averaging on finite intervals, we establish an averaging principle on the whole real axis, i.e. the so-called second Bogolyubov theorem, for semilinear stochastic ordinary differential equations in Hilbert space with Poisson stable (in particular, periodic, quasi-periodic, almost periodic, almost automorphic etc) coefficients. Under some appropriate conditions we prove that there exists a unique recurrent solution to the original equation, which possesses the same recurrence property as the coefficients, in a small neighborhood of the stationary solution to the averaged equation, and this recurrent solution converges to the stationary solution of averaged equation uniformly on the whole real axis when the time scale approaches zero.



    Fractional calculus, the study of derivatives and integrals of non-integer orders, has increasingly become a powerful tool in modeling physical phenomena that defy conventional descriptions under ideal conditions [51]. It addresses the limitations inherent in classical models by providing a more accurate representation of systems with anomalous properties. Despite a relatively slow historical progression, the past decade has seen a renaissance, with fractional-order models advancing alongside traditional integer-order counterparts.

    These fractional-order models, encompassing both ordinary and partial differential equations with non-integer order terms have been instrumental in advancing our understanding of a diverse array of physical phenomena. They have been successfully applied in fields ranging from signal processing and control theory to more complex systems in diffusion processes, thermodynamics, biophysics, and blood-flow dynamics, as well as in electrodynamic, electrochemical, and electromagnetic theories, not to mention their significant roles in continuum and statistical mechanics, along with dynamical systems, see References[5,12,48,56].

    Substantial progress in the numerical approximation of fractional-order models have been made, enhancing the precision and applicability of these models in computational simulations. Beyond the traditional finite difference schemes for fractional equations [11,54], noteworthy advancements have been made in the Finite Element Methods (FEM). For instance, the work by Acosta and Borthagaray [2] shed light on regularity results for the analytic solution of the fractional Poisson problem and delineated convergence rates for FEM approximations, with these findings further substantiated by computational examples in [1]. Additionally, Ervin and Roop [32] contributed to this field by employing the FEM approach to steady-state fractional advection dispersion equations,

    Da(p0Dβx+qxDβ1)Du+b(x)Du+c(x)u=f,

    where D represents the first-order spatial derivative, 0Dβx, xDβ1 represent the left and right fractional integral operators with 0β<1 and 0p, q1 satisfying p+q=1. The FEM with Galerkin framework was used in [33] to solve fractional diffusion wave equations. Jin et al. described the Petrov-Galerkin FEM for the fractional convection-diffusion equations [35], where they considered the 1D fractional boundary value problem

    {0Dαxu+bu+qu=finD=(0,1),u(0)=u(1)=0,

    where fL2(D) and 0Dαxu denotes Caputo fractional derivative of order α(3/2,2).

    A comprehensive review of numerical approximations for fractional models is available in [30] and the references cited therein. Recently, focussing on the polygonal approximation techniques, Zhang et al. [58] studied a local projection stabilization Virtual Element Method (VEM) for the time-fractional Burger's equation with higher Reynold's number. Recent attempts with the VEM approximation for time-fractional Partial Differential Equations (PDEs) [31,57,58] highlight the need to further develop the studies to more general nonlinear problems.

    Our research extends the VEM to such a time-fractional nonlinear partial differential equation, providing both theoretical analysis and validation through numerical experiments. In this work, we consider the 2D time-fractional nonlinear convection-diffusion equation

    {c0DαtuΔu+b(x)u+f(u)=g(x,t),(x,t)Ω×(0,T),u(x,0)=0,xΩ,u(x,t)=0,(x,t)Ω×(0,T), (1.1)

    where ΩR2 denotes a bounded polygon region with boundary Ω and the order of time derivative varies such that 0<α<1. The convective field is represented by b(x)=(b1(x),b2(x))T, the function f(u) encapsulates the nonlinear term or terms of interest, while g(x,t) is a forcing term. To show theoretical estimates, suitable regularities of the terms in (1.1) will be discussed in Section 3. According to [38], notation c0Dαt denotes the Caputo fractional derivative of order α and is defined as,

    c0Dαtu(t):=1Γ(1α)t0(ts)αu(s)sds.

    Background material on the VEM. The VEM is used for approximation of the solution of PDEs on general polygonal and polyhedral meshes. It is an extension of the finite element method and shares some similarities with it. Like the FEM, VEM also formulates the problem in a weak form, which involves multiplying the PDE by a test function and integrating by parts over the domain. The weak form allows for the use of piecewise smooth approximations and leads to a system of linear equations that can be solved numerically. Differently from FEM, the VEM introduces the concept of "virtual element spaces", which are sets of functions that approximate the exact solution of the PDEs at both the local (elementwise) and global level. Such functions are implicitly defined as the solution to a local PDE problem, and are never explicitly used to compute the coefficients of the global matrix. The global spaces are defined on polygonal or polyhedral meshes and are constructed to be compatible with the underlying mesh structure and possess some degree of global regularity. In depth examinations of the parallelism between the VEM and finite elements for polygonal/polyhedral meshes are provided in [25,45], and its relationship with BEM-influenced FEM approaches in [24]. The VEM is also a variational reinterpretation of the Mimetic Finite Difference (MFD) approach, specifically its nodal variant introduced in [15,17,22,42]. An extensive discussion on MFD is found in the comprehensive review article [41] and the book [20]. A major advantage of the VEM is its ability to handle arbitrary polygonal and polyhedral meshes. Unlike the "traditional" FEM, which typically requires triangular or quadrilateral elements in 2D and tetrahedral or hexahedral elements in 3D, the VEM can handle meshes composed of elements with any number of sides with an upper bound from theoretical level. This flexibility is particularly useful when dealing with complex geometries. The VEM is designed to provide stable and accurate solutions for a wide range of problems. The method incorporates stabilization techniques to handle challenges associated with irregular meshes and distorted elements. Additionally, the VEM can achieve high-order accuracy by incorporating polynomial approximation spaces of higher degree in the virtual element space definition. Initially introduced to address elliptic problems such as the Poisson equation, cf. [6,19], the VEM was later applied to linear and nonlinear diffusion, convection-diffusion, and convection-reaction-diffusion problems [3,4,10,14] the Stokes equations [13,16,43,44], the polyharmonic equation [9], and many other models. Meanwhile, the nonconforming formulation for diffusion problems was proposed in [29] as the finite element reformulation of [40] and later extended to general elliptic problems [26], Stokes problem [23], and the biharmonic equation [8,59]. A recent contributed book documents the state of the art of this methodology, cf. [7].

    Paper's outline. The structure of this paper is as follows. Section 2 outlines some preliminary and established results, the model problem with weak formulation framework is given in Section 3. Section 4 presents the discrete formulation, well-posedness and convergence analysis of a fully discrete VEM. Section 5 offers numerical examples that substantiate our theoretical findings. Section 6 concludes the paper with a summary of our work.

    The concept of fractional derivatives has been explored and delineated extensively within the mathematical literature. In this section, we recall some definitions and background results that will be useful in the paper. For a comprehensive exposition of these concepts, the reader is directed to refer [49,50].

    Let α be a positive real number, m a non-negative integer, and f(x) a univariate function defined on x[a,b]. The left and right hand Riemann-Liouville (R-L) fractional derivatives of f(x) of order α, with m1<αm, are defined as

    RDαa+f(x)=1Γ(mα)dmdxmxaf(t)(xt)αm+1dt,RDαbf(x)=(1)mΓ(mα)dmdxmbxf(t)(xt)αm+1dt,

    where Γ denotes, as usual, the Gamma function [53].

    At the end of the sixties, Caputo introduced the formulation for fractional derivatives, which jointly with Minardi was then applied in their research on viscoelasticity theory [27]. Caputo's definition of the fractional derivative is given by

    cDαx=ImαDmform1<αm,

    where Imα is a corresponding integral of mα fractional order. This derivative can be interpreted as:

    cDαxf(x)={1Γ(mα)x0f(m)(t)(xt)α+1mdtform1<α<m,dmdxmf(x)forα=m.

    On a domain Ω and vL2(Ω), the L2 inner product is defined by v,v=Ωv(x)v(x)dx and the associated L2(Ω) norm is defined by ||v||=(Ω|v(x)|2dx)1/2. For any non-negative integer n, Hn(Ω) denotes a Sobolev space with the related norm

    ||w||n=(0sn||swxs||2dx)1/2

    and seminorm ||n. Let C0(Ω) be a space of infinitely differentiable functions with compact support in Ω and Hm0(Ω) is a closure of C0(Ω) with respect to norm ||||m. Finally, we denote by Pk(Ω) the space of all polynomials over Ω of degree up to k. Also, C is independent of mesh sizes, denotes a generic constant that may vary at different occurrences.

    We consider the following time-fractional nonlinear convection-diffusion-reaction equation:

    {c0DαtuΔu+b(x)u+f(u)=g(x,t)inΩ×(0,T],u(x,0)=0,xΩ,u=0,(x,t)Ω×(0,T]. (3.1)

    Here, 0<α<1, c0Dαtu denotes the α-th order Caputo fractional derivative of u, and g(x,t)L2(Ω) is a forcing/load term. For the purpose of theoretical analysis, we make the assumptions:

    (H1) Let f:RR be a Lipschitz continuous function, that is, there exists L>0 such that,

    |f(v1)f(v2)|L|v1v2|v1,v2R.

    (H2) We also suppose that bL(Ω) and there exists μ00 such that for almost every xΩ,

    μ(x):=112b(x)μ00.

    Then, the following wellposedness theorem states the existence and uniqueness of the solution to problem (3.1). The proof of these results can be found in [36].

    Theorem 1 (Wellposedness). Under Assumption (H1), problem (3.1) admits a unique solution u for 0<α<1, such that,

    (i)uCα([0,T];L2(Ω))C([0,T];H10(Ω)H2(Ω)),(ii)cDαtuC([0,T];L2(Ω)),u(t)tL2(Ω),(iii)||u(t)t||Ctα1fort(0,T],

    where C is a strictly positive, real constant (independent of u).

    Li et al. [39] established a discrete fractional Grönwall type inequality for Caputo type fractional derivative. Taking the motivation from this work, we here establish a discrete fractional Grönwall type inequality (Grunwald-Letnikov approximation) to R-L type fractional derivative. To do so, we consider the relation between Caputo and R-L fractional derivatives expressed as

    c0Dαtu=R0Dαt(uu(x,0)).

    Since u(x,0)=0, we find that c0Dαtu(,t)=R0Dαtu(,t). As our initial model has the Caputo type derivative, it is important to take zero initial conditions to use the Caputo and R-L type interchangeably because they are equal when initial conditions are zero. It is noteworthy that the methodology (Grünwald equality) for a time-fractional derivative used is specific for the R-L type [35,37]. The reason to start with the Caputo-type derivative is because of its versatile and meaningful interpretation of physical phenomena and allows for clear formulations of initial and boundary conditions. Furthermore, the literature indicates that there are no inherent limitations on the initial conditions, as the Grünwald approximation can be extended to the Caputo-type derivative with inhomogeneous initial values. This extension is achieved by adding appropriate correction terms, as referenced in the work by Scherer et al. [52]. Now, we write the model problem (3.1) in terms of the R-L fractional derivative reads as

    {R0DαtuΔu+b(x)u+f(u)=g(x,t),(x,t)Ω×(0,T],u(x,0)=0,xΩ,u(x,t)=0,(x,t)Ω×(0,T]. (3.2)

    The weak/variational formulation of the model fractional elliptic problem (3.2) is:

    Find u(t)H10(Ω) for almost all t(0,T) such that

    {m(R0Dαtu,w)+a(u,w)+b(u,w)+f(u),w=g,wwH10(Ω),t(0,T],u(x,0)=0,xΩ, (3.3)

    where the bilinear forms m(,), a(,) and b(,) are defined as:

    m(w,v)=Ωwvdx,a(w,v)=Ωwvdx,b(w,v)=Ω(bw)vdx,

    and , is the duality product in L2(Ω). The wellposedness of the weak form can be proved along similar lines of the work by Jin et al.[35,36] for the time-fractional nonlinear subdiffusion equation.

    Let {Th}h>0 be a family of polygonal meshes covering Ω, each labelled by the mesh size parameter h which is the maximum of the element diameters hE associated with the mesh Th, i.e., h:=maxEThhE. For mesh regularity, we assume there exists ρ>0 such that for every ETh :

    E is star shaped for a ball of radius ρhE,

    ● distance between any two vertices of E is ρhE.

    Referring to [6], for kN, the local virtual element space VEh of order k is defined as,

    VEh={vhH1(E):ΔvhPk(E),vh|EC0(E),vh|ePk(e)eE},

    where E is the polygonal boundary of E and eE denotes a generic edge.

    Next, we define some useful polynomial projection operators. We consider the elliptic projection operator Π,Ek:H1(E)Pk(E), which is the orthogonal projector onto the space of polynomials of degree k with respect to the H1 semi-norm. Formally, for every vH1(E), the k-degree polynomial Π,Ekv is the solution to the variational problem

    E(Π,Ekvv)qkdx=0qkPk(E),E(Π,Ekvv)ds=0,

    where the second condition is needed to fix the gradient kernel. We also define the L2-orthogonal projection operator Π0,Ek:H1(E)Pk(E), such that for any vH1(E), the polynomial Π0,Ekv satisfies:

    E(Π0,Ekvv)qkdx=0qkPk(E).

    Using the elliptic projection operator, we define the "enhanced" local virtual element space denoted WEh by

    WEh={whVEh:E(Π,Ekwh)qkdx=EwhqkdxqkPk/Pk2(E)}, (4.1)

    where Pk/Pk2(E) denotes the quotient space of equivalence classes of polynomials, where two polynomials are equivalent if their difference is a polynomial of degree at most k2 (roughly speaking, in the implementation, we can consider the union of the linear spaces of homogeneous polynomials of degree exactly k and k1). Accordingly, we define the global virtual element space of order k as:

    Wh={wH10(Ω):w|EWEhETh}.

    A function whWEh is uniquely described by the following sets of values that we can take as the local Degrees of Freedom (DoFs):

    (D1) For k1, the vertex values wh(Vi), where Vi are the vertices of element E;

    (D2) For k>1, the values of wh at the (k1) internal Gauss-Lobatto quadrature points on each edge e subset of E;

    (D3) For k>1, the polynomial moments up to order k2,

    Evhpk2dxpk2Pk2(E).

    The DoFs of the global virtual element space of order k are given by collecting the local DoFs of the elemental spaces WEh from all elements ETh that are glued with C0 continuity. The unisolvence of these degrees of freedom is proved in [6]. A crucial property of definition (4.1) is that the projections Π,Ekwh and Π0,Ekwh are computable using only the DoFs (D1)(D3) of whWEh.

    The semi-discrete VEM approximation of the weak formulation (3.3) is given as:

    Find uh(t)Wh, for almost all t(0,T) such that

    {mh(R0Dαtuh,wh)+ah(uh,wh)+bh(uh,wh)+fh(uh),wh=gh,whwhWh,uh(x,0)=0,xΩ.

    Here, the global discrete bilinear forms ah(,):Wh×WhR,bh(,):Wh×WhRandmh(,):Wh×WhR are defined as:

    ah(vh,wh)=EThaEh(vh,wh)vh,whWh,bh(vh,wh)=EThbEh(vh,wh)vh,whWh,mh(vh,wh)=EThmEh(vh,wh)vh,whWh,

    where the computable discrete local bilinear forms aEh(,):WEh×WEhR, bEh(,):WEh×WEhR, and mEh(,):WEh×WEhR are defined over every ETh as follows:

    aEh(uh,vh)=aE(Π,Ekuh,Π,Ekvh)+sEa((IΠ,Ek)uh,(IΠ,Ek)vh)uh,vhWEh,mEh(uh,vh)=mE(Π0,Ekuh,Π0,Ekvh)+sEm((IΠ0,Ek)uh,(IΠ0,Ek)vh)uh,vhWEh,bEh(uh,vh)=bE(Π0,Ek1uh,Π0,Ekvh)uh,vhWEh,

    where the bilinear forms sEa:VEh×VEhR and sEm:VEh×VEhR are a locally admissible stabilizing terms that are symmetric positive definite and satisfy

    c0aE(vh,vh)sEa(vh,vh)c1aE(vh,vh)vhVEh,d0mE(vh,vh)sEm(vh,vh)d1mE(vh,vh)vhVEh,

    for some positive constants c0,c1,d0andd1 independent of h and E.

    These stabilization terms ensure the local stability of the corresponding elemental bilinear forms aEh(,) and mEh(,) (a formal definition follows below). For the computational purpose, we consider the "dofi-dofi" stabilization, see [46]. For the sake of completeness, we refer to various specific choices of the computable definitions for the stabilizers available in the literature [18,19,21,28].

    We recall that the bilinear forms aEh(,) and mEh(,) possess the fundamental properties of polynomial consistency and stability, that are useful for theoretical estimation.

    Polynomial consistency: ETh and for any whWEh, it holds that

    aEh(p,wh)=aE(p,wh)pPk(E),mEh(p,wh)=mE(p,wh)pPk(E).

    Stability: There exist two pairs of constants independent of h, say, (α,α) and (γ,γ) with 0<αα and 0<γγ, such that for all whWEh, the two following equivalence holds:

    αaE(wh,wh)aEh(wh,wh)αaE(wh,wh),γmE(wh,wh)mEh(wh,wh)γmE(wh,wh).

    The conditions above ensure that the non-polynomial parts sEa(,) and sEm(,) scale as the polynomial parts of aEh(,) and mEh(,), respectively.

    To compute the nonlinear forcing term, we use the L2-orthogonal projector Π0,Ek previously defined. For each element E, we first define fh(uh) as

    fh(uh)|E:=Π0,Ek(f(Π0,Ekuh))on eachETh.

    The orthogonality of Π0,Ek implies that

    fh(uh),vh=EThEfh(uh)vhdx=EThEΠ0,Ek(f(Π0,Ekuh))vhdx=EThEf(Π0,Ekuh)Π0,Ekvhdx.

    Hence, for f(uh)L2(Ω) and uhWh, we state that

    fh(uh):=Π0k(f(Π0kuh)),

    by introducing a global L2-orthogonal projection operator Π0k:L2(Ω)Pk(Th) onto the piecewise polynomial space of degree k built on the mesh partition Th. The global projection operator is such that

    Π0k(f(Π0kuh))|E=Π0,Ek(f(Π0,Ekuh))ETh.

    We outlined that this approximation of the forcing term is computable since the projection Π0,Ekuh is computable from the degrees of freedom of uh in the enhanced space WEh.

    To develop a fully discrete VEM, the next step is to approximate the fractional derivative in the temporal direction. Let 0=t0<t1<<tK=T be a uniform partition of the interval (0,T] with time step size τ=T/K for some positive integer K. By employing the Grunwald-Letnikov approximation, we effectively calculate the R-L type fractional derivative as obtained in the work of Kumar et al. [37]. This computation yields

    RDαtnu=ταni=0w(α)niu(x,ti)+Rn. (4.2)

    Here, the weights are given by

    w(α)i=(1)iΓ(α+1)Γ(i+1)Γ(αi+1),

    where Rn satisfies the estimate ||Rn||cτ, and the approximation (4.2) is also O(τ) accurate.

    Denote

    Un,θh=(1θ2)Unh+θ2Un1h,

    where 0θ1. For any arbitrary choice of θ[0,1] in the discrete scheme, the numerical results show no significant effect on the accuracy, independent of the fractional-order α(0,1). Hence, for convenience and reduced simplified notation, we fix θ=α in the following fully discrete VEM formulation:

    Find UnhWh, n=1,2,,K, such that

    {mh(R0DαtUnh,vh)+ah(Un,αh,vh)+bh(Un,αh,vh)+f(Un,αh),vh)=gh,vhvhWh,U0h(x)=0,xΩ.

    Using the definition of operator RDαtn from (4.2), we rewrite the equation above as:

    ταw(α)0mh(Unh,vh)+ah(Un,αh,vh)+bh(Un,αh,vh)+fh(Un,αh),vh=gh,vhταn1j=1w(α)njmh(Ujh,vh). (4.3)

    This discrete formulation provides a system of algebraic equations that we solve to compute the virtual element approximation.

    To prove the well-posedness of the fully discrete VEM (i.e., the uniqueness and existence of the virtual element approximation), we need the following proposition, which results from an application of the Brouwer's fixed point theorem, see, e.g., [55] for the details.

    Proposition 1. Let H be a finite-dimensional Hilbert space with inner product , and norm |||| and let S:HH be a continuous map such that,

    S(v),v>0vHwith||v||=ρ>0.

    Then, there exists wH such that,

    S(w)=0and||w||<ρ.

    Applying Proposition 1, we can derive the wellposedness result that we state in the next theorem.

    Theorem 2 (Wellposedness). Let U0h,,Un1h be the first n virtual element fields solving the fully discrete VEM for some given initial and boundary condition. Then, a unique solution Unh of the fully discrete VEM exists.

    Proof. We rewrite Eq (4.3) as

    ταw(α)0mh(Unh,vh)+ah(Un,αh,vh)+bh(Un,αh,vh)+fh(Un,αh),vhgh,vh+ταn1j=1w(α)njmh(Ujh,vh)=0, (4.4)

    and multiply each term by (1α2) to obtain

    ταw(α)0mh(Unh,vh)+(1α2)ah(Un,αh,vh)+(1α2)bh(Un,αh,vh)+(1α2)fh(Un,αh),vh+τα(1α2)n1j=1w(α)njmh(Ujh,vh)(1α2)gh,vhτα(α2)mh(Un1h,vh)=0. (4.5)

    Equations (4.4) and (4.5) are equivalent, and their solutions coincide. Then, we define the continuous operator G:WhWh such that

    m(G(Wn,α),V):=ταm(Wn,α,V)+(1α2)a(Wn,α,V)+(1α2)b(Wn,α,V)+(1α2)fh(Wn,α),V+τα(1α2)n1j=1w(α)njm(Ujh,V)(1α2)gh,Vτα(α2)m(Un1h,V) (4.6)

    for Wn,α and V in Wh. Setting V=Wn,α, we find that

    m(G(Wn,α),Wn,α):=ταm(Wn,α,Wn,α)+(1α2)a(Wn,α,Wn,α)+(1α2)b(Wn,α,Wn,α)+(1α2)fh(Wn,α),Wn,α+τα(1α2)n1j=1w(α)njm(Ujh,Wn,α)(1α2)gh,Wn,ατα(α2)m(Un1h,Wn,α). (4.7)

    From Assumption (H1), we have that

    ||fh(Wn,α)||L||Wn,α||+||f(0)||

    from which it follows that

    ||fh(Wn,α)||a(1+||Wn,α||),a>0. (4.8)

    By using the Cauchy-Schwarz inequality in (4.7), Eq (4.8) and noting that w(α)j<0 for 1jn, we find that

    m(G(Wn,α),Wn,α)||Wn,α||2+(1α2)τα||Wn,α||2+(1α2)τα||Wn,α||2+(1α2)ταa(1+||Wn,α||)||Wn,α||+(1α2)n1j=1w(α)nj||Ujh||||Wn,α||(1α2)ταb(1+||Wn,α||)||Wn,α||(α2)||Un1h||||Wn,α||. (4.9)

    Since, τα||Wn,α||>0, we rewrite (4.9) as

    m(G(Wn,α),Wn,α)((12(1α2)ταa)||Wn,α||+(1α2)τα(ab)+(1α2)n1j=1w(α)nj||Ujh||(α2)||Un1h||)||Wn,α||.

    Then, m(G(Wn,α),Wn,α)>0, if and only if it holds that

    (12(1α2)ταa)||Wn,α||+(1α2)τα(ab)+(1α2)n1j=1w(α)nj||Ujh||α2||Un1h||>0.

    Choosing τα<1(1α2)a, there exists a function Wn,α, such that

    ||Wn,α||>1(12(1α2)ταa)((1α2)τα(ab)(1α2)n1j=1w(α)nj||Ujh||+α2||Un1h||).

    This inequality implies that m(G(Wn,α),Wn,α)>0. Thus, for ||Wn,α||=ρ, we have that

    m(G(Wn,α),Wn,α)>0,

    and Proposition 1 implies the existence of the virtual element approximation.

    To prove the uniqueness of a virtual element function Un,αh solving problem (4.3), assume that Un,αh,1 and Un,αh,2 are two solutions. To ease the notation, we denote U1=Un,αh,1 and U2=Un,αh,2. Then, from (4.5) it follows that

    ταm(U1U2,vh)+(1α2)a(U1U2,vh)+(1α2)b(U1U2,vh)=(1α2)f(U1)f(U2),vh. (4.10)

    Setting vh=U1U2=r in (4.10) and using Assumption (H1), we find that

    ||r||2(1α2)ταL(||r||)||r||, (4.11)

    and choosing τα<1(1α2)L sufficiently small, with L>0 being the Lipschitz constant, we find that ||r2||0, which implies the uniqueness of the solution.

    From the theoretical results that we prove in this subsection, some important aspects of the VEM and the fully discrete scheme (4.3) are achieved. These points are given as follows:

    ● The efficiency with which the VEM is combined with a finite difference scheme to form a fully discrete scheme.

    ● The dependence of force/load term and reaction term on both the spatial and temporal domains and extended VEM to such terms nested with both domain discretizations.

    ● The optimal convergence order for a fully discrete scheme is achieved with theoretical analysis and supported with numerical results on several sets of meshes, including non-convex meshes.

    To prove the convergence of the VEM, we derive suitable a priori error bounds. To this end, we need the fractional Grönwall type inequality discussed here. The Grönwall weights w(α)i can be computed as

    w(α)0=1andw(α)i=(1α+1i)w(α)i1fori1.

    Let g(α)n=ni=1w(α)i, so that

    g(α)0=w(α)0andw(α)i=g(α)ig(α)i1for1in.

    Note that w(α)i satisfy

    w(α)0=1,1<w(α)1<w(α)2<<w(α)i<<0,i=0w(α)i=0.

    Therefore, g(α)i1>g(α)ifori=1,,n. Using such a definition of g(α)n, we rewrite the R-L derivative expansion as

    RDατu(x,tn)=ταni=1(g(α)ig(α)i1)u(x,tni)+ταg(α)0u(x,tn).

    Since u(x,t0)=0, we have that

    RDατu(x,tn)=ταni=1g(α)niδu(x,ti),

    where δu(x,ti)=u(x,ti)u(x,ti1) for every i=1,,n.

    Furthermore, we consider the following technical lemmas that we will use in the next derivations. Their proof can be found in [39].

    Lemma 1. Consider the sequence {ϕn} defined by

    ϕ0=1,ϕn=ni=1(g(α)i1g(α)i)ϕniforn1. (4.12)

    Then, {ϕn} satisfies

    0<ϕn<1,ni=jϕnig(α)ij=1,1jn, (4.13)

    and the following inequalities:

    1Γ(α)ni=1ϕninαΓ(1+α), (4.14)
    1Γ(α)Γ(1+(k1)α)n1i=1ϕnii(k1)αnkαΓ(1+kα) (4.15)

    for any integer k1.

    Lemma 2. Let {cn:n0} and {dn:n0} be two non-negative sequences and λ1 and λ2 be two non-negative constants. Then, for

    c0=0andRDατcnλ1cn+λ2cn1+dn,n1,

    a positive constant τ exists such that for ττ we have that

    cn2(tαnαmax1indi)Eα(2Γ(α)λtαn),1nK,

    where Eα(z)=j=0zjΓ(1+αj) is the Mittag-Leffler function [34] and λ=λ1+λ2.

    Lemma 3. For any sequence {ik}Kk=0Wh, the following inequality holds

    mh(RDατik,(1α2)ik+α2ik1)12RDατ||ik||2for1kK. (4.16)

    Proof. We rewrite Eq (4.16) as

    mh(RDατik,(1α2)ik+α2ik1)=(1α2)mh(RDατik,ik)+α2mh(RDατik,ik1)=τα((1α2)kj=0w(α)kjmh(ij,ik)+α2kj=0w(α)kjmh(ij,ik1))=τα((1α2)w(α)0||ik||2+α2w(α)1||ik1||2+((1α2)w(α)1+α2w(α)0)mh(ik,ik1)=τα(+(1α2)k2j=0w(α)kjmh(ij,ik)+α2k2j=0w(α)kjmh(ij,ik1))τα((1α2)w(α)0||ik||2+α2w(α)1||ik1||2+((1α2)w(α)1+α2w(α)0)||ik||2+||ik1||22=τα(+(1α2)k2j=0w(α)kj||ij||2+||ik||22+α2k2j=0w(α)kj||ij||2+||ik1||22).

    Then, use the fact that, (1α2)w(α)1+α2w(α)0<0, and w(α)j<0j1, and we obtain

    mh(RDατik,(1α2)ik+α2ik1)τα(((1α2)w(α)0+12(1α2)w(α)1+α4w(α)0)||ik||2+(α2w(α)1+12(1α2)w(α)1+α4w(α)0)||ik1||2+12k2j=0w(α)kj||ij||2+12(1α2)k2j=0w(α)kj||ik||2+α4k2j=0w(α)kj||ik1||2)=τα(((1α2)w(α)012(1α2)w(α)0+α4w(α)0)||ik||2+(α2w(α)1+12(1α2)w(α)1α4w(α)1)||ik1||2+12k2j=0w(α)kj||ij||2+12(1α2)kj=0w(α)kj||ik||2+α4kj=0w(α)kj||ik1||2)τα2kj=0w(α)kj||ij||2=12RDατ||ik||2,

    which proves inequality (4.16).

    Now, we derive the a priori error bound for the virtual element approximation Unh, which we state in the following theorems.

    Theorem 3 (A priori bound of discrete VEM). Let Unh be the solution to the semi-discrete VEM. Then, there exists a positive constant τ such that for ττ the virtual element field Unh satisfies

    ||Unh||C,n=1,,K, (4.17)

    where C is a positive constant independent of h and τ.

    Proof. The virtual element solution field Unh must satisfy the variational condition

    mh(RDατUnh,vh)+ah(Un,αh,vh)+bh(Un,αh,vh)=f(Un,αh),vhvhWh.

    Setting vh=Un,αh, we obtain

    mh(RDατUnh,Un,αh)+2||Un,αh||212(||f(Un,αh)||2+||Un,αh||2).

    Then, we use Assumption (H1) to find

    mh(RDατUnh,Un,αh)+2||Un,αh||2C((1+||Un,αh||2)+||Un,αh||2),

    where C=max{1,L2}, L being the Lipschitz constant. We apply the standard inequality (a+b)22(a2+b2), a,b0, and we have that

    mh(RDατUnh,Un,αh)C(1+||Un,αh||2).

    Using the results of Lemma 3 yields

    RDατ||Unh||2C(1+||Un,αh||2),

    which implies that

    RDατ||Unh||2C(1+(1α2)2||Unh||2+(α2)2||Un1h||2).

    Finally, in view of Lemma 2, we find a positive constant τ such that ττ and ||Unh||2C, which proves the theorem's assertion. Here C is a positive constant independent of the mesh size h and the time step size τ, and it represents more specifically an upper bound on semi discrete VEM solution Unh. It depends on the assumption of Lipschitz continuity of function f and Lipschitz constant L.

    The last result of this section is about the convergence of the fully discrete VEM, which we prove in the next theorem by deriving an a priori error estimate.

    Theorem 4 (Convergence of the fully discrete VEM). Let u and Unh be the solution to problem (3.1) and the fully discrete VEM (4.3) under Assumption (H1) respectively, with convex regularity of all the elements ETh. Then, there exists a positive constant τ such that

    ||unUnh||C(τ+hk+1),n=1,2,,K, (4.18)

    for ττ, where C is a positive constant independent of τ and h.

    Proof. Let Πh be the projection operator that satisfies

    ah(Πhv,vh)=ah(v,vh)vH10(Ω),vhWh. (4.19)

    According to Cangiani et al. [26] and the references therein, there exists a positive constant C independent of h such that

    ||vΠhv||jChij||v||ivHiH10,j=0k1andi=1k. (4.20)

    The regularity of the domain at j=0 is implied from the convex regularity of the local elements. Now, using this projection operator, we rewrite the approximation error as

    unUnh=(unΠhun)+(ΠhunUnh)=ρnh+Θnh, (4.21)

    with the obvious definition of ρnh and Θnh.

    Assuming that

    Ah(uh,vh)=ah(uh,vh)+bh(uh,vh),

    for any arbitrary vhWh, we find that Θnh satisfies

    mh(RDατΘnh,vh)+Ah(Θn,αh,vh)=mh(RDατ(ΠhunUnh),vh)+Ah((Πhun,αUn,αh),vh)=mh(RDατΠhun,vh)+Ah(Πhun,α,vh)mh(RDατUnh,vh)Ah(Un,αh,vh). (4.22)

    Then, by using (4.3) and (4.19) in (4.22), we obtain that

    mh(RDατΘnh,vh)+Ah(Θn,αh,vh)=mh(RDατΠhun,vh)+Ah(un,α,vh)f(un,α),vh+f(un,α),vhf(Un,αh),vh. (4.23)

    The weak form of (3.1) implies that

    mh(RDαtnun,vh)+Ah(un,vh)=f(un,α),vh. (4.24)

    Using (4.24) in (4.23) yields

    mh(RDατΘnh,vh)+Ah(Θn,αh,vh)=mh((RDατΠhunRDαtnu),vh)+Ah((un,αun),vh)+f(un)f(Un,αh),vh. (4.25)

    Setting vh=Θn,αh in (4.25) and using Cauchy-Schwarz inequality, we obtain

    mh(RDατΘnh,Θn,αh)+||Θn,αh||2||RDατΠhunRDαtnu||||Θn,αh||+||(un,αun)||||Θn,αh||+||f(un)f(Un,αh)||||Θn,αh||L2||unUn,αh||2+L2||Θn,αh||2+L2||Θn,αh||2+12||RDατΠhunRDαtnu||2+12||Θn,αh||2+12||(un,αun)||2L2||unUn,αh||2+(L+12)||Θn,αh||2+12||RDατΠhunRDαtnu||2+12||(un,αun)||2+12||Θn,αh||2. (4.26)

    We add and subtract un,α and Πhun to the argument of ||unUn,αh||; then, we apply the triangular inequality twice and use the definition of ρnh and Θnh from (4.21), and we obtain

    ||unUn,αh||||unun,α||+||ρn,αh||+||Θn,αh||||Θn,αh||+C(τ+hk+1). (4.27)

    We also note that

    ||RDατΠhunRDαtnu||||RDατΠhunRDαtnΠhu||+||RDαtnΠhuRDαtnu||C(τ+hk+1), (4.28)

    and

    ||unun,α||(1α2)(α2)τtntn1||utt(s)||dsCτ. (4.29)

    Using (4.27)–(4.29) in (4.26), we obtain

    mh(RDατΘnh,Θn,αh)3L+12||Θn,αh||2+C(τ+hk+1)2, (4.30)

    which gives

    RDατ||Θnh||2(3L+1)||Θn,αh||2+C(τ+hk+1)2. (4.31)

    From (4.31), we can get

    RDατ||Θnh||2C||Θn,αh||2+C(τ+hk+1)2,

    where C=3L+1. Furthermore, we also can write

    RDατ||Θnh||22C(1α2)2||Θnh||2+2C(α2)2||Θn1h||2+C(τ+hk+1)2. (4.32)

    Applying Lemma 2, there exists a positive constant τ such that ||Θnh||2C(τ+hk+1)2 for ττ, which implies that ||Θnh||C(τ+hk+1). The theorem's assertion follows on applying the triangular inequality along with (4.20) and the definition of Θnh from (4.21).

    A short note on the implementation is discussed in this subsection. We solve the resulting system of nonlinear algebraic equations from (4.3) by incorporating Newton's method.

    Let N be the dimension and {ϕi}Ni=1 be the canonical basis for the global virtual element space Wh. For some βniR,i=1,2,,N, we can write the solution of UnhWh of (4.3) as

    Unh=Ni=1βniϕi. (4.33)

    After defining βn:=(βn1,βn2,,βnN)T, using the value of Unh from (4.33) in (4.3), we get the following nonlinear algebraic equation

    Hi(Unh)=0,1iN, (4.34)

    where

    Hi(Unh)=ταw(α)0mh(Unh,ϕi)+ah(Un,αh,ϕi)+bh(Un,αh,ϕi)+f(Un,αh),ϕi+ταn1j=1w(α)njmh(Ujh,ϕi)g,ϕi.

    Using the Newton's method in (4.34), we obtain the matrix system

    Jβn=H,

    where H=(H1,H2,,HN)T and the entries of the (N×N) Jacobian matrix J are given by

    (J)li=Hiβnl(Unh)=ταw(α)0mh(ϕl,ϕi)+(1α2)ah(ϕl,ϕi)+(1α2)bh(ϕl,ϕi)=+(1α2)f(Un,αh)Unhϕl,ϕi,

    where 1i,lN.

    In this section, we assess the performance of our approximation method by solving problem (3.1) on three different test cases and by using a sequence of refined uniform square, regular Voronoi meshes, and non-convex meshes (sample mesh configurations given in Figure 1). There are a few reasons to choose the different sets of meshes, including the non-convex meshes.

    Figure 1.  (a) Uniform square meshes; (b) Regular Voronoi meshes; (c) Non-convex meshes.

    ● To check the efficiency of the constructed VEM as it is mentioned that the VEM is a generalization of FEM over polygonal meshes.

    ● The use of square meshes was to provide an explanation that VEM can adjust the triangular or quadrilateral elements as well and can provide the accuracy of classical FEMs on these meshes.

    ● The use of non-convex meshes provides the explanation of the point that in VEM, the basis functions are constructed virtually as compared to the implicit definition of basis function in classical FEMs.

    To this end, we measure the approximation errors as the difference between the exact solution u, the L2 orthogonal projection Π0kuh, and the elliptic projection Πkuh of the virtual element approximation uh through the formulas

    e2h,0=ETh||uΠ0kuh||2Eande2h,1=ETh||(uΠkuh)||2E.

    By comparing the errors at two subsequent mesh refinements, we compute the convergence rate. From the theoretical results of Section 3, we expect to see the optimal convergence rates e2h,0=O(hk+1) and e2h,1=O(hk), assuming that the exact solution is smooth enough. Also, the error values in spatial directions for all examples are calculated at T=1 and constant time steps with K=100 that is τ=1/100 wherein time-step τ obeys the CFL condition τO(hk+1).

    The purpose of introducing VEM is to take care of the two-dimensional spatial domain, where it is efficiently combined with a finite difference approach for the temporal direction, and we get a fully discrete scheme. In all the example cases, we have taken zero initial condition then only we are able to use the Grunwald-Letnikov approximation for Caputo type derivatives because with zero initial value, the Caputo type coincides with R-L type, and we can use the Grunwald approximation designed for R-L type derivatives. However, it is important to mention that in Section 5.2, we have taken a case of a non-smooth analytical solution in the temporal direction, which affects the convergence optimality of the temporal direction.

    Consider the two-dimensional time-fractional nonlinear convection-diffusion equation (3.1), which models the anomalous processes, that is, the processes governed by random walks against the ideal Brownian motion [47] by applying the VEM scheme (4.3).

    We let Ω=(0,1)×(0,1), J=[0,1] with b=(1,1)T with the nonlinear term f(u)=u+u2 and the RHS function g(x,t) given by

    g(x,t)=(6t3αΓ(4α)+(2π2+1)t3)sin(πx)sin(πy)+πt3(cos(πx)sin(πy)+sin(πx)cos(πy))+(t3sin(πx)sin(πy))2.

    The virtual element solution and error plots at α=0.8, over the family of uniform square meshes and regular Voronoi meshes are shown in Figures 2 and 3 for k=1, and likewise, for k=2 in Figures 4 and 5. The L2 norm and H1 semi-norm error obtained along with the convergence rates for linear order VEM are presented in Tables 1 and 2 for the uniform square meshes and in Tables 3 and 4 for the regular Voronoi meshes, wherein Table 3 presents the error values in spatial direction with respect to different time step lengths and for different α values. We report the corresponding results for k=2 in Tables 58. In addition, Table 9 demonstrates that the VEM (4.3) is of global order O(τ) at time T=1 and T=0.1. The error and convergence in temporal direction are calculated by refining τ within a range of 12n,n=1,,4 for fixed mesh size h=103, wherein convergence rate in the time direction is calculated by rate=log2(e(τn)(e(τn+1)).

    Figure 2.  Example 1: (a) virtual element solution; (b) plots of the error curves for α=0.8 over uniform square meshes for VEM of order k=1.
    Figure 3.  Example 1: (a) virtual element solution; (b) plots of the error curves for α=0.8 over regular Voronoi meshes for VEM of order k=1.
    Figure 4.  Example 1: (a) virtual element solution; (b) plots of the error curves for α=0.8 over uniform square meshes for VEM of order k=2.
    Figure 5.  Example 1: (a) virtual element solution; (b) plots of the error curves for α=0.8 over regular Voronoi meshes for VEM of order k=2.
    Table 1.  Example 1: L2(Ω)-norm of the approximation error and convergence rate for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 25 4.856 e-02 4.889 e-02
    3 81 1.123 e-02 2.11 1.116 e-02 2.13
    4 289 2.577 e-03 2.12 2.601 e-03 2.10
    5 1089 5.037 e-04 2.35 5.375 e-04 2.27

     | Show Table
    DownLoad: CSV
    Table 2.  Example 1: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 25 6.992 e-01 6.994 e-01
    3 81 3.543 e-01 0.98 3.542 e-01 0.98
    4 289 1.778 e-01 0.99 1.778 e-01 0.99
    5 1089 8.901 e-02 0.99 8.901 e-02 0.99

     | Show Table
    DownLoad: CSV
    Table 3.  Example 1: L2(Ω)-norm of the approximation error over regular Voronoi mesh configuration of unit square domain with mesh size h=2 for the VEM of order k=1 at different time step sizes.
    α=0.4 α=0.8
    DoF τ=1/50 τ=1/100 τ=1/50 τ=1/100
    2 66 2.195 e-02 2.195 e-02 2.178 e-02 2.178 e-02
    3 256 5.003 e-03 5.005 e-03 4.921 e-03 4.923 e-03
    4 999 1.206 e-03 1.208 e-03 1.183 e-03 1.186 e-03
    5 3998 2.895 e-04 2.915 e-04 2.832 e-04 2.858 e-04

     | Show Table
    DownLoad: CSV
    Table 4.  Example 1: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using regular Voronoi meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 66 5.000 e-01 5.000 e-01
    3 256 2.512 e-01 0.99 2.512 e-01 0.99
    4 999 1.260 e-01 0.99 1.260 e-01 0.99
    5 3998 6.296 e-02 1.00 6.295 e-02 1.00

     | Show Table
    DownLoad: CSV
    Table 5.  Example 1: L2(Ω)-norm of the approximation error and convergence rate for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 81 4.079e-03 4.076e-03
    3 289 5.196e-04 2.97 5.195e-04 2.97
    4 1089 6.521e-05 2.99 6.521e-05 2.99
    5 4225 8.169e-06 3.00 8.178e-06 3.00

     | Show Table
    DownLoad: CSV
    Table 6.  Example 1: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 81 1.319e-01 1.319e-01
    3 289 3.358e-02 1.97 3.358e-02 1.97
    4 1089 8.432e-03 1.99 8.432e-03 1.99
    5 4225 2.110e-03 2.00 2.110e-03 2.00

     | Show Table
    DownLoad: CSV
    Table 7.  Example 1: L2(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using regular Voronoi meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 195 1.433e-03 1.433e-03
    3 767 1.792e-04 3.00 1.792e-04 3.00
    4 2997 2.232e-05 3.00 2.232e-05 3.00
    5 11995 2.776e-06 3.01 2.802e-06 2.99

     | Show Table
    DownLoad: CSV
    Table 8.  Example 1: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using regular Voronoi meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 195 6.067e-02 6.068e-02
    3 767 1.479e-02 2.04 1.479e-02 2.04
    4 2997 3.691e-03 2.00 3.691e-03 2.00
    5 11995 9.079e-04 2.02 9.079e-04 2.02

     | Show Table
    DownLoad: CSV
    Table 9.  Example 1: The error and convergence rates at time T=1 and T=0.1, for different values of α in time direction with regular Voronoi mesh configuration in the spatial domain and fixed mesh size h=103.
    α=0.4 α=0.8
    T τ L2-norm rate L2-norm rate
    1 1/2 1.265e-03 1.252e-03
    1/22 6.743e-04 0.91 6.603e-04 0.92
    1/23 3.513e-04 0.94 3.487e-04 0.92
    1/24 1.819e-04 0.95 1.798e-04 0.96
    0.1 1/2 1.652e-03 1.681e-03
    1/22 8.851e-04 0.90 8.932e-04 0.91
    1/23 4.725e-04 0.91 4.683e-04 0.93
    1/24 2.472e-04 0.93 2.405e-04 0.96

     | Show Table
    DownLoad: CSV

    Consider the nonlinear term f(u)=u+u3 with b=(1,1)T and define the function g(x,t) such that the exact solution is given by,

    u=(tα)sin(x)(1x)sin(y)(1y)

    over domain Ω=(0,1)×(0,1) and time interval J=[0,1]. Figure 6 and 7 show the virtual element solution and error plots at α=0.6, over the uniform square and regular Voronoi meshes for k=1. Figure 8 and 9 show the virtual element solution and error plots at α=0.8, over uniform square and regular Voronoi meshes for k=2. We report the approximation errors measured using the L2 norm and H1 semi-norm and the convergence rates for k=1 when applying the VEM for k=1 on the uniform square meshes in Tables 11 and 12, and on the regular Voronoi meshes in Tables 13 and 14, wherein Table 13 presents the error values in spatial direction with respect to different time step lengths and for different α values. We report the results for k=2 in Tables 1518. This example is different in a sense, as in we have assumed the analytical solution in time direction as tα. The optimal order of convergence, that is O(τ), is achieved. The error table 10 provides the corresponding values.

    Figure 6.  Example 2: (a) virtual element solution; (b) plots of the error curves for α=0.6 over uniform square meshes for VEM of order k=1.
    Figure 7.  Example 2: (a) virtual element solution; (b) plots of the error curves for α=0.6 over regular Voronoi meshes for VEM of order k=1.
    Figure 8.  Example 2: (a) virtual element solution; (b) plots of the error curves for α=0.6 over uniform square meshes for VEM of order k=2.
    Figure 9.  Example 2: (a) virtual element solution; (b) plots of the error curves for α=0.6 over regular Voronoi meshes for VEM of order k=2.
    Table 10.  Example 2: The error and convergence rates with fixed T for different values of α in time direction with regular Voronoi mesh configuration in spatial domain and fixed mesh size h=103.
    α=0.4 α=0.8
    T τ L2-norm rate L2-norm rate
    1 1/2 1.463e-03 1.432e-03
    1/22 7.821e-04 0.90 7.576e-04 0.92
    1/23 4.112e-04 0.93 4.011e-04 0.92
    1/24 2.134e-04 0.95 2.104e-04 0.93
    0.1 1/2 1.846e-03 1.823e-03
    1/22 9.853e-04 0.91 9.689e-04 0.91
    1/23 5.176e-04 0.93 5.074e-04 0.93
    1/24 2.687e-04 0.95 2.607e-04 0.96

     | Show Table
    DownLoad: CSV
    Table 11.  Example 2: L2(Ω)-norm of the approximation error and convergence rate for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 25 3.089e-03 3.090e-03
    3 81 7.628e-04 2.02 7.626e-04 2.02
    4 289 1.903e-04 2.00 1.901e-04 2.00
    5 1089 4.772e-05 2.00 4.756e-05 2.00

     | Show Table
    DownLoad: CSV
    Table 12.  Example 2: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 25 4.452e-02 4.452e-02
    3 81 2.257e-02 0.98 2.257e-02 0.98
    4 289 1.133e-02 0.99 1.133e-02 0.99
    5 1089 5.668e-03 1.00 5.668e-03 1.00

     | Show Table
    DownLoad: CSV
    Table 13.  Example 2: L2(Ω)-norm of the approximation error over regular Voronoi mesh configuration of unit square domain with mesh size h=2 for the VEM of order k=1 at different time step sizes.
    α=0.4 α=0.8
    DoF τ=1/50 τ=1/100 τ=1/50 τ=1/100
    2 66 1.481 e-03 1.481 e-03 1.479 e-03 1.479 e-03
    3 256 3.360 e-04 3.358 e-04 3.353 e-04 3.353 e-04
    4 999 8.175 e-05 8.156 e-05 8.132 e-05 8.132 e-05
    5 3998 1.990 e-05 1.971 e-05 1.956 e-05 1.955 e-05

     | Show Table
    DownLoad: CSV
    Table 14.  Example 2: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using regular Voronoi meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 66 3.187e-02 3.187e-02
    3 256 1.604e-02 0.99 1.604e-02 0.99
    4 999 8.032e-03 1.00 8.032e-03 1.00
    5 3998 4.006e-03 1.00 4.006e-03 1.00

     | Show Table
    DownLoad: CSV
    Table 15.  Example 2: L2(Ω)-norm of the approximation error and convergence rate for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 81 2.404e-04 2.404e-04
    3 289 3.060e-05 2.97 3.060e-05 2.97
    4 1089 3.846e-06 2.99 3.842e-06 2.99
    5 4225 5.098e-07 2.92 4.830e-07 2.99

     | Show Table
    DownLoad: CSV
    Table 16.  Example 2: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 81 8.140e-03 8.140e-03
    3 289 2.071e-03 1.97 2.071e-03 1.97
    4 1089 5.199e-04 1.99 5.199e-04 1.99
    5 4225 1.301e-04 2.00 1.301e-04 2.00

     | Show Table
    DownLoad: CSV
    Table 17.  Example 2: L2(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using regular Voronoi meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 195 8.285e-05 8.284e-05
    3 767 1.084e-05 2.93 1.084e-05 2.93
    4 2997 1.332e-06 3.03 1.320e-06 3.04
    5 11995 1.692e-07 2.98 1.671e-07 2.98

     | Show Table
    DownLoad: CSV
    Table 18.  Example 2: H1(Ω)-norm of the approximation error and convergence rate of convergence for α=0.4 and 0.8 using regular Voronoi meshes with mesh size h=2 for the VEM of order k=2.
    α=0.4 α=0.8
    DoFs H1-norm rate H1-norm rate
    2 195 3.814e-03 3.814e-03
    3 767 9.617e-04 1.99 9.617e-04 1.99
    4 2997 2.351e-04 2.03 2.351e-04 2.03
    5 11995 5.701e-05 2.04 5.700e-05 2.04

     | Show Table
    DownLoad: CSV

    In this problem, we choose a variable velocity field b=(y,0)T, the nonlinear term as f(u)=u3 and set g(x,t) from the non-smooth exact solution u=tx1.6 over unit square domain. We report the L2 norm and H1 semi-norm and convergence rates over uniform square meshes using VEM of order k=1, in Tables 19 and 20 respectively, while over non-convex mesh configuration results are reported in Tables 21 and 22. Figure 10 shows the VEM solution and an error plot at α=0.8 for non-convex meshes. Due to the weak regularity of the exact solution, an optimal convergence rate is not achieved for higher-order VEM (k2).

    Table 19.  Example 3: L2(Ω)-norm of the approximation error and rate of convergence for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 25 9.642e-03 9.642e-03
    3 81 2.542e-03 1.92 2.542e-03 1.92
    4 289 6.606e-04 1.94 6.606e-04 1.94
    5 1089 1.702e-04 1.96 1.702e-04 1.96

     | Show Table
    DownLoad: CSV
    Table 20.  Example 3: H1(Ω)-norm of the approximation error and rate of convergence for α=0.4 and 0.8 using uniform square meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 25 1.163e-01 1.163e-01
    3 81 6.098e-02 0.93 6.098e-02 0.93
    4 289 3.167e-02 0.95 3.167e-02 0.95
    5 1089 1.634e-02 0.96 1.634e-02 0.96

     | Show Table
    DownLoad: CSV
    Table 21.  Example 3: L2(Ω)-norm of the approximation error and rate of convergence for α=0.4 and 0.8 using non-convex meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 76 5.900e-03 5.900e-03
    3 301 1.504e-03 1.97 1.504e-03 1.97
    4 1201 3.836e-04 1.97 3.836e-04 1.97
    5 4801 9.763e-05 1.97 9.763e-05 1.97

     | Show Table
    DownLoad: CSV
    Table 22.  Example 3: H1(Ω)-norm of the approximation error and rate of convergence for α=0.4 and 0.8 using non-convex meshes with mesh size h=2 for the VEM of order k=1.
    α=0.4 α=0.8
    DoFs L2-norm rate L2-norm rate
    2 76 1.031e-01 1.031e-01
    3 301 5.309e-02 0.96 5.309e-02 0.96
    4 1201 2.725e-02 0.96 2.725e-02 0.96
    5 4801 1.394e-02 0.97 1.394e-02 0.97

     | Show Table
    DownLoad: CSV
    Figure 10.  Example 3: (a) virtual element solution; (b) plots of the error curves for α=0.8 over non-convex meshes for VEM of order k=1.

    Tables 3 and 13 for Examples 1 and 2 respectively show how the error's concur using VEM of order k=1 over regular Voronoi mesh configurations. For a suitable range of h (0.25-0.01) and τ (0.05-0.01) we see with increasing time steps/decreasing temporal mesh size the error values have very slight changes that infers, as we keep on increasing time steps with constant spatial mesh sizes the error values have very slight changes.

    Time-fractional models are of great significance in process modeling, necessitating the development of efficient numerical techniques for their approximation. This paper provides a comprehensive study of the Virtual Element Method in the context of time-fractional nonlinear convection-diffusion equations. We introduce the VEM, its formulation, and a thorough discretization scheme for the aforementioned equations. Theoretical error estimates are rigorously derived and subsequently validated through numerical examples. This work represents an extension of the VEM approach to address time-fractional PDEs, offering a detailed analysis of both theoretical and computational aspects. In the future, we plan to explore the application of VEM to various fractional models and equations while also comparing their efficiency against existing numerical methods. Our primary focus will be on approximating real-time application problems, including anomalous physical processes, impulse, and fluid transport in complex geometries, and the impact of turbulence on design considerations.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The Laboratory Directed Research and Development (LDRD) program of Los Alamos National Laboratory partially supported the work of GM under project number 20220129ER. Los Alamos National Laboratory is operated by Triad National Security, LLC, for the National Nuclear Security Administration of U.S. Department of Energy (Contract No. 89233218CNA000001). G. Manzini is a member of the Gruppo Nazionale Calcolo Scientifico-Istituto Nazionale di Alta Matematica. Also, Vellore Institute of Technology, Vellore supported the work of M. Chandru under a SEED grant (Sanction Order No. SG20230081).

    The authors declare that they have no conflicts of interest.



    [1] N. N. Bogolyubov, On Some Statistical Methods in Mathematical Physics, Akademiya Nauk Ukrainskoǐ SSR, 1945. (in Russian)
    [2] N. N. Bogolyubov and Y. A. Mitropolsky, Asymptotic Methods in the Theory of Non-Linear Oscillations, Translated from the second revised Russian edition. International Monographs on Advanced Mathematics and Physics Hindustan Publishing Corp., Delhi, Gordon and Breach Science Publishers, New York, 1961.
    [3] A stochastic version of center manifold theory. Probab. Theory Related Fields (1989) 83: 509-545.
    [4] Averaging principle for a class of stochastic reaction-diffusion equations. Probab. Theory Related Fields (2009) 144: 137-177.
    [5] Averaging principle for nonautonomous slow-fast systems of stochastic reaction-diffusion equations: the almost periodic case. SIAM J. Math. Anal. (2017) 49: 2843-2884.
    [6] D. N. Cheban, Asymptotically Almost Periodic Solutions of Differential Equations, Hindawi Publishing Corporation, New York, 2009.
    [7] D. N. Cheban, Global Attractors of Nonautonomous Dynamical and Control Systems, 2nd edition, Interdisciplinary Mathematical Sciences, vol.18, River Edge, NJ: World Scientific, 2015. doi: 10.1142/9297
    [8] D. Cheban and J. Duan, Recurrent motions and global attractors of non-autonomous Lorenz systems, Dyn. Syst., 19 (2004), 41–59. doi: 10.1080/14689360310001624132
    [9] Periodic, quasi-periodic, almost periodic, almost automorphic, Birkhoff recurrent and Poisson stable solutions for stochastic differential equations. J. Differential Equations (2020) 269: 3652-3685.
    [10] Ju. L. Dalec'kiǐ and M. G. Kreǐn, Stability of Solutions of Differential Equations in Banach Space, Translated from the Russian by S. Smith. Translations of Mathematical Monographs, Vol. 43. American Mathematical Society, Providence, R.I., 1974.
    [11] (2014) Stochastic Equations in Infinite Dimensions. Cambridge: 2nd edition, Encyclopedia of Mathematics and its Applications, 152. Cambridge University Press.
    [12] J. Duan and W. Wang, Effective Dynamics of Stochastic Partial Differential Equations, Elsevier Insights. Elsevier, Amsterdam, 2014. doi: 10.1016/B978-0-12-800882-9.00001-9
    [13] (2002) Real Analysis and Probability. Cambridge: Revised reprint of the 1989 original. Cambridge Studies in Advanced Mathematics, 74. Cambridge University Press.
    [14] M. I. Freidlin and A. D. Wentzell, Random Perturbations of Dynamical Systems, Translated from the 1979 Russian original by Joseph Szücs. 3rd edition. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 260. Springer, Heidelberg, 2012. doi: 10.1007/978-3-642-25847-3
    [15] R. Z. Has'minskiǐ, On the principle of averaging the itô's stochastic differential equations, Kybernetika (Prague), 4 (1968), 260–279. (in Russian)
    [16] Weak averaging of semilinear stochastic differential equations with almost periodic coefficients. J. Math. Anal. Appl. (2015) 427: 336-364.
    [17] M. A. Krasnosel'skiǐ, V. Sh Burd and Yu. S. Kolesov, Nonlinear Almost Periodic Oscillations, Nauka, Moscow, 1970 (in Russian). [English translation: Nonlinear Almost Periodic Oscillations. A Halsted Press Book. New York etc.: John Wiley & Sons; Jerusalem- London: Israel Program for Scientific Translations. IX, 326 p., 1973]
    [18] (1943) Introduction to Non-Linear Mechanics. Princeton, N. J.: Annals of Mathematics Studies, no. 11. Princeton University Press.
    [19] (1978) Almost Periodic Functions and Differential Equations. Moscow: Moscow State University Press.
    [20] X. Liu and Z. Liu, Poisson stable solutions for stochastic differential equations with Lévy noise, Acta Math. Sin. (Engl. Ser.), In Press. (also available at arXiv: 2002.00395)
    [21] Almost automorphic solutions for stochastic differential equations driven by Lévy noise. J. Funct. Anal. (2014) 226: 1115-1149.
    [22] Normal form transforms separate slow and fast modes in stochastic dynamical systems. Phys. A (2008) 387: 12-38.
    [23] B. A. Ščerbakov, A certain class of Poisson stable solutions of differential equations, Differencial'nye Uravnenija, 4 (1968), 238–243. (in Russian)
    [24] B. A. Ščerbakov, The comparability of the motions of dynamical systems with regard to the nature of their recurrence, Differencial'nye Uravnenija, 11 (1975), 1246–1255. (in Russian) [English translation: Differential Equations 11 (1975), 937–943].
    [25] G. R. Sell, Topological Dynamics and Ordinary Differential Equations, Van Nostrand-Reinhold, 1971.
    [26] B. A. Shcherbakov, Topologic Dynamics and Poisson Stability of Solutions of Differential Equations, Stiinca, Kishinev, 1972.
    [27] B. A. Shcherbakov, Poisson Stability of Motions of Dynamical Systems and Solutions of Differential Equations, Știinţa, Chişinǎu, 1985. (in Russian)
    [28] K. S. Sibirsky, Introduction to Topological Dynamics, Kishinev, RIA AN MSSR, 1970. (in Russian). [English translation: Introduction to Topological Dynamics. Noordhoff, Leyden, 1975]
    [29] A. V. Skorokhod, Asymptotic Methods in the Theory of Stochastic Differential Equations, Translated from the Russian by H. H. McFaden. Translations of Mathematical Monographs, 78. American Mathematical Society, Providence, RI, 1989. doi: 10.1090/mmono/078
    [30] On large deviations in the averaging principle for SDEs with a "full dependence''. Ann. Probab. (1999) 27: 284-296.
    [31] Weak averaging of stochastic evolution equations. Math. Bohem. (1995) 120: 91-111.
    [32] Average and deviation for slow-fast stochastic partial differential equations. J. Differential Equations (2012) 253: 1265-1286.
  • Reader Comments
  • © 2021 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(2755) PDF downloads(373) Cited by(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog