Adaptive discretization of an integro-differential equation with a weakly singular convolution kernel

https://doi.org/10.1016/j.cma.2003.09.001Get rights and content

Abstract

An integro-differential equation involving a convolution integral with a weakly singular kernel is considered. The kernel can be that of a fractional integral. The integro-differential equation is discretized using the discontinuous Galerkin method with piecewise constant basis functions. Sparse quadrature is introduced for the convolution term to overcome the problem with the growing amount of data that has to be stored and used in each time-step. A priori and a posteriori error estimates are proved. An adaptive strategy based on the a posteriori error estimate is developed. Finally, the precision and effectiveness of the algorithm are demonstrated in the case that the convolution is a fractional integral. This is done by comparing the numerical solutions with analytical solutions.

Introduction

Fractional order operators (integrals and derivatives) have proved to be very suitable for modeling memory effects of various materials and systems of technical interest. In particular, they are very useful when modeling viscoelastic materials, see, e.g., [3], [4], [10]. The drawback of these models is that, when the response is integrated numerically, the whole previous stress or strain history must be included in each time-step. Rather few algorithms for integrating viscoelastic responses (integral equations with singular kernels) are available. Most of them are based on the Lubich convolution quadrature for fractional order operators, see [11] and, e.g., [8]. The Lubich convolution quadrature requires uniformly distributed time-steps. This is a cumbersome restriction, in particular, when analyzing non-linear viscoelastic responses. Furthermore, it is not possible to use adaptivity and goal oriented error estimates. It also restricts the possibility to use sparse time history. In the present work we discuss these difficulties in the context of the following integro-differential equationut(t)+∫0tβ(t−s)Au(s)ds=f(t),t∈(0,T),u(0)=u0,where ut=du/dt and A is a self-adjoint, positive definite, linear operator on a separable Hilbert space H with inner product (·,·) and norm ∥·∥. The operator A may be unbounded with domain of definition D(A)⊂H, but we assume that it has compact inverse, so that the spectral theorem applies. In applications H could typically be L2(Ω) for some spatial domain Ω, and A an elliptic partial differential operator with respect to the spatial variables, or the finite element approximation of such an operator. The abstract Hilbert space framework makes it possible to discuss time discretization without going into details about the spatial approximation. We assume that the data u0H and f(t)∈H are such that the equation has a unique, appropriately regular solution.

The kernel function β may be weakly singular but integrable. More precisely, we assume that β is real-valued, belongs to L1(0,T), and positive definite in the sense that, for any T⩾0,0T0tβ(t−s)ϕ(s)ϕ(t)dsdt⩾0,∀ϕ∈L2(0,T).Our chief example isβ(t)=1Γ(α)1t1−α,0<α⩽1,for which the convolution integral can be interpreted as an integral of fractional order α, see, e.g., the textbook [15], and (1.1) can be written asut(t)+D−α[Au](t)=f(t),t∈(0,T),u(0)=u0,where the operator Dα is the fractional order integral. In the limit α=0 the kernel β approaches the delta function and we obtain the parabolic equationut+Au=f;u(0)=u0,while specializing to α=1 results inut+∫0tAu(s)ds=f,which is equivalent to the hyperbolic equationutt+Au=ft;u(0)=u0,ut(0)=f(0).This means that by varying the fractional integral exponent we obtain a link between parabolic behavior, (e.g., heat conduction) and hyperbolic behavior, (e.g., wave propagation).

Here we develop an adaptive algorithm with a priori and a posteriori error estimates for solving (1.1). The a posteriori error estimate forms the basis for the adaptive strategy. For the numerical integration we adopt the discontinuous Galerkin method with piecewise constant basis functions. To overcome the problem with the growing amount of data, that has to be stored and used in each time-step, we introduce sparse quadrature for the convolution integral. Sparsely distributed time-steps are used in the distant part of the history while small steps are used in the most recent part. The idea is to break up the convolution structure by using piecewise linear interpolants between the large steps in the distant part of the history. This was first studied in [16], [18].

The present work is a further development of the study by McLean et al. [13]. Indeed, they allowed variable time-steps and introduced sparse quadrature, but their study was limited to a priori error estimates.

Similar considerations apply to fractional order differential equations, Dαu+Au=f, which can be rewritten as Volterra integral equations of the second kind by applying the integral operator Dα, see [6]. The kernel function in the resulting integral equation is then of the same kind as in the present study. Fractional differential equations are studied in this way in [5], [6], [7]. The numerical schemes are based on the Lubich convolution quadrature and therefore need equispaced grids, or alternatively logarithmically distributed time-steps, as in [9]. This prevents the use of adaptivity and sparse quadrature. These issues will be addressed in the forthcoming paper [2] using the methods of the present work.

Section snippets

The discontinuous Galerkin method

Let 0=t0<t1<⋯<tn−1<tn<⋯<tN=T be a temporal mesh with time intervals In=(tn−1,tn) and steps kn=tntn−1. We define the finite element space (with subscript D for “discrete”)WD={w:w(t)=wnfort∈In,wn∈D(A),n=1,…,N}.Note that w∈WD may be discontinuous at t=tn; we write wn=w|In=wn=wn−1+ and we let [w]n=wn+wn=wn+1wn denote the jump.

The approximation U∈WD of the solution u of (1.1) is given byU∈WD,withU0=u0,andforn=1,…,N,InUt(t)+∫0tβ(t−s)AU(s)ds−f(t),v(t)dt+([U]n−1,vn−1+)=0∀v∈WD.Here (·,·) denotes

A priori error estimate

The following a priori error estimate is Theorem 6.1 in [13]. We repeat the proof, expressed in our present notation. We recall that ∥·∥ denotes the norm in H.

Theorem 3.1

Let u and U be the solutions of , , respectively. Let ũWD denote the piecewise constant interpolant determined by ũ(t)=u(tn) for tIn. Then, for all tN⩾0,∥UN−u(tN)∥⩽2∥β∥L1(0,tN)tN0∥A(u(t)−ũ(t))∥dt⩽2∥β∥L1(0,tN)n=1NknIn∥Aut(t)∥dt.

Proof

We begin by showing that the bilinear form B is positive. By adding the first and second variants of B(v,v

A posteriori error estimate

In order to obtain a representation of the error we introduce the adjoint problem with arbitrary data φH and T=tN:−ψt(t)+∫tTβ(s−t)Aψ(s)ds=0,t∈(0,T),ψ(T)=φ,By the transformation tT−t, Eq. (4.1) takes the same form as (1.1) with f=0, whose solution we denote by u(t)=E(t)u0 following [12]. The solution of (4.1) is thus ψ(t)=E(Tt)φ.

Eq. (4.1) is the adjoint of (2.12). Using the second variant of (2.7) it can be written in weak form asψ∈W:B(w,ψ)=(wN,φ)∀w∈W.Using w=e in (4.2) and v=ψ in (2.12) we

Sparse quadrature

Sparse quadrature was introduced in [13], but only for the case of a kernel without singularity. Here we use the same procedure, but extend it to the case of the singular kernel with an emphasis on adaptivity.

We introduce time levels 0=M0<M1<M2<⋯ and replace the function sβ(t−s) in the integral qn(AU) in (2.4) by a piecewise linear interpolant,β̃(t,s)=β(t−tMl−11,l(s)+β(t−tMl2,l(s),s∈[tMl−1,tMl],l=1,…,L,β(t−s),s∈[tML,t],whereφ1,l(s)=tMl−sKl,φ2,l(s)=s−tMl−1Kl,Kl=tMl−tMl−1,and L is the

Adaptive strategy

The goal of the adaptive strategy is to provide a solution to the integro-differential equation within a user defined tolerance, TOL. The adaptive strategy is based on a posteriori error estimates of the total error at the final time T=tN. More precisely, we base the strategy on two different estimates of the error, the exact Galerkin error (5.24) given byEG1=sup∥φ∥=1n=1NIn(R,ψ)dt+([U]n−1n−1+),or the estimated Galerkin error (5.26) given byEG2=sup∥φ∥=1n=1Nkn2rG,nsG,n,where, with the choice

Numerical experiments

In the following examples we consider (1.4) with A=1 for different values of the fractional integral exponent α. The equation then readsut(t)+1Γ(α)0tu(s)(t−s)1−αds=f(t),t∈(0,T),u(0)=u0.The weights ωnj in (2.4) and ω̃nj in (5.7) are integrated analytically. The mean value f̄n in (2.4) of the source term f and the integrals over In in the a posteriori estimates are computed using the trapezoidal rule, which means that additional errors are introduced. These errors are not taken into account in

Acknowledgements

The work is supported, partly, by the Swedish Research Council (VR). We would like to thank Professor V. Thomée for pointing out an error in an earlier version of the manuscript.

References (18)

There are more references available in the full text version of this article.

Cited by (31)

  • Error estimates of a high order numerical method for solving linear fractional differential equations

    2017, Applied Numerical Mathematics
    Citation Excerpt :

    Ervin and Roop [10,11] used finite element methods to find the variational solution of the fractional advection dispersion equation, in which the fractional derivative depends on the space, related to the nonlocal operator, but the time derivative term is of first order, related to the local operator. Adolfsson et al. [1,2] considered an efficient numerical method to integrate the constitutive response of fractional order viscoelasticity based on the finite element method. Li et al. [25] considered a time fractional partial differential equation by using the finite element method and obtain the error estimates in both semidiscrete and fully discrete cases.

  • Discontinuous Galerkin method for an integro-differential equation modeling dynamic fractional order viscoelasticity

    2015, Computer Methods in Applied Mechanics and Engineering
    Citation Excerpt :

    See [12] and references therein for examples of application of this approach and a different approach, the so-called sparse quadrature, that was introduced in [15], but only for the case of a kernel without singularity. See [16], where the same procedure has been extended to the case of the singular kernel. We note that, when the exponential decaying kernel, in linear viscoelasticity, is represented as a Prony series, it results in a recurrence formula for history updating; see [10].

View all citing articles on Scopus
View full text