Loading [MathJax]/extensions/TeX/boldsymbol.js
Special Issues

An adjoint-based a posteriori analysis of numerical approximation of Richards equation

  • This paper formulates a general framework for a space-time finite element method for solving Richards Equation in one spatial dimension, where the spatial variable is discretized using the linear finite volume element and the temporal variable is discretized using a discontinuous Galerkin method. The actual implementation of a particular scheme is realized by imposing certain finite element space in temporal variable to the variational equation and appropriate "variational crime" in the form of numerical integrations for calculating integrations in the formulation. Once this is in place, adjoint-based error estimators for the approximate solution from the scheme is derived. The adjoint problem is obtained from an appropriate linearization of the nonlinear system. Numerical examples are presented to illustrate performance of the methods and the error estimators.

    Citation: Victor Ginting. An adjoint-based a posteriori analysis of numerical approximation of Richards equation[J]. Electronic Research Archive, 2021, 29(5): 3405-3427. doi: 10.3934/era.2021045

    Related Papers:

    [1] Victor Ginting . An adjoint-based a posteriori analysis of numerical approximation of Richards equation. Electronic Research Archive, 2021, 29(5): 3405-3427. doi: 10.3934/era.2021045
    [2] Hao Wang, Wei Yang, Yunqing Huang . An adaptive edge finite element method for the Maxwell's equations in metamaterials. Electronic Research Archive, 2020, 28(2): 961-976. doi: 10.3934/era.2020051
    [3] Zuliang Lu, Fei Huang, Xiankui Wu, Lin Li, Shang Liu . Convergence and quasi-optimality of $ L^2- $norms based an adaptive finite element method for nonlinear optimal control problems. Electronic Research Archive, 2020, 28(4): 1459-1486. doi: 10.3934/era.2020077
    [4] Hsueh-Chen Lee, Hyesuk Lee . An a posteriori error estimator based on least-squares finite element solutions for viscoelastic fluid flows. Electronic Research Archive, 2021, 29(4): 2755-2770. doi: 10.3934/era.2021012
    [5] Shuhao Cao . A simple virtual element-based flux recovery on quadtree. Electronic Research Archive, 2021, 29(6): 3629-3647. doi: 10.3934/era.2021054
    [6] Yue Feng, Yujie Liu, Ruishu Wang, Shangyou Zhang . A conforming discontinuous Galerkin finite element method on rectangular partitions. Electronic Research Archive, 2021, 29(3): 2375-2389. doi: 10.3934/era.2020120
    [7] Jun Pan, Yuelong Tang . Two-grid $ H^1 $-Galerkin mixed finite elements combined with $ L1 $ scheme for nonlinear time fractional parabolic equations. Electronic Research Archive, 2023, 31(12): 7207-7223. doi: 10.3934/era.2023365
    [8] Ying Liu, Yanping Chen, Yunqing Huang, Yang Wang . Two-grid method for semiconductor device problem by mixed finite element method and characteristics finite element method. Electronic Research Archive, 2021, 29(1): 1859-1880. doi: 10.3934/era.2020095
    [9] Fenglin Huang, Yanping Chen, Tingting Lin . An error estimator for spectral method approximation of flow control with state constraint. Electronic Research Archive, 2022, 30(9): 3193-3210. doi: 10.3934/era.2022162
    [10] Chunmei Wang . Simplified weak Galerkin finite element methods for biharmonic equations on non-convex polytopal meshes. Electronic Research Archive, 2025, 33(3): 1523-1540. doi: 10.3934/era.2025072
  • This paper formulates a general framework for a space-time finite element method for solving Richards Equation in one spatial dimension, where the spatial variable is discretized using the linear finite volume element and the temporal variable is discretized using a discontinuous Galerkin method. The actual implementation of a particular scheme is realized by imposing certain finite element space in temporal variable to the variational equation and appropriate "variational crime" in the form of numerical integrations for calculating integrations in the formulation. Once this is in place, adjoint-based error estimators for the approximate solution from the scheme is derived. The adjoint problem is obtained from an appropriate linearization of the nonlinear system. Numerical examples are presented to illustrate performance of the methods and the error estimators.



    The subject of investigation in this paper is numerical solutions of the Richards Equation [23]. This equation is a governing mathematical principle for modeling the water flow in an unsaturated porous medium that is driven by the gravity and capillarity that disregards air flow. Since ability to construct closed form solutions to this equation is very limited (see for example [25,18,22,24] for some related effort on the subject), a reliance on numerical approximations is a necessity. However, even with the emergence of many advances of computing technology, this equation remains one of the most challenging problems in porous media flow and transport. Recent review on its numerical solutions can be found in [17].

    There are several outstanding issues attributed to the challenge. Richards Equation is strongly nonlinear, which appears as the dependence of the soil unsaturated hydraulic conductivity (κ) and the water content (ϑ) to the pressure head (u). Note that the presence of the water content in the equation is in terms of its temporal rate of change. Inclusion of gravity in the Darcy's velocity q, written as q=κ(u)z(uz), can potentially create instability in the numerical solutions, in particular, when simulating dry soil conditions. The variable z in the expression of q denotes the vertical spatial coordinate, which is positive in the downward direction. It represents the influence of gravity to the flow. Furthermore, some of the more realistic scenario requires taking into account the soil heterogeneity in the simulations.

    The mixed-form (or coupled-form) of Richards Equation in one dimensional soil column is written as follows:

    {tϑ(u)z(κ(u)z(uz))=0,in(a,b)×(0,T),u(z,0)=u0(z),z(a,b),BoundaryConditions:Bau=ga(t),Bb=gb(t). (1.1)

    Two typical boundary conditions are

    {Bau=κ(u)z(uz)|(a,t),Bbu=u(b,t)}or{Bau=u(a,t),Bbu=u(b,t)}.

    Here we assume that ϑ:(,0](0,1) and κ:(,0](κmin,κs) with κs>κmin>0. The choice (,0] as the domain of ϑ and κ is done to reflect the physical relevance that the pressure head u is always nonpositive and Richards Equation typically models unsaturated flow.

    The major theme in this paper is two-fold. One aspect centers on the development of a numerical approximation of Richards Equation in space-time finite element methods obtained from an appropriate variational formulation. Space-time finite element methods have been previously used for parabolic equations (see for example [20]) and for reaction-diffusion system (see for example [15]). A recent work on application of control volume finite element in combination with method of lines for solving Richards Equation is recorded in [10]. To the best of the author's knowledge, there has not been any attempt to apply space-time finite element methods technique to Richards Equation. In particular, a finite volume element spatial discretization (with linear finite element) is employed due to its inherent local mass conservation property. This is an important trait commonly desired and in some cases imperative in order to produce reliable numerical simulations of flow and transport in porous media (see for example [5,8,16]). The space-time variational formulation in combination with a certain variational crime in the form of numerical integration techniques would in turn yield implementable time marching schemes for approximate solution of Richards Equation.

    The other aspect is concerned with an a posteriori error estimation of the resulting numerical approximation. In this regard, some investigations on a posteriori error analysis of numerical methods for Richards Equation are already available. Bause and Knabner [3] use adaptive mixed hybrid finite element discretizations to solve Richards Equation, where the adaptivity is performed under an a posteriori error indicator that is based on either superconvergence or residual of the approximation. Baron et al. [2] employ Discrete Duality Finite Volume (DDFV) scheme along with second-order backward differentiation formula to solve the equation. They derive an a posteriori error bound of the approximation using the equilibrated fluxes method. Bernardi et al. [4] perform a semi discretization of Richards Equation by finite element method and apply Backward Euler scheme to get the full discretization. Then a posteriori error bounds are derived that aim at distinguishing components of contribution of spatial discretization from temporal discretization.

    In many practical situations, it may not be necessary to measure global property of the approximate solution. More often, an accuracy is desired only for some specified quantities of interest associated with the numerical approximation, which is usually expressed as a functional of the approximate solution. For this purpose, a suitable a posteriori analysis is based on duality, adjoint operators governing the generalized Green's function, and a variational formulation. This approach is adopted in this paper. It is also suitable because, as mentioned before, the numerical approximation of Richards Equation is based on certain variational equations. Utilizations of adjoint equations are not new (see for example [21] for an extensive exposition on their applications). On various roles of adjoint methodologies in performing a posteriori error estimations of numerical methods for differential equations, one can consult [19,15,11,13,12,14,1] and references therein.

    The rest of this paper is organized as follows. A space-time variational equation governing the numerical approximation of Richards Equation is derived in Section 2. The description includes examples of application of numerical integration to produce the time marching schemes. Section 3 carries out the formulation of an adjoint-based a posteriori error analysis of the quantities of interest calculated using the approximate solution. Since the variational equation of the solution is nonlinear, an appropriate linearization is conducted that would make construction of the corresponding adjoint equation amenable. Some numerical examples to demonstrate performance of the numerical methods and the error estimators are shown in Section 4. Here much of the effort is devoted to illustrate global accuracy of numerical methods and reliability of the error estimators in terms of their capability to decompose the total error into relevant components. A comparison to the actual error in some specified quantities of interest is conducted. Finally, the conclusion and future work is discussed in Section 5.

    In what follows, we assume that

    {Bau=κ(u)z(uz)|(a,t)=ga(t),Bbu=u(b,t)=0}. (2.1)

    are supplied to (1.1). Denoting

    H1D={w:[a,b]R:wL2(a,b),wL2(a,b),w(b)=0},

    the solution of (1.1) supplied with (2.1) satisfies

    {tϑ(u),v+A(u;u,v)+ga(t)v(a)=0,vH1D,u(,0),v=u0,v,vH1D. (2.2)

    Here , is the usual scalar product in L2(a,b), and A:C[a,b]×H1D×H1DR is defined as

    A(w;u,v)=κ(w)z(uz),zv.

    The spatial domain (a,b) is partitioned into a collection of M subintervals Th, such that τj=(zj1,zj)Th, with length hj=zjzj1, for j=1,,M and (a,b)=Mj=1τj, where h=max{hj:1jM}. On this Th, let

    Xh={wH1D : winτj islinear τjTh}=span{ϕj}M1j=0,

    where ϕj(z) is the usual 'hat' function such that ϕj(zi)=δij.

    The finite volume approximations rely on a local conservation property associated with the governing equation, in particular with respect to the second order differential operator in (1.1). To fix the idea, consider τ=(zl,zr)(a,b) and apply fundamental theorem of calculus to get

    τz(κ(u)z(uz))dz=κ(u)z(uz)|τ:=κ(u)z(uz)|zrzl. (2.3)

    The above τ is called a control volume. We choose M control volumes. Specifically, given zj, for j=1,,M1, we set τj=(zj1/2,zj+1/2), where zj1/2 is the mid-point of τj and zj+1/2 is the mid-point of τj+1. For z0, we set τ0=(z0,z1/2). Collection of these control volumes is denoted by Th. On this Th, let

    Yh={ηL2(a,b) : ηinτ isconstant τjTh}=span{ϕj}M1j=0,

    where ϕj(z) is a piecewise constant function such that it is equal to 1 in τj and zero elsewhere. Set Jh:H1DYh such that [Jhv](zj)=v(zj), i.e., given vH1D, Jhv is a piecewise constant function over Th. A standard interpolation estimate suggests

    Jhvvh2zv,forvH1D, (2.4)

    where =,. By the Cauchy-Schwarz inequality, this implies

    χ,Jhvvh2χzv,forχL2(a,b),vH1D. (2.5)

    To express (2.3) in a variational setting, define Ah:C[a,b]×H1D×H1DR as

    Ah(w;u,v)=τjThκ(w)z(uz)|τja[Jhv](zj).

    Exclusion of a in the above equation is due to the Neumann boundary condition at that point. Next, we quantify the discrepancy of Ah(;,) from A(;,).

    Proposition 2.1. Let wC[a,b], u,vH1D. Then

    A(w;u,v)=Ah(w;u,v)+εA(w;u,v), (2.6)

    where

    εA(w;u,v)=τjThτjz(κ(w)z(uz))(Jhvv)dz. (2.7)

    Furthermore, when u,wXh and κ(r) is bounded for every r(,0], then

    |εA(w;u,v)|Cκ2hzwz(uz)zv, (2.8)

    where Cκ=supu(,0]|κ(u)|.

    Proof. For any τjTh, integration by parts gives

    τjz(κ(w)z(uz))vdz=τjκ(w)z(uz)zvdzκ(w)z(uz)v|τj,

    which when applied to A(;,) gives

    A(w;u,v)=τjThτjκ(w)z(uz)zvdz=τjTh(τjz(κ(w)z(uz))vdz+κ(w)z(uz)v|τja). (2.9)

    For j=0, set Kj=τj. For j=1,,M1, fix a τjTh and suppose τj,τj+1Th are such that Kj=τjτj=(xj1/2,xj), Kj+1=τj+1τj=(xj,xj+1/2). By fundamental theorem of calculus,

    Kez(κ(w)z(uz))Jhvdz=κ(w)z(uz)|Ke[Jhv](zj), (2.10)

    for e=j,j+1. Recognizing that

    zu|τj=e=j,j+1zu|Ke+zu|x+jxj,

    we may apply (2.10) in Ah(;,) to get

    Ah(w;u,v)=τjTh(τjz(κ(w)z(uz))Jhvdz+κ(w)z(uz)Jhv|τja). (2.11)

    Subtraction of (2.9) from (2.11) and recalling that Jhv(zj)=v(zj) gives (2.6).

    Furthermore, when w,uXh, product rule of differentiation for zτj gives

    z(κ(w)z(uz))=κ(w)zwz(uz)+0,

    so using this identity and the Cauchy-Schwarz inequality,

    |εA(w;u,v)|=|τjThτjκ(w)zwz(uz)(Jhvv)dz|CκτjThzwz(uz)L2(τj)JhvvL2(τj)Cκzwz(uz)JhvvCκ2hzwz(uz)zv.

    This completes the proof.

    Remark 2.1. The foregoing exposition gives an indication that w,Jhvv0 and Ah(w;u,v)A(w;u,v) as h0. This will play a role later on in the a posteriori error analysis. Various estimates such as described in (2.4), (2.5), and (2.8) have been established in several literatures on finite volume element methods (see for example [6,7,9] and references therein).

    In a similar fashion to the spatial variable, we partition [0,T] into a collection of subintervals Ik, such that In=[tn1,tn]Ik with time step kn=tntn1 and [0,T]=Nn=1In, where k=max{kn:1nN}. We denote the jump of a function w(,t) across tn by [w]n=w+nwn, where w+n=limst+nw(,s) and wn=limstnw(,s). On every space-time slab [a,b]×In, the approximate solution is sought in a functional space that contains functions that are piecewise linear polynomial in spatial variable and polynomial of degree q in temporal variable. In particular, we define

    Wqh(In)={v:[a,b]×InR:w(z,t)=qj=0tjvj,n(z),withvj,nXh}. (2.12)

    We denote by Wqh the space of functions defined on [a,b]×[0,T] such that restriction of wWqh to [a,b]×In belongs to Wqh(In). The approximation amounts to finding ˜uWqh that is governed by

    {Nn=1Rh,n(˜u;˜u,w)=0foreverywWqh,˜u0,χ=u0,χforeveryχXh, (2.13)

    where

    Rh,n(˜u;˜u,w)=In(tϑ(˜u),Jhw+Ah(˜u;˜u,w)+ga(t)w(a,t))dt+[ϑ(˜u)]n1,Jhw+n1. (2.14)

    Notice that the first equation in (2.13) is a global formulation in that the integration is over (0,T). While an implementation based on this formulation is possible, it is perhaps more amenable to construct an implementation that is local over In for n=1,,N. The corresponding equation for this formulation can be derived from (2.13) by choosing wWqh such that w|(a,b)×In=vWqh(In) and it is zero everywhere else, which yields

    Rh,n(˜u;˜u,v)=0foreveryvWqh(In). (2.15)

    In what follows, we describe two specific examples that transform (2.15) into computable algebraic schemes.

    Here ˜uW0h, i.e.,

    ˜u|In=˜u+n1=˜un=v0,nXh, (2.16)

    which for every w0Xh is governed by

    knAh(v0,n;v0,n,w0)+w0(a)Inga(t)dt+ϑ(v0,n)ϑ(˜un1),Jhw0=0, (2.17)

    for n=1,,N. Notice that (2.17) is mimicking the Backward Euler difference scheme with v0,nXh being the unknown function to be solved. In particular, setting (U0,n,U1,n,,UM1,n)=UnRM, such that

    v0,n=M1j=0Uj,nϕj, (2.18)

    then Un is governed by

    G(Un)=0, (2.19)

    where G:RMRM such that Gi:RMR, for i=0,1,,M1, is constructed from the left hand side of (2.17) by replacing w0 by ϕi. Here, the dependence on Un is realized through (2.18). Clearly (2.19) is a nonlinear algebraic system of equations governing Un.

    Here ˜uW1h, i.e.,

    ˜u|In=v0,n+tv1,n,tIn,v0,n,v1,nXh, (2.20)

    which implies that ˜u+n1=v0,n+tn1v1,n and ˜un=v0,n+tnv1,n. We may equivalently write

    ˜u|In=tntkn˜u+n1+ttn1kn˜un,tIn,˜u+n1,˜unXh. (2.21)

    Choosing w=tntknψ+n1 with ψ+n1Xh, and using integration by parts along with acknowledging some cancellations,

    Intϑ(˜u),Jhwdt+[ϑ(˜u)]n1,Jhw+n1=k1nInϑ(˜u)ϑ(˜un1),Jhψ+n1dt.

    In a similar fashion, using w=ttn1knψn with ψnXh yields

    Intϑ(˜u),Jhwdt+[ϑ(˜u)]n1,Jhw+n1=k1nInϑ(˜un)ϑ(˜u),Jhψndt.

    Thus ˜uW1h satisfies (2.21) and for every n=1,,N, it is governed by

    {Inϑ(˜u)ϑ(˜un1),Jhψ+n1dt+In(tnt)(Ah(˜u;˜u,ψ+n1)+ga(t)ψ+n1(a))dt=0,Inϑ(˜un)ϑ(˜u),Jhψndt+In(ttn1)(Ah(˜u;˜u,ψn)+ga(t)ψn(a))dt=0, (2.22)

    where

    ˜u+n1=M1j=0U+j,n1ϕj,˜un=M1j=0Uj,nϕj. (2.23)

    Setting

    (U+0,n1,U+1,n1,,U+M1,n1)=U+n1RM

    and

    (U0,n,U1,n,,UM1,n)=UnRM,

    and Un=(U+n1,Un)R2M, then (2.9) yields

    G(Un)=0, (2.24)

    where G:R2MR2M, with G=(G+,G), and G+:R2MRM and G:R2MRM such that G+i:R2MR and Gi:R2MR are respectively constructed from the left hand side of (2.9) by setting ψ+n1=ψn=ϕi, for i=0,1,,M1.

    The preceding description is a derivation of algebraic equations governing the approximation that is faithful to the variational equation (2.15) and the choice of polynomial degree of the temporal variable. Still, for a completely implementable scheme, one must rely on further approximation of the integrations appeared in (2.17) and (2.9). In the current setting, there are two integrations that need to be approximated: the spatial integration , and the temporal integration Indt. Utilization of various numerical integration techniques are pretty common. In particular, in the standard finite element methods for typical steady state problems, forms/functionals in the variational equations, which are expressed as integrations of spatial variables, are approximated by various Gaussian quadratures. This is clearly applicable for ,. To minimize the associated pollution to the global accuracy of the approximation, the numerical integrations must be chosen such that the degree of their errors is of similar order to the errors corresponding to Wqh.

    Furthermore, what is more crucial in this case is how Indt is to be approximated. We note that the only temporal integration in (2.17) is the one associated with the Neumann condition ga(t), and for this a right hand point rule resulting in

    Inga(t)dtknga(tn)

    is adequate.

    Derivation of numerical integrations for (2.9) is a bit more involved. A viable option is the following two point Gaussian quadrature

    Inf(t)dtkn22=1f(t,n),wheret1,n=kn23+tn1+tn2andt2,n=kn23+tn1+tn2. (2.25)

    With this, set

    ˜u1,n=˜u(,t1,n)=γ˜u+n1+(1γ)˜un,˜u2,n=˜u(,t2,n)=(1γ)˜u+n1+γ˜un, (2.26)

    where γ=1+323. The resulting approximations G+n,iG+i and Gn,iGi are expressed as

    G+n,i(Un)=122=1ϑ(˜u,n)ϑ(˜un1),ϕi+knc+(Ah(˜u,n;˜u,n,ϕi)+ga(t,n)δi0)Gn,i(Un)=122=1ϑ(˜un)ϑ(˜u,n),ϕi+knc(Ah(˜u,n;˜u,n,ϕi)+ga(t,n)δi0)

    where c+l=cr=γ, c+r=cl=1γ, and δij is the usual Kronecker delta.

    The approach proceeds with the construction of algebraic equations for (˜u1,n,˜u2,n) where ˜un appearing on the second equation above is represented as

    ˜un=1γ12γ˜u1,nγ12γ˜u2,n, (2.27)

    which is obtained from (2.26). Thus, with (G+n,i,Gn,i) replacing (G+i,Gi), (2.24) is solved to get (˜u1,n,˜u2,n), after which ˜un is recovered from (2.27).

    In many realistic situations, it is often desirable to achieve an acceptable level of accuracy of a numerical approximation in some quantities of interest. Relevant examples include average water content over a certain region and at some time instances or the water content at some locations. Along this line of argument, it may be computationally infeasible as well as very inefficient to attempt to control the error in a global fashion when all that is required is accuracy on those aforementioned quantities. A practical alternative is to estimate the error of the numerical approximation in the specified quantity of interest, whose representation is expressed as a functional of u:

    [Q(u)](T)=ϑ(u(,T)),ψT+T0(ϑ(u),ψ+u(a,t)ψa)dt, (3.1)

    for given data ψT:(a,b)R, ψ:(a,b)×(0,T)R, and ψa:(0,T)R. If one wants to quantify the (averaged) water content at time t=T, then (3.1) uses ψ=0, ψa=0, and ψT is set to be a piecewise constant function in the spatial variable that reflects the desired nature of the average quantity. On the other hand, if an accumulated water content is the quantity of interest to be approximated, then ψT=0, ψa=0, and ψ=1 in (3.1).

    To derive the error in approximating Q(u), we use a generalized Green's function that solves the adjoint problem corresponding to a special choice of (adjoint) data ψT, ψ, and ψa, as illustrated by the description in the previous paragraph. As alluded to earlier, the formulation of an adjoint problem broadens the applications of Green's functions (see for example [19,15,1] and references therein). However, an adjoint operator formally corresponds to a linear operator. Since Richards Equation and the associated numerical approximation are nonlinear problem, we must perform a linearization, after which the adjoint problem is built to correspond to that linearized representation.

    The nonlinearity in Richards Equation stems from θ(u) and κ(u), and that is where the linearization effort is concentrated on. To this end, letting

    ˜uσ=˜u+σ(u˜u),forσ[0,1], (3.2)

    the Mean Value Theorem for integral gives

    ϑ(u)ϑ(˜u)=¯ϑ(u˜u),where¯ϑ=10ϑ(˜uσ)dσ. (3.3)

    Furthermore, setting F:H1(a,b)R by

    F(w)=κ(w)z(wz), (3.4)

    its Fréchet derivative is F(w):H1(a,b)R such that

    F(w)v=κ(w)zv+(κ(w)z(wz))v (3.5)

    Utilizing again the Mean Value Theorem for integral, one gets

    F(u)F(˜u)=10F(uσ)(u˜u)dσ=¯κz(u˜u)+¯v(u˜u), (3.6)

    where

    ¯κ=10κ(˜uσ)dσand¯v=10κ(˜uσ)z(˜uσz)dσ. (3.7)

    At this stage, we are in a position to formulate the adjoint problem. For t[T,0], let φ(,t)H1D satisfy

    {w,¯ϑtφ+zw,¯κzφ+w,¯vzφ=w,¯ϑψ+w(a,t)ψa,t<T,w(,T),¯ϑφ(,T)=w(,T),¯ϑψT, (3.8)

    for every w(,t)H1D. Here φ is solution to the adjoint problem, which is governed by a linear problem as stated in (3.8). The two theorems below state the quantification of error in the approximation of Q(u), which is written in terms of residuals of ˜u weighted against φ. In what follows, we use

    Rn(˜u;˜u,w)=In(tϑ(˜u),w+A(˜u;˜u,w)+ga(t)w(a,t))dt+[ϑ(˜u)]n1,w+n1. (3.9)

    Theorem 3.1. For ˜uWqh in (2.13) and u in (2.2),

    [Q(u)](T)[Q(˜u)](T)=E0+E1+E2+E3, (3.10)

    where

    E0=ϑ(u0)ϑ(˜u0),φ0,E1=Nn=1Rh,n(˜u;˜u,φ),E2=Nn=1εA,n(˜u;˜u,φ)dt,E3=Nn=1εh,n(˜u;φ), (3.11)

    with Rh,n(˜u;˜u,φ) as expressed in (2.14), and

    εA,n(˜u;˜u,w)=InεA(˜u;˜u,w)dt,εh,n(˜u;w)=Intϑ(˜u),wJhwdt+[ϑ(˜u)]n1,(wJhw)+n1. (3.12)

    Proof. Substitute w=e=u˜u in (3.8) so that

    e,¯ϑtφ+ze,¯κzφ+e,¯vzφ=ϑ(u)ϑ(˜u),tφ+κ(u)z(uz)κ(˜u)z(˜uz),zφ=(tϑ(˜u),φ+A(˜u;˜u,φ)+g(t)φ(a,t))tϑ(u)ϑ(˜u),φ, (3.13)

    where we have used the first equation in (2.2). Since

    e,¯ϑψ+e(a,t)ψa=ϑ(u)ϑ(˜u),ψ+(u(a,t)˜u(a,t))ψa, (3.14)

    integration of (3.8) over In yields

    ϑ(un)ϑ(˜un),φn+In(ϑ(u)ϑ(˜u),ψ+(u˜u))(a,t)ψa)dt=ϑ(un1)ϑ(˜un1),φ+n1Rn(˜u;˜u,φ), (3.15)

    where Rn is as expressed in (3.9). Next we sum up (3.15) over n=1,,N and take advantage of the continuity of φ in t to get

    [Q(u)](T)[Q(˜u)](T)=ϑ(u0)ϑ(˜u0),φ0Nn=1Rn(˜u;˜u,φ). (3.16)

    The residual Rn(˜u;˜u,φ) can be decomposed into three components, which is obtained by adding and subtracting Rh,n(˜u;˜u,φ):

    Rn(˜u;˜u,φ)=Rh,n(˜u;˜u,φ)+δRn(˜u;˜u,φ), (3.17)

    where

    δRn(˜u;˜u,w)=Rn(˜u;˜u,w)Rh,n(˜u;˜u,w)=Intϑ(˜u),wJhwdt+[ϑ(˜u)]n1,(wJhw)+n1+In(A(˜u;˜u,w)Ah(˜u;˜u,w))dt=εh,n(˜u;w)+InεA(˜u;˜u,w)dt=εh,n(˜u;w)+εA,n(˜u;˜u,w). (3.18)

    Putting(3.18) to (3.17) and in turn to (3.16) completes the proof.

    Theorem 3.2. For ˜uWqh in (2.15) and u in (2.2),

    [Q(u)](T)[Q(˜u)](T)=E0+E1+E2+E3, (3.19)

    where

    E0=ϑ(u0)ϑ(˜u0),φ0,E1=Nn=1Rn(˜u;˜u,φπqhφ),E2=Nn=1εA,n(˜u;˜u,Πqhφ),E3=Nn=1εh,n(˜u;πqhφ), (3.20)

    and πqhφWqh is the usual projection of φ onto Wqh.

    Proof. Most derivation steps follow the proof of Theorem 3.1 up to (3.16):

    [Q(u)](T)[Q(˜u)](T)=ϑ(u0)ϑ(˜u0),φ0Nn=1Rn(˜u;˜u,φ). (3.21)

    At this stage, we intend to insert (2.15), which is valid when the test function is πqhφWqh. To do so, add and subtract Rn(˜u;˜u,πqhφ) so that

    Rn(˜u;˜u,φ)=Rn(˜u;˜u,φπqhφ)+Rn(˜u;˜u,πqhφ)=Rn(˜u;˜u,φπqhφ)+Rn(˜u;˜u,πqhφ)Rh,n(˜u;˜u,πqhφ)=Rn(˜u;˜u,φπqhφ)+εA,n(˜u;˜u,πqhφ)+εh,n(˜u;πqhφ), (3.22)

    where similar equation to (3.18) has been used, and εA,n and εh,n are as in (3.12). Putting all these results back to (3.21) completes the proof.

    Several numerical examples are presented in this section to achieve two goals: 1) to investigate the global/norm-based accuracy of the proposed approximation, and 2) to validate the robustness of error estimators that are derived from Theorem 3.1 and Theorem 3.2. While the former cannot satisfactorily substitute for a rigorous a priori error analysis, at least it should give an illustrative indicator on the global convergence property of the approximation. With respect to the latter, we only concentrate on the estimators' accuracy in predicting error and their capability to decompose it into relevant components. Various pertinent applications of the proposed error estimators to other aspects in numerical simulation of Richards Equation, such as its role in adaptivity, will be a subject of future work.

    A uniform set of discretization parameters hi=h=(ba)/M and kn=k=T/N is used to construct the algebraic equations (2.19). The time marching is executed by solving this system using the standard Newton's method of iteration.

    While the proposed procedures enjoy a flexibility in their implementation, a closed form solution of Richards Equation is needed for the purpose of assessing their performance. As alluded to in the introduction, it is only on a very rare occasion that a closed form solution of Richards Equation is available. One such instance is when the constitutive relations are expressed as

    κ(u)=κseαu,andϑ(u)=ϑr+(ϑsϑr)eαu, (4.1)

    where κs is the saturated hydraulic conductivity, ϑr and ϑs are respectively the residual and saturated water content, and α is the reciprocal of vertical height associated with the capillary fringe. When g(t)=g=constant, =a,b, then the closed form solution can be expressed as a series representation:

    u(z,t)=α1ln(κ1sw(z,t)), (4.2)

    where

    w(z,t)=C1+C2eαz+eαz/2n=1wn(t)ϕn(z),with (4.3)
    wn(t)=wn(0)eμnt,μn=κs(α2+4λ2n)4α(ϑsϑr)>0,wn(0)=1ϕn,ϕnL0(κseαu0(z)C1C2eαz)eαz/2ϕn(z)dz. (4.4)

    The pair {ϕn,λn}n=1 constitutes an eigenfunction and an eigenvalue that satisfies

    {ϕn=λ2nϕnin(a,b),˜Baϕn=0,˜Bbϕn=0, (4.5)

    where ˜B are boundary conditions for w, which are appropriately derived from B via the relation w(z,t)=κseαu(z,t). The constants {C1,C2} are obtained from imposing the boundary conditions Bu=g.

    Two examples are considered in the numerical experiments whose data are listed in Table 4.1. Solution profiles of these examples are shown in Figure 4.1 and Figure 4.2. The axes on these figures are flipped to follow the plotting style for profiles associated with Richards Equation (see for example [25,24]). The initial condition is

    Table 4.1.  Data for all examples with closed form solution.
    Ex. 1 Ex. 2
    (a,b) (0,60)cm (0,2)m
    T 1000s 0.5hr
    Bau u(a,t) κ(u)z(uz)|(a,t)
    Bbu u(b,t) u(b)
    ga 65cm 0.15m/hr
    gb 0cm 0m
    u0 (4.6) & (4.7) (4.6) & (4.8)
    α 0.01cm1 4m1
    κs 0.001cm/s 0.1m/hr
    θs 0.3 0.6
    θr 0.08 0.02

     | Show Table
    DownLoad: CSV
    Figure 4.1.  Ex. 1: u(z,t) (top) and ϑ(u(z,t)) (bottom).
    Figure 4.2.  Ex. 2: u(z,t) (top) and ϑ(u(z,t)) (bottom).
    u0(z)=α1ln(κ1sf(z)), (4.6)

    where for Ex. 1,

    f(z)=C1+C2eαz+Aeαz/2sin(λ1z),λ1=π/b,C2=κs(eαgaeαgb)1eαb,C1=κseαgaC2,A=4λ1(κseα(65+b/2)C1eαb/2C2eαb/2)((α/2)2+λ21)b, (4.7)

    and for Ex. 2,

    f(z)=C1+C2eαz+eα(bz)/26000n=1Ansin(λn(bz)),λnisgovernedbytan(λnb)+2λnα=0,C1=ga,C2=κseαgbC1eαb,An=αcosh(αb/2)sin(λnb)2λncos(λnb)sinh(αb/2)(α/2)2+λ2n4λnga2λnbsin(2λn). (4.8)

    In this subsection, a set of numerical experiments to investigate accuracy of the approximation is presented. We solve the two examples whose data are listed in Table 4.1.

    Table 4.2 and Table 4.3 list the errors of approximation ϑ(˜u(T)) in L2(a,b)-norm for Ex. 1 and Ex. 2, respectively. Four different number of elements (M=12,24,48,96) and four different number of time steps (N=1,2,4,8) are used to collect the error data in Table 4.2, while for error data in Table 4.3, M=5,10,20,40 and N=4,8,16,32 are used.

    Table 4.2.  Ex. 1: Error of ϑ(˜u(T)) quantified in L2(a,b)-norm.
    M N FVEM-dG0 FVEM-dG1
    12 1 58.0082e-03 14.5267e-03
    24 1 58.1606e-03 15.1200e-03
    48 1 58.1991e-03 15.2688e-03
    96 1 58.2087e-03 15.3060e-03
    12 2 32.5818e-03 0.6939e-03
    24 2 32.5477e-03 0.8597e-03
    48 2 32.5396e-03 0.9086e-03
    96 2 32.5376e-03 0.9211e-03
    12 4 17.3786e-03 0.2823e-03
    24 4 17.2600e-03 0.0916e-03
    48 4 17.2310e-03 0.1222e-03
    96 4 17.2238e-03 0.1344e-03
    12 8 9.0728e-03 0.3586e-03
    24 8 8.9160e-03 0.0793e-03
    48 8 8.8779e-03 0.0154e-03
    96 8 8.8685e-03 0.0153e-03

     | Show Table
    DownLoad: CSV
    Table 4.3.  Ex. 2: Error of ϑ(˜u(T)) quantified in L2(a,b)-norm.
    M N FVEM-dG0 FVEM-dG1
    5 4 4.9527e-02 5.0631e-02
    10 4 2.0846e-02 1.8016e-02
    20 4 1.0534e-02 0.4128e-02
    40 4 0.8508e-02 0.0992e-02
    5 8 4.9824e-02 5.0641e-02
    10 8 1.8868e-02 1.7991e-02
    20 8 0.6907e-02 0.4088e-03
    40 8 0.4671e-02 0.0965e-02
    5 16 5.0159e-02 5.0645e-02
    10 16 1.8208e-02 1.7988e-02
    20 16 0.5197e-02 0.4079e-02
    40 16 0.2637e-02 0.0959e-02
    5 32 5.0382e-02 5.0647e-02
    10 32 1.8027e-02 1.7988e-02
    20 32 0.4504e-02 0.4077e-02
    40 32 0.1646e-02 0.0958e-02

     | Show Table
    DownLoad: CSV
    Table 4.4.  True Value of Quantities of Interest.
    Ex. 1 Ex. 2
    Q in (4.9) 0.231831624739998887 0.129975678959476710
    Q in (4.10) 0.221196137487056291 0.111225678959476525

     | Show Table
    DownLoad: CSV

    First, the accuracy of FVEM-dG1 generally outperforms that of FVEM-dG0, which is especially evident when solving Ex. 1 (see Table 4.2). The approximation error for Ex. 1 seems to be dominated by component of the temporal discretization. For a fixed N, refining M does not quite improve the accuracy. However, for a fixed M, refining N by two roughly reduces the error by two for FVEM-dG0 and by seven to ten for FVEM-dG1, especially for larger M. A strikingly different result is observed for Ex. 2 (see Table 4.3), for which the spatial discretization error component is more dominant. For a fixed N, the error of FVEM-dG1 shows a quadratic convergence with respect to M. On the other hand, when N is still small, the error of FVEM-dG0 resembles a first order convergence with respect to M. As N is increased, a better convergence rate is obtained.

    As mentioned, the error equation for a quantity of interest Q stated in Theorems 3.1 and 3.2 can be used to derive fully computable error estimators for the approximate solution ˜u. Notice that adjoint equation (3.8) is formulated based on linearization that utilizes Mean Value Theorem on the path ˜uσ=˜u+σ(u˜u), for σ[0,1], cf. (3.2). Since in reality u is not available, calculation of solution to the adjoint equation (3.8) must be done using the only available information, namely, the approximate solution ˜u, so in practice, ˜uσ˜u. The adjoint φ is approximated preferably using higher order approximation than the one used to produce ˜u.

    To test the proposed error estimators, we consider two quantities of interest:

    ● The spatial average of water content at time T, which is represented as

    [Q(u)](T)=1babaϑ(u(z,T))dz. (4.9)

    To calculate the adjoint solution associated with this quantity, the corresponding adjoint data is ψT=1/(ba), and the rest are zero.

    ● The total average of water content over (a,b)×(0,T), which is expressed as

    [Q(u)](T)=1TT01babaϑ(u(z,t))dzdt. (4.10)

    The corresponding adjoint data is ψ=1/(ba)/T and the rest are zero.

    The true values of these quantities of interest for the two examples are listed in Table 4.4.

    The approximate solution is ˜uW0h (FVEM-dG0). The adjoint solution is solved by continuous and piecewise quadratic finite element in spatial variable and continuous piecewise linear in temporal variable. Profiles of φ corresponding to each of these quantities of interest for each of the problems are shown in Figure 4.3 and Figure 4.4, respectively.

    Figure 4.3.  Ex. 1: φ(z,t) for Q in (4.9) (top) and for Q in (4.10) (bottom). Each of them is obtained from numerical approximation of (3.8), with ˜uW0h, h=(ba)/96, and k=T/8.
    Figure 4.4.  Ex. 2: φ(z,t) for Q in (4.9) (left) and for Q in (4.10) (right). Each of them is obtained from numerical approximation of (3.8), with ˜uW0h, h=(ba)/50, and k=T/8.

    Table 4.5 and Table 4.6 demonstrate performance of the error estimator for the above quantities of interest when solving Ex. 1. In these tables, the error has been decomposed according to the components of error listed in Theorem 3.1. As in the accuracy assessment results, four different number of elements (M=12,24,48,96) and four different number of time steps (N=1,2,4,8) are used. The last column, which is labeled by Eff. denotes the ratio of error estimator (Err. Est.) to the actual error, so the closer Eff. is to 1 indicates a more accurate error estimator.

    Table 4.5.  Ex. 1: Performance of the Error Estimator in Theorem 3.1 for Q in (4.9).
    M N E0 E1 E2 E3 Err. Est. Eff.
    12 1 -5.5601e-05 6.6298e-03 2.6571e-04 2.2485e-05 6.8623e-03 1.149
    24 1 -1.3843e-05 6.6974e-03 1.3440e-04 5.3567e-06 6.8233e-03 1.139
    48 1 -3.4568e-06 6.7142e-03 6.7378e-05 1.3220e-06 6.7794e-03 1.131
    96 1 -8.6396e-07 6.7184e-03 3.3706e-05 3.2943e-07 6.7515e-03 1.126
    12 2 -4.4365e-05 3.8268e-03 7.1941e-05 3.6142e-05 3.8906e-03 1.120
    24 2 -1.1046e-05 3.8548e-03 3.9509e-05 8.5556e-06 3.8918e-03 1.120
    48 2 -2.7579e-06 3.8617e-03 2.0874e-05 2.1051e-06 3.8820e-03 1.120
    96 2 -6.8914e-07 3.8635e-03 1.0730e-05 5.2410e-07 3.8740e-03 1.115
    12 4 -3.9779e-05 2.0010e-03 2.5003e-05 4.3906e-05 2.0310e-03 1.067
    24 4 -9.9212e-06 2.0128e-03 1.2165e-06 1.0291e-05 2.0254e-03 1.069
    48 4 -2.4774e-06 2.0158e-03 6.9258e-06 2.5084e-06 2.0227e-03 1.069
    96 4 -6.1911e-07 2.0165e-03 3.7657e-06 6.2266e-07 2.0203e-03 1.068
    12 8 -3.8042e-05 1.0175e-03 1.9493e-05 4.6704e-05 1.0457e-03 1.037
    24 8 -9.4764e-06 1.0227e-03 5.0975e-06 1.1390e-05 1.0297e-03 1.036
    48 8 -2.3676e-06 1.0240e-03 2.3602e-06 2.7559e-06 1.0268e-03 1.036
    96 8 -5.9172e-07 1.0244e-03 1.3885e-06 6.8063e-07 1.0258e-03 1.036

     | Show Table
    DownLoad: CSV
    Table 4.6.  Ex. 1: Performance of the Error Estimator in Theorem 3.1 for Q in (4.10).
    M N E0 E1 E2 E3 Err. Est. Eff.
    12 1 -7.6641e-05 -5.9702e-03 6.5348e-06 2.7273e-05 -6.0131e-03 1.290
    24 1 -1.9082e-05 -5.9818e-03 1.6059e-06 6.7751e-06 -5.9925e-03 1.290
    48 1 -4.7653e-06 -5.9847e-03 3.9971e-07 1.6907e-06 -5.9874e-03 1.290
    96 1 -1.1910e-06 -5.9854e-03 9.9817e-08 4.2248e-07 -5.9861e-03 1.290
    12 2 -7.4772e-05 -3.0291e-03 1.6147e-05 2.5436e-05 -3.0623e-03 1.123
    24 2 -1.8616e-05 -3.0342e-03 3.9483e-06 6.3250e-06 -3.0426e-03 1.123
    48 2 -4.6488e-06 -3.0355e-03 9.8092e-07 1.5784e-06 -3.0376e-03 1.123
    96 2 -1.1619e-06 -3.0358e-03 2.4483e-07 3.9441e-07 -3.0363e-03 1.123
    12 4 -7.4397e-05 -1.5584e-03 2.3449e-05 2.4137e-05 -1.5852e-03 1.057
    24 4 -1.8523e-05 -1.5609e-03 5.7189e-06 6.0235e-06 -1.5677e-03 1.057
    48 4 -4.6257e-06 -1.5615e-03 1.4158e-06 1.5036e-06 -1.5632e-03 1.057
    96 4 -1.1561e-06 -1.5617e-03 3.5295e-07 3.7571e-07 -1.5621e-03 1.057
    12 8 -7.4326e-05 -7.9524e-04 2.7824e-05 2.3365e-05 -8.1837e-04 1.027
    24 8 -1.8504e-05 -7.9647e-04 6.8634e-06 5.8679e-06 -8.0225e-04 1.027
    48 8 -4.6210e-06 -7.9678e-04 1.6962e-06 1.4661e-06 -7.9824e-04 1.027
    96 8 -1.1549e-06 -7.9686e-04 4.2208e-07 3.6638e-07 -7.9722e-04 1.027

     | Show Table
    DownLoad: CSV

    Notice that values in the tables give an indication that the error in Q is dominated by the contribution from temporal discretization, as refinement of the spatial mesh for a fixed time step only causes negligible reduction in Err. Est. Reducing the time step by two seems to reduce the error by two, which demonstrates an asymptotic first order convergence with respect to the time step, i.e., Err.Est.=O(k). Prominence of temporal discretization effect to the total error makes sense due to the longer time simulation.

    The component E0 quantifies the quality of representation of the true initial condition u0 in the simulation. Representation of u0 is realized through the projection of u0 onto Xh. Thus, E0 measures the discrepancy attributed to this choice, which shows a second order convergence with respect to h (i.e., E0=O(h2)) for a fixed time step. Comparison of values in the tables indicates that relative contribution of this component to the overall error is less significant.

    The component E1, which measures the residual of the finite volume element discretization weighted by the adjoint solution and integrated over time, shows a decrease with respect to time step as well. In fact, E1 is clearly the main contributor to the total error with asymptotic behavior E1=O(k). The components E2 and E3 measures the discrepancy between the variational setting associated with finite volume element and that of standard continuous Galerkin finite element. In the realm of a priori error analysis, these two components are bounded in terms of the spatial mesh size h (see Proposition 2.1 and (2.5)). For every fixed N, there is a reduction of E2 and E3 as h is refined, with mostly E2=O(h) in Table 4.5, E2=O(h2) in Table 4.6, and E3=O(h2) in both tables. However, these two components are less dominant in relative comparison to E1. Notice also that E0 has a different sign than the rest of components. A capability to distinguish components of error and to recognize potential cancellation is arguably one of the strong advantages of the adjoint-based error estimation techniques. Finally, as illustrated by the Eff., as refinement is performed, the error estimator gives a more accurate prediction.

    Next we utilize the proposed error estimator to the numerical solution of Ex. 2. Table 4.7 and Table 4.8 show the breakdown of error for each of the quantities of interest. Again, following the accuracy assessment results, four different number of elements (M=5,10,20,40) and four different number of time steps (N=4,8,16,32) are used. As in the results for Ex. 1, the proposed estimator performs really well in predicting the error. However, upon a closer observation, the detailed situation is quite different from what happened in Ex. 1. In Table 4.7 (error associated with Q in (4.9)), we notice that the error caused by discretization of the spatial variable is more dominant, so refining the spatial mesh reduces the error. In particular, E3 is seen to be the main contributor to the total error with asymptotic behavior E3=O(h2), which results in Err.Est.=O(h2). Notice also that time step refinement does not seem to improve the accuracy.

    Table 4.7.  Ex. 2: Performance of the Error Estimator in Theorem 3.1 for Q in (4.9).
    M N E0 E1 E2 E3 Err. Est. Eff.
    5 4 -3.6102e-05 -6.2164e-08 -3.6456e-07 1.0290e-02 1.0253e-02 0.977
    10 4 -1.3302e-05 3.3879e-09 -4.6917e-09 3.9111e-03 3.8978e-03 0.993
    20 4 -5.7876e-06 9.9041e-10 -2.2824e-10 1.0090e-03 1.0032e-03 0.998
    40 4 -2.7053e-06 5.6955e-10 -3.8568e-11 2.4965e-04 2.4695e-04 0.999
    5 8 -3.6107e-05 -4.3218e-08 -3.5113e-07 1.1133e-02 1.1096e-02 0.974
    10 8 -1.3302e-05 1.8737e-10 -4.7768e-10 4.3243e-03 4.3110e-03 0.993
    20 8 -5.7876e-06 1.9063e-11 -6.0320e-12 1.0842e-03 1.0784e-03 0.998
    40 8 -2.7053e-06 5.5338e-12 -4.5001e-13 2.6644e-04 2.6373e-02 0.999
    5 16 -3.6121e-05 6.8028e-08 -4.5382e-07 1.1595e-02 1.1559e-02 0.973
    10 16 -1.3302e-05 1.6801e-11 -8.8583e-11 4.5811e-03 4.5678e-03 0.992
    20 16 -5.7876e-06 5.2323e-13 -3.2271e-13 1.1290e-03 1.1232e-03 0.998
    40 16 -2.7053e-06 4.0884e-14 -2.5154e-14 2.7647e-04 2.7376e-04 0.999
    5 32 -3.6089e-05 -4.4295e-08 4.5726e-08 1.1837e-02 1.1801e-02 0.972
    10 32 -1.3302e-05 2.5330e-12 -2.9891e-11 4.7280e-03 4.7144e-03 0.992
    20 32 -5.7876e-06 2.6583e-14 -9.4480e-14 1.1545e-03 1.1487e-03 0.998
    40 32 -2.7053e-06 -1.0436e-14 -1.8777e-14 2.8211e-04 2.7941e-04 0.999

     | Show Table
    DownLoad: CSV
    Table 4.8.  Ex. 2: Performance of the Error Estimator in Theorem 3.1 for Q in (4.10).
    M N E0 E1 E2 E3 Err. Est. Eff.
    5 4 -3.6102e-05 -4.6874e-03 -7.6435e-08 7.2999e-03 2.5763e-03 0.921
    10 4 -1.3302e-05 -4.6875e-03 -8.4587e-10 3.3265e-03 -1.3743e-03 1.026
    20 4 -5.7876e-06 -4.6875e-03 -3.4359e-11 9.8963e-04 -3.7036e-03 1.001
    40 4 -2.7054e-06 -4.6875e-03 -4.2724e-12 2.5707e-04 -4.4331e-03 1.000
    5 8 -3.6102e-05 -2.3437e-03 -4.6414e-08 7.2458e-03 4.8659e-03 0.950
    10 8 -1.3302e-05 -2.3437e-03 -6.5517e-11 3.5887e-03 1.2316e-03 0.961
    20 8 -5.7876e-06 -2.3437e-03 -7.0707e-13 1.0899e-03 -1.2596e-03 1.004
    40 8 -2.7054e-06 -2.3437e-03 -4.9054e-14 2.8339e-04 -2.0631e-03 1.000
    5 16 -3.6105e-05 -1.1719e-03 -7.9249e-08 7.1892e-03 5.9812e-03 0.956
    10 16 -1.3302e-05 -1.1719e-03 -1.0129e-11 3.7626e-03 2.5774e-03 0.977
    20 16 -5.7876e-06 -1.1719e-03 -6.6506e-14 1.1637e-03 -1.3957e-05 1.823
    40 16 -2.7054e-06 -1.1719e-03 -1.1789e-14 3.0287e-04 -8.7171e-04 1.001
    5 32 -3.6103e-05 -5.8593e-04 -2.1637e-08 7.1541e-03 6.5320e-03 0.959
    10 32 -1.3302e-06 -5.8594e-04 -3.1392e-12 3.8643e-03 3.2650e-03 0.979
    20 32 -5.7876e-06 -5.8594e-04 -4.6973e-14 1.2132e-03 6.2149e-04 0.988
    40 32 -2.7054e-06 -5.8594e-04 -1.1491e-14 3.1628e-04 -2.7236e-04 1.002

     | Show Table
    DownLoad: CSV

    The result in Table 4.8 shows that E1 and E3 are two competing error components with E1=O(k) and E3=O(h2). Due to their different sign, they tend to cancel each other, especially when M and N are still smaller. This in turn lowers the magnitudes of total error. However, as N is increased, E3 tends to be more dominant than E1, especially for small M. Since there is an intertwinement of the error components stemming from temporal and spatial variables, Err. Est. in this table does not indicate a clear pattern of convergence with respect to any discretization parameter. However, as the two discretizations are simultaneously refined, there is an observable reduction.

    Tables 4.9 to 4.12 present the decomposition of error for Ex. 1 and Ex. 2 into various components as dictated by Theorem 3.2. The columns for E0=E0, Err. Est., and Eff. are not included since they are the same as in the corresponding columns in Tables 4.5 to 4.8, respectively. The component E1 seems to be the main contributor to the total error. For results associated with Ex. 1 (see Tables 4.9 and 4.10), the asymptotic behavior is roughly E1=O(k). For Ex. 2 with Q as stated in (4.9) (see Table 4.11), the asymptotic behavior is E1=O(h2). However, it is not quite the case for Q in (4.10) (see Table 4.12).

    Table 4.9.  Ex. 1: Decomposition of Error according to Theorem 3.2 for Q in (4.9).
    M N E1 E2 E3
    12 1 6.9104e-03 7.5953e-06 0
    24 1 6.8352e-03 1.8846e-06 6.9389e-18
    48 1 6.7824e-03 4.6724e-07 3.4694e-18
    96 1 6.7523e-03 1.1620e-07 6.9389e-18
    12 2 3.8392e-03 6.1045e-06 8.9582e-05
    24 2 3.8789e-03 1.5228e-06 2.2428e-05
    48 2 3.8787e-03 3.7986e-07 5.6090e-06
    96 2 3.8732e-03 9.4827e-08 1.4024e-06
    12 4 1.9444e-03 5.6811e-06 1.1982e-04
    24 4 2.0038e-03 1.4190e-06 3.0033e-05
    48 4 2.0173e-03 3.5462e-07 7.5129e-06
    96 4 2.0189e-03 8.8625e-08 1.8785e-06
    12 8 9.4839e-04 5.5176e-06 1.2981e-04
    24 8 1.0052e-03 1.3769e-06 3.2575e-05
    48 8 1.0206e-03 3.4421e-07 8.1499e-06
    96 8 1.0243e-03 8.6050e-08 2.0378e-06

     | Show Table
    DownLoad: CSV
    Table 4.10.  Ex. 1: Decomposition of Error according to Theorem 3.2 for Q in (4.10).
    M N E1 E2 E3
    12 1 -5.9364e-03 0 0
    24 1 -5.9734e-03 0 0
    48 1 -5.9826e-03 0 0
    96 1 -5.9849e-03 0 0
    12 2 -3.0501e-03 2.5710e-06 5.9998e-05
    24 2 -3.0397e-03 6.4271e-07 1.5112e-05
    48 2 -3.0369e-03 1.6068e-07 3.7850e-06
    96 2 -3.0361e-03 4.0169e-08 9.4670e-07
    12 4 -1.6024e-03 4.3468e-06 8.7251e-05
    24 4 -1.5722e-03 1.0865e-06 2.1970e-05
    48 4 -1.5644e-03 2.7161e-07 5.5023e-06
    96 4 -1.5624e-03 6.7902e-08 1.3762e-06
    12 8 -8.5030e-04 5.5588e-06 1.0069e-04
    24 8 -8.1048e-04 1.3899e-06 2.5351e-05
    48 8 -8.0031e-04 3.4750e-07 6.3488e-06
    96 8 -7.9774e-04 8.6875e-08 1.5879e-06

     | Show Table
    DownLoad: CSV
    Table 4.11.  Ex. 2: Decomposition of Error according to Theorem 3.2 for Q in (4.9).
    M N E1 E2 E3
    5 4 1.0289e-02 -7.6351e-08 3.0166e-07
    10 4 3.9111e-03 -5.9250e-10 4.6158e-09
    20 4 1.0090e-03 -1.7676e-11 1.7694e-10
    40 4 2.4965e-04 -1.9490e-12 2.1193e-11
    5 8 1.1134e-02 7.3138e-08 -1.3466e-06
    10 8 4.3243e-03 -6.0938e-11 1.1118e-09
    20 8 1.0842e-03 -4.8331e-13 1.1840e-11
    40 8 2.6644e-04 -2.7270e-14 6.6398e-13
    5 16 1.1601e-02 8.7291e-08 -5.8431e-06
    10 16 4.5811e-03 -1.1412e-11 3.7774e-10
    20 16 1.1290e-03 -3.7491e-14 1.0044e-12
    40 16 2.7647e-04 -4.7254e-15 1.7396e-14
    5 32 1.1833e-02 2.1590e-08 3.9613e-06
    10 32 4.7277e-03 -3.8970e-12 1.8867e-10
    20 32 1.1545e-03 -1.9623e-14 1.5073e-13
    40 32 2.8211e-04 -4.4201e-15 -2.7756e-17

     | Show Table
    DownLoad: CSV
    Table 4.12.  Ex. 2: Decomposition of Error according to Theorem 3.2 for Q in (4.10).
    M N E1 E2 E3
    5 4 2.6123e-03 -7.7482e-09 1.1954e-07
    10 4 -1.3610e-03 -5.0964e-11 1.2934e-09
    20 4 -3.6979e-03 -1.4624e-12 4.4327e-11
    40 4 -4.4304e-03 -1.6102e-13 5.1326e-12
    5 8 4.9021e-03 6.9429e-09 -1.0375e-07
    10 8 1.2450e-03 -5.4195e-12 2.1318e-10
    20 8 -1.2539e-03 -4.3805e-14 1.8653e-12
    40 8 -2.0604e-03 -3.6394e-15 9.6166e-14
    5 16 6.0183e-03 1.5114e-08 -1.0262e-06
    10 16 2.5907e-03 -1.0270e-12 5.4566e-11
    20 16 -8.1695e-06 -9.7700e-15 1.0732e-13
    40 16 -8.6900e-04 -2.1094e-15 1.2386e-15
    5 32 6.5684e-03 3.4467e-09 -3.0913e-07
    10 32 3.2783e-03 -3.6607e-13 2.2585e-11
    20 32 6.2727e-04 -8.8020e-15 1.0911e-14
    40 32 -2.6966e-04 -2.1545e-15 -4.1980e-16

     | Show Table
    DownLoad: CSV

    This paper investigates the application of adjoint-based a posteriori error analysis for numerical approximation of Richards Equation. Construction of the approximate solution is cast into space-time variational formulation, specifically using the finite volume element in spatial variable and discontinuous Galerkin finite element in temporal variable. The resulting error estimators have the capability to predict components of error in certain quantities of interest that are expressed as a functional of the solution. The two examples give a strong indication that the error estimators are robust and capable to predict the error satisfactorily.

    As for future work, we are interested in exploring further applications of the error estimators, in particular as to how they are applied to the setting of multidimensional problems. Owing to the various challenges persistent in the approximations of Richards Equation, a utilization of adaptivity is perhaps the only judicious route. Here the adaptivity is multi-faceted, not only as it pertains to local spatial refinement and dynamic time stepping, but also as it relates to determining optimal number of iterations when solving the nonlinear algebraic system. In this regard, the prospect of adjoint-based approach to estimate the components of error seems to be very promising.

    Another interesting subject, which is not pursued in the present work, is a rigorous mathematical analysis of the proposed approximation. It must begin with establishing the existence of an approximate solution of (2.15). Here a potentially useful tool is either the Banach Fixed Point Theorem or the Brouwer Fixed Point Theorem. It should then be followed by a careful convergence analysis with the ultimate goal of showing the existence of a limit of the sequence of approximate solutions as (h,k)(0,0), and confirming that the limit satisfies a weak formulation of the Richards Equation. This can then be supplied with a study convergence rate of the approximate solution with respect to h and k.



    [1] W. Bangerth and R. Rannacher, Adaptive Finite Element Methods for Differential Equations, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2003. doi: 10.1007/978-3-0348-7605-6
    [2] Adaptive multistep time discretization and linearization based on a posteriori error estimates for the Richards equation. Appl. Numer. Math. (2017) 112: 104-125.
    [3] Computation of variably saturated subsurface flow by adaptive mixed hybrid finite element methods. Advances in Water Resources (2004) 27: 565-581.
    [4] A posteriori analysis of a space and time discretization of a nonlinear model for the flow in partially saturated porous media. IMA J. Numer. Anal. (2014) 34: 1002-1036.
    [5] A general mass-conservative numerical solution for the unsaturated flow equation. Water Resour. Res. (1990) 26: 1483-1496.
    [6] Finite volume methods for elliptic PDE's: A new approach. M2AN Math. Model. Numer. Anal. (2002) 36: 307-324.
    [7] A finite volume element method for a non-linear elliptic problem. Numer. Linear Algebra Appl. (2005) 12: 515-546.
    [8] Z. Chen, G. Huan and Y. Ma, Computational Methods for Multiphase Flows in Porous Media, vol. 2 of Computational Science & Engineering, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2006. doi: 10.1137/1.9780898718942
    [9] Error estimates in L2,H1 and L in covolume methods for elliptic and parabolic problems: A unified approach. Math. Comp. (2000) 69: 103-120.
    [10] A mass-conservative control volume-finite element method for solving Richards'equation in heterogeneous porous media. BIT (2011) 51: 845-864.
    [11] (1996) Computational Differential Equations. Cambridge: Cambridge University Press.
    [12] (1995) Introduction to adaptive methods for differential equations. Cambridge: in Acta Numerica, 1995, Acta Numer., Cambridge Univ. Press.
    [13] (1995) Introduction to computational methods for differential equations. New York: in Theory and Numerics of Ordinary and Partial Differential Equations (Leicester, 1994), Adv. Numer. Anal., IV, Oxford Univ. Press.
    [14] A posteriori error bounds and global error control for approximation of ordinary differential equations. SIAM J. Numer. Anal. (1995) 32: 1-48.
    [15] D. J. Estep, M. G. Larson and R. D. Williams, Estimating the error of numerical solutions of systems of reaction-diffusion equations, Mem. Amer. Math. Soc., 146 (2000), no. 696. doi: 10.1090/memo/0696
    [16] The finite volume method for Richards equation. Comput. Geosci. (1999) 3: 259-294.
    [17] Numerical solution of Richards' equation: A review of advances and challenges. Soil Science Society of America Journal (2017) 81: 1257-1269.
    [18] Some steady-state solutions of the unsaturated moisture flow equation with application to evaporation from a water table. Soil Science (1958) 85: 228-232.
    [19] Adjoint methods for PDEs: A posteriori error analysis and postprocessing by duality. Acta Numer. (2002) 11: 145-236.
    [20] Galerkin-type approximations which are discontinuous in time for parabolic equations in a variable domain. SIAM J. Numer. Anal. (1978) 15: 912-928.
    [21] G. I. Marchuk, Adjoint Equations and Analysis of Complex Systems, vol. 295 of Mathematics and its Applications, Kluwer Academic Publishers Group, Dordrecht, 1995, Translated from the 1992 Russian edition by Guennadi Kontarev and revised by the author. doi: 10.1007/978-94-017-0621-6
    [22] Semianalytical solution to Richards' equation for layered porous media. Journal of Irrigation and Drainage Engineering (1998) 124: 290-299.
    [23] Capillary conduction of liquids through porous mediums. Physics (1931) 1: 318-333.
    [24] Analytical solutions for one-dimensional, transient infiltration toward the water table in homogeneous and layered soils. Water Resour. Res. (1991) 27: 753-762.
    [25] An analytical solution to Richards' equation for time-varying infiltration. Water Resour. Res. (1991) 27: 763-766.
  • 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(2257) PDF downloads(179) Cited by(0)

Figures and Tables

Figures(4)  /  Tables(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog