Iterative schemes for coupled flow and transport
in porous media - Convergence and truncation errors

Nicolae Suciu1 and Florin A. Radu2 and Emil Cătinaş1
(Date: May 17, 2024; accepted: June 26, 2024; published online: July 3, 2024.)
Abstract.

Nonlinearities of coupled flow and transport problems for partially saturated porous media are solved with explicit iterative L-schemes. Their behavior is analyzed with the aid of the computational orders of convergence. This approach allows highlighting the influence of the truncation errors in the numerical schemes on the convergence of the iterations. Further, by using manufactured exact solutions, error-based orders of convergence of the iterative schemes are assessed and the convergence of the numerical solutions is demonstrated numerically through grid-convergence tests.

Key words and phrases:
Richards’ equation, coupled flow and transport, finite differences, global random walk, iterative schemes, convergence order
2005 Mathematics Subject Classification:
35G31, 76S05, 76R50, 65M06, 65C35.
1Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy, Fântânele str. no. 57, Cluj-Napoca, Romania, e-mail: nsuciu@ictp.acad.ro, ecatinas@ictp.acad.ro
2Center for Modeling of Coupled Subsurface Dynamics, University of Bergen, Allégaten 41, Bergen, Norway, e-mail: florin.radu@uib.no

1. Introduction

Flow and reactive transport in porous media are modeled by systems of coupled nonlinear partial differential equations. The seepage velocity through the bulk of porous medium is proportional to the hydraulic conductivity and to the gradient of the pressure head, according to Darcy’s law, which combined with the conservation equation results in the Richards’ equation. In its general, mixed formulation, Richards’ equation contains two unknown functions, the volumetric water content and the pressure head. Therefore, it needs to be closed by constitutive relationships for the dependence of the water content on the pressure head and the dependence of the hydraulic conductivity on the water content. Such relationships are highly nonlinear, e.g., the exponential model or the van Genuchten-Mualem model, with discontinuous derivatives when the negative pressure, in unsaturated flow regime, changes into the positive pressure characterizing the saturated regime with constant water content and hydraulic conductivity. Consequently, the resulting Richards’ equation for the pressure head is non-linear and parabolic-elliptic degenerate.

The concentration of the chemical compounds transported in soils and aquifers is governed by parabolic advection-diffusion-reaction equations, with flow velocity provided by the solution of the Richards’ equation. In the presence of capillarity effects, for instance in problems of surfactant transport, the water content depends on pressure as well as on concentration and the flow and transport equations are coupled in a nonlinear way in both directions. Solving these coupled nonlinear problems requires iterative schemes that use different linearization techniques, such as Newton’s method, (modified) Picard method, L-scheme or their combinations. While Newton’s method is second order (under usual circumstances) but only locally convergent, which means that it converges only if the initial choice is not too far from the solution, the L-scheme is a robust method unconditionally first order convergent, the robustness of the Picard method being somewhere in between [8]. A significant improvement of the convergence can be achieved by starting with L-scheme iterations and, once an appropriate initial solution is found, switching to the faster and higher order convergent Newton’s method [13].

The L-scheme can be thought as a quasi-Newton method, with derivative of the water content with respect of the pressure head replaced by a positive constant L. This makes it robust and independent of the initial choice, at the price of being only first order convergent. Given its simplicity, the L-scheme is particularly suitable in solving intricate and fully coupled nonlinear problems. Most of the implementations of the L-scheme are implicit linearizations based on finite element [9, 8, 13] or finite volume [10, 5] discretizations. Explicit L-schemes based on finite difference (FD) and random walk methods, introduced in [17], have been successfully used to solve coupled problems of flow and biodegradation in partially saturated soils and saturated heterogeneous aquifers [18]. The reason of using explicit L-schemes is that numerical solutions of the flow equations are obtained on the same regular grid on which reactive transport is solved with random walk algorithms free of numerical diffusion [17].

Theoretical proofs of convergence of the linearization schemes essentially show that the errors of the iterative solutions with respect to the exact solution of the nonlinear problem can be made arbitrarily small [8]. However, even if the numerical analysis of the scheme provides estimates of the convergence rates, in general it does not produce a thorough and comprehensive characterization of the type of convergence behavior, by predicting the upper/lower limits of the rates as in the theory of the convergent sequences [3, 4]. The analysis can be completed either by assessing computational orders of convergence, with sequences of successive corrections, or by assessing error-based orders, when exact solutions of the problem are available [20]. This approach could be mainly useful when the constraints required by the theoretical proofs are hard to meet, or when there are no convergence proofs at all, but, in practice, the convergence of the scheme is demonstrated numerically.

We investigate in this framework the convergence of the explicit L-scheme through computational and error-based orders of convergence on three numerical examples: solutions of Richards’ equation, fully coupled flow and surfactant transport, and coupled flow and transport with van Genuchten-Mualem parametrization.

2. Coupled flow and transport in porous media

We consider in the following the fully coupled one-dimensional problem of flow governed by Richards’ equation and transport of a single chemical species governed by an advection-diffusion-reaction equation,

(1) tθ(ψ,c)z[K(θ(ψ,c))z(ψ+z)]=0,
(2) t[θ(ψ,c)c]z[Dzcqc]=R(c),

where ψ(z,t) is the pressure head, with z being the height against the gravitational direction, θ is the water content, q=K(θ(ψ,c))z(ψ+z) is the water flux (also called Darcy velocity), K stands for the hydraulic conductivity of the porous medium, c is the concentration, D is the diffusion/dispersion coefficient, and R is the reaction term. The system (1)-(2) is closed via constitutive relationships defining the dependencies θ(ψ,c) and K(θ(ψ,c)).

Equations (1)-(2) are fully coupled through the terms θ(ψ,c) and [θ(ψ,c)c]. Richards’ equation (1) is parabolic-elliptic degenerate, with variable θ(ψ,c) if ψ<0 and θ=const if ψ0. Since the system is strongly nonlinear, through the functional dependencies θ(ψ,c) and K(θ(ψ,c)), linearization methods have to be used to compute numerical solutions.

2.1. Explicit L-schemes for one-dimensional Richards’ equation

Following [17], we start with the staggered FD scheme with backward discretization in time of the Richards’ equation (1) decoupled111In the fully coupled problem θ(ψi,k) is replaced by θ(ψi,k,ci,k), where ci,k is the solution of (2) and the coupled scheme consists of alternating flow and transport steps. from the transport equation (2),

θ(ψi,k)θ(ψi,k1)=
=ΔtΔz2{[K(θ(ψi+1/2,k))(ψi+1,kψi,k)K(θ(ψi1/2,k))(ψi,kψi1,k)]
(3) +(K(θ(ψi+1/2,k))K(θ(ψi1/2,k)))Δz}.

To obtain an iterative method for solving Richard’s equation, we denote by ψi,ks the approximate solution after s iterations and add to the l.h.s. a stabilization term L(ψi,ks+1ψi,ks), where L is a constant which has the dimension of an inverse length. Whenever the iterative method converges, this stabilization term vanishes and the limit is the solution of the nonlinear scheme (2.1). Defining the dimensionless numbers ri±1/2,ks=K(ψi±1/2,ks)Δt/(LΔz2), one obtains

ψi,ks+1=[1(ri+1/2,ks+ri1/2,ks)]ψi,ks+ri+1/2,ksψi+1,ks+ri1/2,ksψi1,ks
(4) +(ri+1/2,ksri1/2,ks)Δz(θ(ψi,ks)θ(ψi,k1))/L.

In practice, we run successive iterations s=1,2, until a threshold condition is fulfilled, e.g., 𝝍ks𝝍ks1εa+εr𝝍ks, where 𝝍ks denotes the solution vector of components ψi,ks, i=1,,I, is the l2 norm, and εa0 and εr0 are the absolute and relative tolerances (see e.g., [8]). The relation (2.1) is an explicit L-scheme that allows the computation of the approximate solution recursively and avoids the need to solve systems of linear equations, as in implicit finite element [8] or finite volume [10] L-schemes. Consequently, the explicit scheme (2.1) vas found to be one order of magnitude faster than the implicit finite volume L-scheme when solving typical problems for partially saturated flows [17].

To transform the FD scheme (2.1) into a random walk scheme we represent the solution ψi,,ks by 𝒩 particles distributed over the lattice sites, ψi,ksni,ksa/𝒩, where a is a unit length that will be disregarded in the following. This results in the following relation summing contributions from neighboring sites to the number of particles at the site i and time step k,

(5) ni,ks+1=[1(ri+1/2,ks+ri1/2,ks)]ni,ks+ri+1/2,ksni+1,ks+ri1/2,ksni1,ks+𝒩fs,

where fs=(ri+1/2,ksri1/2,ks)Δz[θ(ni,ks)θ(ni,k1)]/L and is the floor function.

In order for (5) to be a random walk scheme, the coefficients multiplying numbers of particles at lattice sites [1(ri+1/2,ks+ri1/2,ks)], ri+1/2,k, and ri1/2,k should be normalized probabilities. This implies the following restriction [1(ri+1/2,ks+ri1/2,ks)]1. A sufficient condition for that is

(6) ri±1/2,ks1/2.

The first three terms in r.h.s. of (5) represent groups of particles jumping on the site i form the right and from the left, and the ratio of the initial number of particles at the site i which do not undergo jumps,

(7) δnii±1,k=ri±1/2,ksni±1,ks,δnii,k=[1(ri+1/2,ks+ri1/2,ks)]ni,ks.

With (7), (5) becomes

(8) ni,ks+1=δnii,k+δnii+1,k+δnii1,k+𝒩fs.

The numbers δn of random walkers are binomially distributed random variables. Unlike in classical random walk approaches, consisting of generating random walk trajectories and counting particles at lattice sites, we directly evaluate the binomial variables δn and construct a global random walk (GRW) [22]. Since there are different probabilities rj±1/2,k for the right/left jumps, one obtains a biased global random walk (BGRW) [15]. The resulting BGRW algorithm moves the particles from a given lattice site j according to the rule

(9) nj,ks=δnjj,ks+δnj+1j,ks+δnj1j,ks,

where

(10) δnj±1j,k=rj±1/2,ksnj,ks,δnii,k=[1(rj+1/2,ks+rj1/2,ks)]ni,ks.

The random variable δnj+1j,k=rj+1/2,ksnj,ks representing the number of right jumps in (9) is binomially distributed, with parameters nj,ks, p=rj+1/2,ks/ri,ks, and q=1p=rj1/2,ks/ri,ks, where ri,ks=rj+1/2,ks+rj1/2,ks. The number of left jumps is the difference between the total number of jumps and the number of right jumps δnj1j,k=ri,ksnj,ksδnj+1j,k. Since the computation of the exact binomial distribution could be computationally prohibitive for very large nj,ks, one uses its approximation via a “reduced-fluctuations algorithm” similar to that proposed for the unbiased GRW algorithm in [22] as follows:

algorithm 1.

Reduced fluctuations algorithm
(i) initialize a real variable rest=0;
(ii) compute δn with the relations (10);
(iii) split δn into δn=δn+{δ} at every lattice site;
(iv) approximate δn by δn;
(v) sum up the remainders into the variable rest={δ} over all the sites;
(vi) allocate one particle to the first lattice site where rest1;
(vii) save rest=rest1 and repeat the sequence (ii)-(vi) at every k.

Specifically, the algorithm is implemented in the Matlab code as follows:

Code 1.

Flow-BGRW with reduced fluctuations

restr=0; rest1=0; rest2=0;
D=K(tht); D=D(1:I-1); % conductivity function of water content
r=dt*D/dx^2/L;
rloc=[1-2*r(1),1-(r(1:I-2)+r(2:I-1)),1-2*r(I-1)];
restr=rloc.*n+restr; nn=floor(restr); restr=restr-nn;
rest1=r(2:I-1).*n(3:I)+rest1;
njumpleft=floor(rest1); rest1=rest1-njumpleft; % left jumps
nn(2:I-1)=nn(2:I-1)+njumpleft;
rest2=r(1:I-2).*n(1:I-2)+rest2;
njumpright=floor(rest2); rest2=rest2-njumpright; % right jumps
nn(2:I-1)=nn(2:I-1)+njumpright;

1 computes the solution in the interior of the domain. The Darcy velocity on the boundaries can be computed in several ways: using an approximate forward finite difference discretization of Darcy’s law, extending on the boundary the velocity from the first neighboring interior site, computing the exact velocity analytically when a manufactured solution for the pressure head is available [17, Sect. 5.2.1].

Remark 1.

Since the BGRW is an explicit scheme, the source term can be evaluated first, then the first tree terms of (8) which model the transport will be computed according to 1. The reduced-fluctuations algorithm preserves the indivisibility of the particles by summing up to unity the fractional parts {} of all the terms defined by (7). Therefore, the floor function in (8) is no longer necessary and the source term becomes 𝒩fs. Dividing (8) by 𝒩, using (7), and taking the average, one retrieves the FD L-scheme (2.1) for the unknown ψi,ks=ni,ks¯/𝒩. A condition of kind KΔt/Δz21/2 ensures the stability of the forward-time central-space scheme for the heat equation with constant coefficient K and is assumed (without carrying out the stability analysis) to hold in case of variable coefficients as well [14]. In the present context, the condition (6) guaranties that particles cannot be created or destroyed otherwise than through boundary conditions or in the presence of an internal source/sink, that is, it ensures the stability of the BGRW L-scheme (5). Since the FD L-scheme (2.1) is the ensemble average of the BGRW scheme, its stability is also ensured by the constraint (6).

The FD scheme is also retrieved when all floor operators are removed in 1. The resulting code sequence can be recast as the FD scheme presented in 2 below, where p=n/𝒩 and pp=nn/𝒩.

Code 2.

Flow-FD scheme

D=K(tht); D=D(1:I-1); % conductivity function of water content
r=dt*D/dx^2/L;
rloc=[1-2*r(1),1-(r(1:I-2)+r(2:I-1)),1-2*r(I-1)];
pp=rloc.*p;
pp(2:I-1)=pp(2:I-1)+r(2:I-1).*p(3:I)+r(1:I-2).*p(1:I-2);
Remark 2.

In the particular case of the saturated regime, θ=const, with space-variable hydraulic conductivity K and a given source term f, after setting L=1 and disregarding the time index k in (2.1) one obtains a transient scheme to solve the equation for the hydraulic head h=ψ+z,

1ahsz[Khz]=f.

This scheme has been used to solve coupled flow and transport in saturated aquifers with moderate heterogeneity [16]. However, because the transient scheme may be very slow for highly heterogeneous coefficients K [2], a more efficient strategy to solve coupled problems for saturated porous media is the coupling between an implicit FD flow scheme [2] and a GRW transport solver, see https://github.com/PMFlow/Coupled_FDM_GRW.

Matlab codes for one- and two-dimensional solutions of Richards’ equation, verification tests, and benchmark problems presented in [17] are posted at https://github.com/PMFlow/RichardsEquation.

2.2. Explicit BGRW scheme for reactive transport

Similarly to the derivation of the flow scheme in Section 2.1, we start with a FD scheme with backward time discretization of the equation (2) with a constant diffusion coefficient D,

θ(ψi,k,ci,k)ci,kθ(ψi,k1,ci,k1)ci,k1=
=tfracΔt2Δz(qi+1,kci+1,kqi1,kci1,k)+DΔtΔz2(ci+1,k2ci,k+ci1,k)+R(ci,k)=
(11) 2DΔtΔz2ci,k+(DΔtΔz2Δt2Δzqi+1,k)ci+1,k+(DΔtΔz2+Δt2Δzqi1,k)ci1,k+ΔtR(ci,k),

where ψi,k is the solution of the coupled flow solver. Further, we denote by ci,ks the approximate solution after s iterations, add to the l.h.s. of (2.2) a stabilization factor L(ci,ks+1ci,ks), where L is now a dimensionless constant, and define the dimensionless parameters

(12) r=2DΔtLΔz2,vi±1,ks=ΔtLΔzqi±1,ks.

Representing the concentration by the distribution of 𝒩 particles on the lattice, ci,ksni,ks/𝒩, and using the parameters (12), the stabilized version of the scheme (2.2) becomes an explicit L-scheme,

(13) ni,ks+1=(1r)ni,ks+12(rvi+1,ks)ni+1,ks+12(r+vi1,ks)ni1,ks+gs,

where gs=𝒩{ΔtR(ni,ks)/L[θ(ψi,ks,ni,ks)ni,ksθ(ψi,k1,ni,k1)ni,k1]/L}.

With the constrains

(14) r1,|vi,ks|r,

which ensure the normalization of the jump probabilities, the explicit L-scheme (13) for reactive transport is defined as a BGRW algorithm. The gathering and the spreading of groups of particles at/from a lattice point are described by relations similar to (8) and (9). The binomial random variables δn are defined by

(15) δnj±1j,ks=12(r±vj,ks)nj,ks,δnjj,ks=(1r)nj,ks.
Remark 3.

The biased jump probabilities 12(r±vj,ks) in (15) simulate the advective displacement by moving larger numbers of particles in the positive flow direction. With an unbiased GRW, the advection is simulated by moving groups of particles over vj,ks lattice sites, i.e., δnj+vj,ks=(1r)nj,ks, where v=vj,ks [17, 18]. This procedure is faster but yields overshooting errors when there are important changes in velocity over a time step. The BGRW algorithm which allows only first-neighbor jumps is accurate but can be much slower than the unbiased GRW when a fine discretization is required to describe the spatial variability of the solution. Moreover, the second constraint (14) and the definitions of the parameters (12) imply an upper bound for the local Péclet number, =qi,ksΔz/D2.

The binomial random variables δn defined by (15) are generated with the Algorithm 1. The Matlab implementation of the advection-diffusion steps of the transport solver is presented in 3 below.

Code 3.

Transport-BGRW with reduced fluctuations

u=q*dt/Lc/dx; % computation of the BGRW parameters (12)
ru=2*D*dt/Lc/dx^2*ones(1,I);
restr=0; restjump=0;
for x=1:I
if n(x) > 0
    r=ru(x);
    restr=n(x)*(1-r)+restr; nsta=floor(restr);
    restr=restr-nsta; njump=n(x)-nsta;
    nn(x)=nn(x)+nsta;
    if(njump)>0
        restjump=njump*0.5*(1-u(x)/r)+restjump;
        nj(1)=floor(restjump); restjump=njump-nj(1);
        nj(2)=floor(restjump); restjump=restjump-nj(2);
        if x==1
            nn(2)=nn(2)+nj(2);
        elseif x==I
            nn(I-1)=nn(I-1)+nj(1);
        else
            for i=1:2
                xd=x+(2*i-3);
                nn(xd)=nn(xd)+nj(i);
            end
        end
    end
end
end
Remark 4.

The ensemble average of the BGRW L-scheme for reactive transport also leads to an equivalent FD scheme. However, we do not pursue this approach which gives up the particle indivisibility. The numerical schemes developed in [17, 18] aim at providing a “microscopic” kinematic description of the reactive transport with piecewise-analytic time functions associated to each particle. Using a space-time averaging over this microscopic description we are then able to construct a macroscopic description through almost everywhere continuous fields that can be used to model experimental measurements or to achieve a spatio-temporal upscaling [19].

Matlab codes for coupled flow and transport and test problems are posted in the repository https://github.com/PMFlow/RichardsEquation. BGRW/GRW solvers for decoupled reactive transport with applications to biodegradation processes in soils and aquifers, presented in [18], may be found in the repository https://github.com/PMFlow/MonodReactions.

2.3. Discussion on convergence

We look for convergence conditions by investigating the evolution of the difference eψs=ψsψ of the solutions of (2.1) and (2.1) and of the difference ecs=csc of the solutions of (13) and (2.2) at fixed position i and time k.

Let Lψ and Lc be the stabilization constants flow the flow and transport schemes, respectively. After denoting by ΔtΔz2𝒟ψ(ψ) the r.h.s. of (2.1) and subtracting (2.1) from (2.1), we have

(16) Lψ(eψs+1eψs)+θ(ψs,cs)θ(ψ,c)=ΔtΔz2(𝒟ψ(ψs)𝒟ψ(ψ)).

Similarly, with Δt2Δz𝒟q(c) and ΔtΔz2𝒟D(c) denoting the advection and diffusion operators in the r.h.s. of (2.2), converting the numbers of particles to concentrations, ci,ks=ni,ks/𝒩, and subtracting (2.2) from (13) we have

(17) Lc(ecs+1ecs)+θ(ψs,cs)csθ(ψ,c)c=Δt2Δz(𝒟q(cs)𝒟q(c))+ΔtΔz2(𝒟D(cs)𝒟D(c)).

Assuming

(A1)θ/ψ>0,θ/c>0,(θc)/ψ>0,(θc)/c>0,
(A2)|𝒟ψ(ψs)𝒟ψ(ψ)|L𝒟ψ|eψs|,𝒟D(cs)𝒟D(c)|L𝒟D|ecs|,

(16) and (17) give

(18) (eψs+1)2(eψs)2(11Lψθ(ψ,c)ψ+L𝒟ψΔtLψΔz2)2+(ecs)2(1Lψθ(ψ,c)c)2,
(19) (ecs+1)2(ecs)2(11Lc(θ(ψ,c)c)c+L𝒟qΔt2LcΔz+L𝒟DΔtLcΔz2)2+(eψs)2(1Lc(θ(ψ,c)c)ψ)2.

Let F denote one of the factors of (eψs)2 and (ecs)2 in (18) and (19). Assuming for all factors the condition F1ϵ2, where ϵ(0,1], one obtains

infψ{θ(ψ,c)/ψ} L𝒟ψΔtΔz2+11ϵ2,
supψ{θ(ψ,c)/ψ} L𝒟ψΔtΔz2+1+1ϵ2,
(20) 1Lψθ(ψ,c)c 1ϵ2.
infc{(θ(ψ,c)c)/c} L𝒟qΔt2Δz+L𝒟DΔtΔz2+11ϵ2,
supc{(θ(ψ,c)c)/c} L𝒟qΔt2Δz+L𝒟DΔtΔz2+1+1ϵ2,
(21) 1Lc(θ(ψ,c)c)ψ 1ϵ2.

Adding the inequalities (18) and (19) and using (2.3) and (2.3) one gets

(22) (eψs+1)2+(ecs+1)2(1ϵ)[(eψs)2+(ecs)2].

With 𝒆s2=i=1I[(ψi,ksψi,k)2+(ci,ksci,k)2] being the squared error norm of the solution vector 𝒙s=(𝝍ks,𝒄ks)2I, the inequality (22) implies

(23) 𝒆s+12(1ϵ)𝒆s2,

and thus a contractive property of the coupled L-schemes. Further, (23) can be used to show that {𝒙s} is a convergent Cauchy sequence (see e.g. [7, Sect. 8.1]) and its limit is unique, because 2I is a complete metric space.

For given Δz, the time step Δt is constrained by the stability conditions of the flow and transport schemes. With the positive constant α defined as α=max(2ri±1/2,ks), the stability condition (6) for the flow scheme becomes α1. Further, from α=2max(K(𝝍ks)Δtψ/(LψΔz2), we determine the corresponding time step Δtψ by

(24) Δtψ=αLψΔz22max(K(𝝍ks)).

Similarly, with β=max(r), the stability condition (14) becomes β1, β=2DΔtc/LcΔz2, and the time step Δtc is given by

(25) Δtc=βLcΔz22D.

The schemes (2.1) and (13) are coupled by alternating flow and transport steps at every time iteration [5, 17], which require a unique time step, defined as

(26) Δt=min(Δtψ,Δtc).

Using (24)-(26) we introduce the factor γ,

(27) γ=ΔtΔz2=min(αLψ2max(K(𝝍ks)),βLc2D).

Inserting (27) and γΔz in (2.3) and (2.3), the convergence conditions explicitly depend on the parameters Lψ and Lc and become more robust, with only one term (coming from the advection term of (2)) depending on discretization through Δz. By relating the parameters Lψ and Lc to the physical parameters of the problem, the conditions (2.3) and (2.3) also provide a general frame for an adaptive L-scheme.

We note that, even though the conditions (2.3) and (2.3) show that the coupled L-schemes should converge, they are of limited use in practice: actually they only can be verified if the solution of the problem is known. Therefore, in the following we investigate the convergence by using sequences of iterates and the definitions of the computational and error-based orders of convergence.

3. Numerical examples

We illustrate the explicit iterative methods presented in Section 2 by three numerical examples. First, we solve a problem with transition between unsaturated and saturated flow regimes for the degenerate Richards’ equation with both the FD and the BGRW scheme presented in Section 2.1. Then, we consider a fully coupled flow and surfactant transport problem [6] with an academic example of constitutive laws θ(ψ) and Kθ(ψ), and a one way coupled flow and transport problem with the realistic van Genuchten-Mualem parametrization of θ and K. The last two problems are solved by the coupled FD-flow and BGRW-transport schemes presented in Section 2.1 and Section 2.2.

The convergence of the iterative schemes is investigated numerically with tools from the theory of abstract convergent sequences [3]. The convergence of an arbitrary sequence of real numbers xsx is characterized by the behavior of the successive errors es=|xxs|. The sequence {xs} converges with the (classical) C-order p1 if

(28) (C)limses+1(es)p=Qp{(0,),p>1(0,1),p=1,

which implies the (weaker) order222This is usually denoted by QL (from logarithm) to distinguish from other types of Q orders. Since only definition (29) will be used in the following, we disregard the subscript L. Q (see [4]),

(29) (Q)limslnes+1lnes=p.

While Eq. (29) is a common way to identify the Q-order of convergence, Eq. (28) can subsequently be used to check whether a classical convergence order exists.

If the limit x is not known, the errors es are replaced by the corrections |xs+1xs| and (28)-(29) define “computational orders of convergence”, denoted by a prime, C and Q. If p>1, computational and error-based orders of convergence are equivalent, denoting this by using curly braces, and they are related by

{C,C}{Q,Q}.

However, when p=1, the above equivalences do not hold in general [4] and the existence of the error-based orders has to be checked on a case by case basis.

Results from the theory of convergent sequences in n [3] cannot be directly extrapolated to flow and transport iterative schemes for porous media, where the convergence is investigated in function spaces with norms that often depend on the particular problem formulation (e.g., [8, 10, 13]). Therefore we rely on the sequences of positive real numbers provided by iterative methods, e.g., 𝝍s𝝍s1 and 𝒄s𝒄s1. If the method converges, which in a restricted sense means that the successive corrections become arbitrarily small (e.g., [10]), the corrections form sequences {xs} converging to zero that can be analyzed with the definitions of the error-based orders of convergence (28) and (29) [20]. Further, based on numerical evaluations of the Q or C orders of convergence of the sequences {xs} we may assume the Q or C convergence of the solution (ψs,cs) of the coupled problem.

Remark 5.

Let s be the iteration count for which (28) written for the sequence of successive corrections 𝛙s+1𝛙s defines a constant quotient Q1(s)=Q within a given absolute tolerance εa. Then, from (28), we have

(30) 𝝍s+1𝝍s Q𝝍s𝝍s1Q2𝝍s1𝝍s2
Qss𝝍s+1𝝍s.

Using (30) we also have, for any s>s, m,

𝝍s+m𝝍s=
=𝝍s+1𝝍s+𝝍s+2𝝍s+1++𝝍s+m𝝍s+m1
(Qss+Qss+1++Qss+m1)𝝍s+1𝝍s
=Qss(1+Q++Qm1)𝝍s+1𝝍s
Qss(m=0Qm)𝝍s+1𝝍s
(31) =Qsslimm1Qm1Q𝝍s+1𝝍s=Qss1Q𝝍s+1𝝍s.

According to (5), lims𝛙s+m𝛙s=0, hence {ψs} is a convergent Cauchy sequence and has a unique limit in the complete space I. The convergence of the iterates of the L-scheme implies the cancelation of the stabilization terms, e.g., L(𝛙s+1𝛙s), so that the limit of the sequence {ψs} coincides with the solution of the nonlinear scheme (compare (2.1) and (2.1)). Thus, the C-linear convergence of the L-scheme is equivalent to the convergence of the successive approximations {ψs} to the unique solution of the nonlinear scheme.

Remark 6.

From a theoretical perspective the C-linear convergence of the L-scheme assumes the convergence of the L-scheme to the exact solution. However, the numerical solution is only an approximation of the exact solution depending on the fineness of the discretization. It follows that a limit quotient Q1<1 in (28), which defines the C-linear convergence, cannot be practically reached in numerical simulations which, necessarily use finite discretizations. For instance, the errors with respect to the exact solution of the second order accurate schemes used in the examples below are of the order 𝒪(Δz2). As the errors es in (28) reach this threshold, they cannot be decreased by further iterations. One obtains thus estimations of the quotient Q11, corresponding to the sublinear convergence of the iterations [3]. The approach to the exact solution is instead quantified by the decay of the errors with respect to the discretization. The common approach [12, 2, 17, 20] is to assess estimated orders of convergence (EOC) by the slope of the decay of the errors in the logarithmic scale after successively halving the discretization parameter Δz,

(32) EOC=log(elel+1)/log(2),l=1,2,

3.1. Example 1: Richards’ equation

For the beginning, let us consider the one-dimensional Richards’ equation (1) closed by one of the simplest models, with exponential dependence of the water content on the pressure head and linear dependence of the hydraulic conductivity on the water content [17],

(33) Θ(ψ)={eαψ,ψ<01,ψ0,
(34) K(θ(ψ))={KsatΘ(ψ),ψ<0Ksat,ψ0,

where Θ=(θθres)/(θsatθres) is the normalized water content, θ=θsat and K=Ksat denote the constant water content respectively the constant hydraulic conductivity in the saturated regions (ψ0), and θres is the residual water content.

We solve a typical problem modeling the experimental study of the flow in variably saturated soil columns, formulated in the domain [0,2]. The boundary conditions are specified by a constant pressure ψ(0,t)=ψ0 at the bottom of the soil column and a constant water flux q0 on the top. Together, these constant conditions determine the initial pressure distribution ψ(z,0) as solution of the steady-state flow problem. For t>0, the pressure ψ0 is kept constant at the bottom of the column and on the top the water flux is increased linearly from q0 to q1 until tt1 and is kept constant for t>t1. The texture of an homogeneous soil, representative for a column filled with sand, is parameterized with Ksat=2.77106, θres=0.06, θsat=0.36, α=10. The water fluxes on the top boundary are given by q0=2.77107 and q1=2.50106. To capture the transition from unsaturated to saturated regime, the pressure at the bottom boundary is fixed at ψ0=0.5. For the parameters specified above, we consider meters as length units and seconds as time units. The simulations are conducted up to T=104 (about 2.78 hours) and the intermediate time is taken as t1=T/102.

We use a regular lattice with Δz=102. The BGRW computations are initialized by multiplying the initial condition by the total number of particles 𝒩. Since the hydraulic conductivity varies in time according to (34), the time step is determined according to (24) for the maximum of K at every time iteration and by specifying a maximum α=0.8 of the parameters ri±1/2,k. The parameter of the regularization term in the L-scheme is set to L=1 for the computation of the initial condition (solution of the stationary problem, i.e. for θ/t=0 in (1)) and to L=2 for the solution of the non-stationary problem. In both cases, the convergence criterion is verified by choosing εr=0 and a fixed absolute tolerance εa=1018.

To illustrate the consistency of the representation of the solution by ensembles of random walkers and the equivalence of the BGRW and FD schemes, we compare their solutions for increasing 𝒩 between 103 and 1024 at the final time T (Fig. 2). The errors |ψBGRW(z)ψFD(z)| of the BGRW solution with respect to the FD solution shown in Fig. 2 decrease as 𝒩1 and approach the machine precision 1016 for 𝒩1018. The l2 norm 𝝍BGRW𝝍FD of the difference of the two solution vectors with I=2/Δz+1=101 components decreases at a rate of 10𝒩1 and reaches a plateau at about 1014 (see Table 1). Since the FD solution is the ensemble average of the BGRW solution (see Remark 1), this behavior also illustrates the self-averaging property of the GRW algorithms [22].

𝒩 1e3 1e6 1e10 1e18 1e24
|ψBGRW(z)ψFD(z)|¯ 1.70e-03 4.24e-06 2.52e-10 7.13e-16 9.21e-16
𝝍BGRW𝝍FD 2.27e-02 6.52e-05 3.32e-09 1.13e-14 1.40e-14
Table 1. Mean errors and error norms for Section 3.1, Fig. 2.

Let {ψs} be the sequence of iterates of the BGRW L-scheme and 𝝍BGRW=lims𝝍s. By the triangle inequality 𝝍s𝝍s1𝝍s𝝍FD+𝝍FD𝝍s1 we have lims𝝍s𝝍s12𝝍BGRW𝝍FD. Hence, the order 20𝒩1 of convergence of the BGRW solution to the FD solution is an upper limit for corrections. Fig. 4 and Fig. 4 show the decrease of the correction norms for 𝒩=1010 and 𝒩=1018, respectively. In the first case the corrections reach a plateau of about 109, consistent with the upper bound 20𝒩1. In the second case, the corrections show a plateau at the machine precision, which, as indicated by Fig. 2, is the same as the plateau for the FD scheme.

Remark 7.

The convergence behavior illustrated above can be explained with the reduced fluctuations Algorithm 1. The sum of fractional parts rest=i=1I{δi} can be used to evaluate the truncation errors of the algorithm. At a given lattice site i, {δi}[0,1) with a mean {δi}¯0.5. The sum over the lattice sites is thus rest0.5I. The total numbers of particles on the lattice is ntot=i=1Ini,ks=i=1I|ψi,ks|𝒩=|ψ|¯𝒩I. With these, the truncation error can be evaluated as rest/ntot0.5/|ψ|¯𝒩1, which is independent of the discretization333This can be easily verified by doubling the number of lattice sites and taking the appropriate parameter L=100 in the Matlab code. (i.e., the number of lattice sites I). For verification purposes, we remove all the floor functions in the 1, add the truncation sequence

rest=nn+rest; nn=floor(rest); rest=rest-nn;

at the end, and perform the computation with 𝒩=1010. The relative error of the solution 𝛙T/Δt obtained in this way with respect to the solution of the original 1, evaluated with the l2 norm, is of about 4.25104. At the middle of the time interval [0,T] we evaluate |ψ|¯=6.45102 and obtain the truncation error 0.5/|ψ|¯𝒩1=7.751010, which corresponds to the average value of the plateau shown by the corrections 𝛙s𝛙s1 in this case. Since the original 1 contains three truncation sequences, we take rest/I1.5 and obtain the truncation error 1.5/|ψ|¯𝒩1=2.30109. This error is almost the same as the plateau of the correction from Fig. 4. The plateau arises when the correction norms reach the threshold given by the truncation errors and cannot be further decreased by the iterative method. We note that this threshold is also close to the relative errors of the BGRW solution with respect to the FD solution at the final time T shown in Fig. 2 and Table 1. Thus, the accuracy of the BGRW L-scheme is limited by the truncation errors produced by the reduced fluctuations Algorithm 1.

The plateau of the order 10𝒩1 also arises if one adds to the three terms depending on the parameter r in the FD 2 a noisy term randn/𝒩, where randn is a random variable drawn from the standard normal distribution. The results (not shown here) are similar to those from Fig. 4 and Fig. 6. One expects therefore that the accuracy of the random walk methods based on random number generators is limited as well by truncation errors.

Fig. 6 and Fig. 6 show that the length of the sequence of corrections for 𝒩=1010 not affected by oscillations is too short to allow the estimation of the order of convergence p with (29) and the evaluation of the quotient Q1 according to (28). For 𝒩=1018, the sequence not affected by oscillations is three times longer and the behavior of the estimate p(s)1 and of the quotient Q1(s)<1 suggest the C-linear convergence (see Fig. 8 and Fig. 8). This is, according to Remark 5, a numerical argument for the convergence of the iterates to the unique solution which solves the nonlinear scheme (2.1).

Refer to caption
Figure 1. Example 3.1: Comparison of the BGRW (5) and FD (2.1) solutions.
Refer to caption
Figure 2. Example 3.1: Errors of the BGRW solutions.
Refer to caption
Figure 3. Example 3.1: Corrections (N=1010).
Refer to caption
Figure 4. Example 3.1: Corrections (N=1018).
Refer to caption
Figure 5. Example 3.1: Estimated order of convergence (N=1010).
Refer to caption
Figure 6. Example 3.1: Estimation of the quotient Q1 (N=1010).
Refer to caption
Figure 7. Example 3.1: Estimated order of convergence (N=1018).
Refer to caption
Figure 8. Example 3.1: Estimation of the quotient Q1 (N=1018).
Refer to caption
Figure 9. Example 3.1: Errors of the BGRW solution (N=1016).
Refer to caption
Figure 10. Example 3.1: Error-based quotient Q1 of the BGRW solution.

In the absence of an exact solution for this example, we investigate the existence of the “error-based” order C on sequences of errors with respect to the FD solution taken as a reference. We compute BGRW solutions for 𝒩=1016 and form the sequence {𝝍s𝝍FD}. The errors presented in Fig. 10 maintain a linear trend in semi-logarithmic representation over 130 iterations s and allow the evaluation, according to (28), of a subunit quotient Q1 in the same interval, shown in Fig. 10. We can therefore assume that the solution of the BGRW scheme converges with the linear C-order to the reference FD solution.

3.2. Example 2: fully coupled flow and surfactant transport

The next example is the fully coupled problem of variably saturated flow, with transition from saturated to unsaturated regime, and surfactant transport governed by Eqs. (1) and (2) with the exact manufactured solutions

ψm(z,t)=tz(z1)+x4,cm(z,t)=tz(z1)+ 1,

and the water contend parameterized by

θ(ψ,c)={11ψc/10,ψ<00.3,ψ0.

The hydraulic conductivity is constant K(θ(ψ,c))=1 and the diffusion coefficient in Eq. (2) is set to D=1 (in arbitrary units). The manufactured solutions verify Eqs. (1) ans (2) with supplementary source terms [11, 2, 17, 18]. These source terms, generated by inserting the manufactured solutions into the equations, are calculated analytically with the Matlab symbolic toolbox.

We first solve the problem in the domain [0,1], on a regular lattice with δz=2.5102, with the coupled FD scheme (2.1) and BGRW transport scheme (13). The time step Δt=1.56103 is determined according to (26) with Lψ=Lc=100, α=1, and β=0.5. The corrections are evaluated in the l2 norm with the absolute tolerance εa=1018 and the convergence is assessed with the condition max(𝝍s𝝍s1,𝒄s𝒄s1)εa.

The results obtained for 𝒩=1010 particles used in the transport scheme presented in Figs. 12 and 12 show that the truncation errors generated by the transport 3 yield a plateau 1010 of the concentration corrections and also, due to the coupling, a plateau higher than the machine precision of about 1013 of the pressure corrections. The oscillations arising when the corrections reach the plateau do not allow reliable estimations of the order of convergence and of the trend of the quotient Q1, for both pressure (Figs. 14 and 14) and concentration (Figs. 16 and 16). Increasing the number of particles in the transport scheme to 𝒩=1024 lowers the plateau to the machine precision for both pressure and concentration corrections (Figs. 18 and 18), renders more reliable the estimates of the order of convergence p=1 (Figs. 20 and 22), and indicates limits Q1<1, consistent with the C linear convergence (Figs. 20 and 22). According to Remark 5, this result is an indication of the convergence to the solution of the nonlinear schemes (2.1) and (2.2).

Tests (not presented here) with a similar two dimensional problem [17, Sec. 5.2.1] also show the plateau 1013 of the pressure corrections produced by the coupling of flow and transport for 𝒩=1010 particles in the transport code and a plateau at the machine precision for 𝒩=1024 particles.

Refer to caption
Figure 11. Example 3.2: Corrections, of the pressure solution (N=1010).
Refer to caption
Figure 12. Example 3.2: Corrections of the concentration solution (N=1010).
Refer to caption
Figure 13. Example 3.2: Estimation of the convergence order p, pressure (N=1010).
Refer to caption
Figure 14. Example 3.2: Estimation of the quotient Q1, pressure (N=1010).
Refer to caption
Figure 15. Example 3.2: Estimation of the convergence order p, concentration (N=1010).
Refer to caption
Figure 16. Example 3.2: Estimation of the quotient Q1, concentration (N=1010).
Refer to caption
Figure 17. Example 3.2: Corrections of the pressure solution (N=1024).
Refer to caption
Figure 18. Example 3.2: Corrections of the concentration solution (N=1024).
Refer to caption
Figure 19. Example 3.2: Estimation of the convergence order p, pressure (N=1024).
Refer to caption
Figure 20. Example 3.2: Estimation of the quotient Q1, pressure (N=1024).
Refer to caption
Figure 21. Example 3.2: Estimation of the convergence order p, concentration (N=1024).
Refer to caption
Figure 22. Example 3.2: Estimation of the quotient Q1, concentration (N=1024).

The convergence of the iterations does not yet ensure the convergence of the numerical solution to the manufactured solution. This can be verified by a grid convergence test performed by successively halving the space step Δz. To do that, we fix the number of particles in the transport code to 𝒩=1024, which, as seen above, ensures that the truncation errors do not affect the convergence of the corrections, and perform iterations of the linearization schemes (2.1) and (13) until the corrections fall below the threshold εa=1018. The solutions for successive Δz are further used to evaluate the EOC by the rate of logarithmic decay of the deviations from the manufactured solutions measured with the discrete L2 norm Δzl2. The results presented in Table 2 indicate the second order convergence of the solution {𝝍,𝒒,𝒄}.

Δz ψψm EOC qmm EOC ccm EOC
1.00e-1 4.59e-02 1.23e-01 3.43e-02
5.00e-1 1.14e-02 2.01 3.34e-02 1.89 1.02e-02 1.75
2.50e-2 2.85e-03 2.00 8.68e-03 1.94 2.76e-03 1.88
1.25e-2 7.11e-04 2.00 2.21e-03 1.98 7.21e-04 1.94
Table 2. Orders of convergence of the numerical solution to Example 3.2.

The knowledge of the exact solution allows us to compute error-based orders of convergence as well. Using a finer discretization, Δz=1.25102, and a larger tolerance εa=103 we form sequences of discrete L2 norms of errors and corrections for pressure (Figs. 24 and 24) and concentration (Figs. 26 and 26). We note that as the convergence criterion for errors is fulfilled, the corrections are four to five orders of convergence smaller. Estimations (not shown here) of the order of convergence p with (29) were found to be consistent with the definition of the Q linear convergence, for both errors and the corresponding corrections. The quotient Q1 computed with (28) from the sequences of successive errors approaches the unity, indicating the C sublinear convergence of the linearization schemes for both flow and transport (Figs. 28 and 30). We have thus an illustration of the observation made in Remark 6 that C-linear convergence cannot be obtained in numerical simulations. For the corresponding corrections, Q1<1 indicates the C-linear convergence (Figs. 28 and 30), which, according to Remark 5, indicates the convergence of the iterative schemes to the solutions of the coupled nonlinear schemes (2.1) and (2.2).

Refer to caption
Figure 23. Example 3.2: Errors of the pressure solution.
Refer to caption
Figure 24. Example 3.2: Corrections of the pressure solution.
Refer to caption
Figure 25. Example 3.2: Errors of the concentration solution.
Refer to caption
Figure 26. Example 3.2: Corrections of the concentration solution.
Refer to caption
Figure 27. Example 3.2: Estimation of the quotient Q1, pressure errors.
Refer to caption
Figure 28. Example 3.2: Estimation of the quotient Q1, pressure corrections.
Refer to caption
Figure 29. Example 3.2: Estimation of the quotient Q1, concentration errors.
Refer to caption
Figure 30. Example 3.2: Estimation of the quotient Q1, concentration corrections.

3.3. Example 3: One-way coupled flow and transport problem with van Genuchten-Mualem parameterization

In the following, we consider the one-way coupled system (1)-(2), i.e., θ independent of c, with an experimentally based parameterization model. The relationships defining the water content θ(ψ) and the hydraulic conductivity K(θ(ψ)) are given by the van Genuchten-Mualem model

(35) Θ(ψ)={(1+(αψ)n)m,ψ<01,ψ0,
(36) K(Θ(ψ))={KsatΘ(ψ)12[1(1Θ(ψ)1m)m]2,ψ<0Ksat,ψ0,

where Θ=(θθres)/(θsatθres) is the normalized water content defined in the same way as for the exponential model (33)-(34), α, n, m=11/n are model parameters depending on the soil type, and Ksat is the constant hydraulic conductivity in the saturated regime. In this example, we consider the parameters of a silt loam used in previously published studies [8, 17, 18], θsat=0.396, θres=0.131, α=0.423, n=2.06, Ksat=4.96102.

We chose the manufactured solutions

ψm(z,t)=tz(z1)x4+x2,cm(z,t)=tz(z1)+ 1,

which illustrate again the degeneracy of Richards’ equation, with transitions from saturated to unsaturated flow regime. The diffusion coefficient in the transport equation (2) is set to D=0.01. The source terms generated by the manufactured solutions are calculated as in Section 3.2 above with the Matlab symbolic toolbox.

Since the flow is now independent of transport it is not influenced the truncation produced by the 3. Therefore, we focus here on the error-based orders of convergence of the linearization schemes. The coupled problems are solved in the domain [0,1], with Δz=1.25102 and Δt=3.91103. The convergence is assessed with discrete L2 norms of pressure and concentration errors with respect to the manufactured solutions for the tolerance εa=103. Preliminary tests with increasing numbers of particles show that the transport scheme (13) fails to converge if 𝒩<108. To completely eliminate the effect to the truncation errors, the computations are carried out with 𝒩=1018 particles.

As in the previous example, we found that the correction norms reach values that are systematically smaller by several orders of magnitude than the error norms, for both pressure and concentration and the estimates obtained with (29) indicate the Q order of convergence p=1. The quotient Q1 computed with (28) approaches the unity for the errors of both the pressure and the concentration (Figs. 32 and 34) indicating the sublinear convergence. For the corresponding corrections (Figs. 32 and 34) Q1<1, hence the C linear convergence of both iterative schemes may be assumed. These results illustrate again the observations made in Remarks 5 and 6.

Refer to caption
Figure 31. Example 3.3: Estimation of the quotient Q1, pressure errors.
Refer to caption
Figure 32. Example 3.3: Estimation of the quotient Q1, pressure corrections.
Refer to caption
Figure 33. Example 3.3: Estimation of the quotient Q1, concentration errors.
Refer to caption
Figure 34. Example 3.3: Estimation of the quotient Q1, concentration corrections.

The grid convergence test, conducted in the same way as in Section 3.2, presented in Table 3 indicates the quadratic convergence of the solution {𝝍,𝒒,𝒄} towards the exact manufactured solution.

Δz ψψm EOC qmm EOC ccm EOC
1.00e-1 5.85e-02 2.28e-03 1.23e-02
5.00e-1 1.42e-02 2.05 5.69e-04 2.00 3.54e-03 1.80
2.50e-2 3.27e-03 2.11 1.48e-04 1.94 9.19e-04 1.95
1.25e-2 8.09e-04 2.02 4.50e-05 1.72 2.38e-04 1.95
Table 3. Orders of convergence of the numerical solution to Example 3.3.

4. Conclusions

In this study, we demonstrated through numerical examples the convergence of the explicit L-schemes for nonlinear and degenerate problems of coupled flow and transport processes in porous media. The convergence issue is two-fold. First it is the convergence of the iterative method which solves the nonlinearity of the problem and provides a solution of the nonlinear numerical scheme, for instance the L-scheme for Richards’ equation (2.1) which approximates the solution of the nonlinear scheme (2.1). The second aspect is the convergence of the numerical solution to the exact solution of the problem, provided that there exists a unique solution.

The convergence of the iterative schemes has been investigated through estimations of computational and error-based orders of convergence. The three examples presented in Section 3 provide numerical arguments for the C-linear convergence. According to Remark 5, the C-linear convergence indicates the uniqueness of the iterative solutions and their convergence to the solutions of the nonlinear schemes. The sequences of errors with respect to exact manufactured solutions of the coupled problem presented in Section 3.2 and Section 3.3 are only sublinearly C convergent and illustrate the impossibility of achieving the C-liner convergence via numerical simulations, pointed out in Remark 6. However, with the FD solution as reference, we have shown that the random walk BGRW L-scheme for Richards’ equation could be C-linear convergent (Fig. 10). Using grid convergence tests and manufactured solutions we get strong numerical evidence of second order convergence ((Δz)2) of the L-schemes for coupled flow and transport problems (Table 2 and Table 3).

The random walk method for Richards’ equation is rather of academic interest. It has been used, for instance, to prove the stability of the equivalent FD scheme (Remark 1). Instead, the random walk approach is essential in modeling the transport because it facilitates a discrete description at the molecular level of the chemical reactions [18] and provides a basis for the space-time upscaling of the reactive transport in porous media [19]. An issue of concern of this approach are the truncation errors produced by the Algorithm 1 which preserves the indivisibility of the particles. We have shown that the truncation errors induce a plateau in the decrease of the successive corrections of the linearization method (Remark 7). The plateau is inversely proportional to the number of particles, 𝒩1, and may hinder the evaluation of the convergence orders if 𝒩 is not large enough. In modeling reactive transport, where the amount of molecules is of the order of Avogadro’s number, the plateau reaches the machine precision. However, the truncation-limited precision becomes significant in applications for small systems of particles (e.g., [21]), for instance in modeling the isotopic separation of heavy water where the number of deuterium isotopes can be of the order 𝒩106 or even smaller.

Acknowledgements.

F.A. Radu acknowledges the support of the VISTA program, The Norwegian Academy of Science and Letters, and Equinor.

References