Volume 3, Issue 6 e1140
SPECIAL ISSUE PAPER
Full Access

A stabilized semi-Lagrangian finite element method for natural convection in Darcy flows

Loubna Salhi

Corresponding Author

Loubna Salhi

Laboratory of Complex Systems Engineering & Human Systems, University Mohammed VI Polytechnic, Benguerir, Morocco

Correspondence Loubna Salhi, Laboratory of Complex Systems Engineering & Human Systems, University Mohammed VI Polytechnic, Benguerir 43150, Morocco.

Email: [email protected]

Search for more papers by this author
Mofdi El-Amrani

Mofdi El-Amrani

Department of Applied Mathematics, University of Rey Juan Carlos, Madrid, Spain

Search for more papers by this author
Mohammed Seaid

Mohammed Seaid

Department of Engineering, University of Durham, Durham, UK

International Water Research Institute, University Mohammed VI Polytechnic, Benguerir, Morocco

Search for more papers by this author
First published: 01 December 2020
Citations: 4

Abstract

We present an accurate semi-Lagrangian finite element method for the numerical solution of groundwater flow problems in porous media with natural convection. The mathematical model consists of the Darcy problem for the flow velocity and pressure subject to the Boussinesq approximation of low density variations coupled to a convection–diffusion equation for the concentration. The main idea is to combine the semi-Lagrangian method for time integration with finite element method for space discretization, so that the standard Courant–Friedrichs–Lewy condition is relaxed and the time truncation errors are reduced in the diffusion part of the governing equations. We also use a local L2-projection stabilization technique in order to improve the accuracy of the presented method. Numerical simulations are carried out for a test example of a natural convection in an aquifer system with natural boundaries. The obtained results demonstrate the ability of the proposed semi-Lagrangian finite element method to offer efficient and accurate simulations for natural convection in Darcy flows.

1 Introduction

When a saturated aquifer is subject to an internal motion due only to gravity, heat and mass transfer can take place between different regions as a result of temperature and concentration differences. This type of transfer is well known in the literature by natural convection. Physically speaking, the concept refers to the process of moving energy from or to a material by means of an adjacent fluid in motion in the presence of a gravitational field. This fluid movement is generated exclusively by differences in the density as a result of temperature and concentration gradients within the fluid itself but not by any external source such as a pump, extractor or fan. Natural convection is an important topic that has received growing attention in theoretical and applied researches over recent decades because of its presence in nature and many industrial processes. For example, in the process of mixing ocean waters and storms. In industry, the use of natural convection for heat dissipation is a determining factor for some electronic equipments. The effect of natural convection on the fluid flow was first studied experimentally in concentric spheres and horizontal tubes of various dimensions in References 1-3. It has been shown that this natural convection depends on the variation in some dimensionless control parameters such as Reynolds, Prandtl, Grashof, and Nusselt numbers. In other geometric configurations considered later in References 4, 5, it was shown that this complex process involves also the Frank–Kamenetskii parameter along with the Rayleigh number and the stationary and oscillating regimes on the thermal transfer behavior can occur. Authors showed how complex regimes appeared through successive bifurcations leading from a stable stationary temperature distribution without convection to a stationary symmetric and asymmetric convective solutions and later to a periodic and aperiodic oscillations. In Reference 6, the impact of the critical value of the thickness of the porous layers on the characteristics of natural convection in an open-ended squared cavity was discussed numerically. In Reference 7, the existence of an adiabatic obstacle in a horizontal enclosure of fluid was taken into account, and numerical simulations were developed for different Rayleigh numbers at which the dynamic behavior is evolved from a steady-state to a chaotic pattern. In Reference 8, the case of natural convective flow was studied within an inclined rectangular enclosure by using the linear stability analysis. It was found that, the natural convection generates more stable inclined propagating fronts than those obtained without inclination. In all these works, natural convection is considered in a gaseous or liquid medium with their motion described either by the Navier–Stokes equations or by the Darcy equation in a quasi-stationary form under the Boussinesq approximation. In underground applications, one key issue is how injection of water or other solutes into aquifers affects the natural convection along with the flowing groundwater in the presence of underground services such as pipes or piles, and how they can contribute to wearing down these services. The answer may help to troubleshoot many concerns such as the use of suitable materials for underground services, the detection of parameters and factors affecting performance, the selection of the optimum installation location, and thus to prevent from wrong decisions that can give rise to undesired and nonreversible effects in terms of underground materials lifetime.

The purpose of this article is to present a stabilized semi-Lagrangian finite element method for the numerical study of natural convection effects on groundwater flows and solute transport. Consider for example an unconfined aquifer which is recharged primarily from surface water deposits such as snow melt and precipitation, that flow through underground pipes and drain gradually toward a discharge area such as a river, sea, or lake. This study is intended to be useful not only to simulate the solute transport and groundwater flow but also to show its variation in case of the introduction of solutes other than water or contaminants within the recharge area, that can cause aquifer contamination and then lead to important changes of critical hydrological and medium properties. Examples of groundwater flow models with natural convection that have been studied in similar ways can be found in References 9-14 among others. The governing equations consist of a coupled Darcy problem for the flow and a convection–diffusion equation for the concentration. This includes a wide variety of difficulties which typically arise in the numerical approximation of partial differential equations which describe and consequently determine their dynamics. Current trends in incompressible viscous flows are to combine these equations with complex components to simulate various problems in many applications such as fluid mechanics, ocean circulation, turbulence and multiphase flow models. All these applications require a very efficient and robust numerical solvers for the coupled systems computing the numerical solutions is very often not easy due to the nonlinear form and the presence of the convective terms. For many viscous hydrodynamics problems, the convective term is distinctly more important than the diffusive term, particularly when the Reynolds numbers reach high values, as it becomes a source of computational difficulties and oscillations. Furthermore, It is well known that the solutions of these equations present steep fronts, which need to be resolved accurately in applications and often cause severe numerical difficulties. Hence, semi-Lagrangian methods as known in meteorological community make use of the transport nature of the governing equations by combining the fixed Eulerian grids with a particle tracking along the characteristic curves of the problem under study, see for instance References 15-19. The main idea in these approaches is to rewrite the governing equations in terms of Lagrangian coordinates as defined by the particle trajectories (or characteristic curves) associated with the problem under consideration. The time derivative and the convection terms are combined as a directional derivative along the characteristics leading to a characteristic time-stepping procedure. The Lagrangian treatment in these methods significantly reduces the time truncation errors in the Eulerian methods. Moreover, the semi-Lagrangian method offers the possibility of using time steps that exceed those permitted by the Courant–Friedrichs–Lewy (CFL) stability condition in the Eulerian methods for convection-dominated flows. A class of semi-Lagrangian methods has been investigated in References 16, 20-23 among others. The authors in Reference 24 have studied a semi-Lagrangian method for the incompressible Navier–Stokes equations in the finite element framework. In this reference, the solutions at characteristic feet are approximated by interpolation from finite element basis functions. A first-order semi-Lagrangian approach combined with finite element method has been analyzed for the Navier–Stokes equations in Reference 22. It has been shown that the method is unconditionally stable provided the characteristics are transported by a divergence-free field that is deduced from the flow velocity. The case where the characteristics are transported by a discrete velocity field which is not divergence-free has been studied in Reference 23. Analysis of a semi-Lagrangian method using finite difference discretization has been presented in Reference 20 for convection–diffusion equations.

In most numerical methods using finite element discretizations, mixed formulations in which two finite element spaces are used to approximate the velocity and pressure variables have been well established. However, in the current study, the proposed finite element method consists of using the same approximation spaces for all the variables including the pressure, velocity and concentration. We also the L2-projection to stabilize the finite element method for the coupled system. A similar unified finite element method has been developed and analyzed in Reference 25 for several Darcy flow problems. It is based on regularization of the mixed variational form of the Darcy equations by using the same low-order finite element spaces to approximate both velocity and pressure variables. The method is an extension of the proposed polynomial pressure projection stabilization of the Stokes equations presented in References 26, 27. The results have shown that the method is stable and its accuracy is competitive with that of other mixed finite element formulations since the matrix condition numbers are better and it is less sensitive to the choice of the stabilization parameter used in the stabilized finite element discretizations. This article is organized as follows. The description of the mathematical model is presented in Section 2. Section 3 is devoted to the formulation of the unified semi-Lagrangian finite element method. This section includes the finite element solution of Darcy problem, calculation of departure points, and the formulation of semi-Lagrangian method for solving the convection–diffusion equation. The numerical results for a groundwater benchmark problem are presented in Section 4. Section 5 contains some conclusions and perspectives.

2 Mathematical model

In the present study, we consider natural convection by Darcy flows in two-dimensional domains. An illustration of the physical model considered is shown in Figure 1. Here, a groundwater flow problem in an unconfined aquifer with a well is recharged by a solute with an inlet concentration and subject to a thermal variation ( c c 0 ) , where c is the concentration at saturation and c 0 is the reference concentration. Let Ω R 2 be an open bounded domain with piecewise smooth boundaries Γ as shown in Figure 1. The mathematical formulation consists of coupling Darcy problem for velocity and pressure to a convection–diffusion equation for the change in the concentration. To close the coupled system, a state equation is used to update the density distribution. These equations involve various thermophysical parameters namely the velocity u, pressure p′, concentration of the solute c′, and the mass density ρ which depend continuously on the location x = ( x , z ) Ω and the time t′. The fluid is taken to be Darcian and variations in its properties other than density are ignored. Under the Boussinesq approximation, it is assumed that the density variation has no influence on the flow field except that it gives rise to buoyancy forces. With these constitutive assumptions, the governing equations in a two-dimensional domain are:

Details are in the caption following the image
Illustration of a groundwater flow problem in an aquifer with a well and subject to recharge and discharge boundary conditions
Darcy equations:
u = K p ϵ μ p x , v = K p ϵ μ p z g g c ρ , u x + v z = Φ . ()
Energy equation:
c t + u c x + v c z κ 2 c x 2 + 2 c z 2 = Q ϵ γ c . ()
State equation:
ρ = ρ 0 ( 1 β c ( c c 0 ) ) . ()

The areas of recharge and discharge along with the other areas are considered as boundary conditions. These can be specified as Neumann or Dirichlet conditions depending on the case to be studied. In the Darcy equation (1), u = (u′, v′) is the velocity field, p′ the pressure, ρ the density, ϵ the porosity, μ the dynamic viscosity of the fluid, Φ the volumetric flow rate source or sink terms, K p the permeability, g the gravitational force, and gc is a conversion constant that can be changed so as to increase or decrease the influence of gravity in the system References 28, 29. In the transport equation (2), c′ is the concentration, Q′ the source term, γ the rate of concentration distribution, and κ the diffusion coefficient which may depend on the molecule size and other properties of the diffusing substance as well as on concentration and pressure. The state equation (3) describes the relationship between density and concentration by neglecting any changes in the temperature. Here, c 0 and ρ 0 are, respectively, the initial or reference concentration and density of the solute, and β c is the solutal expansion coefficient which depends on the solute type dissolved in the groundwater.

Since it is convenient to work with dimensionless formulations, we define the following variables
x = x L , t = κ t L 2 , u = L u ϵ κ , p = L 2 p ϵ 2 κ μ , c = c c Δ c , c 0 = c c 0 Δ c , c = c Δ c ,
where L is a characteristic length, p = p + ρ 0 g g c e z is the driving pressure and Δ c is the concentration variation. We also define the following dimensionless parameters
K p = K p L 2 , R a = g β c Δ c L 3 κ ν , R p = R a ϵ 2 g c , Φ = L 2 Φ ϵ κ , Q = L 2 Q ϵ κ Δ c , γ = L 2 γ κ ,
where Ra is the Rayleigh number formulated using the kinematic viscosity ν = μ ρ 0 . In the absence of mechanical diffusion, the solute concentration distribution in groundwater flow is generally assumed to be controlled by the value of Rayleigh number. Hence, Equations (1)–(3) can be rewritten in the dimensionless form as
u = K p ( p R p ( c + c 0 ) e z ) , ()
· u = Φ , D c D t 2 c = Q γ ( c + c ) , ()
where = x , z denotes the gradient operator and D c D t is the material derivative defined as
D c D t = c t + u · c . ()

Note that the system (4)–(5) has to be solved in the bounded spatial domain Ω with given boundary conditions on Γ and given initial conditions. Initially, the concentration is set to zero and a source of the solute is assumed to generate a propagating plume inside the domain. In this study, we propose a numerical technique combining the semi-Lagrangian method and finite element discretization for solving the coupled problem (4)-(5). We also consider a fractional step procedure where the convective part in (5) is decoupled from the generalized system of equations in the time integration process. First, the velocity and pressure are obtained by solving the Darcy problem (4) at each time step then, the concentration is updated at each time step by solving the convection equation (5).

3 A stabilized semi-Lagrangian finite element method

In order to formulate the semi-Lagrangian finite element method we first proceed with the discretization of the space domain Ω = Ω Γ . We generate a quasi-uniform partition Ω h Ω of small elements 𝒦 j . The conforming finite element spaces for concentration and pressure can be defined as
V h = { v h C 0 ( Ω ) : v h | 𝒦 j P k ( 𝒦 j ) , 𝒦 j Ω h } ,
where P k ( 𝒦 j ) is the space of complete polynomials of degree k defined in 𝒦 j . We also define the finite element space Vh for the velocity field as
V h = V h × V h .
Next, we divide the time interval into subintervals [tn, tn + 1] with length Δ t = t n + 1 t n for n = 0, 1, …. We use the notation wn to denote the value of a generic function w at time tn. Then, we formulate the finite element solutions to un(x), cn(x) and pn(x) as
u h n ( x ) = j = 1 M 𝒰 j n φ j ( x ) , c h n ( x ) = j = 1 M 𝒞 j n ϕ j ( x ) , p h n ( x ) = j = 1 M 𝒫 j n ϕ j ( x ) . ()

The symbol ∘ denotes the Hadamard product that produces vectors through element-by-element multiplication of the original two vectors. The functions 𝒰 j n , 𝒞 j n , and 𝒫 j n are respectively, the corresponding nodal values of the velocity u h n ( x ) , concentration c h n ( x ) and pressure p h n ( x ) defined as 𝒰 j n = u h n ( x j ) , 𝒞 j n = c h n ( x j ) and 𝒫 j n = p h n ( x j ) , where { x j } j = 1 M is the set of mesh points in the partition Ω h . In (7), { φ j } j = 1 M = { ( ϕ j , ϕ j ) } j = 1 M and { ϕ j } j = 1 M are the basis vectors and functions of Vh and Vh respectively given by the Kronecker delta symbol.

3.1 Solution of Darcy problem

In this section, we describe the unified finite element method used to solve the Darcy problem, Hence, we rewrite Equation (4) in the spatial domain Ω as
u + K p p = f , · u = Φ , ()
subject to the following boundary condition u · n = 0, where n is the unit outward normal vector on Γ . In (8), f = K p R p ( c + c 0 ) e z and the function Φ satisfies the compatibility condition Ω Φ d Ω = 0 .
Hence, the weak formulation for Equation (8) is: Find ( u , p ) H 0 ( Ω , d i v ) × L 2 ( Ω ) such that
Ω u v d Ω Ω K p p v d Ω = Ω f v d Ω , v H 0 ( Ω , d i v ) , Ω ( · u ) q d Ω = Ω Φ q d Ω , q L 2 ( Ω ) . ()
The weak formulation (9) can be written in a compact form as
𝒜 ( u , v ) + ( p , v ) = f ( v ) , v H 0 ( Ω , d i v ) , 𝒞 ( u , q ) = Φ ( q ) , q L 2 ( Ω ) , ()
where 𝒜 , , and 𝒞 are bilinear forms defined as
𝒜 ( u , v ) = Ω u v d Ω , ( p , v ) = Ω K p p v d Ω , 𝒞 ( u , q ) = Ω ( · u ) q d Ω , ()
and f and Φ are linear forms defined as
f ( v ) = Ω f v d Ω , Φ ( q ) = Ω Φ q d Ω . ()
To approximate the solutions in (11)–(12), the proposed unified finite element method consists of using the same polynomial basis for velocity, concentration, and pressure solutions on the element 𝒦 j of the discrete domain Ω h . Thus, the conforming finite element spaces that we use are Pk (k = 1, 2), that is, polynomials of the same degree k for velocity, concentration and pressure defined on the element 𝒦 j . Hence, to approximate the solution of (10), we define the pair ( S h , Q h ) where
S h = V h H 0 ( Ω , d i v ) and Q h = V h L 2 ( Ω ) . ()
Then, the discrete weak form of (11)–(12) becomes: Find ( u h , p h ) S h × Q h such that for all ( v h , q h ) S h × Q h ,
𝒜 ( u h , v h ) + ( p h , v h ) = f ( v h ) , 𝒞 ( u h , q h ) = Φ ( q h ) . ()
Note that the pair ( S h , Q h ) does not verify the inf–sup condition associated with the mixed form (14), then the discrete weak problem (14) is not stable. Thus, we use the polynomial pressure-projection stabilization method introduced in Reference 25 to stabilize (14). This approach consists in defining a local L2-projection operator onto the discontinuous polynomial space as
[ P ] k 1 = { q h L 2 ( Ω ) : q h | 𝒦 j P k 1 ( 𝒦 j ) , 𝒦 j Ω h } .
The projection operator Π k 1 : L 2 ( Ω ) [ P ] k 1 is defined by
Π k 1 ( p ) = arg min 1 2 Ω ( Π k 1 q p ) 2 d Ω .
Thus, the modified polynomial pressure-projection stabilized method yields: Find ( u h , p h ) S h × Q h such that for all ( v h , q h ) S h × Q h ,
𝒜 ( u h , v h ) + ( p h , v h ) = f ( v h ) , 𝒞 ( u h , q h ) + β 𝒟 ( p h , q h ) = Φ ( q h ) , ()
where 𝒟 is a bilinear form defined by
𝒟 ( p h , q h ) = Ω ( p h Π k 1 p h ) ( q h Π k 1 q h ) d Ω .
In (15), the parameter β = β 1 L 2 with L is a characteristic scale for the length and β is a dimensionless positive real parameter independent of the space discretization parameter h. More details on the selection of this stabilization parameter can be found in References 25-27. It should be stressed that, as in most finite element methods for Darcy problems, the solution pair (u, p) is obtained by solving a linear system of algebraic equations from (15). The stabilized unified method described above has been evaluated in Reference 25 for the Darcy flow problem. The authors presented an error analysis and comparison to examples with known analytical solutions. The results presented have shown that the method is accurate, efficient, and stable. In addition, compared with other finite element methods, the condition numbers associated with the linear systems for the current method are less than other finite element methods. Moreover, the considered unified finite element method has been found to be not strongly sensitive to the choice of the dimensionless parameter β in (15). On the basis of these results, the stabilized unified method can be applied to other problems in which the Darcy problem is coupled with the transport equation.

3.2 Solution of convection–diffusion equation

In the current work, the convection–diffusion problem (5) is solved using a semi-Lagrangian finite element method. The present method requires two steps namely, the computation of characteristic trajectories and the interpolation procedure. In practice, both steps are crucial to the overall accuracy of semi-Lagrangian methods. We consider for each mesh point xj, j = 1, … , M, the characteristic curves X h j n = X h ( x j , t n ) associated with the material derivative (6) are the unique solutions of the ordinary differential equations
d X h ( x j , t ) d t = u h ( t , X h ( x j , t ) ) , t [ t n , t n + 1 ] , X h ( x j , t n + 1 ) = x j . ()
The solutions of (16) are known by departure points at time t of a particle passing through the mesh point xj at time t = tn + 1. To compute the departure points { X h j n } , j = 1, … , M, we use the algorithm proposed in Reference 19 which accurately solves (16) with a second-order accuracy. Let us write the solution of (16) as
X h j n = x j d h j , j = 1 , , M , ()
where the displacement dhj is calculated by an iterative procedure as
d h j ( 0 ) = Δ t 2 ( 3 u h n ( x j ) u h n 1 ( x j ) ) , d h j ( k + 1 ) = Δ t 2 3 u h n x j 1 2 d h j ( k ) u h n 1 x j 1 2 d h j ( k ) , k = 0 , 1 , . ()

We first identify the mesh element 𝒦 ^ j where x j 1 2 d h j ( k ) resides to evaluate values of the velocities u h n x j 1 2 d h j ( k ) and u h n 1 x j 1 2 d h j ( k ) . Then, a finite element interpolation on 𝒦 ^ j is carried out according to (7). This technique was first proposed in Reference 19 for combined finite difference and semi-Lagrangian methods, and it was also used for finite element methods in other studies such as References 16, 21. In our numerical simulations, the iterations in (18) were continued until the trajectory changed by less than 10−7.

We assume that, for all j = 1, … , M, the pairs ( X h j n , 𝒦 ^ j ) and the mesh point values 𝒞 j n are known, we compute the values 𝒞 ^ j n as
𝒞 ^ j n : = c h n ( X h j n ) = k = 1 M 𝒞 k ϕ k ( X h j n ) . ()
The solution c ^ h n ( x ) of the convection equation (6) is then given by
c ^ h n ( x ) = j = 1 M 𝒞 ^ j n ϕ j ( x ) . ()
Hence, the solution of convection part in (5) to approximate solution c h n ( x ) is obtained using the following steps:
  1. Calculate the departure points Xh(xj, tn) using (17) along with the iteration procedure (18).
  2. Identify the mesh element 𝒦 ^ j where the departure points Xh(xj, tn) are located.
  3. Evaluate the gridpoint approximations c ^ h n ( x ) using (20).

In practice, the above convection fractional step allows to avoid grid distortion problems present in the forward tracking methods, since it follows the flow by tracking the characteristics backward from a point x in a fixed grid at the time tn + 1 to a point X at the previous time tn.

3.3 Time integration procedure

For simplicity in the presentation, the time integration of the coupled system is carried out using a first-order implicit Euler scheme. The method consists of the following two steps:
  • Step 1. Solve the Darcy problem at the next time step for u h n + 1 and p h n + 1 :

    u h n + 1 + K p p h n + 1 = f h n , · u h n + 1 = Φ , ()
    where f h n = K p R p ( c h n + c 0 ) e z .

  • Step 2. Solve the convection–diffusion problem for c h n + 1 :

    c h n + 1 c ^ h n Δ t 2 c h n + 1 + γ c h n + 1 = Q γ c . ()

Based on the above variational formulation along with the local L2-projection stabilization technique used, Equations (21)–(22) can be written as a system of linear equations that may be represented in a matrix form as
A U h n + 1 + B P h n + 1 = F h n , C U h n + 1 + β D P h n + 1 = Φ , ( [ M ] + Δ t [ S ] + Δ t γ [ M ] ) C h n + 1 = [ M ] C ^ h n + Δ t ( Q + C ) , ()
where U h n + 1 = ( 𝒰 1 n + 1 , , 𝒰 M n + 1 ) , P h n + 1 = ( 𝒫 1 n + 1 , , 𝒫 M n + 1 ) , C h n + 1 = ( 𝒞 1 n + 1 , , 𝒞 M n + 1 ) , and C ^ h n = ( 𝒞 ^ 1 n , , 𝒞 ^ M n ) according to (7). Here, A, B, C, and D are matrices whose elements are given respectively by
a i j = Ω φ j φ i d Ω , b i j = Ω K p ϕ j φ i d Ω , c i j = Ω ( · φ j ) ϕ i d Ω , d i j = Ω ϕ j ϕ i d Ω , i , j = 1 , , M .
The right-hand sides F 0 h n and Φ are vectors whose components are given by
F h n = Ω f h n φ 1 d Ω , , Ω f h n φ M d Ω , Φ = Ω Φ ϕ 1 d Ω , , Ω Φ ϕ M d Ω .
In (23), [ M ] and [ S ] are respectively the mass and stiffness matrices whose elements are
m i j = Ω ϕ j ϕ i d Ω and s i j = Ω ϕ j ϕ i d Ω , i , j = 1 , , M .
The terms Q and C on the right-hand side are given by
Q = Ω Q ϕ 1 d Ω , , Ω Q ϕ M d Ω and C = Ω γ c ϕ 1 d Ω , , Ω γ c ϕ M d Ω .

Note that in the current study, the same finite elements are used for solving both the Darcy problem and the convection–diffusion–reaction equations without requiring mixed formulations widely used in the literature. This results in a simple and efficient implementation of the method without need for mixed finite element techniques. In our simulation, linear systems of algebraic equations are solved using the conjugate gradient solver with incomplete Cholesky decomposition, and all stopping criteria for iterative solvers are set to 10−7 which is small enough to guarantee that the algorithm truncation error dominates the total numerical error.

Details are in the caption following the image
Unstructured mesh (left) and initial concentration used in our simulations
Details are in the caption following the image
Results for velocity field (first column), pressure (second column) and concentration (third column) obtained at times t = 0.1 (first row), t = 0.3 (second row), t = 0.5 (third row), t = 0.7 (fourth row), and t = 1.8 (fifth row) using Rp = 0
Details are in the caption following the image
Results for velocity field (first column), pressure (second column) and concentration (third column) obtained at times t = 0.1 (first row), t = 0.3 (second row), t = 0.5 (third row), t = 0.7 (fourth row), and t = 1.8 (fifth row) using Rp = 1
Details are in the caption following the image
Results for velocity field (first column), pressure (second column) and concentration (third column) obtained at times t = 0.1 (first row), t = 0.3 (second row), t = 0.5 (third row), t = 0.7 (fourth row), and t = 1.8 (fifth row) using Rp = 10
Details are in the caption following the image
Results for velocity field (first column), pressure (second column) and concentration (third column) obtained at times t = 0.1 (first row), t = 0.3 (second row), t = 0.5 (third row), t = 0.7 (fourth row), and t = 1.8 (fifth row) using Rp = 100

4 Numerical Results

We present numerical results for a groundwater flow problem in an aquifer with a well and subject to recharge and discharge boundary conditions shown in Figure 1. This test example has been proposed in Reference 30 as a benchmark for Darcy flows and widely used in the literature for validation purposes, see References 13, 31 among others. For example, in Reference 13 this test example has been as a prototype for the realistic case of Berrechid aquifer located in Morocco which has been the subject of various geological and hydrogeological studies. This is due to its socioeconomic value as an important local natural water resource of the region. In the last decades, its improper operation on account of the agricultural and industrial development associated with a high population growth led to a considerable increase in conflict and significant economic losses. In the present study, we are interested in a transport problem for which the Darcy flow problem is coupled to a convection–diffusion equation for the solute. Therefore, this test problem is more challenging than those studied in References 13, 30, 31. Hence, the problem statement consists of solving the coupled system (4)–(5) in the computational domain shown in Figure 1. Here, the unconfined aquifer is bounded on the east and west by Neumann boundary conditions, and on the north and south by Dirichlet boundary conditions. On the surface of the well, no-slip boundary conditions are imposed. Initially, the concentration is considered as a Gaussian function centered in the domain with an amplitude set to unity as
c ( x , y , 0 ) = exp ( x 0 . 5 ) 2 + ( y 0 . 5 ) 2 0 . 002 .
An unstructured mesh of 7180 elements and 14,614 nodes is used in our simulations and quadratic P2 elements are used to approximate the concentration, velocity, and pressure based on the stabilized unified technique described above in Section 3. In Figure 2 we illustrate the computational mesh and the initial concentration considered in this study. Note that, in practice, computations should be considered over large areas (in kilometers) and over large time periods (in years), but it would be inefficient to perform simulations on this scale. Therefore, we use dimensionless quantities to reduce computation times. Hence, the dimensionless permeability Kp = 1, the rate of concentration distribution γ = 1 , the reference concentration c0 and the source terms Φ and Q are set to zero. The time step Δ t = 0 . 001 is used in our simulations and numerical results are displayed at five different instants namely, t = 0.1, 0.3, 0.5, 0.7, and 1.8. Effects of the solute Rayleigh number Rp on the flow velocity, pressure and solute concentration are examined in this section. To this end, we run the simulations for Rp = 0, Rp = 1, Rp = 10, and Rp = 100 which cover a large variety of applications for transport in groundwater flows. It should be mentioned that, in the absence of physical diffusion, the variation of solute concentration in groundwater flows is basically supposed to be controlled by the values of the Rayleigh number which is used to characterize the flow regime and to also quantify the natural convection. Recall that by definition R p = R a / ( ϵ 2 g c ) , where ϵ is the porosity and gc a conversion constant. In the case considered in this study, ϵ 2 = 1 and gc = 1 for which the two Rayleigh numbers coincide, that is, Rp = Ra.

In Figures 3-6 we depict velocity fields and snapshots of the pressure and concentration obtained at five different instants using Rp = 0, Rp = 1, Rp = 10, and Rp = 100, respectively. It should be stressed that the selection of these values for the Rayleigh number Rp is adopted here to check the performance of the proposed stabilized semi-Lagrangian finite element method for diffusion-dominated flows at Rp = 0 and convection-dominated flows at Rp = 100. In both cases, the results presented in Figures 3-6 demonstrate that the proposed method is able to simulate solute transport in groundwater at different flow regimes without requiring complicated mixed finite elements and stabilization techniques. In all considered cases, the concentration plume is transported toward the well and both velocity fields and pressure structures are altered by the solute concentration. At low Rayleigh numbers, the distribution of the solute concentration is characterized by high momentum diffusion and low momentum convection with a high local pressure around the solute. At higher Rayleigh numbers, convective term dominates and as a consequence, the solute concentration is mainly transported by the natural convection rather than the diffusion and the local pressure becomes less important. Notice that the main physical structures expected for this type of problems have been accurately captured by our stabilized semi-Lagrangian finite element method. In addition, the performance of our method is very attractive since the computed solutions remain stable without nonphysical oscillations appearing in the vicinity of sharp gradients in the solute concentration and pressure distributions. These features are resolved without requiring small time steps or solvers for nonlinear systems of algebraic equations.

5 Conclusions

In this article, a stabilized semi-Lagrangian finite element method is developed to numerically study natural convection in porous medium using coupled Darcy flow problem and a convection–diffusion equation. This can be useful in several instances such as: to accurately simulate the dispersion of solutes or pollutants released in groundwater aquifers in order to assess and predict their deplorable impact on the quality of water, to efficiently help understanding the effect of these solutes in the presence of underground services, and to assist in decision-making for optimizing the choice of underground materials. The numerical results obtained for the groundwater flow problem in an aquifer with a well are very encouraging, since they demonstrate the capability of the method to provide insight into the groundwater flow behavior in the presence of solutes or pollutants. As future works, we intend to extend the numerical method to more generalized groundwater flow problems. We will focus on studying the equations governing the hydrodynamic dispersion of a solute in heterogeneous and anisotropic porous media. In this case, the dispersion term is treated as a tensor that is characterized by the principal direction that coincides with that of the groundwater flow. Then, a spatially varying velocity field leading to an eventual change of the anisotropy direction, is expected to appear due to heterogeneity in the porous medium. This results in a system subject to a wide variety of computational difficulties which generally arise in the numerical approximation of the associated partial differential equations that should capture the effects of these spatial changes.

Biography

  • Loubna Salhi received the Ph.D. degree in Applied Mathematics in 2017 from the Faculty of Science & Technology of the University Hassan II Casablanca, Morocco. Broadly, her methodological research focuses on theoretical and numerical methods for partial differential equations applied to fluid flow problems in science and engineering. She has worked on the asymptotic analysis, linear stability analysis, Galerkin-characteristic finite element methods, and more recently on meshless methods. She is currently a postdoctoral researcher, at the laboratory of Complex Systems Engineering and Human Systems of the University Mohammed VI Polytechnic, Benguerir, Morocco. Her research interests lie in various fields such as propagating reaction–diffusion fronts, groundwater flows with natural convection and multiphase flow problems.

    The full text of this article hosted at iucr.org is unavailable due to technical difficulties.