Solutions of the Elder problem for a wide range of the Rayleigh number

Similar documents
A note on benchmarking of numerical models for density dependent flow in porous media

Lattice Boltzmann model for the Elder problem

Speed of free convective fingering in porous media

Transience of Free Convective Transport in Porous Media

Rupelton layer. 1 An introduction to the problem and some basic definitions and an introduction to the problem

Studies on flow through and around a porous permeable sphere: II. Heat Transfer

Thermohaline Flow and Reactive Solute Transport in Fractured Rock

Numerical modelling of coupled TH processes to facilitate analysis of geothermal reservoirs - Groß Schönebeck as an example-

Scaling Behavior of Convective Mixing, with Application to Geological Storage of CO 2

EFFECTIVE DISPERSION IN SEAWATER INTRUSION THROUGH HETEROGENEOUS AQUIFERS

RATE OF FLUID FLOW THROUGH POROUS MEDIA

Role of salt sources in density-dependent flow

Onset of convection of a reacting fluid layer in a porous medium with temperature-dependent heat source

Numerical Simulation of Natural Convection in Heterogeneous Porous media for CO 2 Geological Storage

Thermal Analysis Contents - 1

Effect of Variable Viscosity on Convective Heat and Mass Transfer by Natural Convection from Vertical Surface in Porous Medium

Flow of Non-Newtonian Fluids within a Double Porosity Reservoir under Pseudosteady State Interporosity Transfer Conditions

An Introduction to COMSOL Multiphysics v4.3b & Subsurface Flow Simulation. Ahsan Munir, PhD Tom Spirka, PhD

ENTROPY GENERATION IN HEAT AND MASS TRANSFER IN POROUS CAVITY SUBJECTED TO A MAGNETIC FIELD

Tank Experiments, Numerical Investigations and Stochastic Approaches of Density-Dependent Flow and Transport in Heterogeneous Media

Salt transport on islands in the Okavango Delta: Numerical investigations

Effect of Buoyancy Force on the Flow Field in a Square Cavity with Heated from Below

Improving the worthiness of the Henry problem as a benchmark for density-dependent groundwater flow models

Where does free convection (buoyancy and density driven) flow occur?

Simulation study of density-driven natural convection mechanism in isotropic and anisotropic brine aquifers using a black oil reservoir simulator

A breakthrough curve analysis of unstable density-driven flow and transport in homogeneous porous media

Manfred Koch and Bettina Starke. Department of Geohydraulics and Engineering Hydrology University of Kassel, Germany

J. Environ. Res. Develop. Journal of Environmental Research And Development Vol. 8 No. 1, July-September 2013

pifreeze A Freeze / Thaw Plug-in for FEFLOW User Guide

The Saltwater Intrusion Problem: Concepts, Challenges, Modeling, Experiments and Case Studies

INFLUENCE OF VARIABLE PERMEABILITY ON FREE CONVECTION OVER VERTICAL FLAT PLATE EMBEDDED IN A POROUS MEDIUM

SIMILARITY SOLUTION FOR DOUBLE NATURAL CONVECTION WITH DISPERSION FROM SOURCES IN AN INFINITE POROUS MEDIUM

11280 Electrical Resistivity Tomography Time-lapse Monitoring of Three-dimensional Synthetic Tracer Test Experiments

Towards a Numerical Benchmark for 3D Low Mach Number Mixed Flows in a Rectangular Channel Heated from Below

Eigenvalue Analysis of Waveguides and Planar Transmission Lines Loaded with Full Tensor Anisotropic Materials

Ping Cheng Department of Mechanical Engineering University of Hawaii Honolulu, Hawaii 96822

Available online at Energy Procedia 1 (2009) (2008) GHGT-9

2. Governing Equations. 1. Introduction

Introduction. Statement of Problem. The governing equations for porous materials with Darcy s law can be written in dimensionless form as:

Numerical Solution of the Two-Dimensional Time-Dependent Transport Equation. Khaled Ismail Hamza 1 EXTENDED ABSTRACT

Dispersion relations for the convective instability of an acidity front in Hele-Shaw cells

Evaluation of the hydraulic gradient at an island for low-level nuclear waste disposal

MHD and Thermal Dispersion-Radiation Effects on Non-Newtonian Fluid Saturated Non-Darcy Mixed Convective Flow with Melting Effect

O.R. Jimoh, M.Tech. Department of Mathematics/Statistics, Federal University of Technology, PMB 65, Minna, Niger State, Nigeria.

Chapter 2 Thermal Processes

Pollution. Elixir Pollution 97 (2016)

Investigation of different wall profiles on energy consumption and baking time in domestic ovens

Numerical simulation of double-diffusive finger convection

MODELING THE SALTWATER INTRUSION PHENOMENON IN COASTAL AQUIFERS - A CASE STUDY IN THE INDUSTRIAL ZONE OF HERAKLEIO IN CRETE

Introduction. Vincent E. A. Post & Craig T. Simmons

Effect of an adiabatic fin on natural convection heat transfer in a triangular enclosure

(C) Global Journal Of Engineering Science And Researches

CHARACTERIZATION OF HETEROGENEITIES AT THE CORE-SCALE USING THE EQUIVALENT STRATIFIED POROUS MEDIUM APPROACH

Advanced numerical methods for transport and reaction in porous media. Peter Frolkovič University of Heidelberg

Carbon dioxide dissolution in structural and stratigraphic traps

NUMERICAL STUDIES OF TRANSITION FROM STEADY TO UNSTEADY COUPLED THERMAL BOUNDARY LAYERS

Anisotropic Modelings of the Hydrothermal Convection in Layered Porous Media

1.061 / 1.61 Transport Processes in the Environment

MATLAB Solution of Flow and Heat Transfer through a Porous Cooling Channel and the Conjugate Heat Transfer in the Surrounding Wall

Study on MHD Free Convection Heat and Mass Transfer Flow past a Vertical Plate in the Presence of Hall Current

Numerical Investigation of Fluid and Thermal Flow in a Differentially Heated Side Enclosure walls at Various Inclination Angles

Introduction to Heat and Mass Transfer. Week 12

Addition of simultaneous heat and solute transport and variable fluid viscosity to SEAWAT $

NATURAL CONVECTIVE BOUNDARY LAYER FLOW OVER A HORIZONTAL PLATE EMBEDDED

Improving the worthiness of the Elder problem as a benchmark for buoyancy driven convection models

Comparison of Heat and Mass Transport at the Micro-Scale

Field Scale Modeling of Local Capillary Trapping during CO 2 Injection into the Saline Aquifer. Bo Ren, Larry Lake, Steven Bryant

Available online at ScienceDirect

F59: Active Solute Transport in Saturated Porous Media

Numerical techniques for simulating groundwater flow in the presence of temperature gradients

Natural Convection in Vertical Channels with Porous Media and Adiabatic Extensions

Frozen Ground Containment Barrier

Flow patterns and heat transfer in square cavities with perfectly conducting horizontal walls: the case of high Rayleigh numbers ( )

PHYS 432 Physics of Fluids: Instabilities

ES265 Order of Magnitude Phys & Chem Convection

GG655/CEE623 Groundwater Modeling. Aly I. El-Kadi

Appendix: Nomenclature

Chapter 3 Permeability

Sunday September 28th. Time. 06:00 pm 09:00 pm Registration (CAAS) IP: Invited Presentation (IS or SP) 55 mn. CP: Contributed Presentation 25 mn

On steady hydromagnetic flow of a radiating viscous fluid through a horizontal channel in a porous medium

APPLICATION OF 1D HYDROMECHANICAL COUPLING IN TOUGH2 TO A DEEP GEOLOGICAL REPOSITORY GLACIATION SCENARIO

Turbulent Natural Convection in an Enclosure with Colliding Boundary Layers

THREE-DIMENSIONAL DOUBLE-DIFFUSIVE NATURAL CONVECTION WITH OPPOSING BUOYANCY EFFECTS IN POROUS ENCLOSURE BY BOUNDARY ELEMENT METHOD

Convection Workshop. Academic Resource Center

ABSTRACT INTRODUCTION

Applications of Partial Differential Equations in Reservoir Simulation

NUMERICAL SIMULATION OF SUSPENDED SEDIMENT TRANSPORT AND DISPERSAL FROM EVROS RIVER INTO THE NORTH AEGEAN SEA, BY THE MECHANISM OF TURBIDITY CURRENTS

Relative Permeability Measurement and Numerical Modeling of Two-Phase Flow Through Variable Aperture Fracture in Granite Under Confining Pressure

Seawater circulation in sediments driven by interactions between seabed topography and fluid density

METHODOLOGY (3) where, x o is the heat source separation and α is the. entrainment coefficient α.

11/22/2010. Groundwater in Unconsolidated Deposits. Alluvial (fluvial) deposits. - consist of gravel, sand, silt and clay

Soils, Hydrogeology, and Aquifer Properties. Philip B. Bedient 2006 Rice University

Natural Convection and Entropy Generation in a Porous Enclosure with Sinusoidal Temperature Variation on the Side Walls

Chapter 1. Introduction to Nonlinear Space Plasma Physics

6. THE BOREHOLE ENVIRONMENT. 6.1 Introduction. 6.2 Overburden Pressures

A HYBRID SEMI-ANALYTICAL AND NUMERICAL METHOD FOR MODELING WELLBORE HEAT TRANSMISSION

APPENDIX Tidally induced groundwater circulation in an unconfined coastal aquifer modeled with a Hele-Shaw cell

FREE CONVECTION AROUND A SLENDER PARABOLOID OF NON- NEWTONIAN FLUID IN A POROUS MEDIUM

NATURAL CONVECTION IN THE BOUNDARY LAYER OF A CONSTANT TEMPERATURE AND CONCENTRATION VERTICAL WALL EMBEDDED IN A DARCY DOUBLY STRATIFIED POROUS MEDIUM

Geophysical Surveys for Groundwater Modelling of Coastal Golf Courses

Transcription:

European Water 55: 31-4, 216. 216 E.W. Publications Solutions of the Elder problem for a wide range of the Rayleigh number J.N.E. Papaspyros *, K.N. Moutsopoulos ** and E. Tzavellas Laboratory of Ecological Engineering and Technology, Department of Environmental Engineering, School of Engineering, Democritus University of Thrace, 12 Vasilissis Sofias Str., 671, Xanthi, Greece * e-mail: ipapaspi@env.duth.gr ** e-mail: kmoutso@env.duth.gr Abstract: Key words: Variations of fluid density often affect the flow field and transport phenomena. Typical examples concerning groundwater flow in porous media include intrusions of seawater, geothermal applications, and subsurface carbon sequestration. A classic benchmark used to test the capability of numerical codes to handle density-driven flow fields is the Elder problem. In the present work, the open-source numerical code OpenGeosys is used for the solution of the Elder problem. The simulations that are presented here include the standard case of Rayleigh number Ra equal to 4, and cases with lower values of Ra. The sensitivity of the steady-state solutions on the grid density is tested. Further, the results of the simulations are compared with solutions published in earlier works. The steady-state solutions of the standard case Ra=4 are found to be dependent on the time step that is used in the simulations. Based on this fact, a complete scan of the range of Ra from 6 to 4 with different time steps of the simulations is undertaken. The results of this scan are checked against the bifurcation diagram that was provided in an earlier study. groundwater flow, density-driven flow, Elder problem, OpenGeosys, stability analysis 1. INTRODUCTION In this work, numerical solutions of buoyancy-driven flow in porous media and aquifers are investigated. This type of flow appears in many problems of practical interest, like fresh-saltwater interactions and saline contamination (Kaleris, 26; Zimmermann et al., 26; Lin et al., 29), and carbon dioxide sequestration (Farajzadeh et al., 27; Hassanzadeh et al., 27). A classic benchmark to test the capability of numerical codes to handle this type of problems is the Elder- Voss-Souza problem which was proposed by Voss and Souza (1987) as a modification of the heat transfer problem examined by Elder (1967). A description of this problem is provided in Figure 1, which depicts the rectangular cross section of an aquifer containing a homogeneous and isotropic porous medium saturated with freshwater (density ρ =1 kg/m 3 ). In the central upper part there is a source of saltwater (density ρ s =12 kg/m 3 ). These conditions correspond to the initial state of the examined problem, or equivalently to the Initial Conditions for t=. For t> mass transfer occurs from the top of the aquifer downwards due to molecular diffusion and natural convection. The pressure in the upper part of the aquifer is considered to be zero (p=). All boundaries are impermeable. The dynamic viscosity of the water is equal to µ=1-3 kg/m/s, the permeability k=4.845x1-13 m 2, the coefficient of molecular diffusion D =3.565x1-6 m 2 /s and the porosity ε=.1. The dispersion process is considered to be independent of the velocity field, so that no hydrodynamic dispersion occurs (Bear, 1979; Koch and Brady, 1986; Moutsopoulos and Koch, 1999). The evolution of the salt concentration c, presented in non-dimensional form (i.e., rescaled from to 1) for the problem described above, will be examined in the present study. As shown in Figure 1, the flow field is symmetric, and in many studies, including the present one, the problem is solved in half domain to save computing resources.

32 J.N.E. Papaspyros et al. Figure 1. Flow field and boundary conditions of the Elder-Voss-Souza problem considered in the present study. Unlike in the problem depicted in Figure 1, in the original problem introduced and studied (both numerically and experimentally) by Elder, density differences were produced by heating the bottom of the field (which was a Hele-Shaw cell in that case). Nevertheless, the two problems are equivalent for the same value of the Rayleigh number (Ra) defined by the relation (Prasad and Simmons, 23): Flux caused by density difference gkβρ ( Δc) H Ra = = (1) Flux caused by diffusion εµ D where g is acceleration due to gravity, Δc is the variation of the non-dimensional density of saltwater and freshwater (in this study Δc=1), H is the aquifer depth (in this study H=15 m), and β is the linear expansion coefficient of fluid density with changing fluid concentration, defined by the relation = 1 β ρ ρ c (in this study β=.2). The rest of the parameters were defined above. Since the Elder and Elder-Voss-Souza problems are equivalent, they will be lumped together hereafter for simplicity under the term Elder problem, and they will be distinguished by the type of scalar. Complementary to Ra the non-dimensional Nusselt number (Nu) is used to assess system behavior of density-driven flows. The Nusselt number is defined by the following relation (Prasad and Simmons, 23): Nu = Total mass flux Mass flux induced by diffusion QH = WL D Δc s (2) where Q is the flow rate of saltwater at the saltwater source, and W and L s are the width and the length of the saltwater source. For the case of Ra=4 considered by Voss and Souza (1987), Elder (1967) and also several other researchers including Frolkovič and De Schepper (2), Doulgeris and Zissis (24), van Reeuwijk et al. (29), concluded that the problem is unstable, so that small disturbances, i.e., small variations in the Initial Conditions and the Boundary Conditions, can result to important changes in the flow and mass transfer behavior. Therefore, distinct solutions can occur for almost the same Boundary and Initial Conditions. Examples of such disturbances are dog barking (laboratory case), and natural heterogeneities in geologic systems or small pressure changes in a distant aquifer location (field case). Correspondingly, it has been proven that for the numerical simulation case, such disturbances can be caused by truncation errors induced by the spatial discretisation processes (e.g., Frolkovič and De Schepper, 2). Interestingly enough, the solutions occurring in physical systems are equivalent to the aforementioned numerical solutions. Although the disturbances or system perturbations mentioned above influence the flow and

European Water 55 (216) 33 mass transfer behavior continuously, the distinct solutions are classified on the basis of the steadystate behavior, e.g., Diersch and Kolditz (22), Doulgeris and Zissis (24), van Reeuwijk et al. (29). For the Elder problem and for Ra=4, until today three distinct forms of the concentration field have been identified and are characterized as S1, S2, and S3, on the basis of the number of fingers (lobe-shaped instabilities) appearing at steady-state. In S1, S2, and S3, one, two, and three fingers, respectively, appear (van Reeuwijk et al., 29). While the S1 concentration field, in which one single finger located at the vertical symmetry axis of the aquifer appears, has been reported for the whole range to 4 of Ra, the S2 and S3 fields appear for Ra>76 and Ra>172, respectively. For Ra<76, the system is diffusion-dominated and a single solution exists. While for the theoretical case of an infinitely extending strip, convection occurs for Ra>4π 2 (Horton and Rogers, 1945), for the Elder problem in which the horizontal extent is finite, convection occurs for Ra> (van Reeuwijk et al., 29). 2. NUMERICAL SIMULATION In the present work, the Elder problem was simulated using the open-source software OpenGeosys (Kolditz et al., 212), while the open-source software Paraview (www.paraview.org) was used for the post-processing of the results. OpenGeosys uses the Finite Element Method (the classical Galerkin method) for the space discretization, and the Finite Difference Method for the time discretization. Non-linear equations are solved by the Picard scheme. The Rockflow software, developed in the mid-eighties by the Institute of Hydromechanics of the University of Hannover, can be considered an earlier version of OpenGeosys. As noted in the Introduction, the symmetry of the problem was taken into account and the simulations were run for the left half domain of Figure 1. The reflection filter Reflect of the Paraview software was used in order to show the results in the entire flow domain. The spatial discretization is typically characterized according to Frolkovič and De Schepper (2), who suggest structured, uniform grids consisting of 2 2l+1 elements for simulations of the half domain. Grids having a value of l equal to at least 6 are typically considered to be reliable. In the present work, grids similar to these used by Ataie-Ashtiani et al. (214) were employed, i.e., grids consisting of 1x5, 2x1 and 4x2 square elements. For the latter most dense grid, the coefficient l takes a value between 7 and 8. Concerning the time discretization, typically non-constant automatic time steps are used to solve the problem examined herein. Nevertheless, some researchers used uniform time steps, for example Frolkovič and De Schepper (2) used a time step Δt=9 d, and Voss and Souza (1987) and Jamshidzadeh et al. (213) used Δt=3 d. In this study, at first, the simulations used the PI-control (Proportional and Integral feedback control) routine of OpenGeosys, which automatically generates time steps of non-uniform size. The time steps of this routine ranged from 1 minute up to 7 days, with a typical average value of 3-4 days. However, as it will be presented later, it was necessary to also perform simulations by using time steps of uniform size, with Δt varying between 5 d and 6 d, because this parameter was found to influence the form of the concentration field, and depending on its choice S1, S2, or S3 steady-state concentration fields can appear. 3. RESULTS AND DISCUSSION According to van Reeuwijk et al. (29), a successful simulation software for the Elder problem should fulfil the following requirements: 1. Agreement of the numerical results with an existing semi-analytical solution for the pure diffusion problem (Ra=). 2. Reproduction of the unique steady-state solution and the transients leading to it, for the Ra=6 case. 3. Reproduction of the S1 and S2 concentration fields for the Elder problem (case Ra=4).

34 J.N.E. Papaspyros et al. As it will be shown below, the OpenGeosys software fulfils all these requirements. 3.1 Ra=4 Most studies of the Elder problem use a simulation time period of 2 years, assuming that the concentration field has reached its steady-state at this point. In Figure 2, the concentration fields obtained by OpenGeosys for a simulation time period of 2 years, for the three grids mentioned in the previous section (i.e., grids consisting of 1x5, 2x1, and 4x2 square elements), and for both the case of a non-constant time step determined by the PI-control and the case of a constant time step Δt=3 d, are presented. Figure 2. The concentration field at t=2 years, for the three grids used in this study, and for time steps either determined by the PI-control procedure (a, b, c) or constant Δt=3 d (d, e, f). Continuous lines are concentration contours between c=.1 and c=1 with a step of.1. For the former case of PI-controlled time step and for all grids, S1 concentration fields were obtained. Nevertheless, in the study of Ataie-Ashtiani et al. (214), who used the same grid resolution with the present work but performed simulations using FEFLOW and not OpenGeosys, a S2 field was obtained for the coarse grid case, while for the denser grids, S1 profiles were obtained (see their Figure 2). This result indicates that unlike in the case of FEFLOW, for simulations performed with OpenGeosys, the concentration field is not sensitive to the spatial discretization. Nevertheless, as it is depicted in Figure 2, the form of the concentration field depends on the time discretization: using a constant time step Δt=3 d, and not the PI-control routine, S2-type and not S1-type concentration fields were obtained. Figure 3. The concentration field at t=2 years for uniform time step (a) Δt=1 d, (b) Δt=6 d. Continuous lines are concentration contours between c=.1 and c=1 with a step of.1. To further investigate the influence of the time discretization, a series of simulations of time period equal to 2 years for Ra=4 with the dense grid 4x2 and constant time steps of Δt=5 d, 1 d, 3 d and 6 d were performed. For the first three time steps, the same concentration field of S2-type was obtained, but for Δt=6 d the resulting concentration field was of the S1-type. These two forms of the concentration field are depicted in Figure 3 for Δt=1 d and Δt=6 d. This result is

European Water 55 (216) 35 considered to be important: to the best of the present authors knowledge, the dependency of the type of the resulting concentration field on the time step used has not been reported in earlier simulations of the Elder problem. Accordingly, a complete scan of the range 6 to 4 of Ra with different values of constant time steps was undertaken, and it will be presented in Section 4 of the present work. In Figure 4, the S1-type concentration fields obtained from OpenGeosys in the present work for simulation time periods of 1 and 2 years are compared with both the experimental data of Elder (1967) and the numerical results of Voss and Souza (1987), which were computed using the SUTRA software (obviously, the data of Elder are inverted in order to be compared with the concentration fields). The agreement of the results of the present study with both reference data sets is considered satisfactory. Interestingly enough, as it is depicted in Fig. 4a, three fingers may occur at early times for a S1-type profile. Figure 4. Validation of the S1-type scalar fields obtained in the present work: the experimental results of Elder (small white dots) are compared with the numerical results of Voss and Souza (1987) (large black dots), and with the results of OpenGeosys (continuous lines), for simulation time periods (a) t=1 years and (b) t=2 years. Lines and dots are concentration contours c=.2 and c=.6. Concerning the case of S2-type concentration fields, the results of the present work for simulation time periods of 1 and 2 years are compared with the numerical results of Prasad and Simmons (23), and Jamshidzadeh et al. (213) in Figure 5. Again the agreement of the OpenGeosys results with both reference data sets is considered satisfactory. Figure 5. Validation of the S2-type scalar fields obtained in the present work: the numerical results of Prasad and Simmons (23) (small white dots) are compared with the numerical results of Jamshidzadeh et al. (213) (large black dots), and with the results of OpenGeosys (continuous lines), for simulation time periods (a) t=1 years and (b) t=2 years. Lines and dots are concentration contours c=.2 and c=.6. In conclusion, for Ra=4 OpenGeosys delivered concentration fields of both S1-type and S2- type, depending on the time step used. Both types of concentration field were successfully validated against the results of earlier simulations. 3.2 Ra= A semi-analytical solution of the Elder problem in the Laplace space for Ra= (pure diffusion case) has been developed by van Reeuwijk et al. (29). This solution was inverted during the present study to the real time space by the Stehfest algorithm (Stehfest, 197) and it was compared with the corresponding steady-state OpenGeosys solution, which was obtained by setting the permeability k= in Eq. (1). As depicted in Figure 6, the agreement between the numerical and the semi-analytical concentration fields is satisfactory.

36 J.N.E. Papaspyros et al. Figure 6. Comparison of the steady-state concentration field obtained by OpenGeosys for Ra= (left) with the corresponding semi-analytical solution (right). Continuous lines are concentration contours between c=.1 and c=1 with a step of.1. 3.3 Ra=6 According to van Reeuwijk et al. (29), in the case of Ra<76 the Elder problem has a unique solution, which has the form of a S1-type concentration field. The steady-state of the case Ra=6 was simulated by OpenGeosys in the present study, again by setting the permeability k to the proper value in Eq. (1). The concentration field obtained by OpenGeosys is compared in Figure 7 with the corresponding field that was numerically obtained by van Reeuwijk et al. (29) by using a gridindependent pseudospectral code. The agreement of the two concentration fields is considered satisfactory. Figure 7. Concentration field for Ra=6: numerical results obtained by van Reeuwijk et al. (29) (white dots) and by OpenGeosys (continuous lines). Lines and dots are concentration contours between c=.1 and c=1 with a step of.1. 3.4 Temporal evolution of the Nusselt number for the Ra=4 case As it has been demonstrated above, OpenGeosys fulfils all the criteria imposed by van Reeuwijk et al. (29) for the Elder problem. In an additional validation check, the Nusselt number (defined by Eq. 2) was evaluated from the concentration fields obtained by OpenGeosys for several simulation time periods. The values of Nu are compared in Figure 8 with the experimental results of Elder (1967) and with the numerical results of Jamshidzadeh et al. (213). The results of the present study are remarkably close to the experimental results of Elder (1967), while they also agree with the results of Jamshidzadeh et al. (213). Concerning the points that are marked by letters in Figure 8, it has been suggested by Prasad and Simmons (23) that between points A and B a boundary layer is developed near the scalar source. At point C, the plume has reached the bottom boundary, while at point D, the denser part of the plume is in the lower part of the domain. In the region between points D and E, diffusion prevails over mass advection.

European Water 55 (216) 37 2 18 16 A C Elder Jamshidzadeh et al. Present work Nu 14 12 D 1 8 6 E B 2 4 6 8 1 12 14 16 18 2 time (years) Figure 8. Evolution of the Nusselt number in simulation time for Ra=4: comparison of the results of OpenGeosys with those of Elder (1967) and Jamshidzadeh et al. (213). 4. STUDY OF BIFURCATION As noted in Section 3.1, the dependency of the form of the concentration field on the simulation time step is considered an important result of the present study. Accordingly, a complete scan of the range of Ra from 6 to 4 for different constant time steps was undertaken, in an attempt to better characterize this dependency. Since the results of OpenGeosys were found to be independent from the grid density, at least for the three grids used in the present study, the grid of middle density (2x1 square elements) was used here, to save computer time. The range of Ra from 6 to 4 was covered with a step of 2 (i.e., the simulations were run for Ra=6, then Ra=8, then Ra=1, and so on up to Ra=4). For every simulation, the desired value of Ra was obtained by properly setting the permeability k in Equation 1. The simulations were repeated for three values of constant time step Δt=1 d, Δt=3 d, and Δt=6 d. The work that is presented in this section is actually a repeat of the pioneering work of van Reeuwijk et al. (29), who used a grid-independent pseudospectral method to construct the bifurcation diagram of the three stable steady-state forms S1, S2, and S3 of the concentration field in the Ra range -4. These authors used different initial conditions to obtain the three forms S1, S2, and S3 at Ra=4 and then used the obtained steady-state concentration fields as initial conditions for lower Ra, and so on, thereby determining a continuous line for every form in the bifurcation diagram. On the contrary, in the present work all simulations were run with the same initial conditions that were described in the Introduction, and the forms S1, S2, and S3 were obtained by changing the time step Δt of the simulations. An early result of the study that is described in this section was the conclusion that a simulation time period of 2 years is too small to ensure steady-state behavior. Simulation time periods of at least 6 years are mandatory for meaningful steady-state results. There were even cases where the two simulation time periods (2 and 6 years) resulted in different forms of the concentration field. Accordingly, all the simulations reported here had a time period of 6 years. According to van Reeuwijk et al. (29), it is very difficult to obtain the form S3 of the concentration field. Indeed, these authors could not find a set of initial conditions that would result to S3, and they concluded that the basin of attraction of S3 is small, with only a small subset of initial conditions leading to S3. Interestingly, in the present work the form S3 was obtained for

38 J.N.E. Papaspyros et al. Δt=1 d at Ra=28 and Ra=3, as depicted in Figure 9. This could provide some information about the conditions of appearance of the form S3. Figure 9. The form S3 of the steady-state concentration field, obtained for Δt=1 d at Ra=3. Continuous lines are concentration contours between c=.1 and c=1 with a step of.1. A contour map of the form of the concentration field as a function of Ra and the time step Δt of the simulation is provided in Figure 1. This map is based on a grid that was constructed from the triplets (Ra, Δt, and resulting form of the concentration field) of the simulations. The grid was constructed with the nearest neighbor gridding method. The most interesting range of this grid appears to be around Ra=3, where every possible stable steady-state form of the concentration field can be obtained with a proper value of Δt. Figure 1. Contour map of the form of the steady-state concentration field as a function of Ra and Δt. As suggested by van Reeuwijk et al. (29), a good indicator of the system studied is the vertical solute flux at the top wall, which is expressed as a local Sherwood number Sh x, defined by: dc Sh x ( x) = (3) dz z= 1 Sh x could be integrated along the top wall to provide the average Sherwood number Sh: 2 1 Sh = 2 Shx dx (4) Notice that this definition of Sh is valid for normalized dimensions of the flow field, i.e., when z is ranged from to 1 and x from to 4. The definition of Sh should be transformed properly when applied to non-normalized coordinates, like those in Figure 1. The suggestion of van Reeuwijk et al. (29) was verified in the present study, as depicted in Figure 11, where the profiles of the local Sherwood number Sh x along the top wall are plotted for the three forms of the steady-state concentration field that were obtained at Ra=3. Every form of the concentration field appears to have its own characteristic profile of Sh x which corresponds to a characteristic integral Sh along the top wall. In conclusion, Sherwood number is an efficient indicator of the form of the concentration fields, and could be used as the vertical axis in the bifurcation diagram, like in van Reeuwijk et al. (29).

European Water 55 (216) 39 The study of bifurcation that is described in this section is concluded with a check of the results of the present work against the bifurcation diagram that is provided by van Reeuwijk et al. (29). This check is depicted in Figure 12, where the lines of S1, S2, and S3 were digitized from Figure 6 of van Reeuwijk et al. (29) and are plotted together with the pairs (Ra, Sh) of the simulations of the present study, which are grouped by the form of the steady-state concentration field. The agreement is striking, to the point that it could be suggested that future tests of the capability of numerical codes to handle density-driven flow fields which solve the Elder problem should include a check of agreement with the bifurcation diagram of van Reeuwijk et al. (29). It is stressed that the present work was able to reproduce the three stable steady-state forms of the concentration field simply by using different time steps, while all simulations started from the classic initial conditions of the Elder problem. This could provide information on the mechanism of appearance of the different concentration fields. -1 Sh x -2 S1 S2 S3-3 1 2 3 4 5 6 x (m) Figure 11. Profiles of the local Sherwood number Sh x along the top wall of the studied flow field for the three forms of the steady-state concentration field that were obtained at Ra=3. 5 4 Sh 3 2 1 S1 S2 S3 S1 S2 S3 5 1 15 2 25 3 35 4 Ra Figure 12. Check of the results of the present study against the bifurcation diagram of van Reeuwijk et al. (29).

4 J.N.E. Papaspyros et al. 5. CONCLUSIONS The Elder problem was investigated by performing simulations with the open-source OpenGeosys software. It was demonstrated that OpenGeosys can reliably simulate density driven flows. Further, the results of the simulations of the present study indicate that in the unstable flow regimes, truncation errors induced by temporal discretisation can strongly influence the flow regimes and produce different solutions. A phase diagram describing the dependence of the different solutions on the time step of the simulations was provided. The results of the present study were found to agree with the bifurcation diagram of van Reeuwijk et al. (29). To the best of the present authors knowledge, the fact that the distinct forms of the steady-state concentration fields of the Elder problem can be produced by choosing different time steps has not been reported before, and it should be confirmed by further investigations. ACKNOWLEDGEMENT An initial shorter version of the paper has been presented in Greek at the 3 rd Common Conference (13 th of Hellenic Hydrotechnical Association, 9 th of Hellenic Committee on Water Resources Management and 1 st of the Hellenic Water Association) "Integrated Water Resources Management in the New Era", Athens, Greece, December 1-12, 215. REFERENCES Ataie-Ashtiani, Β., Simmons, C.T., Werner, A.D., 214. Influence of boundary condition types on unstable density-dependent flow. Groundwater; 52(3): 378-387. Bear, J., 1979. Groundwater Hydraulics. McGraw-Hill, New York, 567p. Diersch, H.J.G., Kolditz, O., 22. Variable-density flow and transport in porous media: approaches and challenges. Advances in Water Resources; 25(8 12); 899 944. Doulgeris, C., Zissis, T., 24. Simulation of variable density flow and solute transport in porous media. Hydrotechnika; 14(1): 3-15. (in Greek). Elder, J. W., 1967. Transient convection in a porous medium. Journal of Fluid Mechanics; 27(3); 69-623. Farajzadeh, R., Salimi, H., Zitha, P. L., Bruining, H., 27. Numerical simulation of density-driven natural convection in porous media with application for CO 2 injection projects. International Journal of Heat and Mass Transfer; 5(25): 554-564. Frolkovič, P., De Schepper, H., 2. Numerical modelling of convection dominated transport coupled with density driven flow in porous media. Advances in Water Resources; 24(1): 63-72. Hassanzadeh, H., Pooladi-Darvish, M., Keith, D. W., 27. Scaling behavior of convective mixing, with application to geological storage of CO 2. AIChE journal; 53(5): 1121-1131. Horton, C. W., Rogers Jr, F. T., 1945. Convection currents in a porous medium. Journal of Applied Physics; 16(6): 367-37. Jamshidzadeh, Z., Tsai, F.T.-C., Mirbagheri, S.A., Ghasemzadeh, H., 213. Fluid dispersion effects on density-driven thermohaline flow and transport in porous media. Advances in Water Resources; 61: 12-28. Kaleris, V., 26. Submarine groundwater discharge: effects of hydrogeology and of near shore surface water bodies. Journal of Hydrology; 325(1): 96-117. Koch, D. L., Brady, J. F., 1985. Dispersion in fixed beds. Journal of Fluid Mechanics; 154: 399-427. Kolditz, O., 21. Non-linear flow in fractured rock. International Journal of Numerical Methods for Heat and Fluid Flow; 11(5-6): 547-575. Kolditz, O., Bauer, S., Bilke, L., Böttcher, N., Delfs, J. O., Fischer, T., Park, C. H., 212. OpenGeoSys: an open-source initiative for numerical simulation of thermo-hydro-mechanical/chemical (THM/C) processes in porous media. Environmental Earth Sciences; 67(2): 589-599. Lin, J., Snodsmith, J. B., Zheng, C., Wu, J., 29. A modeling study of seawater intrusion in Alabama Gulf Coast, USA. Environmental Geology; 57(1): 119-13. Moutsopoulos, K. N., Koch, D. L., 1999. Hydrodynamic and boundary-layer dispersion in bidisperse porous media. Journal of Fluid Mechanics; 385: 359-379. Prasad, A., Simmons, C.T., 23. Unstable density-driven flow in heterogeneous porous media: a stochastic study of the Elder (1967b) short-heater problem. Water Resources Research; 39(1): 17-128. Stehfest, H., 197. Algorithm 368: Numerical inversion of Laplace transforms [D5]. Communications of the ACM; 13(1), 47-49. van Reeuwijk, M., Mathias, S.A., Simmons, C.T., Ward, J.D., 29. Insights from a pseudospectral approach to the Elder problem. Water Resources Research; 45(4): 1-13. Voss, C. I., Souza, W.R., 1987. Variable density flow and solute transport simulation of regional aquifers containing a narrow freshwater-saltwater transition zone. Water Resources Research; 23(1): 1851-1866. Zimmermann, S., Bauer, P., Held, R., Kinzelbach, W., Walther, J. H., 26. Salt transport on islands in the Okavango Delta: numerical investigations. Advances in Water Resources; 29(1): 11-29.