Numerical simulation of three-dimensional saltwater-freshwater fingering instabilities observed in a porous medium Klaus Johannsen a,∗, Sascha Oswald b, Rudolf Held c, Wolfgang Kinzelbach d, a IWR/Technical b Department c Statoil d Institute

Simulation, University of Heidelberg, Im Neuenheimer Feld 368, 69120 Heidelberg, Germany

of Hydrogeology, UFZ Leipzig-Halle, D-04138 Leipzig, Germany

ASA, Arkitekt Ebbellsvei 10, Rotvoll, N-7005 Trondheim, Norway f¨ ur Hydromechanik und Wasserwirtschaft (IHW), ETH H¨ onggerberg, CH-8093 Z¨ urich, Switzerland

Abstract We consider saltwater-freshwater fingering instabilities in a saturated porous medium. In the first part, we present three-dimensional results obtained from a laboratory experiment using non-invasive imaging. In the second part, we define a set of model problems in which the performed laboratory experiments can be ranged in. Due to its highly nonlinear behavior and inevitable modeling errors, a detailed numerical reproduction of the physical concentration measurements cannot be expected. Nevertheless, four criteria have been identified, two quantitative and two qualitative, which facilitate a substantiated comparison of the physical experiment and the numerical simulation. With respect to these criteria a high degree of similarity could be observed. The use of these features allows a deeper understanding of the physical processes and the influence of the initial conditions. Key words: porous media, variable-density flow, saltwater, instability, multigrid methods, non-invasive imaging, NMR

∗ Corresponding author. Tel.: +49 6221 548828; fax: +49 6221 548860 Email addresses: [email protected] (Klaus Johannsen), [email protected] (Sascha Oswald), [email protected] (Rudolf Held), [email protected] (Wolfgang Kinzelbach).

Preprint submitted to Elsevier Science

18 December 2005

1

Introduction

In groundwater environments, fluid density and viscosity are generally variables related to the concentration of solutes, temperature, or pressure conditions. Both fluid properties may alter the flow patterns and prevailing flow regimes. For variable density flow conditions, this becomes evident in a number of field situations. Among those are convective flows in geothermal systems, flows around salt domes, leaching and disposal of dense solutes or brines, or saltwater fingering below playa lakes and salt pans. These situations have been subject to long-standing investigation and numerical analysis. The variety of findings and current challenges are stated in, e.g., [45, 46], whereas a recent review of the literature on the numerical treatment is given in [11]. We refer to these reviews for details and references. Due to the nonlinear coupling of the equations for density-dependent flow and solute transport, their solution is more complex than for the case of passive tracer transport. Density-dependent flows are rotational and hydrodynamic instability leading to a non-uniqueness of the solution is inherent in the mathematical and physical models of such systems. The criterion for instability that is investigated theoretically, experimentally and numerically is typically expressed in form of a Rayleigh number Ra. This dimensionless ratio of buoyancy to diffusion and dispersion is derived under a steady-state flow assumption. The experimental investigation of gravitationally unstable flow configurations, i.e. light fluid underlying a heavier fluid, mainly employed pseudo-two-dimensional Hele-Shaw setups to study the thermal instability in Rayleigh-B´enard convection [13, 43, 54], the Rayleigh-Taylor instability in porous media [14, 40] or instability induced below an evaporation boundary [55, 56]. The number of experiments in real porous media in three-dimensional setups (as opposed to Hele-Shaw cells) is very limited to date [6, 29, 31, 42]. This paper presents a fully three-dimensional data set of solute concentrations during the process of purely density-driven miscible fingering in a benchmark scale setup. The experiment was designed especially for subsequent numerical simulation, providing all required parameters. High-accuracy, three-dimensional numerical studies were carried out to investigate the mechanisms affecting hydrodynamic instabilities in miscible displacement for well defined geometries, e.g. resolving the gap of Hele-Shaw cells [14, 27], for rectilinear geometries [49, 57] or for quarter five-spot configurations [38]. Difficulties in the numerical handling were found to be closely connected to the geometries of the domain, the spatial variability of the base flow and the presence of line sources and sinks, cf. [28]. The numerical issues concerned the inaccurate representation of velocity variations and grid2

orientation effects that superimpose numerical artifacts on the model predictions. This can even lead to unphysical numerical results, particularly in highly nonlinear systems. We provide a numerical simulation approach that is able to reproduce the physical experiment with sufficient numerical accuracy. Notably, such detailed simulations of miscible fingering are to date rare in three dimensions. The early analytical work [9,21,40] treated instability of miscible fluid displacement in porous media with a first-order perturbation approach. The motion of two incompressible fluids separated by a sharp interface in a porous medium was based on Hele-Shaw equations, or Darcy’s law. The high sensitivity of the problem to perturbations is caused by singularity in the governing equations in the absence of surface tensions or dispersive mechanisms. In miscible displacements in realistic porous media, flow-induced dispersion plays a dominant role by widening the front and thereby yielding a cut-off length scale for the potential growth of instabilities [36]. The cross-over between diffusive and linear growth regimes was analyzed by, e.g., [48, 53]. Subsequently, [10] or [2] proposed an extended theory for the prediction of a stability criterion for miscible displacement in porous media taking a velocity-dependent dispersion tensor into consideration. In a recent treatment, the linear stability analysis was performed on the basis of the three-dimensional Stokes equations [16], thus accounting for Taylor dispersion at the interface due to gradients in the velocity field. This enabled the prediction of a wavelength selection for HeleShaw configurations across high Ra and low Ra regimes. Linear stability analysis of dominant modes that evolve from interface perturbations considers the exponential growth regime of the characteristic initial wavelengths. Such analysis applies strictly to the short time scales. A weakly nonlinear analysis of the formation of finger patterns was very recently presented by [1, 8] for dynamical system conditions. A formal gradient expansion of the nonlocal problem allowed the exact determination of a sub-critical bifurcation for the Saffman-Taylor instability. Such dynamic analytical analysis is beyond the scope of this contribution. Our setup also included a marked heterogeneity in the porous medium, a situation which has not been fully incorporated in stability analyses. Here, we analyze a reproducible setting of dynamically evolving finger instabilities by comparing computations in analogy to a three-dimensional, physical pattern evolution. Previously, we presented comparisons between an experimental investigation and numerical simulation for stable density-dependent flow configurations, or upconing conditions [25, 33]. The study provided a mathematical benchmark for numerical modeling of density-dependent flow, in a three-dimensional setting and for the case of strong density coupling. The nonlinearity in the density unstable model system can be significantly higher, and poses in itself numerical challenges. Benchmarking issues for unstable convection in two-dimensional 3

configurations have been addressed more recently by [52] under steady-state conditions in finite and infinite, bound horizontal or inclined layers, or by [47] for transient conditions for the two-dimensional salt lake problem with an unstable saline boundary layer formed by evaporation. In the latter test case, the hydrodynamic instability proved to be far more sensitive to the initial perturbations than for the purely diffusive regime studied in [52]. The conditions pertaining in field problems necessitate the further development and testing of numerical tools for three-dimensional, transient instabilities. In this contribution we examine saltwater-freshwater fingering in three dimensions in a porous medium. We present, for the first time to our knowledge, a fully threedimensional, high-quality data set of solute concentrations during the process of purely density driven miscible fingering in a porous medium for a bench scale set-up. Moreover, this also comprises the effects of a permeability heterogeneity and was designed especially for subsequent numerical simulation, thus delivering all the required parameters. We provide a numerical simulation approach that is able to represent the situation and processes of the physical experiment with a reasonable numerical accuracy. Notably, such detailed simulations of miscible fingering are quite common in 2D, but are so far very rare in 3D. We develop a framework for comparing experimental and numerical results in a meaningful way. This is not trivial for such a highly nonlinear behaviour, which strongly amplifies perturbations, and cannot be performed as a direct comparison on a point-to-point basis as is typical for other situations. Finally, we demonstrate the capability of the numerical model to capture the essential behaviour of the physical process to a high degree, and then to look at the influence of initial conditions, which could not be done so easily in a physical experiment.

The remainder of the paper is organized as follows. In section 2, we give a description of the experimental setup and the magnetic resonance technique used to obtain the experimental results. In section 3, we define a reproducible set of mathematical model problems, with boundary and initial conditions that enclose the experimental setting. Issues concerning the numerical discretization and solution method are addressed in section 4. In section 5, we present the numerical results and a comparison of experimental and numerical finger dynamics. We show the influence of the initial conditions on the flow modes. A summary of the findings and insights is given in section 6. 4

2

Fingering experiment

2.1 Experimental methods and procedures Density-driven fingering in a saturated, artificial porous medium was observed by magnetic resonance imaging (MRI) in a laboratory experiment. As an initial situation a saltwater layer above a freshwater layer was created in a closed box, which subsequently started to show density fingering of salt water as well as freshwater. The objective was to study this unstable situation of flow solely driven by density gradients. The MRI method allowed the retrieval of three-dimensional, time-dependent information about the evolution of the salt concentration distribution during the experiments. The technique of magnetic resonance imaging has become widespread for medical applications, but has also been applied during the last decade to observe flow and transport behavior in porous media. For example, [35] observed the convection and fingering driven by density forces on a cm scale for homogeneous conditions, [20, 22] measured the influence of velocity variability in a sand column during advective transport using high-contrast agent concentrations, and a three-dimensional test case for stable variable-density flow in porous media was recently proposed based on MRI measurements [31, 33]. A similar method has been used in this study to create a three-dimensional data set for a non-stable situation on the benchscale, providing a time-dependent concentration distribution for miscible fingering by internal density forces in and around an inclusion of higher permeability. With MRI, a three-dimensional volume can be scanned completely in a series of 2D slices consisting of cuboidal volumes (voxels) in rows and columns. For the apparatus used in this study, the measured signal originated from water protons and their local environment. One approach in MRI applications that yields an enhanced signal contrast is to use a contrast agent [7], i.e. paramagnetic ions that change the relaxation times of the water protons in their vicinity. Here Cu2+ ions were used. A brief introduction to MRI using contrast agents is given in, for example, [7, 50]. 2.1.1 Experimental setup The experiment was carried out in a cubic Plexiglass tank containing an artificial porous medium saturated with water. Flow could be induced by injection and discharge through small ports in the bottom centre (O0 ) and the top corners of the tank (O1 -O4 ), which were used for the preparation of a layering of salt water below freshwater. The container could be rotated around a horizontal axis through the volume centre of the porous medium. This allowed the 5

subsequent preparation of an unstable density-layering. See Fig. 1, left, for a sketch of the experimental setup. The inner dimensions of the cubic container were L = 0.2m, and thus the bulk volume of the porous medium was L3 = 8·10−3m3 . This volume contained two zones, each of which was composed of almost uniform glass beads (see Fig. 1, right). The central zone, henceforth called inclusion, was a vertical, quadratic column of beads of average diameter d1 = 1.2mm whereas the rest of the container was filled with beads of average diameter d2 = 0.7mm. The inclusion of larger sized beads was three times more permeable than the ‘background’ (see Table 1 for details).

0 40 4

saltwater

mm

B0 O0

O2

O4 4

O3 200 mm

Saltwater with NMR-Tracer

O1

O0

O1

O4

B0 O2

O3

ca. 2 m

200 mm

mm 0

20

freshwater

Fig. 1. Sketch of the experimental setup. Left: at the beginning of the MRI measurements (preparatory phase); right: a schematic close-up after turning the container over (beginning of main phase).

2.1.2 Procedure and boundary conditions The experiment consisted of two preparatory phases, followed by the main phase of the fingering process. During all three phases, MRI images were taken. (1) To begin with, the porous medium was filled with freshwater. Then a salt solution with a salt mass fraction of 0.30% of NaCl and MRI-tracer was injected at the centre of the bottom of the porous medium. To allow an outflow of water from the porous medium the four top outlets were opened and kept at a fixed hydraulic head. A saltwater dome formed at the bottom following the injection of salt water. Due to the higher permeability in the inclusion, the salt water moved further up in the central part of the porous medium, in and around the inclusion, than in the outer zone. Then all ports were closed to end this phase. 6

(2) During the second phase, no external flow or driving force was applied, and the saltwater dome could settle downwards. The aim of this phase was to prepare a saltwater interface that was more or less horizontal, equilibrated by internal density-driven flow. In spite of a duration of about 150 minutes, the saltwater-freshwater interface did not settle fully to a perfectly horizontal position. Also, the interface was significantly wider close to the corners than in the centre of the container. In the inclusion itself, the saltwater/freshwater interface was a few millimeters further up, and more dispersed, than in the surrounding of the inclusion. (3) A rotation of the container by 180 degrees started the third and main phase of the experiment. This rotation was carefully performed within a few seconds. After this, there was still no external flow or driving force, and an internal density-driven fingering took place. The unstable layering of salt water above freshwater was not an ideal layering, because the saltwater-freshwater interface was not perfectly horizontal, and the transition zone had a detectable extension, especially close to the corners. But the salt-concentration distribution showed almost the same symmetry as the setup, and the interface was not perturbed by rotating it to the unstable configuration. The parameters that characterize the porous medium and the freshwater properties were determined as already reported in detail in [31,33], and the relevant parameters are listed in Table 1. The parameters not yet mentioned are the coefficient of molecular diffusion Dm , the porosity ni , the permeability ki , the dispersion lengths αl,i , αt,i , the density of fresh- resp. salt water ρf , ρs and the viscosity of fresh- resp. salt water µf , µs . Finally, ~g denotes the vector of gravitational acceleration. The index i takes the value 1 for the central part (inclusion) and 2 for the outer part (background) of the domain.

2.1.3 MRI technique This study was conducted using the spin echo technique. The experiments were carried out on a Philips 1.5 Tesla MR whole body system (Philips Gyroscan ACS/NT, Best, The Netherlands) at the university hospital in Z¨ urich, Switzerland, based on a proton spin frequency of 64 MHz. The echo and repetition times were adapted to the experimental conditions of this setup (cf. [34]). The porous medium was positioned horizontally in the MRI apparatus in such a way that one horizontal axis of the container was parallel to the axis of the apparatus. MRI images (matrix size 128 × 128; field of view, F OV = 320mm, reduced FOV 65%) were acquired from 50 adjacent vertical slices of 4mm thickness covering one horizontal length and the vertical length of the cubic container. The voxels of the slice were 2.5mm × 2.5mm in the vertical 7

Table 1 Parameters of experiment and model problem; index 1 denotes the central part (inclusion), index 2 the outer part of the porous medium (background) Parameter

Physical experiment

Fingering

units

min.

max.

model problem

0.29%

0.31%

0.30%



7126

s

ωmax

0.30%

T

241 + 7126

L

200

199

201

200

10−3

m

Dm

10

7

12

10

10−10

m2 s−1

d1

1.2

1.0

1.3



10−3

m

d2

0.7

0.6

0.8



10−3

m

n1

0.392

0.342

0.442

0.392



n2

0.385

0.375

0.395

0.385



k1

10

8.9

13

10

10−10

m2

k2

3.3

3.0

4.1

3.3

10−10

m2

αl,1

1.2

0.6

1.5

1.2

10−3

m

αt,1

0.12

0.03

0.25

0.12

10−3

m

αl,2

0.7

0.3

0.9

0.7

10−3

m

αt,2

0.07

0.01

0.15

0.07

10−3

m

ρs /ρf − 1

2.9

2.1

3.7

2.9

10−3



µf

1.002

0.93

1.02

1.002

10−3

kgm−1 s−1

µs

1.0075

0.94

1.03

1.0075

10−3

kgm−1 s−1

|~g |

9.8065

9.8064

9.8066

9.8065

ms −2

cross-section, and thus the horizontal resolution was 4mm in the y-direction and 2.5mm in x-direction, whereas it was 2.5mm in the vertical z-direction. The imaging via the two-dimensional spin-echo sequence used at least two excitations per image. The salt concentration was observed over a period of more than four hours in total with a maximum temporal resolution of about 3 minutes. During data processing, the outermost rows and columns of the slices were removed if they were located outside the porous medium. For a more detailed description of the specific MRI principles applied, we refer to [32–34]. The mass fraction of Cu2+ was chosen to be in the linear range of the calibration curve with good contrast properties, in accordance with former studies [33, 34]. These mass fractions were topped up with N aCl, summing up to a total of 0.30%, to obtain a salt water with higher density. The resulting den8

Sfrag replacements

sity gradient to freshwater initiated a fingering regime, quick enough to reach the bottom of the container during the time available for the measurement. The MRI signal also represents the salt concentration, because the Cu2+ ions do not separate significantly from the N a+ ions for this setup, as long as the Cu2+ ions do not sorb. The total salt mass fraction is then proportional to the mass fraction of Cu2+ and therefore to the measured signal intensity. In the MRI experiments, the mean signal of freshwater was determined from the image before the experiment was started, representing ω = 0% salt mass fraction. Furthermore, the mean signal of the salt solution, containing 10mmol/L CuSO4 , and ωmax = 0.30% salt mass fraction in total, was evaluated in a region of the porous medium with undiluted salt water. Then a linear calibration relation between the total salt concentrations and the measured signal intensity was used for the interpretation of the images during the experiment (Figure 2). 0.30

ω [%]

Conc. [%]

0.25 0.20 0.15 0.10

B ackground porous medium

0.05

Inclusion

0.00 0

100

200 300 Signal amplitude

400

Signal amplitude

Fig. 2. Measured signal of salt water and freshwater in the two types of porous medium, and linear function used for calibration in each region. The value representing maximum salt mass fraction in the insertion has been adjusted to account for outliers.

The signal to noise ratio is quite low in the images. The noise level of the images for a given type of MRI method can be reduced by decreasing the spatial or temporal resolution, or by increasing the porosity. Since the porosity is given by the porous medium, a compromise between resolution and noise level had to be found, which was chosen mainly in favor of achieving a higher spatial and temporal resolution. A Gaussian filter was applied twice to reduce the effect of noise on the MRI images. This smoothes the signal or concentration and thus yields more realistic isosurfaces or isolines. The possible disadvantage of this smoothing, i.e. a smearing of the concentration gradient at the salt-waterfreshwater interface, proved to be of minor impact here. The values of the salt mass fraction could be determined with reasonable accuracy between 0.03% 9

2.2 Experimental results

ω/ωmax

200

150

Z

100

0.90 0.80 0.70 0.60 0.50 0.40 0.30 0.20 0.10

200 50

X

Sfrag replacements

and close to 0.30%, which covers the transition zone between freshwater and salt water during the fingering process.

100 0 0

150

100

50

0

Y

Fig. 3. 3D view of the vertical cross-sections used for comparison, i.e. a vertical diagonal and a vertical transverse cross-section of the cubic container. The color bar indicates the values of the salt mass fraction in percent.

Images are presented here in vertical cross-sections of the measured salt mass fraction for a particular time. These can be diagonal cross-sections from one edge of the container to the other, or transverse cross-sections at the centre of the porous medium, as illustrated in Fig. 3. Nevertheless, from the measured data any cross-section or any isosurface could be plotted in three dimensions for each measurement time. For a clear perception of the fingering and to allow a straightforward comparison to simulation results, we present some three-dimensional views and movies in the supplementary material [30]. The full data set is used for the quantitative evaluation of characteristic properties in section 5. Fig. 4 shows the salt mass fraction on vertical cross-sections (diagonal and transverse) at four different times (t = 241, 241 + 2549, 241 + 4883 and 241 + 7126s) from the beginning of the third phase. An offset of 241s is used explicitly in anticipation of the numerical simulations. Their initial conditions were set at t = 241s, defining the start of the simulated time. As can be seen from Fig. 4, the movement of salt water and freshwater occurs mainly 10

t = 241 s

t = 241 + 2549 s

t = 241 + 4883 s

t = 241 + 7126 s

Fig. 4. Salt mass fractions measured during the main phase of the MRI experiment, shown in a vertical, diagonal cross-section (left) and a vertical transverse cross-section (right). Twofold Gaussian smoothing has been applied to reduce noise to show the concentrations. Times are taken relative to the beginning of the third phase. 11

at two types of location, one being the inclusion of higher permeability, the other being the vertical edges of the container, where the saltwater-freshwater transition zone is higher up than in the centre. The salt water starts to finger downwards in the inclusion at one side of the square cross-section. At the same time, a freshwater finger starts moving upward on the opposite site. In the background porous medium just at the border of the inclusion, and at the origin of the saltwater fingers, the freshwater dents into the saltwater body (Fig. 4). These small freshwater fingers constrict the saltwater finger in its width, without being capable of creating additional freshwater fingers. At later times, instabilities also develop along the interface outside of the more permeable inclusion.

3

Definition of model problems

In this section, we define a set of model problems related to the physical experiment described in the previous section. Due to parameter uncertainties and the inherent instabilities in the experiment we cannot expect to reproduce the physical results in detail. The aim is to provide a range of model problems within which the physical experiment can be categorized. In order to set up the model problems, two types of idealizations are introduced. The first is concerned with the formulation of the equations to be solved, the boundary conditions and the representation of the dispersion. The second is related to the initial conditions. They were derived from the physical experiment. The model problems we propose differ only in the initial conditions. The governing equations for the coupled density-driven flow in porous media are derived from mass-conservation principles. On the assumption of incompressibility of the fluid, the following system of equations results (see e.g. [5, 23, 26]) ∂(nρ(ω)) + ∇ · (ρ(ω)~v ) = 0, ∂t ∂(nρ(ω)ω) + ∇ · (ρ(ω)(ω~v − D∇ω)) = 0. ∂t

(1a) (1b)

The flow equation (1a) describes conservation of fluid mass, whereas the transport equation (1b) describes conservation of salt mass. Note that the righthand sides of (1a) and (1b) vanish due to the absence of sources and sinks. Density is denoted by ρ = ρ(ω), the diffusion-dispersion tensor by D = D(~v ) and the mass-averaged velocity ~v is given by Darcy’s law ~v = −k/µ(ω)(∇p − ρ(ω)~g ),

(2)

where µ = µ(ω) is the dynamic viscosity and ~g is the vector of gravitational 12

acceleration. Inserting (2) into (1) leads to a nonlinear system of two partial differential equations for the unknown salt mass fraction ω and the pressure p. To close the system, equations of state for ρ and µ, and an expression for the diffusion-dispersion tensor have to be provided. For ρ and µ, various dependencies are provided in the literature (see [19, 23]). We choose an ideal fluid relationship for ρ, and a real, fitted function for µ with respect to the salt mass fraction ω 1 ω 1 1 + , := 1 − ρ ωmax ρf ωmax ρs µ := µf (1 + 1.85ω − 4.1ω 2 + 44.5ω 3 ). 



(3) (4)

The density of freshwater is denoted by ρf = ρ(0), the corresponding value at the maximum salt mass fraction ω = ωmax by ρs , see Table 1. For the diffusion-dispersion tensor, we use Scheidegger’s dispersion model [41] D(~v ) := nDm I + αt |~v |I + (αl − αt )~v~v /|~v |, where I denotes the unit tensor. We assume the validity of Fick’s law and Darcy’s law for our modeling approach. It has been proposed that a nonlinear dispersion approach should be used for high concentration gradients [18, 44]. Due to the low maximum density ρs , we do not follow this approach here. We describe the geometry by a cube of side length L = 200mm with an inclusion of higher permeability of 40 × 40mm2 extending from the bottom to the top of the cube, see section 2. The parameters describing the porous medium differ. The permeability in the inclusion, denoted by k1 , is higher than in the rest of the domain (k2 ). Accordingly, the values of the porosities and dispersion lengths take a corresponding subscript, see Table 1. The cube is closed and therefore no-flux boundary conditions for both ω and p are used on the entire boundary of the domain ~n · (~v ω − D∇ω) = 0,

~n · ~v = 0.

(5)

For a detailed discussion of consistent boundary conditions in density-dependent flow problems, see [17]. The setting described so far applies to all our model problems. The problems differ only in the choice of the initial conditions. Due to the assumed incompressibility, the pressure can be obtained from an elliptic constraint condition, a linear combination of (1a) and (1b). Therefore (1) is a differential-algebraic system and initial conditions have to be provided only for the salt mass fraction. For details, see e.g. [24]. Let (x, y, z) ∈ [−L/2, L/2]3 , then we define the 13

initial salt mass fraction by    0  

ω0 (x, y, z) = ωmax 1    1

2

+

z−h(x,y) w(x,y)

if z ≤ h(x, y) − if z ≥ h(x, y) + else.

w(x,y) , 2 w(x,y) , 2

(6)

with w(x, y) = c2 q

L2

, L4 + 16c21 (x2 + y 2 ) 2(x2 + y 2 ) h(x, y) = h0 + c1 + c3 p(x, y), L2  2  2   y y x 1 1  x −1 + − 1 − l2 l 2 l2 l 5 p(x, y) = p0 0 L L p0 = 2.48262m, l = , h0 = 5 10

if |x|, |y| ≤ l else

,

and ωmax , L from Table 1. Note that p(x, y) is continuous. The initial condition depends on the three parameters c1 , c2 and c3 , which have the dimension of a length. It realizes an instable initial salt distribution with ω0 < ωmax /2 ω0 > ωmax /2

for z < h(x, y), for z > h(x, y).

ω0 = ωmax /2

for z = h(x, y),

Therefore, the isosurface ω = 0.5ωmax is located at (x, y, h(x, y))∩[− L2 , L2 ]3 . Its position depends on c1 and c3 , in the background only on c1 . At the horizontal center x = y = 0, its vertical position is h0 +c3 /10, independent of c1 , c2 . At the horizontal corners x, y ∈ {− L2 , L2 }, its vertical position is h0 +c1 . The parameter c2 is the width of the transition zone perpendicular to the 0.5ωmax -isosurface, in which the salt mass fraction varies linearly from 0 to 1. Within the inclusion, the vertical position of the isosurface is perturbed by c3 p(x, y), a function without symmetries. The maximum of |p(x, y)| is approximately 1; therefore, c3 corresponds to the size of the perturbation in the inclusion. It enables the model problems to develop asymmetric solutions, as has been observed in the laboratory experiments (see section 2). In our numerical investigations, the following values of the parameters are used (c1 , c2 , c3 ) ∈ {0, 20, 40, 80} × {8, 16, 24} × {0, 2.5}.

(7)

For ease of notation, the unit mm is omitted here and in the following. The choice of c1 and c2 covers the presumed values of the experimental setting. The value for c3 is chosen to be the presumed experimental perturbation (vertical extension of the voxels, 2.5mm ) or the value for symmetric simulations (c3 = 0). 14

g replacements 0.1

h0 + c 1 h0 0

-0.05

-0.1

Fig. 5. Illustration of the initial condition (6) for (c 1 , c2 , c3 ) = (20, 16, 2.5). The ω = 0.5ωmax isosurface is displayed. The lifting of the isosurface at the horizontal corners is indicated on the left side, the width of the transition zone on the right.

For the purpose of illustration, we consider the initial condition for (c1 , c2 , c3 ) = (20, 16, 2.5). The corresponding isosurface (x, y, h(x, y)) is displayed in Fig. 5. The lifting of the isosurface at the corners is indicated by labels on the left side of the figure. On the right side, the surfaces (x, y, h(x, y) ± w(x, y)/2) indicate the width of the transition zone. The small perturbation of maximum size 2.5mm can be seen in the inclusion.

4

Numerical treatment

The numerical investigations presented in this paper are carried out with the software package d3 f, a software toolbox designed for the simulation of variable-density flow problems in porous media. In its full version, it contains a preprocessor for interactively designing geometry and physical parameters of model problems, a simulator for discretizing and solving the equations of the variable-density flow and a postprocessor for visualization and data extraction. The simulator is based on the software package UG [4], a toolbox for discretizing and solving partial differential equations. The postprocessor is based on the software package GRAPE [39]. Due to the simple geometry of the model problems, the preprocessor has not been used here. Next, we briefly describe the discretization and solution method applied to the model equations. Further details can be found in [12, 24, 25]. For the spatial discretization of (1) we use a grid hierarchy, which is created by uniform mesh refinement [3]. Starting from the coarsest mesh consisting of 500 hexahedra of size 20 × 20 × 40mm3 , the subsequent grids are generated by halving the mesh 15

Table 2 The hierarchy of grids Level i

# of hexahedra

# of grid points

0

500

726

1

4, 000

4, 851

2

32, 000

35, 301

3

256, 000

269, 001

4

2, 048, 000

2, 099, 601

size in each spatial direction. The number of hexahedra and grid points of the multigrid hierarchy is given in Table 2. Numerical simulations are carried out for grid level 3 and 4 corresponding to horizontal and vertical grid spacings of hx = hy = hz /2 = 2.5mm and hx = hy = hz /2 = 1.25mm, resp. The equations are discretized using a vertex-centered finite volume scheme. The convective terms are discretized without stabilisation. On the assumption of an essential front width of ≥ 8mm, i.e. a front resolution of n ≥ 2 elements, this choice is preferable to stabilized discretizations. A consistent velocity approximation is applied, see [15]. The discretization is locally mass-conserving and secondorder consistent for the unknowns ω and p. To solve the discrete equations, a fully implicit solution technique is applied. The arising nonlinear discrete equations are solved using Newton’s method, for the linear subproblems a linear multigrid method is applied. The components of the multigrid scheme are standard. The prolongation used arises from the natural embedding of one grid into the next finer one. The restriction is chosen to be adjoint to the prolongation. The matrices of the linear subproblems on the coarser grid levels are chosen to be the Jacobi matrices of the corresponding nonlinear problems. The points of linearization are obtained by restriction. On the coarsest grid, the linear equations are solved exactly. As a smoother, we use a block variant of the symmetric SOR method, as described in [24]. The multigrid scheme is used as a preconditioner for the Bi-CGStab method [51]. For the time discretization, the fractional-step-Θ scheme [37] is employed. It is a three-step θ-scheme, which is second-order accurate and especially adapted to convection dominated transport problems. For a detailed discussion in the context of variable-density flow, see [24, 25]. Two issues specific to our investigation require special attention and will be discussed below in more detail. The first is concerned with the realization of the initial conditions (6), the second with the time-step control. In general, the initial conditions (6) cannot be represented exactly on either of the grids of the hierarchy. We derive the initial condition ωl,0 on grid level l by the L2 -projection of ω0 , which is given by kωl,0 − ω0 kL2 = min .

16

The resulting linear system of equations involves the mass matrix, which is well-conditioned. It is solved by a Gauss-Seidel iteration. For the boundary conditions (5) used in our model problems, the total initial salt mass is represented exactly on every grid Z

ωl,0 dV =

Z

ω0 dV,

∀l ≥ 0.

(8)

Note that the L2 -projection leads to a non-monotonous initial salt distribution. If a pointwise interpolation of the initial conditions (6) onto the nodes of the grid together with tri-linear interpolation is used, eq. (8) is fulfilled only up to an order O(h), leading to a serious deterioration of the approximation to the continuous solution. Due to the density-driven instabilities, the model problems show strongly varying dynamics in time. In order to control the discretization error introduced by the time integration, we chose the time-step size to fulfill a Courant criterion. To this end we define a Courant number nc for each hexahedral element e nc (e) = |~v (e)|∆t/h(e), where ~v (e) denotes the Darcy velocity at the midpoint of e, h(e) the shortest edge of e and ∆t the time-step size of the last time step executed. The new time-step size is chosen according to ∆tnew = ∆t/ max e nc (e), i.e. an extrapolation of the time-step size in order to fulfill nc (e) ≤ 1, ∀e. This criterion ensures a stable and accurate discretization of the convection and an accurate time integration. The initial time-step size was chosen to be 0.1s leading to a maximum Courant number smaller than one on all grids we considered.

5

Comparison of experiment and numerical simulations

In this section, we compare the laboratory experiments with the numerical simulations. Due to the high nonlinearity of the problem, an exact reproduction of the experimental findings cannot be expected. Instead, we aim to gain a qualitative understanding of the problem. To this end, we define four criteria to carry out the comparison. Two of these criteria are quantitative and therefore allow a quantitative comparison. As the physical results are reproduced only qualitatively, the considerations will lead merely to qualitative results. The other two criteria are qualitative and lead to corresponding results. Next, we give the definitions. 17

t=0s

t = 2549 s

t = 4883 s

t = 7126 s

Fig. 6. Simulated isolines of the salt mass fraction ω/ω max = 1/8, 2/8, . . . , 7/8 for the model problem with parameters (c 1 , c2 , c3 ) = (20, 16, 2.5), shown in diagonal cross-sections (left) and parallel cross-sections (right). The section planes as well as the time sequence correspond to Fig. 4.

18

1. The first quantitative criterion is given by the scalar functional on the solution defined by Z 1 ρωdV, mi (t) := ωmax ρs Vinc Vinc

(9)

where Vinc denotes the volume of the inclusion. It can be evaluated for the experimental and the numerical findings. This quantity represents the normalized mass of salt within the inclusion. The experimental evaluation will be denoted by mi,e , the numerical by mi,s . The latter depends on the configuration parameters c1 , c2 , c3 of the model problem as well as on the discretization used. For the case that the flow in the inclusion is disconnected from the flow outside the inclusion, mi (t) is constant in time. Its dynamic behavior is therefore a measure for the coupling of the two flow regimes. We refer to this phenomenon as the localization of flow. 2. The second quantitative criterion is the finger-tip velocity. Let zs (t):= min{z|ω(t, x, y, z) = 0.5ωmax , (x, y, z) ∈ Vinc }

(10)

be the finger-tip position of the saltwater finger. The minimum z-coordinate of the ω = 0.5ωmax -isosurface serves as the z- component of the saltwater finger. It is not only applicable for the case of one finger but also for two or more fingers. Its time derivative is called the finger-tip velocity of the saltwater finger. Note that the finger-position may jump, if the finger concentration drops below 0.5ωmax . In that case, the next finger is tracked. Correspondingly, the finger-tip velocity of the freshwater finger is defined by using the time derivative of zf (t):= max{z|ω(t, x, y, z) = 0.5ωmax , (x, y, z) ∈ Vinc }.

(11)

3. The first qualitative criterion is the effect of a sharpening of the saltwaterfreshwater interface as opposed to the dispersive mixing effect. Since this effect appears only in localized regions, it is called constriction effect. 4. The second qualitative criterion is the observation of instabilities outside the inclusion. They can be observed as the development of saltwater-freshwater fingers. We call this the effect of secondary instabilities. Before we discuss the criteria in detail, we want to illustrate the extent to which the physical results can be reproduced by numerical simulations. In Fig. 6, we display the numerical results obtained for the parameters (c1 , c2 , c3 ) = (20, 16, 2.5) on grid level 4. The section planes as well as the time sequence correspond to the ones chosen for the experimental results given in Fig. 4. As can be seen clearly, the qualitative features as well as the speed of the finger 19

Fig. 7. Isolines of the salt mass fraction ω/ω max = 1/8, 2/8, . . . , 7/8 on a cross-section at t = 4883s for horizontal initial front (c 1 , c2 , c3 ) = (0, 8, 2.5) (left), curved initial front (c1 , c2 , c3 ) = (80, 8, 2.5) (right). Details are described in the text.

development are reproduced well, whereas the details of the salt distribution differ significantly. Remark 5.1 The limitations of reproducibility illustrated by the comparison of Figs. 4 and 6 are due to the high nonlinearities and the modeling insufficiencies. They are not due to numerical limitations. Now we will compare the experimental and the numerical results with respect to the criteria 1-4. We focus on non-symmetric solutions (c3 = 2.5), which correspond to the experiment. The results presented were obtained on grid level 4. Comments on the comparison of solutions obtained on grid levels 3 and 4 as well as on symmetric and non-symmetric solutions are given in subsection 5.5.

5.1 The localization of the flow

First, we illustrate two extreme flow patterns. The one extreme is obtained for the parameters (c1 , c2 , c3 ) = (0, 8, 2.5) and is shown in Fig. 7 (left). Here, the flow field is localized mainly in the inclusion. The initially horizontal interface remains horizontal in the background, even for intermediate times (t = 4883s). The other extreme is obtained for the parameters (c1 , c2 , c3 ) = (80, 8, 2.5) and is shown in Fig. 7 (right). In this case, the flow field extends over the entire box and the initially curved interface is bent even more after intermediate times (t = 4883s). A systematic evaluation of the salt mass in the inclusion (9) as a function of time is displayed in Fig. 8. The experimental results are indicated by (x). They document a moderate increase of the salt mass in the inclusion, which is due to an inflow of salt water into the inclusion in the upper part 20

g replacements

1

0.8

mi

0.6

0.4 experiment c1 = 80 c1 = 40 c1 = 20 c1 = 0

0.2

0 0

1000

2000

3000

4000

5000

6000

7000

time [s]

Fig. 8. The normalized salt mass in the inclusion. Experimental results (x), numerical results are grouped by parameter c 1 . Within the groups, corresponding to c2 ∈ {8, 16, 24}, c3 = 2.5 and displayed with the same linetype, the variations are small.

of the domain. The numerical results are indicated by dashed lines, which are grouped together by parameter c1 . Within the groups, the different graphs (not distinguished by line-types) correspond to different values of the initial front width (c2 ∈ {8, 16, 24}). This second parameter is of minor importance for the localization of the flow. An initial interface with zero curvature (c1 = 0) leads to an essentially localized flow, see Fig. 7 (left). Only at large times can a small increase of the salt mass be observed, when boundary effects become noticeable. An increasing initial curvature leads to an increasingly global flow pattern. In the extreme case c1 = 80, the velocity in the inclusion is only pointing downward and the upward recirculation occurs in the less permeable zone. Note that in this case no ascending freshwater finger can be seen, cf. Fig. 7, (right). The transition from local to global flow is essentially controlled by c1 . The experimental results can be ranged in in this one-dimensional field, corresponding to mi,s (c1 ), c1 ∈ [10, 20]. Note that the experimental and the numerical results are in good qualitative agreement for small and medium times.

5.2 The finger-tip velocity First, we compare the experimental findings with the numerical results obtained for (c1 , c2 , c3 ) = (20, 16, 2.5). The evaluation of the finger-tip positions 21

g replacements 0.2 exp. sim.

zs , zf

0.15

0.1

0.05

0 0

1000

2000

3000

4000

5000

6000

7000

time [s]

Fig. 9. The finger-tip positions zs , zf derived from the experiment and the numerical simulation with (c1 , c2 , c3 ) = (20, 16, 2.5). The jump of zf at t = 5000s corresponds to a change of the freshwater finger tracked.

(10), (11) is displayed in Fig. 9. For a large time interval, the finger-tip positions show a linear variation with time for the experimental data as well as for the numerical simulation. This behavior allows us to define a finger-tip velocity for the experiment and for the simulation, which is independent of time. It is denoted by vs for the saltwater finger and vf for the freshwater finger. This is true for almost all simulations using parameters from the range (7). The finger-tip velocities for the experiment and for the simulation with the parameters c1 ∈ {0, 20, 40, 80}, c2 ∈ {8, 16, 24} are displayed in Fig. 10. The numerical results are grouped by the parameter c1 . The groups contain the velocities for a varying initial front width c2 ∈ {8, 16, 24}, which does not have a significant influence. A velocity vf for c1 = 40 has not been included, since it could not be identified clearly. Note that for c1 = 80 the freshwater velocity is negative, i.e. no freshwater finger was ascending. The velocities from the experiment are indicated by solid lines. We see that the experimental value for vf can be matched best by simulations using c1 = 20, the experimental value for vs by simulations using c1 = 0. Both values might be matched to some extent by an intermediate parameter c1 = 10, which is in good agreement with the result given in the previous subsection. 22

g replacements experiment c2 = 8 c2 = 16 c2 = 24

4e-05

vs , vf [ms−1 ]

2e-05

0

-2e-05

-4e-05 0

20

40

80

c1

Fig. 10. The finger-tip velocities v s , vf derived from the experiment and the numerical simulations. The numerical results are grouped by the parameter c 1 . The results were obtained on grid level 4 using c 3 = 2.5.

5.3 The constriction effect The constriction of the width of the saltwater-freshwater interface with time, which is opposed to the dispersive mixing effect, is shown in detail in Fig. 11 at t = 7126s. In the figure, the salt mass fraction of the experiment is displayed on the left and the numerical solution for (c1 , c2 , c3 ) = (20, 16, 2.5) on the right. The width of the interface at the left side of the inclusion is reduced by a factor of 0.33 relative to the actual front width. The velocity field obtained from the numerical simulation explains this effect.

5.4 Secondary instability The secondary instabilities arise in the region with low permeability and can be observed within a small parameter range. They can occur only for a sufficiently narrow and sufficiently horizontal saltwater-freshwater interface. In the range of the parameters used for our model problem, an initial curvature c1 ≤ 20 and an initial interface width c2 ≤ 16 are necessary. We show the results for (c1 , c2 , c3 ) = (20, 16, 2.5) at time t = 7126 on a diagonal cross-section on the right-hand side of Fig. 12 together with the experimental results in the corresponding cross-section. As can be seen such secondary instabilities are clearly observed under the experimental conditions. Our simulations with 23

Fig. 11. Constriction of the saltwater-freshwater interface on a transverse cross-section at t = 7126s for the experimental results (left) and numerical simulation with (c1 , c2 , c3 ) = (20, 16, 2.5) (right). The isolines of the simulation correspond to of the salt mass fraction ω/ωmax = 1/8, 2/8, . . . , 7/8.

Fig. 12. Secondary instability of the saltwater-freshwater interface on a diagonal cross-section at t = 7126 for the experimental results (left) and numerical simulation with (c1 , c2 , c3 ) = (20, 16, 2.5) (right). The isolines of the simulation correspond to of the salt mass fraction ω/ωmax = 1/8, 2/8, . . . , 7/8.

c1 = 0 did not show the secondary instabilities due to lack of perturbation of the interface.

5.5 Further remarks Up to now, we have only considered numerical simulations with an initial perturbation (c3 = 2.5). This leads to non-symmetric numerical solutions, which corresponds to the experiment. The choice c3 = 0 specifies a symmetric initial condition and, due to the symmetry of the model problem, will lead 24

Fig. 13. Comparison of the numerical results for (c 1 , c2 ) = (20, 16) at t = 2549s: symmetric (c3 = 0, left), non-symmetric solution (c 3 = 2.5, right). The isolines of the salt mass fraction ω/ωmax = 1/8, 2/8, . . . , 7/8 are shown.

to a symmetric solution for all times, provided that non-symmetric numerical perturbations can be neglected. In Fig. 13, the numerical solutions at t = 2549s for (c1 , c2 ) = (20, 16) are displayed. On the left, the symmetric solution corresponding to c3 = 0 is shown, on the right the non-symmetric solution corresponding to c3 = 2.5. The small perturbation is amplified strongly during the time evolution leading to an essentially non-symmetric flow regime, which seems to be physically prevailing. The experimental conditions yielded such non-symmetric fingering in a symmetric experimental setup, which are due to inhomogeneities in the porous medium or slight disturbances in an otherwise controlled experimental system. We conclude this section with two remarks on the precision of the numerical results. First, we look at the distribution of the salt mass fraction in space at a certain time. Fig. 14 shows the isosurface ω = 0.5ωmax of the numerical solutions for (c1 , c2 , c3 ) = (20, 16, 2.5) at t = 2549s. The solution obtained on grid level 3 is shown on the left, the one obtained on grid level 4 on the right. While qualitatively no differences can be observed, the exact size of the fingers differs on the two grid levels. Due to the dynamics of the system, these differences increase in time, with the largest differences at the end of the simulation. The differences, though significant, are much smaller than the ones between the experimental and the numerical results. We claim that a further improvement of the numerical accuracy will not change this situation. Secondly, we consider the finger-tip velocity vs of the saltwater finger. We denote its value obtained on grid level l by vsl . Let δsl := |vsl − vsl−1 |/|vsl | denote the relative deviation of the velocities on grid level l and l − 1 and els := |vsl − vs∞ |/|vs∞| the relative error of the velocity on grid level l. On the assumption of second-order convergence, from δsl ≤ 1 follows els ≤ 0.5δsl . The numerical evaluation of the finger-tip velocities leads to e4s ≤ 0.5% for (c1 , c2 ) ∈ (20, 40, 80) × (16, 24), covering the solution depicted in Figs. 6. This 25

Fig. 14. Comparison of numerical solutions for (c 1 , c2 , c3 ) = (20, 16, 2.5) at t = 2549s. The isosurface ω/ωmax = 0.5 on grid level 3 (left) and level 4 (right) is shown.

accuracy is sufficient. For c1 = 0 or c2 = 8, errors up to 18% result. The findings for the freshwater finger velocity are less reliable, showing errors up to 28%. We do not consider this to be sufficiently accurate. A clear explanation cannot be given.

6

Discussion

In this paper, we have investigated saltwater-freshwater instabilities. We have provided results from a laboratory experiment as well as from numerical simulations. Due to the highly nonlinear character of the problem under consideration and the inevitable modelling errors, an exact reproduction of the experimental findings could not be obtained. The aim of this paper was to show the extent to which a numerical simulation of unstable fingering processes is feasible. With respect to the model setup investigated in this paper, a detailed answer can be given. The trivial result is that a detailed reproduction of the experimental findings cannot be obtained. Furthermore, we claim that the limiting factor is given by the mathematical model and the uncertainties of the initial conditions. In our opinion, the numerical errors introduced by the discretization contribute to only a minor extent to the mismatch observed between physical experiment and numerical simulation. On the other hand, we have identified structural aspects of the solution which coincide. These are the global flow pattern, referred to as localization of the flow and the finger-tip velocity, an almost constant velocity of the saltwater-freshwater propagation. In both cases, a qualitative agreement could be observed. The relative errors, even though significantly large, seem to be smaller than the inevitable modelling error. Furthermore, two qualitative phenomena could be 26

observed, the constriction effect and the secondary instabilities. Experiment and simulation show similarities with respect to these phenomena, indicating that basic features of the laboratory experiment have been captured.

Acknowledgements We gratefully acknowledge financial support for travel from the European Commission under INTAS project no. 00-1056. This work was partially funded by the Swiss Federal Office for Education and Science (BBW) in the framework of European Commission grant EVK1-CT-2000-00062. All computations were carried out on an MIMD cluster at the Scientific Computing Center (IWR) at the University of Heidelberg. Finally, we would like to thank the reviewers for their constructive comments.

References [1] E. Alvarez-Lacalle, J. Ortin, and J. Casademunt. Nonlinear saffman-taylor instability. Phys. Rev. Lett., 92(5):Art. no. 054501, 2004. [2] J.-C. Bacri, D. Salin, and Y.C. Yortsos. Analyse lin´eaire de la stabilit´e de l’´ecoulement de fluides miscibles en milieux poreux. C. R. Acad. Sci. Paris, 314(II):139 – 144, 1992. [3] R.E. Bank, A.H. Sherman, and A. Weiser. Refinement algorithms and data structures for regular local mesh refinement. In: Scientific computing IMACS, North-Holland, Amsterdam, 1983. [4] P. Bastian, K. Birken, K. Johannsen, S. Lang, N. Neuss, H. Rentz-Reichert, and C. Wieners. UG - a flexible software toolbox for solving partial differential equations. Computing and Visualization in Science, 1:27–40, 1997. [5] J. Bear and Y. Bachmat. Introduction to Modeling of Transport Phenomena in Porous Media. Kluwer Academic Publishers, 1991. [6] J.L. Beck. Convection in a box of porous material saturated with fluid. Physics of Fluids, 15:1377 – 1383, 1972. [7] P.T. Callaghan. Principles of Nuclear Magnetic Resonance Microscopy. Clarendon Press, Oxford, 1991. [8] J. Casademunt. Viscous fingering as a paradigm of interfacial pattern formation: recent results and new challenges. Chaos, 14(3):809 – 824, 2004. [9] R.L. Chuoke, P. van Meurs, and C. van der Poel. The instability of slow, immiscible, viscous liquid-liquid displacements in permeable media. Trans. Inst. Min. Metall. Pet. Eng., 216:188 – 194, 1959.

27

[10] G. Coskuner and R.G. Bentsen. An extended theory to predict the onset of viscous instabilities for miscible displacements in porous media. Transport in Porous Media, 5:473 – 490, 1990. [11] H.-J.G. Diersch and O. Kolditz. Variable-density flow and transport in porous media: approaches and challenges. Advances in Water Resources, 25:899 – 944, 2002. [12] E. Fein (ed). d3f - Ein Programmpaket zur Modellierung von Dichtestr¨omungen. GRS, Braunschweig, GRS - 139, ISBN 3 - 923875 - 97 - 5, 1998. [13] J.W. Elder. Steady free convection in a porous medium heated from below. J. Fluid Mechanics, 27:29 – 48, 1967. [14] J. Fernandez, P. Kurowski, P. Petitjeans, and E. Meiburg. Density-driven unstable flows of miscible fluids in a Hele-Shaw cell. J. Fluid Mechanics, 451:239 – 260, 2002. [15] P. Frolkovic. Consistent velocity approximation for density driven flow and transport. Advanced Computational Methods in Engineering, eds. R. van Keer et. al., pages 603–611, 1998. [16] F. Graf, E. Meiburg, and C. H¨artel. Density-driven instabilities of miscible fluids in a Hele-Shaw cell: linear stability analysis of the three-dimensional Stokes equations. J. Fluid Mechanics, 451:261 – 282, 2002. [17] S.M. Hassanizadeh and A. Leijnse. On the modelling of brine transport in porous media. Water Resources Research, 24:321–330, 1988. [18] S.M. Hassanizadeh and A. Leijnse. A non-linear theory of high-concentrationgradient dispersion in porous media. Advances in Water Resources, 18(4):203 – 215, 1995. [19] A.W. Herbert, C.P. Jackson, and D.A. Lever. Coupled groundwater flow and solute transport with fluid density strongly dependent upon concentration. Water Resources Research, 24:1781–1795, 1988. [20] K.-H. Herrmann, A. Pohlmeier, S. Wiese, N.J. Shah, O. Nitzsche, and H. Vereecken. Three-dimensional nickel ion transport through porous media using magnetic resonance imaging. Journal of Environmental Quality, 31:506 – 514, 2002. [21] M.A. Hill. Channeling in packed columns. Chem. Eng. Sci., 1:247 – 253, 1952. [22] F. Hoffman, D. Ronen, and Z. Pearl. Evaluation of flow characteristics of a sand column using magnetic resonance imaging. Journal of Contaminant Hydrology, 22:95 – 107, 1996. [23] E.O. Holzbecher. Modeling Density-Driven Flow in Porous Media. SpringerVerlag, Berlin, Heidelberg, 1998. [24] K. Johannsen. Numerische Aspekte dichtegetriebener Stroemung in poroesen Medien. Habilitationsschrift, Heidelberg University, 2004.

28

[25] K. Johannsen, W. Kinzelbach, S.E. Oswald, and G. Wittum. The saltpool benchmark problem - numerical simulation of saltwater upconing in a porous medium. Advances in Water Resources, 25:335 – 348, 2002. [26] A. Leijnse. Three-dimensional modeling of coupled flow and transport in porous media. PhD thesis, University of Notre Dame, Indiana, 1992. [27] J. Martin, N. Rakotomalala, and D. Salin. Gravitational instability of miscible fluids in a Hele-Shaw cell. Physics of Fluids, 14(2):902 – 905, 2002. [28] E. Meiburg and C.-Y. Chen. High-accuracy implicit finite-difference simulations of homogeneous and heterogeneous miscible porous media flows. SPE J., 5(2):129 – 137, 2000. [29] T. Menand and A.W. Woods. Dispersion, scale and time dependence of mixing zones under gravitationally stable and unstable displacements in porous media. Water Resources Research, 41:W05014: doi, 10.1029/2004WR003701, 2005. [30] S. Oswald. http://www.ihw.ethz.ch/url/connect?ihw=fingering EN. [31] S. Oswald. Dichtestr¨ omungen in por¨ osen Medien: Dreidimensionale Experimente und Modellierung. PhD thesis, Institut f¨ ur Hydromechanik und Wasserwirtschaft, ETH Z¨ urich, Switzerland, no. 12812, 1998. [32] S. Oswald, W. Kinzelbach, A. Greiner, and G. Brix. Observation of flow and transport processes in artificial porous media via magnetic resonance imaging in three dimensions. Geoderma, 80(3-4):417 – 429, 1997. [33] S.E. Oswald and W. Kinzelbach. Three-dimensional physical benchmark experiments to test variable-density flow models. Journal of Hydrology, 290(12):22 – 42, 2004. [34] S.E. Oswald, M.B. Scheidegger, and W. Kinzelbach. Time-dependent measurement of strongly density-dependent flow in a porous medium via nuclear magnetic resonance imaging. Transport in Porous Media, 47(1-2):169 – 193, 2002. [35] Z. Pearl, M. Magaritz, and P. Bendel. Nuclear magnetic resonance imaging of miscible fingering in porous media. Transport in Porous Media, 12:107 – 123, 1993. [36] P. Petitjeans, C.-Y. Chen, E. Meiburg, and T. Maxworthy. Miscible quarter fivespot displacements in a Hele-Shaw cell and the role of flow induced dispersion. Physics of Fluids, 11(7):1705 – 1716, 1999. [37] R. Rannacher. Numerical analysis of nonstationary fluid flow. Technical Report SFB 123, Preprint 492, University of Heidelberg, Germany, November 1988. [38] A. Riaz and E. Meiburg. Three-dimensional miscible displacement simulations in homogeneous porous media with gravity override. J. Fluid Mechanics, 494:95 – 117, 2003.

29

[39] M. Rumpf and A. Wierse. GRAPE, eine objektorientierte Visualisierungs- und Numerikplattform. Informatik Forschung und Entwicklung, 7:145–151, 1992. [40] P.G. Saffman and G.I. Taylor. The penetration of a fluid into a porous medium or Hele-Shaw cell containing a more visous fluid. Proc. Royal Society of London A, 245:312 – 329, 1958. [41] A.E. Scheidegger. General theory of dispersion in porous media. Journal of Geophysical Research, 66:3273, 1961. [42] R.A. Schincariol and F.W. Schwartz. An experimental investigation of variable density flow and mixing in homogenous and heterogeneous porous media. Water Resources Research, 26(10):2317 – 2329, 1990. [43] K.J. Schneider. Investigation of the influence of free thermal convection on heat transfer through granular material. In Proc. 11th International Congress on Refrigeration, Oxford, pages 247 – 253. Pergamon Press, 11-4 1963. [44] R.J. Schotting, H. Moser, and S.M. Hassanizadeh. High-concentrationgradient dispersion in porous media: experiments, analysis and approximations. Advances in Water Resources, 22(7):665 – 680, 1999. [45] C.T. Simmons. Variable-density groundwater flow: From current challenges to future possibilities. Hydrogeol. J., 13:116 – 119, 2005. [46] C.T. Simmons, T.R. Fenstemaker, and J.M. Sharp Jr. Variable-density groundwater flow and solute transport in heterogeneous porous media: approaches and challenges. J. Contam. Hydrology, 52(1-4):245 – 275, 2001. [47] C.T. Simmons, K. A. Narayan, and R.A. Wooding. On a test case for density-dependent groundwater flow and solute transport models: The salt lake problem. Water Resources Research, 35(12):3607 – 3620, 1999. [48] C.T. Tan and G.M. Homsy. Stability of miscible discplacement in porous media: Rectilinear flow. Physics of Fluids, 29:3549 – 3556, 1986. [49] H.A. Tchelepi and F.M.J. Orr. Interaction of viscous fingering, permeability heterogeneity, and gravity segregation in three dimensions. SPE Res. Engng., 9(4):266 – 271, 1994. [50] H. van As and D. van Dusschoten. NMR methods for imaging of transport processes in micro-porous systems. Geoderma, 80(3-4):389 – 403, 1997. [51] H. van der Vorst. Bi-CGSTAB: A fast and smoothly converging variant of bicg for the solution of nonsymmetric linear systems. SIAM J. Sci. Stat. Comp., 13:631 – 644, 1992. [52] D. Weatherill, C.T. Simmons, C.I. Voss, and N.I. Robinson. Testing density-dependent groundwater models: two-dimensional steady state unstable convection in infinite, finite and inclined porous layers. Advances in Water Resources, in press, 2004.

30

[53] R.A. Wooding. The stability of a viscous liquid in a vertical tube containing porous material. Proc. R. Soc. London, Ser. A, 252:120 – 134, 1959. [54] R.A. Wooding. Growth of fingers at an unstable diffusing interface in a porous medium or Hele-Shaw cell. J. Fluid Mechanics, 39(3):477 – 495, 1969. [55] R.A. Wooding, S.W. Tyler, and I. White. Convection in groundwater below an evaporating salt lake: 1. onset of instability. Water Resources Research, 33(6):1199 – 1217, 1997. [56] R.A. Wooding, S.W. Tyler, I. White, and P.A. Anderson. Convection in groundwater below an evaporating salt lake: 2. evolution of fingers or plumes. Water Resources Research, 33(6):1219 – 1228, 1997. [57] W.B. Zimmerman and G.M. Homsy. Three-dimensional viscous fingering: a numerical study. Physics of Fluids, 4(9):1901 – 1914, 1992.

31

Numerical simulation of three-dimensional saltwater ...

Dec 18, 2005 - of numerical tools for three-dimensional, transient instabilities. In this con- ..... used as a preconditioner for the Bi-CGStab method [51]. For the time dis- ..... Steady free convection in a porous medium heated from below. J.

2MB Sizes 3 Downloads 222 Views

Recommend Documents

Numerical simulation of three-dimensional saltwater ...
Dec 18, 2005 - subject to long-standing investigation and numerical analysis. ..... the software package d3f, a software toolbox designed for the simulation of.

Numerical simulation of saltwater upconing in a porous ...
Nov 9, 2001 - Grid convergence in space and time is investigated leading to ... transient, density-dependent flow field, and the experimental data are obtained ..... tured meshes is inferior to hexahedral meshes both with respect to memory.

Numerical simulation of three-dimensional saltwater ...
Dec 18, 2005 - roscan ACS/NT, Best, The Netherlands) at the university hospital in Zürich, ..... center x = y = 0, its vertical position is h0+c3/10, independent of c1, c2. At the horizontal .... fingers. We call this the effect of secondary instabi

1630 Numerical Simulation of Pharmaceutical Powder Compaction ...
1630 Numerical Simulation of Pharmaceutical Powder Compaction using ANSYS - A Baroutaji.pdf. 1630 Numerical Simulation of Pharmaceutical Powder Compaction using ANSYS - A Baroutaji.pdf. Open. Extract. Open with. Sign In. Main menu. Displaying 1630 Nu

Numerical Simulation of Nonoptimal Dynamic ...
Computation of these models is critical to advance our ..... Let us now study the model specification of Example 2 in Kubler and Polemarchakis ..... sequentially incomplete markets, borrowing constraints, transactions costs, cash-in-advance.

Numerical Simulation of Fuel Injection for Application to ...
Simulate high speed reacting flow processes for application to Mach 10-12 .... 12. 14 x/d y/d. McDaniel&Graves. Rogers. Gruber et al. Musielak. Log. (Musielak) ...

Development and Numerical Simulation of Algorithms ...
Development and Numerical Simulation of. Algorithms to the Computational Resolution of. Ordinary Differential Equations. Leniel Braz de Oliveira Macaferi. 1.

Hybrid symbolic and numerical simulation studies of ...
Hybrid symbolic and numerical simulation studies of ... for Self-Organizing and Intelligent Systems (CSOIS), Dept. of Electrical and Computer Engineering,.

Numerical simulation of coextrusion and film casting
where oi, i = 1, 2, is the Cauchy stress tensor on each side of the interface and ii is .... which means that the momentum equations are satisfied in the distribution ...

On numerical simulation of high-speed CCD/CMOS ...
On numerical simulation of high-speed CCD/CMOS-based. Wavefront Sensors for Adaptive Optics. Mikhail V. Konnik and James Welsh. School of Electrical ...

Numerical Simulation of 3D EM Borehole ...
exponential convergence rates in terms of a user prescribed quantity of interest against the ... interpolation operator Π defined in Demkowicz and Buffa (2005) for.

On numerical simulation of high-speed CCD ... - Semantic Scholar
have a smaller photosensitive area that cause higher photon shot noise, higher ... that there are i interactions per pixel and PI is the number of interacting photons. .... to the random trapping and emission of mobile charge carriers resulting in ..

Numerical simulation of coextrusion and film casting
This relation is valid in the entire domain R. In other words, the solution of this transport equation (12) ... the name pseudoconcentration). .... computed at the price of very-small-amplitude wiggles in the vicinity of the discontinuities. These ..

Numerical Simulation of the Filling and Curing Stages ...
testados vários esquemas transitórios e advectivos, com vista ao reconhecimentos de quais os que .... Quotient between the old and the new time step. [–] μ. Viscosity. [kg/(m⋅s)] ρ. Density ..... where the part thickness changes abruptly, or

Numerical simulation and experiments of burning ...
Available online 25 July 2009. Keywords: CFD ...... classes (up to 10 mm in diameter roundwood) two 2 m tall trees ...... hr Б qr;biVb ј jb;eЅ4pIbрTeЮ А UЉ.

Numerical Simulation and Experimental Study of ...
2007; published online August 29, 2008. Review conducted by Jian Cao. Journal of Manufacturing Science and Engineering. OCTOBER ... also used to fit the experimental data: M (t) = i=1 ... formed using a commercial FEM code MSC/MARC.

Direct Numerical Simulation of Pipe Flow Using a ...
These DNS are usually implemented via spectral (pseu- dospectral ... In this domain, the flow is naturally periodic along azimuthal (θ) direction and taken to be ...

Numerical Simulation and Experimental Study of ...
Numerical Simulation and. Experimental Study of Residual. Stresses in Compression Molding of Precision Glass Optical. Components. Compression molding of glass optical components is a high volume near net-shape pre- cision fabrication method. Residual

fundamentals of numerical reservoir simulation pdf
fundamentals of numerical reservoir simulation pdf. fundamentals of numerical reservoir simulation pdf. Open. Extract. Open with. Sign In. Main menu.

Numerical Simulation of the Filling and Curing Stages ...
utilização do software de CFD multi-objectivos CFX, concebido para a ... combination with the free-slip boundary condition for the air phase. ...... 2.5 MHz processor and 512 MB RAM memory, using Windows® 2000 operating system. mesh1 ..... V M. V