II. Dealing with Physical Laws: A Stochastic Flowpath Analysis Example

Similar documents
A GENERALIZED CONVECTION-DIFFUSION MODEL FOR SUBGRID TRANSPORT IN POROUS MEDIA

BME STUDIES OF STOCHASTIC DIFFERENTIAL EQUATIONS REPRESENTING PHYSICAL LAW

Streamline calculations. Lecture note 2

Applications of Partial Differential Equations in Reservoir Simulation

A PSEUDO FUNCTION APPROACH IN RESERVOIR SIMULATION

Homogenization and numerical Upscaling. Unsaturated flow and two-phase flow

B005 A NEW FAST FOURIER TRANSFORM ALGORITHM FOR FLUID FLOW SIMULATION

On the solution of incompressible two-phase ow by a p-version discontinuous Galerkin method

Ability of Darcy s Law for Extension in Two- Phase Flow for Sedimentary Medium in Capillary Non-equilibrium Situations

Chapter Seven. For ideal gases, the ideal gas law provides a precise relationship between density and pressure:

dynamics of f luids in porous media

The effect of heterogeneity on unsteady-state displacements

Local Time Step for a Finite Volume Scheme I.Faille F.Nataf*, F.Willien, S.Wolf**

1 Modeling Immiscible Fluid Flow in Porous Media

Numerical Simulation of Single-Phase and Multiphase Non-Darcy Flow in Porous and Fractured Reservoirs

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

INJECTION, CONDUCTION AND PRODUCTION

In all of the following equations, is the coefficient of permeability in the x direction, and is the hydraulic head.

V (r,t) = i ˆ u( x, y,z,t) + ˆ j v( x, y,z,t) + k ˆ w( x, y, z,t)

Multilevel Monte Carlo for two phase flow and transport in random heterogeneous porous media

Waterflooding Performance of Communicating Stratified Reservoirs With Log-Normal Permeability Distribution

Chapter 4: Fluid Kinematics

A New Method for Calculating Oil-Water Relative Permeabilities with Consideration of Capillary Pressure

New Fast Kalman filter method

Macro-Scale Dynamic Effects in Homogeneous and Heterogeneous Porous Media

2. Conservation of Mass

Numerical methods for the Navier- Stokes equations

pressure-saturations relationships for two-phase flow in 3D porous media: implications micro-heterogeneities

Generalised Separable Solution of Double Phase Flow through Homogeneous Porous Medium in Vertical Downward Direction Due to Difference in Viscosity

Upscaling of Richards equation for soil moisture dynamics to be utilized in mesoscale atmospheric models

Chapter 5. The Differential Forms of the Fundamental Laws

Lecture Notes to Accompany. Scientific Computing An Introductory Survey. by Michael T. Heath. Chapter 9

Scientific Computing: An Introductory Survey

Review for Exam 2 Ben Wang and Mark Styczynski

REDUCING ORDER METHODS APPLIED TO RESERVOIR SIMULATION

Getting started: CFD notation

Msc Thesis: Modeling Single & Multi-phase flows in petroleum reservoirs using Comsol Multiphysics: ''Pore to field-scale effects''

Course Outline. 2. Vectors in V 3.

Two-Fluid Model 41. Simple isothermal two-fluid two-phase models for stratified flow:

" = ˆ i # #x + ˆ j # #y + ˆ k # #z. "t = D #2 h

Analysis of oil displacement by water in oil reservoirs with horizontal wells

Appendix: Orthogonal Curvilinear Coordinates. We define the infinitesimal spatial displacement vector dx in a given orthogonal coordinate system with

Simulation of Unsaturated Flow Using Richards Equation

Pore-scale modeling extension of constitutive relationships in the range of residual saturations

Basic Aspects of Discretization

Darcy's Law. Laboratory 2 HWR 531/431

Fundamentals of Fluid Dynamics: Ideal Flow Theory & Basic Aerodynamics

A FRONT-TRACKING METHOD FOR HYPERBOLIC THREE-PHASE MODELS

CS 450 Numerical Analysis. Chapter 9: Initial Value Problems for Ordinary Differential Equations

Simple closed form formulas for predicting groundwater flow model uncertainty in complex, heterogeneous trending media

CURRENT MATERIAL: Vector Calculus.

Modeling of 1D Anomalous Diffusion In Fractured Nanoporous Media

Numerical Simulations of Two-Phase Flow in Rigid Porous Media

Multi-rate mass transfer modeling of two-phase flow in highly heterogeneous fractured and porous media

Assessment of Hydraulic Conductivity Upscaling Techniques and. Associated Uncertainty

Limits and Continuity. 2 lim. x x x 3. lim x. lim. sinq. 5. Find the horizontal asymptote (s) of. Summer Packet AP Calculus BC Page 4

Characterization of reservoir simulation models using a polynomial chaos-based ensemble Kalman filter

Core A-level mathematics reproduced from the QCA s Subject criteria for Mathematics document

Modeling seismic wave propagation during fluid injection in a fractured network: Effects of pore fluid pressure on time-lapse seismic signatures

Chapter 9: Differential Analysis

Numerical Treatment of Two-Phase Flow in Porous Media Including Specific Interfacial Area

Modeling Seepage Control in Hydraulic Structures

Numerical Algorithms as Dynamical Systems

LECTURE NOTES - III. Prof. Dr. Atıl BULU

Algebra and functions; coordinate geometry in the (x, y) plane; sequences and series; differentiation; integration; vectors.

ENGINEERING MATHEMATICS I. CODE: 10 MAT 11 IA Marks: 25 Hrs/Week: 04 Exam Hrs: 03 PART-A

Fracture-Matrix Flow Partitioning and Cross Flow: Numerical Modeling of Laboratory Fractured Core Flood

PRECONDITIONING MARKOV CHAIN MONTE CARLO SIMULATIONS USING COARSE-SCALE MODELS

Chapter 9: Differential Analysis of Fluid Flow

Pore-Scale Analysis of Dynamics of Two-Phase Flow. in Porous Media

Correlation Between Resistivity Index, Capillary Pressure and Relative Permeability

TENSOR RELATIVE PERMEABILITIES: ORIGINS, MODELING AND NUMERICAL DISCRETIZATION

CURRENT MATERIAL: Vector Calculus.

Vectors. January 13, 2013

Proceedings of the ASME nd International Conference on Ocean, Offshore and Arctic Engineering OMAE2013 June 9-14, 2013, Nantes, France

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

Investigating the Effect of Heterogeneity on Buckley-Leverett Flow Model

Introduction to Fluid Dynamics

Chapter 7. Seismic imaging. 7.1 Assumptions and vocabulary

uy Chavent, Jer^ome Jare and Sophie Jegou-Jan The generalized Darcy law says that for each phase the Darcy velocity is proportional to the gradient of

An Overly Simplified and Brief Review of Differential Equation Solution Methods. 1. Some Common Exact Solution Methods for Differential Equations

Math and Numerical Methods Review

Fraunhofer Institute for Computer Graphics Research Interactive Graphics Systems Group, TU Darmstadt Fraunhoferstrasse 5, Darmstadt, Germany

1 Fundamentals. 1.1 Overview. 1.2 Units: Physics 704 Spring 2018

Machine Learning Applied to 3-D Reservoir Simulation

AN UNFITTED METHOD FOR TWO-PHASE FLOW IN FRACTURED POROUS MEDIA

Advances in Locally Varying Anisotropy With MDS

Math 302 Outcome Statements Winter 2013

Analysis of a drainage efficiency in stratified porous media

Contents. Motivation. 1 di 7 23/03/ :41

ENGI Duffing s Equation Page 4.65

Energy minimization for the flow in ducts and networks

Directional Derivative and the Gradient Operator

Fractional flow in radial flow systems: a study for peripheral waterflood

Introduction to Aspects of Multiscale Modeling as Applied to Porous Media

Parameter Estimation in Reservoir Engineering Models via Data Assimilation Techniques

B024 RESERVOIR STREAMLINE SIMULATION ACCOUNTING

Multivector Calculus

Reservoir Simulator Practical

Steps in the Finite Element Method. Chung Hua University Department of Mechanical Engineering Dr. Ching I Chen

Transcription:

II. Dealing with Physical Laws: A Stochastic Flowpath Analysis Example In this chapter we will examine a physical problem where the significance of observing the underlying principles will be displayed. In this mapping application the role of the stochastic approach in physical problems is acknowledged, as well as the computational complexity that accompanies their numerical solution. The BME method has not been used in the stochastic flowpath analysis presented in this chapter; however, the way is paved for understanding of the value of stochasticity in physical problems, which is a key concept in BME. In addition to this, we will encounter some basic computational issues that will be present in our forthcoming BME applications. The area of interest in the following case is multiphase flow in random porous media. One of the most challenging problems of flow modelling in porous media is the solution of the partial differential equations (PDE) governing multiphase flow (see, e.g., Celia et al., 1993; Pankow and Cherry, 1996; Miller et al., 1998; Bensoussan et al., 1978). The main limiting factors in this problem are the heterogeneity of the flow parameters, the coupling of the flow equations for the different phases, and the nonlinearity of the constitutive relations. In addition, laboratory experiments are not sufficient for an understanding of the heterogeneities that influence flow at the field scale. Finally, field-scale simulators require averaging of the fluid properties within grid blocks of considerable sizes, thus implying an additional level of approximation (Ewing, 1989; Hassanizadeh and Gray, 1993).

In light of the above considerations, both analytical and numerical treatments of the multiphase flow equations are faced with serious complications. In the case of analytical solutions, these complications include singular Green's functions, complicated boundary and/or initial conditions, and inputs that vary in both space and time (Peyret and Taylor, 1983; McWhorter and Sunada, 1990). Most numerical methods introduce approximations and simplifications, the significance of which is not always clear or well-justified. In addition to that, they are usually very expensive to implement for two and three-dimensional domains, and often they do not provide a deeper understanding of the physical mechanisms (Allen, 1985; Kueper and Frind, 1991a and 1991b). An important aspect of multiphase flow in porous media is the inherent randomness that is evident in both the geometric structure of the porous formation and the fluid flow patterns (Papanicolaou, 1987a). Deterministic approaches (analytical and numerical) are not sufficient in this case (although they can be used to investigate individual realizations), but stochastic methods can incorporate randomness into multiphase flow modeling. Thus, they account for the lack of a complete characterization of heterogeneity (due to practically finite number of measurements) and provide estimates of the level of uncertainty involved in porous media characterization (Miller et al., 1998; Papanicolaou, 1987a and 1987b). In the stochastic context, multiphase flow equations can be studied: (a) by means of simulation methods that generate an ensemble of random field realizations (Essaid and Hess, 1993; Abdin et al., 1995); and (b) in terms of low-order perturbation expansions of the corresponding statistical moments (Chang et al., 1995). The macroscopic equations of flow in heterogeneous porous media are obtained by means of the representative elementary volume (REV) and the homogenization approach (Hornung, 1996). In the case of single-phase flow it can be shown rigorously that Darcy's 29

law is obtained from the microscopic Stokes equations by means of the homogenization approach (Allaire and Mikelic, 1996). The same techniques can be used in the case of twophase flow (Bourgeat, 1996) but, due to the complexities of the two-phase system, the derivation of macroscopic laws is based on heuristic arguments. Macroscopic equations for two-phase flow in randomly heterogeneous media have also been obtained by the approach of stochastic homogenization (Bourgeat et al., 1995). The work presented here is concerned with a stochastic flowpath (SFP) method for studying multiphase flow of immiscible fluids. Multiphase flow is governed by partial differential equations, in which the physical quantities are represented by means of random fields. The SFP method formulates the multiphase flow problem in terms of ordinary differential equations (ODEs) using concepts from differential geometry. The flow variables are determined by solving ODEs along the flowpaths, the trajectories of which are determined self-consistently by the flow equations. The SFP method incorporates the natural heterogeneity of flow in porous media, and it allows for randomness in the initial and boundary conditions. In addition, it does not require the approximations of perturbation methods, and it does not involve Green's functions. The second-order differential flowpath equations [see Eq.(2.11) below] satisfied by each phase are similar to the equations of twopoint seismic-ray tracing in an isotropic heterogeneous medium (Ben-Menahem, 1981; Jordan and Smith, 1987). The seismic-ray tracing problem has been solved in Ben- Menahem (1981) by transforming the equations into a system of first-order ODEs in an expanded space and using a finite difference scheme. In the case of the two-phase problem, the equations are further complicated by the presence of the gravitational term and the coupling of the two phases by means of the nonlinear relative permeability relations. While the SFP method bears some resemblance to the streamtube (Lenormand, 1995) and streamline (Blunt et al., 1996; Thiele et al., 1996) approaches for solving 30

multiphase flow equations, it differs from them in certain important ways. These approaches, which are very efficient computationally, provide numerical solutions for the multiphase system assuming fractional flow conditions. In this case the capillary pressure is ignored, thus leading to equations for fractional flow variables which depend only on the saturation of the respective phase (Lake, 1989). In contrast, the SFP approach proposed here does not neglect the capillary pressure. The streamtube/streamline simulators achieve additional improvements in computational efficiency by ignoring the gravitational force, and assuming uniform initial conditions. These assumptions permit mapping analytical solutions along streamlines, using the explicit Buckley-Leverett expressions (Lake, 1989; Leverett, 1941). Since our approach determines the flowpaths numerically, it admits nonuniform initial conditions. In addition, the streamline method calculates a single streamline corresponding to the total velocity operator, and then solves the balance equations for each phase along this streamline. The SFP approach, on the other hand, focuses on calculating the flowpaths for all phases. 2.1 Steady-state two-phase flow In the following we study multiphase flow of two immiscible and incompressible fluids within a heterogeneous porous medium, the physical variables of which are represented by means of stochastically homogeneous random fields (Christakos, 1992). In addition, we consider only strong mixing random fields with finite correlation lengths. The permeability sample functions are assumed to be continuous and differentiable. Furthermore, permeability fields that are non-zero everywhere are considered. The last condition is satisfied by log-normal permeability distributions that are often used to model intrinsic permeability. Mathematical formulations based on the log-permeability of the porous medium are commonly used for single-phase flow (Dagan, 1989) and multiphase flow (Douglas Jr. et al., 1995), as well. The flow equations for the phases under steady-state conditions follow from the local Darcy's law for each phase (Bourgeat, 1996) and the 31

incompressibility condition q a (s) = 0 (Miller et al., 1998), where the subscript a is the phase index and q a is the flow vector of the phase a. The incompressibility condition can be expressed as follows j a (s) + ln K a (s) m a j a (s) = 0 (2.1) where s = (s 1,s 2,s 3 ) denotes a point within the flow domain. The piezometric pressure gradient j a is given by j a (s) = - p a (s) - r a gˆ x 3 (2.2) where p a is the fluid pressure, r a is the density of the phase a, and ˆ x 3 is the unit vector in the vertical direction. The spatial random fields K a (s) represent the intrinsic permeabilities of the phases given by Chavent and Jaffre (1986) and Muskat (1949). K a (s) = k ra (s)k(s) (2.3) where the k ra denote the relative permeabilities, and K(s) denotes the intrinsic permeability of the porous medium that is experimentally determined (Chavent and Jaffre, 1986; Dullien, 1992). Finally, the symbol m a will be used in this chapter to denote the dynamic viscosity coefficients of the phases. The formulation of the incompressibility equation in terms of the log-permeability random fields is standard in stochastic hydrology; it is motivated by the fact that the experimental permeability distributions are often lognormal. In reality, the subsurface is an anisotropic medium, and thus a scalar K(s) is a poor assumption for deterministic multiphase models. In stochastic models, however, the anisotropy of the medium is expressed by means of the permeability correlation function (Hristopoulos and Christakos, 1997). Of course, this implies that the anisotropic effects become evident only 32

after ensemble averaging. Extensions of the present model to anisotropic tensor K(s) are also possible. This can be done easier if the SFP theory is formulated in terms of the phase permeabilities instead of their logarithms. Eq.(2.1) should be solved with respect to j a (s) subject to the boundary conditions and the constitutive relations (2.3). The capillary pressure p c (s) is given by the difference between the non-wetting phase (nw) and the wetting phase (w) p c (s) = p nw (s) - p w (s) (2.4) The system of Eqs.(2.2)-(2.4) is usually closed with empirical pressure-saturation ( p - S) relations. These relations are based on the assumption that the capillary pressure depends only on the relative saturations of the two fluids, denoted by S w (s) and S nw (s). These functions represent the fraction of the fluid volume in each phase over the total fluid volume; thus, S nw (s) + S w (s) =1 everywhere. Various p - S relations are in use for different types of fluids and porous media. Specific examples will be presented in a later section. The p - S relation is generally expressed as p c (s) =F c [ S w (s)] (2.5) It is usually assumed, neglecting hysteresis effects, that F c is a non-linear, monotone function of S w (Bourgeat, 1996). F c is a decreasing function of the wetting fluid saturation if p c is defined by Eq.(2.4); if p c is defined as p c = p w - p nw, then F c is a increasing function of the wetting fluid saturation. The relative permeabilities of the wetting and the non-wetting phase are given as functions of the wetting-fluid saturation by means of the k - S relations 33

k r,w (s) = f w [ S w (s)] (2.6) and k r,nw (s) = f nw [ S w (s)] (2.7) While f w is an increasing function of the wetting-phase saturation S w, the f nw is a monotonically decreasing function (Chavent and Jaffre, 1986). The p - S expression (2.5) and the k - S expressions (2.6) and (2.7) supplement the constitutive relations (2.3). The above set of equations is not unique, and it is worth noting that hysteretic constitutive relations have also been observed (Dullien, 1992). However, the specific choice of constitutive relations does not influence the general formulation of the SFP approach introduced in this work, at least if hysteresis is ignored. Eqs.(2.1)-(2.7) constitute the governing system of the two-phase flow equations; the pressure gradients are coupled via the constitutive relations that introduce a nonlinear dependence on saturation. 2.2 Geometric description of flowpath trajectories The SFP analysis is based on a transformation of the multiphase flow system into a set of first-order differential equations that provide a geometric representation of the flow (Figure 2.1). While in the conventional analysis the pressure gradients at the nodes of the numerical grid are considered as the outputs of the multiphase flow system, in the SFP analysis the pressure gradients and the coordinates of the a-flowpaths ( a-fp) are the outputs of the flow system. Naturally, the a-fp trajectories depend upon the heterogeneity of the medium and the physical properties of the fluid phases. 34

BC s a (Cartesian coordinates) Æ (a) Multiphase Flow Equations Æ j a BC Æ Multiphase Flow Equations Æ s a (a - FP coordinates) Æ j a (b) Figure 2.1: (a) Conventional analysis vs. (b) SFP analysis of the multiphase flow system (BC: Boundary conditions, FP: flowpath) (from Christakos et al., 1998). In the SFP method the pressure gradients j a (s) are analyzed into the scalar z a (s) and the direction vector e a (s) as follows j a (s) = z a (s)e a (s) (2.8) The vector e a (s) determines the direction of the a -FP trajectory and its unit Euclidean norm is equal to one, i.e., e a (s) =1. According to the definition (2.8) the a -FP curves coincide with the streamlines (Bear, 1988) of each phase (i.e., the curves that are at every point tangent to the instantaneous phase velocity) for a medium with a scalar, though possibly statistically anisotropic, saturated permeability.the flowpaths do not necessarily coincide with the pathlines (Bear, 1988) (i.e.,the loci of the positions of the fluid particles in space as time progresses), in the case of unsteady flow. 35

a - FP P s a s 3 s 1 s 2 Figure 2.2: An a-fp in R 3. A parametrized representation of the a-fp s a (l) in R n, (where l Œ U à R 1 is a parameter that varies along the a-fp) is obtained in terms of the following mapping (Nutbourne and Martin, 1988): l Æ s a (l) = ( s 1a (l),...,s na (l)) Œ R n To avoid redundancy we replace s a with s when it is clear which phase is denoted by s, e.g., z a (s). Equivalently, the direction of the a-fp at each point in space is identified in terms of n -1 independent parameters g i,a (i =1,..,n -1). An illustration is given in Figure 2.2 for the case of a three-dimensional Euclidean space ( n =3). A set of different values for the parameters g i,a leads to a family of flowpaths (Figure 2.3). Different choices of the parameters l and g i,a (i =1,..,n -1) are possible. A natural choice is l = l, where l measures the total path length measured from an initial point s 0 at the boundary (Figure 2.4). The parameters g i,a usually represent the angles of the direction vector with respect to the axes of the fixed coordinate frame, i.e., the polar and the azimuthal angles in R 3. 36

s a 0 (1) s a 0 (2) g 2, a g 1, a (1), g 2, a (1) g 1, a (2), g 2, a (2) g 1, a Figure 2.3: A family of a-fps in R 3. s 3 q 3,a (s) dl s 0 FP s(l a ) s e a (s) q 1, a ( s) q 2, a (s) s 2 f a (s) s 1 Figure 2.4: A parametrized a-fp in R 3. In a fixed, Cartesian frame the direction vector e a (s) is given by e a (s) = Â 3 e i=1 i,a (s) ˆ x i where: ˆ x i are the unit vectors along the axes, and e i,a (s) are the direction cosines of the a-fp 37

e i,a (s) = cosq i,a (s) = ds i,a (l) dl (2.9) where dl denotes the arc length of an infinitesimal element of the a-fp at point s along the tangential direction e a (s), so that dl 2 2 = ds i,a. This leads to the constraint  n i=1 n ds i,a 2 2  =  e dl 2 i,a =1 (2.10) i=1 n i=1 Note that along an a-fp the direction vector e a (s) depends on a single parameter, i.e., the path length l. The one-dimensional derivative of a function f (s) along the flowpath is given df (s) by = f (s) e a. Using the geometric representations above, it can be shown that the dl rate of change of the pressure gradient along the a-fp is given by (see Appendix A) d Ê dl z (s) ds ˆ a Á a = z Ë dl a (s) + r a g e 3,a (s) (2.11) Note that according to Eq.(2.11) all coordinates of the flowpaths are influenced by the gravitational force by an amount proportional to the rate of change of e 3,a in the specific direction. 2.3 Flowpath formulation of two-phase flow In this section we develop a space transformation (Christakos, 1992) representation of the multiphase flow along the a-fp for the case of the initial value problem. In view of Eqs. (2.8) above, the incompressibility conditions Eqs.(2.1) can be expressed by means of stochastic ordinary differential equations as follows 38

dz a (s) dl + [ e a (s) + d Ê dl ln K (s) ˆ a Á ] z a (s) = 0 (2.12) Ë m a Note that neighboring flowpaths of each phase are coupled via the term that involves the divergence of the direction vector e a. The flux along the flowpath is defined by q a (s) = K (s) a z a (s)e a (s), and the scalar m a flux q a (s) is denoted by q a (s). For the initial value problem the general solution of Eq.(2.12) along the a-fp then is q a (s) = q a (s 0 )exp[-h a (s)] (2.13) where h a (s) is the space transformation of the divergence of e a along the a -FP, given by h a (s) = h a (l) = s a (l ) Ú e a ( s )d l (2.14) s a (0) In the special case that the a -FP is a straight line in a two-dimensional domain, Eq.(2.14) is the standard Radon transform (Helgason, 1980). More generally, if the a-fp is an arbitrary curve, Eq.(2.14) defines a non-linear extension of the Radon transform (Christakos et al., 1998). An equivalent expression for h a (s) is as follows h a (s) = n  i=1 s i,a (l ) Ú s i,a (0) 1 e i,a (s u i ) du i (2.15) e i,a (s u i ) u i where the (s u i ) denotes the vector ( s 1,...,s i-1,u i,s i+1,...,s n ). Eqs.(2.13)-(2.15) provide closed-form expressions for z a in terms of the initial conditions z a (s 0 ), the permeability 39

field K a and the space transformation h a. However, explicit solutions for the z a require the flowpath equations. The incompressibility Eqs.(2.12) are complemented by a set of equations that determine the flowpaths. In particular, Eqs.(2.11) lead to the following 2 n first-order, non-linear differential equations d [ dl z (s)cosq (s) a i,a ] = z (s) a + r a g cosq (s) 3,a, i =1,...,n. (2.16) s i s i Note that for n >1 the variables q i,a are not independent, since they are constrained by  n i=1 q 2 i,a =1. Accordingly, only 2 (n -1) of the Eqs.(2.16) are independent. Eqs.(2.12) and (2.16) form the SFP system, which is closed by the constitutive relationships (2.3)- (2.7) and subject to the initial and boundary flow conditions. Since the p - S relation involves the pressure difference of the two phases, additional relations are required relating the phase pressures to the pressure gradients and the flowpaths; these are obtained by integrating Eq.(A.2) in Appendix A as follows p a (s) = p a (s 0 ) - s a (l ) Ú d l z a ( s ) + r a ge 3,a ( s ) s a (0) [ ] (2.17) Either deterministic or random flow conditions can be considered with the SFP method; the type of boundary conditions is expected to affect the flow solutions as shown in Oliver and Christakos (1996) in the case of single-phase flow. Finally, note that another useful formulation of the multiphase flow equations along the a-fp is obtained in terms of the Jacobian of the transformation from the Cartesian to the flowpath coordinates (Christakos et al., 1998). 40

2.4 Two-phase flow in a layered heterogeneous medium In this section we consider flow in a porous medium with uniaxial heterogeneity in the vertical direction. This scenario applies, e.g., to layered media that exhibit homogeneous intrinsic permeabilities within the layers with variations only along s 3. In this case, Eq.(2.14) that determines the variation of e a between an initial point s 0 and an endpoint s f of an a-fp leads to the algebraic expression Ê h a (s f ) = ln cosq (s ) ˆ 3,a f Á (2.18) Ë cosq 3,a (s 0 ) Eq.(2.13) governs the variation of the flux along the flowpath. In view of Eqs.(2.13) and (2.18) the pressure gradient at the endpoint is determined by z a (s f ) z a (s 0 ) = K (s ) a f cosq 3,a (s 0 ) K a (s 0 ) cosq 3,a (s f ) (2.19) The flowpath equations (2.16) give z a (s)cosq i,a (s) = c i, for i =1,2. It then follows that the magnitude of the pressure gradient is z a (s) = j^,a sinq 3,a (s) where j^,a = c 1 2 + c 2 2. In view of this relationship, Eq.(2.19) leads to tanq 3,a (s f ) K a (s f ) = tanq 3,a (s 0 ) K a (s 0 ) (2.20) Therefore, the flowpaths are determined by the initial conditions and the intrinsic permeabilities. The pressures along the flowpaths are given by the integral in Eq.(2.17). Thus, we obtain the following non-linear system of integral equations for the phase pressures 41

p a (s f ) = p a (s 0 ) - r a gˆ x 3 (s f - s 0 ) - s f Ú d s 3 s 0 m a q 3,a (s 0 ) K a ( s )cosq 2 3,a ( s ) (2.21) The intrinsic permeabilities depend implicitly on the saturation, and thus the integral can not be directly evaluated. In fact, Eqs.(2.20) and (2.21) along with the appropriate k - S relationships and the p - S relation form a system of 3a +1 equations containing an equal number of unknowns, e.g., p a,q 3,a,K a, and S e. Note that the differential equations (2.12) and (2.16) have been replaced with the algebraic equations (2.19) and (2.20). Finally, in the case of saturated, single-phase flow, Eq.(2.20) yields the well-known law of streamline refraction (Bear, 1988). 2.5 Waterflooding of a heterogeneous oil reservoir In the following we consider injection of water (wetting phase) in a two-dimensional, heterogeneous oil reservoir. We consider boundary conditions for the saturation coefficient and the pressure gradients [magnitude z a (s 0 ) and incidence angles q 1,a (s 0 )] of the a-fps given along a boundary line of a square grid. The gravitational term is ignored in two dimensions. The two-phase flow satisfies Eqs.(2.12) and (2.16). The intrinsic logpermeability, lnk s, is modeled using stochastically homogeneous random fields (Christakos, 1992). We use the Brooks-Corey relations for the p - S and the k - S curves. These are expressed in terms of the reduced water saturation S e (s), defined by S e (s) = S w (s) - S w,res S w,sat - S w,res (2.22) where S w,res denotes the residual and S w,sat the maximum saturation of the wetting phase. The reduced saturation takes values in [0,1]. Correspondingly, the reduced oil saturation is 42

given by 1- S e (s). Explicit expressions for the relative permeabilities of the oil-water system are given by the following k - S relations (Bear, 1988): Relative Permeabilities 0 0.2 0.4 0.6 0.8 <--k o k w --> 0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 1 S e Figure 2.5: Plots of the Brooks-Corey k - S curves for l =1 (dashes), l=8 (dots). l=4 (solid line), and 0 k rw (s) = k rw [ S e (s)] (2+3l) l (2.23) and 0 k ro (s) = k ro (2+l) [ 1- S l e (s)] [ 1- S e (s)] 2 (2.24) 43

Capillary Pressure (p c /p e ) 1 1.5 2 2.5 3 3.5 4 0 0.2 0.4 0.6 0.8 1 S e Figure 2.6: Plots of the p - S curves for l=1 (dashes), l=4 (solid line), and l=8 (dots). where k 0 0 ro, k rw and l are empirical parameters that characterize the medium. The relation between the capillary pressure and the water saturation is modeled by means of S e (s) = [ p c (s) p e ] -l, if p c > p e (2.25.1) and S e (s) =1, if p c p e (2.25.2) 44

where p c denotes the capillary pressure at maximum water saturation. Plots of the k - S relations are shown in Figure 2.5 for three different values of the parameter l. For a fixed l, the k rw is an increasing function of S e while the k ro is decreasing. For a fixed S e, the k rw is an increasing function of l while the k ro is decreasing. The p - S behavior is shown in Figure 2.6: The p - S curve is a monotonically decreasing function of S e. In addition, for a given S e the capillary pressure decreases with increasing l. In the following, k 0 ro =0.9, k 0 rw =0.5, S w,res =0.1, S w,sat =0.9. The dynamic viscosity of the water is m w = 5 10-3 poise, and the viscosity of oil m o =1 10-1 poise. The pressure is expressed in units of p e. The complete system of two-phase flow equations involves two incompressibility conditions given by Eqs.(2.12), two independent flowpath equations as shown in Eqs.(2.16), four constitutive relationships from Eqs.(2.4) and (2.21)-(2.23), and two pressure integrals provided by Eqs.(17). Thus, the total number of equations is ten matching the number of unknowns: z a, q 1,a, p a, k r,a ( a = w and o), S e and p c. The above two-phase flow system is investigated by means of numerical simulations. 2.6 Numerical solutions The variables of the flowpath system are the pressure gradients and the direction of the flowpaths that correspond to the two phases. The flowpath equations represent the evolution of each phase along different flowpaths; however, the relative permeability of the phases at every point depends on the capillary pressure, and thus on the pressure of the other phase at this point. This coupling of the flowpaths prevents a straightforward solution of the ODEs by e.g., Runge-Kutta or other first-order nonlinear ODE solvers. Instead, we approximate the ODEs by means of difference equations that we solve using the Newton-Raphson method (e.g., Press et al., 1997). The saturation is, also, an unknown variable to be determined from the solution of the flowpath system. 45

We propose to solve the steady-state flow equations using an iterative numerical approach that assumes a profile S e,0 (s) as an initial guess for the saturation coefficient. This determines the initial capillary pressure and the intrinsic permeabilities, and a reduced system of four first-order nonlinear differential equations is obtained -- which includes the incompressibility conditions Eqs.(2.12) and the independent flowpath equations (2.16). Solution of these equations yields the oil and water pressure gradients and the local direction of the flowpaths throughout the entire domain. The oil and water pressure along the flowpaths are evaluated by the discretized integral of the pressure gradients in Eqs.(2.17), and an updated capillary pressure profile is derived by the difference between the oil and the water pressure; based on this an updated saturation profile is obtained via Eq.(2.23). The reduced system is solved iteratively until a stable saturation profile is obtained. This is determined by achieving convergence on more than 98% of all the included nodes. The steady-state solution should depend on the conditions at the injection boundary. It should be independent of the initial guess of the saturation profile. However, the convergence of the Newton-Raphson method in general depends on the initial guess. Experimenting with different forms of the initial profile (guess) has shown that there is a layer near the boundary within which the saturation rises from the boundary value towards one (fully water-saturated medium). Outside this layer, there are regions of trapped oil with lower water saturation. Different functional dependences for the initial saturation have been tried: a constant profile, linear increase, and exponential increase towards a maximum saturation. The method was found to converge fastest when the latter profile is used as an initial guess. In addition, the saturation towards which the solution converges is independent of the maximum saturation of the initial profile, at least within the range of values that were 46

investigated. This method accomplishes a decoupling of the flow equations that allows their solution independently along the flowpaths. We use a numerical grid with 101 101 nodes and the distance between nodes is taken equal to 1 m. The position on the grid is here denoted by s = (x,y). The saturated permeability K s (s) is assumed to be log-normally distributed with arithmetic mean K s =10-9 m 2 and coefficient of variation (COV) m K = s K K s =0.2. A Gaussian covariance model is assumed for the log-permeability fluctuations f (x, y), i.e., Ê c f (s) = s 2 f exp{- x 2 2 b + y 2 ˆ Á 2 Ë x b } (2.26) y We investigate two cases, an isotropic correlation model with b x = b y =20m, and an anisotropic model with b x =20m and b y =40m. The initial conditions include the directions of the pressure gradients, q a (s 0 ) = u a,s0, where u a,s0 is a zero-mean, uniformly distributed random noise, the magnitude of the pressure gradients z nw (s 0 ) =0.01 and z w (s 0 ) =0.03, and the boundary saturation S e =0.745. The parameter l in the Brooks-Corey relations is assumed to have the value l =0.45 in both cases. The initial guess for the saturation profile increases exponentially from the boundary and approaches a maximum level of 0.92. We discretize the equations using finite differences along the flowpaths. The length of the flowpaths is increased by the finite step dl a. Here we consider dl o = dl w = dl. The size of the step is bounded below by the grid resolution and above by the correlation length, i.e., e dl << x, where e is the lattice step and x is the correlation length of the permeability fluctuations. We use forward finite differences for the derivatives. For function evaluations the Trapezoidal rule (see Appendix C) has been employed due to the step-wise advance of the flowpaths. The length increments in the orthogonal directions of the grid are given by 47

dx a,m x a,m +1 - x a,m = dl ( 2 cosq + cosq a,m +1 a,m) (2.27a) dy a,m y a,m +1 - y a,m = dl ( 2 sinq + sinq a,m +1 a,m) (2.27b) where the index m = 0,1,...,N f,i denotes the m-th point on the flowpath. The integer N f,i denotes the total number of increments in the i-th flowpath. In view of Eq.(2.25), Eq.(2.12) is discretized as follows (z a,m +1 -z a,m ) + [(F s,m +1 - F s,m ) + (lnk r,a,m +1 - lnk r,a,m ) - (q a,m +1 -q a,m )w(q a,m,q a,m +1 )] z a,m + z a,m +1 Ê Á Ë 2 ˆ = 0 (2.28) where F m = lnk m (s). The function w(q a,m,q a,m +1 ) depends only on the local direction of the flowpaths and is given by w(q a,m,q a,m +1 ) = sinq a,m +1 + sinq a,m cosq a,m +1 + cosq a,m - cosq a,m +1 + cosq a,m sinq a,m +1 + sinq a,m (2.29) Eqs.(2.16) lead to the following difference equations z a,m +1 cosq a,m +1 -z a,m cosq a,m = 2(z a,m +1 -z a,m ) cosq a,m +1 + cosq a,m (2.30) and z a,m +1 sinq a,m +1 -z a,m sinq a,m = 2(z a,m +1 -z a,m ) sinq a,m +1 + sinq a,m (2.31) Note that when q a @ 0 the denominator of the right-hand-side term of Eq.(2.31) tends to zero; this indicates that the flowpath is almost parallel to the x-axis, so that the flowpath 48

variables do not vary along the y-axis. Similarly, the denominator of Eq.(2.30) tends to zero around q a @ p 2, i.e., when the flowpath is almost parallel to the y-axis. Hence, when mod(q a,p 2) p 4 (where 'mod' denotes the modulo function), Eq.(2.30) is automatically selected, while Eq.(2.31) is used when mod(q a,p 2) > p 4. Relative Saturation Change 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 isotropic anisotropic 0 10 20 30 40 50 60 70 Newton-Raphson Iterations Figure 2.7: Plot of the relative change in the saturation profile between consecutive Newton- Raphson iterations. 49

Based on this process, the non-linear system of either Eqs.(2.28) and (2.30) or Eqs.(2.28) and (2.31) is appropriately solved by Newton-Raphson. The non-linear solver makes also use of the discretized derivatives of the equations with respect to each of the variables. The derivatives are directly obtained from the non-discretized system equations. Mean Saturation 0.89 0.9 0.91 0.92 0.93 0.94 0.95 0.96 Isotropic Anisotropic 0 10 20 30 40 50 60 70 Newton-Raphson Iterations Figure 2.8: Plots of the mean water saturation in the domain midpoint vs. the number of Newton-Raphson iterations. Eqs.(2.28)-(2.31) involve the unknown variables q a,m +1 and z a,m +1. The q a,m and z a,m values are obtained from the previous step or the initial conditions at the injection boundary. The log-permeability field is known on the grid. Since the traces of the flowpaths 50

do not pass from the grid nodes, estimates of the log-permeability at off-grid points are required. In general, such estimates can be derived by a linear kriging estimator (Christakos, 1992). However, if the correlation length significantly exceeds the length increment, i.e., dl << x, the log-permeability off the grid is accurately estimated by the arithmetic mean of the permeability at the four nearest grid points. Fraction of Updated Points 0.1 0.15 0.2 0.25 0.3 0.35 Isotropic Anisotropic 0 10 20 30 40 50 60 70 Newton-Raphson Iterations Figure 2.9: Plot of the fraction of the grid points at which the saturation updated vs. the number of Newton-Raphson iterations. 51

We denote by N m * the number of updated nodes at the m-th iteration. To estimate the relaxation of the saturation profile we use the relative change in the saturation coefficient DS e,m defined by means of N m * DS e,m = 1 S (s ) - S (s e,m i e,m-1 i * Â ) N m S e,m-1 (s i ) i=1 (2.32) A plot of DS e,m vs. the Newton-Raphson iteration number is shown in Figure 2.7. At the end of the Newton-Raphson cycle the relative saturation change is below 5% for 99.9% of the updated points, in both the isotropic and anisotropic cases. In Figure 2.8 we plot the mean saturation of the domain vs. the number of iterations. The percentage of nodes where the saturation is updated is plotted vs. the number of iterations in Figure 2.9. It is seen that this quantity essentially relaxes to a value about 0.16 after 20 iterations, although significant fluctuations persist. The resulting isopressure contours for the isotropic cases are shown in Figures 2.10 (water) and 2.11 (oil). The constant saturation contours are shown in Figure 2.12. The code generates the pressures at the nodes visited by the flowpaths, e.g., 7792 nodes for the water phase and 2622 nodes for the oil phase in the isotropic case. These numbers are similar in the anisotropic case. We use a kriging estimator to obtain the pressures at the nodes that are not visited by the flowpaths. These values are used for generating the contours in Figures (2.10)-(2.12). The water-phase pressure contours closely resemble single-phase flow. This is expected, since after water injection most of the reservoir is flooded with water except for areas of trapped oil. In the saturation contour plot (Figure 2.12) there is a narrow layer near the boundary within which the water saturation orderly increases towards a maximum close to 0.99. 52

Figure 2.10: Wetting phase isopressure contours for the isotropic intrinsic permeability field (distance in m, pressure in units of the entry pressure p e ). 53

Figure 2.11: Oil phase isopressure contours for the isotropic intrinsic permeability field (distance in m, pressure in units of the entry pressure p e ). 54

Figure 2.12: The reduced water saturation contours for the isotropic intrinsic permeability field (distance in m). 55

Figure 2.13: Wetting phase isopressure contours for the anisotropic intrinsic permeability field (distance in m, pressure in units of the entry pressure p e ). 56

Figure 2.14: Oil phase isopressure contours for the anisotropic intrinsic permeability field (distance in m, pressure in units of the entry pressure p e ). 57

Figure 2.15: The reduced water saturation contours for the anisotropic intrinsic permeability field (distance in m). 58

In the anisotropic case the water phase pressure pattern (shown in Figure 2.13) does not exhibit any significant differences from the isotropic case. The same is true of the saturation contours (shown in Figure 2.15). This indicates that the anisotropy of the intrinsic permeability is not a significant factor in determining the residual distribution of oil. On the other hand, the oil isopressure contours for the anisotropic intrinsic permeability field, shown in Figure 2.14, exhibit significant changes with respect to the isotropic case (this behavior however, may be an artifact, due to the relatively low number of nodes where the oil pressure is calculated). 2.6 Transient effects in multiphase flow In the transient case, the dynamic flow equations for multiphase flow are given by f S a (s,t) + j a (s,t) + [ ln K (s,t) a ] j a (s,t) = 0 (2.33) K a (s,t) t m a where f denotes the porosity of the medium and S a the saturation of each phase. In view of Eq.(2.33), Eq.(2.12) is replaced by f S a (s,t) + z (s,t) a K a (s,t) t l + [ e a (s,t) + Ê l ln K (s,t) ˆ a Á ] z a = 0 (2.34) Ë m a The flowpaths develop a time dependence induced by the change of the saturation field (Christakos and Hristopulos, 1998). The SFP system should be then solved subject to an initial saturation profile. 59

2.7 A few remarks on the stochastic flowpaths method A SFP method was presented in this chapter for studying multiphase flow in heterogeneous porous formations, that is based on established notions of differential geometry. Steadystate conditions have been considered in the previous, but the SFP method is applicable in the case of transient flow as well. Other techniques based on the idea of studying multiphase flow along flowpaths (or streamlines) have also been recently introduced in petroleum engineering (Blunt et al., 1996; Thiele et al., 1996). These techniques, however, address the solution of the simpler fractional flow model often used in oil recovery. In particular, the fractional flow model is based on the high flow-velocity approximation that neglects the capillary pressure; the multiphase flow equations can then be formulated in terms of a total velocity that is determined from an effective Darcy's law. The SFP method is considerably more general than these techniques. Indeed, the SFP method allows for nonzero values of the capillary pressure. It can, also, incorporate various sets of constitutive relationships that are of interest in environmental sciences. In principle, hysteretic p - S relations can be accommodated, as well. In the computational field, a numerical algorithm based on discretization of the flowpath equations and the Newton-Raphson method has been implemented for twodimensional flow. An iterative approach has been used to update an initial saturation profile progressively until the saturation relaxes to a stable value. In the following chapters we will have to deal with non-linear equation systems that will feature a more complex form than Eqs.(2.16). In the light of the increased numerical implications in the coming investigations, Newton-based non-linear solvers proved to be the preferred techniques for application. The problem that was presented in this chapter provided us with valuable experience in the domain of dealing with non-linear systems of equations that will be of use in the following. 60