advances in water resources · element (mfe) method which provides, in addition to the gridcell...

15
An efficient numerical model for incompressible two-phase flow in fractured media Hussein Hoteit a,1 , Abbas Firoozabadi a,b, * a Reservoir Engineering Research Institute, Palo Alto, CA, USA b Yale University, New Haven, CT, USA article info Article history: Received 1 October 2007 Received in revised form 1 February 2008 Accepted 19 February 2008 Available online 7 March 2008 Keywords: Two-phase flow Water injection Fractured media Heterogeneous media Capillary pressure Discrete fractured model Mixed finite element method Discontinuous Galerkin IMPES methods abstract Various numerical methods have been used in the literature to simulate single and multiphase flow in fractured media. A promising approach is the use of the discrete-fracture model where the fracture enti- ties in the permeable media are described explicitly in the computational grid. In this work, we present a critical review of the main conventional methods for multiphase flow in fractured media including the finite difference (FD), finite volume (FV), and finite element (FE) methods, that are coupled with the dis- crete-fracture model. All the conventional methods have inherent limitations in accuracy and applica- tions. The FD method, for example, is restricted to horizontal and vertical fractures. The accuracy of the vertex-centered FV method depends on the size of the matrix gridcells next to the fractures; for an acceptable accuracy the matrix gridcells next to the fractures should be small. The FE method cannot describe properly the saturation discontinuity at the matrix–fracture interface. In this work, we introduce a new approach that is free from the limitations of the conventional methods. Our proposed approach is applicable in 2D and 3D unstructured griddings with low mesh orientation effect; it captures the satura- tion discontinuity from the contrast in capillary pressure between the rock matrix and fractures. The matrix–fracture and fracture–fracture fluxes are calculated based on powerful features of the mixed finite element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure discontinuities at impermeable faults in a simple way. To reduce the numerical dispersion, we use the discontinuous Galerkin (DG) method to approximate the sat- uration equation. We take advantage of a hybrid time scheme to alleviate the restrictions on the size of the time step in the fracture network. Several numerical examples in 2D and 3D demonstrate the robust- ness of the proposed model. Results show the significance of capillary pressure and orders of magnitude increase in computational speed compared to previous works. Ó 2008 Elsevier Ltd. All rights reserved. 1. Introduction Modeling of multiphase fluid flow in fractured media is of inter- est in many of the environmental and energy problems. Examples include tracing a NAPL migration, radioactive waste management in the subsurface, and the enhanced oil recovery in naturally frac- tured reservoirs. Geological characterization and multiphase flow simulation are challenging tasks in complex fractured media. Var- ious conceptual models have been used to describe multiphase flow in fractured media. The main models include the dual-poros- ity and its variations, single-porosity, and the discrete-fractured approaches. The dual-porosity model has traditionally been used to simulate the flow in fractured hydrocarbon reservoirs [1–6]. In this model, the matrix–fracture mass transfer is described by empirical func- tions that incorporate ad hoc shape factors which are based on the- oretical derivations in single phase. In two-phase with capillary and gravity, there is no theory for the calculation of shape factors despite the fact that the accuracy of flow calculations depend on their values. The problem may be even more fundamental. The driving force in the expression of the transfer function in two- phase flow may not have a sound basis. We should point out that advanced variations of the dual-porosity model such as the multi- ple interaction continua method [7] have better accuracy and fea- tures than the conventional model. A more accurate and physics-based approach is the single- porosity model that describes fractures explicitly in the medium. This model requires the fracture-network characterization and the fracture–fluid interaction functions. Modeling fractures with the single-porosity model can be classified as a complex case of heterogeneous porous media, where the fracture gridcells are trea- ted similarly to the matrix gridcells. The implementation of this model is, however, almost impossible in field-scale problems because of excessive fine gridding and severe restrictions on the time steps due to the contrast in geometrical scales between the 0309-1708/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.advwatres.2008.02.004 * Corresponding author. Address: Reservoir Engineering Research Institute, Palo Alto, CA, USA. E-mail address: [email protected] (A. Firoozabadi). 1 Now at ConocoPhillips Co. Advances in Water Resources 31 (2008) 891–905 Contents lists available at ScienceDirect Advances in Water Resources journal homepage: www.elsevier.com/locate/advwatres

Upload: others

Post on 23-Mar-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Advances in Water Resources 31 (2008) 891–905

Contents lists available at ScienceDirect

Advances in Water Resources

journal homepage: www.elsevier .com/ locate/advwatres

An efficient numerical model for incompressible two-phase flow in fractured media

Hussein Hoteit a,1, Abbas Firoozabadi a,b,*

a Reservoir Engineering Research Institute, Palo Alto, CA, USAb Yale University, New Haven, CT, USA

a r t i c l e i n f o

Article history:Received 1 October 2007Received in revised form 1 February 2008Accepted 19 February 2008Available online 7 March 2008

Keywords:Two-phase flowWater injectionFractured mediaHeterogeneous mediaCapillary pressureDiscrete fractured modelMixed finite element methodDiscontinuous GalerkinIMPES methods

0309-1708/$ - see front matter � 2008 Elsevier Ltd. Adoi:10.1016/j.advwatres.2008.02.004

* Corresponding author. Address: Reservoir EngineAlto, CA, USA.

E-mail address: [email protected] (A. Firoozabadi).1 Now at ConocoPhillips Co.

a b s t r a c t

Various numerical methods have been used in the literature to simulate single and multiphase flow infractured media. A promising approach is the use of the discrete-fracture model where the fracture enti-ties in the permeable media are described explicitly in the computational grid. In this work, we present acritical review of the main conventional methods for multiphase flow in fractured media including thefinite difference (FD), finite volume (FV), and finite element (FE) methods, that are coupled with the dis-crete-fracture model. All the conventional methods have inherent limitations in accuracy and applica-tions. The FD method, for example, is restricted to horizontal and vertical fractures. The accuracy ofthe vertex-centered FV method depends on the size of the matrix gridcells next to the fractures; for anacceptable accuracy the matrix gridcells next to the fractures should be small. The FE method cannotdescribe properly the saturation discontinuity at the matrix–fracture interface. In this work, we introducea new approach that is free from the limitations of the conventional methods. Our proposed approach isapplicable in 2D and 3D unstructured griddings with low mesh orientation effect; it captures the satura-tion discontinuity from the contrast in capillary pressure between the rock matrix and fractures. Thematrix–fracture and fracture–fracture fluxes are calculated based on powerful features of the mixed finiteelement (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcellinterfaces and can readily model the pressure discontinuities at impermeable faults in a simple way. Toreduce the numerical dispersion, we use the discontinuous Galerkin (DG) method to approximate the sat-uration equation. We take advantage of a hybrid time scheme to alleviate the restrictions on the size ofthe time step in the fracture network. Several numerical examples in 2D and 3D demonstrate the robust-ness of the proposed model. Results show the significance of capillary pressure and orders of magnitudeincrease in computational speed compared to previous works.

� 2008 Elsevier Ltd. All rights reserved.

1. Introduction

Modeling of multiphase fluid flow in fractured media is of inter-est in many of the environmental and energy problems. Examplesinclude tracing a NAPL migration, radioactive waste managementin the subsurface, and the enhanced oil recovery in naturally frac-tured reservoirs. Geological characterization and multiphase flowsimulation are challenging tasks in complex fractured media. Var-ious conceptual models have been used to describe multiphaseflow in fractured media. The main models include the dual-poros-ity and its variations, single-porosity, and the discrete-fracturedapproaches.

The dual-porosity model has traditionally been used to simulatethe flow in fractured hydrocarbon reservoirs [1–6]. In this model,the matrix–fracture mass transfer is described by empirical func-

ll rights reserved.

ering Research Institute, Palo

tions that incorporate ad hoc shape factors which are based on the-oretical derivations in single phase. In two-phase with capillaryand gravity, there is no theory for the calculation of shape factorsdespite the fact that the accuracy of flow calculations depend ontheir values. The problem may be even more fundamental. Thedriving force in the expression of the transfer function in two-phase flow may not have a sound basis. We should point out thatadvanced variations of the dual-porosity model such as the multi-ple interaction continua method [7] have better accuracy and fea-tures than the conventional model.

A more accurate and physics-based approach is the single-porosity model that describes fractures explicitly in the medium.This model requires the fracture-network characterization andthe fracture–fluid interaction functions. Modeling fractures withthe single-porosity model can be classified as a complex case ofheterogeneous porous media, where the fracture gridcells are trea-ted similarly to the matrix gridcells. The implementation of thismodel is, however, almost impossible in field-scale problemsbecause of excessive fine gridding and severe restrictions on thetime steps due to the contrast in geometrical scales between the

Page 2: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

892 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

fracture and the adjacent rock-matrix gridcells [8]. The matrixinversion due to ill-conditioning becomes a challenge.

The discrete-fracture model [9,10] describes the fracturesexplicitly in the medium similarly to the single-porosity model.However, unlike the single-porosity model, the fractures gridcellsare geometrically simplified by using ðn� 1Þ-dimensional gridcellsin an n-dimensional domain. In other words, in 3D space, the frac-tures are represented by the matrix gridcell interfaces, which are2D. This simplification removes the length-scale contrast causedby the explicit representation of the fracture aperture as in the sin-gle-porosity model. As a result, computational efficiency is im-proved considerably. However, the discrete-fracture model canonly handle a limited number of fractures (of the order of thou-sands) for reasons related to computational resources. One ap-proach to overcome this limitation is to model explicitly only themajor fractures and use an upscaling technique to mimic the effectof the minor fractures [11].

Various numerical methods based on the discrete-fracturemodel have been used to simulate single and multiphase flow infractured media. Methods that are developed for single-phase flowmay not be trivially extended to two-phase flow. In two-phaseflow, high permeable fractures may not be the main flow conductsdue to the contrast in capillary pressure between the fracture andthe rock matrix. Capillary pressure heterogeneity may also causefluid trapping (discontinuous saturation) and, in some cases, pres-sure discontinuously [12–15]. The fluid streamlines may, therefore,change significantly with time. The discrete-fracture model hasbeen used in the finite difference (FD), finite volume (FV), andthe Galerkin finite element (FE). These methods may have someinherent limitations, as discussed below.

Slough et al. [16,17] used the FD method to simulate the migra-tion of a NAPL in 2D and 3D fractured media in structured grids.They used distinct degrees of freedom for the saturation and pres-sure in the fracture and matrix control-volumes. In Cartesian grid-ding, the intersection of horizontal and vertical fracturescorresponds to a control-volume that has the dimensions of thefracture thicknesses (see Fig. 1a). Such a control-volume may re-duce significantly the size of the time step in the numerical model.To eliminate the degrees of freedom at the fracture junctions,Slough et al. used a technique based on steady-state flow equa-

a b

c d

Fig. 1. Different spatial discretizations of the fractures and the adjacent matrixgridcells in FD, FE, CVFE, and MFE. (a) FD; (b) FE; (c) CVFE; (d) MFE.

tions. They suggested to use two-point upstream weighting toapproximate the mobilities. The extension of their technique to afracture gridcell that has several feeding fractures was not dis-cussed. When the control-volume at the fractures junction inFig. 1a is eliminated, one fracture gridcell may have up to three up-stream fractures. Lee et al. [11] used a hierarchical approach basedon fracture lengths such that ‘‘short” fractures are treated implic-itly by increasing the effective matrix permeability while ‘‘long”fractures are described explicitly. Similarly to the model by Sloughet al., the fractures were aligned with the matrix gridcells; that is,either vertical or horizontal. They used a technique for well pro-ductivity indices for single-phase flow (capillary pressure was ne-glected) to calculate the transmissibilities between the fractureand the matrix gridcells.

The Galerkin FE method has been used to model single-phase[9,10,18–20] and two-phase flow [21–23] in fractured media. In2D space, the fractures are represented by 1D entities and the de-grees of freedom are located at the mesh nodes (see Fig. 1b). Thefractures and the adjacent matrix gridcells share the same degreesof freedom, therefore, there is an implicit assumption of the conti-nuity of pressure and saturation at the matrix–fracture interface.As a result, the FE suffers from a weakness in two-phase flow be-cause it cannot account properly for saturation discontinuity fromcapillary pressure contrast at the matrix–fracture interface.

Two FV methods with different spatial discretizations havebeen used for fractured media: (1) the cell-centered FV, and (2)the vertex-centered FV, which is also known as the control-volumefinite element (CVFE) method. In the cell-entered FV, similarly tothe FD (Fig. 1a), the fracture gridcells are located at the boundariesof the matrix gridcells. This approach can describe heterogeneitieswithin the mesh, however, there is numerical complexity in defin-ing the matrix–fracture and the fracture–fracture fluxes. Karimi-Fard et al. [24] introduced a simplified cell-centered FV methodfor two-phase flow in fractured media. The authors, however, men-tioned that their method could create numerical errors in non-orthogonal grids because the fluxes are calculated by the two-pointflux approximation method. They used an analogy between theconductance through resistors and flow in porous media to elimi-nate the control-volumes at fracture intersections, where capillarypressure and gravity effects are neglected. Grant et al. [25] alsoused the cell-centered FV method to model two-phase flow in 2Dfractured media. To calculate the fracture–fracture flux, they intro-duced special nodes at the fracture intersections where they calcu-lated the saturation at an intermediate time step.

Many authors have used the CVFE method and some have com-bined it with the Galerkin FE method to solve the two-phase flowequations in fractured media [26–31]. In the CVFE method, thefracture entities are embedded within the matrix control-volume(see Fig. 1c). In this approach, the calculation of the matrix–frac-ture flux is avoided and there is no difficulty in computing the frac-ture–fracture flux. However, consistent modeling of the saturationdiscontinuity from the capillary pressure contrast at the fracture–matrix interface has not been fully examined. Monteagudo and Fir-oozabadi [32] used the capillary pressure continuity equation atthe matrix–fracture interface and the capillary pressure–saturationfunctions to calculate the saturations in the fracture and the matrixblocks within the same control-volume. In other words, the satura-tion within a control-volume is distributed between the matrixand fracture blocks so that the capillary-pressure continuity is pre-served at the interface. This approach may become inefficientwhen the fracture capillary pressure is zero. Reichenberger et al.[33] used discontinuous approximation functions for the satura-tion in the fractures to account for the saturation discontinuity atthe matrix–fracture interface. Similarly to the method of Monte-agudo and Firoozabadi [32], this approach suffers from zero capil-lary pressure in the fracture.

Page 3: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Cell-average potentialFace-average potential

Simplification

Fig. 2. Modeling of flow in fractured media using the cross-flow equilibrium in thefractures and the adjacent matrix gridcells.

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 893

There is another concern related to the matrix gridding thatmay affect the accuracy of the CVFE method in which one invokesa cross-flow equilibrium between the fracture and the adjacentmatrix blocks. Here, the pressure in the fracture and that in thematrix are assumed equal within the same control-volume. Inthe CVFE approaches that ignore the capillary pressure heterogene-ity, the same saturation is assumed within the control-volume.These assumptions are justifiable if the matrix gridcells next tothe fractures are small enough. For an acceptable accuracy for anenhanced oil recovery problem (gas injection), Hoteit and Firooza-badi [34] found that the matrix-gridcells size next to the fracturesshould be of the order of ten centimeters in a fractured reservoir oflength-scale in kilometers. Reichenberger et al. [33] also used veryfine griddings in their numerical experiments.

The mixed finite element (MFE) method [35,36] has receivedmuch attention for its powerful features in approximating thevelocity field in highly heterogeneous and anisotropic media andhas been widely used to model single-phase flow in fractured med-ia. In Refs. [37,38], the method is used to simulate steady-state sin-gle-phase flow in fracture-networks and the mass transferbetween the fractures and the rock matrix was neglected. Alboinet al. [39] employed the MFE to solve the transport equation infractured media. Martin et al. [40] extended the work of Alboinet al. to model the cases when the fracture is less permeable thanthe matrix.

In this work, we extend the MFE method to two-phase flow infractured media with gravity and capillarity and account for ma-trix–fracture interactions. In our algorithm, the use of the hybrid-ized MFE method [41,36] overcomes the limitations of thenumerical models discussed earlier. In our work, the degrees offreedom for the saturation and pressure in a fracture gridcell arenot the same as those in the neighboring matrix gridcells (seeFig. 1d). As a result, one can describe readily the discontinuity insaturation from capillary pressure heterogeneity. The fractionalflow (global pressure) formulation [42] in the MFE does not allowproper modeling of saturation discontinuity. We use instead a con-sistent formulation from Ref. [43] in which phase pressure servesas the primary unknown.

Most of the conventional approaches in the literature use two-point upstream weighting to approximate the mobilities. For someapplications, this technique is known to provide poor accuracy[44,45]. In this work, we use the discontinuous Galerkin (DG)method [42,46,47] stabilized by a multidimensional slope limiterintroduced by Chavent and Jaffré [42] to reduce the numerical dis-persion in the saturation. The two-phase immiscible flow equa-tions are solved by an implicit-pressure–explicit-saturation(IMPES) approach. The IMPES method may perform more effi-ciently than the fully-implicit method [48]. Its stability is, however,restricted to a CFL condition, which is inversely proportional to thegridcell pore-volumes. Therefore, the fracture gridcells may imposea severe CFL condition. To allow for high computational efficiency,we implement an implicit scheme in the fracture network and anexplicit scheme in the matrix.

With above literature review on two-phase flow in fracturedmedia, we are now set to first present the essence of the proposedmodel and to discuss the governing equations for two-phaseimmiscible flow in porous media. We then rewrite the flow equa-tions in an equivalent form which is appropriate for the MFE for-mulation. The MFE formulations in the matrix and fracturedomains are then applied to approximate the volumetric flux andthe wetting-phase potential, where the capillary pressure is takeninto account. This is followed by a description of the DG formula-tion and the proposed time scheme. Several examples in 2D and3D are presented to show the applicability of our method.

In Ref. [43], we presented the formulation and the numericalmethod for two-phase flow in unfractured media. We focused on

modeling capillary pressure continuity/discontinuity and verifiedthe accuracy of our method with analytical solutions in heteroge-neous media. Some details presented in that work are not repeatedin this paper.

2. Proposed method

In the conventional methods, the flux through gridcell inter-faces is calculated in a post-processing step from the pressure field.In the MFE method, the flux and the pressure unknowns areapproximated simultaneously. The original MFE method with thelowest order Raviart–Thomas space [35,49] leads to a saddle pointproblem, where the linear system to solve is indefinite. The hybrid-ized MFE method [36,41] overcomes this problem by adding newdegrees of freedom (Lagrange multipliers) that represent the pres-sure averages (traces of the pressure) on the gridcell interfaces. Theoriginal MFE and the hybridized MFE are algebraically equivalent[41,50]. The difference is only in the selection of the primary vari-ables. In this work, we use the hybridized MFE, and for the sake ofbrevity we denote it also by MFE. Its primary variables are thetraces of pressure; the secondary variables are the gridcell averagepressures and the averages fluxes across the gridcell interfaces (seeFig. 2).

In our work, we assume capillary pressure continuity (that is,pressure continuity) at the matrix–fracture interface, except whenthe pressure of the non-wetting phase entering a different mediumsaturated with the wetting-phase is less than the threshold (entry)pressure [13–15]. The non-wetting phase becomes immobile and,therefore, there may be discontinuity in the non-wetting phasepressure. Modeling this phenomenon in heterogeneous media bythe MFE method is discussed in Ref. [43]. For the sake of clarity,the discontinuity in capillary pressure is not discussed further inthis work; we assume that the wetting phase is always displacingthe non-wetting phase (imbibition) and so the pressures are al-ways continuous.

In a previous work [51,52], we used the MFE method to simu-late single and two-phase compositional flow in 2D fractured med-ia but with neglected capillary pressure. We also invoked cross-flow equilibrium between the fracture and the adjacent matrixgridcells in a similar fashion as the CVFE method. In other words,the fracture pressure and the adjacent matrix gridcells pressuresare assumed equal (see Fig. 2). As discussed in Section 1, this sim-plification is justifiable if the matrix gridcells next to the fractureare small.

A central idea in our work is a new formulation that does notinvoke the cross-flow equilibrium across the fractures and theneighboring matrix blocks. We assume a constant potential alongthe width of the fracture gridcells (see Fig. 3) and thereforecross-flow equilibrium is assumed across the fractures only. Assketched in Fig. 3, the degrees of freedom for the potential and sat-

Page 4: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Cell-average potentialFace-average potential

Simplification

Fig. 3. Modeling of flow in fractured media using our proposal without the cross-flow equilibrium.

894 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

uration in the matrix and fracture gridcells are distinct. This newapproach alleviates the size constraint on the matrix gridcells nextto the fractures and allows accurate calculation of the saturationsin the fracture and in the adjacent matrix gridcells without a needfor small neighboring matrix blocks. Our proposal has similaritiesto the conventional single-porosity model, however there is a sig-nificant advantage as it allows accurate and efficient calculation ofthe matrix–fracture and fracture–fracture flux, as will be discussedlater.

3. Governing equations

The governing equations in porous media for the flow of twoincompressible and immiscible phases, a wetting phase (referredto by the subscript w), and a non-wetting phase (referred to bythe subscript n), are given by the saturation equation and the gen-eralized Darcy law of each phase. Saturation constraint and thecapillary pressure relation complete the formulation:

/oSa

otþr � ðvaÞ ¼ Fa; a ¼ n;w; ð1Þ

va ¼ �kra

la

Kðrpa þ qagrzÞ; a ¼ n;w; ð2Þ

Sn þ Sw ¼ 1; ð3ÞpcðSwÞ ¼ pn � pw: ð4Þ

In the above equations, K is the absolute permeability tensor, Sa,pa, qa, kra, Fa, and la are the saturation, pressure, density, relativepermeability, sink/source term, and viscosity of phase a, respec-tively, g is the gravity acceleration, z is the depth, and pc is the cap-illary pressure.

As discussed in Ref. [43], the MFE method was initially pro-posed for single-phase flow in porous media. A number of authorshave used the fractional flow formulation (also known as the globalpressure formulation) to extend the MFE for two-phase flow. TheMFE method with the fractional flow formulation, however, be-comes inconsistent for the simulation of flow in heterogeneousmedia with different capillary functions. A consistent formulation,that overcomes the fractional flow deficiency, is provided in Ref.[43]. A brief review of the formulation follows.

We define the flow potential, Ua, of phase a as follows:

Ua ¼ pa þ qagz; a ¼ n;w: ð5Þ

The capillary potential, Uc, becomes:

Uc ¼ Un � Uw ¼ pc þ ðqn � qwÞgz: ð6Þ

Then, Eqs. (1) and (2) are rewritten in the following equivalentforms:

r � ðva þ vcÞ ¼ Ft; ð7Þ

/oSw

otþr � ðfwvaÞ ¼ Fw; ð8Þ

where the velocity variables va and vc, whose sum is the total veloc-ity, are given by:

va ¼ �ktKrUw; ð9Þvc ¼ �knKrUc: ð10Þ

In the above equations, ka ¼ kra=la is the mobility of phase a,kt ¼ kn þ kw is the total mobility, fw ¼ kw=kt is the wetting-phasefractional function, and Ft ¼ Fn þ Fw is the total sink/source term.

The system of Eqs. (7) and (8) are subject to Neumann (flow ratecontrol) and Dirichlet (pressure control) boundary conditions. Thesystem of Eqs. (7)–(10) and the initial and boundary conditions areour working equations with the MFE and DG methods.

4. Approximation of the volumetric flux

The pressure and the saturation equations are decoupled andsolved sequentially in an IMPES-like approach. However, we sug-gest the use of a different temporal scheme for the saturation equa-tion in the fractures, as will be discussed later. The MFE method isused to approximate implicitly the pressure and the velocity in thematrix and the fractures. Since the matrix and fracture gridcellshave different geometrical dimensions (matrix gridcells are n-dimensional and fracture gridcells are ðn� 1Þ-dimensional), thegoverning equations are approximated separately. In the matrixdomain (referred to by superscript m), the equations describingthe volumetric balance and velocity are given by:

vma ¼ �km

t KmrUmw; ð11Þ

r � ðvma þ vm

c Þ ¼ Fmt : ð12Þ

In the fractures, the governing equations are integrated along thefracture width �. The integrals are readily computed since the vari-ations in potential and saturation are averaged across the fracturewidth. The fracture width is taken into consideration in the frac-ture–fracture flux and the accumulation term in the fracture grid-cells. The equations in the fractures (referred to by the superscriptf) in ðn� 1Þ-dimensional are expressed by:

vfa ¼ �kf

tKfrUf

w; ð13Þr � ðvf

a þ vfcÞ ¼ Q f þ Ff

t: ð14Þ

The transfer function Q f in Eq. (14) accounts for the volumetrictransfer across the matrix–fracture boundaries. This transfer func-tion will be evaluated in the following section.

4.1. Discretization in the matrix

The governing equations of the volumetric flow in the matrixexpressed in Eqs. (11) and (12) are discretized separately by thehybridized MFE method with the lowest-order Raviart–Thomasspace. The MFE discretization of the velocity expression in Eq.(11) in matrix gridcell K yields the explicit expression for the fluxqm

a;K;E across a gridcell-face E in terms of the of the cell-average po-tential Um

w;K and all the face-average potentials Kmw;K;E in all the faces

of K , that is:

qma;K;E ¼ am

K;EUmw;K �

XE02oK

bmK;E;E0K

mw;K;E0 ; E 2 oK; ð15Þ

where the terms amK;E and bm

K;E are constants, independent of the po-tential and flux variables. Details are provided in Ref. [43].

The flux in Eq. (15) is written locally for all faces within eachmatrix gridcell. To link the gridcells in the mesh together, the fol-lowing continuity conditions for the flux and the potential are im-posed at each interface E of two neighboring gridcells K and K 0

(that is, E ¼ K \ K 0):

Page 5: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

fracture

matrix

matrix

matrix

fracture

matrix

matrix

matrix

fracture

a

b

c

Fig. 4. Modeling of the flow in fractures in our proposed approach.

Fig. 5. Mesh with five intersecting fractures at I; the fractures k1, k2, k3 are in theupstream, and k4, k5 are in the downstream.

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 895

� If E is neither a fracture nor a barrier, the continuity of flux andpotential is imposed.

qma;K;E þ qm

a;K 0 ;E ¼ 0;

Kmw;K;E ¼ Km

w;K 0 ;E ¼ Kmw;E:

(ð16Þ

� If E is a fracture, the total flux across both sides of the matrix–fracture interface defines the transfer function Q f

a;E at E, whichacts as a sink/source term. There is, however, continuity of thepotential across the matrix–fracture interface E (see Fig. 3).

qma;K;E þ qm

a;K 0 ;E ¼ Q f

a;E;

Kmw;K;E ¼ Km

w;K0;E ¼ Ufw;E:

(ð17Þ

� If E is a barrier, the flux across E is zero and face potentials Kmw;K;E

and Kmw;K 0 ;E may not be equal.

qma;K;E ¼ qm

a;K 0 ;E ¼ 0;

Kmw;K;E 6¼ Km

w;K 0 ;E:

(ð18Þ

We like to make the comment that Eq. (18) takes care of a bar-rier in simple way. In addition to the above conditions, the flux andpotential at the domain boundaries are described by the Neumannor Dirichlet boundary conditions (see Ref. [43]). For the sake ofsimplicity, we assume that the domain boundaries are imperme-able. From Eqs. (15)–(18), one can eliminate the flux variablesand construct a system consisting of the unknowns; the fracturecell-potentials, Uf

w, matrix cell-potentials, Umw, and the matrix

face-potentials, Kmw. The system of equations corresponding to the

matrix gridcells connected to fractures and those not connectedto fractures are written separately as follows:

RT;m;m

RT;m;f

!Um

w �Mm;m Mm;f

Mf;m Mf;f

!Km

w

Ufw

� �¼

0Q f

� �: ð19Þ

The system given in the above equations is split into two parts toexplicitly show the connectivity at the matrix–fracture interfaces.Note that Mm;f ¼ Mf;m in Eq. (19). The matrix–fracture volumetrictransfer function Q f will be eliminated later when we couple theflow between the matrix and the fractures. The definitions of thematrices in Eq. (19) are provided in Appendix A.

The next step is the integration of Eq. (12) over K and to obtainflux unknowns in terms of the potentials given in Eq. (15) [43], thatis,

amK Um

w;K �XE2oK

amK;EK

mw;K;E ¼ Fm

K ; ð20Þ

where amK ¼

PE2oKam

K;E, and FmK ¼ �

PE2oK qm

c;K;E þR R R

KðFmt Þ is a func-

tion of capillary pressure that will be discussed in section ‘‘Calcula-tion of the capillary flux”.

In matrix form, Eq. (20) and the potential continuity equationsgiven in Eqs. (16)–(18), yield:

DmUmw � ðR

m;m Rm;fÞKm

w

Ufw

� �¼ Fm: ð21Þ

The diagonal matrix Dm in Eq. (21) is defined in Appendix A, andRm;m and Rm;f appear in Eq. (19).

4.2. Discretization in the fracture

In n� D space ðn ¼ 2;3Þ, the fracture network is ðn� 1Þ � D. TheMFE formulation in the fracture network is basically similar to theconventional formulation in unfractured ðn� 1Þ � D domain. Thereare, however, two differences: (1) the transfer function which actsas a sink/source in the fracture (see Fig. 4), (2) the fracture–fractureconnectivity that may have multiple fracture-gridcells intersectingat the same interface as in Fig. 5. Similarly to the discretization in

the matrix previously discussed, one discretizes the volumetricflow given in Eqs. (13) and (14) in the fractures. With the MFE dis-cretization of Eq. (13) in a fracture gridcell k (k 2 ðn� 1Þ � D space),the flux qf

a;k;e across an interface e (e 2 ðn� 2Þ � D space) of thefracture gridcell k becomes:

qfa;k;e ¼ af

k;eUfw;k �

Xe02ok

bfk;e;e0K

fw;k;e0 ; e 2 ok: ð22Þ

The constants afk;e and bf

k;e are similar to those defined in Eq.(15). Let ne be the number of intersecting fractures at the interfacee. The fracture–fracture interface e has a negligible volume, there-fore, no accumulation is assumed and the potential continuity isimposed at e, that is,

Page 6: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

896 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

Pne

i¼1qf

a;ki ;e¼ 0;

Kfw;ki ;e

¼ Kfw;e; i ¼ 1; . . . ;ne:

8><>: ð23Þ

Using Eqs. (22) and (23), the flux unknown is eliminated and asystem with the unknowns consisting of the fracture-gridcellspotentials, Uf

w, and the fracture-interfaces potentials, Kfw, is

obtained:eRT;fUfw � ~MfKf

w ¼ 0: ð24Þ

The volumetric balance given in Eq. (14) is integrated over eachfracture gridcell. In matrix form, a system similar to Eq. (21) isobtained:eDfUf

w � eRfKfw ¼ Ff þ Q f: ð25Þ

In the above equation, Q f is the matrix–fracture transfer previouslydefined. The matrices eD, eRf, and Ff in Eq. (25), and eRT;f and eM f in Eq.(24) are defined in Appendix A.

4.3. Combining the discretization in the matrix and fractures

The MFE formulations of the volumetric flow in the matrix andfractures are coupled through the matrix–fracture transfer func-tion and the continuity of the potential given in Eq. (17). By sub-stracting Eq. (25) from the second equation in Eq. (19), thetransfer function Q f cancels out. The resulting equation from thefirst equation in Eq. (19) and Eqs. (21) and (24) becomes:

Dm �Rm;m �Rm;f 0�RT;m;m Mm;m Mm;f 0�RT;m;f Mm;f ðMf;f þ eDfÞ �eRf

0 0 �eRT;f eM f

0BBBB@1CCCCA

Umw

Kmw

Ufw

Kfw

0BBB@1CCCA ¼

Fmw

0�Ff

w

0

0BBB@1CCCA: ð26Þ

The matrix Dm is diagonal and invertible, therefore, a simpleGaussian eliminate can be performed to eliminate the unknownUm

w. The final linear system reads as:

Am;m Am;f 0Am;f Af;f �eRf

0 �eRT;f eM f

0B@1CA Km

w

Ufw

Kfw

0B@1CA ¼ Vm

V f

0

0B@1CA: ð27Þ

The sought unknowns are, therefore, the traces of the wetting-phase potential in the matrix, Km

w, the fracture gridcell potentials,Uf

w, and the traces of the potential in the fracture network, Kfw.

The system in Eq. (27) has a size equal to the number of meshinterfaces plus the number of fracture–fracture interfaces; it issparse, symmetric and positive definite. There is no particular dif-ficulty in constructing the system in Eq. (27). A simple example fora 2D fractured domain is given in Appendix B to show the structureof the linear system. After solving Eq. (27), the matrix potentialsand the flux in the matrix and fracture network are calculated lo-cally from the first equation in Eq. (26), and Eqs. (15) and (22).

4.4. Calculation of the capillary flux

The degenerate flux from the capillary pressure and gravity, qc,in the right-hand side of Eqs. (15) and (22) is written in terms ofthe cell capillary potential Uc and the face capillary potential Kc

in the matrix and fractures. In a matrix gridcell, K , the flux acrossan interface E is given by Hoteit and Firoozabadi [43]:

qmc;K;E ¼ am

K;EUmc;K �

XE02oK

bmK;E;E0K

mc;K;E0 ; E 2 oK: ð28Þ

Similarly to Eq. (28), in a fracture gridcell, k, the flux across an inter-face e is given by:

qfc;k;e ¼ af

k;eUfc;k �

Xe02ok

bfk;e;e0K

fc;k;e0 ; e 2 ok: ð29Þ

By imposing the continuity of capillary pressure and flux at the ma-trix–matrix and fracture–fracture interfaces, the flux unknownsqm

c;K;E and qfc;k;e are eliminated, and the following system is obtained:

bMm 00 bM f

!Km

c

Kfc

� �¼

bV mbV f

!�

bRT;mUmcbRT;fUfc

!: ð30Þ

The matrices in the above equation are defined in Appendix A.The capillary potentials Um

c and Ufc in the right-hand side of Eq.

(30) are calculated by using the capillary pressure functions andthe wetting-phase saturations from the previous time step. Thecapillary pressure at the fracture gridcells are also calculated fromthe corresponding saturations. The system in Eq. (30) is diagonalper block and the matrices bMm, bM f are symmetric and positive def-inite, therefore, solvers such as preconditioned-conjugate-gradient(PCG) can be used efficiently.

5. Approximation of the wetting-phase saturation

As we did for the volumetric flow equation, the saturation equa-tion describing the mass conservation is written separately in thematrix and fractures. In the matrix, the wetting-phase saturationis defined by Sm

w and the governing equation is given by:

/m oSm

w

otþr � ðf m

w vma Þ ¼ Fm

w: ð31Þ

The discontinuous Galerkin (DG) method is used to approxi-mate Eq. (31) with an explicit, second-order Runge–Kutta timescheme. A multidimensional slope limiter [42] is applied in apost-processing step to remove potential oscillations in the satura-tion. A detailed description is provided in Ref. [43].

The saturation equation in the fracture network is written as:

/f oSf

w

otþr � ðf f

wvfaÞ ¼ bQ f

w þ Ffw: ð32Þ

The sink/source term, bQ fw, in the above equation is equal to the

volumetric transfer function, Q f in Eq. (14), multiplied by the wet-ting-phase fractional flow function fw. From the assumption ofincompressibility of the fluids, bQ f

w represents the mass transferfunction through the matrix–fracture interfaces. Because the flowin the fracture network can be much faster than that in the matrix,a first-order implicit time scheme is used to approximate the timeoperator in Eq. (32). This approach removes the CFL restriction onthe time step in the fracture network and allows different timesteps in the matrix and the fractures. Unlike the approximationsin the matrix, a first-order spatial scheme is used for the saturationequation in the fractures. As a result, one may expect more numer-ical dispersion in fracture saturation, however, this may not have asignificant influence as a second-order method is used in the ma-trix. We do not use an implicit DG method in the fracture for itsexpensive computational overhead. The first-order temporal andspatial discretizations of Eq. (32) in a fracture gridcell k at a timestep n reads as:

/fk

Sf;nw � Sf;n�1

w

DtþXe2ok

f f;nw;eqf;n

a;k;e ¼ bQ fþ ;n�1w þ bQ f� ;n

w|fflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflffl}þFf;nw;k: ð33Þ

The second term on the left-hand side of Eq. (33) represents thenet fracture–fracture flux. The volumetric flux qf;n

a;k;e is calculated ina preprocessing step from the MFE approximations, previously dis-cussed. The matrix–fracture mass transfer function on the right-hand side of Eq. (33) is split into two parts that are approximateddifferently according to the flow direction between the fractureand the two adjacent matrix gridcells. The notation bQ fþ ;n�1

w refersto a source term and bQ f� ;n

w refers to a sink term for the fracture grid-cell. We distinguish three possible cases as follows:

Page 7: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

X (m)

Y(m

)

0 20 40 60 80 1000

20

40

60

80

1000.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

X (m)

Y(m

)

0 20 40 60 80 1000

20

40

60

80

1000.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

Fig. 6. Water saturation contours at PVI ¼ 0:5: with and without the cross-flow equilibrium, j ¼ 0:4; Example 1. (a) With cross-flow equilibrium; (b) without cross-flowequilibrium.

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 897

� If the two adjacent matrix gridcells are in the upstream of thefracture, as appears in Fig. 4a, then bQ f� ;n

w ¼ 0 and the total sourceterm for the fracture, bQ fþ ;n�1

w , is approximated explicitly in time.� If the two adjacent matrix gridcells are in the downstream of the

fracture, as appears in Fig. 4b, then bQ fþ ;n�1w ¼ 0 and the total sink

term for the fracture, bQ f� ;nw , is approximated implicitly in time.

� If one matrix gridcell is in the upstream and the other in thedownstream (see Fig. 4c), then the source term is approximatedexplicitly and the sink term implicitly.

The approach presented above is locally conservative in the ma-trix and fractures. The system of equations given in Eq. (33) is lin-earized by the Newton–Raphson (NR) method. If the NR methodfails to converge for a certain time step, one can cut the time stepand does multiple iterations in the fractures without reducing theglobal time step in the matrix. The time step is our method is con-strained by a CFL condition in the matrix where the explicit timemethod is used and by the nonlinearity (Newton iterations) owingto the flow interaction at the matrix–fracture interfaces. A simpledynamic time-step algorithm is adopted, where the time step iscontrolled by the change in saturation in the gridblocks withinthe previous time step and the number of Newton iterations re-quired for convergence.

The mobilities at the fracture–fracture interface are approxi-mated by the conventional single-point upstream-weighting tech-nique. An important point to be addressed next is how to definethe upstream-weighting in case of multiple intersecting fractures,where one fracture gridcell may have multiple upstream fractures.

5.1. Upstream-weighting at multiple intersection fractures

In the past, we developed a technique for multiple intersectingfractures in 2D space [52,53]; the extension to fractures in 3D do-main is straightforward. Let I be the interface (line) connecting NI

Table 1Relevant data for Examples 1 and 2

Domain dimensions: 100 m� 100 mMatrix properties: /m ¼ 0:2, Km ¼ 1 mdFracture properties: /f ¼ 1:0, K f ¼ 106 md, � ¼ 1 mmFluid properties: lw ¼ ln ¼ 1 cp, qw ¼ qn ¼ 1000 kg=m3

Relative permeabilities: linear (Eq. (39))Capillary pressure: neglectedResidual saturations: Srw ¼ 0:0, Srn ¼ 0:0Injection rate: 0.05 PV/dayMesh number: �1000 triangles

fracture gridcells, ki; i ¼ 1; . . . ;NI , as sketched in Fig. 5. The purposeis to evaluate the mobility at I which is given by fractional flow func-tion fw;I (see Eq. (33), for simplicity we drop the notation f). In eachfracture cell ki, the fractional flow function and the volumetric flux

Fig. 7. Matrix elements with different aspect ratios; Example 1. (a) j ¼ 2hb ¼ 1:0; (b)

j ¼ 2hb ¼ 0:1.

Page 8: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

898 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

at I are denoted by fw;kiand qa;w;ki

. According to the signs of the fluxes,influx or efflux at I, we define an integer ‘ ð0 < ‘ < NIÞ, such that:

Fig. 8. Oil recovery and water production versus PVI for different mesh aspect ratios w

Fig. 9. Oil recovery and water production versus PVI for different mesh aspect ratios wproduction.

X (m)

Y (

m)

0 20 40 60 80 1000

20

40

60

80

100

fracture

barrier

Fig. 10. Gridding of the domain with one fracture and one barrier; Case I: the fracture doe(b) Case II.

qa;w;ki> 0 for 0 < i � ‘ ðinfluxÞ;

qa;w;ki6 0 for ‘ < i < NI ðeffluxÞ:

ð34Þ

ith the cross-flow equilibrium; Example 1. (a) Oil recovery; (b) water production.

ith and without the cross-flow equilibrium; Example 1. (a) oil recovery; (b) water

X (m)0 20 40 60 80 100

Y (

m)

0

20

40

60

80

100

fracture

barrier

barrier

s not cross the barrier; Case II: the fracture crosses the barrier; Example 2. (a) Case I;

Page 9: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

X(m

)

0

20

40

60

80

100Y (m)0 20 40 60 80 100

Pre

ssu

re (

bar

)

Pre

ssu

re (

bar

)

200

400

600

X

Y

Z

X(m

)

0

20

40

60

80

100Y (m)0 20 40 60 80 100

200

400

600

X

Y

Z

Fig. 11. Pressure contours for Cases I and II; PVI ¼ 0:5; Example 2. (a) Case I; (b) Case II.

Length (m)

Hei

gh

t (m

)

0 5 10 15 20 250

Length (m)0 5 10 15 20 25

0

5

10

15

20

25

Hei

gh

t (m

)

5

10

15

20

25

Fig. 13. Triangulations of the fracture domain; Example 3. (a) Coarse: Nmk ¼ 1718, nf

j ¼ 104; (b) fine: Nmk ¼ 4667, Nf

j ¼ 191.

X (m)

Y(m

)

0 20 40 60 80 1000

20

40

60

80

100

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

X (m)

Y(m

)

0 20 40 60 80 1000

20

40

60

80

100

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

Fig. 12. Water saturation contours for Cases I and II; PVI = 0.5; Example 2. (a) Case I; (b) Case II.

Table 2Relevant data for Example 3

Domain dimensions: 25 m� 25 mMatrix properties: /m ¼ 0:2, Km ¼ 1 mdFracture properties: /f ¼ 1:0, K f ¼ 8� 105 md, � ¼ 0:1 mmFluid properties: lw ¼ 1 cp, qw ¼ 1000 kg=m3, ln ¼ 0:45 cp, qn ¼ 660 kg=m3

Injection rate: 0.0027 PV/dayMesh number: 1718 (coarse) and 4667 (fine) triangles

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 899

Note that, an integer ‘ that satisfies Eq. (34) does always exist.By writing the total volumetric balance and the total mass balanceat I, one gets,

X‘i¼1

qa;w;ki¼ �

XNI

i¼‘þ1

qa;w;ki; ð35Þ

X‘i¼1

fw;kiqa;w;ki

¼ �XNI

i¼‘þ1

fw;Iqa;w;ki¼ �fw;I

XNI

i¼‘þ1

qa;w;ki: ð36Þ

Page 10: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Table 3Relative permeabilities and capillary pressure in the matrix: Example 3

Sw kw kn pc ðbarÞ

0 0 0.6 10.1 4.5E�6 0.412 0.6420.2 1.4E�4 0.264 0.4340.3 0.001 0.163 0.31240.4 0.004 0.089 0.22610.5 0.014 0.041 0.15920.6 0.035 0.015 0.10450.7 0.075 0.003 0.05820.8 0.148 1.2E�4 0.0180.85 0.200 0 0

Table 4Relative permeabilities and capillary pressure in the fractures: Example 3

Sw kw kn pc ðbarÞ

0 0 1 10.1 0.118 0.882 0.0800.2 0.235 0.765 0.0400.3 0.353 0.647 0.0250.5 0.588 0.412 0.0100.6 0.706 0.294 0.0050.7 0.824 0.176 1.E�30.85 1 0 0

900 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

In Eq. (36), fw;I , which refers to the mobility at I, is equal to the up-stream mobility for the effluxes. From Eqs. (35) and (36), fw;I can bereadily calculated:

fw;kI ¼P‘

i¼1fw;kiqa;w;kiP‘

i¼1qa;w;ki

: ð37Þ

The expression for the mobility in Eq. (37) is a generalization of theconventional two-point upstream-weighting technique.

6. Numerical examples

The combined MFE-DG method with capillary effect was vali-dated using analytical solutions for two-phase flow in unfracturedmedia. We demonstrated that the algorithm enjoys accuracy androbustness especially in highly heterogeneous media with unstruc-tured griddings [43]. In fractured media, the MFE-DG associatedwith the cross-flow equilibrium has been validated with the sin-gle-porosity model [53]. In this work, we provide four examplesof various degrees of complexity. In the first example, we presentresults demonstrating that our method is not sensitive to the sizeof the matrix gridcells next to the fractures. The previous worksin the literature invoking the concept of cross-flow equilibrium be-tween the fractures and the adjacent rock matrix depend on thematrix gridding. The second example shows the flexibility of ourproposed model in numerical simulation of two-phase flow in por-ous media composed of a fracture and a barrier. The third exampleinvestigates the effect of the fracture capillary pressure on the fluidflow. The last example demonstrates the efficiency of our methodin a large 3D problem with layering, barriers, and fractures. Wewould like to point that the gridding for all the problems is byno means optimal. The selection shows powerful features of thealgorithm. All runs were performed on an 1.8 GHz Intel-CentrinoPC with 1 GB of RAM.

6.1. Example 1: mesh sensitivity

In this example, we examine the dependency of our model onthe size of the matrix gridcells in the vicinity of the fractures. Weconsider a 2D horizontal porous medium of dimensions

ð100 m� 100 mÞ with one fracture along the diagonal, as shownin Fig. 6. The domain is initially saturated with oil. Water is in-jected at the lower left corner to produce oil at the opposite corner.The fluid and medium properties are provided in Table 1. The do-main is discretized into a mesh with about 1000 triangular matrixgridcells and 20 fracture gridcells (segments). We define an aspectratio j for the matrix gridcells next to the fractures by j ¼ 2h=b,where h is the triangle height, and b is the base. Fig. 7a shows amesh where the aspect ratio is one. In Fig. 7b, the aspect ratio is re-duced to 0.1 by dragging the nodes of the triangles facing the frac-tures closer to its base (that is, the fracture). The aspect ratio of themesh is, therefore, modified by changing the locations of somenodes without changing their number.

Fig. 6 shows the water saturation profiles with and without thecross-flow equilibrium assumption. The former appears to be moredispersive close to the fracture. Oil recovery and water break-through obtained with the cross-flow equilibrium assumption fordifferent aspect ratios are plotted in Fig. 8. Large matrix gridcellsnext to the fractures, that correspond to large aspect ratios, delaywater breakthrough and result in overestimation of the oil recov-ery. The results show that convergence is obtained with a small as-pect ratio (about 0.1) indicating that the matrix gridcells should berelatively small. On the other hand, our new model, shows almostidentical results for aspect ratios of 1.0 and 0.1 (Fig. 9). Fig. 9 alsoshows that the results obtained with cross-flow equilibrium withfine matrix gridcells ðj ¼ 0:1Þ converges to the results by ourmethod with coarse matrix gridcells ðj ¼ 1Þ. We also note thatboth methods need comparable CPU time on the same mesh, how-ever, our method with j ¼ 1 is nearly one order of magnitude fas-ter than the method with j ¼ 0:1.

6.2. Example 2: flexibility in modeling fractures and barriers

In this example, we demonstrate the flexibility of the MFE-DGmethod in modeling fractures and barriers. We consider a 2D rect-angular domain (100 m� 100 m) initially saturated with oil. Theinjector and producer locations are as in Example 1; the relevantdata are provided in Table 1. One fracture and one barrier areembedded in the porous medium; two cases, I and II, are distin-guished. In Case I, the fracture touches the barrier but does notcross it (see Fig. 10a). In Case II, the fracture intersects the barrierand crosses it (Fig. 10b).

The fracture–barrier intersection point in Case I has a dual role.The precise modeling by the Galerkin FE method with consider-ation of the degrees of freedom at the mesh nodes, is not straight-forward. In Case II, the injected fluid flows through the barrier via asingle point located at the fracture–barrier intersection. Modelingthis case is a challenge for the CVFE method due to considerationof the degrees of freedom at the center of the control-volumes.We model in a simple way both cases without a special treatment.The implementation of simple expressions in Eqs. (17) and (18) inthe MFE-DG method provides the tool to model two-phase flowwith capillarity and gravity in complex media containing fractures(Eq. (17)) and barriers (Eq. (18)). Fig. 11 depicts the pressure con-tours with discontinuity at the barrier (Fig. 11a, Case I) and pres-sure continuity for the fracture crossing the barrier (Fig. 11b,Case II). Results for water saturations at PVI ¼ 0:5 for Cases I andII are plotted in Fig. 12a and b, respectively.

6.3. Example 3: capillary pressure effect

In the past, Karimi-Fard and Firoozabadi [23] have used the do-main in Fig. 13 to study the effect of capillary pressure on waterinjection in fractured media. In this example, we use our methodto provide a comprehensive study in terms of model efficiencyand capillary pressure effect. The domain with the fracture net-

Page 11: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Length (m)

Hei

ght(

m)

0 5 10 15 20 250

5

10

15

20

25

0.80

0.72

0.64

0.57

0.49

0.41

0.33

0.26

0.18

0.10

Length (m)

Hei

gh

t (m

)

0 5 10 15 20 250

5

10

15

20

25

0.80

0.72

0.64

0.57

0.49

0.41

0.33

0.26

0.18

0.10

Length (m)

Hei

gh

t (m

)H

eig

ht

(m)

Hei

gh

t (m

)H

eig

ht

(m)

0 5 10 15 20 250

5

10

15

20

25

0.80

0.72

0.64

0.57

0.49

0.41

0.33

0.26

0.18

0.10

Length (m)0 5 10 15 20 25

0

5

10

15

20

25

0.80

0.72

0.64

0.57

0.49

0.41

0.33

0.26

0.18

0.10

Length (m)0 5 10 15 20 25

0

5

10

15

20

25

0.80

0.72

0.64

0.57

0.49

0.41

0.33

0.26

0.18

0.10

Length (m)0 5 10 15 20 25

0

5

10

15

20

25

0.80

0.72

0.64

0.57

0.49

0.41

0.33

0.26

0.18

0.10

Fig. 14. Water saturation contours and water velocity field for cases I, II, and III at PVI ¼ 0:25 and PVI ¼ 0:50; Example 3. (a) PVI ¼ 0:25: pmc ¼ 0, pf

c ¼ 0 (Case I); (a0)PVI ¼ 0:50: pm

c ¼ 0, pfc ¼ 0 (Case I); (b) PVI ¼ 0:25: pm

c 6¼ 0, pfc 6¼ 0 (Case II); (b0) PVI ¼ 0:50: pm

c 6¼ 0, pfc 6¼ 0 (Case II); (c) PVI ¼ 0:25: pm

c 6¼ 0, pfc ¼ 0 (Case III); (c0) PVI ¼ 0:25:

pmc 6¼ 0, pf

c ¼ 0 (Case III).

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 901

work is sketched in Fig. 13a and b corresponding to a coarse and afine grid. The relative permeability and capillary pressure curves inthe matrix and fractures are given in Tables 3 and 4, respectively.Other relevant data are provided in Table 2. We investigate the ef-fect of capillary pressure on flow by considering three cases. InCase I, the capillary pressures in the matrix ðpm

c Þ and in the fractureðpf

cÞ are neglected (that is, pmc ¼ pf

c ¼ 0). In Case II, both capillarypressures are nonzero, as provided in Tables 3 and 4. In Case III,only the fracture capillary pressure is zero and that of the matrixis given in Table 3.

In Fig. 14, we show the water saturation distributions and watervelocity profiles for the three cases at PVI ¼ 0:25 and PVI ¼ 0:5 onthe fine grid. In Case I. (Fig. 14a and a0), water flows very quicklythrough the fractures and leads to early breakthrough. In Case II

(Fig. 14b and b0), however, the effect of capillary-pressure contrastat the matrix–fracture interface delays breakthrough. Note thatdue to capillary cross-flow, the displacement process becomesmuch more efficient in Case II than in Case I.

As the contrast of capillary pressure increases in Case III(Fig. 14c and c0), the fractures permeability has little influence onwater mobility due to very large capillary effect when water satu-ration in the matrix is much lower in the matrix than in the frac-ture. Although, the fracture absolute permeability is orders ofmagnitudes greater than that in the matrix, water in the fracturesand the surrounding matrix flows with comparable speed. The con-trast in capillary pressure at the fracture–matrix interface creates asignificant cross-flow via the water velocity vectors laterally cross-ing the fractures (see Fig. 14c and c0). The oil recovery and water

Page 12: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Fig. 15. Oil recovery and water production versus PVI for the three cases I, II, and III; Example 3. (a) Oil recovery; (b) water production.

Table 5Relevant data for Example 4

Domain dimensions: 1000 m� 300 m� 250 mMatrix properties: /m ¼ 0:2, Km ¼ f20 md;1 md;10 mdga

Fracture properties: /f ¼ 1:0, K f ¼ 105 md, � ¼ 1 mmFluid properties: as in Example 3Relative permeabilities (layers): Eq. (39) with mw ¼ f2;3;2g;

a mn ¼ f3;3;2ga Srn ¼ 0; Srw ¼ 0:25 (all layers)Relative permeabilities linear (fractures): (Eq. (39)); Srn ¼ Srw ¼ 0Capillary pressure (Case I): neglectedCapillary pressure (Case II): Bf ¼ 0:01, Bm ¼ f1;5;2ga barInjection rate: 10�4 PV/dayMesh number: �8500 prisms

a In layers 1 (top), 2 and 3, respectively.

Table 6CPU time in minutes for Examples 3 and 4

Example 3 (coarse) Example 3 (fine) Example 4

Zero capillarity 0.9 3.7 56Nonzero capillarity 1.5 7.6 98

902 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

breakthrough versus PVI for the three cases are plotted in Fig. 15. Itis obvious that capillary pressure contrast improves the recoveryand delays the breakthrough. The capability to explicitly describethe matrix–fracture cross-flow due to capillary pressure contrastis a significant feature of our method. The results from the fineand coarse grids are almost the same. The CPU times on both grids

X(m

)

0

200

400

600

800

1000

Y (m)0100200300

Z(m

)

0

125

250

XY

Z

Layer 1

Layer 3

Layer 2

Fig. 16. Discretization of the layered domain and the embedded fractures and

for Cases I and II are shown in , Table 6. The CPU time in Case III isabout 20% more than in Case II.

6.4. Example 4: complex 3D media with layers, faults, and fractures

In this example, we consider a 3D layered tilted domainð1000 m� 300 m� 250 mÞ, as appears in Fig. 16a. The mediumhas two barriers and several intersection fractures (Fig. 16b). Thelayer thicknesses from top to bottom are 30 m, 40 m, and 30 m,respectively. The injector is located at the bottom corner of themedium (at ð0;0;0Þ) and the producer at the opposite top corner(at (1000 m, 300 m, 250 m)). The domain is discretized into anunstructured mesh of 8500 matrix gridcells (prisms) (Fig. 16a).The number of fracture gridcells (quadrilaterals) is nearly 1700.In Ref. [43], we showed the applicability of the MFE-DG methodon different discretizations in 3D. The rock properties, and relativepermeability functions used in the layers and the fractures are pro-vided in Table 5. The relative permeabilities and the capillary pres-sure are given by:

krw ¼ Smwe ; krn ¼ ð1� SeÞmn ; ð38Þ

pc ¼ �B lnðSeÞ; ð39Þ

where mw and mn are defined in Table 5, and Se is the normalizedsaturation defined in terms of Srw and Srn, which are the residualsaturations for the wetting and non-wetting phases, respectively:

Se ¼Sw � Srw

1� Srw � Srn: ð40Þ

X(m

)

0

200

400

600

800

1000

Y (m)0100200300

Z( m

)

0

125

250

Fracture

Barrier

barriers; Example 4. (a) 3D layered domain; (b) 2D fractures and barriers.

Page 13: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

X(m

)

0

200

400

600

800

1000

Y (m)0100200300

Z( m

)

0

125

2500.70.60.50.40.30.20.1

X(m

)

0

200

400

600

800

1000

Y (m) 0100200300

Z( m

)

0

125

250

0.9

0.7

0.5

0.3

0.1

a b

Fig. 17. Water saturation profiles in the matrix and fractures at PVI ¼ 0:25, zero capillary pressure; Example 4. (a) Matrix; (b) fractures.

X(m

)

0

200

400

600

800

1000

Y (m)0100200300

Z( m

)

0

125

2500.70.60.50.40.30.20.1

X(m

)

0

200

400

600

800

1000

Y (m) 0100200300

Z( m

)

0

125

250

0.9

0.7

0.5

0.3

0.1

a b

Fig. 18. Water saturation profiles in the matrix and fractures at PVI ¼ 0:25, nonzero capillary pressure; Example 4. (a) Matrix; (b) fractures.

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 903

Two cases with zero and non-zero capillary pressures are con-sidered. Water saturation profiles in the matrix and fractures forboth cases are shown in Figs. 17 and 18. Capillary pressure reduceswater velocity in the fractures and, therefore, improves recovery,as shown in Fig. 19. We note that the use of the implicit timescheme with the desirable feature of smaller time steps in the frac-tures without reducing the time step in the matrix has improvedsignificantly the computational efficiency. In addition to the aboveexamples, we have tested the proposed method for a variety of

Fig. 19. Oil recovery versus PVI with zero and nonzero capillary pressure;Example 4.

other problems. All the examples show powerful features of thealgorithm. We have also compared the simulation results and mea-sured data from water injection in 3D oil-saturated media. Theagreement is excellent. The comparison between simulation anddata will be made in a future publication.

7. Conclusions

A numerical model based on the combined MFE and DG meth-ods is introduced to simulate incompressible two-phase flow inmultidimensional fractured media. The main features of our newmethod are:

� The saturation discontinuity at the matrix–fracture interfacefrom the capillary pressure contrast is properly described.

� A new approach is introduced without the use of the cross-flowequilibrium between the fractures and the adjacent matrixblocks. The continuity of pressure is only imposed at thematrix–fracture interface. This approach allows relatively largematrix gridcells next to the fractures.

� The hybrid time scheme (implicit in the fracture and explicit inthe matrix) alleviates the restriction on the size of the time stepfrom the CFL condition.

� We provide a generalized upstream weighting technique toapproximate the mobilities at the fracture intersections.

� The method has the flexibility to describe complicated fractureddomains with barriers.

� The MFE-DG method has other advantages in reducing thenumerical dispersion and convergence on unstructured meshesof low qualities [43].

Page 14: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

Fig. B.1. 2D fractured media with four matrix gridcells and three intersectingfractures.

Table B.1Mesh labels used in Fig. B.1

Number Labels

Matrix gridcells NmK ¼ 4 K1;K2;K3;K4

Fracture gridcells Nfk ¼ 3 k1; k2; k3

Matrix interfaces NmE ¼ 6 E1; E2; E3; E4; E5; E6

Fracture interfaces Nfe ¼ 4 e1; e2; e3; e4

904 H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905

Acknowledgement

This work was supported by the member companies of the Res-ervoir Engineering Research Institute (RERI). Their support isgreatly appreciated.

Appendix A. The MFE discretization

In the MFE method, the 3D fractured domain is discretized intoa mesh composed of hexahedrons, tetrahedrons, or prisms. Thefractures can, therefore, be quadrilaterals or triangles. Hybrid meshwith different geometrical gridcells can be readily handled by theMFE method (see Appendix B).

We define the following notation in the matrix and fractures:

NmK : number of gridcells in the matrix.

nmK : number of faces of the matrix gridcell K.

NmE : number of non-fracture interfaces in mesh.

Nfk: number of gridcells in the fracture network.

nfk: number of edges of the fracture gridcell k.

Nfe: number of fracture–fracture interfaces in the fracture

network.

The local integration of the velocity equation in the matrix andfracture gridcells yield to two types of elementary matrices:

In a matrix gridcell K , one gets an nmK � nm

K symmetric positivedefinite (SPD) matrix Am

K in terms of the permeability tensor eK m

and the RT0 basis function wmK [43], such that:

Am

K ¼ ½Am

K;E;E0 �E;E02oK ; Am

K;E;E0 ¼Z Z Z

Kwm

K;EeK�1;m

K wmK;E0 :

Similarly, in a fracture gridcell k, one gets an nfk � nf

k SPD matrixAf

k, such that:

Af

K ¼ ½Af

k;e;e0 �e;e02ok; Af

k;e;e0 ¼Z Z

kwf

k;eeK�1;f

k wfk;e0 :

The coefficients bmK;E, am

K;E, and amK in Eqs. (28) and (20) are defined

by:

bmK;E;E0 ¼ A�1;m

K;E;E0 ; amK;E ¼

XE02oK

bmK;E;E0 ; am

K ¼XE2oK

amK;E: ðA:1Þ

In the fracture network, afk, af

k;e, and afk are defined similarly to

those in Eq. (A.1).The different matrices used to construct the linear system in Eq.

(27) are defined in Table A.1, and the matrices appearing in Eq. (27)are given by:

Table A.1Definition of the matrices used to construct the linear system by the MFE method; thesymbols K and k refer to matrix and fracture gridcells, respectively, and E and e referto matrix–matrix and fracture–fracture interface, respectively

Matrix Dimensions Nonzero entries

Dm NmK � Nm

K DmK;K ¼ am

KeDf Nfk � Nf

k Dfk;k ¼ af

kMm;m Nm

E � NmE Mm;m

E;E0 ¼P

E;E03oK bmK;E;E0

Mm;f Nfk � Nf

K Mm;f

E;E0 ¼P

E;E03oK bmK;E;E0

Mf;f Nfe � Nf

e Mf;f

E;E0 ¼P

E;E03oK bmK;E;E0eMf Nf

e � Nfe

eMfe;e0 ¼

Pe;e03okb

fk;e;e0bMm Nm

E � NmE

bMmE;E0 ¼

Pe;e03okb

mK;E;E0bMf Nf

e � Nfe

bMfe;e0 ¼

Pe;e03okb

fk;e;e0

Rm;m NmK � Nm

E Rm;mK;E ¼ am

K;ERm;f Nm

K � Nfk Rm;f

K;E ¼ amK;EeRf Nf

k � Nfe

eRm;fk;e ¼ af

k;ebRm NmK � Nm

EbRm

k;e ¼ amK;EbRf Nf

k � Nfe

bRm;fk;e ¼ af

k;e

Am;m ¼ ðMm;m � RT;m;mD�1;mRm;mÞ;Am;f ¼ ðMm;f � RT;m;mD�1;mRm;fÞ;Af;m ¼ ðMf;m � RT;m;fD�1;mRm;mÞAf;f ¼ ðMf;f þ eDf � RT;m;fD�1;mRm;fÞ:

ðA:2Þ

The matrices Am;m and Af;f are symmetric and Am;f ¼ Af;m.

Appendix B. Structure of the linear system

Consider the 2D fractured media sketched in Fig. B.1. The matrixand fracture gridcells in Fig. B.1 are defined in Table B.1. The linearsystem in Eq. (27) for the calculation of the wetting-phase poten-tial for the 2D mesh in Fig. B.1 has the following structure.

E1 E2 E3 E4 E5 E6 k1 k2 k3 e1 e2 e3 e4

E1 � 0 0 0 0 � � 0 � 0 0 0 0E2 � 0 0 0 0 � � 0 0 0 0 0E3 � � 0 0 0 � 0 0 0 0 0E4 � � 0 0 � � 0 0 0 0E5 � 0 0 0 � 0 0 0 0E6 � � 0 � 0 0 0 0k1 � � � � � 0 0k2 � 0 � 0 � 0k3 � � 0 0 �e1 � � � �e2 � 0 0e3 � 0e4 �

0BBBBBBBBBBBBBBBBBBBBBBBBBBBBB@

1CCCCCCCCCCCCCCCCCCCCCCCCCCCCCA

References

[1] Barenblatt G, Zheltov Y, Kochina I. Basic concepts in the theory of seepage ofhomogeneous fluids in fissurized rocks. J Appl Math Mech1983;24(5):1286–303.

[2] Warren J, Root P. The behavior of naturally fractured reservoirs. SPE J1963:245–55.

Page 15: Advances in Water Resources · element (MFE) method which provides, in addition to the gridcell pressures, the pressures at the gridcell interfaces and can readily model the pressure

H. Hoteit, A. Firoozabadi / Advances in Water Resources 31 (2008) 891–905 905

[3] Kazemi H, Gilman J. Pressure transient analysis of naturally fracturedreservoirs with uniform fracture distribution. SPE J 1969(December):451–62.

[4] Thomas L, Dixon T, Pierson R. Fractured reservoir simulation. SPE J1983:42–54.

[5] Kazemi H, Gilman J. Analytical and numerical solution of oil recovery fromfractured reservoirs with empirical transfer functions. SPE J 1992(May):219–27.

[6] Kazemi H, Merrill L. Numerical simulation of water imbibition in fracturedcores. SPE J 1979(June):175–82.

[7] Pruess K, Narasimhan T. A practical method for modeling fluid and heat flow infractured porous media. SPE J 1985;25(1):14–26.

[8] Ghorayeb K, Firoozabadi A. Numerical study of natural convection anddiffusion in fractured porous media. SPE J 2000;5(1):12–20.

[9] Noorishad J, Mehran M. An upstream finite element method for solution oftransient transport equation in fractured porous media. Water Resour Res1982;18(3):588–96.

[10] Baca R, Arnett R, Langford D. Modeling fluid flow in fractured porous rockmasses by finite element techniques. Int J Num Meth Fluids 1984;4:337–48.

[11] Lee S, Jensen C, Lough M. Efficient finite-difference model for flow in areservoir with multiple length-scale fractures. SPE J 2000;5(3):268–75.

[12] Yortsos Y, Chang J. Capillary effects in steady-state flow in heterogeneouscores. Transport Porous Med 1990;5(4):399–420.

[13] Van Duijn C, Molenaar J, De Neef M. The effect of capillary forces on immiscibletwo-phase flow in heterogeneous porous media. Transport Porous Med1995;21(1):71–93.

[14] Van Duijn C, De Neef M. Similarity solution for capillary redistribution of twophases in a porous medium with a single discontinuity. Adv Water Res1998;21(6):451–61.

[15] Correa A, Firoozabadi A. Concept of gravity drainage in layered porous media.SPE J 1996(March):101–11.

[16] Slough K, Sudicky E, Forsyth P. Grid refinement for modeling multiphase flowin discretely fractured porous media. Adv Water Res 1999;23:261–9.

[17] Slough K, Sudicky E, Forsyth P. Numerical simulation of multiphase flow andphase partitioning is discretely fractured geologic media. J Contam Hydrol1999;40:107–36.

[18] Granet S, Fabrie P, Lemmonier P, Quitard M. A single phase flow simulation offractured reservoir using a discrete representation of fractures. In: Proceedingsof the 6th European conference on the mathematics of oil recovery (ECMORVI), September 8–11, Peebles, Scotland, UK; 1998.

[19] Juanes R, Samper J, Molinero J. A general and efficient formulation of fracturesand boundary conditions in the finite element method. Int J Numer Meth Eng2002;45:1751–74.

[20] Zhang K, Woodbury A. A krylov finite-element approach for multi-speciescontaminant transport in discretely fractured porous media. Adv Water Res2002;25:705–21.

[21] Kim J, Deo M. Comparison of the performance of a discrete fracture multiphasemodel with those using conventional methods. In: Reservoir simulationsymposium, 14–17 February, Houston, TX; 1999. No. SPE51928.

[22] Kim J, Deo M. Finite element discrete fracture model for multiphase flow inporous media. AIChE J 2000;46(6):1120–30.

[23] Karimi-Fard M, Firoozabadi A. Numerical simulation of water injection in 2Dfractured media using discrete-fracture model. SPEREE J2003;4(March):117–26.

[24] Karimi-Fard M, Durlofsky L, Aziz K. An efficient discrete-fracture modelapplicable for general-purpose reservoir simulators. SPE J 2004;9(2):227–36.

[25] Granet S, Fabrie P, Lemonnier P, Quitard M. A two-phase flow simulation of afractures reservoir using a new fissure element method. J Petrol Sci Eng2001;32(1):35–52.

[26] Bastian P, Helmig R. Efficient fully-coupled solution techniques for two-phaseflow in porous media. parallel multigrid solution and large scale computations.Adv Water Res 1999;23:199–216.

[27] Bastian P, Helmig R, Jakobs H, Reichenberger V. Numerical simulation ofmultiphase flow in fractured porous media. In: Numerical treatment ofmultiphase flows in porous media. In: Chen, Ewing, Shi, editors. Lecture notesin physics, vol. 552. Springer-Verlag; 2000. p. 1–18.

[28] Bogdanov I, Mourzenko V, Thovert J, Adler P. Two-phase flow throughfractured porous media. Phys Rev E 2003;68(2):1–24.

[29] Monteagudo J, Firoozabadi A. Control-volume method for numericalsimulation of two-phase immiscible flow in two- and three-dimensionaldiscrete-fractured media. Water Resour Res 2004;40(7):W07405.

[30] Geiger S, Roberts S, Matthäi S, Zoppou C, Burri A. Combining finite element andfinite volume methods for efficient multiphase flow simulations in highlyheterogeneous and structurally complex geologic media. GeoFluids2004;4(4):284–99.

[31] Matthäi S, Mezentsev A, Belayneh M. Control-volume finite-element two-phase flow experiments with fractured rock represented by unstructured 3Dhybrid meshes. In: Reservoir simulation symposium, 31 January–2 February,The Woodlands, TX; 2005.

[32] Monteagudo J, Firoozabadi A. Control-volume model for simulation of waterinjection in fractured media: Incorporating matrix heterogeneity and reservoirwettability effects. SPE J 2007;12(3):355–66.

[33] Reichenberger V, Jakobs H, Bastian P, Helmig R. A mixed-dimensional finitevolume method for two-phase flow in fractured porous media. Adv Water Res2006;29:1020–36.

[34] Hoteit H, Firoozabadi A. Numerical modeling of diffusion in fractured mediafor gas injection and recycling schemes. In: Annual technical conference andexhibition, 4–27 September, San Antonio, TX; 2006. No. SPE103292.

[35] Raviart P, Thomas J. A mixed hybrid finite element method for the secondorder elliptic problem. Lectures notes in mathematics, vol. 606. NewYork: Springer-Verlag; 1977.

[36] Brezzi F, Fortin M. Mixed and hybrid finite element methods. Environmentalengineering. New York: Springer-Verlag; 1991.

[37] Maryška J, Severyn O, Vohralík M. Numerical simulation of fracture flow with amixed-hybrid FEM stochastic discrete fracture network model. Comput Geosci2003;8(3):1499–573.

[38] Vohralík M, Maryška J, Severyn O. Mixed and nonconforming finite elementmethods on a system of polygons. Appl Numer Math 2007;57:176–93.

[39] Alboin C, Jaffré J, Roberts J, Serres C. Modeling fractures as interfaces for flowand transport in porous media. In: Chen, Ewing, editors. Fluid flow andtransport in porous media, vol. 295. American Mathematical Society; 2002. p.13–24.

[40] Martin V, Jaffré J, Roberts J. Modeling fractures and barriers as interfaces forflow in porous media. SIAM 2005;26:1667–91.

[41] Chavent G, Roberts J. A unified physical presentation of mixed, mixed-hybridfinite element method and standard finite difference approximations for thedetermination of velocities in water flow problems. Adv Water Res1991;14(6):329–33.

[42] Chavent G, Jaffré J. Mathematical models and finite elements for reservoirsimulation. Studies in mathematics and its applications. North-Holland: Elsevier; 1986.

[43] Hoteit H, Firoozabadi A. Numerical modeling of two-phase flow inheterogenous permeable media with different capillarity pressures. AdvWater Res 2008;31:56–73.

[44] Coats K. An equation of state compositional model. SPE J1980:363–76.

[45] Hoteit H, Firoozabadi A. Compositional modeling by the combineddiscontinuous Galerkin and mixed methods. SPE J 2006;11(1):19–34.

[46] Cockburn B, Shu C. TVB Runge–Kutta local projection discontinuous Galerkinfinite element method for conservative laws II: general frame-work. MathComp 1989;52:411–35.

[47] Cockburn B, Shu C. The Runge–Kutta discontinuous Galerkin method forconservative laws V: multidimensional systems. J Comput Phys1998;141:199–224.

[48] Monteagudo J, Firoozabadi A. Comparison of fully-implicit and IMPESformulations for simulation of water injection in fractured and unfracturedmedia. Inter J Numer Meth Eng 2006;69(4):698–728.

[49] Thomas J. Sur l’analyse numérique des méthodes d’élément finis hybrides etmixtes. Ph.D. thesis, Univ. de Pierre et Marie Curie, France; 1977.

[50] Hoteit H, Erhel J, Mosé R, Bernard P, Ackerer P. Numerical reliability for mixedmethods applied to flow problems in porous media. Comput Geosci2002;6(2):161–94.

[51] Hoteit H, Firoozabadi A. Compositional modeling by the combineddiscontinuous Galerkin and mixed methods. SPE J 2006;11(1):19–34.

[52] Hoteit H, Firoozabadi A. Compositional modeling of discrete-fractured mediawithout transfer functions by the discontinuous Galerkin and mixed methods.SPE J 2006;11(3):341–52.

[53] Hoteit H, Firoozabadi A. Multicomponent fluid flow by discontinuous Galerkinand mixed methods in unfractured and fractured media. Water Resour Res2005;41(11):W11412.