Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

SPE-169152-MS

Diffusion and Matrix-Fracture Interactions during Gas Injection in Fractured


Reservoirs
Hasan Shojaei and Kristian Jessen, University of Southern California

Copyright 2014, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Improved Oil Recovery Symposium held in Tulsa, Oklahoma, USA, 12–16 April 2014.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessar ily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohi bited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Molecular diffusion can play a significant role in oil recovery during gas injection in fractured reservoirs. Diffusion of gas
components from a fracture into the matrix extracts oil components from matrix and delays, to some extent, the gas
breakthrough. This in turn increases both sweep and displacement efficiencies.
In current simulation models, molecular diffusion is commonly modeled using a classical Fick’s law approach with
constant diffusion coefficients. In the classical Fick’s law approach, the dragging effects (off-diagonal diffusion coefficients)
are neglected. In addition, the gas-oil diffusion at the fracture-matrix interface is normally modeled by assuming an average
composition at the interface which does not have a sound physical basis.
In this paper, we present a dual-porosity model in which the generalized Fick’s law is used for molecular diffusion to
account for the dragging effects; and gas-oil diffusion at the fracture-matrix interface is modeled based on film theory in
which the gas in fracture and oil in the matrix are assumed to be at equilibrium. A novel shape factor is also introduced for
gas-oil diffusion based on film theory. Diffusion coefficients are calculated using the Maxwell-Stefan model and are
pressure, temperature and composition dependent. A time-dependent transfer function is used for matrix-fracture exchange in
which the shape factor is adjusted using a boost factor to differentiate between the transfer rate at early and late times.
Field-scale examples are used to demonstrate that the dragging effects (off-diagonal diffusion coefficients) can
significantly impact the oil recovery during gas injection in fractured reservoirs. It is also shown that using proper physical
models for matrix-fracture interactions (film theory for gas-oil diffusion and transfer function with boost factor) can
considerably affect the simulation results as compared to conventional models. We also show that miscibility is not
developed in the matrix blocks even at pressures above minimum miscibility pressure (MMP) when molecular diffusion is
the main recovery mechanism during gas injection in fractured reservoirs.
The work presented in this paper is directly applicable to the study and design of gas injection processes in fractured
reservoirs through an improved understanding of the effect of diffusion and matrix-fracture interactions on these processes.

Introduction
Gravity drainage, molecular diffusion and viscous displacement are known as the main recovery mechanisms during gas
injection in naturally fractured reservoirs. The relative significance of these mechanisms depends on several factors including
matrix permeability, fracture intensity, fluids properties, injection rate and reservoir pressure and temperature.
Viscous flow does not contribute directly to oil recovery because the injected gas channels through high-permeability
fractures which comprise only a few percentage of the total pore volume. Gravity drainage, which is driven by density
difference between oil and gas, plays a significant role when matrix permeability is high. Molecular diffusion may become
the dominant recovery mechanism in cases with low matrix permeability and high fracture intensity.
Contrary to conventional reservoirs where the impact of molecular diffusion is generally small, molecular diffusion can
play an important role in fractured reservoirs because of the large surface area available for diffusion. Different investigators
have demonstrated the efficiency of molecular diffusion in fractured reservoirs (e.g. da Silva and Belery, 1989; Hu and
Whitson, 1991; Darvish et al., 2006; Hoteit and Firoozabadi, 2009; Vega et al., 2010; Jamili et al., 2010).
Diffusion of gas components from a fracture into the matrix extracts oil components from matrix and delays, to some
extent, the gas breakthrough via high-permeability fractures. As a result, both displacement and sweep efficiencies are
increased. Gravity drainage has a similar effect on displacement and sweep efficiencies. In this paper we focus on molecular
diffusion and matrix-fracture exchange while more details on other drive mechanisms in fractured reservoirs can be found
2 SPE 169152

elsewhere (e.g. Lemonnier and Bourbiaux, 2010; Chordia and Trivedi, 2010; Rezaveisi et al., 2012).
Diffusive mass transfer in porous media can be attributed to three main mechanisms: Knudsen (free-molecule) diffusion,
molecular (continuum) diffusion and surface diffusion. Knudsen diffusion occurs when the pore size is smaller than the mean
free path of the molecules: Knudsen diffusion is dominant when the pore size is in the range of nanometers. Molecular
diffusion takes place when the pore sizes are relatively large compared to the mean free path of the molecules. Surface
diffusion describes the transport of matter in an adsorbed layer and involves interactions between surface and molecules
(Mason and Malinauskas, 1983). In this work Knudsen and surface diffusions are neglected because it is assumed that the
pores are in the range of micrometers and that there is no adsorbed layer on the walls of the pores.
Molecular diffusion is driven by gradients in concentrations (or more generally by gradients in the chemical potential).
Diffusion will oftentimes not play a significant role in regions of a porous media where fluids velocities are high (e.g.
fractures) because the characteristic time for diffusion is relatively large. However, in regions of the porous media where
flow velocities are low (e.g. matrix blocks) diffusion can dominate mass transfer over viscous flow (Perkins and Johnston,
1963; Shojaei et al., 2012).
Single-phase multicomponent molecular diffusion can be modeled using different approaches: The classical Fick’s law,
the generalized Fick’s law and the Maxwell-Stefan (MS) model. In the classical Fick’s law approach, the diffusion flux of
each component is only a function of its own concentration gradient. In other words, the interactions among different species
(dragging effects) are neglected. The generalized Fick’s law approach, in contrary, takes into account the component
interactions; i.e. the diffusion flux of each component depends on the concentration gradients of other components as well. In
the MS model, diffusion is driven by a gradient in the chemical potential and component interactions are accounted for using
component velocities (friction). It can be shown that under certain conditions the generalized Fick’s law and MS model are
equivalent (Taylor and Krishna, 1993).
In most reservoir simulation models (e.g. da Silva and Belery, 1989; Coats, 1989; Darvish et al., 2006; Vega et al., 2010;
Jamili et al., 2010), multicomponent molecular diffusion is modeled using a classical Fick’s law approach with effective
diffusion coefficients. This means the diffusion flux of each component will always be in the opposite direction of its
concentration gradient. However, it has been shown that because of the dragging effects, diffusion may occur from a region
of low to high concentration (reverse diffusion) (Duncan and Toor, 1962). Therefore ignoring the dragging effects (off-
diagonal diffusion coefficients) may lead to significant errors in simulation results as will be shown later.
Effective diffusion coefficients for multicomponent mixtures (classical Fick’s law) are usually calculated using the
approach of da Silva and Belery (1989) which is an extension of Sigmund’s correlation for binary mixtures (Sigmund, 1976).
These coefficients correspond to diagonal elements of the diffusion coefficient matrix. The full-matrix diffusion coefficients
(generalized Fick’s law) can be calculated using the MS model (e.g. Taylor and Krishna, 1993; Ghorayeb and Firoozabadi,
2000; Leahy-Dios and Firoozabadi, 2007) as will be explained later.
The gas-oil diffusion at the fracture-matrix interface (cross-phase diffusion) is usually modeled by assuming an average
composition at the interface (equipartition hypothesis). da Silva and Belery (1989) argue that for practical purposes, this is a
reasonable assumption for matrix blocks surrounded by interconnected fractures. However, the equipartition hypothesis and
gas-oil diffusion coefficient calculations do not have a sound physical basis.
In chemical engineering literature, the cross-phase diffusion has been modeled based on film theory in which oil and gas
are assumed to be in equilibrium at the interface, and component fluxes to be continuous across the interface (e.g. Krishna
and Standart, 1976; Taylor and Krishna, 1993). This approach has also been used in petroleum engineering to model gas-oil
diffusion in lab experiments (e.g. Hu and Whitson, 1991; Irani et al., 2009; Guo et al., 2009; Haugen and Firoozabadi, 2009;
Hoteit 2013).
In this paper, we present a dual-porosity model in which the generalized Fick’s law is used to represent the molecular
diffusion; and gas-oil diffusion at the fracture-matrix interface is modeled based on film theory. A novel shape factor is also
introduced for gas-oil diffusion based on film theory. Diffusion coefficients are calculated using the MS model and are
pressure, temperature and composition dependent. A time-dependent transfer function is used for the matrix-fracture
interactions in which the shape factor is adjusted using a boost factor to differentiate between the transfer rate at early and
late times.
Field-scale examples are used to demonstrate that the dragging effects (off-diagonal diffusion coefficients) can
significantly impact the oil recovery during gas injection in fractured reservoirs. It is also shown that using proper physical
models for matrix-fracture interactions (film theory for gas-oil diffusion and transfer function with boost factor) can
considerably affect the simulation results as compared to conventional models.
The remainder of this manuscript is organized as follows: The governing equations for transport of fluid in a dual-
porosity reservoir are first presented. Intra-phase and cross-phase diffusion are then discussed based on generalized Fick’s
law and film theory. A generalized matrix-fracture transfer function is subsequently presented. Different physical models for
molecular diffusion and matrix-fracture interactions are then compared using field-scale examples. A discussion and analysis
of the presented results concludes the manuscript.
SPE 169152 3

Mathematical Model
We use a dual-porosity approach to model gas injection in fractured reservoirs. Dual-porosity models assume there are two
communicating domains in a fractured reservoir: a flowing domain (fracture) and a stagnant domain (matrix). Mass (or
volume) balance equations are solved independently for these two domains. Transfer of fluids between these two overlapped
domains is accounted for using a source/sink term in the mass (or volume) balance equations (Barenblatt et al., 1960; Warren
and Root, 1963).
We choose dual-porosity formulation because it provides a practical representation of fractured reservoirs with
reasonable accuracy and computational efficiency. Other approaches such as fine-grid single-porosity and discrete fracture
models are computationally expensive and need extensive data which are not generally available (e.g. fracture distribution in
the reservoir). In addition, a dual-porosity model can be constructed from detailed discrete fracture characterization using
proper upscaling techniques (e.g. Karimi-Fard et al., 2006; Gong et al., 2008; Evazi and Mahani, 2010).
The mass conservation equation for a component i in the flowing domain (fracture) containing np phases and nc
components, where advection, diffusion and gravity are the main physical mechanisms, can be written as (Jessen et al., 2008)
(1)
while the mass conservation equation for a component i in the stagnant domain (matrix) can be written as
(2)
where the overall molar density Ci is given by

(3)

the overall molar advective flux Fi is given by

(4)

and the overall molar diffusive flux Hi is given by

(5)

where Hi,j denotes the molar diffusive flux of component i within phase j (intra-phase diffusion) and Hi,jk represents the molar
diffusive flux of component i at the interface between phases j and k (cross-phase diffusion). More details on calculation of
molar diffusive fluxes due to intra-phase and cross-phase diffusion are given in the subsequent section.
In Equations (1) through (4), ϕ is the porosity; xij is the mole fraction of component i in phase j; ρj and Sj represent the
molar density and saturation of phase j respectively; vj is the velocity vector of phase j; qi is the source term of component i;
and Γi represents the transfer of component i from matrix to fracture.
The phase velocities in Eq. (1) are evaluated from the pressure field which is obtained by solving the following volume-
balance equation

(6)

where ct is the total fluid compressibility; p is the pressure; Vt is the total fluid volume; Vcell is the pore volume; and ni is the
number of moles of component i. Capillarity is not included because we focus on gas injection processes in which IFT effects
are less significant.
We use a finite volume IMPES (implicit pressure explicit saturation/composition) formulation with a Cartesian grid in
this work. This means at each new time level, Eq. (6) is solved with coefficients fixed at the old time level. The discrepancy
between cell volume and fluid volume is minimized by carrying errors forward in time (Trangenstein and Bell, 1989; Jessen
et al., 2008).
Once the pressure field is obtained at the new time level using Eq. (6), phase velocities at the gridblock interfaces are
obtained using Darcy’s law
(7)
where K is the permeability tensor; λj and ρmj are the mobility and mass density of phase j; and g is the gravitational vector.
The total number of moles of each component in a gridblock l is then updated using

(8)

where Alm denotes the area connecting gridblocks l and m; (F+H)i,lm is the total molar flux (advective+diffusive) of
component i out of gridblock l at the interface lm, and qi,l represent the source term. The phase-equilibrium calculations are
performed using Peng-Robinson (PR) equation of state (EOS). The advective flux is calculated using the standard single
4 SPE 169152

point upwind (SPU) scheme.

Molecular Diffusion
Molecular diffusion may occur within a single phase (intra-phase) or between two different phases (cross-phase). For two
neighboring gridblocks (l and m) that contain oil and gas phases (Fig. 1), the cross sectional areas that are available for gas-
gas, oil-oil and gas-oil diffusion are Ag=Alm×min(Sg,l,Sg,m), Ao=Alm×min(So,l,So,m) and Ac=Alm-Ag-Ao respectively.

Figure 1: Schematic of two neighboring gridblocks containing oil and gas phases

In the following subsections we explain how to calculate intra-phase and cross-phase molecular diffusion based on
generalized Fick’s law and film theory.

Intra-phase Diffusion
The molar diffusive flux of a component i in phase j (generalized Fick’s law) is given by

(9)

where Dik,j are the multicomponent diffusion coefficients in phase j. The sum of diffusive fluxes of all components in phase j
must be zero ( ); and the diffusive flux for the last component (Hnc,j) is obtained accordingly (Taylor and
Krishna, 1993). The diagonal elements in the diffusion coefficients matrix D are called the main diffusion coefficients; while
the off-diagonal elements are called the cross-diffusion coefficients which are generally non-zero and non-symmetric (i.e. Dik
≠ Dki for i≠k). There are (nc-1)2 Fickian diffusion coefficients which are dependent on the numbering of components.
Multicomponent diffusion coefficients can be calculated using the following equation which is obtained by comparing
the generalized Fick’s law and MS model
(10)
where matrix B is a function of the inverse of the MS coefficients and γ represents the non-ideal behavior of the mixture. The
elements of the matrices B and γ are given by the following equations

(11)

and
(12)
where Đik represent the MS coefficients; δik is the Kronecher delta function and φi is the fugacity coefficient of component i.
The MS coefficients matrix is symmetric and its diagonal elements do not exist. Therefore there are nc(nc-1)/2 MS
coefficients which can be obtained using (Wesselingh and Krishna, 1990; Kooijman and Taylor, 1991)

(13)

where Đoik is the diffusion coefficient of component i infinitely diluted in component j. In this work we use the correlation by
Leahy-Dios and Firoozabadi (2007) to calculate the infinite dilution diffusion coefficients for both vapor and liquid phases.

Cross-phase Diffusion
Let us consider two different phases which have been placed in contact as shown in Fig. 2. According to the film theory,
thermodynamic equilibrium will prevail at the interface between the two phases. In addition, the interface is assumed to offer
no resistance to mass transfer. In other words, there will be no accumulation at the interface, provided there are no interface
SPE 169152 5

chemical reactions (continuity of molar fluxes). In Fig. 2, xi,I and yi,I are the interface compositions; and xi,b and yi,b denote the
bulk phase compositions (Taylor and Krishna, 1993).

Figure 2: Composition profiles near the interface during inter-phase mass transfer, after Taylor and Krishna (1993)

Therefore the cross-phase mass transfer for a stationary interface is governed by two sets of equations: continuity of molar
fluxes at the interface
(14)
and thermodynamic equilibrium
(15)
where, assuming the positive direction is from left to right, denotes the total molar flux of component i from “x” phase
into the interface; denotes the total molar flux of component i from interface into “y” phase; and ( ) is the chemical
potential of component i in “x” (“y”) phase at the interface.
The molar flux consists of advective and diffusive fluxes. Since the advective flux is computed using an upwind scheme,
the continuity of advective flux will always be honored. Therefore the continuity of the molar flux (Eq. 14) reduces to the
continuity of diffusive flux which, based on the generalized Fick’s law, can be written in the following finite difference form

(16)

where Lx (Ly) is the distance between the interface and the center of the gridblock containing the “x” (“y”) phase. To
calculate the molar cross-phase diffusive fluxes, Equations (15) and (16) need to be solved simultaneously. The interface
equilibrium calculations (Eq. (15)) can be done using an EOS at interpolated temperature and pressure. Multicomponent
diffusion coefficients are obtained using Eq. (10).
The solution method described in the above paragraph, is an iterative and hence time-consuming approach. Hoteit (2013)
proposed a non-iterative solution based on gradients of the chemical potential (instead of concentration gradients) which does
not need explicit knowledge of interface compositions. In this approach, which is used in our work, the continuity of
diffusive fluxes is given by

(17)

where ; B is the matrix given by Eq. (11); X is a diagonal matrix with diagonal elements [1/xk]k=1,…,nc-1;
ψk=ln(fk); and fk is the fugacity of component k. We note that because of the chemical equilibrium condition at the interface
(Eq. (15)), we have = . Eq. (17) can be rearranged to obtain the following expression in matrix form
(18)
where and . Once is calculated, it can be substituted in either left-hand side or right-hand side
of Eq. (17) to calculate the molar diffusive flux of each component across the interface. The sum of diffusive fluxes of all
components across the interface must be zero; and the diffusive flux of the last component is obtained accordingly.

Transfer Function
Since the introduction of dual-porosity models in the 1960s, various transfer functions with different levels of sophistication
have been suggested for matrix-fracture interactions. Warren and Root (1963) proposed a pseudo-steady state transfer
function for single-phase flow between matrix and fracture. In their model, the transfer rate per unit bulk volume is obtained
by multiplying the average pressure difference between matrix and fracture by mobility and a parameter which is commonly
referred to as the shape factor. The shape factor has a dimension of reciprocal area and is a characteristic of the fractured
rock.
The application of the shape factor in reservoir simulation was first introduced by Kazemi et al. (1976) for three-phase
flow subject to a pseudo-steady state assumption. A gravity term was later included in the transfer function by Gilman and
6 SPE 169152

Kazemi (1983), Gilman (1986) and Sonier et al. (1988). However, it was added for flow between matrix and fracture in all
three directions, which could lead to an overestimation of the transfer rate. Quandallet and Sabathier (1989) resolved this
issue by treating flow in horizontal and vertical directions separately.
Coats (1989) extended the dual-porosity model for compositional simulation and included molecular diffusion in the
transfer function. He suggested the use of shape factors that are exactly twice those of Kazemi et al. (1976). By solving the
single-phase pressure diffusion equation and approximating the exact solution with simplified but reasonably accurate
functions, Zimmerman et al. (1993) and Lim and Aziz (1995) derived matrix-fracture transfer functions without assuming a
pseudo-steady state condition. While pseudo-steady state transfer functions are only good for late times, the transfer functions
developed by Zimmerman et al. (1993) and Lim and Aziz (1995) are able to provide reasonable accuracy for both early and
late times.
Based on the work of Zimmerman et al. (1993), Lu et al. (2008) proposed a general transfer function that accounts for
both early- and late-time behavior of multi-phase matrix-fracture interactions. Their approach, which is used in our work with
some modifications, is explained in the following paragraphs. The transfer rate of a component i from matrix to fracture is
given by
(19)
where Γi,e, Γi,d and Γi,gd are transfer rates due to expansion, diffusion and gravity drainage respectively. We note again that
capillary effects are assumed to be negligible for gas injection processes. The expansion term is given by

(20)

where σ1 is the shape factor; Be is the boost or correction factor for the expansion term; Km is the (equivalent isotropic) matrix
permeability; λjm is the mobility of phase j in the matrix; xij,m is the mole fraction of component i in phase j in the matrix; ρjm
is the molar density of phase j in the matrix; and pm and pf are the pressure in matrix and fracture respectively. The shape
factor proposed by Lim and Aziz (1995) for systems with anisotropic rectangular matrix blocks is used in this work
(21)
where Km,x, Km,y and Km,z are the matrix permeabilities in x-, y-, and z-direction respectively; and Lx, Ly, Lz are the dimensions
of matrix block in x-, y-, and z-direction respectively. The equivalent isotropic matrix permeability is defined as
(22)
The boost factor for expansion is given by the following equation (Zimmerman et al., 1993; Lu et al., 2008)

(23)

where pm,init stands for the initial pressure in the matrix. At late times, when matrix and fracture pressures are similar, the
boost factor approaches unity and Eq. (20) reduces to Kazemi’s pseudo-steady state transfer function. However, at early
times when the matrix pressure is close to its initial value, the correction factor is larger than one.
To calculate the transfer rate of each component from matrix to fracture due to molecular diffusion, we sum the intra-
phase and cross-phase diffusion rates of that component. For a two-phase gas-oil system, the transfer rate of a component i is
calculated from
(24)
where Γi,d,g, Γi,d,o and Γi,d,c are transfer rates due to gas-gas (intra-phase), oil-oil (intra-phase) and gas-oil (cross-phase)
molecular diffusion respectively. The transfer rate of a component i due to gas-gas diffusion is given by

(25)

where Bd,g is the boost factor for gas-gas diffusion; ϕm is the matrix porosity and Sg=min(Sgm,Sgf) is the fraction of the total
matrix-fracture surface area that is available for gas-gas diffusion. Since the molecular diffusion equation is analogous to the
pressure diffusion equation, the following expression can be used to calculate the boost factor for gas-gas diffusion

(26)

The transfer rate of a component i due to oil-oil diffusion is calculated similarly to gas-gas diffusion. The transfer rate of
each component from matrix to fracture due to gas-oil diffusion is calculated based on the film theory as explained earlier.
For the matrix-fracture system, the continuity of diffusive flux across the interface can be written as

(27)

where Sgo=1- min(Sgm,Sgf) - min(Som,Sof) is the fraction of the total matrix-fracture surface area that is available for gas-oil
diffusion; σ1 is obtained from an isotropic version of Eq. (21); and σ2 is given by
SPE 169152 7

(28)
where bx, by, and bz are the fracture openings in x-, y-, and z-direction respectively. The derivations of Eqs. (27) and (28) are
presented in Appendix A.
By rearranging Eq. (27), we arrive at an equation similar to Eq. (18) in matrix form
(29)
where and . Once is calculated, it can be substituted in either left-hand side or right-
hand side of Eq. (27) to calculate the cross-phase molar diffusive flux of each component from matrix to fracture. The boost
factor for gas-oil diffusion can be obtained from

(30)

The transfer rate of each component from matrix to fracture due to gravity drainage is calculated using the equations
presented in Appendix B.

Results
Two numerical examples are presented in this section to demonstrate the significance of proper modeling of molecular
diffusion and matrix-fracture interactions during gas injection in fractured reservoirs. In both cases, the size of reservoir in x-,
y-, and z-direction is 250 m, 250 m, and 10 m respectively. 25×25×1 gridblocks are used for both flowing and stagnant
domains. The flowing domain has a permeability of 500 md in all three directions; and a porosity of 1%. In the stagnant
domain, the permeabilities in horizontal directions (K x and Ky) are 1 md while the vertical permeability is 0.001 md. The
stagnant domain has a porosity of 20%. The fracture spacing and opening in both x- and y-direction are 10 m and 1 mm
respectively.
In both examples, one injection and one production well are located at opposite corners of the reservoir. The bottomhole
pressure is kept constant for both injection and production wells, with values that will be specified for each example. Linear
relative permeabilities are used for the flowing domain (fractures); while Corey-type relative permeability curves, with
parameters given in Table 1, are used for the stagnant domain (matrix).

Table 1 Parameters for Corey-type relative permeabilities


Sgc Sor ng no krg,e kro,e
Flowing Domain 0.00 0.00 1.0 1.0 1.0 1.0
Stagnant Domain 0.00 0.25 3.0 4.0 1.0 1.0

Example 1
In this example, the initial reservoir pressure is 157 bar. The bottomhole pressure for the production well is kept constant at
157 bar and gas is injected at an average rate of 128 Rm3/day. The injected gas is composed of 90% CO2 and 10% CH4 by
mole. The reservoir is initially saturated with oil of 30% CH 4, 40% nC4H10 and 30% nC12H26 by mole. The reservoir
temperature is 373.15 °K at which the minimum miscibility pressure (MMP) of the oil/gas pair is 150.5 bar. The oil has a
bubble point pressure of 91.6 bar at the reservoir temperature.
Fig. 3 compares the simulation results for example 1 for the cases when diffusion is not considered (blue), when
diffusion is modeled using classical Fick’s law (green), when diffusion is modeled using generalized Fick’s law (red), and
when diffusion is modeled using generalized Fick’s law with boost factors for matrix-fracture interactions (magenda). For all
of these four cases, the gas-oil diffusion is modeled based on film theory. The left tab shows liquid recovery (%) and the right
tab shows producing gas-oil ratio (GOR) versus time for ten years. In the absence of diffusion, the injected gas mainly
sweeps the oil residing in the flowing domain (fracture) which in turn leads to very low liquid recovery and very high
producing GOR. However, when molecular diffusion is present, diffusion of gas components from fracture into the matrix
extracts oil components from matrix and leads to substantially higher liquid recoveries and lower producing GOR.
Significant differences are observed between predictions by the generalized Fick’s and the classical Fick’s approaches
which mean the dragging effects (off-diagonal diffusion coefficients) can significantly impact the molecular diffusion during
gas injection in fractured reservoirs. The generalized Fick’s law approach predicts higher liquid recoveries during the first ten
years of production because the larger contributions by the positive off-diagonal coefficients in this case. Inclusion of boost
factors in the matrix-fracture transfer function can also considerably affect the simulation results as observed from Fig. 3:
Significantly higher liquid recoveries are predicted during the first ten years of production for this case.
A common approach in commercial simulators is to model molecular diffusion using classical Fick’s law with constant
diffusion coefficients. In addition, the gas-oil diffusion is normally modeled by assuming an average composition at the
interface. By assuming an average interface composition in Eq. (16), one can obtain expressions similar to Eqs. (27) and (29)
for gas-oil diffusion. The results obtained from this simplified approach for example 1 are also included in Fig. 3 (black). The
diagonal elements of the diffusion coefficient matrix obtained from Eq. (10) at initial and injection conditions are used for the
entire simulation (constant diffusion coefficients) for oil and gas phases respectively. Although this approach initially
8 SPE 169152

provides acceptable results, it is unable to capture the behavior of the molecular diffusion and matrix-fracture interactions at
later times as can be seen from Fig. 3. We defer further analysis of the differences between predictions by the simplified
approach and more sophisticated models until the discussion section.

Figure 3: Liquid recovery (left) and producing GOR (right) versus time for example 1 with no diffusion (blue), diffusion
from classical Fick’s law and film theory for gas-oil diffusion (green), diffusion from generalized Fick’s law and film theory
for gas-oil diffusion (red), diffusion from generalized Fick’s law and film theory for gas-oil diffusion with boost factor
(magenda), diffusion from classical Fick’s law with constant diffusion coefficients and assuming average composition at the
gas-oil interface (black).

The CO2 composition (mole %) in the stagnant domain (matrix) after ten years (example 1) is shown in Fig. 4 for the case
with no diffusion (left) and the case with diffusion based on generalized Fick’s law and film theory with boost factors for
matrix-fracture interactions (right). It is observed that, in the absence of molecular diffusion, very small amounts of CO2 have
entered the stagnant domain due to viscous displacement and gravity drainage. However when molecular diffusion is present,
considerable amounts of CO2 enter the stagnant domain and extract the oil components. We note that in this figure the
injection well is located at the lower-left corner while the producer is located at the upper-right corner.

Figure 4: The CO2 composition (%) in the stagnant domain after 10 years (example 1) for the case with no diffusion (left) and
the case with diffusion based on generalized Fick’s law and film theory with boost factors.
SPE 169152 9

Example 2
In this example, the initial reservoir pressure is 150 bar, the bottomhole pressure of the production well is kept constant at
150 bar and gas is injected at an average rate of 158 Rm3/day. The injected gas is composed of 50% CO2 and 50% CH4 by
mole. The reservoir is initially saturated with oil of the same compostion as in example 1 (30% CH4, 40% nC4H10 and 30%
nC12H26 by mole) at a temperature of 336.15 °K. The MMP of the oil/gas pair is 178.8 bar; and the oil has a bubble point
pressure of 78.6 bar. Fig. 5 compares the simulation results with different physical models for molecular diffusion and
matrix-fracture interactions for example 2. The same color codes as in Fig. 3 are used to represent each physical model. The
left tab shows liquid recovery (%) while the right tab presents producing GOR versus time for ten years.
Simliar to the previous example, low liquid recovery and high producing GOR is observed in the absence of molecular
diffusion. However, when molecular diffusion is present, substantially higher liquid recoveries and lower producing GOR are
obtained. The generalized Fick’s law approach (magenta) again predicts considerably higher liquid recoveries during the first
ten years as compared to the classical Fick’s law approach (green). Significantly higher liquid recoveries are predicted during
the first ten years of production when boost factors are included in the matrix-fracture transfer function as observed from Fig.
5. Similar to what we observed for example 1, the simplified approach (classical Fick’s law with constant diffusion
coefficients and assuming average composition at the gas-oil interface) is not able to predict the behavior of the molecular
diffusion and matrix-fracture interactions with reasonable accuracy.

Figure 5: Liquid recovery (left) and producing GOR (right) versus time for example 2 with no diffusion (blue), diffusion
from classical Fick’s law and film theory for gas-oil diffusion (green), diffusion from generalized Fick’s law and film theory
for gas-oil diffusion (red), diffusion from generalized Fick’s law and film theory for gas-oil diffusion with boost factor
(magenda), diffusion from classical Fick’s law with constant diffusion coefficients and assuming average composition at the
gas-oil interface (black).

Fig. 6 shows the CO2 composition (mole %) in the stagnant domain (matrix) after ten years (example 2) for the case with no
diffusion (left) and for the case with diffusion based on generalized Fick’s law and film theory with boost factors for matrix-
fracture interactions (right). Very small amounts of CO2 have entered the stagnant domain (due to viscous displacement and
gravity drainage) when molecular diffusion is absent. However, considerable amounts of CO 2 enter the stagnant domain and
extract the oil components when molecular diffusion is present.
10 SPE 169152

Figure 6: The CO2 composition (%) in the stagnant domain after 10 years (example 2) for the case with no diffusion (left) and
the case with diffusion based on generalized Fick’s law and film theory with boost factors.

Discussion
To understand how the oil is recovered by molecular diffusion during gas injection in fractured reservoirs, we analyze the
saturation and composition profiles of a matrix block and its corresponding fracture. Fig. 7 shows the saturation profile
observed during the first ten years of production in a matrix block (dotted line) near the injection well and its corresponding
fracture (solid line) for example 1 as calculated by the model with generalized Fick’s law and film theory with boost factors.
The fracture becomes fully saturated with gas in a short period of time; while the gas saturation in the matrix remains zero for
almost one year, after which it gradually increases.

Figure 7: The saturation profile of a matrix block (dotted line) near the injection well and its corresponding fracture (solid
line) for example 1 as calculated by the case with generalized Fick’s law and film theory with boost factors.

Fig. 8 shows the composition routes (mass fraction) of a matrix block (green) near the injection well and its corresponding
fracture (red) for example 1 during the first ten years of production as calculated by the model with generalized Fick’s law
and film theory with boost factors. The blue curves show the two-phase boundaries in each ternary plane at 157 bar and
373.15 °K; while the dashed brown line represents the dilution line.
SPE 169152 11

Figure 8: Composition profiles (mass fraction) of a matrix block (green) near the injection well and its corresponding fracture
(red) for example 1 during the first 10 years as calculated by the generalized Fick’s law and film theory with boost factors.
The blue curves show the two-phase boundaries in each ternary plane at 157 bar and 373.15 °K. The dashed brown line
represents the dilution line.

Although the pressures in example 1 are above the MMP, miscibility is not developed in the matrix block. As observed from
Fig. 8, molecular diffusion generates a composition profile deep into the two-phase region. Initially, the CH4 composition
increases with a slight icrease in CO2 composition; while the nC4H10 (nC4) composition decreases and nC12H26 (C12)
composition remains almost unchanged. Later, mainly CO2 composition increases and C12 composition decreases until the
composition in the matrix approaches injection composition. It is also observed that the composition profile in the fracture
departs from the multi-contact miscibility (MCM) condition due to exchange with the matrix (and the limited number of cells
used in the representation of the domain).
Fig. 9 shows the saturation profile observed during the first ten years of production at the same matrix block (dotted line)
as in Fig. 7 and its corresponding fracture (solid line) as calculated by the simplified approach (classical Fick’s law with
constant coefficients and average composition at the gas-oil interface). As before, the fracture becomes fully saturated with
gas in a short period of time; while the gas saturation in the matrix remains zero for almost eight years, after which it
gradually increases. We note that a gaseous phase was formed in the matrix after ~ one year when the sophisticated model
was used (see Fig. 7).

Figure 9: The saturation profile of a matrix block (dotted line) near the injection well and its corresponding fracture (solid
line) for example 1 as calculated by the simplified approach (classical Fick’s law with constant coefficients and assuming
average composition at the gas-oil interface).
12 SPE 169152

Fig. 10 can explain why a gaseous phase does not form in the matrix until after ~ eight years as calculated by the simplified
approach. During the first eight years, the CO 2 composition largely increases with small changes in compositions of other
components. This keeps the fluid inside the matrix in the single phase region for nearly 8 years. The composition profile in
the matrix, as calculated by the simplified approach, differs significantly from what is calculated using the model based on
generalized Fick’s law and film theory with boost factor. This can explain why the recovery calculations by the simplified
approach significantly differ from the more sophisticated models. We note that the composition profile in Fig. 10 represents
the first 20 years of production.

Figure 10: Composition profiles (mass fraction) of a matrix block (green) near the injection well and its corresponding
fracture (red) for example 1 during the first 20 years as calculated by the simplified approach (classical Fick’s law with
constant coefficients and assuming average composition at the gas-oil interface). The blue curves show the two-phase
boundaries in each ternary plane at 157 bar and 373.15 °K. The dashed brown line represents the dilution line.

Finally, the CO2 composition (mole %) in the stagnant domain (matrix) after ten years (example 1) as calculated by the
simplified approach (left) is compared with calculations based on generalized Fick’s law and film theory with boost factor
(right) in Fig. 11. It is observed that the two different approaches predict significantly diffenet distributions of CO 2
composition in the stagnant domain after 10 years.

Figure 11: The CO2 composition (%) in the stagnant domain after 10 years (example 1) as calculated by the simplified
approach (left) calculations based on generalized Fick’s law and film theory with boost factors (right).
SPE 169152 13

Conclusions
In this paper, we presented a dual-porosity model in which the generalized Fick’s law is used to represent molecular
diffusion; and the gas-oil diffusion at the fracture-matrix interface is modeled based on film theory. A novel shape factor was
also introduced for gas-oil diffusion based on film theory. In the matrix-fracture transfer function, a boost factor is used to
adjust the shape factor in order to account for the differences between the transfer rate at early and late times. The following
conclusions can be drawn based on the results presented in this work:
 Molecular diffusion can substantially enhance the oil recovery by gas injection in fractured reservoirs. Diffusion of gas
components from fracture into the matrix extracts oil components from matrix and leads to substantially higher liquid
recoveries and lower producing GOR.
 The dragging effects (off-diagonal diffusion coefficients) can significantly impact the oil recovery during gas injection in
fractured reservoirs.
 Miscibility is not developed in the matrix block even at pressures above MMP when molecular diffusion is the main
recovery mechanism during gas injection in fractured reservoirs. In this case, molecular diffusion pushes the composition
profile deep into the two-phase region.
 Using proper physical models for matrix-fracture interactions (film theory for gas-oil diffusion and transfer function with
boost factor) can considerably affect the simulation results as compared to conventional models.
 The conventional approach (classical Fick’s law with constant diffusion coefficients and assuming an average
composition at the gas-oil interface) is unable to accurately predict the behavior of molecular diffusion and matrix-
fracture interactions during gas injection in fractured reservoirs.

Nomenclature
A Cross sectional area
Bd Boost factor for diffusion term
Be Boost factor for expansion term
Bgd Boost factor for gravity drainage term
ct Total fluid compressibility
C Overall molar density
D Diffusion coefficient
Đ MS coefficient
Đo Infinite dilution diffusion coefficient
f Fugacity
F Molar advective flux
g Gravitational vector
H Molar diffusive flux
kr Relative permeability
krg,e End-point relative permeability of gas
kro,e End-point relative permeability of oil
K Absolute permeability
n Number of moles
ng Corey-exponent for gas
no Corey-exponent for oil
N Total molar flux
p Pressure
q Source term in the mass conservation equation
S Saturation
Sgc Critical gas saturation
Sor Residual oil saturation
t Time
Vt Total fluid volume
xij Mole fraction of component i in phase j
Γd Matrix-fracture transfer rate due to diffusion
Γe Matrix-fracture transfer rate due to expansion
Γgd Matrix-fracture transfer rate due to gravity drainage
δ Kronecher delta function
λ Mobility
v Velocity
ρ Density
σ Shape factor
φ Fugacity coefficient
14 SPE 169152

ϕ Porosity
ψ Natural log of fugacity

References
Barenblatt, G.I., Zheltov, 1u.P. and Kochina, I.N., 1960. Basic concepts on the theory of seepage of homogeneous liquids in
fissured rocks. Prikladnaya Matematikai Mekhanika, Akad. Nauk, SSSR 24(5): 852-864.
Chordia, M., and Trivedi, J. 2010. Diffusion in Naturally Fractured Reservoirs-A-Review. Paper SPE 77339 presented at
the SPE Asia Pacific Oil and Gas Conference and Exhibition, Brisbane, Australia, 18-20 October.
Coats, K.H. 1989. Implicit Compositional Simulation of Single-Porosity and Dual-Porosity Reservoirs. Paper SPE 18427
presented at the SPE Symposium on Reservoir Simulation, Houston, TX, 6-8 February.
da Silva, F.V., and Belery, P. 1989. Molecular Diffusion in Naturally Fractured Reservoirs: A Decisive Recovery
Mechanism. Paper SPE 19672 presented at the SPE Annual Technical Conference and Exhibition, San Antonio, TX, 8-
11 October.
Darvish, G., Lindeberg, E., Holt, T., Kleppe, J., and Arild Utne, S. 2006. Reservoir Conditions Laboratory Experiments of
CO2 injection into Fractured Cores. Paper SPE 99650 presented at the SPE Europec/EAGE Annual Conference and
Exhibition, Vienna, Austria, 12-15 June.
Di Donato, G., Tavassoli, Z., and Blunt, M.J. 2006. Analytical and Numerical Analysis of Oil Recovery by Gravity
Drainage. Journal of Petroleum Science and Engineering 54(1): 55-69.
Duncan, J.B., and Toor, H.L. 1962. An Experimental Study of Three Component Gas Diffusion. AIChE Journal 8(1): 38-41.
Evazi, M., and Mahani, H. 2010. Generation of Voronoi Grid Based on Vorticity for Coarse-Scale Modeling of Flow in
Heterogeneous Formations. Transport in Porous Media 83(3): 541-572.
Ghorayeb, K., and Firoozabadi, A. 2000. Molecular, Pressure, and Thermal Diffusion in Nonideal Multicomponent
Mixtures. AIChE Journal 46(5): 883-891.
Gilman, J., and Kazemi, H. 1983. Improvements in Simulation of Naturally Fractured Reservoirs. SPE Journal 23(4): 695-
707.
Gilman, J.R. 1986. An Efficient Finite-Difference Method for Simulating Phase Segregation in the Matrix Blocks in Double-
Porosity Reservoirs. SPE Reservoir Engineering 1(4): 403-413.
Gong, B., Karimi-Fard, M., and Durlofsky, L. 2008. Upscaling Discrete Fracture Characterizations to Dual-Porosity, Dual-
Permeability Models for Efficient Simulation of Flow with Strong Gravitational Effects. SPE Journal 13(1): 58-67.
Guo, P., Wang, Z., Shen, P., and Du, J. 2009. Molecular Diffusion Coefficients of the Multicomponent Gas–Crude Oil
Systems under High Temperature and Pressure. Ind. Eng. Chem. Res. 48(19): 9023–9027.
Haugen, K., Firoozabadi, A. 2009. Composition at the Interface between Multicomponent Non-Equilibrium Phases. J. Chem.
Phys. 130(6): 64707.
Hoteit, H., and Firoozabadi, A. 2009. Numerical Modeling of Diffusion in Fractured Media for Gas-Injection and –Recycling
Schemes. SPE Journal 14: 323-337.
Hoteit, H. 2013. Modeling Diffusion and Gas-Oil Mass Transfer in Fractured Reservoirs. Journal of Petroleum Science and
Engineering 105: 1-17.
Hu, H., Whitson, C.H., and Qi, Y. 1991. A Study of Recovery Mechanisms in a Nitrogen Diffusion Experiment. Paper SPE
21893 presented at the 6th European Symposium on Improved Oil Recovery.
Irani, M., Bozorgmehry R., Pishvaie, M.R., and Tavasoli, A. 2009. A Study on the Effects of Thermodynamic Non-Ideality
and Mass Transfer on Multi-Phase Hydrodynamics Using CFD Methods. World Academy of Science, Engineering and
Technology 34: 627-632.
Jamili, A., Willhite, G.P., and Green, D. 2011. Modeling Gas-Phase Mass Transfer between Fracture and Matrix in Naturally
Fractured Reservoirs. SPE Journal 16(4): 795-811.
Jessen, K., Gerritsen, M., and Mallison, B. 2008. High-Resolution Prediction of Enhanced Condensate Recovery
Processes. SPE Journal 13(2): 257-266.
Karimi-Fard, M., Gong, B., and Durlofsky, L.J. 2006. Generation of Coarse-Scale Continuum Flow Models from Detailed
Fracture Characterizations. Water Resources Research 42(10): W10423.
Kazemi, H., Merrill, L.S., Porterfield, K.L., and Zeman, P.R. 1976. Numerical Simulation of Water-Oil Flow in Naturally
Fractured Reservoirs. SPE Journal 16(6): 317-326.
Kooijman, H.A., and Taylor, R. 1991. Estimation of Diffusion Coefficients in Multicomponent Liquid Systems. Industrial &
Engineering Chemistry Research 30(6): 1217-1222.
Krishna, R., and Standart, G. 1976. Determination of Mass and Energy Fluxes for Multicomponent Vapour–Liquid Systems.
Letts Heat Mass Transfer 3: 173–182.
Leahy-Dios, A., and Firoozabadi, A. 2007. Unified Model for Nonideal Multicomponent Molecular Diffusion
Coefficients. AIChE Journal 53(11): 2932-2939.
Lemonnier, P., and Bourbiaux, B. 2010. Simulation of Naturally Fractured Reservoirs. State of the Art-Part 1–Physical
Mechanisms and Simulator Formulation. Oil & Gas Science and Technology–Revue de l’Institut Français du
Pétrole, 65(2): 239-262.
SPE 169152 15

Lim, K.T., and Aziz, K. 1995. Matrix-Fracture Transfer Shape Factors for Dual-Porosity Simulators. Journal of Petroleum
Science and Engineering 13(3): 169-178.
Lu, H., Di Donato, G., and Blunt, M. 2008. General Transfer Functions for Multiphase Flow in Fractured Reservoirs. SPE
Journal 13(3): 289-297.
Mason, E.A., and Malinauskas, A.P. 1983. Gas Transport in Porous Media: The Dusty-Gas Model. Elsevier, Amsterdam.
Perkins, T.K., and Johnston, O.C. 1963. A Review of Diffusion and Dispersion in Porous Media. SPE Journal 3: 70-84.
Quandalle, P., and Sabathier, J.C. 1989. Typical Features of a Multipurpose Reservoir Simulator. SPE Reservoir
Engineering 4(4): 475-480.
Rezaveisi, M., Ayatollahi, S., and Rostami, B. 2012. Experimental Investigation of Matrix Wettability Effects on Water
Imbibition in Fractured Artificial Porous Media. Journal of Petroleum Science and Engineering 86: 165-171.
Shojaei, H., Rastegar, R., and Jessen, K. 2012. Mixing and Mass Transfer in Multicontact Miscible Displacements. Transport
in Porous Media 94(3): 837-857.
Sigmund, P. 1976. Prediction of Molecular Diffusion at Reservoir Conditions. Part 1-Measurement and Prediction of Binary
Dense Gas Diffusion Coefficients. Journal of Canadian Petroleum Technology, 15(2): 48-57.\
Sonier, F., Souillard, P., and Blaskovich, F.T. 1988. Numerical Simulation of Naturally Fractured Reservoirs. SPE Reservoir
Engineering 3(4): 1114-1122.
Taylor, R., and Krishna, R. 1993. Multicomponent Mass Transfer. John Wiley & Sons, New York.
Trangenstein, J.A., and Bell, J.B. 1989. Mathematical Structure of Compositional Reservoir Simulation. SIAM Journal on
Scientific and Statistical Computing 10(5): 817-845.
Vega, B., O’Brien, W.J., Kovscek, A.R. 2010. Experimental Investigation of Oil Recovery from Siliceous Shale by Miscible
CO2 Injection. Paper SPE 135627 presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19-
22 September.
Warren, J.E., and Root, P.J. 1963. The Behavior of Naturally Fractured Reservoirs. SPE Journal 3(3): 245-255.
Wesselingh, J.A., and Krishna, R. 1990. Mass Transfer. Ellis Horwood, Chichester, England.
Zimmerman, R.W., Chen, G., Hadgu, T., and Bodvarsson, G.S. 1993. A Numerical Dual-Porosity Model with Semianalytical
Treatment of Fracture/Matrix Flow. Water resources research 29(7): 2127-2137.

Appendix A: Shape Factors for Gas-Oil Diffusion


Consider a matrix block that is surrounded by fractures (Fig. A-1). Assume matrix and fractures are filled with different
phases (e.g. oil in the matrix and gas in the fractures); and there is gas-oil equilibrium at the fracture-matrix interface at all six
faces of the matrix block. Also assume the fluids in all fractures surrounding the matrix block have the same pressure and
composition. The transfer rate of each component due to molecular diffusion from the center of matrix block to the matrix-
fracture interface at each face is given by

Figure A-1: Schematic of a matrix block surrounded by fractures

(A-1)
16 SPE 169152

The total transfer rate of each component from the matrix block to the matrix-fracture interface is obtained by summing the
transfer rates in Eq. (A-1)

(A-2)

Dividing both sides of Eq. (A2) by will give us the transfer rate per unit volume of matrix block

(A-3)

where σ1 is given by

(A-4)

The transfer rate of each component due to molecular diffusion from the center of each fracture to the matrix-fracture
interface is given by

(A-5)

The total transfer rate of each component from the fracture to the matrix-fracture interface is obtained by summing the
transfer rates in Eq. (A-5)

(A-6)

Dividing both sides of Eq. (A-6) by will give us the transfer rate per unit volume of matrix block

(A-7)

where σ2 is given by

(A-8)
SPE 169152 17

Based on the film theory, the transfer rates given by Eqs. (A-3) and (A-7) must be equal. This leads to Eq. (27). Since the
molecular diffusion equation is analogous to the pressure diffusion equation, one can replace 4 with π2 in Eqs. (A-4) and (A-
8) and multiply the transfer function by a boost factor based on the derivations of Zimmerman et al. (1993) and Lim and Aziz
(1995).

Appendix B: Transfer Rate due to Gravity Drainage


The transfer rate of a component i from matrix to fracture due to gravity drainage is given by

(B-1)

where Tjk represents the displacement of phase j by phase k due to gravity drainage which acts only in the vertical direction.
Since we are looking at two-phase gas-oil systems, we need to calculate Tgo and Tog. Lu et al. (2008) proposed the following
equations based on the observations of Di Donato et al. (2006)
(B-2)
where Bgd is the correction factor for the gravity drainage term; Sgm,init is the initial gas saturation in the matrix block; S*gm is
the final gas saturation in the matrix at the end of displacement; b is the exponent of the oil relative permeability function;
and βgo and F(Sgf) are parameters that are defined as follows
(B-3)
where λgm, λom and λtm are the gas, oil and total mobilities in the matrix respectively; ∆ρmgo is the difference between mass
densities of gas and oil; g is the acceleration of gravity; and h denotes the hight of the matrix block. Since the gas is much
more mobile than the oil, λgmλom/λtm can be approximated by where is the maximum oil relative permeability.
The parameter F(Sgf) is given by

(B-4)

where Sgf is the gas saturation in the fracture; and Kf is the fracture permeability. Since the gravity drainage acts in the vertical
direction, vertical permeability is used in Eqs. (B-3) and (B-4).
The correction factor for the gravity drainage term is given by
(B-5)
We note that the correction factor for gravity drainage has a value less than unity and will approach unity at late times. We
also note that Tog = ─Tgo for incompressible flow settings (Lu et al., 2008).

You might also like