You are on page 1of 11

Modeling of Geomechanics in Naturally

Fractured Reservoirs
M. Bagheri,* SPE, and A. Settari, SPE, University of Calgary

Summary porosity poroelastic theory of Biot (1941, 1955). In the literature,


Conventional modeling of fractured reservoirs treats fracture- different approaches have been proposed to extend Biot’s single-
system permeability and porosity as static (or pressure-dependent) porosity theory to dual-porosity models.
data. Recent attempts at coupling geomechanics focused on the Valliappan and Khalili-Naghadeh (1990) and Khalili-
permeability but used crude empirical relations and treated the Naghadeh and Valliappan (1991) accounted in their coupled dual-
fluid flow as single porosity. This study takes advantage of the porosity formulations for the effect of rock deformation on the
joint-mechanics theory to develop general, rigorous coupling be- pressure of both media. In these formulations, various coefficients
tween the fluid-flow equation and deformation of fractured media. are involved and defined in terms of measurable physical parameters.
Both porosity and permeability coupling are considered. Ghafouri and Lewis (1996) developed a formulation for de-
The geomechanical part uses the equivalent-continuum ap- formable porous media. In this formulation, the compressibility of
proach, considering both rock- and fracture-deformation proper- fractures is assumed not to alter the compressibility of the whole
ties. Multiple sets of fractures with any dip and strike angle can be system, and the effect of fracture pressure on total deformation
defined. The stiffness of fractures varies with the effective stress was ignored.
according to a law typical for joints. Chen and Teufel (1997) proposed a new formulation out of
The main novelty of this work is that the geomechanics solu- Biot’s theory of poroelastisity for coupling geomechanics and fluid
tion is decomposed into matrix and fracture parts and used to flow in deformable dual media. They added a term to account for
compute their dynamic porosity and permeability separately. This the effect of the pressure of the secondary porosity on volumetric
approach rigorously captures the effect of fractured-media defor- strain, bulk volume, and total pore volume. They derived the
mation on the dual-porosity-flow part of the coupled system and changes of individual-fracture and matrix pore volumes in terms of
allows the permeability and porosity variations to be based on total stress and the pressure of the individual medium. Their final
measurable joint properties. Generally, fracture deformations pro- governing equations were similar to those of Valliappan and Kha-
duce changes of the permeability tensor in both magnitude and lili-Naghadeh (1990). The main difference is the way that the
orientation, which in turn influences reservoir flow and compac- coefficients are defined.
tion behavior. Other models addressing the coupling are also found in litera-
The main issue studied was the variation in the permeability of ture. Chen and Teufel (2000) presented a comparison between
the fracture system. The examples show that fracture deformation some of these models.
has a significant effect on productivity or injectivity and that an- Most—if not all—of these models are based on a fully coupled
isotropy of the permeability tensor develops from deformation. approach, which is only one of the coupling strategies. Other
The results provide an initiative for implementing the case of schemes of coupling geomechanics and reservoir engineering are
full-tensor permeability. also available, and their advantages and disadvantages have been
discussed in the literature. One of the disadvantages of the fully
Introduction coupled approach served as the incentive to develop this research
into an iterative-coupling scheme for a naturally fractured reser-
Similar to other petroleum reservoirs, naturally fractured reservoirs voir. The main objective was to address the effect of geomechanics
can be greatly influenced by the geomechanical behavior of rocks. on existing fracture-flow properties (i.e., porosity and permeabil-
However, under similar conditions, the role of geomechanics is ity). Other effects—such as creation of new fractures by matrix
even more crucial because of the presence of fractures, which may failure, fracture slip, and fault reactivation—are not treated here.
be more stress sensitive than the rock matrix. These fractures are In this model, fracture-flow properties are calculated internally,
affected by stress disturbances because of fluid production and/or provided that the initial distribution of fractures and their mechani-
injection, which results in the opening and closure and the reori- cal properties are known. Depending on the distribution of frac-
entation of fractures. These variations in geomechanical properties tures, a full permeability tensor may result. However, in the ex-
of fractures affect their permeability (both magnitude and direc- amples of this paper, fractures are assumed to be parallel to the
tion), which is a controlling factor in the management of naturally coordinate axes, which results in a diagonal permeability tensor.
fractured reservoirs. In the case of oblique fractures, the geomechanical model de-
To capture this behavior, it is inevitable to consider geome- scribed here must be coupled with a flow model capable of han-
chanical factors in the modeling of fluid flow in naturally fractured dling full permeability tensor. In this case, changes in fracture
reservoirs. Acknowledging a few attempts at coupling fluid-flow permeability will result in a change in the magnitude as well as the
behavior in naturally fractured reservoirs, dual-porosity models direction of the principal permeabilities. This rigorous treatment
used in the industry fail to account for deformability of rock and has also been developed, and its application will be the subject of
fractures. These models use simple pressure-dependent relations for a future study.
rock compressibility, while fracture permeabilities are typically
treated statically throughout the simulation of the entire reservoir life. Modeling of Fractured-Rock Deformation
The theory of coupling geomechanics and reservoir engineer- To model deformation of rock and fractures of a naturally frac-
ing in fractured rocks published in literature is built on the single- tured reservoir accurately, one can use the mechanics of deforma-
tion of single fractures and porous rock to develop the concept of
equivalent continuum in the geomechanics module of the coupled
* Now with Sproule International, Calgary.
simulator. While the dual-porosity nature of fluid flow is retained
in the flow module, the overall deformation of a dual media (frac-
Copyright 2008 Society of Petroleum Engineers
tured reservoir) is calculated with an equivalent single medium.
This paper (SPE 93083) was accepted for presentation at the 2005 SPE Reservoir Simu- Individual deformation of rock and fractures is then calculated by
lation Symposium, The Woodlands, Texas, 31 January–2 February, and revised for publi-
cation. Original manuscript received for review 27 July 2006. Revised manuscript received
the decomposition of the overall deformation, again using the de-
for review 14 December 2006. Paper peer approved 18 March 2007. formation mechanics of fractures and rocks individually.

108 February 2008 SPE Reservoir Evaluation & Engineering


Interaction of the dual media within the flow module with the Equivalent-Continuum Concept. The concept of an equivalent
equivalent single medium in the geomechanics module will be medium of a jointed rock mass (composed of two materials) is to
explained in the next section. Here the focus is to briefly explain introduce a third material that behaves the same way as the jointed
the deformation mechanics of fractures and the concept of the rock mass does, under normal and shear stresses. The moduli of
equivalent continuum. this third material are functions of the moduli of the fractures and
intact rock. Normal and shear stiffness, along with joint orientation
Deformation Mechanics of a Single Fracture. A single fracture and spacing, are key parameters for constructing the constitutive
(also called a joint) deforms under stress differently from an intact equation of the equivalent material. Constitutive relation of frac-
rock. Also, its deformation under normal stress is different from tures, which is described by these parameters, is very similar to
when it is loaded by shear stress. Fracture behavior under shear that of intact rock. For fractures, however, displacements of frac-
stress is more complex; the fracture undergoes several stages of ture planes are the variables rather than strains.
deformation and requires a complex mathematical relation to be Constitutive relation for intact rock obeys Hooke’s law and
modeled (Bandis et al. 1983a). Models of shear behavior involve mathematically has the following form in index notation:
many parameters that require extensive laboratory work, and dif- ␧ij = Cijkl␴kl, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
ficulties exist in translating the laboratory work into the simulation
of fracture shear in underground conditions. Therefore, we limited where ␧ is the strain vector (1×6), ␴ is the stress vector (1×6), and
this study to considering only the normal deformation of fractures, C is the stiffness matrix (6×6). For saturated porous rocks how-
which is more important for fracture permeability and porosity. ever, ␴ in Eq. 7 is replaced by effective stress ␴⬘ defined as:
However, shear-related parameters have been retained in the for- ␴⬘ = ␴ − ␣p, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
mulation; and instead of modeling the shear, values were assigned
where ␣ is Biot’s coefficient and p is the pore pressure.
to these parameters to disable the shear deformation of fractures.
A similar expression governs the constitutive relation between
Strong nonlinear behavior of fracture deformation under nor-
fracture deformation and traction:
mal stress (resembling a hyperbola) has been acknowledged uni-
versally in literature (Bandis et al. 1983b; Youshinaka and Yamabe ⌬␦I = DIJ ⌬⌫J, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)
1986; Goodman 1976). In this work, the fracture-mechanics model where ␦ is the displacement vector (1×3), D is the fracture stiffness
of Bandis et al. (1983b) is implemented in the geomechanics mod- matrix (3×3), and ⌫ is the traction vector (1×3). The traction vector
ule. Their model is based on laboratory investigation of inter- is a vector that bears the direction and magnitude of the stress acting
locked-block samples with natural unfilled joints, which were sub- on the joint surface and can be expressed by Cauchy’s formula:
jected to a series of loading/unloading cycles. A large hysteretic
behavior was observed in the first load cycle, and it diminished on ⌬⌫j = ⌬␴ijni, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (10)
subsequent cycles. where ni represents the components of the vector normal to the
It can be argued that, under highly stressed in-situ conditions, joint surface. Note that the lowercase subscripts in this discussion
fracture has already experienced loading cycles and the hysteresis indicate the global coordinate system. Each joint set has its own
effect is not dominant. A single stress/strain function is then ap- local coordinate system, which is indicated by capital subscripts.
plied for the deformation of a fracture in loading and unloading, Often it is assumed that deformation characteristics of the joint
which is assumed to best represent the fracture-deformation path. in different directions at the joint surface are the same. Neglecting
However, the theory of the equivalent media developed here will the effect of dilatancy behavior (shear-induced volume contraction
apply also to hysteretic behavior, if formulated incrementally. or dilation) and shear strain because of normal stress, Eq. 9 can be
Bandis et al. (1983b) proposed the following equation for a expanded in a simple form:
hyperbolic path of fracture normal deformation:

冤 冥
1
⌬v 0 0
kn
␴n = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1) ⌬␦N ⌬⌫N

冦 冧 冦 冧
a − b⌬v 1
where ␴n is the normal stress, ⌬v is the change in fracture aperture, ⌬␦S = 0 0 ⌬⌫S . . . . . . . . . . . . . . . . . . . . . . . (11)
ks
and a and b are the constants that have physical meaning. For value ⌬␦T ⌬⌫N
of ␴n→⬁, the closure reaches its maximum: 0
1
0
ks
a
⌬vmax = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2) Capital indices represent the local coordinate system of the
b
fracture, which consists of normal, strike, and dip directions (de-
Joint normal stiffness, which is the ratio of change of normal noted by N, S, and T ). This equation will be used to find the
stress to normal closure, is defined as constitutive relation of jointed rocks next.
d␴n For fluid-saturated fractures, ⌫ in Eq. 9 is replaced by effective
kn = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3) stress traction vector ⌫⬘, defined as
d⌬v
⌫⬘ = ⌫ − p. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (12)
The physical meaning of a is found from the derivative of
This form of effective stress is different than that of the matrix
Eq. 1.
shown in Eq. 8, which includes Biot’s coefficient ␣. Difference in
⭸␴n 1 structure of porous materials and joints is the reason for recom-

冉 冊
kn = = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)
⭸⌬v b 2 mending this simple form of effective stress.
a 1 − ⌬v Huang et al. (1995) developed the constitutive relation for a dry
a rock mass represented by an equivalent continuum, using a com-
At zero normal stress, normal stiffness is called initial normal bination of both constitutive relations of matrix and fractures and
stiffness and is shown to be based on the principle of energy conservation. For saturated rock


masses, this approach is valid for cases where pressure difference
⭸␴n 1
kni = = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5) in fractures and matrix (with high-grain compressibility) is small
⭸⌬v ␴n=0 a (Bagheri 2006). The following paragraphs explain this method for
Substituting Eqs. 2 and 5 into Eq. 4 will result in saturated fractured rocks.
Change in the total strain of the continuum because of effec-
kni tive-stress changes consists of two components: one from the in-

冉 冊
kn = . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
⌬v 2 tact rock and the other because of fractures:
1−
⌬vmax ⌬␧ij = ⌬␧ijI + ⌬␧ijJ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (13)

February 2008 SPE Reservoir Evaluation & Engineering 109


Eq. 13 can be rewritten for both media to show the relation of each needed. However, it is possible to study the behavior of rock itself
component in response to the effective-stress change. This is rig- with a single-porosity coupling as discussed below.
orous if we assume a small difference between matrix and fracture
pressures, small matrix capillary pressures (in multiphase-flow Single-Porosity Approach. To study the deformation of fractured
cases), and high grain compressibilities (i.e., ␣ close to 1): reservoirs and their impact on the environment, for example for a
subsidence problem, a single-porosity flow model can do the job.
⌬␧ijI ≅ C ijkl
I
⌬␴ ⬘kl, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (14) The objective here is to see how and to what degree fractures are
⌬␧ijJ ≅ C ijkl
J
⌬␴ ⬘kl, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (15) participating in the overall deformation of fractured rock. This
requires a realistic “continuum pore pressure” for the equivalent
where ␴⬘ is the effective stress in the continuum. For the con- continuum. In other words, regardless of the nature of fluid flow,
tinuum, the definition of effective stress of the matrix is applied any flow simulators that provide a close estimate of reservoir
because fractures can be considered as big pores in the continuum. pressure are useful for this purpose.
Because the fractured gridblocks are treated as continuum in To get a realistic dual-porosity reservoir pressure out of a
the geomechanical module, an average “continuum pore pressure” single-porosity model, one approach could be converting the com-
must be defined, which should have the same affect on deforma- pressibility of fractures to some measure of porosity change of
tion as fracture pressure and matrix pressure do. The continuum equivalent continuum. The pseudoporosity of the resulting con-
pore pressure is calculated on the basis of the following equation: tinuum single-porosity rock, which has a higher compressibility, is
␾m ␾f used for the single-porosity model to provide a more realistic
p= p + p , . . . . . . . . . . . . . . . . . . . . . . . . . . (16) pressure to the geomechanics module. This is accomplished by the
␾m + ␾f m ␾m + ␾f f
calculation of porosity out of overall strain that resulted from
where ␾m and ␾f are matrix and fracture porosities, respectively. deformation of rock and fractures. As a result, compressibility of
For multiphase problems, matrix and fracture pressures can be fractures increases the porosity changes of the pseudorock. A de-
taken as some average of the individual-phase pressures. The ef- tailed description of this approach and results is presented in
fective stresses of the pseudocontinuum are calculated by Eq. 8, in Bagheri and Settari (2006).
which p is given by Eq. 16 and Biot’s coefficient is that of the
matrix. The continuum pore pressure as defined by Eq. 16 is an Dual-Porosity Approach. The inclusion of effects of fracture de-
approximation, and further research should be conducted to inves- formation in stress/strain behavior for fractured rocks provides the
tigate if there is a more representative equation. capability of rigorous modeling of naturally fractured rocks. In this
On the basis of the principle of energy conservation, strain approach, all fractures in the reservoir are represented by several
energy stored in fractures because of stress is equal to the total fracture sets. Each fracture set consists of equally spaced fractures
work done by the traction at all joints that caused joint move- of the same orientation and properties (such as aperture and stiff-
ment. Thus, ness). This concept parallels the geological characterization of
M fractured reservoirs and allows the parameters of the fracture sets
兺 共␴ ⬘ n 兲⌬␦ A , . . . . . . . . . . . . . . . . . . . . . . . . . . (17)
1
␴ ⬘ij⌬␧ijJ = ij
f
i
f
j
f to vary spatially. The flow equations are solved for rock and frac-
V f=1 ture domains in a 3D multiphase, dual-porosity flow simulator. On
where M is the total number of fracture sets, f indicates individual the geomechanics side, the concept of the previously described
fracture sets, and Af is the total area of the f th fracture set, which equivalent continuum simplifies the stress/strain equations to a set
is approximately equal to the volume of the representative block of equations for a single continuum. The solution of these equa-
divided by the average spacing between the fractures in the set. tions leads to an overall stress and strain vector for the continuum
Thus, the strain of fractured rock because of the movement of the media; however, the corresponding deformation of each medium
fractures is given by can be recovered. An iterative coupling scheme is then used to
couple the interaction between fluid and rock.
M
Several parameters including matrix porosity, matrix perme-
兺 n ⌬␦ S ,
1
⌬␧ijJ = f
i
f
j f
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (18) ability, fracture porosity, and fracture permeability can be coupled
f=1
between the flow and geomechanics parts of the model. The main
where S is the spacing between the fractures. focus of this research is to study the effect of fracture-permeability
Effective traction and displacement vectors can be transformed variations on fluid flow in naturally fractured reservoirs. Coupling
between the global and local coordinate systems as follows: of matrix porosity between stress/strain and flow models is per-
formed with the Settari and Mourits (1998) approach, which has
⌬␦ fj = LjJf ⌬␦ fJ, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (19) been derived for conventional reservoirs. In this new model for
⌬⌫⬘I = LfIi⌬⌫⬘i , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (20) naturally fractured reservoirs, however, matrix porosity is related
to the strains associated with the matrix part of the fractured rock.
where L contains the direction cosines between local and global Decomposition of the overall strains to rock and fractures is dis-
coordinate systems. Note that for the fractures in the continuum, cussed later.
the effective stress is obtained by substitution of continuum effec- Model Initialization. Fig. 1 shows the fracture-related calcu-
tive stress into Cauchy’s formula. So the effective stresses in Eq. lations in the initialization part of the model. Bulk volume, initial
20 are somewhat different from those shown in Eq. 12. Substitu- fracture, and matrix porosities along with initial fracture and ma-
tion of the counterpart of Eq. 9 for saturated joints and Eqs. 19, 20, trix pressures of each gridblock are transferred from the flow mod-
and Cauchy’s formula into Eq. 18 gives the general constitutive ule to the geomechanics module to provide initial information for
equation for joints: calculation of continuum pore pressure, fracture aperture, and frac-
M ture porosity.
兺nL D
1
J
C ijkl = f f
i jJ
f f f
JLLLl n k . . . . . . . . . . . . . . . . . . . . . . . . . . . (21) Initial aperture of each fracture set in a gridblock is calculated
f=1 Sf using the Bandis et al. (1983b) technique. As mentioned earlier,
This equation must be added to the constitutive equation for the hysteresi in fracture deformation is neglected in this study. All
matrix to obtain the general-constitutive relation for the equivalent fractures in a given fracture set have the same orientation and
rock continuum. properties within a gridblock because the same normal stress and
fracture pressure are exerted on all of them. Effective stress ex-
Coupling Strategies erted on the planes of a fracture is the net effect of gridblock
The geomechanical module, which is developed to model defor- fracture pressure and normal stress on the fracture plane.
mation of fractured rocks, can be coupled to a single-porosity or a Once initial aperture was calculated for all fracture sets in a
dual-porosity flow model. To study the behavior of fluid flow in a gridblock, the total fracture pore volume is calculated. This is the
deformable naturally fractured reservoir, a dual-porosity model is summation of pore volumes for all fracture sets, each of which in

110 February 2008 SPE Reservoir Evaluation & Engineering


Fig. 1—Fracture-related calculations in the initialization phase.

turn is the product of aperture and the total fracture area of the
fracture set. The exact calculation of the porosity is nontrivial.
However, regardless of fracture orientation, total fracture area Af
can be approximated by:
Lx Ly Lz
Af = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (22)
S
where S is fracture spacing of a fracture set and Lx , Ly , and Lz are
the length, width, and height of the gridblock, respectively. The
accuracy of this approximation was tested in a separate code that
calculates the exact total area of each fracture set (Bagheri 2006).
Considering the uncertainty regarding the properties and the dis-
tribution of fractures in reservoirs, it is observed that this approxi-
mation is sufficiently accurate even for low-intensity fracture sets.
The error inherent in this approximation approaches zero when
fracture intensity increases.
The final part of initialization concerns fracture-permeability
calculations. Permeability of fractures in fractured blocks is cal-
culated on the basis of the information we have about the orien-
tation and spacing of fracture sets, and the aperture of fracture sets
from the previous step. Details of the method of calculation of
fracture permeability used in this study can be found in Gupta et al. Fig. 2—Dual-media treatment in the solution phase.
(2001). To account for deviation from the cubic law (resulting
mainly from fracture roughness and tortuosity), permeability can
be multiplied by a correction factor. porosity changes, and fracture apertures are used for the calcula-
Calculated fracture porosity and fracture permeability, which tion of fracture permeability and porosity.
now represent characterization of the geological structure of the At the end of the geomechanical iteration, matrix porosity ␾m
fractured reservoir, are passed back to the flow model, where these is always updated using the new effective stress associated with
properties are used for flow calculations. the matrix, which is ␴−␣pm. However, the fracture porosity ␾f and
Dynamic Coupling. The dual-porosity iterative coupling ap- fracture and matrix permeabilities km and kf , respectively, may or
proach of this study is shown schematically in Fig 2. This figure may not be updated. Fig. 2 shows the case where the geomechani-
highlights the stages of calculation of fracture and matrix flow cal iteration involves only ␾m.
properties, and the way in which they are coupled to the geome- New fracture porosities are based on the current gridblock bulk
chanics module. For example, in the initialization stage, matrix and volume. Bulk volume is a dynamic parameter, which changes be-
fracture porosity and pressure are transferred to the geomechanics cause of applied stress by the amount of volumetric strain. Math-
module in each geomechanical iteration loop to build a new con- ematically, this is shown as:
tinuum pore pressure using Eq. 16. This pressure is treated as an
external force in the equations. BV = 共1 + ␧v兲 ⭈ BVi , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (23)
New fracture-normal stiffness and flow properties are calcu-
lated in each geomechanical iteration loop using the effective where BV, BVi, and ␧v are the current and initial bulk volume and
stress on the fracture plane, which is a function of the new fracture the total volumetric strain (compressive strains are negative), re-
pressure transferred from the flow model and total stresses. It spectively. The same adjustment of bulk volume must be per-
should be noted that the fracture flow properties are calculated formed in the flow model as well.
before the new total stress field is calculated. As a result, the total Coupling of fracture-flow properties can be treated in two
stresses used in these calculations are lagging one geomechanical ways. Like matrix porosity, fracture-flow properties can be up-
iteration behind the pore pressures. dated in every geomechanical iteration and passed back to the flow
After the calculation of continuum pore pressure, an equivalent model. The highly nonlinear nature of these properties makes the
constitutive matrix is built using the last normal stiffness of frac- convergence process slow. To alleviate convergence difficulties,
tures. This matrix is composed of static Young’s modulus and an alternative approach for coupling fracture-flow properties was
Poisson’s ratios of the rock, static fracture spacing, and dynamic found to be effective. In this method (shown in Fig. 2), fracture
normal stiffness of fractures. Using this matrix in stress/strain porosity and permeability are updated only at the end of each
calculations provides the total stress and strain of the pseudocon- timestep and used in the next timestep. In other words, these prop-
tinuum. The overall deformation is then decoupled as described erties do not partake in coupling iterations; they use the converged
below. Matrix strains are the basis for the calculation of matrix- coupling variables (such as pressure and stress).

February 2008 SPE Reservoir Evaluation & Engineering 111


Decoupling of Deformations. Individual fracture deformations The stress model is constructed of 29 gridblocks in the x-direction
and matrix strains are required to obtain corresponding flow prop- over a length of 11 km, five gridblocks in the y-direction over a
erties of either media. To obtain these parameters, the calculated width of 1.5 km and 23 gridblocks in the z-direction over a height
total stress from the solution of the finite-element model is con- of 5.2 km. The reservoir model is the steeply dipping finely grid-
verted to appropriate effective stress associated with each media. ded region contained inside the stress model shown in darker color
This is done by invoking the proper relation for effective stress and in Fig. 3. It has 19 gridblocks in the x-direction, five gridblocks in
substitution of total stress and pressure associated with each me- the y-direction, and five gridblocks in the z- direction. The reser-
dium as detailed below. voir grid is a part of the stress-model grid (i.e., has the same
For a joint set characterized by its normal vector, n៝ , strike dimensions where the grids coincide). The difference between the
vector, s៝, and dip vector, ៝t, the local components of total stress are depth of the first and last gridblocks in the x-direction for all
obtained from a simple transformation of global stress tensor: reservoir layers is 2313 m.
The initial in-situ matrix porosity is 10%, and the matrix per-
兵␴nn ␶ns ␶nt其T = Tm 兵␴xx ␴yy ␴zz ␶xy ␶yz ␶zx其T, meabilities are 1 md in all directions. Oil and water have constant
. . . . . . . . . . . . . . . . . . . . . . . . . . (24) compressibilities of 1.45×10−6 kPa−1 and 6×10−7 kPa−1, densities
where Tm is the transformation matrix and is equal to of 752.7 kg/m3 and 999.5 kg/m3, and constant viscosities of 2 cp
and 0.1626 cp, respectively. The initial fluid pressure is 112 812.6

冤 冥
n2x n2y n2z 2nxny 2nynz 2nznx kPa at the depth of 7262.5 m.
Different elastic moduli and Poisson’s ratios were assigned to
nxsx nysy nzsz nysx + nxsy nzsy + nysz nzsx + nxsz different parts of the model. The model was divided into seven
nxtx nyty nztz nytx + nxty nzty + nytz nztx + nxtz regions with different rock properties, as shown in Table 1. All of
the regions have the same grain modulus of 2.07×109 kPa. Initial
The entries of the transformation matrix Tm are the components horizontal stresses are 109 137 kPa, and initial vertical stress is
of normal, strike, and dip vectors of the joint set in the global 110 240 kPa at a depth of 4876 m. The gradients of horizontal and
coordinate system. The normal deformation of fractures is then vertical stresses are 22.378 kPa/m and 22.604 kPa/m, respectively.
obtained using Eq. 9 where ⌬⌫J is substituted for the fracture Three identical fracture sets, parallel to the coordinate axes are
normal effective stress, ⌫⬘N⳱␴nn−pf (the shear terms are ignored). assumed in all reservoir gridblocks. They have constant spacing of
This effective stress is also used to update the normal stiffness of 2 m, initial stiffness of 39.62 GPa/m, joint aperture at zero effec-
fractures (for the next geomechanical iteration) in every gridblock. tive stress of 0.44 mm, and minimum aperture of 0.022 mm (maxi-
This approach satisfies Eq. 13 and does not require explicit cal- mum of 95% closure). These parameters are used in calculation of
culation of fracture strains. equivalent moduli of fractured rock according to Eq. 21 and to
Matrix effective stress is obtained by the substitution of pm and calculate shape factor in the transfer function in the flow model
the total stress in Eq. 8. In the next step, matrix strains are obtained (for example, see Kazemi et al. 1976).
by substituting the static-rock-constitutive-matrix moduli and the Differences in effective stresses normal to these fractures in
calculated effective stress in Hooke’s law. It was shown by ex- different locations result in a distribution of fracture porosity and
amples for simple cases—the sum of the two deformations is equal permeability with depth, decreasing from the top to the bottom of
to the total deformation of the pseudocontinuum (Bagheri 2006). the reservoir. For instance, fracture permeability covers a range
from 88.17 md (kz) to 16.77 md (kx, ky) at the first layer. The
Results and Discussion minimum permeability in the entire reservoir is the horizontal
To demonstrate the importance of modeling fracture permeability permeability of 15.43 md at the bottom of the deepest layer, where
as a dynamic parameter, two independent dual-porosity tests are effective stress is the highest. The distribution is shown as a func-
explained here. In both examples, there are three sets of fractures, tion of reservoir block number from left to right in Fig. 4.
each parallel to one coordinate axis. In the first example, these Depletion Scenario. For the depletion scenario, a vertical well
fracture sets have the same deformation characteristics with the is located in the cell (18, 3, 2–5) at the top of reservoir, producing
same fracture spacing. In the second example, fractures deform 8000 m3/d of oil at surface conditions until the reservoir pressure
differently. is depleted to 20 000 kPa. The well is completed through the
bottom four layers of the reservoir. The model is run for 20 years.
Example 1. This example is loosely modeled after the Shearwater Three gridblocks—(1, 3, 1) at the bottom, (9, 3, 1) at the middle,
field (Kenter et al. 1998). The data, presented at the SPE Forum on and (18, 3, 1) at the top of the reservoir—are chosen to monitor the
Reservior Geomechanics (2000), are a good example of an over- behavior of the reservoir.
pressured, high-pressure/high-temperature field with large deple- To study only the effect of fracture closure on permeability,
tion. The actual field is not fractured, but the data were modified fracture porosity was decoupled from the geomechanical code and
to represent a hypothetical dual-porosity dead-oil reservoir. The its variation was calculated by use of the ordinary compressibility
structure of the reservoir and the stress model are shown in Fig. 3. method. Then the results of this model with dynamic permeability
were compared with the results of the same model using a static
fracture permeability of the same initial distribution. These cases
are referred to as “dynamic” and “static” throughout this paper.

Fig. 3—Structure of stress and the reservoir model (Example 1).

112 February 2008 SPE Reservoir Evaluation & Engineering


Fig. 4—Initial Distribution of fracture permeability.
Fig. 5—Effect of fracture closure on production.

Fig. 5 compares cumulative oil production and oil rate in the


Calculation of WI in dynamic-permeability modeling is a com-
depletion case. The dynamic treatment results in a drastic reduc-
plex issue. The difference between BHP-D and GBP-D, shown in
tion in cumulative oil production. Rate of oil production also de-
Fig. 6, reveals the large pressure drop inside the producing grid-
creases more rapidly than in the case of static-fracture permeabil-
block. As a result, finding the proper permeability for the WI
ity. The closure of fractures resulting from an increase in effective
equation is difficult because of variations of pressure (effective
stress (production) reduces overall fracture permeability. The
stress) inside the wellblock.
amount of permeability reduction is higher in the vicinity of the
In this study, the wellblock permeability is used for WI calcu-
production wells. As a result, the producing well experiences a
lations. The geometric part of WI in this approach includes the
higher pressure drop as seen in Fig 6. In very early times, when
variation of the permeability of the gridblock and results in a dynamic
static and dynamic cases produce with the same rate of 8000 m3/d,
geometric part of WI, which is a constant value in the static case.
higher reduction in gridblock pressure of the producing well in the
In another words, during the simulation, new permeabilities are
dynamic case (GBP-D) is seen because of lower permeability re-
used to calculate WI from Eq. 25. This results in smaller well
sulting from fracture closure. In this period, when the well pro-
indices with more production as shown in Fig. 7. The bigger gap
duces on rate control, a large pressure drop in bottomhole pressure
between BHP and GBP in the dynamic case in all times causes this
(BHP) happens in the dynamic case, causing the well to switch pro-
result. Note that the effect of dynamic permeability was ignored in
duction constraint from rate to BHP. This results in early rate drop in
Rb calculation in this study, as it has a negligible effect on the WI.
the dynamic case (case Oil rate−D in Fig. 5). Thereafter, the rate of
Using a more sophisticated method incorporating the variations
pressure drop in the well gridblock in the dynamic case reduces.
of effective stress inside the wellblock in lieu of the current ap-
For rectangular gridblocks and anisotropic permeabilities, the
proach (based on the effective stress using well block pressure)
well index (WI) is:
causes the permeability to reduce more than on the plot shown in
Fig. 8, when we move toward the wellbore. More rigorous treat-
2␲ 公kxkyh ment of well index is left for future studies. This treatment will

冉 冊
WI = , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (25)
Rb demonstrate higher influence of dynamic permeability modeling in
ln +s reservoir performance.
Rw
As initial oil in place is the same in both cases, ultimate re-
where h is the thickness perpendicular to the flow direction, kx and covery would also be the same if enough time is devoted to pro-
ky are the permeabilities, Rw is the well radius, s is skin factor, and duction. However, the production rate is lower in the dynamic
Rb is the effective well radius and is given case. If production is continued, the rate of the pressure decline
will decrease more in the dynamic case, as a result of lower rate of
production. Eventually, the plots of production rate for the static
关共ky Ⲑ kx兲1 Ⲑ 2⌬x2 + 共kx Ⲑ ky兲1 Ⲑ 2⌬y2兴1 Ⲑ 2
Rb = 0.28 × . . . . . . . . . . . . (26)
关共ky Ⲑ kx兲1 Ⲑ 4 + 共kx Ⲑ ky兲1 Ⲑ 4兴

Fig. 7—Time (effective-stress) dependency of geometric part of


Fig. 6—Effect of fracture closure on producer’s pressure. WI of the producer.

February 2008 SPE Reservoir Evaluation & Engineering 113


Fig. 8—Effect of fracture closure on fracture permeability.
Fig. 9—Fracture permeability as a function of effective stress.

and dynamic cases cross. This is the point of maximum difference


between cumulative oil productions, which is approximately 33.9 (Eq. 16) was used for effective-stress calculations because of a
or 51.3%, depending on the chosen reference. Thereafter, rate of small difference between them. (Obviously, internal calculations
oil production in the dynamic case is higher because of dominating in the flow model are performed using pf and pm.) This plot is
higher pressures around the low-permeability well. important because it demonstrates that the fracture permeability is
The drastic change in normal fracture permeability (kz) with indeed a function of effective stress (rather than pressure).
time resulting from this approach is shown in Fig. 8. In the dy- Injection Scenario. All of these concepts also rule the behavior
namic case, permeability reduces up to four times of the initial of a reservoir in injection situations. Fractures around the injector
(static) values. At different locations in the reservoir, the amounts open as effective stress is reduced because of increased pressure.
of permeability reduction are different. At the top of the reservoir, To demonstrate this, water was injected in the model at the low-
permeability of fractures reduces the most, while permeability of permeability bottom part of the reservoir under constant pressure
those at the bottom reduces the least. In addition, there are delays constraint of 150 000 kPa while the production well was shutin.
in permeability changes of the middle and bottom blocks, which The injection well is located in gridblock (1, 3, 2–5).
correspond to the time necessary to propagate the changes in pres- Injection rates and cumulative water injection along with well
sure and, therefore, stress. Evidently, this delay is longest for the pressures are shown in Figs. 10 and 11. Dynamic modeling of
bottom gridblock. fracture permeability allows higher injection rate because of lower
In the upper part of the reservoir where effective stress is lower, rate of increase in well pressure and higher injectivity as shown in
fractures are more deformable. This is the main reason that highest Fig. 12. Similar to the depletion case, the injection rate in dynamic
permeability reduction is observed at the top of the reservoir modeling will eventually cross that of the static case and cumula-
(Fig. 9). Fig. 9 shows a plot of vertical permeability vs. the mean tive-injection plots merge together. This is because higher injec-
horizontal effective stress for the same gridblocks shown in Fig. 8. tion rate eventually increases the reservoir pressure and reduces
All three curves fall on the same trend, as the deformation the injection pressure potential. Cumulative water injection calcu-
characteristics of all the fractures are the same. The reason for lated using dynamic and static methods for fracture permeability at
choosing fracture permeability in the z-direction is that it is a the crossing point is different by approximately 29.6 or 42% if
function of the two horizontal stresses, which vary almost identi- dynamic or static cases are chosen as the reference, respectively.
cally throughout the depletion because of the chosen boundary Vertical fracture permeability (kz) as a function of time and
conditions for deformation. This makes a plot of permeability vs. effective stress is shown in Figs. 13 and 14 for the three monitored
mean effective stress correct. The situation is more complex for the gridblocks. Highest and lowest permeability changes in this case
horizontal permeability, which is a function of both vertical and are also seen in the top and bottom gridblocks, respectively. Simi-
horizontal stress. Also, an average pressure of matrix and fractures lar to the depletion case, a delay is observed in the permeability

Fig. 10—Effect of fracture opening on injection. Fig. 11—Effect of fracture opening on the injector’s pressure.

114 February 2008 SPE Reservoir Evaluation & Engineering


Fig. 12—Time (effective-stress) dependency of geometric part Fig. 13—Effect of fracture opening on fracture permeability.
of WI of the injector.

response for the top and the middle of the reservoir because of the reservoir is horizontal and single layer). The only parameters that
time required for these blocks to see the pressure front. Changes in have been changed in the flow model are the permeability and
fracture permeability are larger than in the depletion case, for the porosity of fractures. Fracture permeability is reduced from its
same amount of effective-stress changes, as seen by comparing original value of 10,000 md to 161 md, and fracture porosity is
Figs. 9 and 14. Again, all gridblocks form a complete fracture- reduced from 1 to 0.14%. Also, shape factor was calculated on the
deformation path, showing the fundamental dependency on effec- basis of the fracture spacings shown in Table 2, and matrix per-
tive stress. meability was multiplied by 0.1. We have extended the test to a
At high injection rates, it is possible to generate negative (ten- coupled model using the stress model shown in Fig. 16. The stress
sile) effective stresses. Although it is clear from Fig.14 that hori- model has seven layers of gridblocks, which are under compres-
zontal effective stresses in our example remain compressive, this sive stresses. Initial homogeneous stress of 27 200 kPa, at the
situation is common in injection processes. Fracture-deformation reference depth of 1196.6 m with a gradient of 22.6 kPa/m, is used
equations used here are developed for positive effective stresses. If to initialize all the stresses.
one attempts to use them for negative effective stresses, caution is Three sets of fractures with different deformation characteris-
required because fracture deformation is very sensitive to change tics are used to populate all reservoir grids. Under normal stress,
in stress and this makes the problem highly nonlinear. However, in these fracture sets deform according to Fig. 17. Table 2 shows the
this case the joint theory is likely not valid and induced (propa- deformation data of these fracture sets, which were chosen such
gating) fractures will form, starting from existing natural fractures. that the fracture permeability at initial stress is isotropic. Fractures
Modeling of such problems that combine fractured media and can close to a maximum of 90% of their aperture at zero effective
induced fractures is the subject of future work. stress.
Dynamic modeling of fracture permeability also improves the An injection well that injects water at the rate of 31.8 m3/d is
run times. Fig. 15 shows the run times for all four cases explained located at gridblock (1, 1, 4) of the stress model and a production
above. well at the opposite corner of the reservoir produces oil at the rate
of 33.4 m3/d, until the reservoir pressure is depleted to 3447.4 kPa.
Example 2. In the previous example, we focused on the effect of Pressure distribution inside the reservoir will be shown for early
dynamic fracture permeability on well productivity or injectivity. times when average reservoir pressures are almost the same for the
This example demonstrates its effect on pressure distribution and static and dynamic cases because of approximately equal amounts
creation of anisotropy because of differences in properties of the of fluid injections and withdrawals. The difference is caused by
fracture sets. The data are based on the classical example of Ka- slight differences in water production (water cut).
zemi et al. (1976). The effects of gravity are not present (the Fig. 18 depicts the pressure distribution at three different times
along the diagonal that connects the injector and producer. At early

Fig. 15—Effect of static and dynamic modeling of fracture per-


Fig. 14—Fracture permeability as a function of effective stress. meability on run times.

February 2008 SPE Reservoir Evaluation & Engineering 115


times, pressures fall in the same trend, except around the producer, Fig. 21 depicts the similar behavior on the two other edges of
where reduction of permeability causes a sharper pressure gradient the reservoir intersecting at the producer gridblock after 450 days.
in the dynamic case. Despite the effect of injection on fracture In this case, pressure in the x-edge is lower than that in y-edge,
opening, difference in production and injection rates causes a re- admitting the same concept again. Evidently, the same pressures in
duction in overall reservoir pressure, which results in fracture clo- the other two corners of reservoir where wells are not located can
sure around the injector (except for the times less than 1 day). This be read either from Fig. 20 or from Fig. 21.
fact can be inferred from the increasingly higher pressure gradient Although the significance of dynamic fracture permeability is
around the injector at later times. It is worth noting that in a reverse primarily in the improvement of the reliability of the simulation
condition (pressurizing), it is expected that the pressure gradients results, lowering the run times is an added value. The improvement
would reduce in the dynamic case. in the run time as a result of dynamic modeling of fracture per-
Fig. 19 shows the change in aperture and the corresponding meability was observed in Example 2 as well.
effective stress on the planes of all three fracture sets in the injector
gridblock. Different stiffnesses and spacings of the joint sets create Conclusions
an anisotropic equivalent medium for the solid modeling. How- In this study, a full-field, iteratively coupled geomechanics and
ever, in terms of fracture permeability, joint properties were cho- dual-porosity reservoir simulator has been successfully developed
sen such that the medium is initially isotropic for fluid flow, with that is able to model fracture permeability in a dynamic fashion.
permeability of 161 md in all directions. The anisotropic nature of The model development includes the effect of coupling through
deformation causes less change in effective stress in the x-direction rock deformation and matrix porosity as well as the effect of
in comparison to change in effective stress in the y-direction be- deformation on fracture permeability. However, the examples dis-
cause of higher intensity of soft fractures normal to the x-direction cussed in this paper have focused on the demonstration of the
(note that the boundary conditions are identical in these two di- effect of dynamic-permeability calculation on fluid-flow behavior
rections). Less change in effective stress in the x-direction results of naturally fractured reservoirs.
in a very small variation in aperture of the fractures with planes The fracture-deformation phenomenon has a great influence on
normal to this direction. performance of naturally fractured reservoirs because of strong
The impact of different fracture deformations on permeability dependency of fracture permeability on fracture aperture. Dynamic
is to decrease the ky /kx ratio from the initial value of 1 to approxi- treatment of fracture permeability does indeed require a coupled
mately 0.902 at 450 days at the injector gridblock, which slightly formulation, since stress is a key parameter in controlling fracture
elongates the tensor toward the x-direction. The reason for rela- deformation. Compared to conventional modeling, the inclusion of
tively small anisotropy despite the appreciable difference in frac- modeling fracture permeability dynamically leads to
ture closures is the number of fractures in each set. In other words, • Large changes in fracture permeability
in spite of small closure of fractures normal to the x-axis, their high • Lower recovery in depletion (or longer time to reach certain
intensity causes higher reduction in permeability in the y-direction recovery)
than in the x-direction. • Higher injectivity and lower productivity of wells
The effect of the permeability anisotropy on flow behavior is • Steeper pressure gradients in depletion
shown in Fig. 20. This plot shows the pressure distribution vs. • Less-steep pressure-gradient distribution in pressurization
distance along the two edges of the reservoir from the injector after • Permeability anisotropy
450 days. In the static case, pressures in the x- and y-direction are • Improving the run times
equal. However, for the dynamic case, the pressure at the edge In addition to the fluid-flow-behavior aspect of fractured rocks
parallel to the x-axis is somewhat higher than along the y-edge. dealt with here, the more realistic modeling of fractured-rock de-
This is a direct result of higher permeability in the x-direction. formation developed in this work will also be important for prob-

Fig. 16—Structure of stress and the reservoir model (Example 2). Fig. 17—Deformation of fracture sets under normal stress.

116 February 2008 SPE Reservoir Evaluation & Engineering


Fig. 19—Fracture aperture variation with time and effective
stress.
Fig. 18—Diagonal pressure distribution inside the flow model.
␣ ⳱ Biot’s coefficient
lems in which the geomechanical aspects are important (e.g. sub- ⌫ ⳱ traction vector
sidence, fault reactivation, and wellbore stability). These applica- ⌬v ⳱ fracture normal deformation
tions will be the subject of future investigations. ␦ ⳱ joint deformation vector
␧ ⳱ strain tensor, normal strain
Nomenclature
␴ ⳱ stress tensor, normal stress
a ⳱ constant in the joint deformation model
␶ ⳱ shear stress
A ⳱ total area of a fracture set in a block
␾ ⳱ porosity
b ⳱ constant in the joint deformation model
BV ⳱ bulk volume Superscripts
C ⳱ stiffness matrix f ⳱ individual fracture set
D ⳱ joint constitutive matrix I ⳱ intact rock
h ⳱ thickness J ⳱ joint
kn ⳱ joint stiffness M ⳱ number of joint sets
k ⳱ permeability T ⳱ joint
L ⳱ transformation matrix ⬘ ⳱ effective
Lx, Ly, Lz ⳱ gridblock dimensions
n ⳱ joint normal vector Subscripts
p ⳱ pressure f ⳱ fracture
Rb ⳱ effective well radius i ⳱ initial
Rw ⳱ wellbore radius i, j, k, l, I, J, L ⳱ tensor indices
s ⳱ skin factor j ⳱ joint
s៝ ⳱ strike vector of a fracture set m ⳱ matrix
S ⳱ joint spacing n ⳱ normal
៝t ⳱ dip vector of a fracture set N, S, T, n, s, t ⳱ joint local coordinate system
Tm ⳱ transformation matrix s ⳱ shear
v ⳱ fracture aperature t ⳱ total
V ⳱ volume v ⳱ volumetric
WI ⳱ well index x,y,z ⳱ coordinate direction

Fig. 20—Comparison of pressure distribution at the reservoir Fig. 21—Comparison of pressure distribution at the reservoir
edges near injector. edges near producer.

February 2008 SPE Reservoir Evaluation & Engineering 117


References Kenter, C.J., Blanton, T.L., Schreppers, G.M.A., Baaijens, M.N., and Ra-
Bagheri, M. 2006. Modeling Geomechanical Effects on the Flow Proper- mos, G.G. 1998. Compaction Study for Shearwater Field. Paper SPE
ties of Fractured Reservoirs. PhD dissertation, U. of Calgary, Calgary. 47280 presented at the SPE/ISRM Rock Mechanics in Petroleum En-
Bagheri, M. and Settari, A. 2004. Fracture Contribution in Reservoir Com- gineering Conference, Trondheim, Norway, 8–10 July. DOI: 10.2118/
paction. Paper presented at the ISRM Regional Symposium EUROCK 47280-MS.
and Geomechanics Colloqium, Salzburg, Austria, 7–9 October. Settari, A. and Mourits, F.M. 1998. A Coupled Reservoir and Geomechani-
Bagheri, M. and Settari, A. 2006. Effect of Fractures on Reservoir Com- cal Simulation System. SPEJ 3 (3): 219–226.
paction and Flow Modeling. Canadian Geotechnical J. 43 (6): 574– SPE Forum III—Reservoir Geomechanics. 2000. Breckenridge, Colorado.
586. Valliappan, S. and Khalili-Naghadeh, N. 1990. Flow Through Fissured
Bandis, S.C., Lumsden, A.C., and Barton, N.R. 1983a. Experimental Stud- Porous Media With Deformable Matrix. International J. for Numerical
ies of Scale Effects on the Shear Behaviour of Rock Joints. Interna- Methods in Engineering 29 (5): 1079–1094.
tional J. of Rock Mechanics and Mining Science & Geomechanics Youshinaka, R. and Yamabe, T. 1986. Joint Stiffness and the Deformation
Abstracts 18 (1): 1–21. Behavior of Discontinuous Rock. International J. of Rock Mechanics
Bandis, S.C., Lumsden, A.C., and Barton, N.R. 1983b. Fundamentals of and Mining Science & Geomechanics Abstracts 23 (1): 19–28
Rock Joint Deformation. International J. of Rock Mechanics and Min-
ing Science & Geomechanics Abstracts. 20 (6): 249–268.
Biot, M.A. 1941. General Theory of Three Dimensional Consolidation. J. SI Metric Conversion Factors
of Applied Physics 12 (2): 155–164.
cp × 1.0* E–03 ⳱ Pa⭈s
Biot, M.A. 1955. Theory of Elasticity and Consolidation for a Porous
Anisotropic Solid. J. of Applied Physics 26 (2): 182–185. ft × 3.048* E–01 ⳱ m
Chen, H.-Y. and Teufel, L.W. 1997. Coupling Fluid-Flow and Geome- ft3 × 2.831 685 E–02 ⳱ m3
chanics in Dual Porosity Modeling of Naturally Fractured Reservoirs. psi × 6.894 757 E+00 ⳱ kPa
Paper SPE 38884 presented at the SPE Annual Technical Conference *Conversion factor is exact.
and Exhibition, San Antonio, Texas, 5–8 October. DOI: 10.2118/
38884-MS.
Chen, H.-Y. and Teufel, L.W. 2000. Coupling Fluid-Flow and Geome- Mohammad Ali Bagheri is a reservoir engineer who works for
chanics in Dual Porosity Modeling of Naturally Fractured Reservoirs— Sproule International in Calgary, specializing in reservoir mod-
Model Description and Comparison. Paper SPE 59043 presented at the eling. E-mail: m.bagheri@sproule.com. His research interests
SPE International Petroleum Conference and Exhibition in Mexico, are numerical modeling of fluid flow and geomechanics in
Villahermosa, Mexico, 1–3 February. DOI: 10.2118/59043-MS. porous media. Bagheri holds a BSc degree in petroleum engi-
Ghafouri, H.R. and Lewis, R.W. 1996. A Finite Element Double Porosity neering from the Petroleum University of Technology and an
Model for Heterogeneous Deformable Porous Media. International J. MSc degree in reservoir engineering from the University of Te-
for Numerical and Analytical Methods in Geomechanics 20 (11): 831– hran, both in Iran. In 2006, he earned his PhD degree in petro-
leum engineering from the University of Calgary. A. (Tony) Set-
844.
tari holds the PanCanadian/Petroleum Society of CIM En-
Goodman, R.E. 1976. Methods of Geological Engineering in Discontinu- dowed Chair in Petroleum Engineering at the University of
ous Rocks. New York City: West Publication Company. Calgary, where he has taught for more than 25 years. He is also
Gupta, A., Penuela, G., and Avila, R. 2001. An Integrated Approach to the one of the principals in TAURUS Reservoir Solutions Ltd., a high
Determination of Permeability Tensors for Naturally Fractured Reser- technology simulation consulting firm in Calgary. Settari is one
voirs. J. Cdn. Pet. Tech. 40 (12): 43–48. of the leading developers of simulation technology for reser-
Huang, T.H., Chan, C.H., and Yang, Z.Y. 1995. Elastic Moduli for Frac- voir modeling, hydraulic fracturing, and geomechanics. He
tured Rock Mass. Rock Mechanics and Rock Engineering 28 (3): 135– holds a BSc degree from the Technical University of Brno,
144. Czechoslovakia and a PhD degree in mechanical engineering
from the University of Calgary. Settari has been involved in a
Kazemi, H., Merrill, L.S. Jr., Porterfield, K.L., and Zeman, P.R. 1976.
wide range of both research and development, and applica-
Numerical Simulation of Water-Oil Flow in Naturally Fractured Res- tion simulation projects, including naturally fractured reservoirs,
ervoirs. SPEJ 16 (6): 317–326. enhanced recovery projects, hydraulic fracturing and acidiz-
Khalili-Naghadeh, N. and Valliappan, S. 1991. Flow Through Fissured ing, in-situ thermal processes in oil sands, perforation mechan-
Porous Media with Deformable Matrix: Implicit Formulation. Water ics, and, more recently, reservoir geomechanics. He served on
Resources Research 27 (7): 1703–1709. the editorial board of JPT and was a Technical Editor for SPE.

118 February 2008 SPE Reservoir Evaluation & Engineering

You might also like