Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Petroleum Science and Engineering 149 (2017) 281–291

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Development and validation of an explicitly coupled geomechanics module MARK


for a compositional reservoir simulator

Mahdi Haddada, , Kamy Sepehrnoorib
a
The University of Texas at Austin, 200 E. Dean Keeton, Stop C0304, CPE 4.148, Austin, TX 78712-1575, USA
b
The University of Texas at Austin, 200 E. Dean Keeton, Stop C0304, CPE 4.104A, Austin, TX 78712-1575, USA

A R T I C L E I N F O A BS T RAC T

Keywords: Pore pressure-stress analyses in stress-sensitive reservoirs investigate interactions between the in-situ stress
Geomechanics and fluid flow; these interactions help or resist production, or conclude surface subsidence during production.
Compositional reservoir simulator Among the tools for these analyses, an explicitly coupled geomechanics and fluid flow model provides an
Coupled pore pressure-stress analysis essential, reliable, and fast production estimate for field planning and development. In this work, we
Finite element method
implemented this model in an in-house, three dimensional, compositional reservoir simulator, UTCOMP,
Chin's iterative coupling
using Chin's iterative coupling method. This development integrated a stand-alone geomechanics module based
on finite element method with the reservoir simulator, an advantage of our coupling algorithm, and improved
our understanding of the production through various enhanced oil recovery processes such as water and CO2
flooding processes previously coded in UTCOMP. Benefiting from the higher time scales of solution variations
due to the geomechanics module, we lowered the frequency of calling this computationally expensive module.
Also, we reduced the order of the finite element shape functions for displacement from quadratic to linear,
which majorly mitigated the high computational cost of our geomechanics studies while we almost maintained
the solution accuracy. To validate our implementation, we investigated a primary oil production case and
compared the results from UTCOMP with those from two other simulators: (1) CMG software program using
different coupling methods; and (2) another pre-validated in-house reservoir simulator, GPAS. In order to
evaluate our improvements in this work, we compared our results with those from a pre-validated in-house
reservoir simulator, GPAS. We observed a minor discrepancy between the solutions at very early times which
originates from the different structures in these two reservoir simulators, IMPEC in UTCOMP and fully implicit
in GPAS.

1. Introduction Method (FEM). This simulator can solve highly nonlinear problems;
however, this advantage hardly justifies the computational expenses of
In stress-sensitive reservoirs, the interaction between in-situ stress this development. The remedies for the high computational expenses
and fluid flow has some complications, which might be in favor of can be listed as the following: (1) using reduced-order elements for
production or against it (Pan, 2009). Also, due to production from displacement in the geomechanics module; (2) calling geomechanics
some extra loose formations, surface subsidence might occur, which module at lower frequencies; and (3) using IMPEC (IMplicit Pressure
suggests coupling a geomechanics model with the reservoir fluid flow Explicit Concentration) scheme instead of fully implicit scheme for
for the sake of a better simulation of this subsidence. An explicitly reservoir simulation. In this work, we attempted to accomplish these in
coupled geomechanics and fluid-flow model is one of the methods to another reservoir simulator with IMPEC scheme.
provide us with an accurate, reliable, and fast prediction for better well The theme of this work is to develop and validate a geomechanics
planning and reservoir management decisions. module in an in-house reservoir simulator; UTCOMP, which is one of
Pan (2009) integrated a geomechanics module with a fully-implicit the 3D compositional reservoir simulators developed at The University
reservoir simulator, GPAS, using fully coupling and iterative coupling of Texas at Austin (Chang, 1990). This simulator, at its current form,
schemes in conjunction with quadratic elements in Finite Element does not entail any geomechanics module despite the fact that various

Abbreviations: UTCOMP, University of Texas COMPositional oil reservoir simulator; GPAS, General Parallel Adaptive Simulator; IMPEC, IMplicit Pressure Explicit Concentration;
3D, Three Dimensional; EOS, Equation of State; VWP, Virtual Work Principle

Corresponding author.
E-mail addresses: Mahdi.haddad@utexas.edu (M. Haddad), kamys@mail.utexas.edu (K. Sepehrnoori).

http://dx.doi.org/10.1016/j.petrol.2016.10.044
Received 13 April 2016; Received in revised form 26 August 2016; Accepted 27 October 2016
Available online 29 October 2016
0920-4105/ © 2016 Elsevier B.V. All rights reserved.
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

Nomenclature ν Poisson's ratio


ρj Mass density of phase j (lbm / ft 3)
ϕ Reservoir porosity ε(⃗ x, y, z ) Strain vector composed of six independent components of
k Reservoir permeability strain tensor at (x,y,z)
ϕ* True porosity εi The ith component of strain on the plane normal to i
ϕ0 Initial reservoir porosity at reference pressure direction
εv Volumetric strain γij The jth component of strain on the plane normal to i
∀b Reservoir bulk volume for a cell ( ft 3) direction
∀ 0b Initial reservoir bulk volume for a cell ( ft 3) U Displacement vector
∀*b , ∀′b True bulk volume for a cell ( ft 3) u x-component of U
∀p Reservoir pore volume for a cell ( ft 3) v y-component of U
∀*p True pore volume for a cell ( ft 3) w z-component of U
t Time (day) ui X-component of displacement in node i of an element in
Ni Number of moles for component i geomechanics model
np Number of phases vi Y-component of displacement in node i of an element in
ξj Molar density of phase j (lbmole / ft 3) geomechanics model
xij Mole fraction of component i in phase j wi Z-component of displacement in node i of an element in
K Permeability tensor or a set of history dependent tensorial geomechanics model
quantities (mD) φ1 to φ8 Shape functions in 8-node brick element
krj Relative permeability for phase j (dimensionless) (ξ , η , ζ ) Finite element local coordinate system
P Pore fluid pressure (psi)
Pj Pressure of phase j (psi) Subscripts and Superscripts
γj Gravity term for phase j, defined as ρj g ( psi / ft )
D Depth from a reference datum plane (ft) 0 First time step
qi Molar injection (positive) or production (negative) rate for n Nth time step or power-law exponent
component i (lbmole / day ) max Maximum value
nc Number of hydrocarbon components T Transpose
c∼p Equivalent compressibility
E Young's modulus

reservoir engineering processes require geomechanics calculations for simulator for the next time step calculations. This procedure is based
a better understanding of production and field development. Moreover, on true Biot's consolidation theory and simplifies Chin's iterative
the IMPEC scheme used in UTCOMP allows easier integration of coupling method (Chin et al., 2002).
various physical phenomena in this reservoir simulator in comparison We simplified the Chin's iterative loop in a time step between the
with fully-implicit reservoir simulators such as GPAS (Pan, 2009). geomechanics module and reservoir simulator to a single execution of
Therefore, in this work, we implement this module and couple it with the geomechanics module per time step. Thereby, we call our im-
the existing code of the UTCOMP reservoir simulator; this coupling is plementation method as “explicit coupling” procedure due to the time-
based on the modified Chin's Iterative coupling method (Chin et al., marching nature of running the reservoir simulator and geomechanics
2002). We validate our results comparing them with the results from module. This simplification may not significantly influence the solution
CMG software program at various coupling procedures (GEM User knowing that UTCOMP is an IMPEC reservoir simulator that applies
Guide, 2015), and GPAS which is a pre-validated in-house reservoir small time steps for the stability and convergence of the solution.
simulator developed at The University of Texas at Austin (Pan, 2009; However, calling the geomechanics module in every small time step in
Wang et al., 1997). The CMG's coupling procedures define different UTCOMP significantly increases the computational expenses.
explicit dependence of porosity on deformation, pore pressure, tem- Therefore, we also attempted to reduce the computational expenses
perature, volumetric strain, and total mean stress (GEM User Guide, by lowering the frequency of calling the geomechanics module.
2015). These functionalities provide various degrees of coupling of We have two porosity definitions in this work: (1) reservoir
geomechanical effects with the reservoir simulator, and thereby, the porosity, ϕ (=∀p /∀b where ∀p and ∀b denote the reservoir pore and bulk
range of solutions to be compared with the results from our imple- volumes, respectively), used in the reservoir fluid-flow model; and (2)
mentation. true porosity, ϕ* (=∀*p /∀*b where ∀*p and ∀*b denote true pore and bulk
volumes, respectively), calculated from the stress analysis model. Eqs.
(1) and (2) relate these two porosity definitions together via the
2. Method volumetric strain, εν , and initial porosity, ϕ0 in virtue of grain
incompressibility and true and reservoir pore volume equality (Pan,
UTCOMP is a non-isothermal EOS compositional reservoir simu- 2009; Chin et al., 2002; Pan et al., 2007):
lator coded in FORTRAN and enables researchers to investigate a wide
variety of important enhanced oil recovery processes (Ghasemi Doroh,
Table 1
2012; Chang, 1990). In this simulator, IMPEC scheme is used for
Equivalent compressibility, c∼p (Pan, 2009).
solving the governing system of equations for fluid flow in porous
media; pressure equation is solved implicitly and concentration values Dimension (n) Equivalent Compressibility (c∼p, n,n = 1, 2, 3)
are calculated explicitly. More details about the IMPEC scheme in
UTCOMP can be found elsewhere (Ghasemi Doroh, 2012). In our 1 (1 − 2ν )(1 + ν )
E (1 − ν ) ϕ0
implementation, geomechanics module is developed and called right 2 2(1 − 2ν )(1 + ν )
after solving the system of equations for pressure every single or Eϕ0
3 3(1 − 2ν )
multiple time step(s), and the geomechanics output, which includes Eϕ0
new porosity and permeability values, is passed to the reservoir

282
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

ϕ*=1 − (1 − ϕ0 ) eεν (1) acceptable (Bathe, 2001; Hughes, 1987). Moreover, calling the geo-
mechanics module in every single time step unnecessarily increases the
⎛ 1−ϕ0 ⎞
ϕ=⎜ ⎟ ϕ* computational expenses without much gains in the solution accuracy.
⎝ 1−ϕ* ⎠ (2) Thereby, we attempted to resolve these issues in our implementation
without compromising the solution accuracy and convergence.
Here, the volumetric strain, εν , is positive in compaction which is in
In order to improve the computational efficiency of the geomecha-
agreement with the well-established geomechanics stress and strain
nics module, we further improved the code with the following features:
sign convention. Notably, this sign convention is different from that in
(1) implementation of first-order elements besides the second-order
the work done by Pan (2009) and Chin et al. (2002). Substituting Eq.
elements; and (2) reducing the frequency of calling the geomechanics
(1) in Eq. (2) results in Eq. (3), an explicit equation for reservoir
module. The formulation of the lower order finite element method is
porosity:
elaborated in the following section which is followed by the results for
ϕ = e−εν −(1 − ϕ0 ) (3) various test cases. Also, lowering the frequency of calling the geome-
chanics module can be implemented by accumulating the pore pressure
Also, reservoir porosity can be expressed in terms of true porosity
change during the time steps between the old and new geomechanics
after substituting the term (1 − ϕ0 ) from Eq. (1) in Eq. (3):
calls and passing the pore pressure change to the geomechanics module
ϕ ≅ ϕ* (1 − εν ) (4) to calculate new porosity and permeability values.
For clarity, the previous implementation by Pan (2009) in GPAS
A detailed derivation of Eqs. (1) through (4) can be found in
can be outlined as below:
Appendix A. These equations include the effect of the volumetric strain
in porosity, and are implemented after the calculation of the volumetric
1. Use true porosity, ϕ* in mass balance equation instead of reservoir
strain in the geomechanics module. Moreover, as expected, Eqs. (1)
porosity.
and (3) show that increasing the volumetric strain (which coincides
2. Use the approximate equivalent compressibility, c∼p as shown in
with the bulk volume contraction according to the signs convention for
Table 1 in order to improve the stability and robustness of the solver
the strains) results in the reduction of true and reservoir porosities or
of the fluid-flow system.
the pore volume. Considering Eqs. (1) through (4), we need to modify
3. (Optional) use Eq. (7) to update the absolute permeability based on
the mass balance equation as Eq. (5) (Pan et al., 2007):
the variation of porosity (Chin et al., 1998):
D [(1 − εν ) ϕ*Ni ] np ⎡ K krj ⎤ D (ϕ*Ni )
∀b − ∀b ∇. ∑ j =1 ⎢ξ j xij μ (∇Pj − γj ∇D ) ⎥+ ∀b εν Dt − qi=0,
Dt ⎣ j ⎦
k / ki=(ϕ*/ ϕi*)n , (7)
i=1, 2, …, nc +1
(5)
where k and ki are the current and initial permeability values, and ϕ*
Since we assume infinitesimal strains, we can neglect the last term, and ϕi* denote the current and initial true porosities, respectively. Eq.
D (ϕ*N )
∀b εν Dt i . Using Eq. (4) we have: (1) provides the current true porosity based on the initial true porosity
np ⎡ K krj ⎤ and the volumetric strain which is obtained from the geomechanics
D (ϕNi ) ⎢ξ x (∇Pj − γj ∇D ) ⎥− qi=0, module. The power-law exponent n is an empirical constant and
∀b − ∀b ∇. ∑ ⎣ j ij μj ⎦
Dt j =1 strongly depends on chemical and mechanical compaction processes
i=1, 2, …, nc +1 (6)
(David et al., 1994). The typical value of n is 3.0 after P & M theory
where ϕ is calculated using Eq. (4) depending on the true porosity, ϕ* (GEM User Guide, 2015) and for chemical compaction processes
and the volumetric strain,εν ; these two parameters are evaluated in the (David et al., 1994). Generally, this power-law relationship between
geomechanics module and transferred to the reservoir simulator for the permeability and porosity is established for stress-sensitive, weak, or
subsequent time step(s) calculations. unconsolidated rocks, and the exponent n is considered as a matching
Eq. (6) uses the same terms as in the mass balance equation in parameter (Chin et al., 1998) and is obtained empirically through
traditional reservoir simulators without the geomechanics effect, which curve-fitting of the permeability reduction as a function of rock
simplifies the integration of a reservoir simulator with a stress analysis porosity with experimental data. This procedure can conclude n in
simulator using minimal modifications. One important modification the wide range of 4.6–25.4 for sandstones (David et al., 1994). For
(for the sake of the porosity initialization phase) is to replace the instance, for a high permeability reservoir at Gulf of Mexico deep-water
traditional constant pore compressibility scalar with the so-called
equivalent compressibility, c∼p , as shown in Table 1 for different
geometric dimensions.
The general time-marching procedure can be expressed as the
following: in every time step, the reservoir porosity and permeability,
and the pore pressure are passed from the reservoir simulator to the
geomechanics module and the updated porosity and permeability
values are returned from the geomechanics module to the reservoir
simulator for the next time step(s). The updated porosity and perme-
ability may be used for multiple subsequent steps in order to reduce the
computational cost of calling the geomechanics module. Moreover, a
type of coupling updates only porosity values after the geomechanics
calculation and permeability remains constant through the simulation.
For further details on the iterative coupling procedure, readers are
referred to Chin et al. (2002) and Tran et al. (2002).
We primarily implemented higher order (quadratic) finite elements
for displacement in the geomechanics module, which caused heavy
computational expenses. Using quadratic elements sufficiently satisfies
severe inf-sup condition for mathematical convergence of our problem;
however, a lower order element not fulfilling this condition may still be Fig. 1. 8-node “brick” element (Smith and Griffiths, 2004).

283
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

turbidite sands, n was obtained to be equal to 5.6 using curve-fitting by 1


φi(e)= (1 + ξξi )(1 + ηηi )(1 + ζζi ),
Chin et al. (1998). Also, Pan (2009) used generic n values equal to 6 8 (10)
and 15 for 3D three-component CO2 injection and 3D two-component
where ξi , ηi , and ζi are the coordinates of the ith corner node in the local
water injection cases, respectively. Considering the fact that n always
coordinate system (ξ, η, ζ ).
holds positive values with high magnitudes in Eq. (7), even small
The displacement vector in Eq. (8) can be re-written as
increments of porosity lead to significant permeability increases, which
{∑ φi ui ∑ φi vi ∑ φi wi }T , or Eq. (11):
models the nonlinear contribution of pore throat size in the perme-
ability during pore body expansion (porosity increment). Notably, ⎧u⎫
⎨v⎬ =[Φ1 Φ2 ⋯ Φ8 ](x, y, z ) ×{U1 U2 … U8 }T ,
originating from field observations, Eq. (7) nonlinearly and mechan- ⎩ w ⎭(x, y, z ) (11)
istically updates permeability as a function of true porosity (or
volumetric strain according to Eq. (1)) which inherently depends on where Φi denotes a 3×3 matrix equal to the product of φi and the unit
the temporal pore pressure changes or the in situ effective stresses. matrix, and Ui is a 3-component vector equal to { ui vi wi } where i
refers to the corner node number. The index (x, y, z ) denotes the spatial
Our implementation has this capability to skip modifying perme- value of the associated vector or tensor. Combining Eqs. (8) and (11)
ability during an explicit coupling procedure whereas the porosity provides the strain-displacement matrix which directly relates the
values still get updated after the geomechanics calculations. This spatial strain matrix to the element corner node displacements. This
capability can be used to study the explicit effect of permeability is the basis for computing stresses and strains in any point based on
modification on production results. Moreover, it is especially applicable their values in the element nodes. To complete the system of equations,
for the matrix loading situations above the critical pressure where grain we used the isotropic fully elastic constitutive law to calculate stresses
crushing and compaction creep occur and consequently, permeability from strains.
behaves independent of porosity variation (David et al., 1994). In this In UTCOMP, an appropriate time increment is calculated in every
work, we attempted to evaluate the significance of including this time step based on the criteria on the solution variation in the previous
permeability modification in a 3D water injection case using a big time step in order to achieve a stable and converged time marching
power-law exponent, n=15. Also, we set n=10 in the other cases in procedure (Chang, 1990). Furthermore, due to the new involved
order to test the viability of our implementation. physics after our implementation, another criterion on time increment
must be imposed specifically for the geomechanics module. One simple
criterion for better convergence of the solution in the later time steps
2.1. Lower order finite element method can be expressed in forms of the maximum allowable reservoir porosity
alteration through the domain as Eq. (12):
One approach to derive finite element discretization is to multiply if (Δϕmax > 10−4) Then Δt =Δt. (10−4 /Δϕmax ), (12)
shape function in the equilibrium equations and integrate over the
domain. The other approach that is more physical is to use virtual work where Δϕmax denotes the maximum porosity alteration in all grids
principle (VWP) (Zienkiewicz, 1972). The finite element derivation during the previous time step. This proposed criterion holds the
here is based on VWP for the deformable system. This method means following characteristics: (1) it assumes linear porosity variation with
that when loads are applied to a solid body and result in the time during a time step, which is a reasonable assumption due to the
deformations, the external work done by the external loads will be tiny time increments in UTCOMP; (2) it is activated when we turn on
converted into the internal work done by the stresses inside the body. the geomechanics model; (3) the constant 10−4 is selected empirically
This internal work is called strain energy, which is caused by the in order to obtain a stable solution and a favorable convergence rate;
effective stresses (Terzaghi, 1936; Biot, 1941) and is stored in the body. and (4) it uses the same time-step control as the CMG's convergence
According to the current work's convention for positive compressions, criterion on the maximum allowable pore pressure change in a time
the effective stress tensor is defined equal to the total stress tensor step. This latest keyword requires a user-defined value for pore
minus the product of pore pressure and unit tensor. pressure change in all grids above which the time step is cut back to
The pore space is occupied not only by the water phase but also the satisfy this convergence condition.
oil and gas phases. However, for the stress analysis purposes and We validated our implementation by comparing our simulation
simplicity, we neglect the capillary pressure (the hydraulic pressure results with those of CMG software program (GEM User Guide, 2015)
difference across phase interfaces) and proceed with the water phase for a simple primary production case. For comparison purposes, we
pressure as the so-called “pore pressure”. According to the definition of tested three different geomechanics coupling types in CMG:
strain, we can write the strain vector →
ε(x, y, z ) as the multiplication of an
operator matrix and displacement vector, as shown in Eq. (8). This
defines full strain-displacement relations.

{ε}={ εx εy εz γxy γyz γzx }T =


⎡ 0 0 ⎤T
⎢ ∂/∂x 0 ∂/∂y ∂/∂z ⎥ ⎧ u ⎫
⎢ 0 ∂/∂y 0 ∂/∂x ∂/∂z 0 ⎥ ⎨ v ⎬=[A]{U},
⎢⎣ 0 0 ∂/∂z 0 ∂/∂y ∂/∂x ⎥⎦ ⎩ w ⎭
(8)

where [A] is the operator matrix and {U} denotes the displacement
vector. The aforementioned volumetric strain can be expressed as the
summation of the first three entries of the above strain vector,
according to Eq. (9):
εv=εx +εy +εz (9)

We used linear 8-node brick-shaped elements as shown in Fig. 1,


and the associated linear shape functions can be summarized in Eq. Fig. 2. Geometry of the primary production case and the final pore pressure contours
(10): from CMG software program.

284
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

depending on volumetric strain in GCOUPLING 1 versus total mean


stress and Biot coefficient in GCOUPLING 2 (GEM User Guide, 2015).
Biot coefficient is assumed equal to the default value of 1.0, and the
total mean stress is internally calculated in CMG depending on the
geomechanical model's boundary conditions, Young's modulus,
Poisson's ratio, and bulk and rock compressibility values (Tran et al.,
2002; GEM User Guide, 2015). The vertical boundaries around the
reservoir and the bottom of the reservoir are constrained for normal
displacements and the top boundary is loaded vertically and can move
freely. These boundary conditions can be implemented in CMG using
the geomechanical coupling factor equal to 1 (GEM User Guide, 2015).
Notably, bulk compressibility is the reciprocal of bulk modulus and this
Fig. 3. Comparison of producing oil rate from different CMG couplings, UTCOMP, and
modulus can be expressed as E /3(1 − 2ν ), where E and ν denote
GPAS for the primary production case. Young's modulus and Poisson's ratio, respectively. However, this work
corrects porosity and permeability based on only volumetric strain.
Table 2 We validated the results systematically in only a primary produc-
Summary of input data for three-dimensional primary production problem. tion case where the overburden stress or geomechanical effect acts as
the dominating production mechanism. First, we tried to match the
Model Parameter Value
results without geomechanics for both UTCOMP and CMG cases.
Reservoir model Length (ft)×Width 1680 × 1680 × 60 Second, we attempted to investigate the difference between results
(ft)×Thickness (ft) from UTCOMP and CMG integrating geomechanics in the simulation.
Number of grid blocks 21 × 21 × 3
Grid block size (ft3) 80 × 80 × 20 3. Results and discussion
Porosity 0.2
Permeability (mD) 10
Initial water saturation 0.0000001 We investigated the performance of our geomechanics module
Simulation time (day) 500 implementation and coupling through the comparison of the results
Reservoir fluid composition C10 (100%) from our simulations with those from CMG software program (GEM
Initial Reservoir Pressure 1500
User Guide, 2015) and in-house reservoir simulator GPAS (Pan, 2009;
(psi)
Production well pressure 1200 Wang et al., 1997). The studied cases are listed as the following: (1)
(psi) primary production; (2) 3D water injection; and (3) 3D CO2 injection.
Production well grid 11 × 11 × (1 to 3) These cases were selected as they demonstrate various production
location mechanisms with variable coupling strengths with the geomechanics
Relative permeability Linear model (Table 3)
effects. For instance, geomechanical forces act as the main production
model
mechanism in the primary production case whereas the effect of these
Geomechanics model Number of elements 21 × 21 × 3 forces in production during CO2 injection may be counteracted due to
Element type 8-node brick the existing gas phase and compression.
Element size (ft3) 80 × 80 × 20
Young's modulus (ksi) 145
Poisson's ratio 0.3
3.1. Primary production case
Rock compressibility 3 × 10−6
( psi−1) The first investigated case is a 3D single-component primary
Permeability Power-law 10 production problem with a constant overburden pressure, as shown
Exponent, n in Fig. 2. This problem is solved using two reservoir simulators,
UTCOMP (Chang, 1990), and CMG (GEM User Guide, 2015), and
the oil production rate from these two simulators is compared with and
Table 3
Linear relative permeability parameters.
without geomechanics in Fig. 3. In this case, both porosity and
permeability are modified and the power-law relationship is used for
Water Oil the modification of permeability using porosity. The parameters of this
case are listed in Tables 2 and 3.
Residual saturation 0.25 0.15
End point relative permeability 1.0 1.0
The geomechanics boundary condition in this case is zero normal
Exponent of relative permeability 1.0 1.0 displacement on all of the boundary surfaces except the upper surface
on which the overburden pressure is exerted. This overburden pressure
is equal to a unit load.
GCOUPLING equal to 0 through 2. According to Fig. 3, the results of all of the simulators without
Briefly, using GCOUPLING 0 which is a one-way coupling method, geomechanics module match very well for producing oil rate.
fluid flow porosity does not depend on deformation, and fluid pressures Furthermore, the CMG case using GCOUPLING 0 gives the same
and temperatures are used by the geomechanics module to update result as the cases without geomechanics since this CMG coupling only
formation strains and stresses. This method is useful when the fluid- modifies Young's modulus. The different oil production rates from
flow calculations use a porosity model that cannot be approximated UTCOMP, GPAS, and CMG when activating the geomechanics calcula-
well by the other "two-way" coupling models, GCOUPLING 1 and 2. In tions originate from their different numerical schemes: IMPEC for
our work in UTCOMP, rock compressibility is modified based on UTCOMP and CMG, and fully implicit for GPAS. The first scheme
Young's modulus, Poisson's ratio, and the initial porosity as in modifies saturation in the next time step based on pore pressure
Table 1, whereas GCOUPLING 0 in CMG modifies Young's modulus change in the current time step whereas the second scheme obtains
based on rock compressibility, Poisson's ratio, and initial porosity. both saturation and pore pressure changes at the current time step.
However, these two modifications are consistent with each other. This difference may conclude negligible discrepancies between the
Both coupling Types 1 and 2 use pressure and temperature for results in cases without the geomechanics calculations considering the
porosity modification; however, the pressure coefficient is calculated smaller time steps in the IMPEC scheme. However, the spatial

285
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

Table 4 plete description can be found in GEM User Guide (2015). The higher
Summary of input data for the 3D water injection case. production rate associated with GCOUPLING 2 is in good agreement
with the UTCOMP results, which validates our geomechanics imple-
Model Parameter Value
mentation in UTCOMP.
Reservoir model Length (ft)×Width 350 × 350 × 60 Moreover, CMG simulator is capable of updating geomechanics
(ft)×Thickness (ft) quantities with a constant frequency indicating the number of elapsed
Number of grid blocks 7×7×3
time steps per geomechanics calls (GEM User Guide, 2015). We used
Grid block size (ft3) 50 × 50 × 20
Porosity 0.3
this capability in the primary production case for GCOUPLING=2 in
Permeability (mD) 100 CMG to increase the mentioned frequency value from 1 to 2, which
Initial water saturation 0.2 reduced the CPU time by a factor of 0.58. This computational time
Simulation time (day) 2000.0 reduction should be almost identical between different cases consider-
Injection well water 100
ing the following: (1) the majority of the overall computational time is
content (%)
Reservoir fluid composition C10 (100%) spent during geomechanics calculations; and (2) the geomechanics
Water injection rate (STB/ 400 computational expenses are proportional to the number of geomecha-
day) nics module calls. This hypothesis is later investigated by comparing
Production well pressure 600 the computational time saving in the current case, primary production
(psi)
Injection well grid location 1×1×1 to 1×1×3
in CMG, with that in the water injection and CO2 injection cases when
Production well grid 7×7×1 to 7×7×3 the frequency of the geomechanics calls is reduced.
location From the mechanical point of view, in our implementation where
Relative permeability Corey's model we have no normal displacement in the lateral walls, a higher oil rate is
model (Table 5)
reasonable as a result of active geomechanics effects; pushing against
Geomechanics model Number of elements 7×7×3 the reservoir through the overburden pressure. This simple case is
Element type 8-node brick similar to the case of squeezing a sponge fully saturated with water. As
Element size (ft3) 50 × 50 × 20 expected, leaving the sponge on a table does not change the sponge's
Young's modulus (ksi) 13 water saturation whereas holding it in hand and squeezing it drains
Poisson's ratio 0.3
most of the water out of the sponge.
Modified Rock 0.00030312691
compressibility ( psi−1)
In order to check the performance of our implementation for other
Permeability Power-law 15 major reservoir engineering processes, we applied the developed
Exponent, n program for the geomechanics module in UTCOMP on 3D cases of
water and CO2 injection. Notably, the above comparison with CMG
results sufficiently validated our implementation. Thereby, in the
Table 5 following cases, we only compared our results with those from a pre-
Corey's model relative permeability parameters for the 3D water injection case.
validated in-house reservoir simulator, GPAS, in order to evaluate our
Water Oil improvements in this work.

Residual saturation 0.2 0.25


End point relative permeability 1.0 1.0 3.2. 3D water injection case
Exponent of relative permeability 2.0 2.0
Water injection is a common secondary recovery method which is
performed in most oil reservoirs before application of enhanced oil
recovery methods. In this method, the characteristics of the region in
the vicinity of the wellbore are of great importance for the water
injectivity as an extremely high injection pressure may apply on the
injection perforations and thereby, locally alter e.g. the porosity and
permeability of the reservoir. Therefore, in this case, we attempted to
couple water injection process simulated by UTCOMP to our geome-
chanics module in order to further understand the realistic behavior of
the reservoir under high mechanical and hydraulic loads. We assumed
a cubic reservoir isolated hydraulically and mechanically from the
vertical surroundings, fixed at the bottom wall, and loaded on top with
the overburden pressure. The detailed input data for the model re-
Fig. 4. Comparison of cumulative oil production in GPAS and UTCOMP with and
without geomechanics module for the 3D water injection case. “por” in the legend
represents porosity coupling whereas “por+perm” denotes porosity and permeability
coupling.

modification of porosity after the geomechanics calculations leads to


more pore pressure changes which influence saturation in the current
or next time step using fully implicit or IMPEC scheme, respectively.
Thereby, the effect of the geomechanics calculations on saturation lags
until the next time step using IMPEC scheme whereas this effect is
observed at the same time step using the fully implicit scheme.
Therefore, the fully implicit scheme leads to slightly higher oil
production rates compared to the IMPEC scheme.
CMG coupling Types 1 and 2 give lower and higher oil rates with
respect to the case without geomechanics module, respectively, due to Fig. 5. Comparison of cumulative oil production in GPAS and UTCOMP with and
the different coupling procedures of these two methods whose com- without geomechanics module for the 3D water injection case (closer view).

286
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

time steps with the variation of pore pressure within the current time
step or reservoir porosity within the previous geomechanics call. As a
complementary, UTCOMP reservoir simulator contains multiple cri-
teria to increase or decrease the time step size depending on the change
of the solution variables at the current time step (Chang, 1990). These
criteria may indirectly take into account the effect of the porosity
change in both the convergence rate and the next time step size, e.g.
through the mass balance equation, Eq. (6), where the reservoir
porosity is modified using the geomechanics module.
Furthermore, in another study, we attempted to demonstrate the
grid independency of the solution. For that purpose, we refined the grid
Fig. 6. Comparison of oil production rate in GPAS and UTCOMP with and without used for the 3D water injection case from 7 × 7 × 3 to 25 × 25 × 10
geomechanics module for the 3D water injection case. keeping all other model parameters the same as those in Tables 4 and
5. Having compared the average reservoir pressure, oil production rate,
generation can be found in Tables 4 and 5. and cumulative oil production from the investigated cases, we con-
Fig. 4 compares the cumulative oil production in the 3D water cluded that the discrepancy between the results due to grid refinement
injection case from two simulators, GPAS and UTCOMP with and is negligible, which proved the independence of the solution from the
without geomechanics, and with or without permeability modifications. grid.
As mentioned before, in an explicit coupling procedure, updating
permeability values can be performed optionally while the porosity 3.3. CO2 injection case
modification provides always a geomechanics coupling. Fig. 4 shows
the perfect match between different methods, which verifies our The last investigated case examines the geomechanics effect on
implementation and the negligible geomechanics effects on production. production where the main driving mechanism is oil swelling, reduc-
This negligible effect is reasonable as the reservoir is fully saturated tion of oil viscosity, and IFT through achieving miscibility conditions
with a single component liquid with low compressibility values, which depending on thermodynamic properties including pore pressure.
does not allow the contraction of the pore space or the porosity and Notably, pore pressure is one of the transferring parameters from the
permeability alteration under the overburden loads. Moreover, Fig. 5 reservoir simulator to the geomechanics module. Considering the effect
shows the small difference between the cumulative oil production of pore pressure alteration on porosity and permeability, the indirect
profiles by a closer look into the curves shown in Fig. 4. The difference influence of our geomechanics coupling on thermodynamic calcula-
is around 0.3% of the average value after 2000 s of injection or tions can be easily inferred. Moreover, CO2 injection has recently
production. attracted significant attention and field application as it provides a
Fig. 6 compares the oil production rate from different simulators great potential for carbon sequestration besides enhancing the oil
and coupling methods, and is in agreement with the results in Figs. 4 recovery. Tables 6 and 7 provide the detailed model description for this
and 5. Noticeably, the oil rate from “UTCOMP+GEOM (por+perm)”
curve maintains the highest level right before the abrupt reduction Table 6
Summary of input data for the 3D CO2 injection case.
around 750 days while it drops to the lowest level right after this sharp
reduction. The integration of this variation in time, as shown in Figs. 4 Model Parameter Value
and 5, smooths the observed transient behavior in oil rate and thereby,
the geomechanics effects. Therefore, the oil rate comparison provides a Reservoir model Length (ft)×Width 400 × 400 × 45
(ft)×Thickness (ft)
better method to evaluate the effect of the geomechanics coupling on
Number of grid blocks 10 × 10 × 3
production. Grid block size (ft3) 40 × 40 × 15
We attempted to reduce the frequency of geomechanics calls for the Porosity 0.3
3D water injection case. This concluded a significant CPU time Permeability (mD) 10
reduction by a factor of 0.6 when calling geomechanics module every Initial water 0.17
saturation
two time steps instead of every time step. Moreover, this reduction in
Simulation time (day) 4000.0
the frequency of the geomechanics calls did not influence the results as Injection well content CO2 (95%) and C1 (5%)
a comparison for the average reservoir pressure, oil production rate, Reservoir fluid CO2 (1%), C1 (29%), NC16
and cumulative oil production from different calling frequencies shows composition (70%)
a perfect match between the results. Total gas injection 10
rate (MSCF/day)
Interestingly, this CPU time reduction factor, 0.6, is close to the Production well 500
CPU time reduction factor, 0.58, in the primary production case which pressure (psi)
was obtained using GCOUPLING=2 in CMG. It should be noted that Injection well grid 1 × 1 × 1 to 1 × 1 × 3
both mentioned cases executed the geomechanics module every other location
Production well grid 10 × 10 × 1 to 10 × 10 × 3
time steps, which led to these CPU time reduction factors close to 0.5.
location
However, the difference between these CPU time reduction factors and Relative permeability Corey's model (Table 7)
0.5 reflects the contribution of the reservoir simulation calculations in model
the overall CPU time.
However, we could not obtain numerical convergence for geome- Geomechanics Number of elements 10 × 10 × 3
model Element type 8-node brick
chanics calls falling three or more time steps apart keeping the
Element size (ft3) 40 × 40 × 15
minimum and maximum time step size as before. This constant user- Young's modulus (ksi) 13
defined frequency is case-dependent, is obtained by trial and error Poisson's ratio 0.3
especially for complicated problems, and depends on the user-defined Modified Rock 0.00030312691
compressibility
allowable minimum and maximum time step size. This constant
(psi−1)
frequency should be replaced by a more rigorous criterion which would Permeability Power- 10
correlate the frequency of calling the geomechanics module for the next law Exponent, n

287
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

Table 7 the ideal CPU time reduction factor which is the inverse of the number
Corey's model relative permeability parameters for the 3D CO2 injection case. of time steps between the geomechanics module calls. For instance, this
normalized value in this CO2 injection case is 1.1 which is close to 1.16
Water Oil Gas
in primary production case or 1.2 in water injection case. Table 8
Residual saturation 0.2 0.1 0 summarizes our analysis on CPU time reduction due to lower calling
End point relative permeability 0.4 0.7 0.9 frequencies of the geomechanics module in all investigated cases.
Exponent of relative permeability 3.0 2.0 2.0
Fig. 10 compares the time step size through time for the 3D CO2
injection case in GPAS and UTCOMP with and without geomechanics
case where all boundaries except the top boundary are constrained for module. This comparison shows that UTCOMP increases time steps
normal displacement and a unit load is exerted on the top boundary. more conservatively mainly due to the IMPEC scheme where the
Figs. 7 through 9 compare respectively the cumulative oil produc- saturation is calculated explicitly through smaller time steps compared
tion, oil production rate, and average reservoir pressure for the 3D CO2 to those in a fully implicit and unconditionally stable reservoir
injection case in GPAS and UTCOMP with and without geomechanics simulator such as GPAS.
module where both porosity and permeability are modified. These
figures show that activating geomechanics slightly increases the 4. Conclusions
cumulative oil production and oil production rate, and slightly
decreases the average reservoir pressure. These slight differences can We developed a geomechanics module and explicitly coupled that
be justified considering the following observations: (1) the preferential with the UTCOMP simulator. We validated this implementation by
compression of the gas phase via the overburden loading, concluding a comparing the results from several cases against GPAS and CMG. The
small pore pressure increase and consequently, a slight porosity and main findings of this work can be summarized as follows:
permeability modification; (2) the positive effect of the overburden
pressure on production enhancement via porosity reduction; and (3) • Due to different coupling mechanisms, the approaches used in the
the transient porosity, and permeability distribution in the reservoir CMG simulator give higher or lower oil production rates with respect
specially in the vicinity of the injector and producer, which builds up to the case where the geomechanics module is inactive. The higher
the average reservoir pressure before the CO2 breakthrough and production rate is in good agreement with the UTCOMP results,
depletes that after the breakthrough. which validates our geomechanics implementation in UTCOMP.
Moreover, the discrepancy between the results from GPAS and • For most of the investigated case studies, UTCOMP results are in a
UTCOMP in the cumulative oil production and the oil production rate, good agreement with those from GPAS and Coupling 2 of CMG. The
as shown in Figs. 7 and 8, may originate from the different numerical slight difference between UTCOMP and GPAS is due to the different
schemes in these two simulators as explained comprehensively regard- structures of these two reservoir simulators; the former is IMPEC
ing the results in Fig. 3. Nevertheless, this discrepancy is transient and and the latter is fully implicit.
at later times, e.g. after 4000 days, the results converge to each other. • Running the geomechanics module every N time steps (where N is
Furthermore, we tried to reduce the number of geomechanics calls greater than one) can significantly reduce the CPU time while
in order to reduce the total CPU time while maintaining the accuracy of maintaining the solution accuracy. The number of time steps per
the solution. For this purpose, we built a new model similar to the CO2 geomechanics calls, however, requires a rigorous criterion other
injection case except that the frequency of geomechanics calls was than a constant user-defined number.
decreased from every time step to every 10 time steps. This calling • Water flooding is more sensitive to the number of time steps per
frequency reduction significantly reduced the computational cost by a geomechanics calls than the gas flooding cases such as CO2 injection.
factor of 0.11, almost proportional to the inverse of the number of time
steps per geomechanics calls. Furthermore, a comparison for the SI Metric Conversion Factors
average reservoir pressure, oil production rate, and the cumulative
oil production from the investigated cases in Figs. 7 through Fig. 9
ft×3.048 E−01=m
shows that calling geomechanics module every 10 time steps does not
ft 3×2.831685 E−02=m3
influence the accuracy of the solution.
mD×9.869233 E−16 = m2
This CPU time reduction factor can be compared with that in the
primary production or water injection cases after normalizing that by day×8.640000 E+04=s

Fig. 7. Comparison of cumulative oil production in GPAS and UTCOMP with and without geomechanics module, and various calling frequencies of geomechanics module for the 3D
CO2 injection case.

288
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

Fig. 8. Comparison of oil production rate in GPAS and UTCOMP with and without geomechanics module, and various calling frequencies of geomechanics module for the 3D CO2
injection case.

Fig. 9. Comparison of average reservoir pressure in GPAS and UTCOMP with and without geomechanics module, and various calling frequencies of geomechanics module for the 3D
CO2 injection case.

Table 8
CPU time reduction using lower calling frequencies of the geomechanics module. The
denominator values in the far right column represent the ideal CPU time reduction factor
and equal to the inverse of the number of time steps between geomechanics calls. .

Case Number of time CPU time Normalized


steps between reduction CPU time
geomechanics factor reduction
calls factor

Primary 2 0.58 0.58/0.5=1.16


production in
CMG-
GCOUPLING=2 Fig. 10. Comparison of time steps in GPAS and UTCOMP with and without geomecha-
3D water injection 2 0.6 0.6/0.5=1.2 nics module for the 3D CO2 injection case.
in UTCOMP
CO2 injection in 10 0.11 0.11/0.1=1.1
UTCOMP

Acknowledgments
psi×6.894757 E+03=Pa
ksi×6.894757 E+06=Pa The authors acknowledge the participants of Reservoir Simulation
psi−1×1.450377 E−04=Pa−1 Joint Industry Project (RS-JIP) at the Center for Petroleum and
bbl/day ×1.840131 E−06=m3/s Geosystems Engineering at The University of Texas at Austin for their
STB/day ×1.840131 E−06=m3/s financial support of the current work. Drs. Jalil Varavei and Feng Pan
STB×1.589873 E−01=m3 are highly acknowledged for their collaboration during the early stages
of our work on GPAS and UTCOMP. Also, we appreciate Dr. Lee Y.

289
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

Chin for his comments on our work in UTCOMP and the derivation of Engineering are highly appreciated for their valuable and meticulous
the porosity equations. We express our deep gratitude to Dr. Emad W. comments to improve the technical and verbal content of this paper.
Al-Shalabi who critically reviewed the manuscript. Furthermore, the Moreover, Computer Modeling Group, LLC is recognized for providing
anonymous reviewers of the Journal of Petroleum Science and academic CMG license for comparison and verification purposes.

Appendix A. True and reservoir porosities

In this appendix, we attempt to demonstrate the detailed derivation of Eqs. (1) through (4) starting from the definition of the volumetric strain
according to the geomechanics sign convention for strains; compaction strains are positive. The differential volumetric strain can be expressed as
Eq. (A-1):
d ∀′b
dεv=− ,
∀′b (A-1)
where ∀′b represents the true bulk volume, and the minus sign on the right hand side is due to the strains sign convention. Integrating Eq. (A-1) from
the initial state (with zero volumetric strain and initial true bulk volume of ∀ 0b ) to an arbitrary state (at volumetric strain equal to εv and true bulk
volume of ∀*b ) leads to Eq. (A-2):
εv ∀*b d ∀′b
∫0 dεv= ∫∀ 0

∀′b
.
(A-2)
b

Therefore, the volumetric strain can be expressed as Eq. (A-3):


∀*b ⎛ ∀* ⎞
εv=−ln ( ∀′b ) | =−ln ⎜ b0 ⎟ .
∀ 0b ⎝ ∀b ⎠ (A-3)
Thus, the bulk volume can be written as a function of the volumetric strain as Eq. (A-4):
∀*b = ∀ 0b e−εv . (A-4)
Eq. (A-4) shows that increasing volumetric strains (equivalent to further compaction of the bulk volume) results in smaller true bulk volumes as
expected.
Furthermore, assuming that the grain compressibility is zero, an expression can be established between the true porosity, ϕ*, and the initial
reservoir porosity, ϕ0 , using Eq. (A-5):

1−ϕ* ∀*g /∀*b ∀ 0b


= = ,
1−ϕ0 ∀ 0g /∀ 0b ∀*b (A-5)
where ∀*g and ∀ 0g
are true grain volume and the initial grain volume, respectively, which are equal to each other due to the grain incompressibility
assumption.
Combining Eqs. (A-4) and (A-5), we obtain Eq. (A-6):
1−ϕ* εv
=e ,
1−ϕ0 (A-6)
which can be re-arranged for ϕ* as a function of the volumetric strain as Eq. (A-7):
ϕ*=1 − (1 − ϕ0 ) eεv . (A-7)
Eq. (A-7) means that more volumetric strain leads to smaller true porosity as expected.
Also, assuming that the reservoir and true pore volumes, ∀p and ∀*p , are equal to each other, we can relate reservoir porosity to true porosity
through Eq. (A-8):
0
ϕ ∀p /∀ b ∀*b
= = .
ϕ* ∀*p /∀*b ∀ 0b (A-8)
Combining Eqs. (A-4), (A-6), and (A-8), we can write Eq. (A-9):
1−ϕ0 ϕ
= .
1−ϕ* ϕ* (A-9)
Thereby, an expression for reservoir porosity as a function of the volumetric strain can be obtained by substituting Eq. (A-7) in Eq. (A-9), which
leads to Eq. (A-10):
⎛ 1−ϕ0 ⎞ ⎛ 1−ϕ0 ⎞ ⎛ 1−ϕ ⎞
⎟=(1 − (1 − ϕ0 ) eεv ) ⎜ ⎟=(1 − (1 − ϕ0 ) eεv ) ⎜ ⎟.
0
ϕ=ϕ* ⎜
⎝ 1−ϕ* ⎠ ⎝ 1−(1−(1−ϕ0 ) e ) ⎠
εv ⎝ (1−ϕ0 ) e ⎠
εv
(A-10)
Simplifying Eq. (A-10) concludes Eq. (A-11) for reservoir porosity as a function of the volumetric strain:
ϕ=(1 − (1 − ϕ0 ) eεv ) e−εv=e−εv −(1 − ϕ0 ) (A-11)
Also, substituting the term (1 − ϕ0 ) in Eq. (A-11) from Eq. (A-6) leads to Eq. (A-12) in order to obtain an explicit relationship between true and
reservoir porosities:

290
M. Haddad, K. Sepehrnoori Journal of Petroleum Science and Engineering 149 (2017) 281–291

ϕ=e−εv −(1 − ϕ0 )=e−εv −(1 − ϕ*) e−εv=ϕ*e−εv . (A-12)


Using the first two terms in the Taylor series of e−εv in the vicinity of εv=0 concludes a simpler expression for reservoir porosity as Eq. (A-13):
ϕ ≅ ϕ* (1 − εv ) (A-13)

References Terzaghi, K., 1936. Theoretical Soil Mechanics. John Wiley and Sons, Inc., New York,
USA, ISBN: 978-0471853053.
Tran, D., Settari, A., Nghiem, L., 2002. New iterative coupling between a reservoir
Bathe, K., 2001. The Inf-sup condition and its evaluation for mixed finite element simulator and a geomechanics module. Paper SPE/ISRM 78192 Presented at the
methods. Comput. Struct. 79 (2), 243–252. http://dx.doi.org/10.1016/S0045- SPE/ISRM Rock Mechanics Conference, Irving, Texas, USA. DOI: http://dx.doi.org/
7949(00)00123-1. 10.2118/78192-MS.
Biot, M.A., 1941. General theory of three-dimensional consolidation. J. Appl. Phys. 12 Wang, P., Yotov, I., Wheeler, M.F., Arbogast, T., Dawson, C., Parashar, M., Sepehrnoori,
(2), 155–164. http://dx.doi.org/10.1063/1.1712886. K., 1997. A new generation EOS compositional reservoir simulator: Part I –
Chang, Y., 1990. Development and Application of an Equation of State Compositional formulation and discretization. Paper SPE 37979 Presented at the SPE Reservoir
Simulator (Ph.D. dissertation). The University of Texas at Austin, Austin, Texas. Simulation Symposium, Dallas, Texas, USA. DOI: http://dx.doi.org/10.2118/37979-
Chin, L.Y., Raghavan, R., and Thomas, L.K. 1998. Fully-Coupled Geomechanics and MS.
Fluid-Flow Analysis of Wells with Stress-Dependent Permeability. Paper SPE 48857 Zienkiewicz, O.C., 1972. The Finite Element Method in Engineering Science. McGraw-
Presented at the SPE International Oil and Gas Conference and Exhibition, Beijing, Hill, New York, ISBN: 0070941386.
China. DOI: http://dx.doi.org/10.2118/48857-MS.
Chin, L.Y., Thomas, L.K., Sylte, J.E., Pierson, R.G., 2002. Iterative coupled analysis of Mahdi Haddad is a graduate research and teaching assistant and Ph.D. degree
geomechanics and fluid flow for rock compaction in reservoir simulation. Oil Gas Sci. candidate in petroleum engineering at The University of Texas at Austin. He has in-
Technol. – Rev. IFP 57 (5), 485–497. http://dx.doi.org/10.2516/ogst:2002032. depth knowledge and experience in developing of 3D, multi-stage hydraulic fracturing
David, C., Wong, T., Zhu, W., Zhang, J., 1994. Laboratory measurement of compaction- models in shale resources using fully coupled pore pressure–stress analyses, Cohesive
induced permeability change in porous rocks: implications for the generation and Zone Model (CZM), and eXtended Finite Element Method (XFEM). Moreover, he
maintenance of pore pressure excess in the crust. PAGEOPH 143 (1), 425–456. developed a wax deposition model in pipelines using a mesh-less method, Smoothed
http://dx.doi.org/10.1007/BF00874337. Particle Hydrodynamics (SPH) in his master's research. He holds two bachelor's degrees
GEM User Guide, 2015. Computer Modelling Group Ltd., Calgary, Alberta, Canada. in mechanical and petroleum engineering and a master's degree in mechanical
Ghasemi Doroh, M., 2012. Development and Application of a parallel Compositional engineering, Energy Conversion, all from Sharif University of Technology. Haddad's
Reservoir Simulator (MS thesis). The University of Texas at Austin, Austin, Texas. current research focuses on the simulation of natural fracture reactivation and micro-
Hughes, T.J.R., 1987. The Finite Element Method: Linear Static and Dynamic Finite seismicity during hydraulic fracturing including strong stress interactions of intersecting
Element Analysis. Prentice-Hall, Englewood Cliffs, NJ, ISBN: 0-13-317025-X. fractures using CZM and Material Point Method (MPM).
Pan, F., Sepehrnoori, K., Chin, L.Y., 2007. Development of a coupled geomechanics
model for a parallel compositional reservoir simulator. Paper SPE 109867 Presented
at the SPE Annual Technical Conference and Exhibition, Anaheim, California, USA. Kamy Sepehrnoori is W.A. ‘‘Monty’’ Moncrief Centennial Chair in the Department of
DOI: http://dx.doi.org/10.2118/109867-MS. Petroleum and Geosystems Engineering at The University of Texas at Austin. His
Pan, F., 2009. Development and Application of a Coupled Geomechanics Model for a research areas include computational methods, reservoir simulation development and
parallel Compositional Reservoir Simulator (Ph.D. dissertation). The University of application, enhanced-oil-recovery modeling, naturally fractured reservoirs, and uncon-
Texas at Austin, Austin, Texas. ventional resources. Sepehrnoori holds a Ph.D. degree in petroleum engineering from
Smith, I.M., Griffiths, D.V., 2004. Programming the Finite Element Method. John Wiley The University of Texas at Austin.
& Sons, Ltd, Chichester, England, ISBN: 0-470-84969-X.

291

You might also like