You are on page 1of 9

Petroleum Reservoir Simulation Coupling

Fluid Flow and Geomechanics


M. Gutierrez, SPE, Virginia Tech, and R.W. Lewis and I. Masters, SPE, U. of Wales, Swansea

Summary
This paper presents a discussion of the issues related to the interaction between rock deformation and multiphase fluid flow behavior
in hydrocarbon reservoirs. Pore-pressure and temperature changes
resulting from production and fluid injection can induce rock deformations, which should be accounted for in reservoir modeling.
Deformation can affect the permeability and pore compressibility of
the reservoir rock. In turn, the pore pressures will vary owing to
changes in the pore volume. This paper presents the formulation of
Biots equations for multiphase fluid flow in deformable porous
media. Based on this formulation, it is argued that rock deformation
and multiphase fluid flow are fully coupled processes that should be
accounted for simultaneously, and can only be decoupled for predefined simple loading conditions. In general, it is shown that reservoir simulators neglect or simplify important geomechanical
aspects that can impact reservoir productivity. This is attributed to
the fact that the only rock mechanical parameter involved in reservoir simulations is pore compressibility. This parameter is shown to
be insufficient in representing aspects of rock behavior such as
stress-path dependency and dilatancy, which require a full tensorial
constitutive relation. Furthermore, the pore-pressure changes
caused by the applied loads from nonpay rock and the influence of
nonpay rock on reservoir deformability cannot be accounted for
simply by adjusting the pore compressibility.
Introduction
In the last two decades, there has been a strong emphasis on the
importance of geomechanics in several petroleum engineering
activities such as drilling, borehole stability, hydraulic fracturing,
and production-induced compaction and subsidence. In these
areas, in-situ stresses and rock deformations, in addition to fluidflow behavior, are key parameters. The interaction between geomechanics and multiphase fluid flow is widely recognized in
hydraulic fracturing. For instance, Advani et al.1 and Settari et al.2
have shown the importance of fracture-induced in-situ stress
changes and deformations on reservoir behavior and how hydraulic
fracturing can be coupled with reservoir simulators. However, in
other applications, geomechanics, if not entirely neglected, is still
treated as a separate aspect from multiphase fluid flow. By treating
the two fields as separate issues, the tendency for each field is to
simplify and make approximate assumptions for the other field. This
is expected because of the complexity of treating geomechanics and
multiphase fluid flow as coupled processes.
Recently, there has been a growing interest in the importance of
geomechanics in reservoir simulation, particularly in the case of
heavy oil or bituminous sand reservoirs,3,4 water injection in fractured and heterogeneous reservoirs,5-7 and compacting and subsiding fields.8,9 Several approaches have been proposed to implement geomechanical effects into reservoir simulation. The
approaches differ on the elements of geomechanics that should be
implemented and the degree to which these elements are coupled
to multiphase fluid flow.
The objective of this paper is to illustrate the importance of geomechanics on multiphase flow behavior in hydrocarbon reservoirs.
An extension of Biots theory10 for 3D consolidation in porous
media to multiphase fluids, which was proposed by Lewis and
Copyright 2001 Society of Petroleum Engineers
This paper (SPE 72095) was revised for publication from paper SPE 50636, first presented
at the 1998 SPE European Petroleum Conference, The Hague, 2022 October. Original
manuscript received for review 15 February 1999. Revised manuscript received 11 April
2001. Paper peer approved 16 April 2001.

164

Sukirman,11 will be reviewed and used to clarify the issues


involved in coupling fluid flow and rock deformation in reservoir
simulators. It will be shown that for reservoirs with relatively
deformable rock, fluid flow and reservoir deformation are fully
coupled processes, and that such coupled behaviors cannot be represented sufficiently by a pore-compressibility parameter alone, as is
done in reservoir simulators. The finite-element implementation of
the fully coupled equations and the application of the finite-element
models to an example problem are presented to illustrate the
importance of coupling rock deformation and fluid flow.
Multiphase Fluid Flow in Deformable
Porous Media
Fig. 1 illustrates the main parameters involved in the flow of multiphase fluids in deformable porous media and how these parameters
ideally interact. The main quantities required to predict fluid movement and productivity in a reservoir are the fluid pressures (and
temperatures, in case of nonisothermal problems). Fluid pressures
also partly carry the loads, which are transmitted by the surrounding
rock (particularly the overburden) to the reservoir. A change in fluid
pressure will change the effective stresses following Terzaghis12
effective stress principle and cause the reservoir rock to deform
(additional deformations are induced by temperature changes in
nonisothermal problems). These interactions suggest two types of
fluid flow and rock deformation coupling:
Stress-permeability coupling, where the changes in pore structure caused by rock deformation affect permeability and fluid flow.
Deformation-fluid pressure coupling, where the rock deformation affects fluid pressure and vice versa.
The nature of these couplings, specifically the second type, are
discussed in detail in the next section.
Stress-Permeability Coupling
This type of coupling is one where stress changes modify the pore
structure and the permeability of the reservoir rock. A common
approach is to assume that the permeability is dependent on
porosity, as in the Carman-Kozeny relation commonly used in basin
simulators. Because porosity is dependent on effective stresses, permeability is effectively stress-dependent. Another important effect,
in addition to the change in the magnitude of permeability, is on the
change in directionality of fluid flow. This is the case for rocks with
anisotropic permeabilities, where the full permeability tensor can be
modified by the deformation of the rock.
Examples of stress-dependent reservoir modeling are given by
Koutsabeloulis et al.6 and Gutierrez and Makurat.7 In both examples,
the main aim of the coupling is to account for the effects of in-situ
stress changes on fractured reservoir rock permeability, which in
turn affect the fluid pressures and the stress field. The motivation for
the model comes from the field studies done by Heffer et al.5 showing that there is a strong correlation between the orientation of the
principal in-situ stresses with the directionality of flow in fractured
reservoirs during water injection. There is also growing evidence
that the earths crust is generally in a metastable state, where most
faults and fractures are critically stressed and are on the verge of further slip.13 This situation will broaden the range of cases with strong
potential for coupling of fluid flow and deformation.
Deformation-Fluid Pressure Coupling:
Biots Theory for Multiphase Flow in
Deformable Porous Media
The coupled deformation and fluid-flow problem was first analyzed by Terzaghi12 in 1925 as a consolidation problem. Since then,
June 2001 SPE Reservoir Evaluation & Engineering

Terzaghis 1D consolidation theory has been used widely in settlement problems in saturated soils. Biot10 extended the theory into a
more general 3D case, based on a linear stress-strain relation and a
single-phase fluid flow. Here, we present an extension of Biots
equations for two-phase immiscible and isothermal flow.
Equations for three-phase flow can be found in Lewis and
Sukirman.11 In the following, tensorial notation is used and summation is implied for repeated indices.
For two-phase fluid flow, the generalized Darcys law is given as

HH

kij kr
v i =
p + gh .

x j

Fluid Pressure
(Temperature)

In-Situ Stresses

. . . . . . . . . . . . . . . . . . . . . . (1)

Rock Deformation

Permeability

In addition to Darcys law, Biots theory includes


(a) Terzaghis effective stress principle,
HH
HH
ij = ij ij p ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
(b) the stress-strain constitutive relation, including the compressibility of the solid grains,
HH
HH
dp
; . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
dij = Dijkl dkl + kl
3K s
(c) the strain-displacement compatibility relation,
HH 1 u u
dij = i + j
2 xi x j

; . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)

and (d) the static-equilibrium equation,


HH
ij
+ Fi = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)
x j
Eqs. 2 through 5 constitute the geomechanical part of Biots
equation. The equations relate the applied internal and external
loads Fi from the static equilibrium condition (Eq. 5) and the pore
pressure p from the effective stress equation (Eq. 2) to the deformation of the rock (Eqs. 3 and 4). The final equation is for mass
balance, which is written as

1
1
vi =
S

xi B
t B

HH

ij Dijkl kl
So
+ o kl
B
3K s t

HH

So 1 ij Dijklkl
+ o

2
B K s
( 3K s )

p + q .

. . . . . . . . . . . . . . . . . . . . (6)

The fluid accumulation term on the right side of Eq. 6 consists


of the following contributions:
(a) the rate of change of fluid volume and saturation for each
phase p,

1
S ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (7)

t B

(b) the rate of rock volumetric change,


HH
ij
v
;
= ij
t
t

Fig. 1Schematic of the interaction between rock deformation,


fluid flow, and temperature in a deformable reservoir.

(c) the rate of change of solid particle volume owing to pore


pressure change,

(1 ) p
; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (9)
K s t
and (d) the rate of solid-particle-volume change caused by the
change in mean effective stress,

HH
ij ij

3K s t

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (10)

Contributions (c) and (d) account for poroelastic effects by


including the grain compressive modulus Ks of the reservoir rock.
For reservoir rocks, poroelastic effects can be significant when the
matrix bulk modulus K has the same order of magnitude as Ks.
The initial formulation of Biots theory emphasizes mechanical
issues over fluid-flow issues. Because of this, the theory is less
compatible with conventional fluid-flow models in terms of the
parameters involved. A reformulation of the theory along the line
of conventional fluid-flow modeling can be found in Chen et al.14
Dual-porosity coupled models also have been proposed for fractured reservoirs by Chen and Teufel15 and Ghafouri.16 It should
also be noted that the theory is not restricted to elastic response of
the rocks and has been extended to thermoporoelastoplasticity.17,18
Reservoir Simulation
The main purpose of reservoir simulation is to model multiphase
fluid flow and heat transfer in porous media. The more advanced
reservoir simulators can handle multicomponent three-phase fluids
with complicated pressure/volume/temperature (PVT) relations
and relative permeabilities.
The equations governing the behavior of two immiscible fluids
flowing in a porous medium can be obtained by combining
Darcys law (Eq. 1) with the mass-balance equations for each flowing phase. In contrast to Eq. 6, the mass-balance law in reservoir
simulators is written simply as

xi

1
1

vi = S + q .
B
t
B

As in Biots equations, these two equations are supplemented


by the equations of state for various fluid properties.
= ( p ) ,

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (12)

k r = k r ( S o , S w ) ,

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)

June 2001 SPE Reservoir Evaluation & Engineering

. . . . . . . . . . . . . . . . . . (11)

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (13)

Pc = po pw = pc ( So , S w ) ,

. . . . . . . . . . . . . . . . . . . . . . . . . (14)
165

and So + S w = 1 .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (15)

Expanding the time derivative on the right side of Eq. 11 gives

S
B

1 1
1 S p
.
+
B
+

p B p S p t

. . . . . . . . . . . . . . . . . . . . . . (16)
The first term on the right side of Eq. 16 is the change in the
volume factor Bp with pressure, giving the fluid compressibility of
phase p as

1
1
.
= cf =
p
p B

then the second term on the right side of Eq. 16 may be rewritten as
. . . . . . . . . . . . . . . . . . . . . . . . (19)

where cp=pore compressibility defined as

cp =

1
.
p

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (20)

S
1 S p
= c f + S c p +
+ q .

B
S p t

. . . . . . . . . . . . . . (21)

Reduction of the Coupled Equations to a


Hydraulic Diffusivity Equation
The hydraulic diffusivity equation from Biots theory is obtained
by introducing Darcys law (Eq. 1) into the mass-balance equation
(Eq. 6). For the sake of simplicity, single-phase flow under isothermal conditions will be considered.
H
HH
kH

ij Dijkl
ij p

+ gh = kl

3K s
x j

kl
t

HH

1 ij Dijklkl

+
+

2
Kf
Ks
( 3K s )

p + q ,
t

. . . . . . . . . . . . . . . . . . . . . . (22)
where K f =1/cf is the bulk modulus of the fluid. This equation
should be compared to the single-phase version of the fluid diffusivity equation (Eq. 21) used in reservoir simulation,
166

. . . . . . . . . . . . . (24)

where K and G=the shear and bulk moduli, respectively, and are
related to the Youngs modulus E and Poissons ratio n as
K=

E
3 (1 2 )

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (25a)

E
.
2 (1 + )

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (25b)

Substituting Eq. 24 into Eq. 3 and the resulting equation into Eq. 2
yields the poroelastic
relates the total
 stress-strain relation, which

stress increment dsij to the strain increment deij and pore-pressure
increment dp.
HH
HH
2
dij = K G dvij + 2Gdij dpij ,
3

. . . . . . . . . . . . . (26)

where a=Biots poro-elastic constant defined as

=1

Substituting Eqs. 1, 17, and 19 into Eq. 11 yields the two-phase


hydraulic diffusivity equation,
HH

kij kr
p + gh

x j

xi

and G =

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (18)

1 1 p
=
= S c p ,
p p p

HH
2

Dijkl = K G ijkl + G ik jl + il jk ,
3

. . . . . . . . . . . . . . . . . . . . . . (17)

If the average fluid pressure p causing pore-volume change in


the effective stress law (Eq. 2) is calculated from11
p = pw S w + po S o ,

. . . . . . . . . . . . . . . . . . . . (23)

where ct=cf +cp is the total compressibility. Again, for the sake of
simplicity, elastic stress-strain behavior
will be considered, in

which case the constitutive tensor Dijkl equals

p
1
1
S =
S
p
B
t B

t
=

HH

kij
p
p + gh = ct
+q,

x j
t

K
.
Ks

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (27)

Substituting Eq. 24 in Eq. 22 yields the poroelastic hydraulic


diffusivity equation,18,19
H
H

kij p

1 p

+ gh = v +
+ q , . . . . . . . . . . . . (28)

t
M
xi x j

B t

where MB=the Biot modulus defined as



1
=
+
.
MB
Ks
Kf

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . (29)

Note that the fluid diffusivity equation (Eq. 28) is coupled to


the poroelastic stress-strain relation (Eq. 26) by the volumetric
strain increment den. Under specific conditions, these two equations can be decoupled and the problem reduced to that commonly
used in reservoir engineering. To do this, two assumptions must
be made to relate the volumetric strain den in Eq. 26 to the porepressure change dp:
The changes in the total stresses.
The loading condition (called stress path in geomechanics).
It should be noted that these assumptions are defined locally and
therefore applied to every point, as opposed to boundary conditions,
which are applied at the boundary of the domain of interest.
Consequently, the strain-displacement compatibility relation (Eq. 4)
and the equilibrium equation (Eq. 5) do not have to be invoked in
making these assumptions.
Hydrostatic Compaction. To decouple Eqs. 26 and 28 locally, one
assumption that can be made is that the reservoir rock is subjected
to a hydrostatic loading, with equal horizontal and vertical deformations under constant total
stresses. Applying the conditions

dex=dey = dez=den and dsij=0 in Eq. 26 yields
d z = Kdv dp = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . (30)

Substituting this into Eq. 28 gives the decoupled flow equation


June 2001 SPE Reservoir Evaluation & Engineering

H
H
2
kij p

1 p
+ gh =
+
+ q . . . . . . . . . . . . . (31)
K M B t
xi x j

Comparing this equation with Eq. 23 gives the total compressibility as

1 2
1
+

.
K MB

( ct )hydro =

. . . . . . . . . . . . . . . . . . . . . . . . . (32)

The subscript hydro is used in the above equation to indicate that


the pore compressibility was calculated assuming hydrostatic loading conditions. In the case of incompressible fluids, Kf and
cf =0; hence, cp=ct gives

(c )
p

hydro

1 2
+

. . . . . . . . . . . . . . . . . . . . . . . . . (33)
Ks
K

Neglecting further poroelastic effects by assuming that Ks


gives a=1, and the pore-compressibility parameter for hydrostatic
loading condition becomes

(c )

p hydro

1
.
K

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (34)

Oedometric Deformation. Another stress path commonly


assumed in determining the pore-compressibility parameter is uniaxial strain compaction (also called Ko compaction in geomechanics), in which the pore pressure is varied with constant total stresses
and the horizontal displacements are blocked. Uniaxial strain
compaction is usually assumed to be a good approximation of the
20
conditions undergone by a reservoir
 during depletion. Applying
the conditions dex=dey = 0 and dsij=0 in Eq. 26 yields
4

d z = K + G dv dp = 0 .
3

. . . . . . . . . . . . . . . . . . . . (35)

Substituting Eq. 35 in Eq. 28 yields


H
H
2
kij p

1 p

+ gh =
+

+q .
M M B t
xi x j

. . . . . . . . . . . . (36)

Correspondingly, the oedometric total compressibility is

( ct )oedo =

1 2
1
+

,
M MB

. . . . . . . . . . . . . . . . . . . . . . . . . (37)

where M=the constrained modulus defined as

E (1 )
4
. . . . . . . . . . . . . . . . . . . . . (38)
M =K+ G=
3
(1 2 )(1 + )
Again, neglecting fluid-compressibility and poroelastic effects
yields the pore-compressibility parameter for oedometric condition, which is

(c )

p oedo

1
.
M

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (39)

Consequences for Reservoir Simulation. The previous discussion has shown that the equations of poroelasticity can be reduced,
under specific local assumptions, to a hydraulic-diffusion-type
equation. Whereas the compressibility of a fluid can be considered
an intrinsic property under constant temperature, however, the pore
compressibility depends on the local conditions assumed. It was
shown that assuming local hydrostatic or oedometric conditions
gives two different values of pore compressibilities. The differences in the pore compressibilities for these two conditions can be
significant. It can be shown that the ratio of the two compressibility values is equal to
June 2001 SPE Reservoir Evaluation & Engineering

(c )
(c )

p hydro
p oedo

3 (1 )
1+

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (40)

For instance, a Poissons ratio of n=0.2 (a typical value for the


elastic response of typical reservoir rocks), the hydrostatic pore
compressibility is twice the oedometric pore compressibility.
Obviously, depending on the loading condition, a wide range of
pore-compressibility values is possible even for idealized elastic
materials. The deformation of reservoir rocks is, however, nonlinear and elastoplastic. Consequently, an even wider range of pore
compressibilities can be expected for reservoir rocks than from
elastic behavior. Pore compressibility can be infinite for nonstrainhardening perfectly plastic rocks. In the extreme case, pore compressibility can be negative when shear loading causes the rock
volume to increase under constant or increasing pore pressure.
In practice, rock compressibility will have an impact when its
absolute value is comparable to or greater than that of the fluid. The
magnitude of pore compressibility in relation to fluid compressibility should provide an initial indication of the possible importance of
fluid flow and deformation coupling on pore-pressure distribution and productivity. The bulk modulus of many reservoir rocks
rarely exceeds 10 GPa. In comparison, the bulk modulus of water
is about 2.25 GPa at standard conditions. It is expected, therefore,
that reservoir rock pore compressibility will generally be greater
than, or at least of the same order of magnitude as, the compressibility of reservoir fluids in oil/water systems. However, for cases
involving pressure maintenance by gasflooding or in case of a
solution-gas drive, the gas can increase the total system compressibility considerably, making the effects of pore compressibility less important.
Two remarks can be made at this point. First, it should be noted
that pore compressibility as defined in reservoir engineering is measured under constant total stresses. Also, from the previous discussion, the measured value of the compressibility pertains only to a
specific stress path and cannot be used for other stress paths.
Boutca19 emphasized the need to measure the pore compressibility
under oedometric conditions, on the basis that most of the reservoir
deforms under oedometric conditions during depletion.
Second, the assumption of oedometric or uniaxial strain condition for every point in a reservoir is only valid, strictly speaking,
for a horizontally infinite reservoir under uniform pressure drawdown. Reservoirs are, however, bounded laterally and do not
deform uniformly even under uniform pressure drawdown. This is
illustrated by the results of a stress analysis of an axisymmetric
(disk-shaped) thin reservoir, whose radius is the approximately the
same as its depth, subjected to a uniform pore-pressure reduction
(Fig. 2). Because of the stiffness and bending of the overburden,
the reservoir deforms nonuniformly, as shown by the displacement
vectors. Close to the reservoir centerline, the displacement vectors
are vertical; hence, the rocks are under uniaxial strain conditions.
Near the flanks, the horizontal displacements are comparable to the
vertical components depicting hydrostatic loading of the rock. In
general, the displacement and stress fields in a reservoir will
depend on the reservoir geometry, boundary conditions, and porepressure distribution, and will be different from the idealized uniaxial strain condition. The presence of discontinuities (e.g., faults
and fractures) and material inhomogeneities (e.g., layering of the
reservoir) will also affect the stress distribution.
Petroleum Reservoir Simulation Coupling
Multiphase Flow and Deformation
Biots formulation constitutes a fully coupled system of equations
following the definition of Zienkiewicz.21 Based on this definition,
fluid flow (Eqs. 1 and 6) and geomechanics (Eqs. 2 through 5)
form a set of separate domains that cannot be analyzed separately.
Consequently, the dependent variables (e.g., fluid pressures and
rock displacements) cannot be explicitly eliminated except, as seen
previously, when assuming specific local conditions. By comparing the governing equations for reservoir simulation (Eq. 21) and
Biots theory (Eqs. 2 through 6 and 22), the following differences
may be observed:
167

6.5 km

Reservoir centerline

11 km

3 km
0.3 km
0.7 km

Hydrocarbon
reservoir

N
38.5 km
Fig. 2Rock displacements caused by a uniform pressure
drawdown in a disk-shaped reservoir.

Reservoir simulation does not include any of the equations


from the geomechanical relations. This means, for instance, that
the calculated pore pressures may not be in equilibrium with the
overburden loads because the effective stress principle and the
equilibrium condition are not accounted for.
The only rock mechanical parameter involved in reservoir
simulation is the pore compressibility cp. This is a scalar quantity
and cannot sufficiently represent the behavior of rocks. Rock
stress-strain behavior is not only nonlinear but also stress-path
dependent and requires a full tensorial relation. As shown earlier,
even for linearly elastic and isotropic materials, different compressibility parameters are obtained depending on the loading path.
Gutierrez8 has analyzed numerically and theoretically the validity of the uncoupled approach. With a finite-element formulation,
the single-phase fluid flow diffusion equation was discretized as

( C + t[])[p] = ( t[][p] + [q] ) ,

. . . . . . . . . . . . (41)

where [Dp]=the matrix of pore-pressure change, [p]=the current pressure matrix, [C]=the compressibility matrix, [F]=the
permeability matrix, [Dq]=the matrix of fluid fluxes, and
Dt=the time increment. From this equation, the pore-pressure
changes can be solved as

p = C + t

) ( t p + q ) .


. . . . . . . . . . (42)

For the single-phase, fully coupled formulation, the finite-element


discretization is given as
K
F
L u

=
t
L t p t p + q

. . . . . . . (43)

where [K]=the stiffness matrix, [L]=the coupling matrix, and


[DF]=the matrix of boundary and self-weight loads. As shown, the
displacement increment matrix [Du] and the pore-pressure-change
matrix [Dp] can be solved simultaneously from this equation.
Solving the displacement field [Du] from the first equation of Eq.
43, substituting [Du] in the second, and solving for [Dp] gives

p = L K L + t

t p + q + L K F . . . . . . . . . (44)
t

The main differences between Eqs. 42 and 43 are:


Pore-pressure change is a function of the full-rock-stiffness
matrix [K] in the fully coupled formulation, while it is a function
only of the rock-compressibility matrix [C] in the uncoupled formulation. The former accounts for the full constitutive behavior of
168

23 km

4 km

Fig. 3Fully coupled model of an idealized reservoir.

the reservoir and nonpay rock system, while the latter is a diagonal
matrix that can only be made dependent on the pore pressure.
Pore-pressure change is also a function of the applied load
[L]t[K]-1[DF], caused by the total stress changes in the fully coupled analysis. Such total stress changes come from the weight of
the overburden, which is transferred nonuniformly to the reservoir
again, according to the pore-pressure distribution in the reservoir.
Application to a Field Case
The extended Biot equations for three-phase fluid flow in
deformable porous media were discretized by Lewis and
Sukirman,11 and the discretized equations were implemented in the
finite-element code CORES (COupled REservoir Simulator).
CORES is a 3D black-oil (three-phase compressible and immiscible fluid flow) simulator. The reservoir rock is modeled by elastic
and/or elastoplastic constitutive models, and the physical properties of the fluids depend on fluid pressures and saturations. In the
finite-element implementation, implicit procedures are used to
solve the fully coupled governing equations where the rock displacements and fluid pressures are the primary unknowns.
To illustrate the importance of analyzing fluid-flow and geomechanical behavior as fully coupled processes, CORES is
applied to the simulation of an idealized North Sea reservoir. The
reservoir has an area of about 6.5 by 11 km and a thickness of 300
m. The complete model, which includes the nonpay rock, has an
area of about 23 by 38.5 km and a thickness of 4 km (of which 3
km is the overburden, 0.3 km the reservoir, and 0.7 km the underburden). A simplified view of the whole model is shown in Fig.
3. A rough mesh with 8 by 20 by 14 eight-noded brick elements
was used in the simulation. However, the results of the 3D model
were also verified by a 2D model with a much finer discretization. No vertical displacements are allowed at the base of the
model, and no lateral displacements are allowed at the four sides
of the model. It is a common practice in geomechanical modeling
to extend the lateral boundaries as far away as computationally
possible from the main loaded region to simulate infinitely horizontal boundaries and minimize local boundary effects. The top
of the model, which corresponds to the seabed, is allowed to
deform freely.
The initial reservoir pressure is assumed to be 48 MPa, which is
uniformly distributed in the reservoir. The initial effective vertical
stress distribution vs. depth is integrated from the self-weights of
June 2001 SPE Reservoir Evaluation & Engineering

TABLE 1MATERIAL PROPERTIES USED IN MODELING


Overburden/Sideburden
Youngs modulus, E (GPa)
Poissons ratio,
Absolute permeability, k (md)

2.5
0.45
0

Underburden
Youngs modulus, E (GPa)
Poisson's ratio,
Absolute permeability, k (md)

13.5
0.45
0

Reservoir

Stiff

Soft

Youngs modulus, E (GPa)


Poissons ratio,
Porosity,
Pore compressibility, cp* (/MPa)
Permeability, k (md)

0.85
0.25
35
0.0028
150

0.05
0.25
35
0.0476
150

* Used only for the uncoupled flow model and calculated from Eq. 39 with
the same E and used in the fully coupled models. For comparison,
cf =4.4104/MPa at initial reservoir conditions (used for both uncoupled
and fully coupled simulation).

the different rock layers, while the initial effective horizontal stresses
are assumed to be one-half of the total vertical stresses. Only the
water and oil phases are considered in the reservoir to simplify the
analysis. Undrained conditions (i.e., no fluid flow or zero permeability) are used for the surrounding nonpay rocks. Realistic relative
permeability and capillary pressure curves, and standard oil and
water formation volume factor curves, were used for the fluid-flow
part of the flow simulation. Elastic rock properties were used in the
geomechanical simulation. The rock properties used are given in
Table 1. The model is analyzed for a 14-year production scenario
by specifying production rates in the finite-element nodes corresponding to production wells within the reservoir. The production
rates applied in the production wells are based on recorded production data in the field.
The calculated pore-pressure distribution at the end of simulation in the top reservoir layer is shown in Fig. 4. After 14 years of
production, the reservoir pore pressure has been reduced to approximately 25 MPa in much of the reservoir. However, despite the
continuous production, the pore pressures have increased to

Fig. 5Reservoir compaction and seabed subsidence at the


end of 14 years of production.
June 2001 SPE Reservoir Evaluation & Engineering

Fig. 4Pore pressure (in MPa) distribution in the reservoir after


14 years of production.

approximately 55 MPa in areas close to the reservoir flanks, which


is higher than the initial pressure of 48 MPa.
The predicted seabed subsidence and reservoir compaction
after 14 years of production are shown in Fig. 5. The maximum
calculated subsidence is about 6 m, which is approximately 85% of
the maximum reservoir compaction. An interesting result is the
expansion of the reservoir and the corresponding heave of the
seabed around the reservoir flanks. This result is an outcome of the
increased pore pressures shown in Fig. 4.
The increase in pore pressure close to the reservoir flanks is
also observed in the case of pressure-driven production. This can
be seen in Fig. 6, which shows the pressure distribution along a
north/south section of the reservoir caused by a pressure drawdown
in a single well at the center of the reservoir. In this figure, the bottomhole pressure history used in the simulation to drive the production is based on the measured bottomhole pressures in the field.
The prescribed bottomhole pressure decreased almost linearly with
time from 48 MPa at the start of production to about 24 MPa after
14 years of production. The results of two analyses are shown in
this figure: one for a soft reservoir with a Youngs modulus of
E=50 MPa, and another for a stiff reservoir with E=850 MPa
(n=0.25 for both cases). The Youngs modulus of 50 MPa for the
soft reservoir corresponds to the post-yield elastoplastic response
of the reservoir rock, while the value for the stiff reservoir corresponds to the elastic response of the reservoir rock.
For the case of the coupled simulation of the soft reservoir, the
pore pressures have increased to about 52 MPa (which is higher
than the initial reservoir pore pressure of 48 MPa) from a distance
of about 1.7 km from the central producing well, despite the continuous drawdown in the producing well. The pore-pressure distributions obtained from uncoupled reservoir simulations are also
shown in the same figure. In the uncoupled reservoir simulation,
the pore compressibilities used (see Table 1) correspond to a uniaxial strain loading condition and were calculated from Eq. 39 with
the same Youngs modulus and Poissons ratio used in the fully
coupled simulation. Again, this is based on the assumption that the
expected horizontal displacements in the reservoir will be negligible in comparison to the horizontal dimensions of the reservoir. As
such, these values are the best a priori estimates of the pore compressibilities for the uncoupled simulations.
As shown, the predicted pore pressures never exceeded the initial pore pressure of 48 MPa. For the case of the stiff reservoir, no
increase in pore pressure above the initial pressure is observed, as
in the case of the soft reservoir, for both the coupled and uncoupled
simulations. However, there are still significant differences in the
predicted pore-pressure distributions from the two simulations.
The differences in the predicted pore pressures shown in Fig. 6,
in the case of both soft and stiff reservoirs, are caused by the lack
of geomechanical terms and the deficiency of using rock compressibility to account for the geomechanical effects in the reservoir simulation. In the reservoir simulation, the rock deformation
was constrained and assumed to be a priori uniaxial. However, the
stress paths followed by the different points in the reservoir are
169

Reservoir pressure

50
Initial reservoir
pressure

40

30
Fully coupled modeling
Reservoir simulation

0
1000
2000
3000
Distance from production well, m
Initial reservoir
pressure

50
Reservoir pressure

Fully coupled modeling


Reservoir simulation

40

30

1000

2000

3000

Distance from production well, m


Fig. 6Comparisons of reservoir pressures from fully coupled
and standard reservoir simulations. Pressure-controlled production. Top: soft reservoir; bottom: stiff reservoir.

controlled by the interaction between the reservoir and the surrounding nonpay rock, and by the constitutive behavior of both the
reservoir and the nonpay rocks.
The increase in pore pressure above initial value during production is an important effect that cannot be predicted by existing
reservoir simulators. This increase in pore pressure is caused by the
load-term load [L]t[K]-1[DF] from the total stress changes and the
interaction of reservoir and overburden in the fully coupled analysis. The pore-pressure reduction resulted in the compaction of the
central part of the reservoir. In turn, the compaction resulted in the
downward movement and bending of the overburden, causing the
reservoir fluids to be squeezed and the pore pressures to increase
toward the reservoir flanks. This coupled response, which comes
from the interaction between the reservoir and the overburden, is a
complicated process. The deformation of the overburden is
dependent on the pore-pressure distribution in the reservoir; on the
other hand, the pore-pressure distribution in the reservoir is also
controlled by deformation of the overburden. This structural indeterminacy in rock deformation is one of the main reasons why it is
not always easy to decouple rock deformation from fluid flow.22
The increase in pore pressure above initial value during fluid
extraction is analogous to the so-called Mandel-Cryer effect
observed in one of the first applications of Biots 3D consolidation
170

theory. Cryer23 showed that on withdrawal of fluid in a consolidating layer of a fluid-saturated medium, the pore pressure instantaneously jumped, then continued to increase for some time, before
pressure dissipation commenced. The pore-pressure increase is
attributed to the downward movement of the layer above the consolidating layer and the increase in total stresses as the weight of
the layer above the consolidating layer is transferred to the fluids,
causing the pore pressure to increase.
The increase in pore pressure owing to rock deformation has
also been referred to as compaction drive in reservoir engineering.
In standard reservoir simulation, the main mechanism accounting
for the compaction drive is the pore-volume reduction of the
reservoir rock. In fully coupled simulation, the downward movement of the overburden also contributes to the compaction drive.
This contribution, particularly when the pore pressures increase
above the initial reservoir pressure, cannot be accounted for simply by adjusting the pore compressibility in reservoir simulations.
The compaction drive will be very pronounced for soft reservoirs,
but it can also be significant for the case of relative stiff reservoirs,
as shown in Fig. 6. Note that this increase is only to be expected
for reservoirs with low-permeability aquifers. Otherwise, the
increase in reservoir pressure from compaction drive will be dissipated into the aquifer. On the other hand, a compressible aquifer
might also contribute to the increase in pore pressure from the
compaction drive.
Several schemes have been proposed in the literature to couple
the stress-strain behavior of rock and multiphase fluid flow.3,4
Settari and Mourits,24 for instance, present an approach where the
porosity is used as a coupling parameter between a finite-element
stress-analysis code and a reservoir simulator. The geomechanical
and reservoir simulators are used in a staggered manner. Porepressure changes are calculated from the reservoir simulation and
converted to nodal loads. From these nodal loads, the in-situ stress
changes and rock displacements are calculated in the geomechanical simulation. An iterative algorithm is used to ensure that the
porosities calculated from the geomechanical simulator are the
same as those calculated from the reservoir simulator.
The iterative approach, however, does not rigorously address
the coupling of geomechanics and reservoir simulations. One possible drawback of such an iterative approach is that there appears
to be no proof that the approach will converge to a unique solution.
For instance, it is not clear whether the approach can be used in the
case where the rock tends to increase in volume with a reduction in
pore pressure (e.g., owing to dilation during shearing). Such a volume increase would require a negative pore compressibility in the
reservoir simulation and may cause numerical instability.
Conclusions
The issues related to the interaction between fluid flow and rock
deformation in reservoir simulation have been discussed in this
paper. A primary type of interaction concerns stress- induced permeability changes, which in turn affect the fluid-pressure distribution. This type of coupling is particularly important in fractured
and faulted reservoirs, where fracture- and fault-permeability
changes can be orders of magnitude greater than those of the bulk
matrix. Moreover, fracture- and fault-permeability changes can
also influence fluid-flow directionality and sweep efficiency.
A comparison of the governing equations used for reservoir
simulations and Biots theory for multiphase fluid flow in
deformable porous media was made. Based on this comparison and
on the results of a simple case study, it was shown that geomechanics and multiphase fluid flow in hydrocarbon reservoirs
should be analyzed as fully coupled processes. As fully coupled
processes, fluid flow and geomechanics form a set of separate
domains that cannot be analyzed separately. The dependent variables in each domain (e.g., fluid pressures and rock displacements)
cannot be eliminated explicitly, except for simple cases corresponding to specific stress paths. However, in general, the deformation
of the reservoir during depletion and recovery is a complicated
process for which a simple stress path cannot be assumed.
It is shown that reservoir simulators, by not taking these coupling
phenomena into consideration, simplify important geomechanical
June 2001 SPE Reservoir Evaluation & Engineering

aspects that can impact reservoir productivity. This is attributed to the


fact that reservoir simulation does not include the governing equations for geomechanics. Moreover, the only rock mechanical parameter involved in reservoir simulations is pore compressibility. This
parameter is not sufficient in representing aspects of rock behavior
such as stress-path dependency and dilatancy, which require a full
constitutive relation, and the influence of the surrounding nonpay
rock on reservoir deformability. Furthermore, the pore-pressure
changes caused by the applied loads from the nonpay rock cannot be
accounted for by simply adjusting the pore compressibility.
Nomenclature
Bp = volume factor for phase p
cf = fluid compressibility
cf p = fluid compressibility of phase p
cp = pore compressibility
(cp) hydro = pore compressibility under hydrostatic loading
(cp)oedo = pore compressibility under oedometric loading
ct = total compressibility
[C] = compressibility matrix
Dijkl = constitutive tensor
E = Youngs modulus
[DF] = matrix of incremental boundary or self-weight loads
Fi = boundary or self-weight loads
g = acceleration owing to gravity
G = shear modulus
h = fluid height
k = absolute permeability
k ij = permeability tensor
krp = relative permeability for phase p
K = bulk modulus
Kf = fluid modulus
Ks = grain compressive modulus
[K] = stiffness matrix
[L] = coupling matrix
M = constrained modulus
MB = Biots modulus
p = pore pressure
pp = fluid pressure for phase p
Pc = capillary pressure
[Dp] = incremental pore-pressure matrix
qp = fluid source or sink
[Dq] = matrix of incremental fluid sources or sinks
Sp = saturation for phase p
t = time
ui = rock displacement
[Du] = displacement increment matrix
npi = fluid velocity for phase p
xi = spatial coordinate
a = Biots coefficient (1-K/Ks)
dij = Kronecker delta function

eij = strain tensor
en = volumetric strain
[F] = permeability matrix
f = porosity
m = fluid viscosity
mp = viscosity for phase p
r = density
rp = density for phase p

s ij = total stress tensor

sij = effective stress tensor
n = Poissons ratio
Subscripts
i, j,k,l = spatial coordinates x,y, z
o = oil
w = water
June 2001 SPE Reservoir Evaluation & Engineering

z = vertical axis
p = fluid phase (o,w)
Acknowledgment
The assistance of Dr. Hamid Ghafouri in carrying out the fully
coupled modeling of the idealized North Sea reservoir is gratefully acknowledged.
References
1. Advani, S.H. et al.: Fluid Flow and Structural Response Modeling
Associated with the Mechanics of Hydraulic Fracturing, SPEFE (June
1986) 309.
2. Settari, A., Ito, Y., and Jha, K.N.: Coupling of a Fracture Mechanics
Model and a Thermal Reservoir Simulator for Tar Sands, J. Cdn. Pet.
Tech. (November 1992) 31, No. 9, 20.
3. Fung, L.S.K., Buchanan, L., and Wan, R.G.: Coupled GeomechanicalThermal Simulation of Deforming Heavy-Oil Reservoirs, paper 361 presented at the 1992 CIM Annual Technical Conference, Calgary, June 1992.
4. Tortike, W.S. and Ali, S.M.: Reservoir Simulation Integrated with
Geomechanics, paper 391 presented at the 1992 CIM Annual
Technical Conference., Calgary, June 1992.
5. Heffer, K.J. et al.: The Influence of Natural Fractures, Faults and Earth
Stresses on Reservoir PerformanceAnalysis by Numerical
Modelling, Proc., 3rd Intl. Conference on North Sea Oil and Gas
Reservoirs, Trondheim, Norway (December 1992) 201.
6. Koutsabeloulis, N.C., Heffer, K.J., and Wong, S.: Numerical
Geomechanics in Reservoir Engineering, Proc., Intl. Conference on
Computer Methods and Advances in Geomechanics (1994) 2097.
7. Gutierrez, M. and Makurat, A.: Coupled HTM Modelling of Cold
Water Injection in Fractured Hydrocarbon Reservoirs, Intl. J. Rock
Mech. Min. Sci. (1997) 34, 429.
8. Gutierrez, M.: Fully Coupled Analysis of Reservoir Compaction and
Subsidence, paper SPE 28900 presented at the 1994 SPE European
Petroleum Conference, London, 2527 October.
9. Lewis, R.W. and Sukirman, Y.: Finite Element Modelling for
Simulating the Surface Subsidence Above a Compacting Hydrocarbon
Reservoir, Intl. J. Num. Analy. Meth. Geomech. (1993) 18, No. 9, 618.
10. Biot, M.A: General Theory of Three-Dimensional Consolidation, J.
Appl. Phys. (1941) 12, 155.
11. Lewis, R.W. and Sukirman, Y.: Finite Element Modelling of ThreePhase Flow in Deforming Saturated Oil Reservoirs, Intl. J. Num.
Analy. Meth. Geomech. (1993) 17, No. 8, 577.
12. Terzaghi, K.: Erdbaumechanik auf Bodenphysikalicher Grundlage,
Franz Deutike, Vienna (1925).
13. Sornette, D., Davy, P., and Sornette, A.: Structuration of the
Lithosphere in Plate Tectonics as a Self-Organized Critical
Phenomenon, J. Geophys. Res. (1990) 95, No. 11, 353.
14. Chen, H.-Y., Teufel, L.W., and Lee, R.L.: Coupled Fluid Flow and
Geomechanics in Reservoir StudyI. Theory and Governing
Equations, paper SPE 30752 presented at the 1995 SPE Annual
Technical Conference and Exhibition, Dallas, 2225 October.
15. Chen, H.-Y. and Teufel, L.W.: Coupling Fluid-Flow and Geomechanics
in Dual-Porosity Modeling of Naturally Fractured Reservoirs, paper
SPE 38884 presented at the 1997 SPE Annual Technical Conference and
Exhibition, San Antonio, Texas, 58 October.
16. Ghafouri, H.R.: Finite element modelling of multi-phase flow through
deformable fractured porous media, PhD thesis, U. of Wales, Swansea,
Wales (1996).
17. Coussy, O.: A General Theory of Thermoporo-elastoplasticity.
Transport in Porous Media (1989) 4, 281293.
18. Charlez, Ph.: Rock Mechanics, Vol. 2 Petroleum Applications, Technip
Ed., Paris (1997).
19. Boutca, M.: Elements of Poro-elasticity for Reservoir Engineering,
Revue de lInstitut Franais du Ptrole, (1992) 47, No. 4, 479.
20. Morita, N. et al.: A Quick Method To Determine Subsidence,
Reservoir Compaction, and In-Situ Stress Induced by Reservoir
Depletion, JPT (January 1989) 71.
21. Zienkiewicz, O.C.: Coupled Problems and Their Numerical Solutions,
Numerical Methods in Coupled Systems, Wiley, New York City (1984) 35.
22. Lewis, R.W., Schrefler, B.A., and Simoni, L.: Coupling vs.
Uncoupling in Soil Consolidation, Intl. J. Num. and Anal. Meth.
Geomech. (1991) 15, No. 8, 533.
171

23. Cryer, C.W.: A Comparison of the Three-Dimensional Consolidation


Theories of Biot and Terzaghi, Q.J. Mech. Appl. Math. (1963) 16, 401.
24. Settari, A. and Mourits, F.M.: Coupling of Geomechanics and
Reservoir Simulation Models, Proc., Intl. Conference on Computer
Methods and Advances in Geomechanics, Morgantown, West Virginia
(1994) 2151.

SI Metric
bar
ft
mile

Conversion Factors
1.0*
E + 05 = Pa
3.048*
E - 01 = m
1.609 344*
E + 00 = km

*Conversion factor is exact.

172

SPEREE

M. Gutierrez is Associate Professor in the Dept. of Civil and


Environmental Engineering at the Virginia Polytechnic Inst.
and State U. in Blacksburg, Virginia. e-mail: magutier@
vt.edu. He was formerly a senior engineer responsible for
reservoir mechanics at the Norwegian Geotechnical Inst. in
Oslo. Gutierrez holds a PhD degree in civil engineering from
the U. of Tokyo. R.W. Lewis is Professor and Head of
Mechanical Engineering at the U. of Wales, Swansea, U.K.
An editor of three international journals on numerical modeling, he has cowritten a book on the finite-element
method in the deformation and consolidation of porous
media and has written or collaborated on more than 250
papers. Lewis holds PhD and DSc degrees from the U. of Wales.
I. Masters is Lecturer in the Dept. of Mechanical Engineering at
the U. of Wales, Swansea. He holds a PhD degree from the U.
of Wales.

June 2001 SPE Reservoir Evaluation & Engineering

You might also like