Anda di halaman 1dari 10

ARTICLE IN PRESS

International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778

Contents lists available at ScienceDirect

International Journal of
Rock Mechanics & Mining Sciences
journal homepage: www.elsevier.com/locate/ijrmms

Finite element analysis of coupled chemo-poro-thermo-mechanical effects


around a wellbore in swelling shale
Xiaoxian Zhou, Ahmad Ghassemi 
Department of Petroleum Engineering, 3116 TAMU-401E Richardson Building, Texas A&M University, College Station, TX 77843, USA

a r t i c l e in fo abstract

Article history: Maintaining wellbore stability is a critical problem when drilling for petroleum production in high
Received 7 July 2008 pressure and temperature shales that tend to swell and deteriorate in the presence of the water phase of
Received in revised form a drilling fluid. In this paper, a finite element model is developed for coupled linear and non-linear
20 November 2008
chemo-poro-thermoelasticity. The model is first verified through comparisons with the analytical
Accepted 25 November 2008
Available online 15 January 2009
results for the cases of linear poroelasticity, poro-thermoelasticity, and chemo-poroelasticity. Then, the
non-linear finite element model is used to analyze the problem of a wellbore in a chemo-poro-
Keywords: thermoelastic formation with non-hydrostatic in situ stress field. The results indicate that the coupling
Chemo-poro-thermoelasticity between temperature and salinity has an important impact on wellbore stability, namely heating and a
Chemical potential
lower solute mass fraction in mud increase pore pressures, the radial and tangential total stresses in the
Chemical osmosis
formation, and vice versa. Finally, the prediction of the non-linear chemo-poroelastic model are
Shale stability
Wellbore failure compared to those obtained using a linear model. Results show that for mud properties within the range
of interest, the linear model provides good results for the purpose of assessing shale instability.
& 2008 Elsevier Ltd. All rights reserved.

1. Introduction consider hydration swelling of water-absorbing rocks. This model


has the advantage that it considers the osmosis, swelling, and
The physico-chemical interactions between the mud and the solute transport between the water phase of the drilling mud and
formation significantly influence the resulting stress and pore the pore fluid on the basis of thermodynamic laws and the
pressure fields in swelling shales. It is thus necessary to consider Gibbs–Duhem equation. On the other hand, because of its coupled
their influence when conducting a rigorous wellbore stability and non-linear nature, a numerical solution is needed making it
analysis, particularly when drilling in high-pressure, high-tem- cumbersome for implementation in wellbore stability models.
perature (HPHT), environments, where the temperature exceeds Ghassemi and Diek [13] developed a linearized version; this
150 1C and reservoir pressure exceeds 68.90 MPa (see e.g., [1]). The chemo-poroelasticity theory for swelling shales in which the
effect of chemo-mechanical process on shale deterioration and chemical potential is expressed as a linear function of the solute
borehole instability under isothermal conditions has been mass fraction. The theory is able to capture the important
extensively studied [2–6]. It has been found that hydraulic fluid phenomena observed in the laboratory and the filed, and allows
transport is often several times lower than the fluid transport analytical treatment of the field equations for a wellbore problem.
induced by chemical and thermal gradients. Certain phenomena Thereafter, Ghassemi et al. [14] presented a linear chemo-poro-
related to thermal processes in shale and other rocks have also thermoelastic model that accounts for both the temperature and
been investigated using poro-thermoelastic theory [7,8]. The chemical potentials based on the theoretical development of Diek
significance of thermal stress and pore pressure around a wellbore and Ghassemi [15]. This is of particular need when drilling in
and their impact on borehole stability has been shown in Wang HPHT shale formations. The coupled non-linear chemo-poro-
and Papamichos [9] and Li et al. [10]. Ghassemi and Diek [11] thermoelsatic model satisfies the laws of thermodynamics and
indicated that under certain condition, thermo-osmosis become has been developed within the framework of Biot poroelasticity.
several times larger than hydraulic flow in shale and should be In this paper, a finite element method is developed for solving
considered. Based on non-equilibrium thermodynamics, Heidug the field equations of chemo-poro-thermoelasticity. The numer-
and Wong [12] developed a fully coupled Biot-like model to ical model is first verified through comparisons with the
analytical results for poroelasticity, poro-thermoelasticity, and
chemo-poroelasticity. Thereafter, the fully coupled chemo-poro-
 Corresponding author. Tel.: +1 979 845 2206; fax: +1 979 862 6579. thermoelastic finite element model is used to solve an example
E-mail address: ahmad.ghassemi@pe.tamu.edu (A. Ghassemi). problem of a wellbore in shale to investigate the interaction

1365-1609/$ - see front matter & 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijrmms.2008.11.009
ARTICLE IN PRESS

770 XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778

between the three processes. And finally, the results of the linear where sij and eij are the components of the total stress and strain
chemo-poroelastic model are compared to those of the non-linear tensors, CS and T are solute mass fraction and temperature,
version to assess the accuracy and the range of applicability of the respectively. The other coefficients appearing in Eqs. (2) and (3)
more computationally efficient linear model. are now defined as follows:
!
oD a  f f oD ða  1Þ
a0 ¼ a  D
; b¼ þ þ (4)
2. Chemo-poro-thermoelastic model r̄f C Ks Kf r̄f KC D
 
The fundamental rock constitutive equations for total stresses oS oD RT 0 a1
and the variation of fluid content are constructed from the
w¼ S
 D
; w0 ¼ w (5)
C C MS K
potential energy density of the wetted clay matrix. Here we shall
only provide a brief overview of the governing equations of the s 0 oD s0 oD ða  1Þ
g1 ¼ K am þ ; g2 ¼ aam þ ðaf  am Þf þ (6)
non-linear chemo-poro-thermoelastic model. C D
KC D
where a is Biot’s coefficient; CS and CD the solute and diluent mass
2.1. Constitutive equations fraction, respectively; r̄f the fluid mass density, f the porosity; Ks
and Kf the modulus of bulk solid matrix and fluid, respectively; am
The shale is assumed to be linearly elastic material, saturated and af the thermal expansion coefficients of solid and fluid,
with a binary electrolyte solution, and swells/shrinks in response respectively; s0 the reference value of the specific fluid entropy at
to variation in the chemical potential of the diluent and solute the average system temperature and solute mass fraction. The
components. The fundamental constitutive relations for the above non-linear system is based on the notion that the chemical
temporal evolution of stress and pore volume fraction are derived potential of the solute is given by mS ¼ RT/MS ln CS, in which MS is
from the free energy density of the wetted mineral matrix the molar mass of the solute, R the universal gas constant and T
containing bound fluid. The free energy density is defined such the absolute temperature. In addition, it is assumed that the
that the differential of the potential represents the work swelling parameters associated with the diluent and the solute
associated with the elastic deformation and the work done to are equal [12], i.e.,
alter the masses of all fluid components, and the rate of change of
entropy per unit referential volume in an infinitesimal macro- MS
oS ¼ oD ¼ o0 (7)
scopic transformation [12]. RT
The potential energy density of the wetted clay matrix is where o0 is the swelling coefficient. Therefore, the coefficients in
obtained from the difference between the internal energy of the the constitutive equations simplify to
saturated rock and that of the fluid. For the case of isotropy, the !
potential is [15] M S o0 a  f f o0 ða  1ÞMS
a0 ¼ a  D
; b¼ þ þ (8)
0 1 r̄f RTC Ks Kf r̄f KRC D T
  X
4G 2 !
2W ¼ K þ I þ GI2  2I1 @ ap  ob mb A  Qp2
3 1 o0 CS a1
0 1
b w¼ 1 ; w0 ¼ w (9)
CS CD K
X b X b
 2p@ B m þ XT A  2T
b
Y mb
b b s 0 o0 M S
X ab g1 ¼ K am þ ,
 2T Z mb  YT 2  2T Wbound (1) RTC D
b s0 o0 ða  1ÞM S
g2 ¼ aam þ ðaf  am Þf þ (10)
where T is temperature, p is the pore pressure and mb is the KRTC D
chemical potential of the b component (solute or diluent). Wbound is The constitutive equations can be linearized with respect to the
the entropy of the bound water; K and G are the bulk modulus and chemical terms to provide a linear chemo-poro-thermoelastic
shear modulus of porous media, respectively; the other thermo- model. We do this by re-writing the chemical potential of the
dynamic response functions Yb, Zba, B, Q, X, a, ob and Y are solute as
defined by: Y b ¼ ðqmbbound =qTÞeij ;mb ;T ; Z ba ¼ ðqmbbound =qma Þeij ;p;T ; RT
mS  ðaS C S  bÞ for all C S 2 ½C min ; C max  and aS ; b40 (11)
B¼ ðqz=qmb Þ eij ;p;T ; Q ¼ ðqz=qpÞeij ;mb ;T ; N ¼ ðqz=qTÞeij ;mb ;p , a¼ MS
S S
ðqs=qpÞ;mb ;T ; ob ¼ ðqsij =qmb Þeij ;p;T ; Y ¼ ðqWsolid =qTÞeij ;mb ;T . I1, I2 are where aS  ðC̄ Þ1 such that C̄ is the mean value of the solute
the first and second invariants of the strain tensor, sij. e is the mass fraction and b is a reference constant. For the details of the
volumetric strain, s is the mean stress, and z is the variation of models the reader is referred to [13–15].
fluid volume per unit reference pore volume.
The stress components and the variation in fluid content are 2.2. Transport equations
obtained by differentiating W with respect to strains and pore
pressure, respectively. Then, differentiation with respect to The equations expressing the fluxes of fluid, solute, and heat in
time yields the constitutive equations for time evolution of the terms of their respective driving forces are derived from the
above variables. dissipation function through thermodynamic arguments of irre-
Following the above steps, the general constitutive equations versible processes. In addition to direct flows, the theory
for the non-linear chemo-poro-thermoelastic model can be incorporates indirect flows of diluent and solute by chemical
written in the following from (compression is considered positive): osmosis and thermal diffusion, respectively. Thus, temperature is
  coupled to ion transfer, pore pressure, and formation stresses.
2G S
s_ ij ¼ 2G_ ij þ K  _ kk dij þ a0 p_ dij  wC_ dij þ g1 T_ dij (2) It is assumed that pore fluid flux is due to hydraulic pressure
3
and chemical osmosis, solute flux is due to the diffusion and
thermal filtration, and heat transfer is due only to the gradient of
S
z_ ¼ a_ ij þ bp_ þ w0 C_  g2 T_ (3) temperature. Therefore, the pore fluid, solute and heat transport
ARTICLE IN PRESS

XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778 771

equations could be described as order interpolation of the displacement can ensure the stability of
  the pore pressure, solute mass fraction and temperature in the
k <r̄ RT
J f ¼ r̄f rp  S f S D rC S (12) spatial domain [19]. Also, this kind of elements can be used for
Z M C C material non-linear problems if the reduced/selective integration
is adopted [20]. The following approximations are used for the
J S ¼ r̄f DS rC S  r̄f DT rT (13)
fields of u, p, CS and T:
T
J T ¼ k rT (14) S
u ¼ Nu ũ; p ¼ Np p̃; C S ¼ NC S C̃ ; T ¼ NT T̃ (19)
where k is the permeability, Z is the fluid viscosity, < is the
reflection coefficient, DS is the solute diffusion coefficient, where Nu, Np, NC S and NT are the shape functions for the solid
diffusivity, DT is the thermal diffusion coefficient and kT is the displacement, pore pressure, solute mass fraction and tempera-
S
thermal conductivity. The reflection coefficient can be described ture fields, respectively, and ũ, p̃, C̃ and T̃ are the vectors of the
as the ratio of the actual osmotic pressure to its theoretical value nodal displacements, nodal pore pressures, nodal solute mass
which is calculated based on the assumption of a perfect ion fractions and nodal temperatures, respectively.
exclusion membrane. It is a phenomenological constant to Substituting Eq. (19) into Eqs. (15)–(18), and then using the
consider the influence of ion transfer on the osmotic pressure Galerkin technique, we could obtain
[16]. Note that for the purpose of focusing on ion transfer effects,
the thermal osmosis is neglected. This issue has previously been _ þ Ap̃_  WC̃_ þ VT̃_ ¼ f_
Kũ (20)
studied in [11].

S
2.3. Field equations AT ũ ^ C̃_ þ N
_ þ Sp̃_ þ M ^ T̃_ þ HH p̃ þ DH C̃S ¼ 0 (21)

The above constitutive and transport equations can be


combined with the momentum, fluid mass, solute mass, and _S S
MC̃ þ DD C̃ þ Q D T̃ ¼ 0 (22)
energy balance equations, to yield the following field equations:
 
G
Kþ rðr  uÞ þ Gr2 u þ mða0 rp  wrC S þ g1 rTÞ ¼ 0 (15) _
RT̃ þ UT̃ ¼ 0 (23)
3

S k k where
_ þ bp_ þ w0 C_ þ g2 T_  r2 p þ LD r2 C S ¼ 0
aðr  uÞ (16)
Z Z Z Z
K¼ BT DB dV; A¼ BT a0 mNp dV,
fC_ S  DS r2 C S  C S DT r2 T ¼ 0 (17) Z
Ve Ve

W¼ BT wmNC S dV (24)
T_  cT r2 T ¼ 0 (18) Ve

in which u is the displacement vector, m ¼ [1,1,0]T for plane strain Z Z


problems and m ¼ [1,1,1,0,0,0]T for three-dimensional problems, V¼ BT g1 mNT dV; S¼ NTp bNp dV,
LD ¼ <r̄f RT=M S C S C D , and the other notations are the same as Ve Ve
Z
those previously defined. ^ ¼
M NTp w0 NC S dV (25)
Eqs. (15)–(18) are non-linear since the coefficients include the Ve
unknowns p, CS (or CD) and T. Therefore, a numerical method
should be adopted to solve the above field equations subjected to Z Z
boundary and initial conditions. Instead of the unknown variables ^ ¼
N NTp g2 NT dV; M¼ NTC S fNC S dV (26)
CS(t), CD(t) and T(t), where t is the elapsed time, Ghassemi and Ve Ve

Diek [13] and Ghassemi et al. [14] used the mean values of the
S D Z Z
solute mass fraction C̄ , diluent mass fractions C̄ and tempera-
ture T̄ in the mud-shale system, to make the system amenable to R¼ NTT NT dV; U¼ ðrNT ÞT cT ðrNT Þ dV (27)
Ve Ve
analytical solution.
It is worth noting that the above field equations have been
Z
developed by extending the work of Heidug and Wong [12] to
HH ¼ ðrNp ÞT ðk=ZÞðrNp Þ dV,
include temperature effects. Thus, our work is based on Ve
Z
irreversible thermodynamic which ensures consistency with the
second law of thermodynamics. This feature sets our work apart DH ¼ ðrNp ÞT LD ðrNp Þ dV (28)
Ve
from other works such as Ekbote and Abousleiman [17,18].
Another distinction is that our expression for the variation of Z
the fluid components does conform to the Gibbs–Duhem relation. DD ¼ ðrNC S ÞT DS ðrNC S Þ dV,
Ve
Z
3. Finite element method for chemo-poro-thermoelastic model QD ¼ ðrNT ÞT ðC S DS ÞðrNT Þ dV (29)
Ve

In the following, the finite element method for the linear plane In the above expressions, Ve is the spatial area of an
strain chemo-thermo-poroelastic problems is presented. The non- element, f is the external applied loads, B is the strain-
linear formulation is similar and will not be shown for brevity. We displacement matrix, and the other notations are the same as
adopt eight-node quadrilateral elements for the solid displace- those defined before.
ment u, four-node quadrilateral elements for the pore fluid In this work, the Crank–Nicolson type of approximation [20] is
pressure p, solute mass fraction CS and temperature T. The higher used to discretize the temporal domain. Finally, we obtain the
ARTICLE IN PRESS

772 XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778

following finite element formula: 4. Verification of finite element method


2 38 9
K A W V > Dũ > To verify the finite element model, the finite element results
>
> >
>
6 T 7> >
6A ðS þ y DtHH Þ ^ þ y DtDH Þ
ðM N^ 7>< Dp̃ >= are compared with the analytical solutions for the poroelastic,
6 7
6 7 S poro-thermoelastic and chemo-poroelastic wellbore problems.
6 0 0 ðM þ y DtDD Þ y DtQ D 7 > DC̃ >
4 5>>
>
>
>
> The problem is analyzed by assuming plane strain conditions and
> >
0 0 0 ðR þ y DtUÞ : DT̃ ; ‘‘instantaneous’’ drilling of the wellbore. The drilling of the
8 9 borehole is simulated by removing the stresses acting on the wall
>
> Df >
>
>
> >
> of the borehole at t ¼ 0. Only a quarter of the borehole is
>
> S >
< DtHH p̃ðt n1 Þ þ DtDH C̃ ðt n1 Þ >
= considered in the finite element model because of the symmetry
¼ S (30)
>
> DtDD C̃ ðtn1 Þ þ DtQ D T̃ðtn1 Þ >
> of the problem. Fig. 1 shows the finite element mesh, which
>
> >
>
>
> >
> consists of 500 quadrilateral eight-noded elements and 1591
: DtUT̃ðtn1 Þ ;
element nodes. The shale parameters used in the paper are
presented in Table 1.
where y is a scalar parameter which can vary between 0.5 and 1.0, We assume that before drilling, the field stress in the x- and y-
Df is the change of the external loads between successive steps, directions, pore pressure, formation temperature and solute mass
tn1 and tn denotes the time of the last step and the current step,
respectively, and Dt ¼ tntn1. It should be noted that the above
element ‘‘stiffness’’ matrix is unsymmetrical. Therefore, the
unsymmetrical solver provided in Smith and Griffiths [20]
is adopted in this work to solve the unsymmetrical linear
equation system.
In the equation system (30), Dp, DCS and DT are the primary
unknowns we need to solve. The system is nonlinear because
p(tn1)+Dp, CS(tn1)+DCS and T(tn1)+DT are included in the
‘‘stiffness’’ matrix so that they should be known in order to
evaluate the ‘‘stiffness’’ matrix. Therefore, an iterative method
should be used, in which the values of the unknowns from a
previous time step are used to calculate the initial ‘‘stiffness’’
matrix in current time step to solve the non-linear equation
50a

system. In any time step of the numerical procedure, initially we


assume Dp, DCS and DT to be zero to evaluate the ‘‘stiffness’’
matrix and then solve the linear equation system to obtain new
Dp, DCS and DT. Then, we use the new Dp, DCS and DT to evaluate
the ‘‘stiffness’’ matrix and solve for the next Dp, DCS and DT. The
iteration continues until Dp, DCS and DT are close enough to those
obtained in the last iteration within a prescribed tolerance, i.e.,

S S
jjDp̃  Dp̃old jj jjT̃  DT̃old jj jjDC̃  DC̃old jj
o; o and S
o
jjDp̃old jj jjDT̃old jj jjDC̃old jj
(31) a
50a
where Dp̃old is the pore pressure increment in the last iteration,
Dp̃ is the pore pressure increment in the current iteration, the Fig. 1. Finite element mesh used in the computations, where a is the radius of the
other notations have similar meanings. We set e ¼ 0.2% in borehole.

this work.
On the other hand, for the linear models proposed by Ghassemi
and Diek [13] and Ghassemi et al. [14], the ‘‘stiffness’’ matrix is Table 1
Properties of the shale used in the computations.
constant during the computation process by using the mean
values of p, CS and T of the mud-shale system. As a result, no Young’s modulus, E 1853 MPa
iteration is required. Drained Poisson’s ratio, n 0.219
S
In the finite element method, ũ, p̃, C̃ and T̃ are primary Undrained Poisson’s ratio, nu 0.461
Skempton’s coefficient, B 0.915
unknowns and they could be obtained directly by solving the
Permeability, k 1.0  109 darcy
equation system. The variations of the effective stresses at Gauss Porosity, f 0.2989
integral points depend on the deformation of rock skeleton and Fluid mass density, r̄f 1111.11 kg/m3
can be expressed as Fluid bulk modulus, Kf 3291 MPa
Fluid viscosity, Z 3.0  104 Pa s
  Reflection coefficient, < 0.2
2G
s_ 0ij ¼ 2G_ ij þ K  _ kk dij (32) Molar mass of solute (NaCl), Ms 0.0585 kg/mol
3
Swelling coefficient, o0 1.5 MPa
Solute diffusion coefficient, DS 2.0  109 m2/s
where the notations in the right hand side are the same as those Pressure diffusion coefficient, L 0
in Eq. (2). Eq. (2) clearly shows the components of the total Thermal expansion coefficient of solid, am 1.8  105 K1
stresses and can be used to obtain the total stresses at Gauss Thermal expansion coefficient of fluid, af 3.0  104 K1
Thermal diffusivity, cT 1.6  106 K1
integral points. The total and effective stresses at other points may
Thermal diffusion coefficient, DT 6.0  1012 m2/(s K)
be obtained by interpolation or extrapolation according to their Specific entropy (NaCl, CS ¼ 0.15), s0 3686 J/(kg K)
values at Gauss integral points.
ARTICLE IN PRESS

XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778 773

fraction are sx0, sy0, p0, T0 and C S0 . Then, upon instantaneous 2


drilling of the well, the wellbore mud pressure, temperature and Solid Curves: analytical solutions
solute mass fraction are pm, Tm and C Sm , respectively and remain Numerical solutions
constant with time. The shear stress component is zero at the wall t* = 10-4
1.5
and the radial stress equals the mud pressure. These conditions t* = 10-3
for use in the FEM are presented in Table 2. t* = 10-2
t* = 10-1
t* = 100

p/So
1
4.1. Poroelastic model

Usually the loading in the poroelastic model is decom-


posed into three fundamental modes: (1) a far-field isotropic 0.5
stress; (2) a virgin pore pressure; (3) a far-field stress deviator.
Here the finite element model is used for the modes 2 and 3
[21,22], and because of space limitation we only show some of 0
the results. 1 1.5 2 2.5
Fig. 2 shows the variations of the tangential total stresses with r/a
radius for pore pressure loading of the wellbore wall, where
Fig. 3. Comparison of the pore pressure variations with radius at y ¼ 0 or p for
Z ¼ a(1v)/2(1+v) and the other notations can be found in Table
mode 3 from numerical and analytical methods.
1. The figure shows that tensile tangential stresses are generated
near the borehole wall due to pore pressure drainage. Figs. 3 and 4
show the variations of the induced pore pressure and tangential
total stress with distance from the borehole for mode 3. From the 0
Solid Curves: analytical solutions
figures, it is found that all the numerical results agree well with
Numerical solutions
the analytical results.
t* = 10-4
-1
t* = 10-3
4.2. Thermo-poroelastic model t* = 10-2
t* = 10-1
t* = 100
σθθ/So

In the thermo-poroelastic model, the initial temperature of the -2


formation is Tsh ¼ 65 1C, and the wellbore wall is suddenly heated
and then maintained at Tm ¼ 115 1C. Figs. 5 and 6 illustrate the
profiles of the induced radial and tangential total stresses around
the wellbore. Analytical solutions are also shown for comparison. -3

Table 2 -4
Problem boundary conditions for FEM. 1 1.1 1.2 1.3 1.4 1.5
r/a
Wellbore wall Infinity or far distance
Fig. 4. Comparison of the tangential stress variations with radius at y ¼ 0 or p for
Force/displacement tx ¼ Pm cos y; Assume displacements in both x mode 3 from numerical and analytical methods.
ty ¼ Pm sin y and y directions are zero
Pore pressure p ¼ pmud p ¼ p0
Temperature T ¼ Tmud T ¼ T0
It can be seen that heating induces compressive radial stresses
Solute mass fraction C S ¼ C Sm C S ¼ C S0
and much more significant compressive tangential stress around
the wellbore. As can be observed from the figures, the numerical
results are in good agreement with the analytic solutions.

2
Solid Curves: analytical solutions 4.3. Chemo-poroelastic model
Numerical solutions
t* = 0.01
1.5 t* = 0.1 The initial solute mass fraction of the formation C Ssh is assumed
t* = 1
t* = 10 to be 0.1, and the solute mass fraction of the mud C Sm is maintained
t* = 1000 at 0.2. Only the chemically induced pore pressure and stresses are
1
considered. As the analytical solutions are available only for the
σθθ/ηpo

linear problem for which the mean values of p, CS and T of the


0.5 mud-shale system are used to evaluate the ‘‘stiffness’’ matrix in
Eq. (30), here we use the linear finite element model. Fig. 7 shows
the induced pore pressure distributions around the borehole
0 corresponding to various times. The induced pore pressure is zero
on the wall of the wellbore and negative inside the formation
(osmotic effect). The peak value decreases with time and moves
-0.5
into the formation. Fig. 8 illustrates that drilling with a higher
1 1.5 2 2.5 3
salinity mud induces a tensile total tangential stress at and in the
r/a vicinity of the wall, and the maximum values occurs inside
Fig. 2. Comparison of analytical and numerical results for tangential stress the formation. The induced radial stress also grows from zero at
variations with radius due to pore pressure loading of the well. the well to become tensile in the formation. The analytical
ARTICLE IN PRESS

774 XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778

4 0.5
Solid Curves: analytical solutions
Numerical solutions

Tangential Total Stress (MPa)


1 hour
0
Radial Total Stress (MPa)

3 24 hours
120 hours

2 -0.5
Solid Curves: analytical solutions
Numerical solutions
1 hour
-1 12 hours
1 24 hours
120 hours

-1.5
0
1 2 3 4 5 6
1 2 3 4 5
r/a
r/a
Fig. 8. Comparison of the tangential total stress variations with radius due to the
Fig. 5. Comparison of the radial total stress variations with radius caused by chemical loading using the numerical and analytical methods.
temperature loading using numerical and analytical methods.

Note that the induced total stresses extend beyond the zone of
6 solute and pore pressure changes. This is observed for chemo-
Solid Curves: analytical solutions poroelasticity as well as in thermo-poroelasticity. The boundaries
5 Numerical solutions of the radial and tangential stress perturbations do not coincide
1 hour with those of temperature and pore pressure perturbations.
Tangential Total Stress (MPa)

24 hours The induced stresses in the zone of pore pressure and temperature
4
120 hours change decay at a different rate than the pore pressure/
3 temperature. In fact, the strain compatibility requirement
causes the induced tangential stress to change sign away from
2 the wellbore wall.

1
5. Finite element analyses of fully coupled chemo-poro-
0 thermoelastic wellbore problem

-1 In this section, we will use the non-linear chemo-poro-


1 2 3 4 5 thermoelastic finite element model to analyze a fully coupled
r/a wellbore stability problem. We adopt the shale properties
presented in Table 1 and the finite element mesh in Fig. 1. We
Fig. 6. Comparison of the tangential total stress variations with radius due to
assume that the wellbore is subjected to a non-hydrostatic
temperature loading using numerical and analytical methods.
horizontal in situ stress field: 20 MPa in the x-direction (y ¼ 0
and p) and 30 MPa in the y-direction (y ¼ p/2 and 3p/2). The mud
0 pressure after the drilling is 15 MPa. We assume Tm ¼ 115 1C and
Tsh ¼ 65 1C for the heating case, and Tm ¼ 65 1C and Tsh ¼ 115 1C for
the cooling case. The initial solute mass fractions of the mud ðC Sm Þ
-0.5 and shale ðC Ssh Þ are assumed to be 0.2 and 0.1 for the case with
Pore Pressures (MPa)

C Sm 4C Ssh , and 0.1 and 0.2 for the case with C Sm oC Ssh . The results
at times of 102, 104 and 106 s, and in the directions of x and y
are presented.
-1

Solid Curves: analytical solutions


Numerical solutions 5.1. Pore pressure distributions
-1.5 1 hour
12 hours The pore pressure redistributions at and in the vicinity of the
24 hours wall due to various combinations of thermal, chemical and
120 hours
mechanical boundary conditions are shown in Figs. 9 and 10. It
-2
is found that heating and C Sm oC Ssh increase the pore pressure
1 1.5 2 2.5 3
while C Sm 4C Ssh (as well as cooling) decrease the pore pressure in
r/a
the formation. The effect of the chemical loading on the pore
Fig. 7. Comparison of the pore pressure variations with radius caused by chemical pressure is more pronounced than the effect of thermal loading.
loading using numerical and analytical methods. The difference of the pore pressures near the wellbore wall is large
at a short time after drilling, but will decrease with the elapse of
the time. The effect of the mechanical loading on the pore
solutions provided in Ghassemi et al. [14] are also presented in the pressure in the formation appears rather fast, while the effects of
figures for comparison. We find that the numerical and analytical thermal and chemical appear relatively slow as the temperature
results are in good agreement. conduction and solute transportation from the mud to the
ARTICLE IN PRESS

XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778 775

25 55
Solid Curves:CmS>CshS 102s 102s
104s Solid Curves: CmS>CshS
Dashed Curves:CmS<CshS 104s

Tangential Total Stress (MPa)


106s 50 Dashed Curves: CmS<CshS
106s
20
Pore Pressure (MPa)

45
15
40

10
35

5 30
1 1.1 1.2 1.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
r/a r/a
Fig. 9. Influence of heating and salinity on pore pressures around the borehole in Fig. 11. Influence of heating and salinity on tangential total stress around the
the direction of y ¼ 0. borehole in the direction of y ¼ 0.

25 45
S>CshS 102s 102s
Solid Curves: Cm
104s Solid Curves: CmS>CshS
104s
Dashed Curves: CmS<CshS Dashed Curves: CmS<CshS
Tangential Total Stress (MPa)
106s 106s
20
Pore Pressure (MPa)

40

15

35
10

5 30
1 1.1 1.2 1.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
r/a r/a

Fig. 10. Influence of heating and salinity on pore pressures around the borehole in Fig. 12. Influence of cooling and salinity on tangential total stress around the
the direction of y ¼ p/2. borehole in the direction of y ¼ 0.

formation take some times. From the figures, we can also observe shale/mud properties would need to be considered. This moti-
the significant influence of the initial non-hydrostatic stress field vated Ghassemi and his co-workers [13–15] to use a linearized
on the pore pressure in the formation. Under the conditions of version of the model to analytically calculate the wellbore
heating the highest pore pressure in the formation appears in the stresses. In the linear model, the chemical potential is linearized
x-direction and C Sm oC Ssh , while the lowest pore pressure appears and the coefficients in Eqs. (15)–(18) are calculated using the
in y direction for C Sm 4C Ssh. average values of the temperature and salinity in the mud-
formation system (instead of the spatial and temporal dependent
5.2. Distribution of tangential total stress unknown temperature and salinity values).
The impact of heating/cooling on the solute mass fraction in
The redistributions of the tangential total stress in the the non-linear and linear numerical models is shown in Fig. 13
formation around the wellbore are shown in Figs. 11 and 12. (using C Sm ¼ 0:2 and C Ssh ¼ 0:1 and the properties of the shale in
Heating and C Sm oC Ssh increase the tangential total stress; while Table 1, along with mesh in Fig. 2). It is found that heating will
cooling and C Sm 4C Ssh decrease the tangential total stress. The expedite the solute mass transport process while cooling slows
maximum total tangential stress at the wellbore wall appears at the process. This phenomenon is only slightly affected by the
the instant of drilling in the x direction under the conditions of linearization. The figure shows that the magnitude of the thermal
heating and C Sm oC Ssh . diffusion coefficient DT affects the results significantly.
Next, we consider the effect of pressure diffusion which was
neglected in all the previous cases. The problem is assumed to be
6. Comparisons between linear and non-linear theory isothermal with Tm ¼ Tsh ¼ 90 1C and the initial in situ stresses and
mud pressures are assumed to be zero. The variations of the
In the previous section, the borehole problem in shale was stresses in the shale are induced only by the chemical loading
numerically studied using a fully coupled non-linear model. A DC S ¼ C Sm  C Ssh . The properties of the shale in Table 1, and the
linear model is less computationally intensive for wellbore finite element mesh in Fig. 1 are still used. Fig. 14 shows the
stability analysis where a range of well orientations, time, and pressure diffusion effect on the distributions of the pore pressure
ARTICLE IN PRESS

776 XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778

0.104 0
Time: 1 hour
Heating, DT = 6.0E-12 m2/(s.K)
0.103 Heating, DT = 2.4E-11 m2/(s.K)
Cooling, DT = 6.0E-12 m2/(s.K)
Solute Mass Fraction

-0.5
Cooling, DT = 2.4E-11 m2/(s.K)

Pore Pressure (MPa)


Results from linear model
0.102

-1
0.101

0.1 -1.5 1 hour


12 hours
Solid Curves: Linear model results
24 hours
Dashed Curves: Non-linear model results 120 hours
0.099
1.15 1.2 1.25 1.3 -2
1 1.5 2 2.5 3
r/a
r/a
Fig. 13. Impact of heating/cooling on the distributions of the solute mass fraction.
Fig. 15. Distribution of induced pore pressure by chemical loading ðC Ssh ¼ 0:1; C Sm ¼
0:2Þ under thermal isothermal condition (Tsh ¼ Tm ¼ 90 1C).
0

-0.4
Pore Pressure (MPa)

-0.1
Radial Total Stress (MPa)

-0.8
-0.2
Solid Curves: L = 0
L = 10-12m2/s
1 hour -0.3
-1.2 12 hours
24 hours
120 hours 1 hour
-0.4 12 hours
-1.6 Solid Curves: Linear model results 24 hours
1 1.5 2 2.5 3 Dashed Curves: Non-linear model results 120 hours
r/a -0.5
1 1.5 2 2.5 3
Fig. 14. Effect of pressure diffusion on the distributions of the induced pore r/a
pressure.
Fig. 16. Distribution of induced total radial stress by chemical loading ðC Ssh ¼
0:1; C Sm ¼ 0:2Þ under thermal isothermal condition (Tsh ¼ Tm ¼ 90 1C).
in the formation (assuming that the pressure diffusion coefficient
L ¼ 1012 m2/s). As expected, the pressure diffusion increases the
0.5
solute transport process and decrease the magnitude of the
induced osmotic pore pressure.
To examine the impact of chemical non-linearity we use
Tangential Total Stress (MPa)

the previous case but without considering the pressure diffusion 0


effect, and compare the prediction of the non-linear and
linear numerical model with respect to the chemo-poroelastic
phenomena. -0.5

6.1. Case with relatively small DC S


-1 1 hour
We let C Ssh ¼ 0:1 and C Sm ¼ 0:2. The distributions of the pore Solid Curves: Linear model results
12 hours
pressures along the radial direction of the wellbore are shown in 24 hours
Dashed Curves: Non-linear model results 120 hours
Fig. 15. Both the results from the linear model and the non-linear
model are presented at t ¼ 1, 12, 24 and 120 h after the chemical -1.5
1 1.5 2 2.5 3
loading is applied. It is found that the magnitudes of the linear
results are lower than those of the non-linear results by about r/a
15%, and the peak of the linear results propagates slower than that Fig. 17. Distribution of induced total tangential stress by chemical loading ðC Ssh ¼
of the non-linear results. The accuracy of the pore pressures 0:1; C Sm ¼ 0:2Þ under thermal isothermal condition (Tsh ¼ Tm ¼ 90 1C).
obtained from the linear model are acceptable when the
difference of C Ssh and C Sm is relatively small. The distributions of
the radial and tangential total stresses for various times are shown Because of their significance on the wellbore stability, we
in Figs. 16 and 17. The linear results are very close to the non- summarize the total stresses (srr,syy,szz) and effective stresses
linear results, particularly for short times after the chemical ðs0rr ; s0yy ; s0zz Þ at the wellbore wall in Tables 3 and 4, which are,
loading is applied. respectively, obtained from the linear and non-linear models. The
ARTICLE IN PRESS

XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778 777

stresses at the wellbore wall are obtained though extrapolation The total and effective stresses at the wellbore wall from the
from the stresses at Gauss integral points. It is observed that the linear and non-linear models are presented in Tables 5 and 6. The
difference between the linear results and non-linear results is stresses from the linear model are significantly overestimated
small. The stresses from the linear model are slightly over- compared to those from the non-linear model. Therefore, when
estimated compared to those from the non-linear model. the difference between C Ssh and C Sm is relatively large, the linear
model does not perform well for stability analysis. However, such
6.2. Case with relatively large DC S large differences are not consistent with the fundamental
assumptions of the theory regarding a dilute solution.
In the following, we assume C Ssh ¼ 0:1 and C Sm ¼ 0:5. Fig. 18
shows the pore pressures profiles around the wellbore for various
7. Conclusions
times. It is observed that a higher salinity mud can induce higher
negative pore pressures in the shale. However, the discrepancy of
the pore pressures from the linear model and those from the non- A finite element method has been developed to implement the
linear model is much larger in this case, where the pore pressures fully coupled chemo-poro-thermoelastic linear and non-linear
could be overestimated as high as 50% in the linear model.
Therefore, it may not be appropriate to use the linear model to 0
obtain the pore pressures in the shale when the difference of C Ssh
and C Sm is relatively large. But, the results can be improved by -0.2
using a better approximation in Eq. (11). Figs. 19 and 20 show the

Radial Total Stress (MPa)


distributions of the total radial and tangential stresses around -0.4
the wellbore for various times. The total stresses from the linear
and non-linear models are in better agreement compared to the
pore pressures. -0.6

-0.8
Table 3
Stresses at the wellbore wall using linear model for the case with C Ssh ¼ 0:1 and 1 hour
12 hours
C Sm ¼ 0:2. -1 Solid Curves: Linear model results 24 hours
Dashed Curves: Non-linear model results 120 hours
Time (h) srr (MPa) syy (MPa) szz (MPa) s0rr (MPa) s0yy (MPa) s0zz (MPa)
-1.2
1 0.005 0.743 0.748 1.034 0.296 0.291 1 1.5 2 2.5 3
12 0.005 0.633 0.637 0.880 0.252 0.249 r/a
24 0.004 0.619 0.623 0.861 0.246 0.243
120 0.004 0.601 0.606 0.837 0.240 0.236 Fig. 19. Distribution of induced total radial stress by chemical loading ðC Ssh ¼
0:1; C Sm ¼ 0:5Þ under thermal isothermal condition (Tsh ¼ Tm ¼ 90 1C).

Table 4
Stresses at the wellbore wall using non-linear model for the case with C Ssh ¼ 0:1
0
and C Sm ¼ 0:2.
Tangential Total Stress (MPa)

Time (h) srr (MPa) syy (MPa) szz (MPa) s0rr (MPa) s0yy (MPa) s0zz (MPa)
-1
1 0.004 0.577 0.580 0.803 0.230 0.226
12 0.003 0.511 0.513 0.710 0.203 0.200
24 0.003 0.503 0.506 0.699 0.199 0.197
120 0.003 0.493 0.496 0.686 0.196 0.193 -2

1 hour
0 -3 12 hours
Solid Curves: Linear model results
24 hours
Dashed Curves: Non-linear model results 120 hours
-1 -4
1 1.5 2 2.5 3
Pore Pressure (MPa)

r/a
-2
Fig. 20. Distribution of induced total tangential stress by chemical loading ðC Ssh ¼
0:1; C Sm ¼ 0:5Þ under thermal isothermal condition (Tsh ¼ Tm ¼ 90 1C).
-3

1 hour Table 5
-4 12 hours Stresses at the wellbore wall using linear model for the case with C Ssh ¼ 0:1 and
Solid Curves: Linear model results 24 hours
C Sm ¼ 0:5.
120 hours
Dashed Curves: Non-linear model results
-5 Time (h) srr (MPa) syy (MPa) szz (MPa) s0rr (MPa) s0yy (MPa) s0zz (MPa)
1 1.5 2 2.5 3
1 0.007 1.259 1.266 1.751 0.501 0.493
r/a 12 0.006 0.944 0.950 1.312 0.375 0.370
24 0.006 0.906 0.911 1.259 0.359 0.354
Fig. 18. Distribution of induced pore pressure by chemical loading ðC Ssh ¼ 0:1; C Sm ¼ 120 0.006 0.855 0.861 1.191 0.341 0.335
0:5Þ under thermal isothermal condition (Tsh ¼ Tm ¼ 90 1C).
ARTICLE IN PRESS

778 XX. Zhou, A. Ghassemi / International Journal of Rock Mechanics & Mining Sciences 46 (2009) 769–778

Table 6 borehole wall, overestimating the required safe mud density


Stresses at the wellbore wall using non-linear model for the case with C Ssh ¼ 0:1 (conservative result). More realistic stress concentration around
and C Sm ¼ 0:5. the boreholes could be predicted by elasto-plastic models. There-
fore, FEM analysis similar to the one presented in the paper are
Time (h) srr (MPa) syy (MPa) szz (MPa) s0rr (MPa) s0yy (MPa) s0zz (MPa)
needed to provide a means for a comprehensive analysis with
1 0.003 0.681 0.682 0.946 0.268 0.266 consideration of non-linear chemical effects and plastic deforma-
12 0.002 0.469 0.469 0.649 0.183 0.182 tions are necessary.
24 0.002 0.443 0.442 0.613 0.171 0.171
120 0.002 0.410 0.410 0.567 0.160 0.159
References

models. In the finite element method, the nodal displacements,


[1] Maldonado B, Arrazola A, Morton B. Ultra-deep HPHT completions:
nodal pore pressures, nodal solute mass fractions and nodal classification, design methodologies, and technical challenges. In: Proceed-
temperatures are taken as primary unknowns. The finite element ings of the offshore technical conference, Houston, 1–4 May 2006, Paper OTC
model was validated by comparing to the available analytical 17927.
[2] Mody FK, Hale AH. A borehole stability model to couple the mechanical and
solutions for the poroelastic, thermo-poroelastic and chemo- chemistry of drilling fluid shale interaction. SPE paper 25728; 1993.
poroelastic wellbore problems in shale. [3] Onaisi A, Audibert A, Bieber MT, Bailey L, Denis J, Hammond PS. X-ray
The non-linear chemo-thermal-poroelasticity finite element tomography visuation and mechanical modeling of swelling shale around the
wellbore. J Petrol Sci Eng 1993;9:313–29.
model was used to analyze a wellbore in a non-hydrostatic stress [4] Sherwood JD. Biot poroelasticity of a chemically active shale. Proc R Soc Lond
field. The numerical results showed that when the mud 1993;A440:365–77.
temperature is higher than the formation temperature (heating) [5] Sherwood JD, Bailey L. Swelling of shale around a cylindrical well-bore. Proc R
Soc Lond 1994;A444:161–84.
and the mud solute concentration is lower than that in the [6] Ghassemi A, Diek A, Wolfe AC, Roegiers JC. A chemo-chemical model for
formation ðC Sm oC Ssh Þ, the pore pressure, radial total stress, borehole stability analyses. In: Proceedings of the 37th US rock mechanical
tangential total stress increase while the radial effective stress, symposium, Vail, Colo. Rotterdam: Balkema; 1999. p 239–46.
[7] McTigue D. Thermoelastic response of fluid-saturated porous rock. J Geophys
tangential effective stress decrease and vice versa. The redistribu-
Res 1986;91:9533–42.
tions of the pore pressures, total stresses and effective stresses are [8] Kurashige M. A thermoelastic theory of fluid-filled porous materials.
also significantly influenced by the in situ stress field. Based on the Int J Solids Struct 1989;25:1039–52.
[9] Wang Y, Papamichos E. Conductive heat flow and thermally induced fluid flow
numerical results, it may be concluded that the mud temperature
around a wellbore in a poroelastic medium. Int J Rock Mech Miner Sci
and solute concentration are two important factors in the 1995;32:262A.
wellbore stability analysis and their appropriate combination [10] Li X, Cui L, Roegiers JC. Thermoporoelastic modeling of wellbore stability in
may be utilized to control the wellbore stability. non-hydrostatic stress field. Int J Rock Mech Miner Sci 1998;35:4–5.
[11] Ghassemi A, Diek A. Poro-thermoelasticity for swelling shales. J Petrol Sci Eng
The influence of linearizing the chemical potential has also 2002;34:123–35.
been examined. The numerical results showed that the linear [12] Heidug WK, Wong SW. Hydration swelling of water-absorbing rocks:
model tends to overestimate the pore pressures, effective stresses a constitutive model. Int J Numer Anal Methods Geomech 1996;20:403–30.
[13] Ghassemi A, Diek A. Linear chemo-poroelasticity for swelling shales: theory
and total stresses around the wellbore compared to the non-linear and application. J Petrol Sci Eng 2003;38:199–212.
model. When the difference of the mud salinity and the shale [14] Ghassemi A, Tao Q, Diek A. Influence of coupled chemo-poro-thermoelastic
salinity is relatively small (e.g., 0.1–0.2), the results from the linear processes on pore pressure and stress distributions around a wellbore in
swelling shale. J Petrol Sci Eng 2008, submitted for publication.
model are satisfactory compared to those from the non-linear [15] Diek A, Ghassemi A. A chemo-poro-thermoelastic model for swelling shales.
model, particularly for short loading times. However, when the Department of Geology and Geological Engineering, University of North
difference of the mud salinity and the shale salinity is relatively Dakota. Int Rep NDRM-WSR-04-2001. 38pp.
[16] Fritz SJ. Ideality of clay membranes in osmotic processes: a review. Clay Clay
large (e.g., 0.4), the magnitudes of the pore pressures around the
Miner 1986;34:214–23.
wellbore from the linear model is significantly larger than those [17] Ekbote S, Abousleiman Y. Porochemoelastic solution for an inclined borehole
from the non-linear model, and both the total stresses and in a transversely isotropic formation. J Eng Mech ASCE 2006;132:754–63.
[18] Ekbote S, Abousleiman Y. Porochemothermoelastic solution for an inclined
effective stresses from the linear model are significantly over-
borehole in a transversely isotropic formation. J Eng Mech ASCE 2005;
estimated at the wellbore wall. However, such large differences 131:522–33.
are not congruent with the fundamental assumptions of the [19] Lewis RW, Schrefler BA. The finite element method in the static and dynamic
theory regarding a dilute solution. Therefore, the linear model is deformation and consolidation of porous media. 2nd ed. New York: Wiley;
1998.
an acceptable approximation of the non-linear model in the range [20] Smith IM, Griffiths DV. Programming the finite element method. 3rd ed.
of concentrations and swelling coefficients of interest. It can be New York: Wiley; 1997.
used in wellbore stability analysis with the understanding that the [21] Carter JP, Booker JR. Elastic consolidation around a deep circular tunnel.
Int J Solids Struct 1982;18:1059–74.
results tend to be conservative. [22] Detournay E, Cheng AH-D. Poroelastic response of a borehole in a non-
Finally, we have assumed linear elastic-brittle rock behavior hydrostatic stress field. Int J Rock Mech Miner Sci Geomech Abstr 1988;
for simplicity. This results in higher stress concentrations at the 25:171–82.

Anda mungkin juga menyukai