Sunteți pe pagina 1din 21

ACTA MECHANICA

Acta Mechanica 119, 199-219 (1996)


9 Springer-Verlag 1996

Poroviscoelastic analysis of borehole


and cylinder problems
Y. Abousleiman, Byblos, Lebanon, A. H.-D. Cheng, Newark, Delaware, C. Jiang, Wuhan,
People's Republic of China, and J.-C. Roegiers, Norman, Oklahoma

(Received June 18, 1995)

Summary. This paper addresses the phenomena of mechanical creep and deformation in rock formations,
coupled with the hydraulic effects of fluid flow. The theory is based on Biot's poroelasticity, generalized to
encompass viscoelastic effectsthrough the correspondence principle. Based on the resultant poroviscoelastic
theory, stress and deformation analyses are performed. The interactions between the fluid pore pressure
diffusion and the elastic/viscoelastic rock matrix deformation are illustrated via two important examples.
First, the problem of a borehole subject to a non-hydrostatic stress state, but deforming under plane strain
condition, is examined. Second, a cylinder under generalized plane strain conditions is solved. Three rocks,
Berea Sandstone, Danian Chalk, and a deep water Gulf of Mexico Shale, covering a wide range of
permeabilities, are considered. The significance of poro- and viscoelastic time-dependent effects is discussed.

1 Introduction

The coupling between pore pressure diffusion and elastic matrix deformation in fluid
impregnated rocks gives rise to a variety of time-dependent phenomena. The underlying
physical mechanisms and their engineering implications have been successfully studied as
the theory of poroelasticity [1], [2]. However, in addition to the effect of viscous fluid
diffusion, the rock mass itself may exhibit viscoelastic behavior. These time-dependent
characteristics may be attributed to the intergranular frictional sliding, the intrinsic solid
grain creep deformation, and even the solid-fluid interactions in fissures at the sub-granular
scale [3], [4]. A coupled poroviscoelastic theory is deemed necessary for the study of fluid
infiltrated rocks.
The groundwork for formulating viscoelasticity in fluid permeated porous media was laid out
by Biot [5], [6], and re-examined by a few others in the quasi-static [7] - [9] and dynamic ranges
[10]. However, not all the earlier work has implemented a specific rheological model and few hT~s
explored the engineering consequences.
With the simultaneous existence of poro- and visco-elasticity, the response of rocks displays
two time scales. The characteristic time of a viscoelastic response is an apparent material
property, while that for poroelasticity is dependent not only on the material itself, but also on
a global geometry length scale, namely the average distance to the drainage surface. To correctly
model the creep and relaxation phenomena in porous rocks, kowledge of their rheological
behavior should be gained not at the bulk material level, but at the micromechanical level.
Following these concepts, a 'micromechanically consistent' poroviscoelastic model was recently
proposed [11]. This particular model will be further examined in this paper with its engineering
consequences explored.
200 Y. Abousleiman et al.

The coexistence of poro- and viscoelastic phenomena has been observed in the laboratory
and in the field for a range of rock types [12]-[15]. The simultaneous modeling of these
mechanisms should help to interpret those observations. In field applications, one of the major
problems encountered is borehole instability. Drilling difficulties are associated mainly with
shales and tufaceous sediments. Typical problems encountered in shales are drill-tool sticking,
sloughing, and creep or swelling in the borehole wall. The cause of these problems can be of
chemical and/or mechanical nature. However, even in oil-base mud which is free from chemical
effects, borehole instability remains a problem. Part of the difficulty hence can be attributed to
mechanical effects, that is, of poro- and visco-elastic origin [15], [16]. The fluid flow in and out of
the drilled formation will impact upon pore pressure, effective stress state, and formation
apparent strength. A poroviscoelastic solution can help elucidate these phenomena and the
associated creep events.
Another area where poroviscoelasticity could be of importance is the determination of in-situ
stresses. Knowledge of in-situ stresses (directions and magnitudes) at great depths is important
for many geophysical and engineering applications. Hydraulic fracturing is so far the most
reliable method for in-situ stress determination in deep holes [17]. However, this operation may
add substantial costs to the overall project. An alternative technique is the strain-relaxation
method which was primarily used for the determination of stress orientation. This technique
relies on the measurement of preferential strairr recovery of rock core samples, due to the relief of
confining in-situ stresses at the time of drilling. Recent development has led to two techniques, the
anelastic strain recovery (ASR) method and the differential strain curve analysis (DSCA) [18],
which have allowed the measurement of not only stress directions but also magnitudes. At
present, neither of these methods accounts for the pore pressure effect which is simultaneously
present.
In this paper simple linear constitutive (rheological) laws of poroviscoelasticity are proposed
in a consistent manner at the micromechanical level. Applications are provided for two problems
of great importance to geophysical and petroleum applications, namely (i) a borehole in
a non-hydrostatic stress state under plane strain deformation; and (ii) a cylinder under
generalized (complete) plane strain condition.

2 Governing equations

The constitutive equations for poroelasticity can be written as follows [1], [2], [19]:

3K - 2G
a i i - eghjP = 2Ge~j + ~ 6~je (1)

Cp = r - ~e (2)

where cqj is the total stress tensor, p the pore pressure, e~j the solid frame strain tensor, e = e~ the
solid frame dilation, ( the variation of fluid volume per unit reference volume, and 6~j the
Kronecker delta. In the above, the sign conventions for stress and strain follow that of elasticity,
namely, tensile stresses and extensile strains are denoted positive. There are four independent
bulk material constants defined above: the shear modulus G, the drained bulk modulus K, Biot's
effective stress coefficient c~, and a compressibility coefficient for the solid-fluid system C.
The connection between ~ and C with some familiar engineering constants may be expressed
Poroviscoelastic analysis of borehole and cylinder problems 201

as

K u q g
- (3.1)
BKu

1 9(vu - v) (1 - 2v.)
C - M - 2GB2(1 - 2v) (1 + v,) 2 (3.2)

where Ku is the undrained bulk modulus, v and v, are the drained and undrained Poisson's ratios,
B is Skempton's pore pressure coefficient, and M is Blot's modulus [21].
Following the spirit of the generalized Hooke's law, the linear poroviscoelastic model can be
established by correlating the stress and the rate of stress ~ij, dlj, #~j ..., p, p,/~ ..., to the strain and
the rate of strain co., dij, +',j .... ~, (, ( .... Alternatively, we may utilize the hereditary integrals [20]
such that

3/((t) - 2G(t)
alj(t) -- 6o~(t) | p(t) = 2(~(t) | eo{t ) + 61j | e(t) (4)

C(t) | p(t) = ((t) - ~(t) | e(t) (5)

where/s d(t), ~(t) and C(t) are relaxation moduli (which have different dimensions as K, G,
etc.), and | denotes the Riemann-Stieltjes integral

Cr(t) | eo{t ) = S G(t - t') deij(t'). (6)


-oo

With the application of the Laplace transform, we have the correspondence relations

3K - 2(~
61j - ~6ij~ = 2CJg.lj + - - (~ij~ (7)

e~ = ( - a~ (8)

in which the tilde "denotes Laplace transform. The quantities K, G, c~, C are the Laplace
transform of K, G, etc. We have avoided the "notation to emphasize the fact that they carry the
same dimensions as the physical constants K, G, etc. These coefficients are the 'correspondence
coefficients' to those of poroelasticity.
Based on this correspondence principle, all equations of poroviscoelasticity in the Laplace
transform domain are formally the same as those for poroelasticity. However, the 'material
coefficients' for poroviscoelasticity carry different meanings. They are quotients of polynomials
of the Laplace transform parameter s. Their exact forms are determined from the specific
spring-dashpot model imposed. Following the same formality, the correspondence coefficients
for the engineering constants can be represented as

~2
/(, = / s + = (9.1)
C

~= R . - R
~R.- (9.2)
202 Y. Abousleiman et al.

3/( - 2G
- 6K + 2G (9.3)

3/s - 2G
% - 6/(. + 26" (9.4)

To complete the poroviscoelasticity model, other governing equations are needed. With the
application of the Laplace transform, they become

equilibrium equation

~ij,j = 0 (10)

Darcy's law

qi = - ~ , (11)

and continuity equation

s( + ql,i = 0 (12)

where ql is the specific discharge vector, s is the Laplace transform parameter, and ~ = k/~y is the
permeability coefficient, with k the intrinsic permeability and #s the fluid viscosity. For
simplicity, the permeability coefficient is assumed to be free from viscoelastic effects. In the above
we have neglected the body force terms and the initial stress and pressure fields for the simplicity
of presentation.

3 Micromechanics considerations

Although we have set up in the preceding Section the governing equations of poroviscoelasticity
through the use of correspondence principle, this formal construction procedure sheds little light
into the physical nature of the corresponding materials coefficients. At that level of understand-
ing, it is difficult to rationalize the use of a proper combination of spring-dashpot system to
represent a creep phenomenon based on physical or intuitive ground. Following these
considerations, the constitutive relations are reconstructed herein based on micromechanics.
Consider a 'sample' of porous material of volume V = Vp + V~, where Vp and V~ are the
volumes of the interconnected pore spaces and the solid grains, respectively. The porosity is
defined as the ratio 4) = Vp/V. We first examine the volumetric response of the porous solids
under an incremental hydrostatic compressive stress, P = --akk/3, and an incremental pore
pressure, p, exerted independently on the sample element. Each of the force components is
responsible for a portion of the volumetric strain, e = A V/V. To facilitate the presentation below,
the force components are reorganized into a pair (P', p) in which

P' = e - p (13)

is the well-known Terzaghi's effective stress. Consider that the force components P' and p each
inflict a volumetric strain characterized by the linear constitutive relations
p!
e (1) - (14.1)
K

e (2) = ----.P (14.2)


Ks
Poroviscoelastic analysis of borehole and cylinder problems 203

The total volumetric strain is thus


p!
e = e (t) + e (2) - - P (15)
K Ks

Based on (15), if a test is conducted under drained condition, i.e., p = 0, the relation reduces to

P
e = e (1) - . (16)
K

K is therefore the 'drained bulk modulus' of the porous frame. In an unjacketed test setup [21],
the sample is submerged in a fluid chamber, and the all-around fluid pressure is raised by p.
Under a pressure equilibrium, P and p are equal; thus P ' = 0, and (15) becomes

e = e (2) -- P . (17)
Ks

It has been argued that under the assumption of microscopic homogeneity and isotropy Ks is the
bulk modulus of the solid grain [22], [23]. We shall adopt that assumption henceforth.
Once the fundamental deformation mechanisms are identified and isolated as (16) and (17), it
becomes possible to visualize certain conceptual spring/dashpot systems based either on the
understanding of the underlying physics, or on the observation of controlled experiments. N o w
utilizing the correspondence principle, (14.1) can be formally expressed as

e~i) .... (18)


K

Similarly, it can be postulated that creep exists in the solid grains due to the existence of various
mechanisms, such as fluid penetrated microcracks and isolated pores, or the intrinsic creep of the
solid. Equation (14.2) can therefore be written in the same form as Eq. (18):

e~2) /~ (19)
K,

Although it is not necessary to specialize at this juncture, for easy understanding, we shall
assume that the volumetric creep of both the solid matrix and the solid grain can be described by
a three-parameter generalized Kelvin model which has found wide application for geomaterials
[18], [24], [25]. This conceptual spring-dashpot arrangement is illustrated in Fig. 1. When the

Fig. 1. A three-parameter generalized Kelvin model


204 Y. Abousleiman et al.

system is subjected to a step load, it instantly deforms in an elastic state characterized by the
spring constant K~. As time progresses, the resistance offered by the dashpot diminishes and the
system softens. At large times, the apparent spring constant becomes K~/(1 + K1/K2), which is
smaller than K~. The speed of the creep is regulated by the dashpot viscosity la. A characteristic
time scale for the creep can be defined as la/K2.
For the above process, the constitutive relation can be written as [20], [26]

K1K2 e(1) + KI#~ (1) = - ( K 1 + K2) P' - laP'. (20)

Performing the Laplace transform to the above and rearranging terms, we find the correspon-
ding bulk modulus defined in (18) as

~(= KIK2 + sKlla


(K 1 + K2) + sla" (21)

By the same token, we define k s as

~2s = Ks1Ks2 + sKsllas


(22)
(Ks1 + K,2) + slts"

Summing (18) and its analog (19), a relation equivalent to (15) is found:

P' p
~=~-+ _---. (23)
K s'

The above can be rearranged to obtain an 'effective stress law'

1
= -~ (P - ap-) (24)

in which

k
c7 -- 1 - / ~ (25)

is the corresponding Biot effective stress coefficient. Assuming that the solid matrix material
remains elastic in compression, Biot obtained a similar expression [6]. An equivalent expression
for c~ was also obtained by Coussy [9]. The relation in (25) indicates that the time-dependent
behavior of c~is likely to be complex. A direct modeling without micromechanical insight may not
be appropriate. This point will be further expounded in the ensuing Section.
Following the same correspondence principle, other micromechanical constitutive laws of
poroelasticity [2], [27] can be generalized for viscoelasticity. For example, an effective stress law
can be built for the fluid strain:

P- (26)

where /~ is the counterpart of Skempton's pore pressure coefficient B in poroelasticity.


It is found to be

(27)
/(AKs - K) + 4 R ( / ~ - Ks)"
Poroviscoelastic analysis of borehole and cylinder problems 205

In the above Ks is the fluid bulk modulus and q~the porosity. For simplicity, both are assumed to
be time-independent.
To complete the constitutive law, consider a sample subject to shear deformation without
volumetric strain. It is clear from elasticity theory that the following relation holds:

r = 2Geij (28)

where i # j. In poroelasticity, it is understood that the fluid does not resist shear deformation;
hence, the same law as the above, with G as the shear modulus of the skeleton, applies. For
poroviscoelasticity, the fluid can resist the rate of shear deformation; it may contribute to the
dashpot viscosity. The viscosity of the fluid, however, is much less than that of the solid. Its effect
vanishes rapidly and can be ignored under the assumption of quasi-static deformation. The
extension to poroviscoelasticity becomes

6i~ = 2Ggli (29)

where G can also be represented by a generalized Kelvin model,

~j = G1G2 + sGd~o (30)


(G1 + G2) + s# o"

The above constitutive equations, (24), (26), and (29), formed under micromechanics
considerations, are consistent with (7) and (8). The time-dependent behaviors are imposed at the
level of K, Ks and G through certain spring-dashpot analogies. All other bulk material coefficient
behaviors are dictated by these three fundamental creep mechanisms. Some of the relations for
bulk material coefficients are explicitly given below:

8 = Kf(s - g) + ~G(G - Ks)


(31.1)

/s - / ( ) + ~bK(K~ - KI)] (31.2)


((u = Ks(~ s _ s + @Ks(Ks - Ks)

For some geomaterials, the viscoelastic effect of/s may be ignored. This special case may be
achieved by replacing/(~ in all the above equations by Ks.

4 Laboratory test

Based on the above derived poroviscoelastic theory, we shall address the issues of laboratory
measurement of material coefficients in this Section. In poroelasticity, the drained bulk modulus
K is normally measured in a jacketed, but drained test [21] by ensuring p = 0. The same test can
be utilized for poroviscoelastic materials. It is however necessary to use a sample that is small
enough such that the time scale for pore pressure dissipation L2/c, where L is the average
drainage path and c is the diffusivity coefficient, is much smaller than that of viscoelastic creep. If
the confining stress of the sample is raised by the amount P, yet the pore pressure is kept at its
original value, the volumetric deformation of the sample A V(t) is monitored for an extended
period of time. Following viscoelasticity convention, a creep compliance function J(t) can be
206 Y. Abousleiman et al.

5.5

K1/K2=2.0
5.0

2.5

1
::~ 2.0

1.5

1.0
0.0 1.0 2.0 5.0 4.0 5.0
K2t/b~ Fig. 2. Creep compliance of a drained sample

defined as
A V(t)
J(t)- PV" (32)

Based on the generalized Kelvin model (21), the transient behavior of J(t) should be
I ( K 1 K 1 )
J(t) = ~ 1 + K2 g a e-KZt/U " (33)

In Fig. 2, the normalized compliance function J(t)/(1 is presented with a few KI/K2 values versus
the dimensionless time K2t/p. These curves can be utilized for the determination of the
coefficients K~, K2, and # from test data. By the same token,/s can be measured in a traditional
unjacketed test (P = p) for poroelasticity [21] by recording the material compliance as the
function of time. The shear compliance can be measured without the interference of the fluid.
For poroelastic materials the bulk coefficients, such as e, B, and K,, can be directly measured.
These concepts however m a y not work well when they are extended to viscoelasticity. Take for
example the direct measurement of c~.In the same jacketed, drained test as described above, if the
amount of fluid expelled from the sample, A VI, is monitored, the poroelasticity c~ can be
determined by the relation [2]
~v~
- (34)
AV"
As an analogy in poroviscoelasticity, we can define an instantaneous apparent effective stress
coefficient as

~(t)- AVAt) (35)


AV(t)
(note that c~(t) r c2(t)). By theoretical prediction, 7(t) behaves as
J~(t)
~(t) = 1 - - - (36)
J(t)
where J~(t) is the compliance function for /~. To examine the expected behavior of c~(t), we
evaluate the function and plot it versus the dimensionless time Kat/ps as Fig. 3. For convenience,
we have fixed the ratio of K1/Ks~ to 0.2. The other variables are presented in a few combinations
Poroviscoelastic analysis of borehole and cylinder problems 207

1.0
9

0.9

0.8

~jO.7

0.6

0.5 Fig. 3. The apparent effective stress coefficient


under a drained test. (K1/K2, K,I/K~2,
Ksz#/K2,us) = case 1: (0.5, 0.5, 10); 2: (0.5, 0.5,
0.4 ~ ,,;,,,,, i ,,~,,,,, ~ , , ,,,,, j , , ~ ..... ~ , , g .... 0.1); 3: (0.5, 2, 10); 4: (0.5, 2, 0.1); 5: (2, 0.5, 10);
o.ool O.Ol o.1 ~ ~o lOO 6: (2, 0.5, 0.1); 7: (2, 2, 10); 8: (2, 2, 0.1); 9: (2,
K2t//~ - , 0)
1.0

0.9 7
5,6

2 9
_+.a
0.8
CI3
4

0.7
Fig. 4. The apparent Skempton pore pres-
sure coefficient under an undrained test.
(K1/K2, K~l/g~z,K~2tz/Kz#s) = case 1 :(0.5,
0.6 ~ ,,~, ..... ~ , , ~ .....
k "&'"" J "~, ..... ~ " ~ , ' " 0.5, 10); 2:(0.5, 0.5, 0.1); 3:(0.5, 2, 10); 4: (0.5,
0.001 0.01 0.1 1 10 100 2, 0.1); 5: (2, 0.5, 10); 6: (2, 0.5, 0.1);
K2t//z 7: (2, 2, 10); 8: (2, 2, 0.1); 9: (2, - , 0)

ofK1/K2 = 0.5 or 2, K~I/Ks2 = 0.5 or 2, and K2ps/K~2p = 0.1 or 10. We also present in dashed
line the special case in which K~ is assumed to be time-independent (#~ ~ ~). Except for the
special case, these curves are non-monotonic. It is generally difficult to extract the information of
the six coefficients K1, Kz, K~t, K~2, li, and #s from these behaviors of c~(t). Hence the best way to
define ~ is t h r o u g h the separate measurement of the compliances of the skeleton a n d the grain.
A n o t h e r standard test c o n d u c t e d for poroelastic material is the undrained test. The ratio of
pore pressure rise to the applied confining stress defines the Skempton pore pressure coefficient
B. F o r viscoelastic behavior, the pore pressure rise should be continuously m o n i t o r e d as
a function of time. The apparent S k e m p t o n coefficient becomes

B(t)- P(t) - 2P-~ (37)

w h e r e / 6 - a denotes the Laplace inverse. By fixing K1/K~ = 0.2, K,/K~a = 0.1 and ~b = 0.2, the
same 9 cases as in Fig. 3 are calculated and presented in Fig. 4. We again observe some
n o n - m o n o t o n i c behavior.
208 Y. Abousleiman et al.

5 Applications

To illustrate the interplay between the poroelastic and viscoelastic effects, three rocks, a Berea
Sandstone [19], a Danian Chalk [13] and a deep water Gulf of Mexico Shale [16] are considered
with their poroelastic properties shown in Table 1. Water is used as the pore fluid with
K s = 3.3 x l 0 9 Pa and gf = 1.0 x 10- 3 Pa 9 s. The selection is made such that the permeabili-
ties, hence the poroelastic time scales, span a wide range. The results should provide a fair idea
about a whole range of rock poroviscoelastic responses.

Table 1. Poroelastic constants


Rock type G (Pa) KI(Pa) Ksl(Pa) q~ k (darcy)

Berea sandstone 6.0 x 10 9 8.0 x 10 9 3.6 x 101~ 0.19 1.9 x 10 -1


Danian chalk 2.2 x 109 3.3 x 109 1.2 x 10x~ 0.23 1.0 x 10- s
Shale 7.6 x 108 1.1 • 10 9 3.4 x 101~ 0.30 1.0 x 10 -7

Borehole

Consider a long and straight borehole with its longitudinal axis coinciding with one of the
principle stress axes, z. It is further assumed that the borehole deforms under plane strain, namely
that Uz = 0. This poroelastic problem has been investigated before [28] with analytical solutions
provided in the Laplace transform domain. By the correspondence principle, those solutions are
applicable to poroviscoelasticity.
To isolate the physical phenomena, and to facilitate the solution by superposition of
problems with different boundary conditions, the borehole problem was studied in three basic
'loading modes'. At the borehole wall, r = a, where a is the borehole radius, the boundary
conditions are summarized in Table 2, in which r and 0 are the polar coordinates, and the 1Is
factor is derived from the step change of the stress application. These loading modes have their
origin as (i) Po stands for a far-field isotropic stress; (ii) Po is derived from a virgin pore pressure;
and (iii) So is a far-field stress deviator. Whereas modes I and II are axisymmetric, mode III is not.
The solutions of these problems are available in Detournay and Cheng [28] and are listed in the
Appendix for easy reference.
For the mode I problem, it has been demonstrated [28] that the formation deforms only in
deviatoric strain; hence, no pressure is generated and p = 0 at all times. The response is purely
viscoelastic. In particular, the stresses
a 2
a~r = - ~roo = - Po -~2 (38)

are time-independent. The creep of the radial displacement, u,, is entirely controlled by the
time-dependent behavior of the shear modulus G. The trend is quite straightforward, hence is not
elaborated here.

Table 2. Loading modes for borehole problem


Mode ~,r 6,o /~

I - Po/s 0 0
II 0 0 po/s
III - S o cos 20/s So sin 20Is 0
Poroviscoelastic analysis of borehole and cylinder problems 209

1.0 1.0
poroviscoelastic poroviscoelastic
poroelostic poroelostic
0.8
0.8
Berea
00.6
0.6 O_

~:~0.4
[
0.4
0.2

0.2
0.0

0.0 .................. ..........


-0.2 . . . . i . . . . i . . . . i . . . . i . . . . I . . . . i . . . . i . . . . i . . . .

10 102 10 ~ lO" 105 106 107 0.1 0.2 0.3 0.5 0.6 0.8 0.9
t (sec)
Fig. 5 Fig. 6
Fig. 5. Borehole pore pressure history at r = 1 m, mode II

Fig. 6. Borehole tangential stress as a function of radial distance at t = 2 days, mode II

Examine next the mode II problem, in which a pore pressure is applied at the borehole wall to
induce a radial diffusion. For the pore pressure expression (46), the material coefficients enter this
solution in the form of a diffusivity coefficient 6, defined in (A. 20.3). The behavior of 6 is
somewhat complex as indicated by its micromechanics definition

zKs/s + 3/~)
6 = Kz(/s - / s (4G + 3/(s) + ~b/s163 - K~) (4G + 3/s (39)

As a first example, we assume that only the shear modulus exhibits viscoelastic properties and
choose Ga = G1 such that the shear modulus is reduced to one-half of its original value at the end
of the creep. Little is known about the viscosity for these rocks. Loosely based on the field
observation of some retrieved cores, we choose here the PG values such that the time scale #a/G2 is
exactly two days for all three rocks. For a borehole with radius a = 0.1 m, Fig. 5 presents the
normalized pressure versus time at the location r = 1 m, in solid lines. For comparison, the
poroelastic solutions are shown by dashed lines. The solutions are presented for the time range of
10 sec to about 100 days. Within that range, the difference in pressure response with and without
viscoelasticity is insignificant.
As a second example, we turn our attention to the viscoelastic properties of K. Choosing
K2 = K1 and the viscoelastic time scale [IlK 2 to 2 days (with G regarded as time independent),
Fig. 6 demonstrates the normalized compressional stress -aoo/po (it is reminded that positive
stress stands for tension), as a function of radial distance at t = 2 days, for the mode II problem.
The viscoelastic effect is clearly observable in Danian Chalk and Berea Sandstone, while for
Shale, with its relatively low permeability, the two solutions track each other closely. Notice that
the dimensionless time for Shale is still small at 2 days, hence part of the solution is in the tensile
range (-aoo/Po < 0).
Figure 7 depicts the tangential stress a00 at the wall at 0 = 0 for the mode III problem. The
viscoelastic and poroelastic responses completely overlap for Berea Sandstone and Danian
210 Y. Abousleiman et al.

4.2 I
bJ
Berea Sandstone eq
poroviscoelastic
4.0 poroelast;c / ....

5.8
o
u') 3.6

i:~ 3.4
///•O nia n Chalk

h o l e
/
/ ~ z

FS
I kaJ

3.2 (./3
Danian Ch a Ikj...---'f~
3.0
oroviscoelastic
2.8 poroelast[c
CD
O Berea Sandstone
O
2.6 , , ,,,,,,i ........ i ........ i , ,,,,,-i , ,,,,,-i , ,,,,,-
0 10 2 10 3 1
0 4
10 5
10 6
10 7 wlO 102 103 10' 105 10 B 107
t (sec) o t (see)

Fig. 7 Fig. 8
Fig. 7. Borehole tangential stress as a function of time at r = a, 0 = 0, mode III

Fig. 8. Borehole radial displacement as a function of time at r = a, 0 = 0, mode III

Chalk, but not for Shale. At large time, the asymptotic value of a00 = - 4So is reached for all
cases. The borehole radial displacement u, at 0 = 0 and r = a is presented in Fig. 8. By noting the
difference between the solid and the dashed lines as the amount of viscoelastic creep, we can
clearly see the two stage consolidation in the Danian Chalk and the Shale cases.

Cylinder

We assume that an infinitely long, circular cylinder is subjected to a prescribed rate of strain
along its longitudinal axis. On its surface, stress and pressure conditions similar to those for the
borehole problem apply. The mathematical solution is attempted here under the assumption of
generalized plane strain [29], [30] (or complete plane strain [31]). In other words, the axial strain is
only a function of time (e= = g(t)). The strain g(t) can take the form of any general function to
simulate different stress or strain loading rates. For the convenience of presentation, we shall use
g(t) = -Aot, where Ao is a constant.
Following the plane strain borehole problem, this problem is also investigated through
loading mode decomposition. In addition to modes I, II, and III, a mode IV is added to repre-
sent the axial strain. The boundary conditions are summarized in Table 3, where the con-
ditions for g,r, #,0, and/~ are applied at r = a, while that for ~zz is enforced throughout the
cylinder.
An examination of mode I, II, and III reveals that they are plane strain problems (as a special
case of generalized plane strain). The plane strain mode I and II solutions have been derived
earlier [2]. The mode III and IV results are presented for the first time in this paper. All these are
listed in the Appendix. In the numerical evaluation of these solutions, we continue to restrict the
viscoelastic behavior to/s and assume that K2 = K1. The viscoelastic time scale is again fixed at
2 days for all rocks. The size of the cylinder is given as a = 10 cm.
Poroviscoelastic analysis of borehole and cylinder problems 211

Table 3. Loading modes for cylinder problem

Mode 6rr 6r0 .5 ~zz

I -- Po/s 0 0 0
II 0 0 Po/S 0
III --So cos 20/s So sin 20/s 0 0
IV 0 0 0 - Ao/s 2

C)

1.2 S
I
poroviscoelostic
poroelastic porovlscoelostic
poroelastic
1.0 0
s
~ . i -4
0.8
Shale
S h a l e
..~0.6
d)_ a_%J
0.4 ~0
tS Chalk
,,,
0.2 (N

0
(:3
0 Berea S a n d s t o n e
0.0 " "' Jr
0 10 2 10 3 10 4 10 5 10 6 10 7 w1, 102 103 10 " 10 ~ 106 107
0
t (sec) t (See)
Fig. 9 Fig. 10

Fig. 9. Cylinder pore pressure history at r = 0, mode I

Fig. 10. Cylinder radial displacement history at r = a, mode I

Figure 9 presents the normalized pressure versus time at r = 0 for mode I. The poroelastic
and poroviscoelastic responses closely follow each other. For Berea Sandstone, the pore pressure
is rapidly diffused due to its relatively large diffusivity coefficient, c, and the small specimen size.
Also of interest to observe is the nonmonotonic pressure behavior, known as the Mandel-Cryer
effect [32], [33] demonstrated in the curve for Shale. Figure 10 illustrates the inward radial
displacement at the cylinder surface, r = a. The consolidation is observed in two stages, with the
poroelastic effect manifesting first and then followed by the viscoelastic one. Figure 11 shows the
axial stress azz at the center of the cylinder, r = 0. The stress is created by the Poisson effect. At
small times, the effective Poisson ratio is dominated by the undrained value v,. Once the pore
pressure is dissipated, it is characterized by the drained value v. As time progresses, the material
continues to soften du6 to the viscoelastic effect, and the axial stress further drops. This stress
relaxation phenomenon is non-monotonic, as evident from the curve for Shale.
Next, the mode II loading, which exposes the cylinder surface to a pore pressure increment,
yet without the change of total stress, is examined. Figure 12 sketches the pore pressure solution
at the center of the cylinder, r = 0. Similar to the earlier cases, the pore pressure responses are
insensitive to the viscoelastic effect. For Berea Sandstone, the pore pressure is immediately
equilibrated. For Shale, the pressure at the center becomes negative for a period of time. The
radial displacement at the outer surface, r = a, is displayed in Fig. 13. This diagram is similar to
212 Y. Abousleiman et al.

1.2 1.2

Shale
1.0 1.0

0_0
~0.6
0.8
:ionCa: . . . . . . . . . .

0.6
b D-0.4 I /Donion Chalk ~
I 0.4 0.4

0.2

0.0
_ _ _ poroelastic
01/
0.0 ~ ~ L ~ porovisco.elastic

-0.2 -0.2 - ~
10 102 10s 10' 10 ~ 108 107 10 102 103 1 0 " 1 0 5 10 8 10
t: (sec) t (see)
Fig. 11 Fig. 12
Fig. 11. Cylinder vertical stress history at r = 0, m o d e I

Fig. 12. Cylinder pore pressure history at r = 0, mode II

O
0
S 0
I +
Ld I, I
00
_ _ poroviscoelastic r~ poroviscoelastic
_ _ _ poroelastic Doroelastic
O
s
I
0
0+ -

#-'N :
E2 /
I
oLLI

DO
o_y_
i, i
SI ' / ~ n ~"/ - Chalk
bJ
ol

0
J Danlan Chalk
o ~ -
0
Berea S-ah-dston-&
O
O
O
J Berea Sandstone
4- , ,,,.,.i , ,,,-.i , ,,,,,,.i . ,,,,,.i ........ i . . . . . . + , ,,,,,,,i , ,,,,,,,i , , ,,,,,,i , ,,,,,,,i . . . . . . . .

~10 102 103 10" 105 106 107 102 103 10" 105 1
o
t (see) t (sec)
Fig. 13 Fig. 14
Fig. 13. Cylinder radial displacement history at r = a, m o d e I i

Fig. 14. Cylinder pore pressure history at r = 0, m o d e I V

Fig. 10. We observe that the m o v e m e n t is outward because pressure increase causes the cylinder
to swell. Despite the mathematically imposed condition that there is no total stress acting on the
cylinder, a viscoelastic creep is nevertheless present.
Mode III normally does not naturally arise as a direct loading condition. It is rather
a mathematical artifact derived from the loading decomposition of the deviatoric part of stress
Poroviscoelastic analysis of borehole and cylinder problems 213

0
0
g
0
4 -t-

poroviscoelastic c,,i _ _ poroviscoelastic


poroelastic ___ poroelastic
c'4
E
i1
hi
/
E~
z
Z
v q+~ Berea S" W ~
o: b
I
+-
cO
0
o ..~" Danian C h a l k ~

Ld
g // Danian Chalk
......... ~ . . . B e , re'o..S.'o.ndsto.n..e..
~0 102 103 10" 105 1 + I

lu~ + O 0 0
I I I I I I

2E+005
I I I . . . . . I

4E+005
. . . . . .

6E+005
. . . I

8E+005
i

1E+006
0 t o t (sec)
Fig. 15 Fig. 16
Fig. 15. Cylinder tangential stress history at r = a, mode IV

Fig. 16. Cylinder vertical stress history at r = 0, mode IV

unloading of a retrieved cylindrical core. The solution (see Appendix) indicates the cylinder
deforms as pure shear in plane strain, hence no pressure is generated at all times. The behavior is
purely viscoelastic, and is not elaborated.
M o d e IV can originate either as a displacement controlled axial loading in a l a b o r a t o r y
setting, or as a consequence of relaxation of in situ stress in a retrieved cylindrical core. F o r the
same rocks and the same viscoelastic properties as before, the axial strain rate is fixed at
~=z = - 1 0 -8 sec -1. Figure 14 depicts the pore pressure rise at r = 0. The pore pressure
generated in Berea Sandstone and D a n i a n Chalk appears to be negligible. F o r Shale, the pore
pressure rises sharply due to the continuous development of pore pressure by the constant strain
rate. After a certain time, the rate of increase in pore pressure is counter-balanced by the pore
pressure diffusion as a result of Darcy's flow. The pressure then levels off to reach a steady state.
F o r the case ofporoviscoelasticity, the viscoelastic creep further contributes to the pore pressure
generation. A larger pore pressure asymptotic value is thus observed.
Figure 15 exemplifies the development of a tensile tangential stress (hoop stress) at
r = a. If the material is purely elastic or viscoelastic, a uni-axial stress condition develops,
and aoo is expected to be zero. The presence of pore pressure however effectively makes
the material inhomogeneous with a hardened core. A tensile stress thus develops near
the skin.
The compressive vertical stress is illustrated in Fig. 16 at r = 0. The stress increment is
basically linear for all cases. The rate of rise is higher for stiffer than for softer rocks. The
poroviscoelastic stresses are lower than their poroelastic counterparts because viscoelastic
materials are m o r e compliant.
Figure 17 exhibits the radial displacements. N o t surprisingly, they increase with time roughly
in the linear fashion. The poroelastic magnitudes are significantly greater than poroviscoelastic
ones.
214 Y. Abousleiman et al.

0
0
0
c5
_ _ poroviscoelastic
___ p0r0elasuc

//
o4 //
0
f.-~O 9 // /
Eo
j-"
~ j
3 ,s f
/ /
f /

O"
0 "~///

..i ~ Shale
. / ~ Ik

. . . . . . . , ,i , nl , , , ,i i i . . . . . . ,, ii ,, , ,, ,, , ,I ,i . . . . . .

OR+O00 2E+005 4[+005 6E+005 8[+005 1E+006 Fig. 17. Cylinder radial displacement history
o t (sec) at r = a, mode IV

6 Conclusions

In this work we have discussed in some detail a linear poroviscoelastic theory for the modeling of
fluid infiltrated porous rocks. The salient features are summarized as follows:
(i) By the use of correspondence principle, the linear constitutive relations of poroelasticity
are immediately transformed to those for poroviscoelasticity.
(ii) By operating in the Laplace transform domain, all poroelastic solutions are generalized
to poroviscoelastic ones.
(iii) The spring-dashpot analogues are best constructed at the micromechanical rather than
the bulk material level. Three fundamental deformation mechanisms that can exhibit viscoelastic "
creep are identified, namely the solid frame volumetric deformation, the solid constituent
volumetric deformation, and the solid frame shear deformation. The mechanical/theological
analogues are applied at the three coefficients K, K~, and G.
(iv) Based on the theoretical prediction, the time-dependent behavior of bulk material
coefficients can be complicated. The possible non-monotonic behaviors of c~ and B are
demonstrated, but are yet to be confirmed in the laboratory.
(v) Two practical problems, a borehole and a cylinder, are investigated with data from three
rocks of drastically different permeabilities. The interaction between the poroelasticity and
viscoelasticity ranges from weak to strong for different problems and different quantities of
interest.
(vi) For the isolation of the physical phenomenon, the problems are investigated through
loading mode decomposition. Due to the linear nature of the problems, these solutions can be
assembled in a variety of fashions to model problems of borehole excavation, borehole
pressurization, fluid production, cylinder loaded by solid contacting force, and by fluid pressure,
stress relaxation of retrieved cores, etc.
(vii) Although the generalized Kelvin model is used as the rheological analog based on past
observation of geomaterials, more complicated spring-dashpot models can be easily adopted
once more experimental evidences are established.
(viii) The boundary conditions used in the various problems are either constant step or
simple linear time functions. More general time behaviors can be adopted by replacing the
constant loading magnitudes by their equivalent Laplace transform functions.
Poroviscoelastic analysis of borehole and cylinderproblems 215

Appendix

Borehole solution

Mode I

2GsG a
(A.1)
Poa r

S f f rr a 2

Po r2
(A.2)

S f f o0 a 2

Po r2
(A.3)

Mode II

(A.4)
poa L~o(~) CKo(~)l

sG, _ 2q F K~(~) a Kdfl)~ (A.5)


Po L~o~) r ~Ko(fl)J

Sffoo _ 2q ~ KI(O a Ki(fl) + K o ( 0 ]


(A.6)
po Lr r CKo(fl) K~J
s~ Ko(r
(A.7)
po Ko(fl)

Mode III

2Gs~Tr_ --C1 KI(0 + K2(0 + C 2 -- -{- C3 cos 20 (A.8)


Soa fl ~ r ~-

2Gsao _ -2C1 1 K2(O 1 -- 2~.


- - C 2 -- -t- C 3 sin 20 (A.9)
Soa ~ 2(1 - G) r

a2 a'*}
So - c~ KI(O + ~ K dO - - -1 -- ~u C 2 ~ -- 3C3 ~ cos 20 (A.IO)

S~oo
So- {-Cj [88KI(~) + ( 1 + ~6~2) K2(01 +3C3 ~} cos 20 (A.11)
216 Y. Abousleiman et al.

So - 2C~ KI(~) + ~5 K2(~)


] 1
2(1 - G~ Ca ~ - 3C3 ~- sin 20 (A.12)

sp {1 q C az}
So = ~ C 1 K 2 ( r ~ 2~5 cos 20 (A.13)

in which

4tiff. - 9)
C~ - (A.14)
D2 - D1

4(1 -- 9,) D2
C2 - (A.15)
D2 - D1

fl(D2 + D1) + 807. - ~7)K2(fi)


C3 = (A.16)
fl(D~ - n ~ )

and

D1 = 207, - ~) Kl(fi) (A.17)

D2 = fi(1 - 9) K2(fl). (A.18)

The following definitions were also used: Kn are modified Bessel functions of the second kind of
order n,

=r ~ (A.19.1)

/3 = a ] / ~ , (A.19.2)

S, q, and 5 are respectively the storativity coefficient, the poroelastic stress coefficient, and the
diffusivity coefficient given by

907, - 9)(1 - G)
= 2d/~2(1 + ~,)2 (1 - ~) (A.20.1)

3(ft. - ~)
q = _ (A.20.2)
2B(1 + 7,) (1 - 9)

= -=. (A.20.3)
S

Cylinder solution

Mode I

2GsG r (1 - 2 9 , ) (1 - 9)lo(/3) + 207, - ~) r


(A.21)
Poa a D3

stir, 2(G - "~) ~- 111({) - (1 - "~) Io(/3)


(A.22)
Po D3
Poroviscoelastic analysis of borehole and cylinder problems 217

SSoo 2(~u - ~7) [lo(4) - 4-11,(4)1 - (1 - 9) Io(fl)


(A.23)
Po D3

S~rO
=0 (A.24)
Po
s~ gu --~ Io(fl) --lo(4)
(A.25)
Po q D3

M o d e II

2GsG r 4-111(4) + (1 -- 2~u) fi- lIl(fl)


- 20(1 - ~) (A.26)
poa a D3

SGr fl- lla(fl) -- 4 - 111(4)


= 20(1 -- ~) (A.27)
Po D3

S~O0 4-111(4) -- 10(4) + fl-111(fl)


- 20(1 - ~) (A.28)
po D3

Sr
=0 (A.29)
Po

s/~ (1 - ~7)lo(4) - 207. - ~)fl-111(fl)


(A.30)
Po Da

Mode III

2GsG r
- cos 20 (A.31)
Xoa a

2GSgo r
- - - sin 20 (A.32)
Soa a

Sffrr
= cos 20 (A.33)
So

S~oo
- cos 20 (A.34)
So

S~rO
- sin 20 (A.35)
So

sp
--~0 (A.36)
So

Mode IV

2sZ~, r ~u(a - ~) Io(fl) - (Vu - v) [fi- 111(fl) + 4-111(4)1


(A.37)
Aoa a D3
218 Y. Abousleiman et al.

S20"rr fl- l/l(fl ) -- ~ - 1]l({ )


_ - 2ft. - 9) (A.38)
AoG Da

SZgoo fl- ~Ii(fl) + ~- ~I~(~) - Io(~)


-2 (A.39)
AoG D3

sZaz~ (9, -- ~) [I0(r + 2fl- all(fl)] - (1 + ~,) (1 - ~) I0(fl)]


-2 (A.40)
Aod D3

S2 ff rO
- :0 (A.41)
AoG

s2p (f)u -- ~) Io(fl) -- Io(~)


(A.42)
AoG q D3

In the above,

D3 = (1 - "~)Io(fl) - 2(9. - 9) fl- lla(fi), (A.43)

and I . are modified Bessel functions of the first kind of order n.

References

[1] Biot, M. A.: General theory of three-dimensional consolidation, J. Appl. Phys. 12, 155-164 (1941).
[2] Detournay, E., Cheng, A. H.-D.: Fundamentals of poroelasticity. In: Comprehensive rock engineering:
principles, practice & projects. Vol. II, Analysis and design method (Fairhurst, C., ed.) pp. 113-171.
Pergamon 1993.
[3] Blot, M. A.: Mechanics of deformation and acoustic propagation in porous media. J. Appl. Phys. 33,
1482-1498 (1962).
[4] Cleary, M. R: Elastic and dynamic regimes of fluid-impregnated solids with diverse micro-structure. Int.
J. Solids Struct. 14, 795-819 (1978).
[5] Blot, M. A.: Theory of deformation of a porous viscoelastic anisotropic solid. J. Appl. Phys. 27,
459-467 (1956).
[6] Biot, M. A.: Nonlinear and semilinear rheology ofporous solids. J. Geophys. Res, 78, 4924-- 4937 (1973).
[7] Tabaddor, E, Little, R. W.: Interacting continuous medium composed of an elastic solid and
incompressible Newtonian fluid. Int. J. Solids Struct. 7, 825-841 (1971).
[8] Wilson, R. K., Aifantis, E. C.: On the theory of consolidation with double porosity. Int J. Eng. Sci. 20,
1009-1035 (1982).
[9] Coussy, O.: Mecanique des milieux poreux. Paris: Editions Technip 1991.
[10] Vgenopoulou, I., Beskos, D. E.: Dynamic behavior of saturated poroviscoelastic media. Acta Mech. 95,
185-195 (1992).
[11] Abousleiman, Y., Cheng, A. H.-D., Jiang, C., Roegiers, J.-C.: A micromechanically consistent
poroviscoelasticity theory for rock mechanics applications. Int. J. Rock Mech. Mining Sci. Geomech.
Abstr. 30, 1177-1180 (1993).
[12] E1 Rabaa, A. W., Meadows, D. L.: Laboratory and field application of the strain relaxation method. SPE
15072, 56th California Regional Meeting, Oakland, CA, April 2 - 4 , 1986.
[13] E1Rabaa, A. W.: Determination of the stress field and fracture direction in the Danian chalk, in: Rock at
great depth (Maury, V., Fourmaintraux, D., eds.) pp. 1017-1024. Rotterdam: Balkema 1989.
[14] Warpinski, N. R., Teufel, L. W.: Author's reply to discussion of: A viscoelastic constitutive model for
determining in-situ stress magnitudes from anelastic strain recovery of core. SPE Production Eng.
287-289, August 1989.
[15] Nakken, S. J., Chritensen, T. L., Marsden, R., Holt, R. M.: Mechanical behavior of clays at high stress
levels for wellbore stability applications. In: Rock at great depth (Maury, V., Fourmaintraux, D., eds.).
Rotterdam: Balkema 1989.
Poroviscoelastic analysis of borehole and cylinder problems 219

[16] Mody, F. K., Hale, A. H.: A borehole stability model to couple the mechanics and chemistry of drilling
fluid/shale interaction. SPE/IADC 25728, Amsterdam, February 1993.
[17] Haimson, B. C., Roegiers, J.-C., Zoback, M. D.: (eds.) Proc. 2nd Int. Workshop hydraulic fracturing
stress measurements, University of Wisconsin-Madison, Minneapolis, 1988.
[18] Warpinski, N. R., Teufel, L. W.: A viscoelastic constitutive model for determining in-situ stress
magnitudes from anelastic strain recovery of core. SPE Production Eng. 272-280, August 1989.
[19] Rice, J. R., Cleary, M. P.: Some basic stress-diffusion solutions for fluid saturated elastic porous media
with compressible constituents, Rev. Geophys. Space Phys. 14, 227-241 (1976).
[20] Fliigge, W.: Viscoelasticity. Waltham: Blaisdell 1967.
[21] Blot, M. A., Willis, D. G.: The elastic coefficients of the theory of consolidation. J. Appl. Mech. 78,
91--96 (1956).
[22] Nur, A., Byerlee, J. D.: An exact effective stress law for elastic deformation of rock with fluids. J.
Geophys. Res. 76, 6414-6419 (1971).
[23] Zimmerman, R., Somerton, W., King, M: Compressibility of porous rocks. J. Geophys. Res. 91,
12765-12777 (1986).
[24] Taylor, D. W., Merchant, W.: A theory of clay consolidation accounting for secondary compression. J.
Math. Phys. 19, 167-185 (1940).
[25] Corapcioglu, M. Y., Brutsaert, W.: Viscoelastic aquifer model applied to subsidence due to pumping.
Water Resour. Res. 13, 597-604 (1977).
[26] Christensen, R. M.: Theory of viscoelasticity, 2nd. ed. New York: Academic Press 1982.
[27] Caroll, M. M.: Mechanical responce of fluid-saturated porous materials. In: Theoretical and applied
mechanics. Proceedings of the 15th Int. Cong. of Theoretical and Applid Mechanics, Toronto 1980
(Rimrott, E P. J., Tabarrok, B. eds.) pp. 251-262.
[28] Detournay, E., Cheng, A. H.-D.: Poroelastic response of a borehole in a non-hydrostatic stress field. Int.
J. Rock Mech. Mining Sci. Geomech. Abstr. 25, 171-182 (1988).
[29] Lekhnitskii, S. G.: Theory of elasticity of an anisotropic elastic body. San Francisco: Holden-Day
1963.
[30] Saad, A.: Elasticity theory and applications. New York: Pergamon Press 1974.
[31] Brady, B. H. G., Bray, J. W.: The boundary element method for determining stresses and displacements
around long openings in a triaxial stress field. Int. J. Rock. Mech. Min. Sci. Geomech. Abstr. 15, 21 - 2 8
(1978).
[32] Cryer, C. W.: A comparison of the three-dimensional consolidation theories of Biot and Terzaghi. Q. J.
Mech. Appl. Math. 16, 401-412 (1963).
[33] Abousleiman, Y., Cheng, A. H.-D., Cui, L., Detournay, E., Roegiers, J.-C.: Mandel's problem revisited.
G6otechnique (to appear).

Authors' addresses: Y. Abousleiman, School of Engineering and Architecture, Lebanese American


University, Byblos, Lebanon; A. H.-D. Cheng, Department of Civil Engineering, University of Delaware,
Newark, DE 19716, U.S.A.; C. Jiang, Institute of Architecture and Civil Engineering, Wuhan University of
Technology, Wuhan, Hubei 430070 P.R. China, and J.-C. Roegiers, School of Petroleum and Geological
Engineering, The University of Oklahoma, Norman, OK73019-0628 U.S.A.

S-ar putea să vă placă și