You are on page 1of 25

CHAPTER 19

A Theory for Coupled Stress and Fluid


Flow Analysis in Jointed Rock Masses
MASANOBU aDA

Department of Foundation Engineering, Saitama University, Japan

ABSTRACT
Rock masses, which commonly contain a large number of discon-
tinuities such as faults and joints, are treated as anisotropic, elastic
porous media in order to get a complete set of equations dealing with
the coupled stress and fluid flow analysis. Two basic assumptions are
adopted: (1) geological discontinuities behave like tiny cracks in an
elastic continuum; (2) each crack is modeled by a set of parallel planar
plates connected by two springs. Elastic compliance and permeability
tensors are successfully formulated in terms of crack tensors, which
depend only on the geometry of the related cracks, plus the shear and
normal stiffness values equivalent to the elasticity of the cracks.

RESUME
Les masses rocheuses, qui comprennent en general beaucoup de
discontinuites, comme les fautes et les joints, sont considerees ici comme
des milieux anisotropes, elastiques et poreux afin d'obtenir un ensemble
compler d'equations qui pefmet d'analyser les contraintes en couplage
avec l'ecoulement fluide. On adopte deux hypotheses fondamentales:
(1) les discontinuites geologiques se comportent comme des fissures
minuscules dans un continuum elastique; (2) on modetise chaque fissure
par ie moyen d'un ensemble de plaques planes paralteles reliees par
deux ressorts. On a reussi a formuler des {enseurs de complaisance
elastique et de permeabilite en fonction des tenseurs de fissuration qui
dependent seulement de la geomerrie de ces fissures et en outre des
349
J. Gittus et al. (eds.), Large Deformations of Solids: Physical Basis and Mathematical Modelling
Elsevier Applied Science Publishers Ltd 1986
350 MASANOBU ODA

valeurs de rigidite normale et sous cisaillement equivalentes a I'elasticite


des fissures.
1. INTRODUCTION

In recent studies of rock mechanics much attention has been focused


on the problems concerning the deep underground burial of high-level
nuclear waste. The repository will be about 300-1000 m below the
surface in crystalline rock masses (e.g. Runchal and Maini, 1980). One
of the most serious problems to be solved is the isolation of the
high-level nuclear waste from the biosphere. Ground water flow
through various geological discontinuities is believed to be the most
significant mechanism of radionuclide migration (e.g. Marsily et at.,
1977).
Geological discontinuities such as faults and joints are of widespread
occurrence in crystalline rocks. To treat these discontinuities in
numerical analyses of rock masses, the so-called joint element has
been developed by Goodman et al. (1968). Recently, Noorishad et al.
(1982) have developed a finite element method for the coupled stress
and fluid flow analysis in which the joint elements are used in place of
the discontinuities. The joint element is really a powerful tool to treat
a few major discontinuities like faults. It is not practical, however, to
replace all visible discontinuities (joints) by the joint elements. As a
matter of fact, individual joints cannot playa dominant role, but tend
to be well connected to make major flow paths. In such cases,
disregard of such joints (called cracks) will lead to a false result in the
flow analyses.
A theory dealing with the coupled stress and fluid flow analyses is
given here with two basic assumptions: (1) cracks behave just like tiny
flaws in an elastic continuum; and (2) each crack can be replaced by a
hydro-mechanical equivalent of parallel planar plates connected by
two springs (Fig. 1). The former assumption may be true because
cracks are usually small enough as compared with the extremely large
size of related rock masses.

2. SPRING MODEL OF DISCONTINUITIES

2.1. Definition of Overall Stress and Strain Tensors


A representative elementary volume V is used in the sense of Bear
(1972). No geological bodies are homogeneous in the strict sense.
COUPLED STRESS AND FLUID FLOW ANALYSIS 351

However, it can always be expected to find an elementary volume


which appears homogeneous on the macro scale.
Overall stress aij and strainEij tensors are defined by taking the
averages of local stress aij and strain fij tensors over a representative
elementary volume, respectively:

aij=i Iv aijdV; Eij=i Iv fijdV (1)

The solid matrix of crystalline rocks is assumed to be impermeable


so that fluid only flows through cracks. Let p be the overall fluid
pressure, defined by.

p = V~c) JV(C)
( p dV (2)

where p is the local fluid pressure and V(c) is the void volume
associated with cracks in the elementary volume V. In accordance with
a:
the common definition, the effective stress tensor j is given by
-, - - so.
aij = aij - PUij (3)
where {)ij is Kronecker's delta.

2.2. Shear Stitfness


Since each crack is replaced by a set of parallel planar plates
connected by two springs (Fig. 1), the elasticity is represented by the
corresponding stiffness values. Bandis et al. (1983) have performed
direct shear tests on several natural joints under normal stresses an in
the range of engineering interest. It has been shown that a hyperbolic
function can be well fitted to the observed non-linear relations
between the shear stress and the shear displacement. According to the
study by Barton and Choubey (1977), such a non-linear curve is

Fig. 1. Spring model of a crack.


352 MASANOBU ODA

simplified by a straight line having an equivalent stiffness G up to the


peak:

G = 1~0 an tan[JRCIoglOC~nS) + <PrJ (4)

where r is the size of a crack, <Pr is the friction angle, and JRC and JCS
are the coefficients showing the joint roughness and the mean joint
compressive strength, respectively. The shear stiffness decreases in-
versely with increase of the crack size, and also depends on the applied
normal stress in a very complicated manner. In order to avoid the
complexity, eqn (4) is linearized with respect to the normal stress, as
follows:

(5)

where g is a non-dimensional scalar, independent of the stress and the


size.
Bandis et al. (1983) have collected 450 shear stiffness values from
various kinds of cracks, including earthquake faults. These stiffness
values have been approximated by an assembly of straight lines with a
unique negative slope in a logarithmic plot of G against r (solid lines in
Fig. 2). Each line corresponds to a series tested under a specified
normal stress. The broken lines, on the other hand, are derived from
eqn (5) in which the non-dimensional scalar g is set to 200. The broken
lines have a slightly steeper negative slope than the solid lines.
However, since the collected values are widely scattered, both lines
2
10 - experiments
(Bandis, et aU
E
E --- Present model
"- 10'
&
::;
10

'"'"
<Il
: 10'
~
ro 162
<Il
.c
(/)
3
16

Fig. 2. Shear stiffness as a function of normal stress and crack size.


COUPLED STRESS AND FLUID FLOW ANALYSIS 353

can be said to be in fairly good agreement especially when the normal


stress is kept larger than 01 MPa.
Let 0 be a unit vector normal to a crack with components ni with
respect to axes Xi (i = 1,2,3). (Note that orientation of a crack is
indicated by two unit vectors 0(+) and 0(-) normal to the major plane,
with parallel but opposite directions (Fig. 1). Here 0 stands for both
o( +) and 0(-) , and is oriented over the entire solid angle Q
corresponding to the surface of a unit sphere.) Using aij as the overall
stress tensor in a representative elementary volume, the normal stress
an acting on the crack is given by aijninj' To handle equations easily in
the sequel, the shear stiffness is averaged over the entire solid angle Q
on the assumption that r is statistically independent of 0:

G-=i (2
GE(o) dQ =-g-(JijNij =-1_g
r r
(6)

where

Here a probability density function E(o) is introduced such that


E(o) dQ gives the probability of the unit normals being oriented inside
a small solid angle dQ. Note that /V;j is a symmetric tensor depending
only on the distribution of the unit normals.

2.3. Normal Stiffness


Bandis et al. (1983) have tested fresh and weathered joints of five
different rock types to investigate the relation between the applied
normal stress an and the corresponding displacement f> due to the
closure of the crack aperture. A typical example is taken from their
paper and is reproduced in Fig. 3. Such a curve is well fitted by the
following hyperbolic function:
f>
an = a - bf> (7)

where a and b are coefficients to be experimentally determined. Let H


be the secant normal stiffness defined by ani f>. Rearrangement of eqn
(7) yields
(8)
Introducing to and Ho as the initial aperture and the tangential normal
354 MASANOBU ODA

60
3,d 2nd '$I cycle
50 I
I
-; I
n. I
::E 40 I
I
'":!' 30
'" j
I

;;
j
"iii 20 ~ II

:! /;
,'",
0
z
10 '-.'.1 I I'.
.0.-/ 1", 'I V
.-:::-... /

0 002 0.04 0.06 0.08 0.10 0.12


Normal displacemenl
(mm)

Fig. 3. Normal stress vs. normal displacement (closure of aperture) curve for a
dolerite joint under repeated loading cycles. (Data are taken from Bandis et
ai., 1983.)

stiffness at an =0 respectively, the coefficients a and b then become


1 b=_I_
a=-' (9)
H o' toHo

The initial aperture exactly accords with the maximum closure that is
attained by increasing the normal stress to infinity.
An aspect ratio c is introduced as a measure of crack shape:
r
c=- (10)
to
A larger crack tends to have a wider initial aperture to (e.g. Mimuro et
ai., 1984). Accordingly it can be assumed that the ratio is kept
constant through cracks being in a rock mass.
No experimental evidence has been published to show the size effect
of the normal stiffness. It seems reasonable, however, to think that the
normal stiffness loses its value with increasing size in a manner similar
to the shear stiffness. Here the tangential normal stiffness Hoat an = 0
is assumed to be inversely proportional to the size r; i.e. Ho = hlr
where h is a constant. This assumption, together with eqns (8)-(10),
yields

(11)
COUPLED STRESS AND FLUID FLOW ANALYSIS 355

The normal stiffness is now averaged over the entire solid angle:

-i
H=
11
HE(o) dQ =- (h + caijNij ) =- h-
o r r
(12)

where
fi = h + caijNij
Referring to the experiments by Bandis et af., the values of hare
within the range from 2 x l(f MPa to 2 x 104 MPa. It depends mark-
edly on the number of cyclic loadings. If the first cycle only is
concerned, the range reduces to 2 x 102-2>< l(Y MPa in spite of the
variety of rock types tested.
In the derivation of eqns (6) and (12), fluid pressure along cracks is
considered to be zero. If this is not the case,' the corresponding
effective stress, aij, must be substituted for the total stress, aij . It is
easy to see that the parallel increase of both an and p causes no change
in the elastic deformation of springs.

3. ELASTICITY OF CRACKED ROCK MASSES

3.1. Descriptioo of Crack Geometry


Consider a representative elementary volume in a given rock mass. It
is homogeneously cut by m(V) cracks whose centers are distributed at
random in it. Then, the concentration of the centers is expressed by
m(V)
p=- (13)
V
Since there is little reliable information about the general shape of
actual cracks, disk-shaped ones with diameter r and aperture tare
assumed here. For simplicity, (0, r, t)-cracks are used if the cracks are

characterized by the following: the unit vectors normal to the cracks
are oriented inside a small solid angle dQ around 0, and the diameters
and the apertures range from r to r + dr and from t to t + dt,
respectively. Now, a probability density function E(o, r, t) is intro-
duced such that 2E(0, r, t) dQ dr dt gives the probability of (0, r, t)-
cracks and it satisfies
m
r E(o, r, t) dQ dr dt = LIm Jr r
Lo
Im Lrm
0 Jo 0 0 JO/2
2E(0, r, t) dQ dr dt = 1 (14)
356 MASANOBU ODA

where Q/2 is the half of Q corresponding to the surface of a


hemisphere, and rm and f m are the maximum sizes of rand f,
respectively.
Three features are pointed out here in relation to eqn (14): (1) If n
is statically independent of rand f, then the function E(n, r, f)
becomes E(n)/(r, f); the term E(n) has appeared in eqn (6). (2) Each
crack produces two unit normal vectors with opposite directions; then
E(n, r, f) equals E( -n, r, f). (3) Note that E(n, r, f) dQ dr df gives the
probability of the unit normals of (n, r, f)-cracks. In this case,
E(n, r, f) is defined over the entire solid angle Q. On the other hand,
2E(n, r, f) dQ dr df gives the probability of (n, r, f)-cracks where
E(n, r, f) is only defined over the half solid angle Q/2.

3.2. Elastic Strain Tensor due to the Presence of Cracks


A scanline ab of length x(j), which is parallel to a selected base vector
j, is set in a representative elementary volume. The scanline crosses
cracks at points PI' Pz, ... ,P,vOl (Fig. 4). The subscript N(j) cor-
responds to the number of cracks crossed by the scanline.
With increasing applied stress, the scanline is displaced to a new
position due to the elastic deformation. Let u be the corresponding
displacement vector (with components Uj). It is not continuous, but
admits discontinuous jumps at cracks. Two sources of elastic strain are
distinguished: ~m) is the elastic strain of the solid matrix, and ~c) is
the additional one due to the discontinuous jumps at cracks. The
corresponding total strain is assumed to be additive:
(15)

Fig. 4. Scanline parallel to a unit vector j and its associated cracks.


COUPLED STRESS AND FLUID FLOW ANALYSIS 357

Fig. 5. Discontinuous jump vector associated with a crack.

Let ben) and b(s) be jump vectors which are parallel to the normal
and maximum shear stresses on an (n, r, t)-crack, respectively (Fig. 5).
Note that ben) is parallel to n and its magnitude is proportional to the
effective normal stress. Accordingly, its component D~n) is given by

(16)

Similarly, b(s) is parallel to the direction of the maximum shear stress


with the following magnitude:
(s) 1 _, _,
D; = G (oi/n/- ok/n;nk n /) (17)

where fI and G have already been given in eqns (12) and (6),
respectively. Summing up these two vectors yields a total jump b
related to the (n, r, t)-crack.
The next step is to sum all the jumps associated with the scanline.
To this end, a column is introduced (Fig. 6): its center axis

scanline

_. a typical
(n . r)- crack

Fig. 6. A column associated with a scanline and (0, r, t)-cracks.


358 MASANOBU ODA

corresponds to the scanline, and its upper and bottom planes consist
of (n, r, t)-cracks. The area of the cross-section therefore equals
(l/4) (np7rr 2 ). Suppose that xO) is long enough to make a large volume.
Then the total number of cracks whose centers are located in the
column is estimated by multiplying the volume by p (eqn 13). It is
further multiplied by 2E(n, r, t) dQ dr dt to give the number dNO) of
(n, r, t)-cracks, as follows:

(18)

It is clearly seen that (n, r, t)-cracks are crossed by the scanline only if
their centers are placed inside the column. Accordingly, dNO) can also
be interpreted as the number of (n, r, t)-cracks crossed by the scanline.
Now the jump vector of an (n, r, t)-crack is multiplied by the
number of eqn (18). It yields the displacement induced by the
jumps of (n, r, t)-cracks along the scanline:

tl.f!:..Ol
L.J D; -"4 x p
_:rr 0) [(1It - g1) n;njnknl + 1]3 _,
gnjnlD;k r 2(n, r, t) dQ dr dt akl
(19)
To obtain the total displacement by all cracks, not restricted to
(n, r, t)-cracks, eqn (19) is integrated over Q/2, 0 ~ r ~ rm and
O~t~tm:

(20a)

where
:rrp r r r 3
m m

F;j...k = 4"" J Jo Jg r n;nj ... nkE(n, r, t) dQ dr dt (20b)


o
is a non-dimensional tensor related to the crack geometry.
In continuum mechanics, the strain tensor E;j is defined by
E..
')
=12(U I,]
.. +UJ,l
..)

By analogy, the strain tensor E~t) equivalent to the displacement of eqn


(20a) is given by

(21)
COUPLED STRESS AND FLUID FLOW ANALYSIS 359

Substituting eqn (20a) into eqn (21) yields

E~c) = [ G-~) F ijkl +4~ (DikFjI + DjkFil + DilFjk + DjlF;k) ] a~l


= Cijkla~l (22)
where C ijkl is an elastic compliance tensor equivalent to the elasticity
of cracks, satisfying the following symmetry conditions:
(23)
It is interesting to point out that eqn (22) accords with the elastic
compliance tensor obtained by Oda (1984) when a simplification of
fi = g is accepted.
3.3. Total Elastic Compliance
In order to obtain the total strain, the elastic strain of the matrix is
added to eqn (22). It is clearly shown in Fig. 7 that the elastic strain is
produced not only by the loads transmitted through the springs
(effective stress) but also by fluid pressure. In other words, the total
stress is responsible for the elastic strain. Since the matrix is assumed
to be isotropically elastic, then E~m) becomes
1
E~m) = E [(1 + V)DikDjl - VDijDkl]akl = Mijklakl (24)

where E and v are Young's modulus and Poisson's ratio of the solid
matrix, respectively, and M ijk1 is the corresponding elastic compliance
tensor.
Using eqns (22) and (24) in eqn (15), an elastic constitutive equation

Fig. 7. Elastic strain in a block (A). Note that not only forces transmitted
through contact to contact but fluid pressure contribute to the elastic
deformation.
360 MASANOBU ODA

is finally formulated by taking into account the overall effect of cracks


on the elasticity, as follows:
-
f ij = -(m)
f ij
+ f-(c)
ij =
Mijkl(Jkl
- + Cijkl(Jkl
- I
(25)
Using eqn (3), eqn (25) becomes
Eij = (M ijkl + Cijkl)i'Jkl - Cii

=Tijkli'Jkl - CijP (26)


where
(27)

Here, Tifll is introduced as the inverse tensor of T ijkl , with the


following definition:
T klmnT;;'~ij = ~(Dkl>lj + DkjD/i) (28)
It also satisfies the following symmetry conditions:
- 1
T ijkl = T-l
jikl = ijlk = k/ij
T- 1 T- 1 (29)
If both sides of eqn (26) are multiplied by T;;'~ij' we have
-
(Jij = T-ijklEkl
1- + T-1C -
ijkl kiP (30)

4. PERMEABILITY TENSOR FOR CRACKED ROCK


MASSES

If a rock mass is assumed to be an anisotropic porous medium, it


obeys Darcy's law in which the apparent flow velocity Vi is related to
the gradient (- a</>/aXi) of total hydraulic head </> through a linking
coefficient k ij called the permeability tensor:
_ - (l' a</> (l'
Vi = - k ij - = - kili (31)
Vo aXj Vo

where (l' is the gravitational acceleration, Vo is the kinematic viscosity,


and ~ is (-a</>/aXj) (e.g. Bear, 1972). It is not always guaranteed, of
course, that any rock mass can be simulated by an equivalent porous
medium having a symmetric permeability tensor. However, the study
by Long et al. (1982) has suggested that a rock mass behaves more like
a porous medium if many cracks make a sufficient number of flow
paths.
COUPLED STRESS AND FLUID FLOW ANALYSIS 361

4.1. Permeability Tensor


Let us consider a flow region having a representative elementary
volume V, and assume that water only flows through cracks (imperme-
able matrix). This assumption is commonly accepted particularly when
crystalline rocks are concerned. Then the apparent flow velocity Vi is
given by taking the average of local velocity v~e) over the associated
crack volume Vee) (Oda, 1985):

v. =.!.
I
r v =.!.V JV(C)
V Jv I
dV r v(e) dv(e)
I
(32)

Let !!!iN be a number of (n, r, t)-cracks whose centers are located


inside the flow region. To estimate the number, the probability of
(n, r, t)-cracks is multiplied by the total number m(V):
!!!iN = 2m(V)E(n, r, t) dQ dr dt (33)
Since each of the (n, r, t)-cracks carries a void volume equal to
(n/4)r 2t, the total void volume !!!iv(e) associated with (n, r, t)-cracks
becomes

!!!iVee) ="4n r 2t!!!iN ="2n m(V)r2tE(n, r, t) dQ dr dt (34)

Let J(e) be the head gradient along an (n, r, t)-crack. If the overall
head gradient J is uniformly distributed over the entire flow \~gion,
J~e) is given by (Fig. 8)

J~e) = (D ij - ninj)~ (35)


Fluid flow along (n, r, t)-cracks is idealized as laminar flow between
parallel planar plates with aperture t, and the local fluid velocity v~e)
along the crack is assumed to be given by

(36)

Fig. 8. Hydraulic gradient along a crack.


362 MASANOBU aDA

This is the so-called cubic law, except that a non-dimensional scalar )..
appears instead of 1/12. If the cracks are large enough to make full
connection among them, then the scalar may approach 1/12. If this is
not the case, the scalar is much smaller than 1/12.
Using eqns (34)-(36), eqn (32) becomes

4 1
[lrP 0 lr rrt ]
1m m
a 0 J
2 3
Vj =).. Vo ( Djj - njnJE(n, r, t) dQ dr dt ~ (37)
Q

Comparison between eqns (31) and (37) yields an equivalent per-


meability tensor kjj :
(38)
where

~j = 4
lrp l'mlrml r t njn E(n, r, t) dQ dr dt
0 0 Q
2 3
j (39)

is a symmetric, second-rank tensor depending only on the crack


geometry. Similar permeability tensors have been reported by Snow
(1969), Dienes (1982) and ada (1985).

4.2. Determination of A
Let K{p) be the permeability in a direction p (directional permeability)
which is parallel to a field gradient J, and J be the magnitude of J
(J=Jp). Then K{p) must satisfy (e.g. Bear, 1972)

(40)

Substituting Darcy's law, eqn (40) becomes


K{p) = kjjpjpj (41)
Once the permeability tensor is fixed, eqn (41) makes it possible to
calculate the directional permeability in any. direction. A reverse
relation is required when the permeability tensor is deduced from the
directional permeabilities K{p(k which are experimentally (or numeri-
cally) determined in various directions p{k) (k = 1, 2, 3, ... , m). This
problem has been solved by Kanatani (1984) so as to minimize the sum
of the square error, giving the following equation for two-dimensional
cases:

(i, j = 1,2) (42)


COUPLED STRESS AND FLUID FLOW ANALYSIS 363

Using this tensor k~c) for k ij in eqn (41), the directional permeability
[(p) is again calculated.
Long et al. (1982) have numerically determined the directional
permeabilities .K(p(k) (k = 1, 2, 3, ... , m) of two-dimensional crack
systems. In their simulation, two-dimensional cracks were treated as
line elements with flux related to aperture by. the cubic law, and the
intact rock was assumed to be impermeable. The flow region consisted
of two constant head boundaries (cf>1 > cf>2), and two boundaries with
the same linear variation in head from cf>1 to cf>2' For each crack
system, the direction of the field gradient was rotated to examine the
change of the directional permeability .K(p(k).
Two crack systems A and B are taken from the paper by Long et al.,
and are reproduced in Figs 9 and 10, together with polar diagrams
showing variation of the directional permeability. Since all information
relating to the crack geometry is available in these cases, the

I<-- 30cm ~

(a ) Crack system of sample (A)

( b) Di reel ianal permeabililies

Fig. 9. Crack system of sample A and the corresponding directional per-


meabilities. (Directional permeabilities K(p) (full curve) are taken from Long
et aI., 1982.)
364 MASANOBU ODA

(a) Crack system of sample (B)

5 (-10 '0
cm 2 )

(b) Directional permeabilities

Fig. 10. Crack system of sample B and the corresponding directional


permeabilitieso

corresponding tensors P~ and P~ for samples A and B respectively are


given in matrix form, with dimensions mm2 , as follows:

~= [ 420 -2 57] x 10-7.


0
375 -2.63] 7
P IJ
-2.57 173 ' P~ = [ -2.63 1.57 x 10- (43)

Three kinds of directional permeability are distinguished in the


polar diagrams of Figs 9 and 10: F.(p) (full curve) is the directional
permeability determined from the numerical analyses by Long et al.
(1982); [(p) (broken curve) is the directional permeability calculated
by substituting the estimated permeability tensor khC) in eqn (41); K(p)
(dotted curve) is the directional permeability calculated by substituting
the crack tensor of eqn (43) in eqn (38) on the assumption that
}, = 1/12.
If the flow region is well divided by cracks, in the case of sample B,
the substitution of }, = 1/12 in eqn (13) provides a reasonable
prediction of the directional permeability ([(p) = K(p)o For sample A,
COUPLED STRESS AND FLUID FLOW ANALYSIS 365

which is less permeable than sample B, on the other hand, K(p) differs
greatly from K(p). It is worthy of note, however, that the dotted curves
(K(p) can be well fitted to the broken curves (K(p) if A. is set to 1/184
for sample A and 1/128 for sample B. In the following, the symbol X
will be used instead of A. when A. is chosen such that eqn (38) is best
fitted to the numerically determined directional permeability.
The above examples suggest that X. depends markedly on the
connectivity among cracks. Bearing in mind that X. is a non-
dimensional scalar, assume the following relation:
(44)
where ~j is the non-dimensional, second-rank tensor of eqn (20b). In
other words, the tensor is regarded as an index measure to show the
connectivity among cracks.
Now the following symbols are used: Po is the first invariant of ~j,
and FiI and Fin are the second and third invariants of the deviatoric
tensor F;j defined by
(45)
Here Po is indicative of the isotropic density of cracks whereas FiI is an
index to show the anisotropy due to the preferred alignment of cracks
(Oda, 1982). Then eqn (44) can be rewritten as
(46)
Oda and Hatsuyama (1985) have calculated X values for various
crack systems by using the numerical analyses developed by Long et ai.
(1982). Their results are reproduced in Fig. 11 in which X is plotted
against the corresponding values of Po, using different symbols
according to the values of FiI' Based on this, some conclusions are
obtained.
When Po is less than 6, there are no unique values of X suitable for
the given crack systems. This is because the connectivity among cracks
is not enough to make flow paths. In other words, any permeability
tensors are not hydro-mechanically equivalent to the given crack
systems. When Fo is between 6 and 17, X. appears as a linear function
of Po:
(47)
where the proportional coefficient a(FiI) depends slightly on the
second invariant. From a practical point of view, however, a(FiI) can
366 MASANOBU ODA

0.10r-----------------...,

, 0.08 - ----- - - ~:_~'_1_2 ---~- r- -:J -..,-~ ----


-.~7-
" .. ....
006
....
0.04

/.::
..
'.
.
/ ..

0.02 /
t
isotropic
"
(0 - 0.2)
/
Intermediate (0.2 - 0.8)
/
anisotropic (0.8 - 1.0)
/
/

o 5 10 15 20 25

Fig. 11. Dependence of a dimensionless coefficient X on the first invariant of


Po which is indicative of connectivity among cracks.
be regarded as constant, and be set to 49 x 10-2 . When Po is greater
than 17, X essentially equals 1/12. In this case, therefore, cracks are
fully connected to give a sufficient number of flow paths.

4.3. Stress-dependent Permeability Tensor


The permeability tensor must be stress-dependent since the crack
aperture t changes in response to the effective normal stress. Using
eqns (10) and (12), we have

(48)

If the aperture is used in eqn (39), the tensor P;j becomes


1 3 _I 3 _I _I 1 _I _I _I
P;j = C3hj - c2F/ijklOkl + cfizfijklmnOk/Omn - fi3fijklmnOpOkiOmnOop (49)

where

fij ... k =4
np Itm ITm i
0 0 Q r5ninj' .. nkE(n, r, t) dQ dr dt (50)

is a tensor depending only on crack geometry and it is quite similar to


the tensor Fij ... k except that r 5 appears in place of r 3
COUPLED STRESS AND FLUID FLOW ANALYSIS 367

In the preceding sections, the four tensors N, F, f and P have been


introduced to describe the corresponding crack geometry. On the basis
of the stereological approach, ada has shown that these tensors can be
determined in terms of in situ measurable quantities of cracks such as
the crack orientation data and the crack traces on rock exposures. This
is an important aspect of the present theory since our interest is in the
practical application (ada, 1984).

4.4. Field Evidence


It becomes clear that eqn (38), together with eqns (47) and (49), can
be used as a permeability tensor which is hydraulically equivalent to a
given crack system. Some field evidence is reported here to see if the
permeability tensor provides a reasonable estimate for in situ rock
masses. To do this, a modification of these equations is given by
considering a rather special case.
A rock mass with isotropic crack geometry is stressed under an
isotropic stress tensor (a;j = a'Dij ). In such a case, h .. k is isotropic and
hence P;j is given by

(51)

where

npI0 rY(r)dr
10 =Iii =4
Tm

From eqn (12), Ii becomes


Ii = h + ea;jN;j = h + ea' (52)
Substitution of eqns (51) and (52) into eqn (38) yields a permeability
tensor for an isotropic medium:
2)./0 ( a')3
kij = 3e 3 1- hie + a' Dij (53)

It can be rewritten as
a
K = Ko( 1- hle ; air (54)

where K is the corresponding hydraulic conductivity equal to akii/3vo


and Ko is Kat a' = O. The effective pressure can be related to depth z
368 MASANOBU aDA

by assuming that earth pressure is hydrostatic (i.e. ii' = y' z where y' is
the unit weight of rock mass submerged in water):

K (
= K o 1- ')3
Yz (55)
hlc+y'z

This is an expression for predicting the change of hydraulic conduc-


tivity with depth. For the past two decades, such relations have been
investigated by many authors (e.g. Snow, 1965, 1968; Gangi, 1978).
Note that eqn (55) is characterized by a relatively simple parameter
hlc with a clear physical meaning. Either if the normal stiffness h is
small or if the aspect ratio c is large, the conductivity sharply decreases
with increasing depth. This seems quite reasonable because a flat crack
having a sinall stiffness tends to reduce the aperture easily with small
increase in effective normal stress:

( y'Z)
t = to 1 - --'--- (56)
hlc+y'z
Bianchi and Snow (1968) have measured the actual crack aperture
by means of macrophotographyand fluorescent liquid penetrants.
They have sampled cracks from various depths at ten different dam
sites and tunnels where granite and gneiss were exposed. The results
of sub-surface measurements are reproduced in Fig. 12. (The mean
aperture at the surface, i.e. at z = 0, is not included in the figure
because it is too large, 905/-lm, to be extrapolated from the sub-
surface apertures.) The broken curve is obtained if to = 200 /-lm and
hl(cy') = 20 (m) are substituted in eqn (56). Note the fairly good
agreement between the broken curve and the measurements. It can be
said therefore that eqn (56) provides a reasonable basis to estimate the
change of crack apertures with increasing depth.
Snow (1968) has analyzed the data of packer injection tests to
determine the hydraulic conductivities in each depth zone at four dam
sites (Fig. 13). Rocks of the investigated dam sites were gneiss and
granite. If K = 11 X 10-6 (m S-I) and hl(cy') = 20 (m) are used in eqn
(55), we obtain the broken curve in the figure. It is worthy of note that
a common value of hl(cy/) is used in both Figs 12 and 13. Snow has
also proposed the following empirical relation to explain the data:
log K = a + b log z (57)
where a and b are coefficients to be determined at each site. There
Fracture aperture (microns)
o 50 100 150 200 250 300
I
. ,. ./'1 I

10 .... / ..
J
/

f
20 f
.i..
I :
30 I

E .,.
. 40
I
J
0
~::>
-
,.
.1
-
'"
,I
SO
;l
0
Qj I
n 60
I
s::;
a.., I
I
0 70
I
., I
80 ,
I
I
90
,
I
I.
1= 200 ( 1 - - ' - )
20 + z
100 -'.,I -

Fig. 12. Change of crack aperture with depth. (Data are taken from Bianchi
and Snow, 1968.)

Hydraulic COnduClivity (m / S)
., -7 al 6 -. -3 .2
10 10 10 10 10 10 10

s::;
a.
850 - o upper crystal
IWO forks C1amslte
o matfOW' pOtnl dat'nstte
100 - v blue masa datns.lte
gr-atllie gneiSS

Fig. 13. Change of hydraulic conductivity with depth. (Data are taken from
Snow, 1968.)
370 MASANOBU ODA

seems to be no marked difference between eqns (55) and (57) if the


coefficients are properly chosen. The difference becomes distinct,
however, if the depth is not restricted to the range 10-300 m.

5. BASIC EQUATIONS FOR COUPLED STRESS


AND FLUID FLOW ANALYSIS

This section is to give a complete set of equations for the coupled


stress and fluid flow analysis of cracked rock masses. Biot (1941, 1955)
has published a general theory of consolidation by treating soils as
elastic porous media. Here a set of equations is given by using the
following hydro-mechanical equivalents instead of explicitly consider-
ing cracks in rock masses:

a ij = Tijk\Ekl + TijllCkl P (58)


k ij = ),.(PkJ)ij - P;j) (59)
Bars over the stress and strain tensors are omitted in these equations
since the cracked rock mass is now idealized by an equivalent
continuum.
Let u be a displacement vector of the matrix. Governing equations
are obtained from the compatibility and equilibrium conditions:
E
I)
. . + u,),1.)
= -Zl(U I,) (60)
aij,j + ptf; = 0 (61)
where Pb is the bulk mass density and h is the body force per unit
mass.
An additional equation is obtained from the mass conservation law
of fluids, together with the permeability tensor of eqn (59). Let n be
the porosity defined by V(c) IV. The mass conservation law leads to

(62)

where D/Dt is the material derivative, Pro is the fluid density and Vi is
the fluid velocity relative to the matrix. As usual, the density of the
fluid is assumed to be constant.
Using Darcy's law, the fluid velocity Vi is given by

nV i a k ij (P
= Vi_ = -- -.- + z ) (63)
Vo apro ,j
COUPLED STRESS AND FLUID FLOW ANALYSIS. 371

Substitution of eqn (63) into eqn (62) yields

(64)

Now consider the porosity in a statistically homogeneous rock mass


in which m(V) cracks exist. Each (0, r, t)-crack produces a void volume
equal to (1/4)nr 2t. Since the number of (0, r, t)-cracks is given by
2m(V) E(o, r, t) dQ dr dt, the void volume associated with them is given

by
1
:4 nm(V)r2t2E(0, r, t) dQ dr dt (65)

In order to calculate the total void volume V(c), eqn (65) is integrated
over Q/2, 0 ~ r"'; rm , and 0 ~ t ~ tm Then the porosity is given by

V(C) 1 1
n =-=-Ro--=Ka~
V c hl}l}
(66)

Finally, eqn (64) becomes

-Pro :t [~(aij - P<\)F;j] = [v~ kij(p + Proa'Z),jt (67)

This is the additional equation derived from the mass conservation


law.

6. CONCLUDING REMARKS

In relation to coupled stress and fluid flow analysis in rock masses, the
governing equations have been studied with special emphasis on the
hydro-mechanical model of geological discontinuities. On the assump-
tion that each crack can be replaced by a set of parallel planar plates
connected by springs, a rock mass is modeled as an anisotropic, elastic,
porous medium with the equivalent elastic compliance and per-
meability tensors. Then the governing equations to solve the coupled
stress and fluid flow are formulated by using the hydro-mechanical
equivalents, and some field evidence is also examined to see if the
present theory is applicable for practical purposes.
372 MASANOBU ODA

REFERENCES

Bandis, S. C., Lumsden, A. C. and Barton, N. R. (1983). Fundamentals of


rock joint deformation, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr.,
20(6),249.
Barton, N. R. and Choubey, V. (1977). The shear strength of rock joints in
theory and practice, Rock Mech., 10, 1.
Bear, J. (1972). Dynamics of Fluids in Porous Media, Elsevier, New York.
Bianchi, L. and Snow, D. T. (1968). Permeability of crystalline rock
interpreted from measured orientation and apertures of fractures, Ann.
Arid Zone, 8(2), 231.
Biot, M. A. (1941). General theory of three-dimensional consolidation, J.
Appl. Phys., 12, 155.
Biot, M. A. (1955). Theory of elasticity and consolidation for a porous
anisotropic media, J. Appl. Phys., 26, 182.
Dienes, J. K. (1982). Permeability, percolation and statistical crack mechanics.
In Issues in Rock Mechanics, R. E. Goodman and F. E. Heuze (Eds)
(Proc. 23rd Symp. on Rock Mech., Berkeley, California).
Gangi, A. F. (1978). Variation of whole and fractured porous rock per-
meability with confining pressure, Int. J. Rock Mech. Min. Sci. &
Geomech. Abstr., 15,249.
Goodman, R. E., Taylor, R. L. and Brekke, T. L. (1968). Model for the
mechanics of jointed rock, J. Soil Mech. Fdn. Div. (ASCE), 94(SM3), 637.
Kanatani, K. (1984). Distribution of directional data and fabric tensors, Int. J.
Engng Sci., 22(2), 149.
Long, J. C. S., Remer, J. S., Wilson, C. R. and Witherspoon, P. A. (1982).
Porous media equivalents for networks of discontinuous fractures, Water
Resources Res., 18(3), 645.
Marsily, G., Ledoux, E., Barbreau, A. and Margat, J. (1977). Nuclear waste
disposal: can the geologist guarantee isolation?, Science, 197(4303),519.
Mimuro, T., Kobayashi, T., Kikuchi, K., Nagai, H., Inou, M., Katoh, K. and
Veno, 1. (1984). A study on the modeling and quantitative estimation of
joint distribution in rock, Proc. 5th Japan Symp. on Rock Mech., Kyoto,
p.127.
Noorishad, 1., Ayatollahi, M. S. and Witherspoon, P. A. (1982). A finite-
element method for coupled stress and fluid flow analysis in fractured rock
masses, Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 19,185.
Oda, M. (1982). Fabric tensor for discontinuous geological materials, Soils and
Foundations, 22(4),96.
Oda, M. (1984). Sinrilarity rule of crack geometry in statistically homogeneous
rock masses, Mechanics of Materials, 3, 119.
Oda, M. (1985). Permeability tensor for discontinuous rock masses,
Geotechnique, 35(4),483.
Oda, M. and Hatsuyama, Y. (1985). Permeability tensor for jointed rock
masses, Proc. Int. Symp. on Fundamentals of Rock Joints, Sweden, p. 303.
Runchal, A. and Maini, T. (1980). The impact of a high level nuclear waste
repository on the regional ground water flow, Int. J. Rock Mech. Min. Sci.
& Geomech. Abstr., 17,253.
COUPLED STRESS AND FLUID FLOW ANALYSIS 373

Snow, D. T. (1965). A parallel-plate model of fractured permeable media,


Ph.D. thesis, University of California.
Snow, D. T. (1968). Hydraulic characteristics of fractured metamorphic rock
of the front range and implications to the Rocky Mountain Arsenal well,
Colo. Sch. Mines Q., 63, 167.
Snow, D. T. (1969). Anisotropy permeability of fractured media, Water
Resources Res., 5(6), 1273.

You might also like