You are on page 1of 10

1.

INTRODUCTION
Flow in fractures was initially estimated using the
conceptually parallel plate model [1]. In parallel
plate model an individual fracture is represented
by two infinite smooth parallel plates separated by
constant distance (aperture) between them. The
flow is assumed to be laminar with a parabolic
velocity profile across the aperture. This led to the
well-know cubic law [2] relating fluid flux to
aperture as follows:
p
D
W Q
12
3

= (1)
where Q is the volumetric flow rate, W is the
width of the fracture perpendicular to flow, D is
the aperture size, is the fluid viscosity and p is
the fluid driving pressure.
The important implication of the cubic law is that
the fluid is characterized by the separation
distance (aperture) although the velocity varies
across the distance. Real fractures, however, have
rough (irregular) surface walls. Therefore, variable
apertures in addition, locations exist where the two
surfaces of the fracture may come into contact,
thereby creating zero aperture size.
In considering the effect of aperture variation, a
simplified form of Navier-Stokes equations the
Reynolds equation has always been used as
follows [3 ~ 6]:
0
3
) , (
3
) , (
=
|
|
.
|

\
|

+
|
.
|

\
|

y
p
D
y x
p
D
x
y x y x
(2)
where p is the pressure and D
(x,y)
is the aperture
size at coordinates x, y.
Theoretically speaking, solving the Navier-Stokes
equations under complicated fracture surfaces will
provide details on pressure and flow velocity
distributions in fractures and avoid restrictions
involved in using the cubic law and Reynolds
equation and thus estimate the flow in fractures
more correctly. The solution of Navier-Stokes
equations, however, is by no means
computationally straight forward.
This paper is concerned with developing the depth
averaged flow model for estimating flows in
single fractured joint of the laboratory experiments.
The authors were motivated to perform this
research basing on the fact that the herein derived
model includes inertia term, viscous term and
fracture surface variation components which could
not be incorporated in the previous models (i.e.
cubic law and Reynolds equation). We believe this
approach avoids the restrictions involved in the
using the Reynolds equation and is
computationally tractable. We also expect that, by

ARMA/USRMS 05-867

Application of depth averaged flow model in estimating the flows in a
single rock joint

Kishida, K., Mgaya, P. and Hosoda, T.
Kyoto University, Kyoto, Japan

Copyright 2005, ARMA, American Rock Mechanics Association

This paper was prepared for presentation at Alaska Rocks 2005, The 40th U.S. Symposium on Rock Mechanics (USRMS): Rock Mechanics for Energy, Mineral and Infrastructure
Development in the Northern Regions, held in Anchorage, Alaska, June 25-29, 2005.
This paper was selected for presentation by a USRMS Program Committee following review of information contained in an abstract submitted earlier by the author(s). Contents of the paper,
as presented, have not been reviewed by ARMA/USRMS and are subject to correction by the author(s). The material, as presented, does not necessarily reflect any position of USRMS,
ARMA, their officers, or members. Electronic reproduction, distribution, or storage of any part of this paper for commercial purposes without the written consent of ARMA is prohibited.
Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgement of where
and by whom the paper was presented.
ABSTRACT: In this paper, a 2-D depth averaged flow model has been developed taking into account the effects of the aperture
variation on the natural rock joint. First, we confirm the model verification using the 1-D sinusoidal model. Then, we carry out the
estimation of flow behavior in a single rock joint under direct shear processes using our developed 2-D flow model. The model is
verified by comparing the calculated results to 1-D analytical solution of an idealized sinusoidal surface joint roughness under the
simple hydraulic conditions. Through this comparison, it is confirmed that there exists good agreement between numerical and
analytical solution. And, an increase of amplitude-to-wavelength ratio (increase of surface roughness) caused a proportionate
decrease of discharge can be confirmed.

including the friction variation component, the
resistance to flow is correctly estimated and
therefore better flow characteristics can be
determined.
2. NUMERICAL MODEL
2.1. Governing equations
The basic equations governing the flow model
herein developed, consists of continuity and
momentum equations of plane 2-D flows, obtained
by integrating the 3-D Navier-Stokes equations [7].
Eqs.3 and 4 based on Cartesian coordinate system.
0 = u (3)
( ) u P F u u
t
u
2
1
+ = +

(4)
Continuity equation
0 =

y
VD
x
UD
(5)
Momentum equations
x-direction
( ) ( ) ( )
y
UVD
x
D U
t
UD


2

( )
|
.
|

\
|

|
|
.
|

\
|

=
x
U
D
x x
D z
gD
P
x
D
b D


2
2
1
|
|
.
|

\
|

+
|
.
|

\
|

+
|
|
.
|

\
|

+
y
z
x
z
y
U
D
y
b b bx


2
2
1
|
|
.
|

\
|

+
|
.
|

\
|

+
y
z
x
z
b b sx

(6)
y-direction
( ) ( ) ( )
y
D V
x
UVD
t
VD

2

( )
|
.
|

\
|

|
|
.
|

\
|

=
x
V
D
x y
D z
gD
P
y
D
b D


2
2
1
|
|
.
|

\
|

+
|
.
|

\
|

+
|
|
.
|

\
|

+
y
z
x
z
y
V
D
y
b b
by


2
2
1
|
|
.
|

\
|

+
|
.
|

\
|

+
y
z
x
z
b b
sy

(7)
where
bx
,
by
,
sx
and
sy
are shear vectors on the
bottom and the top walls, (U, V) are averaged
velocities in x and y directions, P
D
is the pressure
at the top wall, is a momentum correction factor
(in this study, = 1.2), is coefficient of
kinematics viscosity, D is aperture depth and g is
gravitational acceleration. The wall shear stresses
are calculated from the resistance law of laminar
flow as follows:
D
U
D
V
sx bx
sy by

6
;
6
= = = = . (8)
2.2. Procedure of numerical
The standard numerical method for
incompressible fluids (HSMAC) is used [8]. It is
assumed that at the initial condition the hydraulic
variables M (= UD), N (=VD) and pressure P
D
at
time t = n.t are known. Then, the hydraulic
variables M
*
and N
*
at time step t = (n + 1)t are
calculated, explicitly. This procedure utilizes the
concept of pressure (P
D
) correction as follows:
*
,
*
, D
n
j i D j i D
P P P + = (9)
where
5 . 0 ,
1 1
2
2 2 ,
*
,
,
*
=
|
|
.
|

\
|
+
=

y x
t gD
P
j i
j i
j i D
and
y
N N
x
M M
j i j i j i j i

*
,
*
1 ,
*
,
*
, 1 *

+

=
+ +
.
M
*
and N
*
are then recalculated using the new
value of P
D
*
. This process is repeated until the
criteria for the error
*
is satisfied, then, M
*
and
N
*
are considered to be the hydraulic variables at
time step t = (n + 1)t (i.e. M
n+1
, N
n+1
).
3. ONE DIMENSIONAL SIMULATION
3.1. Sinusoidal aperture variation
One of the simplest aperture profile function that
captures some of the geometrical properties of the
roughed surface walled fracture is a duct with a
sinusoidal aperture perturbation in Fig.1. In this
study, we consider two sinusoidal surfaces put
together separated by the constant average
aperture (). Different fracture configuration is
obtained by imposing phase angle (
s
) to the
equation defining the top sinusoidal surface (Z
s
).
The phase angles,
s
= 0 and , are for in phase
and out phase configuration, respectively (Fig.1b)

(a) An idealized flow set-up

(b) Illustration of misalignment
s
of roughness
Fig.1. Geometry of sinusoidal rough-wall channel
The equations defining the top wall (Z
s
), the
bottom wall (Z
b
) and the aperture (D) are
expressed as follows:

+
|
.
|

\
|
+ =
s s
x
Z
2
cos (10)
|
.
|

\
|
=

x
Z
b
2
cos (11)
b s
Z Z D = (12)
where is the amplitude (surface roughness) of
the top and the bottom surfaces and is the wave
length of the surfaces roughness variations.
3.2. Governing equations
The governing equations consists of 1-D
continuity and momentum equations reduces from
2-D equations (Eqs.5, 6 and 7) assuming that, the
x axis is chosen so as to coincide with the
microscopic pressure gradient.
Continuity equation
( )
0 =

x
UD
(13)
Momentum equation
( ) ( )
|
.
|

\
|

x
U
D
x x
D Z
gD
P
x
D
x
D U
s D

2

2 2
1 1 |
.
|

\
|

+ |
.
|

\
|

+
x
Z
x
Z
s sx b bx

(14)
3.3. Method of solution
The solution of 1-D model is obtained from
analytical and numerical integration approaches
under simple hydraulic conditions and some
assumptions. It is assumed that the flow is under
the fully developed steady state condition. We also
consider relatively small amplitude to wave length
ratio of the order less than 0.0125. Based on these
assumptions, we utilize the concept of linear
response for pressure variation [9], which is
defined by the following expression.
'
0
P P P
D
+ = (15)
C
x
g
P
p p
+ |
.
|

\
|
+ =

2
cos
'
(16)
where P
0
is the pressure for undisturbed flow
(parallel plate flow), P is the wave induced
pressure,
p
is the amplitude of the pressure
variation and
p
is the pressure phase angle by
which the maxima precede the wave crest.

(i) Analytical solution
The expressions of the idealized sinusoidal
profiles (Z
s
, Z
b
) and the wavy induced pressure are
substituted in Eq.14. After rearranging the like
terms containing cos(nx), sin(nx), cos(2nx),
sin(2nx), cos(3nx), sin(3nx), cos(4nx) and sin(4nx),
Eq.14 is written in the following general form.
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) 0 4 sin 4 cos
3 sin 3 cos 2 sin
2 cos sin cos
0 8 7
6 5 4
3 2 1
= + + +
+ + +
+ +
f nx f nx f
nx f nx f nx f
nx f nx f nx f
(17)
where n = 2 and f
k
= f(q,
s
,
p
,
p
, , and S).
The solution of Eq.17 is obtained by equating the
coefficients of the constant terms, sins and cosine
of Eq.14 to zero, yielding a set of nine equations is
obtained. Since three unknowns have to be solved
in this case, equations f
0
, f
1
and f
2
are considered.

= + + + +
= + + + +
= + + +
0 sin cos :
0 sin cos :
0 sin cos :
5 4 3 2
2
1 2
5 4 3 2
2
1 1
4 3 2 1 0
c c c q c q c f
a b b q b q b f
a a a q a f
p p p p
p p p p
p p p p



(18)
Keeping the pressure gradient of undisturbed flow
(S) constant, a set of equations 18 is solved for q,

p
and pressure phase angle
p
. The constant C in
the expression of pressure variation is obtained by
applying the pressure boundary condition: at x = 0
and P = 0 to Eq.16 from which the constant C is -

p
cos
p
.
(ii) 1-D integral model
Eq.14 is first non-dimensionalized using the
following non dimensional variables:

= = =

=
b
b
s
s
Z
Z
Z
Z
q
q
q
x
x
D
D
' '
0
; ; ' ; ' ; '

(19)
where q = UD and
12
3
0
gS
q

= is the discharge of
the equivalent parallel plate flow with aperture
size equal to the average aperture size ().
Then, the 1-D integral model is obtained through
integrating the non-dimensionalized form of Eq.14
over one period of sinusoidal aperture variation as
follows:
0 '
'
'
'
1
'
'
'
'
1
'
'
'
1
'
1
'
1
'
6
'
'
'
'
1
'
1
0
1
0
2
2
2
2
3
0
1
0
2
2
'
2
2
'
3 3
0
1
0
3
2
2
'
0
=
(
(

|
.
|

\
|

+
(
(

|
.
|

\
|
|
|
.
|

\
|

+ + |
.
|

\
|
|
|
.
|

\
|

|
|
.
|

\
|

dx
x
D
D
dx
x
D
D
q
q
dx
x
Z
x
Z
D
q
q
gS dx
x
D
D
q
q
s b




(20)
where S is constant pressure gradient
|
|
.
|

\
|
|
|
.
|

\
|

0
'
P
x

of undisturbed parallel plate flow.
We further rewrite Eq.20 into a more general form,
for the sake of the following discussion on the
effects of few fracture parameters on flow.
0 ' '
6
'
3
0
2 3
0
1
2
2
0
=

+ |
.
|

\
|

I q
q
I q
q
gS I q
q


(21)
where

=
1
0
3 1
'
'
'
'
1
dx
x
D
D
I

(
(

|
.
|

\
|
|
|
.
|

\
|

+ + |
.
|

\
|
|
|
.
|

\
|

+ =
1
0
2
2
'
2
2
'
3 2
'
'
1
'
1
'
1
dx
x
Z
x
Z
D
I
s b


|
.
|

\
|

=
1
0
1
0
2
2
2
2
3 3
'
'
'
'
1
'
'
'
'
1
dx
x
D
D
dx
x
D
D
I .

Fig.2. Contribution of parameters I
2
and I
3
on resistance to
flow ( = 0.002 mm and phase angle
s
= )
Since Eq.20 is not practical, the results are
obtained through numerical integration. It should
be noted that, when the aperture variation is
defined by a periodic function, the integration
term I
1
is zero. For parallel plate flow, I
2
= 2 and
parameters I
1
and I
3
vanish. Under this condition,
Eq.22 reduces to cubic law of parallel plate flow.
We, then, investigate the effect of parameters I
2

and I
3
on hydraulic conductivity of a single
fracture by first rewriting Eq.22 in more tractable
form as follows:
(
(

|
.
|

\
|

=
3
2
2
6
12
'
I I
q

. (22)
Based on Fig.2 and Eq.22, it is evident that wall
friction plays a great role on flow resistance in this
case of the parameter, I
2
. The parameter, I
2
,
depends on the amplitude, , and the characteristic
length, , (wave-length) of the fractures aperture
variation. The larger the magnitude of parameter,
I
2
, is, the higher the resistance to flow occurs.
3.4. 1-D results and comparison
Since we considered small amplitude to wave-
length ratio of the order smaller than 0.0125, it is
expected that the flow behavior is close to
unidirectional laminar flow [9]. In this section,
therefore, we compare the results of the 2-D
numerical model, 1-D analytical solution and 1-D
integral model. Also the comparison is made with
Zimmerman et al.[4] results based on Reynolds
equation of the sinusoidal aperture variation of 1-
D model, which assumes that, the resistance due to
each aperture element are in series Eq.23.

(a) q'-
s
relation

(b) q'-roughness relation
Fig.3. Normalized hydraulic conductivity plotted against
fractures parameters


(a) The q and relative roughness relation

(b) The q and I
2
relation
Fig.4. Non-dimensional hydraulic conductivity plotted
against fracture parameters
2
2 5
2
2
2
1
1
2
1
'
|
.
|

\
|

+
|
|
.
|

\
|
|
.
|

\
|

q . (23)
where ( )
3
' =
h
d q .
The variation of hydraulic conductivity for
different phase angles (different fracture
geometry), while keeping the amplitude-to-
wavelength ratio, (2/) = 0.002, is shown in Fig.3
(a). It is observed that the geometry with phase
angle,
s
=0.0rad, has less resistance to hydraulic
conductivity ( ) 0 . 1
3

h
d . This implies that the
resistance to flow is approximately equal to that of
the parallel plate flow with aperture size, = 2
mm. Similar results were reported by Ge [10]
when the effect of tortuosity was not taken into
account. Thereafter, the hydraulic conductivity
decreases with increase of phase angle '
s
' and
become minimum when the phase angle
s
=
(out of phase configuration). There exists good
agreement between 2-D numerical model 1-D
analytical solution and 1-D integral model results.
This trend is in agreement with the results
obtained by Brown and Stockman [11], who
considered the numerical solution of the Reynolds
equation.
Of a particular interest is the observation that
increase of wall roughness expressed by
amplitude-to-wavelength ratio, 2/, from 0.0 to
0.08, while keeping the phase angle
s
= caused
a proportionate decrease of hydraulic conductivity
as depicted in Fig.3 (b). This is in accordance with
the practical effects of fracture's surface roughness,
that is, increases of amplitude-to-wavelength ratio
of the sinusoidal variation while keeping the
average aperture, , constant, has an implication
on an increase of fracture surface roughness which
results into an increase of resistance to flow [10].
In order to make comparison with results of
Zimmerman et al.[4], the hydraulic conductivity
was plotted against a non-dimensional roughness
parameter (/) and the skin friction parameter, I
2
,
for
s
= . Fig.4 shows the hydraulic conductivity
and relative roughness relation used by the
Reynolds equation [4], the 1-D integral and the 2-
D numerical simulation. It is observed that the
Reynolds equation overestimates the magnitude of
hydraulic conductivity as was expected in both
cases.
A Clear difference is observed when normalized
hydraulic conductivity is plotted against the skin
friction parameter, I
2
, in Fig.4. The difference of
the results keeps on increasing for higher values of
the parameter, I
2
, (i.e. when the roughness
increases). This implies that when the effect of
wall roughness becomes significant, Reynolds
equation tend to overestimate the magnitude of
hydraulic conductivity.
In Fig.5, the effect of characteristic length to
hydraulic conductivity is expressed as a ratio /
for keeping consistence with the previous studies
[3, 5]. It is observed that the results of Reynolds
equation based on the equation derived by
Zimmerman and Bodvarsson [4] is not influenced
by the variation of the characteristic length, while
the results of 1-D model herein considered shows
sensitivity.

Fig.5 Dependence of hydraulic conductivity on characteristic
length () of aperture variation ( = 2 mm, = 0.2 mm)
4. FLOW SIMULATION ON THE
MEASURED APERTURES OF THE
EXPERIMENTAL ROCK JOINTS
The 2-D model was used in the simulation of the
two dimensional computations of fluid flow within
the measured aperture fields of the experimental
rock joints. The simulated domain consisted of
specimens with dimension, 80 x 120 mm. The
conditions for simulation were based on the
experimental set up explained below.
4.1. Aperture and joint surface roughness
distribution
The key to successful simulation of flow on rock
joints is the application of the correct aperture
distribution and geometry of joint surface
roughness. In this research work, the geometry of
joint surface roughness and the aperture
distribution on the rock joint during the shearing
process are obtained through the shear mechanical
model [12]. This model utilizes the discrete
profiling data of the joint surface roughness
obtained through the no-contact roughness
profiling system. In this study, the grid size, which
depends on the measured interval of 0.25 mm was
used.
The model can simulate the shear behavior of rock
joints and express the variations of shear stress
and dilation through the profiling data on the joint
surface roughness. Moreover, it can simulate
geometry of joint surface roughness at different
stages of shearing process. Figs.6 and 7 show
examples of the joint roughness and aperture size
distribution respectively, obtained using this
model.
4.2. Description of laboratory experiments
The simultaneous direct shear and permeability
tests were carried in consideration of joint surface
roughness and material properties under constant
normal confining conditions [13]. The
experiments were conducted according to the
standard method used in several studies for
permeability test, where shear displacement,
dilation, shear stress, normal stress, transmissivity
and pore pressure can be determined. In this study,
the hydraulic head and the normal confining stress
were kept on constant at 1.0 m and 1.0 MPa,
respectively. The permeability tests were
performed at each predetermined shear
displacement up to 3 mm.
4.3. Comparison of 2-D models to experimental
results
In this section, comparison is made between the 2-
D numerical model and experimental results using
the effective transmissivity (T) of each joint, as its
use does not require knowledge of the mean
aperture. Flow rate (Q) through each fractured
rock joint of the specimens was computed using
the numerical procedure explained previously. The
effective transmissivity is calculated based on the
following equation:


Fig.6. Examples of surface roughness distribution


Fig.7. Examples of aperture size distribution [(a) shear displacement =0.25mm, (b) shear displacement =2.0mm]

H W
QL
T

= (24)
where L and W are the length and the width of
specimen, respectively, and H is Hydraulic head.
Fig.8 shows the experimental and simulation results
of shear displacement and transmissibility relation.
It is observed that for specimens A, B, C, E, F and G
the results of 2-D flow model exhibits fairly good
agreement with those obtained through experiments.
For specimen D, however, there exists a
considerable disagreement for entire range of shear
displacement, where the 2-D model underestimates
the transmissivity.
In fractures, fluids tend to follow paths with bigger
aperture sizes. Therefore, one of the important
measures of the models applicability is how best it
simulates the channeling effect with respect to the
aperture size distribution. Fig.9 shows the flow
vector distribution simulated by the 2-D model at
2.5 mm shear displacement.
(b) (a)


Fig.8. Comparison between 2-D simulation and experimental results on transmissivity of fractured rock joint

5. DISCUSSION AND CONCLUSION
On basis of the flow results of an idealized rock
joint with sinusoidal aperture variation, it is evident
that the developed 2-D model is not restricted by
the characteristic length of aperture variation ().
The effect of the characteristic length is included in
the terms expressing the surfaces' variation of the
rock joint (
s
Z and
b
Z ). This is in contrary to the
Reynolds equation which has been reported to give
reasonably accurate results up to a certain value of
the ratio (/). Zimmerman et al. [4] proposed that
Reynolds equation can be used to estimate flows in
fracture as long as the ratio (/) < 0.4, while
Brown [3] proposed (/) < 0.03. However, in real
fractures, the criteria proposed by Zimmerman et al.
[4] were not always met [10, 14]. Therefore, the
applicability of Reynolds equation is significantly
restricted.
It should be noted that the dependence of hydraulic
conductivity on characteristic length shown in
Fig.5, is not aimed to give the quantitative value
below which the Reynolds equation is applicable,
but rather to show how the 2-D model herein
developed takes into account the effect of
characteristic length.
Normally, during permeability tests in saturated
rock fractures, using the shear mechanical model,
t ransmi ssi vi t y t end t o i ncrease wi t h shear
displacement. This is based on the fact that, the
average mechanical aperture increases as result of
dilation [15]. This phenomenon is also observed in
the results presented in Fig.8. One possible
expl anat i on of t he model t o underest i mat e

(a) Aperture distribution of Specimen A (b) Velocity distribution of Specimen A

(c) Aperture distribution of Specimen B (d) Velocity distribution of Specimen B

(e) Aperture distribution of Specimen D (f) Velocity distribution of Specimen D
Fig.9. Aperture and flow vector distribution

transmissivity for specimen D is that, it is strongly
believed that, the aperture distribution was
underestimated which resulted into
underestimating of the transmissivity.
As shown in Fig.9, the brightest color refers to the
largest aperture size and vice versa. It is observed
that the model has been able to simulate the
channeling effect across the rock joint of
specimens with comparison to aperture
distribution.
In fact, the use of Reynolds equation in simulating
flow in fractures has several advantages that: it is
computationally easier and analytical solution for
simple geometry can be obtained. However, when
applied to the measured fracture apertures, where
zero apertures (contact regions) are obvious, the
Reynolds equation has severe limitations. Thus the
Reynolds equation can only be used successfully
in unobstructed areas, whereas the obstructed
areas have to be treated differently [4]. This
problem is avoided by using 2-D flow model
developed in this work, and therefore, the
channeling phenomena obtained is sufficiently
accurate.
In this study we have applied the 2-D flow model
to simulate flow in a single saturated rock
fractures. The model has shown its applicability in
simulating the flows herein considered especially
for big range of shear displacement. However,
there exists noticeable disagreement between the
experiment and simulation results for higher shear
displacement of few fractures. The authors believe
that the disagreement is caused by the effect of
aperture deformation that could not be properly
accommodated for in aperture and roughness
distribution. On the other hand since the model
takes into consideration the effect of inertia,
viscosity, wall shear and characteristic length of
aperture variation, the model introduced in this
study can serve a purpose in testing other models.

REFERENCES
1. Snow, D. T. 1965. A Parallel plate model of fractures
Permeable media, Ph.D. Dissertation, University of
Califonia, Berkeley, Califonia
2. Witherspoon, P.A., Wang, J.S.Y., Iwai. K. and Gale, J.
E. 1980. Validity of cubic law for fluid flow in a
deformable rock fracture. Water Resour. Res., 16,
1016-1024.
3. Brown, S. R. 1987. Fluid flow through rock joint: The
effect of surface roughness. J. Geophys. Res. 92, B2:
1337 1347.
4. Zimmerman, R. W., S. Kumar, and G.S. Bodvarsson.
1991. Lubrication theory analysis of the permeability
of rough-walled fractures. Int. J. Rock Mech. Min. Sci.
Geomech. Abstr. 28, No.4: 325 331.
5. Zimmerman R. W. and G. S. Bodvarsson. 1996.
Hydraulic conductivity of rock fractures. Transport in
Porous Media 23: 1 30.
6. Renshaw, C. E. 1995. On the relationship between
mechanical and hydraulic apertures in rough-walled
fractures. J. Geophys. Res. 100, No.B12: 24629
24636.
7. Batchelor, G. K. 1967 An Introduction to fluid
dynamics, pp.147 150. Cambridge University Press.
8. Hosoda, T., K. Inoue, and A. Tada. 1993. Hydraulic
transient with propagation of interface between open-
channel free surface flow and pressurized flow. Proc.
Int. Symp. On Comp. Fluid Dynamics, Sendai, Vol. 1:
291 296.
9. Patel, V.C., J.T. Chon and J.Y.Yoon. 1991. Laminar
flow over wavy walls. J. of Fluid Eng., 113: 574
578.
10. Ge, S. 1997. A governing equation for fluid flow in
rough fractures. Water Resources Research, 33, No.1:
53 61.
11. Brown, S. R. and H. W. Stockman. 1995.
Applicability of Reynolds equation for modeling fluid
flow between rough surfaces. Geophysical Research
Letters, 22, No. 22: 2537 2540.
12. Kishida, K., T. Adachi, and K. Tsuno. 2001.
Modeling of the shear behavior of rock joints under
constant normal confining conditions. Rock
Mechanics in the National Interest, Proceedings of
the 38th U.S. Rock Mechanics Symposium, Elsworth,
Tinucci & Heasly, (eds). 791 - 798. Rotterdam:
Balkema.
13. Kishida, K., T. Hosoda and A. Yamamoto. 2004.
Estimation of the Hydro-Mechanical Behavior on
Rock Joints through the Shear Model. Gulf Rock 2004,
Proceedings of the 6th North American Rock
Mechanics Symposium.
14. Gentier, S., D. Billaux and L. van Vliet. 1989.
Laboratory testing of the voids of a fracture. Rock
Mech. Rock Eng., 22: 149 157.
15. Yeo, I. W., M.H. De Freitas and R. W. Zimmerman.
1998. Effect of shear displacement on the aperture
and permeability of a rock fracture. Int. J. Rock Mech.
Min. Sci. Geomech. Abstr. 35, No.8: 1051 1070.

You might also like